Skip to main content
Philosophical transactions. Series A, Mathematical, physical, and engineering sciences logoLink to Philosophical transactions. Series A, Mathematical, physical, and engineering sciences
. 2018 Nov 12;376(2135):20180110. doi: 10.1098/rsta.2018.0110

Stabilizing a homoclinic stripe

Theodore Kolokolnikov 1,, Michael Ward 2, Justin Tzou 3, Juncheng Wei 2
PMCID: PMC6232602  PMID: 30420550

Abstract

For a large class of reaction–diffusion systems with large diffusivity ratio, it is well known that a two-dimensional stripe (whose cross-section is a one-dimensional homoclinic spike) is unstable and breaks up into spots. Here, we study two effects that can stabilize such a homoclinic stripe. First, we consider the addition of anisotropy to the model. For the Schnakenberg model, we show that (an infinite) stripe can be stabilized if the fast-diffusing variable (substrate) is sufficiently anisotropic. Two types of instability thresholds are derived: zigzag (or bending) and break-up instabilities. The instability boundaries subdivide parameter space into three distinct zones: stable stripe, unstable stripe due to bending and unstable due to break-up instability. Numerical experiments indicate that the break-up instability is supercritical leading to a ‘spotted-stripe’ solution. Finally, we perform a similar analysis for the Klausmeier model of vegetation patterns on a steep hill, and examine transition from spots to stripes.

This article is part of the theme issue ‘Dissipative structures in matter out of equilibrium: from chemistry, photonics and biology (part 2)’.

Keywords: reaction–diffusion systems, pattern formation, stability of patterns

1. Introduction

Consider a typical reaction–diffusion model, such as the Schnakenberg model, which admits a one-dimensional spike solution corresponding to a homoclinic orbit of the underlying slow-diffusing variable. When such a one-dimensional spike is extended trivially to the second dimension, the result is what we shall call a ‘homoclinic stripe’. In common situations, it is well established that a ‘homoclinic stripe’ is unstable and breaks up into spots, unless a domain is very ‘thin’ [14]. This break-up is illustrated in figure 1 row B and figure 2 row 1. In this paper, we examine two situations that can potentially stabilize a homoclinic stripe: adding anisotripy or drift to the fast-diffusing variable. We show that a sufficient amount of anisotropy or drift can stabilize a homoclinic stripe of any length.

Figure 1.

Figure 1.

Top: Boundaries for zigzag and break-up instabilities in the db parameter plane with ε = 0.025, (x, y)∈ [ − 1, 1] × [0, 1], A = 1. Note a good agreement between asymptotics and numerics. Rows A–C show snapshots of full numerical simulations of (1.1) in different parameter regions. Parameters A are in the stable region. Parameters B: stripe exhibits a break-up instability. Parameter C: The stripe is stable with respect to break-up instability, but unstable with respect to a zigzag instability. As a result, the stripe starts to bend (note the slow time scale). Eventually, this is followed by a break-up of a bended stripe. (Online version in colour.)

Figure 2.

Figure 2.

Simulation of (5.1) with ε = 0.05, A = L = 1, c = 10, and with d as follows: (a) d = 1.8ε; (b) d = 1.3ε; (c) d = 1.2ε; (d) d = 1.2ε; (e) d = 1.3ε. Initial conditions are close to those shown in the first column.

For concreteness, we concentrate on the well-studied Schnakenberg model [57] and its variants, but we anticipate these techniques can be extended to other settings. We first examine the anisotropic version of Schankenberg model, using the following scaling in two dimensions:

1. 1.1

We will assume Neumann boundary conditions in the x-direction, with the stripe extending to infinity in the y-direction (Inline graphic), which is the most potentially destabilizing case anyway, although our results can be generalized easily to the case of finite stripe.

We further assume that εO(1) and Inline graphic, whereas A and L are O(1). In this case, u and v become weakly coupled and the one-dimensional spike solution to (1.1) is asymptotically close to a sech2-type profile.

Equations (1.1) model the following process: a fast-diffusing substrate v is consumed by a slowly diffusing activator u, which decays with time. The substrate is being pumped into the system at rate A. The reaction kinetics for u and v occur at different scales: u reacts much slower than v, so that v is effectively slave to u. Note that we scaled the system so that the activator u diffuses isotropically; the anisotropy is only present in the substrate (the case of anisotropic diffusion in u can be reduced to (1.1) by scaling the y coordinate appropriately).

Figure 1 illustrates our results. For sufficiently small b, the stripe is fully stable (row A). As b is increased, there are two types of instabilities that appear, depending on how large d is. These instabilities are triggered by either large (O(1)) or small (O(ε2)) eigenvalues, and are illustrated in rows B and C, respectively. We will refer to them as break-up and zigzag instabilities, respectively. As shown in figure 1, the two instability boundaries (indicated by red and blue curves) cross each-other. The crossing point is shown in green. The main result of this paper is to characterize this crossing analytically.

We now summarize the main result as follows. Consider the stripe solution, where u(x, y, t) is localized along a vertical line. Such a solution is shown row A (last snapshot) and has the asymptotic profile u(x, y, t)∼(LA/2) sech2(x/2ε) (see below for derivation). The stripe is stable provided that Inline graphic where b = b1 and b = b2 are thresholds for break-up and zigzag instabilities, respectively. The stripe is unstable when Inline graphic. Refer to figure 1 (top).

Suppose further that

1. 1.2

Then the asymptotics for stability boundaries b1 and b2 have explicit asymptoics

1. 1.3

and

1. 1.4

Break-up instabilities occur when b > b1. Zigzag instabilities happen when b > b2 (the formula (1.4) is actually valid for any Inline graphic whereas formula (1.3) requires further restriction Inline graphic).

As figure 1 shows, the two stability boundaries intersect. Let (dc, bc) be the point in (d, b) space at which b1 = b2. From (1.3), (1.4), this point is asymptotically given by

1. 1.5

If d > dc then break-up instabilities occur as b is increased above b1. If ε1/2d < dc, then zigzag instabilities occur as b is increased above b2.

Note that there is a relatively large error in (dc, bc) when ε = 0.05. This is due to O(ε1/3) scaling in (1.5). Table 1 shows the error behaviour as is halfed. It shows that the relative error does indeed decrease as Inline graphic

Table 1.

Error in dc and bc as a function of ε.

numerics
asymptotcs
rel. err
ε dc bc dc bc Inline graphic Inline graphic
0.05 0.2201 0.1375 0.2720 0.1861 19.1% 26%
0.025 0.18648 0.09255 0.2159 0.1172 13.6% 21%
0.0125 0.1536 0.06102 0.1714 0.0738 10.3% 17%

2. Stripe steady state

We begin by constructing a stripe steady state. Such a solution independent of y or t, so that (1.1) reduce to

2. 2.1

with Neumann boundary conditions ux( ± L) = vx( ± L) = 0. The asymptotic construction of a spike profile to (2.1) is by now very standard (e.g. [7,8]) but we briefly review it here for completeness and to settle the notation. We assume the spike has a maximum at x = 0. In the inner region, we scale u(x) = U(z), v(x) = V (z), which yields

2. 2.2

Assume that

2. 2.3

Then to leading order we have Vzz∼0 so that V is a constant: Vv0. Then to leading order, U satisfies Uw/v0 where w is the ground state satisfying

2. 2.4

with a well-known explicit solution given by Inline graphic. It remains to determine v0. To do so, we integrate the second equation in (2.1) which yields

2. 2.5

where we used the fact that Inline graphic In conclusion, we obtained

2. 2.6

3. Stability: large eigenvalues

We linearize the stripe using

3.

which results in the following one-dimensional eigenvalue problem,

3. 3.1a

and

3. 3.1b

with Neumann boundary conditions ϕ( ± L) = ψ( ± L) = 0.

In the inner region x = εz, to leading order, (3.1a) can be written as

3. 3.2

where

3. 3.3

is the local linear operator of the associated ground state w(z), and ψ0 = ψ(0). To determine ψ0, we approximate Inline graphic so that

3. 3.4

where G satisfies

3. 3.5

The solution to (3.5) is given by

3.

and, in particular,

3.

Upon plugging in x = 0 into (3.4) we obtain

3.

Now consider the critical scaling

3.

so that (3.2) becomes

3. 3.6

where we used Inline graphic Problem (3.6) is equivalent to solving

3. 3.7

where

3. 3.8

The function f(μ) can be expressed in terms of hypergeometric functions (e.g. [9]). However, the resulting expression is unwieldy and at the end still needs to be evaluated numerically. Here, we compute f(μ) directly by solving numerically the associated BVP (L0 − μ)ϕ = w2 and integrating numerically. Figure 3 shows the graph of f(μ).

Figure 3.

Figure 3.

f(μ) versus μ.

The vertical asymptote corresponds to the eigenvalue Inline graphic of the local operator L0, and asymptotic expansion near Inline graphic shows that f blows up to the left of this asymptote, in agreement with the graph above. Moreover, Inline graphic.

Simple sketching show that there is always an instability band when Inline graphic is sufficiently large, and there is full stability when Inline graphic is sufficiently small. The instability threshold happens when there is a double root for the equation

3. 3.9

Table 2 gives the double-root boundary for selected values of β.

Table 2.

Instability thresholds for large eigenvalues.

Inline graphic Inline graphic Inline graphic Inline graphic
0.05 30.416 1.5208 608.319
0.1 15.2277 1.52277 152.277
0.2 7.65315 1.53063 38.2658
0.5 3.16896 1.58448 6.33792
1 1.76 1.76 1.76
1.4 1.40648 1.96907 1.00463
2 1.17 2.35 0.588
4 0.998 4 0.249
8 0.975 7.8 0.121

Of particular interest are the following limits:

  • d small: In this limit, Inline graphic so that tanh in (3.9) can be replaced by 1, and the instability threshold corresponds to the double-root of the equation Inline graphic which numerically yields α∼0.975 as Inline graphic In the original variables, this yields the threshold (1.3).

  • d large: then Inline graphic and the instability threshold corresponds to the double-root of the equation Inline graphic which numerically yields αβ∼1.5208 as Inline graphic In the original variables, this yields the threshold
    graphic file with name rsta20180110-e27.jpg 3.10
    We will come back to (3.10) in the context of incorporating large drift, see below.

4. Small eigenvalues

Small eigenvalues arise from the translation invariance of the inner problem for the spike [8]. They correspond to an odd solution of the linearized equation (3.1). To leading order in ε, the eigenvalue is zero and a higher-order expansion is necessary to resolve the stability. Proceeding as in [3], we, therefore, expand both the steady state and the eigenfunction in the inner region using the stretched variable z = x/ε:

4.

The resulting equations are

4. 4.1
4. 4.2
4. 4.3

To leading order in ε, the solution to (4.2) is given by ΦUy, λ = 0. To resolve the next order, we start by multiplying (4.2) by Uz and integrating to obtain

4. 4.4

Estimating ΦUy and integrating the right-hand side by parts yields

4. 4.5

and using (4.1) we further estimate

4.

Finally, from (4.3) we estimate

4.

so that

4. 4.6

and (4.6) becomes

4. 4.7

Next we compute the outer solution to and match Ψz(∞). In the outer variables, we get

4.

Recall that ψ satisfies (3.1b). The last two terms decay exponentially and are taken are zero, so that

4.

To get C, multiply by x and integrate, assuming that Φ, Ψ is odd to get

4.

so that

4.

and finally

4. 4.8

Upon substituting (4.8) into (4.7) we obtain, after some algebra,

4. 4.9

where we used Inline graphic

Rewrite (4.9) as

4. 4.10

where

4. 4.11

From (4.10), there is a double root when α = 0.23716 corresponding to Inline graphic Instability exists when α > 0.23716, or b > b2, where b2 is given by (1.4).

This concludes the derivation of (1.3) and (1.4).

5. Vegetation patterns on a sloping ground

Finally, we extend this analysis to a variant of the Schnakenberg model which incorporates the water flowing down the hill. The equations are

5. 5.1a

and

5. 5.1b

(as in (1.1) and for the same reasons, we have replaced vt by zero in (5.1b) to simplify the exposition). For simplicity, we will assume periodic boundary conditions in the x-direction:

5.

with the stripe extending to infinity in the y-direction (Inline graphic). Mathematical models of vegetation patterns have a long history, starting with the work by Lefever and Lejeune on Tiger Bush patterns, [10]. The link between localized vegetation spots and localized structures was established in [11]. Vegetation spots can also be unstable through self-replication [12]. The particular model (5.1) that we chose is a variant of the Klausmeier vegetation model [13]. There, v represents water concentration in the soil whereas u is the plant density. In addition to the usual diffusion (assumed to be isotropic here for simplicity), the water flows down the slope. This is represented by the cvx term (positive c assumes the water flows from right to left). In fact equations (5.1) are special case of Klausmeier model when water evaporation is assumed small (water evaporation is modelled by adding −ev term to the right-hand side of (5.1b)).

There are numerous papers studying the transition between homogeneous vegetation, vegetation spots and vegetation stripes, as well as self-replication. For a very incomplete list of references, see for example [4,1422] and references therein. Generally speaking, spots transition to stripes, and stripes to uniform state, when precipitation rate A is increased.

Here, we show the transition of spots to stripes analytically in the regime of large slope (represented by parameter c) and small diffusion d. The stripes extend in the y-direction, so that they are perpendicular to the slope. The existence of stable stripes for large c was observed numerically in [4,20], matching observations in nature [23,24]. By contrast, in [4] the authors show analytically that a stripe is unstable (in particular when the slope is sufficiently small) and breaks up into spots.

Generally speaking, we find a spot-to-stripe transition as d is sufficiently decreased (and with c large). We restrict the analysis of (5.1) to the regime

5. 5.2

With this restriction, the vegetation density has a sech2-type profile in the x-direction, which we extend into a two-dimensional stripe. This makes it possible to determine the stability thresholds of the stripe analytically.

As shown in appendix A, with restriction (5.2), the cross-section of the stripe has the spike-type profile of the form (2.6), except that x is replaced by x − st, where s is the (slow) speed of motion. This speed is easily computed using solvability condition (as was already done in [25], or see appendix A) with the result that

5. 5.3

Repeating the analysis for large eigenvalue yields the problem (3.6) but with

5.

As a consequence, the analysis leading to (3.10) applies, and it follows that stripe is stable provided d < d1 where

5. 5.4

This threshold is illustrated in figure 2. In the first two rows, d > d1 and the stripe breaks up into spots, which then travel together forming a spotted-stripe pattern. In the second row, we took d = 1.3AL1/2ε. This is very close to the stable regime; the stripe still breaks up but not completely, indicating that the break-up instability is supercritical (i.e. reversible). For d = 1.2AL1/2ε (so that d = 0.97d1, the stable side of the threshold), the stripe is seen to be stable. Stripe stability appears to be very robust: even initial conditions consisting of a single spot evolve into a stable stripe or spotted stripe as shown in rows (d,e), depending on whether d < d1 or d > d1.

6. Discussion

For the Klausmeier model, we have shown that adding sufficient anisotropy can lead to stable stripes. Unlike the previous works [14] which show that only a stripe of very small O(ε) length can be stabilized in the isotropic case, here we show that the stripe of infinite length can become stable under sufficient anisotropy.

We also showed that the boundaries for zigzag and break-up instabilities cross when d = dc = O(ε1/3), as b is increased. This analysis has similarities to the analysis of double-Hopf point of a spike in the Gray–Scott model—see for example [2628]—although the resulting NLEP problem is has a somewhat different form, and it is enough to consider real solutions here (whereas full complex solutions arise when discussing a Hopf bifurcations). In these works, it is shown that a single spike in one dimension can become unstable due to either oscillations in position or in height, when the zero in the left-hand side in (1.1) is replaced by τvt and as τ is increased. These two instabilities also cross each-other when A = O(ε1/6) (with d = 1, b = 0). We remark that this hopf-hopf crossing has been recently extended to a two-dimensional spike in [29], where an unusual loglogε scaling for the crossing point is derived.

The observed patterns are closely connected to the concept of dissipative structures in out of equilibrium matter introduced by Ilya Prigogine and his collaborators [3033]. See also [34] for more recent special issue on dissipative pattern formation in general, and localized structures specifically.

An entirely different mechanism to stabilize a stripe is by adding saturation as described in [35]. Adding sufficient saturation can turn homoclinic (unstable) stripe into a heteroclinic (stable) stripe having a mesa-type profile consisting of back-to-back interfaces.

In the limit of large d (or equivalently, the limit of large c for the drift problem (5.1)), v only depends on y. We may integrate out the x-direction for (1.1) to obtain the following limiting system

6. 6.1

This is analogous to a so-called shadow limit, but in one direction only. In particular, the break-up instability threshold (3.10) can be derived directly from (6.1). It would be interesting to study spike dynamics for this system. The same system also arises in the limit of large drift Inline graphic for (5.1).

There are many open questions. For example, even when the stripe breaks up, the resulting spots can still align, forming a ‘spotted stripe’. As the anisotropy is further decreased, the spotted stripe itself undergoes secondary bifurcation, eventually resulting in a more uniform spot distribution. This is illustrated in figure 4. We plan to address this transition in future work.

Figure 4.

Figure 4.

Simulation of (1.1) on a square 2 × 2 domain with ε = 0.025, A = 1 and with d and b as indicated. For each value of (d, b), the eventual steady state is shown (at t = 107). As anisotropy is decreased, first, the stripe bifurates into a spotted stripe, then the spotted stripe breaks up resulting in a more uniform spot distribution.

Acknowledgments

T.K. would like to thank the hospitality of Wuhan University, where a part of this paper was written.

Appendix A. Speed of propagation

In this appendix, we derive the formula (5.3) for the speed of propagation of a stripe. Similar formula was previously derived in [25]. Assuming stripe moves in the x-direction only, equations (5.1) reduce to

graphic file with name rsta20180110-e54.jpg A 1

Assume that c scales like

graphic file with name rsta20180110-e55.jpg

and expand

graphic file with name rsta20180110-e56.jpg

Then expand

graphic file with name rsta20180110-e57.jpg

At the leading order, we obtain

graphic file with name rsta20180110-e58.jpg A 2

and the next order equations are

graphic file with name rsta20180110-e59.jpg A 3

Equation (A 2) yield

graphic file with name rsta20180110-e60.jpg

so that (A 3) simplifies to

graphic file with name rsta20180110-e61.jpg

Multiplying the equation for U1 by wz and integrating then yields

graphic file with name rsta20180110-e62.jpg

As before, to compute V0 we integrate the equation for v in (A 1) using periodic boundary conditions to obtain

graphic file with name rsta20180110-e63.jpg

so that

graphic file with name rsta20180110-e64.jpg

Finally, using Inline graphic we obtain (5.3).

Data accessibility

This article has no additional data.

Competing interests

We declare we have no competing interests.

Funding

T.K. and M.W. are supported by NSERC, Canada.

References

  • 1.Doelman A, van derPloeg H. 2002. Homoclinic stripe patterns. SIAM J. Appl. Dyn. Syst. 1, 65–104. ( 10.1137/S1111111101392831) [DOI] [Google Scholar]
  • 2.Morgan DS, Kaper TJ. 2004. Axisymmetric ring solutions of the 2D Gray–Scott model and their destabilization into spots. Physica D 192, 33–62. ( 10.1016/j.physd.2003.12.012) [DOI] [Google Scholar]
  • 3.Kolokolnikov T, Ward MJ, Wei J. 2006. Zigzag and breakup instabilities of stripes and rings in the two-dimensional Gray–Scott model. Stud. Appl. Math. 116, 35–95. ( 10.1111/j.1365-2966.2005.0333.x) [DOI] [Google Scholar]
  • 4.Sewalt L, Doelman A. 2017. Spatially periodic multipulse patterns in a generalized Klausmeier–Gray–Scott model. SIAM J. Appl. Dyn. Syst. 16, 1113–1163. ( 10.1137/16M1078756) [DOI] [Google Scholar]
  • 5.Schnakenberg J. 1979. Simple chemical reaction systems with limit cycle behaviour. J. Theor. Biol. 81, 389–400. ( 10.1016/0022-5193(79)90042-0) [DOI] [PubMed] [Google Scholar]
  • 6.Iron D, Wei J, Winter M. 2004. Stability analysis of Turing patterns generated by the Schnakenberg model. J. Math. Biol. 49, 358–390. ( 10.1007/s00285-003-0258-y) [DOI] [PubMed] [Google Scholar]
  • 7.Ward MJ, Wei J. 2002. The existence and stability of asymmetric spike patterns for the Schnakenberg model. Stud. Appl. Math. 109, 229–264. ( 10.1111/1467-9590.00223) [DOI] [Google Scholar]
  • 8.Iron D, Ward MJ, Wei J. 2001. The stability of spike solutions to the one-dimensional Gierer–Meinhardt model. Physica D 150, 25–62. ( 10.1016/S0167-2789(00)00206-2) [DOI] [Google Scholar]
  • 9.Kolokolnikov T, Wei J, Winter M. 2009. Existence and stability analysis of spiky solutions for the Gierer–Meinhardt system with large reaction rates. Physica D 238, 1695–1710. ( 10.1016/j.physd.2009.05.009) [DOI] [Google Scholar]
  • 10.Lefever R, Lejeune O. 1997. On the origin of tiger bush. Bull. Math. Biol. 59, 263–294. ( 10.1007/BF02462004) [DOI] [Google Scholar]
  • 11.Lejeune O, Tlidi M, Couteron P. 2002. Localized vegetation patches: a self-organized response to resource scarcity. Phys. Rev. E 66, 010901 ( 10.1103/PhysRevE.66.010901) [DOI] [PubMed] [Google Scholar]
  • 12.Bordeu I, Clerc MG, Couteron P, Lefever R, Tlidi M. 2016. Self-replication of localized vegetation patches in scarce environments. Sci. Rep. 6, 33703 ( 10.1038/srep33703) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13.Klausmeier CA. 1999. Regular and irregular patterns in semiarid vegetation. Science 284, 1826–1828. ( 10.1126/science.284.5421.1826) [DOI] [PubMed] [Google Scholar]
  • 14.Lejeune O, Tlidi M. 1999. A model for the explanation of vegetation stripes (tiger bush). J. Vegetation Sci. 10, 201–208. ( 10.2307/3237141) [DOI] [Google Scholar]
  • 15.von Hardenberg J, Meron E, Shachak M, Zarmi Y. 2001. Diversity of vegetation patterns and desertification. Phys. Rev. Lett. 87, 198101 ( 10.1103/PhysRevLett.87.198101) [DOI] [PubMed] [Google Scholar]
  • 16.Meron E, Gilad E, von Hardenberg J, Shachak M, Zarmi Y. 2004. Vegetation patterns along a rainfall gradient. Chaos Solitons Fractals 19, 367–376. ( 10.1016/S0960-0779(03)00049-3) [DOI] [Google Scholar]
  • 17.Couteron P, Anthelme F, Clerc M, Escaff D, Fernandez-Oto C, Tlidi M. 2014. Plant clonal morphologies and spatial patterns as self-organized responses to resource-limited environments. Phil. Trans. R. Soc. A 372, 20140102 ( 10.1098/rsta.2014.0102) [DOI] [PubMed] [Google Scholar]
  • 18.Siero E, Doelman A, Eppinga M, Rademacher JD, Rietkerk M, Siteur K. 2015. Striped pattern selection by advective reaction-diffusion systems: resilience of banded vegetation on slopes. Chaos Interdiscip. J. Nonlinear Sci. 25, 036411 ( 10.1063/1.4914450) [DOI] [PubMed] [Google Scholar]
  • 19.Gowda K, Chen Y, Iams S, Silber M. 2016. Assessing the robustness of spatial pattern sequences in a dryland vegetation model. Proc. R. Soc. A 472, 20150893 ( 10.1098/rspa.2015.0893) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20.Sherratt JA. 2016. When does colonisation of a semi-arid hillslope generate vegetation patterns? J. Math. Biol. 73, 199–226. ( 10.1007/s00285-015-0942-8) [DOI] [PubMed] [Google Scholar]
  • 21.Tzou J, Xie S, Kolokolnikov T, Ward MJ. 2017. The stability and slow dynamics of localized spot patterns for the 3-d Schnakenberg reaction-diffusion model. SIAM J. Appl. Dyn. Syst. 16, 294–336. ( 10.1137/16M108121X) [DOI] [Google Scholar]
  • 22.Kolokolnkov T, Wei J. 2018. Pattern formation in a reaction-diffusion system with space-dependent feed rate. SIAM Rev. 60, 626–645. ( 10.1137/17M1116027) [DOI] [Google Scholar]
  • 23.Deblauwe V, Couteron P, Lejeune O, Bogaert J, Barbier N. 2011. Environmental modulation of self-organized periodic vegetation patterns in Sudan. Ecography 34, 990–1001. ( 10.1111/j.1600-0587.2010.06694.x) [DOI] [Google Scholar]
  • 24.Deblauwe V, Couteron P, Bogaert J, Barbier N. 2012. Determinants and dynamics of banded vegetation pattern migration in arid climates. Ecol. Monogr. 82, 3–21. ( 10.1890/11-0362.1) [DOI] [Google Scholar]
  • 25.Sherratt JA. 2013. Pattern solutions of the Klausmeier model for banded vegetation in semiarid environments IV: slowly moving patterns and their stability. SIAM J. Appl. Math. 73, 330–350. ( 10.1137/120862648) [DOI] [Google Scholar]
  • 26.Doelman A, Kaper TJ, Eckhaus W. 2000. Slowly modulated two-pulse solutions in the Gray–Scott model I: asymptotic construction and stability. SIAM J. Appl. Math. 61, 1080–1102. ( 10.1137/S0036139999354923) [DOI] [Google Scholar]
  • 27.Muratov CB, Osipov V. 2002. Stability of the static spike autosolitons in the Gray–Scott model. SIAM J. Appl. Math. 62, 1463–1487. ( 10.1137/S0036139901384285) [DOI] [Google Scholar]
  • 28.Kolokolnikov T, Ward MJ, Wei J. 2005. The existence and stability of spike equilibria in the one-dimensional Gray–Scott model: the pulse-splitting regime. Physica D 202, 258–293. ( 10.1016/j.physd.2005.02.009) [DOI] [Google Scholar]
  • 29.Xie S, Kolokolnikov T. 2017. Moving and jumping spot in a two-dimensional reaction–diffusion model. Nonlinearity 30, 1536–1563. ( 10.1088/1361-6544/aa60b2) [DOI] [Google Scholar]
  • 30.Prigogine I, Nicolis G. 1967. On symmetry-breaking instabilities in dissipative systems. J. Chem. Phys. 46, 3542–3550. ( 10.1063/1.1841255) [DOI] [Google Scholar]
  • 31.Prigogine I, Lefever R. 1968. Symmetry breaking instabilities in dissipative systems. II. J. Chem. Phys. 48, 1695–1700. ( 10.1063/1.1668896) [DOI] [Google Scholar]
  • 32.Prigogine I, Nicolis G. 1971. Biological order, structure and instabilities. Quat. Rev. Biophys. 4, 107–148. ( 10.1017/S0033583500000615) [DOI] [PubMed] [Google Scholar]
  • 33.Glansdorff P, Prigogine I. 1973. Thermodynamic theory of structure, stability and fluctuations. Am. J. Phys. 41, 147–148. ( 10.1119/1.1987158) [DOI] [Google Scholar]
  • 34.Tlidi M, Taki M, Kolokolnikov T. 2007. Introduction: dissipative localized structures in extended systems. Chaos 17, 037101. [DOI] [PubMed] [Google Scholar]
  • 35.Kolokolnikov T, Sun W, Ward M, Wei J. 2006. The stability of a stripe for the Gierer–Meinhardt model and the effect of saturation. SIAM J. Appl. Dyn. Syst. 5, 313–363. ( 10.1137/050635080) [DOI] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Data Availability Statement

This article has no additional data.


Articles from Philosophical transactions. Series A, Mathematical, physical, and engineering sciences are provided here courtesy of The Royal Society

RESOURCES