
Marine Predators Algorithm (MPA) is a newly nature-inspired meta-heuristic algorithm, which is proposed based on the Lévy flight and Brownian motion of ocean predators. Since the MPA was proposed, it has been successfully applied in many fields. However, it includes several shortcomings, such as falling into local optimum easily and precocious convergence. To balance the exploitation and exploration ability of MPA, a modified marine predators algorithm hybridized with teaching-learning mechanism is proposed in this paper, namely MTLMPA. Compared with MPA, the proposed MTLMPA has two highlights. Firstly, a kind of teaching mechanism is introduced in the first phase of MPA to improve the global searching ability. Secondly, a novel learning mechanism is introduced in the third phase of MPA to enhance the chance encounter rate between predator and prey and to avoid premature convergence. MTLMPA is verified by 23 benchmark numerical testing functions and 29 CEC-2017 testing functions. Experimental results reveal that the MTLMPA is more competitive compared with several state-of-the-art heuristic optimization algorithms.
Citation: Yunpeng Ma, Chang Chang, Zehua Lin, Xinxin Zhang, Jiancai Song, Lei Chen. Modified Marine Predators Algorithm hybridized with teaching-learning mechanism for solving optimization problems[J]. Mathematical Biosciences and Engineering, 2023, 20(1): 93-127. doi: 10.3934/mbe.2023006
[1] | Muntaser Safan . Mathematical analysis of an SIR respiratory infection model with sex and gender disparity: special reference to influenza A. Mathematical Biosciences and Engineering, 2019, 16(4): 2613-2649. doi: 10.3934/mbe.2019131 |
[2] | Guodong Li, Wenjie Li, Ying Zhang, Yajuan Guan . Sliding dynamics and bifurcations of a human influenza system under logistic source and broken line control strategy. Mathematical Biosciences and Engineering, 2023, 20(4): 6800-6837. doi: 10.3934/mbe.2023293 |
[3] | Jacek Banasiak, Eddy Kimba Phongi, MirosŁaw Lachowicz . A singularly perturbed SIS model with age structure. Mathematical Biosciences and Engineering, 2013, 10(3): 499-521. doi: 10.3934/mbe.2013.10.499 |
[4] | Muntaser Safan, Klaus Dietz . On the eradicability of infections with partially protective vaccination in models with backward bifurcation. Mathematical Biosciences and Engineering, 2009, 6(2): 395-407. doi: 10.3934/mbe.2009.6.395 |
[5] | Andrea Pugliese, Abba B. Gumel, Fabio A. Milner, Jorge X. Velasco-Hernandez . Sex-biased prevalence in infections with heterosexual, direct, and vector-mediated transmission: a theoretical analysis. Mathematical Biosciences and Engineering, 2018, 15(1): 125-140. doi: 10.3934/mbe.2018005 |
[6] | Abdessamad Tridane, Yang Kuang . Modeling the interaction of cytotoxic T lymphocytes and influenza virus infected epithelial cells. Mathematical Biosciences and Engineering, 2010, 7(1): 171-185. doi: 10.3934/mbe.2010.7.171 |
[7] | Yicang Zhou, Zhien Ma . Global stability of a class of discrete age-structured SIS models with immigration. Mathematical Biosciences and Engineering, 2009, 6(2): 409-425. doi: 10.3934/mbe.2009.6.409 |
[8] | Gigi Thomas, Edward M. Lungu . A two-sex model for the influence of heavy alcohol consumption on the spread of HIV/AIDS. Mathematical Biosciences and Engineering, 2010, 7(4): 871-904. doi: 10.3934/mbe.2010.7.871 |
[9] | Shuyang Xue, Meili Li, Junling Ma, Jia Li . Sex-structured wild and sterile mosquito population models with different release strategies. Mathematical Biosciences and Engineering, 2019, 16(3): 1313-1333. doi: 10.3934/mbe.2019064 |
[10] | Baojun Song, Wen Du, Jie Lou . Different types of backward bifurcations due to density-dependent treatments. Mathematical Biosciences and Engineering, 2013, 10(5&6): 1651-1668. doi: 10.3934/mbe.2013.10.1651 |
Marine Predators Algorithm (MPA) is a newly nature-inspired meta-heuristic algorithm, which is proposed based on the Lévy flight and Brownian motion of ocean predators. Since the MPA was proposed, it has been successfully applied in many fields. However, it includes several shortcomings, such as falling into local optimum easily and precocious convergence. To balance the exploitation and exploration ability of MPA, a modified marine predators algorithm hybridized with teaching-learning mechanism is proposed in this paper, namely MTLMPA. Compared with MPA, the proposed MTLMPA has two highlights. Firstly, a kind of teaching mechanism is introduced in the first phase of MPA to improve the global searching ability. Secondly, a novel learning mechanism is introduced in the third phase of MPA to enhance the chance encounter rate between predator and prey and to avoid premature convergence. MTLMPA is verified by 23 benchmark numerical testing functions and 29 CEC-2017 testing functions. Experimental results reveal that the MTLMPA is more competitive compared with several state-of-the-art heuristic optimization algorithms.
Influenza A is a highly contagious, respiratory, infectious, viral disease and is potentially lethal. The virus is transmitted from human to human through direct contacts (mainly by coughing, sneezing, or close contact). The genetic materials of influenza viruses are composed of single-stranded RNA, and frequent mistakes happen while copying themselves. Due to their high mutation rates, influenza viruses evolve rapidly. Some of their new generations quickly adapt with the new conditions, which helps them succeed in causing new epidemics and (sometimes) pandemics. Therefore, influenza is both seasonal and pandemic. Seasonal influenza infects more than one billion people annually [1]. Recently, it was estimated that seasonal influenza is associated with 294,000 to 518,000 annual respiratory deaths [2]; according to the World Health Organization (WHO) [1], the estimation is even larger. Moreover, Chaves et al. [3] found that influenza is associated with an increased risk of ischemic heart disease (IHD) mortality. The authors attributed the death of about 300,000 IHD adults of ages over 50 years old (every year and globally) to influenza. In their study, they reported a 4% reduction in the global IHD deaths if no influenza was present [3]. Therefore, influenza A viruses do significantly impact human health and, consequently, the global economy [4].
It is evident that sex is a risk factor to influenza incidences and outcomes, as both males and females differ in their respones to the infection [5]. This disparity may reflect genetic and hormonal differences between the two sexes [5]. Based on the repeated influenza outbreaks and pandemics, the morbidity and mortality of females are significantly different from those of males [6]. The extent to which they are different is associated with other risk factors, including age and the chronic medical conditions from which they suffer [5,6]. However, for the purpose of this work, we will focus on mathematically studying the role of sex and gender risk factors on the transmission dynamics of influenza.
Mathematical models have been extensively used to study the dynamics of influenza A infection at both the population level [7,8,9,10,11] and at the cellular level [12,13]. This study is designed to help predict the spread of influenza A infection and to estimate its burden rather than studying the dynamics of cellular interaction and disease progression; therefore, we focus on models that treat the problem at the population level. In this respect, the literature shows the publication of various works that focus on exploring the dynamics in both the absence and in the presence of influenza interventions. For example, Casagrandi et al. [7] extended the classical SIR model by including a fourth compartment that represented an intermediate state between the fully susceptible state (S) and the fully protected state (R). This intermediate state is called cross-immune and is denoted by C. Therefore, an SIRC model was developed and mathematically analyzed. This SIRC model describes the dynamics of influenza A infection in a demographically stationary closed population in the absence of influenza-induced mortality, with the fact that exposing cross-immune individuals to a different strain (other than the running one) of influenza either boosts their immunity or reinfects them. The model assumes that the natural immunity acquired by experiencing the infection decays with time and, in consequence, R-compartment individuals move to the C-compartment. Additionally, the model assumes that a further decay of the immunity of C-compartment individuals moves them to the fully susceptible state. Also, motivated by spanish flu data in Australia, Samsuzzoha et al. [14] employed two deterministic models (precisely, a Susceptible-Infected-Recovered (SEIRS) and a Susceptible-Vaccinated-Exposed-Infected-Recovered-Susceptible (SVEIRS)) to capture the main characteristic of influenza transmission. The authors fitted the models to the data and estimated the parameters involved in both systems.
Other mathematical models were employed to include quarantine as a preventive intervention to control influenza. Vivas-Barber et al. [10] focused on exploring the role of asymptomatic (mild) infections on the long-term transmission dynamics of influenza, in the presence of a fully protective quarantine-based intervention, by introducing a compartment (A) to denote asymptomatically infected individuals to an extended SIQR model (here, Q denotes quarantine individuals). Therefore, an SAIQR model was obtained. The model assumes that the quarantine intervention is completely perfect, while the asymptomatic individuals are capable of transmitting (although less infectious than I-individuals) influenza to the totally susceptible individuals (S). Moreover, the model considers an open population, but ignores the infection lethality. The authors performed a standard analysis of the model and numerically showed (with parameter values typical to influenza) the existence of damped oscillations that described recurring epidemics. Moreover, Erdem et al. [11] extended the SIQR model (with perfect quarantine) to consider the fact that no quarantine program is completely perfect. Therefore, a so-called "σ-quarantine" SIQR model was introduced and mathematically analyzed, where σ∈[0,1] denotes the effectiveness of a quarantine-based intervention program. Other models that were used to describe the dynamics of influenza A infection include (but not limited to) those published by Nuño et al. [15], Alexander et al. [16], and Krishnapriya et al. [17]. However, the role of sex and gender disparities on the influenza dynamics is less explored mathematically.
The classical SIR model has been extended in a previous work to include sex and gender disparities to describe the dynamics of a single outbreak of influenza [4]. The model was mathematically analyzed and was employed to assess the impact of these disparities on the influenza A disease outcomes (i.e., the basic reproduction number and the prevalence of the infection at the endemic situation). The author further extended the model to include the impact of applying an imperfect vaccine-based intervention strategy to contain/eliminate influenza A infection. Basically, the author considered a demographically stationary closed population, with the assumption that both male and female sub-populations were equally recruited solely by births. Moreover, the author neglected the possible repeated exposure to influenza A infection as well as its lethality. However, in this work, we consider various biologically meaningful extensions. First, we extend the classical Susceptible-Infected-Susceptible (SIS) model to include the sex and gender disparities in an open population, where the individuals are recruited by births as well as immigrants, where the immigration rate is assumed to depend on the total population size and some kind of carrying capacities. Additionally, the influenza-induced mortality is taken into account, with the assumption that both the male and female sub-populations have different case fatalities. Therefore, the model will be formulated and proven to be well-posed in Section 2. The equilibrium and bifurcation analyses, where the model is shown to exhibit backward bifurcation, is presented in Section 3. Motivated by simulations, the asymptotic stability of the endemic equilibria is presented in Section 4. The endemic prevalence of the infection in the overall population and in each sex-structured sub-population are given in Section 5. The paper closes with a summary and conclusion in Section 6.
The total population is assumed to be structured based on the individuals' sex and gender into males (having subscript 1) and females (having subscript 2). Denoting the total population size at time t with N(t), the total male (female) population is denoted by N1(t) (N2(t)), so that N(t)=N1(t)+N2(t). It is assumed that new births occur at an average rate ν from the females; therefore, the total number of births at time t is νN2(t), where a proportion q0 of them are females. Moreover, the natural death rate is assumed to be μ. Then, in the absence of infection, the demography of the closed population, is governed by the following system of equations:
dN1(t)dt=(1−q0)νN2(t)−μN1(t),dN2(t)dt=q0νN2(t)−μN2(t),dN(t)dt=νN2(t)−μN(t). | (2.1) |
If this population is assumed to be demographically stationary, in the sense that its size remains roughly constant over time, then νN2(t)=μN(t). Furthermore, assume that more vacancies are created due to developmental reasons so that the population becomes open and new demographic recruitments through immigration are allowed at an average (per unit time) number of new recruitments Λ(K−N(t))/K, where K is a kind of carrying capacity, K−N(t) represents the extended extra opportunities (more than needed to the original population), and Λ is the maximum rate of immigration. If q2 is the proportion of females amongst the immigrants, then the population dynamics, in the absence of infection, is described by the following dynamical system:
dN1(t)dt=(1−q2)Λ(1−N(t)/K)+(1−q0)μN(t)−μN1(t),dN2(t)dt=q2Λ(1−N(t)/K)+q0μN(t)−μN2(t),dN(t)dt=Λ(1−N(t)/K). | (2.2) |
It is noteworthy that the human carrying capacity refers to the number of people a place can sustainably support. It mainly depends on the size of the population, the availability of resources, and how the people use these available resources. Knowing the carrying capacity is vital for sustainable growth without major setbacks due to environmental degradation. In practice, developed countries use the advancement of technology to increase the carrying capacity of a place, to expand economies of scale, and to use the natural resources in an efficient and effective way without degrading the natural environment [18]. Therefore, pull factors are created to pave the way for immigrations. The closer the population size is to the carrying capacity, the lower the immigration rate. This is clear in the right hand side of system (2.2).
We are interested in modeling the transmission dynamics of a potentially lethal non-immunizing respiratory infection (with special reference to influenza A) in an open population, whose size changes over time. Therefore, both the male and female populations are assumed to split into two mutually exclusive categories (for each) according to the individuals' epidemiological status: susceptible (with time-dependent size S(t)) and infected (with time-dependent size I(t)). Precisely, the number of susceptible males and females at time t are denoted by S1(t) and S2(t) respectivel, while the number of infected males and females are given by I1(t) and I2(t), respectively, so that N1(t)=S1(t)+I1(t) and N2(t)=S2(t)+I2(t). A schematic diagram for the transfer between the subpopulation model states is shown in Figure 1, and a brief description of the model state variables is presented in Table 1. Accordingly, new recruitments (either by births or immigrations) are susceptible.
State variable | Description |
S1(t) | Number of susceptible males. |
I1(t) | Number of infected males. |
S2(t) | Number of susceptible females. |
I2(t) | Number of infected females. |
N1(t) | Total male population size. |
N2(t) | Total female population size. |
N(t)=N1(t)+N2(t) | Total population size. |
Definitely, susceptible males are recruited due to births at the rate (1−q0)μN(t) and due to immigration at the rate (1−q2)Λ(1−N(t)/K), while they die naturally at the rate μ and acquire the influenza A infection at a male force of infection λ1(t) and become infected. Infected males either die due to the infection at the rate c1γ or recover from the infection (without acquiring immunity) and become susceptible again at the rate (1−c1)γ, where c1 is the male case fatality. Hence, the spread of the infection in the male population is described by the following equations:
dS1(t)dt=(1−q2)Λ(1−N(t)/K)+(1−q0)μN(t)−(λ1(t)+μ)S1(t)+(1−c1)γ1I1(t),dI1(t)dt=λ1(t)S1(t)−(μ+γ1)I1(t). | (2.3) |
Similarly, susceptible females are recruited due to births at the rate q0μN(t) and due to immigration at the rate q2Λ(1−N(t)/K), while they decline either due to natural death at the rate μ or due to acquiring the infection and becoming infected and capable of transmitting the infection at an incidence rate λ2(t). Infected females either die naturally at the rate μ or die due to the infection at the rate c2γ2, where c2 is the infection case fatality in females and γ2 is the removal rate of infected females. They recover without immunity at the rate (1−c2)γ2. Therefore, the dynamics of the female population is described by the following two differential equations:
dS2(t)dt=q2Λ(1−N(t)/K)+q0μN(t)−(λ2(t)+μ)S2(t)+(1−c2)γ2I2(t),dI2(t)dt=λ2(t)S2(t)−(μ+γ2)I2(t). | (2.4) |
The male and female forces of infections, λ1(t) and λ2(t), respectively, are derived in a way similar to that shown in Safan [4]. More precisely, we denote the average number of contacts that an individual has with other individuals in the total population by ˜β. Additionally, g1 is the susceptibility of males (which is proportional to the probability of success that a susceptible male becomes infected due to contacts with infected males and females), while g2 is the susceptibility of females (which is proportional to the probability that a susceptible female becomes successfully infected as a result of contacts with infected individuals). Assume that rij accounts for the transmissibility of influenza from infected individuals in the population i to susceptible individuals in the population j, for all i,j∈{1,2}. Hence, the rate at which susceptible males acquire the infection λ1(t) and that at which susceptible females acquire infection λ2(t) read as follows:
λ1(t)=g1˜β(r11I1(t)N(t)+r21I2(t)N(t))andλ2(t)=g2˜β(r12I1(t)N(t)+r22I2(t)N(t)). | (2.5) |
Furthermore, assume that the infected males transmit the infection to either males or females at the same potential, in the sense that the effective rate at which infected males infect susceptible males/females is equal (i.e., r11=r12=r1). Similarly, females transmit the infection to either males or females at the same potential, in the sense that the effective rate at which infected females infect susceptible males/females is equal (i.e., r21=r22=r2). Hence, β=g1r11˜β=g1r1˜β:=g1ˉβ accounts for the effective contact rate at which susceptible males acquire influenza and g2r12˜β=g2r1˜β=g2ˉβ=(g2/g1)g1ˉβ:=gβ is the effective contact rate at which susceptible females acquire the influenza virus. Here, g=g2/g1 accounts for the relative susceptibility of females with respect to males. Hence, (2.5) reads as follows:
λ1(t)=β(I1(t)N(t)+rI2(t)N(t)):=λ(t)andλ2(t)=gβ(I1(t)N(t)+rI2(t)N(t)):=gλ(t), | (2.6) |
where r=r2/r1 is the relative transmissibility of females with respect to male infections.
The removal (either by recovery or due to disease-induced death) rates γ1 and γ2 are rescaled by assuming that γ1=γ and γ2=aγ, where a is a dimensionless rescaling parameter which accounts for the relative recoverability of females with respect to males. A brief description of the model parameters is shown in Table 2.
Parameters | Description | Value | Dim. | Ref. |
Λ | The maximum number of immigrating individuals per unit time. | 11,000 | individuals/week | Assumed |
K | Some kind of carrying capacity. | 35×106 | individuals | Assumed |
μ | Per-capita death rate. | 1/(70×52) | Week−1 | [11] |
ˉβ | Per-capita contact rate between individuals. | – | Week−1 | – |
β | Per-capita effective contact rate at which susceptible males acquire influenza. | Computed to adapt with the value of R0 | Week−1 | Assumed |
γ=γ1 | Per-capita removal (by recovery or disease-induced death) rate for infected males. | 7/3.38 | Week−1 | [4,19] |
γ2=aγ | Per-capita removal (by recovery or disease-induced death) rate for infected females. | – | Week−1 | – |
R0 | The basic reproduction number for model. | 1.525 | – | [4,19] |
a | A rescaling parameter accounting for the relative removability (due to recovery or disease-induced death) of infected females with respect to infected males. | 1.1 | – | Assumed |
q0 | The proportion of female new births. | 0.48∈[0.45,0.55] | – | Assumed |
q2 | The proportion of female immigrants. | 0.45∈[0.45,0.55] | – | Assumed |
g1 | The susceptibility of males. | – | – | – |
g2 | The susceptibility of females. | – | – | – |
g=g2/g1 | Relative susceptibility of females with respect to males. | ∈(0,2) | – | Assumed |
r | Relative transmissibility of the infection by females with respect to males. | ∈(0,2) | – | Assumed |
c1 | Infection-case fatality among males. | 0.007∈[0,0.1] | – | Assumed |
c2 | Infection-case fatality among females. | 0.005∈[0,0.1] | – | Assumed |
Notes: Dim. = Dimension, Ref. = References. |
Motivated by the above-detailed assumptions, the populations' overall dynamics is described by the following system of ordinary differential equations:
dS1(t)dt=(1−q2)Λ(1−N(t)/K)+(1−q0)μN(t)−(λ(t)+μ)S1(t)+(1−c1)γI1(t), | (2.7) |
dI1(t)dt=λ(t)S1(t)−(μ+γ)I1(t), | (2.8) |
dS2(t)dt=q2Λ(1−N(t)/K)+q0μN(t)−(gλ(t)+μ)S2(t)+(1−c2)aγI2(t), | (2.9) |
dI2(t)dt=gλ(t)S2(t)−(μ+aγ)I2(t), | (2.10) |
dN(t)dt=Λ(1−N(t)/K)−c1γI1(t)−c2aγI2(t), | (2.11) |
where
λ(t)=β(I1(t)N(t)+rI2(t)N(t)), | (2.12) |
is the male force of infection and the model is defined on the set
Ω={(S1(t),I1(t),S2(t),I2(t),N(t))T∈R5+,S1(t)+I1(t)+S2(t)+I2(t)=N(t),0≤N(t)≤K}. | (2.13) |
It is worth assuring that the letter T in (2.13) denotes a vector transpose.
The model is well-posed in the sense that solutions starting with the initial conditions (S1(0),I1(0),S2(0),I2(0),N(0))T∈Ω remain in Ω for all positive times. The following proposition, whose proof is deferred to Appendix 10.1, summarizes the results on the existence and uniqueness of the above model time-dependent solutions.
Proposition 1. The set Ω is positively invariant and attracts all solutions in R5+. In particular, any solution (S1(t),I1(t),S2(t),I2(t),N(t))T of the dynamical system (2.7)–(2.11), starting with non-negative initial values (S1(0),I1(0),S2(0),I2(0),N(0))T∈Ω, remains in Ω for all t≥0 and is unique.
To analyze the equilibrium for models (2.7)–(2.11), we put the derivatives in its left hand side equal zero and solve the resulting nonlinear algebraic system of equations, together with (2.12), in the model-state variables. Particularly, the Eqs (2.8) and (2.10) imply that the number of infected males and females, at equilibrium, are given respectively by the following:
I1=λS1γ+μandI2=gλS2aγ+μ. | (3.1) |
Unless otherwise stated, in (3.1) and throughout the rest of this work, the quantities S1,I1,S2,I2,N1,N2,N, and λ denote their equilibrium status. By using (3.1) in the equilibrium equations of (2.7), (2.9), (2.11) and performing some rearrangements, we botain the following:
(μ+(1−(1−c1)γγ+μ)λ)S1N−(1−q2)Λ(1N)=(1−q0)μ−(1−q2)ΛK, | (3.2) |
(μ+(1−(1−c2)aγaγ+μ)gλ)S2N−q2Λ(1N)=q0μ−q2ΛK, | (3.3) |
c1γλγ+μS1N+c2aγgλaγ+μS2N−Λ(1N)=−ΛK. | (3.4) |
Now, we use (3.1) in (2.12) at equilibrium to obtain the following:
λ=βλ(1γ+μ⋅S1N+rg1aγ+μ⋅S2N). | (3.5) |
Equation (3.5) implies that two cases arise: either λ=0 or λ≠0. By (3.1), the first case implies that I1=0 and I2=0. Thus, by (3.2)–(3.4), we get S1=(1−q0)K,S2=q0K and N=K. Therefore, the model has an influenza-free equilibrium E0, where
E0=(S01,I01,S02,I02,N0)T=((1−q0)K,0,q0K,0,K)T. | (3.6) |
The basic reproduction number of the models (2.7)–(2.11) is computed by following the approach shown in [20]. To this end, we consider the Eqs (2.8) and (2.10) in computing the non-negative matrix for the new-infection term T and the non-singular matrix for the remaining transfer term Σ as follows
T=((1−q0)β(1−q0)rβgq0βgq0rβ)andΣ=(γ+μ00aγ+μ). | (3.7) |
Consequently, the basic reproduction number R0 is the spectral radius of the matrix TΣ−1. Therefore,
R0=(1−q0)βγ+μ+rgq0βaγ+μ. | (3.8) |
It is noteworthy that the basic reproduction number R0 is not influenced by the immigration parameters q2,Λ, and K. However, it is affected by the entire population parameters q0,r,g,a,β,γ, and μ. Definitely, R0 increases with the increase of the proportion of newborn females q0 if and only if
rg>aγ+μγ+μ:=DmDf, | (3.9) |
where
● Dm=1γ+μ is the length of males' infectious period, and
● Df=1aγ+μ is the length of females' infectious period.
Equation (3.9) says that the contagiousness and transmissibility of influenza A increases with the increase of the newborn females proportion if and only if the product of the relative susceptibility and the relative transmissibility (of females with respect to males) is higher than the ratio between the male and female infectious periods.
The local stability analysis of the infection-free equilibrium E0 has been established based on linearization. The analysis revealed that E0 is locally asymptotically stable if and only if R0<1 and the proof of this result is deferred to in Appendix 10.2.
Based on the above results, we show the following proposition.
Proposition 2. Models (2.7)–(2.11) have an infection-free equilibrium, given by E0=((1−q0)K,0,q0K,0,K)T, which is locally asymptotically stable if and only if R0<1, where R0 is the basic reproduction number.
If λ≠0, then Eq (3.5) implies that
1=β(1γ+μ⋅S1N+rg1aγ+μ⋅S2N). | (3.10) |
On solving the algebraic system of Eqs (3.2)–(3.4) in terms of S1N,S2N and ΛN, we obtain the following:
S1N=Δ1Δ,S2N=Δ2Δ,ΛN=Δ3Δ, | (3.11) |
where
Δ=−gλ2((1−^c1γγ+μ)(1−~c2aγaγ+μ)−(1−q2)c1γγ+μ(1−^c2aγaγ+μ))−μλ(1−(1−q2c1)γγ+μ+g(1−~c2aγaγ+μ))−μ2, | (3.12) |
Δ1=−(1−q0)μ2−gλμ((1−q0)μaγ+μ+(1−q2)c2aγaγ+μ), | (3.13) |
Δ2=−q0μ2−μλ(q0−(q0−q2c1)γγ+μ), | (3.14) |
Δ3=−μ2ΛK−μλ((1−q0)μc1γγ+μ+(1−(1−q2c1)γγ+μ)ΛK+g((1−~c2aγaγ+μ)ΛK+q0μc2aγaγ+μ))−gλ2(ΛK((1−γγ+μ)(1−~c2aγaγ+μ)+q2c1γγ+μ(1−aγaγ+μ))+q0μc2aγaγ+μ(1−^c1γγ+μ)+(1−q0)μc1γγ+μ(1−^c2aγaγ+μ)), | (3.15) |
and
^c1=1−c1,^c2=1−c2,~c2=1−(1−q2)c2. |
A complete derivation of (3.11)–(3.15) is deferred to in Appendix 10.3. It is noteworthy that the formulas in (3.11) imply that
S1=Δ1Δ3ΛandS2=Δ2Δ3Λ. | (3.16) |
By using (3.16) into (3.1), we obtain the following:
I1=λγ+μ⋅Δ1Δ3⋅ΛandI2=gλaγ+μ⋅Δ2Δ3⋅Λ. | (3.17) |
Now, we substitute (3.11) into (3.10) to obtain the following:
Δ=β(1γ+μΔ1+rg1aγ+μΔ2). | (3.18) |
By using (3.12)–(3.14) in (3.18) and rearranging the terms, we arrive at a second-degree polynomial equation (in λ) in the following form:
F(β,λ)=A2λ2+A1μλ+A0μ2=0, | (3.19) |
where
A2=g(1−^c1γγ+μ)(1−~c2aγaγ+μ)−(1−q2)c1γγ+μ(1−^c2aγaγ+μ),A1=1−(1−q2c1)γγ+μ+g(1−~c2aγaγ+μ)−gβ((1−q0)μ+(1−q2)c2aγ(γ+μ)(aγ+μ)+raγ+μ(q0(1−γγ+μ)+q2c1γγ+μ)),A0=1−β((1−q0)γ+μ+rgq0aγ+μ)=1−R0. |
Equation (3.19) is the endemic force of the infection equation. It is quadratic in λ and may have up to two feasible solutions. Here, the feasibility means that the solution values of λ satisfy λ∈[0,∞). Once a solution of (3.19) is obtained, we substitute it in (3.13)–(3.15) and then in (3.16) and (3.17) to obtain the corresponding equilibrium components, and consequently the corresponding equilibrium point.
The polynomial in the left hand side of Eq (3.19) could be considered a function in the variable λ and the parameter β, given that the other model parameters are kept fixed. Therefore, Eq (3.19) could be seen as a bifurcation equation, with β being the bifurcation parameter. Hence, at λ=0, there is a bifurcation point (β0,0) in the plane (β,λ), where
β0=(γ+μ)(aγ+μ)(1−q0)(aγ+μ)+rgq0(γ+μ). | (3.20) |
To investigate the direction of bifurcation at the bifurcation point (β0,0), we make use of the implicit function theorem by following the same approach shown in [21]. Consequently, we compute (and study the sign of) the following expression:
dλdβ|(β0,0)=−∂F/∂β∂F/∂λ|(β0,0), | (3.21) |
where
∂F∂β|(β0,0)=−μ2(1−q0γ+μ+rgq0aγ+μ)<0, | (3.22) |
∂F∂λ|(β0,0)=μA1|β=β0. | (3.23) |
Thus, the model exhibits the existence of a backward bifurcation if and only if A1|β=β0<0. The following two propositions, whose proofs are deterred to in Appendixes 10.4 and 10.5, summarize the conditions required for the existence of a backward bifurcation.
Proposition 3. Models (2.7)–(2.11) exhibit a backward bifurcation if and only if the following condition holds:
rM1(ℓ1−g)>M2(ℓ2−g), | (3.24) |
where
M1=q0g(γ+μ)2(μ+(1−q2)c2aγ),M2=q0(1−q2)c2aγ(γ+μ)(aγ+μ),ℓ1=(1−q0)q2c1γ(aγ+μ)q0(γ+μ)(μ+(1−q2)c2aγ),ℓ2=(1−q0)(aγ+μ)(μ+q2c1γ)q0(1−q2)c2aγ(γ+μ). | (3.25) |
The following proposition presents a more specific equivalent set of inequalities to inequality (3.24).
Proposition 4. The inequality (3.24) is equivalent to either of the following two sets of inequalities (see Figure 2):
r>M2(ℓ2−g)M1(ℓ1−g):=r1andg<ℓ1, | (3.26) |
orr<M2(g−ℓ2)M1(g−ℓ1):=r2andg>ℓ2. | (3.27) |
Figure 2 shows the regions in the plane (g,r) for which a backward bifurcation occurs. The figure is drawn with parameter values as shown in Table 2. The figure shows that the backward bifurcation phenomenon (i.e., the existence of subcritical endemic states) does possibly exist if a high enough r is chosen (i.e., r>r1), while a small enough g is chosen (i.e., g<ℓ1) and vice versa (i.e., r<r2&g>ℓ2), given that the remaining parameters are kept fixed.
It is worth mentioning that if the inequality (3.24) does not hold, then models (2.7)–(2.11) show the existence of a forward bifurcation (i.e., supercritical endemic states), which, in sense, means that the model has a unique endemic equilibrium that exists if and only if β>β0, or equivalently, if and only if R0>1.
By the effective contact rate threshold [22], we refer to a critical value of the successful contact rate β below which the infection dies out without any effort. In case the model parameters are selected such that the model shows only a forward bifurcation, the effective threshold is the value of β at which the basic reproduction number R0=1 (i.e., β⋆=β0). However, if the backward bifurcation condition (3.24) holds, then two positive solutions of the quadratic equation (3.19) exist for values of R0<1. In this case, both solutions closely approach each other with the decrease of the value of β until they coalesce at the turning point, see the schematic diagram in Figure 3.
The value of β at the turning point, say β1, is the effective threshold. To derive the formula of β1, we proceed as follows. From the implicit function theorem, the conditions for the turning point are as follows:
F(β,λ)=0and∂F∂λ=0. | (3.28) |
The two conditions (3.28) are equivalent to the following:
A21−4A0A2=0. | (3.29) |
Hence, the critical rate β1 is the solution of the Eq (3.29) with respect to the contact rate β. By performing some computations and rearranging the terms, we obtain the following:
β1=K2K1+√(K2K1)2−K3K1, | (3.30) |
where
K1=g2((1−q0)γ+μ(1−aγaγ+μ(1−(1−q2)c21−q0))+rq0aγ+μ(1−(1−q2c1q0)γγ+μ))2,K2=g(1−γγ+μ(1−q2c1)+g(1−~c2aγaγ+μ))((1−q0)γ+μ(1−aγaγ+μ(1−(1−q2)c21−q0))+rq0aγ+μ(1−(1−q2c1q0)γγ+μ))−2g((1−q0)γ+μ+rgq0aγ+μ)((1−^c1γγ+μ)(1−~c2aγaγ+μ)−(1−q2)c1γγ+μ(1−^c2aγaγ+μ)),K3=(1−γγ+μ(1−q2c1)+g(1−~c2aγaγ+μ))2−4g(1−^c1γγ+μ)(1−~c2aγaγ+μ)+4(1−q2)c1γγ+μ(1−^c2aγaγ+μ). |
In summary, we show the following proposition.
Proposition 5. The effective contact rate threshold β⋆ is given by the following:
β⋆={β0,if the condition (3.24) does not hold ,β1,if the condition (3.24) holds, | (3.31) |
where β0 and β1 are defined in (3.20) and (3.30), respectively.
As a function of the relative transmissibility parameter r, the effective contact rate threshold β⋆ is depicted in Figure 4. The left Subfigure 4(a) corresponds to the backward bifurcation set of inequalities (3.26), while the right Subfigure 4(b) corresponds to the inequality set (3.27). In each subfigure, the solid curve represents β=β0 (i.e., R0=1), while the broken curve represents β=β1. In the region above the solid curve, the model has a unique endemic equilibrium. However, in the region that lies in between the solid and broken curves, the model has two endemic equilibria. Otherwise, the model has no endemic equilibrium.
In terms of the basic reproduction number R0, the effective basic reproduction threshold R⋆0 is the value below which the infection disappears and does not persist. In case the model shows a forward bifurcation, this effective threshold is R⋆0=1. However, if the model exhibits a backward bifurcation, the effective threshold is as follows:
R⋆0=β1/β0:=R10<1. | (3.32) |
In summary, we have the following proposition.
Proposition 6. The effective basic reproduction number threshold is as follows:
R⋆0={1,if the condition (3.24) does not hold,R10,if the condition (3.24) holds. | (3.33) |
The effective basic reproduction threshold R⋆0 is drawn in the plane (r,R0) and shown in Figure 5. Figure 5(a) shows that, for small enough values of the relative susceptibility g (of females with respect to males), multiple endemic equilibria do exist for values of R10≤R0<1 in a very narrow region on the right of the plane (r,R0). In this case, the multiple equilibria region becomes wide with the increase of the relative transmissibility parameter r. However, for high enough values of the relative susceptibility parameter g, the multiple equilibria region exists on the left of the plane (r,R0) and diminishes with the increase of r, see Figure 5(b).
In case the model undergoes a backward bifurcation, it should be underlined that the effective contact rate threshold β1 depends on the proportion of immigrating females q2. Consequently, the effective basic reproduction number threshold R10 becomes influenced by any increase or decrease in q2.
Motivated by the results shown in Sections 3.3 and 3.4, Eq (3.19) has two feasible solutions if the condition (3.24) holds. The first solution is given by
λ−=μ2A2(−A1−√A21−4A0A2), | (3.34) |
and exists if and only if the condition (3.24) holds together with the inequality β1≤β≤β0. It is worth mentioning that the inequality β1≤β≤β0 is equivalent to R10≤R0≤1, where R0=β/β0 and R10=β1/β0. The other solution of Eq (3.19) is given by
λ+=μ2A2(−A1+√A21−4A0A2), | (3.35) |
and exists if and only if β>β1 (i.e., if and only if R0≥R10).
Both solutions are depicted in the plane (R0,λ) and shown in Figure 6(a). The solution λ− is represented by the broken curve, while the solution λ+ is represented by the solid curve. They collide at the turning point.
However, if the backward bifurcation condition (3.24) does not hold, then the Eq (3.19) has a unique feasible solution that definitely exists if and only if β>β0 (or equivalently, R0>1). This unique solution is given by the formula (3.35) and is depicted in Subfigure 6(b).
By using (3.34) and (3.35) within (3.13)–(3.15) and then with (3.16) and (3.17), we obtain the corresponding equilibrium points whose formulas are determined in the following proposition.
Proposition 7. The equilibrium analysis of models (2.7)–(2.11) reveals the following results on the existence of endemic equilibria:
● If the condition (3.24) does not hold, then the model has a unique endemic equilibrium that exists if and only if R0>1 and is given by
E+=(S+1,I+1,S+2,I+2,N+)T, | (3.36) |
where
S+1=Λ⋅(Δ1Δ3)|λ=λ+,I+1=Λ⋅λ+γ+μ⋅(Δ1Δ3)|λ=λ+,S+2=Λ⋅(Δ2Δ3)|λ=λ+,I+2=Λ⋅gλ+aγ+μ⋅(Δ2Δ3)|λ=λ+,N+=Λ⋅(ΔΔ3)|λ=λ+. | (3.37) |
● If the backward bifurcation condition (3.24) holds, then the model has two endemic equilibria. One of them is in the form (3.36) and exists if β≥β1 (i.e., R0≥R10). The other one exists if R10≤R0<1 and is given by
E−=(S−1,I−1,S−2,I−2,N−)T, | (3.38) |
where
S−1=Λ⋅(Δ1Δ3)|λ=λ−,I−1=Λ⋅λ−γ+μ⋅(Δ1Δ3)|λ=λ−,S−2=Λ⋅(Δ2Δ3)|λ=λ−,I−2=Λ⋅gλ−aγ+μ⋅(Δ2Δ3)|λ=λ−,N−=Λ⋅(ΔΔ3)|λ=λ−. | (3.39) |
● Otherwise, the model has no endemic equilibrium.
In case the model undergoes a backward bifurcation, it is worth noting that the solution λ− lies between λ=0 and λ=λ+. Moreover, the equilibrium point E− is unstable, while the equilibrium point E+ is locally asymptotically stable, whenever it exists. Therefore, the curve representing the solution λ=λ− in the plane (R0,λ) of Figure 6(a) is drawn as a broken line to distinguish it from the solid curve representing the solution λ=λ+ that corresponds to the stable endemic equilibrium.
Due to the complicated terms in the formulas of the endemic equilibria, simulations have been performed to study the asymptotic stability of the model's equilibrium points. To this end, the function ode45 in Matlab has been used to numerically solve the models (2.7)–(2.10) with various randomly selected initial conditions chosen so that they lie in the set of definition Ω. Although the state variables in the model represent the number of individuals, we draw the solutions in the form of proportions to better present the results, as shown in Figures 7–9. These figures are produced with parameter values as shown in Table 2; the parameters r and g are given the values r=2 and g=0.8, while the contact rate β is chosen so that the basic reproduction number R0 takes three different values to explain different scenarios for the fate of the trajectory solution (i.e., the attracting equilibrium point(s)). It is worth confirming that the state variable notations in the legend of the vertical axis in Figures 7–9 denote proportions rather than numbers.
Based on the chosen parameter values, the model undergoes a backward bifurcation at R0=1 and the turning point occurs at the point (R10,λ1)=(0.9973,0.0163) in the plane (R0,λ). Therefore, Figure 7 has been produced with a value of R0=0.8<R10. In this case, no endemic equilibrium exists, and, in consequence, the figure shows that all trajectory solutions are attracted by the infection-free equilibrium (I1=I2=0). However, the solutions are shown in Figure 8 with a value of R0=0.998∈(R10,1). The figure shows that the solutions are attracted either by the infection-free equilibrium (I1=I2=0) or by an endemic equilibrium (i.e., I1>0 and I2>0). Finally, the model has been solved with a value of R0=1.3>1 and the solutions are shown in Figure 9. The figure shows that all solutions are attracted by a unique non-trivial endemic equilibrium (i.e., I1>0 and I2>0).
Motivated by the results shown in Figures 7–9, our in silico simulations show three scenarios for the evolution of influenza infection if the model undergoes backward bifurcation at R0=1. The first scenario is that both of the time-dependent proportions of infected males and infected females eventually approach zero (i.e., their values at the influenza-free equilibrium as shown in Figure 7(b), (d)); therefore, the infection washes out without any further efforts. This scenario is ensured if the combination of the model parameters is chosen so that R0<R10. Another scenario is that both of the time-dependent proportions of infected males and infected females eventually approach a positive level (i.e., their correspondents at the endemic equilibrium as shown in Figure 9(b), (d)); therefore, the infection persists in the population. This scenario happens if the combination of the model parameters used in the simulations satisfies R0>1. The third scenario mixes the above-mentioned two scenarios, where some of the solutions eventually approach the influenza-free equilibrium, while the others eventually approach an influenza endemic equilibrium (as shown in Figure 8(b), (d)). This scenario occurs if the combination of the model parameters used to simulate the model are chosen to satisfy R10≤R0≤1. In the third scenario, the infection elimination depends on the initial conditions. It is worth noting that in case the model only undergoes a forward bifurcation, then there are only the first two scenarios.
In the absence of an endemic infection (i.e., λ=0), the formulas (3.12)–(3.15) imply that
Δ=−μ2,Δ1=−(1−q0)μ2,Δ2=−q0μ2,Δ3=−μ2ΛK. | (5.1) |
Therefore, by (3.11) and (3.17), we obtain the subpopulation proportions
S1N=N1N=1−q0,I1N=0,S2N=N2N=q0,I2N=0, | (5.2) |
and the total population size N=K.
However, in the presence of an endemic infection (i.e., λ≠0), the formulas (3.11)–(3.15) and (3.17) help compute the following expressions.
● The total population size at equilibrium in the endemic situation is given by the following:
N=Λ⋅ΔΔ3. | (5.3) |
● The male equilibrium proportion in the endemic situation, say pm, is given by the following:
pm=N1N=(1+λγ+μ)Δ1Δ. | (5.4) |
● The female equilibrium proportion in the endemic situation, say pf, is given by the following:
pf=1−N1N=N2N=(1+gλaγ+μ)Δ2Δ. | (5.5) |
● The endemic prevalence of infection in the male population, say pI1, is given by the following:
pI1=I1N1=λλ+γ+μ. | (5.6) |
● The endemic prevalence of infection in the female population, say pI2, is given by the following:
pI2=I2N2=gλgλ+aγ+μ. | (5.7) |
● The endemic prevalence of infection in the overall population, say pI, is given by the following:
pI=I1+I2N=λγ+μ⋅Δ1Δ+gλaγ+μ⋅Δ2Δ. | (5.8) |
It is worth mentioning that λ is the males' equilibrium force of infection and is(are) the feasible solution(s) of the Eq (3.19).
Based on various values of the relative transmissibility parameter r and the relative susceptibility parameter g selected from the different regions in the plane (g,r), the expressions in the formulas (5.3)–(5.8) have been drawn as functions of the basic reproduction number R0 and presented in Figures 10–13, while keeping the other parameters fixed as shown in Table 2. The values of the pair (g,r) are shown in the head of each subfigure.
Figures 10 and 11 have been produced with values of r and g such that the model undergoes a forward bifurcation at R0=1. The curves in both figures correspond to the solution λ=λ+ (i.e., they are connected to the endemic equilibrium E+). Each figure is produced with three different values of the pair (g,r), where r is kept fixed for the same figure, though g is allowed to change. The solid curves are drawn with higher values of the relative susceptibility (g=1.5>ℓ2=1.2932) than in the cases of the dotted curve (g=0.8<ℓ1=1.1888) and the broken curve (g=1.10<ℓ1=1.1888).
From a demographic perspective, it is clear that the Figure 10(a), (d) shows that higher values of the relative susceptibility level g implies a reduction in the equilibrium total population size and the equilibrium proportion of female subpopulation, while the contrast is remarkable for the proportion of males at equilibrium pm as shown in the Figure 10(c). Figure 10(c), (d) shows that, in the case of a high enough relative susceptibility level, the proportion of males pm (females pf) at equilibrium strictly initially increases (decreases) with an increase of R0 till reaching a maximum (minimum) and then decreases (increases) with an increase of R0. However, contrasting qualitative behaviors of both pm and pf are remarkable in the case of low enough relative susceptibility levels (g=0.8<ℓ1=1.1888 and g=1.1<ℓ1=1.1888).
From an epidemiological perspective, Figure 10(b), (e), (f) shows that the high relative susceptibility of females with respect to males increases the endemic prevalence of infection in the overall population pI, the endemic prevalence of infection within males pI1, and the endemic prevalence of infection within females pI2, respectively. These prevalences strictly increase with an increase of the basic reproduction number R0.
It is worth explaining that even if the relative transmissibility parameter r is increased (but the model still undergoes only forward bifurcation at R0=1), then the qualitative behavior doesn't differ from that shown in Figure 10, see Figure 11.
The above-described qualitative behavior is ensured as long as the males case fatality is higher than that of the females (i.e., if c1>c2). However, in the opposite case (i.e., if c1<c2), this behavior is different as shown in Figure 14, where two contrasting behaviors are remarkable for values of R0>1. Definitely, for values of R0 in the right-neighbourhood of R0=1, the equilibrium total population size N and the equilibrium proportion of females pf decrease with an increase of the relative susceptibility g. However, for high enough values of R0, they increase with an increase of the relative susceptibility parameter g. On the contrary, the endemic prevalence levels pI,pI1, and pI2 decrease with an increase of the relative susceptibility parameter g for values of R0 slightly above one, while they increase for high enough values of R0. Based on values of r and g that ensure the model undergoes a backward bifurcation at R0=1, the expressions in (5.3)–(5.8) are drawn as functions of the basic reproduction number R0 and presented in Figures 12 and 13. There are two curves for values of R0 slightly less than one. The broken curve corresponds to the unstable endemic equilibrium (computed based on a value of λ=λ−), while the solid one corresponds to the stable endemic equilibrium (computed based on a value of λ=λ+), as implicated by the occurrence of a backward bifurcation. Only the solid curves are of interest, as they represent the stable endemic equilibrium computed based on a value of the endemic force of the infection solution λ+. Both figures show that the qualitative behavior of the functions based on the stable endemic equilibrium remains the same as the behavior shown in Figure 10 and is detailed in the above description. It is worth mentioning that several simulations have been performed to explore the qualitative behavior if the females' case fatality is higher than the males' one. The simulations show that the qualitative behavior of the demographic expressions N,pm, and pf is similar to their correspondences in Figure 14, while the endemic prevalences pI,pI1, and pI2 keep the same behavior as in the case c1>c2.
Motivated by the above analysis, we come up with the following.
1) If the male's case fatality is higher than the female's one, then we have the following results:
● Reducing the relative susceptibility of females with respect to males reduces the endemic prevalence of the infection in the total population and in each sex-structured subpopulation, though it increases for both the female proportion pf and the total population size at equilibrium.
● For high enough levels of the relative susceptibility (g>ℓ2), reducing the basic reproduction number R0 reduces the equilibrium proportion of females until a minimum close to the right-neighbourhood of R0=1 is reached, and then increases again, while the converse is true for small levels of the relative susceptibility parameter g<ℓ1.
2) However, if the male's case fatality is lower than the female's one, then the two contrasting scenarios are remarkable. For small values of R0 which lie in the right neighbourhood of R0=1, the above-mentioned implications are reversed, while those implications do still work for higher values for R0.
The dynamics of influenza A infection has been extensively explored based on mathematical models. However, mathematical models that take the inequalities due to differences in sex and gender into account are less investigated. Motivated by a previous work [4], a SIS model was extended and adapted to describe the dynamics of influenza A in an open population with varying size, where the infection lethality was taken into account. From an epidemiological perspective, the model was adapted to consider differences in susceptibility, infectivity, infection-induced mortality (i.e., lethality of the infection), and recoverability between males and females. However, from a demographic point of view, the model was adapted to consider an open population with a population-size-structured immigration rate. Additionally, the inequality of the birth ratio of females and males was included.
The model has been mathematically analyzed. Definitely, the well-posedness of the model was shown, where the existence and uniqueness of time-dependent solutions and the positive invariance of the model's definition set was proven. The model's equilibrium analysis revealed that the model has an influenza-free equilibrium that was proven to be locally asymptotically stable if and only if R0<1, where R0 is the model's basic reproduction number. Moreover, the bifurcation analysis showed that the model underwent a backward bifurcation at R0=1 for a certain space-set of the model parameters. The conditions for the occurrence of a backward bifurcation was determined and presented in the form of either of the two inequality sets (r>r1,g<ℓ1) or (r<r2,g>ℓ2), where r1,r2,ℓ1 and ℓ2 were defined within the text (formulas (3.25)–(3.27)). The occurrence of a backward bifurcation made the model's behavior more complicated than in case of a forward bifurcation, especially when discussing the possibility to eliminate the infection.
The mathematical implication of the backward bifurcation phenomenon in epidemic models is that a two influenza-endemic (i.e., with positive levels of the infection's state variables) equilibria co-exists with the influenza-free equilibrium for values of R10≤R0<1, where the endemic equilibrium with a higher infection level (i.e., with λ=λ+) is locally asymptotically stable, while the endemic equilibrium with a lower level of the infection (i.e., with λ=λ−) is unstable. The asymptotic local stability of the model's equilibrium solutions was numerically investigated.
The epidemiological implication of the existence of a backward bifurcation is that reducing the basic reproduction number R0 to values slightly less than one is a necessary but no longer a sufficient condition to eliminate the infection. In other words, in the case of a backward bifurcation, the value R0=1 is no longer a threshold, while R0=R10 is the effective threshold (whose formula is given in (3.32)), and strategies aiming to eliminate the infection would be based on reducing R0 to slightly below the effective threshold value R10. Therefore, the minimum effort required to eliminate the infection becomes increased [21,23].
Our analysis showed that the proportion of female new-immigrants didn't affect the value of the basic reproduction number, but it affected the value of the effective basic reproduction number threshold and its influence was associated with the value of the relative susceptibility parameter g. However, the basic reproduction number was affected by the value of the female newborns proportion q0 and the impact's type was determined by the inequality (3.9).
Some demographic quantities (precisely, the total population size and proportion of male and female subpopulations at equilibrium) and some epidemiological expressions (precisely, the endemic prevalence of the infection in the total population, as well as in both male and female subpopulations) were computed and numerically investigated. The effect of changes in the relative transmissibility and susceptibility parameters (r and g, respectively), as well as in the infection's case fatality in both the male and female subpopulations on these demographical and epidemiological expressions, were numerically investigated.
The authors declare they have not used Artificial Intelligence (AI) tools in the creation of this article.
The authors declare that this research is not supported by funding. And, the authors are grateful to the anonymous reviewers for their constructive comments.
The authors declare there is no conflict of interest.
To show that the model is well-posed, we first consider the N-equation in (2.11), where
dN(t)dt=Λ(1−N(t)/K)−c1γI1(t)−c2aγI2(t)≥−(ΛK+c1γ+c2aγ)N. |
Hence, by the comparison theorem, we obtain the following:
N(t)≥N(0)exp(−(ΛK+c1γ+c2aγ)t)≥0∀N(0)≥0, | (A1.1) |
where N(0) is the total population size at t=0. Additionally, we have
dN(t)dt=Λ(1−N(t)/K)−c1γI1(t)−c2aγI2(t)≤Λ(1−N(t)/K), |
and using the comparison theorem, we obtain the following:
N(t)≤K(1−(1−N(0)K)exp(−ΛKt))→Kast→∞. | (A1.2) |
Hence,
0≤N(t)≤K, | (A1.3) |
i.e., N(t) is upper-bounded.
In the same manner, we may use (2.7)–(2.11) to write
dS1(t)dt≥−(λ(t)+μ)S1(t),dI1(t)dt≥−(μ+γ)I1(t),dS2dt≥−(gλ(t)+μ)S2(t),dI2(t)dt≥−(μ+aγ)I2(t), |
and thereby we get
S1(t)≥S1(0)exp(−∫t0(λ(τ)+μ)dτ)≥0∀S1(0)≥0,I1(t)≥I1(0)exp(−(γ+μ)t)≥0∀I1(0)≥0,S2(t)≥S2(0)exp(−∫t0(gλ(τ)+μ)dτ)≥0∀S2(0)≥0,I2(t)≥I2(0)exp(−(aγ+μ)t)≥0∀I2(0)≥0. |
Now, since 0≤S1(t)+I1(t)+S2(t)+I2(t)=N(t), and since both N(t) is upper-bounded, then S1(t),I1(t),S2(t),I2(t) are all upper-bounded and non-negative for all t≥0. Thus, the set Ω is positively invariant.
To show the uniqueness of the time-dependent solutions of models (2.7)–(2.11), we note that the right hand side of these equations are all continuous in the models state variables S1,S2,I1, and I2. Moreover, it is easy to check that all partial derivatives of the functions in the right-hand side of models (2.7)–(2.11) are also continuous in the model state variables. Therefore, they are locally Lipschitz, and, hence, any time-depending solution starting with initial conditions in Ω is unique.
We apply the linearization principle to establish the local stability analysis of the influenza-free equilibrium E0. Rather than traditionally using the S1,I1,S2, and I2 equations in models (2.7)–(2.11), we consider the N1,N2,I1, and I2 equations, as S1=N1−I1 and S2=N2−I2. Therefore, we consider the following model:
dN1(t)dt=(1−q2)Λ(1−N1(t)+N2(t)K)+(1−q0)μ(N1(t)+N2(t))−μN1(t)−c1γI1:=f1(N1,N2,I1,I2),dN2(t)dt=q2Λ(1−N1(t)+N2(t)K)+q0μ(N1(t)+N2(t))−μN2(t)−c2aγI2:=f2(N1,N2,I1,I2),dI1(t)dt=β(I1(t)+rI2(t)N1(t)+N2(t))(N1(t)−I1(t))−(μ+γ)I1(t):=f3(N1,N2,I1,I2),dI2(t)dt=gβ(I1(t)+rI2(t)N1(t)+N2(t))(N2(t)−I2(t))−(μ+aγ)I2(t):=f4(N1,N2,I1,I2), |
which has the following influenza-free equilibrium ˆE0=((1−q0)K,q0K,0,0)T that corresponds to E0. The local stability of ˆE0 implies the local stability of E0. The Jacobean matrix evaluated at ˆE0 is a block matrix in the following triangular form:
J|ˆE0=[AB0D], | (A2.1) |
where
A=[∂f1∂N1∂f1∂N2∂f2∂N1∂f2∂N2]=[−(1−q2)ΛK−q0μ−(1−q2)ΛK+(1−q0)μ−q2ΛK+q0μ−q2ΛK+q0μ−μ],B=[∂f1∂I1∂f1∂I2∂f2∂I1∂f2∂I2]=[−c1γ00−c2aγ],D=[∂f3∂I1∂f3∂I2∂f4∂I1∂f4∂I2]=[(1−q0)β−(γ+μ)(1−q0)rβgq0βq0rgβ−(μ+aγ)]. |
It is noteworthy that the characteristic polynomial of the block triangular matrix (A2.1) is the product of the characteristic polynomials of the matrices A and D. Therefore, the local stability of the influenza-free equilibrium ˆE0 is ensured if and only if the following conditions on the matrices A and D hold:
det(A)>0,Tr(A)<0,det(D)>0,andTr(D)<0. | (A2.2) |
To this end, we compute the following:
det(A)=(−(1−q2)ΛK−q0μ)(−q2ΛK+q0μ−μ)−(−q2ΛK+q0μ)(−(1−q2)ΛK+(1−q0)μ)=ΛKμ>0,Tr(A)=−(1−q2)ΛK−q0μ−q2ΛK+q0μ−μ=−ΛK−μ<0,det(D)=((1−q0)β−(γ+μ))(q0rgβ−(μ+aγ))−gq0(1−q0)rβ2=(γ+μ)(aγ+μ)(1−β(1−q0γ+μ+rgq0aγ+μ))=(γ+μ)(aγ+μ)(1−R0),Tr(D)=((1−q0)β−(γ+μ))+(q0rgβ−(μ+aγ)). |
It is clear that det(D)>0 if and only if R0<1. However, if R0<1, then we have (1−q0)β<(γ+μ) and q0rgβ<(μ+aγ), which implies that Tr(D)<0. Thus, the condition (A2.2) holds if and only if R0<1.
To derive Eq (3.19), we rewrite the algebraic system of equations (3.2)–(3.4) in the following simple form:
B11(S1N)+B13(ΛN)=E1, | (A3.1) |
B22(S2N)+B23(ΛN)=E2, | (A3.2) |
B31(S1N)+B32(S2N)+B33(ΛN)=E3, | (A3.3) |
where
B11=μ+λ(1−(1−c1)γγ+μ),B13=−(1−q2),E1=(1−q0)μ−(1−q2)ΛK,B22=μ+gλ(1−(1−c2)aγaγ+μ),B23=−q2,E2=q0μ−q2ΛK,B31=c1γγ+μλ,B32=c2aγaγ+μgλ,B33=−1,E3=−ΛK. |
Now, we solve the systems (A3.1)–(A3.3) in terms of S1N,S2N, and ΛN to obtain the following:
S1N=Δ1Δ,S2N=Δ2Δ,ΛN=Δ3Δ, | (A3.4) |
where
Δ=|B110B130B22B23B31B32B33|=B11(B22B33−B23B32)−B13B31B22=(μ+(1−(1−c1)γγ+μ)λ)(−μ−(1−(1−c2)aγaγ+μ)gλ+c2aγaγ+μq2gλ)+c1γγ+μ(1−q2)λ(μ+(1−(1−c2)aγaγ+μ)gλ)=−gλ2((1−(1−c1)γγ+μ)(1−(1−(1−q2)c2)aγaγ+μ)−(1−q2)c1γγ+μ(1−(1−c2)aγaγ+μ))−μλ(1−(1−q2c1)γγ+μ+g(1−(1−(1−q2)c2)aγaγ+μ))−μ2,=−gλ2((1−^c1γγ+μ)(1−~c2aγaγ+μ)−(1−q2)c1γγ+μ(1−^c2aγaγ+μ))−μλ(1−(1−q2c1)γγ+μ+g(1−~c2aγaγ+μ))−μ2, | (A3.5) |
where
^c1=1−c1,^c2=1−c2,~c2=1−(1−q2)c2. |
Δ1=|E10B13E2B22B23E3B32B33|=E1(B22B33−B23B32)−B13(E2B32−E3B22)=((1−q0)μ−(1−q2)ΛK)(−μ−gλ(1−(1−c2)aγaγ+μ)+c2aγaγ+μq2gλ)+(1−q2)(ΛK(μ+gλ(1−(1−c2)aγaγ+μ))−(q0μ−q2ΛK)c2aγaγ+μgλ)=((1−q0)μ−(1−q2)ΛK)(−μ−gλ(1−(1−(1−q2)c2)aγaγ+μ))+(1−q2)(−μΛK−gλ(ΛK+q0c2μaγaγ+μ−((1−(1−q2)c2)ΛK)aγaγ+μ))=−(1−q0)μ2−gλ((1−q0)μ−(1−q0)μ(1−(1−q2)c2)aγaγ+μ+(1−q2)c2q0μaγaγ+μ)=−(1−q0)μ2−gλμ(1−q0−(1−(1−q2)c2−q0+q0(1−q2)c2−q0(1−q2)c2))aγaγ+μ=−(1−q0)μ2−gλμ((1−q0)−(1−q0)aγaγ+μ+(1−q2)c2aγaγ+μ)=−(1−q0)μ2−gλμ((1−q0)μaγ+μ+(1−q2)c2aγaγ+μ), | (A3.6) |
Δ2=|B11E1B130E2B23B31E3B33|=B11(E2B33−E3B23)+B31(E1B23−E2B13)=(μ+λ(1−(1−c1)γγ+μ))(−q0μ+q2ΛK−q2ΛK)+c1γγ+μλ(−q2(1−q0)μ+q2(1−q2)ΛK+(1−q2)q0μ−(1−q2)q2ΛK)=−q0μ2+μλ((q0−q2)c1γγ+μ−q0+q0(1−c1)γγ+μ)=−q0μ2+μλ(−q0+(q0−q2c1)γγ+μ), | (A3.7) |
Δ3=|B110E10B22E2B31B32E3|=B11(E3B22−E2B32)−E1B22B31=(μ+(1−(1−c1)γγ+μ)λ)(−(μ+(1−(1−c2)aγaγ+μ)gλ)ΛK−(q0μ−q2ΛK)c2aγaγ+μgλ)−((1−q0)μ−(1−q2)ΛK)(μ+(1−(1−c2)aγaγ+μ)gλ)c1γγ+μλ=(μ+(1−(1−c1)γγ+μ)λ)(−μΛK−gλ((1−(1−(1−q2)c2)aγaγ+μ)ΛK+q0μc2aγaγ+μ))−((1−q0)μ−(1−q2)ΛK)(μλ+(1−(1−c2)aγaγ+μ)gλ2)c1γγ+μ=−μ2ΛK−μλ((1−q0)μc1γγ+μ+(1−(1−q2c1)γγ+μ)ΛK+g((1−(1−(1−q2)c2)aγaγ+μ)ΛK+q0μc2aγaγ+μ))−gλ2(ΛK((1−(1−c1)γγ+μ)(1−(1−(1−q2)c2)aγaγ+μ)−(1−q2)c1γγ+μ(1−(1−c2)aγaγ+μ))+q0μc2aγaγ+μ(1−(1−c1)γγ+μ)+(1−q0)μc1γγ+μ(1−(1−c2)aγaγ+μ))=−μ2ΛK−μλ((1−q0)μc1γγ+μ+(1−(1−q2c1)γγ+μ)ΛK+g((1−~c2aγaγ+μ)ΛK+q0μc2aγaγ+μ))−gλ2(ΛK((1−γγ+μ)(1−~c2aγaγ+μ)+q2c1γγ+μ(1−aγaγ+μ))+q0μc2aγaγ+μ(1−^c1γγ+μ)+(1−q0)μc1γγ+μ(1−^c2aγaγ+μ)). | (A3.8) |
To prove Proposition 3, we make use of the formulas (3.21)–(3.23). They imply that our model exhibits a backward bifurcation if and only if A1|β=β0<0. Now, we have the following:
A1|β=β0=1+g−(1−q2)c1γγ+μ−(1−c1)γγ+μ−gaγ(1−(1−q2)c2)aγ+μ−g((1−q0)μaγ+μ+(1−q2)c2aγaγ+μ)aγ+μ(1−q0)(aγ+μ)+rgq0(γ+μ)+rg(γ+μ)(1−q0)(aγ+μ)+rgq0(γ+μ)(−q0+(q0−q2c1)γγ+μ)=1+g−(1−q2c1)γγ+μ−gaγ(1−(1−q2)c2)aγ+μ−g((1−q0)μ+(1−q2)c2aγ+rq0μ+rq2c1γ)(1−q0)(aγ+μ)+rgq0(γ+μ)=μ+q2c1γγ+μ+g(μ+(1−q2)c2aγ)aγ+μ−g((1−q0)μ+(1−q2)c2aγ+r(q0μ+q2c1γ))(1−q0)(aγ+μ)+rgq0(γ+μ)=(aγ+μ)(μ+q2c1γ)+g(γ+μ)(μ+(1−q2)c2aγ)(γ+μ)(aγ+μ)−g((1−q0)μ+(1−q2)c2aγ+r(q0μ+q2c1γ))(1−q0)(aγ+μ)+rgq0(γ+μ). |
Hence, A1|β=β0<0 if and only if
g(γ+μ)(aγ+μ)((1−q0)μ+(1−q2)c2aγ+r(q0μ+q2c1γ))>((1−q0)(aγ+μ)+rgq0(γ+μ))((aγ+μ)(μ+q2c1γ)+g(γ+μ)(μ+(1−q2)c2aγ)), |
i.e.,
rg(γ+μ)((1−q0)q2c1γ(aγ+μ)−gq0(γ+μ)(μ+(1−q2)c2aγ))>(1−q0)(aγ+μ)2(μ+q2c1γ)+(1−q0)(aγ+μ)(1−q2)c2aγg(γ+μ)−(1−q2)c2aγg(γ+μ)(aγ+μ), |
i.e.,
rgq0(γ+μ)2(μ+(1−q2)c2aγ)((1−q0)q2c1γ(aγ+μ)q0(γ+μ)(μ+(1−q2)c2aγ)−g)>q0(1−q2)c2aγ(γ+μ)(aγ+μ)((1−q0)(aγ+μ)(μ+q2c1γ)q0(1−q2)c2aγ(γ+μ)−g), |
i.e.,
rM1(ℓ1−g)>M2(ℓ2−g), | (A4.1) |
where
M1=q0g(γ+μ)2(μ+(1−q2)c2aγ),M2=q0(1−q2)c2aγ(γ+μ)(aγ+μ),ℓ1=(1−q0)q2c1γ(aγ+μ)q0(γ+μ)(μ+(1−q2)c2aγ),ℓ2=(1−q0)(aγ+μ)(μ+q2c1γ)q0(1−q2)c2aγ(γ+μ). |
To prove Proposition 4, we notice that the numerator of ℓ1 is less than the numerator of ℓ2, while the denominator of ℓ1 is bigger than the denominator of ℓ2. Therefore, ℓ2>ℓ1. Hence, (ℓ1−g) and (ℓ2−g) have the same sign (either negative or positive) if and only if either g<min(ℓ1,ℓ2)=ℓ1 or g>max(ℓ1,ℓ2)=ℓ2. Therefore, the inequality (A4.1) holds if either g<ℓ1 or g>ℓ2. In the first case, we obtain the following condition:
r>M2(ℓ2−g)M1(ℓ1−g)andg<ℓ1, |
while in the second case, we obtain the following condition:
r<M2(g−ℓ2)M1(g−ℓ1)andg>ℓ2. |
This completes the proof.
[1] | J. Kennedy, R. Eberhart, Particle swarm optimization, in Proceedings of ICNN'95 - International Conference on Neural Networks, 4 (1995), 1942–1948. https://doi.org/10.1109/ICNN.1995.488968 |
[2] |
A. H. Gandomi, A. H. Alavi, Krill herd: a new bio-inspired optimization algorithm, Commun. Nonlinear Sci. Numer. Simul., 17 (2012), 4831–4845. https://doi.org/10.1016/j.cnsns.2012.05.010 doi: 10.1016/j.cnsns.2012.05.010
![]() |
[3] |
A. H. Gandomi, X. Yang, A. H. Alavi, Cuckoo search algorithm: a metaheuristic approach to solve structural optimization problems, Eng. Comput., 29 (2013), 17–35. https://doi.org/10.1007/s00366-011-0241-y doi: 10.1007/s00366-011-0241-y
![]() |
[4] |
S. Mirjalili, S. M. Mirjalili, A. Lewis, Grey wolf optimizer, Adv. Eng. Software, 69 (2014), 46–61. https://doi.org/10.1016/j.advengsoft.2013.12.007 doi: 10.1016/j.advengsoft.2013.12.007
![]() |
[5] |
R. V. Rao, V. J. Savsani, D. P. Vakharia, Teaching–learning-based optimization: a novel method for constrained mechanical design optimization problems, Comput.-Aided Des., 43 (2011), 303–315. https://doi.org/10.1016/j.cad.2010.12.015 doi: 10.1016/j.cad.2010.12.015
![]() |
[6] |
S. Mirjalili, A. Lewis, The whale optimization algorithm, Adv. Eng. Software, 95 (2016), 51–67. https://doi.org/10.1016/j.advengsoft.2016.01.008 doi: 10.1016/j.advengsoft.2016.01.008
![]() |
[7] |
A. Faramarzi, M. Heidarinejad, S. Mirjalili, A. Gandomi, Marine predators algorithm: a nature-inspired metaheuristic, Expert Syst. Appl., 152 (2020), 113377. https://doi.org/10.1016/j.eswa.2020.113377 doi: 10.1016/j.eswa.2020.113377
![]() |
[8] |
J. Sarvaiya, D. Singh, Selection of the optimal process parameters in friction stir welding/processing using particle swarm optimization algorithm, Mater. Today: Proc., 62 (2022), 896–901. https://doi.org/10.1016/j.matpr.2022.04.062 doi: 10.1016/j.matpr.2022.04.062
![]() |
[9] |
Z. Hu, H. Norouzi, M. Jiang, S. Dadfar, T. Kashiwagi, Novel hybrid modified krill herd algorithm and fuzzy controller based MPPT to optimally tune the member functions for PV system in the three-phase grid-connected mode, ISA trans., 2022 (2022). https://doi.org/10.1016/j.isatra.2022.02.009 doi: 10.1016/j.isatra.2022.02.009
![]() |
[10] |
Q. Bai, H. Li, The application of hybrid cuckoo search-grey wolf optimization algorithm in optimal parameters identification of solid oxide fuel cell, Int. J. Hydrogen Energy, 47 (2022), 6200–6216. https://doi.org/10.1016/j.ijhydene.2021.11.216 doi: 10.1016/j.ijhydene.2021.11.216
![]() |
[11] |
C. Song, X. Wang, Z. Liu, H. Chen, Evaluation of axis straightness error of shaft and hole parts based on improved grey wolf optimization algorithm, Measurement, 188 (2022), 110396. https://doi.org/10.1016/j.measurement.2021.110396 doi: 10.1016/j.measurement.2021.110396
![]() |
[12] |
H. Abaeifar, H. Barati, A. R. Tavakoli, Inertia-weight local-search-based TLBO algorithm for energy management in isolated micro-grids with renewable resources, Int. J. Electr. Power Energy Syst., 137 (2022), 107877. https://doi.org/10.1016/j.ijepes.2021.107877 doi: 10.1016/j.ijepes.2021.107877
![]() |
[13] |
V. K. Jadoun, G. R. Prashanth, S. S. Joshi, K. Narayanan, H. Malik, F. García Márquez, Optimal fuzzy based economic emission dispatch of combined heat and power units using dynamically controlled Whale Optimization Algorithm, Appl. Energy, 315 (2022), 119033. https://doi.org/10.1016/j.apenergy.2022.119033 doi: 10.1016/j.apenergy.2022.119033
![]() |
[14] |
M. Al-qaness, A. Ewees, H. Fan, L. Abualigah, M. Elaziz, Boosted ANFIS model using augmented marine predator algorithm with mutation operators for wind power forecasting, Appl. Energy, 314 (2022), 118851. https://doi.org/10.1016/j.apenergy.2022.118851 doi: 10.1016/j.apenergy.2022.118851
![]() |
[15] |
M. Al-qaness, A. Ewees, H. Fan, A. Airassas, M. Elaziz, Modified aquila optimizer for forecasting oil production, Geo-spatial Inf. Sci., 2022 (2022), 1–17. https://doi.org/10.1080/10095020.2022.2068385 doi: 10.1080/10095020.2022.2068385
![]() |
[16] |
A. Dahou, M. Al-qaness, M. Elaziz, A. Helmi, Human activity recognition in IoHT applications using Arithmetic Optimization Algorithm and deep learning, Measurement, 199 (2022), 111445. https://doi.org/10.1016/j.measurement.2022.111445 doi: 10.1016/j.measurement.2022.111445
![]() |
[17] |
M. Elaziz, A. Ewees, M. Al-qaness, L. Abualigah, R. Ibrahim, Sine–Cosine-Barnacles Algorithm Optimizer with disruption operator for global optimization and automatic data clustering, Expert Syst. Appl., 207 (2022), 117993. https://doi.org/10.1016/j.eswa.2022.117993 doi: 10.1016/j.eswa.2022.117993
![]() |
[18] |
X. Chen, X. Qi, Z. Wang, C. Cui, B. Wu, Y. Yang, Fault diagnosis of rolling bearing using marine predators algorithm-based support vector machine and topology learning and out-of-sample embedding, Measurement, 176 (2021), 109116. https://doi.org/10.1016/j.measurement.2021.109116 doi: 10.1016/j.measurement.2021.109116
![]() |
[19] |
P. H. Dinh, A novel approach based on three-scale image decomposition and marine predators algorithm for multi-modal medical image fusion, Biomed. Signal Process. Control, 67 (2021), 102536. https://doi.org/10.1016/j.bspc.2021.102536 doi: 10.1016/j.bspc.2021.102536
![]() |
[20] |
M. A. Sobhy, A. Y. Abdelaziz, H. M. Hasanien, M. Ezzat, Marine predators algorithm for load frequency control of modern interconnected power systems including renewable energy sources and energy storage units, Ain Shams Eng. J., 12 (2021), 3843–3857. https://doi.org/10.1016/j.asej.2021.04.031 doi: 10.1016/j.asej.2021.04.031
![]() |
[21] |
A. Faramarzi, M. Heidarinejad, S. Mirjalili, A. Gandomi, Marine predators algorithm: a nature-inspired metaheuristic, Expert Syst. Appl., 152 (2020), 113377. https://doi.org/10.1016/j.eswa.2020.113377 doi: 10.1016/j.eswa.2020.113377
![]() |
[22] |
M. A. Elaziz, D. Mohammadi, D. Oliva, K. Salimifard, Quantum marine predators algorithm for addressing multilevel image segmentation, Appl. Soft Comput., 110 (2021), 107598. https://doi.org/10.1016/j.asoc.2021.107598 doi: 10.1016/j.asoc.2021.107598
![]() |
[23] |
M. Ramezani, D. Bahmanyar, N. Razmjooy, A new improved model of marine predator algorithm for optimization problems, Arabian J. Sci. Eng., 46 (2021), 8803–8826. https://doi.org/10.1007/s13369-021-05688-3 doi: 10.1007/s13369-021-05688-3
![]() |
[24] |
M. Abdel-Basset, D. El-Shahat, R. K. Chakrabortty, M. Ryan, Parameter estimation of photovoltaic models using an improved marine predators algorithm, Energy Convers. Manage., 227 (2021), 113491. https://doi.org/10.1016/j.enconman.2020.113491 doi: 10.1016/j.enconman.2020.113491
![]() |
[25] |
K. Zhong, G. Zhou, W. Deng, Y. Zhou, Q. Luo, MOMPA: multi-objective marine predator algorithm, Comput. Methods Appl. Mech. Eng., 385 (2021), 114029. https://doi.org/10.1016/j.cma.2021.114029 doi: 10.1016/j.cma.2021.114029
![]() |
[26] |
R. Sowmya, V. Sankaranarayanan, Optimal vehicle-to-grid and grid-to-vehicle scheduling strategy with uncertainty management using improved marine predator algorithm, Comput. Electr. Eng., 100 (2022), 107949. https://doi.org/10.1016/j.compeleceng.2022.107949 doi: 10.1016/j.compeleceng.2022.107949
![]() |
[27] |
E. H. Houssein, I. E. Ibrahim, M. Kharrich, S. Kamel, An improved marine predators algorithm for the optimal design of hybrid renewable energy systems, Eng. Appl. Artif. Intell., 110 (2022), 104722. https://doi.org/10.1016/j.engappai.2022.104722 doi: 10.1016/j.engappai.2022.104722
![]() |
[28] |
D. Yousri, A. Ousama, Y. Shaker, A. Fathy, T. Babu, H. Rezk, et al., Managing the exchange of energy between microgrid elements based on multi-objective enhanced marine predators algorithm, Alexandria Eng. J., 61 (2022), 8487–8505. https://doi.org/10.1016/j.aej.2022.02.008 doi: 10.1016/j.aej.2022.02.008
![]() |
[29] |
Y. Ma, X. Zhang, J. Song, L. Chen, A modified teaching–learning-based optimization algorithm for solving optimization problem, Knowledge-Based Syst., 212 (2020), 106599. https://doi.org/10.1016/j.knosys.2020.106599 doi: 10.1016/j.knosys.2020.106599
![]() |
[30] |
N. E. Humphries, N. Queiroz, J. Dyer, N. Pade, M. Musyl, K. Schaefer, et al., Environmental context explains Lévy and Brownian movement patterns of marine predators, Nature, 465 (2010), 1066–1069. https://doi.org/10.1038/nature09116 doi: 10.1038/nature09116
![]() |
[31] |
D. W. Sims, E. J. Southall, N. E. Humphries, G. Hays, C. Bradshaw, J. Pitchford, et al., Scaling laws of marine predator search behaviour, Nature, 451 (2008), 1098–1102. https://doi.org/10.1038/nature06518 doi: 10.1038/nature06518
![]() |
[32] |
G. M. Viswanathan, E. P. Raposo, M. Luz, Lévy flights and superdiffusion in the context of biological encounters and random searches, Phys. Life Rev., 5 (2008), 133–150. https://doi.org/10.1016/j.plrev.2008.03.002 doi: 10.1016/j.plrev.2008.03.002
![]() |
[33] |
F. Bartumeus, J. Catalan, U. L. Fulco, M. Lyra, G. Viswanathan, Optimizing the encounter rate in biological interactions: Lévy versus Brownian strategies, Phys. Rev. Lett., 88 (2002), 097901. https://doi.org/10.1103/PhysRevLett.88.097901 doi: 10.1103/PhysRevLett.88.097901
![]() |
[34] |
A. Einstein, Investigations on the theory of the brownian movement, DOVER, 35 (1956), 318–320. https://doi.org/10.2307/2298685 doi: 10.2307/2298685
![]() |
[35] |
J. D. Filmalter, L. Dagorn, P. D. Cowley, M. Taquet, First descriptions of the behavior of silky sharks, Carcharhinus falciformis, around drifting fish aggregating devices in the Indian Ocean, Bull. Mar. Sci., 87 (2011), 325–337. https://doi.org/10.5343/bms.2010.1057 doi: 10.5343/bms.2010.1057
![]() |
[36] |
D. Yousri, H. M. Hasanien, A. Fathy, Parameters identification of solid oxide fuel cell for static and dynamic simulation using comprehensive learning dynamic multi-swarm marine predators algorithm, Energy Convers. Manage., 228 (2021), 113692. https://doi.org/10.1016/j.enconman.2020.113692 doi: 10.1016/j.enconman.2020.113692
![]() |
[37] |
M. Abdel-Basset, R. Mohamed, S. Mirjalili, R. Chakrabortty, M. Ryan, An efficient marine predators algorithm for solving multi-objective optimization problems: analysis and validations, IEEE Access, 9 (2021), 42817–42844. https://doi.org/10.1109/ACCESS.2021.3066323 doi: 10.1109/ACCESS.2021.3066323
![]() |
[38] |
T. Niknam, R. Azizipanah-Abarghooee, M. R. Narimani, A new multi objective optimization approach based on TLBO for location of automatic voltage regulators in distribution systems, Eng. Appl. Artif. Intell., 25 (2012), 1577–1588. https://doi.org/10.1016/j.engappai.2012.07.004 doi: 10.1016/j.engappai.2012.07.004
![]() |
[39] |
T. Niknam, F. Golestaneh, M. S. Sadeghi, θ-Multiobjective teaching–learning-based optimization for dynamic economic emission dispatch, IEEE Syst. J., 6 (2012), 341–352. https://doi.org/10.1109/JSYST.2012.2183276 doi: 10.1109/JSYST.2012.2183276
![]() |
[40] |
R. V. Rao, V. Patel, An improved teaching-learning-based optimization algorithm for solving unconstrained optimization problems, Sci. Iran., 20 (2013), 710–720. https://doi.org/10.1016/j.scient.2012.12.005 doi: 10.1016/j.scient.2012.12.005
![]() |
[41] |
P. K. Roy, S. Bhui, Multi-objective quasi-oppositional teaching learning based optimization for economic emission load dispatch problem, Int. J. Electr.Power Energy Syst., 53 (2013), 937–948. https://doi.org/10.1016/j.ijepes.2013.06.015 doi: 10.1016/j.ijepes.2013.06.015
![]() |
[42] |
H. Bouchekara, M. A. Abido, M. Boucherma, Optimal power flow using teaching-learning-based optimization technique, Electr. Power Syst. Res., 114 (2014), 49–59. https://doi.org/10.1016/j.epsr.2014.03.032 doi: 10.1016/j.epsr.2014.03.032
![]() |
[43] |
M. Liu, X. Yao, Y. Li, Hybrid whale optimization algorithm enhanced with Lévy flight and differential evolution for job shop scheduling problems, Appl. Soft Comput., 87 (2020), 105954. https://doi.org/10.1016/j.asoc.2019.105954 doi: 10.1016/j.asoc.2019.105954
![]() |
[44] |
D. Tansui, A. Thammano, Hybrid nature-inspired optimization algorithm: hydrozoan and sea turtle foraging algorithms for solving continuous optimization problems, IEEE Access, 8 (2020), 65780–65800. https://doi.org/10.1109/ACCESS.2020.2984023 doi: 10.1109/ACCESS.2020.2984023
![]() |
[45] |
K. Zhong, Q. Luo, Y. Zhou, M. Jiang, TLMPA: teaching-learning-based marine predators algorithm, AIMS Math., 6 (2021), 1395–1442. https://doi.org/10.3934/math.2021087 doi: 10.3934/math.2021087
![]() |
State variable | Description |
S1(t) | Number of susceptible males. |
I1(t) | Number of infected males. |
S2(t) | Number of susceptible females. |
I2(t) | Number of infected females. |
N1(t) | Total male population size. |
N2(t) | Total female population size. |
N(t)=N1(t)+N2(t) | Total population size. |
Parameters | Description | Value | Dim. | Ref. |
Λ | The maximum number of immigrating individuals per unit time. | 11,000 | individuals/week | Assumed |
K | Some kind of carrying capacity. | 35×106 | individuals | Assumed |
μ | Per-capita death rate. | 1/(70×52) | Week−1 | [11] |
ˉβ | Per-capita contact rate between individuals. | – | Week−1 | – |
β | Per-capita effective contact rate at which susceptible males acquire influenza. | Computed to adapt with the value of R0 | Week−1 | Assumed |
γ=γ1 | Per-capita removal (by recovery or disease-induced death) rate for infected males. | 7/3.38 | Week−1 | [4,19] |
γ2=aγ | Per-capita removal (by recovery or disease-induced death) rate for infected females. | – | Week−1 | – |
R0 | The basic reproduction number for model. | 1.525 | – | [4,19] |
a | A rescaling parameter accounting for the relative removability (due to recovery or disease-induced death) of infected females with respect to infected males. | 1.1 | – | Assumed |
q0 | The proportion of female new births. | 0.48∈[0.45,0.55] | – | Assumed |
q2 | The proportion of female immigrants. | 0.45∈[0.45,0.55] | – | Assumed |
g1 | The susceptibility of males. | – | – | – |
g2 | The susceptibility of females. | – | – | – |
g=g2/g1 | Relative susceptibility of females with respect to males. | ∈(0,2) | – | Assumed |
r | Relative transmissibility of the infection by females with respect to males. | ∈(0,2) | – | Assumed |
c1 | Infection-case fatality among males. | 0.007∈[0,0.1] | – | Assumed |
c2 | Infection-case fatality among females. | 0.005∈[0,0.1] | – | Assumed |
Notes: Dim. = Dimension, Ref. = References. |
State variable | Description |
S1(t) | Number of susceptible males. |
I1(t) | Number of infected males. |
S2(t) | Number of susceptible females. |
I2(t) | Number of infected females. |
N1(t) | Total male population size. |
N2(t) | Total female population size. |
N(t)=N1(t)+N2(t) | Total population size. |
Parameters | Description | Value | Dim. | Ref. |
Λ | The maximum number of immigrating individuals per unit time. | 11,000 | individuals/week | Assumed |
K | Some kind of carrying capacity. | 35×106 | individuals | Assumed |
μ | Per-capita death rate. | 1/(70×52) | Week−1 | [11] |
ˉβ | Per-capita contact rate between individuals. | – | Week−1 | – |
β | Per-capita effective contact rate at which susceptible males acquire influenza. | Computed to adapt with the value of R0 | Week−1 | Assumed |
γ=γ1 | Per-capita removal (by recovery or disease-induced death) rate for infected males. | 7/3.38 | Week−1 | [4,19] |
γ2=aγ | Per-capita removal (by recovery or disease-induced death) rate for infected females. | – | Week−1 | – |
R0 | The basic reproduction number for model. | 1.525 | – | [4,19] |
a | A rescaling parameter accounting for the relative removability (due to recovery or disease-induced death) of infected females with respect to infected males. | 1.1 | – | Assumed |
q0 | The proportion of female new births. | 0.48∈[0.45,0.55] | – | Assumed |
q2 | The proportion of female immigrants. | 0.45∈[0.45,0.55] | – | Assumed |
g1 | The susceptibility of males. | – | – | – |
g2 | The susceptibility of females. | – | – | – |
g=g2/g1 | Relative susceptibility of females with respect to males. | ∈(0,2) | – | Assumed |
r | Relative transmissibility of the infection by females with respect to males. | ∈(0,2) | – | Assumed |
c1 | Infection-case fatality among males. | 0.007∈[0,0.1] | – | Assumed |
c2 | Infection-case fatality among females. | 0.005∈[0,0.1] | – | Assumed |
Notes: Dim. = Dimension, Ref. = References. |