Processing math: 59%
Research article Special Issues

Stabilization and pattern formation in chemotaxis models with acceleration and logistic source


  • We consider the following chemotaxis-growth system with an acceleration assumption,

    {ut=Δu(uw)+γ(uuα),xΩ, t>0,vt=Δvv+u,xΩ, t>0,wt=Δww+χv,xΩ, t>0,

    under the homogeneous Neumann boundary condition for u,v and the homogeneous Dirichlet boundary condition for w in a smooth bounded domain ΩRn (n1) with given parameters χ>0, γ0 and α>1. It is proved that for reasonable initial data with either n3, γ0, α>1 or n4, γ>0, α>12+n4, the system admits global bounded solutions, which significantly differs from the classical chemotaxis model that may have blow-up solutions in two and three dimensions. For given γ and α, the obtained global bounded solutions are shown to convergence exponentially to the spatially homogeneous steady state (m,m,0) in the large time limit for appropriately small χ, where m=1|Ω|Ωu0(x) if γ=0 and m=1 if γ>0. Outside the stable parameter regime, we conduct linear analysis to specify possible patterning regimes. In weakly nonlinear parameter regimes, with a standard perturbation expansion approach, we show that the above asymmetric model can generate pitchfork bifurcations which occur generically in symmetric systems. Moreover, our numerical simulations demonstrate that the model can generate rich aggregation patterns, including stationary, single merging aggregation, merging and emerging chaotic, and spatially inhomogeneous time-periodic. Some open questions for further research are discussed.

    Citation: Chunlai Mu, Weirun Tao. Stabilization and pattern formation in chemotaxis models with acceleration and logistic source[J]. Mathematical Biosciences and Engineering, 2023, 20(2): 2011-2038. doi: 10.3934/mbe.2023093

    Related Papers:

    [1] Chichia Chiu, Jui-Ling Yu . An optimal adaptive time-stepping scheme for solving reaction-diffusion-chemotaxis systems. Mathematical Biosciences and Engineering, 2007, 4(2): 187-203. doi: 10.3934/mbe.2007.4.187
    [2] Changwook Yoon, Sewoong Kim, Hyung Ju Hwang . Global well-posedness and pattern formations of the immune system induced by chemotaxis. Mathematical Biosciences and Engineering, 2020, 17(4): 3426-3449. doi: 10.3934/mbe.2020194
    [3] Xiaomei Bao, Canrong Tian . Turing patterns in a networked vegetation model. Mathematical Biosciences and Engineering, 2024, 21(11): 7601-7620. doi: 10.3934/mbe.2024334
    [4] Fabrizio Clarelli, Roberto Natalini . A pressure model of immune response to mycobacterium tuberculosis infection in several space dimensions. Mathematical Biosciences and Engineering, 2010, 7(2): 277-300. doi: 10.3934/mbe.2010.7.277
    [5] Marcello Delitala, Tommaso Lorenzi . Emergence of spatial patterns in a mathematical model for the co-culture dynamics of epithelial-like and mesenchymal-like cells. Mathematical Biosciences and Engineering, 2017, 14(1): 79-93. doi: 10.3934/mbe.2017006
    [6] Fiona R. Macfarlane, Mark A. J. Chaplain, Tommaso Lorenzi . A hybrid discrete-continuum approach to model Turing pattern formation. Mathematical Biosciences and Engineering, 2020, 17(6): 7442-7479. doi: 10.3934/mbe.2020381
    [7] Swadesh Pal, Malay Banerjee, Vitaly Volpert . Spatio-temporal Bazykin’s model with space-time nonlocality. Mathematical Biosciences and Engineering, 2020, 17(5): 4801-4824. doi: 10.3934/mbe.2020262
    [8] Nazanin Zaker, Christina A. Cobbold, Frithjof Lutscher . The effect of landscape fragmentation on Turing-pattern formation. Mathematical Biosciences and Engineering, 2022, 19(3): 2506-2537. doi: 10.3934/mbe.2022116
    [9] Benedetto Bozzini, Deborah Lacitignola, Ivonne Sgura . Morphological spatial patterns in a reaction diffusion model for metal growth. Mathematical Biosciences and Engineering, 2010, 7(2): 237-258. doi: 10.3934/mbe.2010.7.237
    [10] Lin Zhang, Yongbin Ge, Zhi Wang . Positivity-preserving high-order compact difference method for the Keller-Segel chemotaxis model. Mathematical Biosciences and Engineering, 2022, 19(7): 6764-6794. doi: 10.3934/mbe.2022319
  • We consider the following chemotaxis-growth system with an acceleration assumption,

    {ut=Δu(uw)+γ(uuα),xΩ, t>0,vt=Δvv+u,xΩ, t>0,wt=Δww+χv,xΩ, t>0,

    under the homogeneous Neumann boundary condition for u,v and the homogeneous Dirichlet boundary condition for w in a smooth bounded domain ΩRn (n1) with given parameters χ>0, γ0 and α>1. It is proved that for reasonable initial data with either n3, γ0, α>1 or n4, γ>0, α>12+n4, the system admits global bounded solutions, which significantly differs from the classical chemotaxis model that may have blow-up solutions in two and three dimensions. For given γ and α, the obtained global bounded solutions are shown to convergence exponentially to the spatially homogeneous steady state (m,m,0) in the large time limit for appropriately small χ, where m=1|Ω|Ωu0(x) if γ=0 and m=1 if γ>0. Outside the stable parameter regime, we conduct linear analysis to specify possible patterning regimes. In weakly nonlinear parameter regimes, with a standard perturbation expansion approach, we show that the above asymmetric model can generate pitchfork bifurcations which occur generically in symmetric systems. Moreover, our numerical simulations demonstrate that the model can generate rich aggregation patterns, including stationary, single merging aggregation, merging and emerging chaotic, and spatially inhomogeneous time-periodic. Some open questions for further research are discussed.



    This study considers the system

    {ut=Δu(uw)+γ(uuα),xΩ, t>0,vt=Δvv+u,xΩ, t>0,τ0wt=Δww+χv,xΩ, t>0, (1.1)

    in a bounded domain ΩRn (n1) with smooth boundary, where the parameters χ>0, α>1, γ0 and τ0{0,1}. The functions u(x,t) and v(x,t) denote the organism density and the concentration of the signal emitted by the organism itself respectively, the vector function w(x,t)=(w1,w2,,wn)(x,t) represents the velocity of u. The system (1.1) is a generalized version of the classical Keller-Segel model (cf. [1])

    {ut=Δu(uv),τ0vt=Δvv+u,

    which was used to describe the aggregation movement of cells. For half a century, this model and its variants have attracted a lot of attention and yielded rich results. (cf. the surveys [2,3,4,5] for instance).

    In the classical chemotaxis model or other models containing taxis mechanisms such as the prey-taxis model (cf. [6]), it is generally assumed that the advective velocity of the organism is proportional to the concentration gradient of the stimulus/resource. However, this kind of assumption would neglect some details of the movement and is sometimes inappropriate to describe some observations, such as schooling fish (cf. [7,8]), swarms of flying insects (cf. [9]) and flea-beetles (cf. [10]), where the advective velocity variation (i.e., acceleration) rather than the velocity itself of the organism is supposed to be proportional to the gradient of stimuli/resources. This acceleration assumption was used in mathematical modeling of some predator-prey models, such as [11,12], where the model admits Spatio-temporal patterns consistent with experimental observations, while classical predator-prey models can not (cf. [13,14]). By introducing the acceleration assumption into chemotaxis models, [15] considered (1.1) with τ0=γ=0 under the boundary condition

    un=vn=wn=0, nw×n=0,xΩ,

    where n is the outward normal vector on Ω and nw:=(nw1,nw2,,nwn). It is proved in [15] that for reasonable initial data, the system (1.1) with the above boundary condition admits global bounded solutions for n3. This result is significantly different from classical parabolic-parabolic/parabolic-elliptic chemotaxis models, which have the well-known critical mass phenomenon in the two-dimensional case (cf. [16,17,18,19,20]) and admits blow-up solutions in three-dimensional case (cf. [18,21]). For critical mass phenomena in generalized versions of chemotaxis models, we refer the reader to [22,23,24,25,26,27,28,29,30,31,32,33] and the references therein.

    Our main purpose is to study whether the fully parabolic (τ0=1) system (1.1) can preclude blow-up solutions (and has no critical mass phenomena for n3) under the boundary condition

    un=vn=0, w=0,xΩ. (1.2)

    Specifically, we shall consider two cases: with/without the logistic source term. In the case of γ=0, the main task focuses on whether there exist global bounded solutions for all reasonable initial data in two/three dimensions. When γ>0, motivated by the works of the classical chemotaxis growth system (see Remark 1.1), our goal is to find the condition on α such that the model precludes blow-up solutions in high dimensions (n4). Moreover, for global bounded solutions, we shall investigate the large time behavior.

    We assume that initial data (u0,v0,w0)(x):=(u,v,w)(x,0) conforms to

    0u0C0(¯Ω), 0v0W1,(Ω), w0W1,(Ω;Rn)andu0,v00. (1.3)

    The first result is about the global boundedness of solutions, as follows.

    Theorem 1.1. Suppose that τ0=1 and (1.3) holds with

    either(H1):n3, γ0, α>1,or(H2):n4, γ>0, α>12+n4,

    then the system (1.1)-(1.2) possesses a unique classical solution (u,v,w) satisfying

    u,vC0(¯Ω×[0,))C2,1(¯Ω×(0,)),wC0(¯Ω×[0,);Rn)C2,1(¯Ω×(0,);Rn),

    and u,v>0 in Ω×(0,). Moreover, the solution is uniformly bounded in time, i.e.,

    supt[0,)(u(,t)L(Ω)+v(,t)W1,(Ω)+w(,t)W1,(Ω))<C(1+χ)β, (1.4)

    where β>1 and C>0 are two constants independent of χ and t.

    Remark 1.1. Recall the classical chemotaxis growth system

    {ut=Δuχ(uv)+huμuα,xΩ,t>0,τ0vt=Δvv+u,xΩ,t>0, (1.5)

    with the homogeneous Neumann boundary condition and reasonable initial data. The parameters hR, χ,α>0 and τ0{0,1}, and ΩRn is a smooth bounded domain. In the case of α=2, (1.5) admits global bounded solutions for any μ>0 if n=2 and for μ>μ0(Ω) with some positive number μ0(Ω) if n3 (cf. [34,35,36,37,38]). In the case of τ0=0, it is proved in [39] that there exist initial data such that solutions blow up in finite time if

    α<{76 if n{3,4},1+12(n1) if n5.

    From this point of view, Theorem 1.1 shows that the introduction of the acceleration assumption into chemotaxis models can improve the condition for α in some sense, such as (1.1) admits global bounded solutions for α>1 (or γ=0) in three dimensions. Indeed, the condition for α in Theorem 1.1 is more correlated with that in the indirect chemotaxis model, see Remark 1.2 for a detailed reason.

    Remark 1.2. Assume τ0=1. Now we consider a special case: w is a conservative vector field with w=χϕ for some scalar potential function ϕC1(¯Ω). Then the system (1.1) is related to the chemotaxis model with indirect signal production

    {ut=Δuχ(uϕ)+γ(uuα),xΩ,t>0,ϕt=Δϕϕ+v,xΩ,t>0,vt=Δvv+u,xΩ,t>0. (1.6)

    If γ=0, [40] shows that n=4 is the critical dimension for (1.6) with the homogeneous Neumann boundary condition or a mixed boundary condition: no blow-up occurs if Ωu0<(8π2)χ (the radially symmetric setting is needed if the boundary condition is of the Neumann type). When n2, γ>0 and α>12+n4, [41] proved that (1.6) admits global bounded solutions which convergence to (1,1,1) in L(Ω) under the homogeneous Neumann boundary condition. For the system (1.1)-(1.2), Theorem 1.1 has the same restriction for α when n4, but when n=2,3, the restriction can be improved to α>1 (or γ=0). When ignore the random diffusion of v in (1.6) (i.e., the third equation turns into vt=v+u), assuming n2 and α>n2, [42] obtained global bounded solutions to (1.6) under the boundary condition un=ϕn=0 on Ω.

    The global boundedness of solutions is given in Theorem 1.1. Next, we are dedicated to investigating the large time behavior of the obtained solution. The system (1.1)-(1.2) has a non-trivial constant equilibrium (m,m,0) with

    m:={1|Ω|Ωu0(x)>0,if γ=0,1,if γ>0. (1.7)

    We have the following asymptotic dynamics of the system (1.1)-(1.2).

    Theorem 1.2. Suppose that the conditions in Theorem 1.1 hold, then there exists a number ˜χ>0 such that for all χ(0,˜χ), the solution (u,v,w) of the system (1.1)-(1.2) obtained in Theorem 1.1 satisfies

    u(,t)mW1,(Ω)+v(,t)mW1,(Ω)+w(,t)W1,(Ω)Ceσt as t, (1.8)

    where C, σ are two positive constants independent of t. Moreover, if γ>0 and α2, then we can take ˜χ=γ.

    The above theorem gives the global stability of the equilibrium (m,m,0) under the condition that χ is small. However, outside the stable parameter regime, the behavior of solutions is unknown. It is well-known that the one-dimensional classical chemotaxis growth model can generate rich patterns, such as stationary, periodic, and chaotic (cf. [43]). To investigate whether the system (1.1)-(1.2) can generate similar patterns, we shall conduct linear analysis, weakly nonlinear analysis, and perform numerical simulations to see possible patterns.

    Remark 1.3. The weakly nonlinear analysis in Sect. 4.2 shows that the asymmetric system (1.1)-(1.2) admits pitchfork bifurcations (which occur generically in symmetric systems).

    Notations. For brevity, we shall use C or Ci (i=1,2,3,) denote a generic positive constant which may vary from line to line in the context, and use CP to denote the well-known Poincaré constant

    CP:={infC>0C|φL2(Ω)CφL2(Ω) for all φW1,2(Ω) with φΩ=0}. (1.9)

    The rest of this paper is organized as follows. Sect. 2 is devoted to establishing global bounded solutions to (1.1) and the proof of Theorem 1.1. By constructing appropriate Lyapunov functionals, the global stability stated in Theorem 1.2 is proved in Sect. 3. Outside the stable parameters regime, in Sect. 4, we first conduct linear analysis to specify the condition on parameters that allow pattern formation, then use a standard perturbation expansion approach (cf. [44,Sect. 6]) to conduct weakly nonlinear analysis to show pitchfork bifurcations, and finally perform numerical simulations for fully nonlinear parameter regimes to show that stationary patterns, spatially inhomogeneous time-periodic patterns, merging and emerging chaotic patterns are all possible. Moreover, we discussed some open questions for further research.

    This section contributes to the proof of Theorem 1.1. To begin with, we give the existence of local solutions.

    Lemma 2.1. Suppose that the conditions in Theorem 1.1 hold. Then there exists Tmax(0,] such that the system (1.1) has a unique classical solution (u,v,w) satisfying

    {u,vC0(¯Ω×[0,Tmax))C2,1(¯Ω×(0,Tmax)),wC0(¯Ω×[0,Tmax);Rn)C2,1(¯Ω×(0,Tmax);Rn),

    and u,v>0 in Ω×(0,Tmax). Moreover,

     if Tmax<,thenlimtTmaxu(,t)L(Ω)=.

    Proof. This proof is based on standard arguments involving the contraction mapping principle. We omit the detailed proof here and refer the reader to [15] (the main difference here is that w satisfies a parabolic equation, however, the homogeneous Dirichlet boundary for w will provide sufficient regularities in the proof, see Lemma 2.4 below).

    The following facts are basic but crucial.

    Lemma 2.2. Suppose that the conditions in Theorem 1.1 hold. If γ=0,

    u(,t)L1(Ω)=u0L1(Ω) for all t(0,Tmax). (2.1)

    If γ>0,

    u(,t)L1(Ω)max{u0L1(Ω),|Ω|} for all t(0,Tmax) (2.2)

    and

    t+stΩuαC for all t(0,Tmaxs), (2.3)

    where s:=min{1,Tmax2}.

    Proof. Integrating the first equation of (1.1) with respect to xΩ, one has

    ddtΩu=γΩ(uuα). (2.4)

    If γ=0, (2.1) follows immediately. If γ>0, by Hölder's inequality uL1(Ω)|Ω|11αuLα(Ω) thanks to α>1, we obtain

    ddtuL1(Ω)γuαL1(Ω)(u1αL1(Ω)|Ω|1α),

    which proves (2.2). Integrating (2.4) in time and using (2.2), one has (2.3).

    To derive a priori estimates for v, we shall need the following property (cf. [46,Lemma 4.1] and [49,Lemma 1]), which is derived based on smooth properties of the Neumann heat semigroup.

    Lemma 2.3. Let Ω be a bounded domain in Rn, n1, with smooth boundary. Let T(0,] and suppose that zC0(¯Ω×[0,T))C2,1(¯Ω×(0,T)) is a solution of

    {zt=Δzz+g,xΩ,t(0,T),zn=0,xΩ,t(0,T),

    where gL((0,T);Lq(Ω)) with some q1. Then there exists a constant C>0 such that

    z(,t)W1,r(Ω)Cwithr{[1,nqnq) if qn,[1,] if q>n.

    In addition to the well-known smooth properties of the Neumann heat semigroup [63,Lemma 1.3], we will use the Lp-Lq-estimates of the Dirichlet heat semigroup (cf. [55,Lemma 2.4 (i)], where the proof is based on [58,Proposition 48.4*,48.5 and 48.7*] and similar arguments as in the proof of [63,Lemma 1.3]).

    Lemma 2.4. ([55,Lemma 2.4 (i)]). Let etΔ be the Dirichlet heat semigroup in ΩRn (n1), λ1>0 denote the first nonzero eigenvalue of Δ in Ω under the Dirichlet boundary condition. If 1qp, then for any zLq(Ω), it holds that

    etΔzLp(Ω)C(1+tn2(1q1p))eλ1tzLq(Ω) for all t>0,

    and

    etΔzLp(Ω)C(1+t12n2(1q1p))eλ1tzLq(Ω) for all t>0.

    We are in a position to derive the following boundedness criterion.

    Lemma 2.5. Suppose that n1, γ0 and α>1. If there exists p>n such that

    supt(0,Tmax)w(,t)Lpc0 (2.5)

    for some positive constant c0, then there exist two constants β>1 and C>0 independent of χ and t such that

    u(,t)L(Ω)+v(,t)W1,(Ω)+w(,t)W1,(Ω)C(1+χ)β (2.6)

    for all t(0,Tmax).

    Proof. This proof can follow similar arguments developed in [59,Lemma 4.2] (see also [2,Lemma 3.2], [47,Lemma 3.1] and [48,Lemma 3.2] for instance). However, in order to give a clear dependency between the boundedness of supt(0,Tmax)u(,t)L and χ, we shall sketch the proof here. Now for each T(0,Tmax), define

    M(T):=supt(0,T)u(,t)L,

    which is finite thanks to the local existence results in Lemma 2.1. Since p>n, for any q(n,p), it holds that

    pp+1<1n<qandδ:=1pqpq(0,1). (2.7)

    Then by Hölder's inequality, the interpolation inequality, Lemma 2.2, (2.5) and (2.7), we have

    u(,t)w(,t)Lq(Ω)u(,t)Lpqpq(Ω)w(,t)Lp(Ω)c0(1+χ)u(,t)δL(Ω)u(,t)1δL1(Ω)C1(1+χ)Mδ(T)for all t(0,T).

    For t0:=(t1)+ and each t(0,T), the variation-of-constants formula for u(,t) gives

    u(,t)=e(tt0)Δu(,t0)+tt0e(ts)Δ(u(,s)w(,s))+tt0e(ts)Δγ(u(,s)uα(,s)). (2.8)

    Since α>1, we have γ(uuα)γ, then using (2.8), the maximum principle, Lp-Lq estimates of the Neumann heat semigroup (cf. [63,Lemma 1.3]) and 12+n2q<1, one has

    u(,t)L(Ω)u0L(Ω)+tt0e(ts)Δ(u(,s)w(,s))L(Ω)ds+tt0e(ts)ΔγL(Ω)dsC2+C2t0(1+(ts))12n2qeλ2(ts)u(,s)w(,s)Lq(Ω)ds+γ(tt0)C3(1+χ)(Mδ(T)+1) for all t(0,T), (2.9)

    where (etΔ)t0 is the Neumann heat semigroup in Ω and λ2 denotes the first nonzero eigenvalue of Δ in Ω under the homogeneous Neumann boundary condition. In view of (2.9), we arrived at

    M(T)C3(1+χ)(Mδ(T)+1)for all T(0,Tmax),

    which along with δ(0,1) gives

    M(T)max{1,(2C3(1+χ))11δ}for all T(0,Tmax).

    Since the above Ci (i=1,2,3) is independent of T, we have

    u(,t)L(Ω)max{1,(2C3(1+χ))11δ}for all t(0,Tmax). (2.10)

    With (2.10), we can apply Lemma 2.3 to prove

    v(,t)W1,(Ω)C4(1+χ)11δfor all t(0,Tmax), (2.11)

    which together with Lemma 2.4 shows that

    W(,t)W1,(Ω)C5(1+χ)11δfor all t(0,Tmax). (2.12)

    The combination of (2.10)–(2.12) proves (2.6) by letting β=11δ.

    The following conclusion follows from Lemma 2.2 and Lemma 2.5 immediately.

    Corollary 2.6. If (H1) holds, then there exist two constants β>1 and C>0 independent of χ and t such that (2.6) holds.

    Proof. It holds that nn132 thanks to n3. This along with Lemma 2.2 and Lemma 2.3 indicates that

    vW1,32(Ω)C1for all t(0,Tmax). (2.13)

    The variation-of-constants formula for w yields

    w(,t)=et(Δ1)w0+χt0e(ts)(Δ1)v(,s)dsfor all t(0,Tmax), (2.14)

    where (etΔ)t0 is the Dirichlet heat semigroup in Ω. Using Lemma 2.4, (2.13) and the fact that n2(2314)<1 for n3, for all t(0,Tmax), we have

    w(,t)L4(Ω)C2+C2χt0(1+(ts)n2(2314))eλ1tvL32(Ω)C3(1+χ). (2.15)

    In view of Lemma 2.5 and n3<4, the proof is obtained immediately from (2.15).

    Lemma 2.7. Suppose that n2, γ>0, α>12+n4, then there exist two constants β>1 and C>0 independent of χ and t such that (2.6) holds.

    Proof. For ε>0, multiplying (2.4) by eεt yields

    ddt(eεtΩu(,t))+γeεtΩuα(,t)=(ε+γ)eεtΩu(,t)for all t(0,Tmax).

    For t0(0,Tmax), integrating the above equation in time and using (2.2), we have

    γtt0eεsΩuα(,s)dsmax{u0L1(Ω),|Ω|}(1+γε)(eεteεt0)+eεt0Ωu(,t0)max{u0L1(Ω),|Ω|}(1+γε)eεt for all t(t0,Tmax),

    which implies that

    tt0eε(ts)Ωuα(,s)dsmax{u0L1(Ω),|Ω|}(1ε+1γ) for all t(t0,Tmax). (2.16)

    Based on (2.16) and the smoothing properties of the Neumann heat semigroup, we can use similar arguments as in [41,Lemma 3.4] to find some ρ1>n2 with α>(n+2)ρ1n+2ρ1 such that

    vLρ1(Ω)Cfor all t(0,Tmax). (2.17)

    Letting ρ2:=nρ1(nρ1)+>n, then using (2.14), (2.17) and Lemma 2.4, we have

    w(,t)Lρ2(Ω)Cw0L(Ω)+Cχt0(1+(ts)12n2(1ρ11ρ2))eλ1(ts)vLρ1(Ω)dsC(1+χ)for all t(0,Tmax).

    This alongside Lemma 2.5 completes the proof.

    Proof of Theorem 1.1. If (H1) holds, then by Corollary 2.6 we arrive at (2.6). If (H2) holds, then an application of Lemma 2.7 also gives (2.6). In view of Lemma 2.1 and (2.6), we obtain Tmax=. Then (1.4) follows from (2.6) immediately, which completes the proof.

    In this section, we shall construct appropriate Lyapunov functionals to prove the large time behavior of solutions stated in Theorem 1.2. In the following, we suppose that the conditions in Theorem 1.1 hold and (u,v,w) is the solution of the system (1.1)-(1.2) obtained in Theorem 1.1. The proof of Theorem 1.2 is mainly divided into two cases: with/without the logistic source term. We first give the following higher order regularities of the solution.

    Lemma 3.1. For any θ(0,1), there exists a positive constant C(θ) depending on θ such that

    uC2+θ,1+θ2(¯Ω×[1,))+vC2+θ,1+θ2(¯Ω×[1,))+wC2+θ,1+θ2(¯Ω×[1,))C(θ).

    Proof. The conclusion can be proved by standard bootstrap arguments based on (1.4), the Lp estimate and the Schauder estimate (cf. [51]). We omit the details here for brevity and refer the reader to [60,Theorem 2.1] for instance. The following auxiliary lemma will be used to prove global stability.

    Lemma 3.2. ([61,Lemma 1.1]) Let a0 and b>0 be two constants, F(t)0,aH(t)dt<. Assume that EC1([a,)),E is bounded from below and satisfies

    E(t)bF(t)+H(t) in [a,).

    If either FC1([a,)) and F(t)k in [a,) for some k>0, or FCθ([a,)) and FCθ([a,))k for some constants 0<θ<1 and k>0, then

    limtF(t)=0.

    With the aid of the Poincaré inequality (1.9), we can construct the following Lyapunov functional in the case of γ=0.

    Lemma 3.3. Suppose that (H1) holds with γ=0, then there exists a positive constant ˜χ1>0 such that for all χ(0,˜χ1), the function

    E1(t):=CPΩ(um)2+Ω(vm)2+1χ2Ω|w|2 for all t>0

    satisfies

    ddtE1(t)η1E1(t) for all t>0, (3.1)

    where η1:=min{13CP,12}, m and CP are given by (1.7) and (1.9) respectively.

    Proof. Integrating the equations in (1.1) by parts, using (1.9) and Young's inequality, for all t>0, we obtain

    ddtΩ(um)2=2Ω|u|2+2Ωuwu12Ω|u|21CPΩ(um)2+2Ωu2|w|2 (3.2)

    and

    ddtΩ(vm)2=2Ω|v|2+2Ω(um)(vm)2Ω(vm)22Ω|v|2+23Ω(um)212Ω(vm)2, (3.3)

    as well as

    ddtΩ|w|2=2ni=1Ω|wi|22Ω|w|2+2χΩwvΩ|w|2+χ2Ω|v|2. (3.4)

    The combination of (3.2)–(3.4) implies that

    ddtE1(t)13Ω(um)212Ω(vm)2Ω(1χ22CPu2)|w|2for all t>0.

    According to (1.4), we know there exists ˜χ1>0 such that

    1χ22CPu212χ2+12˜χ212CPu2>12χ2for all χ(0,˜χ1),

    which indicates that

    ddtE1(t)13Ω(um)212Ω(vm)212χ2Ω|w|2for all t>0.

    This proves (3.1).

    In the presence of logistics sources, i.e., γ>0, we can construct the following Lyapunov functional.

    Lemma 3.4. Suppose that γ>0 and either (H1) or (H2) holds, then there exists a positive constant ˜χ2>0 such that for all χ(0,˜χ2), the function

    E2(t):=Ω(u1lnu)+χ2Ω(v1)2+Ω|w|2 for all t>0

    satisfies

    ddtE2(t)min{χ23,12}F(t) for all t>0, (3.5)

    where

    F(t):=Ω(u1)2+Ω(v1)2+Ω|w|2 for all t>0.

    Proof. Multiplying the first equation in (1.1) by 11u, integrating the result by parts and using Young's inequality, for all t>0, we obtain

    ddtΩ(u1lnu)=Ω|u|2u2+Ωw|u|uγΩ(uα11)(u1)12Ω|u|2u2+12Ω|w|2γΩφ(u,α)(u1)2, (3.6)

    where φ(s,α):=sα11s1 for s0 and α>1 with φ(1,α):=lims1φ(s,α)=α1. Similarly, as in deriving (3.3), we have

    ddtΩ(v1)22Ω|v|2+23Ω(u1)212Ω(v1)2for all t>0. (3.7)

    For all t>0, the combination of (3.4), (3.6) and (3.7) implies that

    ddtE2(t)Ω(γφ(u,α)23χ2)(u1)2χ22Ω(v1)212Ω|w|2. (3.8)

    {Case1: 1<α<2}. It holds that φ(u,α)=(α1)sα20 with some s0 between u and 1. By (1.4) we know that there exists a constant C1>1 independent of χ(0,1) such that u(x,t)<C1 for all (x,t)Ω×(0,). Therefore, if we take ˜χ2=min{1,(γ(α1)Cα21)12}, then

    γφ(u,α)23χ2γ(α1)Cα21˜χ22+χ23χ23for all χ(0,˜χ2),

    which along with (3.8) proves (3.5).

    {Case2: α2}. We first note that φ(0,α)=1 and φ(1,α)=α11. Since ddαφ(s,α)=sα1lnss1>0 for s0,1, we have

    φ(s,α)φ(s,2)=1for all s0 and α2. (3.9)

    Let ˜χ2=γ, then

    γφ(u,α)23χ2γ23χ2χ23for all χ(0,˜χ2), (3.10)

    which alongside (3.8) also arrives at (3.5).

    We are now in a position to prove the convergence rate of solutions stated in Theorem 1.2.

    Proof of Theorem 1.2. If (H1) holds with γ=0, then it follows from Lemma 3.3 that (3.1) holds for 0<χ<˜χ:=˜χ1, hence we have

    E1(t)E1(0)eη1tfor all t>0. (3.11)

    We have from (3.11) that

    u(,t)mL2(Ω)+v(,t)mL2(Ω)+wL2(Ω)Ceη1tfor all t>0,

    which along with Lemma 3.1 and the Gagliardo-Nirenberg inequality

    ϕW1,(Ω)Cϕn+2n+4W2,(Ω)ϕ2n+4L2(Ω) for all ϕW2,(Ω)

    proves (1.8) with σ:=2η1n+4.

    We next assume that γ>0 and either (H1) or (H2) holds, then Lemma 3.4 indicates that (3.5) holds for 0<χ<˜χ:=˜χ2. Noting that E2(t)0 for all t>0 due to s1lns0 for s0. Using Lemma 3.2, Lemma 3.1 and Lemma 3.4, we obtain F(t)0 as t, which implies that

    u(,t)1L2(Ω)+v(,t)1L2(Ω)+w(,t)L2(Ω)0as t.

    The L'H{ˆo}pital's rule shows that limu1u1lnu(u1)2=limu112u=12, hence we can find t1>1 such that

    14Ω(u1)2Ω(u1lnu)Ω(u1)2for all t>t1.

    This together with Lemma 3.4 shows that there exists a constant η2>0 such that

    ddtE2(t)η2E2for all t>t1.

    We are now at the same position as in (3.11), which also proves (1.8) with σ:=2η2n+4. Moreover, the proof of Lemma (3.4) (see (3.9) and (3.10)) implies that if γ>0 and α2, then ˜χ can be taken as ˜χ=γ.

    A variety of patterns (including stationary, periodic, and chaotic) are observed in the one-dimensional classical Keller-Segel model with the logistic term (cf. [43]). Theorem 1.2 shows that the equilibrium (m,m,0) of (1.1) is globally asymptotic stable for appropriately small χ>0. This raises a natural question, whether the considered system (1.1) with the acceleration assumption can generate aggregated patterns for large χ? To this end, we shall first conduct linear analysis outside the stable parameters regime to specify the condition on parameters such that the equilibrium (m,m,0) is linearly unstable, and then perform numerical simulations to see possible patterns. Throughout this section, we assume that n=1, Ω=(0,L) is an interval of length L>0 and τ0=1 in (1.1).

    It is well known that the Neumann and Dirichlet eigenvalue problems

    {Δφ+ρ(N)φ=0,in (0,L),φn=0,on {0,L},and{Δψ+ρ(D)ψ=0,in (0,L),ψ=0,on {0,L},

    have eigenvalues 0=ρ(N)0<ρ(N)1ρ(N)2ρ(N)k and 0<ρ(D)1ρ(D)2ρ(D)k with corresponding eigenvectors {φk}k=0 and {ψk}k=0 respectively, where {φk}k=0 and {ψk}k=1 respectively form an orthonormal basis of L2(Ω), and φ0=1L. Moreover,

    {φk(x)=2Lcos(ωkx),ψk(x)=2Lsin(ωkx),andρ(N)k=ρ(D)k=ω2kfor all kN+,

    where ωk:=kπL for all kN. Then a solution of (1.1)-(1.2) can be written as

    (u(x,t)v(x,t)w(x,t))=(mm0)+k=0(ak(t)φk(x)bk(t)φk(x)dk(t)ψk(x)), (4.1)

    where m is given by (1.7), and we have supplemented the definition

    ψ0(x):=sin(ω0x)0for all x(0,L) (4.2)

    so that the forms of three component solutions are uniform (noting that k=0dk(t)ψk(x)=k=1dk(t)ψk(x) for any function d0(t) of t).

    For linear stability, we assume |ak|, |bk|, |dk+1|1 for each mode kN to represent (u,v,w) has a small perturbation to the equilibrium (m,m,0) and substitute (4.1) into (1.1) to obtain the following ODE

    ddt(ak(t)bk(t)dk(t))=Jk(ak(t)bk(t)dk(t))+O(umL2(Ω)+vmL2(Ω)+wL2(Ω)),

    where

    Jk:=(ω2kI10mωk1ω2k100χωkω2k1)withI1:=γ(αmα11)0. (4.3)

    Clearly, by (1.7) we have

    I1=0if γ=0andI1>0if γ>0. (4.4)

    The matrix Jk has three eigenvalues, denoted by λk (i=1,2,3), then

    λ3k+D2λ2k+D1λk+D0=0,kN,

    where Di=Di(k2), i=0,1,2, are given by

    {D2(k2):=3ω2k+2+I1,D1(k2):=3ω4k+2(2+I1)ω2k+2I1+1,D0(k2):=ω6k+(I1+2)ω4k+(2I1mχ+1)ω2k+I1,kN. (4.5)

    Define

    \begin{align} D_3(k^2): = D_1D_2-D_0 = 8 \omega_k^6+8( I_1+2) \omega_k^4+\left(2 I_1^2+12 I_1+m \chi +10\right)\omega_k^2 +2( I_1+1)^2, \quad k\in\mathbb N. \end{align} (4.6)

    If the real parts Re(\lambda_k) < 0 for all k\in\mathbb N , then the equilibrium (m, m, 0) is linearly stable. It follows from (4.4) that I_1\geq0 , which alongside (4.5) and (4.6) implies that

    \begin{align} D_i(k_0^2) > 0\quad\text{for }i\in\left\{{1, 2, 3}\right\}, \ k\in\mathbb N. \end{align} (4.7)

    Therefore, by the Routh-Hurwitz criterion (cf. [56,Appendix B]) we know that (m, m, 0) is linearly stable if and only if D_0(k^2) > 0 for all k\in\mathbb N . If there exist some k_0\in \mathbb N such that D_0(k_0^2) < 0 , then (m, m, 0) is linearly unstable. Specifically, we have the following linear stability.

    Lemma 4.1. For the system (1.1)-(1.2) with \chi > 0 , \gamma\geq0 , \alpha > 1 and m given by (1.7), the constant solution (m, m, 0) is linearly unstable if and only if

    \begin{align} \chi > \chi^*: = \inf\limits_{k\in\mathbb N^+}\chi^*_k \quad\mathit{\text{with}}\quad \chi^*_k: = \frac{\left(\omega_k^2+1\right)^2 \left(I_1+\omega_k^2\right)}{m\omega_k^2}, \end{align} (4.8)

    where I_1 is given by (4.3).

    Proof. By (4.7) and the Routh-Hurwitz criterion, we see that (m, m, 0) is linearly unstable if and only if D_0(k^2) < 0 for some k\in\mathbb N^+ (noting that D_0(k^2)\mid_{k = 0} = I_1\geq0 ). The proof is completed by noting that D_0(k^2) < 0 if and only if \chi > \chi^*_k for k\in\mathbb N^+ .

    In this subsection, we assume that

    \begin{align} \gamma = 1, \ \alpha = 2, \ m = 1, \ \chi = \chi^*+ \varepsilon \end{align} (4.9)

    with 0 < \varepsilon\ll1 (the simpler case \gamma = 0 can be discussed similarly), then the system (1.1) becomes

    \begin{align} \left\{ \begin{array}{ll} u_t = \Delta u -\nabla \cdot\left(u w\right)+u-u^2, & x\in(0, L), \ t > 0, \\ v_t = \Delta v- v+u, & x\in(0, L), \ t > 0, \\ w_t = \Delta w -w+\chi\nabla v, & x\in(0, L), \ t > 0, \\ \frac{\partial u}{\partial\bf n} = \frac{\partial v}{\partial\bf n} = w = 0, \quad &x\in\left\{{0, L}\right\}. \end{array} \right. \end{align} (4.10)

    Our goal is to perform nonlinear analysis in a weak sense under the condition 0 < \varepsilon\ll 1 . In this respect, we shall derive the amplitude equations for (4.10) by following a standard perturbation expansion approach (cf. [44,Sect. 6]), see also [50,Sect. 3.2] for an application in a symmetric two-species competition chemotaxis model).

    We have from (4.3) and Lemma 4.1 that I_1 = 1 and hence (4.8) gives \chi^* = \inf_{k\in\mathbb N^+}\frac{\left(\omega_k^2+1\right)^3}{\omega_k^2} . Since \varphi_1(s): = \frac{(s^2+1)^3}{s^2} is a convex function in s > 0 , and \min\limits_{s > 0}\varphi_1(s) = \varphi_1(\frac{\sqrt2}2) = \frac{27}{4} . Therefore, for any L > 0 , \chi^* = \inf\limits_{k\in\mathbb N^+}\chi^*_k attains its infimum at either one or two modes. Without loss of generality, we assume there is only one mode k^*\in\mathbb N^+ such that \chi^* = \chi^*_{k^*} = \inf\limits_{k\in\mathbb N^+}\chi^*_k . Furthermore, for brevity (see Remark 4.2), we assume

    \begin{align} L < \sqrt2\pi, \end{align} (4.11)

    then \frac{k\pi}L > \frac{\sqrt2}2 for all k\in\mathbb N^+ , which indicates that k^* = 1 and

    \begin{align} \chi^* = \frac{\left(\omega_1^2+1\right)^3}{\omega_1^2 }. \end{align} (4.12)

    Moreover, the following result is a direct consequence of Lemma 4.1.

    Corollary 4.2. \mathcal J^*_k is invertible with three eigenvalues have negative real parts for all k\in\mathbb N except at k = 1 , where \mathcal J^*_1 has an eigenvalue 0 with the corresponding eigenvector

    \begin{align} \bm\xi = \left({\xi_1, \xi_2, \xi_3}\right)^T: = \left({-\frac{\omega_1}{1+\omega_1^2}, -\frac{\omega_1}{(1+\omega_1^2)^2}, 1}\right)^T. \end{align} (4.13)

    Proof. Using (4.3) with k = 1 , \chi = \chi^* and (4.9), one can obtain three eigenvalues of \mathcal J^*_1 : 0 and -\frac{\sqrt3 \left(1+\omega_1^2\right)\left({\sqrt3\pm\rm i}\right)}{2} , and the corresponding eigenvector of the eigenvalue 0 is \bm\xi given by (4.13). The rest of the proof follows from the proof of Lemma 4.1 immediately.

    For 0 < \varepsilon\ll1 and \chi = \chi^*+ \varepsilon , we define the slow time variable \tau: = \varepsilon t and (U, V, W)(x, \tau): = (u, v, w)(x, t) , then by (4.10) we know that (U, V, W)(x, \tau) satisfies

    \begin{align} \left\{ \begin{array}{ll} \varepsilon\partial_\tau U = \partial_{xx} U- \partial_{x} (UW)+U-U^2, \quad&x\in(0, L), \ \tau > 0, \\ \varepsilon\partial_\tau V = \partial_{xx} V-V+U, \quad&x\in(0, L), \ \tau > 0, \\ \varepsilon\partial_\tau W = \partial_{xx} W-W+(\chi^*+ \varepsilon) \partial_{x} V, \quad&x\in(0, L), \ \tau > 0, \\ \partial_{x} U = \partial_{x} V = W = 0, \quad&x\in\left\{{0, L}\right\}.\\ \end{array} \right. \end{align} (4.14)

    We shall use the following asymptotic expansions of (U, V, W) for 0 < \varepsilon\ll1 :

    \begin{align} \left( \begin{array}{c} U(x, \tau)\\ V(x, \tau)\\ W(x, \tau)\\ \end{array} \right) = \left( \begin{array}{c} 1\\ 1\\ 0\\ \end{array} \right) + \sum\limits_{j = 1}^\infty \varepsilon^{\frac j2} \left( \begin{array}{c} U^{(j)}(x, \tau)\\ V^{(j)}(x, \tau)\\ W^{(j)}(x, \tau)\\ \end{array} \right). \end{align} (4.15)

    The reason the expansion to be powers of \varepsilon^\frac 12 is that a pitchfork bifurcation can be expected although there is no symmetry of the system, see Remark 4.1 for a detailed reason.

    Remark 4.1. Pitchfork bifurcations occur generally for symmetric models, such as symmetric chemotaxis models as in [50,52]. Here, for the asymmetric system (4.14), a pitchfork bifurcation may also occur at \varepsilon = 0 . To see this, we denote

    \begin{align*} \mathcal L \left( \begin{array}{c} U\\ V\\ W\\ \end{array} \right) : = \left( \begin{array}{c} \mathcal L_1(U, W)\\ \partial_{xx} V-V+U\\ \partial_{xx} W-W+(\chi^*+ \varepsilon) \partial_{x} V\\ \end{array} \right) \end{align*}

    with \mathcal L_1(U, W): = \partial_{xx} U- \partial_{x} (UW)+U-U^2 , and assume that for small 0 < \varepsilon\ll1 , there exists a steady state which can be appropriately expressed as (\overline U_f, \overline V_g, \overline W_h): = (1+f(x), 1+g(x), h(x)) for some small f, g, h , i.e., \mathcal L \left({(\overline U_f, \overline V_g, \overline W_h)^T}\right) = (O(\varepsilon), O(\varepsilon), O(\varepsilon)). We claim that if

    \begin{align} -f''(x)+f(x)+h'(x) = 0, \end{align} (4.16)

    then

    \begin{align*} \mathcal L \left({(\overline U_f, \overline V_g, \overline W_h)^T}\right) = \mathcal L \left({(\overline U_{-f}, \overline V_{-g}, \overline W_{-h})^T}\right), \end{align*}

    which suggests a pitchfork bifurcation. By linearity we only need to verify \mathcal L_1(\overline U_f, \overline W_h) = \mathcal L_1(\overline U_{-f}, \overline W_{-h}) . In fact, it follows from (4.16) that

    \begin{align*} \mathcal L_1(\overline U_f, \overline W_h)- \mathcal L_1(\overline U_{-f}, \overline W_{-h}) & = -2 \left(-f''(x)+f(x)+h'(x)\right) = 0, \end{align*}

    which confirms our claim. There is (f(x), g(x), h(x)) satisfies (4.16), for example,

    \begin{align} (f(x), g(x), h(x)): = \varepsilon^{p_0}p_1 \left({ -\frac{p_2 p_3}{1+p_3^2} \cos (p_3 x), p_4\cos (p_3 x), p_2 \sin (p_3 x) }\right) \end{align} (4.17)

    with positive constants p_j , j = 0, 1, 2, 3, 4 . The above analysis suggests the possible existence of a pitchfork bifurcation, and we shall see that it indeed exists (see (4.37)).

    As in deriving (4.1), we have the extension

    \begin{align} \left( \begin{array}{c} U^{(j)}(x, \tau)\\ V^{(j)}(x, \tau)\\ W^{(j)}(x, \tau)\\ \end{array} \right) = \sum\limits_{k = 0}^\infty \mathcal M_k \left( \begin{array}{c} a_k^{(j)}(\tau)\\ b_k^{(j)}(\tau)\\ d_k^{(j)}(\tau)\\ \end{array} \right) \quad\text{for }j\in\mathbb N^+, \end{align} (4.18)

    where

    \begin{align} \mathcal M_k: = \left( \begin{array}{ccc} \cos \omega_k x & 0 & 0 \\ 0 & \cos \omega_k x & 0 \\ 0& 0&\sin \omega_k x \\ \end{array} \right)\quad\text{for }k\in\mathbb N. \end{align} (4.19)

    Noting that there is no need to specify d_0^{(j)}(\tau) for all j\in\mathbb N^+ due to (4.2). Substituting (4.18) into (4.14), then we can collect the terms of order \varepsilon^\frac j2 for j\in\mathbb N^+ (noting that there are no terms of order \varepsilon^0 ). To begin with, we consider the case of j = 1 .

    \bullet Order \varepsilon^\frac12 terms: collecting the terms of order \varepsilon^\frac 12 , one has

    \begin{align} \left\{ \begin{array}{l} 0 = \partial_{xx} U^{(1)}-{\partial_x}W^{(1)}-U^{(1)}, \\ 0 = \partial_{xx} V^{(1)}-V^{(1)}+U^{(1)}, \\ 0 = \partial_{xx} W^{(1)}-W^{(1)}+\chi^*{\partial_x}V^{(1)}.\\ \end{array} \right. \end{align} (4.20)

    Substituting (4.18) with j = 1 into (4.20), one has

    \begin{align} \sum\limits_{k = 0}^\infty \mathcal J^*_k\mathcal M_k \left( \begin{array}{c} a_k^{(1)}(\tau) \\ b_k^{(1)}(\tau) \\ d_k^{(1)}(\tau) \\ \end{array} \right) = \left( \begin{array}{c} 0 \\ 0\\ 0\\ \end{array} \right), \end{align} (4.21)

    where \mathcal M_k is defined by (4.19) and \mathcal J^*_k: = \mathcal J_k\mid_{\chi = \chi^*} in (4.3) with \chi^* given by (4.12). It follows from Corollary 4.2 and the orthogonality of sine and cosine series that

    \begin{align} a_i^{(1)}(\tau)\equiv b_i^{(1)}(\tau)\equiv d_j^{(1)}(\tau) \equiv 0 \quad\text{for all }i\in\mathbb N/\left\{{1}\right\}, \ j\in\mathbb N^+/\left\{{1}\right\}. \end{align} (4.22)

    For k = 1 , (a_1^{(1)}(\tau), b_1^{(1)}(\tau), d_1^{(1)}(\tau))^T is a multiple of \bm\xi given in (4.13), this along with (4.2), (4.18) and (4.22) indicates that

    \begin{align} \left( \begin{array}{c} U^{(1)}\\ V^{(1)}\\ W^{(1)}\\ \end{array} \right) = P(\tau)\mathcal M_1\bm\xi = P(\tau) \left( \begin{array}{c} \cos(\omega_1x)\xi_1\\ \cos(\omega_1x)\xi_2\\ \sin(\omega_1x)\xi_3\\ \end{array} \right), \end{align} (4.23)

    where P(\tau) is the unknown amplitude function of \tau . Our goal is to find the differential equation that P(\tau) satisfies. To this end, we shall derive the equations that higher order \varepsilon terms satisfy.

    \bullet Order \varepsilon terms: the order \varepsilon terms satisfy the following equations

    \begin{align} \left\{ \begin{array}{l} 0 = \partial_{xx} U^{(2)}-{\partial_x}W^{(2)}-U^{(2)}- \partial_{x}(U^{(1)}W^{(1)})-(U^{(1)})^2, \\ 0 = \partial_{xx} V^{(2)}-V^{(2)}+U^{(2)}, \\ 0 = \partial_{xx} W^{(2)}-W^{(2)}+\chi^*{\partial_x}V^{(2)}.\\ \end{array} \right. \end{align} (4.24)

    As in deriving (4.21), substituting (4.18) with j = 2 into (4.24) and using (4.23), we have

    \begin{align} \sum\limits_{k = 0}^\infty \mathcal J^*_k\mathcal M_k \left( \begin{array}{c} a_k^{(2)}(\tau) \\ b_k^{(2)}(\tau) \\ d_k^{(2)}(\tau) \\ \end{array} \right) = \left( \begin{array}{c} \partial_{x}(U^{(1)}W^{(1)})+(U^{(1)})^2\\ 0\\ 0\\ \end{array} \right) = -P(\tau)^2 \left( \begin{array}{c} c_1\cos (2 \omega_1 x)-\frac{\xi_1^2}{2}\\ 0\\ 0\\ \end{array} \right), \end{align} (4.25)

    where (\xi_1, \xi_2, \xi_3) is given by (4.13) and

    \begin{align} c_1: = -\frac{\xi_1(2 \xi_3\omega_1+\xi_1)}{2} = \frac{\omega_1^2(1+2\omega_1^2)}{2(1+\omega_1^2)^2} > 0. \end{align}

    Again by Corollary 4.2 and the orthogonality of sine and cosine series, one has

    \begin{align} \begin{cases} &(a_1^{(2)}(\tau), b_1^{(2)}(\tau), d_1^{(2)}(\tau))^T = Q(\tau)\bm\xi, \\ &a_i^{(2)}(\tau)\equiv b_i^{(2)}(\tau)\equiv d_j^{(2)}(\tau) \equiv 0 \quad\text{for all }i\in\mathbb N/\left\{{0, 1, 2}\right\}, \ j\in\mathbb N^+/\left\{{1, 2}\right\}, \end{cases} \end{align} (4.26)

    where Q(\tau) is some unknown function. For k = 0 and k = 2 , (4.25) turns into the following two equations

    \begin{align} \mathcal J^*_0 \left( \begin{array}{c} a_0^{(2)}(\tau) \\ b_0^{(2)}(\tau) \\ 0\\ \end{array} \right) = \frac{\xi_1^2}{2}P(\tau)^2 \left( \begin{array}{c} 1 \\ 0 \\ 0 \\ \end{array} \right) \quad\text{and}\quad \mathcal J^*_2 \left( \begin{array}{c} a_2^{(2)}(\tau) \\ b_2^{(2)}(\tau) \\ d_2^{(2)}(\tau) \\ \end{array} \right) = -c_1P(\tau)^2 \left( \begin{array}{c} 1 \\ 0 \\ 0 \\ \end{array} \right), \end{align} (4.27)

    respectively. Using (4.3) with \chi = \chi^* given in (4.12), one can easily obtain (\mathcal J^*_0)^{-1} and (\mathcal J^*_2)^{-1} , which alongside (4.2), (4.18), (4.26) and (4.27) gives

    \begin{align} \left( \begin{array}{c} U^{(2)} \\ V^{(2)} \\ W^{(2)}\\ \end{array} \right) & = P(\tau)^2 \left\{ \frac{\xi_1^2}{2} \left( \begin{array}{c} -1 \\ -1 \\ 0 \\ \end{array} \right) -c_1\mathcal M_2 \bm\eta \right\} + Q(\tau)\mathcal M_1\bm\xi, \end{align} (4.28)

    where \bm\xi and \mathcal M_k ( k\in\mathbb N ) are given by (4.13) and (4.19) respectively, and

    \begin{align} \bm\eta: = \left( \begin{array}{c} \eta_1\\ \eta_2 \\ \eta_3 \\ \end{array} \right) = (\mathcal J^*_2)^{-1} \left( \begin{array}{c} 1 \\ 0 \\ 0 \\ \end{array} \right) = \frac{1}{\omega_2^6+3 \omega_2^4-\omega_2^2 (\chi^*-3)+1} \left( \begin{array}{c} - \left(\omega_2^2+1\right)^2\\ -\left({\omega_2^2+1}\right) \\ \omega_2 \chi^* \\ \end{array} \right). \end{align} (4.29)

    \bullet Order \varepsilon^\frac32 terms: we are now in a position to collect the terms of order \varepsilon^\frac32 which provides \frac d{d\tau}P(\tau) and allows us to establish an ODE for P(\tau ). As before, we have

    \begin{align} \left\{ \begin{array}{l} \partial_\tau U^{(1)} = \partial_{xx} U^{(3)}-{\partial_x}W^{(3)}-U^{(3)}-2U^{(1)}U^{(2)}- \partial_{x}(U^{(2)}W^{(1)}+U^{(1)}W^{(2)}), \\ \partial_\tau V^{(1)} = \partial_{xx} V^{(3)}-V^{(3)}+U^{(3)}, \\ \partial_\tau W^{(1)} = \partial_{xx} W^{(3)}-W^{(3)}+\chi^*{\partial_x}V^{(3)}+{\partial_x}V^{(1)}, \\ \end{array} \right. \end{align}

    which along with similar procedures as in deriving (4.21) and (4.25) implies

    \begin{align} P'(\tau)\mathcal M_1\bm\xi & = \sum\limits_{k = 0}^\infty \mathcal J^*_k\mathcal M_k \left( \begin{array}{c} a_k^{(3)}(\tau) \\ b_k^{(3)}(\tau) \\ d_k^{(3)}(\tau) \\ \end{array} \right) + \left( \begin{array}{c} -2U^{(1)}U^{(2)}- \partial_{x}(U^{(2)}W^{(1)}+U^{(1)}W^{(2)}) \\ 0\\ {\partial_x}V^{(1)}\\ \end{array} \right). \end{align} (4.30)

    Substituting the expressions (4.23) and (4.28) for the lower oder terms into the above equation, and using \omega_{2} = 2\omega_1 and the orthogonality of sine and cosine series, we can collect the \cos(\omega_1x) terms for the first two components of (4.30) and the \sin(\omega_1x) terms for the third component of (4.30) to arrive at

    \begin{align} P'(\tau)\bm\xi & = J^*_1\mathcal M_1 \left( \begin{array}{c} a_1^{(3)}(\tau) \\ b_1^{(3)}(\tau) \\ d_1^{(3)}(\tau) \\ \end{array} \right) +P(\tau)^3 \left( \begin{array}{c} c_2 \\ 0\\ 0\\ \end{array} \right) +P(\tau) \left( \begin{array}{c} 0 \\ 0\\ -\omega_1\xi_2\\ \end{array} \right), \end{align} (4.31)

    where

    \begin{align} c_2: = \frac{c_1 (-\eta_1 \omega_1 \xi_3+2 \eta_1 \xi_1+\eta_3 \omega_1 \xi_1)+\xi_1^2 (\omega_1 \xi_3+2 \xi_1)}{2}, \end{align}

    and \bm\xi , \bm\eta are given by (4.13) and (4.29) respectively. Let

    \begin{align} \bm\beta : = (\beta_1, \beta_2, \beta_3) = \left({-\frac{L}{\pi }-\frac{\pi }{L}, -\frac{\left(L^2+\pi ^2\right)^2}{\pi L^3}, 1}\right), \end{align}

    then \bm\beta is in the left kernel of \mathcal J^*_1 . Multiplying (4.31) from left by \bm\beta and using \bm\beta\cdot\bm\xi = 3 , one has

    \begin{align} P'(\tau) & = c_3 P(\tau)^3 +c_4 P(\tau) \end{align} (4.32)

    with

    \begin{align} \begin{cases} &c_3: = \frac{\beta_1c_2}3 = \frac{\pi ^2 \left(7 L^8+15 \pi ^2 L^6+16 \pi ^4 L^4+66 \pi ^6 L^2+148 \pi ^8\right)}{36 \left(L^2+\pi ^2\right)^2 \left(L^6-12 \pi ^4 L^2-20 \pi ^6\right)} < 0, \\ &c_4: = -\frac{\omega_1\xi_2\beta_3}3 = \frac{\pi ^2 L^2}{3 \left(L^2+\pi ^2\right)^2} > 0, \end{cases} \end{align} (4.33)

    where c_3 < 0 use the fact that L^6-12 \pi ^4 L^2-20 \pi ^6 < 0 due to (4.11).

    By introducing A(t) = \varepsilon^\frac12P(\tau) , (4.32) leads to

    \begin{align} \frac d{dt} A(t) = \varepsilon c_4 A(t)+c_3 A(t)^3, \end{align} (4.34)

    where c_3, c_4 are given by (4.33). Since c_3 < 0 and c_4 > 0 , we know that (4.34) has a supercritical pitchfork bifurcation at \varepsilon = 0 and (4.34) has three steady states:

    \begin{align} 0 \text{(unstable)} \quad\text{and}\quad \pm \varepsilon^\frac12\sqrt{\frac{ c_4}{-c_3}} \text{ (stable)}. \end{align} (4.35)

    Now plugging (4.23) into (4.15) and using A(t) = \varepsilon^\frac12P(\tau) , we can obtain the appropriated solution for 0 < \varepsilon\ll1 :

    \begin{align} \left( \begin{array}{c} U(x, \tau)\\ V(x, \tau)\\ W(x, \tau)\\ \end{array} \right) = \left( \begin{array}{c} 1\\ 1\\ 0\\ \end{array} \right) + A(t) \left( \begin{array}{c} \cos(\omega_1x)\xi_1\\ \cos(\omega_1x)\xi_2\\ \sin(\omega_1x)\xi_3\\ \end{array} \right)+O( \varepsilon), \end{align} (4.36)

    where A(t) solves (4.34). For 0 < \varepsilon\ll1 , we have from \lim\limits_{t\rightarrow +\infty}\tau = +\infty (due to \tau = \varepsilon t ), (4.35) and (4.36) that besides the unstable equilibrium (1, 1, 0) , the system (4.10) also has two other stable steady states with very small amplitude in the form of

    \begin{align} \left( \begin{array}{c} u_s(x)\\ v_s(x)\\ w_s(x)\\ \end{array} \right) = \left( \begin{array}{c} 1\\ 1\\ 0\\ \end{array} \right) \pm \varepsilon^\frac12\sqrt{\frac{c_4}{-c_3}} \left( \begin{array}{c} \cos(\omega_1x)\xi_1\\ \cos(\omega_1x)\xi_2\\ \sin(\omega_1x)\xi_3\\ \end{array} \right)+O( \varepsilon). \end{align} (4.37)

    This shows that a pitchfork bifurcation indeed occurs for the asymmetric (4.10) at \varepsilon = 0 . Moreover, it is easy to verify that (4.17) is satisfied thanks to

    \begin{align*} \frac{\xi_3\omega_1}{1+\omega_1^2} = \frac{\omega_1}{1+\omega_1^2} = -\xi_1. \end{align*}

    Remark 4.2. In this subsection, we have only discussed the case L < \sqrt2\pi in (4.11) which is devoted to letting k^* = 1 and giving a brevity discussion. For a general L > 0 , \chi^* in (4.12) is instead by \chi^* = \frac{\left(\omega_{k^*}^2+1\right)^3}{\omega_{k^*}^2 } for some k^*\in\mathbb N^+ (which can be specified according to the given L > 0 ). Then the rest of the discussion is similar to the above.

    In this subsection, we shall perform numerical simulations to demonstrate possible patterns generated by the system (1.1)-(1.2) for large \chi > \chi^* . We shall consider two cases: with/without the logistic source term. To begin with, we consider the case of \gamma = 0 .

    \bullet Case 1: \gamma = 0 . With \Omega = (0, L) , Lemma 4.1 shows that patterns can be expected for \chi > \chi_* = \frac{\left(1+\omega_1^2\right)^2 }m with \omega_1 = \frac\pi L . We let L = 50 and the initial data (u_0, v_0, w_0) be a small random perturbation of the equilibrium \left({m, m, 0}\right) (without loss of generality, the species mass m is assumed to be 1):

    \begin{equation} (u_0, v_0, w_0) = (m+R, m+R, 0)\quad\text{with }m = 1, \end{equation} (4.38)

    where R\in(-0.01, 0.01) is a random number generated by MATLAB, then we have from (4.8) that

    \chi^* = \chi^*_1 = \left.\left(k^2+1\right)^2\right|_{k = \frac{\pi}{50}}\approx1.008.

    We observe two kinds of patterns generating from the equilibrium \left({1, 1, 0}\right) for \chi > \chi_* , as shown in Figure 1, where (a) shows that stationary patterns appear for \chi = 2 (not too large than \chi^*\approx1.008 ), while single merging aggregation patterns appear for \chi = 20 . Moreover, the later kind of patterns typically appear for large \chi , such as \chi = 50 , we omit the pattern here since it is similar to Figure 1 (b) (and the difference is that the larger \chi is, the more aggregation of (u, v, w) ). Similar stationary patterns as shown in Figure 1 (a) are observed for (1.1) with \tau_0 = \gamma = 0 in one-dimensional case (cf. [15]). Actually, these two kinds of patterns shown in Figure 1 have been also observed in classical chemotaxis models with/without the logistic source term such as [43,57,62], where the works indicate that besides the above two kinds, more patterns will appear in the existence of the logistic source term. From this point of view, we shall next consider the case \gamma = 1 and numerically investigate whether other types of patterns (such as periodic and chaotic) can arise for the system (1.1)-(1.2).

    Figure 1.  Spatio-temporal patterns generated by the system (1.1)-(1.2) in the interval \Omega = (0, 50) with \gamma = 0 and (a) \chi = 2 ; (b) \chi = 20 . The initial data is given by (4.38).

    \bullet Case 2: \gamma > 0 . Without loss of generality, we take \gamma = m = 1 and \alpha = 2 as in (4.9). We first take L = 50 , then we have from (4.12) that

    \begin{align*} \chi^* = \chi^*_{11} \approx6.755. \end{align*}

    Figure 2 shows four kinds of patterns generated from (1, 1, 0) of the system (1.1)-(1.2) for \chi = 8, 10, 12, 30 respectively. Figure 2 (a) shows stationary patterns for \chi = 8 , while the case of \chi = 10 gives single merging aggregation patterns in Figure 2 (b). These two patterns are similar as that shown in Figure 1 for the case of \gamma = 0 . As \chi goes on increasing, other kinds of patterns are observed as desired. When \chi = 12 , Figure 2 (c1) shows the spatially inhomogeneous periodic patterns which are "almost" strip patterns (as the strip patterns shown in Figure 2 (a)) with the eventual time periodicity (at a fixed position x = 25 ) shown explicitly in Figure 2 (c2). Figure 2 (d) considers the case of \chi = 30 and shows chaotic patterns corresponding to merging (two local maxima join together if they are close to each other) and emerging (a new maximum form in the space between two local maxima separated from each other, cf. [45]). The persistent pattern formation process (for large domain) of merging and emerging as shown in Figure 2 (d) is called "coarsening process" in [34].

    Figure 2.  Spatio-temporal patterns generated by the system (1.1)-(1.2) in the interval \Omega = (0, 50) with \gamma = 1 , \alpha = 2 and (a) \chi = 8 ; (b) \chi = 10 ; (c1) \chi = 12 ; (d) \chi = 30 . The subplot in (c2) shows the temporal profile of (u, v, w) at x = 25 for t\in[1000,1500] corresponding to the spatio-temporal patterns in (c1). The initial data is given by (4.38).

    We then let L = 20 , then (4.12) implies

    \begin{align*} \chi^* = \chi^*_{5} \approx6.852. \end{align*}

    In the supercritical case \chi = 30 , Figure 3 (a) shows a spatially inhomogeneous time-periodic pattern which is more obvious than that shown in Figure 2 (c1). Although these patterns are spatially inhomogeneous, however, at a fixed time, say t = 600 , we see that the organism u prefers to cluster in places where the signal is highly concentrated, see Figure 3 (b). Furthermore, when x is fixed, e.g., x = 10 , it is clear that Figure 3 (c) shows that the solution is eventually periodic in time.

    Figure 3.  Spatially inhomogeneous time-periodic patterns generated by the system (1.1)-(1.2) in the interval \Omega = (0, 20) with \gamma = 1 , \alpha = 2 and \chi = 30 . The initial data is given by (4.38).

    Remark 4.3. The above numerical simulations show that for the chemotaxis system (1.1)-(1.2) which is considered under acceleration assumption, large \chi would not destroy the aggregation patterns (two kinds of patterns shown in Figure 1) generated from the nontrivial constant steady state, and can also not produce spatially inhomogeneous periodic patterns or chaotic patterns without the logistic source term. While the introduction of the logistic source term can largely enrich the diversity of patterns as shown in Figure 2 and Figure 3. These observations are consistent with the results on pattern formation for classical chemotaxis models with/without logistic source terms (or volume-filling effect), see [43,53,54,57,62] for instance.

    The above numerical simulations show that a variety of patterns are observed for the system (1.1)-(1.2). The rich dynamics of solutions leave many interesting questions to be explored, below we discuss some of them.

    1. Fully parabolic systems with nonclassical boundary conditions. This paper consider the fully parabolic system (1.1) (for n\geq1 ) under the boundary condition (1.2), however, the global well-posedness (for n\leq3 ) is unknown if the boundary condition instead by

    \begin{equation*} \frac{\partial u}{\partial\bf n} = \frac{\partial v}{\partial\bf n} = \bf w \cdot \bf n = 0, \ \partial_\bf n\bf w\times\bf n = \bf0, \quad x\in\partial \Omega. \end{equation*}

    This is also one of the unsolved questions proposed in [15], where the key is the lack of parabolic regularity with the above boundary condition.

    2. Stationary problems. Theorem 1.2 asserts that the constant equilibrium (m, m, \textbf{0}) is exponentially asymptotic stable for small \chi > 0 , and hence no patterns can be expected in this stable parameter regime. In one-dimensional case \Omega = (0, L) with given L > 0 , for fixed \gamma\geq0 , \alpha > 1 , our linear analysis has specified the critical number \chi^* and patterns can possibly arise for the supercritical case \chi > \chi^* . The numerical simulations suggest that the system (1.1)-(1.2) has non-constant steady states whether or not \gamma = 0 . The existence and stability of non-constant steady states is an interesting and challenging question. In this paper, we only take a very small step in this direction: still in the one-dimensional case, for the weakly nonlinear case \chi = \chi^*+ \varepsilon with 0 < \varepsilon\ll1 , the solutions of the stable non-constant steady states are expressed approximately and the occurrence of pitchfork bifurcations at \varepsilon = 0 are confirmed for the asymmetric system (1.1)-(1.2). However, for the fully nonlinear case or for n\geq2 , a rigorous analysis of the existence and stability of non-constant steady states is blank.

    3. Eventually periodic solutions. For \chi^*\ll\chi and \gamma > 0 , the above numerical simulations indicate that the dynamics of solutions will be more complex, the spatially inhomogeneous time-periodic patterns and chaotic patterns corresponding to merging and emerging are both observed. For the former case, how to give a comprehensive mathematical explanation of the spatially inhomogeneous time-periodic patterns (at least in the one-dimensional case) is a challenging problem.

    The authors are grateful to the anonymous reviewers for useful comments which improved the exposition of this paper. The research of C. Mu is partially supported by the National Natural Science Foundation of China (Nos. 11971082, 12271064), the Chongqing Talent Support program (No. cstc2022ycjh-bgzxm0169), the Natural Science Foundation of Chongqing (No. cstc2021jcyj-msxmX1051), the Fundamental Research Funds for the Central Universities (Nos. 2019CDJCYJ001, 2020CDJQY-Z001), and Chongqing Key Laboratory of Analytic Mathematics and Applications. The research of W. Tao is partially supported by the National Natural Science Foundation of China (No. 12201082), the fellowship of China Postdoctoral Science Foundation (No. 2020M683232), the Fundamental Research Funds for the Central University (No. 2021CDJQY-047) and PolyU Postdoc Matching Fund Scheme Project ID P0030816/B-Q75G and P0034904 (Primary Work Programme W15F).

    The authors declare there is no conflict of interest.



    [1] E. F. Keller, L. A. Segel, Initiation of slime mold aggregation is viewed as an instability, J. Theor. Biol., 26 (1970), 399–415. https://doi.org/10.1016/0022-5193(70)90092-5 doi: 10.1016/0022-5193(70)90092-5
    [2] N. Bellomo, A. Bellouquid, Y. Tao, M. Winkler, Toward a mathematical theory of Keller-Segel models of pattern formation in biological tissues, Math. Models Methods Appl. Sci., 25 (2015), 1663–1763. https://doi.org/10.1142/S021820251550044X doi: 10.1142/S021820251550044X
    [3] N. Bellomo, N. Outada, J. Soler, Y. Tao, M. Winkler, Chemotaxis and cross-diffusion models in complex environments: models and analytic problems toward a multiscale vision, Math. Models Methods Appl. Sci., 32 (2022), 713–792. https://doi.org/10.1142/S0218202522500166 doi: 10.1142/S0218202522500166
    [4] T. Hillen, K. Painter, A user's guide to PDE models for chemotaxis, J. Math. Biol., 58 (2009), 183–217. http://dx.doi.org/10.1007/s00285-008-0201-3 doi: 10.1007/s00285-008-0201-3
    [5] D. Horstmann, From 1970 until present: the Keller-Segel model in chemotaxis and its consequences. II, Jahresber. Deutsch. Math.-Verein., 106 (2004), 51–69.
    [6] P. Kareiva, G. Odell, Swarms of predators exhibit "preytaxis" if individual predators use area-restricted search, Amer. Natur., 130 (1987), 233–270. https://doi.org/10.1086/284707 doi: 10.1086/284707
    [7] G. R. Flierl, D. Grünbaum, S. A. Levins, D. B. Olson, From individuals to aggregations: the interplay between behavior and physics, J. Theor. Biol., 196 (1999), 397–454. https://doi.org/10.1006/jtbi.1998.0842 doi: 10.1006/jtbi.1998.0842
    [8] A. Okubo, H. C. Chiang, C. C. Ebbesmeyer, Acceleration field of individual midges, anarete pritchardi (diptera: Cecidomyiidae), within a swarm, Can. Entomol., 109 (1977), 149–156. https://doi.org/10.4039/Ent109149-1 doi: 10.4039/Ent109149-1
    [9] J. K. Parrish, P. Turchin, Individual decisions, traffic rules, and emergent pattern in schooling fish, Animal groups in three dimensions, Cambridge University Press, Cambridge, 126–142.
    [10] P. Kareiva, Experimental and mathematical analyses of herbivore movement: quantifying the influence of plant spacing and quality on foraging discrimination, Ecol. Monogr., 52 (1982), 261–282. https://doi.org/10.2307/2937331 doi: 10.2307/2937331
    [11] R. Arditi, Y. Tyutyunov, A. Morgulis, V. Govorukhin, I. Senina, Directed movement of predators and the emergence of density-dependence in predator-prey models, Theor. Popul. Biol., 59 (2001), 207–221. https://doi.org/10.1006/tpbi.2001.1513 doi: 10.1006/tpbi.2001.1513
    [12] N. Sapoukhina, Y. Tyutyunov, R. Arditi, The role of prey taxis in biological control: a spatial theoretical model, Amer. Nat., 162 (2003), 61–76. https://doi.org/10.1086/375297 doi: 10.1086/375297
    [13] H.-Y. Jin, Z.-A. Wang, Global dynamics and Spatio-temporal patterns of predator-prey systems with density-dependent motion, European J. Appl. Math., 32 (2021), 652–682. https://doi.org/10.1017/s0956792520000248 doi: 10.1017/s0956792520000248
    [14] F. Yi, J. Wei, J. Shi, Bifurcation and spatiotemporal patterns in a homogeneous diffusive predator-prey system, J. Differ. Equ., 246 (2009), 1944–1977. https://doi.org/10.1016/j.jde.2008.10.024 doi: 10.1016/j.jde.2008.10.024
    [15] W. Tao, Z.-A. Wang, On a new type of chemotaxis model with acceleration, Commun. Math. Anal. Appl., 1 (2022), 319–344.
    [16] M. A. Herrero, J. J. L. Velázquez, Chemotactic collapse for the Keller-Segel model, J. Math. Biol., 35 (1996), 177–194. https://doi.org/10.1007/s002850050049 doi: 10.1007/s002850050049
    [17] D. Horstmann, G. Wang, Blow-up in a chemotaxis model without symmetry assumptions, European J. Appl. Math., 12 (2001), 159–177. https://doi.org/10.1017/S0956792501004363 doi: 10.1017/S0956792501004363
    [18] T. Nagai, Blow-up of radially symmetric solutions to a chemotaxis system, Adv. Math. Sci. Appl., 5 (1995), 581–601.
    [19] T. Nagai, Blowup of nonradial solutions to parabolic-elliptic systems modeling chemotaxis in two-dimensional domains, J. Inequal. Appl., 6 (2001), 37–55. https://doi.org/10.1155/S1025583401000042 doi: 10.1155/S1025583401000042
    [20] T. Nagai, T. Senba, K. Yoshida, Application of the Trudinger-Moser inequality to a parabolic system of chemotaxis, Funkcial. Ekvac., 40 (1997), 411–433.
    [21] M. Winkler, Finite-time blow-up in the higher-dimensional parabolic-parabolic Keller-Segel system, J. Math. Pures Appl., 100 (2013), 748–767. https://doi.org/10.1016/j.matpur.2013.01.020 doi: 10.1016/j.matpur.2013.01.020
    [22] A. Blanchet, J. A. Carrillo, P. Laurençot, Critical mass for a Patlak-Keller-Segel model with degenerate diffusion in higher dimensions, Calc. Var. Part. Differ. Equ., 35 (2009), 133–168. https://doi.org/10.1007/s00526-008-0200-7 doi: 10.1007/s00526-008-0200-7
    [23] A. Blanchet, J. Dolbeault, B. Perthame, Two-dimensional Keller-Segel model: optimal critical mass and qualitative properties of the solutions, Electron. J. Differ. Equ., 2006 (2006), 1–33. https://ejde.math.txstate.edu
    [24] V. Calvez, B. Perthame, M. Sharifi tabar, Modified Keller-Segel system and critical mass for the log interaction kernel, in Stochastic analysis and partial differential equations, vol. 429 of Contemp. Math., Amer. Math. Soc., Providence, RI, 2007, 45–62. https://doi.org/10.1090/conm/429/08229
    [25] J. Fuhrmann, J. Lankeit, M. Winkler, A double critical mass phenomenon in a no-flux-Dirichlet Keller-Segel system, J. Math. Pures Appl., 162 (2022), 124–151. https://doi.org/10.1016/j.matpur.2022.04.004 doi: 10.1016/j.matpur.2022.04.004
    [26] K. Fujie, J. Jiang, Comparison methods for a Keller-Segel-type model of pattern formations with density-suppressed motilities, Calc. Var. Part. Differ. Equ., 60 (2021), Paper No. 92, 37, https://doi.org/10.1007/s00526-021-01943-5
    [27] H.-Y. Jin, Z.-A. Wang, Boundedness, blowup and critical mass phenomenon in competing for chemotaxis, J. Differ. Equ., 260 (2016), 162–196. https://doi.org/10.1016/j.jde.2015.08.040 doi: 10.1016/j.jde.2015.08.040
    [28] H.-Y. Jin, Z.-A. Wang, Critical mass on the Keller-Segel system with signal-dependent motility, Proc. Amer. Math. Soc., 148 (2020), 4855–4873. https://doi.org/10.1090/proc/15124 doi: 10.1090/proc/15124
    [29] Y. Tao, M. Winkler, Critical mass for infinite-time aggregation in a chemotaxis model with indirect signal production, J. Eur. Math. Soc., 19 (2017), 3641–3678. https://doi.org/10.4171/JEMS/749 doi: 10.4171/JEMS/749
    [30] J. I. Tello, M. Winkler, Reduction of critical mass in a chemotaxis system by external application of a chemoattractant, Ann. Sc. Norm. Super. Pisa Cl. Sci., 12 (2013), 833–862.
    [31] M. Winkler, How unstable is spatial homogeneity in Keller-Segel systems? A new critical mass phenomenon in two- and higher-dimensional parabolic-elliptic cases, Math. Ann., 373 (2019), 1237–1282. https://doi.org/10.1007/s00208-018-1722-8 doi: 10.1007/s00208-018-1722-8
    [32] M. Winkler, Can fluid interaction influence the critical mass for taxis-driven blow-up in bounded planar domains?, Acta Appl. Math., 169 (2020), 577–591. https://doi.org/10.1007/s10440-020-00312-2 doi: 10.1007/s10440-020-00312-2
    [33] M. Winkler, A family of mass-critical Keller-Segel systems, Proc. Lond. Math. Soc., 124 (2022), 133–181. https://doi.org/10.1112/plms.12425 doi: 10.1112/plms.12425
    [34] K. Osaki, T. Tsujikawa, A. Yagi, M. Mimura, Exponential attractor for a chemotaxis-growth system of equations, Nonlinear Anal., 51 (2002), 119–144. https://doi.org/10.1016/S0362-546X(01)00815-X doi: 10.1016/S0362-546X(01)00815-X
    [35] R. B. Salako, W. Shen, Global existence and asymptotic behavior of classical solutions to a parabolic–elliptic chemotaxis system with logistic source on rn, J. Differ. Equ., 262 (2017), 5635–5690. https://doi.org/10.1016/j.jde.2017.02.011 doi: 10.1016/j.jde.2017.02.011
    [36] J. I. Tello, M. Winkler, A chemotaxis system with logistic source, Comm. Part. Differ. Equ., 32 (2007), 849–877. https://doi.org/10.1080/03605300701319003 doi: 10.1080/03605300701319003
    [37] M. Winkler, Boundedness in the higher-dimensional parabolic-parabolic chemotaxis system with logistic source, Comm. Part. Differ. Equ., 35 (2010), 1516–1537. https://doi.org/10.1080/03605300903473426 doi: 10.1080/03605300903473426
    [38] T. Xiang, Boundedness and global existence in the higher-dimensional parabolic-parabolic chemotaxis system with/without growth source, J. Differ. Equ., 258 (2015), 4275–4323. https://doi.org/10.1016/j.jde.2015.01.032 doi: 10.1016/j.jde.2015.01.032
    [39] M. Winkler, Finite-time blow-up in low-dimensional Keller-Segel systems with logistic-type superlinear degradation, Z. Angew. Math. Phys., 69 (2018), Paper No. 69, 40. https://doi.org/10.1007/s00033-018-0935-8
    [40] K. Fujie, T. Senba, Application of an Adams type inequality to a two-chemical substances chemotaxis system, J. Differ. Equ., 263 (2017), 88–148. https://doi.org/10.1016/j.jde.2017.02.031 doi: 10.1016/j.jde.2017.02.031
    [41] W. Zhang, P. Niu, S. Liu, Large time behavior in a chemotaxis model with logistic growth and indirect signal production, Nonlinear Anal. Real World Appl., 50 (2019), 484–497. https://doi.org/10.1016/j.nonrwa.2019.05.002 doi: 10.1016/j.nonrwa.2019.05.002
    [42] H. Li, Y. Tao, Boundedness in a chemotaxis system with indirect signal production and generalized logistic source, Appl. Math. Lett., 77 (2018), 108–113. https://doi.org/10.1016/j.aml.2017.10.006 doi: 10.1016/j.aml.2017.10.006
    [43] K. J. Painter, T. Hillen, Spatio-temporal chaos in a chemotaxis model, Physica D, 240 (2011), 363–375, https://doi.org/10.1016/j.physd.2010.09.011. doi: 10.1016/j.physd.2010.09.011
    [44] M. Cross, H. Greenside, Pattern formation and dynamics in nonequilibrium systems, Cambridge University Press, 2009.
    [45] T. Hillen, J. Zielinski, K. J. Painter, Merging-emerging systems can describe spatio-temporal patterning in a chemotaxis model, Discrete Contin. Dyn. Syst. Ser. B, 18 (2013), 2513–2536. https://doi.org/10.3934/dcdsb.2013.18.2513 doi: 10.3934/dcdsb.2013.18.2513
    [46] D. Horstmann, M. Winkler, Boundedness vs. blow-up in a chemotaxis system, J. Differ. Equ., 215 (2005), 52–107. https://doi.org/10.1016/j.jde.2004.10.022 doi: 10.1016/j.jde.2004.10.022
    [47] H.-Y. Jin and Z.-A. Wang, Global stability of prey-taxis systems, J. Differ. Equ., 262 (2017), 1257–1290. https://doi.org/10.1016/j.jde.2016.10.010 doi: 10.1016/j.jde.2016.10.010
    [48] H.-Y. Jin, Z.-A. Wang, Global stabilization of the full attraction-repulsion Keller-Segel system, Discrete Contin. Dyn. Syst., 40 (2020), 3509–3527. https://doi.org/10.3934/dcds.2020027 doi: 10.3934/dcds.2020027
    [49] R. Kowalczyk, Z. Szymańska, On the global existence of solutions to an aggregation model, J. Math. Anal. Appl., 343 (2008), 379–398. https://doi.org/10.1016/j.jmaa.2008.01.005 doi: 10.1016/j.jmaa.2008.01.005
    [50] G. Li, Y. Yao, Two-species competition model with chemotaxis: well-posedness, stability and dynamics, Nonlinearity, 35 (2022), 1329–1359. https://doi.org/10.1088/1361-6544/ac4a8d doi: 10.1088/1361-6544/ac4a8d
    [51] G. M. Lieberman, Second order parabolic differential equations, World Scientific Publishing Co., Inc., River Edge, NJ, 1996. https: //doi.org/10.1142/3302
    [52] P. Liu, J. Shi, Z.-A. Wang, Pattern formation of the attraction-repulsion Keller-Segel system, Discrete Contin. Dyn. Syst. Ser. B, 18 (2013), 2597–2625. https://doi.org/10.3934/dcdsb.2013.18.2597 doi: 10.3934/dcdsb.2013.18.2597
    [53] M. Ma, C. Ou, Z.-A. Wang, Stationary solutions of a volume-filling chemotaxis model with logistic growth and their stability, SIAM J. Appl. Math., 72 (2012), 740–766. https://doi.org/10.1137/110843964 doi: 10.1137/110843964
    [54] M. Ma, Z.-A. Wang, Global bifurcation and stability of steady states for a reaction-diffusion-chemotaxis model with volume-filling effect, Nonlinearity, 28 (2015), 2639–2660. https://doi.org/10.1088/0951-7715/28/8/2639 doi: 10.1088/0951-7715/28/8/2639
    [55] C. Mu, W. Tao, Z.-A. Wang, Global dynamics and spatiotemporal heterogeneity of accelerated preytaxis models, preprint.
    [56] J. D. Murray, Mathematical Biology I: An introduction, vol. 17 of Interdisciplinary Applied Mathematics, 3rd edition, Springer-Verlag, New York, 2002.
    [57] K. J. Painter, T. Hillen, Volume-filling and quorum-sensing in models for chemosensitive movement, Can. Appl. Math. Q., 10 (2002), 501–543.
    [58] P. Quittner, P. Souplet, Superlinear parabolic problems. Blow-up, global existence and steady states, Birkhäuser, Basel, 2019.
    [59] Y. Tao, M. Winkler, Boundedness and decay enforced by quadratic degradation in a three-dimensional chemotaxis–fluid system, Z. Angew. Math. Phys., 66 (2015), 2555–2573. https://doi.org/10.1007/s00033-015-0541-y doi: 10.1007/s00033-015-0541-y
    [60] J. Wang, M. Wang, The dynamics of a predator-prey model with diffusion and indirect prey-taxis, J. Dyn. Differ. Equ., 32 (2020), 1291–1310. https://doi.org/10.1007/s10884-019-09778-7 doi: 10.1007/s10884-019-09778-7
    [61] M. Wang, Note on the Lyapunov functional method, Appl. Math. Lett., 75 (2018), 102–107. https://doi.org/10.1016/j.aml.2017.07.003 doi: 10.1016/j.aml.2017.07.003
    [62] Z. Wang, T. Hillen, Classical solutions and pattern formation for a volume filling chemotaxis model, Chaos, 17 (2007), 037108, 13. https://doi.org/10.1063/1.2766864 doi: 10.1063/1.2766864
    [63] M. Winkler, Aggregation vs. global diffusive behavior in the higher-dimensional Keller-Segel model, J. Differ. Equ., 248 (2010), 2889–2905. http://dx.doi.org/10.1016/j.jde.2010.02.008 doi: 10.1016/j.jde.2010.02.008
  • This article has been cited by:

    1. Wenjie Zhang, Lu Xu, Qiao Xin, Global boundedness of a higher-dimensional chemotaxis system on alopecia areata, 2023, 20, 1551-0018, 7922, 10.3934/mbe.2023343
  • 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(2358) PDF downloads(161) Cited by(1)

Figures and Tables

Figures(3)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog