Skip to main content
Springer logoLink to Springer
. 2023 Oct 27;87(6):78. doi: 10.1007/s00285-023-02008-1

Linking within- and between-host scales for understanding the evolutionary dynamics of quantitative antimicrobial resistance

Martin L Mann-Manyombe 1,2, Abdoulaye Mendy 1,2, Ousmane Seydi 2, Ramsès Djidjou-Demasse 1,2,
PMCID: PMC10611892  PMID: 37889337

Abstract

Understanding both the epidemiological and evolutionary dynamics of antimicrobial resistance is a major public health concern. In this paper, we propose a nested model, explicitly linking the within- and between-host scales, in which the level of resistance of the bacterial population is viewed as a continuous quantitative trait. The within-host dynamics is based on integro-differential equations structured by the resistance level, while the between-host scale is additionally structured by the time since infection. This model simultaneously captures the dynamics of the bacteria population, the evolutionary transient dynamics which lead to the emergence of resistance, and the epidemic dynamics of the host population. Moreover, we precisely analyze the model proposed by particularly performing the uniform persistence and global asymptotic results. Finally, we discuss the impact of the treatment rate of the host population in controlling both the epidemic outbreak and the average level of resistance, either if the within-host scale therapy is a success or failure. We also explore how transitions between infected populations (treated and untreated) can impact the average level of resistance, particularly in a scenario where the treatment is successful at the within-host scale.

Keywords: Antimicrobial resistance, Evolutionary dynamics, Nested models, Non-linear dynamical system

Introduction

Antimicrobial resistance (AMR) is one of the major challenges we face in the modern area (Larsson and Flach 2022). An antimicrobial substance is a chemical agent interacting with the physiology of a bacterial cell. The antimicrobial activity on a given bacterium’s (S) is an increasing function of its concentration in the medium (C), such that S0=0 and SCSsat as CCsat, where Ssat and Csat are saturating constants. This intuitive approach implies that there exists C in 0,Csat such that SC is equal to the intrinsic rate of increase and reverses the growth of a bacterial population. Such a threshold concentration at which a bacterial population does not grow (at least in in vitro) is called the Minimum Inhibitory Concentration (MIC). The level of resistance to a given antimicrobial is then a continuous trait by nature referred to as antimicrobial quantitative resistance (qAMR), at least at the population level, and qAMR is key to better understanding the evolutionary dynamics of AMR (Djidjou-Demasse et al. 2023). Here, we introduce a quantitative descriptor xR—a label of the bacterial strain with resistance level x– describing the level of resistance. Most of the modelling approaches devoted to AMR tackling the case of qualitative (or “binary”) resistance are generally based on the dynamical interaction between two parasite strains resulting in a discrete and finite formulation of MICs (Blanquart 2019). This analysis ignores the evolutionary short-term transient dynamics which lead to the emergence of resistance (e.g., Lipsitch and Levin 1997; Kepler and Perelson 1998; Day and Read 2016; Djidjou-Demasse et al. 2021; Tazzyman and Bonhoeffer 2014; Millan et al. 2014; D’Agata et al. 2008).

Here, we proposed a nested (or embedded) model explicitly linking the within- and between-host evolutionary dynamics. Such a nested structure is particularly important because, over the past few decades, it is clear that ecological and evolutionary dynamics are influenced by processes operating across scales (Elderd et al. 2022). Very few studies considered the continuous nature of AMR in the context of this work (e.g., Djidjou-Demasse et al. 2023), and few studies have implemented a nested model in this context so far (e.g., Beardmore et al. 2017; Shen et al. 2019). The bacterial population is assumed to be phenotypically (and genetically) diverse through the level of antimicrobial resistance x. This quantitative trait affects different components of the bacterial population life cycle, such as growth and death rates. In addition to those effects on the death and birth rates, bacterial population resistance level also mitigates the antimicrobial efficiency with respect to that population. From a theoretical point of view, properties of the within-host model proposed here are based on previous analytical quantitative genetics results developed in Djidjou-Demasse et al. (2017) and Burie et al. (2020).

An integro-differential equation is used to model the within-host dynamics of the bacterial population. Such a within model formulation is previously proposed in Djidjou-Demasse et al. (2023). Each host individual is classified as either a treated host, labeled as T, or an untreated host, labeled as U. The model tracks the dynamics of a bacterial population within a treated host (bT) or an untreated host (bU). At time τ, the bacteria density with resistance level yR within a treated and untreated host is quantified by biT(τ,y) and biU(τ,y) respectively. The subscript “i” (with iI={1,2,,n}) represents an individual immune system and then allows taking into account the immune system heterogeneity in the host population. A bacteria with resistance level y generate offspring with resistance level x at a per-capita rate J(x-y)p(y)biϑ(τ,y), where p(y) is the bacterial intrinsic growth rate, and J(x-y) is the probability for a bacterial population with resistance level y to mutate towards a level x during the reproduction process. Therefore, the total number of bacteria produced at time τ with a resistance level x is quantified by 1+Rbiϑ(τ,x)dx-κRJ(x-y)p(y)biϑ(τ,y)dy, where κ is a positive parameter. The parameter κ>0 is introduced to impose the bacterial population homeostasis. The within-host model reads as

τbiϑ(τ,x)=11+Biϑ(τ)κRJ(x-y)p(y)biϑ(τ,y)dy-ξiϑ(x)biϑ(τ,x),biϑ(0,·)=b0(·), 1.1

where Biϑ(τ)=Rbiϑ(τ,x)dx, is the total bacteria load. The term ξiϑ accounts for the individual clearance of bacterial cells with resistance level x, either by the immune system (μi) or by the efficiency of antimicrobial pressure (k). Thus, ξiU(x)=μi(x), for untreated host, and ξiT(x)=μi(x)+k(x), for treated host. Here, it is assumed that bacteria are subject to a biocidal antimicrobial pressure, .ie. killing and not diminishing the birth rate of bacteria. Note that, the within-host model (1.1) allows to follow evolutionary parameters such as the average level of resistance for treated (x¯iT(τ)) and untreated (x¯iU(τ)) individuals τ-time post infection given by

x¯iϑ(τ)=Rxbiϑ(τ,x)Biϑ(τ)dx,ϑ{T,U}.

At the between-host scale, the host population is subdivided into three states. At any time t, an individual—with the immune system’s response level iI—can be susceptible to the infection Si(t) or infected Iiϑ(t,τ,x¯iϑ(τ)), ϑ={T,U}. The variables τ and x¯iϑ(τ) respectively represent the time post-infection and the average resistance level of the infected host. It is important to clearly understand the meaning of infected individuals Iϑ(t,τ,x¯iϑ(τ)). Indeed, each infected individuals is potentially infected with multiple bacteria strains with variable frequencies and resistance levels. Therefore, x¯iϑ(τ) represent the individual resistance level quantified by the within-host dynamics through the above formula. However, for simplicity, and without loss of generality, we will note Iiϑ(t,τ,x¯iϑ(τ))Iiϑ(t,τ) for ϑ{T,U} and iI.

Individual transmission and loss rates at the between-host scale, τ-time since infection, βiϑ(τ) and αiϑ(τ), are linked to the within-host dynamics at time τ. As an example, these parameters can be represented as Holling functions of type II (or similarly the Beddington-DeAngelis functional response) such that, for all ϑ{T,U},

βiϑ(τ)=β0Biϑ(τ)r0+Biϑ(τ)andαiϑ(τ)=α0Biϑ(τ)r0+Biϑ(τ)+γϑ(τ), 1.2

where β0 and α0 are scaling constants, and r0 is the half-saturation constant for the total bacterial load Biϑ. Note that, for the loss rate of infected individuals αiϑ, the term α0Biϑ/(r0+Biϑ) represents the loss due to the disease induced mortality while γϑ is the loss due to recovery. We can assume that the function γϑ is of the form

γiϑ(τ)=0,ifBiϑ(τ)>Bmin,1,ifBiϑ(τ)Bmin, 1.3

where Bmin is the threshold below which the infection becomes undetectable such that the infected individuals is considered as recovered.

The force of infection induced by infected individuals at time t is then given by

λ(t)=iI0[βiT(τ)IiT(t,τ)+βiU(τ)IiU(t,τ)]dτ.

The nested model proposed here then makes it possible to simultaneously track the epidemiological dynamics of the host population as well as evolutionary quantities such as the average level of resistance at both the within- and between-host scales. Such an approach is original and to our knowledge, no study has considered nested models for the evolutionary dynamics of AMR, viewed as a continuous quantitative trait.

The dynamics of newly infected individuals (i.e. τ=0) in each group (treated or untreated) is thus defined by (for ϑ{T,U})

Iiϑ(t,τ=0)=qiϑλ(t)Si(t), 1.4

where qiT(0,1) is the treatment rate in the host population and qiU=1-qiT. During their infection, treated individuals can stop the treatment at rate ωUT(τ), and untreated infections can join the treated group at rate ωTU(τ). The loss rate of infected individuals τ-time post infection occurs at rate αiϑ(τ). Susceptible individuals are recruited at a constant rate Λi and the natural death rate of the host population is μh. The between-host model then reads

S˙i(t)=Λi-Si(t)λ(t)-μhSi(t),t+τIiT(t,τ)=-αiT(τ)+ωUT(τ)+μhIiT(t,τ)+ωTU(τ)IiU(t,τ),t+τIiU(t,τ)=-αiU(τ)+ωTU(τ)+μhIiU(t,τ)+ωUT(τ)IiT(t,τ),Si(0)=Si,0,Iiϑ(0,τ)=Ii,0ϑ(τ). 1.5

Finally, the nested within-host (1.1) and between-host model (1.4)–(1.5) are summarised by Fig. 1. The main variables and parameters are listed in Table 1. We emphasize that the dynamical properties (that we will recall later) of within-host model (1.1) are precisely analyzed in Djidjou-Demasse et al. (2023). Therefore, our main objective here is devoted to the analysis of the nested model (1.1)–(1.5).

Fig. 1.

Fig. 1

Flow diagram of the nested model. Within-host model: the number of bacteria produced at time τ with resistance level x is 11+-+biϑ(τ,x)dxκ-+J(x-y)p(y)biϑ(τ,y)dy, where J(x-y) is the probability for a bacterial with resistance level yR to mutate towards a level xR and p(y) is the bacterial intrinsic growth rate. Bacterial cells with resistance level x, within an individual with immune system level i are cleared either by the immune system at rate μi(x) or by the antimicrobial efficiency at rate k(x). Between-host model: susceptible individuals are recruited at a constant rate Λi. IiT(t,τ) and IiU(t,τ) are respectively treated and untreated infected individuals at time t, which are infected since time τ. The force of infection in the whole population at time t is λ(t)=iI0[βiT(τ)IiT(t,τ)+βiU(τ)IiU(t,τ)]dτ, with βiϑ(τ) the disease transmission rate of an infected individual τ-time post infection. At the time t, new infections occur at rate λ(t)Si(t), and are either treated with a probability qiT or untreated with a probability qiU=1-qiT. The natural death rate of individuals is μh. If infected since time τ, the loss rate is αiϑ(τ). Untreated individuals, and infected since time τ start the treatment at rate ωTU(τ) while treated individuals stop the treatment at rate ωUT(τ)

Table 1.

Within- and between-host model variables and parameters

Notations
t Time
τ Time since infection (infection age)
x Resistance level
i host immune response level
Model state variables
Within-host
   biT(τ,x) Bacteria density with resistance level x at time τ in a treated host
   biU(τ,x) Bacteria density with resistance level x at time τ in an untreated host
Between-host
   Si(t) Susceptible humans at time t
   IiT(t,τ) Treated infected humans, infected since time τ
   IiU(t,τ) Untreated infected humans, infected since time τ
Model parameters
Within-host
   κ Limitation on bacterial growth factor
   J(x-y) Mutation probability from resistance level x to y per cell division
   p(x) Intrinsic growth rate of bacterial population with resistance level x
   k(x) Killing rate of bacterial population with resistance level x due to drug
   μi(x) Individual clearance rate of the bacterial cell with resistance level x due to the immune response
   ξiU(x)=μi(x) Bacteria clearance rate with resistance level x in untreated hosts
   ξiT(x)=k(x)+μi(x) Bacteria clearance rate with resistance level x in treated hosts due to the immune response and drug
Between-host
   βiϑ(τ), ϑ{T,U} Transmission rate of infected individuals
   αiϑ(τ), ϑ{T,U} Loss rate of infected individuals
   ωUT(τ) Rate of treatment cessation
   ωTU(τ) Rate of treatment start
   qiT Proportion of treated hosts
   qiU=1-qiT Proportion of untreated hosts
   Λi Recruitment rate of susceptibles
   μh Human natural death rate

The rest of this work is organized as follows. In Sect. 2, we state the main results of the nested model that are obtained in this work. These include the existence of the globally defined non-negative semiflow and the existence of the unique positive equilibrium for the within-host model, and global threshold analysis results for the between-host model. The model’s typical dynamics are provided in Sect. 3. This includes the within- and between-host models parameterization and the characterization of the evolutionary parameters such as the average levels of resistance. In Sect. 4, we delve into the effects of various parameters on the equilibrium structure of the host population, along with addressing the parameterization issue within nested models. Section 5 focuses on providing preliminary results. Specifically, it addresses the existence and uniqueness of solutions, derivation of the basic reproduction number, and the existence of a unique endemic equilibrium for System (1.4)–(1.5). Finally, Sect. 6 is devoted to the proof of the global asymptotic results.

Main results

This section is devoted to the main results of the nested model (1.1)–(1.5). Such results include the existence of the unique maximal bounded semiflow, and a precise description of the unique positive equilibrium of Model (1.4)–(1.5). By providing global stability results, we will also conduct a precise threshold analysis of the between-host model (1.4)–(1.5).

First of all, for biological feasibility of the nested model (1.1)–(1.5), we make use of the following assumptions. More precisely, the within-host model (1.1) is formulated based on the following assumption

Assumption 2.1

  1. Functions μi, k, ξiϑ, and p are always positive on R, with ϑ{T,U}. Furthermore, p is a bounded function on R and κ>0. Finally, the function pξiϑ is continuous on R and satisfies pξiϑ>0 and lim|x|pξiϑ(x)=0.

  2. The mutation kernel J is bounded and integrable on R+, positive almost everywhere, and satisfies R+J(x)dx>0, J(-x)=J(x), for all x.

  3. The mutation kernel J decays rather rapidly towards infinity in the sense that J(x)=O1|x| as |x|. In other words, lim|x||x|nJ(x)=0, for all nN.

Furthermore, the between-host model’s parameters satisfy the following assumption.

Assumption 2.2

  1. Recruitment rate Λi (iI) and natural death rate μh are positive constants.

  2. The treatment rates qiϑ (iI, ϑ{T,U}) are positive constants.

  3. The rates ωUT, ωTU belongs to L(R+), with respective essential upper bounds ω¯UT, ω¯TU and positive essential lower bounds ω_UT, ω_TU.

  4. Parameters βiϑ and αiϑ (iI, ϑ{T,U}) are such that βiϑ,αiϑL(R+).

  5. The transmission rates βiϑ(·),s are Lipschitz continuous almost everywhere on R+.

Summary key findings on the within-host dynamics

The dynamical properties of the within-host model (1.1) have been precisely investigated in Djidjou-Demasse et al. (2023). The first result of Model (1.1) is about the existence of the unique maximal bounded semiflow. Such a result reads as,

Theorem 2.3

Let Assumption 2.1 be satisfied. Let bi0ϑL+1. Then,

  1. There exists a unique global solution v(·,bi0ϑ):[0,)L+1(R) of (1.1) with v(0,bi0ϑ)=bi0ϑ and v(τ,bi0ϑ)=biϑ(τ,·) for all τ>0.

  2. The semi-flow defined by {v(τ,bi0ϑ)}τ is bounded dissipative and asymptotically smooth, and hence, its admits a global attractor in L+1(R).

  3. The semi-flow {v(τ,bi0ϑ)}τ is such that for any bi0ϑL+1(R)\{0}, biϑ(τ,x)>0, for all τ>0, xR.

The basic reproduction number Ni0ϑ—defined as the expected number of bacteria arising from one bacterium in a bacteria-free environment—of the bacteria population with resistance level x, within an individual with immune system level i, is calculated as

Ni0ϑ(x)=p(x)ξiϑ(x),forϑ{U,T}. 2.1

Next, a non-trivial equilibrium of Model (1.1) is strongly related to the principal eigenpair of the below linear integral operator Hiϑ defined on Lp(R) (for any p1), by

Hiϑ[viϑ](x)=Ni0ϑ(x)RJ(x-y)Ni0ϑ(y)viϑ(y)dy. 2.2

We then have the following result.

Theorem 2.4

Let r(Hiϑ), the spectral radius of the operator Hiϑ, and ϕ>0 the associated eigenfunction normalized such that ||ϕ||L1=1.

  1. When r(Hiϑ)1, the bacteria-free equilibrium Fi0ϑ is the unique equilibrium of Model (1.1).

  2. When r(Hiϑ)>1, in addition to Fi0ϑ, Model (1.1) has a unique equilibrium F¯iϑ>0 such that
    F¯iϑ(x)=r(Hiϑ)1κ-1Rϕpξiϑdyϕ(x)p(x)ξiϑ(x). 2.3
    Furthermore, the semi-flow {v(τ,bi0ϑ)}τ is uniformly persistent, that is, there exists a constant η such that for any bi0ϑL+1(R)\{0}, the unique solution v(τ,bi0ϑ)=biϑ(τ,·) of Model (1.1) with initial data bi0ϑ satisfies limτinfbiϑ(τ,·)L1>η.
  3. The bacteria-free equilibrium Fi0ϑ of Model (1.1) is asymptotically stable if r(Hiϑ)<1 and unstable if r(Hiϑ)>1.

  4. When r(Hiϑ)<1, the bacteria-free equilibrium Fi0ϑ is globally asymptotically stable in L+1(R), that is, for any solution biϑ(τ,·) with initial bi0ϑL+1(R)\{0}, we have biϑ(τ,·)0 in L+1(R), as τ.

We recall that the within-host model (1.1) is precisely analyzed in Djidjou-Demasse et al. (2023). We then refer to Appendices F–I in Djidjou-Demasse et al. (2023) for the detailed proof of Theorems 2.3 and 2.4. Note that the linear operator Hiϑ naturally emerges when characterizing the positive equilibrium of the within-host model (1.1) (Djidjou-Demasse et al. 2023).

Furthermore, the estimate (2.3) gives that the endemic equilibrium F¯iϑ of the within-host model (1.1) basically relied to the principal eigenfunction of the linear operator Hiϑ for any given probability kernel J satisfying Assumption 2.1. However, the profile of the endemic equilibrium F¯iϑ with respect to xR can be precisely described when the mutation kernel J depends on a small positive parameter (let say ε<<1) with the scaling form

Jε(x)=ε-1Jε-1x, 2.4

where ε>0 represents the mutation variance in the phenotypic space. More precisely, when ε>0 is small, then the endemic equilibrium F¯iϑ concentrates on the set Siϑ defined by

Siϑ=xR:Ni0ϑ(x)=Ni0ϑ.

The set Siϑ is referred to as the set of Evolutionary Attractors (or dominant strains) of the within-host model in the classical adaptive dynamics theory ( e.g., Geritz et al. 1997; Metz et al. 1996). Furthermore, when the function Ni0ϑ is at least of class C1, with a finite number of maximum, it is shown in Djidjou-Demasse et al. (2017) that these dominant strains coincide with the set Siϑ. Denoting by Hi,εϑ, the operator Hiϑ—by replacing the kernel J by Jε—by results in Djidjou-Demasse et al. (2017) (Theorem 2.2), the spectral radius r(Hi,εϑ) of Hi,εϑ satisfied, for ε sufficiently small

r(Hi,εϑ)=Ni0ϑ(x)2+O(ε),for allxSiϑ.

By the above estimate, signrHi,εϑ-1=signNi0ϑ(x)-1, for all x and ε sufficiently small. Furthermore, if ε1, Siϑ={xi} and Ni0ϑ(xi)>1, then the unique positive stationary state F¯iϑF¯i,εϑ, given by (2.3), of the within-host model (1.1) is concentrated around the evolutionary attractor xi in the space of resistance level R. In other words, xi is the average bacterial resistance level at the within-host scale equilibrium and we have limε0Ru(x)F¯i,εϑ(x)x=uxi for any continuous function uCR. We refer to Theorem 2.3 in Djidjou-Demasse et al. (2017) for such a concentration phenomenon.

Key findings of the nested within- and between-host dynamics

At the between-host scale, by setting S(t)=(Si(t))iI, Ii(t,τ)=(IiT(t,τ),IiU(t,τ)), I(t,τ)=(Ii(t,τ))iI, αi(τ)=diag(αiT(τ),αiU(τ)), βi(τ)=(βiT(τ),βiU(τ)), qi=(qiT,qiU), ω(τ)=0ωTU(τ)ωUT(τ)0, and e=0110, System (1.4)–(1.5) rewrites into the following compact form,

S˙(t)=Λ-S(t)λ(t)-μhS(t),I(t,τ=0)=λ(t)diag(S(t))q,t+τI(t,τ)=-diag(Φ(τ))+diag(ω(τ))I(t,τ), 2.5

where λ(t)=i0βi(τ),Ii(t,τ)dτ, Λ=(Λi)iI, q=(qi)iI, Φ(τ)=(Φi(τ))iI, with Φi(τ)=eω(τ)+αi(τ)+μh.

Using the next-generation operator approach (e.g., Diekmann et al. 1990; Inaba 2012), the basic reproduction number R0i of the whole infected individuals of group i, is given by

R0i=ΛiμhkI0βk(τ),Πk(τ,0)qkdτ=ΛiμhkIχk,

where

χk=0βk(τ),Πk(τ,0)qkdτ,

and where Πi(τ2,τ1), 0τ1τ2<, is the evolutionary system generated by the linear operator -Φi(τ)+ω(τ); see Remark 2.5 for some details on Πi. Moreover, the basic reproduction number R0 at the whole between-host scale is such that

R0=iIΛiμhχi=iIχikIχkOverall relative infectiousnessof individuals of groupi.×R0i. 2.6

We refer to Sect. 5.2 for details of the computation of R0i,s and R0.

Note that the parameter χk quantifies the overall infectiousness of the whole infected individuals of group kI. A more explicit expression of the infectiousness χk is difficult to obtain in general. However, one can go further steps in some particular configurations of the treatment status transition rates ω(τ). Indeed, assume that we can find τ0>0 and τ1>0 such that

ωTU(τ)=0,forτ<τ0,ω¯1,forτ0<τ<τ0+τ1,andωUT(τ)=0,for,τ<τ0+τ1. 2.7

In the above scenario, the regimen (0,τ0)-post infection may corresponds to the initial phase where each infections, either treated or untreated, remain to their initial treatment status. The second regimen (τ0,τ0+τ1)-post infection may corresponds to the phase during which previously untreated infections becomes treated while treated infections remain to their initial status. In such a configuration, we have (see Sect. 5.2 for details)

χk=qkTχkT+qkUχkU+Oe-ck(τ0+τ1),

with ck=μh+infταkU(τ)+infταkT(τ),

χkT=0τ0+τ1βkT(τ)Γ0k,T(τ)dτ+Γ0k,T(τ0)ω¯1τ0τ0+τ1βkT(τ)τ0τΓ1k,T(τ)Γ1k,T(η)Γ1k,U(η)dηdτ,χkU=0τ0βkU(τ)Γ0k,U(τ)dτ+Γ0k,U(τ0)τ0τ0+τ1βkU(τ)Γ1k,U(τ)Γ1k,U(τ0)dτ,

and

Γ0k,ϑ(τ)=e-μhτe-0ταkϑ(σ)dσ,ϑ{T,U},Γ1k,U(τ)=e-(ω¯1+μh)τe-0ταkU(σ)dσ,Γ1k,T(τ)=e-μhτe-0ταkT(σ)dσ.

Note that parameters Γ0k,ϑ,s and Γ1k,ϑ,s are survival probabilities during phases (0,τ0) and (τ0,τ0+τ1)-post infection of infected individuals of group kI, treated (ϑ=T) or untreated (ϑ=U).

Remark 2.5

Let Πi(τ2,τ1), 0τ1τ2<, the evolutionary system generated by the linear operator Ai(τ):=-Φi(τ)+ω(τ). It means that Πi is generated from the following evolutionary system

t+τIi(t,τ)=Ai(τ)Ii(t,τ). 2.8

If, for example, the linear operator Ai is diagonal, we have

Ii(t,τ)=e0τAi(η)dηIi(t-τ,0)=Πi(τ,0)Ii(t-τ,0);t>τeτ-tτAi(η)dηIi(0,τ-t)=Πi(τ,τ-t)Ii(0,τ-t);t<τ.

In such a configuration we explicitly have Πi(τ2,τ1)=eτ1τ2Ai(η)dη. However, obtaining an explicit expression for Πi may not always be straightforward or possible in general. A naive approach would be to solve problem (2.8) as above, but it is well known that such an exponent formula does not give a solution to the problem at hand.

In addition to the disease-free equilibrium—the DFE—E0=(S0,0L1((0,),R2n)), with S0=Λi/μhiI, which is always an equilibrium of Model (1.4)–(1.5), this model also exhibits an endemic equilibrium given by the following result

Theorem 2.6

Let Assumptions 2.1 and 2.2 hold. If R0>1, then system (1.4)–(1.5) has a unique endemic equilibrium E=(S,I(τ)), such that iI,

Si=Si0R0andIi(τ)=λSiΠi(τ,0)qi,

where λ=μh(R0-1).

Therefore, the threshold dynamics of Model (1.4)–(1.5) is summarized as follows

Theorem 2.7

Let Assumptions 2.1 and 2.2 hold. Then,

  • (i)
    If R01 or iI0βi(τ),Ii0(τ)dτ=0, then the disease-free equilibrium E0=(S0,0L1((0,),R2n))t of system (1.4)-(1.5) is globally asymptotically stable in the sens that
    limtSi(t),IiT(t,·),IiU(t,·)iI=E0,
    where the above convergence holds for the topology of Rn×L1((0,),R2n).
  • (ii)
    If R0>1 and iI0βi(τ),Ii0(τ)dτ>0, then the endemic equilibrium E of system (1.4)-(1.5) is globally asymptotically stable, that is,
    limtSi(t),IiT(t,·),IiU(t,·)iI=E,
    for the topology of Rn×L1((0,),R2n).

Numerical illustrations

Here, we present a series of numerical simulations employing semi-explicit finite difference numerical schemes. We refer to Djidjou-Demasse (2021) for an example of a code repository within the context of the model proposed here. We illustrate an example of typical dynamics that can be simulated by the nested model (1.1)–(1.5). The model simultaneously captures the outbreak dynamics as well as the evolutionary dynamics of the average resistance level within the host population. The within-host model parameters are basically the same as in Djidjou-Demasse et al. (2023). Intuitively there exist two threshold levels, assumed here 0 and 1 (called reference “sensitive” and “resistant” strains) such that, a strain with resistance level x can be classically referred to as “sensitive”, “intermediate”, or “resistant” depending on whether x<0, 0<x<1, or x>1. For sake of simplicity, we assume that the host population is homogeneous in terms of immune system level, i.e., card(I)=1. For all illustrative scenarios, we will have R0>1 such that the disease is persistent at the between-host scale (Theorem 2.7). The probability density function at the within-host scale (JJε) is assumed of type (2.4). Specifically, we define Jε as a Gaussian distribution Jε(x)=1ε2πe-12xε2, where ε>0 represents a small parameter that signifies the mutation variance within the phenotypic space.

Within-host parameterization The antimicrobial killing rate function k(·) is a decreasing function with respect to the resistance level x such that, k(x)=k0k1k0x, where k0 and k1 are the antimicrobial activity undergone by the reference sensitive and resistant strains. Moreover, knowing p0 and p1, respectively the intrinsic growth rate of reference strains 0 and 1, a suitable expression for function of p is p(x)=pm1+pm-p0p0p0p1·pm-p1pm-p0x-1, where pm is the upper bound of the intrinsic growth rate p and 0<p1<p0<pm. The qualitative behaviour of functions k and p can be found in (Djidjou-Demasse et al. 2023, Fig. 2). We assume that the clearance rate of the bacteria cell due to the immune response, μ(·), is a constant function given by μ(x)=μ. Furthermore, the average fitness cost-benefit ratio of resistance within a bacterial population can be expressed as cb=log(Δ)log(1+δ). Here, Δ=(pm-p1)/p1(pm-p0)/p0>1 quantifies the relative cost of resistance, while δ=k0-k1k1>0 measures the fitness advantage of the reference resistant strain (see Djidjou-Demasse et al. 2023 for details).

Between-host parameterization Parameters βiϑ and αiϑ are defined using Holling type functional responses introduce by (1.2). For all simulations, the threshold Bmin, introduced by (1.3), below which the infection becomes undetectable such that the infected individuals is considered as recovered is fixed as Bmin=10-3B0, with B0=B(0) the initial total bacteria load. The total bacteria load (Biϑ(τ)) and the recovery probability (1-exp(-0τγiϑ(s)ds), τ-time post infection, are illustrated in Fig. 2.

Fig. 2.

Fig. 2

Illustration of the within- and between-host parameterization of infected individuals, either untreated—U, under treatment failure—TF, or under treatment with success—TS. A The total bacteria load Bϑ(τ). B The recovery probability 1-exp(-0τγϑ(s)ds)

Furthermore, an untreated infected individual joins the treated compartment when her total bacteria load is above a threshold (1+θ)B0, with θ0. Therefore, the influx rate from untreated to treated is assumed to be a function with respect to time τ and is defined as follows

ωTU(τ)=0,ifBU(τ)(1+θ)B0,1,ifBU(τ)>(1+θ)B0. 3.1

Similarly, we assume that an infected individual under treatment can drop down such a treatment when the bacteria load reach the same range as before the treatment. Therefore, the influx rate from treated to untreated is given by

ωUT(τ)=0,ifBT(τ)B0,1,ifBT(τ)>B0. 3.2

Initial conditions and model outputs The initial bacterial population b0ϑ(x) is assumed to be composed by a sensitive bacterial population with average resistance level x=0. Hence, we set b0ϑ(x)=m0×N(0,σ0,x), where N(0,σ0,x) stands for the normalized density function of the Gaussian distribution at x with mean 0 and variance σ02. This means that the initial bacterial population is mostly composed of the reference “sensitive” strain. At the between-host scale, the initial condition of the epidemiological model is taken such that the susceptible population starts close to its disease-free equilibrium. More precisely, assuming an initial infection prevalence denoted as Prev=10%, we derive the initial susceptible population as S0=(1-Prev)Λμh, along with the initial distribution of infectives which consists of I0T(τ)=0 and I0U(τ)=PrevΛμh×L(τ) for all τ0. Here, L(τ)=10ln(10)×10-10τ, and it is important to note that L symbolizes the arbitrary initial distribution of individuals who have been infected since time τ. This distribution is scaled so that RL(τ)dτ=1.

The average level of resistance at within-host scale (η(t)) of the host population at time t is such that

η(t)=0x¯T(τ)IT(t,τ)I(t)+x¯U(τ)IU(t,τ)I(t)dτ,

where x¯ϑ,s are the individual average level of resistance and I(t)=0(IT(t,τ)+IU(t,τ))dτ.

Simulated scenarios Two simulated scenarios are considered, the first when the treatment is successful at the within-host level, and the second when the treatment failed at the within-host level. For all our simulated scenarios, the infection is assumed here to be always successful for untreated individuals, i.e., the immune system alone is no more enough to control the infection such that maxxRN0U(x)>1, leading to the bacterial persistence for untreated infections.

Our first scenario is for the case where the treatment is successful at the within-host level, i.e., the basic reproduction number of treated individuals N0T is such that maxxRN0T(x)<1 (Fig. 3C). In such a situation, the bacterial load is under control in the relatively short term for treated individuals (Fig. 3A), while it remains persistent for untreated individuals (Fig. 3B). At the between-host scale, the treatment rate have a strong effect on the epidemic outbreak (Fig. 3F–H). More precisely, increasing the treatment rate qT in the host population strongly reduce the overall epidemic size (Fig. 3F–H), with R0=4.7815, 2.6750 and 0.5685, respectively. Furthermore, the average resistance level in the host population rapidly reach an equilibrium for which the level of resistance is moderately high compared to the initial resistance level of the host population (Fig. 3E).

Fig. 3.

Fig. 3

Dynamics of the nested model when the therapeutic treatment is successful. A, B The bacterial population (treated and untreated) with respect to time τ and resistance level x. C The basic reproduction numbers at the within-host level N0T(x) and N0U(x), with and without drug respectively. D The average resistance level x¯ϑ(τ), ϑ{T,U} of the infected host. E The average resistance level η(t) of the host population. FH The between-host dynamics with respect to the treatment rate where R0=4.7815, 2.6750 and 0.5685, respectively. Parameter values are (σ0,m0,k0,p1/p0,k1/k0,μ)=(0.05,0.05,20,0.5,0.3,0.8598) and Λ=5e4, μh=5.2675e-2, β0=1.2e-3, α0=7.5e-2, θ=1, r0=9e3

In the second scenario, the treatment is assumed unsuccessful at the within-host level, i.e., the basic reproduction number of treated individuals N0T is such that maxxRN0T(x)>1 (Fig. 4C). In such a situation, the bacterial load remains persistent for both treated and untreated infections (Fig. 4A,B). Indeed, while we can observe an apparent decreasing of the bacteria load for some period of time for treated infections (Fig. 4A), at the end, we have the re-emergence of the bacteria population at within-host scale. Such a transient dynamics is explain by the fact that the initiation of treatment modifies the fitness landscape by shifting the maximum point of the within-host basic reproduction number N0T to the point x=x>0 (Fig. 4C). In contrast to the treatment success scenario (Fig. 3), increasing the treatment rate qT in the host population have marginal effect in controlling the epidemic outbreak (Fig. 4F–H). In fact, with qT=0.1, 0.5 and 0.9, the outbreak remains persistent with R0=10.9451, 9.5855 and 8.2259, respectively. Significantly, it is worth noting that although the average resistance level in the host population continues to rise with the treatment rate qT at equilibrium (Fig. 4E), there is a substantial increase in the range of resistance levels compared to the initial resistance level. This stands in contrast to the treatment success scenario (Fig. 3E), where the range of resistance levels remained relatively low. Additionally, in the treatment success scenario (Fig. 3E), the average resistance level in the host population quickly reaches equilibrium. However, in the treatment failure scenario (Fig. 4E), there is a comparatively longer transient period before the average resistance level in the host population reaches equilibrium. Overall, during the transient regimen, there is an initial subsequent increase in the average level of resistance to significantly higher levels, followed by a small decrease (Fig. 4E). This behavior is mostly attributed to the alteration of the fitness landscape caused by the treatment (Fig. 4C).

Fig. 4.

Fig. 4

Dynamics of the nested model when the therapeutic treatment failed. A, B The bacterial population (treated and untreated) with respect to time τ and resistance level x. C The basic reproduction numbers at the within-host level N0T(x) and N0U(x), with and without drug respectively. D The average resistance level x¯ϑ(τ), ϑ{T,U} of the infected host. E The average resistance level η(t) of the host population. FH The between-host dynamics with respect to the treatment rate where R0=10.9451, 9.5855 and 8.2259, respectively. Parameter values are (σ0,m0,k0,p1/p0,k1/k0,μ)=(0.05,0.05,3,0.5,0.01,0.8598) and Λ=5e4, μh=5.2675e-2, β0=1.2e-3, α0=7.5e-2, θ=1, r0=9e3

Discussion

Optimizing the treatment rate in the host population is key to controlling both the epidemic outbreak and the average level of resistance Increasing the treatment rate in the host population contributes to reducing the epidemic size at the between-host scale, although the effect is quite marginal in the treatment failure scenario at the within-host scale (Figs. 3F–H, 4F–H). This can be primarily attributed to the fact that, even in cases of treatment failure, the within-host infection remains controlled for a certain period (Fig. 4A). However, the subsequent phase is characterized by an increase in bacterial population density (Fig. 4A). Conversely, raising the treatment rate within the host population leads to an elevation in the average resistance level of that population, regardless of the treatment scenario at the within-host scale (Figs. 3E, 4E). However, this increase in the average resistance level is particularly pronounced in the case of treatment failure (Fig. 4E), in comparison to the case of treatment success (Fig. 3E). In the case of treatment success, the resistance level remains highly similar to that of the initial bacterial population. Consequently, depending on the treatment regimen implemented within the host population, it becomes essential and intriguing to determine an optimal treatment rate to effectively manage both the outbreak and the average level of antimicrobial resistance.

The delay for treating infections can impact the epidemic outbreak as well as average level of resistance Assume the scenario where the treatment is successful at the within-host scale (Fig. 3A) and at least 50% of infected individuals are under treatment (qT0.5). Two configurations are introduced. In the first configuration, untreated individuals begin treatment at a rate ωTU [defined by (3.1)] when their total bacterial load reaches a threshold value of 1.5×B0, i.e., θ=0.5. In the second configuration, untreated individuals start treatment at a rate ωTU when their total bacterial load reaches a threshold value of 2×B0, i.e., θ=1. The case of θ=0.5 indicates a situation where the delay before initiating the treatment is very short. On the other hand, in the case of θ=1, the delay before starting the treatment is relatively more significant (Fig. 2A). In general, the average resistance level in the host population decreases as the delay before initiating treatment increases (Fig. 5A, B). One possible explanation is that untreated infected individuals do not significantly contribute to the increase in the average resistance level within the host population. Nevertheless, although early treatment effectively controls the epidemic outbreak for both moderate and high treatment rates (Fig. 5C, D)—with R0=0.3922 and 0.1120, respectively—delaying the treatment of infected individuals leads to an epidemic that is out of control, except in cases where the treatment rate is exceptionally high (Fig. 5E, F)—with R0=2.6750 and 0.5685, respectively.

Fig. 5.

Fig. 5

The effect of delay for treating infections on the epidemic outbreak and the average level of resistance when the therapeutic treatment is successful. Line 1 The treatment rate qT=0.5. A The average resistance level η of the host population for θ{0.5,1}. C, E The between-host dynamics for θ=0.5 and 1, where R0=0.3922 and 2.6750 respectively. Line 2 As in line 1 for qT=0.9 where R0=0.1120 and 0.5685 respectively. Other parameter values are (σ0,m0,k0,p1/p0,k1/k0,μ)=(0.05,0.05,20,0.5,0.3,0.8598) and Λ=5e4, μh=5.2675e-2, β0=1.2e-3, α0=7.5e-2, r0=9e3

Nested models parameterization issue Explicitly connecting the within- to and between-host scales are crucial to gain a more realistic picture aiming to integrate into the same modelling framework the epidemic dynamics and the evolutionary dynamics of antimicrobial resistance. However, such an approach requires making assumptions about the parameters at the within-host scale that are equally unknown as the parameters at the between-host scale and hence leading to uncertainty about the appropriate parametrization (Uecker and Bonhoeffer 2021). Such uncertainty is amplified by our approach where the level of resistance is considered a continuous quantitative trait, compared to the classical qualitative (or “binary”) approaches (Djidjou-Demasse et al. 2023).

Time-scale separation hypothesis Within the context of nesting within- and between-host scales, for the modelling of the epidemiology and evolution of pathogens, some studies assumed that the epidemiological and evolutionary time scales are distinct, i.e., the within-host dynamic is fast relative to the between-host dynamic such that the within-host model remains at equilibrium, e.g. (Gilchrist and Coombs 2006; Xue and Bloom 2020; Almocera et al. 2018; Boldin and Diekmann 2008; Coombs et al. 2007; André and Gandon 2006). Although such an assumption on the within-host equilibrium dynamic’ might be appropriate for chronic infections, it leads to a population-scale model that does not explicitly account for the individual time-dependent infectiousness dynamics (Hart et al. 2020). Furthermore, our illustrative examples (Figs. 3, 4) strongly highlight the infectiousness’ time-dependency of infected individuals and show that the between-host dynamic is not necessarily faster compared to the within-host dynamic.

Preliminaries and technical materials

We will go through details on the proof of our main results, namely Theorem 2.7. We will first discuss the existence of a positive global solution of the nested model. Next, we will give details on the derivation of the basic reproduction number of System (1.4)–(1.5) for individuals with an immune system of level i. We will also derive the existence of a unique endemic equilibrium of System (1.4)–(1.5) when R0>1, as well as the long-term persistence of the epidemic in such a case.

Existence of the semiflow

We establish the existence of a positive global solution of the system (2.5). We first formulate system (2.5) in an abstract Cauchy problem. For that, we introduce the Banach space X=Rn×R2n×L1((0,),R2n), endowed with the usual product norm ·X as well as its positive cone X+. Let A:D(A)XX be the linear operator defined by D(A)=Rn×{0R2n}×W1,1((0,),R2n) and

A(S,0R2n,I)=(-μhS,-I(0),-τI+-diag(Φ(τ))+diag(ω(τ))I). 5.1

Let us introduce the non-linear map F:D(A)¯X defined by

F(S,0R2n,I)=Λ-S(t)λ(t),λ(t)diag(S(t))q,diag(ω(·))I.

By identifying φ(t) together with (S(t),0L1,I(t,·))t and by setting φ0=(S0,0L1,I0(·))t the associated initial condition, system (2.5) becomes

dφ(t)dt=Aφ(t)+F(φ(t)),φ(0)=φ0. 5.2

By setting X0=D(A) and X0+=X0X+, the positivity and boundedness of the solutions of system (2.5) are provided by the following result.

Theorem 5.1

There exists a unique strongly continuous semiflow {Ψ(t,·):X0X0}t0 such that, for each φ0X0+, the map φC[0,),X0+ defined by φ=Ψ(·,φ0) is a mild solution of (5.2). That is, 0tφ(s)dsD(A) and φ(t)=φ0+A0tφ(s)ds+0tFφ(s)ds for all t0. Moreover, Ψ(t,·)t satisfies the following properties:

  1. Let Ψ(t,φ0)=S(t),0R2n,I(t,·)t, then the following Volterra formulation holds true for all iI
    Ii(t,τ)=Πi(τ,τ-t)Ii0(τ-t),iftτ,λ(t-τ)Si(t-τ)Πi(τ,0)qi,ift>τ, 5.3
    coupled with the Si(t) equation of (2.5), and where Πi(τ2,τ1), 0τ1τ2<, is the evolutionary system generated by the linear operator -Φi(τ)+ω(τ).
  2. For all φ0X0+, and for all t0, one has
    iISi(t)+0(IiT(t,τ)+IiU(t,τ))dτmaxΛ¯μh,N0, 5.4
    where Λ¯=iIΛi and N0=iISi0+0(Ii0T(τ)+Ii0U(τ))dτ. Furthermore, the subset of the phase space
    (S,I)Rn×L1((0,),R2n)|iISi(t)+0(IiT(t,τ)+IiU(t,τ))dτΛ¯μh,
    is positively invariant and attracts all nonnegative solutions.
  3. The semiflow Ψ(t,·)t generated by (2.5) is bounded dissipative, that is, there exists a bounded set BX0 such that for any bounded set UX0, we can find σ=σ(U,B)0 such that Ψ(t,U)B for tσ.

Proof

It is easy to check that the operator A is a Hille-Yosida operator. Then standard results apply to provide the existence and uniqueness of a mild solution to (2.5) [we refer to Magal and Ruan (2009) and Thieme (2011) for more details]. The Volterra formulation is also standard and we refer to Iannelli (1995), Webb (1985) for more details.

The Si equation of (2.5) gives S˙i(t)Λi-μhSi(t), that is

Si(t)maxΛiμh,Si0.

Next, for estimate (5.4), let φ0X0+, then adding up the Si,s equation together with the IiT,s and IiU,s equations of (1.5) yields for all iI

ddtSi(t)+R0(IiT(t,τ)+IiU(t,τ))dτΛi-μhSi(t)-μhR0(IiT(t,τ)+IiU(t,τ))dτ.

It comes

lim suptiISi(t)+R0(IiT(t,τ)+IiU(t,τ))dτΛ¯μh,

with Λ¯=iIΛi. From where one deduces estimate (5.4) and which ends item 2. of the theorem.

The bounded dissipativity of the semiflow Ψ(t,·)t is a direct consequence of estimate 2.

The following result is straightforward.

Lemma 5.2

Let 0τ1τ2<. By setting Πi(τ2,τ1)=(Πi,jk(τ2,τ1))k,j, with k,j{T,U}, we have

γ0e-(μh+α0)(τ2-τ1)Πi,jk(τ2,τ1)γ1e-μh(τ2-τ1), 5.5

where γ0, γ1>0 and α0=maxi(supαiT,supαiU).

Proof

Let 0τ1τ2<. Note that, for all τ0, we have -ω1ωUT(τ)ω1 and -ω2ωTU(τ)ω2, with ωi>0. Then, for all τ0,

-ω1-α0-μhω2ω1-ω2-α0-μh-Φi(τ)+ω(τ)-ω1-μhω2ω1-ω2-μh.

Therefore,

e-(μh+α0)(τ2-τ1)Γ(τ2,τ1)Πi(τ2,τ1)e-μh(τ2-τ1)Γ(τ2,τ1),

where

Γ(τ2,τ1)=ω1e-(τ2-τ1)(ω1+ω2)+ω2ω1+ω2ω2ω1+ω21-e-(τ2-τ1)(ω1+ω2)ω1ω1+ω21-e-(τ2-τ1)(ω1+ω2)ω2e-(τ2-τ1)(ω1+ω1)+ω2ω1+ω2.

Note that

Γ_Γ(τ2,τ1)Γ¯,whereΓ¯=1ω2ω1+ω2ω1ω1+ω21andΓ_=ω2ω1+ω200ω1ω1+ω2.

Thus,

e-(μh+α0)(τ2-τ1)Γ_Πi(τ2,τ1)e-μh(τ2-τ1)Γ¯,

from where inequality (5.5) follows, and this ends the proof of the lemma.

The basic reproduction number

In the absence of infection, that is I(t,τ)=0L1((0,),R2n), the system (2.5) has a disease-free equilibrium (DFE) given by E0=(S0,0L1((0,),R2n)), with S0=Λi/μhiI. Let Θi(t) be the number of new infections in the host population of group i at time t. Then in an initially infection-free population, by (2.5), we have

Θi(t)=λ(t)Si0=kI0βk(τ),Ik(t,τ)dτSi0,

where ·;· is the usual scalar product.

Linearizing the Volterra formulation (5.3) at the DFE, it comes

Ik(t,τ)=Πk(τ,τ-t)Ik0(τ-t),iftτ,λ(t-τ)Sk0Πk(τ,0)qk,ift>τ.

From where,

Θi(t)=Si0kI0tλ(t-τ)Sk0βk(τ),Πk(τ,0)qkdτ+fi(t),t0,=Si0kI0tβk(τ),Πk(τ,0)qkΘk(t-τ)dτ+fi(t),t0,

where fi(t) is the number of new infections produced by the initial population. Therefore, the basic reproduction number R0i of individuals of group i is calculated as

R0i=Si0kI0βk(τ),Πk(τ,0)qkdτ=Si0kIχk,

with

χk=0βk(τ),Πk(τ,0)qkdτ.

The term χk(τ)=βk(τ),Πk(τ,0)qk quantifies the infectiousness at τ-time post infection of the whole infected individuals of group kI.

Next, let Θ(t)=(Θi(t))iI, the number of new infections in all groups at time t and, f(t)=(fi(t))iI the number of new infections produced by the initial population. We have

Θ(t)=Si0kI0tχk(τ)Θk(t-τ)dτiI+f(t)=0tSi0χk(τ)i,kIΘ(t-τ)dτ+f(t),t>0,

Due to the above formulation, the basic reproduction number R0 of all individuals is calculated as the spectral radius of the matrix (ai,k)i,kI, where

ai,k=Si00χk(τ)dτ=Si0χk.

Some calculations give

R0=kISk00χk(τ)dτ=kISk0χk.

A more explicit expression of the infectiousness χk is difficult to obtain in general. However, one can go further steps in some particular configurations of the treatment status transition rates ω(τ)=0ωTU(τ)ωUT(τ)0. Indeed, assume that (2.7) holds. In such a configuration, we have

Πk(τ,0)=Πk0(τ,0),if0<τ<τ0,Πk0(τ0,0)Πk1(τ,τ0),ifτ0<τ<τ0+τ1,Oe-ckτ,ifτ>τ0+τ1,

with ck=μh+infταkU(τ)+infταkT(τ), and

Πk0(τ,0)=e-μhτdiage-0ταkT(σ)dσ,e-0ταkU(σ)dσ,Πk1(τ,0)=diagω¯10τe-μh(τ-η)e-ηταkT(σ)dσe-(ω¯1+μh)ηe-0ηαkU(σ)dσdη,e-(ω¯1+μh)τe-0ταkU(σ)dσ.

From where, by setting

Γ0k,ϑ(τ)=e-μhτe-0ταkϑ(σ)dσ,ϑ{T,U},Γ1k,U(τ)=e-(ω¯1+μh)τe-0ταkU(σ)dσ,Γ1k,T(τ)=e-μhτe-0ταkT(σ)dσ,

it comes

χk=qkTχkT+qkUχkU+Oe-ck(τ0+τ1),

where

χkT=0τ0+τ1βkT(τ)Γ0k,T(τ)dτ+Γ0k,T(τ0)ω¯1τ0τ0+τ1βkT(τ)τ0τΓ1k,T(τ)Γ1k,T(η)Γ1k,U(η)dηdτ,χkU=0τ0βkU(τ)Γ0k,U(τ)dτ+Γ0k,U(τ0)τ0τ0+τ1βkU(τ)Γ1k,U(τ)Γ1k,U(τ0)dτ.

Proof of Theorem 2.6

The equilibrium of system (2.5) is obtained by solving the following system for all iI

0=Λi-Siλ-μhSi,Ii(τ=0)=λSiqi,τIi(τ)=(-eω(τ)-αi(τ)-μh+ω(τ))Ii(τ), 5.6

where

λ=iI0βi(τ),Ii(τ)dτ. 5.7

Solving (5.6) for Si and Ii yields

Si=Λiμh+λandIi(τ)=λSiΠi(τ,0)qi. 5.8

Replacing (5.8) in (5.7) leads to 1+μh-1λλ=R0λ and since λ>0, we have λ=μh(R0-1).

It follows that system (2.5) has a unique positive endemic equilibrium when R0>1, such that iI,

Si=Si0R0andIi(τ)=λSiΠi(τ,0)qi,

where λ=μh(R0-1).

Technical materials

Before proceed to the proof Theorem 2.7, we introduce some technical materials including the existence of a global compact attractor for the solution semiflow of Model (2.5), the spectral properties of the linearized semiflow of Model (2.5) at any given equilibrium, and the uniform persistence of Model (2.5) when R0>1.

Global compact attractor

To derive the global properties of the solution dynamics, it is necessary to show that the semiflow generated by system (2.5) has a global compact attractor. Denote by

Y=Rn×L1((0,),R2n)andY+=R+n×L+1((0,),R2n),

and endow the set Y with the norm

(S,I(·))=iI|Si|+0(|IiT(t,τ)|+|IiU(t,τ)|)dτ.

For any initial condition φ0Y, the solution semiflow of system (2.5) in Y+ is denoted by Ψ(t,φ0)=(S(t),I(t,·))t. From the Volterra formulation (5.3), we rewrite system (2.5) as follows for all iI:

S˙i(t)=Λi-Si(t)λ(t)-μhSi(t),IiT(t,τ)=Πi,TT(τ,τ-t)Ii0T(τ-t)+Πi,UT(τ,τ-t)Ii0U(τ-t),iftτ,ΠiT(τ,0)λ(t-τ)Si(t-τ),ift>τ, 5.9
IiU(t,τ)=Πi,TU(τ,τ-t)Ii0T(τ-t)+Πi,UU(τ,τ-t)Ii0U(τ-t),iftτ,ΠiU(τ,0)λ(t-τ)Si(t-τ),ift>τ, 5.10

where

ΠiT(τ,0)=Πi,TT(τ,0)qiT+Πi,UT(τ,0)qiUandΠiU(τ,0)=Πi,TU(τ,0)qiT+Πi,UU(τ,0)qiU.

We need to prove the following claim.

Claim 5.3

Let Assumption 2.2 be satisfied. Then, function λ(·) is Lipschitz continuous on R+.

Proof of Claim 5.3

Let C0maxΛ¯μh,φ0, βi=max{βiT,βiU} and β=maxiIβi. Then, |λ(t)|C0β. Let t>0 and h>0. It comes that

λ(t+h)-λ(t)=iI0hβi(τ),Ii(t+h,τ)dτ+iIhβi(τ),Ii(t+h,τ)dτ-iI0βi(τ),Ii(t,τ)dτ=iI0hβi(τ),Πi(τ,0)Ii(t+h-τ,0)dτ+iIhβi(τ),Ii(t+h,τ)dτ-iI0βi(τ),Ii(t,τ)dτC02β2h+iI0βi(τ+h),Ii(t+h,τ+h)dτ-iI0βi(τ),Ii(t,τ)dτ.

Recalling (5.3) and combining the integrals, we obtain

λ(t+h)-λ(t)C02β2h+iI0βi(τ+h),(Πi(τ+h,τ)-I)Ii(t,τ)dτ+iI0βi(τ+h)-βi(τ),Ii(t,τ)dτC02β2h+iI0βi(τ+h),(γ1e-μhh-1)Ii(t,τ)dτ+iI0βi(τ+h)-βi(τ),Ii(t,τ)dτ.

We have |γ1e-μhh-1||e-μhh-1|μhh. Using the Lipschitzianity of βiϑ, we find a positive constant Cβ such that

|λ(t+h)-λ(t)|C02β2h+C0βμhh+CβC0hCλh,

where Cλ=C02β2+C0βμh+CβC0.

Next, we will show that system (2.5) has a global attractor. By using the similar method as in Martcheva and Thieme (2003) and Cheng et al. (2018), we can state the following result.

Lemma 5.4

There exists A0, a compact subset of Y+, which is a global attractor for the solution semiflow of system (2.5). Moreover, A0 is invariant under the solution semiflow, that is

Ψ(t,φ0)A0,foreveryφ0A0,t0.
Proof

We show that Ψ satisfies the assumptions of Lemma 3.2.3 and Theorem 3.4.6 in Hale (2010). To this end, we split the solution semiflow into two parts. For any initial condition φ0Y+, we let Ψ(t,φ0)=Ψ^(t,φ0)+Ψ~(t,φ0), where

Ψ^(t,φ0)=(0Rn,0L1),t>τ,(0Rn,I(t,τ)),tτ,andΨ~(t,φ0)=(S(t),I(t,τ)),t>τ,(S(t),0L1),tτ.

In such a way, we need to prove the following claim:

Claim 5.5
  1. Ψ^(t,φ0)0 as t for every φ0 in Y.

  2. For a fixed t and any bounded set B in Y, the set {Ψ~(t,φ0):φ0B} is precompact.

Proof of Claim 5.5

Now, we show that the first claim holds.

From (5.9) and Lemma 5.2, we have

Ψ^(t,φ0)=iI0(|IiT(t,τ)|+|IiU(t,τ)|)dτ=iI0[Πi,TT(τ,τ-t)+Πi,TU(τ,τ-t)]Ii0T(τ-t)dτ+iIR0[Πi,UT(τ,τ-t)+Πi,UU(τ,τ-t)]Ii0U(τ-t)dτ2γ1e-μhtφ0,

Note that for any bounded φ0, 2γ1e-μhtφ00 as t. This completes the first claim.

To show that the second claim holds, let BY be a bounded subset such that Ψ(t,·)BB. Choose C0>0 such that φ0C0 for all φ0B. From Theorem 5.1 Item 2, φ0B{S(t)} is bounded in Rn and then is precompact in Rn. Hence, to show the compactness, it suffices to show that the set Ψ~(t,φ0)B is precompact for

I~(t,τ)=I(t,τ),t>τ,0L1,tτ. 5.11

By Frechet–Kolmogorov theorem [see Theorem B.2 in Smith and Thieme (2011)], it is sufficient to verify the following conditions:

  • (i)

    supφBiI0(I~iT(t,τ)+I~iU(t,τ))dτ<,

  • (ii)

    limhiIh(|I~iT(t,τ)|+|I~iU(t,τ)|)dτ=0 uniformly with respect to φ0B.

  • (iii)

    limh0iI0(|I~iT(t,τ)-I~iT(t,τ+h)|+|I~iU(t,τ)-I~iU(t,τ+h)|)dτ=0 uniformly with respect to φ0B.

  • (iv)

    limh0iI0h(|I~iT(t,τ)|+|I~iU(t,τ)|)dτ=0 uniformly with respect to φ0B.

By (5.11) we have for all iI

I~i(t,τ)=λ(t-τ)Si(t-τ)Πi(τ,0)qi,t>τ,0L1,tτ.

It follows that above conditions (i), (ii) and (iv) are satisfied.

Now, we show that condition (iii) holds. We have for iI and ht,

0|I~iT(t,τ)-I~iT(t,τ+h)|dτ=0t-h|ΠiT(τ,0)λ(t-τ)Si(t-τ)-ΠiT(τ+h,0)λ(t-τ-h)Si(t-τ-h)|dτ+t-ht|ΠiT(τ,0)λ(t-τ)Si(t-τ)|dτ:=C1(t,h)+C2(t,h).

By Lemma 5.2, and the boundedness of the semiflow, we can find a positice constant C0 such that

C2(t,h)=t-ht|ΠiT(τ,0)λ(t-τ)Si(t-τ)|dτC02βγ1h0ash0. 5.12

Again by Lemma 5.2, we have

C1(t,h)0t-hΠiT(τ,0)Si(t-τ)|λ(t-τ)-λ(t-τ-h)|dτ+0t-hλ(t-τ-h)|ΠiT(τ,0)Si(t-τ)-ΠiT(τ+h,0)Si(t-τ-h)|dτ0t-hΠiT(τ,0)Si(t-τ)|λ(t-τ)-λ(t-τ-h)|dτ+0t-hλ(t-τ-h)Si(t-τ)|ΠiT(τ,0)-ΠiT(τ+h,0)|dτ+0t-hλ(t-τ-h)ΠiT(τ+h,0)|Si(t-τ)-Si(t-τ-h)|dτC02β0|ΠiT(τ,0)-ΠiT(τ+h,0)|dτ+γ1C00t-he-μhτ|λ(t-τ)-λ(t-τ-h)|dτ+γ1C0β0t-he-μhτ|Si(t-τ)-Si(t-τ-h)|dτC02βτΠiT(τ,0)|h|+(γ1C0Cλ+γ1C0CSiβ)h0t-he-μhτdτC02βτΠiT(τ,0)|h|+γ1C0μh(Cλ+CSiβ)h0ash0, 5.13

with CSi=Λi+Ci0(C0β+μh). By (5.12) and (5.13) one concludes that the criterion (iii) holds, and then the second claim holds.

This completes the proof of the lemma.

Spectral properties of the linearized semiflow

The next result is concerned with spectral properties of the linearized semiflow Ψ of Model (2.5) at a given equilibrium point φ~X0+. The associated linearized system (2.5) at the point φ~ reads as

dφ(t)dt=(A+G[φ~])φ(t),

where A is the linear operator defined in (5.1) while G[φ~]L(X0,X) is the bounded linear operator defined by:

G[φ~]φ=-Sλ~-S~λλ~diag(S)q+λdiag(S~)q0L1((0,),R2n),

where λ~=i0βi(τ),I~i(τ)dτ and λ=i0βi(τ),Ii(τ)dτ. We then have the following lemma.

Lemma 5.6

Let us set Ω={νC:Re(ν)>-μh}. Then, the spectrum σ(A+G[φ~])Ω only consists of the point spectrum and one has

σ(A+G[φ~])Ω={νΩ:Δ(ν,φ~)=0},

where function Δ(·,φ~):ΩC is defined by

Δ(ν,φ~)=1-Rν[φ~]+Rν[φ~]λ~ν+μh+λ~, 5.14

with Rν[φ~]=iIS~i0βi(τ),Πi(τ,0)qie-ντdτ.

Proof

Let us denote by A0:D(A0)X0X0 the part of A in X0=D(A), which is defined by

A0φ=Aφ,φD(A0)={φD(A):AφD(A)}.

Then, it is the infinitesimal generator of a C0-semigroup on X0 denoted by {TA0(t)}t0. Let φ=(S,0R2n,I(·))t. We find that

TA0(t)φ(τ)=(e-μhtS,0R2n,Π(τ,τ-t)I(τ-t))t,tτ,(e-μhtS,0R2n,0L1((0,),R2n))t,t>τ.

Then, for tτ0, we have TA0(t-τ0)φXe-μh(t-τ0)φX,tτ0. We deduce that the growth rate ω0(A0)=limt+ln(TA0(t)L(X))t of this semigroup satisfies ω0(A0)-μh. Since operator Gi[φ~] is compact, the results in Arino et al. (1998) or Ducrot et al. (2008) apply and provided that the essential growth rate of {T(A+G[φ~])0(t)}t0-the C0-semigroup generated by the part of (A+G[φ~]) in X0 satisfies

ω0,ess((A+G[φ~])0)ω0,ess(A0)<ω0(A0)-μh.

By results in Engel and Nagel (2001) and Webb (1987), the latter inequality ensures that Ωσ(A+G[φ~]), and it is only composed of point spectrum of (A+G[φ~]).

It remains to derive the characteristic equation. Let νρ(A+G[φ~]), where ρ(·) stands for the resolvent. For φ^=(S^,u^,I^(·))tX and φ=(S,0L1,I(·))tD(A), we have (νI-A-G[φ~])φ=φ^, that is (νI-A)φ-G[φ~]φ=φ^, and from where

φ=(νI-A)-1φ^+(νI-A)-1G[φ~]φ. 5.15

Since

(νI-A)-1φ^=S^iν+μhiI,0R2n,e-ν·Πi(·,0)u^i+0·Πi(·,s)I^i(s)e-ν(·-s)dsiIt,

we find that

(νI-A)-1G[φ~]φ=-Siλ~-S~iλν+μhiI0R2ne-ν·Πi(·,0)Siλ~qi+S~iλqiiI.

Thus, for all iI, equality (5.15) rewrites as

1+λ~ν+μhSi+S~iλν+μh=S^iν+μh,Si=S^iν+μh+λ~-S~iλν+μh+λ~Ii(τ)=e-ντΠi(τ,0)Siλ~qi+S~iλqi+e-ντΠi(τ,0)u^i+0τΠi(τ,s)I^i(s)e-ν(τ-s)ds. 5.16

Substituting (5.16) into expression for λ, it comes

1-Rν[φ~]+Rν[φ~]λ~ν+μh+λ~λ=Rν[φ^]λ~ν+μh+λ~+iI0βi(τ),y^i(τ)dτ, 5.17

where Rν[φ~]=iIS~i0βi(τ),Πi(τ,0)qie-ντdτ and y^i(τ)=e-ντΠi(τ,0)u^i+0τΠi(τ,s)I^i(s)e-ν(τ-s)ds. Therefore, we can isolate λ in system (5.17) if and only if Δ(ν,φ~)=1-Rν[φ~]+Rν[φ~]λ~ν+μh+λ~0.

Uniform persistence

Our next technical material concerns the uniform persistence of Model (2.5) when R0>1 by using the method developed in Theorem 5.2 in Smith and Thieme (2011).

For the invariant sets of uniform persistence, we introduce

M=Rn×{0R2n}×MI,andM=X0+\M,

where

MI=IL+1((0,),R2n):iI0βi(τ),Ii(τ)dτ>0.

For the unique solution φ=(S,0R2n,I) of system (2.5) associated to the initial condition φ0=(S0,0R2n,I0(·))M, we define Ψ(t,φ0)=(S(t),0R2n,I(t,·)) the semiflow of Model (2.5) passing through φ0. Next, we first claim that

Claim 5.7

The subsets M and M are positively invariant with respect to the semiflow Ψ(t,·) generated by system (2.5). Furthermore, limtΨ(t,φ0)=(S0,0R2n,0L+1((0,),R2n))T for each φ0M.

Proof of Claim 5.7

Let φ0=(S0,0R2n,I0(·))TM be given and Ψ(t,φ0)=(S,0R2n,I(t,·))T, the orbit passing through φ0. Since φ0M, then λ(0)>0. Through a direct calculation, we have

tλ(t)=iI0βi(τ),tIi(t,τ)dτ,-(α0+ω0+μh)λ(t),

where α0=maxiI{supαiT,supαiU} and ω0=max{ω¯TU,ω¯UT}. Thus, one obtains that

λ(t)e-(α0+ω0+μh)λ(0)>0,

for t0. This complete the fact that M is positively invariant.

Now, let φ0M. Since for all iI, Si(t)Si0 as t is large enough, the comparison principle implies that

Ii(t,τ)I~i(t,τ),iI, 5.18

where I~i(t,τ) is the solution of the following system

I~i(t,τ=0)=λ~(t)Si0qi,I~i(0,τ)=Ii0(τ),t+τI~i(t,τ)=(-eω(τ)-αi(τ)-μh+ω(τ))I~i(t,τ). 5.19

By the Volterra formulation, we have from (5.19) that

I~i(t,τ)=Πi(τ,τ-t)Ii0(τ-t),iftτ,Πi(τ,0)qiSi0λ~(t-τ),ift>τ,

where λ~(t) satisfies

λ~(t)=kISk00tβk(τ),Πk(τ,0)qkλ~(t-τ)dτ+kItβk(τ),Πk(τ,τ-t)Ik0(τ-t)dτ.

The initial condition with kI0βk(τ),Ik0(τ)dτ=0, leads to

λ~(t)=kISk00tβk(τ),Πk(τ,0)qkλ~(t-τ)dτ.

Since λ~(0)=0, we have λ~(t)=0 for all t0, and then I~i(t,·)=0 for all iI and t0. The comparison in (5.18) implies that Ii(t,·)=0 for all iI and t0 and then M is positively invariant under the semiflow Ψ(t,·). In addition, it is clear for the solution remaining in M, we have for all iI, SiSi0. Hence, limtΨ(t,φ0)=(S0,0R2n,0L+1((0,),R2n))T for each φ0M. This ends the proof of Claim 5.7.

Finally, we end this technical material section by establishing the uniform persistence of system (2.5).

Theorem 5.8

The semiflow {Ψ(t,·)}t0 generated by system (2.5) is uniformly persistent in M with respect to (M,M), that is, there exists a constant η>0 such that for each φ0M,

lim inftS(t)η,

and

lim inftI(t,·)L+1ηwheneverR0>1.

Furthermore, there exists compact global attractor A1 in M for the semiflow {Ψ(t,·)}t0.

Proof

In the following, we will prove that WS({E0})M=, where

WS({E0})={φ0X0+:limt+Ψ(t,φ0)=E0}.

Since from Claim 5.7 the disease-free equilibrium E0 is globally asymptotically stable in M, we need only to study the behavior of the solution starting in M in some neighborhood of E0. To this end, it is sufficient to show that there exists σ>0 satisfying for each φ{vM:E0-vσ} there exists t00 such that Ψ(t,φ0)-E0>σ.

By the way of contradiction, suppose that for each integer n0 there exists a φ0n=(S0n,0L1,I0n){vM:E0-vσ} such that

E0-Ψ(t,φ0n)1n+1,t0.

Denote Ψ(t,φ0n)=(Sn(t),0R2n,In(t,·)), then for all t0 we have

|Sn(t)-S0|1n+1,t0. 5.20

It follows that for all i, we have Sin(t)Si0-1n+1 for all t0. Consider the following system

t+τIn(t,τ)=(-diag(Φ(τ))+diag(ω(τ)))In(t,τ),In(t,0)=λn(t)diag(Sn(t))q,Sn(0)=S0n,In(0,·)=I0n(·),(S0n,I0n)M,

where λn(t)=iI0tβi(τ),Iin(t,τ)dτ, S0n=(Si0n)iI and I0n(·)=(Ii0n(·))iI. By the comparison principle, we have

In(t,·)I~n(t,·), 5.21

where I~n(t,·) is the solution of the following auxiliary system

t+τI~n(t,τ)=(-diag(Φ(τ))+diag(ω(τ)))I~n(t,τ),I~n(t,0)=λ~n(t)diagS0-1n+1Iq,I~n(0,·)=I0n(·),

which gives for all i,

t+τI~in(t,τ)=(-Φi(τ)+ω(τ))I~in(t,τ),I~in(t,0)=λ~n(t)Si0-1n+1qi,I~in(0,·)=Ii0n(·). 5.22

For ease of notation, let us rewrite the system (5.22) as the following form:

dv~in(t)dt=A~in+L~inv~in(t),t0, 5.23

v~in(0)D(A~in)¯, the closure of D(A~in)=0R2×W1,1((0,),R2), where v~in(t)=(0R2,I~in(t,·))t and the operators A~in and L~in are defined as

A~in(0R2,I~in)=(-I~in(0),-τI~in+(-Φi(τ)+ω(τ))I~in)

and

L~in(0R2,I~in)=λ~n(t)Si0-1n+1qi,0L1((0,),R2).

Similarly to the proof of Lemma 5.6, we can derive the characteristic equation Δ(νn)=0 for system (5.22), where

Δ(νn)=1-iISi0-1n+10βi(τ),Πi(τ,0)qie-νnτdτ.

Since R0>1, there exists n0>0 large enough such that for nn0,

R0n=iISi0-1n+10βi(τ),Πi(τ,0)qidτ>1.

The largest eigenvalue νn of system (5.23) satisfies the characteristic equation Δ(νn)=0. Furthermore, R0n>1 implies the existence of a dominant eigenvalue νn>0 such that Δ(νn)=0. Therefore, νn>0 is a simple dominant eigenvalue of (A~in+L~in). From Lemma 5.6, we have shown that ω0,ess(A~in+L~in)-μh and since the semigroup T(A~in+L~in)(t)t0 is irreducible, it follows from Corollary 4.6.8 in Magal and Ruan (2018) that T(A~in+L~in)(t)t0 has asynchronous exponential growth with intrinsic growth constant νnR. Therefore, using Theorem 3.9 in Magal and McCluskey (2013), we have

T(A~in+L~in)0(t)Π~νn=Π~νnT(A~in+L~in)0(t)=eνntΠ~νn,t0,

and there exist constants ε0>0 and η0>0 such that

T(A~in+L~in)0(t)(I-Π~νn)η0e(νn-ε0)t(I-Π~νn),t0,

where Π~νn is the projector on the generalized eigenspace associated with the largest eigenvalue νn>0. We deduce that

Π~νnv~in(t)=eνntΠ~νnv~i0n.

Since νn>0, it follows that limtΠ~νnv~in(t)L1=+. Therefore, limtI~in(t,·)L1=+ and from (5.21), we have limtIin(t,·)L1=+, which is a contradiction to the boundedness of the solution. Thus, WS({E0})M= and we derive from Theorem 4.2 in Hale and Waltman (1989) that the semiflow {Ψ(t,·)}t0 is uniform persistent with respect to the pair (M,M). Moreover, by Theorem 3.7 in Magal and Zhao (2005), there exists a compact global attractor A1M for the semiflow {Ψ(t,·)}t0.

Proof of Theorem 2.7

The proof of Theorem 2.7 is decomposed into two parts. The first part is devoted to the global stability of the disease-free equilibrium, while the second part is devoted to the global stability of the endemic equilibrium.

Proof of Theorem 2.7 (i): global stability of the disease-free equilibrium

When the initial condition of System (2.5) satisfies φ0M, i.e., iI0βi(τ),Ii0(τ)dτ=0, by Claim 5.7, it comes that the semiflow Ψ(t,φ0) generated by system (2.5) is such that limtΨ(t,φ0)=(S0,0R2n,0L+1((0,),R2n)). It then remain to prove the global stability of the disease-free equilibrium when R01.

Theorem 6.1

If R01 then, the disease-free equilibrium E0=(S0,0L1((0,)×R,R2n)) of system (2.5) is globally asymptotically stable.

Proof

By Theorem 5.8, we introduce the following well defined Lyapunov functional V(t)=V1(t)+V2(t), with

V1(t)=kISk-Sk0-Sk0lnSkSk0andV2(t)=kI0ckT(τ)IkT(t,τ)+ckU(τ)IkU(t,τ)dτ,

where ck(τ)=(ckT(τ),ckU(τ))T is a vector of positive constants such that

ck(τ)=S¯0τtΠk(σ,τ)βk(σ)dσ, 6.1

with S¯0=kISk0 and ck(τ)0 as τ. From (6.1), we have

S¯0βkT(τ)+τckT(τ)-γkT(τ)ckT(τ)+ωUT(τ)ckU(τ)=0,S¯0βkU(τ)+τckU(τ)-γkU(τ)ckU(τ)+ωTU(τ)ckT(τ)=0, 6.2

with

γkT(τ)=αkT(τ)+ωUT(τ)+μhandγkU(τ)=αkU(τ)+ωTU(τ)+μh.

Differentiating V1(t) and using Λk=μhSk0, we have

dV1(t)dt=kI1-Sk0SkΛk-Sk(t)λ(t)-μhSk(t)=kIΛk2-Sk0Sk-SkSk0+kIλ(t)Sk0-λ(t)Sk=kIΛk2-Sk0Sk-SkSk0-kIλ(t)Sk+kIS¯00(βkT(τ)IkT(t,τ)+βkU(τ)IkU(t,τ))dτ.

Differentiating V2(t), we have

dV2(t)dt=kI0ckT(τ)tIkT(t,τ)+ckU(τ)tIkU(t,τ)dτ=-kI0ckT(τ)τIkT(t,τ)+γkT(τ)IkT(t,τ)-ωTU(τ)IkU(t,τ)dτ-kI0ckU(τ)τIkU(t,τ)+γkU(τ)IkU(t,τ)-ωUT(τ)IkT(t,τ)dτ.

By integrating by parts, we have

kI0ckT(τ)τIkT(t,τ)dτ=-kIckT(0)IkT(t,0)-kI0τckT(τ)IkT(t,τ)dτ

and

kI0ckU(τ)τIkU(t,τ)dτ=-kIckU(0)IkU(t,0)-kI0τckU(τ)IkU(t,τ)dτ.

Replacing these expressions in dV2(t)dt and using the fact that IkT(t,0)=qkTλ(t)Sk(t) and IkU(t,0)=qkUλ(t)Sk(t), we have

dV2(t)dt=kIλ(t)Sk(ckT(0)qkT+ckU(0)qkU)+kI0(τckT(τ)-γkT(τ)ckT(τ)+ωUT(τ)ckU(τ))IkT(t,τ)dτ+kI0(τckU(τ)-γkU(τ)ckU(τ)+ωTU(τ)ckT(τ))IkU(t,τ)dτ.

Finally, combining dV1(t)dt and dV2(t)dt, gathering some terms and using (6.2), it follows that

dV(t)dt=kIΛk2-Sk0Sk-SkSk0+kIλ(t)Sk(ckT(0)qkT+ckU(0)qkU-1)+kI0(S¯0βkT(τ)+τckT(τ)-γkT(τ)ckT(τ)+ωUT(τ)ckU(τ))IkT(t,τ)dτ+kI0(S¯0βkU(τ)+τckU(τ)-γkU(τ)ckU(τ)+ωTU(τ)ckT(τ))IkU(t,τ)dτ=kIΛk2-Sk0Sk-SkSk0+kIλ(t)Sk(ck(0),qk-1).

Using (6.1), we have

kISk0ck(0),qk=S¯0kISk00βk(τ),Πk(τ,0)qkdτ=S¯0R0=kISk0R0kISk0,sinceR01.

That implies that for all k, ck(0),qk1, when R01. Therefore, we have dV(t)dt0 when R01. The strict equality holds only if Sk(t)=Sk0 hold simultaneously with either R0=1 or Ik(t,0)=0. It is easy to verify that largest invariant set in dVdt=0 is the singleton {E0}. Thus, all solutions of system (2.5) converge to the disease-free equilibrium E0. Hence, E0 is globally asymptotically stable when R01.

Proof of Theorem 2.7 (ii): global stability of the endemic equilibrium

Theorem 6.2

Assume R0>1, then the endemic equilibrium E=(S,I)t of system (2.5) is globally asymptotically stable in Y+.

Proof

By Theorem 5.8, we introduce the following well defined Lyapunov functional L(t)=L1(t)+L2(t)+L3(t), where

L1(t)=kISkhSkSk,L2(t)=kI0dkT(τ)IkT(τ)hIkT(t,τ)IkT(τ)dτandL3(t)=kI0dkU(τ)IkU(τ)hIkU(t,τ)IkU(τ)dτ,

with h the function defined by h(z)=z-1-lnz (zR+), and dk(τ)=(dkT(τ),dkU(τ))T a vector of positive constants given by

dk(τ)=S¯τtΠk(σ,τ)βk(σ)dσ, 6.3

where S¯=kISk and dk(τ)0 as τ. From (6.3), we have

τdk(τ)-(-teω(τ)-αk(τ)-μh+ω(τ))dk(τ)=-S¯βk(τ). 6.4

By using the property of function h, we find that the function L(t) is nonnegative with its global minimum point E.

Step 1: Differentiating L1(t) along the solution of system (1.5) and using Λk=Skλ+μhSk, we obtain

dL1(t)dt=kI1-SkSkΛk-Sk(t)λ(t)-μhSk(t)=kIμhSk2-SkSk-SkSk+kIλSk1+λ(t)λ-λ(t)SkλSk-SkSk=kIμhSk2-SkSk-SkSk+kIλSkhλ(t)λ-hSkSk-hλ(t)Sk(t)λSk.

Step 2: Note that

τIkT(τ)=-γkT(τ)IkT(τ)+ωTU(τ)IkU(τ)andτIkU(τ)=-γkU(τ)IkU(τ)+ωUT(τ)IkT(τ). 6.5

Using (6.5), we have

1-IkT(τ)IkT(t,τ)τIkT(t,τ)=IkT(τ)τhIkT(t,τ)IkT(τ)-1-IkT(τ)IkT(t,τ)γkT(τ)-ωTU(τ)IkU(τ)IkT(τ)IkT(t,τ). 6.6

Differentiating L2(t) and, using (6.6) and integration by parts, we obtain

dL2(t)dt=-kI0dkT(τ)1-IkT(τ)IkT(t,τ)τIkT(t,τ)+γkT(τ)IkT(t,τ)-ωTU(τ)IkU(t,τ)dτ=-kI0dkT(τ)IkT(τ)τhIkT(t,τ)IkT(τ)-ωTU(τ)IkU(τ)-ωTU(τ)IkU(t,τ)+ωTU(τ,x)IkU(τ)IkT(t,τ)IkT(τ)+ωTU(τ)IkT(τ)IkU(t,τ)IkT(t,τ)dτ=-kIdkT(τ)IkT(τ)hIkT(t,τ)IkT(τ)τ=0τ=+kI0dkT(τ)ωTU(τ)IkU(τ)1+IkU(t,τ)IkU(τ)-IkT(t,τ)IkT(τ)-IkT(τ)IkU(t,τ)IkT(t,τ)IkU(τ)dτ+kI0hIkT(t,τ)IkT(τ)(IkT(τ)τdkT(τ)+dkT(τ)τIkT(τ))dτ=kIdkT(0)IkT(0)hIkT(t,0)IkT(0)+kI0dkT(τ)ωTU(τ)IkU(τ)1+IkU(t,τ)IkU(τ)-IkT(t,τ)IkT(τ)-IkT(τ)IkU(t,τ)IkT(t,τ)IkU(τ)dτ+kI0hIkT(t,τ)IkT(τ)(IkT(τ)τdkT(τ)+dkT(τ)τIkT(τ))dτ.

Since IkT(0)=qkTλSk and IkT(t,0)=qkTλ(t)Sk(t), then we have

dL2(t)dt=kIλSkdkT(0)qkThλ(t)Sk(t)λSk+kI0dkT(τ)ωTU(τ)IkU(τ)1+IkU(t,τ)IkU(τ)-IkT(t,τ)IkT(τ)-IkT(τ)IkU(t,τ)IkT(t,τ)IkU(τ)dτ+RkI0hIkT(t,τ)IkT(τ)(IkT(τ)τdkT(τ)+dkT(τ)τIkT(τ))dτ.

By a similarly manner, the derivative of L3(t) gives

dL3(t)dt=kIλSkdkU(0)qkUhλ(t)Sk(t)λSk+kI0dkU(τ)ωUT(τ)IkT(τ)1+IkT(t,τ)IkT(τ)-IkU(t,τ)IkU(τ)-IkU(τ)IkT(t,τ)IkU(t,τ)IkT(τ)dτ+kI0hIkU(t,τ)IkU(τ)(IkU(τ)τdkU(τ)+dkU(τ)τIkU(τ))dτ.

Step 3: Finally, combining dL1(t)dt, dL2(t)dt and dL3(t)dt, we obtain

dL(t)dt=kIμhSk2-SkSk-SkSk+kIλSkhλ(t)λ-hSkSk-hλ(t)Sk(t)λSk+kIλSkdk(0),qkhλ(t)Sk(t)λSk+kI0hIkT(t,τ)IkT(τ)(IkT(τ)τdkT(τ)+dkT(τ)τIkT(τ))dτ+kI0hIkU(t,τ)IkU(τ)(IkU(τ)τdkU(τ)+dkU(τ)τIkU(τ))dτ+kI0dkT(τ)ωTU(τ)IkU(τ)1+IkU(t,τ)IkU(τ)-IkT(t,τ)IkT(τ)-IkT(τ)IkU(t,τ)IkT(t,τ)IkU(τ)dτ+kI0dkU(τ)ωUT(τ)IkT(τ)1+IkT(t,τ)IkT(τ)-IkU(t,τ)IkU(τ)-IkU(τ)IkT(t,τ)IkU(t,τ)IkT(τ)dτ.

We observe that dkT(τ) and dkU(τ) satisfy

dkU(τ)ωUT(τ)IkT(τ)=dkT(τ)ωTU(τ)IkU(τ). 6.7

By using (6.7), (6.5), and (6.4), we obtain

τdkϑ(τ)Ikϑ(τ)+dkϑ(τ)τIkϑ(τ)=-S¯βkϑ(τ)Ikϑ(τ),ϑ={T,U}. 6.8

Moreover, by using (5.7) and (5.8), we have

λ=λkISk0βk(τ),Πk(τ,0)qkdτ. 6.9

Thus, using (6.3) and (6.9), it can be verified that

kIλSkdk(0),qk=λS¯kISk0βk(τ),Πk(τ,0)qkdτ=λS¯=kIλSk, 6.10

which implies that for all k, dk(0),qk=dkT(0)qkT+dkU(0)qkU=1. In addition, note that

kIλSkhλ(t)λ=kI0S¯(βkT(τ)IkT(τ)+βkU(τ)IkU(τ))hλ(t)λdτ. 6.11

Replacing (6.7), (6.8), (6.10) and (6.11) in dL(t)dt and gathering some terms, we obtain

dL(t)dt=kIμhSk2-SkSk-SkSk-kIλSkhSkSk+kI0S¯βkT(τ)IkT(τ)hλ(t)λ-hIkT(t,τ)IkT(τ)dτ+kI0S¯βkU(τ)IkU(τ)hλ(t)λ-hIkU(t,τ)IkU(τ)dτ+kI0dkT(τ)ωTU(τ)IkU(τ)2-IkT(τ)IkT(t,τ)IkU(t,τ)IkU(τ)-IkU(τ)IkU(t,τ)IkT(t,τ)IkT(τ)dτ.

Note that

hλ(t)λ-hIkϑ(t,τ)Ikϑ(τ)=-hIkϑ(t,τ)λIkϑ(τ)λ(t)+λ(t)λ-11-Ikϑ(t,τ)λIkϑ(τ)λ(t),ϑ={T,U}.

Hence, we have

dL(t)dt=kIμhSk2-SkSk-SkSk-kI0S¯βkT(τ)IkT(τ)hIkT(t,τ)λIkT(τ)λ(t)dτ-kIλSkhSkSk-kI0S¯βkU(τ)IkU(τ)hIkU(t,τ)λIkU(τ)λ(t)dτ+kI0dkT(τ)ωTU(τ)IkU(τ)2-IkT(τ)IkT(t,τ)IkU(t,τ)IkU(τ)-IkU(τ)IkU(t,τ)IkT(t,τ)IkT(τ)dτ+λ(t)λ-1kI0S¯βkT(τ)IkT(τ)1-IkT(t,τ)λIkT(τ)λ(t)+S¯βkU(τ,)IkU(τ)1-IkU(t,τ)λIkU(τ)λ(t)dτ.

Note that

kI0S¯βkT(τ)IkT(τ)1-IkT(t,τ)λIkT(τ)λ(t)+S¯βkU(τ)IkU(τ)1-IkU(t,τ)λIkU(τ)λ(t)dτ=0.

Finally, we have

dL(t)dt=kIμhSk2-SkSk-SkSk-kI0S¯βkT(τ)IkT(τ)hIkT(t,τ)λIkT(τ)λ(t)dτ-kIλSkhSkSk-kI0S¯βkU(τ)IkU(τ)hIkU(t,τ)λIkU(τ)λ(t)dτ+kI0dkT(τ)ωTU(τ)IkU(τ)2-IkT(τ)IkT(t,τ)IkU(t,τ)IkU(τ)-IkU(τ)IkU(t,τ)IkT(t,τ)IkT(τ)dτ.

Thus, dL(t)dt0 with equality if and only if Sk(t)=Sk, IkT(t,τ)=IkT(τ) and IkU(t,τ)=IkU(τ). Then, it can be verified that largest invariant set in dLdt=0 is the singleton {E}. It follows that the compact global attractor A0, stated by Lemma 5.4, is such that A0={E}. Therefore, the endemic equilibrium E is globally asymptotically stable in Y+ when R0>1.

Acknowledgements

This work was supported by the ANR (https://anr.fr/en/) grant QUASAR (grant agreement ANR-21-CE45-0004). The funders had no role in study design, data collection and analysis, decision to publish, or preparation of the manuscript.

Declarations

Conflict of interest

The authors of this article declare that they have no financial conflict of interest with the content of this article.

Footnotes

Publisher's Note

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

References

  1. Almocera AES, Nguyen VK, Hernandez-Vargas EA. Multiscale model within-host and between-host for viral infectious diseases. J Math Biol. 2018;77(4):1035–1057. doi: 10.1007/s00285-018-1241-y. [DOI] [PubMed] [Google Scholar]
  2. André J-B, Gandon S. Vaccination, within-host dynamics, and virulence evolution. Evol; Int J Org Evol. 2006;60(1):13–23. [PubMed] [Google Scholar]
  3. Arino O, Axelrod D, Kimmel M, Capasso V, Fitzgibbon W, Jagers P, Kirschner D, Mode C, Novak B, Sachs R, Stephan W, Swierniak A, Thieme H, Boussouar A (1998) Advances in mathematical population dynamics? Molecules, cells and man. In: Advances in mathematical population dynamics ? Molecules, cells and man, volume 6 of series in mathematical biology and medicine. World Scientific, pp 1–910
  4. Beardmore RE, Peña-Miller R, Gori F, Iredell J. Antibiotic cycling and antibiotic mixing: Which one best mitigates antibiotic resistance? Mol Biol Evol. 2017;34(4):802–817. doi: 10.1093/molbev/msw292. [DOI] [PMC free article] [PubMed] [Google Scholar]
  5. Blanquart F. Evolutionary epidemiology models to predict the dynamics of antibiotic resistance. Evol Appl. 2019;12(3):365–383. doi: 10.1111/eva.12753. [DOI] [PMC free article] [PubMed] [Google Scholar]
  6. Boldin B, Diekmann O. Superinfections can induce evolutionarily stable coexistence of pathogens. J Math Biol. 2008;56(5):635–672. doi: 10.1007/s00285-007-0135-1. [DOI] [PubMed] [Google Scholar]
  7. Burie J-B, Djidjou-Demasse R, Ducrot A. Asymptotic and transient behaviour for a nonlocal problem arising in population genetics. Eur J Appl Math. 2020;31(1):84–110. doi: 10.1017/S0956792518000487. [DOI] [Google Scholar]
  8. Cheng C-Y, Dong Y, Takeuchi Y. An age-structured virus model with two routes of infection in heterogeneous environments. Nonlinear Anal Real World Appl. 2018;39:464–491. doi: 10.1016/j.nonrwa.2017.07.013. [DOI] [Google Scholar]
  9. Coombs D, Gilchrist MA, Ball CL. Evaluating the importance of within- and between-host selection pressures on the evolution of chronic pathogens. Theor Popul Biol. 2007;72(4):576–591. doi: 10.1016/j.tpb.2007.08.005. [DOI] [PubMed] [Google Scholar]
  10. D’Agata EMC, Dupont-Rouzeyrol M, Magal P, Olivier D, Ruan S. The impact of different antibiotic regimens on the emergence of antimicrobial-resistant bacteria. PLoS One. 2008;3(12):e4036. doi: 10.1371/journal.pone.0004036. [DOI] [PMC free article] [PubMed] [Google Scholar]
  11. Day T, Read AF. Does high-dose antimicrobial chemotherapy prevent the evolution of resistance? PLoS Comput Biol. 2016;12(1):e1004689. doi: 10.1371/journal.pcbi.1004689. [DOI] [PMC free article] [PubMed] [Google Scholar]
  12. Diekmann O, Heesterbeek JAP, Metz JAJ. On the definition and the computation of the basic reproduction ratio R0 in models for infectious diseases in heterogeneous populations. J Math Biol. 1990;28(4):365–382. doi: 10.1007/BF00178324. [DOI] [PubMed] [Google Scholar]
  13. Djidjou-Demasse R (2021) Antimicrobial-Quantitative-Resistance-v1.0. Zenodo
  14. Djidjou-Demasse R, Ducrot A, Fabre F. Steady state concentration for a phenotypic structured problem modeling the evolutionary epidemiology of spore producing pathogens. Math Models Methods Appl Sci. 2017;27(02):385–426. doi: 10.1142/S0218202517500051. [DOI] [Google Scholar]
  15. Djidjou-Demasse R, Alizon S, Sofonea MT. Within-host bacterial growth dynamics with both mutation and horizontal gene transfer. J Math Biol. 2021;82(3):16. doi: 10.1007/s00285-021-01571-9. [DOI] [PubMed] [Google Scholar]
  16. Djidjou-Demasse R, Sofonea MT, Choisy M, Alizon S (2023) Within-host evolutionary dynamics of antimicrobial quantitative resistance. Math Model Nat Phenom 18:24
  17. Ducrot A, Liu Z, Magal P. Essential growth rate for bounded linear perturbation of non-densely defined Cauchy problems. J Math Anal Appl. 2008;341(1):501–518. doi: 10.1016/j.jmaa.2007.09.074. [DOI] [Google Scholar]
  18. Elderd BD, Mideo N, Duffy MA. Looking across scales in disease ecology and evolution. Am Nat. 2022;199(1):51–58. doi: 10.1086/717176. [DOI] [PubMed] [Google Scholar]
  19. Engel K-J, Nagel R. One-parameter semigroups for linear evolution equations. Semigroup Forum. 2001;63(2):278–280. doi: 10.1007/s002330010042. [DOI] [Google Scholar]
  20. Geritz SAH, Metz JAJ, Kisdi É, Meszéna G. Dynamics of adaptation and evolutionary branching. Phys Rev Lett. 1997;78(10):2024–2027. doi: 10.1103/PhysRevLett.78.2024. [DOI] [Google Scholar]
  21. Gilchrist MA, Coombs D. Evolution of virulence: interdependence, constraints, and selection using nested models. Theor Popul Biol. 2006;69(2):145–153. doi: 10.1016/j.tpb.2005.07.002. [DOI] [PubMed] [Google Scholar]
  22. Hale JK (2010) Asymptotic behavior of dissipative systems. American Mathematical Society
  23. Hale JK, Waltman P. Persistence in infinite-dimensional systems. SIAM J Math Anal. 1989;20(2):388–395. doi: 10.1137/0520025. [DOI] [Google Scholar]
  24. Hart WS, Maini PK, Yates CA, Thompson RN. A theoretical framework for transitioning from patient-level to population-scale epidemiological dynamics: influenza A as a case study. J R Soc Interface. 2020;17(166):20200230. doi: 10.1098/rsif.2020.0230. [DOI] [PMC free article] [PubMed] [Google Scholar]
  25. Inaba H. On a new perspective of the basic reproduction number in heterogeneous environments. J Math Biol. 2012;65(2):309–348. doi: 10.1007/s00285-011-0463-z. [DOI] [PubMed] [Google Scholar]
  26. Kepler TB, Perelson AS. Drug concentration heterogeneity facilitates the evolution of drug resistance. Proc Natl Acad Sci USA. 1998;95(20):11514–11519. doi: 10.1073/pnas.95.20.11514. [DOI] [PMC free article] [PubMed] [Google Scholar]
  27. Larsson DGJ, Flach C-F. Antibiotic resistance in the environment. Nat Rev Microbiol. 2022;20(5):257–269. doi: 10.1038/s41579-021-00649-x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  28. Lipsitch M, Levin BR. The population dynamics of antimicrobial chemotherapy. Antimicrob Agents Chemother. 1997;41(2):363–373. doi: 10.1128/AAC.41.2.363. [DOI] [PMC free article] [PubMed] [Google Scholar]
  29. Iannelli M (1995) Mathematical Theory of Age-Structured Population Dynamics. Giardini Editori e Stampatori in Pisa
  30. Magal P, McCluskey C. Two-group infection age model including an application to nosocomial infection. SIAM J Appl Math. 2013;73(2):1058–1095. doi: 10.1137/120882056. [DOI] [Google Scholar]
  31. Magal P, Ruan S. On semilinear Cauchy problems with non-dense domain. Adv Differ Equ. 2009;14(11–12):1041–1084. [Google Scholar]
  32. Magal P, Ruan S. Theory and applications of abstract semilinear cauchy problems, applied mathematical sciences. Cham: Springer; 2018. [Google Scholar]
  33. Magal P, Zhao X-Q. Global attractors and steady states for uniformly persistent dynamical systems. SIAM J Math Anal. 2005;37(1):251–275. doi: 10.1137/S0036141003439173. [DOI] [Google Scholar]
  34. Martcheva M, Thieme HR. Progression age enhanced backward bifurcation in an epidemic model with super-infection. J Math Biol. 2003;46(5):385–424. doi: 10.1007/s00285-002-0181-7. [DOI] [PubMed] [Google Scholar]
  35. Metz JAJ, Geritz SAH, Meszena G, Jacobs FJA, van Heerwaarden JS. Adaptive dynamics: a geometrical study of the consequences of nearly faithful reproduction. Amsterdam: North-Holland; 1996. [Google Scholar]
  36. Millan AS, Peña-Miller R, Toll-Riera M, Halbert ZV, McLean AR, Cooper BS, MacLean RC. Positive selection and compensatory adaptation interact to stabilize non-transmissible plasmids. Nat Commun. 2014;5(1):5208. doi: 10.1038/ncomms6208. [DOI] [PMC free article] [PubMed] [Google Scholar]
  37. Shen M, Xiao Y, Rong L, Zhuang G. Global dynamics and cost-effectiveness analysis of HIV pre-exposure prophylaxis and structured treatment interruptions based on a multi-scale model. Appl Math Model. 2019;75:162–200. doi: 10.1016/j.apm.2019.05.024. [DOI] [Google Scholar]
  38. Smith HL, Thieme HR (2011) Dynamical systems and population persistence. American Mathematical Society
  39. Tazzyman SJ, Bonhoeffer S. Plasmids and evolutionary rescue by drug resistance. Evolution. 2014;68(7):2066–2078. doi: 10.1111/evo.12423. [DOI] [PubMed] [Google Scholar]
  40. Thieme HR. Global stability of the endemic equilibrium in infinite dimension: Lyapunov functions and positive operators. J Differ Equ. 2011;250(9):3772–3801. doi: 10.1016/j.jde.2011.01.007. [DOI] [Google Scholar]
  41. Uecker H, Bonhoeffer S. Antibiotic treatment protocols revisited: the challenges of a conclusive assessment by mathematical modelling. J R Soc Interface. 2021;18(181):20210308. doi: 10.1098/rsif.2021.0308. [DOI] [PMC free article] [PubMed] [Google Scholar]
  42. Webb GF (1985) Theory of nonlinear age-dependent population dynamics. CRC Press
  43. Webb GF. An operator-theoretic formulation of asynchronous exponential growth. Trans Am Math Soc. 1987;303(2):751–763. doi: 10.1090/S0002-9947-1987-0902796-7. [DOI] [Google Scholar]
  44. Xue KS, Bloom JD. Linking influenza virus evolution within and between human hosts. Virus Evol. 2020;6(1):veaa010. doi: 10.1093/ve/veaa010. [DOI] [PMC free article] [PubMed] [Google Scholar]

Articles from Journal of Mathematical Biology are provided here courtesy of Springer

RESOURCES