Loading [MathJax]/jax/output/SVG/jax.js
Research article Special Issues

Optimal control analysis of malware propagation in cloud environments


  • Cloud computing has become a widespread technology that delivers a broad range of services across various industries globally. One of the crucial features of cloud infrastructure is virtual machine (VM) migration, which plays a pivotal role in resource allocation flexibility and reducing energy consumption, but it also provides convenience for the fast propagation of malware. To tackle the challenge of curtailing the proliferation of malware in the cloud, this paper proposes an effective strategy based on optimal dynamic immunization using a controlled dynamical model. The objective of the research is to identify the most efficient way of dynamically immunizing the cloud to minimize the spread of malware. To achieve this, we define the control strategy and loss and give the corresponding optimal control problem. The optimal control analysis of the controlled dynamical model is examined theoretically and experimentally. Finally, the theoretical and experimental results both demonstrate that the optimal strategy can minimize the incidence of infections at a reasonable loss.

    Citation: Liang Tian, Fengjun Shang, Chenquan Gan. Optimal control analysis of malware propagation in cloud environments[J]. Mathematical Biosciences and Engineering, 2023, 20(8): 14502-14517. doi: 10.3934/mbe.2023649

    Related Papers:

    [1] Maeve L. McCarthy, Dorothy I. Wallace . Optimal control of a tick population with a view to control of Rocky Mountain Spotted Fever. Mathematical Biosciences and Engineering, 2023, 20(10): 18916-18938. doi: 10.3934/mbe.2023837
    [2] Rocio Caja Rivera, Shakir Bilal, Edwin Michael . The relation between host competence and vector-feeding preference in a multi-host model: Chagas and Cutaneous Leishmaniasis. Mathematical Biosciences and Engineering, 2020, 17(5): 5561-5583. doi: 10.3934/mbe.2020299
    [3] Yan-Xia Dang, Zhi-Peng Qiu, Xue-Zhi Li, Maia Martcheva . Global dynamics of a vector-host epidemic model with age of infection. Mathematical Biosciences and Engineering, 2017, 14(5&6): 1159-1186. doi: 10.3934/mbe.2017060
    [4] Yunbo Tu, Shujing Gao, Yujiang Liu, Di Chen, Yan Xu . Transmission dynamics and optimal control of stage-structured HLB model. Mathematical Biosciences and Engineering, 2019, 16(5): 5180-5205. doi: 10.3934/mbe.2019259
    [5] Xinli Hu, Yansheng Liu, Jianhong Wu . Culling structured hosts to eradicate vector-borne diseases. Mathematical Biosciences and Engineering, 2009, 6(2): 301-319. doi: 10.3934/mbe.2009.6.301
    [6] Djamila Moulay, M. A. Aziz-Alaoui, Hee-Dae Kwon . Optimal control of chikungunya disease: Larvae reduction, treatment and prevention. Mathematical Biosciences and Engineering, 2012, 9(2): 369-392. doi: 10.3934/mbe.2012.9.369
    [7] Fahad Al Basir, Yasuhiro Takeuchi, Santanu Ray . Dynamics of a delayed plant disease model with Beddington-DeAngelis disease transmission. Mathematical Biosciences and Engineering, 2021, 18(1): 583-599. doi: 10.3934/mbe.2021032
    [8] Ling Xue, Caterina Scoglio . Network-level reproduction number and extinction threshold for vector-borne diseases. Mathematical Biosciences and Engineering, 2015, 12(3): 565-584. doi: 10.3934/mbe.2015.12.565
    [9] Rundong Zhao, Qiming Liu, Huazong Zhang . Dynamical behaviors of a vector-borne diseases model with two time delays on bipartite networks. Mathematical Biosciences and Engineering, 2021, 18(4): 3073-3091. doi: 10.3934/mbe.2021154
    [10] Rong Ming, Xiao Yu . Global dynamics of an impulsive vector-borne disease model with time delays. Mathematical Biosciences and Engineering, 2023, 20(12): 20939-20958. doi: 10.3934/mbe.2023926
  • Cloud computing has become a widespread technology that delivers a broad range of services across various industries globally. One of the crucial features of cloud infrastructure is virtual machine (VM) migration, which plays a pivotal role in resource allocation flexibility and reducing energy consumption, but it also provides convenience for the fast propagation of malware. To tackle the challenge of curtailing the proliferation of malware in the cloud, this paper proposes an effective strategy based on optimal dynamic immunization using a controlled dynamical model. The objective of the research is to identify the most efficient way of dynamically immunizing the cloud to minimize the spread of malware. To achieve this, we define the control strategy and loss and give the corresponding optimal control problem. The optimal control analysis of the controlled dynamical model is examined theoretically and experimentally. Finally, the theoretical and experimental results both demonstrate that the optimal strategy can minimize the incidence of infections at a reasonable loss.



    Vector-borne diseases exhibit a common feature: a two-way transmission of the pathogen between human hosts and various species of arthropods, most often mosquitos, that serve as carriers (or vectors) of the pathogen, but direct host-host transmission is impossible for most of them. Diseases caused by pathogens transmitted by mosquito bites (malaria, dengue fever, yellow fever, zika, chikungunya) cause periodic outbreaks in the countries located in tropical climate zones, and pose a growing burden on their healthcare systems, as seen, for instance, in a 30-fold increase in dengue fever's incidence since the 1970s [1]. Control campaigns against them are based on two broad types of measures: suppression of transmission and elimination of the vector. They rely on the use of long-lasting insecticidal nets (LLINs), indoor residual spraying with insecticide, use of larvicide, as well as use of personal protection means with repellent activity that include spray-on repellents or repellent-treated textiles and household items.

    The success of control measures aimed at the vector's demography, like bed nets and insecticide spraying, in reducing the disease burden has been limited for various reasons. On the one hand, there is evidence for changing patterns in the transmission of malaria to outdoor environments [2,3,4,5], while the main vectors of dengue, yellow fever and zika viruses, female mosquitoes of the Aedes genus are active during daytime hours. Indiscriminate use of insecticide and larvicide, on the other hand, poses danger to human health and the ecosystem. Furthermore, insecticide resistance has contributed to the failure of the Global Malaria Eradication Programme [6] and has been noted among Aedes mosquitos [7,8].

    Use of repellent-treated clothes and household items as preventive measure could serve to suppress transmission of pathogens from vector to host and vice versa. Changing mosquito host-seeking behavior by repellent application could reduce the biting rate of the female mosquitoes, and this strategy has gained attention among entomologists. This includes studies of the effect of spray-on repellents and wearable repellent devices on mosquito host-seeking behavior [9], and tests of fibers for controlled release of volatile mosquito repellents to prevent bites in the context of malaria control [10]. Development of mosquito repellents based on organic substances from plant oils [11], and of methods for estimation of efficacy of repellents used in textiles and paints [12] reveal new perspectives for such measures.

    Mathematical modeling can help analyze and predict the performance of such interventions for meeting policy objectives. We study a simple Ross-Macdonald type model [6] for a vector-borne disease with control based on distribution of textiles and household items with repellent properties. The constraints on an intervention campaign of this kind encompass, first, the economic cost of production, distribution, etc. of the repellent-treated textiles, which influences its extent, that is, the maximum fraction ˉu of the target population employing products with mosquito repellent property. Second, due to technological and physical limitations, such as evaporation, washing, UV radiation exposure, etc., the repellent property is assumed to be lost after a period of duration T days..

    In this study we propose a method to analyze whether a control campaign under those constraints is able to reduce and maintain the size of the infected host compartment below a certain level, denoted henceforth as the infection cap.

    The infection cap ˉI is a constraint on a state variable of the model (infected hosts) and its value depends on factors, such as the availability of treatment, healthcare system capacity or societal or political tolerances, etc. For the purpose of our analysis, we state the following Questions.

    Q.1 Given ˉu,ˉI and initial data on infected hosts and vectors, establish whether the size of the infected host compartment can remain below the infection cap for all future times and find the optimal strategy to maintain it.

    Q.2 If the initial data are such that the objective from Q.1 can not be met, we address the following: given ˉu,ˉI,T, find the optimal strategy to bring the size of the infected host compartment below the infection cap ˉI in a minimal time not exceeding T, and maintain it below the cap ˉI until the campaign's end.

    Both questions concern characterization of initial data inside the model's state space unlike standard control analysis in the context of vector-borne diseases, which addresses optimal resource allocation [13,14,15,16] These questions can be addressed using the toolbox of optimal control theory.

    Question Q.1 is a viability problem and we study the existence of viable trajectories of the dynamical system, namely those that satisfy the constraint stated in Q.1. The maximum set of initial data for Q.1 where this constraint is met, is the viability kernel. There has been some analysis of viable controls in epidemiology: in reference [17] viability kernels for a model with control on the mosquito population by fumigation have been computed and validated with data from a dengue outbreak in Cali, Colombia. In a previous work [18] we have characterized the viability kernel of a model for a vector-borne disease with susceptible, infected and removed compartments for the host population. Even though this problem has an infinite time horizon, its solution shall be used for the solution of Question Q.2.

    For Question Q.2 we study a reachability problem for a target set (the viability kernel), and compute the minimal entry time function to it. Such analysis can serve as a guide to policy makers to determine determine whether and how fast the repellent-based control campaign constrained by its coverage ˉu and duration T is able to reduce the size of the infected host compartment below the infection cap ˉI. If the minimal time lies within the given time horizon (0,T), the campaign is considered effective. Else, the campaign should be considered ineffective because it cannot meet this objective. The reachability problem framed by Question Q.2 is important as it can provide an indicator of the control campaign's performance under the constraints to decision makers.

    We use a variational approach to compute numerically approximations of the viability kernel of the controlled system and the minimal entry time function to the viability kernel, using sub-zero level set of a value function solving a Hamilton-Jacobi equation associated to the model system. The viability kernel can also be described analytically by its boundary in the phase plane of the state variables (infected hosts and infected vectors) [17]. However, the numerical approximation of the value function from the variational approach {allows us} to reconstruct the control function which solves the viability and reachability problems. We illustrate the application of this method from optimal control theory using parameterized models for malaria transmission in Botswana and Zimbabwe [19] and discuss the efficacy of the repellent-based control campaign in meeting the stated objectives.

    We use the compartmental model for a vector-borne disease as presented in [20], which follows the dynamics of susceptible and infected hosts Sh(t),Ih(t) and susceptible and infected female mosquitoes Sv(t),Iv(t) that serve as vectors for the pathogen, at time t0. Susceptible hosts become infected after receiving a bite from an infectious mosquito. A susceptible mosquito becomes infected after biting an infected host, and after an incubation period of length τ becomes infectious and is able to transmit the pathogen. We model the incubation period without a specific compartment as done by [19,21]. If the within-vector incubation period τ and the expected life-time of a mosquito 1/μ are of the same order of magnitude, the probability of death of an infected mosquito during the pathogen incubation is not negligible. Hence, merely a fraction exp(μτ) of the infected mosquitoes Iv is capable of transmitting the pathogen to the host [21]. In this model no protective immunity is assumed. Hosts from the infected compartment after a short infectious period re-enter immediately the susceptible compartment [22,23], the length of the period is 1/γ, with γ the recovery rate.

    Let u(t) denote the proportion of the target population at time t using the measures provided by the control campaign. With k being the repellent efficacy, the modified mosquito biting rate due to use of control measures is am(1ku(t)). Such modified infection rates are used in reference [15] in the context of an age-structured model for malaria and controls by long-lasting insecticidal nets. We assume that the cost of control effort C(u(t)) (production and distribution of the repellent-treated textiles or spray-on repellents) is borne by decision makers and is a linear function of u(t) (compare reference [38]), and that C(u(t))Cmax for all t0. This means that u(t)ˉu for all t0 for some maximum coverage ˉu(0,1] of the target population. The set of control functions we consider is

    U={u(t):R+[0,ˉu],upiecewise continuous }.

    Observe that for u=ˉu we have the maximum available reduction of the biting rate, while for u=0 there is no control used, and hence no reduction of the biting rate.

    The host population is assumed to be constant over time, Sh(t)+Ih(t)=N. We assume that the repellent-based control does not lead to changes in the total populations of female mosquitoes over time, M=Sv(t)+Iv(t). Hence, we can reduce the number of state variables by letting Sh=NIh,Sv=MIv and non-dimensionalizing x1=Ih/N,x2=Iv/M. In our notation, x(t) denotes the vector of state variables describing the different compartments of the epidemiological model at time t, and f(x,u) denotes the model dynamics subject to a control function uU.

    Letting ν=M/N be the average number of female mosquitoes per host and setting for shortness's sake α=amphexp(μτ),β=ampv, we work with the 2-dimensional non-autonomous system:

    ddtx=f(x,u) withf1(x,u)=(1ku(t))ανx2(1x1)γx1f2(x,u)=(1ku(t))βx1(1x2)μx2, (2.1)

    subject to initial conditions x0=(x1(0),x2(0))Ω=[0,1]2. The domain of definition Ω corresponds to the states x with relevant epidemiological meaning for (2.1). On the segments on the boundary Ω defined by x1=0,x2=0,x2=1,x1=1, one can verify that the flow defined by f points towards intΩ. Thus, Ω is forward invariant under f.

    The associated trajectory of (2.1) for a given control function uU with initial condition x0 will be denoted by xu(t;x0)={(xu1(t),xu2(t)),t>0}. The set of feasible trajectories on the time interval [0,+) starting at x0Ω at time t=0 will be denoted by

    S(x0,U):={xu(t;x0)|uU}.

    The system (2.1) is cooperative (or quasimonotone [20]) because xifj0,ij. Hence, tools from the theory of ordinary differential equations such as comparison principles [24] are at our disposal to establish properties of the feasible trajectories.

    We conclude this section with

    Lemma 1. The function f in system (2.1) is Lipschitz continuous on Ω with Lipschitz constant

    L=[(max{2αν,αν+γ})2+(max{2β,β+μ})2]1/2 (2.2)

    Proof. The estimate follows from direct computation of the Euclidean norm for |f(x,u)f(x,u)|,x,xΩ,u,u[0,ˉu].

    Using the Lipschitz continuity of f on Ω (Lemma 1) as well as the facts that a) the set {f(x,u),0uˉu} is convex for all xΩ and b) we can choose C>0 so f(x,u)C(1+|x|) on Ω (f has linear growth), we establish that solutions to system (2.1) exist for t0 [25,Chapter III.5]. In addition, since the domain and graph of f are closed, f is a Marchaud map on Ω [26,Corollary 2.2.5].

    In Question Q.1 we are interested in finding the initial conditions x0 of (2.1) such that for a given maximum extent ˉu and infection cap ˉI>0, there exists u(t)U such that xu1(t)ˉI,t0. Let A(ˉI)={(x1,x2)Ω|x1ˉI}, which is a closed subset of Ω with a compact boundary. A feasible trajectory xS(x0,U) is called viable for A(ˉI) if the constraint xu1(t)ˉI,t0 is met, in other words, x(t)A(ˉI) for t[0,+]. A set DΩ is a viability domain for f if from any initial state x0D, at least one trajectory xS(x0,U) is viable for D [26,27].

    To answer Question Q.1, we characterize the viability kernel associated to AˉI,ˉu, that is the following set of initial data for (2.1):

    V(ˉI,ˉu):={x0Ω|xS(x0,U) with x(t)A(ˉI),t0}. (3.1)

    Evidently, V(ˉI,ˉu) is the set of all initial data x0A(ˉI) from which viable trajectories for A(ˉI) for t0 exist. It is the largest closed viability domain of f inside A(ˉI) [27]. The set A(ˆI) being closed, and f being a Marchaud map, the viability kernel V(ˉI,ˉu) for system (2.1) is well-defined [26,Theorem 4.1.2]. The viability kernel will depend on the constraint on the control ˉu, and on the state constraint (the infection cap ˉI).

    We introduce the effective basic reproduction number for control intervention u(t)ˉu

    Rˉu=αβν(1k˜u)2γμ (3.2)

    and the critical threshold of infection

    Icrit=Rˉu1Rˉu+βμ(1kˉu), (3.3)

    whenever Rˉu>1,Icrit>0. Following reference [18], Icrit equals the share of infected hosts at the endemic equilibrium E of system (3.4) at the maximum coverage with repellent-treated products.

    Here we briefly recall stability properties of the model equilibria under constant controls u(t)=ˉu. In that case the system (2.1) becomes autonomous,

    ddtx1=αν(1kˉu)x2(1x1)γx1,ddtx2=β(1kˉu)x1(1x2)μx2. (3.4)

    System (3.4) has a trivial, disease-free equilibrium at the origin O and an endemic equilibrium:

    E=(x1,x2)=(Icrit,Rˉu1Rˉu+ανγ(1kˉu)). (3.5)

    Note that EintΩ* if and only if Rˉu>1. At Rˉu=1, a transcritical bifurcation occurs and E and O coincide [23].

    *By intX we denote the interior of a set X.

    Lemma 2. The system (3.4) has the following asymptotic behavior:

    (i) Rˉu1 implies that the disease-free equilibrium O is globally asymptotically stable for (3.4) in Ω.

    (ii) Rˉu>1 implies that the endemic equilibrium E (3.5) is globally asymptotically stable for (3.4) in ΩO.

    The proofs of Lemma 2 and further statements are in the Supplementary Material.

    Using ˉI and Icrit, we characterize whether the viability kernel for system (2.1) has a positive Lebesgue measure in R2:

    Proposition 1. Consider the model with constant vector population (2.1). Let

    (i) ˉImax{0,Icrit}. Then intV(ˉI,ˉu) has positive Lebesgue measure in R2.

    (ii) Icrit>0 and ˉI<Icrit. Then V(ˉI,ˉu)=O.

    Based on Proposition 1, the description of V(ˉI,ˉu) can be refined by defining its boundaries. This has been the approach in reference [17], and we recall the details for model (2.1). We start by a technical result:

    Lemma 3. Let ˉI>max{0,Icrit}. Set ˉv=γˉIαν(1kˉu)(1ˉI) and assume ˉv<1.The initial value problem

    ddsz=f1(z,s,ˉu)f2(z,s,ˉu)=αν(1kˉu)(1z)sγzβ(1kˉu)z(1s)μs,s>ˉv,z(ˉv)=ˉI, (3.6)

    has a unique non-negative solution which is monotone decreasing on the interval (ˉv,min{1,v}), where v is such that z(v)=0.

    Due to the inverse function theorem, Lemma 3 implies the existence of an inverse map z1 to z defined on [0,ˉI] such that z1(z(s))=s. We denote by Z the solution curve (z(s),s) in Ω of the problem in Lemma 3 to describe explicitly the boundary of the viability kernel for the case ˉI>max{0,Icrit}.

    Proposition 2. For given ˉI,ˉu the viability kernel V(ˉI,ˉu) takes one of the following forms

    (i) Let ˉI>(1kˉu)αν(1kˉu)αν+γ, then V(ˉI,ˉu)=A(ˉI).

    (ii) Let ˉI(max{0,Icrit},(1kˉu)αν(1kˉu)αν+γ]. Then

    V(ˉI,ˉu)={[0,ˉI]×[0,ˉv]}{(x1,x2)|ˉvx2min{v,1},0x1z(x2)}, (3.7)

    where z(x) is the solution to the initial value problem (3.6) in Lemma 3, and v such that z(v)=0.

    Unfortunately, this "direct" approach does not work for the case ˉI=Icrit. While Proposition 1 shows that the viability kernel has positive measure, the approach from Lemma 3 cannot work, because the right-hand side of (3.6) is not defined at the initial condition z(ˉv)=ˉI, which is precisely at the endemic equilibrium E.

    To avoid this problem, we state an alternative characterization of the viability kernel using a variational formulation. Following reference [28], we consider an infinite horizon minimization problem with exact penalization of the state constraint. The constraint is expressed via the cost function, defined by the signed distance function Γ

    Γ(x)={infyA(ˉI)|xy|,xΩintA(ˉI)infyΩA(ˉI)|xy|,xΩintA(ˉI). (3.8)

    Note that Γ(x)0 if and only if xA(ˉI). Following the steps in reference [28], we consider for >0 the value function

    v(x)=infuUsupt(0,+)etΓ(xu(t;x)),xR2. (3.9)

    Note that v takes finite values on Ω. From (3.9), it follows v(x)0 if and only if xu(t;x)V(ˉI,ˉu) for all t(0,+). Therefore, the viability kernel may be characterized using sub-zero-level sets of the value function v:

    V(ˉI,ˉu)={xΩ|v(x)0}. (3.10)

    Let >L with L being the Lipschitz constant of f (Lemma 1). Then the value function v satisfies the following dynamic programming principle

    v(x)=infuUmax{etv(x),sups(0,t]esΓ(xux(s))},t>0

    and is Lipschitz continuous [28].

    Let v denote the gradient of v, v=(x1v,x2v). The results in reference [28,Section 4] imply that v is the unique continuous viscosity solution [33,Definition 2] of the Hamilton-Jacobi-Bellman equation

    min{v(x)+maxuUH(x,u,v),v(x)Γ(x)}=0,xR2. (3.11)

    with as above and Hamiltonian

    H(x,u,v)=f(x,u),v. (3.12)

    We refer the reader to reference [25,Chapter 1.3] for the definitions of viscosity solutions to partial differential equations.

    Hence, using well-known numerical methods for Hamilton-Jacobi equations based on finite difference discretization of the spatial derivative [29,30], one can compute the solution of (3.11) and find an approximation of the viability kernel V(ˉI,ˉu) for given values of target infection cap ˉI and maximum coverage ˉu.

    In our case the quantity maxuUH(x,u,v) has an explicit form :

    maxuUH(x,u,v)=(γx1αν(1x1)x2)x1v+(μx2βx1(1x2))x2v+max{0,ανkˉu(1x1)x2x1v}+max{0,βkˉu(1x2)x1x2v}. (3.13)

    Using the computed value function v, we can answer Q.1, compute the optimal control function u(t) and reconstruct the optimal trajectory for a given initial condition x0V(ˉI,ˉu) using the reconstruction algorithm from reference [31,32]. This algorithm is applied in reference [18] to approximate the viability kernel for another model of a vector-borne disease.

    Once we have computed the viability kernel, we turn our attention to the reachability problem from Question Q.2. Let the initial data x0=(x1(0),x2(0)) on infected hosts and vectors, the maximum extent ˉu, the duration T, and the infection cap ˉI be given. We ask if there exists an optimal strategy u(t) to bring x1 population below the infection cap ˉI in a minimal time τ<T and maintain x1(t)<ˉI for t(τ,T].

    To this purpose we define the minimal entry time function TX(x) for initial data x to the target set X as follows:

    TX(x)={+, if {t0|xu(t;x)X}=,infuUmin{t0|xu(t;x)X},else. (4.1)

    The objective in Question Q.2 is to compute the minimal entry time function for the target set being the viability kernel V(ˉI,ˉu). This makes sense unless the viability kernel is trivial (in other words, we are interested in computing T for V(ˉI,ˉu)O). By definition, V(ˉI,ˉu)Ω is a compact set, hence V(ˉI,ˉu) is compact. We shall characterize the minimal entry time function using the value function v from (3.9) based on the approach presented in references [28,33].

    Using the value function v which solves (3.11) and defines the viability kernel V(ˉI,ˉu) via its sub-zero level set (3.10) from the previous section we formulate a variational problem for the minimal entry time function. The backwards reachable set for the target set (which is viability kernel V(ˉI,ˉu)) at time T, denoted by W(T), is defined as the set of those initial data x for system (2.1) such that there exists a feasible trajectory xS(x,U) with xu(t;x)V(ˉI,ˉu) for tT.

    In other words,

    W(T)={xR2|t[0,T],uU:xu(t;x)V(ˉI,ˉu)}. (4.2)

    Note that {W(t),t0} is a family of increasing closed sets [33,Remark 2]. We remark that even though the biologically relevant domain for system (2.1) is Ω, the proof of Lemma 2 implies Ω attracts the trajectories of (2.1) starting at any x0R2+, so in principle there exist T>0 with ΩW(T).

    The set W(T) can be characterized using a level-set approach based on the computed value function v by avoiding any controllability assumption for (2.1). We formulate a minimization problem (4.3) for a value function w, which is the unique continuous viscosity solution of a Hamilton-Jacobi-Bellman equation [33,Definition 2 and Theorem 2].

    To impose a restriction W(T)Ω into the variational formulation, we introduce a penalization term GΩ(x), which is a Lipschitz-continuous function which satisfies the condition for sub-zero-level set, GΩ(x)0 if and only if xΩ.

    To compute the minimal entry time function, we consider the solution w of the minimization problem

    w(t,x)=minuU{max{v(xu(t;x)),max0stGΩ(xu(s;x))}}, (4.3)

    Note that w is defined over R2, but the penalization term in (4.3) makes the value function w positive if the trajectory leaves the domain Ω. The backwards reachable set W(T) for (2.1) is thus defined as the sub-zero-level set of the function w, W(T)={xR2|w(T,x)0}, and (4.3) implies W(T)Ω.

    Following reference [33,Lemma 1], the value function w satisfies the following dynamic programming principle

    w(t+s,x)=infuU{max{w(t,xu(s;x),maxτ[0,s]GΩ(xu(τ;x))}}w(0,x)=max{v(x),GΩ(x)}, (4.4)

    and w is the unique continuous viscosity solution to the Hamilton-Jacobi-Bellman equation

    min{tw(t,x)+maxuUH(x,u,w),w(t,x)GΩ(x)}=0,t0w(0,x)=max{v(x),GΩ(x)},xR2. (4.5)

    with w=(x1w,x2w), and Hamiltonian H provided by (3.12).

    Using w, we define the minimal entry time function (4.1) TV(ˉI,ˉu)(x) for V(ˉI,ˉu) for a trajectory with initial condition x as

    TV(ˉI,ˉu)(x)=inf{t0|xW(t)}=inf{t0|w(t,x)0}.

    As there is no controllability assumption for (2.1), TV(ˉI,ˉu) need not be continuous over Ω. In fact, the proof of part (ii) of Proposition 2 reveals that the minimal entry time function exhibits a discontinuity across the boundary of the viability kernel V(ˉI,ˉu) traced by the solution curve Z defined in Lemma 3. Observe that even though the biologically relevant domain for (2.1) is just Ω, TV(ˉI,ˉu)(x) may take finite values for points xΩ, while TV(ˉI,ˉu)(x)=+ for x:x1,x2<0.

    For the purpose of our study we are interested in minimal entry times not exceeding a given time horizon T. As Ω is positively invariant under f, we employ the penalization term

    GΩ(x)={1,xΩ,1,xΩ.

    To achieve higher accuracy of approximation for w(t,x),t[0,T] close to the boundary Ω we solve numerically (4.5) on a uniform stencil for a domain which contains Ω in its interior using a second-order Runge-Kutta scheme for the time integration. We find an approximation of the minimal entry time function on Ω by using the obtained solution values for w(t,x),xΩ, as well as the optimal control using the reconstruction algorithm given in references [31,32].

    Mathematical analysis of models for vector-borne diseases tends to concern the asymptotic rather than the transient behavior of their solutions using the basic reproduction number R0 as a criterion for the stability of the disease-free equilibrium [34,35,36,37]. Of interest are the parameter values that keep R0 below unity. Applications of optimal control theory often focus on optimal allocation of resources (by minimizing cost functionals of different form) without addressing future transient dynamic behavior [38].

    In this work we study the latter question by considering a control problem with restricted state space, defined by a cap on the size of the infected host compartment, and address two major questions. In Question Q.1 we seek the viability kernel V for the model (2.1), the subset of the state space Ω that comprises those initial data such that a viable control u(t) exists to maintain the size of infected host compartment x1(t) below the infection cap ˉI for all future times. Then, we address a reachability problem formulated as Question Q.2. If the interventions start from an initial condition x0V which is outside the viability kernel, the objective stated in Question Q.1 cannot be met, so instead we are interested in finding the optimal control strategy u(t) which reduces x1(t) below ˉI in the shortest time. The motivation behind Question Q.2 relates to the fact that control measures studied here are restricted in duration due to technological and design limitations leading to a loss of product's repellent property.

    Analysis of viability and reachability allows decision makers to assess the effect of this control intervention on future transient behavior given the epidemic situation at the start. This framework gives an opportunity to analyze the dynamic behavior of the controlled model on its entire domain of definition, or large subsets thereof, by describing existence of viable controls and of sets reachable in finite time. In particular, it is possible to estimate whether under the constraints of cost Cmax (which determines the population coverage ˉu) and duration of repellent action T, the objective of reducing the size of the infected host compartment below the infection cap ˉI can be met.

    The reachability analysis could serve as a predictive indicator for the performance of the intervention with parameters ˉu,ˉI,T by providing analysis of the model's trajectories. If a trajectory which reaches the target (viability kernel) within the time interval (0,T) exists, the campaign is considered effective. Else, the intervention strategy is to be considered insufficient and may need to be supplemented by additional measures (use of repellents with higher efficacy, use of LLINs, indoor residual spraying, vaccinations, preventive treatment, etc.).

    We use parameter values for a malaria model fitted for Botswana and Zimbabwe from reference [19] (listed in Table 1) as a basis for the numerical simulations to exemplify this method. Viability kernels and minimal entry time functions are computed and compared across different values of ˉu.

    Table 1.  Model parameters based on reference [19] used in the numerical simulation for Botswana and Zimbabwe. The values in parentheses refer to Zimbabwe.
    Parameter Description Value Unit
    am Mosquito biting rate 0.082 (0.241) day-1
    (average number of blood meals per day)
    pv Probability of the pathogen transmission
    from a human host to mosquito 0.1
    ph Probability of the pathogen transmission
    from a mosquito to a human host 0.5
    ν Ratio of female mosquitoes to hosts 10
    γ Recovery rate for hosts 114 day\textsuperscript{-1}
    τ Within-vector incubation period 10 day
    μ Mosquito death rate 130 (110) day\textsuperscript{-1}
    Control
    k Efficacy of repellents 60%
    ˉu Campaign extent (maximum coverage) as indicated
    T Campaign duration 100 day
    ˉI Infection cap 0.1

     | Show Table
    DownLoad: CSV

    The main differences between the parameters is the mosquito biting rate, am=0.082 for Botswana and am=0.241 for Zimbabwe, and the expected life-time of a mosquito 1/μ, which is three times higher in Botswana. According to reference [19] and therein cited references, the parameter μ reflects a difference in overall coverage of indoor residual spraying in the two countries. This intervention influences the vector's demography and ecology by direct insecticidal or repellent action, but different spraying solutions are used in homes built from traditional materials (wood, clay or mud bricks) and in Western-style housing [39]. While it has suppressed the transmission of malaria in the past, societal refusal to indoor spraying due to pungency, home wall staining, bug infestation, etc. is on the rise [40], and the positive trends in disease control have been reversed.

    The results from Proposition 1 enable us to compute the threshold for population coverage minˉu such that for a given infection cap ˉI the viability kernel V(ˉI,ˉu) is not trivial:

    minˉu=1kβγˉI+(βγˉI)2+4αβγμν(1ˉI)2αβγν(1ˉI)k. (5.1)

    If minˉu0, then the viability kernel V(ˉI,ˉu) has a positive Lebesgue measure in R2 for all 0ˉu1.

    Figure 1 plots the trade-off relationship (5.1) between ˉI,minˉu for both countries for two values of repellent efficacy k. It shows the minimal coverage ˉu required to produce a viability kernel of positive Lebesgue measure for the given infection cap ˉI. An immediate consequence is that in the case of Botswana, for ˉI0.01, any extent of the control measures will suffice to produce a non-trivial viability kernel. In contrast, more control efforts must be employed in the case of Zimbabwe, where at least 20% population coverage is required (minˉu>0.2) to produce a non-trivial viability kernel. Of course, we must look at the area of the viability kernel to evaluate better the potential for keeping the epidemic peak below ˉI with the repellent-based strategy under consideration.

    Figure 1.  Trade-off relationship between ˉI and minˉu to keep the viability kernel V(ˉI,ˉu) with positive Lebesgue measure. Range of ˉI:(103,101) for two values of the repellent efficacy k. Parameters used for a) Botswana, b) Zimbabwe.

    The infection cap is ˉI=0.1 and repellent efficacy k=0.6. The epidemiological parameters for Botswana indicate that the viability kernels are not trivial for all 0ˉu1. Figure 2a) displays the area of the viability kernel for a range of values of ˉu. There is a positive correlation between ˉu and the area of V(ˉI,ˉu). For the given epidemiological parameters in Zimbabwe, the situation is different: for ˉI=0.1 the viability kernel is trivial if ˉu<0.2152; in other words, under this assumption the model predicts it is not possible to maintain the size of infected host compartment below ˉI for any initial data in Ω if the population coverage is less than 21.5%. The area of the viability kernel for a range of values of ˉu0.215 is plotted in Figure 2b).

    Figure 2.  Area of the viability kernel V(ˉI,ˉu) as function of the campaign's extent ˉu. Parameters for a) Botswana; b) Zimbabwe.

    Examples of viability kernels' shape for Botswana are shown in Figure 3. The viability kernels for Zimbabwe have a similar shape as those for Botswana, displayed in the panels of Figure 3, and are not shown.

    Figure 3.  Numerical approximation of the viability kernel V(ˉI,ˉu) for the epidemiological model. Parameters for Botswana with maximum coverage: a) ˉu=0.6, b) ˉu=0.7, c) ˉu=0.8, d) ˉu=0.9.

    An example for a reconstructed optimal trajectory inside Ω is given in Figure 4a), and the control effort – in Figure 4b) – it is computed using the algorithm from reference [31]. Note that for the same initial data, without intervention, the size of infected host compartment x1 exceeds the cap ˉI (the trajectory follows the purple curve in Figure 4a).

    Figure 4.  a) An example of a reconstructed optimal trajectory (red) maintained inside the viability kernel versus the trajectory with no control u(t)0 (purple). b) Plot of the corresponding control function. Parameters for Botswana with ˉu=0.6.

    The numerical approximations for the minimal entry time functions TV(ˉI,ˉu) are shown in Figure 5 for Botswana and Figure 6 for Zimbabwe. In the simulations we use an infection cap of 10% of the population (ˉI=0.1) and a time horizon of three months (T=100 days). The results show that with these constraints, under the parameter values for Botswana, an outbreak would be successfully brought under the infection cap within the given time horizon, unless the proportion of infected mosquitos is very high. In fact, for ˉu=0.6, the set of initial data x0 such that TV(x0)<100 days (those such that the size of the infected host compartment can be reduced to less than 0.1 within 100 days) is contained within the rectangle [0,1]×[0,0.25] and for ˉu=0.9, it coincides almost fully with the rectangle [0,1]×[0,0.5]. That means that increasing the coverage from 60% to 90% of the population approximately doubles the area of the set {x0|TV(x0)<100 days}. However, at very high levels of infected mosquitos at the start of the intervention (x2(0)>0.5) the intervention cannot be considered effective as for all different extents of coverage ˉu[0.6,0.9] the minimal time needed to reduce the outbreak size below the prescribed cap exceeds 100 days.

    Figure 5.  Numerical approximation of the minimum entry time function TV(ˉI,ˉu) (plot of isolines indicating days). Parameters for Botswana with maximum coverage: a) ˉu=0.6, b) ˉu=0.7, c) ˉu=0.8, d) ˉu=0.9.
    Figure 6.  Numerical approximation of the minimum entry time function TV(ˉI,ˉu) (plot of isolines indicating days). Parameters for Zimbabwe with maximum coverage: a) ˉu=0.6, b) ˉu=0.7, c) ˉu=0.8, d) ˉu=0.9.

    For Zimbabwe, the minimal entry time functions show that for the same maximum population coverage with repellent ˉu less time is required to bring the outbreak under the chosen infection cap ˉI. The isolines in Figure 6 follow a radial pattern rather than parallel shifts like those for Botswana. Even at ˉu=0.6 for the entire set of initial conditions x0[0,1]×[0,0.5] the minimal entry time to the viability kernel is less than 90 days. Reconstructed optimal trajectories for the reachability problem in Question Q.2 for an initial condition x0=(0.3,0.2) are plotted in Figure 7.

    Figure 7.  Reconstructed optimal trajectories for the minimal time needed to bring the size of the infected compartment I below the cap ˉI=0.1 (dashed line in the left panels) if the intervention starts at initial data x0=(0.3,0.2) for a) Botswana the minimal entry time T(x0)=86.7 days, b) Zimbabwe the minimal entry time T(x0)=63.4 days. Parameter of maximum coverage ˉu=0.6.

    It may seem paradoxical that for the model with parameters for Zimbabwe the minimal entry time to the respective viability kernel is shorter than for Botswana; in other words, a scenario with a higher basic reproduction number R0 appears more elastic in its response to the campaign of repellent-based interventions modeled by (2.1). That elasticity could be explained with the difference in the vector's ecology: the model is parameterized with a threefold higher mortality rate for the mosquitos in Zimbabwe (μ=110) relative to Botswana (μ=130), despite the fact that the pathogen transmission rate from vector to host α, and R0 for a baseline scenario of no control (u(t)=0) are higher for Zimbabwe [19]. This simple example shows how transient dynamical properties of the controlled system (2.1) in the case of endemicity stay decoupled from the value of a traditional epidemiological metric R0.

    These simulations agree with {entomological} observations of the importance of interventions which target the vector's demography such as indoor residual spraying, LLINs, etc, for suppressing pathogen transmission [41]. Different societal behavior and inconsistent following of measures such as indoor spraying [19] can potentially have an impact on the performance of interventions such as those considered here and their ability to suppress sporadic outbreaks in the shortest time. The extent of compliance with a measure that is considered primary for malaria control by the WHO may influence the performance of alternative interventions reliant on {repellent-based personal protection}. This result stresses the importance for a multi-modal approach in controlling vector-borne disease outbreaks.

    Our model has several limitations: first, it assumes that the host population is homogeneous, and does not consider age structure or differentiation in terms of disease status (such as asymptomatic, mild, severe and in need of treatment). Second, the host and vector populations are taken constant in time, ignoring effects of seasonal changes on the transmission dynamics. Third, it assumes that all hosts receive the same level of protection by the control effort, but population heterogeneity due to variability in repellent action may exist. Finally, (2.1) does not model immunity in the population. These issues can be addressed by increasing model complexity, including seasonal variations in the mosquito population, adding more compartments for the host population, etc. The state constraint (infection cap) can also be redefined to the respective class of infected (e.g., those in need of treatment). The proposed analytical framework can be adapted for extensions of the model (2.1), such as a time-delay system, a combination of measures (use of insecticides or vaccinations), or a system with an infected and infectious vector compartments [42].

    We have used optimal control theory to solve problems of existence of viable controls and of reachability for a simple model of a vector-borne disease. Available mathematical tools exist that allow decision and policy makers to assess effects of the control intervention on transient dynamics of the model in finite times, rather than just to determine asymptotic properties of equilibria. This approach reveals that, sometimes, depending on the epidemiological parameters and the state of the epidemic (determined by the size of infected host and vector compartments), even extending control efforts to the maximum may not suffice to meet policy objectives such as reduction of the size of the infected compartment below a given level in a fixed time span. Hence, existence of controls that satisfy the constraints of the intervention can be interpreted as an efficiency metric, which suggest whether the current intervention is sufficient or a combination of additional intervention modes may be required. Of course, to be able to make reliable predictions for the performance of a given control intervention, the model should use parameters rightful for the provided epidemiological and/or entomological data.

    The presented model is based on assumptions of population homogeneity, simple disease dynamics and linear cost function for the control, and bringing the model closer to real-world scenarios will undoubtedly increase its complexity. Adding additional state variables means adding an extra dimension to the Hamilton-Jacobi-Bellman equation, and increases the computational complexity for the numerical approximation of the value function and the minimal entry time function. This obstacle should be addressed by development of appropriate, more efficient numerical methods for solving the associated optimal control problem.

    The author would like to thank Bob W. Kooi for helpful discussion during the preparation of the manuscript. This work is supported by the Bulgarian National Science Fund (FNI) within the National Scientific Program Petar Beron i NIE of the Bulgarian Ministry of Education [contract KP-06-DB-5]

    The author declares there is no conflict of interest.

    Proof of Lemma 2. For part (i), consider the Lyapunov candidate function on Ω:

    L=β(1kˉu)γy1+y2.

    Note that on ΩR2+, L0 with equality attained only at the disease-free equilibrium at the origin O.

    Note that

    dLdt=(αβ(1kˉu)2νγμ)y2αβ(1kˉu)2νγy1y2β(1kˉu)y1y2<0

    everywhere on R2+O because

    αβ(1kˉu)2νγμ=μ(αβ(1kˉu)2νγμ1)0.

    Hence L is positive definite, |L|+ on R2+ and dL/dt is negative definite on R2+O. Then Lyapunov's stability theorem [43,Chapter 5.26] implies that the disease-free equilibrium O is globally asymptotically stable.

    For part (ii), the global asymptotic stability of the endemic equilibrium (3.5) for the 2-dimensional system (2.1) is shown using the Lyapunov candidate function:

    L=β(1y2)y1(y1y1y1lny1y1)+αν(1y1)y2(y2y2y2lny2y2).

    Note that L>0 and |L|+ on Ω (for y10,y20). After some algebraic transformations we arrive to

    dLdt=αβν(1y2)y1y2y1(y1y1)2αβν(1y1)y2y1y2(y2y2)2αβν(1y1)(1y2)(y1y2y1y2y1y2)20.

    As y1,y2>0, equality dL/dt=0 can be achieved only at E. Thus L is positive definite and dL/dt is negative definite on ΩO. Lyapunov's stability theorem implies that the endemic equilibrium E is globally asymptotically stable.

    Proof of Proposition 1. The proof of case (i) is based on the local asymptotic stability of the disease-free or the endemic equilibrium of the system (2.1) with constant controls. In particular, we distinguish between 2 cases: first, if Icrit0, then O is globally asymptotically stable for (2.1) with u(t)ˉu (see Lemma 2). This means that trajectories starting from any neighbourhood of O that is sufficiently small and contained inside A(ˉI) will converge to the equilibrium, maintaining the constraint on x1(t)<ˉI for all t>0.

    Second, if Icrit>0, then the endemic equilibrium E is globally asymptotically stable for (2.1) with u(t)ˉu, and EA(ˉI) (see Lemma 2). To show that intV has a positive Lebesgue measure in R2, we consider the lens-shaped set D0 bounded by the x1- and x2-nullcline inside A(ˉI), and show it is a viability domain for (2.1) (see Figure S1). The x1-nullcline is given by

    N1={(x1,x2)|0x1Icrit,x2=γx1α(1kˉu)ν(1x1)} (6.1)

    with the outward-pointing normal

    nN1=(1,γα(1kˉu)ν(1x1)2)T.

    For xN1, it holds f2(x,ˉu)=x1(αβν(1kˉu)2γμ)x21(αβν(1kˉu)2+γβ(1kˉu))α(1kˉu)ν(1x1)0 with equality at O and E, so f(x,ˉu),nN1=γf2(x,ˉu)α(1kˉu)ν(1x1)20, and the curve (6.1) is impermeable from intD0 for the flow f(x,ˉu). The x2-nullcline is given by

    N2={(x1,x2)|0x1Icrit,x2=β(1kˉu)x1β(1kˉu)x1+μ} (6.2)

    with the outward-pointing normal

    nN2=(1,β(1kˉu)μ(β(1kˉu)x1+μ)2)T.

    For xN2, it holds f1(x,ˉu)=x1(αβν(1kˉu)2γμ)x21(αβν(1kˉu)2+γβ(1kˉu))β(1kˉu)x1+μ0 again with equality at O and E, so f(x,ˉu),nN2=f1(x,ˉu)0 and the curve (6.2) is impermeable from intD0 for the flow f(x,ˉu). We conclude that this lens-shaped set D0 is forward invariant under f(x,ˉu) and D0A(ˉI) is a viability domain, completing the proof of case (i).

    The proof of (ii) uses the componentwise inequality f(x,u)f(x,ˉu),xΩ to invoke the comparison theorem [20] and obtain

    xu(t;x0)xˉu(t;x0),t>0 (6.3)

    so xˉu(t;x0) is a subsolution for (2.1) for all x0. Lemma 2 establishes that for any initial condition x0ΩO, this subsolution converges to EΩA(ˉI). In other words, there exists τ>0 such that for the subsolution xˉu(t;x0), xˉu1(t)>ˉI for all t>τ. The comparison (6.3) implies that for the any feasible trajectory xu(t;x0) of the control system (2.1), xu1(t)xˉu1(t)>ˉI for all t>τ. Thus, the only forward invariant set inside V(ˉI,ˉu) is the disease-free equilibrium O.

    Proof of Lemma 3. Denote for the sake of shortness ˜α=αν(1kˉu),˜β=β(1kˉu), and ˜f1(z,s)=f1(z,s,ˉu),˜f2(z,s)=f2(z,s,ˉu), and H(s,z)=˜f1(z,s)˜f2(z,s) the right-hand side of (3.6).

    Figure 1.  Phase portrait of the model with nullclines f1 = f2 = 0.

    Note that due to the choice of ˉv, the numerator ˜f1 of H is 0 at s=ˉv. Further, the denominator of H at (ˉv,ˉI) satisfies

    ˜f2(ˉI,ˉv)=˜βˉI(1ˉv)μˉv=ˉI˜α(1ˉI)(˜α˜βγμˉI˜β(˜α+γ))<0, (6.4)

    so long as ˜α˜βγμ0 (when Icrit0) or when ˉI>Icrit,˜α˜βγμ>0. Therefore,

    ddsz(ˉv)=0. (6.5)

    We have

    ˜f1(z,s)>˜f1(˜z,˜s),˜f2(z,s)<˜f2(˜z,˜s) if s>˜s,z<˜z, (6.6)

    and in particular, ˜f1(z,s)>˜f1(ˉI,ˉv) and ˜f2(z,s)<˜f2(ˉI,ˉv) if s>ˉv,z<ˉI. Also we have max˜f1(z,s)<+,min˜f2(z,s)> on Ω.

    This implies that the denominator of H is negative and bounded away from 0 on [ˉv,1]×[0,ˉI] due to (6.4). Consequently, being a rational function, H is continuous in s and uniformly Lipschitz in z on [ˉv,1]×[0,ˉI]. By the Picard-Lindelöf theorem, there exists ε>0 such that (3.6) has a unique C1-solution z(s) for s[ˉv,ˉv+ε). We show now that this solution z(s) exists on the entire interval [ˉv,+).

    We verify that z(s) is monotone decreasing on [ˉv,ˉv+ε) based on (6.5). Since z is C1, its Taylor expansion around ˉI=z(ˉv) reads

    z(s)=ˉI+O((sˉv)2).

    Therefore, using the relation ˜f1(ˉI,ˉv)=0:

    ˜f1(z(s),s)=˜α(sˉv)(1ˉI)(γ+˜αs)(z(s)ˉI)=˜α(sˉv)(1ˉI)(γ+˜αs)O((sˉv)2),˜f2(z(s),s)˜f2(ˉI,ˉv)=μ(sˉv)˜βz(s)(sˉv)+˜β(1ˉv)O((sˉv)2)=(μ+˜βˉI)(sˉv)+β(1ˉv)O((sˉv)2).

    Thus, there exists ε>0 such that for values ˉv<s<ˉv+ε, ˜f1(z(s),s)>0, while ˜f2(z(s),s)<0 due to (6.4). Hence, the solution z is monotone decreasing on [ˉv,ˉv+ε), and due to the monotonicity of ˜f1,˜f2 provided by (6.6), H<0 strictly on (ˉv,+)×[0,ˉI). Because the solution z(s),s>ˉv is bounded by z(ˉv)=ˉI irrespectively of ε, it can be continued on the whole interval [ˉv,+, and remains monotone decreasing because dz/ds<0.

    Note that it might happen that z(v)=0 for some ˉv<v1. Then we restrict the solution curve Z to the domain [ˉv,v]. This completes the proof.

    Proof of Proposition 2. The proof follows the ideas from reference [17]. To show that the proposed sets are viability kernels, we have to verify their forward invariance under f(x,u) (that is, that they are viability domains for f) and their maximality.

    Case (i). Consider the segment s={ˉI}×[0,1], which forms the eastern boundary of A(ˉI) in the (x1,x2)-state space and the outward-pointing normal to s, ns=(1,0)T. Then f(ˉI,x2,ˉu),ns=α(1kˉu)νx2(1ˉI)γˉI(1kˉu)αν(1ˉI)γˉI<0. This means there exists a control function uU such that the entire set A(ˉI) is forward invariant under (2.1), and therefore, V(ˉI,ˉu)=A(ˉI).

    Case (ii). Let D={[0,ˉI]×[0,ˉv]}{(x1,x2)|ˉvx2min{v,1},0x1z(x2)}. Recall ˉv from Lemma 3. Observe that on the segment s2={ˉI}×(ˉv,1], the inner product of the velocity field f and the outward-pointing normal n2 to s2, f(ˉI,x2,u),n2=α(1ku)νx2(1ˉI)γˉI>(1ku1kˉu1)γˉI0. Hence, by continuity of f, a neighbourhood of s2 in A(ˉI) is not contained inside V(ˉI,ˉu). On the other hand, on the segment s1={ˉI}×[0,ˉv] the inner product of the velocity field f and the outward-pointing normal n1 to s1, f(ˉI,x2,ˉu),n10. Finally, the outward-pointing normal to Z is ˜n=(1,z(x))T. Then f(z(x2),x2,ˉu),˜n)=0. Hence, by choosing u=ˉu, Z is impermeable to the flow of f(x,ˉu). Thus, there exists u such that the interior intD is forward invariant under f(x,u).

    Finally to conclude that D is forward invariant, we show that the curve ZD, itself being a feasible trajectory for (2.1) with u=ˉu. Indeed, the inequality x02>ˉv,x01<ˉI holds for any initial data x0=(x01,x02)Z. Therefore, ˜f2(x01,x02)<˜f2(ˉI,ˉv)<0 by (6.4), and since ddtx2(x0)<0, it follows x2(t)<x02,t>0.

    Let τB(x0) be the minimal entry time function (4.1) to the target set B={(x1,x2)| x1>ˉIx2<ˉv} for the control u(t)=ˉu, or

    τB(x0)=inft>0{t|xˉu(t;x0)B}.

    Since B contains the globally asymptotically stable endemic equilibrium E (3.5), τ(x0)<+ for all x0Z.

    We choose any initial condition x0Z. Then the trajectory of (2.1) for u=ˉu satisfies x2(t)<ˉv for t>τ(x0). Thus, {x2(t)|t[0,τ(x0)]}[ˉv,min{v,1}] and x2(t) takes values inside the domain of the function z which solves the problem (6.4). On t[0,τ(x0)] the difference between x1(t) and z(x2(t) remains constant in t because

    ddt(x1(t)z(x2(t))=f1(x,ˉu)f2(x,ˉu)dzdx2=0,

    Due to the initial condition x0Z, x1(0)z(x2(0))=0, so the above implies x1(t)z(x2(t))=0,t(0,τ(x0)) implying that the curve Z is part of a feasible trajectory for (2.1).

    As it holds by construction, either x1(τ(x0))=ˉI or x2(τ(x0))=ˉv, by the inverse function theorem for z we obtain that xˉux0(τ(x0))=(ˉI,ˉv). Since x0Z was chosen arbitrary, the entire curve Z as defined by the Lemma consists of a trajectory for (2.1). We have demonstrated that the set D as defined in (3.6) is closed, forward invariant under f(x,u) and it is a viability domain for f(x,u).

    It remains to show that D is the maximal viability domain for f(x,u). Consider an initial condition y0A(ˉI)D for f and {claim} the trajectory xˉu(t,y0) leaves A(ˉI) in finite time. It holds y01<ˉI, but y02>z1(y01) due to the definition of D. Denote ˜y0=(y01,z1(y01)), meaning ˜y0y0 componentwise. Since the map f is quasimonotone, we can use the comparison principle [24,Chapter 3] to obtain the componentwise inequalities

    xu(t;y0)xˉu(t;y0)xˉu(t;˜y0),t>0.

    Therefore, the first element in the difference vector Δ(t)=xˉu(t;y0)xˉu(t;˜y0), which we denote by Δ1(t) satisfies Δ1C1(0,+),Δ1(t)0.

    Denote τ0=τ(˜y0), which means ˜y1(τ0)=[xˉu(τ0;˜y0)]1=ˉI. Assume y1(τ0)=[xˉu(τ0;y0)]1=ˉI, which translates to Δ1(τ0)=0, and Δ1 having a local minimum at t=τ0. Hence,

    ddtΔ1(τ0)=0f1(xˉu(τ0;y0),ˉu)=f1(xˉu(τ0;˜y0),ˉu). (6.7)

    Plugging in the equality y1(τ0)=˜y1(τ0)=ˉI into (6.7), we see that

    ˜α(1ˉI)(y2(τ0)˜y2(τ0))=0y2(τ0)=˜y2(τ0),

    and obtain xˉu(τ0;y0)=xˉu(τ0;˜y0). With y˜y, this equality is a contradiction to the uniqueness of the solution trajectory to the autonomous system ddty=f(y,ˉu). Therefore, y1(τ0)>˜y1(τ0)=ˉI, showing that y0V(ˉI,ˉu). This argument shows D is maximal, and we conclude V(ˉI,ˉu)=D.



    [1] X. Zhu, J. Wang, H. Guo, D. Zhu, L. T. Yang, L. Liu, Fault-tolerant scheduling for real-time scientific workflows with elastic resource provisioning in virtualized clouds, IEEE Trans. Parallel Distrib. Syst., 27 (2016), 3501–3517. https://doi.org/10.1109/TPDS.2016.2543731 doi: 10.1109/TPDS.2016.2543731
    [2] E. Pluzhnik, E. Nikulchev, Virtual laboratories in cloud infrastructure of educational institutions, in 2014 2nd 2014 2nd International Conference on Emission Electronics (ICEE), (2014), 1–3.
    [3] M. Ali, S. U. Khan, A. V. Vasilakos, Security in cloud computing: Opportunities and challenges, Inform. Sci., 305 (2015), 357–383. https://doi.org/10.1016/j.ins.2015.01.025 doi: 10.1016/j.ins.2015.01.025
    [4] P. D. Ezhilchelvan, I. Mitrani, Evaluating the probability of malicious co-residency in public clouds, IEEE Trans. Cloud Comput., 5 (2015), 420–427. https://doi.org/10.1109/TCC.2015.2451633 doi: 10.1109/TCC.2015.2451633
    [5] H. El Merabet, A. Hajraoui, A survey of malware detection techniques based on machine learning, Int. J. Adv. Comput. Sci. Appl., 10 (2019). https://doi.org/10.14569/IJACSA.2019.0100148
    [6] K. Lu, J. Cheng, A. Yan, Malware detection based on the feature selection of a correlation information decision matrix, Mathematics, 11 (2023), 961. https://doi.org/10.3390/math11040961 doi: 10.3390/math11040961
    [7] T. Li, Y. Liu, Q. Liu, W. Xu, Y. Xiao, H. Liu, A malware propagation prediction model based on representation learning and graph convolutional networks, Digital Commun. Networks, 2022. https://doi.org/10.3390/math11040961
    [8] Y. Ye, T. Li, D. Adjeroh, S. S. Iyengar, A survey on malware detection using data mining techniques, ACM Comput. Surv., 50 (2017), 1–40. https://doi.org/10.1145/3073559. doi: 10.1145/3073559
    [9] T. Li, Y. Liu, X. Wu, Y. Xiao, C. Sang, Dynamic model of malware propagation based on tripartite graph and spread influence, Nonlinear Dyn., 101 (2020), 2671–2686. https://doi.org/10.1007/s11071-020-05935-6 doi: 10.1007/s11071-020-05935-6
    [10] F. Mira, A systematic literature review on malware analysis, in 2021 IEEE International IOT, Electronics and Mechatronics Conference (IEMTRONICS), (2021), 1–5. https://doi.org/10.1109/IEMTRONICS52119.2021.9422537
    [11] Q. Zhu, Y. Liu, X. Luo, K. Cheng, A malware propagation model considering conformity psychology in social networks, Axioms, 11 (2022). https://doi.org/10.3390/axioms11110632
    [12] X. Ye, S. Xie, S. Shen, Sir1r2: Characterizing malware propagation in wsns with second immunization, IEEE Access, 9 (2021), 82083–82093. https://doi.org/10.1109/ACCESS.2021.3086531 doi: 10.1109/ACCESS.2021.3086531
    [13] N. P. Dong, H. V. Long, N. T. K. Son, The dynamical behaviors of fractional-order se1e2iqr epidemic model for malware propagation on wireless sensor network, Commun. Nonlinear Sci. Numerical Simul., 111 (2022), 106428. https://doi.org/10.1016/j.cnsns.2022.106428 doi: 10.1016/j.cnsns.2022.106428
    [14] S. M. Al-Tuwairqi, W. S. Bahashwan, The impact of quarantine strategies on malware dynamics in a network with heterogeneous immunity, Math. Model. Anal., 27 (2022), 282–302. https://doi.org/10.3846/mma.2022.14391 doi: 10.3846/mma.2022.14391
    [15] A. Martin del Rey, G. Hernandez, A. Bustos Tabernero, A. Queiruga Dios, Advanced malware propagation on random complex networks, Neurocomputing, 423 (2021), 689–696. https://doi.org/10.1016/j.neucom.2020.03.115 doi: 10.1016/j.neucom.2020.03.115
    [16] J. R. C. Piqueira, M. A. Cabrera, C. M. Batistela, Malware propagation in clustered computer networks, Phys. A Stat. Mech. Appl., 573 (2021), 125958. https://doi.org/10.1016/j.physa.2021.125958 doi: 10.1016/j.physa.2021.125958
    [17] W. Zhang, Z. Wang, Z. Zhang, J. Zou, Delay effect on a malware propagation model incorporating user awareness, in 2022 International Conference on Cyber-Physical Social Intelligence (ICCSI), (2022), 555–560. https://doi.org/10.1109/ICCSI55536.2022.9970556
    [18] L. Li, J. Cui, R. Zhang, H. Xia, X. Cheng, Dynamics of complex networks: Malware propagation modeling and analysis in industrial internet of things, IEEE Access, 8 (2020), 64184–64192. https://doi.org/10.1109/ACCESS.2020.2984668 doi: 10.1109/ACCESS.2020.2984668
    [19] M. N. Aman, U. Javaid, B. Sikdar, Iot-proctor: A secure and lightweight device patching framework for mitigating malware spread in iot networks, IEEE Syst. J., 16 (2022), 3468–3479. https://doi.org/10.1109/JSYST.2021.3070404 doi: 10.1109/JSYST.2021.3070404
    [20] S. Hosseini, M. A. Azgomi, Dynamical analysis of a malware propagation model considering the impacts of mobile devices and software diversification, Phys. A Stat. Mech. Appl., 526 (2019), 120925. https://doi.org/10.1016/j.physa.2019.04.161 doi: 10.1016/j.physa.2019.04.161
    [21] S. Hosseini, Defense against malware propagation in complex heterogeneous networks, Cluster Comput., 24 (2021), 1199–1215. https://doi.org/10.1007/s10586-020-03181-4 doi: 10.1007/s10586-020-03181-4
    [22] R. Hassan, S. Rafatirad, H. Homayoun, S. M. P. Dinakarrao, Performance-aware malware epidemic confinement in large-scale iot networks, in ICC 2021 - IEEE International Conference on Communications, (2021), 1–6. https://doi.org/10.1109/ICC42927.2021.9500476
    [23] S. Shen, H. Zhou, S. Feng, J. Liu, H. Zhang, Q. Cao, An epidemiology-based model for disclosing dynamics of malware propagation in heterogeneous and mobile wsns, IEEE Access, 8 (2020), 43876–43887. https://doi.org/10.1109/ACCESS.2020.2977966 doi: 10.1109/ACCESS.2020.2977966
    [24] L. Miao, S. Li, Stochastic differential game-based malware propagation in edge computing-based iot, Secur. Commun. Networks, 2021 (2021), 1–11. https://doi.org/10.1155/2021/8896715 doi: 10.1155/2021/8896715
    [25] V. S. Varma, Y. Hayel, I.-C. Morarescu, A non-cooperative resource utilization game between two competing malware, IEEE Control Syst. Lett., 7 (2023), 67–72. https://doi.org/10.1109/LCSYS.2022.3186620 doi: 10.1109/LCSYS.2022.3186620
    [26] L. Wang, S. S. Iyengar, A. K. Belman, P. Śniatała, V. V. Phoha, C. Wan, Game theory based cyber-insurance to cover potential loss from mobile malware exploitation, Digital Threats Res. Pract., 2 (2021), 1–24. https://doi.org/10.1145/3409959 doi: 10.1145/3409959
    [27] H. Zhou, S. Shen, J. Liu, Malware propagation model in wireless sensor networks under attack-defense confrontation, Comput. Commun., 162 (2020), 51–58. https://doi.org/10.1016/j.comcom.2020.08.009 doi: 10.1016/j.comcom.2020.08.009
    [28] Z. Benomar, C. Ghribi, E. Cali, A. Hinsen, B. Jahnel, Agent-based modeling and simulation for malware spreading in d2d networks, preprint, arXiv: 2201.12230.
    [29] F. Abazari, M. Analoui, H. Takabi, Effect of anti-malware software on infectious nodes in cloud environment, Comput. Secur., 58 (2016), 139–148. https://doi.org/10.1016/j.cose.2015.12.002 doi: 10.1016/j.cose.2015.12.002
    [30] C. Gan, Q. Feng, X. Zhang, Z. Zhang, Q. Zhu, Dynamical propagation model of malware for cloud computing security, IEEE Access, 8 (2020), 20325–20333. https://doi.org/10.1109/ACCESS.2020.2968916 doi: 10.1109/ACCESS.2020.2968916
    [31] M. I. Kamien, N. L. Schwartz, Dynamic optimization: the calculus of variations and optimal control in economics and management, Courier Corporation, 2012.
    [32] E. Pluzhnik, E. Nikulchev, S. Payain, Optimal control of applications for hybrid cloud services, in 2014 IEEE World Congress on Services, 2014,458–461. https://doi.org/10.1109/SERVICES.2014.88
    [33] Q. Zhu, X. Yang, L. X. Yang, C. Zhang, Optimal control of computer virus under a delayed model, Appl. Math. Comput., 218 (2012), 11613–11619. https://doi.org/10.1016/j.amc.2012.04.092 doi: 10.1016/j.amc.2012.04.092
    [34] L. Chen, K. Hattaf, J. Sun, Optimal control of a delayed slbs computer virus model, Phys. A Stat. Mech. Appl., 427 (2015), 244–250. https://doi.org/10.1016/j.physa.2015.02.048 doi: 10.1016/j.physa.2015.02.048
    [35] L. X. Yang, M. Draief, X. Yang, The optimal dynamic immunization under a controlled heterogeneous node-based sirs model, Phys. A Stat. Mech. Appl., 450 (2016), 403–415. https://doi.org/10.1016/j.physa.2016.01.026 doi: 10.1016/j.physa.2016.01.026
    [36] R. C. Robinson, An introduction to dynamical systems: Continuous and discrete, American Mathematical Soc., 2012.
    [37] J. Stewart, Multivariable calculus: Concepts and contexts, Cengage Learning, 2018.
    [38] D. Liberzon, Calculus of variations and optimal control theory: A concise introduction, Princeton university press, 2011.
  • Reader Comments
  • © 2023 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

Metrics

Article views(2036) PDF downloads(61) Cited by(7)

Figures and Tables

Figures(4)  /  Tables(1)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog