Lecture Notes 156.01556
Lecture Notes 156.01556
Lecture Notes 156.01556
by
G. Seregin
University of Oxford
G. Seregin
Oxford University
March 2, 2014
2
Context
Chapter I Preliminaries
1.1. Notation
1.2. Newtonian potential
1.3. Equation div u = b
1.4. Nečas imbedding theorem
1.5. Spaces of solenoidal vector fields
1.6. Linear functionals vanishing on divergence free vector fields
1.7. Helmholtz-Weyl decomposition
Chapter II. Linear stationary problem
2.1. Existence and uniqueness of weak solutions
2.2. Coercive estimates
2.3. Local regularity
2.4. Further local regularity results, n = 2, 3
2.5. Stokes operator in bounded domains
Chapter III. Non-linear stationary problem
3.1. Existence of weak solutions
3.2. Regularity of weak solutions
Chapter IV. Linear non-stationary problem
4.1. Derivative in time
4.2. Explicit solution
4.3. Cauchy problem
4.4. Pressure field. Regularity
4.5. Uniqueness results
4.6. Local interior regularity
4.7. Local boundary regularity
Chapter V. Non-linear non-stationary problem
5.1. Compactness results for non-stationary problems
5.2. Auxiliary problem
3
Foreword
The Lecture Notes are based on the TCC course given by me in Trinity
Terms of 2009-2011. Chapters I-III contains material discussed in Trinity
Term of 2009 (16 hours in total), Chapters IV-V contains lectures of 2010
(16 hours), and finally, lectures of 2011 are covered by Chapter VI (16 hours).
Chapters I-V can be regarded as an Introduction to the Mathematical
Theory of the Navier-Stokes equations, relying mainly on the classical PDE’s
approach. First, the notion of weak solutions is introduced, then their exis-
tence is proven (where it is possible), and, afterwards, differentiability prop-
erties are analyzed. In other words, we treat the Navier-Stokes equations
as a particular case, maybe very difficult, of the theory of nonlinear PDE’s.
From this point of view, the Lectures Notes do not pretend to be a complete
mathematical theory of the Navier-Stokes equations. There are different ap-
proaches, for example, more related to harmonic analysis, etc. Corresponding
list of references (incomplete, of course) is given at the end of the Lecture
Notes.
Finally, Chapters VI and VII contains more advanced material, which
reflects my scientific interests.
Chapter 1
Preliminaries
1.1 Notation
Let us denote by Ω a domain (open connected set) in Rn . Then, C0∞ (Ω; Rm ) is
set of all infinitely differentiable functions from Ω into Rm , having a compact
support in Ω. If m = 1, we use abbreviation C0∞ (Ω).
The Lebesgue space is be denoted as Lp (Ω) and it is be endowed with the
standard norm Z p1
p
kf kp,Ω = |f (x)| dx
Ω
if 1 ≤ p < ∞ and
kf k∞,Ω = ess sup |f (x)|
x∈Ω
if p = ∞.
Lemma 1.1. Let 1 ≤ p < ∞. Then, Lp (Ω) = [C0∞ (Ω)]Lp (Ω) , i.e., Lp (Ω) is
the completion of C0∞ (Ω) in Lp (Ω).
In what follows, we always assume that the exponent of integrability is
finite unless otherwise is specially indicated.
We say that a distribution u, defined in Ω, belongs to the Sobolev space
Wsk (Ω) if and only if all its weak derivatives up to order k are integrable in
Ω with the power s. The norm in this space is defined as
k
X
kukWsk (Ω) = k∇i uks,Ω < ∞.
i=0
5
6 CHAPTER 1. PRELIMINARIES
We also let ◦
k ∞ W (Ω) k
W s (Ω) = [C0 (Ω)] s .
It is said that a distribution u, defined in Ω, belongs to the space Lks (Ω)
if and only if all its weak derivatives of order k are integrable in Ω with the
power s. The norm in this space is defined as
kukLks (Ω) = k∇k uks,Ω < ∞.
A non-trivial statement about spaces Lks (Ω) is as follows.
Theorem 1.2. Lks (Ω) ⊂ Ls,loc (Ω).
Proof We have a distribution T such that
Z
g(x)ϕ(x)dx = −T (∇ϕ)
Ω
for ϕ ∈ C0∞ (Ω) with g = (gi ) ∈ Ls (Ω). Our aim is to show that is in fact a
regular distribution and moreover, there exists a function u ∈ Ls,loc (Ω) such
that T = Tu .
Consider a subdomain Ω0 b Ω, i.e., a bounded subdomain Ω0 ⊂ Ω such
that the closure of Ω0 belongs to Ω. Let 0 < % < dist(Ω0 , ∂Ω). Define a
linear functional l : L1 (Ω0 ) → R in the following way
l(ψ) := T (ψ% )
for ψ ∈ L1 (Ω0 ) where
Z
ψ% (x) = $% (x − y)ψ(y)dy
Ω0
By Poincare inequality
u%0 * u0
0 ≤ ϕ(x) ≤ 1 x ∈ Rn ,
as R → ∞. ◦
As usual, the equivalence classes are introduced in L22 (Ω) so that
(iii) 4u = −f
n
in R .
Proof (i) follows from the theory of singular integrals. (iii) follows from
(i),(ii), and from the classical PDE theory.
Let us prove (ii), assuming that n ≥ 3. By Lemma 1.1, there exists a
sequence fm ∈ C0∞ (Rn ) such that fm → f in Lp (Rn ). Since supp fm is a
compact set in Rn ,
c(m, i)
|∇i um (x)| ≤ n−2+i
|x|
for all x ∈ Rn , for i = 0, 1, 2, for all m = 1, 2, ..., and for some positive c(m, i).
Here, um = E ? fm .
We claim that
◦
um ∈ L2p (Rn ).
Indeed, we have
Z h Z
2 p
|∇ (ϕR um − um )| dx ≤ c |∇2 um |p dx+
Rn Rn \B(R)
Z Z
1 p 1 i
+ p |∇um | dx + 2p |um |p dx ≤
R R
B(2R)\B(R) B(2R)\B(R)
10 CHAPTER 1. PRELIMINARIES
Z h 1 Rn 1 Rn i
≤c |∇2 um |p dx + C(m) p (n−1)p + 2p (n−2)p → 0
R R R R
Rn \B(R)
and u, u,n = ∂u/∂xn , and u,nn are in Lp,loc (Rn ) that implies u(x0 , 0) = 0.
So, the Newtonian potential u solves the following Dirichlet problem in half-
space:
4u = −f (1.2.1)
in Rn+ = {x = (x0 , xn ) : xn > 0},
u(x0 , 0) = 0
for any x0 .
2. The same arguments show that if f (x0 , xn ) = f (x0 , −xn ) then u solves
the Neumann boundary value problem, i.e., it satisfies (1.2.1) and the Neu-
mann boundary condition
u,n (x0 , 0) = 0
for any x0 .
and
div u = −div ∇h = −4h = b.
In the case of the half-space, i.e., Ω = Rn+ := {x = (x0 , x3 ) : x0 ∈ R2 , x3 >
0}, we have
and
k∇2 hk2,R3+ ≤ ckbk2,R3+ . (1.3.2)
The idea is to look for u in the form
u = ∇h + rot A,
div u = 4h = b in R3+ .
12 CHAPTER 1. PRELIMINARIES
Equations for A is coming from condition u|x3 =0 = 0 that leads to the fol-
lowing relations
rot A = −∇h
at x3 = 0. We are seeking A, satisfying additional assumptions:
A|x3 =0 = 0, A3 ≡ 0 in R3+ .
where B1 (x0 ) = h2 (x0 , 0) and B2 (x0 ) = −h,1 (x0 , 0) are known functions.
Now, we are going to exploit arguments, which are quite typical for the
theory of traces for functions from Sobolev spaces. This theory suggests to
seek A in the form:
Z2
Aα (x0 , x3 ) = x3 Bα (x0 + y 0 x3 )K(y 0 )dy 0 ,
R
u1 = h,1 + A2,3 ,
u2 = h,2 − A1,3 ,
u3 = h,3 + A1,2 − A2,1 ,
with the described above A, satisfy all the requirements. This can be done
by direct calculations.
Indeed,
Z Z
∂Bα 0 0
Aα,3 (x) = x3 (x +y x3 )yβ K(y )dy + Bα (x0 +y 0 x3 )K(y 0 )dy 0 . (1.3.3)
0 0
∂zβ
R2 R2
1.3. EQUATION DIV U = B 13
Observing that
∂Bα ∂Bα 1
=
∂zβ ∂yβ x3
and integrating by parts with respect yβ , we can transform the right hand
side of (1.3.3) to the form
Z h i ∂
Aα,3 (x) = − Bα (x0 + y 0 x3 ) − Bα (x0 ) (yβ K(y 0 ))dy 0 +
∂yβ
R2
Z
+ Bα (x0 + y 0 x3 )K(y 0 )dy 0 .
R2
Lemma 3.4. For any smooth function f : R3+ → R, vanishing for sufficiently
large |x|, the following inequality is valid:
Z Z 0 0 Z
2 0 0 2 dx dy
kf k 21 := |f (x , 0) − f (y , 0)| 0 0 3
≤ c |∇f |2 dx (1.3.5)
2
L2 (R ) |x − y |
R2 R2 R3+
dz 0
Z Z
2
kf k 21 = |f (x0 + z 0 , 0) − f (x0 , 0)|2 dx0 .
L2 (R2 ) |z 0 |3
R2 R2
According to (1.3.6), we shall evaluate three integrals. In the first one, the
polar coordinates z 0 = (% cos ϕ, % sin ϕ) are used so that:
dz 0
Z Z
I1 = |f (x0 , |z 0 |) − f (x0 , 0)|2 dx0 =
|z 0 |3
R2 R2
Z Z∞
0 1
= 2π dx 2
|f (x0 , %) − f (x0 , 0)|2 d%.
%
R2 0
The right hand side of the latter inequality can be bounded from above with
the help of Hardy’s inequality
Z∞ p p Z∞
−p
t |g(t) − g(0)| dt ≤ |g 0 (t)|p dt
p
p−1
0 0
dz 0
Z Z Z
0 0 0 2 0
= |f (y , |z |) − f (y , 0)| dy ≤ 8π |∇f |2 dx.
|z 0 |3
R2 R2 R3+
Z1 ∂
|f (x0 + z 0 , |z 0 |) − f (x0 , |z 0 |)| = (x0 + tz 0 , |z 0 |)dt ≤
∂t
0
Z1 21
0 0 0 0 2
≤ |z | |∇x f (x + tz , |z |)| dt ,
0
1.3. EQUATION DIV U = B 15
which give us
dz 0
Z Z
I3 = |f (x0 + z 0 , |z 0 |) − f (x0 , |z 0 |)|2 dx0 ≤
|z 0 |3
R2 R2
Z1
dz 0
Z Z
≤ dt |∇x0 f (x0 + tz 0 , |z 0 |)|2 dx0 =
|z 0 |
R2 0 R2
dz 0
Z Z
= |∇y0 f (y 0 , |z 0 |)|2 dy 0
|z 0 |
R2 R2
Indeed, since
c(b)
|∇h(x)| ≤
|x|2
for |x| 1 and, by (1.3.5) and by (1.3.2), one can conclude that
Then statement (ii) of Proposition 3.2 for this particular class of b follows.
Now, let us prove (1.3.8), directly working out the second derivatives of
A, Z
0 ∂ ∂
Aα,β3 (x , x3 ) = − Bα (x0 + y 0 x3 ) (yγ K(y 0 ))dy 0 +
∂xβ ∂yα
R2
Z
∂
+ Bα (x0 + y 0 x3 )K(y 0 )dy 0 .
∂xβ
R2
Obviously,
∂ ∂ 1
Bα (x0 + y 0 x3 ) = Bα (x0 + y 0 x3 ) .
∂xβ ∂xβ x3
Then Z
1
S = Aα,β3 (x) = δBα (x0 , y 0 , x3 )Kβ (y 0 )dy 0 ,
x3
R2
where
δBα (x0 , y 0 , x3 ) := Bα (x0 + y 0 x3 ) − Bα (x0 )
and
Kβ (y 0 ) := (yγ K(y 0 )),βγ − K,β (y 0 ).
Now, we have
Z Z∞ Z 1 Z 2
2 0 0 0 0 0
S dx ≤ c dx3 dx |δBα (x , y , x3 )||Kβ (y )|dy
x3
R3+ 0 R2 R2
1.3. EQUATION DIV U = B 17
Z Z∞ Z
0 0 dx3
≤c |Kβ (y )|dy |Bα (x0 + y 0 x3 ) − Bα (x0 )|2 dx0 .
x23
R2 0 R2
Z Z1 Z2π Z∞ Z
dx3
2
S dx ≤ c %d%dϕ |Bα (x0 + %(cos ϕ, sin ϕ)x3 ) − Bα (x0 )|2 dx0 .
x23
R3+ 0 0 0 R2
Z1
dz 0
Z Z Z
2
S dx ≤ c %d% |Bα (x0 + z 0 %) − Bα (x0 )|2 dx0 .
|z 0 |3
R3+ 0 R2 R2
Letting y 0 = z 0 %, we show
Z1
|Bα (x0 + y 0 ) − Bα (x0 )|2 0 0
Z Z Z
2
S dx ≤ c d% dx dy ≤
|y 0 |3
R3+ 0 R2 R2
|Bα (z 0 ) − Bα (y 0 )|2 0 0
Z Z
≤ 0 0 3
dz dy ≤ ck∇h(·, 0)k2 21 . (1.3.9)
|z − y | L2 (R2 )
R2 R2
where
e 0 ) := (yγ (yβ K(y 0 )),β )),γ − (yγ K(y 0 )),γ ,
K(y
18 CHAPTER 1. PRELIMINARIES
and Z
1
Aα,βγ (x) = (Bα (x0 + y 0 x3 ) − Bα (x0 ))(K(y 0 )),βγ dy 0 .
x3
R2
for any b ∈ C0∞ (R3+ ). The proof of (1.3.10) consists of two parts.
Step 1. Let us show first that ∇h ∈ L2 (R3+ ). Indeed, since ∇2 h ∈
L2 (R3+ ), ∇h ∈ L2 (B+ (1)), where B+ (R) := {x ∈ B(R) : x3 > 0}. We know
that
c(b)
|∇h(x)| ≤
|x|2
for |x| 1. So,
Z Z Z ZR
d%
|∇h|2 dx ≤ |∇h|2 dx + |∇h|2 dx ≤ ... + c(b) ≤ c(b)
%2
B+ (R) B+ (1) B+ (R)\B+ (1) 1
where
K0 (y 0 ) = K(y 0 ) − (yβ K(y 0 )),β .
Let α = 1. Then B1 (x0 ) = h,2 (x0 , 0) and
∂ 1
B1 (x0 + y 0 x3 ) = h(x0 + y 0 x3 , 0) .
∂y2 x3
1.3. EQUATION DIV U = B 19
So, Z
1 ∂
A1,3 (x) = − (h(x0 + y 0 x3 , 0) − h(x0 , 0)) K0 (y 0 )dy 0 .
x3 ∂y2
R2
Repeating the evaluation of Aα,β3 , we find
Z
A21,3 dx ≤ ckh(·, 0)k 1 .
L22 (R(2)
R3+
Since |h(x)| ≤ c(b)/|x| for |x| 1, one can derive with the help of Lemma
3.4 the inequality
kh(·, 0)k 12 ≤ ck∇hk2,R3+ .
L2 (R(2)
This means that, by Step 1, A1,3 ∈ L2 (R3+ ). The same is true for A2,3 . The
proof of the fact that Aα,β ∈ L2 (R3+ ) is an exercise. So, it has been proven
that
u ∈ L2 (R3+ ) (1.3.11)
provided b ∈ C0∞ (R3+ ).
Now, we wish to finish the proof of (1.3.10). Letting uR = ϕR u, we
observe that
Z Z Z
2 2 1
|∇(u − uR )| dx ≤ c |∇u| dx + c 2 |u|2 dx → 0
R
R3+ R3+ \B+ (R) B+ (2R)\B+ (R)
as R → ∞.
The function uR is not compactly supported in R3+ and we need to cut
it in the direction of x3 . To this end, let us introduce the following cut-off
function: χ(t) = 0 if −∞ < t ≤ ε/2, χ(t) = 2(t − ε/2)/ε if ε/2 < t ≤ ε, and
χ(t) = 1 if t > ε. Considering uR,ε (x) = uR (x)χ(x3 ), we have
Z Zε Z Z Zε
2 2 1 0 0
|∇(uR − uR,ε )| dx ≤ c dx3 |∇uR | dx + 2 dx |uR |2 dx3 .
ε
R3+ 0 R2 R2 0
The first integrals on the right hand side of the last inequality tends to zero
as ε → 0. To show that the second term does the same, we are going to use
two facts. Firstly, uR (x0 , 0) = 0 and secondly, by the Friedreich inequality,
Zε Zε
∂
2
0 2 2 0
|uR (x , x3 )| dx3 ≤ cε uR (x , x3 ) dx3 .
∂x3
0 0
20 CHAPTER 1. PRELIMINARIES
in R3+ and
k∇u(m) k2,R3+ ≤ ckb(m) k2,R3+ .
Moreover, by construction
◦
Then, for any b ∈ L̄p (Ω), there exists u ∈ L12 (Ω) with the following properties:
div u = b
in Ω and
k∇ukp,Ω ≤ c(p, n, Ω)kbkp,Ω .
Remark 3.6. For bounded domain domains, we need the restriction on b:
Z
b(x)dx = 0.
Ω
and
kvf0 kV 0 ≤ ckf kH ∀f ∈ H.
So, we have a bounded linear operator τ : H → V 0 (one-to-one by density)
defined by the identity τ f = vf0 for f ∈ H.
Obviously, τ (H) is a linear manifold V 0 . Moreover, it is dense in V 0 .
To see that it is true, assume it is not, i.e., there exists v00 ∈ V 0 but v00 ∈ /
V0 00 00 0 0
[τ (H)] . By the Hahn-Banach theorem, there exists v ∈ V := (V ) with
the properties < v 00 , v00 >= 1 and < v 00 , v 0 >= 0 for any v 0 ∈ τ (H). Since
V is reflexive, there should be v ∈ V so that < v 00 , v 0 >=< v 0 , v > for any
v 0 ∈ V 0 . This gives us: < v00 , v >= 1 and < vf0 , v >= (f, v) = 0 for any
f ∈ H. Therefore, v = 0 and we get a contradiction. The latter allows us to
identify V 0 with the closure of τ (H) in V 0 . But we can go further and identify
duality relation between V and V 0 with the scalar product (·, ·) on H. Very
often, we call such an identification of V 0 the space dual to V relative to the
Hilbert space H.
So, under our standing assumptions, v 0 ∈ V 0 means that there exists a
sequence sequence fm ∈ H such that
as k, n → ∞ and (v 0 , v) is just notation for lim (fk , v) that exists for all
k→∞
v ∈ V . Moreover,
∇p ∈ L−1
r (Ω).
Ω0
1.4. NEČAS IMBEDDING THEOREM 23
K := k∇pkL−1
r (Ω)
and Z
1
[q]ω = q(x)dx.
|ω|
ω
◦
By Theorem 3.5, there exists u ∈ L1r0 (Ω0 ) such that
div u = q
in Ω0 and
k∇ukr0 ,Ω0 ≤ c(r, n, Ω0 )kqkr0 ,Ω0 .
Functions u and q are supposed to be extended by zero outside Ω0 . Let us
mollify u in a standard way
Z Z
(u)% (x) = ω% (x − y)u(y)dy = ω% (x − y)u(y)dy
Ω0 Rn
24 CHAPTER 1. PRELIMINARIES
with the help of a smooth mollifier ω% . So, (u)% ∈ C0∞ (Ω) for 0 < % <
%0 (Ω0 , Ω). Moreover, we know that
∇(u)% = (∇u)%
and thus
div (u)% = (q)% ,
k∇(u)% kr0 ,Ω ≤ k∇ukr0 ,Ω ≤ ckqkr0 ,Ω ,
by the known mollification properties.
Now, we have (in the sense of distributions)
< ∇p, (u)% >= − < p, div (u)% >= − < p, (q)% >,
which implies
∇k (q)% → ∇k q
k∇ukr0 ,Ω ≤ ckqkr0 ,Ω .
◦
By the definition of L1r0 (Ω), there exists a sequence u(m) ∈ C0∞ (Ω) such that
and thus
q (m) := div u(m) → q in Lr0 (Ω)
as m → ∞. Then, as it is pointed out above, we should have
Z Z
(m)
< ∇p, u >= − pdiv u dx = − pq (m) dx ≤ cKk∇u(m) kr0 ,Ω .
(m)
Ω Ω
kpkr,Ω ≤ cK.
|w|2
Z Z Z
xk 2 xk
2 ww,k 2 dx = (|w| ),k 2 dx = dx ≤
|x| ln |x| |x| ln |x| |x|2 ln2 |x|
|x|>1 |x|>1 |x|>1
Z |w|2 21 Z 21
2
≤2 dx |∇w(x)| dx
|x|2 ln2 |x|
|x|>1 |x|>1
− ln ln |x| + ln ln R3
ψR (x) =
− ln ln R + ln ln R3
if R ≤ |x| ≤ R3 .
Given v ∈ V̂ (R2 ), let
Z Z
IR := |∇(v − vψR )| dx = |∇(1 − ψR )v|2 dx.
2
Ω Ω
|v(x)|2
Z Z
IR ≤ c dx + c |∇v|2 dx. (1.5.1)
|x|2 ln2 |x|
R<|x|<R3 R3 \B(R)
If
|v(x)|2
Z
dx < ∞ (1.5.2)
|x|2 ln2 |x|
3<|x|
is valid, then the right hand side of (5.2) tends to zero as R → ∞ and thus
vψR → v (1.5.3)
in L12 (R2 ).
28 CHAPTER 1. PRELIMINARIES
◦
Remark 5.3. Show that 1 ∈ L12 (R2 ).
In order to prove the validity of (1.5.2), we find a sequence v (k) ∈ C0∞ (R2 )
such that
∇v (k) → ∇v
in L2 (R2 ) and
v (k) → v
a.e. in R2 , and sequence v (k) is bounded in L2 (B(3)). Now, let a smooth
cut-off function η satisfy the properties: η(x) = 0 if |x| < 2 and η(x) = 1 if
|x| > 3. Then, by Leray’s inequality, we have
Z Z
(k) 2
≤c |∇v | dx + c |v (k) |2 dx.
1<|x| 3>|x|
The right hand side of the latter inequality is bounded uniformly in k. And
then (1.5.2) follows from Fatou’s lemma. So, (1.5.3) is proven.
Unfortunately, vψR is not divergence free, in fact, div (vψR ) = v · ∇ψR ,
and we should correct it with he help of results of the previous section. We
apply Theorem 3.5 in the ball B(R3 ), which reads that there exists a function
◦ ◦
wR ∈ L12 (B(R3 )) = W 12 (B(R3 ))
such that
div wR = v · ∇ψR
in B(R3 ) and
k∇wR k2,B(R3 ) ≤ ckv · ∇ψR k2,B(R3 ) .
It is worthy to notice that a constant c in the latter inequality is independent
of v and of R (by scaling). We extend wR by zero to the whole R2 . By the
choice of cut-off function ψR , we have
|v(x)|2
Z
2
k∇wR k2,R2 ≤ c dx → 0
|x|2 ln2 |x|
R<|x|<R3
1.5. SPACES OF SOLENOIDAL VECTOR FIELDS 29
as R → ∞.
We let uR = v · ∇ψR − wR . New function has two important properties:
uR ≡ 0 out of B(R3 ), div uR = 0 in R2 , and
as R → ∞.
∞
On the other hand, the mollification (uR )ε ∈ C0,0 (R2 ) and
R2 R2+
as λ → 0.
Next, we introduce the new function vR,λ = uλ ψR,λ , where ψR,λ (x) =
ψR (x1 , x2 − 2λ). Then
Z Z
|∇vR,λ − ∇uλ | dx = |∇vR,λ − ∇uλ |2 dx =
2
R2 R2+
Z
|∇(ψR ũ − ũ)|2 dx → 0
R2
One should correct the divergence of uR,λ as in the first part. We state that
there exists ◦
wR,λ ∈ L12 (B+ ((0, 2λ), R3 )),
30 CHAPTER 1. PRELIMINARIES
∇uR,λ → ∇uR
in L2 (R2+ ) as λ → 0,
div uR,λ = 0
in R2 ,
uR,λ = 0
out of semi-disk B+ ((0, 2λ), R3 ),
∇uR,λ ∈ L2 (R2 ).
∞
Then it remain to take the mollification (uR,λ )ε that belongs to C0,0 (R2+ ) for
sufficiently small ε, for example, 0 < ε < λ. The well-known property of the
mollification
∇(uR,λ )ε → ∇uR,λ in L2 (R2+ )
completes the proof.
In fact, we have the following statement
Theorem 5.4. Let 1 < m < ∞ and let Ω be Rn , or Rn+ , or a bounded domain
with Lipschitz boundary. Then
◦
J m (Ω) = Jˆm (Ω).
1 1
|l(v)| ≤ ck∇vks,Ω
◦
for any v ∈ L1s (Ω) and
l(v) = 0
for any v ∈ Jˆ1s (Ω).
Then there exists a function p ∈ Ls0 (Ω), s0 = s/(s − 1), such that
Z
l(v) = pdiv vdx
Ω
◦
for any v ∈ L1s (Ω).
k∇uks,Ω ≤ ckqks,Ω .
for any q ∈ Ls (Ω). So, the functional G is bounded on Ls (Ω) and by Riesz
theorem, there exists p ∈ Ls0 (Ω) such that
Z
G(q) = pqdx
Ω
◦
for any q ∈ Ls (Ω). Now, for any u ∈ L1s (Ω), we have
Z
l(u) = G(div u) = pdiv udx.
Ω
For bounded Lipschitz domains, one should replace the space Ls (Ω) with
its subspace
L̄s (Ω) = {q ∈ Ls (Ω) : [q]Ω = 0}
and use the same arguments as above.
◦
However, we can assume that our functional vanish on J 1s (Ω) only.
◦
for any v ∈ L1s (Ω).
Ωm ⊂ Ωm+1
1.6. LINEAR FUNCTIONALS 33
and ∞
[
Ω= Ωm .
m=1
◦
Given v ∈ L1s (Ωm ), define v m = v in Ωm and v m = 0 outside Ωm . Obvi-
◦ ◦
ously, v m ∈ L1s (Ω). We also define a linear functional lm : L1s (Ωm ) → R as
follows:
lm (v) := l(vm )
◦
for any ∈ L1s (Ωm ). It is a bounded functional, i.e.,
|lm (v)| ≤ k∇vks,Ωm
with a constant c independent of m.
By mollification, we can show the following fact: if v ∈ Jˆ1s (Ωm ), then
◦
v m ∈ J 1s (Ω). This immediately implies that lm (v) = 0 for any v ∈ Jˆ1s (Ωm ).
According to Proposition 6.1, there exists pm ∈ Ls0 (Ωm ) such that
Z
lm (v) = pm div vdx
Ωm
◦
for any v ∈ L1s (Ωm ). Obviously, pm is defined up to an arbitrary constant.
Moreover, pm+1 − pm = c(m) = constant in Ωm if k > m. So, we can change
pm+1 adding a constant to achieve the identity pm+1 = pm in Ωm . This
allow us to introduce a function p ∈ Ls0 ,loc (Ω) so that p = pm on Ωm . By
construction, it satisfies identity
Z
l(v) = pdiv vdx, (1.6.1)
Ω
l∗ (v∗ ) = 1
and
l∗ (v) = 0
◦
for any v ∈ J 1s (Ω). By Theorem 6.2, there exists p ∈ Ls0 (Ω) such that
Z
l∗ (v) = pdiv vdx
Ω
◦
for all v ∈ L1s (Ω). However,
Z
l∗ (v∗ ) = pdiv v∗ dx = 0
Ω
since Z
v · ∇pdx = 0
Ω
∞
for any p ∈ G(Ω) and for any v ∈ C0,0 (Ω). Now, assume
◦
u ∈ (J (Ω))⊥ ,
≤ c(Ω)kuk2,Ω k∇vk2,Ω
◦ ◦ ◦
for any v ∈ L1s (Ω). So, l : L12 (Ω) = W 12 (Ω) → R is bounded and l(v) = 0 for
◦
any v ∈ J 12 (Ω) =: V (Ω).
By Theorem 6.2, there exists p ∈ L2 (Ω) such that
Z
l(v) = pdiv vdx
Ω
◦
for any v ∈ L1s (Ω). Therefore, u = ∇p and thus p ∈ G(Ω) and
◦
(J (Ω))⊥ ⊆ G(Ω).
Ω̃ Ω̃ B∗
Ωs Ωs
l l
p(jk ) * pl , ∇p(jk ) * ∇pl
p = pl
p(js ) * p, ∇p(js ) * ∇p
ω Ω
1.8 Comments
The main aim for writing up Chapter I is to show author’s preferences how
the theory of functional spaces related to Navier-Stokes equations can be
developed. In our approach, the basic things are estimates of certain solutions
to the equation div u = f and their applications to the derivation of the Nečas
embedding theorem. Each part of this theory can be given in either more
compact way or even in a different way. For example, in Section 3, one could
apply very nice Bogovskii’s approach, see [1], based on the theory of singular
integrals. For more generic and detailed investigation of spaces arising in the
Navier-Stokes theory, we refer the reader to the monographs [29], [68], and
[17].
Chapter 2
u|∂Ω = 0 (2.1.2)
and if n = 3 and Ω is unbounded then u(x) → u0 as |x| → ∞.
In what follows, we always consider the case
u0 = 0.
Let Z
(f, g) := f (x)g(x)dx.
Ω
∞
If u and p are smooth, then, for any v ∈ C0,0 (Ω), integration by parts
gives the following identity:
Z Z
(−4u + ∇p) · vdx = ∇u : ∇vdx = (∇u, ∇v) = (f, v),
Ω Ω
39
40 CHAPTER 2. LINEAR STATIONARY PROBLEM
f ∈ L−1
2 (Ω).
Remark 1.2. Boundary conditions are understood in the sense of trace, see
◦
the definition of spaces L12 (Ω) and V̂ (Ω) in Sections 1 and 5 of Chapter I. If
Ω is unbounded and n = 3, then condition u(x) → 0 as |x| → ∞ holds in the
following sense:
Z 61 Z 21
6 2
|u| dx ≤ c |∇u| dx .
Ω Ω
Lemma 1.3. (Existence). Given f , there exist at least one weak solution to
boundary value problem (2.1.1) and (2.1.2) that satisfies the estimate
k∇uk2,Ω ≤ kf kL−1
2 (Ω)
.
Proof Assume that there are two different solutions u1 and u2 . Then
(∇(u1 − u2 ), ∇v) = 0
2.2. COERCIVE ESTIMATES 41
∞
for and v ∈ C0,0 (Ω) and, by assumption (2.1.3),
k∇(u1 − u2 )k2,Ω = 0.
holds.
42 CHAPTER 2. LINEAR STATIONARY PROBLEM
div 4h u = 4h g
Now, our aim is to evaluate the right hand side of the latter inequality.
By the definition, we have
1
k4h f + 4wh kL−1 n =
2 (R+ )
|h|
n 1 Z 1
Z o
= sup 4h f · vdx + 4wh · vdx : v ∈ C0∞ (R3+ ), k∇k2,R3+ ≤ 1 =
|h| |h|
R3+ R3+
Z Z
n 1 1 o
= sup − f · 4h vdx − ∇wh : ∇vdx : ... ≤
|h| |h|
R3+ R3+
n 1 Z 12 o
2 ∞ 3
≤ kf k2,R3+ sup |4h v| dx : v ∈ C0 (R+ ), k∇k2,R3+ ≤ 1 +
|h|
R3+
2.2. COERCIVE ESTIMATES 43
1 1
+ k∇wh k2,R3+ ≤ kf k2,R3+ + k∇wh k2,R3+ .
|h| |h|
In the last line, there has been used the following fact (exercise)
n 1 Z 12 o
sup |4h v|2 dx : v ∈ C0∞ (R3+ ), k∇vk2,R3+ ≤ 1 ≤ 1.
|h|
R3+
So, we have
1 h 1 i
k∇(4h u)k2,R3+ ≤ c kf k2,R3+ + k4h gk2,R3+ .
|h| |h|
Tending h to zero, we find the bound for tangential derivatives of v
where α = 1, 2 and
I = kf k2,R3+ + k∇gk2,R3+ .
Step 2 Let us start with evaluation of terms u3,33 and p,3 . u3,33 can be
estimated simply with the help of the equation div u = uα,α + u3,3 = 0. This
gives us u3,33 = −uα,3α and thus
As to the second term, the above estimate, the equation p,3 = f3 + 4u3 , and
bounds for tangential derivatives lead to the inequality
g ∈ L2 (Ω) ⇒ k∇gkL−1
2 (Ω)
≤ ckgkL2 (Ω) .
and
p,α3 = f3,α + u3,iiα = f3,α + (u3,iα ),i ∈ L−1 3
2 (R+ ).
44 CHAPTER 2. LINEAR STATIONARY PROBLEM
So, by the aforesaid observation and by the latter estimates, we can conclude
that
k∇p,α kL−1 3
2 (R+ )
≤ cI.
On the other hand, by Theorem 4.1 of Chapter I,
k∇pk2,R3+ ≤ cI.
The remaining part of the second derivatives can be estimated with the
help of the equation uα,33 = −uα,ββ + p,α − fα and previous bounds. Indeed,
we have
k∇2 uk2,R3+ ≤ cI
and this completes the proof.
Theorem 2.3. Assume that all assumptions of Proposition 2.1 are fulfilled.
Let Ω be a bounded domain with sufficiently smooth boundary. In addition,
assume that
f ∈ Wrk (Ω), g ∈ Wrk+1 (Ω)
with [g]Ω = 0 and with integer k. Then
h i
k∇2 ukWrk (Ω) + k∇pkWrk (Ω) ≤ c(n, r, k, Ω) kf kWrk (Ω) + k∇gkWrk (Ω) .
∇2 v, ∇q ∈ L2 (B+ (τ ))
in R3+ ,
div v = g̃ = ϕg + v · ∇ϕ ∈ W21 (R3+ )
◦
By assumptions, u ∈ L12 (R3+ ) and p ∈ L2 (R3+ ) and thus we are in a position
to apply Proposition 2.1, which reads that ∇2 u, ∇p ∈ L2 (R3+ ) and
h i
k∇2 uk2,R3+ + k∇pk2,R3+ ≤ c k∇g̃k2,R3+ + kf˜k2,R3+ .
satisfying
−4v + ∇q = f
in B.
div v = g
∇2 v, ∇q ∈ L2 (B(τ ))
Step 1 Let v̄ = v − [v]B and fix 1/2 < τ1 < 1. By previous results, see
Proposition 3.2,
Z hZ Z Z i
2 2 2
(|∇ v| + |∇q| )dx ≤ c(τ1 , n) |v̄| dx + |∇v| dx + |q|2 dx .
2 2
B(τ1 ) B B B
B B
B(τ1 ) B
Step 2 Now, obviously, functions v,k and q,k obey the system
−4v,k + ∇q,k = 0
in B(τ1 )
div v,k = 0
As a result,
Z l
X l−1
X
i 2
|∇ v| + |∇i q|2 dx ≤ c(l, n)I.
i=1 i=1
B(1/2)
The proof of the following statement is slightly more complicated but still
can be made along similar lines.
48 CHAPTER 2. LINEAR STATIONARY PROBLEM
Proposition 4.2. Assume that a divergence free vector field v ∈ W21 (B+ )
satisfies the boundary condition
v|x3 =0 = 0
Fix 1/2 < τ1 < 1. By Proposition 3.1, we have additional regularity so that
Z Z h i
2 2 2
(|∇ v| + |∇q| )dx ≤ c(τ1 , n) |v|2 + |∇v|2 + |q|2 dx.
B+ (τ1 ) B+
B+ B+
and, by (2.4.2),
Z Z
2 2 2
(|∇ v| + |∇q| )dx ≤ c(τ1 , n) |∇v|2 dx ≡ cI.
B+ (τ1 ) B+
2.4. FURTHER LOCAL REGULARITY RESULTS, N = 2, 3 49
and
v,α |x3 =0 = 0.
Assume that n = 3 (n = 2 is an exercise). We have for 1/2 < τ2 < τ1
Z Z
(|∇ v,α | + |∇q,α | )dx ≤ c(τ1 , n) |∇v|2 dx ≡ cI.
2 2 2
B+ (τ2 ) B+
It remains to evaluate vi,333 and q,33 . To this end, we are going to exploit
the incompressibility condition: v3,333 = −vα,α33 ∈ L2 (B+ (τ2 )), which gives
us the bound Z
|∇3 v3 |2 dx ≤ cI.
B+ (τ2 )
Now, we are going to use the equations −4vα,3 + q,α3 = 0 one more time and
find
vα,333 = −vα,ββ3 + qα3 ∈ L2 (B+ (τ2 )).
The latter implies Z
|vα,333 |2 dx ≤ cI.
B+ (τ2 )
u|∂Ω = 0
can be transformed in the following way
◦
−4u + ∇p1 = f1 ∈ J (Ω)
in Ω,
div u = 0
u|∂Ω = 0,
where f1 = P f and p1 = p − q. So, without loss of generality, we always may
◦
assume that the right hand side in the Stokes system belongs to J (Ω).
We know that
k∇2 uk2,Ω + k∇pk2,Ω ≤ ckf k2,Ω .
We can also re-write the Dirichlet problem in the operator form
4u
e = f,
where ◦ ◦
4
e := P 4 : J (Ω) → J (Ω)
◦ ◦
e : J 1 (Ω) ∩ W 2 (Ω) → J (Ω)
4 2 2
◦
is a bijection. Given u ∈ J 12 (Ω) ∩ W22 (Ω), we have
(−4u,
e v) = (−4u + ∇p, v) = (−4u, v) = (∇u, ∇v)
∞
for any v ∈ C0,0 (Ω). From the latter identity, we immediately derive the
following estimate
k4uk
e ◦ ≤ k∇uk2,Ω = kuk ◦ 1 .
(J 12 (Ω))0 J 2 (Ω)
52 CHAPTER 2. LINEAR STATIONARY PROBLEM
◦
Here, we use the identification of the dual space (J 12 (Ω))0 described in Section
◦ ◦
4 of Chapter I with V = J 12 (Ω) and H = J (Ω) and in what follows we are
not going to use any special notation for this particular identification. Since
◦ ◦
the space J 12 (Ω) ∩ W22 (Ω) is dense in J 12 (Ω), there exists a unique extension of
◦ ◦
the Stokes operator 4 e (denoted again by 4) e from J 1 (Ω) ∩ W 2 (Ω) to J 1 (Ω).
2 2 2
Moreover, we have the following statement:
◦ ◦
e : J 1 (Ω) → (J 1 (Ω))0 is a bijection.
Proposition 5.1. (i) The extension 4 2 2
◦
(ii) If f ∈ (J 12 (Ω))0 , then
∞
X
kf k2◦ 1 = fk2 /λk ,
(J 2 (Ω))0
k=1
where fk = (f, ϕk ).
◦ ◦
e : J 1 (Ω) → 4(
Proof of Proposition 5.1 Obviously, 4 e J 1 (Ω)) is a
2 2
bijection. Our aim is to show that
◦ ◦
e J 1 (Ω)) = (J 1 (Ω))0 .
4( (2.5.1)
2 2
◦
Lemma 5.2. (i) for f ∈ (J 12 (Ω))0 , we have
∞
X
2
kf k ◦ ≤ fk2 /λk .
(J 12 (Ω))0
k=1
(ii) if
∞
X
fk2 /λk < ∞,
k=1
P∞ ◦ ◦
then the series k=1 fk ϕk converges to f in (J 12 (Ω))0 , f ∈ 4(
e J 1 (Ω)), and
2
∞
X
2
kf k ◦ = fk2 /λk .
(J 12 (Ω))0
k=1
2.5. STOKES OPERATOR IN BOUNDED DOMAINS 53
◦
Proof Fix an arbitrary function a ∈ J 12 (Ω)), then
N
X
N
a = ak ϕk → a
k=1
◦
in J 12 (Ω)). So,
N
X
N N
(f, a) = lim (f, a ) = lim (f, a ) = lim f k ak ≤
N →∞ N →∞ N →∞
k=1
N
X N
21 X 12 ∞
X 12
≤ fk2 /λk a2k λk ≤ fk2 /λk k∇ak2,Ω .
k=1 k=1 k=1
N
X
2
kfN − fM k ◦ ≤ fk2 /λk → 0
(J 12 (Ω))0
k=M +1
as M, N → 0.
◦
We denote by f ∈ (J 12 (Ω))0 the sum of our series. Then, by (i),
∞
X
kf − fN k2◦ 1 ≤ fk2 /λk → 0
(J 2 (Ω))0
k=N +1
and thus
kfN k ◦ → kf k ◦ .
(J 12 (Ω))0 (J 12 (Ω))0
◦
e J 1 (Ω)). Indeed, we have
Now, we are going to prove that f ∈ 4( 2
N
X N
X
fN = fk ϕk = 4
e fk ϕk /λk = 4u
e N,
k=1 k=1
where
N
X ◦
uN = fk ϕk /λk ∈ J 12 (Ω) ∩ W22 (Ω).
k=1
54 CHAPTER 2. LINEAR STATIONARY PROBLEM
By direct computations,
N
X
k∇uN − ∇uM k22,Ω = fk2 /λk → 0.
k=M +1
4u
e N → 4u
e = f.
Next, we have
N
X
kfN k2◦ 1 e N k2 ◦
= k4u 1
= k∇uN k22,Ω = fk2 /λk → kf k2◦ 1 .
(J 2 (Ω))0 (J 2 (Ω))0 (J 2 (Ω))0
k=1
Lemma 5.3.
∞
◦ ◦ X
e J 1 (Ω)) = {f ∈ (J 1 (Ω))0 :
4( fk2 /λk < ∞} =: U.
2 2
k=1
◦ ◦
e J 1 (Ω)), i.e., f = 4
Now, assume that f ⊆ 4( e u for some u ∈ J 1 (Ω). Then
2 2
we have
fk = (f, ϕk ) = (4e u, ϕk ) = (u, 4
e ϕk ) = λk uk .
Since ∞
X
k∇uk22,Ω = u2k λk < ∞,
k=1
we find ∞
X
fk2 /λk < ∞.
k=1
◦
e J 1 (Ω)).
So, f ∈ U and thus U ∈ 4( 2
Now, we proceed with the proof of Proposition 5.1. We are done, if show
that ∞
◦ X
1 0
f ∈ (J 2 (Ω)) ⇒ fk2 /λk < ∞.
k=1
2.6. COMMENTS 55
Next, we let
N
X
N
ak = fk /λk , a = ak ϕk .
k=1
Then
N
X N
X
k∇aN k22,Ω = |ak | 2
k∇ϕk k22,Ω = fk2 /λk .
k=1 k=1
Next, we have
N
X N
X 21
N
(f, a ) = fk2 /λk ≤ kf k ◦
N
k∇a k2,Ω = kf k ◦ fk2 /λk
(J 12 (Ω))
0 (J 12 (Ω))0
k=1 k=1
which implies
N
X
fk2 /λk ≤ kf k2◦ 1
(J 2 (Ω))0
k=1
2.6 Comments
Chapter 2 contains standard results on linear stationary Stokes system in-
cluding the notion of Stokes operator in smooth bounded domains. In ad-
dition, various global and local interior and boundary regularity results are
discussed.
56 CHAPTER 2. LINEAR STATIONARY PROBLEM
Chapter 3
u|∂Ω = 0 (3.1.2)
and if n = 3 and Ω is unbounded then u(x) → 0 as |x| → ∞. Here, ν is a
positive parameter called viscosity. We always assume that
f ∈ L−1
2 (Ω).
57
58 CHAPTER 3. NON-LINEAR STATIONARY PROBLEM
u ∈ L4 (Ω).
Proof Let us reduce our boundary value problem to a fixed point prob-
lem and try to apply the celebrated Leray-Schauder principle.
u = λB(u)
div w × w ∈ L−1
2 (Ω).
According to our results in Chapter II, given w ∈ V (Ω), there exists a unique
u ∈ V (Ω) such that
[A(w), v] := (w × w, ∇v)
and
[F, v] := (f, v).
3.1. EXISTENCE OF WEAK SOLUTIONS 59
w(k) * w
in V (Ω). Then, the compactness of the imbedding of V (Ω) into L4 (Ω) gives
us:
w(k) ⊗ w(k) → w ⊗ w
in L2 (Ω). From the main identity, it follows that
for any v ∈ V (Ω). It remains to plaque v = u(k) −u(m) into the latter relation
and make use of the fact that
νuλ = λA(uλ ) + F,
and thus
1
k∇uλ k2,Ω ≤kf kL−1
2 (Ω)
.
ν
The right hand side of the latter inequality is independent of λ and thus
the existence of at least one fixed point follows from the Leray-Schauder
principle.
Regarding the uniqueness of weak solutions, we have the following
60 CHAPTER 3. NON-LINEAR STATIONARY PROBLEM
Lemma 1.4. Assume that all assumptions of Proposition 1.2 hold. Let in
addition
c20 (n, Ω)|
kf kL−1
2 (Ω)
< 1,
ν2
where c0 (n, Ω) is a constant in the inequality
1
k∇u1 k2,Ω ≤ kf kL−1 (Ω),
ν 2
we find
νk∇(u1 − u2 )k22,Ω ≤ c20 k∇u1 k22,Ω k∇(u1 − u2 )k22,Ω ≤
c20
≤ k∇(u1 − u2 )k22,Ω kf kL−1 (Ω).
ν 2
1
k∇uk2,Ω ≤ kf kL−1 (Ω).
ν 2
3.1. EXISTENCE OF WEAK SOLUTIONS 61
for any v ∈ C0∞ (Ωm ). Moreover, we can fix pm so that pm = pm+1 in Ωm . So,
now, if we introduce a function p, letting p = pm in Ωm , then p ∈ L2,loc (Ω)
and the following identity is valid:
m+s 1
≤ τk 2 Φ(R0 ) + Cτ s R0s
m−s .
1−τ 2
Given 0 < % ≤ R0 , we find an integer number k such that
R0 τ k+1 < % ≤ R0 τ k .
Then
1 % s % s 1
k
Φ(%) ≤ Φ(τ R0 ) ≤ Φ(R0 ) + C .
τR τ 1 − τ m−s
2
Lemma 2.2. Let a divergence free vector-valued function u ∈ W21 (B(R)) and
a tensor-valued function F ∈ Lr (B(R)), with r > n = 3, satisfy the identity
Z Z Z
∇u : ∇vdx = u ⊗ u : ∇vdx + F : ∇vdx
B(R) B(R) B(R)
∞
for any v ∈ C0,0 (B(R)).
Then,
Z % 3 Z 31 Z
2 6
|∇u| dx ≤ c +R |u| dx |∇u|2 dx+
R
B(%) B(R) B(R)
2
Z r2
+cR3(1− r ) |F |r dx
B(R)
∞
for any v ∈ C0,0 (B+ (R)).
Then,
Z % 3 Z 31 Z
2 6
|∇u| dx ≤ c +R |u| dx |∇u|2 dx+
R
B+ (%) B+ (R) B+ (R)
64 CHAPTER 3. NON-LINEAR STATIONARY PROBLEM
Z r2
3(1− r2 ) r
+cR |F | dx
B+ (R)
Multiplying the first equation in (3.2.3) by ũR and integrating the product
by parts, we find
Z Z Z
2
|∇ũR | dx = (u ⊗ u − [u ⊗ u]B(R) ) : ∇ũR dx + F : ũR dx
B(R) B(R) B(R)
Next, we estimate the first term on the right hand side of the latter relation
with the help of Galliardo-Nirenberg inequality and, then, with the help of
Hölder inequality. As a result, we have
Z Z 6
35
2
|u ⊗ u − [u ⊗ u]B(R) | dx ≤ c |∇(u ⊗ u)| dx ≤
5
B(R) B(R)
Z 6 6
53 Z 31 Z 3
43
6
≤c |u| |∇u| dx
5 5 ≤c |u| dx |∇u| dx2 ≤
B(R) B(R) B(R)
Z 31 Z
≤c |u|6 dx R |∇u|2 dx,
B(R) B(R)
3.2. REGULARITY OF WEAK SOLUTIONS 65
∞
for any v ∈ C0,0 (B(R)). By the results of Chapter II, see Section 2, we have
the following estimate
Z % 3 Z
2
|∇uR | dx ≤ c |∇uR |2 dx,
R
B(%) B(R)
and, then, Hölder’s inequality for the term, containing F , in order to get the
estimate of Lemma 2.2.
Lemma 2.4. (Ch.-B. Morrey) Let u ∈ Wm1 (Ω) satisfy the condition
Z
|∇u|2 dx ≤ K%n−m+mα
B(%)
for some 0 < α < 1 and for any B(x0 , %) ⊂ Ω such that 0 < % < %0 with two
positive constants K and %0 .
α
Then u ∈ Cloc (Ω), i.e., u ∈ C α (Ω1 ) for any subdomain Ω1 b Ω.
Lemma 2.5. Assume that all assumptions of Lemma 2.2 hold with R = a.
Then
1− 3
u ∈ Cloc r (B(a)).
66 CHAPTER 3. NON-LINEAR STATIONARY PROBLEM
Lemma 2.6. Assume that all assumptions of Lemma 2.3 hold with R = a.
Then
u ∈ C α (B̄+ (b))
for any 0 < b < a with α = 1 − 3/r.
Proof We have two types of estimates. The first one is so-called ”inte-
rior”. For any b1 ∈]b, a[, the following estimate is valid:
for x0 = (x00 , 0), |x00 | < 21 (a − b1 ), 0 < % ≤ R0 , and K+ depends on the same
arguments as K.
Now, let us denote by ũ extension of u to the whole ball B(a) by zero
and let Z
Φ̃(x0 , %) := |∇ũ|2 dx
B(a)
≤ K+ %1+2α .
So, the statement of the lemma follows from Morrey’s condition on Hölder
continuity, see Lemma 2.4.
68 CHAPTER 3. NON-LINEAR STATIONARY PROBLEM
∞
for any v ∈ C0,0 (B(2a)). If f is of class C ∞ in B(2a), then u is of class C ∞
in B(a).
Proof It is not so difficult to check that there exists a tensor-valued
function F of class C ∞ such that f = −div F . Then, the identity from the
statement of the proposition can be re-written in the following way
Z Z Z
∇u : ∇vdx = u ⊗ u : ∇v + F : ∇vdx
B(2a) B(2a) B(2a)
∞
for any v ∈ C0,0 (B(2a)). From Lemma 2.5, it follows that u belongs, at least,
to C(B(3a/2)). Using the same arguments as in Section 1, we can recover a
pressure p ∈ L2 (B(2a)) (exercise) so that
−4u + ∇p = −div G := −div (u ⊗ u + F )
in B(2a).
div u = 0
Since ∇2 G ∈ L2 (B(a1 )), we can use the linear theory one more time and get:
for any a < a3 < a2 . Proceeding, further, in the same way, we complete the
proof of the lemma.
∞
for any v ∈ C0,0 (B+ (2a)). If f is of class C ∞ in B(2a) ∩ {x3 ≥ 0}, then u is
∞
of class C in B(a) ∩ {x3 ≥ 0}.
Proof We start with our proof in a similar way as in the latter propo-
sition, i.e., we find F of class C ∞ in B(2a) ∩ {x3 ≥ 0} so that f = −div F .
Then, we recover the pressure p ∈ L2 (B+ (2a)), which gives us:
−4u + ∇p = −div G := −div (u ⊗ u + F )
in B+ (2a),
div u = 0
u|x3 =0 = 0.
By Lemma 2.6, u ∈ C(B + (3a/2)) and, by the linear theory,
Next, for α = 1, 2 and for k = 1, 2, 3, we have functions u,k ∈ W21 (B+ (a1 ))
and p,k ∈ L2 (B+ (a1 )) satisfying the system
−4u,k + ∇p,k = −div G,k
in B+ (a1 )
div u,k = 0
u,α |x3 =0 = 0,
where ∇2 G ∈ L2 (B+ (a1 )). Then, again, we apply the linear theory and
conclude that
for any a < a2 < a1 . We need to establish the same properties for ∇2 u,3
and ∇p,3 . To achieve this goal, it is sufficient to evaluate uk,333 and p,33 for
k = 1, 2, 3, which is, in fact, not so difficult. Indeed, denoting gik := −Gij,jk ,
we first use the incompressibility condition:
and
uα,333 = −gα3 + p,α3 − uα,ββ3 ∈ L2 (B+ (a2 )).
So, we can state
u,αβ |x3 =0 = 0,
where ∇3 G ∈ L2 (B+ (a2 )). Here, we are going to proceed as in the case of
the third derivatives. We let
hijk = −Gim,mjk .
So,
∇3 u3,α ∈ L2 (B+ (a3 )).
Then,
p,33α = h33α + u3,jj3 ∈ L2 (B+ (a3 ))
and thus
∇2 p,α ∈ L2 (B+ (a3 )).
Next,
uβ,333α = −uβ,γγ3α + p,β3α − hβα3 ∈ L2 (B+ (a3 )).
So, we have
∇3 u,α ∈ L2 (B+ (a3 )).
Now, let us go back to the incompressibility condition:
Finally,
3.3 Comments
Chapter 3 contains standard results on the existence and regularity of solu-
tions to non-linear stationary boundary value problem. The main point of
the chapter is the local regularity theory, which differs a bit from the theory
for standard elliptic systems.
Chapter 4
Zb
< u∗ , v > (χ) = − < v ∗ (·, t), v(·) > ∂t χ(t)dt
a
73
74 CHAPTER 4. LINEAR NON-STATIONARY PROBLEM
As usual, the left hand side of the above identity is written in the same
way as the right hand side, i.e.,
Zb Zb
< ∂t v ∗ (·, t), v(·) > χ(t)dt = − < v ∗ (·, t), v(·) > ∂t χ(t)dt
a a
for any v ∈ V and for any χ ∈ C0∞ (a, b), although the left hand side might
make no sense as Lebesgue’s integral.
Let us discuss the relationship between the introduced notion and and
the Sobolev derivatives. Assume that
Z
∗
V, V ∈ L1,loc (Ω), C0∞ (Ω) ⊂ V, ∗
< v , v >= v ∗ vdx,
Ω
∗
v ∈ L1,loc (a, b; L1,loc (Ω)) = L1,loc (Ω×]a, b[), (4.1.1)
∂t v ∗ ∈ L1,loc (Ω×]a, b[).
Lemma 1.2. Given ε > 0 and ϕ ∈ C0∞ (Ω×]a, b[), there exist positive integer
number N and functions ϕk ∈ C0∞ (Ω), χk ∈ C0∞ (a, b), k = 1, 2, ..., N such
that
N
X
kϕ − ϕk χk kC 1 (Ω×[a,b]) < ε.
k=1
Let us assume that Lemma 1.2 has been proved. Suppose that ∂t v ∗ is
the derivative in the sense of Definition 1.1 and satisfies assumptions (4.1.1).
Our aim is to show that it is Sobolev’s derivative as well. Take an arbitrary
ε > 0 and an arbitrary function ϕ ∈ C0∞ (Ω×]a, b[) and fix them. Clearly,
ϕ ∈ C0∞ (Ω0 ×]a0 , b0 [) for some Ω0 b Ω and for some a < a0 and b0 < b. Let
number N (ε) and functions ϕk and χk be from Lemma 1.2 for the domain
4.1. DERIVATIVE IN TIME 75
a0 Ω0 a0 Ω0
b0 N (ε)
Z Z X
∗
≤ v (x, t) ∂t ϕ(x, t) − ∂t ϕk (x)χk (t) dxdt+
a0 Ω0 k=1
b0 N (ε)
Z Z X
∗
+ ∂t v (x, t) ϕ(x, t) − ϕk (x)χk (t) dxdt ≤
a0 Ω0 k=1
N (ε)
X
≤ ckϕ − ϕk χk kC 1 (Ω0 ×[a0 ,b0 ]) kv ∗ kL1 (Ω0 ×]a0 ,b0 [) + k∂t v ∗ kL1 (Ω0 ×]a0 ,b0 [)
k=1
≤ cε kv ∗ kL1 (Ω0 ×]a0 ,b0 [) + k∂t v ∗ kL1 (Ω0 ×]a0 ,b0 [) .
where Z
1 x·m
cm (t) = ϕ(x, t)e−iπ l dx.
(2l)n
Cl
76 CHAPTER 4. LINEAR NON-STATIONARY PROBLEM
The Fourier series converges very well. So, after taking real and imaginary
parts, given ε > 0, we find the number N (ε) such that
PN (ε)
where ΦN (ε) (x, t) = k=1 ϕk (x)χk (t). Assume that there exist functions
∞ ∞
ϕ0 ∈ C0 (Ω), χ0 ∈ C0 (]a, b[) with the following property
Φ
e N (ε) = ΦN (ε) ϕ0 χ0
and show
Let t1 = inf t and t2 = sup t. We claim that a < t1 ≤ t2 < b. Assume that
t∈Λ t∈Λ
t2 = b. Then, by the definition, there exists a sequence (xk , tk ) ∈ supp ϕ
with tk → b as k → ∞. Selecting if necessary a subsequence, we have a
contradiction for the limit point (x, b) ∈ supp ϕ. Now, let us show that
Assume that ◦
a ∈J (Ω). (4.2.2)
This problem can be written in the operator form
◦
e = f ∈ L2 (0, T ; (J 12 (Ω))0 ),
∂t u − ∆u
◦
u|t=0 = a ∈J (Ω), (4.2.3)
see notation for the Stokes operator ∆e and for the dual space in the last
section of Chapter 2.
Our task is to construct an explicit solution provided eigenvalues and
eigenfunctions of the Stokes operator ∆
e in the domain Ω are known. So, we
have
−∆ϕ
e k = λk ϕk in Ω,
ϕk = 0 on Ω, (4.2.4)
where k = 1, 2, ....
First, we expand functions f and a, using eigenfunctions and eigenvalues
of the Stokes operator,
∞
X
f (x, t) = fk (t)ϕk (x),
k=1
where
fk (t) = (f (·, t), ϕk (·))
and ∞
X
a(x) = ak ϕk (x), ak = (a, ϕk ).
k=1
78 CHAPTER 4. LINEAR NON-STATIONARY PROBLEM
By our assumptions,
ZT X
∞
2 1
kf k ◦ = |fk (t)|2 dt,
L2 (0,T ;(J 12 (Ω))0 )
k=1
λk
0
∞
X
kak22,Ω = a2k . (4.2.5)
k=1
∞
X
u(x, t) = ck (t)ϕk (x). (4.2.6)
k=1
Assume that
ck (0) = ak , k = 1, 2, .... (4.2.7)
Our further calculations are going to be formal. Later on, we will explain
in what sense the formal solution is a solution to problem (4.2.3). So, if we
plague (4.2.6) into (4.2.3), then the identity
∞
X ∞
X
c0k ϕk + λk ck ϕk = fk ϕk
k=1 k=1
Zt
−λk t
ck (t) = e ak + eλk τ fk (τ )dτ . (4.2.9)
0
So, we have got a formal solution of form (4.2.6). Let us analyze its
4.2. EXPLICIT SOLUTION 79
Zt 2
2 −2λk t 2 −λk (t−τ )
ck (t) ≤ 2e ak + 2 e fk (τ )dτ
0
Zt Zt
≤ 2e−2λk t a2k + 2 e−2λk (t−τ ) dτ fk2 (τ )dτ
0 0
Zt Zt
1 1 −2λk t
≤ 2e−2λk t a2k + fk2 (τ )dτ − e fk2 (τ )dτ.
λk λk
0 0
So, finally,
Zt
1
c2k (t) ≤ 2e−2λk t a2k + fk2 (τ )dτ. (4.2.10)
λk
0
∞ ∞ ∞ Zt
X
−2λ1 t
X X 1
ku(·, t)k22 = c2k (t) ≤ 2e a2k + fk2 (τ )dτ
k=1 k=1 k=1
λk
0
−2λ1 t
≤ 2e kak22 + kf k 2
◦ (4.2.11)
L2 (0,T ;(J 12 (Ω))0 )
or
So,
fk2 (t)
(c2k (t))0 + λk c2k (t) ≤ .
λk
80 CHAPTER 4. LINEAR NON-STATIONARY PROBLEM
ZT ZT
fk2 (t)
c2k (T ) + λk c2k (t)dt ≤ c2k (0) + dt
λk
0 0
ZT
fk2 (t)
= a2k + dt.
λk
0
ZT X
∞
k∇uk2L2 (0,T ;L2 (Ω)) = λk c2k (t)dt
0 k=1
≤ kak22 + kf k2 ◦ . (4.2.13)
L2 (0,T ;(J 12 (Ω))0 )
ZT X
∞
2 |c0k (t)|2
k∂t uk ◦ = dt
L2 (0,T ;(J 12 (Ω))0 )
k=1
λk
0
ZT X
∞ ZT X
∞
fk2 (t)
≤2 λk c2k (t)dt +2 dt
k=1
λk
0 0 k=1
≤ 2k∇uk2L2 (0,T ;L2 (Ω)) + 2kf k 2
◦ .
L2 (0,T ;(J 12 (Ω))0 )
Now, we wish to figure out in which sense (4.2.3) holds. Let us take an
◦
arbitrary function w ∈ L2 (0, T ; J 12 (Ω)) and expend it as a Fourier series
∞
X
w(x, t) = dk (t)ϕk (x).
k=1
4.2. EXPLICIT SOLUTION 81
Obviously,
ZT X
∞
k∇wk2L2 (0,T ;L2 (Ω)) = λk d2k (t)dt < ∞.
0 k=1
Hence,
ZT Z ZT X
∞
∂t u · wdxdt = c0k (t)dk (t)dt,
0 Ω 0 k=1
ZT Z ZT Z X
∞
∇u : ∇wdxdt = ck (t)dk (t)|∇ϕk |2 dxdt =
0 Ω 0 Ω k=1
ZT X
∞
= λk ck (t)dk (t)dt,
0 k=1
ZT Z ZT X
∞
f · wdxdt = fk (t)dk (t)dt,
0 Ω 0 k=1
and, by (4.2.8),
ZT Z
∂t u · w + ∇u : ∇w − f · w dxdt =
0 Ω
ZT X
∞
= c0k (t) + λk ck (t) − fk (t) dk (t)dt = 0.
0 k=1
◦
Taking w(x, t) = χ(t)v(x) with v ∈ J 12 (Ω) and χ ∈ C01 (0, T ), we get that, for
a.a. ∈]0, T [, the identity
Z Z
∂t u(x, t) · v(x) + ∇u(x, t) : ∇v(x) dx = f (x, t) · v(x)dx (4.2.15)
Ω Ω
◦
holds for all v ∈ J 12 (Ω). To be more precise, (4.2.15) is fulfilled at all
Lebesgue’s points of the following functions t 7→ ∂t u(·, t), t 7→ ∇u(·, t), and
82 CHAPTER 4. LINEAR NON-STATIONARY PROBLEM
t 7→ f (·, t). Identity (4.2.15) is called the weak form of the first equation in
(4.2.3).
It remains to establish in what sense the initial data in (4.2.3) are satisfied.
◦
Lemma 2.1. Function t 7→ u(·, t) ∈ (J 12 (Ω))0 can be modified on a zero-
◦
measure subset of [0, T ] so that, for each v ∈ J 12 (Ω), the function
Z
t 7→ u(·, t) · v(·)dx
Ω
is continuous on [0, T ].
◦
Proof Since u ∈ L2 (0, T ; (J 12 (Ω))0 ), a.a. points t0 ∈ [0, T ] are Lebesgue’s
points of t 7→ u(·, t) in the following sense
tZ0 +ε
1
ku(·, t) − u(·, t0 )k ◦ dt → 0
2ε (J 12 (Ω))0
t0 −ε
as ε → 0.
Denote by S the set of al Lebesgue’s points of t 7→ u(·, t). We know that
|S| = T . Let t0 < t1 be two points from S. By the definition of the derivative
∂t u,
ZT Z ZT Z
∂t u(x, t) · v(x)χ(t)dxdt = − u(x, t) · v(x)∂t χ(t)dxdt
0 Ω 0 Ω
◦
for any v ∈ J 12 (Ω) and for any χ ∈ C01 (0, T ). We can easily extend the latter
◦
identity to functions χ ∈ W 12 (0, T ). Take function χ = χε so that χε (t) = 0
if 0 < t ≤ t0 − ε or t1 + ε ≤ t < T , χε (t) = 1 if t0 + ε ≤ t ≤ t1 − ε,
χε (t) = (t − t0 + ε)/(2ε) if t0 − ε < t < t0 + ε, and χε (t) = (t1 + ε − t)/(2ε)
if t1 − ε < t < t1 + ε. Then, we have
ZT Z tZ1 +εZ
1
∂t u(x, t) · v(x)χε (t)dxdt = u(x, t) · v(x)dxdt
2ε
0 Ω t1 −ε Ω
tZ0 +εZ
1
− u(x, t) · v(x)dxdt. (4.2.16)
2ε
t0 −ε Ω
4.2. EXPLICIT SOLUTION 83
Obviously,
1 tZ0 +εZ
u(x, t) − u(x, t0 ) · v(x)dxdt
2ε
t0 −ε Ω
tZ0 +ε
1
≤ ku(·, t) − u(·, t0 )k ◦ dtkv(·)k ◦ 1 →0
2ε (J 12 (Ω))0 ) J 2 (Ω)
t0 −ε
for a.a. t1 ∈ [0, T ]. Since the right-hand side of the latter identity is a
continuous function with respect to t1 , the left-hand side is continuous in t1
as well.
Now, coming back to our function u, we note that u ∈ L∞ (0, T ; L2 (Ω)).
Therefore, we can state that
Z
t 7→ u(x, t) · v(x)dx is continuous in t on [0, T ]
Ω
◦
for each function v ∈J (Ω) and even for each function v ∈ L2 (Ω). The
latter follows from the fact that for any v ∈ L2 (Ω) we have Helmoholtz-
Weyl decomposition in the Ladyzhenskaya form so that v = v1 + ∇p, where
◦
v1 ∈J (Ω) and p ∈ W21 (Ω). Moreover,
Z Z
u(x, t) · v(x)dx = u(x, t) · v1 (x)dx,
Ω Ω
◦
since u(·, t) ∈J (Ω).
So, our initial data are satisfied at least in the following sense. Since
u(·, 0) = a(·) by construction u, we have
Z Z
lim u(x, t) · v(x)dx = a(x) · v(x)dx
t→+0
Ω Ω
where ∞
X
u(x, t) = dk (t)ϕk (x),
k=1
and
ZT X
∞
2 1 0
k∂t uk ◦ = (d (t))2 dt.
L2 (0,T ;(J 12 (Ω))0 )
k=1
λk k
0
In a view of Lemma 2.1, it is sufficient to show that the function t 7→
ku(·, t)k2,Ω is continuous. We know that functions t 7→ dk (t) are continuous
on [0, T ]. Therefore, the function t 7→ gN (t) = N 2
P
k=1 dk (t) is continuous on
2
[0, T ] as well. We know also that gN (t) → ku(·, t)k2,Ω as N → ∞. So, we need
to show that the sequence gN (t) is uniformly bounded and the convergence
is uniform.
First, we show uniform boundedness. We have
Zt X
N
gN (t) − gN (t1 ) = 2 d0k (τ )dk (τ )dτ (4.2.18)
t1 k=1
ZT
1
gN (t) ≤ gN (t1 )dt1 + 2k∇ukL2 (0,T ;L2 (Ω)) k∂t uk ◦
T L2 (0,T ;(J 12 (Ω))0 )
0
1
≤ k∇uk2L2 (0,T ;L2 (Ω)) + 2k∇ukL2 (0,T ;L2 (Ω)) k∂t uk ◦ .
T λ1 L2 (0,T ;(J 12 (Ω))0 )
Zt 12
2
|gN (t) − gN (t1 )| ≤ 2 k∇u(·, t)k2,Ω dt k∂t uk ◦ .
L2 (0,T ;(J 12 (Ω))0 )
t1
Now, gN (t) converges to ku(·, t)k22,Ω uniformly, which means that the function
t 7→ ku(·, t)k22,Ω is continuous on [0, T ].
At last, (4.2.17) follows directly from (4.2.18) if N → +∞.
Actually, the analogue of Theorem 2.2 takes places. Moreover, an abstract
version of it is valid:
V = [S]V , H = [S]H .
Zt
kv(·, t)k2H − kv(·, t1 )k2H =2 (∂t v(·, τ ), v(·, τ ))H dτ
t1
Proof We start with general facts. Let t 7→ v(·, t), where v ∈ Lp (0, T ; V ).
We extend v by zero outside [0, T ]. The first fact is the integral continuity:
for any ε > 0, there is a number δ(ε) > 0 such that
ZT
kv(·, t + h) − v(·, t)kpV dt < ε
0
whenever |h| < δ(ε). This property provides the following. Let
ZT
vε (·, t) = ωε (t − τ )v(·, τ )dτ,
0
ZT
kvε (·, t) − v(·, t)kpV dt ≤
0
Z∞
≤ kvε (·, t) − v(·, t)kpV dt =
−∞
Z∞ Z∞
= k ωε (t − τ )(v(·, τ ) − v(·, t))dτ kpV dt ≤
−∞ −∞
Z∞ Z∞
≤ ωε (t − τ )kv(·, τ ) − v(·, t))kpV dτ dt =
−∞ −∞
Z Z∞
∞
∂t vε → ∂t v in Lp,loc (0, T ; V ∗ ).
4.2. EXPLICIT SOLUTION 87
Further, we can use the same trick as in the case of star-shaped domains.
Without loss of generality, we may replace the interval ]0, T [ with ] − 1, 1[.
We take λ > 1 and define
t
v λ (·, t) = v(·, ), |t| ≤ λ
λ
and thus
1
∂t v λ (·, t) = ∂s v(·, s)|s= λt .
λ
Here, the crucial things are as follows:
kv λ − vkLp (−1,1;V ) + kv λ − vkL2 (−1,1;H) → 0
and
k∂t v λ − ∂t vkLp0 (−1,1;V ∗ ) → 0
as λ → 1. Moreover, for fixed λ > 1,
kv λ − (v λ )ε kLp (−1,1;V ) + kv λ − (v λ )ε kL2 (−1,1;H) → 0
and
k∂t v λ − ∂t (v λ )ε kLp0 (−1,1;V ∗ ) → 0
as ε → 0. Summarizing these two properties, we may construct a sequence
v (k) that is differentiable in t and satisfies:
kv (k) − vkLp (−1,1;V ) + kv (k) − vkL2 (−1,1;H) → 0
and
k∂t v (k) − ∂t vkLp0 (−1,1;V ∗ ) → 0
as k → ∞.
Now, let u = v (k) − v (m) , we have the identity
Z t
2
ku(·, t)kH = 2 (∂t u(·, τ ), u(·, τ )H dτ + ku(·, t1 )k2H , (4.2.19)
t1
◦ ◦
Theorem 2.4. Assume that a ∈J (Ω) and f ∈ L2 (0, T ; (J 12 (Ω))0 ). There
exists a unique function u called weak solution to (4.2.1) such that:
◦ ◦
u ∈ L2 (0, T ; J 12 (Ω)), ∂t u ∈ L2 (0, T ; (J 12 (Ω))0 ); (4.2.20)
for a.a. t ∈ [0, T ],
Z h i Z
∂t u(x, t) · v(x) + ∇u(x, t) · ∇v(x) dx = f (x, t) · v(x)dx (4.2.21)
Ω Ω
◦
for any v ∈ J 12 (Ω);
u(·, 0) = a(·) (4.2.22)
and (4.2.22) is fulfilled in the L2 -sense, i.e., ku(·, t)−a(·)k2,Ω → 0 as t → +0.
Moreover,
u ∈ C([0, T ]; L2 (Ω)).
Proof Existence has been already proven. It remains to show unique-
ness. Assume that u1 is another solution satisfying (4.2.20)–(4.2.22). Then
for w = u − u1 we have
Z Z
∂t w(x, t) · v(x)dx + ∇w(x, t) : ∇v(x)dx = 0
Ω Ω
◦
for a.a. t ∈ [0, T ] and for any v ∈ J 12 (Ω) and thus
Z Z
∂t w(x, t) · w(x, t)dx + |∇w(x, t)|2 dx = 0.
Ω Ω
Assume that ◦ ◦
a ∈J ≡J (Rn ). (4.3.2)
It is supposed also that
◦ ◦
f ∈ L2 (0, T ; (J 12 )0 ) := L2 (0, T ; (J 12 (Rn ))0 ), divf = 0. (4.3.3)
In this case, the Cauchy problem can be reduced to the Cauchy problem for
the heat equation
∂t u − ∆u = f in QT ,
u(x, 0) = a(x) x ∈ Rn . (4.3.4)
Indeed, assume that u is a solution to the Cauchy problem (4.3.4). Take the
divergence of equations in (4.3.4). Then we have
∂t divu − ∆divu = 0 in QT ,
divu(x, 0) = 0 x ∈ Rn .
By the unique solvability of the Cauchy problem for the heat equation, one
can claim that divu = 0 in QT . The pressure field is an arbitrary function of
t.
Solution to (4.3.4) can be given in an explicit form with the help of the
fundamental solution to the heat equation:
Z Zt Z
u(x, t) = Γ(x − y, t)a(x)dx + Γ(x − y, t − τ )f (y, τ )dydτ,
Rn 0 Rn
where
1 |x|2
− 4t
Γ(a, t) =
n e
(4πt) 2
n
for x ∈ R and t > 0. This formula is a good source for understanding
properties of solutions to (4.3.1).
Then,
ZT X
∞
2
kf k ◦ = kf k22,QT = fk2 (t)dt < ∞,
L2 (0,T ;J (Ω))
0 k=1
∞
X
k∇ak22,Ω = λk a2k < ∞. (4.4.4)
k=1
Next, let us multiply the first equation in (4.4.2) by ck , sum up the result
from 1 to N , integrate the sum in time over the interval ]0, t[, and find
N Zt X
N
1X
λk c2k (t) + |c0k (τ )|2 dτ
2 k=1 k=1
0
N Zt N
1X X
= λk a2k + fk (τ )c0k (τ )dτ,
2 k=1 k=1
0
4.4. PRESSURE FIELD. REGULARITY 91
N
X Zt X
N N
X Zt X
N
λk c2k (t) + |c0k (τ )|2 dτ ≤ λk a2k + fk2 (τ )dτ
k=1 0 k=1 k=1 0 k=1
≤ k∇ak22,Ω + kf k22,QT
Zt
k∇u(·, t)k22,Ω + k∂t u(·, τ )k22,Ω dτ ≤ k∇ak22,Ω + kf k22,QT (4.4.5)
0
◦
for any L12 (Ω). For it, we have the estimate
|lt (v)| ≤ k∇u(·, t)k2,Ω k∇vk2,Ω + (k∂t u(·, t)k2,Ω + kf (·, t)k2,Ω )kvk2,Ω .
So, we have
Z Z
∇u(x, t) : ∇v(x)dx − p(x, t)divv(x)dx
Ω
Z ZΩ
= f (x, t) · v(x)dx − ∂t u(x, t) · v(x)dx (4.4.8)
Ω Ω
◦
for any v ∈ L12 (Ω) and for a.a. t ∈ [0, T ].
For those t, i.e., for which (4.4.8) holds, we may apply the regularity
theory developed for the linear stationary Stokes system. More precisely,
one can estimate higher derivatives in spatial variables:
k∇2 u(·, t)k2,Ω + k∇p(·, t)k2,Ω ≤ c(kf (·, t)k2,Ω + k∂t u(·, t)k2,Ω
and thus
k∇2 uk2,QT + k∇pk2,QT ≤ c(kf k2,QT + k∂t uk2,QT ).
Combining the latter estimate with (4.4.5), we get the final bound:
∂t u − ∇u = f − ∇p, divu = 0
1,0
Ws,l (QT ) = {v ∈ Ll (0, T ; Ws1 (QT )},
4.4. PRESSURE FIELD. REGULARITY 93
and
Ws2,1 (QT ) = Ws,s
2,1
(QT ), Ws1,0 (QT ) = Ws,s
1,0
(QT ).
Proof of Theorem 4.1 We need to show ∇u ∈ C([0, T ]; L2 (Ω)). To
◦
this end, we are going to use Theorem 2.3. Introducing H = J 12 (Ω) with
◦
scalar product (u, v)H = (∇u, ∇v), V = J 12 (Ω) ∩ W22 (Ω) with the norm
◦
e 2,Ω , let us verify that V ∗ = J (Ω) is dual to V with respect H.
kvkV = k∆vk
Indeed, let l ∈ V ∗ . So, we have |l(v)| ≤ ck∆vk
e 2,Ω for any v ∈ V . Since
◦ ◦
e ) = J (Ω), for any p ∈ J (Ω), one can define G(p) = l(v), where p = −∆v.
∆(V e
Obviously, |G(p)| ≤ kpk2,Ω klk. By Riesz theorem, there exists a unique
◦
v ∗ ∈ J (Ω) such that Z
G(p) = v ∗ · pdx,
Ω
∗
kv k2,Ω = klk and
Z Z
∗
l(v) = − v · ∆(v)dx
e = ∇v ∗ : ∇vdx.
Ω Ω
◦
Now, any v ∗ from J (Ω) defines a linear functional on V by formula
Z Z
− v · ∆(v)dx = ∇v ∗ : ∇vdx ≤ kv ∗ k2,Ω k∆(v)k
∗ e e ∗
2,Ω ≤ kv k2,Ω kvkV .
Ω Ω
◦
It is easy to prove that in fact klk = kv ∗ k2,Ω . So, V ∗ ' J (Ω), i.e., spaces are
isometrically isomorphic.
By Theorem 2.3, ∇u ∈ C([0, T ]; L2 (Ω)). Theorem 4.1 is proved.
Theorem 4.2. Let Ω be a bounded domain with smooth boundary. Consider
the following initial boundary value problem
Let f ∈ Ls,l (QT ) := Ll (0, T ; Ls (Ω)) for some finite numbers s > 1 and
2,1
l > 1. Then problem (4.4.10) has a unique solution such that u ∈ Ws,l (QT )
1,0
and p ∈ Ws,l (QT ), satisfying the following coercive estimate
kukW 2,1 (QT ) + kpkW 1,0 (QT ) ≤ c(Ω, s, l, n)kf kLs,l (QT ) .
s,l s,l
for w(x, t) = χ(t)W (x) with any function χ ∈ C 1 ([0, T ]) and any divergence
free field W ∈ C 2 (Ω) subject to the end condition χ(T ) = 0 and to the
boundary condition W = 0 on ∂Ω, respectively.
Then v is identically zero in QT . Here, QT = Ω×]0, T [.
and thus
ZT
vk (t)(χ0 (t) − χ(t)λk ) = 0 (4.5.1)
0
where Z
ck = u · ϕk dx.
Ω
We know that SN → u in L2 (Ω) as N → ∞. But it is not sufficient to justify
(4.5.3) by taking the limit below
Z Z
0 = v(x, t) · SN (x)dx → v(x, t) · u(x)dx, N → ∞,
Ω Ω
If we let −∆S
e N = fN ∈ L2 (Ω), then simply, by definition of the Stokes
operator, the partial sum S N solves the following boundary value problem
Now, we are again in a position to apply the regularity theory, developed for
the stationary Stokes system, that gives the estimate
So, we have
N
X
k∇2 SN k2,Ω ≤ c(Ω)k∆S
e N k2,Ω ≤ c(Ω) λ2k c2k
k=1
e 2,Ω ≤ c(Ω)k∆uk2,Ω ≤ c(Ω)k∇2 uk2,Ω .
≤ c(Ω)k∆uk
And thus boundedness of kSN km0 ,Ω has been proven in the case n = 3 as
well.
Now, the aim is to show that v is identically zero in QT . Fix t ∈ [0, T ]
and consider a linear functional
Z
◦
l(w) = v · wdx, w ∈ L1m0 (Ω).
Ω
4.5. UNIQUENESS RESULTS 97
◦
By Poincaré inequality, it is bounded in L1m0 (Ω) and, by (4.5.3), vanishes on
◦
J 1m0 (Ω), i.e.,
◦
l(w) = 0 ∀w ∈ J 1m0 (Ω).
As we know, any functional, possessing the above properties, can be presented
in the form Z
◦
l(w) = p div wdx, ∀w ∈ L1m0 (Ω)
Ω
for some p ∈ Lm (Ω). The latter means that v = −∇p.
Our next step is to show that p is a solution to the classical Neumann
problem: ∆p = 0 in Ω and ∂p/∂ν = 0 on ∂Ω, where ν is the unit outward
normal to the surface ∂Ω, in the following sense
Z
∇p · ∇qdx = 0 ∀q ∈ Wm1 0 (Ω). (4.5.4)
Ω
◦
∞
Indeed, ∇p = −v ∈ J m (Ω). Therefore, there exists a sequence w(k) ∈ C0,0 (Ω)
(k)
such that w → ∇p in Lm (Ω). So,
Z Z
(k)
w · ∇qdx = 0 → ∇p · ∇qdx = 0
Ω Ω
for any w(x, t) = χ(t)W (x), where χ ∈ C 1 ([0, T ]) such that χ(T ) = 0 and
∞
W ∈ C0,0 (Ω).
Then v is identically zero in QT .
Proof By density arguments, (4.5.5), of course, holds for any W ∈
◦
J 1m0 (Ω).
Take any function W ∈ C 2 (Ω) with W = 0 on Ω and div W = 0 in Ω.
◦
We know that W ∈ J 1m0 (Ω), see Chapter I, Theorem 4.3. So, v satisfies all
assumptions of Lemma 5.1 and therefore v ≡ 0 in QT .
The above proof works well under additional assumption on n if 1 <
m < 2. However, we can give an alternative proof that does not need extra
4.6. LOCAL INTERIOR REGULARITY 99
in B and Z
q(x, t)dx = 0, w(·, t) = 0
B
∂t V − ∆V = F − ∇P, div V = 0
in Q,
V =0
4.6. LOCAL INTERIOR REGULARITY 101
where
A = kf ks,n,Q + kvks,n,Q + k∇vks,n,Q + kpks,n,Q .
So, our task is to evaluate the last term on the right hand side of (4.6.6).
The key point here is duality arguments proposed by V. A. Solonnikov.
Introducing new notation u = ∂t w and r = ∂t q, we can derive from the
equations for w and q
∆u(·, t) − ∇r(·, t) = 0
div u(·, t) = ∂t v(·, t) · ∇ϕ(·, t) + v(·, t) · ∇ ∂t ϕ(·, t) (4.6.7)
in B, Z
r(x, t)dx = 0, u(·, t) = 0 (4.6.8)
B
on ∂B.
Given g ∈ Ls0 (B) with s0 = s/(s − 1), let us define a function u
e as a
unique solution to the boundary value problem
u − ∇e
∆e r = g, div u
e=0 (4.6.9)
in B, Z
re(x)dx = 0, u
e=0 (4.6.10)
B
ke
rks0 ,B + k∇e
rks0 ,B ≤ ckgks0 ,B . (4.6.11)
and thus
k∂t wks,n,Q ≤ cA.
Proposition 6.1 is proved.
Keeping in mind the 3D non-stationary non-linear problem, one cannot
expect that the number n is big. In such cases, the following embedding
result can be useful.
2,1
Proposition 6.2. Assume that v ∈ Ws,n (Q) with
2 3
1 < n ≤ 2, µ=2− − > 0.
n s
Then
1
|v(z) − v(z 0 )| ≤ c(m, n, s)(|x − x0 | + |t − t0 | 2 )µ (kvks,n,Q
+k∇vks,n,Q + k∇2 vks,n,Q + k∂t vks,n,Q )
for all z = (x, t) ∈ Q(1/2) and for all z 0 = (x0 , t0 ) ∈ Q(1/2). In other words,
v is Hölder continuous with exponent µ relative to parabolic metric in the
closure of Q(1/2).
Finally, using bootstrap arguments, we can prove the following statement
which in a good accordance with the above example.
Proposition 6.3. Assume that conditions (4.6.2) hold with 1 < n < 2 and
f = 0. Let u and p be an arbitrary solution to system (4.6.1). Then for any
0 < τ < 1 and for any k = 0, 1, ..., the function (x, t) 7→ ∇k u(x, t) is Hölder
continuous with any exponent less than 2 − 2/n in the closure of the set Q(τ )
relative to the parabolic metric.
4.7. LOCAL BOUNDARY REGULARITY 103
∂t u − ∆ u = f − ∇ p, div v = 0 in Q+ (2),
u(x0 , 0, t) = 0.
Then u ∈ Wm2,11 ,n (Q+ (1)) and p ∈ Wm1,01 ,n (Q+ (1)) with the estimate
k∂t ukLm1 ,n (Q+ (1)) + k∇2 ukLm1 ,n (Q+ (1)) + k∇pkLm1 ,n (Q+ (1)) ≤
≤ c(kukLm,n (Q+ (2)) + k∇ukLm,n (Q+ (2)) + kpkLm,n (Q+ (2)) + kf kLm1 ,n (Q+ (2)) ).
∂t v − ∆ v = −∇ q, div v = 0 (4.7.1)
104 CHAPTER 4. LINEAR NON-STATIONARY PROBLEM
Here R+3
= {x = (x0 , x3 ) : x3 > 0 }.
Taking an arbitrary function f (t), we seek a non-trivial solution to (4.7.1)
–(4.7.3) in the form of shear flow, say, along x1 -axis:
Here, a scalar function u solves the following initial boundary value problem
w(0, t) = 0, (4.7.5)
w(y, −4) = 0, (4.7.6)
where 0 < y < +∞ and −4 < t < 0 and wyy = ∂ 2 w/∂y 2 .
It is not so difficult to solve (4.7.4)–(4.7.6) explicitly:
√ y
Zt Z4(τ +4)
2 2
w(y, t) = √ f (t − τ − 4)dτ e−ξ dξ. (4.7.7)
π
−4 0
Keeping in mind that our aim is to construct irregular but summable solution,
we choose the function f as follows
1
f (t) = , 0 < α < 1/2. (4.7.8)
|t|1−α
(iii) Let s, s1 , l, and l1 be numbers greater than 1 and satisfy the condition
n1 1 1o 1
K = max 1− ,1 − <α< . (4.7.9)
2 s l1 2
Then
1,0
v ∈ Ws,l (C+ (3)×] − 9/4, 0[), q ∈ Ls1 ,l1 (C+ (3)×] − 9/4, 0[).
Assume we are given numbers 1 < m < +∞ and 1 < n < 2. Letting
s = s1 = m and l = l1 = n and choosing α so that inequality (4.7.9)
holds. The functions v and q constructed above for the chosen α meet all
the conditions of Proposition 7.1 with f = 0. However, ∇v is unbounded
in any neighborhood of the space-time point z = (x, t) = 0. This is a
counter-example of Seregin-Sverak, which is an essential simplification of the
counter-example given by K. Kang.
4.8 Comments
Chapter 4 contains standard material about existence, uniqueness, and regu-
larity of solutions to the non-stationary Stokes system. A bit new results for
introductory course are in the last three sections. In particular fine unique-
ness theorems and local regularity issues are discussed in Sections 5–7. They
are needed for the local regularity analysis in Chapter 6.
106 CHAPTER 4. LINEAR NON-STATIONARY PROBLEM
Chapter 5
Non-linear Non-Stationary
Problem
107
108 CHAPTER 5. NON-LINEAR NON-STATIONARY PROBLEM
where vn0 = vn /kvn kV . The sequence vn0 is bounded in a reflexive space, and
thus without loss of generality we may assume that
vn0 * v0
in V0 and thus
vn0 → v0
in V and V1 . Since nkvn0 kV1 is bounded and therefore kvn0 kV1 → 0 = kv0 kV1 .
Hence, by assumption (iv), kv0 kV = 0. However, 1 = kvn kV → kv0 kV . This
leads to contradiction. Lemma 1.1 is proved.
Proposition 1.2. (Aubin-Lions lemma) Let 1 < p0 , p1 < ∞, V1 is reflexive,
and define
n o
W ≡ kvkW = kvkLp0 (0,T ;V0 ) + k∂t vkLp1 (0,T ;V1 ) < ∞ .
u(j) * u
v (j) → 0
kv (j) (·, t)kV ≤ ηkv (j) (·, t)kV0 + C(η)kv (j) (·, t)kV1
and thus
kv (j) kLp0 (0,T ;V ) ≤ ηkv (j) kLp0 (0,T ;V0 ) + C(η)kv (j) kLp0 (0,T ;V1 )
for a.a. t ∈ [0, T ] and Lebesgue’s theorem. So, our goal is to prove (5.1.3)
and (5.1.4).
To prove (5.1.3), we exploit the following formula (it is a simple conse-
quence of the definition of ∂t v)
Zt
(j)
v (·, t) = ∂t v(·, τ )dτ + v (j) (·, s) (5.1.5)
s
Zs1 Zt Zs1
(j)
(s1 − t)v (·, t) = ds ∂t v(·, τ )dτ + v (j) (·, s)ds.
t s t
After integration by parts in s in the first term of the right hand side, we
find
v (j) (·, t) = a(j) (·, t) + b(j) (·, t),
where
Zs1
1
a(j) (·, t) = v (j) (·, s)ds
s1 − t
t
110 CHAPTER 5. NON-LINEAR NON-STATIONARY PROBLEM
and
Zs1
1
b(j) (·, t) = (s1 − s)∂t v (j) (·, s)ds.
s1 − t
t
ZT p1 1
(j) p1 0
× k∂t v (·, s)kV1 ds 1 ≤ c|s1 − t| p1 < ε
0
a(j) (·, t) * 0
So, given s1 , (5.1.6) is true and we may find N (s1 , t) such that
for any j ≥ N1 (s1 , t). This proves (5.1.4) and completes the proof of Propo-
sition 1.2.
5.2. AUXILIARY PROBLEM 111
∂t v − ∆v + div v ⊗ w + ∇q = f, div v = 0 in QT ,
v|∂Ω×[0,T ] = 0, (5.2.4)
v|t=0 = a
◦
for all ve ∈ J 12 (Ω);
kv(·, t) − a(·)k2,Ω → 0 (5.2.6)
as t → +0.
◦
for all ve ∈ J 12 (Ω);
kv(·, t) − a(·)k2,Ω → 0 (5.2.9)
as t → +0. Here, fe = f − u ⊗ w. Such a function v exists and is unique (for
given u) according to Theorem 2.3 of Chapter 4 since
◦
fe ∈ L2 (0, T ; (J 12 (Ω))0 ).
So, operator A is well defined. Let us check that it satisfies all the assump-
tions of Theorem 1.3 of Chapter III.
Continuity: Let v 1 = A(u1 ) and v 2 = A(u2 ). Then
Z Z
1 2 1 2
v )dx = (u1 − u2 ) ⊗ w : ∇e
(∂t (v − v ) · ve + ∇(v − v ) : ∇e v dx
Ω Ω
Combining the latter estimates, we see that sets which are bounded in X
remain to be bounded in
◦ ◦
W = {w ∈ L2 (0, T ; J 12 (Ω)), ∂t w ∈ L2 (0, T ; (J 12 (Ω))0 )}.
◦ ◦ ◦
By Proposition 1.2 for V0 = J 12 (Ω), V = J (Ω), and V1 = (J 12 (Ω))0 , such a set
is precompact.
Now, we wish to verify the second condition in Theorem 1.3 of Chapter
III. For v = λA(v), after integration by parts, we find that, for a.a. t ∈ [0, T ],
Z Z
(∂t v · ve + ∇v : ∇e
v )dx = λ (f · ve − (w · ∇v) · ve)dx
Ω Ω
◦
for any ve ∈ J 12 (Ω). If we plague ve(·) = v(·, t) into the latter relation, then
the identity Z
(w · ∇v) · vdx = 0,
Ω
and thus
Z
kvk2QT ≤ T sup |v(x, t)|2 dx ≤ cT (kf k2 ◦ + kak22,Ω ) = R2 .
0<t<T L2 (0,T ;(J 12 (Ω))0 )
Ω
Now, all the statements of Proposition 2.1 follow from the Leray-Schauder
principle. Proposition 2.1 is proved.
Let ω% be the usual mollifying kernel and let
Z
(v)% (x, t) = ω% (x − x0 )v(x0 , t)dx0 .
Ω
114 CHAPTER 5. NON-LINEAR NON-STATIONARY PROBLEM
◦
It is easy to check if t 7→ v(·, t) ∈ J (Ω) then div(v)% (·, t) = 0 (Exercise).
Now, we wish to show that there exists at least one function v % such that:
◦ ◦
v % ∈ C([0, T ]; L2 (Ω)) ∩ L2 (0, T ; J 12 (Ω)), ∂t v % ∈ L2 (0, T ; (J 12 (Ω))0 ); (5.2.10)
◦
for all ve ∈ J 12 (Ω);
kv % (·, t) − a(·)k2,Ω → 0 (5.2.12)
as t → +0.
We note that (5.2.10)-(5.2.12) can be regarded as a weak form of the
following initial boundary value problem
∂t v % − ∆v % + (v % )% · ∇v % + ∇q % = f, div v ρ = 0 in QT ,
v % |∂Ω×[0,T ] = 0, (5.2.13)
v % |t=0 = a.
◦
for all ve ∈ J 12 (Ω);
kv(·, t) − a(·)k2,Ω → 0 (5.2.17)
as t → +0. By Proposition 2.1, such a function exists and is unique. We
need to check that all the assumptions of Theorem 1.3 of Chapter III hold
for our operator A.
Continuity: Do the same as in Proposition 2.1:
1
∂t kv 2 − v 1 k22,Ω + k∇(v 2 − v 1 )k22,Ω
Z 2
= v 2 ⊗ (u2 )% − v 1 ⊗ (u1 )% : ∇(v 2 − v 1 )dx
Ω
Z
= (v 2 − v 1 ) ⊗ (u2 )% : ∇(v 2 − v 1 )dx
Ω
Z
+ v 1 ⊗ (u2 − u1 )% : ∇(v 2 − v 1 )dx.
Ω
The first integral in the right hand side of the latter is zero and the second
one I can be bounded as follows
and thus
∂t v − ∆v + div v ⊗ v + ∇q = f, div v = 0 in QT ,
v|∂Ω×[0,T ] = 0, (5.3.1)
v|t=0 = a
f ∈ L2 (0, T ; V 0 ) (5.3.2)
and
a ∈ H, (5.3.3)
◦ ◦
where V = J 12 (Ω) and H = J (Ω).
Theorem 3.2. Under assumptions (5.3.2) and (5.3.3), there exists at least
one weak Leray-Hopf solution to (5.3.1).
for all ve ∈ V ;
kv % (·, t) − a(·)k2,Ω → 0 (5.3.6)
%
as t → +0. Moreover, v has uniformly bounded energy
|v % |22,QT ≡ sup kv % (·, t)k22,Ω + k∇v % k22,QT ≤ A, (5.3.7)
0<t<T
Therefore
k∂t v % k ◦ ≤ c(Ω) |v % |2,QT kv % k2,QT + |v % |2,QT + kf kL2 (0,T ;V 0 ) .
L2 (0,T ;(J 32 (Ω))0 )
1
Since kv % k2,QT ≤ T 2 |v % |2,QT , we get
k∂t v % k ◦ ≤ A1 , (5.3.8)
L2 (0,T ;(J 32 (Ω))0 )
5.3. WEAK LERAY-HOPF SOLUTIONS 119
v% * v in L2 (0, T ; V ), (5.3.10)
v% → v in L2 (0, T ; H). (5.3.11)
Now, let us show
Z
D% = |v % ⊗ (v % )% − v ⊗ v|dz → 0
QT
QT QT
Z
% %
≤ kv − vk2,QT k(v )% k2,QT + |v ⊗ (v % − v)% |dz
QT
Z
1
+ |v ⊗ ((v)% − v)|dz ≤ kv % − vk2,QT T 2 |v % |2,QT
QT
1 1
+T |v|2,QT kv % − vk2,QT + T 2 |v|2,QT k(v)% − vk2,QT → 0
2
Now, our goal is to show that for each fixed ve, the set of functions fve% is
precompact in C([0, T ]). Indeed, it is uniformly bounded since
sup |fve% (t)| ≤ |v % |2,QT ke
v k2,Ω ≤ c|v % |2,QT ke
vk ◦3 ≤ cAke
vk ◦3 .
0<t<T J 2 (Ω) J 2 (Ω)
t Ω
t+∆t
Z
≤ k∂t v % (·, τ )k ◦ ke
v (·)k ◦ 3 dτ
(J 32 (Ω))0 J 2 (Ω)
t
p p
≤ |∆t|k∂t v % k ◦ ke
vk ◦3 ≤c |∆t|A1 ke
vk ◦3 .
L2 (0,T ;(J 32 (Ω))0 ) J 2 (Ω) J 2 (Ω)
◦
Now, let ve(k) be a countable set that is dense in J 32 (Ω). Applying thedi-
agonal Cantor procedure, we can select a subsequence such that
Z Z
v (x, t) · ve (x)dx → v(x, t) · ve(k) (x)dx
% (k)
Ω Ω
we have
Z Zt Z Z
1 % 2 % 2 1 0
|v (x, t)| dx + |∇v | dxdt = |a(x)|2 dx
2 2
Ω 0 Ω Ω
Zt Z
+ f · v % dxdt0 (5.3.13)
0 Ω
and
Zt Z Zt Z
lim inf f · v % dxdt0 = f · vdxdt0 (5.3.16)
%→0
0 Ω 0 Ω
for all t ∈ [0, T ]. So, (v) of Definition 3.1 follows from (5.3.13)-(5.3.16).
It remains to prove validity of (iv) of Definition 3.1. To this end, notice
that by (5.3.12)
a(·) = v % (·, 0) * v(·, 0)
in L2 (Ω). So, v(·, 0) = a(·). Moreover, according to (ii) of Definition 3.1
which together with week convergence gives (iv) of Definition 3.1. Theorem
3.2 is proved.
Zx1
|u(x1 , x2 )|2 = 2 u(t, x2 )u,1 (t, x2 )dt
−∞
Z∞ 21 Z∞ 21
2 2
≤2 |u(t, x2 )| dt |u,1 (t, x2 )| dt .
−∞ −∞
5.4. MULTIPLICATIVE INEQUALITIES 123
And then
Z∞ Z∞ Z∞ Z∞
4
|u(x1 , x2 )| dx1 dx2 = |u(x1 , x2 )|2 |u(x1 , x2 )|2 dx1 dx2
−∞ −∞ −∞ −∞
Z∞ n Z∞ 21 Z∞ 21
2
≤4 dx1 dx2 |u(t, x2 )| dt |u,1 (t, x2 )|2 dt
−∞ −∞ −∞
Z∞ 21 Z∞ 12 o
× |u(x1 , s)|2 ds |u,2 (x1 , s)|2 ds
−∞ −∞
Z∞ Z∞ 21 Z∞ 21
2 2
= |u(t, x2 )| dt |u,1 (t, x2 )| dt dx2
−∞ −∞ −∞
Z∞ Z∞ 12 Z∞ 12
× |u(x1 , s)|2 ds |u,2 (x1 , s)|2 ds dx1
−∞ −∞ −∞
Proof We have
kuks,Ω ≤ kuk1−α α
2,Ω kuk6,Ω
3(s−2)
with α = 2s
.
Corollary 4.4. Let u ∈ L∞ (0, T ; H) ∩ L2 (0, T ; V ). Then
kuks,l,QT ≤ c(s)|u|2,QT
and
ZT 1l ZT 1l
1−α
ku(·, t)kls,Ω dt ≤ c(s)|u|2,Q T
k∇u(·, t)k αl
2,Ω dt .
0 0
ku · ∇uks,l,QT ≤ c(s)|u|22,QT
2s
with s1 = 2−s
and thus after integration in t and application of Hölder
inequality
ZT Z s 2l 2−l
s1 1 (2−l) 2l
ku · ∇uks,l,QT ≤ k∇uk2,QT |u| dx dt .
0 Ω
It is easy to verify
3 2 3 2l
+ = , l1 = . (5.4.2)
s1 l1 2 l−2
The required inequality follows from Corollary 4.4.
Let us discuss some consequences of (5.4.2):
3 2 3 2−l 3(2 − s) 2 3
+ = 2s + = + −1= ,
s1 l1 2−s
l 2s l 2
which implies
3 2
+ = 4.
s l
for all w ∈ V ;
u(x, 0) = a(x), x ∈ Ω. (5.5.3)
126 CHAPTER 5. NON-LINEAR NON-STATIONARY PROBLEM
Recalling Definition 3.1, part (ii), we get from (5.5.2) that v = v − u satisfies
the identity
Z h i
− v(x, t) · W (x)∂t χ(t) + ∇v(x, t) : ∇W (x)χ(t) dz = 0 (5.5.4)
QT
∞
for all W ∈ C0,0 (Ω) and for any χ ∈ C01 (0, T ). Our aim is to get rid of the
assumption that χ vanishes in a neighborhood of 0. To this end, we are going
to use the following fact:
kv(·, t)k2,Ω → 0 (5.5.5)
as t → +0. Take any function χ ∈ C 1 ([0, T ]) so that χ(T ) = 0 and a function
ϕε having the properties: ϕε (t) = 1 if t ≥ ε, ϕε (t) = 0 if 0 < t ≤ ε/2, and
ϕε (t) = (2t − ε)/ε if ε/2 < t < ε. Then, by (5.5.4), with ϕε χ as χ,
Z h i
ϕε (t) − v(x, t) · W (x)∂t χ(t) + ∇v(x, t) : ∇W (x)χ(t) dz
QT
Zε Z
2
= v(x, t) · W (x)χ(t)dz = Iε .
ε
ε Ω
2
as ε → 0.
So,
Z h i
− v(x, t) · W (x)∂t χ(t) + ∇v(x, t) : ∇W (x)χ(t) dz = 0
QT
∞
for all W ∈ C0,0 (Ω) and any χ ∈ C 1 ([0, T ]) with χ(T ) = 0. This means that
v ≡ 0 in QT and any weak Leray-Hopf solution v has the following properties:
v ∈ C([0, T ]; H) ∩ L2 (0, T ; V ), ∂t v ∈ L2 (0, T ; V 0 ); (5.5.6)
for a.a. t ∈ [0, T ],
Z h i
∂t v(x, t) · w(x) + ∇v(x, t) : ∇w(x) dx
Ω
Z h i
= v(x, t) ⊗ v(x, t) : ∇w(x) + f (x, t) · w(x) dx
e (5.5.7)
Ω
5.5. UNIQUENESS 127
for all w ∈ V ;
v(x, 0) = a(x), x ∈ Ω. (5.5.8)
Now, assume that we have two different solutions v 1 and v 2 . Letting
u = v 2 − v 1 , one deduce from (5.5.7) that:
Z h i
∂t u(x, t) · w(x) + ∇u(x, t) : ∇w(x) dx =
Ω
Z
= (v 2 (x, t) ⊗ v 2 (x, t) − v 1 (x, t) ⊗ v 1 (x, t)) : ∇w(x)dx
Ω
kv 1 (·, t)k44,Ω ≤ 2kv 1 (·, t)k22,Ω k∇v 1 (·, t)k22,Ω ≤ 2|v 1 |22,QT k∇v 1 (·, t)k22,Ω ≤
≤ c(a, f )k∇v 1 (·, t)k22,Ω
and
ku(·, t)k44,Ω ≤ 2ku(·, t)k22,Ω k∇u(·, t)k22,Ω .
And thus
1
∂t ku(·, t)k22,Ω + k∇u(·, t)k22,Ω ≤
2
3 1 1
≤ c(a, f )k∇u(·, t)k2,Ω
2
ku(·, t)k2,Ω
2
k∇v 1 (·, t)k2,Ω
2
.
and thus
∂t ku(·, t)k22,Ω ≤ c0 y(t)ku(·, t)k22,Ω ,
128 CHAPTER 5. NON-LINEAR NON-STATIONARY PROBLEM
where y(t) := k∇v 1 (·, t)k22,Ω . From this differential inequality, it is not diffi-
cult to derive
−c0 Rt y(τ )dτ
∂t e 0 ku(·, t)k22,Ω ≤ 0
and
Rt
−c0 y(τ )dτ
e 0 ku(·, t)k22,Ω ≤ ku(·, 0)k22,Ω = 0.
Therefore, ku(·, t)k22,Ω ≡ 0.
Let us discuss further regularity of 2D weak Leray-Hopf solutions.
Theorem 5.2. Assume that a ∈ V , and f ∈ L2 (QT ). Let v be a unique
solution to initial boundary value problem (5.3.1). Then
∂t v + v · ∇v − ∆v = f − ∇q, div v = 0
fe = f − (v % )% · ∇v % ∈ L2 (QT );
a.e. in QT and u% (·, 0) = a(·). Using the same arguments as in the proof
of the previous statement, we can claim that v % = u% . Multiplying then the
equation
∂t v % + (vk% )% v,k
%
− ∆v % = f % − ∇ p%
e % and integrating each term in the product with respect to x, we find
by ∆v
Z Z
% e % e % |2 dx,
∆v · ∆v dx = |∆v
Ω Ω
Z
∇ p% · ∆v
e % dx = 0,
Ω
Z Z
% % 1
∂t v · ∆v
e dx = ∂t v % · ∆v % dx = − ∂t k∇ v % k22,Ω
2
Ω Ω
To estimate the first and the second factors in the last term of the right
hand side of the above inequality, we are going to exploit Ladyzhenskaya’s
inequality one more time
and
k∇v % k44,Ω ≤ c(Ω)k∇v % k22,Ω (k∇2 v % k22,Ω + k∇v % k22,Ω ).
We also need the Cattabriga-Solonnikov inequality
where function y(t) = k∇v % (·, t)k22,Ω obeys the initial condition y(0) = k∇ak22,Ω .
The latter, together with energy estimate (5.3.7), gives two estimates
But
kfek2,QT ≤ kf k2,QT + k(v % )% · ∇v % k2,QT ≤
≤ kf k2,QT + k(v % )% k4,QT k∇v % k4,QT ≤
h i
% %
≤ c(a, f ) 1 + kv k4,QT k∇v k4,QT .
The right hand side of the above inequality can be evaluated with the help
of Ladyzhenskaya’s inequality and (5.5.12).
Finally, we can pass to the limit as % → 0 and get all the statements of
Theorem 5.2. Theorem 5.2 is proved.
f ∈ L2 (QT ), a ∈ H. (5.6.1)
5.6. FURTHER PROPERTIES 131
u ∈ L∞ (0, T ; H) ∩ L2 (0, T ; V ).
132 CHAPTER 5. NON-LINEAR NON-STATIONARY PROBLEM
∞
Next, plague χw with w ∈ C0,0 (QT ) into identity (iii) of Definition 3.1 and
get for u: Z Z
(−u · ∂t w + ∇u : ∇w)dz = fe · wdz,
QT QT
where
fe = χf − χdiv v ⊗ v − ∂t χv = χf − χv · ∇v − ∂t χv.
By Corollary 4.5,
fe ∈ Ls,l (QT )
for any s, l > 1 satisfying 3/s + 2/l = 4. Moreover, u = 0 for sufficiently
small t. On the other hand, the linear theory ensures that, for such fe, there
exist functions ve and qe such that
2,1 1,0
ve ∈ Ws,l (QT ), qe ∈ Ws,l (QT )
∂t ve − ∆e
v + ∇e
q = fe, div ve = 0
in QT , Z
ve(x, t) = 0 x ∈ ∂Ω, qe(x, t)dx = 0
Ω
for t ∈ [0, T ],
ve(·, 0) = 0
in Ω.
Using essentially the same arguments as in 2D-case, we can show that for
vb = u − ve Z
v · ∂t w − ∇b
(b v : ∇w)dz = 0
QT
∞
for w = χeW with W ∈ C0,0 (Ω) and with χ e ∈ C 2 ([0, T ]) and χ
e(T ) = 0.
Now, the uniqueness results for the linear theory imply that u b = 0. Hence,
2,1
χv ∈ Ws,l (QT ) and
χ(∂t v + v · ∇v − ∆v − f ) = −∇e
q.
Next, take any δ > 0 and assume that χ(t) = χδ (t) = 1 if t > δ and
χ(t) = χδ (t) = t/δ if < t < δ. A pressure qe corresponding to chosen δ is
5.6. FURTHER PROPERTIES 133
1,0
denoted by qδ . Obviously, qδ ∈ Ws,l (Qδ,T ) with required s and l. Assuming
that δ1 > δ2 then qδ1 = qδ2 in Qδ1 ,T . This allows us to introduce the function
q so that
q(·, t) = qδ (·, t)
if t > δ > 0. It is well defined and satisfies the required properties. So, the
first part of the theorem is proved.
Part II. Now, let us go back to the proof of Theorem 3.2 on the existence
of weak Leray-Hopf solutions and try to apply the procedure, described in
the proof of Part I, to regularized problem. Letting u% = χv % , where χ is a
function of t from C02 (0, ∞), we state that u% is a solution to the problem:
for all ve ∈ V ;
ku% (·, t)k2,Ω → 0 (5.6.6)
as t → +0.
Here,
fe% = fe1% + fe2%
with
fe1% (x, t) = χ(t)f (x, t) + ∂t χ(t)v % (x, t),
fe2% (x, t) = −χ(t)(v % )% (x, t) · ∇v % (x, t).
Simply repeating the proof of Corollary 4.4 and Corollary 4.5, we estimate
the second part
with s0 , l0 > 1 and 3/s0 + 2/l0 = 4. Moreover, we can claim that the whole
right hand side fe% is estimated similarly:
e%
Now, according to Theorem 5.2 of Chapter 4, there exists functions u
and qχ% satisfy the relations:
e% ∈ Ws2,1
u 0 ,l0 (QT ), qχ% ∈ Ws1,0
0 ,l0 (QT );
e% − ∆ u
∂t u e% = fe% − ∇ qχ% , e% = 0
div u
in QT ;
e% |∂ 0 QT = 0,
u [qχ% (·, t)]Ω = 0.
By the same theorem, these functions have the bound
k∇2 u
e% ks0 ,l0 ,QT + k∂t u
e% ks0 ,l0 ,QT + k∇qχ% ks0 ,l0 ,QT
≤ c(s0 , Ω)kfe% ks0 ,l0 ,Q ≤ c(s0 , Ω, a, f, χ)
T
k∇2 v % ks0 ,l0 ,Qδ,T + k∂t v % ks0 ,l0 ,Qδ,T + k∇qδ% ks0 ,l0 ,Qδ,T
≤ c(s0 , Ω, a, f, δ). (5.6.7)
in L∞ (0, T ; H),
∇v % * ∇v
in L2 (0, T ; V ),
v% → v
in L2 (QT ), Z Z
%
v (x, t) · w(x)dx → v(x, t) · w(x)dx
Ω Ω
with 3/s00 + 2/l00 = 3/2, see Corollary 4.4. In particular, we may assume that
(take s00 = l00 = 10/3)
v% * v
in L 10 (QT ) and therefore
3
v% → v (5.6.8)
in L3 (QT ) and
(v % )% · ∇v % * v · ∇v
in L1 (QT ).
Given s, l > 1 with 3/s+2/l=3, we find l0 = l and s0 = 3s/(s + 3). It
is easy to check, 3/s0 + 2/l0 = 4. Using the diagonal Cantor procedure and
bounds (5.6.7), one can ensure that
∂t v % * ∂t v, ∇2 v % * ∇2 v
in Ls0 ,l0 (Qδ,T ) for each δ > 0. As to the pressure, the diagonal Cantor proce-
dure can be used one more time to show that
∇qδ% * ∇qδ
∇qδ = f − ∂t v − v · ∇v + ∆v
belongs Ws1,0
0 ,l0 (QT ) and Ls,l (QT ) for each δ > 0.
C0∞ (R3 ×]0, ∞[, we choose δ so small that spt ϕ ∈ R3 ×]δ, ∞[. After mul-
tiplication of the equation
∂t v % + (v % )% · ∇v % − ∆v % = f − ∇qδ%
by ϕv % and integration of the product by parts (which is legal for the regu-
larized solution)
Z Zt0 Z Zt0 Z
|v % (x, t0 )|2 ϕ(x, t0 )dx + ϕ|∇v % |2 dxdt = |v % |2 (∂t ϕ + ∆ϕ)+
Ω 0 Ω 0 Ω
+(v % )% · ∇ϕ(|v % |2 + 2qδ% ) + 2f · v dxdt (5.6.9)
for any t ∈ [δ, T ].
We also know that
qδ% * q
in L 3 (Qδ,T ). (Indeed, qδ% * q in L 5 (Qδ,T ) since 3/(5/3) + 2/(5/3) = 3).
2 3
Taking into account (5.6.8) and using the same arguments as in the proof of
Theorem 3.2, one can pass to the limit in (5.6.9) as % → 0 and get required
local energy inequality (5.6.3).
Theorem 7.2. (Global existence of strong solutions for ”small” data). There
exists a constant c0 (Ω) such that if
π
arctan(k∇ak22,Ω ) + c0 (Ω)(kak22,Ω + kf k22,QT ) < , (5.7.2)
2
then there exists a strong solution to initial boundary value problem (5.3.1)
Proof Let us go back to problem (5.3.4)-(5.3.6), see the proof of Theorem
3.2,
v % ∈ C([0, T ]; H) ∩ L2 (0, T ; V ), ∂t v % ∈ L2 (0, T ; V 0 ); (5.7.3)
5.7. STRONG SOLUTIONS 137
for all ve ∈ V ;
kv % (·, t) − a(·)k2,Ω → 0 (5.7.5)
as t → +0. Here,
f % = f − (v % )% · ∇v % ∈ L2 (QT ).
Using the similar arguments as in the proof of Theorem 5.1 of this section
and Theorems 6.2, 5.1 of Section 4 on uniqueness and regularity for non-
stationary Stokes problem, we may conclude that
Moreover, there exists a pressure field p% ∈ W21,0 (QT ) such that the regu-
larised Navier-Stokes equations
∂t v % − ∆v % = f % − ∇p% , div v % = 0
hold a.e. in QT . This is the starting point for the proof of our theorem.
We know that sequence v % converges to a weak Leray-Hopf solution to corre-
sponding initial boundary value problem (5.3.1). So, what we need is to get
uniform estimates of ∇v % . Let
Z
y(t) := |∇v % (x, t)|2 dx.
Ω
e %
We proceed as in the proof of Theorem 5.2, multiplying the equation by ∆v
and arguing exactly as it has been done there. As a result, after obvious
applications of Cauchy and Hölder inequalities, we find
Z Z
0
y + |∆v | dx ≤ c |(v % )% |2 |∇v % |2 dx + ckf k22,Ω ≤
e % 2
Ω Ω
≤ ckv % k26,Ω k∇v % k23,Ω + ckf k22,Ω . (5.7.6)
138 CHAPTER 5. NON-LINEAR NON-STATIONARY PROBLEM
The first term on the right hand side of the latter inequality can be evaluated
with the help of the Galiardo-Nirenberg inequality
kv % k26,Ω ≤ cy,
where Z
ck (t) = v % (x, t) · ϕk (x)dx.
Ω
with
y(0) = k∇ak22,Ω , g(t) = kf (·, t)k22,Ω
A weaker versions of (5.7.9) is
y 0 (t) h i
≤ c1 (Ω) y(t) + g(t)
1 + y 2 (t)
5.7. STRONG SOLUTIONS 139
y(t) ≤ C, t ∈ [0, T ],
a ∈ V, f ∈ L2 (QT ).
Then there exists T 0 ∈]0, T ] such that initial boundary value problem (5.3.1)
has a strong solution in QT 0 .
Proof Arguing as in the proof of Theorem 7.2, let us go back to inequal-
ity (5.7.9). We need to show that there exists T 0 ≤ T , where y(t) has an
upper bound independent of % and t ∈ [0, T 0 ]. To achieve this goal, we make
a substitution z(t) = y(t) − y(0) and, after application of Young’s inequality,
arrive at the following modification of estimate (5.7.9)
ZT 0 h r
3
i 1 λ1
c1 y (0) + g(t) dt ≤ .
2 c1
0
The latter implies t% ≥ T 0 for all % > 0 and the required estimate
ZT 0 h r
i 1 λ1
y(t) ≤ k∇ak22,Ω + c1 3 2
y (0) + g(t) dt ≤ k∇ak2,Ω +
2 c1
0
Proof In this case, one can deduce from inequality (5.7.9) the following:
y0
≤ c1
y3
for 0 ≤ t ≤ T . The integration gives us:
1 1
− ≤ 2c1 t
y 2 (0) y 2 (t)
and thus
y 2 (t)(1 − 2c1 ty 2 (0)) ≤ y 2 (0).
√
We let T00 = 4c1 y12 (0) . Then 1 − 2c1 ty 2 (0) ≥ 1/2 and y(t) ≤ 2y(0) for
0 < t ≤ T00 which implies T 0 ≥ T00 and we get the required estimate with an
appropriated constant.
Remark 7.5. Solutions, constructed in Theorems 7.2 and 7.3, have the fol-
lowing regularity properties: v ∈ W22,1 (QT ) and there exists a function q such
that ∇q ∈ L2 (QT ) and
∂t v + v · ∇v − ∆v = f − ∇q, div v = 0
∂t v % − ∆v % + ∇q % = f − (v % )% · ∇v % , div v % = 0 in QT
% %
v |∂Ω×[0,T ] = 0, v |t=0 = a,
∂t v + v · ∇v − ∆v = f − ∇q, div v = 0 in QT
v|∂Ω×[0,T ] = 0, v|t=0 = a.
∂t v + v · ∇v − ∆v = f − ∇q, div v = 0 in QT
v|∂Ω×[0,T ] = 0, v|t=0 = a,
∂t u ∈ L2 (0, T ; V 0 ).
Indeed,
Z Z
(u · ∇u) · wdx = u ⊗ u : ∇wdx ≤ kuk24,Ω k∇wk2,Ω
Ω Ω
≤ c(Ω)k∇uk22,Ω k∇wk2,Ω ≤ C(Ω, u)k∇wk2,Ω
∞
for any w ∈ C0,0 (Ω) and for any χ ∈ C0∞ (0, T ). It is easy to see
Z h
∂t u(x, t) · w(x) − (u ⊗ u)(x, t) : ∇w(x)
Ω
i Z
+∇u(x, t) : ∇w(x) dx = f (x, t) · w(x)dx
Ω
Z
1
∂t kvk2,Ω + k∇vk2,Ω = (u1 ⊗ u1 − u2 ⊗ u2 ) : ∇vdx
2 2
2
Ω
Z Z
= (v ⊗ u + u ⊗ v) : ∇vdx = u2 ⊗ v : ∇vdx
1 2
Ω Ω
Z
=− v ⊗ v : ∇u2 dx ≤ k∇u2 k2,Ω kvk24,Ω . (5.7.11)
Ω
which implies
Z Z
−C1 t 2
e |v(x, t)| dx ≤ |v(x, 0)|2 dx = 0.
Ω Ω
∂t u + u · ∇u − ∆u = f − ∇p, div u = 0 in QT
u|∂Ω×[0,T ] = 0, u|t=0 = a,
∞
for any w ∈ C0,0 (QT ).
We would like to test (5.7.14) with u2 but it should be justified. Indeed,
we know that
u1 ∈ L 10 (QT )
3
and, by density arguments, (5.7.14) must be true for w(x, t) = χ(t)v(x) with
◦ ◦
v ∈ J 15 (Ω) and χ ∈ W 11 (0, T ) = {χ ∈ W11 (0, T ) : χ(0) = 0, χ(T ) = 0}. For
2
domains with sufficiently smooth domains
◦ ◦
J 5 (Ω) = Jˆ5 (Ω) = {v ∈ L 5 (Ω) : div v = 0}.
1 1 1
2 2 2
7 3
k∇v(·, t)k 5 ,Ω ≤ c(Ω)k∇v(·, t)k2,Ω
10
(k∇2 (·, t)k2,Ω + k∇(·, t)k2,Ω ) 10 ,
2
7 3
k∇vk 5 ,QT ≤ c(Ω, T ) sup k∇vk2,Ω
10
(k∇2 vk2,QT + k∇vk2,QT ) 10 .
2
0<t<T
where
Zα Z Zα Z
1 1
−Iα = u1 · u2 dxdt = (u1 − a) · (u2 − a)dxdt
α α
0 Ω 0 Ω
Zα Z Zα Z
1 2 1
+ a · (u − a)dxdt + a · (u1 − a)dxdt
α α
0 Ω 0 Ω
Zα Z
1
+ |a|2 dxdt.
α
0 Ω
Since ku1 (·, t) − a(·)k2,Ω and ku2 (·, t) − a(·)k2,Ω go to zero as t → 0, we can
observe that Z
Iα → − |a|2 dx
Ω
as α → 0.
The analogous result takes place at the right end point:
tZ
0 +βZ
1
Iβ = u1 · u2 dxdt
β
t0 Ω
tZ
0 +βZ
1
= u1 (x, t) · (u2 (x, t) − u2 (x, t0 ))dxdt
β
t0 Ω
tZ
0 +βZ
1
+ (u1 (x, t) − u1 (x, t0 )) · u2 (x, t0 )dxdt
β
t0 Ω
Z
+ u1 (x, t0 ) · u2 (x, t0 )dx
Ω
By strong continuity in L2 (Ω) of the strong solution u2 , the first term on the
right hand side goes to zero and by weak continuity of weak solution u1 the
second term there goes to zero as well. So,
Z
Iβ → u1 (x, t0 ) · u2 (x, t0 )dx
Ω
5.7. STRONG SOLUTIONS 147
as β → 0. Finally, we have
Zt0 Z
(−u1 · ∂t u2 − u1 ⊗ u1 : ∇u2 + ∇u1 : ∇u2 − f · u2 )dz
0 Ω
Z Z
1 2
= u (x, t0 ) · u (x, t0 )dx − |a|2 dx (5.7.15)
Ω Ω
Zt0 Z Z Z
2 1 1 2 2 1
∂t u (x, t) · u (x, t)dxdt − |u (x, t0 ) dx + |a|2 dx
2 2
0 Ω Ω Ω
Zt0 Z
+ − u2 ⊗ u2 : ∇(u1 − u2 ) + ∇u1 : ∇(u1 − u2 )
0 Ω
1 2
−f · (u − u ) dxdt = 0.
Z Zt0 Z Z Zt0 Z
1
|u1 (x, t0 )|2 dx + |∇u1 |2 dxdt ≤ |a|2 dx + f · u1 dxdt.
2
Ω 0 Ω Ω 0 Ω
148 CHAPTER 5. NON-LINEAR NON-STATIONARY PROBLEM
The rest of the proof is similar to the proof of Theorem 7.2. Indeed,
Zt0 Z
I= (u2 ⊗ (u1 − u2 )) : ∇(u1 − u2 )dxdt
0 Ω
Zt0 Z
− (u1 − u2 ) ⊗ (u1 − u2 ) : ∇u2 dxdt
0 Ω
Zt0
≤ k∇u2 k2,Ω ku2 − u1 k24,Ω dt
0
Zt0
≤ sup k∇u2 k2,Ω ku2 − u1 k24,Ω dt.
0<t<T
0
Since u2 is a strong solution, the quantity sup0<t<T ku2 k4,Ω = k∇u2 k2,∞,QT is
finite. Applying multiplicative inequality, we have
Zt0 1 3
I ≤ ck∇u2 k2,∞,QT ku2 − u1 k2,Ω
2
k∇(u2 − u1 )k2,Ω
2
dt
0
Zt0 Z 41 Zt0 Z 34
2 2 1 2 2 1 2
≤ ck∇u k2,∞,QT |u − u | dxdt |∇(u − u )| dxdt .
0 Ω 0 Ω
6.1 Notation
In this chapter, we are going to exploit the following notation:
R+ = {t ∈ R : t > 0}, R− = {t ∈ R : t < 0};
Rd+ = {x = (x0 , xd ) : x0 = (xi ), i = 1, 2, ..., d − 1, xd > 0};
Q− = Rd × R− , Q+ = Rd × R+ ;
Qδ,T = Ω×]δ, T [, QT = Ω×]0, T [, Ω ⊂ Rd ;
B(x, r) is the ball in Rd of radius r centered at the point x ∈ Rd , B(r) =
B(0, r), B = B(1);
B+ (x, r) = {y = (y 0 , yd ) ∈ B(x, r) : yd > xd } is a half ball, B+ (r) = B+ (0, r),
B+ = B+ (1);
Q(z, r) = B(x, r)×]t−r2 , t[ is the parabolic ball in Rd ×R of radius r centered
at the point z = (x, t) ∈ Rd × R, Q(r) = Q(0, r), Q = Q(1);
Q+ (r) = Q+ (0, r) = B+ (r)×] − r2 , 0[;
Ls (Ω) and Ws1 (Ω) are the usual Lebesgue and Sobolev spaces, respectively;
Ls,l (QT ) = Ll (0, T ; Ls (Ω)), Ls (QT ) = Ls,s (QT );
151
152 CHAPTER 6. LOCAL REGULARITY THEORY
1,0 1,0
Ws,l (QT ) = {|v| + |∇v| ∈ Ls,l (QT )} and Ws,l (QT ) = {|v| + |∇v| + |∇2 v| +
|∂t v| ∈ Ls,l (QT )} are parabolic Sobolev spaces;
∞
C0,0 (Ω) = {v ∈ C0∞ (Ω) : div v = 0};
◦ ◦
∞
J (Ω) is the closure of the set C0,0 (Ω) in the space L2 (Ω), J 12 (Ω) is the closure
of the same set with respect to the Dirichlet integral;
BM O is the space of functions having bounded mean oscillation;
C(Ω) is the space of continuous function, C α (QT ) is the space of Hölder
continuous with respect to the parabolic metrics;
c is a generic positive constant.
holds for a.a. t ∈]T1 , T [ and all nonnegative functions ϕ ∈ C0∞ (ω×]T1 , ∞[).
One of the main results of the theory of suitable weak solutions reads
6.2. ε-REGULARITY THEORY 153
Lemma 2.2. There exist absolute positive constants ε0 and c0k , k = 1, 2, ...,
with the following property. Assume that a pair U and P is a suitable weak
solution to the Navier-Stokes equations in Q and satisfies the condition
Z
3
3
|U | + |P | dz < ε0 .
2 (6.2.5)
Q
Remark 2.3. For k = 1, Lemma 2.2 has been proven essentially in [2], see
Corollary 1. For alternative approach, we refer the reader to [31], see Lemma
3.1. The case k > 1 was treated in [40], see Proposition 2.1, with the help of
the case k = 1 and regularity results for linear Stokes type systems.
Proposition 2.4. Given numbers ϑ ∈]0, 1/2[ and M > 3, there are two
constants ε1 (ϑ, M ) > 0 and c1 (M ) > 0 such that, for any suitable weak
solution v and q to the Navier-Stokes equations in Q, satisfying the additional
conditions
|(v),1 | < M, Y1 (v, q) < ε1 , (6.2.7)
the following estimate is valid:
2
Yϑ (v, q) ≤ c1 ϑ 3 Y1 (v, q). (6.2.8)
1 Z 31
1
Y (z0 , R; v) = |v − (v)z0 ,R |3 dz ,
|Q(R)|
Q(z0 ,R)
1 Z
3
23
2
Y (z0 , R; q) = R |q − [q]z0 ,R | 2 dz ,
|Q(R)|
Q(z0 ,R)
154 CHAPTER 6. LOCAL REGULARITY THEORY
Z Z
1 1
(v)z0 ,R = v dz, [q]x0 ,R = q dx,
|Q(R)| |B(R)|
Q(z0 ,R) B(x0 ,R)
Y1 (v k , q k ) = ε1k → 0 (6.2.9)
as k → +∞,
3
Yϑ (v k , q k ) > c1 ε1k ϑ 2 (6.2.10)
for all k ∈ N. The constant c1 will be specified later. Let us introduce
functions
Y1 (uk , pk ) = 1, (6.2.11)
2
Yϑ (uk , pk ) > c1 ϑ 3 , (6.2.12)
and the system
1
∂t uk + div ((v k ),1 + ε1k uk ) ⊗ ((v k ),1 + ε1k uk )
ε1k in Q (6.2.13)
−∆ uk = −∇ pk , div uk = 0
and
∂t u + div u ⊗ b − ∆ u = −∇ p
in Q (6.2.15)
div u = 0
6.2. ε-REGULARITY THEORY 155
kuk2,∞,Q(3/4) + k∇ uk2,Q(3/4) ≤ c2 (M ).
Z0 23
3
+ sup |∂t u(x, t)| dt ≤ c3 (M ).
2
x∈B(2/3)
−(2/3)2
Let us estimate the first derivative in time with the help of duality argu-
ments. Indeed, we have from (6.2.13) and (6.2.16)
k∂t uk k ◦ ≤ c5 (M ). (6.2.20)
L 3 (−(3/4)2 ,0;(W 22 (B(3/4)))0 )
2
◦
Here, W 22 (B(3/4)) is the completion of C0∞ (B(2/3)) in W22 (B(2/3)). By the
compactness arguments used in the previous section, a subsequence can be
selected so that
uk → u in L3 (Q(3/4)). (6.2.21)
Now, taking into account (6.2.21) and (6.2.18), we pass to the limit in
(6.2.12) and find
2 2
c1 ϑ 3 + lim sup Yϑ2 (pk ).
c1 ϑ 3 ≤ e (6.2.22)
k→∞
To pass to the limit in the last term of the right hand side in (6.2.22), let us
decompose the pressure pk so that (see [46, 51, 52])
Here, the first function pk1 is defined as a unique solution to the following
boundary value problem: find pk1 (·, t) ∈ L 3 (B) such that
2
Z Z
pk1 (x, t)∆ψ(x) dx = −ε1k uk (x, t) ⊗ uk (x, t) : ∇2 ψ(x) dx
B B
for all smooth test functions ψ subjected to the boundary condition ψ|∂B = 0.
It is easy to see that
∆pk2 (·, t) = 0 in B (6.2.24)
and, by the coercive estimates for Laplace’s operator with the homogeneous
Dirichlet boundary condition, we get the bound for pk1 :
Z 3
Z
3
|p1 (x, t)| 2 dx ≤ cε1k |uk (x, t)|3 dx.
k 2
(6.2.25)
B B
Passing to the limit in (6.2.22), we show with the help of (6.2.25) that:
2 2
c1 ϑ 3 + lim sup Yϑ2 (pk2 ).
c1 ϑ 3 ≤ e (6.2.26)
k→∞
6.2. ε-REGULARITY THEORY 157
We know that the function pk2 (·, t) is harmonic in B and thus the following
estimate is valid:
Z
3 3
k
sup |∇ p2 (x, t)| ≤ c |pk2 (x, t)| 2 dx
2
x∈B(3/4)
B
The latter inequality, together with (6.2.25), allows us to take the limit in
(6.2.27). As a result, we show that
2 2 2
c1 ϑ 3 ≤ e
c1 ϑ 3 + cϑ 3 . (6.2.28)
If, from the very beginning, c1 is chosen so that
c1 = 2(e
c1 + c),
we arrive at the contradiction. Proposition 2.4 is proved.
Proposition 2.4 admits the following iterations.
Proposition 2.5. Given numbers M > 3 and β ∈ [0, 2/3[, we choose ϑ ∈
]0, 1/2[ so that
2−3β
c1 (M )ϑ 6 < 1. (6.2.29)
Let ε1 (ϑ, M ) = min{ε1 (ϑ, M ), ϑ5 M/2}. If
|(v),1 | < M, Y1 (v, q) < ε1 , (6.2.30)
then, for any k = 1, 2, ...,
ϑk−1 |(v),ϑk−1 | < M, Yϑk−1 (v, q) < ε1 ≤ ε1 ,
2+3β (6.2.31)
Yϑk (v, q) ≤ ϑ 6 Yϑk−1 (v, q).
158 CHAPTER 6. LOCAL REGULARITY THEORY
Then, let M = 2002, β = 1/3, and let ϑ be chosen according to (6.2.29) and
fix.
First, we observe that
and
1 1 2 1 1
Y (z0 , 1/4; U, P ) ≤ c(A 3 + A 3 ), |(U )z0 , 1 | ≤ cA 3 .
4 4 4
Selecting ε0 so that
1 2 1
c(ε03 + ε03 ) < ε1 , cε03 < 2002.
for all z0 ∈ Q(3/4) and for all k = 1, 2, .... Hölder continuity of v on the set
Q(2/3) follows from Campanato’s type condition. Moreover, the quantity
sup |v(z)|
z∈Q(2/3)
We are also going to use abbreviations like A(r) = A(v; 0, r), etc.
Our aim is to prove a version of the Caffarelli-Kohn-Nireberg theorem
(Here, we follow F.-H. Lin’s arguments, see [38]).
Theorem 2.7. Let v and q be a suitable weak solution to the Navier-Stokes
equations in Q. There exists a positive universal constant ε such that if
sup E(r) < ε,
0<r<1
1
Z 23 i
2
+ 3 |v| dx ≤ see (6.2.35) ≤
r2
B(r)
n 3 3 Z 34
≤ c % 4 A 4 (%) |∇ v|2 dx +
B(r)
1h 3 1
Z 21 r 3 i 32 o
+ 3 c% 2 A 2 (%) |∇ v|2 dx + %A(%) ≤
r2 %
B(%)
n r 3 Z 43 h 9
3
2 3 %4 i 3
o
≤c A (%) +
2 |∇ v| dx % +
4
3 A 4 (%) .
% r2
B(%)
h 3 % 94 i 3 Zt0 Z 43 o
+ % 4 + 3 A 4 (%) dt |∇ v|2 dx ≤
r2 2 t0 −r B(x0 ,%)
n r 3 3 h 3 % 94 i 3 1
Z 34 o
2
≤c r A 2 (%) + % 4 + 3 A 4 (%)r 2 |∇ v|2 dz ≤
% r2
Q(%)
n r 3 3 h 3 % 94 i 3 1 3 3
o
2
≤c r A (%) + % + 3 A (%)r E (%) % .
2 4 4 2 4 4
% r2
It remains to notice that
9 h % 32
h 3 %4 i 1 3
% 3 i % 3
% + 4
3 r % =
2 4 + r2 ≤ 2 r2
r2 r r r
and then complete the proof of Lemma 2.8.
162 CHAPTER 6. LOCAL REGULARITY THEORY
1
Z
3
23 Z 31 o
+ 2 |q − [q],R | 2 dz |v|3 dz .
R
Q(R) Q(R)
Since Z
1 2
|v|2 dz ≤ cC 3 (R),
R3
Q(z0 ,R)
we find
2
n 2
1
A(R/2) + E(R/2) ≤ c C 3 (R) + C 3 (R) D03 (R)+
R 2 (6.2.37)
o
+ R12 |v| − [|v|2
] |v| dz .
,R
Q(z0 ,R)
we have
Z0 Z 12 Z 21 Z 31
2 2
S≤c dt |∇ v| dx |v| dx |v|3 dx ≤
−R2 B(R) B(R) B(R)
Z0 Z 21 Z 31
1 1
2 3
≤ c R A (R)
2 2 dt |∇ v| dx |v| dx ≤
−R2 B(R) B(R)
1 1
Z 13 Z0 Z 43 23
3 2
≤ c R A (R)
2 2 |v| dz dt |∇ v| dx ≤
Q(R) −R2 B(R)
1
Z 21
+ 32 1 1 1
≤ R 2 A (R)C (R)R
2 3 3 |∇ v|2 dz ≤
Q(R)
1 1 1
2
≤ c R A 2 (R)C 3 (R)E 2 (R).
Now, (6.2.36) follows from the latter relation and from (6.2.37). Lemma 2.9
is proved.
Now, our goal is to work out an estimate for the pressure.
Lemma 2.10. Let 0 < % ≤ 1. Then
h r 52 % 2 1 i
D0 (r) ≤ c D0 (%) + A 2 (%)E(%) (6.2.38)
% r
for all r ∈]0, %].
Proof We split the pressure in two parts
q = p1 + p2 (6.2.39)
By the mean value theorem for harmonic function p2 , we have for 0 <
r ≤ %/2
3 3 3
sup |p2 (x, t) − [p2 ],r (t)| 2 ≤ cr 2 sup |∇p2 (x, t)| 2
x∈B(r) x∈B(%/2)
6.2. ε-REGULARITY THEORY 165
r Z 32
≤c 4 |p2 (x, t) − [p2 ],% (t)|dx (6.2.43)
%
B(%)
c r 23
Z
3
≤ 3 |p2 (x, t) − [p2 ],% (t)| 2 dx.
% %
B(%)
Z0
c 1 r 32
Z Z
c 3
3 3
≤ 2 |p1 | dz + 2 3
2 r |p2 (x, t) − [p2 ],% (t)| 2 dx
r r % %
Q(r) −r2 B(%)
% 2 1
r 52 1 Z 3
≤c E(%)A 2 (%) + c |p 2 − [p 2 ],% | 2 dz
r % %2
Q(%)
% 2 1
r 52 h 1 Z 3
≤c E(%)A (%) + c
2
2
|q − [q],% | 2 dz
r % %
Q(%)
Z
1 3
i
+ 2 |p1 − [p1 ],% | dz 2
%
Q(%)
h r 25 % 2 1
i
≤c D0 (%) + E(%)A 2 (%)
% r
So, inequality (6.2.38) is shown. Lemma 2.10 is proved.
Proof of Theorem 2.7 It follows from (6.2.34), (6.2.38), and the as-
sumptions of Theorem 2.7 that:
h % 3 3 3
r 3 3 i
C(r) ≤ c A 4 (%)ε 4 + A 2 (%) (6.2.44)
r %
and h r 25 % 2 i
1
D0 (r) ≤ c D0 (%) + A (%)ε .
2 (6.2.45)
% r
Introducing the new quantity
3
E(r) = A 2 (r) + D02 (r),
166 CHAPTER 6. LOCAL REGULARITY THEORY
h 3 1 3
i
≤ c C(2r) + D02 (2r) + A 4 (2r)C 2 (2r)ε 4 . (6.2.46)
for any natural numbers k and for any 0 < % ≤ 1. Letting % = 1, we find
k
E(ϑk ) ≤ ϑ 2 E(1) + cG (6.2.51)
for the same values of k. It can be easily deduced from (6.2.51) that
1
E(r) ≤ c(r 2 E(1) + G(ε)) (6.2.52)
where ε0 is a number of Lemma 2.2. Since v and q − [q],r0 are a suitable weak
solution in Q(r0 ), Lemma 2.2 and the Navier-Stokes scaling yield required
statement. Theorem 2.7 is proved.
Now, we are in a position to speculate about ε-regularity theory. Quan-
tities that are invariant with respect to the Navier-Stokes scaling
play the crucial role in this theory. By the definition, such quantities are
defined on parabolic balls Q(r) and have the property
F (v, q; r) = F (v λ , q λ ; r/λ).
There are two types of statements in the ε-regularity theory for suitable
weak solutions to the Navier-Stokes equations and the first one reads:
Suppose that v and q are a suitable weak solution to the Navier-Stokes
equations in Q. There exist universal positive constants ε and {ck }∞ k=1 such
that if F (v, q; 1) < ε then |∇k v(0)| < ck , k = 0, 1, 2, .... Moreover, the
function z 7→ ∇k v(z) is Hölder continuous (relative to the parabolic metric)
with any exponent less 1/3 in the closure of Q(1/2).
An important example of such kind of quantities appears in Lemma 2.2
and is as follows:
Z
1 3
F (v, q; r) = 2 |v|3 + |q| 2 dz.
r
Q(r)
provided
3 2
+ =1
s l
and s ≥ 3. Local regularity results connected with those quantities have been
proved partially by J. Serrin in [63] and then by M. Struwe in [65] for the
6.3. BOUNDED ANCIENT SOLUTIONS 169
velocity field v having finite energy even with no assumption on the pressure.
However, in such a case, we might loose Hölder continuity.
Energy scale-invariant quantities present an important example of the
second kind of quantities. Some of them have been listed above. For more
examples of scaled energy quantities, we refer to the paper [19]. It is worthy
to note that the second statement applied to the scaled dissipation E is the
famous Caffarelli-Kohn-Nirenberg theorem, which is Theorem 2.7. It gives
the best estimate for Hausdorff’s dimension of the singular set for a class of
weak Leray-Hopf solutions to the Cauchy problem. A certain generalization
of the Caffarelli-Kohn-Nirenber theorem itself has been proved in [50] and is
formulated as follows.
Proposition 2.11. Let v and q be a suitable weak solution to the Navier-
Stokes equations in Q. Given M > 0, there exists a positive number ε(M )
having the property: if two inequalities lim supr→0 E(r) < M and
lim inf E(r) < ε(M )
r→0
∞
for any w ∈ C0,0 (Q− ).
170 CHAPTER 6. LOCAL REGULARITY THEORY
Lemma 3.2. For any F = L∞ (Rn ; Mn×n ), there exists a unique function
qF ∈ BM O(Rn ) such that [qF ]B(1) = 0 and
To state Lemma 3.2, the following notation has been used. [f ]Ω is the
mean value of a function f over a spatial domain Ω ∈ Rn . The mean value
of a function g over a space-time domain Q is denoted by (g)Q .
Lemma 3.2 is proved with the help of the singular integral theory, see
[64]. Proof of Lemmata 3.3 and 3.4 can be found, for example, in [34] and
[33].
6.3. BOUNDED ANCIENT SOLUTIONS 171
If we let
F (·, t) = u(·, t) ⊗ u(·, t),
then, by Lemma 3.2, there exists an unique function
which satisfies the condition [pu⊗u ]B(1) (t) = 0 and the equation
for all t ≤ 0.
To state the main result of this section, we introduce the space
If we let ut0 (x, t) = u(x, t) + bt0 (t) in Qt0 = Rn ×]t0 − 1, t0 [, then, for any
number m > 1 and for any point x0 ∈ Rn , the uniform estimate
is valid and, for a.a. z = (z, t) ∈ Qt0 , functions u and ut0 obey the system of
equations
Remark 3.6. The first equation of the latter system can be rewritten in the
following way
in Qt0 in the sense of distributions. So, the real pressure field in Qt0 is the
following distribution pu⊗u + b0t0 · x.
172 CHAPTER 6. LOCAL REGULARITY THEORY
ω = ∇⊥ u = u2,1 − u1,2 ∈ Wm
2,1
(Q− ) := {ω, ∇ω, ∇2 ω, ∂t ω ∈ Lm (Q− )}
and
∂t ω + u · ∇ω − ∆ω = 0 a.e. in Q− .
If n = 3, then
2,1
ω = ∇ ∧ u ∈ Wm (Q− ; R3 )
and
∂t ω + u · ∇ω − ∆ω = ω · ∇u a.e. in Q− .
Remark 3.10. We could analyse smoothness of solutions to the vorticity
equations further and it would be a good exercise. However, regularity results
stated in Theorem 3.9 are sufficient for our purposes.
Remark 3.11. By the embedding theorems, see [33], functions ω and ∇ω
are Hölder continuous in Q− and uniformly bounded there.
6.3. BOUNDED ANCIENT SOLUTIONS 173
It is easy to see that in our case there exists a smooth function pε with the
following property
pε = pF ε + peε . (6.3.2)
It is not difficult to show that the function ∇pF ε is bounded in Qt−0 (exercise).
So, it follows from (6.3.1) and (6.3.2) that
So, we have
in Qt−0 .
Now, let us introduce new functions
Zt
bεt0 (t) = aε (τ )dτ, t0 − 1 ≤ t ≤ t0 ,
t0 −1
in Qt−0 .
Fix an arbitrary cut-off function ϕ so that
Z Zt Z
= ϕ2x0 (x)|vε (x, t0 − 1)| dx + 2
∆ϕ2x0 |vε |2 dxdt0 +
Rn t0 −1 Rn
Zt Z
+ (pF ε − [pF ε ]B(x0 ,2) )vε · ∇ϕ2x0 dxdt0 +
t0 −1 Rn
Zt Z
+ (F ε − [F ε ]B(x0 ,2) ) : ·∇(ϕ2x0 vε )dxdt0 .
t0 −1 Rn
and taking into account that vε (·, t0 − 1) = uε (·, t0 − 1) and |uε (·, t0 − 1)| ≤ 1,
we can estimate the right hand side of the energy identity in the following
way
Zt
I(t) ≤ c(n) + c(n) αε (t0 )dt0 +
t0 −1
Zt0 Z 12 Z t 12
+c(n) |pF ε − [pF ε ]B(x0 ,2) | dxdt 2
αε (t0 )dt0 +
t0 −1 B(x0 ,2) t0 −1
Zt0 Z 21 Z t Z
+c(n) ε ε
|F − [F ]B(x0 ,2) | dxdt 2
ϕ2x0 |∇vε |2 dxdt0 + (6.3.5)
t0 −1 B(x0 ,2) t0 −1 Rn
Zt 21
+ αε (t0 )dt0 , t0 − 1 ≤ t ≤ t0 .
t0 −1
ε
Next, since |F | ≤ c(n), we find two estimates
Zt0 Z
|F ε − [F ε ]B(x0 ,2) |2 dxdt ≤ c(n)
t0 −1 B(x0 ,2)
and
Zt0 Z
|pF ε − [pF ε ]B(x0 ,2) |2 dxdt ≤ c(n)kpF ε k2L∞ (−∞,t0 ;BM O(Rn ))
t0 −1 B(x0 ,2)
and
Zt0 Z Zt0
sup |∇vε |2 dxdt ≤ c(n) 1 + αε (t)dt .
x0 ∈Rn
t0 −1 B(x0 ,1) t0 −1
176 CHAPTER 6. LOCAL REGULARITY THEORY
the estimate
Zt0 Z
kbt0 kL∞ (t0 −1,t0 ) + sup |∇u|2 dxdt ≤ c(n) < +∞ (6.3.7)
x0 ∈Rn
t0 −1 B(x0 ,1)
f = divF = u · ∇u ∈ L2 (Q− ; Rn ).
∇pu⊗u ∈ L2 (Q− ; Rn ).
Using the invariance with respect to shifts and Lemma 3.4, one can conclude
that
ut0 ∈ W22,1 (Q(z0 , τ2 ); Rn ), 1/2 < τ2 < τ1 = 1,
and, moreover, the estimate
for
1 1 1
= − , m1 = 2.
m2 m1 n + 2
By lemma 3.3, by shifts, and by scaling, for 1/2 < τ30 < τ2 , we have the
following estimate
Z h Z i
m2 0
|∇pu⊗u (·, t)| dx ≤ c(n, τ2 , τ3 ) |∇u(·, t)|m2 dx + 1 .
B(x0 ,τ30 ) B(x0 ,τ30 )
and
kut0 kWm2,1 (Q(z0 ,τ3 )) ≤ c(n, τ3 , τ30 ).
2
for
1 1 1
= − .
m3 m2 n + 2
Now, let us take an arbitrary large number m > 2 and fix it. Find α as
an unique solution to the equation
1 1 α
= − .
m 2 n+2
178 CHAPTER 6. LOCAL REGULARITY THEORY
Next, we let k0 = [α] + 1, where [α] is the entire part of the number α. And
then we determine the number mk0 +1 satisfying the identity
1 1 k0
= − .
mk0 +1 2 n+2
and
kut0 kWm2,1 (Q(z0 ,τk0 +1 )) ≤ c(n, m).
k0 +1
Thanks to the inequality τk > 1/2 for any natural numbers k, we complete
the proof of Theorem 3.5.
Proof of Theorem 3.9 Let us consider the case n = 3. The case n = 2
is in fact easier. So, we have
∂t ω − ∆ω = ω · ∇u − u · ∇ω ≡ f.
Take an arbitrary number m > 2 and fix it. By Theorem 3.5, the right hand
side has the following property
for some r ∈]r1 , 1[. It tends to infinity as time t goes to zero from the left
since the origin is a singular point of v. Thanks to the obvious properties of
the function M , one can choose parameters of the scaling in a particular way
letting λk = 1/Mk , where a sequence Mk is defined as
with x(k) ∈ B(r1 ) for sufficiently large k. Before discussing what happens if
k tends to infinity, let us introduce a subclass of bounded ancient (backward)
solutions playing an important role in the regularity theory of the Navier-
Stokes equations.
∂t u + div u ⊗ u − ∆u + ∇p = 0,
div u = 0
Z0
3
max |∇q(x, t)| 2 dt ≤ c1 < ∞. (6.4.1)
x∈B 2
−τ22
Proof of the first statement can be done by induction and found in [13], [31],
and [40]. The second statement follows directly from the first one and the
pressure equation: ∆q = −vi,j vj,i .
Now, let us decompose the pressure q = q1 + q2 . For q1 , we have
h i
∆q1 (x, t) = −div div χB (x)v(x, t) ⊗ v(x, t) , x ∈ R3 , −1 < τ < 0,
Z0
3
max |∇q1 (x, t)| 2 dt ≤ c2 < ∞, (6.4.2)
x∈B 3
−τ22
where B3 = {r2 < r3 < |x| < a3 < a2 }. From (6.4.1) and (6.4.2), it follows
that
Z0
3
max |∇q2 (x, t)| 2 dt ≤ c3 < ∞. (6.4.3)
x∈B 3
−τ22
Z0 5
3
sup |∇y pk2 (y, s)| 2 ds ≤ c3 λk2 . (6.4.5)
y∈B(−xk /λk ,r4 /λk )
−(τ22 −tk )/λ2k
∆y pk1 (y, s) = −divy divy (χB(−xk /λk ,1/λk ) (y)u(k) (y, s) ⊗ u(k) (y, s)), y ∈ R3 ,
for all possible values of s. For such a function, we have the standard estimate
for all s ∈]−(1−tk )/λ2k , 0[. It is valid since |u(k) | ≤ 1 in B(−xk /λk , 1/λk )×]−
(1 − tk )/λ2k , 0[.
6.4. MILD BOUNDED ANCIENT SOLUTIONS 183
pk1 (y, s) = pk1 (y, s) − [pk1 ]B(1) (s) pk2 (y, s) = pk2 (y, s) − [pk2 ]B(1) (s)
Using the same bootstrap arguments, we can show that the following estimate
is valid:
ku(k) kC α (Q(a/2) ≤ c5 (c2 , c3 , c4 , a)
for some positive number α < 1/3. Indeed, the norm ku(k) kC α (Q(a/2)) is
estimated with the help of norms ku(k) kL∞ (Q(a))) and kpk kL 3 (Q(a)) , where
2
pk = pk1 + pk2 . Hence, using the diagonal Cantor procedure, we can select
subsequences such that for some positive α and for any positive a
u(k) → u in C α (Q(a)),
Hence, ∇p2 = 0 in Q(a) for any a > 0. So, p2 (y, s) is identically zero. This
allows us to conclude that the pair u and p1 is a solution to the Navier-Stokes
equations in the sense of distributions and thus u is a nontrivial mild bounded
ancient solution satisfying the condition |u(0, 0)| = 1 and the estimate |u| ≤ 1
in Q− .
184 CHAPTER 6. LOCAL REGULARITY THEORY
(ii) If
g 0 = min{ sup A(v; r), sup C(v; r), sup E(v; r)} < +∞,
0<r<1 0<r<1 0<r<1
for any divergence free function w from C0∞ (Q− ). Assume that
Z0 Z sl
sup Ms,l (u; r) = |u(x, t)|s dx <∞
0<r<∞
−∞ R3
Proof Let us consider the simplest case of the regular LPS quantity
M5,5 . By the pressure equation, we may assume
Z0 Z
5
(|u|5 + |p| 2 )dxdt < +∞.
∞ R3
for any x0 ∈ R3 , any R > 0, and any t0 ≤ T with some universal constant c.
In turn, the ε-regularity theory ensures the inequality
c
|u(x0 , t0 )| <
R
with another universal constant c. Tending R → ∞, we get u(·, t) = 0 as
t ≤ T . One can repeat more or less the same arguments in order to show
that in fact u is identically zero on R3 ×] − ∞, 0].
6.5.2 2D case
In two-dimensional case, we have the following Liouville type theorem.
Theorem 5.2. Assume that n = 2 and u is an arbitrary bounded ancient
solution. Then u(x, t) = b(t) for any x ∈ R2 .
To prove the above statement, we start with an auxiliary lemma.
Lemma 5.3. Let functions
2,1 2,1
ω ∈ Wm (Q− ) = {u ∈ Wm,loc (Q− ) : sup kukWm2,1 (Q(z0 ,1)) < ∞},
z0 ∈Q−
6.5. LIOUVILLE TYPE THEOREMS 187
∂t ω + u · ∇ω − ∆ω = 0 in Q−
Then
w(z) = w(z0 ) in Q(z0 , R).
Proof of Lemma 5.3 ( [27]) In fact, we shall prove even a stronger
result. Let zk be a sequence of points in Q− such that
ω(zk ) → M.
We state that
inf ω(z) → M.
z∈Q(zk ,R)
Indeed, assume that this statement is false. Then, we can find a number
ε > 0 and a sequence of points zk0 ∈ Q(zk , R) such that
ω(zk0 ) ≤ M − ε.
188 CHAPTER 6. LOCAL REGULARITY THEORY
kω k kWm2,1 (Q(R)) ≤ c,
|uk | ≤ 1 in Q(R),
with a constant c that is independent of k. Moreover, we have
∂t ω k + uk · ∇ω k − ∆ω k = 0 in Q(R),
ωk * ω in Wm2,1 (Q(R)),
?
uk *u in L∞ (Q(R); R2 ),
ωk → ω in C(Q(R)),
ω(z) ≤ ω(0) = M z ∈ Q(R), (6.5.8)
ω(z∗ ) ≤ M − ε, (6.5.9)
where z∗ ∈ Q(R). Clearly, ω ∈ Wm2,1 (Q(R)) and
∂t ω + u · ∇ω − ∆ω = 0 in Q(R).
ω(z) = M z ∈ Q(R),
sup ω(z) = M ≤ 0.
z∈Q−
6.5. LIOUVILLE TYPE THEOREMS 189
To this end, assume that the latter statement is wrong and in fact
M > 0.
ϕ≡1 in B(R/2).
By Lemma 5.3, for an arbitrary number R > 0, there exists a point z0R =
(x0R , t0R ) with t0R ≤ 0 such that
On the other hand, since ω = u2,1 − u1,2 , we have after integration by parts
Z
A(R) = (ϕx0R ,2 u1 − ϕx0R ,1 u2 )dz ≤
Q(z0R ,R)
≤ cR3 ,
Z0 Z h i m2
2 2 2 2 2
|ωφ,%% | + 2|ωϕ,%ϕ | + |ωϕ,33 | + 2|(ωϕ /%),% | + 2|(ωϕ /%),3 | %d%dϕ ≤
−T C(a)
where
1 1
∆η = (%η,% ),% + η,33 = η,%% + η,33 + η,% .
% %
Let us make the change of variables
y = (y 0 , y5 ) ∈ R5 , y 0 = (y1 , y2 , y3 , y4 ),
q
% = |y 0 | = y12 + y22 + y32 + y42 , y5 = x3 .
Then after simple calculations, we see that a new function
U (y, t) = (U1 (y, t), U2 (y, t), U3 (y, t), U4 (y, t), U5 (y, t)),
where
u% (%, x3 , t)
Ui (y, t) = yi , i = 1, 2, 3, 4, U5 (y, t) = u3 (%, x3 , ).
%
Obviously, the function U is bounded in Q5− . However, previous arguments
show that ∇5 U is a bounded function as well. Indeed,we have
|∇5 U (y, t)| ≤ c(|∇u(x, t)| + |u% (%, x3 , t)|/%) ≤ c|∇u(x, t)| ≤ c < +∞
|U |, |∇5 U | ∈ L∞ (Q5− ).
for any g ∈ C0∞ (Q5− ). In the way, explained in the previous section, one can
show that, for any m > 1,
2,1
f ∈ Wm (Q5− )
192 CHAPTER 6. LOCAL REGULARITY THEORY
We wish to show that M ≤ 0. Assume that it is not so, i.e., M > 0, and we
can apply Lemma 5.3 in our five-dimensional setting. Then, for any R > 0,
there exists a point yR in R5 and a moment of time tR ≤ 0 such that
f (y, t) ≥ M/2, (y, t) ∈ Q((yR , tR ), R) = B(yR , R)×]tR − R2 , tR [,
where B(yR , R) = {|y − yR | < R}.
By our assumptions,
0 < M0 = sup ωϕ (|x0 |, x3 , t) < +∞.
x∈R3 , t≤0
By the assumptions,
q
sup g = 1, u2% + u23 ≤ A/%, |g| ≤ A/M in Q
e− (6.5.13)
Q
e−
and
(%u% ),% + (%u3 ),3 = 0 in Q
e− .
To formulate the lemma below, we abbreviate
Π = Π(%1 , %2 ; h1 , h2 ), e = Π×]t − t1 , t[.
Q
194 CHAPTER 6. LOCAL REGULARITY THEORY
δ = δ(Π, t, t1 , A, M, ε) ≤ ε
such that if
sup g(x, t) > 1 − δ,
x∈Π
then
inf g(z) > 1 − ε.
z∈Q
e
Proof If we assume that the statement of the lemma is false, then there
must exist a number ε0 > 0 such that, for any natural k, one can find
sequences with the following properties:
in Q
e− and the relations
q
sup g k = 1, |uk% |2 + |uk3 |2 ≤ A/%, |g k | ≤ A/M in Q
e− .
Q
e−
By (6.5.14), there are points (%k , xk3 , t), with (%k , xk3 ) ∈ Π, and (%0k , x0k3 , t0k ) ∈
Q
e such that
?
uk * u e− ; R2 ),
in L∞ (Q
6.5. LIOUVILLE TYPE THEOREMS 195
gk * g in Wm2,1 (Q
e2 ),
e2 = Π2 ×]t − t2 , t[, Π2 = Π(%2 , %2 ; h2 , h2 ) c Π, t2 > t1 , and m >> 1.
where Q 1 2 1 2
Then we have
gk → g in C(Q e2 ) (6.5.16)
and
q
sup g ≤ 1, |u% |2 + |u3 |2 ≤ A/%, |g| ≤ A/M in Q
e2 , (6.5.17)
Q
e2
and
∂t g + (u% + 1/%)g ,% + u3 g ,3 − g ,%% − g ,33 = 0 in Q
e2 .
According to (6.5.15) and (6.5.16),
where
(%k , xk3 , t) → (%0 , x03 , t), (%0k , x0k3 , t0k ) → (%00 , x003 , t00 )
and points (%0 , x03 , t) and (%00 , x003 , t00 ) belong to the closure of the set Q.
e
Clearly, by (6.5.17),
1−ε≤g ≤1 e0 = Π0 ×] − T, 0[,
on Q (6.5.19)
where Π0 = Π(1, R; −L, L). To explain this, we let δ∗ = δ(Π0 , 0, −T, A, M, ε).
Obviously, there exists a point (%0 , x03 , t0 ) ∈ Q
e− such that
It is easy to see
and, therefore, %0 > M/2 > 0. Then one can scale our functions so that
1 − gλ (1, 0, 0) < δ∗ ,
By Lemma 5.8,
1 − ε ≤ gλ ≤ 1 on Q
e0 .
It is always deemed that this operation has been already made and script λ
is dropped. It is important to note two things. Numbers R, L, T , and ε are
in our hands and we cannot use the fact |u| ≤ 1 any more since after scaling
|u| ≤ %0 (R, T, L, A, M, ε).
We choose a cut-off function
ψ(%) = 1 0 ≤ % ≤ R − 1, ψ(%) = 0 % ≥ R,
So, we have
Z
I0 = ∂t g + u% g,% + u3 g,3 − ∆g Φ%d%dx3 dt = I00 =
Q
e−
Z
g,%
= −2 Φ%d%dx3 dt (6.5.20)
%
Q
e−
We replace g with g − 1 in the left hand side of (6.5.20) and, after integration
by parts, have
Z
1
I0 = ∂t Φ + u% Φ,% + u3 Φ,3 + ∆Φ (1 − g)dxdt.
4π
Q−
We know that 1 − g ≤ ε in Q
e0 . Then, by 6.5.13,
Z0 ZL Z1
I0 ≥ (1 − g) ∂t Φ + u3 Φ,3 + Φ,33 %d%dx3 dt + εC0 (R, T, L, A, M )
−T −L 0
Z0 ZL Z0 ZL ZR
I00 = −2 g(0, x3 , t)Φ(0, x3 , t)dx3 dt + 2 Φ,% gd%dx3 dt.
−T −L −T −L 0
The first term on the right hand side of the latter identity is equal to zero.
The second one is estimated from above
Z0 ZL ZR Z0 ZL ZR
I00 =2 Φ,% d%dx3 dt + 2 Φ,% (g − 1)d%dx3 dt ≤
−T −L 0 −T −L 0
Z0 ZL ZR
≤2 Φ,% d%dx3 dt + εC00 (R, T, L, A, M ) =
−T −L 0
198 CHAPTER 6. LOCAL REGULARITY THEORY
Z0 ZL
= −2 Φ(0, x3 , t)dx3 dt + εC00 (R, T, L, A, M ) <
−T −L
inf %uϕ = m ≥ 0.
Q
e−
v = v% e% + v3 e3 vb = vϕ eϕ
for v = v% e% + vϕ eϕ + v3 e3 .
Here, we follow paper [59], where results are stated for the canonical
domain Q = Q(1). The general case can be deduced by re-scaling.
where Z 103
10
M= |v| dz
3 + 1.
Q(3/4)
Remark 6.4. Under the assumptions of Lemma 6.3, the pair v and q is a
suitable weak solution to the Navier-Stokes equations in Q. Hence, the right
hand side of (6.6.4) is bounded from above.
|x0 − x00 | < R, |x3 − x03 | < R}, C(R) = C(0, R), C = C(1);
Q(z0 , R) = C(x0 , R)×]t0 − R2 , t0 [, Q(R) = Q(0, R), Q = Q(1),
200 CHAPTER 6. LOCAL REGULARITY THEORY
Zt0 Z sl
1 s
Ms,l (z0 , r; v) = κ |v| dx dt,
r
t0 −r2 C(x0 ,r)
where κ = l( 3s + 2l − 1) and s ≥ 1, l ≥ 1.
The following statement is proven in a similar way as Proposition 4.4, see
details in [59].
Lemma 6.5. Under assumptions of Theorem 6.1, we have the estimate
A(zb , r; v) + E(zb , r; v) + C(zb , r; v) + D(zb , r; q) ≤ C1 < +∞ (6.6.5)
for all zb and for all r satisfying conditions
1 1
zb = (be3 , 0), b ∈ R,
|b| ≤ , 0<r< . (6.6.6)
4 4
A constant C1 depends only on the constant C in (6.6.1), kvkL3 (Q) , and
kqkL 3 (Q) .
2
Proposition 6.7. Assume that all conditions of Theorem 6.1 hold. Then
C1
|v(x, t)| ≤ (6.6.7)
|x0 |
Pr10 = {r0 < |x0 | < 2r0 , |x3 | < r0 }, Pr20 = {r0 /4 < |x0 | < 3r0 , |x3 | < 2r0 }.
+cC(zb0 , 3r0 ; v) + cD(zb0 , 3r0 ; q) ≤ Φ 4cC1 .
It remains to apply Lemma 6.5 and complete the proof of the proposition.
Now, we proceed with proof of Theorems 6.1 and 6.2. Using Lemmata 6.3,
6.5, 6.6, Remark 6.4, Proposition 6.7 and scaling arguments, we may assume
202 CHAPTER 6. LOCAL REGULARITY THEORY
(without loss of generality) that our solution v and q have the following
properties:
sup A(0, r; v) + E(0, r; v) + C(0, r; v) + D(0, r; q) = A1 < +∞, (6.6.9)
0<r≤1
We may also assume that the function v is Hölder continuous in the closure
of the set C×] − 1, −a2 [ for any 0 < a < 1.
Introducing functions
We scale our functions v and q so that scaled functions keeps axial symmetry:
1
uk (y, s) = λk v(λk y 0 , x3k + λk y3 , tk + λ2k s), λk = ,
Mk
According to (6.6.10),
|yk0 | ≤ A2
for all k ∈ N. Thus, without loss of generality, we may assume that
and by (6.6.9), we can select subsequences (still denote as the entire sequence)
such that
?
uk * u in L∞ (Q(a)), (6.6.14)
and
pk * p in L 3 (Q(a))
2
sup A(0, r; u) + E(0, r; u) + C(0, r; u) + D(0, r; p) ≤ A1 ,
0<r<+∞
Now, our aim is to show that u and p satisfy the Navier-Stokes equations
Q− and u is smooth enough to obey the identity
To this end, we fix an arbitrary positive number a > 0 and consider numbers
k so big that a < Mk /4. We know that uk satisfies the nonhomogeneous heat
equation of the form
where F k = uk ⊗ uk + pk I and
where f k = −uk · ∇uk is the right hand side having the property
kf k k 3 ,Q(3a) ≤ c2 (a).
2
Then, according to the local regularity theory for the Stokes system, see
Chapter IV, we can state that
k∂t uk k3, 3 ,Q(2a) + k∇2 uk k3, 3 ,Q(2a) + k∇pk k3, 3 ,Q(2) ≤ c5 (a).
2 2 2
k∂t uk k6, 3 ,Q(a) + k∇2 uk k6, 3 ,Q(a) + k∇pk k6, 3 ,Q(a) ≤ c7 (a).
2 2 2
p(k) * p
A proof of this proposition and similar facts can be found in [13], [54],
[59], and [55]. Let us comment the last statement of Proposition 7.1. Indeed,
if z = 0 is a singular point of v, the ε-regularity theory gives us
Z
1 3
2
(|v|3 + |q| 2 )dz > ε > 0
r
Q(r)
for all 0 < r < 1 and for some universal constant ε. Making the inverse
change of variables, we find
1
R 3
a2
(|u(k) |3 + |p(k) | 2 )dyds =
Q(a)
1
R 3
λ2k a2
(|v|3 + |q| 2 )dxds > ε > 0
Q(λk a)
for each fixed radius a > 0 and for sufficiently large natural number k. We
cannot simply pass to the limit in the latter identity since it is not clear
whether the pressure p(k) converges strongly. This is quite typical issue when
one works with sequences of weak solutions to the Navier-Stokes equations.
In order to treat this case, let us split the pressure p(k) into two parts. The
first part is completely controlled by the velocity field u(k) while the second
one is a harmonic function with respect to the spatial variables. This, to-
gether with a certain boundedness of the sequence p(k) , implies (6.7.1). For
more details, we recommend papers [54] and [55].
We do not know whether local energy ancient solutions with bounded
scaled energy quantities are identically equal to zero. However, there are
some interesting cases for which the answer is positive. Let us describe them.
Our additional standing assumption of this section can be interpreted as
a restriction on the blowup profile of v and has the form
Z
1 9
15 |v(x, 0)| 8 dx → 0 (6.7.2)
r8
B(r)
u(·, 0) = 0, (6.7.3)
Now, by Proposition 7.1 and by (6.7.2), the right hand side of the latter
inequality tends to zero and this completes the proof of (6.7.3).
In a view of (6.7.3), one could expect that our local energy ancient solution
is identically equal to zero. We call this phenomenon a backward uniqueness
for the Navier-Stokes equations. So, if the backward uniqueness takes place
or at least our ancient solution is zero on the time interval ] − 3/4, 0[, then
(6.7.1) cannot be true and thus, by Proposition 7.1, the origin z = 0 is not a
singular point of the velocity field v.
The crucial point for understanding the backward uniqueness for the
Navier-Stokes equations is a similar phenomenon for the heat operator with
lower order terms. The corresponding statement for the partial differential
inequality involving the backward heat operator with lower order terms has
been proved in [13] and reads:
Theorem 7.2. Assume that we are given a function ω defined on Rn+ ×]0, 1[,
where Rn+ = {x = (xi ) ∈ Rn , xn > 0}. Suppose further that they have the
properties:
ω and the generalized derivatives ∇ω, ∂t ω, and ∇2 ω are square integrable
over any bounded subdomain of Rn+ ×]0, 1[;
for all x ∈ Rn+ , for all 0 < t < 1, and for some M > 0;
ω(x, 0) = 0 (6.7.6)
The interesting feature of Theorem 7.2 is that there has been made no
assumption on ω on the boundary xn = 0. In order to prove the theorem,
two Carleman’s inequalities have been established, see details in [13] and
[12] and the appendix. For the further improvements of the above backward
uniqueness result, we refer to the interesting paper [10].
Theorem 7.2 clearly indicates what one should add to (6.7.3) in order
to get the backward uniqueness for ancient solutions to the Navier-Stokes
equations. Obviously, we need more regularity for sufficiently large x and
a right decay at infinity. One can hope then to apply Theorem 7.2 to the
vorticity equation
∂t ω − ∆ω = ω · ∇u − u · ∇ω, ω = ∇ ∧ u,
for all |x| > R, for all −1 < t < 0, and for some constant c and try to
figure out what follows from (6.7.7). It is not difficult to see that (6.7.3) and
(6.7.7) implies (6.7.6) and (6.7.4), (6.7.5), respectively. At last, the linear
theory ensures the validity of first condition in Theorem 7.2, see details in
[47]. So, Theorem 7.2 is applicable and by it, ω(x, t) = 0 for all |x| > R and
for −1 < t < 0. Using unique continuation across spatial boundaries, see,
for instance, [13], we deduce ω(x, t) = ∇ ∧ u(x, t) = 0 for all x ∈ R3 and,
say, for −5/6 < t < 0. Since u is divergence free, it is a harmonic functionp in
3
R depending on t ∈] − 5/6, 0[ as a parameter. Therefore, for any a > 5/6
and for any x0 ∈ R3 , by the mean value theorem for harmonic functions, we
have Z
2 1
sup |u(x0 , t)| ≤ c sup 3
|u(x, t)|2 dx
−5/6<t<0 −5/6<t<0 a
B(x0 ,a)
a + |x0 |
Z
1
≤c sup 3
|u(x, t)|2 dx ≤ c A(u, a + |x0 |).
−5/6<t<0 a a3
B(|x0 |+a)
6.7. BACKWARD UNIQUENESS 209
Applying Proposition 7.1 once again and taking into account properties of
harmonic functions, one can conclude that
Our further arguments rely upon the ε-regularity theory. Indeed, letting,
say, T = 4, one can find R > 4 so that
Z0 Z
3
(|u|3 + |p| 2 )dxdt < ε.
−4 R3 \B(R/2)
6.8 Comments
The section is essentially the context of my lectures on the local regularity
theory given in Summer School, Cetraro, Italy, 2010, see [58]. It contains
an introduction to the so-called ε-regularity theory in the spirit of the paper
[13], see also [50] for some generalizations. A big part of this section is
an alternative approach to derivation of mild bounded ancient solutions and
Liouville type theorems for them presented in [27]. Here, we follow the paper
[59] although proofs of Liouville type theorems is essentially the same as in
[27].
Chapter 7
Behaviour of L3-Norm
has been proven in the previous chapter. The aim of this chapter is to
improve (7.1.5). At the moment, the best improvement of (7.1.5) is given by
the following theorem.
211
212 CHAPTER 7. BEHAVIOUR OF L3 -NORM
holds true.
Let us briefly outline our proof of Theorem 1.1 that relays upon ideas
developed in [54]-[56]. In particular, in [54], a certain type of scaling has
been invented, which, after passing to the limit, gives a special non-trivial
solution to the Navier-Stokes equations provided there is a finite time blow
up. In [55] and [56], it has been shown that the same type of scaling and
blowing-up can produce the so-called Lemarie-Rieusset local energy solutions,
introduced and carefully studied in the monograph [35], see Appendix B for
details. It turns out to be that the backward uniqueness technique is still
applicable to those solutions. Although the theory of backward uniqueness
itself is relatively well understood, its realization is not an easy task and based
on delicate regularity results for the Navier-Stokes equations. Actually, there
are two main points to verify: solutions, produced by scaling and blowing-up,
vanish at the last moment of time and have a certain spatial decay. The first
property is easy when working with L3 -norm while the second one is harder.
However, under certain restrictions, the required decay is a consequence of the
Lemarie-Rieusset theory. So, the main technical part of the whole procedure
is to show that scaling and blowing-up lead to local energy solutions. On that
way, a lack of compactness of initial data of scaled solutions in L2,loc is the
main obstruction. This is why the same theorem for a stronger scale-invariant
1
norm of the space H 2 is easier. The reason for that is a compactness of the
corresponding embedding, see [41] and [55].
In this chapter, we are going to show that, despite of a lack of compactness
in L3 -case, the limit of the sequence of scaled solutions is still a local energy
solution, for which a spatial decay takes place. Technically, this can be done
by splitting each scaled solution into two parts. The first one is a solution
to a non-linear problem but with zero initial data while the second one is a
solution of a linear problem with weakly converging nonhomogeneous initial
data.
We also prove (7.1.4) as a by-product of the proof of Theorem 1.1, see
Section 4.
7.2. ESTIMATES OF SCALED SOLUTIONS 213
u(k) (y, s) = λk v(x, t), p(k) (y, s) = λ2k q(x, t), (7.2.3)
x = λk y, t = T + λ2k s,
r
T − tk
λk =
S
and a positive parameter S < 10 will be defined later.
By the scale invariance of L3 -norm, u(k) (·, −S) is uniformly bounded in
L3 (R3 ), i.e.,
sup ku(k) (·, −S)k3 = M < ∞. (7.2.4)
k∈N
Let us decompose our scaled solution u(k) into two parts: u(k) = v (k) +w(k) .
Here, w(k) is a solution to the Cauchy problem for the Stokes system:
Obviously, (7.2.5) can be reduced to the Cauchy problem for the heat equa-
tion so that the pressure r(k) = 0 and w(k) can be worked out with the help
of the heat potential. The estimate below is well-known, see, for example
[24],
sup{kw(k) kL5 (R3 ×]−S,0[ + kw(k) kL3,∞ (R3 ×]−S,0[ } ≤ c(M ) < ∞. (7.2.6)
k
214 CHAPTER 7. BEHAVIOUR OF L3 -NORM
Now, our aim is to show that, for a suitable choice of −S, we can prove
unform estimates of v (k) and p(k) in certain spaces, pass to the limit as k → ∞,
and conclude that the limit functions u and p are a local energy solution to the
Cauchy problem for the Navier-Stokes system in R3 ×] − S, 0[ associated with
the initial data, generated by the weak L3 -limit of the sequence u(k) (·, −S).
Let us start with estimates of solution to (7.2.7). First of all, we know
the formula for the pressure:
Z
(k) 1 (k) 2 1
p (x, t) = − |u (x, t)| + K(x − y) : u(k) (y, t) ⊗ u(k) (y, t)dy, (7.2.8)
3 4π
R3
where
Z
1 1
px1(k)
0
(x, t) = − |u(k) (x, t)|2 + K(x − y) : u(k) (y, t) ⊗ u(k) (y, t)dy,
3 4π
B(x0 ,2)
Z
1
p2(k)
x0 (x, t) = (K(x − y) − K(x0 − y)) : u(k) (y, t) ⊗ u(k) (y, t)dy,
4π
R3 \B(x0 ,2)
Z
1
c(k)
x0 (t) = K(x0 − y) : u(k) (y, t) ⊗ u(k) (y, t)dy.
4π
R3 \B(x0 ,2)
7.2. ESTIMATES OF SCALED SOLUTIONS 215
Using the similar arguments as in [35], see also Appendix B, one can
1(k) 2(k)
derive estimates of px0 and px0 . Here, they are:
kp1(k)
x0 (·, t)kL 3 (B(x0 ,3/2)) ≤ c(M )(kv
(k)
(·, t)k2L3 (B(x0 ,2)) + 1), (7.2.10)
2
sup |px2(k)
0
(x, t)| ≤ c(M )(kv (k) (·, t)k2L2,unif + 1), (7.2.11)
B(x0 ,3/2)
where
kgkL2,unif = sup kgkL2 (B(x0 ,1)) .
x0 ∈R3
We further let
α(s) = α(s; k, S) = kv (k) (·, s)k22,unif ,
Zs Z
β(s) = β(s; k, S) = sup |∇v (k) |2 dydτ.
x∈R3
−S B(x,1)
h Z0 i
3
δ(0) ≤ c(M ) γ(0) + (1 + α 2 (s))ds , (7.2.12)
−S
with some positive constant c(M ) independent of k and S. Here, γ and δ are
defined as
Zs Z
γ(s) = γ(s; k, S) = sup |v (k) (y, τ )|3 dydτ
x∈R3
−S B(x,1)
and
Zs Z
3
δ(s) = δ(s; k, S) = sup |p(k) (y, τ ) − cx(k) (τ )| 2 dy dτ,
x∈R3
−S B(x,3/2)
Zs Z h i
= |v (k) |2 ∆ϕ2x0 + v (k) · ∇ϕ2x0 (|v (k) |2 + 2px(k)
0
) dxdτ +
−S R3
Zs Z h
+ w(k) · ∇ϕ2x0 |v (k) |2 + 2ϕ2x0 w(k) ⊗ (w(k) + v (k) ) : ∇v (k) +
−S R3
i
+2w(k) · v (k) (w(k) + v (k) ) · ∇ϕ2x0 dxdτ = I1 + I2 .
The first term I1 is estimated with the help of the Hölder inequality,
multiplicative inequality (7.2.13), and bounds (7.2.10), (7.2.11). So, we find
h Zs 3
I1 ≤ c(M ) (1 + α(τ ) + α 2 (τ ))dτ +
−S
Zs 14 Zs 34 i
3
+ α (τ )dτ β(s) + α(τ )dτ .
−S −S
Zs Z 51 Z 54
5 5
(k) 5
+c |w | dx |v (k) | 4 |∇v (k) | 4 dx dτ +
−S B(x0 ,3/2) B(x0 ,3/2)
7.2. ESTIMATES OF SCALED SOLUTIONS 217
Zs
1
Z 12
+cβ (s) 2 |w(k) |4 dxdτ dτ +
−S B(x0 ,3/2)
Zs
+c kv (k) (·, τ )kL3 (B(x0 ,3/2)) kw(k) (·, τ )k2L3 (B(x0 ,3/2)) dτ.
−S
Taking into account (7.2.6) and applying Hölder inequality several times, we
find
2 1
I2 ≤ c(M )γ 3 (s)(s + S) 3 +
Zs Z 15 Z 12
(k) 5 (k) 2
+c |w | dx |∇v | dx ×
−S B(x0 ,3/2) B(x0 ,3/2)
Z
10
103 1 1
(k)
× |v | dx
3 dτ + c(M )β 2 (s)(s + S) 10 +
B(x0 ,3/2)
1 2
+c(M )γ 3 (s)(s + S) 3 .
It remains to use another known multiplicative inequality
Z 10
103 Z 51
(k) (k) 2
|v (x, s)| dx
3 ≤c |v (x, s)| dx ×
B(x0 ,3/2) B(x0 ,3/2)
Z 103
× (|∇v (k) (x, s)|2 + |v (k) (x, s)|2 dx
B(x0 ,3/2)
Zs 54 Zs 51
(k) 5
+c β(s) + α(τ )dτ ) × α(τ )kw (·, τ )kL5,unif dτ .
−S −S
Finally, we find h 1
α(s) + β(s) ≤ c(M ) (s + S) 5 +
218 CHAPTER 7. BEHAVIOUR OF L3 -NORM
Zs i
+ α(τ )(1 + kw(k) (·, τ )k5L5,unif ) + α3 (τ ) dτ , (7.2.14)
−S
which is valid for any s ∈ [−S, 0[ and for some positive constant c(M ) inde-
pendent of k, s, and S.
It is not so difficult to show that there is a positive constant S(M ) such
that
1
α(s) ≤ (7.2.15)
10
for any s ∈] − S(M ), 0[. In turn, the latter will also imply that
1
α(s) ≤ c(M )(s + S) 5 (7.2.16)
α(s) ≤ 1 (7.2.17)
−S τ
Z0 Z
+ sup |∇w(k) (y, s)|2 dyds ≤ c(M ) < ∞.
x0 ∈R3
−S B(x0 ,1)
Obviously, w(k) and all its derivatives converge to w and to its correspond-
ing derivatives uniformly in sets of the form B(R) × [δ, 0] for any R > 0 and
220 CHAPTER 7. BEHAVIOUR OF L3 -NORM
for any δ ∈] − S, 0[. The limit function satisfies the same representation
formula
1
Z |x − y|2
w(x, t) = 3 exp − w0 (y)dy,
(4π(s + S)) 2 4(s + S)
R3
in which w0 is the weak L3 (R3 )-limit of the sequence u(k) (·, −S). The function
w satisfies the uniform local energy estimate
Z0 Z
+ sup |∇w(y, s)|2 dyds ≤ c(M ) < ∞.
x0 ∈R3
−S B(x0 ,1)
Next, the uniform local energy estimate for the sequence u(k) (with respect
to k) can be deduced from the estimates above. This allows us to exploit
the limiting procedure explained in [25], see Appendix B, in details. As a
result, one can selected a subsequence, still denoted by u(k) , with the following
properties:
for any a > 0,
u(k) → u (7.3.2)
weakly-star in L∞ (−S, 0; L2 (B(a))) and strongly in L3 (B(a)×] − S, 0[) and
in C([τ, 0]; L 9 (B(a))) for any −S < τ < 0;
8
∇u(k) → ∇u (7.3.3)
pe(k)
n ≡p (k)
− c(k)
n * p (7.3.5)
in L 3 (−S, 0; L 3 (B(n))).
2 2
So, arguing in the same way as in [25], see Appendix B, one can show
that u and p satisfy the following conditions:
Z0 Z
sup sup ku(·, s)k2L2 (B(x0 ,1)) + sup |∇u(y, s)|2 dyds < ∞; (7.3.6)
−S<s<0 x0 ∈R3 x0 ∈R3
−S B(x0 ,1)
the function Z
s 7→ u(y, s) · w(y)dy (7.3.8)
R3
is continuous on [−S, 0] for any compactly supported w ∈ L2 (R3 );
∂t u + u · ∇u − ∆u = −∇p, div u = 0 (7.3.9)
in R3 ×] − S, 0[ in the sense of distributions;
for any x0 ∈ R3 , there exists a function cx0 ∈ L 3 (−S, 0) such that
2
2
(|u | + |ep | )dy ds = 2
(|v|3 + |q − b(k) | 2 )dx dt
a (aλk )
Q(a) Q(zT ,aλk )
p p
for all 0 < a < a∗ = inf{1, S/10, T /10} and for all λk ≤ 1. Here,
(k) (k)
zT = (0, T ), pe(k) ≡ pe2 , and b(k) (t) = λ−2k c2 (s). Since the pair v and q − b
(k)
2
(|u(k) |3 + |e
p(k) | 2 )dy ds > ε (7.3.13)
a
Q(a)
−a2∗
for all < s < 0.
For any 0 < a < a∗ /2,
Z
1 3
ε≤ 2 p(k) | 2 + |u(k) |3 )dy ds ≤
(|e
a
Q(a)
Z
1 (k) 3 (k) 3
≤c (|p1 | 2 + |p2 | 2 + |u(k) |3 )dy ds ≤
a2
Q(a)
Z
1 (k) 3
≤c 2 (|p1 | 2 + |u(k) |3 )dy ds+
a
Q(a)
Z0
1 3 (k) 3
+ca 2 sup |p2 (y, s)| 2 ds.
a y∈B(a∗ /2)
−a2
for all 0 < a < a∗ /2. After passing to the limit and picking up sufficiently
small a, we find Z
2
0 < cεa ≤ |u|3 dy ds (7.3.18)
Q(a∗ )
for some positive 0 < a < a∗ /2. So, the limit function u is non-trivial.
Proof Theorem 1.1 The limit function w0 ∈ L3 and, hence,
as |x0 | → ∞. The latter, together with Theorem 1.4 from Appendix B, and
ε-regularity theory for the Navier-Stokes equations, gives a required decay at
infinity. To be more precise, there are positive numbers R, T ∈]a∗ , S[, and
ck with k = 0, 1, ... such that
(7.3.10) in order to get better estimates for the pressure, say, in the domain
(R3 \ B(R))×] − T, 0[. Indeed, using estimates of type (7.2.10) and (7.2.11)
for the parts of the pressure p1x0 and p2x0 in (7.3.10), we show
In a view of (7.3.19), the regularity theory for the stationary Stokes system
gives the estimates
|∇k w(x,
e t)| + |∇k re(x, t)| + |∇k ∂t w(x,
e t)| < c2k (7.3.22)
being valid for all x ∈ B(4R), for all t ∈] − T, 0[, and for all k = 0, 1, ....
Letting U = w − w e and P = r − re, where w = ϕu and r = ϕp, we find
∂t U + div(U ⊗ U ) − ∆U + ∇P = F = −div(U ⊗ w e ⊗ U ) + G,
e+w
div U = 0
226 CHAPTER 7. BEHAVIOUR OF L3 -NORM
in Q∗ = B(4R)×] − T, 0[,
U |∂B(4R)×]−T,0[ = 0.
Here, G = −div(w
e ⊗ w)
e + g − ∂t w
e and
g = (ϕ2 − ϕ)div(u ⊗ u) + uu · ∇ϕ2 + p∇ϕ − 2∇ϕ · ∇u − u∆ϕ.
Since u and p are a local energy solution, it follows from its definition that
there exists a set Σ ⊂] − T, 0[ of full measure, i.e., |Σ| = T , such that U is a
weak Leray-Hopf solution to initial boundary problem for the above system in
B(4R)×]t0 , 0[ for each t0 ∈ Σ. The rest of the proof is based upon estimates
(7.3.21) and (7.3.22) and unique continuation across spatial boundaries for
parabolic differential inequalities and goes along lines of arguments in the
last section of Chapter 6. Theorem 1.1 is proved.
Let us outline the proof of (7.1.4), which is much easier than the proof
of Theorem 1.1. Indeed, arguing as in the main case, we find a sequence
tk → T − 0 such that
m−3
lim kv(·, tk )km (T − tk ) 2m = 0.
k→∞
For solutions u(k) , we may use local energy estimates proved in Appendix B.
In particular, the y give the estimate
ku(k) (·, t)k22,unif ≤ 2cku(k) (·, −S)k22,unif
Z0 Z
sup ku(k) (·, t)k22,unif + sup x0 ∈ R3 |∇u(k) |2 dxdt → 0
−S<t<0
−S B(x0 ,1)
7.4. COMMENTS 227
7.4 Comments
This section is essentially based on my paper [57], which in turn summarizes
all previous attempts made in [54]-[56] to solve the problem about behaviour
of L3 -norm of the velocity field as time approaches possible blow up time.
228 CHAPTER 7. BEHAVIOUR OF L3 -NORM
Appendix A
Proposition 1.1. For any function u ∈ C0∞ (Rn ×]0, 2[; Rm ) and for any
positive number a, the inequality
|x|2
h−2a (t)e− a
R
4t
t
|u|2 + |∇u|2 dxdt
Rn ×]0,2[
(A.1.1)
|x|2
− 4t
h−2a (t)e
R
≤ c0 |∂t u + ∆u|2 dxdt.
Rn ×]0,2[
1−t
is valid with an absolute positive constant c0 and a function h(t) = te 3 .
229
230APPENDIX A. BACKWARD UNIQUENESS AND UNIQUE CONTINUATION
tL = S + A, (A.1.2)
where
1
Sv := t(∆v + (|∇φ|2 − ∂t φ)v) − v (A.1.3)
2
and
1
Av := (∂t (tv) + t∂t v) − t(div(v ⊗ ∇φ) + ∇v∇φ). (A.1.4)
2
Obviously,
R 2 2φ R
t e |∂t u + ∆u|2 dxdt = t2 |Lv|2 dxdt
R R R (A.1.5)
= |Sv|2 dxdt + |Av|2 dxdt + [S, A]v · v dxdt,
In our case,
h0 (t)
|∇φ|2 − ∂t φ = −|∇φ|2 + (a + 1) .
h(t)
The latter relation (together with (A.1.7)) implies the bound
R
t2 (|∇v|2 + |v|2 |∇φ|2 ) dxdt
R R (A.1.10)
≤ 3I − t2 v · Lv dxdt ≤ b1 t2 |Lv|2 dxdt
|u|2
Z |x|2
h−2a (t)(th−1 (t))2 (a + 1) + |∇u|2 e− 4t dxdt
t
Z
|x|2
≤ b2 h−2a (t)(th−1 (t))2 |∂t u + ∆u|2 e− 4t dxdt.
Proof Let u ∈ C0∞ (Q1+ ; Rm ), where Q1+ = (Rn+ +en )×]0, 1[. We are going
to use formulae (A.1.2)–(A.1.6) for new functions u, v, and φ. All integrals
in those formulae are taken now over Q1+ .
First, we observe that
∇φ = ∇φ(1) + ∇φ(2)
(A.1.13)
0
(1)
∇φ (x, t) = − x4t , (2)
∇φ (x, t) = 2αa 1−t x2α−1 en .
tα n
Therefore,
∇φ(1) · ∇φ(2) = 0, |∇φ|2 = |∇φ(1) |2 + |∇φ(2) |2 . (A.1.14)
Moreover,
∇2 φ = ∇2 φ(1) + ∇2 φ(2) ,
δij
− 4t if 1 ≤ i, j ≤ n − 1
(1)
φ,ij = ,
0 if i = n or j = n (A.1.15)
0 if i 6= n or j 6= n
(2)
φ,ij = .
2α(2α − 1)a 1−t x2α−2 if i = n and j = n
tα n
R
+ t |v| ∂t2 φ(s) − 2∂t |∇φ(s) |2 − ∆2 φ(s)
2 2
− 1t |∇φ(s) |2 + 1
∂ φ(s)
t t
dxdt, s = 1, 2.
A.1. CARLEMAN-TYPE INEQUALITIES 233
and, therefore, Z
I= t|v,n |2 dxdt + I2 . (A.1.18)
Now, our aim is to estimate I2 from below. Since α ∈]1/2, 1[, we can drop
the fist integral in the expression for I2 . As a result, we have
Z
I2 ≥ t2 |v|2 (A1 + A2 + A3 ) dxdt, (A.1.19)
where
A1 = −∂t |∇φ(2) |2 ,
1
A2 = A1 − ∆2 φ(2) − |∇φ(2) |2 ,
t
1
A3 = ∂t2 φ(2) + ∂t φ(2) .
t
For A2 , we find
x2α
n
A3 ≥ a(2α − 1) . (A.1.21)
tα+2
On the other hand,
1 1−t
−∂t |∇φ(2) |2 − |∇φ(2) |2 ≥ (2α − 1) 2α+1 4α2 a2 x2(2α−1)
n ≥0
t t
234APPENDIX A. BACKWARD UNIQUENESS AND UNIQUE CONTINUATION
and thus
1
A1 ≥ |∇φ(2) |2 . (A.1.22)
t
Combining (A.1.20)–(A.1.22), we deduce from (A.1.5) the estimate
R 2
t |Lv|2 dxdt ≥ I
R x2α R
≥ a(2α − 1) n
tα
|v|2 dxdt + t|v|2 |∇φ(2) |2 dxdt (A.1.23)
R R
≥ a(2α − 1) |v|2 dxdt + t|v|2 |∇φ(2) |2 dxdt.
Using (A.1.8), we can find the following analog of (A.1.9)
t|∇v|2 dxdt = − 21 |v|2 dxdt − tv · Lv dxdt
R R R
R (A.1.24)
2 2
+ t|v| (|∇φ| − ∂t φ) dxdt.
Due to special structure of φ, we have
|∇φ|2 − ∂t φ = |∇φ(1) |2 − ∂t φ(1) + |∇φ(2) |2 − ∂t φ(2)
= −|∇φ(1) |2 + |∇φ(2) |2 − ∂t φ(2)
and, therefore, (A.1.24) can be reduced to the form
R
t|∇v|2 + t|v|2 (|∇φ(1) |2 + |∇φ(2) |2 ) dxdt
R
t |∇v|2 + |v|2 |∇φ|2 dxdt = − 21 |v|2 dxdt (A.1.25)
R
=
R R R
− tv · Lv dxdt + 2 t|v|2 |∇φ(2) |2 dxdt − t|v|2 ∂t φ(2) dxdt.
But
x2α
n
−t∂t φ(2) ≤ a
tα
and, by (A.1.11) and (A.1.25),
1
R 2φ R
2
te |∇u|2 ≤ − v · (tLv) dxdt
(A.1.26)
R 2 (2) 2
R x2α
+2 t|v| |∇φ | dxdt + a n
tα
|v|2 dxdt.
The classical Cauchy-Scwartz inequality, (A.1.23), and (A.1.26) yield re-
quired inequality (A.1.12).
A.2. UNIQUE CONTINUATION ACROSS SPATIAL BOUNDARIES 235
for all k = 0, 1, ..., for all (x, t) ∈ Q(R, T ), and for some positive constants
Ck . Here,
Condition (A.2.3) means that the origin is zero of infinite order for the func-
tion u.
Theorem 2.1. Assume that a function u obeys conditions (A.2.1)–(A.2.3).
Then, u(x, 0) = 0 for all x ∈ B(R).
Without loss of generality, we may assume that T ≤ 1. Theorem 2.1 is
an easy consequence of the following lemma.
Lemma 2.2. Suppose that all conditions of Theorem 2.1 hold. Then, there
exist a constant γ = γ(c1 ) ∈]0, 3/16[ and absolute constants β1 and β2 such
that
|x|2
|u(x, t)| ≤ c2 (c1 )A0 (R, T )e− 4t (A.2.4)
for all (x, t) ∈ Q(R, T ) satisfying the following restrictions:
Here, √
A0 ≡ max |u(x, t)| + T |∇u(x, t)|.
(x,t)∈Q( 4 R, 34 T )
3
236APPENDIX A. BACKWARD UNIQUENESS AND UNIQUE CONTINUATION
|y|2
h−2a (s)e−
R
≤ c6 A2 4s χ(y, s) dyds (A.2.10)
Q(%,2)
|y|2
+c6 ε12 h−2a (s)e−
R
4s |v|2 dyds.
Q(%,2ε)
|y|2
h−2a (s)e−
R
≤ c6 A2 4s χ(y, s) dyds (A.2.11)
Q(%,2)
R2 −
(%−1)2
≤ c06 A2 −2a
h (3/2) + ρ n−1
h −2a
(s)e 4s ds .
0
β%2
a= . (A.2.13)
2 ln h(3/2)
This is legal, since h(3/2) > 1. Hence, by (A.2.13), inequality (A.2.12) can
be reduced to the form
Z2 2
2 −βρ2 n−1 −β%2 −2a 2β%2 − %8s
D ≤ c7 A e 1+ρ e h (s)e ds .
0
We fix β ∈]0, 1/64[, say, β = 1/100. Then, the latter relation implies the
estimate
Z2 2
2 %
D≤ c07 (c1 , n)A2 e−β% 1+ h−2a
(s)e − 16s
ds . (A.2.14)
0
ln(3/2)
It is easy to check that β < 12
and therefore g 0 (s) ≥ 0 if s ∈]0, 2[, where
%2
− 16s
g(s) = h−2a (s)e and a and % satisfy condition (A.2.13). So, we have
2
D ≤ c8 (c1 , n)A2 e−β% , (A.2.15)
2 √
Observing that |y|2 ≤ 2µ2 |x|
λ2
+ 2 if y ∈ B(µ λx , 1) and letting µ = 2β, we
derive from (A.2.15) and (A.2.16) the following bound
Z
µ2 |x|2 |x|2
|v|2 dyds ≤ c08 A2 e(−2β+ 2 ) t = c08 A2 e−β t . (A.2.17)
Q
e
A.3. BACKWARD UNIQUENESS FOR HEAT OPERATOR IN HALF SPACE239
On the other hand, the regularity theory for linear parabolic equations give
us: Z
|v(µx/λ, 1/2)| ≤ c9 (c1 , n) |v|2 dyds.
2
(A.2.18)
Q
e
for all (x, t) ∈ Q+ and for some M > 0. Natural regularity assumptions,
under which (A.3.1)–(A.3.3) may be considered are, for example, as follows:
u and weak derivatives ∂t u, ∇u, and ∇2 u are square
(A.3.4)
integrable over bounded subdomains of Q+ .
We can formulate the main result result of this section.
Theorem 3.1. Using the above notation introduced, assume that u satisfies
conditions (A.3.1)–(A.3.4). Then u ≡ 0 in Q+ .
This extends the main result of [11] and [12], where an analogue of Theo-
rem 3.1 was proved for (Rn \ B(R))×]0, T [ instead of Q+ . Similarly to those
papers, the proof of Theorem 3.1 is based on two Carleman-type inequalities,
see (A.1.1) and (A.1.12).
We start with proofs of several lemmas. The first of them plays the crucial
role in our approach. It enables us to apply powerful technique of Carleman’s
inequalities.
Lemma 3.2. Suppose that conditions (A.3.1), (A.3.2), and (A.3.4) hold.
There exists an absolute positive constant A0 < 1/32 with the following prop-
erties. If
2
|u(x, t)| ≤ eA|x| (A.3.5)
for all (x, t) ∈ Q+ and for some A ∈ [0, A0 ], then there are constants β(A) >
0, γ(c1 ) ∈]0, 1/12[, and c2 (c1 , A) > 0 such that
0 2 x2
n
|u(x, t)| ≤ c2 e4A|x | e−β t (A.3.6)
where Z
|y|2
I1 = χ(y, s)h−2a (s)e− 4s (|v|2 + |∇v|2 ) dyds.
Qρ
I ≤ c5 (c1 )I1 .
242APPENDIX A. BACKWARD UNIQUENESS AND UNIQUE CONTINUATION
1 1
8Aλ2 − <− (A.3.12)
4s 8s
for s ∈]0, 2]. By (A.3.9) and (A.3.12), we have
2 R2 R |y|2
I1 ≤ c23 e8A|x| χ(y, s)h−2a (s)e− 8s dyds
0 B(ρ)
(A.3.13)
2
h R2 (ρ−1)2
i
≤ c6 e8A|x| h−2a (3/2) + h−2a (s)e− 8s ds .
0
Z Z1 Z Z1
2
D≡ |w| dyds = |v|2 dyds
B(1) 1 B(1) 1
2 2
Z
|y|2
≤ c7 h−2a (s)e− 4s (|w|2 + |∇w|2 ) dyds
Qρ
Z2
ρ2
h i
8A|x|2
≤ c8 (c1 )e h (3/2) + h−2a (s)e− 32s ds
−2a
Z2
ρ2
h i
8A|x|2 −2βρ2 −2a 2βρ2 −2a 2βρ2 − 32s
= c8 e h (3/2)e + h (s)e ds .
0
a = βρ2 / ln h(3/2).
h Z2 i
8A|x0 |2 −βρ2
D ≤ c8 e e 1 + g(s) ds ,
0
A.3. BACKWARD UNIQUENESS FOR HEAT OPERATOR IN HALF SPACE243
ρ2
where g(s) = h−2a (s)e− 64s . It is easy to check that g 0 (s) ≥ 0 for s ∈]0, 2[ if
1
β < 96 ln h(3/2). So, we have
0 2 2 0 2 βx2
n
D ≤ 2c8 e8A|x | e−βρ ≤ 2c8 e8A|x | e− 12t . (A.3.14)
Lemma 3.3. Suppose that the function u obeys conditions (A.3.1), (A.3.2),
(A.3.4), and (A.3.5). There exists a number γ1 (c1 , c? ) ∈]0, γ/2] such that
u(x, t) = 0 for all x ∈ Rn+ and for all t ∈]0, γ1 [.
v(y, s) = 0 (A.3.18)
for all y ∈ Rn+ and for all s ∈]0, 1/2[;
βλ2 y 2
n 2
yn
8Aλ2 |y 0 |2 − 2(λ2 s−γ1 ) 2 |y 0 |2
|∇v(y, s)| + |v(y, s)| ≤ c9 e e ≤ c9 e8Aλ e−β 2s (A.3.19)
√
for all 1/2 < s < 1 and for all y ∈ Rn+ + λ3 en . Since A < 1/32 and λ ≤ γ≤
√
1/ 12, (A.3.19) can be reduced to the form
2
|y 0 |2 yn
|∇v(y, s)| + |v(y, s)| ≤ c11 e 48 e−β 2s (A.3.20)
and
1 r > −1/2
ψ2 (r) = .
0 r < −3/4
We set (see Proposition 1.2 for the definition of φ(1) and φ(2) )
1 (2) yn2α
φB (yn , s) = φ (yn , s) − B = (1 − s) α − B,
a s
where α ∈]1/2, 1[ is fixed, B = a2 φ(2) ( λ3 + 2, 1/2), and
η(yn , s) = ψ1 (yn )ψ2 (φB (yn , s)/B), w(y, s) = η(yn , s)v(y, s).
Z
(1)
≤ c? s2 e2φ e2aφB |∂s w + ∆w|2 dyds.
Q1+
Arguing as in the proof of Lemma 3.2, we can select γ1 (c1 , c? ) so small that
Z
|y 0 |2
I≡ s2 e2aφB (|w|2 + |∇w|2 )e− 4s dyds
Q1+
Z
|y 0 |2
≤ c10 (c1 , c? ) χ(yn , s)(syn )2 e2aφB (|v|2 + |∇v|2 )e− 4s dyds,
3
(Rn
+ +( λ +1)en )×]1/2,1[
Passing to the limit as a → +∞, we see that v(y, s) = 0 if 1/2 ≤ s < 1 and
φB (yn , s) > 0. Using unique continuation across spatial boundaries, we show
that v(y, s) = 0 if y ∈ Rn+ and 0 < s < 1.
Now, Theorem 3.1 follows from Lemmas 3.2 and 3.3 with the help of more
or less standard arguments. We shall demonstrate them just for complete-
ness.
Lemma 3.4. Suppose that the function u meets all conditions of Lemma 3.3.
Then u ≡ 0 in Q+ .
n
Proof By Lemma 3.3, u(x, t) = 0 for x ∈ R √+ and for t ∈]0, γ1 [. By
(1)
scaling, we introduce the function u (y, s) = u( 1 − γ1 y, (1 − γ1 )s + γ1 ).
It easy to check that function u(1) is well-defined in Q+ and satisfies all
conditions of Lemma 3.3 with the same constants c1 and A. Therefore,
u(1) (y, s) = 0 for yn > 0 and for 0 < s < γ1 . The latter means that u(x, t) = 0
for xn > 0 and for 0 < t < γ2 = γ1 + (1 − γ1 )γ1 . Then, we introduce the
function
p
u(2) (y, s) = u( 1 − γ2 y, (1 − γ2 )s + γ2 ), (y, s) ∈ Q+ ,
and apply Lemma 3.3. After k steps we shall see that u(x, t) = 0 for xn > 0
and for 0 < t < γk+1 , where γk+1 = γk + (1 − γk )γ1 → 1.
A0
Proof of Theorem 3.1 Assume that A0 < M . Then λ2 ≡ 2M <
1 2
2
. Introducing function v(y, s) = u(λy, λ s), (y, s) ∈ Q + , we see that this
function satisfies all conditions of Lemma 3.4 with constants c1 and A = 21 A0 .
A0
Therefore, u(x, t) = 0 for xn > 0 and for 0 < t < 2M . Now, we repeat
A0
arguments of Lemma 3.4, replacing γ1 to 2M and A to M , and end up with
the proof of the theorem.
A.4 Comments
The whole chapter is essentially due to a part of the paper [13]
246APPENDIX A. BACKWARD UNIQUENESS AND UNIQUE CONTINUATION
Appendix B
B.1 Introduction
In this Chapter, we deal with the the Cauchy problem for the 3D Navier-
Stokes equations with initial data belonging to the special Morrey class
L2,unif . In contrast to the space L2 , this class does not exclude, for example,
interesting homogeneous functions of order minus one. The main goal is to
show the existence of weak solutions that satisfy the local energy inequality.
Since the global energy inequality, which is the crucial point in the defini-
tion of weak Leray-Hopf solutions, is not valid any more, we are forced to
work with local energy estimates only. However, local estimates involve the
pressure, while, in the global L2 -case, the pressure does not appear at all.
For the Cauchy problem, one would hope to use a nice solution formula for
the pressure in terms of singular integrals. This formula is well-defined for
weak Leray-Hopf solutions, but it should be modified somehow in order to be
useful for functions with very weak decay at the spatial infinity. The problem
of the existence of weak solutions for the initial data from L2,unif has been
essentially solved by P. G. Lemarie-Riesset [35] and our aim is to give our
interpretation of his interesting and important results.
Let us consider the classical Cauchy problem for the Navier-Stokes equa-
tions:
247
248 APPENDIX B. LEMARIE-RIESSET LOCAL ENERGY SOLUTIONS
It is supposed that
◦ ◦
a ∈E 2 , g ∈ G2 (0, T ). (B.1.3)
◦ ◦
Here, spaces E m and Gm (0, T ) with m ≥ 1 are defined as follows:
◦
3
E m = {u ∈ Em : div u = 0 in R },
◦
3
Gm (0, T ) = {u ∈ Gm (0, T ) : div u = 0 in QT = R ×]0, T [ },
Z
Em = {u ∈ Lm,unif : |u(x)|m dx → 0 as |x0 | → +∞},
B(x0 ,1)
ZT Z
Gm (0, T ) = {u ∈ Lm,unif (0, T ) : |u(x, t)|m dxdt → 0
0 B(x0 ,1)
as |x0 | → +∞},
Z 1/m
m
Lm,unif = {u ∈ Lm,loc : kukLm,unif = sup |u(x)| dx < +∞},
x0 ∈R3
B(x0 ,1)
◦
As it has been shown in [35] (see also references there), the space E m is in
fact the closure of the set
◦
∞ 3 ∞ 3 3
C 0 (R ) = {u ∈ C0 (R ) : div u = 0 in R }
B.1. INTRODUCTION 249
with respect to the norm of the space Lm,unif . For the readers’ convenience,
we give the proof of this fact in the last section of this chapter, see Lemma
6.1.
In monograph [35], P. G. Lemarie-Riesset proved that, for g = 0, problem
(B.1.1)–(B.1.3) has at least one weak solution v with the following properties
(see Definition 32.1 in [35]): for any T > 0,
ZT Z
v ∈ L∞ (0, T ; L2,unif ), sup |∇ v|2 dxdt < +∞,
x0 ∈R3
0 B(x0 ,1)
Z Zt Z Zt Z
2 2
ϕ|v(x, t)| dx + 2 ϕ|∇v| dxdt ≤ |v|2 (∂t ϕ + ∆ϕ)
R3 0 R3 0 R3
+v · ∇ϕ(|v|2 + 2p) + 2ϕg · v dxdt (B.1.8)
px0 (x, t) ≡ p(x, t) − cx0 (t) = p1x0 (x, t) + p2x0 (x, t), (B.1.9)
Z
1
p2x0 (x, t) = (K(x − y) − K(x0 − y)) : v(y, t) ⊗ v(y, t) dy
4π
R3 \B(x0 ,2)
Remark 1.2. It is easy to see that (B.1.4), (B.1.6)–(B.1.8) imply the fol-
lowing inequality:
Z Zt Z Z
2 2
ϕ(x)|v(x, t)| dx + 2 ϕ|∇ v| dxds ≤ ϕ(x)|v(x, t0 )|2 dx
R3 t0 R3 R3
Zt Z h i
+ |v|2 ∆ ϕ + ∇ ϕ · v |v|2 + 2p + 2ϕg · v dxds. (B.1.10)
t0 R3
It is valid for any t ∈ [0, T ], for a.a. t0 ∈ [0, T ], including t0 = 0, and for
any nonnegative function ϕ ∈ C0∞ (R3 ).
Remark 1.3. In turn, from (B.1.4), (B.1.6), and (B.1.10), it follows that if
v and p are a local energy solution on the set R3 ×]0, T [, then they are a local
energy solution on the set R3 ×]t0 , T [ for a.a. t0 ∈ [0, T ], including t0 = 0.
B.1. INTRODUCTION 251
We are going to prove the followings statements. The first of them shows
that our information about pressure is sufficient to prove decay for both
velocity v and p.
Theorem 1.4. Assume that conditions (B.1.3) hold. Let v and p be a local
energy solution to the Cauchy problem (B.1.1), (B.1.2). Then v and p satisfy
the following additional properties:
◦
v(·, t) ∈E 2 (B.1.11)
ZT Z
3
sup |p(x, t) − cx0 (t)| 2 dxdt < +∞,
x0 ∈R3
0 B(x0 ,3/2)
ZT Z
3
sup I{|x|>R} |p(x, t) − cx0 (t)| 2 dxdt → 0 (B.1.14)
x0 ∈R3
0 B(x0 ,3/2)
Theorem 1.5. Assume that conditions (B.1.3) hold. There exists at least
one local energy solution to the Cauchy problem (B.1.1), (B.1.2).
kgkL2,unif (0,T ) only, and two functions v and p, being a local energy solution
to the Cauchy problem:
∂t v(x, t) + div v(x, t) ⊗ v(x, t) − ∆ v(x, t) = g(x, t) − ∇ p(x, t),
(B.1.15)
div v(x, t) = 0
p2x0 (x, t) = px0 ,R (x, t) + p̄x0 ,R (x, t) (x, t) ∈ B(x0 , 3/2)×]0, T [, (B.2.1)
where
Z
1
p̄x0 ,R (x, t) = (K(x − y) − K(x0 − y)) : v(y, t) ⊗ v(y, t) dy.
4π
R3 \B(x0 ,2R)
Lemma 2.1. For any x0 ∈ R3 , for any t ∈]0, T [, and for any R ≥ 1, the
following estimate is valid:
c
sup |p̄x0 ,R (x, t)| ≤ kv(·, t)k2L2,unif . (B.2.2)
B(x0 ,3/2) R
∞ Z
X 1
=c |v(y, t)|2 dy
i=0
|x0 − y|4
B(x0 ,2i+2 R)\B(x0 ,2i+1 R)
B.2. PROOF OF THEOREM 1.4 253
∞ Z
X 1
≤c |v(y, t)|2 dy
i=0
(2i+1 R)4
B(x0 ,2i+2 R)
∞
X 1
≤c (2i+2 R)3 kv(·, t)k2L2,unif .
i=0
(2i+1 R)4
Lemma 2.1 is proved.
We let
Zt Z
α(t) = kv(·, t)k2L2,unif , β(t) = sup |∇ v|2 dxds,
x0 ∈R3
0 B(x0 ,1)
Zt Z
γ(t) = sup |v|3 dxds.
x0 ∈R3
0 B(x0 ,1)
Zt 14 Zt 34
3
γ(t) ≤ c α (s) ds β(t) + α(s) ds . (B.2.3)
0 0
Zt Z ZT Z
γR (t) = sup |χR v|3 dxds, GR = sup |χR g|2 dxds
x0 ∈R3 x0 ∈R3
0 B(x0 ,1) 0 B(x0 ,1)
254 APPENDIX B. LEMARIE-RIESSET LOCAL ENERGY SOLUTIONS
Zt Z
3
δR (t) = sup |χR px0 | 2 dxds.
x0 ∈R3
0 B(x0 ,3/2)
Zt 14 Zt
1
Zt 34
3
γR (t) ≤ c αR (s) ds βR (t) + αR (s)ds + 2 α(s)ds . (B.2.5)
R
0 0 0
Lemma 2.2. Assume that v and p are a local energy weak Leray-Hopf solu-
tion to the Cauchy problem (B.1.1)–(B.1.3) on the space-time cylinder QT .
Then we have the estimate
2 4 h
sup αR (t) + βR (T ) + γR (T ) + δR (T ) ≤ C(T, A) kχR ak2L2,unif
3 3
0<t<T
1 i
+GR + . (B.2.6)
R2/3
Proof. To simplify our notation, we let p̂ = px0 .
We fix x0 ∈ R3 and a smooth nonnegative function ϕ such that
and let ϕx0 (x) = ϕ(x − x0 ). For ψ = χ2R ϕx0 , we find from inequality (B.1.10):
Z Zt Z 5
X
2
L≡ ψ(x)|v(x, t)| dx + 2 ψ|∇ v|2 dxds = Ii , (B.2.7)
0 R3 i=1
R3
where
Z Zt Z
I1 = ψ|a|2 dx, I2 = |v|2 ∆ ψ dxds,
R3 0 R3
Zt Z Zt Z
2
I3 = ∇ ψ · v|v| dxds, I4 = 2 ∇ ψ · v p̂ dxds,
0 R3 0 R3
Zt Z
I5 = 2 ψg · v dxds.
0 R3
B.2. PROOF OF THEOREM 1.4 255
Obviously,
I1 ≤ ckχR ak2L2,unif , (B.2.8)
Zt
1
I2 ≤ c αR (s) ds + C(T, A) , (B.2.9)
R
0
Zt
I5 ≤ c αR (s) ds + GR . (B.2.10)
0
The term I3 is evaluated with the help of Hölder inequality in the following
way:
1/3
2/3 c 2/3
I3 ≤ cγ (t) γR (t) + γ (t) .
R
So, by (B.2.4),
2/3 1
I3 ≤ C(T, A) γR (t) + . (B.2.11)
R
Next, we let
I4 = I 0 + I 00 ,
where
Zt Z
00
I =4 χR ϕx0 ∇ χR · v p̂ dxds.
0 B(x0 ,3/2)
where
Zt Z
3
23
J= |χR p̂| 2 dxds .
0 B(x0 ,3/2)
Obviously, J ≤ J1 + J2 + J3 , where
Zt Z
3
23 Zt Z
3
23
J1 = |χR p1x0 | 2 dxds , J2 = |χR px0 ,ρ | 2 dxds ,
0 B(x0 ,3/2) 0 B(x0 ,3/2)
Zt Z
3
23
J3 = |χR p̄x0 ,ρ | 2 dxds ,
0 B(x0 ,3/2)
√
where ρ = R. We start with evaluation of J1 . Letting
χR p1x0 = q1 + q2 ,
where
1
q1 (x, t) = − χR (x)|v(x, t)|2
3
Z
1
+ K(x − y)(χR (x) − χR (y)) : v(y, t) ⊗ v(y, t) dy,
4π
B(x0 ,2)
Z
1
q2 (x, t) = K(x − y)χR (y) : v(y, t) ⊗ v(y, t) dy,
4π
B(x0 ,2)
we use the theory of singular integrals and find the estimate for q2 :
Zt Z Zt Z
3 3
|q2 | dxds ≤ c
2 |χR | 2 |v|3 dxds
0 B(x0 ,3/2) 0 B(x0 ,2)
1/2
≤ C(T, A)γR (t). (B.2.15)
B.2. PROOF OF THEOREM 1.4 257
Since
1
q1 (x, t) = − χR (x)|v(x, t)|2
3
Z
1
+ K(x − y)(χR (x) − χR (x0 )) : v(y, t) ⊗ v(y, t) dy
4π
B(x0 ,2)
Z
1
+ K(x − y)(χR (x0 ) − χR (y)) : v(y, t) ⊗ v(y, t) dy,
4π
B(x0 ,2)
Zt Z Zt Z
3 1/2 c
|q1 | dxds ≤ cγ 1/2 (t)γR (t) + 3/2
2 |v|3 dxds
R
0 B(x0 ,3/2) 0 B(x0 ,2)
Zt Z
3
+c |χR (x0 ) − χR (x))| 2 |v(x, s)|3 dxds
0 B(x0 ,2)
1
1/2
≤ C(T, A)
+ γR (t) .
R3/2
Combining the latter estimate with (B.2.15), we find
1
1/3
J1 ≤ C(T, A) + γR (t) . (B.2.16)
R
Next, we let
χR px0 ,ρ = q3 + q4 ,
where
Z
1
q3 (x, t) = (K(x − y) − K(x0 − y))(χR (x)
4π
B(x0 ,2ρ)\B(x0 ,2)
h 1 Zt 16 i
3
≤ C(T, A) √ + αR (s) ds . (B.2.19)
R
0
and
h1 1 Zt 16 i
1/3
I 0 ≤ C(T, A)γR (t) γR (t) + √ +
3 3
αR (s) ds
R
0
Zt 13 i
h
2/3 1 3
≤ C(T, A) γR (t) + + αR (s) ds . (B.2.22)
R
0
B.2. PROOF OF THEOREM 1.4 259
h 1 Zt 13 i
2/3 3
+C(T, A) γR (t) + + αR (s) ds
R
0
h 1 Zt i
3
αR (t) ≤ ckχR ak6L2,unif + cGR 3 + C(T, A) 3 + αR3
(s) ds .
R
0
So, the first estimate in (B.1.14) follows from (B.2.2) and from (B.2.6). Fi-
nally, the second estimate in (B.1.14) is one of the statements of Lemma
B.2.2, see (B.2.6). Theorem 1.4 is proved.
260 APPENDIX B. LEMARIE-RIESSET LOCAL ENERGY SOLUTIONS
∂t v ε + Fε (v ε ) · ∇ v ε − ∆ v ε = g ε − ∇ pε ,
(B.3.1)
div v ε = 0
where
Z
1
pεx0 ,r (t) ≡ K(x0 − x̄) : v ε (x̄, t) ⊗ Fε (v ε )(x̄, t) dx̄,
4π
R3 \B(x0 ,r)
1 ε
p1ε ε
x0 ,r (x, t) ≡ − v (x, t) · Fε (v )(x, t)+
3
Z
1
K(x − x̄) : v ε (x̄, t) ⊗ Fε (v ε )(x̄, t) dx̄,
4π
B(x0 ,r)
p2ε
x0 ,r,R (x, t) ≡
Z
1
K(x − x̄) − K(x0 − x̄) : v ε (x̄, t) ⊗ Fε (v ε )(x̄, t) dx̄,
4π
B(x0 ,2R)\B(x0 ,r)
p3ε
x0 ,R (x, t) ≡
Z
1
K(x − x̄) − K(x0 − x̄) : v ε (x̄, t) ⊗ Fε (v ε )(x̄, t)dx̄.
4π
R3 \B(x0 ,2R)
Lemma 3.1. For any x0 ∈ R3 and for any R ≥ 1, we have the following
estimate
cr ε
sup |p3ε
x0 ,R (x, t)| ≤ kv (·, t)kL2,unif kFε (v ε )(·, t)kL2,unif . (B.3.8)
x∈B(x0 ,r) R
Taking into account the standard estimates for singular integrals, Lemma
3.1, and inequality (B.3.9), we find:
kp1ε ε 2
x0 ,r (·, t)kL 3 (B(x0 ,r)) ≤ ckv (·, t)kL3 ((B(x0 ,2)) , (B.3.10)
2
sup |p2ε ε 2
x0 ,r,R (x, t)| ≤ C1 (r, R)kv (·, t)kL2,unif , (B.3.11)
x∈B(x0 ,3r/4)
r ε
sup |p3ε 2
x0 ,R (x, t)| ≤ c kv (·, t)kL2,unif . (B.3.12)
x∈B(x0 ,r) R
We let
Zt Z
ε
αε (t) = kv (·, t)k2L2,unif , βε (t) = sup |∇ v ε |2 dxds,
x0 ∈R3
0 B(x0 ,1)
Zt Z ZT Z
γε (t) = sup |v ε |3 dxds, G = sup |g(x, t)|2 dxdt.
x0 ∈R3 x0 ∈R3
0 B(x0 ,1) 0 B(x0 ,1)
Zt 41 Zt 34
γε (t) ≤ c αε3 (s) ds βε (t) + αε (s)ds . (B.3.13)
0 0
h Zt i
αε (t) + βε (t) ≤ c kak2L2,unif +G+ (αε (s) + αε3 (s)) ds . (B.3.14)
0
From the system (B.3.1) and (B.3.2), it is easy to derive the identity
Z Zt Z Z
E≡ ϕ2x0 (x)|v ε (x, t)|2 dx +2 ϕ2x0 |∇ v ε |2 dxds = ϕ2x0 |aε |2 dx
R3 0 R3 R3
Zt Z h
ε 2
+ |v | ∆ ϕ2x0 + ∇ ϕ2x0 ε ε 2 ε
· Fε (v ) |v | + 2p̂x0 ,2 (B.3.15)
0 R3
i
+2ϕ2x0 g ε · v ε dxds.
Taking into account (B.3.13), we derive from (B.3.16) and (B.3.17) the fol-
lowing estimate
h Zt Zt 3
αε (t) + βε (t) ≤ c kak2L2,unif +G+ αε (s) ds + αε2 (s) ds
0 0
Zt 14 Zt 34 i
3
+ αε (s) ds βε (t) + αε (s)ds .
0 0
where
Zt Z
3
δε (t) = sup |p̂εx0 ,2 | 2 dxds.
x0 ∈R3
0 B(x0 ,3/2)
Indeed, let
n ln 2 o
T0 := min T, . (B.3.19)
c(1 + (2c(kak2L2,unif + G))2 )
We claim that if 0 ≤ t < T0 , then α(t) < 2c(kak2L2,unif + G). Otherwise, there
should exist T 0 < T0 such that
αε (t) < 2c(kak2L2,unif ) + G
for 0 ≤ t < T 0 and
αε (T 0 ) = 2c(kak2L2,unif + G).
The main inequality implies the following estimate
Zt
αε (t) ≤ c kak2L2,unif + G + (1 + (2c(kak2L2,unif + G))2 ) αε (τ )dτ
0
Z 23 ZT0 Z 1 ε 2
ε 2 ε 2
c sup |v (x, t)| dx |∇v | + 2 |v | dxdt
0<t<T0 n
B(n) 0 B(n)
and thus
ZT0 Z
10 5
|v ε | 3 dxdt ≤ cn5 A 3 . (B.4.2)
0 B(n)
ZT0 Z
= v ε ⊗ Fε (v ε ) : ∇w − ∇v ε : ∇w + pεn div w + g ε · w dxdt
0 B(n)
ZT0 Z 12 ZT0 Z 12
+ |∇v ε |2 dxdt |∇w|2 dxdt
0 B(n) 0 B(n)
266 APPENDIX B. LEMARIE-RIESSET LOCAL ENERGY SOLUTIONS
ZT0 Z 3
23 ZT0 Z 13
ε 2 3
+ |pn | dxdt |∇w| dxdt
0 B(n) 0 B(n)
ZT0 Z 12 ZT0 Z 12
ε 2 2
+c |g | dxdt |w| dxdt
0 B(n) 0 B(n)
ZT0 Z ZT0 Z
ε 3
|Fε (v )| dxdt ≤ |v ε |3 dxdt,
0 B(n) 0 B(2n)
we have
ZT0 Z ZT0 Z 31
∂t v ε · w dxdt ≤ C(n, T0 , A, G) |∇w|3 dxdt
0 B(n) 0 B(n)
?
v 1,k * v 1 in L∞ (0, T0 ; L2 (B(1))),
v 1,k * v 1 in L2 (0, T0 ; W21 (B(1))),
v 1,k → v 1 in L3 (0, T0 ; L3 (B(1))),
F1,k (v ) → v 1
1,k
in L3 (0, T0 ; L3 (B(δ))), ∀δ < 1,
p1,k
1 * p1 in L 3 (0, T0 ; L 3 (B(1)))
2 2
and the local energy inequality in B(1)×]0, T0 [. The latter means that
Z Zt Z Zt Z
ϕ(x, t)|v(x, t)|2 dx + 2 ϕ|∇v|2 dxdt ≤ |v|2 (∂t ϕ + ∆ϕ)
B(1) 0 B(1) 0 B(1)
+v · ∇ϕ(|v|2 + 2p) + 2ϕg · v dxdt
?
v 2,k * v 2 in L∞ (0, T0 ; L2 (B(2))),
2,k 2
v *v in L2 (0, T0 ; W21 (B(2))),
v 2,k → v 2 in L3 (0, T0 ; L3 (B(2))),
F2,k (v ) → v 2
2,k
in L3 (0, T0 ; L3 (B(δ))), ∀δ < 2,
p2,k
2 * p2 in L 3 (0, T0 ; L 3 (B(2))).
2 2
The functions v 2 and p2 satisfy the Navier-Stokes equations and the local
energy inequality in the space-time cylinder B(2)×]0, T0 [. Obviously, that
v 2 = v on B×]0, T0 [. So, we may extend v by letting v = v 2 on B(2)×]0, T0 [.
As to the function p2 , it follows from the Navier-Stokes equations that ∇p2 =
∇p on B×]0, T0 [. This means that p2 (x, t) − h2 (t) = p(x, t) for x ∈ B and
for t ∈]0, T0 [. Since both p2 and p belong to L 3 (0, T0 ; L 3 (B)), we conclude
2 2
the h2 ∈ L 3 (0, T0 ). This allows to extend the function p to B(2)×]0, T0 [ so
2
that p = p2 − h2 on B(2)×]0, T0 [. Clearly, p ∈ L 3 (0, T0 ; L 3 (B(2))) and the
2 2
functions v and p satisfies the Navier-Stokes equations and the local energy
inequality on the space-time cylinder B(2)×]0, T0 [.
In the case n = 3, we repeat the above arguments choosing a subsequence
of the sequence v 2,k and replacing balls B and B(2) with balls B(2) and B(3),
respectively. Continuing this process, we arrive at the following result. There
exist two functions v and p defined on R3 ×]0, T0 [ such that
p{k} n,rk
n = pn .
{k} ∞
Obviously, pn is a subsequence of the sequence {pn,k
n }k=n . For these new
sequences and for any n ∈ N, we have
?
v {k} * v in L∞ (0, T0 ; L2 (B(n))),
v {k} * v in L2 (0, T0 ; W21 (B(n))),
v {k} → v in L3 (0, T0 ; L3 (B(n))), (B.4.6)
{k}
F{k} (v ) → v in L3 (0, T0 ; L3 (B(δ))), ∀δ < n,
p{k}
n * pn in L 3 (0, T0 ; L 3 (B(n)))
2 2
and
k∂t vkXn ≤ C(n, T0 , A, G), (B.4.7)
ZT0 Z
ess sup kv(·, t)k2L2,unif + sup |∇v|2 dxdt ≤ 2A. (B.4.9)
0<t<T0 x0 ∈R3
0 B(x0 ,1)
Next, we note that for the solution of the regularised problem we have
the following identity:
Z Zt Z Z
{k} {k} 2
ϕ(x)|v 2
(x, t)| dx + 2 ϕ|∇ v | dxds = ϕ|a{k} |2 dx
R3 0 R3 R3
Zt Z h
+ |v {k} |2 ∆ ϕ + ∇ ϕ · F{k} (v {k} ) |v {k} |2 + 2p{k}
n (B.4.11)
0 R3
i
{k} {k}
+2ϕg ·v dxds,
which is valid for any function ϕ ∈ C0∞ (R3 ). Taking into account (B.3.4),
(B.4.6)–(B.4.8), and (B.4.10), we deduce from (B.4.11) the inequality
Z Zt Z Z
2 2
ϕ(x)|v(x, t)| dx + 2 ϕ|∇ v| dxds ≤ ϕ|a|2 dx
R3 0 R3 R3
Zt Z h
2 2
+ |v| ∆ ϕ + ∇ ϕ · v |v| + 2p (B.4.12)
0 R3
i
+2ϕf · v dxds.
The latter holds for any t ∈ [0, T0 ] and for any nonnegative function ϕ ∈
C0∞ (R3 ). On the other hand, from (B.4.10) and from (B.4.12) it follows that
Z
ϕ|v(x, t) − a(x)|2 dx → 0 as t → +0 (B.4.13)
R3
for all ϕ ∈ C0∞ (R3 ). So, v meets (B.1.7). The validity of (B.1.8) follows from
(B.4.6). It remains to establish decomposition (B.1.9).
Thanks to (B.3.18), we have
ZT0 Z
3
|p̂εx0 ,2 | 2 dxdt ≤ A. (B.4.14)
0 B(x0 ,3/2)
We would like to emphasize that the constant on the right hand sides of
{k}
(B.4.14) is independent of ε and x0 . Let p̂x0 ,2 be the sequence generated by
270 APPENDIX B. LEMARIE-RIESSET LOCAL ENERGY SOLUTIONS
{k},x0
v {k} via (B.3.7). For each x0 ∈ R3 , we can find subsequences p̂x0 ,2 and
v {k},x0 such that
{k},x0
p̂x0 ,2 * px0 in L 3 (B(x0 , 3/2)×]0, T0 [).
2
Obviously,
2{k},x
px0 ,2,R0 → p2x0 ,R in L 3 (B(x0 , 3/2)×]0, T0 [),
2
where
Z
1
p2x0 ,R (x, t) = (K(x − y) − K(x0 − y)) : v(y, t) ⊗ v(y, t) dx.
4π
B(x0 ,2R)\B(x0 ,2)
It is supposed that
◦ ◦
a ∈ E 3, g ∈ G3 (0, T ), f ∈ G 5 (0, T ). (B.5.3)
2
Theorem 5.1. Assume that conditions (B.5.3) hold. There exists a unique
pair of functions v and p having the following properties:
p
v ∈ L∞ (0, T ; L3,unif ), (1 + |v|)|∇v| ∈ L2,unif (0, T ),
Z
1
p2x0 (x, t) = (K(x − y) − K(x0 − y)) : f (y, t) dy
4π
R3 \B(x0 ,2)
and
ZT Z
5
sup |p(x, t) − cx0 (t)| 2 dxdt < +∞,
x0 ∈R3
0 B(x0 ,3/2)
ZT Z
5
sup I{|x|>R} |p(x, t) − cx0 (t)| 2 dxdt → 0
x0 ∈R3
0 B(x0 ,3/2)
as R → +∞.
Using Theorem 5.1 and successive approximations, see, for example, [13]
and [17], we can prove the following theorems about solvability of the Cauchy
problem:
Theorem 5.2. Suppose that conditions (B.5.6) hold. There exists a number
T0 ∈]0, T [ with the following property. Given a and g, there exists a pair of
functions v and p, forming a local energy solutions in the space-time cylinder
QT0 = R3 ×]0, T0 [, such that
◦ ◦
v ∈ C([0, T0 ]; E 3 ) ∩ G5 (0, T0 ),
p
|v||∇v| ∈ L2,unif (0, T0 ), p ∈ L 5 (0, T0 ; L 5 ,loc (R3 ));
2 2
ZT0 Z
5
sup |p(x, t) − cx0 (t)| 2 dxdt < +∞,
x0 ∈R3
0 B(x0 ,3/2)
ZT0 Z
5
sup I{|x|>R} |p(x, t) − cx0 (t)| 2 dxdt → 0
x0 ∈R3
0 B(x0 ,3/2)
as R → +∞.
Moreover, assume that a pair u and q is a local energy solution to the
Cauchy problem
Then, u = v.
Theorem 5.3. Suppose that conditions (B.5.6) hold. Given T > 0, there
exists a constant ε(T ) with the following property. If
then there exists a pair of functions v and p, forming a local energy solutions
in the space-time cylinder QT , such that:
◦ ◦
v ∈ C([0, T ]; E 3 ) ∩ G5 (0, T ),
274 APPENDIX B. LEMARIE-RIESSET LOCAL ENERGY SOLUTIONS
p
|v||∇v| ∈ L2,unif (0, T ), p ∈ L 5 (0, T ; L 5 ,loc (R3 ));
2 2
ZT Z
5
sup I{|x|>R} |p(x, t) − cx0 (t)| 2 dxdt → 0
x0 ∈R3
0 B(x0 ,3/2)
as R → +∞.
Moreover, assume that a pair u and q is a local energy solution to the
Cauchy problem (B.5.4), (B.5.5). Then, u = v.
Now, let us outline the proof of Theorem 1.5.
So, according to Proposition 1.6, we can find a number T0 ∈]0, T ] and a
pair of functions v and p that are a local energy solution in the space-time
cylinder QT0 . If T0 = T , then we are done. Assume that it is not. By
Theorem 1.4, we can find t0 ∈]0, T0 [ so that
◦
v(·, t0 ) ∈E 3 ,
v and p are a local energy solution in R3 ×]t0 , T0 [.
Next, there exist T1 and a pair of functions u and q which is a local
◦
energy solution in R3 ×]t0 , T1 [ and u ∈ C([t0 , T1 ]; E 3 ) with u(·, t0 ) = v(·, t0 ).
However, we know that there must be v = u in R3 ×]t0 , T1 [. Without loss of
generality, we may assume that T1 < T . Using density of smooth functions,
let us decompose v(·, t0 ) = a1 + a2 and g = g1 + g2 so that
kg1 kL3,unif (0,T ) + ka1 kL3,unif ≤ ε(T − t0 ),
◦
a2 ∈ C ∞ 3
0 (R ),
and g2 is a function of class C ∞ in QT and there exists R2 > 0 such that the
support of g2 (·, t) lies in B(R2 ) for all t ∈]t0 , T [. According to Theorem 5.3,
there exists a pair u1 and q1 which is a local energy solutions to the Cauchy
problem:
∂t u1 (x, t) + div u1 (x, t) ⊗ u1 (x, t) − ∆ u1 (x, t) + ∇ q1 (x, t) = g1 (x, t),
div u1 (x, t) = 0
B.5. PROOF OF THEOREM 1.5 275
u1 (x, t0 ) = a1 (x), x ∈ R3 .
Moreover,
ku1 kL∞ (t0 ,T ;L3,unif ) ≤ cε(T − t0 ). (B.5.9)
We seek functions u2 and q2 , solving the following Cauchy problem:
∂t u2 (x, t) + div (u2 (x, t) ⊗ u2 (x, t) + u1 (x, t) ⊗ u2 (x, t)+
u2 (x, t0 ) = a2 (x), x ∈ R3 .
We state that this problem has a weak Leray-Hopf solution with the finite
global energy satisfying the local energy inequality. To see that it is really
possible, let us comment the crucial term in proving a priori global energy
estimate. This term has the form
Zt Z
I0 = (u2 ⊗ u1 + u1 ⊗ u2 ) : ∇u2 dxds.
t0 R3
Zt Z
≤ cku1 k2L∞ (t0 ,T ;L3,unif ) |∇u2 |2 + |u2 |2 dxds
t0 B(x0 ,1)
276 APPENDIX B. LEMARIE-RIESSET LOCAL ENERGY SOLUTIONS
Zt Z
2 2 2
≤ cε (T − t0 ) |∇u2 | + |u2 | dxds.
t0 B(x0 ,1)
and therefore
Zt Z
I0 ≤ cε(T − t0 ) |∇u2 |2 + |u2 |2 dxds.
t0 R3
This latter allows us to hide I0 into the left hand side of the global energy
inequality by choosing ε(T − t0 ) sufficiently small and to find
Z Zt Z Z Zt Z
|u2 (x, t)|2 dx + |∇u2 |2 dxds ≤ |a2 (x)|2 dx + 2 g2 · u2 dxds.
R3 t0 R3 R3 t0 R3
Using this estimate and suitable approximations, we can easily prove our
statements about u2 . In addition, all above arguments show that pressure q2
may be taken in the form
1
q2 = K ∗ (u2 ⊗ u2 + u1 ⊗ u2 + u2 ⊗ u1 )
4π
and, moreover,
q2 ∈ L 3 (R3 ×]t1 , T [)
2
1 1
q2x 0
= − (|u2 (x, t)|2 + 2u1 (x, t) · u2 (x, t))+
3
Z
1
+ K(x − y) : (u2 ⊗ u2 + u1 ⊗ u2 + u2 ⊗ u1 )(y, t)dy,
4π
B(x0 ,2)
B.6. DENSITY 277
Z
2 1
q2x = (K(x−y)−K(x0 −y)) : (u2 ⊗u2 +u1 ⊗u2 +u2 ⊗u1 )(y, t)dy,
0
4π
R3 \B(x0 ,2)
Z
1
c2x0 (t) = K(x0 − y) : (u2 ⊗ u2 + u1 ⊗ u2 + u2 ⊗ u1 )(y, t)dy.
4π
R3 \B(x0 ,2)
Now, we let
u = u1 + u2 , q = q1 + q2 .
Our task is to verify that this new pair forms a local energy solution to the
Cauchy problem:
div u(x, t) = 0
u(x, t0 ) = a(x), x ∈ R3 .
The most difficult part of this task is to show that u and q satisfy the local
energy inequality. It can be done essentially in the same way as the corre-
◦
sponding part of the proof of the uniqueness for C([0, T ]; E 3 )-solutions. And
this immediately implies that u = v in the R3 ×]t0 , T1 [. Since p(x, t)−q(x, t) =
c(t) ∈ L 3 (t0 , T1 ), we can change function cx0 in a suitable way and assume
2
that q = p in the R3 ×]t0 , T1 [. So, the pair u and q can be regarded as a
required extension of v and p to the whole space-time cylinder QT . Theorem
1.5 is proved.
B.6 Density
◦ ◦
Lemma 6.1. For any f ∈E m and for any ε > 0, there exists fε ∈ C ∞ 3
0 (R )
such that
kf − fε kLm,unif < ε. (B.6.1)
Proof Let
Bk = B(xk , 2), xk ∈ Z3 .
278 APPENDIX B. LEMARIE-RIESSET LOCAL ENERGY SOLUTIONS
than N2 balls Bk ,
For this covering, we can find a partition of unity such that
X
ϕk ∈ C0∞ (R3 ), spt ϕk ⊂ Bk , ϕk ≥ 0, ϕk = 1.
k
and observe that, for each R > 0, the function v R has a compact support
and, moreover,
div v R = f · ∇ χR in R3 ,
X cN2
kv R kB(x0 ,1) ≤ kv k kB(x0 ,1) ≤ kf kLm,unif , ∀x0 ∈ R3 .
k
R
Next, we let
uR = f χR − v R .
B.7. COMMENTS 279
Obviously, we have
div uR = 0 in R3 ,
and uR has a compact support. Since f ∈ E2 , we see that, for an arbitrary
ε > 0, we can find R > 0 such that
cN2
kf kLm,unif < ε.
≤ kf − f χR kLm,unif +
R
To complete the proof of the lemma, it is enough to smooth uR which is
easy. Lemma 6.1 is proved.
B.7 Comments
The main source for the content of Appendix B is the monograph of P.-
G. Lemarie-Riesset [35]. Our interpretation of his results is given in the
paper [25] and we follow it here. We wish to emphasize that the Lemarie-
Riesset conception of local energy weak Leray-Hopf solution1 is heavily used
in Chapter 7.
1
In fact, G.-P. Lemarie-Riesset himself calls them simply local Leray solutions
280 APPENDIX B. LEMARIE-RIESSET LOCAL ENERGY SOLUTIONS
Bibliography
[2] Caffarelli, L., Kohn, R.-V., Nirenberg, L., Partial regularity of suit-
able weak solutions of the Navier-Stokes equations, Comm. Pure Appl.
Math., Vol. XXXV (1982), pp. 771–831.
[3] Cannone, M., Harmonic analysis tools for solving the incompressible
Navier-Stokes equations, Handbook of Mathematical Fluid Dynamics,
vol. 3, 2002.
[4] Chen, C.-C., Strain, R. M., Tsai, T.-P., Yau, H.-T., Lower bound on
the blow-up rate of the axisymmetric Navier-Stokes equations, Int.
Math. Res. Not., (2008) Vol. 2008 : article ID rnn016, 31 pages,
doi:10.1093/imrn/rnn016.
[5] Chen, C.-C., Strain, R. M., Tsai, T.-P., Yau, H.-T., Lower bounds on
the blow-up rate of the axisymmetric Navier-Stokes equations II, Com-
munications in Partial Differential Equations, 34: 203–232, 2009.
[6] Chen Z.-M., Price W. G., Blow-up rate estimates for weak solutions of
the Navier–Stokes equations, R. Soc. Lond. Proc. Ser. A Math. Phys.
Eng. Sci. 457 (2001), 2625–2642.
[7] Choe, H. L., Lewis, J. L., On the singular set in the Navier-Stokes
equations, J. Functional Anal., 175(2000), pp. 348–369.
281
282 BIBLIOGRAPHY
[9] Escauriaza, L., Carleman inequalities and the heat operator,Duke Math-
ematical Journal, 104(2000), pp. 113 – 126.
[10] L. Escauriaza, c. E. Kenig, G. Ponce, L. Vega, Decay at infinity of caloric
functions within characteristic hyperplanes, Math. Res. Lett., 13 (2006),
441-453.
[11] Escauriaza,L., Seregin, G., Šverák, V., On backward uniqueness for
parabolic equations, Zap. Nauchn. Seminar. POMI, 288(2002), 100–103.
[12] Escauriaza,L., Seregin, G., Šverák, V., On backward uniqueness for
parabolic equations, Arch. Rational Mech. Anal., 169(2003)2, 147–157.
[13] Escauriaza, L., Seregin, G., Šverák, V., L3,∞ -solutions to the Navier-
Stokes equations and backward uniqueness, Russian Mathematical Sur-
veys, 58(2003)2, pp. 211-250.
[14] Escauriaza, L., Vega, L., Carleman inequalities and the heat operator
II, Indiana university Mathematics Journal, 50(2001), pp. 1149 – 1169.
[15] Fefferman, C., http://www.claymath.org/millennium/Navier-Stokes
equations.
[16] Galdi, G. P., An Introduction to the Mathematical Theory of the
Navier-Stokes Equations: Steady-State Problems, Springer Monographs
in Mathematics, Second edition, 2011, 1032p.
[17] Galdi, G.P., An introduction to the Navier-Stokes initial-boundary
value problem, Fundamental directions in mathematical fluid mechan-
ics/Giovanni P. Galdi...ed.,-Basel; Boston; Berlin: Birkhäuser 2000, pp.
1-70.
[18] Giga, Y., Solutions for semilinear parabolic equations in Lp and regu-
larity of weak solutions of the Navier-Stokes equations, J. of Diff. Equa-
tions, 62(1986), pp. 186–212.
[19] Gustafson, S., Kang, K., Tsai, T.-P., Interior regularity criteria for suit-
able weak solutions of the Navier-Stokes equations, Commun. Math.
Phys., 273(2007), 161–176.
[20] Hopf, E., Über die Anfangswertaufgabe für die hydrodynamischen
Grundgleichungen, Math. Nachrichten, 4 (1950-51), 213–231.
BIBLIOGRAPHY 283
[22] Kang, K., Unbounded normal derivative for the Stokes system near
boundary, Math. Ann. 331(2005), pp. 87–109.
[25] Kikuchi, N., Seregin, G., Weak solutions to the Cauchy problem for
the Navier-Stokes equations satisfying the local energy inequality. AMS
translations, Series 2, 220, 141-164.
[26] Kim H., Kozono H., Interior regularity criteria in weak spaces for the
Navier-Stokes equations, manuscripta math., 115 (2004), 85–100.
[27] Koch, G., Nadirashvili, N., Seregin, G., Sverak, V., Liouville theorems
for the Navier-Stokes equations and applications, Acta Mathematica,
203 (2009), 83–105.
[36] Leray, J., Sur le mouvement d’un liquide visqueux emplissant l’espace,
Acta Math. 63(1934), pp. 193–248.
[41] Rusin, W., Sverak, V., Miminimal initial data for potential Navier-
Stokes singularities, arXiv:0911.0500.
[43] Scheffer, V., Hausdorff measure and the Navier-Stokes equations, Com-
mun. Math. Phys., 55(1977), pp. 97–112.
[46] Seregin, G.A., Interior regularity for solutions to the modified Navier-
Stokes equations, J. of math. fluid mech., 1(1999), no.3, pp. 235–281.
[47] Seregin, G., Local regularity theory of the Navier-Stokes equations,
Handbook of Mathematical Fluid Mechanics, Vol. 4, Edited by Fried-
lander, D. Serre, pp, 159-200.
[48] Seregin, G. A., Estimates of suitable weak solutions to the Navier-Stokes
equations in critical Morrey spaces, J. Math. Sci. 143:2 (2007), 2961–
2968.
[49] Seregin, G., A note on local boundary regularity for the Stokes system,
Zapiski Nauchn. Seminar. POMI., v. 370(2009), pp. 151-159.
[50] Seregin, G., Local regularity for suitable weak solutions of the Navier-
Stokes equations, Russian Math. Surveys 62:3 595–614.
[51] Seregin, G. A. On the number of singular points of weak solutions to
the Navier-Stokes equations, Comm. Pure Appl. Math., 54(2001), issue
8, pp. 1019-1028.
[52] Seregin, G.A., Local regularity of suitable weak solutions to the Navier-
Stokes equations near the boundary, J. math. fluid mech., 4(2002),
no.1,1–29.
[53] Seregin, G.A., Differentiability properties of weak solutions to the
Navier-Stokes equations, Algebra and Analysis, 14(2002), No. 1, pp.
193-237.
[54] Seregin, G.A., Navier-Stokes equations: almost L3,∞ -cases, Journal of
mathematical fluid mechanics, 9(2007), pp. 34-43.
[55] Seregin, G., A note on necessary conditions for blow-up of energy solu-
tions to the Navier-Stokes equations, Progress in Nonlinear Differential
Equations and Their Applications, 2011 Springer Basel AG, Vol. 60,
631-645. 98.
[56] Seregin, G., Necessary conditions of potential blow up for the Navier-
Stokes equations, Zapiski Nauchn.Seminar. POMI, 385(2010), 187-199.
[57] Seregin, G., A certain necessary condition of potential blow up for
Navier-Stokes equations, Commun. Math. Phys. 312, 833-845.
286 BIBLIOGRAPHY
[58] Seregin, G., Selected topics of local regularity theory for the Navier-
Stokes equations. Topics in mathematical fluid mechanics, 239313, Lec-
ture Notes in Math., 2073, Springer, Heidelberg, 2013.
[59] Seregin, G., Šverák, V., On Type I singularities of the local axi-
symmetric solutions of the Navier-Stokes equations, Communications
in PDE’s, 34(2009), pp. 171-201.
[60] Seregin, G., Šverák, V., On a bounded shear flow in half space, Zapiski
Nauchn. Seminar POMI, 385(2010), pp. 200-205.
[63] Serrin, J., On the interior regularity of weak solutions of the Navier-
Stokes equations, Arch. Ration. Mech. Anal., 9(1962), pp. 187–195.
[65] Struwe M., On partial regularity results for the Navier-Stokes equations,
Comm. Pure Appl. Math., 41 (1988), pp. 437–458.
[66] Takahashi S., On interior regularity criteria for weak solutions of the
Navier-Stokes equations, Manuscripta Math., 69 (1990), 237-254.