
This study introduced a novel SVIR epidemic model incorporating environmental white noise to account for stochastic fluctuations in disease transmission. The model was analyzed to determine conditions for disease persistence and extinction, with outcomes linked to the basic reproduction number. A numerical approach was employed to facilitate computational analysis, and simulations were conducted using data from existing literature to generate realistic predictions. The stochastic model was further evaluated against its deterministic counterpart to assess predictive accuracy. The results highlight the significant role of randomness in epidemiological dynamics, providing valuable insights into disease spread and control strategies.
Citation: Shah Hussain, Naveed Iqbal, Elissa Nadia Madi, Mohsen Bakouri, Ilyas Khan, Wei Sin Koh. On the stochastic modeling and forecasting of the SVIR epidemic dynamic model under environmental white noise[J]. AIMS Mathematics, 2025, 10(2): 3983-3999. doi: 10.3934/math.2025186
[1] | Shah Hussain, Naveed Iqbal, Elissa Nadia Madi, Thoraya N. Alharthi, Ilyas Khan . Vaccination strategies in a stochastic SIVR epidemic model. AIMS Mathematics, 2025, 10(2): 4441-4456. doi: 10.3934/math.2025204 |
[2] | Zhengwen Yin, Yuanshun Tan . Threshold dynamics of stochastic SIRSW infectious disease model with multiparameter perturbation. AIMS Mathematics, 2024, 9(12): 33467-33492. doi: 10.3934/math.20241597 |
[3] | Yubo Liu, Daipeng Kuang, Jianli Li . Threshold behaviour of a triple-delay SIQR stochastic epidemic model with Lévy noise perturbation. AIMS Mathematics, 2022, 7(9): 16498-16518. doi: 10.3934/math.2022903 |
[4] | Yousef Alnafisah, Moustafa El-Shahed . Deterministic and stochastic model for the hepatitis C with different types of virus genome. AIMS Mathematics, 2022, 7(7): 11905-11918. doi: 10.3934/math.2022664 |
[5] | Xiaodong Wang, Kai Wang, Zhidong Teng . Global dynamics and density function in a class of stochastic SVI epidemic models with Lévy jumps and nonlinear incidence. AIMS Mathematics, 2023, 8(2): 2829-2855. doi: 10.3934/math.2023148 |
[6] | Roshan Ara, Saeed Ahmad, Zareen A. Khan, Mostafa Zahri . Threshold dynamics of stochastic cholera epidemic model with direct transmission. AIMS Mathematics, 2023, 8(11): 26863-26881. doi: 10.3934/math.20231375 |
[7] | Ishtiaq Ali, Sami Ullah Khan . Dynamics and simulations of stochastic COVID-19 epidemic model using Legendre spectral collocation method. AIMS Mathematics, 2023, 8(2): 4220-4236. doi: 10.3934/math.2023210 |
[8] | Xiangming Zhao, Jianping Shi . Dynamic behavior of a stochastic SIR model with nonlinear incidence and recovery rates. AIMS Mathematics, 2023, 8(10): 25037-25059. doi: 10.3934/math.20231278 |
[9] | Ying He, Bo Bi . Conditions for extinction and ergodicity of a stochastic Mycobacterium tuberculosis model with Markov switching. AIMS Mathematics, 2024, 9(11): 30686-30709. doi: 10.3934/math.20241482 |
[10] | Tingting Wang, Shulin Sun . Dynamics of a stochastic epidemic model with quarantine and non-monotone incidence. AIMS Mathematics, 2023, 8(6): 13241-13256. doi: 10.3934/math.2023669 |
This study introduced a novel SVIR epidemic model incorporating environmental white noise to account for stochastic fluctuations in disease transmission. The model was analyzed to determine conditions for disease persistence and extinction, with outcomes linked to the basic reproduction number. A numerical approach was employed to facilitate computational analysis, and simulations were conducted using data from existing literature to generate realistic predictions. The stochastic model was further evaluated against its deterministic counterpart to assess predictive accuracy. The results highlight the significant role of randomness in epidemiological dynamics, providing valuable insights into disease spread and control strategies.
The study of epidemic models has gained increasing importance due to the rising frequency and impact of infectious diseases on global health. These models provide essential frameworks for understanding disease spread and devising effective control strategies. Among these, the susceptible-vaccinated-infectious-recovered (SVIR) model plays a crucial role in analyzing diseases where vaccination is a key factor in mitigating outbreaks.
Traditional deterministic epidemic models assume homogeneous populations and predictable disease spread. However, real-world epidemics exhibit significant randomness due to environmental fluctuations, human behavior, and uncertainties in vaccination coverage. To address these complexities, researchers have incorporated stochasticity into epidemic models, leading to more realistic predictions. Recent advancements in stochastic SVIR models have explored various aspects, including optimal vaccination strategies, the influence of media on disease spread, and the impact of different types of noise [1,2,3].
In this work, we focus on a stochastic SVIR model that accounts for non-constant population sizes and random fluctuations in transmission and vaccination rates. By integrating stochastic elements into key parameters, we aim to capture the inherent randomness of real-world epidemics. Our numerical analyses highlight how vaccination strategies can effectively control disease spread, even under unpredictable disturbances. These findings contribute to understanding the interplay between stochastic influences, vaccination, and population dynamics, offering valuable insights for epidemic control and management [4].
Building upon previous research, including the foundational works of Allen [5], Brauer [6], Gray et al., Rozhnova et al., and Witbooi [7,8,9], we examine how noise and other stochastic factors affect the long-term behavior of SVIR models.
This study addresses these gaps by developing a stochastic SVIR model that incorporates environmental white noise, capturing the inherent variability in epidemic dynamics. Unlike previous studies, the model explicitly accounts for random fluctuations in both disease transmission and vaccination rates while considering a dynamic population size. This approach provides a more comprehensive framework to evaluate the interplay between stochasticity, vaccination strategies, and population dynamics, ultimately advancing the field of stochastic epidemic modeling.
This study builds on the work of Joko Harianto and Titik Suparwati, who initially developed an SVIR model to assess the effectiveness of various vaccination strategies [10]. They structured the model into four distinct compartments: St for susceptible individuals, Vt for vaccinated individuals, It for infected individuals, and Rt for recovered individuals.
In our model, we define β as the transmission rate necessary for disease spread and assume a constant recovery rate denoted by γ. The rate at which susceptible individuals are vaccinated is represented by α, while disease-induced mortality is captured by ω. The natural mortality rates, which are unrelated to the disease, are given as μ1 for the susceptible group, μ2 for the vaccinated, μ3 for the infected, and μ4 for the recovered individuals. Newborns are added to the susceptible class at a constant rate μ, ensuring a continuous supply of individuals who can potentially contract the disease.
Furthermore, γ1 represents the rate at which susceptible individuals acquire immunity and transition into the recovered compartment. The model assumes that immunity gained through vaccination is long-lasting and equivalent to natural immunity. However, vaccinated individuals may still be susceptible to infection, albeit at a reduced rate β1, reflecting partial immunity or lower effective contact with infected individuals. This comprehensive model is formulated as follows [10]:
ddt(SVIR)=(μ−S(μ1+βI+α)αS−V(β1I+γ1+μ2)I(βS+β1V)−I(γ+μ3+ω)γ1V+γI−μ4R). | (1.1) |
We begin our analysis with the following initial conditions: S0>0, V0>0, I0>0, and R0>0. These initial values represent the population sizes for the susceptible, vaccinated, infected, and recovered groups, respectively, at the initial time point t=0, reflecting the population's status at the start of the observational period.
Importantly, the initial population satisfies the relation:
S0+V0+I0+R0=N, | (1.2) |
where N represents the total population size, ensuring that all individuals in the population are accounted for. In the deterministic SVIR model, the total population remains constant as births and deaths balance out. However, in the stochastic version, the total population may decrease over time due to the influence of stochastic perturbations.
To maintain biological plausibility and consistency, all model parameters are assumed to be strictly positive. This assumption prevents biologically unrealistic values and ensures the system remains well-defined.
Following the theoretical framework outlined in [10], the system exhibits key properties that govern disease transmission dynamics over time. While the total population remains constant in the deterministic setting, the distribution of individuals among the compartments evolves dynamically based on the model's parameters.
By ensuring that all parameters are strictly positive and that the population distribution is comprehensive, the model remains consistent with real-world epidemiological scenarios. This approach allows for an accurate representation of disease spread and the impact of intervention strategies.
● The disease-free equilibrium, denoted as E0=(S0,V0,I0,R0), is given by:
E0=(μα+μ1,αμ(α+μ1)(γ1+μ2),0,γ1αμμ4(α+μ1)(γ1+μ2)). |
This equilibrium exists and is globally asymptotically stable if the basic reproduction number R0 satisfies the condition:
R0=βμ(α+μ1)(μ3+γ+ω)+β1αμ(α+μ1)(μ3+γ+ω)(γ1+μ2)≤1. |
● If the basic reproduction number R0 is less than or equal to one (R0≤1), the system stabilizes at a disease-free equilibrium. In this case, the infection cannot sustain itself within the population and eventually dies out. When R0≤1, any initial outbreak will dissipate over time as the number of infected individuals declines to zero.
Conversely, if R0 exceeds one (R0>1), the system's dynamics change significantly. The model then predicts the existence of an endemic equilibrium, where the disease persists in the population at a stable level. This means that the infection will not be eradicated but will remain within the population over time. The infected compartment, represented by I, will only maintain a non-zero value when the condition R0>1 holds. Therefore, the establishment of an endemic state is directly linked to this threshold, mathematically represented as follows:
E+=(S+,V+,I+,R+)=(μα+μ1+βI+,αμ(α+μ1+βI+)(γ1+μ2+β1I+),I+,γ1V++γI+μ4) |
where I+ is the positive root of θ(I)=O1I2+O2I+O3(1−R0), where:
● O1=(μ3+γ+ω)β1β>0,
● O2=(μ3+γ+ω)[(α+μ1)β1+(γ1+μ2)β]−β1βμ,
● O3=(μ3+γ+ω)(α+μ)(γ1+μ)>0.
Environmental factors significantly impact biomathematical models in real-world scenarios, as evidenced by recent studies [11,12]. To accurately capture the dynamics of epidemics within such unpredictable environments, it is essential to employ stochastic differential equations. Recent literature has extensively explored epidemic models influenced by environmental white noise [13,14,15].
This paper aims to elucidate the effects of white noise on epidemic dynamics by analyzing a stochastic SVIR model. We hypothesize that stochastic white noise directly influences the compartments St, Vt, It, and Rt, as explored in [16,17]. Building on this perspective, we develop and analyze the following stochastic SVIR model by introducing perturbations in the parameters α and γ.
While the deterministic model offers valuable insights, it does not capture the randomness inherent in real-world epidemic scenarios. To address this, we introduce stochastic elements into the transmission and vaccination processes by perturbing the parameters α→α+σ1dB1 and γ→γ+σ2dB2, resulting in:
dS=(μ−μ1S−βSI−αS)dt−σ1SdB1,dV=(αS−β1VI−γ1V−μ2V)dt+σ1SdB1,dI=(βSI+β1VI−γI−μ3I−ωI)dt−σ2IdB2,dR=(γ1V+γI−μ4R)dt+σ2IdB2. | (1.3) |
Within our stochastic SVIR model, the terms B1(t) and B2(t) represent independent standard Brownian motion processes. These stochastic processes are essential components of the model, introducing randomness into the system and allowing us to simulate the unpredictable fluctuations that occur in real-world epidemic dynamics. The independence of B1(t) and B2(t) ensures that the random influences they represent do not correlate, thereby accurately capturing the inherent variability in both disease transmission and vaccination rates.
The coefficients σ21 and σ22 represent the variances of the stochastic perturbations associated with α and γ, respectively. Specifically, σ21 corresponds to the variance of the noise affecting the vaccination rate α, while σ22 pertains to the variance of the noise influencing the recovery rate γ. These variances determine the intensity of the stochastic fluctuations: higher values imply a greater degree of randomness within the system.
Introducing stochastic perturbations into epidemiological models allows for a more nuanced representation of disease dynamics. These perturbations account for random environmental factors and inherent variability that traditional deterministic models may overlook. By incorporating stochastic elements, the model better reflects real-world complexities, such as unexpected fluctuations in disease spread due to environmental changes, social behaviors, or other unpredictable influences. This approach ensures that the model not only captures the average behavior of an epidemic but also the potential deviations from expected outcomes, providing a more comprehensive and realistic understanding of disease progression [18,19].
The methodology we adopt in this study involves the development of novel Lyapunov functions and the construction of a rectangular set, both designed to be independent of the positive equilibrium E+ found in the deterministic counterpart of the model. This independence is crucial for accurately analyzing the stability and long-term behavior of the system under stochastic influences [20,21].
In epidemiological studies, it is well-recognized that many infectious diseases exhibit temporal variations that often correspond with seasonal changes. These variations significantly impact the spread and intensity of epidemics, making it essential to incorporate time-periodic coefficients into the analysis of disease dynamics. The inclusion of these periodic factors aligns the model more closely with real-world scenarios, where disease incidence may rise and fall cyclically due to factors such as temperature changes, population movements, or seasonal human activities [22,23,24].
In this study, we work within the framework of a complete probability space, which we denote by (Ω,F,{Ft}t≥0,P). This probability space is structured to include a filtration {Ft}t≥0 that satisfies the standard assumptions typically required in stochastic processes, such as being right-continuous and containing all P-null sets. These conditions are essential for ensuring that the filtration appropriately models the flow of information over time, which is critical for the analysis of stochastic dynamics in the model.
Let R+=[0,∞) be the non-negative real line, and R4+={(ℏ1,ℏ2,ℏ3,ℏ4)∈R4:ℏi>0,i=1,2,3,4} represents the positive orthant in four-dimensional space.
For an integrable function ℓ(t) over the interval [0,∞), we define its time-averaged value as ⟨ℓ⟩t=1t∫t0ℓ(s)ds. When dealing with a bounded function ℓ(t) defined on [0,∞), we use the notations ˆℓ=inft∈[0,∞)ℓ(t) and ˇℓ=supt∈[0,∞)ℓ(t) to indicate its essential infimum and supremum, respectively.
We now present the following theorem, which establishes the existence and uniqueness of positive solutions for the proposed dynamical system.
Theorem 1.1. Assume that the stochastic system described by Eq (1.3) has an initial condition (S0,V0,I0,R0)∈R4+. Then, for all t≥0, the system admits a unique and positive solution (St,Vt,It,Rt)∈R4+ almost surely.
Moreover, the solution (St,Vt,It,Rt) satisfies the following long-term properties:
limt→∞Stt=0,limt→∞Vtt=0,limt→∞Itt=0,limt→∞Rtt=0,a.s. | (1.4) |
Additionally, the following upper bounds hold:
lim supt→∞logStt≤0,lim supt→∞logVtt≤0,lim supt→∞logItt≤0,lim supt→∞logRtt≤0,a.s. | (1.5) |
Furthermore, if μ(t) satisfies the condition
μ(t)>σ21(t)∨σ22(t)2, | (1.6) |
then the following limits hold:
limt→∞1t∫t0S(r)dB1(r)=0,limt→∞1t∫t0I(r)dB2(r)=0,a.s. | (1.7) |
The proof of Theorem (1.1) follows a methodology similar to that used in Theorem 2.1 [18] and is supported by Lemmas 2.1 and 2.2 in [25]. As the reasoning is analogous, we will not reproduce the full details of the proof here.
This section delves into the conditions that determine whether the disease It will die out (extinction) or continue to persist over time. We begin by establishing a theorem that provides the conditions under which the disease will go extinct.
Theorem 2.1. Let the parameter μ be such that μ>max(σ212,σ222). Consider a solution (St,Vt,It,Rt) to the system given by Eq (1.3), where the initial values (S0,V0,I0,R0) are within the positive orthant R4+. Let us consider that the basic reproduction number, denoted as R∗0, is characterized by the following expression:
R∗0:=βμ(μ1+α)(μ3+γ+ω+σ222)+β1μμ2(μ3+γ+ω+σ222), |
and further suppose that this quantity satisfies the condition
R∗0<1. |
Under these assumptions, the solution (St,Vt,It,Rt) satisfies the following asymptotic behavior:
lim supt→∞logItt≤(μ3+γ+ω+σ222)(R∗0−1)<0,almost surely. |
This result indicates that the disease will eventually die out with probability one.
Proof. We start by analyzing the first equation in the system described by Eq (1.3). This equation gives the dynamics of the susceptible population St. Dividing the equation by t and rearranging terms, we have
St−S0t=μ−μ1⟨S⟩t−β⟨SI⟩t−α⟨S⟩t−σ1t∫t0S(r)dB1(r). |
Next, we solve for ⟨S⟩t explicitly:
μ−μ1⟨S⟩t−α⟨S⟩t−β⟨SI⟩t−σ1t∫t0S(r)dB1(r)=St−S0t. |
Rearranging this equatio n yields
μ=(μ1+α)⟨S⟩t+β⟨SI⟩t+St−S0t+σ1t∫t0S(r)dB1(r). |
Thus, solving for ⟨S⟩t gives
⟨S⟩t=μμ1+α−βμ1+α⟨SI⟩t−ð1(t), |
where the term ð1(t) is defined as
ð1(t)=1μ1+α(σ1t∫t0S(r)dB1(r)+St−S0t). |
Utilizing the strong law of large numbers applicable to martingales, along with the fundamental characteristics of stochastic integrals, it can be established that the stochastic component converges to zero as t→∞. Therefore, we conclude that
limt→∞ð1(t)=0,almost surely. |
This leads to the following upper bound for ⟨S⟩t:
lim supt→∞⟨S⟩t≤μμ1+α,almost surely. | (2.1) |
Next, we consider the second equation in the system described by Eq (1.3). This equation gives the dynamics of the vaccinated population Vt. Dividing the equation by t and rearranging terms, we have
Vt−V0t=α⟨S⟩t−β1⟨VI⟩t−(γ1+μ2)⟨V⟩t+σ1t∫t0S(r)dB1(r), |
Next, we solve for ⟨V⟩t explicitly:
α⟨S⟩t−β1⟨VI⟩t−(γ1+μ2)⟨V⟩t+σ1t∫t0S(r)dB1(r)=Vt−V0t. |
Rearranging this equation yields
(γ1+μ2)⟨V⟩t=α⟨S⟩t−β1⟨VI⟩t−Vt−V0t+σ1t∫t0S(r)dB1(r). |
Thus, solving for ⟨V⟩t gives
⟨V⟩t=α⟨S⟩tγ1+μ2−β1γ1+μ2⟨VI⟩t+ð2(t), |
where the term ð2(t) is defined as
ð2(t)=1γ1+μ2(σ1t∫t0S(r)dB1(r)−Vt−V0t). |
Utilizing the strong law of large numbers applicable to martingales, along with the fundamental characteristics of stochastic integrals, it can be established that the stochastic component converges to zero as t→∞. Therefore, we conclude that
limt→∞ð2(t)=0,almost surely. |
This leads to the following upper bound for ⟨V⟩t:
⟨V⟩t=α⟨S⟩tγ1+μ2−β1γ1+μ2⟨VI⟩t. |
From (2.1), one can get
lim supt→∞⟨V⟩t≤αμ(γ1+μ2)(μ1+α),almost surely. | (2.2) |
Now, considering the entire system in (1.3), we sum up the differential equations to get the total population equation:
Rt−R0t+St−S0t+Vt−V0t+It−I0t=μ−μ1⟨S⟩t−μ2⟨V⟩t−(μ3+ω)⟨I⟩t−μ4⟨R⟩t. |
This equation can be rearranged as
μ1⟨S⟩t+μ2⟨V⟩t+(μ3+ω)⟨I⟩t+μ4⟨R⟩t=μ. | (2.3) |
Next, we apply Ito's lemma to the third equation of the system (1.3), which governs the dynamics of the infected population It. This yields
logIt−logI0t=β⟨S⟩t+β1⟨V⟩t−μ3−γ−ω−σ222+σ2B2(t)t. |
By inserting the expression for ⟨V⟩t provided in Eq (2.3) into the current equation, we obtain the following result:
logIt−logI0t=(β−β1μ1μ2)⟨S⟩t+β1μμ2−(μ3+γ+ω+σ222)−β1(μ3+γ)μ2⟨I⟩t−β1μ4μ2⟨R⟩t+ð3(t),≤βμμ1+α+β1μμ2−(μ3+γ+ω+σ222)−β1μ1μ2⟨S⟩t−β1(μ3+γ)μ2⟨I⟩t−β1μ4μ2⟨R⟩t+ð3(t), | (2.4) |
where ð3(t) is a term that converges to zero almost surely as t→∞.
Taking the limit superior of both sides of Eq (2.4), and combining the results from (2.1) and (2.3), we obtain
lim supt→∞logItt≤(μ3+γ+ω+σ222)(R∗0−1),almost surely. |
Therefore, if R∗0<1, then limt→∞It=0, almost surely, indicating that the disease will become extinct with probability one.
Theorem 2.2. If
R∗0=βμ(μ+α+σ21/2)(γ+μ3+ω+σ22/2)+β1μα(μ+α+σ21/2)(μ2+γ1+σ21/2)(γ+μ3+ω+σ22/2)>1, |
then for any initial value (S0,V0,I0,R0)∈R4+, the infected population It of the model has the property
lim inft→∞⟨I⟩t≥(γ+μ3+ω+σ22/2)(R∗0−1)(a1+b1)β+b2β1,a.s, |
where
a1=βμμ+α+σ21/2,b1=β1μα(μ+α+σ21/2)(μ2+γ1+σ21/2),b2=β1μα(μ+α+σ21/2)(μ2+γ1+σ21/2). |
In other words, this result indicates that the disease will continue to spread and persist in the population if the basic reproduction number R∗0 exceeds the critical threshold of 1. Specifically, when R∗0>1, each infected individual, on average, generates more than one new infection, which leads to sustained transmission and the continued presence of the disease within the population.
Proof. Set
V(S,V,I,R)=−logI−(a1+b1)logS−b2logV−c1logR, |
where positive constants a1, b1, b2, and c1 will be established in subsequent calculations. By utilizing Ito's lemma, we derive the following result.
dV=Vdt−(a1+b1)σ1dB1(t)−b2σ1dB1(t)−c1σ2dB2(t)−σ2dB2(t), | (2.5) |
where
LV=L(−logI−(a1+b1)logS−b2logV−c1logR)=−βS−β1V+γ+μ3+ω+σ222−a1μS+a1(μ+α+σ212)+a1βI−b1μS+b1(μ+α+σ212)+b1βI−b2αSV+b2β1I+b2(μ2+γ1+σ212)−c1μ4R+c1γ1+c1γ+c1σ222+γ+μ3+ω+σ222+((a1+b1)β+b2β1)I. |
Let
a1=βμμ+α+σ21/2,b1=β1μα(μ+α+σ21/2)(μ2+γ1+σ21/2),b2=β1μα(μ+α+σ21/2)(μ2+γ1+σ21/2),c1=γ+μ3+ω+σ22/2γ1. |
Then
LV≤−βμμ+α+σ21/2−β1μα(μ+α+σ21/2)(μ2+γ1+σ21/2)+γ+μ3+ω+σ222+((a1+b1)β+b2β1)I. | (2.6) |
This simplifies to
LV≤−(γ+μ3+ω+σ222)(R∗0−1)+((a1+b1)β+b2β1)I. | (2.7) |
By replacing the expression given in Eq (2.7) into Eq (2.5), and then performing integration on both sides of Eq (2.5), we derive the following outcome:
1t[V(St,Vt,It,Rt)−V(S0,V0,I0,R0)]≤−(γ+μ3+ω+σ222)(R∗0−1)+[(a1+b1)β+b2β1]⟨I⟩t−M(t)t, | (2.8) |
where M(t) is a martingale given by
M(t)=(a1+b1)∫t0σ1dB1(s)+b2∫t0σ1dB1(s)+c1∫t0σ2dB2(s)+∫t0σ2dB2(s). |
Applying the strong law of large numbers for martingales, we can deduce that as t approaches infinity, the expression
limt→∞M(t)t=0,a.s. |
converges to zero. This implies that the contribution of the martingale term becomes negligible in the long run.
lim inft→∞1t[V(St,Vt,It,Rt)−V(S0,V0,I0,R0)]≤lim inft→∞[−(γ+μ3+ω+σ222)(R∗0−1)+[(a1+b1)β+b2β1]⟨I⟩t]. |
As t grows indefinitely, the difference term 1t[V(St,Vt,It,Rt)−V(S0,V0,I0,R0)] gradually diminishes. This is due to the application of the strong law of large numbers for martingales, which guarantees that the limit of this difference as t tends to infinity is zero almost surely. Therefore, we can assert:
limt→∞1t[V(Rt,St,Vt,It)−V(R0,S0,V0,I0)]=0. |
Given this result, it follows directly that:
lim inft→∞1t[V(Rt,St,Vt,It)−V(R0,S0,V0,I0)]=0. |
Therefore, the inequality simplifies to:
0≤−(γ+μ3+ω+σ222)(R∗0−1)+[(a1+b1)β+b2β1]lim inft→∞⟨I⟩t. |
Furthermore,
lim inft→∞⟨I⟩t≥(γ+μ3+ω+σ222)(R∗0−1)(a1+b1)β+b2β1. |
This completes the proof.
The deterministic reproduction number and the stochastic threshold values are connected, as stochastic perturbations modify the effective threshold. High stochastic intensities can drive disease extinction by destabilizing infection dynamics. Stochastic perturbations introduce variability, altering disease persistence and extinction probabilities. These insights highlight the need for stochastic modeling in realistic epidemic predictions. The deterministic model uses a basic reproduction number to determine disease persistence. In the stochastic model, this threshold is modified due to the presence of random perturbations. Stochastic fluctuations lower the effective reproduction number, making disease extinction more likely even when the deterministic model predicts persistence. Stronger stochastic fluctuations increase the probability of disease extinction. When random noise intensity reaches a certain level, it disrupts transmission dynamics enough to drive the infection out of the population. This highlights how real-world uncertainties, such as environmental or behavioral changes, can influence disease spread and control. These findings emphasize the importance of incorporating stochasticity into epidemic models to better reflect real-world uncertainties and improve disease prediction and control strategies.
After completing the examination of disease eradication and long-term persistence, we shift our focus to performing numerical simulations. We will execute simulations based on the model described in Eq (1.3) to illustrate the practical implications of our findings. For this purpose, we will use the following parameter values: μ=0.03, μ1=0.04, μ2=0.035, μ3=0.06, μ4=0.045, β=0.5, β1=0.05, γ=0.1, γ1=0.1, α=0.2, and ω=0.1. The initial conditions for the simulation are set as follows: S(0)=0.5, V(0)=0.3, I(0)=0.2, and R(0)=0. For the numerical simulations, we employ the technique of Higham [26]. Let us consider the discretized equation of the model in which ξ is a normally distributed random variable that introduces randomness into the numerical scheme and Δt is the time step controlling how frequently the system updates, which is chosen very small, equals to 0.01 or 0.001.
Sk+1=Sk+(μ−μ1Sk−βSkIk−αSk)Δt−σ1Sk√Δtξk−12σ21Sk(Δtξ2k−Δt),Vk+1=Vk+(αSk−β1VkIk−γ1Vk−μ2Vk)Δt+σ1Sk√Δtξk+12σ21Sk(Δtξ2k−Δt),Ik+1=Ik+(βSkIk+β1VkIk−γIk−μ3Ik−ωIk)Δt−σ2Ik√Δtζk−12σ22Ik(Δtξ2k−Δt),Rk+1=Rk+(γ1Vk+γIk−μ4Rk)Δt+σ2Ik√Δtζk+12σ22Ik(Δtξ2k−Δt). | (3.1) |
Following the incorporation of white noise into the system (1.3), we analyze the behavior of the susceptible class, denoted as S in Figure 1. The addition of stochastic perturbations introduces randomness into the dynamics of the susceptible population, causing fluctuations that deviate from the deterministic model. The stochastic differential equation governing the S class can be expressed as a modified version of the original system, incorporating a noise term. This term reflects the inherent unpredictability in the transmission dynamics due to various external factors, leading to a more realistic representation of the system under study. The resulting model provides valuable insights into the potential variability and uncertainty in the population dynamics, emphasizing the importance of accounting for random effects in epidemiological modeling.
When white noise is introduced into the system (1.3), the dynamics of the vaccinated class, denoted by V, experience stochastic fluctuations in Figure 2. The incorporation of this noise term into the differential equation for V adds a layer of randomness to the vaccination process, simulating real-world uncertainties such as variability in vaccine efficacy or coverage. This stochastic perturbation causes the size of the vaccinated population to fluctuate around its deterministic path, reflecting possible variations in the rate at which individuals are vaccinated or the effectiveness of the vaccine over time. The modified model for V thus captures the inherent unpredictability in the vaccination dynamics, providing a more nuanced understanding of how random factors might influence the overall success of a vaccination program.
The inclusion of white noise into the model significantly impacts the infected class, denoted by I. The stochastic term introduces random fluctuations into the infection dynamics, which can represent various sources of uncertainty such as changes in contact rates, transmission variability, or other external factors affecting the spread of the disease. These random perturbations cause the number of infected individuals to deviate from the deterministic trajectory, capturing the unpredictable nature of disease transmission in a real-world setting. As a result, the modified differential equation for the infected class I not only accounts for the expected progression of the disease but also reflects the possible short-term increases or decreases in the number of infected individuals due to random influences. This stochastic approach provides a more realistic portrayal of the infection dynamics, emphasizing the role of randomness in disease spread and control in Figure 3.
The recovered class, denoted by R, is influenced by the introduction of white noise in a manner that reflects the uncertainties associated with recovery rates. The stochastic perturbations in the model account for the randomness in the recovery process, which could stem from factors such as varying immune responses among individuals, differences in treatment efficacy, or external influences that affect recovery times. As a result, the number of recovered individuals R fluctuates around its expected value, capturing the inherent variability in the recovery process in Figure 4. This addition of white noise ensures that the model more accurately represents the unpredictable nature of recovery, providing a more comprehensive understanding of how recovered individuals contribute to the overall disease dynamics in a stochastic environment.
In classical epidemic modeling, deterministic approaches provide insights into disease dynamics but fail to account for real-world randomness. Our study transitions from a deterministic SVIR model to a stochastic framework by introducing environmental white noise into the transmission and vaccination rates, leading to system (1.3). These stochastic perturbations capture unpredictable fluctuations in infection spread, vaccination uptake, and recovery rates, making the model more biologically realistic than its deterministic counterpart.
From the theoretical results, we establish the stochastic reproduction number R∗0. Unlike the deterministic threshold, stochastic effects introduce an additional extinction condition, meaning that even when R∗0>1, random fluctuations may drive the disease to extinction. Numerical simulations confirm that increasing noise intensity (σ1,σ2) can lower R∗0 below 1, altering long-term epidemic outcomes. This effect is absent in deterministic models, where R0>1 always predicts persistence.
Our findings align with Rozhnova and Nunes [8], who observed that randomness increases infection peak variability, emphasizing the importance of adaptive vaccination. Similarly, Gray et al [7] reported noise-induced disease resurgence, supporting our results.
The implications of these findings are substantial for public health policies. Unlike deterministic models, stochastic simulations show that small fluctuations in transmission or vaccination rates can significantly alter epidemic outcomes. These results reinforce the need for adaptive epidemic control measures, particularly in dynamic environments where real-time responses to disease fluctuations are essential.
One limitation of this model is that it assumes constant rates, whereas real-world epidemiological parameters fluctuate over time. Future research should integrate time-dependent transmission and vaccination rates, spatial effects, and real epidemiological data validation. Further exploration of stochastic control measures, such as adaptive vaccination and quarantine policies, could refine epidemic management strategies.
Overall, this study demonstrates that stochastic SVIR models provide a more realistic representation of epidemic behavior than deterministic models. The stochastic framework captures randomness in disease spread, reinforcing the necessity of integrating uncertainty into epidemic modeling for better forecasting and intervention planning.
This study investigates the dynamics of disease transmission under the influence of stochastic perturbations, providing a comprehensive analysis of the system's behavior. By incorporating white noise into the SVIR model, we have effectively captured the inherent uncertainties present in real-world scenarios, particularly in infection, recovery, and vaccination processes. The stochastic model not only aligns more closely with observed epidemiological data but also offers deeper insights into the conditions for disease persistence and extinction. Our findings highlight the critical role of randomness in epidemiological models, demonstrating its significant impact on disease dynamics and predictive outcomes. This work contributes valuable insights into the complex interplay between deterministic and stochastic factors in epidemic modeling, offering a robust framework for future research and public health strategies aimed at controlling infectious diseases. The implications of this study are broad, paving the way for refinements in epidemiological modeling and improved disease management practices.
Shah Hussain: Writing original draft, software, methodology, formal analysis; Naveed Iqbal: Writing review and editing, visualization, validation, supervision; Elissa Nadia Madi: Investigation, formal analysis, conceptualization; Mohsen Bakouri: Visualization, validation, supervision; Ilyas Khan: Methodology, analysis, writing review and editing, supervision; Wei Sin Koh: Investigation, formal analysis, conceptualization.
All authors have agreed to and given their consent for the publication of this research paper.
All authors declare that no Artificial Intelligence (AI) tools were used in the creation of this article.
The author Mohsen Bakouri extends his appreciation to the Deanship of Postgraduate Studies and Scientific Research at Majmaah University for funding this research work through the project number (R-2025-1590).
All authors declare no conflicts of interest in this paper.
[1] |
A. Alkhazzan, J. Wang, Y. Nie, H. Khan, J. Alzabut, A stochastic Susceptible Vaccinees Infected Recovered epidemic model with three types of noises, Math. Method. Appl. Sci., 47 (2024), 8748–8770. https://doi.org/10.1002/mma.10042 doi: 10.1002/mma.10042
![]() |
[2] |
A. Alkhazzan, J. Wang, Y. Nie, H. Khan, J. Alzabut, An effective transport-related SVIR stochastic epidemic model with media coverage and Lévy noise, Chaos Soliton. Fract., 175 (2023), 113953. https://doi.org/10.1016/j.chaos.2023.113953 doi: 10.1016/j.chaos.2023.113953
![]() |
[3] |
A. Alkhazzan, J. Wang, Y. Nie, K. Hattaf, A new stochastic split-step θ-nonstandard finite difference method for the developed SVIR epidemic model with temporary immunities and general incidence rates, Vaccines, 10 (2022), 1682. https://doi.org/10.3390/vaccines10101682 doi: 10.3390/vaccines10101682
![]() |
[4] |
H. J. Alsakaji, F. A. Rihan, A. Hashish, Dynamics of a stochastic epidemic model with vaccination and multiple time-delays for COVID-19 in the UAE, Complexity, 2022 (2022), 4247800. https://doi.org/10.1155/2022/4247800 doi: 10.1155/2022/4247800
![]() |
[5] |
L. J. S. Allen, A primer on stochastic epidemic models: Formulation, numerical simulation, and analysis, Infectious Disease Modelling, 2 (2017), 128–142. https://doi.org/10.1016/j.idm.2017.03.001 doi: 10.1016/j.idm.2017.03.001
![]() |
[6] |
F. Brauer, Backward bifurcations in simple vaccination/treatment models, J. Biol. Dynam., 5 (2011), 410–418. https://doi.org/10.1080/17513758.2010.510584 doi: 10.1080/17513758.2010.510584
![]() |
[7] |
A. Gray, D. Greenhalgh, L. Hu, X. Mao, J. Pan, A stochastic differential equation SIS epidemic model, SIAM J. Appl. Math., 71 (2011), 876–902. https://doi.org/10.1137/10081856X doi: 10.1137/10081856X
![]() |
[8] | G. Rozhnova, A. Nunes, Stochastic effects in a seasonally forced epidemic model, Phys. Rev. E., 82 (2010), 041906. https://doi.org/10.1103/PhysRevE.82.041906 |
[9] |
P. J. Witbooi, Stability of a stochastic model of an SIR epidemic with vaccination, Acta Biotheor., 65 (2017), 151–165. https://doi.org/10.1007/s10441-017-9308-5 doi: 10.1007/s10441-017-9308-5
![]() |
[10] |
J. Harianto, T. Suparwati, SVIR epidemic model with non constant population, CAUCHY-Jurnal Matematika Murni dan Aplikasi, 5 (2018), 102–111. https://doi.org/10.18860/ca.v5i3.5511 doi: 10.18860/ca.v5i3.5511
![]() |
[11] |
Y. Cai, Y. Kang, M. Banerjee, W. Wang, A stochastic epidemic model incorporating media coverage, Commun. Math. Sci., 14 (2016), 893–910. https://doi.org/10.4310/CMS.2016.v14.n4.a1 doi: 10.4310/CMS.2016.v14.n4.a1
![]() |
[12] |
M. Liu, M. Fan, Permanence of stochastic Lotka-Volterra systems, J. Nonlinear Sci., 27 (2017), 425–452. https://doi.org/10.1007/s00332-016-9337-2 doi: 10.1007/s00332-016-9337-2
![]() |
[13] |
W. Wang, Y. Cai, Z. Ding, Z. Gui, A stochastic differential equation SIS epidemic model incorporating Ornstein–Uhlenbeck process, Physica A, 509 (2018), 921–936. https://doi.org/10.1016/j.physa.2018.06.099 doi: 10.1016/j.physa.2018.06.099
![]() |
[14] |
A. Lahrouz, L. Omari, Extinction and stationary distribution of a stochastic SIRS epidemic model with non-linear incidence, Stat. Probabil. Lett., 83 (2013), 960–968. https://doi.org/10.1016/j.spl.2012.12.021 doi: 10.1016/j.spl.2012.12.021
![]() |
[15] |
X. Zhang, K. Wang, Stochastic model for spread of AIDS driven by Lévy noise, J. Dyn. Diff. Equat., 27 (2015), 215–236. https://doi.org/10.1007/s10884-015-9459-5 doi: 10.1007/s10884-015-9459-5
![]() |
[16] |
Y. Zhao, D. Jiang, The threshold of a stochastic SIS epidemic model with vaccination, Appl. Math. Comput., 243 (2014), 718–727. https://doi.org/10.1016/j.amc.2014.05.124 doi: 10.1016/j.amc.2014.05.124
![]() |
[17] |
Q. Yang, D. Jiang, N. Shi, C. Ji, The ergodicity and extinction of stochastically perturbed SIR and SEIR epidemic models with saturated incidence, J. Math. Anal. Appl., 388 (2012), 248–271. https://doi.org/10.1016/j.jmaa.2011.11.072 doi: 10.1016/j.jmaa.2011.11.072
![]() |
[18] |
C. Ji, D. Jiang, Q. Yang, N. Shi, Dynamics of a multigroup SIR epidemic model with stochastic perturbation, Automatica, 48 (2012), 121–131. https://doi.org/10.1016/j.automatica.2011.09.044 doi: 10.1016/j.automatica.2011.09.044
![]() |
[19] |
D. Jiang, J. Yu, C. Ji, N. Shi, Asymptotic behavior of global positive solution to a stochastic SIR model, Math. Comput. Model., 54 (2011), 221–232. https://doi.org/10.1016/j.mcm.2011.02.004 doi: 10.1016/j.mcm.2011.02.004
![]() |
[20] |
R. Rudnicki, Long–time behavior of a stochastic prey–predator model, Stoch. Proc. Appl., 108 (2003), 93–107. https://doi.org/10.1016/S0304-4149(03)00090-5 doi: 10.1016/S0304-4149(03)00090-5
![]() |
[21] |
F. A. Rihan, H. J. Alsakaji, Dynamics of a stochastic delay differential model for COVID-19 infection with asymptomatic infected and interacting people: Case study in the UAE, Results Phys., 28 (2021), 104658. https://doi.org/10.1016/j.rinp.2021.104658 doi: 10.1016/j.rinp.2021.104658
![]() |
[22] |
Y. Lin, D. Jiang, T. Liu, Nontrivial periodic solution of a stochastic epidemic model with seasonal variation, Appl. Math. Lett., 45 (2015), 103–107. https://doi.org/10.1016/j.aml.2015.01.021 doi: 10.1016/j.aml.2015.01.021
![]() |
[23] |
X. Zhang, D. Jiang, A. Alsaedi, T. Hayat, Periodic solutions and stationary distribution of mutualism models in random environments, Physica A, 460 (2016), 270–282. https://doi.org/10.1016/j.physa.2016.05.015 doi: 10.1016/j.physa.2016.05.015
![]() |
[24] |
Y. Zhao, D. Jiang, The threshold of a stochastic SIS epidemic model with vaccination, Appl. Math. Comput., 243 (2014), 718–727. https://doi.org/10.1016/j.amc.2014.05.124 doi: 10.1016/j.amc.2014.05.124
![]() |
[25] |
C. Zhu, G. Yin, Asymptotic properties of hybrid diffusion systems, SIAM J. Control Optim., 46 (2007), 1155–1179. https://doi.org/10.1137/060649343 doi: 10.1137/060649343
![]() |
[26] |
D. J. Higham, An algorithmic introduction to numerical simulation of stochastic differential equations, SIAM Rev., 43 (2001), 525–546. https://doi.org/10.1137/S0036144500378302 doi: 10.1137/S0036144500378302
![]() |