Interior regularity to the steady incompressible shear thinning fluids with non-Standard growth

  • Received: 01 November 2017 Revised: 01 March 2018
  • Primary: 35Q35; Secondary: 35B65, 76A05

  • We consider weak solutions to the equations of stationary motion of a class of non-Newtonian fluids which includes the power law model. The power depends on the spatial variable, which is motivated by electrorheological fluids. We prove the existence of second order derivatives of weak solutions in the shear thinning cases.

    Citation: Hyeong-Ohk Bae, Hyoungsuk So, Yeonghun Youn. Interior regularity to the steady incompressible shear thinning fluids with non-Standard growth[J]. Networks and Heterogeneous Media, 2018, 13(3): 479-491. doi: 10.3934/nhm.2018021

    Related Papers:

    [1] Hyeong-Ohk Bae, Hyoungsuk So, Yeonghun Youn . Interior regularity to the steady incompressible shear thinning fluids with non-Standard growth. Networks and Heterogeneous Media, 2018, 13(3): 479-491. doi: 10.3934/nhm.2018021
    [2] Grigory Panasenko, Ruxandra Stavre . Asymptotic analysis of a non-periodic flow in a thin channel with visco-elastic wall. Networks and Heterogeneous Media, 2008, 3(3): 651-673. doi: 10.3934/nhm.2008.3.651
    [3] Changli Yuan, Mojdeh Delshad, Mary F. Wheeler . Modeling multiphase non-Newtonian polymer flow in IPARS parallel framework. Networks and Heterogeneous Media, 2010, 5(3): 583-602. doi: 10.3934/nhm.2010.5.583
    [4] Iryna Pankratova, Andrey Piatnitski . Homogenization of convection-diffusion equation in infinite cylinder. Networks and Heterogeneous Media, 2011, 6(1): 111-126. doi: 10.3934/nhm.2011.6.111
    [5] María Anguiano, Renata Bunoiu . Homogenization of Bingham flow in thin porous media. Networks and Heterogeneous Media, 2020, 15(1): 87-110. doi: 10.3934/nhm.2020004
    [6] Andro Mikelić, Giovanna Guidoboni, Sunčica Čanić . Fluid-structure interaction in a pre-stressed tube with thick elastic walls I: the stationary Stokes problem. Networks and Heterogeneous Media, 2007, 2(3): 397-423. doi: 10.3934/nhm.2007.2.397
    [7] Rinaldo M. Colombo, Graziano Guerra . A coupling between a non--linear 1D compressible--incompressible limit and the 1D $p$--system in the non smooth case. Networks and Heterogeneous Media, 2016, 11(2): 313-330. doi: 10.3934/nhm.2016.11.313
    [8] Alessia Marigo . Equilibria for data networks. Networks and Heterogeneous Media, 2007, 2(3): 497-528. doi: 10.3934/nhm.2007.2.497
    [9] Leda Bucciantini, Angiolo Farina, Antonio Fasano . Flows in porous media with erosion of the solid matrix. Networks and Heterogeneous Media, 2010, 5(1): 63-95. doi: 10.3934/nhm.2010.5.63
    [10] Tom Freudenberg, Michael Eden . Homogenization and simulation of heat transfer through a thin grain layer. Networks and Heterogeneous Media, 2024, 19(2): 569-596. doi: 10.3934/nhm.2024025
  • We consider weak solutions to the equations of stationary motion of a class of non-Newtonian fluids which includes the power law model. The power depends on the spatial variable, which is motivated by electrorheological fluids. We prove the existence of second order derivatives of weak solutions in the shear thinning cases.



    Traditionally, the Navier-Stokes equations have received quite a bit of attention. Recently, attention to the behavior of fluids with various viscosities has been increasing dramatically. It is because we can find such fluids everywhere. For example, water, yogurt, lubricants, sand in water, ink, gum solutions, nail polish, ketchup, molasses, ice, paint, custard, paper pulp, even blood in our body. The behavior of many of them can be described in the power law model. In that sense, we are interested in the power law model, which is a generalized Navier-Stokes system.

    As mentioned in [16], electrorheological fluids are viscous liquids, that are characterized by their ability to undergo significant changes in their mechanical properties when an electric field is applied. The motion is governed by a system of partial differential equations consisting of electric field $E$, polarization, density $\rho$, velocity $u$, pressure $\pi$, and deviatoric stress tensor $S$. Refer to [11] for the detail description. In [11] the viscosity is the form of power law model, and the power $p$ depends on the electric field $p(|E|^2)$.

    In this article we consider the following stationary system related with non-Newtonian fluids:

    $\left\{\label{pde1} div{S(x,D(u))}+uu+π=g,divu=0
    \right. \text{ in }\Omega, $
    (1)

    where $g\in L^{\infty}(\Omega)$ is a given exterior force, and $\Omega$ is a bounded domain in $\mathbb{R}^{n}$ with $n = 2, 3$. Here, $D$ denotes the symmetric part of $\nabla$, given by

    $ D^{ij}(u)(x) = \dfrac{1}{2}\left(\partial_{i}u^{j}+\partial_{j}u^{i}\right)(x)\in \mathbb{R}^{n \times n}_{sym}. $

    In addition, $S(x, D(u))$ denotes the deviatoric stress. Then $T : = S- \pi I$ is called the full shear stress, such that $-{\rm{div}}\, T$ represents the sum of the internal forces due to friction, which depends mostly on the material of the fluid.

    The continuous deviatoric stress tensor $S:\Omega\times\mathbb{R}^{n\times n}\to\mathbb{R}^{n\times n}$ is assumed to be $C^{1}$-regular in the gradient variable $z$ for every $z\in \mathbb{R}^{n}\setminus \{0\}$, with $S_{z}(\cdot)$ being ${\rm{Carathéodory}}$ regular and satisfying the following nonstandard growth, monotonicity and continuity assumptions:

    $\left\{ |S(x,z)|+|Sz(x,z)|(|z|2+μ2)12L(|z|2+μ2)p(x)12,ν(|z1|2+|z2|2+μ2)p(x)22|z1z2|2S(x,z1)S(x,z2),z1z2,|S(x1,ξ)S(x2,ξ)|Lω(|x1x2|)[1+|log(|ξ|2+μ2)|]×[(|ξ|2+μ2)p(x1)12+(|ξ|2+μ2)p(x2)12]
    \right. $
    (2)

    for every $z\in \mathbb{R}^{n}\setminus \{0\}$, $z_{1}, z_{2}$, $\xi\in\mathbb{R}^{n}$ and $x, x_{1}, x_{2}\in\Omega$, where $0 <\nu\leq L$ and $\mu\ge 0$ are fixed numbers. The variable exponent function $p(\cdot):\Omega\to[0, +\infty)$ is continuous with modulus of continuity $\omega:[0, \infty)\to [0, \infty)$ satisfying

    $ \gamma_{1}\leq p(x) \leq \gamma_{2} \leq 2 ~~~~ {\rm{and}}~~~~ |p(x)-p(y)|\leq\omega(|x-y|) ~~~~ \textrm{for} ~~~~x, y\in\Omega, $ (3)

    where $\gamma_{1}>3/2$ if $n = 2$ and $\gamma_{1}>9/5$ if $n = 3$. We assume that the variable exponent $p(\cdot)$ is Lipschitz continuous, i.e.

    $ \omega(r)\leq c_pr, $

    for some constant $c_p>0$. For simplicity, we set

    $ p_{0} = p(x_{0}), ~~~~ p_{1}: = \inf\limits_{B_{4R}(x_{0})}p(x) ~~~~{\rm{and}}~~~~ p_{2}: = \sup\limits_{B_{4R}(x_{0})}p(x) $

    for some fixed $B_{4R}(x_{0})\subset\subset \Omega$.

    Here, we denote the variable exponent Lebesgue space $L^{p(\cdot)}(\Omega)$, by the set of all measurable functions $f:\Omega \to \mathbb{R}^{n}$ satisfying

    $ \|f\|_{L^{p(\cdot)}(\Omega)}: = \inf\left\{\lambda > 0 : \int_{\Omega}\left(\dfrac{|f|(x)}{\lambda}\right)^{p(x)}dx\leq 1\right\} < \infty, $

    and the space $W^{1, p(\cdot)}_{0, \sigma}(\Omega)$ is the set of all divergence free $f\in W^{1, 1}_{0}(\Omega)$ with $\|f\|_{L^{p(\cdot)}(\Omega)} + \|\nabla f\|_{L^{p(\cdot)}(\Omega)} <\infty$. For more details, we refer to [6,9]. We say that $u \in W^{1, p(\cdot)}_{loc}(\Omega)^{n}$ is a weak solution of (1), if $u$ satisfies

    $ \int_{\Omega}S(x, D(u)):D(\varphi)dx -\int_{\Omega} u\otimes u:D(\varphi)dx = \int_{\Omega}g\cdot\varphi dx, \, ^{\forall} \varphi\in W^{1, p(\cdot)}_{0, \sigma}(\Omega)^n, $

    where $u\otimes v: = \{u^{i}v^{j}\}_{i, j}$ is a tensor product of $u$ and $v$ in $\mathbb{R}^{n}$.

    In [13], the existence of weak solutions was provided for constant $p > \frac{2n}{n+2}$. In [14,15], the existence of strong solution of (1) was proved when $p$ is constant. In [14], the solution belongs to $C^{1, \alpha}$ for $p>\frac{3}{2}$ when $n = 2$ (up to boundary), and for $p>\frac{6}{5}$ when $n = 3$ (interior regularity). In [15], for $\frac{3n}{n+2} <p <2$, it is shown that $(1+|D(u)|)^{\frac{p-2}{2}}\nabla D(u)\in L^2_{loc}(\Omega)^{n\times n}$, and $u\in W^{2, t}_{loc}(\Omega)^{n}$ for all $1\le t <2$ when $n = 2$, and $u\in W^{2, \frac{3p}{1+p}}_{loc}(\Omega)^n$ when $n = 3$. In [4], for $S(z) = (1+|z|^{p-2})z$ under slip or no-slip boundary conditions a regularity is provided for shear thickening fluid $p>2$.

    In case of anisotropic dissipative potential $f$, where $S = \nabla f$ and

    $ \lambda (1+|z|^2)^{\frac{p_1-2}{2}}|A|^2\le \nabla^2 f(z)(A, A) \le \Lambda (1+|z|^2)^\frac{p_2-2}{2}|A|^2 $

    with exponents $1 <p_1\le p_2 <\infty$ and $2\le p_2 <p_1\frac{n+2}{n}$, the existence of weak solutions is given in [3]. For the anisotropic fluid, it is shown in [5] that $u\in W^{1, p_{1}}_{0}(\Omega)^{n}\cap W^{2, s}_{loc}(\Omega)^{n}$ for some $s>1$ for $p_1>6/5$ when $n = 2$, and $p_1>9/5$ when $n = 3$, and $p_2 <p_1\frac{n+1}{n}$. Also in that article, there is a good introduction about isotropic and anisotropic cases, and the power depending on $x$, $p(x)$.

    For a variable $p(x)$ depending on $E$, in [16] it is proved that a weak solution of (1) exists and it has the second weak derivative for $ \gamma_1 > \dfrac{3n}{n+2}$ and $\mu >0$. Nonetheless, it is not shown there that the second weak derivative of a weak solution belongs to Lebesgue space with variable exponent. It is just shown that a weak solution belongs to $W^{2, \gamma_1}(\Omega)^{n}$.

    The lower bound of $p(\cdot)$, $\gamma_1$, that a weak solution of (1) exists is decreased to $\frac{2n}{n+2}$ in [10]. But if $\gamma_1 > \frac{2n}{n+2}$, we cannot say that $u \otimes u : D(\varphi) \in L^1$ for all $u, \varphi \in W^{1, p(\cdot)}_0(\Omega)^{n}$. In this case, $u$ is not able to be taken as a test function, hence we just consider $ \gamma_1 > \frac{3n}{n+2}$.

    On the other hand, in [7] it was proved that a solution of $p(x)$-Laplace equation has weak second derivative in $L^2$. In that paper, $\mu = 0$ is allowed, but the model is not related with fluid directly. The existence of strong solution of (1) for $S(x, D(u)) = (\mu+|D(u)|)^{p(x)-2}D(u)$ with $\mu>0$ was proved in [8] for $\gamma_1>2$. And the same result was proved in [11] for $\gamma_1>\frac{9}{5}$.

    In this paper, we handle non-Newtonian problem (1) where $\mu$ is allowed to be $0$ and $p(\cdot)$ is Lipschitz continuous. The result of this paper is the following theorem.

    Theorem 1.1. Let $\mu\in [0, \infty)$. We assume that (2) are satisfied and $p(\cdot)$ is Lipschitz continuous function with (3). Let $u$ be a weak solution of (1) with (2). Then $u$ has the following properties:

    $ (1+|D(u)|^2)^{\frac{p(x)-2}{4}}\nabla D(u) \in L^2_{loc}(\Omega)^{n^3} $ (4)

    and for all $1\leq t < 2$,

    $ \left\{ uW2,tloc(Ω)n    when    n=2,uW2,min{3p()1+p(),t}loc(Ω)n    when    n=3.
    \right. $
    (5)

    To simplify the notation, the letter $c$ will always denote any positive constant, which may vary throughout the paper. Moreover, we denote $p' = \frac{p}{p-1}$ as the Hölder conjugate exponent of $p$ and $p^* = \dfrac{n p}{n-p}$ as the Sobolev exponent of $p$ for every $p \in (1, n)$.

    We recall a useful lemma about higher integrability from [1].

    Lemma 2.1. Let $u \in W^{1, p(\cdot)}_{loc}(\Omega)^n$ be a weak solution of (1) and assume that $S$ fulfills conditions $(2)$ with Lipschitz continuous variable exponent, $p(\cdot)$. Then there exist constants $c, \sigma >0$, both depending on $n, \gamma_1, \gamma_2, c_p$ such that if $B_{2R} \subset \subset \Omega$, then

    $ (BR(|D(u)|+μ)(1+σ)p(x)dx)11+σcB2R(|D(u)|+μ)p(x)dx+cB2R(|u|γ1+|u|γ1+1)dx.
    $

    Here, ${{\rlap{-} \smallint }_{B_{R}}} f \, dx$ means the average of $f\in L^{1}(B_{R})$ over $B_R$. We take a constant $R_{0}\leq 1$ such that

    $ 16\omega(4R_{0}) \leq \sigma $ (6)

    and assume $0 <R\leq R_{0}$, throughout this paper. This assumption will be frequently used in the proof, for instance (11) and (12).

    Remark 1. Although the authors only considered the case $\mu>0$ in [1], the statement is still valid for $\mu = 0$, since the proof is also available for $\mu = 0$. In this paper, we can remove the dependence of $c$ on $\gamma_{2}$ since $\gamma_{2}\leq 2$.

    Since $\lim_{t \to 0^+} t^\epsilon \log t = 0$ and $\lim_{t \to +\infty} \tfrac{\log t}{t^\epsilon} = 0$ for any $\epsilon >0$, we have the following useful estimate.

    Lemma 2.2. There exists a constant $c(\epsilon)>0$ depending only on $\epsilon$ such that

    $ \log t \leq c(\epsilon)+ t^\epsilon + t^{-\epsilon} $

    for all $t>0$.

    Our proof is mainly based on that of [15]. We divide our proof of main theorem in three steps. At first, we derive estimation related to finite difference of $D(u)$. After then, applying fractional Sobolev embedding theorem, we can show that $u$ is locally bounded. Finally, we prove that $u$ has second derivatives using difference quotient method.

    Step 1. Estimation of $(|D(u)(x+\lambda {e_k})|^{2}+|D(u)(x)|^{2}+\mu^{2})^{\frac{p(x)-2}{2}}|{\Delta _{\rm{\lambda }}} D(u)|^{2}$.

    From the modulus of continuity, there exists a radius $R_{0}>0$ such that $\omega(8R_{0})\leq \frac{\sigma}{16}$. We use the following weak formulation

    $ \int_{\Omega}S(x, D(u)):D(\varphi)dx = -\int_{\Omega} u_{i}\dfrac{\partial u_{j}}{\partial x_{i}}\varphi_{j} dx + \int_{\Omega} \pi \, {\rm{div}}{\varphi}dx +\int_{\Omega}g\cdot\varphi dx $ (7)

    for all $\varphi \in W^{1, p(x)}_{0}(\Omega)^n$, where $\pi \in L^{p(\cdot)'}(\Omega)$. Let $\eta\in C^{\infty}_{c}(B_{2R}(x_{0}))$, where $B_{2R}(x_{0})\subset \Omega$ such that $\eta\equiv 1$ in $B_{R}(x_{0})$, $|\nabla\eta|\leq \dfrac2R$ and $|\nabla^{2}\eta|\leq \dfrac{4}{R^{2}}$ for some $0 <R\leq R_{0}$. Now we choose a test function $\varphi = \Delta_{-\lambda, k}(\eta^{2}\Delta_{\lambda, k}u)$, where $\Delta_{\lambda, k}f(x) = f(x+\lambda e_{k})-f(x)$. For simplicity, we will denote ${\Delta _{\rm{\lambda }}}: = \Delta_{\lambda, k}$ for $\lambda \in [-R/4, R/4]$ and $B_{2R}: = B_{2R}(x_{0})$. By simple computation, we see

    $ B2RS(x,D(u)):D(Δλ(η2Δλu))dx=B2RΔλS(x,D(u)):D(η2Δλu)dx=B2Rη2ΔλS(x,D(u)):D(Δλu)dx+B2R2ηΔλS(x,D(u)):D(η)Δλudx.
    $
    (8)

    The ellipticity condition in $(2)_2$ implies that

    $ B2Rη2ΔλS(x,D(u)):D(Δλu)dx=B2Rη2[S(x+λek,D(u)(x+λek))S(x,D(u)(x+λek))]ΔλD(u)dx+B2Rη2[S(x,D(u)(x+λek))S(x,D(u)(x))]ΔλD(u)dxB2Rη2[S(x+λek,D(u)(x+λek))S(x,D(u)(x+λek))]ΔλD(u)dx+νO+1η2(|D(u)(x+λek)|2+|D(u)(x)|2+μ2)p(x)22|ΔλD(u)|2dx,
    $
    (9)

    where $O_{1}^{+}: = \{x\in B_{2R} : |Du|(x+\lambda e_{k}) + |Du|(x) + \mu > 0\}$.

    Note that the set $O_{1}^{+} = B_{2R}$ whenever $\mu>0$. Indeed, we here introduced the set $O_{1}^{+}$ to prove Theorem 1.1 for $\mu = 0$, in a simple and unified way with respect to $\mu\in [0, \infty)$.

    Combining (7)-(9), we obtain

    $ I0:=νO+1η2(|D(u)(x+λek)|2+|D(u)(x)|2+μ2)p(x)22|ΔλD(u)|2dxB2Rη2[S(x,D(u)(x+λek))S(x+λek,D(u)(x+λek))]D(Δλu)dxB2R2ηΔλS(x,D(u)):D(η)ΔλudxB2Ruiujxi(Δλ(η2Δλuj))dx+B2Rπdiv(Δλ(η2Δλu))dx+B2RgΔλ(η2Δλu)dx=:I1+I2+I3+I4+I5.
    $
    (10)

    For simplicity, we set $\nu = 1$ since it does not make any trouble for the proof.

    Estimation of $I_{1}$. To estimate $I_{1}$, we introduce a measurable set $O_{2}^{+}$:

    $ O_{2}^{+} = \{x\in B_{2R} : |Du|(x+\lambda e_{k}) > 0\}\subset O_{1}^{+}. $

    We use the continuity assumption in $(2)_3$ to see

    $ |I1|cω(λ)O+2η2|D(Δλu)|[1+|log(|D(u)(x+λek)|2+μ2)|]×[(|D(u)(x+λek)|2+μ2)p(x+λek)12+(|D(u)(x+λek)|2+μ2)p(x)12]dxc|λ|O+2η2|D(Δλu)|(|D(u)(x+λek)|2+|D(u)(x)|2+μ2)p(x)24×[1+|log(|D(u)(x+λek)|2+μ2)|]
    $
    $ ×[(|D(u)(x+λek)|2+|D(u)(x)|2+μ2)2p(x+λek)p(x)4+(|D(u)(x+λek)|2+|D(u)(x)|2+μ2)p(x)4]dxϵO+1η2(|D(u)(x+λek)|2+|D(u)(x)|2+μ2)p(x)22|D(Δλu)|2dx+c(ϵ)|λ|2O+2η2[1+|log(|D(u)(x+λek)|2+μ2)|]2×[(|D(u)(x+λek)|2+|D(u)(x)|2+μ2)2p(x+λek)p(x)2+(|D(u)(x+λek)|2+|D(u)(x)|2+μ2)p(x)2]dx.
    $

    Therefore, Lemma 2.2 and (6) reveals

    $ |I1|ϵI0+c(ϵ)|λ|2B2Rη2[(|D(u)(x+λek)|2+|D(u)(x)|2+μ2)2p(x+λek)p(x)2+(|D(u)(x+λek)|2+|D(u)(x)|2+μ2)p(x)2]×[1+(|D(u)(x+λek)|2+μ2)σγ116+(|D(u)(x+λek)|2+μ2)σγ116]2dxϵI0+c(ϵ)|λ|2B2Rη2(1+|D(u)(x)|)(1+σ)p(x)dx.
    $
    (11)

    Estimation of $I_{2}$. Noting that (6) implies

    $ p_{2}\leq p_{1}+\omega(4R_{0}) \leq p_{1}(1+\sigma) $

    and

    $ \frac{p_{1}(p_{2}-1)}{p_{1}-1}\leq p_{1}+\frac{\omega(4R_{0})p_{1}}{p_{1}-1}\leq p_{1}+\frac{\omega(4R_{0})\gamma_{1}}{\gamma_{1}-1} \leq p_{1}+8\omega(4R_{0}), $ (12)

    we discover

    $ |I2|2(B2R|S(x,D(u))|p1dx)1p1(B2R|Δλ(ηD(η)Δλu)|p1dx)1p1c|λ|(B2R(1+|u|)p1(p(x)1)dx)1p1×(B2R|(ηD(η)Δλu)|p1dx)1p1c|λ|(B2R(1+|u|)(1+σ)p(x)dx)1p1×(|λ|p1R2p1B2R(1+|u|p1)dx+1Rp1B2R|η(Δλu)|p1dx)1p1.
    $
    (13)

    Note that

    $ \eta\nabla(\Delta_{\lambda}u) = \nabla(\eta\Delta_{\lambda}u)-(\nabla\eta)\cdot\Delta_{\lambda}u. $ (14)

    Applying (14) and Korn's inequality, we have

    $ B2R|η(Δλu)|p1dxcB2R|D(ηΔλu)|p1dx+cRp1B2R|Δλu|p1dxcO+1|ηD(Δλu)|p1dx+cRp1|λ|p1B3R|u|p1dx.
    $
    (15)

    And the first integral term in the last line is estimated as follows:

    $ (O+1|η|p1|D(Δλu)|p1dx)1p1c(O+1|η|2(|D(u)(x+λek)|2+|D(u)(x)|2+μ2)(p(x)2)2|D(Δλu)|2dx)12×(B2R(|D(u)(x+λek)|2+|D(u)(x)|2+μ2)p1(2p(x))2(2p1)dx)2p12p1c[I0]12(B3R(1+|u|)(1+σ)p(x)dx)2p12p1.
    $
    (16)

    Combining (13), (15) and (16), we obtain

    $ |I2|2c|λ|2R2B3R(1+|u|)(1+σ)p(x)dx+|λ|R(B3R(1+|u|)(1+σ)p(x)dx)1p1(B2R|η|p1|D(Δλu)|p1dx)1p1c(ϵ)|λ|2R2B3R(1+|u|)(1+σ)p(x)dx+ϵI0.
    $

    Estimation of $I_3$ will be postponed to Step 2.

    Estimation of $I_{4}$. Recalling Lemma 2.1 and the De Rham theory in [10], we see $\pi\in L^{q(\cdot)}(\Omega)$ for $q(x) = (1+\sigma)p'(x)$. Observe that

    $ \min\limits_{ B_{2R}}q(x) = \dfrac{p_{2}(1+\sigma)}{p_{2}-1} > \dfrac{p_{1}}{p_{1}-1} $

    by the assumption $\omega(4R_{0}) <\dfrac{\sigma}{16}$. Via a similar estimating procedure of $I_{2}$, we compute

    $ |I4|=|B2Rπ(Δλ,k(ηηxiΔλui))dx|c|λ|RπLp1(B2R|(ηΔλu)|p1dx)1p1.
    $
    (17)

    In light of Korn's inequality, it holds that

    $ (B2R|(ηΔλu)|p1dx)1p1c|λ|2R2(B2R(1+|u|)(1+σ)p(x)dx)1p1+c(O+1|η|p1|D(Δλu)|p1dx)1p1.
    $
    (18)

    Combining (16), (17) and (18), we obtain

    $ |I4|c|λ|2R2πLp1(B2R)(B2R(1+|u|)(1+σ)p(x)dx)1p1+c(ϵ)|λ|2R2π2Lp1(B2R)(B3R(1+|u|)(1+σ)p(x)dx)2p1p1+ϵI0.
    $

    Estimation of $I_{5}$. Korn's inequality yields

    $ |I5|cgL(B2R)B2R|Δλ(η2(Δλu))|dxc|λ|O+1|D(η2Δλu)|dxc|λ|2RB3R|D(u)|dx+c|λ|O+1|η|2|ΔλD(u)|dxc(ϵ)|λ|2R2B2R(1+|D(u)|)(1+σ)p(x)dx+ϵI0.
    $

    Consequently, it follows from combining the estimates of $I_{1}, \, I_2, \, I_4 \textrm{ and }\, I_{5}$, and (10) that

    $ I0I3+c|λ|2R2B2R(1+|u|)(1+σ)p(x)dx+c|λ|2R2(B3R(1+|u|)(1+σ)p(x)dx)2p1p1.
    $
    (19)

    Step 2. Boundedness of $u$.

    In this step, we shall prove that $u$ is locally bounded. Note that Lemma 2.1 implies that $u$ is locally Hölder continuous by Morrey embedding in case of $n = 2$ and $\gamma_{1} = 2$. So it is not necessary to consider this case. We recall $I_3$ and apply integration by parts formula for finite difference to see

    $ I3=B2RΔλuiujxi(x+λek)η2Δλujdx+B2RuiΔλujxiη2Δλujdx=:I3,1+I3,2.
    $

    We now assume that

    $ u\in W^{1, s}_{loc}(\Omega)^n \textrm{ for some }s\in[\gamma_{1}, n). $

    Next we estimate $I_{3, 1}$

    $ |I_{3, 1}| \leq \int_{B_{2R}}|{\Delta _{\rm{\lambda }}} u|^{2}|\nabla u(x+\lambda {e_k})|dx \leq \|{\Delta _{\rm{\lambda }}} u\|_{L^{2s'}(B_{2R})^n}^{2}\| u \|_{W^{1, s}(B_{2R})^n}. $

    Defining $\theta: = \frac{s(n+2)-3n}{2n}$, we see $0 <\theta <1$ and

    $ \dfrac{1-\theta}{s^{*}}+\dfrac{\theta}{s} = \dfrac{1}{2s'}, $

    where $s'$ is the Hölder conjugate exponent of $s$ and $s^{*}$ is the Sobolev exponent of $s$. It then follows from interpolation that

    $ \|{\Delta _{\rm{\lambda }}} u\|_{L^{2s'}(B_{2R})^n}^{2}\leq \|{\Delta _{\rm{\lambda }}} u\|_{L^{s^{*}}(B_{2R})^n}^{2(1-\theta)}\|{\Delta _{\rm{\lambda }}} u\|_{L^{s}(B_{2R})^n}^{2\theta}\leq c|\lambda|^{2\theta}\|u\|_{W^{1, s}(B_{3R})^n}^{2} $

    and

    $ |I_{3, 1}|\leq c|\lambda|^{2\theta}\|u\|_{W^{1, s}(B_{3R})^n}^{3}. $ (20)

    On the other hand, integration by parts formula for finite difference reveals

    $ |I3,2|12|B2R|ui|(Δλuj)2η2xidx|cRΔλu2L2s(B2R)nuW1,s(B2R)ncR|λ|2θu3W1,s(B3R)n.
    $
    (21)

    Merging up (20) and (21), we obtain

    $ |I_{3}|\leq c(1+\dfrac{1}{R})|\lambda|^{2\theta}\|u\|_{W^{1, s}(B_{3R})^n}^{2}. $ (22)

    Finally, since $p(\cdot)$ is Lipschitz continuous, (19) and (22) give the boundedness of $I_0$:

    $ \label{31} I_{0} \leq c \dfrac{|\lambda|^{2\theta}}{R^{2}}, $ (23)

    where the constant $c$ depends on $\|u\|_{W^{1, s}(B_{3R})^n}$ and $\|\pi\|_{L^{p_{1}'}(B_{2R})}$. Set

    $ \hat{s}: = \dfrac{2s}{s-\gamma_{1}+2}. $

    Then $\gamma_{1} \leq \hat{s} <2$ and $\frac{s(\hat{s}-2)}{\hat{s}} = \gamma_{1}-2$. Thus, H$\ddot{\mathit{\boldsymbol{o}}}$lder's inequality and (23) reveal

    $ BR|ΔλD(u)|ˆsdx=O+1(|D(u)(s+λek)|2+|D(u)(x)|2+μ2)p(x)22ˆs2|ΔλD(u)(x)|ˆsηˆs×(|D(u)(s+λek)|2+|D(u)(x)|2+μ2)2p(x)2ˆs2dx[I0]ˆs2[B2R(|D(u)(s+λek)|2+|D(u)(x)|2+μ2)2p(x)2ˆsˆs2dx]2ˆs2c|λ|θˆsR2[B2R(1+|D(u)(x)|)sdx]2ˆs2.
    $
    (24)

    This implies

    $ D(u)\in W^{t, \hat{s}}(B_{R/2})^{n \times n} ~~~~ \textrm{for any } ~~t\in[0, \theta) $

    and then we obtain by Nikolskii's embedding theorem in [2] that

    $ D(u)\in L^{\frac{n\hat{s}}{n-t\hat{s}}}(B_{R/2})^{n \times n} ~~~~ \textrm{for any }~ t\in[0, \theta). $ (25)

    We set

    $ \sigma(s): = \dfrac{n\hat{s}}{n-\theta\hat{s}} = \dfrac{2ns}{(5-\gamma_{1})n-2s}, $

    then it follows that

    $ \sigma(s)-s\geq \tau_{0}: = \dfrac{\gamma_{1}((n+2)\gamma_{1}-3n)}{(5-\gamma_{1})n-2\gamma_{1}} > 0. $

    Hence (25) can be written as

    $ \nabla u\in L^{\tau}(B_{R/2})^{n \times n} ~~~~ \textrm{for any } \tau\in[1, \sigma(s)). $

    Here, we have used Korn's inequality. According to Sobolev embedding theorem and the fact that $\sigma(s)\leq s^{*}$, we see

    $ u\in W^{1, \tau}(B_{R/2}) ^n ~~~~\textrm{for any }\tau\in[1, \sigma(s)) $

    and so, by covering, we also have

    $ u\in W^{1, \tau}_{loc}(\Omega)^n ~~~~\textrm{for any }\tau\in[1, \sigma(s)). $

    By bootstrap argument, we can conclude that

    $ u\in W^{1, \tau}_{loc}(\Omega)^n ~~~~ \textrm{for any }\tau\in[1, \sigma(n)), $

    where $n < \sigma(n) = \frac{2n}{3-\gamma_{1}}$. By Morrey embedding theorem, $u$ is locally Hölder continuous, and so $u$ is locally bounded.

    Step 3. Completing the proof of Theorem 1.1.

    For $1 <s <\frac{2n}{3-\gamma_{1}}$, we estimate $I_{3}$ again

    $ |I3||B2Ruiujxi(Δλ,k(η2Δλuj))dx||λ|uL(B2R)nuLs(B2R)n×nD(η2Δλu)Ls(B2R)n×n|λ|uL(B2R)nuLs(B2R)n×n×(η2D(Δλu)Ls(B2R)n×n+|λ|RuLs(B2R)n×n).
    $
    (26)

    We select $s = 4-\gamma_{1}$. Then $2\leq s < \frac{2n}{3-\gamma_{1}}$ and

    $ \dfrac{(2-p(x))s'}{2-s'} = \dfrac{2-p(x)}{2-\gamma_{1}}(4-\gamma_{1})\leq (4-\gamma_{1}) = s. $

    By following the calculations in (24), we also have

    $ \label{44-1} η2D(Δλu)Ls(B2R)n×n[I0]12(B2R(|D(u)(x+λek)|2+|D(u)(x)|2+μ2)(2p(x))s2(2s)dx)2s2s[I0]12(B2R(1+|u|)sdx)2s2s.
    $
    (27)

    Applying (27) and Young's inequality to (26), we find

    $ \label{44} |I_{3}|\leq c(\epsilon)\dfrac{|\lambda|^{2}}{R}+\epsilon I_{0}. $ (28)

    Combining (19), (28) and the assumption $p(x) \leq 2$, we obtain

    $ \label{s1} BR(1+|D(u)(x+λek)|2+|D(u)(x)|2)p(x)22|ΔλD(u)λ|2dx=O+(1+|D(u)(x+λek)|2+|D(u)(x)|2)p(x)22|ΔλD(u)λ|2dxO+1(|D(u)(x+λek)|2+|D(u)(x)|2+μ2)p(x)22|ΔλD(u)λ|2dxc,
    $
    (29)

    where

    $ O^{+}: = B_{R}\cap \{x\in \Omega : |Du|(x+\lambda e_{k}) + |Du|(x) + \mu > 0\}\subset O_{2}^{+}. $

    We now divide (27) by $|\lambda|$ to deduce

    $ \frac{1}{|\lambda|}\|{\Delta _{\rm{\lambda }}} \nabla u\|_{L^{s'}(B_{R})^{n \times n}} \leq \frac{1}{|\lambda|}\|\eta^{2}D({\Delta _{\rm{\lambda }}} u)\|_{L^{s'}(B_{2R})^{n \times n}}\leq c $

    for all $\lambda \in \left(-\frac{R}{4}, \frac{R}{4}\right)$, where we have used Korn's inequality and (19).

    By difference quotient method (See [12,Section 5.8.2]), there exists the weak derivative of $\nabla u$ such that

    $ \label{s2} \frac1\lambda {\Delta _{\rm{\lambda }}} \nabla u \rightharpoonup \frac{\partial}{\partial x_k}\nabla u \textrm{ in } L^{s'}(B_R)^{n \times n}. $ (30)

    For the time being, we suppose that

    $ \label{s3} (1+|D(u)(x+λek)|2+|D(u)(x)|2)p(x)24(1+2|D(u)(x)|2)p(x)24
    \textrm{ in } L^q(B_R) $
    (31)

    for all $q \in [1, \infty)$. Then we see that (30) deduce

    $ (1+|D(u)(x+λek)|2+|D(u)(x)|2)p(x)24(ΔλD(u)λ)(1+2|D(u)(x)|2)p(x)24(xkD(u))
    \textrm{ in } L^1(B_R)^{n\times n}. $

    Since there exists a function $w_k \in L^2(B_R)$ by (29) such that

    $ (1+|D(u)(x+\lambda {e_k})|^{2}+|D(u)(x)|^{2})^{\frac{p(x)-2}{4}}\left( \dfrac{{\Delta _{\rm{\lambda }}} D(u)}{\lambda} \right) \rightharpoonup w_k \textrm{ in }L^2(B_R)^{n\times n}, $

    we find $\left(1+2|D(u)(x)|^2 \right)^{\frac{p(x)-2}{4}} \left|\frac{\partial}{\partial x_k}\nabla u \right| = w_k$.

    Verification of (31). To do this, we write $h(x) = \frac 12 +|D(u)(x)|^2$ and $h_\lambda(x) = \frac 12 +|D(u)(x+\lambda {e_k})|^2$, and compute

    $ BR|1(h(x)+hλ(x))2p(x)41(2h(x))2p(x)4|qdx=BR|(2h(x))2p(x)4(h(x)+hλ(x))2p(x)4(h(x)+hλ(x))2p(x)4(2h(x))2p(x)4|qdxcBR((2h(x))2p(x)4+(h(x)+hλ(x))2p(x)4)q1((h(x)+hλ(x))2p(x)4(2h(x))2p(x)4)q×(2h(x))2p(x)4(h(x)+hλ(x))2p(x)4)dxcBR(2h(x))2p(x)4(h(x)+hλ(x))2p(x)4dx.
    $

    And we estimate

    $ (2h(x))2p(x)4(h(x)+hλ(x))2p(x)4=10ddt(h(x)+hλ(x)+t(h(x)hλ(x)))2p(x)4dt2p(x)4|h(x)hλ(x)|10(h(x)+hλ(x)+t(h(x)hλ(x))2p(x)4dt2p(x)2|h(x)hλ(x)|=2p(x)2||D(u)(x)|2|D(u)(x+λek)|2|.
    $

    Hence we have

    $BR|(1+|D(u)(x+λek)|2+|D(u)(x)|2)p(x)24(1+2|D(u)(x)|2)p(x)24|qdxcBR||D(u)(x)|2|D(u)(x+λek)|2|dx.
    $

    Since the integral on the right-hand side above inequality converges to 0 as $\lambda \to 0$, (31) is valid, and (4) is proved.

    To verify (5), we put $V = \left(1+|D(u)|^2 \right)^{\frac{p(x)}{4}}$. Then

    $ \left| \frac{\partial V}{\partial x_k} \right| \leq c \left( (1+|D(u)|^2 )^{\frac{p(x)}{4}}\log(1+ |D(u)|^2)+ (1+|D(u)|^2 )^{\frac{p(x)-2}{4}}\left|\frac{\partial}{\partial x_k}D(u)\right| \right). $

    Using (4), the higher integrability of $D(u)$ and Lemma 2.2, we know $V \in W^{1, 2}_{loc}(\Omega)$. Therefore,

    $ \label{s6} V \in L^s_{loc}(\Omega) \textrm{ for all } s \in [1, \infty) \textrm{ if } n = 2, ~~~~V \in L^6_{loc}(\Omega) \textrm{ if } n = 3. $ (32)

    Note that following inequality holds by Young's inequality:

    $ |a|^t = |b|^{\frac{t(p-2)}{2}}|a|^t |b|^{\frac{t(2-p)}{2}} < |b|^{p-2}|a|^2 + |b|^{\frac{t(2-p)}{2-t}} $

    for any $p \in [1, 2], t \in [1, 2) \textrm{ and } |b| >0.$ For the case $n = 2$ with $t\in [1, 2)$, we estimate

    $ BR|D(u)xk|tdxBR(1+|D(u)|2)p(x)22|D(u)xk|2dx+BR(1+|D(u)|2)t(2p(x))2(2t)dx.
    $

    Then, we immediately deduce $(5)_{1}$ from (32) and Korn's inequality. One can prove $(5)_{2}$ in a similar way. This completes our proof.

    Bae was supported by the Basic Science Research Program through the National Research Foundation of Korea (NRF) funded by the Ministry of Education, Science and Technology (2015R1D1A1A01057976 and 2016K2A9A2A06005080) and Youn was supported by NRF (2015R1A4A1041675). The authors would like to thank referees for valuable comments and suggestions. The authors would like to express their sincere gratitudes Jihoon Ok for helpful discussions.

    [1] Regularity results for stationary electro-rheological fluids. Arch. Ration. Mech. Anal. (2002) 164: 213-259.
    [2] R. A. Adams, Sobolev Spaces, Academic Press, New York-London, 1975.
    [3] Steady states of anisotropic generalized Newtonian fluids. J. Math. Fluid Mech. (2005) 7: 261-297.
    [4] Navier-Stokes Equations with shear-thickening viscosity. Regularity up to the boundary. J. Math. Fluid Mech. (2009) 11: 233-257.
    [5] On strong solutions of the differential equations modeling the steady flow of certain incompressible generalized Newtonian fluids. Algebra i Analiz (2006) 18: 1-23.
    [6] On $W^{1, q(·)}$-estimates for elliptic equations of $p(x)$-Laplacian type. J. Math. Pures Appl. (2016) 106: 512-545.
    [7] Second order regularity for the $p(x)$-Laplace operator. Math. Nachr. (2011) 284: 1270-1279.
    [8] On the $C^{1, γ}(\bar{Ω}) \cap W^{2, 2} (Ω)$ regularity for a class of electro-rheological fluids. J. Math. Anal. Appl. (2009) 356: 119-132.
    [9] L. Diening, P. Harjulehto, P. Hästö and M. Růžička, Lebesgue and Sobolev Spaces with Variable Exponents, Springer, Heidelberg, 2011. doi: 10.1007/978-3-642-18363-8
    [10] An existence result for non-Newtonian fluids in non-regular domains. Discrete Contin. Dyn. Syst. Ser. S (2010) 3: 255-268.
    [11] Existence of local strong solutions for motions of electrorheological fluids in three dimensions. Comput. Math. Appl. (2007) 53: 595-604.
    [12] L. C. Evans, Partial Differential Equations, 2$^{nd}$ edition, Graduate Studies in Mathematics, 19. American Mathematical Society, Providence, RI, 2010. doi: 10.1090/gsm/019
    [13] On analysis of steady flows of fluids with shear-dependent viscosity based on the Lipschitz truncation method. SIAM J. Math. Anal. (2003) 34: 1064-1083.
    [14] $C^{1, α}$-solutions to a class of nonlinear fluids in two dimensions-stationary Dirichlet problem. Zap. Nauchn. Sem. S.-Peterburg. Otdel. Mat. Inst. Steklov. (POMI) (1999) 259: 89-121.
    [15] Interior differentiability of weak solutions to the equations of stationary motion of a class of non-Newtonian fluids. J. Math. Fluid Mech. (2005) 7: 298-313.
    [16] M. Růžička, Electrorheological Fluids: Modeling and Mathematical Theory, Springer-Verlag, Berlin, 2000.
  • This article has been cited by:

    1. Cholmin Sin, P. Areias, The Existence of Strong Solution for Generalized Navier-Stokes Equations withpx-Power Law under Dirichlet Boundary Conditions, 2021, 2021, 1687-9139, 1, 10.1155/2021/6755411
  • Reader Comments
  • © 2018 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

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

Metrics

Article views(6260) PDF downloads(285) Cited by(1)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog