Fact-checked by Grok 2 weeks ago

Planar graph

A planar graph is a graph that can be embedded in the such that its edges intersect only at vertices where they are incident, with no crossings elsewhere. Planar graphs form a fundamental class in , distinguished by their embeddability properties and connections to geometric and combinatorial structures. For a connected planar graph with v vertices, e edges, and f faces in a plane embedding, states that v - e + f = 2. This relation underpins many bounds on planar graphs, such as the fact that a simple planar graph satisfies e \leq 3v - 6 for v \geq 3. A key characterization of planarity is provided by Kuratowski's theorem, which asserts that a finite graph is planar if and only if it contains no that is a subdivision of the K_5 or the K_{3,3}. Equivalent formulations, such as Wagner's theorem using graph minors instead of subdivisions, highlight the structural obstructions to planarity. Planar graphs also exhibit bounded chromatic numbers; the proves that every planar graph can be properly vertex-colored with at most four colors, resolving a longstanding conjecture from the 19th century. These properties have profound implications in areas like network design, VLSI layout, and computational geometry, where avoiding crossings optimizes embeddings and algorithms. Testing planarity is efficiently solvable in linear time using algorithms based on these theorems.

Fundamentals

Definition

A planar graph is an undirected graph that can be embedded in the Euclidean plane such that no two edges intersect except possibly at their endpoints, which are the vertices. Formally, given a graph G = (V, E), it is planar if there exists a mapping of vertices to points in \mathbb{R}^2 and edges to Jordan curves connecting those points, where curves only meet at shared endpoints and not otherwise. In such an embedding, known as a plane graph, the vertices represent points, the edges represent the connecting curves, and the curves divide the plane into connected regions called faces, one of which is the unbounded outer face. Edges in a planar embedding may be represented as curves, but by Fáry's theorem, every simple planar graph admits an equivalent embedding using straight-line segments between vertices without crossings. Graphs that cannot be embedded in this manner are non-planar; the complete graph K_5 on five vertices and the complete bipartite graph K_{3,3} on two sets of three vertices each are the minimal examples of non-planar graphs, as any proper of either is planar.

Planar embeddings

A planar embedding of a graph is a representation of its vertices as distinct points in the and its edges as Jordan arcs connecting the corresponding points, such that no two arcs intersect except possibly at their endpoints, ensuring no edge crossings occur. This drawing divides the into connected regions called faces, with the unbounded outer region designated as the exterior face. Combinatorial embeddings abstract this geometric realization by focusing on the topological structure, specifying a rotation system that assigns a to the edges incident to each . This rotation system uniquely determines the embedding up to of the , capturing the relative arrangement of edges around vertices without regard to specific positions or shapes. In such an , the boundaries of faces are well-defined walks along the edges, and for 2-connected planar graphs, each face boundary forms a simple , ensuring no bridges or cut vertices disrupt the facial structure. Geometric embeddings extend the combinatorial framework by assigning concrete positions and paths to vertices and edges while preserving the non-crossing property. Polyline drawings allow edges to consist of multiple straight segments, providing flexibility in visualization. Fáry's theorem establishes that every simple planar graph admits a straight-line , where all edges are represented as single straight line segments between vertices, equivalent in topological type to any combinatorial . For 2-connected planar graphs, embeddings are unique up to reflection in the sense that the cyclic orders at vertices can be consistently reversed across the entire graph, corresponding to a of the drawing. Whitney's strengthens this for 3-connected planar graphs, asserting that any two embeddings are equivalent up to a of the that may include , meaning the combinatorial structure is rigidly determined.

Planarity Criteria

Kuratowski's theorem

Kuratowski's theorem provides a fundamental characterization of planar graphs in terms of forbidden subgraphs. It states that a finite graph is planar if and only if it contains no subgraph that is a subdivision of the complete graph K_5 or the complete bipartite graph K_{3,3}. This criterion, published in 1930, offers a precise topological condition for planarity without requiring explicit embeddings. The theorem emerged from Kazimierz Kuratowski's work on topological problems in his "Sur le problème des courbes gauches en topologie," where he addressed the embeddability of graphs on the by identifying obstructions to planarity. Kuratowski's contribution built on earlier efforts to classify non-planar graphs, providing the first complete forbidden-subgraph theorem for planarity and influencing subsequent developments in theory. A subdivision of a graph H is formed by replacing one or more edges of H with paths of positive length, where the internal vertices of these paths have degree two in the resulting graph. Such subdivisions, also known as topological minors, preserve the connectivity structure of H while allowing for "stretched" edges. In contrast, a graph minor is a more general concept obtained from a graph by a sequence of edge deletions, vertex deletions, and edge contractions, where contracting an edge merges its endpoints into a single vertex and removes any resulting loops or multiple edges. Every subdivision is a minor, but not vice versa, as contractions can simplify structures beyond mere path replacements. Kuratowski's theorem specifically uses subdivisions to capture topological obstructions to planarity. The proof of Kuratowski's theorem proceeds in two directions. The necessity—that any subdivision of K_5 or K_{3,3} renders a non-planar—follows from the fact that K_5 and K_{3,3} themselves are non-planar, and subdivisions preserve this property under planar embeddings. For sufficiency, consider a minimal non-planar G with no such subdivisions; the proof uses edge deletions and contractions to reduce G to a base case. Specifically, deleting vertices or contracting edges in G yields a smaller that remains non-planar but must eventually contradict the absence of forbidden subdivisions, leading to an embedding via on size. This approach highlights how deletions remove unnecessary structure and contractions handle branching, ensuring the holds. Illustrative examples of non-planar graphs often involve subdivisions of K_5 or K_{3,3}. The complete graph K_5 itself is a trivial subdivision and cannot be drawn without crossings due to its high density. Similarly, K_{3,3}, with two partite sets of three vertices each connected by all possible edges, forces crossings in any attempted plane drawing. The , a famous 3-regular graph on 10 vertices, contains a subdivision of K_{3,3} but no subdivision of K_5, demonstrating its non-planarity via the theorem. In the , one partite set corresponds to three outer vertices, the other to three inner vertices, with connecting paths formed by the graph's edges and spokes. This has significant implications for recognizing non-planarity: to show a is non-planar, it suffices to identify a subdivision of K_5 or K_{3,3} as a , providing a direct and constructive method without needing full algorithms. Such identifications are often feasible for small or symmetric graphs, aiding theoretical and computational analysis.

Wagner's theorem

Wagner's theorem provides a characterization of planar graphs in terms of graph minors. Specifically, a finite undirected 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. A minor of a graph G is obtained by a sequence of operations: deleting edges or isolated vertices, or contracting edges (merging two adjacent vertices into one and removing any resulting loops or multiple edges). This contrasts with subdivisions, which involve only inserting degree-2 vertices along edges to form a topological minor, without allowing contractions or deletions beyond the subgraph itself. The minor relation is thus broader and more flexible for structural analysis, as contractions can simplify graphs in ways subdivisions cannot. Wagner's work built on Kuratowski's 1930 theorem by shifting from forbidden subdivisions to forbidden minors, proving their equivalence for planarity. The proof proceeds in two directions: if a graph contains a subdivision of K_5 or K_{3,3} as a subgraph, contracting the subdivided edges yields K_5 or K_{3,3} as a minor, implying nonplanarity; conversely, any nonplanar graph contains such a subdivision by Kuratowski's theorem, which in turn implies a minor of K_5 or K_{3,3}. Wagner's early insights into minors also anticipated the broader theory, including his that the minor relation forms a well-quasi-order on s, later proven by Robertson and Seymour in their extensive series of papers from 1983 to 2004. The theorem's minor-based formulation has significant applications in the study of minor-closed graph families—classes of graphs excluding a fixed minor set—and their computational aspects. It identifies K_5 and K_{3,3} as the finite forbidden minor set for planarity, aligning with the general principle that every minor-closed family has a finite obstruction set, as established by the Robertson-Seymour theorem. Algorithmically, testing whether a contains a fixed minor like K_5 or K_{3,3} can be done in O(n^3) time, enabling efficient and membership queries for minor-closed families, though no single explicit algorithm covers all such families.

Other characterization theorems

One fundamental necessary condition for a connected to be planar is provided by , which relates the number of vertices v, edges e, and faces f (including the outer face) in any planar : v - e + f = 2. This criterion arises from the of the and holds for all connected plane graphs, though it is not sufficient on its own for determining planarity, as some non-planar graphs may also satisfy it when considering hypothetical embeddings. Planar graphs also admit structural decompositions characterized by parameters like arboricity and thickness. The arboricity of a , defined via the Nash-Williams theorem as the minimum number of forests needed to partition its edges, is at most 3 for any planar graph; this follows from the density bound e \leq 3v - 6 for simple planar graphs with v \geq 3, implying \lceil e(H)/(v(H)-1) \rceil \leq 3 over all subgraphs H. While this provides a necessary condition, graphs with arboricity at most 3 are not necessarily planar. Thickness, the minimum number of planar subgraphs whose union covers all edges, equals 1 precisely for planar graphs, offering a direct but operational characterization tied to edge-partitioning algorithms. Spectral methods offer algebraic characterizations of planarity. The Colin de Verdière invariant \mu(G) of a graph G, defined as the maximum corank of certain symmetric matrices associated with G satisfying specific strong Arnold properties, satisfies \mu(G) \leq 3 if and only if G is planar. This invariant captures topological restrictions through the multiplicity of the second-smallest eigenvalue of generalized Laplacians, providing a computable that aligns with minor-closed properties. Separator theorems further characterize the recursive structure of planar graphs. The Lipton-Tarjan theorem states that every n-vertex planar graph has a of size at most O(\sqrt{n}) that partitions the graph into two components each with at most $2n/3 vertices, enabling divide-and-conquer algorithms for planar graph problems. This property distinguishes planar graphs by their balanced separability, contrasting with denser graphs requiring larger cuts. Poset-theoretic approaches provide combinatorial characterizations. Schnyder's theorem asserts that a graph is planar the of its incidence poset of vertices and edges is at most 3. This links planarity to the minimum number of linear extensions needed to represent the poset, with applications to straight-line drawings. Geometric realizations offer another avenue. The Koebe–Andreev–Thurston theorem characterizes connected simple planar graphs as precisely those that admit a in the plane where circles correspond to vertices and tangencies to edges. For maximal planar graphs, such packings are unique up to transformations, facilitating conformal mappings and rigidity results. Post-2000 algebraic developments include criteria based on even subgraphs and cocycle spaces. A is planar it admits a where every pair of non-adjacent edges crosses an even number of times, which can be tested algebraically via the vanishing of certain Pfaffians over the cycle space. This extends Hanani-Tutte theorems and provides a parity-based for planarity detection.

Structural Properties

Euler's formula

Euler's formula establishes a fundamental relationship among the vertices, edges, and faces of a connected , where a is a planar graph equipped with a specific in the that realizes its planarity. Let v denote the number of vertices, e the number of edges, and f the number of faces, including the unbounded outer face. The formula states that v - e + f = 2. This relation, originally derived by Leonhard Euler in 1752 for the graphs of convex polyhedra, was generalized to arbitrary by in 1813 through of polyhedra onto the . The formula can be proved by induction on the number of edges e. For the base case, consider a tree, which is a connected acyclic plane graph. A tree has e = v - 1 and bounds exactly one face (the outer face), so f = 1. Substituting yields v - (v - 1) + 1 = 2. Now assume the formula holds for all connected plane graphs with fewer than e edges. For a connected plane graph G with e edges, two cases arise. If G has a vertex of degree 1, remove that vertex and its incident edge; the resulting graph G' is connected with v' = v - 1, e' = e - 1, and f' = f, so by the inductive hypothesis, v' - e' + f' = 2, which implies v - e + f = 2. Otherwise, G has minimum degree at least 2 and thus contains a cycle; remove an edge that lies on a cycle. Such an edge is incident to two distinct faces, and removing it merges those faces, yielding G' with v' = v, e' = e - 1, f' = f - 1. By induction, v - (e - 1) + (f - 1) = 2, so v - e + f = 2. This covers all cases for connected plane graphs with at least one edge, completing the induction. For disconnected plane graphs with c connected components, the formula extends to v - e + f = 1 + c, where each additional component beyond the first contributes an extra 1 to the right-hand side (accounting for the single outer face shared among all components). More generally, for embeddings of connected graphs on closed orientable surfaces of g (where the plane corresponds to g = 0), the is \chi = v - e + f = 2 - 2g; multiply-connected surfaces, such as the with g = 1, thus have \chi = 0. A key application derives a density bound for connected plane graphs (no loops or multiple edges) with v \geq 3. Since each face is bounded by at least three edges and each edge bounds at most two faces, it follows that $2e \geq 3f, so f \leq \frac{2e}{3}. Substituting into gives v - e + \frac{2e}{3} \leq 2, or v - \frac{e}{3} \leq 2, hence e \leq 3v - 6. This bound highlights the sparsity of planar graphs compared to non-planar ones.

Degree and density bounds

One of the key consequences of for connected simple planar graphs with v \geq 3 vertices and e edges is an upper bound on the number of edges. In any planar embedding, the faces satisfy the for faces: the sum of the face degrees is $2e. Assuming no multiple edges, each face has degree at least 3, so $2e \geq 3f, where f is the number of faces. Substituting Euler's relation f = e - v + 2 yields $2e \geq 3(e - v + 2), which simplifies to e \leq 3v - 6. This bound implies that the average degree $2e/v < 6, so every planar graph has average less than 6 and thus contains a of at most 5. For maximal planar graphs, where e = 3v - 6, the sum of is $6v - 12. Considering \sum (6 - \deg(v)) = 12, and noting that vertices of at most 5 contribute at least 1 to this sum while those of at least 6 contribute at most 0, there can be at most 12 vertices of at most 5. For bipartite planar graphs, the absence of odd cycles ensures every face has degree at least 4, leading to $2e \geq 4f and thus e \leq 2v - 4 for v \geq 3. These density bounds highlight the sparsity of planar graphs compared to non-planar ones; for example, the complete graph K_5 has 10 edges but violates e \leq 3 \cdot 5 - 6 = 9, confirming its non-planarity. Asymptotically, the maximum number of edges in a simple planar graph on v vertices is $3v - 6, or approximately $3v. Coin graphs, formed as tangency graphs of non-overlapping equal-radius disks in the , provide examples of dense . Such packings achieve hexagonal arrangements where each disk touches at most 6 others, yielding maximum 6 and up to nearly $3v , saturating the bound.

Dual graphs

In the context of a graph, the G^* of a connected plane G is constructed by creating a for each face of G, including the outer face, and placing an between two vertices in G^* for every in G that separates the corresponding faces; if an in G is incident to the same face on both sides, it corresponds to a in G^*, and multiple in G sharing the same pair of faces yield multiple in G^*. This ensures a between the of G and G^*, with vertices of G^* corresponding to faces of G and faces of G^* corresponding to vertices of G. The dual G^* is always a plane multigraph, inheriting a planar embedding from G by placing vertices inside their respective faces and routing dual edges across primal edges without crossings. If G is 3-connected, its dual G^* is simple (no loops or multiple edges) and unique up to , as the embedding of G is unique in this case. Moreover, every connected plane multigraph admits a plane dual, and the double dual (G^*)^* is isomorphic to G. Duality preserves : if G has v vertices, e edges, and f faces, then G^* has v^* = f vertices, e^* = e edges, and f^* = v faces, yielding v^* - e^* + f^* = f - e + v = 2 for embeddings on (or adjusted for the ). Duality facilitates applications in optimization, such as relating minimum cuts in G (as bonds) to maximum flows or cycles in the G^*, enabling shortest path algorithms in the to solve min-cut problems in the primal via duality between cycles and bonds. In , duals of 3-connected correspond to , as exemplified by the (whose has square faces) and its the (with triangular faces). Self- graphs, where G \cong G^*, include the , which is isomorphic to its own .

Special Families

Maximal planar graphs

A maximal planar graph is a simple planar graph to which no edge can be added without creating a non-planar graph. These graphs are equivalently defined as triangulations of the plane, meaning that in any planar , every face, including the unbounded outer face, is bounded by exactly three s. Maximal planar graphs exhibit several key structural properties. For a with v \geq 3 vertices, the number of s is precisely e = 3v - 6. Every such with at least four vertices is 3-connected, and the minimum of any is at least 3. These properties ensure that the achieves the maximum possible edge density among simple planar graphs while remaining embeddable without crossings. Regarding embeddings, Whitney's theorem establishes that every 3-connected planar graph has a unique combinatorial in the up to reflection; this applies directly to maximal planar graphs with at least four vertices, as they are 3-connected. Furthermore, Steinitz's theorem characterizes these graphs as the 1-skeletons of polyhedra: a graph is isomorphic to the vertex-edge graph of a 3-dimensional if and only if it is 3-connected and planar. Consequently, maximal planar graphs admit straight-line drawings in the that realize the vertices as points in position corresponding to such polyhedra. Prominent examples include the tetrahedral graph, which is the complete graph K_4 and represents the simplest non-trivial maximal planar graph with four vertices. Another is the icosahedral graph, the 1-skeleton of the with 12 vertices, all faces triangular, illustrating the connection to solids. The of maximal planar graphs has been extensively studied, particularly for unlabeled instances. The number of unlabeled maximal planar graphs on v vertices forms a known sequence, starting with 1 for v=3, 1 for v=4, 2 for v=5, 5 for v=6, and 14 for v=7, with asymptotic growth analyzed through symmetries and recursive constructions.

Outerplanar graphs

An is a planar graph that admits a planar in which all lie on the boundary of the outer face. This ensures that the graph can be drawn without crossings and with every incident to the unbounded region. Outerplanar graphs exhibit several key properties that distinguish them as a proper subclass of planar graphs. For a simple outerplanar graph with v \geq 3 vertices, the maximum number of is e \leq 2v - 3, a stricter bound than the e \leq 3v - 6 for general planar graphs. They are also series- graphs, meaning they can be constructed from a single by repeated series and compositions, and they contain no K_4 . Additionally, outerplanar graphs are precisely the graphs that are both K_4--free and K_{2,3}--free, where K_{2,3} is the with partitions of sizes 2 and 3. Characterizations of outerplanar graphs extend beyond minor exclusions. A planar graph is outerplanar if and only if it has an whose weak —formed by contracting each inner face to a and connecting dual vertices if their faces share an edge—is . For 2-connected outerplanar graphs, this weak dual is specifically a . In such embeddings, the outer face has degree equal to the number of vertices, and no inner face has degree exceeding 3 in maximal cases. Examples of outerplanar graphs include cycles, which form the simplest non-trivial instances with all vertices on a single face; trees, which are outerplanar as they have no cycles and can be embedded with all vertices on the outer boundary; and fan graphs, consisting of a central connected to all vertices of a , embeddable with the path on the outer face and the central vertex also on it. These graphs find applications in VLSI design, where their series-parallel structure allows efficient, fault-tolerant layouts in small areas with bounded wire lengths. All outerplanar graphs are planar, as the required is a special case of planar , but the converse does not hold; for instance, K_4 is planar but not outerplanar, since any places at least one inside the outer face.

Halin graphs

A Halin graph is formed by a with no degree-two vertices and at least four vertices in the without edge crossings, then connecting its leaves in determined by the to form a that avoids crossing any edges. This construction, introduced by Rudolf Halin in his investigation of minimally 3-connected graphs, yields a plane consisting of the and the added . Halin graphs are planar by their and 3-vertex-connected, as removing any two vertices leaves the connected due to the structure and the encircling . Key properties of Halin graphs include strong Hamiltonicity: every Halin graph on at least four vertices admits a cycle, and moreover, it is 1-Hamiltonian, remaining after deletion of any single . They are also Hamiltonian-connected, possessing a between any pair of distinct vertices. Halin graphs form a minor-closed family under the operation of taking graph ; while they exclude the non-planar K_{3,3} , they frequently contain the K_4 , as seen when the tree includes a of at least three adjacent to leaves. Representative examples of Halin graphs include wheel graphs, constructed from a star where the central connects to all leaves forming the outer , and the triangular prism graph, derived from a path of length two with three leaves connected cyclically. These examples illustrate the versatility of the construction, ranging from highly symmetric structures like wheels to more elongated forms based on elongated trees. Halin graphs find applications in algorithms, where their tree-plus-cycle structure enables efficient straight-line drawings with bounded height relative to the number of vertices, facilitating compact visualizations of hierarchical data.

Advanced Results

Graph enumeration

The enumeration of planar graphs involves counting both labeled and unlabeled instances, as well as their embeddings as planar maps, often leveraging recursive decompositions and generating functions derived from structural properties like levels. For labeled planar graphs on n vertices, exact counts can be obtained through recursive decompositions into 1-connected, 2-connected, and 3-connected components, allowing computation up to n=12. These decompositions rely on the unique embedding of 3-connected planar graphs (by Whitney's theorem) and map enumeration techniques. The total number g_n grows asymptotically as g_n \sim g \cdot n^{-7/2} \gamma^n n!, where \gamma \approx 27.22688 and g \approx 4.2609 \times 10^{-6}; this estimate, including limit laws for parameters like the number of edges, was established using singularity analysis of generating functions. For unlabeled planar graphs, exact enumeration is more challenging due to isomorphism considerations, but computational methods using orderly generation algorithms have determined the counts up to n=11. These numbers, obtained via programs like plantri for generating non-isomorphic instances, are as follows:
nNumber of unlabeled planar graphs
11
21
32
44
511
633
7142
8822
96966
1079853
111140916
The asymptotic growth is with base \gamma_u satisfying $27.22 < \gamma_u < 30.06, though the precise constant and subexponential terms remain open; upper bounds follow from bijections to 4-regular multigraphs, while lower bounds use labeled counts divided by n!. Planar maps, which are embeddings of graphs on or , are enumerated using generating functions based on Tutte's method and theorems from the . For rooted planar maps with f faces, the count relates to the number of edges e and vertices v via v - e + f = 2, enabling multivariate generating functions that solve functional equations for subclasses like 2-connected or triangulated maps. For 3-connected planar maps, Tutte's methods yield explicit generating functions, such as for rooted 3-connected triangulations, with asymptotic growth \sim c \cdot (256/27)^n n^{-5/2} for the number with n faces. These approaches, extended by transfer-matrix methods for restricted families (e.g., bipartite or Eulerian maps), support exact computations for maps with up to around 20 faces, with post-2010 refinements providing asymptotics for subclasses like 4-regular or bipartite 3-connected maps.

Crossing number and minors

The crossing number of a graph G, denoted \operatorname{cr}(G), is the minimum number of edge crossings in any drawing of G in the , where edges are represented by curves that intersect only at their endpoints or at proper crossings, and no three edges meet at a single interior point. A is planar if and only if \operatorname{cr}(G) = 0. The crossing number provides a quantitative measure of non-planarity, as non-planar s necessarily have \operatorname{cr}(G) \geq 1. The crossing number is minor-monotone, meaning that if H is of G, then \operatorname{cr}(H) \leq \operatorname{cr}(G). Consequently, the presence of a non-planar , such as K_5 or K_{3,3}, implies \operatorname{cr}(G) \geq 1. To relate crossing numbers to minor density, the minor crossing number \operatorname{mcr}(G) is defined as the minimum \operatorname{cr}(H) over all graphs H that contain G as . This satisfies \operatorname{mcr}(G) \leq \operatorname{cr}(G) and provides lower bounds on \operatorname{cr}(G) via the structure of supergraphs. For instance, excluding a fixed H bounds the minor crossing number of graphs without H-minors by c |V(G)|^2 for some constant c depending on H, thereby limiting their overall crossing numbers. A fundamental lower bound on the crossing number arises from Székely's crossing lemma, which states that for a simple G with n vertices and e \geq 4n , \operatorname{cr}(G) \geq \frac{e^3}{100 n^2}. This relates edge density to crossings and extends to bounds via , as contracting to form a dense preserves or increases the implied crossing density. For complete graphs K_n, the exact crossing number remains conjectured to equal \frac{1}{4} \lfloor n/2 \rfloor \lfloor (n-1)/2 \rfloor \lfloor (n-2)/2 \rfloor \lfloor (n-3)/2 \rfloor, but recent 2020s results have tightened bounds on related variants, such as the crossing number, providing asymptotic insights into the general case. The class of graphs with \operatorname{cr}(G) \leq k is minor-closed only for k=0; for k \geq 1, it is not, so no finite set of forbidden minors characterizes bounded crossing number beyond planarity. However, excluding certain minors, such as those embeddable on the torus, can enforce polynomial bounds on the crossing number, as toroidal minors introduce unavoidable crossings in planar drawings. Computing the crossing number is NP-hard, even for 3-connected cubic graphs. Recent hardness results extend this to graphs of constant path-width 12. Due to this , practical relies on heuristics, such as branch-and-cut methods that model crossings as linear programs and iteratively refine drawings, or evolutionary algorithms that minimize crossings through local optimizations. These approaches yield good approximations for graphs up to hundreds of vertices but do not guarantee optimality.

Generalizations and Extensions

Higher genus embeddings

Higher genus embeddings generalize the concept of planarity to surfaces beyond or , allowing graphs to be drawn without edge crossings on more complex topological spaces such as or higher-genus manifolds. For an orientable surface of g \geq 1, which can be visualized as a with g handles (e.g., the for g=1), the \chi is given by \chi = 2 - 2g. In a cellular of a connected G with v vertices, e , and f faces on such a surface, states that v - e + f = 2 - 2g. Embeddings can also occur on non-orientable surfaces, which lack a consistent "inside" and "outside," such as the (demigenus 1, equivalent to one cross-cap) with \chi = 1. For a non-orientable surface with k cross-caps, \chi = 2 - k, and applies similarly: v - e + f = 2 - k. The , for instance, allows embeddings of graphs like K_6 that are non-planar, but K_7 requires a higher-genus surface. These distinctions between orientable and non-orientable embeddings affect the minimal surface needed for crossing-free drawings, with non-orientable surfaces often permitting denser graphs for the same \chi. From and the assumption of simple graphs with faces of degree at least 3 (via $2e \geq 3f), a bound emerges: for a connected embeddable on a surface with \chi, the number of edges satisfies e \leq 3(v - \chi). For orientable g, this yields e \leq 3(v - 2 + 2g), generalizing the planar bound e \leq 3v - 6 (where g=0). On the (g=1, \chi=0), for example, e \leq 3v, enabling denser embeddings than in the . A classic example is the K_7, which is non-planar but embeds cellularly on the with 7 vertices, 21 edges, and 14 triangular faces, satisfying $7 - 21 + 14 = 0. This embedding is unique up to symmetry and achieves the toroidal density bound. More generally, the minimal for embedding the complete graph K_m (with m \geq 3) on an orientable surface is \lceil (m-3)(m-4)/12 \rceil, providing the exact threshold beyond which crossings are unavoidable on lower-genus surfaces. The precise determination of this genus formula for complete graphs was established by the Ringel-Youngs theorem in the 1960s, resolving the Heawood map-coloring conjecture for all surfaces except the plane (where the four-color theorem applies). Ringel and Youngs constructed explicit embeddings achieving this minimal , using current graphs and voltage assignments to triangulate surfaces optimally, confirming that K_m embeds on the surface of \lceil (m-3)(m-4)/12 \rceil without crossings. This seminal result not only bounds chromatic numbers but also underpins much of modern .

Abstract planarity concepts

Abstract planarity concepts extend the notion of planarity beyond traditional geometric embeddings in the plane, incorporating algebraic, combinatorial, and topological generalizations that capture planarity-like properties in more abstract settings. These concepts include representations, embeddings in higher dimensions without specific linkages, directed variants for hierarchical s, and measures of decomposability into planar components. Such abstractions facilitate deeper insights into and have applications in optimization, VLSI , and . Planar matroids arise as the cycle matroids of planar graphs, providing an algebraic framework where the independent sets correspond to forests in the graph. The cycle matroid of any graph is binary, meaning it is representable over the finite field \mathrm{GF}(2) using the graph's incidence matrix, where dependencies reflect even-degree subgraphs or cycles modulo 2. For planar graphs specifically, this representation aligns with their embeddability, and the matroid inherits properties like the absence of certain forbidden minors analogous to Kuratowski's theorem. Seminal work by Tutte characterized binary matroids via excluded minors, establishing the foundation for understanding how planar graphic matroids fit within this class. Linkless embeddings generalize planarity to , where a is linkless if no two disjoint cycles form a nontrivial link, meaning they are unknotted and unlinked like in a flat embedding. This concept, introduced in the , allows non-planar s such as K_5 to admit linkless embeddings while forbidding more complex intertwinings. Robertson, Seymour, and Thomas proved that a admits a linkless embedding if and only if it has no from the Petersen family, a set of seven graphs including the and complete graphs K_6 with certain subdivisions. This characterization parallels Wagner's theorem for planarity and is part of the broader Robertson-Seymour theory of graph minors. Upward planar graphs extend planarity to directed acyclic graphs (DAGs), requiring a planar where all edges point upward, typically from lower to higher y-coordinates, to visualize hierarchies without crossings. This property is useful for organizational charts or graphs, ensuring a clear . A DAG is upward planar if it admits such a , and testing this is NP-complete in general, though polynomial-time algorithms exist for restricted cases like st-graphs with a designated source and sink. The concept was formalized in the late , with early algorithms focusing on single-source DAGs to compute upward planar efficiently. Convex planar graphs require a straight-line where all faces, including the outer face, are polygons, enhancing aesthetic and computational properties in . By Fáry's , every planar graph admits a straight-line , but imposes stricter conditions on positions to ensure no reflex angles in faces. Tutte's barycentric algorithm guarantees such drawings for 3-connected planar graphs by solving a that positions interior vertices as barycenters of neighbors, resulting in all interior faces being . This approach, developed in the 1960s, underpins many modern straight-line drawing methods. Word-representable planar graphs intersect the classes of planar and word-representable graphs, where the latter are graphs that can be associated with a word over the alphabet such that adjacent vertices alternate in the word, while non-adjacent do not. Introduced in the , word-representable graphs generalize permutation graphs (which correspond to 2-letter representations), but do not include all planar graphs as a subclass; for example, odd graphs with at least five vertices are planar but non-word-representable. The intersection reveals structural overlaps like bounded width in some cases. Recent work in the has explored their connections to semi-transitive orientations and pattern avoidance, providing bounds on representation lengths and algorithmic recognizability. The theory, motivated by algebraic combinatorics on words, has grown to encompass bounds on induced subgraphs and . Additional abstract measures include thickness and arboricity, which quantify how close a graph is to being planar via decompositions. The thickness of a graph is the minimum number of planar subgraphs needed to cover its edges, with planar graphs having thickness 1 and complete graphs K_n having thickness \lfloor (n + 7)/6 \rfloor except for n=9 and n=10, where it is 3. Defined by Tutte in the , thickness relates to multilayer planar layouts in . Arboricity, the minimum number of forests whose union covers the edges, is at most 3 for planar graphs by the Nash-Williams theorem and , as the edge density satisfies m \leq 3n - 6, implying a forest decomposition bound of \lceil \max (m/(n-1)) \rceil \leq 3. This follows from the theorem's condition that arboricity is the maximum over subgraphs of \lceil e(H)/(|V(H)|-1) \rceil.

References

  1. [1]
    Planar Graphs - Discrete Mathematics - An Open Introduction
    When a connected graph can be drawn without any edges crossing, it is called planar. When a planar graph is drawn in this way, it divides the plane into ...
  2. [2]
    [PDF] Planar Graphs, part 1 25.1 Introduction
    Dec 2, 2009 · A graph is planar if it can be drawn in the plane without any crossing edges. That is, each vertex is located at one point of the plane, and a ...
  3. [3]
    Planar Graphs
    Theorem: [Euler's Formula] For a connected planar simple graph G=(V,E) with e=|E| and v=|V|, if we let r be the number of regions that are created when drawing ...
  4. [4]
    [PDF] Chapter 18 Planar Graphs
    We can now prove Euler's formula (v − e + f = 2) works in general, for any connected planar graph. Proof: by induction on the number of edges in the graph. Base ...
  5. [5]
    [PDF] Planar Graphs - Rutgers University
    This is Kuratowski's theorem from 19351. Theorem 5. A graph G is planar if and only if it contains a topological embedding of K5 or a topological embedding ...
  6. [6]
    [PDF] Kuratowski's Theorem - UChicago Math
    Aug 28, 2017 · Theorem 3.5. (Euler's Formula) For G a connected plane graph, the following relationship holds: ν − + φ = 2. Proof. We prove this by ...
  7. [7]
    The Chromatic Number of Planar Graphs - Oscar Levin
    Theorem 5.2.1. The Four Color Theorem. Any planar graph has chromatic number at most 4. This theorem has a long and interesting history.
  8. [8]
  9. [9]
    [PDF] Planar Graphs - Faculty Web Pages - Kennesaw State University
    A planar graph is still just an abstract object: a set of vertices and a set of edges. We haven't picked a particular drawing for it, and there could be many ...Missing: definition | Show results with:definition
  10. [10]
    [PDF] Lecture 12 1 Introduction to Planar Graphs 2 Generalized Laplacian ...
    Definition 1 (Planar) A graph G = (V,E) is planar if. • for each vertex i ∈ V there exists a point xi ∈ R2. • for each edge (i, j) ∈ E there exists a curve ...
  11. [11]
    [PDF] 3 Planar graphs - Purdue Computer Science
    A graph is an ordered pair of sets,. G = (V,E). The elements in V are called vertices and the elements in E are vertex pairs and called edges. We consider only ...
  12. [12]
    [PDF] Polyhedral Graphs, Complement Graphs, and Graph ... - Digital WPI
    We will also introduce Fary's theorem, without proof. Theorem 3.3 (Fary's theorem). All simple planar graphs can be drawn in the plane with straight line.<|control11|><|separator|>
  13. [13]
    [PDF] Math 228: Kuratowski's Theorem
    Suppose, to the contrary, that K3,3 is planar. Then there is a plane embedding of K3,3 satisfying v − e + f = 2, Euler's formula. Note that here, v = 6 and ...
  14. [14]
    [PDF] (Chapter 4) Planar graphs: definitions and elementary ... - TU Graz
    two difierent Jordan arcs intersect in at most one endpoint. A plane graph is a planar embedding of some planar graph. Recall: A Jordan arc is the image of ...
  15. [15]
    [PDF] Chapter 7 Planar graphs - UCSD Math
    Definition. A planar embedding of a graph is a drawing of the graph in the plane without edges crossing. A graph is planar if a planar embedding of it exists.
  16. [16]
    09-planar-graphs
    A graph is an abstract combinatorial structure that models pairwise relationships. You probably have a good intuitive idea what a graph is already.
  17. [17]
    [PDF] Combinatorial Concepts and Algorithms for Drawing Planar Graphs
    graph G together with a rotation system is called a combinatorial embedding of G. The notion of planar drawings extends to other surfaces. A graph is embeddable.
  18. [18]
    [PDF] Chapter 2 Plane Embeddings - ETH Zürich
    If G is a connected plane graph, then (G∗)∗ = G. We will see later in Section 2.3 that the dual of a 3-connected planar graph is unique (up to isomorphism) ...
  19. [19]
    [PDF] Lecture 5 5.1 Introduction 5.2 Geometric Embeddings
    Sep 16, 2004 · A graph is planar if there exists an embedding of the vertices in IR2, f : V → IR2 such that for all pairs of edges (a, b) and (c, d) in E, with ...
  20. [20]
    [PDF] Chapter 10. Planar Graphs
    Mar 12, 2023 · In an attempt to maximize what rigor we do have, we state a formal definition of a planar embedding in terms of homeomorphisms (in Note.
  21. [21]
    Fáry Theorem -- from Wolfram MathWorld
    Fáry's theorem states that any simple planar graph can be drawn in a planar straight line embedding, ie, using straight line segments for edges, none of which ...<|separator|>
  22. [22]
    [PDF] On 1-Planar Graphs with Rotation Systems
    Hence, in 1-planar embeddings, an edge may occur in up to four faces. As with planar graphs, a 1-planar embedding uniquely implies a 1-planar rotation system.<|separator|>
  23. [23]
    [PDF] Chapter 2 Plane Embeddings - ETH Zürich
    In this form, a compact way to describe a com- binatorial embedding is as a so-called rotation system that consists of a permutation π and an involution ρ ...
  24. [24]
    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/ ...
  25. [25]
    [PDF] Planar Graphs
    Our aim in this section is to prove the surprising converse, a classic theorem of Kuratowski: any graph without a topological K5 or K3,3 minor is planar. Before ...
  26. [26]
    The theorem on planar graphs - ScienceDirect.com
    A theorem for characterizing planar graphs. The proof of such a theorem was published in 1930 by Kazimierz Kuratowski.
  27. [27]
    [PDF] A short proof of Kuratowski's graph planarity criterion - TTIC
    In 1930, K. Kuratowski published his well-known graph planarity criterion [1]: a graph is planar if and only if it does not contain a subgraph, homeomorphic ...Missing: original | Show results with:original
  28. [28]
    [PDF] Section 10.5. Kuratowski's Theorem
    Apr 6, 2023 · In 1930, Kazimierz (aka. “Casimir”) Kuratowski gave the classification of nonplanar graphs in terms of subgraphs related to K5 and. K3,3 in “Sur ...Missing: original | Show results with:original
  29. [29]
    [PDF] 12 Graph Minors - Jeff Erickson
    Theorem 12.1 (Wagner [43]). A graph G is planar if and only if K5 and K3,3 are not minors of G. Wagner's thesis continued with a characterization of all ...
  30. [30]
    [PDF] Planar Graphs and Wagner's and Kuratowski's Theorems
    Aug 29, 2015 · In 1930, Kazimierz Kuratowski proved a theorem that provides a way to tell whether a graph is planar simply by checking whether it contains a ...Missing: sketch deletions
  31. [31]
    Planar Graphs and Euler's Formula
    The equation v−e+f=2 v − e + f = 2 is called Euler's formula for planar graphs . To prove this, we will want to somehow capture the idea of building up more ...
  32. [32]
    On graph thickness, geometric thickness, and separator theorems
    We investigate the relationship between geometric thickness, thickness, outerthickness, and arboricity of graphs. In particular, we prove that all graphs ...Missing: characterizations | Show results with:characterizations
  33. [33]
    [PDF] Section 10.3. Euler's Formula
    Apr 5, 2020 · Bondy and Murty state that Euler first established the formula for polyhe- dra graphs in 1752. Additional historical details can be found in ...
  34. [34]
    [PDF] Planar graphs : a historical perspective. - ThinkIR
    In 1936, he established that any planar graph G can be drawn in the plane where all the edges of G are straight lines. This important theorem is often. 40. Page ...
  35. [35]
    15.2 Euler's Formula
    Leonhard Euler (1707—1783) came up with a formula that holds true for any planar embedding of a connected graph. 🔗. Theorem 15.2.1 (Euler's Formula). If ...
  36. [36]
    [PDF] Euler Characteristic (and the Unity of Mathematics)
    Apr 9, 2016 · Euler's theorem: All planar graphs have Euler characteristic 2! ... ▷ All (orientable) surfaces without boundary are. i.e. spheres and ...
  37. [37]
    [PDF] Diestel, Graph Theory (3rd ed'n)
    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 ...
  38. [38]
    CIRCLE PACKINGS IN DIFFERENT GEOMETRIES - Project Euclid
    Oct 18, 1989 · equal size is a degree 6 circle packing of C We must show that for any degree k circle packing of C, k = 6, and as an immediate consequence ...
  39. [39]
    [PDF] Section 10.2. Duality
    Dec 3, 2023 · We introduce the idea of a face of a planar graph and define the dual G. ∗ of plane graph G. Several properties of the dual graph are given.
  40. [40]
    [PDF] Chapter 7: Planar Graphs 7.1 Plane Graphs
    A maximal planar graph is one where every face is a triangle; equivalently one where E = 3V − 6. If a graph is bipartite, one can improve the bound, since every ...
  41. [41]
    A simple and elementary proof of Whitney's unique embedding ...
    May 11, 2020 · In this note we give a short and elementary proof of a more general version of Whitney's theorem that 3-connected planar graphs have a unique embedding in the ...Missing: uniqueness | Show results with:uniqueness
  42. [42]
    [PDF] Steinitz Theorems for Orthogonal Polyhedra
    The main content of Steinitz's the- orem lies in the other direction, the statement that every. 3-connected planar graph can be represented as a polyhe- dron.
  43. [43]
    [PDF] Section 10.2. Planar Graphs Revisited
    Mar 18, 2023 · Definition. A maximal planar graph is a planar graph with the property that if any pair of nonadjacent vertices are connected by an edge, the ...
  44. [44]
    [PDF] Plane integral drawings of the platonic solid graphs with triangle ...
    Mar 11, 2021 · Abstract. The planar graphs of the platonic solids the tetrahedron, octahedron, and icosahedron can be drawn as triangulations of the plane.<|control11|><|separator|>
  45. [45]
    [PDF] Characterisation of symmetries of unlabelled triangulations and its ...
    Sep 2, 2015 · A better understanding of the symmetries of 3-connected planar graphs is therefore the key for the enumeration of unlabelled planar graphs.
  46. [46]
    Outerplanar Graph -- from Wolfram MathWorld
    An outerplanar graph is a graph that can be embedded in the plane such that all vertices lie on the outer face. Outerplanar graphs are planar and, ...
  47. [47]
    [PDF] On the maximum number of edges of outer k-planar graphs - arXiv
    May 30, 2025 · many directions. By applying Euler's Polyhedra Formula, one can easily derive an upper bound of 2n − 3 edges for an n-vertex outerplanar graph.
  48. [48]
    [PDF] Andrzej Proskurowski - Computer Science - University of Oregon
    Oct 15, 1991 · K is also a minimal forbidden minor for the outerplanar graphs. By the definition, so is the graph K. a. To see that these two graphs constitute ...
  49. [49]
    Outerplanar Graphs and Weak Duals - Semantic Scholar
    Dec 1, 1974 · Outerplanar Graphs and Weak Duals · H. Fleischner, D. Geller, F. Harary · Published 1 December 1974 · Mathematics · Journal of the Indian ...Missing: dual forest
  50. [50]
  51. [51]
    Embedding Outerplanar Graphs in Small Books - SIAM.org
    The significance for VLSI design is that any outerplanar graph can be implemented in small area in a fault-tolerant fashion. MSC codes. 05C45 · 68Q35 · 94C15 ...
  52. [52]
    Generating labeled planar graphs uniformly at random - ScienceDirect
    Jun 15, 2007 · Our approach uses recursive formulae for the exact number of labeled planar graphs with vertices and edges, based on a decomposition into 1-, 2 ...
  53. [53]
    AMS :: Journal of the American Mathematical Society
    Asymptotic enumeration and limit laws of planar graphs. HTML articles powered by AMS MathViewer. by Omer Giménez and Marc Noy;: J. Amer. Math. Soc.
  54. [54]
    [PDF] Counting planar graphs and related families of graphs
    Let gn be the number of labelled planar graphs with n vertices, and let Rn be a graph taken uniformly at random among all gn labelled planar graphs with n ...
  55. [55]
    A005470 - OEIS
    A005470 Number of unlabeled planar simple graphs with n nodes. (Formerly M1252) 36 1, 1, 2, 4, 11, 33, 142, 822, 6966
  56. [56]
    Enumeration of labelled 4-regular planar graphs II: asymptotics - arXiv
    Jan 16, 2020 · We also obtain asymptotic estimates for the number of 2- and 3-connected 4-regular planar graphs, and for the number of 4-regular simple maps, ...Missing: labeled | Show results with:labeled
  57. [57]
    [PDF] Crossing Numbers and Hard Erd}os Problems in Discrete Geometry
    We show that an old but not well-known lower bound for the crossing number of a graph yields short proofs for a number of bounds in discrete plane geometry, ...
  58. [58]
    The Minor Crossing Number | SIAM Journal on Discrete Mathematics
    The minor crossing number of a graph G is defined as the minimum crossing number of all graphs that contain G as a minor. Basic properties of this new ...
  59. [59]
    The Minor Crossing Number of Graphs with an Excluded Minor
    Aug 10, 2025 · We study the topological structure of graphs with bounded minor crossing number and obtain a new strong version of a lower bound based on the ...
  60. [60]
    [2404.13155] On the rectilinear crossing number of complete ... - arXiv
    Apr 19, 2024 · The rectilinear crossing number of a graph is the minimum number of pairs of edges that cross over all rectilinear drawings of the graph.
  61. [61]
    [PDF] The Graph Crossing Number and its Variants: A Survey
    Dec 20, 2011 · Abstract. The crossing number is a popular tool in graph drawing and visualization, but there is not really just one crossing number; ...
  62. [62]
    Crossing number is hard for cubic graphs - ScienceDirect.com
    We prove here that it is NP-hard to determine the crossing number of a simple 3-connected cubic graph. In particular, this implies that the minor-monotone ...
  63. [63]
    Crossing Number is NP-hard for Constant Path-width (and Tree-width)
    Jun 27, 2024 · In this paper we prove that computing exactly the crossing number is NP-hard even for graphs of path-width 12 (and as a result, even of tree-width 9).
  64. [64]
    A branch-and-cut approach to the crossing number problem
    Almost all graphs with a crossing number of two were detected as such by the heuristic. Only two graphs with an actual crossing number of 1 have been estimated ...
  65. [65]
    [PDF] Section 10.6. Surface Embeddings of Graphs
    Apr 30, 2021 · The sphere S0 has Euler characteristic 2 − 2k = 2 − 2(0) = 2. By ... An embedding˜G of a graph G on a surface is a circular embedding if.
  66. [66]
    [PDF] Math 228: Embedding graphs in surfaces
    As we saw in the text, a planar graph is one that can be embedded into the plane (or sphere) in such a way that no edges cross each other. For example, the ...
  67. [67]
    [PDF] Embeddings of Small Graphs on the Torus - Computer Science
    This paper studies embeddings of graphs on the torus, constructing 2-cell embeddings of vertex-transitive graphs with 12 or fewer vertices.Missing: source | Show results with:source
  68. [68]
    [PDF] The contributions of W.T. Tutte to matroid theory - MATRIX
    A GF(2)-representable matroid is called binary. For example, over. GF(2), let ... since, over GF(2), the three corresponding vectors are linearly dependent ...
  69. [69]
    [math/9301216] Linkless embeddings of graphs in $3$-space - arXiv
    Jan 1, 1993 · Abstract: We announce results about flat (linkless) embeddings of graphs in 3-space. A piecewise-linear embedding of a graph in 3-space is ...Missing: cycles 1990s Petersen family original
  70. [70]
    [PDF] Upward Planar Drawing of Single Source Acyclic Digraphs
    Abstract. An upward plane drawing of a directed acyclic graph is a plane drawing of the graph in which each directed edge is represented as a curve monotone ...Missing: seminal | Show results with:seminal
  71. [71]
    A Comprehensive Introduction to the Theory of Word-Representable ...
    May 16, 2017 · This paper offers a comprehensive introduction to the theory of word-representable graphs including the most recent developments in the area.
  72. [72]
    THE THICKNESS OF THE COMPLETE GRAPH
    The thickness t(G) of a graph G is defined as the minimum number of planar subgraphs whose union is G\ this term was proposed by Tutte (7).Missing: original | Show results with:original