Loading [MathJax]/jax/output/SVG/jax.js

Stability estimates for scalar conservation laws with moving flux constraints

  • Received: 01 October 2016 Revised: 01 January 2017
  • Primary: 35L65; Secondary: 35B35

  • We study well-posedness of scalar conservation laws with moving flux constraints. In particular, we show the Lipschitz continuous dependence of BV solutions with respect to the initial data and the constraint trajectory. Applications to traffic flow theory are detailed.

    Citation: Maria Laura Delle Monache, Paola Goatin. Stability estimates for scalar conservation laws with moving flux constraints[J]. Networks and Heterogeneous Media, 2017, 12(2): 245-258. doi: 10.3934/nhm.2017010

    Related Papers:

    [1] Kenneth H. Karlsen, Süleyman Ulusoy . On a hyperbolic Keller-Segel system with degenerate nonlinear fractional diffusion. Networks and Heterogeneous Media, 2016, 11(1): 181-201. doi: 10.3934/nhm.2016.11.181
    [2] Piotr Biler, Grzegorz Karch, Jacek Zienkiewicz . Morrey spaces norms and criteria for blowup in chemotaxis models. Networks and Heterogeneous Media, 2016, 11(2): 239-250. doi: 10.3934/nhm.2016.11.239
    [3] José Antonio Carrillo, Yingping Peng, Aneta Wróblewska-Kamińska . Relative entropy method for the relaxation limit of hydrodynamic models. Networks and Heterogeneous Media, 2020, 15(3): 369-387. doi: 10.3934/nhm.2020023
    [4] Panpan Xu, Yongbin Ge, Lin Zhang . High-order finite difference approximation of the Keller-Segel model with additional self- and cross-diffusion terms and a logistic source. Networks and Heterogeneous Media, 2023, 18(4): 1471-1492. doi: 10.3934/nhm.2023065
    [5] Raul Borsche, Axel Klar, T. N. Ha Pham . Nonlinear flux-limited models for chemotaxis on networks. Networks and Heterogeneous Media, 2017, 12(3): 381-401. doi: 10.3934/nhm.2017017
    [6] Elisabeth Logak, Isabelle Passat . An epidemic model with nonlocal diffusion on networks. Networks and Heterogeneous Media, 2016, 11(4): 693-719. doi: 10.3934/nhm.2016014
    [7] Patrick Henning, Mario Ohlberger . The heterogeneous multiscale finite element method for advection-diffusion problems with rapidly oscillating coefficients and large expected drift. Networks and Heterogeneous Media, 2010, 5(4): 711-744. doi: 10.3934/nhm.2010.5.711
    [8] Toyohiko Aiki, Kota Kumazaki . Uniqueness of solutions to a mathematical model describing moisture transport in concrete materials. Networks and Heterogeneous Media, 2014, 9(4): 683-707. doi: 10.3934/nhm.2014.9.683
    [9] Junjie Wang, Yaping Zhang, Liangliang Zhai . Structure-preserving scheme for one dimension and two dimension fractional KGS equations. Networks and Heterogeneous Media, 2023, 18(1): 463-493. doi: 10.3934/nhm.2023019
    [10] Jinyi Sun, Weining Wang, Dandan Zhao . Global existence of 3D rotating magnetohydrodynamic equations arising from Earth's fluid core. Networks and Heterogeneous Media, 2025, 20(1): 35-51. doi: 10.3934/nhm.2025003
  • We study well-posedness of scalar conservation laws with moving flux constraints. In particular, we show the Lipschitz continuous dependence of BV solutions with respect to the initial data and the constraint trajectory. Applications to traffic flow theory are detailed.



    In this paper, we consider the global solution to the Cauchy problem of fractional drift diffusion system with power-law nonlinearity,

    {tv+Λαv=(vmϕ),t>0,xRN,tw+Λαw=(wmϕ),t>0,xRN,Δϕ=vw,t>0,xRN,v(x,0)=v0(x),w(x,0)=w0(x),xRN, (1.1)

    where m1 is an integer, v(x,t),w(x,t) are the densities of negatively and positively charged particles, ϕ(x,t) is the electric potential determined by the Poisson equation Δϕ=vw. The difficulties mainly come from higher-order nonlinear couplings.

    By the fundamental solution of Laplacian:

    ΦN(x)={12|x|,N=1,12πln|x|,N=2,1N(N2)ω(N)|x|N2,N3, (1.2)

    where ω(N) denotes the volume of the unit ball in RN, the electric potential ϕ can be expressed by the convolution:

    ϕ=(Δ)1(wv)=ΦN(wv)=RNΦN(xy)(wv)(y)dy. (1.3)

    Λ=Δ is the Calderón-Zygmund operator, and the fractional Laplacian Λα=(Δ)α2 with 1<α<2N is a non-local fractional differential operator defined as Eq (1.4)

    Λαv(x)=F1|ξ|αFv(ξ), (1.4)

    where F and F1 are the Fourier transform and its inverse [1].

    In probabilistic terms, replacing the Laplacian Δ with its fractional power Λα=(Δ)α2, it leads to interesting and largely open questions of extensions of results for Brownian motion driven stochastic equations to those driven by Lévy αstable flights.

    In the physical literature, such fractal anomalous diffusions have been recently enthusiastically embraced by a slew of investigators in the context of hydrodynamics, acoustics, trapping effects in surface diffusion, statistical mechanics, relaxation phenomena, and biology [2].

    An important technical difficulty is that the densities of the semigroups generated by Λα=(Δ)α2 do not decay rapidly in xRN as is the case of the heat semigroup S(t)=etΔ (α=2), the Gauss-Weierstrass kernel Kt(x)=F1(et|ξ|2) decays exponentially while the densities F1(et|ξ|α)(0<α<2) of non-Gaussian Lévy αstable semigroups Sα(t)=et(Δ)α2 have only an algebraic decay rate |x|Nα.

    For a more general nonlinear term in Eq (1.1), the motivation is the Keller-Segel model [3,4], a prototype of cross-diffusion models related to pattern formation, it describes the time and space dynamics of the density of cells (or organisms) n(t,x) interacting with a chemoattractant S(t,x) according to the following system:

    {tn=x(Dn(n,s)xnχ(n,s)nxs)+F(n,s),ts=Ds(n,s)Δs+G(n,s), (1.5)

    where F and G are the source terms related to interactions [5]. The positive definite nonlinear terms Dn(n,s) and Ds(n,s) are the diffusivity of the chemoattractant and of the cells, respectively. In many applications the cross-diffusion function χ(n,s) has a complicated structure, and even it has a very simple structure, for example, a polynomial χ(n,s)=nm, it fails to satisfy a global Lipschitz condition.

    For m=1, Eq (1.1) becomes a fractional drift-diffusion system Eq (1.6),

    {tv+Λαv=(vϕ),t>0,xRN,tw+Λαw=(wϕ),t>0,xRN,Δϕ=vw,t>0,xRN,v(x,0)=v0(x),w(x,0)=w0(x),xRN, (1.6)

    Zhao-Liu [6] established global well-posedness and asymptotic stability of mild solutions for the Cauchy problem Eq (1.5) with small initial data in critical Besov spaces, and proved the regularizing-decay rate estimates which imply that mild solutions are analytic in space variables. Ogawa-Yamamoto [7] considered the global existence and asymptotic behavior of solutions for the Cauchy problem Eq (1.5), they showed that the time- global existence of the solutions with large initial data in Lebesgue space Lp(RN) and Sobolev space Wα,p(RN) and obtained the asymptotic expansion of the solution up to the second terms as t+.

    For α=2, Eq (1.6) corresponds to the usual drift-diffusion system,

    {tvΔv=(vϕ),t>0,xRN,twΔw=(wϕ),t>0,xRN,Δϕ=vw,t>0,xRN,v(x,0)=v0(x),w(x,0)=w0(x),xRN, (1.7)

    it has been studied widely [8,9,10,11,12,13,14]. Karch [15] considered the Cauchy problem of a scalar equation with a bilinear operator B

    {tu=Δu+B(u,u),t>0,xRN,u(x,0)=u0(x),xRN.

    For w=0, Eq (1.6) corresponds to the generalized Keller-Segel model of chemotaxis:

    {tv+Λαv=(vϕ),t>0,xRN,Δϕ=v,t>0,xRN,v(x,0)=v0(x),xRN. (1.8)

    For 1<α<2, Escudero [16] proved that Eq (1.8) admits a one-dimensional global solution (the same result also holds for α=2), Biler-Karch [17] studied the Blowup solutions to generalized Keller-Segel model, and Biler-Wu [18] considered two-dimensional chemotaxis models with fractional diffusion. For α=2, Biler-Boritchev-Karch et al., considered the concentration phenomena [19] and gave sharp Sobolev estimates for concentration of solution [20] to the diffusive aggregation model:

    tvεΔv=(vKv)

    with the Poisson kernel function K from the equation Δϕ=v.

    Wu-Zheng [21] considered the parabolic-parabolic system corresponding to the parabolic-elliptic system Eq (1.8), the Keller-Segel system with fractional diffusion generalizing the Keller-Segel model of chemotaxis

    {tu+Λαu=±(uϕ),t>0,xRN,εtϕ+Λαϕ=u,t>0,xRN,u(x,0)=u0(x),v(x,0)=v0(x),xRN, (1.9)

    for initial data (u0,v0) in the critical Fourier-Herz space ˙B22αq(RN)×˙B2αq(RN) with 2q for ε>0 and 1<α2.

    For the fractional evolution equations with higher order nonlinearity, Miao-Yuan-Zhang [22] studied the Cauchy problem for the semilinear fractional power dissipative equation

    {tu+Λαu=F(u),t>0,xRN,u(x,0)=u0(x),xRN, (1.10)

    with the nonlinear term F(u)=f(u) or F(u)=Q(D)f(u), where Q(D) is a homogeneous pseudo differential operator and f(u)=|u|bu or |u|b1u+|u|b2u with b>0,b1>0 and b2>0. For example, the equation in Eq (1.10) contains the semilinear fractional power dissipative equation tu+Λαu=±|u|bu, the generalized convection-diffusion equation tu+Λαu=a(|u|bu), and so on.

    Following the idea of Karch [15], due to the fractional heat semigroup Sα(t)=etΛα and the well-known Duhamel principle, we rewrite the system Eq (1.1) as a system of integral equations

    {v(t)=Sα(t)v0+B(v,,v,w),w(t)=Sα(t)w0+B(w,,w,v), (1.11)

    where

    B(v,,vm,w)=t0Sα(tτ)(vmϕ)(τ)dτ,ϕ=(Δ)1(wv). (1.12)

    A solution of Eq (1.11) and Eq (1.12) is called a mild solution of Eq (1.1).

    Inspired by the contributions summarized in the above items, we aim to extend the results to the system Eq (1.1) with higher-order nonlinear terms (vmϕ) and (wmϕ). The goal of this article is to prove the global well-posedness of mild solutions to the Cauchy problem Eq (1.1) with small initial data in critical Besov spaces. When m=1 in the higher order nonlinear term (vmϕ), we recover the result proved in [6]. The outline of the rest of the article is as follows. In Section 2 we give the definition of homogeneous Besov space by the fractional heat semigroup operator and present some useful estimates. In Section 3 we establish the global existence and uniqueness of the mild solution. In Section 4 we discuss the asymptotic stability of the mild solution. In Section 5 we give the regularizing-decay rate estimates of the mild solution. In Section 6 we consider a fractional drift diffusion system with a generalized electric potential equation and we also give the global existence and asymptotic stability of the mild solution.

    Let S(RN) be the Schwartz space and S(RN) be its dual. Now, we introduce a definition of the homogeneous Besov space by the semigroup operator Sα(t)=etΛα.

    Definition 2.1. [6] Let l>0 and 1p. Define

    ˙Blp,(RN)={fS(RN):SαfC((0,+),Lp),supt>0tlα||Sαf||Lp<} (2.1)

    with the norm

    ||f||˙Blp,(RN)=supt>0tlα||Sα(t)f||Lp. (2.2)

    (˙Blp,(RN),||||˙Blp,) is a Banach space.

    If (v(x,t),w(x,t)) is a solution of the Cauchy problem Eq (1.1), for any λ>0, denote

    vλ(x,t)=λαmv(λx,λαt),wλ(x,t)=λαmw(λx,λαt), (2.3)

    (vλ(x,t),wλ(x,t)) is also a solution of the Cauchy problem Eq (1.1) with the initial data

    (vλ(x,0),wλ(x,0))=(λαmv0(λx),λαmw0(λx)),

    then (vλ(x,t),wλ(x,t)) is called a self-similar solution to Eq (1.1). We can verify that ˙Bαm+npp,(Rn) is a critical space, i.e., the self-similar solution is invariant under the norm ||||˙Bαm+npp,, which defined in [6], for initial data (v0(x),w0(x)) of the system Eq (1.1). In the case the index sc:=npαm provides the minimal regularity for the initial data to ensure the well-posedness of the Cauchy problem Eq (1.1). In order to find a critical space for the solutions of the Cauchy problem Eq (1.1), we define some time-weighted space-time space.

    Let X be a Banach space and I be a finite or infinite interval. We define the time-weighted space-time Banach space,

    Cσ(I;X)={fC(I;X):supt>0t1σ||f(t)||X<} (2.4)

    with the norm ||f||Cσ(I;X)=supt>0t1σ||f(t)||X. The corresponding homogeneous time-weighted space-time Banach space,

    ˙Cσ(I;X)={fCσ(I;X):limt0t1σ||f(t)||X=0}. (2.5)

    We denote C([0,);X) by the set of bounded maps from [0,) to X which are continuous for t>0 and weakly continuously for t=0.

    For initial data (v0(x),w0(x)) in critical homogeneous Besov space ˙Bαm+Npp,(RN) with minimal regularity, we want to find a mild solution of the Cauchy problem Eq (1.1) and discuss the global existence of mild solution in the following critical space,

    X=C([0,),˙Bαm+Npp,(RN))CmαpαpmN([0,),Lp(RN)) (2.6)

    with the norm

    ||u||X=supt>0||u(t)||˙Bαm+Npp,(RN)+supt>0t1mNαp||u(t)||Lp(RN). (2.7)

    For the Laplacian operator Δ and the Calderón-Zygmund operator Λ=Δ, we have the following classical Hardy-Littlewood-Sobolev inequality.

    Lemma 2.2. [23,24] Let 1<p<N, the nonlocal operator (Δ)12 is bounded from Lp(RN) to LNpNp(RN), i.e., fLp(RN),

    ||(Δ)12f||LNpNp(RN)C(N,p)||f||Lp(RN), (2.8)
    ||(Δ)1f||LNpNp(RN)C(N,p)||f||Lp(RN). (2.9)

    For the fractional power operator Λα=(Δ)α2 and the semigroup operator Sα(t)=etΛα, we first consider the Cauchy problem for the homogeneous linear fractional heat equation

    {tu+Λαu=0,t>0,xRN,u(x,0)=u0(x),xRN. (2.10)

    By the Fourier transform the solution can be written as:

    u(t,x)=F1(et|ξ|αFu0(ξ))=F1(et|ξ|α)u0(x)=Kt(x)u0(x)=Sα(t)u0(x), (2.11)

    where the fractional heat kernel Eq (2.12),

    Kt(x)=(2π)N2RNeixξet|ξ|αdξ=tNαK(xt1α), (2.12)

    the function K(x)L(RN)C0(RN), where C0(RN) denotes the space of functions fC(RN) satisfying that lim|x|f(x)=0.

    For the semigroup operator Sα(t) we have LpLq estimates

    Lemma 2.3. [9] Let 1pq. Then, fLp(RN),

    ||Sα(t)f||LqC(N,α)tNα(1p1q)||f||Lp, (2.13)
    ||ΛγSα(t)f||LqC(N,α)tγαNα(1p1q)||f||Lp, (2.14)

    for α>0 and γ>0.

    Following the work of Kato [25,26] and Lemarie-Rieusset [23] for the Navier-Stokes problem, Miao-Yuan [27] gave a general existence and uniqueness result for an abstract operator equation via a contraction argument.

    Lemma 2.4. [27] Let X be a Banach space and B:X×X××XX be a (m+1)linear continuous operator satisfying

    ||B(u1,u2,,um+1)||XK||u1||X||u2||X||um+1||X, (2.15)

    u1,u2,,um+1X for some constant K>0. Let ε>0 be such that (m+1)(2ε)mK<1. Then for every yX with ||y||Xε the equation

    u=y+B(u,u,,u) (2.16)

    has a unique solution uX satisfying that ||u||X2ε. Moreover, the solution u depends continuously on y in the sense that, if ||y||Xε and v=y1+B(v,v,,v), ||v||X2ε, then

    ||uv||X11(m+1)(2ε)mK||yy1||X. (2.17)

    We will use the Lemma to prove the global-in-time existence and uniqueness of the mild solutions to the Cauchy problem Eq (1.1) in the mixed time-space Besov space.

    In this section we give the global existence and uniqueness of mild solution to the Cauchy problem Eq (1.1).

    Theorem 3.1. Let N2 be a positive integer, 1<α2N and

    max{1,mNα}<p<min{N,m(m+1)Nα}. (3.1)

    If (v0,w0)˙Bαm+Npp,(RN), there exists ε>0 such that if ||(v0,w0)||˙Bαm+Npp,ε, the Cauchy problem Eq (1.1) has a unique global mild solution (v,w)X such that ||(v,w)||X2ε. Moreover, the solution depends continuously on initial data in the following sense: let (˜v,˜w)X be the solution of Eq (1.1) with initial data (˜v0,˜w0) such that ||(˜v0,˜w0)||˙Bαm+Npp,(RN)ε, then there is a constant C such that

    ||(v˜v,w˜w)||XC||(v0˜v0,w0˜w0)||˙Bαm+Npp,(RN).

    For the integral system Eqs (1.11) and (1.12) we first consider the term Sα(t)v0=etΛαv0.

    Lemma 3.2. Let v0(x)˙Bαm+Npp,(RN) and Eq (3.1) hold true. Then Sα(t)v0X and

    ||Sα(t)v0||XC(N,α)||v0||˙Bαm+Npp,(RN). (3.2)

    Proof. According to the definition of the norm ||||˙Bαm+Npp,(RN) and LpLq estimates for the semigroup operator Sα(t)=etΛα, we have

    ||Sα(t)v0||˙Bαm+Npp,(RN)=sups>0s1mNαp||Sα(s)Sα(t)v0||Lp=sups>0s1mNαp||Sα(t)Sα(s)v0||LpC(N,α)sups>0s1mNαp||Sα(s)v0||Lp=C(N,α)||v0||˙Bαm+Npp,(RN),

    and

    supt>0t1mNαp||Sα(t)v0||Lp=||v0||˙Bαm+Npp,(RN).

    Therefore, we have

    Sα(t)v0L((0,),˙Bαm+Npp,(RN)),t1mNαpSα(t)v0L((0,),Lp(RN)).

    Moreover, following the method of [23] (Proposition 4.4, P33) we obtain that

    Sα(t)v0C([0,),˙Bαm+Npp,(RN)).

    On the other hand, from v0(x)˙Bαm+Npp,(RN) and Definition 2.1, we have

    Sα(t)v0C((0,),Lp(RN)),t1mNαpSα(t)v0C((0,),Lp(RN)).

    Hence, we have Sα(t)v0X and Eq (3.2) holds true.

    Lemma 3.3. Let (v,w)X and Eq (3.1) hold true. Then B(v,,v,w)X and

    ||B(v,,v,w)||XC(N,α,p)||v||mX||vw||X. (3.3)

    Proof. According to the definition of the norm ||||˙Bαm+Npp,(RN), we have

    ||B(v,,v,w)(t)||˙Bαm+Npp,(RN)=sups>0s1mNαp||Sα(s)B(v,,v,w)(t)||Lp,

    by the expression Eq (1.12) of B(v,,v,w)(t), that is,

    B(v,,vm,w)=t0Sα(tτ)(vmϕ)(τ)dτ,ϕ=(Δ)1(wv), (3.4)

    hence, by the Minkowski inequality, we get

    ||B(v,,v,w)(t)||˙Bαm+Npp,(RN)=sups>0s1mNαp||Sα(s)t0Sα(tτ)(vmϕ)(τ)dτ||Lpt0sups>0s1mNαp||Sα(s)Sα(tτ)(vmϕ)(τ)||Lpdτ. (3.5)

    For 0<stτ, using the LpLq estimates Eq (2.13) and Eq (2.14) for the semigroup operator Sα(t)=etΛα, we have

    sup0<stτs1mNαp||Sα(s)Sα(tτ)(vmϕ)(τ)||LpC(N,α)(tτ)1mNαp||Sα(tτ)(vmϕ)(τ)||Lp=C(N,α)(tτ)1mNαp||Sα(tτ)(vmϕ)(τ)||LpC(N,α,p)(tτ)1mNαp(tτ)mNαp||(vmϕ)](τ)||LNp(m+1)NpC(N,α,p)(tτ)1m(m+1)Nαp||v||mLp||ϕ(τ)||LNpNp,

    the last inequality comes from the Hölder inequality for the product vvv(vw) and mp+NpNp=(m+1)NpNp. Using the classical Hardy-Littlewood-Sobolev inequality Eq (2.8) and Eq (2.9), we have Eq (3.6):

    sup0<stτs1mNαp||Sα(s)Sα(tτ)[vm(Δ)1(vw)](τ)||LpC(N,α,p)(tτ)1m(m+1)Nαp||v(τ)||mLp||(vw)(τ)||Lp. (3.6)

    For s>tτ, using the LpLq estimates Eq (2.13) and Eq (2.14) for the semigroup operator Sα(t)=etΛα, we have

    sups>tτs1mNαp||Sα(s)Sα(tτ)(vmϕ)(τ)||Lp=sups>tτs1mNαp||Sα(t+sτ)(vmϕ)(τ)||LpC(N,α)sups>tτs1mNαp(t+sτ)mNαp||vmϕ(τ)||LNp(m+1)NpC(N,α)sups>tτs1mNαp(t+sτ)mNαp||v||mLp||ϕ(τ)||LNpNp.

    From the condition Eq (3.1): max{1,mNα}<p<min{N,m(m+1)Nα} and s>tτ, the function f(s)=s1mNαp(t+sτ)mNαp has the maximum

    maxs>tτf(s)=f(1mNαp(m+1)Nαp1m(tτ))=C(tτ)1m(m+1)Nαp,

    where C is a constant, by Eq (2.9) we have

    sups>tτs1mNαp||Sα(s)Sα(tτ)[vm(Δ)1(vw)](τ)||LpC(N,α,p)(tτ)1m(m+1)Nαp||v(τ)||mLp||(vw)(τ)||Lp. (3.7)

    Together with Eq (3.6) and Eq (3.7) we have:

    sups>0s1mNαp||Sα(s)Sα(tτ)[vm(Δ)1(vw)](τ)||LpC(N,α,p)(tτ)1m(m+1)Nαp||v(τ)||mLp||(vw)(τ)||Lp. (3.8)

    Putting Eq (3.8) into Eq (3.5), we have

    ||B(v,,v,w)(t)||˙Bαm+Npp,(RN)C(N,α,p)t0(tτ)1m(m+1)Nαp||v(τ)||mLp||(vw)(τ)||LpdτC(N,α,p)supτ>0(τ1mNαp||v(τ)||Lp)msupτ>0(τ1mNαp||(vw)(τ)||Lp)×t0(tτ)1m(m+1)Nαpτ(m+1)Nαp1m1dτC(N,α,p)||v||mX||vw||Xt0(tτ)1m(m+1)Nαpτ(m+1)Nαp1m1dτC(N,α,p)||v||mX||vw||X,

    in the last inequality we use the fact that the Beta function

    t0(tτ)1m(m+1)Nαpτ(m+1)Nαp1m1dτ=B(m+1m(m+1)Nαp,(m+1)Nαp1m)

    converges to a constant, since the condition Eq (3.1) implies that

    m+1m(m+1)Nαp=m+1mp(pmNα)>0,(m+1)Nαp1m=1mp(m(m+1)Nαp)>0.

    Therefore, we have

    ||B(v,,v,w)(t)||˙Bαm+Npp,(RN)C(N,α,p)||v||mX||vw||X. (3.9)

    Next, we consider the estimate of ||B(v,,v,w)(t)||Lp. From Eq (1.12) we have

    ||B(v,,v,w)(t)||Lp=||t0Sα(tτ)(vmϕ)(τ)dτ||LpC(N,α)t0(tτ)mNαp||vm(Δ)1(vw)](τ)||LNp(m+1)NpdτC(N,α)t0(tτ)mNαp||v||mLp||(Δ)1(vw)](τ)||LNpNpdτC(N,α,p)||v||mX||vw||Xt0(tτ)mNαpτ1m1+(m+1)NαpdτC(N,α,p)||v||mX||vw||Xt1m+Nαp,

    thus,

    supt>0t1mNαp||B(v,,v,w)(t)||LpC(N,α,p)||v||mX||vw||X. (3.10)

    In order to prove that B(v,,v,w)X, it suffices to prove that B(v,,v,w) is continuous for t>0 and weakly continuous for t=0 in ˙Bαm+Npp,(RN), and it is continuous for t0 in Lp(RN).

    For any 0<t1<t2, due to Eq (3.4) we have

    B(v,,v,w)(t2)B(v,,v,w)(t1)=t10[Sα(t2τ)Sα(t1τ)][vm(Δ)1(vw)](τ)dτ+t2t1Sα(t2τ)[vm(Δ)1(vw)](τ)dτ:=I(t1,t2)+II(t1,t2). (3.11)

    Similar to the estimate of ||B(v,,v,w)(t)||˙Bαm+Npp,(RN), we have

    ||II(t1,t2)||˙Bαm+Npp,(RN)=sups>0s1mNαp||Sα(s)II(t1,t2)||Lpt2t1sups>0s1mNαp||Sα(s)Sα(t2τ)[vm(Δ)1(vw)](τ)||LpdτC(N,α,p)||v||mX||vw||Xt2t1(t2τ)1m(m+1)Nαpτ(m+1)Nαp1m1dτC(N,α,p)||v||mX||vw||Xt11m+(m+1)Nαp1t2t1(t2τ)1m(m+1)NαpdτC(N,α,p)||v||mX||vw||Xt11m+(m+1)Nαp1(t2t1)1+1m(m+1)Nαp,

    the condition Eq (3.1) implies that 1+1m(m+1)Nαp>0, hence as t2t1,

    ||II(t1,t2)||˙Bαm+Npp,(RN)=sups>0s1mNαp||Sα(s)II(t1,t2)||Lp0. (3.12)

    According to the property of semigroup,

    Sα(t2τ)Sα(t1τ)=[Sα(t2t1)I]Sα(t1τ), (3.13)

    for ϕ=(Δ)1(wv) we get

    ||I(t1,t2)||˙Bαm+Npp,(RN)=sups>0s1mNαp||Sα(s)I(t1,t2)||Lpt10sups>0s1mNαp||Sα(s)[Sα(t2t1)I]Sα(t1τ)(vmϕ)(τ)||Lpdτ=t10sups>0s1mNαp||t2t10ΛαSα(μ)Sα(s)Sα(t1τ)(vmϕ)(τ)dμ||Lpdτ=t10sups>0s1mNαp||t2t10Sα(μ)ΛαSα(s)Sα(t1τ)(vmϕ)(τ)dμ||Lpdτt10sups>0s1mNαpt2t10||Sα(μ)ΛαSα(s)Sα(t1τ)(vmϕ)(τ)||Lpdμdτ, (3.14)

    by the LpLq estimates Eq (2.13) and Eq (2.14) for the semigroup operator Sα(t)=etΛα, we have

    t2t10||Sα(μ)ΛαSα(s)Sα(t1τ)(vmϕ)(τ)||LpdμC(N,α)t2t10μmNαpdμ||ΛαSα(s)Sα(t1τ)(vmϕ)(τ)||LNp(m+1)Np=C(N,α)(t2t1)1mNαp||ΛαSα(s)Sα(t1τ)(vmϕ)(τ)||LNp(m+1)Np. (3.15)

    For 0<st1τ, we have

    sup0<st1τs1mNαp||ΛαSα(s)Sα(t1τ)(vmϕ)(τ)||LNp(m+1)Np=sup0<st1τs1mNαp||Sα(s)ΛαSα(t1τ)(vmϕ)(τ)||LNp(m+1)NpC(N,α)sup0<st1τs1mNαp(t1τ)1||(vmϕ)(τ)||LNp(m+1)NpC(N,α)(t1τ)1mNαp1||v||mLp||ϕ||LNpNpC(N,α,p)(t1τ)1mNαp1||v||mLp||vw||Lp. (3.16)

    For s>t1τ, we have

    sups>t1τs1mNαp||ΛαSα(s)Sα(t1τ)(vmϕ)(τ)||LNp(m+1)Np=sups>t1τs1mNαp||ΛαSα(t1τ+s)(vmϕ)(τ)||LNp(m+1)NpC(N,α)sups>t1τs1mNαp(t1τ+s)1||vmϕ||LNp(m+1)NpC(N,α,p)(t1τ)1mNαp1||v||mLp||vw||Lp. (3.17)

    Putting Eqs (3.15)–(3.17) into Eq (3.14), we have

    ||I(t1,t2)||˙Bαm+Npp,(RN)C(t2t1)1mNαpt10(t1τ)1mNαp1||v(τ)||mLp||(vw)(τ)||LpdτC(t2t1)1mNαpsupτ>0(τ1mNαp||v(τ)||Lp)msupτ>0(τ1mNαp||(vw)(τ)||Lp)×t10(t1τ)1mNαp1τ(m+1)Nαp1m1dτC(t2t1)1mNαp||v||mX||vw||XBtmNαp11, (3.18)

    where C=C(N,α,p) the Beta function B=B(1mNαp,(m+1)Nαp1m) converges due to the condition Eq (3.1), thus we have

    ||I(t1,t2)||˙Bαm+Npp,(RN)C||v||mX||vw||X(t2t1)1mNαptmNαp11, (3.19)

    that is,

    ||I(t1,t2)||˙Bαm+Npp,(RN)=sups>0s1mNαp||Sα(s)I(t1,t2)||Lp0ast2t1. (3.20)

    Putting Eq (3.12) and Eq (3.20) into Eq (3.11) we have

    ||B(v,,v,w)(t1)B(v,,v,w)(t2)||˙Bαm+Npp,(RN)0ast2t1. (3.21)

    This means that B(v,,v,w) is continuous for t>0 in ˙Bαm+Npp,(RN).

    Similarly, we can prove that B(v,,v,w) is weakly continuous for t=0 in ˙Bαm+Npp,(RN) and it is continuous for t0 in Lp(RN). Therefore, we have

    B(v,,v,w)C([0,),˙Bαm+Npp,(RN))CmαpαpmN([0,),Lp(RN)), (3.22)

    that is, B(v,,v,w)X and Eq (3.3) holds true, i.e.,

    ||B(v,,v,w)||XC(N,α,p)||v||mX||vw||X. (3.23)

    This ends the proof of Lemma 3.3.

    The proof of Theorem 3.1. Now for the integral system Eq (1.11) and Eq (1.12) from the Cauchy problem Eq (1.1), we have

    (v(t),w(t))=Sα(t)(v0,w0)+(B(v,,v,w),B(w,,w,v)), (3.24)

    in Lemma 3.2 and Lemma 3.3 we deal with the terms Sα(t)(v0,w0) and

    B(v,,v,w)=t0Sα(tτ)[vm(Δ)1(vw)](τ)dτ,B(w,,w,v)=t0Sα(tτ)[wm(Δ)1(wv)](τ)dτ,

    respectively. For the Banach space X and multi-linear operator B(v,,v,w), which satisfies the estimate Eq (3.23), following the Lemma 2.4, for every (v0,w0)˙Bαm+Npp,(RN), there exists ε>0 such that (m+1)(2ε)mC(N,α,p)<1, then Eq (3.24) has a unique solution (v,w)X such that ||(v,w)||X2ε. Therefore, the Cauchy problem Eq (1.1) has a unique global-in-time mild solution in the mixed time-space Besov space. This completes the proof of Theorem 3.1.

    Theorem 4.1. Let N be a positive integer, 1<α2N and Eq (3.1) hold true and (v,w) and (˜v,˜w) be two mild solutions of the Cauchy problem Eq (1.1) described in Theorem 3.1 corresponding to initial conditions (v0,w0) and (˜v0,˜w0), respectively. If (v0,w0),(˜v0,˜w0)˙Bαm+Npp,(RN) such that

    limt||Sα(t)(v0˜v0,w0˜w0)||˙Bαm+Npp,(RN)=0, (4.1)

    then, we have the following asymptotic stability

    limt(||(v˜v,w˜w)||˙Bαm+Npp,(RN)+tαmNp||(v˜v,w˜w)||Lp(RN))=0. (4.2)

    Proof. Since (v0,w0),(˜v0,˜w0)˙Bαm+Npp,(RN), by Theorem 3.1, there exists a constant ε>0 such that if ||(v0,w0),(˜v0,˜w0)||˙Bαm+Npp,ε, then the mild solutions (v,w) and (˜v,˜w) satisfy that ||(v,w),(˜v,˜w)||X2ε. From Eq (1.11) and Eq (1.12) we have

    {v˜v=Sα(t)(v0˜v0)+m1k=0Bk(v˜v,v,˜v,vw)+Bm(˜v,(v˜v)(w˜w)),w˜w=Sα(t)(w0˜w0)+m1k=0Bk(w˜w,w,˜w,wv)+Bm(˜w,(w˜w)(v˜v)),

    where

    Bk(v˜v,v,˜v,vw)=B(v˜v,v,,vk,˜v,,˜vm1k,vw)=t0Sα(tτ)[(v˜v)vk˜vm1k(Δ)1(vw)](τ)dτ, (4.3)
    Bm(˜v,(v˜v)(w˜w))=B(˜v,,˜vm,(v˜v)(w˜w))=t0Sα(tτ)[˜vm(Δ)1((v˜v)(w˜w))](τ)dτ. (4.4)

    By the definition of ˙Bαm+Npp,(RN)norm, we have

    ||v˜v||˙Bαm+Npp,(RN)||Sα(t)(v0˜v0)||˙Bαm+Npp,(RN)+m1k=0Ik+Im, (4.5)

    where

    (Ik,Im)=||(Bk(v˜v,v,˜v,vw),B(˜v,(v˜v)(w˜w)))||˙Bαm+Npp,(RN).

    For a constant θ(0,1) determined in later we have

    Ik=sups>0s1mNαp||Sα(s)t0Sα(tτ)[(v˜v)vk˜vm1k(Δ)1(vw)](τ)dτ||Lpt0sups>0s1mNαp||Sα(s)Sα(tτ)[(v˜v)vk˜vm1k(Δ)1(vw)](τ)||Lpdτ(θt0+tθt)sups>0s1mNαp||Sα(t+sτ)[(v˜v)vk˜vm1k(Δ)1(vw)]||Lpdτ:=Ik1+Ik2. (4.6)

    In the procedure of estimate of Eq (3.5), instead of the product vvv(vw) with m+1 exponents such that mp+NpNp=(m+1)NpNp, use the Hölder inequality for the product (v˜v)vk˜vm1k(vw) with m+1 exponents such that 1p+kp+m1kp+NpNp=(m+1)NpNp, we can prove that

    Ik1Cθt0(tτ)1m(m+1)Nαp||v˜v||Lp||v||kLp||˜v||m1kLp||vw||LpdτCεmθ0(1η)1m(m+1)Nαpη11m+(m+1)Nαp(tη)1mNαp||v(tη)˜v(tη)||Lpdη, (4.7)

    and

    Ik2Ctθt(tτ)1m(m+1)Nαp||v˜v||Lp||v||kLp||˜v||m1kLp||vw||LpdτCεmtθt(tτ)1m(m+1)Nαpτ11m+(m+1)Nαp(τ1mNαp||v˜v||Lp)dτCεm[supθtτtτ1mNαp||v(τ)˜v(τ)||Lp]. (4.8)

    Together Eq (4.7) with Eq (4.8) we have

    IkCεmθ0(1η)1m(m+1)Nαpη11m+(m+1)Nαp((tη)1mNαp||v(tη)˜v(tη)||Lp)dη+Cεm[supθtτtτ1mNαp||v(τ)˜v(τ)||Lp],k=1,2,,m1. (4.9)

    Similarly we have

    ImCεmθ0(1η)1m(m+1)Nαpη1+1m(m+1)Nαp((tη)1mNαp||((v˜v)(tη),(w˜w)(tη))||Lp)dη+Cεm[supθtτtτ1mNαp||((v˜v)(τ),(w˜w)(τ))||Lp]. (4.10)

    We next consider the term ||v˜v||Lp(RN):

    ||v˜v||Lp(RN)||Sα(t)(v0˜v0)||Lp(RN)+m1k=0Jk+Jm, (4.11)

    where

    (Jk,Jm)=||(Bk(v˜v,v,˜v,vw),B(˜v,(v˜v)(w˜w)))||Lp(RN).

    For the first term we have

    t1mNαp||Sα(t)(v0˜v0)||Lp(RN)21mNαpsupt>0(t2)1mNαp||Sα(t2)(v0˜v0)||Lp(RN)21mNαp||Sα(t)(v0˜v0)||˙Bαm+Npp,(RN). (4.12)

    For the term Jk and ϕ=(Δ)1(wv), we have

    Jk=||t0Sα(tτ)[(v˜v)vk˜vm1kϕ](τ)dτ||LpC(θt0+tθt)(tτ)mNαp||v˜v||Lp||v||kLp||˜v||m1kLp||ϕ||LNpNpdτC(θt0+tθt)(tτ)mNαp||v˜v||Lp||v||kLp||˜v||m1kLp||vw||LpdτCεm(θt0+tθt)(tτ)mNαpτ11m+(m+1)Nαp(τ1mNαp||v˜v||Lp)dτCεmt1m+Nαpθ0(1η)mNαpη11m+(m+1)Nαp((tη)1mNαp||v(tη)˜v(tη)||Lp)dη+Cεmt1m+Nαp[supθtτtτ1mNαp||v(τ)˜v(τ)||Lp],k=1,2,,m1. (4.13)

    Similarly, for the term Jm we have

    JmCεmt1m+Nαpθ0(1η)mNαpη1+1m(m+1)Nαp((tη)1mNαp||((v˜v)(tη),(w˜w)(tη))||Lp)dη+Cεmt1m+Nαp[supθtτtτ1mNαp||((v˜v)(τ),(w˜w)(τ))||Lp]. (4.14)

    Together Eq (4.5) with Eq (4.11) we have

    ||v˜v||˙Bαm+Npp,(RN)+t1mNαp||v˜v||Lp(RN)C||Sα(t)(v0˜v0)||˙Bαm+Npp,(RN)+Cεmθ0(1η)1m(m+1)Nαpη1+1m(m+1)Nαp((tη)1mNαp||((v˜v)(tη),(w˜w)(tη))||Lp)dη+Cεmθ0(1η)mNαpη1+1m(m+1)Nαp((tη)1mNαp||((v˜v)(tη),(w˜w)(tη))||Lp)dη+Cεm[supθtτtτ1mNαp||((v˜v)(τ),(w˜w)(τ))||Lp]. (4.15)

    For w˜w we can get the same estimate similar to Eq (4.15).

    For the convenience we denote

    Q(θ)=θ0(1η)1m(m+1)Nαpη11m+(m+1)Nαpdη+θ0(1η)mNαpη11m+(m+1)Nαpdη,F(t)=||Sα(t)(v0˜v0,w0˜w0)||˙Bαm+Npp,(RN),G(t)=||v˜v||˙Bαm+Npp,(RN)+t1mNαp||v˜v||Lp(RN).

    Due to the condition Eq (3.1), max{1,mNα}<p<min{N,m(m+1)Nα}, we have

    1+1m(m+1)Nαp=m+1mp(pmNα)>0,1m+(m+1)Nαp=1mp(m(m+1)Nαp)>0,1mNαp=1p(pmNα)>0,

    then, we obtain that Q(θ) converges and limθ0Q(θ)=0.

    Due to the condition Eq (4.1) we have limt+F(t)=0 and F(t)L[0,+). Passing the limit in Eq (4.15) we get

    M=lim supt+G(t)C(N,α,p)εm(Q(θ)+1)M, (4.16)

    Choosing θ and ε small enough such that Q(θ)<1 and 2C(N,α,p)εm<1 respectively, then Eq (4.16) implies that M=0. That is, Eq (4.2) holds true. The proof is complete.

    In this section we consider the regularizing decay rate estimates of the mild solutions to the system Eq (1.1). Compared to the case m=1, the main difficulty is caused by the power-law nonlinearity term vm as m>1 in the first two equations of Eq (1.1). To overcome this difficulty, we will apply multiple Leibniz's rule. For the regularizing-decay rate estimates of mild solutions to the Navier-Stokes equations, we refer the reader to [6,28,29,30].

    In what follows, for x=(x1,,xN)RN and β=(β1,,βN)NN0, where N0=N{0} and N={1,2,}, we denote βx=β1x1βNxN and |β|=β1++βN.

    We first describe the main result on regularizing-decay rate estimates of the mild solutions to the system Eq (1.1).

    Theorem 5.1. Let N2 be a positive integer, 1<α2N. Assume that p satisfies Eq (3.1) and (v0,w0)˙Bαm+Npp,(RN), and (v,w) is the mild solution to the system Eq (1.1) with initial data (v0,w0). Furthermore, assume that there exist two positive constants M1 and M2 such that

    sup0t<T(v(t),w(t))˙Bαm+Npp,(RN)M1, (5.1)
    sup0<t<Tt1mNαp(v(t),w(t))Lp(RN)M2. (5.2)

    Then, there exist two positive constants K1 and K2 depending only on M1, M2, N, α, m and p, such that

    (βxv(t),βxw(t))Lq(RN)K1(K2|β|)2|β|t|β|α1m+Nαq (5.3)

    for all pq, t(0,T) and βNN0.

    Remark 1. In fact, Eq (5.3) is equivalent to the claim

    (βxv(t),βxw(t))LqK1(K2|β|)2|β|δt|β|α1m+Nαq (5.4)

    for some δ(1,2] and sufficiently large constants K1 and K2.

    Let us first prepare the refined LpLq estimate for semigroup operator Sα(t).

    Lemma 5.2. Let 1pq. Then for any f˙Bαm+Npp,(RN), we have

    βxSα(t)fLq(RN)C|β|0|β||β|αt|β|α1m+Nαqf˙Bαm+Npp,(RN) (5.5)

    for all t>0,βNN0, and C0 is a constant depending only on N and α.

    Proof. As Sα(t) is the convolution operator with fractional heat kernel Kt(x)=F1(et|ξ|α), by scaling we see that

    Kt(x)=(2π)n2RNeixξet|ξ|αdξ=tNαK(xt1α),

    where K(x)=(2π)n2RNeixξe|ξ|αdξ. It is clear that [22] (Lemma 2.2)

    K(x)Lp(RN),Kt(x)Lp(RN),t(0,),p[1,],

    thus, the Young inequality implies that

    xSα(t)fLqxKt(x)L1fLqC0(N,α)t1αfLq. (5.6)

    By the semigroup property of Sα(t) and the commutativity between semigroup and differential operators, we get

    βxSα(t)f=Ni=1(xiSα(t2|β|))βiSα(t2)f. (5.7)

    Combining Eq (5.6) and Eq (5.7), and using Definition 2.1, we obtain

    βxSα(t)fLq(RN)Ni=1xiSα(t2|β|)βiL(Lq,Lq)Sα(t2)fLq(C0(N,α)(t2|β|)1α)|β|(t4)Nα(1p1q)Sα(t4)fLpC0(N,α)|β||β||β|αt|β|α1m+Nαqsupt>0(t4)1mNαpSα(t4)fLpC0(N,α)|β||β||β|αt|β|α1m+Nαqf˙Bαm+Npp,(RN),

    where TL(Lp,Lq) denotes the norm of linear operator T from Lp to Lq. This proves the Lemma 5.2.

    Next we recall some useful results.

    Lemma 5.3. [31,Lemma 2.1] Let δ>12. Then there exists a positive constant C depending only on δ, such that

    α<β(βα)|α||α|δ|βα||βα|δC(δ)|β||β|δ,βNN0. (5.8)

    Here the notation α<β means that αi<βi,iN, (βα)=Ni=1βi!αi!(βiαi)!, and the dependence of C(δ) on δ is merely of the form j=1jδ12.

    Lemma 5.4. [28] Let ψ0 be a measurable and locally bounded function in (0,) and {ψj}j=1 be a sequence of measurable functions in (0,). Assume that αR and μ,ν>0 satisfying μ+ν=1. Let Bη>0 be a number depending on η(0,1) and Bη be non-increasing with respect to η. Assume that there is a positive constant σ such that

    0ψ0(t)Bηtα+σt(1η)t(tτ)μτνψ0(τ)dτ, (5.9)
    0ψj+1(t)Bηtα+σt(1η)t(tτ)μτνψj(τ)dτ (5.10)

    for all j0, t>0 and η(0,1). Let η0 be a unique positive number such that

    I(η0)=min{12σ,I(1)}withI(η)=11η(1τ)μτανdτ.

    Then, for any 0<ηη0, we have

    ψj(t)2Bηtα,j0,t>0.

    We now prove the Theorem 5.1. Following the idea in Giga-Sawada [28], we first prove the Remark 1, a variant of Theorem 5.1 under extra regularity assumption.

    Proposition 1. Under the same assumptions in Theorem 5.1. Assume further that

    (βxv(t),βxw(t))C((0,T),Lq(RN)) (5.11)

    for all pq and βNN0. Then for any δ(1,2], there exist two positive constants K1 and K2 depending only on M1, M2, N, α, m and p, such that

    (βxv(t),βxw(t))LqK1(K2|β|)2|β|δt|β|α1m+Nαq (5.12)

    for all pq, t(0,T) and βNN0.

    Proof. We split the proof into the following two steps by an induction |β|=m.

    Step 1. We will prove Eq (5.12) for m=0. Equation (5.2) implies that Eq (5.12) is trivial if q=p, thus it suffices to consider q(p,]. Let η(0,1) be a constant to be determined later, we take Lqnorm of the first equation in Eq (1.11) and split the time integral into two parts as follows:

    v(t)LqSα(t)v0Lq+(t(1η)0+tt(1η))Sα(tτ)[vm()1(wv)(τ)]Lqdτ:=E1+E2+E3. (5.13)

    We will estimate term by term.

    For E1, by Lemma 5.2 and Eq (5.1), one can easily see that

    E1C1(N,α)tαm+Nαqv0˙B1m+Npp,C1(N,α,M1)t1m+Nαq. (5.14)

    For E2 and E3, by Lemma 2.2, Lemma 2.3 and Eq (5.2), we have

    E2=t(1η)0Sα(tτ)[vm()1(wv)(τ)]LqdτC2(N,α,p)t(1η)0(tτ)(m+1)Nαp+Nαqv(τ)mLp(v(τ),w(τ))LpdτC2(N,α,p)Mm+12t(1η)0(tτ)(m+1)Nαp+Nαqτ11m+(m+1)NαpdτC2(N,α,p,M2)η11mt1m+Nαq, (5.15)
    E3=tt(1η)Sα(tτ)[vm()1(wv)(τ)]LqdτC3(N,α,p)tt(1η)(tτ)mNαpv(τ)Lqv(τ)m1Lp(v(τ),w(τ))LpdτC3(N,α,p,M2)tt(1η)(tτ)mNαpτ1+mNαpv(τ)Lqdτ. (5.16)

    Combining Eqs (5.14)–(5.16), and setting ˉBη=C1(N,α,M1)+C2(N,α,p,M2)η11m, the inequality Eq (5.13) yields that

    v(t)LqˉBηt1m+Nαq+C3tt(1η)(tτ)mNαpτ1+mNαpv(τ)Lqdτ. (5.17)

    The estimate for w(t) can be done analogously as Eq (5.17). Hence, we have

    (v(t),w(t))LqBηt1m+Nαq+C4tt(1η)(tτ)mNαpτ1+mNαp(v(τ),w(τ))Lqdτ, (5.18)

    where Bη=2ˉBη and C4=2C3(N,α,p,M2).

    By applying Lemma 5.4, we get the desired estimate Eq (5.12) for |β|=k=0 with K1=2Bη0 for some η0=η0(N,α,p,m,M1,M2)(0,1).

    Step 2. Next we prove Eq (5.12) for |β|=k1. Due to the appearance of nonlocal function ϕ, we use a different argument to prove Eq (5.12) for pq<N and Nq, thus we split the proof into the following two cases.

    Case 1: pq<N. In this case, we first differentiate the first equation of Eq (1.11) to obtain the identity

    βxv(t)=βxSα(t)v0t0βxSα(tτ)[vm(Δ)1(wv)(τ)]dτ. (5.19)

    We take the Lqnorm of βxv, for some η(0,1) to be chosen later, we split the time integral into the following two parts:

    βxv(t)LqβxSα(t)v0Lq+(t(1η)0+tt(1η))βxSα(tτ)[vm(Δ)1(wv)(τ)]Lqdτ:=F1+F2+F3. (5.20)

    We next estimate Fi(i=1,2,3) term by term.

    For F1, Lemma 5.2 implies that

    F1Ck0kkαtkα1m+nαqv0˙Bαm+npp,M1Ck0kkαtkα1m+nαq. (5.21)

    For F2, using Lemma 5.2, Lemma 2.3 and Eq (5.2), we have

    F2=t(1η)0βxSα(tτ)[vm(Δ)1(wv)(τ)]LqdτC5(N,α)t(1η)0(tτ2)1αβxSα(tτ2)[vm(Δ)1(wv)(τ)]LqdτC5(N,α)t(1η)0(tτ2)1αNi=1xiSα(tτ4k)kiL(Lq,Lq)×Sα(tτ4)[vm(Δ)1(wv)(τ)]LqdτC5(N,α)t(1η)0(tτ2)1α[C0(tτ4k)1α]k(tτ4)(m+1)Npαp+Nαq×vm(Δ)1(wv)(τ)LNp(m+1)NpdτC5(N,α,p)Ck0kkαt(1η)0(tτ4)kαNα(m+1p1q)v(τ)mLp(v(τ),w(τ))LpdτC5(N,α,p)Mm+12Ck0kkαt(1η)0(tτ4)kαNα(m+1p1q)τ11m+(m+1)NαpdτC5(N,α,p,M2)Ck0kkαηkα11mtkα1m+Nαq, (5.22)

    where k=k1+k2++kN and ki=|βi|(i=1,2,,N).

    Using Leibniz's rule, we split F3 into the following three parts:

    F3=tt(1η)βxSα(tτ)[vm(Δ)1(wv)(τ)]LqdτC6(N,α)tt(1η)(tτ2)1αSα(tτ2)βx[vm(Δ)1(wv)(τ)]LqdτC6(N,α)tt(1η)(tτ2)1αSα(tτ2)[(βxvm)(Δ)1(wv)(τ)]Lqdτ+C6(N,α)tt(1η)(tτ2)1αSα(tτ2)0<γ<β(βγ)(γxvm)(βγx(Δ)1(wv)(τ))Lqdτ+C6(N,α)tt(1η)(tτ2)1αSα(tτ2)[vmβx(Δ)1(wv)(τ)]Lqdτ:=F31+F32+F33. (5.23)

    Here, the notation γ<β means that γβ and |γ|<|β|.

    Now, we establish the estimates for F3j(j=1,2,3). For F31, using Leibniz's rule again, we can split F31 into two parts as follows:

    F31=C7(N,α)tt(1η)(tτ2)1αSα(tτ2)[(βxvm)(Δ)1(wv)]Lqdτ=C7(N,α)tt(1η)(tτ2)1αSα(tτ2)[β(βmβm1)(βm1βm2)(β2β1)×(β1xv)(β2β1xv)(βmβm1xv)+mvm1(βxv)(Δ)1(wv)]Lqdτ=C7(N,α)βmi=1(βiβi1)tt(1η)(tτ2)1αSα(tτ2)mi=1(βiβi1xv)(Δ)1(wv)(τ)Lqdτ+C7(N,α,m)tt(1η)(tτ2)1αSα(tτ2)vm1(βxv)(Δ)1(wv)Lqdτ:=G1+G2, (5.24)

    where we denote β=0=β0β1βm1<βm=β.

    For G2, using Lemma 2.2, Lemma 2.3 and Eq (5.2), we have

    G2C8(N,α,m,p)tt(1η)(tτ2)mNαpvm1LpβxvLq(v(τ),w(τ))LpdτC8(N,α,m,p)Mm2tt(1η)(tτ)mNαpτ1+mNαpβxvLqdτ. (5.25)

    For G1, using Lemma 2.2, Lemma 2.3, Lemma 5.3, Eq (5.2) and Eq (5.12), we have

    G1C9(N,α,p)βmi=1(βiβi1)tt(1η)(tτ)(m1)NαqNαp×mi=1βiβi1xvLq(v(τ),w(τ))LpdτC9(N,α,p)M2βmi=1(βiβi1)tt(1η)(tτ)(m1)NαqNαp×mi=1[K1(K2|βiβi1|2|βiβi1|δ)τ|βiβi1|α1m+Nαq]τ1m+NαpdτC9(N,α,p,M2)βmi=1(βiβi1)mi=1[K1(K2|βiβi1|2|βiβi1|δ)]×tt(1η)(tτ)(m1)NαqNαpτkα1+mNαq1m+NαpdτC9(N,α,p,M2)(C(δ))2(m1)k2kδKm1K2kmδ2I(η)tkα1m+Nαq, (5.26)

    where

    I(η)=11η(1τ)(m1)NαqNαpτkα1+mNαq1m+Nαpdτ. (5.27)

    For F32, using the same arguments as G1, we have

    F32C10(N,α)tt(1η)(tτ2)1αSα(tτ2)[0<γ<β(βγ)(γxvm)×(βγx(Δ)1(wv)(τ))]LqdτC10(N,α)0<γ<β(βγ)tt(1η)(tτ2)1αSα(tτ2)(γxvm)×(βγx(Δ)1(wv)(τ))Lqdτ=C10(N,α)0<γ<β(βγ)tt(1η)(tτ2)1αSα(tτ2)[γmi=1(γiγi1)×mj=1(γjγj1xvm)](βγx(Δ)1(wv)(τ))LqdτC10(N,α)0<γ<β(βγ)γmi=1(γiγi1)tt(1η)(tτ2)1α×Sα(tτ2)mj=1(γjγj1xvm)(βγx(Δ)1(wv)(τ))Lqdτ,

    according to the property of semigroup we get

    F32C10(N,α,p)0<γ<β(βγ)γmi=1(γiγi1)tt(1η)(tτ)N(m1)αqNαp×mj=1γjγj1xvLqβγx(v(τ),w(τ))LpdτC10(N,α,p)0<γ<β(βγ)γmi=1(γiγi1)tt(1η)(tτ)N(m1)αqNαp×mj=1[K1(K2|γjγj1|)2|γjγj1|δτ|γjγj1|α1m+Nαq]×[K1(K2|βγ|)2|βγ|δτ|βγ|α1m+Nαp]dτC10(N,α,p)(C(δ))mKm+11K2k(m+1)δ2k2kδI(η)tkα1m+Nαq, (5.28)

    where γ is defined the same as that in estimating F31 and I(η) is defined in Eq (5.27).

    For F33, analogously we have

    F33C11tt(1η)(tτ)N(m1)αqNαpvmqβx()1(wv)(τ)LNpNpdτC11(N,α)tt(1η)(tτ)N(m1)αqNαpvmqβ1x(v(τ),w(τ))LNpNpdτC11(N,α)tt(1η)(tτ)N(m1)αqNαp[K1τ1m+Nαq]m×[K1(K2(k1))2(k1)δτk1α1m+N(Np)αNp]dτC11(N,α)Km+11K2(k1)δ2k2kδI(η)tkα1m+Nαq, (5.29)

    where I(η) is defined in Eq (5.27).

    Combining the above estimates Eqs (5.20)(5.29) and setting ˉBη by

    ˉBη=M1Ck0kkα+C5Ck0kkαηkα11m+C12k2kδI(η),

    and

    C12=C9Km1K2kmδ2+C10Km+11K2k(m+1)δ2+C11Km+11K2(k1)δ2, (5.30)

    we obtain

    βxv(t)LqˉBηtkα1m+Nαq+C8tt(1η)(tτ)mNαpτ1+mNαpβxv(τ)Lqdτ. (5.31)

    Similarly, we can deal with βxw(t). Hence, we conclude that

    (βxv(t),βxw(t))LqBηtkα1m+Nαq+C13tt(1η)(tτ)mNαpτ1+mNαp(βxv(τ),βxw(τ))Lqdτ, (5.32)

    where Bη=2ˉBη and C13=2C8(N,α,m,p).

    Let ηk=12k. It is clear that I(ηk) is strictly monotone decreasing in k and I(ηk)0 as k. Choosing k0 sufficiently large, such that I(12k)12C13 for all kk0, applying Lemma 5.4, we get

    (βxv(t),βxw(t))Lq2B12ktkα1m+Nαq (5.33)

    for all t>0 and |β|=k. Note that from Eq (5.33), we can choose K1 and K2 sufficiently large such that Eq (5.12) holds for all β satisfying |β|k0. Hence, it suffices to prove that it is possible to choose K1 and K2 such that 2B12kK1(K2k)2kδ for all k>k0. Since

    I(12k)=1112k(1τ)(m1)NαqNαpτkα1+mNαq1m+Nαpdτ(112k)kα11me12α(112k)11m16,

    we can calculate 2B12k as follows:

    2B12k=4ˉB12k4[M1Ck0kkα+C5Ck0kkα(2k)kα+1+1m+16C12k2kδ]4[M1Ck0+2kα+1+1mC5Ck0k1+1m+δ+16C12]k2kδ.

    Obviously, there exists a constant C14>C0 such that Ck0+2kα+1+1mCk0k1+1m+δC2kδ14. Hence,

    2B12k4[(M1+C5)C2kδ14+16C12]k2kδ, (5.34)

    where C12 is defined in Eq (5.30).

    Choosing K1:=8(M1+C5) and K2:=max{C14,32(C9+C10)K1,32C11Km21}, we obtain Eq (5.12). This completes the proof of Proposition 1 for pq<N.

    Case 2: Nq. Now we are in a position to establish the estimate of βxv(t)Lq for Nq. For p satisfying Eq (3.1), using the Gagliardo-Nirenberg inequality [32], we have

    βxv(t)LqC(N,p)βxv(t)θLp2xβxv(t)1θLp,θ=1N2p+N2q. (5.35)

    Now, from Eq (5.35) and the result of Case 1 we see that

    βxv(t)LqC(N,p)[K1(K2k)2kδtkα1m+Nαp]θ[K1(K2(k+2))2(k+2)δtk+2α1m+Nαp]1θC(N,p)K1(K2(k+2))2k+4δtkα1m+Nαq. (5.36)

    It is clear that there exists a constant C152 such that k4C2kδ15, thus we have

    (K2(k+2))2k+4δ=K42k4(1+2k)2k+4δ(K2k)2kδ81e4K42(C15K2k)2kδ.

    Hence, we can choose K1 and K2 sufficiently large such that Eq (5.12) holds for all pq. This completes the proof of Proposition 1.

    Finally, let us show that under the assumptions of Theorem 5.1, the mild solution (v(t),w(t)) of Eq (1.1) always satisfies the regularity condition Eq (5.12).

    Proposition 2. Under the assumptions of Theorem 5.1, the mild solution (v(t), w(t)) satisfies that

    t|β|α+1mNαq(βxv(t),βxw(t))Lq˜K1(˜K2|β|)2|β|δ (5.37)

    for all pq, t(0,T) and βNN0, where ˜K1 and ˜K2 are constants depending only on M1,M2,m,N,α,p and δ.

    Proof. Since the mild solution (v(t),w(t)) is the limit function of the sequence (vj(t),wj(t)) of appropriate Picard iterations as follows:

    (v1(t),w1(t))=(Sα(t)v0,Sα(t)w0),forj2,vj(t)=Sα(t)v0+t0Sα(tτ)[vmj1(Δ)1(vj1wj1)](τ)dτ,wj(t)=Sα(t)w0+t0Sα(tτ)[wmj1(Δ)1(wj1vj1)](τ)dτ.

    Step 1. We first show that

    supj1sup0<t<Tt1mNαp(vj(t),wj(t))LpM2. (5.38)

    When j=1, following from Eq (5.1) we have

    (v1,w1)Lp=(Sα(t)v0,Sα(t)w0)Lpt1m+Nαpsup0<t<Tt1mNαp(Sα(t)v0,Sα(t)w0)Lpt1m+Nαp(v0,w0)˙Bαm+Npp,M1t1m+Nαp. (5.39)

    Hence Eq (5.38) holds for j=1.

    When j2, using Lemma 2.2 and Lemma 2.3, we have

    vj(t)LpSα(t)v0Lp+t0Sα(tτ)[vmj1(Δ)1(vj1wj1)]Lp(τ)dτM1t1m+Nαp+C(N,α,p)t0(tτ)mNαpvj1(τ)mLp(vj1(τ),wj1(τ))LpdτM1t1m+Nαp+C(N,α,p)[sup0<s<Ts1mNαp(vj1(s),wj1(s))Lp]m+1t1m+NαpB,

    where B=10(1τ)mNαpτ11m+(m+1)Nαpdτ=B(1mNαp,1m+(m+1)Nαp) is the standard Beta function which is obviously finite.

    For wj(t) we have the analogous estimate. Then, for j=2,3,, we get

    (vj(t),wj(t))LpC(N,α,p,m,M1,B)t1m+Nαp:=M2t1m+Nαp, (5.40)

    where the constant C(N,α,p,m,M1,B) is always finite. Therefore Eq (5.38) holds true.

    Step 2. To apply the Lemma 5.4, we need to show that (βxv1(t),βxw1(t))Lq is locally bounded in (0,T). Using Lemma 2.3 and Eq (5.1), we have

    βxv1(t)Lq=βxSα(t2)Sα(t2)v0LqC(N,α)(t2)|β|αNα(1p1q)Sα(t2)v0LpC(N,α)(t2)|β|αNα(1p1q)(t2)1m+Nαpsupt>0(t2)1mNαpSα(t2)v0LpC(N,α)M1(t2)|β|α1m+Nαq.

    Similarly, we have a similar estimate on wj(t). Then (βxv1(t),βxw1(t))Lq is locally bounded in (0,T).

    Step 3. Similarly to the proof of Proposition 1, let ψj(t)=βxvj(t)Lq, for all j1 and t(0,T), we have

    ψj+1(t)ˉBηt|β|α1m+Nαq+C8tt(1η)(tτ)mNαpτ1+mNαpψj(τ)dτ. (5.41)

    Using Lemma 5.4 (the version of sequences), we can choose appropriate constants ˜K1 and ˜K2 such that

    ψj(t)˜K1(˜K2|β|)2|β|δt|β|α1m+Nαq. (5.42)

    For wj(t) we have the similar estimate. Hence we complete the proof of Proposition 2.

    The proof of Theorem 5.1. Now Theorem 5.1 follows immediately from Proposition 1 and Proposition 2. We complete the proof of Theorem 5.1.

    In this section, we consider a fractional drift diffusion system with generalized electric potential equation

    {tv+Λαv=(vmϕ),t>0,xRN,tw+Λαw=(wmϕ),t>0,xRN,ϕ=K(vw)(x)=cRNb(x,y)(vw)(y)dy,t>0,xRN,v(x,0)=v0(x),w(x,0)=w0(x),xRN, (6.1)

    where c is a constant and b(x,y) is the kernel function of nonlocal linear integral operator K.

    For K=(Δ)1 which comes from the Poisson equation Δϕ=vw, Eq (6.1) becomes the fractional drift diffusion system Eq (1.1). For instance,

    K(u)(x)=cRN(xy)u(y)|xy|Ndy, (6.2)

    where c is a constant. If c<0, the equation ut=Δu+(uK(u)) models the Brownian diffusion of charge carriers interacting via Coulomb forces. If c>0, the operator K reflects the mutual gravitational attraction of particles. Furthermore, Biler-Woyczynski [33] considered the equation ut=Λαu+(uK(u)).

    We also give the global existence and asymptotic stability of the mild solution to the Cauchy problem Eq (6.1).

    Theorem 6.1. Let N be a positive integer, 1<α2N and Eq (3.1) hold true. Assume that (v0,w0)˙Bαm+Npp,(RN). If the derivative of kernel function b(x,y) satisfies

    |Db(x,y)|C|xy|N+1, (6.3)

    then there exists ε>0 such that if ||(v0,w0)||˙Bαm+Npp,ε, the Cauchy problem Eq (6.1) has a unique global mild solution (v,w)X such that ||(v,w)||X2ε. Moreover, the solution depends continuously on initial data in the following sense: let (˜v,˜w)X be the solution of Eq (6.1) with initial data (˜v0,˜w0) such that ||(˜v0,˜w0)||˙Bαm+Npp,(RN)ε, then there is a constant C such that

    ||(v˜v,w˜w)||XC||(v0˜v0,w0˜w0)||˙Bαm+Npp,(RN). (6.4)

    Proof. After a few modifications of the proof to Theorem 3.1, we can prove this theorem. Here we just give the main difference in the proof.

    By the fractional heat semigroup Sα(t)=etΛα, we rewrite the system Eq (6.1) as a system of integral equations

    {v(t)=Sα(t)v0+B(v,,v,w),w(t)=Sα(t)w0+B(w,,w,v), (6.5)

    where

    B(v,,vm,w)=t0Sα(tτ)[vmK(vw)](τ)dτ. (6.6)

    Similar to Eqs (3.4)–(3.8), we have

    ||B(v,,v,w)(t)||˙Bαm+Npp,(RN)=sups>0s1mNαp||Sα(s)t0Sα(tτ)[vmK(vw)](τ)dτ||Lpt0sups>0s1mNαp||Sα(s)Sα(tτ)[vmK(vw)](τ)||LpdτC(N,α)t0(tτ)1m(m+1)Nαp||vmK(vw)(τ)||LNp(m+1)NpdτC(N,α)t0(tτ)1m(m+1)Nαp||v(τ)||mLp||K(vw)(τ)||LNpNpdτ, (6.7)

    due to the condition Eq (6.3): |Db(x,y)|C|xy|N+1, use Hardy-Littlewood-Sobolev inequality for the integral RN|xy|N+1|vw|dy, we have

    ||K(vw)||LNpNpC(N,p)||vw||Lp. (6.8)

    then

    ||B(v,,v,w)(t)||˙Bαm+Npp,(RN)C(N,α,p)supτ>0(τ1mNαp||v(τ)||Lp)msupτ>0(τ1mNαp||(vw)(τ)||Lp)t0(tτ)1m(m+1)Nαpτ(m+1)Nαp1m1dτC(N,α,p)||v||mX||vw||Xt0(tτ)1m(m+1)Nαpτ(m+1)Nαp1m1dτC(N,α,p)||v||mX||vw||X, (6.9)

    therefore, we have

    ||B(v,,v,w)(t)||˙Bαm+Npp,(RN)C(N,α,p)||v||mX||vw||X. (6.10)

    Similarly, we have

    supt>0t1mNαp||B(v,,v,w)(t)||LpC(N,α,p)||v||mX||vw||X. (6.11)

    Following the main estimates Eq (6.10) and Eq (6.11) and the proof of Theorem 3.1, the Cauchy problem Eq (6.1) has a unique global-in-time mild solution in the mixed time-space Besov space. This completes the proof of Theorem 6.1.

    Using the same method we can prove that the mild solution of the Cauchy problem Eq (6.1) has the following asymptotic stability.

    Theorem 6.2. Let N2 be a positive integer, 1<α2N, Eq (3.1) and Eq (6.3) hold true. Assume that (v,w) and (˜v,˜w) are two mild solutions of the Cauchy problem Eq (6.1) described in Theorem 6.1 corresponding to initial conditions (v0,w0) and (˜v0,˜w0), respectively. If (v0,w0),(˜v0,˜w0)˙Bαm+Npp,(RN) such that

    limt||Sα(t)(v0˜v0,w0˜w0)||˙Bαm+Npp,(RN)=0, (6.12)

    then, we have the following asymptotic stability

    limt(||(v˜v,w˜w)||˙Bαm+Npp,(RN)+tαmNp||(v˜v,w˜w)||Lp(RN))=0. (6.13)

    Theorem 6.3. Let N2 be a positive integer, 1<α2N, Eq (3.1) and Eq (6.3) hold true. Assume that (v0,w0)˙Bαm+Npp,(RN), and (v,w) is the mild solution to the system Eq (6.1) with initial data (v0,w0). Furthermore, assume that there exist two positive constants M1 and M2 such that

    sup0t<T(v(t),w(t))˙Bαm+Npp,(RN)M1, (6.14)
    sup0<t<Tt1mNαp(v(t),w(t))Lp(RN)M2. (6.15)

    Then, there exist two positive constants K1 and K2 depending only on M1, M2, N, α, m and p, such that

    (βxv(t),βxw(t))Lq(RN)K1(K2|β|)2|β|t|β|α1m+Nαq (6.16)

    for all pq, t(0,T) and βNN0.

    The authors are grateful to the anonymous referees for helpful comments and suggestions that greatly improved the presentation of this paper. The research of C. Gu is partially supported by the CSC under grant No. 202006160118. The research of C. Gu and Y. Tang is supported by the NNSF of China (Nos. 12171442 and 11971188).

    The authors have no conflicts in this paper.

    [1] Existence and nonexistence of TV bounds for scalar conservation laws with discontinuous flux. Comm. Pure Appl. Math. (2011) 64: 84-115.
    [2] Riemann problems with non-local point constraints and capacity drop. Math. Biosci. Eng. (2015) 12: 259-278.
    [3] A second-order model for vehicular traffics with local point constraints on the flow. Math. Models Methods Appl. Sci. (2016) 26: 751-802.
    [4] Finite volume schemes for locally constrained conservation laws. Numer. Math. (2010) 115: 609-645, With supplementary material available online.
    [5] A theory of \begin{document}$ L^1$\end{document}-dissipative solvers for scalar conservation laws with discontinuous flux. Arch. Ration. Mech. Anal. (2011) 201: 27-86.
    [6] Well-posedness for a one-dimensional fluid-particle interaction model. SIAM J. Math. Anal. (2014) 46: 1030-1052.
    [7] Kružkov's estimates for scalar conservation laws revisited. Trans. Amer. Math. Soc. (1998) 350: 2847-2870.
    [8] A. Bressan, Hyperbolic Systems of Conservation Laws, vol. 20 of Oxford Lecture Series in Mathematics and its Applications, Oxford University Press, Oxford, 2000, The one-dimensional Cauchy problem.
    [9] Uniqueness for discontinuous ODE and conservation laws. Nonlinear Anal. (1998) 34: 637-652.
    [10] A well posed conservation law with a variable unilateral constraint. J. Differential Equations (2007) 234: 654-675.
    [11] A Hölder continuous ODE related to traffic flow. Proc. Roy. Soc. Edinburgh Sect. A (2003) 133: 759-772.
    [12] Scalar conservation laws with moving constraints arising in traffic flow modeling: an existence result. J. Differential Equations (2014) 257: 4015-4029.
    [13] The Aw-Rascle traffic model with locally constrained flow. J. Math. Anal. Appl. (2011) 378: 634-648.
    [14] M. Garavello and S. Villa, The Cauchy problem for the Aw-Rascle-Zhang traffic model with locally constrained flow, Preprint, 2016.
    [15] First order quasilinear equations with several independent variables. Mat. Sb. (N.S.) (1970) 81: 228-255.
    [16] S. Villa, P. Goatin and C. Chalons, Moving bottlenecks for the Aw-Rascle-Zhang traffic flow model, 2016, URL https://hal.archives-ouvertes.fr/hal-01347925, preprint.
  • This article has been cited by:

    1. Yongqiang Zhao, Yanbin Tang, Approximation of solutions to integro-differential time fractional wave equations in Lpspace, 2023, 18, 1556-1801, 1024, 10.3934/nhm.2023045
    2. Junlong Chen, Yanbin Tang, Homogenization of nonlinear nonlocal diffusion equation with periodic and stationary structure, 2023, 18, 1556-1801, 1118, 10.3934/nhm.2023049
    3. Yongqiang Zhao, Yanbin Tang, Critical behavior of a semilinear time fractional diffusion equation with forcing term depending on time and space, 2024, 178, 09600779, 114309, 10.1016/j.chaos.2023.114309
    4. Caihong Gu, Yanbin Tang, Well-posedness of Cauchy problem of fractional drift diffusion system in non-critical spaces with power-law nonlinearity, 2024, 13, 2191-950X, 10.1515/anona-2024-0023
    5. Yongqiang Zhao, Yanbin Tang, Approximation of solutions to integro-differential time fractional order parabolic equations in Lp-spaces, 2023, 2023, 1029-242X, 10.1186/s13660-023-03057-2
  • Reader Comments
  • © 2017 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(5617) PDF downloads(190) Cited by(12)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog