Vous êtes sur la page 1sur 29

Ann. I. H.

Poincar AN 28 (2011) 159187


www.elsevier.com/locate/anihpc
An alternative approach to regularity for the NavierStokes
equations in critical spaces
Carlos E. Kenig
1
, Gabriel S. Koch

Department of Mathematics, University of Chicago, Chicago, IL 60637, USA


Received 20 August 2009; accepted 14 October 2010
Available online 7 December 2010
Abstract
In this paper we present an alternative viewpoint on recent studies of regularity of solutions to the NavierStokes equations in
critical spaces. In particular, we prove that mild solutions which remain bounded in the space

H
1
2
do not become singular in nite
time, a result which was proved in a more general setting by L. Escauriaza, G. Seregin and V. verk using a different approach.
We use the method of concentration-compactness + rigidity theorem using critical elements which was recently developed
by C. Kenig and F. Merle to treat critical dispersive equations. To the authors knowledge, this is the rst instance in which this
method has been applied to a parabolic equation.
2010 Elsevier Masson SAS. All rights reserved.
Rsum
Dans cet expos, nous prsentons un point de vue diffrent sur les tudes rcentes concernant la rgularit des solutions des
quations de NavierStokes dans les espaces critiques. En particulier, nous dmontrons que les solutions faibles qui restent bornes
dans lespace

H
1
2
ne deviennent pas singulires en temps ni. Ce rsultat a t dmontr dans un cas plus gnral par L. Escau-
riaza, G. Seregin et V. verk en utilisant une approche diffrente. Nous utilisons la mthode de concentration-compacit +
thorme de rigidit utilisant des lments critiques qui a t rcemment dveloppe par C. Kenig et F. Merle pour traiter les
quations dispersives critiques. la connaissance des auteurs, cest la premire fois que cette mthode est applique une quation
parabolique.
2010 Elsevier Masson SAS. All rights reserved.
0. Introduction
In recent studies, the idea of establishing the existence of so-called critical elements (or the earlier minimal
blow-up solutions) has led to signicant progress in the theory of critical dispersive and hyperbolic equations such
as the energy-critical nonlinear Schrdinger equation [2,3,9,26,47,55], mass-critical nonlinear Schrdinger equation
[3032,52,53],

H
1
2
-critical nonlinear Schrdinger equation [28], energy-critical nonlinear wave equation [27], energy-
critical and mass-critical Hartree equations [4044] and energy-critical wave maps [10,36,50,51].
*
Corresponding author.
E-mail addresses: cek@math.uchicago.edu (C.E. Kenig), koch@math.uchicago.edu (G.S. Koch).
1
Supported in part by NSF grant DMS-0456583.
0294-1449/$ see front matter 2010 Elsevier Masson SAS. All rights reserved.
doi:10.1016/j.anihpc.2010.10.004
160 C.E. Kenig, G.S. Koch / Ann. I. H. Poincar AN 28 (2011) 159187
In this paper we exhibit the generality of the method of critical elements by applying it to a parabolic system,
namely the standard
2
NavierStokes equations (NSE):
u
t
u +(u )u +p =0,
u =0. (0.1)
Typically, u is interpreted as the velocity vector eld of a uid lling a region in space, and p is the associated scalar
pressure function.
The critical spaces are those which are invariant under the natural scaling of the equation. For NSE, if u(x, t )
is a solution, then so is u

(x, t ) :=u(x,
2
t ) for any > 0. The critical spaces are of the type X, where u

X
=
u
X
. For the NavierStokes equations, one can take, for example, X =L

((0, +); L
3
(R
3
)) or the smaller space
X=L

((0, +);

H
1
2
(R
3
)). In fact, one has the chain of critical spaces given by the continuous embeddings

H
1
2
_
R
3
_
L
3
_
R
3
_


B
1+
3
p
p,
_
R
3
_
(p<)
BMO
1
_
R
3
_


B
1
,
_
R
3
_
.
These are the spaces in which the initial data of solutions in the critical settings live, and we will also refer to them as
critical spaces that is, spaces of functions on R
3
whose norms satisfy f () =f .
In the recent important paper [14], L. Escauriaza, G. Seregin and V. verk showed that any LerayHopf weak
solution which remains bounded in L
3
(R
3
) cannot develop a singularity in nite time. Their proof used a blow-up
procedure and reduction to a backwards uniqueness question for the heat equation, and was then completed using
Carleman-type inequalities and the theory of unique continuation. Here, we approach the same problem using the
method of critical elements. Although we do also use the main tools appearing in [14] to complete our proof, we
hope that it is more intuitively clear in our exposition why those particular tools are needed.
The precise statement of the main result we address in this paper is the following:
Theorem 0.1. Let u
0


H
1
2
satisfy u
0
=0 and let u =NS(u
0
) be the associated mild solution to NSE satisfying
u(0) =u
0
. Suppose that there is some A>0 such that u(t )

H
1
2 (R
3
)
A for all t >0 such that u is dened. Then u
is dened (and smooth) for all positive times.
(Theorem 0.1 of course follows from the result in [14].) We believe the methods given below will work as well if
we replace

H
1
2
by L
3
in Theorem 0.1 and hence can be used to give an alternative proof of the result in [14] in the
case of mild solutions. For technical reasons (described below) we start with the above result and plan to return to the
more general case in a future publication.
3
The mild solutions to NSE which we consider in our approach have the form
u(t ) =L(t )u
0
+
t
_
0
L(t s)f
_
u(s)
_
ds (0.2)
for some divergence-free initial datum u
0
, with the linear solution operator L and nonlinearity f given by
L(t ) =e
t
, f
_
u(s)
_
=P(u u)(s). (0.3)
Here e
t
u
0
is the convolution of u
0
with the heat kernel, and Eqs. (0.2), (0.3) comes formally from applying the
Helmholtz projection operator P to (0.1) which xes u
t
u and eliminates the termp, and then solving the result-
ing nonhomogeneous heat equation by Duhamels formula. Under sufcient regularity assumptions, mild solutions
are in fact classical solutions, and the existence of such a solution on some time interval is typically established via
the contraction-mapping principle in an appropriate function space.
We follow the method of C. Kenig and F. Merle. In a series of recent works [28,26,27], they use the method of
critical elements to approach the question of global existence and scattering (approaching a linear solution at
2
For simplicity, we have set the coefcient of kinematic viscosity =1.
3
At the time of publication this has in fact already been accomplished, see [22].
C.E. Kenig, G.S. Koch / Ann. I. H. Poincar AN 28 (2011) 159187 161
large times) for nonlinear hyperbolic and dispersive equations in critical settings. For example, for the 3D nonlinear
Schrdinger equation iu
t
+u =f (u) (NLS), they considered (0.2) with
L(t ) =e
it
, (0.4)
the free Schrdinger operator, and both
f (u) =|u|
4
u
_
X=L

_
(0, +);

H
1
_
R
3
___
(0.5)
(the focusing case), and
f (u) =|u|
2
u
_
X=L

_
(0, +);

H
1
2
_
R
3
___
(0.6)
(the defocusing case), the exponents needed for the

H
1
-critical and

H
1
2
-critical settings, respectively. Note that
in the case (0.6) of cubic nonlinearity, the equation is invariant under the same scaling as the NavierStokes equations.
In all cases, the general strategy was essentially the same, which well describe now in the

H
1
2
setting:
For any u
0


H
1
2
(R
3
), one uses xed-point arguments to assign a maximal time T

(u
0
) +such that a solution
u to (0.2) which remains in

H
1
2
for positive time exists and is unique in some scaling-invariant space X
T
for any xed
T <T

(u
0
), where X
T
denotes a space of functions dened on the spacetime region R
3
(0, T ). Dene
|||u||| := sup
t [0,T

(u
0
))
_
_
u(t )
_
_

H
1
2
.
The type of result proved in [28] (variants of which were proved in [26,27] and which we will prove here as well) is
that |||u||| <+ implies that T

(u
0
) =+ and u(t ) L(t )u
+
0


H
1
2 (R
3
)
0 as t + for some u
+
0


H
1
2
(R
3
).
In other words, u exists globally and scatters. Typically this is known to be true for |||u||| <
0
for sufciently small

0
> 0. In the case of NSE, the scattering condition is replaced by decay to zero in norm in other words, we set
u
+
0
=0. For globally dened

H
1
2
-valued solutions, such decay was proved in [20] (see also [21]).
It is worth pointing out that the

H
1
2
decay, which is proved quite easily in [20] by decomposing the initial data into
a large part in L
2
and a small part in

H
1
2
, solving the corresponding equations and employing the standard energy
arguments, is actually used heavily in our proofs, and signicantly reduces the difculty from the NLS case treated
in [28].
The general method of proof, which can be referred to as concentration-compactness + rigidity, is comprised
of the following three main steps for a proof by contradiction:
1. Existence of a critical element:
Assuming a nite maximal threshold A
c
> 0 for which |||u||| < A
c
implies global existence and scattering but
such a statement fails for any A > A
c
, there exists a solution u
c
with |||u
c
||| =A
c
, for which global existence or
scattering fails.
2. Compactness of critical elements:
Such a critical element u
c
produces a compact family that is, up to norm-invariant rescalings and translations
in space (and possibly in time), the set {u
c
(t )} is pre-compact in a critical space.
3. Rigidity:
The existence of the compact family produces a contradiction to known results.
Steps 12 are accomplished by considering a minimizing sequence of solutions {u
n
} with initial data {u
0,n
} such that
|||u
n
||| A
n
, A
n
A
c
for which global existence or scattering fails (typically quantied by u
n

X
T

=+).
Then the main tool for realizing this program is a prole decomposition associated to that sequence, which
explores the lack of compactness in the embedding

H
1
2
(R
3
) L
3
(R
3
). For example, in [28] the following decom-
position (based on [29]) was used for treatment of the NLS case: There exists a sequence {V
0,j
}

j=1


H
1/2
, with
associated linear solutions V
l
j
(x, t ) =e
it
V
0,j
, and sequences of scales
j,n
R
+
and shifts x
j,n
R
3
and t
j,n
R,
such that (after a subsequence in n)
u
0,n
(x) =
J

j=1
1

j,n
V
l
j
_
x x
j,n

j,n
,
t
j,n

2
j,n
_
+w
J
n
(x) (0.7)
162 C.E. Kenig, G.S. Koch / Ann. I. H. Poincar AN 28 (2011) 159187
where the proles in the sum are orthogonal in a certain sense due to the choices of
j,n
, x
j,n
and t
j,n
, and w
J
n
is a
small error for large J and n. (From this, a similar decomposition is then established for the time evolution of u
0,n
.)
Using this setup, it is shown that in fact the solution to NLS with initial datum V
0,j
0
is a critical element for some
j
0
N, which establishes Step 1. Compactness (Step 2) is established using the same tools. (We will discuss Step 3
momentarily.)
Our ultimate goal here is to give such a proof in the context of the NavierStokes equations, where we would con-
sider (0.2), (0.3) for u
0
L
3
(R
3
), with X
T
=C([0, T ]; L
3
(R
3
)) L
5
(R
3
(0, T )) and T

(u
0
) dened accordingly,
and
|||u||| := sup
t [0,T

(u
0
))
_
_
u(t )
_
_
L
3
(R
3
)
.
Since a prole decomposition has already been established by I. Gallagher in [19] for solutions to the NavierStokes
equations evolving from a bounded set in

H
1
2
(based on the decomposition for the initial data in [23]), we restrict
ourselves in this paper to the case of data in

H
1
2
. We expect that the result in [19] can be extended to the L
3
setting,
which would then allow for an extension of our approach to that case. We plan to return to this in a future publication.
4
We take
X
T
:=C
_
[0, T ];

H
1
2
_
R
3
__
L
2
_
(0, T );

H
3
2
_
R
3
__
and as before let
|||u||| := sup
t [0,T

(u
0
))
_
_
u(t )
_
_

H
1
2 (R
3
)
.
The prole decomposition of [19] for solutions corresponding to a bounded sequence {u
0,n
}

H
1
2
(R
3
) of divergence-
free elds takes the form (after a subsequence in n)
u
n
(x, t ) =
J

j=1
1

j,n
U
j
_
x x
j,n

j,n
,
t

2
j,n
_
+e
t
w
J
n
(x) +r
J
n
(x, t ),
where u
n
and U
j
are the solutions to the NavierStokes equations with initial data u
0,n
and V
0,j
, respectively, and
w
J
n
and r
J
n
are again small errors for large J and n. (We remark that the absence in the above decomposition of the
time shifts t
j,n
which appeared in (0.7) greatly simplies matters on a technical level; in [28], this was a signicant
consideration.) The above program is then completed in Theorems 3.1, 3.2 and 3.3 as follows:
By known local-existence and small-data results, there exists a small
0
> 0 such that |||u||| <
0
implies that the
solution exists for all time and tends to zero in the

H
1
2
norm. We thereby assume a nite critical value A
c

0
> 0
(as in Step 1) such that any solution u with |||u||| < A
c
must exist globally and decay to zero in

H
1
2
, and A
c
is the
maximum such value.
The failure of the global existence and decay property, which occurs for some solution u with |||u||| = A for any
A>A
c
, is expressed by u
E
T

(u(0))
=+, where we dene for T >0
u
E
T
=
_
sup
t (0,T )
_
_
u(t )
_
_
2

H
1
2 (R
3
)
+
_
_
D
3/2
u
_
_
2
L
2
(R
3
(0,T ))
_1
2
(so u
E
T

(u(0))
< + whenever |||u||| < A
c
and therefore also, by standard embeddings, u L
5
(R
3
(0, T )) so
such solutions are smooth by the LadyzenskajaProdiSerrin condition, see e.g. [14]).
5
In Theorems 3.1 and 3.2, we establish the existence of a solution (critical element) u
c
with initial datum u
0,c
and
T

(u
0,c
) <+ such that
|||u
c
||| = sup
t [0,T

(u
0,c
))
_
_
u
c
(t )
_
_

H
1
2
=A
c
and u
c

E
T

(u
0,c
)
=+
4
At the time of publication of this article, an L
3
(R
3
) prole decomposition has been established by the second author in [33], and the program
in L
3
(R
3
) has been completed in the collaboration [22] of the second author with I. Gallagher and F. Planchon.
5
Strictly speaking, this condition applies to LerayHopf weak solutions, but in fact the local theory gives

t u L

which implies smooth-


ness. See also [11] for higher regularity of solutions in L
5
x,t
.
C.E. Kenig, G.S. Koch / Ann. I. H. Poincar AN 28 (2011) 159187 163
and, further, that for any such solution, there exist functions x(t ) R
3
, 0 < (t ) + as t T

(u
0,c
) and s(t )
[t, T

(u
0.c
)) such that, for u =u
c
and u
0
=u
0,c
,
K :=
_
1
(t )
u
_
x x(t )
(t )
, s(t )
_
, t
_
0, T

(u
0
)
_
_
(0.8)
is pre-compact in L
3
(R
3
).
The pre-compactness in L
3
of such a K is shown to be inconsistent with T

(u
0
) < + in Theorem 3.3, by
using the backwards uniqueness results for parabolic equations established in [12,13] as well as the theory of unique
continuation for parabolic equations to show that in fact u
c
0. (In general, Step 3 requires something specic to
the particular case being studied for example the Morawetz-type estimate for NLS used in [28] as opposed to the
methods used to establish Steps 1 and 2 which are fairly general in nature.)
One interesting scenario in which such a K would be compact in L
3
is if, in (0.8), one could take (t ) =(T

t )

1
2
for some u
0
L
3
(with T

=T

(u
0
) <+), x(t ) 0 and s(t ) =t , and one imposes that K ={U} for some given
nonzero U L
3
(i.e., u(x, t ) =
1

t
U(
x

t
)). This is the case of a self-similar solution which was rst ruled
out in the important paper [46] by J. Ne cas, M. R ui cka and V. verk in the L
3
(R
3
) setting (see [54] for more general
results), which is in fact the natural setting for self-similar LerayHopf weak solutions. Theorem 3.3 can therefore be
thought of as a generalization of the result in [46]. (Such solutions are of course ruled out as well by the more recent
paper [14], but the proof here is much simpler for that purpose.)
1. Preliminaries
Well say that u is a mild solution of NSE on [t
0
, t
0
+T ] for some t
0
R and T >0 if, for some divergence-free
initial datum u
0
, u solves (in some function space) for t [t
0
, t
0
+T ] the integral equation
u(t ) =e
(t t
0
)
u
0
+
t
_
t
0
e
(t s)
P
_
u(s) u(s)
_
ds. (1.1)
We have used the following notation: For a tensor F =(F
ij
) we dene the vector F by ( F)
i
=

j

j
F
ij
, and
for vectors u and v, we dene their tensor product u v by (u v)
ij
=u
i
v
j
.
Well consider solutions in spatial dimension three (x R
3
), so u = (u
1
, u
2
, u
3
), u
i
= u
i
(x, t ), 1 i 3.
In that case, the projection operator P onto divergence-free elds is dened on a vector eld f by (Pf )
j
=
f
j
+

3
k=1
R
j
R
k
f
k
, 1 j 3, and the Riesz transform R
j
is dened on a scalar g via Fourier transforms by
(R
j
g)

() =
i
j
||
g(). (One can also formally write this as Pf = f
1

( f ).) The heat kernel e


t
is dened
by e
t
g(x) =[e
||
2
t
g()]

(x) = ((4t )
3/2
exp{| |
2
/4t } g)(x), and extended to act component-wise on vector
elds.
Formally, (1.1) comes from applying P to the classical NavierStokes equations which one can write as
u
t
u +p = (u u),
u =0 (1.2)
(since (uu) =(u )u due to the condition u =0) and solving the resulting heat equation (since P(p) =0)
by Duhamels formula.
In what follows, well set L
p
= L
p
(R
3
) and g
p
= g
L
p for any p [1, +],

H
s
=

H
s
(R
3
) = {g S

|
(D
s
g)

() =| |
s
2
g() L
2
} for any s R where S

denotes the space of tempered distributions, and for f =f (x, t )


and any Banach space X, f L
p
((a, b); X) f (t )
X
= f (, t )
X
L
p
(a, b). For any collection of Banach
spaces (X
m
)
M
m=1
and X := X
1
X
M
, well always set g
X
= (

M
m=1
g
2
X
m
)
1
2
. Similarly, for vector-valued
f =(f
1
, . . . , f
M
), we dene f
X
=(

M
m=1
f
m

2
X
)
1
2
. For easy reference, we collect and state here the denitions
of the main spaces to which we refer throughout the paper:
E
T
=L

_
(0, T );

H
1
2
_
L
2
_
(0, T );

H
3
2
_
;
E
(1/2)
T
:=C
_
[0, T );

H
1
2
_
L
2
_
(0, T );

H
3
2
_
; E
(3)
T
:=C
_
[0, T ); L
3
_
L
5
_
R
3
(0, T )
_
.
164 C.E. Kenig, G.S. Koch / Ann. I. H. Poincar AN 28 (2011) 159187
2. Local theory
In this section, we consolidate various known local existence results for the NavierStokes equations. We also
unify the various theories through known persistency results. Although the results presented in this section are well-
known to experts, it seems to us that simple, self-contained proofs are often difcult to locate, so we present them here
for the convenience of the reader and for easy future reference. Moreover, we in no way claim that the proofs given
below are optimal, but we hope they are more or less self-contained. Our main results are given in Section 3, and the
expert reader may prefer to skip directly to that section now.
The goal in what follows is to establish the existence of local solutions to (1.1) in some (spacetime) Banach
space X=X
T
of functions dened on R
3
[0, T ) for some possibly small T >0, with divergence-free initial datum
u
0
in a Banach space X. In what follows, we will let X equal L
3
or

H
1
2
. (See, e.g., [6] and [35] respectively for local
well-posedness in

B
1+
3
p
p,
and BMO
1
, and the recent ill-posedness result [4] for

B
1
,
.) Well re-write (1.1) as
x =y +B(x, x) (2.1)
and, under the assumption that y X, try to solve the equation for some x X (where X will be chosen so that
u
0
X implies e
t
u
0
X). This will be accomplished by the following abstract lemma, using the contraction mapping
principle:
Lemma 2.1. Let X be a Banach space with norm x =x
X
, and let B : XXX be a continuous bilinear form
such that there exists =
X
>0 so that
_
_
B(x, y)
_
_
xy (2.2)
for all x and y in X. Then for any xed y Xsuch that y <1/(4), Eq. (2.1) has a unique
6
solution x Xsatisfying
x R, with
R :=
1

1 4y
2
>0. (2.3)
Proof. Let F(x) =y +B(x, x). Using (2.2) and the triangle inequality, one can verify directly that F maps B
R
:=
{x X| x R} into itself. Moreover, F is a contraction on B
R
as follows: Suppose x, x

B
R
. Then
_
_
F(x) F
_
x

__
_
=
_
_
B(x, x) B
_
x

, x

__
_
=
_
_
B
_
x x

, x
_
+B
_
x

, x x

__
_

_
_
x x

_
_
x +
_
_
x

_
_
_
_
x x

_
_
2R
_
_
x x

_
_
,
and clearly 2R < 1 by (2.3). Hence the contraction mapping principle guarantees the existence of a unique xed-
point x B
R
of the mapping F satisfying F( x) = x, which proves the lemma. 2
2.1. Local theory
7
in

H
1
2
Suppose u
0


H
1
2
, and let E
T
=L

((0, T );

H
1
2
) L
2
((0, T );

H
3
2
), with norm
f
E
T
=
_
f
2
L

((0,T );

H
1
2 )
+f
2
L
2
((0,T );

H
3
2 )
_ 1
2
. (2.4)
Note that E
T
F
T
:=L
4
((0, T );

H
1
), since Hlders inequality gives
f
F
T
f
1
2
L

((0,T );

H
1
2 )
f
1
2
L
2
((0,T );

H
3
2 )
. (2.5)
6
In fact, the uniqueness can be improved to the larger ball of radius
1
2
, see e.g. [7], formula (122).
7
This version of the local theory for initial data in

H
1
2
can be found in [38]. For other versions, see for example the classical paper [16] and a
more modern exposition in [6].
C.E. Kenig, G.S. Koch / Ann. I. H. Poincar AN 28 (2011) 159187 165
By the well-known (see, e.g. [8], p. 120) inequality h
1
h
2


H
1
2
h
1


H
1 h
2


H
1 , we have
T
_
0
f g
2

H
1
2
ds
T
_
0
f
2

H
1
g
2

H
1
ds
_
T
_
0
f
4

H
1
_1
2
_
T
_
0
g
4

H
1
ds
_1
2
and hence
f g
L
2
((0,T );

H
1
2 )
f
F
T
g
F
T
. (2.6)
Denoting
B(f, g)(t ) :=
t
_
0
e
(t s)
P
_
f (s) g(s)
_
ds,
we can write
D
3
2
B(f, g)(t ) =
t
_
0
e
(t s)
F(s) ds
where F(s) :=P ()
1
D
3
2
(f (s) g(s)). Now by the maximal regularity theorem for e
t
(see, e.g., [38], The-
orem 7.3), we have
_
_
D
3
2
B(f, g)
_
_
L
2
(R
3
(0,T ))
F
L
2
(R
3
(0,T ))
,
and so since F D
1
2
(f g), (2.6) gives
_
_
B(f, g)
_
_
L
2
((0,T );

H
3
2 )
f
F
T
g
F
T
. (2.7)
Lets recall the following lemma (see [38], Lemma 14.1):
Lemma 2.2. Let T (0, +] and 1 j 3. If h L
2
(R
3
(0, T )), then
_
t
0
e
(t s)

j
hds C
b
([0, T ); L
2
).
(C
b
indicates bounded continuous functions.) For f, g F
T
, (2.6) shows that D
1
2
(f g) L
2
(R
3
(0, T )), so
Lemma 2.2 gives D
1
2
B(f, g) C([0, T ); L
2
) and hence B(f, g) C([0, T );

H
1
2
). Moreover, the proof of Lemma 2.2
also gives the estimate
_
_
B(f, g)
_
_
L

((0,T );

H
1
2 )
f g
L
2
((0,T );

H
1
2 )
,
which, together with (2.6) gives
_
_
B(f, g)
_
_
L

((0,T );

H
1
2 )
f
F
T
g
F
T
. (2.8)
Using (2.5), (2.7) and (2.8), we now conclude that
_
_
B(f, g)
_
_
F
T
f
F
T
g
F
T
(2.9)
for some > 0. (We remark that is independent of T .) By the standard L
2
energy estimates for the heat equation,
u
0


H
1
2
implies that U E
T
F
T
, where U(t ) :=e
t
u
0
. Therefore we can take T small enough that U
F
T
<
1
4
(or, if u
0


H
1
2
is small enough one may take T =+, giving a small data global existence result), and hence by
Lemma 2.1 with X = F
T
, there exists a unique small mild solution u F
T
of NSE on [0, T ). Note that (2.7) and
(2.8) show that B(u, u) E
T
as well, so that more specically u =U +B(u, u) E
T
. Moreover, u C([0, T );

H
1
2
)
by Lemma 2.2 and the standard theory for the heat equation (see, e.g., [37]).
166 C.E. Kenig, G.S. Koch / Ann. I. H. Poincar AN 28 (2011) 159187
2.2. Local theory in

H
1
2
L

The above result can be rened to give a solution which not only remains in

H
1
2
, but belongs to L

as well for
t >0. This will be shown by considering the spaces
E

T
:=E
T

_
g

t g(x, t ) L

_
R
3
(0, T )
__
,
F

T
:=F
T

_
g

t g(x, t ) L

_
R
3
(0, T )
__
.
We will show that there exists some

>0 such that


_
_
B(f, g)
_
_
F

f
F

T
g
F

T
, (2.10)
and hence there exists a unique small solution u F

T
by Lemma 2.1 so long as U(t ) =e
t
u
0
satises
U
F

T
<
1
4

. (2.11)
Note that Youngs inequality gives
_
_
e
t
u
0
_
_

t
1/2
u
0

3
, (2.12)
hence u
0


H
1
2
L
3
implies that indeed U F

T
. Moreover, as before, the resulting solution in the case of (2.11)
will belong more specically to E

T
.
We claim now that
lim
t 0
_
_

t e
t
u
0
_
_

=0 u
0
L
3
, (2.13)
which will show that (2.11) will hold for any u
0


H
1
2
for T sufciently small. To prove (2.13), for any >0, take R
and M large enough that u
0
(1
M,R
)
3
< /2, where
M,R
(x) =1 for x {|x| < R} {x | |u
0
(x)| M} and 0
otherwise. Then by Youngs inequality we have
_
_

t e
t
u
0
_
_

_
_

t e
t
_
u
0

M,R
__
_

+
_
_

t e
t
_
u
0
_
1
M,R
___
_

t
_
_
e
t
_
_
1
M +

2
<
for small enough t > 0. The bilinear estimate (2.10) is a consequence of the continuous embedding

H
1
2
L
3
, esti-
mates (2.8) and (2.9) and the following claim:
Claim 2.3.
sup
t (0,T )

t
_
_
B(f, g)(t )
_
_

_
_
B(f, g)
_
_
L

((0,T );L
3
)
+ sup
t (0,T )

t
_
_
f (t )
_
_

sup
t (0,T )

t
_
_
g(t )
_
_

.
Proof. Well need the following facts:
(i)
_
_
e
t
u
0
_
_

1
2
u
0

3
,
(ii) e
t
P h =
1
t
2
H
_

t
_
h, where

H(y)

_
1 +|y|
_
4
. (2.14)
(i) is just Youngs inequality, and (ii) can be found for example in [38], Proposition 11.1 on The Oseen Kernel. (See
also [48], translated in [1].) Now write
B(f, g) =e
(t /2)
B(f, g)(t /2) +
t
_
t /2
e
(t s)
P (f g)(s) ds.
C.E. Kenig, G.S. Koch / Ann. I. H. Poincar AN 28 (2011) 159187 167
By (ii) and a change of variables we have

e
(t s)
P (f g)(x, s)

1
(t s)
1/2
_
R
3
H(z)(f g)(x z

t s, s) dz

f (s)

g(s)

(t s)
1/2
_
R
3
1
(1 +|z|)
4
dz =C
f (s)

g(s)

(t s)
1/2
,
and hence by (i)
_
_
B(f, g)(t )
_
_

(t /2)
1/2
_
_
B(f, g)(t /2)
_
_
3
+
t
_
t /2
(t s)
1/2
_
_
f (s)
_
_

_
_
g(s)
_
_

ds
t
1/2
_
_
B(f, g)(t /2)
_
_
3
+t
1/2
sup
s(
t
2
,t )
_
_
f (s)
_
_

sup
s(
t
2
,t )
_
_
g(s)
_
_

.
Therefore, for t (0, T ),
t
1/2
_
_
B(f, g)(t )
_
_

_
_
B(f, g)(t /2)
_
_
3
+t
1/2
sup
s(
t
2
,t )
_
_
f (s)
_
_

t
1/2
sup
s(
t
2
,t )
_
_
g(s)
_
_

sup
t (0,T )
_
_
B(f, g)(t )
_
_
3
+2 sup
t (0,T )
t
1/2
_
_
f (t )
_
_

sup
t (0,T )
t
1/2
_
_
g(t )
_
_

which proves the claim. (We remark that the constant in the claim does not depend on T .) 2
2.3. Local theory in L
3
(and L
3
L

)
Take u
0
L
3
and let D
T
= R
3
(0, T ) for some T (0, +]. We will show local existence
8
in the space
L
5
(D
T
). Note rst that U L
5
(D
T
) where U(x, t ) =e
t
u
0
(x) as follows: Since e
t
u
0
=e
t
(u
0
)
+
e
t
(u
0
)

, we
can assume u
0
0. Since
(U
t
U) U
p1
=0
for any p >1, integration by parts yields
_
R
3

U
p/2
(x, t )

2
dx +
4(p 1)
p
t
_
0
_
R
3

_
U
p/2
_
(x, t )

2
dx dt =
_
R
3

U
p/2
(x, 0)

2
dx.
Taking p =3 and using the inequality
g
L
10/3
(D
T
)
g
2/5
L

((0,T );L
2
)
g
3/5
L
2
((0,T );

H
1
)
which is due to the Hlder
9
and Sobolev inequalities, we have
_
_
U
3/2
_
_
L
10/3
(D
T
)

_
_
U
3/2
(0)
_
_
2
which is exactly
U
L
5
(D
T
)
u
0

3
. (2.15)
8
This version of the local theory for initial data in L
3
was presented in a course on mathematical uid mechanics given by Prof. Vladimr verk
at the University of Minnesota in the spring of 2006, and can be found as well in [11]. For other versions, see e.g. the classical paper [25] or the
more modern treatment in [6].
9
Note that Hlders inequality implies the following interpolation inequality: for p < r < q, =
1
r

1
q
1
p

1
q
(0, 1) and g
r
=|g|

|g|
1

r

g

p
g
1
q
.
168 C.E. Kenig, G.S. Koch / Ann. I. H. Poincar AN 28 (2011) 159187
Recall now that
B(f, g)(x, t ) =
t
_
0
K(, t s)
_
f (, s) g(, s)
_
(x) ds,
where we can write (see (2.14))
K(x, t ) =
_
1
t
2
H(
x

t
), t >0,
0, t 0
for some smooth H L
1
L

. With a slight abuse of notation for simplicity, we now make the following claim:
Claim 2.4.
_
_
_
_
_
t
_
0
K(t s) h(s) ds
_
_
_
_
_
L
5
(R
3
R)
h
L
5/2
(R
3
R)
,
whenever the right-hand side is nite.
Proof.
_
_
_
_
_
_
_
_
_
_
t
_
0
K(t s) h(s) ds
_
_
_
_
_
L
5
x
_
_
_
_
_
L
5
t
(R)

_
_
_
_
_
t
_
0
_
_
K(t s) h(s)
_
_
L
5
x
ds
_
_
_
_
_
L
5
t
(R)

_
_
_
_
_
t
_
0
_
_
K(t s)
_
_
L
5/4
x
_
_
h(s)
_
_
L
5/2
x
ds
_
_
_
_
_
L
5
t
(R)
=
_
_
_
_
_
t
_
0
(t s)
4/5
H
L
5/4
x
_
_
h(s)
_
_
L
5/2
x
ds
_
_
_
_
_
L
5
t
(0,+)

_
_
_
_
_
+
_

h(s)
L
5/2
x
|t s|
1
ds
_
_
_
_
_
L
5
t
(R)

_
_
_
_
h(s)
_
_
L
5/2
x
_
_
L
5/2
t
(R)
where =1/5 and we have used Youngs inequality
10
in the x variable with
1
5
=
1
(5/4)
+
1
(5/2)
1 in the second line,
and one-dimensional fractional integration
11
in the t variable with
1
5
=
1
(5/2)
in the last. Now since for t (0, T )
we have
t
_
0
K(t s) h(s) ds =
t
_
0
K(t s) (f g)(s) ds
for h :=
[0,T ]
f g where
[0,T ]
is the indicator function for the interval [0, T ], for any elds f, g L
5
(D
T
)
Claim 2.4 gives
_
_
B(f, g)
_
_
L
5
(D
T
)

5
f g
L
5/2
(D
T
)

5
f
L
5
(D
T
)
g
L
5
(D
T
)
(2.16)
10
f g
r
f
p
g
q
for 1 p, q, r + whenever
1
p
+
1
q
1 =
1
r
>0.
11
In dimension n, D

f
q
=C

| |
n+
f ()
q
C
p,q
f
q
for 1 <p <q <+ and 0 < <n whenever
1
q
=
1
p


n
(see [49]).
C.E. Kenig, G.S. Koch / Ann. I. H. Poincar AN 28 (2011) 159187 169
for some
5
> 0 independent of T . Hence there exists a unique solution u L
5
(D
T
) by Lemma 2.1 so long as
U
L
5
(D
T
)
<
1
4
5
which is true for small enough T due to (2.15) (or, for u
0

3
sufciently small, one may take
T =+). Moreover, estimate (7.26) of [14] implies
12
that
_
_
B(f, g)
_
_
L

((0,T );L
3
)
f
L
5
(D
T
)
g
L
5
(D
T
)
, (2.17)
the proof of which shows moreover that t B(f, g)(t )
3
is continuous. This, along with the fact that B(f, g)(t )
is weakly continuous in L
3
on (0, T ) (see [14], (7.27)) implies that B(f, g) C((0, T ); L
3
). The standard theory of
the heat equation implies now that u =U +B(u, u) C([0, T ); L
3
). 2
Remark 2.5. We can furthermore assume that

t u(x, t ) L

(D
T
) (for a possibly smaller T ) in a manner similar to
the proof of (2.10), using Claim 2.3 and estimate (2.17). For a construction of local solutions with higher regularity in
similar spaces, see [11].
At this point, we also mention the following uniqueness theorem for solutions in the class C([0, T ); L
3
) established
in [17] (see also [45] for a simplied proof):
Theorem 2.6. Let u
1
, u
2
C([0, T ); L
3
) both satisfy (1.1) for a xed u
0
L
3
. Then u
1
u
2
on [0, T ).
Note that there is no size restriction on u
0

3
.
2.4. Unication of the theories and further properties
For any u
0
L
3
, the local theory guarantees the existence of a mild solution u
(T )
E
(3)
T
on [0, T ) for some T >0
with u
(T )
(0) =u
0
, where we dene
E
(3)
T
:=C
_
[0, T ); L
3
_
L
5
_
R
3
(0, T )
_
. (2.18)
By Theorem 2.6, there can be at most one mild solution in C([0, T ); L
3
) with initial datum u
0
for a xed T >0, and
hence there can be at most one
13
such u
(T )
. Dene T

3
=T

3
(u
0
) >0 by
T

3
(u
0
) :=sup
_
T >0

! mild solution u
(T )
E
(3)
T
on [0, T ) s.t. u
(T )
(0) =u
0
_
.
Again, by the L
3
uniqueness theorem, u
(T
1
)
(0) =u
(T
2
)
(0) implies that u
(T
1
)
(t ) =u
(T
2
)
(t ) for all t [0, min{T
1
, T
2
}),
hence there exists a unique mild solution u
(3)
on [0, T

3
) such that u
(3)
(0) =u
0
and u
(3)
E
(3)
T
for any T <T

3
. Well
call T

3
the maximal time of existence in L
3
of the mild solution associated to u
0
.
Similarly, for u
0


H
1
2
, dening
E
(1/2)
T
:=C
_
[0, T );

H
1
2
_
L
2
_
(0, T );

H
3
2
_
(2.19)
and T

1/2
=T

1/2
(u
0
) >0 by
T

1/2
(u
0
) :=sup
_
T >0

! mild solution u
(T )
E
(1/2)
T
on [0, T ) s.t. u
(T )
(0) =u
0
_
,
there exists a unique mild solution u
(1/2)
on [0, T

1/2
) such that u
(1/2)
(0) =u
0
and u
(1/2)
E
(1/2)
T
for any T < T

1/2
.
Well call T

1/2
the maximal time of existence in

H
1
2
of the solution associated to u
0
.
The standard embeddings and interpolation inequalities give the estimate
g
L
5
(R
3
(0,T ))
g
3/5
L

((0,T );

H
1
2 )
g
2/5
L
2
((0,T );

H
3
2 )
,
12
Strictly speaking, the estimate was derived under the assumption that u satises the differential NSE for a suitable pressure p; however since
the mild solution is smooth, we can reduce to this case, see the proof of Theorem 3.3.
13
In fact, since u E
(3)
T
contains the additional information that u L
5
x,t
, one can derive uniqueness from the local theory without using
Theorem 2.6, but we proceed in this way for simplicity.
170 C.E. Kenig, G.S. Koch / Ann. I. H. Poincar AN 28 (2011) 159187
and hence E
(1/2)
T
E
(3)
T
for any T > 0. This implies that T

1/2
T

3
for any u
0


H
1
2
L
3
. We will now proceed to
show that in fact T

1/2
=T

3
for u
0


H
1
2
.
The following lemma will show that a solution which is continuous in

H
1
2
for some time cannot leave

H
1
2
while
remaining continuous in L
3
(or, equivalently, while remaining in L
5
x,t
), and hence that the strict inequality T

1/2
<T

3
is impossible for u
0


H
1
2
. Furthermore, Theorem 2.6 implies that u
(1/2)
=u
(3)
on [0, T

) where T

=T

1/2
=T

3
. In
what follows, for u
0
L
3
(in particular, for u
0


H
1
2
) well denote u
(3)
(=u
(1/2)
) =: NS(u
0
).
Lemma 2.7. Suppose u C([0, T ]; L
3
) is a mild solution for NSE on [0, T ]. Then
(a) u L
5
(R
3
(0, T ));
(b) If moreover u(0) =u
0


H
1
2
, then u C([0, T ];

H
1
2
) L
2
((0, T );

H
3
2
).
In order to prove the lemma, we will need the following claim:
Claim 2.8. Suppose u C([0, T ]; L
3
), and x some a >0. Then
(i) s sup
t (0,a)

t e
t
u(s)

C[0, T ] and
(ii) s e
t
u(x, s)
L
5
x,t
(R
3
(0,a))
C[0, T ].
If, moreover, u C([0, T ];

H
1
2
), then
(iii) s e
t
u(s)
L
4
t
((0,a);

H
1
)
C[0, T ].
Proof. The claim is a simple consequence of the linearity of the operator e
t
, estimates (2.14(i)) and (2.15), and
the estimate e
t
g
L
4
((0,a);

H
1
)
g

H
1
2
which follows from the L
2
energy inequality for the heat equation and the
standard embeddings. Writing any one of these as e
t
g
X
a
g
X
where u C([0, T ]; X), we have

_
_
e
t
u(s
2
)
_
_
X
a

_
_
e
t
u(s
1
)
_
_
X
a

_
_
e
t
_
u(s
2
) u(s
1
)
__
_
X
a

_
_
u(s
2
) u(s
1
)
_
_
X
which proves the claim. 2
Proof of Lemma 2.7. Note rst that we may actually assume u is a solution in C([0, T + ); L
3
) on [0, T + )
for some > 0. This is guaranteed by the local theory in L
5
x,t
for initial data (e.g., u(T )) in L
3
. By the same local
existence theory, we are guaranteed a solution u
(3)
E
(3)
T
1
=C([0, T
1
); L
3
) L
5
(R
3
(0, T
1
)) with initial datum u(0)
for some small T
1
>0. By the uniqueness guaranteed by Theorem 2.6, we know that in fact u
(3)
=u on [0, T
1
).
Set T :=sup{T

(0, T +) | u E
(3)
T

}. If T > T then we have proved (a), so suppose T T . Since u(T ) L


3
,
(2.15) shows that e
t
u(T ) L
5
(R
3
(0, +)), hence one may take some t
1
(0, ) small enough so that
_
_
e
t
u(T )
_
_
L
5
(R
3
(0,t
1
))
<

0
2
, (2.20)
where
0
> 0 is such that e
t
v
L
5
(R
3
(0,t
1
))
<
0
guarantees an E
(3)
t
1
-solution on (0, t
1
) with initial datum v by
Lemma 2.1 and (2.16). Now, by Claim 2.8(ii), there exists (0, t
1
) such that
_
_
e
t
u(T )
_
_
L
5
(R
3
(0,t
1
))
<
0
. (2.21)
Therefore by the local existence theory, we are guaranteed a solution u
(3)
C([T , T +); L
3
) L
5
(R
3
(T
, T +)), where :=t
1
>0. Again by Theorem 2.6, u
(3)
=u on (T , T +). But now clearly u L
5
(R
3

(0, T +)), which is a contradiction to the denition of T , and (a) is proved.


Lets now turn to (b). First suppose that we can show that u
0


H
1
2
implies u C([0, T ];

H
1
2
). Then by replacing
E
(3)
T
above by F
(1/2)
T
:= C([0, T );

H
1
2
) L
4
((0, T );

H
1
) and replacing Claim 2.8(ii) by Claim 2.8(iii), the method
C.E. Kenig, G.S. Koch / Ann. I. H. Poincar AN 28 (2011) 159187 171
used to prove (a) can be adjusted to show that u L
4
((0, T );

H
1
). This fact, along with standard estimates (see the
proof of the local theory in

H
1
2
), yields the conclusion in (b). It therefore remains only to show that
u C
_
[0, T ]; L
3
_
, u
0


H
1
2
u C
_
[0, T ];

H
1
2
_
. (2.22)
In order to do this, we turn to the local solutions u
(KT )
of Koch and Tataru [35] for (possibly large, but somewhat
regular) initial data in the space bmo
1
, and a persistency theorem given in [18] which states that whenever the
initial datum belongs to various spaces of regular data, including

H
1
2
, the solution with that datum remains there for
positive times.
14
To complete the proof, we rely as before on the uniqueness provided by Theorem 2.6.
We recall briey the space E
(KT )
T
in which the KochTataru solutions were constructed:
E
(KT )
T
:=
_
f

sup
t (0,T )

t
_
_
f (t )
_
_

<+
_

_
f

sup
x
0
R
3
,t (0,T )
1
t
3/2
t
_
0
_
B

t
(x
0
)

f (x, s)

2
dx ds <+
_
,
where for r > 0 we denote B
r
(x
0
) = {x | |x x
0
| < r}. Recall also the denition (see, e.g., [38]) that v bmo
1
if
and only if
15
v
bmo
1 := sup
x
0
R
3
,t (0,T )
1
t
3/2
t
_
0
_
B

t
(x
0
)

e
s
v(x)

2
dx ds <+.
By (2.12), e
t
u
0
E
(KT )
T
for any u
0
L
3
bmo
1
. We know from (2.13) that lim
t 0

t e
t
u
0

=0, and it can


moreover be shown that lim
0
u

bmo
1 =0, where v

(x) =v(x), for any u


0
L
3
. Note also that e
t
v

(x) =
(e

2
t
v)(x), and hence by a change of variables we have
sup
t (0,T )

t
_
_
e
t
u

0
_
_

= sup
t (0,
2
T )

t
_
_
e
t
u
0
_
_

which therefore tends to zero as well as 0 by the preceding statement.


The local result in [35] is that for any T > 0 there exists some small
T
> 0 such that e
t
v
0

E
(KT )
T
<
T
implies
the existence of a unique small solution v
(KT )
E
(KT )
T
on [0, T ) satisfying v(0) =v
0
. By the statements above, for
any u
0
L
3
one can take
1
> 0 small enough so that sup
t (0,1)

t e
t
u

1
0

, u

1
0

bmo
1 <
1
. Letting v
0
= u

1
0
we obtain a solution v on [0, 1) with initial datum u

1
0
. Due to the natural scaling of the equation, u
(KT )
(x, t ) :=
1

1
v(
x

1
,
t

2
1
) is a solution on [0, (
1
)
2
) satisfying u
(KT )
(0) =u
0
.
We return now to the proof of (2.22). Suppose u
0


H
1
2
. By the local existence theory, there exists a solution
u
(1/2)
E
(1/2)
T
1
on [0, T
1
) with u
(1/2)
(0) = u
0
for some small T
1
> 0. By uniqueness (Theorem 2.6), u
(1/2)
= u on
[0, T
1
) hence, in particular, u C([0, T
1
);

H
1
2
).
Dene T
1
:=max{T
1
(0, T +) | u C([0, T
1
);

H
1
2
)}, and suppose T
1
T .
Since u(T
1
) L
3
bmo
1
, there exists some
1
>0 small enough that
sup
t (0,1)

t
_
_
e
t
_
u(T
1
)
_

1
_
_

,
_
_
_
u(T
1
)
_

1
_
_
bmo
1
<
1
/2,
guaranteeing a solution u
(KT )
on [T
1
, T
1
+ t
1
) satisfying u
(KT )
(T
1
) = u(T
1
), where t
1
= (
1
)
2
. Due to the em-
bedding u
bmo
1 u
3
, we have u C([0, T ); bmo
1
), and in particular there exists some

T
1
< T
1
such that for
:=T
1


T
1
small enough we have
14
We remark that it is possible to complete the proof within the framework of the L
3
-theory (see, e.g., the appendix in [21]), without resort-
ing to the stronger tools in [35] and [18] which we originally used here in the interest of expediency. However the following proof may be of
bibliographical interest.
15
Strictly speaking, one should denote this space by bmo
1
T
, but we consider some xed T >0.
172 C.E. Kenig, G.S. Koch / Ann. I. H. Poincar AN 28 (2011) 159187
_
_
_
u(

T
1
)
_

_
u(T
1
)
_

_
_
bmo
1

_
_
_
u(

T
1
)
_

_
u(T
1
)
_

_
_
3
=
_
_
u(

T
1
) u(T
1
)
_
_
3
<
1
/2,
so that [u(

T
1
)]

bmo
1 <
1
. Also, for small enough, we have by (i) of Claim 2.8
sup
t (0,1)

t
_
_
e
t
_
u(

T
1
)
_

1
_
_

= sup
t (0,t
1
)

t
_
_
e
t
u(

T
1
)
_
_

(i)
sup
t (0,t
1
)

t
_
_
e
t
u(T
1
)
_
_

+
1
/2 = sup
t (0,1)

t
_
_
e
t
_
u(T
1
)
_

1
_
_

+
1
/2 <
1
.
Hence we are also guaranteed a solution u
(KT )
on [

T
1
,

T
1
+ t
1
) satisfying u
(KT )
(

T
1
) = u(

T
1
), where t
1
= (
1
)
2
.
Taking < t
1
, we now have a KochTataru solution u
(KT )
on [

T
1
, T
1
+ ) satisfying u
(KT )
(

T
1
) = u(

T
1
), where
=t
1
>0.
Since

T
1
< T
1
, we have u(

T
1
)

H
1
2
by our denition of T
1
. Now, the persistency principle of [18] states that a
KochTataru solution obtained by the xed-point argument on an interval with initial datum in

B
s,q
p
for 1 p, q
+, s >1 will remain in

B
s,q
p
on that interval and is moreover continuous in that norm.
16
Since
17

B
1
2
,2
2
=

F
1
2
,2
2
=

H
1
2
, we see that u
(KT )
C([

T
1
, T
1
+);

H
1
2
). Theorem 2.6 then implies that u
(KT )
=u on [

T
1
, T
1
+), and hence
u C([0, T
1
+);

H
1
2
). This, however, contradicts our assumption on T
1
, hence necessarily T
1
> T , and (2.22) is
proved, completing the proof of Lemma 2.7. 2
Finally, we state here for convenience some known a priori results regarding globally-dened mild solutions to
NSE evolving from possibly large data:
Theorem 2.9 (Decay). Let u C([0, ); X) be a global-in-time solution to (1.1) for some divergence-free u
0
X,
where X is either

H
1
2
(R
3
) or L
3
(R
3
). Then lim
t +
u(t )
X
=0.
This is proved for X =

H
1
2
(R
3
) in [20], and for X = L
3
(R
3
) in [21]. The following result is classical (see, e.g.,
[19] for a proof):
Theorem 2.10 (Energy inequality). Let u
0


H
1
2
(R
3
) be any divergence-free eld, and let u =NS(u
0
). There exists a
small universal constant
0
>0 such that u
0

L
3
(R
3
)

0
implies that T

(u
0
) =+and for any t 0 the following
energy inequality holds:
_
_
u(t )
_
_
2

H
1
2 (R
3
)
+
t
_
0
_
_
u(s)
_
_
2

H
1
2 (R
3
)
ds u
0

H
1
2 (R
3
)
.
Together with Theorem 2.9, we now have the following:
Corollary 2.11. Let u
0


H
1
2
(R
3
), u
0
= 0 and u = NS(u
0
), and suppose T

(u
0
) = +. Then for some large
T
0
>0 we have, for all t T
0
,
_
_
u(t )
_
_
2

H
1
2 (R
3
)
+
t
_
T
0
_
_
u(s)
_
_
2

H
1
2 (R
3
)
ds
_
_
u(T
0
)
_
_
2

H
1
2 (R
3
)
.
Finally, we remark that the following fact follows easily from the local theory and Corollary 2.11 (recall E
T
=
L

((0, T );

H
1
2
) L
2
((0, T );

H
3
2
)):
16
The theorem in [18] includes persistence in inhomogeneous spaces and Lebesgue spaces as well, including L

.
17
These are the standard Besov and TriebelLizorkin spaces; see, e.g., [24].
C.E. Kenig, G.S. Koch / Ann. I. H. Poincar AN 28 (2011) 159187 173
Theorem 2.12. For T (0, +], take (X
0
, X
T
) to be either (

H
1
2
(R
3
), E
T
) or (L
3
(R
3
), L
5
(R
3
(0, T ))). Then for
any u
0
X
0
,
T

(u
0
) <+
_
_
NS(u
0
)
_
_
X
T

(u
0
)
=+.
3. Main results
The following blow-up criterion for the NavierStokes equations was proved in [14]: If u is a weak Leray
Hopf solution of the NavierStokes equations (essentially, a distributional solution to (0.1) satisfying a natural energy
estimate) and if (0, T

) is the maximal interval on which u is smooth, then if T

<+, we must have


limsup
t T

_
_
u(t )
_
_
3
=+.
Alternatively, one can say that u
L

((0,T

);L
3
)
<+ implies that T

=+.
We would like to suggest a different approach to this statement, using the method of critical elements introduced
in [28,26,27], in the setting of mild solutions. We consider here a simpler special case of the above, namely the
statement that for any u
0


H
1
2
,
_
_
NS(u
0
)
_
_
L

((0,T

(u
0
));

H
1
2 )
<+ T

(u
0
) =+ (3.1)
where NS(u
0
)
_
T <T

(u
0
)
E
(1/2)
T
(see (2.19)) is the associated mild solution. Theorem 2.9 then implies moreover that
lim
t +
u(t )

H
1
2
= 0, so that (3.1) may be thought of as a statement of global existence and scattering (time-
asymptotic convergence to a solution of the linear equation), such as was considered in [28] for nonlinear Schrdinger
equations. We start with a setup following [28] as follows:
Suppose (3.1) is false; then there exists some maximal nite A
c
>0 (see (3.2) below) such that
u
L

((0,T

(u
0
));

H
1
2 )
<A
c
T

=+
while, for any A > A
c
, there exists some initial datum u
0,A
and solution u
A
with T

(u
0,A
) < + and
u
A

((0,T

(u
0,A
));

H
1
2 )
A. Note that the local theory implies A
c
> 0 is well-dened since T

(u
0
) = + (and
moreover u E
(1/2)
T
for T =+ as well by Corollary 2.11) for u
0


H
1
2
small enough.
Specically, we dene the critical value A
c
by
A
c
:=sup
_
A>0;
_
_
NS(u
0
)
_
_
L

((0,T

(u
0
));

H
1
2 )
AT

(u
0
) =+
_
. (3.2)
Under the assumption that A
c
< +, well prove the following two theorems, whose analogs were used to prove
statements similar to (3.1) in [28,26,27]:
Theorem 3.1. Suppose A
c
< +. Then there exists some u
0,c


H
1
2
with associated mild solution u
c
such that
T

(u
0,c
) <+ and u
c

((0,T

(u
0,c
));

H
1
2 )
=A
c
.
We call u
c
a critical element.
Theorem 3.2. Suppose A
c
<+, and let u
0


H
1
2
with associated mild solution u =NS(u
0
) satisfy T

(u
0
) <+
and u
L

((0,T

(u
0
));

H
1
2 )
= A
c
. For any {t
n
} [0, T

(u
0
)) such that t
n
T

(u
0
), there exist sequences {s
n
} with
t
n
s
n
<T

(u
0
), {x
n
} R
3
and {
n
} (0, ) with
n
+ such that a subsequence of
1

n
u(
x
n

n
, s
n
) converges
in L
3
.
This is a slightly weaker version of the compactness in L
3
of the closure of a family K, where
K :=
_
1
(t )
u
_
x x(t )
(t )
, t
_
, 0 t <T

(u
0
)
_
(3.3)
174 C.E. Kenig, G.S. Koch / Ann. I. H. Poincar AN 28 (2011) 159187
for some (t ) >0, x(t ) R
3
, as was proved in [28,26,27]. Note, however, that for any sequence t
n
T

(u
0
), setting
x(t
n
) 0 and (t
n
) 1, a convergent subsequence exists by the L
3
-continuity of u. The only remaining scenario is
therefore treated by the weaker statement in Theorem 3.2.
The idea is that A
c
<+ implies the existence of a critical element which produces a type of compactness (see
(0.8)) in the family K. If one can rule out such compactness, then one has proved (3.1). (The nonexistence of a
particular example of a fully compact family K was established in [46] and [54] regarding the so-called self-similar
solutions.) The program is therefore completed with the following ridigity theorem:
Theorem 3.3. Any u
0
L
3
(R
3
) satisfying the conditions of Theorem 3.2 must be identically zero.
Theorem 3.3 has the immediate corollary that no such element u
0
can exist, since T

(0) = + (i.e., the zero


solution is global) whereas by assumption T

(u
0
) <+. This implies, due to Theorems 3.1 and 3.2, that A
c
=+,
which proves the regularity criterion (3.1) (Theorem 0.1) as desired.
3.1. Preliminary lemmas
In order to prove Theorem 3.1 (and also Theorem 3.2), we will need the following prole decomposition which
was proved in [19]. Recall the notation NS(u
0
) for the mild solution in
_
T <T

(u
0
)
E
(1/2)
T
associated to an initial datum
u
0


H
1
2
.
Theorem 3.4 (Prole decomposition). Let {u
0,n
} be a bounded sequence of divergence-free vector elds in

H
1
2
. Then
there exist {x
j,n
} R
3
and {
j,n
} (0, +) which are orthogonal in the sense that, for j, j

N, j =j

,
either lim
n+

j,n

,n
+

j

,n

j,n
=+ or

j,n

,n
1 and lim
n+
|x
j,n
x
j

,n
|

j,n
=+,
and a sequence of divergence-free
18
vector elds {V
0,j
}

H
1
2
with T

(V
0,j
) <+ for an at most nite number of
j N such that the following is true: after possibly taking a subsequence in n,
u
0,n
(x) =
J

j=1
1

j,n
V
0,j
_
x x
j,n

j,n
_
+w
J
n
(x) (3.4)
and
u
n
(x, t ) =
J

j=1
1

j,n
U
j
_
x x
j,n

j,n
,
t

2
j,n
_
+w
l,J
n
(x, t ) +r
J
n
(x, t ) (3.5)
for any J N, for x R
3
and t (0, T
n
) where T
n
:= min{
2
j,n
T
j
| T

(V
0,j
) < +} (or T
n
+ if T

(V
0,j
) =
+ for all j) for any xed numbers T
j
< T

(V
0,j
), u
n
=NS(u
0,n
), U
j
=NS(V
0,j
) and w
l,J
n
(t ) =e
t
w
J
n
for some
w
J
n


H
1
2
and r
J
n
E
T
n
which are small remainders in the sense that
lim
J+
_
limsup
n+
_
_
w
J
n
_
_
3
_
=0, (3.6)
and
lim
J+
_
limsup
n+
_
_
r
J
n
_
_
E
T
n
_
=0, (3.7)
and the decomposition satises the following orthogonality and stability property:
u
0,n

H
1
2
=
J

j=1
V
0,j

H
1
2
+
_
_
w
J
n
_
_
2

H
1
2
+
J
n
, lim
n+

J
n
=0 J N. (3.8)
18
Note that this property of each V
0,j
is a consequence of (3.4) and the orthogonality of the scales-cores. Indeed, u
0,n
V
0,j
where u
0,n
(x) =

j,n
u
0,n
(
j,n
x +x
j,n
) is divergence-free.
C.E. Kenig, G.S. Koch / Ann. I. H. Poincar AN 28 (2011) 159187 175
Recall that E
T
is dened for T > 0 by (2.4), and note that T
n
is simply some time such that for t (0, T
n
),
(3.5) avoids the nite blow-up times of any of the V
0,j
s due to the natural scaling of the equation: if v is a solution
on (0, T ) then v

(x, t ) :=v(x,
2
t ) is a solution on (0,
T

2
) for any >0.
Well also need the following fact regarding the proles U
j
which appear in Theorem 3.4:
Claim 3.5. If T

(V
0,j
) =+ for all j 1, then there exists some n
0
1 such that T

(u
0,n
0
) =+.
This is in fact implicit in the statement of Theorem 3.4. The proof can be found in [19], for example if one follows
the proof of Theorem 2, part (iii) in Section 3.2.2 of that paper. One sees that, under the assumptions of Claim 3.5
above, the remainder r
J
n
belongs to E
T

(u
0,n
)
for large J and n and hence u
n
belongs to the same space due to (3.5)
and standard heat estimates. Theorem 2.12 then implies that T

(u
0,n
) =+ for such n.
Finally, we will need
19
the following backwards uniqueness type of lemma (a stronger version of which is proved
in the last section):
Lemma 3.6. Let u(t ) be a mild solution of NSE with initial datum u(0) = u
0
L
3
. Suppose u(t
1
) = 0 for some
t
1
(0, T

(u
0
)). Then u 0. In particular, u
0
=0 and T

(u
0
) =+.
Lemma 3.6 will follow from the following backwards uniqueness theorem for systems of parabolic equations,
which was proved in [12,13] (for the more general situation of a half-space see also [14], but we do not require such
generality):
Theorem 3.7 (Backwards uniqueness). Fix any R, , M and c
0
>0. Let Q
R,
:=(R
3
\B
R
(0)) (, 0), and suppose
a vector-valued function v and its distributional derivatives satisfy v, v
t
, v,
2
v L
2
() for any bounded subset
Q
R,
, |v(x, t )| e
M|x|
2
for all (x, t ) Q
R,
, |v
t
v| c
0
(|v| + |v|) on Q
R,
and v(x, 0) = 0 for all
x R
3
\B
R
(0). Then v 0 in Q
R,
.
Proof of Lemma 3.6. We will apply Theorem 3.7 with v = =curl
x
u, the associated vorticity to the velocity eld u.
Suppose for the moment that satises the assumptions of Theorem 3.7 (in the whole space, in fact). Then u(t
1
) =0
implies (t
1
) = 0, and hence 0 in R
3
(t
1
, t
1
), for any (0, t
1
). Since mild solutions to NSE satisfy
div
x
u = 0 in the distributional sense, we see that
x
u(t ) = 0, u(t ) L
3
for each t (t
1
, t
1
) which implies that
u 0 in R
3
(t
1
, t
1
). Taking t
1
and using continuity at t =0, the theorem follows. (In the forward direction,
one uses the uniqueness of Theorem 2.6.)
The vorticity satises the equation

t
+(u ) ( )u =0
in the sense of distributions. The assumptions of Theorem 3.7 follow therefore from the fact that u L

(R
3

(, T

(u
0
) )) for any > 0, which follows from Remark 2.5 and arguments similar to those in the proof of
Lemma 2.7. The bounds on the derivatives of u (and hence of ) are then standard, see for example [34]. 2
3.2. Existence of a critical element
In this section well prove Theorem 3.1, which establishes the existence of a critical element.
Proof of Theorem 3.1. Dene A
c
by (3.2) and assume that A
c
< +. (Note that A
c
is well-dened by global
existence for small data.) By denition of A
c
, we can pick A
n
A
c
as n + and u
0,n


H
1
2
, u
0,n
=0 such
that
T

(u
0,n
) <+ and
_
_
NS(u
0,n
)
_
_
L

((0,T

(u
0,n
));

H
1
2 )
A
n
(3.9)
19
In fact, it has now been pointed out in [22] that one does not actually need to use a result such as Lemma 3.6 to prove Theorems 3.1 and 3.2.
176 C.E. Kenig, G.S. Koch / Ann. I. H. Poincar AN 28 (2011) 159187
holds for all n. One cannot prove, as one would naturally hope, that a subsequence of these u
0,n
converges to some
u
0,c
satisfying the assertion of the theorem, but we will see that something similar is true. We may assume that
u
0,n


H
1
2
A
n
2A
c
, (3.10)
and hence we may apply the prole decomposition (Theorem 3.4) to the sequence {u
0,n
}. For the remainder of this
section, V
0,j
, U
j
, u
n
etc. will refer to this particular sequence. We will complete the proof of the theorem by showing
that V
0,j
0
satises the assertion of the theoremfor some j
0
, and the prole U
j
0
=NS(V
0,j
0
) will be our critical element.
Dene T

j
:=T

(V
0,j
), and note that (3.10) and (3.8) imply that

j=1
V
0,j

H
1
2
<+, (3.11)
so that lim
j
V
0,j


H
1
2
=0 and therefore by the small data result, T

j
= + for large enough j. Corollary 2.11
then implies that U
j

E
(+)
<+ for large j as well.
Note that Theorem 2.12 and Claim 3.5 together imply that there exists some j 1 such that U
j

E
T

j
=+, and
hence we may re-order the proles in the decomposition such that for some J
1
1,
U
j

E
T

j
=+ for 1 j J
1
(3.12)
and
U
j

E
T

j
<+ for j >J
1
.
Note also that Theorem 2.12 and Corollary 2.11 imply that
T

j
<+ for 1 j J
1
and T

j
=+ for j >J
1
. (3.13)
To prove Theorem 3.1, it sufces therefore to show that
U
j
0

((0,T

j
0
);

H
1
2 )
=A
c
(3.14)
for some j
0
{1, . . . , J
1
}. This will be accomplished using the following claim, which extends the orthogonality
property (3.8):
Claim 3.8. Set T

j,k
:= T

j

1
k
and t
n
k
:= min
1jJ
1

2
j,n
T

j,k
for j, k N. Fix some k 1, and let {t
n
} R be any
sequence such that t
n
[0, t
n
k
] for all n. Then there exist subsequences n(m, k) and J(m, k) depending on k and
indexed by m such that for n =n(m, k) and J =J(m, k),
_
_
u
n
(t
n
)
_
_
2

H
1
2
=
J

j=1
_
_
U
j,n
(t
n
)
_
_
2

H
1
2
+
_
_
w
l,J
n
(t
n
)
_
_
2

H
1
2
+(k, m),
where lim
m+
(k, m) =0 for each k and we have set

U
j,n
(x, t ) :=
1

j,n
U
j
_
xx
j,n

j,n
,
t

2
j,n
_
.
Assuming Claim 3.8 momentarily, we prove (3.14) (and hence Theorem 3.1) as follows:
Let t
n
j,k
:=
2
j,n
T

j,k
, so that t
n
k
=min
1jJ
1
t
n
j,k
. For any xed k, there exists some j
k
0
{1, . . . , J
1
} such that
t
n
k
=t
n
j
k
0
,k
for innitely many n. Also, there exists some j
0
{1, . . . , J
1
} such that j
k
0
= j
0
for innitely many k, say for a
subsequence k

as +. For each xed , take m


,
to be a subsequence of ms indexed by +
such that
t
n(m
,
,k

)
k

=t
n(m
,
,k

)
j
0
,k

(3.15)
(where n(m, k), J(m, k) are as in Claim 3.8), and m
,
+ as +.
C.E. Kenig, G.S. Koch / Ann. I. H. Poincar AN 28 (2011) 159187 177
Since 1 j
0
J
1
, we have U
j
0

E
T

j
0
=+ by (3.12). Let
A
(j
0
)
:= sup
0t T

j
0
_
_
U
j
0
(t )
_
_

H
1
2
and A
(j
0
)
k
:= sup
0t T

j
0
,k
_
_
U
j
0
(t )
_
_

H
1
2
so that A
(j
0
)
k
A
(j
0
)
as k +. Note that A
(j
0
)
A
c
, else we would have T

j
0
= + by the denition of A
c
which would contradict (3.13). Theorem 3.1 will be proved if we show A
(j
0
)
=A
c
(whereby we can set u
0,c
=V
0,j
0
,
u
c
=U
j
0
), so it remains only to show that A
(j
0
)
A
c
, which we will now do.
By continuity in

H
1
2
, we may take t
k
[0, T

j
0
,k
] such that U
j
0
(t
k
)

H
1
2
=A
(j
0
)
k
. Set t
k,n
=
2
j
0
,n
t
k
, and note that
for k =k

and n =n(m
,
, k

) we have
0 t
k,n

2
j
0
,n
T

j
0
,k
=t
n
k
by (3.15). We may therefore apply Claim 3.8 with t
n
=t
k,n
=t
k

,n(m
,
,k

)
to conclude that for xed , there exists a
subsequence of s such that for n =n(m
,
, k

), J =J(m
,
, k

) and k =k

we have
_
_
u
n
(t
k,n
)
_
_
2

H
1
2
=
J

j=1
_
_
U
j,n
(t
k,n
)
_
_
2

H
1
2
+
_
_
w
l,J
n
(t
k,n
)
_
_
2

H
1
2
+(, ), (3.16)
where (, ) 0 as +. Recall from (3.9) that we dened A
n
so that sup
0t T

(u
0,n
)
u
n
(t )

H
1
2
A
n
.
Therefore, by (3.16), we have
A
2
n(m
,
,k

_
A
(j
0
)
k

_
2
+(, ).
Keeping xed and letting +, this gives A
2
c
(A
(j
0
)
k

)
2
. Finally, letting +, we see A
2
c
(A
(j
0
)
)
2
as
desired proving Theorem 3.1. 2
Remark 3.9. These arguments can moreover be used to show that in fact j
0
=J
1
=1. That is, there is one and only
one prole U
1
which blows up in nite time. This is due to the fact that we have now shown that A
(j
0
)
k

A
c
as
+, hence (3.16) can be used to show that

U
j,n
(t
k,n
)

H
1
2
can be made arbitrarily small for any j =j
0
, which
implies that T

j
=+ by the global existence for small data result and the natural scaling of the equations.
Proof of Claim 3.8. For J N, dene

U
J,n
:=
J

j=1

U
j,n
and

U
J
J
2
,n
:=

U
J,n


U
J
2
,n
for 1 J
2
J.
Fix some k N. We now claim that the following three statements hold, for some subsequences n =n(m) and J =
J(m) indexed by m:
1. For any j, j

N, j =j

,
_
D
1
2
U
j,n
(t
n
), D
1
2
U
j

,n
(t
n
)
_
0 as m. (3.17)
2. For any J J
2
1,
_
_
U
J
J
2
,n
(t
n
)
_
_
2

H
1
2
2
J

j=J
2
+1
_
_
U
j,n
(t
n
)
_
_
2

H
1
2
for all m. (3.18)
3. There exists some M >0 such that
_
_
U
J,n
(t
n
)
_
_

H
1
2
M for all m. (3.19)
178 C.E. Kenig, G.S. Koch / Ann. I. H. Poincar AN 28 (2011) 159187
The easiest to show is (3.19): Since

U
j,n
(t
n
) is dened for all j 1 by assumption, the properties of the prole
decomposition imply
20
(see [19]) that t
n
< T

(u
0,n
) for all n, and so u
n
(t
n
)

H
1
2
2A
c
by (3.9). Using (3.8) and
heat estimates, we see
_
_
w
l,J
n
(t
n
)
_
_

H
1
2

_
_
w
J
n
_
_

H
1
2
3A
c
for large enough n. For each m, use (3.7) to take J large enough, and a corresponding sufciently large n (both in an
increasing fashion) such that
sup
t [0,t
n
k
]
_
_
r
J
n
(t )
_
_

H
1
2
1,
whereupon (3.5) gives

U
J,n
(t
n
)

H
1
2
2A
c
+3A
c
+1 =: M.
To prove (3.17), we take the following diagonal-type subsequence: If t
1,n
:= t
n
/
1,n
is bounded, pass to a sub-
sequence such that t
1,n
t
1
[0, +). Take n(1, k) to be the rst element of this subsequence (or of the original
subsequence if t
1,n
is unbounded). (This is m = 1.) Assume now that the rst m 1 values of n are chosen. If
t
m,n
:=t
n
/
m,n
is bounded, pass to a subsequence of the (m1)st subsequence such that t
m,n
t
m
[0, +). Take
n(m, k) to be the rst element of this sequence which comes after n(m1, k) (or of the original (m1)st subse-
quence if t
m,n
is unbounded). In this way, t
j,n
t
j
[0, +) as m (n = n(m, k)) for any j such that t
j,n
is
bounded.
Note that for 1 j J
1
, t
j,n
=t
n
/
j,n
T

j,k
<+. Therefore if t
j,n
+, necessarily j >J
1
and U
j
E

.
Hence by Theorem 2.9, U
j
(t )

H
1
2
0 as t +, and hence
_
_
U
j,n
(t
j,n
)
_
_

H
1
2
=
_
_
U
j
(t
n
)
_
_

H
1
2
0 as n . (3.20)
In the expression in (3.17), whenever t
j,n
is bounded, using the

H
1
2
-continuity of U
j
in a neighborhood of t
j
, we can
replace
D
1
2
_
1

j,n
U
j
_
x x
j,n

j,n
, t
j,n
__
by D
1
2
_
1

j,n
U
j
_
x x
j,n

j,n
, t
j
__
.
We can also assume that D
1
2
U
j
(t
j
) C

0
by approximating in L
2
x
. If both t
j,n
and t
j

,n
are unbounded, then

_
D
1
2
U
j,n
(t
n
), D
1
2
U
j

,n
(t
n
)
_

_
_
U
j,n
(t
n
)
_
_

H
1
2
_
_
U
j

,n
(t
n
)
_
_

H
1
2
=
_
_
U
j
(t
j,n
)
_
_

H
1
2
_
_
U
j
(t
j

,n
)
_
_

H
1
2
0 as n
by (3.20). If t
j,n
is bounded and t
j

,n
is unbounded, we estimate the above expression by
_
_
U
j
(t
j
)
_
_

H
1
2
_
_
U
j
(t
j

,n
)
_
_

H
1
2
0
as n since we have replaced t
j,n
by the constant t
j
in the rst term, and the second term tends to zero. If both
t
j,n
and t
j

,n
are bounded, replace t
j,n
by t
j
and t
j

,n
by t
j
, and note by a change of variables that
_
D
1
2
_
1

j,n
U
j
_
x x
j,n

j,n
, t
j
__
, D
1
2
_
1

,n
U
j

_
x x
j

,n

,n
, t
j

___
=
_

j,n

,n
_3
2
_
D
1
2
U
j
(y, t
j
)D
1
2
U
j

_

j,n

,n
y +
x
j,n
x
j

,n

,n
, t
j

_
dy
which, since weve assumed the functions in the integrand are compactly supported, is easily seen to tend to zero as
n if
j,n
/
j

,n
0 or if
j,n
=
j

,n
and |(x
j,n
x
j

,n
)/
j

,n
| +. The case
j

,n
/
j,n
0 is handled
using a similar change of variables and (3.17) is proved.
(3.18) is proved using (3.17) as follows: Note that
_
_
U
J
J
2
,n
(t
n
)
_
_
2

H
1
2
=
J

j=J
2
+1
_
_
U
j,n
(t
n
)
_
_
2

H
1
2
+

J
2
j=j

J
_
D
1
2
U
j,n
(t
n
), D
1
2
U
j

,n
(t
n
)
_
.
20
This fact is used again below in Claim 3.10, see (3.22).
C.E. Kenig, G.S. Koch / Ann. I. H. Poincar AN 28 (2011) 159187 179
Suppose the sum on the right has C(J
2
, J) terms in it. For a xed J, (3.17) allows us to take n large enough
(that is, we take a further subsequence of n(m, k)) so that all the terms are smaller than /C(J
2
, J), where
=

J
j=J
2
+1


U
j,n
(t
n
)
2

H
1
2
. As J increases with m, we let n increase sufciently rapidly with m as well so that
this is true for all m, and (3.18) is proved.
Returning now to the proof of Claim 3.8, we want to show that for any >0, there exists some M
1
>0 such that,
for all mM
1
, |A| < where
A:=
_
_
u
n
(t
n
)
_
_
2

H
1
2

j=1
_
_
U
j,n
(t
n
)
_
_
2

H
1
2

_
_
w
l,J
n
(t
n
)
_
_
2

H
1
2
.
We write u
n
=

U
J
2
,n
+

U
J
J
2
,n
+w
l,J
n
+r
J
n
for a sufciently large J
2
to be chosen, expand the rst term in A and cancel
the equal terms. Well then split A into I +II, where II contains all terms with

U
j,n
for j > J
2
(see below). Note
that for j >J
2
>J
1
, U
j
E

and hence, for J


2
sufciently large, Theorem 2.10 and (3.11) give the energy estimate
U
j

V
0,j


H
1
2
. Then, using (3.8), (3.11) and (3.18), for any >0 we have
_
_
U
J
J
2
,n
(t
n
)
_
_
2

H
1
2
2
J

j=J
2
+1
_
_
U
j,n
(t
n
)
_
_
2

H
1
2

j=J
2
+1
V
0,j

H
1
2
< (3.21)
for J
2
=J
2
( ) sufciently large. Now let
II :=2
_
D
1
2
U
J
J
2
,n
(t
n
), D
1
2
_

U
J,n
(t
n
) +w
l,J
n
(t
n
) +r
J
n
(t
n
)
__
+
_
_
U
J
J
2
,n
(t
n
)
_
_
2

H
1
2

j=J
2
+1
_
_
U
j,n
(t )
_
_
2

H
1
2
.
We estimate
|II| 2
_
_
U
J
J
2
,n
(t
n
)
_
_

H
1
2

__
_
U
J,n
(t
n
)
_
_

H
1
2
+
_
_
w
l,J
n
(t
n
)
_
_

H
1
2
+
_
_
r
J
n
(t
n
)
_
_

H
1
2
_
+
_
_
U
J
J
2
,n
(t
n
)
_
_
2

H
1
2
+
J

j=J
2
+1
_
_
U
j,n
(t )
_
_
2

H
1
2
.
Note that we have
_
_
U
J,n
(t
n
)
_
_

H
1
2
M,
_
_
w
l,J
n
(t
n
)
_
_

H
1
2

_
_
w
J
n
_
_

H
1
2
3A
c
and
_
_
r
J
n
(t
n
)
_
_

H
1
2
1
by (3.19), heat estimates and (3.8), and by our choice of subsequences, respectively. Using these and (3.21), for any
>0 we can make |II| </2 for large enough J
2
, which we now x. We are now left to estimate
I :=
_
_
U
J
2
,n
(t
n
) +w
l,J
n
(t
n
) +r
J
n
(t
n
)
_
_
2

H
1
2

J
2

j=1
_
_
U
j,n
(t
n
)
_
_
2

H
1
2

_
_
w
l,J
n
(t
n
)
_
_
2

H
1
2
=

1j=j

J
2
_
D
1
2
U
j,n
(t
n
), D
1
2
U
j

,n
(t
n
)
_
+2
_
D
1
2
U
J
2
,n
(t
n
n), D
1
2
_
w
l,J
n
(t
n
) +r
J
n
(t
n
)
__
+2
_
w
l,J
n
(t
n
), r
J
n
(t
n
)
_
+
_
_
r
J
n
(t
n
)
_
_
2

H
1
2
.
The rst term is handled by (3.17). Since we may pass to a subsequence such that r
J
n
(t
n
)

H
1
2
r
J
n

E
t
n
k
0 as
m by (3.7), using (3.19) and the fact that w
l,J
n
(t
n
) 3A
c
we see that the terms with r
J
n
are small for large m.
The only remaining issue is
_
D
1
2
U
J
2
,n
(t
n
), D
1
2
w
l,J
n
(t
n
)
_
.
Consider D
1
2

U
j,n
(t
n
), D
1
2
w
l,J
n
(t
n
) for 1 j J
2
. If t
j,n
+, then, since w
l,J
n
(t
n
)

H
1
2
3A
c
, we can use
again the fact that U
j
(t
j,n
)

H
1
2
0 as n to see that the term is small for large m (recall that n = n(m)).
Otherwise, it sufces to consider (by replacing t
j,n
by t
j
as before)
180 C.E. Kenig, G.S. Koch / Ann. I. H. Poincar AN 28 (2011) 159187
_
D
1
2
_
1

j,n
U
j
_
x x
j,n

j,n
, t
j
__
, D
1
2
w
l,J
n
(t
n
)
_
=
_
D
1
2
U
j
(y, t
j
)D
1
2
e
t
j,n

j,n
w
J
n
(
j,n
+x
j,n
)
_
(y) dy.
Dene h
n
(y) :=e
t
j,n

[
j,n
w
J
n
(
j,n
+x
j,n
)](y). We claim that h
n
0 in

H
1
2
and hence this term is small for large
m as well (for each j [1, . . . , J
2
]), and Claim 3.8 is proved. Indeed, note that h
n


H
1
2
3A
c
hence h
n
has a weak
accumulation point in

H
1
2
. Note also that h
n

3
w
J
n

3
0 as m (after passing to a subsequence) by (3.6),
so in particular h
n
0 in L
3
. Any weak limit in

H
1
2
for any subsequence is therefore also zero due to the embedding
(L
3
)

(

H
1
2
)

, hence zero is the only weak accumulation point of h


n
in

H
1
2
and thus h
n
0 in

H
1
2
. This concludes
the proof of Claim 3.8 (and Theorem 3.1). 2
3.3. Compactness of critical elements
In this section, well prove Theorem 3.2 which establishes the compactness of any critical element arising from
Theorem 3.1 under the assumption that A
c
<+.
Proof of Theorem 3.2. We will need the following claim:
Claim 3.10. Dene A
c
by (3.2), and suppose that A
c
< +. Suppose u
0


H
1
2
and u = NS(u
0
) are such that
T

(u
0
) < + and u
L

((0,T

(u
0
));

H
1
2 )
= A
c
. For any sequence {t
n
} such that t
n
T

(u
0
), let V
0,j
, U
j
be the
proles associated with the sequence u
0,n
:=u(t
n
). Then, after re-ordering, T

(V
0,j
) <+j =1, and
T

(u
0
) t
n

2
1,n
T

(V
0,1
) (3.22)
for all n. Moreover, after passing to a subsequence in n, for j 2 either U
j
0 or

1,n

j,n
+ as n +. (3.23)
Proof. First note that the following considerations hold for any sequence {t
n
} [0, T

(u
0
)):
Letting u
0,n
=u(t
n
), u
n
(t ) =NS(u
0,n
)(t ) =u(t
n
+t ) and noting T

(u
0
) <+ implies that u
n

E
T

(u
0,n
)
=+
for all n N by Theorem 2.12, we can think of u
0,n
as a minimizing sequence for A
c
, since u(t
n
)

H
1
2
A
c
: A
n
for all n. Proceeding as before and applying the prole decomposition, we see that, for a subsequence,
u(t
n
) =
J

j=1
1

j,n
V
0,j
_
x
j,n

j,n
_
+w
J
n
()
and, by uniqueness,
u(t
n
+t ) =u
n
(t ) =
J

j=1
1

j,n
U
j
_
x
j,n

j,n
,
t

2
j,n
_
+w
l,J
n
(, t ) +r
J
n
(, t )
for t [0, t
n
k
], where t
n
k
:=
2
1,n
(T

(U
1
)
1
k
) for k N with the orthogonality properties (3.6)(3.8). This is justied
since we saw before (see (3.13), (3.14), Remark 3.9, etc.) that we may re-order the elements so that T

(V
0,1
) <+,
T

(V
0,j
) =+ for all j 2 and
U
1

E
T

(V
0,1
)
=+, U
j

<+ for all j 2,


sup
t [0,T

(V
0,1
))
_
_
U
1
(t )
_
_

H
1
2
=A
c
and it is noted moreover in [19] that Eq. (3.22) holds for all n. (In fact, (3.22) is a simple consequence of the properties
of the prole decomposition.)
Suppose now that t
n
T

(u
0
). Then (3.22) gives

2
1,n

(u
0
) t
n
T

(V
0,1
)
0 as n . (3.24)
C.E. Kenig, G.S. Koch / Ann. I. H. Poincar AN 28 (2011) 159187 181
We will now show that moreover, for j 2, either U
j
0 or, after possibly passing to a subsequence, the limit (3.23)
holds as follows: Take t
k
[0, T

(V
0,1
)
1
k
] such that
_
_
U
1
(t
k
)
_
_

H
1
2
= sup
0t T

(V
0,1
)
1
k
_
_
U
1
(t )
_
_

H
1
2
=: A
k
A
c
.
Then, letting t
k,n
:=
2
1,n
t
k
, we can apply Claim 3.8 (where actually J
1
= 1) to see that there exist subsequences
n =n(m, k), J =J(m, k) indexed by m for xed k such that
_
_
u(t
n
+t
k,n
)
_
_
2

H
1
2
=
J

j=1
_
_
U
j,n
(t
k,n
)
_
_
2

H
1
2
+
_
_
w
l,J
n
(t
k,n
)
_
_
2

H
1
2
+(k, m)
where (k, m) 0 as m for xed k. Therefore, as in Remark 3.9, one can take a subsequence m=m(k) such
that, for j 2,
_
_
_
_
U
j
_

2
1,n(m(k),k)

2
j,n(m(k),k)
t
k
__
_
_
_

H
1
2
=
_
_
U
j,n(m(k),k)
(t
k,n(m(k),k)
)
_
_

H
1
2
0 as k .
If along this subsequence n = n(m(k), k) we had that
2
1,n
t
k
/
2
j,n
were bounded, one could therefore extract a con-
vergent subsequence with limit

t < , and therefore conclude that U
j
(

t ) = 0 by the

H
1
2
-continuity and hence that
U
j
0 by Lemma 3.6. Since 0 t
k
< T

(V
0,1
) < , if U
j
= 0 then the limit (3.23) must therefore hold for some
subsequence, and Claim 3.10 is proved. 2
Returning now to the proof of Theorem 3.2, x a sequence {t
n
} such that t
n
T

. We dene a corresponding
sequence {s
n
} by
21
s
n
t
n

2
1,n
=
1
2
T

(V
0,1
) (3.25)
for each n. Note that, for j 2 such that U
j
=0, (3.23) gives (after passing to a subsequence in n)
s
n
t
n

2
j,n
=
T

(V
0,1
)
2

2
1,n

2
j,n
+ as n +. (3.26)
Note also that (3.22) implies that s
n
(t
n
, T

(u
0
)). Indeed, we have
0 <t
n
<s
n
=t
n
+
1
2

2
1,n
T

(V
0,1
) <T

(u
0
)
1
2

2
1,n
T

(V
0,1
) <T

(u
0
).
Notice that, since u
n
=NS(u
0,n
) = NS(u(t
n
)), we may write u(s
n
) =u(t
n
+(s
n
t
n
)) =u
n
(s
n
t
n
) and hence, by
(3.5) we have (for a further subsequence in n)
u(s
n
) =
J

j=1
1

j,n
U
j
_
x
j,n

j,n
,
s
n
t
n

2
j,n
_
+w
l,J
n
(s
n
t
n
) +r
J
n
(s
n
t
n
). (3.27)
Fix any > 0. Since s
n
t
n
=
1
2
T

(V
0,1
)
2
1,n
< T

(V
0,1
)
2
1,n
for all n, the orthogonality properties (3.6) and (3.7)
and Youngs inequality allow us to x some J so large that
_
_
r
J
n
(s
n
t
n
)
_
_
3

_
_
r
J
n
(s
n
t
n
)
_
_

H
1
2


4
(3.28)
and
_
_
w
l,J
n
(s
n
t
n
)
_
_
3

_
_
w
J
n
_
_
3


4
(3.29)
21
The specic factor 1/2 in the denition of s
n
is chosen only for simplicity; one may in fact replace 1/2 by for any xed (0, 1).
182 C.E. Kenig, G.S. Koch / Ann. I. H. Poincar AN 28 (2011) 159187
for large enough n. For such a J, we may take n large enough that
max
2jJ
_
_
_
_
U
j
_
s
n
t
n

2
j,n
__
_
_
_
3


2J
, (3.30)
due to (3.26) and the fact that T

(V
0,j
) =+ for j 2 which implies that lim
t +
U
j
(t )
3
=0 by Theorem 2.9.
We have therefore shown (due to (3.27)) that for any >0, there exists N N such that
_
_
_
_
u(s
n
)
1

1,n
U
1
_
x
1,n

1,n
,
s
n
t
n

2
1,n
__
_
_
_
3
=
_
_
_
_
u(s
n
)
1

1,n
U
1
_
x
1,n

1,n
,
1
2
T

(V
0,1
)
__
_
_
_
3
<
for all n N. A simple change of variables y =
xx
1,n

1,n
shows that
_
_
_
_

1,n
u
_

1,n
_
+
x
1,n

1,n
_
, s
n
_
U
1
_
1
2
T

(V
0,1
)
__
_
_
_
3
<
for n N, and hence setting
n
=(
1,n
)
1
(note that
n
+ by (3.24)) and x
n
=x
1,n
/
1,n
, we see that
22
1

n
u
_
x
n

n
, s
n
_
U
1
_
1
2
T

(V
0,1
)
_
in L
3
as n and Theorem 3.2 is proved. 2
3.4. Rigidity
In this section, well prove Theorem 3.3 which establishes that the compactness of a critical element as established
in Theorem 3.2 is not possible. This, due to Theorems 3.1 and 3.2, gives an alternative proof of the regularity criterion
(3.1) in the case of mild solutions (which follows from the more general result in [14]), and concludes the critical
element proof of Theorem 0.1.
We will use three key known results. The rst two concern systems of parabolic equations: the backwards unique-
ness theorem, Theorem 3.7, proved in [12,13] (see also [14]), and the following unique continuation theorem, a
proof of which can be found in [14] but which was already a well-known fact from the unique continuation theory of
differential inequalities (see also [15]):
Theorem 3.11 (Unique continuation). Let Q
r,
:= B
r
(0) (, 0) for some r, > 0, and suppose a vector-valued
function v and its distributional derivatives satisfy v, v
t
, v,
2
v L
2
(Q
r,
) and there exist c
0
, C
k
>0 (k N) such
that |v
t
v| c
0
(|v| +|v|) a.e. on Q
r,
and |v(x, t )| C
k
(|x| +

t )
k
for all (x, t ) Q
r,
. Then v(x, 0) 0
for all x B
r
(0).
The third result we will need is a local regularity criterion (with estimates) originating in [5] and generalized in
[46] for the so-called suitable weak solutions of the NavierStokes equations, which are dened as follows:
Denition 3.12 (Suitable weak solutions). Let O be an open subset of R
n
, < T
1
< T
2
< +. We call the
pair (u, p) a suitable weak solution of the NavierStokes equations in O (T
1
, T
2
) if u L

((T
1
, T
2
); L
2
(O))
L
2
((T
1
, T
2
); H
1
(O)), p L

((T
1
, T
2
); L
3/2
(O)), u and p satisfy NSE in distributions and the following local energy
inequality holds:
_
O
|u|
2
dx +2
t
_
T
1
_
O
|u|
2
dx ds
t
_
T
1
_
O
_
|u|
2
( +
t
) +u
_
|u|
2
+2p
__
dx ds
for almost all t (T
1
, T
2
) and for all nonnegative cut-off functions C

0
which vanish in a neighborhood of the
parabolic boundary (O{T
1
}) (O[T
1
, T
2
]).
22
As noted previously, for any (0, 1) one can, of course, nd a similar sequence which converges to U
1
( T

(V
0,1
)).
C.E. Kenig, G.S. Koch / Ann. I. H. Poincar AN 28 (2011) 159187 183
We take the statement of the local regularity criterion from [14] (where a self-contained proof, based on the alter-
native method in [39], is presented). It is the following:
Lemma 3.13 (Local smallness regularity criterion). There exist positive absolute constants
0
and c
k
for k N with
the following property: If a suitable weak solution (u, p) of NSE on Q
1
, where Q
r
:= B
r
(0) (r
2
, 0) for r > 0,
satises the condition
_
Q
1
_
|u|
3
+|p|
3
2
_
dx dt <
0
,
then
k1
u is Hlder continuous on Q1
2
for any k N, and for each k we have the estimate
max
Q
1
2

k1
u

c
k
.
We will now prove Theorem 3.3 by using the backwards uniqueness, unique continuation and smallness criterion
results (Theorem 3.7, Theorem 3.11 and Lemma 3.13) to establish the following stronger version of Lemma 3.6:
Lemma 3.14. Suppose u
0
L
3
and set u := NS(u
0
). Suppose there exist some T R such that 0 < T < + and
T T

(u
0
) and a sequence of numbers {s
n
} such that 0 <s
n
T such that the following two properties hold:
(1) sup
t (0,T )
_
_
u(t )
_
_
3
<+ and
(2) lim
n
_
|x|<R

u(x, s
n
)

2
dx =0 for any R >0.
Then u
0
=u =0.
(The case T <T

(u
0
) implies Lemma 3.6 due to the continuity properties of mild solutions.) Note the immediate
corollary that there exists no u
0
L
3
such that T

(u
0
) < + and (1) and (2) hold with T = T

(u
0
), since the
conclusion of Lemma 3.14 implies that T

(u
0
) =T

(0) =+by Theorem 2.6. This will exactly rule out the critical
element produced in Theorem 3.1 and prove the regularity criterion (3.1). Let us postpone the proof of Lemma 3.14
for the moment.
Proof of Theorem 3.3. As in Theorem 3.2, dene A
c
by (3.2) and suppose A
c
<+, and let u
0


H
1
2
with associ-
ated mild solution u :=NS(u
0
) satisfy T

:=T

(u
0
) <+ and u
L

((0,T

(u
0
));

H
1
2 )
=A
c
.
Fix any {t
n
} [0, T

) such that t
n
T

, and let s
n
T

,
n
+ and {x
n
} R
3
be the associated sequences
guaranteed by Theorem 3.2. Using the conclusion of the same theorem, dene
v
n
(x) =
1

n
u
_
x x
n

n
, s
n
_
for any x R
3
and pass to a subsequence in n so that v
n
v in L
3
for some v L
3
. We now make the following
important claim:
Claim 3.15. For s
n
T

as above and for any R >0,


lim
n
_
B
R
(0)

u(x, s
n
)

2
dx =0.
Proof. Fix R >0. We calculate:
_
|x|R

u(x, s
n
)

2
dx =
_
|x|R

n
v
n
(
n
x +x
n
)

2
dx =(
n
)
1
_
|yx
n
|
n
R

v
n
(y)

2
dy =: (
n
)
1
v
n

2
L
2
(B
n
)
.
184 C.E. Kenig, G.S. Koch / Ann. I. H. Poincar AN 28 (2011) 159187
For any small >0, dene B

n
R
:=B

n
R
(0). We then estimate
(
n
)
1
v
n

2
L
2
(B
n
)
=(
n
)
1
v
n

2
L
2
(B
n
B

n
R
)
+(
n
)
1
v
n

2
L
2
(B
n
B
c

n
R
)
(
n
)
1
v
n

2
L
3
(R
3
)
|B

n
R
|
1
3
+(
n
)
1
v
n

2
L
3
(B
n
B
c

n
R
)
|B
n
|
1
3
C(A
c
)
2
R +4CR v
2
L
3
(B
c

n
R
)
for sufciently large n. The rst term of the last line is arbitrarily small for small , and for xed the second term is
arbitrarily small for sufciently large n due to the fact that
n
+. This proves the claim. 2
Assuming Lemma 3.14 holds and taking T =T

, Claim 3.15 clearly completes the proof of Theorem 3.3. There-


fore it remains only to establish the following:
Proof of Lemma 3.14. Since T =T

=T

(u
0
) is the application we seek, we will consider only that case (the case
T <T

(u
0
) is even easier). Assume therefore that u
0
L
3
is such that T

(u
0
) <+,
sup
t (0,T

(u
0
))
_
_
u(t )
_
_
3
<+
where u =NS(u
0
), and there exists {s
n
} with s
n
T

as n such that
lim
n
_
|x|<R

u(x, s
n
)

2
dx =0
for any xed R >0.
Recall that PF = F
1

F for a vector eld F, and for a tensor G we can write


1

( ( G)) =
(

G)) =R
i
R
j
G
ij
, where (R
j
g)

() =
i
j
||
g() and we sum over repeated indices. Therefore, letting
G=u u, F = G and dening p :=R
i
R
j
u
i
u
j
, we have
P (u u) = (u u) +p.
Note that p is well-dened as a distribution since R
k
: L
3
2
L
3
2
boundedly for each k by the CalderonZygmund
theory, so that
_
_
p(t )
_
_
3
2
C
_
_
u(t )
_
_
2
3
, (3.31)
hence
23
p L3
2
,
. Moreover, p satises the equation p(t ) =
i

j
u
i
(t )u
j
(t ) in distributions and therefore p(t ) is
smooth in x for any t (0, T

) since u(t ) is smooth for such t . Moreover, setting B


r
=B
r
(x
0
) for any xed x
0
R
3
and any r >0, we have the following estimates from the classical elliptic theory (see, e.g., [46])
_
_
p(, t )
_
_
C
k,
(B
r
)
c
k
_

i,j
_
_
u
i
(, t )u
j
(, t )
_
_
C
k,
(B
2r
)
+
_
_
p(, t )
_
_
L
3
2 (R
3
)
_
. (3.32)
For any t (0, T

), we now have
u(t ) =e
t
u
0

t
_
0
e
(t s)
_
(u u) +p
_
(s) ds.
It is clear that u
_
0<T <T
L
2
loc unif,x
L
2
t
(R
3
(0, T )), due to the fact that u L
3,
. Therefore u satises (see [38],
Theorem 11.2)
u
t
u + (u u) +p =0, u =0 (3.33)
23
We adopt the notation of [14], setting L
p,
=L

t
L
p
x
for p 1, and denote the corresponding norm by
p,
.
C.E. Kenig, G.S. Koch / Ann. I. H. Poincar AN 28 (2011) 159187 185
in distributions. We now claim that the pair (u, p) locally forms a suitable weak solution, in the sense of Deni-
tion 3.12.
Recall that, by construction, u L
5
(R
3
(0, T )) for any T < T

, so that the LadyzenskajaProdiSerrin reg-


ularity condition (see, e.g., [14]) implies that u is smooth on R
3
(0, T

) (or, alternatively, the construction gives


u L

(R
3
(, T

)) for any >0 and hence is smooth, see, e.g., [38], Proposition 15.1). Therefore (3.31) and
(3.32) together imply that p(t ) is smooth in x for 0 <t <T

and
k
x
p L

loc
(R
3
(0, T

)) for any k.
We clearly have u L

((0, T

); L
2
(O)) for any bounded OR
3
. Since u and p are sufciently smooth on R
3

(0, T

) we may multiply Eq. (3.33) by u for as in Denition 3.12 and integrate by parts to derive the local energy
inequality in

O (0, T

) for a suitably larger region such that O



O, which implies that u L
2
((0, T

); H
1
(O)).
(Indeed, the integrations by parts happen only at the level of integration in x, and then (u, p) L
3,
L3
2
,
allows
integration in t .) Noting also that we have
T

_
0
_
R
3
_
|u|
3
+|p|
3
2
_
dx dt T

_
u
3
3,
+p
3
2
3
2
,
_
<, (3.34)
we see that p L
3
2
(O (0, T

)). The pair (u, p) now clearly satises all the requirements to form a suitable weak
solution to NSE in O(0, T

) for any OR
3
.
Eq. (3.34) implies moreover that for any >0 one can nd R >0 large enough that
T

_
0
_
B

(x
0
)
_
|u|
3
+|p|
3
2
_
dx dt
for all x
0
such that |x
0
| R. Therefore, due to Lemma 3.13 (and an appropriate scaling argument), we see that u
is smooth on :=(R
3
\B
R
0
(0)) [(3/4)T

, T

] for some sufciently large R


0
, and by shifting spatial regions, we
have the global bounds
max

k1
u

c
k
, k =1, 2, 3, . . . . (3.35)
Since u is continuous up to T

outside B
R
0
, Claim 3.15 now implies that u(x, T

) 0 for all x such that |x| R


0
.
We therefore also have :=
x
u 0 on (R
3
\B
R
0
) {T

}. Taking the curl of Eq. (3.33), we see that the vorticity


satises the inequality
|
t
| =

(u ) ( )u

c
_
|| +||
_
in for some c > 0 due to (3.35). Since is bounded and smooth in , we can apply the backwards uniqueness
theorem, Theorem 3.7, to conclude that 0 in .
Now, in

:= R
3
((3/4)T

, (7/8)T

] we have u, and hence , is smooth, hence D


m
t
D
k
x
L
2
loc
(

) for all
m, k 0. We can therefore apply the theorem of unique continuation, Theorem 3.11, to , since 0 in

, to
conclude (by appropriate shifting of local regions) that 0 in

. Therefore, due to the divergence-free condition,
u is harmonic in

which, along with the fact that u L
3,
, implies u 0 in

. Lemma 3.6 (which is essentially
just another application of Theorem 3.7) then gives u
0
=u =0 which proves Lemma 3.14 and concludes the proof of
Theorem 3.3 and Theorem 0.1. 2
Acknowledgements
We would like to thank Professors Isabelle Gallagher and Vladimr verk for many helpful comments and dis-
cussions. The second author would like to extend a personal and special thank you as well to Vladimr verk for his
continuing help, guidance and support.
References
[1] Seven Papers on Equations Related to Mechanics and Heat, American Mathematical Society Translations, Series 2, vol. 75, American Math-
ematical Society, Providence, RI, 1968.
186 C.E. Kenig, G.S. Koch / Ann. I. H. Poincar AN 28 (2011) 159187
[2] J. Bourgain, Global Solutions of Nonlinear Schrdinger Equations, American Mathematical Society Colloquium Publications, vol. 46, Amer-
ican Mathematical Society, Providence, RI, 1999.
[3] J. Bourgain, Global wellposedness of defocusing critical nonlinear Schrdinger equation in the radial case, J. Amer. Math. Soc. 12 (1) (1999)
145171.
[4] Jean Bourgain, Nataa Pavlovi c, Ill-posedness of the NavierStokes equations in a critical space in 3D, J. Funct. Anal. 255 (9) (2008) 2233
2247.
[5] L. Caffarelli, R. Kohn, L. Nirenberg, Partial regularity of suitable weak solutions of the NavierStokes equations, Comm. Pure Appl.
Math. 35 (6) (1982) 771831.
[6] Marco Cannone, Ondelettes, Paraproduits et NavierStokes, Diderot Editeur, Paris, 1995. With a preface by Yves Meyer.
[7] Marco Cannone, Harmonic analysis tools for solving the incompressible NavierStokes equations, in: Handbook of Mathematical Fluid
Dynamics, vol. III, North-Holland, Amsterdam, 2004, pp. 161244.
[8] Jean-Yves Chemin, Le systme de NavierStokes incompressible soixante dix ans aprs Jean Leray, in: Actes des Journes Mathmatiques
la Mmoire de Jean Leray, in: Smin. Congr., vol. 9, Soc. Math. France, Paris, 2004, pp. 99123.
[9] J. Colliander, M. Keel, G. Staflani, H. Takaoka, T. Tao, Global well-posedness and scattering for the energy-critical nonlinear Schrdinger
equation in R
3
, Ann. of Math. (2) 167 (3) (2008) 767865.
[10] Raphal Cte, Carlos E. Kenig, Frank Merle, Scattering belowcritical energy for the radial 4DYangMills equation and for the 2Dcorotational
wave map system, Comm. Math. Phys. 284 (1) (2008) 203225.
[11] Hongjie Dong, Dapeng Du, On the local smoothness of solutions of the NavierStokes equations, J. Math. Fluid Mech. 9 (2) (2007) 139152.
[12] L. Escauriaza, G. Seregin, V. verk, On backward uniqueness for parabolic equations, Zap. Nauchn. Sem. S.-Peterburg. Otdel. Mat. Inst.
Steklov. (POMI) 288 (Kraev. Zadachi Mat. Fiz. i Smezh. Vopr. Teor. Funkts. 32) (2002) 100103, 272.
[13] L. Escauriaza, G. Seregin, V. verk, Backward uniqueness for parabolic equations, Arch. Ration. Mech. Anal. 169 (2) (2003) 147157.
[14] L. Escauriaza, G.A. Seregin, V. verk, L
3,
-solutions of NavierStokes equations and backward uniqueness, Uspekhi Mat. Nauk 58 (2(350))
(2003) 344.
[15] Luis Escauriaza, Francisco Javier Fernndez, Unique continuation for parabolic operators, Ark. Mat. 41 (1) (2003) 3560.
[16] Hiroshi Fujita, Kato Tosio, On the NavierStokes initial value problem. I, Arch. Ration. Mech. Anal. 16 (1964) 269315.
[17] Giulia Furioli, Pierre G. Lemari-Rieusset, Elide Terraneo, Unicit dans L
3
(R
3
) et dautres espaces fonctionnels limites pour NavierStokes,
Rev. Mat. Iberoamericana 16 (3) (2000) 605667.
[18] Giulia Furioli, Pierre Gilles Lemari-Rieusset, Ezzedine Zahrouni, Ali Zhioua, Un thorme de persistance de la rgularit en norme despaces
de Besov pour les solutions de Koch et Tataru des quations de NavierStokes dans R
3
, C. R. Acad. Sci. Paris Sr. I Math. 330 (5) (2000)
339342.
[19] Isabelle Gallagher, Prole decomposition for solutions of the NavierStokes equations, Bull. Soc. Math. France 129 (2) (2001) 285316.
[20] Isabelle Gallagher, Drago s Iftimie, Fabrice Planchon, Non-explosion en temps grand et stabilit de solutions globales des quations de Navier
Stokes, C. R. Math. Acad. Sci. Paris 334 (4) (2002) 289292.
[21] Isabelle Gallagher, Drago s Iftimie, Fabrice Planchon, Asymptotics and stability for global solutions to the NavierStokes equations, Ann. Inst.
Fourier (Grenoble) 53 (5) (2003) 13871424.
[22] Isabelle Gallagher, Gabriel S. Koch, Fabrice Planchon, A prole decomposition approach to the L

t
(L
3
x
) NavierStokes regularity criterion,
arXiv:1012.0145, 2010.
[23] Patrick Grard, Description du dfaut de compacit de linjection de Sobolev, ESAIMControl Optim. Calc. Var. 3 (1998) 213233 (electronic).
[24] Loukas Grafakos, Modern Fourier Analysis, second ed., Graduate Texts in Mathematics, vol. 250, Springer, New York, 2009.
[25] Kato Tosio, Strong L
p
-solutions of the NavierStokes equation in R
m
, with applications to weak solutions, Math. Z. 187 (4) (1984) 471480.
[26] Carlos E. Kenig, Frank Merle, Global well-posedness, scattering and blow-up for the energy-critical, focusing, non-linear Schrdinger equa-
tion in the radial case, Invent. Math. 166 (3) (2006) 645675.
[27] Carlos E. Kenig, Frank Merle, Global well-posedness, scattering and blow-up for the energy-critical focusing non-linear wave equation, Acta
Math. 201 (2) (2008) 147212.
[28] Carlos E. Kenig, Frank Merle, Scattering for H
1/2
bounded solutions to the cubic, defocusing NLS in 3 dimensions, Trans. Amer. Math.
Soc. 362 (4) (2010) 19371962.
[29] Sahbi Keraani, On the defect of compactness for the Strichartz estimates of the Schrdinger equations, J. Differential Equations 175 (2) (2001)
353392.
[30] Sahbi Keraani, On the blow up phenomenon of the critical nonlinear Schrdinger equation, J. Funct. Anal. 235 (1) (2006) 171192.
[31] R. Killip, T. Tao, M. Visan, The cubic nonlinear Schrdinger equation in two dimensions with radial data, J. Eur. Math. Soc. (JEMS) 11 (6)
(2009) 12031258.
[32] Rowan Killip, Monica Visan, Xiaoyi Zhang, The mass-critical nonlinear Schrdinger equation with radial data in dimensions three and higher,
Anal. PDE 1 (2) (2008) 229266.
[33] Gabriel Koch, Prole decompositions for critical Lebesgue and Besov space embeddings, Indiana Univ. Math. J., in press, http://www.iumj.
indiana.edu/IUMJ/forthcoming.php, arXiv:1006.3064, 2010.
[34] Gabriel Koch, Nikolai Nadirashvili, Gregory A. Seregin, Vladimr verk, Liouville theorems for the NavierStokes equations and applica-
tions, Acta Math. 203 (1) (2009) 83105.
[35] Herbert Koch, Daniel Tataru, Well-posedness for the NavierStokes equations, Adv. Math. 157 (1) (2001) 2235.
[36] Joachim Krieger, Wilhelm Schlag, Concentration Compactness for Critical Wave Maps, Monographs of the European Mathematical Society,
in press, http://www.math.upenn.edu/~kriegerj/Papers.html, arXiv:0908.2474.
[37] O.A. Ladyzenskaja, V.A. Solonnikov, N.N. Uralceva, Linear and Quasi-Linear Equations of Parabolic Type, American Mathematical Society,
Providence, RI, 1968.
C.E. Kenig, G.S. Koch / Ann. I. H. Poincar AN 28 (2011) 159187 187
[38] P.G. Lemari-Rieusset, Recent Developments in the NavierStokes Problem, Chapman & Hall/CRC Research Notes in Mathematics, vol. 431,
Chapman & Hall/CRC, Boca Raton, FL, 2002.
[39] Fanghua Lin, A new proof of the CaffarelliKohnNirenberg theorem, Comm. Pure Appl. Math. 51 (3) (1998) 241257.
[40] Changxing Miao, Xu. Guixiang, Lifeng Zhao, Global well-posedness and scattering for the energy-critical, defocusing Hartree equation in
R
1+n
, arXiv:0707.3254.
[41] Changxing Miao, Guixiang Xu, Lifeng Zhao, On the blow-up phenomenon for the mass-critical focusing Hartree equation in R
4
, Colloq.
Math. 119 (1) (2010) 2350.
[42] Changxing Miao, Guixiang Xu, Lifeng Zhao, Global well-posedness and scattering for the energy-critical, defocusing Hartree equation for
radial data, J. Funct. Anal. 253 (2) (2007) 605627.
[43] Changxing Miao, Guixiang Xu, Lifeng Zhao, Global well-posedness and scattering for the mass-critical Hartree equation with radial data,
J. Math. Pures Appl. (9) 91 (1) (2009) 4979.
[44] Changxing Miao, Guixiang Xu, Lifeng Zhao, Global well-posedness, scattering and blow-up for the energy-critical, focusing Hartree equation
in the radial case, Colloq. Math. 114 (2) (2009) 213236.
[45] Sylvie Monniaux, Uniqueness of mild solutions of the NavierStokes equation and maximal L
p
-regularity, C. R. Acad. Sci. Paris Sr. I
Math. 328 (8) (1999) 663668.
[46] J. Ne cas, M. R ui cka, V. verk, On Lerays self-similar solutions of the NavierStokes equations, Acta Math. 176 (2) (1996) 283294.
[47] E. Ryckman, M. Visan, Global well-posedness and scattering for the defocusing energy-critical nonlinear Schrdinger equation in R
1+4
,
Amer. J. Math. 129 (1) (2007) 160.
[48] V.A. Solonnikov, Estimates for solutions of a non-stationary linearized system of NavierStokes equations, Trudy Mat. Inst. Steklov. 70 (1964)
213317.
[49] E.M. Stein, Singular Integrals and Differentiability Properties of Functions, Princeton University Press, Princeton, NJ, 1970.
[50] Terence Tao, Global regularity of wave maps III. Large energy from R
1+2
to hyperbolic spaces, arXiv:0805.4666.
[51] Terence Tao, Global regularity of wave maps VI. Minimal-energy blowup solutions, arXiv:0906.2833.
[52] Terence Tao, Monica Visan, Xiaoyi Zhang, Global well-posedness and scattering for the defocusing mass-critical nonlinear Schrdinger
equation for radial data in high dimensions, Duke Math. J. 140 (1) (2007) 165202.
[53] Terence Tao, Monica Visan, Xiaoyi Zhang, Minimal-mass blowup solutions of the mass-critical NLS, Forum Math. 20 (5) (2008) 881919.
[54] T.-P. Tsai, On Lerays self-similar solutions of the NavierStokes equations satisfying local energy estimates, Arch. Ration. Mech.
Anal. 143 (1) (1998) 2951.
[55] Monica Visan, The defocusing energy-critical nonlinear Schrdinger equation in higher dimensions, Duke Math. J. 138 (2) (2007) 281374.

Vous aimerez peut-être aussi