77
The Geometry of Maximum Principles and a Bernstein Theorem in Codimension 2 Der Fakult¨ at f¨ ur Mathematik und Informatik der Universit¨ at Leipzig eingereichte DISSERTATION zur Erlangung des akademischen Grades DOCTOR RERUM NATURALIUM (Dr.rer.nat.) im Fachgebiet Mathematik vorgelegt von Renan Assimos Martins geboren am 22.05.1989 in Rio de Janeiro, Brasilien Leipzig, den 08.03.2019

The Geometry of Maximum Principles and a Bernstein Theorem ... · a minimal p-submanifold in Rp+2, which is the graph of a smooth function de ned on the entire Rp with bounded slope,

  • Upload
    others

  • View
    2

  • Download
    0

Embed Size (px)

Citation preview

The Geometry of Maximum Principles

and a Bernstein Theorem

in Codimension 2

Der Fakultat fur Mathematik und Informatik

der Universitat Leipzig

eingereichte

D I S S E R T A T I O N

zur Erlangung des akademischen Grades

DOCTOR RERUM NATURALIUM

(Dr.rer.nat.)

im Fachgebiet

Mathematik

vorgelegt

von Renan Assimos Martins

geboren am 22.05.1989 in Rio de Janeiro, Brasilien

Leipzig, den 08.03.2019

Bibliographische Daten

The Geometry of Maximum Principles and a Bernstein Theorem in Codimension

Two

(Die Geometrie der Maximumprinzipien und ein Bernstein Theorem in Kodimen-

sion Zwei )

Assimos Martins, Renan

Universitat Leipzig, Dissertation, 2019

77 Seiten, 6 Abbildungen, 32 Referenzen

Abstract

We develop a general method to construct subsets of complete Riemannian

manifolds that cannot contain images of non-constant harmonic maps from

compact manifolds. We apply our method to the special case where the harmonic

map is the Gauss map of a minimal submanifold and the complete manifold

is a Grassmannian. With the help of a result by Allard [All72], we can study

the graph case and have an approach to prove Bernstein-type theorems. This

enables us to extend Moser’s Bernstein theorem [Mos61] to codimension two, i.e.,

a minimal p-submanifold in Rp+2, which is the graph of a smooth function defined

on the entire Rp with bounded slope, must be a p-plane.

Keywords: Bernstein theorem, minimal graph, harmonic map, Grassmannian,

Gauss map, maximum principle.

v

Acknowledgements

There are a number of people that I would like to thank, who helped me with

this thesis, with my PhD, and with my mathematical work in general:

• I would like to thank Prof. Jurgen Jost for the trust, the support, the

patience and for sharing with me his knowledge on the analytic aspects of

geometry during these years.

• I would especially like to thank Ruijun Wu who has spent several hours

with me explaining from basic analysis to bubbling phenomena and beyond.

After the group discussions, he would have the patience to explain me again

the comments and insights given by Prof. Jost which I could not understand.

• I would also like to thank for all the interesting discussions Kostas Zemas,

who basically made me like PDE even more than geometry (I have never

said that!). Together with him, I thank the new brothers the Max Planck

Institute and the apartment in Stieglitzstarsse 89 has given me: Paolo

Perrone, Ramon Urquijo Novella and Sharwin Rezagholi.

• Among MPI friends, I would like to thank Caio Teodoro and Vanessa

Brandao for joining us on Sundays for guitar playing, caipirinha drinking

and a lot of feijao tropeiro. This has literally given us a lot of energy to work

on the next day! I thank Niccolo Pederzani for his friendship and for being

a honorary member of our apartment as well as for all the insane amount

of cooking he has done. I thank Stefano Steffenoni for all the cocktails he

has made me prepare for him and for being the best caipirinha assistant of

all times.

• I thank Gerardo Sosa, Giovanni Frigeri, Gian Luca Lauteri, (Florio) Maria

Ciaglia, Souhayl Sadik, Pau Aceituno and all the MPI people that have

shared very instructive (or ignorant) moments during those years.

• I would like to thank all the other members of the Max-Planck Institute that

have helped me during my PhD years. Specially Lei Liu, Jingyong Zhu and

all the other members of the group of geometric analysis. Special thanks to

Slava Matveev, who has contributed to the mathematical development of

every single member of the MPI in an immeasurable way.

vii

• I would like to thank Carlos Santana and Eric Clapton for providing the

necessary (and sufficient) soundtracks during the preparation of this work.

• I would like to thank Antje Vandenberg for all the precious support, time,

and patience which she gave to me, and to all the people in our department.

• Last, but not least, I want to immensely thank my family, my girlfriend, my

friends from Brazil, my professors from UFRJ, for the support, the patience

(not my girlfriend here), the amazing times and everything.

Renan Assimos.

viii

Contents

Title page . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . i

Bibliographische Daten (Bibliographical Data) . . . . . . . . . . . . . . iii

Abstract . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . v

Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . vii

Contents ix

1 Introduction 1

1.1 Minimal Surfaces . . . . . . . . . . . . . . . . . . . . . . . . . . . 1

1.1.1 Bernstein’s Theorem . . . . . . . . . . . . . . . . . . . . . 4

1.2 Harmonic maps . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6

1.2.1 The harmonic map flow . . . . . . . . . . . . . . . . . . . 10

1.3 Minimal Submanifolds . . . . . . . . . . . . . . . . . . . . . . . . 18

1.4 Main Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22

1.4.1 Domains that do not admit images of non-constant har-

monic maps . . . . . . . . . . . . . . . . . . . . . . . . . . 22

1.4.2 Bernstein-type theorems in higher codimensions . . . . . . 24

2 Preliminaries 29

2.1 Maximum principles . . . . . . . . . . . . . . . . . . . . . . . . . 29

2.2 Sampson’s maximum principle . . . . . . . . . . . . . . . . . . . . 32

2.3 Grassmannian Manifolds . . . . . . . . . . . . . . . . . . . . . . . 33

2.3.1 Closed geodesics in G+p,n. . . . . . . . . . . . . . . . . . . . 37

2.3.2 Geodesically convex sets in G+p,n . . . . . . . . . . . . . . . 39

2.3.3 The slope of a graph and the Gauss map . . . . . . . . . . 42

3 Proof of main results 45

3.1 The Geometry of the Maximum Principle . . . . . . . . . . . . . . 45

3.1.1 An application of Sampson’s maximum principle . . . . . . 45

3.1.2 Barriers to the existence of harmonic maps . . . . . . . . . 49

3.2 Bernstein-type theorems for codimension 2 . . . . . . . . . . . . . 52

ix

CONTENTS

4 Open problems 57

4.1 Convex supporting domains . . . . . . . . . . . . . . . . . . . . . 57

4.2 Bernstein theorems for codimension c ≥ 3 . . . . . . . . . . . . . 58

4.2.1 Non-smooth setting . . . . . . . . . . . . . . . . . . . . . . 59

Bibliography 61

Selbststandigkeitserklarung (Statement of Authorship) 65

Daten zum Autor (Information about the author) 67

x

1 Introduction

1.1 Minimal Surfaces

In this section we present basic definitions and some equations for minimal sur-

faces as well as a proof of the classical Bernstein theorem for surfaces in R3. We

collect some material from presentations of Xin [Xin03], Osserman [Oss02] and

Dierkes-Hildebrand-Sauvigny [DHS10]. The topic of minimal surfaces has kept

the attention of mathematicians for many years, therefore, apart from the above

mentioned references, there is a vast amount of literature which presents the

concepts of this section in greater detail.

Define a regular surfaceS to be a map χ(u) ∈ Cr(D,Rn), r ≥ 2, on a domain

D ⊂ R2. Let Γ := ∂Ω be a closed curve in D that bounds a subdomain Ω and let

Σ be the surface χ(u)|Ω. We suppose that the area of Σ is less than or equal to

the area of any other surface Σ defined parametrically by another map χ(u) in Ω

with the same boundary condition, that is, for each u ∈ Γ, we have χ(u) = χ(u).

What can we say about the surface Σ?

We can write χ(u) = (x1(u1, u2), ..., xn(u1, u2)) and by ∂χ/∂uj will denote the

vector∂χ

∂uj=

(∂x1

∂uj, ...,

∂xn∂uj

), j ∈ 1, 2.

We define the matrix G = (gij) where

gij =n∑k=1

∂xk∂ui

∂xk∂uj

=∂x

∂ui· ∂x∂uj

and assume that detG > 0 because we will only consider regular surfaces.

Since Ω is a subdomain of D with Ω ⊂ D and Σ := χ(u)|Ω, we define the area

of Σ as

A(Σ) =x

Ω

√detGdu1du2. (1.1.1)

1

1 Introduction

Another very important notion when studying submanifolds of Euclidean spaces

is the one of the normal bundle, the bundle that at each point is orthogonal to

the respective tangent plane. In other words, ∂χ∂u1, ∂χ∂u2

is a basis of the tangent

space of S. Let (TxS)⊥ be defined as the (n− 2)-dimensional vector space such

that

Rn = (TxS)⊥ ⊕ (TxS),

and

NS := (TS)⊥ =∐x∈S

(TxS)⊥.

We call a vector N ∈ NS = (TS)⊥ a normal vector to S, and since N is, by

definition, orthogonal to the basis of the tangent space, it can be computed as

follows.

ds=

∑i

duids

∂χ

∂ui,

d2χ

ds2=

∑i

d2uids2

∂χ

∂ui+∑i,j

duids

dujds

∂2χ

∂ui∂uj,

d2χ

ds·N =

∑bij(N)

duids

dujds

where we have defined

bij(Nu) =∂2χ

∂ui∂uj·Nu, (1.1.2)

where u ∈ D fulfills χ(u) = x.

One can introduce, for the general case of a surface S in Rn, the notion of normal

curvature at a point x ∈ S in the direction of a given unit tangent vector T ∈ TxSand a normal vector N ∈ NxS as (it is a well defined function with respect to T

and N)d2χ

ds·N = κ(N, T ), (1.1.3)

whereas, keeping N fixed but varying the unit vector T , since the set of unit

tangent vectors T 1xS is diffeomorphic to S1 → TxS, we may define the quantities

κ1(N) = maxT

κ(N, T ), κ2(N) = minTκ(N, T ), (1.1.4)

and those are called principal curvatures of S at a point with respect to N .

2

1.1 Minimal Surfaces

Definition 1.1.1. The mean curvature of a regular surface S with respect to a

normal vector N is defined as the average between the principal curvatures

H(N) =κ1(N) + κ2(N)

2. (1.1.5)

There are several explicit expressions for H(N) which we are going to use without

further explanations, but we remark that κ1(N) and κ2(N) are the roots of the

equation

det(bij(N)− λgij) = 0.

Therefore

H(N) =g22b11(N) + g11b22(N)− 2g12b12(N)

2 det(gij)(1.1.6)

Since H(N) is linear in N ∈ NS, there exists a unique vector H ∈ NxS such that

H(N) = H ·N for all N ∈ NxS.

Now, to obtain equations that describe in greater detail how surfaces that minimize

the area look like, we must consider normal variations of the C2-surface Σ, namely

Σλ = χ(u) = χ(u) + λh(u)N(u), u ∈ D, where h(u) is a C2 function and

N(u) ∈ C1 is normal to Σ at χ(u).

Definition 1.1.2 (Euler-Lagrange equations for the area functional). Let Σ =

χ(u) be a C2-surface and take a variation Σλ = χ(u) + λN(u) of Σ. Then the

Euler-Lagrange equation for the area functional is

A′(0) = −2x

Σ

H(N)dA = 0 (1.1.7)

This motivates the following definition.

Definition 1.1.3. A C2-surface Σ is said to be a minimal surface if and only if

H ≡ 0.

For a C2-surface z(x1, x2) = (z3(x1, x2), ..., zn(x1, x2)), where each of the n − 2

functions zk are C2, we have the minimal surface equation for non-parametric

surfaces:

(1 +

∣∣∣∣ ∂z∂x2

∣∣∣∣2)∂2z

∂x21

− 2

(∂z

∂x1

· ∂z∂x2

)∂2z

∂x1∂x2

+

(1 +

∣∣∣∣ ∂z∂x1

∣∣∣∣2)∂2z

∂x22

= 0. (1.1.8)

3

1 Introduction

1.1.1 Bernstein’s Theorem

In this section we present a short proof of the classical theorem of Bernstein

[Ber15]. It is important to remark that the proof given by Bernstein in his original

work is considered to be wrong. It is unclear who was the first person to give

the first correct proof of the statement, but we present here a proof given by

Osserman in [Oss02].

Theorem 1.1.4 (Bernstein’s theorem). Let Σ be a smooth minimal surface in

R3 given by the graph of a smooth function f : R2 −→ R defined on the whole of

R2. Then Σ is a plane.

To prove this theorem we introduce a couple of definitions and use a result by

Osserman that yields Bernstein’s theorem as a corollary.

Definition 1.1.5. Let u1, u2 be parameters of a surface Σ = χ(u1, u2) such that

the respective G matrix satisfies the following conditions when defining their

related G matrix,

g11 = g22, g12 = 0.

Or equivalently, there exists as function λ : S → R with λ(u) > 0 such that

gij = λ2δij

Then the surface χ(u1, u2) is said to be given in isothermal parameters.

The existence of (local) isothermal parameters around each point of a surface is

known since Gauß. We use also the following notation(definition):

φk(ζ) =∂xk∂u1

− i∂xk∂u2

; ζ = u1 + iu2 (1.1.9)

One can prove that φk is analytic in ξ if and only if xk is harmonic in u1, u2.

Moreover, one has

u1, u2 are isotermal parameters ⇐⇒n∑k=1

φ2k(ζ) = 0. (1.1.10)

With these notions we can prove the following theorem by Osserman.

Theorem 1.1.6 (Osserman [Oss02]). Let z(x1, x2) be a solution of the minimal

surface equation (1.1.8) in the whole plane. Then there exists a nonsingular linear

transformation

x1 = u1 (1.1.11)

4

1.1 Minimal Surfaces

x2 = au1 + bu2 with b > 0 (1.1.12)

such that (u1, u2) are global isothermal parameters for the surface defined by

xk = zk(x1, x2), k = 3, ..., n. (1.1.13)

First we prove Bernstein’s theorem using the above result.

Proof of Bernstein’s theorem: By Osserman’s theorem 1.1.6, we can pick the

linear transformations x1 = u1 and x2 = au1 + bu2 with isothermal parameters

u1, u2. Using the functions φk(ζ) defined in Equation (1.1.9) we see that φ1 and

φ2 are constant. But since Equation (1.1.1) holds, φ3 must also be constant and

the theorem follows.

Proof of Osserman’s theorem: By a theorem of Gauß, we can pick isothermal

parameters (ζ1, ζ2) for xk = zk(x1, x2), k = 3, ..., n. Since (ζ1, ζ2) are isothermal,

the functions

φk(ζ) =∂xk∂ζ1

− i∂xk∂ζ2

k = 1, ..., n

are analytic. As the surface xk = zk(x1, x1) is regular we also have∣∣∣∣∂(x1, x2)

∂(ζ1, ζ2)

∣∣∣∣ 6= 0,

and therefore

Imφ1φ2 = −∂(x1, x2)

∂(ζ1, ζ2)6= 0

implying that φ1 6= 0, φ2 6= 0 everywhere and

Im

φ2

φ1

=

1

|φ1|2Imφ1φ2 < 0.

This implies that the function φ2/φ1 is analytic in the whole ζ-plane. It has

negative imaginary part and must therefore be constant by Liouville’s theorem,

thus

φ2 = (a− ib)φ1, b > 0.

Taking the imaginary part of this equation we get

∂x2

∂ζ1

= a∂x1

∂ζ1

− b∂x1

∂ζ2

,

∂x2

∂ζ2

= b∂x1

∂ζ1

+ a∂x1

∂ζ2

.

(1.1.14)

5

1 Introduction

Now we pick x1 = u1, x2 = au1 + bu2, with b > 0 and get, by Equation (1.1.14),

∂u1

∂ζ1

=∂u2

∂ζ2

,∂u2

∂ζ1

= −∂u1

∂ζ2

which are the Cauchy-Riemann equations for u1 and u2, which imply that u1 and

u2 are both harmonic with respect to the parameters (ζ1, ζ2). Hence u1, u2 are

isothermal parameters.

1.2 Harmonic maps

In this section we introduce basic notions and properties of harmonic maps to

establish the necessary definitions, results and notations for the next chapters.

We are closely following the books by Xin [Xin12] and Lin and Wang [LW08].

Let (M, g) and (N, h) be Riemannian manifolds without boundary, of dimension

m and n, respectively. By Nash’s Theorem there exists an isometric embedding

N → RL.

Definition 1.2.1. A map u ∈ W 1,2(M,N) is called harmonic if and only if it is

a critical point of the energy functional

E(u) :=1

2

∫M

‖du‖2dvolg (1.2.1)

where ‖.‖2 = 〈., .〉 is the metric over the bundle T ∗M ⊗ u−1TN induced by g and

h.

Recall that the Sobolev space W 1,2(M,N) is defined as

W 1,2(M,N) =

v : M −→ RL; ||v||2W 1,2(M) =

∫M

(|v|2 + ‖dv‖2) dvolg < +∞ and

v(x) ∈ N for almost every x ∈M

Remark 1.2.2. If u ∈ W 1,2(M,N) is harmonic and φ : M −→M is a conformal

diffeomorphism, i.e. φ∗g = λ2g where λ is a smooth function, we have, by

Eq (1.2.1), that

E(u φ) =

∫M

λ2−m‖du‖2dvolg, (1.2.2)

6

1.2 Harmonic maps

which means that the energy is a conformal invariant if m = 2. This means that

harmonic maps defined on surfaces are invariant under conformal diffeomorphisms

of the domain. This remark motivates the study of stationary harmonic maps.

We do not consider those in the present work. See Helein [Hel02] and Lin and

Wang [LW08].

Let X, Y ∈ Γ(TM) and define for, u ∈ W 1,2(M,N), the map

du : Γ(TM) −→ u−1(TN)

given by

du(X) := u∗X

where du is called the differential 1-form of u. We denote by ∇du the gradient

of du over the induced bundle T ∗M ⊗ u−1(TN), that is ∇du satisfies ∇Y du ∈Γ(T ∗M ⊗ u−1(TN)), for each Y ∈ Γ(TM).

Definition 1.2.3 (Second fundamental form). The second fundamental form of

the map u : M −→ N is the map defined by

BXY (u) = (∇Xdu)(Y ) ∈ Γ(u1TN). (1.2.3)

Note that B is bilinear and symmetric in X, Y ∈ Γ(TM). Indeed, we can see this

map as

B(u) ∈ Γ(Hom(TM TM, u−1TN)).

The second fundamental form is often denoted by A(u) or ∇du. Curiously, B is

mostly used among geometers, while A and ∇d are used among analysts.

Definition 1.2.4. The tension field of a map u : M −→ N is the trace of second

fundamental form

τ(u) = Beiei(u) = (∇eidu)(ei) (1.2.4)

seen as a cross section of the bundle u−1TN .

Taking a C1 variation(u(., t)

)‖t‖<ε, we have

d

dtE(u(., t)) =

d

dt

∫M

e(u)dvolg

=

∫M

∂te(u(x, t))dvolg

7

1 Introduction

=

∫M

∂t

(1

2hij(u(x))

∂ui

∂xα∂uj

∂xβ

)gαβ(x)

√det g dx

= −∫M

hij(u(x, t)

)τ(u(x, t)

)i∂uj∂t

dvolg

= −⟨τ(u(x, t)),

∂u

∂t

⟩L2

Therefore the first variation of the energy functional is:

d

dtE(u(., t)) = −

∫M

⟨∂ut∂t

, τ(ut)

⟩dvolg. (1.2.5)

In local coordinates

e(u) = ‖du‖2 = gij∂uβ

∂xi∂uγ

∂xjhβγ

τ(u) =

(∆gu

α + gijΓαβγ∂uβ

∂xi∂uγ

∂xjhβγ

)∂

∂uα,

where the Γαβγ denote the Christoffel symbols of N .

By Equation (1.2.5) the Euler-Lagrange equations for the energy functional are

τ(u) =

(∆gu

α + gijΓαβγ∂uβ

∂xi∂uγ

∂xjhβγ

)∂

∂uα= 0. (1.2.6)

Hence we can define (weakly) harmonic maps equivalently as maps that (weakly)

satisfy Equation(1.2.6). Equation (1.2.6) is called the harmonic map equation.

Definition 1.2.5 (Totally geodesic maps). A map u ∈ W 1,2(M,N) is called

totally geodesic if and only if its second fundamental form identically vanishes,

i.e. B(u) ≡ 0.

Harmonic maps are interesting mathematical objects. They naturally appear

in the study of minimal immersions, closed geodesics, and other central problems.

Understanding them is crucial for classical and modern Riemannian Geometry.

Lemma 1.2.6 (Composition formulas). Let u : M −→ N and f : N −→ P ,

where (P, i) is another Riemannian manifold. Then

∇d(f u) = df ∇du+∇df(du, du), (1.2.7)

τ(f u) = df τ(u) + tr∇df(du, du). (1.2.8)

8

1.2 Harmonic maps

Remark 1.2.7. When u is harmonic, i.e. τ(u) = 0, the above formula is

particularly useful. In particular, if P = R and f is a (strictly) convex function,

then τ(f u) ≥ 0 (> 0). That is, f u : M −→ R is a (strictly) subharmonic

function on M.

Given two Riemannian manifolds (M, g) and (N, h) a natural question is: Is there

a harmonic map u : (M, g) → (N, h)? If M = S1, this question is about the

existence of closed geodesics. Moreover, if u is a Riemannian immersion, we ask

whether M is a minimal submanifold of N .

There are many results on the existence or non-existence of such maps.

Remark 1.2.8. Assume that M and N are both closed.

• If dim(M) = 1, are there harmonic u : M → N? Yes. Every homotopy

class of maps of a circle into N contains a closed geodesic.

• If dim(N) = 1, are there harmonic u : M → N? Yes.

• If dim(M) = 2 and π2(N) = 0, are there u : M → N harmonic? Yes. This

is a famous result of Sacks-Uhlenbeck [SU81].

• If RiemN ≤ 0, are there u : M → N harmonic? Yes. If KM ≥ 0, the map

is actually totally geodesic.

• Are there u : T→ S2 harmonic with deg(u) = ±1? No.

• Are there u : Sm → Sn harmonic? Yes if n ≤ 1. This is an open problem

in case n is large.

Example 1.2.9 (Lemaire [Lem78]). Given an integer p > 1 and a surface Σp of

genus p there exists a harmonic map φ between Σp and the sphere S2. His proof

uses the following interesting argument: He constructs a map φ : Σg → S1 and a

totally geodesic embedding S1 → S2 such that the following diagram commutes.

Σg S2

S1

φ:= (tot.geod) φ

φ tot. geod

Clearly, the image φ(Σg) ⊂ S2 is contained in the image of a closed geodesic of

S2. In the next section we give an example of a harmonic map from a minimal

surface of even genus into S2 that does not contain closed geodesics in its image.

This will be important later on.

9

1 Introduction

1.2.1 The harmonic map flow

A classical method to find harmonic maps between two Riemannian manifolds is

to look at their evolution equation. More precisely, given a map u0 ∈ W 1,2(M,N),

Equation (1.2.5) above indicates that we can evolve the map u0 according to

∂tu(x, t) = τ

(u(x, t)

)(1.2.9)

with initial value u|t=0 = u0.

Supposing we get a C2 solution u(x, t) for∂tu(x, t) = τ(u(x, t)

)u( · , 0) = u0.

(1.2.10)

Then

d

dtE(u(x, t)) = −

∫M

⟨τ(u(x, t)

), ∂tu(x, t)

⟩u∗TN

dvolg

= −∫M

∣∣τ(u(x, t))∣∣2dvolg (1.2.11)

Since∣∣τ(u(x, t))

∣∣2 ≥ 0 the energy does not increase along the flow.

It was proven by Struwe [Str85], in the case where the dimension of M is 2 and N

is a compact manifold, that there exists a global weak solution of the heat flow of

harmonic maps. Such a solution is smooth with the possible exception of finitely

many singular points.

Theorem 1.2.10 (Struwe). For any initial value u0 ∈ W 1,2(M,N), there exists

a unique solution u of ∂tu(x, t) = τ(u(x, t)

)u( · , t) = u0

on M × [0,+∞) which is regular on M × (0,+∞) except for at most finitely many

points (xl, T l), 1 ≤ l ≤ L characterized by the condition that

lim supT→T l−

ER(u( · , T ), xl

)> ε1 for all R ∈ (0, R0].

Theorem 1.2.11 (Struwe). Let u be the solution of∂tu(x, t) = τ(u(x, t)

)u( · , t) = u0

10

1.2 Harmonic maps

Γ1

Γ2

Γ3

Figure 1.1: A region in S2 without closed geodesics.

constructed in the above theorem. Suppose (x0, T ), T ≤ ∞, is a point where

lim supt→T,t<T

ER(u(., t), x0) > ε1 for all R ∈ (0, R0]

Then there exist sequences xm → x0, tm → T (tm < T ) and Rm → 0 (Rm ∈(0, R0]) and a regular harmonic map u0 : R2 → N such that

uRm,(xm,tm)( · , 0) −→m→∞

u0 locally in H2,2(R2, N)

Furthermore u0 has finite energy and extends to a smooth harmonic map S2 → N .

We use the harmonic map flow and Struwe’s result to give an example of a

harmonic map between a surface of even genus and a sphere of dimension 2 that

contains no closed geodesics in its image.

We start by considering a region on S2 described by Figure 1.1.

Notation 1.2.12. By ‘Eq’ we denote the embedding S1 → S2 with (x1, x2, x3) ∈S2; x3 = 0.

11

1 Introduction

u0

smooth

T3T1 T2

S2

S2

N

S

C1C2

C3

p3

p1

p2

Figure 1.2: The Riemann surface Σ2 and the map u0

If p ∈ Eq\ (Γ1∪Γ2∪Γ3), then A(p) /∈ Eq\ (Γ1∪Γ2∪Γ3) where A is the antipodal

map. Therefore it is clear that S2 \ (Γ1∪Γ2∪Γ3) does not admit closed geodesics.

We define a compact Riemann surface, called Σ2, as follows: around each segment

Γi ⊂ S2 we put a circle Ci; Ci ∩ Cj = ∅, as in the picture

Consider the poles S = (0, 0,−1) and N = (0, 0, 1). Let S2 be another sphere

such that N , S belong to the axis x3 and S2 is a reflection of S2 centered at a

point of the axis and denote by d > 0 the distance between N and S. In S2,

consider also the segments Γi and the circles Ci as before.

We connect the two spheres by three tubes (cylinders), called T 1, T 2, T 3 such

that ∂T i = Ci ∪ Ci, i.e., they glue them exactly in the removed circles (see the

Σ2 in Figure 1.2).

It is important to note two things about this construction:

• A Z3 symmetry around the x3 axis is preserved, i.e. the tubes and the

circles are all equal.

• A Z2 symmetry around the midpoint N and S is preserved

12

1.2 Harmonic maps

These symmetries play an important role in the construction of the desired

harmonic map φ. Clearly, the procedure above generates a genus 2 compact

Riemann surface Σ2 and we consider the induced metric from R3 as its Riemannian

metric.

The next step is to define the smooth initial condition between Σ2 and S2 and

explore the properties of the harmonic map flow.

Define u0 ∈ C∞(Σ2, S

2)

as

u0 : Σ2 −→ S2 (1.2.12)

x 7→

N if x ∈ S2 \ (C1 ∪ C2 ∪ C3)

S if x ∈ S2 \ (C1 ∪ C2 ∪ C3)

(∗) if x ∈ (T 1 ∪ T 2 ∪ T 3)

(1.2.13)

(∗): Consider the following geodesic connecting the south and the north poles in

S2

γ : [−R,+R] −→ S2

t 7−→

(sin(π(t+R)

2R

), 0,− cos

(π(t+R)

2R

))where 2R > 0 is the height of the tubes T i in our Riemann surface Σ2. Now,

given θ ∈ [0, 2π)

Γθ : [−R,R] −→ T i

t 7−→ (r cos θ, r sin θ, t)

where r is the radius of T i.

Up to the diffeomorphism that makes the cylinder straight, define u0|T i as a map

in C∞(Σ2, S2) such that

u0 Γθ(t) = γ(t) for every θ ∈ [0, 2π). (1.2.14)

It is easy to verify that the image of Σ2 under u0 has the same Z2 and Z3

symmetries of Σ2. We point out that the harmonic map flow may preserve

symmetries of the initial condition.

Remark 1.2.13. (Symmetry property of the flow) Let ϕ be the Z3 action on Σ2

given by the rotation by 3π2

over the axis connecting the poles. Let ψ be the ‘same’

Z3 action over the sphere S2. Obviously u0 ϕ = ψ u0, hence we have∂t(ψ u

)= τ(ψ u

)ψ u( · , t) = u0 ϕ

(1.2.15)

13

1 Introduction

or equivalently ∂t(u ϕ

)= τ(u ϕ

)u ϕ( · , t) = ψ u0.

(1.2.16)

Moreover, if u is a solution for Eq (1.2.10) with initial condition (1.2.14), as ψ

is obviously totally geodesic, we have

∂t(ψ u)− τ(ψ u) = dψ(u( · , t))∂tu− dψ(u( · , t))τ(u)

= dψ(u( · , t))(∂tu− τ(u)

)= 0

therefore ψ u is a solution to the same problem.

The same argument applies to the Z2 action (a, b, c) 7→ (a, b,−c).

By the above remark, the harmonic map u∞ : Σ2 → S2 given by the solution of

Equation (1.2.10) with initial condition given by (1.2.14) is Z3 and Z2 equivariant

with respect to the actions defined above.

Another important remark about the initial condition u0 given by Equation (1.2.14)

is that we can control its energy by changing the length and radius of the tubes

T i.

Remark 1.2.14 (A control on the energy of u0). Since ‖γ(t)‖ = π2R

, we have

0 < ‖du0(p)‖ ≤ π

2R

and therefore

0 < E(u0) ≤ 1

2

∫T i

( π

2R

)2

dvolT i

=π2

8R2

∫T idvolT i

=π3r2

4R

By making the tubes connecting S2 and S2 in Σ2 thinner and longer, we can make

the energy of u0 arbitrarily small. More precisely, given any ε > 0, we can pick

r ∈ (0, 1) and R ∈ (1,+∞) such that E(u0) ≤ ε. By Equation (1.2.11), we know

that the energy decreases along the flow and therefore we have a control on the

energy of ut for every t ∈ R+.

14

1.2 Harmonic maps

Theorem 1.2.10 and Theorem 1.2.11 above roughly tell us that given any initial

condition u0 ∈ W 1,2(M,N), there exist a regular solution to the harmonic map

flow with the exception of some finite points on which the energy is controlled.

In other words, we have a harmonic map

u∞ : M\q1, ..., ql → N

and around the singularities this map can be extended to a harmonic sphere

h : S2 → N .

The interesting fact is that (when KN ≤ 1) each harmonic two-sphere which

bubbles out carries some minimum amount of energy. By this we mean that,

if K : S2 → R is the Gauss curvature of S2 and K0 the maximum value of all

sectional curvatures of N, the Gauss-Bonnet theorem implies that

4π =

∫S2

Kdσ ≤ K0

(Area of h(S2)

)≤ K0E(h)

E(h) ≥ 4π

K0

.

In our specific case, N = S2, so K0 = 1. Thus 4π is the minimum amount of

energy needed to form a bubble.

By the definition of the initial condition u0 and Remark 1.2.14, we can avoid

the formation of bubbles by taking Σ2 as the compact Riemann surface with

tubes of length R ∈ (1,+∞) and radius r ∈ (0, 1) such that for a given ε0 > 0,

we have E(u0) ≤ ε0 4π.

Now, given the non-existence of bubbles, we have the harmonic map

u∞ : Σ2 → S2.

It remains to be proven that u∞(Σ2) ⊂ S2 \ (Γ1 ∪ Γ2 ∪ Γ3).

To do so, we need the following lemma due to Courant.

Lemma 1.2.15 (Courant-Lebesgue). Let f ∈ W 1,2(D,Rd), E(f) ≤ C, δ < 1

and p ∈ D = (x, y) ∈ C; x2 + y2 = 1. Then there exists some r ∈ (δ,√δ) for

which f |∂B(p,r)∩D is absolutely continuous and

|f(x1)− f(x2)| ≤ (8πC)12

(log

1

δ

)− 12

(1.2.17)

for all x1, x2 ∈ ∂B(p, r) ∩D.

15

1 Introduction

Proof. We follow [Jos91]. The fact that f |∂B(p,r)∩D is absolutely continuous, at

least for almost every r, follows from the fact that f is in the Sobolev space

W 1,2(D,Rd). Taking polar coordinates (ρ, θ) around p we have, by the intermedi-

ate value theorem, that, for all x1, x2 ∈ ∂B(p, r) ∩D,

|f(x1)− f(x2)| ≤∫

(ρ,θ)∈∂B(p,r)∩D|fθ(x)|dθ

≤ (2π)12

(∫|fθ(x)|2dθ

) 12

But the energy of f on B(p, r) ∩D is

E(f ;B(p, r) ∩D) =1

2

∫B(p,r)∩D

(|fρ|2 +

1

ρ2|fθ|2

)ρ dρ dθ

Hence, there exists a number r ∈ (δ,√δ) such that

1

2

∫(ρ,θ)∈∂B(p,r)∩D

|fθ|2dθ ≤C∫ √δ

δρ−1dρ

=2C

log(δ−1).

The lemma follows from these two equations.

We point out that the Z2 symmetry of u∞(Σ2) implies that there are two

antipodal points, called N and S, on the image of u∞(Σ2). On the other hand,

the Z3 symmetry and the fact that the image is connected, there exist three points

pi ∈ Ti, i = 1, 2, 3, each of them Z2-invariant, such that u∞(pi) ∈ Eq. Moreover,

we have that ϕ(u∞(p1)) = u∞(p2), ϕ(u∞(p2)) = u∞(p3) and ϕ(u∞(p3)) = u∞(p1).

Now we use the Courant-Lebesgue 1.2.15 to show that a small neighborhood of

the tube around pi is mapped into a small neighborhood of u∞(pi) of controlled

size. Namely, since E(u∞) ≤ E(u0) < r2

R, where r is the radius of the tubes and

R their heights. Suppose we take r ∈ (0, 1) and R ∈ (1,+∞) such that u∞ has

no bubbles and r.π << π/6.

Take δ < r2, by Courant-Lebesgue there exists some s ∈ (r2, r) such that

|u∞(x1)− u∞(x2)| ≤

(8π.r2

R

) 12(

log(1

δ

))− 12

(1.2.18)

for every x1, x2 ∈ ∂Bg(pi, s) ∩Bg(pi, 1). But since δ < r2, we get

1δ> 1

s⇒ log

(1δ

)> log

(1s

)⇒

(log(

1s

))− 12 >

(log(

))− 12 .

16

1.2 Harmonic maps

Therefore by Equation (1.2.18),

|u∞(x1)− u∞(x2)| ≤(

8π.sR

) 12(log(

1s

))− 12

=√s

(log( 1s))

12.√

8πR<<< π

6.

In particular, if we call S1i the circle given by the intersection of the x, y-plane in

R3 with Ti the above argument shows that the image of a δ-neighborhhod of S1i

under u∞ is contained in a neighborhood of u∞(pi) that does not intersect any of

the removed curves Γi, because since pi is invariant under the Z2 action, so is S1i .

Now we have to prove that for any other point a ∈ Ti such a /∈ B(S1i , δ) := x ∈

Ti ; d(x, S1i ) ≤ δ, we have that u∞(a) /∈ Eq.

We argue by contracdiction. Obviously, u∞(Σ2) ( S2, because Area(u∞) ≤E(u∞) < r2

R<< 1. As Σ2 is compact, then u∞(Σ2) is compact. Now, if

∂u∞(Σ2) =, then u∞(Ti) is totally geodesic and we have that u∞(a) /∈ Eq.

Suppose ∂u∞(Σ2) 6=. Therefore there exists another point a ∈ Ti\B(S1i , δ) such

that u∞(a) ∈ ∂u∞(Σ2) and u∞(a) ∈ Eq. But now consider a C1-variation, for

|t| < ε, (u∞

)t(x) = u∞(x) + tη(x),

where η : Σ2 −→ S2 is smooth, η ≡ 0 outside a neighborhood of a and(u∞)t(a) ∈

HemN (where we are assuming without loss of generality that d(a, S2) < d(a, S2)).

This map(u∞)t(x) clearly satisfies E((u∞)t) < E(u∞), but since u∞ is a harmonic

smooth map, we have that such a point a does not exist. This concludes the

proof that u∞(Σ2) ⊂ S2\(Γ1 ∪ Γ2 ∪ Γ3).

To summarize, we have shown that the solution u∞ of the harmonic map flow∂tu(x, t) = τ(u(x, t)

)u( . , t = 0) = u0

where u0 : Σ2 → S2 is given by Definition 1.2.14 is a harmonic map from the

compact genus two Riemann surface Σ2 into the sphere S2 and the image u∞(Σ2)

does not contain closed geodesics.

Remark 1.2.16. With the same argument of this example, we could construct

compact Riemann surfaces of genus 2p for any p > 1. The only change in the

proof would be that the Z3 symmetry would be replaced by a Z2p+1 symmetry.

Moreover,if we replace the target by an other 2-dimensional surface with the

17

1 Introduction

appropriate symmetries, we could try to construct a u0-type initial condition

respecting the necessary symmetries.

1.3 Minimal Submanifolds

In this section we introduce the other main object of interest of this work, namely

Riemannian submanifolds of Euclidean space. We introduce concepts like the

second fundamental form from a more geometrical viewpoint and show under

which circumstances these concepts are equivalent to those introduced either for

harmonic maps or for the classical case of surfaces.

Let (Mn, g) be a Riemannian manifold. Suppose there exist another Riemannian

manifold, that we denote by (Mn, g,∇), where n > n and ∇ is the Levi-Civita

connection, and an immersion F : (M, g,∇) −→ (M, g,∇) with the property

g = F ∗g. We denote by k the codimension of the immersion, k := n− n > 0.

Definition 1.3.1 (Normal bundle). For every point p ∈ M ⊂ M we can split

the tangent space of M as

TpM = TpM ⊕NpM

where the latter is called the normal space to M in M at the point p ∈ M .

This shows that we can define the normal bundle with respect to the immersion

M →M as

TM |M =∐p∈M

TM ⊕NM.

Notation 1.3.2. Throughout the text we use the following notation

( . )> : orthogonal projection on TM,

( . )N = ( . )⊥ : orthogonal projection on NM.

Obviously, as the metric in M is the one induced by the metric in M , we have

that for every V,W ∈ Γ(TM), and ν ∈ Γ(NM), the induced connections on TM

and NM are

∇VW := (∇VW )>, and (∇V ν) := (∇ν)⊥,

where on the right-hand side of the equations we are considering extensions of

the vectors V,W ∈ Γ(TM), and ν ∈ Γ(NM) to M . One can easily prove that ∇is the Levi-Civita connection for the Riemannian manifold (M, g).

18

1.3 Minimal Submanifolds

Definition 1.3.3 (Second fundamental form). The second fundamental form

of the manifold M with respect to the embedding F : (M, g) → (M, g) is the

symmetric, bilinear form

B : TM ⊗ TM −→ NM

(X, Y ) 7−→ BXY

(1.3.1)

defined by BXY := ∇XY −∇XY =(∇XY

)N.

Definition 1.3.4 (Shape operator). For each ν ∈ Γ(NM) we define a self-adjoint

operator associated with the second fundamental form, called the shape operator

Aν : TM −→ TM, X 7−→ Aν(X) (1.3.2)

where Aν(X) = −(∇Xν

)>.

Definition 1.3.5. If the second fundamental form of the immersion B is identi-

cally zero at every point of M , we say that M →M is a totally geodesic immersion

and M is a totally geodesic submanifold of M .

This definition is equivalent to the following geometric characterization: M is a

totally geodesic submanifold of M if and only if every geodesic in M is also a

geodesic in M .

Based on the above characterization of totally geodesic submanifolds, one can

easily see that the only totally geodesic submanifolds of Euclidean spaces are their

affine subspaces. In this case, one can understand Bernstein’s problem as the

following question: Are globally non-parametric minimal surfaces of Euclidean

3-space always totally geodesic immersions?

Changing the words ‘surfaces’ to ‘submanifolds’ and ‘Euclidean’ to ‘M ’, we

may try to answer the generalized Bernstein problem.

This question does not always have a positive answer. We will address more

suitable Bernstein-type problems that have appeared in the literature for the last

hundred years.

Definition 1.3.6. The mean curvature vector of M in M is defined as the trace

of the operator B. That is

~H :=1

ntr(B) =

1

n

n∑i=1

Beiei (1.3.3)

where e1, ..., en is a local orthonormal frame for M .

19

1 Introduction

Definition 1.3.7. A submanifold (M, g) in (M, g) is called minimal if and only

if ~H ≡ 0 and is called of parallel mean curvature if ∇ ~H ≡ 0, that is, if ~H is a

parallel cross-section of the normal bundle.

It is straightforward to see that

M totally geodesic =⇒ M minimal =⇒ M parallel m.c.

(B ≡ 0) (H ≡ 0) (∇H ≡ 0)(1.3.4)

Remark 1.3.8. Suppose k = n− n = 1, then Mn →Mn+1

is an embedding of a

hypersurface and NM = spanν. We define the second fundamental form as

A := Aν : TM −→ TM (the shape operator), which is a self-adjoint operator on

each tangent space TpM . This implies that there are real eigenvalues κ1, ..., κn,

called the principal curvatures.

We can define the Gauss and mean curvatures in the same way as in the

classical case of surfaces in R3,

K(p) = κ1(p)...κn(p) is the Gauss curvature (1.3.5)

H(p) =1

n(κ1(p) + ...+ κn(p)) is the mean curvature (1.3.6)

We denote by I(M,M) the set of all immersions F : M −→M . Then the volume

Vol(F (M)) =

∫M

dvolg

is a functional on this space. When considering the immersion with a compact

manifold as domain, we have the following theorem.

Theorem 1.3.9 (First variation of the area functional). Let M be a compact

Riemannian manifold, and F : M −→ M an isometric immersion with mean

curvature vector H. Take a smooth variation Ft of F , for |t| < ε, satisfying

F0 = F and Fy|∂M = F∂M . Then

d

dtVol(FtM)

∣∣∣t=0

= −∫M

⟨nH,

∂Ft∂dt

⟩dvolg (1.3.7)

The proof of this theorem is very similar to the first variation for the energy

functional in the previous section. The computation is a little more involved since

we have to differentiate the square root of a determinant. A careful proof of this

theorem can be found in Xin [Xin03] or any other book on Riemannian geometry.

20

1.3 Minimal Submanifolds

Remark 1.3.10. By Equation (1.3.7), we have that H = 0 is the Euler-Lagrange

equation for the area functional and, as in the case of surfaces, this justifies

the defintion of minimal submanifolds as critical points of this functional or as

submanifolds with vanishing mean curvature vector.

Remark 1.3.11. If the target manifold of the immersion is the Euclidean space

Rn and i : M −→ Rn is an isometric immersion, the mean curvature vector can

be expressed by the equation

∆i = mH, where m = dimM.

It is immediate that M is a minimal submanifold of Rn if and only if each

component of the immersion i is a harmonic function on M .

Example 1.3.12 (Minimal graph in Rn+1). Let M = graph(f) be a submanifold

of Rn+1 with the metric induced from the Euclidean one, where f : M −→ R is a

smooth function. We write

ds2 = gijdxidxj,

where

gij = δij + fifj.

Let w =√

1 +∑

i f2i be the volume element of M . We have gij = δij − 1

w2fifj

and ν ∈ N1M is given by

ν =1

w(f1, ..., fn,−1).

Since

∇ ∂

∂xi

∂xj= (0, ..., 0, fij)

and ⟨B ∂

∂xi∂

∂xj, ν⟩

= − 1

wfij

we obtain gij = fij if M is a minimal hypersurface. Using these equations to

compute the Laplacian of the immersion we get(1 +

n∑i=1

f 2i

)fjj − fifjfij = 0. (1.3.8)

If n = 2, the above becomes the classical equation for minimal surfaces in

Euclidean 3-space

(1 + f 2y )fxx − 2fxfyfxy + (1 + f 2

x)fyy = 0. (1.3.9)

21

1 Introduction

One of the most important tools to study the geometry of submanifolds is the

Gauss map.

Definition 1.3.13 (Gauss Map). Let Mn → Rn be an embedded n-dimensional

oriented submanifold of Euclidean space. For any x ∈M , the tangent space TxM

can be moved to the origin by parallel translation in Rn, to obtain an n-subspace

of Rn, that is, a point in the oriented Grassmannian manifold G+n,n. This defines

a map γ : M −→ G+n,n called the Gauss map of the embedding M → Rn.

The Gauss amp gives information about the geometry of the embedding and is

particularly interesting when we are dealing with submanifolds of parallel mean

curvature. This is explained in the next theorem.

Theorem 1.3.14 (Ruh-Vilms [RV70]). Let Mn be a submanifold in Rn and let

γ : M −→ G+n,n be its Gauss map. Then γ is harmonic if and only if M has

parallel mean curvature.

In the original paper [RV70] the tension field of the Gauss map is explicitly

computed.

1.4 Main Results

1.4.1 Domains that do not admit images of non-constant

harmonic maps

In the next chapter we revise maximum principles for harmonic maps. Classically,

such results provide us with a technique to prove that in a given region V ⊂ N

there are no images of non-constant harmonic maps, where (N, h) is a complete

Riemannian manifold. More precisely, there are no non-constant harmonic maps

u : M −→ N , where M is a compact manifold, such that u(M) ⊂ V . The

technique consists in finding on V a strictly convex function f : V −→ R, then

use composition formulas for harmonic maps and maximum principles to prove

that u must be constant. This tool has been used for many years to prove

non-existence of harmonic maps.

In this work we modify this approach. Instead of using analytical arguments to

obtain a subset that admits a strictly convex function, we explore the geometry

of regions that can contain the image of a non-constant harmonic map.

To economize notation we introduce the following definition to be used through-

out the text.

22

1.4 Main Results

Definition 1.4.1 (Property (?)). We say that an open connected subset R ⊂(N, h), where (N, h) is a complete Riemannian manifold, has property (?) if there

is no pair (M,φ), where (M, g) is a compact manifold and φ : M −→ N is a

non-constant harmonic map with φ(M) ⊂ R.

We use Sampson’s maximum principle to define regions in arbitrary complete

manifolds that have property (?). Then we use geometric intuition and Sampson’s

maximum principle to develop an intrinsic understanding for subsets of a manifold

which cannot contain harmonic maps defined on compact manifolds. More

precisely, we will prove the following theorem.

Theorem. Let (N, h) be a complete Riemannian manifold and Γ : [a, b] −→ N

a smooth embedded curve. Consider a smooth function r : [a, b] −→ R+ and the

region

R =⋃t∈[a,b]

B(Γ(t), r(t)), (1.4.1)

where B(., .) is the geodesic ball and, for every t, r(t) is smaller than the convexity

radius of N. If, for each t0 ∈ (a, b), R\B(Γ(t0), r(t0)) is a disconnected set,

then there exists no compact manifold (M, g) and non-constant harmonic map

φ : M −→ N such that φ(M) ⊂ R. In other words, R has the property (?) of

Definition 1.4.1.

In addition, we prove another theorem on a more abstract notion of what it

means to be a barrier to the existence of harmonic maps. To do so, we need the

following

Condition 1. For every p ∈ R and every v ∈ T 1pN , the geodesic Γ such that

Γ(0) = p and Γ′(0) = v satisfies, for some tv ∈ [0, K], Γ(tv) ∈ ∂R, where K is

independent of p and v.

Notation 1.4.2. For p ∈ N and ν ∈ ∂Bg(p, r) with r 1, let tmax(p, ν) ∈ R+

be the first time at which Γp,ν(t) ∈ ∂R.

Condition 2. For every p ∈ R and every ν ∈ T 1pN , if t < tmax(p, ν), then the

set of directions η ∈ T 1Γp,ν(t)N such that d(p, γRΓp,ν(t),η(ε)) < d(p,Γp,ν(t)) is properly

contained in T 1Γp,ν(t) and connected. Here γR denotes a geodesic in R with ε ∈ R+

sufficiently small (that is, a curve that minimizes the length in R).

The meaning of these conditions will be explained in the respective section.

Under these conditions we obtain the following result.

23

1 Introduction

Theorem. Let (N,h) be a complete Riemannian manifold and suppose that for a

region R ⊂ N , R \ ∂R is open and connected. In addition, suppose R satisfies the

Conditions 1 and 2 above. Then R has property (?) of Definition 1.4.1.

We show that if N is a symmetric space, the conditions are particularly

interesting and give some examples. In particular, we compute the largest subset

of (Sn, g) that contains no image of non-constant harmonic maps defined on

compact manifolds, namely Sn\[Sn−1/ ∼], where ∼ is the equivalence relation

given by the antipodal map and Sn−1 is embedded as an equator in Sn. This was

proven in [Jos12] by defining a strictly convex function in each compact properly

contained subset of Sn\[Sn−1/ ∼], but our proof is simpler and, as we observe

later, a bit more general, since we get some flexibility on the set [Sn−1/ ∼] that

we remove.

We use the results of Allard [All72] and Fisher-Colbrie [FC80] to study the

case where the harmonic map, instead of being defined on a compact manifold, is

defined over the graph of a smooth map. This will be of extreme importance to

study Bernstein-type theorems in higher codimensions.

1.4.2 Bernstein-type theorems in higher codimensions

A central result in the theory of minimal surfaces is Bernstein’s theorem. Profound

methods in analysis and geometric measure theory were developed to generalize

Bernstein’s theorem to higher dimensions, culminating in the theorem of J. Simons

[Sim68] stating that an entire minimal graph has to be planar for dimension

d ≤ 7. This dimension constraint is optimal: Bombieri, de Giorgi, and Giusti

[BDGG69] constructed a counter-example to such an assertion in dimension 8

and higher. This reveals the subtlety and difficulty of the problem. Under the

additional assumption that the slope of the graph is uniformly bounded, Moser

[Mos61] proved a Bernstein-type result in arbitrary dimension.

All the preceding results consider minimal hypersurfaces, that is, minimal

graphs in Euclidean space of codimension one. For higher codimension, the

situation is more complicated. On one hand, Lawson-Osserman [LO77] have

given explicit counterexamples to Bernstein-type results in higher codimension.

Namely, the cone over a Hopf map is an entire Lipschitz solution to the minimal

surface system. Since the slope of the graph of such a cone is bounded, even a

Moser-type result for codimension higher than one cannot hold. On the other

hand, there are also some positive results in higher codimension, although, in view

of the Lawson-Osserman examples, they necessarily require additional constraints.

24

1.4 Main Results

We can describe the main development as a sequence of steps. Those results all

depend on the fact that, by the theorem of Ruh-Vilms in [RV70], the Gauss map of

a minimal submanifold in Euclidean space is harmonic. This Gauss map takes its

values in a Grassmann manifold G+p,n (which is a sphere in the case of codimension

n − p = 1). Therefore, the geometry of Grassmann manifolds is the key to

understanding the scope of Bernstein theorems in higher codimension. Since the

composition of a harmonic map (such as the Gauss map) with a convex function

is a subharmonic function, if such a convex function is found the maximum

principle can be applied to show that, when the domain of the harmonic map is

compact, the resulting subharmonic function is constant. If the convex function

is nontrivial, for instance strictly convex, then the harmonic map is constant, and

the minimal graph is therefore linear. A key technical problem emerges since in

our application the domain is Rp, which is not compact, so that the maximum

principle cannot be applied directly. We postpone the discussion of this issue and

return to the geometric steps.

1. Hildebrandt-Jost-Widman [HJW80] identified the largest ball in the Grass-

mannian on which the squared distance function from the center is convex.

Thus, if the Gauss image is contained in such a ball, that is, when the slopes

of the tangent planes of our minimal submanifold do not deviate too much

from a given direction, the minimal graph is linear and a Bernstein result

holds.

2. Why consider only distance balls? Jost-Xin [JX99] constructed regions in

G+p,n which are larger than convex metric balls but on which the squared

distance function is still convex. After all, even though G+p,n is a symmetric

space, the distance function behaves differently in different directions. Thus

the Gauss region implying the Bernstein results is larger.

3. Why consider only the squared distance function? Jost-Xin-Yang [JXY12],[JXY16]

went further by constructing larger regions in G+p,n that support strictly

convex functions. Thus the Bernstein theorem was further extended. In

particular in codimension 1, the classical case, even minimal hypersurfaces

that might be more general than graphs are shown to be linear. Still, it is

not clear whether the Lawson-Osserman example is sharp or whether there

exist other examples that come closer to the bounds obtained by [JXY16]

in higher codimension.

25

1 Introduction

4. The level sets of convex functions are convex. The starting idea of the

present thesis is that this is the key property: to have a family of convex

hypersurfaces. But why do we need convex functions? We show that a

foliation by convex hypersurfaces suffices, which does not need to come

from a convex function. This clarifies the geometric nature of the maximum

principle that was the cornerstone of the reasoning just described. As far

as we can tell, this seems to be the ultimate conceptual step in this line of

research. Theorem 1.4.3, to be stated shortly, seems to be the optimal result

in codimension 2. It remains to explore our scheme in higher codimension.

In fact, all those results apply more generally to graphs of parallel mean curvature,

since the Gauss map remains harmonic in such cases by the Ruh-Vilms theorem.

It was proven by Chern [Che65] that a hypersurface in Euclidean space that is

an entire graph of constant mean curvature is a minimal hypersurface. Thus, by

Simons’ result, it is a hyperplane for dimension d ≤ 7. See also Chen-Xin [CX92]

for a generalization of Chern’s result.

A maximum principle, however, only applies if the domain is compact, but

the domain of an entire minimal graph is Rp. Therefore, one needs to turn

the qualitative maximum principle into quantitative Harnack-type estimates, a

technique pioneered by Moser [Mos61]. In the proof of Hildebrandt-Jost-Widman

[HJW80], the analytical properties of such convex functions were used to derive

Holder estimates for harmonic maps with values in Riemannian manifolds with

an upper bound for the sectional curvature. By a scaling argument, they could

conclude a Liouville-type theorem for harmonic maps under assumptions including

the harmonic Gauss map. In the same setting, Jost-Xin-Yang refined the tools

developed in [HJW80] and [JX99] to obtain a-priori estimates for harmonic maps,

improving higher codimension Bernstein results. Here, we use the results stated

in the previous section to construct subsets of Grassmannian manifolds that

have property (?) . This method does not allow us to obtain Holder estimates,

but fortunately we can replace them by the result of Allard [All72], which is a

seminal result in geometric measure theory, to study the graph case and obtain

Bernstein-type results. For our purposes, Allard’s theorem reduces the case of

minimal submanifolds of Euclidean space to that of minimal submanifolds of

spheres. The corresponding Gauss map for minimal submanifolds of spheres is

still harmonic, so that the reasoning just described still works.

More concretely, while Lawson-Osserman cones appear in codimension 3 or

26

1.4 Main Results

higher, we prove a Moser-type result for codimension 2.

Theorem 1.4.3 (Moser’s theorem in codimension 2). Let zi = f i(x1, ..., xp),

i = 1, 2, be smooth functions defined everywhere in Rp. Suppose the graph

M = (x, f(x)) is a submanifold with parallel mean curvature in Rp+2. Suppose

that there exists a number β0 < +∞ such that

∆f (x) ≤ β0 for all x ∈ Rp,

where

∆f (x) :=

det

(δαβ +

∑i

f ixα(x)f ixβ(x)

) 12

.

Then f 1, f 2 are linear functions on Rp representing an affine p-plane in Rp+2.

This is the main result of this thesis and it appears, together with section 2.3 and

Chapter 3, in [AJ18].

27

2 Preliminaries

The aim of this chapter is to provide all the material necessary for the proofs of

our main results stated in the previous chapter. We start with classical maximum

principles, a very important tool to study harmonic maps. Then, we dedicate

an entire section to a maximum principle due to Sampson, because it will play

the most significant role in this thesis. Our final paragraph is dedicated to

Grassmannian manifolds, where we will carefully present a comprehensive survey

of the geometry of these manifolds.

2.1 Maximum principles

Once more, we follow Xin [Xin12]. To study the existence of non-trivial harmonic

maps into some prescribed region of a manifold, we use several versions of the

maximum principle. For instance the following.

Theorem 2.1.1 (Maximum principle for functions). Let U be a domain of (N, h)

and f : (N, h) → R a subharmonic function having a maximum at an interior

point of U . Then f is constant.

A proof of this theorem can be found in Jost [Jos08] or Xin [Xin12].

Definition 2.1.2. A function f : U → R defined on an open set of the manifold

(N, h) is called convex (resp. strictly convex) if, and only if, ∇df ≥ 0 (resp. > 0).

If c : (−ε, ε)→ N is a geodesic arc, then the composition formula implies

(f c)′′ = (f c)/dt2 = ∇d(f c)

= ∇df(c′, c′) + df ∇dc (2.1.1)

= ∇df(c′, c′)

This gives us the following geometric interpretation of convex functions defined

on Riemannian manifolds.

29

2 Preliminaries

Proposition 2.1.3. Let f : N → R be a C2-function. Then f is convex (resp.

strictly convex) if and only if, for every geodesic arc c : (−ε, ε) → N , we have

(f c)′′ ≥ 0 (resp. > 0).

Definition 2.1.4 (Convex supporting domain). Let (N, h) be a complete Rie-

mannian manifold. A domain V ⊂ (N, h) is called a convex supporting domain if

there exists a C2-function f : V → R that is strictly convex.

Convex functions can be used to characterize harmonic and totally geodesic

maps.

Proposition 2.1.5. Let u : M −→ N be C2. The map u is totally geodesic if and

only if the pullback f u of any convex function f : U −→ R on a neighborhood of

the image of M under the map u is convex in f−1(U). The map u is harmonic if

and only if the pullback f u of any convex function f : U −→ R in a neighborhood

of u(M) is a subharmonic function.

Proof. From Equation (1.2.7), if the map u is totally geodesic and f is convex,

then∇d(f u) = df ∇du+∇df(du, du)

= 0 +∇df(du, du) ≥ 0

Therefore f u is convex. On the other hand, if u is not totally geodesic, there

exists a point x0 ∈M and a vector v ∈ Tx0M satisfying Bvv(u) 6= 0. We denote

by w the vector Bvv(f)|x0 ∈ Tu(x0)N and choose normal coordinates (yα) around

u(x0), to define

f(y) := bαyα +

∑(yα)2,

where bα will be determined later. Near u(x0), we have Hess(f)|u(x0) = Id, where

Id denotes the identity matrix. Now, taking bα such that bαwα < −|du(v)|2,

where wα is given by w = wα ∂∂yα

, we have

df(w)|u(x0) = bαwα.

Thus Equation (1.2.7) gives

Bvv(f u)|x0 = Hess(f)(u∗v, u∗v)|u(x0) + df(w)|u(x0)

= |du(v)|2 + bαwα < 0,

which contradicts the hypothesis of f u being convex near x0.

The second part requires a similar argument. For instance, if f is convex and

u harmonic, using Equation (1.2.7), we have

∆(f u) = Bu∗ei u∗ei(f) ≥ 0

30

2.1 Maximum principles

Hence (f u) is a subharmonic function. For the converse, if u is not harmonic,

then we can pick a point x0 ∈ M with τ(u)(x0) = w 6= 0. We choose a convex

function f near x0 such that

df(w) < 2 exp(u)(x0)

and

Hess(f)∣∣u(x0)

= Id .

The composition formula gives

∆(f u)|x0 = τ(f u)|x0 = 2 exp(u)(x0) + df(τ(u))x0= 2 exp(u)(x0) + df(w) < 0

which means that f u is not subharmonic near x0.

With these results and definitions in mind, we reach the following conclusion.

Theorem 2.1.6. Let (M, g) be a closed and (N, h) be a compact Riemannian

manifolds. If u : M → N is a harmonic map whose image is contained in a

domain V ⊂ N admitting a strictly convex function f , then u is constant.

Proof. By Proposition 2.1.5, we have

−∆(f u) = tr∇df(du, du) ≥ 0,

Therefore f φ is subharmonic. Since f u(M) ⊂ R is compact, u attains a

minimum and a maximum in M . But ∂M = ∅, so there exists a maximum in the

interior of a domain U ⊂M ; u(U) ⊂ V . Therefore f c is constant. Now, since

−∆(f u) = 0, (2.1.2)

we have

−∆(f u) = tr∇df(du, du) = 0. (2.1.3)

Therefore, if u is not a constant, we may find a vector X ∈ TpM such that dup.X ∈Tu(p)V and ∇df(dup.X, dup.X) > 0. This contradicts Equation (2.1.2).

Remark 2.1.7. The theorem above tells us that the image of a harmonic map

cannot be contained in a convex supporting domain. Of course, a closed geodesic

γ : S1 → N is a harmonic map and so γ(S1) * V ⊂ N , once V admits a strictly

convex function. Then the non-existence of closed geodesics on an open, connected

subset of a complete Riemannian manifold is a necessary condition for this subset

to support a strictly convex function. This remark will be of importance in the

next chapters.

31

2 Preliminaries

2.2 Sampson’s maximum principle

A beautiful result in the theory of harmonic maps is Sampson’s maximum principle.

Theorem 2.2.1 (Sampson’s maximum principle). Let u : (M, g) −→ (N, h) be

a non-constant harmonic map, where M is a compact Riemannian manifold, N

is a complete Riemannian manifold, and S ⊂ N is a hypersurface with definite

second fundamental form at a point y = u(x). Then no neighborhood of x ∈M is

mapped entirely to the concave side of S.

To prove this theorem, we need the following lemma.

Lemma 2.2.2. Let S be a hypersurface in N with definite second fundamental

form at a point y0 ∈ S. Then there exists a strictly convex function f : V −→ Rin a neighborhood V ⊂ N of y0 with f−1(0) = S ∩ V and f < 0 on the concave

side.

Proof. We can pick local coordinates (u1, ..., un) on V around y0 such that S∩V is

defined by un = 0 and the coordinate curve along un is a geodesic with arc length

|un| starting from the hypersurface S. The side to which the second fundamental

form B points, called the concave side, is such that the coordinate un is negative.

Moreover, the (u1, ..., un) induce coordinates (u1, ..., un−1) in S ∩ V . Denoting

by i : (u1, ..., un−1) 7−→ (u1, ..., un−1, 0) the standard embedding of S ∩ V into

N and by un the function un : (u1, ..., un) 7−→ un, we note that for any nonzero

X ∈ Ty0SHess(un)(i∗X, i∗X) = −dun(B(X,X)). (2.2.1)

But since un was chosen such that its derivative is negative in the direction

B(X,X), that is dun(B(X,X)) < 0, we have Hess(un)(i∗X, i∗X) > 0 and

Hess(un)

(∂

∂un,∂

∂un

)= 0

Define the function f = (un + 1)2 − 1 on V . Then for any Y ∈ Ty0N , we have

Hess(f)(Y, Y )|y0 = 2(∇Y un)2 + 2 Hess(un)(Y, Y ).

By taking a nonzero X ∈ Ty0S we get

Hess(f)(X,X)∣∣y0

= 2 Hess(un)(X,X) > 0 (2.2.2)

and

Hess(f)

(∂

∂un,∂

∂un

)= 2.

32

2.3 Grassmannian Manifolds

This shows that f is a strictly convex function. But obviously, f(y) = 0 if and

only if u2n(y) + 2un(y) + 1 − 1 = 0, which happens if and only if un(y) = 0 or

un(y) = −2, but we can choose V such that we have f(y) = 0 if and only if

un(y) = 0. Therefore f−1(0) = S ∩ V . Finally, on the concave side of V we have

(un + 1)2 < 1.

Proof of Theorem 2.2.1: By the previous lemma, there exists a strictly convex

function f on a neighborhood V of the point y. Suppose there exists a neigh-

borhood U of x such that u(U) lies in the concave side of S ∩ V . Then for any

x ∈ U , we have f(u(x)) ≤ 0 = f u(x), which shows that x is a maximum point

of f u in U . By the maximum principle for maps, 2.1.6, f u is constant in U .

Since u is harmonic, the composition formula (1.2.7) gives

0 = ∆(f u) = Hess(f)(u∗ei, u∗ei). (2.2.3)

Since f is strictly convex, u is constant in V , which contradicts the assumption

that u is a nonconstant harmonic map.

2.3 Grassmannian Manifolds

We closely follow the excellent presentation of Kozlov [Koz97] in the first part of

this section. The remaining material can be found in Jost-Xin[JX99].

Let V n be an n-dimensional vector space over R with inner product 〈., .〉. One

defines a 2n-dimensional algebra with respect to the exterior product ∧ by

Λ(V ) = ⊕∞i=0Λi(V )

such that Λ0(V ) = R, Λ1(V ) = V and Λi>n(V ) = 0. Λ(V ) is called the Grassmann

algebra and the elements of Λp(V ) are called p-vectors. If V = Rn, we denote

Λp(Rn) by Λp.

Let eini=1 be a basis for V and λ = (i1, ..., ip) ∈ Λ(n, p) = (i1, ...ip); 1 ≤ i1 <

... < ip ≤ n. We denote by eλ the p-vector

eλ := ei1 ∧ ... ∧ eip .

eλλ∈Λ(n,p) is a basis for the(np

)-dimensional vector space Λp(V ) ⊂ Λ(V ). For

any w ∈ Λp(V ),

w =∑

λ∈Λ(n,p)

wλeλ.

33

2 Preliminaries

Definition 2.3.1 (Simple vector). A p-vector w ∈ Λp(V ) is called simple if it can

be represented as the exterior product of p elements of V, that is, w = af1∧ ...∧fp,with a ∈ R and fi ∈ V for every i ∈ 1, ..., p.

Definition 2.3.2 (Scalar product of p-vectors). The scalar product of two p-

vectors w and w is the number

〈w, w〉 :=∑

λ∈Λ(n,p)

wλwλ.

This is a scalar product and does not depend on the choice of the orthonormal

basis. We also have |w| = 〈w,w〉12 .

Definition 2.3.3 (Inner multiplication of p-vectors). An operation of inner

multiplication is a bilinear map

x: Λp(V )× Λq(V ) −→ Λp−q(V ), p ≥ q ≥ 0

(ω, ξ) 7→ wxξ

such that for any ω ∈ Λp(V ), ξ ∈ Λq(V ), and ϕ ∈ Λp−q(V ) it satisfies

〈ωxξ, ϕ〉 = 〈ω, ξ ∧ ϕ〉.

Definition 2.3.4 (The rank space). The rank space of a p-vector w ∈ Λp(V ),

p ≥ 1, is defined as

Vw = wxΛp−1(V ) = e ∈ V : e = wxv, v ∈ Λp−1(V ) ⊂ V. (2.3.1)

That is, Vw ⊂ V is the minimum subspace W ⊂ V such that w ∈ Λp(W ).

Remark 2.3.5. A nonzero w ∈ Λp(V ) is simple if and only if dimVw = p.

Therefore there is a one-to-one correspondence between the set of oriented p-

planes and the set of unit simple p-vectors. This will be important when we

define the so called Plucker embedding of the Grassmannian manifold.

Definition 2.3.6 (Grasmannian manifold G+p,n). Let Rn be the n-dimensional

Euclidean space. The set of all oriented p-subspaces (called p-planes) consti-

tutes the Grassmannian manifold G+p,n, which is the irreducible symmetric space

SO(n)/(SO(p)× SO(q)), where q = n− p.

34

2.3 Grassmannian Manifolds

For each p-plane V0 ∈ G+p,n, consider the open neighborhood U(V0) of all oriented

p-planes whose orthogonal projection onto V0 is one-to-one. That is, if eipi=1 is

an orthonormal basis of V0, and nαqα=1 is an orthonormal basis of V ⊥0 , one can

write, for V ∈ U(V0),

ei(V ) = ei +Ni (2.3.2)

where Ni ∈ V ⊥0 and ei(V ) ∈ V is such that ei = PrV0(ei(V )).

Decomposing the vector Ni using the basis nα we get a matrix a = (aαi ) ∈ Rp.q

defined by

Ni = aαi nα (2.3.3)

that can be regarded as local coordinates of the p-plane V in U(V0).

Using the one-to-one correspondence between unit simple vectors and oriented

p-planes V ∈ G+p,n, we define the Plucker embedding as

ψ : G+p,n −→ Λp

ψ(V ) :=ψ(V )

|ψ(V )|,

(2.3.4)

where

ψ(V ) := (e1 +N1) ∧ ... ∧ (ep +Np). (2.3.5)

Via the Plucker embedding, G+p,n can be viewed as a submanifold of the Euclidean

space R(np). The restriction of the Euclidean inner product denoted by ω :

G+p,n ×G+

p,n −→ R is

ω(P,Q) =〈ψ(P ), ψ(Q)〉

〈ψ(P ), ψ(P )〉 12 〈ψ(Q), ψ(Q)〉 12(2.3.6)

If e1, ..., ep is an oriented orthonormal basis of P and f1, ..., fp is an oriented

orthonormal basis of Q, then

w(P,Q) = 〈e1 ∧ ... ∧ ep, f1 ∧ ... ∧ fp〉 = detW, (2.3.7)

where W = (〈ei, fj〉).

Remark 2.3.7. Note that ψ(G+p,n) = Kp ∩ S(np)−1 ⊂ R(np) ∼= Λp.

Remark 2.3.8. Let w ∈ ψ(G+p,n) and eipi=1, nαqα=1 be orthonormal bases for

Vw and V ⊥w , respectively. The system of vectors ηiαp , qi=1,α=1 where

ηiα = e1 ∧ ... ∧ ei−1 ∧ nα ∧ ei+1 ∧ ... ∧ ep (2.3.8)

for i ∈ 1, ..., p and α ∈ 1, ..., q is an orthonormal basis of the tangent space

Tw(ψ(G+p,n

)).

35

2 Preliminaries

We write G+p,n instead of ψ(G+

p,n) for the image of the Grassmannian manifold

under the Plucker embedding.

Let (w,X) ∈ TG+p,n be an element of the tangent bundle and ηiαα=1,...,q

i=1,...,p a basis

for TwG+p,n like above. There exist mi ∈ V ⊥w such that

X = m1 ∧ e2 ∧ ... ∧ ep + ...+ e1 ∧ ... ∧ ep−1 ∧mp (2.3.9)

These mi are not necessarily pairwise orthogonal.

Theorem 2.3.9 (Kozlov). Let w ∈ G+p,n and X ∈ TwG

+p,n, X 6= 0. Then

there exists an orthonormal basis eipi=1 in Vw and a system miri=1, with

1 ≤ r ≤ minp, q, of non-zero pairwise orthogonal vectors in V ⊥w , such that

w =e1 ∧ ... ∧ ep, (2.3.10)

X =(m1 ∧ e2 ∧ ... ∧ er + ...+ e1 ∧ ... ∧ er−1 ∧mr) ∧ (er+1 ∧ ... ∧ ep). (2.3.11)

Proof. If we fix an orthonormal basis for Vw, we can write the vector X ∈ TwG+p,n

in the form of Equation (2.3.9), where some of the mj might be zero. As X 6= 0

we can consider the set of all orthonormal bases for Vw such that the number

of non-zero elements in the expression of X is minimal. Among such bases, we

consider just the ones that also minimize the product |m1|...|mr|.If r = 1 the theorem is proven. Let r ≥ 2 and note that if we rotate (e1, e2) by

the angle θ we get

e1 = e1 cos(θ) + e2 sin(θ), e2 = −e1 cos(θ) + e2 sin(θ).

Thereofre e1 ∧ e2 = e1 ∧ e2 and

m1 ∧ e2 + e1 ∧m2 = (m1 cos(θ)−m2 sin(θ)) ∧ e2

+ e1 ∧ (m1 sin(θ) +m2 cos(θ)).

Hence,

m1 = m1 cos(θ)−m2 sin(θ), m2 = m1 sin(θ) +m2 cos(θ),

and mi = mi for i > 2, therefore m1∧m2 = m1∧m2. It remains to be proven that

m1,m2 are orthogonal when their product attains the minimum, when |m1| · |m2|.This follows from the simple fact that

|m1 ∧m2|2 = |m1|2|m2|2 − 〈m1,m2〉2

36

2.3 Grassmannian Manifolds

The minimum of |m1|.|m2| corresponds to the one of |〈m1,m2〉|. But taking the

derivative with respect to θ of the equation

〈m1, m2〉 = 〈m1,m2〉 cos(2θ) +|m1|2 − |m2|2

2sin(2θ)

we have that the vectors m1 and m2 are orthogonal exactly when |m1| · |m2|attains its minimum.

Notation 2.3.10. Based on Equation (2.3.10) and Equation (2.3.11), we write

λi := |mi|, ni :=mi

λi,

ei(s) := ei cos(s) + ni sin(s), ni(s) := −ei sin(s) + ni cos(s),

X0 := er+1 ∧ ... ∧ ep.

Note that e′i(s) = ni(s) and n′i = −ei(s).

2.3.1 Closed geodesics in G+p,n.

Consider the curve

w(t) := e1(λ1t) ∧ ... ∧ er(λrt) ∧X0. (2.3.12)

Theorem 2.3.11. The curve w(t) is a normal geodesic on the manifold G+p,n and

expwX = w(1). (2.3.13)

Sketch of the proof. For each t ∈ R, ei(λit), ni(λit)ri=1 are pairwise orthogonal

vectors and

e′i(λit) =λini(λ

it), n′i(λit) = −λiei(λit),

X(t) :=w′(t) =r∑i=1

λini(λit) ∧ (w(t)xei(λ

it)), (2.3.14)

w(0) = w, w′(0) = X, |X(t)| =

[r∑i=1

(λi)2

] 12

.

Therefore w(t) is parametrized proportionally to the arc-length. Moreover, taking

an additional derivative, we have

w′′(t) = −|X|2w(t) + ξ(t)

37

2 Preliminaries

where ξ(t) is a term like Equation (2.3.14) with ni replacing the ei and therefore

orthogonal to the tangent plane Tw(t)G+p,n.

We will denote the above geodesic by

wX(t) := e1(λ1t) ∧ ... ∧ er(λrt) ∧X0, (2.3.15)

where X = (λ1n1∧e2∧ ...∧er +λ2e1∧n2∧ ...∧er + ...+λre1∧ ...∧er−1∧er)∧X0.

Thus a geodesic in G+p,n between two p-planes is obtained by rotating one into

the other in the Euclidean space, by rotating corresponding basis vectors. As an

example take the 2-plane spanned by e1, e2 in R4: one tangent direction in G+2,4

would be to move e1 into e3 and keep e2 fixed, and the other tangent direction

would be to move e2 into e4 and keep e1 fixed. In other words, we are taking

X, Y ∈ TwG+2,4, for some w ∈ G+

2,4, where X = e3 ∧ e2 and Y = e1 ∧ e4 and

the respective geodesics are given by equation (2.3.15). We can also consider

Z ∈ TwG+2,4 where Z :=

(1√2e3 ∧ e2 + 1√

2e1 ∧ e4

)and the respective geodesic

wZ(t) given by Equation (2.3.15) is obtained when we simultaneously rotate e1

into e3 and e2 into e4. This geometric picture helps to understand our subsequent

constructions. Later we compute the length of these different types of geodesics

on a general Grassmannian.

Theorem 2.3.12. For G+p,n, we have

diam(G+p,n) = max

π,√r0π/2

, (2.3.16)

where r0 = minp, q. More explicitly,

r0 < 4⇒ diam(G+p,n) = π,

r0 ≥ 4⇒ diam(G+p,n) =

√r0π

2

For the next examples, we have to understand the closed geodesics in G+p,n.

Theorem 2.3.13. For geodesics in G+p,n with parametrization (2.3.15), we have

that

wX(t1) = wX(t2) ⇐⇒ ∃k ∈ Zr; λi(t2 − t1) = πki, andr∑i=1

ki ≡ 0 mod 2.

A proof of this statement can be found in Kozlov [Koz97].

Remark 2.3.14. Every geodesic loop in G+p,n is a closed geodesic.

This assertion can be used to compute lengths of closed geodesics in G+p,n, because

L(wX) = |X| =

[r∑i=1

(λi)2

] 12

. (2.3.17)

38

2.3 Grassmannian Manifolds

2.3.2 Geodesically convex sets in G+p,n

We are interested in the convex subsets of a Grassmannian manifold.

Let w ∈ G+p,n and let X ∈ TwG+

p,n be a unit tangent vector. We know that there

exists an orthonormal basis ei, nαα=1,...,qi=1,...,p of Rn and an integer r ≤ min(p, q)

such that w = spanei, n = p+ q,

X =(λ1n1 ∧ e2 ∧ ... ∧ er + ...+ λre1 ∧ ... ∧ er−1 ∧ nr

)∧X0, (2.3.18)

and |X| =(∑r

α=1|λα|2) 1

2 = 1.

Definition 2.3.15. Let X ∈ TwG+p,n be given by Equation (2.3.18). Define

tX :=π

2(|λα′|+ |λβ′|)(2.3.19)

where |λα′| := maxλα and |λβ′| := maxλα; λα 6= λα′.

Definition 2.3.16. Let w ∈ G+p,n and X ∈ TwG+

p,n be a unit tangent vector as in

Equation (2.3.18). Define

BG(w) :=wX(t) ∈ G+

p,n; 0 ≤ t ≤ tX.

Theorem 2.3.17 (Jost-Xin [JX99]). The set BG(w) is a (geodesically) convex

set and contains the largest geodesic ball centered at w.

A detailed proof of this statement can be found in Jost-Xin [JX99].

Now we are going to present some explicit examples of geodesics in G+p,n and

compute at which time they intersect the region we have defined above. Those

examples clarify how geodesics behave in Grassmannian manifolds, and thereby

the geometry of BG(w).

Example 2.3.18. Let w = e1 ∧ ...∧ ep ∈ G+p,n and denote X1 = n1 ∧ e2 ∧ ...∧ ep.

From the notation above λ1 = 1 and λi = 0 for every i 6= 1. Moreover |X1| = 1

and tX1 = π2. Thus the length of this closed geodesic is 2π. Moreover, denoting

by |L(wX(t))| the length of the segment connecting wX(0) and wX(t), we have

|L(wX1(tX1))| = π2. Clearly, wX1(π) = −w and, as G+

p,n is a symmetric space,

wX1(−ε) = w−X1(ε), so wX1 : R −→ G+p,n is a geodesic such that γ|[−π2 ,π2 ] ⊂

BG(w).

It is known that G+p,n is a submanifold of S(np)−1 ⊂ R(np) with the induced metric

of the sphere. So in the direction X1, BG(w) contains half of a great circle

connecting two antipodal points.

39

2 Preliminaries

Example 2.3.19. Consider w = e1 ∧ ... ∧ ep ∈ G+p,n and X2 = 1√

2(n1 ∧ e2 +

e1 ∧ n2) ∧ e3 ∧ ... ∧ ep. Therefore X2 ∈ TwG+p,n with |X2| = 1. Moreover in the

definition of BG(w):

tX2 =π

2( 1√2

+ 1√2)

=

√2π

4=

√2

2

π

2<π

2(2.3.20)

Thus, to make tX as small as possible, we need an equidistribution of the

eigenvalues λα of X satisfying∑

α|λα| = 1 and we maximize λmax + λmax′ . In

this case, wX2

(√2π4

)∈ ∂BG(w) is the point in the boundary closest to w. Also,

by Theorem 2.3.13

wX2(t) = wX2(0) = w ⇐⇒ ∃k1, k2 ∈ Z2; λi(t− 0) = π.ki

and

k1 + k2 = 0 (mod 2)

But

λ1 = λ2 = 1/√

2 , therefore t =√

2k1π =√

2k2π

Then k1 = k2 = 1 solves the equation and tells us that wX2

(√2π)

= w, and, as

wX2 is parametrized by arc length,

L(wX2) =√

2π = 4tX2 . (2.3.21)

It is proven in [Koz97] that this is the smallest non-trivial closed geodesic in G+p,n,

when r0 > 1.

It will be crucial that the equation for the geodesic wX2 is

wX2(t) =(e1 cos

(t√2

)+ n1 sin

(t√2

))∧(e2 cos

(t√2

)+ n2 sin

(t√2

))∧(e3 ∧ ... ∧ ep)

(2.3.22)

So we compute, for any t ∈ R,

40

2.3 Grassmannian Manifolds

〈wX2(t), w〉 =⟨(e1 cos(t/

√2) + n1 sin(t/

√2))

∧(e2 cos(t/

√2) + n2 sin(t/

√2)), (e1 ∧ e2)

⟩·

⟨(e3 ∧ ... ∧ ep), (e3 ∧ ... ∧ ep)

⟩=

⟨e1 cos

(t/√

2), e1

⟩·⟨e2 cos

(t/√

2), e2

⟩= cos2

(t√2

)≥ 0

(2.3.23)

Therefore, using the fact that G+p,n ⊂ S(np)−1 and denoting the antipodal point in

S(np) by −w, the geodesic wX2 never leaves the hemisphere of S(np) that has w as

a pole. In particular wX2(t ∈ R+) does not intersect BG(−w).

Example 2.3.20. Consider w = e1∧...∧ep ∈ G+p,n andXr0 = 1√

r0

(n1 ∧ e2 ∧ ... ∧ er0 + ...+ e1 ∧ ... ∧ er0−1 ∧ nrr0

)∧

er0+1 ∧ ... ∧ ep, so that Xr0 ∈ TwG+p,n with |Xr0| = 1. Therefore

tXr0 =π

2(

1√r0

+ 1√r0

) =

√r0π

4(2.3.24)

and the geodesic wXr0 lies inside BG(w). G+p,n is a symmetric space, thus the curve

wXr0 : (−√r0π

4,√r0π

4) −→ G+

p,n is a geodesic with w−Xr0

(−√r0π

4

)= wXr0

(√r0π

4

)and

wXr0 (t) = wXr0 (0) = w ⇐⇒ ∃k1, ..., kr0 ∈ Zr0 ; λi(t− 0) = π.ki

and

r0∑r0=1

ki = 0 mod 2(2.3.25)

Thus λi = 1/√r0 and t =

√r0k

iπ for every i ∈ 1, ...r0.Consequently, if r0 is an even number, then ki = 1 solves the equation and

wXr0(√

r0π)

= w, and, as wXr0 is parametrized by arc length,∣∣L(wXr0 )

∣∣ =√r0π (= 4tr0). If r0 is an odd number, then ki = 2 solves the equation and

wXr0(√

r0π)

= w, and, as wXr0 is parametrized by arc length,∣∣L(wXr0 )

∣∣ =

2√r0π (= 8tr0).

The equation for the geodesic wXr0 is

41

2 Preliminaries

wXr0 (t) =(e1 cos

(t√r0

)+ n1 sin

(t√r0

))∧

(e2 cos

(t√r0

)+ n2 sin

(t√r0

))∧

...∧(er0 cos

(t√r0

)+ nr0 sin

(t√r0

))∧ (er0+1 ∧ ... ∧ ep),

(2.3.26)

so we compute, for any t ∈ R+,

〈wXr0 (t), w〉 = 〈(e1 cos(t/√r0) + n1 sin(t/

√r0)), e1〉·

〈(e2 cos(t/√r0) + n2 sin(t/

√r0)), e2〉 · · ·

〈(ei cos(t/√r0) + ni sin(t/

√r0)), ei〉 · · ·

〈(er0 cos(t/√r0) + nr0 sin(t/

√r0)), 〉·

〈(er0+1 ∧ ... ∧ ep), (er0+1 ∧ ... ∧ ep)〉

= 〈e1 cos(t/√r0), e1〉 · · · 〈er0 cos(t/

√r0), er0〉

= cosr0(t/√r0).

(2.3.27)

Note that, if r0 is odd, then wXr0 (0) and wXr0(√

r0π)

are antipodal points in

S(np)−1. Also, 〈wXr0 (t), w〉 ≥ 0 and wX0 does not intersect BG(−w) if and only if

r0 is even.

Notation 2.3.21 (Eigendirections on the tangent space). The elements of the

tangent space Y ∈ TwG+p,n given by the rotation of k ∈ 1, ..., r0 basis vectors

into k normal vectors with the same speed (i.e., an integer multiple of λk) will be

called vectors of type k. For example, in Example 2.3.18, X1 ∈ TwG+p,n is a vector

of type 1. In Example 2.3.19, X2 ∈ TwG+p,n is a vector of type 2.

2.3.3 The slope of a graph and the Gauss map

In our main theorem we use the concept of slope of a graph, defined by

∆f (x) :=

det

(δαβ +

∑i

f ixα(x)f ixβ(x)

) 12

, (2.3.28)

42

2.3 Grassmannian Manifolds

Its relation to harmonic Gauss maps plays an important role which we now

describe.

Let fα(x1, ..., xn) : Rn −→ R be smooth functions with α = 1, ...,m and let

M =(x1, ..., xn, f 1(x1, ..., xn), ..., fm(x1, ..., xn)

)⊂ Rn+m,

that is M = graph(f) ⊂ Rn+m.

Denote by F : Rn −→ M the diffeomorphism x 7→ (x, f(x)), where f =

(f 1, ..., fn). M is an n-dimensional submanifold of the Euclidean space Rn+m

with the induced metric g = gijdxidxj. With a basis ei, ηα for Rn+m, where

ei spans Rn, we can write this metric as

gij =

⟨F∗

∂xi, F∗

∂xj

⟩=

⟨ei +

∂fα

∂xiηα, ej +

∂fα

∂xjηα

⟩= δij +

∂fα

∂xi∂fα

∂xj,

so denoting by Df :=(∂fα

∂xi

), a (n×m)-matrix,

(gij) = In +Df(Df)T (2.3.29)

But also, taking w, v ∈ Gn,n+m such that w = e1∧ ...∧en, v = en+1∧e2∧ ...∧enand using that

ψ γ =

(e1 +

∂fα

∂x1ηα

)∧ ... ∧

(en +

∂fα

∂xnηα

). (2.3.30)

where ψ : G+n,n+m −→ R( n

n+m) is the Plucker embedding, we see that

ω(γ, w) =〈ψ γ, w〉

〈ψ γ, ψ γ〉 12 〈w,w〉 12= ∆−1

f , (2.3.31)

ω(γ, v) =〈ψ γ, v〉

〈ψ γ, ψ γ〉 12 〈v, v〉 12=∂f 1

∂x1∆−1f , (2.3.32)

where ω(., .) is the restriction of the Euclidean inner product to the image of the

Plucker embedding, as defined in section 3.

The above equations tell us that Moser’s condition of bounded slope of graph(f),

∆f <∞, measures, in terms of the geometry of Grassmannians, that there exists

w ∈ γ(M) such that γ(M) ⊂ G+n,n+m ∩ Hem

( nn+m)−1w , where Hem

( nn+m)−1w is the

hemisphere in S( nn+m)−1 centered at w. Obviously, as in the codimension 1 case,

G+n,n+1 = Sn and the hemisphere of a point is a convex set. We have a proof for

Moser’s Bernstein-type theorem [Mos61].

43

3 Proof of main results

3.1 The Geometry of the Maximum Principle

3.1.1 An application of Sampson’s maximum principle

The aim of this section is to use Sampson’s maximum principle to derive a general

method for proving that a given region in a manifold has the property (?). Since

the mentioned maximum principle deal with convex hypersurfaces, we start by

recalling a standard property of convex geodesic balls.

Remark 3.1.1. Take a geodesic ball B(p, r) in a complete manifold N such that

r is smaller than the convexity radius of N at p. Then ∂B(p, r) is a hypersurface

of N with definite second fundamental form for every point q ∈ ∂B(p, r).

The main result of this thesis can be seen as a corollary of Sampson’s maximum

principle.

Theorem 3.1.2. Let (N, h) be a complete Riemannian manifold and Γ : [a, b] −→N a smooth embedded curve. Consider a smooth function r : [a, b] −→ R+ and a

region

R :=⋃t∈[a,b]

B(Γ(t), r(t)), (3.1.1)

where B(·, ·) is the geodesic ball and r(t) is smaller than the convexity radius of

N for any t. If, for each t0 ∈ (a, b), the set R\B(Γ(t0), r(t0)) is the union of two

disjoint connected sets, namely the connected component of Γ(a) and the one of

Γ(b), then there exists no compact manifold (M, g) and non-constant harmonic

map φ : M −→ N such that φ(M) ⊂ R. In other words, R has the property (?)

of Definition 1.4.1.

Before presenting the proof of this statement, we make two remarks on the

geometry of the region R

45

3 Proof of main results

φ harmonic

M cpt, ∂M = ∅

R :=⋃t∈[a,b] Bg(Γ(t), r)

r r r y1 = φ(x1)y = φ(x)

x1x

Γ(a) Γ(t)Γ(b)

Figure 3.1: A region R as the union of convex balls

Remark 3.1.3. By hypothesis, for each time t0 ∈ [a, b], the set R\B(Γ(t0), r(t0))

is the union of two disjoint connected components. This condition implicitly rules

out a lot of possibilities for both Γ and r. Namely, suppose we have a manifold

with large convex radius. We could define R as the union of balls over a geodesic

of unit speed Γ of a very small length ε with radius varying smoothly from ε/2

at t = 0, to 10ε at t = ε/2, and back to ε/2 when t = ε. This region does not

satisfies the necessary condition for our theorem.

On the other hand, suppose we have a curve Γ with the shape of a very steep ‘U ’.

Our condition implies that, in this case, the radius of the balls near the turning

point of the ‘U’ must be very small, otherwise there will be some intersection

between the connected components of Γ(a) and Γ(b).

Lemma 3.1.4. Let R be a region defined as in Theorem 3.1.2 satifying the same

hypothesis. Let t0 ∈ (a, b) and consider the open set

Cb(t) :=⋃

t∈[t0,b)

B(Γ(t), t(t))\(B(Γ(t0), r(t0))

)(3.1.2)

This is the connected component of Γ(b) in R\(B(Γ(t0), r(t0))). Then, for every

point y ∈ Cb(t0), there exists t1 > t0 such that y ∈ Cb(t1) and y ∈ ∂B(Γ(t1), r(t1)).

46

3.1 The Geometry of the Maximum Principle

Proof. By definition of Cb(t), whenever s′ > s, we have Cb(s′) ( Cb(s). Since

Cb(t0) is an open set, there exists δ > 0 such that B(y, δ) ⊂ Cb(t0). Therefore, as

r is a smooth function of t, there exists δ(δ, t0) > 0 such that y ∈ Cb(t′), where

t′ := t0 + δ. Moreover, the set

A =t ∈ [t0 + δ, b− δ] | y ∈ Cb(t)

is non-empty, since t′ ∈ A. Now set t1 = max(A) and the lemma follows.

Proof of Theorem 3.1.2: (See Figure 3.1) We argue by contradiction. Suppose

that there exists a compact manifold (M, g) and a non-constant harmonic map

φ : M −→ N such that φ(M) ⊂ R. Let φ(x) = y ∈ R be a point in φ(M). By

definition of R and the fact that φ is non-constant, there exists t0 ∈ [a, b] such

that,

y ∈ ∂B(Γ(t0), r(t0)) ∩ Cb(t0),

where Cb(t) is defined by Equation (3.1.2). Therefore, by Sampson’s maximum

principle, for every open neighborhood Ux of x in M ,

φ(Ux) * B(Γ(t0), r(t0))

That is, there exists x1 ∈ Ux such that φ(x1) = y1 is mapped to the convex side

of B(Γ(t0, r(t0))). Moreover, as we assume that R\B(Γ(t0), r(t0)) is disconnected,

y1 ∈ Cb(t0). Therefore, by Lemma 3.1.4, there exists a t1 > t0 such that

y1 ∈ ∂B(Γ(t1), r(t1)) ∩ Cb(t1)

Again, by Sampson’s maximum principle, every open neighborhood Ux1 of x1 in

M satisfies

φ(Ux1) * B(Γ(t1), r(t1)),

that is, there exists x2 ∈ Ux1 such that φ(x2) =: y2 /∈ B(Γ(t1), r(t1)), y2 is in the

convex side of ∂B(Γ(t1), r(t1)) and in the same connected component as Γ(b). By

induction, we construct a sequence yi, tii∈N, such that for every i ∈ N we have

ti > ti+1, limi→+∞ ti = b, yi ∈ φ(M), and yi ∈ Cb(ti). Since

Cb(ti) ( Cb(ti−1),

it follows that, denoting y∞ = limi→+∞ yi, that y∞ ∈ ∂B(Γ(b), r(b)) ∩ ∂R. Once

more, by Sampson’s maximum principle at y∞ = φ(x∞), every neighborhood

Ux∞ of x∞ in M has a point x′ such that φ(x′) = y′ is in the convex side of

∂B(Γ(p), r(p)) ∩ ∂R. Therefore y′ = φ(x′) /∈ R.

47

3 Proof of main results

p

A(p)

Figure 3.2: S2\(

12Eq)ε

This result has interesting applications.

Example 3.1.5. Let (N, h) = (S2, g) ⊂ (R3, geuc) and let Γ : [0, 1] −→ S2

be the unit speed geodesic such that Γ(0) = (−1, 0, 0), Γ(1/2) = (0, 0, 1) and

Γ(1) = (1, 0, 0). For a given ε > 0, let

R :=⋃t∈[0,1]

B(

Γ(t),π

2− ε

2

)(3.1.3)

and note that R contains the complement of an ε-neighborhood of the half equator

containing the points (0, 0,−1), (0,−1, 0) and (0, 1, 0), denoted by 12Eq. We write

R as S2\(

12Eq)ε. Theorem 3.1.2 implies that S2\

(12Eq)ε

has property (?).See

Figure 3.2.

Remark 3.1.6. The above argument also works for spheres Sn when we take out

a closed half equator, denoted by 12Sn−1. This result appears in [JXY12], where a

strictly convex function on each compact subset properly contained in Sn\12Sn−1

is constructed.

Remark 3.1.7. The main idea of the proof of Theorem 3.1.2 is that ∂R is a

barrier to the existence of non-constant harmonic maps. With the help of the

Sampson’s maximum principle, from now on abbreviated by SMP and the definition

48

3.1 The Geometry of the Maximum Principle

of R as a union of convex balls, we push the image of the harmonic map to this

barrier.

One can make the theorem more general by replacing ∂R ∩ ∂B(Γ(b), r(b)) with

some more flexible barrier. For example, with the family of balls defined in

Example 3.1.5, one would prove that there are no non-constant harmonic maps

whose image is contained in an open subset of S2\γ where γ is a connected curve

connecting the antipodal points (0, 1, 0), (0,−1, 0) ∈ S2.

Note that the classical method used to prove non-existence of such non-constant

harmonic maps, the method of looking for a strictly convex function f : R −→ R(in the geodesic sense), is not very flexible. Once one changes the boundary of R

slightly, one can no longer guarantee that there exists a strictly convex function

f : R −→ R.

3.1.2 Barriers to the existence of harmonic maps

Our proof of Theorem 3.1.2 gave the barrier for the existence of harmonic maps

as the boundary of the union of a family of balls whose centers are given by a

smooth embedded curve.

By studying barriers to the existence of harmonic maps abstractly, we find that

the result does not rely on the existence of such a family of balls along a curve.

The region R defined in Example 3.1.5 contains no closed geodesics, and since

geodesics are harmonic maps, it is necessary to avoid them for the non-existence

of non-constant harmonic maps. Therefore the first condition we are looking for

is one that avoids such curves.

Condition 1. For every p ∈ R and every v ∈ T 1pN , the geodesic Γ such that

Γ(0) = p and Γ′(0) = v satisfies, for some tv ∈ [0, K], Γ(tv) ∈ ∂R, where K is

independent of p and v.

Condition 1 is a qualitative way of imposing the non-existence of closed geodesics

in our region. But imposing the non-existence of closed geodesics in a given

region is not enough to avoid existence of harmonic maps, therefore we need an

extra condition to be stated later. But first, let us introduce some notation.

Notation 3.1.8. For p ∈ N and v ∈ TpN , we denote by γp,v : R −→ N the

geodesic that satisfies γp,v(0) = p and γ′p,v(0) = v.

As an example, suppose we have a point in the image of some geodesic, namely

Γp,v(t0). Take a tangent vector η ∈ TΓp,v(t0)N such that η 6= Γ′p,v(t0). If we

49

3 Proof of main results

consider another geodesic passing through Γp,v(t0) in the direction η at t = 0 ,

we write γΓp,v(t0),η(·) : R −→ N .

Notation 3.1.9. For p ∈ N , ν ∈ ∂Bg(p, r) with r 1, let tmax(p, ν) ∈ R+ be

the first time at which Γp,ν(t) ∈ ∂R.

We impose the following condition on the region R.

Condition 2. For every p ∈ R and every ν ∈ T 1pN , if t < tmax(p, ν), then the set

of directions η ∈ T 1Γp,ν(t)N such that d

(p, γRΓp,ν(t),η(ε)

)< d(p,Γp,ν(t)) is properly

contained in T 1Γp,ν(t) and connected. Here γR denotes a geodesic in R with ε ∈ R+

sufficiently small (that is, a curve that minimizes the length in R).

On one hand, Condition 2 is difficult to check unless N is a symmetric space.

On the other hand, in view of the example given in Chapter 1, Section 1.2.1 on

the existence of a non-trivial harmonic map from a genus two surface into S2,

we must have a very restrictive condition on R, namely Condition 2, to obtain a

result on the existence of harmonic maps as general as the one stated below.

Theorem 3.1.10. Let (N,h) be a complete Riemannian manifold and suppose

that for a region R ⊂ N , we have that R \ ∂R is open and connected. In addition,

suppose R satisfies the Conditions 1 and 2 above. Then R has property (?) of

Definition 1.4.1.

Proof. We argue by contradiction. Suppose that there exists a compact manifold

(M, g) and a non-constant harmonic map φ : M −→ N such that φ(M) ⊂ R.

Therefore, there exist points p = φ(x), q ∈ R such that d(q, p) 1 and a geodesic

Γ with Γ(0) = p, Γ′(0) = vp,q, and Γ(tp) = q. By Condition 1 there exists

tvp,q ∈ R+ such that Γ intersects ∂R. Define

Q0 := ∪t∈R+B(Γ(t), δ) with δ > δ0 > 0

and note that it satisfies the conditions of Theorem 3.1.2.

Therefore, when walking along Γ with the ball B(Γ(t), δ), there exists q1 /∈ Q0

such that q1 = φ(y1), y1 ∈M , and Condition 2 tells us that

dR(q1, p) > dR(q, p).

Let γR1 be a curve in R, connecting p and q1, of minimal length. Note that it

may not be a geodesic in N itself. By Condition 2, there exists η1 ∈ T 1q1N such

that Γq1,η1 increases the distance to p, that is

dR(Γq1,η1(t), p) > dR(q1, p)

50

3.1 The Geometry of the Maximum Principle

for sufficiently small t. As before, Γq1,η1 hits the boundary at some point tq1 by

Condition 1.

Let

Q1 := ∪t∈R+B(Γq1,η1(t), δ1), δ1 > δ0.

By Theorem 3.1.2, there exist q2 ∈ Q1 such that q2 = φ(y2) and

dR(q2, p) > dR(q1, p) > dR(q, p).

Inductively, we build a sequence of points q1, q2, q3, ... such that qi = φ(yi) with

yi ∈M for all i ∈ N and

dR(qi, p) = dR(qi, φ(x)) > dR(qi−1, φ(x)).

We claim that there exists k ∈ N such that qk /∈ R. Namely, by the compactness

of φ(M)∩R, the only possibility for this sequence to stay inside R is if there is q ∈ R

such that qi −→ q. Suppose this happens, then we can take a sufficiently small

ε > 0 such that dR(qi, q) < ε and consider the geodesic Γqi→q in N with Γqi(0) = qi

and Γ′qi(0) = νqi→q. The construction of a region Q1 := ∪t∈R+B(Γqi→q(t), δ) would

give us q1 = φ(y1) such that dR(p, q1) > dR(p, q), and we can continue invoking

this argument. Induction implies that such a k ∈ N exists.

Example 3.1.11. One can use the above theorem to find regions in symmetric

spaces for which property (?) holds. If one has some information on the geometry

of such a space, Conditions 1 and 2 may be easy to check. For instance, let

p ≥ 1 and (Σp, g) be a compact surface of genus p with a hyperbolic metric of

constant curvature. Let γ1, ..., γ2p be smooth curves that generate H1(Σp,R), the

first homology group of Σp. For a given ε > 0, define

R := Σp\(γ1 ∪ ... ∪ γ2p)ε (3.1.4)

We see that R satisfies Conditions 1 and 2, therefore it has property (?). See

Figure 3.3.

An alternative proof of this statement could be given by noting that a harmonic

map defined on a compact manifold and taking values in a polygon with 4p sides

in the Poincare hyperbolic space such that its image is contained in the interior

of this polygon must be constant. The example above illustrates the relation

between properties of harmonic maps and maximum principles.

Remark 3.1.12 (Property (?) for graphs). To use the previous results to obtain

Bernstein-type theorems we have to overcome the compactness of M . We will

51

3 Proof of main results

φ

harmonic

M compact

∂M = ∅

Figure 3.3: Example of barriers for the existence of harmonic maps in immersed

surfaces in R3 of genus one, two and three.

be able to do so for the graph case. More precisely, we will impose the following

conditions on M : Let M = graph(f) be a minimal submanifold, where f : Rp −→Rn, and γ : M −→ G+

p,n is its Gauss map. Suppose γ(M) ⊂ R, where R is a

region in G+p,n that has property (?). As pointed out by Fischer-Colbrie in [FC80],

one can use Allard’s general work on varifolds to prove that the tangent cone at

infinity of a manifold M given by the graph of a smooth map is the cone over

a compact minimal submanifold M of the sphere [All72]. The Gauss map of

M is again harmonic and satisfies γ(M) ⊂ R. Once more, by an argument of

Fischer-Colbrie [FC80], the compactness of M implies that γ is constant. Hence,

M is the graph of a linear map.

We study the geometry of Grassmannians in detail because we want to explicitly

understand regions in G+p,n that have Property (?) and conditions on the Gauss

map of a manifold that guarantee that its image is contained in such a region.

3.2 Bernstein-type theorems for codimension 2

In this section we construct regions in Grassmannian manifolds that have property

(?) for graphs and use them to prove Bernstein-type theorems. First, we need to

explain the relation of the geometry of Gauss maps with the slope of graphs. To

show the gist of our method, we first explain how Moser’s Bernstein theorem for

52

3.2 Bernstein-type theorems for codimension 2

p

A(p)

p

A(p)

Figure 3.4: (12Eq)ε × (1

2Eq)ε remains as a barrier in S2 × S2.

codimension 1 follows from our method. Then we proceed to derive an analogous

result in codimension 2.

We want to use the geometry of Grassmannians to construct regions in G+p,p+2

satisfying Conditions 1 and 2 of Theorem 3.1.10, to conclude that these regions

have property (?).

Let us start with the simplest case, the manifold (S2 × S2, g × g) which is

isometric to G+2,4 with its standard metric. By Example 3.1.5, we know that

S2\(12Eq)ε has property (?). We define R :=

(S2\(1

2Eq)ε × (S2\(1

2Eq)ε)

)and

look at a given geodesic in S2 × S2:

γ :R −→ S2 × S2

t 7→ (γ1(t), γ2(t)).(3.2.1)

Here γ1, γ2 : R −→ S2 are two geodesics in S2. Therefore R satisfies Conditions 1

and 2 of Theorem 3.1.10 (actually, Theorem 3.1.2 already suffices to see that R is

a barrier) and therefore R has property (?) as well. See Figure 3.4.

By Equation (2.3.31), if Σ2 = (x, f(x)) is an embedded minimal surface in R4

where f : R2 −→ R2 is a smooth mapping with slope ∆f < β0 < +∞, then the

harmonic Gauss map γ : Σ2 −→ G+2,4 maps into the above region, i.e., γ(Σ2) ⊂ R.

Now, by Remark 3.1.12, we have that γ is constant, implying that Σ2 is a plane

in R4.

For the more general case where n = p + 2, we proceed as follows. Write

w ∈ G+p,p+2 ⊂ S(p+2

p )−1 as w = e1 ∧ ... ∧ ep, where eipi=1 is an orthonormal basis

for Vw and n1, n2 an orthonormal basis for V ⊥w . In TwG+p,p+2 there are p(p−1)

2

53

3 Proof of main results

vectors of type 2 (as ∧ anti-commutes), we denote them by X i2, i = 1, ..., p(p−1)

2.

For a given ε > 0, define

R :=

p(p−1)2⋃i=1

⋃−tXi2

+ε<t<tXi2−ε

BG

(wXi

2(t)). (3.2.2)

Since G+p,p+2 is a homogeneous space of non-negative curvature and the geodesics

are given by Equation (2.3.15), we have that R satisfies Condition 2 in section

3, that is, for a given w ∈ R, up to time t < tXj , j = 1, 2, there is always a

direction where the distance to w increases. On the other hand, as we take the

union over −tX2 + ε < t < tX2 − ε (i.e. strictly smaller than the critical time),

we are excluding the points where 〈w,wXi2(t)〉 = cos2(t/

√2) = 0. Therefore

〈w,wXi2(t)〉 > 0 for all i = 1, ..., p(p− 1)/2 and there are no closed geodesics in R

centered at any point w ∈ R. Once more, as G+p,p+2 ⊂ S(p+2

p )−1 is a symmetric

space, w−Xi2(t) = wXi

2(−t), and we have that the only points in wXi

2(R) we are

excluding are the ones that are ε-close to the equator with respect to w in S(p+2p )−1,

thus R =(G+p,p+2\Hem(−w)

has property (?) and hence, by Remark 3.1.12,

property (?) for graphs, for every ε > 0.

Once more, by Equation (2.3.31), if Mp = (x, f(x)) is an embedded minimal

submanifold in Rp+2 and f is a smooth mapping with slope ∆f < β0 < +∞,

then the harmonic Gauss map γ : Mp −→ G+p,p+2 satisfies γ(Mp) ⊂ R. Thus γ is

constant and Mp is a plane in Rp+2.

Remark 3.2.1. If n ≥ p + 3, the definition of such a region R would have to

include not only the union over directions of type 3 (and more), but unions over

type 2 vectors in directions of type 1, and so on. For example, once we transport

a type 1 vector along a geodesic with tangent vector of type 3, we may obtain a

type 2 vector and hence we have to verify whether this would give us a closed

geodesic in R. It can be proven that in general a closed geodesic exists inside R if

we take the union over all directions in higher codimensions. Therefore a more

refined definition of such a region in higher codimension is necessary, as follows

from the existence of the Lawson-Osserman cone.

Returning to codimension 2, the above discussion yields the following theorem,

which extends Moser’s result from codimension 1 to codimension 2.

Theorem 3.2.2. Let zi = f i(x1, ..., xp), i = 1, 2, be smooth functions defined

everywhere in Rp. Suppose their graph M = (x, f(x)) is a submanifold with

54

3.2 Bernstein-type theorems for codimension 2

parallel mean curvature in Rp+2. Suppose that there exists a number β0 < +∞such that

∆f (x) ≤ β0 for all x ∈ Rn, (3.2.3)

where

∆f (x) :=

det

(δαβ +

∑i

f ixα(x)f ixβ(x)

) 12

. (3.2.4)

Then f 1, f 2 are linear functions on Rp representing an affine p-plane in Rp+2.

55

4 Open problems

This chapter presents some problems that arise from this dissertation and that

the author intends to work on in the future.

4.1 Convex supporting domains

In Chapter 2, we defined the concept of a convex supporting domain. Namely,

let (N, h) be a Riemannian manifold and V ⊂ N . We say that V is a convex

supporting domain if there exists a strictly convex function f : V −→ R. The

maximum principle 2.1.6 implies the following:

For V ⊂ N, there exists a striclty convex function f : V −→ R⇓ ⇑?

There is no compact M and harm. non-const. φ : M → N such that φ(M) ⊂ V

⇓ ⇑?

There is no n ≥ 1 and harm. non-const. φ : Sn → N such that φ(Sn) ⊂ V

⇓ ⇑?

There is no harm. non-const. φ : S1 → N such that φ(M) ⊂ V

Emery conjectured that if a subset V ⊂ N admits no closed geodesics, that is,

if every harmonic map φ : S1 → N such that φ(S1) ⊂ V is a constant map,

then there exists a strictly convex function f : V → R [Ken90], [CK99]. This

conjecture is refuted by the example we have given in Chapter 1 regarding the

existence of a non-constant harmonic map into S2\(Γ1 ∪ Γ2 ∪ Γ3). What about

all the other reverse implications?

We conjecture that the first reverse implication holds true. Namely, suppose

V ⊂ has property (?), then there exists a strictly convex function f : V → N .

A complete answer to all the implications on this diagram would be a powerful

result. On one hand, it is a problem interesting on its own. On the other hand,

since in this thesis we have given abstract ways of showing that a region has

property (?), this would give us a conceptual way of verifying whether a region

does admit a strictly convex function.

57

4 Open problems

4.2 Bernstein theorems for codimension c ≥ 3

Our methods to obtain regions without images of harmonic maps can be applied

in general complete manifolds, therefore in Grassmannian manifolds of arbitrary

dimensions. In this case it is important to understand the behavior of closed

geodesics in such manifolds. Of course, as we explained in detail, if the codimension

is one, the Grassmannian is just a sphere and closed geodesics are great circles. If

the codimension is two, there are closed geodesics with more interesting geometry.

For instance, the Grassmannian

G+p,p+2 ⊂ S(p+2

p )−1

contains closed geodesics in the north hemisphere of S(p+2p )−1.

The geodesic equations obtained in [Koz97] for Grassmannians, which we used

in [AJ18] to obtain Moser’s Bernstein theorem in codimension 2, can also be used

to study the case of c ≥ 3. The natural questions are:

• Can we obtain bounds closer to the counterexamples of Lawson and Os-

sermann [LO77], by a systematical study of geodesics in Grassmannians of

arbitrary dimensions?

• Based on [JXY16], it seems that Lawson-Osserman cones are not a sharp

counterexample to Bernstein’s theorem in higher codimension. Can we use

similar methods to the ones in [AJ18] to obtain ‘sharper cones’?

Let us give a very brief description of why the problem becomes more difficult in

codimension 3.

To deal with the closed geodesics that appear in the hemisphere of a given

point, we counted all the possible directions that could occur and since the type

2 vectors were the worse case, of which there are just(p+2

2

), we could handle the

problem. Now, we have an extra codimension. Therefore we must control what

happens to the closed geodesics that have as tangent vector the parallel transport

of a type 2 vector along some other curve (geodesic) that has a type 3 vector as

the tangent direction. On the other hand, the same can happen if we parallel

transport a type 3 vector in a type 1 direction and suddenly obtain a problematic

type 2 vector. This could give us closed geodesics inside our region R.

This is a paper in preparation. We are computing examples similar to Exam-

ples 2.3.19 and 2.3.20, but now also carrying out the parallel translation along

58

4.2 Bernstein theorems for codimension c ≥ 3

other type k vectors to obtain the bounds of the desired region and hence the

bounds for a Bernstein-type theorem.

We have geometrical evidences suggesting that Lawson-Osserman cones are

not the optimal counterexamples to Bernstein theorems in higher codimensions.

Hopefully, further work will be done in this direction.

4.2.1 Non-smooth setting

So far, we have only dealt with smooth manifolds. As pointed out in the previous

section, the main tool used in [AJ18] was the Sampson maximum principle.

Thanks to White [Whi09], we know that there exists an equivalent of this theorem

in the case of minimal varifolds.

A powerful method to study minimal submanifolds is to consider their cone at

infinity, which happens to be a varifold.

• Can we develop similar techniques to the ones in [AJ18] to understand the

existence of stable varifolds on a given manifold?

• How can this result lead to more general Bernstein-type problems?

59

Bibliography

[AJ18] Renan Assimos and Jurgen Jost. The geometry of maximum principles

and a Bernstein theorem in codimension 2. 2018. Submitted. Preprint

available on arXiv:1811.09869.

[All72] William K. Allard. On the first variation of a varifold. Annals of

Mathematics, 95(3):417–491, 1972.

[BDGG69] Enrico Bombieri, Ennio De Giorgi, and Enrico Giusti. Minimal cones

and the Bernstein problem. Inventiones mathematicae, 7(3):243–268,

1969.

[Ber15] Sergei Natanovich Bernstein. Sur un theoreme de geometrie et son

application aux equations aux derivees partielles du type elliptique.

Communications de la Societe mathematique de Kharkow, 15(1):38–45,

1915.

[Che65] Shiing-shen Chern. On the curvatures of a piece of hypersurface in

Euclidean space. In Abhandlungen aus dem Mathematischen Seminar

der Universitat Hamburg, volume 29, pages 77–91. Springer, 1965.

[CK99] Jose Manuel Corcuera and Wilfrid S Kendall. Riemannian barycentres

and geodesic convexity. In Mathematical Proceedings of the Cam-

bridge Philosophical Society, volume 127, pages 253–269. Cambridge

University Press, 1999.

[CX92] Q Chen and YL Xin. A generalized maximum principle and its applica-

tions in geometry. American Journal of Mathematics, 114(2):355–366,

1992.

[DdPW17] Juan Davila, Manuel del Pino, and Juncheng Wei. Singularity for-

mation for the two-dimensional harmonic map flow into s2. arXiv

preprint arXiv:1702.05801, 2017.

61

BIBLIOGRAPHY

[DHS10] Ulrich Dierkes, Stefan Hildebrandt, and Friedrich Sauvigny. Minimal

surfaces. In Minimal Surfaces, pages 53–90. Springer, 2010.

[FC80] Doris Fischer-Colbrie. Some rigidity theorems for minimal submani-

folds of the sphere. Acta Mathematica, 145(1):29–46, 1980.

[Hel02] Frederic Helein. Harmonic maps, conservation laws and moving

frames, volume 150. Cambridge University Press, 2002.

[HJW80] Stefan Hildebrandt, Jurgen Jost, and K-O Widman. Harmonic map-

pings and minimal submanifolds. Inventiones mathematicae, 62(2):269–

298, 1980.

[Jos91] Jurgen Jost. Two-dimensional geometric variational problems. Pure

Appl. Math.(NY), 1991.

[Jos08] Jurgen Jost. Riemannian geometry and geometric analysis, volume

42005. Springer, 2008.

[Jos12] Jurgen Jost. Nonpositive curvature: geometric and analytic aspects.

Birkhauser, 2012.

[JX99] Jurgen Jost and Yuan-Long Xin. Bernstein type theorems for higher

codimension. Calculus of variations and Partial differential equations,

9(4):277–296, 1999.

[JXY12] Jurgen Jost, Yuanlong Xin, and Ling Yang. The regularity of harmonic

maps into spheres and applications to Bernstein problems. Journal

of Differential Geometry, 90(1):131–176, 2012.

[JXY16] Jurgen Jost, Yuan Long Xin, and Ling Yang. The Geometry of

Grassmannian manifolds and Bernstein type theorems for higher

codimension. Annali della Scuola Normale Superiore di Pisa. Classe

di scienze, 16(1):1–39, 2016.

[Ken90] Wilfrid S Kendall. Probability, convexity, and harmonic maps with

small image i: uniqueness and fine existence. Proceedings of the

London Mathematical Society, 3(2):371–406, 1990.

[Koz97] Sergey Emel’yanovich Kozlov. Geometry of the real Grassmannian

manifolds. Parts I, II and III. Zapiski Nauchnykh Seminarov POMI,

246:84–129, 1997.

62

BIBLIOGRAPHY

[Lem78] Luc Lemaire. Harmonic nonholomorphic maps from a surface to a

sphere. Proceedings of the American Mathematical Society, 71(2):299–

304, 1978.

[LO77] H Blaine Lawson and Robert Osserman. Non-existence, non-

uniqueness and irregularity of solutions to the minimal surface system.

Acta Mathematica, 139(1):1–17, 1977.

[LW08] Fanghua Lin and Changyou Wang. The analysis of harmonic maps

and their heat flows. World Scientific, 2008.

[Mos61] Jurgen Moser. On Harnack’s theorem for elliptic differential equations.

Communications on Pure and Applied Mathematics, 14(3):577–591,

1961.

[Oss02] Robert Osserman. A survey of minimal surfaces. Courier Corporation,

2002.

[RV70] Ernst A Ruh and Jaak Vilms. The tension field of the Gauss map.

Transactions of the American Mathematical Society, 149(2):569–573,

1970.

[Sim68] James Simons. Minimal varieties in Riemannian manifolds. Annals

of Mathematics, pages 62–105, 1968.

[Str85] Michael Struwe. On the evolution of harmonic mappings of riemannian

surfaces. Commentarii Mathematici Helvetici, 60(1):558–581, 1985.

[SU81] Jonathan Sacks and Karen Uhlenbeck. The existence of minimal

immersions of 2-spheres. Annals of mathematics, pages 1–24, 1981.

[Whi09] Brian White. The maximum principle for minimal varieties of arbitrary

codimension. arXiv preprint arXiv:0906.0189, 2009.

[Xin03] Yuanlong Xin. Minimal submanifolds and related topics, volume 8.

World Scientific, 2003.

[Xin12] Yuanlong Xin. Geometry of harmonic maps, volume 23. Springer

Science & Business Media, 2012.

63

Selbststandigkeitserklarung

Hiermit erklare ich, die vorliegende Dissertation selbstandig und ohne un-

zulassige fremde Hilfe angefertigt zu haben. Ich habe keine anderen als die

angefuhrten Quellen und Hilfsmittel benutzt und samtliche Textstellen, die

wortlich oder sinngemaß aus veroffentlichten oder unveroffentlichten Schriften

entnommen wurden, und alle Angaben, die auf mundlichen Auskunften beruhen,

als solche kenntlich gemacht. Ebenfalls sind alle von anderen Personen bereit-

gestellten Materialien oder erbrachten Dienstleistungen als solche gekennzeichnet.

Leipzig, den 24.01.2019

. . . . . . . . . . . . . . . . . . . . . . . . . . .

(Renan Assimos Martins)

Daten zum Autor

Name: Renan Assimos Martins

Geburtsdatum: 22.05.1989 in Rio de Janeiro (Brasilien)

03/2010 - 08/2013 Bachelorstudium in Mathematik

Universidade Federal do Rio de Janeiro (Rio de Janeiro, Brasilien)

09/2011 - 05/2014 Masterstudium in Mathematik

Universidade Federal do Rio de Janeiro (Rio de Janeiro, Brasilien)

seit 02/2015 Doktorand im Fachgebiet Mathematik

Max-Planck Institut fur Mathematik in den Naturwissenschaften,

Universitat Leipzig