Fact-checked by Grok 2 weeks ago

Graph minor

In , a graph minor is a graph H that can be obtained from another G by deleting some vertices and edges and contracting some edges, where contraction merges two adjacent vertices into a single whose incident edges are the of the originals, possibly with multiples removed. This operation preserves the essential connectivity structure while allowing for simplification, making minors a key tool for analyzing graph properties under topological transformations. The minor relation, denoted HG if H is a minor of G, defines a partial order on the set of all finite undirected , which is transitive and reflexive (every is a minor of itself). A family of is minor-closed if it contains all minors of its members, and such families are central to the study of structure, as they capture hereditary properties invariant under deletions and contractions. The foundational graph minors theorem, proved by and Paul Seymour over a series of 20 papers spanning 1983 to 2004, states that the set of finite graphs under the minor relation forms a well-quasi-order: in every infinite sequence of finite graphs, there exist indices i < j such that the i-th graph is a minor of the j-th. This result implies that every minor-closed family of graphs is characterized by a finite list of forbidden minors, resolving a long-standing conjecture and enabling the finite description of such families. Beyond its theoretical depth, graph minor theory has significant algorithmic implications, including polynomial-time (cubic) algorithms to test whether a given graph contains a fixed minor H, and broader applications in solving problems like the disjoint paths problem and computing treewidth for graphs excluding specific minors. It also underpins structural results, such as the grid theorem linking high treewidth to large grid minors, and influences conjectures like Hadwiger's on chromatic number and clique minors.

Core Concepts

Definitions

In graph theory, a graph H is a minor of a graph G if a graph isomorphic to H can be obtained from G by a sequence of vertex deletions, edge deletions, and edge contractions. This definition applies to simple undirected graphs, which have no loops or multiple edges between the same pair of vertices. Vertex deletion removes a vertex from G along with all edges incident to it. Edge deletion removes a single edge while preserving its endpoints. Edge contraction merges two adjacent vertices u and v into a single vertex, whose incident edges are the union of those originally incident to u or v; any resulting self-loops or multiple edges are then deleted to maintain simplicity. The minor relation is denoted H \preceq G when H is a minor of G, forming a partial order on the set of finite graphs up to isomorphism. Equivalently, G \succsim H indicates the same relation. A model of the minor H in G consists of vertex-disjoint connected subgraphs of G, called branch sets, one for each vertex of H; for every edge in H between vertices x and y, there is an edge in G between the corresponding branch sets V_x and V_y. These branch sets capture the structure obtained after contractions, with edges in H realized by direct connections or paths within G.

Basic Properties

The minor relation on the set of finite undirected graphs, considered up to isomorphism, forms a partial order. This relation is reflexive, as every graph is a minor of itself, obtained without any deletions or contractions. It is transitive: if H is a minor of G and K is a minor of H, then K is a minor of G, since the sequence of edge deletions, vertex deletions, and edge contractions that yield H from G can be followed by those that yield K from H. The relation is also antisymmetric, meaning that if G has H as a minor and H has G as a minor, then G and H are isomorphic. A family of graphs is minor-closed if it contains all minors of its members; that is, whenever G belongs to the family and H is a minor of G, then H also belongs to the family. This closure property follows directly from the definition of the minor relation and ensures that such families are stable under the operations of deletion and contraction. Minor-closed families play a central role in graph theory, as many natural graph properties—such as or bounded —are preserved under taking minors. One key aspect of minor-closed families is their characterization via forbidden minors: a family consists of all graphs that do not contain any graph from a fixed (possibly infinite) set as a minor. This setup provides a structural description, where membership in the family is determined by the absence of certain "obstructing" minors; a foundational result later shows that for any minor-closed family, this set of forbidden minors can be taken to be finite, though the proof of finiteness lies beyond basic properties. Minors also exhibit size relations relative to their host graphs. Specifically, if H is a minor of G, then |V(H)| \leq |V(G)| and |E(H)| \leq |E(G)|, since each operation of edge deletion, vertex deletion, or edge contraction either preserves or strictly decreases the number of vertices and edges.

Illustrative Examples

Introductory Examples

To illustrate the concept of a graph minor, consider path graphs, which provide a simple starting point. A path graph P_n on n vertices is a connected graph consisting of a sequence of n-1 edges forming a chain without branches or cycles. Any shorter path P_m with m < n is a minor of P_n, obtained by deleting the unnecessary vertices and their incident edges from the ends or middle of the path. This process preserves the linear structure while reducing the length, demonstrating how vertex and edge deletions can yield minors. Edge contractions are not typically needed here, as deletions suffice, but in more general cases, they allow merging adjacent vertices along the path to shorten it further if desired. Cycle minors offer another accessible example, highlighting the role of edge deletions. The cycle graph C_4, a 4-vertex cycle with edges forming a square, is a minor of the complete graph K_4 on 4 vertices, where every pair of vertices is connected. To obtain C_4 from K_4, delete the two diagonal edges that cross the square; the remaining four edges form exactly C_4 as a subgraph, which is trivially a minor since no contractions or vertex deletions are required beyond that. This shows how minors can simplify dense graphs by removing excess connections while retaining a cyclic structure. Trees as minors emphasize the use of edge contractions to eliminate cycles. A tree T, being an acyclic connected graph, can be obtained as a minor from a more complex graph G that contains T as a spanning subgraph with additional cycles by contracting edges within those cycles. For instance, if G has cycles attached to the branches of T, contracting each cycle's edges merges its vertices into a single vertex, effectively removing the cycle and preserving the tree's branch structure. This process, repeated as needed, transforms G into T without introducing loops or multiple edges between the same pairs. For a step-by-step visualization of contractions, consider obtaining the complete graph K_3 (a triangle) as a minor from the larger complete graph K_4. Start with K_4, which has vertices a, b, c, d and all six possible edges. Select one edge, say between a and b, and contract it: merge a and b into a new vertex w, remove the self-loop at w, and replace any parallel edges with singles (though none arise here). The resulting graph now has vertices w, c, d with edges w-c, w-d, and c-d, forming exactly K_3. This sequence illustrates how contractions reduce vertex count while maintaining completeness. A key limitation arises with disconnected graphs, underscoring that minors respect component boundaries. If a graph G is disconnected with components of sizes at most k vertices, then G cannot have a connected minor with more than k vertices, as any minor operation—deleting edges or vertices, or contracting edges—applies separately within each component and cannot bridge disconnected parts to form a larger connected structure. For example, two isolated K_2s (each a single edge) have no connected minor larger than K_2, since contractions within components yield at most paths or single vertices per part.

Advanced Examples

One prominent example of a complete graph minor in non-planar graphs is the presence of K_5 as a minor in the , obtained through a sequence of edge contractions that merge vertices into five branch sets, each corresponding to a vertex in K_5, while ensuring complete connectivity between them via the original edges. The , despite its 3-regular structure and girth of 5 (preventing a K_5 subgraph), exhibits this minor, highlighting how contractions can create denser structures in highly connected (3-vertex-connected) graphs. Kuratowski's graphs provide explicit constructions for forbidden minors in planarity: K_{3,3} and K_5. The utility graph, isomorphic to K_{3,3}, directly embeds K_{3,3} as a minor since it is the graph itself, with its three vertices on each side of the bipartition fully connected across partitions, demonstrating non-planarity through this bipartite complete structure. Similarly, K_5 embeds as a minor in non-planar graphs like the Petersen graph, where branch sets are formed by partitioning vertices and contracting paths or edges to achieve the complete 5-vertex connectivity, as no planar graph can contain such a minor. These constructions, rooted in Kuratowski's 1930 characterization of non-planar graphs via subdivisions (equivalent to minors for these graphs), illustrate how explicit vertex partitions and contractions reveal forbidden structures. The Petersen graph further exemplifies advanced minor properties, containing both K_5 and K_{3,3} as minors despite lacking a subdivision of K_5 (though containing one of K_{3,3}) due to its symmetry and degree constraints; for K_{3,3}, deletions followed by contractions yield the bipartite complete graph, underscoring its role as a minimal non-planar example with high connectivity. In graphs embeddable on the torus, grid minors offer insight into structural complexity beyond planarity. For instance, a graph on the torus with sufficient face-width contains a toroidal k \times k-grid as a minor, where the toroidal grid wraps edges around the surface; large such grids (e.g., k \geq 5) are non-planar, as they violate Euler's formula with k^2 vertices and $2k^2 edges exceeding $3k^2 - 6, thus indicating the host graph's non-planarity through this minor. This builds intuition for how surface embeddings allow denser minor structures compared to planar graphs.

Key Theorems and Conjectures

Wagner's Theorem

Wagner's theorem provides a characterization of planar graphs in terms of forbidden minors: a finite graph is planar if and only if it contains neither the complete graph K_5 nor the complete bipartite graph K_{3,3} as a minor. This result was established by Klaus Wagner in 1937, building on Kazimierz Kuratowski's earlier 1930 theorem that characterizes planarity via forbidden subdivisions. The necessity of the condition follows from the fact that the class of s is closed under minors, combined with the non-planarity of K_5 and K_{3,3}. For K_5, with 5 vertices and 10 edges, Euler's formula for s implies that a simple connected on v \geq 3 vertices satisfies e \leq 3v - 6, but here $10 > 3 \cdot 5 - 6 = 9, yielding a . For K_{3,3}, which is bipartite with 6 vertices and 9 edges, the corresponding bound for bipartite s is e \leq 2v - 4, but $9 > 2 \cdot 6 - 4 = 8; alternatively, any drawing requires at least one crossing due to the bipartite structure and edge connectivity. Thus, no can have these as minors. The sufficiency is proved by on the number of vertices. For the case of small graphs, planarity is direct. For larger graphs, consider 3-connected components: by Tutte's theorem, there exists an edge whose yields a smaller 3-connected graph, which is planar by the inductive hypothesis; the original graph is then planar by adding back the edge without crossings. Non-3-connected cases reduce to gluing planar subgraphs along vertices or edges, preserving planarity. Wagner's theorem is equivalent to Kuratowski's theorem, as the notions of subdivision and are interreducible for these forbidden configurations: any subdivision of K_5 or K_{3,3} contracts to the graph itself as a , and conversely, any such can be "expanded" by subdividing edges to form a subdivision . This equivalence highlights that minors generalize subdivisions, providing a more structural characterization. A key implication is that planarity is defined by a finite set of forbidden minors—specifically, just two—demonstrating the minor-closed nature of the family in a concrete and minimal way.

Robertson-Seymour Theorem

The Robertson–Seymour theorem, also known as the graph minors theorem, asserts that the collection of all finite undirected , partially ordered by the relation of one being a minor of another, forms a well-quasi-order. In this context, a well-quasi-order on a set is a quasi-order with no infinite descending chains (sequences where each element is strictly less than the previous) and no infinite antichains (subsets where no two elements are comparable). This property implies that in any infinite sequence of finite , there exist indices i < j such that the i-th is a minor of the j-th . A fundamental of the is that every minor-closed of —meaning a family closed under taking minors and subgraphs—has a finite set of forbidden minors that completely characterizes it: a graph belongs to the if and only if it contains none of these forbidden graphs as a minor. This finite characterization generalizes earlier results, such as Wagner's for planar graphs, to arbitrary minor-closed families. The theorem emerged from an extensive research project by and Paul Seymour, spanning 20 papers published between 1983 and 2004, which collectively resolved Wagner's 1937 that the relation on finite graphs is well-quasi-ordered. The proof, detailed primarily in the series' later installments (notably Graph Minors XVI and XX), builds on foundational work in the earlier papers establishing structural properties of graphs excluding specific . At its core, the proof employs tree decompositions—hierarchical representations of graphs that capture their connectivity via tree-structured bags of vertices—to derive structure theorems for graphs excluding a fixed minor H. These theorems describe such graphs as being built via controlled operations (clique-sums and extensions) from simpler components, such as subgraphs embeddable on surfaces or with bounded tree-width, ensuring the absence of infinite bad sequences under the minor order. The approach culminates in showing that the set of all finite graphs excluding H is well-quasi-ordered, extending inductively to the full result. Among the theorem's significant corollaries are algorithmic implications: for any fixed of forbidden minors, there exists a cubic-time (O(n^3)) to determine whether a given n- belongs to the corresponding minor-closed family. Additionally, the precludes infinite minor-closed antichains, meaning no infinite collection of graphs exists where no one is a minor of another.

Hadwiger's Conjecture

Hadwiger's conjecture, proposed by Hugo Hadwiger in 1943, states that every graph without a K_t as a minor is (t-1)-colorable for any t \geq 1. Equivalently, the conjecture asserts that the Hadwiger number \eta(G), defined as the largest t such that K_t is a minor of the graph G, satisfies \eta(G) \geq \chi(G), where \chi(G) is the chromatic number of G. This formulation links the structural property of excluding clique minors to the graph's coloring requirements. The conjecture arose as an attempt to generalize the four-color theorem, which asserts that planar graphs are 4-colorable and equivalently that graphs without K_5 minors are 4-colorable, a result later proved in 1976. Hadwiger himself verified the conjecture for t \leq 4, noting its equivalence to the four-color theorem for t=5 via Wagner's earlier work on planar graphs and K_5-minor-free graphs. For t=6, Robertson, Seymour, and established the result in 1993 by showing that every K_6-minor-free graph is 5-colorable, building on structural characterizations from graph minor theory. These cases confirm the conjecture for t \leq 6, but despite extensive searches, no counterexamples have been found for larger t, leaving it open for t \geq 7. Partial progress includes connections to the stronger but disproved Hajós conjecture from , which claimed that graphs with chromatic number at least t contain a K_t subdivision (a topological minor); Hadwiger's version weakens this by allowing contractions, providing a potential to colorings via minor exclusions. Additionally, graphs without K_t minors exhibit bounded degeneracy, as every such graph has average degree at most O(t \sqrt{\log t}), implying they are O(t \sqrt{\log t})-colorable by on degenerate orderings. This bound, independently proved by Kostochka in 1984 and Thomason in 1984, establishes that Hadwiger's conjecture holds asymptotically up to a logarithmic factor. As of , the remains unresolved, with no full proof or identified, though algorithmic advances in finding large minors offer approximations for \eta(G) relative to \chi(G). Recent improvements refine the degeneracy bound to O(t (\log \log t)^5)-colorability for K_t-minor-free graphs, narrowing the gap toward the conjectured linear bound.

Other Significant Results

One significant line of research in graph minor theory concerns the extremal function ex(n, H), which denotes the maximum number of edges in an n-vertex graph without H as a minor. Mader established that for any fixed graph H, ex(n, H) = O(n), proving that there exists a constant c(H) such that any graph with more than c(H) n edges contains H as a minor. For the specific case of complete graphs, Kostochka and Thomason independently showed that ex(n, K_r) \sim \frac{r \sqrt{\log r}}{2 \sqrt{2}} n for large r, providing an asymptotic upper bound of O(r \sqrt{\log r} , n). This linear dependence on n holds for fixed r as well, with exact values known for small r; for example, ex(n, K_3) = n-1, achieved by forests. Applications of Szemerédi's regularity lemma have yielded improved bounds on the density of minors in dense graphs. The lemma partitions the vertex set into equitable parts where most bipartite subgraphs between parts are , enabling the extraction of dense substructures that force minors. In the , Thomassen contributed to the study of and their relation to minors, particularly in characterizing linkless embeddings of graphs in 3-space. A linkless embedding is one where no two disjoint cycles form a nontrivial link. Thomassen showed that the class of graphs admitting linkless embeddings is minor-closed, and subsequent work by Robertson, Seymour, and fully characterized it by excluding the seven graphs of the Petersen family as minors; these are the forbidden minors for flat embeddings, a stronger condition where every cycle bounds a disk disjoint from the graph. Thomassen's results on highly connected sets and excluded theorems further linked embedding properties to minor exclusions. The duplication lemma provides a construction tool for graphs with prescribed minor properties by allowing vertex duplication while preserving minor-closed characteristics. Specifically, duplicating a vertex—replacing it with two adjacent vertices connected to its neighbors—maintains the absence of certain forbidden minors in families like perfect graphs or planar graphs, facilitating the building of extremal examples without introducing unwanted minors. This lemma is instrumental in inductive constructions for proving bounds on minor densities. Recent advances post-2020 have improved approximations for the Hadwiger number η(G), the size of the largest complete minor in G. Norin's 2022 ICM survey highlights that K_t-minor-free graphs are O(t \log \log t)-colorable, advancing towards Hadwiger's conjecture by reducing the logarithmic factor from previous O(t (\log t)^{1/4 + \epsilon}) bounds. For graphs with small clique number ω(G) \leq \sqrt{\log t} / (\log \log t)^2, such graphs are O(t)-colorable, supporting linear approximations in restricted cases. While semidefinite programming has been applied to related coloring relaxations via the Lovász theta function, providing O(\sqrt{n})-approximations for chromatic number that indirectly bound η(G) via the conjecture, direct SDP methods for η(G) remain exploratory in surveys up to 2025.

Minor-Closed Graph Families

Characterization and Structure

A proper minor-closed family of graphs—meaning a family closed under taking minors but not containing all graphs—admits a finite by forbidden minors. Specifically, the Robertson–Seymour theorem establishes that every such family is defined by excluding a finite set of graphs as minors, allowing the family to be precisely described as the graphs avoiding these finitely many obstructions. The structure theorem for graphs excluding a fixed minor H provides a decomposition that reveals the hierarchical organization of such graphs. Any H-minor-free graph can be built via clique-sums from subgraphs of bounded treewidth, graphs embeddable on a fixed surface, apex graphs (where removing a single vertex yields a planar graph), and vortex structures (bounded-degree expansions of cycles). This decomposition highlights the bounded complexity and layered nature of minor-closed classes. Within this framework, apex graphs play a key role in capturing near-planar structures, while layered families—those with bounded layered —emerge as essential building blocks for broader minor-closed classes. A minor-closed family has bounded layered if and only if it excludes some apex-forest as , enabling recursive constructions that maintain structural uniformity. For families defined by a fixed finite set of forbidden minors, membership testing is possible in time, as minor containment for each fixed obstruction can be decided efficiently, though the degree of the depends on the size of the forbidden graphs. The property of the minor relation, central to the Robertson–Seymour theorem, forbids infinite antichains in the poset of finite graphs ordered by minors, ensuring that minor-closed families cannot contain infinite incomparable subcollections.

Prominent Examples

Planar graphs form one of the most prominent minor-closed families, characterized precisely as those graphs that exclude both the complete graph K_5 and the K_{3,3} as minors. This finite forbidden minor characterization, established by Wagner in 1936, parallels Kuratowski's earlier theorem for subdivisions but applies directly to the minor relation. A proper subclass of planar graphs, outerplanar graphs—those embeddable in the plane with all vertices on the outer face—are characterized by forbidding K_4 and K_{2,3} as minors. This two-minor obstruction was identified by Sysło, providing a structural basis for efficient and embedding algorithms within this family. Series-parallel graphs, which arise in and have treewidth at most 2, are exactly the graphs excluding K_4 as a minor. This single forbidden minor captures their recursive via series and parallel compositions starting from edges, enabling linear-time optimization for many problems like shortest paths in such networks. The family of forests, consisting of graphs without cycles, is minor-closed and forbids only K_3 (a triangle) as a minor. Any or deletion in preserves acyclicity, making K_3 the unique obstruction that introduces a via minor operations. Toroidal graphs, embeddable on the torus surface (genus 1), form a minor-closed family with a of forbidden minors, as guaranteed by the Robertson-Seymour theorem; the complete set remains unknown, but contains at least 17,523 graphs. This complexity underscores the challenges in fully listing obstructions for higher-genus embeddings. Graphs excluding an r \times r grid as a minor represent families with bounded , a cornerstone of . The grid minor theorem states that any graph with treewidth exceeding a in r (with the current best bound \tilde{O}(r^9) as of 2025) contains an r \times r grid minor, linking minor exclusions to decomposability and algorithmic tractability.

Variants of Graph Minors

Topological Minors

A H is a topological minor of a G if G contains a subdivision of H as a , meaning that H can be obtained from a of G by replacing each with a of arbitrary greater than or equal to 1, where the paths are internally vertex-disjoint except at the endpoints corresponding to vertices of H. This concept arises from the operation of subdividing s, which is a special case of where only paths of greater than 1 are contracted to single s. Every topological of a is also a minor, since subdividing edges and then contracting them back recovers the original structure through allowed minor operations like deletions and contractions. However, the converse does not hold: a may contain H as a minor without containing a subdivision of H as a , particularly when H has of degree greater than 3 and the minor arises from contractions that merge multiple edges incident to a single . For instance, containing K_5 as a topological minor requires the existence of five connected by pairwise internally vertex-disjoint paths, implying that the is 4-connected, whereas a K_5-minor does not necessarily impose such strong connectivity requirements. For complete graphs, the presence of K_n as a topological admits a precise : a G contains K_n as a topological if and only if there exist n distinct vertices in G, each of at least n-1, such that between every pair of these vertices there is an internally vertex-disjoint . This condition ensures the branch vertices and subdivided edges form the required subdivision, and it aligns with connectivity criteria derived from for multiple disjoint paths. Topological minors play a key role in applications involving embeddings and structural graph properties. In particular, the Robertson–Seymour–Thomas theorem characterizes graphs admitting linkless embeddings in 3-dimensional space—embeddings where no two cycles are linked—precisely as those excluding any of the seven graphs in the Petersen family as topological minors. This result connects topological minors to geometric and topological constraints in 3-space, with implications for , where linkless configurations avoid knotted or linked substructures analogous to those in classical knot diagrams. In contrast to standard minor-closed graph families, which are characterized by finite sets of forbidden minors by the Robertson–Seymour graph minor theorem, topological-minor-closed families can require infinite sets of forbidden topological minors for their characterization. For example, certain structurally restricted classes, such as those embeddable on specific surfaces beyond the , may exclude infinitely many minimal topological obstructions due to the properties differing under the topological minor relation.

Induced Minors

An induced minor of a G is a H that can be obtained by first selecting an of G and then contracting a of its . Equivalently, H arises from G via a sequence of deletions followed by contractions. This operation differs from standard minors, as it prohibits deletions, thereby avoiding the creation of new between non-adjacent vertices during contractions and preserving denser structural information from the original . Induced minors form a stricter than ordinary s, implying that every induced minor is also a , but the does not hold. For instance, in chordal graphs, which lack induced cycles of length four or more, the induced-minor well-quasi-orders the subclass of graphs with bounded clique number, ensuring no infinite descending chains or antichains under this partial order. This structural rigidity limits the of induced minors in such families compared to broader minor operations. In perfect graph theory, induced minors play a key role in characterizing contraction-perfect graphs, which are those where every induced minor is perfect; these coincide with the perfect graphs themselves, forbidden by odd holes and odd antiholes as induced subgraphs. Similarly, certain induced-minor-closed graph classes, such as those excluding specific forbidden structures like induced subgraphs of bounded , are \chi-bounded, meaning their chromatic number is controlled by the number. Some classes admit finite forbidden induced minors, enabling structural characterizations analogous to the Robertson-Seymour theorem for minors. For example, trivially perfect graphs, which are (P_4, C_4)-free as induced subgraphs, form an induced-minor-closed family with a finite set of minimal forbidden induced minors, reflecting their closure under contractions.

Minors

An immersion minor of a G, also known as an immersion of a H in G, is defined by an injective mapping \phi: V(H) \to V(G) from the vertices of H to distinct vertices of G, together with a set of edge-disjoint paths in G such that for each edge uv in H, there is a path from \phi(u) to \phi(v) whose internal vertices are not in \phi(V(H)). This model ensures that the edges of H are represented by paths that do not share edges, though they may share internal vertices. Unlike contractions in standard minors, immersions preserve vertex distinctness while allowing path-based edge realizations. Immersion minors form a weaker containment relation compared to topological minors, where the representing paths must also be internally vertex-disjoint from the images of H's vertices; every topological minor is thus an immersion minor, but the converse does not hold. This distinction arises because immersions permit "edge subdivision" via paths that can intersect at non-branch vertices, without allowing vertex splitting as in contractions. The relation is incomparable to the standard minor relation, as neither implies the other in general. In applications, immersion minors extend naturally to directed graphs, where paths respect arc directions, facilitating studies in tournaments and network flow problems such as edge-disjoint . For instance, variants of Hadwiger's for immersions posit that every graph G contains an of the K_{\chi(G)}, linking chromatic number to immersion structure. A seminal result, resolving Nash-Williams' , establishes that the immersion relation well-quasi-orders finite graphs, implying no infinite antichains under immersion containment. For complete graphs, immersion containment is characterized in terms of minimum degree: a graph G with minimum degree at least $11t + 7 contains K_t as a strong immersion (where paths are internally disjoint from branch vertices). This bound improves prior estimates and highlights the role of connectivity in forcing immersions. An example occurs in tournaments, where every tournament with minimum out-degree at least C k (for a constant C) contains a 2-immersion of the complete directed graph on k vertices, with paths of length at most 2.

Shallow Minors

In , a graph H is an r-shallow minor of a G (for some nonnegative r) if H can be obtained from a of G by contracting disjoint connected , each of at most r, and then deleting and edges. The of a connected is the smallest among its , where the of a is the greatest distance to any other in the . This restriction on the size of contracted distinguishes shallow minors from standard minors, where contractions can involve arbitrarily large . The concept was introduced to facilitate the analysis of decompositions and separators, with early work attributing it to explorations in and geometric classes. Shallow minors provide a framework for studying graph sparsity and structural properties, particularly in classes where full minor exclusions are too restrictive for algorithmic or decomposition purposes. They interpolate between subgraphs (when r = 0) and arbitrary minors (as r grows large), allowing approximations of complex minor relations while controlling the "depth" of contractions to preserve local structure. This makes them suitable for metric approximations in graph embeddings, where the bounded radius aligns with notions of coarse geometry, such as Gromov-Hausdorff distances between metric spaces induced by graphs. In sparse graph classes, shallow minors help characterize families with controlled density, enabling efficient algorithms for problems like coloring and embedding that fail under full minor theory. A key property is that classes excluding large shallow minors exhibit improved decompositions, such as small balanced separators of size O(n^{1/2}) for graphs excluding a fixed r-shallow minor, generalizing Lipton-Tarjan separators for planar graphs. Shallow minors are central to the theory of bounded expansion, where a hereditary graph class \mathcal{G} has bounded expansion if the maximum average degree of its r-shallow minors is bounded by a function f_{\mathcal{G}}(r), independent of the graph size. Such classes include minor-closed families like planar graphs but extend to broader sparse structures, such as bounded-degree graphs or unit disk graphs, with applications in kernelization and property testing. In applications, shallow minors underpin results for minor-free graphs with bounded , providing polynomial-time algorithms for problems like independent sets and dominating sets in these classes. For instance, graphs excluding a fixed shallow minor admit low-distortion embeddings into minor-free graphs with controlled parameters. Recent developments in the have leveraged shallow minors to establish product structure theorems for beyond-planar graph classes, such as 1-planar and fan-planar graphs, showing they can be represented as bounded-depth shallow minors of strong products of planar graphs and paths, yielding bounds on parameters like queue-number and . These hierarchies extend to quasi-tree approximations, where shallow minor exclusions imply structural decompositions resembling tree-like hierarchies with bounded branching.

Minors with Additional Constraints

In , parity minors impose restrictions on the standard relation to preserve properties related to even or odd structures, such as cycle lengths modulo 2. Specifically, an odd- of a G is obtained via an odd H-expansion, where a is expanded from H using paths and vertices such that a 2-coloring exists on the branch sets, making inter-branch edges monochromatic and ensuring parities are preserved. This relation maintains closure under bipartiteness and planarity, as odd-minors cannot introduce odd into bipartite graphs. Bipartition parities are preserved through the proper 2-coloring of sets, which aligns with the global bipartiteness of annotated graphs where odd are limited. The seminal graph minor recognition algorithm has been extended to handle conditions, enabling polynomial-time testing for H-minors, where each in the model H in G has a specified (even or length 2). For instance, detecting an K_k-minor requires all corresponding in G to be , which involves modeling with bichromatic paths and monochromatic connections between branch sets. These constraints facilitate solving problems like finding k disjoint or cliques as minors, with runtime O(m \alpha(m,n) n) for fixed k. Signed minors extend the minor concept to signed graphs, where edges are labeled positive or negative, and operations include edge/vertex deletion, , and re-signing to maintain the sign structure. in signed graphs merges vertices while adjusting signs to preserve , defined as cycles with an even number of negative edges (product of signs equal to +1). This preserves the overall of the graph, making signed minors useful for analyzing or in social networks and structural graph properties. Classes of signed graphs embeddable on surfaces like the or are minor-closed under these operations. Oriented minors apply to directed (oriented) graphs, where a directed graph H is an oriented minor of G if H arises from a of G by contracting directed edges, thereby preserving arc directions in the model. This ensures that paths and cycles in H correspond to directed paths in G, maintaining the orientation's asymmetry without reversing arcs during contraction. Such minors are crucial for studying directed graph families, like those excluding tournaments or specific digraphs as minors. These constrained minors have applications in matching theory and orientations, where orientations allow efficient counts via the . A admits a orientation—directing edges so every even central cycle (with a unique ) has an odd number of forward edges—if it avoids certain forbidden . In bipartite graphs, this is equivalent to having no K_{3,3} as a matching minor, obtained by bicontracting degree-two vertices in central subgraphs; such avoidance enables polynomial-time matching enumeration. For near-bipartite graphs, status depends on excluding weak matching like the cubeplex or twinplex. Additional constraints in minor formation, such as even contractions, impose rules to preserve degree parities or structural evenness in specific families. In perfect graphs, even pair contractions—merging non-adjacent vertices with even-symmetric neighborhoods—yield cliques while preserving perfection, avoiding odd-hole introductions. These rules ensure minors remain in families like Eulerian or bipartite graphs without creating odd cycles.

Algorithmic Approaches

Minor Recognition Algorithms

The problem of determining whether a given graph G contains a fixed graph H as a minor, known as minor recognition, was historically challenging. Prior to the work of Robertson and Seymour, the decidability of this problem for arbitrary fixed H remained open, with no general algorithm known, even allowing exponential time; only specific cases, such as testing for small complete graph minors like K_3 (triangle detection), could be handled efficiently using methods like matrix multiplication in O(n^\omega) time where \omega \approx 2.373. Early approaches for general H relied on brute-force enumeration of potential branch sets—disjoint connected subgraphs in G corresponding to vertices of H—followed by verification of inter-set connections via edge contractions and deletions, yielding algorithms exponential in both |V(H)| and n = |V(G)|, such as O(2^{|V(H)|} n^{O(|V(H)|)}) time via recursive subdivision checks. These methods were impractical beyond very small H but provided the foundation for later developments. The Robertson–Seymour theorem resolved this by proving that every minor-closed graph family has a finite set of forbidden minors, implying the existence of a polynomial-time recognition algorithm for any fixed H. The general approach constructs such an algorithm using dynamic programming on a tree decomposition of G with treewidth bounded by a function f(|H|) derived from the theorem's structure theory; the DP table encodes partial minor models over bags of the decomposition, allowing verification of the full model in O(f(|H|) n^3) time, where the cubic dependence arises from processing subproblems involving edge and vertex operations across the decomposition. This method, detailed across their Graph Minors series, applies the finite obstruction characterization to reduce recognition to checking against the forbidden minors, though the function f grows extremely rapidly with |H|, rendering it non-practical for all but tiny H. For small fixed H, simpler naive implementations achieve O(n^3) time by simulating contractions through iterative subgraph searches—enumerating candidate connected components for each vertex of H and testing adjacency via shortest-path computations or union-find structures—avoiding the full decomposition overhead while still leveraging basic minor model definitions. For the special case where H is planar, recognition algorithms achieve linear time O(n). Robertson and Seymour's Graph Minors V provides a seminal O(n) method exploiting the bounded of planar-excluding graphs and techniques to detect minor models via disjoint path routings in surfaces, a result that predates the full theorem and highlights structural advantages for planar H. More recent algorithms, such as the Boyer–Myrvold framework, enable efficient detection of specific planar minors like K_5 or K_{3,3} (topological equivalents for planarity) by adding edges and checking for subdivisions in O(n) time using DFS-based certification and reduction rules, extending naturally to general fixed planar H through similar verifications. These linear-time methods underscore the practical gap between theoretical bounds and implementable efficiency for structured minors. In practice, the theoretical algorithms' enormous constants limit their use, leading to specialized implementations and heuristics. The SageMath software library includes a graph minor recognition function that computes a minor model by searching for branch sets via iterative contraction and isomorphism checks, suitable for graphs up to hundreds of vertices when |H| is small (e.g., |V(H)| \leq 10), often running in seconds for sparse inputs but scaling poorly beyond. For larger sparse graphs and H with hundreds of vertices, heuristic approaches dominate, prioritizing speed over guaranteed correctness for applications like network analysis. These tools emphasize conceptual reliance on minor models while adapting dynamic programming elements for real-world scalability.

Parameterized Complexity and Fixed-Parameter Tractability

The problem of determining whether a G contains a fixed H as a minor can be approached through , where the parameter is typically the size of H, denoted |V(H)|, or the of G. Parameterization by |V(H)| treats H as the small input component, allowing algorithms whose running time is f(|V(H)|) \cdot n^{O(1)} for some function f and input size n = |V(G)| + |E(G)|. In contrast, parameterization by of G exploits structural sparsity in G, even when H is arbitrary and part of the input. The H-minor containment problem is fixed-parameter tractable (FPT) when parameterized by |V(H)|. This seminal result, due to Downey and Fellows, follows from the Robertson-Seymour theorem establishing that the minor relation forms a on finite graphs, which enables dynamic programming over finitely many obstruction sets. The well-quasi-ordering implies that for any fixed H, there are only finitely many minimal forbidden minors, facilitating an FPT algorithm despite the tower-like dependence of f on |V(H)|. When parameterized by the treewidth of G, H-minor containment is also FPT, even with H unrestricted, because the property can be expressed in monadic second-order (MSO) logic. Courcelle's theorem guarantees that any MSO-definable graph property is recognizable in FPT time f(\text{treewidth}(G)) \cdot n^{O(1)} via dynamic programming on a tree decomposition of G. Some variants of minor testing exhibit hardness in parameterized complexity. For instance, the induced topological minor problem—determining if G contains an induced subdivision of H—is W-hard parameterized by |V(H)|, even when restricted to line graphs of G and H. This hardness arises from reductions from multicolored clique problems, highlighting that not all minor-like relations inherit the FPT tractability of standard minors.

References

  1. [1]
    [PDF] Graph Minors
    Its proof, due to Neil Robertson and Paul Seymour, takes well over 500 pages. So we have to be modest: of the actual proof of the graph minor theorem this ...
  2. [2]
    Graph Minor -- from Wolfram MathWorld
    A graph H is a minor of a graph G if a copy of H can be obtained from G via repeated edge deletion and/or edge contraction.
  3. [3]
    [PDF] Graph Minor Theory
    Mar 13, 2017 · 1.1 Minors. By contraction of an edge uv in a graph G we mean identification of u and v, i.e. replacement.
  4. [4]
    [PDF] Reinhard Diestel - Graph Theory
    This fifth edition of the book is again a major overhaul, in the spirit of its first and third edition. I have rewritten Chapter 12 on graph minors to take ...
  5. [5]
    Graph minors. I. Excluding a forest - ScienceDirect
    View PDF; Download full issue. Search ScienceDirect ... Graph minors. I. Excluding a forest. Author links open overlay panelNeil Robertson, P.D. Seymour.
  6. [6]
    [PDF] 12 Graph Minors - Jeff Erickson
    A minor of a graph G is a graph obtained from G by contracting edges, deleting edges, and deleting isolated vertices; a proper minor of G is any minor other ...
  7. [7]
    [PDF] Section 10.5. Kuratowski's Theorem
    Apr 6, 2023 · A minor which is isomorphic to K5 or K3,3 is a Kuratowski minor. A subdivision of K5 or K3,3 is a Kuratowski subdivision. Note 10.5. A.
  8. [8]
    [PDF] Math 228: Kuratowski's Theorem
    A topological minor of a graph G is any graph H that can be transformed into G by a series of contractions, edge deletions, and/or removal of isolated vertices.
  9. [9]
    Sur le problème des courbes gauches en Topologie - EuDML
    Kuratowski, Casimir. "Sur le problème des courbes gauches en Topologie." Fundamenta Mathematicae 15.1 (1930): 271-283. <http://eudml.org/doc/ ...
  10. [10]
    Toroidal grid minors and stretch in embedded graphs - ScienceDirect
    We investigate the toroidal expanse of an embedded graph G, that is, the size of the largest toroidal grid contained in G as a minor.
  11. [11]
    Über eine Eigenschaft der ebenen Komplexe
    Wagner, K. Über eine Eigenschaft der ebenen Komplexe. Math. Ann. 114, 570–590 (1937). https://doi.org/10.1007/BF01594196Missing: Klaus | Show results with:Klaus
  12. [12]
    [PDF] Planar Graphs and Wagner's and Kuratowski's Theorems
    Aug 29, 2015 · A subdivision of K3,3. 3. Wagner's Theorem. In 1937, Klaus Wagner came up with another characterization of planar graphs that is, in fact ...
  13. [13]
    [PDF] The proof of Tutte's and Wagner's theorems; Hadwiger's conjecture
    Oct 22, 2020 · We are now ready to finish the characterization. Theorem 6 (Wagner). A graph G is planar if and only if K5,K3,3 6 m G. Proof.
  14. [14]
    [PDF] Graph minors XX. Wagner's conjecture - Math (Princeton)
    We prove Wagner's conjecture, that for every infinite set of finite graphs, one of its members is isomorphic to a minor of another. Page 3. 1 Introduction. A ...Missing: original | Show results with:original
  15. [15]
    None
    ### Summary of Graph Minors Project by Robertson and Seymour
  16. [16]
    [PDF] Hadwiger's conjecture - Math (Princeton)
    Apr 16, 2015 · Proving that graphs with no K7-minor are 6-colourable is thus the first case of. Hadwiger's conjecture that is still open. Albar and Gonçalves[2] ...
  17. [17]
    A Relative of Hadwiger's Conjecture - SIAM.org
    A hole in $G$ is an induced cycle of length at least four. Hadwiger's conjecture from 1943 states that for every graph $G$, $h(G)\ge \chi(G)$, where $\chi(G)$ ...
  18. [18]
    Hadwiger's conjecture forK 6-free graphs | Combinatorica
    About this article. Cite this article. Robertson, N., Seymour, P. & Thomas, R. Hadwiger's conjecture forK 6-free graphs. Combinatorica 13, 279–361 (1993).
  19. [19]
    [PDF] ON THE CONJECTURE OF HAJÓS
    Various relationships between the Hajós conjecture, the Four colour Theorem and the Hadwiger Conjecture are discussed in [11]. AMS (1980) subject ...
  20. [20]
    [2006.11798] Further progress towards Hadwiger's conjecture - arXiv
    Jun 21, 2020 · In the 1980s, Kostochka and Thomason independently proved that every graph with no K_t minor has average degree O(t\sqrt{\log t}) and hence ...
  21. [21]
    Improved Bound for Hadwiger's Conjecture - ACM Digital Library
    Feb 18, 2025 · Recently, Postle improved it to O ( t ( log log t ) 6 ) -colorable. In this paper, we show that every graph with no K t -minor is ...<|control11|><|separator|>
  22. [22]
  23. [23]
    [PDF] a survey - Math (Princeton)
    Sep 19, 2023 · sharp in another sense. Let us say a class of graphs Fis minor-closed if for all G E ! F, every graph isomorphic to a minor of G is also in ! F.
  24. [24]
    [PDF] Graph Minors. XV. Giant Steps - CORE
    GRAPH MINORS. XV. GIANT STEPS. Page 3. V(S), arrangement, T-local, bridge, H-path in G (called an ``H-pathing'' in [5], due to a typographical error). Let (G ...
  25. [25]
    [PDF] Minor-Closed Graph Classes with Bounded Layered Pathwidth - arXiv
    Oct 19, 2018 · A minor-closed class has bounded layered treewidth if and only if some apex graph is not in the class. Theorem 3 generalises the above-mentioned ...
  26. [26]
    [PDF] Algorithmic Graph Minor Theory: Decomposition, Approximation ...
    This section describes the Robertson-Seymour decom- position theorem characterizing the structure of H-minor- free graphs, which we make algorithmic in this ...
  27. [27]
    Bemerkungen zum Vierfarbenproblem. - EuDML
    Wagner, K.. "Bemerkungen zum Vierfarbenproblem.." Jahresbericht der Deutschen Mathematiker-Vereinigung 46 (1936): 26-32. <http://eudml.org/doc/146109>.Missing: Klaus forbidden minors
  28. [28]
  29. [29]
    [PDF] Diestel: Graph Theory - PRIP
    There are two areas of graph theory which I find both fascinat- ing and important, especially from the perspective of pure mathematics adopted here, but which ...<|control11|><|separator|>
  30. [30]
    [PDF] Induced Minor Models. I. Structural Properties and Algorithmic ...
    A graph H is said to be an induced minor of a graph G if H can be obtained from G by a sequence of vertex deletions and edge contractions. Equivalently,.
  31. [31]
    Induced minors and well-quasi-ordering - ScienceDirect
    A graph H is an induced minor of a graph G if it can be obtained from an induced subgraph of G by contracting edges. Otherwise, G is said to be H-induced ...
  32. [32]
    [PDF] Induced Minors and Region Intersection Graphs
    Each Xu is called a branch set. A graph H is an induced minor of a graph G if H is isomorphic to a graph that can be obtained from an induced subgraph of G ...
  33. [33]
    [PDF] Contractions in perfect graphs - arXiv
    Jan 23, 2024 · characterization of perfect graphs by forbidden induced minors ... Trivially perfect graphs are interval graphs, split and interval graphs are ...
  34. [34]
    Treewidth versus clique number. I. Graph classes with a forbidden ...
    Jun 10, 2020 · Our results yield an infinite family of \chi-bounded induced-minor-closed ... bounded, leading to linear-time algorithms for k-coloring 1 ...
  35. [35]
    [PDF] Complete graph immersions and minimum degree.
    Dec 3, 2015 · An immersion of a graph H in a graph G is a one-to-one mapping ϕ : V (H) → V (G) and a collection of edge-disjoint paths in G, one for each edge.
  36. [36]
    [2305.06204] Immersions of directed graphs in tournaments - arXiv
    May 10, 2023 · If one insists in finding an immersion of a complete directed graph on k vertices then an extra condition on the tournament is necessary.Missing: K_n theory<|control11|><|separator|>
  37. [37]
    On a special case of Hadwiger's Conjecture - ResearchGate
    Aug 9, 2025 · Hadwiger's conjecture for the immersion relation posits that every graph G contains an immersion of the complete graph K χ ( G ) K_{\chi(G)} .
  38. [38]
    [PDF] Graph Minors XXIII. Nash-Williams' immersion conjecture
    We define a quasi-order of the class of all finite hypergraphs, and prove it is a well-quasi-order. This has two corollaries of interest:.
  39. [39]
    Immersions of Directed Graphs in Tournaments - Wiley Online Library
    Nov 20, 2024 · We show that every tournament with minimum out-degree at least must contain a 2-immersion of a complete digraph on vertices.
  40. [40]
    [PDF] Cop-Width, Flip-Width and Strong Colouring Numbers
    May 6, 2025 · Let G be a graph and r ≥ 0 be an integer. A graph H is an r-shallow minor of G if H can be obtained from a subgraph of G by contracting ...<|control11|><|separator|>
  41. [41]
    Shallow excluded minors and improved graph decompositions
    First consider a graph that is a square of a bounded- degree planar graph. The theory of graph minors is of little use for such graphs, since they can have ...
  42. [42]
    Sparsity: Graphs, Structures, and Algorithms | SpringerLink
    Available as PDF; Read on any device; Instant download; Own it forever. Buy eBook ... Jaroslav Nešetřil, Patrice Ossona de Mendez. Pages 3-5. A Few Problems.
  43. [43]
    Shallow Minors, Graph Products and Beyond Planar Graphs - arXiv
    Nov 24, 2021 · In this paper, we extend this line of research by utilizing shallow minors to prove analogous product structure theorems for several beyond planar graph ...Missing: Gromov- Hausdorff
  44. [44]
    [PDF] Odd Minors I - arXiv
    Apr 10, 2023 · Roughly speaking, the odd-minor relation is a restriction of the minor relation of graphs that preserves parities of cycles.
  45. [45]
    [PDF] The Graph Minor Algorithm with Parity Conditions - La Sapienza
    The main algorithmic result of the Graph Minor Theory is a polynomial-time algorithm for testing the existence of a fixed minor [55] which, combined with the ...
  46. [46]
    [PDF] Regular matroid decomposition via signed graphs
    A minor of G;ΣŮ is a signed graph that comes from G;ΣŮ by a series of the following operations: re- signing, deletion of an edge or isolated vertex, and ...<|separator|>
  47. [47]
    Is there a version of Robertson-Seymour's graph minor theorem ...
    Jan 2, 2016 · Here a signed graph is one where each edge is signed either odd or even. A cycle is odd or even according to the sum of the signs of its edges.Is there a version of Robertson-Seymour's graph minor theorem for ...Is there a polynomial-time algorithm to check if a signed graph ...More results from mathoverflow.net
  48. [48]
    [PDF] Homomorphisms of signed graphs: An update - HAL
    Oct 17, 2020 · A signed graph has signed edges. A homomorphism preserves incidence, adjacency, and the signs of closed walks. This extends classical graph ...
  49. [49]
    Some minor-closed classes of signed graphs - ScienceDirect.com
    Feb 28, 2013 · We define four minor-closed classes of signed graphs in terms of embeddability in the annulus, projective plane, torus, and Klein bottle.
  50. [50]
    Directed graph minor theorems - MathOverflow
    Jun 28, 2020 · A directed graph is a minor of another if the first can be obtained from a subgraph of the second by contracting edges.
  51. [51]
    [1711.01806] Directed Graph Minors and Serial-Parallel Width - arXiv
    Nov 6, 2017 · Abstract:Graph minors are a primary tool in understanding the structure of undirected graphs, with many conceptual and algorithmic ...Missing: oriented | Show results with:oriented
  52. [52]
    [PDF] A survey of Pfaffian orientations of graphs - Robin Thomas
    A Pfaffian orientation of a graph has every even cycle with a perfect matching having an odd number of edges directed in either direction.
  53. [53]
    Can every maximal planar graph be obtained as a minor of a planar ...
    Jul 2, 2019 · The question is whether every maximal planar graph G1 (with arbitrary even and/or odd vertex degrees) can be obtained in that way as a minor of ...
  54. [54]
    Undirected graphs - Graph Theory - SageMath Documentation
    Compute the matching polynomial of the graph \(G\). minor(). Return the vertices of a minor isomorphic to \(H\) in the current graph. pathwidth(). Compute the ...
  55. [55]
    [PDF] arXiv:1202.4419v2 [cs.DM] 6 Mar 2014
    Mar 6, 2014 · We prove in Section 5 that Induced Topological Minor is. W[1]-hard when parameterized by |V (H)|, even if G and H are line graphs. This means ...