14
Evolutionary dynamics of infectious diseases in nite populations Jan Humplik a,b,c , Alison L. Hill a,d,n , Martin A. Nowak a,e a Program for Evolutionary Dynamics, Harvard University, One Brattle Square, Cambridge, MA 02138, USA b Institute of Science and Technology Austria, Am Campus 1, 3400 Klosterneuburg, Austria c Faculty of Mathematics and Physics, Charles University in Prague, Czech Republic d Biophysics Program and Harvard-MIT Division of Health Sciences and Technology, Harvard University, Cambridge, MA 02138, USA e Department of Organismic and Evolutionary Biology, Department of Mathematics, Harvard University, Cambridge, MA 02138, USA HIGHLIGHTS We study the effects of nite populations on the evolution of infectious diseases. Our results show that maximization of the basic reproductive ratio may be violated. We calculate a lower bound for the deviation of selection pressure from the innite population case. We nd that mutually invasible strains may exist, and invading strains can be better for the host population. article info Article history: Received 23 October 2013 Received in revised form 17 June 2014 Accepted 30 June 2014 Available online 10 July 2014 Keywords: SIS model Stochastic logistic model Basic reproductive ratio Virulence abstract In infectious disease epidemiology the basic reproductive ratio, R 0 , is dened as the average number of new infections caused by a single infected individual in a fully susceptible population. Many models describing competition for hosts between non-interacting pathogen strains in an innite population lead to the conclusion that selection favors invasion of new strains if and only if they have higher R 0 values than the resident. Here we demonstrate that this picture fails in nite populations. Using a simple stochastic SIS model, we show that in general there is no analogous optimization principle. We nd that successive invasions may in some cases lead to strains that infect a smaller fraction of the host population, and that mutually invasible pathogen strains exist. In the limit of weak selection we demonstrate that an optimization principle does exist, although it differs from R 0 maximization. For strains with very large R 0 , we derive an expression for this local tness function and use it to establish a lower bound for the error caused by neglecting stochastic effects. Furthermore, we apply this weak selection limit to investigate the selection dynamics in the presence of a trade-off between the virulence and the transmission rate of a pathogen. & 2014 Elsevier Ltd. All rights reserved. 1. Introduction Managing infectious diseases of humans, animals and crops requires predicting their dynamics from short-term spread to long-term evolution. Mathematical models provide a framework for this task (Anderson and May, 1991; Diekmann and Heesterbeek, 2000). The classic modeling approach involves tracking specic compartments of the host population with a system of differential equations. These compartments generally divide individuals into different demographic classes (based on, for example, age or exposure) and different stages of infection (such as susceptible, infected, and recovered) with various pathogen strains. Individuals transition between states based on parameters specic to the disease, environ- ment, and mixing patterns. Perhaps the most important insight from such models, when applied deterministically to innite populations, is the existence of a critical value of the parameters necessary for a disease to cause an epidemic. This threshold is generally represented with the basic reproductive ratio, R 0 , which describes the expected number of secondary infections caused by a single infected individual in a fully susceptible population (reviewed in Heffernan et al., 2005). For a pathogen-environment-host scenario with R 0 o1, the infection is subcritical and is guaranteed to die out before infecting a substantial fraction of the population, while for R 0 41, an epidemic can occur. In these deterministic models, R 0 ¼ 1 often corresponds to a trans- critical bifurcation. Contents lists available at ScienceDirect journal homepage: www.elsevier.com/locate/yjtbi Journal of Theoretical Biology http://dx.doi.org/10.1016/j.jtbi.2014.06.039 0022-5193/& 2014 Elsevier Ltd. All rights reserved. n Corresponding author at: Program for Evolutionary Dynamics, Harvard University, One Brattle Square, Cambridge, MA 02138, USA. E-mail addresses: [email protected] (J. Humplik), [email protected] (A.L. Hill). Journal of Theoretical Biology 360 (2014) 149162

Evolutionary dynamics of infectious diseases in …ped.fas.harvard.edu/.../1-s2.0-s0022519314003907-main.pdfEvolutionary dynamics of infectious diseases in finite populations Jan

  • Upload
    others

  • View
    3

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Evolutionary dynamics of infectious diseases in …ped.fas.harvard.edu/.../1-s2.0-s0022519314003907-main.pdfEvolutionary dynamics of infectious diseases in finite populations Jan

Evolutionary dynamics of infectious diseases in finite populations

Jan Humplik a,b,c, Alison L. Hill a,d,n, Martin A. Nowak a,e

a Program for Evolutionary Dynamics, Harvard University, One Brattle Square, Cambridge, MA 02138, USAb Institute of Science and Technology Austria, Am Campus 1, 3400 Klosterneuburg, Austriac Faculty of Mathematics and Physics, Charles University in Prague, Czech Republicd Biophysics Program and Harvard-MIT Division of Health Sciences and Technology, Harvard University, Cambridge, MA 02138, USAe Department of Organismic and Evolutionary Biology, Department of Mathematics, Harvard University, Cambridge, MA 02138, USA

H I G H L I G H T S

� We study the effects of finite populations on the evolution of infectious diseases.� Our results show that maximization of the basic reproductive ratio may be violated.� We calculate a lower bound for the deviation of selection pressure from the infinite population case.� We find that mutually invasible strains may exist, and invading strains can be better for the host population.

a r t i c l e i n f o

Article history:Received 23 October 2013Received in revised form17 June 2014Accepted 30 June 2014Available online 10 July 2014

Keywords:SIS modelStochastic logistic modelBasic reproductive ratioVirulence

a b s t r a c t

In infectious disease epidemiology the basic reproductive ratio, R0, is defined as the average number ofnew infections caused by a single infected individual in a fully susceptible population. Many modelsdescribing competition for hosts between non-interacting pathogen strains in an infinite population leadto the conclusion that selection favors invasion of new strains if and only if they have higher R0 valuesthan the resident. Here we demonstrate that this picture fails in finite populations. Using a simplestochastic SIS model, we show that in general there is no analogous optimization principle. We find thatsuccessive invasions may in some cases lead to strains that infect a smaller fraction of the hostpopulation, and that mutually invasible pathogen strains exist. In the limit of weak selection wedemonstrate that an optimization principle does exist, although it differs from R0 maximization. Forstrains with very large R0, we derive an expression for this local fitness function and use it to establish alower bound for the error caused by neglecting stochastic effects. Furthermore, we apply this weakselection limit to investigate the selection dynamics in the presence of a trade-off between the virulenceand the transmission rate of a pathogen.

& 2014 Elsevier Ltd. All rights reserved.

1. Introduction

Managing infectious diseases of humans, animals and cropsrequires predicting their dynamics – from short-term spread tolong-term evolution. Mathematical models provide a framework forthis task (Anderson and May, 1991; Diekmann and Heesterbeek,2000). The classic modeling approach involves tracking specificcompartments of the host population with a system of differentialequations. These compartments generally divide individuals intodifferent demographic classes (based on, for example, age or exposure)

and different stages of infection (such as susceptible, infected, andrecovered) with various pathogen strains. Individuals transitionbetween states based on parameters specific to the disease, environ-ment, and mixing patterns. Perhaps the most important insight fromsuch models, when applied deterministically to infinite populations, isthe existence of a critical value of the parameters necessary for adisease to cause an epidemic. This threshold is generally representedwith the basic reproductive ratio, R0, which describes the expectednumber of secondary infections caused by a single infected individualin a fully susceptible population (reviewed in Heffernan et al., 2005).For a pathogen-environment-host scenario with R0o1, the infectionis subcritical and is guaranteed to die out before infecting a substantialfraction of the population, while for R041, an epidemic can occur.In these deterministic models, R0 ¼ 1 often corresponds to a trans-critical bifurcation.

Contents lists available at ScienceDirect

journal homepage: www.elsevier.com/locate/yjtbi

Journal of Theoretical Biology

http://dx.doi.org/10.1016/j.jtbi.2014.06.0390022-5193/& 2014 Elsevier Ltd. All rights reserved.

n Corresponding author at: Program for Evolutionary Dynamics, HarvardUniversity, One Brattle Square, Cambridge, MA 02138, USA.

E-mail addresses: [email protected] (J. Humplik),[email protected] (A.L. Hill).

Journal of Theoretical Biology 360 (2014) 149–162

Page 2: Evolutionary dynamics of infectious diseases in …ped.fas.harvard.edu/.../1-s2.0-s0022519314003907-main.pdfEvolutionary dynamics of infectious diseases in finite populations Jan

Pathogen populations are genetically diverse due to highmutation rates and large population sizes. Multiple strains aregenerally competing for the same hosts. Pathogens acquire adap-tations as they move from one environment or species to another,and are continually in an evolutionary arms race with their hosts.As well as describing the spread of a disease, in many traditionalinfectious disease models R0 is sufficient to describe the evolu-tionary trajectory of the disease: the selection gradient is in thedirection of higher R0, and R0 is the quantity that is maximized bythe evolutionarily stable strategy (for example, Anderson and May,1982).

This strategy of maximizing R0 to determine the evolutionarilyoptimal pathogen strategy turns out to have limitations in manysystems (Metz et al., 1996, 2008). Violations of this principle havebeen observed when the simplest models are extended to accountfor biologically relevant population dynamics, such as particulartypes of density-dependent demographic or transmission para-meters, frequency-dependent selection, or host-pathogen co-evo-lution (reviewed in Dieckmann, 2002). More generally, sufficientconditions for R0 maximization have been derived for one of themost common epidemiological models (the well-mixed SIRmodel). These include the absence of genotype-by-environmentinteractions, the presence of only a single transmission pathway,and specific constraints on the density-dependence of mortality(Cortez, 2013). Additionally, the option for pathogens to co-infecthosts or displace resident infections within hosts (“superinfec-tion”) can also result in evolution towards suboptimal R0 (May andNowak, 1994; Bonhoeffer and Nowak, 1994; Nowak and May, 1994;May and Nowak, 1995). Due to the convenience of conclusionsbased on R0, exceptions to the “R0 maximization” rule remainimportant to understand. However, most previous work hasfocused on deterministic models in infinite populations, and thegeneralizability of these conclusions outside this limit is unknown.In this paper we show that even in the context of the simplest“SIS” disease model, R0 maximization may fail when we considerfinite-sized populations, and the direction of selection may besubstantially different in small populations.

The Susceptible-Infected-Susceptible (SIS) Model (Andersonand May, 1991; Diekmann and Heesterbeek, 2000) is one of thesimplest mathematical models describing an infectious processand has been used to describe a variety of phenomena such ascommunicable diseases, computer viruses and peer-influencedbehaviors. The model classifies individuals as either susceptible(healthy) or infected at any point in time. Susceptible individualscan become infected through contact with infected individuals,with a transmission rate β, and once infected, individuals recoverat a rate δ. Recovered hosts are once again susceptible to theinfection. Individual pathogens are not explicitly tracked, butassumed to only survive in living hosts and be transferred throughsome form of contact. This model ignores the detailed time-courseof the disease within a patient (such as latent or exposed phases)and any type of long-term immunity. In simplest form, it alsoassumes that the population consists of randomly and homoge-neously mixing individuals, although extensions to network-basedcontact patterns are common (for example, Eames and Keeling,2002; Cator and Van Mieghem, 2013). Despite these limitations,the small number of parameters of the SIS model (β, δ) means thatdetailed analysis of the phase space is possible, which often yieldsresults that can be generalized to more complex systems.

While deterministic models are by far the most commonly usedin epidemiology, stochastic models are required to answer ques-tions about finite populations and the probability of an epidemicoccurring. Analysis of the stochastic SIS model is complicated bythe fact that the model contains an absorbing state when allindividuals are healthy, which is guaranteed to be reached for allnon-zero values of the parameters as time goes to infinity (Nasell,

1995). As a consequence of this absorbing state, there is no longera critical threshold value of R0 such that the infection reaches anon-zero equilibrium level, since the only true equilibrium is thezero-infection state. The transition is a bit more subtle and is clearonly in the asymptotic limit of large N: for R0o1, the diseaseprevalence decays exponentially, with survival time τ¼Oðlog ðNÞÞ,while for R041, survival time grows at least as eN

αfor some α40.

Also, for large N a long-lived “metastable state” (or “quasi-stationary state”) is reached with an infected level equivalent tothe deterministic model (Ovaskainen, 2001; Ganesh et al., 2005;Castellano and Pastor-Satorras, 2010). Various methods have beendeveloped to analyze the quasi-stationary distribution of thestochastic SIS model by analyzing the limiting behavior of relatedprocesses where the absorbing state has been eliminated. Onerecent approach considers a perturbation to the model to includespontaneous infection of healthy nodes at a rate ϵ and henceeliminates the absorbing state (the ϵ-SIS model). This model hasbeen solved exactly, and for certain small values of ϵ, its thresholdand equilibrium behavior approximates that of the SIS model (VanMieghem and Cator, 2012). Another method considers a related“return process” in which all absorbing states are re-assignedrandomly to transient states according to an arbitrary “return”distribution. By iteratively finding the equilibrium of this systemand using it as the next return distribution, the equilibriumeventually approximates the SIS quasi-stationary distribution(Barbour and Pollett, 2010, 2012). For smaller N, the behavior ofthe stochastic SIS model is further complicated by the fact that theaverage number of secondary infections is no longer described bythe expression for R0 in infinite populations, and is consistentlysmaller (Ross, 2011).

Here we are interested not only in the initial spread but in theevolution of a pathogen obeying SIS dynamics. We will considerthe potential for competition between strains with different valuesof β and δ. In the deterministic, infinite population case, R0 ¼ β=δ,and evolution towards increasing R0 will tend to increase β anddecrease δ when they are uncorrelated. If they are correlated, as isoften considered in studies of the evolution of virulence, evolutionmay proceed towards optimal intermediate values (Antia et al.,1994; Bull, 1994; Lenski and May, 1994; Lipsitch et al., 1996; Levin,1996; Regoes et al., 2000; Ganusov et al., 2002; Ganusov and Antia,2003; Ebert and Bull, 2008). We expect that evolutionary out-comes for infectious diseases in finite populations may be differ-ent, as important effects have been found in other systems. It isonly in finite populations that neutral drift and the fixationprobability become meaningful terms. In evolutionary game the-ory (a framework for studying frequency-dependent selection)(Smith, 1982; Weibull, 1997; Hofbauer and Sigmund, 1998; Nowakand Sigmund, 2004), the conditions for a strategy to be evolutio-narily stable in an infinite population are neither necessary norsufficient in a finite population (Nowak et al., 2004). Here wedemonstrate that similarly interesting results are found for infec-tious disease evolution in finite populations.

2. The SIS model for two diseases

The model we examine is an extension of the stochastic SISmodel that describes the dynamics of two concurrent infections:strain 1, s1, and strain 2, s2. We consider a constant population of Nhosts, where each individual can either be in a susceptible state, orbe infected with one of the two strains. This model is formallydescribed by a continuous-time birth-death process. The numberof individuals infected with the strain si is given by ni. The numberof susceptible individuals is always N�n1�n2. The state space ofthis process consists of pairs of integers ½n1;n2� such thatn1þn2rN (Fig. 1A).

J. Humplik et al. / Journal of Theoretical Biology 360 (2014) 149–162150

Page 3: Evolutionary dynamics of infectious diseases in …ped.fas.harvard.edu/.../1-s2.0-s0022519314003907-main.pdfEvolutionary dynamics of infectious diseases in finite populations Jan

A susceptible individual becomes infected with strain si(i¼ 1 or 2) at a rate equal to the product of βi=N and the numberof infected individuals it contacts. The values of βi will be referredto as the transmission rates. We assume that the population is well-mixed, i.e., that all individuals have equally weighted contact withall other individuals at all times. Infected individuals recover fromthe disease and become susceptible again with a rate δi, which werefer to as the recovery rates. Equivalently, an individual may die,and – to keep population size constant – immediately be replacedby a new susceptible host. Uninfected hosts die at a rate u. Hostsinfected with the strain si die at a rate uþvi, where vi, thevirulence, is defined as the increase in the death rate of a hostcaused by the presence of the infection. The event of either death(with replacement) or recovery of an individual infected with thestrain si constitutes a turnover event and occurs at the turnoverrate ai ¼ uþviþδi. The last assumption we make is that individualsmay only be infected with a single strain at a time, so that neithersuperinfection nor coinfection are allowed.

If we define P½n1 ;n2�ðtÞ as the probability that the system is in astate ½n1;n2� at time t, then we can write a system of differentialequations (the master equations) describing the time evolution ofthis process:

dP½n1 ;n2�ðtÞdt

¼ β1

Nðn1�1ÞP½n1 �1;n2 �ðtÞþ

β2

Nðn2�1ÞP½n1 ;n2 �1�ðtÞ

� ��ðN�n1�n2þ1Þþa1ðn1þ1ÞP½n1 þ1;n2 �ðtÞþa2ðn2þ1ÞP½n1 ;n2 þ1�ðtÞ

� β1

Nn1þ

β2

Nn2

� �ðN�n1�n2Þþa1n1þa2n2

� �P½n1 ;n2 �ðtÞ: ð1Þ

Here we additionally prescribe the function P½n1 ;n2 �ðtÞ to be zerooutside of the domain 0rn1þn2rN. Eq. (1) requires an initialcondition, which is discussed in Section 3. Note that by condition-ing on either n1 ¼ 0 or n2 ¼ 0, the model reduces to a standardone-disease stochastic SIS model.

Analogously to the simple SIS model, the state where allindividuals are susceptible (n1 ¼ n2 ¼ 0) is an absorbing state,and since the total number of states is finite, recovery of theentire population is guaranteed as t-1. Consequently, the sta-tionary solution of Eq. (1) is trivial, and different methods arerequired to understand the important disease dynamics describedby the model.

In particular, we are interested in the following scenario. Afterthe resident infection s1 has spread among the hosts, a singleindividual infected with a mutant strain s2 emerges in thepopulation (see Section 3 for a precise description). Our goal isto decide whether selection favors the invasion of the mutantstrain into the population infected by the resident strain. Toanswer this question, let us define the fixation probability of s2invading s1 as the probability that strain 1 becomes extinct beforestrain 2 does. We will say that invasion of strain 2 into strain 1 isfavored by selection if its fixation probability is greater than that of aneutral mutant. A mutant strain is neutral if its fixation probabilityis the same as that of a mutant with parameters identical to thoseof the resident (i.e., β2 ¼ β1, a2 ¼ a1). In particular, a neutralmutant of s1 added into a population containing n individualsinfected with s1 has a fixation probability 1=ð1þnÞ. These defini-tions follow a standard approach to studying adaptation in finitepopulations (for example Crow and Kimura, 1970; Proulx and Day,2002; Nowak et al., 2004).

To anticipate the results of this analysis, consider a populationof only two hosts. One host is infected with strain 1 and the otherwith strain 2. Since no new infections can occur while both exist inthe population, invasion of strain 2 will be favored by selection ifits average turnover time is greater than the average turnover timeof strain 1 (the turnover times are exponentially distributedrandom variables). This will happen if a14a2. On the other hand,if the population is infinite, the condition for the invasion of strain2 to be favored is that R1oR2, where Ri ¼ βi=ai is the basicreproductive ratio. Therefore, selection dynamics depend heavilyon the population size, with low turnover rates being moreimportant for smaller populations. One of the goals of this paperis to describe the behavior for population sizes between the twoextreme cases of N¼2 and N¼1.

The rest of this paper is organized as follows. In Section 3 weprecisely define the initial condition for the competition betweentwo strains. In Section 4 we proceed to calculate the fixationprobability of a newly introduced strain, and derive a closed formexpression in the limit of large values of the basic reproductiveratios. We use these results in Section 5 to analyze which strainsare favored to invade different resident strains, and we comparethe resulting behavior with known results for infection dynamicsin infinite populations. In Section 6, we analyze the weak selectionlimit to quantify how population size changes the direction of

0 1 2 3 4 5 6 7 8 90

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

Number infected with resident (n1)

Freq

uenc

y ρ n

1 (R1,

N)

R1 = 0.5R1 = 1.5R1 = 3R1 = 5R1 = 10R1 = 40

Number infected with resident (n1)

Num

ber i

nfec

ted

with

mut

ant (

n 2) N

N-1

N-2

1

00 1 N-2 N-1 N

Fig. 1. State space and initial conditions for the two-disease stochastic SIS model. (A) Each dot represents a possible state of the two infections in a population. Transitionscan occur only between neighboring states. Reaching either the lower or left boundary of this state space means that one of the disease strains has gone extinct (and so theremaining strain is said to have “fixed”). The red dots represent the state space in the case of very large basic reproductive ratios (described in Section 4). Restriction of theoriginal Markov process to this state space reduces it to a process on a finite one-dimensional lattice with two absorbing states. (B) The initial condition (Eq. (3)) describesthe number of hosts infected with the resident strain (n1) at the instant the mutant strain (s2) appears. The mutant strain is introduced when a mutation arises in an infectedhost. All graphs use N¼10. Distributions for several values of the basic reproductive ratio of the resident strain (R1) are shown. As R1-1, the distribution becomes non-zeroonly for n1 ¼N�1. (For interpretation of the references to color in this figure caption, the reader is referred to the web version of this article.)

J. Humplik et al. / Journal of Theoretical Biology 360 (2014) 149–162 151

Page 4: Evolutionary dynamics of infectious diseases in …ped.fas.harvard.edu/.../1-s2.0-s0022519314003907-main.pdfEvolutionary dynamics of infectious diseases in finite populations Jan

selection and to identify population sizes at which stochasticeffects become important. We also use this limit to study theselection dynamics when there is a trade-off between the viru-lence and the transmission rate.

3. Quasi-stationary distribution as a description of the initialepidemic

We now describe in detail the initial conditions for the invasionof a new disease strain into a population infected with a residentstrain at an endemic equilibrium. There are two components tothe initial condition: a precise description of the state of theresident infection at this “equilibrium,” and a procedure for theintroduction of a mutant strain.

Prior to the introduction of the mutant (strain 2), the dynamicsof the resident infection (strain 1) are described by the solution ofEq. (1) conditioned on n2 ¼ 0 (the simple stochastic SIS model). Letus denote the resulting probability distribution Pn1 ðt; a1;β1;NÞ. Thequasi-stationary distribution for this process is defined by condi-tioning Pn1 ðt; a1;β1;NÞ to non-extinction:

Qn1ðR1;NÞ ¼ lim

t-1Pn1 ðt; a1;β1;NÞ

1�P0ðt; a1;β1;NÞ; n1A 1;…;Nf g: ð2Þ

We assume that strain 1 is in this quasi-stationary state beforestrain 2 appears. The existence of this limit, and its independenceof the initial condition of Pn1 ðt; a1;β1;NÞ, was proved in previouswork (Mandl, 1960). Although a closed form expression for theparameter dependence of Eq. (2) is not known, many approxima-tions have been derived (Kryscio and Lefévre, 1989; Nåsell, 2001;Ovaskainen, 2001). Since we could use either a1 or β1 to repar-ametrize time in Eq. (1), the quasi-stationary distribution dependsonly on R1.

Throughout the paper, we will refer to the quantities Ri ¼ βi=aias basic reproductive ratios, although, strictly speaking, they arebasic reproductive ratios in infinite populations. Basic reproductiveratios in finite populations calculated as the average number ofsecondary infections differ slightly from the quantities Ri (Ross,2011).

Even though the quasi-stationary distribution exists for any R1,we will mostly restrict our analysis to residents with R141.Roughly speaking, this ensures that the infection can affect anon-negligible fraction of the population, although the distinctionis not as clear as in infinite populations.

We now address the issue of how a mutant is introduced intoan infected population. We consider the following scenario: Theresident infection s1 has reached quasi-stationarity in a populationof N individuals, at which point a pathogen in one of the infectedindividuals undergoes a mutation into a new strain s2. Thiscorresponds to the convention used in evolutionary invasionanalysis. In many cases, a weak selection limit is assumed to hold,and the parameters of s2 differ only infinitesimally from thoseof s1.

Under this scenario, after the mutation occurs, the populationconsists of N hosts in which a single individual is infected with themutant strain s2 and n1 individuals are infected with the residentstrain s1. The variable n1 is random with the probability distribu-tion

ρn1 ðR1;NÞ �Qn1 þ1ðR1;NÞ; n1Af0;…;N�1g: ð3Þ

Fig. 1B shows several examples of this distribution assumingN¼10. Eq. (1) describes a linear system (dp=dt ¼Ap where theelements of p correspond to P½n1;n2�) that, in the case of only onedisease present, contains one absorbing and N transient states. Thequasi-stationary distribution is calculated numerically by evaluat-ing the normalized right eigenvector corresponding to the largest

eigenvalue of the matrix A0, which is obtained from the transitionrate matrix (A) by removing the row and the column correspond-ing to the absorbing state (Keeling and Ross, 2008).

Although we have chosen a precise definition of the initialcondition, there are other alternative definitions that are equallybiologically feasible. For example, strain 2 could appear byimmigration and a subsequent increase in the total host popula-tion size, as opposed to by mutation. This scenario is considered inSection 7.1. Importantly, our results do not change qualitativelywith this alternative definition. Many of the consequences ofdisease competition in finite populations can be inferred fromthe behavior of strains with very large R0, and this behavior iscompletely independent of the choice of the initial condition(detailed in Section 7.1).

4. Calculating the fixation probability

The quantity we are interested in is the fixation probabilityFða2;β2; a1;β1;NÞ of a mutant. Let Fn1 ða2;β2; a1;β1;NÞ denote theconditional fixation probability given that the mutant straininvades a population containing exactly n1 individuals infectedwith the resident strain. By the law of total probability, theunconditional fixation probability is given by

Fða2;β2; a1;β1;NÞ ¼ ∑N�1

n1 ¼ 0ρn1

ðR1;NÞFn1 ða2;β2; a1;β1;NÞ: ð4Þ

The most straightforward way to calculate the conditionalfixation probabilities is to consider the embedded Markov chaincorresponding to the continuous-time Markov process defined byEq. (1). This Markov chain has the same state space as the originalprocess, and the transition probabilities between these states aregiven by

Pð½n1;n2�-½n1þ1;n2�Þ ¼β1n1

N�n1�n2

N

a1n1þa2n2þðβ1n1þβ2n2ÞN�n1�n2

N

;

Pð½n1;n2�-½n1�1;n2�Þ ¼a1n1

a1n1þa2n2þðβ1n1þβ2n2ÞN�n1�n2

N

;

Pð½n1;n2�-½n1;n2þ1�Þ ¼β2n2

N�n1�n2

N

a1n1þa2n2þðβ1n1þβ2n2ÞN�n1�n2

N

;

Pð½n1;n2�-½n1;n2�1�Þ ¼ a2n2

a1n1þa2n2þðβ1n1þβ2n2ÞN�n1�n2

N

:

ð5Þ

First, let us adjust this Markov chain by changing the statespace (see Fig. 1A for the graphical representation). We exclude thestate ½0;0�, and we identify all states with n2 ¼ 0 as one absorbingstate, and all states with n1 ¼ 0 as another one. This reduction instates does not change the outcome of the process since once astrain becomes extinct, it cannot reappear in the population. Thenext step is to enumerate all the possible states ½n1;n2� togetherwith these two absorbing states. Constructing the matrix Q and avector b such that Qi;j is the transition probability from a transientstate i to a transient state j, and bi is the transition probability froma transient state i to the absorbing state n1 ¼ 0, we can calculatethe vector f of probabilities of absorption in the state n1 ¼ 0starting in a state i by solving the linear equation (see Kemenyand Snell, 1976)

ðI�Q Þf ¼ b; ð6Þwhere I is the identity matrix.

The conditional fixation probability Fn1 ða2;β2; a1;β1;NÞ is equalto the element of f that corresponds to the state ½n1;1�. Solving Eq.(6) requires the inversion of an ðN2�NÞ=2�dimensional matrix

J. Humplik et al. / Journal of Theoretical Biology 360 (2014) 149–162152

Page 5: Evolutionary dynamics of infectious diseases in …ped.fas.harvard.edu/.../1-s2.0-s0022519314003907-main.pdfEvolutionary dynamics of infectious diseases in finite populations Jan

and, even for moderate N, the resulting formulas are too compli-cated to be of any use. Therefore, what remains is to start withcertain strain parameters, and numerically solve Eq. (6). Since thequasi-stationary distribution can be calculated numerically as well,we can also obtain the exact fixation probability Fða2;β2; a1;β1;NÞfor given parameters.

For very large Ri, we can obtain a closed form expression for thefixation probability given by

limRi-1

Fða2;β2; a1;β1;NÞ

¼ðN�1Þ β1

β2

a2a1

�1� �2

N�1þa2a1

�Nβ1

β2

a2a1

þ β1

β2

a2a1

� �N

1�Nβ2

β1þa2a1

ðN�1Þ� � : ð7Þ

Derivation of this equation is given in Appendix A. It assumesthat when taking the limit Ri-1, the ratios a2=a1 and β2=β1 arekept constant.

Although strains with very large basic reproductive ratios arenot necessarily realistic, they demonstrate the effects of finitepopulations and will be used to present analytic results through-out the paper. There are several important properties of this limitthat simplify the analysis. For a single infection in the limit Ri-1,the quasi-stationary distribution is non-zero only for ni ¼N. Themean time to extinction from quasi-stationarity then follows froma standard result for the SIS model (Ovaskainen, 2001; Kryscio andLefévre, 1989)

τ¼ 1ai

ðN�1Þ!NN RN�1

i 1þO 1Ri

� �� �: ð8Þ

Therefore, strains with very large basic reproductive ratio exist intheir quasi-stationary state essentially forever, resembling thebehavior of infinite populations. Once a second infection isintroduced, every individual in the population is always infectedwith one of the strains. However, the relative prevalence of thetwo strains can change over time. Even when each individual isinfected, the probability of recovery is non-zero for every indivi-dual, but, with probability one, any such recovery event is followedby a transmission event. Although the whole population is alwaysinfected, every transition between states is accompanied by aninfinitesimally short visit to a state with one uninfected host, andso the state space of the population is effectively the set of the reddots in Fig. 1A.

5. Evolutionary invasion analysis

5.1. Definitions

Each strain is characterized by its turnover and transmissionrates. Since these are positive numbers, the set of all possiblemutants, i.e., the phase space of selection, is the first quadrant ofR2. For a given resident strain, we will refer to the fixationprobability as a function on this phase space as the fixationprobability landscape of the resident.

The first question we want to ask is the following: given aresident strain 1, what is the set of all strains that are favored byselection to invade it? By definition, this set consists of all strainswhose fixation probability is greater than that of strain 1's neutralmutant. The conditional fixation probability of a neutral mutant ofs1 is equal to 1=ð1þn1Þ. Therefore, the fixation probability of aneutral mutant is

FneutralðR1;NÞ ¼ ∑N�1

n1 ¼ 0

11þn1

ρn1ðR1;NÞ: ð9Þ

The boundary that separates the successful and unsuccessfulmutants is given by all the neutral mutants of the strain 1. We willrefer to this boundary as the neutral invasion curve of the resident.For a fixed a1 and β1, it is defined implicitly in the a2�β2 plane bythe equation

Fða2;β2; a1;β1;NÞ ¼ FneutralðR1;NÞ: ð10ÞAnother question we are concerned with is whether knowing

that invasion of strain 2 into strain 1 is favored by selection impliesthat invasion of strain 1 into strain 2 is opposed by selection(and vice versa). For a given resident strain 1, the set of mutantsthat do not satisfy this property is defined by two boundarycurves. One of them is the neutral invasion curve of strain 1(corresponding to the premise of the implication), and the other isa curve of all mutants such that strain 1 is their neutral mutant(corresponding to the conclusion of the implication). We will referto this second curve simply as the second boundary curve, and, forfixed strain 1 parameters, it is defined implicitly in the a2�β2plane by

Fða1;β1; a2;β2;NÞ ¼ FneutralðR2;NÞ; ð11Þwhere R2 also depends on β2 and a2.

Throughout the rest of this section we carry out this analysis forseveral distinct regions of the model parameters.

5.2. Invasion analysis if Ri41 and N-1

In very large populations, strain 1 spreads to eventually reach asteady state in which the fraction of infected individuals is equal to1�1=R1. Therefore, when strain 2 enters the population, it seesN=R1 uninfected hosts. Its basic reproductive ratio with respect tothis smaller population is R2=R1, and so it has a non-zero prob-ability of spreading in this population only if R24R1. If satisfied,this probability of spreading is equal to 1�R1=R2 (derived from theextinction probability for the corresponding birth-death process,see Kendall, 1949; Iwasa et al., 2004). As the number of individualsinfected with strain 2 increases, strain 1 will be driven toextinction. Therefore, assuming Ri41, the fixation probability inthis limit is

limN-1

Fða2;β2; a1;β1;NÞ ¼0 if R2oR1

1�R1

R2if R24R1

:

8<: ð12Þ

When N-1 the fixation probability of a neutral mutant is 0,and so the neutral invasion curve is simply given by β2 ¼ R1a2. Itcoincides with the second boundary curve. If strain 2 is favored toinvade strain 1, strain 1 is not favored to invade strain 2. Therelation R1 ¼ R2 divides the phase space into equivalence classes,and two strains can coexist in a population only if they belong tothe same equivalence class, i.e., if they are neutral mutants of eachother. Fig. 2A shows a contour plot of the fixation probabilitylandscape of Eq. (12) together with the (overlapping) neutralinvasion curve.

5.3. Invasion analysis if N is finite and Ri-1

When the population size is finite but the basic reproductiveratio is very large, the quasi-stationary distribution for the residentstrain is non-zero only for n1 ¼N. As a result, the fixationprobability of a neutral mutant is 1=N (Eq. (9)), independent ofthe resident strain. For other mutants, the fixation probability hasa closed form expression given by Eq. (7).

Fig. 2B,C shows contour plots of the fixation probability land-scape together with the corresponding neutral invasion andsecond boundary curves for two different values of N. Comparingthese results with the one in Fig. 2A, we can immediately infer

J. Humplik et al. / Journal of Theoretical Biology 360 (2014) 149–162 153

Page 6: Evolutionary dynamics of infectious diseases in …ped.fas.harvard.edu/.../1-s2.0-s0022519314003907-main.pdfEvolutionary dynamics of infectious diseases in finite populations Jan

several new features of the evolutionary competition that are notpresent if the population is infinite. Firstly, a mutant can befavored to invade a resident even if its basic reproductive ratio islower than that of the resident (the neutral invasion curves dipsbelow the diagonal in Fig. 2B,C; below line resolution in 2B).Therefore, the basic reproductive ratio may not be maximized byselection in a finite population. We will return to this problem inSection 5.4 when we deal with general values of Ri since in thatcase, this effect has more profound consequences.

A second emergent effect is that the neutral invasion curve maybe located to the left of the second boundary curve. The region inthe phase space bounded by these two curves consists of allmutants which are not favored to invade strain 1, nor is strain1 favored to invade them. This creates a sort of “mutual exclusion”or “status quo” situation between the strains. The existence ofsuch “mutually excluding” strains also implies that, unlike ininfinite populations, being a neutral mutant of a strain does notdefine an equivalence relation on the set of all strains. As aconsequence, it is impossible to assign a value to each strain whosemaximization would determine the likely outcome of an invasion in afinite population.

5.4. Invasion analysis in the case of general N and Ri

5.4.1. Location of the neutral invasion curve and its implicationsFor general values of the population size and basic reproductive

ratio, the fixation probability landscape must be calculatednumerically. Several examples, together with the correspondingneutral invasion and second boundary curves, are shown in Fig. 3.Unlike in Fig. 2 (Ri-1), we do not use the normalized turnoverand transmission rates (a2=a1, β2=β1) because the fixation prob-ability is no longer a function of only these fractions. However, byreparametrizing time in Eq. (1), we can still reduce the number ofrelevant parameters by one (usually, this is achieved by settinga1 ¼ 1Þ.

In addition to the conclusions made in Section 5.3, the finite-ness of Ri is responsible for several new features. Firstly, becausethe mean number of infected individuals in quasi-stationarity is anincreasing function of R0, the fact that invasion of a mutant withR2oR1 can be favored by selection also means that selection canfavor the invasion of a novel strain which is better for the hostpopulation. This is in contrast to infinite populations where afavored invader must have a basic reproductive ratio larger than

that of the resident, and therefore such a selective sweep mustnecessarily lead to a higher fraction of infected individuals. Theextreme case of this behavior is that, given a resident strain, theremay exist mutants which cannot even infect other individuals (i.e.,which have β2 ¼ 0), and yet still be favored by selection to invade.This is evident from the non-zero x-intercepts in Fig. 3A and B.This odd behavior occurs if the timescale of the resident epidemicis sufficiently shorter than the mean turnover time of the mutant.However, this may not be possible in all scenarios because themean turnover time has an upper bound given by 1=u (where u isthe natural death rate of hosts).

Secondly, for finite population sizes and finite basic reproduc-tive ratios, selection can also favor faster times to extinction fromquasi-stationarity (Fig. 4). This is again an example of evolutiontowards strains that are better for the host population.

5.4.2. Mutual invasibility and mutual exclusionWhen Ri is finite, Fig. 3 demonstrates that there may exist

strains that are favored to invade strain 1, but strain 1 is alsofavored to invade them (the second boundary curve is left of theneutral invasion curve). Roughly speaking one can imagine this“mutual invasibility” situation as follows. When the infectionoccurs, there is a pressure towards replacing the old infection inthe population. However, when this new infection has almostcompletely spread, i.e., there is only one individual infected withthe old strain, the pressure is reversed towards spreading of theold infection again. Although intuitive, any such classification iscomplicated in finite populations since the success or failure of aninvasion is probabilistic, and the fixation probability cutoff forbeing “favored by selection” is only a definition relative to theneutral mutant.

Using numerical calculations, we can give a qualitative descrip-tion of the relative positions of the neutral invasion curve and thesecond boundary curve, hence characterizing the mutual invasi-bility properties of different pairs of strains. For simplicity, we willrestrict ourselves to pairs of strains with Ri41. Without loss ofgenerality, we will also assume that β24β1 (mutual invasibility isa symmetric relation, and so invasion behavior for pairs of strainswith β2oβ1 can be inferred by switching the identities of strains1 and 2). By definition, the neutral invasion curve and the secondboundary curve coincide at the resident strain s1 where they caneither cross (Fig. 3A), or touch (Fig. 3B,C). In addition to this point

0 1 2 3 4 50

1

2

3

4

5

a2 / a1

β 2 / β

1

N →∞

1 2 3 4 50

1

2

3

4

5

a2 / a1

β 2 / β

1

N = 10, Ri →∞

0

β 2 / β

1

N = 4, Ri →∞

1 2 3 4 50

1

2

3

4

5

a2 / a1

000.10.20.30.40.50.60.70.80.91

Fig. 2. Fixation probability landscapes in the following cases: (A) N-1; (B) N¼10, Ri-1; C. N¼4, Ri-1. Contour plots show the fixation probability (red¼1, blue¼0) for astrain 2 ðβ2 ; a2Þ invading strain 1 ðβ1; a1Þ. The strains to the left of the neutral invasion curve (solid white line) are favored to invade strain 1 (fixation probability greater thanthat of a neutral mutant), while the strains to the right are not. Strains in between the neutral invasion curve and second boundary curve (dashed white line) are not favoredto invade strain 1, but strain 1 is also not favored to invade them. In the limit N-1 these two curves coincide, and the strain with larger basic reproductive ratio ðRi ¼ βi=aiÞis always favored for invasion. Note that because of the special parameter dependence of the fixation probability in either regime Ri-1 or N-1, these plots areindependent of the choice of the resident strain. (For interpretation of the references to color in this figure caption, the reader is referred to the web version of this article.)

J. Humplik et al. / Journal of Theoretical Biology 360 (2014) 149–162154

Page 7: Evolutionary dynamics of infectious diseases in …ped.fas.harvard.edu/.../1-s2.0-s0022519314003907-main.pdfEvolutionary dynamics of infectious diseases in finite populations Jan

of contact, the curves can also cross in the region β24β1 (Fig. 3B).There is at most one such additional crossing, and its presencedepends on the value of R1. For large values of R1, there is nocrossing, and all strains with β24β1 in between the neutralinvasion curve and the second boundary curve are such thatneither s2 nor s1 are favored to invade the other (“mutualexclusion”, Fig. 3C). Lowering R1 beyond a certain threshold value,a crossing emerges in the region β24β1 (Fig. 3A (out of plotrange), B). Below this crossing but above the resident, there is a setof strains s2 such that s2 and s1 are mutually invasible. Above thecrossing, mutual exclusion remains.

6. Weak selection limit

In this section we study long-term evolutionary dynamics inthe scenario where each mutation induces only small changes inthe strain parameters. This assumption may be violated in parti-cular biologically relevant scenarios, such as within-host evolutionin the presence of strain-specific immune responses, where apoint mutation may result in complete immune escape and a largefitness benefit. The ideal framework for studying evolution in this“weak selection” limit is adaptive dynamics (Geritz et al., 1998;Diekmann, 2004), with the fixation probability (Eq. (4)) assumingthe role of the invasion fitness (Proulx and Day, 2002).

6.1. Selection gradient

By definition, the selection gradient points from the residentstrain (a, β) to the strain with the maximal fixation probability inits infinitesimal neighborhood. Formally, it is given by

g!ða;β;NÞ ¼∇Fða0;β0; a;β;NÞjða0 ;β0 Þ ¼ ða;βÞ: ð13Þ

The operator ∇ represents the gradient with respect to the firsttwo coordinates. We are interested only in the direction of g!, notits magnitude. The first thing to note is that this gradient dependsonly on the parameters of the resident strain and on the popula-tion size. This allowed us to drop the index “1” used to distinguishthe resident strains. Also, as noted before, we can always set a¼1without any loss of generality. Therefore, for a given populationsize, strains with the same basic reproductive ratio have the samedirection of selection.

The vector field of Eq. (13) determines a dynamical system thatwe will refer to as selection dynamics. Assuming it is possible, onecould integrate g!ða;β;NÞ to obtain a local fitness function. Theselection dynamics act to maximize this function. The local fitnessfunction is not unique but its contour lines are (this is also true inan infinite population since the transformation R0-f ðR0Þ, wheref(x) is a strictly monotone function, will not change the outcome of

00.10.20.30.40.50.60.70.80.91

00

N = 15, R1 = 5

0

10

20

50

60N = 5, R1 = 4N = 10, R1 = 1.5

1 2 3 40

2

4

6

8

10

a2 a2 a2

00

5

10

15

20

1 2 3 4 1 2 3 4

40

30

5

β 2

β 2β 2

Fig. 3. Fixation probability landscape for finite Ri and different population sizes. (A) N¼10, R1 ¼ 1:5 and (B) N¼5, R1 ¼ 4, and (C) N¼15, R1 ¼ 5. Contour plots show thefixation probability (red¼1, blue¼0) for strain 2 ðβ2 ; a2Þ invading strain 1 (β1 ¼ R1 ; a1 ¼ 1, marked by white þ). The strains to the left of the neutral invasion curve (solidwhite line) are favored to invade strain 1 (fixation probability greater than that of a neutral mutant), while the strains to the right are not. Only as population size increases isthe strain with larger basic reproductive ratio ðRi ¼ βi=aiÞ always favored for invasion. If the neutral invasion curve is to the right of the second boundary curve (dashed whiteline), the strains in between them are favored to invade strain 1, and strain 1 is also favored to invade them (in A, above the þ; in B, between the þ and the O). If the neutralinvasion curve is to the left of the second boundary curve, the strains in between them are not favored to invade strain 1, but strain 1 is also not favored to invade them (In B,above the O, in C; above the þ). Note that the range of the axes is not the same in each plot. (For interpretation of the references to color in this figure caption, the reader isreferred to the web version of this article.)

1

10

102

103

104

β 2

1 2 3 40

2

4

6

8

10

a2

1.7

4.2

7.923

Fig. 4. Contour plot of the mean time to extinction from quasi-stationarity in theSIS model with N¼5. Time is measured in multiples of 1=a1 ¼ 1. Dashed white lineshighlight mean extinction time contours that include the strains with basicreproductive ratios 0.7, 2, 3, and 5, and turnover rate a1 ¼ 1. Overlaid solid linesshow the neutral invasion curves for these strains. Selection can favor a strain withshorter mean time to extinction.

J. Humplik et al. / Journal of Theoretical Biology 360 (2014) 149–162 155

Page 8: Evolutionary dynamics of infectious diseases in …ped.fas.harvard.edu/.../1-s2.0-s0022519314003907-main.pdfEvolutionary dynamics of infectious diseases in finite populations Jan

selection). These contour lines satisfy the differential equation

dβðaÞda

¼ �∂Fða0;β0

; a;βðaÞ;NÞ∂a0

∂Fða0;β0; a;βðaÞ;NÞ∂β0

��������ða0 ;β0 Þ ¼ ða;βðaÞÞ

: ð14Þ

In the special case R0 ¼ β=a-1, this equation can be solvedanalytically. Using the expression (7) on the right hand side ofEq. (14), we obtain the differential equation

dβada

¼ N � 1N � 2

� �βaa: ð15Þ

This equation can be solved to give

βðaÞ ¼ β1aa1

� �ðN�1Þ=ðN�2Þ; ð16Þ

where we have chosen to describe a contour line that passesthrough the strain ða1;β1Þ. One can deduce from the form of thesecontours that when R0 is very large, a possible choice for the localfitness function is

eR0ða;β;NÞ ¼β

aðN�1Þ=ðN�2Þ: ð17Þ

As expected, limN-1eR0ða;β;NÞ ¼ R0.For strains with generic values of R1, the corresponding contour

lines can be solved for numerically and are more complicated(Fig. 5). They do not pass through the origin, and they are notconvex.

For any R0 value, the major difference between the selectiondynamics in finite versus infinite populations can be understood asfollows. In an infinite population, any contour curve passingthrough a strain with R141 is just a ray with a slope equal toR1. This is not true for contour curves passing through strains withR1o1. The dynamics of these strains can only be described by astochastic model because the number of infected individuals neverreaches macroscopic numbers. The results are that any contourcurve that passes through a strain with R1o1 does not passthrough the origin, is concave, and always lies under the β¼ a line.For finite populations, Fig. 5 demonstrates that the distinctionbetween strains with R1o1 and R141 no longer exists. Anycontour contains strains with all possible values of the basicreproductive ratio. These numerical calculations suggest that forany contour there is some ac (Rc) such that the contour is convex in

the region a4ac (or equivalently for strains with R4Rc). Thethresholds Rc seem to be close to but not equal to 1.

6.2. Rate of convergence towards the N-1 limit

Because the direction of the selection gradient depends only onthe basic reproductive ratio and the population size, it provides agood way to measure the importance of the “finiteness” of thepopulation. For a resident with a basic reproductive ratio R041,we will denote θðN;R0Þ the angle between its selection gradientand the a-axis. For an infinite population, this is simplyarctanð1=R0Þ. To estimate the population sizes for which finite sizeeffects are important, we define the relative difference in selectionpressure as compared to infinite population

δθðN;R0Þ ¼limN-1θðN;R0Þ�θðN;R0Þ

limN-1θðN;R0ÞA ½0;1�: ð18Þ

Furthermore, we define the threshold population size NthðR0Þ as thepopulation size above which δθðN;R0Þo5% (this choice is arbi-trary and was chosen simply to characterize the behavior inregions where an analytic solution is not available). Using Eq.(15), it follows that

limR0-1

δθðN;R0Þ ¼1

N�1: ð19Þ

Consequently, limR0-1NthðR0Þ ¼ 21. Numerical calculations con-firm that the threshold population size increases with decreasingR0. Based on these results we can conclude that, in a population ofN hosts, the relative deviation of the selection pressure from theN-1 case is at least 1=ðN�1Þ. However, for pathogens withsmaller R0, this error can be significantly larger. In Fig. 6 we showthe N dependence of δθðN;R0Þ together with the correspondingthreshold population sizes for several different values of R0.

6.3. Selection dynamics with a trade-off between the transmissionrate and the virulence

In this section we consider the possibility that the virulence(the disease-induced death rate) and the transmission rate of astrain might not be independent parameters. For simplicity, weassume no recovery (δi ¼ 0), so that ai ¼ uþvi. We continue toassume that death is immediately followed by replacement with asusceptible host. The curve that describes all the allowed combi-nations of virulence and transmission rates will be denoted as βðvÞ.

0 2 4 6 8 100

10

20

30

40

a

β

N=5

0 2 4 6 8 100

10

20

30

40

a

β

N=10

0 2 4 6 8 100

10

20

30

40

a

β

N=30

Fig. 5. Contour lines for the local fitness function of an infectious disease in a finite population. The direction of selection of a disease strain with parameters a and β isperpendicular to these curves. (A) N¼4. (B) N¼10. (C) N¼30. The selection gradient is determined by the fixation probabilities of neighboring strains in the phase space, andthe local fitness function is constructed to have a gradient field pointing in the direction of the selection gradient. The contour lines of this function can be calculated bynumerical solution of Eq. (14). The black line represents the boundary R0 ¼ 1 that separates strains that can (above line) from those that cannot (below line) infect asignificant fraction of the population.

J. Humplik et al. / Journal of Theoretical Biology 360 (2014) 149–162156

Page 9: Evolutionary dynamics of infectious diseases in …ped.fas.harvard.edu/.../1-s2.0-s0022519314003907-main.pdfEvolutionary dynamics of infectious diseases in finite populations Jan

The basic condition on βðvÞ is that it is an increasing function. Thisconvention follows from an extensive literature on the evolution ofvirulence (reviewed in Dieckmann et al., 2002), and arises fromthe idea that increasing pathogen numbers within an individualhost can lead to increasing virulence (v) but also lead to anincrease in transmission rate (β).

Considering again the weak selection limit (and hence infini-tesimal mutations), the selection dynamics can be described byrestricting the local fitness function to the line βðvÞ. Singular pointsof the selection dynamics correspond to critical points of the localfitness function. Geometrically, a strain with virulence vn is asingular point if the direction of selection at vn is perpendicular tothe curve βðvÞ at the point vn. This is equivalent to the conditionthat the curve βðvÞ at the point vn shares a tangent with thecontour curve of the local fitness function that passes through thestrain (vn;βðvnÞ). Stability properties of the singular points followfrom relative curvatures of βðvÞ and the contour curve (seeMazancourt and Dieckmann, 2004 for a general analysis of thisgeometrical approach). A short calculation shows that this isequivalent to a standard second derivative test of the local fitnessfunction. Therefore, the local attractors of the selection dynamics(also referred to as convergence stable strategies) correspond tothe virulences, vn, that locally maximize the local fitness function.

A convergence stable strategy can be either an evolutionarilystable strategy, or an evolutionary branching point (Geritz et al.,1998). In the limit of either R0-1 or N-1, mutual invasibilitybetween two neighboring strains is not possible, and so, in thesecases, evolutionary branching is ruled out.

In an infinite population, the convergence stable strategies aresimply local maxima of R0ðvÞ ¼ βðvÞ=ðuþvÞ (Fig. 7A,B; dashedlines). Therefore, taking first and second derivatives, the necessaryconditions for vn to be a convergence stable strategy are

β0ðvnÞ ¼ βðvnÞuþvn

¼ R0ðvnÞ; and β″ðvnÞr0: ð20Þ

In particular, if βðvÞ has a positive second derivative, the selectionpressure either increases the virulence towards infinity or decreasesit towards zero, and there cannot be any convergence stablestrategy with an intermediate vn. An example is depicted graphi-cally in Fig. 7 (compare the concave constraint in A and B, with one

attractor for N-1, to the convex constraint in C and D, with noattractors for N-1).

To quantitatively analyze the situation in a finite population,we assume the limit Ri-1, and use the local fitness function (17)given by

eR0ðvÞ ¼βðvÞ

ðuþvÞðN�1Þ=ðN�2Þ:

By finding the maxima of eR0ðvÞ, we determine that the necessaryconditions for vn to be a convergence stable strategy are

β0ðvnÞ ¼ ðN�1ÞðN�2Þ

βðvnÞðuþvnÞ ¼

ðN�1ÞðN�2ÞR0ðvnÞ; and

β″ðvnÞr ðN�1ÞðN�2Þ2

βðvnÞðuþvnÞ2

: ð21Þ

This result introduces two interesting conclusions about theevolution of virulence in finite populations. Firstly, non-zero localattractors can exist even when βðvÞ has a positive curvature (isconvex) (Fig. 7C,D). Geometrically, the inequality in Eq. (21) meansthat the second derivative of the constraint βðvÞ at vn must be lessthan or equal to the second derivative of the contour curve of thelocal fitness function passing through ðvn;βðvnÞÞ. Therefore, aconstraint that might continuously increase virulence in an infinitepopulation can lead to a finite convergence stable virulence in afinite population.

Secondly, for constraints βðvÞ with negative curvature, where alocal attractor already exists in an infinite population, the locationof the attractor changes depending on the population size. Wedemonstrate this as follows. Denote vn as the local attractor in aninfinite population, i.e., it satisfies the conditions (20). To obtainanalytic results, we assume that ðN�1Þ=ðN�2Þ�1¼ 1=ðN�2Þ issmall. We also assume that the local attractor vn in a finitepopulation is given by a solution of Eq. (21). Since 1=ðN�2Þ issmall, it is expected that the new attractor vn is very close to vn.Linearizing this equation by making a Taylor expansion around vn

on both sides, and keeping only terms linear in v�vn, wedemonstrate that

vn ¼ vn� βðvnÞβðvnÞuþvn

þðN�2ÞðuþvnÞjβ″ðvnÞj� vn� 1

ðN�2ÞR0ðvnÞjβ″ðvnÞj ; ð22Þ

where we used the condition β″ðvnÞo0. Therefore, finite popula-tions shift the convergence stable strategies towards lower virulence.The geometric interpretation of this result is shown in Fig. 7A,B.

The fact that this expectedly small correction is proportional toR0ðvnÞ, when we consider R0 to be infinite, is not inconsistent whenconsidering a specific method of attaining this limit. One can thinkabout the R0-1 limit as setting βðvÞ ¼ cβref ðvÞ, and taking c-1(where βref ðvÞ is some finite reference value of the transmissionrate). Then, the rightmost term in Eq. (22) is manifestly finite sinceit is proportional to βðvnÞ=jβ″ðvnÞj which is independent of c. Onecould also mediate the R0-1 limit by sending the turnover rate(uþv) to zero. This method is additionally complicated by the factthat the derivative is taken with respect to virulence, however, byemploying the same strategy and expressing the turnover rate interms of some finite reference, one can verify that Eq. (22) is againindependent of the scaling factor.

Returning to the case of constraints with positive curvature, weconsider what happens to the attractors as the population size iscontinually reduced from infinity to small values. We determinethat for certain constraints, after this local attractor appears at aparticular population size, it can then bifurcate into two new localattractors. This feature is demonstrated in Fig. 7E using the convex

2 10 20 30 40 50 60 70 800

0.2

0.4

0.6

0.8

1

Population size (N)

Cha

nge

in d

irect

ion

of s

elec

tion

δ θ(

N, R

0 )

R0 = 1.5, Nth = 75R0 = 2, Nth = 43R0 = 3, Nth = 31

R0 = 5, Nth = 26R0 = 10, Nth = 23

R0 = 4, Nth = 28

Fig. 6. Dependence of the selection gradient on the population size. The directionof selection (θðN;R0Þ) for an infectious disease in a finite population depends on thepopulation size (N) and the basic reproductive ratio of the resident strain (R0). Theangle θ is measured in a counterclockwise direction (θ¼ 0 corresponds to selectiontowards lower a). The coefficient δθðN;R0Þ is defined as the relative change in thedirection of selection due to the finite size N of the population. The thresholdpopulation size (Nth) is defined as the population size above which δθðN;R0Þo5%.

J. Humplik et al. / Journal of Theoretical Biology 360 (2014) 149–162 157

Page 10: Evolutionary dynamics of infectious diseases in …ped.fas.harvard.edu/.../1-s2.0-s0022519314003907-main.pdfEvolutionary dynamics of infectious diseases in finite populations Jan

piecewise linear constraint

βðvÞ ¼ 2:6v if vo3:786�5:3þ4v if v43:786

:

(ð23Þ

While unnecessary for our demonstration, one could smoothenthis constraint for example by replacing the sharp step by a Hillfunction. If the constraint is concave, there can be at most oneattractor for any population size, and hence no bifurcation occurs.

Although we have analyzed the finite population case using thelimit Ri-1, based on the shape of the contour curves (see Fig. 5), weexpect similar behavior in the general case as well. However, becausemutual invasibility is possible for certain strains, some local attractorsmight turn out to be evolutionary branching points.

7. Alternative definitions

7.1. Alternative initial conditions

The fixation probability of a mutant is dependent upon thechoice of the initial condition for Eq. (1). Throughout the paper, wemade the choice (3) that a novel strain was introduced into thepopulation by a process that converted an individual infected witha resident strain to one infected with a mutant strain. Populationsize is preserved in this process, and it corresponds to conventionsused in evolutionary invasion analysis. An example of a differentbut equally important initial condition is described by the follow-ing scenario. The resident infection s1 has reached quasi-stationarity in a population of N�1 hosts, and then an individualinfected with a novel strain s2 migrates to the population. Since

0 2 4 6 8 100

5

10

15

20

a = u + v

contour for N = 4

constraint β(v)

β β

0 1 2 30

2

4

6

8

10

a = u + v

constraint β(v) contour for N = ∞contour for N = 5

0 5 10 15 200

1

2

3

4

v

loca

l fitn

ess

N = 3N = 4N = 7N = ∞

constraint: Eq(23) 5

0 2 4 6 8 100

0.2

0.4

0.6

0.8

1

v

loca

l fitn

ess

N = 4N = ∞

constraint: β(v) = cvk

0 1 20

0.5

1

1.5

2

2.5

3

v

loca

l fitn

ess

N = 5N = ∞

constraint: β(v) = cv/(1+v)

Fig. 7. Selection dynamics with a trade-off between the transmission rate and the virulence. In all figures it is assumed that the baseline death rate u¼1, and that therecovery rate δ¼0. Black diamonds mark different local attractors. (A) The local fitness function (R0 ; eR0) versus the virulence, v, for the constraint βðvÞ ¼ cv=ð1þvÞ, with c¼10.There is at most one local attractor for a concave constraint. The convergence stable virulence vn shifts to lower values when the population size is decreased from N ¼1(dashed line) to N¼5 (solid line). (B) The convergence stable virulence occurs at the value of v where the contour line of the local fitness function (blue line) shares a tangentwith the constraint βðvÞ (green line), so that the direction of selection is perpendicular to the constraint. (C) The local fitness function (R0; eR0) versus the virulence, v, for theconstraint βðvÞ ¼ cvk , with c¼1.137 and k¼1.2. There is no convergence stable strategy for N ¼1 (dashed line). When N decreases, a convergence stable strategy appears(solid line). D. The convergence stable virulence occurs at the value of v where the contour line of the local fitness function (blue line) is parallel to the constraint βðvÞ (greenline), so that the direction of selection is perpendicular to the constraint. In finite populations, the point of contact is a local attractor only if the second derivative of theconstraint at this point is less than the one of the contour. E. The local fitness function (R0 ; eR0) versus the virulence, v, for the convex constraint in Eq. (23). For N ¼1, thisconstraint does not lead to a finite convergence stable virulence. As the population size is decreased, a single local attractor emerges, and when the population size reachesN¼4, this attractor bifurcates into two new local attractors. At N¼3, one of these disappears again. (For interpretation of the references to color in this figure caption, thereader is referred to the web version of this article.)

J. Humplik et al. / Journal of Theoretical Biology 360 (2014) 149–162158

Page 11: Evolutionary dynamics of infectious diseases in …ped.fas.harvard.edu/.../1-s2.0-s0022519314003907-main.pdfEvolutionary dynamics of infectious diseases in finite populations Jan

the per contact transmission rate of the resident in a population ofN individuals is β1=N, its transmission rate in a population of N�1individuals is β1ðN�1Þ=N. Therefore, in this scenario, the prob-ability that initially there are n1 individuals infected with theresident strain is

ρ0n1 ðR1;NÞ �Qn1 R1

N�1N

;N�1� �

; n1A 1;…;N�1f g: ð24Þ

Luckily, there are no qualitative differences in the behavior ifone decides to use this initial condition instead of the standardone. In particular, in the limit Ri-1, both initial conditions reduceto the same distribution. This is not surprising. The initial condi-tion with N�1 individuals infected with the resident and 1 indi-vidual infected with the mutant is the only meaningful choicewhen the limit Ri-1 is assumed because any non-infectedindividual must be immediately infected. That this limit eliminatesthe ambiguity in the choice of the initial condition is an importantfact that further strengthens the view of the large Ri limit as anideal indicator of the stochastic effects.

7.2. Alternative invasion conditions

Throughout this paper we have been using a definition ofneutrality based on the fixation probability of a mutant that isindistinguishable from the resident. While this definition has aclear biological interpretation, it is definitely not the only possiblechoice. A different choice could be based on a comparison of timesthe population spends infected with strains 1 and 2 if mutationsbetween the strains occur at some rate u, and the timescale ofmutation is much larger than the timescale of invasion. We couldsay that the strains are neutral mutants of each other if these timesare equal. In the limit Ri-1, the total mutation rate of apopulation is strain-independent because every individual isalways infected, and hence is a source of a possible mutation. Thisallows us to conclude that the times the population spendsinfected with each disease are equal if and only if the fixationprobability of strain 1 invading strain 2 is the same as the fixationprobability of the reverse process. Using Eq. (7), this happens ifand only if

βN�22

aN�12

¼ βN�21

aN�11

: ð25Þ

Therefore, at least in the large Ri limit, the “invasion curves”corresponding to this alternative definition of neutrality wouldcoincide with the contour curves of the weak selection limit.In particular, neutrality would define an equivalence relation, andwe would not need to worry about the existence of pairs of strainsmutually not favored to invade each other. Although this is a niceproperty for a neutrality definition to satisfy, we believe that ouroriginal approach is more fundamental, hence we do not pursuethis alternative any further.

8. Discussion

We have analyzed the competition for hosts between twopathogen strains, each of them obeying the dynamics of astochastic SIS model. We examined whether invasion of a mutantstrain into a resident strain at endemic equilibrium is favored byselection by comparing its fixation probability to that of a neutralmutant. This is one of the simplest models of evolutionarycompetition of infectious diseases, and yet, our results violatethe standard rule of R0 maximization observed for this samemodel in infinite populations. We find that strains with R0 valueslower than the resident strain can be favored to invade. In fact, weshow that in finite populations, there exist pairs of strains such

that neither of them is favored by selection to invade the other(“mutual exclusion”), or that both are favored to invade the other(“mutual invasibility”). This implies that invasion of new patho-gens in finite populations cannot be described simply by compar-ing R0 values (or values of any other function) of each strain.

Since, for any strain, one can find a mutant with a smaller R0value that is favored to invade, evolution can lead to a healthierhost population. This effect is impossible in infinite populations forsimple SIS models. Though related, these results are not a simpleconsequence of the dependence of the number of secondaryinfections on population size (Ross, 2011), but depend on longer-term behavior associated with disease competition.

If we restrict our analysis to mutants whose parameters areinfinitesimally close to those of the resident, the situation simplifiesconsiderably. In this weak selection limit, we can locally define afitness function that is maximized by selection. In the limit of R0-1,this local fitness function is equal to eR0ða;β;NÞ ¼ R0 ð1=aÞ1=ðN�2Þ

(Eq. (17)). Although the assumption R0-1 is not applicable tomost pathogens, understanding it is important for three reasons.First, the results in this limit are independent of the choice of theinitial condition. Second, it provides a bound for the behavior inthe case of general R0. Specifically, in a finite population ofN hosts, the relative deviation of the selection pressure from theone given by R0 maximization (δθðN;R0Þ) is at least 1=ðN�1Þ.Third, the form of the results is very simple, and as such it isuseful for a qualitative understanding of the effects of finitepopulations. If one studies a model based on R0 maximization,and is interested in what happens when the population size isreduced, replacement of R0 with eR0 should provide a first insightinto the stochastic behavior.

The interpretation of the form of eR0 is simple. The smaller thepopulation size, the more important it is to have a large turnovertime rather than a large transmission rate. This is in opposition toinfinite populations, where there is a symmetry between thesetwo attributes, and both are equally important. That turnover timematters more in small populations is also demonstrated by theresult that, when the transmission rate depends on the virulence,lowering the population size also lowers the evolutionarily stablevirulence. Intuitively, slower turnover (longer lifespan, lowerrecovery rate, or lower virulence) is more important than a highertransmission rate in small populations because it allows a strain tobeat a competitor not only by infecting more individuals, but –

because stochastic extinction is inevitable – by “waiting it out.”A natural generalization of our model is to consider a finite but

spatially structured population. This will be the focus of futurework, but intuition suggests that any non-trivial geometry shouldmanifest itself in the exponent in Eq. (17) for the R0-1 limit. Thisin turn would have an effect on the lower bound of the relativedeviation from R0 maximization. The ultimate question is whetherfor certain population structures this lower bound fails toapproach zero as the population size goes to infinity. For example,if one imagines a very large population that consists of many smallweakly interacting subpopulations, it is intuitive that some of thefeatures of the small populations should be relevant on themacroscopic scale. Whether and under what conditions this istrue still needs to be investigated. A large body of related work hasused SIR-type models to study the evolution of virulence inparticular types of structured populations (for example, Bootsand Sasaki, 1999; Haraguchi and Sasaki, 2000; Boots et al., 2004;Webb et al., 2013, reviewed in Messinger and Ostling, 2009).

Another necessary step is to incorporate the possibility ofcontinual mutations into the model. Only by studying the resultingmutation-selection balance can one study disease evolution onmore general timescales and relate the model to empirical data.

Finally, there is an interesting similarity between our model in theregime Ri-1 and the Moran process with frequency-dependent

J. Humplik et al. / Journal of Theoretical Biology 360 (2014) 149–162 159

Page 12: Evolutionary dynamics of infectious diseases in …ped.fas.harvard.edu/.../1-s2.0-s0022519314003907-main.pdfEvolutionary dynamics of infectious diseases in finite populations Jan

selection (Nowak et al., 2004). Both models assume a population of Nhosts. In the Moran process, every individual carries one of twoalleles A or B. In the limit Ri-1 of our model, every individual isinfected with one of two strains s1 or s2. In the Moran process, everyevent consists of choosing a random individual for death, and arandom individual for reproduction. In our model, every eventconsists of choosing a random individual for a recovery (or death),and a random individual for infecting the newly recovered host.There are, however, two important differences. First, in the Moranprocess, all individuals have the same probability of dying. This is notso in our model, where the probability of being chosen for recoverydepends on ai. Second, in the Moran process, the events of choosingan individual for death and choosing an individual for reproductionare independent. In our model, the event of choosing an individualwho will infect the newly recovered host depends on whether it wasan individual infected with s1 or s2 who recovered. This seconddifference is the reason the limit Ri-1 of our model cannot bemapped on a standard Moran process with frequency-dependentselection.

Acknowledgments

The authors thank Daniel Rosenbloom, Benjamin Allen, OliverHauser and Gabriel Leventhal for fruitful discussions, as well as theanonymous reviewers for valuable feedback. J.H. received supportfrom the Zdenek Bakala Foundation and the Mobility Fund ofCharles University in Prague. M.A.N. received support from theJohn Templeton Foundation. M.A.N. and A.L.H. received supportfrom the Foundational Questions in Evolutionary Biology Fund.

Appendix A. Fixation probability for residents and mutantswith large basic reproductive ratios

A.1. Quasi-stationary distribution for large basic reproductive ratio

As a first step towards calculating the fixation probability, weneed an expression for the quasi-stationary distribution (2). In theregion of large R1, it behaves as (Ovaskainen, 2001)

Qn1ðR1;NÞ ¼

1� N2

ðN�1Þ1R1

þO 1

R21

!if n1 ¼N

N2

ðN�1Þ1R1

þO 1

R21

!if n1 ¼N�1

O 1

R21

!if n1oN�1

;

8>>>>>>>>>>><>>>>>>>>>>>:ð26Þ

Looking at the transition probabilities of the single-disease SISmodel (i.e., setting n2 ¼ 0 in Eq. (5)), one can gain an intuitiveinsight into this result. The condition of large R1 means that theprobability of infecting an uninfected host is much larger than theprobability that an infected individual recovers (or dies). There-fore, the population soon reaches a state where all individuals areinfected. Every once in a while, an individual is recovered. Thishappens at a rate Na1. However, this newly recovered host isquickly reinfected with the rate β1ðN�1Þ=N. If we observe thesystem for a long time, the fraction of time it spends in the staten1 ¼N�1 is

1β1

NðN�1Þ

1Na1

þ 1β1

NðN�1Þ

� N2

ðN�1Þ1R1

; ð27Þ

which, assuming ergodicity, can be identified with the probabilityof being in the state n1 ¼N�1, explaining the expression (26).Furthermore, this reasoning is valid only if

Pðn1-n1þ1ÞcPðn1-n1�1Þ 8n1Af1;…;N�1g: ð28Þ

This is equivalent to R1cN, giving us a more useful description ofthe region of validity of the approximation (26).

A.2. Derivation of the fixation probability in the Ri-1 limit

Using a similar reasoning as in the case of only one strainpresent, most of the time every host in the population will beinfected. Occasionally, an individual recovers, however, it is quicklyreinfected with either one of the two strains. Therefore, in theregion of large Ri, the state of the population is constrained to thered dots shown in Fig. 1A. More precisely, if we want this to betrue, the transition probabilities from states with n1þn2 ¼N�1must satisfy

Pð½n1;n2�-½n1þ1;n2�ÞcPð½n1;n2�-½n1�1;n2�Þ;Pð½n1;n2�-½n1þ1;n2�ÞcPð½n1;n2�-½n1;n2�1�Þ;Pð½n1;n2�-½n1;n2þ1�ÞcPð½n1;n2�-½n1�1;n2�Þ;Pð½n1;n2�-½n1;n2þ1�ÞcPð½n1;n2�-½n1;n2�1�Þ; ð29Þ

which is equivalent to

R1cN; R1ca2a1NðN�2Þ;

R2cN; R2ca1a2NðN�2Þ: ð30Þ

Note that this is a much more stringent condition than Ri4N.Assuming the conditions (30) are satisfied, the original process

reduces to a Markov chain on a one-dimensional lattice with2N�1 states (the red dots in Fig. 1A). We will assign the numberi¼0 to the state ½n1 ¼N�1;n2 ¼ 0�, i¼1 to ½n1 ¼N�1;n2 ¼ 1�, i¼2to ½n1 ¼N�2;n2 ¼ 1� and so on until i¼ 2ðN�1Þ is assigned to½n1 ¼ 0;n2 ¼N�1�. The states i¼0 and i¼ 2ðN�1Þ are absorbing atn2 ¼ 0 and n1 ¼ 0 respectively. It follows from Eq. (5) that the jumpprobabilities are

Pði-iþ1Þ ¼a1 N� iþ1

2

� �a1 N� iþ1

2

� �þa2

iþ12

;

Pði-i�1Þ ¼a2iþ12

a1 N� iþ12

� �þa2

iþ12

;

9>>>>>>>>>>>=>>>>>>>>>>>;if i is odd ð31Þ

Pði-iþ1Þ ¼β2

i2

β1 N�1� i2

� �þβ2

i2

;

Pði-i�1Þ ¼β1 N�1� i

2

� �β1 N�1� i

2

� �þβ2

i2

:

9>>>>>>>>>>=>>>>>>>>>>;if i is even ð32Þ

In the limit of R1-1, the initial condition (3) is non-zero onlyfor n1 ¼N�1. The fixation probability is equal to the conditionalfixation probability FN�1ða2;β2; a1;β1;NÞ, which is the same as theprobability of absorption at i¼ 2ðN�1Þ starting from the state i¼1.Note that the jump probabilities, and therefore also the absorptionprobabilities, depend only on the ratios a2=a1 and β2=β1. Prob-abilities of absorption from any state satisfy solvable recurrence

J. Humplik et al. / Journal of Theoretical Biology 360 (2014) 149–162160

Page 13: Evolutionary dynamics of infectious diseases in …ped.fas.harvard.edu/.../1-s2.0-s0022519314003907-main.pdfEvolutionary dynamics of infectious diseases in finite populations Jan

equations that yield

FN�1ða2;β2; a1;β1;NÞ ¼1

1þ∑2N�3k ¼ 1 ∏k

i ¼ 1Pði-i�1ÞPði-iþ1Þ

¼ðN�1Þ β1

β2

a2a1

�1� �2

N�1þa2a1

�Nβ1

β2

a2a1

þ β1

β2

a2a1

� �N

1�Nβ2

β1þa2a1

ðN�1Þ� �: ð33Þ

Calculating the conditional fixation probability starting atn1 ¼N�2, n2 ¼ 1,

FN�2ða2;β2; a1;β1;NÞ ¼ 1þPð1-0ÞPð1-2Þ

� �FN�1ða2;β2; a1;β1;NÞ; ð34Þ

one could also take into account the 1=R1 corrections in Eq. (26).Using Eq. (4), we find

Fða1;β1; a2;β2;NÞ

¼N�1þ N2

N�1a2a1

1R1

þO 1Ri

� � !β1

β2

a2a1

�1� �2

N�1þa2a1

�Nβ1

β2

a2a1

þ β1

β2

a2a1

� �N

1�Nβ2

β1þa2a1

ðN�1Þ� � : ð35Þ

In addition to higher order terms, the Oð1=RiÞ correction alsocontains 1=Ri terms that come from a possible jump to the stateson the n1þn2 ¼N�2 diagonal in the state space. We do notevaluate these corrections since, for our purposes, the mostinteresting case is when Ri-1. Then, the fixation probability issimply

limRi-1

Fða2;β2; a1;β1;NÞ

¼ðN�1Þ β1

β2

a2a1

�1� �2

N�1þa2a1

�Nβ1

β2

a2a1

þ β1

β2

a2a1

� �N

1�Nβ2

β1þa2a1

ðN�1Þ� � ð36Þ

References

Anderson, R.M., May, R.M., 1982. Coevolution of hosts and parasites. Parasitology 85(02), 411–426.

Anderson, R.M., May, R.M., 1991. Infectious Diseases of Humans: Dynamics andControl. Oxford University Press, USA.

Antia, R., Levin, B.R., May, R.M., 1994. Within-host population dynamics and theevolution and maintenance of microparasite virulence. Am. Nat. pp. 457–472.

Barbour, A.D., Pollett, P.K., 2010. Total variation approximation for quasi-stationarydistributions. J. Appl. Probab. 47 (December (4)), 934–946, zentralblatt MATHidentifier 05835859, Mathematical Reviews number (MathSciNet) MR2752899.⟨http://projecteuclid.org/euclid.jap/1294170510⟩.

Barbour, A.D., Pollett, P.K., 2012. Total variation approximation for quasi-equilibrium distributions, II. Stoch. Process. Appl. 122 (November (11)), 3740–-3756 ⟨http://www.sciencedirect.com/science/article/pii/S030441491200155X⟩.

Bonhoeffer, S., Nowak, M.A., 1994. Mutation and the evolution of virulence. Proc. R.Soc. Lond. Ser. B: Biol. Sci. 258 (1352), 133–140 ⟨http://rspb.royalsocietypublishing.org/content/258/1352/133.short⟩.

Boots, M., Hudson, P.J., Sasaki, A., 2004. Large shifts in pathogen virulence relate tohost population structure. Science 303 (February (5659)), 842–844 14764881.⟨http://www.sciencemag.org/content/303/5659/842⟩.

Boots, M., Sasaki, A., 1999. ‘Small worlds’ and the evolution of virulence: infectionoccurs locally and at a distance. Proc. R. Soc. Lond. Ser. B: Biol. Sci. 266 (October(1432)), 1933–1938 ⟨http://rspb.royalsocietypublishing.org/content/266/1432/1933.abstract⟩.

Bull, J.J., 1994. Perspective: virulence. Evolution 48 (October (5)), 1423.Castellano, C., Pastor-Satorras, R., 2010. Thresholds for epidemic spreading in

networks. Phys. Rev. Lett. 105 (November (21)), 218701 ⟨http://link.aps.org/doi/10.1103/PhysRevLett.105.218701⟩.

Cator, E., Van Mieghem, P., 2013. Susceptible-infected-susceptible epidemics on thecomplete graph and the star graph: exact analysis. Phys. Rev. E 87 (1), 012811⟨http://pre.aps.org/abstract/PRE/v87/i1/e012811⟩.

Cortez, M.H., 2013. When does pathogen evolution maximize the basic reproductivenumber in well-mixed host-pathogen systems? J. Math. Biol. 67 (December(6–7)), 1533–1585 ⟨http://link.springer.com/article/10.1007/s00285-012-0601-2⟩.

Crow, J.F., Kimura, M., 1970. An Introduction to Population Genetics Theory. BurgessPub. Co., Harper and Row, New York, USA, 656 pp., ⟨http://books.google.at/books?id=MLETAQAAIAAJ⟩.

Dieckmann, U., 2002. Adaptive dynamics of pathogen–host interactions. In:Dieckmann, U., Metz, J.A.J., Sabelis, M.W., Sigmund, K. (Eds.), AdaptiveDynamics of Infectious Diseases: In Pursuit of Virulence Management. Cam-bridge University Press, Cambridge, UK, pp. 39–58.

Dieckmann, U., Metz, J., Sabelis, M., Sigmund, K., 2002. Adaptive Dynamics ofInfectious Diseases: in Pursuit of Virulence Management. Cambridge Studies inAdaptive Dynamics. Cambridge University Press, Cambridge, UK.

Diekmann, O., 2004. A beginners guide to adaptive dynamics. Banach Center Publ.63, 47–86 ⟨http://igitur-archive.library.uu.nl/math/2006-0727-200409/UUindex.html⟩.

Diekmann, O., Heesterbeek, J.A.P., April 2000. Mathematical Epidemiology ofInfectious Diseases: Model Building, Analysis and Interpretation. John Wiley& Sons.

Eames, K.T.D., Keeling, M.J., 2002. Modeling dynamic and network heterogeneitiesin the spread of sexually transmitted diseases. Proc. Natl. Acad. Sci. USA 99(October (20)), 13330–13335 ⟨http://www.pnas.org.ezp-prod1.hul.harvard.edu/content/99/20/13330.abstract⟩.

Ebert, D., Bull, J.J., 2008. The evolution and expression of virulence. In: Stearns, S.C.,Koella, J.C. (Eds.), Evolution in Health and Disease, 2nd edition OxfordUniversity Press, USA, pp. 153–167, January.

Ganesh, A., Massoulie, L., Towsley, D., 2005. The effect of network topology on thespread of epidemics. In: Proceedings IEEE INFOCOM 2005. 24th Annual JointConference of the IEEE Computer and Communications Societies, vol. 2.pp. 1455–1466.

Ganusov, V.V., Antia, R., 2003. Trade-offs and the evolution of virulence ofmicroparasites: do details matter?. Theor. Popul. Biol. 64 (September (2)),211–220 12948682.

Ganusov, V.V., Bergstrom, C.T., Antia, R., 2002. Within-host population dynamicsand the evolution of microparasites in a heterogeneous host population. Int. J.Org. Evol. 56 (February (2)), 213–223 11926490.

Geritz, S., Kisdi, E., Mesze'NA, G., Metz, J., 1998. Evolutionarily singular strategiesand the adaptive growth and branching of the evolutionary tree. Evol. Ecol. 12(1), 35–57.

Haraguchi, Y., Sasaki, A., 2000. The evolution of parasite virulence and transmissionrate in a spatially structured population. J. Theor. Biol. 203 (2), 85–96.

Heffernan, J.M., Smith, R.J., Wahl, L.M., 2005. Perspectives on the basic reproductiveratio. J. R. Soc. Interface 2 (September (4)), 281–293 16849186. ⟨http://rsif.royalsocietypublishing.org/content/2/4/281⟩.

Hofbauer, J., Sigmund, K., May 1998. Evolutionary Games and Population Dynamics.Cambridge University Press, Cambridge, UK.

Iwasa, Y., Michor, F., Nowak, M.A., 2004. Evolutionary dynamics of invasion andescape. J. Theor. Biol. 226 (January (2)), 205–214 ⟨http://www.sciencedirect.com/science/article/pii/S0022519303003333⟩.

Keeling, M., Ross, J., 2008. On methods for studying stochastic disease dynamics. J.R. Soc. Interface 5 (19), 171–181 ⟨http://rsif.royalsocietypublishing.org/content/5/19/171.abstract⟩.

Kemeny, J., Snell, J., 1976. Finite Markov Chains: with a New Appendix “General-ization of a Fundamental Matrix”. Undergraduate Texts in Mathematics.Springer.

Kendall, D.G., 1949. Stochastic processes and population growth. J. R. Stat. Soc. Ser. B(Methodol.) 11 (January (2)), 230–282 ⟨http://www.jstor.org/stable/2984078⟩.

Kryscio, R.J., Lefévre, C., 1989. On the extinction of the S-I-S stochastic logisticepidemic. J. Appl. Probab. 26 (4), 685–694.

Lenski, R.E., May, R.M., 1994. The evolution of virulence in parasites and pathogens:reconciliation between two competing hypotheses. J. Theor. Biol. 169 (August(3)), 253–265 ⟨http://www.sciencedirect.com/science/article/pii/S0022519384711465⟩.

Levin, B.R., 1996. The evolution and maintenance of virulence in microparasites.Emerg. Infect. Dis. 2 (2), 93–102 8903208PMC2639826.

Lipsitch, M., Siller, S., Nowak, M.A., 1996. The evolution of virulence in pathogenswith vertical and horizontal transmission. Evolution, 1729–1741. ⟨http://www.jstor.org/stable/10.2307/2410731⟩.

Mandl, P., 1960. On the asymptotic behaviour of probabilities within groups ofstates of a homogeneous Markov process. Časopis pro pěstování matematiky 85(4), 448–456 (in Russian).

May, R.M., Nowak, M.A., 1994. Superinfection, metapopulation dynamics, and theevolution of diversity. J. Theor. Biol. 170 (1), 95–114 ⟨http://www.cscs.umich.edu/�rlr/CAFI/uploads/Main/mayNovakSuperinfection.pdf⟩.

May, R.M., Nowak, M.A., 1995. Coinfection and the evolution of parasite virulence.Proc. R. Soc. Lond. Ser. B: Biol. Sci. 261 (August (1361)), 209–215 7568274.⟨http://www.ncbi.nlm.nih.gov/pubmed/7568274⟩.

Mazancourt, C.d., Dieckmann, U., 2004. Trade-off geometries and frequency-dependent selection. Am. Nat. 164 (6), 765–778 ⟨http://www.jstor.org/stable/10.1086/424762⟩.

Messinger, S.M., Ostling, A., 2009. The consequences of spatial structure for theevolution of pathogen transmission rate and virulence. Am. Nat. 174 (October(4)), 441–454 ⟨http://www.jstor.org/stable/27735856⟩.

Metz, J.A., Mylius, S.D., Dieckmann, O., 1996. When Does Evolution Optimise?: Onthe Relation Between Types of Density Dependence and Evolutionarily StableLife History Parameters. IIASA Working Paper WP-96-004. International Insti-tute for Applied Systems Analysis, Laxenburg, Austria.

J. Humplik et al. / Journal of Theoretical Biology 360 (2014) 149–162 161

Page 14: Evolutionary dynamics of infectious diseases in …ped.fas.harvard.edu/.../1-s2.0-s0022519314003907-main.pdfEvolutionary dynamics of infectious diseases in finite populations Jan

Metz, J.A.J., Mylius, S.D., Diekmann, O., 2008. When does evolution optimize? Evol.Ecol. Res. 10 (5), 629–654 ⟨http://www.bio.vu.nl/thb/course/ecol/MetzMyli2008.pdf⟩.

Nasell, I., July 1995. The threshold concept in stochastic epidemic and endemicmodels. In: Mollison, D. (Ed.), Epidemic Models: Their Structure and Relation toData. Cambridge University Press, pp. 71–83.

Nowak, M.A., May, R.M., 1994. Superinfection and the evolution of parasitevirulence. Proc. R. Soc. Lond. Ser. B: Biol. Sci., 81–89.

Nowak, M.A., Sasaki, A., Taylor, C., Fudenberg, D., 2004. Emergence of cooperationand evolutionary stability in finite populations. Nature 428 (April (6983)),646–650 ⟨http://www.nature.com/nature/journal/v428/n6983/abs/nature02414.html⟩.

Nowak, M.A., Sigmund, K., 2004. Evolutionary dynamics of biological games.Science 303 (February (5659)), 793–799 ⟨http://www.sciencemag.org/cgi/content/abstract/sci;303/5659/793⟩.

Nåsell, I., 2001. Extinction and quasi-stationarity in the verhulst logistic model. J.Theor. Biol. 211 (1), 11–27 ⟨http://www.sciencedirect.com/science/article/pii/S0022519301923288⟩.

Ovaskainen, O., 2001. The quasistationary distribution of the stochastic logisticmodel. J. Appl. Probab. 38 (4), 898–907.

Proulx, S.R., Day, T., 2002. What can invasion analyses tell us about evolution understochasticity in finite populations? Selection 2 (04 (January)), 2–15, doi:http://dx.doi.org/10.1556/Select.2.2001.1-2.2.

Regoes, R.R., Nowak, M.A., Bonhoeffer, S., 2000. Evolution of virulence in aheterogeneous host population. Evolution 54 (February (1)), 64–71 ⟨http://www.jstor.org/stable/2640474⟩.

Ross, J.V., 2011. Invasion of infectious diseases in finite homogeneous populations.J. Theor. Biol. 289 (November), 83–89 ⟨http://www.sciencedirect.com/science/article/pii/S0022519311004383⟩.

Smith, J.M., 1982. Evolution and the Theory of Games. Cambridge University Press,Cambridge, UK, October.

Van Mieghem, P., Cator, E., 2012. Epidemics in networks with nodal self-infectionand the epidemic threshold. Phys. Rev. E 86 (July), 016116 ⟨http://link.aps.org/doi/10.1103/PhysRevE.86.016116⟩.

Webb, S.D., Keeling, M.J., Boots, M., 2013. A theoretical study of the role of spatialpopulation structure in the evolution of parasite virulence. Theor. Popul. Biol.84 (March), 36–45 23274478.

Weibull, J.W., 1997. Evolutionary Game Theory. MIT Press, Cambridge, MA.

J. Humplik et al. / Journal of Theoretical Biology 360 (2014) 149–162162