730
Views
0
CrossRef citations to date
0
Altmetric
Special issue In memory of Fred Brauer

Effects of behaviour change on HFMD transmission

, , , &
Article: 2244968 | Received 15 Aug 2022, Accepted 01 Aug 2023, Published online: 15 Aug 2023

ABSTRACT

We propose a hand, foot and mouth disease (HFMD) transmission model for children with behaviour change and imperfect quarantine. The symptomatic and quarantined states obey constant behaviour change while others follow variable behaviour change depending on the numbers of new and recent infections. The basic reproduction number R0 of the model is defined and shown to be a threshold for disease persistence and eradication. Namely, the disease-free equilibrium is globally asymptotically stable if R01 whereas the disease persists and there is a unique endemic equilibrium otherwise. By fitting the model to weekly HFMD data of Shanghai in 2019, the reproduction number is estimated at 2.41. Sensitivity analysis for R0 shows that avoiding contagious contacts and implementing strict quarantine are essential to lower HFMD persistence. Numerical simulations suggest that strong behaviour change not only reduces the peak size and endemic level dramatically but also impairs the role of asymptomatic transmission.

MATHEMATICS SUBJECT CLASSIFICATIONS:

This article is part of the following collections:
An article collection in honour of Fred Brauer

1. Introduction

The emergence and spread of infectious diseases not only pose a great threat to public health, but also hinder economic growth and social stability. Among them is hand, foot and mouth disease (HFMD), a common viral infection in infants and children under 5 years old. More than 20 types of enteroviruses can cause HFMD, of which coxsackievirus A16 (CA16) and enterovirus 71 (EV71) are the two major etiological agents. The disease is primarily spread through close contact with infected people, through the respiratory tract or the faecal-oral route. The incubation period for HFMD is typically around 3–5 days. The virus can be isolated from the faeces or throat secretions of patients during the incubation period. Most infected people are asymptomatic or have mild symptoms, but they can still spread the infection. The clinical manifestations of HFMD usually include fever, mouth sores and skin rash on hands, feet and mouth, lasting 7-10 days. In particular, CA16 is associated with mild cases while EV71 infections could be severe and even fatal. Recovery from infection induces immunity against that specific serotype but is not immune to other serotypes. No specific treatment is yet available for HFMD and pain relievers can be given to reduce discomfort. So far three inactivated monovalent EV71 vaccines have been approved in China, but the coverage is relatively low. Good personal hygiene such as washing hands properly can reduce the risk of infection substantially.

HFMD was first reported in New Zealand in 1957. The disease is now endemic in many parts of the world. In particular, it is prevalent in Southeast Asian countries and regions such as Singapore, Vietnam, Mongolia and Malaysia. For example, an outbreak of HFMD in the state of Sarawak in Malaysia in 1997 caused 2626 cases and 31 deaths [Citation38]. There are also cases of HFMD in developed economies including Japan, the United States and Canada [Citation32]. In China, the earliest recorded HFMD case was from Shanghai in 1981, and the first HFMD outbreak occurred in Tianjin in 1983, which resulted in more than 7,000 confirmed cases of CA16 infection. Since then, cases have emerged in Northeast, East and South of China. In Spring 2008, a large-scale outbreak of HFMD due to EV71 occurred in Fuyang City, Anhui Province, with almost 25,000 reported cases. Subsequently, HFMD was categorized as a legally notifiable disease in China and there are usually more than one million reported cases every year [Citation18] (see Figure ).

Figure 1. Annual number of HFMD cases reported in mainland China, 2008–2021.

Figure 1. Annual number of HFMD cases reported in mainland China, 2008–2021.

Mathematical modelling of HFMD has a relatively short history compared to other common infectious diseases like influenza, measles and malaria, but it has developed rapidly in the past two decades. In 2004, Wang and Sung [Citation41] adopted an SIRS model with seasonally-forced transmission rate to evaluate the impact of weather on severe enterovirus infections in Taiwan from 2000 to 2003. Tiing and Labadin [Citation38] used an SIRS model that consists of natural birth and death and disease-caused death to predict the number of infectives and the duration of the outbreak in Sarawak, Malaysia. Roy and Halder [Citation33] established a HFMD model to distinguish between symptomatic cases and asymptomatic cases. Liu [Citation23] proposed a periodic SEIQRS model and showed that quarantine is beneficial to the control of HFMD. Ma et al. [Citation27] generalized the model of Liu [Citation23] to an SEIAQRS model by adding asymptomatic infection and transmission and fitted the model to the weekly HFMD data of Shandong Province. Li et al. [Citation22] constructed a two-stage-structured (i.e. children and adults) SEIQRS model and estimated the annual basic reproduction number of HFMD in China from 2009 to 2014. To explore the impact of contaminated environments on HFMD dynamics, Wang et al. [Citation43] formulated an SEIARS-W model with both human-to-human and environment-to-human transmission routes where W denotes the concentration of pathogen in the environment. Shi et al. [Citation34] took vaccination, contaminated environments, quarantine and asymptomatic infection into consideration and evaluated the contribution of EV71 vaccine in reducing HFMD cases. Recently, Zhao et al. [Citation48] developed a stage-structured (scattered and school children) SEIRS model where transmission rates and latent periods are temperature-dependent or time-periodic and found that the time-averaged system may underestimate the infection risk and size.

Human behaviour and disease transmission interact with each other strongly [Citation28]. During the 1918 influenza pandemic, a range of nonpharmaceutical interventions such as isolating infected people, closing schools, restricting or banning mass gatherings, and promoting cleaning and disinfection, were tried to mitigate the disease spread across the U.S. and shown to reduce the overall mortality significantly in cities where early and effective interventions were implemented [Citation4]. In response to the SARS epidemic in 2003, residents in Hong Kong quickly adopted preventive measures including mask use, frequent hand washing, and avoiding crowded places [Citation20]. Travel restrictions, stay-at-home orders, personal protection equipment, and social distancing have been the key measures against the spread of SARS-CoV-2 [Citation31]. The inclusion of behavioural changes in epidemiological models has grown fast in recent years [Citation12]. We highlight some typical deterministic models in which behavioural change is incorporated into contact rates since a similar approach will be used in the current study. As early as 1973, London and Yorke [Citation25] explored the mechanism of periodic disease outbreaks by modelling seasonal variation in contact rate due to the opening and closing of school. To consider the psychological effects, Capasso and Serio [Citation7] introduced an SIR epidemic model with a nonlinear force of infection depending on the instantaneous number of infectives. Hsu and Hsieh [Citation17] proposed a SARS model with quarantine and other intervention measures where the contact rate decreases with respect to the cumulative number of probable cases representing behaviour change by individuals. Liu et al. [Citation24] considered another type of SARS model in which the transmission coefficient is an exponential decreasing function of the reported numbers of the exposed, infectious and hospitalized individuals. They showed that the presence of media/psychological impact could induce multiple outbreaks or even sustained periodic oscillations. To evaluate the effects of information and education campaigns on the HIV epidemic in Uganda, Joshi et al. [Citation19] divided the susceptible population into three categories according to behaviour changes and added a compartment to reflect the amount of education campaigns. Cui et al. [Citation10] analysed an SIS model where the contact rate is decreasing in the number of infected individuals and found that contact reduction driven by media coverage cannot change disease dynamics but can reduce the endemic level. Gao and Ruan [Citation16] considered the SIS model with media coverage in a patchy environment and showed that increasing media coverage in any patch will reduce the infection size in each patch. Brauer [Citation5] proposed a simple SIR epidemic model with different constant contact reductions for susceptible and infectious members. Misra et al. [Citation30] established a model that divides the susceptible class into aware and unaware groups and includes the time delay in execution of awareness programmes. Xiao et al. [Citation45] proposed an SIR model where the media effect is described by a piecewise smooth function that depends on the number of cases and its change rate. Based on an SEIR model, Collinson and Heffernan [Citation9] compared the effects of different media functions on epidemic outcomes. Wang et al. [Citation42] incorporated behaviour change into a cholera model via disease contact rates and host shedding rate. The role of behaviour change in containing disease spread has received considerable attention in modelling COVID-19 transmission [Citation15].

Behaviour changes are common in the public health response to HFMD in China. For example, during the high transmission season, more frequent disinfection and sanitation practices and more careful health checks are performed in preschools and kindergartens. Children are encouraged to wash their hands frequently and avoid cross-contamination before eating and after playing outdoors. In the case of a serious epidemic, public health, education and communication departments work closely and run health campaigns intensively and release situation reports regularly. When a case is identified on campus, the patient is required to be treated at hospital or stayed at home and the entire class may be quarantined. The aim of this paper is to investigate the influence of behaviour change on the spread of HFMD. Our paper is structured as follows. In Section 2, we develop a mathematical model with status-dependent behaviour change to describe HFMD transmission. Section 3 establishes the threshold dynamics of the model system in terms of the basic reproduction number. In Section 4, data fitting and numerical simulations are conducted to further analyse the consequence of different control strategies. We conclude the paper in Section 5 with a discussion of our main findings and their implications.

2. Model formulation

The incidence rates of HFMD vary significantly among different age groups [Citation18]. The vast majority of HFMD reported cases are among infants and children. Thus, we only consider the population consisting of children under 12 in this study. Since the case fatality rate of HFMD is very low, especially after the distribution of EV71 vaccines, we ignore the disease-induced mortality. The total population, N(t), is divided into six mutually exclusive classes S(t),E(t),Is(t),Ia(t),Q(t) and R(t) denoting the number of susceptible, exposed, symptomatically infected, mildly or asymptomatically infected, quarantined and recovered individuals at time t, respectively. So, N(t)=S(t)+E(t)+Is(t)+Ia(t)+Q(t)+R(t).A susceptible individual can get infected through contact with symptomatic, asymptomatic or quarantined cases. After the incubation period, an exposed individual may or may not develop symptoms. Symptomatic cases will be quarantined after being diagnosed. Quarantine can reduce transmission but cannot perfectly prevent it. People who recovered from infections are assumed to acquire permanent immunity due to the lower incidence rate in elder children. The HFMD transmission is illustrated in Figure .

Figure 2. Flow chart for the HFMD model.

Figure 2. Flow chart for the HFMD model.
Let US,UE,UIs,UIa,UQ and UR denote the contact rate of susceptible, exposed, symptomatic, asymptomatic, quarantined and recovered individuals, respectively. The normal contact rate under no behaviour change is denoted by c. In reality, an individual's response to an epidemic is affected by the person's health status [Citation5]. People without symptoms or ‘healthy people’ (including susceptible, exposed, asymptomatic and recovered classes) are worried about contracting the disease, so they tend to take more protective measures as the epidemic is getting worse. Meanwhile, symptomatic cases have a high probability of being confirmed and their behaviours may no longer change with the progression of the epidemic. Based on these observations, we assume that US=UE=UIa=UR=f(Is,Q)c,UIs=α1c,UQ=α2c,where f(Is,Q) is a continuously differentiable, positive and decreasing function of Is and Q, and α1(0,1) and α2(0,1) are the relative contact rate of the symptomatically infected and quarantined populations, respectively. In other words, symptomatically infected, quarantined and the remaining individuals decrease their contact rate by a fraction of α1,α2 and f(Is,Q), respectively. The level of behaviour change is governed by the severity of an epidemic which is closely related to the number of reported cases and its change rate [Citation45]. Since asymptomatic infections are hard to detect, only symptomatic cases are assumed to be reported. The cumulative number of reported cases that are still infectious is Q and the number of newly reported cases per unit time is κIs. Hence, we suppose that the behaviour change function f for healthy people depends on Is and Q. In particular, f(0,0)=1, i.e. no behaviour change occurs if there is no symptomatic or quarantined case (no reported case). Denote the transmission probability from an infectious individual to a susceptible individual per contact by p. For convenience, we assume the symptomatic and asymptomatic cases have the same infectivity. The theoretical results throughout this paper remain valid if otherwise. Thus, the usual transmission coefficient is β=cp. The total number of contacts over the entire population per unit time is U=USS+UEE+UIsIs+UIaIa+UQQ+URR=f(Is,Q)c(S+E+Ia+R)+α1cIs+α2cQ=f(Is,Q)cN+(α1f(Is,Q))cIs+(α2f(Is,Q))cQ.Since the sum of Is and Q only accounts for a small proportion of the total population, the approximation US/U1/N holds. Accordingly, the forces of infection driven by symptomatically infected, asymptomatically infected and quarantined classes are USUIsIsUpα1cIsNp=α1βIsN,USUIaIaUpf(Is,Q)cIaNp=f(Is,Q)βIaN,USUQQUpα2cQNp=α2βQN,respectively. The above derivation and the flow chart lead to the following endemic model with nonnegative initial conditions for HFMD transmission (1) dSdt=Λλ(S,E,Is,Ia,Q,R)SμS,dEdt=λ(S,E,Is,Ia,Q,R)S(σ+μ)E,dIsdt=ρσE(κ+γ1+μ)Is,dIadt=(1ρ)σE(γ2+μ)Ia,dQdt=κIs(γ3+μ)Q,dRdt=γ1Is+γ2Ia+γ3QμR,(1) where λ(S,E,Is,Ia,Q,R)=βα1Is+f(Is,Q)Ia+α2QNis the force of infection for the susceptible population. All model parameters except f(Is,Q) are constants with their definitions and ranges summarized in Table .

Adding all equations of model (Equation1) gives dNdt=ΛμN.Since N(t)=Λμ+(N(0)Λμ)eμtN:=Λμ as t, it follows from the theory of asymptotically autonomous systems (see e.g. Castillo-Chavez and Thieme [Citation8]) that system (Equation1) can be reduced to (2) dSdt=Λλ(S,E,Is,Ia,Q)SμS,dEdt=λ(S,E,Is,Ia,Q)S(σ+μ)E,dIsdt=ρσE(κ+γ1+μ)Is,dIadt=(1ρ)σE(γ2+μ)Ia,dQdt=κIs(γ3+μ)Q,(2) where λ(S,E,Is,Ia,Q)=βα1Is+f(Is,Q)Ia+α2QN and N=Λμ. We can easily prove that system (Equation2) is mathematically well-posed and biologically meaningful.

Theorem 2.1

For any initial condition starting in Ω={(S,E,Is,Ia,Q)R+5|S+E+Is+Ia+QΛ/μ}, system (Equation2) has a unique solution that remains in Ω for all time t0.

Table 1. Descriptions and ranges of model parameters (time unit is day).

Proof.

The vector field generated by the right side of model (Equation2) is continuously differentiable in Ω, so it is Lipschitz continuous. Thus, given any nonnegative initial condition, there exists a unique solution for all t0. If S = 0, then S=Λ>0. It follows from Proposition B.7 in Smith and Waltman [Citation36] that S(t)0. The nonnegativity of other state variables can be similarly proved. If S+E+Is+Ia+Q=Λ/μ, then S+E+Is+Ia+Q=(γ1Is+γ2Ia+γ3Q)0. This means that Ω is positively invariant with respect to system (Equation2).

3. Mathematical analysis

In this section, we use the next generation matrix method [Citation11] to define the basic reproduction number of model (Equation2) and then establish the threshold dynamic result for the model.

3.1. Basic reproduction number

Setting the right-hand side of system (Equation2) to zero gives a unique disease-free equilibrium E0=(Λ/μ,0,0,0,0).Using the recipe of van den Driessche and Watmough [Citation39], we get F=(λ(t)S000)andV=((σ+μ)EρσE+(κ+γ1+μ)Is(1ρ)σE+(γ2+μ)IaκIs+(γ3+μ)Q).Linearizing F and V at E0 gives the incidence and transition matrices F=(0α1ββα2β000000000000)andV=(σ+μ000ρσκ+γ1+μ00(1ρ)σ0γ2+μ00κ0γ3+μ),respectively. The basic reproduction number of model (Equation2) is defined as the spectral radius of the next generation matrix FV1, i.e. R0=ρ(FV1)=R1+R2+R3,where R1=σσ+μρβα11κ+γ1+μ,R2=σσ+μ(1ρ)β1γ2+μ,R3=σσ+μρκκ+γ1+μβα21γ3+μrepresent the average number of secondary cases produced by one infected individual in the person's symptomatic, asymptomatic and quarantined stages, respectively. In R1, the ratio σ/(σ+μ) is the probability that an exposed individual will survive the incubation period and become infectious, ρ is the fraction of exposed individuals who will develop symptoms, βα1 is the transmission coefficient for symptomatically infected individuals, and 1/(κ+γ1+μ) is the average infectious period of a symptomatic individual. In R2, the term 1ρ is the fraction of exposed individuals who will be symptomless, β is the transmission coefficient for the asymptomatically infected individuals, and 1/(γ2+μ) is the average infectious period of an asymptomatic individual. In R3, the ratio κ/(κ+γ1+μ) is the probability that a symptomatic individual will be quarantined, βα2 is the transmission coefficient for quarantined individuals, and 1/(γ3+μ) is the average infectious period of a quarantined individual. Clearly, only behaviour responses from symptomatically infected and quarantined people affect the basic reproduction number.

3.2. Threshold dynamics

In what follows, we study the disease dynamics of model (Equation2) by Lyapunov method.

Theorem 3.1

For model (Equation2), if R01, then the disease-free equilibrium E0 is globally asymptotically stable; if R0>1, then the disease-free equilibrium E0 is unstable.

Proof.

Following Theorem 2 in van den Driessche and Watmough [Citation39], the local asymptotic stability of the disease-free equilibrium E0 can be immediately obtained. Thus, it suffices to show the global attractivity of E0 in Ω as R01. We construct a Lyapunov function L=E+h1Is+h2Ia+h3Q,where constants h1,h2 and h3 are yet to be determined. Differentiating L along (Equation2) gives L=E+h1Is+h2Ia+h3Q=βα1Is+f(Is,Q)Ia+α2QNS(σ+μ)E+h1Is+h2Ia+h3Q.The facts SN and f(Is,Q)1 indicate that Lβ(α1Is+Ia+α2Q)(σ+μ)E+h1(ρσE(κ+γ1+μ)Is)+h2((1ρ)σE(γ2+μ)Ia)+h3(κIs(γ3+μ)Q).Combining like terms yields (3) L(h1ρσ+h2(1ρ)σ(σ+μ))E+(βα1h1(κ+γ1+μ)+h3κ)Is+(βh2(γ2+μ))Ia+(βα2h3(γ3+μ))Q.(3) To find appropriate constants h1,h2 and h3 so that the last three terms on the right-hand side of (Equation3) equal zero, we set βα1h1(κ+γ1+μ)+h3κ=0,βh2(γ2+μ)=0,βα2h3(γ3+μ)=0.Solving the above gives (h1,h2,h3)=(βα1κ+γ1+μ+κβα2(κ+γ1+μ)(γ3+μ),βγ2+μ,βα2γ3+μ).Substituting the obtained h1,h2 and h3 into (Equation3), we get L(ρσβα1κ+γ1+μ+(1ρ)σβγ2+μ+ρσκβα2(κ+γ1+μ)(γ3+μ)(σ+μ))E=(R01)(σ+μ)E.If R0<1, then L0 and L=0 if and only if E = 0. The second and third equations of (Equation2) give Is=(κ+γ1+μ)IsandIa=(γ2+μ)Ia,which imply Is=Ia=0. Thus, the last equation of (Equation2) gives Q=(γ3+μ)Q and hence Q = 0. The first equation of (Equation2) becomes S=ΛμS and thus S=Λ/μ. If R0=1, then the equality L=0 implies that βα1Is+f(Is,Q)Ia+α2QNS=β(α1Is+Ia+α2Q).Then S=N and Ia=0 or Is=Q=0, or Is=Ia=Q=0 holds and we can proceed as before. In summary, if R01, then the largest compact invariant subset of {(S,E,Is,Ia,Q)Ω|L=0} is the singleton {E0}. By the LaSalle's invariance principle, the disease-free equilibrium E0 is globally asymptotically stable as R01.

Since the matrix V1F associated to model (Equation2) is reducible, Theorem 2.2 in Shuai and van den Driessche [Citation35] does not work directly. However, by a suitable split of the matrix FV (see Subsection 3.3 in Gao and Cao [Citation14]), we can still use a similar approach to construct an implicit Lyapunov function. Next we consider the disease transmission under R0>1.

Theorem 3.2

Assume that R0>1, then system (Equation2) is uniformly persistent, i.e. there is a constant ε>0 such that each solution φt(x0)=(S(t),E(t),Is(t),Ia(t),Q(t)) with the initial value x0=(S(0),E(0),Is(0),Ia(0),Q(0))Ω satisfies lim inft(E(t),Is(t),Ia(t),Q(t))(ε,ε,ε,ε),where E(0)+Is(0)+Ia(0)+Q(0)>0.

Proof.

We use Theorem 4.6 in Thieme [Citation37] to prove the persistence. Denote Ω0={(S,E,Is,Ia,Q)Ω|E>0,Is>0,Ia>0,Q>0},Ω0=ΩΩ0={(S,E,Is,Ia,Q)Ω|E=0orIs=0orIa=0orQ=0}.Obviously, Ω and Ω0 are positively invariant and Ω0 is a relatively closed subset of Ω. The point dissipativity of system (Equation2) can be seen from Theorem 2.1. Set M={x0Ω0|φt(x0)Ω0,t0},D={(S,0,0,0,0)Ω|S0}.We claim that M=D. Note that DM, so we only need to prove MD. If x0Ω0D, then E(0)+Is(0)+Ia(0)+Q(0)>0. It follows from the irreducibility of FV that φt(x0)Ω0 for t>0. Therefore, we have φt(x0)Ω0 and x0M which implies M=D.

The only equilibrium contained in M=D is E0. So x0Mω(x0)={E0}. Thus, {E0} is a compact and isolated invariant set in M. Let Ws(E0) be the stable manifold of E0. It remains to prove that Ws(E0)Ω0= when R0>1. Suppose not, then there exists x0Ω0 such that φt(x0)E0 as t+. Define M(ξ)=FVξK=((σ+μ)(12ξ)α1β(12ξ)β(12ξ)α2βρσ(κ+γ1+μ)00(1ρ)σ0(γ2+μ)00κ0(γ3+μ)),where K=(02α1β2β2α2β000000000000).Using Theorem 2 in van den Driessche and Watmough [Citation39], we know the spectral bound s(M(0))=s(FV)>0 if and only if R0=ρ(FV1)>1. Since s(M(ξ)) is continuous in ξ, there exists a small ξ1 such that s(M(ξ))>0 for ξ[0,ξ1].

It follows from limtφt(x0)=E0 that, for some T>0, we have φt(x0)E02τ,t>T,where 2 is the Euclidean norm and τ is small enough such that S(t)N1ξ1andf(Is(t),Q(t))1ξ1.Therefore, when t>T, we have dEdt(α1βIs+(1ξ1)βIa+α2βQ)(1ξ1)(σ+μ)E(12ξ1)(α1βIs+βIa+α2βQ)(σ+μ)E,which implies dEdt(12ξ1)(α1βIs+βIa+α2βQ)(σ+μ)E,dIsdt=ρσE(κ+γ1+μ)Is,dIadt=(1ρ)σE(γ2+μ)Ia,dQdt=κIs(γ3+μ)Q.Consider an auxiliary system (4) u=M(ξ1)u,(4) where u=(u1,u2,u3,u4)T=(E,Is,Ia,Q)T. Since M(ξ1) is essentially nonnegative and irreducible, by the Perron–Frobenius theorem, it has a positive eigenvector associated to s(M(ξ1))>0. Thus, any positive solution of system (Equation4) satisfies ui(t)+ as t+, i = 1, 2, 3, 4. Applying the comparison principle [Citation36], we get limt+(E(t),Is(t),Ia(t),Q(t))=(+,+,+,+).This gives a contradiction. So Ws(E0)Ω0=. By Theorem 4.6 in Thieme [Citation37], the system is uniformly persistent if R0>1.

Thus, R0 is a threshold quantity that determines disease extinction and persistence. By Theorem 2.4 in Zhao [Citation47], the uniform persistence of system (Equation2) implies the existence of at least one endemic equilibrium. Furthermore, we can prove the uniqueness of endemic equilibrium in a direct way.

Theorem 3.3

For model (Equation2), there exists exactly one endemic equilibrium if and only if R0>1.

Proof.

By Theorem 3.1, it suffices to show the uniqueness of endemic equilibrium. Denote the endemic equilibrium of (Equation2) by E=(S,E,Is,Ia,Q) which satisfies (5a) Λλ(S,E,Is,Ia,Q)SμS=0,(5a) (5b) λ(S,E,Is,Ia,Q)S(σ+μ)E=0,(5b) (5c) ρσE(κ+γ1+μ)Is=0,(5c) (5d) (1ρ)σE(γ2+μ)Ia=0,(5d) (5e) κIs(γ3+μ)Q=0.(5e) Solving E,Ia and Q from (Equation5c)–(Equation5e) gives (6) E=κ+γ1+μρσIs,Ia=(1ρ)(κ+γ1+μ)ρ(γ2+μ)Is,Q=κγ3+μIs.(6) Substituting (Equation6) into (Equation5b) yields βSIsN(α1+f(Is,Q)(1ρ)(κ+γ1+μ)ρ(γ2+μ)+α2κγ3+μ)=(σ+μ)κ+γ1+μρσIs,which can be rewritten as SN(σ+μ)(κ+γ1+μ)ρσ(R1+f(Is,Q)R2+R3)=(σ+μ)(κ+γ1+μ)ρσ.Thus, we have (7) S=NR1+f(Is,Q)R2+R3.(7) The sum of (Equation5a) and (Equation5b) is (8) ΛμS=(σ+μ)E.(8) Substituting (Equation6) and (Equation7) into (Equation8) yields Is=ΛμSσ+μρσκ+γ1+μ=(ΛμNR1+f(Is,Q)R2+R3)ρσ(σ+μ)(κ+γ1+μ).It follows that f(Is,Q)=(aΛaΛIsR1R3)1R2,where a=ρσ(σ+μ)(κ+γ1+μ).Denote G(Is)=f(Is,κγ3+μIs)(aΛaΛIsR1R3)1R2,Is[0,aΛ).Differentiating G(Is) with respect to Is, we find that G(Is) is strictly decreasing in Is[0,aΛ). Note that G(0)=1(1R1R3)1R2=R1+R2+R31R2=R01R2>0 as R0>1and limIsaΛG(Is)= and aΛ<N=Λμ.Thus, the equation G(Is)=0 has a unique positive root when R0>1.

Remark 3.4

For simple endemic models like SIS, SIAR and SEIRS, the fraction of people being susceptible at the endemic equilibrium (when it exists) usually equals the reciprocal of the basic reproduction number, i.e. S/N=1/R0. However, for model (Equation2), it follows from (Equation7) that SN=1R1+f(Is,Q)R2+R31R1+R2+R3=1R0.Biologically speaking, behaviour change reduces the number of infections. Indeed, (Equation6) indicates that the population size of each nonsusceptible compartment at the endemic equilibrium E becomes smaller when behaviour change occurs. Moreover, it weakens the relative contribution of asymptomatic class in disease transmission. In fact, the ratio of the forces of infections attributed to Is,Ia and Q at E is α1βIsN:f(Is,Q)βIaN:α2βQN=α1Is:f(Is,Q)Ia:α2Q=α1ρσκ+γ1+μE:f(Is,Q)(1ρ)σγ2+μE:α2κγ3+μρσκ+γ1+μEα1ρκ+γ1+μ:1ργ2+μ:κκ+γ1+μα2ργ3+μ=R1:R2:R3,where the second equality is due to the equilibrium Equation (5). Interestingly, in the absence of behaviour change, the relative contributions of infected states Is,Ia and Q to new infections at the endemic equilibrium are the same as their contributions to the basic reproduction number.

Before ending this section, we give a sufficient condition under which the endemic equilibrium is globally attractive. In particular, it is satisfied when f(Is,Q) is constant. The proof is postponed to Appendix A.

Theorem 3.5

Suppose that (f(Is,Q)f(Is,Q))(f(Is,Q)Iaf(Is,Q)Ia)0holds for Is,Ia,Q>0 and Is+Ia+QN=Λ/μ. If R0>1, then the unique endemic equilibrium E of system (Equation2) is globally asymptotically stable in Ω minus the disease-free space.

4. Numerical analysis

In this part, we first fit the proposed epidemiological model (Equation1) to weekly HFMD reported case data in Shanghai, China. Then, we carry out some sensitivity analysis and numerical simulations to compare the effectiveness of different intervention strategies and explore the role of behaviour change on infection control.

4.1. Date fitting

The first HFMD case in mainland China was emerged in Shanghai in 1981. It has been mandatory to report cases of HFMD to Shanghai Municipal Center for Disease Control and Prevention since 2005, three years earlier than the country. Shanghai has established a city-wide communicable disease surveillance system that consists of medical institutions, district and municipal centres for disease control and prevention. Every year tens of thousands of cases are reported in Shanghai and almost all of them are admitted to three specialized hospitals. The usual peak season in Shanghai is from April to July and a smaller peak may occur from September to November. Although Shanghai is one of the richest cities in China, its annual HFMD incidence rate is significantly higher than the national average partially due to meteorological factors and migrant population. For example, the incidence rates of Shanghai and China in 2016 are 237/100,000 and 178/100,000, respectively. Schools and kindergartens implement emergency response plan for childhood infectious diseases and class or school closure is common when a cluster of cases are identified in a short time period.

By using the least-squares method, we fit model (Equation1) to the weekly HFMD case data of Shanghai in 2019 (Zhao et al. [Citation49]). According to the 2019 Shanghai Statistical Yearbook [Citation2], we set Λ=463, μ=2.3×104 and N(0)=2.01×106. To reduce uncertainty, we choose some reasonable parameter values from literature as follows α1=0.5,α2=0.1,ρ=0.4,σ=0.25,κ=0.5,andγ3=0.25,where the time unit is one day. We assume that γ1=γ2=γ, indicating that individuals either symptomatic or asymptomatic will take γ1 days to recover. Since it takes time to put people into quarantine, it is reasonably believed that γ1>γ31, i.e. γ<γ3. The behaviour change function is assumed to take the form (9) f(Is,Q)=11+m1Is+m2Q,(9) where m1>0 and m2>0 measure the influence of newly reported cases and recently cumulative cases on the contact rate of ‘healthy people’, respectively. The initial conditions of model (Equation1) are set as (S(0),E(0),Is(0),Ia(0),Q(0),R(0))=(1.01×106,635,91,682,350,106).Denote the given data set by {(t1,Y1),,(tn,Yn)} where Yi is the newly reported number of cases in week i. The simulated weekly cases are obtained as Zi=ti1tiκIs(t)dt,and the sum-of-squares error is defined as: SSE=i=1n(YiZi)2.We use the open-source R programming language to calibrate the model and estimate the disease transmission coefficient, β, behaviour change parameters, m1 and m2, and recovery rate of infection, γ. Specifically, based on the Levenberg–Marquardt algorithm, we use the nls.lm function from the R package minpack.lm to perform the least-squares fitting. Note that in our model calibration, the transmission coefficient, β, is considered as a time-dependent cubic B-spline function instead of a constant to better capture the seasonality in reported data. The fitted parameter values are m1=3.8×104, m2=7.2×104, and γ=0.14. Figure  illustrates the simulated average weekly cases versus time, and the grey shaded area gives the 95% confidence interval for the number of simulated cases per week, which matches the changing pattern of the reported weekly cases. Hence our model provides a good fit for the time series of the weekly reported HFMD cases. Moreover, the time evolution of the basic reproduction number is estimated and shown in Figure , which indicates that the asymptotically infected individuals contributes the most to the transmission. This is mainly attributed to the rapid and strict quarantine of symptomatic cases. The average basic reproduction number and transmission coefficient are R¯0=2.41 and β¯=0.52, respectively, which indicate that HFMD cannot be eradicated under the current control strategy.

Figure 3. Comparison of reported and simulated weekly HFMD cases of Shanghai in 2019. The red triangles are reported weekly cases, the blue dots are simulated weekly cases, and the grey shaded area gives the 95% confidence interval for the number of simulated cases per week.

Figure 3. Comparison of reported and simulated weekly HFMD cases of Shanghai in 2019. The red triangles are reported weekly cases, the blue dots are simulated weekly cases, and the grey shaded area gives the 95% confidence interval for the number of simulated cases per week.

Figure 4. Time evolution of R0=R1+R2+R3, where R1,R2 and R3 denote the secondary infections produced by one infected individual during symptomatic, asymptomatic and quarantined stages, respectively.

Figure 4. Time evolution of R0=R1+R2+R3, where R1,R2 and R3 denote the secondary infections produced by one infected individual during symptomatic, asymptomatic and quarantined stages, respectively.

4.2. Sensitivity analysis

Since the threshold dynamics of model (Equation2) are wholly governed by the basic reproduction number, reducing R0 to be less than one is desirable for disease eradication. Thus, it is necessary to examine how R0 varies with model parameters. Direct calculations find that R0 is monotone increasing with respect to β,α1,α2 and σ, decreasing in γ1,γ2,γ3 and μ, and independent of Λ and f. Mathematically, the monotonicity of R0 on ρ and κ may change with parameter setting by noting that sgn(R0ρ)=sgn(α1κ+γ1+μ1γ2+μ+κκ+γ1+μα2γ3+μ)and sgn(R0κ)=sgn(α2γ3+μα1γ1+μ).In the real world, the highly possible relations α2<α1<1, γ1<γ3 and γ2<κ+γ1 imply that R0 is probably decreasing in ρ and κ.

To see which way is the best in lowering disease persistence, we have to compare the sensitivity indices of R0 to parameter variation. The normalized forward sensitivity index is widely used and it is defined as the ratio of the relative change of the output to the relative change of the parameter [Citation3]. Specifically, let u be a variable that depends differentially on a parameter s. Then the sensitivity index (SI) of u in terms of s is expressed as Υsu=limΔs0Δuu/Δss=limΔs0ΔuΔssu=ussu.The complex structure of R0 for model (Equation1) makes it difficult to analytically compare the sensitivity indices of different parameters. Therefore, we select a typical set of parameter values based on the above fitting result: (10) β=0.4,α1=0.5,α2=0.1,ρ=0.4,σ=0.25,κ=0.5,γ1=0.14,γ2=0.14,γ3=0.25,μ=2.3×104.(10) The corresponding basic reproduction number is R0=R1+R2+R31.88 with R1=0.12,R2=1.71 and R3=0.05. So, the percentages of contribution of the infected states Is,Ia and Q to R0 are 6.62%, 90.73% and 2.65%, respectively. This suggests that asymptomatic transmission is a leading cause of HFMD persistence. We then calculate their sensitivity indices as shown in Figure (a). Obviously, R0 is most sensitive to the transmission coefficient, β, followed by the recovery rate of asymptomatic infections, γ2, and the proportion of infections being symptomatic, ρ, and is least sensitive to the progression rate from the exposed state to the infectious state, σ, the progression rate leaving children group, μ, and the recovery rate of the symptomatic people, γ1.

Figure 5. Sensitivity analysis for R0 with respect to model parameters: (a) sensitivity indices (SI), and (b) partial rank correlation coefficients (PRCC).

Figure 5. Sensitivity analysis for R0 with respect to model parameters: (a) sensitivity indices (SI), and (b) partial rank correlation coefficients (PRCC).
The conclusion from the sensitivity indices can be biased since it only measures the influence of a single parameter on model output which could be strongly relied on the choice of a specific parameter set. To improve the robustness of sensitivity analysis, we investigate the impact of global parameter variations on model outcome. By using the Latin Hypercube Sampling (LHS) method [Citation29], we generate 106 random parameter sets with ranges in Table . Each input parameter is assumed to be uniformly distributed. Then we calculate the partial rank correlation coefficient (PRCC) of R0 with respect to each involved parameter (see Figure (b)). The global sensitivity analysis gives a similar conclusion, whereas the main difference is that the relative contact rate of the quarantined state, α2, is moderately positively correlated to R0. In reality, it is hard to detect and treat asymptomatic cases and change the symptomatic ratio. This indicates that reducing contacts between susceptible and infectious individuals and implementing strict quarantine are vital to HFMD control.

4.3. Numerical simulations

Example 4.1

Behaviour change on infection size

We notice that the behaviour change function f(Is,Q) does not influence R0. Thus, variable behaviour change has no impact on disease persistence. However, it follows from Remark 3.4 that the change can protect some susceptibles from contracting the disease. To quantitatively examine the role of behaviour change on a HFMD epidemic wave, we consider Λ=463 and the remaining parameter setting is the same as (Equation10). The initial conditions are as follows: (S(0),E(0),Is(0),Ia(0),Q(0),R(0))=(1.11×106,10,0,0,0,9×105).Again the basic reproduction number is R01.88>1. We adopt the same behaviour change function as given in (Equation9) and select three scenarios with different level of behaviour change (BC):

  1. no behaviour change: m1=m2=0;

  2. weak behaviour change: m1=6×104 and m2=3×104;

  3. strong behaviour change: m1=9×103 and m2=4.5×103.

Figure (a) shows the total number of symptomatically infected individuals, i.e. Is+Q, with m1 and m2 increasing from zero to baseline values to 15-fold of them. We can see that the introduction of infectives always leads to an outbreak and the disease spread attains its first peak almost at the same time under no and weak behaviour change. The endemic equilibria E=(S,E,Is,Ia,Q,R) and the nonsusceptible ratios 1SN associated to the three scenarios are: (1.068×106,869,136,929,271,9.427×105)and46.9%,no behaviour change,(1.202×106,746,116,798,233,8.094×105)and40.3%,weak behaviour change,(1.731×106,260,41,278,81,2.819×105)and14.0%,strong behaviour change,respectively. Weak behaviour change can sharply reduce the peak size (the maximum number of symptomatic infections) but moderately lower the endemic level, while strong behaviour change reduces both peak size and endemic level dramatically. When there is no behaviour change, the solution converges to the endemic equilibrium in an oscillating way. Stronger behaviour change results in less or no damped oscillations over time. Figure (b) illustrates that the healthy people eventually reduce their contacts by 12% and 42% through weak and strong behaviour change, respectively. This can also be obtained by noting f(Is(t),Q(t))f(Is,Q) as t. In particular, under strong behaviour change, 42% reduction in the contact rate of healthy people can lead to 70% reduction in symptomatic infections. The respective percentages of contribution of Is,Ia and Q to new infections change from 6.62%, 90.73% and 2.65% (same as R0) to 7.45%, 89.57% and 2.98% to 10.73%, 84.98% and 4.29% as behaviour change scenario changes from (a) to (b) to (c).

Figure 6. Simulations of model (Equation1) for (a) the square root of the total number of symptomatic infections, Is+Q, and (b) the behaviour change function, f(Is,Q), versus time under no behaviour change (red dashed line), weak behaviour change (black solid line), and strong behaviour change (blue dotted line). See main text for parameter values and initial condition.

Figure 6. Simulations of model (Equation1(1) dSdt=Λ−λ(S,E,Is,Ia,Q,R)S−μS,dEdt=λ(S,E,Is,Ia,Q,R)S−(σ+μ)E,dIsdt=ρσE−(κ+γ1+μ)Is,dIadt=(1−ρ)σE−(γ2+μ)Ia,dQdt=κIs−(γ3+μ)Q,dRdt=γ1Is+γ2Ia+γ3Q−μR,(1) ) for (a) the square root of the total number of symptomatic infections, Is+Q, and (b) the behaviour change function, f(Is,Q), versus time under no behaviour change (red dashed line), weak behaviour change (black solid line), and strong behaviour change (blue dotted line). See main text for parameter values and initial condition.

Furthermore, using the same parameter setting and initial condition, the dependences of the total symptomatic infections and the relative contact rate of the healthy people at the endemic equilibrium, i.e. Is+Q and f(Is,Q), against the behaviour change parameters, m1 and m2, are plotted in Figure . Clearly, the more sensitive the public is to an increase in the number of new or recent infections, the lower the number of cases and contacts. Note that the contours of Is+Q and f(Is,Q) are straight lines in the m1-m2 plane. Indeed, given behaviour change parameters m~10 and m~20, denote the corresponding endemic equilibrium by E~=(S~,E~,I~s,I~a,Q~,R~), then f(I~s,Q~)=11+m~1I~s+m~2Q~andκI~s=(γ3+μ)Q~.Therefore, for any pair of parameter values on the line segment L:{(m1,m2)R+2:m1I~s+m2Q~=m~1I~s+m~2Q~},the associated unique endemic equilibrium remains unchanged. More specifically, the line is (m1m~1)I~s+(m2m~2)Q~=0,i.e.(γ3+μ)(m1m~1)+κ(m2m~2)=0.Hence, the two quantities Is+Q and f(Is,Q) are constant on L.

Example 4.2

Quarantine on infection size

We now examine the impact of quarantine on symptomatic cases. Note that quarantine strategy is characterized by κ and α2, which are related to the speed and efficiency of quarantine. There is yet no specific treatment for HFMD, so we assume that quarantine does not accelerate the recovery process.

Figure 7. The contour plots of the total number of symptomatic infections and the behaviour change function at the endemic equilibrium, Is+Q and f(Is,Q), versus parameters m1 and m2. See main text for parameter values.

Figure 7. The contour plots of the total number of symptomatic infections and the behaviour change function at the endemic equilibrium, Is∗+Q∗ and f(Is∗,Q∗), versus parameters m1 and m2. See main text for parameter values.

Consider the same set of parameter values as given in Example 4.1 with the exception of κ and α2. The initial conditions of model (Equation1) are (S(0),E(0),Is(0),Ia(0),Q(0),R(0))=(2.01×106,10,0,0,0,0).For fixed α2=0.1, the total numbers of symptomatic cases Is+Q over time are plotted in Figure (a) under the three behaviour change scenarios (a)–(c) as κ increases from 0.1 to 0.3 to 0.9. The basic reproduction numbers R0 correspond to κ=0.1,0.3 and 0.9 are 2.07, 1.93 and 1.84, respectively. When there is no change in behaviour, fast quarantine reduces the peak size and the epidemic size of symptomatic infections considerably and delays the peak time (the time of the maximum number of infected humans) by 2-4 weeks. In the presence of behaviour change, fast quarantine only has a mild effect on disease propagation. Interestingly, it suggests that a higher but shorter epidemic (red curves) may have a larger epidemic size than a lower but much longer epidemic (blue or black curves). Meanwhile, given κ=0.3, Figure (b) shows how the total numbers of symptomatic infections evolve under different levels of behaviour change as the value of α2 decreases from 0.4 to 0.1 to 0.025. So, improving the efficiency of quarantine plays a similar but weaker role in containing disease spread which means that timely quarantine is more important than strict quarantine. It is worth mentioning that quarantine significantly affects the infection size even though it has little impact on R0.

Figure 8. Simulations of model (Equation1) for the cube root of the total number of symptomatic infections, Is+Q3, under different behaviour change and (a) quarantine rate κ=0.1, 0.3 and 0.9, and (b) relative contact rate for the quarantined individuals α2=0.4, 0.1 and 0.025.

Figure 8. Simulations of model (Equation1(1) dSdt=Λ−λ(S,E,Is,Ia,Q,R)S−μS,dEdt=λ(S,E,Is,Ia,Q,R)S−(σ+μ)E,dIsdt=ρσE−(κ+γ1+μ)Is,dIadt=(1−ρ)σE−(γ2+μ)Ia,dQdt=κIs−(γ3+μ)Q,dRdt=γ1Is+γ2Ia+γ3Q−μR,(1) ) for the cube root of the total number of symptomatic infections, Is+Q3, under different behaviour change and (a) quarantine rate κ=0.1, 0.3 and 0.9, and (b) relative contact rate for the quarantined individuals α2=0.4, 0.1 and 0.025.

5. Discussion

HFMD is a childhood contagious disease that spreads widely in China. In recent years, a number of mathematical modelling studies on HFMD have been done with the consideration of seasonal variation in contact pattern, vaccine introduction, environmental contamination and so on [Citation23, Citation34, Citation43]. During a disease outbreak, changes in behaviour driven by governmental action or individual reaction are very common. Their role in disease control and elimination has been increasingly considered by modellers [Citation12, Citation19, Citation30]. In this paper, we proposed a deterministic endemic model with imperfect quarantine to explore the impact of behaviour change on the spread of HFMD. The total population is split into six compartments where the symptomatic and quarantined compartments obey constant behaviour change and the behaviour change of other compartments depends on the number of newly reported cases and recently cumulative cases. We defined a biologically feasible region and derived the basic reproduction number R0 of the model. Using the Lyapunov functional method and the persistence theory, a threshold dynamic result was obtained, i.e. the disease-free equilibrium is globally asymptotically stable as R01 and the disease is uniformly persistent otherwise. Moreover, there exists a unique endemic equilibrium if R0>1 and a sufficient condition for the global stability of the endemic equilibrium was given.

In addition, we fitted our model to the reported HFMD case data in Shanghai and found that the basic reproduction number is around 2.41. Based on the fitting result and parameter ranges, we conducted both local and global sensitivity analysis for R0 in terms of model parameters. Control measures that reduce close contacts and improve quarantine adherence can be successful in constraining HFMD. Two numerical examples were presented to analyse the influence of behaviour change and quarantine on reducing disease burden. In the first example, we compared the total numbers of symptomatic infections under different behaviour changes. Although incidence-based or prevalence-based behaviour change of healthy people does not affect the disease persistence, it substantially lowers the endemic level and flattens the curve so that the health care capacity can meet the demands during an outbreak. Ignoring behaviour change can potentially overestimate the disease magnitude and the contribution of asymptomatic transmission. The stronger the behaviour change to an epidemic, the smaller the epidemic size. Thus, it is beneficial to run more health education campaigns during high transmission season. In the second example, we demonstrated that fast quarantine of symptomatic cases can significantly curtail the infection size and delay the peak time in the absence of behaviour change. Nevertheless, behaviour change may weaken the benefit of quarantine.

To some extent, we generalize the HFMD model of Ma et al. [Citation27] by considering behaviour change and imperfect quarantine. Although our modelling and analysis are based on HFMD, they are applicable to some other infectious diseases such as measles and mumps. Most existing epidemic models with behaviour change either assume that all individuals have the same behaviour change depending on the number of infected individuals [Citation7, Citation10, Citation42], or divide the total population into multiple groups and different groups have different but constant behaviour change [Citation5, Citation30, Citation46]. The main novelty of our model is that some states/individuals have variable behaviour change while others have constant behaviour change. Meanwhile, we assume that the variable behaviour change is determined by the number of confirmed cases and its change rate. The idea is somewhat similar to that of Xiao et al. [Citation45] but we use the number of symptomatic cases to replace the change rate of infected people which facilitates the mathematical analysis. A related form of contact rate was introduced by Liu et al. [Citation24] but the epidemiological meaning is different. We found that the transmission from mild or asymptomatic patients plays a dominant role in maintaining HFMD spreading. This result is surprising and arguable [Citation34]. On the one hand, symptomatic cases of HFMD are often easy to identify, sharply reducing their transmission potential. On the other hand, the presumed high asymptomatic ratio and the equal infectivity for the asymptomatic and symptomatic individuals may exaggerate the contribution of asymptomatic transmission.

There are many possibilities to improve and generalize the current work. The global asymptotic stability of the endemic equilibrium is generally unclear. Fitting the model to long term rather than one-year data is challenging but more convincing. Seasonal epidemics of HFMD indicate that a periodic epidemic model with time-dependent transmission coefficient may be favourable [Citation23]. Some additional biological factors like environment-to-human transmission route [Citation42], coexistence of multiple enteroviruses [Citation32], EV71 vaccination [Citation34], population movement [Citation16], waning immunity, age structure, stochasticity, and varying population size can be added to the model. A full understanding of how the emergency response plan is implemented in schools and kindergartens can help us to build a more realistic model. How to mechanistically determine and quantify the behaviour change function? To what extent can behaviour change reduce the final size of an epidemic as vital dynamics are ignored (see Brauer [Citation6])? Comparing the constant behaviour change, variable behaviour change and mixed behaviour change approaches with epidemiological and behavioural data is necessary for the selection and application of behaviour change models. It is desirable to develop a general modelling framework with heterogeneities in behaviour change. Since R0 is defined at the disease-free equilibrium at which there is no behaviour change, the reproduction number or disease dynamics are unaffected by behaviour change. In reality, long-term behaviour change may eradicate an infectious disease in a specific region. One way to construct a behaviour change function that still works when the initial conditions are near the disease-free space is to assume that the contact rate depends on the cumulative number of cases [Citation17]. Changes in behaviour are not limited to contact rate but also include vaccination rate, recruitment rate [Citation40], movement rate [Citation26] and so on. Besides the case and death counts, some other factors can strengthen or weaken behaviour change. The ongoing COVID-19 pandemic indicates that individual response to epidemic waves varies significantly with vaccine development and distribution. The biennial cycle of HFMD incidence indicates that there should be more cases in 2020 than in 2019. However, due to the COVID-19 outbreak, there were 5,585 reported cases of HFMD in 2020 compared to 24,615 cases in 2019 in Shanghai while there were 761,355 reported HFMD cases in 2020 compared to 1,918,830 cases in 2019 in China [Citation1, Citation49]. It is attractive to understand the transmission dynamics of HFMD under the COVID-19-induced behaviour change. We must realize that behaviour change like limiting contact and travel could be costly and unsustainable. A cost-effectiveness analysis of behaviour change measures is indispensable. In the era of mobile internet and artificial intelligence, the rapid development in data collection, storage, and extraction methods makes behaviour change models for infectious diseases more applicable but there are still quite a few challenges to overcome [Citation13].

Acknowledgments

We thank the two referees for their valuable comments, Dr. Jim Cushing for handling the submission, and Drs. Jifa Jiang and Yijun Lou for helpful suggestions on the draft.

Disclosure statement

No potential conflict of interest was reported by the author(s).

Additional information

Funding

This work was supported by the Natural Science Foundation of Shanghai [grant numbers 20ZR1440600 and 20JC1413800], the National Natural Science Foundation of China [grant number 12071300], the Key Research Projects of Beijing Natural Science Foundation-Haidian District Joint Fund [grant number L192012], and the CSU Office of Research through a startup grant.

References

  • National Administration of Disease Control and Prevention, Reported Cases and Deaths of National Notifiable Infectious Diseases – China, 2020. http://www.nhc.gov.cn/jkj/s3578/202103/f1a448b7df7d4760976fea6d55834966.shtml, 2021.
  • Shanghai Municipal Bureau of Statistics and Survey Office of the National Bureau of Statistics in Shanghai, 2020 Shanghai Statistical Yearbook. http://tjj.sh.gov.cn/tjnj/index.html, 2020.
  • L. Arriola and J.M. Hyman, Senditivity analysis for uncertainty quantification in mathematical models, in Mathematical and Statistical Estimation Approaches in Epidemiology, G. Chowell, J. M. Hyman, L. M. A. Bettencourt, and C. Castillo-Chavez, eds., Springer, New York, 2009, pp. 195–247.
  • M.C.J. Bootsma and N.M. Ferguson, The effect of public health measures on the 1918 influenza pandemic in U.S. cities, Proc. Natl. Acad. Sci. U.S.A. 104 (2007), pp. 7588–7593.
  • F. Brauer, A simple model for behaviour change in epidemics, BMC Publ. Health 11 (2011), pp. 53.
  • F. Brauer, The final size of a serious epidemic, Bull. Math. Biol. 81 (2019), pp. 869–877.
  • V. Capasso and G. Serio, A generalization of the Kermack-McKendrick deterministic epidemic model, Math. Biosci. 42 (1978), pp. 43–61.
  • C. Castillo-Chavez and H.R. Thieme, Asymptotically autonomous epidemic models, in Mathematical Population Dynamics: Analysis of Heterogeneity, O. Arino, D. E. Axelrod, M. Kimmel, and M. Langlais, eds., Wuerz, Winnipeg, Canada, 1995, pp. 33–50.
  • S. Collinson and J.M. Heffernan, Modelling the effects of media during an influenza epidemic, BMC Publ. Health 14 (2014), pp. 376.
  • J. Cui, X. Tao, and H. Zhu, An SIS infection model incorporating media coverage, Rocky Mountain J. Math. 38 (2008), pp. 1323–1334.
  • O. Diekmann, J.A.P. Heesterbeek, and J.A.J. Metz, On the definition and the computation of the basic reproduction ratio R0 in models for infectious diseases in heterogeneous populations, J. Math. Biol.28 (1990), pp. 365–382.
  • S. Funk, M. Salathé, and V.A.A. Jansen, Modelling the influence of human behaviour on the spread of infectious diseases: A review, J. R. Soc. Interface 7 (2010), pp. 1247–1256.
  • S. Funk, S. Bansal, C.T. Bauch, K.T.D. Eames, W.J. Edmunds, A.P. Galvani, and P. Klepac, Nine challenges in incorporating the dynamics of behaviour in infectious diseases models, Epidemics 10 (2015), pp. 21–25.
  • D. Gao and L. Cao, Vector-borne disease models with Lagrangian approach. under review, 2021.
  • D. Gao and D. He, Special issue: Modeling the biological, epidemiological, immunological, molecular, virological aspects of COVID-19, Math. Biosci. Eng. 18 (2021), pp. 983–985.
  • D. Gao and S. Ruan, An SIS patch model with variable transmission coefficients, Math. Biosci. 232 (2011), pp. 110–115.
  • S. Hsu and Y. Hsieh, Modeling intervention measures and severity-dependent public response during severe acute respiratory syndrome outbreak, SIAM J. Appl. Math. 66 (2006), pp. 627–647.
  • T. Ji, T. Han, X. Tan, S. Zhu, D. Yan, Q. Yang, Y. Song, A. Cui, Y. Zhang, N. Mao, S. Xu, Z. Zhu, D. Niu, Y. Zhang, and W. Xu, Surveillance, epidemiology, and pathogen spectrum of hand, foot, and mouth disease in mainland of China from 2008 to 2017, Biosaf. Health 1 (2019), pp. 32–40.
  • H. Joshi, S. Lenhart, K. Albright, and K. Gipson, Modeling the effect of information campaigns on the HIV epidemic in Uganda, Math. Biosci. Eng. 5 (2008), pp. 757–770.
  • J.T.F. Lau, X. Yang, H. Tsui, and J.H. Kim, Monitoring community responses to the SARS epidemic in Hong Kong: From day 10 to day 62, J. Epidemiol. Community Health 57 (2003), pp. 864–870.
  • Y. Li, J. Zhang, and X. Zhang, Modeling and preventive measures of hand, foot and mouth disease (HFMD) in China, Int. J. Environ. Res. Public Health 11 (2014), pp. 3108–3117.
  • Y. Li, L. Wang, L. Pang, and S. Liu, The data fitting and optimal control of a hand, foot and mouth disease (HFMD) model with stage structure, Appl. Math. Comput. 276 (2016), pp. 61–74.
  • J. Liu, Threshold dynamics for a hfmd epidemic model with periodic transmission rate, Nonlinear Dyn. 64 (2011), pp. 89–95.
  • R. Liu, J. Wu, and H. Zhu, Media/psychological impact on multiple outbreaks of emerging infectious diseases, Comput. Math. Methods Med. 8 (2007), pp. 153–164.
  • W.P. London and J.A. Yorke, Recurrent outbreaks of measles, chickenpox and mumps. I. Seasonal variation in contact rates, Am. J. Epidemiol. 98 (1973), pp. 453–468.
  • M. Lu, D. Gao, J. Huang, and H. Wang, Relative prevalence-based dispersal in an epidemic patch model, J. Math. Biol. 86 (2023), pp. 52.
  • Y. Ma, M. Liu, Q. Hou, and J. Zhao, Modelling seasonal HFMD with the recessive infection in Shandong, China, Math. Biosci. Eng. 10 (2013), pp. 1159–1171.
  • P. Manfredi and A. d'Onofrio, Modeling the Interplay Between Human Behavior and the Spread of Infectious Diseases, Springer, New York, 2013.
  • S. Marino, I.B. Hogue, C.J. Ray, and D.E. Kirschner, A methodology for performing global uncertainty and sensitivity analysis in systems biology, J. Theor. Biol. 254 (2008), pp. 178–196.
  • A. Misra, A. Sharma, and V. Singh, Effect of awareness programs in controlling the prevalence of an epidemic with time delay, J. Biol. Syst. 19 (2011), pp. 389–402.
  • N. Perra, Non-pharmaceutical interventions during the COVID-19 pandemic: A review, Phys. Rep.913 (2021), pp. 1–52.
  • M. Pons-Salort and N.C. Grassly, Serotype-specific immunity explains the incidence of diseases caused by human enteroviruses, Science 361 (2018), pp. 800–803.
  • N. Roy and N. Halder, Compartmental modeling of hand, foot and mouth infectious disease (HFMD), Res. J. Appl. Sci. 5 (2010), pp. 177–182.
  • L. Shi, H. Zhao, and D. Wu, Modelling and analysis of HFMD with the effects of vaccination, contaminated environments and quarantine in mainland China, Math. Biosci. Eng. 16 (2019), pp. 474–500.
  • Z. Shuai and P. van den Driessche, Global stability of infectious disease models using Lyapunov functions, SIAM J. Appl. Math. 73 (2013), pp. 1513–1532.
  • H.L. Smith and P. Waltman, The Theory of the Chemostat: Dynamics of Microbial Competition, Cambridge University Press, Cambridge, 1995.
  • H.R. Thieme, Persistence under relaxed point-dissipativity (with application to an endemic model), SIAM J. Math. Anal. 24 (1993), pp. 407–435.
  • F.C.S. Tiing and J. Labadin, A simple deterministic model for the spread of hand, foot and mouth disease (HFMD) in Sarawak, in Second Asia International Conference on Modeling & Simulation, 2008, pp. 947–952.
  • P. van den Driessche and J. Watmough, Reproduction numbers and sub-threshold endemic equilibria for compartmental models of disease transmission, Math. Biosci. 180 (2002), pp. 29–48.
  • J.X. Velasco-Hernández, F. Brauer, and C. Castillo-Chavez, Effects of treatment and prevalence-dependent recruitment on the dynamics of a fatal disease, IMA J. Math. Appl. Med. 13 (1996), pp. 175–192.
  • Y. Wang and F. Sung, Modeling the infections for enteroviruses in Taiwan. preprint, Available at, 2004.
  • X. Wang, D. Gao, and J. Wang, Influence of human behavior on cholera dynamics, Math. Biosci. 267 (2015), pp. 41–52.
  • J. Wang, Y. Xiao, and R.A. Cheke, Modelling the effects of contaminated environments on HFMD infections in mainland China, Biosystems 140 (2016), pp. 1–7.
  • J. Wang, Z. Teng, X. Cui, C. Li, H. Pan, Y. Zheng, S. Mao, Y. Yang, L. Wu, X. Guo, X. Zhang, and Y. Zhu, Epidemiological and serological surveillance of hand-foot-and-mouth disease in Shanghai, China, 2012–2016, Emerg. Microbes Infect. 7 (2018), pp. 8.
  • Y. Xiao, T. Zhao, and S. Tang, Dynamics of an infectious diseases with media/psychology induced non-smooth incidence, Math. Biosci. Eng. 10 (2013), pp. 445–461.
  • C. Yang, X. Wang, D. Gao, and J. Wang, Impact of awareness programs on cholera dynamics: Two modeling approaches, Bull. Math. Biol. 79 (2017), pp. 2109–2131.
  • X.-Q. Zhao, Uniform persistence and periodic coexistence states in infinite dimensional periodic semiflows with applications, Can. Appl. Math. Quart. 3 (1995), pp. 473–495.
  • H. Zhao, L. Shi, J. Wang, and K. Wang, A stage structure hfmd model with temperature-dependent latent period, Appl. Math. Model. 93 (2021), pp. 745–761.
  • Z. Zhao, C. Zheng, H. Qi, Y. Chen, M.P. Ward, F. Liu, J. Hong, Q. Su, J. Huang, X. Chen, J. Le, X.Liu, M. Ren, J. Ba, Z. Zhang, Z. Chang, and Z. Li, Impact of the coronavirus disease 2019 interventions on the incidence of hand, foot, and mouth disease in mainland China, Lancet Reg. Health West Pac.20 (2022), Article ID 100362.

Appendix. Proof of Theorem 3.5

Proof.

Let E=(S,E,Is,Ia,Q) denote the endemic equilibrium of model (Equation2) which satisfies (5). Define V1=Sg(SS)+Eg(EE),V2=Isg(IsIs),V3=Iag(IaIa),V4=Qg(QQ),where g(x)=x1lnx0 with equality if and only if x = 1.

Consider a Lyapunov function V=l1V1+l2V2+l3V3+l4V4,where positive constants li, i = 1, 2, 3, 4 are to be determined. The derivative of V along the solution of system (Equation2) is V=l1(1SS)S+l1(1EE)E+l2(1IsIs)Is+l3(1IaIa)Ia+l4(1QQ)Q.For convenience, we write f=f(Is,Q) and f=f(Is,Q). Using (Equation5a) and (Equation5b), V1=(1SS)(βα1Is+fIa+α2QNS+μSβα1Is+fIa+α2QNSμS)+(1EE)(βα1Is+fIa+α2QNSβα1Is+fIa+α2QNSEE)=μ(SS)2S+α1βIsSN(2+IsIsSSEESIsESIsE)+fβIaSN(2+fIafIaSSEEfSIaEfSIaE)+α2βQSN(2+QQSSEESQESQE).Then we have 2+IsIsSSEESIsESIsE=(1SIsESIsE)+(1SS)+IsIsEElnSIsESIsE+lnSS+IsIsEE=lnIsEIsE+IsIsEE=lnEElnIsIs+IsIsEE,where the equality holds if and only if S=S and EE=IsIs. Similarly, 2+QQSSEESQESQElnEElnQQ+QQEEwith equality if and only if S=S and EE=QQ. By assumption, 2+fIafIaSSEEfSIaEfSIaE=(1fIafIa)(1ff)+3SSfSIaEfSIaEffEE+IaIa(SS1)(fSIaEfSIaE1)(ff1)EE+IaIaln(SSfSIaEfSIaEff)EE+IaIa=lnEElnIaIa+IaIaEEwith equality if and only if S=S, EE=IaIa and f(Is,Q)=f(Is,Q). Therefore, V1α1βIsSN(lnEElnIsIs+IsIsEE)+fβIaSN(lnEElnIaIa+IaIaEE)+α2βQSN(lnEElnQQ+QQEE).Similarly, by using (Equation5c), V2=(1IsIs)(ρσEρσEIsIs)=ρσE(1+EEIsIsIsEIsE)ρσE(EEIsIslnEE+lnIsIs)with equality if and only if EE=IsIs. By using (Equation5d), V3=(1IaIa)((1ρ)σE(1ρ)σEIaIa)=(1ρ)σE(1+EEIaIaIaEIaE)(1ρ)σE(EEIaIalnEE+lnIaIa)with equality if and only if EE=IaIa. By using (Equation5e), V4=(1QQ)(κIsκIsQQ)=κIs(1+IsIsQQQIsQIs)κIs(IsIsQQlnIsIs+lnQQ)with equality if and only if IsIs=QQ.

Combining the above estimates gives V(l1βSN(α1Is+fIa+α2Q)+l2ρσE+l3(1ρ)σE)(EElnEE)+(l1α1βIsSNl2ρσE+l4κIs)(IsIslnIsIs)+(l1fβIaSNl3(1ρ)σE)(IaIalnIaIa)+(l1α2βQSNl4κIs)(QQlnQQ).Setting the coefficient of each term on the right-hand side of the inequality to zero leads to a system of linear equations with respect to l1,l2,l3 and l4, from which we obtain l1=ρ(1ρ)σE,l2=(1ρ)βS(α1Is+α2Q)N,l3=ρβf(Is,Q)SIaN,l4=ρ(1ρ)σα2βSEQκNIs.Under this choice of li for i = 1, 2, 3, 4, we have V0. Furthermore, V=0 implies that S=S,f(Is,Q)=f(Is,Q),E=qE,Is=qIs,Ia=qIaandQ=qQfor some q>0. Substituting these relations into the first equation of (Equation2) gives 0=Λλ(S,qE,qIs,qIa,qQ)SμS=Λλ(S,E,Is,Ia,Q)SμSλ(S,qE,qIs,qIa,qQ)=λ(S,E,Is,Ia,Q)βα1qIs+fqIa+α2qQN=βα1Is+fIa+α2QNq=1.Thus, V=0 holds only at E. By the standard Lyapunov stability theorem, the global stability of E is established.