Elliptic PDE Duke April 12 2021
Elliptic PDE Duke April 12 2021
Tarek M. Elgindi
Abstract
These are notes on Elliptic PDE following the notes of Q. Han and F.H. Lin for the first
part and then Caffarelli and Cabre for the second part. I also benefited from and have taken
from notes of L. Silvestre, C. Mooney, and X. Ros Oton and X. Fernandez.
Contents
1 Introductory Remarks 3
1.1 Some contexts in which Elliptic PDEs arise . . . . . . . . . . . . . . . . . . . . . 3
1.2 What the course is about . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
1.3 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
1
6 The C α and C 1,α Theory 39
6.1 General results on Morrey-Campanato Spaces . . . . . . . . . . . . . . . . . . . . 40
6.2 C α bounds for u . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44
6.3 C 1,α bounds for u . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48
6.4 Exercises on the Newtonian Potential . . . . . . . . . . . . . . . . . . . . . . . . 50
7 De Giorgi’s Theorem 52
7.1 From L2 to L∞ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53
7.2 From L∞ to C α . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60
7.2.1 Harnack Inequality . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62
7.2.2 Oscillation Lemma . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63
7.3 From C α to C 1,α to C ∞ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65
7.4 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67
9 Viscosity Solutions 73
9.1 Elliptic Equations and the Definition of Viscosity Solutions . . . . . . . . . . . . 74
9.1.1 Basic Properties of Viscosity Solutions . . . . . . . . . . . . . . . . . . . . 76
9.2 The Class of solutions of all uniformly Elliptic Equations and the Pucci Operators 79
2
1 Introductory Remarks
Elliptic PDEs arise naturally in a number of different contexts. Two important examples are:
Indeed, an important branch of the calculus of variations is to study functionals of the form:
Z b
L(u) = F (x, u(x), u0 (x))dx. (1.1)
a
It is an exercise to check that whenever F (x, p, z) is convex in the variable z, the Euler-Lagrange
equation associated to the functional is “elliptic.” These variational problems also include a
number of important geometric problems such as the problems of minimal surfaces and obstacle
problems.
Regarding the second class of problems, solutions to so-called semilinear elliptic PDE such
as:
∆ψ = F (ψ) (1.2)
serve as stationary solutions to the incompressible Euler equation and the non-linear heat,
wave, and Schrodinger equations. Even in the context of evolution equations, the pressure
in the incompressible Euler and Navier-Stokes equations solves an elliptic problem. Solutions
to semilinear problems like (1.2) are also helpful in studying blow-up dynamics of non-linear
equations.
The course is about the qualitative properties of solutions to elliptic PDEs. We will be most
3
2. Classical and modern regularity theory.
1.3 Exercises
Remark 1.1. Exercises in this course are of varying degrees of complexity. Some will require one
line and some will require several pages (of course, those requiring more work are not necessarily
more difficult).
1. Derive the equation for the pressure in the incompressible Euler/Navier-Stokes equation
∂t u + u · ∇u + ∇p = ν∆u,
div(u) = 0.
Z 1
0
L(u) = e−u + u2 ,
0
where the admissible space is X = {u ∈ C 2 ([0, 1]) : u(0) = 0, u(1) = 1}. An interesting
1
fact is that if the boundary condition u(1) = 1 is replaced by u(1) = 2, then there is
3. Consider Pn , the vector space of all (real) polynomials defined on Rn . Let’s define the
4
4. Find ψ ∈ C 2 (B1 (0) \ {0}) with ∆ψ ∈ C(B1 (0)) while D2 ψ is not bounded at 0. Next, find
a ψ ∈ C 2 (B1 (0) \ {0}) with det(D2 ψ) ∈ C(B1 (0)) while D2 ψ is not bounded at 0. Could
5. (Functional Analysis) Let K ∈ L2 ([0, 1] × [0, 1]) and assume that K(x, y) = K(y, x) for all
Z 1
LK (f ) = K(x, y)f (y)dy.
0
for all f, g ∈ L2 .
u(0) = 0, u(1) = 0.
5
2 Notation and a few Definitions
For x, y ∈ Rd ,
d
X
(x, y) := xi yi ,
i=1
and
p
|x| := (x, x).
∂Br (x) = {y ∈ Rd : |x − y| = r}
Br (x) = {y ∈ Rd : |x − y| ≤ r}.
We write ∂xi as the partial derivative with respect to xi . We write ∂xki for k−partial derivatives
with respect to xi , ∂xi xj is the mixed partial derivative with respect to xi and xj , etc. One of
d
X
∆ := ∂x2i .
i=1
partial derivatives of all orders in all directions. In this case we write f ∈ C ∞ (Br (x)).
Definition 2.2. Let f : Br (x) → R be measurable. We say that f ∈ L∞ (Br (x)) if there exists
6
When f is continuous and bounded on Br (x), kf kL∞ is just the supremum of |f |.
order α if
|f (y1 ) − f (y2 )|
sup |f (y)| + sup < ∞.
y∈Br (x) y1 ,y2 ∈Br (x) |y1 − y2 |α
|f (y1 ) − f (y2 )|
kf kC α = sup |f (y)| + sup .
y∈Br (x) y1 ,y2 ∈Br (x) |y1 − y2 |α
Definition 2.4. A function is said to belong to C k,α (Br (x)) if it is k−times differentiable and
monic functions
The first place we usually encounter elliptic equations is through the Dirichlet principle. Physi-
cally, this is the problem of finding the steady temperature distribution of a plate whose bound-
ary has a given temperature distribution. For simplicity, let’s take the plate to be B1 (0) ⊂ R2 .
Now let’s assume that the given temperature distribution on ∂B1 (0) is some function g. If the
Definition 3.1. Let Ω ⊂ Rd be open. If u is twice differentiable on Ω and ∆u(y) = 0 for all
y ∈ Ω, u is called harmonic in Ω.
Upon reflecting upon this equation, there are several questions that come to mind:
1. Given a g, does there always exist a solution u to (3.1)-(3.2)? Is it unique? What about
7
if g is rough?
2. Assuming that the first question is settled, what qualitative properties can we tell about
u given g?
4. How much of the above remains once we change the plate from B1 (0) to a general shape
We will touch upon some aspects of some of these questions during these lectures. First, I
3.1 Minimzation
Z
L(u) = |∇u|2 .
B1 (0)
We let
Remark 3.2. This theorem does not show that such a u exists. However, it shows that if a
minimzer of L existed, it would have to satisfy (3.1)-(3.2), which is called the Euler-Lagrange
8
V is called the set of admissible “variations.” Then, for all v ∈ V and t ∈ R, u + tv ∈ X. This
is because v ∈ C 2 (B1 (0)) vanishes on ∂B1 (0). This implies that L(u) = inf t∈R L(u + tv). It is
easy to check that, as a function of t, L(u + tv) is continuously differentiable (in fact, it is just
d
L(u + tv) = 0.
dt t=0
Z Z
d 2 d
0= |∇u + t∇v| = |∇u|2 + 2t(∇u, ∇v) + t2 |∇v|2
dt B1 (0) t=0 dt B1 (0) t=0
Z
= 2(∇u, ∇v).
B1 (0)
now we write:
Z Z
div(v∇u) = (∇u, ∇v) + v∆u.
B1 (0) B1 (0)
Z Z
v∇u · n = v∆u.
∂B1 (0) B1 (0)
Z
v∆u = 0
B1 (0)
9
Lemma 3.3. Assume f ∈ C(B1 (0)) and
Z
fv = 0
B1 (0)
for all v ∈ V = {v ∈ C 2 (B1 (0)) : v(y) = 0 ∀y ∈ ∂B1 (0)}. Then, f (y) = 0 for all y ∈ B1 (0).
Using this lemma we see ∆u(y) = 0 for all y ∈ B1 (0). This concludes the proof of Theorem
(1).
Minimization problems are a great source of elliptic problems which are relevant both in
theoretical and applied mathematics. We now move to another important aspect of elliptic
equations: the maximum principle. Physically, the maximum principle says that if the steady
state temperature profile of a plate is maximized on the boundary of the plate if there are no
Lemma 3.4. Assume that f : Rd → R is twice differentiable. Assume f attains a local maximum
A consequence of this is the maximum principle which is derived from physical intuition: if
we are heating a plate (no matter what shape it is) on its boundaries, the temperature inside will
be lower than the highest temperature on the boundary and higher than the lowest temperature
on the boundary unless the temperature on the boundary is constant. If the temperature on the
boundary were not constant, the temperature should attain its maximum and minimum values
only on the boundary (since the inside values should be some kind of average of the boundary
10
values). What was just described in words is called the strong maximum principle. We will now
give the weak maximum principle which says the maximum and the minimum must be attained
on the boundary.
and
M = sup f (x)
x∈∂Ω
and
m = inf f (x).
x∈∂Ω
(since ∆|x|2 = 2d). This means that f has no interior minimum, since at an interior minimum
of f , we must have ∆f non-negative by the preceding lemma. This implies that
However,
for all > 0. This implies inf x∈∂Ω f (x) ≤ inf x∈Ω f (x) and so they must be equal (since
11
we trivially have the opposite inequality). Now applying the same reasoning to −f , we get
Remark 3.5. Note that if ∆u ≤ 0 in Ω, then u will satisfy the maximum principle on Ω while
What this means is that there is at most one solution of (3.1)-(??) for a given g (uniqueness).
Moreover, we have that if the boundary values of two harmonic functions are close, then they
are close inside as well (stability). The proof follows simply from the fact that u − v is also
Theorem 3. Assume u is twice differentiable and harmonic on Br (x) for some x ∈ Rd and
r > 0. Then,
Z Z
1 1
u(x) = u(y)dy = u(y)dS(y)
|∂Bδ (x)| ∂Bδ (x) |Br (x)| Br (x)
Proof. First note that, by translation invariance, we may assume x = 0. (By replacing u with
ũ(·) = u(· + x), which is then harmonic on Br (0)). Next, notice that
Z
1
lim u(y)dS(y) = u(0).
δ→0 |∂Bδ (0)| ∂Bδ (0)
Indeed,
Z Z Z
1 1 1
u(y)dS(y)−u(0) = (u(y)−u(0))dS(y) ≤ |u(y)−u(0)|dS(y)
|∂Bδ (0)| ∂Bδ (0) |∂Bδ (0)| ∂Bδ (0) |∂Bδ (0)| ∂Bδ (0)
12
Z
1
≤ |∇u|L∞ δdS(y) = δ|∇u|L∞ → 0
|∂Bδ (0)| ∂Bδ (0)
Z
1
I(δ) := u(y)dS(y).
|∂Bδ (0)| ∂Bδ (0)
We will show that I 0 (δ) = 0 for all 0 < δ < r. This will imply that I(δ) = u(x) for all δ < r
Z Z Z
d−1
∆u(y)dy = ∂n u(y)dS(y) = δ ∂δ u(yδ)dS(y).
Bδ (0) ∂Bδ (0) ∂B1 (0)
The first equality uses the divergence theorem and the second is just using the change of variables
Z Z Z
d−1 d−1 d d−1 d 1
δ ∂δ u(yδ)dS(y) = δ u(yδ)dS(y) = δ u(y)dS(y) ,
∂B1 (0) dδ ∂B1 (0) dδ δ d−1 ∂Bδ (0)
where in the last equality we just the changed variables back again yδ → y. Thus we see:
Z Z
d−1 d 1
∆u(y)dy = δ u(y)dS(y) .
Bδ (0) dδ δ d−1 ∂Bδ (0)
This implies that if ∆u ≡ 0, I 0 (δ) = 0 for all δ > 0. This concludes the proof of the surface
Z
1
u(0) = u(y)ds(ω)
|∂Bs (0)| ∂Bs (0)
Z δ Z δ Z
u(0) |∂Bs (0)| = u(ω)dS(ω)ds
0 0 ∂Bs (0)
13
which then implies
Z
u(0)|Bδ (0)| = u(y)dy.
Bδ (0)
The mean value property has a number of important consequences: the strong maximum
Remark 3.7. Since we will make great use of the mean value property, it will be useful to use
Theorem 4. Let Ω be an open set. Assume u ∈ C(Ω) satisfies the mean value property on Ω.
Remark 3.8. A continuous function f on Ω satisfies the mean value property on Ω if whenever
R
Br (x) ⊂ Ω implies that then u(x) = −Bδ (x) u(y)dy for all δ < r. Note that the end of the
preceding proof gives that the “surface” and “volume” mean value properties are equivalent.
Proof. We will follow the proof in the book of Han and Lin. Let φ ∈ Cc∞ (B1 (0)) with
R
B1 (0) φ =1
Z 1
ωd rd−1 ψ(r)dr = 1.
0
1 z
Now define φ (z) = n φ( ) for > 0. Now, for all x ∈ Ω with < dist(x, ∂Ω) we have:
Z Z Z
1
u(y)φ (y − x)dy = u(x + y)φ(y)dy = n u(x + y)φ(y/)dy
Ω |y|<
Z
= u(x + y)φ(y)dy,
|y|<1
14
where we have just changed variables twice. Now we observe:
Z Z 1 Z
u(x + y)φ(y)dy = rn−1 u(x + rω)φ(rω)dSω dr
|y|<1 0 ∂B1 (0)
Z 1 Z
= ψ(r)rn−1 u(x + rω)dSω dr.
0 |ω=1|
R
Now notice that |ω=1| u(x + rω)dSω is independent of r using the mean-value property. Thus,
Z Z 1
u(y)φ (y − x)dy = u(x)ωn ψ(r)rn−1 dr = u(x).
Ω 0
is a C ∞
R
But notice that since u is continuous and φ is compactly supported, Ω u(y)φ (y − x)dy
function of x. Thus, u ∈ C ∞ (Ω). Now we use the equality from the previous proof:
Z Z
d−1 d 1
∆u(y)dy = r u(y)dS(y)
Br (x) dr rd−1 ∂Br (x)
and by the mean-value property the quantity on the left hand side is identically zero for all r
Lemma 3.9. Suppose that f ∈ L1 (B1 (0)) and that = 0 for all φ ∈ Cc∞ (B1 (0)).
R
B1 (0) f ∆φ
Remark 3.10. This means that any “weakly harmonic” function is smooth and harmonic.
Proof. The idea of the proof is to convolve f with an approximation of the identity. We will
then show that f satisfies the local mean value property. That is, we will show that for all
15
x ∈ B1 (0) there exists rx > 0 so that
Z
− f (y) = f (x),
Br (x)
for all r < rx . It then follows from the preceding proof that f is harmonic and smooth. Let
Φ ∈ Cc∞ and Rd Φ = 1. It is easy to construct a radial function with these two properties.
R
Define
1 x
Φ (x) = d
Φ( ).
Define
Z
f (x) = Φ (x − y)f (y)dy.
B1 (0)
Z
∆f (x) = ∆Φ (x − y)f (y)dy,
B1 (0)
for all x ∈ B1 (0). If is sufficiently small, we then see that ∆f (x) = 0. It follows that f
Now we just send → 0 and observe that f must satisfy the (local) mean value property almost
everywhere. It is then easy to show that f is actually continuous. Indeed, by the mean value
Z
1
|f (x) − f (y)| ≤ |f |
|Br | Br (x)∆Br (y)
16
Now we just observe that for an L1 function, given an > 0, there is a δ > 0 so that if |A| < δ,
Z
|f | < .
A
Theorem 5. Let Ω ⊂ Rd be open and connected. Assume u ∈ C(Ω̄) satisfies the mean-value
property. Then, u assumes its maximum and minimum only on ∂Ω unless u is constant.
Proof. We will only do the case of maximum (just multiply by −1 to get the other case!). Let
Since u is continuous, Σ is closed. Now let’s show that Σ is also open. Indeed, let x ∈ Σ ∩ Ω.
Z
M = u(x) = − u(y)dy ≤ M
Br (x)
by the mean value property. Thus, u(y) = M for all y ∈ Br (x). This means that Σ is open and
closed. Thus Σ = Ω or Σ = φ.
Theorem 6. Any harmonic function on Rd which is bounded from above or below is identically
constant
To prove this theorem, we will need to establish gradient estimates for the Laplacian.
17
Theorem 7. Suppose u ∈ C 1 (BR (x0 )) is harmonic in BR (x0 ). Then,
√
d d
|D(u)(x0 )| ≤ sup |u|.
R BR (x0 )
Proof. Observe that ∂i u is harmonic and so we may apply to it the mean-value formula:
Z Z Z
d
∂i u(x0 ) = − ∂i u(y)dy = − div(0, .., u, 0, ...)dy = − u(y)νi dS(y).
BR (x0 ) BR (x0 ) R ∂BR (x0 )
Thus,
d
|∂i u(x0 )| ≤ sup |u|.
R BR (x0 )
Note that using the above proof, we also have that if u ≥ 0 is harmonic on BR , then
d
|∂i u(x0 )| ≤ u(x0 )
R
Theorem 8. Suppose u is harmonic in Ω (an open and connected subset of Rd ). Then, for any
1
u(y) ≤ u(x) ≤ Cu(y)
C
for all x, y ∈ K.
18
Proof. Let x0 ∈ K and R > 0. Assume B2R (x0 ) ⊂ Ω. Now let y ∈ BR (x0 ). On the one hand,
we have
Z Z
u(y) = − u(z)dz ≤ 2n − u(z)dz = 4n u(x0 ).
BR (y) B2R (x0 )
Z Z
1 1
u(y) = − u(z)dz ≥ − u(z)dz = n u(x0 ).
B2R (y) 2n BR (x0 ) 2
Thus,
1
u(x0 ) ≤ u(y) ≤ 2n u(x0 ).
2n
1
u(y) ≤ u(x) ≤ 4n u(y).
4n
Now take x1 , ..., xN ∈ K with the property that BR (xi ) covers K with 4R < dist(K, ∂Ω). Then
N
we get the claimed inequality for C = 41n .
Theorem 9. Suppose u ∈ C 1 (B1 (0)) is a harmonic function in B1 (0). If u(x) < u(x0 ) for any
2
v(x) = e−α|x| − e−α ≥ 0
19
on B1 (0). Notice that
Thus,
∆v(x) > 0
1
if |x| ≥ 2 and α > 2n. Thus, v is subharmonic in B1 \ B1/2 and it is positive in |x| < 1 while it
Since u is harmonic,
∆h > 0
in B1 \ B 1 . Now, we have h (x0 ) = 0 while h ≤ 0. Now, we have u(x) < u(x0 ) for |x| = 21 . Thus,
2
we may take very small so that h < 0 on |x| = 12 . In particular, h achieves its maximum at
the point x0 by the weak maximum principle (since h is maximum at x0 in ∂(B1 \ B 1 )). Thus,
2
∂n h (x0 ) ≥ 0.
This implies that ∂n u(x0 ) ≥ −∂n v(x0 ) = 2α exp(−α). Now we need to bound from below.
Let w(x) = u(x0 ) − u(x) > 0 in B1 . Since w is harmonic in B1 , we may apply to it the Harnack
inequality to get:
inf w ≥ c(n)w(0).
B1/2
Thus,
20
Now since v ≤ 1, we may take = 12 c(n)(u(x0 ) − u(0)). Thus,
α can be taken to be 2n + 1.
Remark 4.1. In the following, we will use the “summation convention” that repeated indices
where δi,j = 1 if i = j and δi,j = 0, i 6= j. In other words, {δi,j }i,j is just the identity matrix.
∆f = div(A∇f ),
where A = Id. Elliptic equations are just those where A is a symmetric positive definite matrix
If for example,
∂i (ai,j ∂j u) = f,
21
u, f ∈ Cc∞ then we have:
Z Z
∂i (ai,j ∂j u)u = uf.
Thus,
Z Z
ai,j ∂i u∂j u = uf ,
It follows that ∇u can be controlled by u itself! This is what is known as the “energy method.”
Let us now give a localized version which is known as the Cacciopolli inequality.
Z
ai,j Di uDj φ = 0, for any φ ∈ C01 (B1 )
In other words, u is a weak solution of ∂j (aij Di u) = 0. Then, for any function η ∈ C01 (B1 ), we
have
Z Λ 2 Z
η 2 |Du|2 ≤ 4 |Dη|2 u2 ,
B1 λ B1
Λ
Remark 4.4. We will find that many of the results we will prove rely on the ratio λ, which
may be important when studying degenerate elliptic problems when λ vanishes at some points
in the domain.
Z
aij Di uDj (η 2 u) = 0.
Therefore,
Z Z
2
η aij Di uDj u = −2 uηaij Di uDj η.
22
It follows that
Z Z sZ sZ
λ η 2 |Du|2 ≤ 2Λ |Du||Dη||u||η| ≤ 2Λ |Du|2 η 2 |Dη|2 u2 .
It follows that
Z Λ 2 Z
2 2
|Du| η ≤ 4 |u|2 |Dη|2 .
λ
Corollary 4.5. Let u be as in the preceding Lemma. Then For any 0 ≤ r < R ≤ 1 we have that
Z Z
16 Λ 2
2
|Du| ≤ u2 .
Br (R − r)2 λ BR \Br
2
Proof. Take η such that η = 1 in Br and η = 0 outside BR and |Dη| ≤ R−r . Why can we do
this?
Lemma 4.6. Let u ∈ C 1 (Ω) and assume that Ω is bounded. If u = 0 on ∂Ω then we have
Z Z
|u|2 ≤ CΩ |Du|2 .
Ω Ω
In general,
Z Z
2
|u − ū| ≤ C2,Ω |Du|2 .
Ω Ω
Corollary 4.7. Let u be as in the last two results. Then, for any 0 < R ≤ 1, there holds
Z Z
2
u ≤ θ1 u2
BR/2 BR
23
and
Z Z
2
|Du| ≤ θ2 |Du|2 ,
BR/2 BR
3
Proof. Let η ∈ C01 (BR ) with η = 1 on BR/2 and |Dη| ≤ R. Then,
Z Z Z Λ 2 Z
2 2 2 2 2 2
|D(ηu)| = |ηDu + uDη| ≤ 2 |Du| η + u |Dη| ≤ 2(1 + 4 ) |Dη|2 u2 ,
BR BR λ BR
where
Λ 2
C1 = 18(1 + 4 ).
λ
Z Z
2 2
(ηu) ≤ c(n)R |D(ηu)|2 ,
BR BR
where c(n) is a constant depending only on the dimension. How can we remember that we get
Z Z
(ηu)2 ≤ c(n)C1 u2 .
BR BR \BR/2
Z Z
u2 ≤ c(n)C1 u2 .
BR/2 BR \BR/2
Thus,
Z Z
C1 c(n)
2
u ≤ u2 .
BR/2 C1 c(n) + 1 BR
24
Therefore, the first part of the theorem works with
C1 c(n)
θ1 = < 1.
C1 c(n) + 1
Now we move to the second inequality. To achieve this, we apply Corollary (4.5) with r = R/2
R
to u − a, where a = −BR \B u. Then we have:
R/2
Z Z Λ 2 Z
64 Λ 2
2 2
|Du| ≤ 2 |u − a| ≤ 64c2 (n) |Du|2 .
BR/2 R λ BR \BR/2 λ BR \BR/2
2
Λ
64c2 (n) λ
Thus, we can take θ2 = 2 , where c2 (n) is the second Poincaré constant.
Λ
64c(n) λ
+1
We only assume that aij , bi , c are bounded and continuous in Ω̄ and that aij is elliptic. One of the
purposes of this section is to use maximum principles to prove a-priori estimates and establish
symmetry results. Another nice application will be the Alexandroff Maximum Principle.
Remark 5.1. A consequence of ellipticity is that aii ≥ λ for each i, just by choosing ξ = ei in
Lemma 5.2. If u ∈ C 2 (Ω) ∩ C(Ω̄) and Lu > 0 in Ω and c ≤ 0, then u cannot attain a
sup u ≤ sup u+
Ω ∂Ω
Proof. Assume it attains a non-negative maximum at x = x∗ . Then, aij (x∗ )∂ij u(x)∗ ≤ 0. Why
25
Remark 5.3. Is the condition c ≤ 0 necessary?
Proof. The proof is somewhat similar to the case of the Laplacian done in the previous section.
Note that a11 ≥ λ. Define hα (x) = exp(αx1 ). We are going to choose α very large to magnify
Lhα = exp(αx1 )(α2 a11 (x) + αb1 (x) + c(x)) ≥ exp(αx1 )(α2 λ + αb1 (x) + c(x))
where B = |b|L∞ , C = |c|L∞ . Now if α is large, we get that Lhα > 0. Next, define
w = u + hα .
We have that
Lw > 0,
sup w ≤ sup w+
Ω ∂Ω
Consequently,
Lu = f, in Ω
26
u = g, on ∂Ω.
To get uniqueness for the Neumann problem, we need the Hopf Lemma.
Theorem 11. Let B be an open ball in Rn and assume that u ∈ C 2 (B) ∩ C(B \ {x0 }) satisfies
for any x ∈ B and u(x0 ) ≥ 0. Then, for any outward direction ν at x0 with ν · n > 0, we have:
1
lim inf (u(x0 ) − u(x0 − tν)) > 0.
t→0 t
Remark 5.5. That the limit is ≥ 0 is obvious from the assumptions but all the ≥ 0 statement
requires is that u(x0 ) ≥ u(x) for all x rather than the strict inequality. The assumption of
Proof. We may assume that B = Br (0), u ∈ C(B̄), and that u(x) < u(x0 ) for all x ∈ B̄ \ {x0 }.
Why? Just as before, our goal will be to perturb u a little bit so that we keep the maximum
at x0 but so that the normal derivative changes. Note that since u(x) < u(x0 ) throughtout
B̄ \ {x0 },
As in the proof of the Hopf Lemma for the Laplacian, let us define:
Note that h vanishes on ∂Ω. Let us Claim that Lh > 0 in Br \ Br/2 so long as α is sufficiently
v = u + h,
27
we will have that v(x) ≤ v(x0 ), for all x ∈ ∂Br/2 , so long as < u(x0 ) − M . Thus, by the weak
v(x) ≤ v(x0 ),
Thus,
v(x0 ) − v(x0 − tν)
lim inf ≥ −∂ν h(x0 ).
t→0 t
Lh = exp(−α|x|2 ) 4α2 aij (x)xi xj − 2αaii (x) − 2αbi (x)xi + c − c exp(−αr2 ) − c(x) exp(−αr2 )
≥ exp(−α|x|2 ) 4α2 λ|x|2 − 2αaii (x) − 2αbi (x)xi + c ≥ exp(−α|x|2 )(α2 λr2 − (α + 1)K),
for some fixed K depending only on r, c, bi , Λ. Thus, taking α large, we get that Lh > 0.
A nice corollary of the Hopf Lemma is the Strong Maximum Principle (which then justifies
constant.
Proof. As in the proof for harmonic functions, let us define the set
Σ = {x ∈ Ω : u(x) = M },
where M is the non-negative maximum of u on Ω̄. Our goal is to show that Σ is open and closed
28
In this general case, assume toward a contradiction that Σ is a non-empty proper subset of
Ω. Note that since Ω is open and connected, it is also path connected. Note that Ω \ Σ is open.
Now take x1 ∈ Ω \ Σ and x2 ∈ Σ. Draw a continuous path inside of Ω from x1 to x2 and let x∗
be the first time this path hits Σ. Since x∗ ∈ Ω, there is a small ball B1 around x∗ which lies in
Ω. This ball contains members of Σ and Ω \ Σ by construction. Thus, we may take y ∈ B1 for
which δ := d(y, Σ) < d(y, ∂Ω). Thus, there is an x ∈ Σ ∩ B̄δ satisfying the conditions of the Hopf
Lemma. It follows that ∂n u(x) > 0. But ∇u(x) = 0 since x∇Σ. This is a contradiction.
Ω or u ≡ 0 in Ω.
Then, if Lu ≥ 0, we have
Theorem 14. Suppose there exists w ∈ C 2 (Ω) ∩ C 1 (Ω̄) with w > 0 in Ω̄ and Lw ≤ 0 in Ω. If
u
u ∈ C 2 (Ω) ∩ C 1 (Ω̄) satisfies Lu ≥ 0 in Ω, then w cannot assume its non-negative maximum in
u u
Ω unless w ≡ constant. Additionally, if w assumes its non-negative maximum at x0 ∈ ∂Ω and
u
w 6≡ constant, then
u
∂ν (x0 ) > 0.
w
u
Proof. Set v = w. Then, v satisfies
Lw
aij Dij v + Bi Di v + v ≥ 0,
w
29
for some Bi . Indeed,
u 1 2 1
aij Dij = aij Dij u − 2 aij Di uDj w + uaij Dij
w w w w
1 2 Dj w
= aij Dij u − 2 aij Di uDj w − uaij Di 2
w w w
1 u Di Dj w 2 1
= aij Dij u − aij − 2 aij Di uDj w + 2uaij 3 Di wDj w
w w w w w
Moreover,
u 1 u Di w
Di = Di u − .
w w w w
Let’s define
1
bˆi = aij Dj w
w
It follows that
u u Lw
aij Dij + Bi Di + v=
w w w
1 u Di Dj w ˆ 2 2u u Di w u
1
aij Dij u− aij +bi − Di u+ 2 Di w +Bi Di u− + 2 aij Dij w+bi Di w +cv,
w w w w w w w w w
Now we define:
Bi := 2b̂i + bi .
Then we see:
Lw 1
aij Dij v + Bi Di v + v= Lu ≥ 0.
w w
Lw
Note that w ≤ 0. This means we can apply the strong maximum principle and Hopf Lemma
to v.
Remark 5.6. This implies that L has a comparison principle. If Lu ≥ 0 and u ≤ 0 on ∂Ω then
u ≤ 0 on Ω.
Proposition 5.7. Assume that |(x − y) · e| < d for all x, y ∈ Ω, where e is a unit vector and
30
d > 0. There exists d0 > 0 depending on λ, |bi |L∞ , |c+ |L∞ so that if d ≤ d0 , the assumptions of
Proof. Without a loss of generality, let us assume that e = (1, 0, ..., 0) and that Ω̄ ⊂ {0 < x1 <
d}. Assume that sup |bi |L∞ , |c+ |L∞ ≤ N . Define w = eαd − eαx1 > 0 in Ω̄. Observe that
Now we choose α large, depending only on N and λ, so that −α2 λ + αN < −4N . It follows
that
1
Now we just take d0 = α and we are done.
∂xx + k 2 : H 2 → L2
2π
is an isomorphism on spatially periodic functions of period strictly less than k .
Remark 5.8. Later we will show that if the diameter of Ω is just bounded (not necessarily
small) but if the volume |Ω| is small, then we again have a comparison principle.
Lu = f in Ω
u=φ on ∂Ω,
31
for some f ∈ C(Ω̄) and φ ∈ C(∂Ω). If c ≤ 0, then we have:
Proof. Define the positive constants F, Φ by F = |f |L∞ and Φ = |φ|L∞ . Now, suppose that we
Lw ≤ −F in Ω
w≥Φ on ∂Ω.
It would then follow by the comparison principle that L(w + u) ≤ 0 and w + u ≥ 0 on ∂Ω. From
w ≥ |u|
Suppose that Ω ⊂ {0 < x1 < d}. Note that we don’t assume that d is small. Set
≥ (λα2 + b1 α)F ≥ F
|w| ≤ Φ + eαd F,
32
from which the result follows by the comparison principle as we discussed before.
L = aij Dij + bi Di .
Lemma 5.10.
Proof.
Lemma 5.11.
1
L(|Du|2 ) ≥ λ|D2 u|2 − C|Du|2 − |D(Lu)|2
2
Proof.
So that
Thus,
33
≥ λ|D2 u|2 + 2Dk u(Dk (aij Dij u) − Dk aij Dij u) − C|Du||D2 u|
1
≥ λ|D2 u|2 − C|Du|2 − |D(Lu)|2 .
2
Proof. Let
L = aij Dij + bi Di .
34
Next, we have the interior gradient estimate.
Proposition 5.13. Under the same assumptions as the previous proposition, if Ω0 ⊂⊂ Ω then
sup |Du| ≤ C,
Ω0
Thus,
Thus,
1 λγ
Lv ≥ λγ |D2 u|2 + ( − Λ2 |Dγ|2 )|D2 u|2 − C|Du|2 − C .
2 2
|Dγ|2 ≤ Cγ
35
5.2 Boundary Estimates
The preceding results were on interior estimates (or global estimates assuming boundary esti-
mates). Now we will show how to establish boundary estimates. We will assume that u satisfies
Lu = f (x, u) in Ω,
u = φ on ∂Ω,
• f ∈C
• φ ∈ C 2 (Ω̄).
• L is uniformly elliptic
for all x ∈ Ω and x0 ∈ ∂Ω, where C depends only on the sup norm of f, aij , b, u and λ and |φ|C 2 .
Proof. By replacing u → u + φ, we may assume that u = 0 on ∂Ω. We may also assume that
L = aij Dij + bi Di .
L(±u) ≥ −F.
Lw ≤ −F
36
in Ω while w(x0 ) = 0 and w|∂Ω ≥ 0. Once we can show that the Lipschitz norm of w is bounded
Consider the exterior ball BR (y) with BR (y) ∩ Ω̄ = {x0 }. Define d(x) = |x − y| − R for any
ψ(0) = 0
|ψ 0 | ≤ C∗ .
Let us also require that ψ 0 > 0 and ψ 00 < 0. Now let’s compute Lw for x ∈ Ω:
Now,
xi − yi
Di d(x) = ,
|x − y|
Thus,
aii aij (xi − yi )(xj − yj ) nΛ λ
aij Dij d ≤ − 3
≤ −
|x − y| |x − y| |x − y| |x − y|
nΛ − λ
≤ .
R
Thus,
nΛ − λ
Lw ≤ ψ 00 λ + ψ 0 + |b|L∞ .
R
Lw ≤ −F,
37
we need
nΛ − λ
λψ 00 + ψ 0 ( + |b|) + F ≤ 0.
R
ψ 00 + Aψ 0 + B = 0.
Hence,
B C
ψ(d) = − d + (1 − e−Ad ).
A A
Observe that
B
ψ 0 = Ce−Ad −
A
and
ψ 00 = −ACe−Ad .
5.3 Exercises
• Show that eigenvalues of K must be real and that eigenvectors with different eigen-
• Let M := sup|v|=1 |(Kv, v)|. Show that there exists v∗ with |v∗ | = 1 satisfying so that
|(Kv∗ , v∗ )| = M .
38
• Conclude that K has a complete basis of eigenvectors.
• Given q ∈ C([0, 1]), prove that there exists an orthonormal basis en of L2 ([0, 1])
satisfying:
e00n + qen = λn en
en (0) = en (1) = 0.
2. Find a function ψ ∈ C ∞ (R3 ) with lim|x|→∞ ψ(x) = 0 and ∆ψ > 0 on R3 . Is this possible
in R2 ?
3. Assume that ∆ψ = 0 on B1 (0) and that ψ and ∂n ψ both vanish on a segment of ∂B1 (0).
Show that ψ ≡ 0.
∆ψ = 0 in y > 0,
ψ = φ on y = 0,
satisfies |∂y φ(0, y)| → ∞ as y → 0. Extra credit: show that if φ ∈ C 1,α for some α > 0,
We will now study the basics of the Schauder theory. Here, we will assume that the ellipticity
coefficients aij are continuous. We will only consider equations in the following form:
39
Recall that this equation has nice energy properties and we will use these to derive regularity
results on solutions u of the above equation. We will start by showing how local energy control
Z
|u − ux,r |2 ≤ M 2 rn+2α ,
Br (x)
for any Br (x) ⊂ Ω, for some α ∈ (0, 1). Then, u ∈ C α (Ω) and for any Ω0 ⊂⊂ Ω there holds
|u(x) − u(y)|
sup |u| + sup ≤ c M + |u| L2 ,
Ω0 x,y∈Ω0 |x − y|α
where c = c(n, α, Ω, Ω0 ).
|ux0 ,r1 − ux0 ,r2 |2 ≤ 2 |u(x) − ux0 ,r1 |2 + |u(x) − ux0 ,r2 − u(x)|2
R R
Thus, if r1 = 2i+1
and r2 = 2i
, we see that
c(n, α)
|ux0 ,2−h R − ux0 ,2−k R | ≤ M Rα .
2hα
40
Note that c(n, α) blows up as α → 0. This means that {ux0 ,2−i R } is Cauchy and its limit is
independent of R. Call the limit û(x0 ) It follows that for every r we have that
This implies that u is uniformly bounded by CM Rα + ux,R for any x ∈ Ω0 , while the latter can
be estimated by the L2 norm of u. Next, let us give the bound on the Hölder semi-norm of u.
R0
Let x, y ∈ Ω0 and let R = |x − y| < 2 . Then,
The first two are bounded by CRα as before. To estimate the last one we write
and we average over ζ ∈ B2R (x) ∩ B2R (y). Noting that BR (x) ⊂ B2R (x) ∩ B2R (y), we get
Z
|Du|2 ≤ M 2 rn−2+2α
Br (x)
for some α ∈ (0, 1) and for all Br (x) ⊂ Ω. Then u ∈ C α and for any Ω0 ⊂⊂ Ω we have the
41
Proof. By the Poincaré inequality,
Z Z
|u − ux,r |2 ≤ c(n)r2 |Du|2 ≤ c(n)M 2 rn+2α .
Br (x) Br (x)
Z
|Du|2 ≤ M rµ ,
Br (x0 )
for any Br (x0 ) ⊂ Ω, for some µ ∈ [0, n). Then, for any Ω0 ⊂⊂ Ω there holds
Z
|u|2 ≤ c(n, λ, µ, Ω, Ω0 )(M + |u|2L2 )rλ ,
Br
Proof. Note that λ < n in all cases. By the Poincaré inequality, we have that
Z
|u − ux0 ,r |2 ≤ c(n)M rµ+2 ≤ c(n)M rλ .
Br (x0 )
Z ρ n Z
2 λ n 2 λ
u ≤ 2c(n)M r + c(n)ρ |ux0 ,r | ≤ 2c(n)M r + c(n) u2 ,
Bρ r Br
where we used the Cauchy-Schawrz inequality in the second inequality and the triangle inequality
satisfies:
ρ n
φ(ρ) ≤ C M rλ + φ(r) ,
r
42
for any ρ < r ≤ R0 and for some λ ∈ (0, n). If we could replace rλ on the right by ρλ we would
Lemma 6.3. Let φ(t) be nonnegative and nondecreasing and assume that
ρ
φ(ρ) ≤ A ( )α + φ(r) + Brβ ,
r
whenever ρ ≤ r ≤ R. Then, for any γ ∈ (β, α) there is a constant 0 so that if < 0 we have
ρ
φ(ρ) ≤ c ( )γ φ(r) + Bρβ .
r
Choose τ < 1 so that 2Aτ α = τ γ and assume 0 τ −α < 1. Then, for r < R
and therefore,
Bτ kβ rβ
≤ τ (k+1)γ φ(r) + .
1 − τ γ−β
1 ρ γ Bρβ
φ(ρ) ≤ φ(r) + .
τγ r τ 2β (1 − τ γ−β )
43
6.2 C α bounds for u
As we explained before in class, the key idea is that when the coefficients are continuous we
can compare the solution to the solution to the constant-coefficient problem. Let us start by
recalling the basic estimate for harmonic functions. We have already established this (which
follows from the interior regularity results, using the surface mean-value property), but let us
aij Dij w = 0
Z ρ n Z
2
|Dw| ≤ c |Dw|2
Bρ (x0 ) r Br (x0 )
Z ρ n+2 Z
|Dw − (Dw)x0 ,ρ |2 ≤ c |Dw − (Dw)x0 ,r |2
Bρ (x0 ) r Br (x0 )
Proof. Exercise.
Corollary 6.5. Suppose w is as in the preceding Lemma. Then, for any u ∈ H 1 (Br (x0 )) there
Z ρ Z Z
n
|Du|2 ≤ c |Du|2 + |D(u − w)|2
Bρ (x0 ) r Br (x0 ) Br (x0 )
and
Z ρ Z Z
2 n+2 2
|Du − (Du)x0 ,ρ | ≤ c |Du − (Du)x0 ,r | + |D(u − w)|2
Bρ (x0 ) r Br (x0 ) Br (x0 )
44
Proof.
Now we will use the preceding two results to establish the following Theorem.
Dj (aij Di u) + cu = f
Assume that aij ∈ C(B̄1 ) with modulus of continuity τ and that c ∈ Ln (B1 ), f ∈ L1 (B1 ) for some
q ∈ (n/2, n). Then, u ∈ C α (B1 ) with α = 2− nq . Moreover, there exists an R0 = R0 (λ, Λ, τ, |c|Ln )
Z
|Du|2 ≤ Crn−2+2α |f |2Lq + |u|2H 1 ,
Br (x)
Proof. We will focus on the case n > 2. The case n = 2 should be checked separately (and is
Z Z Z
aij (x0 )Di wDj φ = f φ − cuφ + (aij (x0 ) − aij (x))Di uDj φ.
B1
w = u on ∂Br (x0 ) has a unique solution. Thus we may choose φ = u − w in the following:
Z Z
Di vDj φ = f φ − cuφ(aij (x0 ) − aij (x))Di uDj φ.
Br (x0 )
Z 2n
(n−2)/2n Z 1/2
v n−2 ≤ c(n) |Dv|2 .
Br Br (x0 )
45
2n 2n
Observe that the conjugate exponent to n−2 is n+2 . Note that
1 1 n−2
+ + = 1.
n 2 2n
Z Z
|Dv|2 ≤ c τ (r)2 |Du|2 + |c|2Ln |u|2L2 + |f |2L2n/(n+2) .
Br (x0 ) Br (x0 )
Z Z
2 2 ρ n
|Du| ≤ c (τ (r) + ) |Du|2 + |c|2Ln |u|2L2 + |f |2L2n/(n+2) .
Bρ (x0 ) r Br (x0 )
Z Z
2
2 ρ n
|Du| ≤ c (τ (r) + ) |Du|2 + rn−2+2α |f |2Lq + |c|2Ln |u|2L2 .
Bρ (x0 ) r Br (x0 )
Case 1: c ≡ 0.
Z Z
2
2 ρ n 2 n−2+2α 2
|Du| ≤ c (τ (r) + ) |Du| + r |f |Lq .
Bρ (x0 ) r Br (x0 )
Now we apply our technical Lemma and note that there is a R0 for which τ (r) as small as we
Z ρ n−2+2α Z
2
|Du| ≤ C |Du|2 + ρn−2+2α |f |2Lq .
Bρ (x0 ) r Br (x0 )
46
Applied at r = R0 , we arrive at the result.
We have that
Z ρ Z Z
2 n 2 2 n−2+2α 2
|Du| ≤ C ( ) + τ (r) |Du| + r |f |Lq + u2 .
Bρ (x0 ) r Br (x0 ) Br
Z ρ Z h
2 n 2
|Du| ≤ C ( ) + τ (r) |Du|2 + rn−2+2α |f |2Lq + |u|2H 1 .
Bρ (x0 ) r Br (x0 )
Z Z Z
u2 ≤ Crδ1 ( u2 + |Du|2 ),
Br (x0 ) B1 B1
Z ρ Z h
2 n 2
|Du| ≤ C ( ) + τ (r) |Du|2 + max{rn−2+2α , rδ1 } |f |2Lq + |u|2H 1 .
Bρ (x0 ) r Br (x0 )
Z
|Du|2 ≤ Cr2 (|f |2Lq + |u|2H 1 ),
Br
n/2 < q < n, then u ∈ C α with α = 2 − nq . Is this sharp? Check homogeneous examples.
47
6.3 C 1,α bounds for u
−Dj (aij Di u) + cu = f.
We now move to prove the Hölder continuity of Du, under the right assumptions on f and c.
Theorem 17. Let u ∈ H 1 (B1 ) be a weak solution to the above elliptic problem. Assume that
n
aij ∈ C α (B̄1 ) and c, f ∈ Lq (B1 ) with q > n. Let α = 1 − q ∈ (0, 1). Then, Du ∈ C α (B1 ).
Moreover, there exists R0 = R0 (λ, |aij |C α , |c|Lq ) so that for any x ∈ B1/2 and r ≤ R0 , there
holds
Z
|Du − (Du)x,r |2 ≤ Crn+2α (|f |2Lq + |u|2H 1 ),
Br (x)
Proof. As before, we decompose u into v + w where w satisfies the homogeneous equation with
Z Z
aij (x0 )Di vDj φ = f φ − cuφ + (aij (x0 ) − aij (x))Di uDj φ.
Br Br (x0 )
Z Z
|Dv|2 ≤ C τ (r)2 |Du|2 + |c|2Ln |u|2L2 (Br ) + |f |2L2n/(n+2) .
Br (x0 ) Br (x0 )
Now, remember that u = v − w, where w satisfies the constant coefficient homogeneous problem.
Z ρ Z
2 n 2
|Du| ≤ C ( + τ (r) ) |Du|2 + |c|2Ln |u|2L2 (Br ) + |f |2L2n/(n+2)
Bρ r Br
48
and
Z ρ Z Z
n+2
|Du−(Du)x0 ,ρ |2 ≤ C |Du−(Du)x0 ,r |2 +τ (r)2 |Du|2 +|c|2Ln |u|2L2 (Br ) +|f |2L2n/(n+2) .
Bρ r Br Br
while
In the case where c ≡ 0 and aij (x) ≡ aij (x0 ), the result is clear since we then get that
Z ρ Z
2 n+2
|Du − (Du)x0 ,ρ | ≤ C |Du − (Du)x0 ,r |2 + rn+2α |f |2Lq ,
Bρ r Br
from which the result follows using the technical Lemma. Next, let us look at the case
Z ρ Z
n
|Du|2 ≤ C ( + τ (r)2 ) |Du|2 + rn+2α |f |2Lq .
Bρ r Br
Z
|Du|2 ≤ Crn−2δ (|f |2Lq + |u|2H 1 ).
Br
Z ρ Z
2 n+2
|Du − (Du)x0 ,ρ | ≤ C |Du − (Du)x0 ,r |2 + rn+2α−2δ (|f |2Lq + |u|2H 1 ) .
Bρ r Br
49
Now using the technical Lemma again gives us that Du ∈ C α−δ . It follows that
Z
|Du|2 ≤ c(n)rn ,
Br
Z ρ Z
n+2
|Du − (Du)x0 ,ρ |2 ≤ C |Du − (Du)x0 ,r |2 + rn+2α (|f |2Lq + |u|2H 1 ) .
Bρ r Br
The following problems give a different point of view on the regularity theory for the Laplacian
in Rd by studying properties of the Newtonian potential. It turns out that for the scenarios
we studied in the preceding section, all the results can be established perturbatively from the
results we can get using the Newtonian potential. This approach is adopted, for example, in the
1. Define:
cd
K(x) = ,
|x|d−2
Z
ψ(x) = K(x − y)f (y)dy.
Rd
∆ψ = f
on Rd .
50
• Change variables in the integral so that all derivatives fall on f . By translation
• Send → 0 and observe that all terms tend to zero except one boundary term which
∆ψ = f,
• For d ≥ 3, find the formula for K using the Fourier transform (Hint: what must K̂(ξ)
be?)
• Use the preceding idea to find the kernels of the fractional Laplacian operators.
2. We now move to look at a-priori estimates for solutions. In the following, we will take
• Fix p < d < q. Prove that there is a constant Cp,q (independent of f ) so that
51
• Show there is a universal C (independent of f ) so that
Hint: Deduce from the formula that ∇ψ, D2 ψ ∈ L2 and decay. Next, compute
(∆ψ)2 and use that ∆ψ = f . If you integrate by parts, make sure to justify why
R
this is allowed. Note that ψ, ∇ψ, D2 ψ do decay, but not very fast.
• Bonus: If 0 < α < 1, show that we can replace |D2 ψ|L∞ by |D2 ψ|C α in the preceding
estimate.
7 De Giorgi’s Theorem
Hilbert’s 19th problem was to establish the regularity of minimizers to functionals of the form:
Z
F (∇u),
Ω
when F is a smooth convex function. The problem is that the “Direct Method” of the Calculus
div(∇F (∇u)) = 0,
which is just
∂ij F (∇u)∂ij u = 0.
52
By differentiating the equations and using the Schauder theory, it is possible to deduce C ∞
smoothness of u once we know, for example, that u ∈ C 1,α . All what we know from the energy
is that u ∈ H 1 . The De Giorgi-Nash theorem consists of establishing first that any H 1 weak
solution to
aij Dij u = 0
is uniformly bounded by the L2 norm when the coefficients are merely bounded and measurable.
After that, we will establish C α regularity of u and C 1,α regularity in the case that the functional
F is smooth. Higher smoothness follows from the Schauder Theory. We will follow De Giorgi’s
proof.
7.1 From L2 to L∞
Theorem 18. Assume that aij ∈ L∞ (B1 ) and c ∈ Lq (B1 ) for some q > n/2 :
while
aij ξi ξj ≥ λ|ξ|2 .
Z
aij Di uDj φ + cuφ ≤ fφ
with
C = C(n, λ, Λ, p, q).
53
Proof. The proof is somewhat complicated and involves a number of steps. First, define:
v = (u − k)+ ,
φ = vζ 2 .
Hencewhen we take this choice of test function, the domain of integration will be the set {u > k}:
Z
aij Di uDj φ = aij Di uDj vζ 2 + 2aij Di uDj ζvζ
Z Z
≥λ |Dv|2 ζ 2 − 2Λ |Dv||Dζ|vζ
Λ2
Z Z
λ 2 2
≥ |Dv| ζ − 2 |Dζ|2 v 2 ,
2 λ
where we just used the Cauchy-Schwartz inequality. Note that this is just the same trick we
did in the proof of the Cacciopolli inequality. Thus, using the definition of subsolution, we have
that
Z Z Z Z Z
2 2 Λ 2 2 2 2 2
|Dv| ζ ≤ C( ) v |Dζ| + |c|v ζ + k |c|ζ + |f |vζ 2 ,
2
λ
where we just used that |u| ≤ k + v. Again we recall that vζ ∈ H01 (B1 ) so that we can apply
|vζ|L2n/(n−2) ≤ c(n)|D(vζ)|L2 .
Z
1− n−2 − 1q
|f |vζ 2 ≤ |f |Lq |vζ|L2n/(n−2) |{}| 2n
1 q
+n − 1q
≤ c(n)|f |Lq |D(vζ)f |L2 |{vζ 6= 0}| 2
54
c(n) 2 1− 1
≤ δ|D(vζ)|2L2 + |f |Lq |{vζ 6= 0}| q ,
δ
Z Z Z
2 Λ
2 2 2 2 2 1− 1q
|D(vζ)| ≤ C( , n) v |Dζ| + |c|v ζ + k |c|ζ 2 + F 2 |{vζ 6= 0}| ,
λ
with F = |f |Lq (B1 ). Now, we next observe using the same estimate as above on the f term
Z
1− n−2 − 1q
|c|v 2 ζ 2 ≤ c(n)|c|Lq |D(vζ)|2L2 |{vζ 6= 0}| n
2
− 1q
≤ c(n)|c|Lq |D(vζ)|2L2 |{vζ 6= 0}| n .
Z Z
2
|c|ζ ≤ |c| ≤ |c|Lq |{vζ 6= 0}|1−1/q .
Consequently, if |{|vζ 6= 0}| is small enough (depending on n and |c|Lq ), we have that
Z Z
Λ2 2 2 2 2 1− 1q
|D(vζ)| ≤ C( , n) v |Dζ| + (k + F )|{vζ 6= 0}| . (7.1)
λ
Note that if k is large, we must have that |{vζ = 0}| is quite small since v ∈ L2 . Next, with
Z
(vζ)2 ≤ |vζ|2L2n/(n−2) |{vζ 6= 0}|2/n ≤ c(n)|D(vζ)|L2 |{vζ 6= 0}|2/n .
Z Z 2
− 1q
2 1+ n
(vζ) ≤ C v 2 |Dζ|2 |{vζ 6= 0}|2/n + (k + F )2 |{vζ 6= 0}| .
55
In particular, if |{vζ 6= 0}| is small, we have that there is an > 0 so that
Z Z
(vζ)2 ≤ C v 2 |Dζ|2 |{vζ 6= 0}| + (k + F )2 |{vζ 6= 0}|1+ .
We now choose the cut-ff function. For fixed 0 < r < R ≤ 1 we choose ζ ∈ C0∞ (BR ) such
2
that ζ ≡ 1 in Br and 0 ≤ ζ ≤ 1 and |Dζ| ≤ R−r in B1 . Now set
A(k, r) = {x ∈ Br : u ≥ k}.
Z Z
2
1
(u − k) ≤ C 2
|A(k, R)| (u − k)2 + (k + F )2 |A(k, R)|1+ .
A(k,r) (R − r) A(k,R)
The key here is that we have a freedom of choosing two scales when using this inequality, the
Z
1 1 +
|A(k, r)| ≤ u+ ≤ |u |L2 . (7.2)
k A(k,R) k
1
In fact, we have that is bounded by k2
on the right hand side, but we do not need this extra
smallness I guess. Thus (7.1) holds so long as k ≥ k0 = C|u+ |L2 where C = C( Λλ , |c|Lq , n). Our
goal is to now show that there is a k large, precisely when k ≥ k∗ = C(k0 + F ) so that
Z
(u − k)2 = 0.
A(k, 12 )
This will imply that u+ ≤ k on B 1 . Now, let us take h > k ≥ k0 0 < r < 1. Then,
2
Z Z Z
2 2
(u − h) ≤ (u − k) ≤ (u − k)2 ,
A(h,r) A(h,r) A(k,r)
since if u ≥ h ≥ k, we have that (u − h)2 ≤ (u − k)2 gives the first inequality and the A(h, r) ⊂
56
A(k, r) gives the second. Moreover,
Z
1
|A(h, r)| ≤ (u − k)2 .
(h − k)2 A(k,r)
Z Z
1
(u − h)2 ≤ C|A(R, h)| (u − h)2 + (h + F )2 |A(h, R)|
A(h,r) (R − r)2 A(h,R)
(h + F )2
Z
1 1 1+
≤C + (u − k)2 .
(R − r)2 (h − k)2 (h − k)2 A(k,R)
Hence, if we define φ(k, r) = |(u − k)+ |L2 (Br ) , for h > k ≥ k0 and R > r we have
1 h+F 1
φ(h, r) ≤ C + φ(k, R)1+ . (7.3)
R−r h − k (h − k)
Note that φ is increasing in R and decreasing in k as we established before. Our goal is to now
show that
1
φ(k∗ , ) = 0
2
1
kl = k0 + k(1 − )
2l
1 1
rl = + l+1 ,
2 2
Let us thus apply (7.3) for (h, r) = (kl , rl ) and (k, R) = (kl−1 , rl−1 ). Observe that
1 k
rl−1 − rl = kl − kl−1 = .
2l+1 2l
Then we see:
C l(1+) k0 + k + F
φ(kl , rl ) ≤ 2 1 + φ(kl−1 , rl−1 )1+ .
k k
57
Thus,
C l(1+)
φ(kl , rl ) ≤ 2 φ(kl−1 , rl−1 )1+ .
k
φ(k0 , r0 )
φ(kl , rl ) ≤ ,
γl
for all l ≥ 0. Once the claim is established, we send l → ∞ and see that φ(k∗ , 12 ) = 0. We
establish this claim by induction on l. Clearly it is true for l = 0. Assume it is true for l − 1.
Now we have:
Remember that we still have to choose k and γ. We are done if we make a choice of k, γ so that
γ = 21+ .
Now we move to the general case 0 < θ < 1 and p > 0, which will follow essentially by scaling
and interpolation. The first thing to observe is that scaling down does not affect the ellipticity
1
sup u+ ≤ C |u+
| L2 (B ) + R
2−n/q
|f |Lq (B )
BR/2 Rn/2 R R
58
Now let us get an idea of how this works. By iterating this bound, it is thus easy to see that
sup u+ ≤ C 2nj/2 |u+ |L2 (B1 ) + |f |Lq (B1 ) .
B1−2−j
1
sup u+ ≤ C |u+
| L2 (B ) + |f |Lq (B ) .
1
Bθ (1 − θ)n/2 1
1
sup u+ ≤ C |u+
| 2
L (BR ) + |f |L (B1 ) .
q
BθR ((1 − θ)R)n/2
2−p + p
|u+ |2L2 (BR ) ≤ |u+ |L∞ |u |Lp .
1
|u+ |L2 ≤ δ 2/(2−p) |u+ |L∞ + |u+ |Lp ,
δ 2/p
since
p 2−p
+ = 1.
2 2
Thus,
1 1
|u+ |L∞ (BθR ) ≤ |u+ |L∞ (BR ) + C |u+
| p
L (BR ) + |f |L (B1 ) .
q
2 ((1 − θ)R)n/p
Now let’s define f (t) = |f |L∞ (Bt ) . What we have shown is that
1 C
f (r) ≤ f (R) + |u+ |Lp (B1 ) + C|f |Lq (B1 ) .
2 (R − r)n/p
We will now show that we can in fact remove the first term on the right side.
59
Lemma 7.1. Let f ≥ 0 be bounded in [τ0 , τ1 ] with τ0 ≥ 0. Suppose for any τ0 < t < s ≤ τ1 we
have:
A
f (t) ≤ θf (s) + + B,
(s − t)α
A
f (t) ≤ c(α, θ) + B .
(s − t)α
Proof. Exercise.
7.2 From L∞ to C α
Definition 7.2. The function u ∈ H 1 (B1 ) is called a subsolution (supersolution) of the equation
Lu = 0
We now observe the following nice Lemma (that we saw before in the case of the square
function)
0,1
Lemma 7.3. Let Φ ∈ Cloc (R) be convex. Then,
60
2. If u is a super solution and Φ0 ≤ 0 then v = Φ(u) is a subsolution provided v ∈ Hloc
1 (B ).
1
Remark 7.4. In both cases, v is a subsolution. It follows that for any k ∈ Z we have that
2 (R). Then,
Proof. Assume Φ ∈ Cloc
Z Z Z
0 0
aij Di vDj φ = aij Φ (u)Di uDj φ = aij Di uDj (Φ (u)φ) − aij Di uDj uφΦ00 (u) ≤ 0,
B1 B1
since Φ0 (u)φ ≥ 0 and where we used ellipticity on the second term. The general case (Φ ∈ C 1 )
follows by approximation.
Lemma 7.5. For any > 0, there exists C = C(, n) so that for any u ∈ H 1 with
we have
Z Z
u2 ≤ C(, n) |Du|2 .
B1 B1
Proof. Suppose not. Then, there exists > 0 and a sequence un ∈ H 1 with
while
|Dun |L2 (B1 )
→ 0.
|un |L2 (B1 )
61
We may thus normalize un so that we have
|un − c|L2 → 0.
Remark 7.6. Find a direct way to prove this result that gives us the dependence of C on and
n.
Next, we establish the following Lemma, which has the flavor of a Harnack inequality.
infB1/2 u ≥ C.
Proof. By adding a constant to u, we may assume that u ≥ δ > 0. Then we will send δ → 0.
62
L2 → L∞ bound we proved,
sup v ≤ C|Dv|L2 .
B1/2
We will now show that the right hand side is bounded (independent of δ). Choose as a test
ζ2
function u where ζ ∈ C01 (B2 ). Then we have:
ζ2
Z Z Z
aij Di uDj u ζaij Di uDj ζ
0≤ aij Di uDj =− ζ2 +2 .
u u2 u
Z Z
2
|Dv| = |D log u|2 ≤ C.
B1 B1
Λ
Here, C only depends on λ and n. It now follows that
sup v ≤ C.
B1/2
u ≥ exp(−C).
Finally, we will establish the Oscillation Theorem. First let us give a definition.
63
Definition 7.8. Fix u ∈ L∞ (Ω). Define the oscillation of u on Ω by:
Remark 7.9. These are both defined in the “essential” supremum and infimum sense.
oscB1/2 u ≤ γ oscB1 u.
Proof. Let
Observe that
1 u − β1 1
u ≥ (α1 + β1 ) ⇐⇒ ≥ ,
2 α1 − β 1 2
1 α1 − u 1
u ≤ (α1 + β1 ) ⇐⇒ ≥ .
2 α1 − β 1 2
α1 −β1
This might be confusing, but note that we could add or subtract 1 = α1 −β1 and it is clear. Now,
we know that for half of B1 one of the above two inequalities on the left must hold.
Case 1:
n 2(u − β1 ) o 1
x ∈ B1 : ≥ 1 ≥ |B1 |.
α1 − β 1 2
u − β1 1
inf ≥ .
B1/2 α1 − β1 C
64
It follows that
1
β2 ≥ β1 + (α1 − β1 ).
C
α2 ≤ α1 .
Thus we have:
1
(α2 − β2 ) ≤ (1 − )(α1 − β1 ).
C
In Case 2, we have
n 2(α1 − u) o 1
x ∈ B1 : ≥ 1 ≥ |B1 |.
α1 − β 1 2
|u(x) − u(y)| Λ
sup |u| + sup α
≤ c(n, )|u|L2 (B1 ) .
B1/2 x6=y in B1/2 |x − y| λ
We now show how the above Theorem of De Giorgi implies a positive answer to Hilbert’s 19th
problem on the regularity of minimizers of smooth convex functionals. Again, let F be a convex
65
functional. Then, any local minimizer satisfies
Z
∇F (∇u) · ∇φ = 0.
Now let’s assume that φ is compactly supported in B1 . Then, if h is very small, we have:
Z
∇F (∇u(· + h)) · ∇φ = 0.
It follows that
Z
[∇F (∇u(· + h)) − ∇F (∇u(·))] · ∇φ = 0,
Z Z 1
D2 F (t(∇u(· + h)) + (1 − t)∇u(·))(∇u(· + h) − ∇u(·)) · ∇φ = 0.
0
R1
Now define à = 0 D2 F (t(∇u(· + h)) + (1 − t)∇u(·))dt. Then we see:
∇u(· + h) − ∇u(·)
Z
Ã( ) · ∇φ = 0.
h
Thus, the difference quotient is a weak solution for each h small (say we now restrict to B3/4 ).
It follows now from the De Giorgi Theorem that for each h small,
Since this is true for all h and C is independent of h, it follows in a not too difficult way that
From here, using the Schauder theory, it is easy to show that u is in fact C ∞ if F is.
66
7.4 Exercises
2. Find the dependence of c on and n in the Poincaré Sobolev inequality given in Lemma
3. Show how the Oscillation Theorem implies the local C α regularity in De Giorgi’s Theorem.
4. Take a look at the paper of James Serrin “Pathological solutions of elliptic differential
equations.” Explain his examples and show how they relate to the De Giorgi theorem.
5. Let T be a compact self adjoint operator on a Hilbert space H and fix λ 6= 0. Show that
(λI − T )u = f,
for f ∈ H in light of this. Remember to consider carefully the case when the operator is
not injective. It is helpful to show first that the kernel must be finite dimensional.
using the notion of weak derivative, from integration by parts. For simplicity we just
define it formally in one dimension (in higher dimensions you just do the same for each
index). If f ∈ Lp , we say that g is the weak derivative of f if for all φ ∈ C01 we have
Z Z
gφ = − vφ0 .
In this case we write g = Df . It is not difficult to show that when f ∈ Lp ∩ C 1 , its weak
derivative is the same as its usual derivative. We let W k,p be the space of all functions
f ∈ Lp with Dα f ∈ Lp for all |α| ≤ k. Obviously Cc∞ ⊂ W k,p ⊂ Lp , where the inclusion
67
is as sets. We may endow W k,p with the W k,p norm:
X
|f |W k,p = |Dα f |Lp .
|α|≤k
It is a standard result that the set W k,p endowed with this norm is actually a Banach
space for any k ∈ N and any 1 ≤ p ≤ ∞ and that this space is separable when p < ∞, the
class Cc∞ being dense in those cases. There are various embedding Theorems that follow
from the Sobolev and Morrey Inequalities that we will not discuss now. The goal of this
problem as well as the corresponding one on the next homework assignment will be to
then |D2 w|Lp ≤ Cp |f |Lp , when 1 < p < ∞. In a previous assignment we already saw the
case p = 2. This was just an integration by parts argument. In the next assignment we
will show it for p = 1 in a weak sense. Then, an interpolation argument will allow us to
conclude the result for all other values of p in the range. Read Sections 9.3 and 9.4
of Gilbarg and Trudinger’s book. A part of the next assignment will be devoted to
We now move to an explanation of the book of Cafarelli and Cabré. The general framework is
L(x) = l0 + l · x,
68
be written as:
1
P (x) = L(x) + (Ax, x),
2
M 2
P (x) = L(x) ± |x| .
2
Definition 8.3. Given two continuous functions u, v : A → R and x0 ∈ A, we say that v touches
u(x) ≤ v(x), ∀x ∈ A,
u(x0 ) = v(x0 ).
Next,
Definition 8.4.
69
Definition 8.6. For u : Ω → R and h ∈ Rn
for x ∈ Ω with x ± h ∈ Ω.
Remark 8.7. •
u(x0 +h)−u(x0 ) u(x0 )−u(x0 −h)
|h| − |h|
∆2h u(x0 ) =
|h|
• If L is affine, ∆2h L = 0.
1
∆2h P (x0 ) = (Ah, h),
|h|2
∆2h (P ) = M.
• If u is twice differentiable at x0 ,
Lemma 8.8. Assume that a paraboloid of opening M touches u from above. Then,
∆2h u(x0 ) ≤ M.
70
Proof. Since ∆2h L ≡ 0 for any affine L, we may assume that
1
u(x) ≤ M |x − x0 |2
2
−Θ(u, B|h| (x0 ))(x0 ) ≤ ∆2h u(x0 ) ≤ Θ̄(u, B|h| (x0 ))(x0 ).
In particular,
We have already seen above that control of Θ thus gives pointwise control of ∂vv u, for any
fixed vector v ∈ Rn . Now we want to see how control of Θ in Lp actually gives Lp control of
D2 u.
Proposition 8.10. Let 1 < p ≤ ∞ and u ∈ C(Ω). Let > 0 and define
Z
|f |Lp = sup f g.
|g| 0 =1
Lp
71
Thus1 , it suffices to show that
Z
| u∂ij φ| ≤ 2|Θ(u, )|Lp |φ|Lp0 ,
Ω
for all φ ∈ Cc∞ (Ω). Let us start by observing (as we noted above) that if v ∈ Rn is a fixed unit
vector.
Z Z Z
u∂vv φ = lim u∆2δv φ = lim φ∆2δv u ≤ |Θ(u, )|Lp |φ|Lp0 ,
δ→0 δ→0
by Corollary 8.9. Note that in the second equality we used change of variables and the compact
support of φ. Next, let us see that mixed derivatives ∂ij can be bounded by a sum of derivatives
1
∂ij = (2∂vv − ∂ii − ∂jj ),
2
where
ei + ej
v= √ .
2
Remark 8.11. The reason for the 2 on the right hand side of the inequality in the proposition
is just how we wrote ∂ij . My impression is that this is not really necessary.
bounds on Θ.
Proposition 8.12. Let u ∈ C(Ω) and assume that B ⊂⊂ Ω is convex. For sufficiently small,
1
Note that for this estimate, we are just going to assume that u ∈ C 2 (Ω̄)). The result is in fact true, but we
should know a little more about Sobolev spaces to jump to the next point. Another way to think of this is just
as the definition of the norm of |D2 u|Lp .
72
define
for x ∈ B̄. Assume that Θ(x, ) ≤ K for all x ∈ B̄. Then, u ∈ C 1,1 (¯). Moreover,
Now we just take any two points x, y and use the mean value theorem for each component of
Du(x) − Du(y).
9 Viscosity Solutions
Our goal now is to give the definition of viscosity subsolutions and supersolutions for fully non-
linear elliptic PDE, which can be viewed as “weak” solutions. Normally, when we define weak
solutions, we relax the notion of solution by using test functions. This is a property that strong
solutions have that allows us to broaden the notion of solutions to all objects enjoying that
property. This, along with the statement that weak + smooth implies strong, is what allows us
to say that the notion of weak solutions is a reasonable one (at least at first glance). Viscosity
solutions are a different type of relaxation where the property that we extend is based on the
comparison principle, rather than integration by parts. This has many uses. One use is that this
can be used to define solutions to fully nonlinear problems and non-divergence form equations.
A second one is that viscosity solutions do not have to be H 1 from the beginning but can merely
be continuous.
73
9.1 Elliptic Equations and the Definition of Viscosity Solutions
matrices.
Definition 9.1. F is said to be uniformly elliptic if there are two positive constants λ ≤ Λ such
Remark 9.2. Note that in the above, by the properties of symmetric matrices, |N | is just the
Example 9.3. If F (M, x) = tr(M ), this corresponds to the Laplacian. The Laplacian is elliptic
|N | ≤ tr(N ) ≤ n|N |,
for all N ≥ 0.
N = N + − N −,
position is not unique since there could be many eigenvectors with zero eigenvalue. From the
definition, we have the following Lemma (but these do not affect the sizes of |N − | and |N + |.
74
Lemma 9.4. F is uniformly elliptic if and only if
for all M, N ∈ S, x ∈ Ω.
Remark 9.5. It follows that any uniformly elliptic F is monotone increasing and Lipschitz
continuous in M .
F (M + N ) = F (M − N − + N + ).
Hence,
λ|N + | ≤ F (M + N ) − F (M − N − ) ≤ Λ|N + |
Moreover,
λ|N − | ≤ F (M ) − F (M − N − ) ≤ Λ|N − |
Thus,
−Λ|N − | ≤ F (M − N − ) − F (M ) ≤ −λ|N − |.
Keeping the right inequality yields he result. Now let us assume the conclusion is true. if
N = N + , we have the upper bound in the definition of uniform ellipticity for free. Now if we
75
(9.1):
F (D2 u(x), x) = f
in Ω, when the following condition holds: if x0 ∈ Ω and φ ∈ C 2 (Ω) is such that u − φ has a local
maximum at x0 , then
In other words, if φ touches u at x0 from above, then the inequality holds. If u is a supersolution
and if φ touches u at x0 from below, then the opposite inequality holds. A viscosity solution of
3. Same as the previous point with φ ∈ C 2 replaced by φ is a paraboloid (same condition and
same conclusion).
Proof. (1) =⇒ (2) =⇒ (3), by the previous definition. To show that (3) =⇒ (1), assume
that φ ∈ C 2 (Ω) is such that u − φ has a local maximum at x0 . Then, for > 0, define
1
P (x) = u(x0 ) + Dφ(x0 ) · (x − x0 ) + (D2 φ(x0 )(x − x0 ), (x − x0 )) + |x − x0 |2 .
2 2
76
Thus, since we have (3), it follows that
for all > 0. Recall F is Lipschitz continuous in M by the definition of uniform ellipticity. In
Proof. Exercise.
Next, we have that viscosity subsolutions that are also C 2 are true subsolutions.
Lemma 9.9 (Weak =⇒ Strong). Assume that u ∈ C 2 (Ω). Then, u is a viscosity subsolution of
Proof. First, in the forward case, we can take φ = u and we get the conclusion. In the reverse
D2 (u − φ)(x0 ) ≤ 0.
77
Thus,
D2 u(x0 ) ≤ D2 φ(x0 ).
Since F is increasing in its first variable, if F (D2 u(x0 ), x0 ) ≥ f (x0 ) (u is a subsolution in the
Proposition 9.10. Assume that u and v are viscosity subsolutions of (9.1) in Ω. Then, max u, v
Proof. Any φ touching max u, v from above at a point x0 ∈ Ω touches either u or v from above
at x0 . Since both u and v are viscosity subsolutions, we must have F (D2 φ(x0 ), x0 ) ≥ f (x0 ).
Proposition 9.11. Let {Fk }k≥1 be a sequence of continuous uniformly elliptic operators with
Then,
F (D2 u, x) ≥ f (x),
Proof. Let P be a paraboloid that touches u from above at x0 . Take a small r > 0 so that
78
Note that P is fixed, so the first term is arbitrarily small for k large. Now we just want to
estimate from below Fk (D2 P (x0 ), x0 ). Now, we have that P ≥ u on Br (x0 ). Thus, for > 0,
1
Qk, (x) = P (x) + |x − x0 |2 + uk (x0 ) − u(x0 ).
2
Observe that
D2 Qk, = D2 P + I.
It follows from continuity that there exists xk ∈ Br (x0 ) so that Qk, touches uk from above at
xk . Thus,
9.2 The Class of solutions of all uniformly Elliptic Equations and the Pucci
Operators
X X
M− (M, λ, Λ) = M− (M ) = λ ei + Λ ei ,
ei >0 ei <0
X X
M+ (M, λ, Λ) = M+ (M ) = Λ ei + λ ei ,
ei >0 ei <0
79
Let A be a symmetric matrix such that
A ∈ Aλ,Λ .
X X
M− (M, λ, Λ) = λ ei (M ) + Λ ei (M ) = min tr(AM ).
A∈AλΛ
ei (M )>0 ei (M )<0
The minimum is achieved by an A which is diagonalizable in the same basis as M and has
eigenvalue λ when M has a positive eigenvalue and Λ when it has a negative one. Similarly,
X X
M+ (M, λ, Λ) = λ ei (M ) + Λ ei (M ) = max tr(AM ).
A∈AλΛ
ei (M )<0 ei (M )>0
Both equalities are seen by noting that the trace of a symmetric matrix is the sum of its
1. M− (M ) ≤ M+ (M ).
5. M+ (M ) + M− (N ) ≤ M+ (M + N ) ≤ M+ (M ) + M+ (N ).
6. M− (M ) + M− (N ) ≤ M− (M + N ) ≤ M− (M ) + M+ (N ).
80
7. N ≥ 0 =⇒ λ|N | ≤ M− (N ) ≤ M+ (N ) ≤ nΛ|N |.
the viscosity sense in Ω. Similarly S̄(λ, Λ, f ) are the set of viscosity subsolutions u satisfying
Lemma 9.15. Assume that u ∈ S(λ, Λ, f ) ∩ C 2 (Ω). Then, for each x ∈ Ω, there is a uniformly
Remark 9.16. A consequence of this fact is that once we prove some fact about all elements
of the class S (which one could view as essentially a “linear space”), they will automatically be
like that.
81
Now we give a few basic properties of S, S̄, S, S ∗ (when their arguments are clear from
S̄, S, S ∗ . This follows from the second property of the Pucci operators above.
2. u ∈ S(λ, Λ, f ) =⇒ −u ∈ S̄(λ, Λ, −f ). This follows from the third property of the Pucci
operators.
α ·
v ∈ S(λ, Λ, f ( )).
r2 r
This is a fundamental scaling law and it follows from the definition of the Pucci operators.
u − φ ∈ S(λ, Λ, f − g).
This is essentially a linearity property and it is what we alluded to in the previous remark.
Proof. We have already mentioned the proofs of (1)-(3) in the statement. To prove (4), let
However, we know that M+ is subadditive from the fifth property of the Pucci operators.
Therefore,
by the assumption on φ.
Remark 9.18. Since M− is super-additive, a similar result can be established for supersolu-
tions.
82
Combining some of our previous results, we get the following.
• u ∈ S(0) =⇒ u+ ∈ S(0).
• S(f ), S̄(f ), S(f ) are all closed under uniform limits on compact sets.
Proposition 9.20. Let u satisfy F (D2 u, x) ≥ f (x) in the viscosity sense. Then,
λ
u ∈ S( , Λ, f (x) − F (0, x)).
n
λ
u − φ ∈ S( , Λ, f (x) − F (D2 φ(x), x)).
n
Proof. We prove the second statement since the first is just the case φ = 0. Suppose that ψ
X λ X
≤ F (D2 φ(x0 ), x0 ) + Λ ei (ψ) + ei (ψ)
n
ei (ψ)>0 ei (ψ)<0
λ
= F (D2 φ(x0 ), x0 ) + M+ (D2 φ(x0 ), , Λ).
n
Therefore,
λ
M+ (D2 φ(x0 ), , Λ) ≥ f (x0 ) − F (D2 φ(x0 ), x0 )
n
83
10 Alexandroff Maximum Principle and Applications
We start by establishing the Alexandroff maximum principle for classical solutions and then
prove the viscosity case. For this chapter we will mainly follow the exposition of the book of
L = aij (x)Dij ,
with aij ∈ C(Ω). As usual, we assume that the matrix A is positive definite. Define
D∗ := det(A)1/n .
Note that D∗ is just the geometric mean of the eigenvalues of A. Note that since the arithmetic
1
tr(A) ≥ D∗ .
n
Let λ, Λ denote the largest and smallest eigenvalues of A respectively. If A is strictly positive
definite we have
0 < λ ≤ D∗ ≤ Λ.
Γ+
u = {y ∈ Ω : u(x) ≤ u(y) + Du(y) · (x − y), for any x ∈ Ω}.
84
The set Γ+ is called the upper contact set of u and
D2 u ≤ 0 on Γ+ .
Note that the upper contact set can be defined for continuous functions as:
Γ+ n
u = {y ∈ Ω : u(x) ≤ u(y) + p(y) · (x − y), for any x ∈ Ω and some y ∈ R }.
u is concave if and only if Γ+ = Ω. If u ∈ C 1 , then p(y) = Du(y). We can now state the
f
∈ Ln (Ω).
D∗
Then,
f−
sup u ≤ sup u+ + C| | n +,
Ω ∂Ω D∗ L (Γ )
u(x) > 0}. As we noted before, D2 u ≤ 0 on Γ+ . Thus, if we consider the function u = u − |x|2 ,
D2 u < −2Id is negative semi-definite on Γ+ . This implies that ∇u : Γ+ ∩Ω+ → ∇u (Γ+ ∩Ω+ )
Z Z
1= |det(D2 u − 2Id)|.
∇u (Γ+ ∩Ω+ ) Γ+ ∩Ω+
Z Z
+ + 2
aij Dij u n f n
|∇u(Γ ∩ Ω )| ≤ −det(D u) ≤ − ≤| | n +,
Γ+ ∩Ω+ nD∗ nD∗ L (Γ )
1
since n multiplied by the trace always controls the n-th root of the determinant.
85
Now we wish to show that
BM (0) ⊂ ∇u(Γ+ ∩ Ω+ ) ⊂ Rn ,
where
supΩ u − inf Ω u
M= .
diam(Ω)
Assume without loss of generality that u attains its maximum m > 0 at 0 ∈ Ω. Consider
L(x) = m + a · x,
m
where |a| < diam(Ω) . By definition, L(x) > 0 on Ω and L(0) = m. Since u is maximal at 0 we
have that ∇u(0) = 0. Thus, there exists x1 close to 0 with u(x1 ) > L(x1 ) > 0, while u ≤ 0 < L
on ∂Ω. Therefore, there exists x̃ so that ∇u(x̃) = ∇L(x̃). Now we translate the plane up until
we have the highest such position. This point of tangency will be in the contact set. This means
Remark 10.1. To see the last step, it is best to imagine we have a hyperplane with small
“slope” coming down from infinity. The first point it touches has to be inside Γ+ ∩ Ω+ and the
Remark 10.2. The above proof works almost identically when we had a lower order term c(x)u
Theorem 21. Suppose u ∈ C(Ω̄) ∩ C 2 (Ω) satisfies aij ∂ij u + cu ≥ 0 in Ω with u ≤ 0 on ∂Ω.
Assume diam(Ω) ≤ d. Then, there is a constant δ = δ(n, λ, Λ, d, |c+ |L∞ ) > 0 so that if |Ω| < δ
then u ≤ 0 in Ω.
86
Proof. We have
1
sup u+ ≤ C(λ, Λ)|c+ u+ |Ln (Ω) ≤ |c+ |L∞ c(n, λ, d)|Ω|1/n sup u+ ≤ sup u+ .
Ω Ω 2 Ω
We now move to the ABP estimate for viscosity solutions, following the book of Caffarelli and
Cabré. Define:
Theorem 22. Suppose u ∈ S̄(λ, Λ, f ) in B1 with u ≥ 0 on ∂B1 for some f ∈ C(B̄1 ). Then we
have
Z 1/n
sup u− ≤ c(n, λ, Λ) (f + )n ,
B1 B1 ∩{u=Γu }
f (x0 ) ≥ 0
and
1
L(x) ≤ Γu (x) ≤ L(x) + C(f (x0 ) + (x))|x − x0 |2 ,
2
where L is an affine function and x is close to x0 and (x) → 0 as x → x0 . The second inequality
will be established by observing that if a parabola touches Γu from below at a contact point,
then it touches u from below. The key point is that since f is bounded, the opening of the
87
paraboloids touching from below has to be bounded. For example, Γu could not have a corner
singularity at a contact point if u ∈ S + . Note that the convex envelope need only be Lipschitz
in general.
With some work (see below), the above will show that Γu is C 1,1 on the contact set. Off
the contact set, Γu should be affine and uniformly smooth up to the contact set. We will then
be able to conclude that Γu ∈ C 1,1 . Then, we will apply the classical Alexandroff Maximum
Principle we proved above to Γu and it will imply the estimate in the statement of the theorem.
What we mentioned will be established rigorously via series of Lemmas that we now give.
Lemma 10.3. Let u ∈ S̄(λ, Λ, f ) in Bδ = Bδ (0). Assume that f ∈ L∞ (Bδ ) and that φ is convex
on Bδ and that
in Bδ . Then,
Remark 10.4. The purpose of this is to establish the two inequalities we mentioned above.
Note that here we are proving the second inequality we mentioned in the sketch above in the
case where L = 0 (which one could assume WLOG by subtracting from u an affine function).
φ(x0 ) = Cr r2 .
Without loss of generality, assume that x0 = (0, 0, 0, ..., r). Therefore, by a simple argument
88
done first for smooth convex functions, we see that
Our goal is to now to bound C̄ in terms of f . As alluded to in the sketch above, we will show
Cr
h̃ = h.
4
h̃ ≤ φ ≤ u on ∂Rr .
Note that we showed that φ ≥ r2 Cr on the whole of the upper boundary xn = r. Moreover,
observe that
Cr
h̃(0) = > φ(0) = u(0) = 0.
4r2
89
then see that since the eigenvalues of D2 h̃ are just given by Cr /2, − 2C
N2
r
, ..., − 2C
N2
r
, we have
that
Cr Cr
λ − 2Λ(n − 1) 2 ≤ |f |L∞ (Rr ) .
2 N
Therefore, if we choose
8Λ(n − 1)
N2 = ,
λ
we see that
2
Cr ≤ |f | ∞ .
λ L (Rr )
1
This concludes the proof by taking ν = N.
Lemma 10.5. Assume Γ ∈ C(B1 ) and that there exists a closed set A on which Γ can be
approximated in C 1,1 by paraboloids. Assume also that for the open set B = B1 \ A we have that
to each x ∈ B there is a unique connected maximal open set Bx on which Γ is affine with the
Proof. Let x, y ∈ B1 . We want to compute the second order difference quotient with x and
y. Let us denote it by Dx,y (Γ). If y ∈ Bx , then Dx,y (Γ) = 0. If not, then there is a x∗ ∈ A
done.
Finally, we observe that because of Lemma (10.3), Γu satisfies the assumptions of the pre-
ceding Lemma with the closed set A = {x : Γu = u}. Note that if x ∈ Ac , we must have that Γ
estimate.
90
11 Harnack Inequality and Oscillation Lemma
We now move to establish the Harnack inequality for members of S ∗ . As we know, one of the
important consequences of the Harnack inequality is the oscillation lemma, which we use to
establish Hölder regularity. We will start by introducing two tools. The first is the construction
of a “barrier function” and the second is a decomposition of cubes similar to the Calderón-
Zygmund decomposition. Toward this end, we give some notation. Ql (x0 ) is the cube of side
Lemma 11.1. Given constants 0 < λ ≤ Λ, there exists φ ∈ C ∞ (Rn ) and universal constants
φ ≥ 0 in Rn \ B2√n (11.1)
φ ≤ −2 in Q3 (11.2)
This is a function which looks like a bowl with a dip in the center and a rather slow conver-
c :
Proof. We search for a radial φ which in fact takes the following form in B1/4
φ(x) = M1 − M2 |x|−α ,
91
Outside of B2√n , φ > 0 and is uniformly smooth. We can take φ to be radial and connecting
√ √
0 and −2 for the radial variable r ∈ [ 23 n, 2 n]. It does not really matter how we connect
φ. The key thing we need to check now is that φ is superharmonic outside B1/4 in the sense
1
that M+ (D2 φ, λ, Λ) ≤ 0 in |x| ≥ 4. Let us now compute the eigenvalues of D2 φ in B1/4
c .
Note that they must be a function of |x|. Thus, it suffices to compute the eigenvalues at
(r, ..., 0, 0). At such a point, it is not difficult to check that D2 φ is diagonal and the eigenvalues
and
Z
α ≤ − f < 2n α.
Qj
Remark 11.3. The proof is elementary. The use of this result is that we can decompose Q1
into two sets, a good set on which f is uniformly bounded and a “bad” set on which f may not
Proof. We just cut Q1 into 2n dyadic cubes. We keep a cube Q at any generation if −Q f ≥ α.
R
It must of course be that −Q f ≤ 2n α. We keep dissecting the ones we don’t keep. The result
R
Next, we give a slight modification. If Q is a daydic cube, we say that Q̃ is its predecessor
92
Lemma 11.4. Let A ⊂ B ⊂ Q1 be measurable sets and 0 < δ < 1 such that
• |A| ≤ δ, and
A ⊂ ∪∞
j=1 Qj ,
|A ∩ Qj | |A ∩ Q̃j |
δ< ≤ 2n δ < δ,
|Qj | |Q̃j |
for all j. Relabel the Q̃j so that they are pairwise disjoint. Using the second assumption in the
Q̃j ⊂ B.
Thus,
X X
|A| = |A ∩ Q̃j | ≤ δ |Q̃j | ≤ δ|B|.
We now proceed to establish the Harnack inequality, which is the content of the following
Theorem.
93
In fact, it is not difficult to check that, by a scaling and covering argument it suffices to
Lemma 11.5. There exist universal constants 0 , C > 0 so that if u ∈ S ∗ (λ, Λ, f ) in Q4√n and
sup u ≤ C.
Q1/4
Definition 11.6. Assume that g : Ω → R is measurable. Then, λg (t) = |{x ∈ Ω : u(x) > t}|, is
Remark 11.7. Note that λg is a decreasing function of t and that rate of decay of λg as t grows
essentially determines what Lp λg belongs to. sup g ≤ C is equivalent to λg (t) ≡ 0 for t > C.
Our first task will be to show that λu (t) ≤ Ct− where C, are universal constants. For
this, we will use the Barrier argument, the cube decomposition lemma, and the ABP estimate.
A key feature of the ABP estimate we will use is that we integrate over the contact set on the
right hand side of the ABP estimate. The upper bound on the distribution function λu already
implies that
for some small q and C that are universal. After showing this, we will show the same decay of
λu at “all scales,” by which we just mean that we can rescale u and the domain and get the
We will first show a base Lemma and then iterate it along with the cube decomposition Lemma
94
Lemma 11.8. There exist universal constants 0 > 0, 0 < µ < 1 and M > 1 so that if u ∈ S̄(|f |)
u ≥ 0 in Q4√n
inf u ≤ 1
Q3
|f |Ln (Q4√n ) ≤ 0 ,
|{u ≤ M } ∩ Q1 | > µ,
where M = −2.
Remark 11.9. We will apply the ABP estimate to w = u + φ where φ is the barrier function
of Lemma (11.1).
Proof. Define w = u + φ. Note that B2√n ⊂ Q4√n and that M+ (D2 φ) ≤ Cξ (pointless). Thus,
we have that w ∈ S̄(|f | + Cξ) in B2√n . Now, w ≥ 0 on ∂B2√n since φ = 0 there and u ≥ 0. We
also have that inf Q3 (w) ≤ −1 since φ ≤ −2 there. Thus, applying the ABP estimate to w, we
see that:
1
≤ |{w = Γw } ∩ Q1 |.
(2C)n
Observe, however, that since w = u + φ, when w(x) = Γw (x) we have w(x) ≤ 0 (recall that Γw
1
≤ |{u ≤ M } ∩ Q1 |.
(2C)n
95
Now we establish the bounds on the distribution function using the cube decomposition
lemma.
Proof. Observe that we have the bound for k = 1. Suppose that the bound holds for k − 1 and
define
|A| ≤ (1 − µ)|B|.
The key is to use the cube lemma. The first assumptions of the lemma are satisfied. Indeed,
we have A ⊂ B ⊂ Q1 and |A| ≤ (1 − µ). We must now show that if Q is a dyadic cube,
|A ∩ Q| > (1 − µ)|Q| =⇒ Q̃ ⊂ B.
Let us suppose that Q̃ 6⊂ B. That is, take x̃ ∈ Q̃ such that u(x̃) ≤ M k−1 . Then, let us transform:
1
x = x0 + y,
2i
1
ũ(y) = u(x).
M k−1
96
We claim that ũ satisfies the assumptions of the previous lemma. If so, then:
|Q \ A|
µ < |{ũ ≤ M } ∩ Q1 | = 2in |{u(x) ≤ M k } ∩ Q| = .
|Q|
Thus, |Q \ A| > µ|Q|, which contradicts our assumption about Q. To complete the proof, we
just need to show that ũ satisfies the assumptions of the previous Lemma. Observe that by a
Moreover, it is easy to see that |f˜|Ln (Q4√n ≤ |f |Ln (Q4√n ) ≤ 0 . Now we are done.
Note that to establish the above bounds, we only used that u ∈ S̄. We will now observe that
since u ∈ S ∗ (f ), we also have that −u ∈ S̄(|f |). This means that the above considerations also
Lemma 11.11. Let u ∈ S(−|f |) in Q4√n . Assume that |f |Ln ≤ 0 and that
|{u ≥ t} ∩ Q1 | ≤ dt− ,
1
|x0 |∞ ≤
4
and
u(x0 ) ≥ ν j−1 M0 ,
97
then
−/n −j/n
where lj = σM0 ν .
Remark 11.12. What we are saying now is that if there is an x0 with u(x0 ) ≥ M0 , then there is
x1 so that u(x1 ) ≥ νM0 , and an x2 so that u(x2 ) ≥ ν 2 M0 and so on. Furthermore, these points
get closer and closer to each other. We will later show that there is then a sequence of points
making u unbounded, which will contradict the continuity of u. It will follow that u < M0 ,
1 n √
σ > d2 (4 n)n ,
2
−/n 1
σM0 + dM0− ≤ .
2
It follows that Qlj (x0 ) ⊂ Q1 . Now suppose that supQj u < ν j M0 . We will derive a contradiction
M −
0
|{u ≥ ν j M0 /2} ∩ Qlj /4√n (x0 )| ≤ |{u ≥ ν j M0 /2} ∩ Q1 | ≤ dν −j .
2
small. This will be the contradiction. Toward this end, consider the transformation
lj
x = x0 + √ y, y ∈ Q4√n , x ∈ Qj = Qlj (x0 )
4 n
Note that v > 0 on Q4√n by assumption. Let us claim that v satisfies the assumptions of the
98
first Lemma 11.8. We will show this is the case later. Thus we have that
M0 l n
j
|{u(x) < ν j } ∩ Qlj /4√n (x0 )| ≤ √ dM0− .
2 4 n
Thus using the bound on the set of {u ≥ ν j M0 /2} ∩ Qlj /4√n (x0 ), we have that
l n l j n M −
j 0
√ ≤ √ dM0− + dν −j .
4 n 4 n 2
This is a contradiction. To complete the proof, we just have to show that v satisfies the hy-
v ∈ S̄(f˜),
l l
where f˜(y) = ( 4√j n )2 (ν j−1 (ν − 1)M0 )−1 |f (x)|, where x = x0 + √j y.
4 n
Since u(x0 ) ≥ ν j−1 M0 ,
inf Q3 v ≤ 1. Moreover,
l 2 1 lj l 1
j j
|f˜|Ln (Q4√n ) = √ |f (x 0 + √ y)| n √
L (Q4 n ) = √ |f | n
j−1
4 n ν (ν − 1)M0 4 n 4 n ν (ν − 1)M0 L (Qlj (x0 ))
j−1
≤ 0 ,
ν
where we used that (ν − 1)M0 = 2 and ν > 1.
We can now finish the proof of the Harnack inequality by proving 11.5. Note that if M0 is
99
sufficiently large, we have that
∞
X 1
lj ≤ .
4
j=1
Note that taking M0 larger does only strengthens the results of the preceding Lemma. We will
1
now show that supQ1/4 u ≤ M0 . Suppose not. Then, there exists |x1 |∞ ≤ 8 so that u(x1 ) ≥ M0 .
Now the arguments we gave at the end of the proof of the De Giorgi Theorem allow us to
In the first part of this section, we will establish the following important Theorem.
supersolution. Then,
λ
u − v ∈ S( , Λ)
n
in Ω.
The result follows directly from the preceding theorem and the maximum principle for vis-
cosity solutions, which relies on the Alexandroff Maximum Principe. The proof of the theorem
100
will rely on approximate solutions introduced by Jensen.
Definition 12.3. Let u ∈ C(Ω) and assume that H ⊂⊂ Ω is an open set. For > 0 we defined
1
u (x0 ) = sup u(x) + − |x − x0 |2 ,
x∈H̄
for x0 ∈ H.
Observe that u (x0 ) ≥ u(x0 ) + for all x0 ∈ H. We will show the following Theorem.
Example 12.4. It is instructive to consider the following two examples. First, when u1 (x) = |x|,
5
u1 (x0 ) = |x0 | + ,
4
since it is easy to check that the maximum occurs at x0 + 2 when x0 ≥ 0 and x0 − 2 when
x0 < 0. Notice that u1 is an example of a corner that is opening up, so it is already “regular”
from below. Let us now consider the function u2 (x) = −|x|. The situation is now a bit more
5
u2 (x0 ) = −|x0 | + , for |x0 | >
4 2
In the middle region there is a change and the optimal x is actually x = 0. In that case, we see
that
1
u2 (x0 ) = − x0 2 , for |x0 | ≤ .
2
101
Theorem 25 (Aleksandrov). If f is a convex function on Ω ⊂ Rd , then f is twice differentiable
almost everywhere.
Remark 12.5. A proof of this result can be found in a paper of Crandall, Ishii, and Lions on
Corollary 12.6. If f is C 1,1 from below, then f is twice differentiable almost everywhere.
Proof. C 1,1 from below means, in the notation of the first section on viscosity solutions, that
Θ̄(f, ) ≥ −M for some , M > 0. That is, the second order difference quotients are bounded
from below, or there are parabloids of fixed opening M touching f from below at any point
in the domain. Therefore, f + M |x|2 is convex. The result now follows from the Aleksandrov
theorem.
We will use this to say that the upper envelope u is in fact twice differentiable almost
everywhere.
2. For any x0 ∈ H, there is a concave paraboloid of opening 2/ that touches u from below
Toward establishing the theorem, let us mention a few useful properties of the upper envelope.
2. u (x0 ) ≥ u(x0 ) + .
102
3. |u (x0 ) − u (x1 )| ≤ 3 diam(H)|x0 − x1 |.
0
4. If 0 < < 0 u (x0 ) ≤ u (x0 ).
Proof of Lemma 12.7. Points 1,2,4,6 are obvious from the definition. Now let us move to estab-
1 1 1 2
u (x0 ) ≥ u(x) + − |x − x0 |2 ≥ u(x) + − |x − x1 |2 − |x0 − x1 |2 − |x0 − x1 ||x − x1 |
1 3
≥ u(x) + − |x − x1 |2 − diam(H)|x1 − x0 |
Therefore,
3
u (x0 ) ≥ u (x1 ) − diam(H)|x1 − x0 |.
1 ∗
|x − x0 |2 = u(x∗0 ) + − u (x0 ) ≤ u(x0 ∗) − u(x0 ) ≤ osc(u).
0
Proof of the Theorem 26. The Lipschitz continuity of u follows from the third property in
Lemma (12.7). The other properties imply the uniform convergence u → u. Now for fixed
x0 ∈ H, we define
1
P0 (x) = u(x∗0 ) + − |x − x∗0 | ≤ u (x),
for any x while P0 (x0 ) = u (x0 ). It follows that u is C 1,1 from below and thus almost every-
where twice differentiable by the Aleksandrov Theorem. Now we prove that u is a subsolution
103
whenever u is. Assume that F (D2 u) ≥ 0 in the viscosity sense on H. Fix x0 ∈ H1 ⊂⊂ H and
1
Q(x) = P (x + x0 − x∗0 ) + |x0 − x∗0 |2 − .
Note that Q is just a horizontal and vertical translation of P . Now, we can ensure that if
1
u (x + x0 − x∗0 ≥ u(x) + − |x0 − x∗0 |2 ,
1
u(x) ≤ P (x + x0 − x∗0 ) + |x0 − x∗0 |2 − = Q(x),
while u(x∗0 ) = Q(x∗0 ), since P (x0 ) = u (x0 ). Thus, Q touches u from above at x∗0 . Thus,
0 ≤ F (D2 Q) = F (D2 P ).
supersolution. Then,
λ
u − v ∈ S( , Λ)
n
in Ω.
104
u is the upper envelope of u and v is the lower envelope of v. This is because uniform limits
of viscosity subsolutions on compact sets are viscosity subsolutions, as we showed in the past.
Observe that w ≥ 0 on ∂Br (x0 ) and w(x0 ) < 0, since P touches u − v from above. Since
w ∈ C 1,1 almost everywhere, we can apply one step of the Alexandrov Maximum Principle to
Z
−
0 < sup w ≤ det(D2 Γw ).
Br (x0 )∩{w=Γw }
D2 w ≥ 0
for almost every x ∈ {w = Γw }. From the lower bound of the integral, we see that
|{w = Γw }| > 0,
105
≤ F (D2 v (x1 )) + Λ||(D2 P )+ || − λ||(D2 P )− || + 2Λδ
Then,
λ
uh,e,β ∈ S( , Λ).
n
Lemma 12.10. Let 0 < α < 1, 0 < β ≤ 1 and K > 0 be constants. Let u ∈ L∞ ([−1, 1] and
|uh,β |C α ≤ K,
independent of h.
Proof. Exercise.
106
Noting that uh,e,α ∈ L∞ (B1 ) by the Harnack inequality (+oscillation Lemma etc), we can
apply the above corollary finitely many times (a universal number) to deduce that uh,e,1 ∈ L∞
M +N F (M ) + F (N )
F( )≥ .
2 2
Assuming that the solution is sufficiently smooth, this is just the following computation first in
one dimension
∂ee (F (D2 u)) = ∂e (Fij (D2 u)(∂e ∂ij u)) = (Fij )(D2 u)∂ij (∂ee u) + Fij,kl (D2 u)∂ij ∂e u ∂kl ∂e u = 0.
Note that Fij is the derivative of F with respect to the ij variable of F . Now concavity implies
107
that the second quantity is non-positive. To see this in the general viscosity case, we will rely
on the following Theorem applied in a suitable way to the second order difference quotients of
u.
Theorem 28. Let F be concave and let u and v be subsolutions of F (D2 w) = 0 in Ω. Then,
1
2 (u + v) is a viscosity suboslution of F (D2 w) = 0 in Ω.
Corollary 12.13. Assume that F is concave and that F (D2 u) = 0 in the viscosity sense. If e
λ
∂ee u ∈ S( , Λ)
n
in Ω.
The first corollary follows since 12 (u(x+eh)+u(x−eh)−2u(x)) can be written as the difference
The second corollary follows from the first since the class S is closed under uniform limits. Now
we turn to the proof of the Theorem, which again uses the envelope.
Proof of the Theorem. We need to show that 21 (u +v ) is a viscosity subsolution of F (D2 w) = 0.
1
Let P be a paraboloid that touches 2 (u + v ) from above at x0 . We need to show that
1
w(x) = P (x) − (u + v ) + δ|x − x0 |2 − δr2 .
2
Applying the exact same strategy as we did before, we find an x1 where u , v , w are all second
108
order differentiable at x1 and for which
1
D2 (P + δ|x − x0 |2 − (u + v ))(x1 ) = D2 w(x1 ) ≥ 0,
2
while F (D2 u (x1 )) ≥ 0 and F (D2 v (x1 )) ≥ 0. Since F is concave, we have that
1
F (D2 (u + v )(x1 )) ≥ 0.
2
Thus,
F (D2 P + 2δI) ≥ 0.
Now we send δ → 0.
Concave equations
The goal of this section is to establish C 2,α regularity of viscosity solutions to F (D2 u) = 0 when
F (D2 u) = 0
109
Lemma 13.1. Under the hypotheses of the Theorem, there exists a universal constant 0 < δ0 < 1
Note that the assumption that the oscillation of D2 u being bounded by 2 can be easily
1
ensured by dividing u by a constant. Note that t F (t·) has the same ellipticity constants as
F (·). The main tool we will use to establish the Theorem is that we already know that
λ
∂ee u ∈ S( , Λ).
n
We first “recall” the following result that we established in the course of proving the Harnack
inequality:
λ
kM2 − M1 k ≤ k(M2 − M1 )+ k = sup ((M2 − M1 )e, e)+ .
λ+Λ |v|=1
Remark 13.4. The second equality follows from diagonalizing the matrix M2 − M1 . The first
inequality is the key and it is intuitively clear. Indeed, recall that F (·) is “like” a trace operator
and so F (M1 ) = F (M2 ) = 0 implies that the positive and negative eigenvalues of M1 and M2
0 = F (M2 ) ≤ F (M1 )+Λk(M2 −M1 )+ k−λk(M2 −M1 )− k ≤ (Λ+λ)k(M2 −M1 )+ k−λkM2 −M1 k,
110
where we used the triangle inequality in the second inequality.
Lemma 13.5 (Key Lemma). Under the hypotheses of the Theorem, assume also that
Assuming this Lemma holds, let us now prove the main oscillation lemma.
Oscillation Lemma. Since diam(D2 u(B1 )) = 2, we can cover D2 u(B1 ) by N balls of radius 0 ,
where 0 is as in the last Lemma (N is universal since 0 is). By the last Lemma, we know that
let us define
w(x) = 4u(x/2),
D2 w(B1/2 ) can be covered by N − 2 of the original balls. Thus, there exists k ≤ N so that
diam(D2 u(B1/2k )) ≤ 1.
λ c0
Proof of the Key Lemma. Let c0 = Λ+λ (as in Lemma 13.3). Take 0 ≤ 32 . By assumption, we
Mi = D2 u(xi ),
D2 u(B1 ) ⊂ ∪N
i=1 Bc0 /16 (Mi ).
111
In fact, since the radii are now fixed at c0 /16, we have that
0
D2 u(B1 ) ⊂ ∪N
i=1 Bc0 /8 (Mi ),
where N 0 ≤ N is universal. It follows that there exists one of the Mi , let us say it is M1 , for
which
where η is universal. This is just because they are finitely many and {(D2 u)−1 (Bc0 /8 (Mi ))}
1
cover B1 . Since the diameter of D2 u(B1 ) > 1 and since 0 ≤ 32 , there must exist M2 so that
|M2 − M1 | ≥ 41 . Thus, by the Ellipticity Lemma 13.3, there exists a unit vector e ∈ Rn with
c0
uee (x2 ) ≥ uee (x1 ) + .
4
Now define
Using the concavity of F , we have that 0 ≤ v ∈ S̄(λ/n, Λ, 0) in B1 . Since the measure of points
for which D2 u(x) is within c0 /8 of D2 u(x1 ) is larger than η and since uee (x2 ) ≥ uee (x1 ) + c0 /4,
we have that
Now since B2 (Mi ) cover D2 u(B1 ), we can find a single j so that
Observe now that for all x ∈ B1/2 we have that (D2 u(x)e, e) ≤ K − C1 while all matrices in
112
M ∈ B2 (Mj ) satisfy (M e, e) > K − 5. Now we take 50 ≤ C1 and it follows that B2 (Mj ) ∩
13.2 L∞ to C 2
We will now prove a stronger version of the preceding Theorem: that viscosity solutions are C 2,α
and the C 2,α norm of solutions to concave equations can be controlled just by its L∞ norm.
Remark 13.6. The existence of a classical solution, uniqueness of the viscosity solution, and
the Evans-Krylov Theorem imply the above Theorem. The proof given by Caffarelli and Cabre
of the above theorem, however, is quite soft and independently interesting (it also seems to
apply to a wider class of F than what was previously obtained using existence, uniqueness,
and Evans-Krylov). Why does the a-priori bound follow from Evans-Krylov? It is similar to a
Remark 13.7. We may assume that F (0) = 0. Indeed, for a general elliptic F , there exists
|F (0)|
a unique t ∈ R so that F (tI) = 0 while |t| ≤ λ . This just follows from the definition of
ellipticity. Now if we define the paraboloid P (x) = 2t |x − x0 |2 , we see that we can define
Let us start with a sketch of the proof. The first step is to observe that if F (0) = 0, we can
find a supporting hyperplane for F at 0. That is, we can find a linear functional L on the space
L(M ) ≥ F (M ).
113
Since L is linear, we must have L(M ) = tr(AM ). Since F is elliptic, it is easy to check that A
is positive definite with eigenvalues in [λ, Λ] by a rotation, we can assume that L(M ) = tr(M ).
It follows that any solution to F (D2 u) = 0 satisfies ∆u ≥ 0 in the viscosity sense. We will then
for all x0 ∈ B1/2 . From there we will observe that the functions u∗h (x) = 1 −
R
(
h2 ∂Bh (x0 )
u − u(x))
belong to the class S. A simple argument will show that u∗h ∈ L1 uniformly in h, from which
we will see that it is uniformly L∞ in h (Because it is an element of S). This will then allow us
to say that ∆u ∈ L∞ and eventually D2 u ∈ L∞ . From there, we will modify slightly the proof
Proof. The proof will proceed in steps. First observe that by scaling we may as well assume that
|u|L∞ ≤ 1 and so our task will be to show first that |u|C 1,1 (B1/2 ≤ C and then |u|C 2,α (B1/2 ≤ C,
L(M ) ≥ F (M ),
X
L(M ) = aij mij = tr(AM ).
i,j
114
by the definition of ellipticity. Similarly,
It follows that A is symmetric and positive definite. We can thus change coordinates so that A
is diagonal and further change coordinates (by scaling) so that A is the identity. Note that this
change of coordinates is universal since the eigenvalues of A are in [λ, Λ]. It follows that we may
assume that A = Id. Note that |u|C 2,α in this coordinate is equivalent to |u|C 2,α in the usual
coordinates.
We thus see that for any C 2 function, ∆φ ≥ F (D2 φ). Now, if u is a viscosity solution, we
have that for any φ touching u from above, ∆φ ≥ F (D2 φ) ≥ F (D2 u) = 0. Thus, u satisfies
∆u ≥ 0
1
in the viscosity sense. Now we claim that for any x0 ∈ B1/2 and 0 < h < 2 we have that
Z
u(x0 ) ≤ − u.
∂Bh (x0 )
Indeed, for fixed h, let ∆w = 0 in Bh (x0 ) and h = w on ∂Bh (x0 ). By the maximum principle
for viscosity subsolutions (from the ABP estimate), we have that u ≤ w. Thus, in particular,
Z
u(x0 ) ≤ w(x0 ) = − u,
∂Bh (x0 )
where we used the mean value property on the harmonic function w and w = u on ∂Bh (x0 ).
R
Step 3: Properties of u∗h (x) = 1
h2
−
∂Bh (x) u − u(x)
Our goal will be to show that the function u∗h (x) is uniformly in L1 .
Note that the function u∗h is continuous and nonnegative on B1/2 for any 0 < h < 1/2.
115
Observe that if φ ∈ C 2 we have that φ∗h → 1
2n ∆φ on compact sets. This can be seen easily
using a Taylor expansion of u and can just be checked for quadratic polynomials. Note also that
Note now that u∗h ≥ 0 so we can bound |u∗h |L1 as follows. Fix φ ≥ 0 be in Cc∞ (B1/2 ) with
Z Z Z Z Z Z Z
1
|u∗h | = u∗h ≤ u∗h φ = − u− φ−u(x)φ(x) = uφ∗h ≤ C(n)|D2 φ|L∞ ≤ C(n).
B1/3 B1/3 B1/2 h2 B1/2 Bh (x) Bh (x) B1
Note that u∗h is an average of u minus u. We want to see that the average of u is a subsolution
of F (D2 u) = 0. Since F is concave, if u1 , ..., uk are viscosity subsolutions then ki=1 λi ui is also
P
P
a viscosity subsolution whenever λi ≥ 0 and λi = 1. Observe that
Z Z
− u=− u(y + x)dy.
Sh (x) Sh (0)
Now note that u(· + y) is a subsolution for each y ∈ Sh (0). Now by the closeness of of the class
Z
h2 u∗h = (− u − u(x))
for C independent of h.
116
Step 5: u ∈ C 1,1
It follows from the preceding bound that ∆u ∈ L∞ (B1/4 ). Indeed, let φ ∈ Cc∞ (B1/4 ). We
Z Z Z
u∆ψ = 2n lim uψh∗ ≤ 2n sup | u∗h ψ| ≤ C|h|L1 .
h→0 h B1/4
|∆u|L∞ (B1/4 ) ≤ C.
1
∆2he u(x) = (u(x + he) + u(x − he) − 2u(x)]
h2
belong to S(λ/n, Λ), since F is concave. Thus, since u ∈ H 2 (B1/5 ), we have that ∆2he u ∈
1
L2 (B1/10 ) for any 0 < h < 10 . It thus follows from the Harnack inequality that
sup ∆2he u ≤ C.
B1/11
1
for all 0 < h < 10 . Thus, u − C2 |x|2 is concave and so u is twice differentiable almost everywhere
uee (x) ≤ C
almost everywhere for any |e| = 1. Since F (D2 u) = 0 in the viscosity sense and u is twice
differentiable almost everywhere, F (D2 u(x)) = 0 almost everywhere. Using the definition of
117
ellipticity we see that
Thus,
and that u is twice differentiable on a set A of full measure in B1/11 where F (D2 u(x)) = 0 and
118