Skip to main content
Log in

Quasi-neutral dynamics in a coinfection system with N strains and asymmetries along multiple traits

  • Published:
Journal of Mathematical Biology Aims and scope Submit manuscript

Abstract

Understanding the interplay of different traits in a co-infection system with multiple strains has many applications in ecology and epidemiology. Because of high dimensionality and complex feedback between traits manifested in infection and co-infection, the study of such systems remains a challenge. In the case where strains are similar (quasi-neutrality assumption), we can model trait variation as perturbations in parameters, which simplifies analysis. Here, we apply singular perturbation theory to many strain parameters simultaneously and advance analytically to obtain their explicit collective dynamics. We consider and study such a quasi-neutral model of susceptible–infected–susceptible (SIS) dynamics among N strains, which vary in 5 fitness dimensions: transmissibility, clearance rate of single- and co-infection, transmission probability from mixed coinfection, and co-colonization vulnerability factors encompassing cooperation and competition. This quasi-neutral system is analyzed with a singular perturbation method through an appropriate slow–fast decomposition. The fast dynamics correspond to the embedded neutral system, while the slow dynamics are governed by an N-dimensional replicator equation, describing the time evolution of strain frequencies. The coefficients of this replicator system are pairwise invasion fitnesses between strains, which, in our model, are an explicit weighted sum of pairwise asymmetries along all trait dimensions. Remarkably these weights depend only on the parameters of the neutral system. Such model reduction highlights the centrality of the neutral system for dynamics at the edge of neutrality and exposes critical features for the maintenance of diversity.

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

Similar content being viewed by others

Notes

  1. To be precise the slow manifold is an invariant of the full system for small \(\epsilon >0\) with equation \(x = \phi ^{\epsilon }(y,t)\). Here we use for simplicity the name slow manifold with \(\phi =\phi ^0\).

  2. The name neutral system comes from the fact that if \(\epsilon =0\) then the parameters do not depend on the strains as in the neutral theory, and the model describes indistinguishable strains.

  3. From the neutral analysis, we know that the order zero terms will cancel each other out, but there is no reason that the same occurs for the subsisting order 1 terms.

  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.

  5. Recall that \(m=\gamma +r\), \(R_0=\frac{\beta }{m}\) and \(T=1-\frac{1}{R_0}\).

  6. For instance if the attractor domains do not depend on \(\epsilon \).

  7. Recall that the solution \({\bar{x}} = \phi (y,\tau )\) is asymptotically stable if for each \(\tau _0 > 0\), a \(\delta (\tau _0)\) can be found such that: \(\Vert x_0 - \phi (y,\tau _0)\Vert \le \delta (\tau _0)\) yields \(\lim \nolimits _{\tau \rightarrow \infty }\Vert x(\tau ;\tau _0,y_0) - \phi (y,\tau )\Vert = 0\).

References

  • Adler FR, Brunet RC (1991) The dynamics of simultaneous infections with altered susceptibilities. Theor Popul Biol 40(3):369–410

    Article  MATH  Google Scholar 

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

    Article  Google Scholar 

  • Alizon S, de Roode JC, Michalakis Y (2013) Multiple infections and the evolution of virulence. Ecol Lett 16(4):556–567

    Article  Google Scholar 

  • Allesina S, Levine JM (2011) A competitive network theory of species diversity. Proc Natl Acad Sci 108(14):5638–5642

    Article  Google Scholar 

  • Alonso D, McKane AJ, Pascual M (2007) Stochastic amplification in epidemics. J R Soc Interface 4:575–582

    Article  Google Scholar 

  • Balmer O, Tanner M (2011) Prevalence and implications of multiple-strain infections. Lancet Infect Dis 11:868–878

    Article  Google Scholar 

  • Bartlett MS (1957) Measles periodicity and community size. J R Stat Soc A 120:48–70

    Article  Google Scholar 

  • Birger R, Kouyos R, Dushoff J, Grenfell B (2015a) Modeling the effect of HIV co-infection on clearance and sustained virologic response during treatment for hepatitis C virus. Epidemics 12:1–10 (Papers arising from Epidemics 4)

  • Birger RB, Kouyos RD, Cohen T, Griffiths EC, Huijben S, Mina MJ, Volkova V, Grenfell B, Metcalf CJE (2015b) The potential impact of coinfection on antimicrobial chemotherapy and drug resistance. Trends Microbiol 23(9):537–544

  • Bratus AS, Posvyanskii VP, Novozhilov AS (2014) Replicator equations and space. Math Model Nat Phenom 9(3):47–67

    Article  MathSciNet  MATH  Google Scholar 

  • Bremermann HJ, Thieme H (1989) A competitive exclusion principle for pathogen virulence. J Math Biol 27(2):179–190

    Article  MathSciNet  MATH  Google Scholar 

  • Cardin PT, Teixeira MA (2017) Fenichel theory for multiple time scale singular perturbation problems. SIAM J Appl Dyn Syst 16:1425–1452

    Article  MathSciNet  MATH  Google Scholar 

  • Chawanya T, Tokita K (2002) Large-dimensional replicator equations with antisymmetric random interactions. J Phys Soc Jpn 71(2):429–431

    Article  Google Scholar 

  • Chen L, Ghanbarnejad F, Brockmann D (2017) Fundamental properties of cooperative contagion processes. New J Phys 19(10):103041

    Article  MATH  Google Scholar 

  • Cobey S, Lipsitch M (2012) Niche and neutral effects of acquired immunity permit coexistence of pneumococcal serotypes. Science (New York, NY) 335:1376–1380

    Article  Google Scholar 

  • Davies NG, Flasche S, Jit M, Atkins KE (2019) Within-host dynamics shape antibiotic resistance in commensal bacteria. Nat Ecol Evol 3(3):440

    Article  Google Scholar 

  • Dawes JHP, Gog JR (2002) The onset of oscillatory dynamics in models of multiple disease strains. J Math Biol 45:471–510

    Article  MathSciNet  MATH  Google Scholar 

  • Faust K, Raes J (2012) Microbial interactions: from networks to models. Nat Rev Microbiol 10:538–550

    Article  Google Scholar 

  • Fenichel N (1979) Geometric singular perturbation theory for ordinary differential equations. J Differ Equ 31(1):53–98

    Article  MathSciNet  MATH  Google Scholar 

  • Ferguson N, Andreasen V (2002) The influence of different forms of cross-protective immunity on the population dynamics of antigenically diverse pathogens. In: Mathematical approaches for emerging and reemerging infectious diseases: models, methods, and theory. Springer, New York

  • Gjini E, Madec S (2017) A slow–fast dynamic decomposition links neutral and non-neutral coexistence in interacting multi-strain pathogens. Theor Ecol 10:129–141

    Article  Google Scholar 

  • Gjini E, Madec S (2021a) The ratio of single to co-colonization is key to complexity in interacting systems with multiple strains. Ecol Evol. https://doi.org/10.1002/ece3.7259

  • Gjini E, Madec S (2021b) Towards a mathematical understanding of colonization resistance. bioRxiv

  • Gjini E, Valente C, Sá-Leão R, Gomes MGM (2016) How direct competition shapes coexistence and vaccine effects in multi-strain pathogen systems. J Theor Biol 388:50–60

    Article  MATH  Google Scholar 

  • Gog J, Grenfell BT (2002) Dynamics and selection of many-strain pathogens. Proc Natl Acad Sci USA 99:17209–17214

    Article  Google Scholar 

  • Golubitsky M, Stewart I (2002) The symmetry perspective: from equilibrium to chaos in phase space and physical space. Birkhauser, Berlin

    Book  MATH  Google Scholar 

  • Grilli J, Adorisio M, Suweis S et al (2017) Feasibility and coexistence of large ecological communities. Nat Commun 8:14389

    Article  Google Scholar 

  • Hofbauer J, Sigmund K (1998) Evolutionary games and population dynamics. Cambridge University Press, Cambridge

    Book  MATH  Google Scholar 

  • Kucharski AJ, Andreasen V, Gog JR (2016) Capturing the dynamics of pathogens with many strains. J Math Biol 72:1–24

    Article  MathSciNet  MATH  Google Scholar 

  • Kuehn C (2010) Multiple time scale dynamics, vol 191. Applied mathematical sciences book series (AMS). Springer, Berlin

    MATH  Google Scholar 

  • Le TMT, Madec S. Spatiotemporal evolution of coinfection dynamics: a reaction–diffusion model. J Dyn Differ Equ to appear

  • Le T-M-T, Gjini E, Madec S (2021) Disentangling how multiple traits drive 2 strain frequencies in sis dynamics with coinfection. biorXiv. https://doi.org/10.1101/2021.04.29.442023

    Article  MATH  Google Scholar 

  • Lipsitch M (1997) Vaccination against colonizing bacteria with multiple serotypes. PNAS 94(12):6571–6576

    Article  Google Scholar 

  • Lipsitch M, Colijn C, Cohen T, Hanage WP, Fraser C (2009) No coexistence for free: neutral null models for multistrain pathogens. Epidemics 1(1):2–13

    Article  Google Scholar 

  • Lourenço J, Recker M (2013) Natural, persistent oscillations in a spatial multi-strain disease system with application to dengue. PLoS Comput Biol 9:e1003308

    Article  Google Scholar 

  • Ma J, Ma Z (2006) Epidemic threshold conditions for seasonally forced SEIR models. Math Biosci Eng 3:161

    Article  MathSciNet  MATH  Google Scholar 

  • Madec S, Gjini E (2020) Predicting n-strain coexistence from co-colonization interactions: epidemiology meets ecology and the replicator equation. Bull Math Biol 82:142

    Article  MathSciNet  MATH  Google Scholar 

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

    Article  MathSciNet  MATH  Google Scholar 

  • McKane AJ, Newman TJ (2005) Predator–prey cycles from resonant amplification of demographic stochasticity. Phys Rev Lett 94:218102

    Article  Google Scholar 

  • Mosquera J, Adler FR (1998) Evolution of virulence: a unified framework for coinfection and superinfection. J Theor Biol 195(3):293–313

    Article  Google Scholar 

  • Murray JD (2002) Mathematical biology, I. An introduction. Springer, Berlin

    Book  MATH  Google Scholar 

  • Pinotti F, Ghanbarnejad F, Hövel P, Poletto C (2019) Interplay between competitive and cooperative interactions in a three-player pathogen system. arXiv preprint arXiv:1912.07289

  • Teschl G (2010) Ordinary differential equations and dynamical systems, vol 140. Graduate studies in mathematics. American Mathematical Society, Providence

    MATH  Google Scholar 

  • Thieme HR (2007) Pathogen competition and coexistence and the evolution of virulence. In: Mathematics for life science and medicine. Springer, pp 123–153

  • Tikhonov AN (1952) Systems of differential equations containing a small parameter multiplying the derivative. Mat SB (NS) 31:575–586

    Google Scholar 

  • van Baalen M, Sabelis MW (1995) The dynamics of multiple infection and the evolution of virulence. Am Nat 146(6):881–910

    Article  Google Scholar 

  • Verhulst F (1996) Nonlinear differential equations and dynamical systems, 2nd edn. Springer, Berlin

    Book  MATH  Google Scholar 

  • Warren DK, Nitin A, Hill C, Fraser VJ, Kollef MH (2004) 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

    Article  Google Scholar 

  • Wechselberger M (2020) Geometric singular perturbation theory beyond the standard form. Frontiers in applied dynamical systems: reviews and tutorials. Springer, Berlin

    Book  MATH  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Sten Madec.

Additional information

Publisher's Note

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

Appendices

Appendix

Neutral system

In Sect. 2.3, we only give formal proof in the analysis of the Neutral system. This formal proof is seen as a guide for the rigorous analysis of the normal form in Sect. 3. However, since we can do it, we give rigorous proof for some results of the neutral system.

We first start with a classical lemma.

Lemma 15

Let a and b belongs to \(C^1({\mathbb {R}})\). Let y be a solution of the linear differential equation \(y'=a(t)y+b(t)\) with \(y(0)=y_0\).

Assume that \(a(t)\rightarrow -a^*<0\) and \(b(t)\rightarrow 0\). Then \(y(t)\rightarrow 0\) as \(t\rightarrow +\infty .\)

Proof

The proof is classical and we give it for the sake of completeness. We have the explicit formula as follows

$$\begin{aligned} y(t)=e^{A(t)}y_0+\int _0^t e^{A(t)-A(s)}b(s) ds \end{aligned}$$

wherein we have a note

$$\begin{aligned} A(t)=\int _0^t a(s)ds. \end{aligned}$$

Let \(0<\epsilon < a^*\) be fixed and choose \(t_0>0\) such that for any \(t>t_0\), we have \(\left| a(t) + a^*\right| <\epsilon \) and \(\left| b(t)\right| \le \epsilon \).

We have

$$\begin{aligned} \left| y(t)\right| \le e^{(-a^* +\epsilon )(t-t_0)} e^{A(t_0)}\left( y_0+\int _0^{t_0} e^{-A(s)}|b(s)|ds\right) +\epsilon \int _{t_0}^t e^{(-a^* +\epsilon )(t-s)}ds \end{aligned}$$

This yields \(|y(t)|\le C_1e^{(-a^* +\epsilon )\left( t-t_0\right) }+C_2\epsilon \) for some positive constant \(C_1\) and \(C_2\), which ends the proof by taking \(\epsilon \rightarrow 0\) and \(t \rightarrow \infty \). \(\square \)

Lemma 16

Let \(S(t)\rightarrow S^*\) and \(T(t)\rightarrow T^*\). Let I satisfy \(I(0)\ge 0\) and

$$\begin{aligned} \dfrac{dI}{dt} = \beta T S - (m+\beta k T) I \end{aligned}$$

Then \(I(t)\rightarrow I^*:= \frac{mT^*}{m + \beta k T^*}\)

Proof

Indeed, substitute (ST) by \(\left( S^* + (S - S^*), T^* + (T - T^*)\right) \) into the equation of I in (2.6) to obtain

$$\begin{aligned} \dfrac{dI}{dt}= & {} m T^*-(m+\beta k T^*)I + \left[ \beta S^*\left( T - T^*\right) + \beta T^*\left( S - S^*\right) \right. \nonumber \\{} & {} \left. + \beta \left( T - T^*\right) \left( S - S^*\right) \right] . \end{aligned}$$
(A.1)

Define \({\tilde{I}}\) as the solution of

$$\begin{aligned} \dfrac{d{\tilde{I}}}{dt} = \beta T^* S^* - (m+\beta k T^*) {\tilde{I}}. \end{aligned}$$

It is clear that \({\tilde{I}}(t)\rightarrow I^*\). The equation for \(I - {\tilde{I}}\) reads

$$\begin{aligned} \dfrac{d}{dt}\left( I - {\tilde{I}}\right) = g(t)\left( I - {\tilde{I}}\right) + f(t). \end{aligned}$$
(A.2)

Wherein we have set \(g(t)= -(m+\beta k T^*+\beta k (T-T^*))\) and \(f\left( t\right) = \beta S^*\left( T - T^*\right) + \beta T^*\left( S - S^*\right) + \beta \left( T - T^*\right) \left( S - S^*\right) +\beta k (T-T^*)I^*\).

By Proposition 3, we have \(g(t)\rightarrow -(m+\beta k T^*)<0\) and \(f\left( t\right) \rightarrow 0\) as \(t\rightarrow +\infty \). Thus, by the Lemma 15, we have \(I(t)-{\tilde{I}}(t)\rightarrow 0\) as \(t\rightarrow +\infty \) which ends the proof. \(\square \)

Lemma 17

Let \((S(t),T(t),I(t))\rightarrow (S^*,T^*,I^*)\) and define

$$\begin{aligned} A(t)=\begin{pmatrix} -(m+\beta k T(t)&{}\quad m\\ -\frac{\beta k T(t)}{2}&{}\quad -\frac{\beta k I(t)}{2} \end{pmatrix}\quad \text {and}\quad A^*=\begin{pmatrix} -(m+\beta k T^*&{}\quad m\\ -\frac{\beta k T^*}{2}&{}\quad -\frac{\beta k I^*}{2} \end{pmatrix} \end{aligned}$$

Define also

$$\begin{aligned} P = \begin{pmatrix} 2T^*&{}\quad I^* \\ D^*&{}\quad T^* \end{pmatrix}, \qquad P^{-1} = \dfrac{1}{|P|}\begin{pmatrix} T^* &{}\quad -I^* \\ -D^* &{}\quad 2T^* \end{pmatrix} \end{aligned}$$

such that \(P^{-1} A P= \begin{pmatrix} -\xi &{}\quad 0\\ 0&{}\quad 0\end{pmatrix}\) with \(\xi =m+\beta k T^*-\frac{1}{2} \beta k I^*\).

Let \((v_i,z_i)\) be bounded and verify

$$\begin{aligned} \dfrac{d}{dt}\begin{pmatrix} v_i\\ z_i\end{pmatrix}= P^{-1} A(t) P \begin{pmatrix} v_i\\ z_i\end{pmatrix}. \end{aligned}$$

Then \(v_i(t)\rightarrow 0\) and \(z_i'(t)\rightarrow 0\).

Proof

This is a direct application of the Lemma 15 on the equation of \(v_i\) once we note that \(z_i\) is bounded. \(\square \)

Remark 18

Note that it is wrong that \(z_i(t)\rightarrow z_i^*\) for some constant \(z_i^*\). Indeed, \(z_i\) is susceptible to following very slow dynamics. However, for the application in the slow–fast system (\(\epsilon >0\)), this very slow perturbation is below the order \(\epsilon \) and has no effect.

Lemma 19

Let \(I_{ij}\), \(i,j \in \{1,\dots ,N\}\), satisfy the neutral system (2.14) and \(z_i\) verify the system (2.13). Assume that for each (ij), \(\left( {\tilde{I}}_i (t),{\tilde{J}}_j(t)\right) = \left( I^*z_i(t), T^*z_j(t)\right) \), satisfy that \(\left| I_i(t) - {\tilde{I}}_i(t)\right| \rightarrow 0\) and \(\left| J_j(t) - {\tilde{J}}_j(t) \right| \rightarrow 0\). Then \(\left| I_{ij}(t) - \frac{\beta k}{m}I^*T^*z_i(t) z_j(t)\right| \rightarrow 0\) as \(t \rightarrow \infty \).

Proof

Consider the equation

$$\begin{aligned} \dfrac{d{\tilde{I}}_{ij}}{dt} = - m {\tilde{I}}_{ij} + \beta k {\tilde{I}}_i {\tilde{J}}_j \end{aligned}$$
(A.3)

then we can obtain the differential equation for \(I_{ij} - {\tilde{I}}_{ij}\), recall \(I_{ij}\) satisfies (2.14),

$$\begin{aligned} \dfrac{d}{dt}\left( I_{ij} - {\tilde{I}}_{ij}\right)= & {} - m \left( I_{ij} - {\tilde{I}}_{ij}\right) + \beta k \left[ \left( I_i - {\tilde{I}}_i\right) {\tilde{J}}_j \right. \nonumber \\{} & {} \left. + \left( J_j - {\tilde{J}}_j\right) {\bar{I}}_i + \left( I_i - {\tilde{I}}_i\right) \left( J_j - {\tilde{J}}_j\right) \right] . \end{aligned}$$
(A.4)

Since \(\left| I_i(t) - {\tilde{I}}_i(t)\right| \rightarrow 0\) and \(\left| J_i(t) - {\tilde{J}}_i(t)\right| \rightarrow 0\), then

$$\begin{aligned} \left| \left( I_i - {\tilde{I}}_i\right) {\tilde{J}}_j + \left( J_j - {\tilde{J}}_j\right) {\bar{I}}_i + \left( I_i - {\tilde{I}}_i\right) \left( J_j - {\tilde{J}}_j\right) \right| \rightarrow 0. \end{aligned}$$

Using Lemma 15, we have that \(\left| I_{ij}(t) - {\tilde{I}}_{ij}(t)\right| \rightarrow 0\) as \(t \rightarrow +\infty \).

We then compute the solution \({\tilde{I}}_{ij}\) of (A.3) to be

$$\begin{aligned} {\tilde{I}}_{ij}\left( t\right)&= e^{-mt}\left( \beta k I^*T^*\int _{0}^{t}e^{ms}z_i\left( s\right) z_j\left( s\right) ds + C\right) , \qquad C\in {\mathbb {R}}. \nonumber \\&= e^{-mt}\left( \dfrac{\beta k I^*T^*}{m}\left( e^{mt}z_i(t)z_j(t) - z_i(0)z_j(0) \right. \right. \nonumber \\&\quad \left. \left. - \int _{0}^{t}e^{ms}\dfrac{d}{dt}\left[ z_i\left( s\right) z_j\left( s\right) \right] ds\right) + C \right) , \end{aligned}$$
(A.5)

by integrating by parts. Note that \(z_i'(t) \rightarrow 0\), \(\forall i\) (Lemma 17), by a similar proof in Lemma 15, we can prove that \(e^{-mt}\int _{0}^{t}e^{ms}\frac{d}{dt}\left[ z_i\left( s\right) z_j\left( s\right) \right] ds \rightarrow 0\). Hence \({\tilde{I}}_{ij} \rightarrow \frac{\beta k I^* T^*}{m}z_i z_j\) as \(t \rightarrow \infty \), implying \(I_{ij} \rightarrow \frac{\beta k I^* T^*}{m}z_i z_j\), which ends this proof. \(\square \)

Tychonov’s theorems

Theorem 20

(see Tikhonov 1952) Consider the initial value problem

$$\begin{aligned} \left\{ \begin{aligned} \epsilon&\dfrac{dx}{d\tau } = f(x,y,\tau ) + \epsilon \dots , \quad&x(0) = x_0, \quad&x \in D \subset {\mathbb {R}}^n,\\&\dfrac{dy}{d\tau } = g(x,y,\tau ) + \epsilon \dots , \quad&y(0) = y_0,\quad&y \in G \subset {\mathbb {R}}^n. \end{aligned} \right. \end{aligned}$$
(B.1)

For f and g, we take sufficiently smooth vector functions in x, y and t; the dots represent (smooth) higher-order terms in \(\epsilon \).

  1. (a)

    We assume that a unique solution to the initial value problem exists and suppose this also holds for the reduced problem

    $$\begin{aligned} \left\{ \begin{aligned}&0= f(x,y,\tau ),\\&\dfrac{dy}{d\tau } = g(x,y,\tau ), \quad&y(0) = y_0, \end{aligned} \right. \end{aligned}$$
    (B.2)

    with solution \({\bar{x}}(\tau )\), \({\bar{y}}(\tau )\).

  2. (b)

    Suppose that \(0 = f(x,y,\tau )\) is solved by \({\bar{x}} = \phi (y,\tau )\), where \(\phi (y,\tau )\) is a continuous function and an isolated root, i.e., there exists a neighborhood of \(\phi (y,\tau )\) such that there is no other solution for \(0 = f(x,y,\tau )\) in this vicinity. Also, suppose that \({\bar{x}} = \phi (y,t)\) is an asymptotically stable solutionFootnote 7 of the equation \(\frac{dx}{dt} = g(x,y,\tau )\), where \(\tau = \epsilon t\), that is uniform in the parameters \(y \in G\) and \(\tau \in {\mathbb {R}}^+\).

  3. (c)

    x(0) is contained in an interior subset of the domain of attraction of \({\bar{x}} = \phi (y,\tau )\) in the case of initial conditions \(y = y(0)\), \(\tau = 0\).

Then, we have

$$\begin{aligned} \begin{aligned}&\lim \limits _{\epsilon \rightarrow 0} x_\epsilon (\tau ) = {\bar{x}}(\tau ), \quad&0 <\tau _0 \le \tau \le T,\\&\lim \limits _{\epsilon \rightarrow 0} y_\epsilon (\tau ) = {\bar{y}}(\tau ), \quad&0 \le \tau \le T, \end{aligned} \end{aligned}$$
(B.3)

with \(\tau _0\) and T are constants independent on \(\epsilon \).

Besides this Tychonov’s theorem, it needs to use another result that allows us to approximate the original system by the slow–fast form. The following error estimate gives a more precise description of these limits. (theorem 9.1, Verhulst 1996 adapted here for the simple case \(m=0\)).

Theorem 21

(see Verhulst 1996) Consider the initial value problem

$$\begin{aligned} \dfrac{dx}{dt} = f_0(t,x) +\epsilon R(t,x,\epsilon ) \end{aligned}$$
(B.4)

with \(x(t_0) = \eta \) and \(|t - t_0| \le h\), \(x \in D \subset {\mathbb {R}}^n\), \(0 \le \epsilon \le \epsilon _0\). Assume that in this domain we have

  1. (a)

    f(tx) continuous in t and x, 2 times continuously differentiable in x;

  2. (b)

    \(R(t,x,\epsilon )\) continuous in tx and \(\epsilon \), Lipschitz-continuous in x.

Let \(x_0(t)\) be the solution of

$$\begin{aligned} \dfrac{dx}{dt} = f_0(t,x) \end{aligned}$$
(B.5)

with \(x_0(t_0) = \eta \) Let \(T>0\) and assume that both x and \(x_0\) are defined on [0, T] for any \(\epsilon \in (0,\epsilon _0)\). There exist \(C>0\) (depending on T) such that for any \(\epsilon \in (0,\epsilon _0)\), and \(t\in (0,T)\), we have the estimate

$$\begin{aligned} \left\| x(t) -x_0(t) \right\| \le C \epsilon \end{aligned}$$
(B.6)

1.1 Proof of Lemma 7 of error estimate

In this subsection, we estimate error through three steps when \(\epsilon \) is small enough as follows.

  1. 1.

    Lemma 22: approximate the solution S, \(I_i\) of the original dynamics (2.1) using the solution of the slow–fast form (3.9)

  2. 2.

    Lemma 23: approximate the solution of the slow–fast form (3.9) using the solution of the slow system (3.14)

  3. 3.

    Lemma 24: approximate the solution \(\left( I_{ij}\right) _{i,j}\) of (2.1) using the solution \({\textbf{z}}\) of (3.14).

Lemma 22

Under our assumptions, for any initial values of (2.1), there exists \(\tau _0 > 0\) and initial value \({\textbf{z}}\left( \tau _0\right) \) of (3.9), such that for any \(T>\tau _0\), there are \(\epsilon _0 > 0\) and \(C_T > 0\) which satisfy that \(\forall \epsilon < \epsilon _0\)

$$\begin{aligned} \left| S\left( \dfrac{\tau }{\epsilon } \right) - S^*\right| + \sum _{i=1}^{N}\left| I^*z_i(\tau ) - I_i\left( \dfrac{\tau }{\epsilon } \right) \right| + \sum _{i=1}^{N}\left| T^*z_i(\tau ) - J_i\left( \dfrac{\tau }{\epsilon } \right) \right| \le \epsilon C_T,\nonumber \\ \end{aligned}$$
(B.7)

for all \(\tau _0 \le \tau \le T\), where \(\left( S, I_i, J_i\right) _{i = 1,\dots ,N}\) verifies (2.1) and \((z_1, \dots , z_N)\) is the solution of (3.9).

Proof

  • First step, we wish to estimate the error between \(S^*\), \(T^*\), \(I^*\) and the solution \(\left( S,T,I\right) \) of (2.5).

    Denote \(x = \left( S, I_1, \dots , I_N, J_1, \dots , J_N \right) \), by (2.5), we can rewrite the system for x in the form of \(\frac{dx}{dt} = f_0(t,x) + \epsilon f_1(t,x)\), where \(f_0(t,x)\) and \(f_1(t,x)\) are polynomial in multiple variables \(\left( S, T, I_i, J_i \right) \), \(i = 1,\dots ,N\), implying that \(f_0(t,x)\) and \(f_1(t,x)\) satisfy conditions of Theorem 21.

    Hence, if \(x^r = \left( S^r, T^r, I^r_i, J^r_i \right) \) satisfies the neutral system (2.6) and \(x = \left( S, T, I_i, J_i \right) _{1\le i \le N}\) satisfies (2.5), then \(\left\| x - x^r\right\| _{{\mathbb {R}}^{2N+2}} = O(\epsilon )\).

    Combine with the arguments in Sect. 2.3, the approximation of solution (STI) of (2.5) by \(\left( S^*,T^*,I^*\right) \) is accordingly plausible in the sense of \(O\left( \epsilon \right) \). We have done our first step.

  • Second step, we claim that all the algebraic and linear transformations from (2.5) to (3.9) are equivalent with error estimate \(O\left( \epsilon \right) \), including changing \(\left( S,T,I\right) \) to \(\left( X,Y\right) \) using \(S^*, T^*, I^*\) (proved in the first part), changing \(\begin{pmatrix} I_i \\ J_i \end{pmatrix}\) to \(\begin{pmatrix} v_i \\ z_i \end{pmatrix}\) (linear operator) and changing to time scale \(\tau = \epsilon t\) with re-denote \(z\left( \tau \right) \) (see the argument in (3.9)).

    By a similar argument to verify the conditions of Expansion Theorem 21, and note that \(v\left( \tau \right) \rightarrow 0\) asymptotically, by Theorem 21 we have that

    $$\begin{aligned} \sum _{i=1}^{N}\left| I^*z_i(\tau ) - I_i\left( \dfrac{\tau }{\epsilon } \right) \right| + \sum _{i=1}^{N}\left| T^*z_i(\tau ) - J_i\left( \dfrac{\tau }{\epsilon } \right) \right| = O\left( \epsilon \right) , \end{aligned}$$

    for all \(\tau _0 \le \tau \le T\), where \(\left( I_i, J_i\right) _{i=1,\dots ,N}\) verify (2.1) and \((z_1, \dots , z_N)\) is the solution of (3.9).

Combining two parts, we have the conclusion for this lemma. \(\square \)

Lemma 23

The solution \({\textbf{z}} = \left( z_i\right) \) of the slow–fast form system (3.9) tends to the solution of the slow system (3.14) as \(\epsilon \rightarrow 0\) locally uniformly in time on \([\tau _0,T]\), with \(\tau _0 > 0\), \(T > \tau _0\) arbitrarily and independent on \(\epsilon \).

Proof

It suffices to verify the conditions for Tychonov’s theorem; see Theorem (20).

  • Firstly, we prove that (3.9) and (3.14) with initial values possess the unique solution.

    The right-hand side of (3.9) is a vector function with all the components being polynomials of variables \(\left( X,Y,{\textbf{L}},{\textbf{v}},{\textbf{z}}\right) \) (explicitly computed in Sects. 2.4, 3.3 and 3.3.1) in the bounded set [0, T] of time. Hence, it is global Lipschitz and continuously differentiable, implying the uniqueness of solution for (3.9), according to the Picard–Lindelof Theorem (Theorem 2.2 in Teschl 2010). Similarly, (3.14) has the unique solution

  • Secondly, by the proof of Lemma 6, we have that the solution \((X,Y,{\textbf{L}},{\textbf{v}},{\textbf{z}})\) of (3.7) with any initial condition \((X,Y,{\textbf{L}},{\textbf{v}},{\textbf{z}})(0)=(X_0,Y_0,{\textbf{L}}_0,{\textbf{v}}_0,{\textbf{z}}_0)\in {\mathbb {R}}\times {\mathbb {R}}\times \left( {\mathbb {R}}^n\right) ^3\) verifies \({\textbf{z}}(t)={\textbf{z}}_0\) for all \(t\ge 0\) and \(\lim _{t\rightarrow +\infty }(X,Y,{\textbf{L}},{\textbf{v}})(t)={\varvec{\Phi }}({\textbf{z}}_0)\) asymptotically, in which, \(\varPhi \left( {\textbf{z}}\right) = \left( X^*({\textbf{z}}),Y^*({\textbf{z}}),\chi _3{\textbf{L}}^*({\textbf{z}}),0\right) \), satisfy the system (3.9) in slow timescale, with \(\epsilon = 0\) as follows

    $$\begin{aligned} \left\{ \begin{aligned}&0 = -\beta T^*X +\chi _1 F_X^1\left( {\textbf{v}},{\textbf{z}}\right) +\chi _2 F_X^2\left( {\textbf{v}},{\textbf{z}}\right) +\chi _3 F_X^3\left( {\textbf{L}}\right) + O(\epsilon )\\&0 = \beta (S^* - T^* -k I^*)X - (m+\beta k T^*)Y +\chi _1 F_Y^1\left( {\textbf{v}},{\textbf{z}}\right) \\&\qquad \qquad +\chi _2 F_Y^2\left( {\textbf{v}},{\textbf{z}}\right) +\chi _5 F_Y^5\left( {\textbf{v}},{\textbf{z}}\right) \\&0 =-m L_i + \chi _3 F_{L_i} \left( {\textbf{v}},{\textbf{z}}\right) \\&0 = -\xi v_i \end{aligned} \right. \end{aligned}$$
    (B.8)

Applying Tychonov’s Theorem, we have the required conclusion. \(\square \)

By two Lemmas 23 and 24, we have that

$$\begin{aligned} \left| S\left( \dfrac{\tau }{\epsilon } \right) - S^*\right| + \sum _{i=1}^{N}\left| I^*z_i(\tau ) - I_i\left( \dfrac{\tau }{\epsilon } \right) \right| + \sum _{i=1}^{N}\left| T^*z_i(\tau ) - J_i\left( \dfrac{\tau }{\epsilon } \right) \right| \le \epsilon C_T,\nonumber \\ \end{aligned}$$
(B.9)

for all \(\tau _0 \le \tau \le T\), where \(\left( S, I_i, J_i\right) _{i = 1,\dots ,N}\) verifies (2.1) and \((z_1, \dots , z_N)\) is the solution of (3.14).

Finally, we will find an approximation of \(I_{ij}\), \(1 \le i \le N\) and estimate the error.

Lemma 24

Under our assumptions, for any initial values of (2.1), there exists \(\tau _0 > 0\) and initial value \({\textbf{z}}\left( \tau _0\right) \) of (3.14), such that for any \(T>\tau _0\), there is \(\epsilon _0 > 0\) and \(C_T > 0\), which satisfy that \(\forall \epsilon < \epsilon _0\)

$$\begin{aligned} \sum _{i,j = 1}^{N}\left| I_{ij}\left( \dfrac{\tau }{\epsilon }\right) - D^*z_i\left( \tau \right) z_j\left( \tau \right) \right| \le \epsilon C_T, \end{aligned}$$
(B.10)

for all \(\tau _0 \le \tau \le T\), \((I_{ij})_{1 \le i,j \le N}\) satisfying (2.1) and \((z_1, \dots , z_N)\) is the solution of the reduced system (3.14).

Proof

This proof is similar to the proof of Lemma 19. Denoting \({\textbf{z}}\) be the solution of (3.14), for \(1 \le i,j \le N\), denote \(I^r_{ij}\) to be the solution of

$$\begin{aligned} \dfrac{dI^r_{ij}\left( t\right) }{dt} = - mI^r_{ij}\left( t\right) + \beta k I^*T^*z_i\left( \epsilon t\right) z_j\left( \epsilon t\right) , \end{aligned}$$
(B.11)

Then, for each \(\tau _0>0\) and \(T>\tau _0\), after the changing time scale \(\tau = \epsilon t\), from (2.1) and (B.11), we have that

$$\begin{aligned} \begin{aligned} \dfrac{d}{dt}\left( I_{ij}\left( \dfrac{\tau }{\epsilon } \right) - I^r_{ij}\left( \dfrac{\tau }{\epsilon } \right) \right) = -&m\left( I_{ij}\left( \dfrac{\tau }{\epsilon }\right) - I^r_{ij}\left( \dfrac{\tau }{\epsilon } \right) \right) - \epsilon \gamma u_{ij} I_{ij}\left( \dfrac{\tau }{\epsilon }\right) \\&+ \left( \beta _jk_{ij}I_i\left( \dfrac{\tau }{\epsilon } \right) J_j\left( \dfrac{\tau }{\epsilon } \right) - \beta k I^* T^* z_i\left( \tau \right) z_j\left( \tau \right) \right) . \end{aligned}\nonumber \\ \end{aligned}$$
(B.12)

By (B.9), we have that \(\left| \beta _jk_{ij} I_i\left( \frac{s}{\epsilon }\right) J_j\left( \frac{s}{\epsilon }\right) - \beta k I^* T^* z_i(s) z_j(s)\right| = O(\epsilon )\) uniformly for \(s \in [\tau _0,T]\). Note that \(\left| I_{ij}\right| \le 1\), then using the expansion theorem- Theorem 21, we observe that \(\left| I_{ij}\left( \frac{\tau }{\epsilon }\right) - I^r_{ij}\left( \frac{\tau }{\epsilon }\right) \right| = O(\epsilon )\).

We then compute the solution \(\left( I^r_{ij}\right) _{1\le i,j \le N}\) of (B.11) to be

$$\begin{aligned} I^r_{ij}\left( t\right) = e^{-mt}\left( \beta k I^*T^*\int _{0}^{t}e^{ms}z_i\left( \epsilon s\right) z_j\left( \epsilon s\right) ds + C\right) , \qquad C\in {\mathbb {R}}. \end{aligned}$$
(B.13)

Using integrating by parts, (B.13) becomes

$$\begin{aligned} I^r_{ij}(t)= & {} e^{-mt}\left( \beta k I^*T^*\cdot \dfrac{1}{m}\left( e^{mt}z_i(\tau )z_j(\tau ) - z_i(0)z_j(0) \right. \right. \nonumber \\{} & {} \left. \left. - \int _{0}^{t}e^{ms}\dfrac{d}{dt}\left[ z_i\left( \epsilon s\right) z_j\left( \epsilon s\right) \right] ds\right) + C \right) \end{aligned}$$
(B.14)

By (3.14), we deduce that \(\dfrac{d}{dt}\left[ z_i\left( \epsilon s\right) z_j\left( \epsilon s\right) \right] = O\left( \epsilon \right) \), implying \(\left| I^r_{ij}\left( \frac{\tau }{\epsilon }\right) - k\frac{I^*T^*}{S^*}\right. \left. z_i(\tau )z_j(\tau )\right| \rightarrow O\left( \epsilon \right) \) as \(t \rightarrow \infty \). Hence, noting that \(k\frac{I^*T^*}{S^*} = D^*\), we have \(\left| I^r_{ij}\left( t\right) - D^*z_i\left( \epsilon t\right) z_j\left( \epsilon t\right) \right| = O\left( \epsilon \right) \), implying \(\left| I_{ij}\left( \frac{\tau }{\epsilon } \right) - D^*z_i\left( \tau \right) z_j\left( \tau \right) \right| = O\left( \epsilon \right) \), for all \(1 \le i,j \le N\). \(\square \)

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., 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). https://doi.org/10.1007/s00285-023-01977-7

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s00285-023-01977-7

Keywords

Mathematics Subject Classification

Navigation