Skip to main content
Advertisement
Browse Subject Areas
?

Click through the PLOS taxonomy to find articles in your field.

For more information about PLOS Subject Areas, click here.

  • Loading metrics

Dynamical behaviors of a stochastic SIVS epidemic model with the Ornstein-Uhlenbeck process and vaccination of newborns

Abstract

In this paper, we study a stochastic SIVS infectious disease model with the Ornstein-Uhlenbeck process and newborns with vaccination. First, we demonstrate the theoretical existence of a unique global positive solution in accordance with this model. Second, adequate conditions are inferred for the infectious disease to die out and persist. Then, by classic Lynapunov function method, the stochastic model is allowed to obtain the sufficient condition so that the stochastic model has a stationary distribution represents illness persistence in the absence of endemic equilibrium. Calculating the associated Fokker-Planck equations yields the precise expression of the probability density function for the linearized system surrounding the quasi-endemic equilibrium. In the end, the theoretical findings are shown by numerical simulations.

1 Introduction

Epidemic models with vaccination are one of the important models for studying epidemiology. The purpose of epidemiological research is to constrain and avoid disease incidence and transmission. Protection of susceptible population is one of the important means of preventing and dominating infectious diseases, and vaccination is an effective measure to protect susceptible population. Taking vaccination into account, Zhao et al. [1] formulated a stochastic SIS model with vaccination, giving a threshold for the model when the white noise intensity is low. Zhang et al. [2] studied a stochastic SVIR model and found that white noise contributed to disease control. Considering more realistic factors, many models of infectious diseases with vaccination have been explored [36].

However, vaccination of the population does not mean permanent immunity. After a while, they may lose their immunity. This phenomenon has been studied in many literatures [79]. In Ref. [10], authors studied an SIS infectious disease model with vaccination and changing total population size, where the effectiveness of vaccines is time-limited. In Ref. [11], the following SIS model with immunization was also discussed by Li and Ma, (1) where S(t), I(t) and V(t) are individuals who are susceptible to infection, infection, and vaccination, respectively. g(0 < g < 1) is a proportion of newborn vaccinations, A is a constant growth in the population, η is the natural mortality rate for these three compartments, p is the proportion of vulnerable individuals with vaccination, β is the coefficient of transmission between S and I, γ is the recovery rate of infectious individuals, α is the proportion of vaccinated people who have lost immunity, ε is the cause-specific mortality of infectious individuals. The lethality of the disease is less than the proportion of vulnerable individuals with vaccination, based on the realities of the social environment. Assume that all parameters are non-negative and η, A > 0.

By studying model Eq (1), Li et al. [11] obtained several important conclusions as follows:

  1. (1) The threshold is ,
  2. (2) The system exists the disease-free equilibrium and the endemic equilibrium P* = (S*, I*, V*), where
  3. (3) P0 is globally asymptotically stable when R0 < 1, P* is globally asymptotically stable when R0 > 1.

In fact, environmental fluctuations inevitably affect the occurrence and development of infectious diseases. Considering infectious diseases as a stochastic process is therefore more relevant. For this reason, Beddington and May [12] proposed the parameter perturbation method for the first time, which is a classical method to derive stochastic differential equation model from the corresponding deterministic system, and this method has been widely used [1317]. Liu et al. [18] investigated a stochastic SIQR infectious disease model that introduced the effects of two types of environmental noise, namely white noise and telegraph noise. They deduced adequate conditions for I(t) to persist and become extinct. There are sufficient conditions to prove that a positive recurrence solution exists when I(t) persists. Teng et al. [19] studied a stochastic SIS model with non-linear incidence, where white noise interferes with the propagation coefficient β. They obtained the threshold R0 for determining disappearance and prevalence of disease. In order to learn disease dynamics in a random environment, the next SIVS model was studied by Zhao et al. [20], (2)

Clearly, the model Eq (2) is randomly derived from the deterministic model Eq (1) via

B(t) is the standard Brownian movement, θ is the intensity of white noise. After integrating both sides and dividing by t, we get

It’s easy to get as t → 0. This suggests that the fluctuations of β(t) may become very large when the time interval is very small, which is clearly implausible. In Ref. [21], Allen indicated that compared with the linear function of white noise, the mean-reverting process has an advantage in showing the diversity of the environment. The mean-reversion process has a wide range of applications in both finance and physics [2224]. The parameters are set by a linear function of white Gaussian noise when they are perturbed by white Gaussian noise. However, there is another approach where the parameters are associated with a mean-reversion process [2530], namely the Ornstein-Uhlenbeck process, in the form of: where m(t) satisfies (3)

Notice that k, θ > 0. k is the reversion speed, θ is the volatility intensity of process m(t), B(t) represents a standard Brownian movement. By simple calculations, we can obtain , where m0m(0). Unlike white Gaussian noise, with the Ornstein-Uhlenbeck process tends to 0 as t → 0. Moreover, according to Ref. [3133], it is easy to get m(t) to be ergodic and to converge weakly to the invariant density

So, using the ergodic theorem in Ref. [34], one obtains (4)

Then, for a sufficiently small time interval Δt, the correlation coefficient of the Ornstein-Uhlenbeck process m(t) is r(m(t), m(t + Δt)) = 1 − ot) and r(m(t), m(t + Δt)) = 0 for white Gaussian noise [21]. Since the correlation coefficient between neighbouring states is close to 1, this indicates a high degree of correlation between neighbouring states in the Ornstein-Uhlenbeck process. This allows the model to better capture the transmission direction and related structure of epidemics. In contrast, white Gaussian noise has a correlation coefficient close to 0, which does not reflect the correlation of infectious diseases.

At the same time, there are many interacting variables in the environment that affect the infectious disease system, and these variables change in a continuous manner. In contrast to white Gaussian noise, the Ornstein-Uhlenbeck process is continuous (namely, the sample paths are continuous functions), which makes the model more realistic and interpretable and better reflects the continuous nature of infectious diseases. In summary of the comparison, we choose to introduce the Ornstein-Uhlenbeck process instead of white Gaussian noise.

Combining with the Eqs (1) and (3), the stochastic model is as follows: (5)

Assume that is a complete probability space with a filtration {Ft}t≥0 that meets the usual conditions (it is right continuous and F0 includes all -null sets).

The paper is organised below. In Section 2, the existence of a unique positive solution for the system Eq (5) is proved. We obtain the condition that determines whether the disease will become extinct or persistent in Section 3. When , the disease may persist; when , the disease may become extinct. In Section 4, we attain the adequate requirement that can be prove the existence of the stationary distribution by composing appropriate Lyapunov functions. In Section 5, the precise expression for the density function of the linearised system correspondent to the stochastic system Eq (5) surrounding the quasi-equilibrium is derived. In Section 6, the validity of theoretical results will be demonstrated by means of numerical simulations.

2 Existence and uniqueness of the global solution

In analysing the dynamic behaviour of epidemic, considering whether the positive solution is global is a crucial step. In this section, we demonstrate the existence of the unique global positive solution.

Theorem 2.1. For any initial value , on t ≥ 0, there does exist a unique solution (S(t), I(t), V(t), m(t)) to system Eq (5) which stays with probability 1 in .

Proof. Due to the local Lipschitz continuity of the coefficients of model Eq (5), there exists a unique local solution on t ∈ [0, τe), where τe denotes the explosion time [35].

Next we verify that τe = ∞ a.s., with the aim of certificating that the solution is global. Make r0 > 0 sufficiently large so that S(0), I(0), V(0) and em(0) are in . For each integer r > r0, define the stopping time as

Here we define inf ∅ = ∞ in the paper where ∅ is the empty set. As r → ∞, τr increases monotonically. Make , hence ττe a.s.. When τ = ∞ a.s. is satisfied, then τe = ∞ a.s. and (S(t), a.s., on t ≥ 0.

Then, consider the paradox, that is τ < ∞ a.s., then there are the constants h > 0 and δ ∈ (0, 1) that make . Therefore, suppose that an integer r1r0 makes a.s., ∀rr1. Define the C2-function W1: as follows:

Using u − 1 ≥ ln u for all u > 0, it can be seen that the function is non-negative. From equation Eq (5), it is evident that the total population is regulated by then

Suppose that . For ∀rr0 and ∀h > 0, W1 uses the Itô’s formula to produce: (6) where

Here k1 > 0, and k1 is independent of the initial value. Substituting the above inequality into Eq (6) to obtain

Integrating from 0 to τrh and taking expectations, there is

Let Ωr = {τrh} for rr1, then we get , δ ∈ (0, 1). Notice that for every ω ∈ Ωr, S(τr, ω), I(τr, ω), V(τr, ω) and equals either r or . Therefore, where is the indicator function of Ωr. As r → ∞, we have which creates a paradox. It completes the proof.

Remark 2.1. From Theorem 2.1, there is a unique global solution (S(t), I(t), V(t), . Hence and

If , then a.s.. This suggests that the domain is a positive invariant set. Namely,

3 Extinction and persistence

We derive the requirements for when the illness becomes extinct and when it becomes endemic in this section. At the first, we give the definitions

From this, it follows that when m = 0, the threshold condition for disease extinction and persistence for model Eq (5) is the threshold R0 for deterministic model disease extinction and persistence.

Theorem 3.1. If holds, then

I(t) in system Eq (5) is exponentially approaching 0 a.s., that is, the infectiousness of the disease will disappear. Moreover,

Proof. Define the following functions:

By using Itô’s formula, we get where (7)

Integrating each sides of Eq (7) and dividing by t yields

Then, combining Eq (4) and taking limits, we get which hints (8)

From system Eq (5), it can be generated

Then, using Eq (8), we have (9)

From the third equation of system Eq (5), so then

Thus, from Eq (9), we have

The proof of the theory is complete.

Theorem 3.2. If holds, then

Proof. From the first line of Eq (7),

Integrating both sides, we have

Combining Eq (4), we have

This completes the proof.

Remark 3.1. By looking at the expression for , it is clear that can be inferred to be R0 < 1, which suggests that the condition for disease extinction in both deterministic and stochastic systems can be united as . Similarly, can be inferred to be R0 > 1, which suggests that can be viewed as a unifying threshold for the prevalence of disease in both deterministic and stochastic systems.

4 Stationary distribution

In deterministic infectious disease models, disease persistence is usually expressed as the stability of endemic equilibrium points. Since there is no endemic equilibrium in the stochastic system Eq (5), we examine the stationary distribution that represents the persistence of disease in this part.

Lemma 4.1. [3638] For any initial value Z(0) ∈ Γ, if there exists a bounded closed domain Uδ ∈ Γ with a regular boundary, where is the transition probability of Z(t). That is to say, the system has at least one ergodic stationary distribution.

Theorem 4.1. If holds, system Eq (5) exists a stationary distribution π(⋅).

Proof. Using Itô’s formula to system Eq (5), we can have

Denote and where Q is an adequately big and positive constant that fulfills the next inequality:

has the lowest value because , when (S, I, V, m) is near the boundary of Γ. We may therefore construct a non-negative function:

Employing Itô’s formula, we can obtain where

Then, define a closed subset of Uδ by

The following inequalities hold because δ is an adequately small constant.

Divide the complement of Uδ into 5 subsets:

This gives us the following outcomes.

Case 1. , then

Case 2. , then

Case 3. , then

Case 4. , then

Case 5. , then

Given the above five cases, there is a constant δ that causes

In addition, assume that R > 0 such that

For simplicity, we denote Z(t) = (S(t), I(t), V(t), m(t)). For the arbitrary initial value Z(0) ∈ Γ, we can have

By taking the infimum bound for the inequality mentioned above and then associating the results with Eq (4), we get where and are the indicator functions of the set {Z(τ) ∈ Uδ} and {Z(τ) ∈ Γ\Uδ}. This suggests that

Therefore,

Since Γ is invariant under the system Eq (5), the solution (S(t), I(t), V(t), m(t)) is analysed on Γ. Based on the results in Ref. [37] and from the invariance of Γ and the above inequality, it can be inferred that there exists an invariant probability measure on Γ. Therefore, when , there exists a stationary distribution on Γ for system Eq (5).

5 Density function

It is important to note that the probability density function can reveal a variety of characteristics of the disease dynamics. By transforming the stochastic model Eq (5) into a linearized system, we work on solving for an accurate expression for the probability density function surrounding the quasi-endemic equilibrium. If holds, there exists a quasi-endemic equilibrium satisfying the following equations:

When stochastic factors are not taken into account in the model, the deterministic model Eq (2) has the identical endemic equilibrium as the quasi-endemic equilibrium. Let u1 = SS*, u2 = II*, u3 = VV*, u4 = mm*, and we obtain the linearized system (10) where

Theorem 5.1. If , the solution (u1, u2, u3, u4) of system Eq (10) obeys the normal probability density function Φ(u1, u2, u3, u4), and the form of Φ(u1, u2, u3, u4) is as follows: where and where

Proof. Firstly, let du = Audt + GdB(t), the matrix form of system Eq (10) is obtained:

By calculating the polynomial equation for the eigenvalues λ of the matrix A, one obtains where b1 = a11 + a33, b2 = a12a21 + a11a33a13a31, b3 = a12a21a33. The characteristic roots of φA(λ) can be obtained from λ1 = −a44 and λ3 + b1λ2 + b2λ + b3 = 0. Then the characteristic roots all contain negative real parts as a result of b1 > 0, b2 > 0, b3 > 0 and b1b2b3 > 0. It is evident that A is a Hurwitz matrix. According to Ref. [39], the density function Φ(u1, u2, u3, u4) of system Eq (10) can be represented as the following Fokker-Planck equation: which may be approximately expressed as the following Gaussian distribution:

Here D is a real symmetric matrix fulfilling DG2D + ATD + DA = 0. Then suppose that it is positive definite and D−1 = Σ, we have (11)

Then prove that Σ is positive definite. Let then

Next, let we get

Note that by calculations, we have where a1 = −a44, a2 = −a14, a3 = −a11 + a12, a4 = −a11 + a12 + a21, a5 = −a12 , . The transformation matrix M1 is obtained by following the procedure in Ref. [40], where , . Then, by means of the standard R1 transformation matrix, one obtains where c1 = b1 + a44, c2 = b2 + a44b1, c3 = b3 + a44b2, c4 = a44b3. Then, Eq (11) may be changed into the form shown below: which can be expressed as where . Using the Lemma 2.3 in Ref. [40], the form of Σ1 can be expressed as follows:

Moreover, from b1 > 0, b2 > 0, b3 > 0, a44 > 0 and b1b2b3 > 0, it can be inferred that c1 > 0, c3 > 0, c4 > 0 and . As a result, the matrix Σ1 is positive definite. It is possible to determine the precise expression of Σ, which is the positive definite matrix Σ = (m1θ)2(M1J3J2J1)−1Σ1[(M1J3J2J1)−1]T. This proof is complete.

Remark 5.1. By Theorem 5.1, we know that the solution (S(t), I(t), V(t)) of the system Eq (5) obeys the normal density function . Here, we define Σ(3) as the third-order principal minor of Σ. Hence, S(t), I(t) and V(t) will each converge to the marginal density functions: where is the element in row i, column i on Σ. Namely, S(t), I(t) and V(t) will converge to the marginal distributions , and , respectively.

6 Numerical simulation

In this section, various numerical simulations are provided in order to verify the scientific results. Next, by employing the Milstein method to numerically simulate the model, the correctness of the conclusions is demonstrated. The model Eq (5) is discretized, and the matching discretization model is obtained as follows:

Here Δt is the time interval and takes the value of Δt = 1, and hi is a random variable that follows a standard normal distribution.

Example 6.1 Assume the parameters k = 0.65, θ = 0.14 and the starting points in the below examples are all (S(0), I(0), V(0), m(0)) = (0.04, 0.8, 0.03, −0.02). Other parameters are as in Group 1 of Table 1. Note that makes the condition of Theorem 3.1 hold. This shows that I(t) has an exponential tendency to go to zero with probability 1. Then, as evidence for our findings, we present Fig 1.

thumbnail
Fig 1. Computer simulations of S(t), I(t) and V(t) for system Eq (5) in case of disease extinction.

https://doi.org/10.1371/journal.pone.0310175.g001

Example 6.2 Assume the parameters k = 0.7, θ = 0.1, and other parameters are as in Group 2 of Table 1. Note that satisfies the requirement of Theorem 3.2. In other words, the illness is going to prevail. Then, as evidence for our findings, we present Fig 2.

thumbnail
Fig 2. Computer simulations of S(t), I(t) and V(t) for system Eq (5) in the presence of disease persistence.

https://doi.org/10.1371/journal.pone.0310175.g002

Example 6.3 Assume that the parameters in group (a) take the same values as in Example 6.1 and that the parameters in group (b) take the same values as in Example 6.2. Through 10,000 stochastic simulations, we can obtain Fig 3 representing the expectation and standard deviation of S(t), I(t) and V(t). It shows that the disease will quickly die out in case (a), while the disease will become endemic in case (b).

thumbnail
Fig 3. Computer simulations of expectation and standard deviation for the model Eq (5).

https://doi.org/10.1371/journal.pone.0310175.g003

Example 6.4 It is assumed that these parameters are selected with the identical values as in Example 6.2. Based on 10,000 stochastic simulations, the ergodic stationary distribution is proved in Fig 4. It is obviously that the value on the histogram surround P* = (S*, I*, V*) ≈ (0.6111, 0.7333, 2.5555) of the deterministic model. The solution (S(t), I(t), V(t), m(t)) of system Eq (5) obeys the normal density function . The matrix Σ is represented as from which the following three marginal density functions are inferred

thumbnail
Fig 4. Computer simulations of the histograms of frequencies for the model Eq (5).

https://doi.org/10.1371/journal.pone.0310175.g004

Example 6.5 Consider the corresponding discretized deterministic SIVS model

Assume that the parameters take the same values as in Example 6.2. The mean-reverting process is substituted into the deterministic SIVS model to transform it into the stochastic SIVS model. Then we draw a comparison chart between deterministic SIVS model and stochastic SIVS model in Fig 5. It shows that the stochastic SIVS model that includes the mean-reverting process fluctuates around the deterministic model.

thumbnail
Fig 5. Comparison plot between the stochastic model Eq (5) with the mean-reverting process and deterministic model Eq (1).

https://doi.org/10.1371/journal.pone.0310175.g005

Example 6.6 Assume the parameters θ = 0.1, and other parameters are as in Group 2 of Table 1. To study the effect of k size on disease progression. As can be seen in Fig 6, the fluctuation of the disease becomes smaller as k increases.

thumbnail
Fig 6. Computer simulations for different values of k for the model Eq (5).

https://doi.org/10.1371/journal.pone.0310175.g006

Example 6.7 Assume the parameters k = 0.7, and other parameters are as in Group 2 of Table 1. To study the effect of θ size on disease progression. As can be seen in Fig 7, the fluctuation of the disease becomes larger as θ increases.

thumbnail
Fig 7. Computer simulations for different values of θ for the model Eq (5).

https://doi.org/10.1371/journal.pone.0310175.g007

7 Conclusion

We demonstrate necessity and rationale of introducing the Ornstein-Uhlenbeck process by contrasting it with white Gaussian noise. Considering the theoretical results obtained in Ref. [11], we concentrate on a stochastic model with the mean-reverting process. Instead of the typical linear function with white noise, the parameter β is designed to fulfill the mean-reversion process. By doing this, the issue that may become infinite as the time interval becomes smaller is avoided. In this paper, in the situation where the unique global positive solution to the system Eq (5) exists, the conditions for disappearance and persistence of the illness are gained, i.e., when , the illness tends to extinction; when , the illness becomes epidemic. Further, the system Eq (5) exists a stationary distribution when , indicating that the illness will likely end up becoming an epidemic that lasts for a long time. To further study the infectious disease dynamics, by resolving the related equations, the probability density function surrounding the quasi-endemic equilibrium is explored, which reveals many dynamical features. The final step is to demonstrate theoretical results through numerical simulation. Through numerical simulations, we find an interesting conclusion: smaller regression rates or larger fluctuation intensities make the stochastic system more volatile.

In addition, a uniform threshold for disease extinction and prevalence in stochastic systems remains difficult to obtain due to limitations of existing mathematical methods. We expect that this problem can be solved in future studies.

Acknowledgments

Thanks to the reviewers for their helpful comments and suggestions.

References

  1. 1. Zhao YN, Jiang DQ. The threshold of a stochastic SIS epidemic model with vaccination. Appl Math Comput. 2014;243: 718–727.
  2. 2. Zhang XH, Jiang DQ, Hayat T, Ahmad B. Dynamical behavior of a stochastic SVIR epidemic model with vaccination. Phys A. 2017;483: 94–108.
  3. 3. Zhang SH, Xu R. Global stability of an SIS epidemic model with age of vaccination. Differ Equ Dyn Syst. 2022;30: 1–22.
  4. 4. Wang LL, Xu R. Global stability of an SEIR epidemic model with vaccination. Int J Biomath. 2016;9: 1650082.
  5. 5. Wang XJ, Liang Y, Li JH, Liu MX. Modeling COVID-19 transmission dynamics incorporating media coverage and vaccination. Math Biosci Eng. 2023;20: 10392–10403. pmid:37322938
  6. 6. Gan SQ, Wei FY. Study on a susceptible-infected-vaccinated model with delay and proportional vaccination. Int J Biomath. 2018;11: 1850102.
  7. 7. Liu SY, Shen MW, Bi YJ. Global asymptotic behavior for mixed vaccination strategy in a delayed epidemic model with interim-immune. Math Biosci Eng. 2020;17: 3601–3617. pmid:32987546
  8. 8. Zhao X, Zhou WS. Global stability of the deterministic and stochastic SIS epidemic models with vaccination. J Math Resea Appl. 2021;41: 62–68.
  9. 9. Zhao YN, Zhang QM, Jiang DQ. The asymptotic behavior of a stochastic SIS epidemic model with vaccination. Adv Differ Equ. 2015;2015(1): 328.
  10. 10. Li JQ, Ma ZE. Qualitative analyses of SIS epidemic model with vaccination and varying total population size. Math Comput Modelling. 2002;35: 1235–1243.
  11. 11. Li JQ, Ma ZE. Global analysis of SIS epidemic models with variable total population size. Math Comput Modelling. 2004;39: 1231–1242.
  12. 12. Beddington JR, May RM. Harvesting natural populations in a randomly fluctuating environment. Science. 1977;197: 463–465. pmid:17783245
  13. 13. Huo HF, Cui FF, Xiang H. Dynamics of a SAITS alcoholism model on unweighted and weighted networks. Phys A. 2018;496: 249–262.
  14. 14. Ma JY, Lin W. Dynamics of a stochastic COVID-19 epidemic model considering asymptomatic and isolated infected individuals. Math Biosci Eng. 2022;19: 5169–5189. pmid:35430859
  15. 15. Wang WM, Cai YL, Li JL, Gui ZJ. Periodic behavior in a FIV model with seasonality as well as environment fluctuations. J Frankl Inst. 2017;354: 7410–7428.
  16. 16. Yu XW, Yuan SL, Zhang TH. The effects of toxin-producing phytoplankton and environmental fluctuations on the planktonic blooms. Nonlinear Dyn. 2018;91: 1653–1668.
  17. 17. Zhou YL, Zhang WG, Yuan SL. Survival and stationary distribution in a stochastic SIS model. Discr Dyn Nat Soc. 2013;2013: 1–12.
  18. 18. Liu Q, Jiang DQ, Shi NZ. Threshold behavior in a stochastic SIQR epidemic model with standard incidence and regime switching. Appl Math Comput. 2018;316: 310–325.
  19. 19. Teng ZD, Wang L. Persistence and extinction for a class of stochastic SIS epidemic models with nonlinear incidence rate. Phys A. 2016;451: 507–518.
  20. 20. Zhao YN, Jiang DQ, O’Regan D. The extinction and persistence of the stochastic SIS epidemic model with vaccination. Phys A. 2013;392: 4916–4927.
  21. 21. Allen E. Environmental variability and mean-reverting processes. Discr Cont Dyn B. 2016;21: 2073–2089.
  22. 22. Luo J, Lin YC, Wang SJ. Intraday high-frequency pairs trading strategies for energy futures: evidence from China. Applied Economics. 2023;55(56): 6646–6660.
  23. 23. Wang Y, Xu X, Zhang JZ. Optimal investment strategy for DC pension plan with stochastic income and inflation risk under the Ornstein-Uhlenbeck model. Mathematics. 2021;9(15): 1756.
  24. 24. Habtemicael S, SenGupta I. Ornstein-Uhlenbeck processes for geophysical data analysis. Phys A. 2014;399: 147–156.
  25. 25. Zhou YX, Jiang DQ. Dynamical behavior of a stochastic SIQR epidemic model with Ornstein-Uhlenbeck process and standard incidence rate after dimensionality reduction. Commun Nonlinear Sci Numer Simul. 2023;116: 106878.
  26. 26. Shang JX, Li WH. Aynamical behaviors of a stochastic SIRV epidemic model with the Ornstein-Uhlenbeck process. Adv Cont Discr Mod. 2024;2024(1): 9.
  27. 27. Zhou Y, Yuan SL, Ma JL. Survival and stationary distribution analysis of a stochastic competitive model of three species in a polluted environment. Bull Math Biol. 2015;77: 1285–1326.
  28. 28. Zhang ZY, Liang GZ, Chang KK. Stationary distribution of a reaction-diffusion hepatitis B virus infection model driven by the Ornstein-Uhlenbeck process. PLoS ONE. 2023;18(9): e0292073. pmid:37773949
  29. 29. Wang WM, Cai YL, Ding ZQ, Gui ZJ. A stochastic differential equation SIS epidemic model incorporating Ornstein-Uhlenbeck processes. Phys A. 2018;509: 921–936.
  30. 30. Shi ZF, Jiang DQ. Dynamical behaviors of a stochastic HTLV-I infection model with general infection form and Ornstein-Uhlenbeck process. Chaos Solitons & Fractals. 2022;165: 112789.
  31. 31. Cai YL, Kang Y, Banerjee M, Wang WM. A stochastic SIRS epidemic model with infectious force under intervention strategies. J Differ Equ. 2015;259: 7463–7502.
  32. 32. Zhang XF, Yuan R. A stochastic chemostat model with mean-reverting Ornstein-Uhlenbeck process and Monod-Haldane response function. Appl Math Comput. 2021;394: 125833.
  33. 33. Lin YG, Jiang DQ, Xia PY. Long-time behavior of a stochastic SIR model. Appl Math Comput. 2014;236: 1–9.
  34. 34. Kutoyants Y. A. Statistical inference for ergodic diffusion processes. London: Springer London; 2004.
  35. 35. Mao XR. Stochastic differential equations and applications. Chichester: Horwood Publishing; 1997.
  36. 36. Du NH, Nguyen DH, Yin GG. Conditions for permanence and ergodicity of certain stochastic predator–prey models. J Appl Probab. 2016;53(1): 187–202.
  37. 37. Meyn SP, Tweedie RL. Stability of Markovian processes III: Foster-Lyapunov criteria for continuous-time processes. Adv Appl Prob. 1993;25: 518–548.
  38. 38. Dieu NT. Asymptotic Properties of a Stochastic SIR Epidemic Model with Beddington-DeAngelis Incidence Rate. J Dyn Diff Equat. 2018;30(1):93–106.
  39. 39. Roozen H. An asymptotic solution to a two-dimensional exit problem arising in population dynamics. SIAM J Appl Math. 1989;49: 1793–1810.
  40. 40. Zhou BQ, Jiang DQ, Dai YC, Hayat T. Stationary distribution and density function expression for a stochastic SIQRS epidemic model with temporary immunity. Nonlinear Dyn. 2021;105: 931–955. pmid:34121810