Skip to main content
Log in

Epidemic Spreading on Complex Networks as Front Propagation into an Unstable State

  • Original Article
  • Published:
Bulletin of Mathematical Biology Aims and scope Submit manuscript

Abstract

We study epidemic arrival times in meta-population disease models through the lens of front propagation into unstable states. We demonstrate that several features of invasion fronts in the PDE context are also relevant to the network case. We show that the susceptible-infected-recovered model on a network is linearly determined in the sense that the arrival times in the nonlinear system are approximated by the arrival times of the instability in the system linearized near the disease-free state. Arrival time predictions are extended to general compartmental models with a susceptible-exposed-infected-recovered model as the primary example. We then study a recent model of social epidemics where higher-order interactions lead to faster invasion speeds. For these pushed fronts, we compute corrections to the estimated arrival time in this case. Finally, we show how inhomogeneities in local infection rates lead to faster average arrival times.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8

Similar content being viewed by others

Data Availability

The datasets generated during and/or analyzed during the current study are available from the corresponding author on reasonable request.

Notes

  1. \(D_{\textrm{eff}}(\textrm{P})\) as defined in Brockmann and Helbing (2013) is defined by first computing an effective distances between connected nodes defined as \(1-\log (\textrm{P}_{mn})\). Then, for any two notes that are not connecting the effective distance is defined as the minimal sum of effective distances along all paths connecting the two nodes.

  2. In fact, this phenomena occurs more generally for spatially extended systems such as lattice dynamical systems or systems with non-local diffusion in both discrete and continuous time; see Hindes et al. (Jul 2013) and Weinberger et al. (2002) among others for examples.

References

  • Aronson DG, Weinberger HF (1978) Multidimensional nonlinear diffusion arising in population genetics. Adv Math 30(1):33–76

    Article  MathSciNet  MATH  Google Scholar 

  • Balcan D, Gonçalves B, Hu H, Ramasco JJ, Colizza V, Vespignani A (2010) Modeling the spatial spread of infectious diseases: the global epidemic and mobility computational model. J Comput Sci 1(3):132–145

    Article  Google Scholar 

  • Barrat A, Barthelemy M, Vespignani A (2008) Dynamical processes on complex networks. Cambridge University Press, Cambridge

    Book  MATH  Google Scholar 

  • Belik V, Geisel T, Brockmann D (2011) Natural human mobility patterns and spatial spread of infectious diseases. Phys Rev X 1(1):011001

    Google Scholar 

  • Berestycki H, Nadin G (2019) Asymptotic spreading for general heterogeneous Fisher-KPP type equations. Mem Am Math Soc

  • Besse C, Faye G (2021) Spreading properties for sir models on homogeneous trees. arXiv preprint arXiv:2106.04359

  • Besse C, Faye G, Roquejoffre J-M, Zhang M (2022) The logarithmic Bramson correction for Fisher-KPP equations on the lattice \({\mathbb{Z}}\)

  • Bramson M (1983) Convergence of solutions of the Kolmogorov equation to travelling waves. Mem Am Math Soc 44(285):iv+190

    MathSciNet  MATH  Google Scholar 

  • Brockmann D, Helbing D (2013) The hidden geometry of complex, network-driven contagion phenomena. Science 342(6164):1337–1342

    Article  Google Scholar 

  • Chen LM, Holzer M, Shapiro A (2018) Estimating epidemic arrival times using linear spreading theory. Chaos Interdiscip J Nonlinear Sci 28(1):013105

    Article  MATH  Google Scholar 

  • Colizza V, Barrat A, Barthélemy M, Vespignani A (2006) The role of the airline transportation network in the prediction and predictability of global epidemics. Proc Natl Acad Sci 103(7):2015–2020

    Article  MATH  Google Scholar 

  • Ebert U, van Saarloos W (2000) Front propagation into unstable states: universal algebraic convergence towards uniformly translating pulled fronts. Physics D 146(1–4):1–99

    Article  MathSciNet  MATH  Google Scholar 

  • Fu S-C, Guo J-S, Wu C-C (2016) Traveling wave solutions for a discrete diffusive epidemic model. J Nonlinear Convex Anal 17(9):1739–1751

    MathSciNet  MATH  Google Scholar 

  • Gautreau A, Barrat A, Barthélemy M (2007) Arrival time statistics in global disease spread. J Stat Mech Theory Exp 2007(09):L09001–L09001

    Article  MATH  Google Scholar 

  • Gautreau A, Barrat A, Barthélemy M (2008) Global disease spread: statistics and estimation of arrival times. J Theor Biol 251(3):509–522

    Article  MathSciNet  MATH  Google Scholar 

  • Gucwa I, Szmolyan P (2009) Geometric singular perturbation analysis of an autocatalator model. Discrete Contin Dyn Syst Ser S 2(4):783–806

    MathSciNet  MATH  Google Scholar 

  • Hadeler KP, Rothe F (1975) Travelling fronts in nonlinear diffusion equations. J Math Biol 2(3):251–263

    Article  MathSciNet  MATH  Google Scholar 

  • Hindes J, Singh S, Myers CR, Schneider DJ (2013) Epidemic fronts in complex networks with metapopulation structure. Phys Rev E 88:012809

    Article  Google Scholar 

  • Hoffman A, Holzer M (2019) Invasion fronts on graphs: the Fisher–KPP equation on homogeneous trees and Erdős–Rényi graphs. Discrete Contin Dyn Syst Ser B 24(2):671–694

    MathSciNet  MATH  Google Scholar 

  • Hufnagel L, Brockmann D, Geisel T (2004) Forecast and control of epidemics in a globalized world. Proc Natl Acad Sci 101(42):15124–15129

    Article  Google Scholar 

  • Iacopini I, Petri G, Barrat A, Latora V (2019) Simplicial models of social contagion. Nat Commun 10:1–9

    Article  Google Scholar 

  • Iannelli F, Koher A, Brockmann D, Hövel P, Sokolov IM (2017) Effective distances for epidemics spreading on complex networks. Phys Rev E 95(1):012313, 7 (2017)

  • Jamieson-Lane A, Blasius B (2020) Calculation of epidemic arrival time distributions using branching processes. Phys Rev E 102:042301

    Article  MathSciNet  Google Scholar 

  • Jamieson-Lane A, Blasius B (2020) Epidemic arrival times; theory, discussion, and limitations. arXiv preprint arXiv:2004.05557

  • Jones CK (1995) Geometric singular perturbation theory. Dyn Syst 44–118

  • Kiss IZ, Miller JC, Simon PL (2017) Mathematics of epidemics on networks, volume 46 of interdisciplinary applied mathematics. Springer, Cham. From exact to approximate models

  • Openflights.org. Airport, airline and route data. https://openflights.org/data.html. Accessed 22 July 2021

  • Pastor-Satorras R, Castellano C, Van Mieghem P, Vespignani A (2015) Epidemic processes in complex networks. Rev Mod Phys 87(3):925

    Article  MathSciNet  Google Scholar 

  • Rvachev LA, Longini IM Jr (1985) A mathematical model for the global spread of influenza. Math Biosci 75(1):1–22. With an editorial note

  • Shigesada N, Kawasaki K, Teramoto E (1986) Traveling periodic waves in heterogeneous environments. Theor Popul Biol 30(1):143–160

    Article  MathSciNet  MATH  Google Scholar 

  • Taylor D, Klimm F, Harrington HA, Kramár M, Mischaikow K, Porter MA, Mucha PJ (2015) Topological data analysis of contagion maps for examining spreading processes on networks. Nat Commun 6(1):1–11

    Google Scholar 

  • Van den Broeck W, Gioannini C, Gonçalves B, Quaggiotto M, Colizza V, Vespignani A (2011) The GLEaMviz computational tool, a publicly available software to explore realistic epidemic spreading scenarios at the global scale. BMC Infect Dis 11(1):1–14

    Article  Google Scholar 

  • van Saarloos W (2003) Front propagation into unstable states. Phys Rep 386(2–6):29–222

    Article  MATH  Google Scholar 

  • Weinberger HF, Lewis MA, Li B (2002) Analysis of linear determinacy for spread in cooperative models. J Math Biol 45(3):183–218

    Article  MathSciNet  MATH  Google Scholar 

  • Wu C-C (2017) Existence of traveling waves with the critical speed for a discrete diffusive epidemic model. J Differ Equ 262(1):272–282

    Article  MathSciNet  MATH  Google Scholar 

Download references

Acknowledgements

This project was conducted as part of a NSF sponsored REU program. All participants received support from the NSF (DMS-2007759). The authors thank the anonymous referees whose comments improved the paper.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Matt Holzer.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Appendix A Singular Perturbation Analysis of the Local Model 4.2

Appendix A Singular Perturbation Analysis of the Local Model 4.2

We consider (4.2) with the goal of motivating the approximate solution presented in (4.6). Our approach mimics the analysis of a model of an autocatalator chemical reaction model presented in Gucwa and Szmolyan (2009). We begin with the system (4.3) where we wish to track the solution to the initial value problem with initial conditions \(S(0)=1-\kappa \), \(I(0)=\kappa \) in the limit as \(\epsilon =\frac{1}{\rho }\rightarrow 0\). As mentioned in Sect. 4 this system has two slow manifolds defined as curves of equilibrium when \(\epsilon \) is set equal to zero; see (4.4). The slow manifold on the I axis is normally hyperbolic and it follows that the reduced flow on the slow manifold is, to leading order in \(\epsilon \) given by \(I'=-\beta I\) and so we obtain that after some critical time \(\Omega \) the solution of I(t) can be described as in (4.6). The second slow manifold is given by the S axis, but this manifold lacks normal hyperbolicity so we are unable to track the solution of the initial value problem using linearization.

To overcome this lack of normal hyperbolicity we use geometric desingularization techniques or “blow-up” techniques to resolve the flow when I is small. Following Gucwa and Szmolyan (2009) we will change coordinates to

$$\begin{aligned} S=\bar{S}, \ I=r \bar{I}, \ \epsilon =r\bar{\epsilon }, \ \bar{I}^2+\bar{\epsilon }^2=1, \end{aligned}$$

effectively transforming the S axis to a cylinder with polar coordinates for the I and \(\epsilon \) variables. It is often easier to study the flow in coordinate charts and we employ two distinct charts. The first is known as the rescaling chart with coordinates

$$\begin{aligned} S=S_1, \ I=r_1 I_1, \ \epsilon =r_1, \end{aligned}$$

while the second chart has coordinates

$$\begin{aligned} S=S_2, \ I=r_2, \ \epsilon =r_2\epsilon _2. \end{aligned}$$

The two charts can be related via

$$\begin{aligned} S_2=S_1, \ r_2=r_1I_1, \ \epsilon _2=\frac{1}{I_1}. \end{aligned}$$

Our goal is to track an initial condition with \(S(0)=1-\kappa \), \(I(0)=\kappa \) with \(\kappa \) small as it evolves past the non-hyperbolic S axis to the section \(\Sigma _{out}=\{ (S,I) \ | \ I=\eta \}\) for some \(\eta >0\) at which the solution can be effectively described by a fast transition to the I axis followed by a slow relaxation along the I axis until the solution converges to the origin. In contrast to Gucwa and Szmolyan (2009), our estimates here are approximate and not rigorous. We believe that the estimates presented here could be made rigorous, but we do not pursue such an analysis here.

Analysis in first chart The first chart is known as the rescaling chart where \(r_1\) is simply a proxy for \(\epsilon \). Converting (4.3) to the coordinates of the first chart we find,

$$\begin{aligned} \frac{\textrm{d}S_1}{\textrm{d}\tau }= & {} -\alpha r_1^2 S_1 I_1-S_1r_1^2I_1^2 \nonumber \\ \frac{\textrm{d}I_1}{\textrm{d}\tau }= & {} \alpha r_1S_1I_1 -\beta r_1 I_1+S_1r_1 I_1^2 \nonumber \\ \frac{\textrm{d}r_1}{\textrm{d}\tau }= & {} 0 \end{aligned}$$
(A.1)

Rescaling the independent variable to divide the vector field by \(r_1\), we find the desingularized system

$$\begin{aligned} \frac{\textrm{d}S_1}{\textrm{d}t}= & {} -\alpha r_1 S_1 I_1-S_1r_1I_1^2 \nonumber \\ \frac{\textrm{d}I_1}{\textrm{d}t}= & {} \alpha S_1I_1 -\beta I_1+S_1 I_1^2 \nonumber \\ \frac{\textrm{d}r_1}{\textrm{d}t}= & {} 0. \end{aligned}$$
(A.2)

Let \(\eta >0\) and define the section \(\Sigma _1=\{ (S_1,I_1,r_1) \ | \ I_1=\eta \} \). Suppose that we start with initial conditions \(I(0)=\kappa \) and \(S(0)=1-\kappa \) which correspond to initial conditions \(S_1(0)=1-\kappa \) and \(I_1(0)=\frac{\kappa }{\epsilon }\). We therefore require \(\kappa \) to scale smaller than \(\epsilon \) so that \(I_1(0)\) is near zero. To obtain a leading order description of the dynamics, we set \(r_1=0\) in (A.2) and approximate \(S_1(t)=1\). Then \(I_1\) obeys (to leading order in \(\epsilon \))

$$\begin{aligned} \frac{\textrm{d}I_1}{\textrm{d}t} = (\alpha -\beta )I_1 + I_1^2, \quad I_1(t)=\frac{C(\alpha -\beta )e^{(\alpha -\beta )t}}{1-Ce^{(\alpha -\beta )t}}, \ C=\frac{\kappa }{\kappa +\epsilon (\alpha -\beta )}. \end{aligned}$$

Define \(\Omega _1\) such that \(I_1(\Omega _1)=\eta \). Using the leading-order description for \(I_1(t)\), we estimate

$$\begin{aligned} \Omega _1\approx \frac{1}{\alpha -\beta } \log \left( \frac{\eta (\kappa +\epsilon (\alpha -\beta ))}{(\alpha -\beta +\eta )\kappa }\right) \end{aligned}$$

We now convert our solution to the coordinates of the second chart and proceed with tracking the solution.

Analysis in second chart Converting (4.3) to the coordinates of the second chart we find,

$$\begin{aligned} \frac{\textrm{d}S_2}{\textrm{d}\tau }= & {} -\alpha r_2^2\epsilon _2 S_2-r_2^2S_2 \nonumber \\ \frac{\textrm{d}r_2}{\textrm{d}\tau }= & {} \alpha r_2^2\epsilon _2 S_2 -\beta r_2^2 \epsilon _2+S_2r_2^2 \nonumber \\ \frac{\textrm{d}\epsilon _2}{\textrm{d}\tau }= & {} -\alpha r_2\epsilon _2^2 S_2 +\beta r_2 \epsilon _2^2-S_2r_2\epsilon _2 \end{aligned}$$
(A.3)

Rescaling the dependent variable to divide the vector field by the nonzero factor \(\alpha r_2\epsilon _2S_2-\beta r_2\epsilon _2+S_2r_2\) we obtain the desingularized system

$$\begin{aligned} \frac{\textrm{d}S_2}{\textrm{d}s}= & {} -r_2\left( \frac{1}{1-\frac{\beta \epsilon _2}{\alpha \epsilon _2S_2 +S_2}}\right) \nonumber \\ \frac{\textrm{d}r_2}{\textrm{d}s}= & {} r_2 \nonumber \\ \frac{\textrm{d}\epsilon _2}{\textrm{d}s}= & {} -\epsilon _2. \end{aligned}$$
(A.4)

Define \(\Sigma _2=\{ (S_2,r_2,\epsilon _2) \ | \ r_2=\eta \}\) with \(\eta \) defined as before and recall the initial conditions in the section \(\Sigma _1\) which correspond to \(S_2(0)=1-\kappa +\mathcal {O}(\epsilon )\), \(r_2(0)=\eta \epsilon \), \(\epsilon _2(0)=\frac{1}{\eta }\). The transition time between sections can then be evaluated explicitly, it terms of the transformed time-scale s, as \(s=-\log \epsilon \). To determine estimates for the transition time \(\tau _2\) in the \(\tau \) time-scale we note that the timescales are related by the integral

$$\begin{aligned} \tau _2=\int _0^{-\log (\epsilon )} \frac{1}{\alpha r_2\epsilon _2 S_2(\sigma )-\beta r_2\epsilon _2 +S_2(\sigma )r_2(\sigma )} d\sigma \end{aligned}$$

We will obtain an approximation to \(t_2\) by setting \(S_2(\sigma )=1\) in the integral. We are then able to integrate (recalling that \(r_2\epsilon _2=\epsilon \)) and find

$$\begin{aligned} \tau _2\approx & {} \frac{1}{\epsilon (\alpha -\beta )}\left( -\log (\epsilon )+\log \left( \frac{\epsilon (\alpha -\beta )+\epsilon \eta }{\epsilon (\alpha -\beta )+\eta }\right) \right) , \\\approx & {} \frac{1}{\epsilon (\alpha -\beta )}\left( -\log (\epsilon )+\log \left( \frac{\epsilon }{\eta }\frac{(\alpha -\beta )+\eta }{1+\frac{\epsilon }{\eta }(\alpha -\beta )}\right) \right) \\\approx & {} \frac{1}{\epsilon (\alpha -\beta )}\left( -\log (\eta )+\log \left( \frac{\alpha -\beta +\eta }{1+\frac{\epsilon }{\eta }(\alpha -\beta )}\right) \right) \end{aligned}$$

Rescaling the independent variable from \(\tau \) to t we obtain an estimate on the total transit time of the initial condition \(I(0)=\kappa \epsilon \) to \(I(t)=\eta \) as

$$\begin{aligned} \Omega\approx & {} \frac{1}{\alpha -\beta }\left( \log \left( \eta \frac{ (\kappa +\epsilon (\alpha -\beta ))}{(\alpha -\beta +\eta )\kappa }\right) -\log (\eta )+ \log \left( \frac{\alpha -\beta +\eta }{1+\frac{\epsilon }{\eta }(\alpha -\beta )}\right) \right) \\\approx & {} \frac{1}{\alpha -\beta }\left( \log \left( \frac{ (\kappa +\epsilon (\alpha -\beta ))}{\kappa }\right) +\log \left( \frac{1}{1+\frac{\epsilon }{\eta }(\alpha -\beta )}\right) \right) . \end{aligned}$$

Using \(\frac{\kappa }{\epsilon }\) small and \(\epsilon \ll 1\) we find the approximation in (4.5).

Rights and permissions

Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Armbruster, A., Holzer, M., Roselli, N. et al. Epidemic Spreading on Complex Networks as Front Propagation into an Unstable State. Bull Math Biol 85, 4 (2023). https://doi.org/10.1007/s11538-022-01110-7

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s11538-022-01110-7

Keywords

Navigation