A geometry-induced topological phase transition in random graphs

1, 2, 1, 2, 3, 1, 2, Jasper van der Kolk, ∗ M. Angeles´ Serrano, † and Mari´anBogu˜n´a ‡ 1Departament de F´ısica de la Mat`eriaCondensada, Universitat de Barcelona, Mart´ıi Franqu`es1, E-08028 Barcelona, Spain 2Universitat de Barcelona Institute of Complex Systems (UBICS), Universitat de Barcelona, Barcelona, Spain 3Instituci´oCatalana de Recerca i Estudis Avan¸cats (ICREA), Passeig Llu´ıs Companys 23, E-08010 Barcelona, Spain Clustering – the tendency for neighbors of nodes to be connected – quantifies the coupling of a to its underlying latent metric space. In random geometric graphs, clustering undergoes a continuous phase transition, separating a phase with finite clustering from a regime where clustering vanishes in the thermodynamic limit. We prove this geometric-to-nongeometric phase transition to be topological in nature, with atypical features such as diverging free energy and entropy as well as anomalous finite size scaling behavior. Moreover, a slow decay of clustering in the nongeometric phase implies that some real networks with relatively high levels of clustering may be better described in this regime.

Network geometry [1] provides a simple and compre- on a discrete phase space defined by the N(N 1)/2 pairs hensive approach to complex networks. The existence of of distances among the N nodes of the network.− Interest- latent metric spaces underlying complex networks offers ingly, despite the fact that fermions in the model are non- a deft explanation for their intricate topologies, giving interacting particles, we show that the system undergoes at the same time important clues on their functionality. a topological phase transition at a critical temperature 1 The small-world property, high levels of clustering (or tri- Tc = βc− , separating a “geometric” phase, with a finite angles in the graph), heterogeneity in the distri- fraction of triangles attached to nodes (clustering), and bution, and hierarchical organization are all topological a nongeometric phase, where clustering vanishes in the properties observed in real networks that find a simple thermodynamic limit. This transition is not exclusive to explanation within the network geometry paradigm. In the S1/H2 model as it takes place in a very general class adition, this approach has been used to define geometric of spatial networks defined in compact homogeneous and communities [2–4], to understand the self-similar archi- isotropic Riemannian manifolds of arbitrary dimension- tecture of real networks [5–7], the navigability proper- ality [22]. ties of systems such as the Internet [8–10] and the hu- In this letter, we analyze in detail this geometric-to- man brain [11], and the temporal evolution of social net- nongeometric (GNG) phase transition which shows in- works [12]. teresting anomalous scaling behavior as compared with 1 These results are based on the S model [5] and its iso- standard continuous phase transitions, where one ob- 2 morphically equivalent formulation, the H model [13]. serves a power law decay at the critical point and a faster 1 In the S model, nodes are assumed to live in a metric decay in the disordered phase. In fact, the GNG transi- similarity space where distances can be evaluated. At the tion is reminiscent of the Berezinskii-Kosterlitz-Thouless same time, nodes are heterogeneous, with popular and topological phase transition (BKT) [23, 24] in the sense not so popular nodes coexisting within the same system. that there is no symmetry breaking at the critical point A link between a pair of nodes is created with a prob- but a transition between two phases with a different orga- ability that resembles a gravity law, increasing with the nization of topological defects, in this case given by the product of nodes’ popularities and decreasing with the cycles in the network. However, unlike the BKT tran- increase of their in the similarity space. Inter- sition, the GNG transition is driven by the divergence estingly, both similarity and popularity coordinates can of both the free energy and the entropy of the system. be combined to map nodes onto the hyperbolic plane As a result, in the geometric phase β > βc, the order so that the connection probability of the 1 model be- arXiv:2106.08030v1 [physics.soc-ph] 15 Jun 2021 S parameter measuring clustering approaches continuously comes just a function of the hyperbolic distance among the transition with very smooth behavior; at the critical nodes [13]. This allows us to say that hyperbolic geom- point, the finite size behavior of clustering is anomalous, etry is the “natural” geometry of complex networks. In- decaying logarithmically to zero for very large systems; terestingly, many analytic results have been derived for in the nongeometric phase, we discover a weakly geo- the 1/ 2 model, e. g. [5, 13, 14], S H metric region βc0 < β < βc where clustering decays as clustering [13–15], diameter [16–18], percolation [19, 20], a power of the system size with an exponent that de- self-similarity [5], or spectral properties [21]. pends on the temperature; and finally, when β < βc0 , As shown in [13, 22], the S1/H2 model is equivalent to a clustering behaves as in standard random graphs [25]. system of noninteracting fermions at temperature T . The We show that these results can be interpreted as those fermions correspond to the links of the network and live of a standard continuous phase transition with an effec- 2 tive system size Neff = ln N. This leads us to propose which have been calculated from the grand canonical par- the standard finite size scaling hypothesis but with Neff tition function and the grand potential (see SI). Notice instead of N. Beyond the obvious theoretical interest, that both the free energy and entropy per link diverge these results have practical implications since we show at the critical temperature β = βc = 1. This implies that several real complex networks with high levels of that there is a phase transition at βc that is anomalous clustering are better described using the S1/H2 model and cannot be described by Landau’s symmetry-breaking with β . βc, thus justifying studying the large tempera- theory of continuous phase transitions [26] and, thus, the ture behavior. order parameter cannot be defined as a derivative of the The S1 is a model with hidden variables, representing free energy. We conjecture that, as the underlying mech- the location of the nodes in a similarity space and their anisms leading to this anomalous behavior are quite gen- popularity within the network. Specifically, each node eral, these findings are not exclusive to the S1 model. is assigned a random angular coordinate θi distributed We argue that βc separates two distinct phases with uniformly in [0, 2π], fixing its position in a circle of radius different organization of the cycles in the network. More R = N/2π. In this way, in the limit N 1 nodes are specifically, a geometric phase with finite clustering corre-  distributed in a line according to a Poisson point process sponds to β > βc, where the triangle inequality induces a of density one with periodic boundary conditions. Each finite fraction of triangles attached to nodes, and a nonge- node is also given a hidden degree κi, which corresponds ometric phase with zero clustering in the thermodynamic to its ensemble expected degree. Each pair of nodes is limit is reached when β < βc, where links are mainly long connected with probability range and the fraction of triangles vanish. This indicates 1 that clustering is a useful order parameter to character- pij = β , (1) ize this transition, and that the emergence of clustering 1 + xij conditions the abundance of long range connections. At µκˆ iκj   low temperatures, the high energy associated to connect- where xij = R∆θij is the distance between nodes i and ing spatially distant points causes the majority of links j along the circle, and β > 1 andµ ˆ are model parame- attached to a given node to be local. The number of en- ters fixing the average clustering coefficient and average ergetically feasible links connecting very distant pairs of degree of the network, respectively. Interestingly, the nodes grows as temperature increases, diverging at the connection probability in Eq. (1) can be rewritten as the + critical point when β βc and causing the divergence Fermi distribution of the entropy, which is→ a measure of the number of effec- 1 tive link microestates. At β βc the number of available pij = β(ε µ) , (2) ≤ e ij − + 1 long range states becomes suddenly macroscopic due to where the energy of state ij is the logarithmic dependence of the energy on the distance, which causes the divergence of the entropy in this regime. xij In the high temperature regime β < βc, for a fixed εij = ln (3) 1 κiκj set of coordinates (κ , θ ), the model can be defined   i i S and the chemical potential µ = lnµ ˆ. Then, it becomes in different ways depending on the specific constraints and how the connection probability depends on distance clear that the S1 model can be thought of as a system of noninteracting identical fermions, corresponding to links, (while still being geometric and maximally random). If that can occupy the set of N(N 1)/2 states with energies we fix the sequence of expected degrees, so that the de- − gree distribution of the network remains unaltered when given by εij, which grow slowly with the distance. In this representation, parameter β plays the role of the temperature is increased beyond the critical temperature, inverse of temperature, controlling the level of noise in the connection probability is [22] should be the system, whereas the chemical potential µ fixes the 1 pij = , (6) expected number of links. xβ 1 + ij In the low temperature regime β > 1, the chemical po- µκˆ iκj β π tential is µ = ln 2π k sin β [22], with k the network β β 1 h i h i withµ ˆ (1 β)2− N − / k for β < 1 andµ ˆ = average degree. Assuming a node distribution on the cir- ≈ 1 − h i (2 k ln N)− when β = 1. Notice that the model con- cle described by a Poisson point process, the free energy vergesh i to the soft configuration model with a given ex- per link of the system reads pected degree sequence [25, 27–29] in the limit of infinite 2F β π temperature β = 0. As an alternative, one could use the = ln sin 1 (4) N k 2π k β − same connection probability in Eq. (1) and fix the value of h i  h i  µˆ taking into account the finite size of the system. In this and the entropy per link case, the degree distribution becomes progressively more 2S π homogeneous as the temperature increases, converging to = β π cot , (5) 1 N k − β the Erdos-Renyi ensemble (N, p)[30] with p N − in h i G ∼ 3 the infinite temperature limit. In both definitions, the 1 long range connections become ubiquitous, which causes 10− the entropy density to diverge and the clustering vanish in the thermodynamic limit. In the following, we use the 10 2 first approach because, by fixing the degree sequence, the − model can be directly compared with real networks. C We compute the global clustering coefficient, C, as the 3 10− local clustering coefficient averaged over all nodes in a network. The local clustering coefficient for a given node β =0.0 β =0.8 i is defined as the probability that a pair of randomly 4 10− β =0.2 β =1.0 chosen neighbors are neighbors themselves and, using re- β =0.4 β =1.2 sults from [31], it can be computed as β =0.6 β =1.5 103 104 105 j=i k=i pijpjkpik N 6 6 Ci = 2 . (7) P Pj=i pij 6 FIG. 1. Average clustering coefficient vs network size for 1 P  S In the SI we derive analytic results for the behavior of geometric networks with homogeneous degrees. The networks clustering in the neighborhood of the critical point when were generated by applying the DPG technique to a config- hidden degrees follow a power law distribution ρ(κ) uration model network with a homogeneous degree sequence γ α ∼ k = 4, k. Dashed lines are power law fits used to estimate κ− with 2 < γ < 3 and a cutoff κ < κc N , α > 1/2, ∀ ∼ 1/2 the exponent σ. growing faster than the structural cutoff κs N , which sets the onset of structural degree-degree∼ corre- lations [32]. Specifically, C behaves as is a weakly geometric region βc0 < β < βc where cluster- ing decays very slowly, with an exponent that depends 2 (β βc) β > βc − on the temperature. Finally, at β < βc0 , we recover the 2 same result of the soft configuration model for scale-free  [ln N]− β = βc  degree distributions [25]. The results in Eq. (8) around C  , (8) ∼  σ(β) the critical point suggest that Neff = ln N and not N  N − βc0 < β < βc  plays the role of the system size. Indeed, in terms of this σ(β)  N − ln N 0 β < βc0 effective size, we observe a power law decay at the critical  ≤  point and a faster decay in the unclustered phase, as ex- 2  where βc0 = γ and pected for a continuous phase transtion. Consequently, we expect the finite size scaling ansatz of standard con- 1 2(β− 1) β0 < β < β tinous phase transitions to hold with this effective size. − c c σ(β) = . (9) We then propose that, in the neighborhood of the critical ( γ 2 0 β < β0 − ≤ c point, clustering at finite size N can be written as

η 1 If α = 1/2, then the logarithm in the last line of Eq. (8) ν ν C(β, N) = [ln N]− f (β βc) [ln N] , (10) should be removed and if κc grows with N slower than − 2   any power law then βc0 = 3 and σ(β) = 1 when β < βc0 . with η = 2, ν = 1, and where f(x) is a scaling function α See the SI for the general case κc N with α < 1/2. that behaves as f(x) xη for x . ∼ ∼ → ∞ Notice that the behavior in a close neighborhood of βc is We test these results with numerical simulations, and independent of γ. by direct numerical integration of Eq. (7) using Eq. (1) Results in Eq. (8) are remarkable in many respects. for β > βc and Eq. (6) for β βc, see SI. Simula- First, clustering undergoes a continuous transition at tions are performed with the degree-preserving≤ geometric βc = 1, attaining a finite value in the geometric phase Metropolis-Hastings algorithm (DPG) introduced in [33], β > βc and becoming zero in the nongeometric phase that allows us to explore different values of β while pre- β < βc in the thermodynamic limit. Additionally, the ap- serving exactly the degree sequence. Given a network, proach to zero when β β+ is very smooth since both → c the algorithm selects at random a pair of links connect- clustering and its first derivative are continuous at the ing nodes i, j and l, m and swaps them (avoiding multiple critical point. Second, right at the critical point, cluster- links and self-connections) with a probability given by ing decays logarithmically with the system size, and it de- β cays as a power of the system size when β < βc. This is at ∆θij∆θlm odds with traditional continuous phase transitions, where pswap = min 1, , (11) " ∆θil∆θjm # one observes a power law decay at the critical point and   an even faster decay in the disordered phase. Third, there where ∆θ is the angular separation between the corre- 4

8 1.0 6 Heterogen. γ = 2.7 Heterogen. γ = 2.7 Simulations Num. Integration

N 5 ν 6 0.8 / η ]

N 4 ln 0.6 [ 4 C σ 3 η/ν = 1.90 η/ν = 1.90 ν = 0.48 ν = 0.71 0.4 2 2 10 0 10 2 0 2 − − 0.2 Simulations Heterogeneous γ = 2.3 3 γ = Homogen. Homogen. Num. Integration Heterogeneous 2.7 Simulations 4 Num. Integration Theory Homogeneous

0.0 ν / 2 0.0 0.2 0.4 0.6 0.8 1.0 η ] 3

β N ln [

C 1 2 FIG. 2. Exponent σ in the nongeometric phase evaluated η/ν = 2.00 η/ν = 2.00 ν = 0.59 ν = 0.83 from numerical simulations with the DPG algorithm (colored 1 circles), numerical integration of Eq. (7) (dashed lines), and 0 5 0 5 1 0 1 theoretical approach Eq. (8) (solid lines). Networks are gener- − − (β 1)[ln N]1/ν (β 1)[ln N]1/ν ated with a homogeneous distribution of hidden degrees (red − − lines and circles) and a power law distribution with exponents γ = 2.3 and γ = 2.7, blue and green lines and circles, respec- tively. In the heterogeneous case, we fit a function of the form FIG. 3. Finite size scaling Eq. (10) for heterogeneous net- σ N − ln N when β < 2/γ to obtain the exponent σ. works with γ = 2.7 (top row) and homogeneous networks (bottom row). Left column correspond to numerical simula- tions with sizes in the range N (5 102, 105), whereas the right column is obtained from numerical∈ × integration of Eq. (8) sponding pair of nodes. This algorithm maximizes the with sizes in the range N (5 105, 108). likelihood that the network is S1 geometric while pre- ∈ × serving the degree sequence and the set of angular coor- dinates, and it does so both above and below the critical with exponent η/ν 2 in all cases. The exponent ν, temperature. Notice that the algorithm does not depend however, departs from≈ the theoretical value ν = 1 in nu- on the hidden degrees κi, nor on the parameterµ ˆ. merical simulations due to their small sizes but improves Figure1 shows the behavior of the average clustering significantly with numerical integration for bigger sizes. coefficient as a function of the number of nodes for homo- We then expect Eq. (10) to hold, albeit for very large geneous S1 networks with different values of β, showing system sizes. σ a clear power law dependence N − in the nongeomet- The slow decay of clustering in the nongeometric phase ric phase β < βc, with an exponent that varies with implies that some real networks with significant levels of β as predicted by our analysis. These results are used clustering may be better described using the S1 model to measure the exponent σ as a function of the inverse with temperatures in the regime β < βc. Given a real temperature β, which in Fig.2 are compared with the network, the DPG algorithm can be used to find its value theoretical value given by Eq. (8). The agreement is not of β. To do so, nodes in the real network are given very good for values of β close to βc and for very het- random angular coordinates in (0, 2π). Then the DPG erogeneous networks. This discrepancy is expected due algorithm is applied, increasing progressively the value to the slow approach to the thermodynamic limit in the of β until the clustering coefficient of the randomized nongeometric phase, which suggests that the range of our network matches the one measured in the real network. numerical simulations, N [5 102, 105], is too limited. Many real networks have very high levels of clustering ∈ × To test for this possibility, we solve numerically Eq. (7) and lead to values of β > βc. However, there are no- for sizes in the range N [5 105, 108] and measure nu- table examples with values of β below the critical point. merically the exponent σ∈. In× this case, the agreement is As an example, in the SI table we show values of β ob- also very good for heterogeneous networks. The remain- tained for several real networks of interest with values ing discrepancy when β βc is again expected since, as below of slightly above βc. In fact, some of them are shown in Eq. (8), right≈ at the critical point clustering found to be very close to the critical point, like protein- decays logarithmically rather than as a power law. Fi- protein interaction networks of specific human tissues, nally, Fig.3 shows the finite size scaling Eq. (10) both with β 1 or the genetic interaction network of the ≈ for the numerical simulations and numerical integration Drosophila Melanogaster, β 1.1. ≈ of Eq (7). In both cases, we find a very good collapse Our results in this letter show that the coupling of the 5 energies of states and an underlying geometry in systems [6] G. Garc´ıa-P´erez, M. Bogu˜n´a,and M. A.´ Serrano, Nature of noninteracting fermions can lead to anomalous phase Physics 2018 14, 583 (2018). transitions between different topological phases. Despite [7] M. Zheng, A. Allard, P. Hagmann, Y. Alem´an-G´omez, ´ fermions being noninteracting, the set of states that these and M. A. Serrano, Proceedings of the National Academy fermions can occupy are correlated by the triangle in- of Sciences 117, 20244 (2020). equality in the underlying metric space. This correlation [8] M. Bogu˜n´a,D. Krioukov, and K. C. Claffy, Nat Phys 5, 74 (2009). then induces an effective interaction between fermions, [9] M. Bogu˜n´a, F. Papadopoulos, and D. Krioukov, Nat ultimately leading to a clustered phase at low tempera- Commun 1, 1 (2010). tures. Interestingly, the logarithmic dependence of the [10] A. Guly´as,J. J. B´ır´o,A. K˝or¨osi,G. R´etv´ari,and D. Kri- energy of states with the metric distance results in the oukov, Nat Commun 6, 7651 (2015). divergence of the free energy and entropy at a finite tem- [11] A. Allard and M. A.´ Serrano, PLOS Computational Bi- perature βc and, thus, to a different ordering of cycles be- ology 16, e1007584 (2020). [12] M. A. R. Flores and F. Papadopoulos, Physical review low βc, where clustering vanishes in the thermodynamic limit. The finite size behavior of the transition is anoma- letters 121, 258301 (2018). [13] D. Krioukov, F. Papadopoulos, M. Kitsak, A. Vahdat, lous, with ln N and not N playing the role of the system and M. Bogu˜n´a, Phys Rev E 82, 036106 (2010). size. Such slow approach to the thermodynamic limit [14] L. Gugelmann, K. Panagiotou, and U. Peter, in Au- is relevant in real networks in the nongeometric phase, tom Lang Program (ICALP 2012, Part II), LNCS 7392 β . βc, for which high levels of clustering can still be ob- (2012). served. All together, our results describe an anomalous [15] E. Candellero and N. Fountoulakis, Internet Math 12, 2 topological phase transition that cannot be described by (2016). the classic Landau theory but that is, nevertheless, dif- [16] M. A. Abdullah, N. Fountoulakis, and M. Bode, Internet Math 10.24166/im.13.2017 (2017). ferent from other topological phase transitions, such as [17] T. Friedrich and A. Krohmer, SIAM J Discret Math 32, the BKT transition. 1314 (2018). [18] T. M¨ullerand M. Staps, Adv Appl Probab 51, 358 (2019). ´ ACKNOWLEDGMENTS [19]M. A. Serrano, D. Krioukov, and M. Bogu˜n´a, Phys Rev Lett 106, 048701 (2011). [20] N. Fountoulakis and T. M¨uller, Ann Appl Probab 28, We acknowledge support from: the ICREA Academia 607 (2018). award, funded by the Generalitat de Catalunya; Agen- [21] M. Kiwi and D. Mitsche, Ann Appl Probab 28, 941 cia estatal de investigaci´onproject number PID2019- (2018). 106290GB- C22/AEI/10.13039/501100011033; and Gen- [22] M. Bogu˜n´a,D. Krioukov, P. Almagro, and M. A. Serrano, eralitat de Catalunya grant number 2017SGR1064. Phys. Rev. Research 2, 023040 (2020). [23] V. Berezinskii, Sov. Phys. JETP 32, 493 (1971); 34, 610 J. vd K. acknowledges support from the Secretaria (1972). d’Universitats i Recerca de la Generalitat de Catalunya [24] J. M. Kosterlitz and D. J. Thouless, Journal of Physics i del Fons Social Europeu. C: Solid State Physics 6, 1181 (1973). [25] P. Colomer-de Sim´onand M. Bogu˜n´a, Phys Rev E 86, 026120 (2012). [26] P. Hohenberg and A. Krekhov, Physics Reports 572, 1 (2015), an introduction to the Ginzburg–Landau theory ∗ [email protected] of phase transitions and nonequilibrium patterns. † [email protected] [27] J. Park and M. E. J. Newman, Phys Rev E 70, 66117 ‡ [email protected] (2004). [1] M. Boguna, I. Bonamassa, M. D. Domenico, S. Havlin, [28] P. van der Hoorn, G. Lippner, and D. Krioukov, Journal D. Krioukov, and M. A. Serrano, Nature Reviews Physics of Statistical Physics 173, 806 (2018). 3, 114 (2021), arXiv:2001.03241 [physics.soc-ph]. [29] D. Garlaschelli, F. Den Hollander, and A. Roccaverde, [2]M. A.´ Serrano, M. Bogu˜n´a,and F. Sagu´es, Mol Biosyst Journal of Statistical Physics 173, 644 (2018). 8, 843 (2012). [30] P. Erd˝osand A. R´enyi, Publ Math 6, 290 (1959). [3] G. Garc´ıa-P´erez,M. Bogu˜n´a,A. Allard, and M. A. Ser- [31] M. Bogu˜n´aand R. Pastor-Satorras, Phys Rev E 68, rano, Scientific Reports 6, 33441 (2016). 36112 (2003). [4] A. Faqeeh, S. Osat, and F. Radicchi, Phys Rev Lett 121, [32] M. Bogu˜n´a,R. Pastor-Satorras, and A. Vespignani, Eur 098301 (2018). Phys J B - Condens Matter 38, 205 (2004). ´ [5]M. A.´ Serrano, D. Krioukov, and M. Bogu˜n´a, Phys Rev [33] M. Starnini, E. Ortiz, and M. A. Serrano, New Journal Lett 100, 078701 (2008). of Physics 21, 053039 (2019). Supplementary Information for A geometry-induced topological phase transition in random graphs

1, 2, 1, 2, 3, 1, 2, Jasper van der Kolk, ∗ M. Angeles´ Serrano, † and Mari´anBogu˜n´a ‡ 1Departament de F´ısica de la Mat`eriaCondensada, Universitat de Barcelona, Mart´ıi Franqu`es1, E-08028 Barcelona, Spain 2Universitat de Barcelona Institute of Complex Systems (UBICS), Universitat de Barcelona, Barcelona, Spain 3Instituci´oCatalana de Recerca i Estudis Avan¸cats (ICREA), Passeig Llu´ıs Companys 23, E-08010 Barcelona, Spain

CONTENTS

I. Analytics 1 A. Preliminaries 1 B. Free Energy and Entropy 3 C. Scaling Behaviour of Clustering with System Size 3 1. Angular Manipulation 4 2. Case 0 < β < 1 5 3. Case β = 1 11 D. Exponent η 15

II. Real Networks 16

References 16

I. ANALYTICS

A. Preliminaries

The average clustering coefficient for a node with hidden degree κ and angular position θ is defined in the S1 [1] model as N 2 C(κ, θ) = dκ0dκ00dθ0dθ00ρ(κ0)ρ(κ00)F (κ0, κ00, θ0, θ00)F (κ, κ00, θ, θ00)F (κ, κ0, θ, θ0), (S1) k¯(κ, θ)   Z in a network with system size N. Here the function k¯(κ, θ) is the average degree of a node with hidden coodinates (κ, θ), ρ(κ) is the hidden degree density and F (κ, κ0, θ, θ00) is the connection probability between two nodes with hidden coordinates (κ, θ) and (κ0, θ0). The exact form of these functions will be discussed in the following. Note that as the model has rotation symmetry, one only needs to investigate the node at angular coordinate θ = 0. The average clustering coefficient can be computed from C(κ) in the following manner

C = dκ0ρ(κ0)C(κ0). (S2) Z However, as C(κ) is a bounded monotonically decreasing function, is suffices to find the scaling of C(κ) for small arXiv:2106.08030v1 [physics.soc-ph] 15 Jun 2021 κ [2]. In Eq. (S1), ρ(κ) defines the distribution of the hidden degrees. In the following, we always apply a power law distribution. As we are interested in finite-sized scale-free networks, we choose the following distribution γ 1 (γ 1)κ0− γ ρ(κ) = − 1 γ κ− κ0 < κ < κc. (S3) 1 κc − − κ0  

[email protected][email protected][email protected] 2

We choose to not specify the specific form of the cut-offs just yet. We just demand that κ0 is such that the correct average degree is obtained and that, to lowest order, it does not depend on the system size. The average degree of nodes with hidden variable κ and angular position θ is defined as

k¯(κ, θ) = N dκ0dθ0ρ(κ0)F (κ, κ0, θ, θ0). (S4) Z The function F describes the probability of two nodes in the network to be connected and is given by the Fermi-Dirac distribution. As was proved in Ref. [3], the exact form of this function is different for β above and below one. For high temperatures, β < 1, 1 F (κ0, κ00, θ0, θ00) = . (S5) (N∆θ)β 1 + β (2π) µκˆ 0κ00

Here ∆θ = π π θ0 θ00 . The quantityµ ˆ in 1D in this region is given by − | − | − || (1 β) µˆ(N) = − . (S6) 2β k N 1 β h i − The quantityµ ˆ in our model controls the degree distribution of the system. As was explained in the main text, we want the distribution to remain constant when varying β, which is why one needs to introduce different connection probabilities and redefineµ ˆ for the different regimes. For low temperatures, β > 1, we can follow the same logic as was used to derive Eq. (S5) to find 1 F (κ0, κ00, θ0, θ00) = . (S7) N∆θ β 1 + 2πµκˆ κ  0 00  In Ref. [3] it is noted that in this regime one can also define a connection probability in the thermodynamic limit, given in terms of the spatial coordinates x, in our 1D case the coordinates on a infinite line: 1 FN (κ0, κ00, x0, x00) = , (S8) →∞ β x0 x00 1 + | − | µκˆ κ  0 00  In both Eq. (S8) and (S7)µ ˆ is given by

β sin(π/β) µˆ = . (S9) 2π k h i Exactly at the transition pointµ ˆ also takes a different form. As in Ref. [3] this specific case was not investigated we derive it here. One starts with an expression for the expected degree κ. N π dθ κ = dα0ρ(α0) π 1 + Nθ exp(α + α ) Z Z0 2π 0 N 1+ exp(α+α0) 2 dt = 2 dα0ρ(α0) exp( α α0) − − t Z Z1 N = 2 dα0ρ(α0) exp( α α0) ln 1 + exp(α + α0) . (S10) − − 2 Z  Now, for large system sizes one obtains

κ = 2 ln(N) dα0ρ(α0) exp( α α0) = 2 ln(N) exp( α) exp( α) . (S11) − − − h − i Z Using that κ = k , we have h i h i k = 2 ln(N) exp( α) 2, (S12) h i h − i 3 which implies that 1 1 κ = (2 k ln(N))1/2 exp( α) α = ln(κ) + ln (S13) h i − ⇒ − 2 2 k ln(N)   h i  Using now that α = (ln κ + 1 lnµ ˆ)[3] andµ ˆ = exp µ one finally finds − 2 1 µˆ = 2 k ln(N) − . (S14) h i  

B. Free Energy and Entropy

As was explained in the main text, for β > 1, the edges can be taken to represent noninteracting fermions with energy equal to x  = ln ij . (S15) ij κ κ  i j  One can then define the grand canonical partition function of the system as follows

β x − ln = ln 1 + ij , (S16) Z µκˆ κ i,j " i j # X   whereµ ˆ = exp µ, with µ the chemical potential. Now, as similarity space is homogeneous, one can place the i’th node on the origin, leading to N identical terms. Then we assume to be working with a large system size allowing us to approximate the sums in Eq. (S16) by integrals. This leads to the following expression

β x − ln = N dκdκ0ρ(κ)ρ(κ0) dx ln 1 + Z µκκˆ ZZ Z "  0  # β = Nµˆ dκdκ0ρ(κ)ρ(κ0)κκ0 dt ln 1 + t− ZZ Z µˆ k 2π Nβ k   = N h iπ = h i, (S17) sin β 2 where in the last stepµ ˆ was plugged in. Note that the expressions is multiplied with a factor N, which of course makes our quantities diverge in the thermodynamic limit. However, it is not the thermodynamic quantities themselves but their densities we are interested in, thus we can divide away the factor N. In fact, we choose to look at the thermodynamic quantities per link, where the amount of links in the system is N k /2. We can then use the above expression to define the grand potential h i 1 1 Ξ = ln = N k , (S18) −β Z −2 h i which can in turn be used to find the free energy of the system

2F 2Ξ β π β β+ = + µ = µ 1 = ln sin 1 → c ln(β 1) (S19) N k N k − 2π k β − ∼ − h i h i  h i  From this, we can find the entropy of the system 2 2S 2β ∂F π β β+ 1 = = β π cot → c . (S20) N k N k ∂β − β ∼ β 1 h i h i −

C. Scaling Behaviour of Clustering with System Size

In the following section we find the dominant finite size scaling of the clustering coefficient for β 1. Note that for β > 1 we already know that clustering is independent of system size as the connection probability≤ is finite in thermodynamic limit. We will look at the properties of this regime in section (I D). We start by manipulating the angular integrals of Eq. (S1) as to simplify the task at hand later on. We then turn to the scaling when β < 1 and conclude with an analysis of the scaling when β = 1. 4

1. Angular Manipulation

We start by manipulating the angular integrals of Eq. (S1) to make it easier to work with, i.e. get rid of the absolute values in the expressions for ∆θ. The equation has the following form:

dκ0dκ00dθ0dθ00ρ(κ0)ρ(κ00)F (κ, κ0, π π θ0 )F (κ, κ00, π π θ00 )F (κ0, κ00, π π θ0 θ00 )) C(κ) = − | − | || − | − | || − | − | − || . dκ dθ ρ(κ )F (κ, κ , π π θ ) RRRR 0 0 0 0 − | − | 0|| (S21) RR Here, we have used θ = 0. Let us first investigate the trivial case of the denominator, where we only focus on the angular integral

2π π 2π dθ0F (κ, κ0, π π θ0 ) = dθ0F (κ, κ0, π π θ0 ) + dθ0F (κ, κ0, π π θ0 ) 0 − | − | || 0 − | − | || π − | − | || Z Z π 2π Z π = dθ0F (κ, κ0, θ0) + dθ0F (κ, κ0, 2π θ0) = 2 dθ0F (κ, κ0, θ0), (S22) − Z0 Zπ Z0 where in the last step we have performed the transformation t = 2π θ0 and t θ0 on the second integral. The numerator can be rewritten in a similar way to obtain four terms − →

2π 2π dθ0 dθ00F (κ, κ0, π π θ0 )F (κ, κ00, π π θ00 )F (κ0, κ00, π π θ0 θ00 ) − | − | || − | − | || − | − | − || Z0 Z0 π θ0 θ0 =2 dθ0 dθ00F (κ, κ0, θ0)F (κ, κ00, θ00)F (κ0, κ00, θ0 θ00) + dθ00F (κ, κ0, θ00)F (κ, κ00, θ0)F (κ0, κ00, θ0 θ00) − − Z0  Z0 Z0 π θ0 π − + dθ00F (κ, κ0, θ0)F (κ, κ00, θ00)F (κ0, κ00, θ0 + θ00) + dθ00F (κ, κ0, θ0)F (κ, κ00, θ00)F (κ0, κ00, 2π θ0 θ00) . 0 π θ − − Z Z − 0  (S23) The first two terms are not exactly the same. However, as the full expression of the clustering coefficient also contains integrals over the hidden degrees, one can interchange κ0 κ00 together with θ0 θ00. This thus shows that the first two terms contribute equally to the clustering coefficient.↔ All in all, we will thus↔ be working with the following three terms

π θ0 4 dθ0 dθ00F (κ, κ0, θ0)F (κ, κ00, θ00)F (κ0, κ00, θ0 θ00) − Z0 Z0 π π θ0 − +2 dθ0 dθ00F (κ, κ0, θ0)F (κ, κ00, θ00)F (κ0, κ00, θ0 + θ00) 0 0 Z π Z π +2 dθ0 dθ00F (κ, κ0, θ0)F (κ, κ00, θ00)F (κ0, κ00, 2π θ0 θ00). (S24) 0 π θ − − Z Z − 0 Now, before we get started on finding the scaling with respect to the system size of each term individually, it might be that we can avoid doing so by some simple arguments. Indeed, we will show that the first term will always dominate the others in the large N limit, and so we only have to find its scaling. Let us start with the second term

π π θ0 − 2 dκ0dκ00ρ(κ0)ρ(κ00) dθ0 dθ00F (κ, κ0, θ0)F (κ, κ00, θ00)F (κ0, κ00, θ0 + θ00) 0 0 ZZ Z π Z π 2 dκ0dκ00ρ(κ0)ρ(κ00) dθ0 dθ00F (κ, κ0, θ0)F (κ, κ00, θ00)F (κ0, κ00, θ0 + θ00). (S25) ≤ ZZ Z0 Z0 The above statement is true as the integrand is strictly positive and so extending the integration domain will only make the integral larger. Now, we can split the θ00 integral and perform θ0 θ00 and κ0 κ00 on the second term to obtain ↔ ↔ π π 2 dκ0dκ00ρ(κ0)ρ(κ00) dθ0 dθ00F (κ, κ0, θ0)F (κ, κ00, θ00)F (κ0, κ00, θ0 + θ00) ZZ Z0 Z0 π θ0 = 4 dκ0dκ00ρ(κ0)ρ(κ00) dθ0 dθ00F (κ, κ0, θ0)F (κ, κ00, θ00)F (κ0, κ00, θ0 + θ00) ZZ Z0 Z0 5

π θ0 4 dκ0dκ00ρ(κ0)ρ(κ00) dθ0 dθ00F (κ, κ0, θ0)F (κ, κ00, θ00)F (κ0, κ00, θ0 θ00). (S26) ≤ − ZZ Z0 Z0 In the final step we use the functional form of F with respect to the angular coordinate is 1 F (x) = . (S27) 1 + xβ As xβ is monotonously increasing, and 1/(1+x) is monotonously decreasing, F (x) is monotonously decreasing. Thus, it is largest when x is smallest. Obviously, θ0 + θ00 > θ0 θ00 for all (θ0, θ00) [0, π] [0, θ0]. We have thus proven that the first term in Eq. (S24) dominates the second term.− We can follow similar∈ steps× for the third term. We we will now only clarify steps if they are new.

π π

2 dκ0dκ00ρ(κ0)ρ(κ00) dθ0 dθ00F (κ, κ0, θ0)F (κ, κ00, θ00)F (κ0, κ00, 2π θ0 θ00) − − κZZ,κ Z0 π Zθ 0 00 − 0 π θ0

4 dκ0dκ00ρ(κ0)ρ(κ00) dθ0 dθ00F (κ, κ0, θ0)F (κ, κ00, θ00)F (κ0, κ00, 2π θ0 θ00). (S28) ≤ − − κZZ0,κ00 Z0 Z0

Now, one knows that 2π θ0 θ00 θ0 θ00 (θ ,θ ) [0,π] [0,θ ]. For the same reasons as before, this then implies − − ≥ − ∀ 0 00 ∈ × 0

π θ0

4 dκ0dκ00ρ(κ0)ρ(κ00) dθ0 dθ00F (κ, κ0, θ0)F (κ, κ00, θ00)F (κ0, κ00, 2π θ0 θ00) − − κZZ0,κ00 Z0 Z0

π θ0

4 dκ0dκ00ρ(κ0)ρ(κ00) dθ0 dθ00F (κ, κ0, θ0)F (κ, κ00, θ00)F (κ0, κ00, θ0 θ00), (S29) ≤ − κZZ0,κ00 Z0 Z0 so this term is also dominated by the first term in Eq. (S24).

2. Case 0 < β < 1

The first step is to perform the transformation x = κ0 and y = κ00 , where we define κ2 N β/((2π)βµˆ). This leads κs κs s to ≡

κc/κs κc/κs π θ0 γ dx dy dθ0 dθ00(xy)− F (κ, x, θ0)F (κ, y, θ00)F (x, y, θ0 θ00) − κ0/κs κ0/κs 0 0 C(κ) = 2 R R R R 2 + h.o.t., (S30) κc/κs π γ dx dθ0x− F (κ, x, θ0) κ0/κs 0 ! R R We investigate the numerator and denominator separately and define

κc/κs κc/κs π θ0 γ A = dx dy dθ0 dθ00(xy)− F (κ, x, θ0)F (κ, y, θ00)F (x, y, θ0 θ00). (S31) − − κ0Z/κs κ0Z/κs Z0 Z0

κc/κs π γ B = dx dθ0x− F (κ, x, θ0). (S32)

κ0Z/κs Z0 It is also useful to define

κc/κs κc/κs π θ0 γ A+ = dx dy dθ0 dθ00(xy)− F (κ, x, θ0)F (κ, y, θ00)F (x, y, θ0 + θ00). (S33)

κ0Z/κs κ0Z/κs Z0 Z0 6

Our investigation will focus on finding upper and lower bounds for these integrals. Using the fact that 1 1 β < β , θ0,θ00,x,y, (S34) (θ0 + θ00) (θ0 θ00) ∀ 1 + 1 + − xy xy we can conclude that A+ < A . As numerical investigation leads us to expect that both have the same scaling, this implies that we do not need to− worry about an upper bound for A+ nor the lower bound for A . If the functions f(N) and g(N) in equation −

f(N) < A+ < A < g(N) (S35) − have the same dominant scaling, one can immediately conclude that A also has that exact dominant scaling. One − might ask why we introduce A+ in the first place, when in the end we are only interested in the scaling of A . The − answer to this is that A+ in general has nicer properties due to the lack of (θ0 θ00), as it is thus easier to find a lower bound for it than for A . − − We start with the simplest integral, the B-term, which can be solved exactly. To this end we first need to rewrite it a bit. By performing two substitutions

κs θ0 x0 = x x0 x, t = t θ0, (S36) κc → π → one obtains

1 1 1 γ γ κc − x− B = π dθ0 dx β 2 . (S37) κs (πθ0) κs   Z0 κ0Z/κc 1 + κcκx This then gives the following expression

1 γ β 2 π κ0 − 1, 1/β π κ B = (γ 1)β F ; s (β(γ 1) 1)(γ 1) − κ 2 1 1 + 1/β − κκ − − − (  s   0  1 γ β 2 κc − 1, 1/β π κ (γ 1)β F ; s − − κ 2 1 1 + 1/β − κκ  s   0  1 γ β 2 κ0 − 1, γ 1 π κs κ0 2F1 − ; − κ γ − κκ  s   0  1 γ β 2 κc − 1, γ 1 π κs + κ0 2F1 − ; . (S38) κ γ − κκ  s   0  ) a, b Here, F ; z is the ordinary hypergeometric function [4] To lowest order, using κ √N, one finds that B 2 1 c s ∼ then scales as 

γ 3 B N −2 (S39) ∼

Next we turn to the A+ term. From the form of the standard connection probability given in Eq. (S27), we see that it is smallest when x is largest, which is the case when θ0, θ00 are largest, so when they are both π. Plugging this in we obtain

γ 1 1 1 A+ dxdy(xy)− ≥ πβκ πβκ (2π)β Z 1 + s 1 + s 1 + κx κy xy 2 3β κ 2 γ 1 1 1 = π− dxdy(xy) − κx κy xy . (S40) κs   Z 1 + β 1 + β 1 + β π κs π κs (2π) 7

Now, this is exactly the same integral (with the exception of the π’s, but they will obviously not change scaling) as the one evaluated in Ref. [2]. As was found in the reference, the scaling depends on whether how we set κc relative to κ . We distinguish two regimes. First there, is the regime where κ κ κ . In this case, the scaling is s 0  s  c 2 A c κ− ln(κ /κ ). (S41) + ≥ +,1 s c s Then there is the region where κ κ κ (κ κ must be required to hold) where one obtains 0 ≤ c ≤ s 0  s 2γ 8 6 2γ A c κ − κ − . (S42) + ≥ +,2 s c This, however, does not give the full scaling behaviour, as numerical results show us that for large β there is different scaling. To find where this different scaling comes from we take a step back and look at the full integral A+ as given in Eq. (S33). One might be tempted to, as in Ref. [2], expand the first two connection probabilities to first order. However, the presence of the angular coordinate makes this impossible. The argument of these connection probabilities β θ κs has the form x = κx . It becomes clear that for small enough θ, x no longer is large and the approximation thus breaks down. We thus expect different scaling behaviour to arise as a result of small angular coordinates. To investigate this further, we split the integration domain in a convenient way and investigate the domain = [0, (xy)1/β] [0, t]. D1 × However, this is only possible in the case that both κc κs, as only then the angular coordinates remain small for ≤ 2/β all x and y. For the case that κc κs we define the more restrictive domain 2 = [0, (κ0/κs) ] [0, t]. Starting with the case κ κ one finds  D × c ≤ s 1 2/β γ 1 1 A+ β dxdy(xy) − κ y κ x ≥ 1 + 2 1 + s 1 + s Z κ κ 4/β+2γ 2 2 (κs/κ)− − 2 2 2 2 = B κ γ , 1 γ + B κ γ , 1 γ + 1 + 2β κ0+κ − β − β − κc+κ − β − β      4/β+2γ 2 4/β+2γ 2 4/β 2γ c κ− − + c κ− − κ − , (S43) ∼ +,s,1 s +,s,2 s c where B [a, b] is the incomplete beta function. For the case κ κ one obtains z c  s 4/β κ0 γ 1 1 1 A+ dxdy(xy)− ≥ κ κ κ2 κ κ2 2β κ2  s  Z 1 + s 0 1 + s 0 0 2 2 1 + 2 κx κs κy κs xy κs 4/β κ0 γ 4/β+2γ 2 dxdy(xy)− c κ− − , (S44) ≈ κ ∼ +,s,3 s  s  Z where in the first step it was noted that irrespective of the value of x and y, the argument of the connection probabilities is small. We now have five different scaling behaviours. Which terms dominate will depend on the value of β as well on κc. α/2 To quantify how the scaling varies with κc we introduce the exponent α such that κc N . The different regimes of κ described above correspond to α [0, 1] for κ κ and α (1, 2 ] for κ ∼κ . Using these definitions we c ∈ c ≤ s ∈ γ 1 c  s can conclude that −

2/β+γ 1 1 C N − − + C N − ln N if κ κ +,1 +,2 c  s A+ 1 γ 2/β (1 α)(γ 2/β) (1 α)(γ 3) , (S45) ≥ N − C N − + C N − − + C N − − if κ κ +,3 +,4 +,5 c ≤ s    where C+,i are constants. Now obviously this is just a lower bound. To show that the clustering indeed scales like this we must also find an 2/β upper bound, which we do by turning to the A term. We divide the integration domain in two: s = [0, (κ0/κs) ] 2/β − D × [0, θ0] and = [(κ /κ ) , π] [0, θ0]. We first turn to region . Dl 0 s × Dl

γ 1 1 1 A ,l = dθ0dθ00 dxdy(xy)− β β β − θ0 κs θ00 κs (θ0 θ00) ZZDl ZZ 1 + 1 + 1 + − κx κy xy 2 κ 2 γ β 1 dθ0dθ00 dxdy (xy) − (θ0θ00(θ0 θ00))− xy ≤ κs l − 1 + (θ θ )β   ZZD ZZ 0− 00 8

θ00

X D

a

Z Y D D b a θ0

FIG. S1. Integration regions. In the grey region ( Y + Z ) the argument of the Lerch zeta function is bigger than one, in the D D hatched region ( X ) it is not and the black region is s. D D

2 2(3 γ) 2 κ dθ0dθ00 κc − β κc = Φ (θ0 θ00)− , 2, 3 γ κ (θ θ (θ θ ))β κ κ s l 0 00 0 00 s "− − s − #   ZZD −     2 2(3 γ) 2 κ dθ0dθ00 κ0 − β κ0 + Φ (θ0 θ00)− , 2, 3 γ κ (θ θ (θ θ ))β κ κ s l 0 00 0 00 s "− − s − #   ZZD −     2 3 γ κ dθ0dθ00 κ0κc − β κ0κc 2 β 2 Φ (θ0 θ00)− 2 , 2, 3 γ . (S46) − κs (θ0θ00(θ0 θ00)) κs − − κs −   ZZDl −     Here, Φ[z, a, b] is the Lerch zeta function. One sees that these three terms are similar, and so we treat the general integral

β β Φ (θ0 θ00)− ζ, 2, 3 γ Φ θ00− ζ, 2, 3 γ − − − 3 γ − − 3 γ Iζ = dθ0dθ00 β ζ − = dθ0dθ00 β ζ − , (S47) (θ0θ00(θ0 θ00)) (θ0θ00(θ0 θ00)) ZZDl  −  ZZDl  −  where the transformation θ000 = θ0 θ00, θ000 θ00 was performed. Now, the argument of the Lerch zeta function can − → β γ 3 in principle be smaller and larger than one. If it is smaller, it can be shown that Φ[ (θ0 θ00)− ζ, 2, 3 γ] < 2 − . If it is bigger one can use the identity described in Ref. [2] − − −

2 2(3 γ) 1 1 Φ[ z , 2, 3 γ] = z− − 2ψ(γ) ln z + ϑ(γ) + Φ , 2, γ 2 , (S48) − − z2 z2 −     where

ψ(γ) = Φ[ 1, 1, 3 γ] + Φ[ 1, 1, γ 2] and ϑ(γ) = π2 cot(πγ) csc(πγ). (S49) − − − − − The transition point between these two cases of course lies at

a = ζ1/β (S50)

2/β We must thus split the integration domain in three regions (where b = (κ /κ ) ): = [a, π] [a, θ0], = Dl 0 s DX × DY [a, π] [0, a] and Z = [b, a] [0, θ0] as depicted in Fig. S1. Now, the grey region is the one where the Lerch zeta function× argumentD is bigger than× one, in the hatched region we can bound the Lerch zeta function away and the black region is s and we thus do not care about it for the moment. Before going any further, let us note that Fig. S1 looks slightly differentD for different κ and ζ. If κ κ and ζ = (κ /κ )2, then ζ 1 and thus so is a. However, a as an c c  s c s  integration limit must be smaller than π and thus in this case the X and Y regions disappear. When κc κs this D D 2 ≤ is not the case as for all ζ, a < π. Finally, irrespective of the value of κc, for ζ = (κ0/κs) , a = b and thus region Z vanishes. Implementing the transformation given by Eq. (S48) in the grey region one obtains D

β β(2 γ) ζ 2 γ 2 Iζ dθ0dθ00(θ0(θ0 θ00))− θ00 − ψ(γ) ln β + ϑ(γ) + ζ − (γ 2)− . (S51) ≤ − ( θ00 − ) ZZ h   i 9

As this leads to three different angular integrals, in the end we have seven different integrals to solve.

β 2 3β 1 a − dθ0dθ00 = B1[2β 1, 1 β] B1[1 β, 1 β] θ0θ00(θ0 θ00) 3β 2 − − − − − ZZDX  −  − 

3β 2 + B a [2β 1, 1 β] + (a/π) − B a [1 β, 1 β] π − − π − −  β 1/2 5/2 3β 4 − π − Γ[1 β] 2 3β + − (a/π) − 1 (S52) Γ[3/2 β](3β 2) − − −  2 3β = cX11 a − + cX12 (S53)

β 2 2β 1 a − dθ0dθ00 = 2(β 1)B a [2β 1, 1 β] 2 π θ0(θ0 θ00) 2(β 1) − − − ZZDY  −  −  1/2 π− (β 1)Γ[1 β]Γ[β 1/2] 1 − − − − −

2(β 1), β 2β 2 + (1 2F1 − ; a/β )(a/π) − (S54) − 2β 1  −   2 2β c a − + c (S55) ∼ Y11 Y12

2 γ β 2 γβ θ00 − a − dθ0dθ00 = B1[1 + 2β γβ, 1 β] θ0(θ0 θ00) γβ 2 − − ZZDY  −  − 

B1[2β 1, 1 β] + B a [2β 1, 1 β] − − − π − −

γβ 2 (a/π) − B a [1 + 2β γβ, 1 β] (S56) − π − −  2 γβ 1+2β γβ c a − + c a − (S57) ∼ Y21 Y22

2 γ β 2 βγ 1 2β θ00 − ζ βa − π − dθ0dθ00 ln β = 2 θ0(θ0 θ00) θ00 (β(γ 2) 1)(βγ 2) ZZDY  −    − − − β 1 4 − 1 3 (β(γ 2) 1)Γ[1 β]Γ β × 2 2β − − − − 2 π −   2β 1 π − Γ[1 β]Γ[ γβ + 2β + 2] + − − Γ[ γβ + β + 2] − 1 + (γβ 2)(Hβ(2 γ) H1+β γβ) − − − − × β(γ 2) 1  − − 

2β 1 (β(γ 2) 1) β, 2β 1 a a − − − 2F1 − ; − 2β 1 2β π  −   (βγ 2) β, γβ + 2β + 1, γβ + 2 a − 3F2 − − ; β(γ 2) 1 β + 1 γβ + 2β + 2, γβ + 2β + 2 π − −  − −  β, β( γ) + 2β + 1 a + 2F1 − ; (S58) β( γ) + 2β + 2 π  −  ! 10

2 γβ 1+2β γβ c a − + c a − (S59) ∼ Y31 Y32

β 2 2β 2 2β 1 a − b − 2 2β 2 2β dθ0dθ00 = − 2 = cZ11 a − + cZ12 b − (S60) θ0(θ0 θ00) 2(β 1) ZZDZ  −  −

2 γ β 2 βγ 2 βγ θ00 − Γ[1 β]Γ[ γβ + 2β + 1] b − a − dθ0dθ00 = − − − (S61) θ0(θ0 θ00) (βγ 2)Γ[ γβ + β + 2] ZZDZ  −  − −  2 γβ 2 γβ = cZ21 a − + cZ22 b − (S62)

2 γ β θ00 − ζ Γ[1 + 2β γβ]Γ[1 β]β 2 γβ 2 γβ dθ0dθ00 ln β = − − a − b − θ0(θ0 θ00) θ00 (βγ 2)Γ[2 + β γβ] − ZZDZ  −    − −  1 1 Hβ(2 γ) H1+β γβ + log(ζ) × − − − γβ 2 − β  − 2 γβ 2 γβ a − log a b − log b + − (S63) a2 γβ b2 γβ − − − 

2 γβ 1 a − cZ31 + cZ32 log(a) β log(ζ) = − (S64) 2 γβ  1   b − c + c log(b) log(ζ) − Z31 Z32 − β     The next step is to organise the different scalings (see Tab. (S1), where we have defined cYi = cY1i + cY2i + cY3i and similarly for Z.) that were found and find which is dominant.

2 2 2 κ0 κ0κc κc κc ζ = ζ = 2 ζ = (κc κs) ζ = (κc κs) κs κs κs κs 4 2  4 ≤ β 2γ γ β 2γ κ0 − κ0κc β −    κc − cX11 + cX11 2 + cX11 + κs κs κs X 2(3 γ) 3 γ 2(3 γ)  κ0  −  κ0κc  −  κc  − cX12 cX12 2 cX12 κs κs κs 4 2 4 β 2γ γ 2γ κ0  − κ0κc β − κc β − cY 1 + cY 1 2 + cY 1 + κs κs κs Y 4+ 2 2γ 2+1/β γ 4+2/β 2γ  κ0  β −  κ0κc  −  κc  − cY 2 cY 2 2 cY 2 κs κs κs 2(2 γ)     2 2β κc −   cZ π − + 11 κs 2 2 γβ 4 γ − 2γ κ0κc β − (cZ21 + cZ31) π + κc β − cZ1 2 + 4 cZ1 + κs β 2γ κs 4 κ0 − 4 2γ (cZ22 cZ32 ) κ 2γ   κ0 β − − s −   κ0 β − (cZ22 cZ31 ) + (cZ22 cZ31 ) + κs 2 2 γβ κc κs Z − 4 cZ π − ln  + − 4 2γ β 32 κs 2γ cZ κ β −  κ 2cZ κ β −  κ 32 0 ln c + 4 2γ 32 0 ln c + β κs κ0 2 κ0 β −  κc β κs κ0 4 β cZ32 κ ln κ + 4 2 γ β 4 s 0 4 2γ β 4 κ0κc − κ0 − 2 γβ κc − κ0  − cZ12 2 − cZ12 κs κs cZ23 π  ln(π)+   κs κs 2(2 γ) 4 4     κc − κ0 β −     cZ 12 κs κs     TABLE S1. The different terms resulting from (S46).

Let us note that as the final results contains I 2 2 2I 2 , the terms containing ln(κ /κ ) cancel. We now κc /κs κ0κc/κs c 0 have many different scaling behaviours, and the question− of which one dominates again depends on the value of β as 2 well as κc. As a matter of fact, if one includes the κs− pre-factor in Eq. (S46), one recovers the same behaviour as was found for the lower bound 2/β+γ 1 1 C ,1N − − + C ,2N − ln N if κc κs − −  A ,l 1 γ 2/β (1 α)(γ 2/β) (1 α)(γ 3) , (S65) − ≤ N − C ,3N − + C ,4N − − + C ,5N − − if κc κs − − − ≤     11 where C ,i are constants. This seems− to go in the right direction. However, we have not explored the full integration domain yet. It turns out though that the integration domain does not lead to any new scaling: Ds

γ 1 1 1 I ,s = dxdy(xy)− β β β − θ0 κs θ00 κs (θ0 θ00) ZZ ZZDs 1 + 1 + 1 + − κx κy xy

γ dxdy(xy)− 1 ≤ ZZ ZZDs 4/β κ0 γ = dxdy(xy)− κ  s  ZZ 4/β 1 γ 1 γ 2 κ 1 κ − κ − = 0 c 0 κ (1 γ)2 κ − κ  s  −  s   s  ! 2(1 γ+2/β) 1 κ0 − ∼ (1 γ)2 κ −  s  1+γ 2/β N − − . (S66) ∼

The contribution of s is thus subleading for small β and equally dominant as the other contributions for large β. We have thus shown thatD for the the upper and lower bound the dominant scaling is the same. We now have the scaling of all distinct parts, B,A ,A , so we can now combine them all. + −

2 2/β 2 γ C1N − + C2N − ln N if κc κs C 2 2/β 2 2/β α(γ 2/β) 1 α(γ 3)  (S67) ∼ C N − + C N − − − + C N − − − if κ κ ( 3 4 5 c ≤ s

Let us discuss the limiting cases of α. When α = 0, κ0 κc, and the network thus has a homogeneous degree 2 2/β 1 ≈ 2 2/β distribution. Then, C (C3 + C4)N − + C5N − . If α = 1, i.e. κc κs, the scaling becomes C C3N − + 2 γ ∼ ≈ ∼ (C4 + C5)N − .

3. Case β = 1

1 1 β We now turn to the limit β = 1. We know that in this case µ scales asµ ˆ (ln N)− instead ofµ ˆ N − , and ∼ ∼ thus κ √N ln N, which of course alters scaling. However,there might be more differences. We will represent all s ∼ integrals evaluated at β = 1 by a tilde (A˜ , A˜+, B˜). We start with B˜: − 1 1 1 γ γ ˜ κc − x− B = π dθ dx 2 κs πθκs   Z0 κ0Z/κc 1 + κcκx πκ2 1 γ log 1 + s 2 γ 2 κ − κκ κcκ 1 κ − κκ πκ = π c c + 0 c log 1 + s κ πκ2 2 γ  γ 2 κ πκ2 κκ  s   s − −  c  s  0  2 1 γ 2 1 1, γ 1 πκs κ0 − 1, γ 1 πκs + 2F1 − ; 2F1 − ; . (S68) (γ 2)(γ 1) γ − κκ − κ γ − κκ − −  c   c   0 !  The second term is dominant and thus B scales as

γ 3 γ 1 γ 3 − − B˜ κ − log(κ ) N 2 (log N) 2 . (S69) ∼ s s ∼ For the lower bound of the numerator of the clustering coefficient we can use the result found in Eq. (S45) as nowhere was it assumed that β < 1. Irrespective of κc this gives us

γ 3 γ 3 A˜ c˜ N − (ln N) − . (S70) + ≤ + 12

For the upper bound of A˜ we cannot follow the same path as was done in the case of general β. This is because the upper bound employed, given− by Eq. (S46), diverges in the β = 1 limit. Thus, we must find a stricter bound. This is 2 done by once again dividing the angular integration domain, this time in four pieces: s = [0, (κ0/κs) ] [0, θ0], 2 = 2 2 2 2 D2 2 × D 2 [(κ0/κs) , π] [0, (κ0/κs) ], 3 = [(κ0/κs) , π] [θ0 (κ0/κs) , θ0] and 3 = [2(κ0/κs) , π] [(κ0/κs) , θ0 (κ0/κs) ], as represented× in Fig. S2 D × − D × −

π b θ00 −

b

b θ0

2 κ0 FIG. S2. Integration regions where b = 2 The black region is region s. The horizontally striped region is region 2. The κs D D vertically striped region is region 3. The grey region is region 4. D D

Note that regions 2 and 3 overlap, but that is not a problem as our integrand is positive and counting a region double just increasesD the valueD of the integral, which in turn work for our purposes as we are only looking for an upper bound. For the region we can use the result (S66): Ds γ 3 γ 3 A˜ ,s c˜ ,sN − (ln N) − . (S71) − ≤ − Turning to we obtain D2

γ dθ0dθ00 1 1 A˜ ,2 = dxdy(xy)− − θ0κs θ00κs θ0 θ00 ZZ ZZD2 1 + 1 + 1 + − κx κy xy

κs/κ0 κs/κ0 κ dθ dθ xγ 3yγ 2 0 00 dx dy − − (S72) ≤ κs 2 θ0 1 + xy(θ0 θ00) ZZD κsZ/κc κsZ/κc − where we have bounded the integral by decreasing the size of the denominators of the first and second terms. We also performed a change of variables of x and y. We now extend the lower bounds of the x and y integrals to zero, which can be done as our integral is positive, and so the resulting integral will be larger or equal to the original one.

κs/κ0 κs/κ0 γ 3 γ 2 κ dθ0dθ00 x − y − A˜ ,2 dx dy − ≤ κs 2 θ0 1 + xy(θ0 θ00) ZZD Z0 Z0 − 2 2 κ 2γ 3 dθ0dθ00 κs κs = (κs/κ0) − Φ 2 (θ0 θ00), 1, γ 2 Φ 2 (θ0 θ00), 1, γ 1 (S73) κs θ0 −κ0 − − − −κ0 − − ZZD2       We know again have the situation that depending on the values of the angular coordinates, the arguments of the Φ’s diverge or go to zero. For the region = [b, 2b] [0, b], θ0 θ00 [0, b], so the argument lies between zero and one. D2s × − ∈ For the region 2l = [2b, π] [0, b], θ0 θ00 [b, π], so the argument is larger than one. We first turn to the second region. Here theD argument can× diverge− and∈ we should thus perform a similar transformation as Eq. (S48). It is not 13 exactly the same as the second argument of the Φ’s is now 1 and not two 2, but the derivation is equivalent. This leads us to

2 2 κ 2γ 3 dθ0dθ00 κs κs (κs/κ0) − Φ 2 (θ0 θ00), 1, γ 2 Φ 2 (θ0 θ00), 1, γ 1 κs θ0 −κ0 − − − −κ0 − − ZZD2l       2 2 γ κ 2γ 3 dθ0dθ00 κs − = (κs/κ0) − 2 (θ0 θ00) Φ[ 1, 1, 3 γ] + Φ[ 1, 1, 2 γ] κs θ0 κ0 − − − − − ZZD2l    2 1 2 1  κs − κs − (θ0 θ00) Φ (θ0 θ00) , 1, 3 γ − κ2 − − κ2 − −  0  "  0  # 2 1 γ κs − + (θ0 θ00) Φ[ 1, 1, 2 γ] + Φ[ 1, 1, 1 γ] κ2 − − − − −  0  2 1 2 1  κs − κs − (θ0 θ00) Φ (θ0 θ00) , 1, 2 γ − κ2 − − κ2 − −  0  "  0  #  2 2 γ κ 2γ 3 dθ0dθ00 κs − (κs/κ0) − 2 (θ0 θ00) Φ[ 1, 1, 3 γ] + Φ[ 1, 1, 2 γ] ≤ κs θ0 κ0 − − − − − ZZD2l    2 1 γ  κs − + (θ0 θ00) Φ[ 1, 1, 2 γ] + Φ[ 1, 1, 1 γ] κ2 − − − − −  0  2 1  κs − 2(γ 3) γ 3 γ 3 2 (θ0 θ00) κ − N − (ln N) − . (S74) − κ2 − ∼ s ∼  0   For we can immediately bound away the Φ to find D2s 2 2 κ 2γ 3 dθ0dθ00 κs κs (κs/κ0) − Φ 2 (θ0 θ00), 1, γ 2 Φ 2 (θ0 θ00), 1, γ 1 κs θ0 −κ0 − − − −κ0 − − ZZD2s       κ 2γ 3 dθ0dθ00 κ 2γ 5 γ 3 γ 3 (κs/κ0) − = (κs/κ0) − ln 2 N − (ln N) − . (S75) ≤ κs θ0 κs ∼ ZZD2s γ 3 γ 3 Combining the two results we find that A˜ ,2 I N − (ln N) − , where I is some integral, as expected. Then we investigate to : − ≤ ∼ D3

γ 1 1 1 A˜ ,3 = dxdy(xy)− dθ0dθ00 − θ0κs θ00κs θ0 θ00 ZZ ZZD3 1 + 1 + 1 + − κx κy xy

γ 1 1 1 = dxdy(xy)− dθ0dθ00 θ0κs (θ0 θ00)κs θ00 ZZ ZZD2 1 + 1 + − 1 + κx κy xy 2 κ 1 γ 1 γ 1 1 dxdyx − y − dθ0dθ00 ≤ κs θ0 θ0 θ00   ZZ ZZD2 − 2 2(2 γ) 1 κ κ − π2 κ2 = 0 Li 0 (2 γ)2 κ κ 6 − 2 κ2π −  s   s    s   2(γ 3) γ 3 γ 3 κ − N − (ln N) − . (S76) ∼ s ∼ Here Li (z) is the dilogarithm. The final region to be studied is : 2 D4

γ 1 1 1 A˜ ,4 = dxdy(xy)− dθ0dθ00 − θ0κs θ00κs θ0 θ00 ZZ ZZD4 1 + 1 + 1 + − κx κy xy 2 κ 1 γ 1 1 dxdy(xy) − dθ0dθ00 θ θ ≤ κs θ0θ00 1 + 0− 00   ZZ ZZD4 xy 14

κ /κ κ /κ 2 s 0 s 0 κ γ 3 1 1 = dx dy(xy) − dθ0dθ00 κs 4 θ0θ00 1 + xy(θ0 θ00)   κsZ/κc κsZ/κc ZZD − κ /κ κ /κ 2 s 0 s 0 κ γ 3 1 1 dx dy(xy) − dθ0dθ00 ≤ κs 4 θ0θ00 1 + xy(θ0 θ00)   Z0 Z0 ZZD − 2 2(γ 2) 2 κ κs − 1 κs = dθ0dθ00 Φ 2 (θ0 θ00), 2, γ 2 κs κ0 θ0θ00 −κ0 − −     ZZD4   2 2 κ 1 2 γ κs dθ0dθ00 (θ0 θ00) − Ψ(.) log 2 (θ0 θ00) + ϑ(.) ≤ κs θ0θ00 − κ0 −   ZZD4      2(γ 3) κs − 1 2 + (θ0 θ00)− (3 γ)− . (S77) κ − −  0   Let us investigate the term with the logarithm first.

2 2 π θ0 κ0/κs 2 − 2 γ 2 κ (θ0 θ00) − κs dθ0 dθ00 − log 2 (θ0 θ00) κs θ0θ00 κ0 −   2Z 2 2Z 2   2κ0/κs κ0/κs 2 2 πκs/κ0 θ0 1 2 2(2 γ) − 2 γ κ κ0 − (θ0 θ00) − = dθ0 dθ00 − log (θ0 θ00) κs κs θ0θ00 −     Z2 Z1 2 2 πκs/κ0 θ0 1 2 2(2 γ) − 2 γ κ κ0 − (θ00) − = dθ0 dθ00 log (θ00) . (S78) κs κs θ0(θ0 θ00)     Z2 Z1 −

This can then be evaluated. The θ00 integral leads to a variety of different terms, that need to be treated separately. Some variable transformations need to be performed, and some special functions need to be expanded to their series representation. It can be shown that the integral to leading order is constant in N, implying that the logarithm term 2(γ 3) of A˜ ,4 scales as κs − . The other two terms in expression (S77) are easier to evaluate: − 2 γ 2 γ 1 2 γ b − + π − dθ0dθ00 (θ0 θ00) − = B b [3 γ, γ 2] B 1 [3 γ, γ 2] 1 2 θ0θ00 − γ 2 − π − − − − − ZZD4 −   2 γ 2b 2(2 γ) b − ln 2 κ − + − π 0 (S79) γ 2 ∼ κ −   s  2b π 2 1 1 2 log 2 π 2 log b 1 κs dθ0dθ00 (θ0 θ00)− = − − 2 . (S80) θ0θ00 − b − π ∼ κ0 ZZD4   γ 3 γ 3 Plugging this back in we find that also the integral over the region scales as N − (ln N) − . D4 Thus, we can finally conclude that for β = 1, the clustering coefficient must scale as

γ 3 γ 3 N − (log N) − 2 C γ 3 γ 1 = (log N)− . (S81) ∼ N − (log N) −

With this we have found the critical exponent η/ν = 2. 15

D. Exponent η

In this section we show that the scaling exponent η that encodes how the clustering approaches zero when β + → βc = 1. As this only requires working on the low temperature side of the transition, we can directly work in the thermodynamic limit. To this end, we denote the general definition of the clustering coefficient with hidden degree κ and (without loss of generality) spacial coordinate r = 0

∞ ∞ ∞ ∞ dκ0 dκ00 dr0 dr00ρ(κ0)ρ(κ00)F (κ, κ0, r0 )F (κ, κ00, r00 )F (κ0, κ00, r0 r00 ) κ κ | | | | | − | 0 0 −∞ −∞ C(κ) = R R R R 2 . (S82) ∞ ∞ dκ0 dr0ρ(κ0)F (κ, κ0, r0 ) κ0 | | ! R −∞R where we can use connection probability (S8) andµ ˆ (S9). Let us first turn to the denominator:

∞ dr0 dκ0ρ(κ0) β = κ, (S83) Z Z 1 + r0 −∞ κκ0µˆ   γ 1 where we have plugged in the definition ofµ ˆ and used that k = γ−2 κ0. h i −

The next step is the numerator. We first perform the transformation t = r0/(κκ0µˆ) and τ = r00/(κκ00µˆ) to obtain

2 1 γ µˆ 2 2γ 2 (κ0κ00) − 1 1 − C(κ) = (γ 1) κ0 dκ0dκ00dtdτ β β β . (S84) 4 − 1 + t 1 + τ 1 + κt κτ ZZZZ | | | | κ00 − κ0 2 2 3 We know thatµ ˆ (β 1) + ((β 1) ). This is exactly the scaling that we expect from numerical investigation for the clustering∼ coefficient.− Thus,O all− we need to prove is that at β = 1, the numerator is finite. If so, its (β 1) n − dependence must be order (1). If the full expression contained (β 1)− terms with n > 0 it would diverge at the critical point and if the dominantO term was ((β 1)n) with n > 0− the numerator would go to zero at the critical point. And indeed, numerical integration showsO that− at β = 1 the numerator is finite, leading to the conclusion that

C(κ) (β 1)2 (S85) ∼ − such that η = 2, which in turn implies that ν = 1. 16

II. REAL NETWORKS

As was stated in the main text, the DPG algorithm can be used to find the temperature of its embedding in the S1 model. We list here a collection of real networks and their corresponding inverse temperatures. We choose to restrict ourselves to models where the inverse temperature lies below or close to the transition point βc.

Network Names Type V E k Target C β | | | | h i CElegans-C [5] Biological - Brain 279 2287 16 0.34 1.5 Drosophila1-C [6] Biological - Brain 350 2887 16 0.25 1.1 Drosophila2-C [6] Biological - Brain 1770 8905 10 0.33 1.1 Arabidopsis-G [7] Biological - Cell 4519 10721 4.7 0.16 1.2 CElegans-G [5] Biological - Cell 3692 7650 4.2 0.11 0.77 Drosophila-G [8] Biological - Cell 8114 38909 9.6 0.12 1.1 Human1-P [9] Biological - Cell 913 7472 16 0.23 1.0 Human2-P [9] Biological - Cell 1090 9369 17 0.20 1.0 Mus-G [8] Biological - Cell 7402 16858 4.6 0.13 1.1 Rattus-G [8] Biological - Cell 2350 3484 3.0 0.22 0.74 Yeast1-P [10] Biological - Cell 1647 2518 3.1 0.10 1.2 Yeast2-P [11] Biological - Cell 1458 1948 2.7 0.14 1.5 Polblogs-H [12] Citation - Hyperlinks 1222 16714 27 0.36 1.1 Wiki-H [13] Citation - Hyperlinks 1872 15367 16 0.42 1.3 Ecological [14] Ecological - Troffic 700 6495 18 0.10 0.15 Commodities [15] Economic - Commodities 374 1090 5.8 0.22 1.2 Friends-OFF [16] Social Offline - Friends 2539 10455 8.2 0.15 1.4 Airports1 [17] Transport - Flights 1572 17214 22 0.64 1.4

TABLE S2. Properties of a selection of networks with the inverse temperature β obtained with the DPG algorithm. Only networks with β < 1.5 are shown.

[1] M. Bogu˜n´aand R. Pastor-Satorras, Phys Rev E 68, 36112 (2003). [2] P. Colomer-de Sim´onand M. Bogu˜n´a,Phys Rev E 86, 026120 (2012). [3] M. Bogu˜n´a,D. Krioukov, P. Almagro, and M. A. Serrano, Phys. Rev. Research 2, 023040 (2020). [4] Note that we use a slightly different form than the standard 2F1[a, b; c; z] for aesthetic purposes. [5] Y.-Y. Ahn, H. Jeong, and B. J. Kim, Physica A: Statistical Mechanics and its Applications 367, 531 (2006). [6] S.-y. Takemura, A. Bharioke, T. Zhao, J. A. Horne, R. D. Fetter, S. Takemura, K. Blazek, L.-A. Chang, O. Ogundeyi, M. A. Saunders, V. Shapiro, C. Sigmund, Z. Lu, G. M. Rubin, L. K. Scheffer, I. A. Meinertzhagen, D. B. Chklovskii, A. Nern, S. Vitaladevuni, P. K. Rivlin, W. T. Katz, D. J. Olbris, S. M. Plaza, and P. Winston, Nature 500, 175 (2013). [7] A. Consortium, Science 333, 601 (2011). [8] C. Stark, B.-J. Breitkreutz, T. Reguly, L. Boucher, A. Breitkreutz, and M. Tyers, Nucleic Acids Research 34, D535 (2006). [9] A. Chang, I. Schomburg, S. Placzek, L. Jeske, M. Ulbrich, M. Xiao, C. W. Sensen, and D. Schomburg, Nucleic Acids Research 43, D439 (2014). [10] H. Yu, P. Braun, M. A. Yıldırım, I. Lemmens, K. Venkatesan, J. Sahalie, T. Hirozane-Kishikawa, F. Gebreab, N. Li, N. Simonis, T. Hao, J.-F. Rual, A. Dricot, A. Vazquez, R. R. Murray, C. Simon, L. Tardivo, S. Tam, N. Svrzikapa, C. Fan, A.-S. de Smet, A. Motyl, M. E. Hudson, J. Park, X. Xin, M. E. Cusick, T. Moore, C. Boone, M. Snyder, F. P. Roth, A.-L. Barab´asi,J. Tavernier, D. E. Hill, and M. Vidal, Science 322, 104 (2008). [11] Protein network dataset, KONECT (2007). [12] L. A. Adamic and N. Glance, Proceedings of the 3rd international workshop on Link discovery , 36–43 (2005). [13] B. Heaberlin and S. DeDeo, Future Internet 8 (2016). [14] J. A. Dunne, C. C. Labandeira, and R. J. Williams, Proceedings of the Royal Society B: Biological Sciences 281, 20133280 (2014). [15] D. Grady, C. Thiemann, and D. Brockmann, Nature communications 3, 864 (2012). [16] J. Moody, Social Networks 23, 261 (2001). [17] T. Opsahl, Why anchorage is not (that) important: Binary ties and sample selection, http://web.archive.org/web/ 20080207010024/http://www.808multimedia.com/winnt/kernel.htm (2011), accessed: June 2021.