Processing math: 100%
Research article Special Issues

Orientable vertex transitive embeddings of Kp

  • Received: 07 February 2023 Revised: 27 March 2023 Accepted: 07 April 2023 Published: 23 April 2023
  • MSC : 20B15, 20B30, 05C25, 05C30

  • In [J. Combin. Theory Ser. B, 99 (2009), 447-454)], Li characterized the classification of vertex-transitive embeddings of complete graphs, and proposed how to enumerate such maps. In this paper, we study the counting problem of orientable vertex-transitive embeddings of Kp, where p5 is a prime. Moreover, we obtain the number of non-isomorphic orientable vertex-transitive complete maps with p vertices.

    Citation: Xue Yu, Qingshan Zhang. Orientable vertex transitive embeddings of Kp[J]. AIMS Mathematics, 2023, 8(7): 15024-15034. doi: 10.3934/math.2023767

    Related Papers:

    [1] Naveed Iqbal, Ranchao Wu, Yeliz Karaca, Rasool Shah, Wajaree Weera . Pattern dynamics and Turing instability induced by self-super-cross-diffusive predator-prey model via amplitude equations. AIMS Mathematics, 2023, 8(2): 2940-2960. doi: 10.3934/math.2023153
    [2] Wei Li, Qingkai Xu, Xingjian Wang, Chunrui Zhang . Dynamics analysis of spatiotemporal discrete predator-prey model based on coupled map lattices. AIMS Mathematics, 2025, 10(1): 1248-1299. doi: 10.3934/math.2025059
    [3] Anna Sun, Ranchao Wu, Mengxin Chen . Turing-Hopf bifurcation analysis in a diffusive Gierer-Meinhardt model. AIMS Mathematics, 2021, 6(2): 1920-1942. doi: 10.3934/math.2021117
    [4] Jing Zhang, Shengmao Fu . Hopf bifurcation and Turing pattern of a diffusive Rosenzweig-MacArthur model with fear factor. AIMS Mathematics, 2024, 9(11): 32514-32551. doi: 10.3934/math.20241558
    [5] Xiao-Long Gao, Hao-Lu Zhang, Xiao-Yu Li . Research on pattern dynamics of a class of predator-prey model with interval biological coefficients for capture. AIMS Mathematics, 2024, 9(7): 18506-18527. doi: 10.3934/math.2024901
    [6] Teekam Singh, Ramu Dubey, Vishnu Narayan Mishra . Spatial dynamics of predator-prey system with hunting cooperation in predators and type I functional response. AIMS Mathematics, 2020, 5(1): 673-684. doi: 10.3934/math.2020045
    [7] Ting Gao, Xinyou Meng . Stability and Hopf bifurcation of a delayed diffusive phytoplankton-zooplankton-fish model with refuge and two functional responses. AIMS Mathematics, 2023, 8(4): 8867-8901. doi: 10.3934/math.2023445
    [8] Jaywan Chung, Yong-Jung Kim, Ohsang Kwon, Chang-Wook Yoon . Biological advection and cross-diffusion with parameter regimes. AIMS Mathematics, 2019, 4(6): 1721-1744. doi: 10.3934/math.2019.6.1721
    [9] Shuo Xu, Chunrui Zhang . Spatiotemporal patterns induced by cross-diffusion on vegetation model. AIMS Mathematics, 2022, 7(8): 14076-14098. doi: 10.3934/math.2022776
    [10] Pan Xue, Cuiping Ren . Spatial patterns for a predator-prey system with Beddington-DeAngelis functional response and fractional cross-diffusion. AIMS Mathematics, 2023, 8(8): 19413-19426. doi: 10.3934/math.2023990
  • In [J. Combin. Theory Ser. B, 99 (2009), 447-454)], Li characterized the classification of vertex-transitive embeddings of complete graphs, and proposed how to enumerate such maps. In this paper, we study the counting problem of orientable vertex-transitive embeddings of Kp, where p5 is a prime. Moreover, we obtain the number of non-isomorphic orientable vertex-transitive complete maps with p vertices.



    Since 1946, biologist Crombic proved the stability effect through experiments [1,2] and more and more scholars analyzed the refuge effect on the population model [3,4,5,6,7,8,9], mainly focused on the self-diffusion effect on dynamic behavior of the population system. In addition to the effect of self-diffusion, cross-diffusion also plays an important role during the population pattern formation. About the predator-prey systems with diffusion terms, many scholars have studied the Turning instability and Hopf bifurcation of its constant equilibrium [10,11,12,13,14,15,16,17]. At present, for the reaction-diffusion predator-prey system, most literatures [18,19,20,21,22,23,24,25] focus on Turing instability of the constant equilibrium, but there are few research results on the stability of the periodic solutions. Therefore, it is significant to study the Turing pattern formation of Hopf bifurcating periodic solutions for the cross-diffusion population model with prey refuge and the Holling Ⅲ functional response.

    In 2015, Yang et al. [9] studied a diffusive prey-predator system in Holling type Ⅲ with a prey refuge:

    {u(x,t)t=D1Δu+auru2α(1m)2u2vβ2+(1m)2u2, xΩ,t>0,v(x,t)t=D2Δvcv+kα(1m)2u2vβ2+(1m)2u2, xΩ,t>0,νu=νv=0, xΩ,t>0. (1.1)

    Here, u,v indicates the quantity of prey and predator respectively; α,β,a,r,c,k are all positive; a is the intrinsic growth rate of the prey; a/r represents the maximum carrying capacity of the environment on the prey; c is the mortality rate of the predator; k represents the conversion rate after the predator eating the prey; m[0,1) indicates the refuge coefficient, i.e., the proportion of the protected prey. Only (1m)u can be caught by the predator. In the real world, the mobility of each species is affected not only by itself but also by the density of other species. Therefore, on the basis of (1.1), we introduce the cross-diffusion terms and establish the population model as follows:

    {u(x,t)t=D11Δu+D12Δv+auru2α(1m)2u2vβ2+(1m)2u2, xΩ,t>0,v(x,t)t=D21Δu+D22Δvcv+kα(1m)2u2vβ2+(1m)2u2, xΩ,t>0,νu=νv=0, xΩ,t>0, (1.2)

    where Ω=(0,lπ) is a bounded domain with smooth boundary Ω in Rn and D11,D22 and D12,D21 are the self-diffusivity and cross-diffusivity of u and v. We assume that the diffusion coefficients satisfy D11D22D12D21>0.

    The organizational structure of the rest is as follows: In section two, we study the stability of Hopf bifurcating periodic solutions for the ordinary differential population model and derive the first derivative formula of the periodic function for the corresponding perturbed model. In section three, we give the conditions of Turing patterns occurring at Hopf bifurcating periodic solutions in the reaction-diffusion population system. In section four, we give a brief conclusion. Finally, the relevant conclusions are verified by numerical simulations.

    In order to research conveniently, we nondimensionalize model (1.2). Let ˆu=uβ,ˆv=vkβ,ˆt=at, and we still replace ˆu,ˆv,ˆt with u,v,t, then model (1.2) becomes

    {ut=d11Δu+d12Δv+upu2s(1m)2u2v1+(1m)2u2, xΩ,t>0,vt=d21Δu+d22Δvθv+s(1m)2u2v1+(1m)2u2, xΩ,t>0,νu=νv=0, xΩ,t>0, (2.1)

    where, d11=D11a,d12=D12a,d21=D21a,d22=D22a,θ=ca and p=rβa,s=kαa.

    The ordinary differential equations corresponding to the reaction-diffusion population model (2.1) are

    {dudt=upu2s(1m)2u2v1+(1m)2u2,t>0,dvdt=θv+s(1m)2u2v1+(1m)2u2, t>0. (2.2)

    By calculation, four equilibria of model (2.2) are P0=(0,0),P1=(1/p,0),P+=(κ,vκ),P=(u,v) with

    κ=11mθsθ,vκ=κθ(1pκ),u=κ,v=(1+pκ)(1+(1m)2κ2)s(1m)2κ.

    Clearly, the equilibrium P=(u,v) has no biological significance, so we do not study its dynamic behavior. Let's make the following assumptions:

    (A1)s>θ,0κ<1p;

    (A2)s<2θ,2θs2θp<κ<1p;

    (A3)s2θ,θsθ<κ<1p;

    (A4)θ<s<2θ,θsθ<κ<2θs2θp.

    Theorem 2.1. Let κ0=2θs2θp and assume that (A1) satisfies. The following results are true for model (2.2).

    (1) If (A2) (or(A3)) holds, then the positive equilibrium P+=(κ,vκ) is locally asymptotically stable. If (A4) holds, then the positive equilibrium P+=(κ,vκ) is unstable.

    (2) If (A3) holds, the positive equilibrium P+=(κ,vκ) is locally asymptotically stable for κ(κ0,1p), while unstable for κ(θsθ,κ0). When κ=κ0, the model undergoes a supercritical Hopf bifurcation at P+=(κ,vκ), a family of periodic solutions (uT(t),vT(t)) bifurcate from P+=(κ,vκ) and the bifurcating periodic solutions are stable.

    Proof. If (A1) holds, then P+=(κ,vκ) is a unique positive equilibrium of (2.2). Setting the Jacobi matrix of (2.2) at (κ,vκ) is

    J(κ):=(a(κ),b(κ)c(κ),0),

    where, a(κ)=2θs(1pκ)1,b(κ)=θ and c(κ)=2(sθ)s(1pκ). The characteristic equation of J(κ) is

    λ2T(κ)λ+D(κ)=0, (2.3)

    with

    T(κ)=2θs(1pκ)1,D(κ)=2θ(sθ)s(1pκ).

    then the roots of Eq (2.3) are

    λ1,2=12[T(κ)±T2(κ)4D(κ)].

    If (A2)(or(A3)) satisfies, then all the eigenvalues of J(κ) have strictly negative real parts according to the stability theory, and P+=(κ,vκ) is locally asymptotically stable. If (A4) is true, then all the eigenvalues of J(κ) have positive real parts, hence, P+=(κ,vκ) is unstable. For an arbitrary κ(θsθ,κ0), model (2.2) is unstable at P+=(κ,vκ), and for an arbitrary κ(κ0,1p), P+=(κ,vκ) (2.2) is locally asymptotically stable. When κ=κ0, J(κ0) has a pair of pure imaginary roots λ=±iω0 with ω0=(sθ)12. Let  λ(κ)=β(κ)±iω(κ) be the roots of Eq (2.3) near κ=κ0, then we have

    β(κ)=θs(1pκ)12,dβ(κ)dκ|κ=κ0=pθs<0.

    According to the Poincaré-Andronov-Hopf bifurcation theorem, system (2.1) undergoes Hopf bifurcation at κ=κ0. Let the eigenvectors of J(κ0) and J(κ0) corresponding to the eigenvalues iω0 and iω0 be q=(1,b0)T,q=(a0,b0)T, satisfying <q,q>=1 and <q,ˉq>=0, where b0=ω0θi,a0=12lπ, b0=θ2lπω0. Denote

    f(κ,u,v)=u+κp(u+κ)2s(1m)2(u+κ)2(v+vκ)1+(1m)2(u+κ)2,g(κ,u,v)=θ(v+vκ)+s(1m)2(u+κ)2(v+vκ)1+(1m)2(u+κ)2

    by [26], and we give the expression of the cubic coefficient c1(κ0) in normal form. Calculating Qqq, Qqˉq and Cqqˉq,

    Qqq=(c0d0),Qqˉq=(e0f0),Cqqˉq=(g0h0),

    with

    c0=2p(2s39s2θ+14sθ28θ3+4is(sθ)θω0)s2(s2θ),d0=2p(sθ)(s26sθ+8θ2+4isθω0)s2(s2θ),e0=2p(2s25sθ+4θ2)s2,f0=2p(s25sθ+4θ2)s2,g0=8p2(sθ)θ2(6(s2θ)2+is(s4θ)ω0)s3(s2θ)2,h0=8p2(sθ)θ2(6(s2θ)2is(s4θ)ω0)s3(s2θ)2,

    as well as

    <¯q,Qqq>=is2(s2θ)ω0p(θ(s37s2θ+14sθ28θ3)+4s(sθ)θω20) +1s2(s2θ)ω0p(2s3+s2θ(9+4θ)+8θ32sθ2(7+2θ))ω0,<q,Cqqˉq>=4p2(sθ)θ2(θ+iω0)(6i(s2θ)2+s(s4θ)ω0)s3(s2θ)2ω0 =4p2(sθ)θ2s3(s2θ)2(sθ(s4θ)6(s2θ)2) +4p2(sθ)θ2s3(s2θ)2ω0(6θ(s2θ)2+s(s4θ)ω02)i,<q,Qqq>=is2(s2θ)ω0(pθ(s37s2θ+14sθ28θ3)4s(sθ)θω20) +1s2(s2θ)ω0(2s3+s2(94θ)θ+8θ3+2sθ2(7+2θ)),<q,Qqˉq>=p(iθ(s25sθ+4θ2)+(2s2+5sθ4θ2)ω0).

    Then, we can obtain

    H20=(c0d0)<q,Qqq>(1b0)<¯q,Qqq>(1¯b0)=0,H11=(e0f0)<q,Qqˉq>(1b0)<¯q,Qqˉq>(1ˉb0)=0,

    so

    c1(κ0)=i2ω0<q,Qqq><q,Qqˉq>+12<ˉq,Cqqˉq>. (2.4)

    Its real and imaginary parts are

    Re(c1(κ0))=p2s2(s2θ)ω02(2s3+s2(94θ)θ+8θ3+2sθ2(7+2θ))θ(s25sθ+4θ2) p2s2(s2θ)ω0(pθ(s37s2θ+14sθ28θ3)4s(sθ)θω20)(2s2+5sθ4θ2) +2p2(sθ)θ2s3(s2θ)2(sθ(s4θ)6(s2θ)2) (2.5)

    and

    Im(c1(κ0))=p2s2(s2θ)ω0(2s3+s2(94θ)θ+8θ3+2sθ2(7+2θ))(2s2+5sθ4θ2) p2s2(s2θ)ω02(pθ(s37s2θ+14sθ28θ3)4s(sθ)θω20)θ(s25sθ+4θ2) +2p2(sθ)θ2s3(s2θ)2ω0(6θ(s2θ)2+s(s4θ)ω02). (2.6)

    If Re(c1(κ0))<0(>0), then the Hopf bifurcation is backward (forward) and the bifurcating periodic solutions (uT(t),vT(t)) are stable (unstable).

    In this subsection, for the perturbed population model, we derive the first derivative formula of the periodic function about the perturbation coefficients. On the basis of model (2.1), we introduce the perturbation term τ and coefficients (k11k12k21k22). The corresponding perturbed population model is

    (I+τ(k11k12k21k22))(dudtdvdt)=(upu2s(1m)2u2v1+(1m)2u2θv+s(1m)2u2v1+(1m)2u2), (2.7)

    where τ is sufficiently small such that (1+τk11τk12τk211+τk22) is reversible, then (2.7) can be reduced to

    (dudtdvdt)=1K(τ)(1+k22τk12τk21τ1+k11τ)(upu2s(1m)2u2v1+(1m)2u2θv+s(1m)2u2v1+(1m)2u2), (2.8)

    where

    K(τ):=|(1+τk11τk12τk211+τk22)|=(k11k22k12k21)τ2+(k11+k22)τ+1>0.

    At P+=(κ,vκ), the Jacobian matrix of (2.8) is

    J(κ,τ):=1K(τ)(J11(κ,τ)J12(κ,τ)J21(κ,τ)J22(κ,τ)), (2.9)

    where,

    J11(κ,τ):=(1+k22τ)a(κ)k12c(κ)τ,J12(κ,τ):=(1+k22τ)b(κ),J21(κ,τ):=(1+k11τ)c(κ)k21a(κ)τ,J22(κ,τ):=k21b(κ)τ,a(κ)=2θs(1pκ)1,b(κ)=θ,c(κ)=2(sθ)s(1pκ). (2.10)

    Let the characteristic equation corresponding to J(κ,τ) be

    λ2H(κ,τ)λ+D(κ,τ)=0, (2.11)

    where

    H(κ,τ)=1K(τ)τ(k22a(κ)k12c(κ)k21b(κ))+1K(τ)a(κ),D(κ,τ)=1K(τ)2θ(sθ)s(1pκ). (2.12)

    When κκτ, let ˉβ(κτ)±iˉω(κτ) be the roots of the characteristic Eq (2.11), then

    ˉβ(κτ)=12H(κ,τ),ˉω(κτ)=124D(κ,τ)H2(κ,τ). (2.13)

    Lemma 2.1. (See [26]) When κκ0, the population model (2.2) has a stable periodic solution (uT(t),vT(t)) bifurcating from P+=(κ,vκ) and T is the minimum positive period of (uT(t),vT(t)). Then there is a positive number τ1 such that for any τ(τ1,τ1), the perturbed population model (2.7) has a periodic solution (uT(t,τ),vT(t,τ)) depending on if τ, T(τ) is the minimum positive periodic function. When τ0, (uT(t,τ),vT(t,τ))(uT(t),vT(t)) and T(τ)T, then

    T(τ)=2πˉω(κτ)(1+(ˉβ(κτ)Im(c1(κτ))ˉω(κτ)Re(c1(κτ))ˉω(κτ)ˉω(κτ))(κκτ)+O((κκτ)2),κκτ,c1(κτ):=i2ˉω(κτ)(g20(τ)g11(τ)2|g11(τ)|213|g02(τ)|2)+g21(τ)2.

    Re(c1(κτ)) and Im(c1(κτ)) are the real and imaginary parts of c1(κτ), and ˉβ(κτ) and ˉω(κτ) are defined by (2.13).

    Theorem 2.2. When κκ0 for the perturbed population model (2.7), the first-order derivative formula of the periodic function, with respect to the perturbation coefficients, is

    T(0)=D(κ0)k11+D(κ0)k22Im(c1(κ0))Re(c1(κ0))c(κ0)k12Im(c1(κ0))Re(c1(κ0))b(κ0)k21,

    where b(κ0)=θ,c(κ0)=sθθ,D(κ0)=sθ. Re(c1(κ0)) and Im(c1(κ0)) are defined in (2.5) and (2.6).

    Proof. By Lemma 2.1, differentiating the periodic function T(τ), we have

    T(τ)=2πˉω2(κτ)dˉω(κτ)dτ2πˉω(κτ)(ˉβ(κτ)Im(c1(κτ))ˉω(κτ)Re(c1(κτ))ˉω(κτ)ˉω(κτ))dκτdτ+O(κκτ).

    If κκτ, then O(κκτ)0, and setting τ=0, then ˉω(κ0)=ω(κ0)=D(κ0).

    We first compute dκτdτ|τ=0. At κ=κτ, by the expression of H(κ,τ) defined in (2.12), we can gain

    τ(k22a(κτ)k12c(κτ)k21b(κτ))+a(κτ)=0. (2.14)

    Differentiating (2.14) with respect to τ, we obtain

    (k22a(κτ)k12c(κτ)k21b(κτ))+a(κτ)dκτdτ=0, (2.15)

    and setting τ=0, we have

    dκτdτ|τ=0=k12c(κ0)+k21b(κ0)a(κ0), (2.16)

    with

    b(κ0)=θ,c(κ0)=sθθ,a(κ0)=2θps.

    Second, we calculate ˉω(κ0). When κκτ, we derive

    ˉω(κ)=124D(κ,τ)H2(κ,τ),

    thereby,

    ˉω(κ)=κD(κ,τ)12H(κ,τ)κH(κ,τ)4D(κ,τ)H2(κ,τ).

    Since H(κτ,τ)=0 and κD(κτ,τ)=1K(τ)D(κτ), we have

    ˉω(κ0)=κD(κτ,τ)2D(κτ,τ)|τ=0=D(κτ)2K(τ)D(κτ)|τ=0=D(κ0)2D(κ0). (2.17)

    At last, we calculate ddτ(ˉω(κτ))|τ=0. By ˉω(κτ)=D(κτ,τ), we can get

    ddτ(ˉω(κτ))=12D(κτ,τ)ddτ(D(κτ,τ)). (2.18)

    According to D(κτ,τ)=D(κτ)K(τ), we have

    ddτ(D(κτ,τ))=K(τ)K2(τ)D(κτ)+ddτ(D(κτ))1K(τ). (2.19)

    Setting τ=0, we can obtain

    K(τ)K2(τ)D(κ0)=(k11+k22)D(κ0),ddτ(D(κτ))1K(τ)|τ=0=D(κ0)dκτdτ(0). (2.20)

    Therefore, from (2.16) and (2.20), we have

    ddτ(D(κτ,τ))|τ=0=(k11+k22)D(κ0)+k12c(κ0)+k21b(κ0)a(κ0)D(κ0). (2.21)

    By (2.18) and (2.21), we obtain

    ddτ(ˉω(κτ))|τ=0=12D(κ0)((k11+k22)D(κ0)+k12c(κ0)+k21b(κ0)a(κ0)D(κ0)). (2.22)

    Again, by (2.16), (2.17) and (2.22), we can derive

    T(0)=D(κ0)k11+D(κ0)k22Im(c1(κ0))Re(c1(κ0))c(κ0)k12Im(c1(κ0))Re(c1(κ0))b(κ0)k21.

    With respect to the population model (1.2) and according to the theory expounded in [27], we study the mathematical mechanisms of Turing patterns occurring at the stable periodic solution (uT(t),vT(t)). By the first derivative formula of the periodic function of the perturbed population model (2.7), we give the following theorem.

    Theorem 3.1. If hypothesis (A4) and Re(c1(κ0))<0 hold, when κκ0, the stable spatially homogeneous Hopf bifurcating periodic solution bifurcates (uT(t),vT(t)) from P+=(κ,vκ). If the domain Ω is large enough and

    D(κ0)d11+D(κ0)d22Im(c1(κ0))Re(c1(κ0))c(κ0)d12Im(c1(κ0))Re(c1(κ0))b(κ0)d21<0,

    then the following conclusions are true:

    (1) The reaction-diffusion population model (1.2) produces Turing patterns at the periodic solution (uT(t),vT(t));

    (2) If Im(c1(κ0))<0(>0), then when k12>M1(k21>M2), the reaction-diffusion population model (1.2) produces Turing patterns. That is, Turing patterns occuring at the periodic solution are determined by the cross-diffusion coefficients k12(k21), where

    b(κ0)=θ,c(κ0)=sθθ,D(κ0)=sθ.

    Re(c1(κ0)) and Im(c1(κ0)) from (2.5) and (2.6):

    M1:=D(κ0)d11+D(κ0)d22Im(c1(κ0))Re(c1(κ0))b(κ0)d21Im(c1(κ0))Re(c1(κ0))c(κ0),M2:=D(κ0)d11+D(κ0)d22Im(c1(κ0))Re(c1(κ0))c(κ0)d12Im(c1(κ0))Re(c1(κ0))b(κ0).

    Proof. Let the linearized vector form of the population model (2.1) at (uT(t),vT(t)) be

    (ϕt,φt)T=diag(DΔϕ,D Δ φ)+JT(t)(ϕ,φ)T), (3.1)

    where,

    JT(t):=(12puT(t)2s(1m)2uT(t)vT(t)(1+(1m)2(uT(t))2)2s(1m)2(uT(t))21+(1m)2(uT(t))22s(1m)2uT(t)vT(t)(1+(1m)2(uT(t))2)2θ+s(1m)2(uT(t))21+(1m)2(uT(t))2)

    is the Jacobian matrix of model (2.1) at (uT(t),vT(t)). D:=(d11d12d21d22), Δ is the Laplace operator. Let αn and ηn(x) be the eigenvalues and eigenfunctions of Δ in region Ω, respectively, and (ϕ,φ)T=(h(t),g(t))Tn=0knηn(x). For the sake of convenience, we set ς:=αn0,nN0:={0}N, then

    (dh(t)dt,dg(t)dt)T=ςD(h(t)g(t))+JT(t)(h(t)g(t)). (3.2)

    If D=0, then Eq (3.2) can be reduced to

    (dh(t)dt,dg(t)dt)T=JT(t)(h(t),g(t))T. (3.3)

    Setting ρ(t) as the fundamental solution matrix of Eq (3.3), it satisfies ρ(0)=I2. Let λi,i=1,2 be the eigenvalues of ρ(T), the corresponding eigenfunctions are (ξi,ηi)T,i=1,2, i.e.,

    ρ(T)(ξi,ηi)T=λi(ξi,ηi)T,

    where λ1 and λ2 are Floquet multipliers. Define

    (ϕi(t),ψi(t))T:=ρ(t)(ξi,ηi)T,

    clearly,

     (ϕi(0),ψi(0))T=(ξi,ηi)T,ρ(T)(ϕi(0),ψi(0))T=λi(ϕi(0),ψi(0))T.

    Differentiating with respect to t in (2.2), we can obtain

    (uv)=(12pu2s(1m)2uv(1+(1m)2u2)2s(1m)2u21+(1m)2u22s(1m)2uv(1+(1m)2u2)2θ+s(1m)2u21+(1m)2u2)(dudt,dvdt)T,

    then λ1=1 is the eigenvalue of ρ(T) and the eigenvector is (ϕ1(t),ψ1(t))T=(duT(t)dt|t=0,dvT(t)dt|t=0)T. Since (uT(t),vT(t)) is stable, |λi|<1. Let ρ(t,ς) be the fundamental solution matrix of Eq (3.2), then we have

    ρ(t,ς)t=ςDρ(t,ς)+JT(t)ρ(t,ς)

    and ρ(t,0)=ρ(t). By the implicit function theorem, there is ς1>0, ς(ς1,ς1) and continuous differentiable functions δi(ς),ξi(ς),ηi(ς), such that

    ρ(T,ς)(ξi(ς),ηi(ς))T=δi(ς)(ξi(ς),ηi(ς))T, (3.4)

    where δ1(ς) and δ2(ς) are Floquet multipliers. Make the following definition

    (ϕi(t,ς),ψi(t,ς))T:=ρ(t,ς)(ξi(ς),ηi(ς))T; (3.5)

    by ρ(0,ς)=I, we have

    (ϕi(0,ς),ψi(0,ς))T=(ξi(ς),ηi(ς))T. (3.6)

    From (3.4) and (3.6), we can gain

    ρ(T,ς)(ϕi(0,ς),ψi(0,ς))T=δi(ς)(ϕi(0,ς),ψi(0,ς))T,

    and especially

    (ϕi(t,0)ψi(t,0))=ρ(t,0)(ξi(0)ηi(0))=ρ(t)(ξiηi)=ρ(t)(ϕi(0)ψi(0))=(ϕi(t)ψi(t)).

    Taking i=1, by (3.5), we know

    (ϕ1(t,ς),ψ1(t,ς))T:=ρ(t,ς)(ξ1(ς),η1(ς))T.

    Hence, we can derive

    (ϕ1(t,ς)t,ψ1(t,ς)t)T=ςD(ϕ1(t,ς),ψ1(t,ς))T+JT(t)(ϕ1(t,ς),ψ1(t,ς))T.

    Differentiating the above equation with respect to ς and setting ς=0, we obtain

    (ϕ1ς(t,0)t,ψ1ς(t,0)t)T=D(ϕ1(t),ψ1(t))T+JT(t)(ϕ1ς(t,0),ψ1ς(t,0))T, (3.7)

    where, ϕ1ς:=ςϕ1,ψ1ς:=ςψ1. On the other hand, from (3.4) and (3.5), we can get

    (ϕ1(T,ς),ψ1(T,ς))T=δ1(ς)(ϕ1(0,ς),ψ1(0,ς))T.

    Differentiating with respect to ς, we have

    (ϕ1ς(T,ς),ψ1ς(T,ς))T=δ1(ς)(ϕ1(0,ς),ψ1(0,ς))T+δ1(ς)(ϕ1ς(0,ς),ψ1ς(0,ς))T.

    Let ς=0 by (3.6) and δ1(0)=λ1=1, and we can derive

    (ϕ1ς(T,0),ψ1ς(T,0))T=δ1(0)(ϕ1(0),ψ1(0))T+(ϕ1ς(0,0),ψ1ς(0,0))T. (3.8)

    According to Lemma 2.1, (uT(t,τ),vT(t,τ)) is the periodic solution of the perturbed population model (2.7), i.e.,

    (1+τd11τd12τd211+τd22)(uT(t,τ)tuT(t,τ)t)=(uT(t,τ)p(uT(t,τ))2s(1m)2(uT(t,τ))2vT(t,τ)1+(1m)2(uT(t,τ))2θvT(t,τ)+s(1m)2(uT(t,τ))2vT(t,τ)1+(1m)2(uT(t,τ))2).

    Differentiating with respect to τ and letting τ=0, we have

    (d(tuT(t,0))dτ,d(tvT(t,0))dτ)T=D(ϕ1(t),ψ1(t))T+JT(t)(duT(t,0)dτ,dvT(t,0)dτ)T, (3.9)

    where tuT(t,0)=ϕ1(t),tvT(t,0)=ψ1(t). Since T(τ) is the minimum positive periodic solution of (uT(t,τ),vT(t,τ)), we have

    (uT(t,τ),vT(t,τ))=(uT(t+T(τ),τ),vT(t+T(τ),τ)).

    Differentiating with respect to τ and letting τ=0,t=0, we can gain

    (duT(T,0)dτ,dvT(T,0)dτ)T=T(0)(ϕ1(0),ψ1(0))T+(duT(0,0)dτ,dvT(0,0)dτ)T, (3.10)

    where uT(t,0)=uT(t),vT(t,0)=vT(t),T(0)=T. Define

    Z(t):=(ϕ1ς(t,0),ψ1ς(t,0))T(duT(t,0)dτ,dvT(t,0)dτ)T,

    and by (3.7)–(3.10), we get

    Z(t)=JT(t)Z(t), (3.11)
    Z(T)Z(0)=(δ1(0)+T(0))(ϕ1(0),ψ1(0))T. (3.12)

    Let Z(t)=A(t)(Z1,Z2)T be the general solution of (3.11), where any vector (Z1,Z2)TR2. Since (ϕ1(0),ψ1(0))T and (ϕ2(0),ψ2(0))T are linearly independent, there exits constants p1 and p2 such that

    (Z1,Z2)T=p1(ϕ1(0),ψ1(0))T+p2(ϕ2(0),ψ2(0))T. (3.13)

    Substituting (3.13) into (3.12), we get δ1(0)+T(0)=0. According to Theorem 2.2, if T(0)<0, then δ1(0)>0. Assuming that Ω is sufficiently large, then there is at least one eigenvalue αn of Δ so that δ1(ς)=δ1(αn)>1. Therefore, the population model (1.2) appears to have Turing patterns at (uT(t),vT(t)). When T(0)<0 by Theorem 2.2, we have

    D(κ0)d11+D(κ0)d22Im(c1(κ0))Re(c1(κ0))c(κ0)d12Im(c1(κ0))Re(c1(κ0))b(κ0)d21<0. (3.14)

    Since (A3) is true, we can obtain b(κ0)=θ<0,c(κ0)=sθθ>0. If Re(c1(κ0))<0, then when Im(c1(κ0))<0,

    Im(c1(κ0))Re(c1(κ0))c(κ0)>0,Im(c1(κ0))Re(c1(κ0))b(κ0)<0.

    From (3.14), we gain

    d12>D(κ0)d11+D(κ0)d22Im(c1(κ0))Re(c1(κ0))b(κ0)d21Im(c1(κ0))Re(c1(κ0))c(κ0):=M1.

    When the cross-diffusion coefficient d12>M1, the cross-diffusion population model (1.2) generates Turing patterns at the periodic solution (uT(t),vT(t)). Similarly, if Im(c1(κ0))>0, then

    Im(c1(κ0))Re(c1(κ0))c(κ0)<0,Im(c1(κ0))Re(c1(κ0))b(κ0)>0,

    so

    d21>D(κ0)d11+D(κ0)d22Im(c1(κ0))Re(c1(κ0))c(κ0)d12Im(c1(κ0))Re(c1(κ0))b(κ0):=M2.

    When the cross-diffusion coefficient d21>M2, the cross-diffusion population model (1.2) produces Turing patterns at the periodic solution (uT(t),vT(t)).

    We shall conduct numerical simulations in three cases to verify the relevant conclusions: The diffusive population model forms Turing patterns at the periodic solutions. Fix the parameters in model (2.1):

    m=0.6,s=0.1,θ=0.09,p=0.0592,xΩ=(0,30),

    then (κ,vκ)=(7.5,46.3) is a unique positive equilibrium. By calculation, κ0=7.505. According to Theorem 2.1, when κκ0, b(κ0)=0.09,c(κ0)=19,D(κ0)=0.01, Rec1(κ0)=3.1316×103<0 and Imc1(κ0)=4.3667×103, simultaneously, hypothesis (A4) is true. Take the initial values as u0=8+0.1cos(x),v0=47+0.1cos(x).

    (1) If d11=1,d22=1,d12=d21=0, then D(κ0)d11+D(κ0)d22=0.2>0. By Theorem 3.1, in model (1.2), Turing patterns do not appear at (uT(t),vT(t)), namely, the same self-diffusion rates do not destroy the stability of the periodic solution(See [28]). If d11d22,d11>0,d22>0 and d12=d21=0, then D(κ0)d11+D(κ0)d22>0 and the periodic solution of diffusion model (1.2) is stable (Figure 1).

    Figure 1.  The periodic solution (uT(t),vT(t)) of the reaction-diffusion equation is stable.

    (2) If d11=0.2,d22=0.5,d12=0.05, then M2=0.6195. Select d21=0.7 by calculating T(0)<0. According to Theorem 3.1, when d21>M2=0.6195, cross-diffusion induces system (1.2) to produce Turing patterns at (uT(t),vT(t)) (Figure 2).

    Figure 2.  Cross-diffusion-induced Turing patterns.

    (3) If d11=1,d22=1,d12=0.02, then M2=1.6184. We choose d21=1.7, through computation and T(0)<0. According to Theorem 3.1, when d21>M2, cross-diffusion induces system (1.2) to produce Turing patterns at (uT(t),vT(t)) (Figure 3).

    Figure 3.  Turing patterns induced by diffusion coefficient d21.

    In this paper, we established a cross-diffusion population model with prey refuge and Holling Ⅲ functional response, and studied the mathematical mechanisms of Turing patterns generated by the diffusion-driven instability of the periodic solutions. The results show that when Im(c1(κ0))<0(>0), the symbol of the diffusivity expression T(0) is actually determined by the cross-diffusion coefficient d21(d12). That is, when d21>M2(d12>M1) and the region Ω is sufficiently large, T(0)<0 and model (1.2) generate Turing patterns at the periodic solutions. Our research more accurately determined the range of cross-diffusion coefficients of Turing patterns occurring at the periodic solutions. This provided a new idea for model (1.2) to generate Turing instability at the periodic solutions.

    The authors declare they have not used Artificial Intelligence (AI) tools in the creation of this article.

    The authors declare that they have no competing interests.



    [1] N. L. Biggs, Cayley maps and symmetrical maps, Proc. Cambridge Philos. Soc., 72 (1972), 381–386. https://doi.org/10.1017/s0305004100047216 doi: 10.1017/s0305004100047216
    [2] N. L. Biggs, Classification of complete maps on orientable surfaces, Rend. Mat.(6), 4 (1971), 645–655.
    [3] N. L. Biggs, A. T. White, Permutation groups and combinatorial structures, vol. 33 of London Mathematical Society Lecture Note Series, Cambridge-New York: Cambridge University Press, 1979.
    [4] A. Devillers, W. Jin, C. H. Li, C. E. Praeger, On normal 2-geodesic transitive Cayley graphs, J. Algebraic Combin., 39 (2014), 903–918. https://doi.org/10.1007/s10801-013-0472-7 doi: 10.1007/s10801-013-0472-7
    [5] J. D. Dixon, B. Mortimer, Permutation groups, vol. 163 of Graduate Texts in Mathematics, New York: Springer-Verlag, 1996. http://dx.doi.org/10.1007/978-1-4612-0731-3
    [6] S. F. Du, J. H. Kwak, R. Nedela, Regular embeddings of complete multipartite graphs, Eur. J. Combin., 26 (2005), 505–519. https://doi.org/10.1016/j.ejc.2004.02.010 doi: 10.1016/j.ejc.2004.02.010
    [7] W. W. Fan, C. H. Li, The complete bipartite graphs with a unique edge-transitive embedding, J. Graph Theory, 87 (2018), 581–586. https://doi.org/10.1002/jgt.22176 doi: 10.1002/jgt.22176
    [8] W. W. Fan, C. H. Li, H. P. Qu, A classification of orientably edge-transitive circular embeddings of Kpe,pf, Ann. Comb., 22 (2018), 135–146. https://doi.org/10.1007/s00026-018-0373-5 doi: 10.1007/s00026-018-0373-5
    [9] W. W. Fan, C. H. Li, N. E. Wang, Edge-transitive uniface embeddings of bipartite multi-graphs, J. Algebr. Comb., 2 (2019), 125–134. https://doi.org/10.1007/s10801-018-0821-7 doi: 10.1007/s10801-018-0821-7
    [10] Y. Q. Feng, J. H. Kwak, J. X. Zhou, Enumerating reflexible 2-cell embeddings of connected graphs, Sci. China Math., 56 (2013), 933–950. https://doi.org/10.1007/s11425-012-4544-2 doi: 10.1007/s11425-012-4544-2
    [11] R. Jajcay, R. Nedela, Half-regular Cayley maps, Graphs Combin., 31 (2015), 1003–1018. https://doi.org/10.1007/s00373-014-1428-y doi: 10.1007/s00373-014-1428-y
    [12] L. D. James, Imbeddings of the complete graph, Ars Combin., 16 (1983), 57–72.
    [13] L. D. James, Edge-symmetric orientable imbeddings of complete graphs, European J. Combin., 11 (1990), 133–144. https://doi.org/10.1016/S0195-6698(13)80067-4 doi: 10.1016/S0195-6698(13)80067-4
    [14] L. D. James, G. A. Jones, Regular orientable imbeddings of complete graphs, J. Combin. Theory Ser. B, 39 (1985), 353–367. https://doi.org/10.1016/0095-8956(85)90060-7 doi: 10.1016/0095-8956(85)90060-7
    [15] V. P. Korzhik, H. J. Voss, On the number of nonisomorphic orientable regular embeddings of complete graphs, J. Combin. Theory Ser. B, 81 (2001), 58–76. https://doi.org/10.1006/jctb.2000.1993 doi: 10.1006/jctb.2000.1993
    [16] V. P. Korzhik, H. J. Voss, Exponential families of non-isomorphic non-triangular orientable genus embeddings of complete graphs, J. Combin. Theory Ser. B, 86 (2002), 186–211. https://doi.org/10.1006/jctb.2002.2122 doi: 10.1006/jctb.2002.2122
    [17] C. H. Li, Vertex transitive embeddings of complete graphs, J. Combin. Theory Ser. B, 99 (2009), 447–454. https://doi.org/10.1016/j.jctb.2008.09.002 doi: 10.1016/j.jctb.2008.09.002
    [18] B. P. Mull, R. G. Rieper, A. T. White, Enumerating 2-cell imbeddings of connected graphs, Proc. Amer. Math. Soc., 103 (1988), 321–330. https://doi.org/10.2307/2047573 doi: 10.2307/2047573
    [19] R. B. Richter, J. Širáň, R. Jajcay, T. W. Tucker, M. E. Watkins, Cayley maps, J. Combin. Theory Ser. B, 95 (2005), 189–245. https://doi.org/10.1016/j.jctb.2005.04.007 doi: 10.1016/j.jctb.2005.04.007
    [20] J. Širáň, T. W. Tucker, Characterization of graphs which admit vertex-transitive embeddings, J. Graph Theory, 55 (2007), 233–248. https://doi.org/10.1002/jgt.20239 doi: 10.1002/jgt.20239
    [21] M. Škoviera, J. Širáň, Regular maps from Cayley graphs. Ⅰ. Balanced Cayley maps, Proc. Discrete Math., 109 (1992), 265–276. https://doi.org/ 10.1016/0012-365X(92)90296-R doi: 10.1016/0012-365X(92)90296-R
    [22] X. Yu, B. G. Lou, The edge-regular complete maps, Open Math., 18 (2020), 1719–1726. https://doi.org/10.1515/math-2020-0115 doi: 10.1515/math-2020-0115
    [23] X. Yu, Enumeration of orientable vertex-transitive embeddings of Kpd. Submitted.
    [24] X. Yu, C. H. Li, B. G. Lou, Orientable vertex primitive complete maps. Submitted.
    [25] X. Yu, B. G. Lou, W. W. Fan, The complete bipartite graphs which have exactly two orientably edge-transitive embeddings, Ars Math. Contemp., 18 (2020), 371–379. https://doi.org/10.26493/1855-3974.1900.cc1 doi: 10.26493/1855-3974.1900.cc1
  • Reader Comments
  • © 2023 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

Metrics

Article views(1468) PDF downloads(54) Cited by(2)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog