Emergence of Geometric Turing Patterns in Complex
Emergence of Geometric Turing Patterns in Complex
(Received 12 December 2022; revised 5 May 2023; accepted 22 May 2023; published 22 June 2023)
Turing patterns, arising from the interplay between competing species of diffusive particles, have long
been an important concept for describing nonequilibrium self-organization in nature and have been
extensively investigated in many chemical and biological systems. Historically, these patterns have been
studied in extended systems and lattices. Recently, the Turing instability was found to produce topological
patterns in networks with scale-free degree distributions and the small-world property, although with an
apparent absence of geometric organization. While hints of explicitly geometric patterns in simple network
models (e.g., Watts-Strogatz) have been found, the question of the exact nature and morphology of geometric
Turing patterns in heterogeneous complex networks remained unresolved. In this work, we study the Turing
instability in the framework of geometric random graph models, where the network topology is explained
using an underlying geometric space. We demonstrate that not only can geometric patterns be observed,
their wavelength can also be estimated by studying the eigenvectors of the annealed graph Laplacian.
Finally, we show that Turing patterns can be found in geometric embeddings of real networks. These results
indicate that there is a profound connection between the function of a network and its hidden geometry, even
when the associated dynamical processes are exclusively determined by the network topology.
DOI: 10.1103/PhysRevX.13.021038 Subject Areas: Complex Systems,
Interdisciplinary Physics,
Nonlinear Dynamics
found in the context of multilayered networks, where particles, coexisting with regions of low concentration.
particles diffuse not only within network layers but also When the dynamics takes place on a metric space, these
between them [20–23]. Here, the dynamical parameters inhomogeneities give rise to geometric patterns. It is
can be chosen such that the particle concentration within important to note that the first assumption of this
a layer is homogeneous whereas the distribution between framework is that diffusion alone, without the activation-
layers is heterogeneous. Similar patterns can be found in inhibition mechanism, should lead to a homogeneous
networks displaying community structure [24], where distribution of particles.
concentrations can be chosen to be homogeneous within In continuous media, diffusion is modeled by
a community and heterogeneous between communities. It Brownian particles and, in degree regular lattices, by
has been shown that these types of patterns can also be standard random walks. Diffusion in complex networks
reproduced in empirical networks, provided their com- is, however, different. Indeed, a simple random walk
munity structure is sufficiently strong. process on a heterogeneous network leads to a steady
While hints of explicitly geometric patterns in simple state where the concentration of particles is proportional
network models were found in several works [14,25–27], to the degree of each node and, so, highly hetero-
no comprehensive study of this phenomenon has been geneous. Thus, before adding the activation-inhibition
undertaken. Most importantly, the network models dynamics, we have to carefully define diffusion on
described in these studies are not suitable to describe networks to ensure that a homogeneous concentration of
real systems. In fact, geometric domains in a classical species is a fixed point of the dynamics. In Appendix A 1
sense, typical of Turing patterns in lattices and continu- we show that this can be achieved by a continuous
ous media, have not been observed in real complex time random walk with jump rates proportional to the
networks. This is a consequence of the apparent lack current node’s degree. Then, the evolution equation for the
of geometric structure in small-world networks, causing average concentration of particles at a given node i, ui ðtÞ,
topological distances between nodes—measured by the can be written as
shortest path lengths on the graph—to collapse around
the average value, scaling logarithmically (or slower) X
dui ðtÞ N
with the system size. ¼ −ϵ Lij uj ðtÞ; ð1Þ
This paradigm has changed recently with the develop- dt j¼1
ment of network geometry [28]. Indeed, a latent metric
space with hyperbolic geometry underlying real complex where ϵ is the diffusion coefficient, N the number of nodes,
networks provides the simplest explanation to many of and Lij the Laplacian matrix defined as in Ref. [50]:
their observed topological properties, including hetero-
geneous degree distributions [29–31], clustering [30–34],
Lij ¼ kj δij − aij : ð2Þ
small-worldness [35–37], and percolation [38,39],
spectral [40], and self-similarity properties [29,41,42].
These models have been extended to growing networks In this last equation aij is the adjacency matrix of the
[43], weighted networks [44], multilayer networks network, that hereafter is assumed to describe a single
[45,46], networks with community structure [47–49], connected component.
and they are also the basis for defining a renormalization
group for complex networks [41,42]. A. Activation-inhibition dynamics
In this paper, we show that the Turing instability triggers
In the diffusion process described by Eq. (1), particles
the emergence of purely geometric patterns that become
do not interact when they meet in the same node.
evident in the latent space of real complex networks. We
However, in realistic settings, particles may undergo all
analyze this phenomenon within the framework of network
sorts of reaction processes. In particular, the Turing
geometry and find the relation between the parameters of
instability arises when two different species, U and V,
the dynamical model and the general properties of the
interact upon meeting at the same node, under an
network, such as the congruency of the network with the
activation-inhibition process. Within each node, species
underlying network space and the level of heterogeneity of
U undergoes an autocatalytic process and, simultaneously,
its degree distribution.
favors the increase of the population of species V. At the
same time, species V inhibits the growth of species U even
II. BACKGROUND
though it cannot survive without it. When the average
The Turing instability arises as a consequence of the number of particles per node is large, we can neglect
different diffusivity rates between activator and inhibitor fluctuations in the number of particles and work under the
species. Starting from a homogeneous concentration of mean field approximation. Let ui ðtÞ and vi ðtÞ be the
particles, upon small perturbations, the system evolves average number of particles at node i of species U and V,
towards a stable state with regions of high concentration of respectively. Their evolution equations are given by
021038-2
EMERGENCE OF GEOMETRIC TURING PATTERNS IN COMPLEX … PHYS. REV. X 13, 021038 (2023)
dui ðtÞ XN If we combine this with the fact that each component is
¼ f½ui ðtÞ; vi ðtÞ − ϵ Lij uj ðtÞ; ð3Þ associated to a given node of the network, we find that the
dt j¼1 dynamics will necessarily evolve toward a state with some
of the nodes containing a high concentration of species U
dvi ðtÞ XN
and the rest a low concentration.
¼ g½ui ðtÞ; vi ðtÞ − σϵ Lij vj ðtÞ; ð4Þ
dt In Ref. [11] it was shown that there exist unstable
j¼1
eigenvalues (i.e., λα > 0) whenever
where functions fðu; vÞ and gðu; vÞ represent the activa- 2 sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi3
tion-inhibition dynamics within the nodes and ϵ and σϵ are
f g 6f g f g 7
the activator and inhibitor diffusion coefficients, respec- σ ≥ σ c ¼ v 2 u 4 u v − 2 − 2 1 − u v 5: ð7Þ
tively. We further assume that the system has a stable f u f v gu f v gu
homogeneous stationary state, that is, usti ¼ ū and vsti ¼ v̄,
for which fðū; v̄Þ ¼ 0 and gðū; v̄Þ ¼ 0, so that ðū; v̄Þ is a
In this regime, all eigenvalues that lie within the interval
stable fixed point of the dynamics. The activation-
Λ ∈ ½Λ− ; Λþ , with
inhibition dynamics and its stability condition impose
constraints on functions fðu; vÞ and gðu; vÞ. In particular, qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
its first derivatives f u ≡ ∂u fðū; v̄Þ, f v ≡ ∂v fðū; v̄Þ, gu ≡ 1
Λ ¼ σf u þ gv ðσf u − gv Þ2 þ 4σf v gu ; ð8Þ
∂u gðū; v̄Þ, and gv ≡ ∂v gðū; v̄Þ must satisfy f u > 0, f v < 0, 2σϵ
gu > 0, gv < 0, and f u þ gv < 0, f u gv − f v gu > 0 (see
Ref. [11] for further details). become unstable, with the most unstable one being
The Turing instability arises when, due to the differences
rffiffiffiffiffiffiffiffiffiffiffiffiffi
in the diffusion coefficients of the two species, the fixed 1 −f v gu
point becomes unstable so that any small perturbation Λmax ¼ f u − gv − ð1 þ σÞ : ð9Þ
1−σ σ
drives the system away from it. Interestingly, the conditions
for this to take place depend only on f u , f v , gu , gv and the
The functional dependence of Eqs. (7) and (8) implies
ratio between the diffusion coefficients of the two species
that, for a given set of parameters f u , f v , gu , gv and for one
σ. Similarly to the case of continuous media, where the
particular eigenvalue Λα , it is always possible to select
perturbation around the fixed point can be expanded in
parameters ϵ and σ such that Λα ∈ ½Λ− ; Λþ . In the case that
Fourier modes—the eigenfunctions of the Laplacian ∇2 —
Λþ ≈ Λ− ≈ Λα , one can assume that the differences in the
and where the eigenvalues are related to the wave numbers,
concentration of species in the nodes are encoded in the
here the perturbation can be expanded in terms of the
eigenvector ϕ ⃗ α.
eigenvectors of the graph Laplacian Lij [11],
Interestingly, these results depend only on the first
X
N derivatives of functions fðu; vÞ and gðu; vÞ coupled to
ui ðtÞ − ū ≈ cα eλα t ϕαi ; ð5Þ the diffusion coefficients. Therefore, there exists a whole
α¼1 class of systems where such instability may emerge, from
chemical reactions [51–53] or biological morphogenesis
where cα and λα are some constants that depend on the [3,54] to ecosystems [55–58] and game theory applied to
parameters of the model Eqs. (3) and (4) and on the ecological systems [59,60]. In ecology, one such model is
eigenvalue Λα (similarly for species V). This expansion the Mimura-Murray model [56]. Here, u and v represent
is possible because the eigenvectors of the Laplacian prey and predator densities, respectively, and the dynamics
matrix fϕ ⃗ α ¼ ðϕα ; ϕα ; …; ϕα Þ; α ¼ 1; …; Ng form a com- is governed by the functions
1 2 N
plete orthonormal basis of RN . If λα < 0; ∀ Λα , then
perturbations around the fixed point are absorbed expo- a þ bu − u2
nentially fast, so that the fixed point is stable. However, if fðu; vÞ ¼ − v u; ð10Þ
c
there exists at least one Λα such that λα > 0, then the fixed
point becomes unstable. In this case, those nodes with
cα ϕαi > 0 will increase the concentration of species U, gðu; vÞ ¼ ½u − ð1 þ dvÞv: ð11Þ
whereas those with negative value will decrease it. Note
that for all eigenvectors associated to nonvanishing eigen- From Eqs. (3), (4), (10), and (11), we see that in the absence
values we have of preys, predators go extinct and in the absence of
predators, preys attain a constant population. In general,
X
N there is a fixed point with positive densities of preys and
ϕαi ¼ 0: ð6Þ predators. In this paper, we use this dynamics as a case
i¼1 study (technical details are given in Appendix A 3).
021038-3
JASPER VAN DER KOLK et al. PHYS. REV. X 13, 021038 (2023)
B. Geometric description of complex networks: choice, the model becomes small-world whenever
The S1 =H2 model 1 < β ≤ 2, or 2 < γ ≤ 3, or both. Additionally, degree
As we have seen, the shape of emerging patterns is heterogeneity is modulated by the exponent γ and the limit
dictated by the arrangements of the signs in the components γ → ∞ is equivalent to a homogeneous distribution
of the eigenvectors of the Laplacian matrix. In continuous of hidden degrees ρðκÞ ¼ δðκ − hkiÞ and a Poisson degree
media or regular lattices, these signs are arranged in the distribution.
metric space in an alternating fashion, so that when the As it is defined, the S1 model represents the only
parameters of the dynamics are tuned to select these maximally random ensemble of geometric graphs that are
eigenvectors as unstable, the concentration of species simultaneously sparse, clustered, small-world, with
also follows the same geometric pattern. In random graphs heterogeneous degree distribution, and with only struc-
models, like the configuration model [61–64] or the tural degree-degree correlations [67]. In the regime β > 1,
Barábasi-Albert model [65], the situation is different. For the connection probability does not depend on the system
these type of models, there is no associated metric space size and, therefore, it generates networks with finite
and the most distinctive feature of nodes is the degree. In clustering in the thermodynamic limit. In this paper, we
this case, the Turing instability leads to patterns that are restrict ourselves to this regime of temperatures. However,
purely topological, with high or low concentration of the model has been extended to the case β < 1, showing
species depending on nodes’ degrees [11]. Real complex interesting quasigeometric properties in the neighborhood
networks, however, are better described by geometric of β ≲ 1 even though clustering vanishes for β ≤ 1 in the
random graph models. In such models, nodes are embedded thermodynamic limit [34]. It is worth mentioning that the
in a metric space and the connection probability depends model is defined in arbitrary dimensions. However, as
on the distances among nodes in this metric space. This shown in Ref. [70], most real networks are well described
approach has led to the development of the field of network by low dimensional spaces. Thus, to keep the analysis
geometry [28], giving rise to the most comprehensive simple, we restrict ourselves to the one-dimensional case.
description of real complex networks. The S1 model is isomorphic to a purely geometric model
To generate geometric networks, we use the S1 =H2 in the hyperbolic plane of constant negative curvature −1,
model [29,30,66], otherwise known as the geometric soft the so-called H2 [30] model. Upon mapping the hidden
configuration model [67]. The model combines a similarity degree κ to a radial coordinate r as
dimension, encoded in a one-dimensional sphere, and a
κ
popularity dimension, quantified by nodes’ degrees. In r ¼ RH2 − 2 ln ; ð13Þ
particular, each node is given a pair of hidden variables κ0
κi ∈ ½κ 0 ; ∞Þ and θi ∈ ½0; 2πÞ. The former accounts for the
ith node’s ensemble average degree, and can be generated the connection probability Eq. (12) becomes
from an arbitrary distribution ρðκÞ. The latter is the 1
angular position of node i on the circle, as an abstraction pij ¼ ðβ=2ÞðdH2 ;ij −RH2 Þ
; ð14Þ
of its position in the similarity space. In the simplest 1þe
version of the model, κ and θ are statistically independent
random variables and θ is homogeneously distributed. where RH2 ¼ 2 ln ðN=μπκ 20 Þ is the radius of a hyperbolic
This final assumption, however, is not critical to our disk and dH2 ;ij is the hyperbolic distance between nodes i
model and, in fact, in real systems the angular distribution and j given by the hyperbolic law of cosines:
is inhomogeneous, inducing the emergence of geometric
communities [47–49,68,69]. The model is fully deter- cosh ðdH2 ;ij Þ ¼ cosh ri cosh rj − sinh ri sinh rj cos Δθij ;
mined by the following connection probability between ð15Þ
two nodes i and j:
where Δθij is the angular separation between the nodes.
1
pij ¼ β ; ð12Þ Because of this isomorphism, we are free to work with one
1þ
xij version or the other indistinctly. In this paper, we use the S1
μκ i κj
model to perform the analytical calculations and the H2
model for visualization.
where xij is the distance between the nodes in the circle of
radius R ¼ N=2π, β > 1 is the inverse of the temperature
of the ensemble [30,34,67], and μ ¼ ðβ=2πhkiÞ sinðπ=βÞ a III. MIMURA-MURRAY DYNAMICS
parameter fixing the average degree hki of the network. ON GEOMETRIC NETWORKS
While the model is defined for arbitrary distributions of To illustrate the role of geometry on the Turing insta-
hidden degrees, here we choose a power law distribution bility, we run the Mimura-Murray dynamics on a real
of the form ρðκÞ ¼ ðγ − 1Þκ γ−1 −γ
0 κ , with κ ≥ κ 0 . With this network, namely that of the connectome of a mouse
021038-4
EMERGENCE OF GEOMETRIC TURING PATTERNS IN COMPLEX … PHYS. REV. X 13, 021038 (2023)
021038-5
JASPER VAN DER KOLK et al. PHYS. REV. X 13, 021038 (2023)
the network ensemble defined by the S1 =H2 model for a where F stands for Fourier transform. Note that μ−1 ¼
given sequence of hidden variables ðκ i ; θi Þ, i ¼ 1; …; N. hkiΨ̂ð0Þ. From Eq. (21), it is easy to see that the integral
Unfortunately, for quenched networks it is not possible to over the spatial coordinate of nontrivial eigenvectors must
get any analytical insight on the structure of the eigen- vanish, that is, ϕ̂ðκ; 0Þ ¼ 0, similarly to the case of the
vectors of the Laplacian matrix. To overcome this problem, quenched Laplacian matrix Eq. (6). Indeed, by setting
we use the annealed approximation, which has been
ω ¼ 0 in Eq. (21), we conclude that ϕ̂ðκ; 0Þ ¼ 0 is the only
extensively used in the literature to tackle a wide variety
solution when Λ ≠ 0. Thus, the annealed approximation
of problems, from neural networks and opinion dynamics
preserves the basic property of the Laplacian matrix given
to epidemic spreading [73–81]. In this approach, the
in Eq. (6).
network structure is resampled from the ensemble at a
rate faster than the diffusion dynamics. This allows us to
B. Homogeneous ensemble
replace the adjacency matrix aij by its ensemble average,
the connection probability pij, and consider the network The analytic solution of Eq. (21) can be found in
not as a quenched system but as a dynamic one. With this particular cases. One such case is a homogeneous ensemble
approximation, the eigenvalue problem of the annealed where all nodes have the same hidden degree, that is,
Laplacian matrix can be written as ρðκÞ ¼ δðκ − hkiÞ, leading to a Poisson degree distribution
X of average hki. In this situation, the eigenvector is just a
X ϕðκ j ; θj Þ 1 function of the frequency ω and Eq. (21) becomes
β ¼ β − Λ ϕðκi ; θi Þ; ð18Þ
xij xij
j≠i 1 þ μκ κ j≠i 1 þ μκ κ
i j i j Ψ̂ðμhki2 ωÞ hki − Λ
− ϕ̂ðωÞ ¼ 0: ð23Þ
where Λ is the eigenvalue and ϕðκ i ; θi Þ the component of Ψ̂ð0Þ hki
the corresponding eigenvector of a node i with hidden
variables ðκi ; θi Þ. Note that, as in the case of the true Assuming that ϕ̂ðωÞ ≠ 0, this equation defines a dispersion
Laplacian matrix, Λ ¼ 0 is an eigenvalue with constant relation between the eigenvalue Λ and its characteristic
eigenvector. frequency ωc as a solution of the transcendent equation:
In the thermodynamic limit, the curvature of the circle
goes to zero and, thus, nodes become distributed in R1 Λðωc Þ Ψ̂ðμhki2 ωc Þ
¼ 1− : ð24Þ
according to a Poisson point process of density one. In this hki Ψ̂ð0Þ
limit, the distance between two nodes can be evaluated as
xij ¼ jxi − xj j, where xi and xj are the positions of nodes i The characteristic frequency ωc in Eq. (24) is constrained
and j in R1 . Finally, we take the continuum limit in Eq. (18) by the boundary conditions and the discretization of nodes
by replacing the sum over index j by a double integral over in the space. Indeed, since nodes are distributed on the line
nodes coordinates ðκ; xÞ. In this way, Eq. (18) can be at density 1, frequencies above π (and so wavelengths of the
written as the following integral equation: order 1) cannot be detected. Additionally, if the system is
Z ∞ Z ∞ finite of length L ¼ N, frequencies below 2π=N will have
ϕðκ 0 ;x0 Þ
ρðκ 0 Þdκ0 dx0 β ¼ ðκ − ΛÞϕðκ; xÞ: ð19Þ associated wavelengths comparable to the system size and,
κ0 −∞ jx−x0 j
1 þ μκκ0 so, will not be detected either. Therefore, we look for
solutions of Eq. (24) in the domain ωc ∈ ½2π=N; π.
The spatial part of Eq. (19) has the form of a convolution It is illustrative to analyze in detail the homogeneous
integral. Thus, we can take advantage of the convolution case with β ¼ ∞, which corresponds to a connection
theorem for Fourier transforms. By defining the Fourier probability being a step function. In this case, function
transform of the eigenvector as Ψ̂ðzÞ takes the form
Z ∞
ϕ̂ðκ; ωÞ ≡ e−iωx ϕðκ; xÞdx; ð20Þ sin z
Ψ̂ðzÞ ¼ 2 ; ð25Þ
−∞ z
Eq. (19) can be written in Fourier space as and μ ¼ 1=ð2hkiÞ. Combining these results, the dispersion
Z relation defining ωc becomes
∞ κ 0 ρðκ 0 Þ Ψ̂ðμκκ 0 ωÞ κ−Λ
ϕ̂ðκ 0 ; ωÞdκ 0 ¼ ϕ̂ðκ; ωÞ; ð21Þ
κ0 hki Ψ̂ð0Þ κ Λðωc Þ 2 sin hkiωc
2
¼ 1− : ð26Þ
hki hkiωc
where we define Ψ̂ as
In the case of β ¼ ∞, nodes connect to nearest neighbors
1
Ψ̂ðzÞ ≡ F ðzÞ; ð22Þ on the line that are below a certain critical distance, so that
1 þ jxjβ the average number of such neighbors is hki. Therefore,
021038-6
EMERGENCE OF GEOMETRIC TURING PATTERNS IN COMPLEX … PHYS. REV. X 13, 021038 (2023)
FIG. 3. (a)–(c),(g)–(i) Examples of three eigenvectors of the quenched Laplacian matrix with low, medium, and high associated
frequencies for a single network generated by the homogeneous ensemble with N ¼ 1000 and hki ¼ 20 and β ¼ ∞ (a)–(c) and β ¼ 1.5
(g)–(i). (d)–(f),(j)–(l) Concentration of species U in the Mimura-Murray dynamics for the eigenvalues in the column to the left of
each panel.
021038-7
JASPER VAN DER KOLK et al. PHYS. REV. X 13, 021038 (2023)
The fact that ðβ − 1Þ−1 < 1=2 implies that, in this regime, a
certain eigenvalue leads to a much lower frequency in the
case of small β’s than it does for high β’s, an observation
that is corroborated by Fig. 4(b).
Figures 3(g)–3(l) show the same analysis as in
Figs. 3(a)–3(f) but for the case β ¼ 1.5, so that networks
are deep into the small-world regime, with many long-
range connections among distant nodes in the metric space.
We observe the same qualitative behavior as in the case
β ¼ ∞, except that, as expected, results have more noise.
FIG. 4. (a) Rescaled eigenvalue as a function of the character-
istic frequency in log-log scale. (b) Characteristic frequency as a
Yet, we can clearly identify geometric patterns both in the
function of the rescaled eigenvalue. In both cases hki ¼ 10 and eigenvectors and in the steady state of the Mimura-Murray
several β’s are investigated. dynamics when the dynamic parameters are properly tuned.
However, we notice that, in agreement with our theoretical
prediction in Eq. (29), eigenvalues with similar associated
matrix corresponding to low, medium, and high frequencies eigenvector frequencies as those in Figs. 3(a) and 3(d) are
of a graph generated by the homogeneous ensemble. The significantly larger.
periodic pattern is very clear in the case of low eigenvalue
with very low frequency. In the medium frequency case, we
can still identify a periodic behavior, although distorted by C. Heterogeneous ensemble
noise. In the case of high frequency, periodic behavior is In real networks, the degree distribution is typically
totally absent. We also run the Mimura-Murray dynamics heterogeneous. Therefore, we have to solve the eigenvalue
on the same network selecting parameters’ values such that problem in Eq. (21) for a heterogeneous distribution of
the same eigenvalues used in Figs. 3(a)–3(c) become hidden degrees ρðκÞ. In particular, we choose a scale-free
unstable (see Appendix A 3 for technical details of the distribution ρðκÞ ∝ κ −γ . For this distribution of hidden
dynamics). Figures 3(d)–3(f) show the concentration of degrees, Eq. (21) can be rewritten as
species U in the three cases. We observe that the patterns in Z
the concentration of species U closely follow the patterns of 1 z0γ−3 ω̃
ðγ − 2Þ Ω̂ Θ̂ðz0 ; ω̃Þdz0 ¼ Θ̂ðz; ω̃Þ; ð30Þ
the corresponding eigenvectors, justifying the relevance 0 1 − Λ̃z 0 zz0
of the structure of eigenvectors for dynamical processes on
networks. where we redefine variables ðκ; Λ; ωÞ as
Based on these observations, we conclude that the
annealed approximation provides good results in quenched κ0 Λ
z≡ ; Λ̃ ≡ ; ω̃ ≡ μκ 20 ω; ð31Þ
networks for small eigenvalues such that Λ ≪ hki, which κ κ0
corresponds to low frequencies, that is, μhki2 ωc ≪ 1. For
arbitrary values of β, we can then take the low frequency and functions as
limit in Eq. (24) to derive the relation between ωc and Λ.
Using the definition of the Fourier transform, it is easy to Ψ̂ðxÞ κ 0 ω̃
Ω̂ðxÞ ≡ ; Θ̂ðz; ω̃Þ ≡ ð1 − Λ̃zÞϕ̂ ; : ð32Þ
see that for small z, Ψ̂ð0Þ z μκ 20
Ψ̂ðzÞ zβ−1 β≤3 When Λ̃ < 1, the singularity of the kernel in the integral of
1− ∼ ð28Þ Eq. (30) falls outside the domain of integration. In this case,
Ψ̂ð0Þ z2 β > 3:
Eq. (30) can be solved numerically by Gaussian quadrature
as the solution of the system of equations,
Figure 4(a) shows the low frequency limit of the dispersion
relation computed numerically for different values of β,
Xn
wj zγ−3
j ω̃
which corroborates this scaling behavior. Using this result, ðγ − 2Þ Ω̂ Θ̂ðzj ; ω̃Þ ¼ Θ̂ðzi ; ω̃Þ; ð33Þ
the characteristic frequency scales as j¼1 1 − Λ̃zj z i zj
8 1=ðβ−1Þ
>
> 1 Λ where zi , with i ¼ 1; …; n, are the zeros of the orthogonal
>
> β≤3
>
< hki hki polynomials used in the quadrature, wi their associated
ωc ∼ sffiffiffiffiffiffiffi ð29Þ weights, and n ≫ 1 the number of points within the domain
>
> 1 Λ of integration used to evaluate the integral.
>
> β > 3:
>
: hki hki Equation (33) defines a homogeneous system of linear
equations with kernel matrix KðΛ; ωÞ given by
021038-8
EMERGENCE OF GEOMETRIC TURING PATTERNS IN COMPLEX … PHYS. REV. X 13, 021038 (2023)
wj zγ−3
j ω̃
K ij ≡ ðγ − 2Þ Ω̂ − δij : ð34Þ
1 − Λ̃zj zi zj
021038-9
JASPER VAN DER KOLK et al. PHYS. REV. X 13, 021038 (2023)
FIG. 7. (a),(c),(e),(g) Spectrum of an ensemble of networks generated by the homogeneous (a),(c) and heterogeneous (e),(g)
ensembles with β ¼ 1.5 (a),(e) and β ¼ 1.5 (c),(g). For the heterogeneous ensemble γ ¼ 2.5 is used. In all cases the average degree is
given by hki ¼ 12. (b),(d),(f),(h) Dispersion relation obtained by applying the fast Fourier transform to the eigenvectors and detecting
the frequency with highest contribution. The opacity of the points is proportional to total amount of eigenvectors with the frequency
corresponding to the point. The more transparent a point is, the fewer eigenvectors have that frequency. The black dashed lines are the
theoretical predictions given by the annealed approximation. The parameter r̂ gives the percentage of eigenvectors classified as periodic
by the procedure described in Appendix A 2. The notation implies that the results are significant with p ¼ 0.01.
combined with the DFT—using a confidence level of 2σ— observable frequencies is small. Third, as shown in
is able to detect only between 8% and 12% of periodic Ref. [70], some real networks may be better represented
eigenvalues. in similarity spaces with dimension higher than D ¼ 1.
Thus, a geometric pattern in a high dimensional space can
V. REAL NETWORKS be distorted in a lower dimensional projection, making its
detection difficult. Finally, the angular distribution in real
To analyze the emergence of Turing patterns in real networks is not perfectly uniform, with significant fluctua-
networks, we first need to find an embedding of the tions defining geometric communities. As shown in
network under study in the hyperbolic plane. This amounts Ref. [24], these communities have an impact on the
to finding, for each node of the network, its hidden degree κ eigenvector structure of the Laplacian matrix, so that in
and angular coordinate θ. For this task, we use Mercator this case we should expect patterns with mixed effects from
[83], a tool which combines machine learning and maxi- both the geometry of the network and its community
mum likelihood methods to find optimal embeddings structure. Despite all these limitations, we expect to find
congruent with the S1 =H2 model. We then run the patterns in real networks associated to small nonvanishing
Mimura-Murray dynamics on the network and look for eigenvalues.
geometric patterns in the angular similarity space inferred Figure 8 shows results for four different real networks
by Mercator. from different domains. Detailed definitions of these
Of course, in the case of real networks, we expect less networks as well as those of several others which are
clean results as compared to synthetic graphs generated by analyzed in Supplemental Material [82] can be found in
the S1 =H2 model. Indeed, there are four main factors Appendix A 5. Figures 8(a)–8(d) show the structure of the
affecting the emergence of Turing patterns in real networks. eigenvector corresponding to a small eigenvalue (and
First, the embedding method is noisy, so that periodic lower frequency) for the four studied networks. The
patterns will also be noisy. Second, even if the embedding geometric patterns in the similarity space found by
is perfect, the degree distribution is generally hetero- Mercator are very clear, with a wavelength of the order
geneous so that, as in the S1 =H2 model, the range of of the system size. The same structure is translated into
021038-10
EMERGENCE OF GEOMETRIC TURING PATTERNS IN COMPLEX … PHYS. REV. X 13, 021038 (2023)
FIG. 8. (a)–(d) Eigenvectors ϕðθÞ corresponding to the eigenvalue Λ as denoted in the figure for the real networks analyzed in
this study, namely, FriendsOFF, an off-line friendship network; WTW 2013, the World-Trade-Web; CElegans-G, a network of
genetic interaction of the nematode C. elegans; and Malaria-G, a network of highly variable genes of the human malaria parasite.
(e)–(h) Concentration of species U for the Mimura-Murray dynamics setting the parameters to make these eigenvalues unstable. The red
dotted line represents the trend obtained using the Savitzky-Golay filter with window size π. Panels (i)–(l) show the same as (e)–(h) but
in the hyperbolic representation of the four networks.
021038-11
JASPER VAN DER KOLK et al. PHYS. REV. X 13, 021038 (2023)
021038-12
EMERGENCE OF GEOMETRIC TURING PATTERNS IN COMPLEX … PHYS. REV. X 13, 021038 (2023)
ζ i usti For each network, we repeat this procedure for all its
¼ const: ðA4Þ eigenvectors, leaving us with a set of pairs ðΛ; ωc Þ of
ki
statistically significant characteristic frequencies, from
Therefore, if particles diffuse at the same rate in all where we can derive the dispersion relation as an average
nodes—so with ζ i ¼ ζ; ∀ i—then the average concen- over many network realizations.
tration at node i is proportional to ki and only when
ζi ∝ ki the average concentration of particles is the same 3. Dynamics on the network
everywhere. We then make the choice ζ i ¼ ϵki and the
diffusion equation finally reads In this paper we use the Mimura-Murray model to perform
numerical simulations. In particular, as in Ref. [11], we
dui ðtÞ XN choose the parameters as a ¼ 35, b ¼ 16, c ¼ 9, and
¼ −ϵ Lij uj ðtÞ: ðA5Þ d ¼ 2=5, which gives the fixed point ðū; v̄Þ ¼ ð5; 10Þ. In
dt j¼1 this case, the critical value of σ above which the Turing
instability arises is σ c ¼ 15.5071. We then set the value to
Similarly to the case of diffusion in continuous media, σ ¼ 15.6, slightly above the critical point. Unstable eigen-
we call ϵ the diffusion coefficient of the species in the values will then lie within the interval ½Λ− ; Λþ with
network, even though, unlike diffusion in the con-
tinuum, it has dimensions of inverse of time.
1.410 26 1.666 67
Λ− ¼ and Λþ ¼ ; ðA9Þ
ϵ ϵ
2. Dispersion relation
To compute the dispersion relation in quenched net- and the most unstable one given by
works, we generate a large number of them from the S1 =H2
model for a given set of parameters β, γ, and hki. For each
such network, we compute its eigenvectors and eigenvalues 1.533 25
Λmax ¼ : ðA10Þ
sorted in increasing order. We then compute the discrete ϵ
Fourier transform of the eigenvectors and measure the
characteristic frequency as the one corresponding to the To select a given eigenvalue of interest Λ , we set it to
highest peak in the Fourier spectrum. To asses whether a Λ ¼ Λmax and use Eq. (A10) to fix the value of ϵ. With
peak in the spectrum is statistically significant, we define a all the parameters of the dynamics fixed, we set every
null model where the entries of the components of the node’s initial condition to be the stationary one, that is,
eigenvector are independently drawn from a normal dis- ðui ð0Þ; vi ð0ÞÞ ¼ ðū; v̄Þ for all i except for a randomly chosen
tribution N ðμ; σ 2 Þ with μ ¼ 0 and σ 2 ¼ N −1 . The first one, j, for which ðuj ð0Þ; vj ð0ÞÞ ¼ ðū þ δ; v̄ þ δÞ, where
condition comes from Eq. (6) and the second from the fact δ ¼ 10−5 is a small perturbation. We then let the system
that eigenvectors are normalized. Denoting ϕ̂rand
i as one of evolve toward its steady state.
the N entries of the DFT of a random eigenvector from our
null model, it can be shown that it satisfies the following
distribution function: 4. Cheeger constant for the S1 =H2 model
It is worth mentioning an interesting property of the
2 −q
Probðjϕ̂rand
i j < qÞ ¼ 1 − e : ðA6Þ S1 =H2 model concerning its spectral gap, defined as the
smallest non-null eigenvalue of the Laplacian Λ2 . Given a
With this in hand, we can find the value q for which there is graph generated by the S1 model, GS1 , Cheeger’s inequal-
a probability p that at least one peak in the white noise ities state that
Fourier spectrum lies above q. We choose a probability
p ¼ 10−2 and take the corresponding value of q to be the ½hðGS1 Þ2
minimal value a peak in the Fourier spectrum of one of the ≤ Λ2 < 2hðGS1 Þ; ðA11Þ
2kc
eigenvectors of a network generated by the model needs to
cross to be considered periodic. Thus,
where hðGS1 Þ is the isoperimetric (or Cheeger) constant
Y
N of the graph [71,72]. This constant defines the optimal
p¼1− Probðjϕ̂i j2 < qÞ; ðA7Þ cut of the graph in two disjoint sets of nodes. For any
i¼1 partition A and B in two disjoint sets of sizes N A and N B ,
the Cheeger constant is defined as the ratio between the
so that number of edges connecting the two sets EAB and the size
of the smallest set, minimized over all possible bipartitions
q ¼ − ln½1 − ð1 − pÞ1=N : ðA8Þ of the graph; that is,
021038-13
JASPER VAN DER KOLK et al. PHYS. REV. X 13, 021038 (2023)
021038-14
EMERGENCE OF GEOMETRIC TURING PATTERNS IN COMPLEX … PHYS. REV. X 13, 021038 (2023)
TABLE I. Properties of the five real networks studied in (xiii) WTW-2013, Refs. [105–107]: The world trade web
Figs. 1 and 8. The parameter r̂ gives the percentage of as of 2013, where nodes represent countries and
eigenvectors classified as periodic by the procedure described edges trade relations.
in Appendix A 2. The notation implies that the results are Table I shows the network properties of these
significant with p ¼ 0.01. networks.
Network N hki β γ r̂
Airports 1226 3.9 1.0 4 7.1%
CElegans-C 279 16.4 1.5 3.3 3.2% [1] A. M. Turing, The Chemical Basis of Morphogenesis, Phil.
CElegans-G 878 7.2 2.6 2.7 5.9% Trans. R. Soc. B 237, 37 (1952).
Commodities 374 5.8 1.1 2.8 2.4% [2] A. Gierer and H. Meinhardt, A Theory of Biological
FriendsOFF 2539 8.2 1.3 9 4.4% Pattern Formation, Kybernetik 12, 30 (1972).
[3] H. Meinhardt, Models of Biological Pattern Formation
Human-C 989 36.1 2.3 8 7.5% (Academic Press, New York, 1982), 10.1016/S0070-2153
List contacts 410 13.5 2.2 7.5 6.1% (07)81001-5.
Malaria-G 307 18.3 2.9 9 3.9% [4] M. C. Cross and P. C. Hohenberg, Pattern Formation
Mouse-C 213 27.9 2.0 3.5 2.3% Outside of Equilibrium, Rev. Mod. Phys. 65, 851 (1993).
Power 4941 2.7 1.3 10 11.0% [5] A. Koch and H. Meinhardt, Biological Pattern Formation:
From Basic Mechanisms to Complex Structures, Rev.
URV-Email 1133 9.6 1.4 7 3.0%
Mod. Phys. 66, 1481 (1994).
Wikipedia 4589 46.4 1.2 2.8 1.0% [6] A. M. Zhabotinsky, M. Dolnik, and I. R. Epstein, Pattern
WTW 2013 189 5.8 1.9 2.4 5.8% Formation Arising from Wave Instability in a Simple
Reaction-Diffusion System, J. Chem. Phys. 103, 10306
(1995).
[7] M. Cross and H. Greenside, Pattern Formation and
(viii) Malaria-G, Ref. [99]: The network of recombinant Dynamics in Nonequilibrium Systems (Cambridge Uni-
antigen genes from the human malaria parasite versity Press, Cambridge, England, 2009).
Plasmodium falciparum, where nodes are var genes [8] S. Kondo and T. Miura, Reaction-Diffusion Model as a
encoding for an antigen protein expressed on the Framework for Understanding Biological Pattern For-
surface of the infected red blood cell. These var mation, Science 329, 1616 (2010).
genes are capable of interchanging bits of genetic [9] D. Walgraef, Spatio-Temporal Pattern Formation: With
Examples from Physics, Chemistry, and Materials Science
information, thus creating a vast amount of slightly
(Springer Science & Business Media, New York, 2012).
different genes that in turn create highly varying [10] H. Othmer and L. Scriven, Instability and Dynamic Pattern
antigens that are hard for the human immune system in Cellular Networks, J. Theor. Biol. 32, 507 (1971).
to detect. Two nodes are connected if they share a [11] H. Nakao and A. S. Mikhailov, Turing Patterns in
substring of nucleotides of significant length, in- Network-Organized Activator–Inhibitor Systems, Nat.
dicating that an interchange of genetic material Phys. 6, 544 (2010).
between the two nodes has taken place. [12] W. Horsthemke, K. Lam, and P. K. Moore, Network
(ix) Mouse-C, Ref. [100]: A connectome of the Topology and Turing Instabilities in Small Arrays of
mouse brain where nodes represent anatomical Diffusively Coupled Reactors, Phys. Lett. A 328, 444
(2004).
subregions of the brain and the edges the physical
[13] P. K. Moore and W. Horsthemke, Localized Patterns in
connections between them. The original network
Homogeneous Networks of Diffusively Coupled Reactors,
was directed. Physica (Amsterdam) 206D, 121 (2005).
(x) Power, Ref. [101]: A network which represents the [14] M. Asllani, J. D. Challenger, F. S. Pavone, L. Sacconi, and
power grid of the western U.S. Here, nodes are D. Fanelli, The Theory of Pattern Formation on Directed
transforms or power relay points and edges represent Networks, Nat. Commun. 5, 4517 (2014).
power lines running between these locations. [15] R. Muolo, M. Asllani, D. Fanelli, P. K. Maini, and T.
(xi) URV-Email, Ref. [102]: A network representing the Carletti, Patterns of Non-normality in Networked Systems,
exchange of emails between members of the Rovira i J. Theor. Biol. 480, 81 (2019).
Virgili University in Spain, in 2003. The nodes [16] J. Petit, B. Lauwens, D. Fanelli, and T. Carletti, Theory of
represent the members and the edges the email Turing Patterns on Time Varying Networks, Phys. Rev.
Lett. 119, 148301 (2017).
traffic between them. The original version of the
[17] R. A. V. Gorder, A Theory of Pattern Formation for
network was directed.
Reaction-Diffusion Systems on Temporal Networks, Proc.
(xii) Wikipedia, Refs. [103,104]: A network where nodes R. Soc. A 477, 20200753 (2021).10.1098/rspa.2020.0753
represent Wikipedia articles and where two nodes [18] R. Muolo, L. Gallo, V. Latora, M. Frasca, and T. Carletti,
are connected if there is a link between them. The Turing Patterns in Systems with High-Order Interactions,
original network was directed. Chaos Solitons Fractals 166, 112912 (2023).
021038-15
JASPER VAN DER KOLK et al. PHYS. REV. X 13, 021038 (2023)
[19] S. Gao, L. Chang, M. Perc, and Z. Wang, Turing Patterns in [39] N. Fountoulakis and T. Müller, Law of Large Numbers for
Simplicial Complexes, Phys. Rev. E 107, 014216 (2023). the Largest Component in a Hyperbolic Model of Complex
[20] M. Asllani, D. M. Busiello, T. Carletti, D. Fanelli, and G. Networks, Ann. Appl. Probab. 28, 607 (2018).
Planchon, Turing Patterns in Multiplex Networks, Phys. [40] M. Kiwi and D. Mitsche, Spectral Gap of Random
Rev. E 90, 042814 (2014). Hyperbolic Graphs and Related Parameters, Ann. Appl.
[21] M. Asllani, D. M. Busiello, T. Carletti, D. Fanelli, and G. Probab. 28, 941 (2018).
Planchon, Turing Instabilities on Cartesian Product Net- [41] G. García-Pérez, M. Boguñá, and M. Á. Serrano, Multi-
works, Sci. Rep. 5, 12927 (2015). scale Unfolding of Real Networks by Geometric Renorm-
[22] N. E. Kouvaris, S. Hata, and A. Díaz-Guilera, Pattern alization, Nat. Phys. 14, 583 (2018).
Formation in Multiplex Networks, Sci. Rep. 5, 10840 [42] M. Zheng, G. García-Pérez, M. Boguñá, and M. Á.
(2015). Serrano, Scaling Up Real Networks by Geometric
[23] D. M. Busiello, T. Carletti, and D. Fanelli, Homogeneous- Branching Growth, Proc. Natl. Acad. Sci. U.S.A. 118,
per-Layer Patterns in Multiplex Networks, Europhys. Lett. e2018994118 (2021).
121, 48006 (2018). [43] F. Papadopoulos, M. Kitsak, M. Á. Serrano, M. Boguñá,
[24] B. A. Siebert, C. L. Hall, J. P. Gleeson, and M. Asllani, and D. Krioukov, Popularity versus Similarity in Growing
Role of Modularity in Self-Organization Dynamics in Networks, Nature (London) 489, 537 (2012).
Biological Networks, Phys. Rev. E 102, 052306 (2020). [44] A. Allard, M. Á. Serrano, G. García-Pérez, and M.
[25] M. Asllani, T. Carletti, and D. Fanelli, Tune the Topology to Boguñá, The Geometric Nature of Weights in Real Com-
Create or Destroy Patterns, Eur. Phys. J. B 89, 260 (2016). plex Networks, Nat. Commun. 8, 14103 (2017).
[26] G. Cencetti, F. Battiston, T. Carletti, and D. Fanelli, [45] K.-K. Kleineberg, M. Boguñá, M. Á. Serrano, and F.
Generalized Patterns from Local and Non Local Reac- Papadopoulos, Hidden Geometric Correlations in Real
tions, Chaos Solitons Fractals 134, 109707 (2020). Multiplex Networks, Nat. Phys. 12, 1076 (2016).
[27] M.-T. Hütt, D. Armbruster, and A. Lesne, Predictable [46] K.-K. Kleineberg, L. Buzna, F. Papadopoulos, M. Boguñá,
Topological Sensitivity of Turing Patterns on Graphs, and M. Á. Serrano, Geometric Correlations Mitigate the
Phys. Rev. E 105, 014304 (2022). Extreme Vulnerability of Multiplex Networks against
[28] M. Boguñá, I. Bonamassa, M. D. Domenico, S. Havlin, D. Targeted Attacks, Phys. Rev. Lett. 118, 218301 (2017).
Krioukov, and M. Á. Serrano, Network Geometry, Nat. [47] K. Zuev, M. Boguñá, G. Bianconi, and D. Krioukov,
Rev. Phys. 3, 114 (2021). Emergence of Soft Communities from Geometric Prefer-
[29] M. Á. Serrano, D. Krioukov, and M. Boguñá, Self- ential Attachment, Sci. Rep. 5, 9421 (2015).
Similarity of Complex Networks and Hidden Metric [48] G. García-Pérez, M. Á. Serrano, and M. Boguñá, Soft
Spaces, Phys. Rev. Lett. 100, 078701 (2008). Communities in Similarity Space, J. Stat. Phys. 173, 775
[30] D. Krioukov, F. Papadopoulos, M. Kitsak, A. Vahdat, and (2018).
M. Boguñá, Hyperbolic Geometry of Complex Networks, [49] A. Muscoloni and C. V. Cannistraci, A Nonuniform
Phys. Rev. E 82, 036106 (2010). Popularity-Similarity Optimization (nPSO) Model to Effi-
[31] L. Gugelmann, K. Panagiotou, and U. Peter, Random ciently Generate Realistic Complex Networks with Com-
Hyperbolic Graphs: Degree Sequence and Clustering, in munities, New. J. Phys. 20, 052002 (2018).
Automata, Languages, and Programming. ICALP 2012,
[50] P. van Mieghem, Graph Spectra for Complex Networks
edited by A. Czumaj, K. Mehlhorn, A. Pitts, and R.
Wattenhofer, Lecture Notes in Computer Science Vol. 7392 (Cambridge University Press, Cambridge, England, 2010),
(Springer, Berlin, Heidelberg, 2012), 10.1007/978-3-642- 10.1017/CBO9780511921681.
31585-5_51. [51] I. Prigogine and R. Lefever, Symmetry Breaking Instabil-
[32] E. Candellero and N. Fountoulakis, Clustering and the ities in Dissipative Systems. II, J. Chem. Phys. 48, 1695
Hyperbolic Geometry of Complex Networks, Internet (1968).
Math. 12, 2 (2016). [52] V. Castets, E. Dulos, J. Boissonade, and P. De Kepper,
[33] N. Fountoulakis, P. van der Hoorn, T. Müller, and M. Experimental Evidence of a Sustained Standing Turing-
Schepers, Clustering in a Hyperbolic Model of Complex Type Nonequilibrium Chemical Pattern, Phys. Rev. Lett.
Networks, Electron. J. Pro 26, 1 (2021). 64, 2953 (1990).
[34] J. van der Kolk, M. Á. Serrano, and M. Boguñá, An [53] Q. Ouyang and H. L. Swinney, Transition from a Uniform
Anomalous Topological Phase Transition in Spatial Ran- State to Hexagonal and Striped Turing Patterns, Nature
dom Graphs, Commun. Phys. 5, 245 (2022). (London) 352, 610 (1991).
[35] M. A. Abdullah, N. Fountoulakis, and M. Bode, Typical [54] M. P. Harris, S. Williamson, J. F. Fallon, H. Meinhardt,
Distances in a Geometric Model for Complex Networks, and R. O. Prum, Molecular Evidence for an Activator–
Internet Math. 1 (2017).10.24166/im.13.2017 Inhibitor Mechanism in Development of Embryonic Feather
[36] T. Friedrich and A. Krohmer, On the Diameter of Hyper- Branching, Proc. Natl. Acad. Sci. U.S.A. 102, 11734 (2005).
bolic Random Graphs, SIAM J. Discrete Math. 32, 1314 [55] L. A. Segel and S. A. Levin, Application of Nonlinear
(2018). Stability Theory to the Study of the Effects of Diffusion on
[37] T. Müller and M. Staps, The Diameter of KPKVB Random Predator-Prey Interactions, AIP Conf. Proc. 27, 123
Graphs, Adv. Appl. Probab. 51, 358 (2019). (1976).
[38] M. Á. Serrano, D. Krioukov, and M. Boguñá, Percolation [56] M. Mimura and J. Murray, On a Diffusive Prey-Predator
in Self-Similar Networks, Phys. Rev. Lett. 106, 048701 Model which Exhibits Patchiness, J. Theor. Biol. 75, 249
(2011). (1978).
021038-16
EMERGENCE OF GEOMETRIC TURING PATTERNS IN COMPLEX … PHYS. REV. X 13, 021038 (2023)
[57] J. L. Maron and S. Harrison, Spatial Pattern Formation in [75] U. Bastolla and G. Parisi, Closing Probabilities in the
an Insect Host-Parasitoid System, Science 278, 1619 Kauffman Model: An Annealed Computation, Physica
(1997). (Amsterdam) 98D, 1 (1996).
[58] J. P. Gibert and J. D. Yeakel, Laplacian Matrices and [76] B. Luque and R. V. Solé, Phase Transitions in Random
Turing Bifurcations: Revisiting Levin 1974 and the Networks: Simple Analytic Determination of Critical
Consequences of Spatial Structure and Movement for Points, Phys. Rev. E 55, 257 (1997).
Ecological Dynamics, Theor. Ecol. 12, 265 (2019). [77] T. Rohlf and S. Bornholdt, Criticality in Random Thresh-
[59] A. K. Fahimipour, F. Zeng, M. Homer, A. Traulsen, old Networks: Annealed Approximation and Beyond,
S. A. Levin, and T. Gross, Sharp Thresholds Limit the Physica (Amsterdam) 310A, 245 (2002).
Benefit of Defector Avoidance in Cooperation on Net- [78] D. Vilone and C. Castellano, Solution of Voter Model
works, Proc. Natl. Acad. Sci. U.S.A. 119, e2120120119 Dynamics on Annealed Small-World Networks, Phys. Rev.
(2022). E 69, 016109 (2004).
[60] A. Brechtel, P. Gramlich, D. Ritterskamp, B. Drossel, and [79] M. Boguná, C. Castellano, and R. Pastor-Satorras, Lange-
T. Gross, Master Stability Functions Reveal Diffusion- vin Approach for the Dynamics of the Contact Process on
Driven Pattern Formation in Networks, Phys. Rev. E 97, Annealed Scale-Free Networks, Phys. Rev. E 79, 036110
032307 (2018). (2009).
[61] E. A. Bender and E. R. Canfield, The Asymptotic Number [80] B. Guerra and J. Gómez-Gardenes, Annealed and Mean-
of Labeled Graphs with Given Degree Sequences, J. Field Formulations of Disease Dynamics on Static and
Comb. Theory Ser. A 24, 296 (1978). Adaptive Networks, Phys. Rev. E 82, 035101(R) (2010).
[62] B. Bollobás, A Probabilistic Proof of an Asymptotic [81] S. C. Ferreira, R. S. Ferreira, and R. Pastor-Satorras,
Formula for the Number of Labelled Regular Graphs, Quasistationary Analysis of the Contact Process on
Eur. J. Combinatorics 1, 311 (1980). Annealed Scale-Free Networks, Phys. Rev. E 83, 066113
[63] M. Molloy and B. Reed, A Critical Point for Random (2011).
Graphs with a Given Degree Sequence, Random Struct. [82] See Supplemental Material at http://link.aps.org/
Algorithms 6, 161 (1995). supplemental/10.1103/PhysRevX.13.021038 for an analy-
[64] M. Molloy and B. Reed, The Size of the Giant Component sis of the periodicity of the eigenvectors of several real
of a Random Graph with a Given Degree Sequence, Comb. networks.
Probab. Comput. 7, 295 (1998). [83] G. García-Pérez, A. Allard, M. Á. Serrano, and M.
[65] A.-L. Barabási and R. Albert, Emergence of Scaling in Boguñá, Mercator: Uncovering Faithful Hyperbolic Em-
Random Networks, Science 286, 509 (1999). beddings of Complex Networks, New J. Phys. 21, 123033
[66] M. A. Serrano and M. Boguñá, The Shortest Path to (2019).
Network Geometry: A Practical Guide to Basic Models [84] P. Hagmann, L. Cammoun, X. Gigandet, R. Meuli, C. J.
and Applications, Elements in Structure and Dynamics Honey, V. J. Wedeen, and O. Sporns, Mapping the Struc-
of Complex Networks (Cambridge University Press, tural Core of Human Cerebral Cortex, PLoS Biol. 6, e159
Cambridge, England, 2022), 10.1017/9781108865791. (2008).
[67] M. Boguñá, D. Krioukov, P. Almagro, and M. Á. Serrano, [85] V. D. Blondel, J.-L. Guillaume, R. Lambiotte, and E.
Small Worlds and Clustering in Spatial Networks, Phys. Lefebvre, Fast Unfolding of Communities in Large Net-
Rev. Res. 2, 023040 (2020). works, J. Stat. Mech. (2008) P10008.
[68] M. Boguñá, F. Papadopoulos, and D. Krioukov, Sustaining [86] E. Ortiz and M. Á. Serrano, Multiscale Voter Model on Real
the Internet with Hyperbolic Mapping, Nat. Commun. 1, 1 Networks, Chaos Solitons Fractals 165, 112847 (2022).
(2010). [87] K.-K. Kleineberg, Metric Clusters in Evolutionary Games
[69] M. Á. Serrano, M. Boguñá, and F. Sagués, Uncovering the on Scale-Free Networks, Nat. Commun. 8, 1888 (2017).
Hidden Geometry Behind Metabolic Networks, Mol. Bio- [88] J. D. O’Brien, K. A. Oliveira, J. P. Gleeson, and M. Asllani,
syst. 8, 843 (2012). Hierarchical Route to the Emergence of Leader Nodes in
[70] P. Almagro, M. Boguñá, and M. Á. Serrano, Detecting the Real-World Networks, Phys. Rev. Res. 3, 023117 (2021).
Ultra Low Dimensionality of Real Networks, Nat. Com- [89] A. Allard, M. Ángeles Serrano, and M. Boguñá, Geometric
mun. 13, 6096 (2022). Description of Clustering in Directed Networks, arXiv:
[71] G. Pólya and G. Szegö, Isoperimetric Inequalities in 2302.09055.
Mathematical Physics (AM-27) (Princeton University [90] E. W. Montroll and G. H. Weiss, Random Walks on
Press, Princeton, NJ, 1951), http://www.jstor.org/stable/j Lattices. II, J. Math. Phys. 6, 167 (1965).
.ctt1b9rzzn. [91] G. H. Weiss, Aspects and Applications of the Random
[72] J. Cheeger, A Lower Bound for the Smallest Eigenvalue Walk (North-Holland, Amsterdam, 1994), 10.1007/
of the Laplacian, in Problems in Analysis, edited by BF02179402.
R. C. Gunning (Princeton University Press, Princeton, [92] J. Kunegis, KONECT: The Koblenz Network Collection, in
NJ, 1969), pp. 195–199, 10.1515/9781400869312-013. Proceedings of the 22nd International Conference on World
[73] B. Derrida and Y. Pomeau, Random Networks of Wide Web (WWW '13 Companion) (ACM, New York,
Automata: A Simple Annealed Approximation, Europhys. 2013), pp. 1343–1350, 10.1145/2487788.2488173.
Lett. 1, 45 (1986). [93] Y.-Y. Ahn, H. Jeong, and B. J. Kim, Wiring Cost in the
[74] S.-i. Amari, N. Fujita, and S. Shinomoto, Four Types of Organization of a Biological Neuronal Network, Physica
Learning Curves, Neural Comput. 4, 605 (1992). (Amsterdam) 367A, 531 (2006).
021038-17
JASPER VAN DER KOLK et al. PHYS. REV. X 13, 021038 (2023)
[94] J. S. Kim and M. Kaiser, From Caenorhabditis elegans to [101] D. J. Watts and S. H. Strogatz, Collective Dynamics of
the Human Connectome: A Specific Modular Organization ‘Small-World’ Networks, Nature (London) 393, 440 (1998).
Increases Metabolic, Functional and Developmental Effi- [102] R. Guimerà, L. Danon, A. Díaz-Guilera, F. Giralt, and
ciency, Phil. Trans. R. Soc. B 369, 20130529 (2014). A. Arenas, Self-Similar Community Structure in a Network
[95] A. Cho, J. Shin, S. Hwang, C. Kim, H. Shim, H. Kim, H. of Human Interactions, Phys. Rev. E 68, 065103(R)
Kim, and I. Lee, WormNet v3: A Network-Assisted (2003).
Hypothesis-Generating Server for, Caenorhabditis ele- [103] R. West, J. Pineau, and D. Precup, Wikispeedia: An Online
gans, Nucleic Acids Res. 42, W76 (2014). Game for Inferring Semantic Distances between Concepts,
[96] D. Grady, C. Thiemann, and D. Brockmann, Robust in Proceedings of the International Joint Conference on
Classification of Salient Links in Complex Networks, Artificial Intelligence, California, 2009.
Nat. Commun. 3, 864 (2012). [104] R. West and J. Leskovec, Human Waynding in Information
[97] J. Moody, Peer Influence Groups: Identifying Dense Networks, in Proceedings of the 21st International
Clusters in Large Networks, Soc. Networks 23, 261 confErence on World Wide Web (WWW '12) (ACM,
(2001). 2012), pp. 619–628, 10.1145/2187836.2187920.
[98] L. Isella, J. Stehlé, A. Barrat, C. Cattuto, J.-F. Pinton, and [105] M. Á. Serrano and M. Boguñá, Topology of the World
W. V. den Broeck, What’s in a Crowd? Analysis of Face- Trade Web, Phys. Rev. E 68, 015101(R) (2003).
to-Face Behavioral Networks, J. Theor. Biol. 271, 166 [106] M. Á. Serrano, M. Boguñá, and A. Vespignani, Patterns of
(2011). Dominant Flows in the World Trade Web, J. Econ.
[99] D. B. Larremore, A. Clauset, and C. O. Buckee, A Network Interact. Coord. 2, 111 (2007).
Approach to Analyzing Highly Recombinant Malaria [107] G. García-Pérez, M. Boguñá, A. Allard, and M. A.
Parasite Genes, PLoS Comput. Biol. 9, e1003268 (2013). Serrano, The Hidden Hyperbolic Geometry of Interna-
[100] S. W. Oh et al., A Mesoscale Connectome of the Mouse tional Trade: World Trade Atlas 1870–2013, Sci. Rep. 6,
Brain, Nature (London) 508, 207 (2014). 33441 (2016).
021038-18