In this paper, the dynamical behaviors of a 2-component coupled diffusive system modeling hair follicle spacing is considered. For the corresponding ODEs, we not only consider the stability and instability of the unique positive equilibrium solutions, but also show the existence of unstable Hopf bifurcating periodic solutions. For the reaction-diffusion equations, we are mainly interested in the Turing instability of the positive equilibrium solution, as well as Hopf bifurcations and steady-state bifurcations. Our results showed that, under certain conditions, the reaction-diffusion system not only has Hopf bifurcating periodic solutions (both spatially homogeneous and non-homogeneous, all unstable), but it also has non-constant positive bifurcating equilibrium solutions. This allows for a clearer understanding of the mechanism for the spatiotemporal patterns of this particular system.
Citation: Zhili Zhang, Aying Wan, Hongyan Lin. Spatiotemporal patterns and multiple bifurcations of a reaction- diffusion model for hair follicle spacing[J]. Electronic Research Archive, 2023, 31(4): 1922-1947. doi: 10.3934/era.2023099
[1] | Nan Xiang, Aying Wan, Hongyan Lin . Diffusion-driven instability of both the equilibrium solution and the periodic solutions for the diffusive Sporns-Seelig model. Electronic Research Archive, 2022, 30(3): 813-829. doi: 10.3934/era.2022043 |
[2] | Weiyu Li, Hongyan Wang . Dynamics of a three-molecule autocatalytic Schnakenberg model with cross-diffusion: Turing patterns of spatially homogeneous Hopf bifurcating periodic solutions. Electronic Research Archive, 2023, 31(7): 4139-4154. doi: 10.3934/era.2023211 |
[3] | Jiani Jin, Haokun Qi, Bing Liu . Hopf bifurcation induced by fear: A Leslie-Gower reaction-diffusion predator-prey model. Electronic Research Archive, 2024, 32(12): 6503-6534. doi: 10.3934/era.2024304 |
[4] | Yuan Xue, Jinli Xu, Yuting Ding . Dynamics analysis of a diffusional immunosuppressive infection model with Beddington-DeAngelis functional response. Electronic Research Archive, 2023, 31(10): 6071-6088. doi: 10.3934/era.2023309 |
[5] | Fengrong Zhang, Ruining Chen . Spatiotemporal patterns of a delayed diffusive prey-predator model with prey-taxis. Electronic Research Archive, 2024, 32(7): 4723-4740. doi: 10.3934/era.2024215 |
[6] | Cunbin An . Global bifurcation and structure of stationary patterns of a diffusive system of plant-herbivore interactions with toxin-determined functional responses. Electronic Research Archive, 2023, 31(4): 2095-2107. doi: 10.3934/era.2023107 |
[7] | Qing Li, Junfeng He . Pattern formation in a ratio-dependent predator-prey model with cross diffusion. Electronic Research Archive, 2023, 31(2): 1106-1118. doi: 10.3934/era.2023055 |
[8] | Xianjun Wang, Huaguang Gu, Bo Lu . Big homoclinic orbit bifurcation underlying post-inhibitory rebound spike and a novel threshold curve of a neuron. Electronic Research Archive, 2021, 29(5): 2987-3015. doi: 10.3934/era.2021023 |
[9] | Qinghua Zhu, Meng Li, Fang Han . Hopf bifurcation control of the ML neuron model with Hc bifurcation type. Electronic Research Archive, 2022, 30(2): 615-632. doi: 10.3934/era.2022032 |
[10] | Ruizhi Yang, Dan Jin . Dynamics in a predator-prey model with memory effect in predator and fear effect in prey. Electronic Research Archive, 2022, 30(4): 1322-1339. doi: 10.3934/era.2022069 |
In this paper, the dynamical behaviors of a 2-component coupled diffusive system modeling hair follicle spacing is considered. For the corresponding ODEs, we not only consider the stability and instability of the unique positive equilibrium solutions, but also show the existence of unstable Hopf bifurcating periodic solutions. For the reaction-diffusion equations, we are mainly interested in the Turing instability of the positive equilibrium solution, as well as Hopf bifurcations and steady-state bifurcations. Our results showed that, under certain conditions, the reaction-diffusion system not only has Hopf bifurcating periodic solutions (both spatially homogeneous and non-homogeneous, all unstable), but it also has non-constant positive bifurcating equilibrium solutions. This allows for a clearer understanding of the mechanism for the spatiotemporal patterns of this particular system.
The development of regularly arranged body parts has attracted the attention of a huge number of experimental biologists and theoreticians alike, and one of the crucial issues is getting to know the underlying mechanism in the formation of epidermal appendages such as feathers and hairs. Many theoretical models, including mathematical models of coupled reaction-diffusion equations, have been used to describe the formation of animal pigmentation patterns and distribution.
In [1], Sick et al. proposed a system of reaction-diffusion equations to model the influence of the Wnt signaling pathway in primary hair follicle initiation in mice. It is suggested that Wnt and Dkk have a primary influence on the spacing patterns of hair follicles in mice. The interactions between Wnt and Dkk are modeled by using a modified Gierer-Meinhardt reaction-diffusion (activator-inhibitor) model. The Wnt-Dkk interaction is modeled by the following 2-component reaction-diffusion equations:
{∂u∂t−d1Δu=ρ1u2(γ+v)(1+ku2)−μu,x∈Ω,t>0,∂v∂t−d2Δv=ρ2u2(γ+v)(1+ku2)−λv,x∈Ω,t>0,∂νu=∂νv=0,x∈∂Ω,t≥0,u(x,0)=u0(x)>0,v(x,0)=v0(x)>0,x∈Ω, | (1.1) |
where Ω is an open bounded domain in RN, N≥1, with the smooth boundary ∂Ω; u=u(x,t) and v=v(x,t) stand for the concentrations of the activator Wnt and the inhibitor Dkk at time t and position x∈Ω, respectively; d1>0 and d2>0 are the diffusion rates of the activator Wnt and the inhibitor Dkk, respectively; γ and k are non-negative saturation parameters. The constants ρ1 and ρ2 scale the production rates of the activator Wnt and the inhibitor Dkk, respectively; The negative terms denote that both chemicals decay linearly with constants μ and λ; u0,v0∈C2(Ω)∩C0(¯Ω) and the Neumann boundary conditions indicate that there is no flux of the chemical substances of u and v on the boundary.
System (1.1) is a modification of the homogeneous Gierer-Meinhardt model of the following form:
{∂u∂t−d1Δu=upvq+σ1(x)−u,x∈Ω,t>0,∂v∂t−d2Δv=urvs+σ2(x)−v,x∈Ω,t>0,∂νu=∂νv=0,x∈∂Ω,t≥0,u(x,0)=u0(x)>0,v(x,0)=v0(x)>0,x∈Ω, | (1.2) |
where u=u(x,t) and v=v(x,t) stand for the concentrations of the activator and the inhibitor at time t and position x∈Ω, respectively. The non-negative functions σ1(x) and σ2(x) are background source terms for the activator and the inhibitor, respectively. The exponents p,q,r and s are non-negative.
In [2], global existence of the solutions of System (1.1) is considered by calculating the uniform bounds. Analysis of the conditions for the emergence of spatially heterogeneous solutions is performed by using a limiting form of the original reaction-diffusion system. The conditions for pattern formation given in [1] are improved by including those subregions in the parameter space where far-from-equilibrium heterogeneous solutions occur.
To simplify the analysis of the Turing mechanism, in [1], a modification of System (1.1) is considered. In [1], the parameter γ is chosen so that γ≈0. Then, they obtained the following reaction-diffusion equations:
{∂u∂t−d1Δu=ρ1u2v(1+ku2)−μu,x∈Ω,t>0,∂v∂t−d2Δv=ρ2u2v(1+ku2)−λv,x∈Ω,t>0,∂νu=∂νv=0,x∈∂Ω,t≥0,u(x,0)=u0(x)>0,v(x,0)=v0(x)>0,x∈Ω. | (1.3) |
As argued in [2], System (1.3) captures the dynamics of (1.1) quite well, and, due to the smaller number of parameters, the conditions for emergence of Turing patterns are easier to analyze.
In [3], Rashkov considered the Turing instability of the positive constant equilibrium solution, as well as the existence and stability of both the regular and the discontinuous non-constant stationary solutions of System (1.3). Veerman and Doelman [4] showed that all of the positive non-constant regular equilibrium solutions (in the one-dimensional spatial domain) are unstable. The results are quite different from the corresponding singularly perturbed system, which has stable non-constant spike solutions.
We would also like to mention that a similar but different model used to describe the hair follicle bulb was proposed by Mooney and Nagorcka [5,6,7]. In particular, Nagorcka and Mooney [7] described a theoretical mechanism for cell differentiation based on the substances X and Y, which constitute a reaction-diffusion system, and the substance Z, which diffuses radially outward from the dermal papilla through the bulb. The model describes the reaction and diffusion of morphogens X and Y, and it is defined by
{∂u∂t−d1Δu=−au+bvp1+vp,x∈Ω,t>0,∂v∂t−d2Δv=−cu+e(vp+r)1+vp,x∈Ω,t>0,∂νu=∂νv=0,x∈∂Ω,t≥0,u(x,0)=u0(x)>0,v(x,0)=v0(x)>0,x∈Ω, | (1.4) |
where u and v can be seen as reduced or dimensionless concentrations of morphogens of X and Y, respectively, with positive parameters a,b,c,e and r and diffusion coefficients d1 and d2. The rates a,b,c and e have units of time−1, and r and p(>1) are dimensionless quantities. The terms −au+bvp/(1+vp) and −cu+e(vp+r)/(1+vp) represent the net rates of the productions u and v, respectively, where the reaction-diffusion system constitutes an activator-inhibitor pair with the morphogens X the inhibitor and Y the activator. For System (1.4), Yi et al. [8] provided some global analyses of the model that were dependent upon some parametric thresholds/constraints. They found that, when one of the dimensionless parameters is less than one, the unique positive equilibrium is globally asymptotically stable. On the contrary, when this threshold is greater than one, the existence of both steady-state and Hopf bifurcations can be observed under further parametric constraints. In [9], Yang and Ju considered the Turing instability of the spatially homogeneous periodic solutions of System (1.4).
In this paper, we are mainly concerned with the dynamics of both the reaction-diffusion equations (1.3) and its corresponding ODE system. The highlights of the paper are as follows:
1. Consider the dynamics of the ODEs. By using the decay rate λ of the inhibitor v as the main bifurcation parameter, we are able to show the existence of Hopf bifurcating periodic solutions. That is, when the decay rate λ crosses the critical value λ0, the densities of the activator u and the inhibitor v will undergo temporal oscillations. Moreover, we are able to prove that, once the Hopf bifurcating periodic solutions occur, they must be unstable. This is one of the novel points of this paper.
2. Consider the dynamics of the PDEs. Three kinds of results are obtained. First, under suitable conditions on the decay rate λ of the inhibitor v and the diffusion rates (d1 and d2), (uλ,vλ) will always be stable in the reaction-diffusion equations. In this case, the dynamics of the PDEs can be accurately determined by the dynamics of the ODEs. This is the so-called lumped parameter phenomenon [10]. Second, we are able to derive precise conditions on λ, as well as the diffusion rates (d1 and d2) such that Turing instability (see [11]) of (uλ,vλ) occurs. In this case, Turing patterns (spotted or striped patterns) of the reaction-diffusion system can be observed. Finally, under suitable conditions, non-constant Hopf bifurcating periodic solutions (both spatially homogeneous and spatially non-homogeneous) and non-constant, positive, steady-state bifurcating solutions can be obtained. These results are new and cannot be found in the existing literature. This is the other novel point of this paper.
The structure of this paper is as follows. In Section 2, we mainly study the dynamic behavior of the corresponding ordinary differential equation system of System (1.3), including the existence, stability and instability of the positive steady-state solution and the existence, stability of the Hopf bifurcating periodic solution. In Section 3, we mainly study the dynamic behavior of the reaction-diffusion system presented as System (1.3), including the existence, stability and instability of the positive steady-state solution and the existence of Hopf bifurcating periodic solutions and non-constant, positive, steady-state bifurcating periodic solutions. In the Appendix section, we list some of our proofs conducted in Section 3.
The corresponding ODE system of System (1.3) takes the following form:
dudt=ρ1u2v(1+ku2)−μu,dvdt=ρ2u2v(1+ku2)−λv. | (2.1) |
System (2.1) has a unique positive equilibrium solution (uλ,vλ), where
uλ:=√1k(ρ21λρ2μ2−1),vλ:=ρ2μρ1λuλ, | (2.2) |
where it is assumed that λ>ρ2μ2/ρ21.
In what follows, we shall always assume that λ>ρ2μ2/ρ21 so that (uλ,vλ) is the positive equilibrium solution of (2.1). We shall fix ρ1,ρ2 and k and choose λ and μ as two main bifurcation parameters.
We have the following results on the stability and instability of (uλ,vλ):
Theorem 2.1. Let λ>ρ2μ2ρ21 so that (uλ,vλ) is the unique positive equilibrium solution of (2.1). Define
λ0:=(√ρ21+16ρ2μ−ρ1)μ4ρ1. | (2.3) |
Then, 0<λ0<2ρ2μ2ρ21. In particular, the following results hold true:
1. Suppose that μ≥ρ212ρ2 or, equivalently that λ0≤ρ2μ2ρ21 holds. Then, for all λ∈(ρ2μ2ρ21,+∞), (uλ,vλ) is locally asymptotically stable with respect to System (2.1);
2. Suppose that μ<ρ212ρ2 or, equivalently that 0<ρ2μ2ρ21<λ0(<2ρ2μ2ρ21) holds. Then, the following conclusions hold:
(a) for any λ∈(ρ2μ2ρ21,λ0), (uλ,vλ) is unstable;
(b) for any λ∈(λ0,+∞), (uλ,vλ) is locally asymptotically stable with respect to System (2.1).
(c) At λ=λ0, System (2.1) undergoes a Hopf bifurcation at λ=λ0. That is, there exists an s∗>0 such that, for s∈(0,s∗), there exists (λ(s),Z(s),u(⋅,s),v(⋅,s)) so that (u(⋅,s),v(⋅,s)) is a periodic solution of (2.1) with the minimum period Z(s)→2π/√D(λ0) and (λ(s),u(⋅,s),v(⋅,s))→(λ0,uλ0,vλ0) as s→0. Moreover, the bifurcation is subcritical (i.e., the bifurcating periodic solution is unstable) and forward.
Proof. It is trivial to check that 0<λ0<2ρ2μ2ρ21 under the assumption of λ>ρ2μ2ρ21.
The linearized operator of System (2.1) evaluated at (uλ,vλ) is given by
J(λ):=((−ρ21λ+2ρ2μ2)μρ21λ−ρ1ρ2λ2ρ22μ3ρ31λ−2λ), | (2.4) |
where we use the fact that λ=ρ2μ2(1+ku2λ)/ρ21.
The eigenvalue problem of J(λ) is governed by γ2−T(λ)γ+D(λ)=0, where
T(λ):=−2ρ21λ2−ρ21μλ+2ρ2μ3ρ21λ,D(λ):=2μ(ρ21λ−ρ2μ2)ρ21. | (2.5) |
Clearly, D(λ)>0 holds for all λ>ρ2μ2ρ21. In particular, T(λ)<0 for all λ>λ0, T(λ)>0 for all 0<λ<λ0 and T(λ0)=0. Since T(2ρ2μ2ρ21)<0, we have that λ0<2ρ2μ2ρ21.
Part 1. Clearly, λ0≤ρ2μ2ρ21 is equivalent to μ≥ρ212ρ2. Since λ0≤ρ2μ2ρ21<λ, we have that T(λ)<0 for all λ>ρ2μ2ρ21. Since D(λ)>0 holds for all λ>ρ2μ2ρ21, we complete the proof.
Part 2. Case (a). Since ρ2μ2ρ21<λ<λ0, we have that T(λ)>0. Thus, (uλ,vλ) is unstable; Case (b) is similar to Part 1.
We now prove Case (c). At λ=λ0, we have that T(λ0)=0 and D(λ0)>0. Then, the eigenvalue problem has a pair of purely imaginary eigenvalues γ=±i√D(λ0). Let β(λ)±iω(λ) be the eigenvalues of the eigenvalue problem satisfying β(λ0)=0 and ω(λ0)=√D(λ0). Then, we have
β′(λ0)=12T′(λ0)=−(1+ρ21+8ρ2μ+ρ1√ρ21+16ρ2μ8ρ2μ)<0, | (2.6) |
which shows that the transversality condition holds. Thus, at λ=λ0, System (2.1) undergoes a Hopf bifurcation at λ=λ0. That is, there exists an s∗>0 such that, for s∈(0,s∗), there exists (λ(s),Z(s),u(⋅,s),v(⋅,s)) so that (u(⋅,s),v(⋅,s)) is a periodic solution of (2.1) with the minimum period Z(s)→2π/√D(λ0) and (λ(s),u(⋅,s),v(⋅,s))→(λ0,uλ0,vλ0) as s→0.
Next, we shall study the bifurcation direction and the stability of the bifurcating periodic solutions. Rewrite (2.1) in the following form:
(u′v′)=((−ρ21λ+2ρ2μ2)μρ21λ−ρ1ρ2λ2ρ22μ3ρ31λ−2λ)(uv)+(F1(λ,u,v)G1(λ,u,v)), | (2.7) |
where
F1(λ,u,v):=√kρ2√ρ21λ−ρ2μ2(ρ2μ4(−3ρ21λ+4ρ2μ2)ρ41λ2u2−2μ3ρ1uv+ρ21λ2ρ22v2)+kρ21λ−ρ2μ2(4ρ22μ5(ρ41λ2−3ρ21ρ2μ2λ+2ρ22μ4)ρ61λ3u3+2μ3λuv2)+kρ21λ−ρ2μ2(ρ2μ4(3ρ21λ−4ρ2μ2)ρ31λu2v−ρ31λ3ρ22v3)+O(|u|4,|u|3|v|),G1(λ,u,v):=√k√ρ2(ρ21λ−ρ2μ2)(ρ32μ4(−3ρ21λ+4ρ2μ2)ρ51λ2u2−2ρ22μ3ρ21uv+ρ1λ2v2)+kρ21λ−ρ2μ2(4ρ32μ5(ρ41λ2−3ρ21ρ2μ2λ+2ρ22μ4)ρ71λ3u3+2ρ2μ3λρ1uv2)+kρ21λ−ρ2μ2(ρ22μ4(3ρ21λ−4ρ2μ2)ρ41λu2v−ρ21λ3ρ2v3)+O(|u|4,|u|3|v|), |
where all of the partial derivatives are evaluated at (λ,uλ,vλ).
For λ close to λ0, we define a real 2-by-2 matrix
Y:=(10β(λ)−J11(λ)J12(λ)−ω(λ)J12(λ)), |
where
J11(λ):=(−ρ21λ+2ρ2μ2)μρ21λ,J12(λ):=−ρ1ρ2λ, |
and, for λ close to λ0,
β(λ)=12T(λ),ω(λ)=12√4D(λ)−T2(λ). |
Clearly, the matrix Y is well defined since J12(λ0)=−ρ1ρ2λ0<0 implies that, for λ close to λ0, J12(λ)≠0.
By using the linear transformation (u,v)T=Y(x,y)T, we can reduce (2.7) to the following system:
(x′y′)=(β(λ)−ω(λ)ω(λ)β(λ))(xy)+(F(λ,x,y)G(λ,x,y)), | (2.8) |
where
F(λ,x,y):=F1(λ,x,−ρ2(ρ21λβ(λ)+ρ21μλ−2ρ2μ3)ρ31λ2x+ρ2ω(λ)ρ1λy),G(λ,x,y):=ρ21λβ(λ)+ρ21μλ−2ρ2μ3ρ21λω(λ)F(λ,x,y)+ρ1λρ2ω(λ)G1(λ,x,−ρ2(ρ21λβ(λ)+ρ21μλ−2ρ2μ3)ρ31λ2x+ρ2ω(λ)ρ1λy). |
With the Taylor expansion of F(λ,u,v), we have
F(x,y)=a20x2+a11xy+a02y2+a30x3+a21x2y+a12xy2+a03y3+O(|x|4,|x|3|y|), | (2.9) |
where
a11:=√kρ2(2λ−μ)ω(λ)√ρ21λ−ρ2μ2,a02:=√kρ2ω2(λ)√ρ21λ−ρ2μ2,a20:=√kρ2(ρ41λ2(2λ−μ)2−12ρ2μ4(ρ21λ−ρ2μ2))4ρ41λ2√ρ21λ−ρ2μ2,a12:=kρ2(−6ρ21λ2+3ρ21μλ−2ρ2μ3)ω2(λ)2ρ21λ(ρ21λ−ρ2μ2),a03:=−kρ2ω3(λ)ρ21λ−ρ2μ2,a21:=−kρ2c1ω(λ)ρ21(ρ21λ−ρ2μ2)−kρ2μ2c2ω(λ)4ρ41λ2(ρ21λ−ρ2μ2).a30:=−kρ2c3λ4ρ21(ρ21λ−ρ2μ2)+kρ2μ3c48ρ41λ(ρ21λ−ρ2μ2)+kρ22μ5c54ρ61λ3(ρ21λ−ρ2μ2), | (2.10) |
where
c1:=3ρ21λ2−3ρ21μλ+2ρ2μ3,c2:=3ρ41λ2−16ρ21ρ2μ2λ+12ρ22μ4,c3:=4ρ21λ2−6ρ21μλ+3ρ21μ2+4ρ2μ3,c4:=32ρ21ρ2μλ+ρ41λ−24ρ22μ3,c5:=9ρ41λ2−30ρ21ρ2μ2λ+20ρ22μ4. |
With the Taylor expansion of G(λ,x,y), we have
G(x,y)=b20x2+b11xy+b02y2+b30x3+b21x2y+b12xy2+b03y3+O(|x|4,|x|3|y|), | (2.11) |
where
bij:=ρ21λβ(λ)+ρ21μλ−2ρ2μ3ρ21λω(λ)aij+ρ1λρ2ω(λ)eij, i,j=0,1,2⋯, | (2.12) |
with
e11:=√kρ2ρ2(2λ−μ)ω(λ)ρ1√ρ21λ−ρ2μ2,e02:=√kρ2ρ2ω2(λ)ρ1√ρ21λ−ρ2μ2,e20:=−√kρ2f14ρ51λ2√ρ21λ−ρ2μ2,e12:=−kρ22(6ρ21λ2−3ρ21μλ+2ρ2μ3)ω2(λ)2ρ31λ(ρ21λ−ρ2μ2),e03:=−kρ22ω3(λ)ρ1(ρ21λ−ρ2μ2),e21:=−kρ22f2ω(λ)4ρ51λ2(ρ21λ−ρ2μ2),e30:=−kρ22f38ρ51λ2(ρ21λ−ρ2μ2)−kρ32μ3f44ρ71λ3(ρ21λ−ρ2μ2), |
where
f1:=ρ41ρ2λ2(2λ−μ)2−12ρ22μ4(ρ21λ−ρ2μ2),f2:=3ρ41λ2(2λ−μ)2+8ρ21ρ2μ3λ(λ−2μ)+12ρ22μ6,f3:=ρ41λ2(2λ−μ)3+12ρ22μ6(2λ+5μ),f4:=ρ41λ2(2λ+μ)(2λ−9μ)−20ρ22μ6. |
Indeed, we have
b11:=√kρ2μ(2λ−μ)(ρ21λ−2ρ2μ2)2ρ21λ√ρ21λ−ρ2μ2,b02:=√kρ2μω(λ)(ρ21λ−2ρ2μ2)2ρ21λ√ρ21λ−ρ2μ2,b20:=√kρ2μ(ρ21λ−2ρ2μ2)h18ρ61λ3√ρ21λ−ρ2μ2ω(λ),b12:=−kρ2μh2ω(λ)4ρ41λ2(ρ21λ−ρ2μ2),b03:=−kρ2μ(ρ21λ−2ρ2μ2)ω2(λ)2ρ21λ(ρ21λ−ρ2μ2),b21:=−kρ2μh38ρ21λ(ρ21λ−ρ2μ2)+kρ32μ6h42ρ61λ3(ρ21λ−ρ2μ2),b30:=−kρ2μh516ρ21λ(ρ21λ−ρ2μ2)ω(λ)+kρ32μ6h62ρ41λ2(ρ21λ−ρ2μ2)ω(λ)+kρ42μ9h7ρ81λ4(ρ21λ−ρ2μ2)ω(λ), | (2.13) |
where
h1:=ρ41λ2(2λ−μ)2−12ρ2μ4(ρ21λ−ρ2μ2),h2:=3ρ41λ2(2λ−μ)−4ρ21ρ2μ2λ(3λ−2μ)−4ρ22μ5,h3:=3ρ21λ(2λ−μ)2−2ρ2μ2(12λ2+16μλ+11μ2),h4:=ρ21λ(4λ−11μ)+6ρ2μ3,h5:=ρ21λ(2λ−μ)3−4ρ2μ2(4λ3−8μλ2+11μ2λ+4μ3),h6:=2λ2−11μλ−12μ2,h7:=ρ21λ(3λ+10μ)−5ρ2μ3. |
Following page 90 of [12], we define
c1(λ0):=i2ω(λ0)(g20g11−2|g11|2−13|g02|2)+g212, | (2.14) |
where
g11:=14(Fxx+Fyy+i(Gxx+Gyy)),g02:=14(Fxx−Fyy−2Gxy+i(Gxx−Gyy+2Fxy)),g20:=14(Fxx−Fyy+2Gxy+i(Gxx−Gyy−2Fxy)),g21:=18(Fxxx+Fxyy+Gxxy+Gyyy+i(Gxxx+Gxyy−Fxxy−Fyyy)), | (2.15) |
where all quantities are to be evaluated at (λ0,uλ0,vλ0).
Substituting (2.15) into (2.14), we have
Re(c1(λ0))=k(2λ0+μ)2(μ−λ0)/(8μ2). |
In Part 2, we assume that μ<ρ212ρ2. Thus, we have that μ<3ρ212ρ2, which is equivalent to λ0<μ. Thus, we have Re(c1(λ0))>0. Together with (2.6), and in consideration of [13], it follows that the bifurcation is subcritical (i.e., the bifurcating periodic solution is unstable) and the bifurcation direction is forward.
Remark 2.2. Biological meaning: Remember that μ and λ are the decay rates of the activator u and the inhibitor v, respectively. Choosing λ as the main bifurcation parameter (by fixing μ) means that we want to know how the dynamics of System (2.1) changes as the decay rate of v changes. Indeed, one can also choose μ as the main bifurcation parameter by fixing λ. (We did not consider this case in the paper, but we expect that System (2.1) can also exhibit similar results.) Our results indicate that, as the decay rate of v changes, System (2.1) might have periodic solutions. That is, the density functions of u and v may undergo temporal oscillations with the evolution of time.
For convenience, we copy System (1.3) here:
{∂u∂t−d1Δu=ρ1u2v(1+ku2)−μu,x∈Ω,t>0,∂v∂t−d2Δv=ρ2u2v(1+ku2)−λv,x∈Ω,t>0,∂νu=∂νv=0,x∈∂Ω,t≥0,u(x,0)=u0(x)>0,v(x,0)=v0(x)>0,x∈Ω. | (3.1) |
We shall study the dynamics of System (3.1). Without loss of generality, we assume that Ω=(0,ℓπ), with ℓ>0.
In this subsection, we shall consider the stability and Turing instability of the unique positive constant equilibrium solution (uλ,vλ) with respect to the reaction-diffusion system presented as System (3.1).
We have the following results on the stability and Turing instability of (uλ,vλ).
Theorem 3.1. Let μ≥ρ212ρ2 or, equivalently, 0<λ0≤ρ2μ2ρ21 hold. Then, the following conclusions hold true:
1.For any λ∈(2ρ2μ2ρ21,+∞), (uλ,vλ) is locally asymptotically stable with respect to the diffusive system given by System (3.1); in this case, the Turing instability of (uλ,vλ) cannot feasibly occur;
2. For any λ∈(ρ2μ2ρ21,2ρ2μ2ρ21),
(a) if, in addition, d1≥(2ρ2μ2−ρ21λ)μℓ2ρ21λ holds, then (uλ,vλ) is locally asymptotically stable with respect to the diffusive system given by System (3.1); in this case, the Turing instability of (uλ,vλ) cannot feasibly occur;
(b) if, in addition, d1<(2ρ2μ2−ρ21λ)μℓ2ρ21λ holds, then we define nd1, with 1≤nd1<+∞, to be the largest positive integer such that, for any integer n∈[1,nd1],
d1<(2ρ2μ2−ρ21λ)μℓ2ρ21λn2. |
Define ˆd2:=min1≤n≤nd1d(n)2, where
d2=d(n)2:=−2λℓ2(μ(ρ21λ−ρ2μ2)+d1ρ21λn2ℓ2)n2(μ(ρ21λ−2ρ2μ2)+d1ρ21λn2ℓ2). |
Then, for any d2<ˆd2, (uλ,vλ) is locally asymptotically stable with respect to the diffusive system given by System (3.1). If d2>ˆd2 holds, then (uλ,vλ) undergoes Turing instability in the reaction-diffusion system given by System (3.1).
Proof. The proof is moved to the Appendix.
Similarly, we have the following results:
Theorem 3.2. Let μ<ρ212ρ2 or, equivalently, ρ2μ2ρ21<λ0<2ρ2μ2ρ21 hold. Then, the following conclusions hold true:
1. For any λ∈(2ρ2μ2ρ21,+∞), (uλ,vλ) is locally asymptotically stable with respect to the diffusive system given by System (3.1); in this case, the Turing instability of (uλ,vλ) cannot feasibly occur;
2. For any λ∈(λ0,2ρ2μ2ρ21),
(a) if, in addition, d1≥(2ρ2μ2−ρ21λ)μℓ2ρ21λ holds, then (uλ,vλ) is locally asymptotically stable with respect to the diffusive system given by System (3.1).
(b) if, in addition, d1<(2ρ2μ2−ρ21λ)μℓ2ρ21λ holds, then, for any d2<ˆd2, (uλ,vλ) is locally asymptotically stable with respect to the diffusive system given by System (3.1). If d2>ˆd2 holds, then (uλ,vλ) undergoes Turing instability in the reaction-diffusion system given by System (3.1).
3. For any λ∈(ρ2μ2ρ21,λ0), (uλ,vλ) is unstable with respect to the reaction-diffusion system given by System (3.1); in this case, the Turing instability of (uλ,vλ) cannot feasibly occur.
Remark 3.3. If one of the following cases holds, i.e., (1) of Theorem 3.1, 2(a) of Theorem 3.1, (1) of Theorem 3.2 or 2(a) of Theorem 3.2, then (uλ,vλ) is locally asymptotically stable with respect to the reaction-diffusion system given by System (3.1). In this case, the Turing instability of (uλ,vλ) will never occur. This means that, when the decay rate of v and the diffusion rates are chosen in certain ranges, Turing patterns are unlikely to occur; If one of the following cases holds, i.e., 2(b) of Theorem 3.1 or 2(b) of Theorem 3.2, then (uλ,vλ) will undergo Turing instability in the reaction-diffusion system given by System (3.1). This means that, when the decay rate of v and the diffusion rates are chosen in certain ranges, Turing patterns can be expected.
In this subsection, we shall consider the occurrence of spatially non-homogeneous periodic solutions bifurcating from Hopf bifurcations.
When λ0≤ρ2μ2ρ21, we have that T(λ)≤0 (and, hence, Tn(λ)≤0,Tn(λ) is defined in (A.1)) for λ>ρ2μ2ρ21. Thus, to expect the spatially non-homogeneous periodic solution bifurcating from Hopf bifurcations, we need to concentrate on the case when T(λ)≥0, or, equivalently,
λ0>ρ2μ2ρ21(or equivalently μ<ρ212ρ2)and λ∈(ρ2μ2ρ21,λ0]. | (3.2) |
In this case, under (3.2), (uλ,vλ) is unstable with respect to the reaction-diffusion system given by System (3.1).
Following [14], we shall identify the Hopf bifurcation values, denoted by λH, which satisfy the following: there exists n∈N0 such that
Tn(λH)=0,Dn(λH)>0, and Tj(λH)≠0,Dj(λH)≠0 for j≠n; | (3.3) |
and, for the unique pair of complex eigenvalues near the imaginary axis α(λ)±iω(λ),
α′(λH)≠0. | (3.4) |
Since λ0>ρ2μ2ρ21 and λ∈(ρ2μ2ρ21,λ0], we have that T(λ)≥0. One can check that T′(λ) is decreasing in (ρ2μ2ρ21,λ0), given that T(λ0)=0 and limλ→0+T(λ)=+∞. Define
T∗=T(ρ2μ2ρ21)=μ(ρ21−2ρ2μ)ρ21>T(λ0)=0 and ℓn=n√d1+d2T∗,with n∈N0∖{0}. | (3.5) |
Then, for any ℓ>ℓ1, there exists an n∈N0∖{0} such that ℓn<ℓ<ℓn+1. Then, for any 1≤j≤n, T(λ)=(d1+d2)j2ℓ2 has a unique root, denoted by λHj, satisfying
ρ2μ2ρ21<λHn<⋯<λHj<⋯<λH1<λ0, | (3.6) |
and if n≠m, λHn≠λHm.
It remains necessary to derive precise conditions so that, for all λ∈(ρ2μ2ρ21,λ0), Dn(λ)>0 for all n∈N0. In fact, since
Dn(λ)=2μ(ρ21λ−ρ2μ2)ρ21+((ρ21λ−2ρ2μ2)μρ21λd2+2d1λ)n2ℓ2+d1d2n4ℓ4, |
we can regard Dn(λ) as the quadratic function of n2/ℓ2. Define
F(x)=d1d2x2+((ρ21λ−2ρ2μ2)μρ21λd2+2d1λ)x+2μ(ρ21λ−ρ2μ2)ρ21. |
Then, we have that Dn(λ)=F(n2/ℓ2). The discriminant of F(x) is given by
ΔF:=μ2d22ρ41λ2(ρ21λ−2ρ2μ2)2+4d1λ(d1λ−d2μ). | (3.7) |
By (3.7), to let ΔF<0, we need first to assume that d1λ−d2μ<0 or, equivalently that d2d1>λμ. Then, solving ΔF<0 from (3.7), we have
2ρ21λ2(ρ21λ−2μ√ρ2(ρ21λ−ρ2μ2))μ(ρ21λ−2ρ2μ2)2<d2d1<2ρ21λ2(ρ21λ+2μ√ρ2(ρ21λ−ρ2μ2))μ(ρ21λ−2ρ2μ2)2. | (3.8) |
If 8−4√3<ρ21λρ2μ2<8+4√3 holds, then we have
2ρ21λ2(ρ21λ−2μ√ρ2(ρ21λ−ρ2μ2))μ(ρ21λ−2ρ2μ2)2<λμ. | (3.9) |
Then, to let ΔF<0, we need
d2d1∈(λμ,2ρ21λ2(ρ21λ+2μ√ρ2(ρ21λ−ρ2μ2))μ(ρ21λ−2ρ2μ2)2). | (3.10) |
If ρ21λρ2μ2>8+4√3 or 1<ρ21λρ2μ2<8−4√3 holds, then we have
λμ<2ρ21λ2(ρ21λ−2μ√ρ2(ρ21λ−ρ2μ2))μ(ρ21λ−2ρ2μ2)2. | (3.11) |
Then, to ΔF<0, we need
d2d1∈(2ρ21λ2(ρ21λ−2μ√ρ2(ρ21λ−ρ2μ2))μ(ρ21λ−2ρ2μ2)2,2ρ21λ2(ρ21λ+2μ√ρ2(ρ21λ−ρ2μ2))μ(ρ21λ−2ρ2μ2)2). | (3.12) |
So far, by Theorem 2.1 of [14], we are in the position to state the following results on Hopf bifurcations:
Theorem 3.4. Suppose that (3.2) holds and that
{(3.10)holdsif ρ21λρ2μ2∈(8−4√3,8+4√3),(3.12)holdsif ρ21λρ2μ2∈(8+4√3,+∞)∪(1,8−4√3). | (3.13) |
Then, for any ℓ>ℓ1 (ℓn is defined in (3.5)), there exists n points λHj(ℓ), 1≤j≤n, satisfying
λHn<⋯<λHj<⋯<λH1<λ0 |
such that the reaction-diffusion system undergoes a Hopf bifurcation at λ=λHj or λ=λ0 and the following holds true:
1. The bifurcating periodic solutions from λ=λH0 are spatially homogeneous and unstable, which coincides with the periodic solution of the corresponding ODE system;
2. The bifurcating periodic solutions from λ=λHj are spatially non-homogeneous and unstable.
Remark 3.5. 1. For the bifurcating spatially homogeneous periodic solutions, it is always unstable in the reaction-diffusion system given by System (3.1) since it is unstable in the corresponding ODE system given by System (2.1). On the other hand, the bifurcating spatially non-homogeneous periodic solutions from λ=λHj are always unstable, since, under (3.2), (uλ,vλ) is unstable with respect to the reaction-diffusion system given by System (3.1).
2. Biological meaning: Our results showed that, with the inclusion of the spatial diffusion, as the decay rate of v changes, the reaction-diffusion system given by System (3.1) might not only undergo temporal oscillations, but also spatiotemporal oscillations with the evolution of time. That is, the densities of the activator u and the inhibitor v will oscillate with respect to t, but they will also depend on the spatial variable x (in the case of spatially non-homogeneous periodic solutions).
In this subsection, we shall use the steady-state bifurcation theory to show the existence of the bifurcating non-constant positive equilibrium solutions.
Following [14], we identify steady-state bifurcation values, denoted by λS, which satisfy the following: there exists n∈N0 such that
Dn(λS)=0,Tn(λS)≠0 and Tj(λS)≠0,Dj(λS)≠0 for j≠n, | (3.14) |
and that
ddλDn(λS)≠0. | (3.15) |
By Theorems 3.1 and 3.2, for any λ∈(2ρ2μ2ρ21,+∞), (uλ,vλ) is locally asymptotically stable with respect to the diffusive system given by System (3.1). Hence, any potential bifurcation point λS must be in the interval (ρ2μ2ρ21,2ρ2μ2ρ21). We shall study the steady-state bifurcation points in this interval.
We rewrite Dn(λ) in the form of
Dn(λ)=μC(λ)−d2A(λ)p+d1d2p2, |
where p:=n2/ℓ2 and
C(λ):=2(ρ21λ−ρ2μ2)ρ21, A(λ):=−2d1ρ21λ2+d2ρ21μλ−2d2ρ2μ3d2ρ21λ. |
Solving p from Dn(λ)=0, we have
p=p±(λ):=d2A(λ)±√C(λ)(d22h(λ)−4d1d2μ)2d1d2, | (3.16) |
where
h(λ):=A(λ)2C(λ)=(2d1ρ21λ2+d2ρ21μλ−2d2ρ2μ3)22d22ρ21λ2(ρ21λ−ρ2μ2). |
By direct calculation, we have
h′(λ):=g1(λ)g2(λ)2d22ρ21λ3(ρ21λ−ρ2μ2)2, | (3.17) |
where
g1(λ):=2d1ρ21λ2+d2ρ21μλ−2d2ρ2μ3,g2(λ):=2d1ρ41λ3−(4d1ρ21ρ2μ2+d2ρ41μ)λ2+6d2ρ21ρ2μ3λ−4d2ρ22μ5. |
Clearly, g1(λ)=0 has a unique positive root, denoted by λg1. In particular, g1(λ)<0 for λ∈(0,λg1), while g1(λ)>0 for λ∈(λg1,+∞). On the other hand, since g2(0)=−4d2ρ22μ5<0 and g2(+∞)=+∞, g2(λ)=0 has at least one positive root in (0,+∞). In particular, g2(λ)=0 may have one unique positive root or three positive roots in (0,+∞).
In what follows, for our particular interests, we assume that g2(λ)=0 has a unique positive root in (0,+∞). (The case of g2(λ)=0 having more than one positive root in (0,+∞) is much more complicated and will be considered in our future investigations.)
We have the following results on the properties of h(λ) in the interval (ρ2μ2ρ21,2ρ2μ2ρ21):
Lemma 3.6. Let λg1 be the unique positive root of g1(λ)=0. Assume, in addition that, g2(λ)=0 also has a unique root, denoted by λg2 in (0,+∞). Then, the following conclusions hold true:
1. Suppose that 0<d2d1<2ρ2μρ21 holds. Then,
(a) λg1<ρ2μ2ρ21<λg2<2ρ2μ2ρ21;
(b) for any λ∈(ρ2μ2ρ21,2ρ2μ2ρ21), h(λ)>0 and
h(ρ2μ2ρ21)=+∞,h(2ρ2μ2ρ21)=8d21ρ2μ2d22ρ21. |
(c) h(λ) is decreasing in λ∈(ρ2μ2ρ21,λg2), while it is increasing in λ∈(λg2,2ρ2μ2ρ21); at λ=λg2, h(λ) attains its positive minimum value h(λg2).
2. Suppose that d2d1>2ρ2μρ21 holds. Then,
(a) λg2<ρ2μ2ρ21<λg1<2ρ2μ2ρ21;
(b) for any λ∈(ρ2μ2ρ21,2ρ2μ2ρ21), h(λ)>0 and
h(ρ2μ2ρ21)=+∞,h(2ρ2μ2ρ21)=8d21ρ2μ2d22ρ21. |
(c) h(λ) is decreasing in λ∈(ρ2μ2ρ21,λg1), while it is increasing in λ∈(λg1,2ρ2μ2ρ21); at λ=λg1, h(λ) attains its positive minimum value h(λg1).
Proof. The proof is moved to the Appendix.
Define
ΣD(λ):=C(λ)(d22h(λ)−4d1d2μ). |
We have the following results on the sign of ΣD(λ) in the interval (ρ2μ2ρ21,2ρ2μ2ρ21):
Lemma 3.7. Let λg1 be the unique positive root of g1(λ)=0. Assume, in addition that, g2(λ)=0 also has a unique root, denoted by λg2 in (0,+∞). Then, the following conclusions hold true:
1. Suppose that 0<d2d1<2ρ2μρ21 holds.
(a) if 4μh(λg2)<d2d1<2ρ2μρ21 holds, then, for all λ∈(ρ2μ2ρ21,2ρ2μ2ρ21), ΣD(λ)>0;
(b) if d2d1<min(4μh(λg2),2ρ2μρ21) holds, then there exist λ_,¯λ∈(ρ2μ2ρ21,2ρ2μ2ρ21), with λ_<¯λ, such that
h(λ_)4μ=h(¯λ)4μ=d1d2, |
and, for any λ∈(λ_,¯λ), ΣD(λ)<0, while, for λ∈(ρ2μ2ρ21,λ_]∪[¯λ,2ρ2μ2ρ21), ΣD(λ)>0.
2. Suppose that d2d1>2ρ2μρ21 holds. Then, there exists a unique point λ∗∈(ρ2μ2ρ21,2ρ2μ2ρ21) such that h(λ∗)=4μ(d1d2). In particular, ΣD(λ)<0 for any λ∈(λ∗,2ρ2μ2ρ21), while ΣD(λ)>0 for any λ∈(ρ2μ2ρ21,λ∗].
Proof. The proof is moved to the Appendix.
For clarity of our later discussions, we divide our discussions into the following cases:
Case 1: 4μh(λg2)<d2d1<2ρ2μρ21 and λ∈(ρ2μ2ρ21,2ρ2μ2ρ21);
Case 2: d2d1<min(4μh(λg2),2ρ2μρ21) and λ∈(ρ2μ2ρ21,λ_]∪[¯λ,2ρ2μ2ρ21);
Case 3: d2d1>2ρ2μρ21 and λ∈(ρ2μ2ρ21,λ∗], where λ∗ is defined in Lemma 3.7.
Theorem 3.8. Suppose that either Case 1 or 2 holds. Then, no steady-state bifurcation around (uλ,vλ) occurs.
Proof. Suppose that either Case 1 or 2 holds. Then, p±(λ) is well defined. Clearly, d2d1<2ρ2μρ21. Then, by (a) of Lemma 3.6, we have that λg1<ρ2μ2ρ21<λg2<2ρ2μ2ρ21. One can check that A(λ) is decreasing since
A′(λ):=−2d1ρ21λ+2d2ρ2μ3d2ρ21λ2. | (3.18) |
Since A(λg1)=0, it follows that, for all λ∈(ρ2μ2ρ21,2ρ2μ2ρ21), A(λ)<0. Thus, p±(λ)<0 whenever they are well defined. In this case, no steady-state bifurcation around (uλ,vλ) occurs.
Lemma 3.9. Suppose that Case 3 holds. Then, there exists λ∗∈(ρ2μ2ρ21,2ρ2μ2ρ21), with λ∗<λg1, such that p+(λ) is decreasing while p−(λ) is increasing in (ρ2μ2ρ21,λ∗]. In particular,
0<p−(ρ2μ2ρ21)<p−(λ)<p−(λ∗)=p+(λ∗)<p+(λ)<p+(ρ2μ2ρ21)<+∞, | (3.19) |
and limλ→λ∗p′−(λ)=+∞ and limλ→λ∗p′+(λ)=−∞.
Proof. The proof is moved to the Appendix.
Therefore, for any n>0, if p−(ρ2μ2ρ21)<n2/ℓ2<p+(ρ2μ2ρ21), then there exists λSn∈(ρ2μ2ρ21,λ∗) such that p−(λSn)=0 or p+(λSn)=0.
Define
ℓ+n:=n/√p+(ρ2μ2ρ21), ℓ−n:=n/√p−(ρ2μ2ρ21). | (3.20) |
Then, for any ℓ∈(ℓ+n,ℓ−n), there exists λSn such that Dn(λSn)=0. These points λSn are potential steady-state points. As remarked in [14], however, it is possible that, for some i<j, p−(λSi)=p+(λSj). In this case, for λ=λSi=λSj, 0 is not a simple eigenvalue of L(λ), and we shall consider bifurcations at such points. Let LE be the set of such points.
Summarizing the aforementioned observations, by Theorem 3.10 of [14], we have the following results on the existence of steady-state bifurcations:
Theorem 3.10. Suppose that Case 3 holds. Let ℓ+n and ℓ−n be defined in (3.20). If, for some n∈N, ℓ∈(ℓ+n,ℓ−n)∖{LE} and there exists a point λSn∈(ρ2μ2ρ21,λ∗) such that p+(λSn)=0 or p−(λSn)=0. Then, there is a smooth curve of positive solutions of the reaction-diffusion system bifurcating from (λ,u,v)=(λSn,uλSn,vλSn).
Remark 3.11. Biological meaning: Our results showed that, for any ℓ>0 (the length of the spatial domain), there exists an n>0 such that, for a suitable λ (the decay rate of the inhibitor v) and suitable diffusion rates d1 and d2, the reaction-diffusion system given by System (3.1) might have positive non-constant steady-state solutions with the eigen-mode cos(nx/ℓ); That is, the densities of the activator u and the inhibitor v have a non-uniform spatial distribution in Ω. From the viewpoint of pattern dynamics, in this case, System (3.1) will undergo spatial patterns which are different from Turing patterns.
In this study, we were mainly concerned with the spatiotemporal patterns and multiple bifurcations of a reaction-diffusion model for hair follicle spacing.
First, we consider the stability and instability of the equilibrium solution of the ODE system. In particular, by using the center manifold theory, normal form methods, as well as the standard Hopf bifurcation theory, we were able to prove the existence of the Hopf bifurcating periodic solutions bifurcating from the equilibrium solution. By calculating the first Lyapunov coefficient, we found that the Hopf bifurcating periodic solutions are always unstable. This is one of the novel points of this paper. Since the bifurcating periodic solutions are unstable in ODEs, they will never undergo Turing instability (see [9,15,16] for more details on the Turing instability of periodic solutions) in the reaction-diffusion equations given by System (3.1). This is different from the equilibrium solutions, which may experience Turing instability under certain suitable conditions on the diffusion rates (d1 and d2) and the decay rate of v. This was shown in the analysis of the reaction-diffusion system.
Second, we studied the stability and instability of the constant equilibrium solution in the reaction-diffusion system. In particular, by using Hopf bifurcation theory and steady-state bifurcation theory, we were capable of showing the existence of spatially non-homogeneous Hopf bifurcating periodic solutions, as well as the non-constant bifurcating steady-state solutions for the reaction-diffusion equations. Moreover, Turing instability of the constant equilibrium solution was investigated in detail. If Turing instability of the constant equilibrium solutions occurs, then Turing patterns emerge.
However, in application, for many reaction-diffusion equations, Turing-Hopf bifurcations can also be observed. In this study, we did not consider the existence of Turing-Hopf bifurcations. Indeed, the analytical analysis of Turing-Hopf bifurcation will be much more difficult than that of either Turing bifurcation or Hopf bifurcation. We will consider Turing-Hopf bifurcations of this particular reaction-diffusion system in our future investigations.
Regarding the numerical simulations, we would like to mention that we did not include numerical simulations in the paper. The reasons are as follows: from our analytical analysis, it was found that the Hopf bifurating periodic solutions are always unstable, so it is hard to simulate numerically; on the other hand, for the steady-state bifurcations, under our conditions, the constant equilibrium solutions and bifurcating steady-state solutions are unstable. Again, it is hard to simulate numerically.
Finally, we ought to remark that our results in the current paper tend to be much more analytical. Although it will make contributions to the field of bifurcation theory with applications, as well as to the field of mathematical biology, further work needs to be done to use our analytical results to understand the factors that affect the hair follicle spacing. We shall study it in our future investigations.
Zhili Zhang, Aying Wan and Hongyan Lin were partially supported by the National Natural Science Foundation of China (12061033).
The authors declare that there is no conflict of interest.
1. Proof of Theorem 3.1. To understand the stability and instability of (uλ,vλ), we need to know the linearized operator of System (3.1) evaluated at (uλ,vλ), which is given by
L(λ):=(d1∂2∂x2+(−ρ21λ+2ρ2μ2)μρ21λ−ρ1ρ2λ2ρ22μ3ρ31λd2∂2∂x2−2λ). |
It is well known that (see [14,17,18,19,20]) the eigenvalue problem
−φ″ |
has eigenvalues (), with corresponding eigenfunctions .
Let
be an eigenfunction for with the eigenvalue . Then, we have
where
Then, the eigenvalues of are given by the eigenvalues of for .
Indeed, the characteristic equation of is
where
(A.1) |
Note that or, equivalently, holds; we have that for any and since . We now consider the sign of , with :
Suppose that holds. Then, for any , we have that . Thus, in this case, is locally asymptotically stable with respect to the diffusive system given by System (3.1). By Theorem 2.1, is also locally asymptotically stable with respect to the ODE system. Thus, Turing instability of does not occur.
Suppose that holds.
If, in addition, holds, then, for any , we have
(A.2) |
In this case, is locally asymptotically stable with respect to the diffusive system given by System (3.1). Again, Turing instability of does not occur.
If, in addition holds, then, by the definition of , for any integer , we have that . Moreover, for any , with . In this case, is locally asymptotically stable with respect to the diffusive system given by System (3.1). If holds, then there exists at least an such that . In this case, is unstable with respect to the diffusive system given by System (3.1). Thus, undergoes Turing instability in the reaction-diffusion system given by System (3.1). We thus complete the proof.
2. Proof of Lemma 3.6. We only prove the first part, since the proof of the second part is similar. Clearly, is increasing in . By direct calculation, we have
(A.3) |
If holds, then, by (A.3), and . This implies that . In this case, the monotonic properties of can be obtained. We thus complete the proof.
3. Proof of Lemma 3.7. We only prove Part 1, since the other parts can be proved similarly. Assume that holds. Then, we have that .
If, in addition, or, equivalently that holds, then, for any , .
If, in addition, holds, then there exist , with , such that
and for any , while for . We thus complete the proof.
4. Proof of Lemma 3.9. Suppose that Case 3 holds. By Lemma 3.7, is well defined in , where . Thus, in this interval, .
Since is decreasing, is increasing in and ; by
(A.4) |
we know that is decreasing in .
On the other hand, by , we know that is increasing in , since is decreasing and is increasing. By direct calculation, it follows that
(A.5) |
Since , we can obtain that and .
Noticing that , we have
Since is increasing and is decreasing, we have that for all . In summary, we have (3.19). We thus complete the proof.
[1] |
S. Sick, S. Reinker, J. Timmer, T. Schlake, WNT and DKK determine hair follicle spacing through a reaction-diffusion mechanism, Science, 314 (2006), 1447–1450. https://doi.org/10.1126/science.1130088 doi: 10.1126/science.1130088
![]() |
[2] |
P. Rashkov, Remarks on pattern formation in a model for hair follicle spacing, Disct. Cont. Dyns. Syst. Ser. B, 20 (2015), 1555–1572. https://doi.org/10.3934/dcdsb.2015.20.1555 doi: 10.3934/dcdsb.2015.20.1555
![]() |
[3] |
P. Rashkov, Regular and discontinuous solutions in a reaction-diffusion model for hair follicle spacing, Biomath, 3 (2014), 1–12. https://doi.org/10.11145/j.biomath.2014.11.111 doi: 10.11145/j.biomath.2014.11.111
![]() |
[4] |
F. Veerman, A. Doelman, Pulses in a Gierer-Meinhardt equation with a slow nonlinearity, SIAM J. Dyn. Syst., 12 (2013), 28–60. https://doi.org/10.1137/120878574 doi: 10.1137/120878574
![]() |
[5] |
J. R. Mooney, Steady states of a reaction-diffusion system on the off-centre annulus, SIAM J. Appl. Math., 44 (1984), 745–761. https://doi.org/10.1137/0144053 doi: 10.1137/0144053
![]() |
[6] |
B. N. Nagorcka, Evidence for a reaction-diffusion system as a mechanism controlling mammalian hair growth, BioSystems, 16 (1984), 323–332. https://doi.org/10.1016/0303-2647(83)90015-1 doi: 10.1016/0303-2647(83)90015-1
![]() |
[7] |
B. N. Nagorcka, J. R. Mooney, The role of a reaction-diffusion system in the formation of hair fibres, J. Theor. Biol., 98 (1982), 575–607. https://doi.org/10.1016/0022-5193(82)90139-4 doi: 10.1016/0022-5193(82)90139-4
![]() |
[8] |
F. Yi, H. Zhang, A. Cherif, W. Zhang, Spatiotemporal patterns of a homogenous diffusive system modeling hair growth: Global stability and multiple bifurcation analysis, Comm. Pure. Appl. Anal., 13 (2014), 347–369. https://doi.org/10.3934/cpaa.2014.13.347 doi: 10.3934/cpaa.2014.13.347
![]() |
[9] |
Y. Yang, X. Ju, Diffusion-driven instability of the periodic solutions for a diffusive system modeling mammalian hair growth, Nonlinear Dyn., 111 (2023), 5799–5815. https://doi.org/10.1007/s11071-022-08114-x doi: 10.1007/s11071-022-08114-x
![]() |
[10] |
E. Conway, D. Hoff, J. Smoller, Large time behavior of solutions of systems of nonlinear reaction-diffusion equations, SIAM J. Appl. Math., 35 (1978), 1–16. https://doi.org/10.1137/0135001 doi: 10.1137/0135001
![]() |
[11] |
A. M. Turing, The chemical basis of morphogenesis, Phil. Trans. Royal Soc. London, B237 (1952), 37–72. https://doi.org/10.2307/92463 doi: 10.2307/92463
![]() |
[12] | B. Hassard, N. Kazarinoff, Y. Wan, Theory and Application of Hopf Bifurcation, Cambridge University Press, Cambridge, 1981. |
[13] | S. Wiggins, Introduction to Applied Nonlinear Dynamical Systems and Chaos, Springer-Verlag, New York, 1991. |
[14] |
F. Yi, J. Wei, J. Shi, Bifurcation and spatiotemporal patterns in a homogenous diffusive predator-prey system, J. Differ. Equations, 246 (2009), 1944–1977. https://doi.org/10.1016/j.jde.2008.10.024 doi: 10.1016/j.jde.2008.10.024
![]() |
[15] |
M. Wang, F. Yi, On the dynamics of the diffusive Field-Noyes model for the Belousov-Zhabotinskii reaction, J. Differ. Equations, 318 (2022), 443–479. https://doi.org/10.1016/j.jde.2022.02.031 doi: 10.1016/j.jde.2022.02.031
![]() |
[16] |
F. Yi, Turing instability of the periodic solutions for reaction-diffusion systems with cross-diffusion and the patch model with cross-diffusion-like coupling, J. Differ. Equations, 281 (2021), 379–410. https://doi.org/10.1016/j.jde.2021.02.006 doi: 10.1016/j.jde.2021.02.006
![]() |
[17] |
J. Jang, W. Ni, M. Tang, Global bifurcation and structure of Turing patterns in the 1-D Lengyel-Epstein model, J. Dyn. Differ. Equations, 16 (2004), 297–320. https://doi.org/10.1007/s10884-004-2782-x doi: 10.1007/s10884-004-2782-x
![]() |
[18] |
W. Ni, M. Tang, Turing patterns in the Lengyel-Epstein system for the CIMA reactions, Trans. Am. Math. Soc., 357 (2005), 3953–3969. https://doi.org/10.1090/S0002-9947-05-04010-9 doi: 10.1090/S0002-9947-05-04010-9
![]() |
[19] |
R. Peng, F. Yi, X. Zhao, Spatiotemporal patterns in a reaction-diffusion model with the Degn-Harrison reaction scheme, J. Differ. Equations, 254 (2013), 2465–2498. https://doi.org/10.1016/j.jde.2012.12.009 doi: 10.1016/j.jde.2012.12.009
![]() |
[20] |
F. Yi, S. Liu, N. Tuncer, Spatiotemporal patterns of a reaction-diffusion Substrate-Inhibition Seelig model, J. Dyn. Differ. Equations, 29 (2017), 219–241. https://doi.org/10.1007/s10884-015-9444-z doi: 10.1007/s10884-015-9444-z
![]() |