The Speed of Epidemic Waves in a One-Dimensional Lattice of SIR Models

20
Math. Model. Nat. Phenom. Vol. ?, No. ?, 200?, pp. ?-? The Speed of Epidemic Waves in a One-Dimensional Lattice of SIR Models Igor Sazonov a1 , Mark Kelbert b and Michael B. Gravenor c a School of Engineering, Swansea University, Singleton Park, SA2 8PP, U.K. b Department of Mathematics, Swansea University b Institute of Life Science, School of Medicine, Swansea University Abstract. A one-dimensional lattice of SIR (susceptible/infected/removed) epidemic centres is considered numerically and analytically. The limiting solutions describing the behaviour of the standard SIR model with a small number of initially infected individuals are derived, and expres- sions found for the duration of an outbreak. We study a model for a weakly mixed population distributed between the interacting centres. The centres are modelled as SIR nodes with interac- tion between sites determined by a diffusion-type migration process. Under the assumption of fast migration, a one-dimensional lattice of SIR nodes is studied numerically with deterministic and random coupling, and travelling wave-like solutions are found in both cases. For weak coupling, the main part of the travelling wave is well approximated by the limiting SIR solution. Explicit formulae are found for the speed of the travelling waves and compared with results of numeri- cal simulation. Approximate formulae for the epidemic propagation speed are also derived when coupling coefficients are randomly distributed, they allow us to estimate how the average speed in random media is slowed down. Key words: spatial epidemic models; dynamic systems; travelling waves AMS subject classification: 92D30 1. Introduction Mathematical models have been long used to study the epidemiology of a wide range of infectious diseases [1]. The classical SIR (susceptible/infected/removed) model is one of the cornerstones 1 Corresponding author. E-mail: [email protected] 1

Transcript of The Speed of Epidemic Waves in a One-Dimensional Lattice of SIR Models

Math. Model. Nat. Phenom.Vol. ?, No. ?, 200?, pp. ?-?

The Speed of Epidemic Waves in a One-Dimensional Latticeof SIR Models

Igor Sazonova1, Mark Kelbertb and Michael B. Gravenorc

a School of Engineering, Swansea University, Singleton Park, SA2 8PP, U.K.b Department of Mathematics, Swansea University

b Institute of Life Science, School of Medicine, Swansea University

Abstract. A one-dimensional lattice of SIR (susceptible/infected/removed) epidemic centres isconsidered numerically and analytically. The limiting solutions describing the behaviour of thestandard SIR model with a small number of initially infected individuals are derived, and expres-sions found for the duration of an outbreak. We study a model for a weakly mixed populationdistributed between the interacting centres. The centres are modelled as SIR nodes with interac-tion between sites determined by a diffusion-type migration process. Under the assumption of fastmigration, a one-dimensional lattice of SIR nodes is studied numerically with deterministic andrandom coupling, and travelling wave-like solutions are found in both cases. For weak coupling,the main part of the travelling wave is well approximated by the limiting SIR solution. Explicitformulae are found for the speed of the travelling waves and compared with results of numeri-cal simulation. Approximate formulae for the epidemic propagation speed are also derived whencoupling coefficients are randomly distributed, they allow us to estimate how the average speed inrandom media is slowed down.

Key words: spatial epidemic models; dynamic systems; travelling wavesAMS subject classification: 92D30

1. IntroductionMathematical models have been long used to study the epidemiology of a wide range of infectiousdiseases [1]. The classical SIR (susceptible/infected/removed) model is one of the cornerstones

1Corresponding author. E-mail: [email protected]

1

I. Sazonov et al. The Speed of Epidemic Waves in a One-Dimensional Lattice of SIR Models

of mathematical epidemiology, describing the dynamic of infection in a single population [18, 5].The analysis of infection spread through linked systems of populations such as urban centres is ofgreat importance [20] and attracts considerable interest, in particular for planning the response toemerging pandemic diseases. An excellent review of a huge literature on spatial epidemic modelscan be found in [15].

The numerical simulation of spatial systems gives insight into different scenarios of epidemicspread and can be used to help plan actual interventions. However, the inclusion of large numbers(potentially many thousands) of interacting nodes can be costly in terms of integrating the SIRprocesses, and simplified models (e.g., with the epidemic processes greatly simplified within thenodes) are widely used. For example, during the recent foot and mouth disease outbreak in theUK [9] ‘real time’ modelling of the spatial spread of the epidemic was performed, on the net-work of almost all sheep/cattle/pig livestock premises in the country. Following this experience,recommendations were made for the construction, in advance, of well defined and understoodquantitative models as part of attempts to prepare for future outbreaks of emerging diseases.

A well-developed simulation technique is based on passing to the continuous media [18, 8,14, 6, 3]. Usually this limit leads to PDE of the reaction diffusion type. A well-known feature isthe existence of the so-called travelling wave solutions, that preserve speed and shape. Analogouspatterns are systematically observed in epidemic data sets [6]. However, situations are possiblewhen the continuous limit is not applicable, although similar patters can occur. Analysis of thediscrete nature of spatial interactions is one of goals our work.

Many recent publications focus on the spatial aspect, and the propagation of disease (or com-puter viruses or rumours) in networks with different topology attracts a lot of attention (see review[10] and papers cited therein, in particular, the ‘Small World’ network [17]). Usually, these mod-els (i.e. the contact processes) apply a simplified description of the epidemic within a single node(down to the simplest binary model: infected or not). As a step towards including more detailedepidemic dynamics into the nodes of these lattice-type spatial models, it is important to consider asystem of locally interacting SIR models and to study their ‘global’ behaviour.

Here, we consider a network with the simplest topology—a 1D lattice with a large number ofinteracting nodes. The dynamics within each node is described by a standard SIR model, withadditional terms accounting for the interaction between nodes. More precisely, we use an approx-imation of weak coupling and fast migration to obtain a relatively simple and explicit descriptionof epidemic propagation. We suggest that the interaction is due to migration of a relatively smallproportion of the population (including those infective) between nodes, and the interaction is pro-portional to the share of time this population spends in the neighbouring nodes. Using an approx-imation of weak coupling, we assume that the share of time to be spent in another node is small.The situation is therefore most applicable when populations are concentrated mainly in their owncommunities, and the spread of an epidemic has a distinctive time delay caused by a lack of strongmixing in the overall population.

We establish the existence of travelling wave type solutions in such systems, under determinis-tic and random coupling and determine its speed of propagation. It is shown that in an interestingparameter range it is impossible to approximate the finite differences by derivatives, i.e. to pass toreaction-diffusion equations, and a discrete spatial approach is essential for the analysis.

2

I. Sazonov et al. The Speed of Epidemic Waves in a One-Dimensional Lattice of SIR Models

The paper is organized as follows. In Section 2, the limiting solutions of a single SIR modelare introduced. They provide a good approximation for an outbreak triggered by a small number ofinitial infectives, and provide the base for describing travelling waves in the lattice of SIR nodes inSection 4. In Section 3 the migration model for spread between nodes is discussed, and equationsof interacting SIR models are described. In Section 4, travelling waves in a 1D lattice of interactingidentical SIR models are investigated. In Section 5, the study is extended for a lattice with randomcoupling. In Section 6, we study the limits of applicability of the approximation used in analysis,estimate the possible errors and consider the continuous case. Finally, in Section 7, we draw ourconclusions, and discuss the possible extensions of the model.

2. Limiting solution of the SIR modelIn the classical SIR model, the numbers of susceptibles S, infectives I and removed (immune/deaddue to disease/vaccinated) R evolve in accordance with the equations

d

dtS = −βSI,

d

dtI = βSI − αI,

d

dtR = αI (2.1)

where β is the infection rate, α the recovery rate (e.g. [18, 5]). Adding these equations andintegrating the sum, we obtain S + I + R = const = N where N is the population number.The typical evolutionary problem is to integrate (2.1) if the initial number of infectives I0 is set att = t0.

We write equations (2.1) in the dimensionless form

d

dτs = −ρsi (2.2)

d

dτi = (ρs− 1)i (2.3)

where s = S/N , i = I/N , are, respectively, the shares of susceptibles and infectives, and τ = αtis the dimensionless time. Here ρ = (β/α) N is called the basic reproduction number usuallydenoted by R0 [18, 1, 7]. The differential equation for r can be omitted as r = 1 − s − i. Theinitial conditions in these variables can be written as

i(τ0) = i0, s0 := s(τ0) = 1− i0. (2.4)

The initial number of infectives grows if ρs0 > 1 [5] and reaches its maximal value [5]

imax = 1− ρ−1 + ρ−1 ln(ρ−1s−10 ) (2.5)

at an instant which we denote as τmax. Then the outbreak time can be defined as [5]

τoutb := τmax − τ0.

In Fig. 1a, the solutions i(τ) are plotted for ρ = 4 and various values of i0. Initial instantsτ0 are chosen to achieve the maximum imax at the same instant. Observe that as i0 decreases, the

3

I. Sazonov et al. The Speed of Epidemic Waves in a One-Dimensional Lattice of SIR Models

-2 -1 0 1 20

0.1

0.2

0.3

0.4

0.5

t - t

i(t)

max

0 0.2 0.4 0.6 0.8 10

0.2

0.4

i

0.6

s

smax

a b

Figure 1: Solutions i(τ) (a) and i(s) (b) for i0 = 1/4, 1/8, 1/16, 1/32 and ρ = 4 (dashed curves:the smaller i0 the thinner is the curve). The thin solid curve indicates the limiting solution.

solutions tend to an unique limiting solution ilim(τ) starting from the point i = 0, s = 1 at theinstant τ = −∞. On the phase diagram {s, i}, the limiting solution ilim(s) is the lower boundaryof all realistic trajectories (Fig. 1b).

The limiting solutions ilim(τ) are shift invariant with respect to τ , and from now on we setτmax = 0. They then form a one parameter family plotted in Fig. 2.

-5 0 510

-8

10-6

10-4

10-2

100

t

10

9

8

7

6

543

2

ilim

Figure 2: Limiting solutions i(τ) for ρ = 2, 3, . . . , 10 in the logarithmic scale. Values of ρ areindicated in circles.

The limiting solutions have simple asymptotic behaviors

ilim = A0eλ0τ if τ → −∞ (2.6)

ilim = A∞eλ∞τ if τ → +∞ (2.7)

4

I. Sazonov et al. The Speed of Epidemic Waves in a One-Dimensional Lattice of SIR Models

where A0 and A∞ are some constants; λ0 = ρ−1 > 0, λ∞ = −1+ρslim∞ < 0 are the growth/decay

rates. Hereslim∞ := lim

τ→∞slim(τ) = −1

ρW0

(−ρe−ρ)

(2.8)

where Wk(z) is the kth branch of the Lambert function [4] (see Appendix for details).Any solution started at τ0 with small enough i0 can be approximated by the limiting curve

i(τ ; ρ, i0, τ0) ≈ ilim(τ − τ0 − τoutb; ρ) (2.9)

The maximal discrepancy between usual and limiting solutions occurs in the vicinity of maxi-mum of i and approximately equals

max |i− ilim| ≈ imax − ilimmax = − ln(s0)− 1

ρ≈ i0

ρ.

To apply approximation (2.9) it is necessary to calculate τoutb for given ρ and i0. An integralformula for the outbreak time is

τoutb =

∫ s0

smax

ds

ρ(s− s2) + s ln(s/s0)(2.10)

wheresmax := s(τmax) = ρ−1. (2.11)

This value can be easily obtained by setting di/dτ = 0 in (2.3); substituting (2.11) into (A1) weobtain (2.5).

When i0 → 0, non-decaying asymptotic terms of (2.10) are

τoutb∼= 1

λ0

ln1

i0+ C + O(i0) (2.12)

where C = C(ρ) is independent of i0. The proof, and the integral formula for C(ρ) are given inthe Appendix. The constant C is related to the constant A0 in (2.6)

C = λ−10 ln A0. (2.13)

3. Interacting SIR processesConsider the migration process between two interacting identical centres. Then the rate of changeof infectives/susceptibles will be determined by migration as well, and additional terms appear inthe equations

d

dτs1 = −ρs1i1 +

d

dτs2→1 − d

dτs1→2 (3.1)

d

dτi1 = ρs1i1 − i1 +

d

dτi2→1 − d

dτi1→2 (3.2)

d

dτs2 = −ρs2i2 +

d

dτs1→2 − d

dτs2→1 (3.3)

d

dτi2 = ρs2i2 − i2 +

d

dτi1→2 − d

dτi2→1 (3.4)

5

I. Sazonov et al. The Speed of Epidemic Waves in a One-Dimensional Lattice of SIR Models

where im→n and sm→n are, respectively, the shares of infectives/susceptibles migrated from nodem to node n.

Let the population from node 1 migrate between the two nodes, spending a fraction of time εin the second node. Suppose that at τ = 0, the share i1 of infectives is concentrated in node 1.Under the assumption that the migration is described by linear equations, the share of infectives atnode 2 can be described by the convolution with a response function g(τ) of a linear system. Forthe problem in hand, it is convenient to write this convolution in the form

i1→2(τ) =d

[i1(τ) ∗ g(τ)

]:=

d

∫ +∞

−∞i1(τ

′) g(τ−τ ′) dτ ′ (3.5)

where g(τ) = 0 if τ < 0 due to the casuality principle. Generally, i1→2(τ) converges exponentiallyto the equilibrium value εi1 with the characteristic time ∆τ , for example, from the analogy with adiffusion process it can be

g(τ) = ε(1− e−τ/∆τ

)θ(τ) (3.6)

where θ(τ) is the Heaviside unit-step function.The approximation of weak coupling considered here means that ε ¿ 1. Then im→n ¿ 1 and

sm→n ¿ 1.If those first infected appear in node 1, then the outbreak develops first in this node. Its in-

fectious individuals migrating to node 2 trigger the outbreak there with a certain delay. A nodeis sensitive to a small amount of infectives only before the outbreak, becoming almost insensitiveduring and after it. Therefore, we can neglect transport terms in (3.1)–(3.2) as they do not changethe essential epidemic dynamics in node 1. For the approximation of weak coupling, we can alsoneglect transport terms i2→1, s2→1 in (3.3)–(3.4). The small term s1→2 in (3.3) also does not impacton the solution of the equation and can be dropped. Thus, in the approximation of weak coupling(ε ¿ 1) and when the outbreak is first initiated in node 1, the only essential transport term is i1→2

in (3.4) which accounts for infectives migrating from node 1 to node 2 and triggering the outbreakthere.

Here, we also restrict our consideration to an approximation of fast migration, when the char-acteristic time of the migration process ∆τ is much smaller than 1/λ0, i.e. the characteristic timeof the epidemic growth on its linear stage

∆τ ¿ 1/λ0. (3.7)

Theng(τ) ≈ ε θ(τ) and i1→2(τ) ≈ εi1(τ). (3.8)

Then equations (3.1)–(3.4) coincide with (2.2)–(2.3) whereas equations (3.3)–(3.4) take theform

d

dτs2 = −ρs2i2 (3.9)

d

dτi2 = (ρs2 − 1)i2 + ε

d

dτi1. (3.10)

6

I. Sazonov et al. The Speed of Epidemic Waves in a One-Dimensional Lattice of SIR Models

Note that the model (3.9)–(3.10) is different from the well-known model of simple epidemics ininteracting groups (see [5], § 2.2) where interaction terms appear without the time derivative. In thefast migration approximation, if the epidemic dynamics are temporarily neglected (spread betweenindividuals and removal are switched off), equations (3.9)–(3.10) imply that i2 ≡ i1→2 = εi1, i.e.the migration provides the only source of infectives in node 2.

Analogous equations can be derived for many interacting nodes with the external forces, (d/dτ)im→n, n, m =1, 2, . . . ,m 6= n, proportional to the rate of change of the number of infectives arriving from othernodes coupled with the nth node. They can also describe networks of interacting nodes of differ-ent topology (e.g. [10, 17], etc); although in this work we consider the simplest network topologyonly.

4. One-dimensional lattice of SIR modelsConsider a lattice of interacting identical SIR nodes. Let the dynamics in the first node (n = 1)be described by (2.2)–(2.3) with initial conditions (2.4), and the dynamics in all subsequent nodes(n = 2, 3, . . .) by (3.9)–(3.10) which, assuming fast migration, take the form

d

dτsn = −ρsnin (4.1)

d

dτin = (ρsn − 1)in +

d

dτin−1→n (4.2)

with in−1→n = εin−1.Numerical integration of the ODEs (4.1)–(4.2) with different ρ, ε and i0 (see Fig. 3) shows

t

in

0 5 10 15 20 25 3010

-20

10-15

10-10

10-5

100

Figure 3: Numerical simulation of lattice with coupled SIR nodes (first 10 nodes) for ρ = 3 andfor ε = 10−3.

that the time difference Tn between outbreaks in the (n−1)th and nth nodes tends to a constantT (ρ, ε) (see Fig. 4). The period of initial growth of Tn is more clearly visible in the logarithmicalscale (see Fig. 4, curve 2). The solution in(τ), sn(τ) for n À 1 almost replicates the solution

7

I. Sazonov et al. The Speed of Epidemic Waves in a One-Dimensional Lattice of SIR Models

100 200 300 400 500 600 700 800 900 10002.1

2.11

2.12

2.13

2.14

2.15

2.16

2.17

2.18

2.19

2.210

010

110

210

3

T

n

n

1

2

n

Tmax

Figure 4: Time interval Tn between outbreaks in two neighbouring nodes versus n for ρ = 3 andε = 10−3 for linear (curve 1) and logarithmic (curve 2) n-scale. The approximation (4.11) is shownby the dashed line.

in−1(τ), sn−1(τ) in shape, with a delay T

in(τ) ∼= in−1(τ + T ), sn(τ) ∼= sn−1(τ + T ). (4.3)

Thus the numerical integration shows that for n → ∞ the solution tends to the universal onedelayed in time by nT : in(τ) → itr(τ − nT ). This indicates that a type of travelling non-linearwave propagates with the velocity c = T−1 in the discretized lattice of individual SIR nodes withunit distances between nodes, whereas the variable n plays the role of a spatial coordinate, say x.Substituting the solution of the form sn(τ) = str(τ − nT ), in(τ) = itr(τ − nT ) into (4.1)–(4.2),we obtain a system of two equations

d

dτstr = −ρstritr (4.4)

d

dτitr = (ρstr − 1)itr + ε

d

dτitr(τ + T ). (4.5)

Its solution must satisfy the initial conditions

str(−∞) = 1, itr(−∞) = 0. (4.6)

Here T = T (ρ, ε) is an eigenvalue of the problem (4.4)–(4.6). As the equations are invariant withrespect to a shift in time, we fix τ = 0 as an instant when itr is maximal.

As the solution itr decays for τ > 0, the second terms on the r.h.s. of (4.5) are essential onlyfor τ < −T . Hence, for τ > −T the solution is close to the solution for the individual nodeconsidered above. In the limit ε → 0 it tends to the limiting solution ilim in (2.9). Numericalsimulation confirms that the travelling wave and limiting solution differ only for τ < −T (see

8

I. Sazonov et al. The Speed of Epidemic Waves in a One-Dimensional Lattice of SIR Models

-T 0

i

t

10-20

10-15

10-10

10-5

100

a

- T 0

l(t)

t

l =

l =

l =

max

8

0

b

Figure 5: Traveling wave itr(τ ; ρ, ε) (solid) and limiting solution ilim(τ ; ρ) (dashed) versus timefor ρ = 3 and ε = 10−3 (a). Plot of the growth rate over time, λ(τ) = (d/dτ) ln(i) (b).

Fig. 5). For τ < −T the value s(τ) differs negligibly from 1, and the equation (4.5) can be wellapproximated by a linear one

d

dτitr = λ0i

tr + εd

dτitr(τ + T ). (4.7)

Here the solution of (4.7) has an exponential form itr = exp{λτ} where the parameter λ is relatedto the eigenvalue T of the non-linear problem

λ = λ0 + ελeλT . (4.8)

A general approach described [22, 16]) can be applied to find T , and provides the resultswell supported by numerical evidence. More precisely, in some non-linear systems, the velocity

9

I. Sazonov et al. The Speed of Epidemic Waves in a One-Dimensional Lattice of SIR Models

of a travelling wave corresponds to the lowest value taken over a range of wavenumbers for thelinearized equation. This should be true for a system in which non-linearity results in slowingdown the wave. In fact, the velocity of travelling wave is proportional to the growth rate λ0 ofinfinitesimal solution (see below). The higher the wave amplitude i, the smaller is the share ofsusceptibles s, hence the smaller is the growth rate which is λ = ρs − 1 < λ0. Therefore wefind the smallest speed of propagation (greatest T ) for the linear solution and compare it with theresults of numerical simulation.

Substituting the plane wave solution of the form

itr(τ, n) = exp (λτ − nT ) (4.9)

to (4.1)–(4.2) and neglecting non-linear terms there, we obtain the characteristic equation (cf. (4.8))

L(λ, T ) := λ− λ0 − ελeλT = 0. (4.10)

Solving the system of equations L(λ, T ) = 0, (∂/∂λ)L(λ, T ) = 0 we obtain explicit formulas forthe maximal value of T and express the correspondent value of λ in terms of the Lambert function

Tmax =1

λ0

(W0 − 1)2

W0

, λmax = λ0W0

W0 − 1. (4.11)

Here W0 = W0(e/ε) is the main branch of the Lambert function, e is Euler’s constant.The solution has the following expansion for small ε

Tmax ≈ 1

λ0

[ln

1

ε− ln ln

1

ε− 1

]if ε ¿ 1. (4.12)

This converges slowly as an expansion with respect to (ln ln(ε−1))k whereas the Lambert functioncan be computed easily by fast converging iterations like W (k+1)(x) = ln x− ln W (k)(x).

Numerical integration shows that the computed outbreak time difference Tn is very close toTmax for sufficiently large n but noticeably exceeds it when n = O(10) (see Fig. 6).

Fig. 5b presents a plot of the current growth/decay rate of i

λ(τ) =d ln(i)

dτ. (4.13)

The limiting solution λ(τ) passes from the value λ0 to λ∞, and the travelling wave λ(τ) is closeto λth if τ is small enough. For ρ = 3 and ε = 10−3 we obtain λth ≈ 2.39, λ0 = 2, λ∞ ≈ −0.82.These asymptotes are shown by horizontal dotted lines.

5. Lattice with randomized interaction coefficientsFinally, we consider a lattice of SIR nodes with the same reproduction numbers ρ for every nodebut randomized interaction coefficients εn, i.e the lattice is described by equations

d

dτsn = −ρsnin (5.1)

d

dτin = (ρsn − 1)in + εn

d

dτin−1. (5.2)

10

I. Sazonov et al. The Speed of Epidemic Waves in a One-Dimensional Lattice of SIR Models

e

10-10

10-9

10-8

10-7

10-6

10-5

10-4

10-3

10-2

10-1

l T

Figure 6: Delay time λ0T versus ε for ρ = 2, 3, . . . , 10. Circles indicate numerical results. In thisscale, all circles for different ρ and the same ε are merged. Dashed line is drawn by use of (4.11).

t

in

0 5 10 15 20 25 3010

-20

10-15

10-10

10-5

100

Figure 7: Numerical simulation of lattice with coupled SIR nodes (first 10 nodes) for ρ = 3 andfor randomized εn with ε̄ = 10−3 and (σε/ε̄)

2 = 2.

Let all εn be statistically independent, positive, with the mean ε̄ and variance σ2ε . We consider

the simplest model, with coefficients εn having the log-normal distribution, i.e. ln εn have a normaldistribution with mean µ and variance σ2. Then Tn are also normally distributed in the main termof its asymptotic, see (4.12). Values ε̄ and σε can be determined from the following relations (e.g.[21])

ε̄ = eµ+σ2/2, σ2ε = ε̄2

(eσ2 − 1

). (5.3)

Resolving (5.3), we find µ and σ

µ = ln ε̄− 1

2ln(1 + v), σ2 = ln(1 + v) (5.4)

where v = σ2ε/ε̄

2 is the variance normalized by the square of the mean value.

11

I. Sazonov et al. The Speed of Epidemic Waves in a One-Dimensional Lattice of SIR Models

Assume that in the case ε̄ ¿ 1 and σε ¿ ε̄ all time intervals Tn between outbreaks in the(n− 1)th and nth nodes are described by (4.11), i.e.

λ0Tn ≈(W0(e/εn)− 1

)2

W0(e/εn). (5.5)

Assumption (5.5) agrees well with numerical simulation and implies that the mean time T̄ intervalcan be calculated via the integral

λ0T̄ =

∫ ∞

0

(W0(e/ε)− 1

)2

W0(e/ε)f(ε) dε, (5.6)

f(ε) =e−(ln ε− µ)2

2 σ2

√2πσε

.

Here f(ε) is the probability density of the log-normal distribution. The substitution ε = eµ+σx =ε̄eσx−σ2/2 in (5.6) gives

λ0T̄ =1√2π

∫ +∞

−∞

[W0

(e1−σx+σ2/2

ε̄)− 1

]2

W0

(e1−σx+σ2/2

ε̄) e−

x2

2 dx. (5.7)

Now we expand the integrand into series with respect to σ and integrate it term-by-term. Termswith odd powers of σ vanish after integration. The first two terms approximate (5.7) very well evenfor σ exceeding unity

λ0T̄ =(W0 − 1)2

W0

+ σ2W0 (W 20 + 2)

8 (W0 + 1)5+ O(σ4). (5.8)

Here W0 = W0(e/ε̄). (See Fig. 8: closed circles are fitted well by the dotted line).In numerical simulations for a lattice of 401 nodes, the last 200 outbreak intervals Tn were taken

for averaging. We performed 50 realizations for every v, thus every mean value Tn was averagedby 104 independent values. The results are plotted in Fig. 8. Note that the above assumption fitssurprisingly well for rather large values of the variance: σε > ε̄.

6. Limits of applicability for the approximationsHere we apply the general approach developed in ([22, 16]) to different modifications of our basicmodel and compare how these modelling assumptions are reflected in changes to the predictedepidemic speed.

First we consider all the interaction terms between neighbouring lattice nodes

(∂/∂τ) sn = − ρinsn + (∂/∂τ) [sn−1→n + sn+1→n − sn→n−1 − sn→n+1] (6.1)(∂/∂τ) in = ρ(sn − 1)in + (∂/∂τ) [in−1→n + in+1→n − in→n−1 − in→n+1] . (6.2)

12

I. Sazonov et al. The Speed of Epidemic Waves in a One-Dimensional Lattice of SIR Models

l T l T

a b

Figure 8: Dependence of λ0T̄ on v = (σε/ε̄)2 for ρ = 3, ε̄ = 10−3 (a) and ρ = 5, ε̄ = 10−2 (b).

Numerical results are indicated by open squares. Dotted line indicates the approximation (5.8),closed circles indicate the direct numerical integration of (5.7).

In the linear approximation the equation for in is independent of sn. The linearized equation (6.2)in the fast migration approximation reads as

(∂/∂τ) in = λ0in + ε (∂/∂τ) [in−1 + in+1 − 2in] . (6.3)

Substituting the solution in the form in(τ) = eλ(τ−nT ), we obtain the following characteristicfunction

L(λ, T ) = λ− λ0 − ελ[eλT + e−λT − 2

]. (6.4)

We solve equations L(λ, T ) = 0, ∂L(λ, T )/∂τ = 0 numerically (as they have no solution in aclosed form) and compare with the expressions (4.11) above. The results for comparison are givenin Table 1 for ε = 10−1, 10−2, 10−3. Note that an error of neglecting the transport terms is O(ε).This error is mainly related to neglecting (outgoing) transport terms in→n±1 which decreases theepidemic growth. Thus, the reduced form of equations (4.1)–(4.2) is justified in the approximationof weak coupling.

Table 1: Comparison of Tmax for different approximations

ε λ0Tmax, (6.4) λ0Tmax, (4.11) error, % λ0Tmax, (6.8) error, %

10−3 4.274 4.263 0.2 12.17 18510−2 2.478 2.415 2.5 3.849 6010−1 1.097 0.833 24 1.217 10

13

I. Sazonov et al. The Speed of Epidemic Waves in a One-Dimensional Lattice of SIR Models

Next we compare the results for the epidemic speed in discrete and continuous spatial models.For this, we introduce the spatial coordinate x instead of n

in(τ) = i(x, τ), in+1(τ) = i(x + ∆x, τ)

and, similarly, for sn. Here ∆x is the distance between nodes (later on we set it to unity). Then wecan write

in±1(τ) = i(x±∆x, τ) = i(x, τ)±∆x∂

∂xi(x, τ) +

(∆x)2

2

∂2

∂x2i(x, τ) + O((∆x)3). (6.5)

Substituting (6.5) into (6.3), neglecting terms O((∆x)3) and setting ∆x = 1, we obtain a linearPDE of the form

∂i

∂τ= λ0i + ε

∂3i

∂2x∂τ. (6.6)

Substituting the solution of the form i = eλ(τ−xT ) into (6.6), we obtain the function

L(λ, T ) = λ− λ0 − ελ3T 2. (6.7)

Note that the same equation emerges by expanding e±λT in series with respect to small λT in (6.4).Solving L(λ, T ) = 0, ∂L(λ, T )/∂τ = 0 we obtain

Tmax =2

3λ0

1√3ε

, λmax =3λ0

2. (6.8)

Note that the asymptotic behaviour of Tmax for small ε is different: ε−1/2 in (6.8) in contrast withln(ε−1) in (4.12). The results for comparison are also given in Table 1.

We conclude that for ε ≤ 10−2 the weak coupling approximation gives good accuracy, whereasthe error for the continuous model becomes large. If ε approaches 0.1, the continuous approxima-tion gives more accurate result for epidemic speed than the small coupling approximation althoughthe latter still has the reasonable accuracy O(ε).

Also based on approach described in [22, 16], we can estimate the accuracy of the fast migra-tion approximation. The linearized equation in this case takes the form

∂τin(τ) = λ0in(τ) +

∂2

∂τ 2

∫ τ

−∞in−1(τ

′)g(τ − τ ′)dτ ′ (6.9)

Here we use the kernel g(τ) described by (3.6). Substituting in = eλ(τ−nT ), we obtain the charac-teristics function

L(λ, T ) = λ− λ0 − ελeλT

λ∆τ + 1(6.10)

The results of numerical solutions for L(λ, T ) = 0, (∂/∂τ)L(λ, T ) = 0 are plotted in Fig 9.Observe that for small ε the increment λ0T (ε, ∆τ)− λ0T (ε, 0) depends slightly on ε (see Fig 9a).This leads to a smaller relative change of the epidemic speed for small ε (see Fig 9b).

14

I. Sazonov et al. The Speed of Epidemic Waves in a One-Dimensional Lattice of SIR Models

0 2 4 6 8 10

1

2

l Dt0

l [ (Dt)- (0)]T T0

1

23

00 2 4 6 8 10

2

1

1

2

3

l Dt0

T T(Dt)/ (0)

3

a b

Figure 9: Change of T cause by finite migration characteristic time ∆τ : absolute (a) and relative(b) increment of T (τ) for ε = 10−1 (curve 1), 10−2 (curve 2) and 10−3 (curve 3).

7. DiscussionA 1D lattice of interacting SIR epidemic models is considered in the approximation of weak in-teractions and fast migration. Numerical simulations confirm that over a wide parameter range(ε = 10−10 ÷ 10−1, ρ = 2 ÷ 10), the solution tends to an universal one which behaves as atravelling wave preserving its shape and speed. In the framework of this model, a simple explicitformula is obtained for the speed of the epidemic spread in a lattice with either constant or randominteraction coefficients. Compared with numerical simulations, the solutions for the randomizedcase demonstrate surprisingly good accuracy over a wide range of applicability, i.e. when the vari-ance of the coefficient exceeds the square of its mean value. In the weak coupling approximationthe developing outbreak in every node of the lattice is well approximated by the limiting solutionobtained for a single SIR model.

A similar analysis is performed for the finite characteristic migration time ∆τ , and responsesof the form given in (3.5)–(3.6). The qualitative behaviour of the travelling waves remains thesame, with the speed reduced compared to the approximation of fast migration. The shape of thetravelling wave around the outbreak will presumably remain the same and be well approximated bythe limiting solution. The numerical simulation of a lattice with the finite ∆τ and random couplingparameters will be a subject of future work.

Note that the results of this paper can be straightforwardly extended to a response function ofform g(τ) = εθ(τ − δ). We simply add δn to Tn obtained numerically (see Fig. 4) or theoretically(see (4.11)), i.e. T δ

n = Tn + δn where δn is the migration delay between the nth and (n+1)thnodes. In the case of a Gaussian distribution of coupling coefficients, the distribution of T δ

n remainsGaussian in the main term of the approximation with the mean value of T δ

n being the sum of T̄ (5.8)and the mean of δn.

In the present work, we concentrate on an asymptotic case when the number of introducedinfectives is small compared to the number of cases generated during the outbreak. In the lattice

15

I. Sazonov et al. The Speed of Epidemic Waves in a One-Dimensional Lattice of SIR Models

model, this is equivalent to small interaction coefficients (weak coupling approximation). In thisapproximation, the outbreak time (in a single SIR) and the time between outbreaks in neighbouringnodes are logarithmically large: see (2.12) for a single SIR model and (4.12) for the lattice. Thenumerical solutions in Section 6 (see Table 1) demonstrate that an approximation of finite differ-ences by derivatives, i.e. the models of reaction-diffusion PDEs [18, 8, 14, 6, 3] lead to significanterrors in the value of propagation speed, and thus a discrete spatial model is essential for the analy-sis (cf. [19, 12]). Clearly, the problem of determining whether the discrete or continuous nature ofthe phenomena should be considered is rather serious (see discussion in the paper ‘The importanceof being discrete and spatial’ [19]). In contrast to this paper we do not consider a population ofdiscrete individuals, however we show that the step from continuous media to discreteness in spacecan itself lead to different predictions as demonstrated here.

Our formulae for the speed and shape of a travelling wave in a lattice of SIR models potentiallyprovides the means to estimate the interaction coefficients between urban populations in relevantecological databases. One of the key examples from epidemiology, that links an extensive databaseto an underlying travelling wave model is based on the records of measles outbreaks in and aroundLondon since the 1940s [6]. In the case of the measles data however, it was concluded that theoutbreaks in large cities were in reality synchronized by strong linkage, endemic status and ashared seasonal forcing term in the contact rates (due to the school year). In contrast, in a latticeof towns and small population centres, travelling waves could persist if there was local extinctionof disease in small centres. The outbreaks in these centres will therefore tend to lag behind theepidemic generated in a larger community until the weak coupling between the towns sparks anew outbreak. We note that our derivation of time to the peak of the outbreak can also provide asimple approximation for the fade-out time of the outbreak (e.g., by defining the extinction timewhen the number of infected individuals is reduced below 1). If travelling waves in epidemiologyare caused by an interaction of weak coupling, and local centre extinction, our model provides auseful analytical framework for investigating the underlying mechanisms of observed travellingwaves using long term historical records such as certain measles scenarios, and other directlytransmitted infections such as the influenza viruses.

The problems addressed here are characterized by different time scales, T , ∆τ and 1/λ0. Amore delicate use of ‘separation of scales’ phenomena and backward influence will be exploited inforthcoming papers. We also intend to study other types of deterministic or random coupling andother distributions of random coupling constants.

Although a network of the simplest topology is considered here (intuitively it corresponds to asequence of towns located along a railway or highway), a more general interpretation of the resultsis required. A 2D system would clearly fit more closely with the geographical spread of realdiseases. From a physical viewpoint, the 1D lattice corresponds to a plane wave approximation,which is standard for a cylindrical wave far from the source. We believe that away from a localizedsource, a cylindrical epidemic follows more and more the pattern described by the 1D modelconsidered in our work. A more specific description of this phenomena will be a subject of futurework.

Although very well studied, the SIR model continues to provide analytical insight into the basicmechanisms of epidemic processes. This basic model cannot represent all the features of real-life

16

I. Sazonov et al. The Speed of Epidemic Waves in a One-Dimensional Lattice of SIR Models

phenomena, it is often found that the more complex simulation-type models are used in the case ofspecific outbreaks and to plan actual interventions. However, there can be concerns over assessingthe validity of complex models, that include many nodes, and how the large number of assumptionsand parameter values determine the model output [13]. There is therefore a need for continuousdevelopment of analytical models of the spatial evolution of epidemics (e.g. [2, 11]) to provide atheoretical basis for those used in planning scenarios.

Acknowledgment. The work was supported by EPSRC Project EP/F014015/1. We would liketo thank anonymous referees for helpful comments.

References[1] R.M. Anderson, R.M. May. Infectious Disease of Humans: Dynamics and Control. Oxford

Univ. Press, Oxford, 1991.

[2] D. Brockmann, L. Hufnagel, T. Geisel. Dynamics of modern epidemics. In SARS: ACase Study in Emerging Infections. Edited by A.R. McLean, R.M. May, J. Pattison, andR.A. Weiss. Oxford Univ. Press, Oxford, 2005.

[3] R.S. Cantrell, C. Cosner. Spatial Ecology via Reaction-Diffusion Models. Mathematical Sur-veys and Monographs. Wiley, Chichester, 2003.

[4] R.M. Corless, G.H. Gonnet, D.E.G. Hare, D.J. Jeffrey, and D.E. Knuth. On the Lambert Wfunction. Advances in Computational Math., 5 (1996), 329–359.

[5] D.J. Daley, J. Gani. Epidemic Modelling. Cambridge Univ. Press, Cambridge, 1996.

[6] B.T. Grenfell, O.N. Bjørnstad, J. Kappey. Traveling waves and spatial hierarchies in measlesepidemics Nature, 414 (2001), No. 13, 716–723.

[7] J.A.P. Heesterbeek. A brief history of R0 and a recipe for its calculation. Acta Biotheor., 50(2002) No. 5, 375–376.

[8] Y. Hosono, B. Ilyas. Traveling waves for a simple diffusive epidemic model. Math. Models &Methods in Appl. Sciences., 5 (1995), No. 7, 935–966.

[9] R.R. Kao. The role of mathematical modelling in the control of the 2001 FMD epidemic inthe UK. Trends in Microbiology, 10 (2002), 279-286.

[10] M.J. Keeling, K.T.D. Eames. Networks and epidemic models. J. R. Soc. Interface, 2 (2005),295-307.

[11] E. Kenah, J.M. Robins. Second look at the spread of epidemics on networks. Phys. Rev. E, 76(2007) No. 3, Pt. 2, 036113.

[12] S.A. Levin, R. Durrett. From Individuals to epidemics Philos. Trans.: Biological Sciences,351 (1996), 1615–1621.

17

I. Sazonov et al. The Speed of Epidemic Waves in a One-Dimensional Lattice of SIR Models

[13] R.M. May. Uses and abuses of mathematics in biology. Science, 303 (2004), 790–793.

[14] V. Mendez. Epidemic models with an infected-infectious period. Phys. Rev. E., 57 (1998)No. 3, 3622–3624.

[15] D. Mollison, ed. Epidemic Models: Their Structure and Relation to Data, Cambridge Univ.Press, Cambridge, 1995.

[16] D. Mollison. Dependence of Epidemic and Population Velocities on Basic parameters. Math.Biosciences, 107 (1991), 255–287.

[17] C. Moore, M. Newman. Epidemics and percolation in small-world networks. Phys. Rev. E,61 (2000), 5678-5682.

[18] J.D. Murray. Mathematical Biology. Springer-Verlag. London, 1989.

[19] R. Durrett, S.A. Levin. The importance of being discrete (and spatial), Theor. Pop. Biol. 46(1994), 363–394.

[20] L. Rass, J. Radcliffe. Spatial deterministic epidemics. Vol. 102 of Mathematical Surveys andMonographs. Amer. Math. Soc. Providence RI, USA, 2003.

[21] Y. Suhov, M. Kelbert. Probability and Statistics by Example. Vol. I. Cambridge Univ. Press,Cambridge, 2005.

[22] F. van den Borsch, J.A.J. Metz, O. Diekmann. The velocity of spatial population expansion,J. Math. Biol., 28 (1990), 529–565.

A Approximations for the outbreak timeDividing (2.3) by (2.2) and integrating the obtained equation, taking into account (2.4), we obtainthe dependence of i versus s (cf. [18])

i = 1− s +1

ρln

s

s0

(A1)

The inverse dependence can be written via the Lambert Wk(x) function, i.e. the kth branch ofsolution of W exp(W ) = x [4] (alternatively called ProductLog in the algebraic package MATH-EMATICA)

s = −1

ρWk

(−ρs0eρ(i−1)

)(A2)

where k = −1 for the growing part of the solution and k = 0 for the decaying part. Setting i = 0we obtain an explicit formula for s∞

s∞ = −1

ρW0

(−ρs0e−ρ

)

18

I. Sazonov et al. The Speed of Epidemic Waves in a One-Dimensional Lattice of SIR Models

which gives (2.8) by setting s0 = 1. Substituting (A1) into (2.2) and integrating τ(s) in the limits[smax, s0], we obtain formula (2.10).

To study the asymptotic behaviour of (2.10) for i0 → 0 we make the substitution s = 1− ξ

τoutb∼=

∫ 1−ρ−1

i0

f(ξ)(A3)

where f(ξ) = (1− ξ)(ρξ + ln

((1− ξ)/(1− i0)

)).

Denote the root of f(ξ) with the minimal absolute value by ξ0 = −i0(ρ − 1)−1 + O(i20).Consider the function

F (ξ; i0) =1

f(ξ)− 1

f ′(ξ0)(ξ − ξ0). (A4)

It is regular in the vicinity of ξ = 0, and it is easy to show that F (ξ; i0) = F (ξ; 0) + O(i0),∀ξ ∈ [0, 1]. Hence, ∫ 1−ρ−1

i0

F (ξ; i0)dξ =

∫ 1−ρ−1

0

F (ξ; 0)dξ + O(i0)

Our goal is to represent τoutb as a term independent of i0 plus a correction

τoutb∼=

[∫ 1−ρ−1

0

F (ξ; 0)dξ + O(i0)

]+

∫ 1−ρ−1

i0

f ′(ξ0)(ξ−ξ0).

To this aim, we estimate the correction represented by the second integral∫ 1−ρ−1

i0

f ′(ξ0)(ξ−ξ0)=

1

λ0

[ln

1

i0+2 ln(1−ρ−1)

]+ O(i0).

This argument also yields an expression for constant C(ρ) in (2.12)

C(ρ) =

∫ 1−ρ−1

0

F (ξ; 0)dξ +2 ln(1− ρ−1)

λ0

. (A5)

0 5 10 15 200

0.1

0.2

0.3

0.4

0.5

r

C

Figure 10: Outbreak time coefficient C0 (a) vs ρ: exact—solid, approximation—dashed line.

19

I. Sazonov et al. The Speed of Epidemic Waves in a One-Dimensional Lattice of SIR Models

The function f(ξ) vanishes at ξ = 0 and ξ = 1. As the integrand in (A3) grows towards thelimits of the integration domain, its boundaries are the main contributors, especially the integrationnear its lower limit (ξ = i0). Approximating f(ξ) by the second order polynomial

P2(ξ) = λ0 ξ (1− ξ)

which vanishes at ξ = 0, 1 and matches the first derivative f ′(0), and neglecting non-essentialterms at i0 → 0, we obtain

C ≈ ln(λ0)

λ0

(A6)

that gives A0 ≈ λ0 in accordance with (2.13). Comparison between C(ρ) computed numericallyand approximated by (A6) is shown in Fig. 10.

20