Skip to main content
Log in

Seasonal Invasion Dynamics in a Spatially Heterogeneous River with Fluctuating Flows

  • Original Article
  • Published:
Bulletin of Mathematical Biology Aims and scope Submit manuscript

Abstract

A key problem in environmental flow assessment is the explicit linking of the flow regime with ecological dynamics. We present a hybrid modeling approach to couple hydrodynamic and biological processes, focusing on the combined impact of spatial heterogeneity and temporal variability on population dynamics. Studying periodically alternating pool-riffle rivers that are subjected to seasonally varying flows, we obtain an invasion ratchet mechanism. We analyze the ratchet process for a caricature model and a hybrid physical–biological model. The water depth and current are derived from a hydrodynamic equation for variable stream bed water flows and these quantities feed into a reaction-diffusion-advection model that governs population dynamics of a river species. We establish the existence of spreading speeds and the invasion ratchet phenomenon, using a mixture of mathematical approximations and numerical computations. Finally, we illustrate the invasion ratchet phenomenon in a spatially two-dimensional hydraulic simulation model of a meandering river structure. Our hybrid modeling approach strengthens the ecological component of stream hydraulics and allows us to gain a mechanistic understanding as to how flow patterns affect population survival.

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.

Institutional subscriptions

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

Similar content being viewed by others

References

  • Allan, J. D., & Castillo, M. M. (2007). Stream ecology: Structure and function of running waters (2nd ed.). Dordrecht: Springer.

    Book  Google Scholar 

  • Anderson KE, Harrison LR, Nisbet RM, Kolpas A (2013) Modeling the influence of flow on invertebrate drift across spatial scales using a 2D hydraulic model and a 1D population model. Ecol Model 265:207–220

  • Anderson, K. E., Paul, A. J., McCauley, E., Jackson, L. J., Post, J. R., & Nisbet, R. M. (2006). Instream flow needs in streams and rivers: the importance of understanding ecological dynamics. Front Ecol Environ, 4, 309–318.

    Article  Google Scholar 

  • Aronson, D. G., & Weinberger, H. F. (1975). Nonlinear diffusion in population genetics, combustion, and nerve pulse propagation. In J. A. Goldstein (Ed.), Partial differential equations and related topics. Lecture notes in mathematics (Vol. 446, pp. 5–49). Berlin: Springer.

  • Bencala, K. E., & Walters, R. A. (1983). Simulation of solute transport in a mountain pool-and-riffle stream: a transient storage model. Water Resour Res, 19(3), 718–724.

    Article  Google Scholar 

  • Berestycki, H., Hamel, F., & Roques, L. (2005). Analysis of the periodically fragmented environment model: II. Biological invasions and pulsating travelling fronts. J Math Pures et Appliquees, 84, 1101–1146.

    Article  MathSciNet  MATH  Google Scholar 

  • Biggs, B. J. F. (1996). Hydraulic habitat of plants in streams. Regul Rivers, 12, 131–144.

    Article  Google Scholar 

  • Booker, D. J. (2003). Hydraulic modelling of fish habitat in urban rivers during high flows. Hydrol Process, 17, 577–599.

    Article  Google Scholar 

  • Bouckaert, F. W., & Davis, J. (1998). Microflow regimes and the distribution of macroinvertebrates around stream boulders. Freshw Biol, 40, 77–86.

    Article  Google Scholar 

  • Bunn, S. E., & Arthington, A. H. (2002). Basic principles and ecological consequences of altered flow regimes for aquatic biodiversity. Environ Manag, 30, 492–507.

    Article  Google Scholar 

  • Chaudhry, M. H. (1993). Open-channel flow. Englewood Cliffs: Prentice-Hall.

    MATH  Google Scholar 

  • Chesson, P. (2000). Mechanisms of maintenance of species diversity. Ann Rev Ecol Syst, 31, 343–366.

    Article  Google Scholar 

  • Clifford, N. J., Harmar, O. P., Harvey, G., & Petts, G. E. (2006). Physical habitat, eco-hydraulics and river design: a review and re-evaluation of some popular concepts and methods. Aquat Conserv, 16, 389–408.

    Article  Google Scholar 

  • Cooper, S. D., Diehl, S., Kratz, K., & Sarnelle, O. (1998). Implications of scale for patterns and processes in stream ecology. Aust J Ecol, 23, 27–40.

    Article  Google Scholar 

  • Dudgeon, D., Arthington, A. H., Gessner, M. O., Kawabata, Z., Knowler, D. J., Lévêque, C., et al. (2006). Freshwater biodiversity: importance, threats, status and conservation challenges. Biol Rev, 81, 163–182.

    Article  Google Scholar 

  • Emery, J. C., Gurnell, A. M., Clifford, N. J., Petts, G. E., Morrissey, I. P., & Soar, P. J. (2003). Classifying the hydraulic performance of riffle-pool bedforms for habitat assessment and river rehabilitation design. River Res Appl, 19, 533–549.

    Article  Google Scholar 

  • Fausch, K. D., Torgersen, C. E., Baxter, C. V., & Li, H. W. (2002). Landscapes to riverscapes: bridging the gap between research and conservation of stream fishes. BioScience, 52, 483–498.

    Article  Google Scholar 

  • Fischer, H. B., List, E. J., Koh, R. C. Y., Imberger, J., & Brooks, N. H. (1979). Mixing in inland and coastal waters. New York: Academic Press Inc.

    Google Scholar 

  • Fisher, R. A. (1937). The wave of advance of advantageous genes. Ann Eugen, 7, 355–369.

    Article  MATH  Google Scholar 

  • Gallien, L., Münkemüller, T., Albert, C. H., Boulangeat, I., & Thuiller, W. (2010). Predicting potential distributions of invasive species: where to go from here? Divers Distrib, 16, 331–342.

    Article  Google Scholar 

  • Goetze, D., Karlowski, U., Porembski, S., Tockner, K., & Watve, A. (2008). Spatial and temporal dimensions of biodiversity dynamics. In W. Barthlott, K. E. Linsenmair, & S. Porembski (Eds.), Biodiversity: structure and function (pp. 166–208). Oxford: Encyclopedia of Life Support System.

    Google Scholar 

  • Gregory, K. J., Gurnell, A. M., Hill, C. T., & Tooth, S. (1994). Stability of the pool-riffle sequence in changing river channels. Regul Rivers, 9, 35–43.

    Article  Google Scholar 

  • Hayes, J. W., Hughes, N. F., & Kelly, L. H. (2007). Process-based modelling of invertebrate drift transport, net energy intake and reach carrying capacity for drift-feeding salmonids. Ecol Model, 207, 171–188.

    Article  Google Scholar 

  • Hilker, F. M., & Lewis, M. A. (2010). Predator-prey systems in streams and rivers. Theor Ecol, 3, 175–193.

    Article  Google Scholar 

  • Ibanez, I., Gornish, E., Buckley, L., Debinski, D., Hellmann, J., Helmuth, B., et al. (2013). Moving forward in global-change ecology: capitalizing on natural variability. Ecol Evol, 3, 170–181.

    Article  Google Scholar 

  • Jackson, S. T., Betancourt, J. L., Booth, R. K., & Gray, S. T. (2009). Ecology and the ratchet of events: climate variability, niche dimensions, and species distributions. Proc Natl Acad Sci USA, 106, 19685–19692.

    Article  Google Scholar 

  • Janauer, G. A. (2000). Ecohydrology: fusing concepts and scales. Ecol Eng, 16, 9–16.

    Article  Google Scholar 

  • Jin, Y., & Lewis, M. A. (2012). Seasonal influences on population spread and persistence in streams: spreading speeds. J Math Biol, 65, 403–439.

    Article  MathSciNet  MATH  Google Scholar 

  • Jin, Y., & Lewis, M. A. (2011). Seasonal influences on population spread and persistence in streams: critical domain size. SIAM J Appl Math, 71, 1241–1262.

    Article  MathSciNet  MATH  Google Scholar 

  • Jungwirth, M., Muhar, S., & Schmutz, S. (1995). The effects of recreated instream and ecotone structures on the fish fauna of an epipotamal river. Hydrobiologia, 303, 195–206.

    Article  Google Scholar 

  • Labbe, T. R., & Fausch, K. D. (2000). Dynamics of intermittent stream habitat regulate persistence of a threatened fish at multiple scales. Ecol Appl, 10, 1774–1791.

    Article  Google Scholar 

  • Lancaster, J., & Downes, B. J. (2010). Linking the hydraulic world of individual organisms to ecological processes: putting ecology into ecohydraulics. River Res Appl, 26, 385–403.

    Article  Google Scholar 

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

    Article  MathSciNet  MATH  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 (Erratum: 61, 137–138 (2008))

  • Liang, X., & Zhao, X.-Q. (2010). Spreading speeds and traveling waves for abstract monostable evolution systems. J Funct Anal, 259, 857–903.

    Article  MathSciNet  MATH  Google Scholar 

  • Lloyd, F., & Sites, R. W. (2000). Microhabitat associations of three species of Dryopoidea (Coleoptera) in an Ozark stream: a comparison of substrate, and simple and complex hydraulic characters. Hydrobiologia, 439, 103–114.

    Article  Google Scholar 

  • Lou, Y., & Zhao, X.-Q. (2010). The periodic Ross-Macdonald model with diffusion and advection. Appl Anal, 89, 1067–1089.

    Article  MathSciNet  MATH  Google Scholar 

  • Lutscher, F., Pachepsky, E., & Lewis, M. A. (2005). The effect of dispersal patterns on stream populations. SIAM Rev, 47, 749–772.

    Article  MathSciNet  MATH  Google Scholar 

  • Lutscher, F., Lewis, M. A., & McCauley, E. (2006). Effects of heterogeneity on spread and persistence in rivers. Bull Math Biol, 68, 2129–2160.

    Article  MathSciNet  MATH  Google Scholar 

  • Lutscher, F., Nisbet, R. M., & Pachepsky, E. (2010). Population persistence in the face of advection. Theor Ecol, 3, 271–284.

    Article  Google Scholar 

  • Lutscher, F., & Seo, G. (2011). The effect of temporal variability on persistence conditions in rivers. J Theor Biol, 283, 53–59.

    Article  MathSciNet  Google Scholar 

  • Lytle, D. A., & Merritt, D. M. (2004). Hydrologic regimes and riparian forests: a structured population model for cottonwood. Ecology, 85, 2493–2503.

    Article  Google Scholar 

  • Milhous RT, Waddle TJ (2012) Physical Habitat Simulation (PHABSIM) Software for Windows (v.1.5.1). Fort Collins, CO: USGS Fort Collins Science Center. Institute of Freshwater, Research Drottningholm, Sweden

  • Neubert, M. G., Kot, M., & Lewis, M. A. (2000). Invasion speeds in fluctuating environments. Proc R Soc B, 267, 1603–1610.

    Article  Google Scholar 

  • Nilsson, C., Reidy, C. A., Dynesius, M., & Revenga, C. (2005). Fragmentation and flow regulation of the world’s large river systems. Science, 308, 405–408.

    Article  Google Scholar 

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

    Book  MATH  Google Scholar 

  • Orrock, J. L., Baskett, M. L., & Holt, R. D. (2010). Spatial interplay of plant competition and consumer foraging mediate plant coexistence and drive the invasion ratchet. Proc R Soc B, 277, 3307–3315.

    Article  Google Scholar 

  • Pachepsky, E., Lutscher, F., Nisbet, R., & Lewis, M. A. (2005). Persistence, spread and the drift paradox. Theor Popul Biol, 67, 61–73.

    Article  MATH  Google Scholar 

  • Pahl-Wostl, C., Arthington, A., Bogardi, J., Bunn, S. E., Hoff, H., Lebel, L., et al. (2013). Environmental flows and water governance: managing sustainable water use. Curr Opin Environ Sustain, 5, 341–351.

    Article  Google Scholar 

  • Palmer, M. A., Reidy-Liermann, C., Nilsson, C., Florke, M., Alcamo, J., Lake, P. S., et al. (2006). Climate change and world’s river basins: anticipating management options. Front Ecol Environ, 6, 81–89.

    Article  Google Scholar 

  • Pasour, V. B., & Ellner, S. P. (2010). Computational and analytic perspectives on the drift paradox. SIAM J Appl Dyn Syst, 9, 333–356.

    Article  MathSciNet  MATH  Google Scholar 

  • Poff, N. L., Allan, J. D., Hart, D. D., Richter, B. D., Arthington, A. H., Rogers, K. H., et al. (2003). River flows and water wars: emerging science for environmental decision making. Front Ecol Environ, 1, 298–306.

    Article  Google Scholar 

  • Poff, N. L., Allan, J. D., Bain, M. B., Karr, J. R., Prestegaard, K. L., Richter, B. D., et al. (1997). The natural flow regime: a paradigm for river conservation and restoration. BioScience, 47, 769–784.

    Article  Google Scholar 

  • Poff, N. L., & Zimmerman, J. K. H. (2010). Ecological responses to altered flow regimes: a literature review to inform the science and management of environmental flows. Freshw Biol, 55, 194–205.

    Article  Google Scholar 

  • Power, M. E., Brozović, N., Bode, C., & Zilberman, D. (2005). Spatially explicit tools for understanding and sustaining inland water ecosystems. Front Ecol Environ, 3, 47–55.

    Article  Google Scholar 

  • Reid, M. A., Thoms, M. C., & Dyer, F. J. (2006). Effects of spatial and temporal variation in hydraulic conditions on metabolism in cobble biofilm communities in an Australian upland stream. J N Am Bentholog Soc, 25, 756–767.

    Article  Google Scholar 

  • Richter, B. D., Mathews, R., Harrison, D. L., & Wigington, R. (2003). Ecologically sustainable water management: managing river flows for ecological integrity. Ecol Appl, 13, 206–224.

    Article  Google Scholar 

  • Schreiber, S. J. (2010). Interactive effects of temporal correlations, spatial heterogeneity and dispersal on population persistence. Proc R Soc B, 277, 1907–1914.

    Article  Google Scholar 

  • Seo, G., & Lutscher, F. (2011). Spread rates under temporal variability: calculation and application to biological invasions. Math Models Methods Appl Sci, 21, 2469–2489.

    Article  MathSciNet  MATH  Google Scholar 

  • Shenton, W., Bond, N. R., Yen, J. D. L., & Nally, R. M. (2012). Putting the “ecology” into environmental flows: ecological dynamics and demographic modelling. Environ Manag, 50, 1–10.

    Article  Google Scholar 

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

    Google Scholar 

  • Simpson, K., McCauley, E., & Nelson, W. A. (2008). Spatial heterogeneity and rates of spread in experimental streams. Oikos, 117, 1491–1499.

    Article  Google Scholar 

  • Speirs, D. C., & Gurney, W. S. C. (2001). Population persistence in rivers and estuaries. Ecology, 82, 1219–1237.

    Article  Google Scholar 

  • Stalnaker, C. B., Bovee, K. D., & Waddle, T. J. (1996). Importance of the temporal aspects of habitat hydraulics to fish population studies. Regul Rivers, 12, 145–153.

    Article  Google Scholar 

  • Stalnaker CB, Lamb BL, Henriksen J, Bovee K, Bartholow J (1995) The instream flow incremental methodology: a primer for IFIM. U.S , Geological Survey Biological, Washington, DC, Report 29

  • Statzner, B., Gore, J. A., & Randesh, V. H. (1988). Hydraulic stream ecology: observed patterns and potential applications. J N Am Benthol Soc, 7, 307–360.

    Article  Google Scholar 

  • Stelter, C., Reich, M., Grimm, V., & Wissel, C. (1997). Modelling persistence in dynamic landscapes: lessons from a metapopulation of the grasshopper Bryodema tuberculata. J Anim Ecol, 66, 508–518.

    Article  Google Scholar 

  • Suren, A. M., Smart, G. M., Smith, R. A., & Brown, S. L. R. (2000). Drag coefficients of stream bryophytes: experimental determinations and ecological significance. Freshw Biol, 45, 309–317.

    Article  Google Scholar 

  • Tharme, R. E. (2003). A global perspective on environmental flow assessment: emerging trends in the development and application of environmental flow methodologies for rivers. River Res Appl, 19, 397–441.

    Article  Google Scholar 

  • Thoms, M. C., & Parsons, M. (2003). Identifying spatial and temporal patterns in the hydrological character of the Condamine-Balonne river, Australia, using multivariate statistics. River Res Appl, 19, 443–457.

    Article  Google Scholar 

  • Vörösmarty, C. J., McIntyre, P. B., Gessner, M. O., Dudgeon, D., Prusevich, A., Green, P., et al. (2010). Global threats to human water security and river biodiversity. Nature, 467, 555–561.

    Article  Google Scholar 

  • Walks, D. J. (2007). Persistence of plankton in flowing water. Can J Fish Aquat Sci, 64, 1693–1702.

    Article  Google Scholar 

  • Wallis, C., Maddock, I., Visser, F., & Acreman, M. (2012). A framework for evaluating the spatial configuration and temporal dynamics of hydraulic patches. River Res Appl, 28, 585–593.

    Article  Google Scholar 

  • Weinberger, H. F. (2002). On spreading speeds and traveling waves for growth and migration models in a periodic habitat. J Mat Biol, 45, 511–548.

    Article  MathSciNet  MATH  Google Scholar 

  • With, K. A. (2002). The landscape ecology of invasive spread. Conserv Biol, 16, 1192–1203.

    Article  Google Scholar 

Download references

Acknowledgments

The authors thank the anonymous referees for many insightful comments and references that helped in improving the paper. The authors thank Julia Blackburn for discussions and for programming the code used in Figs. 7 and 8. We thank Ed McCauley for inspiring us to undertake much of this research. This work was supported by the MITACS NCE Network for Biological Invasions and Dispersal, the Natural Sciences and Engineering Research Council of Canada (Discovery and Accelerator grants), Alberta Sustainable Resource Development, the Alberta Water Research Institute, Killam Research Fellowship and a Canada Research Chair (MAL).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Yu Jin.

Appendices

Appendix 1: Derivation of Single Models from Benthic-Drift Models

1.1 (a) Derivation of (1)

Note that in a river or stream, many species of plankton spend most of the time on the benthos and little time in the flow (Walks 2007). We may divide the whole population into two parts: population on benthos and population in flow, and then describe the population dynamics by the benthic-drift system [(see (Pachepsky et al. 2005)]

$$\begin{aligned} \displaystyle \frac{\partial N_\mathrm{{d}}}{\partial t}&= \mu N_\mathrm{{b}}-\sigma N_\mathrm{{d}} +D \frac{\partial ^2 N_\mathrm{{d}}}{\partial x^2}-v \frac{\partial N_\mathrm{{d}}}{\partial x},\nonumber \\ \displaystyle \frac{\partial N_\mathrm{{b}}}{\partial t}&= -\mu N_\mathrm{{b}}+\sigma N_\mathrm{{d}} +f(N_\mathrm{{b}}) N_b, \end{aligned}$$
(25)

where \(N_\mathrm{{b}}\) and \(N_\mathrm{{d}}\) are population densities on the benthos and in the flow, respectively, \(D\) is the diffusion rate, \(v\) is the flow velocity, \(\mu \) is the transfer rate of individuals from the benthos to the flow, and \(\sigma \) is the transfer rate from the flow to the benthos. In a limiting case, when the transfer rates become strong (i.e., \(\sigma , \mu \rightarrow \infty \) with \(\sigma =\tau \mu \)), the second equation in (25) yields \(N_\mathrm{{b}} =\tau N_\mathrm{{d}}\), and hence, (25) can be combined into a single equation:

$$\begin{aligned} \frac{\partial N_\mathrm{{b}}}{\partial t}=\tilde{f}(N_\mathrm{{b}})N_\mathrm{{b}}+\tilde{D}\frac{\partial ^2 N_\mathrm{{b}}}{\partial x^2}-\tilde{v}\frac{\partial N_\mathrm{{b}}}{\partial x} \end{aligned}$$
(26)

with

$$\begin{aligned} \tilde{f}(N_\mathrm{{b}})=\frac{f(N_\mathrm{{b}})}{1+1/\tau }, \tilde{D}=\frac{D}{\tau +1}, \tilde{v}=\frac{v}{\tau +1}, \end{aligned}$$

where \( f\), \(D\), and \(v\) are the parameters of (25) [see Sect. 6 in (Pachepsky et al. 2005) for details]. Model (1) is of the same form as (26).

1.2 (b) Derivation of (3)

In a spatially heterogeneous habitat, the benthic-drift model becomes

$$\begin{aligned} \displaystyle \frac{\partial N_\mathrm{{d}}}{\partial t}&= \mu N_\mathrm{{b}}- \sigma N_\mathrm{{d}} + \frac{1}{A(x)}\frac{\partial }{\partial x}\left( D(x)A(x)\frac{\partial N_\mathrm{{d}}}{\partial x}\right) - \frac{Q}{A(x)}\frac{\partial N_\mathrm{{d}}}{\partial x},\nonumber \\ \displaystyle \frac{\partial N_\mathrm{{b}}}{\partial t}&= -\mu N_\mathrm{{b}}+ \sigma N_\mathrm{{d}} +f(N_\mathrm{{b}}) N_\mathrm{{b}}, \end{aligned}$$
(27)

where \(A(x)\) is the cross-sectional area at \(x\) (see (Lutscher et al. 2006) for the model details). Similarly as above, when the transfer rates \(\mu \rightarrow \infty \) and \(\sigma \rightarrow \infty \) with \(\sigma =\tau \mu \), (27) can be combined into a single model, which has the same form as (3).

1.3 (c) Derivation of (24)

In a two-dimensional habitat, the benthic-drift model becomes

$$\begin{aligned} \displaystyle \frac{\partial N_\mathrm{{d}}(x,y,t)}{\mathrm{{d}}t}&= \frac{\mu (x,y)}{h(x,y)} N_\mathrm{{b}}(x,y,t) -\sigma (x,y) N_\mathrm{{d}}(x,y,t)\nonumber \\&-\frac{1}{h(x,y)}\left[ \frac{\partial }{\partial x}\left[ v_1(x,y) h(x,y) N_\mathrm{{d}}(x,y,t)\right] \right. \nonumber \\&\left. +\frac{\partial }{\partial y}\left[ v_2(x,y) h(x,y) N_\mathrm{{d}}(x,y,t)\right] \right] \nonumber \\&+\frac{1}{h(x,y)}\left[ \frac{\partial }{\partial x}\left[ D(x,y) h(x,y)\frac{\partial N_\mathrm{{d}}(x,y,t)}{\partial x}\right] \right. \nonumber \\&\left. +\frac{\partial }{\partial y}\left[ D(x,y) h(x,y)\frac{\partial N_\mathrm{{d}}(x,y,t)}{\partial y}\right] \right] ,\nonumber \\ \displaystyle \frac{ \partial N_\mathrm{{b}}(x,y,t)}{\mathrm{{d}}t}&= f(N_\mathrm{{b}}(x,y,t))+\sigma (x,y) N_\mathrm{{d}}(x,y,t) h(x,y) -\mu (x,y) N_\mathrm{{b}}(x,y,t),\nonumber \\ \end{aligned}$$
(28)

where \(h(x,y)\) is the water depth at \((x,y)\), \(v_1\), and \( v_2 \) are the flow velocity in the \(x\) and \(y\) directions, respectively. Similarly as above, when the transfer rates \(\mu \rightarrow \infty \) and \(\sigma \rightarrow \infty \) with \(\sigma =\tau \mu /h\), where \(\tau \) is a constant, we approximately have \(n_\mathrm{{b}} =\tau n_\mathrm{{d}}\). Then, the summation of the first equation multiplied by \(\tau \) and the second equation multiplied by a constant \(\varsigma \ge 0\) yields

$$\begin{aligned} \displaystyle \frac{\partial N_\mathrm{{b}}(x,y,t)}{\mathrm{{d}}t}&= \tilde{f}(N_\mathrm{{b}}) N_\mathrm{{b}}-\frac{1}{h(x,y)}\left[ \frac{\partial }{\partial x}\left[ \tilde{v}_1(x,y) h(x,y)N_\mathrm{{b}}(x,y,t)\right] \right. \nonumber \\&+\left. \frac{\partial }{\partial y}\left[ \tilde{v}_2(x,y) h(x,y) N_\mathrm{{b}}(x,y,t)\right] \right] \nonumber \\&+\frac{1}{h(x,y)}\left[ \frac{\partial }{\partial x}\left[ \tilde{D}(x,y) h(x,y)\frac{\partial N_\mathrm{{b}}(x,y,t)}{\partial x}\right] \right. \nonumber \\&\left. +\frac{\partial }{\partial y}\left[ \tilde{D}(x,y) h(x,y)\frac{\partial N_\mathrm{{b}}(x,y,t)}{\partial y}\right] \right] , \end{aligned}$$
(29)

with

$$\begin{aligned} \tilde{f}(N_\mathrm{{b}})=\frac{\varsigma f(N_\mathrm{{b}})}{1+\varsigma }, \tilde{D}=\frac{D}{\varsigma +1}, \tilde{v}=\frac{v}{\varsigma +1}. \end{aligned}$$

Model (24) has the same form of (29) if \(f(N_\mathrm{{b}})=rN_\mathrm{{b}}(1-N_\mathrm{{b}}/K)\).

Appendix 2: The Upstream and Downstream Extents

Definition 1

Assume that a species is introduced into the river at \(t=0\). The upstream (downstream) extent (or range) \(x_t^{-}\) (\(x_t^{+}\)) at time \(t\) is defined to be the most upstream (downstream) position where individuals are observed at this time.

Remark

In simulations, \(x_t^{-}\) (\(x_t^{+}\)) is represented by the location where the population reaches a threshold detection density \(N_{\mathrm{{thresh}}}\) at time \(t\) in the upstream (downstream) direction. That is, the first \(x\) in the upstream (downstream) direction such that \(N(t,x)=N_{\mathrm{{thresh}}}\) is defined as \(x_t^{-}\) (\(x_t^{+}\)). See Fig. 9.

Fig. 9
figure 9

Upstream and downstream extents of a population with density distribution \(N(t,x)\) in a river

Definition 2

Assume that a species is introduced into the river at \(t=0\). The average speeds of spread of the population in the time interval \([0,t]\) are defined as \((x_0^--x_t^{-})/t\) and \((x_t^{+}-x_0^+)/t\) in the upstream and downstream directions, respectively.

Remark

Note that \((x_0^--x_t^{-})/t \rightarrow c^-\) and \((x_t^{+}-x_0^+)/t \rightarrow c^+\) as \(t\rightarrow \infty \), where \(c^-\) and \(c^+\) are asymptotic spreading speeds as described in Eq. (2). Therefore, we can use \((x_0^--x_t^{-})/t\) and \((x_t^{+}-x_0^+)/t\) to approximate the speeds of spread of the population in the upstream and downstream directions. In the simulations in this paper, we simply use the upstream extent \(x_t^-\) to describe the upstream invasion.

Appendix 3: The Derivation of the Water Depth Eq. (9)

In a gradually varied flow, the water depth \(y\) (unit: m) is non-uniform due to spatially varying slopes of the river bed. The governing equation for the gradually varied flow [see (5–7) in (Chaudhry 1993)] is given as:

$$\begin{aligned} \frac{\mathrm{{d}}y}{\mathrm{{d}}x} = \frac{S_0(x)-S_f(y)}{1-F_r^2(y)}. \end{aligned}$$
(30)

Here, \(S_0\) (unit: m/m) is the slope of the channel bed. It varies in space and is considered as a function of the spatial variable \(x\) (unit: m), i.e., \(S_0=S_0(x)\). Generally, the average values of \(S_0\) are between 0.0002 and 0.008 for big rivers and are slightly larger for small streams (see Table 5). \(S_f\) is the friction slope, i.e., the slope of the energy grade line, or approximation of the water surface slope. It is also spatially varying and can be determined from the Manning equation

$$\begin{aligned} S_f=\frac{n^2 v^2}{k^2 R_h^{4/3}} ~, \end{aligned}$$
(31)

where \(n\) (unit: s/m\(^{1/3}\)) is Manning’s roughness coefficient, varying with basic channel bed mechanism with values in the order of 0.025–0.050 for rivers and representing the resistance to water flows in channels, \(v\) (unit: m/s) is the water flow velocity, \(k=1\) is the conversion factor, \(R_h\) (unit: m) is the hydraulic radius, which is the ratio of wetted area \(A\) (unit: m\(^2\)) and wetted perimeter \(P\) (unit: m), i.e., \(R_h=A/P\). \(F_r\) is the Froude number that is defined as the ratio between the flow velocity and the water wave propagation velocity and is used to determine the resistance of a partially submerged object moving through water. It is a dimensionless parameter. For an arbitrarily shaped channel,

$$\begin{aligned} F_r=\sqrt{\frac{Q^2 W}{g A^3}}, \end{aligned}$$

where \(W\) (unit: m) is the top width at a particular depth, \(Q\) (unit: m\(^3\)/s) is the flow discharge, and \(g=9.8\) (unit: m/s\(^2\)) is the gravitational acceleration.

Table 5 Parameters in this paper

Consider a river (or stream) channel with rectangular cross-sections and fixed width \(B\) along one spatial dimension \(x\). Note that for natural rectangular rivers, the width \(B\) is much larger than the depth \(y\) (i.e., \(B\gg y\)), and therefore, \(R_h= By/(B+2y)\approx y\), so we simply assume \(R_h=y\). It then follows that \(Q = A v = y B v\), \(W\equiv B\),

$$\begin{aligned} S_f=\frac{n^2 v^2}{k^2 R_h^{4/3}}\approx \frac{n^2 v^2}{k^2 y^{4/3}}= \frac{n^2 Q^2}{k^2 B^2 y^{10/3}} \end{aligned}$$
(32)

and

$$\begin{aligned} F_r=\sqrt{\frac{Q^2 W}{g A^3}}=\sqrt{\frac{Q^2 B}{g A^3}}=\frac{Q}{B y\sqrt{g y}}. \end{aligned}$$
(33)

Substituting (32) and (33) into (30), we obtain a first order ODE of the channel depth \(y\):

$$\begin{aligned} \frac{\mathrm{{d}}y}{\mathrm{{d}}x} =\displaystyle \frac{g k^2 B^2 y^{10/3} S_0(x) -n^2 Q^2 g}{k^2 B^2 g y^{10/3}-Q^2 k^2 y^{1/3}} \displaystyle = \frac{g k^2 y^{10/3} S_0(x) -n^2 (Q/B)^2 g}{k^2 g y^{10/3}-(Q/B)^2 k^2 y^{1/3}}. \end{aligned}$$

Appendix 4: The Uniform Flow

Assume that the river has rectangular cross sections and a constant bed slope. If the river channel is long and channel cross sections and the bottom slope do not change with distance, then the flow accelerates or decelerates for a distance until the accelerating and resistive forces are equal (Chaudhry 1993). From that point on, the flow velocity and flow depth remain constant. Such a flow, in which the flow depth does not change with distance, is called a uniform flow, and the corresponding flow depth is called the normal depth, which is actually the critical point of Eq. (9) with a constant bed slope \(S_0\):

$$\begin{aligned} y_n=\left( \frac{Q^2 n^2}{B^2 S_{0} k^2}\right) ^{\frac{3}{10}}. \end{aligned}$$

As a subcritical flow has downstream control, for any downstream boundary condition, in such a flow, the water depth approximates to the normal depth in the upstream end far away from the downstream (see Fig. 10). Therefore, if we just consider the population spreading to the upstream end, then we may assume that the water depth stays at the normal depth, and the associated flow is a uniform flow.

Fig. 10
figure 10

The water depths in a river with a constant bed slope and with different downstream boundary conditions. The solid line represents the normal depth \(y_n=0.5978 m\). Parameters: \(g=9.8~m/s^2\), \(k=1\), \(B=50~m\), \(n=0.03~s/m^{1/3}\), \(S_{0}=0.005\), \(Q_=50~m^3/s\), channel length \(= 700 m\)

Appendix 5: The Periodic Solution to (9)

1.1 Existence, Uniqueness and Stability

Note that a periodic solution to (9) with period \(L\) corresponds to a solution to (19), and that a flow is subcritical if and only if \(F_r<1\), which implies that \(y>y_c=(Q^2/(B^2g))^{1/3}\), where \(y_c\) is called the critical depth of the river. We solve the ODE (19) in the half plane of \(y>y_c\). When the bed slope \(S_0(x)\) is given as in (18), (19) has a solution if and only if there exists some \(y_0>y_c\) such that the solutions to the two problems

$$\begin{aligned}&\left\{ \begin{array}{l} \displaystyle \frac{\mathrm{{d}}y_1}{\mathrm{{d}}x} = \frac{g k^2 y_1^{10/3} S_{0r} -n^2 (Q/B)^2 g}{k^2 g y_1^{10/3}-(Q/B)^2 k^2 y_1^{1/3}}\\ y_1(0)=y_0. \end{array}\right. \mathrm{{and}}\\&\left\{ \begin{array}{l} \displaystyle \frac{\mathrm{{d}}y_2}{\mathrm{{d}}x} = -\left( \frac{g k^2 y_2^{10/3} S_{0p} -n^2 (Q/B)^2 g}{k^2 g y_2^{10/3}-(Q/B)^2 k^2 y_2^{1/3}}\right) ,\\ y_2(0)=y_0. \end{array}\right. \end{aligned}$$

satisfy \(y_1(L_1)=y_2(L_2)\), which is true if and only if there exists some \(y_0>y_c\) such that the solutions to the two problems

$$\begin{aligned}&\left\{ \begin{array}{l} \displaystyle \frac{\mathrm{{d}}y_1}{\mathrm{{d}}x} = \frac{g k^2 y_1^{10/3} S_{0r} -n^2 (Q/B)^2 g}{k^2 g y_1^{10/3}-(Q/B)^2 k^2 y_1^{1/3}}\\ y_1(0)=y_0. \end{array} \right. \mathrm{{and}}\\&\left\{ \begin{array}{l} \displaystyle \frac{\mathrm{{d}}y_3}{\mathrm{{d}}x} = -\left( \frac{g k^2 y_3^{10/3} S_{0p} -n^2 (Q/B)^2 g}{k^2 g y_3^{10/3}-(Q/B)^2 k^2 y_3^{1/3}}\right) \cdot \frac{L_2}{L_1},\\ y_3(0)=y_0. \end{array}\right. \end{aligned}$$

satisfy \(y_1(L_1)=y_3(L_1)\).

We actually look for a solution to (19) in the half plane of \(y>y_n\), where \(y_n=\left( Q^2 n^2/(B^2 S_{0r} k^2)\right) ^{3/10}\) is the normal depth of the river where the bed slope is \(S_{0r}\), because otherwise it is easy to see from the above equivalent relations that we cannot have a solution to (19). Let \(f_1\) and \(f_2\) be the right-hand sides of these two equations, respectively. Then \(f_1\) is an increasing function of \(y\), and \(f_2\) is a decreasing function of \(y\). As \(S_{0r}>0\) and \(S_{0p}<0\), we know that \(f_1<f_2\) for small \(y>y_c\) and \(f_1>f_2\) for big \(y>y_c\). This results in that the solution \(y_1(x,y_0)<y_3(x,y_0)\) at small \(x\) and \(y_1(x,y_0)>y_3(x,y_0)\) at big \(x\) provided that \(y_0\) is not such that \(f_1>f_2\). Therefore, by the continuity of \(y_1\) and \(y_3\) with respect to \(x\), for any \(y_0>y_c\) with \(f_1(y_0)<f_2(y_0)\), there exists an \(x_{y_0}\) such that \(y_1(x_{y_0},y_0)=y_3(x_{y_0},y_0)\). By the continuity and monotonicity of \(f_1\) and \(f_2\), there exists a unique \(y_0>y_n\) such that \(x_{y_0}=L_1\), which corresponds to a unique solution to (19).

For any solution to (9) with some downstream boundary condition, we can show that the solution approaches the periodic solution at the very upstream periods by iterating the solution backward from the downstream to the upstream. As an illustration, an example is shown in Fig. 11, where the solid curve represents the periodic solution. Therefore, the periodic solution to (9) is stable for solutions to (9) with all downstream boundary conditions.

Fig. 11
figure 11

The water depths in a periodic pool-riffle river with different downstream boundary conditions. The solid line is the periodic solution with downstream boundary condition \(y(\mathrm{{end}})=0.7652\) m. Parameters: \(g=9.8~\mathrm{{m/s^2}}\), \(k=1\), \(B=50\) m, \(n=0.03~\mathrm{{s/m^{1/3}}}\), \(L_1=100\) m, \(L_2=200\) m, \(S_{0p}=-0.001\), \(S_{0r}=0.005\), \(Q_=50~\mathrm{{m^3/s}}\), channel length\(= 1,500\) m

1.2 Numerical Calculation of the Periodic Solution

Recall that subcritical flow has a downstream control, which means that to change the flow conditions in a section, flow conditions must be changed at a downstream location. Consequently, when we solve Eq. (19), we start the computation at a downstream control section and proceed in the upstream direction.

In more details, the idea for solving (19) is as follows. Guessing a boundary value \(y^{(0)}(L)\) at the downstream end \(x=L\), one calculates backward and obtains all values of \(y\) on \([0,L]\), especially \(y^{(0)}(0)\) at the upstream end \(x=0\). Then one iterates by taking a new boundary value at the downstream \(y^{(1)}(L)=y^{(0)}(0)\) and obtains \(y^{(1)}(0)\). Then let \(y^{(2)}(L)=y^{(1)}(0)\) and repeat the process till \(y^{(n)}(0)=y^{(n)}(L)\) at the \(n\)-th step. The reason for integrating backward is that the depth is asymptotic to a constant (called normal depth) proceeding upstream. The fixed point iteration scheme will converge very quickly integrating upstream, while it will diverge if integrated downstream. In each iteration, the equation is solved by the Runge–Kutta method.

Appendix 6: Spreading Speeds for the Models

1.1 Spreading Speeds for the Model in a Temporally Constant Flow

1.1.1 The Existence of Spreading Speeds for (3)

Let \(N(t,x)\) be the solution to (3) with initial value \(N_0\in C(\mathbb {R},\mathbb {R})\). Define the solution map \(\Phi _t\) of (3) as

$$\begin{aligned} \Phi _t[N_0](x)=N(t,x), \, t\ge 0. \end{aligned}$$

It follows from the standard theory of solutions to (3) that (3) generates a monotone semiflow \(\{\Phi _t\}_{t\ge 0}\) in the sense that \(\Phi _0[N_0] = N_0\) for all \(N\in C(\mathbb {R}_+,\mathbb {R})\), \(\Phi _t[\Phi _s[N_0]] = \Phi _{t+s}[N_0]\) for all \(t\ge 0\), \(s\ge 0\) and \(N_0\in C(\mathbb {R},\mathbb {R})\), and \(Q[t, N_0] := \Phi _t[N_0]\) is continuous in \((t, N_0)\) for all \(t\ge 0\) and \(N_0\in C(\mathbb {R},\mathbb {R})\). As (3) is a standard parabolic equation, \(\Phi _t\) satisfies all conditions in the abstract theory for spreading speeds for a semiflow defined in a periodic habitat in (Liang and Zhao 2010). Then by ((Liang and Zhao 2010, Theorem 5.2)), we obtain the existence of the upstream and downstream spreading speeds (\(c^-\) and \(c^+\)) [see also e.g., (Berestycki et al. 2005; Liang and Zhao 2007; Weinberger 2002)]. Note that by similar arguments to those in ((Lou and Zhao 2010, Lemma 2.10)), we have \(c^++c^->0\). The results of spreading speeds are included in the following theorem.

Theorem 1

Let \(N(t,x;\varphi )\) be the solution of (3) with \(N(0,x;\varphi )=\varphi (x)\) for all \(x\in \mathbb {R}\). The system (3) admits upstream spreading speed \(c^-\) and downstream spreading speed \(c^+\) in the following sense.

  1. (i)

    For any \(c>c^+\) and \(c^\prime >c^-\), if \(\varphi \in C_K=\{\psi \in C(\mathbb {R},\mathbb {R}),0\le \psi (x)\le K \mathrm{{for all}}x\in \mathbb {R}\}\) with \(\varphi (x)=0\) for \(x\) outside a bounded interval, then

    $$\begin{aligned} \lim \limits _{t\rightarrow \infty , x\ge c t}N(t,x;\varphi )=0, \qquad \lim \limits _{t\rightarrow \infty , x\le -c^\prime t}N(t,x;\varphi )=0. \end{aligned}$$
  2. (ii)

    For any \(c<c^+\) and \(c^\prime <c^-\), there is a positive number \(r\in \mathbb {R}\), such that if \(\varphi \in C_{K}\) and \(\varphi (x)>0\) for \(x\) on an interval of length \( r\), then

    $$\begin{aligned} \lim \limits _{t\rightarrow \infty , -c^\prime t\le x\le c t}(N(t,x;\varphi )-K)=0. \end{aligned}$$

Moreover, the spreading speeds in the upstream and downstream directions are also the minimal wave speeds for spatially periodic traveling waves, respectively. The following result follows from Theorem 5.3 in (Liang and Zhao 2010).

Theorem 2

\(\{\Phi _t\}_{t\ge 0}\) has an \(L\)-periodic rightward traveling wave \(V(x-ct,x)\) connecting \(K\) to \(0\) with \(V(\xi ,x)\) being continuous and non-increasing in \(\xi \in \mathbb {R}\) if and only if \(c\ge c_+\). \(\{\Phi _t\}_{t\ge 0}\) has an \(L\)-periodic rightward traveling wave \(V(x+ct,x)\) connecting \(0\) to \(K\) with \(V(\xi ,x)\) being continuous and increasing in \(\xi \in \mathbb {R}\) if and only if \(c\ge c_-\).

1.1.2 The Estimation of Spreading Speeds for (20)

We can follow the steps in Example 6.2 in (Weinberger 2002) to derive the spreading speeds for (20). The linearized equation for (20) at \(N=0\) is

$$\begin{aligned} \frac{\partial N(t,x)}{\partial t} = \frac{1}{y(x)} \frac{\partial }{\partial x} \left[ D(x) y(x) \frac{\partial N(t,x)}{\partial x} \right] - \frac{Q}{B y(x)} \frac{\partial N(t,x)}{\partial x} + r N(t,x).\qquad \quad \end{aligned}$$
(34)

Let \(N(t,x)=e^{\lambda t-\zeta \rho x}\psi (x)\) with \(\zeta >0\) and \(\rho =\pm 1\), and substitute it into (34). We obtain

$$\begin{aligned} \lambda e^{\lambda t-\zeta \rho x}\psi (x)&= \frac{1}{y(x)} \frac{\partial }{\partial x} \left[ D(x) y(x) \frac{\partial [e^{\lambda t-\zeta \rho x} \psi (x)]}{\partial x} \right] \\&-\frac{Q}{B y(x)} \frac{\partial [e^{\lambda t-\zeta \rho x}\psi (x)]}{\partial x} +r e^{\lambda t-\zeta \rho x}\psi (x), \end{aligned}$$

which can be simplified as

$$\begin{aligned} \lambda \psi (x)&= \displaystyle D(x)\psi ^{\prime \prime }(x) +\left[ \frac{1}{y(x)} \frac{\partial (D(x) y(x))}{\partial x} -2 \zeta \rho D(x)-\frac{Q}{B y(x)}\right] \psi ^\prime (x)\nonumber \\&\displaystyle +\left[ D(x)\zeta ^2-\frac{\zeta \rho }{y(x)}\frac{\partial (D(x) y(x))}{\partial x}+ \frac{Q \zeta \rho }{B y(x)}+r\right] \psi (x). \end{aligned}$$
(35)

Define \(L_\rho \) as

$$\begin{aligned} L_\rho&:= D(x) \frac{\partial ^2}{\partial x^2}\left[ \frac{1}{y(x)} \frac{\partial (D(x) y(x))}{\partial x}-2 \zeta \rho D(x) -\frac{Q}{B y(x)}\right] \\&+\frac{\partial }{\partial x} \left[ D(x) \zeta ^2 -\frac{\zeta \rho }{y(x)}\frac{\partial (D(x) y(x))}{\partial x}+ \frac{Q \zeta \rho }{B y(x)}+r \right] . \end{aligned}$$

It follows that \(L_\rho \) is compact and strongly positive, and hence, it admits a single principal eigenvalue with a positive periodic eigenfunction. Let \(\lambda (\zeta \rho )\) be the principle eigenvalue and \(\psi \) be the associated positive periodic eigenfunction. Then \(\lambda (\zeta \rho )\) and \(\psi \) satisfy (35) with \(\psi (0)=\psi (L)\) and \(\psi ^\prime (0)=\psi ^\prime (L)\). Define

$$\begin{aligned} c^\rho =\inf \limits _{\zeta >0}\frac{\lambda (\zeta \rho )}{\zeta }. \end{aligned}$$

When \(\rho =1\), \(c^\rho \) is the downstream spreading speed and when \(\rho =-1\), \(c^\rho \) is the upstream spreading speed. We can apply the techniques for the principal eigenvalue of Hill’s equations to (35) to obtain \(\lambda \psi (x)\).

This approximation provides a way to estimate spreading speeds for (20). However, we cannot have explicit expressions and will have to follow numerical calculations. This results in difficulties in finding how different factors affect spreading speeds.

1.2 Spreading Speeds for the Model in a Time-Varying Flow Environment (22)

To find the spreading speeds for the model (22) with varying water discharge, let \(N(t,x)=V(t,x)\cdot e^{-\eta \rho x}\) and substitute it into the linearized equation of (22) at \(N=0\):

$$\begin{aligned} \displaystyle \frac{\partial N(t,x)}{\partial t} \displaystyle&= -\frac{\partial y(t,x)}{\partial t} \cdot \frac{N(t,x)}{y(t,x)}\nonumber \\&\displaystyle +\frac{1}{y(t,x)} \frac{\partial }{\partial x} \left[ D(t,x) y(t,x) \frac{\partial N(t,x)}{\partial x} \right] \nonumber \\&- \frac{Q(t)}{B y(t,x)} \frac{\partial N(t,x)}{\partial x} + f(0) N(t,x), \end{aligned}$$
(36)

where \(\rho \) represents the direction with \(\rho =1\) in the downstream direction and \(\rho =-1\) in the upstream direction. We then have

$$\begin{aligned} \displaystyle \frac{\partial V(t,x)}{\partial t} \displaystyle&= -\frac{\partial y(t,x)}{\partial t}\cdot \frac{V(t,x)}{y(t,x)} +D(t,x)\frac{\partial ^2 V(t,x)}{\partial x^2}\nonumber \\&\displaystyle +\left[ -2\eta \rho D(t,x)+ \frac{\partial D(t,x)}{\partial x}+\frac{D(t,x)}{y(t,x)}\frac{\partial y(t,x)}{\partial x}-\frac{Q(t)}{B y(t,x)}\right] \frac{\partial V(t,x)}{\partial x}\nonumber \\&\displaystyle + \left[ \eta ^2 D(t,x)-\eta \rho \left( \frac{\partial D(t,x)}{\partial x}+\frac{D(t,x)}{y(t,x)}\frac{\partial y(t,x)}{\partial x}-\frac{Q(t)}{B y(t,x)}\right) +f(0)\right] \nonumber \\&\times V(t,x), \end{aligned}$$
(37)

Let \(L_{\eta \rho }^T\) be the Poincaré map of (37), where \(T\) is the period of the varying flow. Then the theory in (Weinberger 2002) implies that the spreading speed of \(L_{\eta \rho }^T\) is

$$\begin{aligned} c_\rho ^T= \inf \limits _{\eta >0}\frac{\ln \lambda (\eta \rho )}{\eta }, \end{aligned}$$

where \(\lambda (\eta \rho )\) is the principal eigenvalue of \(L_{\eta \rho }^T\) corresponding to a positive periodic eigenfunction, and the sign of \(\rho \) determines the direction (upstream or downstream). It then follows from the theory in (Liang et al. 2006) that the spreading speed of (22) is

$$\begin{aligned} c_\rho =\frac{1}{T} \inf \limits _{\eta >0} \frac{\ln \lambda (\eta \rho )}{\eta }. \end{aligned}$$

\(c_{1}\) is the downstream spreading speed, and \(c_{-1}\) is the upstream spreading speed.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Jin, Y., Hilker, F.M., Steffler, P.M. et al. Seasonal Invasion Dynamics in a Spatially Heterogeneous River with Fluctuating Flows. Bull Math Biol 76, 1522–1565 (2014). https://doi.org/10.1007/s11538-014-9957-3

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11538-014-9957-3

Keywords

Navigation