Abstract
A stochastic differential game model for animal migration between two habitats under uncertain environment, a new population dynamics model, is formulated. Its novelty is the use of an impulse control formalism to naturally describe migrations with different timings and magnitudes that the conventional models could not handle. Uncertainty of the environment that the population faces with is formulated in the context of the multiplier robust control. The optimal migration strategy to give the maximized minimal profit is found through a Hamilton–Jacobi–Bellman quasi-variational inequality (HJBQVI). A key message from HJBQVI is that its free boundary determines the optimal migration strategy. Solving the HJBQVI is carried out with a specialized stable and convergent finite difference scheme. This paper theoretically suggests that the sub-additivity of the performance index, the index to be optimized through the migration, critically affects the resulting strategy. The computational results with the established scheme are consistent with the theoretical predictions and support importance of the sub-additivity property. Social interaction to reduce the net mortality rate is also quantified, suggesting a linkage between the present and existing population dynamics models.













Similar content being viewed by others
Abbreviations
- A i, j :
-
Diagonal coefficients of the discretized HJBQVI at the (i, j)-th vertex vi,j
- n :
-
Independent variable that represents the total population
- B :
-
The bivariate map that sums the cost and benefit of migration
- B t :
-
1-D standard Brownian motion
- c :
-
A nonnegative, bounded, and smooth univariate function
- C 1, C 2 :
-
Positive constants
- C ψ :
-
Parameter shaping the regularity of the awareness ψ
- E:
-
Expectation
- E :
-
Error norm in the policy iteration algorithm
- f :
-
Generic sufficiently smooth function
- G :
-
A continuous function
- h :
-
Free boundary of the solution to the HJBQVI
- J :
-
Performance index
- k 0, k 1, k 2 :
-
Weighting constants appearing in the migration cost B
- K :
-
Maximum body weight
- m n :
-
Intervals in the n-direction
- m x :
-
Intervals in the x-direction
- N t :
-
The total population in the habitat at the time t
- N max :
-
The maximum value of the total population Nt
- p i, j :
-
Integer to define the optimal control η* at each vertex
- s :
-
Time
- t :
-
Time
- \(u = \left({\tau_{1};\eta_{1},\tau_{2};\eta_{2}, \ldots} \right)\) :
-
Migration strategy
- u * :
-
Optimized u
- v i, j :
-
(i, j)-th vertex
- w t :
-
The uncertainty of the mortality rate at the time t
- w * :
-
Optimized w
- w max :
-
Maximum value of wt
- \(W = \left\{{\left. w \right|0 \le w \le w_{\max}} \right\}\) :
-
Admissible range of wt
- X t :
-
The representative body weight at the time t
- x :
-
Independent variable that represents the body weight
- α, β :
-
Model parameters shaping the functional form of the migration cost B
- γ :
-
Parameter shaping the regularity of the awareness ψ
- δ :
-
Discount rate
- ɛ :
-
Error tolerance in the policy iteration algorithm
- Δn :
-
Cell length in the n-direction
- Δx :
-
Cell length in the x-direction
- φ :
-
Test functions
- \(\Phi\) :
-
Value function
- \(\hat{\Phi}\) :
-
Piece-wise linear interpolation of discretized \(\Phi\)
- η i :
-
Total number of migrants of the ith migration
- η * :
-
Optimized η
- κ :
-
Parameter shaping the accumulated profit in the current habitat
- μ :
-
Deterministic body growth rate
- ψ :
-
Coefficient that represents the state-dependent awareness
- ψ 0 :
-
Parameter shaping ψ
- Ψ:
-
Dummy functions for defining viscosity solutions
- λ :
-
Base mortality rate
- ϑ :
-
Weighting content appearing in the accumulated profit in the current habitat
- ρ :
-
Penalty parameter
- σ :
-
Amplitude of the stochastic growth rate
- τ :
-
The smallest time s such that Ns = 0
- τ i :
-
Timing of the ith migration
- χ i, j :
-
Characteristic function to penalize the non-local term of the HJBQVI in numerical computation
- \(\Omega\) :
-
The domain to define the HJBQVI
- \(\bar{\Omega}\) :
-
Closure of \(\Omega\)
- \(\Omega_{\text{m}}\) :
-
The migration sub-domain. A sub-domain of \(\Omega\)
- \(\Omega_{\text{r}}\) :
-
The residency sub-domain. A sub-domain of \(\Omega\)
- \({\mathcal{F}}_{t}\) :
-
Natural filtration generated by \(B_{t}\). Its subscript is sometimes dropped when there is no confusion
- \({\mathcal{L}}\) :
-
Partial differential operator defining the HJBQVI
- \({\mathcal{L}}_{w}\) :
-
Partial differential operator defining a linear counterpart of the HJBQVI
- \({\mathcal{M}}\) :
-
Non-local operator defining the HJBQVI
- \({\mathcal{U}}\) :
-
The admissible set of u
- \({\mathcal{W}}\) :
-
The admissible set of w
- \({\mathbf{\mathbb{R}}}\) :
-
1-D real space
References
Altarovici A, Reppen M, Soner HM (2017) Optimal consumption and investment with fixed and proportional transaction costs. SIAM J Control Optim 55:1673–1710. https://doi.org/10.1137/15M1053633
Araujo HA, Cooper AB, MacIsaac EA, Knowler D, Velez-Espino A (2015) Modeling population responses of Chinook and coho salmon to suspended sediment using a life history approach. Theor Popul Biol 103:71–83. https://doi.org/10.1016/j.tpb.2015.04.003
Argasinski K, Broom M (2013) Ecological theatre and the evolutionary game: how environmental and demographic factors determine payoffs in evolutionary games. J Math Biol 67:935–962. https://doi.org/10.1007/s00285-012-0573-2
Argasinski K, Broom M (2018) Interaction rates, vital rates, background fitness and replicator dynamics: how to embed evolutionary game structure into realistic population dynamics. Theor Biosci 137:1–18. https://doi.org/10.1007/s12064-017-0257-y
Azimzadeh P (2017a) A zero-sum stochastic differential game with impulses, precommitment, and unrestricted cost functions. Appl Math Optim. https://doi.org/10.1007/s00245-017-9445-x
Azimzadeh P (2017b) Impulse control in finance: numerical methods and viscosity solutions. arXiv preprint arXiv:1712.01647
Ballerini M et al (2008) Empirical investigation of starling flocks: a benchmark study in collective animal behaviour. Anim Behav 76:201–215. https://doi.org/10.1016/j.anbehav.2008.02.004
Baltas I, Xepapadeas A, Yannacopoulos AN (2018) Robust control of parabolic stochastic partial differential equations under model uncertainty. Eur J Control. https://doi.org/10.1016/j.ejcon.2018.04.004
Barles G, Souganidis PE (1991) Convergence of approximation schemes for fully nonlinear second order equations. Asympt Anal 4:271–283. https://doi.org/10.3233/ASY-1991-4305
Barta Z et al (2008) Optimal moult strategies in migratory birds. Philos Trans R Soc B Biol Sci 363:211–229. https://doi.org/10.1098/rstb.2007.2136
Bauer S, Klaassen M (2013) Mechanistic models of animal migration behaviour—their diversity, structure and use. J Anim Ecol 82:498–508. https://doi.org/10.1111/1365-2656.12054
Bauer S et al (2011) Cues and decision rules in animal migration. In: Milner-Gulland EJ, Fryxell JM, Sinclair ARE (eds) Animal migration: a synthesis. Oxford University Press, Oxford, pp 68–87
Bauer S, Lisovski S, Hahn S (2016) Timing is crucial for consequences of migratory connectivity. Oikos 125:605–612. https://doi.org/10.1111/oik.02706
Becker DJ, Streicker DG, Altizer S (2015) Linking anthropogenic resources to wildlife-pathogen dynamics: a review and meta-analysis. Ecol Lett 18:483–495. https://doi.org/10.1111/ele.12428
Bensoussan A, Liu J, Yuan J (2010) Singular control and impulse control: a common approach. Discret Contin Dyn B 13:27–57. https://doi.org/10.3934/dcdsb.2010.13.27
Bensoussan A, Hoe S, Yan Z, Yin G (2017) Real options with competition and regime switching. Math Finance 27:224–250. https://doi.org/10.1111/mafi.12085
Berdahl A, Westley PA, Quinn TP (2017) Social interactions shape the timing of spawning migrations in an anadromous fish. Anim Behav 126:221–229. https://doi.org/10.1016/j.anbehav.2017.01.020
Bod’ová K et al (2018) Probabilistic models of individual and collective animal behavior. PLoS ONE 13:e0193049. https://doi.org/10.1371/journal.pone.0193049
Brönmark C et al (2008) Seasonal migration determined by a trade-off between predator avoidance and growth. PLoS ONE 3:e1957. https://doi.org/10.1371/journal.pone.0001957
Cadenillas A, Zapatero F (2000) Classical and impulse stochastic control of the exchange rate using interest rates and reserves. Math Finance 10:141–156. https://doi.org/10.1111/1467-9965.00086
Capasso V, Bakstein D (2005) An introduction to continuous-time stochastic processes. Birkhauser, Boston
Charmantier A, Gienapp P (2014) Climate change and timing of avian breeding and migration: evolutionary versus plastic changes. Evol Appl 7:15–28. https://doi.org/10.1111/eva.12126
Charnov EL, Turner TF, Winemiller KO (2001) Reproductive constraints and the evolution of life histories with indeterminate growth. PNAS 98:9460–9464. https://doi.org/10.1073/pnas.161294498
Chen W, Du K, Qiu X (2018) Analytic properties of American option prices under a modified Black–Scholes equation with spatial fractional derivatives. Phys A 491:37–44. https://doi.org/10.1016/j.physa.2017.08.068
Chevalier E, Gaïgi MH, Vath VL (2017) Liquidity risk and optimal dividend/investment strategies. Math Financ Econ 11:111–135. https://doi.org/10.1007/s11579-016-0173-9
Christensen S, Irle A, Ludwig A (2017) Optimal portfolio selection under vanishing fixed transaction costs. Adv Appl Probab 49:1116–1143. https://doi.org/10.1017/apr.2017.36
Colombo EH, Anteneodo C (2015) Metapopulation dynamics in a complex ecological landscape. Phys Rev E 92:022714. https://doi.org/10.1103/PhysRevE.92.022714
Conradt L (2011) Models in animal collective decision-making: information uncertainty and conflicting preferences. Interface Focus. https://doi.org/10.1098/rsfs.2011.0090
Córdova-Lepe F, Del Valle R, Ramos-Jiliberto R (2018) The process of connectivity loss during habitat fragmentation and their consequences on population dynamics. Ecol Model 376:68–75. https://doi.org/10.1016/j.ecolmodel.2018.01.012
Cote J et al (2017) Behavioural synchronization of large-scale animal movements—disperse alone, but migrate together? Biol Rev 92:1275–1296. https://doi.org/10.1111/brv.12279
Crandall MG, Ishii H, Lions PL (1992) User’s guide to viscosity solutions of second order partial differential equations. Bull Am Math Soc 27:1–67
Crozier LG et al (2017) High-stakes steeplechase: a behavior-based model to predict individual travel times through diverse migration segments. Ecosphere. https://doi.org/10.1002/ecs2.1965
Dall SRX et al (2005) Information and its use by animals in evolutionary ecology. Trends Ecol Evol 20:187–193. https://doi.org/10.1016/j.tree.2005.01.010
Danchin E et al (2004) Public information: from nosy neighbors to cultural evolution. Science 305:487–491. https://doi.org/10.1126/science.1098254
Dang DM, Forsyth PA (2014) Continuous time mean-variance optimal portfolio allocation under jump diffusion: an numerical impulse control approach. Numer Methods Partial Differ Equ 30:664–698. https://doi.org/10.1002/num.21836
Davis MH, Guo X, Wu G (2010) Impulse control of multidimensional jump diffusions. SIAM J Control Optim 48:5276–5293. https://doi.org/10.1137/090780419
De Leenheer P et al (2017) The puzzle of partial migration: adaptive dynamics and evolutionary game theory perspectives. J Theor Biol 412:172–185. https://doi.org/10.1016/j.jtbi.2016.10.011
Dixit AK, Pindyck RS (1994) Investment under uncertainty. Princeton University Press, Princeton
Dudley PN (2018) A salmonid individual-based model as a proposed decision support tool for management of a large regulated river. Ecosphere. https://doi.org/10.1002/ecs2.2074
El Farouq N, Barles G, Bernhard P (2010) Deterministic minimax impulse control. Appl Math Optim 61:353–378. https://doi.org/10.1007/s00245-009-9090-0
Filin I (2015) The relation between maternal phenotype and offspring size, explained by overhead material costs of reproduction. J Theor Biol 364:168–178. https://doi.org/10.1016/j.jtbi.2014.09.007
Forsyth PA, Labahn G (2008) Numerical methods for controlled HamiltonJacobi-Bellman PDEs in finance. J. Comp. Financ 11(Winter):1–44
Forsyth PA, Vetzal KR (2012) Numerical methods for nonlinear PDEs in finance. In: Duan JC, Härdle WK, Gentle JE (eds) Handbook of computational finance. Springer, Berlin, pp 503–528
Gil MA et al (2018) Social information links individual behavior to population and community dynamics. Trends Ecol Evol. https://doi.org/10.1016/j.tree.2018.04.010
Guo X, Wu G (2009) Smooth fit principle for impulse control of multidimensional diffusion processes. SIAM J Control Optim 48:594–617. https://doi.org/10.1137/080716001
Ha SY, Liu JG (2009) A simple proof of the Cucker–Smale flocking dynamics and mean-field limit. Commun Math Sci 7:297–325. https://doi.org/10.4310/CMS.2009.v7.n2.a2
Hansen L, Sargent TJ (2001) Robust control and model uncertainty. Am Econ Rev 91:60–66. https://doi.org/10.1257/aer.91.2.60
Hedenström A (2008) Adaptations to migration in birds: behavioural strategies, morphology and scaling effects. Philos Trans R Soc B Biol Sci 363:287–299. https://doi.org/10.1098/rstb.2007.2140
Hendriks AJ, Mulder C (2008) Scaling of offspring number and mass to plant and animal size: model and meta-analysis. Oecologia 155:705–716. https://doi.org/10.1007/s00442-007-0952-3
Hofbauer J, Sigmund K (2003) Evolutionary game dynamics. Bull Am Math Soc 40:479–519. https://doi.org/10.1090/S0273-0979-03-00988-1
Hofrichter J, Jost J, Tran TD (2017) Information geometry and population genetics. Springer, Berlin
Iwasa Y, Levin SA (1995) The timing of life history events. J Theor Biol 172:33–42. https://doi.org/10.1006/jtbi.1995.0003
Jansen JE, Van Gorder RA (2018) Dynamics from a predator–prey–quarry–resource–scavenger model. Theor Ecol 11:19–38. https://doi.org/10.1007/s12080-017-0346-z
Johansson J, Brännström Å, Metz JA, Dieckmann U (2018) Twelve fundamental life histories evolving through allocation-dependent fecundity and survival. Ecol Evol 8:3172–3186. https://doi.org/10.1002/ece3.3730
Jonsson N (1991) Influence of water flow, water temperature and light on fish migration in rivers. Nord J Freshw Res 66:20–35. https://doi.org/10.4236/jwarp.2013.55049
Jonsson N, Jonsson B (2002) Migration of anadromous brown trout Salmo trutta in a Norwegian river. Freshw Biol 47:1391–1401. https://doi.org/10.1046/j.1365-2427.2002.00873.x
Kaitala A, Kaitala V, Lundberg P (1993) A theory of partial migration. Am Nat 142:59–81. https://doi.org/10.1086/285529
Kalise D, Kunisch K (2018) Polynomial approximation of high-dimensional Hamilton–Jacobi–Bellman equations and applications to feedback control of semilinear parabolic PDEs. SIAM J Sci Comput 40:A629–A652. https://doi.org/10.1137/17M1116635
Kinnison MT, Unwin MJ, Hendry AP, Quinn TP (2001) Migratory costs and the evolution of egg size and number in introduced and indigenous salmon populations. Evolution 55:1656–1667. https://doi.org/10.1554/0014-3820(2001)055
Koleva MN, Valkov RL (2017) Modified barrier penalization method for pricing American options. In: Ehrhardt M, Günther M, ter Maten EJW (eds) Novel methods in computational finance. Springer, Cham, pp 215–226
Lande R, Engen S, Saether BE (2003) Stochastic population dynamics in ecology and conservation. Oxford University Press, Oxford
Larsson M (2009) Possible functions of the octavolateralis system in fish schooling. Fish Fish 10:344–353. https://doi.org/10.1111/j.1467-2979.2009.00330.x
Larsson M (2012) Incidental sounds of locomotion in animal cognition. Anim Cognit 15:1–13. https://doi.org/10.1007/s10071-011-0433-
Lehtonen J, Jaatinen K (2016) Safety in numbers: the dilution effect and other drivers of group life in the face of danger. Behav Ecol Sociobiol 70:449–458. https://doi.org/10.1007/s00265-016-2075-5
Lemasson BH, Haefner JW, Bowen MD (2014) Schooling increases risk exposure for fish navigating past artificial barriers. PLoS ONE 9:e108220. https://doi.org/10.1371/journal.pone.0108220
Lungu EM, Øksendal B (1997) Optimal harvesting from a population in a stochastic crowded environment. Math Biosci 145:47–75. https://doi.org/10.1016/S0025-5564(97)00029-1
Lv J, Wang K, Jiao J (2015) Stability of stochastic Richards growth model. Appl Math Model 39:4821–4827. https://doi.org/10.1016/j.apm.2015.04.016
Maeda S, Yoshida K, Kuroda H (2018) Turbulence and energetics of fish nest and pool structures in agricultural canal. Paddy Water Environ. https://doi.org/10.1007/s10333-018-0642-2
Magrath RD et al (2015) Eavesdropping on heterospecific alarm calls: from mechanisms to consequences. Biol Rev 90:560–586. https://doi.org/10.1111/brv.12122
Mann RP, Helbing D (2017) Optimal incentives for collective intelligence. PNAS 114:5077–5082. https://doi.org/10.1073/pnas.1618722114
Mariani P, Křivan V, MacKenzie BR, Mullon C (2016) The migration game in habitat network: the case of tuna. Theor Ecol 9:219–232. https://doi.org/10.1007/s12080-015-0290-8
McLaren JD et al (2014) Optimal orientation in flows: providing a benchmark for animal movement strategies. J R Soc Interface 11:20140588. https://doi.org/10.1098/rsif.2014.0588
McNamara JM, Houston AI, Collins EJ (2001) Optimality models in behavioral biology. SIAM Rev 43:413–466. https://doi.org/10.1137/S0036144500385263
Mohapatra A, Ohms HA, Lytle DA, De Leenheer P (2016) Population models with partial migration. J Differ Equ Appl 22:316–329. https://doi.org/10.1080/10236198.2015.1091451
Morales JM et al (2010) Building the bridge between animal movement and population dynamics. Philos Trans R Soc B Biol Sci 365:2289–2301. https://doi.org/10.1098/rstb.2010.0082
Mouri G, Shinoda S, Oki T (2010) Estimating Plecoglossus altivelis altivelis migration using a mass balance model expressed by hydrological distribution parameters in a major limpid river basin in Japan. Ecol Model 221:2808–2815. https://doi.org/10.1016/j.ecolmodel.2010.08.029
Nagaya T et al (2008) Evaluation of suitable hydraulic conditions for spawning of ayu with horizontal 2D numerical simulation and PHABSIM. Ecol Model 215:133–143. https://doi.org/10.1016/j.ecolmodel.2008.02.043
Nowak MA, Sigmund K (2004) Evolutionary dynamics of biological games. Science 303:793–799. https://doi.org/10.1126/science.1093411
Nutz M (2018) A mean field game of optimal stopping. SIAM J Control Optim 56:1206–1221. https://doi.org/10.1137/16M1078331
Oberman AM (2006) Convergent difference schemes for degenerate elliptic and parabolic equations: Hamilton–Jacobi equations and free boundary problems. SIAM J Numer Anal 44:879–895. https://doi.org/10.1137/S0036142903435235
Øksendal B (2003) Stochastic differential equations. In: Øksendal B (ed) Stochastic differential equations. Springer, Berlin, pp 65–84
Øksendal BK, Sulem A (2005) Applied stochastic control of jump diffusions, vol 498. Springer, Berlin
Olsson KH, Gislason H (2016) Testing reproductive allometry in fish. ICES J Mar Sci 73:1466–1473. https://doi.org/10.1093/icesjms/fsw01
Perera S, Long H (2017) An approximation scheme for impulse control with random reaction periods Oper. Res Lett 45:585–591. https://doi.org/10.1016/j.orl.2017.08.014
Peskir G, Shiryaev A (2006) Optimal stopping and free-boundary problems. Basel, Birkhäuser, pp 123–142
Phillips JA et al (2018) An asymmetric producer-scrounger game: body size and the social foraging behavior of coho salmon. Theor Ecol 9:9. https://doi.org/10.1007/s12080-018-0375
Reichard M, Jurajda P, Ondračkovaá M (2002) Interannual variability in seasonal dynamics and species composition of drifting young-of-the-year fishes in two European lowland rivers. J Fish Biol 60:87–101. https://doi.org/10.1111/j.1095-8649.2002.tb02389.x
Reid JM et al (2018) Population and evolutionary dynamics in spatially structured seasonally varying environments. Biol Rev. https://doi.org/10.1111/brv.12409
Reina A, Miletitch R, Dorigo M, Trianni V (2015) Quantitative micro–macro link for collective decisions: the shortest path discovery/selection example. Swarm Intell 9:75–102. https://doi.org/10.1007/s11721-015-0105-y
Sachs G, Lenz J (2011) New modeling approach for bounding flight in birds. Math Biosci 234:75–83. https://doi.org/10.1016/j.mbs.2011.08.005
Sahashi G, Morita K (2018) Adoption of alternative migratory tactics: a view from the ultimate mechanism and threshold trait changes in a salmonid fish. Oikos 127:239–251. https://doi.org/10.1111/oik.03715
Sahashi G, Morita K, Kishi D (2018) Spatial expansion and increased population density of masu salmon parr independent of river restoration. Ichthyol Res. https://doi.org/10.1007/s10228-018-0628-5
Sainmont J, Thygesen UH, Visser AW (2013) Diel vertical migration arising in a habitat selection game. Theor Ecol 6:241–251. https://doi.org/10.1007/s12080-012-0174-0
Salinger DH, Anderson JJ (2006) Effects of water temperature and flow on adult salmon migration swim speed and delay. Trans Am Fish Soc 135:188–199. https://doi.org/10.1577/T04-181.1
Sanz-Aguilar A et al (2012) To leave or not to leave: survival trade-offs between different migratory strategies in the greater flamingo. J Anim Ecol 81:1171–1182. https://doi.org/10.1111/j.1365-2656.2012.01997.x
Sasaki A, Iwasa Y (1991) Optimal growth schedule of pathogens within a host: switching between lytic and latent cycles. Theor Popul Biol 39:201–239. https://doi.org/10.1016/0040-5809(91)90036-F
Satterfield DA, Marra PP, Sillett TS, Altizer S (2018) Responses of migratory species and their pathogens to supplemental feeding. Philos Trans R Soc B Biol Sci B 373:20170094. https://doi.org/10.1098/rstb.2017.0094
Schaefer M, Menz S, Jeltsch F, Zurell D (2018) sOAR: a tool for modelling optimal animal life-history strategies in cyclic environments. Ecography 41:551–557. https://doi.org/10.1111/ecog.03328
Shang Y, Bouffanais R (2014) Influence of the number of topologically interacting neighbors on swarm dynamics. Sci Rep 4:4184. https://doi.org/10.1038/srep04184
Skubic E, Taborsky M, McNamara JM, Houston AI (2004) When to parasitize? A dynamic optimization model of reproductive strategies in a cooperative breeder. J Theor Biol 227:487–501. https://doi.org/10.1016/j.jtbi.2003.11.021
Sumpter DJ, Pratt SC (2009) Quorum responses and consensus decision making. Philos Trans R Soc B Biol Sci 364(1518):743–753. https://doi.org/10.1098/rstb.2008.0204
Sumpter D, Buhl J, Biro D, Couzin I (2008) Information transfer in moving animal groups. Theor Biosci 127:177–186. https://doi.org/10.1007/s12064-008-0040-1
Tago Y (2002) Migration behaviors of sea-run ayu Plecoglossus altivelis (Pisces) in Toyama Bay, Japan. Nippon Suisan Gakkaishi 68:554–563. https://doi.org/10.2331/suisan.68.554 (in Japanese with English Abstract)
Tago Y (2004) Relationship between body size of ayu migrating up rivers flowing into Toyama Bay and water temperature. Aquac Sci 52:315–323. https://doi.org/10.11233/aquaculturesci1953.52.315 (in Japanese with English Abstract)
Takai N et al (2018) The seasonal trophic link between Great Cormorant Phalacrocorax carbo and ayu Plecoglossus altivelis altivelis reared for mass release. Ecol Res. https://doi.org/10.1007/s11284-018-1610-4
Taylor PD, Jonker LB (1978) Evolutionary stable strategies and game dynamics. Math Biosci 40:145–156. https://doi.org/10.1016/0025-5564(78)90077-9
Thomas LH (1949) Elliptic problems in linear difference equations over a network. Watson Scientific Computing Laboratory Report, Columbia University, New York
Tian L et al (2017) The valuation of photovoltaic power generation under carbon market linkage based on real options. Appl Energy 201:354–362. https://doi.org/10.1016/j.apenergy.2016.12.092
Tran TD, Hofrichter J, Jost J (2013) An introduction to the mathematical structure of the Wright–Fisher model of population genetics. Theor Biosci 132:73–82. https://doi.org/10.1007/s12064-012-0170-3
Tran TD, Hofrichter J, Jost J (2014) The evolution of moment generating functions for the Wright–Fisher model of population genetics. Math Biosci 256:10–17. https://doi.org/10.1016/j.mbs.2014.07.007
Tran TD, Hofrichter J, Jost J (2015) The free energy method and the Wright–Fisher model with 2 alleles. Theor Biosci 134:83–92. https://doi.org/10.1007/s12064-015-0218-2
Tsujimura M (2015) Pollutant abatement investment under ambiguity in a two-period model. Int J Real Options Strategy 3:13–26. https://doi.org/10.12949/ijros.3.13
Varpe Ø (2017) Life history adaptations to seasonality. Integr Comp Biol 57:943–960. https://doi.org/10.1093/icb/icx123
Vicsek T, Zafeiris A (2012) Collective motion. Phys Rep 517:71–140. https://doi.org/10.1016/j.physrep.2012.03.004
Wang S, Yang X (2015) A power penalty method for a bounded nonlinear complementarity problem. Optimization 64:2377–2394. https://doi.org/10.1080/02331934.2014.967236
Wang X, Pan Q, Kang Y, He M (2016) Predator group size distributions in predator–prey systems. Ecol Complex 26:117–127. https://doi.org/10.1016/j.ecocom.2016.04.003
Ward A, Webster M (2016) Sociality: the behaviour of group-living animals. Springer, Cham, pp 149–174
Weinan E, Yu B (2018) The Deep Ritz method: a deep learning-based numerical algorithm for solving variational problems. Commun Math Stat 6:1–12. https://doi.org/10.1007/s40304-018-0127-z
Wikelski M et al (2003) Avian metabolism: costs of migration in free-flying songbirds. Nature 423(6941):704. https://doi.org/10.1038/423704a
Yaegashi Y, Yoshioka H, Unami K, Fujihara M (2018) A singular stochastic control model for sustainable population management of the fish-eating waterfowl Phalacrocorax carbo. J Environ Manag 219:18–27. https://doi.org/10.1016/j.jenvman.2018.04.099
Yi B, Viens F, Li Z, Zeng Y (2015) Robust optimal strategies for an insurer with reinsurance and investment under benchmark and mean–variance criteria. Scand Actuar Journal 8:725–751. https://doi.org/10.1080/03461238.2014.883085
Yoshioka H (2016) Mathematical analysis and validation of an exactly solvable model for upstream migration of fish schools in one-dimensional rivers. Math Biosci 281:139–148. https://doi.org/10.1016/j.mbs.2016.09.014
Yoshioka H (2017) A simple game-theoretic model for upstream fish migration. Theor Biosci 136:99–111. https://doi.org/10.1007/s12064-017-0244-3
Yoshioka H (2018a) ‘Mathematical exercise’ on a solvable stochastic control model for animal migration. ANZIAM J 59:C15–C28. https://doi.org/10.21914/anziamj.v59i0.12566
Yoshioka H (2018b) An exactly solvable multiple optimal stopping problem. Adv Differ Equ 2018:173. https://doi.org/10.1186/s13662-018-1626-7
Yoshioka H, Yaegashi Y (2017a) Optimization model to start harvesting in stochastic aquaculture system. Appl Stoch Model Bus 33:476–493. https://doi.org/10.1002/asmb.2250
Yoshioka H, Yaegashi Y (2017b) Robust stochastic control modeling of dam discharge to suppress overgrowth of downstream harmful algae. Appl Stoch Model Bus. https://doi.org/10.1002/asmb.2301
Yoshioka H, Yaegashi Y (2018a) Mathematical analysis for management of released fish. Opt Control Appl Methods 39:1141–1146. https://doi.org/10.1002/oca.2392
Yoshioka H, Yaegashi Y (2018b) An optimal stopping approach for onset of fish migration. Theor Biosci. https://doi.org/10.1007/s12064-018-0263-8
Yoshioka H, Unami K, Fujihara M (2014) Mathematical analysis on a conforming finite element scheme for advection–dispersion–decay equations on connected graphs. J JSCE Ser A2 70:I_265–I_274. https://doi.org/10.2208/jscejam.70.I_265
Yoshioka H, Yaegashi Y, Unami K, Fujihara M (2016) Application of stochastic control theory to biophysics of fish migration around a weir equipped with fishways. In: Zhang L, Song X, Wu Y (eds) Theory, methodology, tools and applications for modeling and simulation of complex systems. Springer, Singapore, pp 190–200
Yoshioka H, Shirai T, Tagami D (2017) Viscosity solutions of a mathematical model for upstream migration of potamodromous fish, In: 12th SDEWES conference, October 4–8, 2017, Dubrovnik, Proceedings, Paper No. 571, pp 571-1–571-12
Yoshioka H, Yaegashi Y, Yoshioka Y, Tsugihashi T (2018) Non-renewable fishery resource management under incomplete information. In: The 20th European conference on mathematics for industry, 18–22 June 2018, Budapest (Accepted)
Yoshioka H, Shirai T, Tagami D (2019) A mixed optimal control approach for upstream fish migration. J Sust Dev Energy Water Environ Syst 7:101–121. https://doi.org/10.13044/j.sdewes.d6.0221
Zhang K, Yang XQ (2017) Pricing European options on zero-coupon bonds with a fitted finite volume method. Int J Numer Anal Model 14:405–418. https://doi.org/10.1007/s11274-015-1903-5
Zhao G, Zhai K, Zong G (2018) On optimal stopping and free boundary problems under ambiguity. Stat Probab Lett 9:9. https://doi.org/10.1016/j.spl.2018.04.005
Zielinski DP, Voller VR, Sorensen PW (2018) A physiologically inspired agent-based approach to model upstream passage of invasive fish at a lock-and-dam. Ecol Model 382:18–32. https://doi.org/10.1016/j.ecolmodel.2018.05.004
Acknowledgements
JSPS Research Grant No. 17K15345 and a grant for ecological survey of a life history of the landlocked ayu Plecoglossus altivelis from MLIT of Japan supports this research. The author thanks the two reviewers for their critical comments and suggestions on biological and mathematical descriptions in this paper. The author is grateful to a reviewer for helpful comments, suggestions, and discussions on the performance index from a viewpoint of evolutionary biology.
Author information
Authors and Affiliations
Corresponding author
Additional information
Publisher's Note
Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.
Appendices
Appendix 1
Proofs of several propositions in the main text are presented in this appendix. As described in the main text, we set K = 1 without any loss of generality.
Proof of Proposition 3-1
Set \(x_{1},x_{2},n_{1},n_{2} \in \left[{0,1} \right]\) and fix \(u \in {\mathcal{U}}\left({n_{1}} \right) \cap {\mathcal{U}}\left({n_{2}} \right)\) and \(w \in {\mathcal{W}}\). The process \(X\) with the initial condition \(x_{i}\) is denoted as \(X^{\left(i \right)}\) (\(i = 1,2\)). Similar notation applies to \(N\). Then, we have
The right-hand side is bounded from above by the quantity
Therefore, we obtain
The three expectations appearing in (56) have to be evaluated. Firstly, by the path-wise uniqueness, the estimate
with a positive constant \(C_{1}\) follows, which leads to
assuming that the expectation and integral are interchanged. Here, the coefficient \(C_{1}\) can be taken sufficiently large so that it is independent of the controls. This is because the intervention (4) is linear with respect to \(\eta_{i}\) as in Guo and Wu (2009).
The first expectation of (56) is evaluated as
By Proposition 1 with the argument in Proof of Lemma 3.1 of Davis et al. (2010), we have
On the other hand, application of Itô’s formula to \(Y_{s}^{\left( i \right)} = \left( {X_{s}^{\left( i \right)} } \right)^{\beta }\) (\(i = 1,2\)) yields
with
Then, again as in Proof of Lemma 3.1 of Davis et al. (2010), we have
By the Fubini’s theorem and Proposition 1 (\(0 \le Y_{t}^{\left( i \right)} ,X_{t}^{\left( i \right)} \le 1\)), the expectations in the right-hand side of (63) are estimated from above as
and
By (64) and (65), (63) reduces to
Therefore, there exists a sufficiently large constant \(C_{2} > 0\) such that
Thus, with a sufficiently large \(\delta\) (Assumption 2), by (60) we get
Substituting (68) into (59) yields
On the last expectation of (56), we have
By (69) and an analogous discussion to (57), (70) leads to
with \(C_{2} > 0\) that is taken to be larger if necessary, by Assumption 2. Combining (58), (69), and (71) leads to
where \(G\) is continuous. Then, we have
and thus
Similarly, exchanging \(\left( {x_{1} ,n_{1} } \right)\) and \(\left( {x_{2} ,n_{2} } \right)\) in (73) gives
The proof is thus completed.□
Proof of Proposition 3-2
Fix an admissible \(w \in {\mathcal{W}}\). Firstly, substituting the null control \(u \equiv 0\) where \(\eta_{i} = 0\) (\(i = 0,1,2,3,\ldots\)) to \(J\) yields \(0 \le \Phi\). Secondly, substituting the control \(\eta_{0} = n\), \(\tau_{0} = 0\), \(\eta_{i} = 0\) (\(i = 1,2,3,\ldots\)) to \(J\) yields \(- B\left({x,n} \right) \le \Phi\), which thus leads to the lower bound \(\hbox{max} \left\{{0, - B\left({x,n} \right)} \right\} \le \Phi \left({x,n} \right)\). Finally, we have the estimate
which is the upper bound.□
Proof of Proposition 3-3
The statement immediately follows from the order property \({\mathcal{U}}\left({n_{1}} \right) \subset {\mathcal{U}}\left({n_{2}} \right)\) for \(0 \le n_{1} \le n_{2} \le 1\) and the increasing property of the performance index with respect to \(N\) and \(X\).□
Proof of Proposition 3-4
The proof of the proposition is just the same with those of Propositions 2.3 through 2.5 of Yoshioka and Yaegashi (2018b).□
Proof of Proposition 4
The proof of \(\alpha = 1\) is trivial, and we therefore assume \(0 < \alpha < 1\). An elementary calculation shows that the bivariate function \(F\left({\eta_{1},\eta_{2}} \right)\) for \(\eta_{1},\eta_{2} > 0\) is concave and takes the maximum value \(2^{\alpha} - 2 < 0\) at \(\left({\eta_{1},\eta_{2}} \right) = \left({1,1} \right)\). This shows that the right-hand side of inequality (28) is negative for \(\eta_{1},\eta_{2} > 0\), which completes the proof.□
Proof of Proposition 5
We have \(F\left({1,1} \right) = 2^{\alpha} - 2 > 0\), implying that the condition (28) fails if \(\frac{{k_{0}}}{{k_{1}}}\) is sufficiently small.□
Proof of Proposition 6
For \(\left({x,\eta} \right) \in \left[{0,1} \right] \times \left({0,1} \right]\), \(B\) is evaluated from below as
By an elementary calculation, we have
If \(k_{0}\) is sufficiently large, then
implying that the null control \(u \equiv 0\) is optimal, which completes the proof.□
Proof of Proposition 7
Substituting \(\Phi \left({x,n} \right) = c\left(x \right)n^{z}\) with a constant \(z\) into (15) under the stated assumptions yields
with
The relationship (81) holds for small \(n\) if \(z = 1\), leading to the desired result.□
Proof of Proposition 8
To show that value function \(\Phi\) satisfies the viscosity property in \(\left({0,1} \right) \times \left({0,1} \right)\) is straightforward, which can be proven essentially following the argument in the proof of Theorem 9.8 in Øksendal and Sulem (2005). This is because of the regularity and boundedness assumptions on the coefficients in the SDEs of the population dynamics and the performance index \(J\). The condition for \(nx = 0\) is straightforward to check since the value function \(\Phi\) equals 0 when \(nx = 0\). Therefore, the condition for the boundaries \(n = 1\) and \(x = 1\) should be checked, which seems to be non-trivial as in the case for \(\left({0,1} \right) \times \left({0,1} \right)\). For the sake of brevity, define
for generic sufficiently regular \(f = f\left({x,n} \right)\) and \(w \in {\mathcal{W}}\). Then, we have
A key in the proof for \(n = 1\) is that the process \(\left({X_{t},N_{t}} \right)\) is valued in \(\bar{\Omega}\) by Proposition 1 and Eqs. (3) and (4), without further constraints. This fact enables us to formally apply the existing proofs based on dynamic programming principles for controlled diffusion processes. The proof for this case proceeds in an essentially the same way with that of Theorem 9.8 in Øksendal and Sulem (2005) with minor modifications specific to the present problem. Notice that we have
This is proven in the same way with Lemma 9.7 of Øksendal and Sulem (2005) based on an argument by contradiction using \(\varepsilon\)-optimal controls.The proof for \(n = 1\) is as follows. The proof for \(x = 1\) is not presented here, since it is essentially the same with that for \(n = 1\). The viscosity sub-solution property is firstly checked. Set a point \(\left({x,n} \right) = \left({x_{0},1} \right)\) with \(0 < x_{0} < 1\) and take a test function \(\varphi\) for viscosity sub-solutions that fulfills the conditions of Definition 2. We should prove the inequality
If \(\Phi - {\mathcal{M}}\Phi = 0\) at \(\left({x_{0},1} \right)\), then (86) trivially holds. Therefore, we assume \(\Phi - {\mathcal{M}}\Phi > 0\) at this point. Hence, we have to show
Since \(\tau_{1}^{*}\) is a stopping time, we have either \(\tau_{1}^{*} = 0\) or \(\tau_{1}^{*} > 0\) a.s.. If \(\tau_{1}^{*} = 0\) a.s., then the process \(N_{t}\) controlled by the optimizer \(u^{*}\) immediately jumps from \(n = 1\) to some \(n = \underset{\raise0.3em\hbox{$\smash{\scriptscriptstyle-}$}}{n}\) with \(0 \le \underset{\raise0.3em\hbox{$\smash{\scriptscriptstyle-}$}}{n} < 1\). Therefore,
where \(\bar{u}^{*} = \left({\tau_{2}^{*};\eta_{2}^{*},\tau_{3}^{*};\eta_{3}^{*}, \ldots} \right)\). This is a contradiction since \(\Phi - {\mathcal{M}}\Phi > 0\) at \(\left({x_{0},1} \right)\). Therefore, \(\tau_{1}^{*} = 0\) a.s. is impossible. If \(\tau_{1}^{*} > 0\) a.s., then choose \(R_{0} < + \infty\) and sufficiently small \(\rho_{0} > 0\) such that
By the definition, we have \({\text{E}}\left[{\bar{\tau}} \right] > 0\). By the dynamic programming principle, for any \(\varepsilon_{0} > 0\), there exists an \(\varepsilon_{0}\)-optimal control \(u_{0} \in {\mathcal{U}}\) such that
where \(u = u_{0}\) in the second line of (90). Since \(\varphi \ge \Phi\), combining (90) and the Dynkin formula (Theorem 1.24 of Øksendal and Sulem 2005) for the process \(e^{- \delta s} \varphi \left({X_{s},N_{s}} \right)\) gives
Using the assumption \(\Phi \left({x_{0},1} \right) = \varphi \left({x_{0},1} \right)\) leads to
and thus
Since \(\varepsilon_{0}\) can be taken arbitrary small, letting \(\rho_{0} \to + 0\) in (93) gives
which is the desired result. This is because, if \({\mathcal{L}}\varphi \left({x_{0},1} \right) > 0\), then the inequality \({\mathcal{L}}\varphi > 0\) holds in a neighborhood of the point \(\left({x_{0},1} \right)\) (This neighborhood should be contained in \(\Omega\)). Then, for sufficiently small \(\varepsilon_{0}\) and \(\rho_{0}\), the inequality (93) is violated. This is a contradiction.
The viscosity super-solution property is secondly checked. Take a test function \(\varphi\) for viscosity super-solutions that fulfills the conditions of Definition 2. We have to show the inequality
Since we always have \(\Phi - {\mathcal{M}}\Phi \ge 0\), it is sufficient to show
Set a null control \(u \in {\mathcal{U}}\) with \(\tau_{1} \to + \infty\). Set \(\bar{\tau} = \hbox{min} \left\{{\tau,\rho_{0}} \right\}\) with a positive constant \(\rho_{0} > 0\). We have \({\text{E}}\left[{\bar{\tau}} \right] > 0\) by \(0 < x_{0} < 1\) and \(n = 1\). Since \(\varphi \le \Phi\), combining the dynamic programming principle and the Dynkin formula gives
and thus
Dividing both sides of (101) by \({\text{E}}\left[{\bar{\tau}} \right]\) and letting \(\rho_{0} \to + 0\) gives the desired result (96).□
Proof of Proposition 10
The monotonicity statement directly follows from the positive coefficient condition by Proposition 9. The consistency follows from the fact that the one-sided first-order and exponential discretization used in the present scheme are consistent for smooth solutions, which is sufficient for guaranteeing the consistency in the above-mentioned sense.□
Appendix 2
Convergence of the present finite difference scheme is carried out to demonstrate that the computational resolution \(m_{x} = m_{n} = 256\) employed in the main text is sufficiently fine for investigations carried out in this paper. Table 2 summarizes the difference between the reference solution with \(m_{x} = m_{n} = 1024\) and numerical solutions having coarser resolutions with \(m_{x} = m_{n}\). The reference solution is used here for analyzing convergence of numerical solutions since the present HJBQVI is not exactly solvable. The difference between the reference and numerical solutions is measured through the \(l^{2}\) and \(l^{\infty}\) norms. Table 2 indicates that the errors between the numerical solution with \(m_{x} = m_{n} = 256\) and the reference solution are sufficiently small compared with the magnitude of \(\Phi\). This computational result justifies the employed resolution \(m_{x} = m_{n} = 256\). The results presented in Table 2 imply that the convergence rate of the present finite difference scheme is almost first order, which is consistent with the theoretical estimate (Oberman 2006).
Next, the computed \(\Phi\) is examined against the theoretical upper and lower bounds. It was found by preliminary investigations that the lower bound is far sharper than the upper bound and the computed \(\Phi\) obviously satisfies the upper bound. Therefore, the computed \(\Phi\) is examined against the lower bound here. Figure 14 shows the difference between the computed \(\Phi\) and the lower bound: namely, \(\Delta \Phi = \Phi_{\text{computed}} - \Phi_{\text{lower bound}}\) with \(m_{x} = m_{n} = 256\), which should not be negative theoretically. Figure 14 shows that the violation of the lower bound is very small and the bound is sharp for small \(x\) or large \(n\) in particular. Table 3 shows the minimum value of \(\Delta \Phi\) over the computational domain. The computational results presented in Fig. 14 and Table 3 demonstrate that the present numerical scheme does not completely satisfy the lower bound, but the violation is almost the order of \(O\left({\rho^{- 1}} \right)\), which can be made sufficiently small. For example, \(\rho = O\left(m \right)\) is a reasonable choice from the viewpoint of computational accuracy of the proposed discretization method.
Rights and permissions
About this article
Cite this article
Yoshioka, H. A stochastic differential game approach toward animal migration. Theory Biosci. 138, 277–303 (2019). https://doi.org/10.1007/s12064-019-00292-4
Received:
Accepted:
Published:
Issue Date:
DOI: https://doi.org/10.1007/s12064-019-00292-4