On the Number of Maximal Cliques in
Two-Dimensional Random Geometric Graphs:
Euclidean and Hyperbolic
Graduate School of Information Science and Technology
The University of Tokyo
Tokyo 113-8656, Japan [email protected]
Abstract: Maximal clique enumeration appears in various real-world networks, such as social networks and protein-protein interaction networks for different applications. For general graph inputs, the number of maximal cliques can be up to . However, many previous works suggest that the number is much smaller than that on real-world networks, and polynomial-delay algorithms enable us to enumerate them in a realistic-time span. To bridge the gap between the worst case and practice, we consider the number of maximal cliques in two popular models of real-world networks: Euclidean random geometric graphs and hyperbolic random graphs. We show that the number of maximal cliques on Euclidean random geometric graphs is lower and upper bounded by and with high probability for any . For a hyperbolic random graph, we give the bounds of and where is the power-law degree exponent between 2 and 3.
Keywords: Maximal Cliques, Random Geometric Graphs, Real-World Networks
1 Introduction
1.1 Background
Detecting all maximal cliques in a graph is a crucial analysis tool for real-world networks from various fields: social networks, protein-protein interaction networks, and web graphs because cliques correspond to meaningful components in the networks [22, 21, 3]. Not only does it have many direct applications, but its algorithms and techniques are used in other clique-related methods such as clique percolation [19] and -clique counting [17]. This is because we can detect all cliques by enumerating only the maximal ones.
For general graph inputs, the number of maximal cliques can be up to [20]. Therefore, enumerating all of them is NP-hard. However, many studies report that in real-world networks, is much smaller than that. Thus, polynomial-delay algorithms, the running time of which is bounded by , can enumerate all maximal cliques in realistic-time span even for networks with millions of vertices [8]. Also, classic Bron-Kerbosch algorithm [5] (plus graph orientation [7]) is known to be efficient in many instances [9], although its worst running time is and not bounded in terms of . Here we strike upon the question: why is the number of maximal cliques small on real-world networks?
In the study of real-world networks, networks that appear naturally in various fields are considered. In terms of the graph structure, it seems that networks from different domains are entirely different from each other. However, it is known that they share specific common properties. For example, they have a power-law degree distribution: the number of nodes with a vertex degree of is proportional to . In many cases, is between two and three, and these networks are called scale-free. Additionally, they have the triadic closure property, meaning that if two vertices have common neighbors, they are likely to be connected. The property is often described with a measure called the clustering coefficient, and real-world graphs often have a high clustering coefficient. Other common properties include tree-like structures, small diameter, and small clique number.
One of the most combinatorially studied models of real-world networks is hyperbolic random graphs [18]. The graph is generated by independently placing vertices according to a particular distribution in a two-dimensional space with negative curvature and connecting two vertices within a certain distance. It is thus classified as a random geometric graph. It is known that a hyperbolic plane naturally induces power-law degree distribution [18]. Also, a hyperbolic random graph satisfies a high clustering coefficient with high probability [14], which distinguishes itself from other power-law models such as Barabási-Albert [2] and Chung-Lu random graphs. Parameters studied in this model include: the number of -cliques, clique number [12], treewidth [4], modularity [6], and diameter [13].
1.2 Our Main Results and Contributions
In this paper, we consider the number of maximal cliques in hyperbolic random graphs. We also consider the number on two-dimensional Euclidean random geometric graphs, which are the Euclidean counterpart to hyperbolic random graphs. Euclidean random geometric graphs are also thought to be good representations of some types of real-world graphs [15, 16], although they do not possess power-law degree distribution. Our findings are as follows:
Theorem 1.1 (Main 1)
Let be a constant. Let be the number of maximal cliques in a two-dimensional Euclidean random geometric graph whose connection distance is . There exists positive constants and such that for all ,
as .
Theorem 1.2 (Main 2)
Let . Let be the number of maximal cliques in a hyperbolic random graph whose power-law degree exponent is . There exist positive constants and such that for all ,
as .
For hyperbolic random graphs, we consider the case when for hyperbolic random graphs. In this case, the graphs are scale-free and have cliques with high probability [12]. In general graphs, the number of maximal cWe and the bound we obtained is much smaller than that.
To prove the main theorem, we consider what is called an octahedral graph . The definition of the graph is the following. Let where and . Therefore, is a graph with pairwise disjoint edges. An octahedral graph is the complement of . We have the following theorems from the previous study.
Theorem 1.3 (Forklore)
If the graph has as a vertex-induced subgraph, then the number of maximal cliques is lower-bounded by .
Theorem 1.4 (M. Farber, M. Hujter, and Z. Tuza [10])
If and there exists no as a vertex-induced subgraph, then the number of maximal cliques is upper-bounded by .
Let be the maximum such that a graph has as its vertex-induced subgraph. With these theorems, all that remains is to bound . If is a constant, then the number of maximal cliques is polynomial. Unfortunately, this is not the case for the two random geometric graphs. However, our bounds on are much smaller than the obvious bound.
Intuitively, is small because any pair of unconnected vertices have common neighbors, which is against the triadic closure property, i.e. two vertices with common neighbors are likely to be connected. As gets larger, the more severe the violation becomes. This explanation can be mathematically justified on Euclidean and hyperbolic random geometric graphs. The arguments on the two different random graphs are basically the same even though the definitions of distance are different, and the parallel postulate does not hold on a hyperbolic plane.
Our contributions are briefly summarized as follows. Firstly, to the best of our knowledge, this is the first work that assesses the number of maximal cliques on random geometric graphs. We shed light on the importance of and develop geometric and probabilistic techniques to determine its size. Those techniques apply to both Euclidean and hyperbolic planes. Secondly, what we have found is yet another result followed by -closed graphs [11] supporting that the triadic closure property plays an essential role in maximal cliques. Lastly, we give an upper bound of on real-world graph datasets. It turns out that is often at most 5-20, even on networks with hundreds of thousands of vertices (See Table 1 at the end of this paper). The upper bound of given by is still far from the actual value. However, it is still surprising how small is in practice.
1.3 Proof Sketch
Here we explain how we bound . For the lower bound, we construct regions so that vertices on them would form . Here we use a technique called Poissonization introduced in [23].
To derive the upper bound, We define a set of segments the distance between and is greater than the connection distance. Since is somewhat like the complement of the graph, a vertex-induced subgraph that is isomorphic to corresponds to segments that are pairwise disjoint (or we call “independent” in this paper). We soon find out that any two independent segments must intersect. Therefore, we can define an “angle” between them. By the pigeonhole principle, if is large enough, then there exists a set of segments with decent cardinality whose “angles” are small (Lemma 2.10 and Lemma 3.14). We also prove that if two intersecting segments have small “angles”, the geometry is very restricted. If we fix one segment, then endpoints of the other segments can only exist in small regions (Corollary 2.8 and Corollary 3.12). By combining the two facts, we derive that cannot get too large because the number of vertices in those small regions cannot be too large with high probability.
Unlike Euclidean random geometric graphs, vertices in hyperbolic random graphs are placed by non-uniform distribution. To deal with this, we classify the segments by the density around their endpoints. We treat the segments in dense regions slightly differently from those in sparse regions. However, the overall strategies of the proof are the same as Euclidean random geometric graphs.


1.4 Organization of the Paper
In Section 2, we define Euclidean random geometric graphs and prove the main theorem. In Section 3.1-3.3, we define hyperbolic random graphs and discuss how the proof of the random geometric graphs on a Euclidean plane can be extended to a hyperbolic plane. For the upper bound, we prove the weaker version of the main theorem. In Section 3.4, we discuss how to treat the non-uniformity of hyperbolic random graphs and prove the complete version of the main theorem. Finally, Section 4 is the conclusion.
2 Euclidean Random Geometric Graphs
2.1 Definitions
Let , and . A two-dimensional Euclidean random geometric graph is obtained as below:
-
•
The vertex set is .
-
•
The vertices are identically and independently distributed on according to a probability density function .
-
•
The edge set is given by
Here, where and are the -coordinates of and respectively. From here, we often identify a vertex with its position.
Given a region on (where we can perform integration), define . is equal to the probability that a vertex lies on . Note that we can assume no three vertices lie on a single line since the probability is 0. This avoids some corner case arguments.
2.1.1 Poissonization
This is introduced in [23]. Let be a Poisson random variable with mean . Consider where the number of vertices is chosen by the Poisson distribution . Then, on such a random graph, the number of vertices on a region follows a Poisson distribution with mean . In particular, the probability that has no vertex on itself is . Moreover, for disjoint regions and , the number of vertices on each region is an independent Poisson distribution with mean and , respectively. Therefore, on Poissonized random geometric graphs, arguments of multiple disjoint regions are easier. Let be some property of a graph. Then
Here, by the Stirling’s approximation. Thus, we get
(2.1) |
Although our results are on the normal random geometric graph , we use this technique to derive the lower bound of the number of maximal cliques.
2.2 Lower Bound of the Number of Maximal Cliques
Let be an integer. Let . Consider taking a polar coordinate system whose origin is . Let . For , Define where and . With some calculations, we can confirm the following.

Proposition 2.1
Let and . For a vertex on and a vertex on ,
Proof: Let and be radial coordinates of and , respectively. If , then
Otherwise,
For , let be 0-1 random variables which are equal to 1 if and only if both and have at least one vertex on themselves. Let . Then, there exists as a vertex-induced subgraph. We are left to lower bound .
Proposition 2.2
as
Proof:
On the third line, we used the Taylor expansion.
Proposition 2.3
If , then there exists a positive constant such that as
Proof: Consider the Poissonization . For each , the number of vertices on follows an independent and identical Poisson distribution with mean . Therefore, are independent and identical random variables. Let . This is equal to the probability that has at least one vertex on itself. By the Proposition 2.2, . Also, we have . Since are independent, we can apply the Chernoff bound and obtain . Now we are back to the normal model . Using (2.1), it follows that . This goes to 0 as .
By combining the proposition with the Theorem 1.3, we obtain the main theorem regarding .
Corollary 2.4
There exists a positive constant such that as .
2.3 Upper Bound of the Number of Maximal Cliques
Let denote the segment between vertices and on a Euclidean plane. Let . Note that is like a complement of the random graph. Two segments and in are called independent if , , , and are not in . Two or more segments in are called independent if any pair of the segments are independent. From the definition, it is obvious that the following proposition holds.
Proposition 2.5
There is no as a vertex-induced subgraph There is no set of independent segments whose cardinality is .
Therefore, we are left to bound such on . With elementary geometry, we can confirm the following.
Proposition 2.6
On a Euclidean plane, two independent segments intersect.
We defer the proof to the next section. Thanks to the proposition, we can define an angle between two independent segments thanks to the proposition.
Let and be independent segments. Suppose that the counter-clockwise order of four endpoints is . Let be the intersection of two segments. Define a directed angle . For convinience, define . It holds that . Therefore, the order of the two segments matters.
From here, we consider conditions that two independent segments mWe angle between them is known. Again, be independent segments. Suppose is the counter-clockwise order of four endpoints. Let be the midpoint of . Consider taking a polar coordinate system whose origin and polar axis are and . Let be the polar coordinate of . Note that and (refer to Figure 3). The next proposition states that if the directed angle is small, then is bounded tight.
Proposition 2.7
Let . If , then where
Proof: Let . Let be the intersection of and . Let . Note that is the length of and holds. The same thing stands for . From the definition of independence of segments,
must hold. From the first inequality, we get
With the same argument, the second inequality yields . Regardless of how are lined up on , we have . We can upper and lower bound as below.

If are lined up in this order, we can bound as . By considering the other case ( is lined up in this order) in the same way, we have the following.
Corollary 2.8
Let . If , then either or must lie on a region or a region where

Then we obtain the next lemma: if the directed angle is bounded small, the expected number of independent segments is also small.
Lemma 2.9
Let be a segment. For a constant , let be a number of segments where and are independent and . Then as .
Proof: By the Corollary 2.8, is at most the number of vertices in . Therefore, . We are left to prove as , and
To derive the fourth line, we use the Taylor expansion.
To fully use the Lemma 2.9, we prove another lemma, which states that if there exists large , we can always take a set of segments so that its size is unneglectable, and the directed angles between them are small.
Lemma 2.10
Let be a positive integer. If there exists a set of independent segments , then there exists a segment and a set of independent segments such that
(2.5) |
Proof:
Let be some segment in . For an integer , define
.
Since , we have
. By the pigeonhole principle, there exists an index
such that . Consider taking the segment in so that
.
If we can prove , we are done.
Let be the intersection of and , be that of and ,
and be that of and . If , then it obviously holds.
We have two other cases: are lined up on in this order, and are lined up on in this order.
For the former case, , and
where , , are the inner angle of . Therefore,
(2.6) | |||
Here, (2.6) follows from the fact that the sum of the inner angles of a triangle is equal to .
The latter case can be shown similarly.
We finally combine the two lemmas to upper bound . After that, we apply Theorem 1.4 to get the upper bound of the number of maximal cliques.
Theorem 2.11
Let be a positive constant. Let . Then,
as .
Proof: Let and . Suppose has as its vertex-induced subgraph. Apply Lemma 2.10 for and . Then there exists a segment and a set of independent segments such that (2.5) holds. We claim that for a certain segment , the probability that there exists a set of independent segments such that (2.5) holds is at most . By Lemma 2.9, the number of segments that satisfy the third condition of (2.5) is at most the number of vertices on , and that expected number is bounded by . However, to satisfy , The cardinarity of needs be at least . This means that the number of vertices on also needs to be at least . Since the vertices are placed independently, we can apply the Chernoff bound to get the probability bound .
Finally, we apply a union bound over all segments of which there are at most . We have
This goes to 0 as .
Corollary 2.12
There exist a positive constant such that for all , as
3 Hyperbolic Random Graphs
3.1 Definitions
Given , , and , let , and . A hyperbolic random graph is obtained as below:
-
•
The vertex set is .
-
•
The vertices are identically and independently distributed on a hyperbolic plane. The probability density function by a polar coordinate is:
-
•
The edge set is given by

In our work, we are interested in the case where the generated graph is “scale-free” with high probability. Let and be polar coordinates of and respectively. Let . Then, the hyperbolic cosine formula suggests that
Again, define . On a hyperbolic plane, the area of is equal to . Let . Since the value of does not depend on , we sometimes denote it as . Intuitively, is the the probability that a vertex lies on a single unit square around . By differentiating , we can confirm that it is monotonically decreasing with respect to . Thus, the graph is dense near the origin of the plane.
Let be the origin of the plane. We define as the set of points such that . In other words, .
3.2 Lower Bound of the Number of Maximal Cliques
To obtain the lower bound of , we do the same thing as the Euclidean random geometric graphs i.e. We take regions so that vertices on them form .
Let be an integer. Let . For , Define where and are determined as
We now assume that . We later confirm this holds under proper and .
Proposition 3.1
Let and . For a vertex on and a vertex on ,
Proof: Let and be radial coordinates of and , respectively. If , then
Otherwise,
For , let be 0-1 random variables which are equal to 1 if and only if both and have at least one vertex on themselves. Let . Then, there exists as a vertex-induced subgraph. Again, we are left to lower bound .
Proposition 3.2
If , then as and
Proof: We first claim that . It is sufficient to prove , and this follows since and . With the same argument, we have . Then, can be calculated as below.
On the third line, we use the fact that is a monotonically decreasing function. Here,
The last equality holds since .
As a byproduct of the Proposition 3.2, we obtain the condition which holds.
Corollary 3.3
If , then with sufficiently large and sufficiently small , .
Proof: Recall that and . Therefore, we have with sufficiently large . Also, recall that as and , proving with sufficiently large and sufficiently small .
Proposition 3.4
If , then there exists a positive constant such that as
Proof: As in the Proposition 2.3, consider the Poissonization . The proof is the same. We only need to comfirm that for sufficiently large and .
Again, by combining the proposition with the Theorem 1.3, we obtain the main theorem regarding .
Corollary 3.5
There exists a positive constant such that as .
3.3 Upper Bound of the Number of Maximal Cliques
To upper bound , we prove the hyperbolic versions of the Lemma 2.9 and 2.10. Again, we start with defining segments. Let denote the segment (i.e., the shortest geodesic) between vertices and on a hyperbolic plane. Let . We define the independence of segments in the same way as for the Euclidean case. We first confirm that two independent segments intersect. To do so, we use the following lemma from [12].
Lemma 3.6 ([12])
Let and be points on a hyperbolic plane where . Let be a line that passes through and . Let and be points on the same halfplane induced by . Suppose , , , and . Then, .
Also, we use the following basic facts of non-Euclidean geometry.
Proposition 3.7 (Euclid’s Proposition I-13)
If a line stood on another line makes angles, it will make two right angles, or the sum of the angles is equal to two right angles.
Proposition 3.8 (Euclid’s Proposition I-19)
In any given triangle, the side for a larger angle is larger.
Proposition 3.9 (Saccheri–Legendre Theorem [24])
The sum of all three inner angles in a triangle is less than or equal to two right angles.
Note that all of those Lemma and Propositions do not rely on the parallel postulate. Therefore, they are valid on both a Euclidean and a hyperbolic plane.
Proposition 3.10
On a hyperbolic plane, Two independent segments intersect.
Proof: Let and be two independent segments. Let be a line that passes through and . We claim that and are located on the different halfplanes induced by . We prove this by contradiction. Suppose and are on the same halfplane. Take points and on the segment so that , , and , where . If we can prove , then we also have , , and by symmetricity. By Lemma 3.5, it follows that . However, this is a contradiction since .
Now the goal is to prove Focus on . By the definition of independent segments, is shorter than . Therefore, by the Proposition 3.9, . We have . Otherwise, and , and these contradict with the Proposition 3.8. With the same argument, . By the Proposition 3.7, . Therefore, on , . Again, by the Proposition 3.8, the length of is greater than . Recall that . Thus, We can conclude that .
Notice that the proof of the Proposition 3.10 does not rely on the parallel postulate. Therefore, this proof is also valid on a Euclidean plane, proving the Proposition 2.6.
We can also define the angle of segments on a hyperbolic plane. Again, be independent segments. Suppose is the counter-clockwise order of four endpoints. Let be the intersection of the two segments. Define . Let be the midpoint of . Consider taking a polar coordinate system whose origin and polar axis are and . Let be the polar coordinate of . Also, define . The definitions are summarized in Figure 3.
Lemma 3.11
Let . For all ,
(3.1) | |||
(3.2) | |||
(3.3) |
where
Proof: By the definition of independent segments, , , , and are all less than or equal to . Therefore,
(3.4) | ||||
(3.5) | ||||
(3.6) | ||||
(3.7) |
From (3.4), we have
(3.8) |
By doing the same thing on (3.5), we obtain
(3.9) |
The last inequalities hold since . Plugging this in to (3.9) yields (3.1).
From (3.7), we have
Multiplying this with (3.9) gives us
For ,
Note that this holds no matter how and are lined up on . We have
Corollary 3.12
Let . If , then either or must lie on or where
Here,
Corollary 3.13
Let be a segment. Let . Let be a number of segments which is independent from and , then
as and .
Proof:
In this proof, we simply denote as .
It is sufficient to prove that
. Here,
Here, we can bound
The inequality follows because is a convex function. The last equality follows by the assumption . Also,
To fully use the Lemma 3.13, we prove the hyperbolic version of the Lemma 2.10. The statement is identical.
Lemma 3.14
Let be a positive integer. If there exists a set of independent segments , then there exists a segment and a set of independent segments such that
(3.13) |
Proof: The proof of Lemma 2.10 is also valid on a hyperbolic plane, except for the part (2.6) where we used the fact that the sum of inner angles of a triangle is equal to . This can be replaced with the Proposition 3.8. Then we get
(3.14) | |||
Thus, we can prove the same statement.
It is not so hard from here to prove the limited version of the main theorem. The next proposition states that if vertices of do not fall on a region near the origin, we can bound its size.
Proposition 3.15
Let . Let be a subgraph of induced by . Let where is arbitrary positive constant. Then,
as .
Proof: We will not go into detail here. However, the statement can be proved in the same way as the Theorem 2.11. We end up claiming when . Since , it follows.
3.4 Dealing With Non-Uniformity
We prove the additional facts about two independent segments. Again, we will use the definitions of points, angles, and length in Figure 3.
Corollary 3.16
For all , .
Proof: Let . Due to symmetricity, (3.3) in the Lemma 3.11 also holds for . Especially, when for some , we have
Since and are monotonic functions on , we have the bound on .
Corollary 3.17
With sufficiently large , and .
Proof: It is sufficient to prove that we have if is large enough. We omit the detail here since it is a simple calculation task.
The following proposition states that can be lower bounded in terms of .
Proposition 3.18
if , then with sufficiently large ,
Proof: If , it clearly holds. Otherwise, Let . Let be the intersection of and . Let . By the sine formula on a hyperbolic plane, it holds (no matter how and are lined up on ) that
Therefore,
For sufficiently large since and
Lemma 3.19
Let and be independent segments. Let be the distance between the midpoint of and the origin. Suppose one of ’s endpoints is in , and , then with sufficiently large , both endpoints of are outside of .
Proof: Note that we would never have the case where both endpoints of a segment in are in . We prove the proposition by contradiction. By symmetricity and the fact that , We can assume that is in without loss of generality.
Recall that . We have . Therefore, either or must be greater than or equal to . First, suppose . By the Lemma 3.18,
Therefore,
This is a contradiction since must be less than or equal to . On the other hand, suppose . Then,
Again, this is a contradiction.

Using the Lemma 3.19, we divide independent segments into some sets so that segments in each set satisfy either condition.
Lemma 3.20
Let be a positive integer. If there exists a set of independent segments , then there exists a segment , a set of independent segments such that,
(3.19) |
or
(3.24) |
for some .
Proof: Take a segment whose endpoint is in . If there is none, then we are done with the Lemma 3.14. Let . For each index , we define
Since , . By the pigeonhole principle, there exists an index such that either , , or has the cardinality greater or equal to . Let be such a set. Consider taking the segment in so that . If or , . The third condition of (3.24) is satisfied. Also, by the Corollary 3.17, endpoints of must be outside of a circle whose radius and center are and . Therefore, the fourth condition of (3.19) is satisfied. We have proved that and fulfill (3.24). If , then we must have so that is not empty. Therefore, . By Lemma 3.19, both endpoints of must be outside of , and (3.19) holds.
Theorem 3.21
Let be a positive constant. Let . Then,
as .
Proof: Let and . Suppose has as its vertex-induced subgraph. Apply Lemma 3.20 for and . As in the proof of 2.11, Then there exists a segment and a set of independent segments such that either (3.19) or (3.24) holds. We first claim that for a certain segment , the probability that there exists a set of independent segments such that (3.19) holds is at most . By the Corollary 3.13, the expected number of segments which is possibly in can be bounded by . However, the size of must be . By the Chernoff bound, we have the probability bound. We next claim that for a certain segment , the probability that there exists a set of independent segments such that (3.24) holds is at most . Again, it is sufficient to prove that the expected number of segments that satisfy the third and the fourth condition of (3.24) is . Let . Here,
Finally, we apply a union bound over all segments and obtain
which goes to 0 as .
Theorem 3.22
There exist a positive constant such that for all ,
as
4 Conclusions
In this paper, we started with the question: why is the number of maximal cliques small on real-world networks? To give an explanation to it, we consider the number of maximal cliques in two models of real-world networks: Euclidean random geometric graphs and hyperbolic random graphs. To bound the number, we focus on vertex-induced subgraphs, which are isomorphic to . Similar geometric and probabilistic techniques apply to both random graphs. For future works, we are interested in whether the property of has positive effects on clique enumeration algorithms on real-world graphs. Also, the generalization to a higher dimensional space is an open question.
References
- [1] W. Aiello, F. Chung, and L. Lu. A random graph model for massive graphs. In Proceedings of the 32nd Annual ACM Symposium on Theory of Computing, pages 171-180, 2000.
- [2] R. Albert, and A. L. Barabási, Statistical mechanics of complex networks. Reviews of Modern Physics, 74(1):47, 2002.
- [3] L. Becchetti, P. Boldi, C. Castillo, and A. Gionis. Efficient semi-streaming algorithms for local triangle counting in massive graphs. In Proceedings of the 14th ACM SIGKDD International Conference on Knowledge Discovery and Data Mining, pages 16-24, 2008
- [4] T. Bläsius, T. Friedrich, and A. Krohmer. Hyperbolic random graphs: separators and treewidth. In Proceedings of the 24th Annual European Symposium on Algorithms, pages 1-16, 2016
- [5] C. Bron, and J. Kerbosch. Algorithm 457: finding all cliques of an undirected graph. Communications of the ACM, 16(9):575-577, 1973
- [6] J. Chellig, N. Fountoulakis, and F. Skerman. On the diameter of hyperbolic random graphs. Journal of Complex Networks, 10(1):cnab051, 2022
- [7] N. Chiba, and T. Nishizeki. Arboricity and subgraph listing algorithms. SIAM Journal on Computing, 14(1):210-223, 1985
- [8] A. Conte, R. Grossi, A. Marino, and L, Versari. Sublinear-space bounded-delay enumeration for massive network analytics: Maximal cliques. In 43rd International Colloquium on Automata, Languages, and Programming (ICALP 2016), volume 55, 148:1-148:15, 2016
- [9] D. Eppstein, M. Löffler, and D. Strash. Listing all maximal cliques in large sparse real-world graphs. Journal of Experimental Algorithmics, 18:1-3, 2013
- [10] M. Farber, M. Hujter, and Z. Tuza. An upper bound on the number of cliques in a graph. Networks, 23(3):207-210, 1993
- [11] J. Fox, T. Roughgarden, C. Seshadhri, F. Wei, and N. Wein. Finding cliques in social networks: A new distribution-free model. SIAM journal on computing, 49(2):448-464, 2020
- [12] T. Friedrich, and A. Krohmer. Cliques in hyperbolic random graphs. In Proceedings of the 34th IEEE Conference on Computer Communications, pages 1544-1552, 2015
- [13] T. Friedrich, and A. Krohmer. On the diameter of hyperbolic random graphs. SIAM Journal on Discrete Mathematics, 32(2):1314-1334, 2018
- [14] L. Gugelmann, K. Panagiotou, and U. Peter. Random hyperbolic graphs: degree sequence and clustering In 39th International Colloquium on Automata, Languages, and Programming (ICALP 2012), pages 573-585, 2012
- [15] M. Haenggi, J. G. Andrews, F. Baccelli, O. Dousse, and M. Franceschetti. Stochastic geometry and random graphs for the analysis and design of wireless networks IEEE Journal on Selected Areas in Communications, 27(7):1029-1046, 2009
- [16] D. J. Higham, M. Rašajski, and N. Pržulj. Fitting a geometric graph to a protein–protein interaction network. Bioinformatics, 24(8):1093-1099, 2008
- [17] S. Jain, and C. Seshadhri. The power of pivoting for exact clique counting. In Proceedings of the 13th International Conference on Web Search and Data Mining, pages 268-276, 2020
- [18] D. Krioukov, F. Papadopoulos, M. Kitsak, A. Vahdat, and M. Boguná. Hyperbolic geometry of complex networks. Physical Review E, 82(3):036106, 2010
- [19] J. M. Kumpula, M. Kivelä, K. Kaski, and J. Saramäki. Sequential algorithm for fast clique percolation. Physical Review E, 78(2):026109, 2008
- [20] J. W. Moon, and L. Moser. On cliques in graphs. Israel Journal of Mathematics, 3(1):23-28, 1965
- [21] T. Nepusz, H. Yu, and A. Paccanaro. Detecting overlapping protein complexes in protein-protein interaction networks. Nature Methods, 9(5):471-472, 2012
- [22] G. Palla, I. Derényi, I. Farkas, and T. Vicsek. Uncovering the overlapping community structure of complex networks in nature and society. Nature, 435(7043):814-818, 2005
- [23] M. Penrose. Random geometric graphs. OUP Oxford, Volume 5, 2003
- [24] P. J. Ryan. Euclidean and non-euclidean geometry: an analytic approach. Cambridge University Press, 1986
Graph | |||
---|---|---|---|
amazon0601 | 403394 | 2443408 | 5 |
as-skitter | 1696415 | 11095298 | 15 |
ca-AstroPh | 18772 | 198050 | 7 |
com-dblp | 317080 | 1049866 | 4 |
com-youtube | 1134890 | 2987624 | 7 |
email-Enron | 36692 | 183831 | 7 |
email-EuAll | 265214 | 364481 | 8 |
facebook_combined | 4039 | 88234 | 19 |
hrg | 180705 | 968205 | 4 |
loc-gowalla_edges | 196591 | 950327 | 8 |
soc-buzznet | 101164 | 2763067 | 14 |
soc-digg | 770800 | 5907133 | 17 |
soc-Epinions1 | 75879 | 405740 | 10 |
web-Google | 875713 | 4322051 | 5 |
web-NotreDame | 325729 | 1090108 | 5 |
web-Stanford | 281903 | 1992636 | 5 |
wiki-Talk | 2394385 | 4659565 | 13 |
wiki-topcats | 1791489 | 25444207 | 10 |
wiki-Vote | 7115 | 100762 | 8 |