Skip to main content
Log in

Spatiotemporal Evolution of Coinfection Dynamics: A Reaction–Diffusion Model

  • Published:
Journal of Dynamics and Differential Equations Aims and scope Submit manuscript

Abstract

This paper investigates the impact of spatial heterogeneity on the interaction between similar strains in a dynamical system of coinfecting strains with spatial diffusion. The SIS model studied is a reaction–diffusion system with spatially heterogeneous coefficients. The study considers two limiting cases: asymptotically slow and fast diffusion coefficients. When the diffusion coefficient is small, the slow system is shown to be a semilinear system of “replicator equations,” describing the spatiotemporal evolution of the strains’ frequencies. This system is of the reaction–advection–diffusion type, with an additional advection term that explicitly involves the heterogeneity of the associated neutral system. In the case of fast diffusion, traditional methods of aggregating variables are used to reduce the spatialized SIS problem to a homogenized SIS system, on which the results of the non-spatial model can be applied directly.

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.

Similar content being viewed by others

Availability of data and materials

Not applicable

Code availability

Not applicable

Notes

  1. Notice that we also include same strain coinfection, as argued in [6, 11, 17]. This allows the system to simulate the exclusion of every strain but one.

  2. This is due to the fact that each equation reads \(\partial _t u_i -d\Delta u_i =f_i(u)\) with \(f_i(u_{|u_i=0})\ge 0\).

  3. The name semi-neutral system comes from the fact that if \(\epsilon =0\), except the coefficients of diffusion terms, then the parameters do not depend on the strains as in the neutral theory.

  4. We use the usual notation abuse. Rigorously speaking, we have to define \(\widetilde{X}(\tau )=X\left( \frac{\tau }{\epsilon }\right) \) and the same for each variables. Here we remove the \(\widetilde{}\) for simplicity.

References

  1. Poggiale, J.C.: Lotka–Volterra’s model and migrations: breaking of the well-known center. Math. Comput. Model. 27(4), 51–61 (1998). https://doi.org/10.1016/S0895-7177(98)00005-3

    Article  MathSciNet  MATH  Google Scholar 

  2. Poggiale, J.C.: Predator–prey models in heterogeneous environment: emergence of functional response. Math. Comput. Model. 27(4), 63–71 (1998)

    Article  MathSciNet  MATH  Google Scholar 

  3. Wang, Z.-C., Wu, J.: Travelling waves of a diffusive Kermack–Mckendrick epidemic model with non-local delayed transmission. Proc R Soc A Math Phys Eng Sci 466, 237–261 (2009)

    MathSciNet  MATH  Google Scholar 

  4. Wang, X.-S., Wang, H., Wu, J.: Traveling waves of diffusive predator–prey systems: disease outbreak propagation. Discrete Contin. Dynam. Syst. 32, 3303–3324 (2012)

    Article  MathSciNet  MATH  Google Scholar 

  5. Adler, F.R., Brunet, R.C.: The dynamics of simultaneous infections with altered susceptibilities. Theor. Popul. Biol. 40(3), 369–410 (1991). https://doi.org/10.1016/0040-5809(91)90061-J

    Article  MATH  Google Scholar 

  6. Alizon, S.: Co-infection and super-infection models in evolutionary epidemiology. Interface Focus 3(6), 20130031 (2013)

    Article  Google Scholar 

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

    Article  MathSciNet  MATH  Google Scholar 

  8. Marchaim, D., Perez, F., Lee, J., Bheemreddy, S., Hujer, A.M., Rudin, S., Hayakawa, K., Lephart, P.R., Blunden, C., Shango, M., Campbell, M.L., Varkey, J., Manickam, P., Patel, D., Pogue, J.M., Chopra, T., Martin, E.T., Dhar, S., Bonomo, R.A., Kaye, K.S.: “swimming in resistance’’: co-colonization with carbapenem-resistant enterobacteriaceae and Acinetobacter baumannii or Pseudomonas aeruginosa. Am. J. Infect. Control 40(9), 830–835 (2012). https://doi.org/10.1016/j.ajic.2011.10.013

    Article  Google Scholar 

  9. Warren, D., Nitin, A., Hill, C., Fraser, V., Kollef, M.: Occurrence of co-colonization or co-infection with vancomycin-resistant enterococci and methicillin-resistant staphylococcus aureus in a medical intensive care unit. Infect. Control Hosp. Epidemiol. 25(2), 99–104 (2004)

    Article  Google Scholar 

  10. Madec, S., Gjini, E.: Predicting n-strain coexistence from co-colonization interactions: Epidemiology meets ecology and the replicator equation. Bull. Math. Biol. 82(11), 142 (2020). https://doi.org/10.1007/s11538-020-00816-w

    Article  MathSciNet  MATH  Google Scholar 

  11. Le, T.M.T., Gjini, E., Madec, S.: Quasi-neutral dynamics in a coinfection system with N strains and asymmetries along multiple traits. J. Math. Biol. 87, 48 (2023)

    Article  MathSciNet  Google Scholar 

  12. Cantrell, R.S., Cosner, C.: Spatial Ecology Via Reaction–Diffusion Equations. Wiley, New York (2004)

    Book  MATH  Google Scholar 

  13. Allen, L.J.S., Bolker, B.M., Lou, Y., Nevai, A.L.: Asymptotic profiles of the steady states for an sis epidemic reaction–diffusion model. Discrete Contin. Dynam. . 21(1), 1–20 (2008)

    Article  MathSciNet  MATH  Google Scholar 

  14. Castella, F., Hoffbeck, J.-P., Lagadeuc, Y.: A reduced model for spatially structured predator-prey systems with fast spatial migrations and slow demographic evolutions. Asymptot. Anal. 61(3–4), 125–175 (2009). https://doi.org/10.3233/asy-2008-0905

    Article  MathSciNet  MATH  Google Scholar 

  15. Castella, F., Madec, S., Lagadeuc, Y.: Global behavior of n competing species with strong diffusion: diffusion leads to exclusion. Appl. Anal. 95(2), 341–372 (2016). https://doi.org/10.1080/00036811.2015.1004320

    Article  MathSciNet  MATH  Google Scholar 

  16. Bratus, A.S., Posvyanskii, V.P., Novozhilov, A.S.: Replicator equations and space. Math. Model. Nat. Phenom. 9(3), 47–67 (2014)

    Article  MathSciNet  MATH  Google Scholar 

  17. van Baalen, M., Sabelis, M.W.: The dynamics of multiple infection and the evolution of virulence. Am. Nat. 146, 881–910 (1995)

    Article  Google Scholar 

  18. Hollis, S.L., Martin, R.H., Jr., Pierre, M.: Global existence and boundedness in reaction–diffusion systems. SIAM J. Math. Anal. 18(3), 744–761 (1987). https://doi.org/10.1137/0518057

    Article  MathSciNet  MATH  Google Scholar 

  19. Morgan, J.: Global existence for semilinear parabolic systems. SIAM J. Math. Anal. 20(5), 1128–1144 (1989). https://doi.org/10.1137/0520075

    Article  MathSciNet  MATH  Google Scholar 

  20. Pao, C.V.: Nonlinear Parabolic and Elliptic Equations. Springer, Berlin (1993)

    Book  Google Scholar 

  21. Biegert, M.: The Neumann Laplacian on spaces of continuous functions. Note di Matematica 22(1), 65–74 (2003)

    MathSciNet  MATH  Google Scholar 

  22. Pazy, A.: Semigroups of Linear Operators and Applications to Partial Differential Equations. Applied Mathematical Sciences, vol. 44. Springer, Berlin (1983)

    MATH  Google Scholar 

  23. Diekmann, O., Heesterbeek, J., Metz, J.: On the definition and the computation of the basic reproduction ratio r0 in models for infectious diseases in heterogeneous populations. J. Math. Biol. 28, 365–382 (1990)

    Article  MathSciNet  MATH  Google Scholar 

  24. Smoller, J.: Shock Waves and Reaction–Diffusion Equations. Grundlehren der mathematischen Wissenschaften, vol. 258. Springer, Berlin (1994)

    Book  Google Scholar 

  25. Anderson, R.M., May, R.: Infectious Diseases of Humans: Dynamics and Control. Oxford University Press, Oxford (1991)

    Google Scholar 

  26. Dushoff, J., Levin, S.: The effects of population heterogeneity on disease invasion. Math. Biosci. 128(1–2), 25–40 (1995)

    Article  MATH  Google Scholar 

  27. Lajmanovich, A., Yorke, J.: A deterministic model for gonorrhea in a nonhomogeneous population. Bellman Prize Math. Biosci. 28, 221–236 (1976)

    Article  MathSciNet  MATH  Google Scholar 

  28. Lloyd, A., May, R.: Spatial heterogeneity in epidemic models. J. Theor. Biol. 179(1), 1–11 (1996)

    Article  Google Scholar 

  29. Mottoni, P., Orlandi, E., Tesei, A.: Asymptotic behavior for a system describing epidemics with migration and spatial spread of infection. Nonlinear Anal. Theory Methods Appl. 3, 663–675 (1979)

    Article  MathSciNet  MATH  Google Scholar 

  30. Pang, D., Xiao, Y.: The sis model with diffusion of virus in the environment. Math. Biosci. Eng. 16, 2852–2874 (2019)

    Article  MathSciNet  MATH  Google Scholar 

  31. Fitzgibbon, W., Langlais, M., Morgan, J.: A reaction–diffusion system modeling direct and indirect transmission of diseases. Discrete Contin. Dyn. Syst. Ser. B 4, 893–910 (2004)

    MathSciNet  MATH  Google Scholar 

  32. Fitzgibbon, W., Langlais, M.: Simple models for the transmission of microparasites between host populations living on noncoincident spatial domains. In: Structured Population Models in Biology and Epidemiology (2008)

  33. Ruan, S.: Spatial-temporal dynamics in nonlocal epidemiological models. In: Mathematics for Life Science and Medicine (2007)

  34. Hofbauer, J., Sigmund, K.: Evolutionary Games and Population Dynamics. Cambridge University Press, London (1998)

    Book  MATH  Google Scholar 

  35. Wu, Y., Zou, X.: Asymptotic profiles of steady states for a diffusive sis epidemic model with mass action infection mechanism. J. Differ. Equ. 261, 4424–4447 (2016)

    Article  MathSciNet  MATH  Google Scholar 

  36. Megretski, A.: Singular Perturbations and Averaging. Lecture Notes, Massachusetts Institute of Technology (2003)

  37. Duan, G.-R., Patton, R.J.: A note on Hurwitz stability of matrices. Automatica 34, 509–511 (1998)

    Article  MathSciNet  MATH  Google Scholar 

Download references

Acknowledgements

We would like to thank Professor Boris ANDREIANOV, Laboratory of Mathematics and Theoretical Physics, University of Tours. Professor Andreianov helped us with several techniques in the proof of Theorem 7.

Funding

We have no funding received for this paper.

Author information

Authors and Affiliations

Authors

Contributions

Two authors have the same contributions, including writing and revising this manuscript.

Corresponding author

Correspondence to Thi Minh Thao Le.

Ethics declarations

Conflict of interest

We declare that the authors have no competing interests as defined by Springer, or other interests that might be perceived to influence the results and/or discussion reported in this paper.

Ethics approval

Not applicable

Consent to participate

Two authors consent to participate.

Consent for publication

Two authors consent to participate.

Additional information

Publisher's Note

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

Appendices

Appendix A Proof for Theorem 7

The idea of our proof bases on the technique mentioned in [36].

Proof

Firstly, we make a convention for the norm using in this proof. For each \(t \in \mathbb {R}^+\), for every \(f_1, f_2 \in L^2\left( \Omega \times \mathbb {R},\mathbb {R}^n\right) \) we denote

$$\begin{aligned} \langle f_1, f_2 \rangle = \int _{\Omega }f_1\left( x,t\right) \cdot f_2\left( x,t\right) dx, \end{aligned}$$

where the \(f_1\cdot f_2\) representing for the usual scalar product \(\sum _{i=1}^{n}f^i_1 f^i_2\) in \(\mathbb {R}^n\). This scalar product \(\langle \cdot ,\cdot \rangle \) induces the norm

$$\begin{aligned} \left\| f\left( \cdot ,t\right) \right\| _2 = \left( \int _{\Omega }f\left( x,t\right) \cdot f\left( x,t\right) dx\right) ^{1/2} \end{aligned}$$

For the sake of convenience in this proof, we only write \(\left\| \cdot \right\| \) instead of \(\left\| \cdot \right\| _2\).

We do the same convention for \(\langle g_1,g_2 \rangle \) and \(\left\| g\left( \cdot ,t\right) \right\| \) for all \(g_1,g_2,g \in C^1\left( \Omega \times \mathbb {R},\mathbb {R}^m\right) \).

Because in the finite dimensional space, all norms are equivalent, we then denote \(|\cdot |\) to be the usual 2-Euclidean norm. Moreover, we recall the notation \(A \prec 0\) for a symmetric matrix A if A is definitely negative, and \(A \succ 0\) for definitely positive symmetric matrix.

First, let us show that the interval \([t_0,t_1]\) can be subdivided into subinterval \(\Delta _k = [\tau _{k-1},\tau _k]\), where \(k \in \{1,2,\dots ,N\}\) and \(t_0 = \tau _0< \tau _1< \dots <\tau _N = t_1\) in such a way that for every k, there exists a symmetric matrix \(P_k = P^T_k \succ 0\) for which

$$\begin{aligned} P_k A(x,t) + A^T(x,t)P_k \prec -I. \end{aligned}$$
(A1)

Indeed, since A(xt) is a Hurwitz matrix for every \(t \in [t_0,t_1]\), according to [37], there exists \(P(x,t) = P^T(x,t) \succ 0\) such that

$$\begin{aligned} P(x,t) A(x,t) + A^T(x,t)P(x,t) \prec -I. \end{aligned}$$

Since A depends continuously on t, there exists an open interval \(\Delta (t)\) such that \(t \in \Delta (t)\) and

$$\begin{aligned} P(x,t)A(x,\tau ) + A^T(x,\tau )P(x,t) \prec -I, \quad \forall \tau \in \Delta (t). \end{aligned}$$

Now the open intervals \(\Delta (t)\) with \(t \in \left[ t_0,t_1\right] \) cover the whole closed bounded interval \(\left[ t_0,t_1\right] \) and taking a finite number of \(\tau _k\), \(k=1,\dots ,N\) such that \(\left[ t_0,t_1\right] \) is completely covered by \(\Delta (\tau _k)\) yields the desired partition subdivision.

We can note that a strictly negative upper bound is not required on the real parts’ eigenvalues uniformly in space because the spatial domain is supposed to be compact.

Note that, from (A1), for all \(y \in \mathbb {R}^m\) we have that

$$\begin{aligned} y^T\left( P_kA + A^TP_k\right) y \prec -y^T y. \end{aligned}$$
(A2)

Second, because FG are continuously differential in x and t, then for every \(\mu >0\) there exists \(C,r >0\) such that

$$\begin{aligned}{} & {} \left\| F\left( f_0(x,t) + \bar{\delta }_f\left( x,t\right) ,g_0(x,t) + \bar{\delta }_g\left( x,t\right) ,x,t\right) - F\left( f_0,g_0,x,t\right) \right\| \nonumber \\{} & {} \quad \le C\left( \left\| \bar{\delta }_f\left( x,t\right) \right\| + \left\| \bar{\delta }_g\left( x,t\right) \right\| \right) \end{aligned}$$
(A3)

for all \(t \in \mathbb {R}\), \(\bar{\delta }_f\left( x,t\right) \in \mathbb {R}^n\), \(\bar{\delta }_g\left( x,t\right) \in \mathbb {R}^m\) satisfying

$$\begin{aligned} \forall t \in [t_0,t_1], \forall x\in \Omega , \qquad |\bar{\delta }_f(x,t)| \le r, \qquad |\bar{\delta }_g (x,t)| \le r. \end{aligned}$$

For the sake of simplicity, we write \(\bar{\delta }_f\) and \(\bar{\delta }_g\) instead of \(\bar{\delta }_f\left( x,t\right) \) and \(\bar{\delta }_g\left( x,t\right) \). We now have the Taylor expansion as follows, noting that \(G\left( f_0(x,t),g_0(x,t),x,t\right) = 0\),

$$\begin{aligned} G\left( f_0(x,t) + \bar{\delta }_f,g_0(x,t) + \bar{\delta _g},x,t\right) = A\left( x,t\right) \bar{\delta _g} + B(x,t)\bar{\delta }_f + o\left( |\bar{\delta _g}|\right) + o\left( |\bar{\delta _g}|\right) , \end{aligned}$$
(A4)

with B(xt) is the Jacobian matrix of \(G\left( \cdot ,\cdot ,t\right) \) with respect to the first variable.

For each \(k = 1,\dots , N\), and \(u \in \mathbb {R}^m\), set \(|u|_k = \left( u^T P_k u\right) ^{1/2}\), then \(|\cdot |_k\) is a norm in \(\mathbb {R}^m\). Indeed, because \(P_k \succ 0\) then \(|\cdot |_k\) is well-defined, it suffices to check the condition \(|u + v|_k \le |u|_k + |v|_k\), which is equivalent to

$$\begin{aligned} \left( u^TP_kv\right) ^2 \le \left( u^TP_ku\right) \left( v^TP_kv\right) . \end{aligned}$$

It now becomes

$$\begin{aligned} \left( \left( L^Tu\right) ^T\left( L^Tv\right) \right) ^2 \le \left( \left( L^Tu\right) ^T\left( L^Tu\right) \right) \left( \left( L^Tv\right) ^T\left( L^Tv\right) \right) , \end{aligned}$$
(A5)

thanks to the Cholesky’s factorization, which states that, if \(P_k \succ 0\), there exist a square matrix such that \(P_k = L^T_kL_k\). Note that, (A5) holds because of the inequality Cauchy-Schwarz. Hence, \(|\cdot |_k\) is a norm in \(\mathbb {R}^m\) and it is equivalent to an arbitrary norm in \(\mathbb {R}^m\).

Then, for \(\delta _f(x,t) = f(x,t) - f_0(x,t)\), \(\delta _g(x,t) = g(x,t) - g_0(x,t)\), we have that

$$\begin{aligned} \left\{ \begin{array}{ll} \dfrac{d}{d t}\left\| \delta _f\right\| ^2 &{}\le C_1 \left( \left\| \delta _f\right\| + \left\| \delta _g\right\| \right) \left\| \delta _f\right\| ,\\ \epsilon \dfrac{d}{d t} \left\| \delta _g\right\| ^2_k &{}\le -q\left\| \delta _g\right\| ^2_kdt + C_1\left( \left\| \delta _f\right\| ^2 + \epsilon \right) dt \end{array} \right. \end{aligned}$$
(A6)

as long as \(\delta _f\), \(\delta _g\) are sufficiently small, where \(C_1\), q are positive constants which do not depend on k.

Initially, for the sake of simplicity, in the following arguments, we write f, g instead of \(f\left( x,t\right) \) and \(g\left( x,t\right) \), respectively. Then, we have the equation for \(\delta _f\left( x,t\right) \) as follows

$$\begin{aligned} \begin{aligned} \dfrac{\partial }{\partial t}\delta _f&= F\left( f_0 + \delta _f,g_0+\delta _g,x,t\right) - F\left( f_0 ,g_0,x,t\right) + K\delta _f \end{aligned}. \end{aligned}$$

By the convention of \(\left\| \cdot \right\| \), we have that

$$\begin{aligned} \begin{aligned} \dfrac{d}{d t}\left\| \delta _f\right\| ^2&= \dfrac{d}{d t}\langle \delta _f,\delta _f\rangle = 2\langle \dfrac{\partial }{\partial t}\delta _f,\delta _f\rangle =\langle F\left( f_0 + \delta _f,g_0+\delta _g,x,t\right) \\&\quad - F\left( f_0 ,g_0,x,t\right) + K\delta _f,\delta _f\rangle \\&= \langle F\left( f_0 + \delta _f,g_0+\delta _g,x,t\right) - F\left( f_0 ,g_0,x,t\right) ,\delta _f\rangle + \langle K\delta _f,\delta _f\rangle \\&\le \left\| F\left( f_0 + \delta _f,g_0+\delta _g,x,t\right) - F\left( f_0 ,g_0,x,t\right) \right\| \left\| \delta _f\right\| \\&\quad + \langle K\delta _f,\delta _f\rangle \le C\left( \left\| \delta _f\right\| + \left\| \delta _g\right\| \right) \left\| \delta _f\right\| + \langle K\delta _f,\delta _f\rangle . \end{aligned} \end{aligned}$$

On the other hand, recalling that \(K_f = a_f\left( x\right) \nabla + \Delta \) implies

$$\begin{aligned} \begin{aligned} \langle K_f \delta _f,\delta _f\rangle = \langle \Delta \delta _f,\delta _f\rangle + \langle a_f\left( x\right) \nabla \delta _f,\delta _f\rangle = -\int _{\Omega }\left| \nabla \delta _f\right| ^2dx + \int _{\Omega }a_f\left( x\right) \nabla \delta _f\cdot \delta _f dx \end{aligned} \end{aligned}$$

which leads to, when we apply the Young inequality for the term \(\int _{\Omega }a\left( x\right) \nabla \delta _f\cdot \delta _f dx\),

$$\begin{aligned} \begin{aligned} \langle K_f \delta _f,\delta _f\rangle \le&-\int _{\Omega }\left| \nabla \delta _f\right| ^2dx + \max \limits _{x\in \Omega } \left( \left| a_f\left( x\right) \right| \right) \\&\left[ \dfrac{1}{\max \limits _{x\in \Omega } \left( \left| a_f\left( x\right) \right| \right) }\int _{\Omega }\left| \nabla \delta _f\right| ^2dx + C\left( \max \limits _{x\in \Omega }\left( \left| a_f\left( x\right) \right| \right) \right) \int _{\Omega }\left| \delta _f\right| ^2dx\right] , \end{aligned} \end{aligned}$$

where \(\left| a_f\left( x\right) \right| \) is the matrix in which entries are absolute values of corresponding coordinates of \(a_f\left( x\right) \).

Accordingly, we have the estimation for \(\dfrac{d}{d t}\left\| \delta _f\right\| ^2\) as follows

$$\begin{aligned} \dfrac{d}{d t}\left\| \delta _f\right\| ^2 \le C_1\left( \left\| \delta _f\right\| + \left\| \delta _g\right\| \right) \left\| \delta _f\right\| \end{aligned}$$
(A7)

Next, we come to control the growth of \(\left\| \delta _g\right\| _k\). We first observe that

$$\begin{aligned} \begin{aligned} \epsilon \dfrac{\partial }{\partial t}\delta _g =&G\left( f_0\left( x,t\right) +\delta _f, g_0\left( x,t\right) +\delta _g,x,t\right) \\&+ \epsilon K_g\delta _g + \epsilon \left[ K_g g_0\left( x,t\right) + \dfrac{\partial }{\partial t}g_0\left( x,t\right) + G_1(x)\cdot \nabla f_0\right] + \epsilon G_1(x)\cdot \nabla \delta _f. \end{aligned} \end{aligned}$$

We denote \(\epsilon \left[ K_g g_0\left( x,t\right) + \dfrac{\partial }{\partial t}g_0\left( x,t\right) + G_1(x)\cdot \nabla f_0\right] \) as \(O\left( \epsilon \right) \), then

$$\begin{aligned} \epsilon \dfrac{\partial }{\partial t}\delta _g = G\left( f_0\left( x,t\right) +\delta _f, g_0\left( x,t\right) +\delta _g,x,t\right) + \epsilon K_g\delta _g + O\left( \epsilon \right) + \epsilon G_1(x)\cdot \nabla \delta _f. \end{aligned}$$
(A8)

Using the Taylor expansion for G in (A4) and the equation (A8), we obtain the following computations

$$\begin{aligned} \begin{aligned} \epsilon \dfrac{d}{d t}\left\| \delta _g\right\| ^2_k&= \epsilon \dfrac{d}{d t}\langle \delta _g, P_k\delta _g\rangle = \epsilon \langle \dfrac{\partial }{\partial t}\delta _g,P_k\delta _g\rangle + \epsilon \langle \delta _g,P_k\dfrac{\partial }{\partial t}\delta _g\rangle \\&= \left( \langle A\delta _g, P_k\delta _g\rangle + \langle \delta _g,P_kA\delta _g\rangle \right) + 2B(x,t)\langle \delta _f,\delta _g\rangle \\&\quad + \langle o\left( \left| \delta _g\right| \right) + o\left( \left| \delta _g\right| \right) + O\left( \epsilon \right) ,P_k\delta _g + \delta _g\rangle \\&\quad + 2\epsilon \langle G_1(x)\cdot \nabla \delta _f,P_k\delta _g + \delta _g\rangle + \epsilon \langle K_g\delta _g, P_k\delta _g + \delta _g\rangle \\&= \langle \delta _g,\left( A^TP_k + P_kA\right) \delta _g\rangle + 2B(x,t)\langle \delta _f,\delta _g\rangle + \langle o\left( \left| \delta _f\right| \right) \\&\quad +o\left( \left| \delta _g\right| \right) + O\left( \epsilon \right) ,P_k\delta _g + \delta _g\rangle \\&\quad + 2\epsilon \langle G_1(x)\cdot \nabla \delta _f,P_k\delta _g + \delta _g\rangle + \epsilon \langle K_g\delta _g, P_k\delta _g + \delta _g\rangle \\&\le -\left\| \delta _g\right\| ^2 + 2C_1\left\| \delta _f\right\| \left\| \delta _g\right\| + \langle o\left( \left| \delta _f\right| \right) + o\left( \left| \delta _g\right| \right) + O\left( \epsilon \right) ,P\delta _g + \delta _g\rangle \\&\quad + 2\epsilon \langle G_1(x)\cdot \nabla ,P_k\delta _g + \delta _g\rangle + \epsilon \langle K_g\delta _g, P\delta _g + \delta _g\rangle . \end{aligned} \end{aligned}$$
(A9)

Using the Young inequality, we have the estimation for \(\langle o\left( \left| \delta _f\right| \right) + o\left( \left| \delta _g\right| \right) + O\left( \epsilon \right) ,P\delta _g + \delta _g\rangle \) as follows

$$\begin{aligned} \langle o\left( \left| \delta _f\right| \right) + o\left( \left| \delta _g\right| \right) + O\left( \epsilon \right) ,P\delta _g + \delta _g\rangle \le O\left( \epsilon \right) + \tilde{C}\left\| \delta _g\right\| ^2, \qquad \text {with} \quad \tilde{C} \ll 1. \end{aligned}$$
(A10)

Alternatively, applying the Young inequality, we have that

$$\begin{aligned} \langle G_1(x)\cdot \nabla \delta _f,P_k\delta _g + \delta _g\rangle \le C\left( G_1\right) \left\| \nabla \delta _f\right\| ^2 + \left\| \delta _g\right\| ^2 \le C\left( G_1\right) + \left\| \delta _g\right\| ^2 \end{aligned}$$
(A11)

since \(\nabla \delta _f\) is bounded in \(\Omega \).

For the term \(\langle K_g\delta _g, P\delta _g + \delta _g\rangle \), we get that

$$\begin{aligned} \begin{aligned} \langle K_g\delta _g, P\delta _g + \delta _g\rangle&=\langle \Delta \delta _g,P\delta _g \rangle + \langle \Delta \delta _g, \delta _g \rangle + \langle a_g\left( x\right) \nabla \delta _g,P\delta _g\rangle + \langle a_g\left( x\right) \nabla \delta _g,\delta _g\rangle \\&= \int _{\Omega }\Delta \delta _g\cdot P\delta _gdx + \int _{\Omega }\Delta \delta _g\cdot \delta _gdx + \int _{\Omega }a_g\left( x\right) \nabla \delta _g\cdot P\delta _gdx \\&\quad + \int _{\Omega }a_g\left( x\right) \nabla \delta _g\cdot \delta _gdx\\&= -\int _{\Omega }\nabla \delta _g\cdot \nabla \left( P\delta _g\right) dx - \int _{\Omega }\left| \nabla \delta _g\right| ^2dx + \int _{\Omega }a_g\left( x\right) \nabla \delta _g\cdot P\delta _gdx \\&\quad + \int _{\Omega }a_g\left( x\right) \nabla \delta _g\cdot \delta _gdx \\&= -\int _{\Omega }\nabla \delta _gP \nabla \delta _gdx - \int _{\Omega }\nabla \delta _g\cdot \left( \nabla P\right) \delta _gdx - \int _{\Omega }\left| \nabla \delta _g\right| ^2dx \\&\quad + \int _{\Omega }a_g\left( x\right) \nabla \delta _g\cdot P\delta _gdx + \int _{\Omega }a_g\left( x\right) \nabla \delta _g\cdot \delta _gdx. \end{aligned} \end{aligned}$$

Note that \(P \succ 0\) then \(\int _{\Omega }\nabla \delta _gP \nabla \delta _gdx \ge \lambda \left\| \nabla \delta _g\right\| ^2\). Applying the Young inequality once more for the terms

$$\begin{aligned} \int _{\Omega }\nabla \delta _g\cdot \left( \nabla P\right) \delta _gdx, \qquad \int _{\Omega }a_g\left( x\right) \nabla \delta _g\cdot P\delta _gdx, \qquad \int _{\Omega }a_g\left( x\right) \nabla \delta _g\cdot \delta _gdx, \end{aligned}$$

we have that

$$\begin{aligned} \begin{aligned} \langle K_g\delta _g, P\delta _g + \delta _g\rangle&\le - \left( 1 + \lambda \right) \left\| \nabla \delta _g\right\| ^2 + \left( 1 + \lambda \right) \left\| \nabla \delta _g\right\| ^2 + C\left( 1 + \lambda \right) \left\| \delta _g\right\| ^2 \end{aligned} \end{aligned}$$

which implies

$$\begin{aligned} \langle K_g\delta _g, P\delta _g + \delta _g\rangle \le C\left( 1 + \lambda \right) \left\| \delta _g\right\| ^2, \end{aligned}$$
(A12)

with \(C\left( 1 + \lambda \right) \) denoting a constant depending on \(1 + \lambda \).

Combining these equations (A9), (A10), (A11), and (A12), and noting that two norms \(\left\| \cdot \right\| _k\) and \(\left\| \cdot \right\| _k\) are equivalent, we observe that

$$\begin{aligned} \epsilon \dfrac{\partial }{\partial t}\left\| \delta _g\right\| ^2_k \le \left( 2\epsilon + \epsilon C\left( 1+\lambda \right) - 1\right) \left\| \delta _g\right\| ^2_k + C_1\left\| \delta _f\right\| \left\| \delta _g\right\| + C_1\epsilon \end{aligned}$$

which implies when \(\epsilon \) small enough

$$\begin{aligned} \epsilon \dfrac{\partial }{\partial t}\left\| \delta _g\right\| ^2_k \le -q\left\| \delta _g\right\| ^2_k + C_1\left\| \delta _f\right\| \left\| \delta _g\right\| + C_1\epsilon \end{aligned}$$
(A13)

Thus, combine (A7) and (A13) and we obtain that

$$\begin{aligned} \dfrac{d}{dt}\left( \left\| \delta _f\right\| ^2 + \epsilon \dfrac{C_1}{q}\left\| \delta _g\right\| ^2\right) \le C_1\left\| \delta _f\right\| ^2 - \left\| \delta _g\right\| ^2_k + C_1\epsilon . \end{aligned}$$
(A14)

for some constant \(C_1\) independent of k.

By the Gronwall’s inequality for \(\left( \left\| \delta _f\right\| ^2 + \frac{\epsilon C_1}{q}\left\| \delta _g\right\| _k^2\right) \), for each \(k \ge 1\), we can regard \(\tau _{k-1}\) as the initial value, and then deduce that

$$\begin{aligned} \left\| \delta _f\left( \tau _{k-1} + \tau \right) \right\| ^2 \le e^{C_3\tau }\left( \left\| \delta _f\left( x,\tau _{k-1}\right) \right\| ^2 + \epsilon \dfrac{C_1}{q}\left\| \delta _g\left( x,\tau _{k-1}\right) \right\| _k^2\right) dx + C_1\epsilon \end{aligned}$$

for \(\tau \in [0,\tau _k - \tau _{k-1}]\). With the aid of this bound for the growth of \(|\delta _f|\), the second inequality of (A6) implies a bound for \(\left\| \delta _g\right\| _k\) as following

$$\begin{aligned} \begin{aligned}&\left\| \delta _g\left( \tau _{k-1} + \tau \right) \right\| ^2_kdx \le e^{-q\tau /\epsilon }\left\| \delta _g\left( \tau _{k-1}\right) \right\| ^2_k\\&\quad + C_4\left( \left\| \delta _f\left( x,\tau _{k-1}\right) \right\| ^2dx + \epsilon \dfrac{C_1}{q}\left\| \delta _g\left( x,\tau _{k-1}\right) \right\| ^2_k\right) \\&\quad + C_4\epsilon . \end{aligned} \end{aligned}$$

We already have that \(\delta _f\left( x,t_0\right) = \delta _f\left( x,\tau _0\right) \le \epsilon \) and \(\delta _g\left( x,t_0\right) = \delta _g\left( x,\tau _0\right) \le \epsilon _0\) for \(\epsilon _0\) small enough. Then, by the compactness of \(\bar{\Omega }\), for \(\tau \in \left[ 0,\tau _1 - \tau _0\right] \), \(\left\| \delta _f\left( \tau \right) \right\| ^2 \le O\left( \epsilon \right) \), for all \(x \in \Omega \). Make a process similarly and successively for \(k = 1,2,\dots \), we have that \(\left\| \delta _f\right\| ^2 \le O\left( \epsilon \right) \) for all \(x\in \Omega \). Analogously, we can also prove that \(\left\| \delta _g\right\| ^2 \le O\left( \epsilon \right) \).

Therefore, \(\int _{\Omega }\left| f(x,t) - f_0(x,t)\right| ^2dx \le C\epsilon \) and \(\int _{\Omega }|g(x,t) - g_0(x,t)|^2dx \le C\epsilon \), and we have the conclusion of the theorem. \(\square \)

Appendix B Proof for theorem 8

Proof

Note that \(\left\| F\left( u_1,x\right) - F\left( u_2,x\right) \right\| \le C\left\| u_1 - u_2\right\| ,\forall u_1,u_2 \in D\left( F\right) \) and \(|G\left( u,x\right) v|\) is bounded, \(\forall u,v\) bounded due to the continuous differentiability of G in a bounded domain. Consider

$$\begin{aligned} \begin{aligned} \dfrac{1}{2}\dfrac{\partial }{\partial t} |u -v|^2&= \left( u-v\right) \dfrac{\partial }{\partial t}\left( u-v\right) \\&= \left( u -v\right) \left[ F\left( u,x\right) - F\left( v,x\right) \right] + \epsilon \left( u -v\right) G(u,x) + \epsilon \left( u - v\right) \Delta \left( u - v\right) \\&\le C|u - v|^2 + O\left( \epsilon \right) + \epsilon \left( u - v\right) \Delta \left( u - v\right) . \end{aligned} \end{aligned}$$
(B1)

Taking the integral of (B1) over \(\Omega \) and using the Neumann boundary condition implies that

$$\begin{aligned} \begin{aligned} \dfrac{1}{2}\dfrac{\partial }{\partial t}\int _{\Omega } |u -v|^2dx \le C\int _{\Omega }|u-v|^2dx + O\left( \epsilon \right) - \epsilon \int _{\Omega }\left\| \nabla \left( u - v\right) \right\| ^2dx, \end{aligned} \end{aligned}$$

which leads to

$$\begin{aligned} \dfrac{\partial }{\partial t}\int _{\Omega } |u -v|^2dx \le C\int _{\Omega }|u-v|^2dx + O\left( \epsilon \right) . \end{aligned}$$

Applying Gronwall’s inequality, and we have that

$$\begin{aligned} \int _{\Omega } |u -v|^2dx \le O\left( \epsilon \right) e^{Ct^2}, \end{aligned}$$

which implies \(\int _{\Omega } |u -v|^2dx = O\left( \epsilon \right) \) for all \(t < \mathcal {T}\) with given \(\mathcal {T} > 0\), by the compactness of \(\bar{\Omega }\). \(\square \)

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

Le, T.M.T., Madec, S. Spatiotemporal Evolution of Coinfection Dynamics: A Reaction–Diffusion Model. J Dyn Diff Equat (2023). https://doi.org/10.1007/s10884-023-10285-z

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s10884-023-10285-z

Keywords

Navigation