Skip to main content

ORIGINAL RESEARCH article

Front. Appl. Math. Stat., 01 March 2024
Sec. Mathematical Biology
This article is part of the Research Topic Modeling Ecological and Epidemiological Interactions for Sustainable Practices View all 5 articles

Optimal control and stability analysis of monkeypox transmission dynamics with the impact of contaminated surfaces

  • 1Department of Mathematics, Faculty of Mathematics and Natural Sciences, Universitas Indonesia, Depok, Indonesia
  • 2Institute of Mathematical Sciences, Faculty of Science, Universiti Malaya, Kuala Lumpur, Malaysia

This study presents a comprehensive analysis of the transmission dynamics of monkeypox, considering contaminated surfaces using a deterministic mathematical model. The study begins by calculating the basic reproduction number and the stability properties of equilibrium states, specifically focusing on the disease-free equilibrium and the endemic equilibrium. Our analytical investigation reveals the occurrence of a forward bifurcation when the basic reproduction number equals unity, indicating a critical threshold for disease spread. The non-existence of backward bifurcation indicates that the basic reproduction number is the single endemic indicator in our model. To further understand the dynamics and control strategies, sensitivity analysis is conducted to identify influential parameters. Based on these findings, the model is reconstructed as an optimal control problem, allowing for the formulation of effective control strategies. Numerical simulations are then performed to assess the impact of these control measures on the spread of monkeypox. The study contributes to the field by providing insights into the optimal control and stability analysis of monkeypox transmission dynamics. The results emphasize the significance of contaminated surfaces in disease transmission and highlight the importance of implementing targeted control measures to contain and prevent outbreaks. The findings of this research can aid in the development of evidence-based strategies for mitigating the impact of monkeypox and other similar infectious diseases.

1 Introduction

The global outbreak of monkeypox in May 2022 has highlighted the significant public health risks associated with zoonotic diseases [1]. Monkeypox, caused by the monkeypox virus, primarily spreads to humans through contact with infected animals, but human-to-human transmission can also occur through various routes [2, 3]. Contaminated surfaces play a crucial role in the transmission of diseases. Studies indicate that viral transmission can take place through contaminated surfaces, and mounting evidence underscores the pivotal role of contaminated fomites or surfaces in the spread of viral diseases [4]. Specifically, the environmental viral load significantly influences the transmission of the monkeypox virus, with direct contact with contaminated items or materials emerging as a major contributing factor to its spread [5]. The disease manifests in distinct phases, including an incubation period, a prodromal phase, and an eruptive stage characterized by skin lesions [2, 6]. The ongoing monkeypox outbreak has become a public health emergency of global concern, with over 63,000 cases reported, including its spread to locations such as the United Kingdom [7, 8]. This outbreak poses a strain on healthcare systems already dealing with the COVID-19 pandemic [9]. Although monkeypox cases are typically mild, at-risk groups such as children, gravid women, and individuals with compromised immune systems may experience severe cases [10, 11]. Historically, monkeypox outbreaks have been prevalent in marginalized communities in Africa, where the disease became endemic [1214]. Concerns have been raised due to waning immunity resulting from the discontinuation of smallpox vaccination and the evolving epidemiology of the virus [15, 16]. To effectively address the monkeypox outbreak, it is crucial to develop robust disease modeling strategies for public health planning and response.

Previously, there has been limited focus on monkeypox, resulting in a lack of understanding of the disease's transmission dynamics. However, a few numbers of studies have made efforts to employ mathematical modeling techniques to gain insights into the monkeypox virus's dynamics. Mathematical modeling has been studied in Al-Shomrani et al. [17] to analyze the interaction between humans and animals in relation to the Monkeypox virus, while Okyere and Ackora-Prah[18] explores the mechanism of human transmission of the virus through a mathematical study. The authors in Peter et al. [19] investigated the transmission dynamics of the monkeypox virus and found that isolating infected individuals in the human population effectively reduces disease transmission. Alharbi et al. [20] explored the effectiveness of treatment and vaccination interventions as containment measures for monkeypox. In a study by Michael et al. [21], the authors examined a mathematical model for monkeypox that incorporated surveillance as a control measure. Additional valuable contributions can be found in El-Mesady et al. [22], Alshehri and Ullah [23], and Alzubaidi et al. [24]. According to the existing literature, it is evident that further research and exploration are necessary to gain a deeper understanding of the phenomenon of monkeypox. This study aims to investigate the dynamics of monkeypox transmission and its control in humans by employing a classical and deterministic model that incorporates contaminated surfaces as a separate compartment.

In our study, we present a novel contribution to the field by adopting a deterministic approach, distinguishing our study from previous studies conducted by Addai et al. [5] and Li et al. [25], who utilized the Caputo fractional derivative. Furthermore, while Madubueze et al. [26] employed a deterministic model and made commendable progress, our study extends beyond by conducting more in-depth analyses in optimal control and numerical simulations for the proposed strategies. Additionally, our research sets itself apart from the study by Alshehri et al. [23], who also employed a deterministic model by incorporating an exposed population variable into our model and conducting a comprehensive cost-effectiveness analysis for the implemented strategies. Our strong emphasis on these aspects showcases the distinctive contributions and advancements made in our study compared to the existing literature.

The subsequent sections of this article are outlined in the following manner: In Section 2, we explain the construction of the mathematical model. Next, Section 3 focuses on analyzing the model, including the disease-free equilibrium, basic reproduction number, endemic equilibrium, and global stability. We conduct a bifurcation analysis in Section 4. Section 5 presents a sensitivity analysis of the model. Section 6, explores an optimal control problem, including its characterization, simulation, and discussion. Finally, we conclude the article by summarizing the key findings in Section 7.

2 Model construction

In this section, we present the construction of our mathematical model for the optimal control and stability analysis of monkeypox transmission dynamics, considering the impact of contaminated surfaces. In developing the model, an assumption is made that the population of humans (N) remains constant over a period of time. The model is developed based on compartmentalization, where individuals are classified into different compartments representing their disease status. We consider the following compartments: susceptible (S), exposed (E), infectious (I), recovered (R), and contaminated surfaces (C). The explanations of the model construction for each compartment are constructed using Figure 1 and given as follows:

Susceptible (S). This compartment represents individuals who are susceptible to contracting monkeypox. They have not been exposed to the virus and can become infected if they come into contact with infectious individuals or contaminated surfaces. The rate of change in the susceptible compartment is determined by multiple factors. First, the birth rate of the human population (Λh) contributes to the increase in the number of susceptible individuals. Second, the term (((βhI)N)+βcC)S represents the combined effect of virus transmission from infectious individuals (I) and contaminated surfaces (C) to susceptible individuals, where βh and βc represent the transmission rate of monkeypox through contact of human to other human or contaminated surfaces, respectively. This term signifies the rate at which susceptible individuals become infected. In addition, it is important to note that the natural mortality rate denoted by (μh) applies to all compartments, including the infected and recovered. This rate contributes to the decrease in the respective populations due to non-disease-related causes.

Exposed (E). The exposed compartment represents individuals who have been exposed to the monkeypox virus but are not yet infectious. During this period, the virus is incubating within their bodies, and they do not exhibit symptoms or pose an immediate risk of transmission. However, they can potentially transmit the virus to susceptible individuals or contaminated surfaces. The rate of change in the exposed compartment depends on various factors. The term (((βhI)N)+βcC)S represents the rate at which exposed individuals become infected, considering virus transmission from infectious individuals (I) and contaminated surfaces (C). Additionally, γE accounts for the transition from the exposed to the infectious stage, where γ represents the transition rate due to infection progression. This term represents the rate at which exposed individuals progress to the infectious compartment.

Infectious (I). The infectious compartment consists of individuals who have become infectious and can spread the monkeypox virus to susceptible individuals or contaminated surfaces. They may exhibit symptoms of the disease and contribute to the transmission dynamics. The rate of change in the infectious compartment is influenced by several factors. First, γE represents the rate at which exposed individuals progress to the infectious stage, reflecting the transition from the exposed compartment to the infectious compartment. Moreover, γ0I signifies the rate at which infectious individuals recover and transition to the recovered compartment, where γ0 represents the recovery rate from monkeypox.

Recovered (R). This compartment represents individuals who have recovered from the monkeypox infection and developed immunity. The rate of change in the recovered compartment is primarily determined by the rate at which infectious individuals recover and transition to the recovered compartment, denoted as γ0I.

Contaminated surfaces (C). The contaminated surfaces compartment represents surfaces or environments that have been contaminated by the monkeypox virus. These surfaces can act as sources of infection if individuals come into contact with them. The virus can persist on these surfaces and potentially contribute to the transmission of monkeypox. The rate of change in the contaminated surfaces compartment is influenced by two main factors. First, αI denotes the rate at which infectious individuals contaminate surfaces, indicating the rate at which the virus is deposited on surfaces. Second, μpC represents the cleaning rate or decay rate of the virus on contaminated surfaces, signifying the rate at which the virus is removed or becomes inactive on surfaces.

Figure 1
www.frontiersin.org

Figure 1. Transmission diagram depicting the spread of the disease in the model.

The model can be represented by the following set of differential equations:

dSdt=Λh-((βhIN)+βcC)S-μhS,dEdt=((βhIN)+βcC)S-(γ+μh)E,dIdt=γE-(γ0+μh)I,dRdt=γ0I-μhR,dCdt=αI-μpC.    (1)

The initial conditions of the model (Equation 1) are non-negative:

S(0)0,E(0)0,I(0)0,R(0)0,C(0)0.    (2)

All parameters in the system are positive. The specific values of these parameters and their descriptions are provided in Table 1. Before doing additional analysis on the model, it is essential to ensure that the solutions of system (Equation 1) exhibit both positivity and boundedness. These properties are stated in the following theorems.

Table 1
www.frontiersin.org

Table 1. Parameter descriptions and values for the basic model (1).

Theorem 1. Given non-negative initial conditions (Equation 2), every solution (S, E, I, R, C) of model (Equation 1) is positive for all t ≥ 0.

Proof. In the boundary region of 05, we have system (Equation 1) equals to:

dSdt|S=0,E0,I0,R0,C0=Λh0,dEdt|S0,E=0,I0,R0,C0=((βhIN)+βcC)S0,dIdt|S0,E0,I=0,R0,C0=γE0,dRdt|S0,E0,I0,R=0,C0=γ0I0,dCdt|S0,E0,I0,R0,C=0=αI0.

The computation above shows that on the boundary planes of the non-negative of 05, all of the rates are non-negative. Consequently, all of the vector field direction is inward from the boundary planes. As a result, initiating the system from non-negative initial conditions ensures that all solutions remain positive.

Theorem 2. The solutions of model (Equation 1) are bounded in the region

Ω={(S,E,I,R,C)+5|Nmax{N(0),Λhμh},CαΛhμhμp}.

Proof. Combining the first four equations within the model (Equation 1) yields an equation that describes the overall population as follows:

dNdt=Λh-μhN,dNdt+μhN=Λh.

By applying the integrating factor method using exp(0tμhdt)=exp(μht), it can be established that N(t) satisfies:

N=Λhμh+N(0)exp(-μht).

If we take the limit as t → ∞, we have NΛhμh. If the initial condition is larger than Λhμh, then the solution will monotonically decrease and tends to Λhμh. If N(0)<Λhμh, then the solution will monotonically increase and tends to Λhμh. On the other hand, if N(0)=Λhμh, then the solution will remain constant at Λhμh. Hence, we have N(t)max{N(0),Λhμh}.

For the variable C, since IN, we got

dCdt=αI-μpCαΛhμh-μpC,dCdt+μpCαΛhμh.

Moreover, employing the integrating factor method with exp(0tμpdt)=exp(μpt), we get

CαΛhμhμp+C(0)exp(-μpt).

With t → ∞, CαΛhμhμp. Hence, the theorem is proven.

To facilitate further analysis of the model, we apply scaling techniques to normalize the variables. By introducing the scaling variables s=SN,e=EN,i=IN,r=RN, and c=CN, where N represents the total population size, we can express the model in dimensionless form. Additionally, we redefine the transmission rate of contaminated surfaces as βp = βcN. Utilizing the relationship s = 1 − eir, the final model is obtained:

dsdt=(1-s)μh-(βhi+βpc)s,dedt=(βhi+βpc)s-(γ+μh)e,didt=γe-(γ0+μh)i,dcdt=αi-μpc.    (3)

In this transformed form, all variables and parameters in model (Equation 3) are dimensionless, enabling a more convenient analysis of the model's behavior and stability properties.

3 Analysis of the model

3.1 The monkeypox-free equilibrium and the basic reproduction number

The monkeypox-free equilibrium is a crucial state in the analysis of the proposed model. It represents the equilibrium point of the system in the absence of the monkeypox disease. The monkeypox-free equilibrium is determined by setting all disease-related variables, including the infected variables (e, i, c) to zero [29]. By solving the resulting system of equations, the monkeypox-free equilibrium, denoted as E0, is derived. The equation for the monkeypox-free equilibrium is given by:

E0=(s0,e0,i0,c0)=(1,0,0,0).

With the disease-free equilibrium established, we proceed to calculate the basic reproduction number, denoted as R0. The R0 represents the expected number of secondary cases generated by a single primary case during its infectious period. E0 serves as an indicator of the qualitative behavior of these models, such as the persistence or extinction of the disease [30]. A higher E0 indicates a higher potential for disease spread, while a lower E0 suggests a decreased likelihood of sustained transmission. To calculate the E0 for our model, we adopt the next-generation matrix approach [31], a well-established method for its determination.

The evaluation of the subsystem of model (Equation 1) that specifically focuses on the infected compartment in E0 is given by

ȷ=[-(γ+μh)βhβpγ-(γ0+μh)00α-μp].

The matrix ȷ can be written as the sum of the transmission matrix T and transition matrix Σ, where

T=[0βhβp000000],Σ=[-(γ+μh)00γ-(γ0+μh)00α-μp].

Since the second and third rows of matrix T consist entirely of zeros, the next-generation matrix can be derived as

K=-ETTΣ-1E=[βhγ(γ+μh)(γ0+μh)+βpγα(γ+μh)(γ0+μh)μp],

where E = [1, 0, 0]T. Consequently, the basic reproduction number, calculated as the spectral radius of K, is given by

R0=γ(μpβh+αβp)(γ+μh)(γ0+μh)μp.    (4)

Based on the expression for R0 and the application of Theorem 2 in Driessche and Watmough [32], we establish the local stability of E0, leading to our summarized findings in the following theorem.

Theorem 3. If R0 < 1, then monkeypox-free equilibrium is locally asymptotically stable and unstable if R0 > 1.

Proof. The disease-free equilibrium stability is examined through the linearization of system (Equation 1) at E0, leading to the derivation of the corresponding Jacobian matrix

J|E0=[-μh0-βhβp0-(γ+μh)βhβp0γ-(γ0+μh)000α-μp].

The first eigenvalue λ1 = −μh is negative, whereas the remaining three eigenvalues can be determined as the roots of the following cubic polynomial

P(λ):=λ3+a1λ2+a2λ+a3=0

where

a1=2μh+μp+γ+γ0,a2=μp(2μh+γ+γ0)+αβpγμp+(γ+μh)(γ0+μh)(1-R0),a3=μp(γ+μh)(γ0+μh)(1-R0).

According to the Routh–Hurwitz criteria for a third-degree polynomial [33], the system (Equation 1) exhibits local asymptotic stability when R0 < 1, as indicated by the positive values of a2 and a3. The additional condition of a1a2 > a3 can be easily checked, since a1a2 is equal to a3 plus to some extra positive terms.

3.2 The monkeypox endemic equilibrium

The determination of the monkeypox endemic equilibrium point provides insights into the long-term dynamics of the system, indicating the persistence of infection within the population [34]. We represent it as E1=(s*,e*,i*,c*), where:

s*=1R0,         e*=μhμp(γ0+μh)(R0-1)γ(μpβh+αβp),i*=μhμp(R0-1)(μpβh+αβp),         c*=αμh(R0-1)(μpβh+αβp).

Given that the expression for the endemic equilibrium point is consistently positive when R0 > 1, we establish the following theorem regarding its existence.

Theorem 4. The monkeypox model within system (Equation 1) has a unique endemic equilibrium E1 whenever R0 > 1.

The next theorem provides a summary of the local stability analysis of the endemic equilibrium.

Theorem 5. The monkeypox endemic equilibrium is locally asymptotically stable if R0 > 1.

Proof. The Jacobian matrix evaluated at the endemic equilibrium E1 is

J|E1=[-μhR00-βhR0-βpR0-μh(1-R0)-(γ+μh)βhR0βpR00γ-(γ0+μh)000α-μp].

The Jacobian's characteristic equation is given as

λ4+a1λ3+a2λ2+a3λ+a4=0

where

a1=μhR0+2μh+μp+γ+γ0,a2=μhR0(2μh+μp+γ+γ0)+μp(2μh+γ+γ0)+γαβpμpR0,a3=μhμpR0(2μh+γ+γ0)+μhγαβpμpR0+μh(γ+μh)(γ0+μh)(R0-1),a4=μhμp(γ+μh)(γ0+μh)(R0-1).

It can be seen that ai for i = 1, 2, 3, 4 is strictly positive when R0 > 1. In addition, it has been verified that a1a2a3>a32+a12a4 by checking that a1a2a3-a32+a12a4>0 when R0 > 1. Hence, based on the Routh–Hurwitz criterion [33], the endemic equilibrium is found to be locally asymptotically stable if R0 > 1.

3.3 Global stability of the monkeypox-free equilibrium

Here, the monkeypox-free equilibrium global stability for model (Equation 3) is examined. Theorem 6 states the result.

Theorem 6. The monkeypox-free equilibrium E0 is globally asymptotically stable if R0 < 1.

Proof. The global stability of E0 is established using the direct Lyapunov technique. Consider the Lyapunov function on R+4, which is presented by,

L=(s-1-lns)+e+γ+μhγi+βpμpc.

Initially, it is evident that L is equal to zero at the monkeypox-free equilibrium. To establish the positivity of L for all (s, e, i, c) = (1, 0, 0, 0), it suffices to observe the expression

(s-1-lns),

as the function g(s) = s − 1 − ln(s) satisfies a global minimum at s = 1 and g(s) = 0. Consequently, g(s) is greater than zero for all s > 0, excluding s = 1. Therefore, the first term is positive. The other terms are obviously positive as well. Additionally, it is clear that L is radially unbounded.

L is differentiated, then substitution is made using the equations of model (Equation 3), which gives us

dLdt=-μh(s-1)2s+(γ+μh)(γ0+μh)γ(R0-1)i.

It is apparent that the first term is non-positive, and similarly, the second term is non-positive as long as R0 < 1. Hence, it can be established that dLdt<0 holds true for all (s, e, i, c) ≠ (1, 0, 0, 0). Therefore, based on Lyapunov's theorem ([35], Theorem 7.1), the monkeypox-free equilibrium is globally asymptotically stable.

3.4 Stability analysis on the endemic equilibrium

Applying the principles of the center manifold theory made by Castillo-Chavez and Song [36], we make the assumption

s=x1,e=x2,i=x3,c=x4.

Therefore, system (Equation 3) can be reformulated as follows:

f1=x1=(1-μh)x1-(βhx3+βpx4)x1,f2=x2=(βhx3+βpx4)x1-(γ+μh)x2,f3=x3=γx2-(γ0+μh)x3,f4=x4=αx3-μpx4.    (5)

In the case where R0 = 1, let us assume that α is chosen as the bifurcation parameter. Solving for α, from Equation (4) we obtain

α*=(γ+μh)(γ0+μh)μp-γμpβhγβp

Evaluating the Jacobian matrix of system (Equation 5) at the monkeypox-free equilibrium E0, where α = α*, leads to

J|(E0,α*)=[-μh0-βh-βp0-(γ+μh)βhβp0γ-(γ0+μh)000α*-μp].

The computation of the eigenvalues for J|(E0,α*) yields a simple zero eigenvalue, while the remaining three eigenvalues have a negative real part. Thus, we obtain the right eigenvector w=(w1,w2,w3,w4)T associated with the zero eigenvalue by solving J|(E0,α*)w=0. We have

w1=-(γ+μh)(γ0+μh)μp,w2=(γ0+μh)μpμh,w3=βhγμh,w4=((γ+μh)(γ0+μh)-γβh)μh.

Similarly, let the left eigenvector v=(v1,v2,v3,v4)T associated with the zero eigenvalue. Solving the equation vJ|(E0,α*)=0 leads to

v1=0,v2=μpγ,v3=μp(γ+μh),v4=βpγ.

As v1 is zero, the derivatives of f1 is not needed. However, considering the derivatives of f2, f3, and f4, the non-zero second derivatives can be expressed as

2f2x1x3=βh,   2f2x1x4=βp,   2f4x3α*=1.

The determination of the bifurcation coefficients a and b, where their signs indicate the direction of bifurcation, is obtained as

a=k,i,j=1nvkwiwj2fkxixj(0,0)   =2v2w1w32f2x1x3+2v2w1w42f2x1x4   =-2μhμpγβp2(γ+μh)2(γ0+μh)2<0b=k,i,j=1nvkwi2fkxiϕ(0,0)   =v4w32f4x3α*   =βp2γ2μh>0

Since a is negative and b is positive, this shows that the system exhibits a forward bifurcation. Hence, the outcome stated in Theorem 7 is verified.

Theorem 7. At R0 = 1, the system (Equation 3) exhibits a forward bifurcation.

In the following theorem, we show the global stability of the endemic equilibrium in system (Equation 3) at E1.

Theorem 8. The global asymptotic stability of the system (Equation 3) is established for the unique endemic equilibrium E1 when R0 > 1.

Proof. We establish the global stability of the endemic equilibrium of system (Equation 3) on the fixed set Ω = {(s, e, i, c) : s ≥ 0, e ≥ 0, i ≥ 0, c ≥ 0, s + e + i ≤ 1}. It is important to note that the solutions of the system are bounded and, as proven in Theorem (5), the endemic equilibrium E1 exhibits local asymptotic stability. The remaining task is to demonstrate the absence of periodic solutions in system (Equation 3).

We will employ the Dulac Bendixson criterion to establish our result. Adopting the notations specified in Theorem 3.6 of Martcheva's book [35], by choosing the Dulac function D = 1, we obtain the following:

f1x1+f2x2+f3x3+f4x4=-μh-(βhx3+βpx4)-(γ+μh)-μp<0

Thus, there are no periodic orbits in the open first quadrant. This implies the global asymptotic stability of the endemic equilibrium.

3.5 Bifurcation diagram and autonomous simulation

In this section, we visualize the result of Theorems 6 and 7 using a bifurcation diagram which is given in Figure 2A. To perform the simulation, we use the parameter values as follows:

μh=179,μp=15,βp=7.6977×10-8,γ0=17,γ=114,                                        α=0.0004,

except βh, which is set to be a bifurcation parameter. The blue and red curves represent the monkeypox-free equilibrium and monkeypox endemic equilibrium points, respectively. The solid and dotted curves represent the stable and unstable equilibrium, respectively. The point BP is the value of βh such that R0 = 1, where βh = 0.183. It can be seen that when βh < BP, then R0 < 1, which indicates a global stable of monkeypox-free equilibrium. As βh increases, R0 also increases. When βh reaches the BP, then the monkeypox-free equilibrium becomes unstable which is indicated by a dotted blue curve. At the same time, a stable monkeypox endemic equilibrium starts to arise and becomes larger as βh increases. To illustrate the stable equilibrium, we consider two sample points of βh at P1 and P2, i.e., βh = 0.1 at P1 which gives us R0 = 0.546. However, at P2, with βh = 0.3, we have R0 = 1.638. Figure 2B shows all trajectories solutions of i for four different initial condition move toward the monkeypox-free equilibrium i* = 0, while Figure 2C shows all trajectories tend to monkeypox endemic equilibrium i* = 0.0269.

Figure 2
www.frontiersin.org

Figure 2. (A) Bifurcation diagram of system (3) in βhi plane, (B) trajectory of solution at point P1 (when βh = 0.1) for various initial condition tends to monkeypox free equilibrium, and (C) trajectory of solution at point P2 (when βh = 0.3) for various initial condition tends to monkeypox endemic equilibrium.

4 Optimal control problem

4.1 Optimal control model

Here, in this section, we extend our model in Equation (1) by implementing three different control variables, namely the self-protection (u1(t)), hospitalization (u2(t)), and disinfectant to reduce virus in the surface (u3(t)).

1. Self-protection (u1(t)). To reduce human-to-human contact and prevent the spread of monkeypox (monkeypox), several measures can be taken. These include social distancing, practicing good hygiene, and wearing a face mask [27, 37, 38]. We assume that u1(t) is the proportion of humans who are using self-protection. We assume that the government consistently promotes a self-awareness campaign to encourage adherence to self-protection measures within the general population. Hence, we define u1(t)N as the proportion of people using self-protection at time t. There are three possible incidences of contact based on the use of self-protection. First, when both individuals employ self-protection, it is indicated by (u1N)2. Second, when only one uses self-protection, then it is indicated by u1N(1 − u1)N. The last is when both the individuals do not use self-protection, which is indicated by contact between (1 − u1)N and (1 − u1)N. Let us assume that this self-protection will reduce the chance of transmission rate β by a factor ξ1 if people who made contact both using self-protection, ξ2 if only one of them uses self-protection, and no reduction if no one uses self-protection. Hence, the total of new infections of monkeypox when there is a proportion of u1 who use self-protection is given by

Λ(S,I)=ξ1βhu1(t)Su1(t)I+ξ2βhu1(t)S(1-u1(t))I+ξ2βh                    (1-u1(t))Su1(t)I                +βh(1-u1(t))S(1-u1(t))I,                =(ξ1u12(t)+2ξ2u1(t)(1-u1(t))+(1-u1(t))2)βhSI.

For the sake of simplification, we denote k1(u1(t))=(ξ1u12(t)+2ξ2u1(t)(1-u1(t))+(1-u1(t))2).

2. Hospitalization (u2(t)). Individuals admitted to the hospital due to monkeypox are typically provided with supportive measures, including pain relief, antiviral medications, fluid replacement, and interventions to address potential complications such as pneumonia or sepsis. While no specific treatments exclusively target monkeypox, certain antiviral medications, such as tecovirimat (TPOXX), brincidofovir, and cidofovir, can be employed to aid in managing the infection [3941]. We assume only a proportion of monkeypox-infected individuals get hospitalized, with a proportion of u2. These hospitalized individuals will get proper treatment which will increase their recovery rate from γ0 to γ1. Hence, the total of recovered infected individuals is given by

f(I)=γ1u2(t)I+γ0(1-u2(t))I,       =(γ1u2(t)+(1-u2(t))γ0)I.

For the sake of simplification, let k2(u2(t))= (γ1u2(t)+ (1 − u2(t)) γ0).

3. Disinfectant (u3(t)). Regularly disinfecting contaminated surfaces is crucial in controlling the spread of monkeypox, as it reduces the viability and transmission of the virus from surfaces to humans. The lessons we have learned from the COVID-19 pandemic highlight the significance of frequent surface disinfection to prevent the spread of viruses. Effective disinfectants, including those proven effective against SARS-CoV-2, can also be beneficial in combating monkeypox. By implementing rigorous disinfection protocols, especially for frequently touched surfaces and in healthcare settings, we can minimize the risk of surface-mediated transmission of the disease [42, 43]. We assume that the use of disinfectant will kill the free viruses on the environment's surface. We assume this disinfectant is implemented with a rate of u3(t). Hence, with this intervention, the C compartment will reduce not only by μp but also by u3(t).

With the above explanation, the model in Equation (1) now reads as follows.

dSdt=Λh-((k1(u1(t))βhIN)+βcC)S-μhS,dEdt=((k1(u1(t))βhIN)+βcC)S-(γ+μh)E,dIdt=γE-(k2(u2(t))+μh)I,dRdt=k2(u2(t))I-μhR,dCdt=αI-μpC-u3(t)C,    (6)

where k1(u1(t))=(ξ1u12(t)+2ξ2u1(t)(1-u1(t))+(1-u1(t))2) and k2(u2(t)) = (γ1u2(t) + (1 − u2(t))γ0).

Assuming s=SN,e=EN,i=IN,r=RN, c=CN, βp = βcN and r = 1 − sei, then we have system (Equation 6) reduced as follows.

dsdt=(1-s)μh-(k1(u1(t))βhi+βpc)s,dedt=(k1(u1(t))βhi+βpc)s-(γ+μh)e,didt=γe-(k2(u2(t))+μh)i,dcdt=αi-(μp+u3(t))c.    (7)

Our aim is to reduce the proportion of infected individuals e and i and also contaminated surfaces with a minimum cost for intervention ui(t) for i = 1, 2, 3 which represent self-protection, hospitalization, and disinfectant, respectively. Mathematically, it can be done by minimizing the following cost function

J(ui(t))=0tf[ω1e+ω2i+ω3c+ϕ1u12+ϕ2u22+ϕ3u32]dt    (8)

where ωi > 0 for i = 1, 2, 3 represent the weight parameter for e, i, and c, respectively. On the other hand, ϕi > 0 for i = 1, 2, 3 represent the weight cost for control variable u1(t), u2(t), and u3(t), respectively.

If we assume that u1(t) = u1, u2(t) = u2, u2(t) = u2, k1(u1(t)) = k1, and k2(u2(t)) = k2, then the control reproduction number of system (Equation 7) can be expressed as

Rc=γ(βhk1(μp+u3)+βpα)(γ+μh)(μp+u3)(k2+μh).

From the provided expression, it is apparent that Rc(u1 = 0, u2 = 0, u3 = 0) = R0.

4.2 Existence of solution

Theorem 9. There exists an optimal control function ui*,i=1,2,3 as well as its corresponding set of solution s*(t), e*(t), i*(t), c*(t), thereby J(ui*)=min{J(ui)},i=1,2,3.

Proof. In this proof, we follow the process in Xu et al. [44]. The theorem's proof can be demonstrated using the Cesari Theorem [45], which fulfills the following conditions:

1. The set of controls and state variables is non-empty.

2. The set of controls is both closed and convex.

3. The right-hand side of system (Equation 7) is limited by a linear function of the state and control variables.

4. The integrand within the objective function exhibits convexity concerning the input controls ui*,i=1,2,3.

5. There is a constant C1 > 1, along with positive numbers C2 and C3, such that:

J(ui(t))C2(i=13|ui(t)|2)C1/2-C3,i=1,2,3.    (9)

Clearly, when the system exhibits uniform Lipschitz continuity, both set U and the set of solutions for initial values are not empty [46], thereby fulfilling condition 1. Condition 2 and condition 4 are confirmed based on the given definition. Condition 3 can be shown to be true by following the discussion below. This argument is similar to classical arguments that have been presented in Fister et al. [47].

The system (Equation 7) can be expressed as:

G(x)dx(t)dt=Ax+F(x),

where x(t) = [s(t) e(t) i(t) c(t)]T represents the vector of state variables, while matrix A and the non-linear function F(x) are defined as follows:

A=[-μh0000-(γ+μh)000γ-(k2+μh)000α-(μp+u3)],F(x)=[μh-k1βhis-βpcsk1βhis+βpcs00].

As a result of applying the Hölder inequality, F(x) fulfills the following condition:

|F(x1)-F(x2)|P(|s1-s2|+|e1-e2|),

implies

|G(x1)-G(x2)|Q|x1-x2|,    (10)

where Q = max{P, ‖A‖} < ∞.

Hence, the function G(x) exhibits uniform Lipschitz continuity, fulfilling the assumption of condition 1, and meets the prescribed bound (Equation 10) specified in the Cesari Theorem. The control signals are confined within the closed set [uimin,uimax], and the objective function's value in (Equation 8) is bounded in the compact interval [0, 1]. Consequently, the function G(x) is linear in ui, i = 1, 2, 3. This satisfies condition 3.

The objective function (Equation 8) meets the condition (Equation 9), by selecting C1=1.5,C2=min{12τi},i=1,2,3, and C3 = Δ(0). Thus, condition 5 is established.

4.3 Optimal control characterization

The application of Pontryagin's minimum principle [48] leads us to define the Hamiltonian function as

H(ui(t))=ω1e+ω2i+ω3c+ϕ1u12+ϕ2u22+ϕ3u32+λ1                        [(1-s)μh-(k1βhi+βpc)s]                        +λ2[(k1βhi+βpc)s-(γ+μh)e]                        +λ3[γe-(k2+μh)i]+λ4[αi-(μp+u3)c],

where the costate variable λi and its coefficients correspond to the right-hand side of the model for the ith state variable.

The adjoint system can be obtained by calculating the partial derivatives of H with respect to each of the state variables, resulting in the following:

λ1t=-Hs        =λ1(iβhk1+cβp+μh)-λ2(iβhk1+cβp),λ2t=-He        =-ω1+λ2(γ+μh)-λ3γ,λ3t=-Hi        =-ω2+λ1k1βhs-λ2k1βhs+λ3(k2+μh)-λ4α,λ4t=-Hc        =-ω3+λ1βps-λ2βps+λ4(μp+u3).

with the transversality conditions λi(tf) = 0, i = 1, 2, 3.

We can determine the optimal solution for u1, u2, and u3 by utilizing the relevant conditions derived from Pontryagin's principle. This involves differentiating H with respect to each control variable and subsequently solving the resulting equations to obtain the expression for the control variables u1, u2, and u3 as follows:

u1*=-isβh(λ1-λ2)(ξ2-1)is(λ1-λ2)(ξ1-2ξ2+1)βh-ϕ1,u2*=-λ3(γ0-γ1)i2ϕ2,u3*=λ4c2ϕ3.

Thus, the optimal control solution can be precisely characterized by the given formula:

u1(t)=min{max{u1min,u1*},u1max},u2(t)=min{max{u2min,u2*},u2max},u3(t)=min{max{u3min,u3*},u3max}.

5 Numerical experiments

5.1 Sensitivity of the control reproduction number

An elasticity sensitivity analysis is conducted to identify the key parameters influencing the transmission and spread of monkeypox. The sensitivity index is used to assess the relative impact of parameters on the basic reproduction number Rc and aids in designing effective control measures [49]. The parameter values are obtained from Table 1. An index value that is negative (or positive) signifies an inverse (or direct) influence of the respective parameter on Rc. The sensitivity index of R0 with respect to the parameter θ is expressed as

EθRc=Rcθ×θRc.

Consequently,

EβhRc=(ξ1u12+2ξ2u1(1-u1)+(1-u1)2)βh(μp+u3)(μp+u3)βh(ξ1u12+2ξ2u1(1-u1)+(1-u1)2)+αβp,EβpRc=αβp(μp+u3)βh(ξ1u12+2ξ2u1(1-u1)+(1-u1)2)+αβp,EαRc=αβp(μp+u3)βh(ξ1u12+2ξ2u1(1-u1)+(1-u1)2)+αβp,EγRc=μhγ+μh,Eγ0Rc=-(1-u2)γ0γ1u2+(1-u2)γ0+μh,EμhRc=((1-u2)γ0+γ1u2+γ+2μh)μh((u2-1)γ0-γ1u2-μh)(γ+μh),EμpRc=-μpαβp((μp+u3)(1+(ξ1-2ξ2+1)u12+(2ξ2-2)u1)βh+αβp)(μp+u3),Eu1Rc=βh(2ξ1u1+2ξ2(1-u1)-2ξ2u1-2+2u1)u1(μp+u3)(μp+u3)βh(ξ1u12+2ξ2u1(1-u1)+(1-u1)2)+αβp,Eu2Rc=-(γ1-γ0)u2γ1u2+(1-u2)γ0+μh,Eu3Rc=-u3αβp((μp+u3)(1+(ξ1-2ξ2+1)u12+(2ξ2-2)u1)βh+αβp)(μp+u3).

Upon substituting the parameter values provided in Table 1, u1 = 0.4, u2 = 0.4, u3 = 0.4 for Rc = 0.6570877605 < 1, and u1 = 0.1, u2 = 0.1, u3 = 0.1 for Rc = 1.503519079 > 1. In addition, the assumed values for γ1, ξ1, and ξ2 are 13, 0.5, and 0.3, respectively. The corresponding elasticity sensitivity indices are presented in Table 2. Furthermore, Figure 3 displays a graphical representation in the form of a bar plot. In the scenario where Rc < 1, signifying the absence of an outbreak in the disease spread, Figure 3A illustrates that the most sensitive parameter influencing Rc is βh, the transmission rate between humans. This underscores the pivotal role of human-to-human transmission in preventing an outbreak. On the contrary, when Rc > 1, indicating the presence of an outbreak, Figure 3B reveals that the sensitivity of Rc is primarily influenced by two key parameters: γ0, the recovery rate of the infected individuals, and βh, the human-to-human transmission rate. This dual sensitivity emphasizes the importance of recovery dynamics and the rate of transmission between humans in understanding and mitigating the outbreak dynamics of monkeypox.

Table 2
www.frontiersin.org

Table 2. Sensitivity index of the parameters in Rc using parameter values in Table 1, except u1, u2, and u3.

Figure 3
www.frontiersin.org

Figure 3. Sensitivity analysis of Rc. (A) Rc < 1. (B) Rc > 1.

We conclude this sensitivity analysis section by presenting a two-parameter sensitivity analysis regarding the control parameters u1, u2, and u3. The results can be observed in Figure 4. In Figure 4A, we depict how the control reproduction number, Rc, varies with the self-protection (u1) and hospitalization (u2) control parameters. It is evident that increasing the values of u1 and u2 can significantly decrease the magnitude of Rc. The sensitivity of both u1 and u2 appears to be equally dominant in reducing Rc. The same observation holds for Figures 4B, C, where increasing the values of parameters u1, u2, and u3 can reduce the value of Rc. However, it is evident from this numerical experiment that the parameter u3 is not as significant in suppressing the magnitude of Rc, as indicated in Figures 4B, C. This finding aligns with the results of the elasticity index analysis shown in Figure 3.

Figure 4
www.frontiersin.org

Figure 4. Sensitivity analysis of Rc with respect to (A) u1 and u2, (B) u2 and u3, and (C) u1 and u3.

5.2 Optimal control simulations

In this section, we investigate the implications of several of optimal control strategies for reducing the prevalence of monkeypox within the human population. Our focus is specifically directed toward assessing the potential of combined control measures to stop the disease's transmission. We aim to provide a demonstrative analysis through numerical simulations of the monkeypox model, both with and without the incorporation of optimized control interventions, to examine the influence of the control variables introduced in the preceding section.

Investigations are conducted into the effects of various combinations of optimal control strategies represented by u1, u2, and u3. These strategies are systematically categorized into three distinct scenarios: single control, double controls, and triple controls, facilitating a coherent grouping of the seven potential control strategies that have been simulated in this study. To clarify, these strategies are defined as follows: Strategy 1 uses only u1, Strategy 2 uses only u2, Strategy 3 uses only u3, Strategy 4 uses both u1 and u2, Strategy 5 uses both of u1 and u3, Strategy 6 uses both of u2 and u3, and Strategy 7 uses of all three controls u1, u2, and u3.

In the following, a comprehensive simulation along with an analysis and discussion of these seven interventions are presented.

1. Single intervention

The influence of employing Strategy 1 (self-protection control, u1 only) on the transmission of monkeypox within the population is illustrated in Figure 5. The implementation of this optimal control results in a 58.7% reduction of the infected human population, as evident in Figure 5C after 50 days. Conversely, the susceptible population exhibits a 46.2% increase, as depicted in Figure 5A. The control profiles that correspond to Strategy 1 are illustrated in Figure 5E. The self-protection control profile shows a constant level of 79% until day 45, at which point it gradually declines monotonically until it reaches its lowest threshold. This pattern suggests an early phase marked by a strong self-defense mechanism deployment, which may be suggestive of a critical phase during the management of the disease. Over time, there is a deliberate decrease in the intensity of these measures.

Figure 7 (as in Appendix A) illustrates the impact of implementing Strategy 2 (use of hospitalization intervention, u2 only) on the dissemination and transmission of monkeypox. The susceptible population experiences a rapid increase with the application of control compared to the scenario without control, as depicted in Appendix Figure 7A, with a 34.6% rise observed after 50 days. When the strategy is used to its fullest potential, as illustrated in Appendix Figure 7C, there is a notable reduction in the population of infected humans, amounting to 51.6%. Appendix Figure 7E analyzes the control profiles of each control variable in strategy 2. The control profile for hospitalization reveals a distinctive curve, commencing at 60%, gradually decreasing until day 10. Following this decline, the curve maintains near-constant levels at 30% until day 30, indicating a stable phase in hospitalization measures. Subsequently, there is a further monotonic decrease reaching zero, indicating a strategic reduction in hospitalization efforts after an initial period of intensity.

Figure 8 (as in Appendix A) displays time series plots showing the effect of applying Strategy 3 (disinfectant control, u3 alone). Using a singular application of disinfectant control results in reduced exposure of susceptible individuals by 48.9% at day 50, as illustrated in Appendix Figure 8B. Consequently, this leads to a decline in the overall number of infected individuals within the population, as demonstrated in Appendix Figure 8C, with a reduction of 46.9%. It is noted that the decrease in contaminated surfaces concentration occurs more rapidly compared to preceding strategies, as represented in Appendix Figure 8D. Following Appendix Figure 8E, we present the control profiles associated with this strategy. A distinct pattern can be seen in the control profile for disinfectant application on contaminated surfaces, which first increases in a curve before peaking at 90%. The curve then starts to progressively decline until it reaches its lowest value. This profile proposes a strategic and efficient use of disinfectant, stressing a thorough application at first and a gradual, deliberate reduction in intensity over time.

2. Combination of two interventions

The impact of employing Strategy 4 (combination of self-protection and hospitalization controls) on the dynamics of monkeypox within the population is depicted in Appendix Figure 9. A greater number of susceptible individuals are shielded during the intervention period, and the size of the infected population notably decreased after a few days of employing this strategy, as shown in Appendix Figures 9A, C, respectively. Specifically, there is a 59.6% increase in susceptible individuals and a remarkable 79% decrease in the infected population.

As in Appendix A, Figure 10, we showcase the simulation of Strategy 5 involving the combination of the use of self-protection and disinfectant controls on the population. There is a remarkable reduction in both the infected human population and contaminated surfaces when the combined interventions are applied, in comparison to the scenario without control, as illustrated in Appendix Figures 10B, D. This results in an 85.5% reduction in the infected population and a significant 93.1% reduction in contaminated surfaces.

Figure 11 (see Appendix A) displays the influence of Strategy 6, optimizing the use of hospitalization and disinfectant controls on the dynamics of monkeypox transmission. The impact of the combined interventions on contaminated surfaces mirrors the findings in Strategy 5 (see Appendix Figures 10D, 11D). It is noteworthy, though, that in this particular intervention, the reduction in the number of infected individuals occurs more rapidly, as Appendix Figure 11C illustrates.

3. Combination of three interventions

The effect of Strategy 7 (combination of self-protection, hospitalization, and disinfectant) on the dynamic behavior of monkeypox is presented in Figure 12 in Appendix A, where Appendix Figure 12E specifically shows the plots displaying the control profiles for all optimal control variables in this strategy. It is observed that the susceptible and exposed individuals, in addition to contaminated surfaces, are in the best results in all strategies as revealed in Appendix Figures 12A, B, D. On the contrary, this is not the case for the number of infected individuals (see Appendix Figure 12C).

Figure 5
www.frontiersin.org

Figure 5. Simulations applying Strategy 1 to the optimal control model. (A) Susceptible individuals, (B) exposed individuals, (C) infected individuals, (D) contaminated surfaces, and (E) control profile.

5.3 Cost-effectiveness analysis

The objective here is to evaluate the most cost-effective strategy among the mentioned monkeypox control techniques in a setting with limited resources. To achieve this, we employ metrics such as the Infection Averted Ratio (IAR) and Average Cost-Effectiveness Ratio (ACER) [50]. Applying the mathematical definitions of these economic assessments as outlined in Asamoah et al. [50], the outcomes of these calculations are presented in Table 3.

Table 3
www.frontiersin.org

Table 3. IAR and ACER for all strategies.

When employing the Infection Averted Ratio, the strategy exhibiting the largest IAR value is considered the most effective one [51]. Figure 6A suggests that the strategy yielding the highest cost-effectiveness is Strategy 7, which involves the combined use of self-protection, hospitalization, and disinfectant.

Figure 6
www.frontiersin.org

Figure 6. (A) IAR plots for all strategies. (B) ACER plots for all strategies except Strategy 1, as its value exceeds the range of the figure's index.

On the other hand, the strategy exhibiting the lowest ACER ratio is deemed the most cost-effective, in accordance with the Average Cost-Effectiveness Ratio methodology [51]. Strategy 2, as can be seen in Figure 6B, emerges as the most cost-effective strategy, involving the use of hospitalization only.

6 Conclusion

A mathematical model for monkeypox transmission considering contaminated surfaces in the transmission process is considered in this article. The model was developed based on an SEIRC model. Given the comprehensive analysis presented in this study on the transmission dynamics of monkeypox, incorporating contaminated surfaces into a deterministic mathematical model, several key findings emerge. The calculated basic reproduction number serves as a critical threshold, with a forward bifurcation occurring when it equals unity, signifying a pivotal point for disease spread. The absence of a backward bifurcation underscores the singular significance of the basic reproduction number as the sole indicator of the model's endemicity.

The stability analysis, focusing on disease-free and endemic equilibria, establishes global stability conditions based on the basic reproduction number. Notably, the study emphasizes the global stability of the monkeypox-free equilibrium when the reproduction number is below one, contrasting with the instability observed when it exceeds this threshold. The existence and global stability of the endemic equilibrium become evident when the reproduction number surpasses one.

To deepen insights into control strategies, sensitivity analysis identifies influential parameters. This leads to the formulation of an optimal control problem, encompassing self-protection, hospitalization, and disinfection as key interventions. Existence of the solution of the optimal control problem is shown analytically. The Pontryagin Maximum Principle is used to characterize the optimal control problem, giving us the set of adjoint equations and showing optimality conditions.

Numerical simulations explore various control scenarios, demonstrating the efficacy of implemented measures in curbing monkeypox spread. Cost-effectiveness analyses, employing IAR and ACER criteria, underscore the superiority of combined interventions, particularly the synergistic effect of self-protection and disinfection. The IAR results indicate that Strategy 7 is the optimal approach, whereas the ACER metric favors Strategy 2.

The findings of this research underscore the critical role of contaminated surfaces in disease transmission, advocating for targeted control measures. These results align with the conclusions reached by previous studies [5, 23], further supporting the existing body of research highlighting the significant contribution of surface contamination to the spread of infectious diseases. The suggested optimal control strategies, validated through rigorous analysis and simulations, provide valuable insights for evidence-based approaches in mitigating the impact of monkeypox. Additionally, future research considerations, such as incorporating human awareness and assessing the impact of animals as potential vectors, offer avenues for further improvement and a more comprehensive understanding of the dynamics involved in combatting the spread of monkeypox.

Data availability statement

The original contributions presented in the study are included in the article/Supplementary material, further inquiries can be directed to the corresponding author.

Author contributions

AH: Data curation, Formal analysis, Investigation, Methodology, Software, Visualization, Writing – original draft, Writing – review & editing. DA: Conceptualization, Funding acquisition, Investigation, Methodology, Project administration, Supervision, Writing – original draft, Writing – review & editing. MN: Data curation, Investigation, Methodology, Resources, Supervision, Validation, Visualization, Writing – review & editing.

Funding

The author(s) declare financial support was received for the research, authorship, and/or publication of this article. This research was funded by the Ministry of Research, Technology and Higher Education Indonesia (ID number: NKB-1132/UN2.RST/HKP.05.00/2023).

Acknowledgments

The authors thank all reviewers for their constructive suggestions.

Conflict of interest

The authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

Publisher's note

All claims expressed in this article are solely those of the authors and do not necessarily represent those of their affiliated organizations, or those of the publisher, the editors and the reviewers. Any product that may be evaluated in this article, or claim that may be made by its manufacturer, is not guaranteed or endorsed by the publisher.

Supplementary material

The Supplementary Material for this article can be found online at: https://www.frontiersin.org/articles/10.3389/fams.2024.1372579/full#supplementary-material

References

1. WHO Recommends New Name for Monkeypox Disease. Available online at: https://www.who.int/news/item/28-11-2022-who-recommends-new-name-for-monkeypox-disease (accessed July 7, 2023).

Google Scholar

2. Gessain A, Nakoune E, Yazdanpanah Y. Monkeypox. N Engl J Med. (2022) 387:1783–93. doi: 10.1056/NEJMra2208860

Crossref Full Text | Google Scholar

3. Beeson A, Styczynski A, Hutson CL, Whitehill F, Angelo KM, Minhaj FS, et al. Mpox respiratory transmission: the state of the evidence. Lancet Microbe. (2023) 4:e277–83. doi: 10.1016/S2666-5247(23)00034-4

PubMed Abstract | Crossref Full Text | Google Scholar

4. Boone SA, Gerba CP. Significance of fomites in the spread of respiratory and enteric viral disease. Appl Environ Microbiol. (2007) 73:1687. doi: 10.1128/AEM.02051-06

PubMed Abstract | Crossref Full Text | Google Scholar

5. Addai E, Ngungu M, Omoloye MA, Marinda E. Modelling the impact of vaccination and environmental transmission on the dynamics of monkeypox virus under Caputo operator. Math. Biosci. Eng. (2023) 20:10174–99. doi: 10.3934/mbe.2023446

PubMed Abstract | Crossref Full Text | Google Scholar

6. Wilson ME, Hughes JM, McCollum AM, Damon IK. Human monkeypox. Clin Infect Dis. (2014) 58:260–7. doi: 10.1093/cid/cit703

Crossref Full Text | Google Scholar

7. Mpox Outbreak Global Map | Mpox | Poxvirus | CDC. Available online at: https://www.cdc.gov/poxvirus/mpox/response/2022/world-map.html (accessed July 7, 2023)

Google Scholar

8. Nuzzo JB, Borio LL, Gostin LO. The WHO declaration of monkeypox as a global public health emergency. JAMA. (2022) 328:615–7. doi: 10.1001/jama.2022.12513

PubMed Abstract | Crossref Full Text | Google Scholar

9. Lai CC, Hsu CK, Yen MY, Lee PI, Ko WC, Hsueh PR. Monkeypox: an emerging global threat during the COVID-19 pandemic. J. Microbiol. Immunol. Infect. (2022) 55:787–94. doi: 10.1016/j.jmii.2022.07.004

PubMed Abstract | Crossref Full Text | Google Scholar

10. Lahariya C, Thakur A, Dudeja N. Monkeypox disease outbreak (2022): epidemiology, challenges, and the way forward. Indian Pediatr. (2022) 59:636–42. doi: 10.1007/s13312-022-2578-2

PubMed Abstract | Crossref Full Text | Google Scholar

11. Okonji OC, Okonji EF. Monkeypox during COVID-19 era in Africa: current challenges and recommendations. Ann Med Surg. (2022) 81:104381. doi: 10.1016/j.amsu.2022.104381

PubMed Abstract | Crossref Full Text | Google Scholar

12. Kraemer MUG, Tegally H, Pigott DM, Dasgupta A, Sheldon J, Wilkinson E, et al. Tracking the 2022 monkeypox outbreak with epidemiological data in real-time. Lancet Infect Dis. (2022) 22:941–2. doi: 10.1016/S1473-3099(22)00359-0

PubMed Abstract | Crossref Full Text | Google Scholar

13. Nakoune E, Olliaro P. Waking up to monkeypox. BMJ. (2022) 377:o1321. doi: 10.1136/bmj.o1321

Crossref Full Text | Google Scholar

14. Quarleri J, Delpino MV, Galvan V. Monkeypox: considerations for the understanding and containment of the current outbreak in non-endemic countries. GeroScience. (2022) 44:2095–103. doi: 10.1007/s11357-022-00611-6

PubMed Abstract | Crossref Full Text | Google Scholar

15. Reynolds MG, Carroll DS, Karem KL. Factors affecting the likelihood of monkeypox's emergence and spread in the post-smallpox era. Curr Opin Virol. (2012) 2:335–43. doi: 10.1016/j.coviro.2012.02.004

PubMed Abstract | Crossref Full Text | Google Scholar

16. Mahase E. Monkeypox: what do we know about the outbreaks in Europe and North America? BMJ. (2022) 377:o1274. doi: 10.1136/bmj.o1274

PubMed Abstract | Crossref Full Text | Google Scholar

17. Al-Shomrani MM, Musa SS, Yusuf A. Unfolding the transmission dynamics of monkeypox virus: an epidemiological modelling analysis. Mathematics. (2023) 11:1121. doi: 10.3390/math11051121

Crossref Full Text | Google Scholar

18. Okyere S, Ackora-Prah J. Modeling and analysis of monkeypox disease using fractional derivatives. Results Eng. (2023) 17:100786. doi: 10.1016/j.rineng.2022.100786

PubMed Abstract | Crossref Full Text | Google Scholar

19. Peter OJ, Kumar S, Kumari N, Oguntolu FA, Oshinubi K, Musa R. Transmission dynamics of Monkeypox virus: a mathematical modelling approach. Model Earth Syst Environ. (2022) 8:3423–34. doi: 10.1007/s40808-021-01313-2

PubMed Abstract | Crossref Full Text | Google Scholar

20. Alharbi R, Jan R, Alyobi S, Altayeb Y, Khan Z. Mathematical modeling and stability analysis of the dynamics of monkeypox via fractional-calculus. Fractals. (2022) 30:2240266. doi: 10.1142/S0218348X22402666

Crossref Full Text | Google Scholar

21. Michael UE, Omenyi LO, Kafayat E, Nwaeze E, Akachukwu OA, Ozoigbo G, et al. Monkeypox mathematical model with surveillance as control. Commun Math Biol Neurosci. (2023) 2023:6. doi: 10.28919/cmbn/7781

Crossref Full Text | Google Scholar

22. El-Mesady A, Elsonbaty A, Adel W. On nonlinear dynamics of a fractional order monkeypox virus model. Chaos Solitons Fractals. (2022) 164:112716. doi: 10.1016/j.chaos.2022.112716

PubMed Abstract | Crossref Full Text | Google Scholar

23. Alshehri A, Ullah S. Optimal control analysis of Monkeypox disease with the impact of environmental transmission. AIMS Math. (2023) 8:16926–60. doi: 10.3934/math.2023865

Crossref Full Text | Google Scholar

24. Alzubaidi AM, Othman HA, Ullah S, Ahmad N, Alam MM. Analysis of Monkeypox viral infection with human to animal transmission via a fractional and Fractal-fractional operators with power law kernel. Math Biosci Eng. (2023) 20:6666–90. doi: 10.3934/mbe.2023287

PubMed Abstract | Crossref Full Text | Google Scholar

25. Li C, Samreen, Ullah S, Nawaz R, AlQahtani SA, Li S. Mathematical modeling and analysis of monkeypox 2022 outbreak with the environment effects using a Cpauto fractional derivative. Phys Scr. (2023) 98:105239. doi: 10.1088/1402-4896/acf88e

Crossref Full Text | Google Scholar

26. Madubueze CE, Onwubuya IO, Nkem GN, Chazuka Z. The transmission dynamics of the monkeypox virus in the presence of environmental transmission. Front Appl Math Stat. (2022) 8:1061546. doi: 10.3389/fams.2022.1061546

Crossref Full Text | Google Scholar

27. Mpox (monkeypox). Available online at: https://www.who.int/news-room/fact-sheets/detail/monkeypox (accessed July 23, 2023).

Google Scholar

28. Morgan CN, Whitehill F, Doty JB, Schulte J, Matheny A, Stringer J, et al. Environmental persistence of monkeypox virus on surfaces in household of person with travel-associated infection, Dallas, Texas, USA, 2021. Emerg Infect Dis. (2022) 28:1982. doi: 10.3201/eid2810.221047

PubMed Abstract | Crossref Full Text | Google Scholar

29. Brauer F, Castillo-Chavez C, Feng Z. Endemic disease models. Math Models Epidemiol. (2019) 69:63. doi: 10.1007/978-1-4939-9828-9_3

Crossref Full Text | Google Scholar

30. Aldila D, Ndii MZ, Anggriani N, Windarto, Tasman H, Handari BD. Impact of social awareness, case detection, and hospital capacity on dengue eradication in Jakarta: a mathematical model approach. Alex Eng J. (2023) 64:691–707. doi: 10.1016/j.aej.2022.11.032

Crossref Full Text | Google Scholar

31. Diekmann O, Heesterbeek JAP, Roberts MG. The construction of next-generation matrices for compartmental epidemic models. J R Soc Interface. (2010) 7:873–85. doi: 10.1098/rsif.2009.0386

PubMed Abstract | Crossref Full Text | Google Scholar

32. Driessche PVD, Watmough J. Reproduction numbers and sub-threshold endemic equilibria for compartmental models of disease transmission. Math Biosci. (2002) 180:29–48. doi: 10.1016/S0025-5564(02)00108-6

PubMed Abstract | Crossref Full Text | Google Scholar

33. Bavafa-Toosi Y. Stability analysis. Introduction Linear Control Syst. (2019) 2019:265–31. doi: 10.1016/B978-0-12-812748-3.00003-3

Crossref Full Text | Google Scholar

34. Alqahtani H, Badshah Q, Sakhi S, ur Rahman G, Gmez-Aguilar JF. Qualitative aspects and sensitivity analysis of MERS-Corona epidemic model with and without noise. Phys Scr. (2023) 98:125018. doi: 10.1088/1402-4896/ad0bb6

Crossref Full Text | Google Scholar

35. Martcheva M. An Introduction to Mathematical Epidemiology, Volume 61. Berlin: Springer US. (2015). doi: 10.1007/978-1-4899-7612-3

Crossref Full Text | Google Scholar

36. Castillo-Chavez C, Song B. Dynamical models of tuberculosis and their applications. Math Biosci Eng. (2004) 1:361–404. doi: 10.3934/mbe.2004.1.361

PubMed Abstract | Crossref Full Text | Google Scholar

37. How to Protect Yourself | Mpox | Poxvirus | CDC. Available online at: https://www.cdc.gov/poxvirus/mpox/prevention/protect-yourself.html (accessed July 23, 2023).

Google Scholar

38. Mpox (monkeypox): How it Spreads, Prevention and Risks - Canada.ca. Available online at: https://www.canada.ca/en/public-health/services/diseases/mpox/risks.htmla3 (accessed July 23, 2023).

Google Scholar

39. Miller MJ, Cash-Goldwasser S, Marx GE, Schrodt CA, Kimball A, Padgett K, et al. Severe monkeypox in hospitalized patients — United States, August 10–October 10, 2022. MMWR Morb Mortal Wkly Rep. (2022) 71:1412–7. doi: 10.15585/MMWR.MM7144E1

PubMed Abstract | Crossref Full Text | Google Scholar

40. Benites-Zapata VA, Ulloque-Badaracco JR, Alarcon-Braga EA, Hernandez-Bustamante EA, Mosquera-Rojas MD, Bonilla-Aldana DK, et al. Clinical features, hospitalisation and deaths associated with monkeypox: a systematic review and meta-analysis. Ann Clin Microbiol Antimicrob. (2022) 21:1–18. doi: 10.1186/s12941-022-00527-1

PubMed Abstract | Crossref Full Text | Google Scholar

41. DeWitt ME, Polk C, Williamson J, Shetty AK, Passaretti CL, McNeil CJ, et al. Global monkeypox case hospitalisation rates: a rapid systematic review and meta-analysis. eClinicalMedicine. (2022) 54:101710. doi: 10.1016/j.eclinm.2022.101710

PubMed Abstract | Crossref Full Text | Google Scholar

42. Kampf G, Todt D, Pfaender S, Steinmann E. Persistence of coronaviruses on inanimate surfaces and their inactivation with biocidal agents. J Hosp Infect. (2020) 104:246–51. doi: 10.1016/j.jhin.2020.01.022

PubMed Abstract | Crossref Full Text | Google Scholar

43. About List N: Disinfectants for Coronavirus (COVID-19) | US EPA. Available online at: https://www.epa.gov/coronavirus/about-list-n-disinfectants-coronavirus-covid-19-0 (accessed August 5, 2023).

Google Scholar

44. Xu D, Xu X, Xie Y, Yang C. Optimal control of an SIVRS epidemic spreading model with virus variation based on complex networks. Commun Nonlinear Sci Numer Simul. (2017) 48:200–10. doi: 10.1016/j.cnsns.2016.12.025

Crossref Full Text | Google Scholar

45. Fleming WH, Rishel RW. Deterministic and Stochastic Optimal Control. Google Books. Available online at: https://scholar.google.com/scholar_lookup?title=Deterministic%20and%20stochastic%20optimal%20control&publication_year=1975&author=W.H.%20Fleming&author=R.W.%20Rishel (Accessed August 9, 2023).

Google Scholar

46. Birkhoff G, Rota G-C. Ordinary Differential Equation, 4th ed. New York, NY: Wiley (1989). References - Scientific Research Publishing. Available online at: https://www.scirp.org/ (accessed August 9, 2023).

Google Scholar

47. Fister KR, Lenhart S, McNally JS. Optimizing chemotherapy in an HIV model. Electron J Differ Equ. (1998) 1998:1–12.

Google Scholar

48. Pontryagin LS. Mathematical Theory of Optimal Processes. London: Routledge (2018) doi: 10.1201/9780203749319

Crossref Full Text | Google Scholar

49. Wu J, Dhingra R, Gambhir M, Remais JV. Sensitivity analysis of infectious disease models: methods, advances and their application. J R Soc Interface. (2013) 10:20121018. doi: 10.1098/rsif.2012.1018

PubMed Abstract | Crossref Full Text | Google Scholar

50. Asamoah JKK, Yankson E, Okyere E, Sun GQ, Jin Z, Jan R, et al. Optimal control and cost-effectiveness analysis for dengue fever model with asymptomatic and partial immune individuals. Results Phys. (2021) 31:104919. doi: 10.1016/j.rinp.2021.104919

Crossref Full Text | Google Scholar

51. Oke SI, Matadi MB, Xulu SS. Cost-effectiveness analysis of optimal control strategies for breast cancer treatment with ketogenic-diet. Far East J Math Sci. (2018) 109:303–42. doi: 10.17654/MS109020303

Crossref Full Text | Google Scholar

Keywords: monkeypox, mathematical model, stability analysis, optimal control, contaminated surfaces

Citation: Hassan AH, Aldila D and Noor Aziz MH (2024) Optimal control and stability analysis of monkeypox transmission dynamics with the impact of contaminated surfaces. Front. Appl. Math. Stat. 10:1372579. doi: 10.3389/fams.2024.1372579

Received: 18 January 2024; Accepted: 06 February 2024;
Published: 01 March 2024.

Edited by:

Nossaiba Baba, University of Hassan II Casablanca, Morocco

Reviewed by:

Asmaa Idmbarek, University of Hassan II Casablanca, Morocco
Imane Agmour, University of Hassan II Casablanca, Morocco

Copyright © 2024 Hassan, Aldila and Noor Aziz. This is an open-access article distributed under the terms of the Creative Commons Attribution License (CC BY). The use, distribution or reproduction in other forums is permitted, provided the original author(s) and the copyright owner(s) are credited and that the original publication in this journal is cited, in accordance with accepted academic practice. No use, distribution or reproduction is permitted which does not comply with these terms.

*Correspondence: Dipo Aldila, aldiladipo@sci.ui.ac.id

Disclaimer: All claims expressed in this article are solely those of the authors and do not necessarily represent those of their affiliated organizations, or those of the publisher, the editors and the reviewers. Any product that may be evaluated in this article or claim that may be made by its manufacturer is not guaranteed or endorsed by the publisher.