You are on page 1of 13

Regularity Criterion to the axially symmetric

Navier-Stokes Equations∗†
Dongyi Wei
School of Math Sciences and BICMR, Peking University,
arXiv:1508.03318v1 [math.AP] 13 Aug 2015

Beijing, 100871, People’s Republic of China.

Abstract

Smooth solutions to the axially symmetric Navier-Stokes equations obey the fol-
lowing maximum principle:kruθ (r, z, t)kL∞ ≤ kruθ (r, z, 0)kL∞ . We first prove the
global regularity of solutions if kruθ (r, z, 0)kL∞ or kruθ (r, z, t)kL∞ (r≤r0 ) is small
compared with certain dimensionless quantity of the initial data. This result im-
proves the one in Zhen Lei and Qi S. Zhang [10]. As a corollary, we also prove the
global regularity under the assumption that |ruθ (r, z, t)| ≤ | ln r|−3/2 , ∀ 0 < r ≤
δ0 ∈ (0, 1/2).
Key words axially symmetric, Navier-Stokes Equations, Regularity
Criterion.
MSC2010 35Q30, 76D05, 76D07, 76N10.

1 Introduction
In the cylindrical coordinate system with (x1 , x2 , x3 ) = (r cos θ, r sin θ, z), an axially
symmetric solution of the Navier-Stokes equations is a solution of the following form

u(x, t) = ur (r, z, t)er + uθ (r, z, t)eθ + uz (r, z, t)ez , p(x, t) = p(r, z, t),

where x
1x2   x x 
2 1
er = ,
, 0 , eθ = − , , 0 , ez = (0, 0, 1).
r r r r
In terms of (ur , uθ , uz , p), the axially symmetric Navier-Stokes equations are as follows
u2



 ∂t ur + u · ∇ur − △ur + ur2r − rθ + ∂r p = 0,


 ∂t uθ + u · ∇uθ − △uθ + u2θ + ur uθ = 0,

r r
(1.1)



 ∂ u
t z + u · ∇u z − △u z + ∂ z p = 0,


∂r (rur ) + ∂z (ruz ) = 0.


E-mail address: jnwdyi@163.com (D.Y. Wei).

1
It is well-known that finite energy smooth solutions of the Navier-Stokes equations satisfy
the following energy identity
Z t
2
ku(t)kL2 + 2 k∇u(s)k2L2 ds = ku0 k2L2 < +∞. (1.2)
0

Denote Γ = ruθ . One can easily check that


2
∂t Γ + u · ∇Γ − △Γ + ∂r Γ = 0. (1.3)
r
A significant consequence of (1.3) is that smooth solutions of the axially symmetric
Navier-Stokes equations satisfy the following maximum principle (see, for instance,[1][2])

kΓkL∞ ≤ kΓ0 kL∞ . (1.4)

We can compute the vorticity

ω = ∇ × u = ωr er + ωθ eθ + ωz ez ,

where
1
ωr = −∂z (uθ ), ωθ = ∂z (ur ) − ∂r (uz ), ωz = ∂r (ruθ ).
r
Denote
ωθ ωr ∂z uθ
Ω= , J= =− ,
r r r
then   
2 uθ
∂t Ω + u · ∇Ω − △ + ∂r Ω + 2 J = 0,


r  r

 (1.5)
2 ur
∂t J + u · ∇J − △ + ∂r J − (ωr ∂r + ωz ∂z ) = 0.


r r

We emphasis that J was introduced by Chen-Fang-Zhang in [4], while Ω appeared much


earlier and can be at least tracked back to the book of Majda-Bertozzi in [12]. Both of
the two new variables are of great importance in our work.
Our goal is to prove that the smallness of kΓkL∞ (r≤r0 ) or kΓ0 kL∞ implies the global
regularity of the solutions. Here is our result.

Theorem 1.1. Let r0 > 0. Suppose that u0 ∈ H 2 such that Γ0 ∈ L∞ . Denote

M1 = (1 + kΓ0 kL∞ )ku0 kL2 and M0 = (kJ0 kL2 + kΩ0 kL2 )M13 .

Then there exists an absolute positive constant C0 > 0 such that if


  n o −3/2
1/4 −1/2
(a) kΓkL (r≤r0 ) ≤ 1 + ln C0 max M0 , r0 M1 + 1
∞ ,

then the axially symmetric Navier-Stokes equations are globally well-posed.

2
1/4 −1/2
Remark 1.1. Choose r0 > 0 such that M0 ≥ r0 M1 and use (1.4), then we can
obtain the following global regularity condition
1/4
(b) kΓ0 kL∞ ≤ (1 + ln(C0 M0 + 1))−3/2 ,

this condition depends only on the initial value and is very useful especially when kΓ0 kL∞
is very small, in this sense it improves the result in [10]. On the other hand, if we take
r0 → 0+ in condition (a) we can obtain an important corollary.

Corollary 1.1. Let δ0 ∈ (0, 1/2), u be the strong solution of the axially symmetric
Navier-Stokes equations with initial value u0 ∈ H 2 and kΓ0 kL∞ < ∞. If

|Γ(r, z, t)| ≤ | ln r|−3/2 , ∀ 0 < r ≤ δ0 , (1.6)

then u is regular globally in time.

Denote q 
2
− 43
K(ε) = exp 2ε − 1 − 1 , K0 (ε) = exp(ε− 3 − 1)

for 0 < ε ≤ 1, then one can easily check that


1 4 4 K0 (ε)
1 + ln K(ε) + (ln K(ε))2 = ε− 3 , ε 3 K(ε) ≥ > 0,
2 C∗
for some absolute positive constant C∗ and 0 < ε ≤ 1. The use of the functions K and
K0 is due to a new important observation in Lemma 2.3.
Throughout this paper, we assume u ∈ C([0, T ∗ ); H 2) to be the unique strong solution
to the Navier-Stokes equations (1.1) with initial value u(0) = u0 , and the maximal
existence time T ∗ > 0. We also assume u to be axially symmetric with ur , uθ , uz , Γ, Ω, J
defined above, and denote

Γ0 = Γ(0), Ω0 = Ω(0), J0 = J(0), kΓkL∞ = kΓkL∞ (R3 ×(0,T ∗ )) .

Due to the regularity of solutions to Navier-Stokes equations, u ∈ C((0, T ∗); H 4 ), and

Ω, J ∈ C([0, T ∗); L2 ) ∩ C((0, T ∗ ); H 2),



uθ ur ∗ 3 ur
, ∈ C((0, T ); H ), ∂r = 0.
r r r r=0
Hence, all calculations below are legal for t ∈ (0, T ∗). (see [10] for more explaination)
Now let us recall some highlights on the study of the axially symmetric Navier-
Stokes equations. If the swirl uθ = 0, global regularity result was proved independently
by Ukhovskii and Yudovich [15], and Ladyzhenskaya [7], also [11] for a refined proof. In

3
the presence of swirl, the global regularity problem is still open. Recently, tremendous
efforts and interesting progress have been made on the regularity problem of the axi-
ally symmetric Navier-Stokes equations [1][2][3][4][5][9][10]. There are many significant
results under the sufficient conditions for regularity of axially symmetric solution of type
ur 2 3 3
ωθ ∈ Lp (0, T ; Lq (R3 )) and ∈ Lp (0, T ; Lq (R3 )), + ≤ 2, < q < +∞
r p q 2

in [1]. It has been shown in [4] that the axially symmetric solution is smooth in R3 ×(0, T ]
when r d uθ ∈ Lp (0, T ; Lq (R3 )) with
    
2 3 2 3
d ∈ [0, 1), (p, q) ∈ ,∞ × ,∞ , + ≤ 1 − d ,
1−d 1−d p q

in particular, global regularity is obtained if |Γ| ≤ Cr α for some α > 0, C > 0. In [10]
global regularity is obtained if |Γ| ≤ C| ln r|−2 for some C > 0. Clearly, our Corollary
1.1 improves the one in [10].
Here global regularity means T ∗ = +∞, and we only need to prove Ω ∈ L∞ (0, T ∗; L2 (R3 )),
hence we can use the results in [4] or [14] to obtain global regularity. We can also use
ur ur
Lemma 2.1 in section 2 to obtain ∇ ∈ L∞ (0, T ∗ ; L2 (R3 )), and ∈ L∞ (0, T ∗; L6 (R3 )),
r r
then the results in [1] or [6] imply the global regularity.
This paper is organized as follows, in section 2 we will give some notations and
3 important Lemmas, in section 3 we will first follow the proof in [10] then use the
Lemmas in section 2 to conclude the proof.

2 Notations and Lemmas


The Laplacian operator △ and the gradient operator ∇ in the cylindrical coordinate are
1 1 eθ
△ = ∂r2 + ∂r + 2 ∂θ + ∂z2 , ∇ = er ∂r + ∂θ + ez ∂z .
r r r
We will use C to denote a generic absolute positive constant whose meaning may change
from line to line. If |f |2 is axially symmetric, we will denote
Z
kf kL2 = |f |2rdrdz, dx = rdrdz.
2

The following estimate will be used very often. (see [5][8][13])


u 2 u 2
r 2 r 2
Lemma 2.1. ∇ ≤ kΩk2L2 , ∇ ≤ k∂z ΩkL2 .
r L2 r L2

4
Proof. By virtue of ∂r (rur ) + ∂z (ruz ) = 0, we can find the stream function ψθ such
that
1
ur = −∂z ψθ , uz = ∂r (rψθ ),
r
then we can compute
   
2 ψθ 2 ur
− △ + ∂r = Ω, △ + ∂r = ∂z Ω.
r r r r

Using integration by parts, we have


 
ur 2 ur u 2
ur
Z Z 2
r
− △ + ∂r dx = ∇ + (0, z, t) dz,

r r r r L2 r

therefore,  
u 2 Z
ur 2 ur
r
∇ 2 ≤ − △ + ∂r dx
r L r r r
ur ur u
Z Z
r
=− ∂z Ωdx = ∂z Ωdx ≤ ∇ kΩkL2 ,

r r r L2
hence, u
r
∇ 2 ≤ kΩkL2 .
r L
Since   2
2 ur
k∂z Ωk2L2

= △ + ∂r

r r L2
1 ur 2
u 2
ur 1 ur
Z
r
= △ + 4 ∂r + 4 △ ∂r rdrdz,

r L2 r r L2 r r r
and u 2 u 2
r 2 r ur
△ 2 = ∇ , ∂r = 0,

r L r L2 r r=0
1 ur 2
Z  
ur 1 ur 1 ur 2 ur 2
Z
△ ∂r rdrdz = ∂r +
∂z − ∂r (0, z, t)dz ≥ 0,
r r r r r L2 2 r r
u 2
r
we can obtain ∇2 ≤ k∂z Ωk2L2 , this completes the proof Lemma 2.1.

r L2
Denote
Z r v
′ ′
v(r, z, t) = |uθ (r , z, t)|dr , for r > 0, a(t) = (t) ,

0 r L∞

then we have the following inequality


u
2 θ
Lemma 2.2. a(t) ≤ kJ(t)kL2 (t) .
r L2

5
Z r′
′ ′
Proof. For r > 0, z ∈ R, t > 0 as v(r , z , t) = ′ ′
|uθ (r, z ′ , t)|dr, by Hölder inequality
0
we have
r′ r′
|uθ (r, z ′ , t)|2
Z Z
′ ′ 2
|v(r , z , t)| ≤ rdr dr
0 0 r

r ′2 r +∞
|uθ (r, z, t)|2
Z Z
= dr −∂z dz
2 0 z r
Z ′ Z +∞
r ′2 r
= dr 2Juθ (r, z, t)dz
2 Z0
u z
θ
≤ r ′2 |J| (r, z, t)rdrdz
r


′2
≤ r kJ(t)kL (t) .
2
r L2
Hence, we get
′ ′ 2
2
v(r , z , t) u
θ

a(t) = sup ≤ kJ(t)kL2 (t) 2.

r ′ r

L
r >0,z ∈R
′ ′

This completes the proof.


εK(ε)
Lemma 2.3. Assume that t > 0, kΓkL∞ (r≤r1 ) ≤ ε ≤ 1, and 0 < r1 ≤ , then
a(t)
1
|uθ (t)| 2 kΓkL∞ + ε− 3
Z Z Z
1
2
|f | dx ≤ ε− 3 |∂r f | dx + C |f |2dx, (2.1)
r r12 r
r≥ 21

2
kΓk2L∞ + ε 3
Z Z Z
2
2 2 2
|uθ (t)| |f | dx ≤ ε 3 |∂r f | dx + C |f |2 dx, (2.2)
r12 r
r≥ 21

for all axially symmetric scalar and vector functions f ∈ H 1 .

Proof. We first prove that if f = 0 for r ≥ r1 , then

|uθ (t)| 2
Z Z
− 31
|f | rdrdz ≤ ε |∂r f |2 rdrdz. (2.3)
r
Z r1

In this case, f (r , z) = − ∂r f (r, z)dr for 0 < r ′ < r1 , by Hölder inequality we have
r′

r1 r1
dr
Z Z
′ 2 2
|f (r , z)| ≤ r|∂r f (r, z)| dr .
r′ r′ r
Consequently,
r1 r1
dr ′
Z Z Z Z
′ ′ 2 ′ 2 ′
|uθ (r , z, t)||f (r , z)| dr ≤ r|∂r f (r, z)| dr |uθ (r , z, t)| dr , (2.4)
0 r′ r

6
by the definition of v we have
Z r1 r1 r1
dr ′ dr
Z Z

|uθ (r , z, t)| dr = v(r, z, t) , (2.5)
0 r′ r 0 r

|Γ| ε
by the definition of a(t) we have v(r, z, t) ≤ ra(t). On the other hand, |uθ | = ≤
r r
ε
for 0 < r < r1 . Hence, if ≤ r ≤ r1 , then
a(t)
  Z r
ε
v(r, z, t) = v , z, t + |uθ (r ′ , z, t)|dr ′
a(t) ε
Z r a(t) 
ε ε ′ ra(t)
≤ a(t) + dr = ε 1 + ln .
a(t) ε
a(t)
r′ ε

The above estimates of v implies


r1 Z ε Z εK(ε)  
dr a(t) ra(t) ra(t) dr
Z
a(t)
v(r, z, t) ≤ dr + ε 1 + ln
0 r 0 r ε ε r
Z K(ε)  a(t)
 (2.6)
dr 1 1
= ε+ ε(1 + ln r) = ε 1 + ln K(ε) + (ln K(ε))2 = ε− 3 ,
1 r 2

εK(ε)
here we used 0 < r1 ≤ . By (2.4), (2.5), (2.6), we have
a(t)
Z Z
2 − 31
′ ′ ′
|uθ (r , z, t)||f (r , z)| dr ≤ ε r|∂r f |2 dr,

integrate in z, we obtain (2.3). Now we discuss general f . Take a smooth cut-off function
1
of r such that (i) φ′ ≤ 0, (ii) φ ≡ 1 if 0 ≤ r ≤ , (iii) φ ≡ 0 if r ≥ 1. Then we have
2
  2  2 !
|uθ (t)| 2 r r
Z Z Z
|f |2 drdz

|f | rdrdz = |uθ (t)| φ f drdz + |uθ (t)| 1 − φ
r r1 r1
Z     2
r |Γ| 2
Z
− 31

≤ε ∂r φ f rdrdz + |f | drdz
r1 r≥ 21 r !
r

C 4kΓkL∞
Z Z Z
− 31 2 2
≤ε |∂r f | rdrdz + 2 |f | rdrdz + |f |2 rdrdz
r1 r≥ r21 r12 r1
r≥ 2
1 Z

kΓkL∞ + ε 3
Z
1
≤ ε− 3 |∂r f |2 rdrdz + C |f |2rdrdz,
r12 r≥ 2
r1

7
here we used (2.3) and the fact that
Z     2 Z "  2   2

∂r φ r r 2 2 ∂r φ r

f rdrdz = φ |∂r f | + |f |
r1 r1 r1
   #
r r
+∂r |f |2φ ∂r φ rdrdz
r1 r1
"  2   2 #      
r r r r
Z Z
2 2 2

= φ |∂r f | + |f | ∂r φ rdrdz − |f | ∂r φ ∂r φ r drdz
r1 r1 r1 r1
C
Z Z
≤ |∂r f |2 + 2 |f |2rdrdz.
r1 r≥ r21

Similarly we have
  2  2 !
r r
Z Z Z
|uθ (t)|2 |f |2 rdrdz = |uθ (t)|2 φ f rdrdz + |uθ (t)|2 1 − φ |f |2 rdrdz

r1 r1
  2
r |Γ|2 2
Z Z

≤ kΓkL∞ (r≤r1 ) |uθ (t)| φ f drdz + 2
|f | rdrdz
r1 r≥
r1 r
!2
C 4kΓk2L∞
Z Z Z
1 2 2
≤ εε −3
|∂r f | rdrdz + 2 |f | rdrdz + |f |2rdrdz
r1 r≥ r21 r12 r1
r≥ 2
2 Z
2
kΓkL∞ + ε
Z
2 3
≤ ε 3 |∂r f |2 rdrdz + C 2
|f |2 rdrdz.
r1 r
r≥ 21

This completes the proof.

3 Proof of the results


Proof of Theorem 1.1. By applying standard energy estimate to J equation, we have
Z  
1d 2 ur
Z Z
2
kJkL2 + J(u · ∇)Jdx − J △ + ∂r Jdx − J(ωr ∂r + ωz ∂z ) dx = 0.
2 dt r r
Using ∇ · u = 0, one has
1
Z Z
J(u · ∇)Jdx = J 2 (∇ · u)dx = 0.
2
On the other hand, by direct calculations, one has
Z  
2
Z
− J △ + ∂r Jdx = k∇JkL2 + |J(0, z, t)|2 dz.
2
r
Consequently, we have
1d ur
Z Z
kJk2L2 + k∇Jk2L2 + 2
|J(0, z, t)| dz = J(ωr ∂r + ωz ∂z ) dx. (3.1)
2 dt r

8
Similarly, by applying the energy estimate to the equation of Ω, one obtains that
1d uθ
Z Z
2 2 2
kΩkL2 + k∇ΩkL2 + |Ω(0, z, t)| dz = −2 JΩdx. (3.2)
2 dt r
Notice that
ur uθ 
Z Z 
J(ωr ∂r + ωz ∂z ) dx = [∇ × (uθ eθ )] · J∇ dx
r  r
uθ 1 1 ur 2
Z 
= uθ eθ · ∇J × ∇ dx ≤ k∇Jk2L2 + uθ ∇ ,
r 2 2 r L2

and by (3.1), we have

d ur
Z
2

kJk2L2 + k∇Jk2L2 + 2 2
|J(0, z, t)| dz ≤ uθ ∇ . (3.3)

dt r L2

Now we estimate the right hand side of (3.2) and (3.3). under the assumption of condition
(b), let
  n o −3/2
1/4 −1/2
ε = 1 + ln C0 max M0 , r0 M1 + 1 , (3.4)

then kΓkL∞ (r≤r0 ) ≤ ε ≤ 1, and we can apply Lemma 2.3 with


 
εK(ε)
r1 = r(t) = min , r0 ,
a(t)

and take f = J, Ω in (2.1), we have

uθ |uθ | 2 |uθ | 2
Z Z Z
1 1

−2 JΩdx ≤ ε 3 |Ω| dx + ε 3 |J| dx
r r 1
r
C(1 + ε 3 kΓkL∞ )
Z Z
2
≤ |∂r Ω| dx + |Ω|2 dx (3.5)
r(t)2 r(t)
r≥ 2
− 32 1
Cε (1 + ε 3 kΓk ∞ )
Z Z
2 2 L
+ε −3
|∂r J| dx + |J|2 dx.
r(t)2 r≥ 2r(t)

ur ur
Choosing f = ∂r , ∂z in (2.2), we have
r r
2 2
ur ur 2 Cε 3 (1 + ε− 3 kΓk2L∞ ) u 2
Z Z
2
2 r
uθ ∇ 2 ≤ ε 3 ∂r ∇ dx + ∇ dx. (3.6)

r L r r(t)2 r(t)
r≥ 2 r

Denote
1 2
M2 = 1 + ε 3 kΓkL∞ + ε− 3 kΓk2L∞ ,

by Lemma 2.1 we have


ur 2 u 2
Z
r 2
∂r ∇ dx ≤ ∇2 2 ≤ k∂z ΩkL2 .

r r L

9
Inserting (3.5), (3.6) into (3.2), (3.3), we have
2
!    
d ε CM2 ur 2
Z
3 2
2 2 2 2
kJkL2 + kΩkL2 ≤ ε 3 ∇ + |Ω| + |J| dx. (3.7)

dt 2 r(t)2 r≥ r(t)
2
r
Denote 2
ε3
A(t) = kJ(t)k2L2 + kΩ(t)k2L2 ,
2
then
A(t) ∈ C[0, T ∗ ) ∩ C 1 (0, T ∗ ).
By Lemma2.1 and the fact that
ur ∇ur ur
∇ = − 2 er ,
r r r
u 2 u 2
2 2 2 r θ 2
|∇u| = |∇ur | + |∇uθ | + |∇uz | + + ,
r r
we obtain
 
ur 2
Z u 2
r
∇ + |Ω| dx ≤ ∇ 2 + kΩk2L2 ≤ 2kΩk2L2 ,
2
r≥ r(t)
2
r r L
 
ur 2 C
Z Z
2 2
∇ + |Ω| + |J| dx ≤ 2
|∇u|2 dx,
r≥ 2
r(t) r r(t)
hence we have
   
d CM2 ur 2
Z
2
2 2
A(t) ≤ ε 3 ∇ + |Ω| + |J| dx

dt r(t)2 r≥ r(t)
2
r
 
CM2 C
Z
2
2 2 2
≤ min 2ε 3 kΩkL2 + kJkL2 , |∇u| dx (3.8)
r(t)2 r(t)2
k∇u(t)k2L2
 
CM2
≤ min A(t), .
r(t)2 r(t)2
εK(ε)
Fix t ∈ (0, T ∗), if r(t) = ≤ r0 , by Lemma 2.2, we have
a(t)
1
a(t)2 ≤ A(t) 2 k∇u(t)kL2 ,
1
1 a(t)2 A(t) 2 k∇u(t)kL2
2
= 2
≤ .
r(t) (εK(ε)) (εK(ε))2
And (3.8) implies
1
( 1
)
d CM2 A(t) 2 k∇u(t)kL2 A(t) 2 k∇u(t)k3L2
A(t) ≤ min A(t),
dt (εK(ε))2 (εK(ε))2
k∇u(t)k3L2
 
CM2 A(t)k∇u(t)kL2 1
= min A(t) 2 ,
(εK(ε))2 (εK(ε))2
4
CM2 A(t) 3 k∇u(t)k2L2
≤ 8 ,
(εK(ε)) 3

10
otherwise, we have r(t) = r0 and

d CM2 k∇u(t)k2L2
A(t) ≤ .
dt r04

Combining the above two cases we have


( 4
)
d A(t) 3 1
A(t) ≤ CM2 k∇u(t)k2L2 max 8 , 4
.
dt (εK(ε)) 3 r0
" ( 4
)#−1
+∞
y3 1
Z
Denote F (y) = max 8 , 4
dy, then
y (εK(ε)) 3 r0

d
F (A(t)) ≥ CM2 k∇u(t)k2L2 ,
dt
and we can use the energy identity to obtain
Z t
F (A(0)) − F (A(t)) ≤ CM2 k∇u(s)k2L2 ds ≤ C1 M2 ku0 k2L2 .
0

Therefore, if the condition (a)′ : F (A(0)) > C1 M2 ku0 k2L2 is satisfied, then

inf F (A(t)) > 0, sup A(t) < +∞, sup kΩ(t)k2L2 < +∞,
0<t<T ∗ 0<t<T ∗ 0<t<T ∗

and these imply the global regularity. n o


p
Now we claim that, if C0 > C∗ max 1, C1 /3 , then (3.4) implies condition (a)′ .
Here C1 , C∗ are absolute positive constants. Notice that

(εK(ε))2
  
F (A(0)) ≥ F max A(0),
r03
− 31
(εK(ε))2

8
= 3 max A(0), 3
(εK(ε)) 3 ,
r0

2
! 21
1 ε 3
A(0) =2 kJ(0)k2L2 + kΩ(0)k2L2 ≤ kJ0 kL2 + kΩ0 kL2 ,
2
from the definition of M2 , M1 , M0 and (1.4) we have
2 2 1
M2 < ε− 3 (1 + kΓkL∞ )2 , M2 ku0k2L2 ≤ ε− 3 M12 , A(0) 2 M13 ≤ M0 .

And (3.4) implies n o


1/4 −1/2
K0 (ε) > C0 max M0 , r0 M1 ,

11
hence we obtain
n 1
o
3/2

M2 ku0 k2L2
 32 ε−1 M13 max A(0) 2 , εK(ε)/r0

F (A(0)) 33/2 (εK(ε))4
n o
3/2
max M0 , εK(ε)M13 /r0

33/2
(ε(εK(ε))
4
)
4 3 3
3 M C
0 ∗ C ∗ M 1
≤ 3− 2 max ,
K0 (ε)4 K0 (ε)3 r03/2
 4 3  3
− 23 C∗ C∗ − 32 C∗
< 3 max 4
, 3 =3
C C C0
r 03 0
3 3 − 3
≤ 3− 2 = C1 2 ,
C1

and F (A(0)) > C1 M2 ku0k2L2 .

Therefore the claim is true, this completes the proof of Theorem 1.1.

Proof of Corollary 1.1. First, we can take r0 ∈ (0, δ0 ) such that


−1/2 1/4 −1/2
r0 M1 ≥ M0 , C0 M1 r0 + 1 < e−1 r0−1 ,

by (1.6), we have | ln r0 |−3/2 ≥ kΓkL∞ (r≤r0 ) , using the property of this r0 we have
  n o −3/2
1/4 −1/2
1 + ln C0 max M0 , r0 M1 + 1
  −3/2 −3/2
−1/2
= 1 + ln C0 r0 M1 + 1 > 1 + ln e−1 r0−1 = | ln r0 |−3/2 ,

Therefore condition (a) in Theorem 1.1 is satisfied, and we can use Theorem 1.1 to get
the global regularity, this completes the proof of Corollary 1.1.

Acknowledgments The author would like to thank the professors Gang Tian and
Zhifei Zhang for some valuable suggestions.

References
[1] D. Chae, J. Lee, On the regularity of the axisymmetric solutions of the Navier-Stokes
equations, Math. Z., 239 (2002), 645-671.

[2] C. C. Chen, R. M. Strain, T. P. Tsai, H. T. Yau, Lower bound on the blow-up rate
of the axisymmetric Navier-Stokes equations, Int. Math Res. Notices (2008), vol. 8,
artical ID rnn016, 31 pp.

12
[3] C. C. Chen, R. M. Strian, T. P. Tsai, H. T. Yau, Lower Bounds on the Blow-Up
Rate of the Axi-Symmetric Navier-Stokes Equations II, Electron. Comm. PDE. 3
(2009) 203-232.

[4] H. Chen, D. Fang, T. Zhang, Regularity of 3D axisymmetric Navier-Stokes equa-


tions. Available online at arXiv:1505.00905

[5] T. Y. Hou, Z. Lei, C. M. Li, Global reuglarity of the 3D axi-symmetric Navier-Stokes


equations with anisotropic data, Comm. P.D.E., 33 (2008), 1622-1637.

[6] A. Kubica, M. Pokorny, W. Zajaczkowski, Remarks on regularity criteria for axially


symmetric weak solutions to the Navier-Stokes equations, Math. Methods Appl. Sci.
35 (2012) no. 3, 360-371.

[7] O. A. Ladyzhenskaya, Unique global solvability of the three-dimensional Cauchy


problem for the Navier-Stokes equations in the presence of axial symmetry, Zap.
Naucn. Sem. Leningrad. Otdel. Math. Inst. Steklov. 7(1968) 155-177.

[8] Z. Lei, On Axially Symmetric Incompressible Magnetohydrodynamics in Three


Dimen- sions. Available online at arXiv:1212.5968, to appear in JDE.

[9] Z. Lei, E. A Navas, Q. S. Zhang, A priori bound on the velocity in Axially symmetric
Navier-Stokes Equations, arXiv:1309.6625v2.

[10] Z. Lei, Q. S. Zhang, Criticality of the Axially Symmetric Navier-Stokes Equations,


arXiv:1505.02628v2.

[11] S. Leonardi, J. Malek, J. Necas, M. Pokorny, On axially symmetric flows in R3 , Z.


Anal. Anwendungen, 18(1999) no.3, 639-649.

[12] A. Majda, A. Bertozzi, Vorticity and incompressible flow. Cambridge Texts in Ap-
plied Mathematics, 27. Cambridge University Press, Cambridge, 2002.

[13] C. Miao, X. Zheng, On the global well-posedness for the Boussinesq system with
horizontal dissipation. Comm. Math. Phys. 321 (2013), no. 1, 33-67.

[14] J. Neustupa, M. Pokorny, An interior regularity criterion for an axially symmet-


ric suitable weak solution to the Navier-Stokes equations, J. Math. Fluid Mech.,
2(2000), 381-399.

[15] M. R. Ukhovskii, V. I. Yudovich, Axially symmetric flows of ideal and vis-


cous fluids filling the whole space, Prikl.Mat.Meh.3259-69(Russian),translated as
J.Appl.Math.Mech. 32(1968) 52-61.

13

You might also like