Lzqlu
Lzqlu
PROFESSOR ZHIQIN LU
Introduction
Let’s state the most important theorem in geometric analysis:
Theorem 0.1. All theorems in geometric analysis are also theorems in Calculus.
In Calculus, we study complicated function theory on relatively simple space: the Euclidean spaces.
In topology, we study relative simple function theory on complicated spaces. In differentially, we
create a kind of Calculus that takes the underlying topological spaces into account. As a consequence,
we have the following
1. Inevitably, the study of geometric analysis will have implications in topology;
2. We only study those Calculus properties that are covariant with the choices of local coordinates.
In this lecture, we will fully conform the above philosophy.
Example of the above setting appears in the Euclidean metric under polar coordinates. For example
in R2 , the Euclidean metric can be written as
ds2 = dr2 + r2 dθ2 .
Of course, the curvature of the above metric is zero. We shall see that, even in the most general
setting, certainly components of the curvature tensor are quite simple.
We let
ω1 = dr
ωk = f (r, θ)ηk , k>1
Then
ds2 = ω12 + · · · + ωn2
Let e1 , · · · , en be the dual basis of ω1 , · · · , ωn , and let ηkl be the connection form for (ds0 )2 . Define
ω1l = −e1 (lg f )ωl
ωkl = ηkl − ek (lg f )ωl + el (lg f )ωk
for k, l > 1. Then we have
dωi = −ωij ∧ ωj .
∂
Note that e1 = ∂r
. A straightforward computation gives
ek ( ∂f
∂r
)
dω1l + ω1k ∧ ωkl = − ωk ∧ ωl
f
Proof. Since
∂ lg f
ω1l = −
ωl .
∂r
We have
∂ ∂
dω1l = −ek (lg f ) ωk ∧ ωl − (lg f )dωl .
∂r ∂r
On the other hand, we have
∂
ω1k ∧ ωkl = − (lg f )ωk ∧ (ηkl − ek (lg f )ωl + el (lg f )ωk )
∂r
∂ ∂
= dηl + (lg f )ek (lg f )ωk ∧ ωl .
∂r ∂r
Since
∂ ∂f ∂
em ( lg f ) = f −1 em − em (lg f ) (lg f ),
∂r ∂r ∂r
we have
em ( ∂f
∂r
)
dω1l + ω1k ∧ ωkl = − ωm ∧ ωl .
f
Thus we have
em ( ∂f
∂r
) en ( ∂f
∂r
)
R1lmn = −δln + δlm .
f f
∂
The Ricci curvature at ∂r
direction is
1 ∂ 2f
R11 = R1l1l = −(n − 1) .
f ∂r2
More generally, we have the following Uhlenbeck’s trick: Let
ds2 = (dr)2 + hij (r, θ)dθi θj .
Ket ω1 = dr and ω2 , · · · , ωn be defined such that
X
hij dθi dθj = ωj2
j>1
To prove rresult, we write the above equations into matrix form. let
ω2
ω = ... , A = (aij )
ωn
Then
∂ω
= Aω
∂r
Let A = B + C be the decomposition of the matrix A into symmetric and skew-symmetric parts.
Let Q be an orthogonal matrix valued function. We consider
∂ ∂Q
(Qω) = ω + QAω
∂r ∂r
Let B̃ be a symmetric matrix-valued function such that
∂
(Qω) = B̃Qω
∂r
If we let B̃ = QBQ> . Then the equation becomes
∂Q
= −QC
∂r
Given the above equation, we can verify that any solution must be orthogonal matrices:
∂(QQ> )
= −QCQ> − QC > Q = 0.
∂r
Now back to the computation of the curvature. Let (ηij ), (i, j > 1) be the connection forms for
Riemannian metric
hij dθi dθj .
Define
∂ωi
ωi1 = . ωij = ηij
∂r
Then we have
dωi = −ωij ∧ ωj
for i > 1 and
∂ωi
ωi1 ∧ ωi = ∧ ωi .
∂r
Since ∂ω
∂r
i
= aij ωj for symmetric (aij ), we have
ωi1 ∧ ωi = aij ωj ∧ ωi = 0
Thus (ωij ) is connection forms of (ds)2 .
We compute
1
dωi1 + ωik ∧ ωk1 = Ri1kl ωk ∧ ωl .
2
For the left hand side of above, the terms with {dr ∧ dθj } should be
∂ωi1 ∂ 2 ωi
dr · = dr · .
∂r ∂r2
In other words, we have
∂ 2 aij
dr ∧ dθj = Ri11l dr ∧ ωl
∂r2
It follows that
∂ ∂ ∂ 2 aij
Ric( , ) = R1i1i = −aij
∂r ∂r ∂r2
4 PROFESSOR ZHIQIN LU
p
Let f = log det hij . Then f = log(det aij ). From the above equation, we get
2
∂ 2f ∂f ∂ ∂
2
+ = −Ric( , )
∂r ∂r ∂r ∂r
Using the above computation, we get the following comparison theorems.
Theorem 1.1. (Laplacian comparison theorem) Let X be a complete Riemannian manifold of di-
mension n. Let
Ric(X) = −(n − 1)k 2
Let N be an n-dimensional simply connected space form of constant sectional curvature −k 2 . Let
ρM , ρN be the distance functions to fixed reference points, respectively. If x ∈ X and y ∈ N such that
ρM (x) = ρN (y)
Then in the sense of distribution, we have
∆ρM (x) ≤ ∆ρN (y).
Proof. Outside the cut-locus and the reference points, the function ρM is smooth. Since ρM is a
distance function, |∇ρM | = 1. The Laplacian can be written as
1 ∂ ij √ ∂
∆= √ (g g )
g ∂xi ∂xj
Under the assumption that ds2 = dρ2 + hij dθi dθj , we have
∂ √ ∂f
∆ρ = lg g =
∂r ∂r
By our computation, we have
∂ 2f ∂f ∂ ∂
2
+ ( )2 = −Ric( , )
∂r ∂r ∂r ∂r
The corresponding function on N satisfies
∂ 2 f0 ∂f0 2
(1.1) 2
+( ) = (n − 1)k 2
∂r ∂r
The initial conditions on f and f0 are
∂f n − 1 ∂f0 n−1
∼ , ∼
∂r r ∂r r
Since
∂ ∂
−Ric( , ) ≤ (n − 1)k 2 .
∂r ∂r
By the maximum principle,
∂f ∂f0
≤ .
∂r ∂r0
This proves the comparison theorem at the smooth points of f .
Solving equation (??), we get
1
f (r) = · sinh kr
sinh kρ
Thus we have
n−1 n−1
∆N ρ = kρ coth kρ ≤ (1 + kρ)
ρ ρ
Thus the Laplacian comparison theorem can be written as
n−1
∆M ρ ≤ (1 + kρ)
ρ
NOTES IN GEOMETRIC ANALYSIS 5
We shall prove that the above inequality is true even at singular points of ρ, in the sense of distribu-
tion. To see this, we let Ω be the domain in X such that ρ is smooth on Ω. Ω is a star-like domain
in X. Apparently
X = Ω ∪ Cut(ρ),
where Cut(ρ) is the cut-locus of X. Since ρ is at least continuous, and since the measure of Cut(ρ)
at least continuous, and since the measure of Cut() is zero we must have
Z Z
ρ∆ψ = ρ∆ψ
X
for any smooth function ψ. Let Ω be an exhaustion of Ω in the sense that Ω ⊂ Ω, Ω ⊂ Ω0 if > 0
and lim→0 Ω = Ω. Then we have
Z Z Z
ρ∆ψ = − ∇ρ∇ψ = lim(−1) ∇ρ∇ψ
X X →0 Ω
Using exactyly the same method, we have the following volume comparison theorem of Bishop
Theorem 1.2. Let X be an n-dimensional complete Riemannian manifold. Let
Ric(X) ≥ −(n − 1)k 2 .
Then
vol(∂B(R))
volk (∂B(R))
is a decreasing function. Here volk (∂B(R)) is the volume of sphere of radius R of space form of
constant sectional curvature −k 2 , up to a constant, it is well defined as
volk (∂B(R)) − k sinh R.
The proof is exactly the same as that of the Laplacian comparison theorem. So we omit it.
Corollary 1.3. (Bishop) Let X be an n-dimensional Riemannian manifold. If Ric(X) ≥ (n − 1)k.
Then for any R > 0,
vol(BX (R))
vol(k, R)
is a decreasing function. In particular
volBX (k) ≤ V (k, R)
where V (k, R) is the volume of ball of radius R in space form of curvature k.
6 PROFESSOR ZHIQIN LU
Yesterday we proved the following: Let M be a complete Riemannian Manifold such taht Ric(M ) ≥
−(n − 1)k 2 . Then
1
∆r ≤ (1 + kr).
r
Here at non-smooth point, the inequality is true in the sense of distribution.
The main computation is as follows: Let the metric of M
ds2 = dr2 + hij dθi dθj
Then we have
2
∂ 2f
1 ∂f ∂ ∂
(1.2) + ≤ −Ric( , )
∂r2 n−1 ∂r ∂r ∂r
p
for f = log det(hij ). We have
∂f
∆r =
∂r
Thus using the equality (??) and the maximal principle we can prove the comparison result.
One important part of the computation is called Uhlenbeck’s trick, which is particularly useful in
computing the curvature of the Riemannian metric
dr2 + gr
where gr is the Riemannian metric on U .
Using Uhlenbeck’s trick, we get
1 ∂
(∆r)2 + ∆r + Ric(∇r, ∇r) ≤ 0
n−1 ∂r
However, the above inequality also follows from the so-called Ricci identity.
Let f be a smooth function, then by Ricci identity, we have
1 ∂
∆|∇f | = |∇2 f |2 + ∆f + Ric(∇f, ∇f )
2 ∂r
If we specialize f to be r, then |∇r| = 1 and ∆|∇r|2 ≡ 0. On the other hand
2
X 1 X 2 1
|∇2 f |2 ≥ fii2 ≥ fii = (∆f )2
i
n − 1 n − 1
and the inequality follows.
We implicitly used the first variational formula, because the special form
dr2 + hij dθi dθj
is implied by the Gauss-lemma, and we use the first variational formula to prove the Gauss Lemma.
Then f is bounded. Furthermore, there is a constant C = C(α, C1 , C2 , ||ψ0 ||C 0 , ∇ψ1 ) such that
f ≤C
Proof. We first assume that f is bounded. That is sup f < +∞. We claim that there is sequence
{xk } in X such taht for any > 0, if k is large enough, we have
f (xk ) > sup f −
|∇f |(xk ) <
∆f (xk ) <
To prove the claim, we first take a sequence {yk } such that
lim f (yk ) = sup f
k→∞
Define the cut-off function r as follows
ψ:R→R
smooth, 0 ≤ ρ ≤ 1, ρ(t) = 1 for 0 ≤ t ≤ 1,ρ(t) = 0 for t > 2 and ρ0 (t) ≤ 0 for all t ∈ R. Let R be a
large number to be determined later.Let d(x) = dist(x, yk ) be the distance function. In general, the
function d(x) is only a continuous function. But let
2
d (x)
g(x) = ρ f (x)
R2
If the maximum point of g(x) happens to be not smooth, we can always perburb the reference point
by a little. So without loss of generality, we assume that g(x) is smooth. Let xk be the maximum
point of g(x) . By the definition of xk , we have
g(xk ) ≥ g(yk ) = f (yk ) ≥ sup f −
Using the definition, we also have
d2 (xk )
1−ρ <
R2 sup f
Since
2
0 = ∇g(xk ) = ρ0 d∇df (xk ) + ρ∇f (xk )
k2
we have
−ρ0 2
|∇f |(xk ) = d|∇d|f (xk )
ρ k2
Because |∇d| = 1, if R is big enough, we have
|∇f |(xk ) <
Finally, we have
0 ≥ ∆g(xk ) = ρf + 2∇ρ∇f + f ∆ρ
In order to prove the claim, we just need to prove that
2
d (x, yk )
∆ρ <
R2
for k large enough. A straight forward computations gives that
2 0 4d2 00 2
∆ρ =2
ρ + 4 ρ + 2 ρ0 d∆d
R R k
Since the Ricci curvature of X is bounded from below, by the Laplacian comparison theorem, we
have
d∆d ≥ −C,
for some constant C depending only on the lower bound of the Ricci curvature. Since ρ0 ≤ 0, for R
large enough we have ∆ρ < for any . The claim is proved.
8 PROFESSOR ZHIQIN LU
By the above theorem, for a complete manifold, the only interesting harmonic functions may be
bounded harmonic functions.
In order to study the bounded harmonic functions, we first introduce the Ricci identity.
Let f be a smooth function of M . Define the derivative of f using the following formula fi ωi = df .
Using the same idea, we define
fij ωj = dfi − fs ωsj .
The matrix (fij ) is called the Hessian matrix. We have
fij ωj ∧ ωi = dfi ∧ ωi − fs ωsj ∧ ωi = 0.
Thus the Hessian matrix is always symmetric.
The 3rd order co-variant derivatives are defined as
fijk ωk = dfij − fis ∧ ωsj − fsj ∧ ωsi
A careful computation gives
1
fijk ωk ∧ ωj = − fs Rsikj ωk ∧ ωj
2
Thus we have
fijk − fikj = fs Rsijk
In particular, we have the following Ricci identity
flik − fkli = −fs Rilsk
Remark 2.4. We have ∆f = fii
With the preparation above, we prove the following.
Theorem 2.5. Let M be a complete Riemannian manifold, dim M = n ≥ 2. Ric(M ) ≥ −(n −
1)k, k ≥ 0. Let u be apositive harmonic function. Then on any geodesic ball Ba (x), we have
1
!
|∇u| 1 + ak 2
≤ Cn
u a
where Cn is a constant only depends on n.
Proof. We first prove that
1 X
(2.2) ∆|∇u|2 ≥ u2ij − (n − 1)k|∇u|2
2 i,j
Since !2
X 1 X 1
u2jj ≥ ujj = u2
j>1
n−1 j>1
n − 1 11
we have
1 1 X 2
∆|∇u| ≥ u − (n − 1)k|u1 |
n − 1 |∇u| j 1j
Now we assume that ψ = |∇u|/u. Then we have
∆|∇u| 1 1
∆ψ = + 2∇|∇u|∇ + |∇u|∆
u u u
Since
1 1 ∇ψ∇u
2∇|∇u|∇ = 2∇(ψu)∇( ) = −2 − 2ψ 3
u u u
We have
∆|∇u| ∇ψ∇u
∆ψ = −2
u u
We have
∇ψ∇u 1
= −ψ 3 − 2∇|∇u|∇
u u
Thus if > 0 is small enough
∆|∇u| ∇ψ∇u
− ≥ ψ 3
u u
Thus we get
∇ψ∇u
∆ψ ≥ −(n − 1)kψ − (2 − ) + ψ 3
u
Using the maximum principle, we have ψ is bounded.
Corollary 2.6. Let M be a complete Riemannian manifold if Ric(M ) ≥ 0. Then any positive
harmonic function is a constant.
Proof. If Ric ≥ 0, then k = 0. We have
∇ψ∇u
∆ψ ≥ −(2 − ) + ψ 3
u
2
For = n−1 . Thus using the generalized maximal principle, ψ ≡ 0
Corollary 2.7 (Harnack Inequality). Let M be n-dimensional Riemannian manifold, Ric(M ) ≥
−(n − 1)k. If u is a positive harmonic function in Ba . Then
sup u ≤ C(n, a, k) inf u
Ba/2 Ba/2
and
2 θi θj θi θj
hij = r gij − g1j − gi1 + g11 2
θ1 θ1 θ1
p
Thus det(hij ) = rn−1 · µ, for µ being a regular function (at least for fixed θi ). Thus
∂f n−1
∆ρ = ∼ + small terms.
∂r r
If we choose (gij ) to be normal, we can actually compute
θi θj 1
det δij + 2 = 2
θi θ1
p
Thus det(hij ) can be extended as a regular function near 0.
3. Eigenvalue problems
We first make different notation of the Laplace operator.
We assume that M is a Riemannian manifold with the Riemannian metric
X
ds2 = gij dxi dxj .
Let (g ij ) be the inverse matrix of (gij ) and let g = det(gij ). Then under the local coordinates
(x1 , · · · , xn ) we define
1 ∂ ij √ ∂
∆= √ g g
g ∂xi ∂xj
Apparently, we can write
∂2 1 √ ∂
∆ = g ij + √ g ij g
∂xi ∂xj g ∂xj
12 PROFESSOR ZHIQIN LU
C ∞ (M ) is not a complete metric space. The complete metric space is L2 (M ). However, there is no
way that we can extend ∆ on L2 (M ).
Proof. The key pint is that any differential operator is a closed graph operator. Thus if ∆ is extend-
able, then by close graph theorem, ∆ has to be a bounded operator. however ∆ is not a bounded
operator, as by the example of Heaviside function
0 x≤0
f (x) = continuous cure connected (0,0) and (1,1)
1 x≥1
Thus we can only extend the operator into a densely defined self-adjoint operator.
Recall that an operator ∆ is self-adjoint, if
Dom(∆) = Dom(∆∗ )
and
h∆f, gi = hf, ∆gi
for any f, g ∈ Dom(∆)
In functional analysis, we have the following theorem: Let
Z
Q(φ, ψ) = ∇φ∇ψ.
Then Q is a non-negative quadratic form defined on H01 (M ). Then there is a unique densely defined
operator A such that
Q(φ, ψ) = −(Aφ, ψ)
such an operator A is in fact called the Dirichlet Laplacian Operator.
As an exercises, we prove that on only manifold, L2 harmonic function must be constant.
Theorem 3.2. Let A be the Dirichlet extension of Laplacian ∆. A function f is called A-harmonic,
if f ∈ Dom(A) and Af = 0. If f ∈ L2 (M ), then f is a constant.
Proof. The key point is that
Q(ρ2 f, f ) = ρ2 f, ∆f = 0
Thus using the same method as before, f is a constant.
If M is a compact manifold with no boundary. we still use ∆ to denote the Dirichlet extension of
the Laplace operator. By the elliptic regularity, the spectrum of ∆ are discrete. That is, there is a
sequence
0 = λ0 < λ1 ≤ λ2 ≤ · · ·
2
such that for any λi , there ∃fi ∈ L such that,
∇fi = −λi fi
We have similar results similar results for manifolds with boundary conditions. To be more precise,
the Laplacians acting on functions with the following boundary conditions.
• Dirichlet boundary condition: ∆ acting on f vanishing on the boundary.
∂f
• Neumann boundary condition: ∆ acting on function such that ∂n = 0.
For eigenvalues, we have the following minimax principle. Assume that M is a closed manifold, then
R
|∇f |2
λ1 = Rinf R 2
f =0 f
R
To prove the above result, we let ψ be any smooth function such that M ψ = 0. Then by the
definition of λ1 , we have
|∇(f + )|2
Z
R ≥ λ1 .
(f + ψ)2
14 PROFESSOR ZHIQIN LU
Note that Z
(∆f + λ1 f ) = 0
Then
∆f + λ1 f ≡ 0
and λ1 is the first eigenvalue.
By elliptic regularity, λ1 > 0. Thus we have the following Poincare-inequality. There exists a
constant C, such that Z Z
|∇f | ≥ C f 2
2
R
for any function with f = 0. Before going further, let’s prove the following C0 -area formula.
Theorem 3.3. Let M be a compact Riemannian manifold with boundary. Let f ∈ H 1 (M ). Then
Z Z+∞Z
g
g= dσ.
M {f =σ} |∇f |
−∞
Proof. Without loss of generality, we assume that |∇f | 6= 0. Thus by the implicit function theorem
{f = σ}is a smooth manifold.
Using cut-off function, we may assume that Supp g is contained in a coordinate chart. Thus we
may assume that the Riemannian metric is given under the global coordinates (x1 , · · · , xn ) as follows
X
ds2 = gij dxi dxj
by definition Z Z q
g= g det(gij )dx1 · · · dxn
M M
Since ∇f 6= 0, we can solve the equation
f =σ
by
x1 = x1 (σ, x2 , · · · , xn )
or in order word, by the implicit function theory (σ, x2 , · · · , xn ) is a local coordinate system as well.
The Jacobian of the transformation is
∂x1
dx1 ∧ · · · ∧ dxn = dσ ∧ dx2 ∧ · · · ∧ dxn
∂σ
On the other side, restricting to f = σ, the Riemann metric can be written as
∂x1 ∂x1 ∂x1 ∂x1
(g11 · + g1l + gk1 + gkl )dxk dxl
∂xk ∂xl ∂xk ∂xl
If we choose local coordinates such that gkl = δkl . Then we have
∂x1 ∂x1
· + δkl dxk dxl
∂xk ∂xl
The volume form of the above is
∂x1 ∂x1
X ∂x1 2 1 X ∂f
2
|∇f |2
(3.1) det δkl + · =1+ = =
∂xk ∂xl ∂xk ∂f ∂xk ∂f
∂x1 ∂x1
Thus we have
Z Z Z
∂x1
q q
g= g det(gij )dx1 · · · dxn = g det(gij ) dσ ∧ dx2 ∧ · · · ∧ dxn
M M M ∂σ
NOTES IN GEOMETRIC ANALYSIS 15
Then using the Sobolev inequality, the isopermetric inequality follows by letting → 0.
In order to prove that the isopermetric inequality implies the Sobolev inequality, we use the co-area
formula. We assume that f ≥ 0. Then
Z Z∞
|∇f | = Area(f = σ)dσ
M
0
We also have
Z Z∞ Z∞
n n n n
|f | n−1 = vol(f n−1 > λ)dλ = vol(f > σ)σ n−1 dσ
M n−1
0 0
Using the isopermetric inequality, we have
Z Z∞ Z∞
n−1
|∇f | = Area(f = σ)dσ ≥ C vol(f > σ) n dσ
M
0 0
Thus in order to prove the Sobolev inequality, we just need to prove that
Z∞ Z∞
n−1 1 n−1
vol(f > σ) n dσ ≥ C( vol(f > σ)σ n−1 dσ) n
0 0
16 PROFESSOR ZHIQIN LU
We let
F (σ) = vol(f > σ)
Zt
n−1
φ(t) = F (σ) n dσ
0
t n−1
n
Z
1
ψ(t) = F (σ)σ n−1 dσ
0
Then φ(0) = ψ(0). Using the monotonicity of F (σ) we can prove that
n
φ0 (t) ≥ ψ 0 (t)
n−1
Thus
n
φ(∞) ≥ ψ(∞)
n−1
Corollary 3.6.
Z n−p
np
Z p1
np
p
f n−p ≤C |∇|
for any p > 1.
Definition 3.7. Let M be a compact Riemannian manifold if ∂M 6= ∅,
vol(∂Ω)
hD (M ) = inf | Ω ⊂⊂ M
vol(Ω)
if ∂M = ∅
vol(H)
hN = inf | H is a hyper surface.
min(vol(M1 ), vol(M2 ))
Theorem 3.8. (Cheeger) For Dirichlet condition, we have
1
λ1 ≥ h2D
4
For Neumann codition
1
λ1 ≥ h2N (M )
4
Proof. We only prove the case for Dirichlet condition. We first observe that, if there is constant µ
such that Z Z
|∇φ| ≥ µ |φ|
M M
1 2
for any φ with φ|∂M = 0. Then λ1 ≥ To see this, we consider φ = f 2
4
µ.
Z Z Z 21 Z 12
2 2 2
µ f ≤2 |f | · |∇f | ≤ 2 f |∇F |
M M M
Thus Z Z
1 2 2
µ f ≤ |∇f |2
4 M M
Since the above is true for any function f , we must have
1
λ1 ≥ µ2
4
Finally, we prove a result which is well known but can’t readily be found in the literature.
NOTES IN GEOMETRIC ANALYSIS 17
Thus we have Z Z
∇f ∇φ = −µ1 fφ
M M
by Green’s formula Z Z Z
∂f
∇f ∇φ = φ − ∆f φ
M ∂M ∂n M
Thus since ∆f = µ1 f , we have Z
∂f
=0
∂M ∂n
∂f
and we must have ∂n
= 0.
4. Eigenvalue Problems (II)
By the variational characterizing of the eigenvalues, we know that it is usually more difficult to
get the lower bound estimate of eigenvalues. Among all the eigenvalues, the lower bound of the first
eigenvalue is particularly important.
The Cheeger’s result did give a lower bound estimate of the first eigenvalues. But the bounds
are not ”computable”. In geometry, ”computable” bounds provide effective versions of Poincare and
Sobolev inequalities.
The following Lichnerowicz therem gives a good lower bound of the first eigenvalue for closed
manifold.
Theorem 4.1. (Lichnerowicz) Let M be a closed n-dimensional Riemannian manifold. Assume that
Ric(M ) ≥ (n − 1)k > 0
Then λ1 ≥ nk.
Proof. One line proof, let µ be the first eigenfunction. Then use the Ricci identity we have
1 X
∆|∇u|2 ≥ u2ij + ∇u∇∆u + Ric(∇u, ∇u)
2
We have
!2
X X 1 X λ1 2
u2ij ≥ u2ii ≥ uii = u
n n
Ric(∇u, ∇u) ≥ (n − 1)k|∇u|2
Thus we have
1 λ2
∆|∇u|2 ≥ 1 u2 − λ1 |∇u|2 + (n − 1)k|∇u|2
2 n
Taking integration on both sides, we get
λ21
− λ21 + (n − 1)kλ1 ≤ 0
n
The theorem follows.
18 PROFESSOR ZHIQIN LU
In 1962. Obata proved, if λ1 = nk, then M has to be the standard sphere. For the rest of this
section, we use the gradient estimates to find ”computable” lower bounds of the first eigenvalue. We
prove the following theorem
k2
Theorem 4.2 (Li-Yau). let M be a closed manifold and Ric(M ) ≥ 0. Then λ1 ≥ 2d2
, d is the
diameter.
Proof. Let u be the first eigenfunction. After normalization, we may assume that
1 = sup u > inf u = −k ≥ −1
for some 1 ≥ k > 0. Let
1−k
u− 2
ũ = 1+k
2
The after this linear change of u, we have
∆ũ = λ1 (ũ + a)
sup ũ = 1
inf ũ = −1
for a = 1−k
1+k
, 1 > a ≥ 0.
1
Let g = 2 (|∇ũ|2 + (λ1 + )ũ2 ) for some > 0 to be determined later. Assume that at x0 .
g(x0 ) = max g.
Using the maximum principle, at x0 , we have
ũj ũji + (λ1 + )ũui = 0
and
0 ≥ ∆g
= ũ2ij + Ric(∇ũ, ∇ũ) + ∇ũ∇∆ũ + (λ1 + )|∇ũ|2 + (λ1 + )ũ∆ũ
≥ ũij − λ1 |∇ũ|2 + (λ1 + )|∇ũ|2 − λ1 (λ1 + )ũ(ũ + a)
If at x0 , ∇ũ = 0. Then we have
|∇ũ|2 + (λ1 + )ũ2 ≤ (λ1 + )
In particular, we have
|∇ũ|2 + λ1 (1 + a)ũ ≤ λ1 (1 + a)
If ∇ũ(x0 ) 6= 0. Then using the Cauchy inequality
(ũi ũj ũij )2
ũ2ij ≥ 4
= (λ1 + )2 ũ2
|∇ũ|
Thus we have
0 ≥ ∇g(x0 )
≥ (λ1 + )ũ + |∇ũ|2 − λ1 (λ1 + )ũ2 − λ1 (λ1 + )a
≥ 2g − λ1 (λ1 + )a
For any > λ1 a, the above gives
|∇ũ|2 + λ1 (1 + a)ũ ≤ λ1 (1 + a)
Let
f (t) = arcsin ũ(σ(t)),
where σ(t) is the arc-length curve connecting the minimal point and the maximum point of ũ. Then
by the above argument, we have p
|f 0 (t)| ≤ λ1 (1 + a)
NOTES IN GEOMETRIC ANALYSIS 19
It is a very interesting result. We end this section by citing a result of Li and Croke.
Theorem 4.5. Let M be a manifold with boundary. Then there is a constant C = C(n, d, V, k) > 0
such taht the Sobolev constant is > C > 0.
Proof. We first assume that p ≥ 2. Let ψ ≥ 0, ψ ∈ C0∞ (B(a)). Then since ∆u = 0, we have
Z
∇u∇ψ = 0
B(a)
Note that
4 p
up−2 |∇u|2 = 2
|∇u 2 |2
p
We have Z Z
p
2 2
ρ |∇u | ≤ C
2 |∇ρ|2 up
B(a) B(a)
If we allow C to be a litter bigger, we shall get
Z Z
p
2
|∇(ρu 2 )| ≤ C |∇ρ|2 up
B(a) B(a)
We let 2∗ = 2n
n−2
> 2. Using the Sobolev-inequality we have
Z p 2∗ 21∗ Z
(5.1) ρu 2 ≤C |∇ρ|2 up
B(a) B(a)
NOTES IN GEOMETRIC ANALYSIS 21
n k
Specializing p = pk , where pk = p( n−2 ) , we have
p1
C · 4k
k
||u||Lpk +1 (B(Rk+1 )) ≤ ||u||Lpk (B(Rk ))
(1 − θ)2 a2
Iterating, we have
p1
Y C · 4k k
||u||Lpk +1 (B(Rk+1 )) ≤ ||u||Lp (B(a))
(1 − θ)2 a2
We need to prove that
p1
Y C · 4k k C
≤ ( Exercises)
(1 − θ)2 a2 ((1 − θ)2 a2 )n/2p
Since the right-hand side is independent of R, we let k → ∞ and get
C
||u||L∞ (B(θa)) ≤ ||u||Lp (B(a))
((1 − θ)a)n/p
This proves the theorem for p ≥ 2.
Now we assume that 0 < p < 2. Using the result for p = 2, we have
Z 12
C 1− p2 p
||u||L∞ (B(θa)) ≤ ||u||L∞ (B(a)) u
((1 − θ)a)n/2 B(a)
1 C
||u||L∞ (B(θa)) ≤ ||u||L∞ (B(a)) + ||u||Lp (B(a))
2 ((1 − θ)a)n/p
Let ψ(s) = ||u||L∞ (B(sa)) , we get
1 C
ψ(s) ≤ ψ(t) + ||u||Lp (B(a))
2 (1 − s)n/p
∀0 < s < t ≤ a. Iterating again, we get
C
ψ(s) ≤ ||u||Lp (B(a)) , ( exercise )
((1 − s)a)n/p
Theorem 5.2. (weak Harnark inequality) There is a constant C > 0, p0 > 0 such that
Z p1
1 p0
0
inf ≥ u
B(aθ) C B(a)
Thus u−1 is a subsolution. Using the above lemma, for any p, we have
Z
−p
sup u ≤ C u−p
B(θ) B(1)
In order to prove the theorem, we just need to prove that for p > 0 small enough
Z Z
−p
u up ≤ C
B(1) B(1)
H
We let ω = log u − β, where β = log u. We shall establish
I
(5.2) ep|ω| ≤ C.
B(1)
for some σ > 1. To see this, let ρ be the cut-off function whose support is within B(σ̄) for some
σ̄ > σ. Since u is harmonic, we have
Z
∇u∇(u−1 ρ2 ) = 0
B(σ̄)
It follows that
Z Z
2 2
ρ |∇ω| ≤ ∇ω∇ρ2
B(σ̄) B(σ̄)
sZ sZ
≤2 ρ2 |∇ω|2 |∇ρ|2
B(σ̄) B(σ̄)
Since ρ only depends on σ, σ̄, we get the desired inequality. By the Poincare inequality, we have
Z
|ω|2 ≤ C
B(σ̄)
To get the estimate (??), we still use the Moser iteration. First observe that
∆ω = −|∇ω|2
Let ρ be a cut-off function to be determined later. Then we have
−ρ2 |ω|2q ∆ω = ρ2 |ω|2q |∇ω|2
It follows that
Z Z
2 2q 2
ρ |ω| |∇ω| = ∇ω∇(ρ2 |ω|2q )
Z Z
= 2 ρ∇ρ∇ω|ω| + 2q ρ2 |ω|2q−1 |∇ω|2
2q
Thus we have Z Z
2 q+1 2 2q 2q+2 2
|∇(ρ |ω| )| ≤ 2Cq (2q) + |ω| |∇ρ|
Of course, we need to prove the existence of these functions when M is complete non-compact.
The pure point spectrum of ∆ are those λ ∈ R such that
• There exist a L2 function f 6= 0 such that ∆f + λf = 0.
• The multiplicity of λ is finite.
• In a neighborhood of λ, it is the only spectrum point.
We define
ρ(∆) = {y ∈ R|(∆ − y)−1 is a bounded operator }
σ(∆) = R−ρ(∆) is the spectrum set of ∆. From the above discussion, σ(∆) decomposes as the union
of pure point spectrum, and the so-called essential spectrum, which is by definition, the complement
of pure point spectrum.
Using the above definition, λ ∈ σ(∆) belongs to the set σess (∆), if either
• λ is an eigenvalue of infinite multiplicity, or
• λ is the limiting point of σ(∆).
The following theorems in functional analysis characterizing the essential spectrum.
Theorem 6.1. A necessary and sufficient condition for the interval (−∞, λ) to intersect the essential
some of an self-adjoint densely defined operator A is that, for all > 0, there exists an infinite
dimensional subspace G ⊂ Dom(A), for which (Af − λf − f, f ) < 0.
Theorem 6.2. A necessary and sufficient condition for the interval (λ − σ, λ + σ) to intersect the
essential spectrum of A is that there exists an infinite dimensional subspace G ⊂ Dom(A) for which
||(A − λI)f || < σ||f ||, f ∈ G.
For reference, see Dormelly [?], Topology 20, 1-14,1981.
Using the above result, we give the following variational characterization of the lower bound of
spectrum and the lower bound of essential spectrum.
NOTES IN GEOMETRIC ANALYSIS 25
Now we assume that K is a compact set. Let K 0 a larger ball containing K. Let ρ be the cut off
function such thatR ρ ≡ 1 on K but ρ ≡ 0 outside K 0 . We claim that (key point) that ∀ > 0, there
is an f ∈ V with f 2 = 1 but
Z
ρ2 f 2 <
M
If the above is not true, then for any f ∈ V
Z Z
2 2
ρ f ≥ 0 , if f2 = 1
M
Since the set f ∈ V is of infinite dimensional, the set ρf is of infinite dimensional. Thus we can find
an orthogonal basis Z
ρ2 fi fj = 0, i 6= j
R
while we can still keeping f 2 = 1. We consider
Z Z Z
|∇(ρfi )| ≤ 2 |∇ρ| fi + 2 ρ2 |∇fi |2
2 2 2
≤ 2C + 2(λ0ess + )
Thus R
∇(ρfi )2 2C + 2(λess + )
R ≤
(ρfi )2 0
for infinitely dimensional space. This is a contradiction because on the compact set K 0 , the eigen-
values go to infinity. R
With the above preparation, we can prove out theorem ∀ > 0, we find an f with f 2 = 1 but
Z
ρ2 f 2 <
26 PROFESSOR ZHIQIN LU
Consider ρ1 = 1 − ρ,
Z Z Z Z
2
|∇(ρ1 f | = ρ21 |∇f |2 +2 ρ1 f ∇ρ1 ∇f + f 2 |∇ρ1 |2 .
This is obvious: let B(x0 , R0 ) be a ball radius R0 with center x0 such that |x0 | > 2R + 1 + R0 . Then
B(x0 , R0 ) ⊂ Rn \ B(R). Let f be the first Dirichlet eigenfunction of B(x0 , R0 ). Then if R0 → ∞
R
|∇f |2
R <
f2
Zero extending f to R \ B(R) we get the result.
Using ??, we can even prove that
σess (∆) = [0, +∞)
q the above, we make the following observation ∀λ > 0, ∀m ∈ Z, we can find a square of
To prove
size mπ nλ such that
r r
λ λ λ
f = sin x1 · sin x2 · · · sin xn
n n n
is an eigenfunction with Dirichlet condition: ∆f + λf = 0.
However, we can’t use Theorem ?? directly, the reason is that f is not second differentiable near
the boundary. Thus we need to use cut off function. Without loss of generality, we may assume that
the square is in the first quadratic, we denote such a square to be S. Let S1 be a square with the
same center as S1 but with smaller size. We assume that the distance of the boundary of S1 to S is
NOTES IN GEOMETRIC ANALYSIS 27
d, which is to be determined later. We choose a cut off function ρ such that ρ ≡ 1 on S1 and ρ ≡ 0
outside S. By the same argument as before we can prove that
||∆(ρf ) + λρf ||L2 < ||ρf ||L2
Thus σess (∆) = [0, ∞).
Unfortunately, the above argument doesn’t apply to the general case. Thus following result of
Wang Jiaping is very surprising and interesting
Theorem 6.5 (J-P. Wang). Let M be a complete manifold with non-negative Ricci curvature. Then
σess (∆) = [0, ∞).
Note that λess was known before, e.g. P.Li-Wang, Books.
As in the case of harmonic functions, for positive solutions of the heat equations
∂
(∆ − )u = 0
∂t
we also have the differentiable Harnack inequality. The theorem is as follows:
28 PROFESSOR ZHIQIN LU
k=1 1 t > d2
NOTES IN GEOMETRIC ANALYSIS 29
We can compare the above eigenvalue estimate with the result of Cheng-Li. From (??), we can
get
∞
n
X
e−λk t ≤ Ct− 2 , t ≤ d2
k=1
Thus we have
Z Z 2+n
∂ 2 2
n
H(x, y, t) dy ≤ −C H(x, y, t)
∂t M M
which implies
gα(n/2) (x, y) ≤ C4 ψ 2 (x)e−βd(x,y) .
Cheeger-Yau’s Heat kernel comparison theorem.
Theorem 7.7. Let M be a complete Riemannian manifold such that Ric(M ) ≥ 0. Fixing x ∈
M, r0 > 0. Then the heat kernel H(x, y, t) in B(x, r0 ) and the heat kernel (r(x, y), t) in V (k, r0 )
satisfies the following
(r(x, y), t) ≤ H(x, y, t)
For both Dirichlet and Neumann conditions.
Proof. Using the property of the heat kernel, we have
Zt Z
d
H(x, y, t) − (x, y, t) = ((x, z, t − s)H(z, y, s))dzds
B(x,r0 ) ds
0
Zt Z
d
=− ( (r(x, z), t − s))H(z, y, s)dzds
B(x,r0 ) ds
0
Zt Z
d
+ (r(x, z), t − s) H(z, y, s)dzds
B(x,r0 ds
0
Zt Z
=− ˜
∆(r(x, z), t − s)H(z, y, s)dzds
B(x,r0 )
0
Zt Z
+ (r(x, z), t − s)∆H(z, y, s)dzds
B(x,r0 )
0
Using the Green’s formula, under either the Dirichlet or Neumann boundary condition, we have
Z Z
(r(x, z), t − s)∆H(z, y, s)dz = ∆(r(x, z), t − s)H(z, y, s)dz.
B(x,r0 ) B(x,r0 )
This essentially follows from the Laplacian comparison theorem: Let x = (r, ξ), ξ ∈ S n−1 . Then
2
˜ = ∂ + m(r) ∂ , ∂ p
∆ m(r) = log det g̃
∂r2 ∂r ∂r
∂2 ∂ ∂ √
∆ = 2 + m(r, ξ) , m(r, ξ) = log det g
∂r ∂r ∂r
Since Ric(M ) ≥ (n − 1)k, using the volume comparison theorem, we have m(r, ξ) ≤ m(r). Since
∂
∂r
< 0, we have
˜
∆(r, t − s) ≤ ∆(r, t − s)
In the above proof, we didn’t take the cut-locus into a count, using some kind of limiting process,
we can overcome the difficulty.
Theorem 7.8. Let M be a complete Riemannian manifold such that Ric(M ) ≥ (n−1)k, n = dim M .
We use B(x0 , r) to denote the ball centered at x0 with radius r. Let V (k, r) be the ball of radius r in
a simply connected space form. Then with the Dirichlet boundary condition we have
λ1 (B(X0 , R)) ≤ λ1 (V (k, r))
Proof. Let H(x, y, t) and (x, y, t) be the corresponding heat kernel. Then we have
X
H(x, y, t) = e−λ1 t φ2i (x)
X
(x, y, t) = eλ̃1 t φ̃2i (x)
This last asymptotical expansion of the Green’s function also implies that
∆x G(x, y) = −δx,y .
From the asymptotical behavior we can find that n = 2 and n > 2, the Green’s functions are very
different. We make the following
Definition 8.1. A complete manifold is said to be parabolic, if and only if it doesn’t admit a positive
Green’s function. Otherwise it is said to be non-parabolic.
Definition 8.2. An End, E, with respect to a compact subset Ω ⊂ M an unbounded connected
component of M \ Ω. The number of ends with respect to Ω, denoted by NΩ (M ), is the number of
unbounded connected component of M \ Ω.
Definition 8.3. An end E is said to be parabolic, if it doesn’t admit a positive harmonic function
f satisfying
f ≡ 1 on ∂E
and
lim f (y) < 1
n→E(∞)
where E(∞) denotes the infinity of E. Otherwise, E is said to be non-parabolic and the function f
is said to be a barrier function of E.
We prove the following result:
Theorem 8.4. Let E be an parabolic end. Let A(R) = E ∩ ∂B(r) where B(R) is the ball of radius
R with respect to some reference point. Let f be a harmonic function on E such that
f |∂E = 1, f |A(R) = 0
Then Z
lim |∇f |2 → 0
R→∞ E
Proof. Using the Green’s formula, we have
Z Z
2 ∂f
|∇f | = −
E E ∂r
where ∂
∂r
is th outer normal direction, we claim that ∂f∂r
→ 0. To see this, we take a sequence
R1 < R2< · · · < Rk → ∞. The corresponding harmonic function fi = fRi . By the maximal
principle, fi are increasing sequence on any compact set of E. Let
lim fi = f
i→∞
Then f must be a positive harmonic function. by the parabolicity, f ≡ 1. By the maximal principle
again
∂fi
→0
∂r
as i → ∞.
Example 8.5. R2 and Rn . Let R > 0 be a big number. Let
F (R) = {f ∈ X0∞ (R) : f = 1 for |x| < R, f rotational symmetric }
If n > 2, then for any c > 0, there exists an R0 such that for any R > R0
Z
|∇f |2 > C
Rn
for any f ∈ F (R). If n = 2, then for any > 0 there exists R0 > 0 such that for any R > R0 , we
can find an fk ∈ F (R) for which Z
|∇f |2 <
R2
NOTES IN GEOMETRIC ANALYSIS 33
Let f be the harmonic function defined on Bp (R) \ Bp (1) satisfying the boundary conditions
f (y) = si (1) on ∂Bp (1)
f (y) = Gi (p, y) on ∂Bp (R)
The maximum principle implies that
f (y) ≥ Gi (p, y) on Bp (R) \ Bp (1)
In particular, we have
∂f ∂Gi
≤ on ∂Bp (R)
∂r ∂r
On the other hand, since f (y) is harmonic, Stokes theorem implies that
Z Z Z
∂f ∂f
0= ∇f = −
Bp (R)\Bp (1) ∂Bp (R) ∂r ∂Bp (1) ∂r
Thus we get Z
∂f
≤ −1
∂Bp (1) ∂r
Let’s now consider h to be the harmonic function defined on Bp (R) \ Bp (1) satisfying the boundary
conditions
h(y) = Si (1) on ∂Bp (1)
h(y) = ii (R) on ∂Bp (R)
Again the maximum principle implies that h(y) ≤ f (y) on Bp (R) \ Bp (1) and
∂h ∂f
≤ on ∂Bp (1)
∂r ∂r
Thus we have Z Z
∂h ∂h
= ≤ −1
∂Bp (R) ∂r ∂Bp (1) ∂r
Define the function
−1
ZR ZR
dt dt
g(r) = (Si (1) − ii (R)) + ii (R)
Ap (t) Ap (t)
1 r
Then g(r(y)) will have the same boundary conditions as h(y). The Dirichlet integral minimizing
property for harmonic functions implies that
Z Z
2
|∇h| ≤ |∇g|2
Bp (R)\Bp (1) Bp (R)\Bp (1)
R −1 2
ZR Z
dt 1
= (Si (1) − ii (R)) Ap (r)dr
Ap (t) Ap (r)
1 1
R −1
Z
dt
= (Si (1) − ii (R))2
Ap (t)
1
Theorem 8.9. Let Ω be a bounded convex domain of Rn . Let u be the first Dirichlet eigenfunction.
Let
∆u = −λ1 u, u > 0.
Then log u must be concave.
Proof. We choose any function u0 > 0, u0 |∂Ω = 0 such that − log u0 is concave. Such a function
always exists. For example, we can take the convex hull of graph of − log u.
Consider the flow
∂u
= ∆u + λ1 u, u|∂Ω = 0
∂t
We assume that ut → u, the first eigenfunction. We are going to use the maximum principle. Let
T be the biggest number such that det(−∇2 log u) is degenerated. Thus there is an x0 ∈ M and a
direction i such that
−(log u)ii (x0 ) = 0
and for other j, (− log u)jj (x0 ) ≥ 0. Let ϕ = − log u. Then the evolution of ϕ is
∂ϕ
= ∆ϕ − |∇ϕ|2 − λ1
∂t
By the maximum principle, ϕiik = 0, ∂ϕ ii
∂t
≤ 0 and ∆ϕii ≥ 0. Thus
∂ϕii
0≥ = ∆ϕii − 2ϕk ϕkii − 2ϕ2kl ≥ −2ϕ2ki .
∂t
However by convexity, ϕ2ki ≤ ϕii · ϕkk = 0, ϕki ≡ 0. The theorem follows from strong maximum
principle.
In fact, if x0 ∈ ∂Ω. By choosing an orthonormal frame {e1 , · · · , en } such that e1 be the out normal
direction. If we let e1 |∂Ω = ∂x∂ 1 . Then
n
∂G X
(x0 ) = 2 vi vi1 − 2λv1 (µ − v)
∂x1 i=1
We claim that
∂G
(x0 ) ≥ 0
∂x1
36 PROFESSOR ZHIQIN LU
∂v
To see this, we fist observed that ∂x1
= 0. This can be proved using the variational principle. Thus
n
∂G X
(x0 ) = 2 vi vi1
∂x1 i=2
Thus we have
∂G X
0≤ (x0 ) = −2 hij vi vj
∂x1
Since Ω is assumed to be convex, all vi = 0. Thus
G(x) ≤ sup λ(µ − v)2
x∈Ω
Since ∇v 6= 0, we can choose local orthonormal frame such that v1 6= 0, vi = 0(i > 1). Thus at x0 ,
we have
v11 (x0 ) = λ(µ − v)
vi1 (x0 ) = 0, 2 ≤ i ≤ n
Thus
X
0 ≥ ∆G = 2 vij2 + 2λ(µ − v)v − 4v12 (log u1 )11
Using a result of Brascamp and Lieb, log u1 is an concave function. Thus (lg u)11 (x0 ) ≤ 0. Thus
X
vij2 + λ2 (µ − v)v ≤ 0
Or in other word
v11 + λ2 v(µ − v)|x0 ≤ 0
which is not possible. Thus we must have ∇v = 0 at x0 and thus
G(x) ≤ sup λ(µ − v)2
we have
√ |∇v|2
λ≥ p
(sup v − inf v)2 − (sup v − v)2
Using the same method as in the estimate of the first eigenvalue, we get
π2
λ1 ≥
4d2
NOTES IN GEOMETRIC ANALYSIS 37
Remark 9.1. Using the method of Zhong-Yang, Zhong-Yu was able to modified the above estimate
to
π2
λ2 − λ1 ≥ 2
d
However, even the above estimate is not optimal, Van der Berg ( see also Yau, Problem session)
conjectured that
3π 2
λ2 − λ1 ≥ 2
d
The estimate is asymptotically accurate for a very thin rectangular.
Not much progress was made in the direction of this conjecture. In JFA, 176, 368-399(2000),
Banuelos and Mendez-Hernandez proved that if Ω is a convex domain in R2 which is symmetric with
respect to both x− and y−axises, then the Van der Berg conjecture is true.
For a triangle, Lu and Rowlett proved the following
Theorem 9.2. Let ∆ be a triangle and let d = d(∆) be the diameter of ∆. That is, d is the longest
side of the triangle. Then ∀c > 0
(λ2 (∆) − λ1 (∆))d2 (∆) ≤ C
is a compact set.
Proof. When a sequence of triangles doesn’t converge, then the smallest angle must go to zero. We
assume that the smallest angle is απ and we assume that the diameter d = 1. The key part of the
proof is the following cutting lemma.
Let P2 , P1 be two points on BC such that |P1 C| = α , |P2 C| = 2α , where < 29 . For the sake
of simplicity, we assume that ∠ACB = π2 . We are going to prove that the eigenvalues of ABC and
AQ2 P2 C are almost the same, thus cutting an acute angle won’t make too much difference.
38 PROFESSOR ZHIQIN LU
Noting that Z Z
2
ρ f1 ≥ ρ2 f12 ≥ 1 − β ≥ 1 − α2/3−r
U0 U
we then have
α2/3−3
λ1 (U 0 ) ≤ λ1 (ABC) + ≤ λ1 (ABC) + α2/3−3
1 − α2/3−
By modifying the above argument, we are able to prove that
λ2 (U 0 ) − λ2 (ABC) ≤ α2/3−3 + (λ2 − λ1 )α2/3−3
Solving the above inequality, we have
λ2 − λ1 ≥ λ2 (U 0 ) − λ1 (U 0 ) − O(α2/3−3 )
By the gap theorem, and using the fact that the diameter of U 0 is at most 4α , we get
π2
λ2 (U 0 ) − λ1 (U 0 ) ≥
16α2
The compactness theorem follows.
Proof. We take r = d(x, y). Then since B1 (y) ⊂ Br+1 (x), we must have
1 1 1 1
ν(B1 (y)))− 2 ≥ ν(Br+1 (x))− 2 ≥ Ce− 2 (r+1) ν(B1 (x))− 2
For any x. Thus the integration in the lemma if less than
Z
1
C e−r · e 2 (r+1) ν(B1 (x))−1 dy
M
Rr
We let f (r) = vol(∂Br (x)) and F (r) = f (t)dt. Then up to a constant, the above expression is less
0
than
Z∞
1
−1
(ν(B1 (x))) e− 2 r f (r)dr
0
n−1
By volume comparison again, f (r) ≤ Cr ν(B1 (x)). The lemma follows.
40 PROFESSOR ZHIQIN LU
Lemma 10.3. For any β > 0, there is an n ∈ N, α < 0, C < ∞ such that the integral kernel
n/2
gα (x, y) of (∆ − α)−n/2 exists and satisfies
gαn/2 (x, y) ≤ Ce−βd(x,y) ϕ(x)2
1
where ϕ(x) = (ν(B1 (x)))− 2 .
The lemma was proved on page 40 theorem 9.6 , using the heat kernel estimates.
Before gonging further, let’s make some remarks on the kernel of an operator. Let A be an operator
on functions. If there is a function g(x, y) such that
Z
Af (x) = g(x, y)f (y)dy
M
then call g(x, y) the kernel of A. However, in general, the kernel doesn’t exist.
To see why the kernel in general not exist, we let ξ ∈ ρ(∆). To be more specific, ∆ is an operator
on L2 (M ), so we assume that ξ ∈ ρ(∆2 ). The operator (∆2 − ξ)−1 is called the resolvent. It is a
bounded operator form L2 (M ) → L2 (M ). (By Hahn-Banach theorem, it can be extended to whole
L2 (M )). However, in general, the kernel doesn’t exist, if not, we let fi → δ be a sequence converges
to the δ−function in L1 (M ). Then ∆2 fi could have been bounded. we need estimate to extend ∆2
form L2 (M ) to L1 (M ).
Lemma 2 told us that for α < 0, (∆ − α)−n/2 has a kernel. The operator can be extended to
L1 (M ). Further more, the kernel exponentially decays.
The Laplacian we used here is the geometric Laplacian. That is, it is a positive operator.
For our purpose, we just need to prove σ(∆1 ) ⊂ σ(∆2 ), which is also the major part of the paper
of Sturm.
Recall that x ∈ ρ(A), if (A−xI)−1 is a bounded operator. We define the spectrum σ(A) = R−ρ(A).
Note that σ(∆2 ) ⊂ [0, ∞). Then for α < 0, (∆2 − αI)−1 is bounded. Since
∆ − αI = (∆ − ξI)(1 − (α − ξ)(∆ − ξI)−1 )
we have
(∆ − ξI)−1 = (1 − (α − ξ)(∆ − ξ)−1 )(∆ − αI)−1
or for any n
(∆ − ξI)−n = (1 − (α − ξ)(∆ − ξ)−1 )n (∆ − αI)−n
From the above identity, the kernel for (∆ − ξI)−n exists. Let
A = (1 − (α − ξ)(∆ − ξ)−1 )n
Then
Ax (gα(n/2) (x, y))
is the kernel of (∆ − ξI)−n
Lemma 10.4. Let g(x, y) be the kernel of (∆ − ξI)−n . Then
|g(x, y)| ≤ Ce−d(x,y) ϕ(x)ϕ(y)
1
where ϕ(x) = ν(B1 (x))− 2 .
We omit the proof of the above lemma.
By lemma ?? Z
sup |g(x, y)|dy < +∞
x
Thus (∆2 − ξ)−n is an operator from L1 → L1 .
Since ∆2 and ∆1 are the same acting on C ∞ -functions, (∆1 − ξ)−n is a bounded operator from
L (M ) → L1 (M ).
1
Proof. Since (∆1 − ξ)−n is bounded, there is a neighborhood of ξ such that for any ξ 0 in the neigh-
borhood, (∆1 − ξ 0 )−n is also bounded. By
(∆1 − ξ)−1 = (∆1 − ξ 0 )−1 (1 − (ξ − ξ 0 )(∆1 − ξ 0 )−1 )−1
and the fact that the latter can be expanded to a convergent series, (∆1 − ξ)−1 is bounded in L1 (M ).
Thus ξ ∈ ρ(∆1 ).
From the above lemma, we have
σ(∆1 ) ⊂ σ(∆2 ) ⊂ [0, ∞)
The theorem thus follows from
Theorem 10.6. (Wang) Let M be a complete Riemannian manifold with non-negative Ricci curva-
ture. Then σ(∆1 ) = [0, ∞)
Before giving the proof, we use the following theorem.
Theorem 10.7. Let M be a complete non-compact Riemannian manifold. Ric(M ) ≥ 0. Then
vol Bp (R) ≥ C(n, volBp (1))R
Proof. Fixing x0 ∈ ∂Bp (R), using the comparison theorem we have ∆ρ2 ≤ 2n. For any ϕ ∈ C0∞ (M ),
ϕ ≥ 0, we have
Z Z
2
ϕ∆ρ ≤ 2n ϕ
M M
We choose a standard cut-off function ϕ = ψ(ρ(x)), where
1, 0≤t≤R−1
1
ψ(t) = (R + 1 − t), R−1≤t≤R+1
2
0, t≥R+1
By Stokes theorem
Z Z
2
ϕ∆ρ = − ∇ϕ∇ρ2
M
Z
= −2 ψ 0 ρ|∇ρ|2
Z
= ρ
Bx0 (R+1) \Bx0 (R−1)
In general, Harnack inequality implies Hölder continuity, which is weaker than the differ-
entiable Harnack inequality.
The above result implies Yau’s theorem on L2 harmonic functions. In the sense of distri-
bution, the positive part and the negative part of a harmonic function must be subharmonic.
44 PROFESSOR ZHIQIN LU
(5) Let M be a complete noncompact Riemann surface and let K be the Gauss curvature. We
assume that Z
|K|dVM < ∞.
M
Prove that M is a parabolic manifold.