Location via proxy:   [ UP ]  
[Report a bug]   [Manage cookies]                

A refractory density approach to a multi-scale SEIRS epidemic model

Anton Chizhov Laurent Pujo-Menjouet pujo@math.univ-lyon1.fr Tilo Schwalger Mattia Sensi
(July 2, 2024)
Abstract

We propose a novel multi-scale modeling framework for infectious disease spreading, borrowing ideas and modeling tools from the so-called Refractory Density (RD) approach. We introduce a microscopic model that describes the probability of infection for a single individual and the evolution of the disease within their body. From the individual-level description, we then present the corresponding population-level model of epidemic spreading on the mesoscopic and macroscopic scale. We conclude with numerical illustrations taking into account either a white Gaussian noise or an escape noise to showcase the potential of our approach in producing both transient and asymptotic complex dynamics as well as finite-size fluctuations consistently across multiple scales. A comparison with the epidemiology of coronaviruses is also given to corroborate the qualitative relevance of our new approach.

keywords:
Refractory Density; Partial Differential Equations; Age-structured model; Time since last infection; Finite-size fluctuations
\affiliation

[inst1]organization=Centre Inria d’Université Côte d’Azur,addressline=2004 Rte des Lucioles, city=Biot, postcode=06410, country=France

\affiliation

[inst2]organization=Universite Claude Bernard Lyon 1, CNRS, Ecole Centrale de Lyon, INSA Lyon, Université Jean Monnet, ICJ, UMR5208, Inria,city=Villeurbanne, postcode=69622, country=France

\affiliation

[inst3]organization=Bernstein Center for Computational Neuroscience,addressline=Philippstr. 13, Haus 6, city=Berlin, postcode=10115, country=Germany \affiliation[inst4]organization=Technische Universität Berlin Institute of Mathematics,addressline=Straße des 17. Juni 136, city=Berlin, postcode=10623, country=Germany

\affiliation

[inst5]organization=Department of Mathematical Sciences “G. L. Lagrange”, Politecnico di Torino,addressline=Corso Duca degli Abruzzi 24, city=Torino, postcode=10129, country=Italy

1 Introduction

The susceptibility of an individual to an infectious disease as well as the infectiousness of an infected individual considerably depends on the time of the person’s last infection, or “infection age”. Such dependence on the infection course and history can be modeled by age-structured population models. These models have been used since the seminal paper by Kermack and McKendrick [17] for a wide variety of mathematical models in epidemics, see e.g. [2, 3, 6, 13, 18]. Taking into account the time since the last infection (TSLI) is extremely convenient to introduce a temporal heterogeneity in the population concerning an ongoing epidemic, and to explicitly model a complex disease evolution with consequences both within and between hosts.

Population models structured by this TSLI are deterministic macroscopic models that relate to the idealized limit of an infinitely large population. Two important questions arise here: first, a fundamental question is how such macroscopic models are linked to “microscopic models”, where the TSLI or other internal variables are tracked for each individual in the population. Microscopic models can be more directly related to the mechanisms underlying the spreading of disease and to parameters measured in clinical studies. However, these models are too complex to understand the collective dynamics of an epidemic. Second, populations in reality (say, a local community relevant to the disease or an age group) are not infinitely large and thus exhibit fluctuations that reflect the infections of single individuals. Therefore, this raises the question of whether there exists an intermediate “mesoscopic model” that combines the simplicity of age-structured population models with the ability to capture fluctuations due to finite population size. Again, the parameters of such a mesoscopic model should be linked to the microscopic parameters. Answering these questions requires a multi-scale modeling framework for epidemic diseases.

We found our inspiration in the neuroscience field, where the age-structured population dynamics models have already been used in the form of the refractory density method [7, 8, 11, 12, 14, 24, 27, 29, 28]. Indeed, in that case, the basic units at the microscopic scale are the neurons, whose probability to fire a spike is largely determined by the time since the last spike as well as the spike input from other neurons in a large network of neurons. Taking into account the time since the last spike is important because neurons exhibit a period of absolute and relative refractoriness after a spike, where the firing probability is strongly reduced. At the macroscopic scale, the evolution of a neuronal population is governed by a system of first-order partial differential equations called refractory-density equations (RDE). The RDE tracks the density of the times since the last spikes within a continuous phase state, thereby avoiding discrete states such as spiking, refractoriness, and resting. Importantly, the RDE is a bottom-up mean-field model that is directly derived from the single neuron equations at the microscopic scale. Simulations using this technique yield precise solutions for both equilibrium states and transient non-equilibrium regimes.

In the last two decades, there have been several advancements in the refractory-density method that offered the possibility for significant progress in the multi-scale modeling of cortical networks in the brain. First, the method has been adapted for the important case where spiking events are triggered by the threshold crossing of a neuronal membrane potential that is driven by biologically plausible Gaussian white [7] or colored noise [8, 26], while the classical RDE is based on a phenomenological hazard-rate-based spike-generation mechanism (“escape noise”) [14]. Second, the RDE has been generalized to finite-size populations [25, 28], resulting in a stochastic RDE that accurately reproduces the finite-size fluctuations at the mesoscopic scale. This advancement offers a unique and consistent description of a neural circuit at three levels of granularity: micro-, meso- and macroscopic scales. Additionally, there has been a further advancement in the generalization of RDE to accommodate complex, multi-dimensional [7], and even two-compartmental [9] neurons, where the neuronal state is given by multiple variables rather than a single membrane potential. These advancements collectively allow the application of RDE for constructing meso- and macroscopic models based on a wide range of microscopic models.

The main concept of this paper is to adopt the recent generalizations of the refractory-density method from neuroscience to epidemiology. In particular, we suggest a novel multi scale modeling framework that connects three different scales (micro, meso, macro). Moreover, the theoretical framework applies to two different noise mechanisms at the microscopic scale (Gaussian white noise and escape noise). We show their distinct effects on the form of the age-structure epidemic population model at the meso- and macroscopic scales. To this end, we consider an infectious disease with a relatively short infection period, followed by temporary immunity and the possibility of multiple subsequent infections (e.g., flu or SARS-CoV-2). Our model can be naturally structured within a classical SEIRS (Susceptible – Exposed – Infected – Recovered – Susceptible) epidemic framework, where we continuously track the interplay between viral load and the immune system within each individual.

Unlike the classical ordinary differential equation (ODE)-based compartmental models of epidemics (see, e.g., [19]), we choose a continuous phase space, focusing on the time since infection and the corresponding evolution of the interaction between the virus and the immune system. In this space, the SEIRS stages can be defined either strictly or loosely; see section 2 for a detailed explanation of our definition. In reality, an individual does not experience strict boundaries between SEIRS stages, and our approach allows us to avoid such additional assumptions. Instead, we propose that two characteristics—- the time since the previous infection and the interplay between viral load and immune response -— sufficiently describe an individual’s state within a population. We believe that avoiding strictly defined compartments and introducing a continuous space of states is a key advantage of our model.

Another advantage of our approach is that the spread of infectious disease at the mesoscopic and macroscopic population level is derived from the equations that stand for the dynamics of infection, disease development, and recovery of a single individual. A similar approach governs the derivation of macroscopic equations starting from their microscopic counterparts in kinetic theory, see e.g. [10, 20]. Our construction differs however significantly, as we explain in sections 2 and 3. In particular, the present paper is, to the best of our knowledge, novel in that it combines such a multi-scale (microscopic, mesoscopic, macroscopic) model with an age (in the sense of “infection age”) structured component, exploiting tools and analytical techniques coming from the so-called Refractory Density approach.

The paper is structured as follows: in Section 2, we present the microscopic model describing the evolution of the disease within a single individual. In Section 3 we introduce the equations governing the macro-scale population, while the mesoscopic ones are given in Section 4. We illustrate our work in Section 5 with numerical simulations and compare them to real data. We finally discuss our work in Section 6.

Refer to caption


Figure 1: Schematic of the micro- and macroscopic models described in sections 2 and 3. A, Microscopic model. The state variable Visubscript𝑉𝑖V_{i}italic_V start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT of an individual i𝑖iitalic_i evolves in time t𝑡titalic_t (bottom trace); its crossings of the threshold VTsuperscript𝑉𝑇V^{T}italic_V start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT determine the onsets of disease, the infection time moments, when the drive D(t)𝐷𝑡D(t)italic_D ( italic_t ) appear, which describes virus production and immune response (top). The time since the last infection tsuperscript𝑡t^{*}italic_t start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT is zeroed at the onset of the disease. B, Macroscopic model. At time t𝑡titalic_t, individuals with different tsuperscript𝑡t^{*}italic_t start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT are distributed with the density ρ𝜌\rhoitalic_ρ in the tsuperscript𝑡t^{*}italic_t start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT-space (bottom). With time, the density transports rightwards (as tsuperscript𝑡t^{*}italic_t start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT increases with t𝑡titalic_t) and undergoes a return flux to the infection state t=0superscript𝑡0t^{*}=0italic_t start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT = 0 according to the hazard function H𝐻Hitalic_H, so the dashed arrows represent subsequent infections. The hazard depends on the mean state variable U𝑈Uitalic_U, which is similar to Visubscript𝑉𝑖V_{i}italic_V start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT in A but is extrapolated over the threshold.

2 Microscopic model of single individuals

We consider a population of N1much-greater-than𝑁1N\gg 1italic_N ≫ 1 individuals, among which an infectious disease is spreading. We do not consider birth and death processes, and consequently, we assume the population to be constant (although this assumption is not critical for our framework). We assume that the risk of infection for an individual depends on two state variables Visubscript𝑉𝑖V_{i}italic_V start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT and tisubscriptsuperscript𝑡𝑖t^{*}_{i}italic_t start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT (see Table 1 for a summary of the variables and parameters of the model). The variable Visubscript𝑉𝑖V_{i}italic_V start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT represents the interplay of the viral load and the immune response within individual i𝑖iitalic_i. We call this variable viral state. The dynamics of Vi(t)subscript𝑉𝑖𝑡V_{i}(t)italic_V start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_t ) are governed by the interaction of multiple processes (Fig.1A, bottom). First, due to the immune system, the number of viral particles relaxes with time constant τ𝜏\tauitalic_τ. Second, it grows proportionally to the fraction of infected individual I(t)𝐼𝑡I(t)italic_I ( italic_t ), with a coefficient of proportionality k𝑘kitalic_k, which is assumed to be proportional to the basic reproduction number R0subscript𝑅0R_{0}italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT of the corresponding disease. This modeling assumption is qualitatively close to the underlying homogeneous mixing assumption of classical ODE models, whereby any individual is equally likely to meet with any other individual in the population. Third, it mimics the time course of the disease composed of a rapid increase of the viral load at disease onset and a subsequent decrease caused by the immune response. This time course is driven by the function D(ti)𝐷superscriptsubscript𝑡𝑖D(t_{i}^{*})italic_D ( italic_t start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ), started at each onset of the disease (Fig.1A, top). Note that, since Visubscript𝑉𝑖V_{i}italic_V start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT represents the interplay between the virus and the immune system, it can become negative at the negative phase of D(ti)𝐷superscriptsubscript𝑡𝑖D(t_{i}^{*})italic_D ( italic_t start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ).

Variable Notation Units
t𝑡titalic_t Time days
tsuperscript𝑡t^{*}italic_t start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT Time since last infection, or “infection age” days
Visubscript𝑉𝑖V_{i}italic_V start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT Interplay of viral load and immune response for individual i𝑖iitalic_i v-units
U𝑈Uitalic_U Mean interplay of viral load and immune response v-units
ρ𝜌\rhoitalic_ρ Density of individuals distributed in tsuperscript𝑡t^{*}italic_t start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT-space people/day
ν𝜈\nuitalic_ν Population rate of new infections people/day
I𝐼Iitalic_I Fraction of infectious individuals in the population nondim.
τ𝜏\tauitalic_τ Time scale of passive immune response days
τrsubscript𝜏𝑟\tau_{r}italic_τ start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT Time scale of disease progress and cure days
τIsubscript𝜏𝐼\tau_{I}italic_τ start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT Infectious period days
τRsubscript𝜏𝑅\tau_{R}italic_τ start_POSTSUBSCRIPT italic_R end_POSTSUBSCRIPT Time from infection to loss of immunity (“Refractory period”) days
D𝐷Ditalic_D Time course of the interplay for disease and treatment v-units/day
N𝑁Nitalic_N Total number of people in the population people
Ncsubscript𝑁𝑐N_{c}italic_N start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT Cumulative cases in the population people
R0subscript𝑅0R_{0}italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT Basic Reproduction Number nondim.
k𝑘kitalic_k Average individual rate of viral load v-units/day
H𝐻Hitalic_H Hazard; HΔt𝐻Δ𝑡H\Delta titalic_H roman_Δ italic_t is the probability of infection in the interval (t,t+Δt)𝑡𝑡Δ𝑡(t,t+\Delta t)( italic_t , italic_t + roman_Δ italic_t ) people/day
VTsuperscript𝑉𝑇V^{T}italic_V start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT Threshold of infection v-units
σ𝜎\sigmaitalic_σ Noise amplitude v-units
Table 1: Notation for the micro- and macroscopic models described in sections 2 and 3.

The state variable tisuperscriptsubscript𝑡𝑖t_{i}^{*}italic_t start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT measures the time relative to the last infection time (onset of the disease) of individual i𝑖iitalic_i. These onsets define an ordered sequence t1,i<t2,i<t3,i<subscript𝑡1𝑖subscript𝑡2𝑖subscript𝑡3𝑖t_{1,i}<t_{2,i}<t_{3,i}<\dotsbitalic_t start_POSTSUBSCRIPT 1 , italic_i end_POSTSUBSCRIPT < italic_t start_POSTSUBSCRIPT 2 , italic_i end_POSTSUBSCRIPT < italic_t start_POSTSUBSCRIPT 3 , italic_i end_POSTSUBSCRIPT < ⋯ of infection times for individual i𝑖iitalic_i. The continuous states of the epidemic in the tsuperscript𝑡t^{*}italic_t start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT-space may be roughly interpreted in terms of discrete SEIRS stages (Fig.1A). To this end, we may partition the population with respect to an ongoing epidemic in the following compartments: S𝑆Sitalic_S, individuals who are Susceptible to the disease; E𝐸Eitalic_E, individuals who have been Exposed to the disease but are not yet infectious; I𝐼Iitalic_I, Infected individuals who may spread the disease; and R𝑅Ritalic_R, Recovered individuals who are temporarily immune from infection. According to these definitions, we can loosely define the stages in the tsuperscript𝑡t^{*}italic_t start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT-axis. When Vi(t)subscript𝑉𝑖𝑡V_{i}(t)italic_V start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_t ) overcomes a threshold VTsuperscript𝑉𝑇V^{T}italic_V start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT (in the case of Gaussian noise, see below) or an infection event is triggered by a probabilistic “soft” threshold (in the case of escape noise), the individual becomes infected and consequently belongs to compartment I𝐼Iitalic_I. During the infectious phase, Vi(t)subscript𝑉𝑖𝑡V_{i}(t)italic_V start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_t ) rises above the threshold because of the viral production during disease and then decreases because of the immune response. The I𝐼Iitalic_I state lasts for a fixed duration τI>0subscript𝜏𝐼0\tau_{I}>0italic_τ start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT > 0. After the time τIsubscript𝜏𝐼\tau_{I}italic_τ start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT has elapsed since the onset of infection, the individual is effectively immune to the disease and consequently belongs to compartment R𝑅Ritalic_R. The sojourn in this compartment lasts until the end of the recovery period τR>τIsubscript𝜏𝑅subscript𝜏𝐼\tau_{R}>\tau_{I}italic_τ start_POSTSUBSCRIPT italic_R end_POSTSUBSCRIPT > italic_τ start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT, measured from the onset of infection. After this time, when Visubscript𝑉𝑖V_{i}italic_V start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT is low, the individual becomes Susceptible and then Exposed until the next infection when the individual becomes Infected and infectious again. Note that only the Infectious and Recovered states are strictly defined by the interval 0<t<τI0superscript𝑡subscript𝜏𝐼0<t^{*}<\tau_{I}0 < italic_t start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT < italic_τ start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT and τIt<τRsubscript𝜏𝐼superscript𝑡subscript𝜏𝑅\tau_{I}\leq t^{*}<\tau_{R}italic_τ start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT ≤ italic_t start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT < italic_τ start_POSTSUBSCRIPT italic_R end_POSTSUBSCRIPT, respectively. However, the duration of these states does not directly affect the evolution of the individual’s state Visubscript𝑉𝑖V_{i}italic_V start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT, but instead determines the periods when the individual affects the population and is in an “absolute refractory period”, respectively.

Taken together, the dynamics for single individuals i𝑖iitalic_i, i=1,,N𝑖1𝑁i=1,\dotsc,Nitalic_i = 1 , … , italic_N, between two subsequent infection times tk,isubscript𝑡𝑘𝑖t_{k,i}italic_t start_POSTSUBSCRIPT italic_k , italic_i end_POSTSUBSCRIPT and tk+1,isubscript𝑡𝑘1𝑖t_{k+1,i}italic_t start_POSTSUBSCRIPT italic_k + 1 , italic_i end_POSTSUBSCRIPT is modeled as

dVidtdsubscript𝑉𝑖d𝑡\displaystyle\frac{\text{d}V_{i}}{\text{d}t}divide start_ARG d italic_V start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG start_ARG d italic_t end_ARG =Viτ+k(t)I(t)+D(ti)+σ(I(t))ξi(t),absentsubscript𝑉𝑖𝜏𝑘𝑡𝐼𝑡𝐷superscriptsubscript𝑡𝑖𝜎𝐼𝑡subscript𝜉𝑖𝑡\displaystyle=-\frac{V_{i}}{\tau}+k(t)I(t)+D(t_{i}^{*})+\sigma(I(t))% \leavevmode\nobreak\ \xi_{i}(t),= - divide start_ARG italic_V start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG start_ARG italic_τ end_ARG + italic_k ( italic_t ) italic_I ( italic_t ) + italic_D ( italic_t start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ) + italic_σ ( italic_I ( italic_t ) ) italic_ξ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_t ) , (1)
dtidt𝑑superscriptsubscript𝑡𝑖𝑑𝑡\displaystyle\frac{dt_{i}^{*}}{dt}divide start_ARG italic_d italic_t start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT end_ARG start_ARG italic_d italic_t end_ARG =1,absent1\displaystyle=1,= 1 , (2)
with the fraction of infected individuals
I(t)𝐼𝑡\displaystyle I(t)italic_I ( italic_t ) =Nc(t)Nc(tτI)N,absentsubscript𝑁𝑐𝑡subscript𝑁𝑐𝑡subscript𝜏𝐼𝑁\displaystyle=\frac{N_{c}(t)-N_{c}(t-\tau_{I})}{N},= divide start_ARG italic_N start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ( italic_t ) - italic_N start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ( italic_t - italic_τ start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT ) end_ARG start_ARG italic_N end_ARG , (3)

where Nc(t)subscript𝑁𝑐𝑡N_{c}(t)\in\mathbb{N}italic_N start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ( italic_t ) ∈ blackboard_N is a global counting variable that keeps track of the cumulative number of cases in the population up to time t𝑡titalic_t. Clearly, this counting variable is constant when no new infection occurs. At the infection times tk,isubscript𝑡𝑘𝑖t_{k,i}italic_t start_POSTSUBSCRIPT italic_k , italic_i end_POSTSUBSCRIPT, however, the overall dynamics is supplemented with an additional jump condition (reset):

Vi(tk,i+)=VT,ti(tk,i+)=0,Nc(tk,i+)=Nc(tk,i)+1,formulae-sequencesubscript𝑉𝑖superscriptsubscript𝑡𝑘𝑖superscript𝑉𝑇formulae-sequencesuperscriptsubscript𝑡𝑖superscriptsubscript𝑡𝑘𝑖0subscript𝑁𝑐superscriptsubscript𝑡𝑘𝑖subscript𝑁𝑐superscriptsubscript𝑡𝑘𝑖1V_{i}(t_{k,i}^{+})=V^{T},\quad t_{i}^{*}(t_{k,i}^{+})=0,\quad N_{c}(t_{k,i}^{+% })=N_{c}(t_{k,i}^{-})+1,italic_V start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_t start_POSTSUBSCRIPT italic_k , italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT ) = italic_V start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT , italic_t start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ( italic_t start_POSTSUBSCRIPT italic_k , italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT ) = 0 , italic_N start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ( italic_t start_POSTSUBSCRIPT italic_k , italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT ) = italic_N start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ( italic_t start_POSTSUBSCRIPT italic_k , italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT ) + 1 , (4)

where tk,isuperscriptsubscript𝑡𝑘𝑖t_{k,i}^{-}italic_t start_POSTSUBSCRIPT italic_k , italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT and tk,i+superscriptsubscript𝑡𝑘𝑖t_{k,i}^{+}italic_t start_POSTSUBSCRIPT italic_k , italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT denote the left- and right-sided limits, respectively. In Eq. (3), we indicate with τIsubscript𝜏𝐼\tau_{I}italic_τ start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT the average time spent in the infectious state or average duration of the infectious window. We note that Eq. (3) can be easily generalized to permit a more fine-grained, graded account of infectiousness depending on infection age tsuperscript𝑡t^{*}italic_t start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT (see Discussion, Eq. (16)). Furthermore,

D(t)𝐷superscript𝑡\displaystyle D(t^{*})italic_D ( italic_t start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ) =4aexp(tτr)aexp(t4τr)absent4𝑎superscript𝑡subscript𝜏𝑟𝑎superscript𝑡4subscript𝜏𝑟\displaystyle=4a\exp\left(-\frac{t^{*}}{\tau_{r}}\right)-a\exp\left(-\frac{t^{% *}}{4\tau_{r}}\right)= 4 italic_a roman_exp ( - divide start_ARG italic_t start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT end_ARG start_ARG italic_τ start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT end_ARG ) - italic_a roman_exp ( - divide start_ARG italic_t start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT end_ARG start_ARG 4 italic_τ start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT end_ARG ) (5)

is a response function that drives the time course of Visubscript𝑉𝑖V_{i}italic_V start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT during disease. The parameter a𝑎aitalic_a is the severity of the disease, and τrsubscript𝜏𝑟\tau_{r}italic_τ start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT is the time of disease progress and cure. The last term in Eq. (4) represents an independent Gaussian white noise with mean 𝔼[ξ(t)]=0𝔼delimited-[]𝜉𝑡0\mathbb{E}[\xi(t)]=0blackboard_E [ italic_ξ ( italic_t ) ] = 0, auto-covariance function 𝔼[ξ(t)ξ(t)]=δ(tt)𝔼delimited-[]𝜉𝑡𝜉superscript𝑡𝛿𝑡superscript𝑡\mathbb{E}[\xi(t)\xi(t^{\prime})]=\delta(t-t^{\prime})blackboard_E [ italic_ξ ( italic_t ) italic_ξ ( italic_t start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) ] = italic_δ ( italic_t - italic_t start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) and strength σ(I)𝜎𝐼\sigma(I)italic_σ ( italic_I ) (δ()𝛿\delta(\cdot)italic_δ ( ⋅ ) denotes the Dirac delta function). As explained below, this term is only present in our model variant with Gaussian noise and is absent in the variant with escape noise where we set σ(I)0𝜎𝐼0\sigma(I)\equiv 0italic_σ ( italic_I ) ≡ 0.

As initial conditions of the model, we assume that before time t=0𝑡0t=0italic_t = 0 there were no cases, Nc(t)=0subscript𝑁𝑐𝑡0N_{c}(t)=0italic_N start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ( italic_t ) = 0 for t<0𝑡0t<0italic_t < 0 and that at time t=0𝑡0t=0italic_t = 0 a fraction I0subscript𝐼0I_{0}italic_I start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT of the population gets infected, i.e. Nc(0)=I0Nsubscript𝑁𝑐0subscript𝐼0𝑁N_{c}(0)=I_{0}Nitalic_N start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ( 0 ) = italic_I start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_N. Thus the state variables for the I0Nsubscript𝐼0𝑁I_{0}Nitalic_I start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_N initially infected individuals are Vi(0)=VTsubscript𝑉𝑖0superscript𝑉𝑇V_{i}(0)=V^{T}italic_V start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( 0 ) = italic_V start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT and ti(0)=0superscriptsubscript𝑡𝑖00t_{i}^{*}(0)=0italic_t start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ( 0 ) = 0, while the state variables of all other (non-infected) individuals (the remaining (1I0)N1subscript𝐼0𝑁(1-I_{0})N( 1 - italic_I start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) italic_N individuals in the population) is initially Vi(0)=0subscript𝑉𝑖00V_{i}(0)=0italic_V start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( 0 ) = 0 and ti(0)=+superscriptsubscript𝑡𝑖0t_{i}^{*}(0)=+\inftyitalic_t start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ( 0 ) = + ∞.

Importantly, it remains to define how the infection events are triggered. We consider two different variants to define these events, and hence the infection times tk,isubscript𝑡𝑘𝑖t_{k,i}italic_t start_POSTSUBSCRIPT italic_k , italic_i end_POSTSUBSCRIPT, corresponding to two different ways of including stochasticity in the model. We refer to these variants as the white- and escape-noise cases, respectively.

Model with escape noise

To introduce stochasticity that phenomenologically captures the influences of various sources of noise, we model the infection events of an individual i𝑖iitalic_i by a probabilistic risk of becoming infected depending on the individual’s current state variables Vi(t)subscript𝑉𝑖𝑡V_{i}(t)italic_V start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_t ) and ti(t)subscriptsuperscript𝑡𝑖𝑡t^{*}_{i}(t)italic_t start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_t ). Mathematically, we keep the dynamics of Visubscript𝑉𝑖V_{i}italic_V start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT between infection events deterministic (i.e. setting σ=0𝜎0\sigma=0italic_σ = 0) and generate infection events stochastically by a state-dependent hazard rate

Hi(t)=H(Vi(t),ti(t)).subscript𝐻𝑖𝑡𝐻subscript𝑉𝑖𝑡superscriptsubscript𝑡𝑖𝑡H_{i}(t)=H(V_{i}(t),t_{i}^{*}(t)).italic_H start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_t ) = italic_H ( italic_V start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_t ) , italic_t start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ( italic_t ) ) . (6)

The hazard rate means that an infection of individual i𝑖iitalic_i occurs in a small time step of length ΔtΔ𝑡\Delta troman_Δ italic_t with conditional probability Hi(t)Δtsubscript𝐻𝑖𝑡Δ𝑡H_{i}(t)\Delta titalic_H start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_t ) roman_Δ italic_t given the individual’s current state (Vi(t),ti(t))subscript𝑉𝑖𝑡subscriptsuperscript𝑡𝑖𝑡(V_{i}(t),t^{*}_{i}(t))( italic_V start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_t ) , italic_t start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_t ) ). More precisely, we have the following conditional probability:

Pr{individual i gets infected in (t,t+Δt)|Vi(t),ti(t)}=Hi(t)Δt+o(Δt)Prconditional-setindividual i gets infected in 𝑡𝑡Δ𝑡subscript𝑉𝑖𝑡superscriptsubscript𝑡𝑖superscript𝑡subscript𝐻𝑖𝑡Δ𝑡𝑜Δ𝑡\text{Pr}\left\{\text{individual $i$ gets infected in }(t,t+\Delta t)\;|\;V_{i% }(t),t_{i}^{*}(t^{-})\right\}=H_{i}(t)\Delta t+o(\Delta t)Pr { individual italic_i gets infected in ( italic_t , italic_t + roman_Δ italic_t ) | italic_V start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_t ) , italic_t start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ( italic_t start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT ) } = italic_H start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_t ) roman_Δ italic_t + italic_o ( roman_Δ italic_t ) (7)

as Δt0Δ𝑡0\Delta t\to 0roman_Δ italic_t → 0. For a concrete choice of the hazard rate in simulations, we use a rectified power function with absolute refractory period τRsubscript𝜏𝑅\tau_{R}italic_τ start_POSTSUBSCRIPT italic_R end_POSTSUBSCRIPT:

H(V,t)=cmax(0,V/VT)m𝟙tτR,H(V,t^{*})=c\max(0,V/V^{T})^{m}\mathbbm{1}_{t^{*}\geq\tau_{R}},italic_H ( italic_V , italic_t start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ) = italic_c roman_max ( 0 , italic_V / italic_V start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT ) start_POSTSUPERSCRIPT italic_m end_POSTSUPERSCRIPT blackboard_1 start_POSTSUBSCRIPT italic_t start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ≥ italic_τ start_POSTSUBSCRIPT italic_R end_POSTSUBSCRIPT end_POSTSUBSCRIPT , (8)

where m>0𝑚0m>0italic_m > 0. We illustrate the effect of varying m𝑚mitalic_m in Figure 2.

Model with white Gaussian noise

An alternative way to model noise is to consider an additive white-noise drive in the dynamics of Visubscript𝑉𝑖V_{i}italic_V start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT. This “diffusive noise” is modeled by the term σ(I(t))ξi(t)𝜎𝐼𝑡subscript𝜉𝑖𝑡\sigma(I(t))\leavevmode\nobreak\ \xi_{i}(t)italic_σ ( italic_I ( italic_t ) ) italic_ξ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_t ) in Eq. (1). The stochastic term may reflect both the fluctuations of the number of viral particles of a given type around a particular individual and the fluctuations of activity of the immune system of this individual. To prevent that in the absence of a viral infection in the population an infection occurs spontaneously by noise, we enforce the noise strength to be zero when I(t)𝐼𝑡I(t)italic_I ( italic_t ) is zero by choosing the specific function σ(I)=σ^𝟙I>0𝜎𝐼^𝜎subscript1𝐼0\sigma(I)=\hat{\sigma}\mathbbm{1}_{I>0}italic_σ ( italic_I ) = over^ start_ARG italic_σ end_ARG blackboard_1 start_POSTSUBSCRIPT italic_I > 0 end_POSTSUBSCRIPT. In the white-noise case, infection events are triggered when the viral state Vi(t)subscript𝑉𝑖𝑡V_{i}(t)italic_V start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_t ) crosses a certain threshold VTsuperscript𝑉𝑇V^{T}italic_V start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT, provided that at least an absolute refractory time τRsubscript𝜏𝑅\tau_{R}italic_τ start_POSTSUBSCRIPT italic_R end_POSTSUBSCRIPT has passed since the last infection. Hence, the infection times tk,isubscript𝑡𝑘𝑖t_{k,i}italic_t start_POSTSUBSCRIPT italic_k , italic_i end_POSTSUBSCRIPT are defined by the condition

V(tk,i)VT,t(tk,i)τR.formulae-sequence𝑉superscriptsubscript𝑡𝑘𝑖superscript𝑉𝑇superscript𝑡superscriptsubscript𝑡𝑘𝑖subscript𝜏𝑅V(t_{k,i}^{-})\geq V^{T},\quad t^{*}(t_{k,i}^{-})\geq\tau_{R}.italic_V ( italic_t start_POSTSUBSCRIPT italic_k , italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT ) ≥ italic_V start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT , italic_t start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ( italic_t start_POSTSUBSCRIPT italic_k , italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT ) ≥ italic_τ start_POSTSUBSCRIPT italic_R end_POSTSUBSCRIPT . (9)

Note that if this condition is fulfilled at some time tk,isuperscriptsubscript𝑡𝑘𝑖t_{k,i}^{-}italic_t start_POSTSUBSCRIPT italic_k , italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT, the reset condition Eq. (4) ensures that the above threshold condition no longer holds at tk,i+superscriptsubscript𝑡𝑘𝑖t_{k,i}^{+}italic_t start_POSTSUBSCRIPT italic_k , italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT, i.e. immediately after this time, as it should be.

The above equations conclude the definition of the microscopic model. Given the following application of the refractory-density (RD) method, we note that the case of white Gaussian noise can be approximately mapped to the case of escape noise [7, 26]. In this case, the noise term in Eq. (1) is omitted and the condition Eq. (9) is substituted by (7), where Hi(t)subscript𝐻𝑖𝑡H_{i}(t)italic_H start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_t ) is now given by the following hazard function (10). The hazard rate H(t,t)𝐻𝑡superscript𝑡H(t,t^{*})italic_H ( italic_t , italic_t start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ) depends on the actual state Visubscript𝑉𝑖V_{i}italic_V start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT, its rate of change dVi/dtdsubscript𝑉𝑖d𝑡\text{d}V_{i}/\text{d}td italic_V start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT / d italic_t, the threshold VTsuperscript𝑉𝑇V^{T}italic_V start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT, and the noise amplitude σ𝜎\sigmaitalic_σ, i.e., H(t,t)=H(Vi(t),V˙i(t),ti(t);σ(I(t)))𝐻𝑡superscript𝑡𝐻subscript𝑉𝑖𝑡subscript˙𝑉𝑖𝑡subscriptsuperscript𝑡𝑖𝑡𝜎𝐼𝑡H(t,t^{*})=H(V_{i}(t),\dot{V}_{i}(t),t^{*}_{i}(t);\sigma(I(t)))italic_H ( italic_t , italic_t start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ) = italic_H ( italic_V start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_t ) , over˙ start_ARG italic_V end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_t ) , italic_t start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_t ) ; italic_σ ( italic_I ( italic_t ) ) ). In [7], this function was found as an approximate solution of a first-time passage problem based on the Kolmogorov-Fokker-Planck equation:

H(U,U˙,t,σ)𝐻𝑈˙𝑈superscript𝑡𝜎\displaystyle H(U,\dot{U},t^{*},\sigma)italic_H ( italic_U , over˙ start_ARG italic_U end_ARG , italic_t start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT , italic_σ ) ={A(T)+B(T,U˙,σ),ift>τR,σ>0; 0,otherwise},\displaystyle=\{A(T)+B(T,\dot{U},\sigma),\leavevmode\nobreak\ \leavevmode% \nobreak\ \hbox{if}\leavevmode\nobreak\ \leavevmode\nobreak\ t^{*}>\tau_{R},% \leavevmode\nobreak\ \leavevmode\nobreak\ \sigma>0;\leavevmode\nobreak\ % \leavevmode\nobreak\ \leavevmode\nobreak\ 0,\leavevmode\nobreak\ \hbox{% otherwise}\},= { italic_A ( italic_T ) + italic_B ( italic_T , over˙ start_ARG italic_U end_ARG , italic_σ ) , if italic_t start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT > italic_τ start_POSTSUBSCRIPT italic_R end_POSTSUBSCRIPT , italic_σ > 0 ; 0 , otherwise } , (10)
A(T)𝐴𝑇\displaystyle A(T)italic_A ( italic_T ) =1τexp(0.00611.12T0.257T20.072T30.0117T4),absent1𝜏0.00611.12𝑇0.257superscript𝑇20.072superscript𝑇30.0117superscript𝑇4\displaystyle=\frac{1}{\tau}\leavevmode\nobreak\ \exp(0.0061-1.12T-0.257T^{2}-% 0.072T^{3}-0.0117T^{4}),= divide start_ARG 1 end_ARG start_ARG italic_τ end_ARG roman_exp ( 0.0061 - 1.12 italic_T - 0.257 italic_T start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - 0.072 italic_T start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT - 0.0117 italic_T start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT ) ,
B(T,U˙,σ)𝐵𝑇˙𝑈𝜎\displaystyle B(T,\dot{U},\sigma)italic_B ( italic_T , over˙ start_ARG italic_U end_ARG , italic_σ ) =2σπ[U˙]+exp(T2)1+erf(T),T=VTUσ.formulae-sequenceabsent2𝜎𝜋subscriptdelimited-[]˙𝑈superscript𝑇21erf𝑇𝑇superscript𝑉𝑇𝑈𝜎\displaystyle=\frac{2}{\sigma\sqrt{\pi}}\bigg{[}\dot{U}\bigg{]}_{+}\frac{\exp(% -T^{2})}{1+\hbox{erf}(T)},\leavevmode\nobreak\ \leavevmode\nobreak\ % \leavevmode\nobreak\ \leavevmode\nobreak\ T=\frac{V^{T}-U}{\sigma}.= divide start_ARG 2 end_ARG start_ARG italic_σ square-root start_ARG italic_π end_ARG end_ARG [ over˙ start_ARG italic_U end_ARG ] start_POSTSUBSCRIPT + end_POSTSUBSCRIPT divide start_ARG roman_exp ( - italic_T start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) end_ARG start_ARG 1 + erf ( italic_T ) end_ARG , italic_T = divide start_ARG italic_V start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT - italic_U end_ARG start_ARG italic_σ end_ARG .

Here and in the following, [x]+=max(0,x)subscriptdelimited-[]𝑥0𝑥[x]_{+}=\max(0,x)[ italic_x ] start_POSTSUBSCRIPT + end_POSTSUBSCRIPT = roman_max ( 0 , italic_x ) denotes the rectified linear function. We refer to [26] for a slightly more elaborate approximation of diffusive noise by escape noise.

3 Macroscopic model of a population

For a large number of individuals N𝑁Nitalic_N, the dynamics of the microscopic model can be either evaluated with a Monte-Carlo simulation or obtained much more efficiently by integrating corresponding macroscopic population equations. To this end, we apply the refractory-density (RD) approach [27] to the epidemiological model described at an individual level in section 2. According to this approach, the state variables of a single “particle”/individual are parameterized by the “age” tsuperscript𝑡t^{*}italic_t start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT (with the total derivative in time substituted by the sum of partial derivatives, i.e. d/dt=/t+/tdd𝑡𝑡superscript𝑡\text{d}/\text{d}t=\partial/\partial t+\partial/\partial t^{*}d / d italic_t = ∂ / ∂ italic_t + ∂ / ∂ italic_t start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT), and the population is characterized by the density of particles in the one-dimensional, semi-infinite tsuperscript𝑡t^{*}italic_t start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT-space, i.e. ρ(t,t)𝜌𝑡superscript𝑡\rho(t,t^{*})italic_ρ ( italic_t , italic_t start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ) (Fig.1B). In our case, in line with other age-dependent models in mathematical epidemiology, the “age” is the time elapsed since infection. This means that at any time t𝑡titalic_t, the number of individuals in the population who became infected between t1subscriptsuperscript𝑡1t^{*}_{1}italic_t start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT and t2subscriptsuperscript𝑡2t^{*}_{2}italic_t start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT time units ago are represented by

t1t2ρ(t,t)dt.superscriptsubscriptsubscriptsuperscript𝑡1subscriptsuperscript𝑡2𝜌𝑡superscript𝑡dsuperscript𝑡\int_{t^{*}_{1}}^{t^{*}_{2}}\rho(t,t^{*})\text{d}t^{*}.∫ start_POSTSUBSCRIPT italic_t start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_t start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUPERSCRIPT italic_ρ ( italic_t , italic_t start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ) d italic_t start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT .

At the macroscopic level, we define the rate of new infections

ν(t)=limΔt0limNNc(t+Δt)Nc(t)NΔt.𝜈𝑡subscriptΔ𝑡0subscript𝑁subscript𝑁𝑐𝑡Δ𝑡subscript𝑁𝑐𝑡𝑁Δ𝑡\nu(t)=\lim_{\Delta t\to 0}\lim_{N\to{}\infty}\frac{N_{c}(t+\Delta t)-N_{c}(t)% }{N\Delta t}.italic_ν ( italic_t ) = roman_lim start_POSTSUBSCRIPT roman_Δ italic_t → 0 end_POSTSUBSCRIPT roman_lim start_POSTSUBSCRIPT italic_N → ∞ end_POSTSUBSCRIPT divide start_ARG italic_N start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ( italic_t + roman_Δ italic_t ) - italic_N start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ( italic_t ) end_ARG start_ARG italic_N roman_Δ italic_t end_ARG . (11)

We also introduce a function U(t,t)𝑈𝑡superscript𝑡U(t,t^{*})italic_U ( italic_t , italic_t start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ) as follows: for the model with escape noise, U(t,t)𝑈𝑡superscript𝑡U(t,t^{*})italic_U ( italic_t , italic_t start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ) is the unique function that maps the infection age ti(t)subscriptsuperscript𝑡𝑖𝑡t^{*}_{i}(t)italic_t start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_t ) to the state variable Vi(t)subscript𝑉𝑖𝑡V_{i}(t)italic_V start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_t ), i.e. Vi(t)=U(t,ti(t))subscript𝑉𝑖𝑡𝑈𝑡subscriptsuperscript𝑡𝑖𝑡V_{i}(t)=U(t,t^{*}_{i}(t))italic_V start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_t ) = italic_U ( italic_t , italic_t start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_t ) ). For the model with Gaussian white noise, there is no such deterministic map. However, we can first map the model to the approximate escape-noise model, Eq. (10), for which we can define again the unique function U(t,t)𝑈𝑡superscript𝑡U(t,t^{*})italic_U ( italic_t , italic_t start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ). Intuitively, this function can be interpreted as the variable Vi(t)subscript𝑉𝑖𝑡V_{i}(t)italic_V start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_t ) averaged across realizations of the Gaussian-white noise for individuals i𝑖iitalic_i that had the time of last infection equal to tsuperscript𝑡t^{*}italic_t start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT, i.e., the same tsuperscript𝑡t^{*}italic_t start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT. The equation for U𝑈Uitalic_U then follows from the corresponding averaging of Eq. (1). The bottom traces in Fig.1B illustrate distributions of U𝑈Uitalic_U and ρ𝜌\rhoitalic_ρ in tsuperscript𝑡t^{*}italic_t start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT-space.

The RD model corresponding to Eqs. (1)-(4) consists of two transport equations for ρ(t,t)𝜌𝑡superscript𝑡\rho(t,t^{*})italic_ρ ( italic_t , italic_t start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ) and U(t,t)𝑈𝑡superscript𝑡U(t,t^{*})italic_U ( italic_t , italic_t start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ), and two integrals for ν(t)𝜈𝑡\nu(t)italic_ν ( italic_t ) and I(t)𝐼𝑡I(t)italic_I ( italic_t ):

ρt+ρt𝜌𝑡𝜌superscript𝑡\displaystyle\frac{\partial\rho}{\partial t}+\frac{\partial\rho}{\partial t^{*}}divide start_ARG ∂ italic_ρ end_ARG start_ARG ∂ italic_t end_ARG + divide start_ARG ∂ italic_ρ end_ARG start_ARG ∂ italic_t start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT end_ARG =ρH(U(t,t),t),absent𝜌𝐻𝑈𝑡superscript𝑡superscript𝑡\displaystyle=-\rho\leavevmode\nobreak\ H(U(t,t^{*}),t^{*}),= - italic_ρ italic_H ( italic_U ( italic_t , italic_t start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ) , italic_t start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ) , (12a)
Ut+Ut𝑈𝑡𝑈superscript𝑡\displaystyle\frac{\partial U}{\partial t}+\frac{\partial U}{\partial t^{*}}divide start_ARG ∂ italic_U end_ARG start_ARG ∂ italic_t end_ARG + divide start_ARG ∂ italic_U end_ARG start_ARG ∂ italic_t start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT end_ARG =Uτ+k(t)I(t)+D(t),absent𝑈𝜏𝑘𝑡𝐼𝑡𝐷superscript𝑡\displaystyle=-\frac{U}{\tau}+k(t)I(t)+D(t^{*}),= - divide start_ARG italic_U end_ARG start_ARG italic_τ end_ARG + italic_k ( italic_t ) italic_I ( italic_t ) + italic_D ( italic_t start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ) , (12b)
ν(t)𝜈𝑡\displaystyle\nu(t)italic_ν ( italic_t ) =0ρ(t,t)H(U(t,t),t)𝑑t+I0δ(t),absentsuperscriptsubscript0𝜌𝑡superscript𝑡𝐻𝑈𝑡superscript𝑡superscript𝑡differential-dsuperscript𝑡subscript𝐼0𝛿𝑡\displaystyle=\int_{0}^{\infty}\rho(t,t^{*})\leavevmode\nobreak\ H(U(t,t^{*}),% t^{*})\,dt^{*}+I_{0}\delta(t),= ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_ρ ( italic_t , italic_t start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ) italic_H ( italic_U ( italic_t , italic_t start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ) , italic_t start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ) italic_d italic_t start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT + italic_I start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_δ ( italic_t ) , rate of new cases (12c)
I(t)𝐼𝑡\displaystyle I(t)italic_I ( italic_t ) =tτItν(s)𝑑s,absentsuperscriptsubscript𝑡subscript𝜏𝐼𝑡𝜈𝑠differential-d𝑠\displaystyle=\int_{t-\tau_{I}}^{t}\nu(s)\,ds,= ∫ start_POSTSUBSCRIPT italic_t - italic_τ start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_t end_POSTSUPERSCRIPT italic_ν ( italic_s ) italic_d italic_s , fraction of infected people (12d)

The boundary condition for ρ𝜌\rhoitalic_ρ, resulting from the conservation of people in the population, is

ρ(t,0)=ν(t).𝜌𝑡0𝜈𝑡\rho(t,0)=\nu(t).italic_ρ ( italic_t , 0 ) = italic_ν ( italic_t ) . (13)

The boundary condition for U𝑈Uitalic_U corresponding to the reset in the microscopic model, Eq. (4), is U(t,0)=VT𝑈𝑡0subscript𝑉𝑇U(t,0)=V_{T}italic_U ( italic_t , 0 ) = italic_V start_POSTSUBSCRIPT italic_T end_POSTSUBSCRIPT. The initial condition for the density ρ𝜌\rhoitalic_ρ corresponding to an initial fraction of infected people I0subscript𝐼0I_{0}italic_I start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT at t=0𝑡0t=0italic_t = 0 and to a situation, where individuals never encountered this viral infection before time t=0𝑡0t=0italic_t = 0 (formally, ti(0)=subscriptsuperscript𝑡𝑖superscript0t^{*}_{i}(0^{-})=\inftyitalic_t start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( 0 start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT ) = ∞) is

ρ(0,t)=δ(t)I0+δ(t)(1I0).𝜌0superscript𝑡𝛿superscript𝑡subscript𝐼0𝛿superscript𝑡1subscript𝐼0\rho(0,t^{*})=\delta(t^{*})I_{0}+\delta(t^{*}-\infty)(1-I_{0}).italic_ρ ( 0 , italic_t start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ) = italic_δ ( italic_t start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ) italic_I start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT + italic_δ ( italic_t start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT - ∞ ) ( 1 - italic_I start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) .

Furthermore, the corresponding initial condition for U𝑈Uitalic_U can be taken as the steady-state solution of Eq.(12b) (with U/t=0𝑈𝑡0\partial U/\partial t=0∂ italic_U / ∂ italic_t = 0 and I(t)=0𝐼𝑡0I(t)=0italic_I ( italic_t ) = 0 for t<0𝑡0t<0italic_t < 0):

U(0,t)=VTexp(t/τ)+0tes/τD(ts)𝑑s.𝑈0superscript𝑡superscript𝑉𝑇superscript𝑡𝜏superscriptsubscript0superscript𝑡superscript𝑒𝑠𝜏𝐷superscript𝑡𝑠differential-d𝑠U(0,t^{*})=V^{T}\exp{(-t^{*}/\tau)}+\int_{0}^{t^{*}}e^{-s/\tau}D(t^{*}-s)\,ds.italic_U ( 0 , italic_t start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ) = italic_V start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT roman_exp ( - italic_t start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT / italic_τ ) + ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_t start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT end_POSTSUPERSCRIPT italic_e start_POSTSUPERSCRIPT - italic_s / italic_τ end_POSTSUPERSCRIPT italic_D ( italic_t start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT - italic_s ) italic_d italic_s .

In the density equation (12a), the density diminishes because of the sink term ρH𝜌𝐻-\rho H- italic_ρ italic_H but the total mass 0ρ(t,t)𝑑t=1superscriptsubscript0𝜌𝑡superscript𝑡differential-dsuperscript𝑡1\int_{0}^{\infty}\rho(t,t^{*})\,dt^{*}=1∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_ρ ( italic_t , italic_t start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ) italic_d italic_t start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT = 1 is conserved at all times because of the boundary condition, Eq. (13). This boundary condition acts as a source term at t=0superscript𝑡0t^{*}=0italic_t start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT = 0 given by the rate of new cases ν(t)𝜈𝑡\nu(t)italic_ν ( italic_t ) that exactly balances the integrated sink term, Eq. (12c). We note that this conservation law reflects the fact that we neglect mortality, and focus on the transmission mechanism.

Furthermore, in the density equation (12a), the risk of illness is evaluated by the hazard function H𝐻Hitalic_H. In Eq. (12), we used the hazard function for the case of escape noise. For the case of white noise, the hazard function H(U(t,t),t)𝐻𝑈𝑡superscript𝑡superscript𝑡H(U(t,t^{*}),t^{*})italic_H ( italic_U ( italic_t , italic_t start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ) , italic_t start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ) should be replaced by H(U(t,t),U˙(t,t),t,σ(I(t)))𝐻𝑈𝑡superscript𝑡˙𝑈𝑡superscript𝑡superscript𝑡𝜎𝐼𝑡H(U(t,t^{*}),\dot{U}(t,t^{*}),t^{*},\sigma(I(t)))italic_H ( italic_U ( italic_t , italic_t start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ) , over˙ start_ARG italic_U end_ARG ( italic_t , italic_t start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ) , italic_t start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT , italic_σ ( italic_I ( italic_t ) ) ), where H𝐻Hitalic_H is given by Eq. (10) and U˙(t,t)=(t+t)U(t,t)˙𝑈𝑡superscript𝑡subscript𝑡subscriptsuperscript𝑡𝑈𝑡superscript𝑡\dot{U}(t,t^{*})=(\partial_{t}+\partial_{t^{*}})U(t,t^{*})over˙ start_ARG italic_U end_ARG ( italic_t , italic_t start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ) = ( ∂ start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT + ∂ start_POSTSUBSCRIPT italic_t start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT ) italic_U ( italic_t , italic_t start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ) is given by the right-hand-side of Eq. (12b).

Once again, note that Eq. (12d) permits a simple generalization to graded infectiousness depending on the course of the infection (see Discussion, Eq. (16)).

4 Mesoscopic model of a finite-size population

At the intermediate mesoscopic scale, the size of (sub-)populations is finite. The finite size causes fluctuations in the fraction of infectious people I(t)𝐼𝑡I(t)italic_I ( italic_t ), which may yield significant finite-size effects through the interaction of finite-size noise and nonlinear population dynamics. In the case of a finite size population consisting of N𝑁N\in\mathbb{N}italic_N ∈ blackboard_N individuals, N1much-greater-than𝑁1N\gg 1italic_N ≫ 1, we apply the corresponding theory from [25, 28], which yields a stochastic generalization of the macroscopic dynamics given in the previous section. To this end, we introduce the pseudo-density ρ(t,t)𝜌𝑡superscript𝑡\rho(t,t^{*})italic_ρ ( italic_t , italic_t start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ) in terms of the survivor function S(t,t)𝑆𝑡superscript𝑡S(t,t^{*})italic_S ( italic_t , italic_t start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ) as

ρ(t,t)=S(t,t)νN(tt).𝜌𝑡superscript𝑡𝑆𝑡superscript𝑡subscript𝜈𝑁𝑡superscript𝑡\rho(t,t^{*})=S(t,t^{*})\leavevmode\nobreak\ \nu_{N}(t-t^{*}).italic_ρ ( italic_t , italic_t start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ) = italic_S ( italic_t , italic_t start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ) italic_ν start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT ( italic_t - italic_t start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ) . (14a)
Here, S𝑆Sitalic_S and νNsubscript𝜈𝑁\nu_{N}italic_ν start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT are given as the solution of the following system of stochastic partial differential equations, generalizing Eq. (12)
St+St𝑆𝑡𝑆superscript𝑡\displaystyle\frac{\partial S}{\partial t}+\frac{\partial S}{\partial t^{*}}divide start_ARG ∂ italic_S end_ARG start_ARG ∂ italic_t end_ARG + divide start_ARG ∂ italic_S end_ARG start_ARG ∂ italic_t start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT end_ARG =SH(U),S(t,0)=1,formulae-sequenceabsent𝑆𝐻𝑈𝑆𝑡01\displaystyle=-S\leavevmode\nobreak\ H(U),\qquad S(t,0)=1,= - italic_S italic_H ( italic_U ) , italic_S ( italic_t , 0 ) = 1 , (14b)
Ut+Ut𝑈𝑡𝑈superscript𝑡\displaystyle\frac{\partial U}{\partial t}+\frac{\partial U}{\partial t^{*}}divide start_ARG ∂ italic_U end_ARG start_ARG ∂ italic_t end_ARG + divide start_ARG ∂ italic_U end_ARG start_ARG ∂ italic_t start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT end_ARG =Uτ+u(t)+kI(t)+D(t),U(t,0)=VT,formulae-sequenceabsent𝑈𝜏𝑢𝑡𝑘𝐼𝑡𝐷superscript𝑡𝑈𝑡0subscript𝑉𝑇\displaystyle=-\frac{U}{\tau}+u(t)+k\leavevmode\nobreak\ I(t)+D(t^{*}),\qquad U% (t,0)=V_{T},= - divide start_ARG italic_U end_ARG start_ARG italic_τ end_ARG + italic_u ( italic_t ) + italic_k italic_I ( italic_t ) + italic_D ( italic_t start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ) , italic_U ( italic_t , 0 ) = italic_V start_POSTSUBSCRIPT italic_T end_POSTSUBSCRIPT , (14c)
I(t)𝐼𝑡\displaystyle I(t)italic_I ( italic_t ) =tτItνN(s)𝑑s,absentsuperscriptsubscript𝑡subscript𝜏𝐼𝑡subscript𝜈𝑁𝑠differential-d𝑠\displaystyle=\int_{t-\tau_{I}}^{t}\nu_{N}(s)\,ds,= ∫ start_POSTSUBSCRIPT italic_t - italic_τ start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_t end_POSTSUPERSCRIPT italic_ν start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT ( italic_s ) italic_d italic_s , (14d)
ν(t)𝜈𝑡\displaystyle\nu(t)italic_ν ( italic_t ) =[0ρ(t,t)H(U(t,t))𝑑t+H¯(t)(10ρ(t,t)𝑑t)]+,absentsubscriptdelimited-[]superscriptsubscript0𝜌𝑡superscript𝑡𝐻𝑈𝑡superscript𝑡differential-dsuperscript𝑡¯𝐻𝑡1superscriptsubscript0𝜌𝑡superscript𝑡differential-dsuperscript𝑡\displaystyle=\left[\int_{0}^{\infty}\rho(t,t^{*})\leavevmode\nobreak\ H(U(t,t% ^{*}))\,dt^{*}+\bar{H}(t)\left(1-\int_{0}^{\infty}\rho(t,t^{*})\,dt^{*}\right)% \right]_{+},= [ ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_ρ ( italic_t , italic_t start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ) italic_H ( italic_U ( italic_t , italic_t start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ) ) italic_d italic_t start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT + over¯ start_ARG italic_H end_ARG ( italic_t ) ( 1 - ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_ρ ( italic_t , italic_t start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ) italic_d italic_t start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ) ] start_POSTSUBSCRIPT + end_POSTSUBSCRIPT , (14e)
νN(t)subscript𝜈𝑁𝑡\displaystyle\nu_{N}(t)italic_ν start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT ( italic_t ) =ν(t)+ν(t)Nξ(t),absent𝜈𝑡𝜈𝑡𝑁𝜉𝑡\displaystyle=\nu(t)+\sqrt{\frac{\nu(t)}{N}}\xi(t),= italic_ν ( italic_t ) + square-root start_ARG divide start_ARG italic_ν ( italic_t ) end_ARG start_ARG italic_N end_ARG end_ARG italic_ξ ( italic_t ) , (14f)
with
H¯(t)=0H(U(t,t))(1S(t,t))ρ(t,t)𝑑t0(1S(t,t))ρ(t,t)𝑑t,¯𝐻𝑡superscriptsubscript0𝐻𝑈𝑡superscript𝑡1𝑆𝑡superscript𝑡𝜌𝑡superscript𝑡differential-dsuperscript𝑡superscriptsubscript01𝑆𝑡superscript𝑡𝜌𝑡superscript𝑡differential-dsuperscript𝑡\bar{H}(t)=\frac{\int_{0}^{\infty}H(U(t,t^{*}))(1-S(t,t^{*}))\rho(t,t^{*})\,dt% ^{*}}{\int_{0}^{\infty}(1-S(t,t^{*}))\rho(t,t^{*})\,dt^{*}},over¯ start_ARG italic_H end_ARG ( italic_t ) = divide start_ARG ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_H ( italic_U ( italic_t , italic_t start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ) ) ( 1 - italic_S ( italic_t , italic_t start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ) ) italic_ρ ( italic_t , italic_t start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ) italic_d italic_t start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT end_ARG start_ARG ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT ( 1 - italic_S ( italic_t , italic_t start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ) ) italic_ρ ( italic_t , italic_t start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ) italic_d italic_t start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT end_ARG , (14g)

where ξ(t)𝜉𝑡\xi(t)italic_ξ ( italic_t ) is the white Gaussian noise of unity amplitude. We remark that, for all t,t𝑡superscript𝑡t,t^{*}italic_t , italic_t start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT, S(t,t),[0,1]S(t,t^{*}),\in[0,1]italic_S ( italic_t , italic_t start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ) , ∈ [ 0 , 1 ]; hence, the function H¯(t)¯𝐻𝑡\bar{H}(t)over¯ start_ARG italic_H end_ARG ( italic_t ) is non-negative for all t𝑡titalic_t. Furthermore, we note that νN(t)subscript𝜈𝑁𝑡\nu_{N}(t)italic_ν start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT ( italic_t ) must be regarded as an abstract quantity that mathematically only makes sense as a distribution, i.e. if it is integrated against some test function over time. Specifically, the empirical rate of new cases ν^Nsubscript^𝜈𝑁\hat{\nu}_{N}over^ start_ARG italic_ν end_ARG start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT measured over some finite time step ΔtΔ𝑡\Delta troman_Δ italic_t (such that the expected number of new cases Nν(t)Δt1much-greater-than𝑁𝜈𝑡Δ𝑡1N\nu(t)\Delta t\gg 1italic_N italic_ν ( italic_t ) roman_Δ italic_t ≫ 1) would be

ν^N(t)=[tt+ΔtνN(s)𝑑s]+.subscript^𝜈𝑁𝑡subscriptdelimited-[]superscriptsubscript𝑡𝑡Δ𝑡subscript𝜈𝑁𝑠differential-d𝑠\hat{\nu}_{N}(t)=\left[\int_{t}^{t+\Delta t}\nu_{N}(s)\,ds\right]_{+}.over^ start_ARG italic_ν end_ARG start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT ( italic_t ) = [ ∫ start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_t + roman_Δ italic_t end_POSTSUPERSCRIPT italic_ν start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT ( italic_s ) italic_d italic_s ] start_POSTSUBSCRIPT + end_POSTSUBSCRIPT . (15)

While in practice it is extremely rare that the integral takes negative values for large N𝑁Nitalic_N, we added the rectification []+subscriptdelimited-[][\cdot]_{+}[ ⋅ ] start_POSTSUBSCRIPT + end_POSTSUBSCRIPT to enforce a non-negative rate. The same statement also holds true for Eq. (14e).

Likewise, ρ𝜌\rhoitalic_ρ is not a normalized probability density [25, 28]. Therefore, it is no longer the strict interpretation of the empirical density of the times since the last infection but ρ(t,t)dt𝜌𝑡superscript𝑡𝑑superscript𝑡\rho(t,t^{*})dt^{*}italic_ρ ( italic_t , italic_t start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ) italic_d italic_t start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT must be regarded as an abstract measure. In comparison with the macroscopic model, this system of equations is stochastic. Its solution corresponds to a single realization of the noise ξ(t)𝜉𝑡\xi(t)italic_ξ ( italic_t ). When Nabsent𝑁N\xrightarrow{}\inftyitalic_N start_ARROW start_OVERACCENT end_OVERACCENT → end_ARROW ∞, the solution converges to the solution of the macroscopic model. The precise numerical method employed for these simulations is described in [25] (see also the provided simulation code in Python).

5 Simulations

Model with escape noise

A first simulation of epidemic waves is shown in Fig. 2. The onset of an epidemic is modeled by a short pulse of magnitude I0subscript𝐼0I_{0}italic_I start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT of the rate of new infections ν(t)𝜈𝑡\nu(t)italic_ν ( italic_t ) at time t=0𝑡0t=0italic_t = 0 reflecting the appearance of contaminating patients. These conditions result in oscillations of the rate of cases ν(t)𝜈𝑡\nu(t)italic_ν ( italic_t ) and the fraction of potentially infected population I(t)𝐼𝑡I(t)italic_I ( italic_t ). These oscillations tend to relax. The dynamics of a single individual state Vi(t)subscript𝑉𝑖𝑡V_{i}(t)italic_V start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_t ) (Fig. 2A, next to bottom panel) follows the epidemic waves of ν(t)𝜈𝑡\nu(t)italic_ν ( italic_t ) and I(t)𝐼𝑡I(t)italic_I ( italic_t ). Each peak of Vi(t)subscript𝑉𝑖𝑡V_{i}(t)italic_V start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_t ) reflects the virus reproduction during the disease. The decreasing phase reflects the immune response, after which an individual becomes again susceptible and might get infected during the next contaminating flux k(t)I(t)𝑘𝑡𝐼𝑡k(t)I(t)italic_k ( italic_t ) italic_I ( italic_t ). Each spike of the rate of cases ν(t)𝜈𝑡\nu(t)italic_ν ( italic_t ) leads to a correspondingly shaped distribution of the density in the tsuperscript𝑡t^{*}italic_t start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT-space.

A                                                                                      B

Refer to caption Refer to caption

Figure 2: Escape noise in the form of (8). Simulation of the response to the rapid change of the rate of viral load k𝑘kitalic_k from 0.6 to 2.4 and the short infecting pulse I0(t)subscript𝐼0𝑡I_{0}(t)italic_I start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( italic_t ), at t=100𝑡100t=100italic_t = 100. A, top to bottom: I0(t)subscript𝐼0𝑡I_{0}(t)italic_I start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( italic_t ), k(t)𝑘𝑡k(t)italic_k ( italic_t ), the fraction of infected individuals I(t)𝐼𝑡I(t)italic_I ( italic_t ), the state variable Vi(t)subscript𝑉𝑖𝑡V_{i}(t)italic_V start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_t ) of a representative individual, and the rate of new cases ν(t)𝜈𝑡\nu(t)italic_ν ( italic_t ). For I(t)𝐼𝑡I(t)italic_I ( italic_t ) and ν(t)𝜈𝑡\nu(t)italic_ν ( italic_t ), the solutions obtained with both macro- (black lines) and microscopic equations for N=20000𝑁20000N=20000italic_N = 20000 (blue lines) are shown. B, the distribution in the tsuperscript𝑡t^{*}italic_t start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT-space of the mean state variable U𝑈Uitalic_U, the source term ρH𝜌𝐻\rho Hitalic_ρ italic_H, and the density ρ𝜌\rhoitalic_ρ, in 4 time moments. Parameters: m=1𝑚1m=1italic_m = 1, c=0.015𝑐0.015c=0.015italic_c = 0.015 days-1, τ=50𝜏50\tau=50italic_τ = 50, τI=10subscript𝜏𝐼10\tau_{I}=10italic_τ start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT = 10, τr=1subscript𝜏𝑟1\tau_{r}=1italic_τ start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT = 1, τR=20subscript𝜏𝑅20\tau_{R}=20italic_τ start_POSTSUBSCRIPT italic_R end_POSTSUBSCRIPT = 20, a=100𝑎100a=100italic_a = 100, and VT=2superscript𝑉𝑇2V^{T}=2italic_V start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT = 2.

The population waves move in the tsuperscript𝑡t^{*}italic_t start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT-space from t=0superscript𝑡0t^{*}=0italic_t start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT = 0 towards the region where U𝑈Uitalic_U is about to cross the threshold VTsuperscript𝑉𝑇V^{T}italic_V start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT and form the peaks of ρH𝜌𝐻\rho Hitalic_ρ italic_H, which is the distribution of falling sick individuals (Fig. 2B). Since the integral over ρH𝜌𝐻\rho Hitalic_ρ italic_H determines the rate of cases ν(t)𝜈𝑡\nu(t)italic_ν ( italic_t ), the increase of the ρH𝜌𝐻\rho Hitalic_ρ italic_H determines the next peak of ν(t)𝜈𝑡\nu(t)italic_ν ( italic_t ) and I(t)𝐼𝑡I(t)italic_I ( italic_t ), and so on.

We assume that at t=t1=700𝑡subscript𝑡1700t=t_{1}=700italic_t = italic_t start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = 700 the basic reproduction rate drops 4 fold, which might be due to any kind of containment measures. As a result, the epidemic wave generation stops through a significant decrease in the rate of cases ν(t)𝜈𝑡\nu(t)italic_ν ( italic_t ) (Fig. 2A).

For the escape noise in the form of (8), the epidemic depends on the steepness m𝑚mitalic_m of the hazard function (Fig. 3). With the increase of m𝑚mitalic_m from 0.5 to 1, the epidemic oscillations increase. However, it is non-trivial that in the case of more steep H𝐻Hitalic_H with m=2𝑚2m=2italic_m = 2 the epidemic shows only one peak. In this case, the rapid illness of the whole population results in a short splash of contamination k(t)I(t)𝑘𝑡𝐼𝑡k(t)I(t)italic_k ( italic_t ) italic_I ( italic_t ) that occurs at the recovery phase of the population when no one is susceptible.

Refer to caption

Figure 3: Escape noise in the form of (8) with m=𝑚absentm=italic_m = 0.5, 1 and 2. The remaining parameter values are as in Fig. 2.

Seasonality

One of the experimentally observed features of epidemics is its seasonal oscillations, as illustrated in Fig. 4 by data from [21]. In Fig. 4, we illustrate the case of seasonal change of k𝑘kitalic_k through a step function alternating between k=2.4𝑘2.4k=2.4italic_k = 2.4 and k=1.2𝑘1.2k=1.2italic_k = 1.2 every 180 days. In response to a meander-like k𝑘kitalic_k, we observe a complex pattern of waves with various amplitudes.

A

Refer to caption

B

Refer to caption

Figure 4: Epidemiology of Seasonal Coronaviruses. A, Data from [21]; specifically, the monthly prevalence of seasonal coronaviruses (sCoVs) detected among patients with respiratory illness virologically tested in NHS Greater Glasgow and Clyde, Scotland, United Kingdom. B, The case of oscillating k𝑘kitalic_k. Escape noise in the form of (8). The remaining parameter values are as in Fig. 2.

Model with white Gaussian noise

When simulating the system with the white Gaussian noise (Fig. 5), we observed a similar behavior. The onset of an epidemic is again modeled by a short pulse of the rate of new infections ν(t)𝜈𝑡\nu(t)italic_ν ( italic_t ) of magnitude I0subscript𝐼0I_{0}italic_I start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT at time t=0𝑡0t=0italic_t = 0 reflecting the appearance of infectious individuals. At the same time, the viral load noise jumps (σ𝜎\sigmaitalic_σ changes from 0.50.50.50.5 to 2222). These conditions result in sustained oscillations of the rate of cases ν(t)𝜈𝑡\nu(t)italic_ν ( italic_t ) and the fraction of potentially infected population I(t)𝐼𝑡I(t)italic_I ( italic_t ). Again, the dynamics of a single individual state Vi(t)subscript𝑉𝑖𝑡V_{i}(t)italic_V start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_t ) (Fig. 5A, next to bottom panel) follows the epidemic waves of ν(t)𝜈𝑡\nu(t)italic_ν ( italic_t ) and I(t)𝐼𝑡I(t)italic_I ( italic_t ). When at t=t1=900𝑡subscript𝑡1900t=t_{1}=900italic_t = italic_t start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = 900 the basic reproduction rate k𝑘kitalic_k drops due to assumed anti-epidemic measures, the epidemic wave generation stops.

A                                                                                      B

Refer to caption Refer to caption

C                                                                   D

Refer to caption Refer to caption

Figure 5: White noise. Simulation of the response to the rapid change of the rate of viral load k𝑘kitalic_k from 3 to 1.2 and the noise amplitude σ(t)𝜎𝑡\sigma(t)italic_σ ( italic_t ) from 0.5 to 2, at t=t0=0𝑡subscript𝑡00t=t_{0}=0italic_t = italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 0. A, top to bottom: k(t)𝑘𝑡k(t)italic_k ( italic_t ), σ(t)𝜎𝑡\sigma(t)italic_σ ( italic_t ), the fraction of potentially infected population I(t)𝐼𝑡I(t)italic_I ( italic_t ), the state variable Vi(t)subscript𝑉𝑖𝑡V_{i}(t)italic_V start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_t ) of a representative individual, and the rate of new cases ν(t)𝜈𝑡\nu(t)italic_ν ( italic_t ). For I(t)𝐼𝑡I(t)italic_I ( italic_t ) and ν(t)𝜈𝑡\nu(t)italic_ν ( italic_t ), the solutions obtained with both macro (black lines) and microscopic equations for N=20000𝑁20000N=20000italic_N = 20000 (blue lines) are shown. B, the distribution in the tsuperscript𝑡t^{*}italic_t start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT-space of the mean state variable U𝑈Uitalic_U, the source term ρH𝜌𝐻\rho Hitalic_ρ italic_H, and the density ρ𝜌\rhoitalic_ρ, in 4 time moments. C, Response to periodically changing k𝑘kitalic_k mimicking change of seasons (VT=4superscript𝑉𝑇4V^{T}=4italic_V start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT = 4). D, Response to periodically changing k𝑘kitalic_k mimicking change of seasons (VT=3.7superscript𝑉𝑇3.7V^{T}=3.7italic_V start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT = 3.7). Parameters: τ=50𝜏50\tau=50italic_τ = 50, τI=10subscript𝜏𝐼10\tau_{I}=10italic_τ start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT = 10, τr=1subscript𝜏𝑟1\tau_{r}=1italic_τ start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT = 1, τR=20subscript𝜏𝑅20\tau_{R}=20italic_τ start_POSTSUBSCRIPT italic_R end_POSTSUBSCRIPT = 20, a=50𝑎50a=50italic_a = 50, VT=4superscript𝑉𝑇4V^{T}=4italic_V start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT = 4, and σ=2π𝜎2𝜋\sigma=2\sqrt{\pi}italic_σ = 2 square-root start_ARG italic_π end_ARG.

Mesoscopic model

In order to consider a finite number of individuals in a population, we apply the mesoscopic model described in section 4, based on Eqs.(14a)-(14g). The solutions are different for different N𝑁Nitalic_N (Fig.6). The solutions for small N𝑁Nitalic_N differ significantly for different realizations of the noise. The model in the limit Nabsent𝑁N\xrightarrow{}\inftyitalic_N start_ARROW start_OVERACCENT end_OVERACCENT → end_ARROW ∞ is equivalent to the macroscopic model. Hence, the solutions for large N𝑁Nitalic_N converge for different realizations and approach the macroscopic model solution shown in Fig. 2.

Refer to caption

Figure 6: Mesoscopic model (8) for different N=200𝑁200N=200italic_N = 200, 2000, and 20000, and different realizations of noise. The parameters are as in Fig. 2. The solutions with Nabsent𝑁N\xrightarrow{}\inftyitalic_N start_ARROW start_OVERACCENT end_OVERACCENT → end_ARROW ∞ converge to the macroscopic model solution shown in Fig. 2.

6 Discussion

In the present work, we have proposed the description of an epidemic that stems from a microscopic model of an individual and results in the meso- and macroscopic models of the entire population. The microscopic evolution is described by the dynamics of two state variables for each individual, the interplay between the viral load and immune response Visubscript𝑉𝑖V_{i}italic_V start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT, and the time since the last infection tisubscriptsuperscript𝑡𝑖t^{*}_{i}italic_t start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT. The corresponding mesoscopic and macroscopic models are then given by deterministic and stochastic refractory-density equations, respectively. We have shown the consistency between the models. The presented simulations remind of epidemic “waves”, for instance, the ones reported in [21]. One of our basic ideas to describe the epidemic in terms of the evolution of viral load depending on the time since the last infection is supported by the experimental data from [16], where the viral load was measured in patients with SARS-CoV-2 as a function of the time since infection. Our multi scale modeling framework allows us to take into account such microscopic parameters measured in individuals in order to constrain the parameters of the meso- or macro-scale models.

The approach and the model that we presented in this manuscript could naturally be extended and generalized. We conclude the manuscript with a few promising leads for future research, which we identified during the writing of this work.

6.1 Treatment

In simulations, we varied the individual rate of viral load k(t)𝑘𝑡k(t)italic_k ( italic_t ), which accounted for the seasonality. This factor can also reflect such social measures as wearing masks and isolation. In contrast, medical treatment of patients is a function of the time since infection tsuperscript𝑡t^{*}italic_t start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT, and it changes the course of the disease. Naturally, this factor can be taken into account through modification of the shape of the function D𝐷Ditalic_D entering the equation for the interplay of viral load and immune response, Eq. (1).

6.2 Modelling infectiousness depending on infection age

It has been shown that the infectiousness of SARS-CoV-2 depends on the time since the last infection [16]. This can be easily modeled in our framework by generalizing Eq. (3), Eq. (12d) and Eq. (14d) for the micro-, macro- and mesoscopic fractions of infected people to

I(t)=1N0tκ(ts)𝑑Nc(s),I(t)=0tκ(ts)ν(s)𝑑sandI(t)=0tκ(ts)νN(s)𝑑s,formulae-sequence𝐼𝑡1𝑁superscriptsubscript0𝑡𝜅𝑡𝑠differential-dsubscript𝑁𝑐𝑠formulae-sequence𝐼𝑡superscriptsubscript0𝑡𝜅𝑡𝑠𝜈𝑠differential-d𝑠and𝐼𝑡superscriptsubscript0𝑡𝜅𝑡𝑠subscript𝜈𝑁𝑠differential-d𝑠I(t)=\frac{1}{N}\int_{0}^{t}\kappa(t-s)\,dN_{c}(s),\quad I(t)=\int_{0}^{t}% \kappa(t-s)\nu(s)\,ds\quad\text{and}\quad I(t)=\int_{0}^{t}\kappa(t-s)\nu_{N}(% s)\,ds,italic_I ( italic_t ) = divide start_ARG 1 end_ARG start_ARG italic_N end_ARG ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_t end_POSTSUPERSCRIPT italic_κ ( italic_t - italic_s ) italic_d italic_N start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ( italic_s ) , italic_I ( italic_t ) = ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_t end_POSTSUPERSCRIPT italic_κ ( italic_t - italic_s ) italic_ν ( italic_s ) italic_d italic_s and italic_I ( italic_t ) = ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_t end_POSTSUPERSCRIPT italic_κ ( italic_t - italic_s ) italic_ν start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT ( italic_s ) italic_d italic_s , (16)

respectively. The kernel κ:+[0,1]:𝜅superscript01\kappa:\mathbb{R}^{+}\to[0,1]italic_κ : blackboard_R start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT → [ 0 , 1 ] describes the infectiousness depending on the infection course. Our equations (3),  (12d) and Eq. (14d) are recovered as the special case κ(t)=𝟙(0,τI)(t)𝜅superscript𝑡subscript10subscript𝜏𝐼superscript𝑡\kappa(t^{*})=\mathbbm{1}_{(0,\tau_{I})}(t^{*})italic_κ ( italic_t start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ) = blackboard_1 start_POSTSUBSCRIPT ( 0 , italic_τ start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT ) end_POSTSUBSCRIPT ( italic_t start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ), allowing only two possible states – infectious and non-infectious. Our infectious period τI=10subscript𝜏𝐼10\tau_{I}=10\leavevmode\nobreak\ italic_τ start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT = 10days well correspond to the half-duration of cell culture infectivity measured in [16] for SARS-CoV-2.

6.3 Network of multiple populations

Certain types of heterogeneity (e.g. different social communities, age groups, etc.) and spatial structure (cities) can be treated by splitting a large heterogeneous population into smaller homogeneous subpopulations, again in analogy to RDE application in neuroscience, for instance, to simulate cortical neuronal populations [9]. Moreover, the methodology can be extended to interacting populations, such as in the scenario of epidemic transmission between countries, consideration of a network of communities, similar to the so-called multi-group approach (see e.g. [1, 23]). Since the multi-group approach naturally operates with smaller-size subpopulations, the proposed mesoscopic approach could be applied in this case.

6.4 Multiple internal variables

The proposed derivation of the population model from that for an individual applies not only to the 1-D individual model described by a single ODE for the viral state Visubscript𝑉𝑖V_{i}italic_V start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT, but also to multidimensional cases, in which one may be interested in including additional characteristics of individuals and/or of populations, and keep track of these as the disease spreads. In this way, additional internal variables can be introduced, describing, for instance, activation and/or inactivation of some processes in the immune system. This concept resembles the consideration of Hodgkin-Huxley-like neurons in neuroscience [7]. If an individual is characterized by n𝑛nitalic_n ODEs, the population model comprises n+1𝑛1n+1italic_n + 1 partial differential equations (PDEs) for the state variable and the density. These equations remain one-dimensional, with time since the infection tsuperscript𝑡t^{*}italic_t start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT is the sole independent variable, besides t𝑡titalic_t. Consequently, this approach maintains computational efficiency even for multi-dimensional individuals. Leveraging this advantage, more intricate models can be developed.

In comparison with the above-mentioned kinetic theory [20] that results in the Fokker-Planck equation describing the evolution of the population in the phase space of a state variable, the proposed approach considers the evolution in the phase space of the time since infection. Once the microscopic models are identical, these two macroscopic approaches would result in very similar numerical simulations, at least in the case of white Gaussian noise [27]. In the more general case of non-scalar state variables, the Fokker-Planck equation becomes multidimensional, whereas the transport equations of the refractory density approach remain 1-D PDEs that are effectively solvable.

6.5 Age-dependence

The epidemics to a different extent affect people of different ages [16]. Our approach allows the consideration of the population distributed by age a^^𝑎\hat{a}over^ start_ARG italic_a end_ARG via an age-dependent hazard function H(V,t,a^)𝐻𝑉superscript𝑡^𝑎H(V,t^{*},\hat{a})italic_H ( italic_V , italic_t start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT , over^ start_ARG italic_a end_ARG ). In this case, the main variables have to be parameterized by a^^𝑎\hat{a}over^ start_ARG italic_a end_ARG as ρ=ρ(t,t,a^)𝜌𝜌𝑡superscript𝑡^𝑎\rho=\rho(t,t^{*},\hat{a})italic_ρ = italic_ρ ( italic_t , italic_t start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT , over^ start_ARG italic_a end_ARG ) and U=U(t,t,a^)𝑈𝑈𝑡superscript𝑡^𝑎U=U(t,t^{*},\hat{a})italic_U = italic_U ( italic_t , italic_t start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT , over^ start_ARG italic_a end_ARG ). Assuming that the hazard function changes much slower with age a^^𝑎\hat{a}over^ start_ARG italic_a end_ARG than with the time since the last infection (i.e. |H/a^||H/t|much-less-than𝐻^𝑎𝐻superscript𝑡|\partial H/\partial\hat{a}|\ll|\partial H/\partial t^{*}|| ∂ italic_H / ∂ over^ start_ARG italic_a end_ARG | ≪ | ∂ italic_H / ∂ italic_t start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT |), the age a^^𝑎\hat{a}over^ start_ARG italic_a end_ARG can be treated quasi-statically as a parameter and not as an independent dynamical variable. Under this plausible assumption, the equations of the macroscopic model remain to be 1-D transport equations. However, the fraction of infected individuals I(t)𝐼𝑡I(t)italic_I ( italic_t ) would be an integral over a^^𝑎\hat{a}over^ start_ARG italic_a end_ARG. This parametrization would allow the consideration of different disease time courses, etc. The infectivity κ(t,a^)𝜅superscript𝑡^𝑎\kappa(t^{*},\hat{a})italic_κ ( italic_t start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT , over^ start_ARG italic_a end_ARG ) can also be dependent on a^^𝑎\hat{a}over^ start_ARG italic_a end_ARG because usually people interact stronger within their age groups.

6.6 Asymptomatic spread

We remark that, for our model, we are interested in the ability of an individual to spread the disease, rather than showing symptoms; hence, the infected and infectious population might be generalized to include asymptomatic individuals, thus allowing to adapt our construction to more complex compartmental models such as the ones presented in [4, 5, 15, 22, 23]. This may be done in multiple ways, depending on which characteristics of symptomatic and asymptomatic spread one wishes to capture in their model.

6.7 Heterogeneous severity

It could be of biological interest to assume a heterogeneous setting, e.g. the severity a𝑎aitalic_a in Eq. (5) could vary among individuals depending on the vaccine status, stronger or weaker immune system, etc.; whereas we can assume that τrsubscript𝜏𝑟\tau_{r}italic_τ start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT is approximately constant and depends only on the specific disease. However, this generalization would carry a non-negligible increase in the computational cost of our simulations, and make mean field approximations which we perform in section 3 and 4 considerably more difficult. One potential strategy to deal with such heterogeneity would be to split the population into several, approximately homogeneous subpopulations and use a multi-population version of our model as described above. Importantly, in this scenario, we could take full advantage of the mesoscopic model because it remains valid for small sub-populations. We leave such a generalization to heterogeneous systems as a promising outlook for future work.

6.8 Complexity and parameters

Despite more mathematical complexity in comparison with the classical SEIRS models (PDEs versus ODEs), the proposed model has a similar number of parameters. The SEIRS model has 5 parameters (β𝛽\betaitalic_β, σ𝜎\sigmaitalic_σ, γ𝛾\gammaitalic_γ, and ξ𝜉\xiitalic_ξ, in traditional notations, and the time scale). Instead, in the case of omitted explicit course of the recovery D(t)𝐷𝑡D(t)italic_D ( italic_t ), the proposed model has also only 5 parameters: τ𝜏\tauitalic_τ, τIsubscript𝜏𝐼\tau_{I}italic_τ start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT, k𝑘kitalic_k, σ𝜎\sigmaitalic_σ, and VTsuperscript𝑉𝑇V^{T}italic_V start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT.

6.9 Inverse problem

The model may be used in future studies for solving reverse problems of finding the parameters from statistical data for a certain disease. For this purpose, one should have a sufficient set of experimental data that would reflect the effects of each of the parameters. For instance, the effect of k𝑘kitalic_k can be studied with the effect of masks; the effect of τIsubscript𝜏𝐼\tau_{I}italic_τ start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT with the effect of care; the effect of VTsuperscript𝑉𝑇V^{T}italic_V start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT with the effect of provisional stimulation of the immune system of the entire population; the effect of σ𝜎\sigmaitalic_σ with a selection of more homogeneous subpopulation; the effect of τ𝜏\tauitalic_τ with the selection of faster (or slower) recovering subpopulation, and so on.

6.10 Limitations

In this first work on the topic, we made several simplifying assumptions. One of the most important is that our model assumes a strictly conserved population: no removal, no mortality, closed community: no external influx, no out-flux. It would be of interest to consider, instead, a similar model in which the population is allowed to vary in time, for any of the reasons listed above.

Declaration of Competing Interest

The authors have no competing interest to declare.

Declaration of equal contribution

All authors contributed equally to this work.

Code availability

The code used to numerically solve the models presented in this manuscript is available at
https://github.com/schwalger/refracdens_epidem.

Acknowledgements

Mattia Sensi was partially supported by the Italian Ministry for University and Research (MUR) through the PRIN 2020 project “Integrated Mathematical Approaches to Socio-Epidemiological Dynamics” (No. 2020JLWP23, CUP: E15F21005420006).

Appendix A Macroscopic vs microscopic models, white Gaussian vs. escape noise

To illustrate the agreement between microscopic and mesoscopic models as well as escape noise and Gaussian white noise, we consider a slightly simplified setup. Whereas the main model equation (1) includes the term k(t)I(t)𝑘𝑡𝐼𝑡k(t)I(t)italic_k ( italic_t ) italic_I ( italic_t ), which is recurrently dependent on the activity of the entire population ν(t)𝜈𝑡\nu(t)italic_ν ( italic_t ), here we consider a simplified problem with a step-wise input signal instead of the term k(t)I(t)𝑘𝑡𝐼𝑡k(t)I(t)italic_k ( italic_t ) italic_I ( italic_t ). This case can be interpreted as a rough approximation of an epidemic problem, where the time course of the epidemic is assumed to be known and shaped as a step function. In this interpretation, we are interested in the probabilistic behavior of an individual, with ν(t)𝜈𝑡\nu(t)italic_ν ( italic_t ) to be a probability for this individual to get ill. We simulate this response for the case of white noise and the escape noise in the form of Eq. (10) (Fig.7).

Refer to caption

Figure 7: Solutions of a test with a stepwise term instead of k(t)I(t)𝑘𝑡𝐼𝑡k(t)I(t)italic_k ( italic_t ) italic_I ( italic_t ) in the Eqs. (1) and (12d), obtained with the microscopic model (N=2000𝑁2000N=2000italic_N = 2000) with the white Gaussian noise and the escape noise described with Eq. (10), and the microscopic model. The value of k(t)I(t)𝑘𝑡𝐼𝑡k(t)I(t)italic_k ( italic_t ) italic_I ( italic_t ) is 0 for t[0,100)𝑡0100t\in[0,100)italic_t ∈ [ 0 , 100 ) and t>500𝑡500t>500italic_t > 500 and 0.40.40.40.4 for t[100,500]𝑡100500t\in[100,500]italic_t ∈ [ 100 , 500 ].

From the methodical point of view, Fig. 7 shows the testing comparison of the macroscopic model with the microscopic models with white Gaussian and escape noise. The microscopic solutions converge to the macroscopic ones.

References

  • [1] M. Adimy, A. Chekroun, L. Pujo-Menjouet, and M. Sensi. A multigroup approach to delayed prion production. Discrete and Continuous Dynamical Systems-B, 2023.
  • [2] J. A. Alfaro-Murillo, Z. Feng, and J. W. Glasser. Analysis of an epidemiological model structured by time-since-last-infection. Journal of differential equations, 267(10):5631–5661, 2019.
  • [3] R. M. Anderson. The epidemiology of HIV infection: variable incubation plus infectious periods and heterogeneity in sexual activity. J. Roy. Stat. Soc.: Series A (Statistics in Society), 151(1):66–93, 1988.
  • [4] S. Ansumali, S. Kaushal, A. Kumar, M. K. Prakash, and M. Vidyasagar. Modelling a pandemic with asymptomatic patients, impact of lockdown and herd immunity, with applications to SARS-CoV-2. Annual reviews in control, 50:432–447, 2020.
  • [5] J. Arino and S. Portet. A simple model for COVID-19. Infectious Disease Modelling, 5:309–315, 2020.
  • [6] F. Brauer and J. Watmough. Age of infection epidemic models with heterogeneous mixing. Journal of biological dynamics, 3(2-3):324–330, 2009.
  • [7] A. V. Chizhov and L. J. Graham. Population model of hippocampal pyramidal neurons, linking a refractory density approach to conductance-based neurons. Physical Review E, 75(1):011924, 2007.
  • [8] A. V. Chizhov and L. J. Graham. Efficient evaluation of neuron populations receiving colored-noise current based on a refractory density method. Phys. Rev. E, 77(1):011910, 2008.
  • [9] A. V. Chizhov and L. J. Graham. A strategy for mapping biophysical to abstract neuronal network models applied to primary visual cortex. PLoS Comput. Biol., 17(8):e1009007, 2021.
  • [10] R. Della Marca, N. Loy, and A. Tosin. An SIR–like kinetic model tracking individuals’ viral load. Networks and Heterogeneous Media, 17(3):467–494, 2022.
  • [11] G. Dumont, J. Henry, and C. O. Tarniceriu. Oscillations in a Fully Connected Network of Leaky Integrate-and-Fire Neurons with a Poisson Spiking Mechanism. J. Nonlin. Sci, 34(1):18, 2024.
  • [12] G. Dumont and C. O. Tarniceriu. Pattern Formation in a Spiking Neural-Field of Renewal Neurons, 2024.
  • [13] A. d’Onofrio, M. Iannelli, P. Manfredi, and G. Marinoschi. Optimal epidemic control by social distancing and vaccination of an infection structured by time since infection: the COVID-19 case study. SIAM Journal on Applied Mathematics, pages S199–S224, 2023.
  • [14] W. Gerstner and W. M. Kistler. Spiking Neuron Models: Single Neurons, Populations, Plasticity. Cambridge University Press, August 2002. Google-Books-ID: Rs4oc7HfxIUC.
  • [15] X. Huo, J. Chen, and S. Ruan. Estimating asymptomatic, undetected and total cases for the COVID-19 outbreak in Wuhan: a mathematical modeling study. BMC Infectious Diseases, 21(1):476, 2021.
  • [16] T. C. Jones, G. Biele, B. Mühlemann, T. Veith, J. Schneider, J. Beheim-Schwarzbach, T. Bleicker, J. Tesch, M. L. Schmidt, L. E. Sander, Fl. Kurth, P. Menzel, R. Schwarzer, M. Zuchowski, J. Hofmann, A. Krumbholz, A. Stein, A. Edelmann, V. M. Corman, and C. Drosten. Estimating infectiousness throughout SARS-CoV-2 infection course. Science, 373(6551):eabi5273, 2021.
  • [17] W. O. Kermack and A. G. 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(772):700–721, 1927.
  • [18] L. L. Li, C. P. Ferreira, and B. Ainseba. Mathematical analysis of an age structured problem modelling phenotypic plasticity in mosquito behaviour. Nonlinear Analysis: Real World Applications, 48:410–423, 2019.
  • [19] M. Y. Li and L. Wang. Global stability in some SEIR epidemic models. In Mathematical approaches for emerging and reemerging infectious diseases: models, methods, and theory, pages 295–311. Springer, 2002.
  • [20] N. Loy and A. Tosin. Boltzmann-type equations for multi-agent systems with label switching. Kinetic and Related Models, 14(5):867–894, 2021.
  • [21] S. Nickbakhsh, A. Ho, D. F. P. Marques, J. McMenamin, R. N. Gunson, and P. R. Murcia. Epidemiology of seasonal coronaviruses: establishing the context for the emergence of coronavirus disease 2019. The Journal of infectious diseases, 222(1):17–25, 2020.
  • [22] S. Ottaviano, M. Sensi, and S. Sottile. Global stability of SAIRS epidemic models. Nonlinear Analysis: Real World Applications, 65:103501, 2022.
  • [23] S. Ottaviano, M. Sensi, and S. Sottile. Global stability of multi-group SAIRS epidemic models. Mathematical Methods in the Applied Sciences, 46(13):14045–14071, 2023.
  • [24] K. Pakdaman, B. Perthame, and D. Salort. Dynamics of a structured neuron population. Nonlinearity, 23(1):55–75, 2009.
  • [25] V. Schmutz, E. Löcherbach, and T. Schwalger. On a Finite-Size Neuronal Population Equation. SIAM J. Applied Dynamical Systems, 22(2):996–1029, 2023.
  • [26] T. Schwalger. Mapping Input Noise to Escape Noise in Integrate-and-fire neurons: A Level-Crossing Approach. Biol. Cybern., 115:539–562, 2021.
  • [27] T. Schwalger and A. V. Chizhov. Mind the last spike—firing rate models for mesoscopic populations of spiking neurons. Current opinion in neurobiology, 58:155–166, 2019.
  • [28] T. Schwalger, M. Deger, and W. Gerstner. Towards a theory of cortical columns: From spiking neurons to interacting neural populations of finite size. PLoS Comput Biol, 13(4):e1005507, 2017.
  • [29] C. O. Tarniceriu. Age-structure in neuronal models. Annals of the Alexandru Ioan Cuza University-Mathematics, 66(2):385–396, 2020.