20.07.2013 Views

RICHARD SCHOEN - SPECTRAL GEOMETRY (MATH 286 ...

RICHARD SCHOEN - SPECTRAL GEOMETRY (MATH 286 ...

RICHARD SCHOEN - SPECTRAL GEOMETRY (MATH 286 ...

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

<strong>RICHARD</strong> <strong>SCHOEN</strong> - <strong>SPECTRAL</strong> <strong>GEOMETRY</strong> (<strong>MATH</strong> <strong>286</strong>)<br />

LECTURE NOTES<br />

NOTES BY OTIS CHODOSH<br />

Contents<br />

1. Classical Eigenvalue Theory 2<br />

1.1. Eigenvalues and Isoperimetric Problems 3<br />

1.2. Questions about Eigenvalues 5<br />

1.3. Upper Bounds on λk 6<br />

1.4. Lower Bounds for λk 8<br />

2. Sharp Lower Bounds on λ1 10<br />

2.1. The Fundamental Gap Problem 14<br />

3. Background on Surfaces and Harmonic Maps 14<br />

3.1. Area and Energy of Maps 15<br />

3.2. Harmonic Maps 16<br />

4. Conformal Balancing of Measures 18<br />

4.1. First Eigenvalue Bounds by Conformal Balancing 19<br />

4.2. Coarse Upper Bounds By Conformal Balancing 23<br />

4.3. Conformal Volume 24<br />

5. Free Boundary Minimal Surfaces and Steklov Eigenvalues 28<br />

5.1. Examples of Steklov-Critical Submanifolds 29<br />

5.2. Structure of Extremal Metrics 30<br />

References 30<br />

Date: February 7, 2013.<br />

1


2 NOTES BY OTIS CHODOSH<br />

For reference, the class website is http://math.stanford.edu/~schoen/math<strong>286</strong>. These notes<br />

are very much a work in progress on my part, caveat emptor.<br />

Please inform me at ochodosh@math.stanford.edu of errors of any magnitude.<br />

1. Classical Eigenvalue Theory<br />

The main reference for this section is [SY94, Chapters 3 and 4]. We’ll first consider Riemannian<br />

manifolds (M n , g). One problem we’ll discuss is maximizing the first eigenvalue of the Laplacian,<br />

λ1(∆g) subject to the area constraint A(g) = 1. Recall that<br />

<br />

(1.1) λ1(g) = min |∇u| 2 <br />

dVg : u 2 <br />

= 1,<br />

This is the first eigenvalue of the Laplacian associated to g<br />

(1.2) ∆gu =<br />

M<br />

1<br />

√ det g<br />

∂<br />

∂x i<br />

M<br />

M<br />

det gg ij ∂u<br />

∂x j<br />

<br />

<br />

u = 0 .<br />

There are several boundary conditions we could associated with this problem, depending on whether<br />

or not ∂M is empty. If ∂M = ∅, then we have an increasing sequence of eigenvalues 0 = λ0 < λ1 ≤<br />

λ2 ≤ . . . (written with multiplicity). If ∂M = ∅, then we have several natural choices of boundary<br />

conditions<br />

Definition 1.1. (Dirichlet Boundary Conditions) Here, we fix u = 0 on ∂M. This results in<br />

eigenvalues<br />

0 < λ D 1 < λ D 2 ≤ λ D 3 . . .<br />

Definition 1.2 (Neumann Boundary Conditions). We consider admissible functions with ∂u<br />

∂ν = 0<br />

on ∂M, where ν is a normal vector along the boundary. This gives eigenvalues<br />

0 = λ N 0 < λ N 1 ≤ λ N 2 ≤ . . .<br />

We will later consider the following eigenvalue problem as well<br />

Definition 1.3 (Steklov Eigenvalues). These are the spectrum of the Dirichlet to Neumann map<br />

where Lu = ∂û<br />

∂ν<br />

L : C ∞ (∂M) → C ∞ (∂M),<br />

(where ν is the outward pointing normal to ∂M) for û solving the Dirichlet problem<br />

<br />

∆û = 0 in M<br />

û = u on ∂M<br />

It is not hard to see that the eigenvalues take the form σ0 = 0 < σ1 ≤ σ2 ≤ . . . .<br />

As briefly mentioned above, we have the following variational characterization of the eigenvalues:<br />

(1) If ∂M = ∅ or we are using Neumann boundary conditions, then<br />

<br />

|∇u| 2<br />

λk = min <br />

M u2<br />

<br />

: u ⊥ u0, . . . , uk−1<br />

(2) For Dirichlet boundary conditions, this is also valid, as long as we only consider u satisfying<br />

u = 0 on ∂M.<br />

(3) The Steklov eigenvalues have a similar characterization<br />

<br />

|∇u| 2 <br />

<br />

σk = min : uui = 0, i = 1, 2, . . . , k − 1 .<br />

u2<br />

∂M


<strong>RICHARD</strong> <strong>SCHOEN</strong> - <strong>SPECTRAL</strong> <strong>GEOMETRY</strong> (<strong>MATH</strong> <strong>286</strong>) LECTURE NOTES 3<br />

1.1. Eigenvalues and Isoperimetric Problems. There are several isoperimetric constants which<br />

we now define<br />

Definition 1.4 (Isoperimetric Constants). We let<br />

<br />

<br />

c1 = inf<br />

|∂Ω|<br />

: Ω ⋐ M, Ω has finite perimeter<br />

|Ω| n−1<br />

n<br />

denote the usual isoperimetric constant. Additionally we define (if ∂M = ∅)<br />

<br />

|∂Ω|<br />

hD(M) = inf : Ω ⋐ M<br />

|Ω|<br />

<br />

<br />

|∂Ω|<br />

hN(M) = inf<br />

: ∂Ω is a hypersurface separating M into two pieces<br />

min{|Ω|, |M\Ω|}<br />

It is not hard to check that if ∂M = ∅, then<br />

<br />

(1.3) c1<br />

M<br />

|u| n<br />

n−1<br />

n<br />

n−1<br />

<br />

≤<br />

M<br />

|∇u|.<br />

Furthermore if ∂M = ∅, then<br />

<br />

hD(M) |u| ≤ |∇u| for all u with u = 0 on ∂M<br />

M M<br />

<br />

hN(M) |u| ≤ |∇u|<br />

M M<br />

for all u ≥ 0 with |{u = 0}| ≥ 1<br />

2 |M|.<br />

(1.4)<br />

These inequalities are ready consequences of the co-area formula<br />

Proposition 1.5 (Co-Area Formula). For M a compact Riemannian manifold, possibly with boundary,<br />

if f ∈ H1 (M), then for any measurable function g on M we have<br />

<br />

∞ <br />

g<br />

g =<br />

dσ.<br />

|∇f|<br />

M<br />

−∞<br />

{f=σ}<br />

Because it is more difficult, we check the second inequality in (1.4). For u ≥ 0, note that<br />

|{u = 0}| ≥ 1<br />

2 |M| implies that |{u ≥ a}| ≤ |{u ≤ a}|. Then, by the co-area formula<br />

<br />

∞ <br />

|∇u| =<br />

1 da<br />

M<br />

0<br />

∞<br />

{u=a}<br />

(1.5)<br />

=<br />

0<br />

|∂{u ≥ a}|da<br />

∞<br />

≥ hN(M) |{u ≥ a}|da<br />

0 <br />

= hN(M) u.<br />

The above isoperimetric inequalities are closely related to the first eigenvalue, as first observed<br />

by Cheeger<br />

Theorem 1.6 (Cheeger). For M compact with nonempty boundary, we have<br />

λ D 1 (M) ≥ 1 2<br />

hD(M)<br />

4<br />

and<br />

λ N 1 (M) ≥ 1<br />

4 hN(M) 2 .<br />

M


4 NOTES BY OTIS CHODOSH<br />

Proof. For the first inequality, we’ll show that for all u with u|∂M = 0 then<br />

<br />

1 2<br />

(1.6)<br />

hD(M) u<br />

4 2 <br />

≤ |∇u| 2 .<br />

To do so, we replace u by u2 in the first equation in (1.4), above. This gives<br />

(1.7)<br />

<br />

hD(M) u 2 <br />

≤<br />

<br />

2|u||∇u| ≤ u 2<br />

1<br />

2 <br />

M<br />

M<br />

M<br />

M<br />

M<br />

|∇u|<br />

M<br />

2<br />

1<br />

2<br />

,<br />

which implies the desired inequality upon rearranging terms.<br />

For the second inequality, we’ll show that for any Lipschitz function f on M<br />

1<br />

(1.8)<br />

4 hN(M)<br />

<br />

|f − f| 2 <br />

≤ |∇f| 2 .<br />

M<br />

We clearly have that |{f < f}| ≤ 1<br />

1<br />

2 |M| and |{f > f}| ≤ 2 |M|. Thus, applying (1.4), we see that<br />

<br />

hN(M) ((f − f)+) 2 <br />

≤ |∇((f − f)+) 2 |<br />

(1.9)<br />

Thus<br />

(1.10)<br />

M<br />

1<br />

4 hN(M)<br />

<br />

<br />

= 2<br />

{f≥f}<br />

{f≥f}<br />

M<br />

(f − f)+|∇f|<br />

<br />

≤ 2 ((f − f)+)<br />

M<br />

2<br />

((f − f)+) 2 <br />

≤<br />

{f≥f}<br />

1 2<br />

|∇f| 2 .<br />

|∇f|<br />

{f≥f}<br />

2<br />

We may similarly show<br />

1<br />

(1.11)<br />

4 hN(M)<br />

<br />

((f − f)−) 2 <br />

≤ |∇f|<br />

{f≤f}<br />

2 .<br />

Adding the two equations, and using the fact that <br />

{f=f} |∇f| = 0, as is easily checked, we may<br />

thus conclude our desired inequality. <br />

Theorem 1.7 (Faber-Krahn). For Ω ⊂ R n a smooth bounded domain, we have that for R > 0 so<br />

that |BR(0)| = |Ω|, then<br />

λ D 1 (Ω) ≥ λ D 1 (BR(0))<br />

with equality if and only if Ω is a round ball.<br />

Proof. Let f be the first eigenfunction, with f > 0 in Ω and <br />

1<br />

2<br />

Ω f 2 = 1. We’ll define a symmetric<br />

rearrangement as follows. Choose a function on BR(0), g(x) which is a function of the radial<br />

variable only, i.e. g(x) = g(|x|), and where we require g = 0 on ∂BR and |{g ≥ c}| = |{f ≥ c}| for<br />

c ≥ 0 (it is clear that we can choose such a g). We claim that<br />

(1) <br />

Ω f 2 = <br />

BR g2<br />

(2) <br />

BR |∇g|2 ≤ <br />

Ω |∇f|2<br />

Taking these for granted for the moment, we see that this finishes the proof, becase<br />

(1.12) λ D <br />

BR<br />

1 (BR) ≤<br />

|∇g|2<br />

<br />

<br />

|∇f| 2 = λ D 1 (Ω).<br />

≤<br />

g2<br />

BR<br />

Now, we prove these two properties. For the first property, we see that<br />

<br />

(1.13)<br />

f 2 ∞<br />

= |{f 2 ∞<br />

≥ c}|dc = |{g 2 <br />

≥ c}|dc = g 2<br />

Ω<br />

0<br />

0<br />

Ω<br />

BR


<strong>RICHARD</strong> <strong>SCHOEN</strong> - <strong>SPECTRAL</strong> <strong>GEOMETRY</strong> (<strong>MATH</strong> <strong>286</strong>) LECTURE NOTES 5<br />

To see the second property, note that because g is radial, it is easy to check that |∇g| is constant<br />

on {g = c}. Thus<br />

(1.14) |{g = c}| =<br />

On the other hand,<br />

(1.15)<br />

<br />

{f=c}<br />

|∇g|<br />

{g=c}<br />

1<br />

2 <br />

{g=c}<br />

<br />

1<br />

|∇f|<br />

{f=c} |∇f| ≥<br />

2 1<br />

{f=c}<br />

= |{f = c}| 2<br />

≥ |{g = c}| 2<br />

<br />

= |∇g|<br />

{g=c}<br />

1<br />

2<br />

1<br />

|∇g|<br />

{g=c}<br />

1<br />

|∇g| .<br />

In the second to last line, we used that the surfaces {f = c} and {g = c} enclose the same area,<br />

and {g = c} is a sphere, so the isoperimetric inequality gives the desired bound. However, by the<br />

co-area formula<br />

(1.16)<br />

<br />

<br />

1 d<br />

d<br />

1<br />

= − |{f ≥ c}| = − |{g ≥ c}| =<br />

{f=c} |∇f| dc dc {g=c} |∇g| ,<br />

so combining the last two steps, we see that<br />

<br />

<br />

(1.17)<br />

|∇f| ≥ |∇g|.<br />

Ω<br />

0<br />

{f=c}<br />

{g=c}<br />

We may finally conclude the proof by using the co-area formula<br />

<br />

(1.18)<br />

|∇f| 2 <br />

∞ <br />

∞ <br />

=<br />

|∇f| dc ≥<br />

|∇g| dc =<br />

{f=c}<br />

0 {g=c}<br />

BR<br />

|∇g| 2 .<br />

We remark that it is easy to see from the proof that equality holds if and only if Ω is a round<br />

ball.<br />

1.2. Questions about Eigenvalues.<br />

Question 1.8. We thus are led to three natural questions<br />

(1) What sort of bounds can we give on the first (low) eigenvalues?<br />

(2) What sort of bounds can we give for high eigenvaluse?<br />

(3) What geometric invariants are relevant?<br />

We remark that for the first eigenvalue, if Ric(g) ≥ (n − 1)k and diam(M) ≤ d then we can<br />

bound λ1 above and below in terms of k and d. Furthermore, the Weyl asymptotics tell us that for<br />

c(n) = (2n) −n ωn (where ωn = |B1|), we have that<br />

(1.19) λk(M) n<br />

2 ∼ c(n) k<br />

|M| .<br />

We further note<br />

Conjecture 1.9 (Polya). It is a classical conjecture by Polya that for Ω ⊂ R 2<br />

and<br />

λ D k<br />

λ N k<br />

(Ω) ≥ c(2) k<br />

|Ω|<br />

(Ω) ≤ c(2) k<br />

|Ω| .


6 NOTES BY OTIS CHODOSH<br />

Remark 1.10. It is known that for Ω ⊂ R n ,<br />

1.3. Upper Bounds on λk.<br />

λk(Ω) n<br />

2 ≥ n k<br />

c(n)<br />

n + 2 |Ω| .<br />

Theorem 1.11. For (M, g) a complete manifold with Ric ≥ (n − 1)k, for x0 ∈ M letting B(x0, r0)<br />

denote the geodesic ball of radius r0 and Vn(k, r0) denote a ball of radius r0 in M n k , the simply<br />

connected manifold with constant sectional curvature k, then we have that<br />

λ D 1 (B(x0, r0)) ≤ λ D 1 (Vn(k, r0))<br />

with equality if and only if B(x0, r0) isometric to Vn(k, r0).<br />

For ϕ(r) the first Dirichlet eigenvalue on Vn(k, r0) (it is radial by symmetry), notice that ϕ(r) > 0<br />

for r < r0 and ϕ ′ (r) ≤ 0 for r ≤ r0. Further note that the eigenvalue equation simplifies because of<br />

the symmetry of ϕ, giving<br />

(1.20)<br />

1<br />

′<br />

′ det g(r)ϕ (r) + λ1ϕ = 0.<br />

det g(r)<br />

The second term is positive, and thus we have that<br />

1<br />

′<br />

′<br />

(1.21)<br />

det g(r)ϕ (r) < 0<br />

det g(r)<br />

for r < r0.<br />

Proof of Theorem 1.11. We let p(x) = dist(x, x0). Consider ϕ ◦ p ∈ Lip(B(x0, r0)) with ϕ ◦ p = 0<br />

on ∂B(x0, r0). We claim that<br />

<br />

(1.22)<br />

|∇(ϕ ◦ p)|<br />

B(x0,r0)<br />

2 ≤ λ D <br />

1 (Vn(k, r0)) (ϕ ◦ p)<br />

B(x0,r0)<br />

2 .<br />

To show this, we will pull the problem back to the tangent bundle by the exponential map (using<br />

expx0 (Br0 (0)) = B(x0, r0)). For each ζ ∈ Tx0M with |ζ| = 1, let a(ζ) ∈ (0, r0] be the maximum<br />

value so that expx0 (tζ) minimizes up to a(ζ). It is a standard fact that a(ζ) is continuous. We now<br />

define Ω = {rζ ∈ Tx0M : r ≤ a(ζ)}. Furthermore, notice that in Ω, p(ξ) = |ξ| is smooth. Now, we<br />

have that<br />

<br />

|∇(ϕ ◦ p)| 2 a(ζ)<br />

=<br />

|∇(ϕ ◦ ρ)| 2 (t, ζ)t n−1 <br />

θ(t, ρ)dt dσ(ζ)<br />

(1.23)<br />

B(x0,r0)<br />

<br />

B(x0,r0)<br />

(ϕ ◦ p) 2 <br />

=<br />

S n−1<br />

S n−1<br />

0<br />

a(ζ)<br />

(ϕ ◦ ρ) 2 (t, ζ)t n−1 <br />

θ(t, ρ)dt dσ(ζ)<br />

0<br />

where θ(t, ζ) is the coefficient of the volume element in these coordinates. By the Volume Comparison<br />

Theorem (cf. [SY94, Theorem 1.2]) we have<br />

−1 ∂θ<br />

(1.24) θ<br />

∂t<br />

Thus<br />

(1.25)<br />

a(ζ)<br />

0<br />

2 dϕ<br />

t<br />

dt<br />

n−1 θ(t, ζ)dt = −<br />

a(ζ)<br />

0<br />

∂θk<br />

≤ θ−1<br />

k ∂t .<br />

<br />

n−1 −1 d<br />

t θ(t, ζ) ϕ t<br />

dt<br />

n−1 θ(t, ζ) dϕ<br />

<br />

t<br />

dt<br />

<br />

n−1 θ(t, ζ)dt<br />

:=I<br />

<br />

<br />

t=a(ζ)<br />

.<br />

t=0<br />

+ ϕ dϕ<br />

dt tn−1θ(t, ζ)<br />

<br />

=0


<strong>RICHARD</strong> <strong>SCHOEN</strong> - <strong>SPECTRAL</strong> <strong>GEOMETRY</strong> (<strong>MATH</strong> <strong>286</strong>) LECTURE NOTES 7<br />

We further compute<br />

(1.26)<br />

<br />

d2ϕ n − 1<br />

I = ϕ +<br />

dt2 t<br />

<br />

d2ϕ n − 1<br />

≥ ϕ +<br />

dt2 t<br />

<br />

dθ dϕ<br />

+ θ−1<br />

dt dt<br />

<br />

dϕ<br />

+ θ−1<br />

k<br />

dθk<br />

dt<br />

dt<br />

<br />

=∆ M n k ϕ<br />

= −λ D 1 (Vn(k, r0))ϕ 2<br />

Inserting this into the above expression gives the desired inequality. Furthermore, if there is equality,<br />

we see that Ω = B(x0, r0) and Jacobi field arguments show that B(x0, r0) has constant curvature.<br />

<br />

We now recall the following general principle<br />

Proposition 1.12. For M compact without boundary and Ω1, . . . , Ωk+1 disjoint domains in M,<br />

we have that<br />

λk(M) ≤ max λ D 1 (Ω1), . . . , λ D 1 (Ωk+1) .<br />

Proof. For ϕ1, . . . , ϕk+1 the first eigenfunctions of Ω1, . . . , Ωk+1, we have that they all have disjoint<br />

support. Now, if {1, u1, . . . , uk−1} are the first k − 1 eigenfunctions for M, then (because it is an<br />

linear algebra problem) we may choose a1, . . . , ak+1 ∈ R (not all zero) so that<br />

k+1<br />

(1.27) ϕ :=<br />

i=1<br />

aiϕi<br />

is orthogonal to the span of {1, u1, . . . , uk−1}. Then, we have that<br />

(1.28)<br />

<br />

M<br />

|∇ϕ| 2 k+1<br />

=<br />

i=1<br />

k+1<br />

=<br />

i=1<br />

a 2 i<br />

<br />

|∇ϕi|<br />

M<br />

2<br />

a 2 i λ D 1 (Ωi)<br />

<br />

ϕ<br />

M<br />

2 i<br />

≤ max<br />

i {λ1(Ωi)}<br />

k+1<br />

= max<br />

i {λ1(Ωi)}<br />

i=1<br />

<br />

M<br />

a 2 i<br />

ϕ 2 ,<br />

<br />

ϕ<br />

M<br />

2 i<br />

which is the desired conclusion. <br />

Using the previous two results, we thus obtain<br />

Theorem 1.13. For M compact and Ric ≥ (n − 1)k, then λm(M) ≤ λ1<br />

Vn<br />

<br />

diam M k, 2m .<br />

d<br />

Proof. If d = diam(M), then we can find m + 1 disjoint balls of radius 2m . Thus, we may apply<br />

the above two results. <br />

Corollary 1.14. For M compact and Ric ≥ (n − 1)k<br />

(1) If k = 0, then<br />

λm(M) ≤ n2 m 2 π 2<br />

d 2


8 NOTES BY OTIS CHODOSH<br />

(2) If k = 1 then<br />

(3) If k = −1 then<br />

λ1 ≤<br />

λ1(M) ≤ nπ2<br />

d 2<br />

(n − 1)2<br />

4<br />

+ 2n(n + 4)<br />

d 2<br />

1.4. Lower Bounds for λk. The following lower bound for the first eigenvalue is classical<br />

Theorem 1.15 (Lichnerowicz). For (M, g) closed, Ric ≥ (n − 1)k (for k > 0) then λ1(M) ≥ nk.<br />

Although we will not prove it, we remark that Obata has proven that equality in the above<br />

theorem implies that M is isometric to a round sphere of sectional curvature k.<br />

Proof. Let u be a first eigenfunction normalized so that <br />

M u2 = 1. Then, we have the following<br />

“Ricci formula” at p ∈ M (letting ei be an orthonormal frame with [ei, ej]|p = 0 at p)<br />

(1.29)<br />

1<br />

2 ∆|∇u|2 = 1<br />

∇ei<br />

∇ei (∇eju)2 2<br />

= ∇ei (∇eju∇ei ∇eju) = |∇ 2 u| 2 + (∇eju)∇ei ∇ei∇ej u<br />

= |∇ 2 u| 2 + (∇eju)∇ei ∇ej∇ei u<br />

= |∇ 2 u| 2 + 〈∇u, ∇∆u〉 + Ric(∇u, ∇u).<br />

We note that until this step, we have not used anything about u or the metric. However, we may<br />

now use our lower Ricci bound, combined with the fact that u is an eigenfunction to see that<br />

1<br />

2 ∆|∇u|2 <br />

<br />

≥ <br />

∇2u − ∆u<br />

n g<br />

<br />

2<br />

<br />

+ 1<br />

n (∆u)2 − λ1(M)|∇u| 2 + (n − 1)k|∇u| 2 (1.30)<br />

.<br />

Integrating this inequality and throwing away the first term on the right 1 , thus gives (using the<br />

fact that <br />

M u2 = 1)<br />

(1.31) 0 ≥ 1<br />

n λ1(M) 2 − λ1(M) 2 + (n − 1)kλ1(M).<br />

Rearranging this gives the desired inequality. <br />

Now we give a lower bound for all of the eigenvalues, which is not sharp but does have the right<br />

asymptotic behavior as compared to the Weyl asymptotics. We’ll do so for M compact of dimension<br />

≥ 3 with ∂M = ∅ and under the Dirichlet boundary conditions, although these restrictions can be<br />

removed by slightly modifying the technique used below. We recall that the isoperimetric inequality<br />

(1.32) c1|Ω| n<br />

n−1 ≤ |∂Ω|<br />

is equivalent to the L1-Sobolev inequality<br />

<br />

(1.33) c1 |ϕ| n<br />

n−1<br />

M<br />

n−1<br />

n<br />

<br />

≤<br />

We will instead consider the L2-Sobolev inequality<br />

(1.34)<br />

<br />

c2 |ϕ| 2n<br />

n−2<br />

n<br />

n−2<br />

<br />

≤<br />

1 It is the vanishing of this term that Obata analyzes<br />

M<br />

|∇ϕ| for ϕ|∂M = 0.<br />

|∇ϕ| 2<br />

for ϕ|∂M = 0,


<strong>RICHARD</strong> <strong>SCHOEN</strong> - <strong>SPECTRAL</strong> <strong>GEOMETRY</strong> (<strong>MATH</strong> <strong>286</strong>) LECTURE NOTES 9<br />

which we easily derive from the L 1 version by plugging in ϕ 2 . This gives the lower bound<br />

(1.35) c2 ≥<br />

2 n − 2 c2 1<br />

n − 1 4 .<br />

We further remark that Croke [Cro80] has shown that if M n has Ric ≥ (n − 1)k, diam ≤ D and<br />

Vol(M) ≥ V , then we have that c1(M) ≥ C(n, d, k, V ).<br />

Theorem 1.16 ([CL81]). For M with nonempty boundary, and of dimension n ≥ 3, we have<br />

λ D k<br />

(M) ≥ c2<br />

e<br />

2<br />

k n<br />

.<br />

|M|<br />

Proof. The proof uses the heat kernel, whose essential properties we now recall. It is a function<br />

H(x, y, t) defined on M × M × (0, ∞) so which the fundamental solution to the heat equation, i.e.<br />

⎧<br />

<br />

⎪⎨<br />

∂<br />

∂t − ∆x H = 0<br />

(1.36)<br />

H|x∈∂M = 0<br />

⎪⎩<br />

limt↘0 H(x, y, t) = δy(x).<br />

In particular, it has the following properties<br />

(1) limt↘0 H(x, x, t) = ∞<br />

(2) H(x, y, t) = H(y, x, t) (this follows from self-adjointness of the heat operator)<br />

(3) H(x, y, t)dy ≤ 1 (because of the Dirichlet boundary conditions)<br />

(4) If ui, λi are the eigenfunctions and eigenvalues of ∆, and we take the ui to be L 2 -orthonormal,<br />

then H(x, y, t) = ∞<br />

i=1 e−λit ui(x)ui(y), where the sum converges (along with the term by<br />

term derivatives) for t > 0.<br />

(5) For 0 < s < t, H(x, y, t) = <br />

We note that the last property implies that<br />

<br />

(1.37) H(x, x, 2t) =<br />

M<br />

H(x, z, s)H(z, y, t − s)dz.<br />

M<br />

(H(x, y, t)) 2 dy.<br />

With this in mind, we compute<br />

<br />

∂<br />

H(x, z, t)<br />

∂t M<br />

2 <br />

= 2 H(x, z, t)<br />

M<br />

∂<br />

(H(x, z, t)) dz<br />

∂t<br />

<br />

= 2 H(x, z, t)∆z (H(x, z, t)) dz<br />

M<br />

= −2 |∇zH(x, z, t)| 2 (1.38)<br />

dz<br />

≤ −2c2<br />

Now, using Hölder’s inequality, we have that<br />

(1.39)<br />

<br />

M<br />

H 2 <br />

dz ≤<br />

H 2n<br />

n+2 H 4<br />

<br />

n+2 ≤<br />

M<br />

<br />

(H(x, z, t))<br />

M<br />

2n<br />

n−2<br />

n<br />

n−2<br />

H 2n<br />

n−2<br />

n+2<br />

n−2<br />

<br />

Combined with the above inequality, this thus shows that<br />

(1.40)<br />

<br />

∂<br />

∂t<br />

H 2 <br />

dz ≤ −2c2<br />

H 2 n+2<br />

n<br />

dz<br />

4<br />

n+2<br />

H<br />

.<br />

<br />

≤<br />

H 2n<br />

n−2<br />

n+2<br />

n−2 .


10 NOTES BY OTIS CHODOSH<br />

Letting f(x, t) = H(x, x, 2t), by (1.37), we thus see that<br />

(1.41)<br />

i=1<br />

∂f<br />

∂t<br />

n+1<br />

≤ −2c1f n<br />

limt↘0 f(x, t) = ∞,<br />

which may be integrated for a fixed x, showing that f(x, t) ≤ 4<br />

nc2t − n<br />

2 . Integrating this in x and<br />

using the fact that H(x, x, 2t) = ∞ i=1 e−2λit (ui(x)) 2 , we thus have that<br />

∞<br />

(1.42)<br />

e −2λit<br />

<br />

4<br />

≤<br />

n c2t<br />

n<br />

− 2<br />

|M|.<br />

Taking tk so that 2λktk = n<br />

2<br />

n<br />

−<br />

(1.43) ke 2 ≤<br />

, we thus see that<br />

∞<br />

e −2λitk<br />

<br />

4<br />

≤<br />

n c2tk<br />

i=1<br />

− n<br />

2<br />

=<br />

c2<br />

λk<br />

− n<br />

2<br />

|M|,<br />

which yields the desired bound. <br />

2. Sharp Lower Bounds on λ1<br />

We’ll now discuss the recent results of Andrews-Clutterbuck, giving sharp lower bounds for the<br />

first eigenvalue on closed manifolds. The main references for this section are the works [AC11,<br />

AC12].<br />

Theorem 2.1 (Andrews-Clutterbuck [AC11, AC12]). If we let<br />

λ1(n, k, D) = inf{λ1(M n , g) : M closed , Ric ≥ (n − 1)k, diam ≤ D},<br />

then<br />

λ1(n, k, D) = µ(n, k, D)<br />

where µ is the first nonzero eigenvalue of the following 1-dimensional operator<br />

1<br />

Lϕ =<br />

(ck) n−1<br />

<br />

(ck) n−1 ϕ ′ ′ ′′<br />

= ϕ − (n − 1)Tkϕ ′<br />

on the interval [−D/2, D/2] with Neumann boundary conditions. Here<br />

⎧<br />

⎪⎨ cos(<br />

ck(τ) =<br />

⎪⎩<br />

√ kτ) k > 0<br />

1 k = 0<br />

cosh( √ −kτ) k < 0.<br />

We have also defined<br />

⎧√<br />

√<br />

⎪⎨ k tan( kτ) k > 0<br />

Tk(τ) = 1<br />

⎪⎩<br />

−<br />

k = 0<br />

√ −k tanh( √ −kτ) k < 0,<br />

in order to have a more convenient expression for L.<br />

We briefly recall some of the history behind this estimate. For k = 0, this was proven by Zhong-<br />

Yang in 1985 [ZY84] (with the sharp constant), but the proof was much more difficult. For k > 0,<br />

we may bound diam ≤ π √ k by Meyer’s Theorem, and as we check below, this recovers Lichnerowicz’s<br />

theorem. However, for a smaller D this is sharper. For k = 0, some progress on this was made<br />

by Bakry-Qian in 2000 [BQ00]. We finally remark that recently Andrews-Ni [AN11] extended this<br />

work to the Bakry-Emry tensor and the associated f-Laplacian.<br />

Example 2.2. If k = 0, then the operator L becomes Lϕ = ϕ ′′ on [−D/2, D/2], so the first<br />

eigenfunction is ϕ(τ) = sin πτ<br />

D<br />

. Thus µ = π 2<br />

D 2 . As such λ1(n, 0, D) = π2<br />

D 2 .


<strong>RICHARD</strong> <strong>SCHOEN</strong> - <strong>SPECTRAL</strong> <strong>GEOMETRY</strong> (<strong>MATH</strong> <strong>286</strong>) LECTURE NOTES 11<br />

Example 2.3. For k > 0, by Meyer’s Theorem, we have that D ≤ π<br />

√ D , as remarked above. Thus,<br />

the operator L becomes<br />

Lϕ = ϕ ′′ − (n − 1) √ k tan( √ kτ)ϕ<br />

so the first eigenfunction is ϕ(τ) = sin( √ kτ) (it is easily seen to be an eigenvalue, and Sturm-<br />

Liouville theory shows that it must be the first one). Thus, we easily see that λ n, k, π<br />

<br />

√ = nk,<br />

D<br />

which is simply Lichnerowicz’s result.<br />

The strategy of the proof is to detect the lowest eigenvalue by knowing how quickly solutions to<br />

the heat equation decay. This is because we may solve<br />

(2.1)<br />

∂u<br />

∂t<br />

= ∆u<br />

u(x, 0) = u0(x)<br />

by expanding u0 = ∞<br />

i=0 aiϕi where ϕi are eigenvalues of the Laplacian. Then, the solution to the<br />

heat equation is u(x, t) = ∞<br />

i=0 e−λit aiϕi. This does not necessarily converge to zero, as λ0 = 0,<br />

but the key idea is that |ϕ(x, t)−ϕ(y, t)| does converge to zero and in fact |ϕ(x, t)−ϕ(y, t)| ≈ e −λ1t .<br />

Thus, the main ingredient in Andrews-Clutterbuck’s proof is showing that<br />

|u(x, t) − u(y, t)| ≤ ce −µt .<br />

for any solution to the heat equation. Then, taking u = e −λ1t ϕ1(x), we see that<br />

|ϕ1(x) − ϕ1(y)| ≤ ce (λ1−µ)t ,<br />

which implies that λ1 ≥ µ.<br />

The main ingredient in the proof of Theorem 2.1 is the following<br />

Theorem 2.4. Assume that (M n , g) is closed with Ric ≥ (n − 1)k and diam ≤ D. Fix a function<br />

ϕ(t) (defined for t ∈ [0, D/2]) with ϕ ′ ≥ 0. Let v(x, y) = 2ϕ<br />

d(x,y)<br />

2<br />

<br />

. Then, the linear operator,<br />

defined on functions M × M → R<br />

L (u) := inf trg(A Hessu) : A ∈ Sym 2 (T ∗ (M × M), A ≥ 0, A = g −1 on TxM and TyM <br />

satisfies<br />

L (v) ≤ 2Lϕ<br />

in the viscosity sense (where both sides are evaluated at d(x, y)/2).<br />

We will define what it means to satisfy the differential inequality “viscosity sense” later, as we<br />

will see that it follows immediately from the proof at smooth points of v. We further remark that<br />

(in coordinates) the assumption on A that we are making is: regarding A as a (symmetric) 2 × 2<br />

matrix of n × n blocks, the on-diagonal blocks are g −1 and the off diagonal blocks need only to be<br />

chosen so that A is symmetric and non-negative. In particular, trg(A Hessu) looks like the Laplacian<br />

on M × M but with additional cross terms.<br />

Proof. Assyme that (x, y) is a smooth point of v, i.e. that there is a unique minimizing geodesic from<br />

x to y, γ(s) for s ∈ [−d/2, d/2] (we will write d = d(x, y)). Pick an orthonormal basis {E (x)<br />

i }n i=1 for<br />

TxM so that E (x)<br />

n = γ ′ (−d/2). We may parallel transport the frame to y, obtaining {E (y)<br />

i }n i=1<br />

orthonormal basis for TyM so that E (y)<br />

n = γ ′ (d/2). These combine to give a basis for T (x,y)M × M.<br />

We fix 2<br />

n−1 <br />

(2.2) A = (E (x)<br />

i , E (y)<br />

i ) ⊗ (E (x)<br />

i , E (y)<br />

i ) + (E (x)<br />

n , −E (y)<br />

n ) ⊗ (E (x)<br />

n , −E (y)<br />

n ).<br />

i=1<br />

2 We are identifying T ∗ M with T M via the metric.<br />

, an


12 NOTES BY OTIS CHODOSH<br />

It is easy to check that A is an allowable choice in the definition of L , and that it suffices to prove<br />

the desired inequality for A. We note that<br />

n−1 <br />

(2.3) trg(A Hessv) = Hessv((E (x)<br />

i , E (y)<br />

i ), (E (x)<br />

i , E (y)<br />

i )) + Hessv((E (x)<br />

n , −E (y)<br />

n ), (E (x)<br />

n , −E (y)<br />

n )).<br />

i=1<br />

Furthermore, from the chain rule<br />

(2.4) Hessv = ϕ ′ Hess d(x,y) + 1<br />

2 ϕ′′ ∇d ⊗ ∇d.<br />

We thus use this to compute trg(A Hessv). We first let σ(r) = (γ(−d/2 + r), γ(d/2 − r)), which<br />

has σ ′ (0) = (E (x)<br />

n , −E (y)<br />

n ). But, d(σ(r)) = d(γ(−d/2 + r), γ(d/2 − r)) = d − 2r. This shows that<br />

Hessd is zero in this direction, and we have that<br />

(2.5) Hessv((E (x)<br />

n , −E (y)<br />

n ), (E (x)<br />

n , −E (y)<br />

n )) = 0 + 1<br />

2 ϕ′′ (−2) −2 = 2ϕ ′′ .<br />

For the other terms, we let σ(r) = (exp x(rE (x)<br />

i ), exp y(rE (y)<br />

i )) and define f(r) = d(σ(r)). Notice that<br />

f(r) ≤ length(γr) where γr is any geodesic between expx(rE (x)<br />

i ) and expy(rE (y)<br />

i )). In particular, we<br />

choose (the choice comes from “transplanting of Jacobi fields” from the constant curvature space)<br />

<br />

.<br />

(2.6) γr(s) = exp γ(s)<br />

−d/2<br />

<br />

rck(s)<br />

ck(d/2) E(γ(s))<br />

i<br />

Because f(r) lies below length(γr) and is tangent at r = 0, we see that<br />

f ′ (0) = d<br />

<br />

<br />

length(γr)<br />

dr r=0<br />

f ′′ (0) ≤ d2<br />

dr2 (2.7)<br />

<br />

<br />

length(γr).<br />

r=0<br />

By the standard second variation of arc length, we see that<br />

d<br />

(2.8)<br />

2<br />

dr2 <br />

<br />

1<br />

length(γr) =<br />

r=0 (ck(d/2)) 2<br />

<br />

d/2 <br />

d<br />

ds ck<br />

2 − (ck) 2 <br />

R(Ei, En, Ei, En) ds.<br />

Summing from i = 1 to n − 1 and using Ric ≥ (n − 1)k, we get that<br />

(2.9)<br />

n−1 <br />

i=1<br />

Hessd((E (x)<br />

i , E (y)<br />

i ), (E (x)<br />

i , E (y)<br />

i )) ≤<br />

n − 1<br />

(ck(d/2)) 2<br />

d/2<br />

−d/2<br />

= −2(n − 1)Tk(d/2)<br />

d<br />

ds ck<br />

2<br />

− (ck) 2 k<br />

The last equality follows from integrating by parts. Combining the above results, the theorem<br />

follows (at smooth points of v).<br />

Now, we recall<br />

Definition 2.5. We say that L (v) ≤ 2Lϕ in the viscosity sense if for any point (x0, y0) ∈ M × M<br />

and a C 2 function ψ(x, y) defined near (x0, y0) with<br />

(1) ψ(x, y) ≤ v(x, y) near (x0, y0) and<br />

(2) ψ(x0, y0) = v(x0, y0),<br />

then L ψ ≤ 2Lϕ at (x0, y0).<br />

However, examining the above proof, it is clear that that this statement holds as well. Thus, we<br />

have proven the theorem. <br />

<br />

ds


<strong>RICHARD</strong> <strong>SCHOEN</strong> - <strong>SPECTRAL</strong> <strong>GEOMETRY</strong> (<strong>MATH</strong> <strong>286</strong>) LECTURE NOTES 13<br />

Figure 1. The top function does not satisfy u ′′ ≤ 0 in the viscosity sense, while<br />

the bottom one does.<br />

We give an example of a viscosity solution to a differential inequality in Figure 1.<br />

Now, to prove Theorem 2.1, we let ϕ1 be the first eigenfunction of L. It is clear that ϕ ′ 1<br />

Furthermore, ϕ(s, t) = e −tµ ϕ1(s) solves<br />

(2.10)<br />

∂ϕ<br />

∂t<br />

= Lϕ<br />

ϕ|t=0 = ϕ1.<br />

We further assume that u(x, t) is a solution to the heat equation on M<br />

(2.11)<br />

∂u<br />

∂t<br />

= ∆u<br />

u|t=0 = u0.<br />

As we have remarked previously, we will finish the proof of Theorem 2.1 if we can show<br />

Claim 2.6. For u and ϕ defined as above, we have that<br />

<br />

d(x, y)<br />

|u(x, t) − u(y, t)| ≤ 2Cϕ , t ,<br />

2<br />

where C is chosen so that the inequality holds at t = 0.<br />

Proof. For ɛ > 0, we consider<br />

<br />

d(x, y)<br />

(2.12) Zɛ(x, y, t) = u(y, t) − u(x, t) − 2Cϕ , t − ɛ(1 + t).<br />

2<br />

Because Zɛ(x, y, 0) ≤ −ɛ < 0, if Zɛ(x, y, t) > 0 for some (x, y, t), then there is (x0, y0) so that<br />

Zɛ(x0, y0, t0) = 0 and Zɛ(x, y, t) < 0 for t < t0. We note that x0 = y0, because Zɛ = −ɛ(1 + t) on<br />

the diagonal. Thus, defining<br />

(2.13) ψ(x, y) = u(y, t0) − u(x, t0) − ɛ(1 + t0),<br />

<br />

we have that ψ(x0, y0) = 2Cϕ<br />

and ψ(x, y) ≤ 2Cϕ<br />

<br />

d(x0,y0)<br />

2 , t0<br />

d(x,y)<br />

2 , t0<br />

> 0.<br />

<br />

for (x, y) near (x0, y0).<br />

Thus, by (the viscosity part of) Theorem 2.4, we have that L (v) ≤ 2CLϕ. However, because v is<br />

of product type, by definition of L , we see that<br />

(2.14) ∆yu(y0, t0) − ∆xu(x0, t0) ≤ 2CLϕ = 2C dϕ<br />

dt<br />

Thus<br />

(2.15)<br />

∂<br />

<br />

<br />

dϕ<br />

(u(y, t) − u(x, t) − ɛ(1 + t)) + ɛ ≤ 2C<br />

∂t (x0,y0,t0) dt ,<br />

or ∂Zɛ<br />

∂t (x0, y0, t0) ≤ −ɛ < 0. This is a contradiction, proving our claim.


14 NOTES BY OTIS CHODOSH<br />

k = 0<br />

k < 0<br />

Figure 2. The manifolds showing sharpness of Theorem 2.1. Notice that the capping<br />

process does not destroy the eigenvalue bounds or the curvature bounds, in<br />

particular because the area of the un-capped surfaces is ∼ a, while the area of the<br />

hemispheres is ∼ a 2 .<br />

Finally, we note that we have only shown that λ1(n, k, D) ≥ µ(n, k, D). To show equality, we<br />

must exhibit (M n i , gi) with Ricgi ≥ (n − 1)k and diam(M n i , gi) ≤ D and λ1(M n i , gi) → µ. We<br />

will briefly discuss the simplest cases k = 0 and k < 0. First, on S n−1 × [−D/2, D/2] we define<br />

ga = ack(s)g S n−1 + ds 2 for some constant a. One may easily check that (if k ≤ 0), this has<br />

Ric ≥ (n − 1)k for any a > 0. Furthermore, one may check that ck(s) used as a test function in<br />

the Rayleigh quotient shows that lim sup a→0 λ D 1 (ga) ≤ µ. Finally, one may cap off the ends with<br />

hemispheres to get a sequence of closed Riemannian manifolds with the desired properties. These<br />

manifolds are illustrated in Figure 2.<br />

2.1. The Fundamental Gap Problem. We briefly remark that the techniques discussed here<br />

were also used by Andrews-Clutterbuck to prove<br />

Theorem 2.7 (The Fundamental Gap Conjecture [AC11]). For Ω ⊂ R n a bounded convex domain<br />

and V a weakly convex potential, define L := ∆ − V . Considering the Dirichlet eigenvalue problem,<br />

we denote the eigenvalues by 0 < λ1 < λ2 ≤ . . . . Then there is a lower bound on the “fundamental<br />

gap”<br />

λ2 − λ1 ≥ 3π2<br />

D 2<br />

where D is the diameter of Ω. Furthermore, this is a sharp bound.<br />

This was a natural conjecture coming from the study of Schrödinger operators, formulated independently<br />

by van den Berg, Ashbaugh-Benguria and Yau around 1983. Various authors obtained<br />

non sharp results, e.g. Wong-Yau obtained a lower bound of π2<br />

4D2 , see [SY94, p. 131] (the paper<br />

[AC11] gives a detailed list of prior works).<br />

3. Background on Surfaces and Harmonic Maps<br />

We will need to recall certain facts about two dimensional geometry. If (M 2 , g) is an oriented<br />

surface with metric g, then the metric and orientation define isothermal coordinates, i.e. coordinates<br />

x 1 , x 2 so that gij = λ(x) 2 δij for some λ(x 1 , x2) > 0. Isothermal coordinates always exist.<br />

Furthermore, given isothermal coordinates, writing z = x 1 + √ −1x 2 , may define a holomorphic<br />

coordinate patch. One may check that the change of coordinates between such patches is conformal<br />

and thus holomorphic.<br />

In particular, we recall<br />

Theorem 3.1 (Uniformization Theorem). If (M 2 , g) is oriented (but not necessarily complete) so<br />

that π1(M) = {1} then (M, g) is conformally equivalent to one of<br />

(1) the sphere S 2 ,


<strong>RICHARD</strong> <strong>SCHOEN</strong> - <strong>SPECTRAL</strong> <strong>GEOMETRY</strong> (<strong>MATH</strong> <strong>286</strong>) LECTURE NOTES 15<br />

(2) the unit disk D (with either the flat metric or the hyperbolic metric—it is irrelevant since<br />

both conformal to the other), or<br />

(3) the complex plane C.<br />

Corollary 3.2. Every Riemann surface has a compatible conformal metric which is complete and<br />

has constant curvature.<br />

3.1. Area and Energy of Maps. Consider F : (M 2 , g) → (N n , h) (which we do not assume to<br />

be a Riemannian immersion). We define the (unsigned) area of (the image of) F to be<br />

<br />

<br />

det(F<br />

(3.1) A(F ) = Area(F ) :=<br />

∗h) det(g) dVg.<br />

Alternatively, choosing coordinates x 1 , x 2 , we may write<br />

<br />

(3.2) A(F ) =<br />

or<br />

<br />

(3.3) A(F ) =<br />

M<br />

M<br />

M<br />

F∗(∂/∂x 1 ) ∧ F∗(∂/∂x 2 )dx 1 ∧ dx 2<br />

F∗(∂/∂x 1 ) 2 F∗(∂/∂x 2 ) 2 − 〈F∗(∂/∂x 1 ), F∗(∂/∂x 2 )〉dx 1 ∧ dx 2 .<br />

We emphasize that this is the unsigned area, so in particular if n = 2 then A(F ) = <br />

We also define the energy of F to be<br />

<br />

(3.4) E(F ) = Energy(F ) := trg(F ∗ h)dVg.<br />

In coordinates x 1 , x 2 , (F ∗ h)ij = F∗(∂/∂x 1 ), F∗(∂/∂x 2 ) , so<br />

(3.5) e(F ) := trg(F ∗ h)<br />

We have<br />

M<br />

2<br />

g ij F∗(∂/∂x 1 ), F∗(∂/∂x 2 ) .<br />

i,j=1<br />

M<br />

| Jac(F )|dV .<br />

Claim 3.3. We have A(F ) ≤ 1<br />

2E(F ) with equality if and only if F is weakly conformal, i.e.<br />

F ∗h = λ2g with λ ≥ 0.<br />

Proof. Let x 1 , x 2 be isothermal coordinates, so gij = u 2 δij. Thus<br />

(3.6) e(F ) = u −2<br />

Furthermore,<br />

2<br />

F∗(∂/∂x i ) 2 .<br />

(3.7) dVg = det(g)dx 1 ∧ dx 2 = u 2 dx 1 ∧ dx 2 .<br />

Combining these two equations shows that<br />

<br />

(3.8) E(F ) =<br />

M<br />

i=1<br />

i=1<br />

2<br />

F∗(∂/∂x i ) 2 dx 1 ∧ dx 2 .


16 NOTES BY OTIS CHODOSH<br />

Thus<br />

(3.9)<br />

<br />

A(F ) =<br />

<br />

≤<br />

≤ 1<br />

2<br />

M<br />

M<br />

<br />

F∗(∂/∂x 1 ) 2 F∗(∂/∂x 2 ) 2 − 〈F∗(∂/∂x 1 ), F∗(∂/∂x 2 )〉dx 1 ∧ dx 2<br />

F∗(∂/∂x 1 ) 2 F∗(∂/∂x 2 ) 2 dx 1 ∧ dx 2<br />

M<br />

i=1<br />

= 1<br />

E(F )<br />

2<br />

2<br />

F∗(∂/∂x i ) 2 dx 1 ∧ dx 2<br />

In the third line, we used the AM-GM inequality. Finally, to check the case of equality, we see<br />

that if equality holds in each step, then F∗(∂/∂x 1 ) and F∗(∂/∂x 2 ) must be orthogonal (by the first<br />

inequality being an equality) and have the same norm (by equality in AM-GM). This easily implies<br />

that F must be weakly conformal. <br />

3.2. Harmonic Maps. As above, we take F : (M 2 , g) → (N n , g). We define<br />

Definition 3.4. The map F is harmonic if F is stationary (i.e a critical point) for E(F ) (under<br />

variations of the map F ).<br />

It is readily computed that the Euler-Lagrange equations in isothermal coordinates take the form<br />

(3.10) ∆F :=<br />

2<br />

i=1<br />

∇ ∂<br />

∂xi F∗<br />

<br />

∂<br />

∂xi <br />

= 0.<br />

We remark that the connection used here is really the Levi-Civita connection on (N, h) pulled back<br />

by F ∗ , i.e. F ∗ (∇ N ), which is a connection on the bundle F ∗ T N → M (clearly this connection ∇<br />

is compatible with the metric h on the bundle F ∗ T N). If we also choose coordinates u 1 , . . . , u n on<br />

N, then writing u α = u ◦ F , we may also write this as<br />

(3.11) ∆δu α +<br />

2<br />

n<br />

Γ<br />

i=1 β,γ=1<br />

α βγ (u(x))∂uβ<br />

∂xi ∂uγ = 0,<br />

∂xi for α = 1, . . . , n. Notice that because the energy functional is metric independent, the Euler-<br />

Lagrange equations are as well (so the Laplacian is taken with respect to the flat metric). To check<br />

(3.11), we note that the Christoffel symbols of ∇ = F ∗ (∇ N ) may be computed as<br />

(3.12)<br />

F ∗ T N Γ γ<br />

iβ<br />

∂<br />

= ∇ ∂<br />

∂uγ = ∇ N<br />

∂x i<br />

∂<br />

∂u β<br />

“<br />

∂<br />

F∗<br />

∂xi ” ∂<br />

∂uβ = ∂uα<br />

∂x i ∇N ∂<br />

∂u α<br />

= ∂uα<br />

∂x i<br />

N Γ δ α,β<br />

∂<br />

∂uβ ∂<br />

,<br />

∂uγ and thus the desired formula follows from the Leibniz rule for connections.<br />

We note<br />

Remark 3.5. The map F is harmonic and conformal if and only if F (M) := Σ is a minimal surface<br />

in N.


<strong>RICHARD</strong> <strong>SCHOEN</strong> - <strong>SPECTRAL</strong> <strong>GEOMETRY</strong> (<strong>MATH</strong> <strong>286</strong>) LECTURE NOTES 17<br />

To see that harmonic and conformal implies minimal, choosing isothermal coordinates we have<br />

that the second fundamental form of Σ is<br />

(3.13) II Σ<br />

<br />

∂<br />

F∗<br />

∂xi <br />

∂<br />

, F∗<br />

∂xj <br />

= ∇ ∂<br />

∂xi <br />

∂<br />

F∗<br />

∂xj ⊥Σ <br />

∂<br />

and if F is conformal, F∗ ∂xi <br />

is an orthogonal basis, so F ∗ (h) = λ2δ. Thus, we have that the<br />

mean curvature of Σ is<br />

<br />

2<br />

(3.14)<br />

H Σ = λ −2<br />

= λ −2<br />

i=1<br />

2<br />

i=1<br />

= λ −2 (∆F ) ⊥Σ<br />

<br />

∇ ∂<br />

∂xi <br />

∂<br />

F∗<br />

∂xi <br />

⊥Σ<br />

∇ ∂<br />

∂xi F∗<br />

<br />

∂<br />

∂xi ⊥Σ where (·) ⊥Σ denotes projection onto the normal bundle of Σ ↩→ N 3<br />

On the other hand, if F is a minimal surface, then choosing g = F ∗h, we have made F into<br />

an isometry (and thus a conformal map). The above computation shows that (∆F ) ⊥Σ = 0, so<br />

we just need to show that the tangential directions vanish as well. However, this always happens,<br />

irrespective of the minimality of F . To see this, we compute (the computation for ∂/∂x2 is identical)<br />

<br />

∂<br />

∆F, F∗<br />

∂x1 <br />

= ∇ ∂<br />

∂x1 <br />

∂<br />

F∗<br />

∂x1 <br />

∂<br />

, F∗<br />

∂x1 <br />

+ ∇ ∂<br />

∂x2 <br />

∂<br />

F∗<br />

∂x2 <br />

∂<br />

, F∗<br />

∂x1 <br />

= 1 ∂<br />

2 ∂x1 <br />

<br />

<br />

F∗ <br />

∂<br />

∂x1 <br />

2 <br />

∂<br />

− F∗<br />

∂x2 <br />

, ∇ ∂<br />

∂x2 <br />

∂<br />

F∗<br />

∂x1 <br />

= 1 ∂<br />

2 ∂x1 <br />

<br />

<br />

F∗ <br />

∂<br />

∂x1 <br />

2 <br />

∂<br />

− F∗<br />

∂x2 <br />

, ∇ ∂<br />

∂x1 <br />

∂<br />

F∗<br />

∂x2 <br />

= 1 ∂<br />

2 ∂x1 <br />

<br />

<br />

F∗ <br />

∂<br />

∂x1 <br />

2<br />

− 1 ∂<br />

2 ∂x1 <br />

<br />

<br />

F∗ <br />

∂<br />

∂x2 (3.15)<br />

<br />

2<br />

= 0.<br />

<br />

∂<br />

The second equality follows from the fact that F is conformal so F∗ ∂x1 <br />

∂ and F∗ ∂x2 <br />

are orthogonal.<br />

The third equality follows from symmetry of the Hessian and finally, conformality is used to<br />

show that<br />

<br />

<br />

(3.16)<br />

<br />

F∗ <br />

∂<br />

∂x1 <br />

<br />

= <br />

F∗ <br />

∂<br />

∂x2 <br />

<br />

.<br />

Two important special cases of harmonic maps are when (N, g) is Euclidean space (R n , δ) or the<br />

round sphere (S n , g). In the first case, we have<br />

Example 3.6. A map F : (M 2 , g) → (R n , δ) has F (M) minimal if and only if ∆F α = 0, for<br />

α = 1, . . . , n and F is conformal.<br />

Example 3.7. In general, a map F : (M 2 , g) → N n ⊂ Rk given in the form F = (F 1 , . . . , F k ) with<br />

F (p) ∈ N for all p ∈ M has the energy functional<br />

k<br />

E(F ) = ∇ N F α 2 dVg<br />

M α=1<br />

3 Note that the normal bundle is naturally a subbundle of F ∗ T N, so our conventions are coherent.


18 NOTES BY OTIS CHODOSH<br />

and Euler-Lagrange equations (∆F ) N = 0 (i.e. the component which is tangent to N vanishes) 4 .<br />

In particular, the Euler-Lagrange equations are equivalent to requiring that ∆F is tangent to N.<br />

In isothermal coordinates, this is equivalent to<br />

2<br />

<br />

∂2F ∆F =<br />

∂xi∂xi ⊥N =<br />

i=1<br />

2<br />

II N<br />

i=1<br />

<br />

∂F ∂F<br />

,<br />

∂xi ∂xi <br />

,<br />

where IIN is the (vector valued) second fundamental form of N ⊂ Rk . In particular, if N = Sn ⊂<br />

Rn+1 , we have that<br />

IIx(v, w) = − 〈v, w〉 x,<br />

so F : M → Sn is harmonic means that<br />

∆δF α 2<br />

<br />

<br />

+ <br />

∂F<br />

∂xi<br />

<br />

2<br />

<br />

F α = 0.<br />

i=1<br />

In particular, if F is conformal as well, then F ∗ (gS n) = λ2 δ, so we easily see that this implies<br />

∆δF α + 2λ 2 F α = 0, or in other words<br />

∆gF α + 2F α = 0.<br />

Thus, we have shown that conformal minimal immersions into a sphere are given in coordinates<br />

with eigenfunctions with respect to the induced metric.<br />

Finally, we remark that given F : (M 2 , g) → (N 2 , h) a conformal map, for u a function on N,<br />

we have that<br />

(3.17) EM(u ◦ F, g) = EN(u, h)<br />

and thus ∆hu = 0 implies that ∆g(u ◦ F ) = 0. To check the above formula, we note that F :<br />

(M, F ∗ (h)) → (N, h) is a (local) isometry, so<br />

(3.18) EN(u, h) = EM(u ◦ F, F ∗ (h))<br />

If we let F ∗ (h) = λ 2 g, then the right hand side is equal to EM(u ◦ F, g) by conformal invariance of<br />

energy.<br />

4. Conformal Balancing of Measures<br />

We will make extensive use of the following result<br />

Theorem 4.1. Let µ be a probability measure on Bn+1 , the closed unit ball. Assume that µ has no<br />

point mass on ∂B, i.e. µ({p}) = 0 for all p ∈ ∂B. Then, there is a conformal diffeomorphism ϕ of<br />

B (extending to a conformal diffeomorphism of the boundary) so that<br />

<br />

(4.1)<br />

ϕ(x)dµ(x).<br />

B n+1<br />

Proof. Associated to any y ∈ Bn+1 , we will choose ϕy : Bn+1 → Bn+1 , a conformal diffeomorphism<br />

(extending to the boundary) with the following three properties: (1) ϕ0 = idB, (2) ϕy(−y) = 0 and<br />

(3) ϕy is a conformal dilation fixing the two points ± y<br />

|y| . It is not hard to see that this uniquely<br />

determines the map ϕy, and an explicit formula is<br />

<br />

1 − |y| 2<br />

∗<br />

(4.2) ϕy(x) = (x + y) + y<br />

|x + y| 2<br />

where z∗ = z<br />

|z| 2 . Alternatively, one may describe ϕy geometrically as the map which is an conformal<br />

inversion in a sphere S−y which is chosen as in Figure 3 to meet ∂Bn+1 orthogonally and at the<br />

4 We emphasize that, contrary to the computations above, we are not projecting onto the tangent and normal<br />

bundle to N ↩→ R k , rather than Σ ↩→ N


<strong>RICHARD</strong> <strong>SCHOEN</strong> - <strong>SPECTRAL</strong> <strong>GEOMETRY</strong> (<strong>MATH</strong> <strong>286</strong>) LECTURE NOTES 19<br />

points of intersection between the plane normal to −y and ∂B n+1 , followed by an inversion with<br />

respect to Sy.<br />

Sy<br />

∂B n+1<br />

y<br />

|y|<br />

y<br />

Figure 3. The map ϕy is simply an inversion with respect to the sphere S−y which<br />

is the unique sphere orthogonal to ∂B n+1 at the points shown (i.e. the points on<br />

∂B n+1 ∩ P where P is the plane through −y which is normal to y), followed by an<br />

inversion with respect to Sy.<br />

Now, we consider the function<br />

<br />

(4.3) F (y) =<br />

B n+1<br />

−y<br />

- y<br />

|y|<br />

ϕy(x)dµ(x).<br />

It is not hard to check that because there are no point masses, F extends continuously to B n+1 and<br />

F |∂B = id∂B. The desired conclusion thus follows from a fixed point argument. <br />

4.1. First Eigenvalue Bounds by Conformal Balancing. Here, we discuss several classical<br />

eigenvalue bounds using Theorem 4.1, the conformal balancing theorem. This first result is due to<br />

J. Hersch in 1970.<br />

Theorem 4.2 ([Her70]). For (M 2 , g) a sphere with an arbitrary metric, we have that<br />

λ1(M 2 , g) Area(M 2 , g) ≤ 8π<br />

with equality if and only (M 2 , g) is a round sphere.<br />

Proof. By the uniformization theorem, there is a conformal map to the round sphere F : (M 2 , g) →<br />

(S2 , g0) ⊂ R3 . Notice that thus, F ∗ (g0) = 1<br />

1<br />

2e(F )g = 2 |∇F |2g. We’ll consider µ := F#(µg), the<br />

pushforward under the map F of the g-volume measure. By the conformal balancing theorem,<br />

there is ϕ : B3 → B3 so that<br />

<br />

(4.4)<br />

ϕ dµ = 0.<br />

We may rewrite this as<br />

(4.5)<br />

<br />

M 2<br />

S 2<br />

ϕ ◦ F dµg = 0,<br />

S−y


20 NOTES BY OTIS CHODOSH<br />

and because ϕ ◦ F is also a conformal map ϕ ◦ F : (M 2 , g) → (S2 , g0), we may assume without loss<br />

of generality that we actually chose F so that<br />

<br />

(4.6)<br />

F dµg = 0.<br />

M 2<br />

Thus, the components of F = (F 1 , F 2 , F 3 ) are orthogonal to the constant functions on M, and are<br />

thus acceptable test functions for λ1(g). As such, we have that<br />

<br />

(4.7) λ1(g) (F i ) 2 <br />

dµg ≤ |∇F i | 2 dµg.<br />

M<br />

Summing over i = 1, 2, 3, we thus have that (because |F | = 1)<br />

<br />

(4.8) λ1(g) Area(g) ≤ |∇F | 2 <br />

dµg = e(F ) dµg = 2A(F ) = 8π.<br />

M<br />

We used the fact that F is conformal in the first equality.<br />

In the case of equality, we see that the F i ’s are first eigenfunctions. By scaling g, we assume<br />

(4.9) ∆gF i = −2F i .<br />

Finally, because |F | = 1,<br />

(4.10)<br />

M<br />

M<br />

0 = 1<br />

∆|F |2<br />

2<br />

= |∇F | 2 + ∆F · F<br />

= |∇F | 2 − 2|F | 2<br />

= |∇F | 2 − 2.<br />

Thus, we see that F ∗ (g0) = 1<br />

2 |∇F |2 g = g, so F is an isometry. <br />

The argument used by Hersch is similar to the argument used by Weinstock in 1954 to prove<br />

Theorem 4.3 ([Wei54]). For Ω ⊂ R 2 simply connected, then<br />

σ1(Ω)L(∂Ω) ≤ 2π<br />

where σ1 is the first (nonzero) Steklov eigenvalue (as defined in Definition 1.3). Furthermore,<br />

equality holds if and only if Ω is a round disk.<br />

Proof. By the Riemann mapping theorem, there is a conformal map F : Ω → D. 5 Letting ds denote<br />

the arc length measure of ∂Ω, by applying the conformal balancing argument as above, we may<br />

assume without loss of generality that<br />

<br />

(4.11)<br />

F ds = 0.<br />

∂Ω<br />

We briefly recall that the variational characterization of the first nonzero Steklov eigenvalue is<br />

<br />

Ω (4.12) σ1(Ω) = inf<br />

|∇u|2<br />

<br />

: u = 0 .<br />

This shows that<br />

<br />

(4.13) σ1(Ω)<br />

∂Ω<br />

∂Ω u2<br />

(F i ) 2 <br />

≤<br />

Ω<br />

∂Ω<br />

|∇F i | 2<br />

5 The possibility of the target being C is precluded by the Liouville theorem, because there are clearly nonconstant<br />

harmonic functions on Ω, obtained by solving a Dirichlet problem.


<strong>RICHARD</strong> <strong>SCHOEN</strong> - <strong>SPECTRAL</strong> <strong>GEOMETRY</strong> (<strong>MATH</strong> <strong>286</strong>) LECTURE NOTES 21<br />

Summing for i = 1, 2 gives<br />

<br />

(4.14) σ1(Ω)L(∂Ω) ≤<br />

Ω<br />

|∇F | 2 = 2 Area(D) = 2π.<br />

The equality case is slightly different. If equality holds, we may assume that the F i are eigenfunctions<br />

with eigenvalue 1. Thus ∆F i ∂F i<br />

= 0 in Ω and ∂ν = F i on ∂Ω. The second fact shows that<br />

∂F<br />

∂ν is a unit vector on ∂Ω. Thus, by conformality<br />

1<br />

(4.15)<br />

2 |∇F |2 <br />

<br />

= <br />

∂F 2<br />

<br />

∂ν = 1<br />

on ∂Ω. However, we would like this on all of Ω. To see this, we apply the maximum principle to<br />

<br />

1<br />

(4.16) ∆ log |∇F |2 = 0,<br />

2<br />

which holds because the log of the modulus of a holomorphic function is always harmonic. <br />

We remark that if we try to generalize this to surfaces with boundary, (M 2 , g), if we let ˆg = λ 2 g<br />

where λ > 0 and λ ≡ 1 on ∂M, then σk(M, g) = σk(M, ˆg). In particular, we define<br />

Definition 4.4. Two surfaces with boundary (M1, g1) and (M2, g2) are σ-isometric if there is a<br />

conformal map F : (M1, g1) → (M2, g2) with 1<br />

2 |∇F |2 = 1.<br />

Clearly in this case, σk(M1, g1) = σk(M2, g2). We may define a slightly more general notion<br />

Definition 4.5. Two surfaces with boundary (M1, g1) and (M2, g2) are σ-homothetic if there is a<br />

conformal map F : (M1, g1) → (M2, g2) with 1<br />

2 |∇F |2 = α > 0 on ∂M, where α is some constant.<br />

Again, under this condition, σk(M1, g1)L(∂M1, g1) = σk(M2, g2)L(∂M2, g2). In particular, using<br />

this definition allows us to generalize (the equality case of) Weinstock’s theorem to surfaces<br />

Theorem 4.6. For (M 2 , g) a simply connected surface,<br />

σ1(M)L(∂M) ≤ 2π<br />

with equality if and only if (M, g) is σ-homothetic to D.<br />

The proof is identical, except we in the equality case all we may conclude is that |∇F | 2 is constant<br />

on ∂M, and thus we have the σ-homothetic property by the same argument as before.<br />

The following theorem due to Szegő (also in 1954) is also proven with a similar, although more<br />

involved, argument, so we discuss it here<br />

Theorem 4.7 ([Sze54]). For Ω ⊂ R 2 a simply connected planar domain, we have a bound on the<br />

first nonzero Neumann eigenvalue of Ω as follows 6<br />

We remark that the right hand side is ≈ 3.4π.<br />

λ N 1 (Ω) Area(Ω) ≤ λ N 1 (D) Area(D).<br />

We remark that on the upper half-sphere S 2 + = S 2 ∩ {z ≥ 0}, the functions x, y restricted to S 2 +<br />

are easily seen to be the first Neumann eigenfunctions, with eigenvalue 2. Thus µ1(S 2 +) Area(S 2 +) =<br />

4π > 3.4π, so this bound cannot generalize to arbitrary surfaces with boundary.<br />

Proof. We recall that by separation of variables, the first Neumann eigenfunctions on D are given<br />

by f(r) cos θ and f(r) sin θ where f(r) is a Bessel function solving the ODE<br />

(4.17) r 2 f ′′ + rf ′ + (r 2 − 1)f = 0<br />

6 One may also characterize the case of equality, but we will not discuss this here


22 NOTES BY OTIS CHODOSH<br />

with boundary condition f ′ (1) = 0, and because we require this to correspond to a smooth eigenfunction<br />

at the origin f ′ (0) = 0. We will later need that (f(r) 2 ) ′ ≥ 0 for r ∈ [0, 1]. To check this,<br />

suppose otherwise. Then, it attains a negative minimum at some point r0 ∈ (0, 1). Thus, at r0,<br />

0 = (f 2 ) ′′ = 2(f ′ ) 2 + 2f ′′ f. Thus at r0,<br />

(4.18) (f ′ ) 2 = −f ′′ f =<br />

f ′<br />

r0f + (r0) 2 − 1<br />

r 2 0<br />

However, this is a contradiction, because both terms on the right hand side are strictly negative.<br />

We define<br />

(4.19) H : D → R 2<br />

(r, θ) ↦→ (f(r) cos θ, f(r) sin θ).<br />

Let F : Ω → D be a conformal map. We claim that by balancing, we may produce a conformal<br />

diffeomorphism ϕ : D → D so that if ψ := H ◦ ϕ, then<br />

<br />

(4.20)<br />

ψ d(F#µ0) = 0,<br />

D<br />

where µ0 is Lebesgue measure on Ω. This is not exactly in the form we proved the balancing<br />

theorem for, but it is clear that the proof also works for this statement. Thus, we may assume that<br />

the conformal map F was chosen so that<br />

<br />

(4.21)<br />

H ◦ F dµ0 = 0.<br />

Ω<br />

Now, we fix Ψ = H ◦ F . The coefficients are thus admissible test functions for the first Neumann<br />

eigenvalue, so we have that<br />

(4.22) λ N <br />

1 (Ω) (Ψ i ) 2 <br />

≤ |∇Ψ i | 2 .<br />

Ω<br />

Summing over i = 1, 2 and using the facts that |H| 2 = f(r) 2 and that the second integral is then<br />

conformally invariant<br />

λ N <br />

1 (Ω) (f ◦ F )<br />

Ω<br />

2 <br />

≤ |∇H|<br />

D<br />

2<br />

= λ N <br />

1 (D) |H|<br />

D<br />

2<br />

= λ N <br />

1 (D) f 2 (4.23)<br />

.<br />

On the other hand,<br />

(4.24)<br />

<br />

Ω<br />

(f ◦ F ) 2 <br />

=<br />

D<br />

Ω<br />

D<br />

f 2 | Jac(F −1 )|.<br />

We may choose coordinates so that F −1 : D → D is given by w = w(z), and thus this is equal to<br />

<br />

(4.25)<br />

f<br />

D<br />

2<br />

<br />

<br />

<br />

∂w <br />

<br />

∂z =<br />

1<br />

f(r)<br />

0<br />

2 2π <br />

r <br />

∂w <br />

<br />

<br />

0 ∂z (reiθ )dθ dr.<br />

<br />

:=G(r)<br />

We claim that <br />

∂w <br />

∂z is subharmonic. To see this, notice that because w is holomorphic<br />

(4.26) 4 ∂2<br />

<br />

<br />

<br />

∂w 2<br />

<br />

<br />

<br />

∂z∂z ∂z = 4 <br />

∂<br />

<br />

2w ∂z2 2<br />

<br />

<br />

≥ 0.


<strong>RICHARD</strong> <strong>SCHOEN</strong> - <strong>SPECTRAL</strong> <strong>GEOMETRY</strong> (<strong>MATH</strong> <strong>286</strong>) LECTURE NOTES 23<br />

By the mean-value property for subharmonic functions, we have that G(r) is non-decreasing with<br />

r. Thus<br />

(4.27)<br />

We thus compute<br />

(4.28)<br />

1<br />

0<br />

r<br />

0<br />

f(r) 2 rG(r)dr =<br />

G(t)tdt = r 2<br />

1<br />

0<br />

d<br />

dr<br />

= f(1) 2<br />

≥<br />

=<br />

1<br />

0<br />

1<br />

0<br />

1<br />

0<br />

r<br />

1<br />

0<br />

= Area(Ω) 1<br />

π<br />

0<br />

G(sr)s ds ≤ r 2<br />

<br />

G(t)t dt<br />

G(t)t dt −<br />

<br />

G(t)t dt f(1) 2 −<br />

<br />

G(t)t dt 2<br />

<br />

|H|<br />

D<br />

2<br />

1<br />

0<br />

G(t)t dt.<br />

1<br />

(f(r)<br />

0<br />

2 ) ′<br />

r<br />

0<br />

1<br />

r<br />

0<br />

2 (f(r) 2 ) ′ dr<br />

1 <br />

0<br />

rf(r) 2 dr<br />

<br />

G(t)t dt dr<br />

We have used (4.27) in the third line (combined with the fact proved earlier that (f(r) 2 ) ′ ≥ 0), and<br />

in the final line, we have used the fact that f(r) 2 = |H| 2 . Thus, combining this with (4.23), yields<br />

the desired result. <br />

We remark that the key here was that <br />

∂w <br />

∂z was subharmonic. By a Bochner formula argument<br />

(cf. [ES64]), one can show that this also holds when Kg ≤ 0, and in particular the same argument<br />

as above yields<br />

Corollary 4.8. For (M 2 , g) with Kg ≤ 0,<br />

λ N 1 (M, g) Area(M, g) ≤ λ N 1 (D) Area(D).<br />

4.2. Coarse Upper Bounds By Conformal Balancing. Here we discuss a generalization of the<br />

conformal balancing argument to bound the first eigenvalue of surfaces with higher genus and/or<br />

number of boundary components. The inequalities will not be sharp (or are not expected to be<br />

sharp) in most cases. The first statement in the following theorem is due to Yang-Yau [YY80] and<br />

the first term in the second statement is due to [FS11]<br />

Theorem 4.9. For M 2 an orientable surface, letting γ denote the genus of M. Then, for any<br />

metric g, we have the eigenvalue bounds<br />

(1) If ∂M = ∅, then<br />

<br />

γ + 3<br />

λ1(M, g) Area(M, g) ≤ 8π .<br />

2<br />

(2) If ∂M = ∅, then letting k = |π0(∂M)| denote the number of boundary components, we have<br />

<br />

<br />

γ + 3<br />

σ1(M, g)L(∂M, g) ≤ min 2π(γ + k), 8π .<br />

2<br />

We remark that if γ = 0 in the first statement, we obtain the Hersch bound, Theorem 4.2, but<br />

for higher genus it is not expected to be sharp. We also remark that a surface of genus γ and with<br />

k boundary components is homeomorphic to a genus k surface with k disks removed.<br />

Proof. To prove statement (1), we choose a conformal branched cover of the round sphere by M,<br />

F : M → S 2 ⊂ R 3 . If the genus γ > 0, then certainly deg(F ) > 1. By conformal balancing, we


24 NOTES BY OTIS CHODOSH<br />

may assume that<br />

(4.29)<br />

<br />

M<br />

F = 0.<br />

Using the components of F as test functions in the Rayleigh quotient and summing yields<br />

<br />

(4.30) λ1(M, g) |F | 2 <br />

≤ |∇F | 2 .<br />

M<br />

Because the right hand side is equal to 2 Area(F ) and this is equal to 8π deg(F ), since we are<br />

considering unsigned area, and almost all points of S 2 are covered deg(F ) times. Thus, we have<br />

(4.31) λ1(M, g) Area(M, g) ≤ 8π deg(F ).<br />

<br />

γ+3<br />

As such, the theorem will follow assuming we can find a cover F with deg(F ) ≤ 2 . This is the<br />

subject of “Brill-Noether theory,” cf. [GH94, p. 261] and [Yan94]<br />

We now will prove the first bound in item (2) of the statement of the theorem, i.e. that<br />

σ1(M, g)L(∂M, g) ≤ 2π(k + γ). A crucial ingredient in the proof is the following result of Ahlfors.<br />

The bound was recently improved by Gabard, so we state this version here<br />

Theorem 4.10 ([Ahl50, Gab06]). For (M 2 , g) a surface of genus γ and with k boundary components,<br />

then there is a conformal branched cover F : M → S 2 with deg(F ) ≤ γ + k.<br />

We fix such a cover F : M → S2 and by balancing pushforward of the ∂M boundary measure<br />

µ0 we may assume that<br />

<br />

(4.32)<br />

F dµ0 = 0.<br />

We may thus apply the variational characterization of eigenvalues to conclude that<br />

(4.33) σ1(M, g)L(∂M, g) ≤ 2 Area(F ) = 2π deg(F ) ≤ 2π(k + γ).<br />

To demonstrate the other bound σ1(M, g)L(∂M, g) ≤ 8π<br />

nents of M with k disks, letting<br />

(4.34)<br />

∂M<br />

M<br />

ˆ M = M ∪ D ∪ · · · ∪ D.<br />

γ+3<br />

2<br />

<br />

, we fill in the boundary compo-<br />

We may extend the metric g into the disks in an arbitrary manner so that ˆg is at least Lipschitz.<br />

Letting µ0 be the measure on ˆ M coming from the boundary measure of ∂M with respect to g, as<br />

in the proof of (1), we may find a conformal branched cover F : M → S2 <br />

with deg(F ) ≤ ,<br />

but in this case we balance F#µ0, so we may assume that F : ˆ M → S2 has<br />

<br />

(4.35)<br />

F dµ0 = 0.<br />

Restricting F to M, the variational characterization argument yields<br />

(4.36) σ1(M, g)L(∂M, g) ≤ 2 Area(F |M) ≤ 2 Area(F ) = 4π deg(F ),<br />

∂M<br />

yielding the desired bound. <br />

4.3. Conformal Volume. We will first need the following result about the conformal group of<br />

S n , which says that other than isometries, the maps ϕy as described in Theorem 4.1 are the only<br />

conformal diffeomorphisms.<br />

Proposition 4.11 (Liouville Theorem). For n ≥ 3, if F : S n → S n a conformal diffeomorphism,<br />

then F = R ◦ ϕy for some y ∈ B n+1 and R some rotation in SO(n).<br />

γ+3<br />

2


<strong>RICHARD</strong> <strong>SCHOEN</strong> - <strong>SPECTRAL</strong> <strong>GEOMETRY</strong> (<strong>MATH</strong> <strong>286</strong>) LECTURE NOTES 25<br />

A proof of this may be found in [SY94, pp. 232-3], where the theorem is proven by considering<br />

the conformal change of Ricci curvature, which yields a PDE constraint on the conformal factor of<br />

such a map. We let G(n) denote the conformal group of S n .<br />

Definition 4.12 (Conformal Volume [LY82]). Suppose that ϕ : M 2 → S n is (weakly) conformal.<br />

We define the conformal volume, due to Li-Yau<br />

Volc(n, ϕ) := sup Area(f ◦ ϕ).<br />

f∈G<br />

and<br />

Volc(n, M) := inf Volc(n, ϕ),<br />

ϕ<br />

where the infimum is taken over ϕ as above.<br />

Using this, we have<br />

Theorem 4.13 ([LY82]). For (M 2 , g) a surface without boundary, we have<br />

λ1(M, g) Area(M, g) ≤ 2 Volc(n, M)<br />

with equality if and only if (after rescaling so that λ1 = 2) there is a isometric minimal immersion<br />

by first eigenfunctions ϕ : M → S n .<br />

Proof. For now, fix a conformal map ϕ : M → Sn . By balancing, we may assume that<br />

<br />

(4.37)<br />

ϕ = 0<br />

and thus<br />

(4.38) λ1(M, g) Area(M, g) ≤ 2 Area(ϕ) ≤ 2 Volc(n, ϕ).<br />

M<br />

Taking the infimum over such ϕ the inequality follows.<br />

For the equality statement, assume that λ1(M, g) = 2. Thus we have that Area(M, g) =<br />

Volc(n, M). Thus, there is a sequence of conformal maps ϕj : M → S n so that <br />

M ϕj = 0 and<br />

(4.39) lim<br />

j→∞ Volc(n, ϕj) = Volc(n, M) = Area(M, g).<br />

Thus<br />

<br />

(4.40) 2 Volc(n, ϕj) ≥ 2 Area(ϕj) = |∇ϕj|<br />

M<br />

2 <br />

≥ 2 |ϕj|<br />

M<br />

2 = 2 Area(g).<br />

As such, we see that ϕj is bounded in W 1,2 (M, Sn ). Extracting a subsequence, we may assume<br />

that ϕj → ϕ weakly in W 1,2 , strongly in L2 and pointwise a.e. We claim that ϕ : M → Sn is an<br />

isometric immersion by eigenfunctions.<br />

First, recall that to show that ϕj converges strongly to ϕ it is enough to show that the norms<br />

converge. Because<br />

<br />

(4.41) Volc(n, ϕj) ≥ |∇ϕj| 2 .<br />

The left hand side converges to 2 Area(g), and by lower semicontinuity of the norm under weak<br />

convergence, we have that<br />

<br />

(4.42) 2 Area(M, g) ≥ |∇ϕ| 2 .<br />

However, because ϕ converges in L2 , we have that <br />

M ϕ = 0 and thus we may use the Poincaré<br />

inequality giving<br />

<br />

(4.43)<br />

|∇ϕ|<br />

M<br />

2 <br />

≥ 2 |ϕ|<br />

M<br />

2 = 2 Area(M, g).<br />

M<br />

M


26 NOTES BY OTIS CHODOSH<br />

Thus, <br />

M |∇ϕ|2 = 2 Area(M, g), so there cannot be energy loss. Thus the ϕj converge in W 1,2 .<br />

Due to this convergence, we have that<br />

<br />

(4.44)<br />

|∇ϕ|<br />

M<br />

2 <br />

= 2 |ϕ|<br />

M<br />

2<br />

which shows that the components of ϕi are eigenfunctions with eigenvalue 2. Finally, using 1 = |ϕ| 2 ,<br />

we have that<br />

(4.45) 0 = 1<br />

2 ∆|ϕ|2 = |∇ϕ| 2 + ϕ · ∆ϕ = |∇ϕ| 2 − 2<br />

Thus |∇ϕ| 2 = 2, and because ϕ is conformal (by the W 1,2 convergence of the ϕj) we have that ϕ<br />

is an isometry. <br />

We now discuss the following theorem due to Li-Yau [LY82]. The rigidity statement is due to<br />

Montiel-Ros [MR85].<br />

Theorem 4.14. If ϕ : M 2 → S n is conformal and minimal then Area(ϕ) = Volc(n, ϕ), i.e.<br />

Area(f ◦ ϕ) ≤ Area(ϕ) for f ∈ G(n). Furthermore, equality holds if and only if f is an isometry,<br />

as long as n ≥ 3 and ϕ(M) ⊆ S n−1 ⊂ S n .<br />

Here, we will only prove the equality Area(ϕ) = Volc(n, ϕ), and not the rigidity statement.<br />

Proof of Theorem 4.14. The proof follows readily from properties of the Willmore functional, so<br />

we define it and briefly discuss its properties. The crucial observation is that if ˚ A is the trace free<br />

second fundamental form for Σ := ϕ(M) ⊂ Sn , then ˚ A2 gd Areag|Σ is invariant under conformal<br />

changes of the background metric g on Sn . Thus, defining<br />

Definition 4.15 (Willmore Functional). For Σ ↩→ Sn , we define<br />

<br />

1<br />

W (Σ) :=<br />

4 ˚ A 2 <br />

+ K d Area .<br />

where K is the Gaussian curvature of Σ.<br />

Σ<br />

Because the second term is a topological invariant by Gauss-Bonnet, the Willmore function is<br />

thus conformally invariant. We note that W (S2 ) = 4π. On the other hand, using the Gauss<br />

equations, one may easily check that<br />

(4.46)<br />

1<br />

4 ˚ A 2 + K = H 2 + 1,<br />

where H is the mean curvature of Σ. Thus, we have<br />

(4.47) W (Σ) ≥ Area(Σ),<br />

with equality if and only if Σ is minimal.<br />

Thus, for any f ∈ G(n), we have that<br />

(4.48) Area(f ◦ ϕ) ≤ W (f ◦ ϕ) = W (ϕ) = Area(ϕ),<br />

as desired. <br />

Corollary 4.16. If ϕ : M 2 → S n is an isometric and minimal embedding by first eigenfunctions,<br />

then Vol(M) = Volc(n, M).<br />

Proof. We may assume that λ1(M, g) = 2. By the Li-Yau eigenvalue estimate, Theorem 4.13,<br />

Area(M) ≤ Volc(n, M). On the other hand, by definition Volc(n, M) ≤ Volc(n, ϕ) = Area(M) by<br />

the previous theorem. <br />

Thus, we may compute the conformal volume in a few cases<br />

Example 4.17. The simplest example is Volc(S 2 , n) = 4π for n ≥ 2.


<strong>RICHARD</strong> <strong>SCHOEN</strong> - <strong>SPECTRAL</strong> <strong>GEOMETRY</strong> (<strong>MATH</strong> <strong>286</strong>) LECTURE NOTES 27<br />

Example 4.18. Another example is given by the projective plane, where we have Volc(RP 2 , n) =<br />

6π for n ≥ 4. To see this, we recall the Veronese embedding. First, we embed ϕ : S 2 → S 4 ⊂ R 5 by<br />

second eigenfunctions<br />

ϕ(x, y, z) =<br />

<br />

xy, xz, yz, 1<br />

2 (x2 − y 2 ), 1<br />

2 √ 3 (x2 + y 2 − 2z 2 <br />

) .<br />

One may check that |ϕ| = 1 and that this map is conformal with |∇ϕ| 2 constant, so it is an embedding.<br />

On the other hand it is manifestly harmonic, because the coordinates are all eigenfunctions<br />

on S 2 . Because it is even, the map descends to RP 2 as a minimal embedding by first eigenfunctions,<br />

so the conclusion easily follows the fact that then 2 Volc(RP 2 , n) = λ1(RP 2 , n) Area(RP n ) = 6 · 2π.<br />

In particular, the second example shows that we have solved the Hersch problem on RP 2 , because<br />

Theorem 4.13 then gives a sharp estimate for the first eigenvalue.<br />

We can also compute the conformal volume for certain conformal classes of tori, starting with<br />

Example 4.19 (Square Torus). The Clifford torus is a square torus S 1 × S 1 inside of S 3 defined<br />

by ϕ(x, y) = (e 2πix , e 2πiy ) ∈ S 3 ⊂ C 2 . One may check that this is an conformal minimal embedding<br />

by first eigenfunctions and Area(ϕ) = 2π 2 . Thus, we see that Volc(square torus) = 2π 2 .<br />

Example 4.20 (Rhombic Torus). The 60◦ rhombic torus is the conformal class containing the flat<br />

torus R2 <br />

/ (1, 0), ( 1<br />

2 ,<br />

√<br />

3<br />

2 )<br />

<br />

. There is an conformal minimal embedding into S5 ⊂ C3 is given by<br />

ϕ(x, y) = 1<br />

<br />

√ e<br />

3<br />

4πi y √<br />

3 , e 2πi<br />

“<br />

x− y ”<br />

√<br />

3 , e 2πi<br />

“<br />

x+ y ”<br />

√<br />

3<br />

One may check that Area(ϕ) = 4π2<br />

√ 3 > 2π 2 , so Volc(rhombic torus) = 4π2<br />

√ 3 .<br />

In fact, as we’ll hopefully discuss later, for any metric g in any conformal class on T 2 , we have<br />

that λ1(g) Area(g) ≤ 2 4π2 √ , which is sharp by the above example.<br />

3<br />

Recall that for Σ 2 ⊂ S n , we have that W (Σ) ≥ Volc(n, Σ), where W (·) is the Willmore functional<br />

as defined in Definition 4.15. In particular, we have<br />

Conjecture 4.21 (Willmore Conjecture). For Σ 2 ⊂ S n is homeomorphic to a torus, W (Σ) ≥ 2π 2 .<br />

This has been recently proven by Marques-Neves when n = 3 [MN12]. In fact, as part of their<br />

proof they have shown that for Σ ⊂ S 3 of genus > 0, then letting Σt := {p : d(p, Σ) = t} for<br />

t ∈ [−π, π], then<br />

(4.49) 2π 2 ≤ sup<br />

t∈[−π,π]<br />

Volc(Σt) ≤ W (Σ).<br />

We now the recall Theorem 4.14, which says that for ϕ : M 2 → S n which is conformal and<br />

minimal, then for all F ∈ G, A(ϕ) ≥ A(F ◦ ϕ) with equality only if F is an isometry (assuming that<br />

n ≥ 3 and ϕ(M) ⊆ S n−1 ⊂ S n ). This rigidity statement will be crucial in the following theorem<br />

Theorem 4.22. Assume that ϕ : M 2 → S n is a conformal minimal embedding by first eigenfunctions<br />

(and that n ≥ 3 and ϕ(M) ⊆ S n−1 ⊂ S n ). Then, every conformal automorphism of M is an<br />

isometry.<br />

Proof. Suppose we have some f : M → M which is a conformal diffeomorphism. We consider<br />

ϕ ◦ f : M → Sn , which is still a conformal minimal embedding. We claim that the components of<br />

the map ϕ ◦ f are first eigenfunctions. To see this, by conformal balancing, there is F ∈ G so that<br />

<br />

(4.50)<br />

F ◦ ϕ ◦ f = 0<br />

M


28 NOTES BY OTIS CHODOSH<br />

and thus<br />

(4.51)<br />

2 Volc(n, M) = 2 Area(ϕ) = 2 Area(ϕ ◦ f)<br />

≥ 2 Area(F ◦ ϕ ◦ f) = E(F ◦ ϕ ◦ f)<br />

≥ λ1(g) Area(M, g) = 2 Volc(n, M).<br />

The second inequality follows because we have balanced ϕ ◦ f, so the components of F ◦ ϕ ◦ f are<br />

admissible test functions in for the first eigenvalue. Furthermore, by Theorem 4.14. equality in the<br />

first inequality holds if and only if F is an isometry. Thus, because equality must hold, we know<br />

that F is an isometry, and thus the components of ϕ ◦ f are first eigenfunctions.<br />

Now, by the same argument as in the end of the proof of Theorem 4.13, this readily implies that<br />

f is an isometry. <br />

5. Free Boundary Minimal Surfaces and Steklov Eigenvalues<br />

Now, we discuss related estimates for surfaces with boundary. It will be crucial for us to discuss<br />

free boundary minimal submanifolds of the ball B n . The simplest example of a free boundary<br />

minimal submanifold is the cone over a minimal submanifold of S n−1 , as in Figure 4. The crucial<br />

ν = x<br />

S n−1<br />

C(Σ k )<br />

Figure 4. The cone over Σ k , a minimal submanifold of S n−1 is an example of a<br />

free boundary minimal surface.<br />

observation is that in this case, x 1 , . . . , x n are Steklov eigenvalues because ∆x i = 0 in C(Σ k ) and<br />

Dνx = x. In general, note that this holds for<br />

Definition 5.1 (Free Boundary Minimal Submanifold of B n ). A minimal submanifold of B n is a<br />

free boundary minimal submanifold if its boundary lies on S n−1 and it meets S n−1 orthogonally.<br />

In particular, just as in the boundaryless case, if M is a surface with nonempty boundary ∂M,<br />

then for ϕ : M → B n (for n ≥ 3) a proper 7 , conformal map, we can define the conformal volume of<br />

ϕ to be<br />

(5.1) Volc(n, ϕ) = sup Area(F ◦ ϕ).<br />

F ∈G<br />

We remark that the conformal group of B n is the same as that of the ball. In particular, by using<br />

the usual balancing arguments we have discussed above, one may show<br />

7 Note that in this case, this is just the requirement that exactly ∂M is mapped to S n−1 .<br />

Σ k


<strong>RICHARD</strong> <strong>SCHOEN</strong> - <strong>SPECTRAL</strong> <strong>GEOMETRY</strong> (<strong>MATH</strong> <strong>286</strong>) LECTURE NOTES 29<br />

Theorem 5.2 (cf. [FS11, Corollary 6.3]). For ϕ : M → S n as above, we have that<br />

σ1(g)L(∂M, g) ≤ Volc(n, ϕ)<br />

with equality if and only if ϕ is a free boundary minimal surface which is isometric (homothetic)<br />

on ∂M.<br />

This suggests the following question in the spirit of Theorem 4.14<br />

Question 5.3. If Σ is a free boundary minimal surface and F ∈ G, is it true that |F (Σ)| ≤ |Σ|?<br />

This seems unlikely to be true, but the following seems more likely<br />

Conjecture 5.4 (Schoen). If Σ k ⊂ B n is a free boundary minimal surface and F : B n → B n is a<br />

conformal diffeomorphism of the ball, then |F (∂Σ)| ≤ |∂Σ|.<br />

5.1. Examples of Steklov-Critical Submanifolds. We can describe several examples of submanifolds<br />

of B n which maximize the first Steklov eigenvalue in the sense of Theorems 4.9 and 5.2.<br />

Let M be of genus γ and having k boundary components. We define<br />

(5.2) σ ∗ (γ, k) := sup{σ1(g)L(∂M, g) : g a smooth metric on M}.<br />

By Theorem 4.9, this is a finite quantity. A natural question is<br />

Question 5.5. Is σ ∗ (γ, k) achieved by a smooth metric? If so, by what?<br />

The simplest case in which we may answer this affirmatively is<br />

Example 5.6. The flat disk achieves σ ∗ (0, 1) = 2π.<br />

The next case which the maximizer is understood is<br />

Example 5.7 (The Critical Catenoid). When γ = 0, k = 2, σ(0, 2) ∗ is achieved by the (unique<br />

up to rotations) catenoid in B 3 which meets S 2 orthogonally. See Figure 5. Recall that a catenoid<br />

Figure 5. The critical catenoid, a free boundary minimal submanifold of B 3 .<br />

may be described by rotating x = cosh z around the z-axis. In particular, the piece in B 3 may be<br />

parametrized by<br />

ϕ : [−T, T ] × S 1 → B 3<br />

ϕ(t, θ) = (cosh t cos θ, cosh t sin θ, t).<br />

Here, T is a uniquely determined by the requirement that the catenoid meet S 2 orthogonally. One<br />

may check that this is a conformal harmonic map consisting of Steklov first eigenfunctions.


30 NOTES BY OTIS CHODOSH<br />

We also may describe<br />

Example 5.8 (The Critical Möbius Strip). While we have mostly focused on orientable surfaces in<br />

the above discussion, we also remark that for the Möbius band R × S 1 /(t, θ) ∼ (−t, θ + π), defining<br />

the map<br />

(5.3) ϕ(t, θ) = (2 sinh t cos θ, 2 sinh t sin θ, cosh 2t cos 2θ, cosh 2t, sin 2θ),<br />

there is a unique T0 so that ϕ| [−T0,T0]×S 1 is a free boundary minimal embedding in to B4 r for some<br />

radius r (we may then rescale the entire picture so that r = 1).<br />

It seems that one should be able to use gluing techniques to desingularize the union of the critical<br />

catenoid and a plane, to obtain free boundary minimal surfaces with three boundary components<br />

and of high genus, but this has not been carried out.<br />

Finally, we remark that Fraser-Schoen have constructed explicit maximizers of σ(0, k) [FS11].<br />

Which look like two planes joined by necks on the boundary, which increase the number of boundary<br />

components but not the genus. Their construction shows that σ(0, k) ∗ is strictly increasing with k<br />

and limk→∞ σ(0, k) ∗ = 4π.<br />

5.2. Structure of Extremal Metrics. In fact, it turns out that maximizers of the scale invariant<br />

first eigenfunction must be minimally embedded by first eigenfunctions, i.e.<br />

Theorem 5.9. For (M 2 , g) with ∂M = ∅, if λ1(g) Area(g) = λ ∗ (M) = sup{λ1(g) Area(g)}, then<br />

there are ϕ1, . . . , ϕn linearly independent first eigenfunctions so that they combine to give a map<br />

ϕ : M → S n−1 which is a homothety and a conformal minimal embedding.<br />

For the boundary case, we have the similar<br />

Theorem 5.10. For (M 2 , g) with ∂M = ∅, if σ1(g)L(∂M, g) = σ ∗ (M) = sup{σ1(g)L(∂M, g)},<br />

then there are ϕ1, . . . , ϕn first Steklov eigenfunctions so that they give a map ϕ : M → B n which<br />

is conformal and harmonic and ϕ(M) is a free boundary minimal surface. Furthermore, g is σhomothetic<br />

to ϕ ∗ (δ).<br />

References<br />

[AC11] Ben Andrews and Julie Clutterbuck, Proof of the fundamental gap conjecture, J. Amer. Math. Soc. 24 (2011),<br />

no. 3, 899–916. MR 2784332 (2012d:35051)<br />

[AC12] Ben Andrews and Julie Clutterbuck, Sharp modulus of continuity for parabolic equations on manifolds and<br />

lower bounds for the first eigenvalue, http://arxiv.org/abs/1204.5079 (2012).<br />

[Ahl50] Lars L. Ahlfors, Open Riemann surfaces and extremal problems on compact subregions, Comment. Math.<br />

Helv. 24 (1950), 100–134. MR 0036318 (12,90b)<br />

[AN11] Ben Andrews and Lei Ni, Eigenvalue comparison on bakry-emery manifolds, http://arxiv.org/abs/1111.<br />

4967 (2011).<br />

[BQ00] Dominique Bakry and Zhongmin Qian, Some new results on eigenvectors via dimension, diameter, and Ricci<br />

curvature, Adv. Math. 155 (2000), no. 1, 98–153. MR 1789850 (2002g:58048)<br />

[CL81] Shiu Yuen Cheng and Peter Li, Heat kernel estimates and lower bound of eigenvalues, Comment. Math.<br />

Helv. 56 (1981), no. 3, 327–338. MR 639355 (83b:58076)<br />

[Cro80] Christopher B. Croke, Some isoperimetric inequalities and eigenvalue estimates, Ann. Sci. École Norm. Sup.<br />

(4) 13 (1980), no. 4, 419–435. MR 608287 (83d:58068)<br />

[ES64] James Eells, Jr. and J. H. Sampson, Harmonic mappings of Riemannian manifolds, Amer. J. Math. 86<br />

(1964), 109–160. MR 0164306 (29 #1603)<br />

[FS11] Ailana Fraser and Richard Schoen, The first Steklov eigenvalue, conformal geometry, and minimal surfaces,<br />

Adv. Math. 226 (2011), no. 5, 4011–4030. MR 2770439 (2012f:58054)<br />

[Gab06] Alexandre Gabard, Sur la représentation conforme des surfaces de Riemann à bord et une caractérisation<br />

des courbes séparantes, Comment. Math. Helv. 81 (2006), no. 4, 945–964. MR 2271230 (2007k:14122)<br />

[GH94] Phillip Griffiths and Joseph Harris, Principles of algebraic geometry, Wiley Classics Library, John Wiley &<br />

Sons Inc., New York, 1994, Reprint of the 1978 original. MR 1288523 (95d:14001)<br />

[Her70] Joseph Hersch, Quatre propriétés isopérimétriques de membranes sphériques homogènes, C. R. Acad. Sci.<br />

Paris Sér. A-B 270 (1970), A1645–A1648. MR 0292357 (45 #1444)


<strong>RICHARD</strong> <strong>SCHOEN</strong> - <strong>SPECTRAL</strong> <strong>GEOMETRY</strong> (<strong>MATH</strong> <strong>286</strong>) LECTURE NOTES 31<br />

[LY82] Peter Li and Shing Tung Yau, A new conformal invariant and its applications to the Willmore conjecture and<br />

the first eigenvalue of compact surfaces, Invent. Math. 69 (1982), no. 2, 269–291. MR 674407 (84f:53049)<br />

[MN12] Fernando C. Marques and André Neves, Min-max theory and the willmore conjecture, http://arxiv.org/<br />

abs/1202.6036 (2012).<br />

[MR85] Sebastián Montiel and Antonio Ros, Minimal immersions of surfaces by the first eigenfunctions and conformal<br />

area, Invent. Math. 83 (1985), no. 1, 153–166. MR 813585 (87d:53109)<br />

[SY94] R. Schoen and S.-T. Yau, Lectures on differential geometry, Conference Proceedings and Lecture Notes in<br />

Geometry and Topology, I, International Press, Cambridge, MA, 1994, Lecture notes prepared by Wei Yue<br />

Ding, Kung Ching Chang [Gong Qing Zhang], Jia Qing Zhong and Yi Chao Xu, Translated from the Chinese<br />

by Ding and S. Y. Cheng, Preface translated from the Chinese by Kaising Tso. MR 1333601 (97d:53001)<br />

[Sze54] G. Szegő, Inequalities for certain eigenvalues of a membrane of given area, J. Rational Mech. Anal. 3 (1954),<br />

343–356. MR 0061749 (15,877c)<br />

[Wei54] Robert Weinstock, Inequalities for a classical eigenvalue problem, J. Rational Mech. Anal. 3 (1954), 745–753.<br />

MR 0064989 (16,368c)<br />

[Yan94] Kichoon Yang, Meromorphic functions, divisors, and projective curves: an introductory survey, J. Korean<br />

Math. Soc. 31 (1994), no. 4, 569–608. MR 1312529 (96a:14031)<br />

[YY80] Paul C. Yang and Shing Tung Yau, Eigenvalues of the Laplacian of compact Riemann surfaces and minimal<br />

submanifolds, Ann. Scuola Norm. Sup. Pisa Cl. Sci. (4) 7 (1980), no. 1, 55–63. MR 577325 (81m:58084)<br />

[ZY84] Jia Qing Zhong and Hong Cang Yang, On the estimate of the first eigenvalue of a compact Riemannian<br />

manifold, Sci. Sinica Ser. A 27 (1984), no. 12, 1265–1273. MR 794292 (87a:58162)

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!