Skip to main content Accessibility help
×
Hostname: page-component-5c6d5d7d68-pkt8n Total loading time: 0 Render date: 2024-08-08T13:26:00.425Z Has data issue: false hasContentIssue false

References

Published online by Cambridge University Press:  14 January 2021

V. M. (Nitant) Kenkre
Affiliation:
University of New Mexico
Luca Giuggioli
Affiliation:
University of Bristol
Get access

Summary

Image of the first page of this content. For PDF version, please use the ‘Save PDF’ preceeding this image.'
Type
Chapter
Information
Theory of the Spread of Epidemics and Movement Ecology of Animals
An Interdisciplinary Approach using Methodologies of Physics and Mathematics
, pp. 272 - 295
Publisher: Cambridge University Press
Print publication year: 2021

Access options

Get access to the full version of this content by using one of the access options below. (Log in options will check for institutional or personal access. Content may require purchase if you do not have access.)

References

Abbott, K. D., Ksiazek, T. G., and Mills, J. N. 1999. Long-term hantavirus persistence in rodent populations in central Arizona. Emerging Infectious Diseases, 5(1), 102.Google Scholar
Abramowitz, M., and Stegun, I. A. 1965. Handbook of Mathematical Functions with Formulas, Graphs, and Mathematical Tables. New York: Dover.Google Scholar
Abramson, G. 2003. Waves of Hanta. Pages 291–299 of: Kenkre, V. M., and Lindenberg, K. (eds.), AIP conference proceedings on Modern Challenges in Statistical Mechanics: Patterns, Noise, and the Interplay of Nonlinearity and Complexity, vol. 658. Melville, NY: American Institute of Physics.Google Scholar
Abramson, G., Bishop, A. R., and Kenkre, V. M. 2001. Effects of transport memory and nonlinear damping in a generalized Fisher’s equation. Physical Review E, 64(6), 066615.Google Scholar
Abramson, G., Giuggioli, L., Kenkre, V. M., Dragoo, J. W., Parmenter, R. R., Parmenter, C. A., and Yates, T. L. 2006. Diffusion and home range parameters for rodents: Peromyscus maniculatus in New Mexico. Ecological Complexity, 3, 6470.Google Scholar
Abramson, G., Giuggioli, L., Parmenter, R. R., and Kenkre, V. M. 2013. Quasi-one-dimensional waves in rodent populations in heterogeneous habitats: a consequence of elevational gradients on spatio-temporal dynamics. Journal of Theoretical Biology, 319, 96101.Google Scholar
Abramson, G., and Kenkre, V. M. 2002. Spatio-temporal patterns in the Hantavirus infection. Physical Review E, 66(5), 011912.CrossRefGoogle Scholar
Abramson, G., Kenkre, V. M., and Bishop, A. R. 2002. Analytic solutions for nonlinear waves in coupled reacting systems. Physica A: Statistical Mechanics and Its Applications, 305(3–4), 427436.Google Scholar
Abramson, G., Kenkre, V. M., Yates, T. L., and Parmenter, R. R. 2003. Traveling waves of infection in the Hantavirus epidemics. Ecological Complexity, 65, 519534.Google Scholar
Abramson, G., and Wio, H. S. 1995. Time behaviour for diffusion in the presence of static imperfect traps. Chaos, Solitons & Fractals, 6, 15.Google Scholar
Acebrón, J. A., Bonilla, L. L., Vicente, C. J. P., Ritort, F., and Spigler, R. 2005. The Kuramoto model: a simple paradigm for synchronization phenomena. Reviews of Modern Physics, 77(1), 137.Google Scholar
Adams, E. S. 2001. Approaches to the study of territory size and shape. Annual Review of Ecology and Systematics, 32, 277303.Google Scholar
Aguirre, M. A., Abramson, G., Bishop, A. R., and Kenkre, V. M. 2002. Simulations in the mathematical modeling of the spread of the Hantavirus. Physical Review E, 66, 041908.Google Scholar
Alberts, A. C. 1992. Constraints on the design of chemical communication systems in terrestrial vertebrates. American Naturalist, 139, S62–S89.Google Scholar
Aldana, M., Dossetti, V., Huepe, C., Kenkre, V. M., and Larralde, H. 2007. Phase transitions in systems of self-propelled agents and related network models. Physical Review Letters, 98, 095702.Google Scholar
Aldana, M., and Huepe, C. 2003. Phase transitions in self-driven many-particle systems and related non-equilibrium models: a network approach. Journal of Statistical Physics, 112(1–2), 135153.Google Scholar
Allee, W. C. 1938. The Social Life of Animals. New York: W.W. Norton &Co.Google Scholar
Aoki, I. 1982. A simulation study on the schooling mechanism in fish. Bulletin of the Japanese Society of Scientific Fisheries, 48.Google Scholar
Aranson, I. S., and Tsimring, L. S. 2006. Patterns and collective behavior in granular media: theoretical concepts. Reviews of Modern Physics, 78(2), 641.Google Scholar
Aronson, D. G., and Weinberger, H. F. 1975. Nonlinear diffusion in population genetics, combustion, and nerve pulse propagation. Pages 5–49 of: Byrne, G. (ed.), Partial Differential Equations and Related Topics. Berlin: Springer.Google Scholar
Aronson, D. G., and Weinberger, H. F. 1978. Multidimensional nonlinear diffusion arising in population genetics. Advances in Mathematics, 30(1), 3376.Google Scholar
Asakura, S., and Oosawa, F. 1958. Interaction between particles suspended in solutions of macromolecules. Journal of Polymer Science, 33(126), 183192.Google Scholar
Astwood, A. M. 2012. Transport Studies of Some Microscopic and Macroscopic Systems. Ph thesis, University of New Mexico, Albuquerque, NM.Google Scholar
Astwood, A., and Kenkre, V. M. 2020. Simple Smoluchowski models for flocking. University of New Mexico preprint.Google Scholar
Attanasi, A., Cavagna, A., Del Castello, L., Giardina, I., Grigera, T. S., Jelić, A., Melillo, S., Parisi, L., Pohl, O., Shen, E., and Viale, M. 2014. Information transfer and behavioural inertia in starling flocks. Nature Physics, 10(9), 691696.Google Scholar
Baglietto, G., and Albano, E. V. 2009. Nature of the order-disorder transition in the Vicsek model for the collective motion of self-propelled particles. Physical Review E, 80(5), 050103.CrossRefGoogle ScholarPubMed
Bailey, N. T. J. 1975. The Mathematical Theory of Infectious Diseases and Its Applications. London: Charles Griffin & Co.Google Scholar
Bajec, I. L., and Heppner, F. H. 2009. Organized flight in birds. Animal Behaviour, 78(4), 777789.Google Scholar
Balescu, R. 1975. Equilibrium and Nonequilibrium Statistical Mechanics. New York: John Wiley & Sons.Google Scholar
Ball, P. 1999. The Self-Made Tapestry: Pattern Formation in Nature. Oxford, UK: Oxford University Press.Google Scholar
Ballard, M., Kenkre, V. M., and Kuperman, M. N. 2004. Periodically varying externally imposed environmental effects on population dynamics. Physical Review E, 70(3), 031912.CrossRefGoogle ScholarPubMed
Ballerini, M., Cabibbo, N., Candelier, R., Cavagna, A., Cisbani, E., Giardina, I., Lecomte, V., Orlandi, A., Parisi, G., Procaccini, A., Viale, M., and Zdravkovic, V. 2008. Interaction ruling animal collective behavior depends on topological rather than metric distance: evidence from a field study. Proceedings of the National Academy of Sciences USA, 105(4), 12321237.Google Scholar
Banavar, J. R., Damuth, J., Maritan, A., and Rinaldo, A. 2002. Supply–demand balance and metabolic scaling. Proceedings of the National Academy of Sciences USA, 99(16), 1050610509.Google Scholar
Barabási, A.-L. 2002. Linked: The New Science of Networks. Cambridge, UK: Perseus.Google Scholar
Barabási, A.-L., and Stanley, H. E. 1995. Fractal Concepts in Surface Growth. Cambridge, UK: Cambridge University Press.Google Scholar
Barkai, E., Metzler, R., and Klafter, J. 2000. From continuous time random walks to the fractional Fokker–Planck equation. Physical Review E, 61(1), 132.Google Scholar
Bartumeus, F., da Luz, M. G. E., Viswanathan, G. M., and Catalan, J. 2005. Animal search strategies: a quantitative random-walk analysis. Ecology, 86(11), 30783087.Google Scholar
Bartumeus, F., Giuggioli, L., Louzao, M., Bretagnolle, V., Oro, D., and Levin, S. A. 2010. Fishery discards impact on seabird movement patterns at regional scales. Current Biology, 20, 215222.Google Scholar
Bazazi, S., Buhl, J., Hale, J. J., Anstey, M. L., Sword, G. A., Simpson, S. J., and Couzin, I. D. 2008. Collective motion and cannibalism in locust migratory bands. Current Biology, 18(10), 735739.Google Scholar
Bechinger, C., Rudhardt, D., Leiderer, P., Roth, R., and Dietrich, S. 1999. Understanding depletion forces beyond entropy. Physical Review Letters, 83(19), 3960.Google Scholar
Be’er, A., and Ariel, G. 2019. A statistical physics view of swarming bacteria. Movement Ecology, 7(9).Google Scholar
Bel, G., and Barkai, E. 2005. Weak ergodicity breaking in the continuous-time random walk. Physical Review Letters, 94(24), 240602.Google Scholar
Ben-Jacob, E., Cohen, I., and Levine, H. 2000. Cooperative self-organization of microorganisms. Advances in Physics, 49(4), 395554.Google Scholar
Benhamou, S. 2007. How many animals really do the Lévy walk? Ecology, 88(8), 19621969.Google Scholar
Benhamou, S., and Cornelis, D. 2010. Incorporating movement behavior and barriers to improve kernel home range space use estimates. Journal of Wildlife Management, 74, 13531360.Google Scholar
Berg, H. C. 1993. Random Walks in Biology. Princeton, NJ: Princeton University Press.Google Scholar
Berg, H. C. 2000. Motile behavior of bacteria. Physics Today, 53(1), 2429.Google Scholar
Berthier, L., Jacquin, H., and Zamponi, F. 2011. Microscopic theory of the jamming transition of harmonic spheres. Physical Review E, 84(5), 051103.CrossRefGoogle ScholarPubMed
Bertin, E., Droz, M., and Grégoire, G. 2006. Boltzmann and hydrodynamic description for self-propelled particles. Physical Review E, 74(2), 022101.Google Scholar
Bertin, E., Droz, M., and Grégoire, G. 2009. Hydrodynamic equations for self-propelled particles: microscopic derivation and stability analysis. Journal of Physics A: Mathematical and Theoretical, 42(44), 445001.Google Scholar
Biagini, F., Hu, Y., Øksendal, B., and Zhang, T. 2010. Stochastic Calculus for Fractional Brownian Motion and Applications. London: Springer.Google Scholar
Bialek, W., Cavagna, A., Giardina, I., Mora, T., Pohl, O., Silvestri, E., Viale, M., and Walczak, A. M. 2014. Social interactions dominate speed control in poising natural flocks near criticality. Proceedings of the National Academy of Sciences USA, 111(20), 72127217.Google Scholar
Binder, K., and Kob, W. 2011. Glassy Materials and Disordered Solids: An Introduction to Their Statistical Mechanics. Second ed. Singapore: World Scientific.CrossRefGoogle Scholar
Bolker, B. M. 2008. Ecological models and data in R. Princeton, NJ: Princeton University Press.Google Scholar
Boltzmann, L. 1872. Weitere studien über das wärme gleichgewicht unter gasmoläkülen. Sitzungsberichte der Akademie der Wissenschaften in Wien, 66, 275370.Google Scholar
Börger, L., Dalziel, B. D., and Fryxell, J. M. 2008. Are there general mechanisms of animal home range behaviour? A review and prospects for future research. Ecology Letters, 11, 637650.Google Scholar
Botten, J., Mirowsky, K., Ye, C., Gottlieb, K., Saavedra, M., Ponce, L., and Hjelle, B. 2002. Shedding and intracage transmission of Sin Nombre hantavirus in the deer mouse (Peromyscus maniculatus) model. Journal of Virology, 76(15), 75877594.Google Scholar
Bovet, P., and Benhamou, S. 1988. Spatial analysis of animals’ movements using a correlated random walk model. Journal of Theoretical Biology, 131(4), 419433.CrossRefGoogle Scholar
Bowen, W., and McTaggart Cowan, I. 1980. Scent marking in coyotes. Canadian Journal of Zoology, 58(4), 473480.CrossRefGoogle Scholar
Bowman, F. 1953. Introduction to Elliptic Functions: With Applications. New York: Dover Publications.Google Scholar
Boyer, D., and Solis-Salas, C. 2014. Random walks with preferential relocations to places visited in the past and their application to biology. Physical Review Letters, 112, 240601.Google Scholar
Boys, C. V. 1959. Soap Bubbles, Their Colours and the Forces Which Mold Them. New York: Courier Corporation.Google Scholar
Brauer, F., and Castillo-Chavez, C. 2012. Mathematical Models in Population Biology and Epidemiology. 2nd ed. New York: Springer.Google Scholar
Brazhkin, V. V., Fomin, Yu, D., Lyapin, A. G., Ryzhov, V. N., Tsiok, E. N., and Trachenko, K. 2013. “Liquid-gas” transition in the supercritical region: fundamental changes in the particle dynamics. Physical Review Letters, 111(13), 145901 1–5.Google Scholar
Brenner, M. P., Levitov, L. S., and Budrene, E. O. 1998. Physical mechanisms for chemotactic pattern formation by bacteria. Biophysical Journal, 74(4), 16771693.Google Scholar
Brown, J. L., and Orians, G. H. 1970. Spacing patterns in mobile animals. Annual Review of Ecology and Systematics, 1, 239262.Google Scholar
Buhl, J., Sumpter, D. J. T., Couzin, I. D., Hale, J. J., Despland, E., Miller, E. R., and Simpson, S. J. 2006. From disorder to order in marching locusts. Science, 312, 14021406.Google Scholar
Burt, W. H. 1943. Territoriality and home range concepts as applied to mammals. Journal of Mammalogy, 24, 346352.Google Scholar
Cagnacci, F., Boitani, L., Powell, R. A., and Boyce, M. S. 2010. Animal ecology meets GPS-based radiotelemetry: a perfect storm of opportunities and challenges. Philosophical Transactions of the Royal Society B, 365, 21572162.Google Scholar
Calisher, C. H., Sweeney, W., Mills, J. N., and Beaty, B. J. 1999. Natural history of Sin Nombre virus in western Colorado. Emerging Infectious Diseases, 5(1), 126.Google Scholar
Camazine, S., Deneubourg, J. L., Franks, N. R., Sneyd, J., Theraulaz, G., and Bonabeau, E. 2001. Self-Organization in Biological Systems. Princeton, NJ: Princeton University Press.Google Scholar
Camelo-Neto, G., Silva, A. T. C., Giuggioli, L., and Kenkre, V. M. 2008. Effect of predators of juvenile rodents on the spread of the Hantavirus epidemic. Bulletin of Mathematical Biology, 70(1), 179188.Google Scholar
Candia, J., Parris, P. E., and Kenkre, V. M. 2007. Transport properties of random walks on scale-free/regular-lattice hybrid networks. Journal of Statistical Physics, 129(2), 323333.Google Scholar
Canosa, J. 1973. On a nonlinear diffusion equation describing population growth. IBM Journal of Research and Development, 17(4), 307313.Google Scholar
Cantrell, R. S., and Cosner, C. 2003. Spatial Ecology via Reaction-Diffusion Equations. Chichester, UK: Wiley & Sons.Google Scholar
Cao, P.-L. 1994. Computer experiments for surface diffusion: the real time in Monte Carlo simulation. Physical Review Letters, 73(19), 2595.Google Scholar
Carpenter, C. R. 1934. A field study of the behavior and social relations of howling monkeys (Alouatta palliata). Comparative Psychology Monograph, 10(2), 1168.Google Scholar
Carslaw, H. S., and Jaeger, J. C. 1959. Conduction of Heat in Solids. Oxford, UK: Clarendon Press.Google Scholar
Cavagna, A., Del Castello, L., Giardina, I., Grigera, T., Jelic, A., Melillo, S., Mora, T., Parisi, L., Silvestri, E., Viale, M., and Walczak, A. M. 2015. Flocking and turning: a new model for self-organized collective motion. Journal of Statistical Physics, 158(3), 601627.Google Scholar
Cavagna, A., Giardina, I., and Grigera, T. S. 2018. The physics of flocking: correlation as a compass from experiments to theory. Physics Reports, 728, 162.Google Scholar
Chaikin, P. M., Lubensky, T. C., and Witten, T. A. 1995. Principles of Condensed Matter Physics. Vol. 10. Cambridge, UK: Cambridge University Press.Google Scholar
Chandrasekhar, S. 1943. Stochastic problems in physics and astronomy. Reviews of Modern Physics, 15, 189.Google Scholar
Chase, M. 2016. Memory Effects in Brownian Motion, Random Walks under Confining Potentials, and Relaxation of Quantum Systems. PhD thesis, University of New Mexico, Albuquerque, NM.Google Scholar
Chase, M., Spendier, K., and Kenkre, V. M. 2016. Analysis of confined random walkers with applications to processes occurring in molecular aggregates and immunological systems. Journal of Physical Chemistry B, 120(12), 30723080.CrossRefGoogle ScholarPubMed
Chaté, H., Ginelli, F., Grégoire, G., and Raynaud, F. 2008. Collective motion of self-propelled particles interacting without cohesion. Physical Review E, 77(4), 046113.Google Scholar
Chepizhko, A. A., and Kulinskii, V. L. 2009. The kinetic regime of the Vicsek model. Pages 25–33 of: AIP Conference Proceedings on Statistical Physics: Modern Trends and Applications, vol. 1198. Melville, NY: American Institute of Physics.Google Scholar
Chepizhko, A. A., and Kulinskii, V. L. 2010. On the relation between Vicsek and Kuramoto models of spontaneous synchronization. Physica A: Statistical Mechanics and Its Applications, 389(23), 53475352.Google Scholar
Chepizhko, A. A., and Kulinskii, V. L. 2014. The hydrodynamic description for the system of self-propelled particles: ideal Viscek fluid. Physica A: Statistical Mechanics and Its Applications, 415, 493502.Google Scholar
Childs, J. E., Ksiazek, T. G., Spiropoulou, C. F., Krebs, J. W., Morzunov, S., Maupin, G. O., Gage, K. L., Rollin, P. E., Sarisky, J., Enscore, R. E., Frey, J. K., Peters, C. J., and Nichol, S. T. 1994. Serologic and genetic identification of Peromyscus maniculatus as the primary rodent reservoir for a new hantavirus in the southwestern United States. Journal of Infectious Diseases, 169(6), 12711280.CrossRefGoogle ScholarPubMed
Christiansen, P. L., and Scott, A. C. 1990. Davydov’s Soliton Revisited: Self-Trapping of Vibrational Energy in Protein. Vol. 243. New York: Plenum Press.Google Scholar
Chupeau, M., Bénichou, O., and Voituriez, R. 2015. Cover times of random searches. Nature Physics, 11(10), 844.Google Scholar
Clayton, R. K. 1980. Photosynthesis: Physical Mechanisms and Chemical Patterns. Cambridge, UK: Cambridge University Press.Google Scholar
Clerc, M. G., Escaff, D., and Kenkre, V. M. 2005. Patterns and localized structures in population dynamics. Physical Review E, 72(5), 056217.Google Scholar
Clerc, M. G., Escaff, D., and Kenkre, V. M. 2010. Analytical studies of fronts, colonies, and patterns: combination of the Allee effect and nonlocal competition interactions. Physical Review E, 82(3), 036210.Google Scholar
Codling, E. A., Planck, M. J., and Benhamou, S. 2008. Random walk models in biology. Journal of the Royal Society Interface, 95(5), 813834.Google Scholar
Cooke, S. J., Hinch, S. G., Wikelski, M., Andrews, R. D., Kuchel, L. J., Wolcott, T. G., and Butler, P. J. 2004. Biotelemetry: a mechanistic approach to ecology. Trends in Ecology and Evolution, 19, 334343.Google Scholar
Courchamp, F., Berec, L., and Gascoigne, J. 2008. Allee Effects in Ecology and Conservation. New York: Oxford University Press.Google Scholar
Courchamp, F., Clutton-Brock, T., and Grenfell, B. 1999. Inverse density dependence and the Allee effect. Trends in Ecology and Evolution, 14(10), 405410.Google Scholar
Couzin, I. D., and Franks, N. R. 2003. Self-organized lane formation and optimized traffic flow in army ants. Proceedings of the Royal Society of London. Series B: Biological Sciences, 270(1511), 139146.Google Scholar
Couzin, I. D., Krause, J., Franks, N. R., and Levin, S. A. 2005. Effective leadership and decision-making in animal groups on the move. Journal of Theoretical Biology, 218, 111.Google Scholar
Couzin, I. D., Krause, J., James, R., Ruxton, G. D., and Franks, N. R. 2002. Collective memory and spatial sorting in animal groups. Nature, 433, 513516.Google Scholar
Cross, M. C., and Hohenberg, P. C. 1993. Pattern formation outside of equilibrium. Reviews of Modern Physics, 65(3), 851.Google Scholar
Czirók, A., Ben-Jacob, E., Cohen, I., and Vicsek, T. 1996. Formation of complex bacterial colonies via self-generated vortices. Physical Review E, 54(2), 1791.CrossRefGoogle ScholarPubMed
Dahmen, K. A., Nelson, D. R., and Shnerb, N. M. 2000. Life and death near a windy oasis. Journal of Mathematical Biology, 41(1), 123.Google Scholar
Davies, N. B., and Houston, A. I. 1984. Territory economics. Pages 148–169 of: Krebs, J. R., and Davies, N. B. (eds.), Behavioural Ecology: An Evolutionary Approach, 2nd ed. Oxford, UK: Blackwell Science.Google Scholar
Dawnkaski, E. J., Srivastava, D., and Garrison, B. J. 1995. Time dependent Monte Carlo simulations of H reactions on the diamond {001}(2×1) surface under chemical vapor deposition conditions. Journal of Chemical Physics, 102(23), 94019411.Google Scholar
DeAngelis, D. L., and Gross, L. J. 2018. Individual-Based Models and Approaches in Ecology: Populations, Communities and Ecosystems. Boca Raton, FL: CRC Press.Google Scholar
DeWit, R. 1971. Relation between dislocations and disclinations. Journal of Applied Physics, 42(9), 33043308.Google Scholar
Doncaster, C. P., and Woodroffe, R. 1993. Den site can determine shape and size of badger territories: implications for group-living. Oikos, 66, 8893.CrossRefGoogle Scholar
Dossetti, V., and Sevilla, F. J. 2015. Emergence of collective motion in a model of interacting brownian particles. Physical Review Letters, 115(5), 058301.Google Scholar
Downs, J. A., Horner, M. W., and Tucker, A. D. 2011. Time-geographic density estimation for home range analysis. Annals GIS, 17(3), 163171.Google Scholar
Dresden, M. 1961. Recent developments in the quantum theory of transport and galvano-magnetic phenomena. Reviews of Modern Physics, 33(2), 265.Google Scholar
Durrett, R., and Levin, S. A. 1994. The importance of being discrete (and spatial). Theoretical Population Biology, 46, 363394.Google Scholar
Ebeling, W., and Schimansky-Geier, L. 2008. Swarm dynamics, attractors and bifurcations of active Brownian motion. European Physical Journal Special Topics, 157(1), 1731.CrossRefGoogle Scholar
Ebert, U., and van Saarloos, W. 2000. Front propagation into unstable states: universal algebraic convergence towards uniformly translating pulled fronts. Physica D: Nonlinear Phenomena, 146(1–4), 199.Google Scholar
Edwards, A. M. 2011. Overturning conclusions of Lévy flight movement patterns by fishing boats and foraging animals. Ecology, 92(6), 12471257.Google Scholar
Ehrenfest, P., and Ehrenfest, T. 1959. The Conceptual Foundations of the Statistical Approach in Mechanics. Ithaca, NY: Cornell University Press.Google Scholar
Engel, M., Anderson, J. A., Glotzer, S. C., Isobe, M., Bernard, E. P., and Krauth, W. 2013. Hard-disk equation of state: first-order liquid-hexatic transition in two dimensions with three simulation methods. Physical Review E, 87, 042134.Google Scholar
Erdélyi, A., Magnus, W., Oberhettinger, F., and Tricomi, F. G. 1953. Higher Transcendental Functions. New York: McGraw-Hill.Google Scholar
Erdmann, U., Ebeling, W., and Mikhailov, A. S. 2005. Noise-induced transition from translational to rotational motion of swarms. Physical Review E, 71(5), 051904.Google Scholar
Ernest, S. K. M., Brown, J. H., and Parmenter, R. R. 2000. Rodents, plants, and precipitation: spatial and temporal dynamics of consumers and resources. Oikos, 88(3), 470482.Google Scholar
Escudero, C., Buceta, J., de La Rubia, F. J., and Lindenberg, K. 2004. Extinction in population dynamics. Physical Review E, 69(2), 021908.Google Scholar
Falcón-Cortés, A., Boyer, D., Giuggioli, L., and Majumdar, S. N. 2017. Localization transition induced by learning in random searches. Physical Review Letters, 119(14), 140603.Google Scholar
Farrell, F. D. C., Marchetti, M. C., Marenduzzo, D., and Tailleur, J. 2012. Pattern formation in self-propelled particles with density-dependent motility. Physical Review Letters, 108(24), 248101.Google Scholar
Fedotov, S. 2001. Front propagation into an unstable state of reaction-transport systems. Physical Review Letters, 86(5), 926.Google Scholar
Fermi, L. 2014. Atoms in the Family: My Life with Enrico Fermi. Chicago: University of Chicago Press.Google Scholar
Feynman, R. P., Leighton, R. B., and Sands, M. 2011. The Feynman Lectures on Physics, Vol. II: The New Millennium Edition: Mainly Electromagnetism and Matter. New York: Basic Books.Google Scholar
Fisher, R. A. 1937. The wave of advance of advantageous genes. Annals of Eugenics, 7(4), 355369.Google Scholar
Förster, T. 1948. Intermolecular energy migration and fluorescence. Annalen der Physik (Leipzig), 2, 5575.Google Scholar
Fuentes, M. A., Kuperman, M. N., and Kenkre, V. M. 2003. Nonlocal interaction effects on pattern formation in population dynamics. Physical Review Letters, 91(15), 158104.Google Scholar
Fuentes, M. A., Kuperman, M. N., and Kenkre, V. M. 2004. Analytical considerations in the study of spatial patterns arising from nonlocal interaction effects. Journal of Physical Chemistry B, 108(29), 1050510508.Google Scholar
Gautrais, J., Ginelli, F., Fournier, R., Blanco, S., Soria, M., Chaté, H., and Theraulaz, G. 2012. Deciphering interactions in moving animal groups. PLoS Computational Biology, 8(9), e1002678.Google Scholar
Gibbs, J. W. 1902. Elementary Principles in Statistical Mechanics: Developed with Especial Reference to the Rational Foundation of Thermodynamics. New York: Scribner & Sons.Google Scholar
Giuggioli, L. 2020. Exact spatio-temporal dynamics of confined lattice random walks in arbitrary dimensions: a century after Smoluchowski and Pólya. Physical Review X, 10(2), 021045.Google Scholar
Giuggioli, L., Abramson, G., Kenkre, V. M., Parmenter, R. R., and Yates, T. L. 2006. Theory of home range estimation from displacement measurements of animal populations. Journal of Theoretical Biology, 240, 126135.Google Scholar
Giuggioli, L., Abramson, G., Kenkre, V. M., Suzán, G., Marcé, E., and Yates, T. L. 2005. Diffusion and home range parameters from rodent population measurements in Panama. Bulletin of Mathematical Biology, 67(5), 11351149.Google Scholar
Giuggioli, L., Arye, I., Heiblum Robles, A., and Kaminka, G. A. 2018. From ants to birds: a novel bio-inspired approach to online area coverage. Pages 31–43 of: Distributed Autonomous Robotic Systems. Berlin: Springer.Google Scholar
Giuggioli, L., and Bartumeus, F. 2010. Animal movement, search strategies and behavioural ecology: a cross-disciplinary way forward. Journal of Animal Ecology, 79, 906909.Google Scholar
Giuggioli, L., and Bartumeus, F. 2012. Linking animal movement to site fidelity. Journal of Mathematical Biology, 64(4), 647656.Google Scholar
Giuggioli, L., Kalay, Z., and Kenkre, V. M. 2008. Study of transients in the propagation of nonlinear waves in some reaction diffusion systems. The European Physical Journal B, 62(3), 341.Google Scholar
Giuggioli, L., and Kenkre, V. M. 2003. Analytic solutions of a nonlinear convective equation in population dynamics. Physica D: Nonlinear Phenomena, 183(3–4), 245259.CrossRefGoogle Scholar
Giuggioli, L., and Kenkre, V. M. 2014. Consequences of animal interactions on their dynamics: emergence of home ranges and territoriality. Movement Ecology, 2(20).Google Scholar
Giuggioli, L., McKetterick, T. J., and Holderied, M. 2015. Delayed response and biosonar perception explain movement coordination in trawling bats. PLoS Computational Biology, 11(3), e1004089.Google Scholar
Giuggioli, L., Potts, J. R., and Harris, S. 2011a. Animal interactions and the emergence of territoriality. PLoS Computational Biology, 7(3), e1002008.Google Scholar
Giuggioli, L., Potts, J. R., and Harris, S. 2011b. Brownian walkers within subdiffusing territorial boundaries. Physical Review E, 83, 061138.Google Scholar
Giuggioli, L., Potts, J. R., and Harris, S. 2012. Predicting oscillatory dynamics in the movement of territorial animals. Journal of the Royal Society Interface, 9(72), 15291543.Google Scholar
Giuggioli, L., Potts, J. R., Rubenstein, D. I., and Levin, S. A. 2013. Stigmergy, collective actions, and animal social spacing. Proceedings of the National Academy of Sciences USA, 110(42), 1690416909.Google Scholar
Giuggioli, L., Sevilla, F. J., and Kenkre, V. M. 2009. A generalised master equation approach to modelling anomalous transport in animal movement. Journal of Physics A: Mathematical and Theoretical, 42, 434004.Google Scholar
Giuggioli, L., Viswanathan, G. M., Kenkre, V. M., Parmenter, R. R., and Yates, T. L. 2007. Effects of finite probing windows on the interpretation of the multifractal properties of random walks. Europhysics Letters, 77, 4004.Google Scholar
Glass, G. E., Cheek, J. E., Patz, J. A., Shields, T. M., Doyle, T. J., Thoroughman, D. A., Hunt, D. K., Enscore, R. E., Gage, K. L., Irland, C., Peters, C. J., and Bryan, R. 2000. Using remotely sensed data to identify areas at risk for hantavirus pulmonary syndrome. Emerging Infectious Diseases, 6(3), 238.Google Scholar
Glass, G. E., Shields, T. M., Parmenter, R. R., Goade, D., Mills, J. N., Cheek, J., Cook, J., and Yates, T. L. 2006. Predicted Hantavirus risk in 2006 for the southwestern United States. Occasional Papers, Museum of Texas Tech University, 255, 116.Google Scholar
Glass, G. E., Yates, T. L., Fine, J. B., Shields, T. M., Kendall, J. B., Hope, A. G., Parmenter, C. A., Peters, C. J., Ksiazek, T. G., Li, C.-S., Patz, J. A., and Mills, J. N. 2002. Satellite imagery characterizes local animal reservoir populations of Sin Nombre virus in the southwestern United States. Proceedings of the National Academy of Sciences USA, 99(26), 1681716822.Google Scholar
Glauber, R. J. 1963. Time-dependent statistics of the Ising model. Journal of Mathematical Physics, 4(2), 294307.Google Scholar
Goodenough, A. E., Little, N., Carpenter, W. S., and Hart, A. G. 2017. Birds of a feather flock together: insights into starling murmuration behaviour revealed using citizen science. PLoS ONE, 12(6).Google Scholar
Gordon, D. M. 1997. The population consequences of territorial behavior. Trends in Ecology and Evolution, 12, 6366.Google Scholar
Gorenflo, R., Luchko, Y., and Mainardi, F. 1999. Analytical properties and applications of the Wright function. Fractional Calculus and Applied Analysis, 2, 383414.Google Scholar
Gorenflo, R., Mainardi, F., Moretti, D., Pagnini, G., and Paradisi, P. 2002. Discrete random walk models for space–time fractional diffusion. Chemical Physics, 284(1–2), 521541.Google Scholar
Gosling, L. M., and Roberts, S. C. 2001. Scent-marking by male mammals: cheat-proof signals to competitors and mates. Advances in the Study of Behavior, 30, 169217.Google Scholar
Grassé, P.-P. 1959. La reconstruction du nid et les coordinations interindividuelles chez Bellicositermes natalensis et Cubitermes sp. La théorie de la stigmergie: essai d’nterprétation du comportement des termites constructeurs. Insectes Sociaux, 6(1), 4180.Google Scholar
Grégoire, G., and Chaté, H. 2004. Onset of collective and cohesive motion. Physical Review Letters, 92(2), 025702.Google Scholar
Grigolini, P. 2006. The continuous-time random walk versus the generalized master equation. Pages 357–474 of: Fractals, Diffusion, and Relaxation in Disordered Complex Systems: Advances in Chemical Physics, Part A, vol. 133. Hoboken, NJ: John Wiley & Sons.Google Scholar
Grimm, V., and Railsback, S. F. 2005. Individual-Based Modeling and Ecology. Princeton, NJ: Princeton University Press.Google Scholar
Grossman, D., Aranson, I. S., and Jacob, E. B. 2008. Emergence of agent swarm migration and vortex formation through inelastic collisions. New Journal of Physics, 10(2), 023036.Google Scholar
Grossmann, R., Schimansky-Geier, L., and Romanczuk, P. 2012. Active Brownian particles with velocity-alignment and active fluctuations. New Journal of Physics, 14(7), 073033.Google Scholar
Grünbaum, D., and Okubo, A. 1994. Modelling social animal aggregations. Pages 296–325 of: Frontiers in Mathematical Biology, vol. 100 Lecture Notes in Biomathematics. Berlin: Springer.Google Scholar
Gueron, S., and Levin, S. A. 1993. Self-organization of front patterns in large wildebeest herds. Journal of Theoretical Biology, 165(4), 541552.Google Scholar
Gueron, S., Levin, S. A., and Rubenstein, D. I. 1996. The dynamics of herds: from individuals to aggregations. Journal of Theoretical Biology, 182(1), 8598.Google Scholar
Haggett, P. 2000. The Geographical Structure of Epidemics. Oxford, UK: Oxford University Press.Google Scholar
Haining, R. P. 2003. Spatial Data Analysis: Theory and Practice. Cambridge, UK: Cambridge University Press.Google Scholar
Halperin, B. I., and Nelson, D. R. 1978. Theory of two-dimensional melting. Physical Review Letters, 41(2), 121124.Google Scholar
Han, Y., Ha, N. Y., Alsayed, A. M., and Yodh, A. G. 2008. Melting of two-dimensional tunable-diameter colloidal crystals. Physical Review E, 77(4), 041406.Google Scholar
Hanski, I. 1999. Metapopulation Ecology. Oxford, UK: Oxford University Press.Google Scholar
Harris, W. F. 1977. Disclinations. Scientific American, 237(6), 130145.Google Scholar
Harrison, R. L. 1992. Toward a theory of inter-refuge corridor design. Conservation Biology, 6(2), 293295.Google Scholar
Heiblum Robles, A., and Giuggioli, L. 2018. Phase transitions in stigmergic territorial systems. Physical Review E, 98(4), 042115.Google Scholar
Helbing, D., Farkas, I., and Vicsek, T. 2000. Simulating dynamical features of escape panic. Nature, 407(6803), 487490.Google Scholar
Hemenger, R. P., Lakatos-Lindenberg, K., and Pearlstein, R. M. 1974 . Impurity quenching of molecular excitons. III. Partially coherent excitons in linear chains. Journal of Chemical Physics, 60(8), 32713277.Google Scholar
Herbert-Read, J. E., Perna, A., Mann, R. P., Schaerf, T. M., Sumpter, D. J. T., and Ward, A. J. W. 2011. Inferring the rules of interaction of shoaling fish. Proceedings of the National Academy of Sciences USA, 108, 18726–18731.Google Scholar
Hilborn, R., and Mangel, M. 1997. The Ecological Detective: Confronting Models with Data. Vol. 28. Princeton, NJ: Princeton University Press.Google Scholar
Hilfer, R., and Anton, L. 1995. Fractional master equations and fractal time random walks. Physical Review E, 51, R848–R851.Google Scholar
Hinde, R. A. 1956. The biological significance of the territory of birds. Ibis, 98, 340369.Google Scholar
Holland, O., and Melhuish, C. 1999. Stigmergy, self-organisation and sorting in collective robotics. Artificial Life, 5(2), 173202.Google Scholar
Holstein, T. 1959a. Studies of polaron motion: Part I. The molecular-crystal model. Annals of Physics, 8(3), 325342.Google Scholar
Holstein, T. 1959b. Studies of polaron motion: Part II. The “small” polaron. Annals of Physics, 8(3), 343389.Google Scholar
Hughes, B. D. 1995. Random Walks and Random Environments: Random Walks. Vol. 1. Oxford, UK: Clarendon Press.Google Scholar
Hughes, B. D., Shlesinger, M. F., and Montroll, E. W. 1981. Random walks with self-similar clusters. Proceedings of the National Academy of Sciences USA, 78(6), 32873291.Google Scholar
Hurst, H. E., Black, R. P., and Simaika, Y. M. 1965. Long-Term Storage: An Experimental Study. London: Constable.Google Scholar
Huth, A., and Wissel, C. 1992. The simulation of the movement of fish schools. Journal of Theoretical Biology, 156(3), 365385.Google Scholar
Hyman, J., Hughes, M., Searcy, W. A., and Nowicki, S. 2004. Individual variation in the strength of territory defense in male song sparrows: correlates of age, territory tenure, and neighbor aggressiveness. Behaviour, 141, 1527.Google Scholar
Ihle, T. 2011. Kinetic theory of flocking: derivation of hydrodynamic equations. Physical Review E, 83(3), 030901.Google Scholar
Izús, G., Deza, R., Borzi, C., and Wio, H. S. 1997. Global analysis of pattern selection and bifurcations in monostable reaction-diffusion systems. Physica A: Statistical Mechanics and Its Applications, 237(1–2), 135149.Google Scholar
Jankowitz, W. J., Van Rooyen, M. W., Shaw, D., Kaumba, J. S., and Van Rooyen, N. 2008. Mysterious circles in the Namib Desert. South African Journal of Botany, 74(2), 332334.Google Scholar
Jiang, L., Giuggioli, L., Perna, A., Escobedo, R., Lecheval, V., Sire, C., Han, Z., and Theraulaz, G. 2017. Identifying influential neighbors in animal flocking. PLoS Computational Biology, 13(11).Google Scholar
Kac, M. 1959. Probability and Related Topics in Physical Sciences. Vol. 1. Providence, RI: American Mathematical Society.Google Scholar
Kac, M., Uhlenbeck, G. E., and Hemmer, P. C. 1963. On the van der Waals theory of the vapor-liquid equilibrium. I. Discussion of a one-dimensional model. Journal of Mathematical Physics, 4(2), 216228.Google Scholar
Kalcounis-Rüppell, M. C., and Ribble, D. O. 2007. A phylogenetic analysis of the breeding systems of Neotomine–Peromyscine rodents. Pages 68–85 of: Wolff, J., and Sherman, P. W. (eds.), Rodent Societies: An Ecological and Evolutionary Perspective. Chicago: University of Chicago Press.Google Scholar
Kantelhardt, J. W., Zschiegner, S. A., Koscielny-Bunde, E., Havlin, S., Bunde, A., and Stanley, H. E. 2002. Multifractal detrended fluctuation analysis of nonstationary time series. Physica A: Statistical Mechanics and Its Applications, 316(1–4), 87114.Google Scholar
Kapfer, S. C., and Krauth, W. 2015. Two-dimensional melting: from liquid-hexatic coexistence to continuous transitions. Physical Review Letters, 114, 035702.Google Scholar
Kassner, K. 1996. Pattern Formation in Diffusion-Limited Crystal Growth. Singapore: World Scientific.CrossRefGoogle Scholar
Katz, Y., Tunstrøm, K., Ioannou, C. C., Huepe, C., and Couzin, I. D. 2011. Inferring the structure and dynamics of interactions in schooling fish. Proceedings of the National Academy of Sciences USA, 108(46), 1872018725.Google Scholar
Kays, R., Crofoot, M. C., Jetz, W., and Wikelski, M. 2015. Terrestrial animal tracking as an eye on life and planet. Science, 348(6240), aaa2478.Google Scholar
Keeling, M. J., and Rohani, P. 2011. Modeling Infectious Diseases in Humans and Animals. Princeton, NJ: Princeton University Press.Google Scholar
Keller, E. F., and Segel, L. A. 1971. Traveling bands of chemotactic bacteria: a theoretical analysis. Journal of Theoretical Biology, 30(2), 235248.Google Scholar
Kenkre, V. M. 1974. Coupled wave-like and diffusive motion of excitons. Physics Letters A, 47, 119120.Google Scholar
Kenkre, V. M. 1977a. The generalized master equation and its applications. Pages 441–461 of: Landman, U. (ed.), Statistical Mechanics and Statistical Methods in Theory and Application. New York: Plenum.Google Scholar
Kenkre, V. M. 1977b. Master-equation theory of the effect of vibrational relaxation on intermolecular transfer of electronic excitation. Physical Review A, 16, 766776.Google Scholar
Kenkre, V. M. 1980. Theory of exciton annihilation in molecular crystals. Physical Review B, 22, 20892098.Google Scholar
Kenkre, V. M. 1981. Validity of the bilinear rate equation for exciton annihilation and expressions for the annihilation constant. Zeitschrift für Physik B Condensed Matter, 43(3), 221227.Google Scholar
Kenkre, V. M. 1982a. The master equation approach: coherence, energy transfer, annihilation, and relaxation. Pages 1–109 of: Hoehler, G. (ed.), Exciton Dynamics in Molecular Crystals and Aggregates. Berlin: Springer.Google Scholar
Kenkre, V. M. 1982b. A theoretical approach to exciton trapping in systems with arbitrary trap concentration. Chemical Physics Letters, 93(3), 260263.Google Scholar
Kenkre, V. M. 1989. The quantum nonlinear dimer and extensions. In: Pnevmatikos, S., Bountis, T., and Pnevmatikos, S. (eds.), Singular Behaviour and Nonlinear Dynamics. London: World Scientific.Google Scholar
Kenkre, V. M. 1994. Proceedings of the Granular Material Symposium. Pages BB 6.5.1–8 of: Sen, S., and Hunt, M. L. (eds.), Polarons and Applications, vol. 627. Warrendale, PA: MRS.Google Scholar
Kenkre, V. M. 1998. Four stages in the study of electron-phonon interactions. Journal of Luminescence, 76 /77, 511517.Google Scholar
Kenkre, V. M. 2001a. Spatial memories and correlation functions in the theory of stress distribution in granular materials. Granular Matter, 3(1–2), 2328.Google Scholar
Kenkre, V. M. 2001b. Theory of stress distribution in granular materials: the memory formalism. Pages BB6.5.1-8 of: Surajit, S., and Hunt, M. L. (eds.), Materials Research Society conference proceedings on the granular state, vol. 627. Cambridge, UK: Cambridge University Press.Google Scholar
Kenkre, V. M. 2003. Memory formalism, nonlinear techniques, and kinetic equation approaches. Pages 63–103 of: Kenkre, V. M., and Lindenberg, K. (eds.), AIP conference proceedings on Modern Challenges in Statistical Mechanics: Patterns, Noise, and the Interplay of Nonlinearity and Complexity, vol. 658. Melville, NY: American Institute of Physics.Google Scholar
Kenkre, V. M. 2004. Results from variants of the Fisher equation in the study of epidemics and bacteria. Physica A: Statistical Mechanics and Its Applications, 342, 242248.Google Scholar
Kenkre, V. M. 2005. Statistical mechanical considerations in the theory of the spread of the Hantavirus. Physica A: Statistical Mechanics and Its Applications, 356, 121126.Google Scholar
Kenkre, V. M., and Campbell, D. K. 1986. Self-trapping on a dimer: time-dependent solutions of a discrete nonlinear Schrödinger equation. Physical Review B, 34(7), 4959.Google Scholar
Kenkre, V. M., Endicott, M. R., Glass, S. J., and Hurd, A. J. 1996. A theoretical model for compaction of granular materials. Journal of the American Ceramic Society, 79(12), 30453054.Google Scholar
Kenkre, V. M., Giuggioli, L., Abramson, G., and Camelo-Neto, G. 2007. Theory of Hantavirus infection spread incorporating localized adult and itinerant juvenile mice. European Physics Journal B, 55, 461470.Google Scholar
Kenkre, V. M., and Knox, R. S. 1974a. Generalized-master-equation theory of excitation transfer. Physical Review B, 9, 52795290.Google Scholar
Kenkre, V. M., and Knox, R. S. 1974b. Theory of fast and slow excitation transfer rates. Physical Review Letters, 33(14), 803.Google Scholar
Kenkre, V. M., and Kumar, N. 2008. Nonlinearity in bacterial population dynamics: proposal for experiments for the observation of abrupt transitions in patches. Proceedings of the National Academy of Sciences USA, 105(48), 1875218757.CrossRefGoogle ScholarPubMed
Kenkre, V. M., and Kuperman, M. N. 2003. Applicability of the Fisher equation to bacterial population dynamics. Physical Review E, 67(5), 051921.Google Scholar
Kenkre, V. M., and Lindenberg, K. 2003. Modern Challenges in Statistical Mechanics: Patterns, Noise, and the Interplay of Nonlinearity and Complexity. Vol. 658, Conference Proceedings. Melville, NY: American Institute of Physics.Google Scholar
Kenkre, V. M., Montroll, E. W., and Shlesinger, M. F. 1973. Generalized master equations for continuous-time random walks. Journal of Statistical Physics, 9, 4550.Google Scholar
Kenkre, V. M., Parmenter, R. R., Peixoto, I. D., and Sadasiv, L. 2005. A theoretical framework for the analysis of the West Nile virus epidemic. Mathematical and Computer Modelling, 42(3–4), 313324.Google Scholar
Kenkre, V. M., and Parris, P. E. 1983. Exciton trapping and sensitized luminescence: a generalized theory for all trap concentrations. Physical Review B, 27(6), 3221.Google Scholar
Kenkre, V. M., and Reineker, P. 1982. Exciton Dynamics in Molecular Crystals and Aggregates. Vol. 94, Springer Tracts in Modern Physics. Berlin: Springer.Google Scholar
Kenkre, V. M., Scott, J. E., Pease, E. A., and Hurd, A. J. 1998. Nonlocal approach to the analysis of the stress distribution in granular systems. I. Theoretical framework. Physical Review E, 57(5), 58415849.Google Scholar
Kenkre, V. M., and Sevilla, F. J. 2007. Thoughts about anomalous diffusion: time-dependent coefficients versus memory functions. Pages 147–160 of: Ali, T. S., and Sinha, K. B. (eds.), Contributions to Mathematical Physics: A Tribute to Gerard G. Emch. New Delhi: Hindustani Book Agency.Google Scholar
Kenkre, V. M., and Sugaya, S. 2014. Theory of the transmission of infection in the spread of epidemics: interacting random walkers with and without confinement. Bulletin of Mathematical Biology, 76(12), 30163027.Google Scholar
Kenkre, V. M., and Tsironis, G. P. 1987. Nonlinear effects in quasielastic neutron scattering: exact line-shape calculation for a dimer. Physical Review B, 35(4), 1473.Google Scholar
Kenkre, V. M., and Van Horn, H. M. 1981. Annihilations of stationary particles on a lattice. Physical Review A, 23(6), 3200.Google Scholar
Kenkre, V. M., and Wong, Y. M. 1981. Effect of transport coherence on trapping: quantum-yield calculations for excitons in molecular crystals. Physical Review B, 23(8), 3748.Google Scholar
Kermack, W. O., and McKendrick, A. G. 1927. A contribution to the mathematical theory of epidemics. Proceedings of the Royal Society of London A, 115(772), 700721.Google Scholar
Kessler, D. A., and Levine, H. 1993. Pattern formation in Dictyostelium via the dynamics of cooperative biological entities. Physical Review E, 48(6), 4801.Google Scholar
Kierstead, H., and Slobodkin, L. B. 1953. The size of water masses containing plankton blooms. Journal of Marine Research, 12(1), 141147.Google Scholar
Kinezaki, N., Kawasaki, K., Takasu, F., and Shigesada, N. 2003. Modeling biological invasions into periodically fragmented environments. Theoretical Population Biology, 64(3), 291302.Google Scholar
Kleman, M., and Laverntovich, O. D. 2007. Soft Matter Physics: An Introduction. New York: Springer.Google Scholar
Kokko, H., and Sutherland, W. J. 2001. Ecological traps in changing environments: ecological and evolutionary consequences of a behaviourally mediated Allee effect. Evolutionary Ecology Research, 3(5), 603610.Google Scholar
Kolmogorov, A. N., Petrovsky, I., and Piskunov, N. 1937 . Étude de l’équation de la diffusion avec croissance de la quantité de matière et son application à un problème biologique. Bulletin of the University of Moskow, Ser. Internat., Sec. A, 1, 125.Google Scholar
Kosterlitz, D. J., and Thouless, J. M. 1973. Ordering, metastability and phase transitions in two-dimensional systems. Journal of Physics C: Solid State Physics, 1181(7), 11811203.Google Scholar
Kot, M. 2001. Elements of Mathematical Ecology. Cambridge, UK: Cambridge University Press.Google Scholar
Kozel, R. M., and Fleharty, E. D. 1979. Movements of rodents across roads. Southwestern Naturalist, 24(2), 239248.Google Scholar
Krebs, J. R., and Davies, N. B. 1987. An Introduction to Behavioral Ecology. Oxford, UK: Blackwell Scientific Publications.Google Scholar
Kuenzi, A. J., Morrison, M. L., Swann, D. E., Hardy, P. C., and Downard, G. T. 1999. A longitudinal study of Sin Nombre virus prevalence in rodents, southeastern Arizona. Emerging Infectious Diseases, 5(1), 113.Google Scholar
Kumar, N., and Kenkre, V. M. 2011. Effects of gradual spatial variation in resources on population extinction: analytic calculations for abrupt transitions. Physica A: Statistical Mechanics and Its Applications, 390(2), 257262.CrossRefGoogle Scholar
Kumar, N., Kuperman, M. N., and Kenkre, V. M. 2009b. Theory of possible effects of the Allee phenomenon on the population of an epidemic reservoir. Physical Review E, 79(4), 041902.Google Scholar
Kumar, N., Parmenter, R. R., and Kenkre, V. M. 2010. Extinction of refugia of hantavirus infection in a spatially heterogeneous environment. Physical Review E, 82(1), 011920.Google Scholar
Kumar, N., Viswanathan, G. M., and Kenkre, V. M. 2009a. Hurst exponents for interacting random walkers obeying nonlinear Fokker–Planck equations. Physica A: Statistical Mechanics and Its Applications, 388(18), 36873694.Google Scholar
Kuramoto, Y. 1984. Chemical Oscillations, Waves, and Turbulence. Berlin: Springer.Google Scholar
Lakatos-Lindenberg, K., Hemenger, R. P., and Pearlstein, R. M. 1972. Solutions of master equations and related random walks on quenched linear chains. Journal of Chemical Physics, 56(10), 48524867.Google Scholar
Landau, L. D. 1933. Electron motion in crystal lattices. Physikalische Zeitschrift der Sowjetunion, 3, 664.Google Scholar
Landman, U., Montroll, E. W., and Shlesinger, M. F. 1977. Random walks and generalized master equations with internal degrees of freedom. Proceedings of the National Academy of Sciences USA, 74(2), 430433.Google Scholar
Lee, C. T., Hoopes, M. F., Diehl, J., Gilliland, W., Huxel, G., Leaver, E. V., McCann, K., Umbanhowar, J., and Mogilner, A. 2001. Non-local concepts and models in biology. Journal of Theoretical Biology, 210(2), 201219.Google Scholar
Lewis, M. A., and Murray, J. D. 1993. Modelling territoriality and wolf-deer interactions. Ecology, 366, 738740.Google Scholar
Likos, C. N. 2001. Effective interactions in soft condensed matter physics. Physics Reports, 348(4–5), 267439.Google Scholar
Lin, A. L. 2003. Resonant Chemical Oscillations: Pattern Formation in Reaction-Diffusion Systems. Pages 315–328 of: Kenkre, V. M., and Lindenberg, K. (eds.), AIP conference proceedings on Modern Challenges in Statistical Mechanics: Patterns, Noise, and the Interplay of Nonlinearity and Complexity, vol. 658. Melville, NY: American Institute of Physics.Google Scholar
Lin, A. L., Mann, B. A., Torres-Oviedo, G., Lincoln, B., Käs, J., and Swinney, H. L. 2004. Localization and extinction of bacterial populations under inhomogeneous growth conditions. Biophysical Journal, 87(1), 7580.Google Scholar
Lindner, M., Nir, G., Vivante, A., Young, I. T., and Garini, Y. 2013. Dynamic analysis of a diffusing particle in a trapping potential. Physical Review E, 87(2), 022716.Google Scholar
Livesay, D. R., Jambeck, P., Rojnuckarin, A., and Subramaniam, S. 2003. Conservation of electrostatic properties within enzyme families and superfamilies. Biochemistry, 42(12), 34643473.Google Scholar
Long, J. A., and Nelson, T. A. 2012. Time geography and wildlife home range delineation. Journal of Wildlife Management, 76(2), 407413.Google Scholar
Lopez, U., Gautrais, J., Couzin, I. D., and Theraulaz, G. 2012. From behavioural analyses to models of collective motion in fish schools. Interface Focus, 2(6), 693707.Google Scholar
Lotka, A. J. 1926. Elements of physical biology. Science Progress in the Twentieth Century (1919–1933), 21(82), 341343.Google Scholar
Ludwig, D., Aronson, D. G., and Weinberger, H. F. 1979. Spatial patterning of the spruce budworm. Journal of Mathematical Biology, 8(3), 217258.Google Scholar
Lukeman, R., Li, Y.-X., and Edelstein-Keshet, L. 2010. Inferring individual rules from collective behavior. Proceedings of the National Academy of Sciences USA, 107(28), 1257612580.Google Scholar
Lyons, A. J., Turner, W. C., and Getz, W. M. 2013. Home range plus: a space-time characterization of movement over real landscapes. Movement Ecology, 1(2).Google Scholar
MacArthur, R. H, and Wilson, E. O. 1967. The Theory of Island Biogeography. Princeton, NJ: Princeton University Press.Google Scholar
Madras, N., and Slade, G. 2013. The Self-Avoiding Walk. Berlin: Springer.Google Scholar
Maher, C. A., and Lott, D. F. 1995. Definitions of territoriality used in the study of variation in vertebrate spacing systems. Animal Behaviour, 49, 15811597.Google Scholar
Mainardi, F., Luchko, Y., and Pagnini, G. 2001. The fundamental solution of the space-time fractional diffusion equation. Fractional Calculus and Applied Analysis, 4, 153192.Google Scholar
Mandelbrot, B. B., and Van Ness, J. W. 1968. Fractional Brownian motions, fractional noises and applications. SIAM Review, 10(4), 422437.Google Scholar
Mann, B. 2001. Spatial phase-transitions in bacterial growth. PhD thesis, University of Texas, Austin, TX.Google Scholar
Manne, K. K., Hurd, A. J., and Kenkre, V. M. 2000. Nonlinear waves in reaction-diffusion systems: The effect of transport memory. Physical Review E, 61(4), 4177.Google Scholar
Marchetti, M. C., Joanny, J.-F., Ramaswamy, S., Liverpool, T. B., Prost, J., Rao, M., and Simha, R. A. 2013. Hydrodynamics of soft active matter. Reviews of Modern Physics, 85(3), 1143.Google Scholar
Marfin, A. A., Petersen, L. R., Eidson, M., Miller, J., Hadler, J., Farello, C., Werner, B., Campbell, G. L., Layton, M., Smith, P., Bresnitz, E., Cartter, M., Scaletta, J., Obiri, G., Bunning, M., Craven, R. C., Roehrig, J. T., Julian, K. G., Hinten, S. R., Gubler, D. J., and the ArboNET Cooperative Surveillance Group. 2001. Widespread West Nile virus activity, eastern United States, 2000. Emerging Infectious Diseases, 7(4), 730.Google Scholar
Maude, G. 2010. The Spatial Ecology and Foraging Behaviour of the Brown Hyaena (Hyaena brunnea). PhD thesis, University of Bristol, Bristol, UK.Google Scholar
May, R. M. 2001. Stability and Complexity in Model Ecosystems. 2nd ed. Princeton, NJ: Princeton University Press.Google Scholar
McInnis, D. 2007. Applications of Nonlinear Science and Kinetic Equations to the Theory of the Spread of Epidemics. PhD thesis, University of New Mexico, Albuquerque, NM.Google Scholar
McKane, A. J., and Newman, T. J. 2004. Stochastic models in population biology and their deterministic analogs. Physical Review E, 70, 041902.Google Scholar
McKean, H. P. Jr. 1970. Nagumo’s equation. Advances in Mathematics, 4(3), 209223.Google Scholar
Méndez, V., Fedotov, S., and Horsthemke, W. 2010. Reaction-Transport Systems: Mesoscopic Foundations, Fronts, and Spatial Instabilities. Berlin: Springer.Google Scholar
Méndez, V., Campos, D., and Bartumeus, F. 2016. Stochastic Foundations in Movement Ecology. Berlin: Springer.Google Scholar
Méndez, V., and Campos, D. 2008. Population extinction and survival in a hostile environment. Physical Review E, 77(2), 022901.Google Scholar
Mermin, N. D. 1968. Crystalline order in two dimensions. Physical Review, 176(1), 250.Google Scholar
Mermin, N. D., and Wagner, H. 1966. Absence of ferromagnetism or antiferromagnetism in one- or two-dimensional isotropic Heisenberg models. Physical Review Letters, 17(22), 1133.Google Scholar
Metzler, R., Jeon, J.-H., Cherstvy, A. G., and Barkai, E. 2014. Anomalous diffusion models and their properties: non-stationarity, non-ergodicity, and ageing at the centenary of single particle tracking. Physical Chemistry Chemical Physics, 16(44), 2412824164.Google Scholar
Metzler, R., and Klafter, J. 2000. The random walks guide to anomalous diffusion: a fractional dynamics approach. Physics Reports, 339, 177.Google Scholar
Mikhailov, A. S., and Zanette, D. H. 1999. Noise-induced breakdown of coherent collective motion in swarms. Physical Review E, 60, 45714575.Google Scholar
Mills, J. N., Ksiazek, T. G., Peters, C. J., and Childs, J. E. 1999a. Long-term studies of hantavirus reservoir populations in the southwestern United States: a synthesis. Emerging Infectious Diseases, 5(1), 135.Google Scholar
Mills, J. N., Yates, T. L., Ksiazek, T. G., Peters, C. J., and Childs, J. E. 1999b. Long-term studies of hantavirus reservoir populations in the southwestern United States: rationale, potential, and methods. Emerging Infectious Diseases, 5(1), 95.Google Scholar
Mogilner, A., and Edelstein-Keshet, L. 1995. Selecting a common direction I. How orientational order can arise from simple contact responses between interacting cells. Journal of Mathematical Biology, 33(6), 619660.Google Scholar
Mogilner, A., and Edelstein-Keshet, L. 1996. Spatio-angular order in populations of self-aligning objects: formation of oriented patches. Physica D: Nonlinear Phenomena, 89(3–4), 346367.Google Scholar
Mogilner, A., and Edelstein-Keshet, L. 1999. A non-local model for a swarm. Journal of Mathematical Biology, 38(6), 534570.Google Scholar
Mogilner, A., Edelstein-Keshet, L., and Ermentrout, G. B. 1996. Selecting a common direction. Journal of Mathematical Biology, 34(8), 811842.Google Scholar
Mollison, D. 1977. Spatial contact models for ecological and epidemic spread. Journal of the Royal Statistical Society: Series B, 39(3), 283313.Google Scholar
Mollison, D. 1991. Dependence of epidemic and population velocities on basic parameters. Mathematical Biosciences, 107, 255287.Google Scholar
Montroll, E. W. 1964. Random walks on lattices. Pages 193–220 of: Bellman, R. (ed.), Stochastic Processes in Mathematical Physics and Engineering. Proceedings of Symposia in Applied Mathematics, vol. 16. Providence, RI: American Mathematical Society.Google Scholar
Montroll, E. W. 1967. On nonlinear processes involving population growth and diffusion. Journal of Applied Probability, 4(2), 281290.Google Scholar
Montroll, E. W. 1969. Random walks on lattices containing traps. Page 6 of: Physical Society of Japan Journal Supplement, Proceedings of the International Conference on Statistical Mechanics, vol. 26.Google Scholar
Montroll, E. W., and Potts, R. B. 1955. Effect of Defects on Lattice Vibrations. Physical Review, 100(2), 525543.Google Scholar
Montroll, E. W., and Weiss, G. H. 1965. Random walks on Lattices II. Journal of Mathematical Physics, 6(2), 167181.Google Scholar
Montroll, E. W., and West, B. J. 1979. On an enriched collection of stochastic processes. Pages 61–175 of: Montroll, E. W., and Lebowitz, J. J. (eds.), Studies in Statistical Mechanics: Vol. VII. Fluctuation Phenomena. Amsterdam: North Holland Publishing.Google Scholar
Moorcroft, P. R., and Lewis, M. A. 2006. Mechanistic Home Range Analysis. Princeton, NJ: Princeton University Press.Google Scholar
Mora, T., and Bialek, W. 2011. Are biological systems poised at criticality? Journal of Statistical Physics, 144(2), 268302.Google Scholar
Moussaïd, M., Helbing, D., and Theraulaz, G. 2011. How simple rules determine pedestrian behavior and crowd disasters. Proceedings of the National Academy of Sciences USA, 108(17), 68846888.Google Scholar
Murray, J. D. 2003a. Mathematical Biology: I. An Introduction. 3rd ed. Berlin: Springer.Google Scholar
Murray, J. D. 2003b. Mathematical Biology: II. Spatial Models and Biomedical Applications. 3rd ed. Berlin: Springer.Google Scholar
Nagy, M., Akos, Z., Biro, D., and Vicsek, T. 2010. Hierarchical group dynamics in pigeon flocks. Nature, 464, 890893.Google Scholar
Nagy, M., Daruka, I., and Vicsek, T. 2007. New aspects of the continuous phase transition in the scalar noise model (SNM) of collective motion. Physica A: Statistical Mechanics and Its Applications, 373, 445454.Google Scholar
Namboodiri, V. M. K., Levy, J. M., Mihalas, S., Sims, D. W., and Shuler, M. G. H. 2016. Rationalizing spatial exploration patterns of wild animals and humans through a temporal discounting framework. Proceedings of the National Academy of Sciences USA, 113(31), 87478752.Google Scholar
Nasci, R. S., Savage, H. M., White, D. J., Miller, J. R., Cropp, B. C., Godsey, M. S., Kerst, A. J., Bennett, P., Gottfried, K., and Lanciotti, R. S. 2001. West Nile virus in overwintering Culex mosquitoes, New York City, 2000. Emerging Infectious Diseases, 7(4), 742.Google Scholar
Nathan, R. M., Getz, W. M., Revilla, E., Holyoak, M., Kadmon, R., Saltz, D., and Smouse, P. E. 2008. A movement ecology paradigm for unifying organismal movement research. Proceedings of the National Academy of Sciences USA, 105, 19052–19059.Google Scholar
Nayfeh, A. H. 2008. Perturbation Methods. New York: John Wiley & Sons.Google Scholar
Nelson, D. R., and Shnerb, N. M. 1998. Non-Hermitian localization and population biology. Physical Review E, 58(2), 1383.Google Scholar
Newell, A. C. 1997. The dynamics and analysis of patterns. Pages 201–268 of: Nijhout, H. F., Nadel, L., and Stein, D. (eds.), Pattern Formation in the Physical and Biological Sciences, vol. 5. Boca Raton, FL: CRC Press.Google Scholar
Newman, K., and Borchers, D. L. 2014. Modelling Population Dynamics: Model Formulation, Fitting and Assessment Using State-Space Methods. Methods in Statistical Ecology. New York: Springer.Google Scholar
Newton, I. 1992. Experiments on the limitation of bird numbers by territorial behaviour. Biological Reviews, 67, 129173.Google Scholar
Nice, M. M. 1941. The role of territory in bird life. American Naturalist, 26(3), 441487.Google Scholar
Nichol, S. T., Spiropoulou, C. F., Morzunov, S., Rollin, P. E., Ksiazek, T. G., Feldmann, H., Sanchez, A., Childs, J., Zaki, S., and Peters, C. J. 1993. Genetic identification of a hantavirus associated with an outbreak of acute respiratory illness. Science, 262(5135), 914917.Google Scholar
Nicolis, G. 1995. Introduction to Nonlinear Science. Cambridge, UK: Cambridge University Press.Google Scholar
Nigg, B. M., MacIntosh, B. R., and Mester, J. 2000. Biomechanics and Biology Of Movement. Champaign, IL: Human Kinetics Publishers.Google Scholar
Nimzowitsch, A. 1930. My System: A Treatise on Chess. Vol. 5. New York: Three Rivers Press.Google Scholar
Niwa, H.-S. 1994. Self-organizing dynamic model of fish schooling. Journal of Theoretical Biology, 171(2), 123136.Google Scholar
Okubo, A. 1980. Diffusion and Ecological Problems: Mathematical Models. Vol. 10 Biomathematics. Berlin: Springer.Google Scholar
Okubo, A., Grünbaum, D., and Edelstein-Keshet, L. 2001. The dynamics of animal grouping. Pages 197–237 of: Diffusion and Ecological Problems: Modern Perspectives. New York: Springer.Google Scholar
Okubo, A., and Levin, S. A. 2001. Diffusion and Ecological Problems: Modern Perspectives. 2nd ed. New York: Springer.Google Scholar
Olfati-Saber, R. 2006. Flocking for multi-agent dynamic systems: algorithms and theory. IEEE Transactions on Automatic Control, 51(3), 401420.Google Scholar
Ovaskainen, O. 2004. Habitat-specific movement parameters estimated using mark– recapture data and a diffusion model. Ecology, 85(1), 242257.Google Scholar
Park, S., Wolanin, P. M., Yuzbashyan, E. A., Lin, H., Darnton, N. C., Stock, J. B., Silberzan, P., and Austin, R. 2003. Influence of topology on bacterial social interaction. Proceedings of the National Academy of Sciences USA, 100(24), 1391013915.Google Scholar
Parmenter, C. A., Yates, T. L., Parmenter, R. R., and Dunnum, J. L. 1999. Long-term field studies of small mammals for hantavirus prevalence: statistical sensitivity for detection of spatial and temporal patterns in rodent population densities. Emerging Infectious Diseases, 5, 118–25.Google Scholar
Parmenter, R. R., and MacMahon, J. A. 1983. Factors determining the abundance and distribution of rodents in a shrub-steppe ecosystem: the role of shrubs. Oecologia, 59, 145156.Google Scholar
Parris, P. E., Candia, J., and Kenkre, V. M. 2008. Random-walk access times on partially disordered complex networks: an effective medium theory. Physical Review E, 77(6), 061113.Google Scholar
Parris, P. E, and Kenkre, V. M. 2005. Traversal times for random walks on small-world networks. Physical Review E, 72(5), 056119.Google Scholar
Parrish, J. K., and Hamner, W. M. 1997. Animal Groups in Three Dimensions: How Species Aggregate. Cambridge, UK: Cambridge University Press.Google Scholar
Pastor, J. 2011. Mathematical Ecology of Populations and Ecosystems. Chichester, UK: John Wiley & Sons.Google Scholar
Pearce, D. J. G., Miller, A. M., Rowlands, G., and Turner, M. S. 2014. Role of projection in the control of bird flocks. Proceedings of the National Academy of Sciences USA, 111(29), 1042210426.Google Scholar
Peixoto, I. D., Giuggioli, L., and Kenkre, V. M. 2005. Arbitrary nonlinearities in convective population dynamics with small diffusion. Physical Review E, 72(4), 041902.Google Scholar
Pekar, S. I. 1954. Untersuchung über die Elektronentheorie der Kristalle, Berlin, Akad. Berlin: Akademie Verlag.Google Scholar
Perry, N. 2005. Experimental validation of a critical domain size in reaction–diffusion systems with Escherichia coli populations. Journal of the Royal Society Interface, 2(4), 379387.Google Scholar
Petrovskii, S. V., and Li, B.-L. 2005. Exactly Solvable Models of Biological Invasion. Boca Raton, FL: Chapman and Hall/CRC.Google Scholar
Pimentel, J. A., Aldana, M., Huepe, C., and Larralde, H. 2008. Intrinsic and extrinsic noise effects on phase transitions of network models with applications to swarming systems. Physical Review E, 77(6), 061138.Google Scholar
Podlubny, I. 1999. Fractional Differential Equations: An Introduction to Fractional Derivatives, Fractional Differential Equations, to Methods of Their Solution and Some of Their Applications. San Diego, CA: Academic Press.Google Scholar
Pope, M., and Swenberg, C. E. 1999. Electronic Processes in Organic Crystals and Polymers. 2nd ed. New York: Oxford University Press.Google Scholar
Porfiri, M., Stilwell, D. J., Bollt, E. M., and Skufca, J. D. 2006. Random talk: Random walk and synchronizability in a moving neighborhood network. Physica D: Nonlinear Phenomena, 224(1–2), 102113.Google Scholar
Poston, T., and Stewart, I. 1978. Catastrophe Theory and Its Applications. London: Pitman Publishing.Google Scholar
Potts, J. R., Harris, S., and Giuggioli, L. 2012. Territorial dynamics and stable home range formation for central place foragers. PLoS ONE, 7(3), e34033.Google Scholar
Potts, J. R., Harris, S., and Giuggioli, L. 2013. Quantifying behavioral changes in territorial animals caused by sudden population declines. American Naturalist, 182(3).Google Scholar
Powell, R. A., and Mitchell, M. S. 2012. What is a home range? Journal of Mammalogy, 93(4).Google Scholar
Press, W. H., Flannery, B. P., Teukolsky, S. A., and Vetterling, W. T. 1989. Numerical Recipes. Vol. 3. Cambridge, UK: Cambridge University Press.Google Scholar
Puri, S. 1991. Singular-perturbation analysis of the Fisher equation. Physical Review A, 43(12), 7031.Google Scholar
Puri, S., and Wiese, K. J. 2003. Perturbative linearization of reaction–diffusion equations. Journal of Physics A: Mathematical and General, 36(8), 2043.Google Scholar
Pyke, G. H. 1979. The economics of territory size and time budget in the golden-winged sunbird. American Naturalist, 114, 131145.Google Scholar
Pyke, G. H. 2015. Understanding movements of organisms: it’s time to abandon the Lévy foraging hypothesis. Methods in Ecology and Evolution, 6(1), 116.Google Scholar
Qi, W., Gantapara, A. P., and Dijkstra, M. 2014. Two-stage melting induced by dislocations and grain boundaries in monolayers of hard spheres. Soft Matter, 10(30), 54495457.Google Scholar
Quinn, G. P., Keough, M.J., et al. 2002. Experimental Design and Data Analysis for Biologists. New York: Cambridge University Press.Google Scholar
Quinn, R. A., and Goree, J. 2001. Experimental test of two-dimensional melting through disclination unbinding. Physical Review E, 64(5), 051404.Google Scholar
Rabinovich, M. I., Ezersky, A. B., and Weidman, P. D. 2000. The Dynamics of Patterns. Singapore: World Scientific.Google Scholar
Raghib, M., Levin, S. A., and Kevrekidis, I. G. 2010. Multiscale analysis of collective motion and decision-making in swarms: an advection–diffusion equation with memory approach. Journal of Theoretical Biology, 264(3), 893913.Google Scholar
Ramaswamy, S. 2010. The mechanics and statistics of active matter. Annual Review of Condensed Matter Physics, 1(1), 323345.Google Scholar
Rappel, W.-J., Nicol, A., Sarkissian, A., Levine, H., and Loomis, W. F. 1999. Self-organized vortex state in two-dimensional Dictyostelium dynamics. Physical Review Letters, 83(6), 1247.Google Scholar
Redner, S. 2001. A Guide to First-Passage Processes. Cambridge, UK: Cambridge University Press.Google Scholar
Redner, S., and Ben-Avraham, D. 1990. Nearest-neighbour distances of diffusing particles from a single trap. Journal of Physics A: Mathematical and General, 23(22), L1169– L1173.Google Scholar
Reichl, L. E. 2009. A Modern Course in Statistical Physics. 3rd ed. Hoboken, NJ: John Wiley & Sons.Google Scholar
Renshaw, E. 1993. Modelling Biological Populations in Space and Time. Vol. 11. Cambridge, UK: Cambridge University Press.Google Scholar
Reynolds, A. 2015. Liberating Lévy walk research from the shackles of optimal foraging. Physics of Life Reviews, 14, 5983.Google Scholar
Reynolds, C. W. 1987. Flocks, herds and schools: a distributed behavioral model. ACM SIGGRAPH Computer Graphics, 21(4), 2534.Google Scholar
Ribble, D. O., and Stanley, S. 1998. Home ranges and social organization of syntopic Peromyscus boylii and P. truei. Journal of Mammalogy, 79(3), 932941.Google Scholar
Risken, H. 1984. The Fokker–Planck Equation: Methods of Solution and Applications. Berlin: Springer Series in Synergetics, vol. 18.Google Scholar
Robertson, G. E., Caldwell, G. E., Hamill, J., Kamen, G., and Whittlesey, S. 2013. Research Methods in Biomechanics. Champaign, IL: Human Kinetics Publishers.Google Scholar
Romanczuk, P., Bär, M., Ebeling, W., Lindner, B., and Schimansky-Geier, L. 2012. Active Brownian particles. The European Physical Journal Special Topics, 202(1), 1162.Google Scholar
Rubenstein, D. I. 1981. Individual variation and competition in the Everglades pygmy sunfish. Journal of Animal Ecology, 50, 337350.Google Scholar
Rubin, R. J., and Weiss, G. H. 1982. Random walks on lattices. The problem of visits to a set of points revisited. Journal of Mathematical Physics, 23(2), 250253.Google Scholar
Ruefenacht, B., and Knight, R. L. 1995. Influences of corridor continuity and width on survival and movement of deermice Peromyscus maniculatus. Biological Conservation, 71(3), 269274.Google Scholar
Saichev, A. I., and Zaslavsky, G. M. 1997. Fractional kinetic equations: solutions and applications. Chaos: An Interdisciplinary Journal of Nonlinear Science, 7(4), 753764.Google Scholar
Sarvaharman, S., Heiblum-Robles, A., and Giuggioli, L. 2019. From micro-to-macro: how the movement statistics of individual walkers affect the formation of segregated territories in the territorial random walk model. Frontiers in Physics, 7, 129.Google Scholar
Satomi, S. 2016. The Smoluchowski Equation in Population Dynamics and the Spread of Infection. PhD thesis, University of New Mexico, Albuquerque, NM.Google Scholar
Scher, H., and Montroll, E. W. 1975. Anomalous transit-time dispersion in amorphous solids. Physical Review B, 12(6), 24552477.Google Scholar
Schmaljohn, C., and Hjelle, B. 1997. Hantaviruses: a global disease problem. Emerging Infectious Diseases, 3(2), 95.Google Scholar
Schneider, W. R., and Wyss, W. 1989. Fractional diffusion and wave equations. Journal of Mathematical Physics, 30(1), 134144.Google Scholar
Scott, J. E., Kenkre, V. M., and Hurd, A. J. 1998. Nonlocal approach to the analysis of the stress distribution in granular systems. II. Application to experiment. Physical Review E, 57(5), 58505857.Google Scholar
Seuront, L., and Stanley, H. E. 2014. Anomalous diffusion and multifractality enhance mating encounters in the ocean. Proceedings of the National Academy of Sciences USA, 111(6), 22062211.Google Scholar
Sevilla, F. J., Dossetti, V., and Heiblum Robles, A. 2014. Synchronization and collective motion of globally coupled Brownian particles. Journal of Statistical Mechanics: Theory and Experiment, 2014(12), P12025.Google Scholar
Shigesada, N., and Kawasaki, K. 1997. Biological Invasions: Theory and Practice. Oxford, UK: Oxford University Press.Google Scholar
Shigesada, N., Kohkichi, K., and Teramoto, E. 1986. Traveling periodic waves in heterogeneous environments. Theoretical Population Biology, 30, 143160.Google Scholar
Shlesinger, M. F. 1974. Asymptotic solutions of continuous-time random walks. Journal of Statistical Physics, 10(5), 421434.Google Scholar
Shlesinger, M. F., and Klafter, J. 1986. Lévy walks versus Lévy flights. Pages 279–283 of: Stanley, H. E., and Ostrowsky, N. (eds.), On Growth and Form: Fractal and Non-Fractal Patterns in Physics. Dordrecht, Netherlands: Martinus Nijhoff Publishers.Google Scholar
Shlesinger, M. F., Klafter, J., and West, B. J. 1986. Lévy walks with applications to turbulence and chaos. Physica A: Statistical Mechanics and Its Applications, 140(1–2), 212218.Google Scholar
Shlesinger, M. F., Klafter, J., and Wong, Y. M. 1982. Random walks with infinite spatial and temporal moments. Journal of Statistical Physics, 27(3), 499512.Google Scholar
Shnerb, N. M. 2000. Extinction of a bacterial colony under forced convection in pie geometry. Physical Review E, 63(1), 011906.Google Scholar
Shugard, W, and Reiss, H. 1976. Derivation of the continuous-time random walk equation in non-homogeneous lattices. Journal of Chemical Physics, 65, 2827.Google Scholar
Sinhuber, M., and Ouellette, N. T. 2017. Phase coexistence in insect swarms. Physical Review Letters, 119(17), 178003.Google Scholar
Sinhuber, M., Van Der Vaart, K., Ni, R., Puckett, J. G., Kelley, D. H., and Ouellette, N. T. 2019. Three-dimensional time-resolved trajectories from laboratory insect swarms. Scientific Data, 6, 190036.Google Scholar
Skellam, J. G. 1951. Random dispersal in theoretical populations. Biometrika, 38, 196218.Google Scholar
Smoluchowski, M. von. 1916. Drei Vorträge über Diffusion, Brownsche Molekularbewegung und Koagulation von Kolloidteilchen II. Physikalische Zeitschrift, 17, 585599.Google Scholar
Sokolov, A., Goldstein, R. E., Feldchtein, F. I., and Aranson, I. S. 2009. Enhanced mixing and spatial instability in concentrated bacterial suspensions. Physical Review E, 80(3), 031903.Google Scholar
Spendier, K., and Kenkre, V. M. 2013. Analytic solutions for some reaction-diffusion scenarios. Journal of Physical Chemistry B, 117(49), 1563915650.Google Scholar
Spendier, K., Sugaya, S., and Kenkre, V. M. 2013. Reaction-diffusion theory in the presence of an attractive harmonic potential. Physical Review E, 88, 062142.Google Scholar
Stamps, J. A. 1994. Territorial behavior: testing the assumptions. Advances in the Study of Behavior, 23, 173232.Google Scholar
Stamps, J. A. 1995. Motor learning and the value of familiar space. American Naturalist, 146, 4158.Google Scholar
Stanley, H. E. 1971. Phase Transitions and Critical Phenomena. Oxford, UK: Clarendon Press.Google Scholar
Steiniger, S., and Hunter, A. J. S. 2013. A scaled line-based kernel density estimator for the retrieval of utilization distributions and home ranges from GPS movement tracks. Ecological Informatics, 13, 18.Google Scholar
Stickel, L. F. 1968. Home range and travels. Pages 373–411 of: King, J. A. (ed.), Biology of Peromyscus (Rodentia), Special Publication No. 2. Stillwater, OK: The American Society of Mammalogists.Google Scholar
Strandburg-Peshkin, A., Farine, D. R., Couzin, I. D., and Crofoot, M. C. 2015. Shared decision-making drives collective movement in wild baboons. Science, 348(6241), 13581361.Google Scholar
Strandburg-Peshkin, A., Twomey, C. R., Bode, N. W., Kao, A. B., Katz, Y., Ioannou, C. C., Rosenthal, S. B., Torney, C. J., Wu, H., Levin, S. A., and Couzin, I. D. 2013. Visual sensory networks and effective information transfer in animal groups. Current Biology, 23(17), R709–711.Google Scholar
Strausbaugh, L. J., Marfin, A. A., and Gubler, D. J. 2001. West Nile encephalitis: an emerging disease in the United States. Clinical Infectious Diseases, 33(10), 17131719.Google Scholar
Strogatz, S. H. 2018. Nonlinear Dynamics and Chaos: With Applications to Physics, Biology, Chemistry, and Engineering. Boca Raton, FL: CRC Press.Google Scholar
Strömbom, D. 2011. Collective motion from local attraction. Journal of Theoretical Biology, 283(1), 145151.Google Scholar
Sugaya, S., and Kenkre, V. M. 2018. Analysis of transmission of infection in epidemics: confined random walkers in dimensions higher than one. Bulletin of Mathematical Biology, 80(12), 31063126.Google Scholar
Sumpter, D. J. T. 2006. The principles of collective animal behaviour. Philosophical Transactions of the Royal Society B: Biological Sciences, 361(1465), 522.Google Scholar
Sumpter, D. J. T. 2010. Collective Animal Behavior. Princeton, NJ: Princeton University Press.Google Scholar
Swinney, H. L., Fitch, V. L., Marlow, D. R., and Dementi, M. A. E. 1996. Emergence and evolution patterns. Pages 51–74 of: Fitch, V. L., and Marlow, D. R. (eds.), Critical Problems in Physics: Proceedings of a Conference Celebrating the 250th Anniversary of Princeton University. Princeton, NJ: Princeton University Press.Google Scholar
Szabo, A., Lamm, G., and Weiss, G. H. 1984. Localized partial traps in diffusion processes and random walks. Journal of Statistical Physics, 34(1–2), 225238.Google Scholar
Szabó, B., Szöllösi, G. J., Gönci, B., Jurányi, Z., Selmeczi, D., and Vicsek, T. 2006. Phase transition in the collective migration of tissue cells: experiment and model. Physical Review E, 74(6), 061908.Google Scholar
Taylor, C. M., and Hastings, A. 2005. Allee effects in biological invasions. Ecology Letters, 8(8), 895908.Google Scholar
Taylor, , L. A, Taylor, G. K., Lambert, B., Walker, J. A., Biro, D., and Portugal, S. J. 2019. Birds invest wingbeats to keep a steady head and reap the ultimate benefits of flying together. PLoS Biology, 17(6).Google Scholar
Terman, C. R. 1968. Population dynamics. Pages 412–450 of: King, J. A. (ed.), Biology of Peromyscus (Rodentia), Special Publication No. 2. Stillwater, OK: The American Society of Mammalogists.Google Scholar
Theophilides, C. N., Ahearn, S. C., Grady, S., and Merlino, M. 2003. Identifying West Nile virus risk areas: the dynamic continuous-area space-time system. American Journal of Epidemiology, 157(9), 843854.Google Scholar
Theraulaz, G., and Bonabeau, E. 1999. A brief history of stigmergy. Artificial Life, 5(2), 97116.Google Scholar
Thompson, D. W. 1961. On Growth and Form. Cambridge, UK: Cambridge University Press.Google Scholar
Tilman, D., and Kareiva, P. (eds.). 1997. Spatial Ecology: The Role of Space in Population Dynamics and Interspecific Interactions. Vol. 89. Princeton, NJ: Princeton University Press.Google Scholar
Tinnin, D.S. 2003. Testing the Refugia Hypothesis: Population Dynamics of Peromyscus and Hantavirus Seroprevalence across an Elevational Gradient. MPhil thesis, University of New Mexico, Albuquerque, NM.Google Scholar
Tobin, P. C., Berec, L., and Liebhold, A. M. 2011. Exploiting Allee effects for managing biological invasions. Ecology Letters, 14(6), 615624.Google Scholar
Toner, J., and Tu, Y. 1995. Long-range order in a two-dimensional dynamical XY model: how birds fly together. Physical Review Letters, 75(23), 4326.Google Scholar
Toner, J., and Tu, Y. 1998. Flocks, herds, and schools: a quantitative theory of flocking. Physical Review E, 58(4), 4828.Google Scholar
Toner, J., Tu, Y., and Ramaswamy, S. 2005. Hydrodynamics and phases of flocks. Annals of Physics, 318(1), 170244.Google Scholar
Touchette, H., Van der Straeten, E., and Just, W. 2010. Brownian motion with dry friction: Fokker–Planck approach. Journal of Physics A: Mathematical and Theoretical, 43(44), 445002.Google Scholar
Toyozawa, Y. 1954. Theory of the electronic polaron and ionization of a trapped electron by an exciton. Progress of Theoretical Physics, 12(4), 421442.Google Scholar
Trepat, X., Wasserman, M. R., Angelini, T. E., Millet, E., Weitz, D. A., Butler, J. P., and Fredberg, J. J. 2009. Physical forces during collective cell migration. Nature Physics, 5(6), 426430.Google Scholar
Tsallis, C. 1988. Possible generalization of Boltzmann–Gibbs statistics. Journal of Statistical Physics, 52(1–2), 479487.Google Scholar
Tunstrøm, K., Katz, Y., Ioannou, C. C., Huepe, C., Lutz, M. J., and Couzin, I. D. 2013. Collective states, multistability and transitional behavior in schooling fish. PLoS Computational Biology, 9(2).Google Scholar
Turchin, P. 1998. Quantitative Analysis of Movement: Measuring and Modelling Population Redistribution in Animals and Plants. Sunderland, MA: Sinauer Associates.Google Scholar
Uhlenbeck, G. E. 1955. The statistical mechanics of non-equilibrium phenomena. Lecture Notes.Google Scholar
Vallée, O., and Soares, M. 2004. Airy Functions and Applications to Physics. Singapore: World Scientific.Google Scholar
van Saarloos, W. 2003. Front propagation into unstable states. Physics Reports, 386(2–6), 29222.Google Scholar
Vicsek, T., Czirók, A., Ben-Jacob, E., Cohen, I., and Shochet, O. 1995. Novel type of phase transition in a system of self-driven particles. Physical Review Letters, 75(6), 1226.Google Scholar
Vicsek, T., and Zafeiris, A. 2012. Collective motion. Physics Reports, 517, 71140.Google Scholar
Viswanathan, G. M., Buldyrev, S. V., Havlin, S., Da Luz, M. G. E., Raposo, E. P., and Stanley, H. E. 1999. Optimizing the success of random searches. Nature, 401(6756), 911.Google Scholar
Viswanathan, G. M., da Luz, M. G. E., Raposo, E. P., and Stanley, H. E. 2011. The Physics of Foraging: An Introduction to Random Searches and Biological Encounters. Cambridge, UK: Cambridge University Press.Google Scholar
Volkov, I., Banavar, J. R., He, F., Hubbell, S. P., and Maritan, A. 2005. Density dependence explains tree species abundance and diversity in tropical forests. Nature, 438(7068), 658.Google Scholar
Volpert, V., and Petrovskii, S. 2009. Reaction–diffusion waves in biology. Physics of Life Reviews, 6(4), 267310.Google Scholar
Volterra, V. 1927. Variazioni e fluttuazioni del numero d’individui in specie animali conviventi. C. Ferrari.Google Scholar
Wade, R. C., Gabdoulline, R. R., Lüdemann, S. K., and Lounnas, V. 1998. Electrostatic steering and ionic tethering in enzyme–ligand binding: Insights from simulations. Proceedings of the National Academy of Sciences USA, 95(11), 59425949.Google Scholar
Wakita, J.-I., Komatsu, K., Nakahara, A., Matsuyama, T., and Matsushita, M. 1994. Experimental investigation on the validity of population dynamics approach to bacterial colony formation. Journal of the Physical Society of Japan, 63(3), 12051211.Google Scholar
Wang, C., Horby, P. W., Hayden, F. G., and Gao, G. F. 2020. A novel coronavirus outbreak of global health concern. Lancet, 395(10223), 470473.Google Scholar
Wang, M. D., Yin, H., Landick, R., Gelles, J., and Block, S. M. 1997. Stretching DNA with optical tweezers. Biophysical Journal, 72(3), 1335.Google Scholar
Wang, Z., Alsayed, A. M., Yodh, A. G., and Han, Y. 2010. Two-dimensional freezing criteria for crystallizing colloidal monolayers. Journal of Chemical Physics, 132(15), 154501.Google Scholar
Ward, A. J. W., Sumpter, D. J. T., Couzin, I. D., Hart, P. J. B., and Krause, J. 2008. Quorum decision-making facilitates information transfer in fish shoals. Proceedings of the National Academy of Sciences USA, 105(19), 69486953.Google Scholar
Waser, P. M. 1981. Sociality or territorial defense? The influence of resource renewal. Behavioral Ecology and Sociobiology, 8, 231237.Google Scholar
Watts, D. J. 2000. Small Worlds: The Dynamics of Networks between Order and Randomness. Princeton, NJ: Princeton University Press.Google Scholar
Wax, N. 1954. Selected Papers on Noise and Stochastic Processes. New York: Courier Dover Publications.Google Scholar
Weber, H., Marx, D., and Binder, K. 1995. Melting transition in two dimensions: a finite-size scaling analysis of bond-orientational order in hard disks. Physical Review B, 51(20), 1463614651.Google Scholar
West, B. J., Bologna, M., and Grigolini, P. 2003. Physics of Fractal Operators. New York: Springer.Google Scholar
West, G. B., Brown, J. H., and Enquist, B. J. 1997. A general model for the origin of allometric scaling laws in biology. Science, 276(5309), 122126.Google Scholar
Wilson, E. O. 2000. Sociobiology: The New Synthesis. 25th ed. Cambridge, MA: Harvard University Press.Google Scholar
Wolff, J. O. 1993. Why are female small mammals territorial? Oikos, 68, 364370.Google Scholar
Wolff, J. O. 1997. Population regulation in mammals: an evolutionary perspective. Journal of Animal Ecology, 66, 113.Google Scholar
Wolff, J. O., and Peterson, J. A. 1998. An offspring-defense hypothesis for territoriality in female mammals. Ethology Ecology and Evolution, 10, 227239.Google Scholar
Yates, T. L., Mills, J. N., Parmenter, C. A., Ksiazek, T. G., Parmenter, R. R., Vande Castle, J, Calisher, R., Nichol, C. H., Abbott, S. T., Young, K. D., Morrison, J. C., Beaty, M. L., Dunnum, B. J., Baker, J. L., Salazar-Bravo, R. J., Peters, J., C. J. 2002. The ecology and evolutionary history of an emergent disease: Hantavirus pulmonary syndrome. BioScience, 52(11), 989998.Google Scholar
Young, A. P. 1979. Melting and the vector Coulomb gas in two dimensions. Physical Review B, 19(4), 18551866.Google Scholar
Zaburdaev, V., Denisov, S., and Klafter, J. 2015. Lévy walks. Reviews of Modern Physics, 87(2), 483.Google Scholar
Zumaya, M., Larralde, H., and Aldana, M. 2018. Delay in the dispersal of flocks moving in unbounded space using long-range interactions. Scientific Reports, 8(1), 19.Google Scholar
Zwanzig, R. 1964. On the identity of three generalized master equations. Physica, 30(6), 11091123.Google Scholar

Save book to Kindle

To save this book to your Kindle, first ensure coreplatform@cambridge.org is added to your Approved Personal Document E-mail List under your Personal Document Settings on the Manage Your Content and Devices page of your Amazon account. Then enter the ‘name’ part of your Kindle email address below. Find out more about saving to your Kindle.

Note you can select to save to either the @free.kindle.com or @kindle.com variations. ‘@free.kindle.com’ emails are free but can only be saved to your device when it is connected to wi-fi. ‘@kindle.com’ emails can be delivered even when you are not connected to wi-fi, but note that service fees apply.

Find out more about the Kindle Personal Document Service.

Available formats
×

Save book to Dropbox

To save content items to your account, please confirm that you agree to abide by our usage policies. If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account. Find out more about saving content to Dropbox.

Available formats
×

Save book to Google Drive

To save content items to your account, please confirm that you agree to abide by our usage policies. If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account. Find out more about saving content to Google Drive.

Available formats
×