Solutions To Partial Differential Equations by Lawrence Evans
Solutions To Partial Differential Equations by Lawrence Evans
Matthew Kehoe
Contents
Chapter 5 Solutions 3
5.10.4 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
5.10.6 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
5.10.7 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
5.10.8 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
5.10.9 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
5.10.11 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
5.10.13 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
5.10.15 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
5.10.17 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
5.10.18 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
Exercise (Hardy’s Inequality on R+ ) . . . . . . . . . . . . . . . . 16
1
Evans Chapters 5 - 9
Chapter 6 Solutions 17
6.6.3 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
6.6.5 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
6.6.7 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
6.6.8 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
6.6.10 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
6.6.13 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
6.6.15 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
Elasticity Exercise . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
Chapter 7 Solutions 33
7.5.1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
7.5.2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34
7.5.5 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
7.5.6 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
7.5.12 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38
7.5.13 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
7.5.14 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40
7.5.15 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
7.5.16 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44
Chapter 8 Solutions 47
8.7.1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47
8.7.2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48
8.7.4 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49
8.7.7 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51
8.7.15 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52
2
Evans Chapters 5 - 9
Chapter 9 Solutions 54
9.7.11 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54
9.7.12 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56
9.7.13 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62
Chapter 5 Solutions
Proof. The sufficiency part of the lemma follows directly from the fundamental
theorem of calculus. That is, if u is integrable on [a, b], and if U is defined by
Z x
U (x) := u(t) dt, a ≤ x ≤ b,
a
then U 0 (x) = u(x) for almost every x in [a, b]. To prove the necessity part, we
let U be an absolutely continuous function on [a, b]. Then U is differentiable
almost everywhere and U 0 is integrable on [a, b]. Let
Z x
G(x) := U (a) + U 0 (t) dt, x ∈ [a, b].
a
3
Evans Chapters 5 - 9
SN
5.10.6 Assume U is bounded and U ⊂⊂ i=1 Vi . Show there exist C ∞
functions ζi (i = 1, 2, . . . , N ) such that
(
0 ≤ ζi ≤ 1, spt ζi ⊂ Vi (i = 1, 2, . . . , N )
PN
i=1 ζi = 1 on U .
Proof. We first complete problem 5.10.5. Let U and V be open sets with
V ⊂⊂ U . We need to show that there exists a smooth function ζ such that
ζ ≡ 1 on V and ζ = 0 near ∂U .
4
Evans Chapters 5 - 9
where Wi denotes an open ball which covers a portion of W and possibly the
boundary. Then we may observe that we can use the mollifier ηε̃i for every
open ball Wi where
ψi (x) := ηε̃i ∗ χW (x).
By defining
N
X ψi (x)
ζ(x) := PN ,
i=1 i=1 ψi (x)
we observe that for any fixed x ∈ U , only three terms in the sum will be
nonzero. As V ⊂⊂ W ⊂⊂ U , it is clear that ζ ≡ 1 on V and ζ = 0 near ∂U .
5
Evans Chapters 5 - 9
SN SN
To complete 5.10.6, we assume U is bounded and U ⊂⊂ i=1 Wi ⊂⊂ i=1 Vi .
So, U has a finite cover {V1 , . . . , Vn } where for every Vi , we have a ψi as
constructed above. The support of ψi is contained entirely in Vi , ψi ≡ 1 on
Wi , and every ψi is smooth by definition. Therefore, we define
ψi (x)
ζi (x) := PN
i=1 ψi (x)
P
where i ζi ≡ 1 by construction and for all x ∈ U , the support of ζi is
contained in Vi . Also, ζi is smooth because the ψi are smooth. Thus, the
collection {ζi }N
i=1 fulfills all of the requirements and is a partition of unity
subordinate to the cover {V1 , . . . , Vn }.
5.10.7 Assume that U is bounded and there exists a smooth vector field α
such that α · ν ≥ 1 along ∂U , where ν as usual denotes the outward unit
normal. Assume 1 ≤ p < ∞.
Apply the Gauss-Green Theorem to ∂U |u|p α · ν dS, to derive a new proof of
R
Therefore, since
∇|u|p = p|u|p−1 (sgn u)Du,
we have for p = 1 Z Z
|u| dS ≤ C |u| + |Du| dx.
∂U U
On the other hand, if p > 1 then we apply Young’s inequality with
a = |Du|, b = |u|p−1 , q = p/(p − 1) to form
Z Z Z
p p−1
|∇|u| | dx ≤ C p|u| |Du| dx ≤ C |Du|p + (p − 1)|u|p dx.
U U U
6
Evans Chapters 5 - 9
The constants above are different at every inequality. We may now observe
that
Z Z
|u|p dS ≤ C |u|p + |Du|p + (p − 1)|u|p dx
∂U
ZU
≤C |u|p + |Du|p dx,
U
as required.
T : Lp (U ) → Lp (∂U )
Then
0 ≤ un (x) ≤ 1,
un (x) = 1, x ∈ ∂U.
Therefore, for every x ∈ U , we see that un (x) → 0 pointwise, so by the
dominated convergence theorem
kun k2L2 (U ) → 0.
which implies that the area of the boundary is equal to zero. As T un = 1 for
every n we see that we have arrived at a contradiction. The same analysis
works for Lp (U ) when 1 ≤ p < ∞.
7
Evans Chapters 5 - 9
which is equivalent to CkukL2 (U ) kD2 ukL2 (U ) . To show that the left-hand side
converges to kDuk2L2 (U ) , we evaluate the difference and once again apply
Cauchy–Schwarz
Z Z
(Dvk · Dwk − Du · Du) dx = Dvk · (Dwk − Du) + Du · (Dvk − Du) dx
U
ZU
≤ |Dvk | · |Dwk − Du| + |Du| · |Dvk − Du| dx
U
≤ kDvk kL2 (U ) kDwk − DukL2 (U ) + kDukL2 (U ) kDvk − DukL2 (U ) ,
8
Evans Chapters 5 - 9
Du = 0 a.e. in U .
uε = ηε ∗ u ∈ C ∞ (Uε ),
9
Evans Chapters 5 - 9
then we will see that the Lipschitz constant will be unbounded. Define r = 1
and θ = ±(π − ε) and consider two points in K where x1 (r, θ) = (1, π − ε) and
x2 (r, θ) = (1, ε − π). Then when ε → 0+
sin π−ε − sin ε−π
|u(x1 ) − u(x2 )| 2 2
sup ≥ lim+
x1 6=x2 |x1 − x2 | ε→0 |cos(π − ε) − sin(ε − π)|
2 cos (ε/2)
= lim
ε→0+ sin(π − ε)
→ ∞.
10
Evans Chapters 5 - 9
5.10.15 Fix α > 0 and let U = B 0 (0, 1). Show that there exists a constant C,
depending only on n and α, such that
Z Z
u2 dx ≤ C |Du|2 dx,
U U
provided
|{x ∈ U | u(x) = 0}| ≥ α, u ∈ H 1 (U ).
where Z Z
1 1
uE := u(x) dx = N
u(x) dx.
|E| E L (E) E
Define
un − (un )E
vn := , (2)
kun − (un )E kL2 (U )
then vn ∈ W 1,2 (U ) and
Z
1
kvn kL2 (U ) = 1, (vn )E = 0, |Dvn |2 dx < ,
U n
where the last statement follows from
1
kDvn k2L2 (U ) < kvn k2L2 (U ) .
n
11
Evans Chapters 5 - 9
(Hint: Use that any sequence that converges in Lp has a subsequence that
converges pointwise a.e.)
ku − uk kW 1,p (U ) → 0
as k → ∞. Now
12
Evans Chapters 5 - 9
We may choose a subsequence so that uk tends a.e. to u and hence F 0 (uk ) a.e.
to F 0 (u). Now
Z
0 0
F (uk )Du k − F (u)Du φ dx
Z
F 0 (uk ) − F 0 (u) Duk + F 0 (u) Duk − Du φ dx
=
Z
≤C kDuk k∞ |F 0 (uk ) − F (u)| + kF 0 (u)k∞ kuk − ukW 1,p (U ) |φ| dx
where
Z ∂u Z
0 n ∂u ∂un ∂u
(u )
F n · − dx ≤ M − dx → 0 as n → ∞.
∂xi ∂xi U ∂xi ∂xi
U
13
Evans Chapters 5 - 9
which is the desired form of vxi = F 0 (u)uxi . As both F and u are sufficiently
smooth, it follows that the right-hand side of (4) is in Lp (U ), which implies
that vxi ∈ Lp (U ). To see that v ∈ W 1,p (U ), we may observe
Z Z Z
p p p
|v| dx = |F (u) − F (0)| dx ≤ M |u|p < ∞.
U U U
14
Evans Chapters 5 - 9
√ z
, if z > 0,
(Fε )0 (z) = z 2 + ε2
0, if z ≤ 0,
which implies that k(Fε )0 kL∞ (R) ≤ 1 for every ε > 0. Also, F (z) =
limε→0 Fε (z), where (
z, if z ≥ 0,
F (z) =
0, if z < 0.
The conditions of the chain rule for W 1,p are satisfied and we have
Z Z
∂φ ∂u
Fε (u) dx = − (Fε )0 (u) φ dx,
U ∂xj U ∂x j
15
Evans Chapters 5 - 9
and
u+ = lim Fε (u) in U ,
ε→0
5.10.17 to find
Z Z
∂φ ∂φ
u+ dx = lim Fε (u) dx
U ∂xj U ε→0 ∂x j
Z
∂φ
= lim Fε (u) dx
ε→0 U ∂xj
Z
∂u
= − lim (Fε )0 (u) φ dx (Chain Rule)
ε→0 U ∂xj
Z
∂u
=− lim (Fε )0 (u) φ dx (DCT)
ε→0 ∂xj
ZU
∂u
=− φ dx.
u>0 ∂xj
Du,
a.e. on {u > 0},
D|u| = 0, a.e. on {u = 0},
−Du, a.e. on {u < 0}.
16
Evans Chapters 5 - 9
Observe that Z t Z 1
1
f (s) ds = f (ts) ds.
t 0 0
Therefore by applying Minkowski’s integral inequality to the left hand side of
(5), we see
Z ∞ Z 1 p !1/p Z 1 Z ∞ 1/p
p
f (ts) ds dt ≤ |f (ts)| dt ds
0 0 0 0
Z 1 Z ∞ 1/p
−1/p p
= s ds |f (t)| dt
0 0
Z ∞ 1/p
p p
= |f (s)| ds ,
p−1 0
as required.
Chapter 6 Solutions
provided Z Z
∆u∆v dx = f v dx
U U
for all v ∈ H02 (U ). Given f ∈ L2 (U ), prove that there exists a unique weak
solution of (∗).
kukL2 (U ) ≤ C1 k∇ukL2 (U ) ,
17
Evans Chapters 5 - 9
2
Therefore k∇ uk2L2 (U ) is an equivalent norm in H02 (U ) and we think of
18
Evans Chapters 5 - 9
Hence Z Z
∆u∆v dx = f v dx ∀v ∈ H02 (U ),
U U
19
Evans Chapters 5 - 9
= f v dx.
U
We need to prove that the bilinear form is continuous and coercive. First, it is
continuous by the Cauchy-Schwarz inequality and the Trace inequality
between H 1 (U ) and L2 (∂U )
Z Z
|B[u, v]| ≤ |Du · Dv| dx + |(T u)(T v)| dS
U ∂U
≤ kDukL2 (U ) kDvkL2 (U ) + kT ukL2 (∂U ) kT vkL2 (∂U )
≤ (C + 1)kukH 1 (U ) kvkH 1 (U ) .
20
Evans Chapters 5 - 9
2
However, since T un → T u in L (∂U ), we see from (10)
Z Z
2
T u dS = lim T u2n dS → 0.
∂U n→∞ ∂U
6.6.7 Let u ∈ H 1 (Rn ) have compact support and be a weak solution of the
semilinear PDE
−∆u + c(u) = f in Rn ,
where f ∈ L2 (Rn ) and c : R → R is smooth, with c(0) = 0 and c0 ≥ 0. Assume
also c(u) ∈ L2 (Rn ). Derive the estimate
(Hint: Mimic the proof of Theorem 1 in §6.3.1, but without the cutoff function
ζ.)
21
Evans Chapters 5 - 9
therefore
Z
|B2 | ≤ |c(u)| · |Dk−h (Dkh u)| dx
Rn
Z
C
≤ C2 ε |Dkh (Du)|2 dx + kc0 k2L∞ kuk2L2 .
R n ε
22
Evans Chapters 5 - 9
The above analysis holds for every k ∈ {1, . . . , n} and h > 0 small. Therefore
we apply Theorem 3 in §5.8.2 of Evans to conclude
Z
|D2 u|2 dx ≤ C̃ kf k2L2 + kuk2L2 .
Rn
6.6.8 Let
Pnu be a smooth solution of the uniformly elliptic equation
Lu = − i,j=1 aij (x)uxi uxj = 0 in U . Assume that the coefficients have
bounded derivatives. Set v := |Du|2 + λu2 and show that
Lv ≤ 0 in U
implies
n
X n
X
D(Lu) = − Daij uxi uxj − aij Duxi uxj = 0.
i,j=1 i,j=1
Therefore
n
X n
X
− Daij uxi uxj = aij Duxi uxj .
i,j=1 i,j=1
23
Evans Chapters 5 - 9
We then compute
n
X
Lv = − aij Du · Du + λu2
xi xj
i,j=1
Xn
= −2 aij Duxi xj · Du + Duxi · Duxj + λuuxi xj + λuxi uxj
i,j=1
n
X n
X n
X n
X
= −2 aij Duxi xj · Du + aij Duxi · Duxj + λu aij uxi xj + λ aij uxi uxj
i,j=1 i,j=1 i,j=1 i,j=1
n
X n
X n
X
= −2 aij Duxi xj · Du − 2 aij Duxi · Duxj − 0 − 2λ aij uxi uxj
i,j=1 i,j=1 i,j=1
n
X X n n
X
=2 Daij uxi uxj · Du − 2 aij Duxi · Duxj − 2λ aij uxi uxj
i,j=1 i,j=1 i,j=1
6.6.10 Assume U is connected. Use (a) energy methods and (b) the
maximum principle to show that the only smooth solutions of the Neumann
boundary-value problem
(
−∆u = 0 in U
∂u
∂ν = 0 on ∂U
24
Evans Chapters 5 - 9
= Dv · Du dx.
U
Letting u = v yields Z
|Du|2 dx = 0,
U
which implies that Du = 0 a.e. in U . As U is connected, we know from 5.10.11
that u ≡ C a.e. in U for some constant C.
For (b), assume by contradiction that u is not a constant. Then there must be
some x0 where u attains its maximum in Ū . If x0 ∈ U , then the strong
maximum principle implies that u is constant in contradiction to our
assumption. Therefore, the maximum can only be obtained at the boundary of
U and we conclude that u(x0 ) > u(x) for all x ∈ U . However, Hopf’s lemma
then implies that
∂u 0
(x ) > 0
∂ν
in contradiction to ∂u
∂ν = 0 on ∂U . So u must obtain its maximum inside U
and the strong maximum principle implies that u ≡ C.
Pn
6.6.13 (Courant minmax principle) Let L = − i,j=1 (aij uxi )xj , where
((aij )) is symmetric. Assume the operator L, with zero boundary conditions,
has eigenvalues 0 < λ1 < λ2 ≤ · · · . Show
λk = max
P min B[u, u] (k=1,2,. . . ).
S∈ k−1 u∈S ⊥
kukL2 =1
25
Evans Chapters 5 - 9
Therefore
min B[u, u] ≤ λk ,
u∈S ⊥
kukL2 =1
max
P min B[u, u] ≤ λk .
S∈ k−1 u∈S ⊥
kukL2 =1
min B[u, u] = λk ,
u∈S ⊥
kukL2 =1
νk = max
P min B[u, u] ≥ λk ,
S∈ k−1 u∈S ⊥
kukL2 =1
hence νk ≥ λk .
26
Evans Chapters 5 - 9
(
−∆w = λw in U (τ )
w=0 on ∂U (τ ),
normalized so that kwkL2 (U (τ )) = 1. Suppose that λ and w are smooth
functions of τ and x. Prove Hadamard’s variational formula
∂w 2
Z
λ̇ = − v · ν dS,
∂U (τ ) ∂ν
d
where · = dτ and v · ν is the normal velocity of ∂U (τ ). (Hint: Use the calculus
formula from §C.4.)
By §C.4 in Evans,
Z
d
λ̇(τ ) = |∇w|2 dx
dτ U (τ )
Z Z
|∇w|2 v · ν dS + |∇w|2 τ dx
=
∂U (τ ) U (τ )
Z Z
= |∇w|2 v · ν dS + 2∇w∇wτ dx
∂U (τ ) U (τ )
Z Z
= |∇w|2 v · ν dS + 2w(−∆wτ ) dx (Integrate by parts).
∂U (τ ) U (τ )
27
Evans Chapters 5 - 9
and because w = 0 on ∂U (τ )
Z
d d
0= kwkL2 (U (τ )) = |w|2 dx
dτ dτ U (τ )
Z Z
|w|2 v · ν dS + |w|2 τ dx
=
∂U (τ ) U (τ )
Z
|w|2 τ dx
=
U (τ )
Z
=2 wwτ dx.
U (τ )
Therefore
Z Z
2
λ̇(τ ) = |∇w| v · ν dS + 2w λ(τ )wτ + λ̇(τ )w dx
∂U (τ ) U (τ )
Z Z Z
2
= |∇w| v · ν dS + 2λ(τ ) wwτ dx + 2λ̇(τ ) w2 dx
∂U (τ ) U (τ ) U (τ )
Z
= |∇w|2 v · ν dS + 2λ̇(τ ).
∂U (τ )
2
The boundary condition w = 0 on ∂U (τ ) implies |∇w|2 is equivalent to ∂w
∂ν
because the gradient is perpendicular to the boundary. This yields
∂w 2
Z
−λ̇(τ ) = v · ν dS,
∂U (τ ) ∂ν
as required.
28
Evans Chapters 5 - 9
defined using the indices i, j, k, l ∈ {1, . . . , n} so that, in the end, the PDE in
(7) denotes the system of equations
n
X
− ∂j (aijkl ekl (u)) = fi , i = 1, . . . , n.
j,k,l=1
for some λ ∈ (0, ∞). These ensure the system is uniformly elliptic. Assume
throughout that aijkl ∈ L∞ (U ).
H01 (U ; Rn ) = Cc∞ (U ; Rn )
Therefore
Z Z Z
Ae(u) : ∇v dx − (Ae(u))ν · v dS = f v dx.
U ∂U U
29
Evans Chapters 5 - 9
Therefore we have Z Z
Ae(u) : e(v) dx = f v dx
U U
(b) Prove that weak solutions always exist for f ∈ L2 (U ; Rn ) and are unique.
Hint: establish Korn’s inequality, which says that
Z Z
C |e(u)|2 ≥ |Du|2 ∀ u ∈ H01 (U ; Rn )
U U
and F : H 1 (U ; Rn ) → R where
Z
F (v) = f v dx.
U
We will apply the Lax-Milgram theorem to show that weak solutions are unique
and exist for f ∈ L2 (U ; Rn ). We first apply Hooke’s law to write
where where µ > 0 and λ ≥ 0 are called the Lamé constants. Then since
I : e(v) = ∇ · v our bilinear form B[u, v] becomes
Z
B[u, v] = Ae(u) : e(v) dx = 2µ(e(u) : e(v)) + λ(∇ · u, ∇ · v).
U
30
Evans Chapters 5 - 9
We then need to show that the last term on the RHS is positive. Integrating by
parts and recognizing that v = 0 on ∂U gives
n Z n Z
∂ 2 vj
Z
X ∂vi ∂vj X ∂vj
dx = − vi dx + νj vi dS
i,j=1 U ∂xj ∂xi i,j=1 U
∂xi ∂xj ∂U ∂xi
n Z Z
X ∂vi ∂vj ∂vj
= dx − νi vi dS
i,j=1 U
∂x i ∂x j ∂U ∂x j
n
! n
X ∂vi X ∂vj
= dx
i=1
∂xi j=1
∂xj
Z
= (∇ · v)2 dx ≥ 0,
U
31
Evans Chapters 5 - 9
Therefore weak solutions for f ∈ L (U ; Rn ) exist and are unique by the Lax-
2
Milgram theorem.
(c) Prove the regularity theorem that if A and f are infinitely differentiable
throughout U , then so must be the unique weak solution u.
Proof. By Hooke’s law, we follow Ciarlet [3] and rewrite the elasticity tensor as
Ae(u) = λ(tr e(u))I + 2µe(u) (13)
where once again µ > 0 and λ ≥ 0 are called the Lamé constants. We then
sketch a proof of the following
Theorem 4. (Regularity of weak solutions to the linear elasticity problem) Let
U be a domain in Rn with a boundary ∂U of class C 2 . Let f ∈ Lp (U ) and
p ≥ 6/5. Then the weak solution u ∈ H01 (U ) of the linear elasticity problem is
in the space W 2,p (U ) and satisfies
−div (λ(tr e(u))I + 2µe(u)) = f in Lp (U ).
Furthermore, let m ≥ 1 be an integer. If the boundary ∂U is of class C m+2 and
if f ∈ W m,p , then the weak solution u ∈ H01 (U ) is in the space W m+2,p (U ).
(2) Following the results of Agmon, Douglis, Nirenberg, and Geymonat, the
uniform ellipticity condition gives the regularity result for m = 0 and p ≥ 6/5.
I refer to Ciarlet [3] as the details of this analysis are technical.
32
Evans Chapters 5 - 9
We will apply Green’s formula for Sobolev Spaces. Let ν = (νi ) be the outward
unit normal vector for ∂U . Then for i = 1, 2, . . . , n Green’s formula gives
Z Z Z
(∂i u)v dx = − u∂i v dx + uvνi dS.
U U ∂U
Applying the formula to the LHS of (14) with Dirichlet boundary conditions
yields
Z Z
{λ(tr e(u))I + 2µe(u)} : e(v) dx = − div {λ(tr e(u))I + 2µe(u)} · v dx,
U U
m,p
by which we then conclude that f ∈ W (U ).
f ∈ W m,p (U ) =⇒ u ∈ W m+2,p (U )
Chapter 7 Solutions
7.5.1 Prove that there is at most one smooth solution of this initial/boundary-
value problem for the heat equation with Neumann boundary conditions
ut − ∆u = f in UT
∂u
∂ν = 0 on ∂U × [0, T ]
u = g on U × {t = 0}.
33
Evans Chapters 5 - 9
34
Evans Chapters 5 - 9
|Du|2 dx
R
B[u, u] U
λ1 = min 2 = min 2 ,
u∈H01 (U ) kukL2 (U ) u∈H01 (U ) kukL2 (U )
u6=0 u6=0
d 1 2
where we may observe that dt 2 ku(·, t)kL2 (U ) = −B[u, u] by the weak formation.
This implies
d
ku(·, t)k2L2 (U ) ≤ −2λ1 kuk2L2 (U ) .
dt
We now apply Gronwall’s inequality to obtain the exponential decay estimate.
Letting η(t) = ku(·, t)k2L2 (U ) we see that
This forms
ku(·, t)k2L2 (U ) ≤ e−2λ1 t kgk2L2 (U ) ,
as desired.
7.5.5 Assume (
uk * u in L2 (0, T ; H01 (U ))
u0k * v in L2 (0, T ; H −1 (U )).
35
Evans Chapters 5 - 9
for every w ∈ H01 (U ). If we didn’t want to use the hint, then we could observe
by (15) and (16) that
36
Evans Chapters 5 - 9
*Z +
T Z T
0
u(t)φ (t) dt, w = huφ0 , wi dt
0 0
Z T
= hu, φ0 wi dt
0
Z T
= lim huk , φ0 wi dt
k→∞ 0
Z T
= lim huk φ0 , wi dt
k→∞ 0
*Z +
T
0
= lim uk (t)φ (t) dt, w
k→∞ 0
* Z +
T
= lim − u0k (t)φ(t) dt, w
k→∞ 0
Z T
= lim −hu0k φ, wi dt
k→∞ 0
Z T
= lim −hu0k , φwi dt
k→∞ 0
Z T
= −hv, φwi dt
0
*Z +
T
= −v(t)φ(t) dt, w .
0
Rb
(Hint: We have a
(v, uk (t)) dt ≤ Ckvk|b − a| for 0 ≤ a ≤ b ≤ T and v ∈ H.)
37
Evans Chapters 5 - 9
we let k → ∞ to deduce
Z b
(v, u(t)) dt ≤ Ckvk|b − a|.
a
Proof. The requirement for λ to be in the resolvent set is that the operator
λI − A maps D(A) bijectively onto X. As λ, µ ∈ p(A) we see that
λI − A : D(A) → X, µI − A : D(A) → X
are both bijective operators onto X and are invertible. This implies that the
operator times its inverse is the identity. By spectral theory,
38
Evans Chapters 5 - 9
which proves the first resolvent identity. The second identity follows from re-
peating the analysis and switching µ and λ. Then
Therefore
Rλ − Rµ
Rλ Rµ =
µ−λ
and we have that
Rµ − Rλ Rλ − Rµ
Rµ Rλ = = = Rλ Rµ .
λ−µ µ−λ
used in (16) of §7.4.1. (Hint: Approximate the integral by a Riemann sum and
recall that A is a closed operator.)
u ∈ D(A), v = Au.
39
Evans Chapters 5 - 9
k Z ∞
X ∗
lim e−λti AS(t∗i )u∆t = eλt AS(t)u dt.
k→∞ 0
i=1
as required.
Proof. For (a) we know from (9) in §2.3 of Evans that the fundamental solution
of the heat equation in Rn may be written as the convolution
Z
1 |x−y|2
− 4t
[S(t)g](x) = e g(y) dy
(4πt)n/2 Rn
40
Evans Chapters 5 - 9
[S(t)g](x) = Jt ∗ g(x).
The space
1. Jt > 0
2. Js ∗ Jt = Js+t
R
3. Rn Jt (x) dx = 1
The first property is clear from the definition of Jt . The second property follows
from
Z Z
1 |x−y|2 1 |x−z|2
− 4(s+t) − 4t
n/2
e g(y) dy = n/2
e ×
Rn (4π(s + t)) n (4πt)
Z R
1 |z−y|2
n/2
e− 4s g(y) dydz
Rn (4πs)
where
Z
1 |x−y|2 1 1 |x−z|2 |z−y|2
− 4(s+t)
e = e− 4t e− 4s dz.
(4π(s + t))n/2 (4πt)n/2 (4πs)n/2 Rn
41
Evans Chapters 5 - 9
analogously as before. These properties imply that the operators {S(t)}t≥0 form
a semigroup where Ss St = Ss+t . The first property implies that the semigroup
is positive and S(t) maps positive functions into positive functions. In fact, it
maps positive functions into strictly positive functions. To verify (6) in Evans we
show that S(t)g → g in X = Lp (Rn ) as t → 0+ if g ∈ Lp (Rn ). For g ∈ Lp (Rn )
we find that
Z Z p
kS(t)g − gkpLp =
Jt (y)g(x − y) dy − g(x) dx
Rn Rn
Z Z
h i p
= n t
J (y) g(x − y) − g(x) dy dx
Rn R
Z Z
h √ i p
= n J1 (v) g(x − tv) − g(x) dv dx
Rn R
Z Z √ p
≤ J1 (v)g(x − tv) − g(x) dv dx
n n
ZR R Z √ p
= J1 (v) g(x − tv) − g(x) dx dv
Rn Rn
42
Evans Chapters 5 - 9
Proof. We first recall the proof of (i) and (ii) in Theorem 1 of §7.4 of Evans.
Let u ∈ D(A). Then
S(s)S(t)u − S(t)u S(t)S(s)u − S(t)u
lim = lim +
s−>0+ s s−>0 s
S(s)u − u
= S(t) lim +
s−>0 s
= S(t)Au.
Thus S(t)u ∈ D(A) and AS(t)u = S(t)Au. We will now assume that u ∈ D(Ak )
and show that S(t)u ∈ D(Ak ). By the induction hypothesis, we have D(Ak ) :=
{u ∈ D(Ak−1 ) | Ak−1 u ∈ D(A)} where k ≥ 2. The base case for k = 1 is
handled through Theorem 1. Let n = k − 1 so that the difference quotient
becomes
S(s)S(t)An u − S(t)An u S(t)S(s)An u − S(t)An u
lim = lim
s−>0+ s s−>0+ s
S(s)An u − An u
= S(t) lim + .
s−>0 s
This shows that S(t)An u ∈ D(A). However, in order to prove that S(t)u ∈
D(Ak ), we need to show that An S(t)u ∈ D(A). Hence we apply part (ii) of
Theorem 1 to see that S(t)An u = An S(t)u. Thus S(t)u ∈ D(Ak ).
43
Evans Chapters 5 - 9
Proof. To get a representation formula for the solution, we apply the Fourier
transform. We denote û(ξ, t) as the Fourier transform of u with respect to the
space variable x. This gives
2
ût = −|ξ| û in ÛT
û = 0 on ∂ Û × [0, T ]
û = ĝ on Û × {t = 0},
2
whose solution is û(ξ, t) = ĝ(ξ)e−|ξ| t . Taking the inverse Fourier transform, we
get u = S(·)g, where the heat semigroup {S(t)}t≥0 is defined by
Z
1 |x−y|2
− 4t
(S(t)g)(x) = e g(y) dy, t > 0, x ∈ Rn .
(4πt)n/2 Rn
Here, (S(0)g)(x) = g(x) and the remainder of the semigroup properties for
(S(t))t≥0 follow from the analysis in 7.5.14. As before, S(t)g = Jt ∗ f where
Z
1 −|x|2 /4t
Jt (x) = e , Jt (x) dx = 1, t > 0,
(4πt)n/2 Rn
44
Evans Chapters 5 - 9
Since g ∈ D(∆k ), we know that u(·, t) = S(t)g ∈ D(∆k ) for all finite time
intervals. It remains that show that D(∆k ) ⊆ H 2k (U ) for all k ∈ N. This
follows from the observation that
for all non-negative v ∈ H01 (U ). If for some compactly contained ball BR (x0 ) ⊂⊂
U we have that
ess sup u = ess sup u,
BR (x0 ) U
Hint: Apply Moser’s weak Harnack inequality. We rely on the following two
lemmas.
Lemma 6. Suppose u ∈ H 1 (U ) is a weak subsolution of L, i.e., Lu ≥ 0 and
Z
hA(x)Du, Dvi ≤ 0,
U
where
λI ≤ A(x) ≤ ΛI a.e. in U
for λ, Λ ∈ (0, ∞). Then u is bounded above on any U0 ⊂⊂ U . Therefore, if u is
a weak solution of Lu = 0, it is bounded above and below on any U0 satisfying
these conditions.
Lemma 7. Let u be a positive supersolution in B4r0 (x0 ) ⊂ Rn . For 0 < p <
n
n−2 , and if n ≥ 3 then
Z !1/p
p C
− u dx ≤ 2 ess inf u,
n
B2r0 (x0 )
n−2 − p Br0 (x0 )
45
Evans Chapters 5 - 9
where C = C(n, Λλ ). In the case n = 2, the estimate is true provided 0 < p < ∞
and the constant C depends on p and Λ C
λ in place of ( n −p)2 .
n−2
it must also be true that there is another ball Br0 (y0 ) with B4r0 (y0 ) ⊂ U such
that
ess sup u = ess sup u. (19)
Br0 (y0 ) U
Moreover, by Lemma 6 we see that we can take ess sup u to be finite since
U
we know that ess sup u < ∞. Define M to be a positive number such that
BR (x0 )
M > ess sup u. Then M − u is a positive supersolution and we can apply our
U
second lemma due to Moser. As M − u is a positive supersolution, we can apply
Lemma 7 to it. Passing to the limit, we see that the inequality holds for
because of (19) and (20). As M is equal to the supremum of u over the domain,
we also know that u ≤ M . Therefore we have that
u=M (21)
a.e. on B2r0 (y0 ). Given that we have found that u is constant on a ball of
radius 2r0 , we now extend this result to the entire domain. Let y ∈ U be
arbitrary. Then there is a sequence of balls Bi := Bri (yi ) for i = 0, . . . , n such
that B4ri (yi ) ⊂ U and Bi−1 ∩ Bi 6= ∅ for i = 1, . . . , n. Furthermore, y ∈ Bn .
Since B0 ∩ B1 6= ∅ and we previously showed that u = M a.e. on B2r0 (y0 ), we
have
ess sup u = M.
B1
By the same reasoning as before, we find that u = M a.e. on the ball B2r1 (y1 ).
We then iterate this process over every ball to obtain
u=M
a.e. on B2rn (yn ). As y ∈ Brn (yn ), we see that u(y) = M . Furthermore, since
y ∈ U was arbitrary, we have that u ≡ M a.e. on all of U .
46
Evans Chapters 5 - 9
Chapter 8 Solutions
Proof. For (a), we apply the Riemann–Lebesgue lemma which states that the
Fourier transform of an L1 function vanishes at infinity. As sin(kx) = (eikx −
e−ikx )/(2i), we observe that for any f ∈ L2 (0, 1), f χ[0,1] ∈ L1 (R), and therefore
1
eikx − e−ikx
Z Z
lim sin(kx)f dx = lim f χ[0,1] dx = 0.
k→∞ 0 k→∞ R 2i
for every g ∈ L2 (0, 1). It suffices to prove this in g ∈ C0∞ (0, 1) because these
functions are dense in L2 (0, 1). Then
k−1 j+1
Z 1 XZ k
uk (x)g(x) dx = uk (x)g(x) dx
j
0 j=0 k
j+λ j+1
k−1 Z Z !
X k k
= ag(x) dx + bg(x) dx .
j j+λ
j=0 k k
As g(x) is continuous, we have that for every j ∈ {0, . . . , k − 1}, there exist µj
and νj in [j/k, (j + 1)/k] such that
Z j+λ
k λ
ag(x) dx = a g(µj ),
j
k
k
j+1
1−λ
Z k
bg(x) dx = b g(νj ),
j+λ
k
k
47
Evans Chapters 5 - 9
as a consequence of the mean value theorem and the continuity of g(x). There-
fore
Z 1 k−1 k−1
1X 1X
uk (x)g(x) dx = aλ g(µj ) + b(1 − λ) g(νj ).
0 k j=0 k j=0
R1
However, both of these are Riemann sums which converge to 0 g(x) dx. We
pass to the limit to deduce
Z 1 Z 1 Z 1
lim uk (x)g(x) dx = aλ g(x) dx + b(1 − λ) g(x) dx
k→∞ 0 0 0
Z 1
= aλ + b(1 − λ) g(x) dx
0
Then Z (j+1)/k
vk (x) dx = 0.
j/k
R
So by defining the characteristic function g = χ[a,b] , we see that gvk dx → 0
(because the rationals are dense). The same is true for a step function, as
these are a linear combination of characteristic functions of an interval. As step
functions are dense in L2 , we see that this would also imply that uk * u :=
λa + (1 − λ)b.
−∆u + Dφ · Du = f in U
R
is the Euler-Lagrange equation corresponding to the fuctional I[w] := U
L(Dw, w, x) dx.
(Hint: Look for a Lagrangian with an exponential term.)
Proof. We are given that L(p, z, x) = L(Du, u, x) and by the hint we consider
the exponential term e−φ(x) . Let
1 2
L(p, z, x) = e−φ(x) |p| − zf (x) .
2
48
Evans Chapters 5 - 9
Then
1
L(Du, u, x) = e−φ(x) |Du|2 − uf (x)
2
and the functional is
Z
1
I[u] = e−φ(x) |Du|2 − uf (x) dx.
U 2
Therefore we can minimize the function by evaluating the derivative at time τ
Z
0 d
I [u] = I[u] = e−φ(x) (Du · Dũ − ũf (x)) dx
dτ U
d
where ũ = dτ u. As in the derivation shown in §8.1.2, evaluating the derivative
at τ = 0 will minimize the function. We compute
Z
I 0 [u] = e−φ(x) (Du · Dũ − ũf (x)) dx
ZU
= e−φ(x) Du · Dũ − ũe−φ(x) f (x) dx
U
Z
= ũe−φ(x) Dφ · Du − ũe−φ(x) ∆u − ũe−φ(x) f (x) dx
U
Z
= ũe−φ(x) (Dφ · Du − ∆u − f (x)) dx
U
= 0.
Hence,
−∆u + Dφ · Du = f
as desired.
8.7.4 Assume η : Rn → R is C 1 .
(a) Show L(P, z, x) = η(z) det P (P ∈ Mn×m , z ∈ Rn ) is a null Lagrangian.
Proof. For (a), we apply two identities from the Lemma (Divergence-free rows)
on page 464 of Evans:
(det P )I = P T (cof P ) (22)
and
n
X
(det P )δij = pki (cof P )kj , (i, j = 1, . . . , n). (23)
k=1
49
Evans Chapters 5 - 9
and then compute the first term in the Euler-Lagrange equations. (22) and (23)
forms
n
X n X
X n
− Lpki (Du, u, x) =− (ηzj (u)) ujxi (cof Du)ki
xi
i=1 i=1 j=1
Xn
+ η(u) (cof Du)ki,xi
i=1
n X
X n
=− (ηzj (u)) ujxi (cof Du)ki
i=1 j=1
n
X
=− ηzj (u)δjk det(Du)
j=1
Pn Pn
because det(Du)δjk = i=1 pji (cof Du)ki and j=1 (cof Du)kj,xj = 0. The sec-
ond term in the Euler-Lagrange equations becomes
Lzk (Du, u, x) = ηzk (u) det(Du).
Therefore since determinants are null Lagrangians,
n
X
− Lpki (Du, u, x) + Lzk (Du, u, x)
xi
i=1
n
X
=− ηzj (u)δjk det(Du) + ηzk (u) det(Du)
j=1
= 0.
Hence, L(P, z, x) = η(z) det P is a null Lagrangian.
50
Evans Chapters 5 - 9
is a null Lagrangian.
Proof. We will match the notation on page 462 of Evans and represent P ∈
Mm×n by 1
p1 . . . p1n
P =
..
.
m m
p1 . . . pn m×n
Pn
where P = (pki ) for i rows, k columns, and tr(A) = i=1 aii . Expanding L(P )
gives
n n
!2 n
X X X
2 2 k i i
tr(P ) − tr(P ) = pi pk − pi = pki pik − pii pkk .
i,k=1 i=1 i,k=1
We then see that the RHS consists of a sum of subdeterminants of the matrix
i
pi pik
pki pkk
obtained from evaluating the larger matrix P ∈ Mm×n . We need to show that
these subdeterminants are null Lagrangian. To do this, we first observe that if
i = k then we are done. On the other hand, if i 6= k then it is easier to apply
the notation shown in lecture. Observing that p is represented by Du and that
we are taking the derivative with respect to subscript gives
∂i ui ∂k ui
.
∂i uk ∂k uk
The determinant of the above matrix is
∂i ui ∂k ui
det = ∂i ui ∂k uk − ∂i uk ∂k ui .
∂i uk ∂k uk
Therefore
W (F ) = Fii Fkk − Fik Fki
where
∂W ∂W
= Fkk , = −Fik ,
∂Fii ∂Fki
∂W ∂W
= −Fki , = Fii .
∂Fik ∂Fkk
The Euler-Lagrange equations are
∂ ∂W
(Du) = ∂i (∂k uk ) − ∂k (∂i uk ) = (∂ik − ∂ki )uk = 0
∂xj ∂F ij
51
Evans Chapters 5 - 9
and
∂ ∂W
(Du) = ∂k (∂i ui ) − ∂i (∂k ui ) = (∂ki − ∂ik )ui = 0.
∂xj ∂F kj
clear that all of the 2 × 2 subdeterminants are null Lagrangian
Therefore, it isP
n
and −L(P ) = i,k=1 ∂i ui ∂k uk − ∂i uk ∂k ui as a sum of subdeterminants is null
Lagrangian.
Alternatively we can consider
and then compute the Euler-Lagrange equations abstractly through the simpli-
fied notation on page 456 and 463. Defining
Dpi L(Du) = The gradient of L(Du) with respect to the index row i
we find that the Euler-Lagrange equations (17) on page 463 may be written as
the divergence
where f ∈ L2 (U ) and
(b) Prove Z Z
Du · D(w − u) dx ≥ f (w − u) dx
U U
for all w ∈ A.
52
Evans Chapters 5 - 9
For (a), the existence of a minimizer follows from choosing a minimizing se-
quence {uk }∞
k=1 ⊂ A with
|Du| ≤ 1 a.e.,
For (b), we fix an arbitrary w ∈ A. Since A is convex we know that for every
0 ≤ τ ≤ 1,
u + τ (w − u) = (1 − τ )u + τ w ∈ A.
Hence by defining
i(τ ) := I[u + τ (w − u)],
53
Evans Chapters 5 - 9
i0 (0) ≥ 0. (24)
Then if 0 < τ ≤ 1,
Chapter 9 Solutions
may be irreversible. (That is, find a Hilbert space H and a convex, proper, lower
semicontinuous function I : H → (−∞, ∞] such that the semigroup solution of
(∗) satisfies
S(t)u = S(t)û
for some t > 0 and u 6= û.)
I : R → (−∞, ∞]
be defined by (
|u| if |u| ≤ 1
I(u) :=
u2 if |u| > 1.
54
Evans Chapters 5 - 9
Figure 2: The piecewise defined lower semi-continuous function I(u). The slope
of I(u) is ±1 when |u| ≤ 1 and ±2 when |u| > 1.
By the theory of gradient flows in Evans, we know that the gradient flow
(
u0 ∈ −∂I[u] (t > 0)
u(0) = u0
55
Evans Chapters 5 - 9
Therefore we can choose two initial conditions u0 and û0 and then go forward
in time t. Eventually there will be a time t > 0 such that S(t)u0 = S(t)û0 = 0.
Thus, the flow is irreversible.
|Du| ≤ 1,
meaning that nowhere can the sandpile have slope greater than 1. As usual,
Dx u = Du. We propose the dynamics
where a = a(x, t) ≥ 0 describes the flow rate of sand rolling downhill, that is,
in the direction −Du. Suppose that
f − ut ∈ ∂I[u],
Proof. We follow the analysis done by Aronsson, Evans, and Wu in [1] and [5].
We first proceed informally and analyze the physical aspects of the model.
Consider the non-linear p-Laplacian
∆p u = div(|Du|p−2 Du)
56
Evans Chapters 5 - 9
for some constant C ∈ (0, ∞). It then follows that there exists a sequence
pk → ∞ such that
∗
upk → u
locally uniformly
Dupk → Du∗ boundedly, a.e.
* a weakly ∗ in L∞ .
p−2
|Dupk |
57
Evans Chapters 5 - 9
|Du| ≤ 1 (28)
everywhere. The constraint has the physical interpretation that the the slope
will not remain in equilibrium and that the sand will flow downhill when the
slope is one. Suppose f ≥ 0 is a source term which represents the rate at
which sand is added to the sandpile whose initial height is zero. This would
initially suggest that the model would be ut = f in any region where the con-
straint (28) is active. However, adding more sand in a particular location would
break the constraint and it would be perfectly reasonable to expect that these
newly added sand particles would instantaneously roll downhill. The particles
would continue rolling downhill until stopping at a rest state where adding new
particles maintains the constraint.
58
Evans Chapters 5 - 9
We start by sending p → ∞. Then for every T > 0 we have that the functions
{up }p≥n+1 are bounded in L∞ (Rn × [0, T ]) and {Dup , up,t }p≥n+1 are bounded
in L2 (Rn × [0, T ]). Further, the functions {up }p≥n+1 have a uniform bound
supported in Rn × [0, T ]. Therefore we can extract a sequence pi → ∞ and a
limit u so that for every T > 0
(
upi → u a.e. and in L2 (Rn × (0, T ))
(29)
Dupi * Du, upi,t * ut weakly in L2 (Rn × (0, T )).
where we will send p → ∞ to recover the original PDE. We make some explicit
assumptions regarding the source terms {fp }p≥n+1 . For this, we select m distinct
points {dk }m n
k=1 ⊂ R and m smooth, nonnegative functions of time {fk (t)}k=1 .
m
for every v ∈ L2 (Rn ) at a.e. t > 0. We also suppose that v is Lipschitz and that
the Lipschitz constant is at most one. Then the expression
Z
f (x, t)(v − u(x, t)) dx
Rn
59
Evans Chapters 5 - 9
may defined as
m
X
fk (t)(v(dk ) − u(dk , t)).
k=1
Proof of Theorem 9: As |Du| ≤ 1, we have that I [u(·, t)] = 0 for a.e. t > 0. We
will also assume that v has compact support. To verify (30) is true, we need to
show that Z
(f (x, t) − ut (x, t))(v − u(x, t)) dx ≤ 0 (31)
Rn
for a.e. t > 0. To do this, we fix two times 0 < t1 < t2 . Our original PDE
becomes
fp − up,t = ∂Ip [up ].
This implies
Z t2 Z t2 Z
(t2 − t1 )Ip [v] ≥ I [up (·, t)] dt + (fp (x, t) − up,t (x, t))(v − up (x, t)) dxdt
t1 t1 Rn
Z t2 Z
≥ (fp (x, t) − up.t (x, t))(v − up (x, t)) dxdt, (32)
t1 Rn
lim Ip [v] = 0.
p→∞
kup (·, t)kC 0,1/2 ≤ C (kDup (·, t)kL2n + kup (·, t)kL∞ ) .
60
Evans Chapters 5 - 9
This gives
Z t2 Z t2
kup (x, t)kC 0,1/2 dt ≤ C + C kDup (x, t)kL2n dt
t1 t1
Z t2 Z 1/2n
2n
=C +C |Dup (x, t)| dxdt
t1 Rn
Z t2 Z 1/p
≤C +C |Dup (x, t)|p dxdt
t1 Rn
≤ C,
where all of the integrations are done over finite values. Therefore we find
Z t2 Z m Z t2
X Z
fpi (x, t)upi (x, t) dxdt = fk (t) dpki (x)upi (x, t) dxdt
t1 Rn k=1 t1 B(dk ,rpi )
Xm Z t2
= fk (t)upi (dk , t) dt+
k=1 t1
m
X Z t2 Z
fk (t) dpki (x) upi (x, t) − upi (dk , t) dxdt
k=1 t1 B(dk ,rpi )
:= A1 + A2 .
≤ Crp1/2
i
,
R t2
because t1
kup (x, t)kC 0,1/2 dt ≤ C. We then fix r > 0 to find
m Z
X t2
A1 = fk (t)upi (dk , t) dt
k=1 t1
Xm Z t2 Z
= fk (t)− upi (dk , t) − upi (x, t) dxdt
k=1 t1 B(dk ,r)
Xm Z t2 Z
+ fk (t)− upi (x, t) dxdt
k=1 t1 B(dk ,r)
:= B1 + B2 .
Proceeding as before,
|B1 | ≤ Cr1/2 .
61
Evans Chapters 5 - 9
which implies
Z Z m Z t2
t2 X
lim sup fpi (x, t)upi (x, t) dxdt − fk (t)u(dk , t) dt ≤ Cr1/2
pi →∞ t1 Rn t1
k=1
for a.e. t ≥ 0 and v. This establishes (31) and completes the proof. The exercise
then follows from the assumptions of |Du| ≤ 1 and spt a ⊆ {|Du| = 1} which
are included in the assumptions of Theorem 9. The last assumption that
ut − div(Adu) = f in R2 × (0, ∞),
is used in defining the degenerate parabolic PDE
(
up,t − div(|Dup |p−2 Dup ) = fp in Rn × (0, ∞)
up = g on Rn × {t = 0}
62
Evans Chapters 5 - 9
Proof. The idea behind this problem is that convex gradient flows converge
exponentially to a minimizer.
We first review the theory of gradient flows. In general, a partial differential
equation is a gradient flow of an energy functional E : X → (−∞, +∞] on a
metric space (X, d) if the equation may be written as
d
u(t) = −∇d E(u(t)), u(0) = u ∈ X,
dt
for a generalized notion of gradient ∇d .
Here E(u) isn’t continuous with respect to the L2 (U ) norm, even on the re-
stricted set G := H01 (U ) ∩ H 2 (U ). However, its directional derivatives are well
defined for every u ∈ G since
63
Evans Chapters 5 - 9
To see this, we impose the additional assumption that for all u ∈ G := H01 (U ) ∩
H 2 (U ), there exists D2 E(u) : G → H satisfying
(∇E(u + hv) − ∇E(u), w)
(D2 E(u)v, w) = lim , ∀v ∈ G.
h→0 h
The convexity inequality then implies
(∇E(u + hv) − ∇E(u), hv) θ |hv|2 θ
(D2 E(u)v, v) = lim 2
≥ lim = |v|2 .
h→0 h h→0 2 h2 2
Therefore
d 1 θ
|∇E(u(t))|2 = −(D2 E(u(t))∇E(u(t)), ∇E(u(t))) ≤ − |∇E(u(t))|2 ,
dt 2 2
and integrating yields the inequality (15). In our case
ut = −∇E(u(t))
64
Evans Chapters 5 - 9
The same analysis follows from applying Gronwall’s inequality after integration.
Remark: Perhaps Evans intended to “take a more direct route to arrive at this
inequality.” Returning back to the energy functional E : L2 (U ) → R ∪ {+∞}
defined by ( R
1
|∇u(x)|2 dx, for u ∈ H01 (U )
E(u) := 2 U
+∞ otherwise.
We integrate by parts to find
Z Z
d
E(u(t)) = ∇u(t) · ∇ut (t) = − |∆u(t)|2 = −k∇E(u)k2L2 (U ) .
dt U U
65
Evans Chapters 5 - 9
References
[1] G. Aronsson, L.C. Evans, and Y. Wu. “Fast/Slow Diffusion and Growing
Sandpiles”. In: Journal of Differential Equations 131.2 (1996), pp. 304–
335. issn: 0022-0396. doi: https : / / doi . org / 10 . 1006 / jdeq . 1996 .
0166. url: http://www.sciencedirect.com/science/article/pii/
S0022039696901667.
[2] Haim Brezis. Functional Analysis, Sobolev Spaces and Partial Differential
Equations. Springer New York, 2010. doi: 10.1007/978-0-387-70914-7.
url: https://doi.org/10.1007%2F978-0-387-70914-7.
[3] P.G. Ciarlet. Three-Dimensional Elasticity. Mathematical Elasticity. Else-
vier Science, 1994. isbn: 9780444817761. url: https://books.google.
com/books?id=sQiOzyTOJXUC.
[4] Klaus-Jochen Engel and Rainer Nagel. “One-parameter semigroups for lin-
ear evolution equations”. In: Semigroup forum. Vol. 63. 2. Springer. 2001,
pp. 278–280.
[5] Lawrence C Evans. “Partial differential equations and Monge-Kantorovich
mass transfer”. In: Current developments in mathematics 1997.1 (1997),
pp. 65–126.
[6] Giovanni Leoni. “A First Course in Sobolev Spaces”. In: (Jan. 2009). doi:
10.1090/gsm/105.
[7] Walter Rudin. Real and Complex Analysis, 3rd Ed. USA: McGraw-Hill,
Inc., 1987. isbn: 0070542341.
[8] ELIAS M. STEIN. Singular Integrals and Differentiability Properties of
Functions (PMS-30). Princeton University Press, 1970. isbn: 9780691080796.
url: http://www.jstor.org/stable/j.ctt1bpmb07.
66