Nash epidemics

Simon K. Schnyder skschnyder@gmail.com Institute of Industrial Science, The University of Tokyo, 4-6-1 Komaba, Meguro-ku, Tokyo 153-8505, Japan    John J. Molina Department of Chemical Engineering, Kyoto University, Kyoto 615-8510, Japan    Ryoichi Yamamoto Department of Chemical Engineering, Kyoto University, Kyoto 615-8510, Japan    Matthew S. Turner m.s.turner@warwick.ac.uk Department of Physics, University of Warwick, Coventry CV4 7AL, UK Institute for Global Pandemic Planning, University of Warwick, Coventry CV4 7AL, UK
(July 5, 2024)
Abstract

Faced with a dangerous epidemic humans will spontaneously social distance to reduce their risk of infection at a socio-economic cost. Compartmentalised epidemic models have been extended to include this endogenous decision making: Individuals choose their behaviour to optimise a utility function, self-consistently giving rise to population behaviour. Here we study the properties of the resulting Nash equilibria, in which no member of the population can gain an advantage by unilaterally adopting different behaviour. We leverage a new analytic solution to obtain, (1) a simple relationship between rational social distancing behaviour and the current number of infections; (2) new scaling results for how the infection peak and number of total cases depend on the cost of contracting the disease; (3) characteristic infection costs that divide regimes of strong and weak behavioural response and depend only on the basic reproduction number of the disease; (4) a closed form expression for the value of the utility. We discuss how these analytic results provide a deep and intuitive understanding into the disease dynamics, useful for both individuals and policymakers. In particular the relationship between social distancing and infections represents a heuristic that could be communicated to the population to encourage, or “bootstrap”, rational behaviour.

Throughout history, epidemics caused by infectious diseases have caused considerable harm to humans. Individuals are typically assumed to be able to adjust their behaviour in reaction to the threat of an epidemic [1, 2, 3, 4, 5, 6, 7, 8, 9, 10]. In order to make choices about their behaviour individuals can weigh the costs and benefits of the outcomes of their behaviour. They may reduce their social activity when infections are high, in order to reduce the probability of becoming infected themselves, provided that the health costs outweigh the social and economic costs. A common assumption is that individual agents act rationally, i.e. to maximise an objective function or economic utility. This remains one of the fundamental assumptions of modern economic theory despite its limitations [11]. Rational individuals, who aim to maximise their individual objective function, end up targeting a Nash equilibrium [1, 9, 2, 12, 13] rather than the global utility maximum, which requires a coordinated effort to maximise a collective objective function  [14, 15, 7, 16]. Our work directly builds on this approach. It is possible to bring a Nash equilibrium into alignment with the global optimum [14, 17, 18, 19], e.g. via tax and subsidy incentives [20] which can be designed to bias rational individual behaviour appropriately.

While these studies tend to employ highly stylised mean-field compartmentalised models, they demonstrate the feasibility of such approaches. Such models can be extended to more accurately represent the complexity of epidemics and the systems in which they occur, such as additional compartment types with different risk and behaviour profiles [21, 2, 22, 23, 24, 25], seasonal effects [26], waning immunity [27, 25] e.g. due to new variants [28], as well as spatial, transmission or behavioural heterogeneity [29, 30, 22, 31, 32]. Other approaches feature spatial [33] and temporal networks [34, 35], and/or agent-based models [36, 37, 38, 39]. Others have worked to incorporate uncertainty and noise, by considering stochastic control [40, 41, 42, 43, 44], decision making under uncertainty [45, 46] and by understanding the robustness of control [47, 48, 49]. There have also been studies on inferring model structure and epidemiological properties from observed data [50, 51, 52, 53]. Finally, we also remark on the intriguing possibility of allowing individual opinions to directly influence policy makers [54].

Nash equilibria are widely believed to occur within such idealised models that incorporate endogenous behaviour during epidemics. However, until this work, solutions have only been accessible numerically. This is because the problem is intrinsically nonlinear, both at the level of the epidemiological dynamics and the objective function, leading to nonlinear control equations. Here, we provide for the first time, an analytic solution to the nonlinear time-dependent equilibrium control equations. This also demonstrates the existence of such a Nash equilibrium. In the limit of vanishing infection cost our results trivially recover the known analytic solutions for compartment models with constant basic reproduction number [55, 56, 57, 58, 59], i.e. without endogenous rational behaviour.

We focus on the case where the cost of infection is constant and where the government takes no role in directing the response to the epidemic. This situation has been already discussed, e.g. by [1] among many others, but only using numerical solutions. We do not investigate other possible policy interventions such as vaccination and treatment strategies, [60, 61, 62, 63, 9, 64, 3, 37, 4, 15, 39, 65, 66, 67, 25, 68], or isolation, testing, and active case-tracing strategies [69, 70]. We also ignore the situation where a vaccine becomes available during the epidemic. While the early arrival of a vaccine would have consequences for both equilibrium and globally optimal behaviour [1, 7, 13, 71], this lies outside of the scope of this work.

1 Epidemic dynamics

We use a standard SIR compartmentalised model [55] for the epidemic. The population is divided into susceptible, infected and recovered compartments, the latter implicitly including fatalities. The compartments evolve over time as

s˙˙𝑠\displaystyle\dot{s}over˙ start_ARG italic_s end_ARG =ksiabsent𝑘𝑠𝑖\displaystyle=-k\>s\>i= - italic_k italic_s italic_i (1)
i˙˙𝑖\displaystyle\dot{i}over˙ start_ARG italic_i end_ARG =ksiiabsent𝑘𝑠𝑖𝑖\displaystyle=k\>s\>i-i= italic_k italic_s italic_i - italic_i (2)
r˙˙𝑟\displaystyle\dot{r}over˙ start_ARG italic_r end_ARG =iabsent𝑖\displaystyle=i= italic_i

Here a dot denotes a time derivative and the time dependence of s(t)𝑠𝑡s(t)italic_s ( italic_t ), i(t)𝑖𝑡i(t)italic_i ( italic_t ), r(t)𝑟𝑡r(t)italic_r ( italic_t ) and k(t)𝑘𝑡k(t)italic_k ( italic_t ) is omitted for brevity. We normalise the compartments, 1=s+i+r1𝑠𝑖𝑟1=s+i+r1 = italic_s + italic_i + italic_r. We use one timescale for both recovery and the duration of infectiousness, for simplicity, and have rescaled the equations so that time t𝑡titalic_t is measured in units of this single timescale. The initial conditions are set as s(0)=s0𝑠0subscript𝑠0s(0)=s_{0}italic_s ( 0 ) = italic_s start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, i(0)=i0𝑖0subscript𝑖0i(0)=i_{0}italic_i ( 0 ) = italic_i start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, r(0)=r0𝑟0subscript𝑟0r(0)=r_{0}italic_r ( 0 ) = italic_r start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, with s0,i0,r00subscript𝑠0subscript𝑖0subscript𝑟00s_{0},i_{0},r_{0}\geq 0italic_s start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT , italic_i start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT , italic_r start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ≥ 0 and s0+i0+r0=1subscript𝑠0subscript𝑖0subscript𝑟01s_{0}+i_{0}+r_{0}=1italic_s start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT + italic_i start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT + italic_r start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 1. In all figures we arbitrarily select a time origin t=0𝑡0t=0italic_t = 0 where the epidemic is in its very early stages according to r(0)=i(0)/(R01)=106𝑟0𝑖0subscript𝑅01superscript106r(0)=i(0)/(R_{0}-1)=10^{-6}italic_r ( 0 ) = italic_i ( 0 ) / ( italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT - 1 ) = 10 start_POSTSUPERSCRIPT - 6 end_POSTSUPERSCRIPT with s(0)=1i(0)r(0)𝑠01𝑖0𝑟0s(0)=1-i(0)-r(0)italic_s ( 0 ) = 1 - italic_i ( 0 ) - italic_r ( 0 ).

The population’s average social activity behaviour is encoded in the current infection rate, assumed to satisfy k(t)0𝑘𝑡0k(t)\geq 0italic_k ( italic_t ) ≥ 0 although our analytic results later suggest a stronger bound k(t)1𝑘𝑡1k(t)\geq 1italic_k ( italic_t ) ≥ 1. We assume that the disease exhibits a natural level of activity in the absence of any behavioural modification that is a constant known as the basic reproduction number R0subscript𝑅0R_{0}italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT. Below we use the case k(t)=R0𝑘𝑡subscript𝑅0k(t)=R_{0}italic_k ( italic_t ) = italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT to establish a non-behavioural baseline dynamics for comparison.

2 Nash equilibrium behaviour

In order to study self-organised behaviour, we imagine an average individual making decisions about their own behaviour. This represents a mean-field game [72, 73], for the Nash equilibrium of which a set of ordinary differential equations can be straightforwardly derived [74, 9].

The individual’s effect on the epidemic is negligible but they can influence their own fate by selecting a strategy κ(t)0𝜅𝑡0\kappa(t)\geq 0italic_κ ( italic_t ) ≥ 0 which it is initially assumed can differ from the population-averaged strategy k(t)𝑘𝑡k(t)italic_k ( italic_t ). The probabilities that an individual is in each of the compartments evolves over time according to

ψ˙ssubscript˙𝜓𝑠\displaystyle\dot{\psi}_{s}over˙ start_ARG italic_ψ end_ARG start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT =κψsiabsent𝜅subscript𝜓𝑠𝑖\displaystyle=-\kappa\psi_{s}i= - italic_κ italic_ψ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT italic_i (3)
ψ˙isubscript˙𝜓𝑖\displaystyle\dot{\psi}_{i}over˙ start_ARG italic_ψ end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT =κψsiψiabsent𝜅subscript𝜓𝑠𝑖subscript𝜓𝑖\displaystyle=\kappa\psi_{s}i-\psi_{i}= italic_κ italic_ψ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT italic_i - italic_ψ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT (4)

Lowering κ(t)𝜅𝑡\kappa(t)italic_κ ( italic_t ) directly increases the probability of the individual remaining susceptible and reduces their probability of becoming infectious. While these equations are similar to eqs. 2, they couple to the infectious compartment of the population i𝑖iitalic_i as the only donor of infection.

We assume that an individual has rational interests that can be captured by an objective function or utility. In general this will depend on both their own and the population behaviours, U(κ(t),k(t))𝑈𝜅𝑡𝑘𝑡U(\kappa(t),k(t))italic_U ( italic_κ ( italic_t ) , italic_k ( italic_t ) ). The individual seek to maximise this objective function. Assuming that the population consist of identical individuals, a Nash equilibrium exists if there is a strategy k=κ(t)𝑘𝜅𝑡k=\kappa(t)italic_k = italic_κ ( italic_t ), adopted by the population, and the individual cannot improve their outcome by unilaterally deviating from the behaviour κ𝜅\kappaitalic_κ,

U(κ~(t),κ(t))U(κ(t),κ(t)) for any κ~(t).𝑈~𝜅𝑡𝜅𝑡𝑈𝜅𝑡𝜅𝑡 for any ~𝜅𝑡\displaystyle U(\tilde{\kappa}(t),\kappa(t))\leq U(\kappa(t),\kappa(t))\quad% \text{ for any }\tilde{\kappa}(t).italic_U ( over~ start_ARG italic_κ end_ARG ( italic_t ) , italic_κ ( italic_t ) ) ≤ italic_U ( italic_κ ( italic_t ) , italic_κ ( italic_t ) ) for any over~ start_ARG italic_κ end_ARG ( italic_t ) . (5)

In order to find this Nash strategy one first maximises U(κ,k)𝑈𝜅𝑘U(\kappa,k)italic_U ( italic_κ , italic_k ) over κ𝜅\kappaitalic_κ for an arbitrary, exogenous k𝑘kitalic_k [9]. This constitutes a standard constrained optimisation problem. To make the strategy self-consistent, one then assumes that all individuals in the population would optimise their behaviour in the same way, and therefore k=κ𝑘𝜅k=\kappaitalic_k = italic_κ. This then automatically results in ψs=ssubscript𝜓𝑠𝑠\psi_{s}=sitalic_ψ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT = italic_s, ψi=isubscript𝜓𝑖𝑖\psi_{i}=iitalic_ψ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = italic_i with dynamics that corresponds to the Nash equilibrium.

In this work, we focus on an idealised individual objective function or utility U𝑈Uitalic_U with, see ref. 19,

U=𝑈absent\displaystyle U=italic_U = 0tfu(t)𝑑t+Ufsuperscriptsubscript0subscript𝑡𝑓𝑢𝑡differential-d𝑡subscript𝑈𝑓\displaystyle\ \int_{0}^{t_{f}}u(t)dt+U_{f}∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_t start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT end_POSTSUPERSCRIPT italic_u ( italic_t ) italic_d italic_t + italic_U start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT (6)
u=𝑢absent\displaystyle u=italic_u = ft[αψiβ(κR0)2]superscript𝑓𝑡delimited-[]𝛼subscript𝜓𝑖𝛽superscript𝜅subscript𝑅02\displaystyle\ f^{-t}\left[-\alpha\>\psi_{i}-\beta\>(\kappa-R_{0})^{2}\right]italic_f start_POSTSUPERSCRIPT - italic_t end_POSTSUPERSCRIPT [ - italic_α italic_ψ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT - italic_β ( italic_κ - italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ] (7)

The average infection cost is given by α𝛼\alphaitalic_α (this also includes the cost of death) with α0𝛼0\alpha\geq 0italic_α ≥ 0.

The social and financial costs of social distancing are parametrised by a constant β>0𝛽0\beta>0italic_β > 0. In what follows we choose to work in units of utility in which β=1𝛽1\beta=1italic_β = 1, without loss of generality. The quadratic form of this social distancing term encodes that it is costly to deviate from one’s default behaviour and ensures that an individual would naturally select behaviour corresponding to κ=R0𝜅subscript𝑅0\kappa=R_{0}italic_κ = italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT if there were no epidemic (or it bore no cost). Motivated by the observation that no individual, regardless of compartment, can socially distance without incurring a cost, we assume that the cost of social distancing always applies, in contrast to other work, e.g. [1, 15] where the cost of social distancing is paid mostly by the s𝑠sitalic_s-compartment. Our choice corresponds to an approximation in which individuals are uncertain about which compartment they find themselves in, e.g. when infections can be asymptomatic.

In general one might truncate the utility integral at a final time tfsubscript𝑡𝑓t_{f}italic_t start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT. One approach here is to assuming that at that time the susceptible compartment becomes completely and perfectly vaccinated. The cost of being infected when the vaccine becomes available is expressed as a so-called salvage term Uf=ftfαψi(tf)/(1+lnf)subscript𝑈𝑓superscript𝑓subscript𝑡𝑓𝛼subscript𝜓𝑖subscript𝑡𝑓1𝑓U_{f}=\ -f^{-t_{f}}\alpha\ \psi_{i}(t_{f})/(1+\ln f)italic_U start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT = - italic_f start_POSTSUPERSCRIPT - italic_t start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT end_POSTSUPERSCRIPT italic_α italic_ψ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_t start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT ) / ( 1 + roman_ln italic_f ), see section 6.1 for a derivation. If tfsubscript𝑡𝑓t_{f}italic_t start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT is comparable to or shorter than the duration of the epidemic then it can have a qualitative effect on rational decision-making and thus the course of the epidemic [1, 7, 71]. However, this lies outside of the scope of this work, and so we choose tfsubscript𝑡𝑓t_{f}\to\inftyitalic_t start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT → ∞. In this case, vaccination plays no role in decision making and i(tf)0𝑖subscript𝑡𝑓0i(t_{f})\to 0italic_i ( italic_t start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT ) → 0.

Since the utility function is convex, we expect that the optimisation problem has a (unique) solution. We directly demonstrate uniqueness and existence by calculating the analytic solution to this problem below.

We use a standard Hamiltonian/Lagrangian approach [74, 9], which in optimal control theory is referred to as Pontryagin’s maximum principle [75], to calculate the optimal behaviour of an individual κ𝜅\kappaitalic_κ in response to an exogenous behaviour k𝑘kitalic_k and the corresponding course of the epidemic. This approach allows for reformulating the optimisation problem as a boundary value problem which is generated from an auxiliary function, the Hamiltonian. The Hamiltonian for the individual can be expressed by

H=𝐻absent\displaystyle H=italic_H = u+vs(κψsi)+vi(κψsiψi)𝑢subscript𝑣𝑠𝜅subscript𝜓𝑠𝑖subscript𝑣𝑖𝜅subscript𝜓𝑠𝑖subscript𝜓𝑖\displaystyle\ u+v_{s}(-\kappa\psi_{s}i)+v_{i}(\kappa\psi_{s}i-\psi_{i})italic_u + italic_v start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( - italic_κ italic_ψ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT italic_i ) + italic_v start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_κ italic_ψ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT italic_i - italic_ψ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) (8)
=\displaystyle== ft[αψi+(κR0)2]superscript𝑓𝑡delimited-[]𝛼subscript𝜓𝑖superscript𝜅subscript𝑅02\displaystyle-f^{-t}\left[\alpha\psi_{i}+\,(\kappa-R_{0})^{2}\right]- italic_f start_POSTSUPERSCRIPT - italic_t end_POSTSUPERSCRIPT [ italic_α italic_ψ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT + ( italic_κ - italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ] (9)
(vsvi)κψsiviψisubscript𝑣𝑠subscript𝑣𝑖𝜅subscript𝜓𝑠𝑖subscript𝑣𝑖subscript𝜓𝑖\displaystyle-(v_{s}-v_{i})\kappa\psi_{s}i-v_{i}\psi_{i}- ( italic_v start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT - italic_v start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) italic_κ italic_ψ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT italic_i - italic_v start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_ψ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT (10)

The Lagrange fields vs(t)subscript𝑣𝑠𝑡v_{s}(t)italic_v start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( italic_t ) and vi(t)subscript𝑣𝑖𝑡v_{i}(t)italic_v start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_t ), expressing the expected present value of the utility of being in each compartment at each point in time [9] enforce the constraint of the dynamics to eqs. (4). Their equations of motion are

v˙ssubscript˙𝑣𝑠\displaystyle\dot{v}_{s}over˙ start_ARG italic_v end_ARG start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT =Hψs=(vsvi)κiabsent𝐻subscript𝜓𝑠subscript𝑣𝑠subscript𝑣𝑖𝜅𝑖\displaystyle=-\frac{\partial H}{\partial\psi_{s}}=(v_{s}-v_{i})\kappa i= - divide start_ARG ∂ italic_H end_ARG start_ARG ∂ italic_ψ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT end_ARG = ( italic_v start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT - italic_v start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) italic_κ italic_i (11)
v˙isubscript˙𝑣𝑖\displaystyle\dot{v}_{i}over˙ start_ARG italic_v end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT =Hψi=ftα+viabsent𝐻subscript𝜓𝑖superscript𝑓𝑡𝛼subscript𝑣𝑖\displaystyle=-\frac{\partial H}{\partial\psi_{i}}=f^{-t}\alpha+v_{i}= - divide start_ARG ∂ italic_H end_ARG start_ARG ∂ italic_ψ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG = italic_f start_POSTSUPERSCRIPT - italic_t end_POSTSUPERSCRIPT italic_α + italic_v start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT (12)

with boundary conditions

vs(tf)=Ufψs,f=0,vi(tf)=Ufψi,f=ftfα1+lnf.formulae-sequencesubscript𝑣𝑠subscript𝑡𝑓subscript𝑈𝑓subscript𝜓𝑠𝑓0subscript𝑣𝑖subscript𝑡𝑓subscript𝑈𝑓subscript𝜓𝑖𝑓superscript𝑓subscript𝑡𝑓𝛼1𝑓\displaystyle v_{s}(t_{f})=\frac{\partial U_{f}}{\partial\psi_{s,f}}=0,\ v_{i}% (t_{f})=\frac{\partial U_{f}}{\partial\psi_{i,f}}=\frac{-f^{-t_{f}}\alpha}{1+% \ln f}.italic_v start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( italic_t start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT ) = divide start_ARG ∂ italic_U start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT end_ARG start_ARG ∂ italic_ψ start_POSTSUBSCRIPT italic_s , italic_f end_POSTSUBSCRIPT end_ARG = 0 , italic_v start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_t start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT ) = divide start_ARG ∂ italic_U start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT end_ARG start_ARG ∂ italic_ψ start_POSTSUBSCRIPT italic_i , italic_f end_POSTSUBSCRIPT end_ARG = divide start_ARG - italic_f start_POSTSUPERSCRIPT - italic_t start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT end_POSTSUPERSCRIPT italic_α end_ARG start_ARG 1 + roman_ln italic_f end_ARG . (13)

Given the exogenous course of the epidemic in the population, the individual can optimise their own utility be choosing the strategy that satisfies 0=H/κ0𝐻𝜅0=\partial H/\partial\kappa0 = ∂ italic_H / ∂ italic_κ. From this we obtain κ=R0ft2(vsvi)ψsi𝜅subscript𝑅0superscript𝑓𝑡2subscript𝑣𝑠subscript𝑣𝑖subscript𝜓𝑠𝑖\kappa=R_{0}-\frac{f^{t}}{2}(v_{s}-v_{i})\psi_{s}iitalic_κ = italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT - divide start_ARG italic_f start_POSTSUPERSCRIPT italic_t end_POSTSUPERSCRIPT end_ARG start_ARG 2 end_ARG ( italic_v start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT - italic_v start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) italic_ψ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT italic_i. Assuming that the population consist entirely of identical individuals who would all independently from each other choose the same strategy, we can conclude that the average population behaviour must be self-consistently given by k(t)=κ(t)𝑘𝑡𝜅𝑡k(t)=\kappa(t)italic_k ( italic_t ) = italic_κ ( italic_t ). Hence, this gives rise to a Nash equilibrium. Then, naturally also s=ψs𝑠subscript𝜓𝑠s=\psi_{s}italic_s = italic_ψ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT and i=ψi𝑖subscript𝜓𝑖i=\psi_{i}italic_i = italic_ψ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT, and

k=κ=R0ft2(vsvi)si.𝑘𝜅subscript𝑅0superscript𝑓𝑡2subscript𝑣𝑠subscript𝑣𝑖𝑠𝑖\displaystyle k=\kappa=R_{0}-\frac{f^{t}}{2}(v_{s}-v_{i})si.italic_k = italic_κ = italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT - divide start_ARG italic_f start_POSTSUPERSCRIPT italic_t end_POSTSUPERSCRIPT end_ARG start_ARG 2 end_ARG ( italic_v start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT - italic_v start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) italic_s italic_i . (14)

The variational approach as stated only yields conditions sufficient to identify extrema. To confirm that a solution is a maximum, 2H/κ2<0superscript2𝐻superscript𝜅20\partial^{2}H/\partial\kappa^{2}<0∂ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_H / ∂ italic_κ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT < 0 is required. Here we find 2H/κ2=2<0superscript2𝐻superscript𝜅220\partial^{2}H/\partial\kappa^{2}=-2<0∂ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_H / ∂ italic_κ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT = - 2 < 0.

3 Analytic solution

The Nash equilibrium k(t)𝑘𝑡k(t)italic_k ( italic_t ) optimising the utility eq. 6 is given by the solution of eq. 2, eqs. 11, 12 and 13, in conjunction with the optimality condition eq. 14. From here, we calculate the analytic solution for this set of equations. We assume the case without economic discounting, f=1𝑓1f=1italic_f = 1.

Firstly, we work with the integrated fraction of infected cases up to time t𝑡titalic_t, i.e. the fraction of recovered cases r𝑟ritalic_r, defined as

r=0ti(t)𝑑t+r0𝑟superscriptsubscript0𝑡𝑖superscript𝑡differential-dsuperscript𝑡subscript𝑟0r=\int_{0}^{t}i(t^{\prime})dt^{\prime}+r_{0}italic_r = ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_t end_POSTSUPERSCRIPT italic_i ( italic_t start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) italic_d italic_t start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT + italic_r start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT (15)

noting that i=r˙𝑖˙𝑟i=\dot{r}italic_i = over˙ start_ARG italic_r end_ARG. In what follows it is convenient to consider an implicit form for the behaviour k(r)𝑘𝑟k(r)italic_k ( italic_r ). Because r(t)𝑟𝑡r(t)italic_r ( italic_t ) is monotonic we can rely on a one-to-one mapping between k(t)𝑘𝑡k(t)italic_k ( italic_t ) and k(r)𝑘𝑟k(r)italic_k ( italic_r ). The second transformation involves defining K˙=ki˙𝐾𝑘𝑖\dot{K}=k\>iover˙ start_ARG italic_K end_ARG = italic_k italic_i, hence K𝐾Kitalic_K obeys

K(r)=r0rk(r)𝑑r𝐾𝑟superscriptsubscriptsubscript𝑟0𝑟𝑘superscript𝑟differential-dsuperscript𝑟K(r)=\int_{r_{0}}^{r}k(r^{\prime})dr^{\prime}italic_K ( italic_r ) = ∫ start_POSTSUBSCRIPT italic_r start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_r end_POSTSUPERSCRIPT italic_k ( italic_r start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) italic_d italic_r start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT (16)

Equations (2) then lead to s˙=sK˙˙𝑠𝑠˙𝐾\dot{s}=-s\dot{K}over˙ start_ARG italic_s end_ARG = - italic_s over˙ start_ARG italic_K end_ARG which integrates to

s=s0eK(r)𝑠subscript𝑠0superscript𝑒𝐾𝑟s=s_{0}e^{-K(r)}italic_s = italic_s start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT - italic_K ( italic_r ) end_POSTSUPERSCRIPT (17)

Using 1=s+i+r1𝑠𝑖𝑟1=s+i+r1 = italic_s + italic_i + italic_r, we obtain directly

i=1rs0eK(r)𝑖1𝑟subscript𝑠0superscript𝑒𝐾𝑟i=1-r-s_{0}e^{-K(r)}italic_i = 1 - italic_r - italic_s start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT - italic_K ( italic_r ) end_POSTSUPERSCRIPT (18)

Since i=r˙𝑖˙𝑟i=\dot{r}italic_i = over˙ start_ARG italic_r end_ARG, we can integrate this equation to obtain

F(r)r0rdr1rs0eK(r)=thencer=F1(t)formulae-sequence𝐹𝑟superscriptsubscriptsubscript𝑟0𝑟𝑑superscript𝑟1superscript𝑟subscript𝑠0superscript𝑒𝐾superscript𝑟𝑡hence𝑟superscript𝐹1𝑡F(r)\equiv\int_{r_{0}}^{r}\frac{dr^{\prime}}{1-r^{\prime}-s_{0}e^{-K(r^{\prime% })}}=t\ \ \ {\rm hence}\ \ r=F^{-1}(t)italic_F ( italic_r ) ≡ ∫ start_POSTSUBSCRIPT italic_r start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_r end_POSTSUPERSCRIPT divide start_ARG italic_d italic_r start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_ARG start_ARG 1 - italic_r start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT - italic_s start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT - italic_K ( italic_r start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) end_POSTSUPERSCRIPT end_ARG = italic_t roman_hence italic_r = italic_F start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT ( italic_t ) (19)

Recalling our assumption that tfsubscript𝑡𝑓t_{f}\to\inftyitalic_t start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT → ∞, we can conclude that i(tf)0𝑖subscript𝑡𝑓0i(t_{f})\to 0italic_i ( italic_t start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT ) → 0. In this limit the cumulative total of infections reaches its final value given by the non-zero root of

rf+s0eK(rf)=1subscript𝑟𝑓subscript𝑠0superscript𝑒𝐾subscript𝑟𝑓1r_{f}+s_{0}e^{-K(r_{f})}=1italic_r start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT + italic_s start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT - italic_K ( italic_r start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT ) end_POSTSUPERSCRIPT = 1 (20)

Equations 15, 16, 17, 18, 19 and 20 above hold irrespective of the form of the objective function.

Refer to caption
Figure 1: Direct plots of the analytic solution. (a) The analytic solution of the Nash equilibrium social distancing problem as obtained in eq. 28 as a function of the recovered r𝑟ritalic_r for an exemplary range of infection costs α𝛼\alphaitalic_α and R0=4subscript𝑅04R_{0}=4italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 4. Initial conditions here and in all following figures are set to r0=106subscript𝑟0superscript106r_{0}=10^{-6}italic_r start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 10 start_POSTSUPERSCRIPT - 6 end_POSTSUPERSCRIPT and i0=3106subscript𝑖03superscript106i_{0}=3\cdot 10^{-6}italic_i start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 3 ⋅ 10 start_POSTSUPERSCRIPT - 6 end_POSTSUPERSCRIPT. (b) The fraction of infectious i𝑖iitalic_i as a function of the susceptible s𝑠sitalic_s for the same range of α𝛼\alphaitalic_α. (c) Deviation of the social distancing behaviour k𝑘kitalic_k from the pre-epidemic default R0subscript𝑅0R_{0}italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT as a function of i𝑖iitalic_i, emphasising their linear relationship as established in eq. 24.

Concerning the Lagrange fields, we can see directly from eq. 12 that vi(t)=αsubscript𝑣𝑖𝑡𝛼v_{i}(t)=-\alphaitalic_v start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_t ) = - italic_α, whereas vssubscript𝑣𝑠v_{s}italic_v start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT follows from eq. 11

v˙s=(vs+α)K˙subscript˙𝑣𝑠subscript𝑣𝑠𝛼˙𝐾\displaystyle\dot{v}_{s}=(v_{s}+\alpha)\dot{K}over˙ start_ARG italic_v end_ARG start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT = ( italic_v start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT + italic_α ) over˙ start_ARG italic_K end_ARG (21)

Integrating we obtain

vs+α=μeK=μs0ssubscript𝑣𝑠𝛼𝜇superscript𝑒𝐾𝜇subscript𝑠0𝑠\displaystyle v_{s}+\alpha=\mu e^{K}=\mu\frac{s_{0}}{s}italic_v start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT + italic_α = italic_μ italic_e start_POSTSUPERSCRIPT italic_K end_POSTSUPERSCRIPT = italic_μ divide start_ARG italic_s start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG start_ARG italic_s end_ARG (22)

with a constant μ𝜇\muitalic_μ. From the boundary condition vs(tf)=0subscript𝑣𝑠subscript𝑡𝑓0v_{s}(t_{f})=0italic_v start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( italic_t start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT ) = 0 we can conclude

vs=αsfsαsubscript𝑣𝑠𝛼subscript𝑠𝑓𝑠𝛼\displaystyle v_{s}=\alpha\frac{s_{f}}{s}-\alphaitalic_v start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT = italic_α divide start_ARG italic_s start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT end_ARG start_ARG italic_s end_ARG - italic_α (23)

with sf=s(tf)subscript𝑠𝑓𝑠subscript𝑡𝑓s_{f}=s(t_{f})italic_s start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT = italic_s ( italic_t start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT ). The optimal behaviour is then given by eq. 14

k𝑘\displaystyle kitalic_k =R0αsf2iabsentsubscript𝑅0𝛼subscript𝑠𝑓2𝑖\displaystyle=R_{0}-\frac{\alpha s_{f}}{2}i= italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT - divide start_ARG italic_α italic_s start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT end_ARG start_ARG 2 end_ARG italic_i (24)

This is tremendously simple: the equilibrium strength of social distancing kR0𝑘subscript𝑅0k-R_{0}italic_k - italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT is proportional to both the number of infectious cases and the cost of infection at any given time, see fig. 1c. With s=s0eK(r)𝑠subscript𝑠0superscript𝑒𝐾𝑟s=s_{0}e^{-K(r)}italic_s = italic_s start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT - italic_K ( italic_r ) end_POSTSUPERSCRIPT we have

rs=srK=sksubscript𝑟𝑠𝑠subscript𝑟𝐾𝑠𝑘\displaystyle\partial_{r}s=-s\partial_{r}K=-sk∂ start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT italic_s = - italic_s ∂ start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT italic_K = - italic_s italic_k (25)

and therefore, inserting eq. 24 and i=1rs𝑖1𝑟𝑠i=1-r-sitalic_i = 1 - italic_r - italic_s,

rssubscript𝑟𝑠\displaystyle\partial_{r}s∂ start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT italic_s =s[R0αsf2(1rs)]absent𝑠delimited-[]subscript𝑅0𝛼subscript𝑠𝑓21𝑟𝑠\displaystyle=-s[R_{0}-\frac{\alpha s_{f}}{2}(1-r-s)]= - italic_s [ italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT - divide start_ARG italic_α italic_s start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT end_ARG start_ARG 2 end_ARG ( 1 - italic_r - italic_s ) ] (26)
=s[a+br+bs]absent𝑠delimited-[]𝑎𝑏𝑟𝑏𝑠\displaystyle=-s[a+br+bs]= - italic_s [ italic_a + italic_b italic_r + italic_b italic_s ] (27)

with a=R0b𝑎subscript𝑅0𝑏a=R_{0}-bitalic_a = italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT - italic_b and b=αsf/2𝑏𝛼subscript𝑠𝑓2b=\alpha s_{f}/2italic_b = italic_α italic_s start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT / 2. This has an analytic solution that satisfies s(rr0)s0𝑠𝑟subscript𝑟0subscript𝑠0s(r\to r_{0})\to s_{0}italic_s ( italic_r → italic_r start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) → italic_s start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT

s(r)=exp[12(rr0)(2a+b(r+r0))]1s0πb2exp[(a+br0)22b](Erf[a+br02b]Erf[a+br2b])𝑠𝑟12𝑟subscript𝑟02𝑎𝑏𝑟subscript𝑟01subscript𝑠0𝜋𝑏2superscript𝑎𝑏subscript𝑟022𝑏Erfdelimited-[]𝑎𝑏subscript𝑟02𝑏Erfdelimited-[]𝑎𝑏𝑟2𝑏\displaystyle s(r)=\frac{\exp\left[-\frac{1}{2}(r-r_{0})(2a+b(r+r_{0}))\right]% }{\frac{1}{s_{0}}-\sqrt{\frac{\pi b}{2}}\exp\left[\frac{(a+br_{0})^{2}}{2b}% \right]\left(\text{Erf}\left[\frac{a+br_{0}}{\sqrt{2b}}\right]-\text{Erf}\left% [\frac{a+br}{\sqrt{2b}}\right]\right)}italic_s ( italic_r ) = divide start_ARG roman_exp [ - divide start_ARG 1 end_ARG start_ARG 2 end_ARG ( italic_r - italic_r start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) ( 2 italic_a + italic_b ( italic_r + italic_r start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) ) ] end_ARG start_ARG divide start_ARG 1 end_ARG start_ARG italic_s start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG - square-root start_ARG divide start_ARG italic_π italic_b end_ARG start_ARG 2 end_ARG end_ARG roman_exp [ divide start_ARG ( italic_a + italic_b italic_r start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 italic_b end_ARG ] ( Erf [ divide start_ARG italic_a + italic_b italic_r start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG start_ARG square-root start_ARG 2 italic_b end_ARG end_ARG ] - Erf [ divide start_ARG italic_a + italic_b italic_r end_ARG start_ARG square-root start_ARG 2 italic_b end_ARG end_ARG ] ) end_ARG (28)

Using rf=1s(rf)subscript𝑟𝑓1𝑠subscript𝑟𝑓r_{f}=1-s(r_{f})italic_r start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT = 1 - italic_s ( italic_r start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT ), we can self-consistently determine rfsubscript𝑟𝑓r_{f}italic_r start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT and thus obtain the solution. We show the result of eq. 28 for a range of infection costs α𝛼\alphaitalic_α in fig. 1a. The analytic solution for the infectious compartment i(r)=1rs(r)𝑖𝑟1𝑟𝑠𝑟i(r)=1-r-s(r)italic_i ( italic_r ) = 1 - italic_r - italic_s ( italic_r ) can be plotted in a natural way on the s-i plane, see fig. 1b.

In our approach, time is parametrised as eq. 19, which can easily be evaluated numerically. The analytic solution can then be plotted in the typical way, fig. 2.

Refer to caption
Figure 2: Analytic solution as a function of time. (a) Equilibrium social activity behaviour of the population k(t)𝑘𝑡k(t)italic_k ( italic_t ) and corresponding dynamics of the disease (b) s𝑠sitalic_s and (c) i𝑖iitalic_i for an exemplary range of infection costs α𝛼\alphaitalic_α and R0=4subscript𝑅04R_{0}=4italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 4. Since infections incur a cost, the equilibrium behaviour seeks to avoid excessive infections by self-organised social distancing. The higher the cost, the more reduced social activity k𝑘kitalic_k becomes.
Refer to caption
Figure 3: Scaling. (a) Excess cases ε(α,R0)𝜀𝛼subscript𝑅0\varepsilon(\alpha,R_{0})italic_ε ( italic_α , italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) vs. infection cost α𝛼\alphaitalic_α for a range of basic reproduction numbers R0subscript𝑅0R_{0}italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT. The high infection cost asymptotes, see eq. 34, are shown as dashed lines and the crossover costs αssubscriptsuperscript𝛼𝑠\alpha^{\star}_{s}italic_α start_POSTSUPERSCRIPT ⋆ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT, see eq. 36, as black stars. Inset: The data collapses onto the low-α𝛼\alphaitalic_α and high infection cost asymptotes by rescaling the cost α𝛼\alphaitalic_α with the crossover cost αssuperscriptsubscript𝛼𝑠\alpha_{s}^{\star}italic_α start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ⋆ end_POSTSUPERSCRIPT, see eq. 36, while rescaling ε(α,R0)𝜀𝛼subscript𝑅0\varepsilon(\alpha,R_{0})italic_ε ( italic_α , italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) with its non-behavioural limit, see eq. 32. (b) The infection peak i^^𝑖\hat{i}over^ start_ARG italic_i end_ARG vs. α𝛼\alphaitalic_α for a range of R0subscript𝑅0R_{0}italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT. The high infection cost asymptotes, see eq. 35, are shown as dashed lines and the crossover costs αisubscriptsuperscript𝛼𝑖\alpha^{\star}_{i}italic_α start_POSTSUPERSCRIPT ⋆ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT, see eq. 37, as grey stars. Inset: The data collapses onto the low-α𝛼\alphaitalic_α and high infection cost asymptotes by rescaling the cost α𝛼\alphaitalic_α with the crossover cost αisubscriptsuperscript𝛼𝑖\alpha^{\star}_{i}italic_α start_POSTSUPERSCRIPT ⋆ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT, eq. 37, while rescaling the peak height with its non-behavioural limit, see eq. 33.
Refer to caption
Figure 4: Behavioural response. Characterisation of the Nash equilibrium response in the R0subscript𝑅0R_{0}italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPTα𝛼\alphaitalic_α parameter space. On the high R0subscript𝑅0R_{0}italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT – low-α𝛼\alphaitalic_α side of the line, the behaviour is well represented by the non-behavioural limit, in which it is not rational to significantly modify one’s behaviour. On the low R0subscript𝑅0R_{0}italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT – high infection cost side, it is rational to strongly modify one’s behaviour. The lines describing the crossover are given by the critical costs αssuperscriptsubscript𝛼𝑠\alpha_{s}^{\star}italic_α start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ⋆ end_POSTSUPERSCRIPT for the transition in the excess cases, see eq. 36, and/or αisuperscriptsubscript𝛼𝑖\alpha_{i}^{\star}italic_α start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ⋆ end_POSTSUPERSCRIPT for the transition in the infection peak, see eq. 37. The parameter values used for some of the curves in figs. 2 and 1 are marked by analogously coloured dots.

4 Results

The higher the infection cost α𝛼\alphaitalic_α, the stronger is the incentive to reduce social activity and hence k𝑘kitalic_k, see fig. 2. The stronger the reduction in k𝑘kitalic_k, the more slowly the epidemic progresses, the lower the peak infection levels are, and the higher the total number of cases 1sf1subscript𝑠𝑓1-s_{f}1 - italic_s start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT becomes.

In what follows, we analyse the epidemic using two key quantities, the excess cases ε𝜀\varepsilonitalic_ε and the peak of the epidemic max(i)𝑖\max(i)roman_max ( italic_i ). For tfsubscript𝑡𝑓t_{f}\to\inftyitalic_t start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT → ∞, herd immunity is always reached. The final number of susceptibles then always satisfies sf1/R0subscript𝑠𝑓1subscript𝑅0s_{f}\leq 1/R_{0}italic_s start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT ≤ 1 / italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, with 1/R01subscript𝑅01/R_{0}1 / italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT the minimum number of cases for which herd immunity is guaranteed. The cases in excess of this threshold are defined as

ε=1/R0sf𝜀1subscript𝑅0subscript𝑠𝑓\displaystyle\varepsilon=1/R_{0}-s_{f}italic_ε = 1 / italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT - italic_s start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT (29)

We will calculate ε𝜀\varepsilonitalic_ε and max(i)𝑖\max(i)roman_max ( italic_i ) in two limiting cases: (1) The Non-behavioural limit in which there is no perceived infection cost α=0𝛼0\alpha=0italic_α = 0. In this case there is no reason to modify one’s behaviour, k=R0𝑘subscript𝑅0k=R_{0}italic_k = italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, see purple lines in figs. 1 and 2. (2) The high-infection-cost asymptote in which infection costs are very high, α/R021much-greater-than𝛼superscriptsubscript𝑅021\alpha/R_{0}^{2}\gg 1italic_α / italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ≫ 1. By matching these solutions we will obtain crossover costs between these scaling results.

Non-behavioural limit

For this edge case only, the analytic solution was known previously [55, 56, 57, 58, 59]. We recover it in our notation as follows. Since α=0𝛼0\alpha=0italic_α = 0, eq. 27 is solved by

s(r)=s0eR0(rr0)𝑠𝑟subscript𝑠0superscript𝑒subscript𝑅0𝑟subscript𝑟0\displaystyle s(r)=s_{0}e^{-R_{0}(r-r_{0})}italic_s ( italic_r ) = italic_s start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT - italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( italic_r - italic_r start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) end_POSTSUPERSCRIPT (30)

Its limit sf=eR0(1sfr0)subscript𝑠𝑓superscript𝑒subscript𝑅01subscript𝑠𝑓subscript𝑟0s_{f}=e^{-R_{0}(1-s_{f}-r_{0})}italic_s start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT = italic_e start_POSTSUPERSCRIPT - italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( 1 - italic_s start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT - italic_r start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) end_POSTSUPERSCRIPT yields

sf=W(s0R0eR0(r01))/R0subscript𝑠𝑓𝑊subscript𝑠0subscript𝑅0superscript𝑒subscript𝑅0subscript𝑟01subscript𝑅0\displaystyle s_{f}=-W(-s_{0}R_{0}e^{R_{0}(r_{0}-1)})/R_{0}italic_s start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT = - italic_W ( - italic_s start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( italic_r start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT - 1 ) end_POSTSUPERSCRIPT ) / italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT (31)

with the product logarithm W𝑊Witalic_W. Hence,

ε=(1+W(s0R0eR0(r01)))/R0.𝜀1𝑊subscript𝑠0subscript𝑅0superscript𝑒subscript𝑅0subscript𝑟01subscript𝑅0\displaystyle\varepsilon=(1+W(-s_{0}R_{0}e^{R_{0}(r_{0}-1)}))/R_{0}.italic_ε = ( 1 + italic_W ( - italic_s start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( italic_r start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT - 1 ) end_POSTSUPERSCRIPT ) ) / italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT . (32)

The peak of the epidemic i^=max(i)=i(t^)^𝑖𝑖𝑖^𝑡\hat{i}=\max(i)=i(\hat{t})over^ start_ARG italic_i end_ARG = roman_max ( italic_i ) = italic_i ( over^ start_ARG italic_t end_ARG ) occurs at the time t^^𝑡\hat{t}over^ start_ARG italic_t end_ARG for which i˙(t^)=0˙𝑖^𝑡0\dot{i}(\hat{t})=0over˙ start_ARG italic_i end_ARG ( over^ start_ARG italic_t end_ARG ) = 0 and thus s(t^)=1/R0𝑠^𝑡1subscript𝑅0s(\hat{t})=1/R_{0}italic_s ( over^ start_ARG italic_t end_ARG ) = 1 / italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, see eq. 2. Inserting this and r=1si𝑟1𝑠𝑖r=1-s-iitalic_r = 1 - italic_s - italic_i into eq. 30, we obtain

i^=max(i)=1r0(1+ln(s0R0))/R0^𝑖𝑖1subscript𝑟01subscript𝑠0subscript𝑅0subscript𝑅0\displaystyle\hat{i}=\max(i)=1-r_{0}-(1+\ln(s_{0}R_{0}))/R_{0}over^ start_ARG italic_i end_ARG = roman_max ( italic_i ) = 1 - italic_r start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT - ( 1 + roman_ln ( italic_s start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) ) / italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT (33)

High-infection-cost asymptote

The final number of cases sfsubscript𝑠𝑓s_{f}italic_s start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT can be calculated in the limit of large αR02much-greater-than𝛼superscriptsubscript𝑅02\alpha\gg R_{0}^{2}italic_α ≫ italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT, where sf=1/R0εsubscript𝑠𝑓1subscript𝑅0𝜀s_{f}=1/R_{0}-\varepsilonitalic_s start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT = 1 / italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT - italic_ε with ε𝜀\varepsilonitalic_ε small and assuming that s0>1/R0subscript𝑠01subscript𝑅0s_{0}>1/R_{0}italic_s start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT > 1 / italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT. We obtain

ε=2R02/α𝜀2superscriptsubscript𝑅02𝛼\displaystyle\varepsilon=2R_{0}^{2}/\alphaitalic_ε = 2 italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / italic_α (34)

from an expansion of eq. 28 in both 1/α1𝛼1/\alpha1 / italic_α and ε𝜀\varepsilonitalic_ε small and by matching order by order. This result is satisfied well, see fig. 3a. For the peak height, we obtain in the same limit, see section 6.2 for the calculation,

i^=max(i)=2R0(R01)/α^𝑖𝑖2subscript𝑅0subscript𝑅01𝛼\displaystyle\hat{i}=\max(i)=2R_{0}(R_{0}-1)/\alphaover^ start_ARG italic_i end_ARG = roman_max ( italic_i ) = 2 italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT - 1 ) / italic_α (35)

which is also satisfied well, see fig. 3b.

Scaling and phase diagram

Observing in fig. 3a that the excess cases are roughly constant at low α𝛼\alphaitalic_α and therefore well described by the non-behavioural limit, we obtain a crossover cost αssuperscriptsubscript𝛼𝑠\alpha_{s}^{\star}italic_α start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ⋆ end_POSTSUPERSCRIPT at which the non-behavioural and high infection cost asymptotes of eqs. 32 and 34, respectively, match

αs=2R03/(1+(W(s0R0eR0(r01)))\displaystyle\alpha_{s}^{\star}=2R_{0}^{3}/\left(1+(W(-s_{0}R_{0}e^{R_{0}(r_{0% }-1)})\right)italic_α start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ⋆ end_POSTSUPERSCRIPT = 2 italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT / ( 1 + ( italic_W ( - italic_s start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( italic_r start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT - 1 ) end_POSTSUPERSCRIPT ) ) (36)

For the infection peak, we similarly obtain a crossover cost αisuperscriptsubscript𝛼𝑖\alpha_{i}^{\star}italic_α start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ⋆ end_POSTSUPERSCRIPT from matching eqs. 33 and 35

αi=2R02(R01)/(R0(1r0)1ln(s0R0))superscriptsubscript𝛼𝑖2superscriptsubscript𝑅02subscript𝑅01subscript𝑅01subscript𝑟01subscript𝑠0subscript𝑅0\displaystyle\alpha_{i}^{\star}=2R_{0}^{2}(R_{0}-1)/\left(R_{0}(1-r_{0})-1-\ln% (s_{0}R_{0})\right)italic_α start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ⋆ end_POSTSUPERSCRIPT = 2 italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT - 1 ) / ( italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( 1 - italic_r start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) - 1 - roman_ln ( italic_s start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) ) (37)

These crossover values and the non-behavioural limits for ε𝜀\varepsilonitalic_ε and max(i)𝑖\max(i)roman_max ( italic_i ) can be used to achieve complete collapse of ε𝜀\varepsilonitalic_ε and max(i)𝑖\max(i)roman_max ( italic_i ) onto master curves, see fig. 3b and fig. 3d, respectively.

Both crossover values, αisuperscriptsubscript𝛼𝑖\alpha_{i}^{\star}italic_α start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ⋆ end_POSTSUPERSCRIPT and αssuperscriptsubscript𝛼𝑠\alpha_{s}^{\star}italic_α start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ⋆ end_POSTSUPERSCRIPT, determine different aspects of the “phase diagram” of social distancing, see fig. 4. The crossover αisuperscriptsubscript𝛼𝑖\alpha_{i}^{\star}italic_α start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ⋆ end_POSTSUPERSCRIPT for the infection peak describes a behavioural transition in the most intuitive signal of an epidemic. The infection peak also corresponds to the most restrictive value of social distancing, see eq. 24. For α<αi𝛼superscriptsubscript𝛼𝑖\alpha<\alpha_{i}^{\star}italic_α < italic_α start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ⋆ end_POSTSUPERSCRIPT social distancing is extremely weak, see e.g. for α=50𝛼50\alpha=50italic_α = 50 in fig. 2a (For R0=4subscript𝑅04R_{0}=4italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 4, αi59superscriptsubscript𝛼𝑖59\alpha_{i}^{\star}\approx 59italic_α start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ⋆ end_POSTSUPERSCRIPT ≈ 59 and αs139superscriptsubscript𝛼𝑠139\alpha_{s}^{\star}\approx 139italic_α start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ⋆ end_POSTSUPERSCRIPT ≈ 139). Social distancing is ultimately aimed at reducing excess cases. For αiααssuperscriptsubscript𝛼𝑖𝛼superscriptsubscript𝛼𝑠\alpha_{i}^{\star}\leq\alpha\leq\alpha_{s}^{\star}italic_α start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ⋆ end_POSTSUPERSCRIPT ≤ italic_α ≤ italic_α start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ⋆ end_POSTSUPERSCRIPT there is social distancing, but still only on a relatively short time frame, see the data for α=100𝛼100\alpha=100italic_α = 100 in fig. 2a. It starts to visibly affect the peak of the epidemic but not its duration, fig. 2c, and has a very limited effect on the total of cases, fig. 2b. This can be a viewed as the consequence of low gearing between the drop in infectivity and excess cases. Only for α>αs𝛼superscriptsubscript𝛼𝑠\alpha>\alpha_{s}^{\star}italic_α > italic_α start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ⋆ end_POSTSUPERSCRIPT is there considerable social distancing for an extended time, which then achieves a significant reduction in excess cases.

Refer to caption
Figure 5: Cost of the epidemic. Total epidemic cost relative to the cost of an infection, U/α𝑈𝛼-U/\alpha- italic_U / italic_α, as a function of infection cost α𝛼\alphaitalic_α under equilibrium social distancing. The corresponding non-behavioural, eq. 39, and high-infection-cost asymptotes, eq. 40 are indicated by dotted and dashed lines, respectively.

Utility

The utility, eq. 6, evaluated at the equilibrium behaviour can be directly calculated using the analytic solution

U𝑈\displaystyle Uitalic_U =α[rfr0+sf2(R0(rfr0)+ln(sf/s0))]absent𝛼delimited-[]subscript𝑟𝑓subscript𝑟0subscript𝑠𝑓2subscript𝑅0subscript𝑟𝑓subscript𝑟0subscript𝑠𝑓subscript𝑠0\displaystyle=-\alpha\left[r_{f}-r_{0}+\frac{s_{f}}{2}\left(R_{0}(r_{f}-r_{0})% +\ln(s_{f}/s_{0})\right)\right]= - italic_α [ italic_r start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT - italic_r start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT + divide start_ARG italic_s start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT end_ARG start_ARG 2 end_ARG ( italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( italic_r start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT - italic_r start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) + roman_ln ( italic_s start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT / italic_s start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) ) ] (38)

noting that sfsubscript𝑠𝑓s_{f}italic_s start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT and rf=1sfsubscript𝑟𝑓1subscript𝑠𝑓r_{f}=1-s_{f}italic_r start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT = 1 - italic_s start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT depend on α𝛼\alphaitalic_α and R0subscript𝑅0R_{0}italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT. The total infection cost is given by α(rfr0)𝛼subscript𝑟𝑓subscript𝑟0-\alpha(r_{f}-r_{0})- italic_α ( italic_r start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT - italic_r start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) with the remainder being the total social distancing cost. Especially for intermediary R0subscript𝑅0R_{0}italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT and high infection costs α𝛼\alphaitalic_α, equilibrium behaviour strongly reduces the total epidemic cost, see fig. 5. Again, we investigate the two limiting cases: For low α𝛼\alphaitalic_α, we obtain with eq. 31

U𝑈\displaystyle Uitalic_U =α[1+1R0W(s0R0eR0(r01))r0]absent𝛼delimited-[]11subscript𝑅0𝑊subscript𝑠0subscript𝑅0superscript𝑒subscript𝑅0subscript𝑟01subscript𝑟0\displaystyle=-\alpha\left[1+\frac{1}{R_{0}}W(-s_{0}R_{0}e^{R_{0}(r_{0}-1)})-r% _{0}\right]= - italic_α [ 1 + divide start_ARG 1 end_ARG start_ARG italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG italic_W ( - italic_s start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( italic_r start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT - 1 ) end_POSTSUPERSCRIPT ) - italic_r start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ] (39)

For high α𝛼\alphaitalic_α, i.e. α/R021much-greater-than𝛼superscriptsubscript𝑅021\alpha/R_{0}^{2}\gg 1italic_α / italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ≫ 1, with sf=1R0ε=1R02R02α1/R0subscript𝑠𝑓1subscript𝑅0𝜀1subscript𝑅02superscriptsubscript𝑅02𝛼1subscript𝑅0s_{f}=\frac{1}{R_{0}}-\varepsilon=\frac{1}{R_{0}}-\frac{2R_{0}^{2}}{\alpha}% \approx 1/R_{0}italic_s start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT = divide start_ARG 1 end_ARG start_ARG italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG - italic_ε = divide start_ARG 1 end_ARG start_ARG italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG - divide start_ARG 2 italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_α end_ARG ≈ 1 / italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, we obtain

U𝑈\displaystyle Uitalic_U =α[32(1r0)3+ln(R0s0)2R0]absent𝛼delimited-[]321subscript𝑟03subscript𝑅0subscript𝑠02subscript𝑅0\displaystyle=-\alpha\left[\frac{3}{2}(1-r_{0})-\frac{3+\ln(R_{0}s_{0})}{2R_{0% }}\right]= - italic_α [ divide start_ARG 3 end_ARG start_ARG 2 end_ARG ( 1 - italic_r start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) - divide start_ARG 3 + roman_ln ( italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_s start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) end_ARG start_ARG 2 italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG ] (40)

5 Discussion and conclusion

In summary, we have identified an analytic solution for the Nash equilibrium behaviour for social distancing during an epidemic. We leveraged this solution to obtain (1) a simple relationship between the strength of rational social distancing and the current number of cases, eq. 24; (2) scaling results for the total number of cases, eq. 34, and the infection peak, eq. 35 which only depend on the basic reproduction number and the cost of contracting the disease; (3) characteristic infection costs, eqs. 36 and 37, that divide regimes of strong and weak social distancing and depend only on the basic reproduction number of the disease; (4) a closed form expression for the value of the utility, eq. 38. These four results represent a remarkable simplification of a complex optimisation problem.

We believe our work to be useful to policy makers because it yields a simple, albeit idealised, classification of the impact of self-organised social distancing during epidemics and thus can serve as guide for policy. Given the basic reproduction number of a given disease R0subscript𝑅0R_{0}italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT and its estimated cost of infection α𝛼\alphaitalic_α, we show that one can either expect negligible social distancing from the population, when the infection cost is below a characteristic cost, or substantial social distancing when it is above.

There is an ongoing debate about the degree to which behaviour of individuals is truly rational, as we (and others) assume. In this context our most significant result is that the rational decision making process seems to be intuitively accessible to most members of the population: rational social distancing is proportional to the infection cost and to the current number of cases. It is remarkable that the rational response we derive can be condensed into such a simple heuristic, understandable to a typical member of the population. While it may indeed be a challenge for such individuals to derive our results for themselves, a policymaker could communicate this simple heuristic, to be adopted by the population in order to assist them in targeting truly rational behaviour. It is not unrealistic to expect this advice to influence the population decision making, especially given that it can be shown to be in each individual’s self interest. In this sense the present work may itself help to “bootstrap” such rational behaviour.

While rational behaviour is not the mathematically optimal solution that maximises utility, as would be accessible under arbitrarily precise government control, it is relatively close to it. Rational behaviour also has the advantage of being stable, in the sense that it suppresses the detrimental behaviour of freeloaders, who are worse off if they deviate from the Nash equilibrium behaviour. The fact that rational behaviour is so desirable means that new tools that enable policymakers to help individuals target rational behaviour, like the ones we provide here, may be extremely valuable.

6 Methods

6.1 Vaccination salvage term

A perfect vaccine applied to the whole population at time tfsubscript𝑡𝑓t_{f}italic_t start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT corresponds to immediately moving the susceptible fraction of the population into the recovered compartment, s(t>tf)=0𝑠𝑡subscript𝑡𝑓0s(t>t_{f})=0italic_s ( italic_t > italic_t start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT ) = 0 and ψs(t>tf)=0subscript𝜓𝑠𝑡subscript𝑡𝑓0\psi_{s}(t>t_{f})=0italic_ψ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( italic_t > italic_t start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT ) = 0. Eqs. (2) reduce to

i˙˙𝑖\displaystyle\dot{i}over˙ start_ARG italic_i end_ARG =iabsent𝑖\displaystyle=-i= - italic_i (41)

The remaining infectious recover exponentially, with i(tf)=if𝑖subscript𝑡𝑓subscript𝑖𝑓i(t_{f})=i_{f}italic_i ( italic_t start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT ) = italic_i start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT,

i(t>tf)=ifexp[(ttf)]𝑖𝑡subscript𝑡𝑓subscript𝑖𝑓𝑡subscript𝑡𝑓\displaystyle i(t>t_{f})=i_{f}\exp[-(t-t_{f})]italic_i ( italic_t > italic_t start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT ) = italic_i start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT roman_exp [ - ( italic_t - italic_t start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT ) ] (42)

Analogously for the individual probabilities, with ψi(tf)=ψi,fsubscript𝜓𝑖subscript𝑡𝑓subscript𝜓𝑖𝑓\psi_{i}(t_{f})=\psi_{i,f}italic_ψ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_t start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT ) = italic_ψ start_POSTSUBSCRIPT italic_i , italic_f end_POSTSUBSCRIPT,

ψi(t>tf)=ψi,fexp[(ttf)]subscript𝜓𝑖𝑡subscript𝑡𝑓subscript𝜓𝑖𝑓𝑡subscript𝑡𝑓\displaystyle\psi_{i}(t>t_{f})=\psi_{i,f}\exp[-(t-t_{f})]italic_ψ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_t > italic_t start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT ) = italic_ψ start_POSTSUBSCRIPT italic_i , italic_f end_POSTSUBSCRIPT roman_exp [ - ( italic_t - italic_t start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT ) ] (43)

Since nobody can get freshly infected, the population selects pre-epidemic behaviour, κ(t>tf)=R0𝜅𝑡subscript𝑡𝑓subscript𝑅0\kappa(t>t_{f})=R_{0}italic_κ ( italic_t > italic_t start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT ) = italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT. The contribution to the utility Ufsubscript𝑈𝑓U_{f}italic_U start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT that arises from the recovery process after tfsubscript𝑡𝑓t_{f}italic_t start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT can be written in analogy to eq. 7

Ufsubscript𝑈𝑓\displaystyle U_{f}italic_U start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT =tfft[αψi(t)(κ(t)R0)2]𝑑tabsentsuperscriptsubscriptsubscript𝑡𝑓superscript𝑓𝑡delimited-[]𝛼subscript𝜓𝑖𝑡superscript𝜅𝑡subscript𝑅02differential-d𝑡\displaystyle=\int_{t_{f}}^{\infty}f^{-t}\left[-\alpha\>\psi_{i}(t)-(\kappa(t)% -R_{0})^{2}\right]dt= ∫ start_POSTSUBSCRIPT italic_t start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_f start_POSTSUPERSCRIPT - italic_t end_POSTSUPERSCRIPT [ - italic_α italic_ψ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_t ) - ( italic_κ ( italic_t ) - italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ] italic_d italic_t

This can be integrated to yield

Uf=ftfαψi,f1+lnfsubscript𝑈𝑓superscript𝑓subscript𝑡𝑓𝛼subscript𝜓𝑖𝑓1𝑓\displaystyle U_{f}=-f^{-t_{f}}\alpha\ \frac{\psi_{i,f}}{1+\ln f}italic_U start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT = - italic_f start_POSTSUPERSCRIPT - italic_t start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT end_POSTSUPERSCRIPT italic_α divide start_ARG italic_ψ start_POSTSUBSCRIPT italic_i , italic_f end_POSTSUBSCRIPT end_ARG start_ARG 1 + roman_ln italic_f end_ARG (44)

6.2 High-infection cost asymptote for the infection peak height

In the large α𝛼\alphaitalic_α limit α/R021much-greater-than𝛼superscriptsubscript𝑅021\alpha/R_{0}^{2}\gg 1italic_α / italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ≫ 1 we have sf1/R0subscript𝑠𝑓1subscript𝑅0s_{f}\approx 1/R_{0}italic_s start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT ≈ 1 / italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT from eq. 29, hence

b=αsf/2α/(2R0)𝑏𝛼subscript𝑠𝑓2𝛼2subscript𝑅0\displaystyle b=\alpha s_{f}/2\approx\alpha/(2R_{0})italic_b = italic_α italic_s start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT / 2 ≈ italic_α / ( 2 italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) (45)

large according to bR0much-greater-than𝑏subscript𝑅0b\gg R_{0}italic_b ≫ italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT. The infection peak i˙=0˙𝑖0\dot{i}=0over˙ start_ARG italic_i end_ARG = 0 occurs at i^=max(i)^𝑖𝑖\hat{i}=\max(i)over^ start_ARG italic_i end_ARG = roman_max ( italic_i ) where eq. 2 yields 0=ks10𝑘𝑠10=ks-10 = italic_k italic_s - 1. Using eq. 24 we have

1=ks=s(R0i^b)s=1R0i^b1𝑘𝑠𝑠subscript𝑅0^𝑖𝑏𝑠1subscript𝑅0^𝑖𝑏\displaystyle 1=ks=s(R_{0}-\hat{i}b)\Rightarrow s=\frac{1}{R_{0}-\hat{i}b}1 = italic_k italic_s = italic_s ( italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT - over^ start_ARG italic_i end_ARG italic_b ) ⇒ italic_s = divide start_ARG 1 end_ARG start_ARG italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT - over^ start_ARG italic_i end_ARG italic_b end_ARG (46)

with the sum rule,

i^=1rs=1r1R0i^b^𝑖1𝑟𝑠1𝑟1subscript𝑅0^𝑖𝑏\displaystyle\hat{i}=1-r-s=1-r-\frac{1}{R_{0}-\hat{i}b}over^ start_ARG italic_i end_ARG = 1 - italic_r - italic_s = 1 - italic_r - divide start_ARG 1 end_ARG start_ARG italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT - over^ start_ARG italic_i end_ARG italic_b end_ARG (47)
(R0i^b)i^=(1r)(R0i^b)1absentsubscript𝑅0^𝑖𝑏^𝑖1𝑟subscript𝑅0^𝑖𝑏1\displaystyle\Rightarrow(R_{0}-\hat{i}b)\hat{i}=(1-r)(R_{0}-\hat{i}b)-1⇒ ( italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT - over^ start_ARG italic_i end_ARG italic_b ) over^ start_ARG italic_i end_ARG = ( 1 - italic_r ) ( italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT - over^ start_ARG italic_i end_ARG italic_b ) - 1 (48)

This yields a quadratic equation for i^^𝑖\hat{i}over^ start_ARG italic_i end_ARG with physical root

i^=^𝑖absent\displaystyle\hat{i}=\ over^ start_ARG italic_i end_ARG = R0+b(1r)2bsubscript𝑅0𝑏1𝑟2𝑏\displaystyle\frac{R_{0}+b(1-r)}{2b}divide start_ARG italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT + italic_b ( 1 - italic_r ) end_ARG start_ARG 2 italic_b end_ARG (49)
(R0+b(1r))2+4b(1(1r)R0)2bsuperscriptsubscript𝑅0𝑏1𝑟24𝑏11𝑟subscript𝑅02𝑏\displaystyle-\frac{\sqrt{(R_{0}+b(1-r))^{2}+4b(1-(1-r)R_{0})}}{2b}- divide start_ARG square-root start_ARG ( italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT + italic_b ( 1 - italic_r ) ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + 4 italic_b ( 1 - ( 1 - italic_r ) italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) end_ARG end_ARG start_ARG 2 italic_b end_ARG (50)
\displaystyle\approx\ (1r)2[4b((1r)R01)2(R0+b(1r))2]1𝑟2delimited-[]4𝑏1𝑟subscript𝑅012superscriptsubscript𝑅0𝑏1𝑟2\displaystyle\frac{(1-r)}{2}\left[\frac{4b((1-r)R_{0}-1)}{2(R_{0}+b(1-r))^{2}}\right]divide start_ARG ( 1 - italic_r ) end_ARG start_ARG 2 end_ARG [ divide start_ARG 4 italic_b ( ( 1 - italic_r ) italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT - 1 ) end_ARG start_ARG 2 ( italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT + italic_b ( 1 - italic_r ) ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ] (51)

For large α𝛼\alphaitalic_α the infection peak occurs early in the epidemic, when r=1si1𝑟1𝑠𝑖much-less-than1r=1-s-i\ll 1italic_r = 1 - italic_s - italic_i ≪ 1, e.g. see fig. 1 for α=400𝛼400\alpha=400italic_α = 400. Using eq. 45 and recalling also that bR0much-greater-than𝑏subscript𝑅0b\gg R_{0}italic_b ≫ italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT we find

i^^𝑖\displaystyle\hat{i}over^ start_ARG italic_i end_ARG 2R0(R01)αabsent2subscript𝑅0subscript𝑅01𝛼\displaystyle\approx\frac{2R_{0}(R_{0}-1)}{\alpha}≈ divide start_ARG 2 italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT - 1 ) end_ARG start_ARG italic_α end_ARG (52)
Acknowledgements.
We would like to dedicate this work to the memory of Prof. George Rowlands, who passed away in early 2021 and who was involved in many of the early discussions leading to this work. We thank Paul François, Shuhei Horiguchi, Tetsuya J. Kobayashi, and Takehiro Tottori for helpful discussions. This work was supported by the Grants-in-Aid for Scientific Research (JSPS KAKENHI) under Grants No. 20H00129 (RY), 20H05619 (RY), 22H04841 (SKS), 22K14012 (SKS), 23H04508 (JJM), and the JSPS Core-to-Core Program “Advanced core-to-core network for the physics of self-organizing active matter” JPJSCCA20230002 (all of us). MST acknowledges the generous support of visiting fellowships from JSPS Fellowship, ID L19547, the Leverhulme Trust, Ref. IAF-2019-019, and the kind hospitality of the Yamamoto group. The funders had no role in study design, data collection and analysis, decision to publish, or preparation of the manuscript. The authors declare that there are no conflicts of interest. Author contributions: All authors designed the research. SKS and MST performed the research. JJM assisted with code development and numerical methods. All authors wrote the paper.

References

  • Reluga [2010] T. C. Reluga, Game Theory of Social Distancing in Response to an Epidemic, PLoS Comput. Biol. 6, e1000793 (2010).
  • Fenichel et al. [2011] E. P. Fenichel, C. Castillo-Chavez, M. G. Ceddia, G. Chowell, P. A. G. Parra, G. J. Hickling, G. Holloway, R. Horan, B. Morin, C. Perrings, M. Springborn, L. Velazquez, and C. Villalobos, Adaptive human behavior in epidemiological models, Proc. Natl. Acad. Sci. 108, 6306 (2011).
  • Wang et al. [2016] Z. Wang, C. T. Bauch, S. Bhattacharyya, A. D’Onofrio, P. Manfredi, M. Perc, N. Perra, M. Salathé, and D. Zhao, Statistical physics of vaccination, Physics Reports 664, 1 (2016)arXiv:1608.09010 .
  • Chang et al. [2020] S. L. Chang, M. Piraveenan, P. Pattison, and M. Prokopenko, Game theoretic modelling of infectious disease dynamics and intervention methods: a review, Journal of Biological Dynamics 14, 57 (2020)arXiv:1901.04143 .
  • Verelst et al. [2016] F. Verelst, L. Willem, and P. Beutels, Behavioural change models for infectious disease transmission: A systematic review (2010-2015), Journal of the Royal Society Interface 1310.1098/rsif.2016.0820 (2016).
  • Bhattacharyya and Reluga [2019] S. Bhattacharyya and T. Reluga, Game dynamic model of social distancing while cost of infection varies with epidemic burden, IMA J. Appl. Math. (Institute Math. Its Appl. 84, 23 (2019).
  • Makris and Toxvaerd [2020] M. Makris and F. Toxvaerd, Cambridge Working Papers in Economics, Cambridge Working Papers in Economics 2097 (2020).
  • Yan et al. [2021] Y. Yan, A. A. Malik, J. Bayham, E. P. Fenichel, C. Couzens, and S. B. Omer, Measuring voluntary and policy-induced social distancing behavior during the COVID-19 pandemic, Proceedings of the National Academy of Sciences of the United States of America 118, 1 (2021).
  • Reluga and Galvani [2011] T. C. Reluga and A. P. Galvani, A general approach for population games with application to vaccination, Math. Biosci. 230, 67 (2011).
  • Lux [2021] T. Lux, The social dynamics of COVID-19, Physica A: Statistical Mechanics and its Applications 567, 125710 (2021).
  • Kahneman [2003] D. Kahneman, Maps of bounded rationality: Psychology for behavioral economics, American economic review 93, 1449 (2003).
  • McAdams [2020] D. McAdams, Nash SIR: An Economic-Epidemiological Model of Strategic Behavior During a Viral Epidemic, Covid Economics 10.2139/ssrn.3593272 (2020).
  • Eichenbaum et al. [2021] M. S. Eichenbaum, S. Rebelo, and M. Trabandt, The Macroeconomics of Epidemics, Review of Financial Studies 34, 5149 (2021).
  • Rowthorn and Toxvaerd [2020] R. Rowthorn and F. Toxvaerd, The optimal control of infectious diseases via prevention and treatment, Cambridge Working Papers in Economics 2027 (University of Cambridge, 2020).
  • Toxvaerd and Rowthorn [2020] F. Toxvaerd and R. Rowthorn, On the management of population immunity, Cambridge Working Papers in Economics 2080 (University of Cambridge, 2020).
  • Li et al. [2017] J. Li, D. V. Lindberg, R. A. Smith, and T. C. Reluga, Provisioning of Public Health Can Be Designed to Anticipate Public Policy Responses, Bull. Math. Biol. 79, 163 (2017).
  • Bethune and Korinek [2020] Z. A. Bethune and A. Korinek, COVID-19 infection externalities: trading off lives vs. livelihoods, Working Paper 27009 (National Bureau of Economic Research, Cambridge, MA, 2020).
  • Aurell et al. [2022] A. Aurell, R. Carmona, G. Dayanikli, and M. Laurière, Optimal Incentives to Mitigate Epidemics: A Stackelberg Mean Field Game Approach, SIAM Journal on Control and Optimization 60, S294 (2022)arXiv:2011.03105 .
  • Schnyder et al. [2023a] S. K. Schnyder, J. J. Molina, R. Yamamoto, and M. S. Turner, Rational social distancing policy during epidemics with limited healthcare capacity, PLOS Computational Biology 19, e1011533 (2023a)arXiv:2205.00684 .
  • Althouse et al. [2010] B. M. Althouse, T. C. Bergstrom, and C. T. Bergstrom, A public choice framework for controlling transmissible and evolving diseases, Proceedings of the National Academy of Sciences of the United States of America 107, 1696 (2010).
  • Acemoglu et al. [2020] D. Acemoglu, V. Chernozhukov, I. Werning, and M. D. Whinston, Optimal targeted lockdowns in a multi-group SIR model, Working Paper 27102 (National Bureau of Economic Research, 2020).
  • Prem et al. [2017] K. Prem, A. R. Cook, and M. Jit, Projecting social contact matrices in 152 countries using contact surveys and demographic data, PLOS Computational Biology 13, e1005697 (2017).
  • Huang et al. [2022] C. I. Huang, R. E. Crump, P. E. Brown, S. E. Spencer, E. M. Miaka, C. Shampa, M. J. Keeling, and K. S. Rock, Identifying regions for enhanced control of gambiense sleeping sickness in the Democratic Republic of Congo, Nature Communications 13, 1 (2022).
  • Tildesley et al. [2022] M. J. Tildesley, A. Vassall, S. Riley, M. Jit, F. Sandmann, E. M. Hill, R. N. Thompson, B. D. Atkins, J. Edmunds, L. Dyson, and M. J. Keeling, Optimal health and economic impact of non-pharmaceutical intervention measures prior and post vaccination in England: a mathematical modelling study, Royal Society Open Science 910.1098/rsos.211746 (2022).
  • Keeling et al. [2022] M. J. Keeling, L. Dyson, M. J. Tildesley, E. M. Hill, and S. Moore, Comparison of the 2021 COVID-19 roadmap projections against public health data in England, Nature communications 13, 4924 (2022).
  • He et al. [2013] D. He, J. Dushoff, T. Day, J. Ma, and D. J. Earn, Inferring the causes of the three waves of the 1918 influenza pandemic in England and Wales, Proceedings of the Royal Society B: Biological Sciences 28010.1098/rspb.2013.1345 (2013).
  • Giannitsarou et al. [2021] C. Giannitsarou, S. Kissler, and F. Toxvaerd, Waning Immunity and the Second Wave: Some Projections for SARS-CoV-2, American Economic Review: Insights 3, 321 (2021).
  • Schwarzendahl et al. [2022] F. J. Schwarzendahl, J. Grauer, B. Liebchen, and H. Löwen, Mutation induced infection waves in diseases like COVID-19, Scientific Reports 12, 1 (2022).
  • Mossong et al. [2008] J. Mossong, N. Hens, M. Jit, P. Beutels, K. Auranen, R. Mikolajczyk, M. Massari, S. Salmaso, G. S. Tomba, J. Wallinga, J. Heijne, M. Sadkowska-Todys, M. Rosinska, and W. J. Edmunds, Social contacts and mixing patterns relevant to the spread of infectious diseases, PLoS Medicine 5, 0381 (2008).
  • Tildesley et al. [2010] M. J. Tildesley, T. A. House, M. C. Bruhn, R. J. Curry, M. O’Neil, J. L. Allpress, G. Smith, and M. J. Keeling, Impact of spatial clustering on disease transmission and optimal control, Proceedings of the National Academy of Sciences of the United States of America 107, 1041 (2010).
  • Sun et al. [2021] K. Sun, W. Wang, L. Gao, Y. Wang, K. Luo, L. Ren, Z. Zhan, X. Chen, S. Zhao, Y. Huang, Q. Sun, Z. Liu, M. Litvinova, A. Vespignani, M. Ajelli, C. Viboud, and H. Yu, Transmission heterogeneities, kinetics, and controllability of SARS-CoV-2, Science 371, eabe2424 (2021).
  • Hill et al. [2023] E. M. Hill, N. S. Prosser, P. E. Brown, E. Ferguson, M. J. Green, J. Kaler, M. J. Keeling, and M. J. Tildesley, Incorporating heterogeneity in farmer disease control behaviour into a livestock disease transmission model, Preventive Veterinary Medicine 219, 106019 (2023).
  • Chandrasekhar et al. [2021] A. G. Chandrasekhar, P. Goldsmith-Pinkham, M. O. Jackson, and S. Thau, Interacting regional policies in containing a disease, Proceedings of the National Academy of Sciences of the United States of America 118, 1 (2021)arXiv:2008.10745 .
  • Holme and Saramäki [2012] P. Holme and J. Saramäki, Temporal networks, Physics Reports 519, 97 (2012)arXiv:1108.1780 .
  • Holme and Masuda [2015] P. Holme and N. Masuda, The basic reproduction number as a predictor for epidemic outbreaks in temporal networks, PLoS ONE 10, 1 (2015)arXiv:1407.6598 .
  • Ferguson et al. [2006] N. M. Ferguson, D. A. T. Cummings, C. Fraser, J. C. Cajka, P. C. Cooley, and D. S. Burke, Strategies for mitigating an influenza pandemic, Nature 442, 448 (2006).
  • Tanimoto [2018] J. Tanimoto, Social Dilemma Analysis of the Spread of Infectious Disease (2018) pp. 155–216.
  • Mellacher [2020] P. Mellacher, COVID-Town: An Integrated Economic-Epidemiological Agent-Based Model, GSC discussion papers 23 (Graz Schumpeter Centre, 2020).
  • Grauer et al. [2020] J. Grauer, H. Löwen, and B. Liebchen, Strategic spatiotemporal vaccine distribution increases the survival rate in an infectious disease like Covid-19, Scientific Reports 10, 1 (2020)arXiv:2005.04056 .
  • Yong and Zhou [1999] J. Yong and X. Y. Zhou, Stochastic controls: Hamiltonian systems and HJB equations, Vol. 43 (Springer Science & Business Media, 1999).
  • Lorch et al. [2018] L. Lorch, A. De, S. Bhatt, W. Trouleau, U. Upadhyay, and M. Gomez-Rodriguez, Stochastic Optimal Control of Epidemic Processes in Networks,   (2018)arXiv:1810.13043 .
  • Tottori and Kobayashi [2022] T. Tottori and T. J. Kobayashi, Memory-Limited Partially Observable Stochastic Control and Its Mean-Field Control Approach, Entropy 24, 1 (2022)arXiv:2203.10682 .
  • Tottori and Kobayashi [2023a] T. Tottori and T. J. Kobayashi, Forward-Backward Sweep Method for the System of HJB-FP Equations in Memory-Limited Partially Observable Stochastic Control, Entropy 2510.3390/e25020208 (2023a).
  • Tottori and Kobayashi [2023b] T. Tottori and T. J. Kobayashi, Decentralized Stochastic Control with Finite-Dimensional Memories: A Memory Limitation Approach, Entropy 25, 1 (2023b).
  • Barnett et al. [2023] M. Barnett, G. Buchak, and C. Yannelis, Epidemic responses under uncertainty, Proceedings of the National Academy of Sciences of the United States of America 120, 1 (2023).
  • Shea et al. [2023] K. Shea, R. K. Borchering, W. J. M. Probert, E. Howerton, T. L. Bogich, S.-L. Li, W. G. van Panhuis, C. Viboud, R. Aguás, A. A. Belov, S. H. Bhargava, S. M. Cavany, J. C. Chang, C. Chen, J. Chen, S. Chen, Y. Chen, L. M. Childs, C. C. Chow, I. Crooker, S. Y. Del Valle, G. España, G. Fairchild, R. C. Gerkin, T. C. Germann, Q. Gu, X. Guan, L. Guo, G. R. Hart, T. J. Hladish, N. Hupert, D. Janies, C. C. Kerr, D. J. Klein, E. Y. Klein, G. Lin, C. Manore, L. A. Meyers, J. E. Mittler, K. Mu, R. C. Núñez, R. J. Oidtman, R. Pasco, A. Pastore y Piontti, R. Paul, C. A. B. Pearson, D. R. Perdomo, T. A. Perkins, K. Pierce, A. N. Pillai, R. C. Rael, K. Rosenfeld, C. W. Ross, J. A. Spencer, A. B. Stoltzfus, K. B. Toh, S. Vattikuti, A. Vespignani, L. Wang, L. J. White, P. Xu, Y. Yang, O. N. Yogurtcu, W. Zhang, Y. Zhao, D. Zou, M. J. Ferrari, D. Pannell, M. J. Tildesley, J. Seifarth, E. Johnson, M. Biggerstaff, M. A. Johansson, R. B. Slayton, J. D. Levander, J. Stazer, J. Kerr, and M. C. Runge, Multiple models for outbreak decision support in the face of uncertainty, Proceedings of the National Academy of Sciences 120, 2017 (2023).
  • Kantner and Koprucki [2020] M. Kantner and T. Koprucki, Beyond just “flattening the curve”: Optimal control of epidemics with purely non-pharmaceutical interventions, Journal of Mathematics in Industry 1010.1186/s13362-020-00091-3 (2020), arXiv:2004.09471 .
  • Köhler et al. [2021] J. Köhler, L. Schwenkel, A. Koch, J. Berberich, P. Pauli, and F. Allgöwer, Robust and optimal predictive control of the COVID-19 outbreak, Annual Reviews in Control 51, 525 (2021)arXiv:2005.03580 .
  • Morris et al. [2021] D. H. Morris, F. W. Rossine, J. B. Plotkin, and S. A. Levin, Optimal, near-optimal, and robust epidemic control, Communications Physics 4, 1 (2021)arXiv:2004.02209 .
  • Adhikari et al. [2020] R. Adhikari, A. Bolitho, F. Caballero, M. E. Cates, J. Dolezal, T. Ekeh, J. Guioth, R. L. Jack, J. Kappler, L. Kikuchi, H. Kobayashi, Y. I. Li, J. D. Peterson, P. Pietzonka, B. Remez, P. B. Rohrbach, R. Singh, and G. Turk, Inference, prediction and optimization of non-pharmaceutical interventions using compartment models: the PyRoss library,   (2020)arXiv:2005.09625 .
  • Pietzonka et al. [2021] P. Pietzonka, E. Brorson, W. Bankes, M. E. Cates, R. L. Jack, and R. Adhikari, Bayesian inference across multiple models suggests a strong increase in lethality of COVID-19 in late 2020 in the UK, PLoS ONE 16, 1 (2021).
  • Li et al. [2021] Y. I. Li, G. Turk, P. B. Rohrbach, P. Pietzonka, J. Kappler, R. Singh, J. Dolezal, T. Ekeh, L. Kikuchi, J. D. Peterson, A. Bolitho, H. Kobayashi, M. E. Cates, R. Adhikari, and R. L. Jack, Efficient Bayesian inference of fully stochastic epidemiological models with applications to COVID-19, Royal Society Open Science 8, 211065 (2021).
  • Molina et al. [2022] J. J. Molina, S. K. Schnyder, M. S. Turner, and R. Yamamoto, Nash Neural Networks : Inferring Utilities from Optimal Behaviour,   (2022)arXiv:2203.13432 .
  • Mellacher [2023] P. Mellacher, The impact of corona populism: Empirical evidence from Austria and theory, Journal of Economic Behavior & Organization 209, 113 (2023).
  • Kermack and McKendrick [1927] W. O. Kermack and A. McKendrick, A contribution to the mathematical theory of epidemics, Proceedings of the Royal Society of London. Series A, Containing Papers of a Mathematical and Physical Character 115, 700 (1927).
  • Miller [2012] J. C. Miller, A Note on the Derivation of Epidemic Final Sizes, Bulletin of Mathematical Biology 74, 2125 (2012).
  • Harko et al. [2014] T. Harko, F. S. Lobo, and M. K. Mak, Exact analytical solutions of the Susceptible-Infected-Recovered (SIR) epidemic model and of the SIR model with equal death and birth rates, Applied Mathematics and Computation 236, 184 (2014)arXiv:1403.2160 .
  • Miller [2017] J. C. Miller, Mathematical models of SIR disease spread with combined non-sexual and sexual transmission routes, Infectious Disease Modelling 2, 35 (2017)1609.08108 .
  • Kröger and Schlickeiser [2020] M. Kröger and R. Schlickeiser, Analytical solution of the SIR-model for the temporal evolution of epidemics. Part A: time-independent reproduction factor, Journal of Physics A: Mathematical and Theoretical 5310.1088/1751-8121/abc65d (2020).
  • Bauch et al. [2003] C. T. Bauch, A. P. Galvani, and D. J. Earn, Group interest versus self-interest in smallpox vaccination policy, Proceedings of the National Academy of Sciences of the United States of America 100, 10564 (2003).
  • Bauch and Earn [2004] C. T. Bauch and D. J. Earn, Vaccination and the theory of games, Proceedings of the National Academy of Sciences of the United States of America 101, 13391 (2004).
  • Reluga et al. [2006] T. C. Reluga, C. T. Bauch, and A. P. Galvani, Evolving public perceptions and stability in vaccine uptake, Mathematical Biosciences 204, 185 (2006).
  • Tildesley et al. [2006] M. J. Tildesley, N. J. Savill, D. J. Shaw, R. Deardon, S. P. Brooks, M. E. Woolhouse, B. T. Grenfell, and M. J. Keeling, Optimal reactive vaccination strategies for a foot-and-mouth outbreak in the UK, Nature 440, 83 (2006).
  • Chen and Toxvaerd [2014] F. Chen and F. Toxvaerd, The economics of vaccination, Journal of Theoretical Biology 363, 105 (2014).
  • Moore et al. [2021] S. Moore, E. M. Hill, L. Dyson, M. J. Tildesley, and M. J. Keeling, Modelling optimal vaccination strategy for SARS-CoV-2 in the UK, PLoS Computational Biology 17, 1 (2021).
  • Moore et al. [2022] S. Moore, E. M. Hill, L. Dyson, M. J. Tildesley, and M. J. Keeling, Retrospectively modeling the effects of increased global vaccine sharing on the COVID-19 pandemic, Nature Medicine 28, 2416 (2022).
  • Hill et al. [2022] E. M. Hill, N. S. Prosser, E. Ferguson, J. Kaler, M. J. Green, M. J. Keeling, and M. J. Tildesley, Modelling livestock infectious disease control policy under differing social perspectives on vaccination behaviour, PLOS Computational Biology 18, e1010235 (2022).
  • Keeling et al. [2023] M. J. Keeling, S. Moore, B. S. Penman, and E. M. Hill, The impacts of SARS-CoV-2 vaccine dose separation and targeting on the COVID-19 epidemic in England, Nature Communications 14, 1 (2023).
  • Kucharski et al. [2020] A. J. Kucharski, P. Klepac, A. J. Conlan, S. M. Kissler, M. L. Tang, H. Fry, J. R. Gog, W. J. Edmunds, J. C. Emery, G. Medley, J. D. Munday, T. W. Russell, Q. J. Leclerc, C. Diamond, S. R. Procter, A. Gimma, F. Y. Sun, H. P. Gibbs, A. Rosello, K. van Zandvoort, S. Hué, S. R. Meakin, A. K. Deol, G. Knight, T. Jombart, A. M. Foss, N. I. Bosse, K. E. Atkins, B. J. Quilty, R. Lowe, K. Prem, S. Flasche, C. A. Pearson, R. M. Houben, E. S. Nightingale, A. Endo, D. C. Tully, Y. Liu, J. Villabona-Arenas, K. O’Reilly, S. Funk, R. M. Eggo, M. Jit, E. M. Rees, J. Hellewell, S. Clifford, C. I. Jarvis, S. Abbott, M. Auzenbergs, N. G. Davies, and D. Simons, Effectiveness of isolation, testing, contact tracing, and physical distancing on reducing transmission of SARS-CoV-2 in different settings: a mathematical modelling study, The Lancet Infectious Diseases 20, 1151 (2020).
  • Piguillem and Shi [2020] F. Piguillem and L. Shi, Optimal COVID-19 Quarantine and Testing Policies, EIEF Working Paper 20/04 (2020).
  • Schnyder et al. [2023b] S. K. Schnyder, J. J. Molina, R. Yamamoto, and M. S. Turner, Rational social distancing in epidemics with uncertain vaccination timing, PLOS ONE 18, e0288963 (2023b)arXiv:2305.13618 .
  • Bensoussan et al. [2013] A. Bensoussan, J. Frehse, and P. Yam, Mean Field Games and Mean Field Type Control Theory (Springer, 2013).
  • Carmona and Delarue [2018] R. Carmona and F. Delarue, Probabilistic Theory of Mean Field Games with Applications I (Springer, 2018).
  • Lenhart and Workman [2007] S. Lenhart and J. Workman, Optimal Control Applied to Biological Models (Chapman and Hall/CRC, 2007).
  • Pontryagin et al. [1986] L. S. Pontryagin, V. Boltyanskii, R. V. Gamkrelidze, and E. F. Mishchenko, The Mathematical Theory of Optimal Processes (Gordon and Breach Science Publishers, 1986).