
This research concerned with a new formulation of a spatial predator-prey model with Leslie-Gower and Holling type II schemes in the presence of prey social behavior. The aim interest here is to distinguish the influence of Leslie-Gower term on the spatiotemporal behavior of the model. Interesting results are obtained as Hopf bifurcation, Turing bifurcation and Turing-Hopf bifurcation. A rigorous mathematical analysis shows that the presence of Leslie-Gower can induce Turing pattern, which shows that this kind of interaction is very important in modeling different natural phenomena. The direction of Turing-Hopf bifurcation is studied with the help of the normal form. The obtained results are tested numerically.
Citation: Fethi Souna, Salih Djilali, Sultan Alyobi, Anwar Zeb, Nadia Gul, Suliman Alsaeed, Kottakkaran Sooppy Nisar. Spatiotemporal dynamics of a diffusive predator-prey system incorporating social behavior[J]. AIMS Mathematics, 2023, 8(7): 15723-15748. doi: 10.3934/math.2023803
[1] | Yanting Xiao, Wanying Dong . Robust estimation for varying-coefficient partially linear measurement error model with auxiliary instrumental variables. AIMS Mathematics, 2023, 8(8): 18373-18391. doi: 10.3934/math.2023934 |
[2] | Remigijus Leipus, Jonas Šiaulys, Dimitrios Konstantinides . Minimum of heavy-tailed random variables is not heavy tailed. AIMS Mathematics, 2023, 8(6): 13066-13072. doi: 10.3934/math.2023658 |
[3] | Alanazi Talal Abdulrahman, Khudhayr A. Rashedi, Tariq S. Alshammari, Eslam Hussam, Amirah Saeed Alharthi, Ramlah H Albayyat . A new extension of the Rayleigh distribution: Methodology, classical, and Bayes estimation, with application to industrial data. AIMS Mathematics, 2025, 10(2): 3710-3733. doi: 10.3934/math.2025172 |
[4] | Huimin Li, Jinru Wang . Pilot estimators for a kind of sparse covariance matrices with incomplete heavy-tailed data. AIMS Mathematics, 2023, 8(9): 21439-21462. doi: 10.3934/math.20231092 |
[5] | Bin Yang, Min Chen, Jianjun Zhou . Forecasting the monthly retail sales of electricity based on the semi-functional linear model with autoregressive errors. AIMS Mathematics, 2025, 10(1): 1602-1627. doi: 10.3934/math.2025074 |
[6] | Chen Chen, Xiangbing Chen, Yi Ai . Convex-structured covariance estimation via the entropy loss under the majorization-minimization algorithm framework. AIMS Mathematics, 2024, 9(6): 14253-14273. doi: 10.3934/math.2024692 |
[7] | Naif Alotaibi, A. S. Al-Moisheer, Ibrahim Elbatal, Salem A. Alyami, Ahmed M. Gemeay, Ehab M. Almetwally . Bivariate step-stress accelerated life test for a new three-parameter model under progressive censored schemes with application in medical. AIMS Mathematics, 2024, 9(2): 3521-3558. doi: 10.3934/math.2024173 |
[8] | Khaled M. Alqahtani, Mahmoud El-Morshedy, Hend S. Shahen, Mohamed S. Eliwa . A discrete extension of the Burr-Hatke distribution: Generalized hypergeometric functions, different inference techniques, simulation ranking with modeling and analysis of sustainable count data. AIMS Mathematics, 2024, 9(4): 9394-9418. doi: 10.3934/math.2024458 |
[9] | Zhanshou Chen, Muci Peng, Li Xi . A new procedure for unit root to long-memory process change-point monitoring. AIMS Mathematics, 2022, 7(4): 6467-6477. doi: 10.3934/math.2022360 |
[10] | Walid Emam . Benefiting from statistical modeling in the analysis of current health expenditure to gross domestic product. AIMS Mathematics, 2023, 8(5): 12398-12421. doi: 10.3934/math.2023623 |
This research concerned with a new formulation of a spatial predator-prey model with Leslie-Gower and Holling type II schemes in the presence of prey social behavior. The aim interest here is to distinguish the influence of Leslie-Gower term on the spatiotemporal behavior of the model. Interesting results are obtained as Hopf bifurcation, Turing bifurcation and Turing-Hopf bifurcation. A rigorous mathematical analysis shows that the presence of Leslie-Gower can induce Turing pattern, which shows that this kind of interaction is very important in modeling different natural phenomena. The direction of Turing-Hopf bifurcation is studied with the help of the normal form. The obtained results are tested numerically.
Coronavirus disease (COVID-19) pandemic is considered the biggest global threat worldwide because of thousands of confirmed infections, accompanied by thousands of deaths [1]. COVID-19 was identified and named by the World Health Organisation (WHO) on January 10, 2020 following an earlier viral infection episode in Wuhan, China in December 2019, and was declared by the WHO to be a public health emergency of international concern in 2020 [2,3]. COVID-19 by its nature is a very contagious disease that spreads easily from person-to-person through direct contact with objects or surfaces that are contaminated with the virus. Moreover, inhalation of respiratory droplets from both asymptomatic and symptomatic infectious individuals cause transmission [6]. The symptoms of COVID-19 appear 2-14 days after exposure and may include fever, dry cough, muscle pain, fatigue, and shortness of breath [7]. The symptoms are mild in 85% of cases, and vary from severe in 10% to critical in 5% of those infected, but a larger proportion of infected individuals exhibit mild or no symptoms [3]. The severity and progression of COVID-19 are known to be exacerbated by the presence of co-morbidities such as diabetes, hypertension and cardio/cerebrovascular diseases [8]. It has also been observed that COVID-19 mortality risk is highly concentrated within the elderly population [9].
Available scientific evidence classify COVID-19 infected individuals into three broad categories; individuals who manifest severe symptoms, individuals who manifest mild symptoms and individuals who do not manifest any COVID-19 symptoms (asymptomatic) and yet remain infectious undetected. The non-manifestation of COVID-19 symptoms in some infected people complicates the epidemiology of the COVID-19 pandemic. Firstly, asymptomatic individuals are unlikely to seek medical care or self-quarantine given that they cannot tell whether they have the disease unless detected through testing or contact tracing. Secondly, they will continue interacting with healthy people thereby, spreading the virus. Although the asymptomatic categories form a large proportion of COVID-19 infections, it is not yet known to what extent they spread the virus relative to categories with severe symptoms which constitute a small proportion of COVID-19 infections. The size of this delay may play an important role in minimizing the spread of the disease in the community. It is therefore essential to gain a better and more comprehensive understanding of the effects of time delay on COVID-19 transmission and control.
Mathematical models have proved to be essential guiding tools for epidemiologists, biologists as well as policymakers. Models can provide solutions to phenomena which are difficult to measure practically. Recently, a number of mathematical models have been proposed to study the spread and control of COVID-19 (see, for example [1,2,6,7,8,9,19,20,21,22,23,24], and references therein) have certainly produced many useful results and improved the existing knowledge on COVID-19 dynamics. In [4] a discrete fractional Susceptible-Infected-Treatment-Recovered-Susceptible (SITRS) model for simulating the coronavirus (COVID-19) pandemic was proposed by taking into account the possibility that people who have been infected before can lose their temporary immunity and get reinfected. In [5] a novel reaction-diffusion coronavirus (COVID- 19) model was employed to investigate the effect of random movements of individuals from different compartments in their environments. A limitation of these studies, however, is the non-inclusion of the time taken before an infectious human is detected and quarantined, despite the fact that in many countries where the disease is endemic, lack of financial and human resources often results on delay in detection and quarantining of infectious individuals. In addition epidemic models with time delay often exhibit periodic solutions and as a consequence understanding the nature of these periodic outbreaks plays a crucial role in designing policies that can successfully control the disease (see, for example [10]). In [11] a mathematical model with time delay was proposed to describe the outbreak of 2019-nCoV in China to show that the novel dynamic system can well predict the outbreak trend of the disease. In [18] Pei and Zhang constructed a SIRD epidemic model (S-Susceptible, I-Infected, R-Recovered, D-Dead) which is a non-autonomous dynamic system with an incubation time delay to study the evolution of the COVID-19 in Wuhan City, Hubei Province and China Mainland. In [25] a system of ordinary differential equations with delays was utilized to describe the evolution of the COVID-19 pandemic.
Although this study is not the first to incorporate discrete delay in studying COVID-19 transmission, the main goal of this article is to explore the dynamics and stability analysis of a COVID-19 model with discrete delay. Hence, we formulated a mathematical model that incorporates a discrete delay that represents the incubation period. In addition, we investigate the impact of the time taken to detect and quarantine infectious individuals on the disease dynamics. The rest of the paper is organised as follows. In Section 2, the propose model present the analytical results. In Section 3, numerical simulations are done to verify the theoretical results presented in the study. Finally, a concluding remark rounds up the paper.
The COVID-19 pandemic remains a major global threat worldwide. This is mainly attributed to several challenges associated with effective control which range from the inadequate use of control measures such as wearing of disposable surgical face masks, regular hand-washing with plenty of soap under running water, the use of alcohol-based hand sanitizers in the absence of soap and water, vaccines among others as recommended by the WHO [16,17]. Although these interventions have succeeded greatly in many countries the major problem in the spread of COVID-19 is human-to-human transmission in a heterogeneous community. The implementation of interventional strategies such as quarantine/isolation during infection remains a big challenge in the fight against the disease because of hunger, poverty, and poor health facilities, especially in developing countries in sub-Saharan Africa where governments lack social securities. Furthermore, these challenges often lead to delay in detection and quarantine/isolation of infectious individuals.
In this article, we propose a model to analyze the impact of delay in treatment and the time needed to detect/diagnose and quarantine individuals infected with COVID-19. Assume that the infected individuals are in two different categories, those with mild symptoms and those with severe illness. Anyone can have mild to severe symptoms. We subdivide the total population, N(t), at time t, into six compartments namely; susceptible individuals S(t), asymptomatic (undetected infectious) individuals I(t), infectious detected and quarantined Q(t) and recovered individuals R(t). The recovered population R(t) is made up of individuals who have successfully recovered from the infection either naturally or through various health support mechanisms (since the disease has no treatment). The two additional compartments Hm(t) and Hs(t) represent the symptomatic (hospitalized) individuals who develop symptoms. The distinction between the two categories of hospitalized individuals represent that Hm are hospitalized individuals who develop mild symptoms while Hs(t) represent hospitalized individuals who develop severe symptoms. The human population at any given time t, is given by N(t)=S(t)+I(t)+Q(t)+Hm(t)+Hs(t)+R(t). The proposed COVID-19 model with a time delay factor is given by:
{dS(t)dt=Λ−β1I(t)S(t)−β2Hm(t)S(t)−β2Hs(t)S(t)−δS(t),dI(t)dt=β1I(t−τ)S(t−τ)+β2Hm(t−τ)S(t−τ)+β2Hs(t−τ)S(t−τ)−(α+δ+σ1)I(t),dQ(t)dt=αI(t)−(γ+δ+σ1)Q(t),dHm(t)dt=(1−p)γQ(t)−(σ2+μ+δ)Hm(t),dHs(t)dt=pγQ(t)−(σ2+μ+δ)Hs(t),dR(t)dt=σ1(I(t)+Q(t))+σ2(Hm(t)+Hs(t))−δR(t), | (2.1) |
where Λ is the recruitment rate, δ denotes natural mortality rate, β1 denotes the contact rate of asymptomatic (undetected infectious) and susceptible humans, β2 denotes the contact rate of symptomatic (hospitalized) and susceptible humans, α is the detection rate of asymptomatic (undetected infectious) patients, σ1 denotes the recovery rate of asymptomatic (undetected infectious) and quarantined individuals, σ2 is the recovery rate of hospitalized individuals, and μ represents the disease-induced death rate of symptomatic humans. Proportions 0<p<1 account for hospitalized individuals with severe symptoms, while the remainder (1−p) accounts for hospitalized individuals with mild symptoms. The quarantined individuals are hospitalized after 1/γ days. τ is a discrete time delay representing the latent period. The model diagram is depicted in Figure 1.
The initial conditions for the model (2.1) are given as follows:
S(θ)=φ1(θ),I(θ)=φ2(θ),Q(θ)=φ3(θ),Hm(θ)=φ4(θ),Hs(θ)=φ5(θ),R(θ)=φ6(θ),θ∈[−τ,0],τ>0,φ=(φ1,φ2,φ3,φ4,φ5,φ6)∈C+⊂C. | (2.2) |
Here, C is the Banach space C([−τ,0],R6) of continuous functions mapping the interval [−τ,0] into R6 with the sup-norm ‖φ‖=supθ∈[−τ,0]‖φi‖, for i=1,2,3,4,5,6 for φ∈C. The non-negative cone of C is defined as C+=C([−τ,0],R6).
Equation (2.3) in theorem (2.1) shows that the model formulated in this study is biologically meaningful. Precisely, the theorem demonstrates that for non-negative initial conditions, the solutions of the proposed model are non-negative and bounded for all t>0.
Theorem 2.1. There exists a unique solution for the COVID-19 model (2.1). Furthermore, the solution is non-negative for all t>0 and lies in the set:
Γ={(S,I,Q,Hm,Hs,R)∈R6+:S+I+Q+Hm+Hs+R≤Λδ} | (2.3) |
Proof. To prove the positivity of the model system (2.1), we investigate the direction of the vector field given by the right-hand side of the model system (2.1) on each space and note whether the vector field points to the interior of R6+ or is tangent to the coordinate space. We observe that:
{(S′)S=0=Λ≥0,(I′)I=0=β2Hm(t−τ)S(t−τ)+β2Hs(t−τ)S(t−τ)≥0,(Q′)Q=0=αI(t)≥0,(H′m)Hm=0=(1−p)γQ(t)≥0,(H′s)Hs=0=pγQ(t)≥0,(R′)R=0=σ1(I(t)+Q(t))+σ2(Hm(t)+Hs(t))≥0. | (2.4) |
It follows that the vector field given by the right-hand side of the model system (2.1) on each coordinate plane is either tangent to the coordinate plane or points to the interior of R6+. Hence, the positivity of the solutions starting in the interior of R6+ is assured. R6+ is a positively invariant set of the SIQHmHsR model system (2.1). Moreover, if the initial conditions φi≥ 0, (i = 1, 2, 3, 4, 5, 6.) are, therefore, the corresponding solutions of the model system (2.1).
Theorem 2.2. The solutions of the SIQHmHsR system (2.1) with the initial conditions of (2.2) are uniformly bounded in the region Γ.
Proof. To prove the boundedness of the model system (2.1), we add all model equations, which gives:
N′(t)=Λ−δN(t)−μ(Hm(t)+Hs(t))−β1I(t)S(t)−β2Hm(t)S(t)−β2Hs(t)S(t)+β1I(t−τ)S(t−τ)+β2Hm(t−τ)S(t−τ)+β2Hs(t−τ)S(t−τ)=Λ−δN(t)−μ(Hm(t)+Hs(t))−∫tt−τddξ{β1I(ξ)S(ξ)+β2Hm(ξ)S(ξ)+β2Hs(ξ)S(ξ)}dξ≤Λ−δN(t). | (2.5) |
Since N′(t)≤Λ−δN(t) for 0≤τ<t, it follows by applying the standard comparison Theorem in [27] that N(t)≤N(0)e−δt+(Λδ)(1−e−δt). In particular, we have N(t)≤(Λδ) if N(0)≤(Λδ). Therefore, we conclude that the population is bounded. Hence, all solutions in R6+ eventually enter Γ.
Since the last equation in the model system (2.1) is independent of the other equations, system (2.1) may be reduced to the following system:
dS(t)dt=Λ−β1I(t)S(t)−β2Hm(t)S(t)−β2Hs(t)S(t)−δS(t), | (2.6) |
dI(t)dt=β1I(t−τ)S(t−τ)+β2Hm(t−τ)S(t−τ)+β2Hs(t−τ)S(t−τ)−m1I(t), | (2.7) |
dQ(t)dt=αI(t)−m2Q(t), | (2.8) |
dHm(t)dt=(1−p)γQ(t)−m3Hm(t), | (2.9) |
dHs(t)dt=pγQ(t)−m3Hs(t), | (2.10) |
where m1=(α+δ+σ1), m2=(γ+δ+σ1) and m3=(σ2+μ+δ).
It can easily be verified that in the absence of the disease in the community, system (2.6)-(2.10) admit a disease-free equilibrium given by E0=(S0=Λδ,0,0,0,0). In addition, from equations (2.8), (2.9) and (2.10) we have:
Q∗=αI∗m2,H∗m=(1−p)γαI∗m2m3,andH∗s=pγαI∗m2m3. | (2.11) |
Substituting these results into equation (2.7) gives:
β1I∗S∗+β2(1−p)γαI∗S∗m2m3+β2pγαI∗S∗m2m3−m1I∗=0. | (2.12) |
From (2.12) we have:
(β1S∗m1+β2(1−p)γαS∗m1m2m3+β2pγαS∗m1m2m3−1)m1I∗=0. | (2.13) |
It follows that:
I∗=0,orS∗=m1m2m3β2m2m3+β2(1−p)γα+β2pγα. | (2.14) |
Substituting the value of S∗ into equation (2.6) gives:
I∗=m2m3δβ1m2m3+β2(1−p)αγ+β2pαγ×(β1Λm1δ+β2(1−p)αγΛm1m2m3δ+β2pαγΛm1m2m3δ−1). | (2.15) |
Substituting (2.15) into (2.11) yields:
Q∗=αm3δβ1m2m3+β2(1−p)αγ+β2pαγ×(β1Λm1δ+β2(1−p)αγΛm1m2m3δ+β2pαγΛm1m2m3δ−1), |
H∗m=(1−p)αγδβ1m2m3+β2(1−p)αγ+β2pαγ×(β1Λm1δ+β2(1−p)αγΛm1m2m3δ+β2pαγΛm1m2m3δ−1), |
H∗s=pαγδβ1m2m3+β2(1−p)αγ+β2pαγ×(β1Λm1δ+β2(1−p)αγΛm1m2m3δ+β2pαγΛm1m2m3δ−1). | (2.16) |
From the computations in equation (2.16), we observe that I∗, Q∗, H∗M and H∗S makes biological sense whenever:
β1Λm1δ+β2(1−p)αγΛm1m2m3δ+β2pαγΛm1m2m3δ>1. | (2.17) |
Therefore, if we let the basic reproduction number of model (2.6)-(2.10) be:
R0=β1Λm1δ+β2(1−p)αγΛm1m2m3δ+β2pαγΛm1m2m3δ. | (2.18) |
It follows that models (2.6)-(2.10) has a second equilibrium E∗(S∗,I∗,Q∗,H∗m,H∗s) point known as the endemic equilibrium which exists whenever R0>1.
Biologically, the basic reproduction number R0 represents the average number of new or secondary COVID-19 infections caused by the introduction of an infectious individual into a totally susceptible population. In fact,
● the term β1Λm1δ is the average number of secondary infections generated as result of contact between susceptible individuals and one asymptomatic (Undetected infectious) COVID-19 patient,
● the term β2(1−p)αγΛm1m2m3δ represents the average number of new COVID-19 cases generated when susceptible individuals come into contact with a hospitalized patient of class Hm,
● the term β2pαγΛm1m2m3δ gives the average number of secondary COVID-19 infections which occur in the community when susceptible individuals come into contact with a hospitalized patient of class Hs.
Then we have the following results:
Theorem 2.3. If R0≤1, then the disease-free equilibrium E0 is globally asymptotically stable.
The detailed proof process can be obtained in Appendix A.
Next, we investigate the global stability of the endemic equilibrium point E∗ of models (2.6)-(2.10) when R0>1.
Theorem 2.4. If R0>1, then model (2.6)-(2.10) has a globally asymptotically stable endemic equilibrium point.
The proofs of Theorem 2.4 is given in Appendix B.
In this section, we perform numerical analysis to explore the behavior of the model system (2.1) and illustrate the stability of the equilibria solutions. We numerically solve the model system (2.1) using dde23 [14] based on Runge-Kutta methods through MATLAB software and parameters values adopted from Table 1, and the initial population levels were assumed as follows: S(0)=10, and I(0)=Q(0)=Hm(0)=Hs(0)=2.
Symbol | Units | Value | Source |
Λ | day −1 | 0.0000433 | [13] |
β1 | day −1 | 0.124 | [13] |
β2 | day −1 | 0.05 | [13] |
δ | day −1 | 0.0000357 | [13] |
α | day −1 | Vary | Assumed |
μ | day −1 | 0.043 | [15] |
σ1 | day −1 | 0.854 | [13] |
σ2 | day −1 | 0.0987 | [13] |
γ | day−1 | Vary | Assumed |
p | unit-less | Vary | Assumed |
The numerical results in Figures 2–4 illustrate the dynamical solutions of the model system (2.1) for different values of τ at the endemic equilibrium point of R0=3.7095. The results were obtained using the parameter values in Table 1, with δ=0.00000357, α=0.001, σ1=0.00124, σ2=0.0182, γ=0.001, and p=5.4×10−10 coupled with delay values τ=5, τ=15, and τ=25 for Figures 2, 3, and 4 respectively. To improve the clarity of the results, the solution for all populations were zoomed in. In all cases for certain parameter values and initial population levels, the model system (2.1) exhibits some periodic oscillation. Precisely, we note that the infected population of class I(t) and Q(t) oscillates with a reduced amplitude from the start for a considerable time frame, thereafter the oscillations dies off and converges to the endemic equilibrium point. Similar patterns are observed for compartment S(t) in Figure 2(a), 3(a), and 3(a). We can also note that the intensity and amplitude of oscillation at τ=15 are high compared to that at τ=5 and 25. In addition, the implication of these results is that the inclusion of the time delay factor destabilizes the endemic equilibrium point for a certain period of time, leading to periodic oscillations which arise due to the existence of Hopf bifurcations. These results agree with the analytical analysis of the global stability for the endemic equilibrium point in Theorem 2.4.
The results in Figure 5 demonstrate the existence of Hopf bifurcations that arise due to the inclusion of the time delay factor τ=15 in the model system (2.1). The results were obtained by using parameter values in Table (1), with δ=0.00000357, α=0.001, σ1=0.00124, σ2=0.0182, γ=0.001, and p=5.4×10−10, leading to R0=3.7095. Overall, these results are in agreement to those depicted in Figures 2–4. Thus, the inclusion of the time delay factor leads to the existence of Hopf bifurcations.
The results in Figure 6 demonstrate the effects of σ1 (recovery rate of asymptomatic and quarantined individuals) and σ2 (recovery rate of hospitalized individuals) on the dynamics of the disease in the population. The parameters and initial values are fixed and provided in Table 1. Overall, we note that increasing the recovery rate of asymptomatic and quarantined individuals (modeled by parameter σ1) reduces the spread of the disease in the population. In particular, when σ1 is greater than 50%, the disease dies in the population and persists when less than 50%.
The results in Figures 7 and 8 demonstrate the solution profile for mild (Hm) and severe (Hs) hospitalized individuals for different values of τ at the endemic equilibrium point of R0=3.7095. The results were obtained using parameters values in Table 1, with δ=0.00000357, α=0.001, σ1=0.00124, σ2=0.0182, γ=0.001, and p=5.4×10−10 coupled with delay values τ=5, and τ=15 for Figures 7 and 8 respectively. Overall, for certain parameter values and initial population levels, the model system (2.1) does not exhibit periodic oscillation. Precisely, we note that the infected population of class Hm and Hs decrease gradually from the start for a considerable time frame, thereafter the solutions converge to the endemic equilibrium point. This has the implication that the inclusion of the time delay factor has less effects on hospitalized individuals. These results agree with the analytical analysis of global stability for the endemic equilibrium point in Theorem 2.4.
Figure 9 demonstrates the convergence of the solution profile of model system (2.1) to the disease-free-equilibrium for R0<1. The parameter values used in simulations are in Table (1), with β1=5.33×10−7, β2=0.089, α=0.0001, σ1=0.00124, σ2=0.0182, γ=0.01, and p=5.4×10−10, leading to R0=0.171. We observe that the variable for epidemiological classes S(t) and I(t) for t≤50 all solutions in Figures 9(a) and 9(b) respectively decrease at the beginning and finally attain stability to the disease-free-equilibrium point. In addition, the variable for quarantine individual in Figure 9(a) increase rapidly during the first 50 days, followed by a gradual decline and stability of solutions at the disease-free equilibrium point. In particular, the disease dies out in the population after 50 days which is in agreement with the analytical results summarized the Theorem 2.3. The results in Figure 10 demonstrate the existence of the bifurcations that arise due to the inclusion of the time delay factor τ=15 in the model system (2.1). The results were obtained by using parameter values in Table (1), with β1=5.33×10−7, β2=0.089, α=0.0001, σ1=0.00124, σ2=0.0182, γ=0.01, and p=5.4×10−10, leading to R0=0.171. We can note that the inclusion of the time delay factor leads to the existence of bifurcations in the model system.
In this article, we have developed and analyzed a mathematical model for COVID-19 that incorporates a discrete delay that accounts for the latent period. We compute the basic reproduction number and demonstrate that it is an important threshold quantity for the stability of equilibria. By constructing suitable Lyapunov functionals, it is shown that the model has a globally asymptotically stable infection-free equilibrium whenever the reproduction number is less than unity. Furthermore, whenever the reproduction number is greater than the unity then the model has a unique endemic equilibrium point which is globally asymptotically stable. Numerical simulations are carried out to illustrate the main results. Although quarantine/isolation of an asymptomatic individual is a relatively easy strategy to implement, some studies suggest that quarantine/isolation of asymptomatic, symptomatic and susceptible individuals maybe more effective (see for example [26]). The rationale being that by decreasing host density, the number of contacts per unit time between humans is low, thereby reducing disease transmission. In [26] it was demonstrated that quarantine/isolation of both asymptomatic and symptomatic individuals only can be effective whenever the number of infected hosts is above a certain critical level [26]. We expect to improve this study in the future by developing (COVID-19) model(s) with a time delay that will enable the comparison of the aforementioned aspects. In addition the bifurcation analysis of epidemic models with more compartments and parameters will be more complex and this is a major challenge for the future.
All authors are grateful to their respective institutions for their support during the preparation of the manuscript. Paride O. Lolika acknowledges the support from the University of Juba, South Sudan.
The authors declare that there are no conflicts of interest.
Proof of Theorem 2.3. We denote by xt the translation of the solution of models (2.6)-(2.10), that is:
xt=(S(t+θ),I(t+θ),Q(t+θ),Hm(t+θ),Hs(t+θ)), |
and consider the Lyapunov function:
V(t)=(β1m1+β2(1−p)αγm1m2m3+β2pαγm1m2m3)I(t)+(β2(1−p)γm2m3+β2pγm2m3)Q(t)+β2m3Hm(t)+β2m3Hs(t)+(β1m1+β2(1−p)αγm1m2m3+β2pαγm1m2m3)×∫tt−τ(β1I(θ)S(θ)+β2Hm(θ)S(θ)+β2Hs(θ)S(θ))dθ. | (4.1) |
Then, the time derivative of V(t) along solutions of models (2.6)-(2.10):
dVdt=(β1m1+β2(1−p)αγm1m2m3+β2pαγm1m2m3)×(β1I(t)+β2Hm(t)+β2Hs(t))S(t)−(β1I(t)+β2Hm(t)+β2Hs(t))=(β1m1+β2(1−p)αγm1m2m3+β2pαγm1m2m3)S(t)−1]×[β1I(t)+β2Hm(t)+β2Hs(t)]. | (4.2) |
Since S(t)≤S0(S0=Λδ) for t≥0, we have:
dVdt≤[(β1m1+β2(1−p)αγm1m2m3+β2pαγm1m2m3)S0−1]×[β1I(t)+β2Hm(t)+β2Hs(t)]=[R0−1][β1I(t)+β2Hm(t)+β2Hs(t)]. | (4.3) |
Therefore, ˙V(t)<0 holds if R0<1. Furthermore, ˙V(t)=0 if R0=1. Thus, the largest invariant set of ˙V(t) is a singleton such that S(t)=S0, I(t)=Q(t)=Hm(t)=Hs(t)=0. From the LaSalle invariance principle [12], the disease-free equilibrium of models (2.6)-(2.10) denoted by E0 is globally asymptotically stable whenever R0≤1. This completes the proof.
Proof of Theorem 2.4. Let us consider the Lyapunov function:
W(t)=W1(t)+W2(t). | (4.4) |
Here,
W1(t)={S(t)−S∗−S∗ln(S(t)S∗)}+{I(t)−I∗−I∗ln(I(t)I∗)}+(β2H∗m+β2H∗s)S∗αI∗×{Q(t)−Q∗−Q∗ln(Q(t)Q∗)}+β2H∗mS∗γ(1−p)Q∗×{Hm(t)−H∗m−H∗mln(Hm(t)H∗m)}+β2H∗sS∗γpQ∗{Hs(t)−H∗s−H∗sln(Hs(t)H∗s)}, | (4.5) |
W2(t)=βS∗I∗∫τ0{I(t−ξ)S(t−ξ)S∗I∗−1}dξ−βS∗I∗∫τ0{ln(I(t−ξ)S(t−ξ)S∗I∗)}dξ+β2S∗H∗m∫τ0{Hm(t−ξ)S(t−ξ)S∗H∗M−1}dξ−β2S∗H∗m∫τ0{ln(Hm(t−ξ)S(t−ξ)S∗H∗m)}dξ+β2H∗SS∗∫τ0{Hs(t−ξ)S(t−ξ)S∗H∗s−1}dξ−β2H∗SS∗∫τ0{ln(Hs(t−ξ)S(t−ξ)S∗H∗s)}dξ. | (4.6) |
The derivatives of W1(t) are given by:
dW1(t)dt=(1−S∗S)dSdt+(1−I∗I)dIdt+(β2H∗m+β2H∗s)S∗αI∗(1−Q∗Q)dQdt+β2H∗mS∗γ(1−p)Q∗(1−H∗mHm)dHmdt+β2H∗sS∗γpQ∗(1−H∗sHs)dHsdt. | (4.7) |
Substituting the appropriate differentials from (2.6)-(2.10), we have:
dW1(t)dt={1−S∗S}{Λ−β1I(t)S(t)−β2Hm(t)S(t)−β2Hs(t)S(t)−δS(t)}+{1−I∗I}{β1I(t−τ)S(t−τ)+β2Hm(t−τ)S(t−τ)+β2Hs(t−τ)S(t−τ)−m1I(t)}+(β2H∗m+β2H∗s)S∗αI∗{1−Q∗Q}{αI(t)−m2Q(t)}+β2H∗mS∗γ(1−p)Q∗{1−H∗mHm}{(1−p)γQ(t)−m3Hm(t)}+β2H∗sS∗γpQ∗{1−H∗sHs}{(pγQ(t)−m3Hs(t)}. | (4.8) |
At endemic equilibrium, we have:
{Λ=(β1I∗+β2H∗m+β2H∗s)S∗+δS∗,m1I∗=(β1I∗+β2H∗m+β2H∗s)S∗,m1Q∗=αI∗,m3H∗m=(1−p)γQ∗,m3H∗s=pγQ∗. | (4.9) |
Using the above constants, we have:
dW1(t)dt=δs∗(2−SS∗−S∗S)+β1I∗S∗(2−S∗S−II∗.SS∗)+β2H∗mS∗×(4−S∗S−QQ∗.H∗mHm−II∗.Q∗Q−HmH∗m.SS∗)+β2H∗sS∗×(4−S∗S−QQ∗.H∗sHs−II∗.Q∗Q−HsH∗s.SS∗)+β1I(t−τ)S(t−τ)(1−I∗I)+β2Hm(t−τ)S(t−τ)(1−I∗I)+β2Hs(t−τ)S(t−τ)(1−I∗I)−β1IS−β2HmS−β2HsS. | (4.10) |
The derivatives of W+2 are given by:
dW2(t)dt=β1S∗I∗ddt∫τ0{I(t−ξ)S(t−ξ)S∗I∗−1}dξ−β1S∗I∗ddt∫τ0{ln(I(t−ξ)S(t−ξ)S∗I∗)}dξ+β2S∗H∗mddt∫τ0{Hm(t−ξ)S(t−ξ)S∗H∗m−1}dξ−β2S∗H∗mddt∫τ0{ln(Hm(t−ξ)S(t−ξ)S∗H∗m)}dξ+β2S∗H∗sddt∫τ0{Hs(t−ξ)S(t−ξ)S∗H∗s−1}dξ−β2S∗H∗sddt∫τ0{ln(Hs(t−ξ)S(t−ξ)S∗H∗s)}dξ,=β1S∗I∗∫τ0ddt{I(t−ξ)S(t−ξ)S∗I∗−1}dξ−β1S∗I∗∫τ0ddt{ln(I(t−ξ)S(t−ξ)S∗I∗)}dξ+β2S∗H∗m∫τ0ddt{Hm(t−ξ)S(t−ξ)S∗H∗m−1}dξ−β2S∗H∗m∫τ0ddt{ln(Hm(t−ξ)S(t−ξ)S∗H∗m)}dξ+β2S∗H∗s∫τ0ddt{Hs(t−ξ)S(t−ξ)S∗H∗s−1}dξ−β2S∗H∗s∫τ0ddt{ln(Hs(t−ξ)S(t−ξ)S∗H∗s)}dξ,=−β1S∗I∗∫τ0ddξ{I(t−ξ)S(t−ξ)S∗I∗−1−ln(I(t−ξ)S(t−ξ)S∗I∗)}dξ−β2S∗H∗m∫τ0ddξ{Hm(t−ξ)S(t−ξ)S∗H∗m−−ln(Hm(t−ξ)S(t−ξ)S∗H∗m)}dξ−β2S∗H∗s∫τ0ddξ{Hs(t−ξ)S(t−ξ)S∗H∗s−1−ln(Hs(t−ξ)S(t−ξ)S∗H∗s)}dξ,=β1S∗I∗{I(t)S(t)S∗I∗−I(t−τ)S(t−τ)S∗I∗+ln(I(t−τ1)S(t−τ)I(t)S(t))}+β2S∗H∗m{Hm(t)S(t)S∗H∗m−Hm(t−τ)S(t−τ)S∗H∗m+ln(Hm(t−τ)S(t−τ)Hm(t)S(t))}+β2S∗H∗s{Hs(t)S(t)S∗H∗s−Hs(t−τ)S(t−τ)S∗H∗s+ln(Hs(t−τ)S(t−τ)Hs(t)S(t))}. | (4.11) |
Combining the derivatives ˙W1(t) and ˙W2(t), we have:
dW(t)dt=δS∗{2−SS∗−S∗S}+β1I∗S∗{2−S∗S(t)−S(t−τ)I(t−τ)S∗I+ln(I(t−τ)S(t−τ)I(t)S(t))}+β2S∗H∗m{4−S∗S(t)−H∗mHm(t).QQ∗−I(t)I∗.Q∗Q(t)−S(t−τ)Hm(t−τ)I∗S∗H∗mI+ln(Hm(t−τ)S(t−τ)Hm(t)S(t))}+β2S∗H∗s{4−S∗S(t)−H∗sHs(t).QQ∗−I(t)I∗.Q∗Q(t)−S(t−τ)Hs(t−τ)I∗S∗H∗sI+ln(Hs(t−τ)S(t−τ)Hs(t)S(t))}=δS∗{2−SS∗−S∗S}+β1I∗S∗{1−S∗S(t)+ln(S∗S(t))}+β1I∗S∗{1−S(t−τ)I(t−τ)S∗I+ln(I(t−τ)S(t−τ)I(t)S∗)}+β2S∗H∗m{1−S∗S(t)+ln(S∗S(t))}+β2S∗H∗m{1−H∗mHm(t).QQ∗+ln(H∗mQ(t)Hm(t)Q∗)}+β2S∗H∗m{1−I(t)I∗.Q∗Q(t)+ln(I(t)Q∗I∗Q(t))}+β2S∗H∗m{1−S(t−τ)Hm(t−τ)I∗S∗H∗mI+ln(Hm(t−τ)S(t−τ)I∗H∗mS∗I)}+β2S∗H∗s{1−S∗S(t)+ln(S∗S(t))}+β2S∗H∗s{1−H∗sHs(t).QQ∗+ln(H∗sQHs(t)Q∗)}+β2S∗H∗s{1−I(t)I∗.Q∗Q(t)+ln(I(t)Q∗I∗Q(t))}+β2S∗H∗s{1−S(t−τ)Hs(t−τ)I∗S∗H∗sI+ln(Hs(t−τ)S(t−τ)I∗H∗sS∗I)}. | (4.12) |
Since the arithmetic mean is greater than or equal to the geometric mean, we have
2≤S(t)S∗+S∗S(t), | (4.13) |
and it follows that
{2−S(t)S∗+S∗S(t)}≤0 | (4.14) |
for all S(t)>0, because the arithmetic mean is greater than or equal to the geometric mean.
Further, note that a continuous and differentiable function G(t)=1−g(t)+lng(t) is always non positive for any function g(t)>0, and g(t)=0 if and only if g(t)=1. Thus we note that
1−S∗S(t)+ln(S∗S(t))=G(S∗S(t))≤0 | (4.15) |
1−S(t−τ)I(t−τ)S∗I+ln(S(t−τ)I(t−τ)S∗I)=G(S(t−τ)I(t−τ)S∗I(t))≤0 | (4.16) |
1−H∗mQHmQ∗+ln(H∗mQHmQ∗)=G(H∗mQHmQ∗)≤0 | (4.17) |
1−Q∗IQI∗+ln(Q∗IQI∗)=G(Q∗IQI∗)≤0 | (4.18) |
1−S(t−τ)Hm(t−τ)I∗S∗H∗mI+ln(S(t−τ)Hm(t−τ)I∗S∗H∗mI)=G(S(t−τ)Hm(t−τ)I∗S∗H∗mI(t))≤0 | (4.19) |
1−H∗sQHsQ∗+ln(H∗sQHsQ∗)=G(Hs∗QHsQ∗)≤0 | (4.20) |
1−S(t−τ)Hs(t−τ)I∗S∗H∗sI+ln(S(t−τ)Hs(t−τ)I∗S∗H∗sI)=G(S(t−τ)Hs(t−τ)I∗S∗H∗sI(t))≤0 | (4.21) |
Hence, it follows that W(t)≤0 and consequently, ˙W(t)≤0. Moreover, the largest invariant set of ˙W(t)=0 is a singleton where S(t)≡S∗, I(t)≡I∗, Q(t)≡Q∗, Hm(t)≡H∗m, and Hs(t)≡H∗s. Using LaSalle's invariance principle [12], we conclude that the endemic equilibrium point E∗ is globally asymptotically stable if R0>1.
[1] | V. Ajraldi, M. Pittavino, E. Venturino, Modelling herd behaviour in population systems. Nonlinear Anal. Real World Appl., 12 (2011), 2319–2338. https://doi.org/10.1016/j.nonrwa.2011.02.002 |
[2] |
P. A. Braza, Predator-prey dynamics with square root functional responses, Nonlinear Anal. Real World Appl., 13 (2012), 1837–1843. https://doi.org/10.1016/j.nonrwa.2011.12.014 doi: 10.1016/j.nonrwa.2011.12.014
![]() |
[3] |
I. M. Bulai, E. Venturino, Shape effects on herd behavior in ecological interacting population models, Math. Comput. Simulat., 141 (2017), 40–55. https://doi.org/10.1016/j.matcom.2017.04.009 doi: 10.1016/j.matcom.2017.04.009
![]() |
[4] | I. Boudjema, S. Djilali, Turing-Hopf bifurcation in Gauss-type model with cross diffusion and its application, Nonlinear Stud., 25 (2018), 665–687. |
[5] |
S. Djilali, Herd behavior in a predator-prey model with spatial diffusion: bifurcation analysis and Turing instability, J. Appl. Math. Comput., 58 (2018), 125–149. https://doi.org/10.1007/s12190-017-1137-9 doi: 10.1007/s12190-017-1137-9
![]() |
[6] |
S. Djilali, Impact of prey herd shape on the predator-prey interaction, Chaos, Solitons Fract., 120 (2019), 139–148. https://doi.org/10.1016/j.chaos.2019.01.022 doi: 10.1016/j.chaos.2019.01.022
![]() |
[7] | S. Djilali, Effect of herd shape in a diffusive predator-prey model with time delay, J. Appl. Anal. Comput., 9 (2019), 638–654. |
[8] |
S. Djilali, S. Bentout, Spatiotemporal patterns in a diffusive predator-prey model with prey social behavior, Acta. Appl. Math., 169 (2020), 125–143. https://doi.org/10.1007/s10440-019-00291-z doi: 10.1007/s10440-019-00291-z
![]() |
[9] |
S. Djilali, Pattern formation of a diffusive predator-prey model with herd behavior and nonlocal prey competition, Math. Meth. Appl. Sci., 43 (2020), 2233–2250. https://doi.org/10.1002/mma.6036 doi: 10.1002/mma.6036
![]() |
[10] |
S. Djilali, Spatiotemporal patterns induced by cross-diffusion in predator-prey model with prey herd shape effect, Int. J. Biomath., 13 (2020), 2050030. https://doi.org/10.1142/S1793524520500308 doi: 10.1142/S1793524520500308
![]() |
[11] | S. Djilali, B. Ghanbari, S. Bentout, A. Mezouaghi, Turing-Hopf bifurcation in a diffusive Mussel-Algae model with time-fractional-order derivative, Chaos, Solitons Fract., 138 (2020) 109954. https://doi.org/10.1016/j.chaos.2020.109954 |
[12] |
B. Ghanabri, S. Djilali, Mathematical and numerical analysis of a three-species predator-prey model with herd behavior and time fractional-order derivative, Math. Meth. Appl. Sci., 43 (2020), 1736–1752. https://doi.org/10.1002/mma.5999 doi: 10.1002/mma.5999
![]() |
[13] |
B. Ghanabri, S. Djilali, Mathematical analysis of a fractional-order predator-prey model with prey social behavior and infection developed in predator population, Chaos, Solitons Fract., 138 (2020), 109960. https://doi.org/10.1016/j.chaos.2020.109960 doi: 10.1016/j.chaos.2020.109960
![]() |
[14] |
B. Ghanbari, S. Kumar, R. Kumar, A study of behaviour for immune and tumor cells in immunogenetic tumour model with non-singular fractional derivative, Chaos, Solitons Fract., 133 (2020), 109619. https://doi.org/10.1016/j.chaos.2020.109619 doi: 10.1016/j.chaos.2020.109619
![]() |
[15] |
J. Gine, C. Valls, Nonlinear oscillations in the modified Leslie-Gower model, Nonlinear Anal. Real World Appl., 51 (2020), 103010. https://doi.org/10.1016/j.nonrwa.2019.103010 doi: 10.1016/j.nonrwa.2019.103010
![]() |
[16] |
E. F. D. Goufo, S. Kumar, S. B. Mugisha, Similarities in a fifth-order evolution equation with and with no singular kernel, Chaos, Solitons Fract., 130 (2020), 109467. https://doi.org/10.1016/j.chaos.2019.109467 doi: 10.1016/j.chaos.2019.109467
![]() |
[17] |
C. S. Holling, The functional response of invertebrate predator to prey density, Mem. Entomol. Soc. Canada, 98 (1966), 5–86. https://doi.org/10.4039/entm9848fv doi: 10.4039/entm9848fv
![]() |
[18] |
C. A. Ibarra, J. Flores, Dynamics of a Leslie-Gower predator-prey model with Holling type II functional response, Allee effect and a generalist predator, Math. Comput. Simul., 188 (2021), 1–22. https://doi.org/10.1016/j.matcom.2021.03.035 doi: 10.1016/j.matcom.2021.03.035
![]() |
[19] |
W. Jiang, Q. An, J. Shi, Formulation of the normal forms of Turing-Hopf bifurcation in reaction-diffusion systems with time delay, J. Differ. Equ., 268 (2020), 6067–6102. https://doi.org/10.1016/j.jde.2019.11.039 doi: 10.1016/j.jde.2019.11.039
![]() |
[20] |
W. Jiang, H. Wang, X. Cao, Turing instability and turing-hopf bifurcation in diffusive Schnakenberg systems with gene expression time delay, J. Dyn. Differ. Equ., 31 (2019), 2223–2247. https://doi.org/10.1007/s10884-018-9702-y doi: 10.1007/s10884-018-9702-y
![]() |
[21] |
S. Kumar, A new analytical modelling for fractional telegraph equation via Laplace transform, Appl. Math. Model., 38 (2014), 3154–3163. https://doi.org/10.1016/j.apm.2013.11.035 doi: 10.1016/j.apm.2013.11.035
![]() |
[22] |
S. Kumar, A. Kumar, D. Baleanu, Two analytical methods for time-fractional nonlinear coupled Boussinesq-Burger's equations arise in propagation of shallow water waves, Nonlinear Dyn., 85 (2016), 699–715. https://doi.org/10.1007/s11071-016-2716-2 doi: 10.1007/s11071-016-2716-2
![]() |
[23] |
S. Kumar, M. M. Rashidi, New analytical method for gas dynamics equation arising in shock fronts, Comput. Phys. Commun., 185 (2014), 1947–1954. https://doi.org/10.1016/j.cpc.2014.03.025 doi: 10.1016/j.cpc.2014.03.025
![]() |
[24] |
S. Kumar, D. Kumar, S. Abbasbandy, M. M. Rashidide, Analytical solution of fractional Navier-Stokes equation by using modified Laplace decomposition method, Ain Shams Eng. J., 5 (2014), 569–574. https://doi.org/10.1016/j.asej.2013.11.004 doi: 10.1016/j.asej.2013.11.004
![]() |
[25] |
S. Kumar, S. Ghosh, B. Samet, An analysis for heat equations arises in diffusion process using new Yang-Abdel-Aty-Cattani fractional operator, Math. Meth. Appl. Sci., 43 (2020), 6062–6080. https://doi.org/10.1002/mma.6347 doi: 10.1002/mma.6347
![]() |
[26] |
S. Kumar, R. Kumar, R. P. Agarwal, B. Samet, A study of fractional Lotka-Volterra population model using Haar wavelet and Adams-Bashforth-Moulton methods, Math. Meth. Appl. Sci., 43 (2020), 5564–5578. https://doi.org/10.1002/mma.6297 doi: 10.1002/mma.6297
![]() |
[27] |
Y. Liu, J. Wei, Spatiotemporal dynamics of a modified Leslie-Gower model with weak allee effect, Int. J. Bifurcat. Chaos, 30 (2020), 2050169. https://doi.org/10.1142/S0218127420501692 doi: 10.1142/S0218127420501692
![]() |
[28] |
Y. Li, F. Zhang, X. Zhuo, Flip bifurcation of a discrete predator-prey model with modified Leslie-Gower and Holling-type III schemes, Math. Biosci. Eng., 17 (2019), 2003–2015. https://doi.org/10.3934/mbe.2020106 doi: 10.3934/mbe.2020106
![]() |
[29] |
C. V. Pao, Dynamics of nonlinear parabolic systems with time delays, J. Math. Anal, Appl., 198 (1996), 751–779. https://doi.org/10.1006/jmaa.1996.0111 doi: 10.1006/jmaa.1996.0111
![]() |
[30] |
M. M. Rashidi, A. Hosseini, I. Pop, S. Kumar, N. Freidoonimehr, Comparative numerical study of single and two-phase models of nano-fluid heat transfer in wavy channel, Appl. Math. Mech.-Engl. Ed., 35 (2014), 831–848. https://doi.org/10.1007/s10483-014-1839-9 doi: 10.1007/s10483-014-1839-9
![]() |
[31] |
F. Souna, A. Lakmeche, S. Djilali, The effect of the defensive strategy taken by the prey on predator-prey interaction, J. Appl. Math. Comput., 64 (2020), 665–690. https://doi.org/10.1007/s12190-020-01373-0 doi: 10.1007/s12190-020-01373-0
![]() |
[32] |
F. Souna, A. Lakmeche, S. Djilali, Spatiotemporal patterns in a diffusive predator-prey model with protection zone and predator harvesting, Chaos, Solitons Fract., 140 (2020), 110180. https://doi.org/10.1016/j.chaos.2020.110180 doi: 10.1016/j.chaos.2020.110180
![]() |
[33] |
F. Souna, S. Djilali, A. Lakmeche, Spatiotemporal behavior in a predator-prey model with herd behavior and cross-diffusion and fear effect, Eur. Phys. J. Plus, 136 (2021), 474. https://doi.org/10.1140/epjp/s13360-021-01489-7 doi: 10.1140/epjp/s13360-021-01489-7
![]() |
[34] | M. E. Taylor, Partial differential equations III–Nonlinear equations, Applied Mathematical Science, Springer-Verlag, 1996. |
[35] |
S. Wang, Z. Xie, R. Zhong, Y. Wu, Stochastic analysis of a predator-prey model with modified Leslie-Gower and Holling type II schemes, Nonlinear Dyn., 101 (2020), 1245–1262. https://doi.org/10.1007/s11071-020-05803-3 doi: 10.1007/s11071-020-05803-3
![]() |
[36] |
D. Xu, M. Liu, X. Xu, Analysis of a stochastic predator-prey system with modified Leslie-Gower and Holling-type IV schemes, Phys. A, 537 (2020), 122761. https://doi.org/10.1016/j.physa.2019.122761 doi: 10.1016/j.physa.2019.122761
![]() |
[37] |
C. Xu, S. Yuan, T. Zhang, Global dynamics of a predator-prey model with defence mechanism for prey, Appl. Math. Lett., 62 (2016), 42–48. https://doi.org/10.1016/j.aml.2016.06.013 doi: 10.1016/j.aml.2016.06.013
![]() |