• 검색 결과가 없습니다.

GLOBAL AXISYMMETRIC SOLUTIONS TO THE 3D NAVIER-STOKES-POISSON-NERNST-PLANCK SYSTEM IN THE EXTERIOR OF A CYLINDER

N/A
N/A
Protected

Academic year: 2021

Share "GLOBAL AXISYMMETRIC SOLUTIONS TO THE 3D NAVIER-STOKES-POISSON-NERNST-PLANCK SYSTEM IN THE EXTERIOR OF A CYLINDER"

Copied!
16
0
0

로드 중.... (전체 텍스트 보기)

전체 글

(1)

https://doi.org/10.4134/BKMS.b200536 pISSN: 1015-8634 / eISSN: 2234-3016

GLOBAL AXISYMMETRIC SOLUTIONS TO

THE 3D NAVIER–STOKES–POISSON–NERNST–PLANCK SYSTEM IN THE EXTERIOR OF A CYLINDER

Jihong Zhao

Abstract. In this paper we prove global existence and uniqueness of ax-isymmetric strong solutions for the three dimensional electro-hydrodyna-mic model based on the coupled Navier–Stokes–Poisson–Nernst–Planck system in the exterior of a cylinder. The key ingredient is that we use the axisymmetry of functions to derive the Lpinterpolation inequalities,

which allows us to establish all kinds of a priori estimates for the velocity field and charged particles via several cancellation laws.

1. Introduction

In this paper, we study the following dissipative system of nonlinear and nonlocal equations modeling the flow of electro-hydrodynamics in a connected open set Ω ⊂ R3 with smooth boundary ∂Ω:

(1)                ut+ (u · ∇)u − ∆u + ∇P = ∆Ψ∇Ψ, div u = 0, n−t + (u · ∇)n−= ∇ · (∇n−− n−∇Ψ), n+t + (u · ∇)n+= ∇ · (∇n++ n+∇Ψ), ∆Ψ = n−− n+,

where u = (u1, u2, u3) and P denote the unknown vector velocity and scalar

pressure of fluid, respectively, n−and n+denote the densities of binary diffuse

negatively and positively charged particles, respectively, and Ψ is the elec-trostatic potential. All physical parameters in (1) have been taken to be 1 for simplicity of presentation. Furthermore, in the domain Ω where the fluid occupies, the system (1) is assumed to supplement with the following initial

Received June 18, 2020; Accepted February 9, 2021.

2010 Mathematics Subject Classification. Primary 35B07, 35D35, 35K55, 76W05. Key words and phrases. Navier–Stokes–Poisson–Nernst–Planck equations, axisymmetric solutions, global existence.

This work was financially supported by the National Natural Science Foundation of China (no. 11961030) and the Natural Science Foundation of Shaanxi Province (no. 2018JM1004).

c

2021 Korean Mathematical Society 729

(2)

conditions: (2) (u, n−, n+)|t=0= (u0, n−0, n + 0), div u0= 0, n−0 > 0, n + 0 > 0 in Ω,

and the velocity field equations is determined by the no slip boundary condition: u = 0 on ∂Ω × (0, T ),

(3)

while the equations for the charged densities are determined by the pure Neu-mann boundary conditions:

∂n− ∂ν = 0, ∂n+ ∂ν = 0, ∂Ψ ∂ν = 0 on ∂Ω × (0, T ), (4)

where ν denotes the unit outward normal vector of ∂Ω.

The system (1)–(4) appears in the context as the Navier–Stokes–Poisson– Nernst–Planck system in electro-hydrodynamics, which was intended to ac-count for the electro-diffusion phenomenon in an incompressible electrical fluid medium, for example, see [14,16]. Generally speaking, the self-consistent charge transport is described by the Poisson–Nernst–Planck equations, while the fluid motion is governed by the incompressible Navier–Stokes equations with forcing terms. We refer the readers to see [16] for the detailed mathematical description and physical background of this fluid-dynamical model.

If the flow is charge-free, i.e., n−= n+= Ψ = 0, then the system (1) reduces

to the following incompressible Navier–Stokes equations: (5)

(

ut+ (u · ∇)u − ∆u + ∇P = 0,

div u = 0.

In their celebrated works, Leray [12] and Hopf [5] proved that the n-dimensional (n ≥ 2) Navier–Stokes equations (5), subject to the initial data of L2-finite

en-ergy, admits at least a global Leray–Hopf weak solution. It is well-known that such a global weak solution is regular and unique in two dimensional case, but in three dimensional case, the regularity and uniqueness of such weak solutions still remains a challenging open problem in mathematical fluid dynamics. On the other hand, many efforts have been made to study various solutions with certain special structures, and the axisymmetric solution is such an important case. For the axisymmetric Navier–Stokes equations without swirl, Ladyzhen-skaya [9] and Ukhovskii–Yudovich [17] independently proved global existence, uniqueness and regularity of axisymmetric weak solutions. Later on, Leonardi et al. [11] gave a refined proof, and Abidi [2] extended this global regularity result to certain initial data in critical space ˙H12. For the axisymmetric Navier– Stokes equations with non-trivial swirl, Ladyzhenskaya [10] and Abe–Seregin [1] proved the global existence of unique axisymmetric strong solution in the ex-terior of a cylinder subject to the no slip and Navier boundary conditions. The crucial points for the analysis in [10] and [1] are the interpolation inequalities and the maximum principle.

(3)

To our knowledge, mathematical analysis of the system (1) was initiated by Jerome [6], where the author established a local well-posedness theory and stability under the inviscid limit based on the Kato’s semigroup framework. Subsequently, local existence with any initial data and global existence with small initial data in various critical functional spaces (e.g., Lebesgue and Besov spaces) were established by [13, 18–20]. On the other hand, the global exis-tence, as well as regularity and uniqueness, of weak solutions for the system (1) adapted with various boundary conditions have been studied in two or three dimensions by [3, 4, 7, 8, 15, 16], some analytical results similar to the Navier– Stokes equations were obtained.

The main goal of this paper is to study global existence of axisymmetric strong solutions for the electro-hydrodynamic system (1) in the exterior of a cylinder subject to the initial boundary-value conditions (2)–(4). The key ingredient is that we shall make use of axisymmetry of functions to derive some Lp interpolation inequalities, which allow us to establish some crucial a

priori estimates for the velocity field and charged particles via several important cancellation laws. Without loss of generality, we assume that

Ω = {x = (x1, x2, x3) ∈ R3,

q x2

1+ x22= δ > 0, x3∈ R}.

Denote a point in Ω by x. Let us consider the cylindrical coordinates: r = q x2 1+ x22, θ = arctan x2 x1 , z = x3,

and denote the three standard basis vectors are as follows: er= ( x1 r , x2 r , 0), eθ= (− x2 r , x1 r , 0), ez= (0, 0, 1).

A function f or a vector function u = (ur, uθ, uz) is said to be axisymmetric if f , ur, uθ and uz are independent of the angular variable θ, i.e., f and u has the following forms:

f (x1, x2, x3) = f (r, z), u(x1, x2, x3) = ur(r, z)er+ uθ(r, z)eθ+ uz(r, z)ez.

Due to the uniqueness of strong solutions, it is clear that if the initial data (u0, n−0, n

+

0) is axisymmetric, then the strong solution (u, n−, n+) to the

prob-lem (1)–(4) is also axisymmetric.

Before we state the main result, let us introduce the following notations. We denote by Lp(Ω), 1 < p < ∞ (or L∞(Ω)) the space of the usual scalar-valued or vector-scalar-valued functions defined on Ω with the p-th power absolutely integrable (or essentially bounded scalar-valued or vector-valued functions) for the Lebesgue measure. For m ∈ N, 1 ≤ p ≤ ∞, the Sobolev space Wm,p(Ω) is the space of functions in Lp(Ω) with derivatives of order less than or equal to

m in Lp(Ω), i.e.,

Wm,p(Ω) = {f ∈ Lp(Ω) : Dαf ∈ Lp(Ω), |α| ≤ m}. In particular, when p = 2, we denote Hm(Ω) = Wm,2(Ω). Let C

0 (Ω) be the

(4)

C0∞(Ω) in Wm,p(Ω) is denoted by Wm,p

0 (Ω) (H0m(Ω) when p = 2). Moreover,

let C0,σ∞(Ω) be the space

C0,σ∞(Ω) = {u ∈ C0∞(Ω) : div u = 0}.

The closure of C0,σ∞(Ω) in H1(Ω) is denoted by H1 0,σ(Ω).

Our main result is stated as follows.

Theorem 1.1. Let Ω = {x = (x1, x2, x3) ∈ R3,px21+ x22 = δ > 0, x3 ∈

R}. Assume that (u0, n−0, n +

0) is an axisymmetric initial data and satisfy the

following regularity conditions: (6) u0∈ H2(Ω)∩H0,σ1 (Ω), n − 0, n + 0 ∈ H 2(Ω)∩L1(Ω), n− 0 > 0, n + 0 > 0 in Ω.

Then for every T > 0, there exists a unique axisymmetric strong solution (u, n−, n+) to the initial-boundary value problem (1)–(4) with

u ∈ C([0, T ], H0,σ1 (Ω)) ∩ L∞(0, T ; H2(Ω)), n−, n+∈ C([0, T ], H2(Ω) ∩ L1(Ω)) and ut, n−t, n + t ∈ L ∞(0, T ; L2(Ω)), ∇u t, ∇n−t, ∇n + t ∈ L 2(0, T ; L2(Ω)).

We shall prove Theorem 1.1 in the next section. Throughout the paper, we shall use the notation k · kLp instead of k · kLp(Ω), and k · kHm instead of k · kHm(Ω)for simplicity. We denote by C the harmless positive constant, which may depend on initial datum and its value may change from line to line, the special dependence will be pointed out explicitly in the text if necessary.

2. The proof of Theorem 1.1

We first prove the following crucial interpolation inequalities for axisymmet-ric functions in Lebesgue spaces.

Lemma 2.1. Let D = {(r, z) : r ≥ δ > 0, z ∈ R}. Then for any axisymmetric vector function u = ur(r, z)e

r+ uθ(r, z)eθ+ uz(r, z)ez satisfying ui ∈ H1(D)

for i = r, θ, z, there exists a constant C depending only on p and δ such that for any 2 ≤ p < ∞, we have

(7) kukLp(Ω)≤ C(p, δ)(kuk 2 p L2(Ω)k∇uk 1−2 p L2(Ω)+ kukL2(Ω)), where Ω = {x ∈ R3:px2

1+ x22= δ > 0, x3∈ R} is the corresponding domain

of D in Cartesian coordinates. We emphasize here that (7) also holds for any axisymmetric scalar function.

Proof. Notice that in the Cartesian coordinates, u1= x1 r u rx2 r u θ, u 2= x2 r u r+x1 r u θ, u 3= uz. Then we have kukpLp= Z Ω |u|pdx =Z Ω  (x1 r u rx2 r u θ)2+ (x2 r u r+x1 r u θ)2+ (uz)2 p 2 dx

(5)

= Z Ω (|ur|2+ |uθ|2+ |uz|2)p2dx ≤ C(p) Z Ω (|ur|p+ |uθ|p+ |uz|p)dx.

By using the Gagliardo–Nirenberg’s inequality in two dimensions, we know that for i = r, θ, z, kui(r, z)kp Lp(D) ≤ C(p)ku i(r, z)k2 L2(D)k e∇ui(r, z)k p−2 L2(D),

where e∇ := (∂r, ∂z) is the two dimensional gradient operator. It follows that

kuikp Lp(Ω)= Z Ω |ui|pdx = 2π Z D |ui|prdrdz = 2π Z D (|ui|rp1)pdrdz ≤ C(p) Z D (|ui|r1p)2drdz  Z D (|∂rui|r 1 p)2+ 1 p2(|u i|r1 p−1)2+(|∂ zui|r 1 p)2drdz p2−1 ≤ C(p) Z D (|ui|r1p)2drdz  Z D (|∂rui|r 1 p)2+ (|∂ zui|r 1 p)2drdz p2−1 + C(p) Z D (|ui|r1p)2drdz  Z D (|ui|r1p−1)2drdz p2−1 . Since 2 ≤ p < ∞, r ≥ δ > 0, it is easily seen that

rp2 = rrp2−1≤ rδ2p−1, r2p−2≤ rδ2p−3. This yields immediately that

kuikp Lp(Ω)≤ C(p, δ) hZ D |ui|2rdrdz Z D (|∂rui|2+ |∂zui|2)rdrdz p2−1 + Z D |ui|2rdrdz p2i ≤ C(p, δ) kuik2L2(Ω)k∇u i kp−2L2(Ω)+ ku i kpL2(Ω).

The proof of Lemma 2.1 is achieved. 

Next we establish the following several crucial a priori estimates. Let us introduce two new functions v := n−+ n+ and w := n− n+, and then the

problem (1) is reduced into the following system of equations:

(8)                ut+ (u · ∇)u − ∆u + ∇P = w∇Ψ, div u = 0, vt+ (u · ∇)v = ∇ · (∇v − w∇Ψ), wt+ (u · ∇)w = ∇ · (∇w − v∇Ψ), ∆Ψ = w,

(6)

while initial conditions (2) and boundary conditions (3)–(4) are correspondingly changed into the following way:

(9) (u, v, w)|t=0= (u0, v0, w0) = (u0, n−0 + n + 0, n − 0 − n + 0), div u0= 0 in Ω and u = 0, ∂v ∂ν = 0, ∂w ∂ν = 0, ∂Ψ ∂ν = 0 on ∂Ω × (0, T ). (10)

It is clear that if (u, n−, n+) is an axisymmetric strong solution of the problem (1)–(4) (P and Ψ can be determined by (u, n−, n+)), then (u, v, w) is an ax-isymmetric strong solution of the problem (8)–(10) (P and Ψ can be determined by (u, v, w)), and vice verse. Therefore, we aim at establishing some crucial a priori estimates of axisymmetric strong solutions for the problem (8)–(10). Lemma 2.2. Let the assumptions (6) be in force, and let (u, v, w) be the cor-responding axisymmetric strong solution to the problem (8)–(10) on [0, T ] for any 0 < T ≤ ∞. Then we have

sup 0≤t≤T k(v(t), w(t))k2L2+ 2 Z T 0 k(∇v(t), ∇w(t))k2L2dt ≤ k(v0, w0)k2L2; (11) sup 0≤t≤T k(u(t), ∇Ψ(t))k2 L2+2 Z T 0 k(∇u(t), ∆Ψ(t))k2 L2dt ≤ k(u0, ∇Ψ(0))k2L2. (12)

Proof. Multiplying the third equation of (8) by v and integrating over Ω, one has (13) 1 2 d dtkvk 2 L2+ k∇vk2L2+ Z Ω v∇w · ∇Ψdx + Z Ω vw2dx = 0,

where we have used integration by parts, the boundary conditions (10) and the divergence free condition div u = 0 to yield

Z Ω (u · ∇)vvdx = 1 2 Z Ω (u · ∇)v2dx = −1 2 Z Ω (∇ · u)v2dx = 0. Repeating the same steps for w, we have

(14) 1 2 d dtkwk 2 L2+ k∇wk2L2+ Z Ω w∇v · ∇Ψdx + Z Ω vw2dx = 0.

After integration by parts, we obtain the following cancellation by consideration of the fifth equation of (8) and the boundary conditions (10):

Z Ω w∇v · ∇Ψdx + Z Ω v∇w · ∇Ψdx = − Z Ω vw2dx. Therefore, adding up the above estimates (13) and (14) yields

(15) 1 2 d dtk(v, w)k 2 L2+ k(∇v, ∇w)k2L2+ Z Ω vw2dx = 0.

(7)

Integrating (15) over [0, t] for all 0 < t ≤ T implies that sup 0≤t≤T k(v, w)k2 L2+ 2 Z T 0 k(∇v, ∇w)k2 L2dτ + 2 Z T 0 Z Ω vw2dxdτ (16) = k(v0, w0)k2L2.

Since v is nonnegative, which can be ensured by the nonnegativity of n− and n+ (see for example [6, 16]), we get (11).

To prove (12), multiplying the first equations of (8) by u, after integration by parts, it can be easily seen that

1 2 d dtkuk 2 L2+ k∇uk2L2 = Z Ω w∇Ψ · udx. (17)

On the other hand, multiplying the fourth equation of (8) by Ψ, after integra-tion by parts and using the fifth equaintegra-tion of (8), one has

(18) 1 2 d dtk∇Ψk 2 L2+ Z Ω |∆Ψ|2dx + Z Ω v|∇Ψ|2dx + Z Ω u · ∇Ψwdx = 0. Adding up the above estimates (17) and (18) yields that

(19) 1 2 d dtk(u, ∇Ψ)k 2 L2+ k(∇u, ∆Ψ)k2L2+ Z Ω v|∇Ψ|2dx = 0.

Then (12) follows by integrating (19) in the time interval [0, t] for any 0 < t ≤ T , and using v is nonnegative. The proof of Lemma 2.2 is achieved.  Lemma 2.3. Let the assumptions (6) be in force, and let (u, v, w) be the cor-responding axisymmetric strong solution to the problem (8)–(10) on [0, T ] for any 0 < T < ∞. Then we have

sup 0≤t≤T k(∇v(t), ∇w(t))k2L2+ Z T 0 k(∆v(t), ∆w(t))k2L2dt ≤ C, (20) where C = C(ku0kL2, k(v0, w0)kH1∩L1, T ).

Proof. Multiplying the third equation of (8) by −∆v and integrating over Ω, after integration by parts, one gets

1 2 d dtk∇vk 2 L2+ k∆vk 2 L2 (21) = Z Ω (u · ∇)v∆vdx + Z Ω ∇w · ∇Ψ∆vdx + Z Ω w2∆vdx.

The three terms on the right hand side of (21) can be bounded in the follow-ing way by usfollow-ing H¨older’s inequality, (11), (12), (7) with p = 4 and Young’s inequality,

Z

(u · ∇)v∆vdx ≤ kukL4k∇vkL4k∆vkL2

(8)

≤ C(kuk12 L2k∇uk 1 2 L2+ kukL2)(k∇vk 1 2 L2k∆vk 1 2 L2+ k∇vkL2)k∆vkL2 ≤ C(1 + k∇uk12 L2)(k∇vk 1 2 L2k∆vk 1 2 L2+ k∇vkL2)k∆vkL2 ≤ C(k∇uk12 L2k∇vk 1 2 L2+k∇vk 1 2 L2)k∆vk 3 2 L2+C(k∇uk 1 2 L2k∇vkL2+k∇vkL2)k∆vkL2 ≤ 1 8k∆vk 2 L2+ C(1 + k∇uk2L2)k∇vk2L2; Z Ω ∇w · ∇Ψ∆vdx ≤ k∇wkL4k∇ΨkL4k∆vkL2 ≤ C(k∇wk12 L2k∆wk 1 2 L2+ k∇wkL2)(k∇Ψk 1 2 L2k∆Ψk 1 2 L2+ k∇ΨkL2)k∆vkL2 ≤ C(k∇wk12 L2k∆wk 1 2 L2+ k∇wkL2)k∆vkL2 ≤ 1 8k∆vk 2 L2+ C(k∇wkL2k∆wkL2+ k∇wk2L2) ≤ 1 8k∆vk 2 L2+ 1 8k∆wk 2 L2+ Ck∇wk2L2; Z Ω w2∆vdx ≤ kwk2L4k∆vkL2 ≤ C(kwkL2k∇wkL2+ kwk2L2)k∆vkL2 ≤ 1 8k∆vk 2 L2+ C(k∇wk2L2+ 1). Taking the above three estimates into (21) yields

d dtk∇vk 2 L2+ 5 4k∆vk 2 L2 (22) ≤ 1 4k∆wk 2 L2+ C(1 + k∇uk2L2)(1 + k∇vk2L2+ k∇wk2L2). Repeating the same calculations for the equation of w, we get

d dtk∇wk 2 L2+ 5 4k∆wk 2 L2 (23) ≤ 1 4k∆vk 2 L2+ C(1 + k∇uk2L2)(1 + k∇vk2L2+ k∇wk2L2). Adding up (22) and (23) provides

d dtk(∇v, ∇w)k 2 L2+ k(∆v, ∆w)k2L2 (24) ≤ C(1 + k∇uk2 L2)(1 + k∇vk2L2+ k∇wk2L2).

Notice that k∇Ψ(0)kL2 can be controlled by kw0kL2∩L1. Therefore, applying the Gronwall’s inequality to (24) yields that

sup 0≤t≤T k(∇v, ∇w)k2 L2+ Z T 0 k(∆v, ∆w)k2 L2dt ≤ C (25)

(9)

for some C = C(ku0kL2, k(v0, w0)kH1∩L1, T ). The proof of Lemma 2.3 is

achieved. 

Lemma 2.4. Let the assumptions (6) be in force, and let (u, v, w) be the cor-responding axisymmetric strong solution to the problem (8)–(10) on [0, T ] for any 0 < T < ∞. Then we have

sup 0≤t≤T k∇u(t)k2 L2+ Z T 0 k∆u(t)k2 L2dt ≤ C(ku0kH1, k(v0, w0)kL2∩L1, T ). (26)

Proof. Multiplying the first equations of (8) by −∆u and integrating over Ω, after integration by parts, we get

1 2 d dtk∇uk 2 L2+ k∆uk2L2 = Z Ω (u · ∇)u∆udx − Z Ω w∇Ψ∆udx. (27)

Similarly, we can estimate two terms on the right hand side of (27) based on the facts (11), (12) and (20):

Z

(u · ∇)u∆udx ≤ kukL4k∇ukL4k∆ukL2 ≤ C(kuk12 L2k∇uk 1 2 L2+ kukL2)(k∇uk 1 2 L2k∆uk 1 2 L2+ k∇ukL2)k∆ukL2 ≤ C(k∇uk12 L2+ k∇ukL2)k∆uk 3 2 L2+ C(k∇ukL2+ k∇uk 3 2 L2)k∆ukL2 ≤ 1 4k∆uk 2 L2+ C(1 + k∇uk2L2)k∇uk2L2; − Z Ω w∇Ψ∆udx ≤ kwkL4k∇ΨkL4k∆ukL2 ≤ C(kwk12 L2k∇wk 1 2 L2+ kwkL2)(k∇Ψk 1 2 L2k∆Ψk 1 2 L2+ k∇ΨkL2)k∆ukL2 ≤ 1 4k∆uk 2 L2+ C(1 + k∇wk2L2).

From the above two estimates, it follows easily from (27) that d

dtk∇uk

2

L2+ k∆uk2L2≤ C(1 + k∇uk2L2+ k∇wk2L2)(1 + k∇uk2L2), (28)

which yields (26) by applying Gronwall’s inequality, (11) and (12). The proof

of Lemma 2.4 is achieved. 

Lemma 2.5. Let the assumptions (6) be in force, and let (u, v, w) be the cor-responding axisymmetric strong solution to the problem (8)–(10) on [0, T ] for any 0 < T < ∞. Then we have

sup 0≤t≤T k(ut, vt, wt, ∇Ψt)k2L2+ Z T 0 k(∇ut, ∇vt, ∇wt)k2L2dt (29)

(10)

≤ C(ku0kH2, k(v0, w0)kH2∩L1, T ).

Proof. Taking the derivative to the system (8) with respect to t, we see that

(30)         

utt+ (ut· ∇)u + (u · ∇)ut− ∆ut+ ∇Pt= wt∇Ψ + w∇Ψt,

vtt+ (ut· ∇)v + (u · ∇)vt= ∇ · (∇vt− wt∇Ψ − w∇Ψt),

wtt+ (ut· ∇)w + (u · ∇)wt= ∇ · (∇wt− vt∇Ψ − v∇Ψt),

∆Ψt= wt.

We first consider the estimates for vt and wt. Multiplying the second

equa-tion of (30) by vt, the third equation of (30) by wt, and integrating over Ω,

respectively, then adding up the resultant two equalities, we get 1 2 d dtk(vt, wt)k 2 L2+ k(∇vt, ∇wt)k2L2 = − Z Ω (ut· ∇)v · vt+ (ut· ∇)w · wtdx − Z Ω ∇ · (wt∇Ψ)vt+ ∇ · (vt∇Ψ)wtdx − Z Ω ∇ · (w∇Ψt)vt+ ∇ · (v∇Ψt)wtdx := I1+ I2+ I3. (31)

Applying H¨older’s inequality, interpolation inequality (7) and Young’s inequal-ity, and using (11), (12), (20) and (26), the right hand side of (31) can be estimated as follows: I1≤ k∇vkL2kutkL4kvtkL4+ k∇wkL2kutkL4kwtkL4 ≤ C(kutk 1 2 L2k∇utk 1 2 L2+ kutkL2)(k(vt, wt)k 1 2 L2k(∇vt, ∇wt)k 1 2 L2+ k(vt, wt)kL2) ≤ 1 4k∇utk 2 L2+ 1 8k(∇vt, ∇wt)k 2 L2+ C(kutk2L2+ k(vt, wt)k2L2); I2= − Z Ω wvtwtdx ≤ kwkL2kvtkL4kwtkL4 ≤ C(kvtk 1 2 L2k∇vtk 1 2 L2+ kvtkL2)(kwtk 1 2 L2k∇wtk 1 2 L2+ kwtkL2) ≤ 1 8k(∇vt, ∇wt)k 2 L2+ Ck(vt, wt)k2L2; I3= Z Ω (w∇Ψt· ∇vt+ v∇Ψt· ∇wt)dx ≤ kwkL4k∇ΨtkL4k∇vtkL2+ kvkL4k∇ΨtkL4k∇wtkL2 ≤ C(k∇Ψtk 1 2 L2k∆Ψtk 1 2 L2+ k∇ΨtkL2)(k∇vtkL2+ k∇wtkL2) ≤ 1 8k(∇vt, ∇wt)k 2 L2+ 1 4k∆Ψtk 2 L2+ C(kwtk2L2+ k∇Ψtk2L2),

(11)

where we have used (20) and (26) to bound k(v, w)kL4 in the derivation of I3: k(v, w)kL4 ≤ C(k(v, w)k 1 2 L2k(∇v, ∇w)k 1 2 L2+ k(v, w)kL2) ≤ C. Taking all above estimates into (31), we obtain

d dtk(vt, wt)k 2 L2+ 5 4k(∇vt, ∇wt)k 2 L2 (32) ≤ 1 2k(∇ut, ∆Ψt)k 2 L2+Ck(ut, ∇Ψt, vt, wt)k2L2.

Next we derive the estimates for utand ∇Ψt. Adding up the first equations

of (30) and the third equation of (30) tested with utand Ψt, respectively, using

the relation ∆Ψt= wt, after integration by parts, we see that

1 2 d dtk(ut, ∇Ψt)k 2 L2+ k(∇ut, ∆Ψt)k2L2 = − Z Ω (ut· ∇)u · utdx + Z Ω wt∇Ψ · utdx + Z Ω (u · ∇wt) · Ψtdx + Z Ω ∇ · (vt∇Ψ + v∇Ψt)Ψtdx := I4+ I5+ I6+ I7, (33)

where we have used the cancellation relation Z Ω w∇Ψt· utdx + Z Ω (ut· ∇)wΨtdx = 0.

Applying (11), (12), (20) and (26) again, we can bound Ii (i = 4, 5, 6, 7) one

by one as follows:

I4≤ k∇ukL2kutk2L4 ≤ C(kutkL2k∇utkL2+ kutk2L2) ≤1 8k∇utk 2 L2+ Ckutk2L2; I5≤ kwtkL2k∇ΨkL4kutkL4 ≤ CkwtkL2(k∇Ψk 1 2 L2k∆Ψk 1 2 L2+ k∇ΨkL2)(kutk 1 2 L2k∇utk 1 2 L2+ kutkL2) ≤1 8k∇utk 2 L2+ C(kutk2L2+ kwtk2L2); I6= − Z Ω wt∇Ψt· udx ≤ kukL2kwtkL4k∇ΨtkL4 ≤ C(kwtk 1 2 L2k∇wtk 1 2 L2+ kwtkL2)(k∇Ψtk 1 2 L2k∆Ψtk 1 2 L2+ k∇ΨtkL2) ≤1 8k∇wtk 2 L2+ 1 8k∆Ψtk 2 L2+ C(kwtk2L2+ k∇Ψtk2L2); I7= Z Ω (vt∇Ψ · ∇Ψt+ v|∇Ψt|2)dx ≤ k∇ΨkL2kvtkL4k∇ΨtkL4+ kvkL2k∇Ψtk2L4

(12)

≤1 8k∇vtk 2 L2+ 1 8k∆Ψtk 2 L2+ C(kvtk2L2+ k∇Ψtk2L2). Taking all above estimates into (33), we have

d dtk(ut, ∇Ψt)k 2 L2+ 3 2k(∇ut, ∆Ψt)k 2 L2≤ 1 4k(∇vt, ∇wt)k 2 L2 (34) + Ck(ut, ∇Ψt, vt, wt)k2L2. Adding up (32) and (34) and using the fourth equation of (30), we finally obtain

d

dtk(ut, vt, wt, ∇Ψt)k

2

L2+k(∇ut, ∇vt, ∇wt)k2L2≤ Ck(ut, ∇Ψt, vt, wt)k2L2. (35)

Applying Gronwall’s inequality yields (29). The proof of Lemma 2.5 is achieved.  Based on Lemmas 2.2–2.5, one can apply the Galerkin approximation and the Aubin–Lions compactness principle to prove that for every T > 0, there exists a global axisymmetric strong solution (u, v, w) to the problem (8)–(10) satisfying

(36) u ∈ C([0, T ], H0,σ1 (Ω))∩L∞(0, T ; H2(Ω)), v, w ∈ C([0, T ], H2(Ω)∩L1(Ω)) and

(37) ut, vt, wt∈ L∞(0, T ; L2(Ω)), ∇ut, ∇vt, ∇wt∈ L2(0, T ; L2(Ω)).

Finally, let us prove the uniqueness of the axisymmetric strong solutions. Let (u1, v1, w1) and (u2, v2, w2) be two solutions of the problem (8)–(10) with the

same initial data and satisfy (36) and (37). Denoting δu = u1−u2, δv = v1−v2,

δw = w1− w2, δP = P1− P2, δΨ = Ψ1− Ψ2. Then we have (38)               

(δu)t+ (δu · ∇)u1+(u2· ∇)δu − ∆δu + ∇δP = δw∇Ψ1+ w2∇δΨ,

∇ · δu = 0,

(δv)t+ (δu · ∇)v1+ (u2· ∇)δv = ∇ · (∇δv − δw∇Ψ1− w2∇δΨ),

(δw)t+ (δu · ∇)w1+ (u2· ∇)δw = ∇ · (∇δw − δv∇Ψ1− v2∇δΨ),

∆δΨ = δw with the initial condition

(δu, δv, δw)|t=0= (0, 0, 0)

and the boundary conditions δu = 0, ∂(δv) ∂ν = 0, ∂(δw) ∂ν = 0, ∂(δΨ) ∂ν = 0 on ∂Ω × (0, T ). Taking the L2-inner product of the third equation of (38) with δv, the fourth

equation of (38) with δw, and adding up two resultant equalities, one has 1 2 d dtk(δv, δw)k 2 L2+ k(∇δv, ∇δw)k2L2 (39)

(13)

= − Z Ω (δu · ∇)v1δv + (δu · ∇)w1δwdx − Z Ω ∇ · (δw∇Ψ1)δv + ∇ · (δv∇Ψ1)δwdx − Z Ω ∇ · (w2∇δΨ)δv + ∇ · (v2∇δΨ)δwdx := J1+ J2+ J3.

For Ji (i = 1, 2, 3), we can derive that

J1≤ Ck(v1, w1)kL2kδukL4k(δv, δw)kL4 ≤ 1 8k∇δuk 2 L2+ 1 8k(∇δv, ∇δw)k 2 L2+ Ck(v1, w1)k2L2k(δu, δv, δw)k 2 L2; J2= − Z Ω w1δvδwdx ≤ Ckw1kL2kδvkL4kδwkL4 ≤ 1 8k(∇δv, ∇δw)k 2 L2+ Ckw1k2L2k(δv, δw)k2L2; J3≤ Ck(v2, w2)kL4k∇δΨkL4k(∇δv, ∇δw)kL2 ≤ 1 8k(∇δv, ∇δw)k 2 L2+ 1 8k∆δΨk 2 L2+ Ck(v2, w2)k2H1k∇δΨk 2 L2. Taking the above estimates into (39), we get

d dtk(δv, δw)k 2 L2+ 5 4k(∇δv, ∇δw)k 2 L2 (40) ≤ 1 4k∇δuk 2 L2+ 1 4k∆δΨk 2 L2+ Ck(v1, w1, v2, w2)k2H1k(δu, δv, δw, ∇δΨ)k 2 L2. Notice that we can rewrite the fourth equation of (38) as

(δw)t+ (u1· ∇)δw + (δu · ∇)w2= ∇ · (∇δw − δv∇Ψ1− v2∇δΨ).

(41)

Therefore, taking the L2-inner product to the first equations of (38) with δu,

(41) with δΨ, then adding up two resultant equalities together, and taking account of ∆δΨ = δw, we see that

1 2 d dtk(δu, ∇δΨ)k 2 L2+ k(∇δu, ∆δΨ)k2L2 (42) = − Z R3

(δu · ∇u1) · δudx+

Z Ω δw∇Ψ1δudx+ Z Ω ∇ · (δv∇Ψ1+v2∇δΨ)δΨdx := J4+ J5+ J6,

where we have used the facts by consideration of integration by parts and ∆δΨ = δw: Z Ω (u1· ∇)δwδΨdx = − Z Ω (u1· ∇) (∇δΨ)2 2 dx = 0

(14)

and

Z

w2∇δΨ · δu + (δu · ∇)w2δΨdx = 0.

Similarly, for Ji (i = 4, 5, 6), we obtain

J4≤ Ck∇u1kL2kδuk2L4 ≤ 1 8k∇δuk 2 L2+ Ck∇u1k2L2kδuk2L2; J5≤ Ck∇Ψ1kL2kδukL4kδwkL4≤ 1 8k∇δuk 2 L2+ 1 8k∇δwk 2 L2 + Ck∇Ψ1k2L2k(δu, δw)k2L2; J6≤ Ck(∇Ψ1, v2)kL2k(δv, ∇δΨ)kL4k∇δΨkL4 ≤ 1 8k∇δvk 2 L2+ 1 8k∆δΨk 2 L2+ Ck(∇u1, ∇Ψ1, v2)k2L2k(δv, ∇δΨ)k 2 L2. Taking all above estimates into (42) yields that

d dtk(δu, ∇δΨ)k 2 L2+ 5 4k(∇δu, ∆δΨ)k 2 L2 (43) ≤ 1 4k(∇δv, ∇δw)k 2 L2+ Ck(∇Ψ1, v2)k2L2k(δu, δv, δw, ∇δΨ)k 2 L2. Summing up (40) and (43), we conclude that

d dtk(δu, δv, δw, ∇δΨ)k 2 L2+ k(∇δu, ∇δv, ∇δw)k2L2 (44) ≤ CY(t)k(δu, δv, δw, ∇δΨ)k2 L2, where Y(t) is defined by

Y(t) :=

2

X

i=1

k(ui, vi, wi, ∇Ψi)k2H1.

Since Y(t) is integrable for the time interval [0, T ] for any 0 < T < ∞ , and Lebesgue dominated convergence theorem ensures thatRt

0Y(τ )dτ is a

continu-ous nondecreasing function which vanishes at zero. Hence, (δu, δv, δw) ≡ (0, 0) on time interval [0, t] for small enough t. Finally, because the function t → k(δu, δv, δw)kL2 is also continuous, a standard connectivity argument enables us to conclude that (δu, δv, δw) ≡ (0, 0) on Ω × [0, T ]. We complete the proof of Theorem 1.1.

References

[1] K. Abe and G. Seregin, Axisymmetric flows in the exterior of a cylinder, Proc. Roy. Soc. Edinburgh Sect. A 150 (2020), no. 4, 1671–1698. https://doi.org/10.1017/prm. 2018.121

[2] H. Abidi, R´esultats de r´egularit´e de solutions axisym´etriques pour le syst`eme de Navier-Stokes, Bull. Sci. Math. 132 (2008), no. 7, 592–624. https://doi.org/10.1016/j. bulsci.2007.10.001

(15)

[3] J. Fan, F. Li, and G. Nakamura, Regularity criteria for a mathematical model for the deformation of electrolyte droplets, Appl. Math. Lett. 26 (2013), no. 4, 494–499. https: //doi.org/10.1016/j.aml.2012.12.003

[4] J. Fan, G. Nakamura, and Y. Zhou, On the Cauchy problem for a model of electro-kinetic fluid, Appl. Math. Lett. 25 (2012), no. 1, 33–37. https://doi.org/10.1016/j. aml.2011.07.004

[5] E. Hopf, ¨Uber die Anfangswertaufgabe f¨ur die hydrodynamischen Grundgleichungen, Math. Nachr. 4 (1951), 213–231. https://doi.org/10.1002/mana.3210040121

[6] J. W. Jerome, Analytical approaches to charge transport in a moving medium, Trans-port Theory Statist. Phys. 31 (2002), no. 4-6, 333–366. https://doi.org/10.1081/TT-120015505

[7] , The steady boundary value problem for charged incompressible fluids: PNP/Navier-Stokes systems, Nonlinear Anal. 74 (2011), no. 18, 7486–7498. https: //doi.org/10.1016/j.na.2011.08.003

[8] J. W. Jerome and R. Sacco, Global weak solutions for an incompressible charged fluid with multi-scale couplings: initial-boundary-value problem, Nonlinear Anal. 71 (2009), no. 12, e2487–e2497. https://doi.org/10.1016/j.na.2009.05.047

[9] O. A. Ladyˇzenskaja, Unique global solvability of the three-dimensional Cauchy problem for the Navier-Stokes equations in the presence of axial symmetry, Zap. Nauˇcn. Sem. Leningrad. Otdel. Mat. Inst. Steklov. (LOMI) 7 (1968), 155–177.

[10] , The mathematical theory of viscous incompressible flow, Second English edition, revised and enlarged. Translated from the Russian by Richard A. Silverman and John Chu. Mathematics and its Applications, Vol. 2, Gordon and Breach, Science Publishers, New York, 1969.

[11] S. Leonardi, J. M´alek, J. Neˇcas, and M. Pokorn´y, On axially symmetric flows in R3, Z.

Anal. Anwendungen 18 (1999), no. 3, 639–649. https://doi.org/10.4171/ZAA/903 [12] J. Leray, Sur le mouvement d’un liquide visqueux emplissant l’espace, Acta Math. 63

(1934), no. 1, 193–248. https://doi.org/10.1007/BF02547354

[13] H. Ma, Global large solutions to the Navier-Stokes-Nernst-Planck-Poisson equations, Acta Appl. Math. 157 (2018), 129–140. https://doi.org/10.1007/s10440-018-0167-0 [14] I. Rubinstein, Electro-diffusion of ions, SIAM Studies in Applied Mathematics, 11, Society for Industrial and Applied Mathematics (SIAM), Philadelphia, PA, 1990. https: //doi.org/10.1137/1.9781611970814

[15] R. J. Ryham, Existence, uniqueness, regularity and long-term behavior for dissipative systems modeling electrohydrodynamics, arXiv:0910.4973v1.

[16] M. Schmuck, Analysis of the Navier-Stokes-Nernst-Planck-Poisson system, Math. Models Methods Appl. Sci. 19 (2009), no. 6, 993–1015. https://doi.org/10.1142/ S0218202509003693

[17] M. R. Ukhovskii and V. I. Iudovich, Axially symmetric flows of ideal and viscous fluids filling the whole space, J. Appl. Math. Mech. 32 (1968), 52–61. https://doi.org/10. 1016/0021-8928(68)90147-0

[18] J. Zhao, C. Deng, and S. Cui, Well-posedness of a dissipative system modeling elec-trohydrodynamics in Lebesgue spaces, Differ. Equ. Appl. 3 (2011), no. 3, 427–448. https://doi.org/10.7153/dea-03-27

[19] J. Zhao and Q. Liu, Well-posedness and decay for the dissipative system modeling electro-hydrodynamics in negative Besov spaces, J. Differential Equations 263 (2017), no. 2, 1293–1322. https://doi.org/10.1016/j.jde.2017.03.015

[20] J. Zhao, T. Zhang, and Q. Liu, Global well-posedness for the dissipative system modeling electro-hydrodynamics with large vertical velocity component in critical Besov space, Discrete Contin. Dyn. Syst. 35 (2015), no. 1, 555–582. https://doi.org/10.3934/dcds. 2015.35.555

(16)

Jihong Zhao

School of Mathematics and Information Science Baoji University of Arts and Sciences

Baoji, Shaanxi 721013, P. R. China Email address: jihzhao@163.com

참조

관련 문서

Modern Physics for Scientists and Engineers International Edition,

We derive a mathematical formulation and a process for an appro- priate control along the portion of the boundary to minimize the vorticity motion due to the flow in the

Ross: As my lawfully wedded wife, in sickness and in health, until

웹 표준을 지원하는 플랫폼에서 큰 수정없이 실행 가능함 패키징을 통해 다양한 기기를 위한 앱을 작성할 수 있음 네이티브 앱과

glen plaids 글렌 플레이드와 캐시미어 카디건, 캐리지 코트, 그리고 케이프 -&gt; 격자무늬의 캐시미어로 된 승마용 바지, 마부용 코트, 말 그림이 수

다양한 번역 작품과 번역에 관한 책을 읽는 것은 단순히 다른 시대와 언어, 문화의 교류를 넘어 지구촌이 서로 이해하고 하나가

The index is calculated with the latest 5-year auction data of 400 selected Classic, Modern, and Contemporary Chinese painting artists from major auction houses..

The “Asset Allocation” portfolio assumes the following weights: 25% in the S&amp;P 500, 10% in the Russell 2000, 15% in the MSCI EAFE, 5% in the MSCI EME, 25% in the