This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Radial boundary layers for the singular Keller-Segel model

Qianqian Hou Institute for Advanced Study in Mathematics, Harbin Institute of Technology, Harbin 150001, People’s Republic of China (qianqian.hou@connect.polyu.hk).

Abstract This paper is concerned with the diffusion limit (as ε0\varepsilon\rightarrow 0) of radial solutions to a chemotaxis system with logarithmic singular sensitivity in a bounded interval with mixed Dirichlet and Robin boundary conditions. We use a Cole-Hopf type transformation to resolve the logarithmic singularity and prove that the solution of the transformed system has a boundary-layer profile as ε0\varepsilon\to 0, where the boundary layer thickness is of 𝒪(εα)\mathcal{O}(\varepsilon^{\alpha}) with 0<α<120<\alpha<\frac{1}{2}. By transferring the results back to the original chemotaxis model via Cole-Hopf transformation, we find that boundary layer profile is present at the gradient of solutions and the solution itself is uniformly convergent with respect to ε>0\varepsilon>0.

MSC: 35A01, 35B40, 35K57, 35Q92, 92C17

Keywords: Chemotaxis, Boundary layers, Logarithmic singularity, Perturbation Method

1 Introduction

Chemotaxis describes the oriented movement of species stimulated by uneven distribution of a chemical substance in the environment. It is a significant mechanism accounting for abundant biological process/phenomenon, such as aggregation of bacteria [34, 44], slime mould formation [15], fish pigmentation [35], tumor angiogenesis [4, 7, 6], primitive streak formation [36], blood vessel formation [12], wound healing [39]. Mathematical models of chemotaxis were first proposed by Keller and Segel in their seminal works [19, 20, 21]. In this paper, we are concerned with the following chemotaxis model:

{ut=[Duχu(lnc)],(x,t)Ω×(0,)ct=εΔcμuc,\displaystyle\left\{\begin{array}[]{lll}u_{t}=\nabla\cdot[D\nabla u-\chi u\nabla(\ln c)],\quad(x,t)\in\Omega\times(0,\infty)\\ c_{t}=\varepsilon\Delta c-\mu uc,\end{array}\right. (1.3)

where Ω\Omega is a domain in n\mathbb{R}^{n} with smooth boundary. System (1.3) was first advocated in [21] to describe the traveling band propagation of bacterial chemotaxis observed in the experiment of Adler [1, 2]. It later appeared in the work by Levine et al [23] to model the initiation of tumor angiogenesis, where u(x,t)u(x,t) represents the density of vascular endothelial cells and c(x,t)c(x,t) denotes the concentration of signaling molecules vascular endothelial growth factor (VEGF). The parameters D>0D>0, ε0\varepsilon\geq 0 are diffusion coefficients of the endothelial cells and the chemical VEGF respectively, χ>0\chi>0 is the chemotactic coefficient measuring the intensity of chemotaxis and μ0\mu\geq 0 is the chemical consumption rate by cells. In particular it was pointed out in [23] that the chemical diffusion process is far less important comparing to its interaction with endothelial cells and thus the diffusion coefficient ε\varepsilon could be small or negligible. Despite of its biological significance, (1.3) is difficult to study mathematically due to the singularity of lnc\ln c at c=0c=0. The well-known way to overcome this singularity was applying the following Cole-Hopf transformation (cf. [22, 31]):

v=lnc=cc\vec{v}=-\nabla\ln c=-\frac{\nabla c}{c} (1.4)

to transform (1.3) into a system of conservation laws:

{utχ(uv)=DΔu,vt(uε|v|2)=εΔv,(u,v)(x,0)=(u0,v0)(x).\left\{\begin{array}[]{ll}u_{t}-\chi\nabla\cdot(u\vec{v})=D\Delta u,\\[2.84526pt] \vec{v}_{t}-\nabla(u-\varepsilon|\vec{v}|^{2})=\varepsilon\Delta\vec{v},\\ (u,\vec{v})(x,0)=(u_{0},\vec{v}_{0})(x).\end{array}\right. (1.5)

The transformed system (1.5) attracts extensive attentions and numerous interesting results have been developed. We briefly recall these results by the dimension of spaces. In the one dimensional case, the global well-posedness along with large time behavior of solutions was investigated when Ω=\Omega=\mathbb{R} in [13, 25] with ε=0\varepsilon=0 and in [37, 33] with ε>0\varepsilon>0. When Ω=(0,1)\Omega=(0,1), authors in [54, 28] obtained the unique global solution under Neumann-Dirichlet boundary conditions for ε=0\varepsilon=0, and the result was later extended to the case ε>0\varepsilon>0 in [49, 43]. The problem with ε0\varepsilon\geq 0 is also globally well-posed [26] with Dirichlet-Dirichlet boundary conditions. Furthermore, the existence and stability of traveling wave solutions were studied in [18, 29, 30, 31, 32, 27, 3]. However to the best of our knowledge, except when it is associated with radially symmetric initial data, the known well-posedness results of problem (1.5) in the multi-dimension are merely confined to local large and global small solutions, cf. [24, 14, 8, 38, 45, 48] for details when Ω=n\Omega=\mathbb{R}^{n} (n2n\geq 2) and [28, 41] when Ωn\Omega\subset\mathbb{R}^{n} (n2n\geq 2) is bounded. If the initial data are radially symmetric and Ω=BR(0):={x=(x1,x2,,xn)n||x|<R}\Omega=B_{R}(0):=\{x=(x_{1},x_{2},\cdots,x_{n})\in\mathbb{R}^{n}\,|\,\,|x|<R\} (n2n\geq 2), Winkler [50] recently proved that (1.3) with ε>0\varepsilon>0 subject to Neumann boundary conditions admits a global generalized (weak) solution which is radially symmetric and smooth away from the origin x=0x=0.

In addition to the above well-posedness results, the asymptotic behavior of solutions as ε0\varepsilon\rightarrow 0 is a particularly relevant issue (mentioned in [23] that ε\varepsilon could be small/negligible) and has been studied in several circumstances for equation (1.5) and thus for the original equation (1.3) via transformation (1.4). For illustration, denote by (uε,vε)(u^{\varepsilon},\vec{v}^{\,\varepsilon}) and (u0,v 0)(u^{0},\vec{v}^{\,0}) the solutions of (1.5) with ε>0\varepsilon>0 and ε=0\varepsilon=0 respectively (when n=1n=1 we shall use the notation vv instead of v\vec{v} since it is a scalar). First in unbounded domains, it has been shown that both traveling wave solutions (cf. [47]) in {\mathbb{R}} and the global small-data solution of the Cauchy problem (cf. [48, 38]) in n(n=2,3){\mathbb{R}}^{n}(n=2,3) are uniformly convergent in ε\varepsilon, namely (uε,vε)(u^{\varepsilon},\vec{v}^{\,\varepsilon}) converge to (u0,v 0)(u^{0},\vec{v}^{\,0}) in LL^{\infty}-norm as ε0\varepsilon\to 0. With Ω=(0,1)\Omega=(0,1), the solutions still converge (cf. [49]) as ε0\varepsilon\rightarrow 0 when (1.5) is endowed with the following mixed homogeneous Neumann-Dirichlet boundary conditions

ux|x=0,1=v|x=0,1=0.u_{x}|_{x=0,1}=v|_{x=0,1}=0.

However if both of uu and vv are subject to the Dirichlet boundary conditions, one can not preassign the boundary value for v0v^{0} when ε=0\varepsilon=0 since it is intrinsically determined by the second equation of (1.5) as v0|x=0,1=v0|x=0,1+0tux0|x=0,1dτv^{0}|_{x=0,1}=v_{0}|_{x=0,1}+\int_{0}^{t}u_{x}^{0}|_{x=0,1}\,d\tau. Thus the plausible Dirichlet boundary conditions should be prescribed as:

{u|x=0,1=u¯0,v|x=0,1=v¯,ifε>0,u|x=0,1=u¯0,ifε=0,\left\{\begin{aligned} &u|_{x=0,1}=\bar{u}\geq 0,\quad v|_{x=0,1}=\bar{v},\ &\mathrm{if}\ \varepsilon>0,\\ &u|_{x=0,1}=\bar{u}\geq 0,\ &\ \ \ \ \mathrm{if}\ \varepsilon=0,\end{aligned}\right. (1.6)

where u¯0,\bar{u}\geq 0, v¯\bar{v}\in\mathbb{R} are constants. In this case, if the boundary values of vv with ε>0\varepsilon>0 and ε=0\varepsilon=0 do not match, then the solution component vv would diverge near the end points x=0,1x=0,1 as ε0\varepsilon\rightarrow 0 and this phenomenon is termed as the boundary layer effect, which has been an important topic in the fluid mechanics [42] when investigating the inviscid limit of the Navier-Stokes equations near a boundary and has attracted extensive studies (cf. [9, 10, 11, 17, 46, 52, 53]) since the pioneering work [40] by Prandtl in 1904. In particular, this boundary layer effect for problem (1.5)-(1.6) has been recently numerically verified in [26] and rigorously proved in [16].

Enlightened by these results, it is natural to expect that (1.5) in multi-dimension (n2n\geq 2) possesses boundary layer solutions as well when prescribing appropriate Dirichlet boundary conditions. In particular we aim to investigate this issue for its radial solutions in the present paper. To this end, we first rewrite (1.3) in its radially symmetric form by assuming that the solutions (u,c)(u,c) are radially symmetric, depending only on the radial variable r=|x|r=|x| and time variable tt. In a domain bounded by two concentric sphere, i.e. Ω={x=(x1,x2,,xn)n|   0<a<|x|<b}\Omega=\{x=(x_{1},x_{2},\cdots,x_{n})\in\mathbb{R}^{n}\,|\,\,\,0<a<|x|<b\}, (1.3) reads as

{ut=1rn1(rn1ur)r1rn1(rn1ucrc)r,(r,t)(a,b)×(0,)ct=ε1rn1(rn1cr)ruc,u(r,0)=u0(r),c(r,0)=c0(r),\displaystyle\left\{\begin{array}[]{lll}u_{t}=\frac{1}{r^{n-1}}(r^{n-1}u_{r})_{r}-\frac{1}{r^{n-1}}\left(r^{n-1}u\frac{c_{r}}{c}\right)_{r},\quad(r,t)\in(a,b)\times(0,\infty)\\ c_{t}=\varepsilon\frac{1}{r^{n-1}}(r^{n-1}c_{r})_{r}-uc,\\ u(r,0)=u_{0}(r),\quad c(r,0)=c_{0}(r),\end{array}\right. (1.10)

where D=χ=μ=1D=\chi=\mu=1 have been assumed without loss of generality. Similar as deriving (1.5) from (1.3), we apply the following Cole-Hopf type transformation

v=(lnc)r=crc,v=-(\ln c)_{r}=-\frac{c_{r}}{c}, (1.11)

which turns (1.10) into

{ut=1rn1(rn1ur)r+1rn1(rn1uv)r,(r,t)(a,b)×(0,)vt=ε1rn1(rn1vr)rεn1r2vε(v2)r+ur,(u,v)(r,0)=(u0,v0)(r).\displaystyle\left\{\begin{array}[]{lll}u_{t}=\frac{1}{r^{n-1}}(r^{n-1}u_{r})_{r}+\frac{1}{r^{n-1}}(r^{n-1}uv)_{r},\qquad\,(r,t)\in(a,b)\times(0,\infty)\\ v_{t}=\varepsilon\frac{1}{r^{n-1}}(r^{n-1}v_{r})_{r}-\varepsilon\frac{n-1}{r^{2}}v-\varepsilon(v^{2})_{r}+u_{r},\\ (u,v)(r,0)=(u_{0},v_{0})(r).\end{array}\right. (1.15)

Similar to (1.6), the Dirichlet boundary conditions for (1.15) are prescribed as

{u|r=a,b=u¯,v|r=a=v¯1,v|r=b=v¯2,ifε>0,u|r=a,b=u¯,ifε=0.\left\{\begin{array}[]{lll}u|_{r=a,b}=\bar{u},\,\,v|_{r=a}=\bar{v}_{1},\,v|_{r=b}=\bar{v}_{2},\quad{\rm if}\,\,\varepsilon>0,\\ u|_{r=a,b}=\bar{u},\quad\qquad\qquad\qquad\qquad\quad\,\,\,{\rm if}\,\,\varepsilon=0.\end{array}\right. (1.16)

In this paper, we shall investigate the asymptotic behavior of solutions to (1.15)-(1.16) as ε0\varepsilon\rightarrow 0 for n2n\geq 2 (if n=1n=1, it coincides with the one-dimensional model (1.5)-(1.6) which has been studied in [16] as aforementioned). In particular, the solution component vv is proved to have a boundary layer due to the mismatch of its boundary values as ε0\varepsilon\rightarrow 0 (see Theorem 2.2).

2 Main results

To study the boundary layer effect, we first present the global well-posedness and regularity estimates for solutions of (1.15)-(1.16) with ε=0\varepsilon=0 in Theorem 2.1. By these estimates, we then state the main result on the convergence for uu and boundary layer formation by vv in Theorem 2.2. Finally, the result is converted to the original chemotaxis model (1.10) via (1.11). We begin with introducing some notations.

Notations. Without loss of generality, we assume 0ε<10\leq\varepsilon<1 since the zero diffusion limit as ε0\varepsilon\rightarrow 0 is our main concern. Throughout this paper, unless specified, we use CC to denote a generic positive constant which is independent of ε\varepsilon and dependent on TT. In contrast, C0C_{0} denotes a generic constant independent of ε\varepsilon and TT. For simplicity, LpL^{p} represents Lp(a,b)L^{p}(a,b) with 1p1\leq p\leq\infty, HkH^{k} denotes Hk(a,b)H^{k}(a,b) with kk\in\mathbb{N} and \|\cdot\| stands for L2\|\cdot\|_{L^{2}}. Moreover, if f(r,t)Lp(a,b)f(r,t)\in L^{p}(a,b) for fixed t>0t>0, we use f(t)Lp\|f(t)\|_{L^{p}} to denote f(,t)Lp\|f(\cdot,t)\|_{L^{p}}.

The first result is on the global well-posedness of (1.15)-(1.16) with ε=0\varepsilon=0.

Theorem 2.1.

Assume that (u0,v0)H2×H2(u_{0},v_{0})\in H^{2}\times H^{2} with u00u_{0}\geq 0 satisfy the compatible conditions u0(a)=u0(b)=u¯u_{0}(a)=u_{0}(b)=\bar{u}. Then the initial-boundary value problem (1.15)-(1.16) with ε=0\varepsilon=0 has a unique solution (u0,v0)C([0,);H2×H2)(u^{0},v^{0})\in C([0,\infty);H^{2}\times H^{2}) such that the following estimates hold true.
(i) If u¯>0\bar{u}>0, there is a constant C0C_{0} independent of tt such that

u0(t)u¯H22+v0(t)H22+0t(u0(τ)u¯H32+u¯(rn1v0)r(τ)H12)𝑑τC0.\begin{split}\|u^{0}(t)-\bar{u}\|_{H^{2}}^{2}+\|v^{0}(t)\|_{H^{2}}^{2}+\int_{0}^{t}\big{(}\|u^{0}(\tau)-\bar{u}\|_{H^{3}}^{2}+\bar{u}\|(r^{n-1}v^{0})_{r}(\tau)\|_{H^{1}}^{2}\big{)}\,d\tau\leq C_{0}.\end{split} (2.1)

Moreover,

limtu0(t)u¯L=0.\lim_{t\rightarrow\infty}\|u^{0}(t)-\bar{u}\|_{L^{\infty}}=0. (2.2)

(ii) If u¯=0\bar{u}=0, for any 0<T<0<T<\infty, there exists a constant CC depending on TT such that

u0L(0,T;H2)+v0L(0,T;H2)+u0L2(0,T;H3)C.\|u^{0}\|_{L^{\infty}(0,T;H^{2})}+\|v^{0}\|_{L^{\infty}(0,T;H^{2})}+\|u^{0}\|_{L^{2}(0,T;H^{3})}\leq C. (2.3)

We proceed to recall the definition of boundary layers (BLs) following the convention of [10, 11].

Definition 2.1.

Denote by (uε,vε)(u^{\varepsilon},v^{\varepsilon}) and (u0,v0)(u^{0},v^{0}) the solution of (1.15)-(1.16) with ε>0\varepsilon>0 and ε=0\varepsilon=0, respectively. If there exists a non-negative function δ=δ(ε)\delta=\delta(\varepsilon) satisfying δ(ε)0\delta(\varepsilon)\rightarrow 0 as ε0\varepsilon\rightarrow 0 such that

limε0uεu0L(0,T;C[a,b])=0,limε0vεv0L(0,T;C[a+δ,bδ])=0,lim infε0vεv0L(0,T;C[a,b])>0,\displaystyle\begin{aligned} &\lim_{\varepsilon\rightarrow 0}\|u^{\varepsilon}-u^{0}\|_{L^{\infty}(0,T;C[a,b])}=0,\\ &\lim_{\varepsilon\rightarrow 0}\|v^{\varepsilon}-v^{0}\|_{L^{\infty}(0,T;C[a+\delta,b-\delta])}=0,\\ &\liminf_{\varepsilon\rightarrow 0}\|v^{\varepsilon}-v^{0}\|_{L^{\infty}(0,T;C[a,b])}>0,\end{aligned}

we say that the initial-boundary value problem (1.15)-(1.16) has a boundary layer solution as ε0\varepsilon\to 0 and δ(ε)\delta(\varepsilon) is called a boundary layer thickness (BL-thickness).

Our main result is as follows.

Theorem 2.2.

Suppose that (u0,v0)H2×H2(u_{0},v_{0})\in H^{2}\times H^{2} with u00u_{0}\geq 0 satisfy the compatible conditions u0(a)=u0(b)=u¯u_{0}(a)=u_{0}(b)=\bar{u} and v0(a)=v¯1,v0(b)=v¯2v_{0}(a)=\bar{v}_{1},v_{0}(b)=\bar{v}_{2}. Let (u0,v0)(u^{0},v^{0}) be the solution obtained in Theorem 2.1. For any 0<T<0<T<\infty, we denote

ε0=min{(8C00TF(t)𝑑t)2,(32C02TeC00TF(t)𝑑t0TF(t)𝑑t)2},\displaystyle\varepsilon_{0}=\min\bigg{\{}\Big{(}8C_{0}\int_{0}^{T}F(t)\,dt\Big{)}^{-2},\,\,\Big{(}32C_{0}^{2}Te^{C_{0}\int_{0}^{T}F(t)\,dt}\int_{0}^{T}F(t)\,dt\Big{)}^{-2}\bigg{\}},

where the function F(t)F(t) is defined in (4.9) by u0(t)H2\|u^{0}(t)\|_{H^{2}}, v0(t)H2\|v^{0}(t)\|_{H^{2}} and the constant C0C_{0} (given in (4.19)) depends only on a,ba,b and nn. Then (1.15)-(1.16) with ε(0,ε0]\varepsilon\in(0,\varepsilon_{0}] admits a unique solution (uε,vε)C([0,T];H2×H2)(u^{\varepsilon},v^{\varepsilon})\in C([0,T];H^{2}\times H^{2}). Furthermore, any function δ=δ(ε)\delta=\delta(\varepsilon) satisfying

δ(ε)0andε1/2/δ(ε)0,asε0\delta(\varepsilon)\rightarrow 0\,\,{\rm and}\,\,\varepsilon^{1/2}/\delta(\varepsilon)\rightarrow 0,\,{\rm as}\,\varepsilon\rightarrow 0 (2.4)

is a BL-thickness of (1.15)-(1.16) such that

uεu0L(0,T;C[a,b])Cε1/4,\|u^{\varepsilon}-u^{0}\|_{L^{\infty}(0,T;C[a,b])}\leq C\varepsilon^{1/4}, (2.5)
vεv0L(0,T;C[a+δ,bδ])Cε1/4δ1/2.\|v^{\varepsilon}-v^{0}\|_{L^{\infty}(0,T;C[a+\delta,b-\delta])}\leq C\varepsilon^{1/4}\delta^{-1/2}. (2.6)

Moreover,

lim infε0vεv0L(0,T;C[a,b])>0\liminf_{\varepsilon\rightarrow 0}\|v^{\varepsilon}-v^{0}\|_{L^{\infty}(0,T;C[a,b])}>0 (2.7)

if and only if

0tur0(a,τ)𝑑τ0or0tur0(b,τ)𝑑τ0,forsomet[0,T].\int_{0}^{t}u^{0}_{r}(a,\tau)\,d\tau\neq 0\qquad{\rm or}\quad\int_{0}^{t}u^{0}_{r}(b,\tau)\,d\tau\neq 0,\quad{\rm for\,\,some}\,\,t\in[0,T]. (2.8)

By employing transformation (1.11), we next convert the above results for (1.15)-(1.16) to the pre-transformed chemotaxis model (1.10). The counterpart of the original model reads as follows:

{ut=1rn1(rn1ur)r1rn1(rn1ucrc)r,ct=ε1rn1(rn1cr)ruc,u(0,r)=u0(r),c(0,r)=c0(r),u|r=a,b=u¯,[cr+v¯1c](a,t)=0,[cr+v¯2c](b,t)=0.\left\{\begin{array}[]{lll}u_{t}=\frac{1}{r^{n-1}}(r^{n-1}u_{r})_{r}-\frac{1}{r^{n-1}}\left(r^{n-1}u\frac{c_{r}}{c}\right)_{r},\\ c_{t}=\varepsilon\frac{1}{r^{n-1}}(r^{n-1}c_{r})_{r}-uc,\\ u(0,r)=u_{0}(r),\,\,\,c(0,r)=c_{0}(r),\\ u|_{r=a,b}=\bar{u},\,\,\,[c_{r}+\bar{v}_{1}c](a,t)=0,\,\,[c_{r}+\bar{v}_{2}c](b,t)=0.\end{array}\right. (2.9)
Proposition 2.1.

Assume c0>0c_{0}>0 and (u0,lnc0)H2×H3(u_{0},\ln c_{0})\in H^{2}\times H^{3}. Suppose that the assumptions in Theorem 2.2 hold with v0=(lnc0)rv_{0}=-(\ln c_{0})_{r}. Let 0<T<0<T<\infty. Then (2.9) with ε[0,ε0]\varepsilon\in[0,\varepsilon_{0}] admits a unique solution (uε,cε)C([0,T];H2×H3)(u^{\varepsilon},c^{\varepsilon})\in C([0,T];H^{2}\times H^{3}) such that

uεu0L(0,T;C[a,b])Cε1/4,cεc0L(0,T;C[a,b])Cε1/4.\begin{split}\|u^{\varepsilon}-u^{0}\|_{L^{\infty}(0,T;C[a,b])}\leq C\varepsilon^{1/4},\\ \|c^{\varepsilon}-c^{0}\|_{L^{\infty}(0,T;C[a,b])}\leq C\varepsilon^{1/4}.\end{split} (2.10)

Moreover, the gradient of cc has a boundary layer effect as ε0\varepsilon\rightarrow 0, that is

crεcr0L(0,T;C[a+δ,bδ])Cε1/4δ1/2,\|c^{\varepsilon}_{r}-c^{0}_{r}\|_{L^{\infty}(0,T;C[a+\delta,b-\delta])}\leq C\varepsilon^{1/4}\delta^{-1/2}, (2.11)

with the function δ(ε)\delta(\varepsilon) defined (2.4) and the following estimate holds

lim infε0crεcr0L(0,T;C[a,b])>0,\liminf_{\varepsilon\rightarrow 0}\|c^{\varepsilon}_{r}-c^{0}_{r}\|_{L^{\infty}(0,T;C[a,b])}>0, (2.12)

if and only if (2.8) is true.

At the end of this section, we briefly introduce the main ideas used in the paper. Although the system (1.15)-(1.16) with n2n\geq 2 is in a similar form to its counterpart with n=1n=1 for which the vanishing diffusion limit has been studied in [16] based on a ε\varepsilon-independent estimate for solutions with ε>0\varepsilon>0, the methods used there can not be applied to study the present problem since when n2n\geq 2 the system (1.15)-(1.16) with ε>0\varepsilon>0 lacks an energy-like structure or a Lyapunov function to provide a preliminary estimate uniformly in ε\varepsilon. Moreover, one can not use the estimates derived in [50] for the present problem either since those estimates depend on ε\varepsilon. The difficulty in our analysis consists in deriving the ε\varepsilon-convergence estimates in (2.5) and (2.6) without any uniform-in-ε\varepsilon priori bounds on solutions (uε,vε)(u^{\varepsilon},v^{\varepsilon}). Inspired by the works [5, 51], this will be achieved in section 4 by regarding (uε,vε)(u^{\varepsilon},v^{\varepsilon}) with small ε>0\varepsilon>0 as a perturbation of (u0,v0)(u^{0},v^{0}) and then estimating their difference (uεu0,vεv0)(u^{\varepsilon}-u^{0},v^{\varepsilon}-v^{0}) by the method of energy estimates and a new Gronwall’s type inequality (see Lemma 4.1) on ODEs. The proof of Theorem 2.1 is standard and will be given in section 3.

3 Proof of Theorem 2.1

This section is to prove Theorem 2.1 based on the following lemmas where the a priori estimates on solution (u0,v0)(u^{0},v^{0}) of (1.15)-(1.16) with ε=0\varepsilon=0 are derived by the energy method. We set off by rewriting (1.15)-(1.16) with ε=0\varepsilon=0 as follows:

{ut0=1rn1(rn1ur0)r+1rn1(rn1u0v0)r,vt0=ur0,(u0,v0)(r,0)=(u0,v0)(r),u0(a,t)=u0(b,t)=u¯.\displaystyle\left\{\begin{array}[]{lll}u^{0}_{t}=\frac{1}{r^{n-1}}\big{(}r^{n-1}u^{0}_{r}\big{)}_{r}+\frac{1}{r^{n-1}}\big{(}r^{n-1}u^{0}v^{0}\big{)}_{r},\\ v^{0}_{t}=u^{0}_{r},\\ (u^{0},v^{0})(r,0)=(u_{0},v_{0})(r),\\ u^{0}(a,t)=u^{0}(b,t)=\bar{u}.\end{array}\right. (3.5)
Lemma 3.1.

Suppose the assumptions in Theorem 2.1 hold and u¯>0\bar{u}>0. Then there exists a positive constant C0C_{0} independent of tt such that

abrn1[(u0lnu0u0)(t)(u¯lnu¯u¯)lnu¯(u0(t)u¯)]dr+12abrn1(v0)2(t)𝑑r+0tabrn1(ur0)2u0𝑑r𝑑τC0\begin{split}\int_{a}^{b}&r^{n-1}[(u^{0}\ln u^{0}-u^{0})(t)-(\bar{u}\ln\bar{u}-\bar{u})-\ln\bar{u}(u^{0}(t)-\bar{u})]dr\\ +&\frac{1}{2}\int_{a}^{b}r^{n-1}(v^{0})^{2}(t)dr+\int_{0}^{t}\int_{a}^{b}r^{n-1}\frac{(u^{0}_{r})^{2}}{u^{0}}drd\tau\leq C_{0}\end{split} (3.6)

and

r(n1)/2[u0(t)u¯]2+0tr(n1)/2ur0(τ)2𝑑τC0.\|r^{(n-1)/2}[u^{0}(t)-\bar{u}]\|^{2}+\int_{0}^{t}\|r^{(n-1)/2}u^{0}_{r}(\tau)\|^{2}d\tau\leq C_{0}. (3.7)
Proof.

Taking the L2L^{2} inner products of the first and second equation of (3.5) with rn1(lnu0lnu¯)r^{n-1}(\ln u^{0}-\ln\bar{u}) and rn1v0r^{n-1}v^{0} respectively, we then add the results and use integration by parts to get

ddtabrn1[(u0lnu0u0)(u¯lnu¯u¯)lnu¯(u0u¯)]𝑑r+12ddtabrn1(v0)2𝑑r+abrn1(ur0)2u0𝑑r=0,\displaystyle\begin{split}\frac{d}{dt}&\int_{a}^{b}r^{n-1}[(u^{0}\ln u^{0}-u^{0})-(\bar{u}\ln\bar{u}-\bar{u})-\ln\bar{u}(u^{0}-\bar{u})]dr\\ +&\frac{1}{2}\frac{d}{dt}\int_{a}^{b}r^{n-1}(v^{0})^{2}dr+\int_{a}^{b}r^{n-1}\frac{(u^{0}_{r})^{2}}{u^{0}}dr=0,\end{split}

which gives rise to (3.6) upon integration over (0,t)(0,t). To prove (3.7), we denote u~(r,t)=u0(r,t)u¯\tilde{u}(r,t)=u^{0}(r,t)-\bar{u} and find from (3.5) that (u~,v0)(r,t)(\tilde{u},v^{0})(r,t) satisfies

{u~t=1rn1(rn1u~r)r+1rn1(rn1u~v0)r+u¯rn1(rn1v0)r,vt0=u~r,(u~,v0)(r,0)=(u0u¯,v0)(r),u~(a,t)=u~(b,t)=0.\displaystyle\left\{\begin{array}[]{lll}\tilde{u}_{t}=\frac{1}{r^{n-1}}(r^{n-1}\tilde{u}_{r})_{r}+\frac{1}{r^{n-1}}(r^{n-1}\tilde{u}v^{0})_{r}+\frac{\bar{u}}{r^{n-1}}(r^{n-1}v^{0})_{r},\\ v^{0}_{t}=\tilde{u}_{r},\\ (\tilde{u},v^{0})(r,0)=(u_{0}-\bar{u},v_{0})(r),\\ \tilde{u}(a,t)=\tilde{u}(b,t)=0.\end{array}\right. (3.12)

Multiplying the first and second equation of (3.12) by rn1u~r^{n-1}\tilde{u} and u¯rn1v0\bar{u}r^{n-1}v^{0}, respectively. Adding the results gives

12ddt(r(n1)/2u~2+u¯r(n1)/2v02)+r(n1)/2u~r2=abrn1u~v0u~r𝑑r12r(n1)/2u~r2+12u~L2r(n1)/2v02.\begin{split}\frac{1}{2}&\frac{d}{dt}\Big{(}\|r^{(n-1)/2}\tilde{u}\|^{2}+\bar{u}\|r^{(n-1)/2}v^{0}\|^{2}\Big{)}+\|r^{(n-1)/2}\tilde{u}_{r}\|^{2}\\ =&-\int_{a}^{b}r^{n-1}\tilde{u}v^{0}\tilde{u}_{r}dr\\ \leq&\frac{1}{2}\|r^{(n-1)/2}\tilde{u}_{r}\|^{2}+\frac{1}{2}\|\tilde{u}\|_{L^{\infty}}^{2}\|r^{(n-1)/2}v^{0}\|^{2}.\end{split} (3.13)

Note that u~L\|\tilde{u}\|_{L^{\infty}} can be estimated as follows

|u~(r,t)|=|u0(r,t)u¯|=|arur0𝑑r|(abu0𝑑r)1/2(ab(ur0)2u0𝑑r)1/2.\displaystyle|\tilde{u}(r,t)|=|u^{0}(r,t)-\bar{u}|=\left|\int_{a}^{r}u^{0}_{r}dr\right|\leq\left(\int_{a}^{b}u^{0}dr\right)^{1/2}\left(\int_{a}^{b}\frac{(u^{0}_{r})^{2}}{u^{0}}dr\right)^{1/2}.

Then substituting the above estimate into (3.13) and integrating the result over (0,t)(0,t) we have

12r(n1)/2u~(t)2+12u¯r(n1)/2v0(t)2+120tr(n1)/2u~r2𝑑τ120tab(ur0)2u0𝑑r𝑑τu0L(0,t;L1)r(n1)/2v0L(0,t;L2)2,\displaystyle\begin{split}\frac{1}{2}&\|r^{(n-1)/2}\tilde{u}(t)\|^{2}+\frac{1}{2}\bar{u}\|r^{(n-1)/2}v^{0}(t)\|^{2}+\frac{1}{2}\int_{0}^{t}\|r^{(n-1)/2}\tilde{u}_{r}\|^{2}\,d\tau\\ \leq&\frac{1}{2}\int_{0}^{t}\int_{a}^{b}\frac{(u^{0}_{r})^{2}}{u^{0}}\,drd\tau\cdot\|u^{0}\|_{L^{\infty}(0,t;L^{1})}\|r^{(n-1)/2}v^{0}\|^{2}_{L^{\infty}(0,t;L^{2})},\end{split}

which, along with (3.6) and the fact

u0L(0,t;L1)C0supτ[0,t]{abrn1[(u0lnu0u0)(τ)(u¯lnu¯u¯)lnu¯(u0(τ)u¯)]𝑑r}\displaystyle\|u^{0}\|_{L^{\infty}(0,t;L^{1})}\leq C_{0}\sup_{\tau\in[0,t]}\left\{\int_{a}^{b}r^{n-1}[(u^{0}\ln u^{0}-u^{0})(\tau)-(\bar{u}\ln\bar{u}-\bar{u})-\ln\bar{u}(u^{0}(\tau)-\bar{u})]dr\right\}

implies (3.7). The proof is completed.

We proceed to derive higher regularity properties for the solution (u~,v0)(\tilde{u},v^{0}) of (3.12).

Lemma 3.2.

Suppose the assumptions in Theorem 2.1 hold and u¯>0\bar{u}>0. Let (u~,v0)(r,t)(\tilde{u},v^{0})(r,t) be the solution of (3.12). Then there is a constant C0C_{0} independent of tt such that

(rn1v0)r(t)2+r(n1)/2u~r(t)2+0t(u¯(rn1v0)r2+r(n1)/2u~t2)𝑑τC0.\|(r^{n-1}v^{0})_{r}(t)\|^{2}+\|r^{(n-1)/2}\tilde{u}_{r}(t)\|^{2}+\int_{0}^{t}\big{(}\bar{u}\|(r^{n-1}v^{0})_{r}\|^{2}+\|r^{(n-1)/2}\tilde{u}_{t}\|^{2}\big{)}d\tau\leq C_{0}. (3.14)
Proof.

We multiply the second equation of (3.12) with rn1r^{n-1} and differentiate the resulting equation with respect to rr. Then from the first equation of (3.12) we obtain

(rn1v0)rt=(rn1u~r)r=rn1u~t(rn1u~v0)ru¯(rn1v0)r.(r^{n-1}v^{0})_{rt}=(r^{n-1}\tilde{u}_{r})_{r}=r^{n-1}\tilde{u}_{t}-(r^{n-1}\tilde{u}v^{0})_{r}-\bar{u}(r^{n-1}v^{0})_{r}. (3.15)

Taking the L2L^{2} inner product of (3.15)\eqref{e9} against 2(rn1v0)r2(r^{n-1}v^{0})_{r} to get

ddt(rn1v0)r2+2u¯(rn1v0)r2=2abrn1u~t(rn1v0)r𝑑r2ab(rn1u~v0)r(rn1v0)r𝑑r:=I1+I2.\begin{split}\frac{d}{dt}&\|(r^{n-1}v^{0})_{r}\|^{2}+2\bar{u}\|(r^{n-1}v^{0})_{r}\|^{2}\\ =&2\int_{a}^{b}r^{n-1}\tilde{u}_{t}(r^{n-1}v^{0})_{r}dr-2\int_{a}^{b}(r^{n-1}\tilde{u}v^{0})_{r}(r^{n-1}v^{0})_{r}dr\\ :=&I_{1}+I_{2}.\end{split} (3.16)

We may rewrite I1I_{1} as

I1=2ddtab(rn1u~)(rn1v0)r𝑑r2ab(rn1u~)(rn1v0)rt𝑑r:=M1+M2,\displaystyle\begin{split}I_{1}=2\frac{d}{dt}\int_{a}^{b}(r^{n-1}\tilde{u})(r^{n-1}v^{0})_{r}dr-2\int_{a}^{b}(r^{n-1}\tilde{u})(r^{n-1}v^{0})_{rt}dr:=M_{1}+M_{2},\end{split}

where M1M_{1} can be reorganized as

M1=ddt(12(rn1v0)r2+2rn1u~212(rn1v0)r2rn1u~2)\displaystyle M_{1}=\frac{d}{dt}\left(\frac{1}{2}\|(r^{n-1}v^{0})_{r}\|^{2}+2\|r^{n-1}\tilde{u}\|^{2}-\left\|\frac{1}{\sqrt{2}}(r^{n-1}v^{0})_{r}-\sqrt{2}r^{n-1}\tilde{u}\right\|^{2}\right)

and M2M_{2} can be estimated by (3.15) and the Poincaré inequality as

M2=2ab(rn1u~)(rn1u~r)r𝑑r=2ab(rn1u~)r(rn1u~r)𝑑rC0r(n1)/2u~r2.\displaystyle\begin{split}M_{2}&=-2\int_{a}^{b}(r^{n-1}\tilde{u})(r^{n-1}\tilde{u}_{r})_{r}dr&=2\int_{a}^{b}(r^{n-1}\tilde{u})_{r}(r^{n-1}\tilde{u}_{r})dr\leq C_{0}\|r^{(n-1)/2}\tilde{u}_{r}\|^{2}.\end{split}

Hence

I1ddt(12(rn1v0)r2+2rn1u~212(rn1v0)r2rn1u~2)+C0r(n1)/2u~r2.\displaystyle I_{1}\leq\frac{d}{dt}\left(\frac{1}{2}\|(r^{n-1}v^{0})_{r}\|^{2}+2\|r^{n-1}\tilde{u}\|^{2}-\left\|\frac{1}{\sqrt{2}}(r^{n-1}v^{0})_{r}-\sqrt{2}r^{n-1}\tilde{u}\right\|^{2}\right)+C_{0}\|r^{(n-1)/2}\tilde{u}_{r}\|^{2}.

To estimate I2I_{2}, we first note that for fixed t>0t>0 if f(r,t)H1f(r,t)\in H^{1} satisfies f|r=a,b=0f|_{r=a,b}=0 it follows that f(r,t)2=2arffr𝑑r2f(t)fr(t),f(r,t)^{2}=2\int_{a}^{r}ff_{r}dr\leq 2\|f(t)\|\|f_{r}(t)\|, which leads to

f(t)L2f(t)1/2fr(t)1/2andf(t)LC0fr(t),\|f(t)\|_{L^{\infty}}\leq\sqrt{2}\|f(t)\|^{1/2}\|f_{r}(t)\|^{1/2}\quad\,\,{\rm and}\,\,\,\,\|f(t)\|_{L^{\infty}}\leq C_{0}\|f_{r}(t)\|, (3.17)

thanks to the Poincaré inequality f(t)C0fr(t)\|f(t)\|\leq C_{0}\|f_{r}(t)\|. Then we deduce from (3.17) and the Sobolev embedding inequality that

I2u¯2(rn1v0)r2+4u¯u~L2(rn1v0)r2+4u¯u~r2rn1v0L2u¯2(rn1v0)r2+C0r(n1)/2u~r2(r(n1)/2v02+(rn1v0)r2).\begin{split}I_{2}\leq&\frac{\bar{u}}{2}\|(r^{n-1}v^{0})_{r}\|^{2}+\frac{4}{\bar{u}}\|\tilde{u}\|_{L^{\infty}}^{2}\|(r^{n-1}v^{0})_{r}\|^{2}+\frac{4}{\bar{u}}\|\tilde{u}_{r}\|^{2}\|r^{n-1}v^{0}\|_{L^{\infty}}^{2}\\ \leq&\frac{\bar{u}}{2}\|(r^{n-1}v^{0})_{r}\|^{2}+C_{0}\|r^{(n-1)/2}\tilde{u}_{r}\|^{2}\big{(}\|r^{(n-1)/2}v^{0}\|^{2}+\|(r^{n-1}v^{0})_{r}\|^{2}\big{)}.\end{split} (3.18)

Substituting the above estimates for I1I_{1} and I2I_{2} into (3.16), one derives

ddt(12(rn1v0)r2+12(rn1v0)r2rn1u~2)+32u¯(rn1v0)r2C0r(n1)/2u~r2(rn1v0)r2+C0r(n1)/2u~r2(r(n1)/2v02+1)+2ddtrn1u~2.\displaystyle\begin{split}\frac{d}{dt}&\left(\frac{1}{2}\|(r^{n-1}v^{0})_{r}\|^{2}+\|\frac{1}{\sqrt{2}}(r^{n-1}v^{0})_{r}-\sqrt{2}r^{n-1}\tilde{u}\|^{2}\right)+\frac{3}{2}\bar{u}\|(r^{n-1}v^{0})_{r}\|^{2}\\ \leq&C_{0}\|r^{(n-1)/2}\tilde{u}_{r}\|^{2}\|(r^{n-1}v^{0})_{r}\|^{2}+C_{0}\|r^{(n-1)/2}\tilde{u}_{r}\|^{2}\big{(}\|r^{(n-1)/2}v^{0}\|^{2}+1\big{)}+2\frac{d}{dt}\|r^{n-1}\tilde{u}\|^{2}.\end{split}

Then applying Gronwall’s inequality to the above result and using Lemma 3.1, we conclude that

(rn1v0)r(t)2+u¯0t(rn1v0)r2𝑑τC0.\|(r^{n-1}v^{0})_{r}(t)\|^{2}+\bar{u}\int_{0}^{t}\|(r^{n-1}v^{0})_{r}\|^{2}d\tau\leq C_{0}. (3.19)

We proceed to estimate r(n1)/2u~r(t)\|r^{(n-1)/2}\tilde{u}_{r}(t)\| by multiplying the first equation of (3.12) with 2rn1u~t2r^{n-1}\tilde{u}_{t} in L2L^{2} and derive

ddtr(n1)/2u~r2+2r(n1)/2u~t2=2ab(rn1u~v0)ru~t𝑑r+2u¯ab(rn1v0)ru~t𝑑r:=I3+I4.\begin{split}\frac{d}{dt}\|r^{(n-1)/2}\tilde{u}_{r}\|^{2}+2\|r^{(n-1)/2}\tilde{u}_{t}\|^{2}&=2\int_{a}^{b}(r^{n-1}\tilde{u}v^{0})_{r}\tilde{u}_{t}dr+2\bar{u}\int_{a}^{b}(r^{n-1}v^{0})_{r}\tilde{u}_{t}dr\\ &:=I_{3}+I_{4}.\end{split} (3.20)

By similar arguments as deriving (3.18), we estimate I3I_{3} as

I312r(n1)/2u~t2+C0r(n1)/2u~r2(r(n1)/2v02+(rn1v0)r2)\displaystyle\begin{split}I_{3}\leq\frac{1}{2}\|r^{(n-1)/2}\tilde{u}_{t}\|^{2}+C_{0}\|r^{(n-1)/2}\tilde{u}_{r}\|^{2}(\|r^{(n-1)/2}v^{0}\|^{2}+\|(r^{n-1}v^{0})_{r}\|^{2})\end{split}

and by the Cauchy-Schwarz inequality, I4I_{4} is estimated as

I412r(n1)/2u~t2+C0(rn1v0)r2.\displaystyle I_{4}\leq\frac{1}{2}\|r^{(n-1)/2}\tilde{u}_{t}\|^{2}+C_{0}\|(r^{n-1}v^{0})_{r}\|^{2}.

Then feeding (3.20) on the above estimates for I3I_{3} and I4I_{4}, we have

ddtr(n1)/2u~r2+r(n1)/2u~t2C0r(n1)/2u~r2(r(n1)/2v02+(rn1v0)r2)+C0(rn1v0)r2.\begin{split}\frac{d}{dt}&\|r^{(n-1)/2}\tilde{u}_{r}\|^{2}+\|r^{(n-1)/2}\tilde{u}_{t}\|^{2}\\ \leq&C_{0}\|r^{(n-1)/2}\tilde{u}_{r}\|^{2}(\|r^{(n-1)/2}v^{0}\|^{2}+\|(r^{n-1}v^{0})_{r}\|^{2})+C_{0}\|(r^{n-1}v^{0})_{r}\|^{2}.\end{split} (3.21)

Integrating (3.21) over (0,t)(0,t) and using (3.7) and (3.19), one arrives at

r(n1)/2u~r(t)2+0tr(n1)/2u~t2𝑑τC0,\displaystyle\|r^{(n-1)/2}\tilde{u}_{r}(t)\|^{2}+\int_{0}^{t}\|r^{(n-1)/2}\tilde{u}_{t}\|^{2}d\tau\leq C_{0},

which, in conjunction with (3.19) gives (3.14). The proof is completed.

Lemma 3.3.

Suppose that the assumptions in Theorem 2.1 hold and u¯>0\bar{u}>0. Then there exists a constant C0C_{0} independent of tt such that

r(n1)/2u~t(t)2+(rn1v0)rr(t)2+0t(r(n1)/2u~rt2+u¯(rn1v0)rr2)𝑑τC0\|r^{(n-1)/2}\tilde{u}_{t}(t)\|^{2}+\|(r^{n-1}v^{0})_{rr}(t)\|^{2}+\int_{0}^{t}\big{(}\|r^{(n-1)/2}\tilde{u}_{rt}\|^{2}+\bar{u}\|(r^{n-1}v^{0})_{rr}\|^{2}\big{)}d\tau\leq C_{0} (3.22)

and

(rn1u~r)r(t)2+0t((rn1u~r)r2+(rn1u~r)rr2)𝑑τC0.\|(r^{n-1}\tilde{u}_{r})_{r}(t)\|^{2}+\int_{0}^{t}\big{(}\|(r^{n-1}\tilde{u}_{r})_{r}\|^{2}+\|(r^{n-1}\tilde{u}_{r})_{rr}\|^{2}\big{)}d\tau\leq C_{0}. (3.23)
Proof.

Differentiating the first equation of (3.12) with respect to tt and multiplying the result with 2rn1u~t2r^{n-1}\tilde{u}_{t}, we get upon integration by parts that

ddtr(n1)/2u~t2+2r(n1)/2u~rt2=2ab(rn1u~v0)tu~rt𝑑r2u¯ab(rn1v0)tu~rt:=I5+I6.\begin{split}\frac{d}{dt}&\|r^{(n-1)/2}\tilde{u}_{t}\|^{2}+2\|r^{(n-1)/2}\tilde{u}_{rt}\|^{2}\\ &=-2\int_{a}^{b}(r^{n-1}\tilde{u}v^{0})_{t}\tilde{u}_{rt}dr-2\bar{u}\int_{a}^{b}(r^{n-1}v^{0})_{t}\tilde{u}_{rt}\\ &:=I_{5}+I_{6}.\end{split} (3.24)

By (3.17) and the second equation of (3.12) we have that

I5C0(u~tLr(n1)/2v0r(n1)/2u~rt+u~Lvt0r(n1)/2u~rt)C0(r(n1)/2u~t1/2r(n1)/2u~rt3/2r(n1)/2v0+r(n1)/2u~ru~rr(n1)/2u~rt)12r(n1)/2u~rt2+C0(r(n1)/2u~t2r(n1)/2v04+r(n1)/2u~r4).\displaystyle\begin{split}I_{5}\leq&C_{0}\left(\|\tilde{u}_{t}\|_{L^{\infty}}\|r^{(n-1)/2}v^{0}\|\|r^{(n-1)/2}\tilde{u}_{rt}\|+\|\tilde{u}\|_{L^{\infty}}\|v^{0}_{t}\|\|r^{(n-1)/2}\tilde{u}_{rt}\|\right)\\ \leq&C_{0}\left(\|r^{(n-1)/2}\tilde{u}_{t}\|^{1/2}\|r^{(n-1)/2}\tilde{u}_{rt}\|^{3/2}\|r^{(n-1)/2}v^{0}\|+\|r^{(n-1)/2}\tilde{u}_{r}\|\|\tilde{u}_{r}\|\|r^{(n-1)/2}\tilde{u}_{rt}\|\right)\\ \leq&\frac{1}{2}\|r^{(n-1)/2}\tilde{u}_{rt}\|^{2}+C_{0}\left(\|r^{(n-1)/2}\tilde{u}_{t}\|^{2}\|r^{(n-1)/2}v^{0}\|^{4}+\|r^{(n-1)/2}\tilde{u}_{r}\|^{4}\right).\end{split}

We use again the second equation of (3.12) and Cauchy-Schwarz inequality to get

I612r(n1)/2u~rt2+2u¯2r(n1)/2u~r2.\displaystyle I_{6}\leq\frac{1}{2}\|r^{(n-1)/2}\tilde{u}_{rt}\|^{2}+2\bar{u}^{2}\|r^{(n-1)/2}\tilde{u}_{r}\|^{2}.

Substituting the above estimates for I5I_{5}-I6I_{6} into (3.24), then integrating the results over (0,t)(0,t) and using Lemma 3.1 along with Lemma 3.2, we conclude that

r(n1)/2u~t(t)2+0tr(n1)/2u~rt2𝑑τC0.\|r^{(n-1)/2}\tilde{u}_{t}(t)\|^{2}+\int_{0}^{t}\|r^{(n-1)/2}\tilde{u}_{rt}\|^{2}d\tau\leq C_{0}. (3.25)

We proceed to estimating the remaining part (rn1v0)rr(t)2+u¯0t(rn1v0)rr2𝑑τ\|(r^{n-1}v^{0})_{rr}(t)\|^{2}+\bar{u}\int_{0}^{t}\|(r^{n-1}v^{0})_{rr}\|^{2}d\tau in (3.22). Differentiating (3.15) with respect to rr and multiplying the resulting equation with 2(rn1v0)rr2(r^{n-1}v^{0})_{rr} we get

ddt(rn1v0)rr2+2u¯(rn1v0)rr2=2ab(rn1u~t)r(rn1v0)rr𝑑r2ab(rn1u~v0)rr(rn1v0)rr:=I7+I8.\begin{split}\frac{d}{dt}&\|(r^{n-1}v^{0})_{rr}\|^{2}+2\bar{u}\|(r^{n-1}v^{0})_{rr}\|^{2}\\ =&2\int_{a}^{b}(r^{n-1}\tilde{u}_{t})_{r}(r^{n-1}v^{0})_{rr}dr\\ &-2\int_{a}^{b}(r^{n-1}\tilde{u}v^{0})_{rr}(r^{n-1}v^{0})_{rr}\\ :=&I_{7}+I_{8}.\end{split} (3.26)

To estimate of I7I_{7}, we note for g(r,t)L2(a,b)g(r,t)\in L^{2}(a,b) with fixed t>0t>0, it follows that

b(n1)r(n1)/2g(t)2g(t)2a(n1)r(n1)/2g(t)2.b^{-(n-1)}\|r^{(n-1)/2}g(t)\|^{2}\leq\|g(t)\|^{2}\leq a^{-(n-1)}\|r^{(n-1)/2}g(t)\|^{2}. (3.27)

Then from Cauchy-Schwarz inequality and (3.27) one derives

I7u¯2(rn1v0)rr2+C0(r(n1)/2u~t2+r(n1)/2u~rt2).\displaystyle I_{7}\leq\frac{\bar{u}}{2}\|(r^{n-1}v^{0})_{rr}\|^{2}+C_{0}(\|r^{(n-1)/2}\tilde{u}_{t}\|^{2}+\|r^{(n-1)/2}\tilde{u}_{rt}\|^{2}).

To bound I8I_{8} we first estimate 0t(rn1u~r)r2𝑑τ\int_{0}^{t}\|(r^{n-1}\tilde{u}_{r})_{r}\|^{2}d\tau by the first equation of (3.12) as follows:

0t(rn1u~r)r2𝑑τ0trn1u~t2𝑑τ+C00tu~r2𝑑τ(rn1v0)L(0,t;H1)2+C0u¯20t(rn1v0)r2𝑑τC0,\begin{split}\int_{0}^{t}\|(r^{n-1}\tilde{u}_{r})_{r}\|^{2}d\tau\leq&\int_{0}^{t}\|r^{n-1}\tilde{u}_{t}\|^{2}d\tau+C_{0}\int_{0}^{t}\|\tilde{u}_{r}\|^{2}d\tau\cdot\|(r^{n-1}v^{0})\|_{L^{\infty}(0,t;H^{1})}^{2}\\ &+C_{0}\bar{u}^{2}\int_{0}^{t}\|(r^{n-1}v^{0})_{r}\|^{2}d\tau\\ \leq&C_{0},\end{split} (3.28)

where (3.17) and Lemma 3.1 - Lemma 3.2 have been used. Then (3.28) along with (3.27) and (3.7) implies that

0tu~rr2𝑑τC00t((rn1u~r)r2+r(n1)/2u~r2)𝑑τC0,\int_{0}^{t}\|\tilde{u}_{rr}\|^{2}d\tau\leq C_{0}\int_{0}^{t}(\|(r^{n-1}\tilde{u}_{r})_{r}\|^{2}+\|r^{(n-1)/2}\tilde{u}_{r}\|^{2})d\tau\leq C_{0}, (3.29)

where the constant C0C_{0} depends on aa and bb. Noting that (rn1u~v0)rr=(rn1v0)rru~+2(rn1v0)ru~r+(rn1v0)u~rr(r^{n-1}\tilde{u}v^{0})_{rr}=(r^{n-1}v^{0})_{rr}\tilde{u}+2(r^{n-1}v^{0})_{r}\tilde{u}_{r}+(r^{n-1}v^{0})\tilde{u}_{rr}, one deduces by (3.17) and the Sobolev embedding inequality that

I8u¯2(rn1v0)rr2+2u¯(rn1u~v0)rr2u¯2(rn1v0)rr2+C0(u~r2(rn1v0)rr2+u~r2(rn1v0)r2+u~rr2(rn1v0)r2+u~rr2v02).\displaystyle\begin{split}I_{8}\leq&\frac{\bar{u}}{2}\|(r^{n-1}v^{0})_{rr}\|^{2}+\frac{2}{\bar{u}}\|(r^{n-1}\tilde{u}v^{0})_{rr}\|^{2}\\ \leq&\frac{\bar{u}}{2}\|(r^{n-1}v^{0})_{rr}\|^{2}+C_{0}\big{(}\|\tilde{u}_{r}\|^{2}\|(r^{n-1}v^{0})_{rr}\|^{2}\\ &+\|\tilde{u}_{r}\|^{2}\|(r^{n-1}v^{0})_{r}\|^{2}+\|\tilde{u}_{rr}\|^{2}\|(r^{n-1}v^{0})_{r}\|^{2}+\|\tilde{u}_{rr}\|^{2}\|v^{0}\|^{2}\big{)}.\end{split}

We feed (3.26) on the above estimates for I7I_{7}-I8I_{8} then apply Gronwall’s inequality, Lemma 3.1 - Lemma 3.2, (3.25) and (3.29) to the result to find

(rn1v0)rr(t)2+u¯0t(rn1v0)rr2𝑑τC0,\displaystyle\|(r^{n-1}v^{0})_{rr}(t)\|^{2}+\bar{u}\int_{0}^{t}\|(r^{n-1}v^{0})_{rr}\|^{2}d\tau\leq C_{0},

which, along with (3.25) yields (3.22). We next prove (3.23). By similar arguments as deriving (3.28) one gets

(rn1u~r)r(t)2rn1u~t(t)2+C0u~r(t)2(rn1v0)(t)H12+C0u¯2(rn1v0)r(t)2C0,\begin{split}\|(r^{n-1}\tilde{u}_{r})_{r}(t)\|^{2}\leq&\|r^{n-1}\tilde{u}_{t}(t)\|^{2}+C_{0}\|\tilde{u}_{r}(t)\|^{2}\|(r^{n-1}v^{0})(t)\|_{H^{1}}^{2}\\ &+C_{0}\bar{u}^{2}\|(r^{n-1}v^{0})_{r}(t)\|^{2}\\ \leq&C_{0},\end{split} (3.30)

where (3.25) and Lemma 3.1 - Lemma 3.2 have been used. We differentiate (3.15) with respect to rr and conclude that

0t(rn1u~r)rr2dτC0(0trn1u~rt2𝑑τ+0tr(n1)/2u~t2𝑑τ+u¯20t(rn1v0)rr2𝑑τ)+C00t(u~r2+u~rr2)𝑑τ(rn1v0)L(0,t;H2)2C0,\begin{split}\int_{0}^{t}&\|(r^{n-1}\tilde{u}_{r})_{rr}\|^{2}d\tau\\ \leq&C_{0}\left(\int_{0}^{t}\|r^{n-1}\tilde{u}_{rt}\|^{2}d\tau+\int_{0}^{t}\|r^{(n-1)/2}\tilde{u}_{t}\|^{2}d\tau+\bar{u}^{2}\int_{0}^{t}\|(r^{n-1}v^{0})_{rr}\|^{2}d\tau\right)\\ &+C_{0}\int_{0}^{t}\big{(}\|\tilde{u}_{r}\|^{2}+\|\tilde{u}_{rr}\|^{2}\big{)}d\tau\cdot\|(r^{n-1}v^{0})\|_{L^{\infty}(0,t;H^{2})}^{2}\\ \leq&C_{0},\end{split} (3.31)

where we have used (3.22), Lemma 3.1 and Lemma 3.2. Finally collecting (3.28), (3.30) and (3.31) we derive (3.23). The proof is finished.

We are now in the position to prove Theorem 2.1 by the above Lemma 3.1 - Lemma 3.3.

Proof of Theorem 2.1. We first prove Part (i) of Theorem 2.1. By Lemma 3.1 and (3.27), one derives

v0(t)2C0r(n1)/2v0(t)2C0,u~(t)2+0tu~H12𝑑τC0,\|v^{0}(t)\|^{2}\leq C_{0}\|r^{(n-1)/2}v^{0}(t)\|^{2}\leq C_{0},\qquad\|\tilde{u}(t)\|^{2}+\int_{0}^{t}\|\tilde{u}\|_{H^{1}}^{2}d\tau\leq C_{0}, (3.32)

where the constant C0C_{0} depends on a,ba,b and nn and the Poincaré inequality u~2C0u~r2\|\tilde{u}\|^{2}\leq C_{0}\|\tilde{u}_{r}\|^{2} has been used. On the other hand, for f(r,t)H1f(r,t)\in H^{1} with fixed tt we have

fr2=r(n1)[(rn1f)r(n1)rn2f]2a2(n2)(rn1f)r2+a2(n1)(n1)b2(n2)f2C0((rn1f)r2+f2).\begin{split}\|f_{r}\|^{2}=&\|r^{-(n-1)}[(r^{n-1}f)_{r}-(n-1)r^{n-2}f]\|^{2}\\ \leq&a^{-2(n-2)}\|(r^{n-1}f)_{r}\|^{2}+a^{-2(n-1)}(n-1)b^{2(n-2)}\|f\|^{2}\\ \leq&C_{0}(\|(r^{n-1}f)_{r}\|^{2}+\|f\|^{2}).\end{split} (3.33)

Then it follows from Lemma 3.2, Lemma 3.1, (3.27) and (3.33) that

vr0(t)2+u~r(t)2+0t(u¯(r(n1)v0)r2+u~t2)𝑑τC0.\|v^{0}_{r}(t)\|^{2}+\|\tilde{u}_{r}(t)\|^{2}+\int_{0}^{t}\big{(}\bar{u}\|(r^{(n-1)}v^{0})_{r}\|^{2}+\|\tilde{u}_{t}\|^{2}\big{)}d\tau\leq C_{0}. (3.34)

Similarly, it follows from Lemma 3.3 and (3.33) that

vrr0(t)2+u~rr(t)2+0t(u¯(r(n1)v0)rr2+u~rr2+u~rrr2)𝑑τC0.\|v^{0}_{rr}(t)\|^{2}+\|\tilde{u}_{rr}(t)\|^{2}+\int_{0}^{t}\left(\bar{u}\|(r^{(n-1)}v^{0})_{rr}\|^{2}+\|\tilde{u}_{rr}\|^{2}+\|\tilde{u}_{rrr}\|^{2}\right)\,d\tau\leq C_{0}. (3.35)

Thus collecting (3.32), (3.34) and (3.35) we derive the desired a priori estimate (2.1), which along with the fixed point theorem implies the existence of solution (u0,v0)(u^{0},v^{0}) in C([0,);H2×H2)C([0,\infty);H^{2}\times H^{2}).

We next prove (2.2). Integrating (3.21) over (0,)(0,\infty) with respect to tt, then using Lemma 3.1 and Lemma 3.2, we have

0ddtr(n1)/2u~r2dtC0r(n1)/2u~rL2(0,;L2)2(r(n1)/2v0L(0,;L2)2+(rn1v0)rL(0,;L2)2)+C0(rn1v0)rL2(0,;L2)2C0,\displaystyle\begin{split}\int_{0}^{\infty}&\frac{d}{dt}\|r^{(n-1)/2}\tilde{u}_{r}\|^{2}dt\\ \leq&C_{0}\|r^{(n-1)/2}\tilde{u}_{r}\|_{L^{2}(0,\infty;L^{2})}^{2}\left(\|r^{(n-1)/2}v^{0}\|_{L^{\infty}(0,\infty;L^{2})}^{2}+\|(r^{n-1}v^{0})_{r}\|_{L^{\infty}(0,\infty;L^{2})}^{2}\right)\\ &+C_{0}\|(r^{n-1}v^{0})_{r}\|_{L^{2}(0,\infty;L^{2})}^{2}\\ \leq&C_{0},\end{split}

which, along with (3.7) implies that r(n1)/2u~r2W1,1(0,).\|r^{(n-1)/2}\tilde{u}_{r}\|^{2}\in W^{1,1}(0,\infty). Hence, it follows that

limtu~rC0limtr(n1)/2u~r=0,\displaystyle\lim_{t\rightarrow\infty}\|\tilde{u}_{r}\|\leq C_{0}\lim_{t\rightarrow\infty}\|r^{(n-1)/2}\tilde{u}_{r}\|=0,

which, along with the Gagliardo-Nirenberg inequality (u0u¯)(t)L2C0(u0u¯)(t)L2(u0u¯)r(t)L2\|(u^{0}-\bar{u})(t)\|_{L^{\infty}}^{2}\leq C_{0}\|(u^{0}-\bar{u})(t)\|_{L^{2}}\|(u^{0}-\bar{u})_{r}(t)\|_{L^{2}} and (3.7), gives (2.2). Part (i) of Theorem 2.1 is thus proved.

We proceed to prove Part (ii). When u¯=0\bar{u}=0, for 0<T<0<T<\infty one can easily deduce the a priori estimates (2.3) by the standard energy method that bootstraps the regularity of the solution (u0,v0)(u^{0},v^{0}) from L2L^{2} to H2H^{2}. We omit this procedure for simplicity and refer readers to [26] for details. Then the existence of solution (u0,v0)(u^{0},v^{0}) follows from (2.3) and the fixed point theorem. The proof is finished.

\Box

4 Proof of Theorem 2.2 and Proposition 2.1.

Let (uε,vε)(u^{\varepsilon},v^{\varepsilon}) and (u0,v0)(u^{0},v^{0}) be the solutions of (1.15)-(1.16) corresponding to ε>0\varepsilon>0 and ε=0\varepsilon=0 respectively. Then the initial-boundary value problem for their differences h:=uεu0h:=u^{\varepsilon}-u^{0}, w:=vεv0w:=v^{\varepsilon}-v^{0} reads:

{ht=1rn1(rn1hr)r+1rn1(rn1hw)r+1rn1(rn1u0w)r+1rn1(rn1hv0)r,wt=ε1rn1(rn1wr)r2εwwr+hr+ε1rn1(rn1vr0)r2ε(wvr0+v0wr+v0vr0)εn1r2(w+v0),(r,t)(a,b)×(0,)(h,w)(r,0)=(0,0),h|r=a,b=0,w|r=a=v¯1v0(a,t),w|r=b=v¯2v0(b,t).\left\{\begin{split}&h_{t}=\frac{1}{r^{n-1}}(r^{n-1}h_{r})_{r}+\frac{1}{r^{n-1}}(r^{n-1}hw)_{r}+\frac{1}{r^{n-1}}(r^{n-1}u^{0}w)_{r}+\frac{1}{r^{n-1}}(r^{n-1}hv^{0})_{r},\\ &w_{t}=\varepsilon\frac{1}{r^{n-1}}(r^{n-1}w_{r})_{r}-2\varepsilon ww_{r}+h_{r}+\varepsilon\frac{1}{r^{n-1}}(r^{n-1}v^{0}_{r})_{r}-2\varepsilon(wv^{0}_{r}+v^{0}w_{r}+v^{0}v^{0}_{r})\\ &\quad\quad\,-\varepsilon\frac{n-1}{r^{2}}(w+v^{0}),\qquad\quad(r,t)\in(a,b)\times(0,\infty)\\ &(h,w)(r,0)=(0,0),\\ &h|_{r=a,b}=0,\,\,\,w|_{r=a}=\bar{v}_{1}-v^{0}(a,t),\,w|_{r=b}=\bar{v}_{2}-v^{0}(b,t).\end{split}\right. (4.1)

To prove Theorem 2.2 we shall invoke an elementary result (see Lemma 4.1) on an ordinary differential equation (ODE) and a series of lemmas on the a priori estimates for solutions of (4.1). In particular, the L2L^{2}-estimate for solution (h,w)(h,w) and higher regularity estimates for the solution component hh will be established in Lemma 4.2 - Lemma 4.5, and Lemma 4.6 will give a weighted L2L^{2}-estimate for the derivative of ww.

We proceed to prove the following Lemma, which gives an upper bound for the solution of an ODE involving a small parameter γ\gamma. It extends a result in [5, 51] with k=2k=2 to any integer k2k\geq 2.

Lemma 4.1.

Let k2k\geq 2 be an integer and 0<T<0<T<\infty. Let C0>1C_{0}>1 be a constant independent of TT and f1(t)f_{1}(t), f2(t)0f_{2}(t)\geq 0 be two continuous functions on [0,T][0,T]. Consider the ODE

{ddty(t)γf1(t)+f2(t)y(t)+C0[y2(t)++yk(t)],y(0)=0.\left\{\begin{aligned} &\frac{d}{dt}y(t)\leq\gamma f_{1}(t)+f_{2}(t)y(t)+C_{0}[y^{2}(t)+\cdots+y^{k}(t)],\\ &y(0)=0.\end{aligned}\right. (4.2)

If we set

γ0=min{[4(k1)]1(0Tf1(t)𝑑t)1,[8TG(k1)2]1(0Tf1(t)𝑑t)1},\gamma_{0}=\min\left\{[4(k-1)]^{-1}\left(\int_{0}^{T}f_{1}(t)\,dt\right)^{-1},\,\,\,\,[8TG(k-1)^{2}]^{-1}\left(\int_{0}^{T}f_{1}(t)\,dt\right)^{-1}\right\}, (4.3)

with G:=C0(e0Tf2(t)𝑑t)k1G:=C_{0}\left(e^{\int_{0}^{T}f_{2}(t)\,dt}\right)^{k-1}. Then for γ(0,γ0]\gamma\in(0,\gamma_{0}], any solution y(t)0y(t)\geq 0 of (4.2) satisfies

y(t)e0Tf2(t)𝑑tmin{3,32T(k1)G,  12(k1)γ0Tf1(t)𝑑t},t[0,T].y(t)\leq e^{\int_{0}^{T}f_{2}(t)\,dt}\cdot\min\left\{3,\,\,\frac{3}{2T(k-1)G},\,\,12(k-1)\gamma\int_{0}^{T}f_{1}(t)\,dt\right\},\quad t\in[0,T]. (4.4)
Proof.

Let U(t)=y(t)e0tf2(τ)𝑑τU(t)=y(t)e^{-\int_{0}^{t}f_{2}(\tau)\,d\tau}. Then (4.2) can be rewritten as

ddtU(t)γf1(t)e0tf2(τ)𝑑τ+C0(e0tf2(τ)𝑑τ)U2++C0(e0tf2(τ)𝑑τ)k1Uk.\displaystyle\frac{d}{dt}U(t)\leq\gamma f_{1}(t)e^{-\int_{0}^{t}f_{2}(\tau)\,d\tau}+C_{0}\left(e^{\int_{0}^{t}f_{2}(\tau)\,d\tau}\right)U^{2}+\cdots+C_{0}\left(e^{\int_{0}^{t}f_{2}(\tau)\,d\tau}\right)^{k-1}U^{k}.

Noting that e0tf2(τ)𝑑τ1e^{\int_{0}^{t}f_{2}(\tau)d\tau}\geq 1 thanks to f2(t)0f_{2}(t)\geq 0, we deduce that

{ddtU(t)γf1(t)+GU2(t)(1+U(t))k2,U(0)=0.\left\{\begin{aligned} &\frac{d}{dt}U(t)\leq\gamma f_{1}(t)+GU^{2}(t)(1+U(t))^{k-2},\\ &U(0)=0.\end{aligned}\right. (4.5)

For later use, we define

σ=min{G,14T2(k1)2G,  16(k1)2γ2G(0Tf1(t)𝑑t)2}.\sigma=\min\left\{G,\,\,\frac{1}{4T^{2}(k-1)^{2}G},\,\,16(k-1)^{2}\gamma^{2}G\left(\int_{0}^{T}f_{1}(t)\,dt\right)^{2}\right\}. (4.6)

Now dividing both sides of (4.5) by (1+Gσ)k\left(1+\sqrt{\frac{G}{\sigma}}\right)^{k}, it follows that

ddtU(t)(1+GσU(t))kγf1(t)+GU2(t)(1+GσU(t))2(1+U(t))k2(1+GσU(t))k2.\displaystyle\begin{split}\frac{\frac{d}{dt}U(t)}{\left(1+\sqrt{\frac{G}{\sigma}}U(t)\right)^{k}}\leq\gamma f_{1}(t)+\frac{GU^{2}(t)}{\left(1+\sqrt{\frac{G}{\sigma}}U(t)\right)^{2}}\cdot\frac{\left(1+U(t)\right)^{k-2}}{\left(1+\sqrt{\frac{G}{\sigma}}U(t)\right)^{k-2}}.\end{split}

Then noting σG\sigma\leq G due to its definition (4.6), we deduce from the above inequality that

ddtU(t)(1+GσU(t))k\displaystyle\frac{\frac{d}{dt}U(t)}{\left(1+\sqrt{\frac{G}{\sigma}}U(t)\right)^{k}}\leq γf1(t)+σ,\displaystyle\gamma f_{1}(t)+\sigma,

which integrated over (0,t)(0,t) with t(0,T]t\in(0,T], yields

σGk11(1+GσU(t))k1σGk1σTγ0Tf1(t)𝑑tσG2(k1)γ0Tf1(t)𝑑t,\begin{split}\frac{\sqrt{\frac{\sigma}{G}}}{k-1}\cdot\frac{1}{\left(1+\sqrt{\frac{G}{\sigma}}U(t)\right)^{k-1}}\geq&\frac{\sqrt{\frac{\sigma}{G}}}{k-1}-\sigma T-\gamma\int_{0}^{T}f_{1}(t)\,dt\\ \geq&\frac{\sqrt{\frac{\sigma}{G}}}{2(k-1)}-\gamma\int_{0}^{T}f_{1}(t)\,dt,\end{split} (4.7)

where we have used the fact σTσG2(k1)\sigma T\leq\frac{\sqrt{\frac{\sigma}{G}}}{2(k-1)}, thanks to the definition of σ\sigma. We shall prove that

γ0Tf1(t)𝑑tσG4(k1),\gamma\int_{0}^{T}f_{1}(t)\,dt\leq\frac{\sqrt{\frac{\sigma}{G}}}{4(k-1)}, (4.8)

of which the proof is split into three cases by the value of σ\sigma.
Case 1, when σ=G\sigma=G, it follows from the definition of γ0\gamma_{0} in (4.3) that

γ0Tf1(t)𝑑tγ00Tf1(t)𝑑t14(k1)=σG4(k1).\displaystyle\gamma\int_{0}^{T}f_{1}(t)\,dt\leq\gamma_{0}\int_{0}^{T}f_{1}(t)\,dt\leq\frac{1}{4(k-1)}=\frac{\sqrt{\frac{\sigma}{G}}}{4(k-1)}.

Case 2, when σ=14T2(k1)2G\sigma=\frac{1}{4T^{2}(k-1)^{2}G}, we have by using (4.3) again that

γ0Tf1(t)𝑑tγ00Tf1(t)𝑑t18TG(k1)2=σG4(k1).\displaystyle\gamma\int_{0}^{T}f_{1}(t)\,dt\leq\gamma_{0}\int_{0}^{T}f_{1}(t)\,dt\leq\frac{1}{8TG(k-1)^{2}}=\frac{\sqrt{\frac{\sigma}{G}}}{4(k-1)}.

Case 3, when σ=16(k1)2γ2G(0Tf1(t)𝑑t)2\sigma=16(k-1)^{2}\gamma^{2}G\left(\int_{0}^{T}f_{1}(t)\,dt\right)^{2}, one immediately get

γ0Tf1(t)𝑑t=σG4(k1).\displaystyle\gamma\int_{0}^{T}f_{1}(t)\,dt=\frac{\sqrt{\frac{\sigma}{G}}}{4(k-1)}.

Hence combining the above Case 1 - Case 3, we conclude that (4.8) holds true and it follows from (4.8) and (4.7) that

(1+GσU(t))k14,t[0,T]\displaystyle\left(1+\sqrt{\frac{G}{\sigma}}U(t)\right)^{k-1}\leq 4,\quad\,t\in[0,T]

thus

U(t)3σG,t[0,T]\displaystyle U(t)\leq 3\sqrt{\frac{\sigma}{G}},\quad\,t\in[0,T]

which, along with (4.6) and the definition of U(t)U(t), yields the desired estimate (4.4). The proof is finished.

In the sequel, for convenience we denote

E(t):=r(n1)/2h(t)2+r(n1)/2w(t)2+εr(n1)/2wr(t)2,\displaystyle E(t):=\|r^{(n-1)/2}h(t)\|^{2}+\|r^{(n-1)/2}w(t)\|^{2}+\varepsilon\|r^{(n-1)/2}w_{r}(t)\|^{2},
F(t):=u0(t)H22+v0(t)H2+v0(t)H22+v0(t)H24+|v¯1|2+|v¯2|2+1.F(t):=\|u^{0}(t)\|_{H^{2}}^{2}+\|v^{0}(t)\|_{H^{2}}+\|v^{0}(t)\|_{H^{2}}^{2}+\|v^{0}(t)\|_{H^{2}}^{4}+|\bar{v}_{1}|^{2}+|\bar{v}_{2}|^{2}+1. (4.9)

The following lemma gives the L2L^{2}-estimate for the solution (h,w)(h,w) of problem (4.1).

Lemma 4.2.

Let 0<t<0<t<\infty. Then there exists a constant C0C_{0} independent of ε\varepsilon and tt, such that

ddt(r(n1)/2h(t)2+r(n1)/2w(t)2)+32r(n1)/2hr(t)2+2εr(n1)/2wr(t)2+2(n1)εr(n3)/2w(t)2C0ε2F(t)+C0F(t)E(t)+C0E2(t)+C0E3(t)+2ε[rn1wrw]|ab.\begin{split}\frac{d}{dt}&(\|r^{(n-1)/2}h(t)\|^{2}+\|r^{(n-1)/2}w(t)\|^{2})+\frac{3}{2}\|r^{(n-1)/2}h_{r}(t)\|^{2}\\ &+2\varepsilon\|r^{(n-1)/2}w_{r}(t)\|^{2}+2(n-1)\varepsilon\|r^{(n-3)/2}w(t)\|^{2}\\ \leq&C_{0}\varepsilon^{2}F(t)+C_{0}F(t)E(t)+C_{0}E^{2}(t)+C_{0}E^{3}(t)+2\varepsilon[r^{n-1}w_{r}w]|_{a}^{b}.\end{split} (4.10)
Proof.

Testing the first equation of (4.1) with 2rn1h2r^{n-1}h in L2L^{2} and using integration by parts, we get

ddtr(n1)/2h2+2r(n1)/2hr2=2abrn1hwhr𝑑r2abrn1(u0w+hv0)hr𝑑r:=J1+J2.\displaystyle\begin{split}\frac{d}{dt}\|r^{(n-1)/2}h\|^{2}+2\|r^{(n-1)/2}h_{r}\|^{2}&=-2\int_{a}^{b}r^{n-1}hwh_{r}\,dr-2\int_{a}^{b}r^{n-1}(u^{0}w+hv^{0})h_{r}\,dr\\ &:=J_{1}+J_{2}.\end{split}

The estimate of J1J_{1} follows from (3.17) and (3.27):

J12r(n1)/2hrhLr(n1)/2wC0r(n1)/2hr32r(n1)/2h12r(n1)/2w18r(n1)/2hr2+C0r(n1)/2h2r(n1)/2w4.\displaystyle\begin{split}J_{1}\leq&2\|r^{(n-1)/2}h_{r}\|\|h\|_{L^{\infty}}\|r^{(n-1)/2}w\|\\ \leq&C_{0}\|r^{(n-1)/2}h_{r}\|^{\frac{3}{2}}\|r^{(n-1)/2}h\|^{\frac{1}{2}}\|r^{(n-1)/2}w\|\\ \leq&\frac{1}{8}\|r^{(n-1)/2}h_{r}\|^{2}+C_{0}\|r^{(n-1)/2}h\|^{2}\|r^{(n-1)/2}w\|^{4}.\end{split}

On the other hand, the Sobolev embedding inequality and Cauchy-Schwarz inequality entail that

J218r(n1)/2hr2+C0u0H12r(n1)/2w2+C0v0H12r(n1)/2h2.\displaystyle J_{2}\leq\frac{1}{8}\|r^{(n-1)/2}h_{r}\|^{2}+C_{0}\|u^{0}\|_{H^{1}}^{2}\|r^{(n-1)/2}w\|^{2}+C_{0}\|v^{0}\|_{H^{1}}^{2}\|r^{(n-1)/2}h\|^{2}.

Collecting the above estimates for J1J_{1} and J2J_{2}, we conclude that

ddtr(n1)/2h(t)2+74r(n1)/2hr(t)2C0F(t)E(t)+C0E3(t).\begin{split}\frac{d}{dt}\|r^{(n-1)/2}h(t)\|^{2}+\frac{7}{4}\|r^{(n-1)/2}h_{r}(t)\|^{2}\leq C_{0}F(t)E(t)+C_{0}E^{3}(t).\end{split} (4.11)

We proceed by taking the L2L^{2} inner product of the second equation of (4.1) with 2rn1w2r^{n-1}w and get

ddtr(n1)/2w2+2εr(n1)/2wr2+2(n1)εr(n3)/2w2= 2ε[rn1wrw]|ab4εabrn1(wwr+wvr0)w𝑑r+2ab(rn1hr+ε(rn1vr0)r)w𝑑r2εab(2rn1v0wr+2rn1v0vr0+(n1)rn3v0)w𝑑r:=2ε[rn1wrw]|ab+J3+J4+J5.\displaystyle\begin{split}\frac{d}{dt}&\|r^{(n-1)/2}w\|^{2}+2\varepsilon\|r^{(n-1)/2}w_{r}\|^{2}+2(n-1)\varepsilon\|r^{(n-3)/2}w\|^{2}\\ =&\,2\varepsilon[r^{n-1}w_{r}w]|_{a}^{b}-4\varepsilon\int_{a}^{b}r^{n-1}(ww_{r}+wv^{0}_{r})wdr\\ &+2\int_{a}^{b}\big{(}r^{n-1}h_{r}+\varepsilon(r^{n-1}v^{0}_{r})_{r}\big{)}wdr\\ &-2\varepsilon\int_{a}^{b}\big{(}2r^{n-1}v^{0}w_{r}+2r^{n-1}v^{0}v^{0}_{r}+(n-1)r^{n-3}v^{0}\big{)}wdr\\ :=\,&2\varepsilon[r^{n-1}w_{r}w]|_{a}^{b}+J_{3}+J_{4}+J_{5}.\end{split}

First Sobolev embedding inequality and (3.27) yield

J34εwLr(n1)/2wrr(n1)/2w+4εvr0Lr(n1)/2w2C0ε(r(n1)/2wr+r(n1)/2w)r(n1)/2wrr(n1)/2w+C0εv0H2r(n1)/2w2.\displaystyle\begin{split}J_{3}\leq&4\varepsilon\|w\|_{L^{\infty}}\|r^{(n-1)/2}w_{r}\|\|r^{(n-1)/2}w\|+4\varepsilon\|v^{0}_{r}\|_{L^{\infty}}\|r^{(n-1)/2}w\|^{2}\\ \leq&C_{0}\varepsilon\big{(}\|r^{(n-1)/2}w_{r}\|+\|r^{(n-1)/2}w\|\big{)}\|r^{(n-1)/2}w_{r}\|\|r^{(n-1)/2}w\|+C_{0}\varepsilon\|v^{0}\|_{H^{2}}\|r^{(n-1)/2}w\|^{2}.\end{split}

It follows from Cauchy-Schwarz inequality and (3.27) that

J414r(n1)/2hr2+C0r(n1)/2w2+C0ε2v0H22.\displaystyle J_{4}\leq\frac{1}{4}\|r^{(n-1)/2}h_{r}\|^{2}+C_{0}\|r^{(n-1)/2}w\|^{2}+C_{0}\varepsilon^{2}\|v^{0}\|_{H^{2}}^{2}.

Moreover Sobolev embedding inequality, Cauchy-Schwarz inequality and (3.27) lead to

J5εr(n1)/2wr2+C0εv0H12r(n1)/2w2+r(n1)/2w2+C0ε2v0H24+C0ε2v02.\displaystyle J_{5}\leq\varepsilon\|r^{(n-1)/2}w_{r}\|^{2}+C_{0}\varepsilon\|v^{0}\|_{H^{1}}^{2}\|r^{(n-1)/2}w\|^{2}+\|r^{(n-1)/2}w\|^{2}+C_{0}\varepsilon^{2}\|v^{0}\|_{H^{2}}^{4}+C_{0}\varepsilon^{2}\|v^{0}\|^{2}.

Collecting the above estimates for J3J_{3}-J5J_{5} and recalling that 0<ε<10<\varepsilon<1, we end up with

ddtr(n1)/2w(t)2+2εr(n1)/2wr(t)2+2(n1)εr(n3)/2w(t)22ε[rn1wrw]|ab+14r(n1)/2hr(t)2+C0ε2F(t)+C0F(t)E(t)+C0E2(t),\displaystyle\begin{split}\frac{d}{dt}&\|r^{(n-1)/2}w(t)\|^{2}+2\varepsilon\|r^{(n-1)/2}w_{r}(t)\|^{2}+2(n-1)\varepsilon\|r^{(n-3)/2}w(t)\|^{2}\\ \leq&2\varepsilon[r^{n-1}w_{r}w]|_{a}^{b}+\frac{1}{4}\|r^{(n-1)/2}h_{r}(t)\|^{2}+C_{0}\varepsilon^{2}F(t)+C_{0}F(t)E(t)+C_{0}E^{2}(t),\end{split}

which, adding to (4.11) gives the desired estimate (4.10). The proof is completed.

We turn to estimate the derivative of ww and the boundary term in (4.10).

Lemma 4.3.

Let 0<t<0<t<\infty. Then there exists a constant C0C_{0} independent of ε\varepsilon and tt, such that

ddt(εr(n1)/2wr(t)2)+12r(n1)/2wt(t)2C0ε2F(t)+C0F(t)E(t)+C0E2(t)+r(n1)/2hr(t)2+2ε[rn1wrwt]|ab\begin{split}\frac{d}{dt}&\big{(}\varepsilon\|r^{(n-1)/2}w_{r}(t)\|^{2}\big{)}+\frac{1}{2}\|r^{(n-1)/2}w_{t}(t)\|^{2}\\ &\leq C_{0}\varepsilon^{2}F(t)+C_{0}F(t)E(t)+C_{0}E^{2}(t)+\|r^{(n-1)/2}h_{r}(t)\|^{2}+2\varepsilon[r^{n-1}w_{r}w_{t}]|_{a}^{b}\end{split} (4.12)

and

2ε[rn1wrw]|ab+2ε[rn1wrwt]|abC0ε1/2F(t)+C0F(t)E(t)+C0E2(t)+14r(n1)/2wt(t)2+14r(n1)/2hr(t)2.\begin{split}2\varepsilon&[r^{n-1}w_{r}w]|_{a}^{b}+2\varepsilon[r^{n-1}w_{r}w_{t}]|_{a}^{b}\\ &\leq C_{0}\varepsilon^{1/2}F(t)+C_{0}F(t)E(t)+C_{0}E^{2}(t)+\frac{1}{4}\|r^{(n-1)/2}w_{t}(t)\|^{2}+\frac{1}{4}\|r^{(n-1)/2}h_{r}(t)\|^{2}.\end{split} (4.13)
Proof.

Taking the L2L^{2} inner product of the second equation of (4.1) with 2rn1wt2r^{n-1}w_{t}, then using integration by parts we have

ddtεr(n1)/2wr(t)2+2r(n1)/2wt(t)2=2ε[rn1wrwt]|ab4εabrn1wwrwt𝑑r4εabrn1(wvr0+v0wr+v0vr0)wt𝑑r+2ab[rn1hr+ε(rn1vr0)rε(n1)rn3wε(n1)rn3v0]wt𝑑r:=2ε[rn1wrwt]|ab+J6+J7+J8.\displaystyle\begin{split}\frac{d}{dt}&\varepsilon\|r^{(n-1)/2}w_{r}(t)\|^{2}+2\|r^{(n-1)/2}w_{t}(t)\|^{2}\\ =&2\varepsilon[r^{n-1}w_{r}w_{t}]|_{a}^{b}-4\varepsilon\int_{a}^{b}r^{n-1}ww_{r}w_{t}dr-4\varepsilon\int_{a}^{b}r^{n-1}(wv^{0}_{r}+v^{0}w_{r}+v^{0}v^{0}_{r})w_{t}dr\\ &+2\int_{a}^{b}\big{[}r^{n-1}h_{r}+\varepsilon(r^{n-1}v^{0}_{r})_{r}-\varepsilon(n-1)r^{n-3}w-\varepsilon(n-1)r^{n-3}v_{0}\big{]}w_{t}dr\\ :=&2\varepsilon[r^{n-1}w_{r}w_{t}]|_{a}^{b}+J_{6}+J_{7}+J_{8}.\end{split}

We first employ Sobolev embedding inequality and (3.27) to deduce that

J6C0εwH1r(n1)/2wrr(n1)/2wt18r(n1)/2wt2+C0ε2(r(n1)/2wr2+r(n1)/2w2)r(n1)/2wr2\displaystyle\begin{split}J_{6}\leq&C_{0}\varepsilon\|w\|_{H^{1}}\|r^{(n-1)/2}w_{r}\|\|r^{(n-1)/2}w_{t}\|\\ \leq&\frac{1}{8}\|r^{(n-1)/2}w_{t}\|^{2}+C_{0}\varepsilon^{2}\big{(}\|r^{(n-1)/2}w_{r}\|^{2}+\|r^{(n-1)/2}w\|^{2}\big{)}\|r^{(n-1)/2}w_{r}\|^{2}\end{split}

and that

J718r(n1)/2wt2+C0ε2(v0H22r(n1)/2w2+v0H12r(n1)/2wr2+v0H14).\displaystyle J_{7}\leq\frac{1}{8}\|r^{(n-1)/2}w_{t}\|^{2}+C_{0}\varepsilon^{2}\big{(}\|v^{0}\|_{H^{2}}^{2}\|r^{(n-1)/2}w\|^{2}+\|v^{0}\|_{H^{1}}^{2}\|r^{(n-1)/2}w_{r}\|^{2}+\|v^{0}\|_{H^{1}}^{4}\big{)}.

Moreover Cauchy-Schwarz inequality and (3.27) entail that

J854r(n1)/2wt2+r(n1)/2hr2+C0ε2(v0H22+r(n1)/2w2).\displaystyle J_{8}\leq\frac{5}{4}\|r^{(n-1)/2}w_{t}\|^{2}+\|r^{(n-1)/2}h_{r}\|^{2}+C_{0}\varepsilon^{2}(\|v^{0}\|_{H^{2}}^{2}+\|r^{(n-1)/2}w\|^{2}).

Then (4.12) follows from the above estimates on J6J_{6}-J8J_{8}. It remains to prove (4.13). By the definition of ww and Gagliardo-Nirenberg interpolation inequality, one deduces that

2ε[rn1wrw]|abC0εwrL(|v¯1|+|v¯2|+v0L)C0ε(wr12wrr12+wr)(|v¯1|+|v¯2|+v0H1)ηε2wrr2+C0(1+1/η)εwr2+C0(ε1/2+ε)(|v¯1|+|v¯2|+v0H1)2,\begin{split}2\varepsilon[r^{n-1}w_{r}w]|_{a}^{b}\leq&C_{0}\varepsilon\|w_{r}\|_{L^{\infty}}(|\bar{v}_{1}|+|\bar{v}_{2}|+\|v^{0}\|_{L^{\infty}})\\ \leq&C_{0}\varepsilon(\|w_{r}\|^{\frac{1}{2}}\|w_{rr}\|^{\frac{1}{2}}+\|w_{r}\|)(|\bar{v}_{1}|+|\bar{v}_{2}|+\|v^{0}\|_{H^{1}})\\ \leq&\eta\varepsilon^{2}\|w_{rr}\|^{2}+C_{0}(1+1/\eta)\varepsilon\|w_{r}\|^{2}+C_{0}(\varepsilon^{1/2}+\varepsilon)(|\bar{v}_{1}|+|\bar{v}_{2}|+\|v^{0}\|_{H^{1}})^{2},\end{split} (4.14)

where η>0\eta>0 is a small constant to be determined. By a similar argument as deriving (4.14) and the second equation of (1.15) with ε=0\varepsilon=0, we further get

2ε[rn1wrwt]|abηε2wrr2+C0(1+1/η)εwr2+C0(ε1/2+ε)vt0H12ηε2wrr2+C0(1+1/η)εwr2+C0(ε1/2+ε)u0H22.\begin{split}2\varepsilon[r^{n-1}w_{r}w_{t}]|_{a}^{b}\leq&\eta\varepsilon^{2}\|w_{rr}\|^{2}+C_{0}(1+1/\eta)\varepsilon\|w_{r}\|^{2}+C_{0}(\varepsilon^{1/2}+\varepsilon)\|v^{0}_{t}\|_{H^{1}}^{2}\\ \leq&\eta\varepsilon^{2}\|w_{rr}\|^{2}+C_{0}(1+1/\eta)\varepsilon\|w_{r}\|^{2}+C_{0}(\varepsilon^{1/2}+\varepsilon)\|u^{0}\|_{H^{2}}^{2}.\end{split} (4.15)

To bound the term wrr2\|w_{rr}\|^{2} in (4.14) and (4.15), we use the second equation of (4.1), Sobolev embedding inequality and (3.27) and derive

ε2wrr2C1(r(n1)/2wt2+r(n1)/2hr2+ε2r(n1)/2wr2+ε2r(n1)/2w2v0H22+ε2r(n1)/2wr4+ε2r(n1)/2w2r(n1)/2wr2+ε2v0H22+ε2r(n1)/2wr2v0H12+ε2v0H14+ε2r(n1)/2w2),\begin{split}\varepsilon^{2}\|w_{rr}\|^{2}\leq C_{1}&\big{(}\|r^{(n-1)/2}w_{t}\|^{2}+\|r^{(n-1)/2}h_{r}\|^{2}+\varepsilon^{2}\|r^{(n-1)/2}w_{r}\|^{2}+\varepsilon^{2}\|r^{(n-1)/2}w\|^{2}\|v^{0}\|^{2}_{H^{2}}\\ &+\varepsilon^{2}\|r^{(n-1)/2}w_{r}\|^{4}+\varepsilon^{2}\|r^{(n-1)/2}w\|^{2}\|r^{(n-1)/2}w_{r}\|^{2}+\varepsilon^{2}\|v^{0}\|^{2}_{H^{2}}\\ &+\varepsilon^{2}\|r^{(n-1)/2}w_{r}\|^{2}\|v^{0}\|^{2}_{H^{1}}+\varepsilon^{2}\|v^{0}\|^{4}_{H^{1}}+\varepsilon^{2}\|r^{(n-1)/2}w\|^{2}\big{)},\end{split} (4.16)

where we have used the notation C1C_{1} to distinguish it from the constant C0C_{0} in (4.14) and (4.15). Finally feeding (4.14) and (4.15) on (4.16) then adding the results, we obtain (4.13) by taking η\eta small enough such that C1η<18C_{1}\eta<\frac{1}{8} and by using 0<ε<10<\varepsilon<1. The proof is completed.

We next apply Lemma 4.1 to the combination of Lemma 4.2 and Lemma 4.3 to obtain the following result.

Lemma 4.4.

Let 0<T<0<T<\infty and ε(0,ε0]\varepsilon\in(0,\varepsilon_{0}] with ε0\varepsilon_{0} defined in Theorem 2.2. Then there exists a constant CC independent of ε\varepsilon, depending on TT such that

hL(0,T;L2)2+wL(0,T;L2)2+εwrL(0,T;L2)2+hrL2(0,T;L2)2+wtL2(0,T;L2)2Cε12\begin{split}\|&h\|_{L^{\infty}(0,T;L^{2})}^{2}+\|w\|_{L^{\infty}(0,T;L^{2})}^{2}\\ &+\varepsilon\|w_{r}\|_{L^{\infty}(0,T;L^{2})}^{2}+\|h_{r}\|_{L^{2}(0,T;L^{2})}^{2}+\|w_{t}\|_{L^{2}(0,T;L^{2})}^{2}\leq C\varepsilon^{\frac{1}{2}}\end{split} (4.17)

and

wrrL2(0,T;L2)Cε3/4.\|w_{rr}\|_{L^{2}(0,T;L^{2})}\leq C\varepsilon^{-3/4}. (4.18)
Proof.

We first add (4.12) and (4.13) to (4.10) and find

ddtE(t)+14r(n1)/2hr(t)2+2(n1)εr(n3)/2w(t)2+14r(n1)/2wt(t)2C0ε12F(t)+C0F(t)E(t)+C0E2(t)+C0E3(t),\begin{split}\frac{d}{dt}&E(t)+\frac{1}{4}\|r^{(n-1)/2}h_{r}(t)\|^{2}+2(n-1)\varepsilon\|r^{(n-3)/2}w(t)\|^{2}+\frac{1}{4}\|r^{(n-1)/2}w_{t}(t)\|^{2}\\ &\leq C_{0}\varepsilon^{\frac{1}{2}}F(t)+C_{0}F(t)E(t)+C_{0}E^{2}(t)+C_{0}E^{3}(t),\end{split} (4.19)

where 0<ε<10<\varepsilon<1 has been used. Then we apply lemma 4.1 to (4.19) by taking k=3,γ=ε1/2k=3,\,\gamma=\varepsilon^{1/2} and f1(t)=f2(t)=C0F(t)f_{1}(t)=f_{2}(t)=C_{0}F(t) to conclude for ε(0,ε0]\varepsilon\in(0,\varepsilon_{0}] that

hL(0,T;L2)2+wL(0,T;L2)2+εwrL(0,T;L2)2(C0eC00TF(t)𝑑t0TF(t)𝑑t)ε12,\|h\|_{L^{\infty}(0,T;L^{2})}^{2}+\|w\|_{L^{\infty}(0,T;L^{2})}^{2}+\varepsilon\|w_{r}\|_{L^{\infty}(0,T;L^{2})}^{2}\leq\big{(}C_{0}e^{C_{0}\int_{0}^{T}F(t)\,dt}\int_{0}^{T}F(t)dt\big{)}\varepsilon^{\frac{1}{2}}, (4.20)

where (3.27) has been used. Then we integrate (4.19) over (0,T) and applying (4.20) to the result to deduce that

hL(0,T;L2)2+wL(0,T;L2)2+εwrL(0,T;L2)2+hrL2(0,T;L2)2+wtL2(0,T;L2)2Cε12,\displaystyle\|h\|_{L^{\infty}(0,T;L^{2})}^{2}+\|w\|_{L^{\infty}(0,T;L^{2})}^{2}+\varepsilon\|w_{r}\|_{L^{\infty}(0,T;L^{2})}^{2}+\|h_{r}\|_{L^{2}(0,T;L^{2})}^{2}+\|w_{t}\|_{L^{2}(0,T;L^{2})}^{2}\leq C\varepsilon^{\frac{1}{2}},

where the constant CC depending on TT and 0TF(t)𝑑t\int_{0}^{T}F(t)dt is finite thanks to Theorem 2.1. We thus derive (4.17) and proceed to prove (4.18). Indeed, it follows from the second equation of (4.1), Sobolev embedding inequality and (4.17) that

εwrrL2(0,T;L2)C(wtL2(0,T;L2)+εwrL2(0,T;L2)+εwL(0,T;H1)wrL2(0,T;L2))+C(hrL2(0,T;L2)+εv0L2(0,T;H2)+εwrL2(0,T;L2)v0L(0,T;H2))+C(εv0L2(0,T;H2)+εwL2(0,T;L2)+εv0L2(0,T;L2))Cε1/4.\displaystyle\begin{aligned} \varepsilon\|w_{rr}\|_{L^{2}(0,T;L^{2})}\leq&C\left(\|w_{t}\|_{L^{2}(0,T;L^{2})}+\varepsilon\|w_{r}\|_{L^{2}(0,T;L^{2})}+\varepsilon\|w\|_{L^{\infty}(0,T;H^{1})}\|w_{r}\|_{L^{2}(0,T;L^{2})}\right)\\ &+C\left(\|h_{r}\|_{L^{2}(0,T;L^{2})}+\varepsilon\|v^{0}\|_{L^{2}(0,T;H^{2})}+\varepsilon\|w_{r}\|_{L^{2}(0,T;L^{2})}\|v^{0}\|_{L^{\infty}(0,T;H^{2})}\right)\\ &+C\left(\varepsilon\|v^{0}\|_{L^{2}(0,T;H^{2})}+\varepsilon\|w\|_{L^{2}(0,T;L^{2})}+\varepsilon\|v^{0}\|_{L^{2}(0,T;L^{2})}\right)\\ \leq&C\varepsilon^{1/4}.\end{aligned}

We thus derive (4.18) and the proof is completed.

Higher regularity estimates for the solution component hh is given in the following lemma.

Lemma 4.5.

Suppose 0<T<0<T<\infty and ε(0,ε0]\varepsilon\in(0,\varepsilon_{0}]. Then there is a constant CC independent of ε\varepsilon, depending on TT such that

hrL(0,T;L2)2+htL(0,T;L2)2+hrtL2(0,T;L2)2Cε1/2.\|h_{r}\|_{L^{\infty}(0,T;L^{2})}^{2}+\|h_{t}\|_{L^{\infty}(0,T;L^{2})}^{2}+\|h_{rt}\|_{L^{2}(0,T;L^{2})}^{2}\leq C\varepsilon^{1/2}. (4.21)
Proof.

We first take the L2L^{2} inner product of the first equation of (4.1) with 2rn1ht2r^{n-1}h_{t} and use integration by parts to get

ddtr(n1)/2hr2+2r(n1)/2ht2=2abrn1(hw+u0w+hv0)hrt𝑑r12r(n1)/2hrt2+C0(hr2w2+u0H12w2+hr2v02).\begin{split}\frac{d}{dt}&\|r^{(n-1)/2}h_{r}\|^{2}+2\|r^{(n-1)/2}h_{t}\|^{2}\\ =&-2\int_{a}^{b}r^{n-1}(hw+u^{0}w+hv^{0})h_{rt}dr\\ \leq&\,\frac{1}{2}\|r^{(n-1)/2}h_{rt}\|^{2}+C_{0}(\|h_{r}\|^{2}\|w\|^{2}+\|u^{0}\|_{H^{1}}^{2}\|w\|^{2}+\|h_{r}\|^{2}\|v^{0}\|^{2}).\end{split} (4.22)

Then differentiating the first equation of (4.1) with respect to tt and multiplying the resulting equation with 2rn1ht2r^{n-1}h_{t} in L2L^{2}, we derive

ddtr(n1)/2ht2+2r(n1)/2hrt2=2abrn1htwhrt𝑑r2abrn1(hwt+ut0w+u0wt+htv0+hvt0)hrt𝑑r:=K1+K2.\displaystyle\begin{split}\frac{d}{dt}&\|r^{(n-1)/2}h_{t}\|^{2}+2\|r^{(n-1)/2}h_{rt}\|^{2}\\ =&-2\int_{a}^{b}r^{n-1}h_{t}wh_{rt}dr-2\int_{a}^{b}r^{n-1}\big{(}hw_{t}+u^{0}_{t}w+u^{0}w_{t}+h_{t}v^{0}+hv^{0}_{t}\big{)}h_{rt}dr\\ :=&K_{1}+K_{2}.\end{split}

The estimate for K1K_{1} follows from (3.17) and (3.27):

K1C0r(n1)/2hrthtLwC0r(n1)/2hrt3/2r(n1)/2ht1/2w14r(n1)/2hrt2+C0r(n1)/2ht2w4.\displaystyle\begin{split}K_{1}\leq&C_{0}\|r^{(n-1)/2}h_{rt}\|\|h_{t}\|_{L^{\infty}}\|w\|\\ \leq&C_{0}\|r^{(n-1)/2}h_{rt}\|^{3/2}\|r^{(n-1)/2}h_{t}\|^{1/2}\|w\|\\ \leq&\frac{1}{4}\|r^{(n-1)/2}h_{rt}\|^{2}+C_{0}\|r^{(n-1)/2}h_{t}\|^{2}\|w\|^{4}.\end{split}

Sobolev embedding inequality and (3.27) entail that

K214r(n1)/2hrt2+C0(r(n1)/2hr2wt2+ut0H12w2)+C0(u0H12wt2+v0H12r(n1)/2ht2+vt0H12h2).\displaystyle\begin{split}K_{2}\leq&\frac{1}{4}\|r^{(n-1)/2}h_{rt}\|^{2}+C_{0}\Big{(}\|r^{(n-1)/2}h_{r}\|^{2}\|w_{t}\|^{2}+\|u^{0}_{t}\|_{H^{1}}^{2}\|w\|^{2}\Big{)}\\ &+C_{0}\Big{(}\|u^{0}\|_{H^{1}}^{2}\|w_{t}\|^{2}+\|v^{0}\|_{H^{1}}^{2}\|r^{(n-1)/2}h_{t}\|^{2}+\|v^{0}_{t}\|_{H^{1}}^{2}\|h\|^{2}\Big{)}.\end{split}

Then collecting the above estimates for K1K_{1} and K2K_{2}, one derives

ddtr(n1)/2ht2+32r(n1)/2hrt2C0(w4+v0H12)r(n1)/2ht2+C0wt2r(n1)/2hr2+C0u0H12wt2+C0(u0H32+u0H22v0H22)w2+C0u0H22h2,\begin{split}\frac{d}{dt}&\|r^{(n-1)/2}h_{t}\|^{2}+\frac{3}{2}\|r^{(n-1)/2}h_{rt}\|^{2}\\ \leq&C_{0}\left(\|w\|^{4}+\|v^{0}\|_{H^{1}}^{2}\right)\|r^{(n-1)/2}h_{t}\|^{2}+C_{0}\|w_{t}\|^{2}\|r^{(n-1)/2}h_{r}\|^{2}\\ &+C_{0}\|u^{0}\|_{H^{1}}^{2}\|w_{t}\|^{2}+C_{0}\left(\|u^{0}\|_{H^{3}}^{2}+\|u^{0}\|_{H^{2}}^{2}\|v^{0}\|_{H^{2}}^{2}\right)\|w\|^{2}+C_{0}\|u^{0}\|_{H^{2}}^{2}\|h\|^{2},\end{split} (4.23)

where we have used inequalities ut0H12C0(u0H32+u0H22v0H22)\|u^{0}_{t}\|_{H^{1}}^{2}\leq C_{0}(\|u^{0}\|_{H^{3}}^{2}+\|u^{0}\|_{H^{2}}^{2}\|v^{0}\|_{H^{2}}^{2}) and vt0H12C0u0H22\|v^{0}_{t}\|_{H^{1}}^{2}\leq C_{0}\|u^{0}\|_{H^{2}}^{2}, thanks to the first and second equations of (3.5). Finally by adding (4.23) to (4.22) and applying Gronwall’s inequality to the result, and then using (4.17) we obtain (4.21). The proof is completed.

We turn to establish a weighted L2L^{2}-estimate (enlightened by [17]) on the derivative of ww.

Lemma 4.6.

For 0<T<0<T<\infty and ε(0,ε0]\varepsilon\in(0,\varepsilon_{0}], there is a constant CC independent of ε\varepsilon, depending on TT such that

(ra)(rb)wrL(0,T;L2)2+ε(ra)(rb)wrrL2(0,T;L2)2Cε1/2.\displaystyle\|(r-a)(r-b)w_{r}\|^{2}_{L^{\infty}(0,T;L^{2})}+\varepsilon\|(r-a)(r-b)w_{rr}\|^{2}_{L^{2}(0,T;L^{2})}\leq C\varepsilon^{1/2}.
Proof.

Taking the L2L^{2} inner product of the second equation of (4.1) with 2(ra)2(rb)2wrr-2(r-a)^{2}(r-b)^{2}w_{rr} and using integration by parts, one gets

ddt(ra)(rb)wr2+2ε(ra)(rb)wrr2=2εab(ra)2(rb)2wrr[1rn1(rn1vr0)rn1r2(w+v0)]𝑑r+4εab(ra)2(rb)2wrr(wvr0+v0wr+v0vr0)𝑑r4ab(2rab)(ra)(rb)wrwt𝑑r2ab(ra)2(rb)2wrrhr𝑑r2εab(ra)2(rb)2wrr(n1rwr2wwr)𝑑r:=i=37Ki.\begin{split}\frac{d}{dt}&\|(r-a)(r-b)w_{r}\|^{2}+2\varepsilon\|(r-a)(r-b)w_{rr}\|^{2}\\ =\,&-2\varepsilon\int_{a}^{b}(r-a)^{2}(r-b)^{2}w_{rr}\Big{[}\frac{1}{r^{n-1}}(r^{n-1}v^{0}_{r})_{r}-\frac{n-1}{r^{2}}(w+v_{0})\Big{]}dr\\ &+4\varepsilon\int_{a}^{b}(r-a)^{2}(r-b)^{2}w_{rr}(wv^{0}_{r}+v^{0}w_{r}+v^{0}v^{0}_{r})dr\\ &-4\int_{a}^{b}(2r-a-b)(r-a)(r-b)w_{r}w_{t}dr\\ &-2\int_{a}^{b}(r-a)^{2}(r-b)^{2}w_{rr}h_{r}dr\\ &-2\varepsilon\int_{a}^{b}(r-a)^{2}(r-b)^{2}w_{rr}\Big{(}\frac{n-1}{r}w_{r}-2ww_{r}\Big{)}dr\\ :=&\sum_{i=3}^{7}K_{i}.\end{split} (4.24)

We next estimate K3K_{3}-K7K_{7}. Indeed Cauchy-Schwarz inequality, Sobolev embedding inequality and (3.27) yield

K3+K418ε(ra)(rb)wrr2+C0ε(v0H22+w2+v0H22w2+v0H12wr2+v0H14)\displaystyle\begin{split}K_{3}+K_{4}\leq&\frac{1}{8}\varepsilon\|(r-a)(r-b)w_{rr}\|^{2}\\ &+C_{0}\varepsilon(\|v^{0}\|_{H^{2}}^{2}+\|w\|^{2}+\|v^{0}\|_{H^{2}}^{2}\|w\|^{2}+\|v^{0}\|_{H^{1}}^{2}\|w_{r}\|^{2}+\|v^{0}\|_{H^{1}}^{4})\end{split}

and

K5+K718ε(ra)(rb)wrr2+C0(1+ε+εw2+εwr2)(ra)(rb)wr2+C0wt2.\displaystyle K_{5}+K_{7}\leq\frac{1}{8}\varepsilon\|(r-a)(r-b)w_{rr}\|^{2}+C_{0}(1+\varepsilon+\varepsilon\|w\|^{2}+\varepsilon\|w_{r}\|^{2})\|(r-a)(r-b)w_{r}\|^{2}+C_{0}\|w_{t}\|^{2}.

For the term K6K_{6}, we use integration by parts and the first equation of (4.1) to get

K6=4ab(2rab)(ra)(rb)wrhr𝑑r+2ab(ra)2(rb)2wrhrr𝑑r=4ab(2rab)(ra)(rb)wrhr𝑑r+2ab(ra)2(rb)2wr(htn1rhr)𝑑r2ab(ra)2(rb)2wr(hw+u0w)r𝑑r2ab(ra)2(rb)2wr[(hv0)r+n1r(hw+u0w+hv0)]𝑑r:=R1+R2+R3+R4.\displaystyle\begin{split}K_{6}=\,&4\int_{a}^{b}(2r-a-b)(r-a)(r-b)w_{r}h_{r}dr+2\int_{a}^{b}(r-a)^{2}(r-b)^{2}w_{r}h_{rr}dr\\ =\,&4\int_{a}^{b}(2r-a-b)(r-a)(r-b)w_{r}h_{r}dr+2\int_{a}^{b}(r-a)^{2}(r-b)^{2}w_{r}\Big{(}h_{t}-\frac{n-1}{r}h_{r}\Big{)}dr\\ &-2\int_{a}^{b}(r-a)^{2}(r-b)^{2}w_{r}(hw+u^{0}w)_{r}dr\\ &-2\int_{a}^{b}(r-a)^{2}(r-b)^{2}w_{r}\Big{[}(hv^{0})_{r}+\frac{n-1}{r}(hw+u^{0}w+hv^{0})\Big{]}dr\\ :=&R_{1}+R_{2}+R_{3}+R_{4}.\end{split}

We proceed to estimate R1R_{1}-R4R_{4}. First it follows from Cauchy-Schwarz inequality that

R1+R2(ra)(rb)wr2+C0(hr2+ht2).\displaystyle R_{1}+R_{2}\leq\|(r-a)(r-b)w_{r}\|^{2}+C_{0}(\|h_{r}\|^{2}+\|h_{t}\|^{2}).

Moveover, we use Cauchy-Schwarz inequality and apply (3.17) to hh and (ra)(rb)w(r-a)(r-b)w to derive

R32(hL+u0L)(ra)(rb)wr2+2(hr+ur0)(ra)(rb)wr(ra)(rb)wLC0(hr+u0H1)(ra)(rb)wr2+C0(hr+ur0)(ra)(rb)wr[(ra)(rb)w]rC0(hr+u0H1+1)2(ra)(rb)wr2+C0w2.\displaystyle\begin{split}R_{3}\leq&2(\|h\|_{L^{\infty}}+\|u^{0}\|_{L^{\infty}})\|(r-a)(r-b)w_{r}\|^{2}\\ &+2(\|h_{r}\|+\|u^{0}_{r}\|)\|(r-a)(r-b)w_{r}\|\|(r-a)(r-b)w\|_{L^{\infty}}\\ \leq&C_{0}(\|h_{r}\|+\|u^{0}\|_{H^{1}})\|(r-a)(r-b)w_{r}\|^{2}\\ &+C_{0}(\|h_{r}\|+\|u^{0}_{r}\|)\|(r-a)(r-b)w_{r}\|\|[(r-a)(r-b)w]_{r}\|\\ \leq&C_{0}(\|h_{r}\|+\|u^{0}\|_{H^{1}}+1)^{2}\|(r-a)(r-b)w_{r}\|^{2}+C_{0}\|w\|^{2}.\end{split}

The estimate for R4R_{4} follows from the Sobolev embedding inequality, (3.17) and Cauchy-Schwarz inequality:

R4C0(ra)(rb)wr(hw+hv0H1+u0w)(ra)(rb)wr2+C0(hr2w2+hr2v0H12+u0H12w2).\displaystyle\begin{split}R_{4}\leq&C_{0}\|(r-a)(r-b)w_{r}\|(\|hw\|+\|hv^{0}\|_{H^{1}}+\|u^{0}w\|)\\ \leq&\|(r-a)(r-b)w_{r}\|^{2}+C_{0}(\|h_{r}\|^{2}\|w\|^{2}+\|h_{r}\|^{2}\|v^{0}\|_{H^{1}}^{2}+\|u^{0}\|_{H^{1}}^{2}\|w\|^{2}).\end{split}

We thus conclude from the above estimates for R1R_{1}-R4R_{4} that

K6C0(hr+u0H1+1)2(ra)(rb)wr2+C0(hr2+ht2+hr2w2+u0H12w2+hr2v0H12+w2).\displaystyle\begin{split}K_{6}\leq&C_{0}(\|h_{r}\|+\|u^{0}\|_{H^{1}}+1)^{2}\|(r-a)(r-b)w_{r}\|^{2}\\ &+C_{0}(\|h_{r}\|^{2}+\|h_{t}\|^{2}+\|h_{r}\|^{2}\|w\|^{2}+\|u^{0}\|_{H^{1}}^{2}\|w\|^{2}+\|h_{r}\|^{2}\|v^{0}\|_{H^{1}}^{2}+\|w\|^{2}).\end{split}

Substituting the above estimates for K3K_{3}-K7K_{7} into (4.24), then applying Gronwall’s inequality, Lemma 4.4, Lemma 4.5 and Theorem 2.1 to the result, we obtain the desired estimate and the proof is finished.

We next prove Theorem 2.2 by the results derived in Lemma 4.4 - Lemma 4.6.
 
Proof of Theorem 2.2. By Lemma 4.4, Lemma 4.5 and Sobolev embedding inequality, we deduce that

uεu0L(0,T;C[a,b])C0uεu0L(0,T;H1)Cε1/4,\displaystyle\|u^{\varepsilon}-u^{0}\|_{L^{\infty}(0,T;C[a,b])}\leq C_{0}\|u^{\varepsilon}-u^{0}\|_{L^{\infty}(0,T;H^{1})}\leq C\varepsilon^{1/4},

which gives (2.5). Clearly δ24(ba)2(ra)2(rb)2\delta^{2}\leq\frac{4}{(b-a)^{2}}(r-a)^{2}(r-b)^{2} holds for δ<b+a2\delta<\frac{b+a}{2} and r(a,b)r\in(a,b), thus it follows from Lemma 4.6 that

δ2a+δbδwr2(r,t)𝑑r4(ba)2a+δbδ(ra)2(rb)2wr2(r,t)𝑑rCε1/2,t[0,T]\displaystyle\delta^{2}\int_{a+\delta}^{b-\delta}w^{2}_{r}(r,t)\,dr\leq\frac{4}{(b-a)^{2}}\int_{a+\delta}^{b-\delta}(r-a)^{2}(r-b)^{2}w^{2}_{r}(r,t)\,dr\leq C\varepsilon^{1/2},\quad t\in[0,T]

which, along with Lemma 4.4 and Gagliardo-Nirenberg inequality entails that

vεv0L(0,T;C[a+δ,bδ])C0(wL(0,T;L2(a+δ,bδ))+wL(0,T;L2(a+δ,bδ))1/2wrL(0,T;L2(a+δ,bδ))1/2)C(ε1/4+ε1/8ε1/8δ1/2)Cε1/4δ1/2,\displaystyle\begin{split}\|v^{\varepsilon}-v^{0}\|_{L^{\infty}(0,T;C[a+\delta,b-\delta])}\leq&C_{0}\big{(}\|w\|_{L^{\infty}(0,T;L^{2}(a+\delta,b-\delta))}\\ &+\|w\|_{L^{\infty}(0,T;L^{2}(a+\delta,b-\delta))}^{1/2}\|w_{r}\|_{L^{\infty}(0,T;L^{2}(a+\delta,b-\delta))}^{1/2}\big{)}\\ \leq&C\big{(}\varepsilon^{1/4}+\varepsilon^{1/8}\cdot\varepsilon^{1/8}\delta^{-1/2}\big{)}\\ \leq&C\varepsilon^{1/4}\delta^{-1/2},\end{split}

provided δ<1\delta<1. Hence we derive (2.6) and we next prove the equivalence between (2.7) and (2.8). We first prove that (2.8) implies (2.7). Assume 0t0ur0(a,τ)𝑑τ0\int_{0}^{t_{0}}u^{0}_{r}(a,\tau)\,d\tau\neq 0 for some t0[0,T]t_{0}\in[0,T]. Then integrating the second equation of (1.15) with ε=0\varepsilon=0 over (0,t0)(0,t_{0}) along with compatible condition v0(a)=v¯1v_{0}(a)=\bar{v}_{1} gives

v0(a,t0)=v¯1+0t0ur0(a,τ)𝑑τ.v^{0}(a,t_{0})=\bar{v}_{1}+\int_{0}^{t_{0}}u^{0}_{r}(a,\tau)\,d\tau. (4.25)

We thus have

lim infε0vεv0L(0,T;C[a,b])lim infε0|v¯1v0(a,t0)|=lim infε0|0t0ur0(a,τ)𝑑τ|>0.\displaystyle\begin{split}\liminf_{\varepsilon\rightarrow 0}\|v^{\varepsilon}-v^{0}\|_{L^{\infty}(0,T;C[a,b])}\geq\liminf_{\varepsilon\rightarrow 0}|\bar{v}_{1}-v^{0}(a,t_{0})|=\liminf_{\varepsilon\rightarrow 0}|\int_{0}^{t_{0}}u^{0}_{r}(a,\tau)\,d\tau|>0.\end{split}

Similar arguments lead to (2.7) when assuming that 0t0ur0(b,τ)𝑑τ0\int_{0}^{t_{0}}u^{0}_{r}(b,\tau)\,d\tau\neq 0 for some t0[0,T]t_{0}\in[0,T]. We thus proved (2.7) provided (2.8). The proof of that (2.7)\eqref{j8} indicates (2.8)\eqref{j9} will follow from the argument of contradiction. Indeed, if we assume that (2.7) holds and the opposite of (2.8) holds, that is

0tur0(a,τ)𝑑τ=0and0tur0(b,τ)𝑑τ=0,forallt[0,T],\displaystyle\int_{0}^{t}u^{0}_{r}(a,\tau)\,d\tau=0\quad{\rm and}\quad\int_{0}^{t}u^{0}_{r}(b,\tau)\,d\tau=0,\quad{\rm for}\,\,{\rm all}\,\,t\in[0,T],

which, along with the second equation of (1.15) with ε=0\varepsilon=0 leads to

(vεv0)(a,t)=[v0(a,t)v¯1]=0tur0(a,τ)𝑑τ=0,(vεv0)(b,t)=[v0(b,t)v¯2]=0tur0(b,τ)𝑑τ=0,\displaystyle\begin{split}(v^{\varepsilon}-v^{0})(a,t)=-[v^{0}(a,t)-\bar{v}_{1}]=-\int_{0}^{t}u^{0}_{r}(a,\tau)d\tau=0,\\ (v^{\varepsilon}-v^{0})(b,t)=-[v^{0}(b,t)-\bar{v}_{2}]=-\int_{0}^{t}u^{0}_{r}(b,\tau)d\tau=0,\end{split}

where the compatible conditions v0(a)=v¯1,v0(b)=v¯2v_{0}(a)=\bar{v}_{1},\,v_{0}(b)=\bar{v}_{2} have been used. Thus w|r=a,b=(vεv0)|r=a,b=0w|_{r=a,b}=(v^{\varepsilon}-v^{0})|_{r=a,b}=0 and wt|r=a,b=[(vεv0)|r=a,b]t=0w_{t}|_{r=a,b}=[(v^{\varepsilon}-v^{0})|_{r=a,b}]_{t}=0 and the terms 2ε[rn1wrw]|ab2\varepsilon[r^{n-1}w_{r}w]|_{a}^{b}, 2ε[rn1wrwt]|ab2\varepsilon[r^{n-1}w_{r}w_{t}]|_{a}^{b} in (4.10) and (4.12) would vanish. Then by similar arguments as deriving (4.20), we conclude that

hL(0,T;L2)2+wL(0,T;L2)2+εwrL(0,T;L2)2Cε2,\displaystyle\|h\|_{L^{\infty}(0,T;L^{2})}^{2}+\|w\|_{L^{\infty}(0,T;L^{2})}^{2}+\varepsilon\|w_{r}\|_{L^{\infty}(0,T;L^{2})}^{2}\leq C\varepsilon^{2},

which, along with Sobolev embedding inequality gives rise to

limε0vεv0L(0,T;C[a,b])C0limε0(wL(0,T;L2)+wrL(0,T;L2))=0,\displaystyle\lim_{\varepsilon\rightarrow 0}\|v^{\varepsilon}-v^{0}\|_{L^{\infty}(0,T;C[a,b])}\leq C_{0}\lim_{\varepsilon\rightarrow 0}(\|w\|_{L^{\infty}(0,T;L^{2})}+\|w_{r}\|_{L^{\infty}(0,T;L^{2})})=0,

which, contradicts with (2.7), thus (2.7) implies (2.8). The proof is completed.

\Box
  We next convert the result of Theorem 2.2 to the initial-boundary value problem (2.9) for the original chemotaxis model to prove Proposition 2.1.

  Proof of Proposition 2.1. Let

cε(r,t)=c0(r)exp{0t[uε+ε(vε)2εvrεεn1rvε](r,τ)𝑑τ},c0(r,t)=c0(r)exp{0tu0(r,τ)𝑑τ},\begin{split}&c^{\varepsilon}(r,t)=c_{0}(r)\exp\left\{\int_{0}^{t}\left[-u^{\varepsilon}+\varepsilon(v^{\varepsilon})^{2}-\varepsilon v^{\varepsilon}_{r}-\varepsilon\frac{n-1}{r}v^{\varepsilon}\right](r,\tau)d\tau\right\},\\ &c^{0}(r,t)=c_{0}(r)\exp\left\{-\int_{0}^{t}u^{0}(r,\tau)d\tau\right\},\end{split} (4.26)

where (uε,vε)(u^{\varepsilon},v^{\varepsilon}) and (u0,v0)(u^{0},v^{0}) are the solutions of problem (1.15)-(1.16) with ε>0\varepsilon>0 and ε=0\varepsilon=0 respectively. It is easy to check that (uε,cε)(u^{\varepsilon},c^{\varepsilon}) and (u0,c0)(u^{0},c^{0}) with cεc^{\varepsilon} and c0c^{0} defined (4.26) are the unique solutions of (2.9) corresponding to ε>0\varepsilon>0 and ε=0\varepsilon=0, respectively.

The first inequality in (2.10) follows directly from the Sobolev embedding inequality, Lemma 4.4 and Lemma 4.5 as following:

uεu0L(0,T;C[a,b])C0uεu0L(0,T;H1)Cε1/4.\displaystyle\|u^{\varepsilon}-u^{0}\|_{L^{\infty}(0,T;C[a,b])}\leq C_{0}\|u^{\varepsilon}-u^{0}\|_{L^{\infty}(0,T;H^{1})}\leq C\varepsilon^{1/4}.

To prove the second inequality in (2.10), we deduce from (4.26) that

|cε\displaystyle|c^{\varepsilon} (r,t)c0(r,t)|\displaystyle(r,t)-c^{0}(r,t)| (4.27)
=\displaystyle= |c0(r,t)||exp{0t[(uεu0)+ε(vε)2εvrεεn1rvε]𝑑τ}1|\displaystyle|c^{0}(r,t)|\cdot\bigg{|}\exp\bigg{\{}\int_{0}^{t}[-(u^{\varepsilon}-u^{0})+\varepsilon(v^{\varepsilon})^{2}-\varepsilon v^{\varepsilon}_{r}-\varepsilon\frac{n-1}{r}v^{\varepsilon}]\,d\tau\bigg{\}}-1\bigg{|}
=\displaystyle= |c0(r,t)||eGε(r,t)1|,\displaystyle|c^{0}(r,t)|\cdot|e^{G^{\varepsilon}(r,t)}-1|,

where we have denoted by Gε(r,t)=0t[(uεu0)+ε(vε)2εvrεεn1rvε]𝑑τG^{\varepsilon}(r,t)=\int_{0}^{t}[-(u^{\varepsilon}-u^{0})+\varepsilon(v^{\varepsilon})^{2}-\varepsilon v^{\varepsilon}_{r}-\varepsilon\frac{n-1}{r}v^{\varepsilon}]\,d\tau for convenience. For (r,t)[a,b]×[0,T](r,t)\in[a,b]\times[0,T], it follows from Lemma 4.4 and the Sobolev embedding inequality that

|Gε(r,t)|\displaystyle|G^{\varepsilon}(r,t)| (4.28)
\displaystyle\leq C0T1/2uεu0L2(0,T;L)+C0TεvεL(0,T;L)2+C0T1/2εvrεL2(0,T;L)\displaystyle C_{0}T^{1/2}\|u^{\varepsilon}-u^{0}\|_{L^{2}(0,T;L^{\infty})}+C_{0}T\varepsilon\|v^{\varepsilon}\|_{L^{\infty}(0,T;L^{\infty})}^{2}+C_{0}T^{1/2}\varepsilon\|v^{\varepsilon}_{r}\|_{L^{2}(0,T;L^{\infty})}
+C0T1/2εvεL2(0,T;L)\displaystyle+C_{0}T^{1/2}\varepsilon\|v^{\varepsilon}\|_{L^{2}(0,T;L^{\infty})}
\displaystyle\leq C0T1/2uεu0L2(0,T;H1)+C0TεvεL(0,T;H1)2+C0T1/2εvrεL2(0,T;H1)\displaystyle C_{0}T^{1/2}\|u^{\varepsilon}-u^{0}\|_{L^{2}(0,T;H^{1})}+C_{0}T\varepsilon\|v^{\varepsilon}\|_{L^{\infty}(0,T;H^{1})}^{2}+C_{0}T^{1/2}\varepsilon\|v^{\varepsilon}_{r}\|_{L^{2}(0,T;H^{1})}
+C0T1/2εvεL2(0,T;H1)\displaystyle+C_{0}T^{1/2}\varepsilon\|v^{\varepsilon}\|_{L^{2}(0,T;H^{1})}
\displaystyle\leq Cε1/4+Cε(v0L(0,T;H1)2+ε1/2)+Cε(vr0L2(0,T;H1)+ε3/4)\displaystyle C\varepsilon^{1/4}+C\varepsilon\left(\|v^{0}\|_{L^{\infty}(0,T;H^{1})}^{2}+\varepsilon^{-1/2}\right)+C\varepsilon\left(\|v^{0}_{r}\|_{L^{2}(0,T;H^{1})}+\varepsilon^{-3/4}\right)
+Cε(v0L2(0,T;H1)+ε1/4)\displaystyle+C\varepsilon\left(\|v^{0}\|_{L^{2}(0,T;H^{1})}+\varepsilon^{-1/4}\right)
\displaystyle\leq C1ε1/4,\displaystyle C_{1}\varepsilon^{1/4},

for some constant C1C_{1} independent of ε\varepsilon (depending on TT) and the assumption 0<ε<10<\varepsilon<1 has been used in the last inequality. Then we apply Taylor expansion to eGε(r,t)e^{G^{\varepsilon}(r,t)} and using (4.28) to conclude that

|eGε(r,t)1|k=11k!|Gε(x,t)|kk=1C1kk!ε1/4C2ε1/4,\left|e^{G^{\varepsilon}(r,t)}-1\right|\leq\sum_{k=1}^{\infty}\frac{1}{k!}|G^{\varepsilon}(x,t)|^{k}\leq\sum_{k=1}^{\infty}\frac{C_{1}^{k}}{k!}\varepsilon^{1/4}\leq C_{2}\varepsilon^{1/4}, (4.29)

where the assumption 0<ε<10<\varepsilon<1 has been used in the second inequality and the constant C2:=eC1C_{2}:=e^{C_{1}} is independent of ε\varepsilon. On the other hand, by the assumptions c0(r)>0c_{0}(r)>0 and lnc0H3\ln c_{0}\in H^{3} in Proposition 2.1 we derive that lnc0Llnc0H1C3\|\ln c_{0}\|_{L^{\infty}}\leq\|\ln c_{0}\|_{H^{1}}\leq C_{3} for some positive constant C3C_{3}, which along with the fact c0(r)=elnc0(r)c_{0}(r)=e^{\ln c_{0}(r)} leads to

eC3c0(r)eC3forr[a,b].e^{-C_{3}}\leq c_{0}(r)\leq e^{C_{3}}\qquad\text{for}\,\,r\in[a,b]. (4.30)

Moreover, from Theorem 2.1 we know that

0tu0(r,τ)𝑑τL(0,T;L)C0Tu0L(0,T;H1)C4fort[0,T],\left\|\int_{0}^{t}u^{0}(r,\tau)d\tau\right\|_{L^{\infty}(0,T;L^{\infty})}\leq C_{0}T\|u^{0}\|_{L^{\infty}(0,T;H^{1})}\leq C_{4}\qquad\text{for}\,\,t\in[0,T], (4.31)

where the constant C4C_{4} depending on TT. Thus we deduce from the second equality of (4.26), (4.30) and (4.31) that

0<C51<c0(r,t)<C5for(r,t)[a,b]×[0,T],0<C_{5}^{-1}<c^{0}(r,t)<C_{5}\qquad\text{for}\,\,(r,t)\in[a,b]\times[0,T], (4.32)

with C5=e(C3+C4)C_{5}=e^{(C_{3}+C_{4})}. Hence, it follows from (4.27), (4.29) and (4.32) that

cεc0L(0,T;L)C6ε1/4,\|c^{\varepsilon}-c^{0}\|_{L^{\infty}(0,T;L^{\infty})}\leq C_{6}\varepsilon^{1/4}, (4.33)

where C6:=C2C5C_{6}:=C_{2}C_{5} is independent of ε\varepsilon. We thus derive the second inequality in (2.10) and proceed to prove (2.11). It follows from transformation (1.11) that

crεcr0=(v0vε)cε+v0(c0cε),c^{\varepsilon}_{r}-c^{0}_{r}=(v^{0}-v^{\varepsilon})c^{\varepsilon}+v^{0}(c^{0}-c^{\varepsilon}), (4.34)

which, in conjunction with (2.6) and (2.10) leads to

crεcr0L(0,T;C[a+δ,bδ])vεv0L(0,T;C[a+δ,bδ])(c0L(0,T;C[a,b])+Cε1/4)+Cε1/4v0L(0,T;C[a,b])Cε1/4δ1/2,\displaystyle\begin{split}\|c^{\varepsilon}_{r}-c^{0}_{r}\|_{L^{\infty}(0,T;C[a+\delta,b-\delta])}\leq&\|v^{\varepsilon}-v^{0}\|_{L^{\infty}(0,T;C[a+\delta,b-\delta])}(\|c^{0}\|_{L^{\infty}(0,T;C[a,b])}+C\varepsilon^{1/4})\\ &+C\varepsilon^{1/4}\|v^{0}\|_{L^{\infty}(0,T;C[a,b])}\\ \leq&C\varepsilon^{1/4}\delta^{-1/2},\end{split}

where δ<1\delta<1 has been used thanks to δ(ε)0\delta(\varepsilon)\rightarrow 0 as ε0\varepsilon\rightarrow 0. We thus derived (2.11). To prove the equivalence between (2.12) and (2.8), we first derives two positive constants C7C_{7} and C8C_{8} independent of ε\varepsilon such that

0<C7cε(r,t)C8 for(r,t)[a,b]×[0,T],0<C_{7}\leq c^{\varepsilon}(r,t)\leq C_{8}\qquad\text{ for}\quad(r,t)\in[a,b]\times[0,T], (4.35)

by choosing ε\varepsilon small enough such that C6ε1/4<12C5C_{6}\varepsilon^{1/4}<\frac{1}{2C_{5}} in (4.33) and using (4.32). With (4.35) in hand, we next prove the equivalence between (2.12) and (2.8). First, it follows from (4.34), (2.10) and (4.35) that

lim infε0crεcr0L(0,T;C[a,b])lim infε0[cεL(0,T;L)vεv0L(0,T;L)v0L(0,T;L)cεc0L(0,T;L)]lim infε0[cεL(0,T;L)vεv0L(0,T;L)]lim supε0[v0L(0,T;L)cεc0L(0,T;L)]C7lim infε0vεv0L(0,T;L).\begin{split}&\liminf_{\varepsilon\rightarrow 0}\|c^{\varepsilon}_{r}-c^{0}_{r}\|_{L^{\infty}(0,T;C[a,b])}\\ \geq&\liminf_{\varepsilon\rightarrow 0}\left[\|c^{\varepsilon}\|_{L^{\infty}(0,T;L^{\infty})}\|v^{\varepsilon}-v^{0}\|_{L^{\infty}(0,T;L^{\infty})}-\|v^{0}\|_{L^{\infty}(0,T;L^{\infty})}\|c^{\varepsilon}-c^{0}\|_{L^{\infty}(0,T;L^{\infty})}\right]\\ \geq&\liminf_{\varepsilon\rightarrow 0}\left[\|c^{\varepsilon}\|_{L^{\infty}(0,T;L^{\infty})}\|v^{\varepsilon}-v^{0}\|_{L^{\infty}(0,T;L^{\infty})}\right]\\ &-\limsup_{\varepsilon\rightarrow 0}\left[\|v^{0}\|_{L^{\infty}(0,T;L^{\infty})}\|c^{\varepsilon}-c^{0}\|_{L^{\infty}(0,T;L^{\infty})}\right]\\ \geq&C_{7}\liminf_{\varepsilon\rightarrow 0}\|v^{\varepsilon}-v^{0}\|_{L^{\infty}(0,T;L^{\infty})}.\end{split} (4.36)

Dividing (4.34) by cεc^{\varepsilon} and applying a similar argument as deriving (4.36), then using (4.35) and (2.10), one deduces that

lim infε0vεv0L(0,T;C[a,b])lim infε0crεcr0cεL(0,T;L)lim supε0v0(1c0cε)L(0,T;L)1C8lim infε0crεcr0L(0,T;L),\displaystyle\begin{split}&\liminf_{\varepsilon\rightarrow 0}\|v^{\varepsilon}-v^{0}\|_{L^{\infty}(0,T;C[a,b])}\\ \geq&\liminf_{\varepsilon\rightarrow 0}\left\|\frac{c^{\varepsilon}_{r}-c^{0}_{r}}{c^{\varepsilon}}\right\|_{L^{\infty}(0,T;L^{\infty})}-\limsup_{\varepsilon\rightarrow 0}\left\|v^{0}\left(1-\frac{c^{0}}{c^{\varepsilon}}\right)\right\|_{L^{\infty}(0,T;L^{\infty})}\\ \geq&\frac{1}{C_{8}}\,\liminf_{\varepsilon\rightarrow 0}\|c^{\varepsilon}_{r}-c^{0}_{r}\|_{L^{\infty}(0,T;L^{\infty})},\end{split}

which, in conjunction with (4.36) indicates the equivalence between (2.12) and (2.7). Then we conclude that (2.12) is equivalent to (2.8) by using Theorem 2.2. The proof is completed.

\Box

Acknowledgements

This work is supported by China Postdoctoral Science Foundation (No.2019M651269), National Natural Science Foundation of China (No.11901139).

References

  • [1] J. Adler. Chemotaxis in bacteria. Science, 153:708–716, 1966.
  • [2] J. Adler. Chemoreceptors in bacteria. Science, 166:1588–1597, 1969.
  • [3] M. Chae, K. Choi, K. Kang, and J. Lee. Stability of planar traveling waves in a Keller-Segel equation on an infinite strip domain. J. Differential Equations, 265:237–279, 2018.
  • [4] M.A.J. Chaplain and A.M. Stuart. A model mechanism for the chemotactic response of endothelial cells to tumor angiogenesis factor. IMA J. Math. Appl. Med., 10(3):149–168, 1993.
  • [5] P. Constantin. Note on loss of regularity for solutions of the 3d incompressible euler and related equations. Commun. Math. Phys., 104:311–326, 1986.
  • [6] L. Corrias, B. Perthame, and H. Zaag. A chemotaxis model motivated by angiogenesis. C. R. Math. Acad. Sci. Paris, 2:141–146, 2003.
  • [7] L. Corrias, B. Perthame, and H. Zaag. Global solutions of some chemotaxis and angiogenesis systems in high space dimensions. Milan  J.  Math., 72:1–29, 2004.
  • [8] C. Deng and T. Li. Well-posedness of a 3D parabolic-hyperbolic Keller-Segel system in the sobolev space framework. J. Differential Equations, 257:1311–1332, 2014.
  • [9] P.C. File. Considerations regarding the mathematical basis for Prandtl’s boundary layer theory. Arch. Ration. Mech. Anal., 28(3):184–216, 1968.
  • [10] H. Frid and V. Shelukhin. Boundary layers for the Navier-Stokes equations of compressible fluids. Comm. Math. Phys., 208:309–330, 1999.
  • [11] H. Frid and V. Shelukhin. Vanishing shear viscosity in the equations of compressible fluids for the flows with the cylinder symmetry. SIAM J. Math. Anal., 31:1144–1156, 2000.
  • [12] A. Gamba, D. Ambrosi, A. Coniglio, A. de Candia, S. Di Talia, E. Giraudo, G. Serini, L. Preziosi, and F. Bussolino. Percolation, morphogenesis, and burgers dynamics in blood vessels formation. Phys. Rev. Lett., 90:118101, 2003.
  • [13] J. Guo, J.X. Xiao, H.J. Zhao, and C.J. Zhu. Global solutions to a hyperbolic-parabolic coupled system with large initial data. Acta Math. Sci. Ser. B Engl. Ed, 29:629–641, 2009.
  • [14] C. Hao. Global well-posedness for a multidimensional chemotaxis model in critical besov spaces. Z. Angew Math. Phys., 63:825–834, 2012.
  • [15] H. Höfer, J.A. Sherratt, and P.K. Maini. Cellular pattern formation during Dictyostelium aggregation. Physica D., 85:425–444, 1995.
  • [16] Q. Hou, Z. Wang, and K. Zhao. Boundary layer problem on a hyperbolic system arising from chemotaxis. J. Differential Equations, 261:5035–5070, 2016.
  • [17] S. Jiang and J. Zhang. On the non-resistive limit and the magnetic boundary-layer for one-dimensional compressible magnetohydrodynamics. Nonlinearity, 30:3587–3612, 2017.
  • [18] H.Y. Jin, J. Li, and Z. Wang. Asymptotic stability of traveling waves of a chemotaxis model with singular sensitivity. J. Differential Equations, 255(2):193–219, 2013.
  • [19] E.F. Keller and L.A. Segel. Initiation of slime mold aggregation viewed as an instability. J. Theor. Biol., 26(3):399–415, 1970.
  • [20] E.F. Keller and L.A. Segel. Model for chemotaxis. J. Theor.  Biol., 30:225–234, 1971.
  • [21] E.F. Keller and L.A. Segel. Traveling bands of chemotactic bacteria: A theoretical analysis. J. Theor. Biol., 30:377–380, 1971.
  • [22] H.A. Levine and B.D. Sleeman. A system of reaction diffusion equations arising in the theory of reinforced random walks. SIAM J. Appl. Math., 57:683–730, 1997.
  • [23] H.A. Levine, B.D. Sleeman, and M. Nilsen-Hamilton. A mathematical model for the roles of pericytes and macrophages in the initiation of angiogenesis. I. the role of protease inhibitors in preventing angiogenesis. Math. Biosci., 168:71–115, 2000.
  • [24] D. Li, T. Li, and K. Zhao. On a hyperbolic-parabolic system modeling chemotaxis. Math. Models Methods Appl. Sci., 21:1631–1650, 2011.
  • [25] D. Li, R. Pan, and K. Zhao. Quantitative decay of a one-dimensional hybrid chemotaxis model with large data. Nonlinearity, 7:2181–2210, 2015.
  • [26] H. Li and K. Zhao. Initial-boundary value problems for a system of hyperbolic balance laws arising from chemotaxis. J. Differential Equations, 258(2):302–338, 2015.
  • [27] J. Li, T. Li, and Z. Wang. Stability of traveling waves of the Keller-Segel system with logarithmic sensitivity. Math. Models Methods Appl. Sci., 24(14):2819–2849, 2014.
  • [28] T. Li, R. Pan, and K. Zhao. Global dynamics of a hyperbolic-parabolic model arising from chemotaxis. SIAM J. Appl. Math., 72(1):417–443, 2012.
  • [29] T. Li and Z. Wang. Nonlinear stability of travelling waves to a hyperbolic-parabolic system modeling chemotaxis. SIAM J. Appl. Math., 70(5):1522–1541, 2009.
  • [30] T. Li and Z. Wang. Nonlinear stability of large amplitude viscous shock waves of a hyperbolic-parabolic system arising in chemotaxis. Math. Models Methods Appl. Sci., 20(10):1967–1998, 2010.
  • [31] T. Li and Z. Wang. Asymptotic nonlinear stability of traveling waves to conservation laws arising from chemotaxis. J. Differential Equations, 250(3):1310–1333, 2011.
  • [32] T. Li and Z. Wang. Steadily propagating waves of a chemotaxis model. Math. Biosci., 240(2):161–168, 2012.
  • [33] V. Martinez, Z. Wang, and K. Zhao. Asymptotic and viscous stability of large-amplitude solutions of a hyperbolic system arising from biology. Indiana Univ. Math. J. , 67:1383–1424, 2018.
  • [34] J.D. Murray. Mathematical Biology I: An Introduction. Springer, Berlin, 3rd edition, 2002.
  • [35] K.J. Painter, P.K. Maini, and H.G. Othmer. Stripe formation in juvenile pomacanthus explained by a generalized Turing mechanism with chemotaxis. Proc. Natl. Acad. Sci., 96:5549–5554, 1999.
  • [36] K.J. Painter, P.K. Maini, and H.G. Othmer. A chemotactic model for the advance and retreat of the primitive streak in avian development. Bull. Math. Biol., 62:501–525, 2000.
  • [37] H. Peng, L. Ruan, and C. Zhu. Convergence rates of zero diffusion limit on large amplitude solution to a conservation laws arising in chemotaxis. Kinetic and Related Models, 5:563–581, 2012.
  • [38] H. Peng, H. Wen, and C. Zhu. Global well-posedness and zero diffusion limit of classical solutions to 3D conservation laws arising in chemotaxis. Z. Angew Math. Phys., 65(6):1167–1188, 2014.
  • [39] G.J. Petter, H.M. Byrne, D.L.S. Mcelwain, and J. Norbury. A model of wound healing and angiogenesis in soft tissue. Math. Biosci., 136(1):35–63, 2003.
  • [40] L. Prandtl. Über Flüssigkeitsbewegungen bei sehr kleiner Reibung. In “Verh. Int. Math. Kongr., Heidelberg 1904”, Teubner, 1905.
  • [41] L.G. Rebholz, D. Wang, Z. Wang, C. Zerfas, and K. Zhao. Initial boundary value problems for a system of parabolic conservation laws arising from chemotaxis in multi-dimensions. Discrete Contin. Dyn. Syst., 39(7):3789–3838, 2019.
  • [42] H. Schlichting. Boundary layer theory. McGraw-Hill company, London-New York, 7th edition, 1987.
  • [43] Y.S. Tao, L.H. Wang, and Z. Wang. Large-time behavior of a parabolic-parabolic chemotaxis model with logarithmic sensitivity in one dimension. Discrete Contin. Dyn. Syst-Series B., 18:821–845, 2013.
  • [44] R. Tyson, S.R. Lubkin, and J. Murray. Models and analysis of chemotactic bacterial patterns in a liquid medium. J.  Math.  Biol., 266:299–304, 1999.
  • [45] D. Wang, Z. Wang, and K. Zhao. Cauchy problem of a system of parabolic conservationlaws arising from the singular keller-segel model inmulti-dimensions. Indiana U. Math. J., 2019.
  • [46] Y.G. Wang and Z. Xin. Zero-viscosity limit of the linearized compressible Navier-Stokes equations with highly oscillatory forces in the half-plane. SIAM  J. Math. Anal., 37(4):1256–1298, 2005.
  • [47] Z. Wang. Mathematics of traveling waves in chemotaxis. Discrete Contin. Dyn. Syst-Series B, 18:601–641, 2013.
  • [48] Z. Wang, Z. Xiang, and P. Yu. Asymptotic dynamics on a singular chemotaxis system modeling onset of tumor angiogenesis. J. Differential Equations, 260:2225–2258, 2016.
  • [49] Z. Wang and K. Zhao. Global dynamics and diffusion limit of a one-dimensional repulsive chemotaxis model. Comm. Pure Appl. Anal., 12:3027–3046, 2013.
  • [50] M. Winkler. Renormalized radial large-data solutions to the higher-dimensional keller-segel system with singular sensitivity and signal absorption. J. Differential Equations, 264:2310–2350, 2018.
  • [51] J. Wu and X. Xu. Well-posedness and inviscid limits of the boussinesq equations with fractional laplacian dissipation. Nonlinearity, 27:2215–2232, 2014.
  • [52] Z. Xin and T. Yanagisawa. Zero-viscosity limit of the linearized Navier-Stokes equations for a compressible viscous fluid in the half-plane. Comm. Pure Appl. Math., 52(4):479–541, 1999.
  • [53] L. Yao, T. Zhang, and C. Zhu. Boundary layers for compressible Navier-Stokes equations with density-dependent viscosity and cylindrical symmetry. Ann. Inst. H. Poincaré Anal. Non Linéaire, 28(5):677–709, 2011.
  • [54] M. Zhang and C. Zhu. Global existence of solutions to a hyperbolic-parabolic system. Proceedings of the American Mathematical Society, 135:1017–1027, 2007.