Skip to main content
Log in

Threshold Dynamics and Probability Density Function of a Stochastic Multi-Strain Coinfection Model with Amplification and Vaccination

  • Published:
Qualitative Theory of Dynamical Systems Aims and scope Submit manuscript

Abstract

Multi-strain infectious diseases, which are usually prevented from spreading widely by vaccination, have two main transmission mechanisms: competitive exclusion and co-existence. In this paper, a stochastic multi-strain coinfection model with amplification and vaccination is developed. For the deterministic model, the basic reproduction number \(R_0\) and fixed points are provided. For the stochastic model, we first prove the existence and uniqueness of the positive solution under any initial value. Then, a portion of those infected with the common strain will always become patients with the amplified strain, which increases the risk of death from the disease. Therefore, we verified that patients with common strains would become extinct if \(R_1^s<1\). Furthermore, by constructing the Lyapunov function, we find that model (3) has a unique ergodic stationary distribution if \(R_0^S>1\). Particularly, we get a concrete form of the probability density of the distribution \(\kappa (\cdot )\) near equilibrium \(E^*\), where \(E^*\) is the quasi-local equilibrium of the stochastic model. Finally, the results are verified by numerical simulation. The results show that vaccination can control disease outbreaks or even eliminate them.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8

Similar content being viewed by others

References

  1. Overview of the Epidemic of National Statutory Infectious Diseases in 2021. National Health Commission

  2. Bonhoeffer, S., Nowak, M.A.: Mutation and the evolution of virulence. Proc. R. Soc. Lond. B. 258(1352), 133–140 (1994)

    Article  ADS  Google Scholar 

  3. Martcheva, M.: A non-autonomous multi-strain SIS epidemic model. J. Biol. Dyn. 3(2–3), 235–251 (2009)

    Article  MathSciNet  PubMed  Google Scholar 

  4. Kuddus, M.A., McBryde, E.S., Adekunle, A.I., White, L.J., Meehan, M.T.: Mathematical analysis of a two-strain disease model with amplification. Chaos Soliton Fract. 143, 110594 (2021) https://doi.org/10.1016/j.chaos.2020.110594

  5. He, D.H., Ali, S.T., Fan, G.H., Gao, D.Z., Song, H.T., Lou, Y.J., Zhao, S., Cowling, B.J., Stone, L.: Evaluation of effectiveness of global COVID-19 vaccination campaign. Emerg. Infect. Dis. 28(9), 1873–1876 (2022)

    Article  PubMed  PubMed Central  Google Scholar 

  6. Kaur, S.P., Gupta, V.: COVID-19 vaccine: a comprehensive status report. Virus Res. 288, 198114 (2020). https://doi.org/10.1016/j.virusres.2020.198114

  7. Yavuz, M., Coşar, F.Ö., Günay, F., Özdemir, F.N.: A new mathematical modeling of the COVID-19 pandemic including the vaccination campaign. Open J. Model. Simul. 9(3), 299–321 (2021)

    Article  Google Scholar 

  8. Wang, X.Y., Yang, J.Y., Han, Y.: Threshold dynamics of a chronological age and infection age structured cholera model with Neumann boundary condition. Z. Angew. Math. Phys. 74(4), 170 (2023)

    Article  MathSciNet  Google Scholar 

  9. Yang, J.Y., Yang, L., Jin, Z.: Optimal strategies of the age-specific vaccination and antiviral treatment against influenza. Chaos Soliton Fract. 168, 113199 (2023) https://doi.org/10.1016/j.chaos.2023.113199

  10. Saha, S., Samanta, G., Nieto, J.J.: Epidemic model of COVID-19 outbreak by inducing behavioural response in population. Nonlinear Dyn. 102, 455–487 (2020)

    Article  PubMed  PubMed Central  Google Scholar 

  11. Saha, S., Samanta, G., Nieto, J.J.: Impact of optimal vaccination and social distancing on COVID-19 pandemic. Math. Comput. Simul. 200, 285–314 (2022)

    Article  MathSciNet  PubMed  PubMed Central  Google Scholar 

  12. Wu, H., Zhang, L., Li, H.L., Teng, Z.D.: Stability analysis and optimal control on a multi-strain coinfection model with amplification and vaccination. Results Phys. 50, 106556 (2023) https://doi.org/10.1016/j.rinp.2023.106556

  13. Khajanchi, S., Bera, S., Roy, T.K.: Mathematical analysis of the global dynamics of a HTLV-I infection model, considering the role of cytotoxic T-lymphocytes. Math. Comput. Simul. 180, 354–378 (2021) https://doi.org/10.1016/j.matcom.2020.09.009

  14. Bera, S., Khajanchi, S., Roy, T.K.: Stability analysis of fuzzy HTLV-I infection model: a dynamic approach. J. Appl. Math. Comput. 69(1), 171–199 (2023)

    Article  MathSciNet  Google Scholar 

  15. Han, B.T., Jiang, D.Q., Zhou, B.Q., Hayat, T., Alsaedi, A.: Stationary distribution and probability density function of a stochastic SIRSI epidemic model with saturation incidence rate and logistic growth. Chaos Soliton Fract. 142, 110519 (2021). https://doi.org/10.1016/j.chaos.2020.110519

  16. Omame, A., Abbas, M., Din, A.: Global asymptotic stability, extinction and ergodic stationary distribution in a stochastic model for dual variants of SARS-CoV-2. Math. Comput. Simulat. 204, 302–336 (2023) https://doi.org/10.1016/j.matcom.2022.08.012

  17. Zhang, X.H., Jiang, D.Q., Alsaedi, A., Hayat, T.: Stationary distribution of stochastic SIS epidemic model with vaccination under regime switching. Appl. Math. Lett. 59, 87–93 (2016). https://doi.org/10.1016/j.aml.2016.03.010

  18. Zhao, Y.N., Jiang, D.Q., O’Regan, D.: The extinction and persistence of the stochastic SIS epidemic model with vaccination. Physica A 392(20), 4916–4927 (2013)

    Article  ADS  MathSciNet  Google Scholar 

  19. Li, Z.M., Zhang, T.L., Li, X.Q.: Threshold dynamics of stochastic models with time delays: A case study for Yunnan, China. Electron Res. Arch. 29(1), 1661–1679 (2021)

    Article  MathSciNet  Google Scholar 

  20. Cai, Y.L., Kang, Y., Wang, W.M.: A stochastic SIRS epidemic model with nonlinear incidence rate. Appl. Math. Comput. 305, 221–240 (2017). https://doi.org/10.1016/j.amc.2017.02.003

  21. Rao, F., Wang, W.M., Li, Z.B.: Stability analysis of an epidemic model with diffusion and stochastic perturbation. Commun. Nonlinear Sci. 17(6), 2551–2563 (2012)

    Article  MathSciNet  Google Scholar 

  22. Zhang, S.Q., Yuan, S.L., Zhang, T.H.: Dynamic analysis of a stochastic eco-epidemiological model with disease in predators. Stud. Appl. Math. 149(1), 5–42 (2022)

    Article  MathSciNet  Google Scholar 

  23. Silver, S.D., Driessche, P., Khajanchi, S.: A dynamic multistate and control model of the COVID-19 pandemic. J. Public Health, 1–14 (2023). https://doi.org/10.1007/s10389-023-02014-z

  24. Mollah, S., Biswas, S., Khajanchi, S.: Impact of awareness program on diabetes mellitus described by fractional-order model solving by homotopy analysis method. Ric. Mat., 1–26 (2022). https://doi.org/10.1007/s11587-022-00707-3

  25. Sarkar, K., Mondal, J., Khajanchi, S.: How do the contaminated environment influence the transmission dynamics of COVID-19 pandemic? Eur. Phys. J. Spec. Top. 231(18–20), 3697–3716 (2022)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  26. Sardar, M., Khajanchi, S.: Is the Allee effect relevant to stochastic cancer model? J. Appl. Math. Comput. 68(4), 2293–2315 (2022)

    Article  MathSciNet  Google Scholar 

  27. Dwivedi, A., Keval, R., Khajanchi, S.: Modeling optimal vaccination strategy for dengue epidemic model: a case study of India. Phys. Scr. 97(8), 085214 (2022). https://doi.org/10.1088/1402-4896/ac807b

    Article  ADS  Google Scholar 

  28. Saha, S., Dutta, P., Samanta, G.: Dynamical behavior of SIRS model incorporating government action and public response in presence of deterministic and fluctuating environments. Chaos Solit. Fract. 164, 112643 (2022)

    Article  MathSciNet  Google Scholar 

  29. Zhao, Y.N., Jiang, D.Q.: The threshold of a stochastic SIS epidemic model with vaccination. Appl. Math. Comput. 243, 718–727 (2014). https://doi.org/10.1016/j.amc.2014.05.124

  30. Ma, Z.E., Zhou, Y.C., Li, C.Z.: Qualitative and Stability Methods for Ordinary Differential Equations. Science Press, Beijing (2015)

  31. Hasminskii, R.: Stochastic Stability of Differential Equations. Sijthoff and Noordhoff. Alphen aan den Rijn, NetherBlands (1980)

  32. Shi, Z.F., Jiang, D.Q., Zhang, X.H., Alsaedi, A.: A stochastic SEIRS rabies model with population dispersal: stationary distribution and probability density function. Appl. Math. Comput. 427, 127189 (2022) https://doi.org/10.1016/j.amc.2022.127189

  33. Zuo, W.J., Jiang, D.Q.: Stationary distribution and periodic solution for stochastic predator-prey systems with nonlinear predator harvesting. Commun. Nonlinear Sci. 36, 65–80 (2016)

    Article  MathSciNet  CAS  Google Scholar 

  34. Zhang, G., Li, Z.M., Din, A.: A stochastic SIQR epidemic model with lévy jumps and three-time delays. Appl. Math. Comput. 431, 127329 (2022). https://doi.org/10.1016/j.amc.2022.127329

  35. Zhao, S.N., Yuan, S.L., Zhang, T.H.: The impact of environmental fluctuations on a plankton model with toxin-producing phytoplankton and patchy agglomeration. Chaos Soliton Fract. 162, 112426 (2022). https://doi.org/10.1016/j.chaos.2022.112426

  36. Zhang, S.Q., Yuan, S.L., Zhang, T.H.: A predator–prey model with different response functions to juvenile and adult prey in deterministic and stochastic environments. Appl. Math. Comput. 413, 126598 (2022) https://doi.org/10.1016/j.amc.2021.126598

  37. Samantaa, G., Bera, S.P.: Analysis of a Chlamydia epidemic model with pulse vaccination strategy in a random environment. Nonlinear Anal-Model. 23(4), 457–474 (2018)

    Article  MathSciNet  Google Scholar 

  38. Nguyen, D.H., Yin, G., Zhu, C.: Long-term analysis of a stochastic SIRS model with general incidence rates. SIAM J. Appl. Math. 80(2), 814–838 (2020)

    Article  MathSciNet  Google Scholar 

  39. Liu, Q., Shi, Z.F.: Analysis of a stochastic HBV infection model with DNA-containing capsids and virions. J. Nonlinear Sci. 33(2), 23 (2023). https://doi.org/10.1007/s00332-022-09883-w

    Article  ADS  MathSciNet  CAS  Google Scholar 

  40. Zhou, B.Q., Jiang, D.Q., Dai, Y.C., Hayat, T.: Threshold dynamics and probability density function of a stochastic avian influenza epidemic model with nonlinear incidence rate and psychological effect. J. Nonlinear Sci. 33(2), 29 (2023)

    Article  ADS  MathSciNet  Google Scholar 

  41. Gardiner, C.W., et al.: Handbook of Stochastic Methods. Springer, Berlin (1985)

  42. Roozen, H.: An asymptotic solution to a two-dimensional exit problem arising in population dynamics. SIAM J. Appl. Math. 49(6), 1793–1810 (1989)

    Article  MathSciNet  Google Scholar 

  43. Higham, D.J.: An algorithmic introduction to numerical simulation of stochastic differential equations. SIAM Rev. 43(3), 525–546 (2001)

    Article  ADS  MathSciNet  Google Scholar 

Download references

Acknowledgements

This work was supported by the National Natural Science Foundation of China (Grant Nos. 12271421, 12371504) and the Shaanxi Province Innovation Talent Promotion Plan Project (Grant No: 2023KJXX-056)

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Qiaoling Chen.

Ethics declarations

Conflict of interest

The authors declare that they have no conflict of interest.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Appendix

Appendix

For the same dimensional symmetric matrices \(Q_1\) and \(Q_2\), we have

\(Q_1\succ 0\): \(Q_1\) is a positive definite matrix,

\(Q_1\succeq 0\): \(Q_1\) is a positive semi-definite matrix,

\(Q_1\succeq Q_2\): \(Q_1-Q_2\) is at least a positive semi-definite matrix.

Moreover, for any invertible matrix Q, it is found that \(Q\Sigma Q^{T}\succ 0\) if \(\Sigma \succ 0\). Furthermore, define \(\Sigma _j\) as the solutions of the following algebraic equations

$$\begin{aligned} \mathcal {N}_j+\mathcal {A}\Sigma _j+\Sigma _j\mathcal {A}^T=0, ~j=1,2,3,4,\end{aligned}$$

where

$$\begin{aligned}{} & {} \mathcal {N}_1=diag\{\sigma _1,0,0,0\},~\mathcal {N}_2=diag\{0,\sigma _2,0,0\},~ \mathcal {N}_3=diag\{0,0,\sigma _3,0\} ~and\\{} & {} \quad \mathcal {N}_4=diag\{0,0,0,\sigma _4\}.\end{aligned}$$

Consider the algebraic equation

$$\begin{aligned} \mathcal {N}_1+\mathcal {A}\Sigma _1+\Sigma _1\mathcal {A}^T=0, \end{aligned}$$
(28)

which the related proof can be divided into two conditions:

$$\begin{aligned}(\aleph _1)~a_{j1}=0~(\forall j=2,3,4),~~(\hbar _1)=a_{21}^2+a_{31}^2+a_{41}^2\ne 0.\end{aligned}$$

Case 1. If \((\aleph _1)\) is satisfied, an intuitive calculation shows that

$$\begin{aligned}\Sigma _1=diag\Big \{\frac{\sigma _1}{2a_{11}},0,0,0\Big \}=\vartheta _{11}.\end{aligned}$$

The characteristic polynomial of \(\mathcal {A}\) is that \(\varphi _\mathcal {A}(\lambda )=(\lambda -a_{11})(\lambda I_4-\mathcal {H}_1)\), where \(\mathcal {H}_1=(a_{ij})_{\{2\le i,j\le 4\}}\) is the matrix. Obviously, \(\mathcal {A}\) has one of the eigenvalue \(\lambda =a_{11}\). Based on the Theorem 2.1, we know that \(a_{11}>0\), which implies that \(\vartheta _{11}\succeq 0\) and

$$\begin{aligned} \Sigma _1\succeq \vartheta _{11}. \end{aligned}$$
(29)

Case 2. If \((\hbar _1)\) is satisfied, \(a_{21}\ne 0\), or \(a_{31}\ne 0\), or \(a_{41}\ne 0\) can be obtained. We must illustrate that the elements \(a_{j1}~(j = 2, 3, 4,)\) have the equivalent status in \(\mathcal {A}\).

Define the invertible matrices \(\mathcal {J}_1\) and \(\mathcal {J}_2\) as follows

$$\begin{aligned}\mathcal {J}_1=\left( \begin{array}{cccc} 1&{}0&{}0&{}0\\ 0&{}0&{}1&{}0\\ 0&{}1&{}0&{}0\\ 0&{}0&{}0&{}1 \end{array}\right) ,\qquad \mathcal {J}_2=\left( \begin{array}{cccc} 1&{}0&{}0&{}0\\ 0&{}0&{}0&{}1\\ 0&{}0&{}1&{}0\\ 0&{}1&{}0&{}0 \end{array}\right) , \end{aligned}$$

which satisfies \(\dot{\mathcal {A}}=\mathcal {J}_1\mathcal {A}\mathcal {J}_1^{-1}\), \(\ddot{\mathcal {A}}=\mathcal {J}_2\mathcal {A}\mathcal {J}_2^{-1}\), \(\dot{\Sigma _1}=\mathcal {J}_1\Sigma _1\mathcal {J}_1^T\) and \(\ddot{\Sigma _1}=\mathcal {J}_2\Sigma _1\mathcal {J}_2^T\).

Then (28) can be equivalently expressed as

$$\begin{aligned} \mathcal {J}_1\mathcal {N}_1\mathcal {J}_1^T+\dot{\mathcal {A}}\dot{\Sigma _1}+\dot{\Sigma _1}\dot{\mathcal {A}}^T=0,~and~ \mathcal {J}_2\mathcal {N}_1\mathcal {J}_2^T+\ddot{\mathcal {A}}\ddot{\Sigma _1}+\ddot{\Sigma _1}\ddot{\mathcal {A}}^T=0.\end{aligned}$$

Now, we can find that \(\Sigma _1\), \(\dot{\Sigma _1}\) and \(\ddot{\Sigma _1}\) not only have the same positive definiteness, but also they are Hurwitz matrix. Additionally, we can obtain that \(\dot{a_{21}}=a_{31}\) and \(\ddot{a_{21}}=a_{41}\). Therefore, we only consider the condition of \(a_{21}\ne 0\).

Let \(\mathcal {B}=\mathcal {J}_3\mathcal {A}\mathcal {J}_3^{-1}\), where the elimination matrix \(\mathcal {J}_3\) is given by

$$\begin{aligned} \begin{array}{rl} \mathcal {J}_3=\left( \begin{array}{cccc} 1&{}0&{}0&{}0\\ 0&{}1&{}0&{}0\\ 0&{}-\frac{a_{31}}{a_{21}}&{}1&{}0\\ 0&{}-\frac{a_{41}}{a_{21}}&{}0&{}1 \end{array}\right) . \end{array} \end{aligned}$$

By calculation, we denote

$$\begin{aligned} \begin{array}{rl} \mathcal {B}=\left( \begin{array}{cccc} b_{11}&{}b_{12}&{}b_{13}&{}b_{14}\\ b_{21}&{}b_{22}&{}b_{23}&{}b_{24}\\ 0&{}b_{32}&{}b_{33}&{}b_{34}\\ 0&{}b_{42}&{}b_{43}&{}b_{44} \end{array}\right) . \end{array} \end{aligned}$$

There is \(b_{21}=a_{21}\ne 0\), equation (28) can be rewritten as:

$$\begin{aligned} \mathcal {J}_3\mathcal {N}_1\mathcal {J}_3^T+\mathcal {B}(\mathcal {J}_3\Sigma _1\mathcal {J}_3^T)+(\mathcal {J}_3\Sigma _1\mathcal {J}_3^T)\mathcal {B}^T=0. \end{aligned}$$
(30)

Next, the parameters \((b_{32}, b_{42})\) are similarly considered for two conditions:

$$\begin{aligned} (\aleph _2)~b_{m2}=0~(\forall m=3,4),~~(\hbar _2)~=b_{32}^2+b_{42}^2\ne 0. \end{aligned}$$

Case 2.1 If \((\aleph _2)\) is satisfied, by direct computation

$$\begin{aligned} \begin{array}{rl} \mathcal {J}_3\Sigma _1\mathcal {J}_3^T=\left( \begin{array}{cccc} \mathcal {H}_2&{}O_{2,2}\\ O_{2,2}&{}O_{2,2} \end{array}\right) =\mathcal {H}_3, \end{array} \end{aligned}$$

where O is zero matrix and \(\mathcal {H}_2=(\varsigma _{ij})_{2\times 2}\) is the symmetric matric, where

$$\begin{aligned}\begin{aligned}&\varsigma _{22}=-\frac{\sigma _1b_{21}^2}{2(b_{11}+b_{22})(b_{11}b_{22}-b_{12}b_{21})},\\&\varsigma _{11}=\frac{\sigma _1\Big (b_{22}^2+b_{11}b_{22}-b_{12}b_{21}\Big )\varsigma _{22}}{b_{21}^2},\\&\varsigma _{12}=\varsigma _{21}=-\frac{\sigma _1b_{22}\varsigma _{22}}{b_{21}}. \end{aligned}\end{aligned}$$

It is easy to obtain the characteristic polynomial of \(\mathcal {A}\):

$$\begin{aligned}\varphi _\mathcal {A}(\lambda )=\varphi _\mathcal {B}(\lambda )=\Big (\lambda ^2-(b_{11}+b_{22})\lambda +(b_{11}b_{22} -b_{12}b_{21})\Big )\Big (\lambda I_2-\mathcal {H}_4\Big )=0,\end{aligned}$$

where \(\mathcal {H}_4=(b_{ij})_{\{3\le i,j\le 4\}}\) is the matrix. Because of the \(\mathcal {A}\) is a Hurwtiz matrix, all of the roots of the equation \(\lambda ^2-(b_{11}+b_{22})\lambda +(b_{11}b_{22}-b_{12}b_{21})=0\) have negative real part. Moreover, we have \(b_{11}+b_{22}<0\) and \(b_{11}b_{22}-b_{12}b_{21}>0\). Since \(b_{21}\ne 0\), we have

$$\begin{aligned}\varsigma _{11}>0,~\varsigma _{11}\varsigma _{22}-\varsigma _{12}^2=\frac{\sigma _1(b_{11}b_{22}-b_{12}b_{21}) \varsigma _{22}^2}{b_{21}^2}>0, \end{aligned}$$

which shows that \(\mathcal {H}_2\succ 0\) and \(\mathcal {J}_3\Sigma _1\mathcal {J}_3^T\succeq 0\).

Define two positive semi-definite matrices \(\vartheta _{12}\) and \(\tilde{\vartheta _{12}}\) by

$$\begin{aligned}\vartheta _{12}=diag\Big \{\frac{\sigma _1(\varsigma _{11}\varsigma _{22}-\varsigma _{12}^2)}{\varsigma _{22}},0,0,0\Big \}, ~\tilde{\vartheta _{12}}=\mathcal {H}_3-\vartheta _{12}.\end{aligned}$$

We have

$$\begin{aligned}\Sigma _1=\mathcal {J}_3^{-1}\mathcal {H}_3(\mathcal {J}_3^{-1})^T=\mathcal {J}_3\vartheta _{12} (\mathcal {J}_3^{-1})^T+\mathcal {J}_3^{-1}\tilde{\vartheta _{12}}(\mathcal {J}_3^{-1})^T =\vartheta _{12}+\mathcal {J}_3^{-1}\tilde{\vartheta _{12}}(\mathcal {J}_3^{-1})^T.\end{aligned}$$

Through \(\vartheta _{12}\succeq 0\), it can be obtained that \(\mathcal {J}_3^{-1}\tilde{\vartheta _{12}}(\mathcal {J}_3^{-1})^T\succeq 0\) and

$$\begin{aligned} \Sigma _1\succeq \vartheta _{12}. \end{aligned}$$
(31)

Case 2.2 If \((\hbar _2)\) is satisfied, Similar to Case 2, we have \(b_{32}\ne 0\) or \(b_{42}\ne 0\) and the elements \(b_{m2}~(m=3,4)\) have the equivalent status in \(\mathcal {B}\). Therefore, we only consider the condition of \(b_{32}\ne 0\).

Define the invertible matrices \(\mathcal {J}_4\) as follows

$$\begin{aligned} \begin{array}{rl} \mathcal {J}_4=\left( \begin{array}{cccc} 1&{}0&{}0&{}0\\ 0&{}1&{}0&{}0\\ 0&{}0&{}0&{}1\\ 0&{}0&{}1&{}0 \end{array}\right) , \end{array} \end{aligned}$$

which satisfies \(\hat{\mathcal {B}}=\mathcal {J}_4\mathcal {B}\mathcal {J}_4^{-1}\) and \(\hat{\Sigma _1}=\mathcal {J}_4\Sigma _1\mathcal {J}_4^T\). That is, (30) also can be rewritten as \(\mathcal {J}_4\mathcal {N}_1\mathcal {J}_4^T+\hat{\mathcal {B}}\hat{\Sigma _1}+\hat{\Sigma _1}\hat{\mathcal {B}}^T=0\).

Let \(\mathcal {C}=\mathcal {J}_5\mathcal {B}\mathcal {J}_5^{-1}\), where \(\mathcal {J}_2\) is called the second elimination metric

$$\begin{aligned} \begin{array}{rl} \mathcal {J}_5=\left( \begin{array}{cccc} 1&{}0&{}0&{}0\\ 0&{}1&{}0&{}0\\ 0&{}0&{}1&{}0\\ 0&{}0&{}-\frac{b_{42}}{b_{32}}&{}1 \end{array}\right) , \end{array} \end{aligned}$$

By calculation, it can be denoted

$$\begin{aligned} \begin{array}{rl} \mathcal {C}=\left( \begin{array}{cccc} c_{11}&{}c_{12}&{}c_{13}&{}c_{14}\\ c_{21}&{}c_{22}&{}c_{23}&{}c_{24}\\ 0&{}c_{32}&{}c_{33}&{}c_{34}\\ 0&{}0&{}c_{43}&{}c_{44} \end{array}\right) . \end{array} \end{aligned}$$

We have \(c_{21}=b_{21}=a_{21}(\ne 0)\) and \(c_{32}=b_{32}(\ne 0)\). Equations (30) or (28) can be rewritten as:

$$\begin{aligned} (\mathcal {J}_5\mathcal {J}_3)\mathcal {N}_1(\mathcal {J}_5\mathcal {J}_3)^T+\mathcal {C}\Big ((\mathcal {J}_5\mathcal {J}_3)\Sigma _1(\mathcal {J}_5\mathcal {J}_3)^T\Big ) +\Big ((\mathcal {J}_5\mathcal {J}_3)\Sigma _1(\mathcal {J}_5\mathcal {J}_3)^T\Big )\mathcal {C}^T=0. \end{aligned}$$
(32)

Next, the parameter \(c_{43}\) is similarly considered for two conditions:

$$\begin{aligned}(\aleph _3)~c_{43}=0,~~(\hbar _3)~c_{43}^2\ne 0.\end{aligned}$$

Case 2.2.1 If \((\hbar _3)\) is satisfied, let \(\mathcal {D}=\mathcal {J}_6\mathcal {C}\mathcal {J}_6^T\), where the standardized transformation matrix

$$\begin{aligned} \begin{array}{rl} \mathcal {J}_6=\left( \begin{array}{cccc} d_{11}&{}d_{12}&{}d_{13}&{}d_{14}\\ 0&{}d_{22}&{}d_{23}&{}d_{24}\\ 0&{}0&{}d_{33}&{}d_{34}\\ 0&{}0&{}0&{}1 \end{array}\right) . \end{array} \end{aligned}$$

where

$$\begin{aligned}\begin{aligned}&d_{11}=c_{21}c_{32}c_{43},~ d_{12}=c_{32}c_{43}(c_{22}+c_{33}+c_{44}),~d_{22}=c_{32}c_{43},\\&d_{13}=c_{43}\Big (c_{23}c_{32}+c_{33}(c_{33}+c_{44}) +c_{34}c_{43}+c_{44}^2\Big ),\\&d_{14}=c_{43}\Big (c_{24}c_{32}+c_{34}(c_{33}+c_{44})\Big )+c_{44}(c_{34}c_{43}+c_{44}^2),\\&d_{22}=c_{32}c_{43}, ~d_{23}=c_{43}(c_{33}+c_{44}), ~d_{24}=c_{34}c_{43}+c_{44}^2, ~d_{33}=c_{43}, ~d_{34}=c_{44}. \end{aligned}\end{aligned}$$

By calculation, it can be obtained

$$\begin{aligned} \begin{array}{rl} \mathcal {D}=\left( \begin{array}{cccc} -\varphi _1&{}-\varphi _2&{}-\varphi _3&{}-\varphi _4\\ 1&{}0&{}0&{}0\\ 0&{}1&{}0&{}0\\ 0&{}0&{}1&{}0 \end{array}\right) , \end{array} \end{aligned}$$

where \(\varphi _1, \varphi _2, \varphi _3, \varphi _4\) are the same as above. Furthermore, equation (32) can be rewritten as:

$$\begin{aligned}{} & {} (\mathcal {J}_6\mathcal {J}_5\mathcal {J}_3)\mathcal {N}_1(\mathcal {J}_6\mathcal {J}_5\mathcal {J}_3)^T+\mathcal {C}\Big ((\mathcal {J}_6\mathcal {J}_5\mathcal {J}_3)\Sigma _1(\mathcal {J}_6\mathcal {J}_5 \mathcal {J}_3)^T\Big )\nonumber \\{} & {} \quad +\Big ((\mathcal {J}_6\mathcal {J}_5\mathcal {J}_3) \Sigma _1(\mathcal {J}_6\mathcal {J}_5\mathcal {J}_3)^T\Big )\mathcal {C}^T=0. \end{aligned}$$
(33)

Moreover, using the Lemma 2.2, we can obtain that the matrix \(\Sigma _1\) is positive definite whose explicit form is

$$\begin{aligned} \begin{array}{rl} \Sigma _1=\left( \begin{array}{cccc} \frac{\sigma _1(\varpi _2\varpi _3-\varpi _1\varpi _4)}{2(\varpi _1\varpi _2\varpi _3-\varpi _3^2-\varpi _1^2\varpi _4)} &{}0&{}-\frac{\sigma _1\varpi _3}{2(\varpi _1\varpi _2\varpi _3-\varpi _3^2-\varpi _1^2\varpi _4)}&{}0\\ 0&{}\frac{\sigma _1\varpi _3}{2(\varpi _1\varpi _2\varpi _3-\varpi _3^2-\varpi _1^2\varpi _4)}&{}0&{}-\frac{\sigma _1\varpi _1}{2(\varpi _1\varpi _2\varpi _3-\varpi _3^2-\varpi _1^2\varpi _4)}\\ -\frac{\sigma _1\varpi _3}{2(\varpi _1\varpi _2\varpi _3-\varpi _3^2-\varpi _1^2\varpi _4)}&{}0&{} \frac{\sigma _1\varpi _1}{2(\varpi _1\varpi _2\varpi _3-\varpi _3^2-\varpi _1^2\varpi _4)}&{}0\\ 0&{}-\frac{\sigma _1\varpi _1}{2(\varpi _1\varpi _2\varpi _3-\varpi _3^2-\varpi _1^2\varpi _4)}&{}0&{} \frac{\sigma _1(\varpi _1\varpi _2-\varpi _3)}{2(\varpi _1\varpi _2\varpi _3-\varpi _3^2-\varpi _1^2\varpi _4)} \end{array}\right) . \end{array} \end{aligned}$$

Compared the \(\mathcal {N}_1^2+\mathcal {D}\Sigma _1+\Sigma _1\mathcal {D}^T=0\), we have \(\Sigma _1=\Xi ^{-2}(\mathcal {J}_6\mathcal {J}_5\mathcal {J}_3)\Sigma _1(\mathcal {J}_6\mathcal {J}_5\mathcal {J}_3)^T\), where \(\Xi =d_{11}\sigma _1\). Thus, the matrix \(\Sigma _1=\Xi ^{-2}(\mathcal {J}_6\mathcal {J}_5\mathcal {J}_3)^{-1}\Sigma _1((\mathcal {J}_6\mathcal {J}_5\mathcal {J}_3)^{-1})^T\) is at least a semi-positive matrix.

Case 2.2.2 If \((\aleph _3)\) is satisfied, then the characteristic polynomial of \(\mathcal {C}\) is

$$\begin{aligned}\phi _\mathcal {C}(\lambda )=\phi _\mathcal {A}(\lambda )=(\lambda -c_{44})(\lambda ^3+ \xi _1\lambda ^2+\xi _2\lambda +\xi _3)=0,\end{aligned}$$

where

$$\begin{aligned}\begin{aligned}&\xi _1=-(c_{11}+c_{22}+c_{33}),~\xi _2=c_{11}c_{22}+c_{11}c_{33}+c_{22}c_{33}-c_{12}c_{21}-c_{23}c_{32},\\&\xi _3=-c_{11}c_{22}c_{33}-c_{13}c_{21}c_{32}+c_{33}c_{21}c_{12}+c_{32}c_{23}c_{11}. \end{aligned}\end{aligned}$$

Using Theorem 1 and \(\mathcal {A}\) is a Hurwitz matrix, the solution of equation \(\lambda ^3+\xi _1\lambda ^2+\xi _2\lambda +\xi _3=0\) have roots with all negative real components, that is

$$\begin{aligned}\xi _1>0,~\xi _2>0,~\xi _3>0,~\xi _1\xi _2-\xi _3>0.\end{aligned}$$

Let \(\mathcal {D}=\mathcal {J}_7\mathcal {C}\mathcal {J}_7^{-1}\) and using the Lemma 2.3, where the standardized transformation matrix \(\mathcal {J}_7\)

$$\begin{aligned} \begin{array}{rl} \mathcal {J}_7=\left( \begin{array}{cccc} c_{21}c_{32}&{}c_{32}c_{22}&{}c_{32}c_{23}+c_{33}^2+c_{34}&{}c_{24}c_{32}+c_{34}c_{33}\\ 0&{}c_{32}&{}c_{33}&{}c_{34}\\ 0&{}0&{}1&{}0\\ 0&{}0&{}0&{}1 \end{array}\right) . \end{array} \end{aligned}$$

By calculation, we have

$$\begin{aligned} \begin{array}{rl} \mathcal {D}=\left( \begin{array}{cccc} -\epsilon _1&{}-\epsilon _2&{}-\epsilon _3&{}-\epsilon _4\\ 1&{}0&{}0&{}0\\ 0&{}1&{}0&{}0\\ 0&{}0&{}0&{}c_{44} \end{array}\right) . \end{array} \end{aligned}$$

where

$$\begin{aligned}\begin{aligned}&\epsilon _1=-(c_{11}+c_{22}),~\epsilon _2=c_{11}c_{22}-c_{33}^2-c_{34}-c_{23}c_{32}\\&\epsilon _3=c_{11}c_{34}+c_{22}c_{34}+(c_{11}+c_{22})c_{33}^2+c_{11}c_{23}c_{32}+c_{12}c_{21}c_{33} -c_{11}c_{22}c_{33}\\&\qquad -c_{13}c_{21}c_{32}-c_{24}c_{32}c_{43}-c_{33}c_{34}c_{43},\\&\epsilon _4=c_{11}c_{24}c_{32}+c_{12}c_{21}c_{34}+c_{11}c_{33}c_{34}+c_{22}c_{33}c_{34} -c_{11}c_{22}c_{34}-c_{14}c_{21}c_{32}-c_{24}c_{32}c_{44}\\&\qquad -c_{24}c_{32}c_{44}-c_{33}c_{34}c_{44}. \end{aligned}\end{aligned}$$

Then, the equation (32) can be rewritten as

$$\begin{aligned}{} & {} (\mathcal {J}_7\mathcal {J}_5\mathcal {J}_3)\mathcal {N}_1(\mathcal {J}_7\mathcal {J}_5\mathcal {J}_3)^T+\mathcal {C}\Big ((\mathcal {J}_7\mathcal {J}_5\mathcal {J}_3)\Sigma _1 (\mathcal {J}_7\mathcal {J}_5\mathcal {J}_3)^T\Big )\nonumber \\{} & {} \quad +\Big ((\mathcal {J}_7\mathcal {J}_5\mathcal {J}_3) \Sigma _1(\mathcal {J}_7\mathcal {J}_5\mathcal {J}_3)^T\Big )\mathcal {C}^T=0. \end{aligned}$$
(34)

where the \(\Sigma _1\) is shown as

$$\begin{aligned} \begin{array}{rl} \Sigma _1=\left( \begin{array}{cccc} \frac{\sigma _1\epsilon _1}{2(\epsilon _1\epsilon _2-\epsilon _3)}&{}0&{}-\frac{\sigma _1}{2(\epsilon _1\epsilon _2-\epsilon _3)}&{}0\\ 0&{}\frac{\sigma _1}{2(\epsilon _1\epsilon _2-\epsilon _3)}&{}0&{}0\\ -\frac{\sigma _1}{2(\epsilon _1\epsilon _2-\epsilon _3)}&{}0&{}\frac{\sigma _1\epsilon _1}{2\epsilon _3(\epsilon _1\epsilon _2-\epsilon _3)} &{}0\\ 0&{}0&{}0&{}0 \end{array}\right) . \end{array} \end{aligned}$$

Compared the \(\mathcal {N}_1^2+\mathcal {D}\tilde{\Sigma _1}+\tilde{\Sigma _1}\mathcal {D}^T=0\), we have \(\tilde{\Sigma _1}=\Xi ^{-2}(\mathcal {J}_7\mathcal {J}_5\mathcal {J}_3)\Sigma _1 (\mathcal {J}_7\mathcal {J}_5\mathcal {J}_3)^T\), where \(\Xi =c_{21}c_{32}\sigma _1\). Thus, the matrix \(\Sigma _1=\Xi ^{-2}(\mathcal {J}_7\mathcal {J}_5\mathcal {J}_3)^{-1}\tilde{\Sigma _1}((\mathcal {J}_7 \mathcal {J}_5\mathcal {J}_3)^{-1})^T\) is at least a semi-positive matrix.

For \(\mathcal {N}_i~(i=2,3,4)\), its approach is similar to that of \(\mathcal {N}_1\). Therefore, we ignore it here. The proof is complete.

Rights and permissions

Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Niu, L., Chen, Q. & Teng, Z. Threshold Dynamics and Probability Density Function of a Stochastic Multi-Strain Coinfection Model with Amplification and Vaccination. Qual. Theory Dyn. Syst. 23, 94 (2024). https://doi.org/10.1007/s12346-024-00957-6

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s12346-024-00957-6

Keywords

Navigation