ArticlePDF Available

Axisymmetric flows in the exterior of a cylinder

Authors:

Abstract

We study an initial-boundary value problem of the three-dimensional Navier-Stokes equations in the exterior of a cylinder $\Pi =\{x=(x_{h}, x_3)\ \vert \vert x_{h} \vert \gt 1\}$ , subject to the slip boundary condition. We construct unique global solutions for axisymmetric initial data $u_0\in L^{3}\cap L^{2}(\Pi )$ satisfying the decay condition of the swirl component $ru^{\theta }_{0}\in L^{\infty }(\Pi )$ .
arXiv:1708.00694v2 [math.AP] 25 Apr 2018
AXISYMMETRIC FLOWS IN THE EXTERIOR OF A CYLINDER
K. ABE AND G. SEREGIN
Abstract. We study an initial-boundary value problem of the three-dimensional Navier-
Stokes equations in the exterior of a cylinder Π = {x=(xh,x3)| |xh|>1}, subject to the
slip boundary condition. We construct unique global solutions for axisymmetric initial data
u0L3L2(Π) satisfying the decay condition of the swirl component ruθ
0L(Π).
1. Introduction
We consider the three-dimensional Navier-Stokes equations:
(1.1) tuu+u· u+p=0
div u=0in Π×(0,).
It is well known that for small initial data u0L3
σ(R3), there exists a unique global solution
uBC([0,); L3) of (1.1) [22]. However, unique existence of a global solution is unknown
in general for large initial data in L3with finite energy. Here, BC([0,); X) denotes the
space of all bounded and continuous functions from [0,) to a Banach space Xand Lp
σ(Π)
denotes the Lp-closure of compactly supported smooth solenoidal vector fields in a domain
ΠR3.
For initial data with finite energy u0L2(R3), it is well known that global Leray-Hopf
weak solutions exist [29], [19]. However, their regularity and uniqueness are unknown. For
large initial data in L3(R3), weak solutions are constructed in [7], [27]. See [40] for weak
L3-solutions.
The purpose of this paper is to construct unique global solutions of (1.1) for large ax-
isymmetric initial data in L3L2. We say that a vector field uis axisymmetric if
u(x)=tRu(Rx)xR3, η [0,2π],
for R=(er(η),eθ(η),ez) and er(η)=t(cos η, sin η, 0), eθ(η)=t(sin η, cos η, 0), ez=
t(0,0,1). We say that a scaler function pis axisymmetric if p(x)=p(Rx) for xR3
and η[0,2π]. We set the cylindrical coordinate (r, θ, z) by x1=rcos θ,x2=rsin θ,x3=z
and decompose the axisymmetric vector field into three terms:
Date: April 26, 2018.
2010 Mathematics Subject Classification. 35K20, 35B07, 35K90.
Key words and phrases. Navier-Stokes equations, Axisymmetric solutions, exterior of a cylinder.
1
2
u(x)=ur(r,z)er(θ)+uθ(r,z)eθ(θ)+uz(r,z)ez.
The azimuthal component uθis called swirl velocity (see, e.g., [34]).
Unique global solutions of (1.1) for axisymmetric initial data without swirl were first
constructed in [24], [42] by the Galerkin approximation. Later on, unique global solutions
are constructed in [28] by a strong solution approach for axisymmetric data without swirl
in H2(R3). See also [1] for H1/2(R3). For axisymmetric solutions of (1.1), the vorticity
ω=curl uis expressed by
ω=ωrer+ωθeθ+ωzez
=(zuθ)er+(zurruz)eθ+ruθ+uθ
rez,
and for v=urer+uzez, the azimuthal component ωθsatisfies the vorticity equation
(1.2) tωθ
r+v· ωθ
r + 2
rrωθ
r=zuθ
r2.
For axisymmetric solutions without swirl, the right-hand side vanishes and the global a priori
estimate
ωθ
r
L2(R3)
ωθ
0
r
L2(R3)t>0,
holds. The above vorticity estimate implies existence of unique global solutions for axisym-
metric data without swirl u0L3L2(R3). (We may assume the condition ωθ
0/rL2(R3)
since local-in-time solutions belong to H2(R3).) In other words, unique global solutions
exist for large axisymmetric initial data in L3L2(R3), provided that without swirl. For
axisymmetric data with swirl, unique existence of global solutions in R3is unknown.
In this paper, we study axisymmetric solutions with swirl in the exterior of a cylinder
Π = {x=(x1,x2,x3)R3| |xh|>1,xh=(x1,x2)},
subject to the slip boundary condition
(D(u)n)tan =0,u·n=0 on Π.(1.3)
Here, n=erdenotes the unit outward normal vector field on Π,D(u)=(u+Tu)/2 is
the deformation tensor and ftan =fn(f·n) is a tangential component of a vector field f
on Π. Since axisymmetric vector fields u=urer+uθeθ+uzezsatisfy
3
ur=0, ruθuθ=0, ruz=0 on {r=1},
subject to the slip boundary condition (1.3), the azimuthal component of vorticity ωθvan-
ishes on the boundary (see Remarks 6.1 (ii) for the Dirichlet boundary condition).
By the partial regularity result [6], it is expected that axisymmetric solutions are smooth
in the interior of Π. Moreover, as noted in [11], they will not develop singularities on the
boundary due to viscosity. See [37], [18] for partial regularity results up to the boundary
subject to the Dirichlet boundary condition. The regularity theory for the slip boundary
condition (1.3) may be simpler than that for the Dirichlet boundary condition. In fact, for a
half space regularity results are deduced from a whole space case by a reflection argument;
see [3]. In this paper, we prove that axisymmetric solutions are suciently smooth in the
exterior of the cylinder Π×(0,), subject to the slip boundary condition (1.3). We impose
the slip boundary condition in order to construct approximate solutions for R3; see Remarks
1.2 (iii).
Our goal is to construct unique global mild solutions of (1.1) for axisymmetric initial data
with swirl in L3L2(Π). Since the boundary of the cylinder ΠR3is uniformly regular,
we construct mild solutions by using the ˜
Lp-theory. We set
˜
Lp(Π)=LpL2(Π)
(resp. ˜
Lp
σ(Π)=Lp
σL2
σ(Π)) for p[2,). It is proved in [14] ( [15]) that that the Helmholtz
projection Pacts as a bounded operator on ˜
Lp(Π). Moreover, it is recently shown in [17] that
the Stokes operator subject to the slip boundary condition A=Pgenerates a C0-analytic
semigroup on ˜
Lp
σ(Π) (see also [14], [16] for the Dirichlet boundary condition). We construct
mild solutions for u0˜
L3
σ(Π) of the form
u(t)=etAu0Zt
0
e(ts)AP(u· u)(s)ds.(1.4)
Since the swirl component satisfies the Robin boundary condition, axisymmetric solutions
of (1.4) satisfy the energy equality
ZΠ|u|2dx+2Zt
0ZΠ|∇v|2+|∇uθ|2+
uθ
r
2dxds+2Zt
0ZΠ|uθ|2dHds=ZΠ|u0|2dx,(1.5)
where dHdenotes the surface element on Π.
We construct unique global solutions for large axisymmetric data with swirl u0˜
L3
σ(Π)
satisfying the decay condition of the swirl component ruθ
0L(Π). The main result of this
paper is the following:
Theorem 1.1. Let u0˜
L3
σ(Π)be an axisymmetric vector field. Assume that ruθ
0L(Π).
Then, there exists a unique axisymmetric mild solution u BC([0,); ˜
L3(Π)) satisfying
(1.5) for t 0.
4
Remarks 1.2.(i) It is unknown in general whether axisymmetric solutions in R3for u0
˜
L3
σ(R3) satisfying ruθ
0L(R3) are globally bounded for all t>0. See [35], [36], [8],
[21] for regularity criteria of axisymmetric solutions. For axisymmetric solutions, an upper
bound of the form |u(x,t)| Cr1,r<1, is called type I condition. It is proved in [9], [10] by
De Giorgi method and [23], [39] by the Liouville-type theorem that axisymmetric solutions
do not develop type I singularities. See [38] about type I singularities. Recently, it is shown
in [26] ( [31]) that axisymmetric smooth solutions in R3×(T,0) for u(·,T)L2(R3)
and ruθ(·,T)L(R3) satisfy an upper bound of the form |u(x,t)| C|log r|1/2r2near
(r,t)=0 with some constant C.
(ii) It is known that solutions of (1.1) in R3are smooth if the direction of vorticity is Lip-
schitz continuous for spatial variables in regions of high vorticity magnitude [13] (called a
geometric regularity criterion). For axisymmetric flows without swirl, vorticity varies only
in the azimuthal direction and is identified with a scalar function. On the other hand, for
axisymmetric flows with swirl vorticity varies also in the radial and vertical directions. We
constructed unique global solutions whose vorticity may become large and vary in three di-
rections. For a half space R3
+, a geometric regularity criterion is proved in [4], subject to the
slip boundary condition. See also [5] for the Dirichlet boundary condition.
(iii) Theorem 1.1 implies existence of approximate solutions for R3. Since the exterior of
the cylinder Πε={r> ε}approaches R3as ε0, axisymmetric solutions in R3can be
viewed as limits of solutions in Πε. Indeed, axisymmetric solutions without swirl in Πε
are uniformly bounded in L
tH1
xfor ε > 0 and approach those in R3[24, p.78, l.7]. See
Remarks 6.1 (iii). For the case with swirl, unique existence of global solutions is proved
in [43] ( [44]) in a bounded cylindrical domain for suciently smooth initial data. It is
unknown whether global solutions with swirl are uniformly bounded for all ε > 0. We
constructed unique global mild solutions for u0˜
L3
σ(Πε) satisfying the uniform estimate
for the swirl component (1.6).
Let us sketch the proof of Theorem 1.1. We first construct local-in-time mild solutions
of (1.4) for u0˜
L3
σand prove that mild solutions are axisymmetric and satisfy the energy
equality (1.5) for axisymmetric initial data. The major step of the proof is to derive a global
L4-bound for axisymmetric solutions u=v+uθeθ. Once we obtain the global bound, it is
not dicult to see that uBC([0,); ˜
L3) by local solvability and the energy equality (1.5).
We first prove the global L-estimate for the swirl component
||ruθ||L(Π) ||ruθ
0||L(Π)t>0.(1.6)
Since r1 in the exterior of the cylinder Π, the L-estimate (1.6) and the energy equality
(1.5) implies the global L4-bound for uθof the form
||uθ||4 ||ruθ
0||
1
2
||u0||
1
2
2t>0.(1.7)
In order to prove (1.6), we study the drift-diusion equation subject to the Robin boundary
condition:
5
(1.8)
tΓ + b· Γ∆Γ + 2
rrΓ = 0 in Π×(0,T),
nΓ + 2Γ = 0 on Π×(0,T),
Γ = Γ0on Π× {t=0}.
Here, n=rdenotes the normal derivative. The function Γ = ruθis a solution of (1.8) for
b=v. We prove the L-estimate
||Γ||L(Π) ||Γ0||L(Π)t>0,(1.9)
for solutions to (1.8). Since the sign of the coecient is plus in the Robin boundary condi-
tion, a maximum principle holds if the coecient band Γare bounded in Π×[0,T]. Then
the L-estimate (1.9) easily follows from a maximum principle (see Lemma 3.1). If Γis
decaying suciently fast as |x| , we are able to obtain (1.9) by estimating Lp-norms
of Γfor p=2mand sending m . Since we assume that ruθ
0is merely bounded, the
function ruθmay not decay as |x| . We shall prove (1.9) for non-decaying solutions Γ.
We apply the L-estimate (1.9) for ruθand obtain (1.6). Note that the boundedness of
ruθdoes not follow from properties of local-in-time solutions to (1.1) for u0˜
L3
σ. For this
purpose, we first extend the L-estimate (1.9) for mild solutions to (1.8) for Γ0Land
the coecient bsuch that t1/23/2pbC([0,T]; Lp) vanishes at time zero for p(3,]. We
then deduce from the integral form (1.4) that ruθis a mild solution to (1.8) (see Lemma 4.7).
We next estimate a global L4-norm of v=urer+uzez. We apply an interpolation inequality
||v||4C||v||
1
4
2(||v||2+||ωθ||2)3
4,(1.10)
and estimate an energy norm of the vorticity ωθ. Since ωθvanishes on the boundary, we
control the external force z(uθ/r)2by using viscosity and estimate
(1.11) ZΠ
ωθ
r
2dx+Zt
0ZΠωθ
r
2dxdsZΠ
ωθ
0
r
2dx+||ruθ
0||2
||u0||2
2
=:E t >0.
Since the above vorticity estimate implies the global bound
(1.12) ZΠ|ωθ|2dx+Zt
0ZΠ |∇ωθ|2+
ωθ
r
2!dxdsZΠ|ωθ
0|2dx
+C(E3
4||u0||
1
2
2+||ruθ
0||2
)||u0||2
2,t>0,
the local-in-time solution u=v+uθeθis globally bounded on L4.
This paper is organized as follows. In Section 2, we state a local existence theorem of
mild solutions for u0˜
L3
σand prove axial symmetry of mild solutions. In Section 3, we
6
study the drift-diusion equation (1.8) for a bounded coecient and prove the L-estimate
(1.9) by a maximum principle. In Section 4, we extend (1.9) for mild solutions to (1.8)
under the weak regularity condition of a coecient, and apply (1.9) for the swirl component
of axisymmetric solutions. In Section 5, we prove the a priori estimates (1.11) and (1.12).
In Section 6, we prove Theorem 1.1. In Appendix A, we give a proof for a local solvability
result stated in Section 2. In Appendix B, we prove some interpolation inequalities used in
Section 5.
2. Local existence of axisymmetric solutions on ˜
L3
In this section, we construct local-in-time axisymmetric solutions of (1.1) for u0˜
L3
σ
satisfying the energy equality (1.5). Local solvability for u0˜
L3
σis known for R3[22,
Theorem 3]. We give a proof for the exterior of the cylinder by using ˜
Lp-theory in Appendix
A.
2.1. Local solvability. Let Cα([δ, T]; X) denote the space of all α-th H¨older continuous
functions fC([δ, T]; X) for a Banach space X. Let Cα((0,T]; X) denote the space of
functions in Cα([δ, T]; X) for all δ(0,T). For the convenience, we denote by ˜
Lp=LpL2
also for p=.
Lemma 2.1. For u0˜
L3
σ, there exist T >0and a unique mild solution of (1.4) satisfying
t3
2(1
31
p)uC([0,T]; ˜
Lp),3p ,(2.1)
t3
2(1
31
r)+1
2uC([0,T]; ˜
Lr),3r<,(2.2)
t3/2(1/31/p)u and t3/2(1/31/r)+1/2u vanish at time zero except for p =3. Moreover,
(2.3) uCα((0,T]; ˜
L3),
uCα
2((0,T]; ˜
L3),0< α < 1.
We show that mild solutions satisfy (1.1) by applying an abstract regularity result [32,
4.3.1 Theorem 4.3.4].
Proposition 2.2. Let B be a generator of an analytic semigroup in a Banach space X with
a domain D(B). Assume that f L1(0,T;X)Cβ((0,T]; X)for β(0,1). Then,
w=Zt
0
e(ts)Bf(s)ds
belongs to Cβ((0,T]; D(B)) C1+β((0,T]; X).
7
Proposition 2.3. The mild solution u in Lemma 2.1 satisfies
uCγ((0,T]; D(A)) C1+γ((0,T]; L2),0< γ < 1
2,(2.4)
for D(A)={uL2
σH2|(D(u)n)tan =0,u·n=0Π}. In particular, u satisfies the
equations (1.1) and (1.3).
Proof. We set f=Pu· u. It follows from (2.1)-(2.3) that
||f||2 ||u· u||2 ||u||3||∇u||6C
t3
4
,
||f(t)f(τ)||2 ||(u(t)u(τ)) · u(t)||2+||u(τ)· (u(t)u(τ))||2
||u(t)u(τ)||3||∇u(t)||6+||u(τ)||6||∇u(t) u(τ)||3
C|tτ|α
t3
4
+|tτ|α
2
τ1
4for 0 < τ < tT.
Thus fL1(0,T;L2)Cα/2((0,T]; L2) for α(0,1). Applying Proposition 2.2 yields
(2.4).
2.2. Axial symmetry. We show that mild solutions are axisymmetric and satisfies the en-
ergy equality (1.5) for axisymmetric initial data.
Lemma 2.4. Assume that u0is axisymmetric. Then, the mild solution u in Lemma 2.1 is
axisymmetric and satisfies
(2.5)
tur+v· ur|uθ|2
r1
r2ur+rp=0
tuθ+v· uθ+ur
ruθ1
r2uθ=0
tuz+v· uzuz+zp=0
rur+ur
r+zuz=0
in Π×(0,T),
ur=0, ruθuθ=0, ruz=0on Π×(0,T),(2.6)
and the energy equality (1.5).
Proposition 2.5. Assume that a vector field u =urer+uθeθ+uzezsatisfies (1.3). Then,
(ur,uθ,uz)satisfies (2.6). The converse also holds.
8
Proof. By fundamental calculations using the cylindrical coordinate, we observe that
D(urer)er=rurer+1
2rθureθ+1
2zurez,
D(uθeθ)er=1
2ruθuθ
reθ,
D(uzez)er=1
2ruzez,
D(u)er=rurer+1
21
rθur+ruθuθ
reθ+1
2(zur+ruz)ez.
By (1.3), (ur,uθ,uz) satisfies (2.6). Conversely, suppose that (2.6) holds. Then,
D(u)er=rurer,u·er=0 on {r=1}.
Thus (1.3) holds for u=urer+uθeθ+uzez.
Proposition 2.6. Set the rotation operator U =Uη:L2(Π) L2(Π)by
f(x)7− tR f (Rx)
and R =(er(η),eθ(η),ez)for η[0,2π]. Then, we have
UetA f=etAU f,(2.7)
UPg=PUg,(2.8)
U(h· h)=(Uh)· (U h),(2.9)
for f L2
σ, g L2and h H1satisfying h · hL2.
Proof. We give a proof for (2.7). We are able to prove (2.8) and (2.9) by a similar way.
We set w=etA fand wη=Uηw. Since the Stokes equations are rotationally invariant, wη
satisfies
twηwη+qη=0
div wη=0in Π×(0,),
with some associated pressure qη. It follows that
wη(x)=tRw(Rx)
=wr(r, θ +η, z)er(θ)+wθ(r, θ +η, z)eθ(θ)+wz(r, θ +η, z)ez.
Since (wr,wθ,wz) satisfies (2.6) by Proposition 2.5, wηsatisfies the slip boundary condition
9
(1.3). Since wηis a unique solution of the Stokes equations for fη=Uηf, we have wη=
etA fη.
Proof of Lemma 2.4. We multiply Uby (1.4). It follows from (2.7)-(2.9) that
Uu =UetAu0Zt
0
Ue(ts)AP(u· u)(s)ds
=etAUu0Zt
0
e(ts)AP(Uu · U u)(s)ds.
Since u0is axisymmetric, u0=Uu0. Hence Uu is a mild solution of (1.1) for u0. By the
uniqueness of the mild solution, we have u=Uηufor η[0,2π]. Thus uis axisymmetric.
Since usatisfies (1.1) and (1.3) by Proposition 2.3, (ur,uθ,uz) satisfies (2.5) and (2.6). The
energy equality (1.5) follows from integration by parts.
3. A maximum principle
We consider the drift-diusion equation (1.8) with a bounded coecient and prove the
L-estimate (1.9) by a maximum principle. Let C(Π×[0,T]) denote the space of all bounded
and continuous functions in Π×[0,T]. Let C2,1(Π×[δ, T]) denote the space of all functions
fC(Π×[δ, T]) such that s
tk
xfC(Π×[δ, T]) for 2s+|k| 2. We denote by C2,1(Π×(0,T])
the space of all functions in C2,1(Π×[δ, T]) for all δ(0,T). The goal of this section is:
Lemma 3.1. Let ΓC2,1(Π×(0,T]) C(Π×[0,T]) be a solution of (1.8). Assume that
bC(Π×[0,T]). Then, the L-estimate (1.9) holds for t 0.
We prove Lemma 3.1 by a maximum principle. When Πis bounded, a maximum princi-
ple with the Robin boundary condition is known [30, Lemma 2.3]. We give a proof for the
unbounded domain Π.
Proposition 3.2 (Maximum principle).Assume that ΓC2,1(Π×(0,T]) C(Π×[0,T])
satisfies
tΓ + b· Γ∆Γ + 2
rrΓ0in Π×(0,T],(3.1)
nΓ + 2Γ0on Π×(0,T],(3.2)
Γ0on Π× {t=0}.(3.3)
Then,
10
Γ0in Π×[0,T].(3.4)
Corollary 3.3. Assume that the reverse inequalities of (3.1)-(3.3) hold. Then, Γ0in
Π×[0,T].
Proof of Lemma 3.1. We set
M=sup
xΠ
Γ0(x),
m=inf
xΠΓ0(x).
We first show (1.9) when m0. We set
Γm=mΓ.
The function Γmsatisfies (3.1) and (3.3). Since m0, it follows that
(n+2)Γm=2m(n+2)Γ
=2m0.
Hence the condition (3.2) is satisfied. Applying Proposition 3.2 implies that
mΓ(x,t) in Π×[0,T].(3.5)
We next estimate Γfrom above. We first consider the case M0. Since Γ0M0, we
apply Proposition 3.2 to Γand observe that Γ0. It follows from (3.5) that
||Γ||=inf
xΠ
Γ(x,t)
m=||Γ0||.
Thus (1.9) holds. We next consider the case M>0. We set
ΓM=MΓ.
Since (n+2)ΓM=2M>0, the reverse inequalities of (3.1)-(3.3) hold for ΓM. Applying
Corollary 3.3 implies that
Γ(x,t)Min Π×[0,T].(3.6)
11
By (3.5) and (3.6), we obtain
||Γ||=max ninf
xΠΓ(x,t),sup
xΠ
Γ(x,t)o
max{−m,M}=||Γ0||.
We proved (1.9) when m0.
It remains to show (1.9) when m>0. Since Γ0m>0, we observe that Γ0 by
Corollary 3.3. Applying Corollary 3.3 for ΓM=MΓimplies that 0 ΓM. Thus (1.9)
holds when m>0. The proof is complete.
We prove Proposition 3.2 from the following:
Proposition 3.4. We set
L=t+b· + 2
rr,
N=n· .
Assume that ΓC2,1(Π×(0,T]) C(Π×[0,T]) satisfies
(L+1)Γ0in Π×(0,T],(3.7)
(N+2)Γ0on Π×(0,T],(3.8)
Γ0on Π× {t=0}.(3.9)
Then,
Γ0in Π×[0,T].
Proof of Proposition 3.2. Applying Proposition 3.4 for ˜
Γ = Γetimplies (3.4).
We first consider the case when the function Γattains a maximum in Π. When Γattains
the maximum as |x| , we modify Γso that it attains a maximum in Π.
Proof of Proposition 3.4. We argue by contradiction. Suppose on the contrary that there
exists a point (x0,t0)Π×[0,T] such that
Γ(x0,t0)>0.(3.10)
12
We set
M=sup Γ(x,t)|xΠ,t[0,T]>0.
Case 1. The function Γattains the maximum in Π×[0,T].
We take a point (x1,t1)Π×[0,T] such that
M= Γ(x1,t1)>0.
By (3.9), we may assume that t1>0. Then, there are two cases whether x1Πor x1Π.
(a) x1Π. We observe that
tΓ(x1,t1)0,
Γ(x1,t1)=0,
∆Γ(x1,t1)0.
Hence we have
((L+1)Γ)(x1,t1)Γ(x1,t1)>0.
This contradicts (3.7). Thus the function Γdoes not attain the maximum in the interior of Π.
(b) x1Π. Since the function Γincreases along the normal direction near the boundary,
we have
Γ
n(x1,t1)0.
It follows that
((N+2)Γ)(x1,t1)2Γ(x1,t1)>0.
This contradicts (3.8). Thus the function Γdoes not attain the maximum on the boundary.
Case 2. The function Γattains the maximum at space infinity.
We modify Γand reduce the problem to Case 1. We set
Γε(x,t)= Γ(x,t)ε(At +|x|2),
13
by positive constants A, ε > 0. We shall show that, by choosing A1and εsuciently small,
depending on b,x0,t0and Γ(x0,t0), the function Γεsatisfies the conditions (3.7)-(3.10).
Once we verify these conditions, it is not dicult to derive a contradiction. In fact, the
function Γεis negative in Π {|x|>R} × [0,T] for R=M. The condition (3.10) for Γε
implies the existence of some point (x1,t1)Π {|x| R} × [0,T] such that
Mε=sup Γε(x,t)|xΠ,t[0,T]
= Γε(x1,t1)>0.
However, by the same way as we have shown in Case 1, the conditions (3.7)-(3.10) for Γε
imply that such the point (x1,t1) does not exist. Thus we are able to conclude that Case 2
does not occur neither.
It remains to show (3.7)-(3.10) for Γε. It follows that
(n+2)(At +|x|2)=(r+2)(At +r2+|z|2)
=2(At +|z|2)+2r(r1)
0,
(N+2)Γε=(N+2)Γε(n+2)(At +|x|2)0.
Thus the conditions (3.8) and (3.9) are satisfied for A, ε > 0. We show that (3.7) holds for
Γεand suciently large A. Since
L(At +|x|2)=t+b· + 2
rr(At +|x|2)
=A+2b·x2,
it follows that
(L+1)Γε=(L+1)Γε(L+1)(At +|x|2)
=(L+1)Γε(A(1 +t)+|x|2+2b·x2).
Since the function Γsatisfies (3.7), the first term of the right-hand side is negative. We set
A0=sup 2+2||b||L(Π×[0,T])|x| |x|2|xΠ>0.
It follows that
14
A(1 +t)+|x|2+2b·x2A(2 +2||b|||x| |x|2)
AA0.
Thus the condition (3.7) holds for Γεand AA0. Since
Γε(x0,t0)= Γ(x0,t0)ε(At0+|x0|2),
the condition (3.10) holds for Γε,ε < ε0and ε0= Γ(x0,t0)(At0+|x0|2)1>0. We proved
that (3.7)-(3.10) holds for Γε. The proof is now complete.
4. An a priori L-estimate for swirl
We prove the a priori L-estimate for the swirl component (1.6) (Lemma 4.7). Since the
boundedness of ruθdoes not follow from properties of local-in-time solutions to (1.1), we
extend the L-estimate (1.9) for mild solutions to (1.8). In the subsequent section, we show
that ruθis a mild solution to (1.8) and obtain the desired estimate (1.6).
4.1. Mild solutions. We define a mild solution of (1.8). We set the elliptic operators by
L0γ= γ1
r2γ,
L1Γ = ∆Γ 2
rrΓ,
subject to the Robin boundary conditions, nγ+γ=0 and nΓ + 2Γ = 0 on Π. We also
set the operator L
0= r2, subject to the Dirichlet boundary condition. By the classical
Lp-estimates for elliptic operators [2], it is known that the operators B=L0,L1,L
0generate
C0-analytic semigroups on Lpfor p(1,) [32, Theorem 3.1.3]. Moreover, the semigroups
are analytic also for p=(see [32, Corollary 3.1.24]). By analyticity of the semigroups,
they satisfy the regularizing estimate
||k
xetB f||C
t3
2p+|k|
2||f||p
(4.1)
for 0 <tT0, 3 <p and |k| 1. By using the semigroup et L1, we consider the integral
equation
Γ = etL1Γ0Zt
0
e(ts)L1(b· Γ)(s)ds.(4.2)
We assume that the coecient bsatisfies the regularity condition
15
(4.3) t1
23
2pbC0([0,T]; Lp) for 3 <p .
Here, C0([0,T]; Lp) denotes the space of all functions in C([0,T]; Lp), vanishing at time
zero. Note that solutions of (1.4) satisfies the condition (4.3) by Lemma 2.1. We prove
the L-estimate (1.9) for mild solutions ΓCw([0,T]; L) of (4.2), where Cw([0,T]; L)
denotes the space of all weakly-star continuous functions from [0,T] to L.
We first recall that mild solutions of (4.2) are H¨older continuous up to second orders in
Π×[0,T] for suciently smooth Γ0and bby the older regularity results for second order
equations [25, Chapter IV], [32, Chapter 5].
Let C(Π) denote the space of all bounded and continuous functions in Π. Let Cm(Π)
denote the space of all functions fC(Π) such that k
xfC(Π) for |k| mwith non-
negative integer m. We denote by C(Π) the space of all functions in Cm(Π) for all m1.
We denote by Cµ(Π) the space of all µ-th older continuous functions fC(Π) for µ
(0,1). For m=[m]+µ,Cm(Π) denotes the space of all functions fC[m](Π) such that
k
xfCµ(Π) for |k|=[m], where [m] is the greatest integer smaller than m>0. We denote
by Cµ,µ/2(Π×[0,T]) the parabolic older space for µ(0,2), which is the space of all
functions fC(Π×[0,T]) such that f(·,t)Cµ(Π) for t[0,T] and f(x,·)Cµ/2[0,T]
for xΠ. We denote by C2+µ,1+µ/2(Π×[0,T]) the space of all functions fC2,1(Π×[0,T])
such that s
tk
xfCµ,µ/2(Π×[0,T]) for 2s+|k| 2.
Proposition 4.1. Let T >0. Let b satisfy (4.3).
(i) For Γ0L, there exists a unique mild solution ΓCw([0,T]; L)of (4.2) such that
t1/2ΓCw([0,T]; L). If Γ0and b are axisymmetric, the mild solution Γis axisymmetric.
(ii) Assume that
bCµ,µ/2(Π×[0,T]), µ (0,1),(4.4)
Γ0C2+µ(Π)and nΓ + 2Γ = 0on Π.(4.5)
Then, the mild solution belongs to C2+µ,1+µ/2(Π×[0,T]). In particular, the L-estimate
(1.9) holds for t 0.
Proof. The assertion (i) follows from a standard iteration argument. We are able to prove
axial symmetry by a similar way as we did in the proof of Lemma 2.4. The assertion (ii)
follows from a H ¨older regularity result for second order equations [32, Theorem 5.1.21,
Corollary 5.1.22]. The L-estimate (1.9) follows from Lemma 3.1.
16
4.2. Approximation of initial data. We prove the L-estimate (1.9) without the conditions
(4.4) and (4.5) by approximation. For this purpose, we prepare H ¨older norms for space-time
functions [25]. We set the µ-th H¨older semi-norm in Q= ×(δ, T] for µ(0,1) by
[f](µ, µ
2)
Q=sup
t(δ,T]
[f](µ)
(t)+sup
x
[f](µ
2)
(δ,T](x),
[f](µ)
(t)=sup (|f(x,t)f(y,t)|
|xy|µ
x,y,x,y),
[f](µ
2)
(δ,T](x)=sup (|f(x,t)f(x,s)|
|ts|µ
2
t,s(δ, T],t,s).
When µ=1, we set
[f](1,1
2)
Q=||∇f||L(Q)+sup
x
[f](1
2)
(δ,T](x).
For m=[m]+µ, we set
[f](m,m
2)
Q=X
2s+|k|=[m]
[s
tk
xf](µ, µ
2)
Q,
|f|(m,m
2)
Q=X
2s+|k|≤[m]||s
tk
xf||L(Q)+[f](m,m
2)
Q.
We first remove the condition (4.5) by approximation of Γ0L.
Proposition 4.2. For Γ0L(Π), there exists a sequence {Γ0} C(Π)supported in Π
such that
(4.6) ||Γ0|| ||Γ0||
Γ0 Γ0a.e. in Π.
Proof. For x=rer(θ)+zez, we set
˜
Γ0(x)=
Γ0((rε)er(θ)+zez)r1+ε,
0 0 r<1+ε.
By mollification of ˜
Γ0, we obtain the desired sequence.
Proposition 4.3. In Proposition 4.1 (ii), the estimate (1.9) holds without the condition (4.5).
17
Proof. For Γ0L, we take a sequence {Γ0}satisfying (4.6). Since Γ0 is smooth in Π
and supported in Π, it satisfies the condition (4.5). Since the estimate (1.9) holds for the
mild solution Γεof (4.2) for Γ0 by Proposition 4.1 (ii), it follows from (4.6) that
||Γε|| ||Γ0||t>0.(4.7)
We shall show that Γεconverges to a mild solution of (4.2) for Γ0. We use the H ¨older
continuity of the coecient bin (4.4). We apply the local H¨older estimate for parabolic
equations [25, Chapter IV, Theorem 10.1] and estimate
|Γε|(2+µ,1+µ
2)
QC||Γε||L(Π×(0,T))
(4.8)
for Q=(BΠ)×(δ, T] and δ > 0 with some constant C, independent of ε. Here, BR3
denotes an open ball satisfying BΠ,. By (4.7) and (4.8), Γεsubsequently converges to
a limit Γlocally uniformly in Π×(0,T] up to second derivatives.
It is not dicult to see that the limit Γis a mild solution of (4.2) for Γ0. In fact, by
choosing a subsequence, we have
etL1Γ0 etL1Γ0locally uniformly in Π×(0,T].
Since Γεconverges to Γlocally uniformly in Π×(0,T], similarly for each 0 <s<t, we
have
eρL1b· ΓεeρL1b· Γlocally uniformly in Π×(0,T].
Hence sending ε0 to (4.2) implies the limit Γis a mild solution for Γ0. The estimate
(4.7) is inherited to the limit Γ.
4.3. Approximation of a coecient. We next remove the condition (4.4).
Proposition 4.4. For b satisfying (4.3), there exists a sequence {bε} C(Π×[0,T]) satis-
fying (4.3) and
lim
ε0sup
0tT
t1
23
2p||bbε||p(t)=0for 3<p<.(4.9)
Proof. We may assume that bis smooth in Πby mollification by spatial variables. Since g=
t1/23/(2p)bvanishes at time zero by (4.3), by shifting gby a time variable, and mollification,
we obtain a sequence {gε} C(Π×[0,T]) such that gε(·,t) is supported in (0,T] and
18
lim
ε0sup
0tT||gεg||p(t)=0.(4.10)
Since gε(·,t) is supported in (0,T], the function bε=t1/2+3/2pgεis smooth in Π×[0,T]
and satisfies (4.3). The convergence (4.9) follows from (4.10).
Lemma 4.5. The estimate (1.9) holds for mild solutions of (4.2) for Γ0Land t >0.
Proof. We shall show the estimate (1.9) between 0 <tT1for some T1>0. Once we
have (1.9) near time zero, it is extendable for all t>0 by taking t=T1as an initial time.
We take a sequence {bε}satisfying (4.9). Since the estimate (1.9) holds for a mild solution
Γεfor Γ0Land the coecient bεby Proposition 4.3, we have
||Γε|| ||Γ0||t>0.(4.11)
We shall show that Γεconverges to a mild solution Γin the sense that
lim
ε0sup
0<tT1n||ΓΓε||+t1
2||∇(ΓΓε)||o=0.(4.12)
The desired estimate follows from (4.11) and (4.12) by sending ε0.
We set ρε= Γ Γεand aε=bbε. It follows from (4.2) that
ρε=Zt
0
e(ts)L1(b· Γbε· Γε)ds
=Zt
0
e(ts)L1(aε· Γ + bε· ρε)ds.
For p(3,), we set the constants
Kε=sup
0<tT1||ρε||+t1
2||∇ρε||,
K=sup
0<tT1||Γ||+t1
2||∇Γ||,
Nε=sup
0tT1
t1
23
2p||aε||p,
Lε=sup
0tT1
t1
23
2p||bε||p.
It follows from (4.1) that
19
||ρε||Zt
0
C
(ts)3
2p||aε||p||∇Γ||+||bε||p||∇ρε||ds
C(NεK+LεKε)Zt
0
ds
(ts)3
2ps13
2p
=C0(NεK+LεKε).
Similarly, we estimate ρεand obtain
KεC0(NεK+LεKε).
We take an arbitrary δ > 0. By (4.3), there exists T1>0 such that
sup
0<tT1
t1
23
2p||b||pδ.
By (4.9), that there exits ε0>0 such that
sup
0<tT1
t1
23
2p||bbε||pδfor εε0.
We estimate
Lε=sup
0<tT1
t1
23
2p||bε||psup
0<tT1
t1
23
2p||bεb||p+sup
0<tT1
t1
23
2p||b||p
2δfor εε0.
By taking δ=(4C0)1, we estimate
Kε2C0NεK.
Since Nε0 as ε0 by (4.9), we proved (4.12).
4.4. An application to axisymmetric solutions. We now prove the a priori L-estimate
for the swirl component (1.6). It suces to show that the swirl component ruθis a mild
solution of (4.2).
Proposition 4.6. The semigroups satisfy
20
eθ·etA f=etL0fθ,(4.13)
eθ·curl etAg=et L
0eθ·curl g,(4.14)
retL0γ=etL1(rγ),(4.15)
for axisymmetric f , g L2
σand γLsatisfying eθ·curl g L2and rγL.
Proof. We set w=etA f. Since wθ=eθ·wsatisfies
twθ1
r2wθ=0 in Π×(0,),
nwθ+wθ=0 on Π×(0,),
wθ=fθon Π× {t=0},
the function wθagrees with etL0fθby the uniqueness of the heat equation. Similarly, we are
able to prove (4.14) and (4.15).
Lemma 4.7 (A priori L-estimate).Let u =v+uθeθbe an axisymmetric mild solution of
(1.4) in Lemma 2.4. Assume that ruθ
0L. Then, ruθis a mild solution of (4.2) for b =v.
In particular, the L-estimate (1.6) holds.
Proof. We set
(4.16) h=u· u=v· ur|uθ|2
rer+v· uθ+ur
ruθeθ+(v· uz)ez,
Φ = (IP)h,
for axisymmetric mild solutions uin [0,T]. Since his axisymmetric, the function Φis
independent of θand we have
eθ·Pu· u=eθ·(h Φ)
=eθ·(herrΦezzΦ)
=v· uθ+ur
ruθ.
We multiply eθby (1.4). It follows from (4.13) that
uθ=etL0uθ
0Zt
0
e(ts)L0v· uθ+ur
ruθds.
21
Since uθ
0Lby ruθ
0L, the above integral form implies that uθCw([0,T]; L) and
t1/2uθCw([0,T]; L).
On the other hand, there exists a unique axisymmetric mild solution Γfor Γ0=ruθ
0and
b=vby Proposition 4.1 (i). We multiply r1by (4.2). It follows from (4.15) that γ= Γ/r
satisfies
γ=etL0uθ
0Zt
0
e(ts)L0v· γ+ur
rγds.
Since γCw([0,T]; L) and t1/2γCw([0,T]; L), it is not dicult to show that uθ
agrees with γby estimating the dierence uθγ. Thus ruθis a mild solution of (4.2). The
proof is now complete.
5. Energy estimates for the azimuthal component of vorticity
We prove the global estimates (1.11) and (1.12).
(5.1)
tωθ+v· ωθur
rωθ1
r2ωθ=z|uθ|2
rin Π×(0,T),
ωθ=0 on Π×(0,T),
ωθ=ωθ
0on Π× {t=0}.
Proposition 5.1. Axisymmetric mild solutions of (1.4) in Lemma 2.4 satisfy
ωθCγ((0,T]; D(L
0)) C1+γ((0,T]; L2),0< γ < 1
2.(5.2)
In particular, ωθsatisfies the vorticity equation (5.1), where D(L
0)=H2H1
0denotes the
domain of the operator L
0= r2on L2.
Proof. We recall that the mild solution uis expressed by
u(η+δ)=eηAu(δ)+Zη
0
e(ηs)Af(δ+s)ds,
for 0 δηTδand f=Pu· u. It follows from (4.16) that
22
g:=eθ·curl f
=eθ·curl (h Φ)
=zhr+rhz
=v· ωθ+ur
rωθ+z|uθ|2
r.
We multiply eθ·curl by u. It follows from (4.14) that
ωθ(η+δ)=eηL
0ωθ(δ)+Zη
0
e(ηs)L
0g(δ+s)ds.
Since uCγ((0,T]; H2) for γ(0,1/2) by (2.4), we have gCγ((0,T]; L2). By Proposi-
tion 2.2, ωθsatisfies (5.2).
Proposition 5.2. The function v =urer+uzezsatisfies
||∇v||2=||ωθ||2,(5.3)
||ur||4C||ur||
1
4
2||ωθ||
3
4
2,(5.4)
||uz||4C||uz||
1
4
2(||uz||2+||ωθ||2)3
4,(5.5)
||ωθ||4C||ωθ||
1
4
2||∇ωθ||
3
4
2,(5.6)
||∇(ωθeθ)||2=||∇ωθ||2
2+
ωθ
r
2
21
2,(5.7)
with some constant C. In particular, the estimate (1.10) holds.
Proof. Since vsatisfies
v=curl curl v div v
=curl (ωθeθ),
it follows from (2.6) that
ZΠ|∇v|2dx=ZΠ
v·vdx+ZΠ
v
n·vdH
=ZΠ
curl (ωθeθ)·vdxZΠ
(rurur+ruzuz)dH
=ZΠ|ωθ|2dx.
23
Thus (5.3) holds. Similarly, we obtain (5.7) by integration by parts. Since vrand ωθvanish
on Π, applying the interpolation inequality (B.1) implies (5.4) and (5.6). We apply (B.2)
for vzand obtain (5.5).
Lemma 5.3. The estimates (1.11) and (1.12) hold for t >0for axisymmetric mild solutions
for u0˜
L3
σsatisfying ruθ
0Land u0L2with some constant C.
Proof. We prove (1.11). Since ωθ/rsatisfies (1.2) and vanishes on Π, by multiplying 2ωθ/r
by (1.2) and integration by parts, we have
d
dtZΠ
ωθ
r
2dx+2ZΠωθ
r
2dx=2ZΠuθ
r2zωθ
rdx.
Since r1, it follows that
uθ
r
4 ||uθ||4=
(ruθ)1
2uθ
r1
2
4 ||ruθ||
1
2
uθ
r
1
2
2.(5.8)
By the Young’s inequality, we estimate
2ZΠuθ
r2zωθ
rdx
uθ
r
4
4+
ωθ
r
2
2
||ruθ||2
uθ
r
2
2+
ωθ
r
2
2.
Hence
d
dtZΠ
ωθ
r
2dx+ZΠωθ
r
2dx ||ruθ||2
uθ
r
2
2.
We integrate the both sides between (0,t). By applying (1.5) and (1.6), we obtain (1.11).
We prove (1.12). We multiply 2ωθby (5.1) to see that
d
dtZΠ|ωθ|2dx+2ZΠ|∇ωθ|2+
ωθ
r
2dx=2ZΠ
ur
rωθωθdx+2ZΠ
z|uθ|2
rωθdx
=:I+II.
It follows from (5.4), (5.6) and (5.7) that
24
|I|=2ZΠ
ur
rωθωθdx2
ωθ
r
2||ur||4||ωθ||4
C
ωθ
r
2||ur||
1
4
2||ωθ||2||∇ωθ||
3
4
2
C
ωθ
r
3
4
2||ur||
1
4
2||ωθ||2||∇(ωθeθ)||2
C
ωθ
r
3
2
2||ur||
1
2
2||ωθ||2
2+1
2||∇(ωθeθ)||2
2.
Since r1, it follows from (5.8) that
|II|=2ZΠ
z|uθ|2
rωθdx2||uθ||2
4||∇ωθ||2
2||ruθ||
uθ
r
2||∇(ωθeθ)||2
2||ruθ||2
uθ
r
2
2+1
2||∇(ωθeθ)||2
2.
By combining the estimates for Iand II, we obtain
d
dtZΠ|ωθ|2dx+ZΠ|∇ωθ|2+
ωθ
r
2dxC
ωθ
r
3
2
2||ur||
1
2
2+||ruθ||2
||ωθ||2
2+
uθ
r
2
2.
We integrate the both sides between (0,t). By (1.11), (1.5) and (1.6), we obtain (1.12).
6. Global bounds on L4
Proof of Theorem 1.1. For an axisymmetric u0˜
L3
σsatisfying ruθ
0L, the axisymmetric
mild solution uC([0,T]; ˜
L3) satisfies (1.6) by Lemma 4.7. It follows from (1.6) and (2.2)
that ruθ
0(·,t0)Land
u(·,t0)L2for t0(0,T].
We may assume that u0L2by taking t=t0as an initial time. It follows from (1.7) and
(1.10)-(1.12) that uL(0,;L4). By (1.5) and Lemma 2.1, the mild solution belongs to
BC([0,); ˜
L3). The proof is now complete.
25
Remarks 6.1.(i) (The Euler equations) We constructed global solutions in the exterior do-
main by using viscosity. For the Euler equations, existence of global solutions is unknown.
We refer to [33], [20] for a one-dimensional blow-up model of axisymmetric Euler flows on
the boundary. See [12], [11] for blow-up results of models.
(ii) (The Dirichlet boundary condition) The statement of Lemma 2.1 is valid also for the
Dirichlet boundary condition. However, in this case unique existence of global solution
is unknown even for axisymmetric data without swirl. Since the azimuthal component of
vorticity ωθdoes not vanish on the boundary subject to the Dirichlet boundary condition,
the global vorticity estimates (1.11) and (1.12) are not available unlike the slip boundary
condition.
(iii) (Uniform estimates) The assertion of Theorem 1.1 is valid also for the exterior of the
cylinder Πε={r> ε}and we are able to construct global solutions u=uεsatisfying (1.6)
and the energy equality
(6.1) ZΠε|u|2dx+2Zt
0ZΠε|∇v|2+|∇uθ|2+
uθ
r
2dxds
+2
εZt
0ZΠε|uθ|2dHds=ZΠε|u0|2dx,t0.
For the case without swirl, the a priori estimates
ZΠε
ωθ
r
2dx+2Zt
0ZΠεωθ
r
2dxdsZΠε
ωθ
0
r
2dx,(6.2)
(6.3) ZΠε|ωθ|2dx+Zt
0ZΠε|∇ωθ|2+
ωθ
r
2dxds
ZΠε|ωθ
0|2dx+C
ωθ
0
r
3
2
L2(Πε)||u0||
5
2
L2(Πε),t>0,
hold with some constant C, independent of ε. Hence we have a uniform bound
sup
εε0||u||L(0,;H1(Πε)) <,
provided that L2-norms of uθ
0,ωθ
0/rand ωθ
0in Πεare uniformly bounded for εε0.
Appendix A. Mild solutions on ˜
L3
We give a proof for local solvability of (1.1) on ˜
L3(Lemma 2.1).
Proposition A.1. The Stokes semigroup satisfies
26
||k
xetA f||˜
LpC
t3
2(1
q1
p)+|k|
2||f||˜
Lq,(A.1)
||k
x(eρA1)eηAf||˜
LqCρα
ηα+|k|
2||f||˜
Lq,(A.2)
for f ˜
Lq
σ,2qp<,0<t, ρ, η T0,α(0,1),|k| 1and T0>0.
Proof. The estimates (A.1) and (A.2) follow from estimates of the Stokes semigroup on
˜
Lq[17, Theorem 1.2] and the interpolation inequality (B.2).
Proposition A.2. For u0˜
L3
σ, there exists T >0and a unique mild solution of (1.4)
satisfying (2.1) and (2.2).
Proof. We set
uj+1=etAu0Zt
0
e(ts)AP(uj· uj)ds,
u1=etAu0,
Kj=sup
0tT
tγ{||uj||˜
Lp+t1
2||∇uj||˜
Lp},
for γ=1/23/(2p) and p(3,). We take q[2,p]. By applying the Young’s inequality,
we estimate
||uj· uj||q ||uj||η||∇uj||p
for 1 =1/q1/p. Since q[2,p] and p>3, we observe that σ=3(1/p1)=
3(2/p1/q)3/p<1. By applying the interpolation inequality (B.2), we estimate
||uj||LηC||uj||1σ
Lp||∇uj||σ
W1,p.
We take T1 and estimate
||uj· uj||LqC||uj||1σ
Lp||∇uj||1+σ
W1,pC Kj
sγ!1σ 2Kj
sγ+1
2!1+σ
CK2
j
s3
2(11
q).
Since the above estimate holds for sT1, we have
||uj· uj||˜
Lq=max{||uj· uj||Lq,||uj· uj||L2}
CK2
j
s3
2(11
q).(A.4)
Applying (A.1) implies
27
||k
xe(ts)APuj· uj||˜
Lp
CK2
j
(ts)3
2(1
q1
p)+|k|
2s3
2(11
q).(A.5)
We estimate Kj+1. We set p0=max{3p/(p+3),2}and fix q(p0,3) so that the right
hand-side of (A.5) is integrable near s=tfor |k| 1. It follows from (A.1) and (A.5) that
||uj+1||˜
Lp ||etAu0||˜
Lp+C
t1
23
2p
K2
j.
Similarly, we estimate uj+1and obtain Kj+1K1+C0K2
j. Since the Stokes semigroup is
strongly continuous on ˜
L3, we have K10 as T0. We take T>0 suciently small so
that K1(4C0)1and
Kj2K1for j=1,2,···.
By a similar way, we estimate the dierence uj+1ujand obtain
sup
0tT
tγ(||uj+1uj||˜
Lp+t1
2||∇(uj+1uj)||˜
Lp)0 as j .
Thus the sequence {uj}converges to a mild solution usatisfying (2.1) and (2.2) for p,r
(3,). In particular, we have
K=sup
0tT
tγ{||u||˜
Lp+t1
2||∇u||˜
Lp} 2K1.(A.6)
The uniqueness follows from the integral form since tγuand tγ+1/2uvanish at time zero.
It remains to show (2.1) and (2.2) at end points. The property (2.1) for p=follows
from the interpolation inequality (B.2). It follows from (A.1), (A.4) and (A.6) that
||uetAu0||˜
L3Zt
0||e(ts)APu· u||˜
L3ds
CK2
1Zt
0
ds
(ts)3
2(1
q1
3)s3
2(11
q)=CK2
1.
Since K10 as T0, the mild solution uis strongly continuous on ˜
L3at time zero.
Thus (2.1) holds for p=3. By a similar way, we estimate t1
2||∇u etAu0||˜
L3CK2
1. Since
t1/2etAu0vanishes on ˜
L3at time zero, (2.2) holds for r=3.
Proof of Lemma 2.1. It remains to show the H ¨older continuity (2.3). We set f=Pu· u.
It follows from (A.4) that
28
||f||˜
LqCK2
s3
2(11
q)for 2 q3.(A.7)
We take an arbitrary δ(0,T) and α(0,1). For δτ < tT, we estimate
||u(t)u(τ)||˜
L3 ||etAu0eτAu0||˜
L3+Zt
τ||e(ts)Af||˜
L3ds+Zτ
0||e(ts)Afe(τs)Af||˜
L3ds
=:I+II +III.
It follows from (A.2), (A.1) and (A.7) that
ICδα(tτ)α||u0||˜
L3,
II CZt
τ||f||˜
L3dsCδ1K2(tτ).
We estimate III. Since
e(ts)Afe(τs)Af=(e(tτ)A1)e(τs)
2Ae(τs)
2Af,
it follows from (A.2), (A.1) and (A.6) that
||e(ts)Afe(τs)Af||˜
L3Ctτ
τsα||e(τs)
2Af||˜
L3
C(tτ)α
(τs)α+3
2(1
q1
3)||f||˜
Lq
C′′K2(tτ)α
(τs)α+3
2(1
q1
3)s3
2(11
q).
We take q[2,3) so that 3/2(1/q1/3) <1αand obtain
III CδαK2(tτ)α.
Thus uCα([δ, T]; ˜
L3) for α(0,1). By a similar way, uCα/2([δ, T]; L2) follows. We
proved (2.3). The proof is now complete.
29
Appendix B. Interpolation inequalities
We give a proof for interpolation inequalities used in Proposition 5.2.
Lemma B.1. The estimates
||ϕ||pC||ϕ||1σ
q||∇ϕ||σ
q, ϕ W1,q
0,(B.1)
||φ||pC||φ||1σ
q||φ||σ
1,q, φ W1,q,(B.2)
hold for 1qp satisfying σ=3(1/q1/p)<1, where W1,q
0denotes the space of
functions in W1,q, vanishing on Π.
Proof. The estimate (B.1) for Π = R3holds by estimates of the heat semigroup. Since the
trace of ϕW1,q
0vanishes on Π, we apply (B.1) to the zero extension of ϕto R3and obtain
the desired estimate for ΠR3. For functions φW1,qwith non-trivial traces, we use an
extension operator E:W1,q(Π) W1,q(R3) acting as a bounded operator also from Lq(Π)
to Lq(R3) [41, Chapter VI, 3.1 Theorem 5]. By applying (B.1) for R3and Eφ, we obtain
(B.2).
Acknowledgements
The first author would like to thank Oxford University for their hospitality from October
2015 to January 2016. The first author was supported by JSPS through the Grant-in-aid
for Research Activity Start-up 15H06312, Young Scientist (B) 17K14217 and Scientific
Research (B) 17H02853. The second author was supported by the Ministry of Education
and Science of the Russian Federation (grant 14.Z50.31.0037).
References
[1] H. Abidi. R ´esultats de egularit´e de solutions axisym´etriques pour le syst`eme de Navier-Stokes. Bull. Sci.
Math., 132(7):592–624, (2008).
[2] S. Agmon, A. Douglis, and L. Nirenberg. Estimates near the boundary for solutions of elliptic partial
dierential equations satisfying general boundary conditions. I. Comm. Pure Appl. Math., 12:623–727,
(1959).
[3] H.-O. Bae and B. Jin. Regularity for the Navier-Stokes equations with slip boundary condition. Proc. Amer.
Math. Soc., 136:2439–2443, (2008).
[4] H. Beir ˜ao da Veiga. Vorticity and regularity for flows under the Navier boundary condition. Commun. Pure
Appl. Anal., 5:907–918, (2006).
[5] H. Beiao da Veiga. Vorticity and regularity for viscous incompressible flows under the Dirichlet boundary
condition. Results and related open problems. J. Math. Fluid Mech., 9:506–516, (2007).
[6] L. Caarelli, R. Kohn, and L. Nirenberg. Partial regularity of suitable weak solutions of the Navier-Stokes
equations. Comm. Pure Appl. Math., 35:771–831, (1982).
[7] C. P. Calder´on. Existence of weak solutions for the Navier-Stokes equations with initial data in Lp.Trans.
Amer. Math. Soc., 318:179–200, (1990).
30
[8] D. Chae and J. Lee. On the regularity of the axisymmetric solutions of the Navier-Stokes equations. Math.
Z., 239:645–671, (2002).
[9] C.-C. Chen, R. M. Strain, T.-P. Tsai, and H.-T. Yau. Lower bound on the blow-up rate of the axisymmetric
Navier-Stokes equations. Int. Math. Res. Not. IMRN, pages Art. ID rnn016, 31, (2008).
[10] C.-C. Chen, R. M. Strain, T.-P. Tsai, and H.-T. Yau. Lower bounds on the blow-up rate of the axisymmetric
Navier-Stokes equations. II. Comm. Partial Dierential Equations, 34:203–232, (2009).
[11] K. Choi, T. Y. Hou, A. Kiselev, G. Luo, V. Sverak, and Y. Yao. On the finite-time blowup of a one-
dimensional model for the three-dimensional axisymmetric Euler equations. Comm. Pure Appl. Math.,
70:2218–2243, (2017).
[12] K. Choi, A. Kiselev, and Y. Yao. Finite time blow up for a 1D model of 2D Boussinesq system. Comm.
Math. Phys., 334:1667–1679, (2015).
[13] P. Constantin and C. Feerman. Direction of vorticity and the problem of global regularity for the Navier-
Stokes equations. Indiana Univ. Math. J., 42:775–789, (1993).
[14] R. Farwig, H. Kozono, and H. Sohr. An Lq-approach to Stokes and Navier-Stokes equations in general
domains. Acta Math., 195:21–53, (2005).
[15] R. Farwig, H. Kozono, and H. Sohr. On the Helmholtz decomposition in general unbounded domains.
Arch. Math. (Basel), 88:239–248, (2007).
[16] R. Farwig, H. Kozono, and H. Sohr. On the Stokes operator in general unbounded domains. Hokkaido
Math. J., 38:111–136, (2009).
[17] R. Farwig and V. Rosteck. Resolvent estimates of the Stokes system with Navier boundary conditions in
general unbounded domains. Adv. Dierential Equations, 21:401–428, (2016).
[18] S. Gustafson, K. Kang, and T.-P. Tsai. Regularity criteria for suitable weak solutions of the Navier-Stokes
equations near the boundary. J. Dierential Equations, 226:594–618, (2006).
[19] E. Hopf. ¨uber die Anfangswertaufgabe ur die hydrodynamischen Grundgleichungen. Math. Nachr.,
4:213–231, (1951).
[20] T. Hou and G. Luo. On the finite-time blow up of a 1d model for the 3d incompressible euler equations,
arxiv:1311.2613.
[21] Q. Jiu and Z. Xin. Some regularity criteria on suitable weak solutions of the 3-D incompressible axisym-
metric Navier-Stokes equations. In Lectures on partial dierential equations, New Stud. Adv. Math., pages
119–139. Int. Press, Somerville, MA, (2003).
[22] T. Kato. Strong Lp-solutions of the Navier-Stokes equation in Rm, with applications to weak solutions.
Math. Z., 187:471–480, (1984).
[23] G. Koch, N. Nadirashvili, G. Seregin, and V. ˇ
Sver´ak. Liouville theorems for the Navier-Stokes equations
and applications. Acta Math., 203:83–105, (2009).
[24] O. A. Lady ˇzenskaya. 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), 7:155–177, (1968).
[25] O. A. Lady ˇzhenskaya, V. A. Solonnikov, and N. N. Uraltseva. Linear and Quasilinear Equations of Par-
abolic Type., volume 23 of Translations of Mathematical Monographs. American Mathematical Society,
Providence, RI, 1968.
[26] Z. Lei, E. A. Navas, and Qi S. Zhang. A priori bound on the velocity in axially symmetric Navier-Stokes
equations. Comm. Math. Phys., 341:289–307, (2016).
[27] P. G. Lemari´e-Rieusset. Recent developments in the Navier-Stokes problem, volume 431. Chapman &
Hall/CRC, Boca Raton, FL, 2002.
[28] S. Leonardi, J. M ´alek, J. Neˇcas, and M. Pokorn´y. On axially symmetric flows in R3.Z. Anal. Anwendungen,
18:639–649, (1999).
[29] J. Leray. Sur le mouvement d’un liquide visqueux emplissant l’espace. Acta Math., 63:193–248, (1934).
[30] G. M. Lieberman. Second order parabolic dierential equations. World Scientific Publishing Co., Inc.,
River Edge, NJ, 1996.
[31] J.B. Loftus and Qi S. Zhang. A priori bounds for the vorticity of axially symmetric solutions to the Navier-
Stokes equations. Adv. Dierential Equations, 15:531–560, (2010).
[32] A. Lunardi. Analytic semigroups and optimal regularity in parabolic problems. Progress in Nonlinear
Dierential Equations and their Applications, 16. Birkh¨auser Verlag, Basel, 1995.
31
[33] G. Luo and T. Hou. Potentially singular solutions of the 3d incompressible euler equations,
arxiv:1310.0497.
[34] A. J. Majda and A. L. Bertozzi. Vorticity and incompressible flow, volume 27 of Cambridge Texts in
Applied Mathematics. Cambridge University Press, Cambridge, 2002.
[35] J. Neustupa and M. Pokorn ´y. An interior regularity criterion for an axially symmetric suitable weak solu-
tion to the Navier-Stokes equations. J. Math. Fluid Mech., 2:381–399, (2000).
[36] J. Neustupa and M. Pokorn´y. Axisymmetric flow of Navier-Stokes fluid in the whole space with non-zero
angular velocity component. In Proceedings of Partial Dierential Equations and Applications (Olomouc,
1999), volume 126, pages 469–481, (2001).
[37] G. Seregin. Local regularity of suitable weak solutions to the Navier-Stokes equations near the boundary.
J. Math. Fluid Mech., 4:1–29, (2002).
[38] G. Seregin. Lecture notes on regularity theory for the Navier-Stokes equations. World Scientific Publishing
Co. Pte. Ltd., Hackensack, NJ, 2015.
[39] G. Seregin and V. ˇ
Sver´ak. On type I singularities of the local axi-symmetric solutions of the Navier-Stokes
equations. Comm. Partial Dierential Equations, 34:171–201, (2009).
[40] G. Seregin and V. ˇ
Sver´ak. On global weak solutions to the Cauchy problem for the Navier-Stokes equations
with large L3-initial data. Nonlinear Anal., 154:269–296, (2017).
[41] E. M. Stein. Singular integrals and dierentiability properties of functions. Princeton Mathematical Series,
No. 30. Princeton University Press, Princeton, N.J., 1970.
[42] M. R. Ukhovskii and V. I. Iudovich. Axially symmetric flows of ideal and viscous fluids filling the whole
space. J. Appl. Math. Mech., 32:52–61, (1968).
[43] W. M. Zaja¸czkowski. Global existence of axially symmetric solutions to Navier-Stokes equations with
large angular component of velocity. Colloq. Math., 100:243–263, (2004).
[44] W. M. Zaja¸czkowski. Global axially symmetric solutions with large swirl to the Navier-Stokes equations.
Topol. Methods Nonlinear Anal., 29:295–331, (2007).
(K. Abe) Departmen t of Math ematics, Graduate School of Sc ience, Osaka City University, Sugimoto 3-3-
138, Sumiyoshi-ku Osaka 558-8585, Japan
E-mail address:kabe@sci.osaka-cu.ac.jp
(G. Seregin) Mathematical Institute, Oxford University, Oxford, 24-29 StGiles OX1 3LB UK and
Voronezh State University, Voronezh, Russia
E-mail address:seregin@maths.ox.ac.uk
... Based on this discovery, the authors in [44,45] proved the global existence and uniqueness of solutions in a bounded cylindrical domain with the axis of symmetry being removed for sufficiently smooth initial data. Recently, in the exterior of a cylinder, Abe and Seregin [2] constructed the unique global mild solution for axisymmetric initial data ...
... where u ν 0θ = u ν θ (x, 0) and a special Navier-slip boundary condition (α = 0), u ν · n = 0, D (u ν ) n · τ = 0, on ∂Ω, (1.6) which is called 'perfect slip' Navier boundary condition. In [2], the authors first established the global L ∞ estimate of the quantity of swirl velocity ...
... The first purpose of this paper is to check that (i) : adding the forcing term f; (ii) : imposing general Navier-slip boundary condition (1.2), if one can reduce the regularity condition on the initial data stated in (1.5), so that the unique global solution still exists. As will be indicated in this paper, we can indeed do it, without destroying the existence and uniqueness of solutions, but the price we need to pay is that the solution does not have as high regularity as that in [2]. ...
Article
Full-text available
In this paper, we study the initial boundary value problem and vanishing viscosity limit for incompressible axisymmetric Navier–Stokes equations with swirls in the exterior of a cylinder under Navier-slip boundary condition. In the first part, we prove the existence of a unique global solution with the axisymmetric initial data u0ν∈Lσ2(Ω) and axisymmetric force f∈L2([0,T];L2(Ω)) . This result improves the initial regularity condition on the global well-posedness result obtained by K Abe and G Seregin (2020 Proc. R. Soc. Edinburgh A 150 1671–98) and extends their boundary condition. In the second part, we make the first attempt to investigate the inviscid limit of unforced viscous axisymmetric flows with swirls and prove that the viscous axisymmetric flows with swirls converge to inviscid axisymmetric flows without swirls under the condition ‖ru0θν‖L2(Ω)=O(ν) . Some new uniform estimates, independent of the viscosity, are obtained here. The second result can be thought as a follow-up work to the previous work by K Abe (2020 J. Math. Pures Appl. 137 1–32), where the inviscid limit for the same equations without swirls in an infinite cylinder was studied.
... Symmetry results concerning the Navier-Stokes equations have been the subject of many mathematical studies, due to their connection with regularity issues [1,62,63] and the solvability of the renowned Leray-flux problem [42,43,44], among others. Whenever the (smooth) obstacle K and the external force f are axisymmetric, in Section 3.3 we address the question of existence of axisymmetric solutions to (1.3), that is, solutions displaying rotational symmetry with respect to the z-axis. ...
... Now, given any scalar function φ ∈ C ∞ 0 (Ω), an integration by parts and the divergence-free condition in (3.65) 1 imply that Ω u n · ∇φ = Ωn u n · ∇φ = ∂Ωn φ(u n · ν) = 0 ∀n ∈ N , since u n vanishes on ∂K n and so does φ on ∂M. Then, along the subsequence (3.90), the weak convergence in (3.90) 1 yields ...
... It can be easily verified that R z (φ) −1 = R z (φ) , and therefore det(R z (φ)) = 1, for every φ ∈ [0, 2π]. In order to determine the existence of generalized solutions to problem (3.13) displaying rotational symmetry with respect to the z-axis, the following definition is given (see [1]): ...
Article
Full-text available
The steady motion of a viscous incompressible fluid in an obstructed finite pipe is modeled through the Navier-Stokes equations with mixed boundary conditions involving the Bernoulli pressure and the tangential velocity on the inlet and outlet of the tube, while a transversal flux rate F is prescribed along the pipe. Existence of a weak solution to such Navier-Stokes system is proved without any restriction on the data by means of the Leray-Schauder Principle, in which the required a priori estimate is obtained by a contradiction argument based on Bernoulli's law. Through variational techniques and with the use of an exact flux carrier, an explicit upper bound on F (in terms of the viscosity, diameter and length of the tube) ensuring the uniqueness of such weak solution is given. This upper bound is shown to converge to zero at a given rate as the length of the pipe goes to infinity. In an axially symmetric framework, we also prove the existence of a weak solution displaying rotational symmetry.
... It is clear that C γ > (γ + 1 2 ) 2 for γ = − 1 2 , so we see the solenoidal and axisymmetric constraint improves the best constant of the Hardy-Leray inequality. However, since the sole assumption of axisymmetry does not change the optimality of the constant (γ + 1 2 ) 2 in (1.1) N =3 , it is expected that the condition of axisymmetry in Theorem 1.1 can be relaxed. Indeed, we shall show in this paper that Theorem 1.1 does hold by imposing only one component of u to be axisymmetric. ...
... where the last term e ϕ u ϕ is called the swirl part of u, which we abbreviate as u ϕ = e ϕ u ϕ . Also, the scalar function u ϕ is called the swirl component of u. 1 A vector field u is called axisymmetric if its three components u ρ , u θ and u ϕ are independent of ϕ. Now, let us assume that u ∈ D γ (R 3 ) 3 , and that its swirl part u ϕ is axisymmetric, i.e., u ϕ is independent of ϕ. ...
... In the context of fluid mechanics, the effect of the swirl component of a velocity vector field is well-studied from various view points; see for example, [11], [15], [1], [9]. [10], [13], [16], to name a few. ...
... The result appears to have two implications. First, it seems to be the first case where the regularity problem for finite energy solutions is solved beyond dimension 2. In the literature ASNS are sometimes considered as 2 1 2 dimensional equations. The second pertains recent interesting non-uniqueness results by Buckmaster and Vicol [5] on some very weak (infinite energy) solutions of the Navier Stokes equations using the technique of convex integration of De Lellis and L. Szekelyhidi [14]. ...
... The interested reader is referred to the relatively recent papers and their references therein: [39] by Xiao and Xin, [11] by G.Q. Chen and Z.M. Qian and [29] by Masmoudi and Rousset. See also the paper [1] by Abe and Seregin on ASNS in the exterior of a cylinder containing the x 3 axis, subject to the Navier slip condition. From (2.2) and Korn's inequality, one can quickly derive a L 2 t L 2 x bound for |∇v|. ...
Preprint
A domain in $\mathbb{R}^3$ that touches the $x_3$ axis at one point is found with the following property. For any initial value in a $C^2$ class, the axially symmetric Navier Stokes equations with Navier slip boundary condition has a finite energy solution that stays bounded for any given time.
... This implies that the singularity of axisymmetric solutions can only happen at the axis. When the domain is the exterior of a cylinder, the global existence of unique axisymmetric strong solution was proved in [33] and [1] when the no slip and Navier boundary conditions were supplemented, respectively. The crucial points for the analysis in [33] and [1] are an interpolation inequality and the maximum principle (1.6), respectively. ...
... When the domain is the exterior of a cylinder, the global existence of unique axisymmetric strong solution was proved in [33] and [1] when the no slip and Navier boundary conditions were supplemented, respectively. The crucial points for the analysis in [33] and [1] are an interpolation inequality and the maximum principle (1.6), respectively. The axisymmetric solutions for inhomogeneous Navier-Stokes system without swirls were studied in [2] and references therein. ...
Preprint
In this paper, the global strong axisymmetric solutions for the inhomogeneous incompressible Navier-Stokes system are established in the exterior of a cylinder subject to the Dirichlet boundary conditions. Moreover, the vacuum is allowed in these solutions. One of the key ingredients of the analysis is to obtain the ${L^{2}(s,T;L^{\infty}(\Omega))}$ bound for the velocity field, where the axisymmetry of the solutions plays an important role.
... 162]. In the case when the slip boundary condition is prescribed, the global unique solution was recently proved in [1] for axisymmetric data for u 0 \in L 2 \cap L 3 and ru \theta 0 \in L \infty by an a priori estimate (1.6) for vorticity. It is worth noting that the stability of axisymmetric solutions in a domain away from the axis under three-dimensional perturbations was studied in [7]. ...
... Since the vector field v can be written as v = v r (r, ϑ, z)e r (ϑ) + v ϑ (r, ϑ, z)e ϑ (ϑ) + v z (r, ϑ, z)e z , where we have set e ϑ := (− sin ϑ, cos ϑ, 0) with ϑ ∈ [0, 2π]. By fundamental calculations using the cylindrical coordinate, we see that v satisfies ∇ × v × n Σt = 0 and v, n Σt = 0 on Σ t if and only if (v r , v ϑ , v z ) satisfies v r = 0, ∂ r v ϑ − v ϑ = 0, ∂ r v z = 0 on Σ t , see, e.g., [1,Prop. 2.5]. ...
Preprint
The aim of this paper is to propose a new formulation of free boundary problems of the Navier--Stokes equations in the three-dimensional Euclidean space with moving contact line in terms of classical balance laws and boundary conditions. The new system does not require any additional boundary conditions on the moving contact line, where a contact angle between a free interface and a rigid surface is oscillating in time. It should be emphasized that the total available energy of the system is conserved along with smooth solutions and the negative total available energy is a strict Lyapunov functional if the velocity field and the free interface are axisymmetric. Furthermore, we show local well-posedness of the formulated system provided that the initial data are axisymmetric. Of crucial importance for the analysis is the property of maximal $L^p - L^q$-regularity for the corresponding linearized problem.
Article
A domain in R3 that touches the x3 axis at one point is found with the following property. For any initial value in a C2 class, the axially symmetric Navier Stokes equations with Navier slip boundary condition have a finite energy solution that stays bounded for any given time, i.e. no finite time blow up of the fluid velocity occurs. The result seems to be the first case where the Navier-Stokes regularity problem is solved beyond dimension 2.
Article
We construct global weak solutions of the Euler equations in an infinite cylinder Π={x∈R3|xh=(x1,x2),r=|xh|<1} for axisymmetric initial data without swirl when initial vorticity ω0=ω0θeθ satisfies ω0θ/r∈Lq for q∈[3/2,3). The solutions constructed are Hölder continuous for spatial variables in Π‾ if in addition that ω0θ/r∈Ls for s∈(3,∞) and unique if s=∞. The proof is by a vanishing viscosity method. We show that the Navier-Stokes equations subject to the Neumann boundary condition is globally well-posed for axisymmetric data without swirl in Lp for all p∈[3,∞). It is also shown that the energy dissipation tends to zero if ω0θ/r∈Lq for q∈[3/2,2], and Navier-Stokes flows converge to Euler flow in L2 locally uniformly for t∈[0,∞) if additionally ω0θ/r∈L∞. The L2-convergence in particular implies the energy equality for weak solutions.
Article
In connection with the recent proposal for possible singularity formation at the boundary for solutions of three-dimensional axisymmetric incompressible Euler's equations (Luo and Hou, Proc. Natl. Acad. Sci. USA (2014)), we study models for the dynamics at the boundary and show that they exhibit a finite-time blowup from smooth data.
Book
The lecture notes in this book are based on the TCC (Taught Course Centre for graduates) course given by the author in Trinity Terms of 2009-2011 at the Mathematical Institute of Oxford University. It contains more or less an elementary introduction to the mathematical theory of the Navier-Stokes equations as well as the modern regularity theory for them. The latter is developed by means of the classical PDE’s theory in the style that is quite typical for St Petersburg’s mathematical school of the Navier-Stokes equations. The global unique solvability (well-posedness) of initial boundary value problems for the Navier-Stokes equations is in fact one of the seven Millennium problems stated by the Clay Mathematical Institute in 2000. It has not been solved yet. However, a deep connection between regularity and well-posedness is known and can be used to attack the above challenging problem. This type of approach is not very well presented in the modern books on the mathematical theory of the Navier-Stokes equations. Together with introduction chapters, the lecture notes will be a self-contained account on the topic from the very basic stuff to the state-of-art in the field. © 2015 by World Scientific Publishing Co. Pte. Ltd. All rights reserved.
Article
Consider the Stokes resolvent system in general unbounded domains Ω Rn, n ≥ 2, with boundary of uniform class C3, and Navier slip boundary condition. The main result is the resolvent estimate in function spaces of the type Lq defined as Lq ∩ L2 when q ≥ 2, but as Lq + L2 when 1 < q < 2, adapted to the unboundedness of the domain. As a consequence, we get that the Stokes operator generates an analytic semigroup on a solenoidal subspace Lqσ(Ω) of Lq(Ω).
Article
The existence of weak solutions for the Navier-Stokes equations for the infinite cylinder with initial data in Lp is considered in this paper.We study the case of initial data in Lp(Rn), 2 < p < n, and n = 3, 4. An existence theorem is proved covering these important cases and therefore, the "gap" between the Hopf-Leray theory (p = 2) and that of Fabes-Jones-Riviere (p > n) is bridged. The existence theorem gives a new method of constructing global solutions. The cases p = n are treated at the end of the paper.
Article
The aim of the note is to discuss different definitions of solutions to the Cauchy problem for the Navier-Stokes equations with the initial data belonging to the Lebesgue space $L_3(\mathbb R^3)$
Article
Long time existence of axially symmetric solutions to the Navier-Stokes equations in a bounded cylinder and with boundary slip conditions is proved. The axially symmetric solutions with nonvanishing azimuthal component of velocity (swirl) are examined. The solutions are such that swirl is small in a neighbourhood close to the axis of symmetry but it is large in some positive distance from it. There is a great difference between the proofs of global axially symmetric solutions with vanishing and nonvanishing swirl. In the first case global estimate follows at once but in the second case we need a lot of considerations in weighted spaces to show it.
Article
It is known that the Stokes operator is not well-defined in L q -spaces for certain unbounded smooth domains unless q = 2. In this paper, we generalize a new approach to the Stokes resolvent problem and to maximal regularity in general unbounded smooth domains from the three-dimensional case, see [7], to the n-dimensional one, n ≥ 2, replacing the space L q , 1 2, satisfies the classical resolvent estimate, generates an analytic semigroup and has maximal regularity.