Skip to main content
Log in

Spread rates of a juvenile-adult population in constant and temporally variable environments

  • ORIGINAL PAPER
  • Published:
Theoretical Ecology Aims and scope Submit manuscript

Abstract

The question of how growth, dispersal, and environmental factors affect the persistence and spread of an invasive species is of great importance in spatial ecology. Motivated by the fact that in a species, different development stages may have different vital rates and dispersal characteristics, we propose and study a reaction-diffusion juvenile-adult model, which is a natural extension of the classical Fisher’s equation. We investigate the spread rates of the population if persistent. By comparing our juvenile-adult model with the physically unstructured Fisher model, we find that Fisher equation can be approximated by our juvenile-adult model in several ways. Accordingly, the spreading speed for Fisher’s model represents a special case of that for the juvenile-adult model. We analyze how the vital rates and different dispersal abilities between juveniles and adults influence the spreading spread of the structured population, the results indicate that the juvenile-adult model provides more insights into population spread than Fisher equation. We then study a reaction-diffusion juvenile-adult model with temporally periodic coefficients. We develop a novel numerical method to calculate the spreading speed under temporal variability. Finally, we utilize the time-periodic model to understand the spatial spread of a population with separate breeding and non-breeding seasons. In particular, we scrutinize how the seasonal variation in vital rates and dispersal rates, and the duration of the breeding season affect the spreading speed of the population. The theory developed here can provide effective strategies to control the spread of invasive species.

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
Fig. 9

Similar content being viewed by others

References

  • Alqawasmeh Y, Lutscher F (2019) Persistence and spread of stage-structured populations in heterogeneous landscapes. J. Math. Biol. 78:1485–1527

    PubMed  Google Scholar 

  • Aronson DG, Weinberger HF (1975) Nonlinear diffusion in population genetics, combustion and nerve propagation, In Proceedings of the tulane program in partial differential equations and related topics Lecture notes in mathematics, vol 466. Springer, Berlin

    Google Scholar 

  • Bastviken DTE, Caraco NF, Cole JJ (1998) Experimental measurements of zebra mussel (Dreissena polymorpha) impacts on phytoplankton community composition. Fresh Water Biol. 39:375–386

    Google Scholar 

  • Bateman AW, Neubert MG, Krkosek M, Lewis MA (2015) Generational spreading speed and the dynamics of ecological invasions. Am. Nat. 186:362–375

    PubMed  Google Scholar 

  • Berger J (1986) Wild horses of the great basin: social competition and population size. University of Chicago Press, Chicago

    Google Scholar 

  • Cantrell RS, Cosner C (2003) Spatial ecology via reaction-diffusion equations. Wiley, New York

    Google Scholar 

  • Cantrell RS, Cosner C, Martínez S (2020) Persistence for a two-stage reaction-diffusion system. Mathematics 8:396

    Google Scholar 

  • Caswell H (2001) Matrix population models. Sinauer Associates Inc., Sunderland

    Google Scholar 

  • Cushing JM (1998) An introduction to structured population dynamics. SIAM, Bangkok

    Google Scholar 

  • Dale PD, Maini PK, Sherratt JA (1994) Mathematical modelling of corneal epithelial wound healing. Math. Biosci. 124:127–147

    PubMed  CAS  Google Scholar 

  • Elliot EC, Cornell SJ (2012) Dispersal polymorphism and the speed of biological invasions. PLoS ONE 7:e40496

    Google Scholar 

  • Fisher RA (1937) The wave of advantageous genes. Ann. Eugen. 7:355–369

    Google Scholar 

  • Garnier J, Lewis MA (2016) Population expansion under climate change: the genetic consequences. Bull. Math. Biol. 78:2165–2185

    PubMed  Google Scholar 

  • Guh YJ, Tamai TK, Yoshimura T (2019) The underlying mechanisms of vertebrate seasonal reproduction. Proc. Jpn. Acad. Ser. B 95:343–357

    CAS  Google Scholar 

  • Gurtin ME, MacCamy RCRC (1981) Diffusion models for age-structured populations. Math. Biosci. 54:49–59

    Google Scholar 

  • Hosono Y, Ilyas B (1995) Traveling waves for a simple diffusive epidemic model. Math. Models Methods Appl. Sci. 5:935–966

    Google Scholar 

  • Hernandez GE (1988) Dynamics of populations with age-difference and diffusion: localization. Appl. Anal. 29:143–163

    PubMed  CAS  Google Scholar 

  • Kolmogorov AN, Petrowsky IG, Piscounov NS (1937) étude de l’équation de la diffusion avec croissance de la quantité de matiére et son application á un probléme biologique, Bull. Univ. Etat Mosc., Ser. Int. A 1:1–26

    Google Scholar 

  • Kot M (2001) Elements of mathematical ecology. Cambridge University Press, Cambridge

    Google Scholar 

  • Lewis MA, Petrovskii S, Potts J (2016) Mathematics behind biological invasions. Springer-Verlag, Berlin

    Google Scholar 

  • Li BT, Weinberger HF, Lewis MA (2005) Spreading speeds as slowest wave speeds for cooperative systems. Math. Biosci. 196:82–98

    PubMed  Google Scholar 

  • Liang X, Yi Y, Zhao X-Q (2006) Spreading speeds and traveling waves for periodic evolution systems. J. Diff. Eqns. 231:57–77

    Google Scholar 

  • Liang X, Zhao X-Q (2007) Asymptotic speeds of spread and traveling waves for monotone semiflows with applications. Commun. Pure Appl. Math. 60:1–40

    Google Scholar 

  • Lutscher F, Lewis MA (2004) Spatially-explicit matrix models: A mathematical analysis of stage-structured integrodifference equations. J. Math. Biol. 48:293–324

    PubMed  Google Scholar 

  • Ma S (2001) Traveling wavefronts for delayed reaction-diffusion systems via a fixed point theorem. J. Diff. Eqns. 171:294–314

    Google Scholar 

  • Metz JAJ, Diekmann O (1986) The dynamics of physiologically structured populations Lecture notes in biomathematics, vol 68. Springer, Berlin

    Google Scholar 

  • Murray JD (2002a) Mathematical biology I: an introduction. Springer, Berlin

    Google Scholar 

  • Murray JD (2002b) Mathematical biology II: spatial models and biomedical applications. Springer, Berlin

    Google Scholar 

  • Nelson WA, Potapov A, Lewis MA, Hundsdorfer A, He F (2008) Balancing ecological complexity in predictive models: A reassessment of risk models in the mountain pine beetle. J. Appl. Ecol. 45:248–257

    Google Scholar 

  • Neubert MG, Caswell H (2000) Demography and dispersal: calculation and sensitivity analysis of invasion speed for structured populations. Ecology 81:1613–1628

    Google Scholar 

  • Okubo A, Levin S (2001) Diffusion and ecological problems. Springer, New York

    Google Scholar 

  • Pysek P, Richardson DM (2010) Invasive species, environmental change and management, and health. Annu. Rev. Environ. and Resour. 35:25–55

    Google Scholar 

  • Seo G, Lutscher F (2011) Spread rates under temporal variability: calculation and application to biological invasions. Math. Models Methods Appl. Sci. 12:2469–2489

    Google Scholar 

  • Sherratt JA, Murray JD (1991) Mathematical analysis of a basic model for epidermal wound healing. J. Math. Biol. 29:389–404

    PubMed  CAS  Google Scholar 

  • Shigesada N, Kawasaki K (1997) Biological invasions: theory and practice. Oxford University, Oxford

    Google Scholar 

  • Skellam JG (1951) Random dispersal in theoretical populations. Biometrika 38:196–218

    PubMed  CAS  Google Scholar 

  • Sun Z, Parvinen K, Heino M, Metz JAJ, de Roos AM, Dieckmann U (2020) Evolution of reproduction periods in seasonal environments, Am. Nat., 196, E-article

  • Tilman D, Kareiva P (1997) Spatial ecology. Princeton University Press, Princeton

    Google Scholar 

  • Wang H (2011) Spreading speeds and traveling waves for non-cooperative reaction-diffusion systems. J. Nonlinear Sci. 21:747–783

    CAS  Google Scholar 

  • Weinberger HF (1982) Long-time behavior of a class of biological models. SIAM J. Math. Anal. 13:353–396

    Google Scholar 

  • Weinberger HF, Lewis MA, Li B (2002) Analysis of linear determinacy for spread in cooperative models. J. Math. Biol 45:183–218

    PubMed  Google Scholar 

  • Webb GF (1985) Theory of nonlinear age-dependent population dynamics. Marcle Dekker, New York

    Google Scholar 

  • Williamson M (1999) Invasion. Ecograph 22:5–12

    Google Scholar 

  • Zhang L, Wang Z-C, Zhao X-Q (2015) Threshold dynamics of a time periodic reaction-diffusion epidemic model with latent period. J. Diff. Eqns. 258:3011–3036

    Google Scholar 

  • Zhang Y, Zhao X-Q (2013) A reaction-diffusion Lyme disease model with seasonality. SIAM J. Appl. Math. 73:2077–2099

    Google Scholar 

  • Zhao X-Q (2017) Dynamical Systems in Population Biology. Second edition, Springer-Verlag, New York

    Google Scholar 

  • Asian carp research at the Minnesota Aquatic Invasive Species Research Center, Learn the facts about Asian carp, https://www.maisrc.umn.edu/sites/maisrc.umn.edu/files/asian\_carp\_factsheet.pdf

Download references

Acknowledgements

We thank Frithjof Lutscher (University of Ottawa) for helpful discussions. We appreciate the two anonymous reviewers for their helpful comments and suggestions.

Funding

QH was partially supported by the NSF of China (11871060), the Venture and Innovation Support Program for Chongqing Overseas Returnees (7820100158), the Fundamental Research Funds for the Central Universities (XDJK2018B031), and the faculty startup fund from Southwest University (20710948). YZ was financially supported by the NSF of China (11701415), and the China Scholarship Council (201906255055).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Yuxiang Zhang.

Appendices

Appendix A

Comparison between the juvenile-adult model and the Fisher’s model

We shall show that the juvenile-adult model (2) reduces to Fisher’s equation in two special scenarios. Accordingly, the spreading speed of the juvenile-adult population is an approximation to that of single-compartment population.

The maturation rate \(\textit{\textbf{g}}\) is very large

We show that the limiting system of (2) as \(g\rightarrow \infty\) (i.e., individuals can reproduce immediately after being born) is Fisher’s equation. Summing the first and the second equation of (2), we obtain

$$\begin{aligned} \frac{\partial (J+A)}{\partial t}= & {} D_1\frac{\partial ^2 J}{\partial x^2}+D_2\frac{\partial ^2 A}{\partial x^2} \\&-m_1J+bA-m_2A-\alpha _1 J^2-\alpha _2 A^2. \end{aligned}$$
(32)

Since \(g\rightarrow \infty\), the first equation of (2) yields \(J\rightarrow 0\), accordingly, system (A1) becomes

$$\begin{aligned} \frac{\partial A}{\partial t}=D_2\frac{\partial ^2 A}{\partial x^2}+(b-m_2)A-\alpha _2 A^2. \end{aligned}$$
(33)

That is, as \(g\rightarrow \infty\), the juvenile-adult model reduces to an unstructured model because all newborns immediately become adults. Clearly, (A2) is Fisher’s equation, which yields an asymptotic spreading speed \(2\sqrt{(b-m_2)D_2}\). Note that as \(g\rightarrow \infty\), the persistence condition (4) becomes \(b>m_2\) as required.

Both stages have the same life characteristics

Due to ‘linear determinacy’ for spread in cooperative models (Weinberger et al. (2002)), which equates the spreading speed in a nonlinear system (2) with the minimum traveling wave speed of the linearized (at zero) system if no Allee effect is involved, we focus on the linearized system

$$\begin{aligned} {\left\{ \begin{array}{ll} \frac{\partial J}{\partial t}=D_1\frac{\partial ^2 J}{\partial x^2}+bA-gJ-m_1J,\\ \frac{\partial A}{\partial t}=D_2\frac{\partial ^2 A}{\partial x^2}+gJ-m_2A. \end{array}\right. } \end{aligned}$$
(34)

We now show that the linearized juvenile-adult model (A3) reduced a linearized Fisher’s equation when the juvenile and the adult have the same mortality rate and the same dispersal ability. Letting \(m_1=m_2=:m\) and \(D_1=D_2{:=}D\), we add the two equations of (A3) together to get

$$\begin{aligned} \frac{\partial (J+A)}{\partial t}=D\frac{\partial ^2 (J+A)}{\partial x^2}+bA-m(J+A). \end{aligned}$$
(35)

We introduce a density-dependent reproduction rate \(\hat{b}(J,A)=\frac{bA}{J+A}\). Then (A4) can be written as

$$\begin{aligned} \frac{\partial (J+A)}{\partial t}= & {} D\frac{\partial ^2 (J+A)}{\partial x^2} \\&+\hat{b}(J,A)(J+A)-m(J+A). \end{aligned}$$
(36)

We then approximate \(\hat{b}(J,A)\) by a constant reproduction rate, which can be done by considering a stable ratio of J to A. The spatially homogeneous system of (A3) with \(m_1=m_2=m\) and \(D_1=D_2{:=}D\) is given by

$$\begin{aligned} {\left\{ \begin{array}{ll} \frac{d J}{d t}=bA-gJ-mJ,\\ \frac{d A}{d t}=gJ-mA. \end{array}\right. } \end{aligned}$$
(37)

Straightforward calculations give

$$\begin{aligned} \frac{d}{dt}\left( \frac{J}{A}\right) =\frac{1}{A}\frac{dJ}{dt}-\frac{J}{A^2}\frac{dA}{dt} =b-g\left( \frac{J}{A}\right) -g\left( \frac{J}{A}\right) ^2. \end{aligned}$$
(38)

Setting \(\frac{d}{dt}\left( \frac{J}{A}\right) =0\), we find a stable ratio of J to A given by

$$\begin{aligned} \sqrt{\frac{1}{4}+\frac{b}{g}}-\frac{1}{2}{:=}\frac{\bar{J}}{\bar{A}}, \end{aligned}$$

which is the same as the asymptotically stable ratio we found in section 2.2 (see (22)). Thus, we approximate \(\hat{b}(J,A)\) by

$$\begin{aligned} \frac{b\bar{A}}{\bar{J}+\bar{A}}=\frac{b}{\bar{J}/\bar{A}+1} =\frac{2b}{\sqrt{1+\frac{4b}{g}}+1}{:=}\bar{b}. \end{aligned}$$

Therefore, (36) can be approximated by

$$\begin{aligned} \frac{\partial (J+A)}{\partial t}=D\frac{\partial ^2 (J+A)}{\partial x^2}+\bar{b}(J+A)-m(J+A), \end{aligned}$$
(39)

which is a linearized Fisher’s equation with a spreading speed \(2\sqrt{(\bar{b}-m)D}\). Notice that when \(m_1=m_2=m\), the persistence condition (4) becomes \(bg>(g+m)m\), which is exactly equivalent to \(\bar{b}>m\) as required.

Appendix B

A numerical method for computing \(\varvec{\rho (\mu )}\)

We divide the interval [0, T] into n equal subintervals \([t_i, t_{i+1}]\), \(i=1,2, \cdots , n\), where \(0=t_1<t_2<\cdots <t_{n+1}=T\). Denote the length of subintervals by \(\Delta t=T/n\). We then discretize system (27) by the system of difference equations:

$$\begin{aligned} {\left\{ \begin{array}{ll} \frac{\mathcal {J}(t_{i+1})-\mathcal {J}(t_i)}{\Delta t}=[D_1(t_i)\mu ^2-g(t_i)+m_1(t_i))]\mathcal {J}(t_i)+b(t_i) \mathcal {A}(t_i),\\ \frac{\mathcal {A}(t_{i+1})-\mathcal {A}(t_i)}{\Delta t}=g(t_i)\mathcal {J}(t_i)+[D_2(t_i)\mu ^2-m_2(t_i)]\mathcal {A}(t_i). \end{array}\right. } \end{aligned}$$
(40)

Introducing the matrix

$$\begin{aligned} \begin{pmatrix} \Delta t[D_1(t_i)\mu ^2-g(t_i)+m_1(t_i))]+1 &{} \Delta t b(t_i) \\ \Delta tg(t_i) &{} \Delta t[D_2(t_i)\mu ^2-m_2(t_i)]+1 \\ \end{pmatrix}, \ \ i=1,2,\cdots , n, \end{aligned}$$

we rewrite (B1) as

$$\begin{aligned} \begin{pmatrix} \mathcal {J}(t_{i+1}) \\ \mathcal {A}(t_{i+1}) \\ \end{pmatrix}=M_i \begin{pmatrix} \mathcal {J}(t_{i}) \\ \mathcal {A}(t_{i}) \\ \end{pmatrix}, i=1,2,\cdots , n. \end{aligned}$$

Thus, if \(\Delta t\) is sufficiently small, we have the following approximation:

$$\begin{aligned} \begin{pmatrix} \mathcal {J}(T) \\ \mathcal {A}(T) \\ \end{pmatrix}\approx M_n\cdots M_2 M_1 \begin{pmatrix} \mathcal {J}(0) \\ \mathcal {A}(0) \\ \end{pmatrix}. \end{aligned}$$

In terms of the definition of the Poincaré map \(Q_\mu ^T\) of (28), we can approximate \(Q_\mu ^T\) by the matrix operator \(M_n\cdots M_2 M_1{:=}\mathcal {M}_n\) by choosing a sufficiently large n. Thus, \(\rho (\mu )\), the spectral radius of \(Q_\mu ^T\), can be approximated by the spectral radius of the matrix \(\mathcal {M}_n\), for very large n.

We observe that if the functions b(t) and g(t) are positive, then when \(\Delta t\) is sufficiently small, all entries of the matrix \(M_i\) are positive, so all entries of the matrix \(\mathcal {M}_i\) are positive as well. By the famous Perron-Frobenius Theorem, we have the following results: (1) \(\mathcal {M}_n\) has a positive real eigenvalue \(\hat{\rho }(\mu )_n\), called the dominant eigenvalue, such that any other eigenvalue (possibly, complex) is strictly smaller than \(\hat{\rho }(\mu )_n\), in absolute value. Therefore, the spectral radius of matrix \(\mathcal {M}_n\) is equal to \(\hat{\rho }(\mu )_n\). (2) There exists an eigenvector \((\xi _{n,1}, \xi _{n,2})\) of \(\mathcal {M}_n\) associated with eigenvalue \(\hat{\rho }(\mu )_n\) such that both components of this eigenvector are positive. In other words, the eigenvector \((\xi _1, \xi _2)\) associated with \(\rho (\mu )\) can be approximated by \((\xi _{n,1}, \xi _{n,2})\) for very large n.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Huang, Q., Zhang, Y. Spread rates of a juvenile-adult population in constant and temporally variable environments. Theor Ecol 14, 145–160 (2021). https://doi.org/10.1007/s12080-020-00485-4

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s12080-020-00485-4

Keywords

Navigation