Skip to main content

Thank you for visiting nature.com. You are using a browser version with limited support for CSS. To obtain the best experience, we recommend you use a more up to date browser (or turn off compatibility mode in Internet Explorer). In the meantime, to ensure continued support, we are displaying the site without styles and JavaScript.

  • Article
  • Published:

Information limitation and the dynamics of coupled ecological systems

Abstract

The dynamics of large ecological systems result from vast numbers of interactions between individual organisms. Here, we develop mathematical theory to show that the rate of such interactions is inherently limited by the ability of organisms to gain information about one another. This phenomenon, which we call ‘information limitation’, is likely to be widespread in real ecological systems and can dictate both the rates of ecological interactions and long-run dynamics of interacting populations. We show how information limitation leads to sigmoid interaction rate functions that can stabilize antagonistic interactions and destabilize mutualistic ones; as a species or type becomes rare, information on its whereabouts also becomes rare, weakening coupling with consumers, pathogens and mutualists. This can facilitate persistence of consumer–resource systems, alter the course of pathogen infections within a host and enhance the rates of oceanic productivity and carbon export. Our findings may shed light on phenomena in many living systems where information drives interactions.

This is a preview of subscription content, access via your institution

Access options

Buy this article

Prices may be subject to local taxes which are calculated during checkout

Fig. 1: Model setup, sensory signals and searcher motion in empirical systems.
Fig. 2: Information, motion and interaction rates.
Fig. 3: Population persistence and collapse in consumer–resource systems.
Fig. 4: Information limitation dictates qualitative dynamics in consumer–resource systems.
Fig. 5: Information limitation can destabilize mutualistic interactions.
Fig. 6: Sensory length scale drives dynamics and productivity.

Similar content being viewed by others

Data availability

No data were generated in this study. Data in Fig. 1c were digitized from published studies19,20,5456. Chlorophyll data shown in Fig. 6b can be freely accessed via NOAA ERDDAP server (https://coastwatch.pfeg.noaa.gov/erddap).

Code availability

Simulations and ODE system solutions were conducted in the R environment (R Development Core Team). Code is available for download as Supplementary Software.

References

  1. Nowak, M. A. & Bangham, C. R. Population dynamics of immune responses to persistent viruses. Science 272, 74–79 (1996).

    CAS  PubMed  Google Scholar 

  2. Anderson, R. M. & May, R. M. Infectious Diseases of Humans: Dynamics and Control (Oxford Univ. Press, 1992).

  3. Suweis, S., Grilli, J., Banavar, J. R., Allesina, S. & Maritan, A. Effect of localization on the stability of mutualistic ecological networks. Nature Commun. 6, 10179 (2015).

    CAS  Google Scholar 

  4. Levine, J. M., Bascompte, J., Adler, P. B. & Allesina, S. Beyond pairwise mechanisms of species coexistence in complex communities. Nature 546, 56–64 (2017).

    CAS  PubMed  Google Scholar 

  5. Woodson, C. B. & Litvin, S. Y. Ocean fronts drive marine fishery production and biogeochemical cycling. Proc. Natl Acad. Sci. USA 112, 1710–1715 (2015).

    CAS  PubMed  Google Scholar 

  6. Stevens, M. Sensory Ecology, Behaviour, and Evolution (Oxford Univ. Press, 2013).

  7. Hein, A. M., Carrara, F., Brumley, D. R., Stocker, R. & Levin, S. A. Natural search algorithms as a bridge between organisms, evolution, and ecology. Proc. Natl Acad. Sci. USA 113, 9413–9420 (2016).

    CAS  PubMed  Google Scholar 

  8. Torres, L. G. A sense of scale: foraging cetaceans use of scale‐dependent multimodal sensory systems. Mar. Mammal Sci. 33, 1170–1193 (2017).

    Google Scholar 

  9. Hein, A. M., Brumley, D. R., Carrara, F., Stocker, R. & Levin, S. A. Physical limits on bacterial navigation in dynamic environments. J. R. Soc. Interface 13, 20150844 (2016).

    PubMed  PubMed Central  Google Scholar 

  10. Nevitt, G. A. et al. Evidence for olfactory search in wandering albatross, Diomedea exulans. Proc. Natl Acad. Sci. USA 105, 4576–4581 (2008).

    CAS  PubMed  Google Scholar 

  11. Hutchinson, J. M. & Waser, P. M. Use, misuse and extensions of ‘ideal gas’ models of animal encounter. Biol. Rev. 82, 335–359 (2007).

    PubMed  Google Scholar 

  12. Dusenbery, D. B. Living at Micro Scale: The Unexpected Physics of Being Small (Harvard Univ. Press, 2009).

  13. Gurarie, E. & Ovaskainen, O. Towards a general formalization of encounter rates in ecology. Theor. Ecol. 6, 189–202 (2013).

    Google Scholar 

  14. Hein, A. M. & McKinley, S. A. Sensory information and encounter rates of interacting species. PLoS Comput. Biol. 9, e1003178 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  15. Redner, S. A Guide to First-Passage Processes (Cambridge Univ. Press, 2001).

  16. Lanoiselée, Y., Moutal, N. & Grebenkov, D. S. Diffusion-limited reactions in dynamic heterogeneous media. Nature Commun. 9, 4398 (2018).

    Google Scholar 

  17. Slessor, K. N., Winston, M. L. & Le Conte, Y. Pheromone communication in the honeybee (Apis mellifera L.). J. Chem. Ecol. 31, 2731–2745 (2005).

    CAS  PubMed  Google Scholar 

  18. McDonald, M. A., Calambokidis, J., Teranishi, A. M. & Hildebrand, J. A. The acoustic calls of blue whales off California with gender data. J. Acoust. Soc. Am. 109, 1728–1735 (2001).

    CAS  PubMed  Google Scholar 

  19. van Breugel, F., Riffell, J., Fairhall, A. & Dickinson, M. H. Mosquitoes use vision to associate odor plumes with thermal targets. Curr. Biol. 25, 2123–2129 (2015).

    PubMed  PubMed Central  Google Scholar 

  20. Lämmermann, T. et al. Neutrophil swarms require LTB4 and integrins at sites of cell death in vivo. Nature 498, 371–375 (2013).

    PubMed  Google Scholar 

  21. Oaten, A. & Murdoch, W. W. Functional response and stability in predator–prey systems. Am. Nat. 109, 289–298 (1975).

    Google Scholar 

  22. Rosenzweig, M. L. Paradox of enrichment: destabilization of exploitation ecosystems in ecological time. Science 171, 385–387 (1971).

    CAS  PubMed  Google Scholar 

  23. Brose, U., Williams, R. J. & Martinez, N. D. Allometric scaling enhances stability in complex food webs. Ecol. Lett. 9, 1228–1236 (2006).

    PubMed  Google Scholar 

  24. Bascompte, J., Melián, C. J. & Sala, E. Interaction strength combinations and the overfishing of a marine food web. Proc. Natl Acad. Sci. USA 102, 5443–5447 (2005).

    CAS  PubMed  Google Scholar 

  25. Murdoch, W. W. Switching in general predators: experiments on predator specificity and stability of prey populations. Ecol. Monogr. 39, 335–354 (1969).

    Google Scholar 

  26. Real, L. A. The kinetics of functional response. Am. Nat. 111, 289–300 (1977).

    Google Scholar 

  27. Sarnelle, O. & Wilson, A. E. Type III functional response in Daphnia. Ecology 89, 1723–1732 (2008).

    PubMed  Google Scholar 

  28. Williams, R. J. & Martinez, N. D. Stabilization of chaotic and non-permanent food-web dynamics. Eur. Phys. J. B 38, 297–303 (2004).

    CAS  Google Scholar 

  29. McCauley, E., Nisbet, R. M., Murdoch, W. W., de Roos, A. M. & Gurney, W. S. Large-amplitude cycles of Daphnia and its algal prey in enriched environments. Nature 402, 653–657 (1999).

    CAS  Google Scholar 

  30. Antia, R., Koella, J. C. & Perrot, V. Models of the within-host dynamics of persistent mycobacterial infections. Proc. R. Soc. Lond. B 263, 257–263 (1996).

    CAS  Google Scholar 

  31. Reddy, R. C. & Standiford, T. J. Effects of sepsis on neutrophil chemotaxis. Curr. Opin. Hematol. 17, 18–24 (2010).

    CAS  PubMed  Google Scholar 

  32. Savoca, M. S. & Nevitt, G. A. Evidence that dimethyl sulfide facilitates a tritrophic mutualism between marine primary producers and top predators. Proc. Natl Acad. Sci. USA 111, 4157–4161 (2014).

    CAS  PubMed  Google Scholar 

  33. Cronin, T. W., Johnsen, S., Marshall, N. J. & Warrant, E. J. Visual Ecology (Princeton Univ. Press, 2014).

  34. Cahil, J. F. & McNickle, G. G. The behavioral ecology of nutrient foraging by plants. Annu. Rev. Ecol. Evol. Syst. 42, 289–311 (2011).

    Google Scholar 

  35. McKenzie, H. W., Lewis, M. A. & Merrill, E. H. First passage time analysis of animal movement and insights into the functional response. Bull. Math. Biol. 71, 107–129 (2009).

    PubMed  Google Scholar 

  36. Hassell, M. P., Lawton, J. H. & Beddington, J. R. Sigmoid functional responses by invertebrate predators and parasitoids. J. Anim. Ecol. 46, 249–262 (1977).

    Google Scholar 

  37. Bartumeus, F. et al. Foraging success under uncertainty: search tradeoffs and optimal space use. Ecol. Lett. 19, 1299–1313 (2016).

    PubMed  Google Scholar 

  38. Benhamou, S. Of scales and stationarity in animal movements. Ecol. Lett. 17, 261–272 (2014).

    PubMed  Google Scholar 

  39. Raposo, E. P. et al. How landscape heterogeneity frames optimal diffusivity in searching processes. PLoS Comput. Biol. 7, e1002233 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  40. Cooper, W. E. Jr & Blumstein, D. T. Escaping from Predators: An Integrative View (Cambridge Univ. Press, 2015).

  41. Gil, M. A. & Hein, A. M. Social interactions among grazing reef fish drive material flux in a coral reef ecosystem. Proc. Natl Acad. Sci. USA 114, 4703–4708 (2017).

    CAS  PubMed  Google Scholar 

  42. Berg, H. C. & Purcell, E. M. Physics of chemoreception. Biophys. J. 20, 193–219 (1977).

    CAS  PubMed  PubMed Central  Google Scholar 

  43. Fenton, A. & Perkins, S. E. Applying predator–prey theory to modelling immune-mediated, within-host interspecific parasite interactions. Parasitology 137, 1027–1038 (2010).

    PubMed  Google Scholar 

  44. Chtanova, T. et al. Dynamics of neutrophil migration in lymph nodes during infection. Immunity 29, 487–496 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

  45. Vandermeer, J. H. & Boucher, D. H. Varieties of mutualistic interaction in population models. J. Theor. Biol. 74, 549–558 (1978).

    CAS  PubMed  Google Scholar 

  46. Goh, B. S. Stability in models of mutualism. Am. Nat. 113, 261–275 (1979).

    Google Scholar 

  47. Holland, J. N., DeAngelis, D. L. & Bronstein, J. L. Population dynamics and mutualism: functional responses of benefits and costs. Am. Nat. 159, 231–244 (2002).

    PubMed  Google Scholar 

  48. Benoit-Bird, K. J., Waluk, C. M. & Ryan, J. P. Forage species swarm in response to coastal upwelling. Geophys. Res. Lett. 46, 1537–1546 (2019).

    Google Scholar 

  49. Kahru, M., Di Lorenzo, E., Manzano-Sarabia, M. & Mitchell, B. G. Spatial and temporal statistics of sea surface temperature and chlorophyll fronts in the California Current. J. Plankton Res. 34, 749–760 (2012).

    CAS  Google Scholar 

  50. de Roos, A. M. & Persson, L. Population and Community Ecology of Ontogenetic Development (Princeton Univ. Press, 2013).

  51. Jensen, F. H., Johnson, M., Ladegaard, M., Wisniewska, D. M. & Madsen, P. T. Narrow acoustic field of view drives frequency scaling in toothed whale biosonar. Curr. Biol. 28, 3878–3885 (2018).

    CAS  PubMed  Google Scholar 

  52. Marten, K. & Marler, P. Sound transmission and its significance for animal vocalization. Behav. Ecol. Sociobiol. 2, 271–290 (1977).

    Google Scholar 

  53. Vergassola, M., Villermaux, E. & Shraiman, B. I. ‘Infotaxis’ as a strategy for searching without gradients. Nature 445, 406–409 (2007).

    CAS  PubMed  Google Scholar 

  54. Ward, S. Chemotaxis by the nematode Caenorhabditis elegans: identification of attractants and analysis of the response by use of mutants. Proc. Natl Acad. Sci. USA 70, 817–821 (1973).

    CAS  PubMed  Google Scholar 

  55. Dekker, T. & Cardé, R. T. Moment-to-moment flight manoeuvres of the female yellow fever mosquito (Aedes aegypti L.) in response to plumes of carbon dioxide and human skin odour. J. Exp. Biol. 214, 3480–3494 (2011).

    PubMed  Google Scholar 

  56. Riffell, J. A. et al. Flower discrimination by pollinators in a dynamic chemical environment. Science 344, 1515–1518 (2014).

    CAS  PubMed  Google Scholar 

Download references

Acknowledgements

We thank members of the Information in Ecology Workshop held at UBC, S. Munch, S. McKinley, R. Nisbet, M. O’Connor, D. Brumley, F. Carrara, S. Redner and M. Gil, for comments and suggestions. A.M.H. and B.T.M. were supported by NSF IOS grant no. 1855956. A.M.H. acknowledges Simons Foundation grant no. 395890.

Author information

Authors and Affiliations

Authors

Contributions

A.M.H. and B.T.M. designed the study, performed analyses and wrote the paper.

Corresponding author

Correspondence to Andrew M. Hein.

Ethics declarations

Competing interests

The authors declare no competing interests.

Additional information

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

Extended data

Extended Data Fig. 1 Shape of interaction rate functions is preserved under continuous variation in the degree of directed motion.

(a) Rate of drift toward target as a function of distance from the target for the two-state model described in the main text with l*=100 body lengths (black line), and for three alternative models in which drift rate toward the target is a saturating function of a signal that decays exponentially (purple), like a Gaussian (orange) or like an inverse-square power law (green) with distance from the target. Signal decay parameters were chosen so that the mean drift speed at distance 100 body lengths are equal for all models. Drift rate, H(l), is calculated as a saturating function of the signal to constrain searchers to a maximum drift speed: H(l)=−vmaxtanh(S(l)), where S(l) is the signal value at a distance l, from a target (see Supplementary Discussion for details). The hyperbolic tangent form is motivated by past work on signal-dependent taxis responses15,16; other saturating functions yield similar results. (b) Interaction rate computed by solving Eq. (1) using the drift rates shown in panel (a). (c) Per-capita interaction rates corresponding to interaction rates shown in panel (b). Note non-monotonic form similar to that shown in Fig. 2c of the Main Text. (d) Range of drift functions with different parameter values. Each curve is one parameterization of the exponential signal function shown in panel (b) with drift rate again given by H(l)=−vmaxtanh(S(l)). Dark colored curves indicate functions in which change in drift rate with distance from target is relatively slow, whereas lighter colors are drift rate functions with more abrupt transition from high to low drift rate as distance from target increases. (e) Per-capita encounter rate as a function of target density for the same curves shown in panel (d). Note that location and height of peak changes but qualitative shape of curve is preserved.

Extended Data Fig. 2 Interaction rates in random target landscapes.

(a) Interaction rate and per-capita interaction rate in a two-dimensional landscape with targets distributed according to a Poisson spatial process, (b) targets distributed according to a Poisson spatial process in which the length scale of sensory information, l*, differs for each target, and (c) targets distributed according to a Poisson spatial process where the searcher depletes targets as it moves through the environment. Points correspond to l* of 5 (squares), 10 (boxes), 20 (triangles), and 40 (diamonds) body lengths. For the simulation with variable l*, these values are averages. See Supplementary Discussion ‘Robustness to spatial arrangement, target variability, and target depletion’ for a description of simulations.

Extended Data Fig. 3 Analytic approximation of interaction rates in a random target landscape.

Per-capita interaction rates in a two-dimensional landscape with targets distributed according to a Poisson spatial process. Lines show analytic approximation (see Supplementary Discussion ‘Analytic approximation for interaction rate in a random target field’). Note that, while approximation overestimates per-capita rate at intermediate density, prediction at high and low densities is accurate, as is the predicted location of peak per-capita interaction rate.

Supplementary information

Supplementary Information

Supplementary discussion of analyses and additional supporting results as well as tables.

Reporting Summary

Supplementary Software

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Hein, A.M., Martin, B.T. Information limitation and the dynamics of coupled ecological systems. Nat Ecol Evol 4, 82–90 (2020). https://doi.org/10.1038/s41559-019-1008-x

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41559-019-1008-x

This article is cited by

Search

Quick links

Nature Briefing

Sign up for the Nature Briefing newsletter — what matters in science, free to your inbox daily.

Get the most important science stories of the day, free in your inbox. Sign up for Nature Briefing