Skip to main content
Springer logoLink to Springer
. 2017 Sep 22;56(5):142. doi: 10.1007/s00526-017-1233-6

Stable spike clusters for the precursor Gierer–Meinhardt system in R2

Juncheng Wei 1, Matthias Winter 2,, Wen Yang 3
PMCID: PMC6961491  PMID: 32009743

Abstract

We consider the Gierer–Meinhardt system with small inhibitor diffusivity, very small activator diffusivity and a precursor inhomogeneity. For any given positive integer k we construct a spike cluster consisting of k spikes which all approach the same nondegenerate local minimum point of the precursor inhomogeneity. We show that this spike cluster can be linearly stable. In particular, we show the existence of spike clusters for spikes located at the vertices of a polygon with or without centre. Further, the cluster without centre is stable for up to three spikes, whereas the cluster with centre is stable for up to six spikes. The main idea underpinning these stable spike clusters is the following: due to the small inhibitor diffusivity the interaction between spikes is repulsive, and the spikes are attracted towards the local minimum point of the precursor inhomogeneity. Combining these two effects can lead to an equilibrium of spike positions within the cluster such that the cluster is linearly stable.

Mathematics Subject Classification: 35B35, 92C15, 35B40, 35B25

Introduction

In 1952, Turing [16] studied how pattern formation could start from a state without patterns. He explained the onset of pattern formation by a combination of two properties of the system:

  • (i)

    presence of spatially varying instabilities

  • (ii)

    absence of spatially homogeneous instabilities.

Since Turing’s pioneering work many models have been proposed and studied to explore the so-called Turing diffusion-driven instability in reaction–diffusion systems to understand biological pattern formation. One of the most popular of these models is the Gierer–Meinhardt system [5, 12, 13], which in two dimensions can be stated as follows:

At=ε2ΔA-μA+A2H,inΩ,τHt=DΔH-H+A2,inΩ,Aν=Hν=0,onΩ. 1.1

We assume that the diffusivities ε and D are small positive constants satisfying 0<ε2D1logDε1 and τ is a nonnegative constant which is independent of ε. Further, ΩR2 is a smooth bounded domain and ν denotes the outward normal derivative at a point on its boundary Ω. In this paper we assume that Ω=BR is a disk around the origin with radius R. For the standard Gierer–Meinhardt system it is assumed that μ(y)1. In this study we consider a precursor inhomogeneity μ(|y|) which is a positive, rotationally symmetric and sufficiently smooth function of the spatial variable y defined in the domain Ω.

The main idea underpinning these stable spike clusters is the following: due to the small inhibitor diffusivity the interaction between spikes is repulsive and the spikes are attracted towards the local minimum point of the precursor inhomogeneity. Combining these two effects can lead to an equilibrium of spike positions within the cluster such that the cluster is linearly stable. The repulsive nature of spikes has been shown in [4]. The attracting feature of the local minimum of a precursor has been established in [17].

Problem (1.1) without precursor has been studied by numerous authors. For the one-dimensional case in a bounded interval (-1,1) with Neumann boundary conditions, the existence of symmetric N-peaked solutions (i.e. spikes of the same amplitude in leading order) was first established by Takagi [15]. The existence of asymmetric N-spikes was first shown by Ward and Wei [18] and Doelman et al. [3] independently. For symmetric N-peaked solutions, Iron et al. [8] studied the stability by using matched asymptotic expansions while Ward and Wei [18] later studied the stability for asymmetric N-spikes. Existence and stability for symmetric spikes in one spatial dimension was then established rigorously by the first two authors [23].

For the Gierer–Meinhardt system in two dimensions, the first two authors rigourously proved the existence and stability of multiple peaked patterns for the Gierer–Meinhardt system in the weak coupling case and the strong coupling case for symmetric spikes [2022]. Here we say that the system is in the weak coupling case if D as ε0 and in the strong coupling case if the parameter D is a finite constant independent of ε. For more results and background on the Gierer–Meinhardt system, we refer to [26] and the references therein.

In fact, already in the original Gierer–Meinhardt system [5, 12, 13], the authors have introduced precursor inhomogeneities. These precursors were proposed to model the localisation of the head structure in the coelenterate Hydra. Gradients have also been used in the Brusselator model to restrict pattern formation to some fraction of the spatial domain [7]. In this example, the gradient carries the system in and out of the pattern-forming region of the parameter range (for example across the Turing bifurcation). Thus it restricts the domain where peak formation can occur. A similar localisation effect has been used to model segmentation patterns for the fruit fly Drosophila melanogaster in [6, 11].

In [24] the existence and stability of N-peaked steady states for the Gierer–Meinhardt system with precursor inhomogeneity has been explored. The spikes in these patterns can vary in amplitude and have irregular spacing. In particular, the results imply that a precursor inhomogeneity can induce instability. Single-spike solutions for the Gierer–Meinhardt system with precursor including spike dynamics have been studied in [17].

Recently, the first two authors in [27] studied the Gierer–Meinhardt system with precursor in one spatial dimension and proved the existence and stability of a cluster, which consists of N spikes approaching the same limiting point. More precisely, they consider the existence of a steady-state spike cluster consisting of N spikes near a nondegenerate local minimum point y0 of the inhomogeneity μ(y), i.e., μ(y0)=0, μ(y0)>0, where y(-1,1),y0(-1,1). Further, they show that this solution is linearly stable.

Now we consider this problem in two dimensions. We shall study the existence and stability of positive k-peaked steady-state spike clusters to (1.1). For simplicity, we shall study the steady-state problem for positive solutions of (1.1) in the disk BR around the origin with radius R, which can be stated as follows:

ε2ΔA-μ(|y|)A+A2H=0,inBR,DΔH-H+A2=0,inBR,Aν=Hν=0,onBR, 1.2

where ν denotes the outward normal derivative at a point on BR.

Inspired by the work [1], where the authors constructed multi-bump ground-state solutions and the centres of these bumps are located at the vertices of a regular polygon, while each bump resembles, up to translation and multiplication with a constant amplitude, the unique radially symmetric solution of

Δw-w+w2=0inR2,0<w(x)0as|x|, 1.3

in this paper we shall prove the existence and stability of a spike cluster located near a nondegenerate minimum point of the precursor such that the positions of the spikes form a regular polygon. We note that the presence of such patterned steady state configurations appears driven by the smallness of the relative size σ2=ε2/D of the diffusion rates of the activating and inhibiting substances. However, there is some difference between our problem and the one considered in [1]. Here, we also need to take into consideration the precursor μ(|y|) and further assume that the inhibitor diffusivity D is very small. After introducing the transformation

y=εx,A^(x)=1ξε2DA(εx),H^(x)=1ξε2DH(εx),

and dropping hats, Eq. (1.2) becomes,

ΔA-μ(|εx|)A+A2H=0,inBR/ε,ΔH-σ2H+ξA2=0,inBR/ε,Aν=Hν=0,onBR/ε, 1.4

where the explicit definition of ξ will be given in (3.6).

Our first result on the existence of k-spike clusters is the following:

Theorem 1.1

(Existence of k-spike clusters). Let k2 be a positive integer. We assume μ=μ(|y|)C3(BR) be a positive, radially symmetric function and μ(0)=1,μ(0)=0,μ(0)>0, where μ denotes the radial derivative. Then, for

maxεD,DlogDε0,

problem (1.2) has a k-spike cluster solution which concentrates at 0. In particular, it satisfies

Aε(y)i=1kξDε2wyε-qi,Hε(qi)ξDε2, 1.5

where ξ is given in (3.6) and q1,,qk are the vertices of a k regular polygon. Further, εqi0,i=1,,k.

Remark 1

Here the assumption on the value of μ(0)=1 is introduced to make the computation and representation convenient. Without loss of generality we can always apply some scaling transformation for the solution (Aε,Hε) to achieve the assumption μ(0)=1.

Remark 2

The limit

maxεD,DlogDε0

is equivalent to the two simultaneous limits

εD0

and

DlogDε0.

The first limit

εD0

means that the diffusivity of the activator A is asymptotically of a higher order than the diffusivity of the inhibitor H. If this is not satisfied the pattern observed will no longer have a spike profile.

For the second limit

DlogDε0

to hold it is necessary that D0. (In fact, this is the condition which occurs in the case of one spatial dimension.) Here, for two spatial dimensions there is a second factor logDε which tends to infinity due to the first limit. The second factor appears due to the logarithmic singularity of the Green’s function, see (3.4). The second limit is exactly the condition which guarantees that the spikes form a cluster, i.e. their distances tend to zero, see (3.1), (3.2).

Remark 3

The exact scaling of qi follows from the balancing condition (4.11) with the radius Rε defined in (3.2). In this balancing condition there are two contributions: the first comes from the interactions of neighbouring spikes, the second stems from the precursor inhomogeneity.

Key steps in the proof of Theorem 1.1:

Proof

Step 1. In Sect. 3 we use an ansatz of an approximate spike cluster solution and consider the linearised operator around this ansatz. We compute the remainder when this ansatz is plugged into the Gierer–Meinhardt system. Then we introduce the linearised operator around this ansatz. Key results for Liapunov–Schmidt reduction will proved in appendix A.

Step 2.

In appendix A key results for the method of Liapunov–Schmidt reduction are proved. It is shown that this linearised operator as well as the conjugate operator are uniformly invertible modulo the kernel and cokernel consisting of the translation modes. Then it is shown that the fully nonlinear system has a solution modulo the kernel and cokernel. This implies that the existence problem can be reduced to a finite-dimensional problem.

Step 3.

In Sect. 4 the reduced problem is solved.

Next we state our second result which concerns the stability of the k-spike cluster steady states given in Theorem 1.1. Whereas the existence result is valid for any number k2 of spikes, the stability or instability of spike clusters depends on k.

Theorem 1.2

(Stability of k-spike clusters). For maxεD,DlogDε sufficiently small, let Aε,Hε be the k-spike cluster steady state given in Theorem 1.1. Then there exists τ0>0 which is independent of ε and D but may depend on k such that the k-spike cluster steady state (Aε,Hε) is linearly stable for 2k3 and unstable for k5 provided that 0τ<τ0.

The main existence and stability results can be extended to a cluster with k spikes located on a regular polygon plus an extra spike in its centre.

Proposition 1.3

For any k2 there also exists a (k+1)-spike cluster steady state similar to the one given in Theorem 1.1 but with an extra spike in the centre of the regular polygon. This (k+1)-spike cluster is linearly stable if k5 provided that 0τ<τ0 for some τ0>0.

Remark 4

These results suggest that placing a spike in the centre of a polygon can stabilise the spike cluster in the following sense: by putting a spike in the centre of the polygon it is possible to get a stable polygonal spike cluster containing six spikes but without a centre the number of spikes for a stable cluster cannot be five or more.

Remark 5

The stability of a spike cluster with four spikes on a regular polygon remains open. The stability problem in this case requires further expansion of an eigenvalue which is zero in the two leading orders. One mode which has to be resolved is that of two opposite spikes moving towards the centre and the other two spikes moving away from the centre without changing the distance between neighbours.

Remark 6

One instability of a spike cluster with five or more spikes on a regular polygon comes from a mode such that the spikes tend to move away from their original positions on a circle, some to the inside and some to the outside, since this increases the distances between neighbouring spikes.

Remark 7

The stability of a spike cluster with spikes on a regular polygon with six vertices plus its centre remains open. The stability problem in this case requires some new analysis since each spike now has three neighbours which have the same smallest distance (two spikes on the circle plus the spike in the centre).

Remark 8

One instability of a spike cluster with seven or more spikes on a regular polygon with centre comes from a mode such that the spikes tend to move away from their original positions on a circle, some to the inside and some to the outside, since this increases the distances between neighbouring spikes.

Key steps in the proof of Theorem 1.2:

Proof

To study the stability of a k-spike cluster we have to study large eigenvalues of order O(1) and small eigenvalues of order o(1) separately.

  • Step 1.

    In Sect. 5 we consider eigenvalues of order O(1). We show, using the results of Sect. 2, that they all have negative real part. This part applies to all k2.

  • Step 2.

    In appendix B we consider small eigenvalues and reduce the stability problem to a finite-dimensional problem. This part applies to all k2.

  • Step 3.

    In Sect. 6 we study this finite-dimensional problem. The computations depend on the number k of spikes in an essential way. At the end of Sect. 6 the spectral properties for different values of k are studied separately.

Key steps in the proof of Proposition 1.3:

Proof

Step 1. The same as Step 1 in the proof of Theorem 1.2.

Step 2.

The same as Step 2 in the proof of Theorem 1.2.

Step 3.

In Sect. 7 we study the finite-dimensional problem. The computations depend on the number k of spikes in an essential way. At the end of Sect. 7 the spectral properties for different values of k are studied separately.

We confirm and illustrate the main results by a few numerical computations (Figs. 1,2,3).

Fig. 1.

Fig. 1

Clustered spiky steady state of (1.2) for ε2=0.001,D=0.01,μ(|y|)=1+5|y|2. Shown is a 3-spike cluster consisting of 3 spikes on a regular polygon. The activator A is displayed in two three-dimensional surface plots from different perspectives in the top two graphs and its projection to the domain plane is shown in the bottom graph

Fig. 2.

Fig. 2

Clustered spiky steady state of (1.2) for ε2=0.001,D=0.01,μ(|y|)=1+5|y|2. Shown is a 2-spike cluster consisting of 2 spikes on a regular polygon. The activator A is displayed in two three-dimensional surface plots from different perspectives in the top two graphs and its projection to the domain plane is shown in the bottom graph

Fig. 3.

Fig. 3

Clustered spiky steady state of (1.2) for ε2=0.0001,D=0.002,μ(|y|)=1+5|y|2. Because smaller diffusivities are chosen, we now get more spikes than in Figs. 1 and 2. Shown is a (4+1)-spike cluster consisting of 4 spikes on a regular polygon plus a spike in the centre. The activator A is displayed in two three-dimensional surface plots from different perspectives in the top two graphs and its projection to the domain plane is shown in the bottom graph

This paper is organised as follows. In Sect. 2 we present some preliminaries on the spectral properties of the nonlocal linear operators which will appear in the existence proof and in the stability proof for the analysis of large eigenvalues of order O(1). We study the existence of a k-spike cluster solution to (1.2) in Sect. 3 and appendix A (Liapunov–Schmidt reduction) and Sect. 4 (solving the reduced problem); in appendix A we prove some key results for Liapunov–Schmidt reduction which are needed in Sect. 3. In Sect. 5 we rigourously study the large eigenvalues of order O(1) for the linearised problem around the steady state spike cluster. The small eigenvalues of order o(1) are investigated in appendix C (general theory) and Sect. 6 (explicit computation of small eigenvalues which decide the stability of spike clusters). In Sect. 7 we sketch how the approach can be adapted to show existence and stability of a cluster for which the spikes are located at the vertices of a regular polygon with centre.

Note that in Sect. 6 the number k of spikes plays an explicit role and different values of k are considered separately. In the same way, in Sect. 7 the number k+1 of spikes plays a role and the treatment depends on the value of k. In particular, the different cases of k are studied towards end of Sects. 6 and 7, respectively. The other sections of the paper apply to all values of k2 in the same way.

Throughout the paper, by Cc we denote generic constants which may change from line to line. Further, h.o.t. stands for higher order terms.

Preliminaries: spectral properties of some eigenvalue problems

In this section we shall provide some preliminaries which will be needed for the existence and stability proofs.

Let w be the ground state solution given in (1.3), i.e, the unique solution of the problem

Δw-w+w2=0,yR2,w>0,w(0)=maxxR2w(x),w(x)0as|x|0. 2.1

Let

L0ϕ=Δϕ-ϕ+2wϕ,ϕH2(R2), 2.2

where

H2(R2)=uL2(R2):uxiL2(R2),2uxixjL2(R2),i,j=1,2

and L2(R2) is the space of all square integrable functions defined on R2.

We first recall the following well known result:

Lemma 2.1

The eigenvalue problem

L0ϕ=λϕ,ϕH2(R2), 2.3

admits the following set of eigenvalues

λ1>0,λ2=λ3=0,λ4<0,. 2.4

The eigenfunction Φ0 corresponding to λ1 can be made positive and radially symmetric, the space of eigenfunctions corresponding to the eigenvalue 0 is

K0:=spanwxj,j=1,2.

Proof

This lemma follows from [10, Theorem 2.1] and [14, Lemma C].

Next we consider the following nonlocal eigenvalue problem:

Δϕ-ϕ+2wϕ-2R2wϕdxR2w2dxw2=α0ϕ,ϕH2(R2). 2.5

Problem (2.5) plays a key role in the study of large eigenvalues (See Sect. 5 below). For problem (2.5), we have the following theorem due to [19, Theorem 1.4].

Theorem 2.2

Let α00 be an eigenvalue of the problem (2.5). Then we have R(α0)-c1 for some c1>0.

We shall also consider the following system of nonlocal eigenvalue problems:

LΦ:=ΔΦ-Φ+2wΦ-2R2wΦdxR2w2dxw2, 2.6

where

Φ:=ϕ1,ϕ2,,ϕkT(H2(R2))k.

Then the conjugate operator of L under the scalar product in L2(R2) is given by

LΦ=ΔΦ-Φ+2wΦ-2R2w2ΦdxR2w2dxw. 2.7

We have the following result:

Lemma 2.3

We have

Ker(L)=K0K0K0,

and

Ker(L)=K0K0K0.

Proof

The system (2.6) is in diagonal form. Suppose

LΦ=0.

For i=1,2,,k the i-th equation of (2.6) is given by

Δϕi-ϕi+2wϕi-2R2wϕidxR2w2dxw2=0. 2.8

We claim that (2.8) admits only the solution wxi,i=1,2. Indeed, we note that ϕi=R2wϕidxR2w2dxw satisfies that

Δϕi-ϕi+2wϕi=R2wϕidxR2w2dxw2.

As a result, ϕi-2ϕi satisfies

Δϕi-2ϕi-ϕi-2ϕi+2wϕi-2ϕi=0,

and we get

ϕi-2ϕi=c1wx1+c2wx2 2.9

by Lemma 2.1. Multiplying by w on both sides of (2.9) and integrating, we have R2wϕidx=0. Hence, ϕi is the solution to

Δϕ-ϕ+2wϕ=0,

and we get the first conclusion of Lemma 2.3. To prove the second statement, we proceed in a similar way for L, and the i-th equation of (2.7) is given as

Δϕi-ϕi+2wϕi-2R2w2ϕidxR2w2dxw=0. 2.10

Multiplying (2.10) by w and integrating, we obtain R2w2ϕidx=0. Then we have

Δϕi-ϕi+2wϕi=0.

By Lemma 2.1 again, we get the second conclusion and the proof is finished.

By the result of Lemma 2.3, we have

Lemma 2.4

The operator

L:(H2(R2))k(L2(R2))k,LΦ=ΔΦ-Φ+2wΦ-2R2wΦdxR2w2dxw2,

is invertible if it is restricted as follows:

L:K0K0(H2(R2))kK0K0(L2(R2))k.

Moreover, L-1 is bounded.

Proof

This results follows from the Fredholm Alternative and Lemma 2.3.

Next we study the eigenvalue problem for L : 

LΦ=αΦ. 2.11

We have

Lemma 2.5

For any nonzero eigenvalue α of (2.11) we have R(α)-c<0.

Proof

Let (Φ,α) satisfy the system (2.11). Suppose R(α)0 and α0. The i-th equation of (2.11) becomes

Δϕi-ϕi+2wϕi-2R2wϕidxR2w2dxϕi=αϕi.

By Theorem 2.2, we conclude that

R(α)-c<0.

It is interesting to observe that for spike clusters the operator in the nonlocal eigenvalue problem (2.6) and its conjugate operator (2.7) take diagonal form. Thus they can be studied very easily by considering the scalar nonlocal eigenvalue problem which arises from the study of a single spike. In other words, for a spike cluster in the nonlocal eigenvalue problem each spike “feels” only itself and not the other spikes.

Existence I: reduction to finite dimensions

In this section, we shall reduce the existence problem to a finite-dimensional problem. In the first step, we choose a good approximation to an equilibrium state. Then we shall use Liapunov–Schmidt reduction to reduce the original problem to a finite-dimensional one. Some key results for Liapunov–Schmidt reduction are proved in appendix A. In the next section, we solve the reduced problem.

First of all, let us set the candidate points for the location of spikes to the activator. Let Qε denote the set of the vertex points of the regular k-polygon:

Qε=q=(q1,,qk)qi=2Rεcos2(i-1)πk,2Rεsin2(i-1)πk, 3.1

where Rε is chosen such that

1CDεlog1DlogDεRεCDεlog1DlogDε 3.2

for some constant C>1 independent of ε and D. If maxεD,DlogDε0, then we can see that εRε0 and DεRε0.

Recall that we want to solve (1.4) which is given by

ΔA-μ(|εx|)A+A2H=0,inBR/ε,ΔH-σ2H+ξA2=0,inBR/ε,Aν=Hν=0,onBR/ε,

where σ2=ε2D.

Next we introduce the cut-off function χε,qj(x)=χ(x-qjRεsinπk), where χ is a function which satisfies

χ(x)=1,|x|12,0,|x|>1,χC0(R2). 3.3

The gap between 12 and 1 is to be filled with an arbitrary function that bridges the two parts infinitely smoothly.

Let (q1,,qk) be defined as in (3.1) and we set

W=j=1kχε,qj(x)w(x-qj).

Let G(xz) be the Green function given by

ΔxG(x,z)-G(x,z)+δz(x)=0inBR/D,G(x,z)ν=0onBR/D.

We have

  1. If 0<|x-z|1, we have
    G(x,z)=12πlog1|x-z|+H(x,z), 3.4
    where H(xz) is a continuous function and xH(x,z)|x=z=0.
  2. If 1|x-z|R, we have
    G(x,z)=C|x-z|-12e-|x-z|1+o(1),|xG(x,z)|=G(x,z)(1+o(1)) 3.5
    for some generic constant C.

We write

ξ-1=BR/εG(σq1,σz)j=1kχ(εz)w(z-qj)2dz=12πlog1σR2w2dz+o(1). 3.6

For a function uH2(BR/ε), let T[u] be the unique solution in HN2(BR/ε) of the following problem:

ΔT[u]-σ2T[u]+ξu2=0inBR/ε, 3.7

where

HN2(BR/ε)=uH2(BR/ε)|uν=0onBR/ε.

Written differently, we have

T[u](x)=ξBR/εGσ(x,z)u2(z)dy, 3.8

where Gσ(x,z) is the Green function which satisfies (Δ-σ2)Gσ(x,z)+δz(x)=0 in BR/ε with Neumann boundary condition. Note that

Gσ(x,z)=G(σx,σz)forx,zBR/ε. 3.9

System (1.4) is equivalent to the following equation in operator form:

Sε(u,v)=S1(A,H)S2(A,H)=0,HN2(BR/ε)×HN2(BR/ε)L2(BR/ε)×L2(BR/ε), 3.10

where

S1(A,H)=ΔA-μ(|εx|)A+A2H:HN2(BR/ε)×HN2(BR/ε)L2(BR/ε),S2(A,H)=ΔH-σ2H+ξA2:HN2(BR/ε)×HN2(BR/ε)L2(BR/ε).

For Eq. (1.4), we choose our approximate solution as follows,

Aε,q=W,Hε,q=T[W]. 3.11

Note that Hε,q satisfies

0=ΔHε,q-σ2Hε,q+ξAε,q2=ΔHε,q-σ2Hε,q+ξj=1kw(x-qj)2+h.o.t..

Further, by our choice of ξ in (3.6), it is easy to see that Hε,q(qi)=1,i=1,,k. We insert our ansatz (3.11) into (1.4) and calculate

S2(Aε,q,Hε,q)=0, 3.12

and

S1(Aε,q,Hε,q)=ΔAε,q-μ(|εx|)Aε,q+Aε,q2Hε,q=j=1k[Δw(x-qj)-μ(|εx|)w(x-qj)]+j=1kw2(x-qj)Hε,q-1+h.o.t.=j=1k(1-μ(|εx|))w(x-qj)+j=1kw2(x-qj)(Hε,q-1-1)+h.o.t. 3.13

On the other hand, we calculate for j=1,,k and x=qj+z with qj=(qj,1,qj,2) in the range |σz|<δ for δ>0 small enough:

Hε,q(qj+z)-1=ξBR/ε(Gσ(qj+z,t)-Gσ(qj,t))Aε,q2dt=ξBR/ε(Gσ(qj+z,t)-Gσ(qj,t))w(t-qj)2dt+ξBR/ε(Gσ(qj+z,t)-Gσ(qj,t))l,ljw(t-ql)2dt+Oe-2Rεsinπk=ξR212πlog|t||z-t|w2(t)dt+ξl=12F(q)qj,lσzlR2w2(t)dt+ξl,m=122F(q)qj,lqj,mσ2zlzmR2w2(t)dt+Oe-2Rεsinπk+Oσ3|z|2+σ2Rσ-12e-Rσ|z|, 3.14

where Rσ=2σRεsinπk and

F(q)=i=1kH(σqi,σqi)+i,j,ijG(σqi,σqj). 3.15

Note that the error term in (3.14) consists of three parts. The first part is very small and comes from the difference between Aε,q and w due to the decay of the activator component between spikes and near the boundary. The second part estimates higher order terms in the expansion of F(q) around q. The third part estimates the smallness of the contribution of non-neighbouring spikes in F(q). The same three types of error terms appear repeatedly throughout the paper.

Substituting (3.14) into (3.13), we have the following key estimate,

Lemma 3.1

For x=qj+z,|σz|<δ and δ>0 small enough, we have

S1(Aε,q,Hε,q)=S1,1+S1,2, 3.16

where

S1,1(z)=ξHε,q(qj)-2R2w2(t)dtw2(z)(σz·qjF(q)+σ2l,m=12zlzm2F(q)qj,lqj,m+h.o.t.)+εl=12zlμ(0)+12μ(0)ε|qj|+O(ε2|qj|2)εqj,li=1kw(x-qi)+Oσ3|z|2+σ2Rσ-12e-Rσ|z|+e-2Rεsinπk, 3.17

and

S1,2(z)=ξw2(z)R(|z|)+ε2Rε2w(z), 3.18

where R(|z|) is a radially symmetric function with the property that

R(|z|)=O(log(1+|z|)).

Further, S1(Aε,q,Hε,q)=e-δσ for |x-qj|δσ,j=1,2,,k.

The above estimates will be very important in the following calculations, where (3.10) is solved modulo kernel and cokernel.

Now we study the linearised operator defined by

L~ε,q:=Sε,qAε,qHε,q,L~ε,q:HN2(BR/ε)×HN2(BR/ε)L2(BR/ε)×L2(BR/ε).

Set

Kε,q:=spanAε,qqj,lj=1,2,,k,l=1,2HN2(BR/ε),

and

Cε,q:=spanAε,qqj,lj=1,2,,k,l=1,2L2(BR/ε).

The operator L~ε,q is not uniformly invertible in σ and Dlog1σ due to its approximate kernel

Kε,q:=Kε,q{0}HN2(BR/ε)×HN2(BR/ε). 3.19

and approximate cokernel

Cε,q:=Cε,q{0}L2(BR/ε)×L2(BR/ε). 3.20

Then we define

Kε,q:=Kε,qHN2(BR/ε)HN2(BR/ε)×HN2(BR/ε), 3.21
Cε,q:=Cε,qL2(BR/ε)L2(BR/ε)×L2(BR/ε), 3.22

where Cε,q and Kε,q denote the orthogonal complement with the scalar product of L2(BR/ε) in the spaces HN2(BR/ε) and L2(BR/ε), respectively.

Let πε,q denote the projection in L2(BR/ε)×L2(BR/ε) onto Cε,q, where the second component of the projection is the identity map. We are going to show that the equation

πε,qSε,qAε,q+ϕε,qHε,q+ψε,q=0

has a unique solution Σε,q=ϕε,qψε,qKε,q if maxεD,DlogDε is small enough (Liapunov–Schmidt reduction).

Set

Lε,q=πε,qL~ε,q:Kε,qCε,q. 3.23

In appendix A we will show the following key results for Liapunov–Schmidt reduction:

  1. The linear operator Lε,q is uniformly invertible.

  2. There exists Σε,q=ϕε,qψε,qKε,q.

Then, in the next section, we will solve the reduced problem and determine the point qQε.

Existence II: the reduced problem

In this section, we solve the reduced problem and complete the proof of Theorem 1.1.

By Lemma 9.2, for each qQε, there exists a unique solution (Φε,q,Ψε,q)Kε,q such that

Sε,qAε,q+Φε,qHε,q+Ψε,p=Ξε,q0Cε,q.

Our idea is to find q such that Sε,qAε,q+Φε,qHε,q+Ψε,qCε,q. Let

Wε,j,i(q):=1ξBR/εS1(Aε,q+Φε,q,Hε,q+Ψε,q)Aε,qqj,idz,

where j=1,2,,k and i=1,2. We set

Wε(q)=Wε,1,1(q),,Wε,k,2(q).

It is easy to see that Wε(q) is a map which is continuous in q, and our problem is reduced to finding a zero of the vector field Wε(q). Since the points q1,q2,,qk are the vertices of a regular k-polygon and μ(ε|x|) is a radially symmetric function, if we can find qQε such that Wε,1,1(q),Wε,1,2(q)=0, then Wε(q)=0. Further, we note that the approximate solution (Aε,q,Hε,q) is invariant under rotation by 2πk. Recall that by (3.1), we have q1=(q1,1,0)=(2Rε,0). Thus, using [1, Corollary 7.1], Wε,1,2 equals 0. So, all that remains is finding q such that Wε,1,1(q)=0.

We calculate the asymptotic expansion of Wε,1,1(q),

BR/εS1(Aε,q+Φε,q,Hε,q+Ψε,q)Aε,qq1,1dz=BR/εΔ(Aε,q+Φε,q)-μ(Aε,q+Φε,q)+(Aε,q+Φε,q)2Hε,q+Ψε,qAε,qq1,1dz=BR/εΔ(Aε,q+Φε,q)-(Aε,q+Φε,q)+(Aε,q+Φε,q)2Hε,qAε,qq1,1dz+BR/ε(Aε,q+Φε,q)2Hε,q+Ψε,q-(Aε,q+Φε,q)2Hε,qAε,qq1,1dz+BR/ε(1-μ)(Aε,q+Φε,q)Aε,qq1,1dz=I1+I2+I3,

where Ii,i=1,2,3 are defined at the last equality.

For I1, we have by Lemma 9.2,

I1=BR/εΔ(Aε,q+Φε,q)-(Aε,q+Φε,q)+(Aε,q+Φε,q)2Hε,q(q1)Aε,qq1,1dz-BR/ε(Aε,q+Φε,q)2Hε,q2(q1)(Hε,q-Hε,q(q1))Aε,qq1,1dz+Oe-2Rεsinπk=-BR/εΔ(w1+Φε,q)-(w1+Φε,q)+(w1+Φε,q)2Hε,q(q1)w1z1dz+BR/ε(w1+Φε,q)2Hε2(q1)2(Hε,q(q1+z)-Hε,q(q1))w1z1dz+Oe-2Rεsinπk, 4.1

where w1(z)=w(q1+z). Note that by Lemma 9.2, we have Φε,q,2 is radially symmetric with respect to z. Then we have

BR/ε[ΔΦε,q-Φε,q+2w1Φε,q]w1z1dz=BR/εΦε,q,1z1[Δw-w+w2]dz=0 4.2

and

BR/ε(Φε,q)2w1z1dz=BR/εΦε,q,1Φε,q,2w1z1dz=Oσlog1σ-2Rσ-12e-Rσ+σ2log1σ-2+log1σ-1ε2Rε. 4.3

From (4.1)–(4.3), we get

I1=BR/εw12(Hε,q(q1+z)-Hε,q(q1))w1z1dz+h.o.t.=ξσk=12F(q)q1,kR2w2zkwz1dzR2w2dz+h.o.t.=-c1ξσF(q)q1,1+h.o.t., 4.4

where F(q) is defined in (3.15), c1=13R2w2dzR2w3dz and h.o.t. represent terms of the order

σlog1σ-2Rσ-12e-Rσ+σ2log1σ-2+log1σ-1ε2Rε.

Next we study the term I2. We recall that Ψε,q satisfies the following equation

ΔΨε,q-σ2Ψε,q+2ξAε,qΦε,q+ξΦε,q2=0. 4.5

As for the perturbation term Φε,q, we can also make a decomposition for Ψε,q=Ψε,q,1+Ψε,q,2, where

ΔΨε,q,1-σ2Ψε,q,1+2ξAε,qΦε,q,1+ξΦε,q,12+2Φε,q,1Φε,q,2=0,

and

ΔΨε,q,2-σ2Ψε,q,2+2ξAε,qΦε,q,2+ξΦε,q,22=0.

Then we can easily see that

Ψε,q,1H2(BR/ε)=Oσlog1σ-1Rσ-12e-Rσ+σ2log1σ-1+ε2Rε

and Ψε,q,2 is radially symmetric with respect to z. Further, from the Green representation formula we get that

Ψε,q,1(q1+z)-Ψε,q(q1)=ξBR/ε(Gσ(p1,q1+z)-Gσ(p1,z))2Aε,qΦε,q+Φε,q2dz=o(1)ξσ|q1F(q)||z|+R1(|z|), 4.6

where R1(|z|) is a radially symmetric function.

Substituting (4.5) and (4.6) into I2, we get

I2=BR/ε(Aε,q+Φε,q)2Hε,q+Ψε,q-(Aε+Φε,q)2Hε,qAε,qq1,1dz=-BR/ε(Aε,q+Φε,q)2Hε,q2Ψε,qAε,qq1,1dz+Oσlog1σ-2Rσ-12e-Rσ+σ2log1σ-2+log1σ-1ε2Rε=-BR/ε13w13y1(Ψ-Ψ(q1))dz+Oσlog1σ-2Rσ-12e-Rσ+σ2log1σ-2+log1σ-1ε2Rε=o(1)ξσ|q1F(q)|+Oσlog1σ-2Rσ-12e-Rσ+σ2log1σ-2+log1σ-1ε2Rε. 4.7

For I3, we have

I3=BR/ε(1-μ)w(x-q1)w(x-q1)q1,1dx+Oe-2Rεsinπk=BR/ε(1-μ(|εq1|))w1w1q1,1dz+BR/ε(μ(|εq1|)-μ(|ε(q1+z)|))w1w1q1,1dz+Oe-2Rεsinπk=R2[μ(|εq1|)z1εz1+μ(|εq1|)z2εz2]w(z)w(z)z1dz+Oe-2Rεsinπk=εμ(|εq1|)z1R2z1wwz1dz+Oe-2Rεsinπk=c2ε2Rεμ(0)+Oe-2Rεsinπk+ε3Rε2, 4.8

where c2=2R2z1wwz1dz<0 is negative and μ(0) denotes the second radial derivative of the radially symmetric function μ at the origin.

From (4.1) to (4.8), we get that Wε,1,1(q) can be represented as follows:

Wε,1,1(q)=-c1ξσF(q)q1,1+c2ε2Rεμ(0)+h.o.t. 4.9

Using the asymptotic behaviour of the Green function Gσ, we have

Wε,1,1(q)=-c1ξσq1,1(Gσ(q1,q2)+Gσ(q1,qk))+c2ε2Rεμ(0)+h.o.t.=c1ξσRσ-12e-Rσq1-q2|q1-q2|+q1-qk|q1-qk|+c2ε2Rεμ(0)+h.o.t.=2c1ξσRσ-12e-Rσsinπk+c2ε2Rεμ(0)+h.o.t.=:W^ε,1,1(q)+h.o.t.. 4.10

Thus, denoting the leading order contribution of Wε,1,1(q) by W^ε,1,1(q), it follows that W^ε,1,1(q) depends only on Rε. Then we have W^ε,1,1(q)=0 for

ξ2σRε,0sinπk-32e-2σRε,0sinπk+c3D=h.o.t., 4.11

where c3=c2μ(0)4c1(sinπk)2 is negative, and we finally get

Rε,0=12σsinπklog1D-32loglog1D-logξc3+Ologlog1Dlog1D.

If ξ-1D is sufficiently small, then we easily get that Eq. (4.11) admits a unique solution and it is nondegenerate. As a consequence, in the neighborhood of Rε,0, we can find R^ε,0 such that Wε,1,1(q)=0. Thus, we have solved the reduced problem and the proof of Theorem 1.1 is complete.

Stability analysis I: study of large eigenvalues

To prove Theorem 1.2, we consider the stability of the solution (Aε,Hε) for (1.2) which was given in Theorem 1.1.

Linearizing the Gierer–Meinhardt system (1.1) around the equilibrium states (Aε,Hε), we obtain the following eigenvalue problem:

Δxϕε-μ(|εx|)ϕε+2AεHεϕε-Aε2Hε2ψε=λεϕε,Δxψε-σ2ψε+2ξAεϕε=τλεσ2ψε, 5.1

Here λε is some complex number and

ϕεHN2(BR/ε),ψεHN2(BR/ε).

In this section, we study the large eigenvalues, i.e., we assume that |λε|c>0 for ε small. The derivation of a matrix characterizing the small eigenvalues will be done in appendix B since this study is quite technical. Finally, in the next section, we discuss the small eigenvalues explicitly by considering these matrices. That part is central to understanding the stability of spike clusters.

If R(λε)-c, we are done. Then λε is a stable large eigenvalue. Therefore we may assume that R(λε)-c and for a subsequence in ε,D, we have λελ00. We shall derive the limiting eigenvalue problem of (5.1) as maxεD,DlogDε0 which reduces to a system of nonlocal eigenvalue problems.

The key references are Theorem 2.2 and Lemma 2.5.

The second equation in (5.1) is equivalent to

Δψε-σ2(1+τλε)ψε+2ξεAεϕε=0. 5.2

We introduce the following:

σλε=σ1+τλε,

where in 1+τλε we take the principal part of the square root. This means that the real part of 1+τλε is positive, which is possible because R(1+τλε)12.

Let us assume that ϕεH2(BR/ε)=1. We cut off ϕε as follows:

ϕε,j(x)=ϕε(x)χε,qj(x),

where the test function χε,qj(x) was introduced in (3.3).

From R(λε)-c and the exponential decay of w, we can derive from (5.1) that

ϕε=j=1kϕε,j+h.o.t.inHN2(BR/ε).

Since ϕεH2(BR/ε)=1, by taking a subsequence, we may also assume that ϕε,jϕj in H2(BR/ε) as maxεD,DlogDε0 for j=1,2,,k. We have by (5.2)

ψε(x)=ξBR/εGσλε(x,z)Aε(z)ϕε(z)dz. 5.3

At x=qi,i=1,2,,k, we calculate

ψε(qi)=ξBR/εGσλε(qi,z)j=1kwj(z)ϕε,j(z)dz+h.o.t.=12πξlog1σλεBR/εwϕdz+h.o.t. 5.4

Substituting the above equation in the first equation of (5.1), taking the limit maxεD,DlogDε0, we get

Δxϕi-ϕi+2wϕi-21+τλ0R2wϕidzR2w2dzw2=λ0ϕi,i=1,2,k, 5.5

where ϕiH2(R2). Then we have

Theorem 5.1

Let λε be an eigenvalue of (5.1) such that R(λε)>-c for some c>0.

  1. Suppose that (for suitable sequence maxεD,DlogDε0) we have λελ00. Then λ0 is an eigenvalue of the nonlocal eigenvalue problem given in (5.5).

  2. Let λ00 with R(λ0)>0 be an eigenvalue of the nonlocal eigenvalue problem given in (5.5). Then for maxεD,DlogDε small enough, there is an eigenvalue λε of (5.1) with λελ0 as maxεD,DlogDε0.

Proof

(1) of Theorem 5.1 follows by asymptotic analysis similar to the one obtained in appendix A.

To prove (2) of Theorem 5.1, we follow a compactness argument of Dancer. For the details we refer to Chapter 4 of [26].

We now study the stability of (5.1), by Lemma 2.5, for any nonzero eigenvalue λ0 in (5.5) we have

R(λ0)-c0<0forsomec0>0.

Thus, by Theorem 5.1, for maxεD,DlogDε small enough, all nonzero large eigenvalues of (5.1) have strictly negative real parts. More precisely, all eigenvalues λε of (5.1) for which λελ00 holds, satisfy R(λε)-c<0.

In conclusion, we have finished studying the large eigenvalues (of order O(1)) and derived results on their stability properties. It remains to study the small eigenvalues (of order o(1)) which will be done in appendix B and the next section.

Stability analysis III: study of the matrix Mμ(q)

In this section, we shall study the matrix Mμ(q) which was derived in appendix B. The eigenvalues of this matrix will determine the stability of small eigenvalues. Up to a constant positive factor, Mμ(q) is given by the hessian matrix of the term

(q)=i,j,ijξ1(σ|qi-qj|)12e-σ|qi-qj|+c3i=1kμ(|εqi|), 6.1

where c3=-c2c1 with c1,c2 given in (4.4) and (4.8) respectively.

Using

|q1-q2|=1σlog1D-32loglog1D-logξc3+Ologlog1Dlog1D, 6.2

(see (4.11)), it is not difficult to see that in leading order

2q1,iq1,j(q)ξσ32e-σ|q1-q2|1|q1-q2|52(q1-q2)i(q1-q2)j,+ξσ32e-σ|q1-qk|1|q1-qk|52(q1-qk)i(q1-qk)j,

where i,j=1,2. By rotational symmetry, it is enough to compute

2q1,iq1,j,2q1,iq2,j,2q1,iqk,j,i,j=1,2.

Terms which depend on qm are obtained by suitable rotation of terms which contain q1. By straightforward computation, we have

(q)q1,i=-ξσ12e-σ|q1-q2|(q1-q2)i|q1-q2|32-ξ12σ12e-σ|q1-q2|(q1-q2)i|q1-q2|52-ξσ12e-σ|q1-qk|(q1-qk)i|q1-qk|32-ξ12σ12e-σ|q1-qk|(q1-qk)i|q1-qk|52+h.o.t.,

and

2(q)q1,iq1,j=ξσ32e-σ|q1-q2|(q1-q2)i(q1-q2)j|q1-q2|52+ξσ32e-σ|q1-qk|(q1-qk)i(q1-qk)j|q1-q2|52+h.o.t.. 6.3

For the terms 2(q)q1,iq2,j,i,j=1,2, we note that

(q)q1,1=ξσ12e-σ|q2-q1|(q2-q1)1|q2-q1|32+ξ12σ12e-σ|q2-q1|(q2-q1)1|q2-q1|52+h.o.t.,

and

(q)q1,2=ξσ12e-σ|q2-q1|(q2-q1)2|q2-q1|32+ξ12σ12e-σ|q2-q1|(q2-q1)2|q2-q1|52+h.o.t..

Then we get

2(q)q1,1q2,i=-ξσ32e-σ|q2-q1|(q2-q1)1(q2-q1)i|q2-q1|52+h.o.t.,i=1,2, 6.4

and

2(q)q1,2q2,i=-ξσ32e-σ|q1-q2|(q2-q1)2(q2-q1)i|q1-q2|52+h.o.t.,i=1,2. 6.5

Similarly, for the terms 2(q)q1,iqk,j,i,j=1,2, we have

2(q)q1,1qk,i=-ξσ32e-σ|qk-q1|(qk-q1)1(qk-q1)i|qk-q1|52+h.o.t.,i=1,2, 6.6

and

2(q)q1,2qk,i=-ξσ32e-σ|qk-q1|(qk-q1)2(qk-q1)i|qk-q1|52+h.o.t.,i=1,2. 6.7

We now compute these expressions in a coordinate system of tangential and normal coordinates around each spike. We remark that these coordinates are the same as in [2]. The spike locations are given by

qj0=Rσσsinπkcosθj,Rσσsinπksinθj,j=1,,k, 6.8

where

θj=(j-1)2πk+α

and αR. Note that the phase shift α appears in the problem due to the rotational invariance of μ=μ(|y|) and we can choose α=0. Then in local coordinates we can write

qj=qj0+qj,1qj|qj|+qj,2qj|qj|,j=1,,k, 6.9

where qj is the radial (normal) vector and the tangential vector qj is obtained from qj by rotation of π/2 in anti-clockwise direction.

From (6.3) to (6.7), using the local coordinate frames and elementary trigonometry, the leading order of the matrix Mμ(q) is

Mμ(q)=ξσ32e-σ|q1-q2|1|q1-q2|52sinπk2(A1+4I)sinπkcosπkA2-sinπkcosπkA2-cosπk2A1+h.o.t., 6.10

where

A1=-210011-21001001-2andA2=0100-1-10100100-10.

Before analyzing the matrix in (6.10), we need some basic facts about circulant matrices. We follow the presentation in [3, 13] and include this material here for completeness. Denote the k-dimensional complex vector space and the ring of k×k complex matrices by Ck and Mk, respectively. Let b=(b1,b2,,bk)Ck, we define a shift operator S:CkCk by

S(b1,b2,,bk)=(bk,b1,,bk-1).

Definition 6.1

The circulant matrix B=circ(b) associated to the vector

b=(b1,b2,,bk)Ck

is the k×k matrix whose nth row is Sn-1b:

B=b1b2bk-1bkbkb1bk-2bk-1b3b4b1b2b2b3bkb1.

We denote by circ(k)Mk the set of all k×k complex circulant matrices.

With this notation, both A1 and A2 are k×k circulant matrices. In fact,

A1=circ{(-2,1,0,,0,1)}andA2=circ{(0,1,0,,0,-1)}.

Let ϵ=e2πik be a primitive kth root of unity, we define

Xl=1k1,ϵl,ϵ2l,,ϵ(k-1)lTCk,forl=0,,k-1,

and

Pk=11111ϵϵk-2ϵk-11ϵk-2ϵ(k-2)2ϵ(k-2)(k-1)1ϵk-1ϵ(k-1)(k-2)ϵ(k-1)2.

For the circulant matrix B=circ(b), let

λl=b1+b2ϵl++bkϵ(k-1)l,forl=0,,k-1.

A simple computation shows that BXl=λlXl. Hence λl is an eigenvalue of B with normalised eigenvector Xl. Since {X1,,Xk} is a linearly independent set of vectors in Ck, all of the eigenvalues of B are given by λl,l=0,,k-1. By direct computation, the eigenvalues of A1 are

λ1,l=-2+ϵl+ϵ(k-1)l=-4sin2lπk,forl=0,,k-1,

and the eigenvalues of A2 are

λ2,l=ϵl-ϵ(k-1)l=2isin2lπk,forl=0,,k-1.

Let diag(a1,a2,,ak) denote the diagonal matrix with diagonal entries a1,a2,,ak and

M=sinπk2(A1+4I)sinπkcosπkA2-sinπkcosπkA2-cosπk2A1.

From the above discussion for the circulant matrix, using

P=Pk0k0kPk

and

0k=diag0,0,0,,0,

we have

P-1MP=Pk-10k0kPk-1sinπk2(A1+4I)sinπkcosπkA2-sinπkcosπkA2-cosπk2A1Pk0k0kPk=4sinπk2(I-D1)isin2πkD2-isin2πkD24cosπk2D1,

where

D1=diag0,sinπk2,sin2πk2,,sin(k-1)πk2,

and

D2=diag0,sin2πk,sin4πk,,sin2(k-1)πk.

Next we divide the matrix P-1MP into k two by two matrices, where the l-th matrix (l=0,1,,k-1) is given by

4sinπk2coslπk2isin2πksin2lπk-isin2πksin2lπk4cosπk2sinlπk2. 6.11

It is easy to see that the determinant of the above matrix is 0 and its trace is positive. Further, we see that the zero eigenvector of the above matrix is

cosπksinlπk,isinπkcoslπkT. 6.12

Since the leading order matrix M admits zero eigenvalues with geometric multiplicity k, we have to expand the matrix Mμ(q) to the next order to determine if these small eigenvalues have positive or negative real part.

Before doing that, we point out a useful fact. Let us consider for example the term 2(q)q2,1q1,1. By direct computation we get

(q)q1,1=ξσ12e-σ|q1-q2|1|q2-q1|32(q2-q1)1=K~(|q1-q2|)(q2-q1)1. 6.13

Computing another derivative of K~(|q1-q2|)(q2-q1)1 with respect to q2,1, we note that there are two types of terms:

K~(|q1-q2|)q2,1(q2-q1)1(q2-q1)1andK~(|q1-q2|)(q2-q1)1q2,1.

The first term is of the same symmetry class as the leading order term (i.e., the higher order term differs from the leading order term only by some small factor). Therefore, this term can be absorbed into the leading-order matrix M.

However, the second term is different and it has to be taken into account. In fact, we will see that this type of terms can be used to resolve the stability problem. We can re-write the second term as follows:

K~(|q1-q2|)(q2-q1)1q2,1=-K~(|q1-q2|)122q2,1q1,1|q2-q1|2. 6.14

Hence, up to some factors it is enough for us to consider the terms 122q2,jq1,i|q1-q2|2,i,j=1,2. These terms together with c3ε2μ(0) are the next order terms in the matrix Mμ(q).

Using the local coordinate frames of q1 and q2 to express Cartesian local coordinates xi,j,i,j,=1,2, we get

x1,1=q1,1,x1,2=q1,2,x2,1=q2,1cos2πk-q2,2sin2πk,x2,2=q2,1sin2πk+q2,2cos2πk.

Using (6.8) and (6.9), this implies

|q1-q2|2=q10+q1,1q1|q1|+q1,2q1|q1|-q20+q2,1q2|q2|+q2,2q2|q2|2=Rσcosπk+x2,2-x1,22+Rσsinπk-x2,1+x1,12=Rσ2+2Rσcos2πkq2,2+sin2πkq2,1-q1,2cosπk+2Rσ-cos2πkq2,1+sin2πkq2,2+q1,1sinπk+cos2πkq2,2+sin2πkq2,1-q1,22+-cos2πkq2,1+sin2πkq2,2+q1,12=Rσ2+2Rσcos2πkq2,2+sin2πkq2,1-q1,2cosπk+2Rσ-cos2πkq2,1+sin2πkq2,2+q1,1sinπk+q1,12+q1,22+q2,12+q2,22-2q1,1q2,1cos2πk+2q1,1q2,2sin2πk-2q1,1q2,2cos2πk-2q1,2q2,1sin2πk.

As a consequence, we have

2|q1-q2|22qi,j=2,i,j=1,2,2|q1-q2|2q1,1q2,1=2|q1-q2|2q1,2q2,2=-2cos2πk, 6.15
2|q1-q2|2q1,1q2,2=2sin2πk,2|q1-q2|2q1,2q2,1=-2sin2πk,2|q1-q2|2q1,1q1,2=2|q1-q2|2q2,1q2,2=0. 6.16

Similarly, in local coordinates q1,q2 we have

|q1|2=q10+q1,1q1|q1|+q1,2q1|q1|2=Rσ2+q1,12+q1,22+2q1,1|Rσ|,

where we used q1·q1=0. This implies

2|q1|2q1,12=2|q1|2q1,22=2,2|q1|2q1,1q1,2=0. 6.17

For the terms c3ε2μ(0), from (4.10) we derive that

4sinπk2K~(|q1-q2|)=c3ε2μ(0). 6.18

From the above discussion and (6.13)–(6.18), expanding the matrix Mμ(q) we get the following second order contribution:

-K~(|q1-q2|)M2=-K~(|q1-q2|)cos2πkA1-sin2πkA2sin2πkA2cos2πkA1. 6.19

By using the matrix Pk, we diagonalise the matrix M2,

Pk-100Pk-1M2Pk00Pk=-K~(|q1-q2|)-4cos2πkD1-2isin2πkD22isin2πkD2-4cos2πkD1.

From the discussion of the leading-order matrix P-1MP, we know that the vectors

vl,1=0,,cosπksinlπkl+1,0,,isinπkcoslπkk+l+1,,0T,(l=0,1,,k-1)

are the eigenvectors with zero eigenvalues of the diagonal form. To show the stability of the eigenvalues in the linear subspace spanned by these eigenvectors, we have to evaluate the bilinear form with respect to these eigenvectors and show that

μl=(P-1M2P)vl,1,vl,1vl,1,vl,10,(l=0,1,,k-1).

If (P-1M2P)vl,1,vl,1=0 some further study is needed. We compute

μl=(P-1M2P)vl,1,vl,1vl,1,vl,1=-4cos2πksinlπk2+4sin2πksin2lπkcosπksinπkcoslπksinlπkcosπk2sinlπk2+sinπk2coslπk2.

Next we discuss when all eigenvalues are positive (linearly stable solution) or some eigenvalues are negative (linearly unstable solution).

For l=0 we have μl=0. This eigenvalue and its eigenvector are connected to rotational invariance of solutions.

For l=1 we compute the numerator in the expression for μ1 as

-8cos2πksinπk4cosπk2+16sinπk4cosπk4=8sinπk4cosπk2>0.

For l=k-1 we compare with the case l=1. The terms sin2lπk and coslπk change sign, the other terms are the same as for l=1. The result is the same as for l=1. The eigenvalues l=1 and l=k-1 together with their eigenvectors correspond to translations and they are stable.

For l=2 we compute the numerator of μ2 as

-4cos2πksin2πk2cosπk2sin2πk2+sinπk2cos2πk2+4sin2πksin4πkcosπksinπkcos2πksin2πk=-4cos2πksin2πk2cosπk2sin2πk2+sinπk2cos2πk2-sin2πk2cosπk2-sinπk2=-4cos2πksin2πk2sinπk2.

Thus μ2>0 for k=3, μ2=0 for k=4 and μ2<0 for k=5,6,.

The eigenvalue for l=2 and k=4 is zero in the first two leading orders. To decide if it possibly contributes to an instability, further expansions are required. This computation is beyond the scope of this paper. We expect that the eigenvalue will be stable and the cluster with 4 spikes on a regular polygon is linearly stable.

The eigenvalue for l=2 and k=5,6, is negative and so the cluster with 5 or more spikes is linearly unstable.

In summary we have considered the small eigenvalues and shown the following: The clusters with 2 spikes or 3 spikes on a polygon are both linearly stable. The clusters with 5 or more spikes are linearly unstable. The borderline case is the cluster with 4 spikes for which one eigenvalue requires further investigation to determine its stability.

Next we consider clusters with spikes located on a regular polygon plus a spike in its centre.

Cluster of spikes on a polygon with centre

In this section, we sketch how the approach can be adapted to show existence and stability of a cluster for which the spikes are located at the vertices of a regular k-polygon with centre. The spike positions are

Qε=q=(q1,,qk,0)qi=2R~εcos2(i-1)πk,2R~εsin2(i-1)πk, 7.1

where R~ε is chosen such that

1CDεlog1DlogDεR~εCDεlog1DlogDε 7.2

for some constant C>1 independent of ε and D.

To get the radius for the equilibrium position, we compute

W~ε,1,1(q)=c1ξσR~σ-1/2e-R~σ+c2ε2R~εμ(0)+h.o.t.=0,

where R~σ=σR~ε. We get

R~ε,0=1σlog1D-32loglog1D-logξc3+Ologlog1Dlog1D.

Due to symmetry we also have W~ε,1,2(q)=0 and W~ε,k+1,1(q)=W~ε,k+1,2(q)=0. From this we get the existence of a steady state of spikes located at the k vertices of a polygon and its centre, where k can be any natural number.

Next we consider the stability of this spike cluster steady state. We assume that k5. We take the same rotated coordinates as above around the vertices of the polygon. For the origin located in the centre of the polygon we keep Cartesian coordinates x1 and x2.

The matrix M~μ(q) is now given as follows:

M~μ(q)=ξσ32e-σ|q1|1|q1|52M~1M~2M~3M~4+h.o.t.=ξσ32e-σ|q1|1|q1|52M~+h.o.t.,

where

M~1=1000-10100-cos2πk0010-cos4πk-1-cos2πk-cos4πk-cos2(k-1)πkk2,M~2=000000000-sin2πk0000-sin4πk00000,M~3=0000000000000000-sin2πk-sin4πk-sin2(k-1)πk0,M~4=0000000000000000000k2.

We multiply M~ from the right by the vector a=(a1,a2,,ak,ak+1,0,0,,0,a2k+2)T, i.e. we assume that the components k+2,k+3,,2k+1 are all zero, and we also multiply M~ from the left by the transpose of this vector. Further, we set ak+1=α,a2k+2=β, where α and β are some real numbers. Then we get

aTM~a=l=1kal-αcos2(l-1)πk-βsin2(l-1)πk2.

This means the matrix M~ is positive semi-definite if it is restricted to the components 1,2,,k, k+1,2k+2. The eigenvalue of any eigenvector in this class is always nonnegative. It is zero if and only if

al=αcos2(l-1)πk+βsin2(l-1)πkforl=1,2,,k,

where α and β are some real numbers which are independent of l. These eigenvectors have positive eigenvalues for the second-order part of the matrix. Similar to the computation for a polygon without centre it can be shown that there are positive contributions to the eigenvalues coming from the components k+1 and 2k+2 which are related to the spike at the centre. Note that these eigenvectors correspond to translations.

In addition we have to study the eigenvalue of any eigenvector orthogonal to this class, i.e. for which the components 1,2,,k,k+1,2k+2 are zero and the components k+2,,2k+1 are arbitrary. The leading-order matrix M for the cluster of the polygon without centre defined in Sect. 6 is the second-order contribution here. It is positive semi-definite in this class (since A1 is positive semi-definite). It is strictly positive definite except for the eigenvector (0,0,,0,1,1,,1)T which has zero eigenvalue (since A1 has zero eigenvector (1,1,,1)T). Note that this eigenvector corresponds to rotations. Here the components k+2,,2k+1 for the polygon with central spike become the components k+1,,2k of the vector for the polygon without centre since the components k+1 and 2k+2 are dropped.

These computations show that the eigenvalues of M~ are nonnegative and they are zero only for eigenvectors which correspond to the rotational invariance of the problem. Together we get the stability of the cluster with spikes located at the vertices of a regular polygon with k5 vertices plus one spike at its centre.

Discussion

We have shown the existence of spike clusters located near a nondegenerate minimum point of the precursor gradient for the Gierer–Meinhardt system such that the spikes are located on regular polygons. We have proved that these solutions are stable for two or three spikes and unstable for five or more spikes. We have considered the problem in the rotationally symmetric case. We have assumed that the precursor and the domain are both rotationally symmetric.

It will be interesting to extend these results to the case that the precursor and the domain are not rotationally symmetric. We are currently studying these effects using the approach in [2], where the existence of spiky patterns for the Schrödinger equation has been extended from the case of a rotationally symmetric potential to the general case. If μ is not rotationally symmetric generically there will be certain possible orientations of the spike cluster and we expect to have stable and unstable equilibrium orientations. If the domain is not a disk higher order terms coming from the regular part of the Green’s function will determine the orientation of the spike cluster and we expect to have stable and unstable equilibrium orientations. Because of the smallness of the inhibitor diffusivity we expect that the influence of μ will dominate that of the domain boundary.

Further analysis is needed to resolve the stability issue for a 4-spike cluster (regular polygon with 4 vertices). Further calculations are required to show that the (k+1)-spike cluster for k6 (regular polygon with k vertices plus a spike in the centre) is stable or unstable. The stability problem in this case requires some new analysis since the interaction between spikes is of a different type from the one considered in this paper. These issues are currently under investigation.

Whereas in one spatial dimension the spikes in a cluster are aligned with equal distance in leading order (although they differ in higher order) [27], in two spatial dimensions a variety of different spike configurations are possible. In this paper we have considered regular polygons and polygons with a spike in the centre. Other arrangements include concentric multiple polygons or positions close to regular polygons. Similar configurations have been studied in [1].

Biologically speaking, the precursor is the information retained from a previous stage of development and the patterns discovered in the reaction–diffusion system at the present will be able to determine the development in the future. The Gierer–Meinhardt system with precursor can be considered a minimal model to describe this behaviour. Generally one has to study larger systems which take into account other effects to make more reliable biological predictions. Therefore it will be interesting to consider reaction–diffusion systems of three and more components and investigate the role which spike clusters play in such systems. One such system is a consumer chain model for which existence and stability of a clustered spiky pattern has been investigated by the first two authors [25]. However, a more systematic approach will be needed to gain a better understanding of the role played by spike clusters in guiding biological development.

Acknowledgements

The research of J. Wei is partially supported by Natural Sciences and Engineering Research Council of Canada. MW thanks the Department of Mathematics at the University of British Columbia for their kind hospitality. The authors are grateful for the referee’s comments which helped to improve the paper.

Appendix A: Some results for Liapunov–Schmidt reduction in Sect. 3

In this appendix we will prove some results which are needed for Liapunov–Schmidt reduction in Sect. 3. In particular, we are going to show that the equation

πε,qSε,qAε,q+ϕε,qHε,q+ψε,q=0

has a unique solution Σε,q=ϕε,qψε,qKε,q if maxεD,DlogDε is small enough.

Recall that

Lε,q=πε,qL~ε,q:Kε,qCε,q.

We will first show that this linear operator is uniformly invertible. Then we will use this to prove the existence of Σε,q=ϕε,qψε,q.

As a preparation, in the following two propositions we show the invertibility of the corresponding linearised operator Lε,q.

Proposition 9.1

Let Lε,q be given in (3.23). There exists a positive constant δ¯ such that for all maxεD,DlogDε(0,δ¯), we can find a positive constant C which is independent of ε,D such that

Lε,qΣL2(BR/ε)CΣH2(BR/ε) 9.1

for arbitrary qQε,ΣKε,q. Further, the map Lε,q is surjective.

Proof of Proposition 9.1

Suppose (9.1) is false. Then there exist sequences {εn},{Dn}, {qn}, {ϕn},{ψn} such that maxεnDn,DnlogDnεn0, qnQε, ϕn=ϕεn,Dn,qn,n=1,2, and

Lεn,qnΣnL2(BR/ε)0asn, 9.2
ϕnH2(BR/ε)+ψnH2(BR/ε)=1,n=1,2,. 9.3

On the other hand, we note ψn satisfies

Δψn-σ2ψn+2Aεn,qnϕn=0.

It is easy to see from the above equation we get ψnH2(BR/ε)CϕnH2(BR/ε). Then we can assume ϕnH2(BR/ε)=1. We define ϕn,i,i=1,2,,k and ϕn,k+1 as follows:

ϕn,i(x)=ϕn(x)χx-qiRεsinπkandϕn,k+1(x)=ϕn(x)-i=1kϕn,i(x), 9.4

where χ(x) is defined in (3.3). Since for i=1,2,,k each sequence {ϕn,i},n=1,2, is bounded in HN2(R2) it has a weak limit in HN2(R2), and therefore also a strong limit in L2(R2) and L(R2). Call this limit ϕi. Then Φ=(ϕ1,,ϕk)T solves the system LΦ=0, where L is given in (2.6). By Lemma 2.4, ΦKer(L)=K0K0. Since ϕnKεn,qn, by taking n we get ΦKer(L). Therefore, Φ=0.

By elliptic estimates we have ϕn,iH2(BR/ε)0 as n for i=1,2,,k. Further, since ϕn,k+1ϕk+1 we get

Δϕk+1-ϕk+1=0inBR/ε.

Therefore we conclude ϕk+1=0 and ϕn,k+1H2(BR/ε)0 as n. This contradicts to ϕnH2(BR/ε)=1.

To complete the proof of Proposition 9.1 we just need to show that conjugate operator to Lε,q (denoted by Lε,q) is injective from Kε,q to Cε,q and the proof for Lε,q follows almost the same process as for Lε,q and we omit it. Thus we finish the proof of Proposition 9.1.

Now we are in position to solve the equation

πε,qSε,qAε,q+ϕHε,q+ψ=0. 9.5

Since Lε,qKε,q is invertible (call the inverse Lε,q-1), we can rewrite (9.5) as

Σ=-Lε,q-1πε,qSε,qAε,qHε,q-Lε,q-1πε,q)(Nε,q(Σ)Mε,q(Σ), 9.6

where

Σ=ϕψ,Nε,q(Σ)=Sε,qAε,q+ϕHε,q+ψ-Sε,qAε,qHε,q-Sε,qAε,qHε,qϕψ,

and the operator Mε,q is defined by (9.6) for ΣHN2(BR/ε)×HN2(BR/ε). We are going to show the operator Mε,q is a contraction map in

Bε,D=ΣHN2(BR/ε)×HN2(BR/ε)ΣH2(BR/ε)<η

provided maxεD,DlogDε is small enough. We have by Lemma 3.1 and Proposition 9.1 that

Mε,q(Σ)H2)N(BR/ε)C(πε,qNε,q(Σ)L2(BR/ε)+πε,qSε,qAε,qHε,qL2(BR/ε))C(c(η)η+cε,D),

where C>0 is independent of η, c(η)0 as η0 and cε,D0 as

maxεD,DlogDε0.

Similarly we can show

Mε,q(Σ)-Mε,q(Σ)H2(BR/ε)Cc(η)Σ-ΣH2(BR/ε).

If we choose η sufficient small, then Mε,q is a contraction map on Bε,D. The existence of the fixed point Σε,q together with an error estimate now follow from the contraction mapping principle. Moreover Σε,q is a solution of (9.6).

Thus we have proved the following lemma.

Lemma 9.2

There exists δ¯>0 such that for

maxεD,DlogDε(0,δ¯)

and qQε, we can find a unique (Φε,q,Ψε,q)Kε,q satisfying

Sε,qAε,q+Φε,qHε,q+Ψε,qCε,q

and

(Φε,q,Ψε,q)H2(BR/ε)C1logDε+ε2Rε2. 9.7

In the following, we need more refined estimates on Φε,q. We recall that S1 can be decomposed into the two parts S1,1 and S1,2, where S1,1 is in leading order an odd function and S1,2 is in leading order a radially symmetric function. Similarly, we can decompose Φε,q:

Lemma 9.3

Let Φε,q be defined in Lemma 9.2. Then for x=qi+z,|σz|<δ and δ>0 small enough, we have

Φε,q=Φε,q,1+Φε,q,2, 9.8

where Φε,q,2 is a radially symmetric function in z and

Φε,q,1H2(BR/ε)=Oε2Rε+σlog1σ-1Rσ-12e-Rσ+σ2log1σ-1. 9.9

Proof

Let S[u]:=S1(u,T[u]). We first solve

S[Aε,q+Φε,q,2]-S[Aε,q]+j=1kS1,2(x-qj)Cε,q, 9.10

for Φε,q,2Kε,q. Then we solve

S[Aε,q+Φε,q,2+Φε,q,1]-S[Aε,q+Φε,q,2]+j=1kS1,1(x-qj)Cε,q, 9.11

for Φε,q,1Kε,q. Using the same proof as in Lemma 9.2, both Eqs. (9.10) and (9.11) have unique solution for maxDε,DlogDε sufficiently small. By uniqueness, Φε,q=Φε,q,1+Φε,q,2. It is easy to see that

S1,1L2(BR/ε)=Oε2Rε+σlog1σ-1Rσ-12e-Rσ+σ2log1σ-1

and S1,2Cε,q. Then we can conclude that Φε,q,1 and Φε,q,2 have the required properties.

Appendix B: Stability analysis II: study of small eigenvalues

In this section, we shall study the small eigenvalues for Eq. (5.1). Namely, we assume λε0 as maxεD,DlogDε0. We shall show that the small eigenvalues are related to μ(0) and the Green function.

Again let (Aε,Hε) be the equilibrium state constructed for equation (1.2). Let

Aε,j=χε,qj(x)Aε(x),j=1,2,,k,

where χε,qj is defined before (3.3). Then it is easy to see that

Aε=j=1kAε,j+h.o.t.inHN2(BR/ε). 10.1

In last section, we have derived the nonlocal eigenvalue (5.5). Let us now set λ0=0 in (5.5), we have that

Δϕi-ϕi+2wϕi-2R2wϕidxR2w2dxw2=0, 10.2

which is equivalent to

L0ϕi-2R2wϕidxR2w2dxw=0,i=1,,k,

where L0 is defined in (2.2). By Lemma 2.1, we have

ϕi-2R2wϕidxR2w2dxwspanwxj,j=1,2,i=1,2,,k. 10.3

Multiplying (10.3) by w and integrating over R2 and summing up, we have R2wϕidx=0, and hence

ϕjK0=spanwxj,j=1,2,i=1,2,,k. 10.4

(10.4) suggests that, at least formally, we should have

ϕεj=1ki=12aj,iεwxi(x-qj), 10.5

where aj,iε are some constant coefficients.

Next we find a good approximation of wxi(x-qj). Note that Aε,j(x)w(x-qj) in H2(BR/ε), and Aε,j satisfies

ΔAε,j-μ(|εx|)Aε,j+Aε,jHε2+h.o.t.=0.

Then we find Aε,jxi satisfies

ΔAε,jxi-μ(|εx|)Aε,jxi+2Aε,jHεAε,jxi-Aε,j2Hε2Hεxi-εμ(|εx|)xirAε,j+h.o.t.=0, 10.6

and we have Aε,jxi=(1+o(1))wxi(x-qj).

We now decompose

ϕε=j=1ki=12aj,iεAε,jxi+ϕε 10.7

with complex numbers aj,iε, where

ϕεK~ε,q:=spanAε,jxij=1,2,,k,i=1,2. 10.8

Our idea is to show that this is a good choice because the error ϕε is small in a suitable norm and thus can be neglected.

For the inhibitor eigenfunction ψε, we make the following decomposition according to ϕε

ψε=j=1ki=12aj,iεψε,j,i+ψε, 10.9

where ψε,j,i is the unique solution of the following problem,

Δψε,j,i-σ2(1+τλε)ψε,j,i+2ξεAε,jAε,jxi=0inBR/ε, 10.10

and

Δψε-σ2(1+τλε)ψε+2ξεAεϕε=0inBR/ε. 10.11

Suppose that ϕεH2(BR/ε)=1. We have aj,iε=BR/εϕεAε,jxidxR2(wx1)2dx, therefore, we get |aj,iε|C.

Substituting the decomposition of ϕε and ψε into the first equation in (5.1), we have

j=1ki=12aj,iεAε,j2Hε2(-ψε,j,i+Hεxi)+εj=1ki=12aj,iεμ(|εx|)yiAε,j+Δϕε-μ(|εx|)ϕε+2AεHεϕε-Aε2Hε2ψε-λεϕε+h.o.t.=λεj=1ki=12aj,iεAε,jxi. 10.12

We set

I1=j=1ki=12aj,iεAε,j2Hε2-ψε,j,i+Hεxi+εj=1ki=12aj,iεμ(|εx|)yiAε,j, 10.13

and

I2=Δϕε-μ(|εx|)ϕε+2AεHεϕε-Aε2Hε2ψε-λεϕε. 10.14

Next we shall first derive the estimate for ϕε. Using (10.12), since ϕεK~ε,q, then similar to the proof of Proposition 9.1, we have that

ϕεH2(BR/ε)CI1L2(BR/ε). 10.15

So our aim is to estimate the term I1. We note that

Hεxi=ξBR/εxiGσ(x,z)Aε2(z)dz=ξBR/εxi(Kσ(x,z)+Hσ(x,z))Aε2(z)dz+h.o.t.

and (for τ=0)

ψε,j,i(x)=2ξBR/εGσ(x,z)Aε,jAε,jzidz=ξBR/ε(Kσ(x,z)+Hσ(x,z)+o(σ2))ziAε,j2dz,

where Kσ(x,z) and Hσ(x,z) refer to the singular and the regular part of the Green function Gσ(x,z), respectively. For τ>0 the last formula should use Gσλε instead of Gσ. Since λε0 it can be shown by comparing the two Green functions that the error from this term does not contribute to the small eigenvalues in leading order. We omit the details.

Then from the above two formulas, we have

Hεxi-ψε,j,i(x)=ξBR/εxiKσ(x,z)Aε,j2(z)-Kσ(x,z)ziAε,j2dz+ξBR/εxiHσ(x,z)Aε,j2(z)-Hσ(x,z)ziAε,j2dz+ξBR/εxil,ljGσ(x,z)Aε,l2dz+h.o.t..

Using the fact that

xiKσ(x,z)+ziKσ(x,z)=h.o.t.,forxz, 10.16

and integrating by parts, we get

Hεxi-ψε,j,i(x)=ξxiFj(x)+h.o.t.R2w2dz, 10.17

where Fj(x)=Hσ(x,qj)+l,ljGσ(x,ql). Then from (10.13), we get

I1=j=1ki=12aj,iεξFj(qj+εx)xiR2w2dz+εμ(|εqj|)xirw(x)+h.o.t.,

where r=x12+x22. Note that

Fj(qj)xi=12F(q)qj,i.

From the proof of Theorem 1.1, we observe that

w2(x)ξFj(x)xiBR/εw2dz+εμ(|εqj|)xirw(x)=o(ε2Rε).

Hence, we have

I1L2(BR/ε=oε2Rεj=1ki=12|aj,iε| 10.18

and

ϕεH2(BR/ε)CI1L(BR/ε)=oε2Rεj=1ki=12|aj,iε|. 10.19

Using the equation for ψε and (10.19), we obtain that

ψεL(Ω)=oε2Rεj=1ki=12|aj,iε|. 10.20

We calculate

BR/εI2Aε,lxmdx=BR/εAε,l2Hε2Hεxmϕε-Aε,lxmψεdx-λεBR/εϕεAε,lxmdx=BR/εAε,l2Hε2Hεxm(ql+x)-Hεxm(ql)ϕεdx+BR/εAε,l2Hε2Hεxm(ql)ϕεdx-λεBR/εϕεAε,lxmdx-BR/εAε,l2Hε2Aε,lxm(ψε(ql+x)-ψε(ql))dx-λεBR/εψε(ql)Aε,l2Hε2Aε,lxmdx=oσε2Rεj=1ki=12|aj,iε|. 10.21

Then we study the algebraic equation for aj,iε. Multiplying both sides of (10.12) by Aε,lxm and integrating over R2, we obtain

r.h.s.=λεj=1ki=12aj,iεBR/εAε,jxiAε,lxmdx=λεj=1ki=12aj,iεδjlδimR2wz12dz(1+o(1))=λεal,mεR2wz12dz(1+o(1)).

For the left hand side, we have

l.h.s.=j=1ki=12aj,iεBR/εAε2Hε2Hεxi-ψε,j,iAε,lxmdx+j=1ki=12aj,iεBR/εεμ(|εqj|)xirAε,jAε,lxmdx+BR/εI2Aε,lxmdx. 10.22

Using (10.17), we obtain

l.h.s.=j=1ki=12aj,iεξ2ql,mqj,iF(q)R2w2wzmzmdzR2w2dz+j=1ki=12aj,iεBR/εε2μ(|εqj|)xiAε,jAε,lxmdx+oσε2Rεj=1ki=12|aj,iε|. 10.23

Note that in (10.23) there is no summation over m. From (10.22) and (10.23), we have

l.h.s.=-c1ξj=1ki=12aj,iε2ql,mqj,iF(q)+c2ε2j=1ki=12aj,iεδjlδmiμ(|εqi|)+h.o.t., 10.24

where c1,c2 have been introduced in (4.4) and (4.8), respectively. Combining the l.h.s. and r.h.s., we have

j=1ki=12aj,iε-c1ξ2ql,mqj,iF(q)+c2ε2δjlδmiμ(|εqj|)+h.o.t.=λεal,mεR2wz12dz(1+o(1)). 10.25

From (10.25), we see that the small eigenvalues with λε0 and are related to the eigenvalues of the 2k×2k matrix Mμ(q), where the (j+ki,l+km)-th component is the following

(Mμ(q))j+ki,l+km=-c1ξ2ql,mqj,iF(q)+c2ε2δjlδmiμ(|εqj|),1j,lk,1i,m2.

In Sect. 6 this matrix has been studied further and its eigenvectors and eigenvalues have been computed explicitly.

Contributor Information

Juncheng Wei, Email: jcwei@math.ubc.ca.

Matthias Winter, Phone: +441895267179, Email: matthias.winter@brunel.ac.uk.

Wen Yang, Email: math.yangwen@gmail.com.

References

  • 1.Del Pino M, Kowalczyk M, Wei J. Multi-bump ground states of the Gierer–Meinhardt system in R2. Ann. Inst. H. Poincaré Anal. Non Linéaire. 2003;20(1):53–85. doi: 10.1016/S0294-1449(02)00024-0. [DOI] [Google Scholar]
  • 2.Del Pino M, Wei J, Yao W. Infinitely many positive solutions of nonlinear Schrödinger equations with non-symmetric potentials. Cal. Var. PDE. 2015;53(1–2):473–523. doi: 10.1007/s00526-014-0756-3. [DOI] [Google Scholar]
  • 3.Doelman A, Kaper TJ, van der Ploeg H. Spatially periodic and aperiodic multi-pulse patterns in the one-dimensional Gierer–Meinhardt equation. Methods Appl. Anal. 2001;8:387–414. [Google Scholar]
  • 4.Ei S-I, Wei JC. Dynamics of metastable localized patterns and its application to the interaction of spike solutions for the Gierer-Meinhardt systems in two spatial dimensions. Jpn. J. Ind. Appl. Math. 2002;19:181–226. doi: 10.1007/BF03167453. [DOI] [Google Scholar]
  • 5.Gierer A, Meinhardt H. A theory of biological pattern formation. Kybernetik (Berlin) 1972;12:30–39. doi: 10.1007/BF00289234. [DOI] [PubMed] [Google Scholar]
  • 6.Harrison LG. Kinetic Theory of Living Pattern. New York: Cambridge University Press; 1993. [Google Scholar]
  • 7.Herschowitz-Kaufman M. Bifurcation analysis of nonlinear reaction–diffusion equations II, steady-state solutions and comparison with numerical simulations. Bull. Math. Biol. 1975;37:589–636. [Google Scholar]
  • 8.Iron D, Ward M, Wei J. The stability of spike solutions to the one-dimensional Gierer–Meinhardt model. Physica D. 2001;150:25–62. doi: 10.1016/S0167-2789(00)00206-2. [DOI] [Google Scholar]
  • 9.Kra I, Simanca SR. On circulant matrices. Not. AMS. 2012;59:368–377. [Google Scholar]
  • 10.Lin, C.S., Ni, W.M.: On the diffusion coefficient of a semilinear Neumann problem. Calculus of variations and partial differential equations (Trento, 1986), Lecture Notes in Mathematics, vol. 1340, pp. 160–174, Springer, Berlin (1988)
  • 11.Lyons MJ, Harrison LG, Lakowski BC, Lacalli TC. Reaction diffusion modelling of biological pattern formation: application to the embryogenesis of Drosophila melanogaster. Can. J. Phys. 1990;68:772–777. doi: 10.1139/p90-112. [DOI] [Google Scholar]
  • 12.Meinhardt H. The Algorithmic Beauty of Sea Shells. 2. Berlin: Springer; 1998. [Google Scholar]
  • 13.Meinhardt H. Models of Biological Pattern Formation. London: Academic Press; 1982. [Google Scholar]
  • 14.Ni WM, Takagi I. On the shape of least energy solution to a semilinear Neumann problem. Commun. Pure Appl. Math. 1991;41:819–851. doi: 10.1002/cpa.3160440705. [DOI] [Google Scholar]
  • 15.Takagi I. Point-condensation for a reaction–diffusion system. J. Differ. Equ. 1986;61:208–249. doi: 10.1016/0022-0396(86)90119-1. [DOI] [Google Scholar]
  • 16.Turing AM. The chemical basis of a morphogenesis. Philos. Trans. R. Soc. Lond. Ser. B. 1952;237:37–72. doi: 10.1098/rstb.1952.0012. [DOI] [Google Scholar]
  • 17.Ward MJ, McInerney D, Houston P, Gavaghan D, Maini PK. The dynamics and pinning of a spike for a reaction–diffusion system. SIAM J. Appl. Math. 2002;62:1297–1328. doi: 10.1137/S0036139900375112. [DOI] [Google Scholar]
  • 18.Ward MJ, Wei J. Asymmetric spike patterns for the one-dimensional Gierer–Meinhardt model: equilibria and stability. Eur. J. Appl. Math. 2002;13:283–320. doi: 10.1017/S0956792501004442. [DOI] [Google Scholar]
  • 19.Wei J. On single interior spike solutions of Gierer–Meinhardt system: uniqueness and spectrum estimates. Eur. J. Appl. Math. 1999;10:353–378. doi: 10.1017/S0956792599003770. [DOI] [Google Scholar]
  • 20.Wei J, Winter M. On the two-dimensional Gierer–Meinhardt system with strong coupling. SIAM J. Math. Anal. 1999;30:1241–1263. doi: 10.1137/S0036141098347237. [DOI] [Google Scholar]
  • 21.Wei J, Winter M. On multiple spike solutions for the two-dimensional Gierer–Meinhardt system: the strong coupling case. J. Differ. Equ. 2002;178:478–518. doi: 10.1006/jdeq.2001.4019. [DOI] [Google Scholar]
  • 22.Wei J, Winter M. Spikes for the two-dimensional Gierer–Meinhardt system: the weak coupling case. J. Nonlinear Sci. 2001;6:415–458. doi: 10.1007/s00332-001-0380-1. [DOI] [Google Scholar]
  • 23.Wei J, Winter M. Existence, classification and stability analysis of multiple-peaked solutions for the Gierer–Meinhardt system in R. Methods Appl. Anal. 2007;14(2):119–163. [Google Scholar]
  • 24.Wei J, Winter M. On the Gierer–Meinhardt system with precursors. Discrete Contin. Dyn. Syst. 2009;25(1):363–398. doi: 10.3934/dcds.2009.25.363. [DOI] [Google Scholar]
  • 25.Wei J, Winter M. Stability of cluster solutions in a cooperative consumer chain model. J. Math. Biol. 2014;68:1–39. doi: 10.1007/s00285-012-0616-8. [DOI] [PubMed] [Google Scholar]
  • 26.Wei J, Winter M. Mathematical Aspects of Pattern Formation in Biological Systems, Applied Mathematical Sciences. London: Springer; 2014. [Google Scholar]
  • 27.Wei J, Winter M. Stable spikes clusters for the one-dimensional Gierer-Meinhardt system. Eur. J. Appl. Math. 2017;28(4):576–635. doi: 10.1017/S0956792516000450. [DOI] [Google Scholar]

Articles from Calculus of Variations and Partial Differential Equations are provided here courtesy of Springer

RESOURCES