Skip to main content
Log in

Spatial movement with distributed memory

  • Published:
Journal of Mathematical Biology Aims and scope Submit manuscript

Abstract

Diffusion has been widely applied to model animal movement that follows Brownian motion. However, animals typically move in non-Brownian ways due to their perceptual judgment. Spatial memory and cognition recently have received much attention in characterizing complicated animal movement behaviours. Explicit spatial memory is modeled via a distributed delayed diffusion term in this paper. The distributed time represents the memory growth and decay over time, and the spatial nonlocality reflects the dependence of spatial memory on location. When the temporal delay kernel is weak under the assumption that animals can immediately acquire knowledge and memory decays over time, the equation is equivalent to a Keller–Segel chemotaxis model. For the strong kernel with learning and memory decay stages, rich spatiotemporal dynamics, such as Turing and checker-board patterns, appear via spatially non-homogeneous steady-state and Hopf bifurcations.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Institutional subscriptions

Fig. 1
Fig. 2
Fig. 3
Fig. 4

Similar content being viewed by others

References

  • Abrahms B, Hazen EL, Aikens EO, Savoca MS, Goldbogen JA, Bograd SJ, Jacox MG, Irvine LM, Palacios DM, Mate BR (2019) Memory and resource tracking drive blue whale migrations. Proc Natl Acad Sci 116(12):5582–5587

    Article  Google Scholar 

  • Amann H (1991) Hopf bifurcation in quasilinear reaction-diffusion systems. In: Delay differential equations and dynamical systems (Claremont, CA, 1990), volume 1475 of Lecture Notes in Mathematics. Springer, Berlin, pp 53–63

  • Bellomo N, Bellouquid A, Tao Y, Winkler M (2015) Toward a mathematical theory of Keller–Segel models of pattern formation in biological tissues. Math Models Methods Appl Sci 25(9):1663–1763

    Article  MathSciNet  MATH  Google Scholar 

  • Britton NF (1990) Spatial structures and periodic travelling waves in an integro-differential reaction–diffusion population model. SIAM J Appl Math 50(6):1663–1688

    Article  MathSciNet  MATH  Google Scholar 

  • Chen S, Yu J (2016) Stability analysis of a reaction–diffusion equation with spatiotemporal delay and Dirichlet boundary condition. J Dyn Differ Equ 28(3–4):857–866

    Article  MathSciNet  MATH  Google Scholar 

  • Cooke KL, Grossman Z (1982) Discrete delay, distributed delay and stability switches. J Math Anal Appl 86(2):592–627

    Article  MathSciNet  MATH  Google Scholar 

  • Crandall MG, Rabinowitz PH (1971) Bifurcation from simple eigenvalues. J Funct. Anal. 8:321–340

    Article  MathSciNet  MATH  Google Scholar 

  • Crandall MG, Rabinowitz PH (1973) Bifurcation, perturbation of simple eigenvalues and linearized stability. Arch Ration Mech Anal 52:161–180

    Article  MathSciNet  MATH  Google Scholar 

  • Ducrot A, Fu X, Magal P (2018) Turing and Turing–Hopf bifurcations for a reaction diffusion equation with nonlocal advection. J Nonlinear Sci 28(5):1959–1997

    Article  MathSciNet  MATH  Google Scholar 

  • Fagan WF (2019) Migrating whales depend on memory to exploit reliable resources. Proc Natl Acad Sci 116(12):5217–5219

    Article  Google Scholar 

  • Fagan WF, Lewis MA, Auger-Méthé M, Avgar T, Benhamou S, Breed G, LaDage L, Schlägel UE, Tang WW, Papastamatiou YP, Forester J, Mueller T (2013) Spatial memory and animal movement. Ecol Lett 16(10):1316–1329

    Article  Google Scholar 

  • Fagan WF, Gurarie E, Bewick S, Howard A, Cantrell RS, Cosner C (2017) Perceptual ranges, information gathering, and foraging success in dynamic landscapes. Am Nat 189(5):474–489

    Article  Google Scholar 

  • Foss-Grant AP (2017) Quantitative challenges in ecology: competition, migration, and social learning. Ph.d thesis, University of Maryland

  • Golledge R (1998) Wayfinding behavior: cognitive mapping and other spatial processes. Johns Hopkins University Press, Baltimore

    Google Scholar 

  • Gourley SA, Ruan S (2000) Dynamics of the diffusive Nicholson’s blowflies equation with distributed delay. Proc R Soc Edinburgh Sect A 130(6):1275–1291

    Article  MathSciNet  MATH  Google Scholar 

  • Gourley SA, So JW-H (2002) Dynamics of a food-limited population model incorporating nonlocal delays on a finite domain. J Math Biol 44(1):49–78

    Article  MathSciNet  MATH  Google Scholar 

  • Hillen T, Buttenschön A (2020) Nonlocal adhesion models for microorganisms on bounded domains. SIAM J Appl Math 80(1):382–401

    Article  MathSciNet  MATH  Google Scholar 

  • Hillen T, Painter KJ (2009) A user’s guide to PDE models for chemotaxis. J Math Biol 58(1–2):183–217

    Article  MathSciNet  MATH  Google Scholar 

  • Jin J, Shi J, Wei J, Yi F (2013) Bifurcations of patterned solutions in the diffusive Lengyel–Epstein system of CIMA chemical reactions. Rocky Mt J Math 43(5):1637–1674

    Article  MathSciNet  MATH  Google Scholar 

  • Kappeler P (2010) Animal behaviour: evolution and mechanisms. Springer, Berlin, Heidelberg

    Book  Google Scholar 

  • Kareiva P, Odell G (1987) Swarms of predators exhibit preytaxis if individual predators use area-restricted search. Am Nat 130(2):233–270

    Article  Google Scholar 

  • Keller EF, Segel LA (1970) Initiation of slime mold aggregation viewed as an instability. J Theoret Biol 26(3):399–415

    Article  MathSciNet  MATH  Google Scholar 

  • Kuto K, Osaki K, Sakurai T, Tsujikawa T (2012) Spatial pattern formation in a chemotaxis-diffusion-growth model. Phys D 241(19):1629–1639

    Article  MathSciNet  MATH  Google Scholar 

  • Lee JM, Hillen T, Lewis MA (2009) Pattern formation in prey-taxis systems. J Biol Dyn 3(6):551–573

    Article  MathSciNet  MATH  Google Scholar 

  • Lewis MA, Murray JD (1993) Modelling territoriality and wolf–deer interactions. Nature 366(6457):738–740

    Article  Google Scholar 

  • Liu P, Shi J (2018) Bifurcation of positive solutions to scalar reaction–diffusion equations with nonlinear boundary condition. J Differ Equ 264(1):425–454

    Article  MathSciNet  MATH  Google Scholar 

  • Liu P, Shi J, Wang Z (2013) Pattern formation of the attraction-repulsion Keller–Segel system. Discrete Contin Dyn Syst Ser B 18(10):2597–2625

    MathSciNet  MATH  Google Scholar 

  • Ma M, Wang Z (2015) Global bifurcation and stability of steady states for a reaction–diffusion-chemotaxis model with volume-filling effect. Nonlinearity 28(8):2639–2660

    Article  MathSciNet  MATH  Google Scholar 

  • Macdonald N (1987) Time lags in biological models: lecture notes in biomathematics, vol 27. Springer, Berlin

    Google Scholar 

  • Mimura M, Tsujikawa T (1996) Aggregating pattern dynamics in a chemotaxis model including growth. Physica A 230(3–4):499–543

    Article  Google Scholar 

  • Moorcroft PR, Lewis MA (2006) Mechanistic home range analysis. Princeton University Press, Princeton

    Google Scholar 

  • Moorcroft PR, Lewis MA, Crabtree RL (1999) Home range analysis using a mechanistic home range model. Ecology 80(5):1656–1665

    Article  Google Scholar 

  • Morales JM, Moorcroft PR, Matthiopoulos J, Frair JL, Kie JG, Powell RA, Merrill EH, Haydon DT (2010) Building the bridge between animal movement and population dynamics. Philos Trans R Soc B Biol Sci 365(1550):2289–2301

    Article  Google Scholar 

  • O’Keefe J, Nadel L (1978) The Hippocampus as a cognitive map. Oxford University Press, Oxford

    Google Scholar 

  • Painter KJ, Hillen T (2011) Spatio-temporal chaos in a chemotaxis model. Phys D 240(4–5):363–375

    Article  MATH  Google Scholar 

  • Potts JR, Lewis MA (2016) How memory of direct animal interactions can lead to territorial pattern formation. J R Soc Interface 13(118):20160059

    Article  Google Scholar 

  • Potts JR, Lewis MA (2019) Spatial memory and taxis-driven pattern formation in model ecosystems. Bull Math Biol 81(7):2725–2747

    Article  MathSciNet  MATH  Google Scholar 

  • Schlägel UE, Lewis MA (2014) Detecting effects of spatial memory and dynamic information on animal movement decisions. Methods Ecol Evol 5(11):1236–1246

    Article  Google Scholar 

  • Shi J (1999) Persistence and bifurcation of degenerate solutions. J Funct Anal 169(2):494–531

    Article  MathSciNet  MATH  Google Scholar 

  • Shi J, Wang C, Wang H (2019) Diffusive spatial movement with memory and maturation delays. Nonlinearity 32(9):3188–3208

    Article  MathSciNet  MATH  Google Scholar 

  • Shi J, Wang C, Wang H, Yan X (2020) Diffusive spatial movement with memory. J Dyn Differ Equ 32(2):979–1002

    Article  MathSciNet  MATH  Google Scholar 

  • Shi J, Wang X (2009) On global bifurcation for quasilinear elliptic systems on bounded domains. J Differ Equ 246(7):2788–2812

    Article  MathSciNet  MATH  Google Scholar 

  • Tao Y (2010) Global existence of classical solutions to a predator-prey model with nonlinear prey-taxis. Nonlinear Anal Real World Appl 11(3):2056–2064

    Article  MathSciNet  MATH  Google Scholar 

  • Tao Y (2013) Global dynamics in a higher-dimensional repulsion chemotaxis model with nonlinear sensitivity. Discrete Contin Dyn Syst Ser B 18(10):2705–2722

    MathSciNet  MATH  Google Scholar 

  • Tello JI, Winkler M (2007) A chemotaxis system with logistic source. Commun Partial Differ. Equ. 32(4–6):849–877

    Article  MathSciNet  MATH  Google Scholar 

  • Turing AM (1952) The chemical basis of morphogenesis. Philos Trans R Soc Lond Ser B 237(641):37–72

    Article  MathSciNet  MATH  Google Scholar 

  • van Schaik CP (2010) Social learning and culture in animals. Springer, Berlin, pp 623–653

    Google Scholar 

  • Wang Z, Zhao K (2013) Global dynamics and diffusion limit of a one-dimensional repulsive chemotaxis model. Commun Pure Appl Anal 12(6):3027–3046

    Article  MathSciNet  MATH  Google Scholar 

  • Winkler M (2010) Boundedness in the higher-dimensional parabolic-parabolic chemotaxis system with logistic source. Commun Partial Differ Equ 35(8):1516–1537

    Article  MathSciNet  MATH  Google Scholar 

  • Winkler M (2014a) Global asymptotic stability of constant equilibria in a fully parabolic chemotaxis system with strong logistic dampening. J Differ Equ 257(4):1056–1077

    Article  MathSciNet  MATH  Google Scholar 

  • Winkler M (2014b) How far can chemotactic cross-diffusion enforce exceeding carrying capacities? J Nonlinear Sci 24(5):809–855

    Article  MathSciNet  MATH  Google Scholar 

  • Winkler M (2017) Emergence of large population densities despite logistic growth restrictions in fully parabolic chemotaxis systems. Discrete Contin Dyn Syst Ser B 22(7):2777–2793

    MathSciNet  MATH  Google Scholar 

  • Wu S, Shi J, Wu B (2016) Global existence of solutions and uniform persistence of a diffusive predator-prey model with prey-taxis. J Differ Equ 260(7):5847–5874

    Article  MathSciNet  MATH  Google Scholar 

  • Zuo W, Shi J (2021) Existence and stability of steady-state solutions of reaction–diffusion equations with nonlocal delay effect. Z Angew Math Phys 72(2):43

    Article  MathSciNet  Google Scholar 

  • Zuo W, Song Y (2015) Stability and bifurcation analysis of a reaction–diffusion equation with spatio-temporal delay. J Math Anal Appl 430(1):243–261

    Article  MathSciNet  MATH  Google Scholar 

  • Zuo W, Song Y Stability and double-Hopf bifurcations of a Gause–Kolmogorov-type predator-prey system with indirect prey-taxis. J Dyn Differ Equ 1–41 (to appear)

Download references

Acknowledgements

We would like to thank the editor and three anonymous reviewers for helpful comments which improve the manuscript. This work was done when the first author visited Department of Mathematics, William & Mary during the academic year 2016–2018, and she would like to thank William & Mary for their support and kind hospitality.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Junping Shi.

Additional information

Publisher's Note

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

Partially supported by a Grant from China Scholarship Council, NSFC Grant-12001240, Natural Science Foundation of Jiangsu Province (No.BK20200589), US-NSF Grant DMS-1715651, NSERC Discovery Grant RGPIN-2020-03911 and Accelerator Grant RGPAS-2020-00090.

Appendix

Appendix

Proof of Theorem 3

We apply Theorem 1.7 of Crandall and Rabinowitz (1971) for the local bifurcation. Fixing \(d_1,\tau >0\), we define a nonlinear mapping \(F:{\mathbb R}^+\times X^2\rightarrow Y^2\) by

$$\begin{aligned} F(d_2,u,v)=\left( \begin{array}{cccc} &{} d_1\Delta u+d_2div(u\nabla v)+f(u)\\ &{} d_1\Delta v+\frac{1}{\tau }(u-v)\\ \end{array}\right) . \end{aligned}$$
(7.1)

It is clear that \(F(d_2,\theta ,\theta )=0\) for any \(d_2>0\). The Fréchet derivative of F with respect to (uv) is

$$\begin{aligned} \begin{aligned}&F_{(u,v)}\left( d_{2,n}^{w}, \theta ,\theta \right) [\varphi ,\psi ] =\left( \begin{array}{cc} &{}d_1\Delta \varphi +d_{2,n}^{w}\theta \Delta \psi +f'(\theta )\varphi \\ &{}d_1\Delta \psi +\frac{1}{\tau }(\varphi -\psi )\\ \end{array}\right) :=L[\varphi ,\psi ].\end{aligned} \end{aligned}$$
(7.2)

Step 1. First we determine the null space of L. From Lemma 4, we have \(D_n(d_{2,s}^w)=0\) so \(\mu =0\) is an eigenvalue of \(J_n^w\) defined in (3.3) thus also an eigenvalue of (3.2) and there exists \(q=(\varphi ,\psi )^T=(1,h_n)\phi _n\in \mathcal {N}(L)\). Moreover as \(T_n(d_{2,s}^w)>0\), \(\mu =0\) is a simple eigenvalue of \(J_n^w\); and since \(\lambda _n\) is a simple eigenvalue of (1.4), and \(d_{2,n}^w\ne d_{2,k}^w\) for any \(k\in {\mathbb N}\) and \(k\ne n\), then \(\mu =0\) is as simple eigenvalue of L and

$$\begin{aligned} \mathcal {N}(L)=\text {Span}\left\{ q=(1,h_n)\phi _n \right\} , \end{aligned}$$

with \(h_n=1/(d_1\lambda _n\tau +1)\), thus \(\dim \left( {\mathcal {N}(L)}\right) =1\).

Step 2. We next consider the range space \(\mathcal {R}(L)\) of L. We can verify that \(\mathcal {R}(L)\) is given by \(\{(f_1,f_2)\in Y^2: \langle q^*, (f_1,f_2)\rangle =0\}\) where \(q^*\in \mathcal {N}(L^*)\) and \(L^*\) is the adjoint operator of L and defined by

$$\begin{aligned} \begin{aligned}&L^*\left[ \varphi ,\psi \right] =\left( \begin{array}{cc} &{}d_1\Delta \varphi +f'(\theta )\varphi +\frac{1}{\tau }\psi \\ &{}d_1\Delta \psi +d_{2,n}^{w}\theta \Delta \varphi -\frac{1}{\tau }\psi \\ \end{array}\right) .\end{aligned} \end{aligned}$$
(7.3)

Since \(\mathcal {N}(L^*)=\text {Span}\left\{ q^*=\left( 1,r_n\right) ^T\phi _n\right\} \), where \(r_n=\tau (d_1\lambda _n-f'(\theta ))\). We obtain

$$\begin{aligned} \mathcal {R}(L)=\left\{ (f_1,f_2)\in Y^2:\ \int _{\Omega }\left( f_1+r_nf_2\right) \phi _ndx=0\right\} , \end{aligned}$$

and \(\text {codim}\left( \mathcal {R}(L)\right) =1\).

Step 3. We show that \(F_{d_2(u,v)}(d_{2,n}^w,\theta ,\theta )[q]\not \in \mathcal {R}(L)\). From (7.1), we have

$$\begin{aligned} F_{d_2(u,v)}\left( d_{2,n}^w,\theta ,\theta \right) [q]=(\theta h_n\Delta \phi _n,0)^T=(-\theta \lambda _n h_n \phi _n,0)^T. \end{aligned}$$
(7.4)

Since

$$\begin{aligned} \int _{\Omega }\left( -\theta \lambda _nh_n\phi _n+0\right) \phi _ndx=-\theta \lambda _nh_n\int _{\Omega }\phi _n^2dx<0, \end{aligned}$$

thus \(F_{d_2(u,v)}\left( d_{2,n}^w,\theta ,\theta \right) [q]\not \in \mathcal {R}(L)\) by the definition of \(\mathcal {R}(L)\). From Step 1, 2 and 3, now we can apply Theorem 1.7 of Crandall and Rabinowitz (1971) to obtain part (i).

Step 4. Now we consider the bifurcation direction and stability of the bifurcating solutions in \(\Gamma _n\). To obtain more detailed information of the bifurcation, we use one-dimensional domain \(\Omega =(0,l\pi )\). In this case, it is known that \(\phi _n=\cos (nx/l)\) and \(\lambda _n=n^2/l^2\), so that \(q=(1,h_n)^{T}\cos (nx/l)\). From Shi (1999), we have

$$\begin{aligned} d'_{2,n}(0)=-\dfrac{\langle l, F_{(u,v)(u,v)}(d_{2,n}^{w},\theta ,\theta )[q,q]\rangle }{2\langle l, F_{d_2(u,v)}(d_{2,n}^{w},\theta ,\theta )[q]\rangle }, \end{aligned}$$

where \(l\in Y\) satisfies \(\mathcal {N}(l)=\mathcal {R}(L)\) and can be calculated as

$$\begin{aligned} \langle l,(f_1,f_2)\rangle =\displaystyle \int _0^{l\pi }\left( f_1+r_nf_2\right) \cos \left( \frac{nx}{l}\right) dx. \end{aligned}$$

By (7.4) and the definition of l, we have

$$\begin{aligned} \left\langle l,\ F_{d_2(u,v)}\left( d_{2,n}^{w}, \theta ,\theta \right) [q]\right\rangle =-\lambda _nh_n\theta \int _{0}^{l\pi }\cos ^2\left( \frac{nx}{l}\right) dx=-\frac{\lambda _nh_n\theta l\pi }{2}. \end{aligned}$$

From (7.1), it can be obtained that

$$\begin{aligned} F_{(u,v)(u,v)}(d_2,u,v)[\varphi ,\psi ][\varphi ,\psi ]=\left( 2d_2\varphi '\psi '+2d_2\varphi \psi ''+f''(u)\varphi ^2,0\right) ^{T}. \end{aligned}$$
(7.5)

This implies that

$$\begin{aligned} \ F_{(u,v)(u,v)}\left( d_{2,n}^{w}, \theta ,\theta \right) [q,q]=\left( \frac{f''(\theta )}{2}+\left( \frac{f''(\theta )}{2}-2d_{2,n}^wh_n\lambda _n\right) \cos \left( \frac{2nx}{l}\right) ,0\right) ^{T}, \end{aligned}$$
(7.6)

and thus

$$\begin{aligned} \begin{aligned}&\left\langle l,\ F_{(u,v)(u,v)}\left( d_{2,n}^{w}, \theta ,\theta \right) [q,q]\right\rangle \\ =&\int _{0}^{l\pi }\left( \frac{f''(\theta )}{2}+\left( \frac{f''(\theta )}{2}-2d_{2,n}^w h_n\lambda _n\right) \cos \left( \frac{2nx}{l}\right) \right) \cos \left( \frac{nx}{l}\right) dx=0. \end{aligned} \end{aligned}$$

Therefore \(d'_{2,n}(0)=0\).

Next we calculate \(d''_{2,n}(0)\) to determine the bifurcation direction by modifying the calculation in Jin et al. (2013). From Shi (1999), \(d''_{2,n}(0)\) takes the form:

$$\begin{aligned} d''_{2,n}(0)=-\dfrac{\left\langle l,\ F_{(u,v)(u,v)(u,v)}\left( d_{2,n}^{w},\theta ,\theta \right) [q,q,q]\right\rangle +3\left\langle l,\ F_{(u,v)(u,v)}\left( d_{2,n}^{w},\theta ,\theta \right) [q,\Theta ]\right\rangle }{3\left\langle l,\ F_{d_2(u,v)}(d_{2,n}^{w},\theta ,\theta )[q]\right\rangle }, \end{aligned}$$

where \(\Theta =(\Theta _1,\Theta _2)\) is the unique solution of

$$\begin{aligned} F_{(u,v)(u,v)}\left( d_{2,n}^{w},\theta ,\theta \right) [q,q]+F_{(u,v)}\left( d_{2,n}^{w},\theta ,\theta \right) [\Theta ]=0. \end{aligned}$$
(7.7)

From (7.5), we have

$$\begin{aligned} F_{(u,v)(u,v)(u,v)}\left( d_2,u,v\right) [\varphi ,\psi ][\varphi ,\psi ][\varphi ,\psi ]=(f'''(u)\varphi ^3,0)^{T}, \end{aligned}$$

thus

$$\begin{aligned} \langle l,F_{(u,v)(u,v)(u,v)}\left( d_{2,n}^{w},\theta ,\theta \right) [q,q,q]\rangle =\int ^{l\pi }_{0}f'''(\theta )\cos ^4\left( \frac{nx}{l}\right) dx=\frac{3l\pi }{8}f'''(\theta ). \end{aligned}$$
(7.8)

In the following, we show the calculation of \(\left\langle l,\ F_{(u,v)(u,v)}\left( d_{2,n}^{w},\theta ,\theta \right) [q,\Theta ]\right\rangle \). By (7.6) and (7.7), we may assume \(\Theta =(\Theta _1,\Theta _2)\) has the following form

$$\begin{aligned} \Theta _1=\Theta _1^0+\Theta _1^2\cos \left( \frac{2nx}{l}\right) ,~\Theta _2=\Theta _2^0+\Theta _2^2\cos \left( \frac{2nx}{l}\right) , \end{aligned}$$
(7.9)

since \(F_{(u,v)(u,v)}\left( d_{2,n}^{w},\theta ,\theta \right) \) consists of only constant and \(\cos \left( \frac{2nx}{l}\right) \) terms. Substituting (7.9) into (7.7), we obtain

$$\begin{aligned} \begin{aligned}&\left( \begin{array}{cc} &{}-4d_1\lambda _n\Theta _1^2\cos \left( \frac{2nx}{l}\right) -4d_{2,n}^{w}\theta \lambda _n\Theta _2^2\cos \left( \frac{2nx}{l}\right) +f'(\theta )\left( \Theta _1^0+\Theta _1^2\cos \left( \frac{2nx}{l}\right) \right) \\ &{}-4d_1\lambda _n\Theta _2^2\cos \left( \frac{2nx}{l}\right) +\frac{1}{\tau }[\Theta _1^0-\Theta _2^0+(\Theta _1^2-\Theta _2^2)\cos \left( \frac{2nx}{l}\right) ] \end{array}\right) \\&=-\left( \frac{f''(\theta )}{2}+\left( \frac{f''(\theta )}{2}-2d_{2,n}^{w}\lambda _nh_n\right) \cos \left( \frac{2nx}{l}\right) ,0\right) ^{T}. \end{aligned} \end{aligned}$$
(7.10)

Form Eq. (7.10), we can solve \(\Theta \) as in (3.10). Thus, we obtain

$$\begin{aligned} \begin{aligned}&\left\langle l,\ F_{(u,v)(u,v)}\left( d_{2,n}^{w},\theta ,\theta \right) [q,\Theta ]\right\rangle \\ =&2d_{2,n}^{w}\lambda _n(\Theta _2^2+h_n\Theta _1^2)\int ^{l\pi }_0\sin \left( \frac{2nx}{l}\right) \sin \left( \frac{nx}{l}\right) \cos \left( \frac{nx}{l}\right) dx\\&+\left( f''(\theta )-d_{2,n}^{w}\lambda _nh_n\right) \Theta _1^0\int ^{l\pi }_0\cos ^2\left( \frac{nx}{l}\right) dx\\&+\left( f''(\theta )\Theta _1^2-d_{2,n}^{w}\lambda _nh_n\Theta _1^2-4d_{2,n}^w\lambda _n\Theta _2^2\right) \Theta _1^2\int ^{l\pi }_0\cos \left( \frac{2nx}{l}\right) \cos ^2\left( \frac{nx}{l}\right) dx\\ =&\frac{l\pi }{2}d_{2,n}^{w}(\Theta _2^2+h_n\Theta _1^2\lambda _n) +\frac{l\pi }{2}\left( f''(\theta )-d_{2,n}^{w}\lambda _nh_n\right) \Theta _1^0\\&+\frac{l\pi }{4}(f''(\theta )\Theta _1^2-4d_{2,n}^w\lambda _n\Theta _2^2-d_{2,n}^w\lambda _nh_n\Theta _1^2). \end{aligned} \end{aligned}$$

Using all above we obtain \(d_{2,n}''(0)\) in Eq. (3.9).

Step 5. By applying Corollary 1.13 and Theorem 1.16 of Crandall and Rabinowitz (1973) or Theorem 5.4 of Liu and Shi (2018), the stability of the bifurcating non-constant steady states can be determined by the sign of \(\mu (s)\) which satisfies

$$\begin{aligned} \lim _{s\rightarrow 0}\frac{-sd'_{2,n}(s)m'(d_{2,n}^{w})}{\mu (s)}=1, \end{aligned}$$
(7.11)

where \(m(d_2)\) and \(\mu (s)\) are the eigenvalues defined as

$$\begin{aligned} \begin{aligned} F_{(u,v)}(d_2,\theta ,\theta )[\varphi (d_2),\psi (d_2)]&=m(d_2)K[\varphi (d_2),\psi (d_2)],\;\;\\&\quad \text {for}~d_2\in (d_{2,n}^{w}-\epsilon ,d_{2,n}^{w}+\epsilon ),\\ F_{(u,v)}(d_{2,n}(s),U_n(s),V_n(s))[\Lambda (s),\Phi (s)]&=\mu (s)K[\Lambda (s),\Phi (s)],\;\;~\text {for}~s\in (-\delta ,\delta ), \end{aligned} \end{aligned}$$

with \(K:X\rightarrow Y\) is the inclusion map \(K(u)=u\), \(m\left( d_{2,n}^{w}\right) =\mu (0)=0\) and \(\left( \varphi \left( d_{2,n}^{w}\right) ,\psi \left( d_{2,n}^{w}\right) \right) =(\Lambda (0),\Phi (0))=(1,h_n)\cos \left( \frac{nx}{l}\right) \).

Now consider the bifurcation at \(d_2=d_{2,N}^w=d_2^*\). From Lemma 4, \((\theta ,\theta )\) is stable and \(m(d_2)<0\) when \(d_2>d_{2,N}^{w}\), and it is unstable and \(m(d_2)>0\) when \(d_2<d_{2,N}^{w}\). One can calculate that

$$\begin{aligned} m(d_2)=\frac{-T_N+\sqrt{T_N^2-4D_N}}{2}, \end{aligned}$$

where \(T_N,D_N\) are defined in (3.5), and this implies that \(m'(d_{2,N}^{w})=-\theta \lambda _N/(T_N\tau )<0\). If \(d''_{2,N}(0)<0\), then \(d'_{2,N}(s)<0\) for \(s\in (0,\delta )\) and \(d'_{2,N}(s)>0\) for \(s\in (-\delta ,0)\). Hence \(-sd'_2(s)m'(d_{2,N}^{w})<0\) for \(s\in (-\delta ,\delta )\backslash \{0\}\), and consequently \(\mu (s)<0\) by(7.11) and the bifurcating solutions are locally asymptotically stable. Similarly when \(d''_{2,N}(0)>0\), the bifurcating solutions are unstable. For any other bifurcation at \(d_2=d_{2,n}^w<d_2^*\), the trivial solution \((\theta ,\theta )\) is already unstable at the bifurcation point, hence all bifurcating solutions are also unstable.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Shi, Q., Shi, J. & Wang, H. Spatial movement with distributed memory. J. Math. Biol. 82, 33 (2021). https://doi.org/10.1007/s00285-021-01588-0

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s00285-021-01588-0

Keywords

Mathematics Subject Classification

Navigation