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

\newsiamremark

remarkRemark \newsiamremarkhypothesisHypothesis \newsiamthmclaimClaim \headersWell-posedness of Cahn-Hilliard model for surface diffusionNiu, Xiang,Yan

Well-posedness of a modified degenerate Cahn-Hilliard model for surface diffusion

Xiaohua Niu School of Mathematics and Statistics, Xiamen University of Technology, Xiamen, China (). [email protected]    Yang Xiang Department of Mathematics, The Hong Kong University of Science and Technollogy, Clear Water Bay, Kowloon, Hong Kong (). [email protected]    Xiaodong Yan Department of Mathematics, The University of Connecticut, Storrs, CT, USA (). [email protected]
Abstract

We study the well-posedness of a modified degenerate Cahn-Hilliard type model for surface diffusion. With degenerate phase-dependent diffusion mobility and additional stabilizing function, this model is able to give the correct sharp interface limit. We introduce a notion of weak solutions for the nonlinear model. The existence result is obtained by approximations of the proposed model with nondegenerate mobilities. We also employ this method to prove existence of weak solutions to a related model where the chemical potential contains a nonlocal term originated from self-climb of dislocations in crystalline materials.

keywords:
Phase field model, degenerate Cahn-Hilliard equation, surface diffusion, well-posedness, weak solutions
{AMS}

35A01, 35G20, 35K25, 74N20, 82C26

1 Introduction

We consider the following modified degenerate Cahn-Hilliard type model

(1) g(u)tu\displaystyle g(u)\partial_{t}u =\displaystyle= (M(u)μg(u)),xΩ2,t[0,)\displaystyle\nabla\cdot(M(u)\nabla\frac{\mu}{g(u)}),\ \ x\in\Omega\subset\mathbb{R}^{2},t\in[0,\infty)
(2) μ\displaystyle\mu =\displaystyle= Δu+1ε2q(u).\displaystyle-\Delta u+\frac{1}{\varepsilon^{2}}q^{\prime}(u).

When g1g\equiv 1, (1)-(2) becomes Cahn-Hilliard (CH) equation with degenerate mobility. The degenerate Cahn-Hilliard equation has been widely studied as a diffuse-interface model for phase separation in binary system [2, 3, 4, 5, 7, 10]. Over the years, the interface motion in the sharp limit has caught a lot of attention for various choice of mobility M(u)M(u) and homogeneous free energy q(u)q(u). When M(u)=1u2M(u)=1-u^{2} and qq being either the logarithmic free energy

q(u)=θ2[(1+u)ln(1+u)+(1u)ln(1u)]+12(1u2)q(u)=\frac{\theta}{2}\left[(1+u)\ln(1+u)+(1-u)\ln(1-u)\right]+\frac{1}{2}(1-u^{2})

with temperature θ=O(εα)\theta=O(\varepsilon^{\alpha}) or the double obstacle potential

q(u)=1u2 for |u|1, q(u)= otherwise,q(u)=1-u^{2}\text{ for }|u|\leq 1,\text{ }q(u)=\infty\text{ otherwise},

Cahn, Elliott, and Novick-Cohen [19] showed via asymptotic expansions that the sharp-interface limit in the time scale O(ε2)O(\varepsilon^{-2}) is interface motion by surface diffusion. Sharp interface limits for different time scales were discussed in [5] for highly disparate diffusion mobility M(u)=1+uM(u)=1+u and smooth double well q(u)=14(1u2)2q(u)=\frac{1}{4}(1-u^{2})^{2}. In particular, the system evolves in t=O(ε2)t=O(\varepsilon^{-2}) time scale according to the combination of a one-sided modified Mullins–Sekerka problem in the phase with nonzero constant mobility and a nonlinear diffusion process that solves a quasi-stationary porous medium equation in the phase with small mobility. A later work by the same authors [6] derived sharp interface limit for O(ε2)O(\varepsilon^{-2}) time scale with diffusion mobility M(u)=|1u2|M(u)=|1-u^{2}| and smooth double well potential q(u)q(u), noting the effect of the diffusion field on the interface motion as a jump of fluxes. The analysis was done on the (unphysical) solution branch with |u|>1|u|>1 on some region. For M(u)=1u2M(u)=1-u^{2} and q=14(1u2)2q=\frac{1}{4}(1-u^{2})^{2}, Lee, Münch and Süli [14] considered the physical branch of solution where |u|<1|u|<1 everywhere and showed that there is an additional nonlinear bulk diffusion term appearing to leading order of the sharp interface limit. Further study in [15] indicates that the leading order sharp-interface motion depends sensitively on the choice of mobility.

The existence of weak solutions for degenerate Cahn-Hilliard equation was proved by Elliotte and Garcke [10] (see [24] for 1D case). Their results include the case M(u)=1u2M(u)=1-u^{2} and qq being the logarithmic free energy. Dai and Du [7] introduced a different notion of weak solutions for degenerate Cahn-Hilliard equation with mobility M(u)=|1u2|mM(u)=|1-u^{2}|^{m} and smooth double well potentials; they showed that their model accommodates the Gibbs-Thomson effect, which was not by the method in [10].

There is a critical issue in modeling surface diffusion by the degenerate Cahn-Hilliard model [11, 21], due to the presence of incompatibility in the asymptotic matching between the outer and inner expansions. Rätz, Ribalta, and Voigt (RRV) [21] fixed this incompatibility by introducing a singular factor 1/g(u)1/g\left(u\right) in front of the chemical potential μ\mu to force it to vanish in the far field. Their model essentially consists of equations (1)-(2) without the g(u)g(u) term on the left side of (1), and other terms for modeling heteroepitaxial growth of thin films. The RRV model with the stabilizing function g(u)g(u) has been validated by numerical simulations [21] and asymptotic analyses [11, 21]. It has been successfully generalized to many applications, e.g., growth of nanoscale membranes [1], dewetting of ultrathin films [17], and grain boundary formation in nanoporous metals [9]. Recently, a phase field model for dislocation self-climb by vacancy pipe diffusion was developed based on degenerate Cahn-Hilliard model with such stabilizing function [20]. However, to the best of our knowledge, well-posedness of these degenerate Cahn-Hilliard models with singular factor that give the correct sharp interface limit for surface diffusion has not been established in the literature.

In this paper, in order to prove the well-posedness of the RRV type Cahn-Hilliard model with correct sharp interface limit for surface diffusion, we propose a modified degenerated Cahn-Hilliard model as given in (1)-(2), and discuss its well-posedness and sharp interface limit. In particular, we have modified the original RRV model so that the equation can be written in the form of gradient flow of the total energy.

Our first result is a sharp interface limit equation for (1) and (2) via formal asymptotic analysis. We obtain the following sharp interface equation

(3) v=λss(ακ)v=\lambda\partial_{ss}(\alpha\kappa)

as ε\varepsilon\rightarrow\infty. Here λ<0\lambda<0, α<0\alpha<0 are constants whose exact forms are derived in section 2. This validates this equation as a diffuse-interface model for surface diffusion.

Our main result concerns the well-posedness of the initial value problem of (1)-(2). For this purpose, we set Ω=[0,2π]n\Omega=[0,2\pi]^{n} and consider the following problem in a periodic setting when n2n\leq 2.

(4) g(u)tu\displaystyle g(u)\partial_{t}u =\displaystyle= (M(u)μg(u)), for xΩ,t[0,)\displaystyle\nabla\cdot(M(u)\nabla\frac{\mu}{g(u)}),\hskip 28.90755pt\text{ for }x\in\Omega,t\in[0,\infty)
(5) μ\displaystyle\mu =\displaystyle= Δu+q(u).\displaystyle-\Delta u+q^{\prime}(u).

Here g(u)=|1u2|mg(u)=|1-u^{2}|^{m} for 2m<2\leq m<\infty, M(u)=M0g(u)M(u)=M_{0}g(u) for some constant M0>0M_{0}>0 and q(u)q(u) satisfies the following assumptions.

  • (i)

    q(u)C2(,)q(u)\in C^{2}(\mathbb{R},\mathbb{R}) and there exist constants Ci>0C_{i}>0, i=1,,10i=1,\cdots,10 such that for all uu\in\mathbb{R} and some 1r<1\leq r<\infty, the following growth assumptions hold.

    (6) C1|u|r+1C2q(u)\displaystyle C_{1}|u|^{r+1}-C_{2}\leq q(u) \displaystyle\leq C3|u|r+1+C4,\displaystyle C_{3}|u|^{r+1}+C_{4},
    (7) |q(u)|\displaystyle|q^{\prime}(u)| \displaystyle\leq C5|u|r+C6,\displaystyle C_{5}|u|^{r}+C_{6},
    (8) Ct|u|r1C8\displaystyle C_{t}|u|^{r-1}-C_{8} \displaystyle\leq q′′(u)C9|u|r1+C10.\displaystyle q^{\prime\prime}(u)\leq C_{9}|u|^{r-1}+C_{10}.

We see that the classical double well potential q(u)=(1u2)2q(u)=(1-u^{2})^{2} satisfies (6)-(8) with r=3r=3.

Our existence proof is obtained via approximations of the proposed model (4)-(5) with positive mobilities. Given any θ>0\theta>0, we define

(9) Mθ(u):=M0gθ(u)M_{\theta}(u):=M_{0}g_{\theta}(u)

with

(10) gθ(u):={|1u2|m if |1u2|>θ,θm if |1u2|θ.g_{\theta}(u):=\left\{\begin{array}[]{cl}|1-u^{2}|^{m}&\text{ if }|1-u^{2}|>\theta,\\ \theta^{m}&\text{ if }|1-u^{2}|\leq\theta.\end{array}\right.

Our first step is to find a sufficiently regular solution for (4)-(5) with mobility Mθ(u)M_{\theta}(u) and stablizing function gθ(u)g_{\theta}(u) together with a smooth potential q(u)q(u).

Theorem 1.1.

Let Mθ,gθM_{\theta},g_{\theta} be defined by (9) and (10), under the assumptions (6)-(8), for any u0H1(Ω)u_{0}\in H^{1}(\Omega) and any T>0T>0, there exists a function uθu_{\theta} such that

  • a)

    uθL(0,T;H1(Ω))C([0,T];Lp(Ω))L2(0,T;W3,s(Ω))u_{\theta}\in L^{\infty}(0,T;H^{1}(\Omega))\cap C([0,T];L^{p}(\Omega))\cap L^{2}(0,T;W^{3,s}(\Omega)), where 1p<1\leq p<\infty, 1s<21\leq s<2,

  • b)

    tuθL2(0,T;(W1,q(Ω)))\partial_{t}u_{\theta}\in L^{2}(0,T;(W^{1,q}(\Omega))^{\prime}) for q>2q>2,

  • c)

    uθ(x,0)=u0(x)u_{\theta}(x,0)=u_{0}(x) for all xΩx\in\Omega,

which satisfies (4)-(5) in the following weak sense

0T<tuθ,ϕ>(W1,q(Ω)),W1,q(Ω)dt\displaystyle\int_{0}^{T}<\partial_{t}u_{\theta},\phi>_{(W^{1,q}(\Omega))^{\prime},W^{1,q}(\Omega)}dt
(11) =0TΩMθ(uθ)Δuθ+q(uθ)gθ(uθ)ϕgθ(uθ)dxdt\displaystyle=-\int_{0}^{T}\int_{\Omega}M_{\theta}(u_{\theta})\nabla\frac{-\Delta u_{\theta}+q^{\prime}(u_{\theta})}{g_{\theta}(u_{\theta})}\cdot\nabla\frac{\phi}{g_{\theta}(u_{\theta})}dxdt

for all ϕL2(0,T;W1,q(Ω))\phi\in L^{2}(0,T;W^{1,q}(\Omega)) with q>2q>2. In addition, the following energy inequality holds for all t>0t>0.

(12) Ω(12|uθ(x,t)|2+q(uθ(x,t)))𝑑x\displaystyle\int_{\Omega}\left(\frac{1}{2}|\nabla u_{\theta}(x,t)|^{2}+q(u_{\theta}(x,t))\right)dx
+0tΩMθ(uθ(x,τ)|Δuθ(x,τ)+q(uθ(x,τ)gθ(uθ(x,τ))|2dxdτ\displaystyle+\int_{0}^{t}\int_{\Omega}M_{\theta}(u_{\theta}(x,\tau)\left|\nabla\frac{-\Delta u_{\theta}(x,\tau)+q^{\prime}(u_{\theta}(x,\tau)}{g_{\theta}(u_{\theta}(x,\tau))}\right|^{2}dxd\tau
\displaystyle\leq Ω(12|u0(x)|2+q(u0(x)))𝑑x.\displaystyle\int_{\Omega}\left(\frac{1}{2}|\nabla u_{0}(x)|^{2}+q(u_{0}(x))\right)dx.

Proof of theorem 1.1 is based on Galerkin approximations. Due to the presence of the stablizing function gθg_{\theta}, it is not obvious how to pass to the limit in the nonlinear term of the Galerkin approximations. Our key observation in this step is strong convergence of uN\nabla u^{N} (up to a subsequence) in L2(ΩT)L^{2}(\Omega_{T}) where ΩT=Ω×[0,T])\Omega_{T}=\Omega\times[0,T]) which allows us to pass to the limit in the nonlinear term.

To obtain the weak solution to (4), we consider the limit of uθiu_{\theta_{i}} for a sequence θi0\theta_{i}\downarrow 0. The key challenge is how to pass to the limit in both sides of (11). In the degenerate Cahn-Hilliard case, the estimates for the positive mobility approximations yield a uniform bound for tuθi\partial_{t}u_{\theta_{i}} and it is straighforward to pass to the limit on the left hand side in the approximating equations. Moreover, the bound on tuθi\partial_{t}u_{\theta_{i}}, together with bound on uθiu_{\theta_{i}} yields strong convergence of Mi(uθi)\sqrt{M_{i}(u_{\theta_{i}})} in C(0,T;Ln(Ω))C(0,T;L^{n}(\Omega)). By this and the weak convergence of Mi(uθi)μθi\sqrt{M_{i}(u_{\theta_{i}})}\nabla\mu_{\theta_{i}} in L2(ΩT)L^{2}(\Omega_{T}), Dai and Du [7] showed (up to a subsequence) that Mθi(uθi)μθiM(u)ξM_{\theta_{i}}(u_{\theta_{i}})\nabla\mu_{\theta_{i}}\rightharpoonup\sqrt{M(u)}\xi weakly in L2(0,T;L2nn+2(Ω))L^{2}(0,T;L^{\frac{2n}{n+2}}(\Omega)) where ξ\xi is the weak limit of Mi(uθi)μθi\sqrt{M_{i}(u_{\theta_{i}})}\nabla\mu_{\theta_{i}}. The main task left is to show M(u)ξ=M(u)(Δu+q′′(u)u)\sqrt{M(u)}\xi=M(u)(-\nabla\Delta u+q^{\prime\prime}(u)\nabla u) and the limit equation becomes a weak form Cahn-Hilliard equation. They [7] proved that this is almost true in the set where u±1u\neq\pm 1. Their main idea is the following. For small numbers δj\delta_{j} monotonically decreasing to 0, they consider the limit in a subset BjB_{j} of ΩT\Omega_{T} where approximate solutions converges uniformly and |ΩT\Bj|<δj|\Omega_{T}\backslash B_{j}|<\delta_{j}. By decomposing Bj=DjD~jB_{j}=D_{j}\cup\tilde{D}_{j} where mobility is bounded from below uniformly in DjD_{j} and controlled above in D~j\tilde{D}_{j} by suitable multiples of δj\delta_{j}, they obtain the weak form equation for the limit function by passing to the limit of uθiu_{\theta_{i}} on DjD_{j} then letting jj goes to \infty. Under further regularity assumptions on Δu\nabla\Delta u, they obtained the explicit expression for ξ\xi in the weak form of the equation.

In this paper, we adapt their idea to our model. There are two main difficulties. The first obtacle is the bound estimate on tuθi\partial_{t}u_{\theta_{i}} blows up when θi\theta_{i} goes to zero and we can not pass to the limit on the left hand side of (11); secondly, due to the presence of the stablizing function gg on the right hand side, it is more complicated to derive an explicit expression of the weak limit of Mi(uθi)μθigθi(uθi){M_{i}(u_{\theta_{i}})}\nabla\frac{\mu_{\theta_{i}}}{g_{\theta_{i}}(u_{\theta_{i}})} in terms of uu on the right hand side of the limit equation. To overcome the first difficulty, we derive an alternative form of (11) by multiplying gθ(uθ)g_{\theta}(u_{\theta}) to both sides (valid due to regularity of uθu_{\theta}, c.f. section 3.4 and equation (85)). From this, we obtain uniform estimates on gθi(uθi)tuθig_{\theta_{i}}(u_{\theta_{i}})\partial_{t}u_{\theta_{i}} which enables us to pass to the limit on the left hand side of the alternate equation (85). To find limit form on the right hand side of (85), we need convergence of Mθi(uθi)\sqrt{M_{\theta_{i}}(u_{\theta_{i}})} in C(0,T;Lp(Ω))C(0,T;L^{p}(\Omega)). Due to the lack of control on tuθi\partial_{t}u_{\theta_{i}}, such convergence can not be derived directly using Aubin-Lions Lemma [7]. Instead, we apply Aubin-Lions lemma to Gi(u)=0ugθi(s)𝑑sG_{i}(u)=\int_{0}^{u}g_{\theta_{i}}(s)ds and derive convergence of gθi(uθi)g_{\theta_{i}}(u_{\theta_{i}}) (consequently on Mθi(uθi)M_{\theta_{i}}(u_{\theta_{i}})) from convergence of GiG_{i} through characterization of compact sets [22] in Lp[0,T;B]L^{p}[0,T;B]. We then follow the idea in [7] to pass to the limit on the right hand side of (85). Finally, we identify an explicit expression of the weak limit of μθigθi(uθi)\nabla\frac{\mu_{\theta_{i}}}{g_{\theta_{i}}(u_{\theta_{i}})} in terms of the weak limit uu under additional integrability assumptions on derivatives of uu.

Theorem 1.2.

For any u0H1(Ω)u_{0}\in H^{1}(\Omega) and T>0T>0, there exists a function u:ΩT=Ω×[0,T]u:\Omega_{T}=\Omega\times[0,T]\rightarrow\mathbb{R} satisfying

  • i)

    uL(0,T;H1(Ω))C([0,T];Ls(Ω))u\in L^{\infty}(0,T;H^{1}(\Omega))\cap C([0,T];L^{s}(\Omega)), where 1s<1\leq s<\infty,

  • ii)

    g(u)tuLp(0,T;(W1,q(Ω)))g(u)\partial_{t}u\in L^{p}(0,T;(W^{1,q}(\Omega))^{\prime}) for 1p<21\leq p<2 and q>2q>2.

  • iii)

    u(x,0)=u0(x)u(x,0)=u_{0}(x) for all xΩx\in\Omega,

which solves (4)-(5) in the following weak sense

  • a)

    There exists a set BΩTB\subset\Omega_{T} with |ΩT\B|=0|\Omega_{T}\backslash B|=0 and a function ζ:ΩTn\zeta:\Omega_{T}\rightarrow\mathbb{R}^{n} satisfying χBPM(u)ζLpp1(0,T;Lqq1(Ω,n))\chi_{B\cap P}M(u)\zeta\in L^{\frac{p}{p-1}}(0,T;L^{\frac{q}{q-1}}(\Omega,\mathbb{R}^{n})) such that

    (13) 0T<g(u)tu,ϕ>((W1,q(Ω)),W1,q(Ω))dt=BPM(u)ζϕdxdt\int_{0}^{T}<g(u)\partial_{t}u,\phi>_{((W^{1,q}(\Omega))^{\prime},W^{1,q}(\Omega))}dt=-\int_{B\cap P}M(u)\zeta\cdot\nabla\phi dxdt

    for all ϕLp(0,T;W1,q(Ω))\phi\in L^{p}(0,T;W^{1,q}(\Omega)) with p,q>2p,q>2. Here P:={(x,t)ΩT:|1u2|0}P:=\{(x,t)\in\Omega_{T}:|1-u^{2}|\neq 0\} is the set where M(u),g(u)M(u),g(u) are nondegenerate and χBP\chi_{B\cap P} is the characteristic function of set BPB\cap P.

  • b)

    Assume uL2(0,T;H2(Ω)).u\in L^{2}(0,T;H^{2}(\Omega)). For any open set UΩTU\in\Omega_{T} on which g(u)>0g(u)>0 and ΔuLp(U)\nabla\Delta u\in L^{p}(U) for some p>1p>1, we have

    (14) ζ=Δu+q′′(u)ug(u)g(u)g2(u)(Δu+q(u))u a.e. in U.\zeta=\frac{-\nabla\Delta u+q^{\prime\prime}(u)\nabla u}{g(u)}-\frac{g^{\prime}(u)}{g^{2}(u)}\left(-\Delta u+q^{\prime}(u)\right)\nabla u\text{ a.e. in }U.

Moreover, the following energy inequality holds for all t>0t>0

(15) Ω(12|u(x,t)|2+q(u(x,t)))𝑑z+ΩrBPM(u(x,τ))|ζ(x,τ)|2𝑑x𝑑τ\displaystyle\int_{\Omega}\left(\frac{1}{2}|\nabla u(x,t)|^{2}+q(u(x,t))\right)dz+\int_{\Omega_{r}\cap B\cap P}M(u(x,\tau))|\zeta(x,\tau)|^{2}dxd\tau
\displaystyle\leq Ω(12|u0(x)|2+q(u0(x)))𝑑x.\displaystyle\int_{\Omega}\left(\frac{1}{2}|\nabla u_{0}(x)|^{2}+q(u_{0}(x))\right)dx.

Adding an additional term 1ε(Δ)12u\frac{1}{\varepsilon}(-\Delta)^{\frac{1}{2}}u to the chemical potential in (2), we can apply the same method to derive existence of weak solutions of the modified model (see section 5 for further details). Such nonlocal model originates from the phase field model for self-climb of dislocation loops [20].

The paper is organized as follows. We shall derive sharp interface limit for (1) and (2) through formal asymptotic expansions in section 2. Section 3 is devoted to the proof of Theorem 1.1 and Theorem1.2 is proved in section 4. Similar existence theorems for the modified model with an additional nonlocal term added to the chemical potential are presented in section 5.

2 Sharp interface limit via asymptotic expansions

In this section, we perform a formal asymptotic analysis to obtain the sharp interface limit of the proposed phase field model (1)- (2)as ε0\varepsilon\rightarrow 0.

2.1 Outer expansions

We first perform expansion in the region far from the dislocations. Assume the expansion for uu is

(16) u(x,y,t)=u(0)(x,y,t)+u(1)(x,y,t)ε+u(2)(x,y,t)ε2+u(x,y,t)=u^{(0)}(x,y,t)+u^{(1)}(x,y,t)\varepsilon+u^{(2)}(x,y,t)\varepsilon^{2}+\ldots

Correspondingly, we have

M(u)=M(u(0))+M(u(0))u(1)ε+(M(u(0))u(2)+12M′′(u(0))(u(1))2)ε2+,\displaystyle M(u)=M(u^{(0)})+M^{\prime}(u^{(0)})u^{(1)}\varepsilon+\left(M^{\prime}(u^{(0)})u^{(2)}+\frac{1}{2}M^{\prime\prime}\left(u^{(0)}\right)\left(u^{(1)}\right)^{2}\right)\varepsilon^{2}+\ldots,
g(u)=g(u(0))+g(u(0))u(1)ε+(g(u(0))u(2)+12g′′(u(0))(u(1))2)ε2+,\displaystyle g(u)=g(u^{(0)})+g^{\prime}(u^{(0)})u^{(1)}\varepsilon+\left(g^{\prime}(u^{(0)})u^{(2)}+\frac{1}{2}g^{\prime\prime}(u^{(0)})\left(u^{(1)}\right)^{2}\right)\varepsilon^{2}+\ldots,
q(u)=q(u(0))+q′′(u(0))u(1)ε+(q′′(u(0))u(2)+12q(3)(u(0))(u(1))2)ε2+,\displaystyle q^{\prime}(u)=q^{\prime}(u^{(0)})+q^{\prime\prime}(u^{(0)})u^{(1)}\varepsilon+\left(q^{\prime\prime}(u^{(0)})u^{(2)}+\frac{1}{2}q^{(3)}(u^{(0)})\left(u^{(1)}\right)^{2}\right)\varepsilon^{2}+\ldots,

We also expand chemical potential μ\mu as

(17) μ=1ε2(μ(0)+μ(1)ε+μ(2)ε2+).\mu=\frac{1}{\varepsilon^{2}}\left(\mu^{(0)}+\mu^{(1)}\varepsilon+\mu^{(2)}\varepsilon^{2}+\ldots\right).

and rewrite equation (1) as

(18) g(u)tu=M0(μμg(u)g(u)u).g(u)\partial_{t}u=M_{0}\nabla\cdot(\nabla\mu-\mu\frac{g^{\prime}(u)}{g(u)}\nabla u).

Set

w=μg(u)g(u)=1ε2(w(0)+w(1)ε+w(2)ε2+).w=-\mu\frac{g^{\prime}(u)}{g(u)}=\frac{1}{\varepsilon^{2}}\left(w^{(0)}+w^{(1)}\varepsilon+w^{(2)}\varepsilon^{2}+\ldots\right).

Plugging the expansions into (18) and (2) and matching the coefficients of ε\varepsilon powers in both equations, the O(1ε2)O(\frac{1}{\varepsilon^{2}}) of (18) and (2) yields

(19) 0=(μ(0)+w(0)u(0))\displaystyle 0=\nabla\cdot\left(\nabla\mu^{(0)}+w^{(0)}\nabla u^{(0)}\right)
(20) μ(0)=q(u(0)).\displaystyle\mu^{(0)}=q^{\prime}(u^{(0)}).

Since

w(0)=μ(0)g(u(0))g(u(0)),w^{(0)}=\mu^{(0)}\frac{g^{\prime}(u^{(0)})}{g(u^{(0)})},

then u(0)=1u^{(0)}=1 or u(0)=1u^{(0)}=-1 satisfies equations (19)-(20). In particular, such choice of u(0)u^{(0)} implies μ(0)=0\mu^{(0)}=0.

The O(1ε)O(\frac{1}{\varepsilon}) equation of (18) and (2) reduces to

(21) 0=(μ(1)+w(0)u(1)+w(1)u(0)),\displaystyle 0=\nabla\cdot\left(\nabla\mu^{(1)}+w^{(0)}\nabla u^{(1)}+w^{(1)}\nabla u^{(0)}\right),
(22) μ(1)=q′′(u(0))u(1).\displaystyle\mu^{(1)}=q^{\prime\prime}(u^{(0)})u^{(1)}.

Since u(0)=1u^{(0)}=1 or 1-1, u(1)=0u^{(1)}=0 satisfies (21)-(22). Moreover, such choice of u(1)u^{(1)} guarantees μ(1)=0\mu^{(1)}=0.

The O(1)O(1) equation of (18) and (2), taking into account of the fact u(0)=±1u^{(0)}=\pm 1, μ(0)=μ(1)=0\mu^{(0)}=\mu^{(1)}=0, reduces to

(23) 0=(μ(2)+w(0)u(2)),\displaystyle 0=\nabla\cdot\left(\nabla\mu^{(2)}+w^{(0)}\nabla u^{(2)}\right),
(24) μ(2)=q′′(u(0))u(2).\displaystyle\mu^{(2)}=q^{\prime\prime}(u^{(0)})u^{(2)}.

Thus u(2)=0u^{(2)}=0 satisfies (23)-(24). Moreover, such choice of u(1)u^{(1)} guarantees μ(2)=0\mu^{(2)}=0.

In general, if u(0)=±1u^{(0)}=\pm 1, u(1)=u(2)==u(k+1)=0u^{(1)}=u^{(2)}=\ldots=u^{(k+1)}=0, the O(εk)O(\varepsilon^{k}) for k1k\geq 1 equation of (18) and (2) yields

(25) 0=(μ(k+2)+w(0)u(k+2))\displaystyle 0=\nabla\cdot\left(\nabla\mu^{(k+2)}+w^{(0)}\nabla u^{(k+2)}\right)
(26) μ(k+2)=q′′(u(0))u(k+2).\displaystyle\mu^{(k+2)}=q^{\prime\prime}(u^{(0)})u^{(k+2)}.

Thus u(k+2)=0u^{(k+2)}=0 satisfies (25) and(26).

In summary, the u=1u=1 or u=1u=-1 in the outer region.

2.2 Inner expansions

For the small inner regions near the dislocations, we introduce local coordinates near the dislocations. Considering a dislocation CC parameterized by arc length parameter ss. We denote a point on the dislocation by 𝐫0(s){\bf r}_{0}(s) with tangent unit vector 𝐭(s)\mathbf{t}(s) and inward normal vector 𝐧(s)\mathbf{n}(s). A point near the dislocation CC is expressed as

𝐫(s,d)=𝐫0(s)+d𝐧(s),{\bf r}(s,d)={\bf r}_{0}(s)+d\mathbf{n}(s),

where dd is the signed distance from point 𝐫{\bf r} to the dislocation. Since the gradients fields are of order O(1ε)O(\frac{1}{\varepsilon}), we introduce ρ=dε\rho=\frac{d}{\varepsilon} and use coordinates (s,ρ)(s,\rho) in the inner region. Under this setting, we write u(x,y,t)=U(s,ρ,t)u(x,y,t)=U(s,\rho,t) and equation (1) can be written as

(27) g(U)(tU+1εvnρU)=M01ερκs(11ερκ(sμμg(U)g(U)sU))\displaystyle\hskip 36.135ptg(U)\left(\partial_{t}U+\frac{1}{\varepsilon}v_{n}\partial_{\rho}U\right)=\frac{M_{0}}{1-\varepsilon\rho\kappa}\partial_{s}\left(\frac{1}{1-\varepsilon\rho\kappa}\left(\partial_{s}\mu-\mu\frac{g^{\prime}(U)}{g(U)}\partial_{s}U\right)\right)
+1ε2M01ερκρ((1ερκ)(ρμμg(U)g(U)ρU)),\displaystyle\hskip 137.31255pt+\frac{1}{\varepsilon^{2}}\frac{M_{0}}{1-\varepsilon\rho\kappa}\partial_{\rho}\left((1-\varepsilon\rho\kappa)\left(\partial_{\rho}\mu-\mu\frac{g^{\prime}(U)}{g(U)}\partial_{\rho}U\right)\right),
(28) μ=11ερκs(11ερκsU)1ε211ερκρ((1ερκ)ρU)+1ε2q(U).\displaystyle\hskip 72.26999pt\mu=-\frac{1}{1-\varepsilon\rho\kappa}\partial_{s}\left(\frac{1}{1-\varepsilon\rho\kappa}\partial_{s}U\right)-\frac{1}{\varepsilon^{2}}\frac{1}{1-\varepsilon\rho\kappa}\partial_{\rho}\left((1-\varepsilon\rho\kappa)\partial_{\rho}U\right)+\frac{1}{\varepsilon^{2}}q^{\prime}(U).

Assume μ\mu takes the same form expansion as (17) and the following expansions hold for UU within dislocation core region:

(29) U(s,ρ,t)=U(0)(ρ)+εU(1)(s,ρ,t)+ε2U(2)+.U(s,\rho,t)=U^{(0)}(\rho)+\varepsilon U^{(1)}(s,\rho,t)+\varepsilon^{2}U^{(2)}+\ldots.

Here we assume the leading order solution U(0)U^{(0)}, which describes the dislocation core profile, remains the same at all points on the dislocation at any time.

Set

W=μg(U)g(U)=1ε2(W(0)+W(1)ε+W(2)ε2+),W=\mu\frac{g^{\prime}(U)}{g(U)}=\frac{1}{\varepsilon^{2}}\left(W^{(0)}+W^{(1)}\varepsilon+W^{(2)}\varepsilon^{2}+\ldots\right),

the leading order for equation (27) and (28) is O(1ε4)O(\frac{1}{\varepsilon^{4}}), which yields

(30) 0=ρ(ρμ(0)W(0)ρU(0)),\displaystyle 0=\partial_{\rho}\left(\partial_{\rho}\mu^{(0)}-W^{(0)}\partial_{\rho}U^{(0)}\right),
(31) μ(0)=ρρU(0)+q(U(0)).\displaystyle\mu^{(0)}=-\partial_{\rho\rho}U^{(0)}+q^{\prime}(U^{(0)}).

Substituting W(0)=μ(0)g(U(0))g(U(0))W^{(0)}=\mu^{(0)}\frac{g^{\prime}(U^{(0)})}{g(U^{(0)})} into (30), we can rewrite (30) as

0=ρ(ρμ(0)μ(0)ρlng(U(0))).0=\partial_{\rho}\left(\partial_{\rho}\mu^{(0)}-\mu^{(0)}\partial_{\rho}\ln g(U^{(0)})\right).

Integrating this equation, we have

(32) ρμ(0)μ(0)ρlng(U(0))=C0(s).\partial_{\rho}\mu^{(0)}-\mu^{(0)}\partial_{\rho}\ln g(U^{(0)})=C_{0}(s).

Since μ(0)=0\mu^{(0)}=0 in the outer region, we must have μ(0)0\mu^{(0)}\rightarrow 0 and ρμ(0)0\partial_{\rho}\mu^{(0)}\rightarrow 0 as ρ±\rho\rightarrow\pm\infty. Therefore C0(s)=0C_{0}(s)=0. Dividing (32) by μ(0)\mu^{(0)} and integrating, we have μ(0)=C~0(s)g(U(0))\mu^{(0)}=\tilde{C}_{0}(s)g(U^{(0)}). Since μ(0)\mu^{(0)} is independent of ss and is 0 in the outer region, we must have C~0(s)=0\tilde{C}_{0}(s)=0. Thus

(33) μ(0)=ρρU(0)+q(U(0))=0.\mu^{(0)}=-\partial_{\rho\rho}U^{(0)}+q^{\prime}(U^{(0)})=0.

Solution U(0)U^{(0)} to (33) subject to far field condition U(0)(+)=1U^{(0)}{(+\infty)}=-1 and U(0)()=1U^{(0)}{(-\infty)}=1 can be found numerically (see [20] for example). In particular, ρU(0)<0\partial_{\rho}U^{(0)}<0 for all ρ\rho.

Next, the O(1ε3)O(\frac{1}{\varepsilon^{3}}) equation of (27) and (28) yields, using μ(0)=0\mu^{(0)}=0, that

(34) 0=ρ(ρμ(1)W(1)ρU(0)),\displaystyle 0=\partial_{\rho}\left(\partial_{\rho}\mu^{(1)}-W^{(1)}\partial_{\rho}U^{(0)}\right),
(35) μ(1)=ρρU(1)+κρU(0)+q′′(U(0))U(1).\displaystyle\mu^{(1)}=-\partial_{\rho\rho}U^{(1)}+\kappa\partial_{\rho}U^{(0)}+q^{\prime\prime}(U^{(0)})U^{(1)}.

When μ(0)=0\mu^{(0)}=0, we have W(1)=μ(1)g(U(0))g(U(0))W^{(1)}=\mu^{(1)}\frac{g^{\prime}(U^{(0)})}{g(U^{(0)})}. Sustituting into (34) and integrating, we have

(36) ρμ(1)μ(1)ρlng(U(0))=C1(s).\partial_{\rho}\mu^{(1)}-\mu^{(1)}\partial_{\rho}\ln g(U^{(0)})=C_{1}(s).

Matching with the outer solutions (ρμ(1),μ(1)0\partial_{\rho}\mu^{(1)},\mu^{(1)}\rightarrow 0 as ρ±\rho\rightarrow\pm\infty), we conclude that C1(s)=0C_{1}(s)=0. Dividing (36) by μ(1)\mu^{(1)} and integrating, we have μ(1)=C~1(s)g(U(0))\mu^{(1)}=\tilde{C}_{1}(s)g(U^{(0)}). Thus (35) can be written as

(37) LU(1)=κρU(0)+C~1(s)g(U(0))LU^{(1)}=-\kappa\partial_{\rho}U^{(0)}+\tilde{C}_{1}(s)g(U^{(0)})

where L=ρρ+q′′(U(0))L=-\partial_{\rho\rho}+q^{\prime\prime}(U^{(0)}) is a linear operator whose kernal is span{ρU(0)}\{\partial_{\rho}U^{(0)}\}. (37) is sovlable iff the right hand side is perpendicular to the kernal of LL, i.e.

+(κρU(0)+C~1(s)g(U(0)))ρU(0)dρ=0.\int_{-\infty}^{+\infty}\left(-\kappa\partial_{\rho}U^{(0)}+\tilde{C}_{1}(s)g(U^{(0)})\right)\partial_{\rho}U^{(0)}d\rho=0.

From this, we conclude

C~1(s)=ακ,\tilde{C}_{1}(s)=\alpha\kappa,

where positive constants α\alpha and β\beta are given by

α=+(ρU(0))2𝑑ρ+g(U(0))ρU(0)dρ<0.\alpha=\frac{\int_{-\infty}^{+\infty}\left(\partial_{\rho}U^{(0)}\right)^{2}d\rho}{\int_{-\infty}^{+\infty}g(U^{(0)})\partial_{\rho}U^{(0)}d\rho}<0.

Therefore

(38) μ(1)=ακg(U(0)).\mu^{(1)}=\alpha\kappa g(U^{(0)}).

Letting μ¯=μg(U)\overline{\mu}=\frac{\mu}{g(U)}, (27) can be written as

(39) g(U)(tU+1εvnρU)\displaystyle g(U)\left(\partial_{t}U+\frac{1}{\varepsilon}v_{n}\partial_{\rho}U\right)
=\displaystyle= M01ερκs(g(U)1ερκ(sμ¯))+1ε2M01ερκρ((1ερκ)g(U)ρμ¯)\displaystyle\frac{M_{0}}{1-\varepsilon\rho\kappa}\partial_{s}\left(\frac{g(U)}{1-\varepsilon\rho\kappa}\left(\partial_{s}\overline{\mu}\right)\right)+\frac{1}{\varepsilon^{2}}\frac{M_{0}}{1-\varepsilon\rho\kappa}\partial_{\rho}\left(\left(1-\varepsilon\rho\kappa\right)g(U)\partial_{\rho}\overline{\mu}\right)

Using μ(0)=0\mu^{(0)}=0, ρμ¯(1)=ρμ(1)g(U(0))=0\partial_{\rho}\overline{\mu}^{(1)}=\partial_{\rho}\frac{\mu^{(1)}}{g(U^{(0)})}=0, the O(1ε2)O(\frac{1}{\varepsilon^{2}}) order equation of (39) reduces to

ρ(g(U(0))ρU(2))=0.\partial_{\rho}\left(g(U^{(0)})\partial_{\rho}U^{(2)}\right)=0.

Integrating with respect to ρ\rho, we have g(U(0))ρU(2)=C2(s)g(U^{(0)})\partial_{\rho}U^{(2)}=C_{2}(s). Matching with outer solutions, we must have C2(s)=0C_{2}(s)=0. Thus ρU(2)=0\partial_{\rho}U^{(2)}=0 which gives U(2)=C~2(s)U^{(2)}=\tilde{C}_{2}(s).

Next we look at the O(1ε)O(\frac{1}{\varepsilon}) equation of (39). Using μ(0)=0\mu^{(0)}=0, ρμ¯(1)=0\partial_{\rho}\overline{\mu}^{(1)}=0 and ρμ¯(2)=0\partial_{\rho}\overline{\mu}^{(2)}=0, we have

g(U(0))vnρU(0)=M0s(g(U(0))sμ¯(1))+M0ρ(g(U(0))ρμ¯(3)).g(U^{(0)})v_{n}\partial_{\rho}U^{(0)}=M_{0}\partial_{s}\left(g(U^{(0)})\partial_{s}\overline{\mu}^{(1)}\right)+M_{0}\partial_{\rho}\left(g(U^{(0)})\partial_{\rho}\overline{\mu}^{(3)}\right).

Integrating this equation with respect to ρ\rho and matching with outer solutions yields

(40) vn=λssμ¯(1)v_{n}=\lambda\partial_{ss}\overline{\mu}^{(1)}

where we used the fact that g(U(0))g(U^{(0)}) is independent of ss and

λ=M0+g(U(0))𝑑ρ+g(U(0))ρU(0)dρ<0.\lambda=\frac{M_{0}\int_{-\infty}^{+\infty}g(U^{(0)})d\rho}{\int_{-\infty}^{+\infty}g(U^{(0)})\partial_{\rho}U^{(0)}d\rho}<0.

By (38), we have μ¯(1)=ακ\overline{\mu}^{(1)}=\alpha\kappa, substitute this into (40), we obtain the sharp interface limit equation

(41) vn=λss(ακ).v_{n}=\lambda\partial_{ss}\left(\alpha\kappa\right).
Remark 2.1.

Notice here the outer and inner expansions are similar to the expansions in [20]. We wrote out all details here for readers’ convenience.

3 Weak solution for phase field model with positive mobilities

In this section we prove existence of weak solutions for phase field model with positive mobilities. Let +\mathbb{Z}_{+} be the set of nonnegative integers and Ω=[0,2π]n\Omega=[0,2\pi]^{n} with n2n\leq 2. We pick an orthonormal basis for L2(Ω)L^{2}(\Omega) as

{ϕj:j=1,2,}={(2π)n/2,Re(πn/2eiξx),Im(πn/2eiξx):ξ+n\{0,,0}}.\displaystyle\{\phi_{j}:j=1,2,\ldots\}=\left\{(2\pi)^{-{n/2}},\text{Re}\left(\pi^{-n/2}e^{i\xi\cdot x}\right),\text{Im}\left(\pi^{-n/2}e^{i\xi\cdot x}\right):\xi\in\mathbb{Z}^{n}_{+}\backslash\{0,\ldots,0\}\right\}.

Observe {ϕj}\{\phi_{j}\} is also orthogonal in Hk(Ω)H^{k}(\Omega) for any k1k\geq 1. Here and throughout the paper, we denote ΩT=(0,T)×Ω\Omega_{T}=(0,T)\times\Omega.

3.1 Galerkin approximations

Define

uN(x,t)=j=1NcjN(t)ϕj(x),μN(x,t)=j=1NdjN(t)ϕj(x),u^{N}(x,t)=\sum_{j=1}^{N}c_{j}^{N}(t)\phi_{j}(x),\hskip 36.135pt\mu^{N}(x,t)=\sum_{j=1}^{N}d^{N}_{j}(t)\phi_{j}(x),

where {cjN,djN}\{c_{j}^{N},d^{N}_{j}\} satisfy

(42) ΩtuNϕjdx\displaystyle\int_{\Omega}\partial_{t}u^{N}\phi_{j}dx =\displaystyle= ΩMθ(uN)μNgθ(uN)ϕjgθ(uN)dx,\displaystyle-\int_{\Omega}M_{\theta}(u^{N})\nabla\frac{\mu^{N}}{g_{\theta}(u^{N})}\cdot\nabla\frac{\phi_{j}}{g_{\theta}(u^{N})}dx,
(43) ΩμNϕj𝑑x\displaystyle\int_{\Omega}\mu^{N}\phi_{j}dx =\displaystyle= Ω(uNϕj+q(uN)ϕj)𝑑x,\displaystyle\int_{\Omega}\left(\nabla u^{N}\cdot\nabla\phi_{j}+q^{\prime}(u^{N})\phi_{j}\right)dx,
(44) uN(x,0)\displaystyle u^{N}(x,0) =\displaystyle= j=1N(Ωu0ϕj𝑑x)ϕj(x).\displaystyle\sum_{j=1}^{N}\left(\int_{\Omega}u_{0}\phi_{j}dx\right)\phi_{j}(x).

(42)-(44) is an initial value problem for a system of ordinary equations for {cjN(t)}\{c_{j}^{N}(t)\}. Since right hand side of (42) is continuous in cjNc_{j}^{N}, the system has a local solution.

Define energy functional

E(u)=Ω{12|u|2+q(u)}𝑑x.E(u)=\int_{\Omega}\left\{\frac{1}{2}|\nabla u|^{2}+q(u)\right\}dx.

Direct calculation yields

ddtE(uN(x,t))=ΩMθ(uN)|μNgθ(uN)|2𝑑x,\frac{d}{dt}E(u^{N}(x,t))=-\int_{\Omega}M_{\theta}(u^{N})\left|\nabla\frac{\mu^{N}}{g_{\theta}(u^{N})}\right|^{2}dx,

integration over tt gives the following energy identity.

Ω(12|uN(x,t)|2+q(uN(x,t)))𝑑x\displaystyle\int_{\Omega}\left(\frac{1}{2}|\nabla u^{N}(x,t)|^{2}+q(u^{N}(x,t))\right)dx
+0tΩMθ(uN(x,τ))|μN(x,τ)gθ(uN(x,τ))|2𝑑x𝑑τ\displaystyle+\int_{0}^{t}\int_{\Omega}M_{\theta}(u^{N}(x,\tau))\left|\nabla\frac{\mu^{N}(x,\tau)}{g_{\theta}(u^{N}(x,\tau))}\right|^{2}dxd\tau
=\displaystyle= Ω(12|uN(x,0)|2+q(uN(x,0)))𝑑x\displaystyle\int_{\Omega}\left(\frac{1}{2}|\nabla u^{N}(x,0)|^{2}+q(u^{N}(x,0))\right)dx
\displaystyle\leq u0L2(Ω)2+C(u0H1Ωr+1+|Ω|)C<\displaystyle\left\Arrowvert\nabla u_{0}\right\Arrowvert_{L^{2}(\Omega)}^{2}+C\left(\left\Arrowvert u_{0}\right\Arrowvert^{r+1}_{H^{1}{\Omega}}+|\Omega|\right)\leq C<\infty

Here and throughout the paper, CC represents a generic constant possibly depending only on TT, Ω\Omega, u0u_{0} but not on θ\theta. Since Ω\Omega is bounded region, by growth assumption assumption (6) and Poincare’s inequality, the energy identity (3.1) implies uNL(0,T;H1(Ω))u^{N}\in L^{\infty}(0,T;H^{1}(\Omega)) with

(46) uNL(0,T;H1(Ω))C for all N,\left\Arrowvert u^{N}\right\Arrowvert_{L^{\infty}(0,T;H^{1}(\Omega))}\leq C\text{ for all }N,

and

(47) Mθ(uN)μNgθ(uN)L2(ΩT)C for all N.\left\Arrowvert\sqrt{M_{\theta}(u^{N})}\nabla\frac{\mu^{N}}{g_{\theta}(u^{N})}\right\Arrowvert_{L^{2}(\Omega_{T})}\leq C\text{ for all }N.

By (46), the coefficients {cjN(t)}\{c_{j}^{N}(t)\} are bounded in time, thus the system (42)-(44) has a global solution. In addition, by Sobolev embedding theorem and growth assumption (7) on q(u)q^{\prime}(u), we have

q(uN)L(0,T;Lp(Ω)),Mθ(uN)L(0,T;Lp(Ω))q^{\prime}(u^{N})\in L^{\infty}(0,T;L^{p}(\Omega)),\hskip 36.135ptM_{\theta}(u^{N})\in L^{\infty}(0,T;L^{p}(\Omega))

for any 1p<1\leq p<\infty with

(48) q(uN)L(0,T;Lp(Ω))C for all N,\displaystyle\left\Arrowvert q^{\prime}(u^{N})\right\Arrowvert_{L^{\infty}(0,T;L^{p}(\Omega))}\leq C\text{ for all }N,
(49) Mθ(uN)L(0,T;Lp(Ω))C for all N.\displaystyle\left\Arrowvert M_{\theta}(u^{N})\right\Arrowvert_{L^{\infty}(0,T;L^{p}(\Omega))}\leq C\text{ for all }N.

3.2 Convergence of uNu^{N}

Given q>2q>2 and any ϕL2(0,T;W1,q(Ω))\phi\in L^{2}(0,T;W^{1,q}(\Omega)), let ΠNϕ(x,t)=j=1N(Ωϕ(x,t)ϕj(x)𝑑x)ϕj(x)\Pi_{N}\phi(x,t)=\sum_{j=1}^{N}\left(\int_{\Omega}\phi(x,t)\phi_{j}(x)dx\right)\phi_{j}(x) be the orthogonal projection of ϕ\phi onto span{ϕj}j=1N\{\phi_{j}\}_{j=1}^{N}. Then

|ΩtuNϕdx|=|ΩtuNΠNϕdx|=|ΩMθ(uN)μNgθ(uN)ΠNϕgθ(uN)dx|\displaystyle\left\arrowvert\int_{\Omega}\partial_{t}u^{N}\phi dx\right\arrowvert=\left\arrowvert\int_{\Omega}\partial_{t}u^{N}\Pi_{N}\phi dx\right\arrowvert=\left\arrowvert\int_{\Omega}M_{\theta}(u^{N})\nabla\frac{\mu^{N}}{g_{\theta}(u^{N})}\cdot\nabla\frac{\Pi_{N}\phi}{g_{\theta}(u^{N})}dx\right\arrowvert
\displaystyle\leq (ΩMθ(uN)|μNgθ(uN)|2𝑑x)12(ΩMθ(uN)|ΠNϕgθ(uN)|2𝑑x)12.\displaystyle\left(\int_{\Omega}M_{\theta}(u^{N})\left\arrowvert\nabla\frac{\mu^{N}}{g_{\theta}(u^{N})}\right\arrowvert^{2}dx\right)^{\frac{1}{2}}\left(\int_{\Omega}M_{\theta}(u^{N})\left\arrowvert\nabla\frac{\Pi_{N}\phi}{g_{\theta}(u^{N})}\right\arrowvert^{2}dx\right)^{\frac{1}{2}}.

Since

ΠNϕgθ(uN)=1gθ(uN)ΠNϕΠNϕgθ(uN)gθ2(uN)uN,\nabla\frac{\Pi_{N}\phi}{g_{\theta}(u^{N})}=\frac{1}{g_{\theta}(u^{N})}\nabla\Pi_{N}\phi-\Pi_{N}\phi\frac{g_{\theta}^{\prime}(u^{N})}{g^{2}_{\theta}(u^{N})}\nabla u^{N},

we have

ΩMθ(uN)|ΠNϕgθ(uN)|2𝑑x\displaystyle\int_{\Omega}M_{\theta}(u^{N})\left\arrowvert\nabla\frac{\Pi_{N}\phi}{g_{\theta}(u^{N})}\right\arrowvert^{2}dx
\displaystyle\leq 2M0Ω(1gθ(uN)|ΠNϕ|2+|gθ(uN)|2gθ3(uN)|ΠNϕ|2|uN|2)𝑑x\displaystyle 2M_{0}\int_{\Omega}\left(\frac{1}{g_{\theta}(u^{N})}\left\arrowvert\nabla\Pi_{N}\phi\right\arrowvert^{2}+\frac{|g^{\prime}_{\theta}(u^{N})|^{2}}{g^{3}_{\theta}(u^{N})}|\Pi_{N}\phi|^{2}|\nabla u^{N}|^{2}\right)dx
\displaystyle\leq C(M0,θ)(ΠNϕL2(Ω)2+ΠNϕL(Ω)2uNL2(Ω)2) here is where we need q>2\displaystyle C(M_{0},\theta)\left(\left\Arrowvert\nabla\Pi_{N}\phi\right\Arrowvert^{2}_{L^{2}(\Omega)}+\left\Arrowvert\Pi_{N}\phi\right\Arrowvert^{2}_{L^{\infty}(\Omega)}\left\Arrowvert\nabla u^{N}\right\Arrowvert^{2}_{L^{2}(\Omega)}\right)\text{ here is where we need }q>2
\displaystyle\leq C(M0,θ)(ΠNϕW1,q(Ω)2)C(M0,θ)ϕW1,q(Ω)2.\displaystyle C(M_{0},\theta)\left(\left\Arrowvert\Pi_{N}\phi\right\Arrowvert^{2}_{W^{1,q}(\Omega)}\right)\leq C(M_{0},\theta)\left\Arrowvert\phi\right\Arrowvert^{2}_{W^{1,q}(\Omega)}.

Therefore

(50) tuNL2(0,T;(W1,q(Ω)))C(M0,θ) for all N.\left\Arrowvert\partial_{t}u^{N}\right\Arrowvert_{L^{2}(0,T;(W^{1,q}(\Omega))^{\prime})}\leq C(M_{0},\theta)\text{ for all }N.

For 1s<1\leq s<\infty, since n2n\leq 2, by Sobolev embedding theorem and Aubin-Lions Lemma (see [22] and Remark 3.1) , the following embeddings are compact :

{fL2(0,T;H1(Ω)):tfL2(0,T;(W1,q(Ω))}L2(0,T;Ls(Ω)),\left\{f\in L^{2}(0,T;H^{1}(\Omega)):\partial_{t}f\in L^{2}(0,T;(W^{1,q}(\Omega))^{\prime}\right\}\hookrightarrow L^{2}(0,T;L^{s}(\Omega)),

and

{fL(0,T;H1(Ω)):tfL2(0,T;(W1,q(Ω))}C([0,T];Ls(Ω)).\left\{f\in L^{\infty}(0,T;H^{1}(\Omega)):\partial_{t}f\in L^{2}(0,T;(W^{1,q}(\Omega))^{\prime}\right\}\hookrightarrow C([0,T];L^{s}(\Omega)).

From this and the boundedness of {uN}\{u^{N}\} and {tuN}\{\partial_{t}u^{N}\}, we can find a subsequence, and uθL(0,T;H1(Ω))u_{\theta}\in L^{\infty}(0,T;H^{1}(\Omega)) such that as NN\rightarrow\infty,

(51) uN\displaystyle u^{N} \displaystyle\rightharpoonup uθ weakly-* in L(0,T;H1(Ω)),\displaystyle u_{\theta}\text{ weakly-* in }L^{\infty}(0,T;H^{1}(\Omega)),
(52) uN\displaystyle u^{N} \displaystyle\rightarrow uθ strongly in C([0,T];Ls(Ω)),\displaystyle u_{\theta}\text{ strongly in }C([0,T];L^{s}(\Omega)),
(53) uN\displaystyle u^{N} \displaystyle\rightarrow uθ strongly in L2(0,T;Ls(Ω)) and a.e. in ΩT,\displaystyle u_{\theta}\text{ strongly in }L^{2}(0,T;L^{s}(\Omega))\text{ and a.e. in }\Omega_{T},
(54) tuN\displaystyle\partial_{t}u^{N} \displaystyle\rightharpoonup tuθ weakly in L2(0,T;(W1,q(Ω)))\displaystyle\partial_{t}u_{\theta}\text{ weakly in }L^{2}(0,T;(W^{1,q}(\Omega))^{\prime})

for 1s<1\leq s<\infty. In addition

uθL(0,T;H1(Ω))C,tuθL2(0,T;(W1,q(Ω)))C(M0,θ).\left\Arrowvert u_{\theta}\right\Arrowvert_{L^{\infty}(0,T;H^{1}(\Omega))}\leq C,\hskip 36.135pt\left\Arrowvert\partial_{t}u_{\theta}\right\Arrowvert_{L^{2}(0,T;(W^{1,q}(\Omega))^{\prime})}\leq C(M_{0},\theta).

By (52), growth assumption (7) on q(uN)q^{\prime}(u^{N}), and general dominated convergence Theorem, we have for any 1s<1\leq s<\infty,

(55) Mθ(uN)Mθ(uθ) strongly in C([0,T];Ls(Ω)),\displaystyle M_{\theta}(u^{N})\rightarrow M_{\theta}(u_{\theta})\text{ strongly in }C([0,T];L^{s}(\Omega)),
(56) Mθ(uN)Mθ(uθ) strongly in C([0,T];Ls(Ω)),\displaystyle\sqrt{M_{\theta}(u^{N})}\rightarrow\sqrt{M_{\theta}(u_{\theta})}\text{ strongly in }C([0,T];L^{s}(\Omega)),
(57) q(uN)q(uθ) strongly in C([0,T];Ls(Ω)).\displaystyle q^{\prime}(u^{N})\rightarrow q^{\prime}(u_{\theta})\text{ strongly in }C([0,T];L^{s}(\Omega)).

By (48) and (57), we have

(58) q(uN)q(uθ) weakly-* in L([0,T];Ls(Ω)).q^{\prime}(u^{N})\rightharpoonup q^{\prime}(u_{\theta})\text{ weakly-* in }L^{\infty}([0,T];L^{s}(\Omega)).
Remark 3.1.

Let XX, YY, ZZ be Banach spaces with compact embedding XYX\hookrightarrow Y and continuous embedding YZY\hookrightarrow Z. Then the embeddings

(59) {fLp(0,T;X);tfL1(0,T;Z)}Lp(0,T;Y)\{f\in L^{p}(0,T;X);\partial_{t}f\in L^{1}(0,T;Z)\}\hookrightarrow L^{p}(0,T;Y)

and

(60) {fL(0,T;X);tfLr(0,T;Z)}C([0,T];Y)\{f\in L^{\infty}(0,T;X);\partial_{t}f\in L^{r}(0,T;Z)\}\hookrightarrow C([0,T];Y)

are compact for any 1p<1\leq p<\infty and r>1r>1 (Corollary 4, [22], see also [16]) . For convergence of uNu^{N}, we apply this for p=2=rp=2=r with X=H1(Ω)X=H^{1}(\Omega), Y=Ls(Ω)Y=L^{s}(\Omega) for 1s<1\leq s<\infty and Z=W1,q(Ω)Z=W^{1,q}(\Omega)^{\prime}.

3.3 Weak solution

By (47) and the lower bound on MθM_{\theta}, we have

(61) μNgθ(uN)L2(ΩT)Cθm2.\left\Arrowvert\nabla\frac{\mu^{N}}{g_{\theta}(u^{N})}\right\Arrowvert_{L^{2}(\Omega_{T})}\leq C\theta^{-\frac{m}{2}}.

By (43), (46) and (48), we have

|ΩμNϕ1gθ(uN)|dx=|ΩμNΠN(ϕ1gθ(uN))|dx\displaystyle\left\arrowvert\int_{\Omega}\frac{\mu^{N}\phi_{1}}{g_{\theta}(u_{N})}\right\arrowvert dx=\left\arrowvert\int_{\Omega}\mu^{N}\Pi_{N}\left(\frac{\phi_{1}}{g_{\theta}(u^{N})}\right)\right\arrowvert dx
=\displaystyle= |ΩuNΠN(ϕ1gθ(uN))𝑑x+Ωq(uN)ΠN(ϕ1gθ(uN))𝑑x|\displaystyle\left\arrowvert\int_{\Omega}\nabla u^{N}\cdot\nabla\Pi_{N}\left(\frac{\phi_{1}}{g_{\theta}(u^{N})}\right)dx+\int_{\Omega}q^{\prime}(u^{N})\Pi_{N}\left(\frac{\phi_{1}}{g_{\theta}(u^{N})}\right)dx\right\arrowvert
=\displaystyle= |ΩuN(ϕ1gθ(uN))dx+Ωq(uN)ΠN(ϕ1gθ(uN))𝑑x|\displaystyle\left\arrowvert\int_{\Omega}\nabla u^{N}\cdot\nabla\left(\frac{\phi_{1}}{g_{\theta}(u^{N})}\right)dx+\int_{\Omega}q^{\prime}(u^{N})\Pi_{N}\left(\frac{\phi_{1}}{g_{\theta}(u^{N})}\right)dx\right\arrowvert
\displaystyle\leq C(θm1uNL2(Ω)2+θmq(uN)L2(Ω)ϕ1L2(Ω))\displaystyle C\left(\theta^{-m-1}\left\Arrowvert\nabla u^{N}\right\Arrowvert_{L^{2}(\Omega)}^{2}+\theta^{-m}\left\Arrowvert q^{\prime}(u^{N})\right\Arrowvert_{L^{2}(\Omega)}\left\Arrowvert\phi_{1}\right\Arrowvert_{L^{2}(\Omega)}\right)
\displaystyle\leq Cθm1.\displaystyle C\theta^{-m-1}.

(61),(3.3) and Poincare’s inequality yield

μNgθ(uN)L2(0,T;H1(Ω))C(θm1+1).\left\Arrowvert\frac{\mu^{N}}{g_{\theta}(u^{N})}\right\Arrowvert_{L^{2}(0,T;H^{1}(\Omega))}\leq C(\theta^{-m-1}+1).

Thus there exists a wθL2(0,T;H1(Ω))w_{\theta}\in L^{2}(0,T;H^{1}(\Omega)) and a subsequence of μNgθ(uN)\frac{\mu^{N}}{g_{\theta}(u^{N})}, not relabeled, such that

(63) μNgθ(uN)wθ weakly in L2(0,T;H1(Ω)).\frac{\mu^{N}}{g_{\theta}(u^{N})}\rightharpoonup w_{\theta}\text{ weakly in }L^{2}(0,T;H^{1}(\Omega)).

Therefore by (55), (63) and Sobolev embedding theorem, we have

(64) μN=gθ(uN)μNgθ(uN)μθ=gθ(uθ)wθ weakly in L2(0,T;W1,s(Ω))\mu^{N}=g_{\theta}(u^{N})\cdot\frac{\mu^{N}}{g_{\theta}(u^{N})}\rightharpoonup\mu_{\theta}=g_{\theta}(u_{\theta})w_{\theta}\text{ weakly in }L^{2}(0,T;W^{1,s}(\Omega))

for any 1s<21\leq s<2. Combining (56), (63)and (64), we have

(65) Mθ(uN)μNgθ(uN)Mθ(uθ)μθgθ(uθ) weakly in L2(0,T;Lq(Ω))\sqrt{M_{\theta}(u^{N})}\nabla\frac{\mu^{N}}{g_{\theta}(u^{N})}\rightharpoonup\sqrt{M_{\theta}(u_{\theta})}\nabla\frac{\mu_{\theta}}{g_{\theta}(u_{\theta})}\text{ weakly in }L^{2}(0,T;L^{q}(\Omega))

for any 1q<21\leq q<2. By (47), we can improve this convergence to

(66) Mθ(uN)μNgθ(uN)Mθ(uθ)μθgθ(uθ) weakly in L2(0,T;L2(Ω)).\sqrt{M_{\theta}(u^{N})}\nabla\frac{\mu^{N}}{g_{\theta}(u^{N})}\rightharpoonup\sqrt{M_{\theta}(u_{\theta})}\nabla\frac{\mu_{\theta}}{g_{\theta}(u_{\theta})}\text{ weakly in }L^{2}(0,T;L^{2}(\Omega)).

By (43), we have

ΩμNuN𝑑x=Ω(|uN|2dx+q(uN)uN)𝑑x.\int_{\Omega}\mu^{N}u^{N}dx=\int_{\Omega}\left(|\nabla u^{N}|^{2}dx+q^{\prime}(u^{N})u^{N}\right)dx.

Integrating with respect to tt from 0 to TT, we have on ΩT=Ω×[0,T]\Omega_{T}=\Omega\times[0,T],

ΩTμN(x,τ)uN(x,τ)dxdτ=ΩT(uN(x,τ)|2dx+q(uN(x,τ))uN(x,τ))dxdτ.\displaystyle\int_{\Omega_{T}}\mu^{N}(x,\tau)u^{N}(x,\tau)dxd\tau=\int_{\Omega_{T}}\left(\nabla u^{N}(x,\tau)|^{2}dx+q^{\prime}(u^{N}(x,\tau))u^{N}(x,\tau)\right)dxd\tau.

Passing to the limit in the equation above, by (53), (57) and (64), we have

(67) ΩTμθuθ𝑑x𝑑τ=limNΩT|uN|2𝑑x𝑑τ+ΩTq(uθ)uθ𝑑x𝑑τ\displaystyle\int_{\Omega_{T}}\mu_{\theta}u_{\theta}dxd\tau=\lim_{N\rightarrow\infty}\int_{\Omega_{T}}|\nabla u^{N}|^{2}dxd\tau+\int_{\Omega_{T}}q^{\prime}(u_{\theta})u_{\theta}dxd\tau

On the other hand,

ΩTμN(x,τ)uθ(x,τ)𝑑x𝑑τ=ΩTμN(x,τ)ΠNuθ(x,τ)𝑑x𝑑τ\displaystyle\hskip 14.45377pt\int_{\Omega_{T}}\mu^{N}(x,\tau)u_{\theta}(x,\tau)dxd\tau=\int_{\Omega_{T}}\mu^{N}(x,\tau)\Pi_{N}u_{\theta}(x,\tau)dxd\tau
=\displaystyle= ΩT(uNΠNuθ(x,τ)+q(uN)ΠNuθ(x,τ))𝑑x𝑑τ\displaystyle\int_{\Omega_{T}}\left(\nabla u^{N}\cdot\nabla\Pi_{N}u_{\theta}(x,\tau)+q^{\prime}(u^{N})\Pi_{N}u_{\theta}(x,\tau)\right)dxd\tau
=\displaystyle= ΩT(uNuθ(x,τ)+q(uN)ΠNuθ(x,τ))𝑑x𝑑τ.\displaystyle\int_{\Omega_{T}}\left(\nabla u^{N}\cdot\nabla u_{\theta}(x,\tau)+q^{\prime}(u^{N})\Pi_{N}u_{\theta}(x,\tau)\right)dxd\tau.

Since ΠNuθuθ\Pi_{N}u_{\theta}\rightarrow u_{\theta} strongly in L2(ΩT)L^{2}(\Omega_{T}), by (51),(58) and (64), as NN\rightarrow\infty, (3.3) yields

(69) ΩTμθuθdxdτ=ΩT(|uθ|2+q(uθ))uθ)dxdτ.\int_{\Omega_{T}}\mu_{\theta}u_{\theta}dxd\tau=\int_{\Omega_{T}}\left(|\nabla u_{\theta}|^{2}+q^{\prime}(u_{\theta}))u_{\theta}\right)dxd\tau.

(67) and (69) gives

(70) limNΩT|uN|2𝑑x𝑑τ=ΩT|uθ|2𝑑x𝑑τ.\lim_{N\rightarrow\infty}\int_{\Omega_{T}}|\nabla u^{N}|^{2}dxd\tau=\int_{\Omega_{T}}|\nabla u_{\theta}|^{2}dxd\tau.

By (46), uNuθ\nabla u^{N}\rightharpoonup\nabla u_{\theta} weakly in L2(ΩT)L^{2}(\Omega_{T}), thus (70) implies

(71) uNuθ strongly in L2(ΩT).\nabla u^{N}\rightarrow\nabla u_{\theta}\text{ strongly in }L^{2}(\Omega_{T}).

Since gθθmg_{\theta}\geq\theta^{m}, (53) implies

(72) g(uN)gθ32(uN)gθ(uθ)gθ32(uθ) a.e in ΩT.\frac{g^{\prime}(u^{N})}{g_{\theta}^{\frac{3}{2}}(u^{N})}\rightarrow\frac{g^{\prime}_{\theta}(u_{\theta})}{g_{\theta}^{\frac{3}{2}}(u_{\theta})}\text{ a.e in }\Omega_{T}.

In addition, g(uN)gθ32(uN)\frac{g^{\prime}(u^{N})}{g_{\theta}^{\frac{3}{2}}(u^{N})} is bounded by

(73) |g(uN)gθ32(uN)|Cθ1m2.\left\arrowvert\frac{g^{\prime}(u^{N})}{g_{\theta}^{\frac{3}{2}}(u^{N})}\right\arrowvert\leq C\theta^{-1-\frac{m}{2}}.

It follows from (71), (72), (73) and generalized dominated convergence theorem (see Remark 3.2) that

(74) g(uN)gθ32(uN)uNgθ(uθ)gθ32(uθ)uθ strongly in L2(ΩT).\frac{g^{\prime}(u^{N})}{g_{\theta}^{\frac{3}{2}}(u^{N})}\nabla u^{N}\rightarrow\frac{g^{\prime}_{\theta}(u_{\theta})}{g_{\theta}^{\frac{3}{2}}(u_{\theta})}\nabla u_{\theta}\text{ strongly in }L^{2}(\Omega_{T}).

Let

fN(t)=g(uN(x,t))gθ32(uN(x,t))uN(x,t)gθ(uθ(x,t))gθ32(uθ(x,t))uθ(x,t)L2(Ω),f^{N}(t)=\left\Arrowvert\frac{g^{\prime}(u^{N}(x,t))}{g_{\theta}^{\frac{3}{2}}(u^{N}(x,t))}\nabla u^{N}(x,t)-\frac{g^{\prime}_{\theta}(u_{\theta}(x,t))}{g_{\theta}^{\frac{3}{2}}(u_{\theta}(x,t))}\nabla u_{\theta}(x,t)\right\Arrowvert_{L^{2}(\Omega)},

by (74), we can extract a subsequence of fNf^{N}, not relabeled, such that fN(t)0f^{N}(t)\rightarrow 0 a.e. in (0,T). By Egorov’s theorem, for any given δ>0\delta>0, there exists Tδ[0,T]T_{\delta}\subset[0,T] with |Tδ|<δ|T_{\delta}|<\delta such that fN(t)f^{N}(t) converges to 0 uniformly on [0,T]\Tδ[0,T]\backslash T_{\delta}.

Given α(t)L2(0,T)\alpha(t)\in L^{2}(0,T), for any ε>0\varepsilon>0, there exists Tδ[0,T]T_{\delta}\subset[0,T] with |Tδ|<δ|T_{\delta}|<\delta such that

(75) Tδα2(t)𝑑t<ε.\int_{T_{\delta}}\alpha^{2}(t)dt<\varepsilon.

Multiplying (42) by α(t)\alpha(t) and integrating in time yield

0Tα(t)ΩtuNϕjdxdt=0Tα(t)ΩMθ(uN)μNgθ(uN)ϕjgθ(uN)dxdt\displaystyle\int_{0}^{T}\alpha(t)\int_{\Omega}\partial_{t}u^{N}\phi_{j}dxdt=\int_{0}^{T}\alpha(t)\int_{\Omega}M_{\theta}(u^{N})\nabla\frac{\mu^{N}}{g_{\theta}(u^{N})}\cdot\nabla\frac{\phi_{j}}{g_{\theta}(u^{N})}dxdt
=\displaystyle= ΩTM0α(t)μNgθ(uN)ϕjdxdt\displaystyle\int_{\Omega_{T}}M_{0}\alpha(t)\nabla\frac{\mu^{N}}{g_{\theta}(u^{N})}\cdot\nabla\phi_{j}dxdt
ΩTα(t)M0ϕjgθ(uN)gθ32(uN)uNMθ(uN)μNgθ(uN)dxdt\displaystyle-\int_{\Omega_{T}}\alpha(t)\sqrt{M_{0}}\phi_{j}\frac{g^{\prime}_{\theta}(u^{N})}{g^{\frac{3}{2}}_{\theta}(u^{N})}\nabla u^{N}\cdot\sqrt{M_{\theta}(u^{N})}\nabla\frac{\mu^{N}}{g_{\theta}(u^{N})}dxdt
=III.\displaystyle=I-II.

Since α(t)ϕjL2(0,T;L2(Ω))\alpha(t)\nabla\phi_{j}\in L^{2}(0,T;L^{2}(\Omega)), by (63) and (64), we have

(77) I=ΩTM0α(t)μNgθ(uN)ϕjdxdtΩTM0α(t)μθgθ(uθ)ϕjdxdt.I=\int_{\Omega_{T}}M_{0}\alpha(t)\nabla\frac{\mu^{N}}{g_{\theta}(u^{N})}\cdot\nabla\phi_{j}dxdt\rightarrow\int_{\Omega_{T}}M_{0}\alpha(t)\nabla\frac{\mu_{\theta}}{g_{\theta}(u_{\theta})}\cdot\nabla\phi_{j}dxdt.

To prove convergence on IIII, observe

ΩTα(t)ϕjgθ(uN)gθ32(uN)uNMθ(uN)μNgθ(uN)dxdt\displaystyle\int_{\Omega_{T}}\alpha(t)\phi_{j}\frac{g^{\prime}_{\theta}(u^{N})}{g^{\frac{3}{2}}_{\theta}(u^{N})}\nabla u^{N}\sqrt{M_{\theta}(u^{N})}\nabla\frac{\mu^{N}}{g_{\theta}(u^{N})}dxdt
ΩTα(t)ϕjgθ(uθ)gθ32(uθ)uθMθ(uθ)μθgθ(uθ)dxdt\displaystyle-\int_{\Omega_{T}}\alpha(t)\phi_{j}\frac{g^{\prime}_{\theta}(u_{\theta})}{g^{\frac{3}{2}}_{\theta}(u_{\theta})}\nabla u_{\theta}\sqrt{M_{\theta}(u_{\theta})}\nabla\frac{\mu_{\theta}}{g_{\theta}(u_{\theta})}dxdt
=\displaystyle= ΩTα(t)ϕj(gθ(uN)gθ32(uN)uNgθ(uθ)gθ32(uθ)uθ)Mθ(uN)μNgθ(uN)dxdt\displaystyle\int_{\Omega_{T}}\alpha(t)\phi_{j}\left(\frac{g^{\prime}_{\theta}(u^{N})}{g^{\frac{3}{2}}_{\theta}(u^{N})}\nabla u^{N}-\frac{g^{\prime}_{\theta}(u_{\theta})}{g^{\frac{3}{2}}_{\theta}(u_{\theta})}\nabla u_{\theta}\right)\cdot\sqrt{M_{\theta}(u^{N})}\nabla\frac{\mu^{N}}{g_{\theta}(u^{N})}dxdt
+ΩTα(t)ϕjgθ(uθ)gθ32(uθ)uθ(Mθ(uN)μNgθ(uN)Mθ(uθ)μθgθ(uθ))𝑑x𝑑t\displaystyle+\int_{\Omega_{T}}\alpha(t)\phi_{j}\frac{g^{\prime}_{\theta}(u_{\theta})}{g^{\frac{3}{2}}_{\theta}(u_{\theta})}\nabla u_{\theta}\cdot\left(\sqrt{M_{\theta}(u^{N})}\nabla\frac{\mu^{N}}{g_{\theta}(u^{N})}-\sqrt{M_{\theta}(u_{\theta})}\nabla\frac{\mu_{\theta}}{g_{\theta}(u_{\theta})}\right)dxdt
=\displaystyle= II1+II2\displaystyle II_{1}+II_{2}

From bound

ΩT|α(t)ϕjgθ(uθ)gθ32(uθ)uθ|2𝑑x𝑑t\displaystyle\int_{\Omega_{T}}\left\arrowvert\alpha(t)\phi_{j}\frac{g^{\prime}_{\theta}(u_{\theta})}{g^{\frac{3}{2}}_{\theta}(u_{\theta})}\nabla u_{\theta}\right\arrowvert^{2}dxdt
\displaystyle\leq Cθ2muθL(0,T;L2(Ω))20Tα2(t)2𝑑t,\displaystyle C\theta^{-2-m}\left\Arrowvert\nabla u_{\theta}\right\Arrowvert_{L^{\infty}(0,T;L^{2}(\Omega))}^{2}\int_{0}^{T}\alpha^{2}(t)^{2}dt,

we conclude that α(t)ϕjgθ(uθ)gθ32(uθ)uθL2(ΩT)\alpha(t)\phi_{j}\frac{g^{\prime}_{\theta}(u_{\theta})}{g^{\frac{3}{2}}_{\theta}(u_{\theta})}\nabla u_{\theta}\in L^{2}(\Omega_{T}). By (66), we can pass to the limit in II2II_{2} and conclude

II2=ΩTα(t)ϕjgθ(uθ)gθ32(uθ)uθ(Mθ(uN)μNgθ(uN)Mθ(uθ)μθgθ(uθ))𝑑x𝑑t0.II_{2}=\int_{\Omega_{T}}\alpha(t)\phi_{j}\frac{g^{\prime}_{\theta}(u_{\theta})}{g^{\frac{3}{2}}_{\theta}(u_{\theta})}\nabla u_{\theta}\cdot\left(\sqrt{M_{\theta}(u^{N})}\nabla\frac{\mu^{N}}{g_{\theta}(u^{N})}-\sqrt{M_{\theta}(u_{\theta})}\nabla\frac{\mu_{\theta}}{g_{\theta}(u_{\theta})}\right)dxdt\rightarrow 0.

To pass to the limit in II1II_{1}, we write

II1\displaystyle II_{1} =\displaystyle= ΩTα(t)ϕj(gθ(uN)gθ32(uN)uNgθ(uθ)gθ32(uθ)uθ)Mθ(uN)μNgθ(uN)dxdt\displaystyle\int_{\Omega_{T}}\alpha(t)\phi_{j}\left(\frac{g^{\prime}_{\theta}(u^{N})}{g^{\frac{3}{2}}_{\theta}(u^{N})}\nabla u^{N}-\frac{g^{\prime}_{\theta}(u_{\theta})}{g^{\frac{3}{2}}_{\theta}(u_{\theta})}\nabla u_{\theta}\right)\cdot\sqrt{M_{\theta}(u^{N})}\nabla\frac{\mu^{N}}{g_{\theta}(u^{N})}dxdt
=\displaystyle= TδΩα(t)ϕj(gθ(uN)gθ32(uN)uNgθ(uθ)gθ32(uθ)uθ)Mθ(uN)μNgθ(uN)dxdt\displaystyle\int_{T_{\delta}}\int_{\Omega}\alpha(t)\phi_{j}\left(\frac{g^{\prime}_{\theta}(u^{N})}{g^{\frac{3}{2}}_{\theta}(u^{N})}\nabla u^{N}-\frac{g^{\prime}_{\theta}(u_{\theta})}{g^{\frac{3}{2}}_{\theta}(u_{\theta})}\nabla u_{\theta}\right)\cdot\sqrt{M_{\theta}(u^{N})}\nabla\frac{\mu^{N}}{g_{\theta}(u^{N})}dxdt
+[0,T]\TδΩα(t)ϕj(gθ(uN)gθ32(uN)uNgθ(uθ)gθ32(uθ)uθ)Mθ(uN)μNgθ(uN)dxdt\displaystyle+\int_{[0,T]\backslash T_{\delta}}\int_{\Omega}\alpha(t)\phi_{j}\left(\frac{g^{\prime}_{\theta}(u^{N})}{g^{\frac{3}{2}}_{\theta}(u^{N})}\nabla u^{N}-\frac{g^{\prime}_{\theta}(u_{\theta})}{g^{\frac{3}{2}}_{\theta}(u_{\theta})}\nabla u_{\theta}\right)\cdot\sqrt{M_{\theta}(u^{N})}\nabla\frac{\mu^{N}}{g_{\theta}(u^{N})}dxdt
=\displaystyle= II11+II12.\displaystyle II_{11}+II_{12}.

By (47), (51), (73) and (75), we can bound II11II_{11} by

|II11|\displaystyle|II_{11}| \displaystyle\leq Tδ|α(t)|gθ(uN)gθ32(uN)uNgθ(uθ)gθ32(uθ)uθL2(Ω)Mθ(uN)μNgθ(uN)L2(Ω)𝑑t\displaystyle\int_{T_{\delta}}|\alpha(t)|\left\Arrowvert\frac{g^{\prime}_{\theta}(u^{N})}{g^{\frac{3}{2}}_{\theta}(u^{N})}\nabla u^{N}-\frac{g^{\prime}_{\theta}(u_{\theta})}{g^{\frac{3}{2}}_{\theta}(u_{\theta})}\nabla u_{\theta}\right\Arrowvert_{L^{2}(\Omega)}\left\Arrowvert\sqrt{M_{\theta}(u^{N})}\nabla\frac{\mu^{N}}{g_{\theta}(u^{N})}\right\Arrowvert_{L^{2}(\Omega)}dt
\displaystyle\leq α(t)L2(Tδ)Mθ(uN)μNgθ(uN)L2(ΩT)gθ(uN)gθ32(uN)uNgθ(uθ)gθ32(uθ)uθL(0,T;L2(Ω))\displaystyle\left\Arrowvert\alpha(t)\right\Arrowvert_{L^{2}(T_{\delta})}\left\Arrowvert\sqrt{M_{\theta}(u^{N})}\nabla\frac{\mu^{N}}{g_{\theta}(u^{N})}\right\Arrowvert_{L^{2}(\Omega_{T})}\left\Arrowvert\frac{g^{\prime}_{\theta}(u^{N})}{g^{\frac{3}{2}}_{\theta}(u^{N})}\nabla u^{N}-\frac{g^{\prime}_{\theta}(u_{\theta})}{g^{\frac{3}{2}}_{\theta}(u_{\theta})}\nabla u_{\theta}\right\Arrowvert_{L^{\infty}(0,T;L^{2}(\Omega))}
\displaystyle\leq C(θ)ε.\displaystyle C(\theta)\varepsilon.

For II12II_{12}, we have

|II12|\displaystyle|II_{12}| \displaystyle\leq [0,T]\Tδ|α(t)|gθ(uN)gθ32(uN)uNgθ(uθ)gθ32(uθ)uθL2(Ω)Mθ(uN)μNgθ(uN)L2(Ω)𝑑t\displaystyle\int_{[0,T]\backslash T_{\delta}}|\alpha(t)|\left\Arrowvert\frac{g^{\prime}_{\theta}(u^{N})}{g^{\frac{3}{2}}_{\theta}(u^{N})}\nabla u^{N}-\frac{g^{\prime}_{\theta}(u_{\theta})}{g^{\frac{3}{2}}_{\theta}(u_{\theta})}\nabla u_{\theta}\right\Arrowvert_{L^{2}(\Omega)}\left\Arrowvert\sqrt{M_{\theta}(u^{N})}\nabla\frac{\mu^{N}}{g_{\theta}(u^{N})}\right\Arrowvert_{L^{2}(\Omega)}dt
=\displaystyle= [0,T]\Tδ|α(t)|fN(t)|Mθ(uN)μNgθ(uN)L2(Ω)dt.\displaystyle\int_{[0,T]\backslash T_{\delta}}|\alpha(t)|f^{N}(t)|\left\Arrowvert\sqrt{M_{\theta}(u^{N})}\nabla\frac{\mu^{N}}{g_{\theta}(u^{N})}\right\Arrowvert_{L^{2}(\Omega)}dt.

Since fN(t)f^{N}(t) converges to 0 uniformly on [0,T]\Tδ[0,T]\backslash T_{\delta}, α(t)L2(0,T)\alpha(t)\in L^{2}(0,T) and Mθ(uN)μNgθ(uN)L2(ΩT)C\left\Arrowvert\sqrt{M_{\theta}(u^{N})}\nabla\frac{\mu^{N}}{g_{\theta}(u^{N})}\right\Arrowvert_{L^{2}(\Omega_{T})}\leq C, letting NN\rightarrow\infty in II12II_{12} yields II120II_{12}\rightarrow 0. Letting ε0\varepsilon\rightarrow 0, we conclude II10II_{1}\rightarrow 0 as NN\rightarrow\infty. Passing to the limit in (3.3), we have

(79) 0Tα(t)Ωtuθ,ϕj(W1,q(Ω)),W1,q(Ω))𝑑t\displaystyle\int_{0}^{T}\alpha(t)\int_{\Omega}\left<\partial_{t}u_{\theta},\phi_{j}\right>_{(W^{1,q}(\Omega))^{\prime},W^{1,q}(\Omega))}dt
=\displaystyle= ΩTα(t)Mθ(uθ)μθgθ(uθ)ϕjgθ(uθ)dxdt.\displaystyle-\int_{\Omega_{T}}\alpha(t)M_{\theta}(u_{\theta})\nabla\frac{\mu_{\theta}}{g_{\theta}(u_{\theta})}\cdot\nabla\frac{\phi_{j}}{g_{\theta}(u_{\theta})}dxdt.

Fix q>2q>2, given any ϕL2(0,T;W1,q(Ω))\phi\in L^{2}(0,T;W^{1,q}(\Omega)), its Fourier series j=1aj(t)ϕj(x)\sum_{j=1}^{\infty}a_{j}(t)\phi_{j}(x) converges strongly to ϕ\phi in L2(0,T;W1,q(Ω))L^{2}(0,T;W^{1,q}(\Omega)). Hence

ΩTMθ(uθ)μθgθ(uθ)ϕΠNϕgθ(uθ)dxdt\displaystyle\int_{\Omega_{T}}M_{\theta}(u_{\theta})\nabla\frac{\mu_{\theta}}{g_{\theta}(u_{\theta})}\cdot\nabla\frac{\phi-\Pi_{N}\phi}{g_{\theta}(u_{\theta})}dxdt
=\displaystyle= ΩTM0μθgθ(uθ)(ϕΠNϕ)dxdt\displaystyle\int_{\Omega_{T}}M_{0}\nabla\frac{\mu_{\theta}}{g_{\theta}(u_{\theta})}\cdot\nabla(\phi-\Pi_{N}\phi)dxdt
ΩT(ϕΠNϕ)M0gθ(uθ)gθ32(uθ)uθMθ(uθ)μθgθ(uθ)dxdt\displaystyle-\int_{\Omega_{T}}(\phi-\Pi_{N}\phi)\sqrt{M_{0}}\frac{g_{\theta}^{\prime}(u_{\theta})}{g_{\theta}^{\frac{3}{2}}(u_{\theta})}\nabla u_{\theta}\cdot\sqrt{M_{\theta}(u_{\theta})}\nabla\frac{\mu_{\theta}}{g_{\theta}(u_{\theta})}dxdt
=\displaystyle= J1J2,\displaystyle J_{1}-J_{2},

where

J1=ΩTM0μθgθ(uθ)(ϕΠNϕ)dxdt0J_{1}=\int_{\Omega_{T}}M_{0}\nabla\frac{\mu_{\theta}}{g_{\theta}(u_{\theta})}\cdot\nabla(\phi-\Pi_{N}\phi)dxdt\rightarrow 0

by(63), (64) and strong convergence of ΠNϕ\Pi_{N}\phi to ϕ\phi in L2(0,T;H1(Ω))L^{2}(0,T;H^{1}(\Omega)). We can bound J2J_{2} by

|J2|=|ΩT(ϕΠNϕ)M0gθ(uθ)gθ3/2(uθ)uθMθ(uθ)μθgθ(uθ)dxdt|\displaystyle|J_{2}|=\left\arrowvert\int_{\Omega_{T}}(\phi-\Pi_{N}\phi)\sqrt{M_{0}}\frac{g_{\theta}^{\prime}(u_{\theta})}{g_{\theta}^{3/2}(u_{\theta})}\nabla u_{\theta}\cdot\sqrt{M_{\theta}(u_{\theta})}\nabla\frac{\mu_{\theta}}{g_{\theta}(u_{\theta})}dxdt\right\arrowvert
\displaystyle\leq M00TϕΠNϕL(Ω)gθ(uθ)gθ3/2(uθ)uθL2(Ω)Mθ(uθ)μθgθ(uθ)L2(Ω)\displaystyle\sqrt{M_{0}}\int_{0}^{T}\left\Arrowvert\phi-\Pi_{N}\phi\right\Arrowvert_{L^{\infty}(\Omega)}\left\Arrowvert\frac{g_{\theta}^{\prime}(u_{\theta})}{g_{\theta}^{3/2}(u_{\theta})}\nabla u_{\theta}\right\Arrowvert_{L^{2}(\Omega)}\left\Arrowvert\sqrt{M_{\theta}(u_{\theta})}\nabla\frac{\mu_{\theta}}{g_{\theta}(u_{\theta})}\right\Arrowvert_{L^{2}(\Omega)}
\displaystyle\leq M0gθ(uθ)gθ3/2(uθ)uθL(0,T;L2(Ω))Mθ(uθ)μθgθ(uθ)L2(ΩT)ϕΠNϕL2(0,T;W1,q(Ω))\displaystyle\sqrt{M_{0}}\left\Arrowvert\frac{g_{\theta}^{\prime}(u_{\theta})}{g_{\theta}^{3/2}(u_{\theta})}\nabla u_{\theta}\right\Arrowvert_{L^{\infty}(0,T;L^{2}(\Omega))}\left\Arrowvert\sqrt{M_{\theta}(u_{\theta})}\nabla\frac{\mu_{\theta}}{g_{\theta}(u_{\theta})}\right\Arrowvert_{L^{2}(\Omega_{T})}\left\Arrowvert\phi-\Pi_{N}\phi\right\Arrowvert_{L^{2}(0,T;W^{1,q}(\Omega))}
\displaystyle\rightarrow 0 as N.\displaystyle 0\text{ as }N\rightarrow\infty.

Consequently (79) and (3.3) imply

(81) 0Ttuθ,ϕ(W1,q(Ω)),W1,q(Ω))𝑑t=ΩTMθ(uθ)μθgθ(uθ)ϕgθ(uθ)dxdt\int_{0}^{T}\left<\partial_{t}u_{\theta},\phi\right>_{(W^{1,q}(\Omega))^{\prime},W^{1,q}(\Omega))}dt=-\int_{\Omega_{T}}M_{\theta}(u_{\theta})\nabla\frac{\mu_{\theta}}{g_{\theta}(u_{\theta})}\cdot\nabla\frac{\phi}{g_{\theta}(u_{\theta})}dxdt

for all ϕL2(0,T;W1,q(Ω))\phi\in L^{2}(0,T;W^{1,q}(\Omega)) with q>2q>2. Moreover, since uN(x,0)=ΠNu0(x)u0(x)u^{N}(x,0)=\Pi_{N}u_{0}(x)\rightarrow u_{0}(x) in H1(Ω)H^{1}(\Omega), we see that uθ(x,0)=u0(x)u_{\theta}(x,0)=u_{0}(x) by (52).

Remark 3.2.

(Generalized dominated convergence theorem) Assume EnE\subset\mathbb{R}^{n} is measurable. gngg_{n}\rightarrow g strongly in Lq(E)L^{q}(E) for 1q<1\leq q<\infty and fnf_{n}, ff: EnE\rightarrow\mathbb{R}^{n} are measurable functions satisfying

fnf a.e. in E;|fn|p|gn|q a.e. in Ef_{n}\rightarrow f\text{ a.e. in }E;\hskip 7.22743pt|f_{n}|^{p}\leq|g_{n}|^{q}\text{ a.e. in }E

with 1p<1\leq p<\infty, then fnff_{n}\rightarrow f in Lp(E)L^{p}(E).

3.4 Regularity of uθu_{\theta}

We now consider the regularity of uθu_{\theta}. Given any aj(t)L2(0,T)a_{j}(t)\in L^{2}(0,T), aj(t)ϕjL2(0,T;C(Ω¯)a_{j}(t)\phi_{j}\in L^{2}(0,T;C(\overline{\Omega})). Integrating (43) from 0 to TT, by (58),(64) and (71), we have

ΩTμθ(x,t)aj(t)ϕj(x)𝑑x𝑑t=ΩT(uθaj(t)ϕj+q(uθ)aj(t)ϕj)𝑑x𝑑t\displaystyle\int_{\Omega_{T}}\mu_{\theta}(x,t)a_{j}(t)\phi_{j}(x)dxdt=\int_{\Omega_{T}}\left(\nabla u_{\theta}\cdot a_{j}(t)\nabla\phi_{j}+q^{\prime}(u_{\theta})a_{j}(t)\phi_{j}\right)dxdt

for all j𝐍j\in{\bf N}. Given any ϕL2(0,T;H1(Ω))\phi\in L^{2}(0,T;H^{1}(\Omega)), its Fouirier series strongly converges to ϕ\phi in L2(0,T;H1(Ω))L^{2}(0,T;H^{1}(\Omega)), therefore

(82) ΩTμθ(x,t)ϕ(x)𝑑x𝑑t=ΩT(uθϕ+q(uθ)ϕ)𝑑x𝑑t.\displaystyle\int_{\Omega_{T}}\mu_{\theta}(x,t)\phi(x)dxdt=\int_{\Omega_{T}}\left(\nabla u_{\theta}\cdot\nabla\phi+q^{\prime}(u_{\theta})\phi\right)dxdt.

Recall μθL2(0,T;Lp(Ω))\mu_{\theta}\in L^{2}(0,T;L^{p}(\Omega)) and q(uθ)L(0,T;Lp(Ω))q^{\prime}(u_{\theta})\in L^{\infty}(0,T;L^{p}(\Omega)) for any 1p<1\leq p<\infty, regularity theory implies uθL2(0,T;W2,p(Ω))u_{\theta}\in L^{2}(0,T;W^{2,p}(\Omega)). Hence

(83) μθ=Δuθ+q(uθ) a.e. in ΩT.\mu_{\theta}=-\Delta u_{\theta}+q^{\prime}(u_{\theta})\text{ a.e. in }\Omega_{T}.

Since growth assumption on qq implies |q′′(u)|C(1+|u|r1)|q^{\prime\prime}(u)|\leq C(1+|u|^{r-1}), pick p>2p>2, we have

Ω|q(uθ)|2𝑑x=Ω|q′′(uθ)|2|uθ|2𝑑x\displaystyle\int_{\Omega}|\nabla q^{\prime}(u_{\theta})|^{2}dx=\int_{\Omega}|q^{\prime\prime}(u_{\theta})|^{2}|\nabla u_{\theta}|^{2}dx
\displaystyle\leq q′′(uθ)L2pp2(Ω)2uθLp(Ω)2\displaystyle\left\Arrowvert q^{\prime\prime}(u_{\theta})\right\Arrowvert_{L^{\frac{2p}{p-2}}(\Omega)}^{2}\left\Arrowvert\nabla u_{\theta}\right\Arrowvert_{L^{p}(\Omega)}^{2}
\displaystyle\leq C(1+uθL2pp2(r1)(Ω)2(r1))uθLp(Ω)2\displaystyle C\left(1+\left\Arrowvert u_{\theta}\right\Arrowvert_{L^{\frac{2p}{p-2}(r-1)}(\Omega)}^{2(r-1)}\right)\left\Arrowvert\nabla u_{\theta}\right\Arrowvert_{L^{p}(\Omega)}^{2}
\displaystyle\leq C(1+uθL(0,T;H1(Ω))2(r1))uθLp(Ω)2\displaystyle C\left(1+\left\Arrowvert u_{\theta}\right\Arrowvert^{2(r-1)}_{L^{\infty}(0,T;H^{1}(\Omega))}\right)\left\Arrowvert\nabla u_{\theta}\right\Arrowvert_{L^{p}(\Omega)}^{2}

Therefore q(uθ)=q′′(uθ)uθL2(ΩT)\nabla q^{\prime}(u_{\theta})=q^{\prime\prime}(u_{\theta})\nabla u_{\theta}\in L^{2}(\Omega_{T}) with

ΩT|q(uθ)|2𝑑x𝑑t(1+uθL(0,T;H1(Ω))2(r1))uθL2(0,T;Lp(Ω))2.\displaystyle\int_{\Omega_{T}}|\nabla q^{\prime}(u_{\theta})|^{2}dxdt\leq\left(1+\left\Arrowvert u_{\theta}\right\Arrowvert^{2(r-1)}_{L^{\infty}(0,T;H^{1}(\Omega))}\right)\left\Arrowvert\nabla u_{\theta}\right\Arrowvert_{L^{2}(0,T;L^{p}(\Omega))}^{2}.

Hence q(uθ)L2(0,T;H1(Ω))q^{\prime}(u_{\theta})\in L^{2}(0,T;H^{1}(\Omega)), combined with μθL2(0,T;W1,s(Ω))\mu_{\theta}\in L^{2}(0,T;W^{1,s}(\Omega)) for any 1s<21\leq s<2, we have uθL2(0,T;W3,s(Ω))u_{\theta}\in L^{2}(0,T;W^{3,s}(\Omega)) and

(84) μθ=Δuθ+q′′(uθ)uθ a.e. in ΩT.\nabla\mu_{\theta}=-\nabla\Delta u_{\theta}+q^{\prime\prime}(u_{\theta})\nabla u_{\theta}\text{ a.e. in }\Omega_{T}.

Regularity of uθu_{\theta} implies uθL(0,T;L2(Ω))L2(0,T;L(Ω))\nabla u_{\theta}\in L^{\infty}(0,T;L^{2}(\Omega))\cap L^{2}(0,T;L^{\infty}(\Omega)). A simple interpolation shows uθL2μμ2(0,T;Lμ(Ω))\nabla u_{\theta}\in L^{\frac{2\mu}{\mu-2}}(0,T;L^{\mu}(\Omega)) for any μ>2\mu>2. Given any ϕLp(0,T;W1,q(Ω))\phi\in L^{p}(0,T;W^{1,q}(\Omega)) with p>2p>2 and q>2q>2, we have gθ(uθ)ϕL2(0,T;W1.r(Ω))g_{\theta}(u_{\theta})\phi\in L^{2}(0,T;W^{1.r}(\Omega)) for any r<min(p,q)r<\min(p,q). From this, we can pick gθ(uθ)ϕg_{\theta}(u_{\theta})\phi as a test function in (81), we have

(85) ΩTtuθgθ(uθ)ϕdxdt=ΩTMθ(uθ)μθgθ(uθ)ϕdxdt\int_{\Omega_{T}}\partial_{t}u_{\theta}g_{\theta}(u_{\theta})\phi dxdt=-\int_{\Omega_{T}}M_{\theta}(u_{\theta})\nabla\frac{\mu_{\theta}}{g_{\theta}(u_{\theta})}\cdot\nabla\phi dxdt

for any ϕLp(0,T;W1,q(Ω))\phi\in L^{p}(0,T;W^{1,q}(\Omega)) with p,q>2p,q>2.

Remark 3.3.

In fact, since Mθ(uθ)L(0,T;Lp(Ω))M_{\theta}(u_{\theta})\in L^{\infty}(0,T;L^{p}(\Omega)) for 1p<1\leq p<\infty, the right hand side of (85) is well defined for any ϕL2(0,T,W1,q(Ω))\phi\in L^{2}(0,T,W^{1,q}(\Omega)) for q>2q>2 and we can extend (85) to hold for all ϕL2(0,T,W1,q(Ω))\phi\in L^{2}(0,T,W^{1,q}(\Omega)).

3.5 Energy Inequality

Since uNu^{N} and μN\mu^{N} satisfies energy identity (3.1), passing to the limit as NN\rightarrow\infty and using strong convergence of uN(x,0)u^{N}(x,0) to u0u_{0} in H1(Ω)H^{1}(\Omega), together with the weak convergence of uNu^{N}, q(uN)q^{\prime}(u^{N}) and Mθ(uN)μNgθ(uN)\sqrt{M_{\theta}(u^{N})}\nabla\frac{\mu^{N}}{g_{\theta}(u^{N})}, the energy inequality (12) follows.

4 Phase field model with degenerate mobility

In this section, we prove theorem 1.2. Fix initial data u0H1(Ω)u_{0}\in H^{1}(\Omega). We pick a montone decreasing positive sequence θi\theta_{i} with limiθi=0\lim_{i\rightarrow\infty}\theta_{i}=0. By theorem 1.1 and (85), for each θi\theta_{i}, there exists

uiL(0,T;H1(Ω))L2(0,T;W3,s(Ω))C([0,T];Lp(Ω))u_{i}\in L^{\infty}(0,T;H^{1}(\Omega))\cap L^{2}(0,T;W^{3,s}(\Omega))\cap C([0,T];L^{p}(\Omega))

with weak derivative

tuiL2(0,T;(W1,q(Ω))),\partial_{t}u_{i}\in L^{2}(0,T;(W^{1,q}(\Omega))^{\prime}),

where 1p<1\leq p<\infty, 1s<21\leq s<2, q>2q>2 such that uθi(x,0)=u0(x)u_{\theta_{i}}(x,0)=u_{0}(x) and for all ϕL2(0,T;W1,q(Ω))\phi\in L^{2}(0,T;W^{1,q}(\Omega)),

(86) 0TΩtuiϕdxdt\displaystyle\int_{0}^{T}\int_{\Omega}\partial_{t}u_{i}\phi dxdt =\displaystyle= 0TΩMi(ui)μigi(ui)ϕgi(ui)dxdt,\displaystyle-\int_{0}^{T}\int_{\Omega}M_{i}(u_{i})\nabla\frac{\mu_{i}}{g_{i}(u_{i})}\nabla\frac{\phi}{g_{i}(u_{i})}dxdt,
(87) μi\displaystyle\mu_{i} =\displaystyle= Δui+q(ui).\displaystyle-\Delta u_{i}+q^{\prime}(u_{i}).

Moreover, for all ψLp(0,T;W1,q(Ω))\psi\in L^{p}(0,T;W^{1,q}(\Omega)) with p,q>2p,q>2, the following holds:

(88) 0TΩgi(ui)tuiψdxdt=0TΩMi(ui)μigi(ui)ψdxdt\int_{0}^{T}\int_{\Omega}g_{i}(u_{i})\partial_{t}u_{i}\psi dxdt=-\int_{0}^{T}\int_{\Omega}M_{i}(u_{i})\nabla\frac{\mu_{i}}{g_{i}(u_{i})}\nabla\psi dxdt

Here we write ui=uθiu_{i}=u_{\theta_{i}}, Mi(ui)=Mθi(uθi)M_{i}(u_{i})=M_{\theta_{i}}(u_{\theta_{i}}), gi(ui)=gθi(uθi)g_{i}(u_{i})=g_{\theta_{i}}(u_{\theta_{i}}) for simplicity of notations.

4.1 Convergence of uiu_{i} and equation for the limit function

Noticing the bound in (46) and (47) only depends on u0u_{0}, we can find a constant CC, independent of θi\theta_{i} such that

(89) uiL(0,T;H1(Ω))C,\displaystyle\left\Arrowvert u_{i}\right\Arrowvert_{L^{\infty}(0,T;H^{1}(\Omega))}\leq C,
(90) Mi(ui)μigi(ui)L2(ΩT)C.\displaystyle\left\Arrowvert\sqrt{M_{i}(u_{i})}\nabla\frac{\mu_{i}}{g_{i}(u_{i})}\right\Arrowvert_{L^{2}(\Omega_{T})}\leq C.

Growth condition on qq^{\prime}, and Sobolev embedding theorem give

q(ui)L(0,T;Lp(Ω))C,\displaystyle\left\Arrowvert q^{\prime}(u_{i})\right\Arrowvert_{L^{\infty}(0,T;L^{p}(\Omega))}\leq C,
Mi(ui)L(0,T;Lp(Ω))C\displaystyle\left\Arrowvert M_{i}(u_{i})\right\Arrowvert_{L^{\infty}(0,T;L^{p}(\Omega))}\leq C

for any 1p<1\leq p<\infty. By (88), for any ϕLp(0,T;W1,q(Ω))\phi\in L^{p}(0,T;W^{1,q}(\Omega)) with p,q>2p,q>2,

|0TΩgi(ui)tuiϕdxdt|=|0TΩMi(ui)μigi(ui)ϕdxdt|\displaystyle\left\arrowvert\int_{0}^{T}\int_{\Omega}g_{i}(u_{i})\partial_{t}u_{i}\phi dxdt\right\arrowvert=\left\arrowvert\int_{0}^{T}\int_{\Omega}M_{i}(u_{i})\nabla\frac{\mu_{i}}{g_{i}(u_{i})}\nabla\phi dxdt\right\arrowvert
\displaystyle\leq 0T(Mi(ui)μigi(ui)L2(Ω)Mi(ui)L2qq2(Ω)ϕLq(Ω))𝑑t\displaystyle\int_{0}^{T}\left(\left\Arrowvert\sqrt{M_{i}(u_{i})}\nabla\frac{\mu_{i}}{g_{i}(u_{i})}\right\Arrowvert_{L^{2}(\Omega)}\left\Arrowvert\sqrt{M_{i}(u_{i})}\right\Arrowvert_{L^{\frac{2q}{q-2}}(\Omega)}\left\Arrowvert\nabla\phi\right\Arrowvert_{L^{q}(\Omega)}\right)dt
\displaystyle\leq Mi(ui)Lpp2(0,T;Lqq2(Ω))12Mi(ui)μigi(ui)L2(ΩT)ϕLp(0,T;Lq(Ω))\displaystyle\left\Arrowvert M_{i}(u_{i})\right\Arrowvert^{\frac{1}{2}}_{L^{\frac{p}{p-2}}(0,T;L^{\frac{q}{q-2}}(\Omega))}\left\Arrowvert\sqrt{M_{i}(u_{i})}\nabla\frac{\mu_{i}}{g_{i}(u_{i})}\right\Arrowvert_{L^{2}(\Omega_{T})}\left\Arrowvert\nabla\phi\right\Arrowvert_{L^{p}(0,T;L^{q}(\Omega))}
\displaystyle\leq CϕLp(0,T;W1,q(Ω)).\displaystyle C\left\Arrowvert\phi\right\Arrowvert_{L^{p}(0,T;W^{1,q}(\Omega))}.

Let

(92) Gi(ui)=0uigi(a)𝑑a.G_{i}(u_{i})=\int_{0}^{u_{i}}g_{i}(a)da.

Thus (4.1) yields tGi(ui)=gi(ui)tuiLp(0,T;(W1,q(Ω)))\partial_{t}G_{i}(u_{i})=g_{i}(u_{i})\partial_{t}u_{i}\in L^{p^{\prime}}(0,T;(W^{1,q}(\Omega))^{\prime}) with p=pp1p^{\prime}=\frac{p}{p-1} and

(93) tGi(ui)Lp(0,T;(W1,q(Ω)))C for all i.\left\Arrowvert\partial_{t}G_{i}(u_{i})\right\Arrowvert_{L^{p^{\prime}}(0,T;(W^{1,q}(\Omega))^{\prime})}\leq C\text{ for all }i.

Moreover, by growth assumption on gg and estimates on uiu_{i}, we have

(94) Gi(ui)L(0,T;W1,s(Ω))C.\left\Arrowvert G_{i}(u_{i})\right\Arrowvert_{L^{\infty}(0,T;W^{1,s}(\Omega))}\leq C.

for 1s<21\leq s<2. By (89), (90), (93)-(94) and Remark 3.1 we can find a subsequence, not relabeled, a function uL(0,T;H1(Ω))u\in L^{\infty}(0,T;H^{1}(\Omega)), a function ξL2(ΩT)\xi\in L^{2}(\Omega_{T}) and a function ηL(0,T;W1,s(Ω))\eta\in L^{\infty}(0,T;W^{1,s}(\Omega)) such that as ii\rightarrow\infty,

(95) uiu weakly-* in L(0,T;H1(Ω)),\displaystyle u_{i}\rightharpoonup u\text{ weakly-* in }L^{\infty}(0,T;H^{1}(\Omega)),
(96) Mi(ui)μigi(ui)ξ weakly in L2(ΩT),\displaystyle\sqrt{M_{i}(u_{i})}\nabla\frac{\mu_{i}}{g_{i}(u_{i})}\rightharpoonup\xi\text{ weakly in }L^{2}(\Omega_{T}),
(97) Gi(ui)η weakly-* in L(0,T;W1,s(Ω))\displaystyle G_{i}(u_{i})\rightharpoonup\eta\text{ weakly-* in }L^{\infty}(0,T;W^{1,s}(\Omega))
(98) Gi(ui)η strongly in Lα(0,T;Lβ(Ω)) and a.e. in ΩT,\displaystyle G_{i}(u_{i})\rightarrow\eta\text{ strongly in }L^{\alpha}(0,T;L^{\beta}(\Omega))\text{ and a.e. in }\Omega_{T},
(99) Gi(ui)η strongly in C(0,T;Lβ(Ω)),\displaystyle G_{i}(u_{i})\rightarrow\eta\text{ strongly in }C(0,T;L^{\beta}(\Omega)),
(100) tGi(ui)tη weakly in Lp(0,T;(W1,qΩ))).\displaystyle\partial_{t}G_{i}(u_{i})\rightharpoonup\partial_{t}\eta\text{ weakly in }L^{p^{\prime}}(0,T;(W^{1,q}\Omega))^{\prime}).

where 1α,β<1\leq\alpha,\beta<\infty. By (99) and (105) from Remark 4.1, we have

Gi(ui(x,t+h))Gi(ui(x,t))C([0,T];Lβ(Ω))0 uniformly in i as h0.\left\Arrowvert G_{i}(u_{i}(x,t+h))-G_{i}(u_{i}(x,t))\right\Arrowvert_{C([0,T];L^{\beta}(\Omega))}\rightarrow 0\text{ uniformly in }i\text{ as }h\rightarrow 0.

Thus given any ε>0\varepsilon>0, there exists hε>0h_{\varepsilon}>0 such that for all 0<h<hε0<h<h_{\varepsilon} and all ii,

Gi(ui(x,t+h))Gi(ui(x,t))C([0,T];Lβ(Ω))β<ε.\left\Arrowvert G_{i}(u_{i}(x,t+h))-G_{i}(u_{i}(x,t))\right\Arrowvert_{C([0,T];L^{\beta}(\Omega))}^{\beta}<\varepsilon.

Given any δ>0\delta>0, let Iδ=(1δ,1+δ)(1δ,1+δ)I_{\delta}=(1-\delta,1+\delta)\cup(-1-\delta,-1+\delta). Consider the interval having ui(x,t)u_{i}(x,t) and ui(x,t+h)u_{i}(x,t+h) as end points. Denote this interval by Ji(x,t;h)J_{i}(x,t;h). We consider three cases.

Case I: Ji(x,t;h)Iδ=\varnothingJ_{i}(x,t;h)\cap I_{\delta}=\varnothing.

In this case, gi(s)max{θim,δ2m}g_{i}(s)\geq\max\{\theta_{i}^{m},\delta^{2m}\} for any sJi(x,t;h)s\in J_{i}(x,t;h) and by (92)

|Gi(ui(x,t+h))Gi(ui(x,t))|=|ui(x,t)ui(x,t+h)gi(s)𝑑s|δ2m|ui(x,t+h)ui(x,t)|.\left|G_{i}(u_{i}(x,t+h))-G_{i}(u_{i}(x,t))\right|=\left|\int_{u_{i}(x,t)}^{u_{i}(x,t+h)}g_{i}(s)ds\right|\geq\delta^{2m}|u_{i}(x,t+h)-u_{i}(x,t)|.

Case II: Ji(x,t;h)Iδ\varnothingJ_{i}(x,t;h)\cap I_{\delta}\neq\varnothing and |ui(x,t+h)ui(x,t)|3δ|u_{i}(x,t+h)-u_{i}(x,t)|\geq 3\delta .

In this case, we have

|Ji(x,t;h)Iδc|13|Ji(x,t;h)||J_{i}(x,t;h)\cap I_{\delta}^{c}|\geq\frac{1}{3}|J_{i}(x,t;h)|

and

|Gi(ui(x,t+h))Gi(ui(x,t))|\displaystyle\left|G_{i}(u_{i}(x,t+h))-G_{i}(u_{i}(x,t))\right| \displaystyle\geq |Ji(x,t;h)Iδcgi(s)𝑑s|\displaystyle\left|\int_{J_{i}(x,t;h)\cap I_{\delta}^{c}}g_{i}(s)ds\right|
\displaystyle\geq δ2m3|ui(x,t+h)ui(x,t)|.\displaystyle\frac{\delta^{2m}}{3}|u_{i}(x,t+h)-u_{i}(x,t)|.

Case III: Ji(x,t;h)Iδ\varnothingJ_{i}(x,t;h)\cap I_{\delta}\neq\varnothing and |ui(x,t+h)ui(x,t)|<3δ|u_{i}(x,t+h)-u_{i}(x,t)|<3\delta

In this case, we have

gi(s)max{(8δ+16δ2)m,θim} for any sJi(x,t;h).g_{i}(s)\leq\max\{(8\delta+16\delta^{2})^{m},\theta_{i}^{m}\}\text{ for any }s\in J_{i}(x,t;h).

Thus

|Gi(ui(x,t+h))Gi(ui(x,t))|3δmax{(8δ+16δ2)m,θim}.\displaystyle\left|G_{i}(u_{i}(x,t+h))-G_{i}(u_{i}(x,t))\right|\leq 3\delta\max\{(8\delta+16\delta^{2})^{m},\theta_{i}^{m}\}.

Pick δ=ε14mβ\delta=\varepsilon^{\frac{1}{4m\beta}} and fix tt. Let

Ωit={xΩ:Ji(x,t:h) satisfies case I or II}.\Omega_{i}^{t}=\{x\in\Omega:J_{i}(x,t:h)\text{ satisfies case I or II}\}.

Then

Ω|ui(x,t+h)ui(x,t)|β𝑑x\displaystyle\int_{\Omega}\left\arrowvert u_{i}(x,t+h)-u_{i}(x,t)\right\arrowvert^{\beta}dx
=\displaystyle= Ωit|ui(x,t+h)ui(x,t)|β𝑑x+Ω\Ωit|ui(x,t+h)ui(x,t)|β𝑑x\displaystyle\int_{\Omega_{i}^{t}}\left\arrowvert u_{i}(x,t+h)-u_{i}(x,t)\right\arrowvert^{\beta}dx+\int_{\Omega\backslash\Omega_{i}^{t}}\left\arrowvert u_{i}(x,t+h)-u_{i}(x,t)\right\arrowvert^{\beta}dx
\displaystyle\leq 3βε12Ωit|Gi(ui(x,t+h))Gi(ui(x,t))|β𝑑x+Ω\Ωit|ui(x,t+h)ui(x,t)|β𝑑x\displaystyle 3^{\beta}\varepsilon^{-\frac{1}{2}}\int_{\Omega_{i}^{t}}\left\arrowvert G_{i}(u_{i}(x,t+h))-G_{i}(u_{i}(x,t))\right\arrowvert^{\beta}dx+\int_{\Omega\backslash\Omega_{i}^{t}}\left\arrowvert u_{i}(x,t+h)-u_{i}(x,t)\right\arrowvert^{\beta}dx
\displaystyle\leq 3βε12+Cε14m\displaystyle 3^{\beta}\varepsilon^{\frac{1}{2}}+C\varepsilon^{\frac{1}{4m}}

Taking maximum on the left side, we have for all ii, any h<hεh<h_{\varepsilon},

(ui(x,t+h)ui(x,t)C([0,T];Lβ(Ω))β3βε12+Cε14m.\left\Arrowvert(u_{i}(x,t+h)-u_{i}(x,t)\right\Arrowvert_{C([0,T];L^{\beta}(\Omega))}^{\beta}\leq 3^{\beta}\varepsilon^{\frac{1}{2}}+C\varepsilon^{\frac{1}{4m}}.

Thus

ui(x,t+h)ui(x,t)C([0,T];Lβ(Ω))β0 uniformly as h0.\left\Arrowvert u_{i}(x,t+h)-u_{i}(x,t)\right\Arrowvert_{C([0,T];L^{\beta}(\Omega))}^{\beta}\rightarrow 0\text{ uniformly as }h\rightarrow 0.

In addition, for any 0<t1<t2<T0<t_{1}<t_{2}<T, (89) implies that for 1β<1\leq\beta<\infty, we have

t1t2ui(x,t)𝑑t is relatively compact in Lβ(Ω).\int_{t_{1}}^{t_{2}}u_{i}(x,t)dt\text{ is relatively compact in }L^{\beta}(\Omega).

Therefore we conclude from Remark 4.1 that

(101) uiu(x,t) strongly in C([0,T];Lβ(Ω)) for 1β<.u_{i}\rightarrow u(x,t)\text{ strongly in }C([0,T];L^{\beta}(\Omega))\text{ for }1\leq\beta<\infty.

Similarly. we can prove

(102) uiu(x,t) strongly in Lα(0,T;Lβ(Ω)) for 1α,β< and a.e. in ΩT.u_{i}\rightarrow u(x,t)\text{ strongly in }L^{\alpha}(0,T;L^{\beta}(\Omega))\text{ for }1\leq\alpha,\beta<\infty\text{ and a.e. in }\Omega_{T}.

Growth condition on M(u)M(u) and (101), (102) yield

Mi(ui)M(u) strongly in C([0,T];Lβ(Ω)) for 1β<,\displaystyle M_{i}(u_{i})\rightarrow M(u)\text{ strongly in }C([0,T];L^{\beta}(\Omega))\text{ for }1\leq\beta<\infty,
Mi(ui)M(u) strongly in Lα(0,T;Lβ(Ω)) for 1α,β<,\displaystyle M_{i}(u_{i})\rightarrow M(u)\text{ strongly in }L^{\alpha}(0,T;L^{\beta}(\Omega))\text{ for }1\leq\alpha,\beta<\infty,
M(i(ui))M(u) strongly in C([0,T];Lγ(Ω)) for 1γ<.\displaystyle\sqrt{M(_{i}(u_{i}))}\rightarrow\sqrt{M(u)}\text{ strongly in }C([0,T];L^{\gamma}(\Omega))\text{ for }1\leq\gamma<\infty.

Hence Gi(ui)G_{i}(u_{i}) converges to G(u)G(u) a.e. in ΩT\Omega_{T} and η=G(u)\eta=G(u). Passing to the limit in (88), we have

(103) 0Tg(u)tu,ϕ((W1,q(Ω)),W1,q(Ω))𝑑t=0TΩM(u)ξϕdxdt\int_{0}^{T}\left<g(u)\partial_{t}u,\phi\right>_{((W^{1,q}(\Omega))^{\prime},W^{1,q}(\Omega))}dt=-\int_{0}^{T}\int_{\Omega}\sqrt{M(u)}\xi\cdot\nabla\phi dxdt

for any ϕLp(0,T;W1,q(Ω))\phi\in L^{p}(0,T;W^{1,q}(\Omega)) with p,q>2p,q>2.

Remark 4.1.

(Compactness in Lp(0,T;B)L^{p}(0,T;B) Theorem 1 in [22]) Assume BB is a Banach space and FLp(0,T;B)F\subset L^{p}(0,T;B). FF is relatively compact in Lp(0,T;B)L^{p}(0,T;B) for 1p<1\leq p<\infty, or in C([0,T],B)C([0,T],B) for p=p=\infty if and only if

(104) {t1t2f(t)𝑑t:fF} is relatively compact in B,0<t1<t2<T,\left\{\int_{t_{1}}^{t_{2}}f(t)dt:f\in F\right\}\text{ is relatively compact in }B,\forall 0<t_{1}<t_{2}<T,
(105) τhffLp(0,T;B)0 as h0 uniformly for fF.\left\Arrowvert\tau_{h}f-f\right\Arrowvert_{L^{p}(0,T;B)}\rightarrow 0\text{ as }h\rightarrow 0\text{ uniformly for }f\in F.

Here τhf(t)=f(t+h)\tau_{h}f(t)=f(t+h) for h>0h>0 is defined on [h,Th][-h,T-h].

4.2 Weak convergence of μigi(ui)\nabla\frac{\mu_{i}}{g_{i}(u_{i})}

We now look for relation between ξ\xi and uu. Following the idea in [7], we decompose ΩT\Omega_{T} as follows. Let δj\delta_{j} be a positive sequence monotonically decreasing to 0. By (96) and Egorov’s theorem, for every δj>0\delta_{j}>0, there exists BjΩTB_{j}\subset\Omega_{T} satisfying |Ωt\Bj|<δj|\Omega_{t}\backslash B_{j}|<\delta_{j} such that

(106) uiu uniformly in Bj.u_{i}\rightarrow u\text{ uniformly in }B_{j}.

We can pick

(107) B1B2BjBj+1ΩT.B_{1}\subset B_{2}\subset\cdots\subset B_{j}\subset B_{j+1}\subset\cdots\subset\Omega_{T}.

Define

Pj:={(x,t)ΩT:|1u2|>δj}.P_{j}:=\{(x,t)\in\Omega_{T}:|1-u^{2}|>\delta_{j}\}.

Then

(108) P1P2PjPj+1ΩT.P_{1}\subset P_{2}\subset\cdots\subset P_{j}\subset P_{j+1}\subset\cdots\subset\Omega_{T}.

Let B=j=1BjB=\cup_{j=1}^{\infty}B_{j} and P=j=1PjP=\cup_{j=1}^{\infty}P_{j}. Then |ΩT\B|=0|\Omega_{T}\backslash B|=0 and each BjB_{j} can be split into two parts:

Dj=BjPj, where |1u2|>δj, and uiu uniformly,\displaystyle D_{j}=B_{j}\cap P_{j},\text{ where }|1-u^{2}|>\delta_{j},\text{ and }u_{i}\rightarrow u\text{ uniformly},
D~j=Bj\Pj, where |1u2|δj, and uiu uniformly .\displaystyle\tilde{D}_{j}=B_{j}\backslash P_{j},\text{ where }|1-u^{2}|\leq\delta_{j},\text{ and }u_{i}\rightarrow u\text{ uniformly }.

(107) and (108) imply

(109) D1D2DjDj+1D:=BP.D_{1}\subset D_{2}\subset\cdots\subset D_{j}\subset D_{j+1}\subset\cdots\subset D:=B\cap P.

For any ΨLp(0,T;Lq(Ω,n))\Psi\in L^{p}(0,T;L^{q}(\Omega,\mathbb{R}^{n})) with p,q>2p,q>2, we have

(110) ΩTMi(ui)μigi(ui)Ψdxdt\displaystyle\int_{\Omega_{T}}M_{i}(u_{i})\nabla\frac{\mu_{i}}{g_{i}(u_{i})}\cdot\Psi dxdt
=\displaystyle= ΩT\BjMi(ui)μigi(ui)Ψdxdt+DjMi(ui)μigi(ui)Ψdxdt\displaystyle\int_{\Omega_{T}\backslash B_{j}}M_{i}(u_{i})\nabla\frac{\mu_{i}}{g_{i}(u_{i})}\cdot\Psi dxdt+\int_{D_{j}}M_{i}(u_{i})\nabla\frac{\mu_{i}}{g_{i}(u_{i})}\cdot\Psi dxdt
+D~jMi(ui)μigi(ui)Ψdxdt\displaystyle+\int_{\tilde{D}_{j}}M_{i}(u_{i})\nabla\frac{\mu_{i}}{g_{i}(u_{i})}\cdot\Psi dxdt

The left hand side of (110) converges to ΩTM(u)ξΨ𝑑x𝑑t\int_{\Omega_{T}}\sqrt{M(u)}\xi\cdot\Psi dxdt. We analyze the three terms on the right hand side separately. To estimate the first term on the right hand side of (110), noticing |ΩT\Bj|0|\Omega_{T}\backslash B_{j}|\rightarrow 0 and

limiΩT\BjMi(ui)μigi(ui)Ψdxdt=ΩT\BjM(u)ξΨ𝑑x𝑑t,\displaystyle\lim_{i\rightarrow\infty}\int_{\Omega_{T}\backslash B_{j}}M_{i}(u_{i})\nabla\frac{\mu_{i}}{g_{i}(u_{i})}\cdot\Psi dxdt=\ \int_{\Omega_{T}\backslash B_{j}}\sqrt{M(u)}\xi\cdot\Psi dxdt,

we have

limjlimiΩT\BjMi(ui)μigi(ui)Ψdxdt=0.\displaystyle\lim_{j\rightarrow\infty}\lim_{i\rightarrow\infty}\int_{\Omega_{T}\backslash B_{j}}M_{i}(u_{i})\nabla\frac{\mu_{i}}{g_{i}(u_{i})}\cdot\Psi dxdt=0.

By uniform convergence of uiu_{i} to uu in BjB_{j}, we introduce subsequence uj,ku_{j,k} such that uj,kuu_{j,k}\rightarrow u uniformly in BjB_{j} and there exists NjN_{j} such that for all kNjk\geq N_{j},

(111) |1uj,k2|>δj2 in Dj,|1uj,k2|2δj in D~j.|1-u^{2}_{j,k}|>\frac{\delta_{j}}{2}\text{ in }D_{j},\hskip 14.45377pt|1-u^{2}_{j,k}|\leq 2\delta_{j}\text{ in }\tilde{D}_{j}.

Thus the third term on the right hand side of (110) can be estimated by

limjlimk|D~jMj,k(uj,k)μj,kgj,k(uj,k)Ψdxdt|\displaystyle\lim_{j\rightarrow\infty}\lim_{k\rightarrow\infty}\left\arrowvert\int_{\tilde{D}_{j}}M_{j,k}(u_{j,k})\nabla\frac{\mu_{j,k}}{g_{j,k}(u_{j,k})}\cdot\Psi dxdt\right\arrowvert
\displaystyle\leq limjlimk{(supD~jMj,k(uj,k))ΨL2(D~j)Mj,k(uj,k)μj,kgj,k(uj,k)L2(D~j)}\displaystyle\lim_{j\rightarrow\infty}\lim_{k\rightarrow\infty}\left\{\left(\sup_{\tilde{D}_{j}}\sqrt{M_{j,k}(u_{j,k})}\right)\left\Arrowvert\Psi\right\Arrowvert_{L^{2}(\tilde{D}_{j})}\left\Arrowvert\sqrt{M_{j,k}(u_{j,k})}\nabla\frac{\mu_{j,k}}{g_{j,k}(u_{j,k})}\right\Arrowvert_{L^{2}(\tilde{D}_{j})}\right\}
\displaystyle\leq (supD~jMj,k(uj,k))|Ω|q22qΨL2(0,T;Lq(Ω)Mj,k(uj,k)μj,kgj,k(uj,k)L2(D~j)\displaystyle\left(\sup_{\tilde{D}_{j}}\sqrt{M_{j,k}(u_{j,k})}\right)|\Omega|^{\frac{q-2}{2q}}\left\Arrowvert\Psi\right\Arrowvert_{L^{2}(0,T;L^{q}(\Omega)}\left\Arrowvert\sqrt{M_{j,k}(u_{j,k})}\nabla\frac{\mu_{j,k}}{g_{j,k}(u_{j,k})}\right\Arrowvert_{L^{2}(\tilde{D}_{j})}
Climjlimkmax{(2δj)m/2,θj,km/2}\displaystyle\leq C\lim_{j\rightarrow\infty}\lim_{k\rightarrow\infty}\max\left\{(2\delta_{j})^{m/2},\theta_{j,k}^{m/2}\right\}
=\displaystyle= 0.\displaystyle 0.

For the second term, we see that

(δj2)mDj|μj,kgj,k(uj,k)|2𝑑x𝑑t\displaystyle\left(\frac{\delta_{j}}{2}\right)^{m}\int_{D_{j}}|\nabla\frac{\mu_{j,k}}{g_{j,k}(u_{j,k})}|^{2}dxdt
\displaystyle\leq DjMj,k(uj,k)|μj,kgj,k(uj,k)|2𝑑x𝑑t\displaystyle\int_{D_{j}}M_{j,k}(u_{j,k})|\nabla\frac{\mu_{j,k}}{g_{j,k}(u_{j,k})}|^{2}dxdt
\displaystyle\leq ΩTMj,k(uj,k)|μj,kgj,k(uj,k)|2𝑑x𝑑tC.\displaystyle\int_{\Omega_{T}}M_{j,k}(u_{j,k})|\nabla\frac{\mu_{j,k}}{g_{j,k}(u_{j,k})}|^{2}dxdt\leq C.

Therefore μj,kgj,k(uj,k)\nabla\frac{\mu_{j,k}}{g_{j,k}(u_{j,k})} is bounded in L2(Dj)L^{2}(D_{j}) and we can extract a further subsequence, not relabeled, which converges weakly to some ξjL2(Dj)\xi_{j}\in L^{2}(D_{j}). Since DjD_{j} is an increasig sequence of sets with limjDj=D\lim_{j\rightarrow\infty}D_{j}=D, we have ξj=ξj1\xi_{j}=\xi_{j-1} a.e. in Dj1D_{j-1}. By setting ξj=0\xi_{j}=0 outside DjD_{j}, we can extend ξj\xi_{j} to a L2L^{2} function ξ~j\tilde{\xi}_{j} defined in DD. Therefore for a.e. xDx\in D, there exists a limit of ξ~j(x)\tilde{\xi}_{j}(x) as jj\rightarrow\infty. Let ξ(x)=limjξ~j(x)\xi(x)=\lim_{j\rightarrow\infty}\tilde{\xi}_{j}(x), we see that ξ(x)=ξj(x)\xi(x)=\xi_{j}(x) for a.e xDjx\in D_{j} and for all jj.

By a standard diagonal argument, we can extract a subsequnce such that

(112) μk,Nkgk,Nk(uk,Nk)ζ weakly in L2(Dj) for all j.\nabla\frac{\mu_{k,N_{k}}}{g_{k,N_{k}}(u_{k,N_{k}})}\rightharpoonup\zeta\text{ weakly in }L^{2}(D_{j})\text{ for all }j.

By strong convergence of Mi(ui)\sqrt{M_{i}(u_{i})} to M(u)\sqrt{M(u)} in C([0,T];Lβ(Ω))C([0,T];L^{\beta}(\Omega)) for 1β<1\leq\beta<\infty, we obtain

χDjMk,Nk(uk,Nk)μk,Nkgk,Nk(uk,Nk)χDjM(u)ζ\chi_{D_{j}}\sqrt{M_{k,N_{k}}(u_{k,N_{k}})}\nabla\frac{\mu_{k,N_{k}}}{g_{k,N_{k}}(u_{k,N_{k}})}\rightharpoonup\chi_{D_{j}}\sqrt{M(u)}\zeta

weakly in L2(0,T;Lq(Ω))L^{2}(0,T;L^{q}(\Omega)) for 1q<21\leq q<2 and all jj. Recall Mi(ui)μigi(ui)ξ\sqrt{M_{i}(u_{i})}\nabla\frac{\mu_{i}}{g_{i}(u_{i})}\rightarrow\xi weakly in L2(ΩT)L^{2}(\Omega_{T}), we have ξ=M(u)ζ\xi=\sqrt{M(u)}\zeta in DjD_{j} for all jj. Hence ξ=M(u)ζ\xi=\sqrt{M(u)}\zeta in DD and consequently

χDMk,Nk(uk,Nk)μk,Nkgk,Nk(uk,Nk)χDM(u)ζ\chi_{D}M_{k,N_{k}}(u_{k,N_{k}})\nabla\frac{\mu_{k,N_{k}}}{g_{k,N_{k}}(u_{k,N_{k}})}\rightharpoonup\chi_{D}M(u)\zeta

weakly in L2(0,T;Lq(Ω))L^{2}(0,T;L^{q}(\Omega)) for 1q<21\leq q<2.

Replacing uiu_{i} by subsequence uk,Nku_{k,N_{k}} in (110) and letting kk\rightarrow\infty then jj\rightarrow\infty, we have

(113) ΩTM(u)ξΨ𝑑x𝑑t\displaystyle\int_{\Omega_{T}}\sqrt{M(u)}\xi\cdot\Psi dxdt =\displaystyle= limjDjM(u)ζΨ𝑑x𝑑t\displaystyle\lim_{j\rightarrow\infty}\int_{D_{j}}M(u)\zeta\cdot\Psi dxdt
=\displaystyle= DM(u)ζΨ𝑑x𝑑t.\displaystyle\int_{D}M(u)\zeta\cdot\Psi dxdt.

It follows from (103) and (113) that

(114) 0Tg(u)tu,ϕ((W1,q(Ω)),W1,q(Ω))𝑑t=DM(u)ζϕdxdt\int_{0}^{T}\left<g(u)\partial_{t}u,\phi\right>_{((W^{1,q}(\Omega))^{\prime},W^{1,q}(\Omega))}dt=-\int_{D}M(u)\zeta\cdot\nabla\phi dxdt

for all ϕLp(0,T;W1,q(Ω))\phi\in L^{p}(0,T;W^{1,q}(\Omega)) where p,q>2p,q>2.

4.3 Relation between ζ\zeta and uu

The desired relation between ζ\zeta and uu is

(115) ζ\displaystyle\zeta =\displaystyle= 1gμμg(u)g2(u)u\displaystyle\frac{1}{g}\nabla\mu-\mu\frac{g^{\prime}(u)}{g^{2}(u)}\nabla u
=\displaystyle= 1g(Δu+q′′(u)u)g(u)g2(u)u(Δu+q(u)).\displaystyle\frac{1}{g}\left(-\nabla\Delta u+q^{\prime\prime}(u)\nabla u\right)-\frac{g^{\prime}(u)}{g^{2}(u)}\nabla u\left(-\Delta u+q^{\prime}(u)\right).

Given the known regularity uL(0,T;H1(Ω))u\in L^{\infty}(0,T;H^{1}(\Omega)) and degeneracy of g(u)g(u), the right hand side of (115) might not be defined as a function. We can, however, under the additional assumption uL2(0,T;H2(Ω))u\in L^{2}(0,T;H^{2}(\Omega)) and suitable assumptions on integrability of Δu\nabla\Delta u, find an explicit expression of ζ\zeta in terms of (115) in suitable subset of ΩT\Omega_{T}.

Claim I: If uL2(0,T:H2(Ω))u\in L^{2}(0,T:H^{2}(\Omega)) and for some jj, the interior of DjD_{j}, denoted by (Dj)(D_{j})^{\circ}, is not empty, then

ΔuL1((Dj))\nabla\Delta u\in L^{1}((D_{j})^{\circ})

and

ζ=Δu+q′′(u)ug(u)g(u)g2(u)(Δu+q(u))u a.e. in (Dj).\zeta=\frac{-\nabla\Delta u+q^{\prime\prime}(u)\nabla u}{g(u)}-\frac{g^{\prime}(u)}{g^{2}(u)}\left(-\Delta u+q^{\prime}(u)\right)\nabla u\text{ a.e. in }(D_{j})^{\circ}.

Proof of the claim I. Since uL2(0,T;H2(Ω))u\in L^{2}(0,T;H^{2}(\Omega)), we can have a subsequence, not relabeled such that, uk,Nku_{k,N_{k}} converges weakly to uu in L2(0,T;H2(Ω))L^{2}(0,T;H^{2}(\Omega)). Since

(116) μk,Nk=Δuk,Nk+q(uk,Nk) in ΩT,\mu_{k,N_{k}}=-\Delta u_{k,N_{k}}+q^{\prime}(u_{k,N_{k}})\text{ in }\Omega_{T},

The right hand side of (116) weakly converges to Δu+q(u)-\Delta u+q^{\prime}(u) in L2(ΩT)L^{2}(\Omega_{T}). Hence

μk,Nkμ=Δu+q(u) weakly in L2(ΩT).\mu_{k,N_{k}}\rightharpoonup\mu=-\Delta u+q^{\prime}(u)\text{ weakly in }L^{2}(\Omega_{T}).

On the other hand, using uk,Nku_{k,N_{k}} and uu as test functions in (82) yield

ΩTμk,Nkuk,Nk𝑑x𝑑t=ΩT(|uk,Nk|2+q(uk,Nk)uk,Nk)𝑑x𝑑t\displaystyle\int_{\Omega_{T}}\mu_{k,N_{k}}u_{k,N_{k}}dxdt=\int_{\Omega_{T}}\left(\left\arrowvert\nabla u_{k,N_{k}}\right\arrowvert^{2}+q^{\prime}(u_{k,N_{k}})u_{k,N_{k}}\right)dxdt
ΩTμk,Nku𝑑x𝑑t=ΩT(uk,Nku+q(uk,Nk)u)𝑑x𝑑t.\displaystyle\int_{\Omega_{T}}\mu_{k,N_{k}}udxdt=\int_{\Omega_{T}}\left(\nabla u_{k,N_{k}}\cdot\nabla u+q^{\prime}(u_{k,N_{k}})u\right)dxdt.

Passing to the limit, by (102), growth assumptions on qq^{\prime} and (116), we have

limkΩT|uk,Nk|2=ΩT|u|2.\lim_{k\rightarrow\infty}\int_{\Omega_{T}}\left\arrowvert\nabla u_{k,N_{k}}\right\arrowvert^{2}=\int_{\Omega_{T}}\left\arrowvert\nabla u\right\arrowvert^{2}.

Therefore

uk,Nku strongly in L2(ΩT).\nabla u_{k,N_{k}}\rightarrow\nabla u\text{ strongly in }L^{2}(\Omega_{T}).

Since uk,NkL2(0,T;W3,s(Ω))u_{k,N_{k}}\in L^{2}(0,T;W^{3,s}(\Omega)), we can differentiate (116) and get

(117) μk,Nk=Δuk,Nk+q′′(uk,Nk)uk,Nk,\nabla\mu_{k,N_{k}}=-\nabla\Delta u_{k,N_{k}}+q^{\prime\prime}(u_{k,N_{k}})\nabla u_{k,N_{k}},

and

(118) μk,Nkgk,Nk(uk,Nk)=1gk,Nk(uk,Nk)μk,Nkμk,Nkgk,Nk(uk,Nk)gk,Nk2(uk,Nk)uk,Nk\nabla\frac{\mu_{k,N_{k}}}{g_{k,N_{k}}(u_{k,N_{k}})}=\frac{1}{g_{k,N_{k}}(u_{k,N_{k}})}\nabla\mu_{k,N_{k}}-\mu_{k,N_{k}}\frac{g^{\prime}_{k,N_{k}}(u_{k,N_{k}})}{g^{2}_{k,N_{k}}(u_{k,N_{k}})}\nabla u_{k,N_{k}}

on DjD_{j}^{\circ}. Thus

(119) μk,Nk=gk,Nk(uk,Nk)μk,Nkgk,Nk(uk,Nk)+μk,Nkgk,Nk(uk,Nk)gk,Nk(uk,Nk)uk,Nk.\nabla\mu_{k,N_{k}}=g_{k,N_{k}}(u_{k,N_{k}})\nabla\frac{\mu_{k,N_{k}}}{g_{k,N_{k}}(u_{k,N_{k}})}+\frac{\mu_{k,N_{k}}}{g_{k,N_{k}}(u_{k,N_{k}})}g^{\prime}_{k,N_{k}}(u_{k,N_{k}})\nabla u_{k,N_{k}}.

Since

gk,Nk(uk,Nk)g(u) uniformly in Dj,\displaystyle g_{k,N_{k}}(u_{k,N_{k}})\rightarrow g(u)\text{ uniformly in }D_{j}^{\circ},
gk,Nk(uk,Nk)gk,Nk(uk,Nk)g(u)g(u) uniformly in Dj,\displaystyle\frac{g^{\prime}_{k,N_{k}}(u_{k,N_{k}})}{g_{k,N_{k}}(u_{k,N_{k}})}\rightarrow\frac{g^{\prime}(u)}{g(u)}\text{ uniformly in }D_{j}^{\circ},
μk,Nkgk,Nk(uk,Nk)ζ weakly in L2(Dj),\displaystyle\nabla\frac{\mu_{k,N_{k}}}{g_{k,N_{k}}(u_{k,N_{k}})}\rightharpoonup\zeta\text{ weakly in }L^{2}(D_{j}^{\circ}),
μk,Nkμ weakly in L2(ΩT),\displaystyle\mu_{k,N_{k}}\rightharpoonup\mu\text{ weakly in }L^{2}(\Omega_{T}),
uk,Nku strongly in L2(ΩT),\displaystyle\nabla u_{k,N_{k}}\rightarrow\nabla u\text{ strongly in }L^{2}(\Omega_{T}),

we have, for any ϕL(Dj)\phi\in L^{\infty}(D_{j}^{\circ}),

Djϕ(gk,Nk(uk,Nk)μk,Nkgk,Nk(uk,Nk)+μk,Nkgk,Nk(uk,Nk)gk,Nk(uk,Nk)uk,Nk)𝑑x𝑑t\displaystyle\int_{D_{j}^{\circ}}\phi\left(g_{k,N_{k}}(u_{k,N_{k}})\nabla\frac{\mu_{k,N_{k}}}{g_{k,N_{k}}(u_{k,N_{k}})}+\frac{\mu_{k,N_{k}}}{g_{k,N_{k}}(u_{k,N_{k}})}g^{\prime}_{k,N_{k}}(u_{k,N_{k}})\nabla u_{k,N_{k}}\right)dxdt
Djϕ(g(u)ζ+g(u)g(u)μu)𝑑x𝑑t,\displaystyle\rightarrow\int_{D_{j}^{\circ}}\phi\left(g(u)\zeta+\frac{g^{\prime}(u)}{g(u)}\mu\nabla u\right)dxdt,

i.e.

μk,Nkη\coloneqg(u)ζ+g(u)g(u)μu weakly in L1(Dj).\nabla\mu_{k,N_{k}}\rightharpoonup\eta\coloneq g(u)\zeta+\frac{g^{\prime}(u)}{g(u)}\mu\nabla u\ \text{ weakly in }L^{1}(D_{j}^{\circ}).

Passing to the limit in (117), we obtain, in the sense of distribution, that

η=Δu+q′′(u)u.\eta=-\nabla\Delta u+q^{\prime\prime}(u)\nabla u.

Since q′′(u)uL2(ΩT)q^{\prime\prime}(u)\nabla u\in L^{2}(\Omega_{T}), we have ΔuL1(Dj)-\nabla\Delta u\in L^{1}(D_{j}^{\circ}), hence

(120) η=Δu+q′′(u)u a.e. in Dj\eta=-\nabla\Delta u+q^{\prime\prime}(u)\nabla u\text{ a.e. in }D_{j}^{\circ}

Since 1gk,Nk(uk,Nk)1g(u)\frac{1}{g_{k,N_{k}}(u_{k,N_{k}})}\rightarrow\frac{1}{g(u)} uniformly in DjD_{j}, we have

1gk,Nkμk,Nk1g(u)η weakly in L1(Dj).\frac{1}{g_{k,N_{k}}}\nabla\mu_{k,N_{k}}\rightharpoonup\frac{1}{g(u)}\eta\text{ weakly in }L^{1}(D_{j}^{\circ}).

Since gk,Nk(uk,Nk)gk,Nk2(uk,Nk)g(u)g2(u)\frac{g^{\prime}_{k,N_{k}}(u_{k,N_{k}})}{g^{2}_{k,N_{k}}(u_{k,N_{k}})}\rightarrow\frac{g^{\prime}(u)}{g^{2}(u)} uniformly in DjD_{j}, we have

gk,Nk(uk,Nk)gk,Nk2(uk,Nk)μk,Nkuk,Nkg(u)g2(u)μu weakly in L1(Dj).\frac{g^{\prime}_{k,N_{k}}(u_{k,N_{k}})}{g^{2}_{k,N_{k}}(u_{k,N_{k}})}\mu_{k,N_{k}}\nabla u_{k,N_{k}}\rightharpoonup\frac{g^{\prime}(u)}{g^{2}(u)}\mu\nabla u\text{ weakly in }L^{1}(D_{j}^{\circ}).

Passing to the limit in (118), we have

ζ\displaystyle\zeta =\displaystyle= 1g(u)ημg(u)g2(u)u=Δu+q′′(u)ug(u)g(u)g2(u)(Δu+q(u))u\displaystyle\frac{1}{g(u)}\eta-\mu\frac{g^{\prime}(u)}{g^{2}(u)}\nabla u=\frac{-\nabla\Delta u+q^{\prime\prime}(u)\nabla u}{g(u)}-\frac{g^{\prime}(u)}{g^{2}(u)}\left(-\Delta u+q^{\prime}(u)\right)\nabla u

on (Dj)(D_{j})^{\circ}. Noticing the value of ζ\zeta on ΩT\D\Omega_{T}\backslash D doesn’t matter since it does not appear on the right hand side of (113).

Claim II: For any open set UΩTU\in\Omega_{T} in which ΔuLp(U)\nabla\Delta u\in L^{p}(U) for some p>1p>1 and g(u)>0g(u)>0, we have

(121) ζ=Δu+q′′(u)ug(u)g(u)g2(u)(Δu+q(u))u.\zeta=\frac{-\nabla\Delta u+q^{\prime\prime}(u)\nabla u}{g(u)}-\frac{g^{\prime}(u)}{g^{2}(u)}\left(-\Delta u+q^{\prime}(u)\right)\nabla u.

in UU.

To prove this, since

(122) μk,Nk=Δuk,Nk+q′′(uk,Nk)uk,Nk in ΩT\nabla\mu_{k,N_{k}}=-\nabla\Delta u_{k,N_{k}}+q^{\prime\prime}(u_{k,N_{k}})\nabla u_{k,N_{k}}\text{ in }\Omega_{T}

and

(123) μk,Nkgk,Nk(uk,Nk)=1gk,Nk(uk,Nk)μk,Nk+μk,Nk1gk,Nk(uk,Nk) on Dj.\nabla\frac{\mu_{k,N_{k}}}{g_{k,N_{k}}(u_{k,N_{k}})}=\frac{1}{g_{k,N_{k}}(u_{k,N_{k}})}\nabla\mu_{k,N_{k}}+\mu_{k,N_{k}}\cdot\nabla\frac{1}{g_{k,N_{k}}(u_{k,N_{k}})}\text{ on }D_{j}.

The right hand side of (122) converges weakly to Δu+q′′(u)u-\nabla\Delta u+q^{\prime\prime}(u)\nabla u in Lq(U)L^{q}(U) for q=min{p,2}>1q=\min\{p,2\}>1. Hence

μk,Nkη=Δu+q′′(u)u weakly in Lq(U).\nabla\mu_{k,N_{k}}\rightharpoonup\eta=-\nabla\Delta u+q^{\prime\prime}(u)\nabla u\text{ weakly in }L^{q}(U).

The right hand side of (123) converges weakly to

ηg(u)g(u)g2(u)μu\frac{\eta}{g(u)}-\frac{g^{\prime}(u)}{g^{2}(u)}\mu\cdot\nabla u

in L1(UDj)L^{1}(U\cap D_{j}) for each jj and therefore

ζ=Δu+q′′(u)ug(u)g(u)g2(u)(Δu+q(u))u\zeta=\frac{-\nabla\Delta u+q^{\prime\prime}(u)\nabla u}{g(u)}-\frac{g^{\prime}(u)}{g^{2}(u)}\left(-\Delta u+q^{\prime}(u)\right)\nabla u

in UDU\cap D. The definition of ζ\zeta can be extended to U\DU\backslash D by our integrability assumption on uu. Define

Ω~T={UΩT:g(u)>0 on U and ΔuLp(U) for some p>1 depending on U}.\tilde{\Omega}_{T}=\{U\subset\Omega_{T}:g(u)>0\text{ on }U\text{ and }\nabla\Delta u\in L^{p}(U)\text{ for some }p>1\text{ depending on }U\}.

Then Ω~T\tilde{\Omega}_{T} is open and ζ\zeta is defined by (121) on Ω~T\tilde{\Omega}_{T}. Since |ΩT\B|=0|\Omega_{T}\backslash B|=0 , M(u)=0M(u)=0 on ΩT\P\Omega_{T}\backslash P and

ΩT\{DΩ~T}{ΩT\B}{ΩT\P},\Omega_{T}\backslash\{D\cup\tilde{\Omega}_{T}\}\subset\{\Omega_{T}\backslash B\}\cup\{\Omega_{T}\backslash P\},

we can take the value of ζ\zeta to be zero outside DΩTD\cup\Omega_{T}, and it won’t affect the integral on the right side of (13).

Lastly the energy inequality (15) follows by taking limit in the energy inequality for uk,Nku_{k,N_{k}}.

Remark 4.2.

In Cahn-Hilliard case, there is convergence of μk\nabla\mu_{k} on L2(Dj)L^{2}(D_{j}), and relation between ξ\xi and uu can be derived directly. Here we only have convergence of μkgk(uk)\nabla\frac{\mu_{k}}{g_{k}(u_{k})} on L2(Dj)L^{2}(D_{j}). In order to obtain convergence of μk\nabla\mu_{k}, we need convergence μk\mu_{k} on Lp(ΩT)L^{p}(\Omega_{T}) for suitable pp, this is where we used the additional assumption uL2(0,T;H2(Ω))u\in L^{2}(0,T;H^{2}(\Omega)).

5 A Modified phase field model for self-climb of prismatic dislocation loops

Dislocations are line defects in crystals [12, 23]. A phase field model [20] was derived based on the pipe diffusion model for self-climb of prismatic dislocation loops [18, 19] that describes the conservative climb of dislocation loops observed in experiments of irradiated materials [13, 12, 8]. In this section, we study the wellposedness of the following modified phase field model for self-climb of prismatic dislocation loops:

(124) g(u)tu\displaystyle g(u)\partial_{t}u =\displaystyle= (M(u)μg(u)) for xΩ2,t[0,)\displaystyle\nabla\cdot(M(u)\nabla\frac{\mu}{g(u)})\text{ for }x\in\Omega\subset\mathbb{R}^{2},t\in[0,\infty)
(125) μ\displaystyle\mu =\displaystyle= Δu+1ε2q(u)+1εfcl\displaystyle-\Delta u+\frac{1}{\varepsilon^{2}}q^{\prime}(u)+\frac{1}{\varepsilon}f_{cl}

Where M(u)=M0g(u)M(u)=M_{0}g(u), g(u)=|1u2|mg(u)=|1-u^{2}|^{m} for 2m<2\leq m<\infty, q(u)q(u) satisfy same assumptions (6)-(7) as those for Eqs. (1)-(2). Here fclf_{cl} is the total climb force with

fcl=fcld+fclappf_{cl}=f_{cl}^{d}+f_{cl}^{app}

where fclappf_{cl}^{app} is the applied climb force, and

(126) fcld(x,y,u)=Gb24π(1ν)Ω(xx¯R3ux¯+yy¯R3uy¯)𝑑x¯𝑑y¯f_{cl}^{d}(x,y,u)=\frac{Gb^{2}}{4\pi(1-\nu)}\int_{\Omega}\left(\frac{x-\overline{x}}{R^{3}}u_{\overline{x}}+\frac{y-\overline{y}}{R^{3}}u_{\overline{y}}\right)d\overline{x}d\overline{y}

represents the climb force generated by all the dislocations. Here Ω2\Omega\subset\mathbb{R}^{2} is a bounded domain, GG is the shear modulus, ν\nu is the Poisson ratio, and R=(xx¯)2+(yy¯)2R=\sqrt{(x-\overline{x})^{2}+(y-\overline{y})^{2}}. In this model, we assume that the prismatic dislocation loops lie and evlove by self-climb in the xyxy plane and all dislocation loops have the same Burgers vector 𝐛=(0,0,b)\mathbf{b}=(0,0,b).

The chemical potential μ\mu comes from variations of the classical Cahn-Hilliard energy and the elastic energy due to dislocations, i.e.

(127) μ=δECHδu+δEelδu,\mu=\frac{\delta E_{CH}}{\delta u}+\frac{\delta E_{el}}{\delta u},

where

(128) ECH(u)=Ω(12|u|2+q(u))𝑑x,\displaystyle E_{CH}(u)=\int_{\Omega}\left(\frac{1}{2}|\nabla u|^{2}+q(u)\right)dx,
(129) Eel=Ω(12ufcld+ufclapp)𝑑x\displaystyle E_{el}=\int_{\Omega}\left(\frac{1}{2}uf^{d}_{cl}+uf_{cl}^{app}\right)dx

are classical Cahn-Hilliard energy and elastic energy, respectively. Under periodic boundary conditions, the climb force generated by the dislocations can be expressed as

(130) fcld(x,y,u)=Gb22(1ν)(Δ)12u.\displaystyle f^{d}_{cl}(x,y,u)=\frac{Gb^{2}}{2(1-\nu)}(-\Delta)^{\frac{1}{2}}u.

Here (Δ)su(-\Delta)^{s}u is a fractional operator defined by

((Δ)sf)=(ξ12+ξ22)s2(f)(ξ)\mathcal{F}((-\Delta)^{s}f)=(\xi_{1}^{2}+\xi_{2}^{2})^{\frac{s}{2}}\mathcal{F}(f)(\xi)

for ξ2\xi\in\mathbb{Z}^{2}. In the analysis below, without loss of generality, we set the coefficient of the climb force Gb22(1ν)=1\frac{Gb^{2}}{2(1-\nu)}=1.

System (124)-(125) is a modified version of the phase field model introduced in [20], which does not have the g(u)g(u) term on the left side of (124). Putting an extra factor h=H0gh=H_{0}g in front of the nonlocal climb force fcldf^{d}_{cl}, the asymptotic analysis in [20] showed that the proposed phase field model yields accurate dislocation self-climb velocity in the sharp interface limit. Moreover, numerical simulations in [20] showed excellent agreement with experimental observations and discrete dislocation dynamics simulation results. Now we prove the wellposedness of the modified model (124)-(125). There is an extra nonlocal term fcldf^{d}_{cl} in this model compared with the model considered in previous sections.

Define

uN(x,t)=j=1NcjN(t)ϕj(x),μN(x,t)=j=1NdjN(t)ϕj(x),u^{N}(x,t)=\sum_{j=1}^{N}c_{j}^{N}(t)\phi_{j}(x),\hskip 36.135pt\mu^{N}(x,t)=\sum_{j=1}^{N}d^{N}_{j}(t)\phi_{j}(x),

where {cjN,djN}\{c_{j}^{N},d^{N}_{j}\} satisfy

(131) ΩtuNϕjdx\displaystyle\int_{\Omega}\partial_{t}u^{N}\phi_{j}dx =\displaystyle= ΩMθ(uN)μNgθ(uN)ϕjgθ(uN)dx,\displaystyle-\int_{\Omega}M_{\theta}(u^{N})\nabla\frac{\mu^{N}}{g_{\theta}(u^{N})}\cdot\nabla\frac{\phi_{j}}{g_{\theta}(u^{N})}dx,
(132) ΩμNϕj𝑑x\displaystyle\int_{\Omega}\mu^{N}\phi_{j}dx =\displaystyle= Ω(uNϕj+q(uN)ϕj+ϕj(Δ)12uN)𝑑x,\displaystyle\int_{\Omega}\left(\nabla u^{N}\cdot\nabla\phi_{j}+q^{\prime}(u^{N})\phi_{j}+\phi_{j}(-\Delta)^{\frac{1}{2}}u^{N}\right)dx,
(133) uN(x,0)\displaystyle u^{N}(x,0) =\displaystyle= j=1N(Ωu0ϕj𝑑x)ϕj(x).\displaystyle\sum_{j=1}^{N}\left(\int_{\Omega}u_{0}\phi_{j}dx\right)\phi_{j}(x).

(131)-(133) is an initial value problem for a system of ordinary equations for {cjN(t)}\{c_{j}^{N}(t)\}. Since right hand side of (131) is continuous in cjNc_{j}^{N}, the system has a local solution.

Define energy functional

F(u)=Ω{12|u|2+q(u)+|(Δ)14u|2}𝑑x.F(u)=\int_{\Omega}\left\{\frac{1}{2}|\nabla u|^{2}+q(u)+|(-\Delta)^{\frac{1}{4}}u|^{2}\right\}dx.

Direct calculation yields

ddtF(uN(x,t))=ΩMθ(uN)|μNgθ(uN)|2𝑑x,\frac{d}{dt}F(u^{N}(x,t))=-\int_{\Omega}M_{\theta}(u^{N})\left|\nabla\frac{\mu^{N}}{g_{\theta}(u^{N})}\right|^{2}dx,

integration over tt gives the following energy identity for any t>0t>0

Ω(12|uN(x,t)|2+q(uN(x,t))+uN(Δ)12uN)𝑑x\displaystyle\int_{\Omega}\left(\frac{1}{2}|\nabla u^{N}(x,t)|^{2}+q(u^{N}(x,t))+u^{N}(-\Delta)^{\frac{1}{2}}u^{N}\right)dx
+0tΩMθ(uN(x,τ))|μN(x,τ)gθ(uN(x,τ))|2𝑑x𝑑τ\displaystyle+\int_{0}^{t}\int_{\Omega}M_{\theta}(u^{N}(x,\tau))\left|\nabla\frac{\mu^{N}(x,\tau)}{g_{\theta}(u^{N}(x,\tau))}\right|^{2}dxd\tau
=\displaystyle= Ω(12|uN(x,0)|2+q(uN(x,0))+uN(x,0)(Δ)12uN(x,0))𝑑x\displaystyle\int_{\Omega}\left(\frac{1}{2}|\nabla u^{N}(x,0)|^{2}+q(u^{N}(x,0))+u^{N}(x,0)(-\Delta)^{\frac{1}{2}}u^{N}(x,0)\right)dx
\displaystyle\leq Ωu0L2(Ω)2+C(u0H1Ωr+1+|Ω|)+12u0L2(Ω)2C<,\displaystyle\int_{\Omega}\left\Arrowvert\nabla u_{0}\right\Arrowvert_{L^{2}(\Omega)}^{2}+C\left(\left\Arrowvert u_{0}\right\Arrowvert^{r+1}_{H^{1}{\Omega}}+|\Omega|\right)+\frac{1}{2}\left\Arrowvert u_{0}\right\Arrowvert_{L^{2}(\Omega)}^{2}\leq C<\infty,

where CC represents a generic constant possibly depending only on TT, Ω\Omega, u0u_{0} but not on θ\theta. Since Ω\Omega is bounded region, by growth assumption assumption (6) and Poincare’s inequality, the energy identity (5) implies uNL(0,T;H1(Ω))u^{N}\in L^{\infty}(0,T;H^{1}(\Omega)) with

(135) uNL(0,T;H1(Ω))C for all N,\left\Arrowvert u^{N}\right\Arrowvert_{L^{\infty}(0,T;H^{1}(\Omega))}\leq C\text{ for all }N,

and

(136) Mθ(uN)μNgθ(uN)L2(ΩT)C for all N.\left\Arrowvert\sqrt{M_{\theta}(u^{N})}\nabla\frac{\mu^{N}}{g_{\theta}(u^{N})}\right\Arrowvert_{L^{2}(\Omega_{T})}\leq C\text{ for all }N.

Repeat the argument in Section 3 and Section 4, replacing energy functional F(u)F(u) by E(u)E(u) when necessary, we can prove the following existence theorem for (124)-(125) with nondegenerate and degenerate mobilities respectively.

Theorem 5.1.

Let Mθ,gθM_{\theta},g_{\theta} be defined by (9) and (10), under the assumptions (6)-(8), for any u0H1(Ω)u_{0}\in H^{1}(\Omega) and any T>0T>0, there exists a function uθu_{\theta} such that

  • a)

    uθL(0,T;H1(Ω))C([0,T];Lp(Ω))L2(0,T;W3,s(Ω))u_{\theta}\in L^{\infty}(0,T;H^{1}(\Omega))\cap C([0,T];L^{p}(\Omega))\cap L^{2}(0,T;W^{3,s}(\Omega)), where 1p<1\leq p<\infty, 1s<21\leq s<2,

  • b)

    tuθL2(0,T;(W1,q(Ω)))\partial_{t}u_{\theta}\in L^{2}(0,T;(W^{1,q}(\Omega))^{\prime}) for q>2q>2,

  • c)

    uθ(x,0)=u0(x)u_{\theta}(x,0)=u_{0}(x) for all xΩx\in\Omega,

which satisfies (4)-(5) in the following weak sense

0T<tuθ,ϕ>(W1,q(Ω)),W1,q(Ω)dt\displaystyle\int_{0}^{T}<\partial_{t}u_{\theta},\phi>_{(W^{1,q}(\Omega))^{\prime},W^{1,q}(\Omega)}dt
(137) =0TΩMθ(uθ)Δuθ+q(uθ)+(Δ)12uθgθ(uθ)ϕgθ(uθ)dxdt\displaystyle=-\int_{0}^{T}\int_{\Omega}M_{\theta}(u_{\theta})\nabla\frac{-\Delta u_{\theta}+q^{\prime}(u_{\theta})+(-\Delta)^{\frac{1}{2}}u_{\theta}}{g_{\theta}(u_{\theta})}\cdot\nabla\frac{\phi}{g_{\theta}(u_{\theta})}dxdt

for all ϕL2(0,T;W1,q(Ω))\phi\in L^{2}(0,T;W^{1,q}(\Omega)) with q>2q>2. In addition, the following energy inequality holds for all t>0t>0.

(138) Ω(12|uθ(x,t)|2+q(uθ(x,t))+uθ(x,t)(Δ)12uθ)𝑑x\displaystyle\int_{\Omega}\left(\frac{1}{2}|\nabla u_{\theta}(x,t)|^{2}+q(u_{\theta}(x,t))+u_{\theta}(x,t)(-\Delta)^{\frac{1}{2}}u_{\theta}\right)dx
+0tΩMθ(uθ(x,τ)|Δuθ(x,τ)+q(uθ(x,τ)+(Δ)12uθgθ(uθ(x,τ))|2dxdτ\displaystyle+\int_{0}^{t}\int_{\Omega}M_{\theta}(u_{\theta}(x,\tau)\left|\nabla\frac{-\Delta u_{\theta}(x,\tau)+q^{\prime}(u_{\theta}(x,\tau)+(-\Delta)^{\frac{1}{2}}u_{\theta}}{g_{\theta}(u_{\theta}(x,\tau))}\right|^{2}dxd\tau
\displaystyle\leq Ω(12|u0(x)|2+q(u0(x))+u0(x)(Δ)12u0)𝑑x.\displaystyle\int_{\Omega}\left(\frac{1}{2}|\nabla u_{0}(x)|^{2}+q(u_{0}(x))+u_{0}(x)(-\Delta)^{\frac{1}{2}}u_{0}\right)dx.

Theorem 5.2.

For any u0H1(Ω)u_{0}\in H^{1}(\Omega) and T>0T>0, there exists a function u:ΩT=Ω×[0,T]u:\Omega_{T}=\Omega\times[0,T]\rightarrow\mathbb{R} satisfying

  • i)

    uL(0,T;H1(Ω))C([0,T];Ls(Ω))u\in L^{\infty}(0,T;H^{1}(\Omega))\cap C([0,T];L^{s}(\Omega)), where 1s<1\leq s<\infty,

  • ii)

    g(u)tuLp(0,T;(W1,q(Ω)))g(u)\partial_{t}u\in L^{p}(0,T;(W^{1,q}(\Omega))^{\prime}) for 1p<21\leq p<2 and q>2q>2.

  • iii)

    u(x,0)=u0(x)u(x,0)=u_{0}(x) for all xΩx\in\Omega,

which solves (4)-(5) in the following weak sense

  • a)

    There exists a set BΩTB\subset\Omega_{T} with |ΩT\B|=0|\Omega_{T}\backslash B|=0 and a function ζ:ΩTn\zeta:\Omega_{T}\rightarrow\mathbb{R}^{n} satisfying χBPM(u)ζLpp1(0,T;Lqq1(Ω,n))\chi_{B\cap P}M(u)\zeta\in L^{\frac{p}{p-1}}(0,T;L^{\frac{q}{q-1}}(\Omega,\mathbb{R}^{n})) such that

    (139) 0T<g(u)tu,ϕ>(W1,q(Ω)),W1,q(Ω)dt=BPM(u)ζϕdxdt\int_{0}^{T}<g(u)\partial_{t}u,\phi>_{(W^{1,q}(\Omega))^{\prime},W^{1,q}(\Omega)}dt=-\int_{B\cap P}M(u)\zeta\cdot\nabla\phi dxdt

    for all ϕLp(0,T;W1,q(Ω))\phi\in L^{p}(0,T;W^{1,q}(\Omega)) with p,q>2p,q>2. Here P:={(x,t)ΩT:|1u2|0}P:=\{(x,t)\in\Omega_{T}:|1-u^{2}|\neq 0\} is the set where M(u),g(u)M(u),g(u) are nondegenerate and χBP\chi_{B\cap P} is the characteristic function of set BPB\cap P.

  • b)

    Assume uL2(0,T;H2(Ω)).u\in L^{2}(0,T;H^{2}(\Omega)). For any open set UΩTU\in\Omega_{T} on which g(u)>0g(u)>0 and ΔuLp(U)\nabla\Delta u\in L^{p}(U) for some p>1p>1, we have

    (140) ζ=Δu+q′′(u)u+(Δ)12ug(u)g(u)g2(u)(Δu+q(u)+(Δ)12u)u.\zeta=\frac{-\nabla\Delta u+q^{\prime\prime}(u)\nabla u+\nabla(-\Delta)^{\frac{1}{2}}u}{g(u)}-\frac{g^{\prime}(u)}{g^{2}(u)}\left(-\Delta u+q^{\prime}(u)+(-\Delta)^{\frac{1}{2}}u\right)\nabla u.

    a.e in UU.

Moreover, the following energy inequality holds for all t>0t>0

(141) Ω(12|u(x,t)|2+q(u(x,t)))𝑑z+ΩrBPM(u(x,τ))|ζ(x,τ)|2𝑑x𝑑τ\displaystyle\int_{\Omega}\left(\frac{1}{2}|\nabla u(x,t)|^{2}+q(u(x,t))\right)dz+\int_{\Omega_{r}\cap B\cap P}M(u(x,\tau))|\zeta(x,\tau)|^{2}dxd\tau
\displaystyle\leq Ω(12|u0(x)|2+q(u0(x)))𝑑x.\displaystyle\int_{\Omega}\left(\frac{1}{2}|\nabla u_{0}(x)|^{2}+q(u_{0}(x))\right)dx.

ACKNOWLEDGEMENTS X.H. Niu’s research is supported by National Natural Science Foundation of China under the grant number 11801214 and the Natural Science Foundation of Fujian Province of China under the grant number 2021J011193. Y. Xiang’s research is supported by the Hong Kong Research Grants Council General Research Fund 16307319. X. Yan’s research is supported by a Research Excellence Grant and CLAS Dean’s Summer Research Grant from University of Connecticut.

References

  • [1] M. Albani, L. Ghisalberti, R. Bergamaschini, M. Friedl, M. Salvalaglio, A. Voigt, F. Montalenti, G. Tütüncüoglu, A. F. i Morral, and L. Miglio, Growth kinetics and morphological analysis of homoepitaxial gaas fins by theory and experiment, Phys. Rev. Mater., 2 (2018), p. 093404.
  • [2] J. W. Cahn and J. E. Hilliard, Free energy of a nonuniform system. i. interfacial free energy, The Journal of Chemical Physics, 28 (1958), pp. 258–267.
  • [3]  , Spinodal decomposition: A reprise, Acta Metallurgica, 19 (1971), pp. 151–161.
  • [4] J. W. Cahn and J. Taylor, Overview no. 113 surface motion by surface diffusion, Acta Metallurgica et Materialia, 42 (1994), pp. 1045–1063.
  • [5] S. Dai and Q. Du, Motion of interfaces governed by the cahn–hilliard equation with highly disparate diffusion mobility, SIAM Journal on Applied Mathematics, 72 (2012), pp. 1818–1841.
  • [6]  , Coarsening mechanism for systems governed by the cahn–hilliard equation with degenerate diffusion mobility, Multiscale Modeling & Simulation, 12 (2014), pp. 1870–1889.
  • [7]  , Weak solutions for the Cahn-Hilliard equation with degenerate mobility, Arch. Ration. Mech. Anal., 219 (2016), pp. 1161–1184.
  • [8] S. Dudarev, Density functional theory models for radiation damage, Annu. Rev. Mater. Res., 43 (2013), p. 35–61.
  • [9] K. L. M. Elder, W. B. Andrews, M. Ziehmer, N. Mameka, C. Kirchlechner, A. Davydok, J.-S. Micha, A. F. Chadwick, E. T. Lilleodden, K. Thornton, and P. W. Voorhees, Grain boundary formation through particle detachment during coarsening of nanoporous metals, Proc. Natl. Acad. Sci., 118 (2021), p. e2104132118.
  • [10] C. M. Elliott and H. Garcke, On the cahn–hilliard equation with degenerate mobility, SIAM Journal on Mathematical Analysis, 27 (1996), pp. 404–423.
  • [11] C. Gugenberger, R. Spatschek, and K. Kassner, Comparison of phase-field models for surface diffusion., Physical review. E, Statistical, nonlinear, and soft matter physics, 78 1 Pt 2 (2008), p. 016703.
  • [12] J. Hirth and J. Lothe, Theory of Dislocations, McGraw-Hill, New York, 1982.
  • [13] F. Kroupa and P. B. Price, Conservative climb of a dislocation loop due to its interaction with an edge dislocation, Philos. Mag., 6 (1961), pp. 243–247.
  • [14] A. A. Lee, A. Münch, and E. Süli, Degenerate mobilities in phase field models are insufficient to capture surface diffusion, Applied Physics Letters, 107 (2015), p. 081603.
  • [15]  , Sharp-interface limits of the cahn-hilliard equation with degenerate mobility, SIAM J. Appl. Math., 76 (2016), pp. 433–456.
  • [16] J.-L. Lions, Quelques méthodes de résolution des problèmes aux limites non linéaires, Dunod; Gauthier-Villars, Paris, 1969.
  • [17] M. Naffouti, R. Backofen, M. Salvalaglio, T. Bottein, M. Lodari, A. Voigt, T. David, A. Benkouider, I. Fraj, L. Favre, A. Ronda, I. Berbezier, D. Grosso, M. Abbarchi, and M. Bollani, Complex dewetting scenarios of ultrathin silicon films for large-scale nanoarchitectures, Sci. Adv., 3 (2017), p. eaao1472.
  • [18] X. Niu, Y. Gu, and Y. Xiang, Dislocation dynamics formulation for self-climb of dislocation loops by vacancy pipe diffusion, Int. J. Plast., 120 (2019), pp. 262 – 277.
  • [19] X. Niu, T. Luo, J. Lu, and Y. Xiang, Dislocation climb models from atomistic scheme to dislocation dynamics, J. Mech. Phys. Solids, 99 (2017), pp. 242 – 258.
  • [20] X. Niu, Y. Xiang, and X. Yan, Phase field model for self-climb of prismatic dislocation loops by vacancy pipe diffusion, Int. J. Plast., 141 (2021), p. 102977.
  • [21] A. Rätz, A. Ribalta, and A. Voigt, Surface evolution of elastically stressed films under deposition by a diffuse interface model, J. Comput. Phys., 214 (2006), p. 187–208.
  • [22] J. Simon, Compact sets in the space lp(o,t;b)lp(o,t;b), Annali di Matematica Pura ed Applicata, 146 (1986), pp. 65–96.
  • [23] Y. Xiang, Modeling dislocations at different scales, Commun. Comput. Phys., 1 (2006), pp. 383–424.
  • [24] J. Yin, On the existence of nonnegative continuous solutions of the cahn-hilliard equation, Journal of Differential Equations, 97 (1992), pp. 310–327.