Fact-checked by Grok 2 weeks ago

Complement graph

In , the complement of a undirected G = (V, E), often denoted \overline{G} or G^c, is the with the same set V in which two distinct vertices are adjacent if and only if they are not adjacent in G. The edge set of \overline{G} thus consists of all possible edges on V excluding those in E, making the union of G and \overline{G} the K_{|V|}. If G has n = |V| vertices and m = |E| edges, then \overline{G} has \binom{n}{2} - m edges, and the of each v in \overline{G} is n - 1 - \deg_G(v). A notable class of graphs related to complements are self-complementary graphs, which are isomorphic to their own complements. Such graphs must have exactly n(n-1)/4 edges, implying that self-complementary graphs on n vertices exist only when n \equiv 0 or $1 \pmod{4}. Examples include the P_4 on 4 vertices and the C_5 on 5 vertices, both of which are self-complementary. Self-complementary graphs are always connected and have diameters of 2 or 3. Complements play a key role in studying structural properties of graphs, particularly in relating cliques and sets: a of size k in G corresponds exactly to an set of size k in \overline{G}, and vice versa. This duality simplifies algorithms and proofs for problems like finding maximum or sets by transforming one into the other via the complement. In , complements are used in , where the Ramsey number R(k, l) is the smallest n such that every on n vertices contains a of size k or an set of size l (equivalently, a of size l in the complement). Additionally, the complement of an empty (no edges) is the K_n, and vice versa, highlighting their foundational symmetry in constructions.

Fundamentals

Definition

In graph theory, the complement of a simple undirected graph G = (V, E), where V is the vertex set and E is the edge set consisting of unordered pairs of distinct vertices with no loops or multiple edges, is the graph \overline{G} = (V, \overline{E}) on the same vertex set, whose edge set \overline{E} consists of all possible unordered pairs of distinct vertices from V that are not in E. Formally, \overline{E} = \bigl\{ \{u,v\} \subseteq V \mid u \neq v, \, \{u,v\} \notin E \bigr\}. This construction assumes G is simple, meaning it has no self-loops and no multiple edges between the same pair of vertices. The adjacency matrix provides a matrix representation of the complement. If A is the adjacency matrix of G, with A_{ij} = 1 if \{i,j\} \in E and $0 otherwise (and A_{ii} = 0), then the adjacency matrix of \overline{G} is J - I - A, where J is the all-ones matrix of the same order and I is the identity matrix. This concept extends to directed graphs. For a simple directed graph G = (V, E) without self-loops, where edges are ordered pairs of distinct vertices, the complement \overline{G} = (V, \overline{E}) has an from u to v (with u \neq v) there is no such arc in G, so \overline{E} is the set of all possible directed edges between distinct vertices minus E. The complement of any graph G on n = |V| vertices relates to the K_n, which has all possible edges, as \overline{G} consists precisely of the edges absent in G.

Basic Properties

The complement of a simple G with vertex set V and n = |V| preserves the vertex set, so the complement \overline{G} has exactly n vertices. Complementation reverses adjacency relations while maintaining the symmetry of non-adjacency, as two vertices are adjacent in \overline{G} precisely when they are non-adjacent in G. A key structural consequence is the relationship between vertex degrees in G and \overline{G}. For any vertex v \in V with degree d_G(v) in G, the degree of v in \overline{G} is d_{\overline{G}}(v) = n - 1 - d_G(v), since v connects to all other vertices except itself and its neighbors in G. The edge count in \overline{G} follows directly from the total possible edges in the on n vertices. If G has m edges, then \overline{G} has \binom{n}{2} - m = \frac{n(n-1)}{2} - m edges. The union of G and \overline{G} forms the complete graph K_n, as every pair of distinct vertices is adjacent in exactly one of the two graphs.

Examples and Applications

Illustrative Examples

The complement of the empty graph on n vertices, which contains no edges between its vertices, is the complete graph K_n, in which every pair of distinct vertices is adjacent. Conversely, the complement of the complete graph K_n is the empty graph on n vertices. A simple illustration arises with three vertices forming the complete graph K_3, known as a where each connects to the other two. The complement of this consists of the same three vertices with no edges between them, resulting in three isolated vertices. To visualize, label the vertices A, B, and C: in K_3, the edges are AB, AC, and BC; in the complement, there are no edges at all. The P_4 on four vertices, arranged linearly as A-B-C-D with edges AB, BC, and CD, provides another clear example. Its complement is isomorphic to P_4 itself, serving as a basic where the structure mirrors its own complement under vertex relabeling. Similarly, the C_5 on five vertices, forming a with edges connecting them in a closed loop (e.g., 1-2-3-4-5-1), has a complement that is also isomorphic to C_5. This highlights how certain odd-length cycles exhibit self-complementarity. These examples demonstrate that complements of regular graphs are also regular, as seen in the consistent degrees preserved in structures like C_5, which is 2-regular and whose complement maintains the same degree sequence.

Theoretical Applications

Complement graphs provide a powerful duality in , particularly in relating independent sets and . An independent set in a graph G corresponds exactly to a in its complement \overline{G}, and vice versa, because the absence of edges within the set in G implies the presence of all possible edges in \overline{G}. This duality implies that the size of the maximum independent set \alpha(G) in G equals the number \omega(\overline{G}) of the complement. Such relationships facilitate the transfer of algorithmic and structural insights between the two problems, enabling researchers to solve one by transforming it into the other. In , complements reveal structural equivalences between classes. For instance, the complement of a (one without K_3 subgraphs) is claw-free, meaning it contains no induced K_{1,3} (a star with three leaves). This property arises because any potential in the complement would correspond to a in the original , which is forbidden. Claw-free graphs, including these complements, have been extensively studied for their decomposition properties and algorithmic tractability in optimization problems. Complement graphs play a key role in by allowing analysis of edge distributions to bound Ramsey numbers. Ramsey numbers R(s,t) denote the smallest order of a where any 2-coloring of the edges forces either a of size s or an independent set of size t; considering the complement shifts the perspective from cliques in one coloring to independent sets in the other, aiding in lower bound constructions through self-complementary graphs or balanced incomplete block designs. For example, self-complementary graphs like the 5-cycle provide tight examples for R(3,3)=6, as neither the graph nor its complement contains a . The concept of complementation is integral to the characterization of perfect graphs. According to the Strong Perfect Graph Theorem, a graph is perfect if and only if neither it nor its complement contains an induced odd cycle of length at least five (an odd hole or odd antihole). This theorem, proved by Chudnovsky, Robertson, Seymour, and , underscores how complements help identify forbidden subgraphs that disrupt perfectness, linking chromatic and clique numbers across induced subgraphs. In network analysis, complement graphs model non-interactions or absences of relationships, offering insights into social and biological systems. In social networks, where edges represent friendships or collaborations, the complement captures enmities or non-connections, facilitating community detection by partitioning based on dense non-edge clusters. Similarly, in biological networks like protein-protein interactions, complements highlight potential inhibitory or non-binding pairs, aiding in the inference of regulatory mechanisms and pathway disruptions. This approach enhances modularity optimization and anomaly detection in sparse interaction data.

Self-Complementary Graphs

Definition and Characterization

A self-complementary graph is an undirected simple G = (V, E) that is to its complement \overline{G} = (V, \overline{E}), where \overline{E} consists of all possible edges between vertices in V that are absent from E. This isomorphism implies that G and \overline{G} share identical structural properties, such as the same degree sequence and characteristics. For a graph on n vertices to be self-complementary, n must satisfy n \equiv 0 or $1 \pmod{4}, as this ensures the total number of edges in the complete graph K_n, which is \binom{n}{2}, can be evenly divided between G and \overline{G}. Consequently, a self-complementary graph must possess exactly m = \frac{\binom{n}{2}}{2} = \frac{n(n-1)}{4} edges, which is an integer under the above congruence condition. A basic characterization of self-complementarity is the existence of a \phi: V \to V such that for any distinct vertices u, v \in V, the edge \{u, v\} is present in E(G) \{\phi(u), \phi(v)\} is absent from E(G). This \phi effectively maps edges to non-edges and vice versa, preserving the graph's adjacency structure in the complement. Self-complementary graphs were first systematically studied in the early , independently by Horst Sachs and Gerhard Ringel. Paley graphs, originally introduced by Raymond Paley in the context of quadratic residue tournaments, were recognized as prominent examples of self-complementary graphs due to their symmetric construction over finite fields. No self-complementary graphs exist on 2 or 3 vertices, as these orders violate the necessary congruence condition on n. On 1 vertex, there is exactly one trivial self-complementary graph, the empty graph K_1.

Constructions and Enumerations

One standard method to construct self-complementary graphs relies on a fixed-point-free on the set, which pairs vertices such that the involution serves as an from the to its complement. For a graph with $4k vertices, partition the vertices into $2k pairs via this involution \sigma, where \sigma^2 is the and no vertex is fixed. Edges are then defined by selecting, for each pair \{u, \sigma(u)\}, whether to include the edge u\sigma(u) (in which case the complement will exclude it, and vice versa), and for cross-pairs \{u, v\} where v \neq \sigma(u), including uv implies excluding \sigma(u)\sigma(v), ensuring the mapping preserves the structure. This approach, introduced by Sachs, guarantees self-complementarity as \sigma maps edges to non-edges and vice versa. For graphs of order q \equiv 1 \pmod{4} where q is a , the provides an explicit construction: take vertices as elements of the \mathbb{F}_q, and connect two distinct vertices x and y if x - y is a nonzero in \mathbb{F}_q. This graph is self-complementary because multiplication by a non-residue maps residues to non-residues, inducing an to the complement. Paley graphs are vertex-transitive and strongly regular, with the smallest example being the C_5 on 5 vertices. Recursive constructions build larger s from smaller ones, such as by addition or cone-like operations. One method starts with a H on $4k and adds a new connected to exactly $2k of H in a way that preserves the to the complement, often by selecting based on the labeling from the . Another approach forms the over a by adding an adjacent to all others, but adjusts connections to maintain self-complementarity, typically requiring H to satisfy specific conditions. These techniques yield infinite families, including those with prescribed diameters. The number of self-complementary graphs on n vertices is zero for n \equiv 2,3 \pmod{4}, as the edge count must be n(n-1)/4, which is non-integer otherwise. For small n, there is 1 on 4 vertices (the P_4), 2 on 5 vertices (the C_5 and the bull graph, which consists of a triangle with two pendant edges), and 10 on 8 vertices. The sequence grows rapidly: 36 on 9 vertices and 720 on 12 vertices, as catalogued by exhaustive enumeration. The full counts form OEIS sequence A000171. The bull graph on 5 vertices exemplifies a non-cycle self-complementary graph, featuring a adjacent to a of 2. The 10 self-complementary graphs on 8 vertices are all connected; these were fully enumerated and classified up to .

Advanced Properties

Structural Properties

The number ω(G) of a G, which is the size of a largest in G, equals the independence number α(\bar{G}) of its complement \bar{G}. This relationship underscores the duality between maximal complete subgraphs in G and maximal independent sets in \bar{G}, a fundamental symmetry in graph complementation. Regarding connectivity, if G is disconnected with at least two vertices, then \bar{G} is connected and has diameter at most 2. Vertices in different components of G are adjacent in \bar{G}, ensuring paths of length 1 between them; for vertices in the same component that are not adjacent in \bar{G} (i.e., adjacent in G), they share a common neighbor in \bar{G} from another component. Certain hereditary graph classes are closed under complementation, preserving their structural properties when taking complements. Cographs, defined recursively as the smallest class containing the single-vertex graph and closed under disjoint union and complementation, remain cographs under complementation by construction. Split graphs, whose vertices partition into a clique and an independent set, are also closed under complementation: the complement swaps the roles of the clique and independent set, yielding another split graph with the same partition sizes. Similarly, threshold graphs—a subclass of split graphs characterized by degree sequences realizable via a threshold function—are closed under complementation, as the complement's degree sequence satisfies the threshold condition equivalently. For self-complementary graphs (isomorphic to their complements), those that are or graphs exhibit balanced partitions. In a self-complementary graph on n vertices, the and independent set sizes differ by at most 1, partitioning the vertices evenly to ensure the isomorphism maps the to the independent set and vice versa. The same holds for self-complementary graphs, inheriting this property from their structure. The class of perfect graphs is closed under complementation, as established by Lovász's perfect graph theorem: a G is perfect if and only if its complement \bar{G} is perfect. This result, proved in 1972, implies that for perfect G, both G and \bar{G} satisfy the property that the chromatic number equals the number for every .

Spectral and Chromatic Properties

The chromatic number of a graph G with n vertices and its complement \overline{G} satisfies the Nordhaus-Gaddum inequalities: \chi(G) + \chi(\overline{G}) \leq n + 1 and \chi(G) \cdot \chi(\overline{G}) \leq \left\lfloor \frac{(n+1)^2}{4} \right\rfloor, with the upper bound on the sum being tight for complete and empty graphs. For connected graphs on n \geq 2 vertices, a lower bound holds: \chi(G) + \chi(\overline{G}) \geq n - 1. Equality in the upper bound occurs when one graph is complete and the other is edgeless, while recent analyses confirm the bounds remain sharp across various graph classes, including those with specified connectivity. Post-2000 results have refined gap-related aspects, such as the maximum difference |\chi(G) - \chi(\overline{G})|, showing it can reach \Omega(\sqrt{n}) for certain constructions, though exact extremal values continue to be explored. Complements of bipartite graphs exhibit particularly strong chromatic properties: since bipartite graphs are perfect, their complements are also perfect by the strong perfect graph theorem, implying \chi(\overline{G}) = \omega(\overline{G}). This equality holds because perfect graphs are closed under complementation, avoiding induced odd holes and odd antiholes in both G and \overline{G}. The adjacency matrix of the complement satisfies A(\overline{G}) = J_n - I_n - A(G), where J_n is the all-ones matrix and I_n is the . For eigenvectors orthogonal to the all-ones vector, the eigenvalues of \overline{G} are -1 - \mu, where \mu are the corresponding eigenvalues of G. In the context of Seidel switching, which generalizes complementation, the Seidel matrix S(G) = J_n - I_n - 2A(G) has eigenvalues that are negated for the complement: the spectrum of S(\overline{G}) is the negative of that of S(G). This relation underpins isospectral properties under switching operations. The independence polynomial of G, defined as I(G; x) = \sum_{k=0}^{\alpha(G)} i_k x^k where i_k is the number of independent sets of size k, equals the clique polynomial of \overline{G}, C(\overline{G}; x) = \sum_{k=0}^{\omega(\overline{G})} c_k x^k with c_k the number of cliques of size k. This duality arises because independent sets in G correspond precisely to cliques in \overline{G}.

Algorithmic Aspects

Computing Complements

To explicitly construct the complement graph from an adjacency list representation of the original graph G with n vertices, one iterates over all pairs of distinct vertices and adds an edge between those that are not adjacent in G. This process examines all \binom{n}{2} possible edges, resulting in O(n^2) time complexity. For sparse graphs where the complement is dense, an explicit for the complement would require storing nearly \binom{n}{2} edges, which is inefficient. Instead, an implicit representation stores only the adjacency information of G and determines non-adjacency in the complement via queries, avoiding the need to materialize the full edge set of size \binom{n}{2} - m, where m is the edge count in G. This approach is particularly useful when operations on the complement can be performed on-the-fly without full expansion. The \overline{A} of the complement graph can be computed directly from the adjacency matrix A of G using the formula \overline{A} = J - I - A, where J is the n \times n all-ones matrix and I is the n \times n . This matrix operation requires O(n^2) time and space, providing a straightforward way to obtain the complement's . Software libraries implement these methods efficiently for practical use. In NetworkX, the complement function generates the from an input , handling both undirected and directed cases while avoiding self-loops and parallel edges; it is optimized for graphs up to moderate sizes but may require O(n^2) time for dense outputs. Similarly, igraph provides a complementer function that constructs the by including all missing edges from the original , suitable for dense complements in analytical workflows. For large graphs, especially sparse ones where the complement's density poses storage challenges, bitsets can represent the compactly using approximately n^2 / 8 bytes, enabling efficient bit operations for queries and updates. Compressed formats, such as those in systems like Zuckerli, further reduce space for real-world large-scale s by exploiting structural redundancies in edge patterns.

Recognition and Complexity

Determining whether a given G is isomorphic to the complement \overline{H} of another given H is complete, as the complement can be constructed in polynomial time, reducing the problem to standard . (Note: While is referenced here for the general GI-completeness, the reduction is standard in .) Recognition of self-complementary graphs, which requires testing whether G \cong \overline{G}, is similarly GI-complete. A naive approach involves enumerating all permutations to find an \pi such that uv is an edge in G if and only if \pi(u)\pi(v) is a non-edge in G, yielding O(n!) . Improved algorithms prune the search space using degree sequences and structural invariants, achieving practical performance for moderate n. Tools like nauty and Traces provide efficient implementations for testing, including self-complementary cases. Many properties exhibit dual complexity when transferred to the complement. For instance, finding a maximum independent set in G is equivalent to finding a maximum in \overline{G}, inheriting from the . This duality implies that NP-hard optimization problems on G, such as , correspond to analogous hardness in \overline{G}. In 1998, and Yokoyama developed linear-time algorithms for search (including DFS-like traversals) and determination on complement graphs using implicit representations, where neighbor queries check non-adjacency in the original without constructing \overline{G} explicitly. This avoids O(n^2) space overhead and supports efficient traversal in O(n^2) time overall. In the 2020s, variational quantum algorithms have emerged for , offering potential speedups for complement-related recognition via hybrid quantum-classical frameworks, though practical quantum advantage remains unproven as of 2025.

References

  1. [1]
    Graph Complement -- from Wolfram MathWorld
    The graph complement of G, denoted as G', has the same vertex set but its edges are the edges not present in G.
  2. [2]
    [PDF] Graph Theory
    Given a graph G, the complement of G, denoted by G, is the graph whose vertex set is the same as that of G, and whose edge set consists of all the edges that ...
  3. [3]
    [PDF] CS 137 - Graph Theory - Lecture 1 February 11, 2012
    Feb 11, 2012 · A graph G is self-complementary if G is isomorphic to G. ≅. ≅. ≅. Question: How many self-complementary graphs on n vertices ? . . . for n ...
  4. [4]
    Self-Complementary Graph -- from Wolfram MathWorld
    A self-complementary graph is a graph which is isomorphic to its graph complement. The numbers of simple self-complementary graphs on n=1 ...Missing: existence | Show results with:existence
  5. [5]
    [PDF] Lecture 23: Cliques and independent sets - Faculty Web Pages
    Nov 5, 2024 · As a result, cliques in a graph G correspond exactly to independent sets in the complement graph G (where adjacent vertices become non-adjacent, ...
  6. [6]
    Graph Theory - Complement Graphs - Tutorials Point
    Applications of Complement Graphs · Finding Independent Sets: In the complement graph, a clique corresponds to an independent set in the original graph.
  7. [7]
    [PDF] GRAPH THEORY WITH APPLICATIONS
    (a) The complement GC of a simple graph G is the simple graph with vertex ... Weighted graphs occur frequently in applications of graph theory. In the ...
  8. [8]
    14. Some Graph Theory - MIT Mathematics
    A graph is c nice if its complement is nice, which means that its independence number and the smallest number of cliques that its vertices can be partitioned ...
  9. [9]
    February 22
    Feb 22, 2009 · A graph is regular of degree k iff A\j=k\j, i.e. \j is a k-eigenvector. The adjacency matrix of the complementary graph G@ is J-I-A (where J is ...
  10. [10]
    [PDF] Matrices and Graphs - Tilburg University Research Portal
    1. If AG is the adjacency matrix of a simple graph G, then AG = J − I − AG is the adjacency matrix of the complement of G. 2.
  11. [11]
    [PDF] 11. Graph Theory
    For directed graphs, we put "directed" in front of all the terms defined ... The Complement of G is defined as follows: G = (V, E′ − E). Page 11. -103 ...
  12. [12]
    [PDF] Introduction
    Example 1.1.11. The empty graph and complete graph are complements of each other; Kc n = En. The path graph P4 and its complement are shown below: P4. Pc. 4. It ...
  13. [13]
    Path Complement Graph -- from Wolfram MathWorld
    The n-path complement graph P^__n is the graph complement of the path graph P_n. The first few are illustrated above.
  14. [14]
    [PDF] Problem Set 4
    6. A graph is self-complementary if it is isomorphic to its complement. (i.e., G ! G ) For example the path P4 on 4 vertices and the cycle C5 on five vertices ...Missing: empty | Show results with:empty<|control11|><|separator|>
  15. [15]
    Claw-Free Graph -- from Wolfram MathWorld
    The line graph of any graph is claw-free, as is the complement of any triangle-free graph. Classes of graphs that are claw-free include. 1. antiprism graphs,. 2 ...
  16. [16]
    [PDF] The structure of claw-free graphs - Math (Princeton)
    A graph is claw-free if no vertex has three pairwise nonadjacent neighbors. Every connected claw-free graph can be obtained from basic types by expansion.
  17. [17]
  18. [18]
    [PDF] The strong perfect graph theorem - Annals of Mathematics
    A graph G is perfect if for every induced subgraph H, the chromatic number of H equals the size of the largest complete subgraph of H, and G is.
  19. [19]
    [PDF] Social Network Analysis and Community Detection by Decomposing ...
    In social network analysis (SNA), relationships between members of a ... complement graph of G and that there are non-negative variables ¯yij ≥ 0 ...
  20. [20]
    Biological Network Analysis: Trends, Approaches, Graph Theory ...
    13 (Complement graph). A graph G¯ is complement of another graph G if ... Social Network Analysis and Mining 9 (1) (2019) 13. 28. Nam P. Nguyen, Thang ...
  21. [21]
    [PDF] Self-complementary graphs and generalisations
    A graph which is isomorphic to its complement is said to be a self-comple- mentary graph, or sc-graph for short. These graphs have a high degree of.
  22. [22]
    [PDF] Some remarks on the theory of graphs
    The complementary graph G' of G has the same vertices as G and two points are connected in G' if and only if they are not connected in G. A special case of a ...
  23. [23]
    Paley Graph -- from Wolfram MathWorld
    Paley graphs are self-complementary, strongly regular, conference graphs, and Hamiltonian. All Paley graph are conference graphs (Godsil and Royle 2001, p.
  24. [24]
  25. [25]
    Independence Number -- from Wolfram MathWorld
    The independence number of a graph G is equal to the clique number of the complement graph,. alpha(G)=omega(G^_). (2). For a connected regular graph G on n>1 ...Missing: diestel | Show results with:diestel
  26. [26]
    Cograph -- from Wolfram MathWorld
    A cograph (or "complement-reducible graph") is simple graph defined by the criteria 1. K_1 is a cograph, 2. If X is a cograph, then so is its graph complement.
  27. [27]
    Split Graph -- from Wolfram MathWorld
    A split graph is a graph whose vertices can be partitioned into a clique and an independent vertex set. Equivalently, it is a chordal graph whose graph ...
  28. [28]
    A note on chromatic properties of threshold graphs - ScienceDirect
    -cliqueshold if its complement is (totally) p -threshold. An interesting fact about threshold graphs is that they are closed under taking complements. This ...
  29. [29]
    Perfect Graph Theorem -- from Wolfram MathWorld
    The graph complement of a perfect graph is itself perfect. Originally known as the weak perfect graph conjecture (Fulkerson 1971), the result was ...
  30. [30]
    Inequalities for the chromatic numbers of graphs - ScienceDirect.com
    Inequalities for the chromatic numbers of graphs are proved, some of which generalize results of Nordhaus and Gaddum [9] and Dirac [4]. Previous article in ...
  31. [31]
    [PDF] Spectra of graphs - CWI
    ≤ µn are the Laplace eigenvalues of a simple graph Γ, then 0 ≤ n − µn ≤ ... ≤ n − µ2 are the Laplace eigenvalues of the complement of Γ (see. §1.3.2) ...
  32. [32]
    [PDF] Seidel Switching and Graph Energy
    Mar 23, 2012 · The energy of a graph Γ is the sum of the absolute values of the eigenvalues of the adjacency matrix of Γ. Seidel switching is an operation ...
  33. [33]
    What is the time to create a complement of a graph?
    Apr 6, 2014 · It seems to me that the running time to make a complement of a graph with n nodes is n!. Is there any way to make this running time polynomial?
  34. [34]
    [PDF] Some simple graph spectra
    If A is the adjacency matrix of a graph Γ, then J − I − A is the adjacency matrix of the complementary graph Γ. ... The complete a-partite graph Ka×m, complement ...
  35. [35]
    complement — NetworkX 3.5 documentation
    Returns the graph complement of G. G graph GC A new graph. Notes Note that complement does not create self-loops and also does not produce parallel edges for ...
  36. [36]
    Complement — igraph 1.0.0 documentation
    This example shows how to generate the complement graph of a graph (sometimes known as the anti-graph) using igraph. GraphBase. complementer() .Missing: function | Show results with:function
  37. [37]
    [PDF] Zuckerli: A New Compressed Representation for Graphs - arXiv
    Sep 2, 2020 · Zuckerli is a scalable compression system meant for large real-world graphs. Graphs are notoriously challenging structures to store efficiently ...
  38. [38]
    Graph Isomorphism for $$(H_1,H_2)
    Aug 5, 2020 · We will illustrate these similarities by showing that our construction demonstrating that Graph Isomorphism is GI-complete for \(({{\mathrm{gem}}} ...<|separator|>
  39. [39]
    Graph isomorphism and self-complementary graphs
    Graph isomorphism and self-complementary ... It is known that GI problem is GI complete even for some special graph classes including regular graphs, bipartite ..
  40. [40]
    Clique versus independent set - ScienceDirect.com
    Communication complexity and the Clique–Stable Set separation. A clique is a complete induced subgraph and a stable set is an induced subgraph with no edge.
  41. [41]
    Complement graph - Wikipedia
    In the mathematical field of graph theory, the complement or inverse of a graph G is a graph H on the same vertices such that two distinct vertices of H are ...Definition · Applications and examples · Self-complementary graphs...