31.12.2014 Views

Constructions of harmonic maps between Hadamard manifolds

Constructions of harmonic maps between Hadamard manifolds

Constructions of harmonic maps between Hadamard manifolds

SHOW MORE
SHOW LESS

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

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

ISSN 1343-9499<br />

TOHOKU<br />

MATHEMATICAL<br />

PUBLICATIONS<br />

Number 22<br />

<strong>Constructions</strong> <strong>of</strong> <strong>harmonic</strong> <strong>maps</strong><br />

<strong>between</strong> <strong>Hadamard</strong> <strong>manifolds</strong><br />

by<br />

Keisuke Ueno<br />

November 2001<br />

c○Tohoku University<br />

Sendai 980-8578, Japan


Editorial Board<br />

Shigetoshi Bando Masanori Ishida Katsuei Kenmotsu<br />

Hideo Kozono Yasuo Morita Tetsuo Nakamura<br />

Seiki Nishikawa Tadao Oda Norio Shimakura<br />

Toshikazu Sunada Izumi Takagi Toy<strong>of</strong>umi Takahashi<br />

Masayoshi Takeda Kazuyuki Tanaka Yoshio Tsutsumi<br />

Eiji Yanagida Takashi Yoshino Akihiko Yukie<br />

This series aims to publish material related to the activities <strong>of</strong> the<br />

Mathematical Institute <strong>of</strong> Tohoku University. This may include:<br />

1. Theses submitted to the Institute by grantees <strong>of</strong> the degree <strong>of</strong> Doctor<br />

<strong>of</strong> Science.<br />

2. Proceedings <strong>of</strong> symposia as well as lecture notes <strong>of</strong> the Institute.<br />

A primary advantage <strong>of</strong> the series lies in quick and timely publication.<br />

Consequently, some <strong>of</strong> the material published here may very likely<br />

appear elsewhere in final form.<br />

Tohoku Mathematical Publications<br />

Mathematical Institute<br />

Tohoku University<br />

Sendai 980-8578, Japan


<strong>Constructions</strong> <strong>of</strong> <strong>harmonic</strong> <strong>maps</strong><br />

<strong>between</strong> <strong>Hadamard</strong> <strong>manifolds</strong><br />

A thesis presented<br />

by<br />

Keisuke Ueno<br />

to<br />

The Mathematical Institute<br />

for the degree <strong>of</strong><br />

Doctor <strong>of</strong> Science<br />

Tohoku University<br />

Sendai, Japan<br />

September 2001


1<br />

Acknowledgments<br />

The author wishes to express his sincere gratitude to Pr<strong>of</strong>essor Seiki Nishikawa for his<br />

valuable discussions and encouragement throughout the preparation <strong>of</strong> this thesis. He also<br />

would like to express his deep appreciation to Pr<strong>of</strong>essor Hajime Urakawa who directed his<br />

interest to the study <strong>of</strong> <strong>harmonic</strong> <strong>maps</strong>, Pr<strong>of</strong>essors Izumi Takagi and Takeyuki Nagasawa for<br />

their invaluable advice on the theory <strong>of</strong> ordinary differential equations, and Pr<strong>of</strong>essor Kazuo<br />

Akutagawa who introduced him to some relevant topics in the field.


2<br />

Abstract<br />

In this thesis, we present several effective methods <strong>of</strong> constructing <strong>harmonic</strong> <strong>maps</strong> <strong>between</strong><br />

<strong>Hadamard</strong> <strong>manifolds</strong>. In particular, in the case <strong>of</strong> real hyperbolic spaces, we reduce<br />

the <strong>harmonic</strong> map equation to an ordinary differential equation defined on the nonnegative<br />

real line, and by showing the existence <strong>of</strong> solutions for a boundary value problem <strong>of</strong> this<br />

equation, we construct a variety <strong>of</strong> new <strong>harmonic</strong> <strong>maps</strong> <strong>between</strong> real hyperbolic spaces.<br />

Furthermore, following the idea due to Donnelly, we establish the existence and uniqueness<br />

result for proper <strong>harmonic</strong> <strong>maps</strong> <strong>between</strong> Damek-Ricci spaces, which are a generalization<br />

<strong>of</strong> rank one symmetric spaces <strong>of</strong> noncompact type. A non-existence result for proper<br />

<strong>harmonic</strong> <strong>maps</strong> from complex hyperbolic spaces to real hyperbolic spaces is also proved.


Contents<br />

3<br />

Introduction 4<br />

1. Ordinary differential equations associated to equivariant <strong>harmonic</strong> <strong>maps</strong> 13<br />

1.1. Local existence <strong>of</strong> solutions to ordinary differential equations 15<br />

1.2. Global properties <strong>of</strong> solutions 30<br />

1.3. Addendum 41<br />

2. Equivariant <strong>harmonic</strong> <strong>maps</strong> 43<br />

2.1. Eigen<strong>maps</strong> 43<br />

2.2. Reduction <strong>of</strong> <strong>harmonic</strong> map equations 50<br />

2.3. Construction <strong>of</strong> equivariant <strong>harmonic</strong> <strong>maps</strong> 53<br />

3. Dirichlet problem at infinity for <strong>harmonic</strong> <strong>maps</strong> <strong>between</strong> Damek-Ricci 59<br />

spaces<br />

3.1. Damek-Ricci spaces 59<br />

3.2. Harmonic <strong>maps</strong> <strong>between</strong> Damek-Ricci spaces 62<br />

4. Non-existence <strong>of</strong> proper <strong>harmonic</strong> <strong>maps</strong> from complex hyperbolic spaces<br />

into real hyperbolic spaces<br />

73<br />

4.1. Pro<strong>of</strong> <strong>of</strong> Theorem 73<br />

4.2. A counter example 78<br />

Bibliography 83


4<br />

Introduction<br />

Let (M,g) and (M ′ ,g ′ ) be complete Riemannian <strong>manifolds</strong>, and u :(M,g) → (M ′ ,g ′ )<br />

a C 2 -map from M to M ′ . For a relatively compact domain D ⊂ M, we define the energy<br />

E D (u) <strong>of</strong>u over D by<br />

E D (u) := 1 ∫<br />

2<br />

D<br />

|d x u| 2 dv g ,<br />

where |d x u| is the Hilbert-Schmidt norm <strong>of</strong> the differential d x u : T x M → T u(x) M ′ <strong>of</strong> u at<br />

x ∈ M, and dv g denotes the volume measure <strong>of</strong> (M,g). We call u a <strong>harmonic</strong> map if it is a<br />

critical point <strong>of</strong> E D , considered as a functional defined on the space C 2 (M,M ′ )<strong>of</strong>C 2 -<strong>maps</strong><br />

from M to M ′ , for all variations with compact support for any D. In other words, u is a<br />

<strong>harmonic</strong> map if and only if it satisfies the Euler-Lagrange equation for the energy functional<br />

E D for any D, that is,<br />

τ(u)(x) :=<br />

n∑<br />

( ˜∇ ei du(e i ) − du(∇ ei e i ))(x) =0, x ∈ M,<br />

i=1<br />

where {e i } n i=1<br />

is an orthonormal frame field <strong>of</strong> M, ∇ is the Levi-Civita connection on the<br />

tangent bundle TM <strong>of</strong> M, and ˜∇ is the induced connection on the pull-back bundle u −1 TM ′<br />

<strong>of</strong> the tangent bundle TM ′ <strong>of</strong> M ′ by u. Geometrically, τ(u) defines a section, called the<br />

tension field <strong>of</strong> u, <strong>of</strong> u −1 TM ′ . It should be remarked that any C 2 <strong>harmonic</strong> map u is smooth,<br />

since τ(u) = 0 is in fact a system <strong>of</strong> semilinear elliptic partial differential equations <strong>of</strong> second<br />

order.<br />

In the case where (M,g) and (M ′ ,g ′ ) are compact and without boundaries, a remarkable<br />

existence result <strong>of</strong> <strong>harmonic</strong> <strong>maps</strong> was established in 1964 by Eells-Sampson [17].<br />

They<br />

proved that if the target manifold (M ′ ,g ′ ) has nonpositive sectional curvature everywhere,<br />

then there exists an energy minimizing <strong>harmonic</strong> map in each homotopy class <strong>of</strong> smooth<br />

<strong>maps</strong> from (M,g) to(M ′ ,g ′ ).<br />

On the other hand, when (M,g) and (M ′ ,g ′ ) are noncompact, complete <strong>manifolds</strong>, not<br />

much has been known for the existence <strong>of</strong> <strong>harmonic</strong> <strong>maps</strong> <strong>between</strong> them. However, in the


5<br />

last decade, substantial progress has been made on the existence, the uniqueness and other<br />

fundamental properties <strong>of</strong> <strong>harmonic</strong> <strong>maps</strong> <strong>between</strong> <strong>Hadamard</strong> <strong>manifolds</strong>, that is, simply<br />

connected, connected, complete Riemannian <strong>manifolds</strong> <strong>of</strong> nonpositive sectional curvature,<br />

which are most typical examples <strong>of</strong> noncompact, complete Riemannian <strong>manifolds</strong>.<br />

The primary object <strong>of</strong> the present thesis is to present several effective methods <strong>of</strong> constructing<br />

<strong>harmonic</strong> <strong>maps</strong> <strong>between</strong> these <strong>Hadamard</strong> <strong>manifolds</strong>.<br />

Now we shall give a brief outline <strong>of</strong> the problems investigated in this thesis and summarize<br />

our results.<br />

1) Since the existence theorem due to Eells and Sampson requires the nonpositivity <strong>of</strong><br />

sectional curvatures on the target manifold, we can not apply it to the existence <strong>of</strong> <strong>harmonic</strong><br />

<strong>maps</strong> <strong>between</strong> standard unit spheres, although they are among the simplest examples <strong>of</strong><br />

Riemannian <strong>manifolds</strong>. However, by modifying the <strong>harmonic</strong> map equation, we see that<br />

several methods are available for constructing <strong>harmonic</strong> <strong>maps</strong> <strong>between</strong> standard unit spheres.<br />

It follows from a well-known theorem <strong>of</strong> Takahashi that the <strong>harmonic</strong> map equation for a<br />

map ϕ :(S m ,g S m) → (S n ,g S n) <strong>between</strong> standard unit spheres is equivalent to the equation<br />

∆ gS m Φ=2e(ϕ)Φ,<br />

where ∆ gS m denotes the Laplace-Beltrami operator <strong>of</strong> (S m ,g S m),e(ϕ) =|dϕ| 2 /2 is the energy<br />

density function <strong>of</strong> ϕ, and Φ = i ◦ ϕ for the inclusion map i : S n → R n+1 . Therefore, when<br />

e(ϕ) is constant, ϕ is a <strong>harmonic</strong> map if each component <strong>of</strong> ϕ is an eigenfunction <strong>of</strong> ∆ gS m .<br />

Such <strong>harmonic</strong> <strong>maps</strong> ϕ are called eigen<strong>maps</strong>, for which a typical example is provided with<br />

the Hopf fibration ϕ : S 3 → S 2 . Not only these are important objects in its own right, but<br />

also, using eigen<strong>maps</strong>, we can obtain various <strong>harmonic</strong> <strong>maps</strong> <strong>between</strong> standard unit spheres,<br />

and <strong>between</strong> real hyperbolic spaces as well. In this regard, in Chapter 2, we shall present a<br />

new method <strong>of</strong> constructing a wealth <strong>of</strong> eigen<strong>maps</strong>.<br />

In order to construct a <strong>harmonic</strong> map <strong>between</strong> standard unit spheres, the notion <strong>of</strong> a<br />

join <strong>of</strong> two eigen<strong>maps</strong> was introduced by Smith ([34]). More precisely, let u : S m−1 → S p−1<br />

and v : S n−1 → S q−1 be eigen<strong>maps</strong>, and r :[0,π/2] → [0,π/2] a smooth function. Noting


6<br />

that for any point z ∈ S m+n−1 , there exist x ∈ S m−1 ,y ∈ S n−1 and t ∈ [0,π/2] such that z<br />

is expressed as<br />

z = ((cos t)x, (sin t)y),<br />

the join u ∗ r v <strong>of</strong> these <strong>maps</strong> is defined by<br />

u ∗ r v : S m+n−1 ∋ ((cos t)x, (sin t)y) ↦→ ((cos r(t))u(x), (sin r(t))v(y)) ∈ S p+q−1 .<br />

Consequently, the <strong>harmonic</strong> map equation for the join map is reduced to an ordinary differential<br />

equation in r = r(t) with a suitable boundary condition. By solving this boundary<br />

value problem, Smith proved that there exists a <strong>harmonic</strong> representative in each homotopy<br />

class π n (S n ) ≃ Z for n ≤ 7. Subsequently, his method was extended by several authors.<br />

In particular, Ding([11]) utilized a variational method to clarify the meaning <strong>of</strong> the dumping<br />

condition, Eells and Ratto([15]) generalized the method to the case <strong>of</strong> ellipsoids, and<br />

Xin([45],[46]) reformulated Smith’s method from the view point <strong>of</strong> Riemannian submersions<br />

(see also [16]).<br />

2) We can generalize the notion <strong>of</strong> a join map to the case <strong>of</strong> real hyperbolic spaces RH m<br />

as follows. First, note that for any point z ∈ RH m+n−1 , there exist x ∈ RH m−1 ,y ∈ S n−1<br />

and t ∈ [0, ∞) such that<br />

z = ((cosh t)x, (sinh t)y).<br />

Let u : RH m−1 → RH p−1 and v : S n−1 → S q−1 be eigen<strong>maps</strong>, and r :[0, ∞) → [0, ∞) a<br />

smooth function. Then the join map u ∗ r v <strong>of</strong> u, v and r is defined to be<br />

u ∗ r v : RH m+n−1 ∋ ((cosh t)x, (sinh t)y) ↦→ ((cosh r(t))u(x), (sinh r(t))v(y)) ∈ RH p+q−1 .<br />

Then we see that u ∗ r v is a <strong>harmonic</strong> map if and only if the function r = r(t) satisfies the<br />

following ordinary differential equation defined on [0, ∞):<br />

{<br />

¨r(t)+ p sinh t<br />

cosh t + m cosh t } { }<br />

µ<br />

2<br />

ṙ(t) −<br />

sinh t cosh 2 t +<br />

ν2<br />

sinh 2 sinh r(t) cosh r(t) =0,<br />

t<br />

where e(u) =µ 2 /2 and e(v) =ν 2 /2 denote the energy density functions <strong>of</strong> u and v, respectively.


On the other hand, the (m + 1)-dimensional real hyperbolic space RH m+1 is represented<br />

as the warped product manifold<br />

([0, ∞) × S m ,dt 2 + (sinh 2 t)g S m).<br />

Thus, for an eigenmap ϕ : S m → S n , we may consider an equivariant map<br />

u : ([0, ∞) × S m ,dt 2 + (sinh 2 t)g S m) ∋ (t, θ)<br />

↦→ (r(t),ϕ(θ)) ∈ ([0, ∞) × S n ,dr 2 + (sinh 2 r)g S n).<br />

Then the <strong>harmonic</strong>ity <strong>of</strong> u is also reduced to the following ordinary differential equation in<br />

r = r(t) defined on [0, ∞):<br />

¨r(t)+m cosh t sinh r(t) cosh r(t)<br />

ṙ(t) − µ2<br />

sinh t sinh 2 t<br />

where e(ϕ) =µ 2 /2 denotes the energy density function <strong>of</strong> ϕ.<br />

In Chapter 1, we shall investigate the following ordinary differential equation defined on<br />

[0, ∞), which stems from the study <strong>of</strong> such <strong>harmonic</strong> <strong>maps</strong>:<br />

{<br />

f<br />

¨r(t)+ p ˙ 1 (t)<br />

f 1 (t) + q f ˙ } {<br />

2 (t)<br />

ṙ(t) − µ 2 h 1(r(t))h ′ 1(r(t))<br />

+ ν 2 h }<br />

2(r(t))h ′ 2(r(t))<br />

=0,<br />

f 2 (t)<br />

f 1 (t) 2<br />

f 2 (t) 2<br />

where p, q, µ and ν are some constants.<br />

We first prove the short time existence <strong>of</strong> solutions following the method due to Baird<br />

([3]). However, it seems to the present author that his pro<strong>of</strong> needs some refinements, in<br />

particular, on the life span <strong>of</strong> solutions. In fact, under appropriate conditions for the given<br />

functions f i (t) and h i (r), we can elaborate his method and obtain a complete pro<strong>of</strong> <strong>of</strong> the<br />

short time existence. Moreover, if we pose a condition on the growth rate <strong>of</strong> the function<br />

f i (t) ast goes to infinity, then such a short time solution can be extended to a global<br />

solution, which ensures the existence <strong>of</strong> an equivariant <strong>harmonic</strong> map as well as a <strong>harmonic</strong><br />

map defined by the join. Applying these results, we can prove that there exists a family<br />

<strong>of</strong> <strong>harmonic</strong> <strong>maps</strong> from the real hyperbolic space RH m onto itself for m ≥ 3, which is<br />

parameterized by Z.<br />

=0,<br />

7


8<br />

Equivariant <strong>harmonic</strong> <strong>maps</strong> have been studied by many authors. For example, Kasue<br />

and Washio ([22]) investigated equivariant <strong>harmonic</strong> <strong>maps</strong> <strong>of</strong> the form<br />

u : ([0, ∞) × S m ,dt 2 + f(t) 2 g S m) ∋ (t, θ)<br />

↦→ (r(t),ϕ(θ)) ∈ ([0, ∞) × S n ,dr 2 + h(r) 2 g S n).<br />

Here ϕ : S m → S n is an eigenmap, and f = f(t) and h = h(r) are warping functions. Since<br />

they constructed a comparison function to show the global existence <strong>of</strong> a solution for the<br />

original equation, the growth order <strong>of</strong> f and h are rather restricted. Thus one can not apply<br />

their argument to the case <strong>of</strong> real hyperbolic spaces.<br />

3) For <strong>Hadamard</strong> <strong>manifolds</strong> (M,g) and (M ′ ,g ′ ), one can consider the Eberlein-O’Neill<br />

compactifications M and M ′ <strong>of</strong> M and M ′ , respectively, by adding their ideal boundaries,<br />

which are defined to be the spheres at infinity given by the asymptotic classes <strong>of</strong> geodesic<br />

rays. Then one can set up the Dirichlet problem at infinity for <strong>harmonic</strong> <strong>maps</strong>, which,<br />

for a given boundary map f : ∂M → ∂M ′ , consists <strong>of</strong> the existence <strong>of</strong> a <strong>harmonic</strong> map<br />

u : M → M ′ which assumes the boundary value f continuously.<br />

This problem can be regarded as a generalization <strong>of</strong> Hamilton’s work [19] on the Dirichlet<br />

problem for <strong>harmonic</strong> <strong>maps</strong> <strong>between</strong> compact Riemannian <strong>manifolds</strong> with boundary to<br />

these noncompact Riemannian <strong>manifolds</strong>. However, since the Riemannian metrics under<br />

consideration blow up at the ideal boundaries, there appear much difficulties in analyzing<br />

the boundary behavior <strong>of</strong> solutions <strong>of</strong> the <strong>harmonic</strong> map equation.<br />

The first progress to the above problem was accomplished around 1990’s by Li-Tam [25]<br />

[26] [27] and Akutagawa [1]. Recall that most typical examples <strong>of</strong> <strong>Hadamard</strong> <strong>manifolds</strong><br />

are the rank one symmetric spaces <strong>of</strong> noncompact type, that is, the real, the complex and<br />

the quaternion hyperbolic spaces and the Cayley hyperbolic plane. In their work, Li and<br />

Tam [25] [26] [27] solved the Dirichlet problem at infinity for <strong>harmonic</strong> <strong>maps</strong> <strong>between</strong> real<br />

hyperbolic spaces. In particular, exploiting the heat equation method, they established a<br />

general theory for the existence and uniqueness <strong>of</strong> solutions to this problem. For example,


9<br />

they proved the following: Suppose that <strong>Hadamard</strong> <strong>manifolds</strong> under consideration satisfy<br />

appropriate geometric conditions on the curvatures, the volume <strong>of</strong> unit balls and the bottom<br />

spectrum <strong>of</strong> the Laplace-Beltrami operator. Then, in order to obtain a global solution to<br />

the equation which converges to the desired <strong>harmonic</strong> map, it suffices to construct a suitable<br />

initial map for the parabolic <strong>harmonic</strong> map equation. On the other hand, by applying the<br />

elliptic method, Akutagawa [1] proved the existence <strong>of</strong> solutions to this problem in the case<br />

<strong>of</strong> the real hyperbolic plane.<br />

In 1994, Donnelly [12] discussed in a general setting the problem for <strong>harmonic</strong> <strong>maps</strong><br />

<strong>between</strong> rank one symmetric spaces <strong>of</strong> noncompact type. He investigated the unbounded<br />

models <strong>of</strong> these spaces realized as upper half space models, and made use <strong>of</strong> their global coordinates<br />

near the ideal boundaries. With these models and coordinates, he constructed good<br />

initial <strong>maps</strong> which enable one to apply Li-Tam’s existence result, and solved the problem for<br />

boundary <strong>maps</strong> having sufficient regularity. A significant step involved is, by investigating<br />

the boundary behavior <strong>of</strong> proper <strong>harmonic</strong> <strong>maps</strong>, to deduce necessary conditions for the<br />

existence <strong>of</strong> solutions, which are expressed in terms <strong>of</strong> the relation <strong>between</strong> geometric structures<br />

around the ideal boundaries. Indeed, a given boundary map is related to the geometric<br />

structure on the boundary, for instance, to the conformal structure in the real case, and to<br />

the contact structure in the complex case.<br />

Since all <strong>manifolds</strong> investigated so far are symmetric spaces and have strictly negative<br />

sectional curvature everywhere, the following question arises naturally: For non-symmetric or<br />

nonpositively curved <strong>manifolds</strong>, can one solve the Dirichlet problem at infinity for <strong>harmonic</strong><br />

<strong>maps</strong> As an answer to this question, in Chapter 3, we shall solve the Dirichlet problem for<br />

<strong>harmonic</strong> <strong>maps</strong> <strong>between</strong> Damek-Ricci spaces.<br />

Damek-Ricci spaces were first introduced by Damek [7] as a semidirect extension <strong>of</strong> the<br />

generalized Heisenberg groups discovered by Kaplan [21]. These spaces may be regarded as<br />

a generalization <strong>of</strong> rank one symmetric spaces <strong>of</strong> noncompact type, since their geometries<br />

are quite similar to each other. However, Damek-Ricci spaces are not symmetric in general<br />

and have nonpositive sectional curvature. In fact, non-symmetric Damek-Ricci spaces admit


10<br />

vanishing sectional curvatures for certain 2-planes (see Theorem 3.1.2). We shall prove that<br />

Donnelly’s existence and uniqueness results can be extended to the case <strong>of</strong> Damek-Ricci<br />

spaces. Thus the geometric conditions such as symmetry and strict negativity <strong>of</strong> sectional<br />

curvatures are proved to be not essential, and it is worthwhile studying the problem for<br />

<strong>Hadamard</strong> <strong>manifolds</strong> in further generality.<br />

In the last section, we shall prove a non-existence result for proper <strong>harmonic</strong> <strong>maps</strong> from<br />

complex hyperbolic spaces to real hyperbolic spaces. To be precise, let B m and D n be<br />

the ball models <strong>of</strong> the m-dimensional complex hyperbolic space and the n-dimensional real<br />

hyperbolic space, respectively. Then we prove that there exists no proper <strong>harmonic</strong> map<br />

u ∈ C 2 (B m , D n ) which has C 1 -regularity up to the ideal boundary, where m, n ≥ 2. Recently,<br />

Li and Ni [24] obtained the same result in the case <strong>of</strong> n =2. Furthermore, we show that<br />

there exists a counter example to this result if we relax the regularity condition up to the<br />

ideal boundary.<br />

It should be remarked that Donnelly [12] proved that, for a suitable boundary map, there<br />

exists a solution to the Dirichlet problem for <strong>harmonic</strong> <strong>maps</strong> from complex hyperbolic spaces<br />

to real hyperbolic spaces, which has sufficiently high regularity up to the ideal boundary.<br />

Hence our theorem appears to contradict his result. However, this is caused by the different<br />

choice <strong>of</strong> the models <strong>of</strong> complex hyperbolic spaces. For the upper half space model, Donnelly<br />

used the one defined by<br />

M 1 =<br />

(R + × R 2m−1 ,h 1 = dy2<br />

y + 1 2 y g 2 1 + 1 )<br />

y g 4 2 ,<br />

where y ∈ R + and g 1 + g 2 is a Riemannian metric on R 2m−1 . On the other hand, our model<br />

is given by<br />

M 2 =<br />

(R + × R 2m−1 ,h 2 = dη2<br />

4η + 1 2 η g 1 + 1 )<br />

η g 2 2 ,<br />

where η ∈ R + . If one define a map f : M 2 → M 1 by<br />

f(η, x i )=( √ η, x i ) for (η, x i ) ∈ R + × R 2m−1 ,


11<br />

then it is easily verified that f ∗ h 1 = h 2 , that is, f : M 2 → M 1 is an isometry, in particular,<br />

a smooth map from M 2 to M 1 . However, a function y = √ η is not smooth at η =0, which<br />

implies that the differentiable structures around the ideal boundaries <strong>of</strong> M 1 and M 2 are<br />

different each other. Thus, even if u : B m → D n is smooth up to the ideal boundary with<br />

respect to the differentiable structure for Donnelly’s model, it may not be the case with<br />

respect to ours.<br />

When a point p ∈ R + × R 2m−1 tends to the ideal boundary, each component <strong>of</strong> the<br />

metric h 1 blows up at least with the order <strong>of</strong> y 2 , and likewise there exists a component <strong>of</strong> the<br />

metric h 2 which blows up with the order <strong>of</strong> η. Thus, we conjecture that, for other general<br />

<strong>Hadamard</strong> <strong>manifolds</strong>, the existence or non-existence <strong>of</strong> proper <strong>harmonic</strong> <strong>maps</strong> is also closely<br />

related to the growth rate <strong>of</strong> Riemannian metrics and the growth order <strong>of</strong> a proper <strong>harmonic</strong><br />

map near ideal boundaries.


CHAPTER 1<br />

Ordinary differential equations associated to<br />

equivariant <strong>harmonic</strong> <strong>maps</strong><br />

In this chapter, we study the existence <strong>of</strong> a positive solution r = r(t) to the following equation<br />

(1.1.1) with the boundary condition (1.1.2):<br />

(1.1.1)<br />

{<br />

f<br />

¨r(t)+ p ˙ 1 (t)<br />

f 1 (t) + q f ˙ }<br />

2 (t)<br />

ṙ(t)<br />

f 2 (t)<br />

{<br />

− µ 2 h 1(r(t))h ′ 1(r(t))<br />

+ ν 2 h }<br />

2(r(t))h ′ 2(r(t))<br />

= 0 on [0, ∞),<br />

f 1 (t) 2<br />

f 2 (t) 2<br />

(1.1.2) lim<br />

t→0<br />

r(t) =0,<br />

where ṙ(t) (resp. h ′ (r)) means (dr/dt)(t) (resp. (dh/ds)(r)), µ and ν are nonnegative numbers,<br />

and p and q are integers. Here we assume that<br />

µ 2 >ν 2 ≥ 0 and p>q≥ 0.<br />

The equation (1.1.1) stems from the study <strong>of</strong> equivariant <strong>harmonic</strong> <strong>maps</strong> <strong>between</strong> complete<br />

noncompact Riemannian <strong>manifolds</strong>, for instance, the real (resp. complex) Euclidean space<br />

(R m ,g R m) (resp. (C m ,g C m)), the real hyperbolic space (RH m ,g RH m) <strong>of</strong> constant negative<br />

sectional curvature −1 and the complex hyperbolic space (CH m ,g CH m) <strong>of</strong> constant holomorphic<br />

sectional curvature −1.<br />

Throughout this chapter, we assume that C ∞ -functions f i = f i (t) (i =1, 2) defined on<br />

13


14 CHAPTER 1. ORDINARY DIFFERENTIAL EQUATIONS<br />

[0, ∞) satisfy the following conditions:<br />

⎧<br />

(F-1)<br />

(F-2)<br />

⎪⎨<br />

(F-3)<br />

⎪⎩ (F-4)<br />

f i (t) > 0on(0, ∞), and ˙ f i (t) ≥ 0on[0, ∞),<br />

lim<br />

t→0<br />

t<br />

f i (t) = a i, where a 1 > 0 and a 2 ≥ 0,<br />

for any t ≥ 0 the following holds:<br />

0 ≤ pt<br />

f˙<br />

1 (t)<br />

f 1 (t) + qt ˙<br />

∫ ∞<br />

{ 1<br />

f 1 (τ) + 1<br />

f 2 (τ)<br />

f 2 (t)<br />

− 1, and<br />

f 2 (t)<br />

}<br />

dτ < ∞.<br />

Moreover, we assume the following conditions for C ∞ -functions h i = h i (s) (i =1, 2) defined<br />

on R:<br />

⎧<br />

(H-1) h i (s) > 0on(0, ∞),<br />

⎪⎨<br />

(H-2) h i (0) = 0, h ′ i(0) > 0, and<br />

⎪⎩ (H-3)<br />

{h i h i ′ } ′ (s) ≥ 0onR.<br />

Our main purpose <strong>of</strong> this chapter is to prove the following theorems.<br />

∫ ∞<br />

ds<br />

Theorem A. If < ∞, then there exists a global, strictly monotone increasing<br />

h(s)<br />

solution to the equation (1.1.1) with the boundary condition (1.1.2). Moreover, at least one<br />

solution is unbounded.<br />

∫ ∞<br />

ds<br />

Theorem B. If = ∞, then every solution to the equation (1.1.1) with the boundary<br />

h(s)<br />

condition (1.1.2) is a global, strictly monotone increasing, bounded solution.<br />

∫ ∞<br />

ds<br />

As a consequence, when = ∞, the limit value lim r(t) <strong>of</strong> a solution r = r(t)<br />

h(s) t→∞<br />

exists. Conversely, for any l ≥ 0, we can construct a global, strictly monotone increasing<br />

solution r = r(t) which converges to l as t →∞(see Theorem 1.2.6).


1.1. LOCAL EXISTENCE OF SOLUTIONS TO (1.1.1) WITH (1.1.2) 15<br />

1.1 Local existence <strong>of</strong> solutions to (1.1.1) with (1.1.2)<br />

In this section, we consider the following ordinary differential equation which includes the<br />

equation (1.1.1) as a special case:<br />

(1.1.3) ¨r(t)+F (t)ṙ(t) − G(t, r(t)) = 0.<br />

Here we assume that F and G satisfy the following conditions:<br />

⎧<br />

(C-1)<br />

F ∈ C ∞ ((0, ∞)),<br />

(C-2) tF ∈ C([0, ∞)), tF(t) − 1 ≥ 0 on [0, ∞),<br />

⎪⎨ (C-3) G ∈ C ∞ ((0, ∞) × R), G(t, 0) = 0 for t>0,<br />

(C-4) G(t, s) > 0 for (t, s) ∈ (0, ∞) × (0, ∞),<br />

lim<br />

t→0<br />

t 2 G(t, s) > 0 for s>0, lim<br />

t→0,s→0 t2 G(t, s) =0, and<br />

⎪⎩ (C-5) s 1 ≤ s 2 =⇒ G(t, s 1 ) ≤ G(t, s 2 ) for any t ∈ (0, ∞).<br />

Let f i and h i be functions satisfying the conditions (F-1) through (F-4) and (H-1) through<br />

(H-3), respectively. Define F (t) and G(t, s) by<br />

f<br />

F (t) =p ˙ 1 (t)<br />

f 1 (t) + q f ˙ 2 (t)<br />

f 2 (t) ,<br />

G(t, s) =µ 2 h 1(s)h ′ 1 (s)<br />

f 1 (t) 2<br />

+ ν 2 h 2(s)h ′ 2 (s)<br />

f 2 (t) 2 .<br />

Then it is easily verified that these functions, F and G, satisfy the conditions (C-1) through<br />

(C-5).<br />

We are going to prove the short time existence <strong>of</strong> a solution to the equation (1.1.3). Our<br />

goal is to show the following<br />

Theorem 1.1.1. Under the conditions (C-1) through (C-5), for any t 0 > 0 and r 0 > 0, there<br />

exists a unique positive solution r = r(t) to the equation (1.1.3) on [0,t 0 ] which satisfies the


16 CHAPTER 1. ORDINARY DIFFERENTIAL EQUATIONS<br />

following conditions:<br />

(1) lim r(t) =0and r(t 0 )=r 0 .<br />

t→0<br />

(2) ṙ(t) > 0 on (0,t 0 ].<br />

In the remainder <strong>of</strong> this section, we choose t 0 > 0 arbitrarily and fix it.<br />

Lemma 1.1.2. Assume that r = r(t) and ρ = ρ(t) are defined on [0,T] and satisfy the<br />

following conditions on (0,T):<br />

⎧<br />

⎪⎨ ¨r(t)+F (t)ṙ(t) =G(t, r(t)),<br />

⎪⎩ ¨ρ(t)+F (t)˙ρ(t) ≤ G(t, ρ(t)).<br />

If r(T )=ρ(T ) and r(0) = ρ(0), then r ≤ ρ on [0,T].<br />

Pro<strong>of</strong>.<br />

Set w(t) :=ρ(t) − r(t). Then we have<br />

ẅ(t)+F (t)ẇ(t) ≤ G(t, ρ(t)) − G(t, r(t)).<br />

Assume that there exists t 1 ∈ (0,T) such that w(t 1 ) < 0. Then, since w(0) = w(T )=0,<br />

we may suppose ẇ(t 1 ) = 0 and ẅ(t 1 ) > 0. On the other hand, since ρ(t 1 )


1.1. LOCAL EXISTENCE OF SOLUTIONS TO (1.1.1) WITH (1.1.2) 17<br />

Corollary 1.1.3. Let r = r(t) and ρ = ρ(t) be solutions to the equation (1.1.3) with the<br />

boundary condition (1.1.2). Ifr(t 0 )=ρ(t 0 ) holds for some t 0 ∈ (0,T), then we have<br />

r(t) =ρ(t) on [0,T),<br />

where [0,T) is the common life span <strong>of</strong> r(t) and ρ(t). In particular, if r(t 0 )=0, then r ≡ 0.<br />

To prove the existence part <strong>of</strong> Theorem 1.1.1, we employ a method due to Baird ([3,<br />

Chapter 6]).<br />

We replace the variable t in the equation (1.1.3) with τ = log t to remove the singularity<br />

at t =0. Then the equation (1.1.3) becomes<br />

(1.1.4)<br />

where<br />

d 2 r dr<br />

(τ)+P (τ) (τ) − Q(τ,r(τ))=0,<br />

dτ<br />

2<br />

dτ<br />

P (τ) =e τ F (e τ ) − 1,<br />

Note that P and Q satisfy the following conditions:<br />

⎧<br />

P ∈ C ∞ (R), P(τ) ≥ 0 on R,<br />

Q(τ,s)=e 2τ G(e τ ,s).<br />

Q ∈ C ∞ (R × R), Q(τ,s) > 0 for (τ,s) ∈ R × (0, ∞)<br />

⎪⎨<br />

lim Q(τ,s) > 0 for any s>0, lim Q(τ,s)=0,<br />

τ→−∞ τ→−∞,s→0<br />

Q(τ,0) = 0<br />

for any τ ∈ R, and,<br />

We also set τ 0 = log t 0 .<br />

⎪⎩<br />

s 1 ≤ s 2 =⇒ Q(τ,s 1 ) ≤ Q(τ,s 2 ) for any τ ∈ R.<br />

Under these conditions, we prove the following<br />

Theorem 1.1.4. For any τ 0 ∈ R and r 0 > 0, there exists a solution r = r(τ) to the equation<br />

(1.1.4) on (−∞,τ 0 ] satisfying the following conditions:<br />

(1) r(τ) > 0,<br />

(2) lim r(τ) =0, lim<br />

τ→−∞<br />

dr<br />

dτ (τ) > 0 on (−∞,τ 0], and<br />

τ→−∞<br />

dr<br />

(τ) =0and<br />

dτ lim<br />

τ→−∞<br />

d 2 r<br />

(τ) =0.<br />

dτ<br />

2


18 CHAPTER 1. ORDINARY DIFFERENTIAL EQUATIONS<br />

It is easy to see that Corollary 1.1.3 together with Theorem 1.1.4 implies Theorem 1.1.1.<br />

To prove Theorem 1.1.4, we begin with the following lemma concerning the life span <strong>of</strong><br />

r = r(τ).<br />

Lemma 1.1.5. Let r = r(τ) be a solution to the equation (1.1.4).<br />

Then (dr/dτ)(τ) is<br />

bounded if so is r(τ). Namely, if (¯τ,τ 0 ] is the life span <strong>of</strong> r = r(τ) and −∞ < ¯τ, then<br />

lim |r(τ)| = ∞.<br />

τ→¯τ+0<br />

Pro<strong>of</strong>.<br />

Let ˜P (τ) = exp ∫ τ P (ρ)dρ. Then, from the equation (1.1.4), we have<br />

d<br />

dτ<br />

{<br />

˜P (τ) dr<br />

dτ (τ) }<br />

= ˜P (τ)Q(τ,r(τ)).<br />

Integrating both sides from τ(< τ 0 )toτ 0 , we obtain<br />

˜P (τ 0 ) dr<br />

dτ (τ 0) − ˜P (τ) dr<br />

dτ (τ) = ∫ τ0<br />

τ<br />

˜P (ρ)Q(ρ, r(ρ))dρ.<br />

Hence<br />

∣ ∫ ˜P (τ)<br />

dr ∣∣∣<br />

∣dτ (τ) ≤ ˜P (τ 0 )<br />

dr<br />

∣dτ (τ τ0<br />

0)<br />

∣ +<br />

Since Q ∈ C ∞ (R × R), the assertion holds.<br />

τ<br />

˜P (ρ)|Q(ρ, r(ρ))|dρ.<br />

Lemma 1.1.6. Let r 1 = r 1 (τ) and r 2 = r 2 (τ) be solutions to the equation (1.1.4) satisfying<br />

r 1 (τ 0 ) ≤ r 2 (τ 0 )<br />

and<br />

dr 1<br />

dτ (τ 0) > dr 2<br />

dτ (τ 0).<br />

Then it holds that<br />

dr 1<br />

r 1 dr 2<br />

dτ<br />

where (¯τ,τ 0 ) is the common life span <strong>of</strong> r 1 and r 2 .<br />

on (¯τ,τ 0 ),<br />

Pro<strong>of</strong>. Let w(τ) =r 1 (τ)−r 2 (τ). Note that w(τ 0 ) ≤ 0 and (dw/dτ)(τ 0 ) > 0. Assume that<br />

there exists τ 1 ∈ (¯τ,τ 0 ) such that w(τ 1 ) ≥ 0. Then there exists a point τ 2 ∈ (τ 1 ,τ 0 ) so that<br />

w(τ 2 ) < 0,<br />

dw<br />

dτ (τ 2) = 0 and d2 w<br />

dτ 2 (τ 2) > 0.


1.1. LOCAL EXISTENCE OF SOLUTIONS TO (1.1.1) WITH (1.1.2) 19<br />

On the other hand, since r 1 and r 2 are solutions to the equation (1.1.4) and the function<br />

s ↦→ Q(·,s) is monotone, it holds that<br />

d 2 w<br />

dτ (τ 2)=−P (τ 2 2 ) dw<br />

dτ (τ 2)+Q(τ 2 ,r 1 (τ 2 )) − Q(τ 2 ,r 2 (τ 2 ))<br />

= Q(τ 2 ,r 1 (τ 2 )) − Q(τ 2 ,r 2 (τ 2 ))<br />

≤ 0,<br />

which yields a contradiction. Hence w(τ) < 0, that is, r 1 (τ) ¯τ) toτ 0 , we obtain<br />

˜P (τ 0 )(r 1 (τ 0 ) − r 2 (τ 0 )) d<br />

dτ (r 1(τ 0 ) − r 2 (τ 0 ))<br />

≥ ˜P (τ)(r 1 (τ) − r 2 (τ)) d<br />

dτ (r 1(τ) − r 2 (τ)) +<br />

> ˜P (τ)(r 1 (τ) − r 2 (τ)) d<br />

dτ (r 1(τ) − r 2 (τ)),<br />

from which it follows that<br />

∫ τ0<br />

˜P (τ)(r 1 (τ) − r 2 (τ)) d<br />

dτ (r 1(τ) − r 2 (τ))<br />

< ˜P (τ 0 )(r 1 (τ 0 ) − r 2 (τ 0 )) d<br />

dτ (r 1(τ 0 ) − r 2 (τ 0 ))<br />

≤ 0.<br />

τ<br />

}<br />

{ } 2 d<br />

˜P (τ)<br />

dτ (r 1(τ) − r 2 (τ)) dτ<br />

Since r 1 (τ) dr 2<br />

dτ (τ)


20 CHAPTER 1. ORDINARY DIFFERENTIAL EQUATIONS<br />

for τ ∈ (¯τ,τ 0 ).<br />

As a corollary <strong>of</strong> this lemma, we obtain the following<br />

Corollary 1.1.7. There is at most one solution to the two point boundary value problem <strong>of</strong><br />

the equation (1.1.4).<br />

For each r 0 > 0 and s 0 ∈ R we shall solve the ordinary differential equation (1.1.4)<br />

backward with the initial condition<br />

dr<br />

(1.1.5) r(τ 0 )=r 0 and<br />

dτ (τ 0)=s 0 .<br />

Our aim is to show that for any r 0 > 0 there exists an s 0 > 0 such that a solution r = r(τ)<br />

exists on (−∞,τ 0 ] and satisfies the condition (1) <strong>of</strong> Theorem 1.1.4.<br />

Define a set A(r 0 )by<br />

A(r 0 ):=<br />

{<br />

s 0 ∈ R<br />

∣<br />

Then we have the following<br />

}<br />

a solution r(τ) to (1.1.4) satisfying (1.1.5) exists, which decreases<br />

.<br />

monotonically to zero within finite time as τ decreases from τ 0<br />

Proposition 1.1.8. Let r 0 > 0. Then one <strong>of</strong> the following two cases occurs:<br />

(1) The set A(r 0 ) is an empty set. In this case, for any s 0 > 0 there exists a positive solution<br />

r = r(τ) on (−∞,τ 0 ] to the equation (1.1.4) with the initial condition (1.1.5), which satisfies<br />

dr<br />

(τ) > 0.<br />

dτ<br />

(2) The set A(r 0 ) is not empty. In this case, A(r 0 ) is an open set, and inf A(r 0 ) > 0.<br />

Pro<strong>of</strong>.<br />

(1) For any fixed r 0 > 0, we take s 0 > 0 arbitrarily. Let r = r(τ) be a solution<br />

to the equation (1.1.4) with the initial condition (1.1.5), and let (¯τ,τ 0 ] be its life span. Then<br />

we have<br />

dr<br />

dτ (τ) > 0 on (¯τ,τ 0].<br />

Indeed, suppose that there exists τ 1 ∈ (¯τ,τ 0 ) such that<br />

dr<br />

dτ (τ 1) = 0<br />

and<br />

dr<br />

dτ (τ) > 0 on (τ 1,τ 0 ].


1.1. LOCAL EXISTENCE OF SOLUTIONS TO (1.1.1) WITH (1.1.2) 21<br />

Since A(r 0 ) is an empty set, r(τ) > 0 for τ ∈ (τ 1 ,τ 0 ]. Choose ¯s 0 so that<br />

{<br />

¯s 0 > max s 0 ,C 1 (τ 0 − τ)+ r }<br />

0<br />

,<br />

τ 0 − τ 1<br />

where C 1 := max τ∈[τ1 ,τ 0 ] Q(τ,r 0 ). We shall show that ¯s 0 ∈A(r 0 ).<br />

Let ρ = ρ(τ) be a solution to the equation (1.1.4) satisfying<br />

dρ<br />

ρ(τ 0 )=r(τ 0 ) and<br />

dτ (τ 0)=¯s 0 > dr<br />

dτ (τ 0),<br />

and (τ 2 ,τ 0 ] the life span <strong>of</strong> ρ. Then from Lemma 1.1.6 it holds that<br />

as long as both r and ρ exist.<br />

ρ(τ) <br />

dτ dτ (τ)<br />

When τ 1 ≤ τ 2 , we have ρ(τ) →−∞as τ → τ 2 +0. Hence there exists τ 3 ∈ (τ 2 ,τ 0 ] such<br />

that<br />

dρ dr<br />

ρ(τ 3 ) = 0 and (τ) ><br />

dτ dτ (τ) > 0 on (τ 3,τ 0 ].<br />

Therefore ¯s 0 ∈A(r 0 ), contradicting the fact A(r 0 )=∅.<br />

On the other hand, when τ 1 >τ 2 ,ρsatisfies (dρ/dτ)(τ) > (dr/dτ)(τ) > 0on(τ 1 ,τ 0 ]. Let<br />

Then C 2 ≥ 0, and it holds that<br />

C 2 :=<br />

min P (τ).<br />

τ∈[τ 1 ,τ 0 ]<br />

d 2 ρ<br />

(τ) =−P (τ)dρ<br />

dτ<br />

2<br />

dτ (τ)+Q(τ,ρ(τ))<br />

≤−P (τ) dρ<br />

dτ (τ)+Q(τ,ρ(τ 0))<br />

dρ<br />

≤−C 2<br />

dτ (τ)+C 1<br />

on [τ 1 ,τ 0 ]. Integrating this inequality from τ ∈ [τ 1 ,τ 0 ]toτ 0 , we have<br />

− dρ (τ) ≤−dρ<br />

dτ dτ (τ 0) − C 2 {ρ(τ 0 ) − ρ(τ)} + C 1 (τ 0 − τ)<br />

≤−¯s 0 + C 1 (τ 0 − τ)<br />

< − r 0<br />

τ 0 − τ 1<br />

.


22 CHAPTER 1. ORDINARY DIFFERENTIAL EQUATIONS<br />

Integrating again both sides from τ 1 to τ 0 , we then obtain<br />

ρ(τ 1 ) < 0.<br />

Therefore, ¯s 0 ∈A(r 0 ), which also contradicts the fact A(r 0 )=∅. Hence<br />

dr<br />

dτ (τ) > 0 on (¯τ,τ 0].<br />

Assume that ¯τ >−∞. Then, it follows from Lemma 1.1.5 that |r(τ)| →∞as τ → ¯τ +0. If<br />

r(τ) →∞, then, by the mean value theorem, there exists a point τ 1 ∈ (¯τ,τ 0 ] such that<br />

dr<br />

dτ (τ 1)=0,<br />

which implies that s 0 ∈A(r 0 ), a contradiction.<br />

Similarly, if r(τ) →−∞, then there exists a point τ 2 ∈ (¯τ,τ 0 ] such that<br />

implying that s 0 ∈A(r 0 ), a contradiction.<br />

As a consequence, ¯τ = −∞, and<br />

r(τ 2 )=0,<br />

r(τ) > 0<br />

and<br />

dr<br />

dτ (τ) > 0 on (−∞,τ 0].<br />

(2) We first prove that A(r 0 ) is an open set. Fix r 0 > 0, and take s 0 ∈A(r 0 ) arbitrarily.<br />

Let r = r(τ) be a solution to the equation (1.1.4) with the initial condition (1.1.5). Then<br />

we can find τ 1 0 such that<br />

r(τ 1 ) = 0 and r(τ) > 0 on (τ 1 ,τ 0 ).<br />

dr<br />

dτ (τ) > 0 on I := [τ 1 − ε, τ 0 ].<br />

Set<br />

η := 1 2 min<br />

τ∈I<br />

dr<br />

dτ (τ), ξ := −1 2 r(τ 1 − ε) and δ := min{η, ξ}.


1.1. LOCAL EXISTENCE OF SOLUTIONS TO (1.1.1) WITH (1.1.2) 23<br />

Since the solutions depend continuously on initial values, it follows that for the above δ>0,<br />

there exists an ¯ε >0 such that if | ¯s 0 − s 0 | < ¯ε, then<br />

∣ sup |ρ(τ) − r(τ)| + sup<br />

dρ dr ∣∣∣<br />

∣ (τ) −<br />

τ∈I<br />

τ∈I dτ dτ (τ) <br />

dτ dτ (τ) − δ>η>0 on [τ 1 − ε, τ 0 ],<br />

ρ(τ 1 − ε) 0. Suppose that inf A(r 0 )=0. For s 0 ∈A(r 0 ), let r = r(τ)<br />

be a solution to the equation (1.1.4) with the initial condition (1.1.5). Then<br />

d 2 r<br />

dτ (τ 0)=−P (τ 2 0 )s 0 + Q(τ 0 ,r 0 ).<br />

Since Q(τ 0 ,r 0 ) > 0, for sufficiently small δ>0, one can take s 0 > 0 so that<br />

d 2 r<br />

dτ (τ 0) > 2δ.<br />

2<br />

Hence there exists an ε = ε(s 0 ) > 0 such that<br />

d 2 r dr<br />

(τ) >δ,<br />

dτ<br />

2<br />

dτ (τ) > 0 and r(τ) > 0 on (τ 0 − 2ε, τ 0 +2ε).<br />

For ¯s 0 ∈A(r 0 ) such that ¯s 0 r(τ) and<br />

Thus we obtain<br />

dρ dr<br />

(τ) <<br />

dτ dτ (τ).<br />

d 2 ρ<br />

(τ) =−P (τ)dρ<br />

dτ<br />

2<br />

dτ (τ)+Q(τ,ρ(τ))<br />

≥−P (τ) dr<br />

dτ (τ)+Q(τ,r(τ)) = d2 r<br />

(τ) >δ<br />

dτ<br />

2


24 CHAPTER 1. ORDINARY DIFFERENTIAL EQUATIONS<br />

on (τ 0 − 2ε, τ 0 +2ε). Note that ε does not depend on the choice <strong>of</strong> ¯s 0 , if ¯s 0 0.<br />

Now, we are in a position to prove Theorem 1.1.4.<br />

Pro<strong>of</strong> <strong>of</strong> Theorem 1.1.4.<br />

(1) Take r 0 > 0 arbitrarily. If A(r 0 ) is an empty set, we have already shown that there<br />

exists a solution r = r(τ) on(−∞,τ 0 ] to the equation (1.1.4) satisfying<br />

dr<br />

r(τ) > 0 and (τ) > 0.<br />

dτ<br />

Thus we assume that A(r 0 ) is not empty. Let r 0 ′ = inf A(r 0 ), and let r = r(τ) be a solution<br />

to the equation (1.1.4) satisfying<br />

dr<br />

r(τ 0 )=r 0 and<br />

dτ (τ 0)=r 0.<br />

′<br />

Note that r 0 ′ ∉A(r 0 ) because A(r 0 ) is an open set.<br />

Suppose that (τ ∗ ,τ 0 ] is the life span <strong>of</strong> r = r(τ). We first show that<br />

r(τ) > 0 on (τ ∗ ,τ 0 ].<br />

We assume that there exists τ 1 ∈ (τ ∗ ,τ 0 ) such that<br />

r(τ 1 ) = 0 and r(τ) > 0 on (τ 1 ,τ 0 ].<br />

It is easily verified that if (dr/dτ)(τ 1 )=0, then r = r(τ) must be the zero solution. Thus,<br />

since r ′ 0 ∉A(r 0 ), there exists τ 2 ∈ (τ 1 ,τ 0 ) such that<br />

r(τ 2 ) > 0<br />

and<br />

dr<br />

dτ (τ 2)=0.


1.1. LOCAL EXISTENCE OF SOLUTIONS TO (1.1.1) WITH (1.1.2) 25<br />

Thus<br />

d 2 r<br />

dτ (τ 2)=−P (τ 2 2 ) dr<br />

dτ (τ 2)+Q(τ 2 ,r(τ 2 ))<br />

= Q(τ 2 ,r(τ 2 )) > 0.<br />

Therefore, r(τ 2 ) is locally a minimum value. On the other hand, we can show that<br />

dr<br />

dτ (τ) < 0<br />

on (τ 1 ,τ 2 ). Indeed, suppose that there exists τ 3 ∈ (τ 1 ,τ 2 ) such that<br />

dr<br />

dτ (τ 3) = 0<br />

and<br />

dr<br />

dτ (τ) < 0 on (τ 3,τ 2 ).<br />

Then we have<br />

d 2 r<br />

dτ (τ 3) ≤ 0.<br />

2<br />

However, from the equation (1.1.4), it holds that<br />

d 2 r<br />

dτ 2 (τ 3)=Q(τ 3 ,r(τ 3 )) > 0,<br />

which is a contradiction. Hence, r(τ 2 ) is positive and is the minimal value <strong>of</strong> r on [τ 1 ,τ 0 ],<br />

which contradicts the assumption r(τ 1 )=0. Thus r>0on(τ ∗ ,τ 0 ].<br />

Second, we show that<br />

dr<br />

dτ (τ) > 0 on (τ ∗,τ 0 ].<br />

Assume that there exists τ 1 ∈ (τ ∗ ,τ 0 ) such that<br />

dr<br />

dτ (τ 1)=0.<br />

Then, by the same argument used for proving r>0, we have<br />

dr<br />

dτ (τ) < 0 on (τ ∗,τ 1 ) and<br />

dr<br />

dτ (τ) > 0 on (τ 1,τ 0 ],<br />

which implies that r(τ 1 ) is the minimal value <strong>of</strong> r on (τ ∗ ,τ 0 ]. Take τ 2 ∈ (τ ∗ ,τ 1 ) and τ 3 ∈ (τ 1 ,τ 0 )<br />

so that r(τ 2 )=r(τ 3 ) and I =[τ 2 ,τ 3 ]. It then follows from the continuous dependence <strong>of</strong>


26 CHAPTER 1. ORDINARY DIFFERENTIAL EQUATIONS<br />

solutions on initial values that for any positive number ε0 such that<br />

∣ sup |ρ(τ) − r(τ)| + sup<br />

dρ dr ∣∣∣<br />

∣ (τ) −<br />

τ∈I<br />

τ∈I dτ dτ (τ) 0 on [τ 2 ,τ 0 ].<br />

On the other hand, we have<br />

ρ(τ 2 ) >r(τ 2 ) − ε>r(τ 1 ) >ρ(τ 1 ) > 0,<br />

ρ(τ 3 ) >r(τ 3 ) − ε>r(τ 1 ) >ρ(τ 1 ) > 0.<br />

Hence there exists τ 4 ∈ (τ 2 ,τ 3 ) such that (dρ/dτ)(τ 4 )=0. As a consequence, ρ = ρ(τ)<br />

satisfies<br />

dρ<br />

ρ>0 on [τ 2 ,τ 0 ] and<br />

dτ (τ 4) = 0 for τ 4 ∈ (τ 2 ,τ 0 ),<br />

which contradict the choice <strong>of</strong> s 0 . Therefore, we have<br />

dr<br />

dτ (τ) > 0 on (τ ∗,τ 0 ].<br />

Finally, we shall prove that τ ∗ = −∞. Ifτ ∗ > −∞, then r = r(τ) blows up at τ ∗ . Then<br />

it follows from Lemma 1.1.5 that<br />

lim r(τ) =+∞.<br />

τ→τ ∗ +0<br />

In this case, there exists ¯τ ∈ (τ ∗ ,τ 0 ) such that (dr/dτ)(¯τ) =0, which contradicts the fact<br />

that (dr/dτ)(τ) > 0on(τ ∗ ,τ 0 ]. Thus τ ∗ = −∞.<br />

(2) Although it can be proved, by using the same argument as that in [22], that<br />

lim<br />

τ→−∞<br />

dr<br />

(τ) =0,<br />


1.1. LOCAL EXISTENCE OF SOLUTIONS TO (1.1.1) WITH (1.1.2) 27<br />

we present here a pro<strong>of</strong> in our context.<br />

Take τ 1 ∈ (−∞,τ 0 ) arbitrarily, and let A be a positive constant such that P (τ) ≤ A for<br />

any τ ∈ (−∞,τ 1 ]. Fix τ 2 ∈ (−∞,τ 1 − 1] arbitrarily, and let ρ be a unique solution to the<br />

following ordinary differential equation with the initial values:<br />

⎧<br />

d<br />

⎪⎨<br />

2 ρ<br />

dτ (τ)+Adρ (τ) =0,<br />

2 dτ<br />

⎪⎩ ρ(τ 2 )=r(τ 2 )<br />

Then it is easy to see that ρ is given by<br />

and<br />

dρ<br />

dτ (τ 2)= dr<br />

dτ (τ 2).<br />

ρ(τ) = 1 A (1 − e−A(τ−τ 2) ) dr<br />

dτ (τ 2)+r(τ 2 ).<br />

Let<br />

R(τ) =ρ(τ) dr (τ) − r(τ)dρ<br />

dτ dτ (τ).<br />

Then R(τ 2 ) = 0 and (dR/dτ)(τ 2 ) > 0. Moreover, R ≥ 0on(τ 2 ,τ 1 ]. Indeed, if this is not the<br />

case, then there exists a point τ ∗ ∈ (τ 2 ,τ 1 ] such that<br />

R(τ ∗ )=0,<br />

dR<br />

dτ (τ ∗) < 0, and R>0 on (τ 2 ,τ ∗ ).<br />

On the other hand, we have<br />

dR<br />

dτ (τ ∗)=ρ(τ ∗ ) d2 r<br />

dτ (τ ∗) − r(τ 2 ∗ ) d2 ρ<br />

dτ (τ ∗)<br />

2<br />

= ρ(τ ∗ ){−P (τ ∗ ) dr<br />

dτ (τ ∗)+Q(τ ∗ ,r(τ ∗ ))} + Ar(τ ∗ ) dρ<br />

dτ (τ ∗)<br />

≥ A{r(τ ∗ ) dρ<br />

dτ (τ ∗) − ρ(τ ∗ ) dr<br />

dτ (τ ∗)} + Q(τ ∗ ,r(τ ∗ ))ρ(τ ∗ )<br />

= Q(τ ∗ ,r(τ ∗ ))ρ(τ ∗ ) > 0,<br />

which leads to a contradiction. Thus R ≥ 0on[τ 2 ,τ 1 ]. Since r>0 and ρ>0, it holds that<br />

r ≥ ρ on [τ 2 ,τ 1 ].<br />

As a consequence, we obtain<br />

(1.1.7) r(τ 2 +1)− r(τ 2 ) ≥ ρ(τ 2 +1)− ρ(τ 2 )= 1 dr<br />

A dτ (τ 2)(1 − e −A )


28 CHAPTER 1. ORDINARY DIFFERENTIAL EQUATIONS<br />

for any τ 2 ∈ (−∞,τ 1 − 1].<br />

Now suppose that<br />

lim sup<br />

τ→−∞<br />

dr<br />

(τ) > 0.<br />

dτ<br />

Then there exist a positive constant δ and a sequence {τ j } satisfying<br />

τ j+1 ≤ τ j − 1 and r(τ j ) ≥ δ.<br />

Then, from the inequality (1.1.7), we have<br />

r(τ j +1)− r(τ j ) ≥ 1 dr<br />

A dτ (τ j)(1 − e −A ) > δ A (1 − e−A ).<br />

Since r is a monotone increasing function, it holds that<br />

r(τ j ) 0. Noting that P (τ) is bounded on (−∞,τ 0 ], we have<br />

Thus there exists δ>0 such that<br />

for τ 1 sufficiently small. Hence we obtain<br />

d 2 r<br />

lim<br />

τ→−∞ dτ (τ) = lim Q(τ,η) > 0.<br />

2 τ→−∞<br />

d 2 r<br />

dτ 2 (τ) ≥ δ on (−∞,τ 1)<br />

dr<br />

dτ (τ) < 0


1.1. LOCAL EXISTENCE OF SOLUTIONS TO (1.1.1) WITH (1.1.2) 29<br />

for sufficiently small τ, which is a contradiction. Therefore<br />

Finally, the equation (1.1.4) implies that<br />

It follows from the condition (C-1) that<br />

lim r(τ) =0.<br />

τ→−∞<br />

d 2 r<br />

dr<br />

(τ) =−P (τ)<br />

dτ<br />

2<br />

dτ (τ)+Q(τ,r(τ)).<br />

lim P (τ) exists and lim<br />

τ→−∞<br />

Q(τ,r(τ))=0,<br />

τ→−∞<br />

which implies that<br />

Thus the pro<strong>of</strong> is completed.<br />

lim<br />

τ→−∞<br />

d 2 r<br />

(τ) =0.<br />

dτ<br />

2<br />

In the remainder <strong>of</strong> this section, by making use <strong>of</strong> a comparison function, we shall investigate<br />

the regularity <strong>of</strong> the solution r = r(t) att =0.<br />

We assume that there exist k, l ≥ 1 such that F and G satisfy the following conditions.<br />

lim<br />

t→0<br />

tF (t) =k,<br />

G(t, s)<br />

lim t2<br />

t→0,s→0 s<br />

= l.<br />

Then we can prove the following<br />

Proposition 1.1.9. Let r = r(t) be a solution to the equation (1.1.3) with the boundary<br />

condition (1.1.2). Then there exist a>0 and t 1 ∈ (0,t 0 ) such that<br />

for t ∈ (0,t 1 ).<br />

0


30 CHAPTER 1. ORDINARY DIFFERENTIAL EQUATIONS<br />

For any ε>0 we can find t 1 ∈ (0,t 0 ) such that<br />

|tF (t) − k| 0 such<br />

that a solution r = r(t) to the equation (1.1.3) satisfies<br />

⎧<br />

⎪⎨ r(t 0 )=r 0 , ṙ(t 0 )=s 0 ,<br />

⎪⎩ lim r(t) =0, and ṙ(t) > 0 on (0,t 0 ].<br />

t→0<br />

We shall show that for a suitable r 0 > 0, there exists a global solution to the equation (1.1.1)<br />

with the initial condition r(t 0 )=r 0 .<br />

The following lemma is an analogue <strong>of</strong> Lemma 1.1.5.<br />

Lemma 1.2.1. Let r = r(t) be a solution to the equation (1.1.3) with the boundary condition<br />

(1.1.2), and let [0,T) be its life span.


1.2. GLOBAL PROPERTIES OF SOLUTIONS 31<br />

(1) Assume that r(t 0 ) > 0 and ṙ(t 0 ) > 0 for some t 0 ∈ (0,T). Then it holds that<br />

ṙ(t) > 0<br />

on (0,T). Hence the solution constructed in the previous section is strictly monotone increasing<br />

as long as it exists.<br />

(2) If T 0on(0,t 0 ]. We shall show that ṙ(t) > 0on(t 0 ,T). If<br />

this is not the case, we have a point t 1 ∈ (t 0 ,T) such that<br />

r(t 1 ) > 0, ṙ(t 1 ) = 0 and ¨r(t 1 ) ≤ 0.<br />

On the other hand, the equation (1.1.3) asserts that<br />

¨r(t 1 )=G(t 1 ,r(t 1 )) > 0,<br />

which is a contradiction. Thus ṙ(t) > 0on[t 0 ,T).<br />

(2) Let ˜F (t) = exp ∫ t F (s)ds. Then, from the equation (1.1.3), we have<br />

d<br />

{ }<br />

˜F (t)ṙ(t) =<br />

dt<br />

˜F (t)G(t, r(t)).<br />

Integrating both sides from t 0 to t(> t 0 ), we obtain<br />

˜F (t) |ṙ(t)| ≤ ˜F (t 0 ) |ṙ(t 0 )| +<br />

∫ t<br />

˜F (t)|G(t, r(t))|dt.<br />

Since G ∈ C ∞ ([0, ∞) × R), if r(t) is bounded when t tends to T, then so is ṙ(t). Thus the<br />

assertion holds.<br />

From now on, we shall consider the original ordinary differential equation (1.1.1).<br />

t 0


32 CHAPTER 1. ORDINARY DIFFERENTIAL EQUATIONS<br />

Lemma 1.2.2.<br />

(1.1.2), then it satisfies<br />

If r = r(t) is a solution to the equation (1.1.1) with the boundary condition<br />

(1.2.1) ṙ(t) 2 ≤ µ2<br />

f 1 (t) h 1(r(t)) 2 +<br />

ν2<br />

2 f 2 (t) h 2(r(t)) 2<br />

2<br />

on [t 0 ,T).<br />

Pro<strong>of</strong>.<br />

It follows from the equation (1.1.1) that<br />

d<br />

dt {f 1(t) p f 2 (t) q ṙ(t)}<br />

= f 1 (t) p−2 f 2 (t) q−2 {µ 2 f 2 (t) 2 h 1 (r(t))h ′ 1(r(t)) + ν 2 f 1 (t) 2 h 2 (r(t))h ′ 2(r(t))}.<br />

Multiplying both sides by f 1 (t) p f 2 (t) q ṙ(t), we obtain<br />

d<br />

dt {f 1(t) p f 2 (t) q ṙ(t)} 2<br />

= µ 2 f 1 (t) 2p−2 f 2 (t) 2q d dt h 1(r(t)) 2 + ν 2 f 1 (t) 2p f 2 (t) 2q−2 d dt h 2(r(t)) 2 .<br />

Integrating both sides <strong>of</strong> this equation from t 1 to t, we have<br />

{f 1 (t) p f 2 (t) q ṙ(t)} 2<br />

= {f 1 (t 1 ) p f 2 (t 1 ) q ṙ(t 1 )} 2<br />

(1.2.2)<br />

+ µ 2 {f 1 (t) 2p−2 f 2 (t) 2q h 1 (r(t)) 2 − f 1 (t 1 ) 2p−2 f 2 (t 1 ) 2q h 1 (r(t 1 )) 2 }<br />

+ ν 2 {f 1 (t) 2p f 2 (t) 2q−2 h 2 (r(t)) 2 − f 1 (t 1 ) 2p f 2 (t 1 ) 2q−2 h 2 (r(t 1 )) 2 }<br />

∫ t<br />

− µ 2 h 1 (r(τ)) 2 d<br />

t 1<br />

dτ {f 1(τ) 2p−2 f 2 (τ) 2q }dτ<br />

−<br />

∫ t<br />

t 1<br />

ν 2 h 2 (r(τ)) 2 d<br />

dτ {f 1(τ) 2p f 2 (τ) 2q−2 }dτ.<br />

The condition (F-1) implies<br />

d<br />

dτ {f 1(τ) m f 2 (τ) n }≥0


1.2. GLOBAL PROPERTIES OF SOLUTIONS 33<br />

for any nonnegative integers m and n. As a consequence, we have<br />

{f 1 (t) p f 2 (t) q ṙ(t)} 2 ≤{f 1 (t 1 ) p f 2 (t 1 ) q ṙ(t 1 )} 2<br />

Letting t 1 tend to 0, we have<br />

+ µ 2 {f 1 (t) 2p−2 f 2 (t) 2q h 1 (r(t)) 2 − f 1 (t 1 ) 2p−2 f 2 (t 1 ) 2q h 1 (r(t 1 )) 2 }<br />

+ ν 2 {f 1 (t) 2p f 2 (t) 2q−2 h 2 (r(t)) 2 − f 1 (t 1 ) 2p f 2 (t 1 ) 2q−2 h 2 (r(t 1 )) 2 }.<br />

{f 1 (t) p f 2 (t) q ṙ(t)} 2 ≤ µ 2 f 1 (t) 2p−2 f 2 (t) 2q h 1 (r(t)) 2 + ν 2 f 1 (t) 2p f 2 (t) 2q−2 h 2 (r(t)) 2 .<br />

Dividing both sides by f 1 (t) 2p f 2 (t) 2q , we obtain the conclusion.<br />

Using this lemma, we present a sufficient condition for the existence <strong>of</strong> a global solution<br />

in terms <strong>of</strong> a relation <strong>between</strong> t 0 and r 0 .<br />

If r(t) = 0 for some t>0, then r(t) ≡ 0 by Corollary 1.1.3. Hence it is a global solution.<br />

In the sequel <strong>of</strong> this section we assume that r(t) > 0 for all t>0.<br />

Let<br />

⎧<br />

⎪⎨ max{h 1 (r),h 2 (r)} (ν >0),<br />

h(r) =<br />

⎪⎩ h 1 (r) (ν =0).<br />

Since ṙ(t) > 0, we have<br />

{<br />

ṙ(t) 1<br />

h(r(t)) ≤ γ f 1 (t) + 1 }<br />

,<br />

f 2 (t)<br />

where γ = max{µ, ν}. Integrating both sides from t 0 to t(∈ [t 0 ,T]), we obtain<br />

∫ r(t) ∫<br />

ds<br />

t<br />

{ 1<br />

(1.2.3)<br />

h(s) ≤ γ t 0<br />

f 1 (τ) + 1 }<br />

dτ.<br />

f 2 (τ)<br />

r 0<br />

Theorem 1.2.3. If r 0 satisfies<br />

∫ ∞ ∫<br />

dr<br />

∞<br />

{ 1<br />

(1.2.4)<br />

r 0<br />

h(r) >γ t 0<br />

f 1 (τ) + 1 }<br />

dτ,<br />

f 2 (τ)<br />

then the solution to the equation (1.1.1) with the boundary condition (1.1.2) exists globally<br />

and is bounded.


34 CHAPTER 1. ORDINARY DIFFERENTIAL EQUATIONS<br />

Pro<strong>of</strong>. Let [0,T) be the life span <strong>of</strong> r, and assume that T0. We shall<br />

(l − ε, l + ε) ∩ L ≠ ∅<br />

for any ε ∈ (0,l). Note that we may choose T 0 > 0 so that<br />

∫ ∞ ∫<br />

ds<br />

∞<br />

{ 1<br />

l−ε h(s) >γ t 0<br />

f 1 (τ) + 1 }<br />

dτ<br />

f 2 (τ)


1.2. GLOBAL PROPERTIES OF SOLUTIONS 35<br />

for any t 0 >T 0 . Then, by Theorem 1.2.3, there exists globally a solution r = r(t) to the<br />

equation (1.1.1) with the boundary condition (1.1.2) satisfying<br />

r(t 0 )=l − ε>0.<br />

Let r(t; t 0 ) denote this solution. Since it is bounded and increasing, the limit r(∞; t 0 ) exists<br />

and<br />

0 ···→l,<br />

lim ¯r j (t) =¯l j .<br />

t→∞


36 CHAPTER 1. ORDINARY DIFFERENTIAL EQUATIONS<br />

It follows from Corollary 1.1.3 that<br />

r 1 (t) 0, and define two positive numbers α and ᾱ by<br />

α = lim<br />

i→∞<br />

r i (t 0 ) and ᾱ = lim<br />

j→∞<br />

¯r j (t 0 ).<br />

Choose α such that α ≤ α ≤ ᾱ, and let r α be a unique solution to the equation (1.1.1) with<br />

the boundary condition (1.1.2) satisfying r α (t 0 )=α. Making use <strong>of</strong> Corollary 1.1.3 again,<br />

we get<br />

r i (t) 0<br />

there exists a solution to the equation (1.1.1) with the boundary condition (1.1.2) satisfying<br />

r(t) → +∞ as t → T.<br />

We suppose that h(s) = max{h 1 (s),h 2 (s)} satisfies<br />

(1.2.5)<br />

∫ ∞<br />

ds<br />

h(s) < ∞.<br />

Note that if the assumption (1.2.5) is false, then any solution is bounded (Corollary 1.2.4).<br />

For t 0 > 0 and r 0 > 0, let r = r(t) be a solution to the equation (1.1.1) with the boundary<br />

condition (1.1.2) satisfying r(t 0 )=r 0 . Such a solution exists uniquely. We denote ṙ(t 0 )by<br />

β(r 0 ,t 0 ), which is uniquely determined by r 0 and t 0 . We define<br />

φ(r 0 ,t 0 )=<br />

where<br />

∫ ∞<br />

r 0<br />

[γ 1 (t 0 ) 2 {h 1 (s) 2 − h 1 (r 0 ) 2 } + γ 2 (t 0 ) 2 {h 2 (s) 2 − h 2 (r 0 ) 2 } + β(r 0 ,t 0 ) 2 ] −1/2 ds,<br />

γ 1 (t 0 )= µ<br />

f 1 (t 0 )<br />

and γ 2 (t 0 )= ν<br />

f 2 (t 0 ).


1.2. GLOBAL PROPERTIES OF SOLUTIONS 37<br />

Lemma 1.2.7. The function φ(r 0 ,t 0 ) is well-defined for r 0 > 0 and t 0 > 0, and<br />

lim φ(r 0,t 0 )=0.<br />

r 0 →∞<br />

Pro<strong>of</strong>. It follows from (H-3) that (h 2 i ) ′ (s) ≥ (h 2 i ) ′ (s − r 0 ) for s>r 0 . Integrating both<br />

sides with respect to s from r 0 to s, we get<br />

Therefore, it holds that for K>0<br />

∫ ∞<br />

r 0 +K<br />

h i (s) 2 − h i (r 0 ) 2 ≥ h i (s − r 0 ) 2 .<br />

[γ 1 (t 0 ) 2 {h 1 (s) 2 − h 1 (r 0 ) 2 } + γ 2 (t 0 ) 2 {h 2 (s) 2 − h 2 (r 0 ) 2 } + β(r 0 ,t 0 ) 2 ] −1/2 ds<br />

≤ C(t 0 )<br />

≤ C(t 0 )<br />

∫ ∞<br />

r 0 +K<br />

∫ ∞<br />

K<br />

[h 1 (s − r 0 ) 2 + h 2 (s − r 0 ) 2 ] −1/2 ds<br />

ds<br />

h(s) .<br />

Note that it follows from the condition (1.2.5) that for any ε>0<br />

for sufficiently large K.<br />

∫ ∞<br />

K<br />

ds<br />

h(s) r 0<br />

where ρ i ∈ (r 0 ,s). It follows from (H-3) that<br />

Therefore it holds that<br />

∫ r0 +K<br />

r 0<br />

h i (s) 2 − h i (r 0 ) 2 =2h i (ρ i )h ′ i (ρ i)(s − r 0 ),<br />

h i (s) 2 − h i (r 0 ) 2 ≥ 2h i (r 0 )h ′ i(r 0 )(s − r 0 ) > 0.<br />

[γ 1 (t 0 ) 2 {h 1 (s) 2 − h 1 (r 0 ) 2 } + γ 2 (t 0 ) 2 {h 2 (s) 2 − h 2 (r 0 ) 2 } + β(r 0 ,t 0 ) 2 ] −1/2 ds<br />

≤ [2(γ 1 (t 0 ) 2 h 1 (r 0 )h ′ 1(r 0 )+γ 2 (t 0 ) 2 h 2 (r 0 )h ′ 2(r 0 ))] −1/2 ∫ r0 +K<br />

≤ C(t 0 ) √ K min { [h 1 (r 0 )h ′ 1 (r 0)] −1/2 , [h 2 (r 0 )h ′ 2 (r 0)] −1/2} .<br />

r 0<br />

ds<br />

√ s − r0


38 CHAPTER 1. ORDINARY DIFFERENTIAL EQUATIONS<br />

We shall show that the right hand side converges to zero as r 0 →∞. Assume that h i (s)h ′ i (s)’s<br />

are bounded functions. Then, by virtue <strong>of</strong> (H-3), we get<br />

h i (s) 2 ≤ Cs<br />

for some positive constant C. As a result, we have<br />

∫ ∞<br />

ds<br />

h(s) ≥ C ∫ ∞<br />

ds<br />

√ s<br />

= ∞,<br />

which contradicts (1.2.5). Hence h i (s)h ′ i (s)’s are unbounded. Taking (H-3) into consideration,<br />

we get<br />

{√<br />

lim max h1 (r 0 )h ′<br />

r 0<br />

1(r 0 ), √ }<br />

h 2 (r 0 )h ′ 2(r 0 ) = ∞,<br />

→∞<br />

and therefore<br />

∫ r0 +K<br />

lim [γ 1 (t 0 ) 2 {h 1 (s) 2 − h 1 (r 0 ) 2 } + γ 2 (t 0 ) 2 {h 2 (s) 2 − h 2 (r 0 ) 2 } + β(r 0 ,t 0 ) 2 ] −1/2 ds =0.<br />

r 0 →∞<br />

r 0<br />

Summing up these estimates for integrations over [r 0 ,r 0 + K] and [r 0 + K, ∞), we obtain<br />

the assertion.<br />

Proposition 1.2.8. The set<br />

{<br />

}<br />

B = T ∈ (0, ∞)<br />

∣ lim r(t) =∞, where r is a solution to (1.1.1) with (1.1.2) t↑T<br />

is dense in (0, ∞).<br />

Pro<strong>of</strong>. Let T>0. We shall show that<br />

(T − ε, T + ε) ∩ B ≠ ∅<br />

for any ε ∈ (0,T). Let r α be a solution to the equation (1.1.1) with the boundary condition<br />

(1.1.2) satisfying r α (T )=α, and [0,T α ) its life-span. By Lemma 1.2.9 below, we have<br />

φ(α, T ) ≥ f 1 (T ) p f 2 (T ) q ∫ Tα<br />

T<br />

dτ<br />

f 1 (τ) p f 2 (τ) q .


1.2. GLOBAL PROPERTIES OF SOLUTIONS 39<br />

Letting α →∞, we obtain from Lemma 1.2.7 that<br />

lim T α = T.<br />

α→∞<br />

Hence T α ∈ (T − ε, T + ε) for sufficiently large α.<br />

Lemma 1.2.9. Let [0,T) be the life span <strong>of</strong> a solution r. Then we obtain<br />

φ(α, t 0 ) ≥ f 1 (t 0 ) p f 2 (t 0 ) q ∫ T<br />

t 0<br />

dt<br />

f 1 (t) p f 2 (t) q .<br />

Pro<strong>of</strong>.<br />

Since ˙ f i (t) ≥ 0 and h i (s) 2 ≤ h i (t) 2 for all s ≤ t, we have<br />

∫ t<br />

and<br />

∫ t<br />

t 0<br />

h 1 (r(τ)) 2 d<br />

dτ {f 1(τ) 2p−2 f 2 (τ) 2q }dτ ≤ h 1 (r(t)) 2 {f 1 (t) 2p−2 f 2 (t) 2q − f 1 (t 0 ) 2p−2 f 2 (t 0 ) 2q },<br />

h 2 (r(τ)) 2 d<br />

t 0<br />

dτ {f 1(τ) 2p f 2 (τ) 2q−2 }dτ ≤ h 2 (r(t)) 2 {f 1 (t) 2p f 2 (t) 2q−2 − f 1 (t 0 ) 2p f 2 (t 0 ) 2q−2 }.<br />

Making use <strong>of</strong> the equation (1.2.2) and these inequalities, we get<br />

which is equivalent to<br />

{f 1 (t) p f 2 (t) q ṙ(t)} 2 ≥{f 1 (t 0 ) p f 2 (t 0 ) q ṙ(t 0 )} 2<br />

+ µ 2 f 1 (t 0 ) 2p−2 f 2 (t 0 ) 2q {h 1 (r(t)) 2 − h 1 (r(t 0 )) 2 }<br />

+ ν 2 f 1 (t 0 ) 2p f 2 (t 0 ) 2q−2 {h 2 (r(t)) 2 − h 2 (r(t 0 )) 2 },<br />

ṙ(t)[γ 1 (t 0 ) 2 {h 1 (r(t)) 2 − h 1 (r 0 ) 2 } + γ 2 (t 0 ) 2 {h 2 (r(t)) 2 − h 2 (r 0 ) 2 } + β(r 0 ,t 0 ) 2 ] −1/2<br />

≥ f 1 (t 0 ) p f 2 (t 0 ) q 1<br />

f 1 (t) p f 2 (t) q .<br />

Integrating both sides from t 0 to T , we obtain<br />

∫ r(T )<br />

r 0<br />

[γ 1 (t 0 ) 2 {h 1 (s) 2 − h 1 (r 0 ) 2 } + γ 2 (t 0 ) 2 {h 2 (s) 2 − h 2 (r 0 ) 2 } + β(r 0 ,t 0 ) 2 ] −1/2 ds<br />

≥ f 1 (t 0 ) p f 2 (t 0 ) q ∫ T<br />

t 0<br />

dt<br />

f 1 (t) p f 2 (t) q .


40 CHAPTER 1. ORDINARY DIFFERENTIAL EQUATIONS<br />

This implies that<br />

φ(r 0 ,t 0 ) ≥ f 1 (t 0 ) p f 2 (t 0 ) q ∫ T<br />

t 0<br />

dt<br />

f 1 (t) p f 2 (t) q .<br />

Note.<br />

For any t ∈ (0,T), we can prove that<br />

φ(r(t),t) ≥ f 1 (t) p f 2 (t) q ∫ T<br />

t<br />

dτ<br />

f 1 (τ) p f 2 (τ) q .<br />

Theorem 1.2.10. For any given T>0, there exists a solution to the equation (1.1.1) with<br />

the boundary condition (1.1.2) satisfying lim r(t) =+∞.<br />

t→T<br />

Pro<strong>of</strong>. By virtue <strong>of</strong> Proposition 1.2.8, there exist sequences {T i }, { ¯T j } and solutions r i ,<br />

¯r j to the equation (1.1.1) with the boundary condition (1.1.2) such that<br />

and<br />

It follows from Corollary 1.1.3 that<br />

0


1.3. ADDENDUM 41<br />

Theorem 1.2.11. There exists a global, positive and strictly monotone increasing solution<br />

r = r(t) satisfying<br />

lim r(t) =∞.<br />

t→∞<br />

Pro<strong>of</strong>. Fix t 0 > 0 arbitrarily. Then, it follows from Theorem 1.2.8 and Theorem 1.2.10<br />

that there exist sequences <strong>of</strong> positive numbers {T i } ∞ i=1, {l j } ∞ j=1 and sequences <strong>of</strong> solutions<br />

{r i } ∞ i=1, {r j } ∞ j=1 to the equation (1.1.1) with the boundary condition (1.1.2) so that<br />

0


42 CHAPTER 1. ORDINARY DIFFERENTIAL EQUATIONS<br />

Under the condition<br />

the identity (1.3.2) becomes<br />

Thus<br />

r(0) = 0 and ṙ(0) = 0,<br />

{f(t)ṙ(t)} 2 = µ 2 h(r(t)) 2 .<br />

f(t)ṙ(t) =µh(r(t)).<br />

As a consequence, we can solve the ordinary differential equation (1.3.1) by making use <strong>of</strong><br />

the method <strong>of</strong> separation <strong>of</strong> variables.


CHAPTER 2<br />

Equivariant <strong>harmonic</strong> <strong>maps</strong><br />

In this chapter, we shall construct equivariant <strong>harmonic</strong> <strong>maps</strong> <strong>between</strong> noncompact, complete<br />

Riemannian <strong>manifolds</strong> by making use <strong>of</strong> the results established in the previous chapter.<br />

We begin with fixing our notation and a brief review <strong>of</strong> relevant theorems for eigen<strong>maps</strong>.<br />

2.1 Eigen<strong>maps</strong><br />

Let (S m ,g S m) be the m-dimensional standard unit sphere. A map ϕ :(S m ,g S m) → (S n ,g S n)<br />

is called an eigenmap if it is a <strong>harmonic</strong> map with constant energy density. It is a well-known<br />

result that from Takahashi’s theorem, we have the following<br />

Lemma 2.1.1. Let ϕ :(S m ,g S m) → (S n ,g S n) be an eigenmap. Then all components <strong>of</strong> the<br />

map Φ=i ◦ ϕ : S m → R n+1 are <strong>harmonic</strong> homogeneous polynomials on R m+1 <strong>of</strong> the same<br />

polynomial degree, where i : S n → R n+1 is the inclusion map.<br />

Thus, if we can find a family <strong>of</strong> <strong>harmonic</strong> homogeneous polynomials <strong>of</strong> the same degree,<br />

say {Φ i } n+1<br />

i=1 , satisfying ∑n+1<br />

Φ 2 i (x) = 1 for x ∈ S m ,<br />

i=1<br />

then we obtain an eigenmap ϕ : S m → S n . An eigenmap ϕ : S m → S n is said to be <strong>of</strong> degree<br />

k if all components <strong>of</strong> ϕ are <strong>harmonic</strong> homogeneous polynomial <strong>of</strong> degree k.<br />

There is no general theory <strong>of</strong> constructing an eigenmap ϕ : S m → S n for a given degree.<br />

It should be remarked that almost all known results have been for the case <strong>of</strong> degree two.<br />

We shall give an algorithm <strong>of</strong> constructing a new eigenmap <strong>of</strong> degree two from the old ones<br />

<strong>of</strong> the same degree.<br />

43


44 CHAPTER 2. EQUIVARIANT HARMONIC MAPS<br />

A bilinear map F : R m × R n → R r is called an orthogonal multiplication if for any<br />

x ∈ R m and y ∈ R n it holds that<br />

‖F (x, y)‖ = ‖x‖‖y‖.<br />

An orthogonal multiplication F : R m × R n → R r is said to be full if its image is not<br />

contained in any hyperplane <strong>of</strong> R r .<br />

In the case <strong>of</strong> m = n, associated with an orthogonal multiplication F : R m × R m → R r ,<br />

we obtain an eigenmap ϕ : S 2m−1 → S r defined by<br />

ϕ(x, y) =(‖x‖ 2 −‖y‖ 2 , 2F (x, y)),<br />

which is called the Hopf construction. Moreover if F is full, then so is ϕ, that is, the image<br />

<strong>of</strong> ϕ is not contained in any totally geodesic submanifold <strong>of</strong> dimension r − 1inS r .<br />

Typical examples <strong>of</strong> a full orthogonal multiplication are given by the complex multiplication<br />

F : C × C ∋ (x, y) ↦→ xy ∈ C and the quaternion multiplication F : H × H ∋<br />

(p, q) ↦→ pq ∈ H. The former induces the Hopf fibration ϕ : S 3 → S 2 , and the latter induces<br />

an eigenmap ϕ : S 7 → S 4 .<br />

We shall now deduce a necessary condition for the existence <strong>of</strong> orthogonal multiplications.<br />

Assume m ≤ n. Since each orthogonal multiplication F : R m × R n −→ R r is a bilinear map,<br />

we can express it as<br />

F (x, y) = ∑ a ij x i y j ,<br />

where x =(x 1 ,... ,x m ) ∈ R m , y =(y 1 ,... ,y n ) ∈ R n and a ij ∈ R r . Since ‖F (x, y)‖ =<br />

‖x‖‖y‖, we have<br />

⎧<br />

〈a ik ,a il 〉 = δ kl ,<br />

⎪⎨<br />

(2.1.1)<br />

〈a ik ,a jk 〉 = δ ij ,<br />

⎪⎩<br />

〈a ik ,a jl 〉 +〈a il ,a jk 〉 =0 (i ≠ j, k ≠ l),<br />

which imply that for each k 0 (1 ≤ k 0 ≤ n) and i 0 (1 ≤ i 0 ≤ m), both {a ik0 } m i=1 and {a i0 k} n k=1<br />

are orthonormal systems in R r , and hence max{m, n} ≤r. Moreover, if F is full, then<br />

r ≤ mn.


2.1. EIGENMAPS 45<br />

Following a method due to M. Parker [30], who applied it to the case m = n, we now<br />

define an mn × mn-matrix G(F )by<br />

⎡<br />

⎤<br />

I n A 12 ··· A 1m<br />

A<br />

G(F ):=<br />

21 I n ··· A 2m<br />

,<br />

⎢ . . . ⎥<br />

⎣<br />

⎦<br />

A m1 A m2 ··· I n<br />

where I n denotes the n × n identity matrix and A ij are n × n-matrix whose entries are<br />

(A ij ) kl = 〈a ik ,a jl 〉, 1 ≤ k, l ≤ n.<br />

By virtue <strong>of</strong> (2.1.1), each A ij is a skew-symmetric matrix and A ji = −A ij . Note that the<br />

determinant <strong>of</strong> G(F ) coincides with Gram’s determinant with respect to the system <strong>of</strong> vectors<br />

{a ij }. Hence rank G(F )=r.<br />

We consider only the case <strong>of</strong> m =2, and prove the following existence result <strong>of</strong> orthogonal<br />

multiplications.<br />

Proposition 2.1.2. There exists a full orthogonal multiplication F : R 2 × R n → R r if and<br />

only if r is even, where n ≤ r ≤ 2n.<br />

Pro<strong>of</strong>. We first prove that rank G(F )(= r) must be even whenever a full orthogonal<br />

multiplication exists. Recall that the characteristic polynomial <strong>of</strong> G(F )is<br />

⎡<br />

⎤<br />

det(G(F ) − µI 2n ) = det ⎣ (1 − µ)I n −A<br />

⎦ ,<br />

A (1 − µ)I n<br />

where A = A 21 . Since<br />

⎡<br />

det ⎣ A<br />

B<br />

−B<br />

A<br />

⎤<br />

⎦ = ∣ ∣ det[A +<br />

√<br />

−1B]<br />

∣ ∣<br />

2<br />

,<br />

A and B being real matrices and |·|denoting the absolute value, we have<br />

det(G(F ) − µI 2n )= ∣ ∣det[(1 − µ)I n + √ −1A] ∣ ∣ 2 .


46 CHAPTER 2. EQUIVARIANT HARMONIC MAPS<br />

Consequently, the multiplicity <strong>of</strong> the zero eigenvalue <strong>of</strong> the matrix G(F ) is even. Hence r<br />

must be even.<br />

Define orthogonal multiplications F 1 : R 2 × R 2 → R 2 and F 2 : R 2 × R 2 → R 4 by<br />

⎧<br />

⎪⎨ F 1 (x, y) =(x 1 y 1 + x 2 y 2 ,x 1 y 2 − x 2 y 1 ),<br />

⎪⎩ F 2 (x, y) =(x 1 y 1 ,x 1 y 2 ,x 2 y 1 ,x 2 y 2 ),<br />

respectively, where x =(x 1 ,x 2 ) and y =(y 1 ,y 2 ). Then, they are full and satisfy<br />

‖F 1 (x, y)‖ = ‖F 2 (x, y)‖ = ‖x‖‖y‖.<br />

In the case <strong>of</strong> n =2k, we decompose R n as the direct sum <strong>of</strong> k-copies <strong>of</strong> R 2 , that is,<br />

R n = R 2 ⊕···⊕R 2 . Hence R 2 × R n =(R 2 × R 2 ) ⊕···⊕(R 2 × R 2 ) (direct sum <strong>of</strong> k-copies<br />

<strong>of</strong> R 2 × R 2 ). Then, for i (0 ≤ i ≤ k),<br />

i-copies (k − i)-copies<br />

{ }} { { }} {<br />

F = F 1 ⊕···⊕F 1 ⊕ F 2 ⊕···⊕F 2<br />

defines an orthogonal multiplication F : R 2 × R n → R 2(n−i) . Thus for even r, where n ≤<br />

r ≤ 2n, there exists an orthogonal multiplication R 2 × R 2k → R r .<br />

In the case <strong>of</strong> n =2k +1, we have the direct sum R n<br />

= R 2k ⊕ R. Associated with<br />

the orthogonal multiplication F : R 2 × R 2k → R r , we obtain an orthogonal multiplication<br />

˜F : R 2 × R 2k+1 → R r+2 defined by<br />

˜F ((x 1 ,x 2 ), (y 1 ,... ,y 2k ,y 2k+1 ))=(F ((x 1 ,x 2 ), (y 1 ,... ,y 2k )), (x 1 y 2k+1 ,x 2 y 2k+1 )),<br />

where (x 1 ,x 2 ) ∈ R 2 and (y 1 ,... ,y 2k ,y 2k+1 ) ∈ R 2k+1 . Hence for even r, where n+1 ≤ r ≤ 2n,<br />

there exists an orthogonal multiplication R 2 × R 2k+1 → R r .<br />

Now we introduce a method <strong>of</strong> constructing a new eigenmap <strong>of</strong> degree two from old<br />

known ones.<br />

Let f : S m → S p and g : S n → S q be eigen<strong>maps</strong> <strong>of</strong> degree two, and F : R m+1 × R n+1 →<br />

R r an orthogonal multiplication. Define a map ϕ : R m+1 × R n+1 → R p+q+r+2 by<br />

(2.1.2) ϕ(x, y) :=(f(x),g(y), √ 2F (x, y)).


2.1. EIGENMAPS 47<br />

Then the restriction ϕ |S m+n+1 <strong>of</strong> ϕ to S m+n+1 gives rise to an eigenmap <strong>of</strong> degree two from<br />

S m+n+1 into S p+q+r+1 . Moreover if f,g and F are full, then so is ϕ.<br />

Most significant case is that <strong>of</strong> m = p =1. In this case, f : S 1 → S 1 is given by the Hopf<br />

map, that is, f(x 1 ,x 2 )=(|x 1 | 2 −|x 2 | 2 , 2x 1 x 2 ), and Proposition 2.1.2 ensures the existence<br />

<strong>of</strong> a full orthogonal multiplication F : R 2 × R n+1 → R r for even r. Thus we obtain the<br />

following<br />

Proposition 2.1.3. Let g : S n → S q be a full eigenmap <strong>of</strong> degree two. Then we have a new<br />

full eigenmap ϕ : S n+2 → S q+r+2 <strong>of</strong> degree two, where r is even and n +1≤ r ≤ 2n +2.<br />

By making use <strong>of</strong> this proposition, we can prove the following<br />

Theorem 2.1.4. (1) (i)<br />

7 ≤ n ≤ 19.<br />

A full eigenmap ϕ : S 5 → S n <strong>of</strong> degree two exists for n =4or<br />

(ii) A full eigenmap ϕ : S 6 → S n <strong>of</strong> degree two exists for 11 ≤ n ≤ 26.<br />

(2) Let k ≥ 3.<br />

(i) A full eigenmap ϕ : S 2k+1 → S n <strong>of</strong> degree two exists for k 2 +3k −10 ≤ n ≤ 2k 2 +5k +1.<br />

(ii) A full eigenmap ϕ : S 2k+2 → S n <strong>of</strong> degree two exists for k 2 +5k − 7 ≤ n ≤ 2k 2 +7k +4.<br />

Note. The space <strong>of</strong> <strong>harmonic</strong> polynomials <strong>of</strong> degree two on R m+1 has the dimension<br />

m(m +3)/2. Thus, in Theorem 2.1.4, the upper bound <strong>of</strong> the dimension <strong>of</strong> target manifold<br />

is best possible to assure the existence <strong>of</strong> a full eigenmap <strong>of</strong> degree two.<br />

Note. Gauchman and Toth ([18]) proved that a full eigenmap <strong>of</strong> degree two ϕ : S 4 → S n<br />

exists for n =4, 7or9≤ n ≤ 13.<br />

In order to prove Theorem 2.1.4, we use the following result.<br />

Lemma 2.1.5 (Gauchman and Toth [18]). From a full eigenmap g : S m → S n <strong>of</strong> degree<br />

two, one can construct a full eigenmap ˜g : S m+1 → S m+n+2 <strong>of</strong> the same degree.


48 CHAPTER 2. EQUIVARIANT HARMONIC MAPS<br />

here.<br />

In fact, they proved an explicit formula for constructing ˜g from g, but we do not need it<br />

Pro<strong>of</strong> <strong>of</strong> Theorem 2.1.4. (1) (i) In [18], it is proved that there exist full eigen<strong>maps</strong><br />

ϕ : S 5 → S n <strong>of</strong> degree two for n =4, 7, 8, 9, 13, 15, 16, 17, 18, 19. On the other hand, it is<br />

well-known that a full eigenmap g : S 3 → S m exists if and only if m = 2 or 4 ≤ m ≤ 8.<br />

Since we have an orthogonal multiplication F : R 2 × R 4 → R 4 , it follows from Proposition<br />

2.1.3 that there exists a full eigenmap ϕ : S 5 → S n <strong>of</strong> degree two for 10 ≤ n ≤ 14.<br />

(ii) From Lemma 2.1.5 and (i) we have a full eigenmap ϕ : S 6 → S n <strong>of</strong> degree two for n =11<br />

or 14 ≤ n ≤ 26. Applying Proposition 2.1.3 to a full eigenmap g : S 4 → S 4 <strong>of</strong> degree two<br />

together with a full orthogonal multiplication F : R 2 × R 5 → R 6 , we can construct a new<br />

eigenmap ϕ : S 6 → S 12 <strong>of</strong> degree two. Moreover, by making use <strong>of</strong> the result due to Wood<br />

([44]), we can find a full orthogonal multiplication F : R 3 × R 4 → R 4 . Hence there exists<br />

a full eigenmap ϕ : S 6 → S 13 <strong>of</strong> degree two defined by ϕ(x, y) =(f(x),g(x), √ 2F (x, y)),<br />

where f : S 3 → S 4 is a full eigenmap <strong>of</strong> degree two and g : S 2 → S 4 is the Veronese map.<br />

(2) We shall proved by an induction argument.<br />

Step 1. In the case <strong>of</strong> k =3.<br />

(i) Since there exists a full eigenmap ϕ : S 6 → S n <strong>of</strong> degree two for 11 ≤ n ≤ 26,<br />

it follows from Lemma 2.1.5 that there exists a full eigenmap ϕ : S 7 → S n <strong>of</strong> degree two<br />

for 19 ≤ n ≤ 34. On the other hand, using Proposition 2.1.3, we have a full eigenmap<br />

ϕ : S 7 → S q+r+2 <strong>of</strong> degree two constructed from a full eigenmap g : S 5 → S q <strong>of</strong> degree<br />

two together with a full orthogonal multiplication F : R 2 × R 6 → R r . Then Proposition<br />

2.1.2 and (1) imply that a full eigenmap ϕ : S 7 → S n <strong>of</strong> degree two exists for n =12or<br />

14 ≤ n ≤ 33.<br />

Let F : R 4 × R 4 → R 4 be the orthogonal multiplication given by the quaternion multiplication,<br />

and let f : S 3 → S p and g : S 3 → S q be full eigen<strong>maps</strong> <strong>of</strong> degree two. Then<br />

Proposition 2.1.3 ensures the existence <strong>of</strong> a full eigenmap ϕ : S 7 → S p+q+5 . When p = q =2,<br />

we have ϕ : S 7 → S 9 , and in the case <strong>of</strong> p =2,q =4, we have ϕ : S 7 → S 11 . Finally, there


2.1. EIGENMAPS 49<br />

exists a full orthogonal multiplication F : R 4 × R 4 → R r for r =8, 11. Therefore the Hopf<br />

construction<br />

ϕ(x, y) =(‖x‖ 2 −‖y‖ 2 , 2F (x, y)) (x, y ∈ R 4 )<br />

gives rise to a full eigenmap ϕ : S 7 → S n <strong>of</strong> degree two for n =8, 11. As a consequence, we<br />

have a full eigenmap ϕ : S 7 → S n <strong>of</strong> degree two for 8 ≤ n ≤ 34. Thus (i) is true for k =3.<br />

(ii) Using Lemma 2.1.5 and the above result, we have a full eigenmap ϕ : S 8 → S n <strong>of</strong><br />

degree two for 17 ≤ n ≤ 43. Thus (ii) is also true for k =3.<br />

Step 2. (i) We assume that the statements (i) and (ii) are true up to k = l(≥ 3). Since<br />

there exists an eigenmap g : S 2l+2 → S q <strong>of</strong> degree two for l 2 +5l − 7 ≤ q ≤ 2l 2 +7l +4, it<br />

follows from Lemma 2.1.5 that we have a new eigenmap<br />

ϕ : S 2l+3 → S n for l 2 +7l − 3 ≤ n ≤ 2l 2 +9l +8.<br />

On the other hand, from our assumption, there exist a full orthogonal multiplication F :<br />

R 2 ×R 2l+2 → R r for 2l +2 ≤ r ≤ 4l +4 and a full eigenmap g : S 2l+1 → S q for l 2 +3l −10 ≤<br />

q ≤ 2l 2 +5l +1. Thus, from Proposition 2.1.3, we have a new eigenmap<br />

ϕ : S 2l+3 → S n for l 2 +5l − 6 ≤ n ≤ 2l 2 +6l +10.<br />

Since<br />

l 2 +5l − 6


50 CHAPTER 2. EQUIVARIANT HARMONIC MAPS<br />

Noting that<br />

2l 2 +10l +15< 2l 2 +11l +13,<br />

we conclude that a full eigenmap ϕ : S 2l+4 → S n exists for l 2 +7l − 1 ≤ n ≤ 2l 2 +11l +13,<br />

that is, (l +1) 2 +5(l +1)− 7 ≤ n ≤ 2(l +1) 2 +7(l +1)+4.<br />

Therefore the statement (ii) is also true for k = l +1.<br />

Note. Let ϕ 1 : S m → S n and ϕ 2 : S n → S p be eigen<strong>maps</strong> <strong>of</strong> degree two. Then the composite<br />

ϕ 2 ◦ ϕ 1 <strong>of</strong> ϕ 1 and ϕ 2 yields an eigenmap <strong>of</strong> degree four from S m to S p . Thus, by an induction<br />

argument, we can obtain an eigenmap <strong>of</strong> degree 2 k .<br />

2.2 Reduction <strong>of</strong> <strong>harmonic</strong> map equations<br />

We begin with the general situation. Let (M 1 ,g 1 ) and (M 2 ,g 2 ) be Riemannian <strong>manifolds</strong>,<br />

and I(⊂ R) an interval. For positive smooth functions f 1 and f 2 defined on I, we consider<br />

the doubly warped product Riemannian manifold defined by<br />

where t ∈ I.<br />

(M,g) =(I × M 1 × M 2 ,dt 2 + f 1 (t) 2 g 1 + f 2 (t) 2 g 2 ),<br />

Lemma 2.2.1. Let Γj i<br />

k be the Christ<strong>of</strong>fel symbols and ∆ g the Laplace-Beltrami operator <strong>of</strong><br />

(M,g). Then we have<br />

⎧<br />

Γj 1<br />

k= −f ˙ 1 (t)f 1 (t)(g 1 ) j−1,k−1 (2 ≤ j, k ≤ m 1 +1),<br />

⎪⎨<br />

Γ 1<br />

j k= − ˙ f 2 (t)f 2 (t)(g 2 ) j−m1 −1,k−m 1 −1 (m 1 +2≤ j, k ≤ m 1 + m 2 +1),<br />

⎪⎩ Γj 1<br />

k = 0<br />

(otherwise),<br />

⎧<br />

Γj i<br />

1 =Γ i f<br />

1 j = ˙ 1 (t)<br />

f 1 (t) δ ij<br />

⎪⎨<br />

(2 ≤ j ≤ m 1 +1),<br />

Γj i<br />

k = 1 Γj i<br />

k (2 ≤ j, k ≤ m 1 +1),<br />

⎪⎩<br />

Γ i<br />

j k = 0<br />

(otherwise),


2.2. REDUCTION OF HARMONIC MAP EQUATIONS 51<br />

⎧<br />

Γj i<br />

1 =Γ i f<br />

1 j = ˙ 2 (t)<br />

f 2 (t) δ ij<br />

⎪⎨<br />

(m 1 +2≤ j ≤ m 1 + m 2 +1),<br />

Γj i<br />

k = 2 Γj i<br />

k (m 1 +2≤ j, k ≤ m 1 + m 2 +1), and<br />

⎪⎩<br />

Γ i<br />

j k= 0<br />

(otherwise),<br />

(<br />

)<br />

∆ g = ∂2<br />

∂t + f˙<br />

1 (t)<br />

m 2 1<br />

f 1 (t) + m f˙<br />

2 (t) ∂<br />

2<br />

f 2 (t) ∂t + 1<br />

f 1 (t) ∆ 2 1 + 1<br />

f 2 (t) ∆ 2.<br />

2<br />

Here x 1 = t ∈ I,(x 2 ,... ,x m1 +1) ∈ M 1 , (x m1 +2,... ,x m1 +m 2 +1) ∈ M 2 ,<br />

m 1 = dimM 1 , and<br />

m 2 = dimM 2 . Also, we denote by p Γj i<br />

k (resp. ∆ p) the Christ<strong>of</strong>fel symbols (resp. the Laplace-<br />

Beltrami operator) <strong>of</strong> (M p ,g p ).<br />

Let ( ˜M 1 , ˜g 1 ) and ( ˜M 2 , ˜g 2 ) be Riemannian <strong>manifolds</strong>, Ĩ(⊂ R) an interval, and h 1 and h 2<br />

nonnegative smooth functions on Ĩ. We consider a product map u defined by<br />

(2.2.1)<br />

u :(I × M 1 × M 2 ,dt 2 + f 1 (t) 2 g 1 + f 2 (t) 2 g 2 ) ∋ (t, x, y)<br />

↦→ (r(t),ϕ(x),ψ(y)) ∈ (Ĩ × ˜M 1 × ˜M 2 ,dr 2 + h 1 (r) 2˜g 1 + h 2 (r) 2˜g 2 ),<br />

where r : I → Ĩ, ϕ : M 1 → ˜M 1 and ψ : M 2 → ˜M 2 are smooth <strong>maps</strong>. The tension field τ(u)<br />

<strong>of</strong> the map u is computed as follows.<br />

In local coordinates, x =(x i ) ∈ M,u(x) =(u α (x)) ∈ Ĩ × ˜M 1 × ˜M 2 , the components <strong>of</strong><br />

the tension field τ(u) is given by<br />

τ(u) α =∆ g u α +<br />

m 1 +m<br />

∑ 2 +1<br />

i,j=1<br />

n 1 +n<br />

∑ 2 +1<br />

β,γ=1<br />

g ij ˜Γ<br />

α<br />

β γ (u(x)) ∂uβ<br />

∂x i<br />

∂u γ<br />

∂x j<br />

.<br />

Then, from Lemma 2.2.1, we have


52 CHAPTER 2. EQUIVARIANT HARMONIC MAPS<br />

Lemma 2.2.2.<br />

⎧<br />

⎪⎨<br />

τ(u) 1<br />

(<br />

)<br />

f˙<br />

1 (t)<br />

=¨r(t)+ m 1<br />

f 1 (t) + m f˙<br />

2 (t)<br />

2 ṙ(t)<br />

f 2 (t)<br />

−2e(ϕ) h′ 1 (r(t))h 1(r(t))<br />

f 1 (t) 2<br />

⎪⎩ τ(u) α = 1<br />

f 2 (t) 2 τ(ψ)α−n 1−1<br />

− 2e(ψ) h′ 2 (r(t))h 2(r(t))<br />

f 2 (t) 2 ,<br />

τ(u) α = 1<br />

f 1 (t) 2 τ(ϕ)α−1 (2 ≤ α ≤ n 1 +1),<br />

(n 1 +2≤ α ≤ n 1 + n 2 +1),<br />

where e(ϕ) and τ(ϕ) denote the energy density function and the tension field <strong>of</strong> ϕ :(M 1 ,g 1 ) →<br />

( ˜M 1 , ˜g 1 ), and e(ψ) and τ(ψ) denote the energy density function and the tension field <strong>of</strong><br />

ψ :(M 2 ,g 2 ) → ( ˜M 2 , ˜g 2 ), respectively.<br />

Proposition 2.2.3. Let u :(I × M 1 × M 2 ,dt 2 + f 1 (t) 2 g 1 + f 2 (t) 2 g 2 ) → (Ĩ × ˜M 1 × ˜M 2 ,dr 2 +<br />

h 1 (r) 2˜g 1 + h 2 (r) 2˜g 2 ) be a map as in (2.2.1). Then u is a <strong>harmonic</strong> map if and only if the<br />

following two conditions hold:<br />

(1) r = r(t) is a solution to the following ordinary differential equation<br />

(2.2.2)<br />

(<br />

)<br />

f˙<br />

1 (t)<br />

¨r(t)+ m 1<br />

f 1 (t) + m f˙<br />

2 (t)<br />

2 ṙ(t)<br />

f 2 (t)<br />

−2e(ϕ) h′ 1 (r(t))h 1(r(t))<br />

f 1 (t) 2<br />

− 2e(ψ) h′ 2 (r(t))h 2(r(t))<br />

f 2 (t) 2 =0.<br />

(2) ϕ :(M 1 ,g 1 ) → ( ˜M 1 , ˜g 1 ) and ψ :(M 2 ,g 2 ) → ( ˜M 2 , ˜g 2 ) are <strong>harmonic</strong> <strong>maps</strong> with constant<br />

energy density, that is, ϕ and ψ are eigen<strong>maps</strong>.<br />

Now, making use <strong>of</strong> Proposition 2.2.3, we shall construct equivariant <strong>harmonic</strong> <strong>maps</strong><br />

<strong>between</strong> noncompact space forms.


2.3. CONSTRUCTIONS OF EQUIVARIANT HARMONIC MAPS 53<br />

2.3 <strong>Constructions</strong> <strong>of</strong> equivariant <strong>harmonic</strong> <strong>maps</strong><br />

2.3.1 Join parameterization<br />

We introduce the join parameterization <strong>of</strong> real hyperbolic spaces, which is an analogue <strong>of</strong><br />

that for standard unit spheres due to Smith([34]).<br />

We consider the hyperboloid model <strong>of</strong> the real hyperbolic space RH m+p+1 , that is,<br />

RH m+p+1 = {(x 0 ,x 1 , ··· ,x m+p+1 ) ∈ R m+p+2 | −x 2 0 + x 2 1 + ···+ x 2 m+p+1 = −1, x 0 > 0}.<br />

For any z ∈ RH m+p+1 there exist x ∈ RH p ,y ∈ S m and t ∈ [0, ∞) such that<br />

z = ((cosh t)x, (sinh t)y).<br />

Note that x and y are uniquely determined for t>0. Define a map f :[0, ∞)×RH p ×S m →<br />

RH m+p+1 by<br />

f(t, x, y) = ((cosh t)x, (sinh t)y).<br />

Then the pull-back metric <strong>of</strong> the standard metric ˜g on RH m+p+1 via f is given by<br />

f ∗˜g = dt 2 + (cosh 2 t)g 1 + (sinh 2 t)g 2 ,<br />

where g 1 (resp. g 2 ) is the standard metric on RH p (resp. S m ).<br />

Let ψ : RH p → RH q ,ϕ : S m → S n be eigen<strong>maps</strong>, and 2e(ψ) =µ 2 , 2e(ϕ) =ν 2 . Then<br />

the map<br />

u : RH m+p+1 ∋ (t, x, y) → (r(t),ψ(x),ϕ(y)) ∈ RH n+q+1<br />

is a <strong>harmonic</strong> map if and only if r = r(t) is a solution to the following ordinary differential<br />

equation:<br />

{<br />

¨r(t)+ p sinh t<br />

cosh t + m cosh t }<br />

ṙ(t)<br />

sinh t<br />

(2.2.3)<br />

{ }<br />

µ<br />

2<br />

−<br />

cosh 2 t +<br />

ν2<br />

sinh 2 sinh r(t) cosh r(t) =0.<br />

t<br />

In order for u to be continuous, we require that r = r(t) satisfies<br />

lim<br />

t→0<br />

r(t) =0.


54 CHAPTER 2. EQUIVARIANT HARMONIC MAPS<br />

The equation (2.2.3) is a special case <strong>of</strong> the equation (1.1.1), where q = m, f 1 (t) =<br />

cosh t, f 2 (t) = sinh t, and h 1 (r) =h 2 (r) = sinh r. Since these functions satisfy the conditions<br />

(F-1) through (F-4) and (H-1) through (H-3), respectively, a solution r = r(t) to the<br />

equation (2.2.3) exists globally. Moreover, since<br />

∫ ∞<br />

dr<br />

sinh r < ∞,<br />

we have a global, strictly monotone increasing, and unbounded solution r = r(t) to the<br />

equation (2.2.3) (see Theorem A in Chapter 1).<br />

In particular, let p = q and ψ : RH p → RH p the identity map. Then, from Theorem<br />

2.1.4, we have the following<br />

Theorem 2.3.1. There exists an equivariant <strong>harmonic</strong> map u : RH m+p+1 → RH n(m)+p+1 .<br />

Here n(m) is the integer for which an eigenmap ϕ : S m → S n(m) exists.<br />

On the other hand, when m =1, there exists a family <strong>of</strong> eigen<strong>maps</strong> S 1 ∋ z ↦→ z k ∈ S 1 ,<br />

where k ∈ Z and z ∈ C. Thus we obtain<br />

Theorem 2.3.2.<br />

itself, which is parameterized by Z.<br />

Pro<strong>of</strong>.<br />

Let m ≥ 3. Then there exists a family <strong>of</strong> <strong>harmonic</strong> <strong>maps</strong> from RH m onto<br />

Since the equation (2.2.3) has a global, strictly monotone increasing, and unbounded<br />

solution r = r(t), if we define u : RH p+2 → RH p+2 by<br />

u : RH p+2 ∋ (t, x, z) ↦→ (r(t),x,z k ) ∈ RH p+2 (x ∈ RH p ,z ∈ S 1 ),<br />

then u is a surjective <strong>harmonic</strong> map.<br />

2.3.2 Warped product <strong>manifolds</strong><br />

Let M = ([0, ∞) × S m ,dt 2 + f(t) 2 g S m) and N = ([0, ∞) × S n ,dr 2 + h(r) 2 g S n), where t, r ∈<br />

[0, ∞). Let f = f(t) and h = h(r) be smooth functions on [0, ∞) satisfying f(t) > 0 and<br />

h(r) > 0 for t>0 and r>0. If f and h satisfy<br />

f(0) = 0,<br />

˙ f(0) = 1, h(0) = 0 and h ′ (0) = 1,


2.3. CONSTRUCTIONS OF EQUIVARIANT HARMONIC MAPS 55<br />

then M and N are smooth Riemannian <strong>manifolds</strong> diffeomorphic to Euclidean space.<br />

Let ϕ : S m → S n be an eigenmap with 2e(ϕ) =µ 2 . Then a product map u(t, x) =<br />

(r(t),ϕ(x)) is a <strong>harmonic</strong> map if and only if r = r(t) is a solution to the following ordinary<br />

differential equation:<br />

f(t)<br />

(2.2.4) ¨r(t)+m ˙<br />

f(t)ṙ(t) − h(r(t))h′ (r(t))<br />

µ2 =0.<br />

f(t) 2<br />

In order for u to be continuous, we require that r = r(t) satisfies<br />

lim r(t) =0.<br />

t→0<br />

Then Theorem A and Theorem B in Chapter 1 imply the following result.<br />

Theorem 2.3.3. Let f = f(t) and h = h(r) be smooth functions on [0, ∞) and R, respectively.<br />

Assume that f and h satisfy the following conditions:<br />

⎧<br />

f(0) = 0, f(0) ˙ = 1, f(t) > 0 and f(t) ˙ ≥ 0 for t>0,<br />

⎪⎨<br />

f(t)<br />

0 ≤ m ˙<br />

∫ ∞<br />

f(t) − 1 for t ≥ 0, dt<br />

f(t) < ∞,<br />

⎪⎩ h(0) = 0, h ′ (0) = 1, h(r) > 0 for r>0 and (h 2 ) ′′ (r) ≥ 0 for r ∈ R.<br />

Then, if there exists an eigenmap ϕ : S m<br />

<strong>harmonic</strong> map<br />

→ S n , then we can construct an equivariant<br />

u : ([0, ∞) × S m ,dt 2 + f(t) 2 g S m) ∋ (t, x) ↦→ (r(t),ϕ(x)) ∈ ([0, ∞) × S n ,dr 2 + h(r) 2 g S n).<br />

<strong>maps</strong>.<br />

As applications <strong>of</strong> this theorem, we now illustrate few examples <strong>of</strong> equivariant <strong>harmonic</strong><br />

Case 1.<br />

Let f(t) = sinh t and h(r) = sinh r. Then M and N are isometric to the real<br />

hyperbolic spaces <strong>of</strong> dimension m + 1 and n +1, respectively. Since these f and h satisfy<br />

the conditions in Theorem 2.3.3 and<br />

∫ ∞<br />

dr<br />

h(r) < ∞,


56 CHAPTER 2. EQUIVARIANT HARMONIC MAPS<br />

there exists a global, strictly monotone increasing, and unbounded solution r = r(t) to the<br />

equation (2.2.4). Therefore, from Theorem 2.1.4, we obtain the following<br />

Theorem 2.3.4. (1) (i) A full equivariant <strong>harmonic</strong> map u : RH 6 → RH n exists for n =5<br />

or 8 ≤ n ≤ 20.<br />

(ii) A full equivariant <strong>harmonic</strong> map u : RH 7 → RH n exists for 12 ≤ n ≤ 27.<br />

(2) Let k ≥ 4.<br />

(i) A full equivariant <strong>harmonic</strong> map u : RH 2k → RH n exists for k 2 +k−12 ≤ n ≤ 2k 2 +k−2.<br />

(ii) A full equivariant <strong>harmonic</strong> map u : RH 2k+1 → RH n exists for k 2 +3k − 11 ≤ n ≤<br />

2k 2 +3k − 1.<br />

Note. From the eigenmap S 1 ∋ z ↦→ z k ∈ S 1 (k ∈ Z), we obtain an equivariant <strong>harmonic</strong><br />

map u : RH 2 → RH 2 . However, u is holomorphic or anti-holomorphic. Indeed, using the<br />

Poincaré disc model D 2 , it is given by D 2 ∋ z ↦→ z k ∈ D 2 .<br />

Case 2. Let f(t) = sinh t and h(r) =r. Then M and N are isometric to the real hyperbolic<br />

space and the real Euclidean space, respectively. Since these f and h satisfy the conditions<br />

in Theorem 2.3.3, there exists a global solution r = r(t) to the equation (2.2.4). Moreover,<br />

it holds that ∫ ∞<br />

dr<br />

h(r) = ∞.<br />

Thus, from Theorem 2.1.4, we obtain the following<br />

Theorem 2.3.5. There exists a family <strong>of</strong> equivariant <strong>harmonic</strong> <strong>maps</strong> from RH m+1<br />

R n(m)+1 , each <strong>of</strong> which has a bounded image.<br />

into<br />

Recently, Nagasawa and Tachikawa studied the case where<br />

∫<br />

√<br />

∞<br />

µ 2<br />

f 1 (t) +<br />

ν2<br />

dt = ∞,<br />

2 f 2 (t)<br />

2<br />

and proved, for example, the following


2.3. CONSTRUCTIONS OF EQUIVARIANT HARMONIC MAPS 57<br />

Theorem 2.3.6 ([28]). Let M =(R + × S m ,dt 2 + f(t) 2 dθ 2 ) and N =(R + × S n ,dr 2 +<br />

h(r) 2 dϕ 2 ). Assume that, in addition to the conditions (F-1) through (F-3) and (H-1) through<br />

(H-3), f and h satisfy<br />

∫ ∞<br />

∫<br />

dt<br />

∞<br />

f(t) dt = ∞ and dr<br />

dr < ∞.<br />

h(r)<br />

Then there exists no equivariant <strong>harmonic</strong> map from M to N except constant <strong>maps</strong>.<br />

As a corollary, there is no equivariant <strong>harmonic</strong> map from R m to RH n , which was<br />

independently proved by Tachikawa([35]). See also [2] and [36].<br />

In consequence, it is observed that the existence or non-existence <strong>of</strong> equivariant <strong>harmonic</strong><br />

<strong>maps</strong> is closely related to the growth orders <strong>of</strong> warping function, f and h, when t and r tend<br />

to ∞, respectively. We shall investigate a related non-existence result in the last chapter.


CHAPTER 3<br />

Dirichlet problem at infinity for <strong>harmonic</strong> <strong>maps</strong><br />

<strong>between</strong> Damek-Ricci spaces<br />

3.1 Damek-Ricci spaces<br />

3.1.1 Generalized Heisenberg algebra<br />

We start this section with a brief review <strong>of</strong> generalized Heisenberg algebras due to Kaplan<br />

[21].<br />

Let (n, 〈 , 〉 Ò<br />

) be a 2-step nilpotent Lie algebra with inner product 〈 , 〉 Ò<br />

, z its center<br />

and v the orthogonal complement <strong>of</strong> z in n with respect to 〈 , 〉 Ò<br />

. Since n is 2-step nilpotent,<br />

ad v| Ú<br />

is a map from v to z for any v ∈ v. We set k v := {u ∈ v | ad v(u) =[v, u] =0} and<br />

consider the orthogonal decomposition v = k v ⊕ v v with respect to 〈 , 〉 Ò<br />

. We call (n, 〈 , 〉 Ò<br />

)a<br />

generalized Heisenberg algebra if ad v| Úv : v v → z is a surjective isometry for any unit vector<br />

v ∈ v.<br />

Generalized Heisenberg algebras can be constructed systematically in the following fashion:<br />

Let (u, 〈 , 〉 Ù<br />

) and (v, 〈 , 〉 Ú<br />

) be real vector spaces equipped with inner product. Let<br />

µ : u × v → v be a composition <strong>of</strong> quadratic forms, that is, µ is a bilinear map satisfying<br />

|µ(u, v)| Ú<br />

= |u| Ù<br />

|v| Ú<br />

for all u ∈ u and v ∈ v. Note that µ : u × v → v satisfies µ(u 0 ,v)=v<br />

for some u 0 ∈ u. Indeed, for any given u 0 ∈ u, set T : v → v by T (v) :=µ(u 0 ,v). Then<br />

µ ′ (u, v) :=µ(u, T −1 (v)) satisfies µ ′ (u 0 ,v)=v. Hence we may suppose µ(u 0 ,v)=v. Define<br />

φ : v × v → u by<br />

〈u, φ(v, v ′ )〉 Ù<br />

= 〈µ(u, v),v ′ 〉 Ú<br />

.<br />

Let z be the orthogonal complement <strong>of</strong> Ru 0 = {ru 0 | r ∈ R} in u, π: u → z the orthogonal<br />

projection and n the direct sum n := v ⊕ z <strong>of</strong> v and z with the natural inner product. Define<br />

59


60 CHAPTER 3. DAMEK-RICCI SPACES<br />

a Lie bracket on n by<br />

(3.1.1) [v + z,v ′ + z ′ ]:=π ◦ φ(v, v ′ ).<br />

Theorem 3.1.1 ([21]). n, equipped with the above inner product and a Lie bracket, is a<br />

generalized Heisenberg algebra. Conversely, any generalized Heisenberg algebra arises in this<br />

manner.<br />

Regarding the existence <strong>of</strong> a composition <strong>of</strong> quadratic forms, we remark the following.<br />

Let ρ be a function defined on positive integers by<br />

ρ(n) :=8p +2 q if n = (odd number) × 2 4p+q , 0 ≤ q ≤ 3.<br />

Then it is known by Hurwitz, Radon and Eckmann that a composition <strong>of</strong> quadratic forms<br />

µ : u × v → v exists if and only if 0 < dim u ≤ ρ(dim v). Therefore we may conclude (see<br />

[21]):<br />

Fact. Given integers m, n > 0, there exists an (n + m)-dimensional generalized Heisenberg<br />

algebra with m-dimensional center if and only if 0


3.1. DAMEK-RICCI SPACES 61<br />

where [ , ] Ò<br />

is the Lie bracket on n defined by (3.1.1). In this way, s becomes a Lie algebra<br />

with an inner product. The simply connected Lie group S associated with s, equipped with<br />

the left invariant metric g S , is called a Damek-Ricci space.<br />

Example. Let v = R 2k , u = R 2 , and µ(z,v) =(−y, x),µ(w, v) =(x, y), where {z,w} is<br />

an orthonormal basis <strong>of</strong> R 2 and v =(x, y) ∈ R k ⊕ R k . Then n is a (classical) Heisenberg<br />

algebra and (S, g S ) is isometric to the complex hyperbolic space whose sectional curvature<br />

K satisfies −4 ≤ K ≤−1. The quaternion hyperbolic space and the Cayley hyperbolic<br />

plane are also Damek-Ricci spaces. These three spaces are called classical (cf. [8]). Thus<br />

Damek-Ricci spaces are generalizations <strong>of</strong> rank one symmetric spaces <strong>of</strong> noncompact type.<br />

Let ∇ be the Levi-Civita connection on (S, g S ). Using the formula<br />

2〈∇ X Y,Z〉 = 〈[X, Y ],Z〉 + 〈[Z, X],Y〉 + 〈[Z, Y ],X〉<br />

for X, Y, Z ∈ s, one obtains<br />

⎧<br />

∇ h X =0, ∇ X h = −[h, X],<br />

⎪⎨ ∇ v v ′ = 〈v, v ′ 〉h + 1 2 [v, v′ ],<br />

∇ v z = ∇ z v = −µ(z,v),<br />

⎪⎩<br />

∇ z z ′ =2〈z,z ′ 〉h,<br />

where X ∈ s,v,v ′ ∈ v,z,z ′ ∈ z and µ is the composition <strong>of</strong> quadratic forms in the definition<br />

<strong>of</strong> n. Then the following results have been proved.<br />

Theorem 3.1.2.<br />

([7], [8]).<br />

(1) A Damek-Ricci space is a symmetric space if and only if it is classical<br />

(2) A Damek-Ricci space has strictly negative sectional curvature if and only if it is classical<br />

([13]).<br />

In fact, Damek constructed a simple example <strong>of</strong> Damek-Ricci space whose sectional<br />

curvature attains zero for some two-dimensional plane. Other examples have been given


62 CHAPTER 3. DAMEK-RICCI SPACES<br />

in [5]. The second assertion was first proved by Boggino [4]. However, his pro<strong>of</strong> needed a<br />

refinement which had been accomplished by Dotti([13]), see also [23]. The above theorem<br />

claims that every non-symmetric Damek-Ricci space has two-dimensional planes for which<br />

the sectional curvature vanishes.<br />

We refer to [5] for further information about the geometry, in particular about the sectional<br />

curvature, <strong>of</strong> the Damek-Ricci spaces.<br />

Finally, we estimate the bottom spectrum λ 1 (S) <strong>of</strong> the Laplace-Beltrami operator for S<br />

with the left invariant metric.<br />

Lemma 3.1.3 ([9]). Let (r, ω) be the polar coordinate on (S, g S ) around an origin. Then<br />

the volume form dv S <strong>of</strong> S is given by<br />

dv S =2 −m (sinh r) n (sinh 2r) m drdσ(ω),<br />

where n = dim v,m = dim z and dσ(ω) is the surface element <strong>of</strong> the standard unit sphere<br />

S n+m .<br />

From this lemma and the fact λ 1 (S) ≥ inf(∆ S r) 2 /4, where ∆ S<br />

Laplace-Beltrami operator, we obtain the following<br />

denotes the positive<br />

Corollary 3.1.4.<br />

λ 1 (S) ≥ 1 4 (n +2m)2 .<br />

Note. Damek and Ricci [10] proved that the equality holds in the above inequality.<br />

3.2 Harmonic <strong>maps</strong> <strong>between</strong> Damek-Ricci spaces<br />

In this section, following Donnelly [12], we shall prove the existence and uniqueness <strong>of</strong><br />

solutions <strong>of</strong> the Dirichlet problem at infinity for <strong>harmonic</strong> <strong>maps</strong>.


3.2. HARMONIC MAPS BETWEEN DAMEK-RICCI SPACES 63<br />

3.2.1 Harmonic map equation<br />

We compute explicitly the tension field <strong>of</strong> a <strong>harmonic</strong> map <strong>between</strong> Damek-Ricci spaces. Let<br />

(S, g S ) be a Damek-Ricci space with left invariant metric g S , and set s = a ⊕ n = a ⊕ v ⊕ z.<br />

Take the unit vector h ∈ a such that ad h(v) =v, ad h(z) =2z for any v ∈ v,z ∈ z and<br />

an orthonormal basis {h, v 1 ,... ,v n ,z 1 ,... ,z m } <strong>of</strong> s, {v i } (resp. {z j }) being an orthonormal<br />

basis <strong>of</strong> v (resp. z). We define a map ϕ : R × N → S by<br />

ϕ(t, n) :=n(exp(th)),<br />

where N is the simply connected Lie group associated with n and exp is the exponential<br />

map on a. Then the induced metric ϕ ∗ g S is given by<br />

ϕ ∗ g S = dt 2 + e −2t g Ú<br />

+ e −4t g Þ<br />

,<br />

where g Ú<br />

+ g Þ<br />

is a left invariant metric on N. Setting y = e t , one can verify that (S, g S )is<br />

isometric to M := R + × N with the Riemannian metric<br />

g M := y −2 dy 2 + y −2 g Ú<br />

+ y −4 g Þ<br />

,<br />

where R + = {r ∈ R | r>0}. A straightforward calculation then yields that the Levi-Civita<br />

connection ∇ on (S, g S ) is given by<br />

⎧<br />

∇ η η = −y −1 η, ∇ η v i = −y −1 v i , ∇ η z k = −2y −1 z k ,<br />

(3.2.1)<br />

⎪⎨<br />

⎪⎩<br />

∇ vi v j = y −1 δ ij η + 1 2<br />

∇ vi z k = 1 n∑<br />

2 y−2<br />

j=1<br />

m∑<br />

k=1<br />

a k<br />

ijz k ,<br />

where a k<br />

ij are the structure constants, [v i ,v j ]=<br />

regarded as left invariant vector fields on S.<br />

a k<br />

j iv j , ∇ zk z l =2y −3 δ kl η,<br />

m∑<br />

k=1<br />

a k<br />

ijz k ,η= ∂/∂y and, v i and z k<br />

Now we compute the tension field. Let (S, g S ) and (S ′ ,g S ′) be Damek-Ricci spaces and<br />

u ∈ C 2 (S, S ′ ). We represent (S, g S ) (resp. (S ′ ,g S ′)) as<br />

(R + × N,y −2 dy 2 + y −2 g Ú<br />

+ y −4 g Þ<br />

) (resp. (R + × N ′ , ȳ −2 dȳ 2 +ȳ −2 g Ú ′ +ȳ −4 g Þ<br />

′)),<br />

are


64 CHAPTER 3. DAMEK-RICCI SPACES<br />

where y, ȳ ∈ R + and n = v + z (resp. n ′ = v ′ + z ′ ) with dim v = n (resp. dim v ′ = n ′ ),<br />

dim z = m (resp. dim z ′ = m ′ ). Take an orthonormal basis {v 1 ,... ,v n ,z 1 ,... ,z m } (resp.<br />

{¯v 1 ,... ,¯v n ′, ¯z 1 ,... ,¯z m ′}) <strong>of</strong>n (resp. n ′ ) as above, and denote their left invariant extensions<br />

on S by the same letters. Let<br />

⎧<br />

e 0 = ∂ ⎧<br />

∂y ,<br />

f 0 = ∂ ∂ȳ ,<br />

⎪⎨<br />

⎪⎨<br />

e i = v i (1 ≤ i ≤ n),<br />

f α =¯v α (1 ≤ α ≤ n ′ ),<br />

⎪⎩ e i = z i−n (n +1≤ i ≤ n + m),<br />

⎪⎩ f α =¯z α−n ′ (n ′ +1≤ α ≤ n ′ + m ′ ),<br />

and<br />

u α j = f ∗ α (du(e i)), u α ij = e j · u α i , τα (u) =f ∗ α (τ(u)),<br />

where f ∗ α is the dual frame <strong>of</strong> f α . Then, since the tension field <strong>of</strong> u is given by<br />

we get<br />

(3.2.2)<br />

⎧<br />

⎪⎨<br />

⎪⎩<br />

τ 0 (u) =<br />

τ α (u) =<br />

τ α (u) =<br />

τ(u) :=<br />

n+m<br />

∑<br />

i=0<br />

n+m<br />

∑<br />

i=0<br />

n∑<br />

( ˜∇ ei du(e i ) − du(∇ ei e i ))=0,<br />

i=1<br />

n+m<br />

∑<br />

g ii u 0 ii +(1− n − 2m)u 0 0y − (u 0 ) −1<br />

n+m<br />

∑<br />

+(u 0 ) −1<br />

i=0<br />

g ii n ′<br />

i=0<br />

∑<br />

n+m<br />

∑<br />

(u β i )2 +2(u 0 ) −3<br />

β=1<br />

i=0<br />

g ii (u 0 i ) 2<br />

∑<br />

g ii n ′ +m ′<br />

β=n ′ +1<br />

n+m<br />

g ii u α ii +(1− n − 2m)uα 0 y − ∑<br />

2(u0 ) −1 g ii u 0 i uα i<br />

n+m<br />

∑ ∑n ′<br />

+(u 0 ) −2 g ii<br />

n+m<br />

∑<br />

i=0<br />

i=0<br />

n<br />

∑<br />

′ +m ′<br />

β=1 γ=n ′ +1<br />

i=0<br />

(u β i )2 ,<br />

u γ i uβ i Γ γ−n′<br />

αβ<br />

(1 ≤ α ≤ n ′ ),<br />

n+m<br />

g ii u α ii +(1− n − 2m)u0 0 y − ∑<br />

4(u0 ) −1 g ii u 0 i uα i<br />

i=0<br />

(n ′ +1≤ α ≤ n ′ + m ′ ),<br />

where u 0 =ȳ(u), and (g ij ) denotes the matrix component <strong>of</strong> the metric g M , (g ij ) its inverse<br />

matrix. Note that u α ij ≠ uα ji because [v i,v j ] ≠0.


3.2. HARMONIC MAPS BETWEEN DAMEK-RICCI SPACES 65<br />

3.2.2 Uniqueness theorem<br />

Let S and S ′ be Damek-Ricci spaces and assume that they are represented as in Section<br />

3.2.1. We denote by ¯S (resp. ¯S ′ ) the Eberlein-O’Neill compactification <strong>of</strong> S (resp. S ′ ). Then<br />

{y =0}×N (resp. {ȳ =0}×N ′ ) represents the ideal boundary, ∂S (resp. ∂S ′ ), except a<br />

point in ∂S (resp. ∂S ′ ) (cf. [6]).<br />

Let u ∈ C 2 (S, S ′ ) be a proper <strong>harmonic</strong> map. First, we investigate a necessary condition<br />

for the existence <strong>of</strong> a C 2 -extension u : ¯S → ¯S ′ <strong>of</strong> u, and then prove the following uniqueness<br />

theorem.<br />

Theorem 3.2.1 (Uniqueness theorem).<br />

Let u and w be proper <strong>harmonic</strong> <strong>maps</strong> <strong>between</strong><br />

Damek-Ricci spaces S and S ′ . Suppose u, w ∈ C 2 ( ¯S, ¯S ′ ) and f := u |∂S = w |∂S . If<br />

then u = w on ¯S.<br />

n+m<br />

∑<br />

n<br />

∑<br />

′ +m ′<br />

j=n+1 γ=n ′ +1<br />

(f γ j )2 > 0 on ∂S,<br />

In order to prove this theorem, the following lemma plays an important role.<br />

Lemma 3.2.2 ([12], [26]). Suppose that ω ∈ C 1 Λ 1 S ∩ C 0 Λ 1 ¯S is a 1-form defined on a<br />

n+m<br />

∑<br />

neighborhood <strong>of</strong> p ∈ ∂S. Let ω = ω i e ∗ i , where {e ∗ i } is the dual c<strong>of</strong>rame <strong>of</strong> {e i }. Then<br />

there exists a sequence <strong>of</strong> points {q k }⊂S such that q k → p and<br />

(<br />

)<br />

∑<br />

g jj ω jj (q k )=0.<br />

lim<br />

k→∞<br />

i=0<br />

y −1 n+m<br />

j=0<br />

Using this lemma together with τ(u) =0, we obtain the following necessary condition.<br />

Lemma 3.2.3. Let u ∈ C 2 ( ¯S, ¯S ′ ) be a proper <strong>harmonic</strong> map. Then at the ideal boundary<br />

we have the following:<br />

(1)<br />

n∑<br />

n<br />

∑<br />

′ +m ′<br />

j=0 β=n ′ +1<br />

(u β j )2 =0,


66 CHAPTER 3. DAMEK-RICCI SPACES<br />

(2) (n +2m)(u 0 0 )4 −<br />

n∑ ∑n ′<br />

(u β j )2 (u 0 0 )2 − 2<br />

j=0 β=1<br />

(3) (1 + n +2m)u α 0 (u 0 0) 2 −<br />

n∑ ∑n ′<br />

n<br />

∑<br />

′ +m ′<br />

j=0 β=1 γ=n ′ +1<br />

n∑<br />

n<br />

∑<br />

′ +m ′<br />

j=0 β=n ′ +1<br />

(4) (2 + n +2m)u 0 0u α 00 =0 (n ′ +1≤ α ≤ n ′ + m ′ ).<br />

(u β j0 )2 − 2<br />

n+m<br />

∑<br />

n<br />

∑<br />

′ +m ′<br />

j=n+1 β=n ′ +1<br />

Γ γ−n′<br />

α β<br />

u β j uγ j0 =0 (1≤ α ≤ n′ ),<br />

(u β j )2 =0,<br />

Pro<strong>of</strong>. Since u ∈ C 2 ( ¯S, ¯S ′ ) is proper, we have u 0 = O(y) and lim y→0 u 0 y −1 = u 0 0.<br />

Multiplying by (u 0 ) 3 y −2 both sides <strong>of</strong> the first equation in (3.2.2), we let y → 0. Then, by<br />

virtue <strong>of</strong> Lemma 3.2.2, the first term on the right hand side tends to 0. Hence we obtain (1).<br />

The rest <strong>of</strong> the statement follows similarly if we multiply the first, second and third<br />

equations in (3.2.2) by (u 0 ) 3 y −4 , (u 0 ) 2 y −3 and u 0 y −3 , respectively.<br />

Since u γ 0j =0(γ ≥ n′ + 1) at the boundary and ddu =0, we have<br />

∑n ′<br />

(3.2.3) u γ j0 = −<br />

µ,ν=1<br />

Γ γ−n′<br />

µν u µ 0u ν j (0 ≤ j ≤ n, n ′ +1≤ γ ≤ n ′ + m ′ ).<br />

Substituting this into the equation (3) <strong>of</strong> Lemma 3.2.3 and adding the result in α, we get<br />

∑n ′<br />

(3.2.4) (1 + n +2m)(u 0 0 )2 (u α 0 )2 +<br />

α=1<br />

Now we can deduce the following<br />

n∑<br />

n<br />

∑<br />

′ +m ′<br />

j=0 γ=n ′ +1<br />

( n ′<br />

∑<br />

α,β=1<br />

Γ γ−n′<br />

αβ<br />

u α 0 uβ j<br />

) 2<br />

=0.<br />

Proposition 3.2.4. Let u ∈ C 2 ( ¯S, ¯S ′ ) be a proper <strong>harmonic</strong> map. If f := u |∂S satisfies<br />

n+m<br />

∑<br />

n<br />

∑<br />

′ +m ′<br />

j=n+1 γ=n ′ +1<br />

(f γ j )2 > 0,<br />

then at the ideal boundary u must satisfy u 0 0 > 0,uα 0 =0(1≤ α ≤ n′ + m ′ ) and u β k0 = uβ 00 =<br />

0(1≤ k ≤ n, n ′ +1≤ β ≤ n ′ + m ′ ).


3.2. HARMONIC MAPS BETWEEN DAMEK-RICCI SPACES 67<br />

Pro<strong>of</strong>. The assumption on f implies that the fourth term in the equation (2) <strong>of</strong> Lemma<br />

3.2.3 never vanishes. Thus u 0 0 > 0, which implies<br />

u α 0 =0, u α 00 = 0 and<br />

∑n ′<br />

α,β=1<br />

Γ γ−n′<br />

αβ<br />

u α j u β j =0<br />

from the equation (4) <strong>of</strong> Lemma 3.2.3 and (3.2.4). Finally, (3.2.3) implies u β k0 =0.<br />

When S and S ′ are real hyperbolic planes, this proposition means that a proper <strong>harmonic</strong><br />

map must be conformal at the ideal boundary.<br />

By Proposition 3.2.4 we get the following<br />

Corollary 3.2.5. We have<br />

⎧<br />

⎪ ⎨<br />

⎪ ⎩<br />

u α 0 = O(y) (1 ≤ α ≤ n′ ),<br />

u α 0 = o(y) (n ′ +1≤ α ≤ n ′ + m ′ ).<br />

We shall now complete the pro<strong>of</strong> <strong>of</strong> Theorem 3.2.1.<br />

If we write<br />

u(y, n) =(ȳ(u), ¯n(u)), w(y, n) =(ȳ(w), ¯n(w)), f(n) =¯n(f),<br />

then it holds that<br />

(3.2.5)<br />

d(u, w) ≤ d((ȳ(u), ¯n(u)), (ȳ(u), ¯n(f)))<br />

+d((ȳ(u), ¯n(f)), (ȳ(w), ¯n(f))) + d((ȳ(w), ¯n(f)), (ȳ(w), ¯n(w))).<br />

From the explicit expression for the metrics and ¯n(f) =¯n(u(0,n)), we see that the first term<br />

on the right hand side <strong>of</strong> (3.2.5) is<br />

∫ y<br />

∣ ∫ ∂¯n ∣∣∣ y<br />

∣<br />

0 ∂t (u(t, n)) ∑n dt ≤<br />

[t ′<br />

n<br />

∑<br />

′ +m ′<br />

−1 |u α 0 | + t −2<br />

0<br />

α=1<br />

α=n ′ +1<br />

]<br />

|u α 0 | dt = o(1).<br />

The last estimate in the above follows from Corollary 3.2.5. Similarly, the third term is O(y).<br />

On the other hand, the second term on the right hand side <strong>of</strong> (3.2.5) is<br />

∣ ȳ(u)<br />

∣∣∣<br />

∣log ȳ(w) ∣ = log u0 (y, n)<br />

w 0 (y, n) ∣ .


68 CHAPTER 3. DAMEK-RICCI SPACES<br />

Since u 0 0 ((0,n)) and w0 0 (w(0,n)) are uniquely determined by f and positive, we have<br />

∣ ȳ(u)<br />

∣∣∣<br />

∣log ȳ(w) ∣ = log u0 0(0,n)y + o(y)<br />

w0(0,n)y 0 + o(y) ∣ = o(1).<br />

Therefore d(u, w) = 0 at the ideal boundary. Hence the maximal principle ([33]) implies<br />

Theorem 3.2.1.<br />

3.2.3 Existence theorem<br />

To show the existence <strong>of</strong> a solution <strong>of</strong> the Dirichlet problem at infinity, we use the heat flow<br />

method due to Li and Tam [25]. As we see in the previous section, Damek-Ricci spaces are<br />

homogeneous spaces <strong>of</strong> nonpositive sectional curvature and have positive bottom spectrum<br />

<strong>of</strong> the Laplace-Beltrami operator. Therefore, we can apply the general existence theory [25,<br />

Theorem 5.2] if there exists a suitable initial map. Indeed, h in Proposition 3.2.7 can be<br />

taken as our initial map.<br />

The decay order <strong>of</strong> the tension field <strong>of</strong> the initial map can be estimated by the following<br />

Lemma 3.2.6. Let h ∈ C 2,ε (∂S,∂S ′ )(0


3.2. HARMONIC MAPS BETWEEN DAMEK-RICCI SPACES 69<br />

Then there exists an h ∈ C 2,ε ( ¯S, ¯S ′ ) such that h = f at the ideal boundary and ‖τ(h)‖ =<br />

O(y ε ), where ‖τ(h)‖ is the norm <strong>of</strong> the tension field in the Riemannian metric.<br />

Pro<strong>of</strong>.<br />

Let φ>0 be a unique solution <strong>of</strong><br />

(n +2m)φ 4 −<br />

n∑ ∑n ′<br />

(f β j )2 φ 2 − 2<br />

j=1 β=1<br />

n+m<br />

∑<br />

n<br />

∑<br />

′ +m ′<br />

j=n+1 β=n ′ +1<br />

(f β j )2 =0.<br />

Since f ∈ C 3,ε , we have φ ∈ C 2,ε (∂S). Set h(y, n) :=(yφ(n),f(n)). Then h ∈ C 2,ε (∂S,∂S ′ ).<br />

Moreover, it is easy to verify that h satisfies<br />

⎧<br />

h 0 0 = φ,<br />

⎪⎨ h α 0 =0 (1≤ α ≤ n ′ + m ′ ),<br />

h β 00 =0 (n ′ +1≤ β ≤ n ′ + m ′ ),<br />

⎪⎩ h γ j0 =0 (1≤ j ≤ n, n′ +1≤ γ ≤ n ′ + m ′ ),<br />

and h = f at the ideal boundary. Lemma 3.2.6 and the explicit expression for the metric<br />

then imply ‖τ(h)‖ = O(y ε ).<br />

Next we construct a comparison function.<br />

Lemma 3.2.8. For sufficiently large r 0 and some constant s, define<br />

⎧<br />

⎪⎨ e −sr 0<br />

(r ≤ r 0 ),<br />

ψ(r) :=<br />

⎪⎩ e −sr (r ≥ r 0 ).<br />

If 0


70 CHAPTER 3. DAMEK-RICCI SPACES<br />

Theorem 3.2.9 (Existence Theorem). Suppose that f ∈ C 3,ε (∂S,∂S ′ )(0


3.2. HARMONIC MAPS BETWEEN DAMEK-RICCI SPACES 71<br />

By virtue <strong>of</strong> [25, Lemma 5.1], we have<br />

‖v t (x, t)‖ ≤c 1 e −c 2t<br />

for some constants c 1 ,c 2 > 0. Then for any T>0<br />

d S ′(u(x),h(x)) ≤<br />

∫ ∞<br />

‖v t (x, t)‖dt ≤<br />

≤ ce −εr(x) T + c 1 e −c2T .<br />

Choosing T = εr, we get<br />

∫ T<br />

‖v t (x, t)‖dt +<br />

∫ ∞<br />

0<br />

0<br />

T<br />

‖v t (x, t)‖dt<br />

ce −εr(x) T and c 1 e −c 2T → 0<br />

as r →∞.<br />

Therefore, d S ′(u(x),h(x)) → 0asx → ∂S and u = h = f at the ideal boundary.<br />

Note. Our result remains true, with necessary modifications, when the target manifold is<br />

replaced by the real hyperbolic space.


CHAPTER 4<br />

Non-existence <strong>of</strong> proper <strong>harmonic</strong> <strong>maps</strong> from complex<br />

hyperbolic spaces into real hyperbolic spaces<br />

In this chapter, we investigate proper <strong>harmonic</strong> <strong>maps</strong> from the m-dimensional complex<br />

hyperbolic space B m to the n-dimensional real hyperbolic space D n . Our goal is to prove<br />

the following non-existence result.<br />

Theorem 4.1.0. Let m, n ≥ 2. Then there is no proper <strong>harmonic</strong> map u ∈ C 2 (B m , D n )<br />

which has C 1 -regularity up to the ideal boundary.<br />

The C 1 -regularity assumption up to the ideal boundary, supposed for proper <strong>harmonic</strong><br />

<strong>maps</strong> in Theorem 4.1.0, plays a crucial role in our pro<strong>of</strong>. Indeed, in the last section, we show<br />

that there exists a counter example to this theorem if we relax its regularity condition up to<br />

the ideal boundary.<br />

4.1 Pro<strong>of</strong> <strong>of</strong> Theorem<br />

Let (B m ,g) and (D n ,g ′ ) be the ball models <strong>of</strong> the m-dimensional complex hyperbolic space<br />

and the n-dimensional real hyperbolic space, respectively. Namely,<br />

(<br />

m<br />

)<br />

B m = {z ∈ C m 1 ∑ {<br />

||z| < 1}, g=<br />

(1 −|z| 2 )δ<br />

(1 −|z| 2 ) 2 ij +¯z i z j} dz i d¯z j ,<br />

(<br />

D n = {x ∈ R n ||x| < 1}, g ′ =<br />

i,j=1<br />

)<br />

4<br />

n∑<br />

(dx i ) 2 ,<br />

(1 −|x| 2 ) 2<br />

where z =(z 1 ,... ,z m ) ∈ B m and x = (x 1 ,... ,x n ) ∈ D n are the canonical Euclidean<br />

coordinates. We endow the Eberlein-O’Neill compactification B m (resp. D n ) with the nat-<br />

73<br />

i=1


74 CHAPTER 4. NON-EXISTENCE OF PROPER HARMONIC MAPS<br />

ural smooth structure inherited from C m (resp. R n ), and regard it as a submanifold with<br />

boundary in C m (resp. R n ). With these understood, we have the following<br />

Lemma 4.1.1. (1) Let g i¯j , Γ i<br />

j k and ∆ B m<br />

be the components, the Christ<strong>of</strong>fel symbols and<br />

the Laplace-Beltrami operator <strong>of</strong> Bergman metric g, respectively. Then we have<br />

⎧<br />

g i¯j =(1−|z| 2 ) −2 {(1 −|z| 2 )δ ij +¯z i z j },<br />

⎪⎨<br />

Γ i<br />

j k =(1−|z|2 ) −1 (¯z j δ ik +¯z k δ ij ),<br />

⎪⎩<br />

∆ B m =(1−|z| 2 )<br />

m∑<br />

(δ ij − z i¯z j ∂ 2<br />

)<br />

∂z i ∂¯z . j<br />

i,j=1<br />

(2) Let g αβ ′ , α ˜Γ β γ and ∆ D n be the components, the Christ<strong>of</strong>fel symbols and the Laplace-<br />

Beltrami operator <strong>of</strong> Poincaré metric g ′ , respectively. Then we have<br />

⎧<br />

g αβ ′ = 1<br />

(1 −|x| 2 ) δ αβ,<br />

2<br />

⎪⎨<br />

˜Γ α<br />

β γ = 2<br />

(1 −|x| 2 ) (xβ δ αγ + x γ δ αβ − x α δ βγ ),<br />

⎪⎩ ∆ D n =(1−|x| 2 ) 2<br />

n∑<br />

α=1<br />

∂ 2<br />

∂x α ∂x α +2(n − 2)(1 −|x|2 )<br />

n∑<br />

α=1<br />

x α<br />

∂<br />

∂x . α<br />

Following [24], we define a family <strong>of</strong> global vector fields N and X j , 1 ≤ j ≤ m, on C m<br />

in the following manner:<br />

⎧<br />

N = ⎪⎨<br />

⎪⎩ X j =<br />

m∑<br />

i=1<br />

m∑<br />

i=1<br />

z i<br />

∂<br />

∂z , i<br />

(δ ij − z i¯z j ) ∂<br />

∂z i =<br />

∂<br />

∂z −〈 ∂<br />

j ∂z ,N〉N,<br />

j<br />

where 〈·, ·〉 denotes the standard Hermitian inner product on C m . Then we get<br />

Lemma 4.1.2 ([24]). (1) N + ¯N is the outer normal vector field on ∂B m .<br />

(2) {X j + ¯X j , √ −1(X j − ¯X j ), √ −1(N − ¯N)} m j=1 is a basis <strong>of</strong> the tangent space T p∂B m at<br />

each p ∈ ∂B m .


4.1. PROOF OF THEOREM 75<br />

Moreover, we obtain the following<br />

m∑<br />

Lemma 4.1.3 ([24]). Let L = (δ ij − z i¯z j ∂ 2<br />

)<br />

∂z i ∂¯z . Then j<br />

L =<br />

=<br />

i,j=1<br />

m∑<br />

X j ¯Xj +(m −|z| 2 ) ¯N +(1−|z| 2 )N ¯N<br />

j=1<br />

m∑<br />

j=1<br />

¯X j X j +(m −|z| 2 )N +(1−|z| 2 ) ¯NN.<br />

In terms <strong>of</strong> our frame fields, the tension field <strong>of</strong> u is expressed as follows.<br />

Lemma 4.1.4. For u ∈ C 2 (B m , D n ) we have<br />

[<br />

]<br />

τ(u) α =(1−|z| 2 ) Lu α 2<br />

+<br />

a(u)(z) {(Nuα )〈 ¯Nu,u〉 +(¯Nu α )〈Nu,u〉−u α |Nu| 2 }<br />

+<br />

2<br />

a(u)(z)<br />

m∑<br />

{(X j u α )〈 ¯X j u, u〉 +(¯X j u α )〈X j u, u〉−u α |X j u| 2 },<br />

j=1<br />

where a(u)(z) =(1−|u(z)| 2 )/(1 −|z| 2 ) and u(z) =(u 1 (z),... ,u n (z)).<br />

Pro<strong>of</strong>.<br />

In our coordinates, the tension field <strong>of</strong> u ∈ C 2 (B m , D n ) is given by<br />

m∑ n∑<br />

τ(u) α =∆ B mu α + g i¯j˜Γ α<br />

β γ (u(z))u β i uγ¯j .<br />

Hence, from Lemma 4.1.1, we have<br />

m∑<br />

τ(u) α =(1−|z| 2 ) (δ ij − z i¯z j ) ∂2 u α<br />

∂z i ∂¯z j<br />

i,j=1<br />

+ 2(1 −|z| 2 )(1 −|u(z)| 2 ) −1 m<br />

∑<br />

=(1−|z| 2 )Lu α +2a(u)(z) −1<br />

=(1−|z| 2 )Lu α<br />

+2a(u)(z) −1<br />

i,j=1 β,γ=1<br />

i,j=1 β,γ=1<br />

n∑<br />

β=1 j=1<br />

{<br />

∑ m<br />

+ (δ ij − z i¯z j )u β i<br />

n∑<br />

i=1<br />

β=1 j=1<br />

n∑<br />

(δ ij − z i¯z j )(u β δ αγ + u γ δ αβ + u α δ βγ )u γ i uβ¯j<br />

[<br />

m∑ {<br />

∑ m<br />

(δ ij − z i¯z j )u α i<br />

i=1<br />

}<br />

u ᾱ j u β +<br />

}<br />

u β u β¯j<br />

{<br />

∑ m<br />

(δ ij − z i¯z j )u β i<br />

i=1<br />

} ]<br />

u α u β¯j<br />

m∑<br />

{(X j u α )u β u β¯j +(X j u β )u ᾱ j u β +(X j u β )u α u β¯j } =(∗1).


76 CHAPTER 4. NON-EXISTENCE OF PROPER HARMONIC MAPS<br />

From the formula ∂/∂¯z j = ¯X j + z j ¯N, we then have<br />

(∗1) = (1 −|z| 2 )Lu α<br />

+2a(u)(z) −1<br />

=(1−|z| 2 )Lu α<br />

n∑<br />

β=1 j=1<br />

+(1−|z| 2 2<br />

)<br />

a(u)(z)<br />

+<br />

2<br />

a(u)(z)<br />

m∑<br />

j=1 β=1<br />

m∑<br />

{(X j u α )u β ( ¯X j u β + z j ¯Nu β )<br />

+(X j u β )u β ( ¯X j u α + z j ¯Nu α )+(X j u β )u α ( ¯X j u β + z j ¯Nu β )}<br />

n∑<br />

{(Nu α )(u β ¯Nu β )+(¯Nu α )(u β Nu β ) − u α (Nu β )( ¯Nu β )}<br />

β=1<br />

n∑<br />

{(X j u α )(u β ¯Xj u β )+(¯X j u α )(u β X j u β ) − u α (X j u β )( ¯X j u β )}.<br />

In order to investigate the asymptotic behavior <strong>of</strong> proper <strong>harmonic</strong> <strong>maps</strong>, Li and Ni<br />

proved the following<br />

Lemma 4.1.5 ([24]). Assume f ∈ C 2 (B m ) ∩ C 1 (B m ). Then we have<br />

1 −|z| 2<br />

lim<br />

(<br />

z→∂B m ε(z)<br />

∫B(z,ε(z))<br />

¯fLf)(ζ)dζ =0,<br />

2m<br />

where ε(z) =(1−|z|)/2 and B(z,ε(z)) is the open ball in (C m ,g C m) with the radius ε(z)<br />

centered at z, and dζ is the volume element <strong>of</strong> (C m ,g C m).<br />

pro<strong>of</strong>.<br />

As a corollary <strong>of</strong> this lemma, we have the following result, which is in essential use in our<br />

Corollary 4.1.6. Assume u ∈ C 2 (B m , C n ) ∩ C 1 (B m , C n ). Then for any point p ∈ ∂B m<br />

there exists a sequence {z j } ∞ j=1 ⊂ B m satisfying the following properties:<br />

(1) z j → p (j →∞).<br />

(2) lim<br />

j→∞<br />

(1 −|z j | 2 )〈Lu, u〉(z j )=0.<br />

We first prove the following result.


4.1. PROOF OF THEOREM 77<br />

Proposition 4.1.7. Let m, n ≥ 2 and u ∈ C 2 (B m , D n ) ∩ C 1 (B m , D n ). If u is a proper<br />

<strong>harmonic</strong> map, then the boundary value <strong>of</strong> u is a constant map.<br />

Pro<strong>of</strong>.<br />

Since u α ∈ R and τ(u) α =0, we have<br />

0=a(u)(z)(1 −|z| 2 )〈Lu, u〉 + 2(1 −|z| 2 )(2|〈Nu,u〉| 2 −|u| 2 |Nu| 2 )<br />

m∑<br />

+2 (2|〈X j u, u〉| 2 −|u| 2 |X j u| 2 ).<br />

j=1<br />

For p ∈ ∂B m , let {z j } ∞ j=1 be a sequence satisfying the conditions in Corollary 4.1.6. Then,<br />

since u ∈ C 1 (B m , D n ), it follows from Corollary 4.1.6 that<br />

Thus we conclude that<br />

and hence<br />

lim a(u)(z j)(1 −|z j | 2 )〈Lu, u〉(z j )=0,<br />

j→∞<br />

lim<br />

j→∞ 2(1 −|z j| 2 )(2|〈Nu,u〉| 2 −|u| 2 |Nu| 2 )(z j )=0.<br />

2<br />

2<br />

m∑<br />

|〈X j u, u〉| 2 =<br />

j=1<br />

m∑<br />

|〈X j u, u〉| 2 =<br />

j=1<br />

On the other hand, on ∂B m , it holds that<br />

m∑<br />

|X j u| 2 at p,<br />

j=1<br />

m∑<br />

|X j u| 2 on ∂B m .<br />

j=1<br />

0=X j |u| 2 =2〈X j u, u〉.<br />

Therefore<br />

which asserts that<br />

m∑<br />

|X j u| 2 =0,<br />

j=1<br />

Since u α ∈ R, we obtain<br />

X j u α =0 on∂B m for 1 ≤ j ≤ m, 1 ≤ α ≤ n.<br />

¯X j u α =0 on∂B m for 1 ≤ j ≤ m, 1 ≤ α ≤ n.


78 CHAPTER 4. NON-EXISTENCE OF PROPER HARMONIC MAPS<br />

Moreover, from Lemma 4.1.3, we have<br />

(m − 1)(N − ¯N) =<br />

m∑<br />

(X j ¯Xj − ¯X j X j ) on ∂B m .<br />

j=1<br />

Hence<br />

(m − 1)(N − ¯N)u<br />

m∑<br />

α = (X j ¯Xj − ¯X j X j )u α =0 on∂B m .<br />

j=1<br />

Thus u is a constant map at the ideal boundary.<br />

Pro<strong>of</strong> <strong>of</strong> Theorem. Let {r j } ∞ j=1 be a sequence such that 0


4.2. A COUNTER EXAMPLE 79<br />

where (y, η, ξ, x) ∈ R + × R 3 and (Y,X) ∈ R + × R. Note that the vector field ∂/∂x is<br />

mapped to the vector field ∂/∂y via the complex structure <strong>of</strong> CH 2 . Let Φ : B 2 → CH 2 and<br />

Ψ:D 2 → RH 2 be the Cayley transforms <strong>of</strong> CH 2 and RH 2 , respectively. Assume that a<br />

map ũ : CH 2 → RH 2 takes the following form:<br />

ũ(y, η, ξ, x) =(f(y),g(x)).<br />

Then the <strong>harmonic</strong> map equation for ũ is reduced to the following system <strong>of</strong> ordinary differential<br />

equations:<br />

⎧<br />

⎪⎨ 4y 2 f yy (y) − 4yf y (y) − y 2 f(y) −1 {4f y (y) 2 − g x (x) 2 } =0,<br />

⎪⎩ g xx (x) =0,<br />

where f y = df/dy, f yy = d 2 f/dy 2 ,g x = dg/dx, g xx = d 2 g/dx 2 .<br />

For any a, b ∈ R, we can easily verify that f(y) =(|a|/2 √ 2)y and g(x) =ax + b are<br />

solutions to this system. Therefore we obtain a family <strong>of</strong> <strong>harmonic</strong> <strong>maps</strong><br />

Let ⎧⎪ ⎨<br />

⎪ ⎩<br />

˜σ (a,b) (y, x) =<br />

ũ (a,b) : CH 2 ∋ (y, η, ξ, x) ↦→<br />

( |a|<br />

2 √ 2 y, ax + b )<br />

∈ RH 2 .<br />

( |a|<br />

2 √ 2 y, a<br />

2 √ 2 x + b )<br />

: RH 2 → RH 2 ,<br />

σ (a,b) =Ψ −1 ◦ ˜σ (a,b) ◦ Ψ:D 2 → D 2 ,<br />

ũ(y, η, ξ, x) =ũ (2<br />

√<br />

2,0) (y, η, ξ, x) =(y, 2 √ 2x) :CH 2 → RH 2 ,<br />

u (a,b) =Ψ −1 ◦ ũ (a,b) ◦ Φ:B 2 → D 2 .<br />

Note that, if a ≠0, each u (a,b) is smooth up to the boundary except for (0, 1) ∈ ∂B 2 . Thus<br />

we need to investigate the boundary behavior <strong>of</strong> u (a,b) near the point (0, 1). To this end, since<br />

any σ (a,b) is an isometry <strong>of</strong> D 2 which fixes the point (0, 1) commonly, it suffices to consider<br />

the case where a =2 √ 2 and b =0. Set u =Ψ −1 ◦ ũ<br />

√<br />

(2 2,0)<br />

◦ Φ.


80 CHAPTER 4. NON-EXISTENCE OF PROPER HARMONIC MAPS<br />

It is then easy to see that u is explicitly given by<br />

u(z 1 ,z 2 )=<br />

1<br />

(1 −|z| 2 +2|1 − z 2 | 2 ) 2 +8(I(z 2 )) 2<br />

× (−8 √ 2|1 − z 2 | 2 I(z 2 ), (1 −|z| 2 ) 2 − 4|1 − z 2 | 4 +8(I(z 2 )) 2 ),<br />

where z =(z 1 ,z 2 ). Now, by a long but straightforward calculation, we can verify that u has<br />

the following properties:<br />

(1) u(0, 1) = (0, 1).<br />

(2) u is Lipschitz continuous at (0, 1).<br />

(3) u is not C 1 at (0, 1).<br />

(4) u is not a constant map at the ideal boundary.<br />

Therefore the C 1 -regularity up to the ideal boundary is an optimal condition for our theorem.<br />

Now, we prove that u satisfies the properties mentioned above. Since the boundary map<br />

<strong>of</strong> ũ : CH 2 → RH 2 is given by ũ(0,η,ξ,x)=(0, 2 √ 2x), it is clear that u : B 2 → D 2 is not<br />

a constant map at the ideal boundary. Also it is easily verified that<br />

(1 −|z| 2 +2|1 − z 2 | 2 ) 2 +8(Iz 2 ) 2 =0, |z| ≤1 ⇐⇒ z 1 =0,z 2 =1.<br />

Claim 1.<br />

Let<br />

lim u(z 1 ,z 2 )=(0, 1).<br />

(z 1 ,z 2 )→(0,1)<br />

z 1 = r(θ 1 + √ −1θ 2 ), z 2 =1+r(θ 3 + √ −1θ 4 ),<br />

where (θ 1 ) 2 +(θ 2 ) 2 +(θ 3 ) 2 +(θ 4 ) 2 =1. Note that<br />

(z 1 ,z 2 ) ≠(0, 1) ⇐⇒ |θ| 2 := (θ 3 ) 2 +(θ 4 ) 2 ≠0.<br />

Then it is easy to see that<br />

1 −|z| 2 = −r(2θ 3 + r), and |1 − z 2 | 2 = r 2 |θ| 2 .


4.2. A COUNTER EXAMPLE 81<br />

Hence we have<br />

u(z 1 ,z 2 )<br />

1<br />

=<br />

(2θ 3 + r − 2r|θ| 2 ) 2 +8(θ 4 ) (−8√ 2r|θ| 2 θ 4 , (2θ 3 + r) 2 − 4r 2 |θ| 4 +8(θ 4 ) 2 )<br />

2<br />

−→ (0, 1) (r → 0).<br />

Consequently, if we define u(0, 1) = (0, 1), then u ∈ C 0 (B 2 , D 2 ).<br />

Claim 2. u is Lipschitz continuous at (0, 1).<br />

We shall prove that there exists a positive constant C such that<br />

‖u(z 1 ,z 2 ) − (0, 1)‖ R 2<br />

sup<br />

≤ C.<br />

‖(z<br />

z∈B 2 ,z≠(0,1)<br />

1 ,z 2 ) − (0, 1)‖ C 2<br />

It is immediate from the definition that<br />

u(z 1 ,z 2 ) − (0, 1)<br />

4|1 − z 2 | 2<br />

=<br />

(1 −|z| 2 +2|1 − z 2 | 2 ) 2 +8(Iz 2 ) (−2√ 2Iz 2 , −2|1 − z 2 | 2 − (1 −|z| 2 ).<br />

2<br />

Hence we obtain<br />

‖u(z 1 ,z 2 ) − (0, 1)‖ R 2 =<br />

4|1 − z 2 | 2<br />

√<br />

(1 −|z|2 +2|1 − z 2 | 2 ) 2 +8(Iz 2 ) 2 .<br />

Using the coordinate used in the pro<strong>of</strong> <strong>of</strong> Claim1, we have<br />

On the other hand, it holds that<br />

‖u(z 1 ,z 2 ) − (0, 1)‖<br />

‖(z 1 ,z 2 ) − (0, 1)‖ = 4|θ| 2<br />

√<br />

(2θ3 + r − 2r|θ| 2 ) 2 +8(θ 4 ) 2<br />

−→<br />

2|θ| 2<br />

√<br />

(θ3 ) 2 +2(θ 4 ) 2 (r → 0).<br />

2|θ| 2<br />

√<br />

(θ3 ) 2 +2(θ 4 ) = √ 2|θ 3 | 2<br />

(θ 3 ) 2 +2(θ 4 ) 2 + √ 2 (θ3 ) 2 +2(θ 4 ) ≤ √ (θ 3 ) 2 +2(θ 4 ) 2 + |θ 3 |.<br />

2<br />

Thus, for some positive constant C, we have<br />

‖u(z 1 ,z 2 ) − (0, 1)‖<br />

lim<br />

(z 1 ,z 2 )→(0,1) ‖(z 1 ,z 2 ) − (0, 1)‖ ≤ C.


82 CHAPTER 4. NON-EXISTENCE OF PROPER HARMONIC MAPS<br />

Hence the function<br />

‖u(z 1 ,z 2 ) − (0, 1)‖<br />

‖(z 1 ,z 2 ) − (0, 1)‖<br />

is bounded on B 2 −{(0, 1)}, which implies that u is Lipschitz continuous at (0, 1).<br />

Claim 3. u is not C 1 at (0, 1).<br />

Let v be the second component <strong>of</strong> u, that is,<br />

v(z 1 ,z 2 )= (1 −|z|2 ) 2 − 4|1 − z 2 | 4 +8(Iz 2 ) 2<br />

(1 −|z| 2 +2|1 − z 2 | 2 ) 2 +8(Iz 2 ) 2 .<br />

Then<br />

∂v<br />

v(0, 1) − v(0, 1 − h) 4(2 − h)<br />

(0, 1) = lim<br />

= lim<br />

∂z2 h→+0 h<br />

h→+0 (2 + h) =2. 2<br />

On the other hand, from a straightforward calculation, we obtain<br />

v z 2 ×{(1 −|z| 2 +2|1 − z 2 | 2 ) 2 +8(Iz 2 ) 2 } 2<br />

= 16(Iz 2 ) 2 [ ¯z 2 {2|1 − z 2 | 2 − (1 −|z| 2 ) − (1 −|z| 2 ) 2 } + 2(1 −|z| 2 )]<br />

−32 √ −1Iz 2 |1 − z 2 | 2 {2|1 − z 2 | 2 +1−|z| 2 }<br />

−2(1 −|z| 2 +2|1 − z 2 | 2 ) 2 { ¯z 2 (1 −|z| 2 ) 2 +4|1 − z 2 | 2 (1 − ¯z 2 )}<br />

+2(2 − ¯z 2 ){(1 −|z| 2 ) 2 − 4|1 − z 2 | 4 }(1 −|z| 2 +2|1 − z 2 | 2 ).<br />

Here we have<br />

{(1 −|z| 2 +2|1 − z 2 | 2 ) 2 +8(Iz 2 ) 2 } 2 = {(2θ 3 + r − 2r‖θ‖ 2 ) 2 +8(θ 4 ) 2 }r 4 .<br />

Hence<br />

v z 2(0, 1 − r) = −2r(2 + r){(1 − r)(2 − r)2 +4r} + 2(1 + r){(2 − r) 2 − 4r 2 }<br />

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

−→ +∞ (r → 0).<br />

Therefore, u is not C 1 at (0, 1).


Bibliography<br />

[1] K. Akutagawa, Harmonic diffeomorphisms <strong>of</strong> the hyperbolic plane, Trans. Amer. Math.<br />

Soc. 342 (1994), 325–342.<br />

[2] K. Akutagawa and T. Tachikawa, Nonexistence results for <strong>harmonic</strong> <strong>maps</strong> <strong>between</strong> noncompact<br />

complete Riemannian <strong>manifolds</strong>, Tokyo J. Math. 16 (1993), 131–145.<br />

[3] P. Baird, Harmonic <strong>maps</strong> with symmetry, <strong>harmonic</strong> morphisms and deformations <strong>of</strong><br />

metrics, Research Notes in Math. 87, Pitman, Boston, MA, 1983.<br />

[4] J. Boggino, Generalized Heisenberg groups and solv<strong>manifolds</strong> naturally associated, Rend.<br />

Sem. Mat. Univ. Politec. Torino 43 (1985), 529–547.<br />

[5] J. Berndt, F. Tricerri and L. Vanhecke, Generalized Heisenberg groups and Damek-Ricci<br />

<strong>harmonic</strong> spaces, Lecture Notes in Math. 1598, Springer-Verlag, Berlin, Heidelberg,<br />

New York, 1995.<br />

[6] M. Cowling, A. H. Dooley, A. Korányi and F. Ricci, H-type groups and Iwasawa decompositions,<br />

Adv. Math. 87 (1991), 1–41.<br />

[7] E. Damek, Curvature <strong>of</strong> a semi-direct extension <strong>of</strong> a Heisenberg type nilpotent group,<br />

Colloq. Math. 53 (1987), 249–253.<br />

[8] E. Damek, The geometry <strong>of</strong> a semi-direct extension <strong>of</strong> a Heisenberg type nilpotent group,<br />

Colloq. Math. 53 (1987), 255–268.<br />

83


84 BIBLIOGRAPHY<br />

[9] E. Damek and F. Ricci, A class <strong>of</strong> nonsymmetric <strong>harmonic</strong> Riemannian spaces, Bull.<br />

Amer. Math. Soc. 27 (1992), 139–142.<br />

[10] E. Damek and F. Ricci, Harmonic analysis on solvable extensions <strong>of</strong> H-type groups, J.<br />

Geom. Anal. 2 (1992), 213–248.<br />

[11] W. Y. Ding, Symmetric <strong>harmonic</strong> <strong>maps</strong> <strong>between</strong> spheres, Comm. Math. Phys. 118<br />

(1988), 641–649.<br />

[12] H. Donnelly, Dirichlet problem at infinity for <strong>harmonic</strong> <strong>maps</strong>: rank one symmetric<br />

spaces, Trans. Amer. Math. Soc. 344 (1994), 713–735.<br />

[13] I. Dotti, On the curvature <strong>of</strong> certain extensions <strong>of</strong> H-type groups, Proc. Amer. Math.<br />

Soc. 125 (1997), 573–578.<br />

[14] P. Eberlein and B. O’Neill, Visibility <strong>manifolds</strong>, Pacific J. Math. 46 (1973), 45–109.<br />

[15] J. Eells and A. Ratto, Harmonic <strong>maps</strong> <strong>between</strong> spheres and ellipsoids, Internat. J. Math.<br />

1 (1990), 1–27.<br />

[16] J. Eells and A. Ratto, Harmonic <strong>maps</strong> and minimal immersions with symmetries, Ann.<br />

<strong>of</strong> Math. Stud. 130, Princeton University Press, Princeton, NJ, 1993.<br />

[17] J. Eells and H. Sampson, Harmonic mappings <strong>of</strong> Riemannian <strong>manifolds</strong>, Amer. J. Math.<br />

86 (1964), 109–160.<br />

[18] H. Gauchman and G. Toth, <strong>Constructions</strong> <strong>of</strong> <strong>harmonic</strong> polynomial <strong>maps</strong> <strong>between</strong><br />

spheres, Geom. Dedicata 50 (1994), 57–79.<br />

[19] R. S. Hamilton, Harmonic <strong>maps</strong> <strong>of</strong> <strong>manifolds</strong> with boundary, Lecture Notes in Math.<br />

471, Springer-Verlag, Berlin-New York, 1975.<br />

[20] P. Hartman, On homotopic <strong>harmonic</strong> mappings, Canad. J. Math. 19 (1967), 673–687.


BIBLIOGRAPHY 85<br />

[21] A. Kaplan, Fundamental solutions for a class <strong>of</strong> hypoelliptic PDE generated by composition<br />

<strong>of</strong> quadratic forms, Trans. Amer. Math. Soc. 258 (1980), 147–153.<br />

[22] A. Kasue and T. Washio, Growth <strong>of</strong> equivariant <strong>harmonic</strong> <strong>maps</strong> and <strong>harmonic</strong> morphisms,<br />

Osaka J. Math. 27 (1990), 899–928.<br />

[23] M. Lanzendorf, Einstein metrics with nonpositive sectional curvature on extensions <strong>of</strong><br />

Lie algebras <strong>of</strong> Heisenberg type, Geom. Dedicata 66 (1997), 187–202.<br />

[24] S.Y. Li and L. Ni, On the holomorphicity <strong>of</strong> proper <strong>harmonic</strong> <strong>maps</strong> <strong>between</strong> unit balls<br />

with the Bergman metrics, Math. Ann. 316 (2000), 333–354.<br />

[25] P. Li and L.-F. Tam, The heat equation and <strong>harmonic</strong> <strong>maps</strong> <strong>of</strong> complete <strong>manifolds</strong>,<br />

Invent. Math. 105 (1991), 1–46.<br />

[26] P. Li and L.-F. Tam, Uniqueness and regularity <strong>of</strong> proper <strong>harmonic</strong> <strong>maps</strong>, Ann. <strong>of</strong> Math.<br />

137 (1993), 167–201.<br />

[27] P. Li and L.-F. Tam, Uniqueness and regularity <strong>of</strong> proper <strong>harmonic</strong> <strong>maps</strong> II, Indiana<br />

Univ. Math. J. 42 (1993), 591–635.<br />

[28] T. Nagasawa and A. Tachikawa, Blow-up solutions for ordinary differential equations<br />

associated to <strong>harmonic</strong> <strong>maps</strong> and their applications, J. Math. Soc. Japan 53 (2001),<br />

485–500.<br />

[29] T. Nagasawa and K. Ueno, Initial-final value problems for ordinary differential equations<br />

and applications to equivariant <strong>harmonic</strong> <strong>maps</strong>, J. Math. Soc. Japan 50 (1998), 545–555.<br />

[30] M. Parker, Orthogonal multiplications in small dimensions, Bull. London Math. Soc.<br />

15 (1983), 368–372.<br />

[31] A. Ratto, Harmonic <strong>maps</strong> <strong>of</strong> spheres and the Hopf construction, Topology 28 (1989),<br />

379–388.


86 BIBLIOGRAPHY<br />

[32] A. Ratto and M. Rigoli, On the asymptotic behaviour <strong>of</strong> rotationally symmetric <strong>harmonic</strong><br />

<strong>maps</strong>, J. Differential Equations 101 (1993), 15–27.<br />

[33] R. Schoen and S. T. Yau Compact group actions and the topology <strong>of</strong> the <strong>manifolds</strong> with<br />

nonpositive curvature, Topology 18 (1979), 361–380.<br />

[34] R. T. Smith, Harmonic mappings <strong>of</strong> spheres, Amer. J. Math. 97 (1975), 364–385.<br />

[35] A. Tachikawa, A nonexistence result for <strong>harmonic</strong> mappings from R n into H n , Tokyo<br />

J. Math. 11 (1988), 311–316.<br />

[36] A. Tachikawa, Harmonic mappings from R m into a <strong>Hadamard</strong> manifold, J. Math. Soc.<br />

Japan 42 (1990), 147–153.<br />

[37] G. Toth, Classification <strong>of</strong> quadratic <strong>harmonic</strong> <strong>maps</strong> <strong>of</strong> S 3 into spheres, Indiana Univ.<br />

Math. J. 36 (1987), 231–239<br />

[38] K. Ueno, Some new examples <strong>of</strong> eigen<strong>maps</strong> from S m into S n , Proc. Japan Acad. 69<br />

(1993), 205–208.<br />

[39] K. Ueno, A study <strong>of</strong> ordinary differential equations arising from equivariant <strong>harmonic</strong><br />

<strong>maps</strong>, Tokyo J. Math. 17 (1994), 395–416.<br />

[40] K. Ueno, The Dirichlet problem for <strong>harmonic</strong> <strong>maps</strong> <strong>between</strong> Damek-Ricci spaces,<br />

Tôhoku Math. J. 49 (1997), 565–575.<br />

[41] K. Ueno, Non-existence <strong>of</strong> proper <strong>harmonic</strong> <strong>maps</strong> from complex hyperbolic spaces into<br />

real hyperbolic spaces, preprint.<br />

[42] H. Urakawa, Equivariant <strong>harmonic</strong> <strong>maps</strong> <strong>between</strong> compact Riemannian <strong>manifolds</strong> <strong>of</strong><br />

cohomogeneity one, Michigan Math. J. 40 (1993), 27–51.<br />

[43] H. Urakawa and K. Ueno, Equivariant <strong>harmonic</strong> <strong>maps</strong> associated to large group actions,<br />

Tokyo J. Math. 17 (1994), 417–437.


BIBLIOGRAPHY 87<br />

[44] R. Wood, Polynomial <strong>maps</strong> from spheres to spheres, Invent. Math. 5 (1968), 163–168.<br />

[45] Y. L. Xin, Riemannian submersion and equivariant <strong>harmonic</strong> <strong>maps</strong>, Differential geometry<br />

(Shanghai, 1991), 272–287, World Sci. Publishing, River Edge, NJ, 1993.<br />

[46] Y. L. Xin, On <strong>harmonic</strong> representative <strong>of</strong> π 2m+1 (S 2m+1 ) , Math. Z. 216 (1994), 389–400.


TOHOKU MATHEMATICAL PUBLICATIONS<br />

No.1<br />

No.2<br />

Hitoshi Furuhata: Isometric pluri<strong>harmonic</strong> immersions <strong>of</strong> Kähler <strong>manifolds</strong><br />

into semi-Euclidean spaces, 1995.<br />

Tomokuni Takahashi: Certain algebraic surfaces <strong>of</strong> general type with irregularity<br />

one and their canonical mappings, 1996.<br />

No.3 Takeshi Ikeda: Coset constructions <strong>of</strong> conformal blocks, 1996.<br />

No.4<br />

No.5<br />

No.6<br />

No.7<br />

No.8<br />

No.9<br />

Masami Fujimori: Integral and rational points on algebraic curves <strong>of</strong> certain<br />

types and their Jacobian varieties over number fields, 1997.<br />

Hisatoshi Ikai: Some prehomogeneous representations defined by cubic forms,<br />

1997.<br />

Setsuro Fujiié: Solutions ramifiées des problèmes de Cauchy caractéristiques<br />

et fonctions hypergéométriques à deux variables, 1997.<br />

Miho Tanigaki: Saturation <strong>of</strong> the approximation by spectral decompositions<br />

associated with the Schrödinger operator, 1998.<br />

Y. Nishiura, I. Takagi and E. Yanagida: Proceedings <strong>of</strong> the International<br />

Conference on Asymptotics in Nonlinear Diffusive Systems — towards the<br />

Understanding <strong>of</strong> Singularities in Dissipative Structures —, 1998.<br />

Hideaki Izumi: Non-commutative L p -spaces constructed by the complex interpolation<br />

method, 1998.<br />

No.10 Youngho Jang: Non-Archimedean quantum mechanics, 1998.<br />

No.11 Kazuhiro Horihata: The evolution <strong>of</strong> <strong>harmonic</strong> <strong>maps</strong>, 1999.<br />

No.12<br />

No.13<br />

No.14<br />

No.15<br />

Tatsuya Tate: Asymptotic behavior <strong>of</strong> eigenfunctions and eigenvalues for<br />

ergodic and periodic systems, 1999.<br />

Kazuya Matsumi: Arithmetic <strong>of</strong> three-dimensional complete regular local<br />

rings <strong>of</strong> positive characteristics, 1999.<br />

Tetsuya Taniguchi: Non-isotropic <strong>harmonic</strong> tori in complex projective spaces<br />

and configurations <strong>of</strong> points on Riemann surfaces, 1999.<br />

Taishi Shimoda: Hypoellipticity <strong>of</strong> second order differential operators with<br />

sign-changing principal symbols, 2000.


No.16 Tatsuo Konno: On the infinitesimal isometries <strong>of</strong> fiber bundles, 2000.<br />

No.17<br />

Takeshi Yamazaki: Model-theoretic studies on subsystems <strong>of</strong> second order<br />

arithmetic, 2000.<br />

No.18 Daishi Watabe: Dirichlet problem at infinity for <strong>harmonic</strong> <strong>maps</strong>, 2000.<br />

No.19<br />

No.20<br />

No.21<br />

No.22<br />

Tetsuya Kikuchi: Studies on commuting difference systems arising from<br />

solvable lattice models, 2000.<br />

Seiki Nishikawa: Proceedings <strong>of</strong> the Fifth Pacific Rim Geometry Conference,<br />

2001.<br />

Mizuho Ishizaka: Monodromies <strong>of</strong> hyperelliptic families <strong>of</strong> genus three curves,<br />

2001.<br />

Keisuke Ueno: <strong>Constructions</strong> <strong>of</strong> <strong>harmonic</strong> <strong>maps</strong> <strong>between</strong> <strong>Hadamard</strong> <strong>manifolds</strong>,<br />

2001.

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

Saved successfully!

Ooh no, something went wrong!