1.
Introduction
Cholera, a pervasive endemic disease, poses substantial risks to global public health, resulting in significant morbidity and mortality. The World Health Organization (WHO) reported that cholera incidence reached approximately 3.1 million cases and 95,000 fatalities in 2022, marking a 145% increase relative to the average of the preceding five years. Transmission of cholera primarily occurs through the ingestion of the pathogen, characterizing it as both a waterborne and foodborne disease. The consumption of water contaminated with sewage along with the ingestion of victuals prepared in unsanitary conditions can significantly facilitate the transmission of the pathogen. Additionally, direct interpersonal contact with infected individuals can transmit the pathogen to susceptible hosts. High-risk individuals carrying the pathogen may spread cholera to their family members through close contact. Individuals who have recovered from cholera may develop a temporary acquired immunity to the pathogen. However, recent studies indicate that this acquired immunity may wane within a few months or weeks. These findings emphasize the critical necessity for ongoing research into the mechanisms of cholera transmission and the development of effective control measures. Various epidemiological frameworks and models have been extensively studied [1,2,3,4]. Gui [5] developed a four-dimensional epidemiological model for the dynamics of cholera transmission, as follows:
where N(S+I+R=N) represents the total population of China. The population is separated into three groups: individuals who are susceptible (S), infected (I), and recovered (R). In addition to these demographic groups, it is important to consider the potential implications of disinfection in controlling cholera, which is closely related to the concentration of vibrios in contaminated water (denoted by state B). In addition, Table 1 provides further information on parameters.
In system (1.1), the third equation is independent of the others, indicating that we only need to study the dynamics of the following subsystems
According to [5], the basic reproduction number is obtained by R0=βhμN(μ+ν)(γ+μ)+ˉβeμNξ(μ+ν)(γ+μ)(δ+c)κ. Moreover, the relevant threshold dynamics of system (1.2) is as follows:
∙ If R0<1, system (1.2) has a disease-free equilibrium E0=(μNμ+ν,0,0), which is globally asymptotically stable.
∙ If R0>1, there exists a unique endemic equilibrium E+=(S+,I+,B+)=(μNξ−(γ+μ)(δ+c)B+(μ+ν)ξ,(δ+c)B+ξ,B+) and it is globally asymptotically stable, where B+ is the unique positive root of the following quadratic equation
and μNξ>(γ+μ)(δ+c)B+. Since it has been proven that the stochastic model can more accurately explain biological processes and infectious illnesses, there is increasing scholarly interest in examining the impact of environmental disturbances on epidemic models [6,7,8,9,10]. As a result, the development and research into stochastic models have intensified. For example, Jiang et al. [11] examined stationary distributions and extinction in non-autonomous logistic equations with random perturbations. In addition, several studies [12,13,14,15] have yielded notable conclusions in this field. One of the most important factors in the epidemic model (1.2) is βe, which is always fluctuating around the average value ˉβe owing to the continuous spectrum of environmental noise. In this sense, βe should be considered a random variable. We assume that βe(t) is an Ornstein–Uhlenbeck process and satisfies the following form to imitate environmental noise's effect on transmission rates:
where α and σ are positive constants indicating the speed of reversion and intensity of volatility, respectively, and ˉβe is a positive constant representing the long-term average transmission rate βe. W(t) is a standard Brownian motion defined on a complete probability space (Ω,F,{Ft}t≥0,P) with a filtration {Ft}t≥0. According to Mao's monograph [16], we can obtain that
The expectation and variance of βe(t) are
As a result, the limit distribution of the Ornstein–Uhlenbeck process βe(t) is N(ˉβe,σ22α). In other words, the probability density of the limit distribution is π(x)=√α√πσe−α(x−ˉβe)2σ2. Furthermore, it is straightforward to deduce that limt→0+E[βe(t)]=βe(0) and limt→0+Var[βe(t)]=0. This result suggests the Ornstein–Uhlenbeck process could be more suitable for describing random perturbations. Moreover, we let β+e(t):=max{β(t),0}, since the transmission rate coefficient should be non-negative. Thus, we obtain the following stochastic model:
In addition, we obtain
Considering the third equation of system (1.3), we have
Accordingly, region
is positively invariant set with respect to model (1.3). As a result, we take the supposition that initial values fulfill S(0)+I(0)<N,B(0)<ξNc, throughout the entire paper. We summarize our main contributions and innovations in comparison with the existing literature as follows: (ⅰ) To obtain the existence of a stationary distribution, which is a probability distribution with some invariant properties, we develop an innovative approach to build some stochastic Lyapunov functions. (ⅱ) During the following discussion, we explore the sufficient conditions for the existence of stationary distribution. The probability density function corresponding to the quasi-equilibrium is established, which is of great significance. (ⅲ) We also study the sufficient conditions for the disease to exterminate exponentially. In summary, the remainder of our paper is structured as follows. The existence and uniqueness of a global solution to system (1.3), as well as other relevant mathematical Lemma, are all presented in Section 2. The sufficient criteria for an ergodic stationary distribution of system (1.3) are derived in Section 3. To be complete, we build sufficient conditions for eliminating viruses and infected cells in Section 4. Lastly, numerical simulations are carried out to demonstrate our theoretical findings.
2.
Preliminaries
An important lemma is introduced in this section to obtain the accurate probability density expression.
Lemma 2.1. [17] For the algebraic equation H20+A0Σ0+Σ0AT0=0, where H0=diag(1,0,0,0), Σ0 is a real symmetric matrix and the standard matrix
If η1>0,η3>0,η4>0 and η1η2η3−η23−η21η4>0, then Σ0 is a positive definite matrix, where
Here, A0 in this form is called the standard R1 matrix.
Theorem 2.1. The system (1.3) has a unique global solution (S(t),I(t),B(t),βe(t)) that will remain in Γ with probability one (a.s.) for any initial value of (S(0),I(0),B(0),βe(0))∈Γ.
Proof. To prove the existence and uniqueness of positive solutions to stochastic models, we usually rely on the standard approach and the classical Khasminskii Lyapunov functional method, which is similar to [18]. Constructing the appropriate Lyapunov function is crucial. We construct a non-negativity C2-function U0(S,I,B,βe) as follows
Since u−1−lnu≥0 for any u>0, then the above function is non-negativity. Applying Itô's formula to U0, we have
where W0 is a positive constant, which is independent of S,I,B and βe.
3.
Stationary distribution
The aim of this section is to analyze whether the stochastic model has a stationary distribution that reflects infectious disease persistence. Define
where ˆβe=(∫∞0x14π(x)dx)4,π(x)=√α√πσe−α(x−ˉβe)2σ2, a1=βhμN(μ+ν)2,b1=μNξˆβe(μ+ν)2(δ+c)κ.
Theorem 3.1. Suppose that Rs0>1, then the stochastic system (1.3) admits at least one ergodic stationary distribution η(⋅) on Γ.
Proof. Using Itô's formula to −lnS,−lnIand−lnB, respectively, we have
Then, define
where positive constants a1,b1,b2,b3 will be approved later. Using Ito's formula on U1 and combining (3.1), we obtain
where
Note that β+e=⏐βe⏐+βe2, we have
Choose
Substituting (3.3) and (3.4) into (3.2), we have
where
Then, define
Making the use of Itô's formula to U2, we have
where
Next, we define
Combining (3.1) and applying Itô's formula to U3, we have
where
Then, we define
where M0 is a sufficiently large constant satisfying
Then, from (3.6)–(3.8), we have
where
Next, we construct a compact set Dε⊂Γ as follows
such that g3(S,I,B,βe)≤−1for any(S,I,B,βe)∈Γ∖Dε:=Dcε. Then, let Dcε=6⋃i=1Dcε,i, where
ε is a small enough constant satisfying the following inequalities
Case 1. If (S,I,B,βe)∈Dcε,1, from (3.10) and (3.11), we have
Case 2. If (S,I,B,βe)∈Dcε,2, from (3.10) and (3.12), we obtain
Case 3. If (S,I,B,βe)∈Dcε,3, from (3.10) and (3.12), we derive
Case 4. If (S,I,B,βe)∈Dcε,4, from (3.10) and (3.12), we get
Case 5. If (S,I,B,βe)∈Dcε,5, from (3.10) and (3.12), we have
Case 6. If (S,I,B,βe)∈Dcε,6, from (3.10) and (3.12), we obtain
In summary, we have
Let
Then, we have
The function U4 has the minimum value U4(S0,I0,B0,β0e), since it tends to +∞ as (S,I,B,βe) approaches the boundary of Γ. Thus, we obtain a non-negative function
Then, applying Itô's formula to U, we have
For any initial value (S(0),I(0),B(0),βe(0))∈Γ and a interval [0,t], using the Itô's integral and then taking mathematical expectation to U, we get
Given the ergodicity of βe(t) and the strong law of large numbers, we have
Taking the inferior limit on both sides of (3.13) and combining with (3.12), (3.14) and (3.15), we get that
Therefore, we have
Making use of Fatou's lemma [19], we have
where P(t,(S,I,B,βe,A) represents the transition probability of (S(t),I(t),B(t),βe(t))T belonging to the set A. According to the Lemma 2.1 in [20], system (1.3) has at least one stationary stochastic distribution when Rs0>1.
4.
Probability density function
The purpose of this section is to derive the explicit expression of probability density function around the quasi-endemic equilibrium based on the corresponding four-dimensional matrix equation.
The quasi-endemic equilibrium P∗=(S∗,I∗,B∗,β∗e) is the solution of the following equations.
By direct calculation, if R0>1, the solution of the above equation is unique, and it is
where S+,I+and B+ are the same as Section 1.
Letting (z1,z2,z3,z4)T=(S−S∗,I−I∗,B−B∗,S−S∗,ˉβe−β∗e)T, system (1.3) can be linearized as follows:
where
Letting
and
then, linearized Eq (4.2) can be expressed by matrix
Based on the continuous Markov processes in [21], system (1.3) has a unique probability density function Φ(z1,z2,z3,z4) surrounding the quasi-endemic equilibrium; according to Fokker–Plank equation, its form is as follows:
As a result, Φ(z1,z2,z3,z4) can be expressed by a quasi-Gaussian distribution since the diffusion matrix G is constant. Therefore
where q is a constant to ensure that the normalized condition is established ∫R4Φ(z1,z2,z3,z4)dz1dz2dz3dz4=1. Then, PG2P+ATP+PA=0 is satisfied by the real symmetric matrix P. If the matrix P is inverse, which denotes P−1=Σ, it can be equivalently changed into
To find the exact expression of the probability density function Φ(z1,z2,z3,z4), we will solve Eq (4.3).
Theorem 4.1. Assuming that if R0>1, then for any initial value (S(0),I(0),B(0),βe(0))∈Γ, the stationary distribution of system (1.3) around P∗ has a normal density function as follows:
where Σ is a positive definite matrix that takes the form
Here,
where ϱ=−σa14m1(a12+a21+a22−a11), m1=a32(a21+a33−a11)a12+a21+a22−a11, m2=m1[a12(a11−a21+a22+a33)+a22(a22+a33)+a13a32+a233], m3=−a13(a12+a21+a22−a11)m1−a333, r1=−a14m1(a12+a21+a22−a11), r2=m1(a12+a22+a21−a11)(−a11−a22−a33), η1=H1+a44,η2=H2+H1a44,η3=H3+H2a44, η4=H3a44,H1=a11+a22+a33, H2=a11(a22+a33)+a12a21+(a22a33−a13a32), H3=a33(a11a22+a12a21)+a32a13(a21−a11).
Proof. The first step we need is to confirm that A is a Hurwitz matrix. Let φA(λ) be the characteristic polynomial of matrix A, which can be written as follows:
where H1=a11+a22+a33,H2=a11(a22+a33)+a12a21+(a22a33−a13a32), H3=a33(a11a22+a12a21)+a32a13(a21−a11). The characteristic roots of φA(λ) can be derived by: λ1=−a44 and λ3+H1λ2+H2λ+H3=0. By calculations, we can obtain
which yields that H1>0,H2>0,H3>0, and H1H2−H3>0. This implies that all the roots of the characteristic equation (4.4) have negative real parts and the matrix A is a Hurwitz matrix.
Next, by solving the equation G2+AΣ+ΣAT=0, we shall find the form of Σ and demonstrate that it is positive definite.
Let A1=J1AJ−11, where the ordering matrix J1 is given by
Then,
Define A2=J2A1J−12, where the elimination J2 takes the form
then, we have
Next, let A3=J3A2J−13, where
by simple calculation, then
where m1=a32(a21+a33−a11)a12+a21+a22−a11.
Let A4=J4A3J−14, where the standardized transform matrix J_4 takes the form
where r_1 = -a_{14}m_1(a_{12}+a_{21}+a_{22}-a_{11}), r_2 = m_1(a_{12}+a_{22}+a_{21}-a_{11})(-a_{11}-a_{22}-a_{33}), m_2 = m_1[a_{12}(a_{11}-a_{21}+a_{22}+a_{33})+a_{22}(a_{22}+a_{33})+a_{13}a_{32}+a_{33}^2], m_3 = -a_{13}(a_{12}+a_{21}+a_{22}-a_{11})m_1-a_{33}^3.
By direct calculation, we obtain
This characteristic polynomial is also invariant according to the invariant of matrix elementary transformations. Consequently, we can obtain
Additionally, Eq (4.3) is equivalently convertible into the following form
where G_0 = diag(1, 0, 0, 0), \, \Sigma_0 = \varrho^{-2}(J_4J_3J_2J_1)\Sigma(J_4J_3J_2J_1)^T, \varrho = -\sigma a_{14}m_1(a_{12}+a_{21}+a_{22}-a_{11}).
Using Lemma 2.1, the form of \Sigma_0 can be given as
More importantly, from H_1 > 0, \, H_2 > 0, \, H_3 > 0, \, a_{44} > 0 and H_1H_2-H_3 > 0, we can deduce that \eta_1 > 0, \, \eta_3 > 0, \, \eta_4 > 0 and \eta_1(\eta_2\eta_3-\eta_1\eta_4)-\eta_3^2 = (H_1+a_{44})(H_1H_2-H_3)a_{44}^2+H_3(H_1H_2-H_3)+H_2(H_1H_2-H_3)a_{44} > 0. So, the matrix \Sigma_0 is a positive definite matrix. Therefore, the matrix \Sigma = \varrho^2(J_4J_3J_2J_1)^{-1}\Sigma_0[(J_4J_3J_2J_1)^{-1}]^T is also positive definite. As a result, the density function around quasi-endemic equilibrium P^* = (S^*, I^*, B^*, \beta^*) is as follows
5.
Extinction
In this section, we explore the condition of disease extinction. Defining
Theorem 5.1. If R_0^E < 1, then the disease of system (1.3) will exponentially die out a.s.
Proof. We define a C^2 -function F as follows
where l_1 = (1+\frac{\nu}{\mu})R_0 \, \text{and}\, l_2 = \frac{\bar{\beta}_e N}{\kappa(\delta+c)}. By using the Itô's formula to \ln F , we obtain
By integrating (5.1) from 0 to t and dividing by t on both sides, we get
Combining (3.14) and letting t\rightarrow +\infty, we know that
Therefore, if R_0^E < 1, then we have \limsup _{t\rightarrow +\infty}\frac{\ln F(t)}{t}\leq \min\bigg\{\delta+c, \frac{\beta_hN\mu}{(\mu+\nu)R_0}\bigg\}(R_0^E-1) < 0, which indicates \lim_{t\rightarrow +\infty}I(t) = 0 and \lim_{t\rightarrow +\infty}B(t) = 0. This suggests that system (1.3) sickness will disappear exponentially.
6.
Numerical simulation
Our analytical results are illustrated in this section through numerical simulations. Assuming x(t) = \beta_e(t)-\bar{\beta}_e , the discretization equation for system (1.3) is:
where the time increment \triangle t > 0, \, \, \zeta_j are random variables and satisfy Gaussian distribution N(0, 1) for j = 1, 2\ldots, n.
Example 6.1. Let (S(0), I(0), B(0), x(0)) = (1.4008\times10^8, 3.944\times10^7, 1.088\times10^8, 1.36\times10^8), \bar{\beta}_e = 5.35\times10^{-2}, \, \beta_h = 2.05\times 10^{-9}, \, \nu = 0.04/365 and the other parameter values are presented in Table 1. Direct computation leads to that (S^*, I^*, B^*, \beta_e^*) = (1.3704\times10^8, 3.8996\times10^7, 9.6683\times10^7, 5.35\times10^{-2}) and
From Figure 1, it can be seen that system (1.3) has a stationary distribution.
Therefore, the solution (S(t), I(t), B(t), \beta_e(t)) of system (1.3) follows the normal density function \Phi(S, I, B, \beta_e)\sim N\big((S^*, I^*, B^*, \beta_e^*)^T, \Sigma\big) = \mathbb{N }\big((1.3704\times10^8, 3.8996\times10^7, 9.6683\times10^7, 5.35\times10^{-2})^T, \Sigma\big).
From Figure 2, we can find that the marginal density functions basically coincide with the corresponding fitting curves.
Example 6.2. Choose \bar{\beta}_e = 3.1\times 10^{-4}, \beta_h = 1.32\times 10^{-5}, \nu = 0.31017/365, \sigma = 0.6, \alpha = 4, and the rest of the parameters unchanged, by direct calculation leads to R_0^E = (1+\frac{\nu}{\mu})R_0+\frac{R_0(\mu+\nu)(\delta+c)\sigma}{\mu\bar{\beta}_e\sqrt{\pi\alpha}\min\bigg\{\delta+c, \frac{\beta_h\mu N}{(\mu+\nu)R_0}\bigg\}} = 0.8489 < 1. Thus, from Theorem 5.1, the disease of the system (1.3) will be extinct exponentially in a long time, which is supported by Figure 3.
7.
Conclusions
In this paper, on the basis of both biological significance and mathematically reasonable hypotheses, we illustrate that the Ornstein–Uhlenbeck process has a more stable variability than the linear perturbation and nonlinear perturbation. In this sense, we establish and analyze a stochastic cholera model with Ornstein–Uhlenbeck process to describe the propagation mechanism of cholera disease in the population. It is proved that the stochastic model has a unique global positive solution. We derive two critical conditions R_0^s and R_0^E; the system has at least one stationary distribution if R_0^s > 1; the virus will be cleared out if R_0^E < 1. The probability density function near the quasi-positive equilibrium is obtained by solving the corresponding Fokker–Planck equation. Mathematically, the existence of a stationary distribution implies the weak stability in the viewpoint of stochastic process. Biologically, the existence of a stationary distribution and probability density indicates the persistence of the disease. Finally, numerical simulations are used to verify our theoretical results.
Author contributions
Ying He: Conceptualization, Investigation, Formal analysis, Writing – review and editing. Bo Bi: Formal analysis, Writing – review and editing, Numerical simulation. All authors have read and approved the final version of the manuscript for publication.
Use of AI tools declaration
The authors declare they have not used Artificial Intelligence (AI) tools in the creation of this article.
Acknowledgments
This work is supported by Hainan Provincial Natural Science Foundation of China (No. 122RC679,121RC554), Talent Program of Hainan Medical University (No. XRC2020030) and Northeast Petroleum University Special Research Team Project (No. 2022TSTD-05).
Conflict of interest
The authors declare there is no conflict of interest.