Download as pdf or txt
Download as pdf or txt
You are on page 1of 29

Minimal Surfaces in H2 × R

Barbara Nelli - Harold Rosenberg

Abstract
In H2× R one has catenoids, helicoids and Scherk-type surfaces. A Jenkins-
Serrin type theorem holds here. Moreover there exist complete minimal graphs in
H2 with arbitrary continuous asymptotic values. Finally, a graph on a domain
of H2 cannot have an isolated singularity.

1 Introduction
In this paper we consider minimal surfaces in H2 × R; particularly, surfaces which are
vertical graphs over domains in H2 . When a convex domain D ⊂ H2 is bounded by
geodesic arcs A1 , . . . An , B1 , . . . , Bm , together with strictly convex arcs C1 , . . . , Cs , we
obtain necessary and sufficient conditions (in terms of the lengths of the boundary arcs
of D) which assure the existence of a unique function u defined in D, whose graph is
a minimal surface of H2 × R, and which takes the values +∞ on the arcs A1 , . . . An ,
−∞ on the arcs B1 , . . . , Bm , and arbitrary prescribed continuous data on the arcs
C1 , . . . , Cs . In R2 × R, this is the theorem of Jenkins and Serrin [JS].
For example, let D be a domain whose boundary is a regular geodesic octagon with
sides A1 , B1 , . . . , A4 , B4 , and suppose the interior angles are 2.π Our theorem yields a
function u in D, whose graph is minimal, taking the values +∞ on each Ai , and −∞
on each Bj . The graph of u is bounded by the eight vertical geodesics passing through
the vertices of D. Rotation of each H2 × {t}, by π about each vertex of D × {t}, extends
the graph of u to a complete embedded minimal surface in H2 × R (one continues the
rotation about all the vertical geodesics that arise). One can take the quotient of H2 ×R
by various Fuchsian groups to obtain interesting quotient surfaces. For example, one
can obtain an 8-punctured sphere in the quotient of total curvature −12π; with four
top ends and four bottom ends.
One also obtains graphs over ideal polygons with vertices at infinity. For example,
consider the polygon which is the boundary of the convex hull of the n roots of unity, n
even and at least four. Then, there is a minimal graph over the interior of this polygon
taking the values plus and minus infinity on adjacent edges (cf. Figure 2(b)).
We prove the existence of entire minimal graphs over H2 (Bernstein’s theorem fails
here). In the model {0 ≤ x21 + x22 < 1} of H2 , the asymptotic boundary of H2 × R is
{x21 + x22 = 1} × R. For any Jordan curve Γ in the asymptotic boundary of H2 × R that

1
has a simple projection on {x21 + x22 = 1}, there is a minimal graph over H2 having Γ
as asymptotic boundary.
In [DN], the existence of such minimal graphs is established when Γ is the boundary
value of a function with very small C 3 -norm on the disk.
Finally, we prove a theorem for minimal graphs defined over a punctured disk in H2 :
the graph extends smoothly to the puncture.

2 Preliminaries
In the three dimensional manifold H2 × R, we take the disk model for H2 . Let x1 , x2
denote the coordinates in H2 and x3 the coordinate in R. The metric in H2 × R is

dx21 + dx22
dσ 2 = + dx23
F
where µ ¶2
1 − x21 − x22
F =
2
The graph of a function u defined over a domain in H2 has constant mean curvature
H if and only if u satisfies the following equation:

µ ¶
∇u 2H
(1) div =
τu F
p
where τu = 1 + F |∇u|2 and the divergence is the divergence in R2 . We list the
principal steps for the computation of (1).
The Christoffel symbols for the metric dσ 2 are the following:
x1
Γ111 = Γ212 = Γ221 = √
F
x2
Γ222 = Γ112 = Γ121 = √
F
x2 x1
Γ211 = − √ , Γ122 = − √
F F
The other Γkij are identically zero.
√ √
Let e1 , e2 , e3 , be the canonical basis of R3 and set ε1 = F e1 , ε2 F e2 , ε3 = e3 , so
¯ be the connection of
that ε1 , ε2 , ε3 is an orthonormal basis for H2 × R. Finally let ∇
2
the metric dσ . We have:

¯ ε ε1 = −x2 ε2 ,
∇ ¯ ε ε2 = −x1 ε1 ,

1 2

2
¯ ε ε2 = x2 ε1 ,
∇ ¯ ε ε1 = x1 ε2 .

1 2

The coordinate vector fields on the graph of u are X1 = √1F ε1 + u1 ε3 ,


√ √
X2 = √1F ε2 + u2 ε3 and N = τ −1 (−u1 F ε1 − u2 F ε2 + ε3 ) is the upward unit normal.
The induced metric on the graph is:
1 1
g11 = + u21 , g12 = u1 u2 , g22 = + u22 .
F F
The coefficients of the second fundamental form are:
µ ¶
¯ 1 x 1 u1 x 2 u2
b11 =< ∇X1 X1 , N >= − √ + √ + u11
τ F F
µ ¶
¯ X2 , N >= 1 x 2 u 1 x 1 u 2
b12 =< ∇ X1 − √ − √ + u12
τ F F
µ ¶
¯ X2 , N >= 1 x 1 u 1 x 2 u 2
b22 =< ∇ X2 √ − √ + u22
τ F F
where < , > is the scalar product for the metric dσ 2 .
Equation (1) is obtained by substituting the quantities just calculated in the following
identity:
b11 g22 + b22 g11 − 2b12 g12
2H = 2
.
g11 g22 − g12

3 Catenoids and Helicoids

Catenoids. We construct a family of minimal rotational surfaces in H2 × R (see also


[PR]).
Let π be a vertical geodesic plane containing the origin and let γ be a curve in π.
Assume γ to be a graph over the x3 axis. Let r be the Euclidean distance between
the point of γ at height t and the x3 axis: r = r(t) is a parametrization of the curve
γ. Consider the surface of revolution S obtained by rotating γ about the x3 axis. S is
minimal if and only if r = r(t) satisfies the following differential equation:

(2) 4r(t)r00 (t) − 4r0 (t)2 − (1 − r(t)4 ) = 0.

A first integral for equation (2) is

3
r02 1 + r4
(3) J(t) = − − = −C
r2 4r2
dJ(t)
where C is a positive constant (i.e. dt
= 0 along the curve γ).
Hence we obtain:

r
0 1 + r4
(4) r =± Cr2 −
4
The allowed values for C and r are
1
C>
2
√ √
2C − 4C 2 − 1 ≤ r2 < 1 ≤ 2C + 4C 2 − 1
i.e.
r r
2C + 1 2C − 1
rmin = − ≤r<1
2 2
Remark that as C−→ 12 then rmin −→1 hence the curve γ disappears at infinity, while
as C−→∞ then rmin −→0.
We can write equation (4) as follows:

dt 2
(5) = ±p
dr 4Cr2 − (1 + r4 )

In order to study the curve γ we can choose the positive sign in (5), i.e. t > 0. In fact
by the symmetries of equation (4) and (5), the curve γ for t < 0 will be the reflection
with respect to the plane x3 = 0 of the curve γ for t > 0. With the choice of the positive
sign, we have the following properties.
dt
(i) dr > 0 hence t is an increasing function of r.
dt
(ii) For r = rmin we have dr = ∞, i.e. the tangent to the curve γ at the point r = rmin
is parallel to the x3 axis and this is the only point where this happens.
(iii) As r−→1, we have:

dt 2
−→ √ .
dr 2C − 1
Hence, when C varies between 12 and +∞ the asymptotic angle of γ varies between π
2
and 0.
d2 t
(iv) dr 2 < 0, hence the concavity does not change.

4
0
1 α
0
x1
0
1 1
0
0
1
3
0
1
0
1
0
1
0 1
0
1
0
1
0
1
0
1
0
1 1
0
0
1
0
1
0
1
0
1
0 1 0
0 1 0
1
1
0
1 rmin
Figure 1

(v) Consider the change of variables w = r2 . Equation (5) becomes

dw
(6) dt = ± q √ √
w(2C + 4C 2 − 1 − w)(w − 2C + 4C 2 − 1)
Hence
Z w ³ √ √ ´− 12
t(w) = ± √
s(2C + 4C 2 − 1 − s)(s − 2C + 4C 2 − 1) ds
2C− 4C 2 −1

This is an elliptic integral. By the properties of elliptic functions lim t(w) is indepen-
√ C→∞
dent of w. For every value of the constant C we have t(4C − 4C 2 − 1) = 0, hence for
the limit value C = ∞, the surface is a horizontal plane (doubly covered).
Using (i)-(v) we have the following theorem (see Figure 1).

Theorem 1. Let Γ± (t) be the two circles at infinity of H2 × R defined by {x21 + x22 =
1, x3 = ±t}. Then, for each t ∈ (0, π2 ) there exists a rotational surface (catenoid)
C(t), whose asymptotic boundary is Γ+ (t) ∪ Γ− (t). As t−→0, C(t) converges to the
doubly covered plane H2 with a singularity at the origin. As t−→∞, C(t) diverges to
the asymptotic boundary of H2 × R. Furthermore for any angle α ∈]0, π2 [ there is a
C(t) whose asymptotic normal vector at boundary points forms an angle equal to α
with the x3 axis.

Added in proof: see Proposition 5.1 in [5], Theorem 15 in [12]).

Helicoids. Let α be a horizontal geodesic passing through the x3 axis. Consider the
surface E obtained by translating α vertically and rotating it around the x3 axis. A
parametrization for E is the following:

5
A1 B1
A1 B1
D D
B2 A2
B2 A2

(a) Geodesic 4 − gon (b) Ideal 4 − gon


Figure 2

X(u, v) = (v cos θ(u), v sin θ(u), u)


v ∈ (−1, 1), u ∈ R and θ : R−→R is a C 2 function representing the angle between α
and the x1 axis at the level u.
E is a minimal surface if and only if

θuu = 0.
Hence the solutions are:
(i) θ(u) = a, a ∈ R. In this case E is a vertical plane forming an angle a with the x1
axis.
(ii) θ(u) = au, a ∈ R \ {0}. In this case the surface E is congruent to a Euclidean
helicoid.

4 Scherk Type Surfaces

Let P be a regular 2k-gon in H2 with (open) edges A1 , B1 , . . . , Ak , Bk (see Figure 2(a),


k = 2). We will construct a minimal graph Σ over the domain D bounded by P such
that the boundary values are alternatively +∞ on the edges Ai and −∞ on the edges
Bi . Σ will be called a Scherk type surface. Also, the Scherk surface exists if the vertices
of Ai and Bj are at infinity; so that P is a ideal polygon whose vertices are the 2k
roots of unity (see Figure 2(b), k = 2).
Choose one edge of P where the desired value is +∞ and call it A. Let B and C be
the two geodesic arcs passing through the center of P and the vertices of A. Let T

6
p1
ε
q
δ 1
βε
p3
d τ A
p8
p4
α
γε
q2
p2
Figure 3

be the (open) triangle with sides A, B and C. Let Γ(n) be the curve obtained by the
union of the following geodesics arcs: B, C together with the arcs obtained by raising
A to height n, and the vertical geodesics joining the vertices of the raised A with the
vertices of A.
Let Σn be a solution of Plateau’s problem for Γ(n). Rado’s theorem is true in H2 × R,
since vertical translation is an isometry, hence Σn is the graph of a function un defined
in the triangle T and

un|A = +n, un|B = un|C = 0.

Theorem 2.The sequence {un } converges to a minimal solution u defined in T such


that

u|A = +∞, u|B = u|C = 0.


Moreover, the gradient of u diverges as one approaches the side A.

Proof. The sequence {un } is non decreasing and positive. Hence, to show that the
function u exists, we will prove that the sequence {un } is uniformly bounded on compact
subsets K of T.
We start by constructing a barrier over the graph of the un in K.
The following construction is represented in Figure 3.
Let α be the horizontal geodesic containing the side A. Denote by q1 and q2 the vertices
of A. For i = 1, 2, let pi the point on α \ A at a distance ε > 0 from qi . Denote by aε
the geodesic arc between p1 and p2 . Let τ be the horizontal geodesic orthogonal to the
edge A passing through the mid-point of A (in order to simplify Figure 3 we assume
that τ passes through the origin of H2 ). Let p∞ be the point at infinity of τ contained
in the same halfplane as T defined by α.

7
bε (h)
cε (h)

p
1


Lh1
p p
3 2

p4
Lh2
Figure 4

Finally denote by βε the horizontal geodesic through p1 and p∞ and by γε the horizontal
geodesic through p2 and p∞ .
Consider a horizontal geodesic δ orthogonal to τ and let p3 = δ ∩ βε , p4 = δ ∩ γε . Call
d the geodesic arc on δ between p3 and p4 , bε the geodesic arc on βε between p1 and
p3 and cε the geodesic arc on γε between p2 and p4 . We can chose ||bε || and ||cε || large
enough such that the quadrilateral with edges aε , bε , cε , d contains the triangle T.
Let h be a positive number and call ∗(h) each object obtained by translating vertically
to height h an object of H2 × {0}. Consider the following curves (see Figure 4):

Lh1 = bε ∪ (p1 × [0, h]) ∪ bε (h) ∪ (p3 × [0, h]),


Lh2 = cε ∪ (p2 × [0, h]) ∪ cε (h) ∪ (p4 × [0, h]).
We claim that there exists a least area, hence stable, minimal annulus bounded by
Lh1 ∪ Lh2 , if ||bε || is sufficiently large. A sufficient condition is given by the Douglas
criteria for the Plateau problem: if there is an annulus bounded by Lh1 ∪ Lh2 with area
smaller than the sum of the areas of the flat geodesic domains bounded by Lh1 and Lh2 ,
then there exists a least area minimal annulus bounded by Lh1 ∪ Lh2 .
By a straightforward computation we obtain that the sum of the areas of the flat
geodesic domains bounded by Lh1 and Lh2 is equal to 2||bε ||h (as ||cε || = ||bε ||).
Now, consider the annnulus that is the union of the four geodesic domains bounded by
the following quadrilaterals (see Figure 5):

Q1 = aε ∪ (p1 × [0, h]) ∪ aε (h) ∪ (p2 × [0, h]),


Q2 = d ∪ (p3 × [0, h]) ∪ d(h) ∪ (p4 × [0, h]),

8
1
0

1
0 Q4
1
0
1
0

Q1
Q2
1
0

Q3
1
0
1
0
1
0
Figure 5

Q3 = aε ∪ bε ∪ d ∪ cε ,
Q4 = aε (h) ∪ bε (h) ∪ d(h) ∪ cε (h).
The area of this annulus is at most 2π + 2||aε ||h, as the area of a hyperbolic triangle is
always smaller than π.
Then, in order to satisfy the Douglas condition, we need:

2π + 2||aε ||h < 2||bε ||h


that is verified as soon as we choose the edge bε long enough. Hence, there exists a
least area minimal annulus Ahε bounded by Lh1 and Lh2 , for any h. By the maximum
principle Ahε is contained in the convex hull of Lh1 ∪ Lh2 .
For each n the annulus Ahε is above the surface Σn (the graph of un ); by above we mean
that if a vertical geodesic meets both surfaces, then the point of Σn is below the points
of Ahε .
To see this, translate vertically Ahε to height n (so, every point of Ahε is above height
n). Then lower the translated Ahε back to height zero. By the maximum principle,
there is no interior contact point between Ahε and Σn before returning to the original
position of Ahε . Moreover, if ² > 0, the boundaries of the two surfaces do not touch.
Letting ε−→0, we conclude that Ah0 = Ah is above Σn and, by the boundary maximum
principle at each interior point of the vertical geodesics q1 × [0, h] and q2 × [0, h] the
tangent plane to Ah is “outside” the tangent plane to Σn (i.e. the angle between the
tangent plane to Σn and the geodesic plane containing either Lh1 or Lh2 is bigger than
the angle between this last plane and the tangent plane to Ah ).
The barrier Ah shows that the sequence {un } is uniformly bounded on compact subsets
K of T such that K is contained in the horizontal projection of Ah . The idea is to show
that the horizontal projections of Ah exhaust T as h−→∞.

9
M
C
t

α
P+

Figure 6

For k > h, one can use Ah as barrier to solve the Plateau problem to find a stable
annulus Ak with boundary Lk1 , Lk2 . So, translating Ah vertically, one sees that the
two surfaces are never tangent (neither at interior points, nor at boundary points).
Hence as k−→∞, the angle the tangent plane of Ak makes along the vertical boundary
segments is controlled by that of Ah .
Now, for each n let M n be the surface A2n translated down a distance n. As each Mn is
stable, one has local uniform area bounds and uniform curvature estimates (see [Sc]).
So, a subsequence of {M n } converges to a minimal surface M ∞ . By the maximum
principle, one can translate Ah up to +∞ and down to −∞ without ever touching
M ∞ . Then, there is some component M of M ∞ whose boundary is the union of the
two vertical geodesics q1 × R and q2 × R. Furthermore the distance between M and
A × R is bounded. In fact this distance is uniformly bounded.
Now, we have to prove that M = A × R.
In H2 , consider the family of equidistant circles {Ct }t≥0 defined as follows: C0 = α,
each Ct is the circle equidistant from the geodesic α, whose curvature vector points
towards the halfplane P+ determined by α, containing the triangle T (see Figure 6).
The family of surfaces Ct × R foliates P+ × R. When t is large one has

(Ct × R) ∩ M = ∅.

Now decrease t : By the maximum principle, one cannot have a first point of contact
between M and Ct × R before t = 0. Then M = A × R and we are through.
Thus {un } has a subsequence converging to a minimal solution u defined on T. The
convergence is uniform on compact subsets of T. Furthermore, as we desired:

u|A = ∞, u|B = u|C = 0.

10
The last assertion of Theorem 2 follows from a more general fact proved in Lemma 1
of next section.
ut

Remark 1. We notice that our construction can be done for any angle, smaller than
π, between the two geodesic arcs where the boundary value is zero. In a 2k-gon P as
above, such angles are πk , k = 2, 3, . . . . Then, we make the symmetry of the graph of
u with respect to one of the two geodesic arcs where u is zero and keep on going with
such symmetries in order to close the surface. The surface thus obtained is the Scherk
type surface Σ, which we were looking for.
Also one can show that the Scherk solutions u in T converge to a Scherk solution in
the ideal triangle T ∗ obtained as limit of the triangles T when the length of the sides
B and C tend to infinity. Reflection then gives a Scherk surface graph over the interior
of a 2k-gon P.

Remark 2. When interior angles of P are choosen to be π2 , (k > 2), then Σ extends
to a complete embedded minimal surface in H2 × R. In fact the surface Σ is bounded
by the 2k vertical geodesics through the vertices of P and one extends Σ by rotation
of π about all the vertical geodesics that arise.

Remark 3. Let P be the regular 2k-gon with π2 angles. Consider the symmetry of P
about each of its vertices. This produces 2k new 2k-gons isometric to P, each having
a vertex in common with P. Consider the hyperbolic isometries identifying alternate
sides of P (that is the translation along the edge between the two chosen sides). The
quotient of the surface Σ by these translations gives a 2k-punctured sphere whose total
curvature is −4π(1 − k).

Remark 4. There is another natural way to obtain a complete surface from Σn


when the interior angles of P are equal to π2 (Σn is the graph of un over the triangle
T ). Assume that the polygon P has 2k sides. Do the symmetry of Σn about all the
geodesic arcs of its boundary. Then continue extending the surface by symmetry in
the geodesic arcs of the boundary. This yields a complete embedded minimal surface
in H2 × R that is invariant by vertical translation by 2n. The quotient of the surface
by this translation gives a compact surface of genus k.

11
5 Jenkins-Serrin Types Theorems

We give necessary and sufficient conditions to solve the Dirichlet problem for the mini-
mal surface equation in H2 × R, over a convex domain of H2 , allowing infinite boundary
values on some arcs of the boundary of the domain.
Let us fix some notation.
We consider an open bounded convex domain D whose boundary ∂D contains two sets
of (open) geodesic arcs A1 , . . . , Ak and B1 , . . . , Bl with the property that no two Ai
and no two Bi have a common endpoint. The remaining part of ∂D is the union of
open convex arcs C1 , . . . , Ch and all endpoints.
We want to find a solution u of the minimal surface equation in D such that

u|Ai = +∞, u|Bj = −∞,


i = 1, . . . , k, j = 1, . . . , l and u takes assigned continuous data on each arc Cs ,
s = 1, . . . , h.
The existence of such a solution depends on a relation between the lengths of the
geodesic arcs of the boundary and the perimeter of polygons inscribed in ∂D whose
vertices are chosen among the vertices of Ai , Bj .
Let P be such a polygon and let
X X
α= ||Ai ||, β = ||Bj ||, γ = Perimeter(P).
Ai ⊂P Bj ⊂P

Theorem 3. Let D be a domain as above and let f s : Cs −→R be continuous functions.


If {Cs } 6= ∅, then the Dirichlet problem in D with boundary values

u|Ai = +∞, u|Bj = −∞, u|Cs = f s


has a solution if and only if

(9) 2α < γ, 2β < γ

for each polygon P as above. If {Cs } = ∅ the result is the same except that if P = ∂D,
then condition (9) should be replaced by α = β.
If it exists, the solution is unique; in the case {Cs } = ∅ uniqueness is up to a constant.

Remark 5. We notice that two convex arcs Cs may have a common endpoint p; there
may be a discontinuity of the data f s at p. It will be clear from the proof of Theorem 3

12
that the minimal surface obtained in this case will contain the vertical segment through
p, between the two limit values at p, of the continuous boundary data.

This result is analogous to that of Jenkins and Serrin for minimal graphs in R3 (cf.
[JS]).
We prove Theorem 3 in 6 steps. Each step, especially 1 and 3, is an interesting result
on its own.
Step 1. Existence when ∂D contains only one geodesic arc A, and one strictly convex
arc C. The function f : C−→R is continuous and positive.
Step 2. Existence when ∂D contains geodesic arcs A1 , . . . , Ak and strictly convex arcs
C1 , . . . , Ch . The functions f s : Cs −→R are continuous and positive.
Step 3. The same as Step 2, with C1 , . . . , Ch convex arcs (not necessarily strictly
convex).
Step 4. Existence when ∂D contains geodesic arcs A1 , . . . , Ak , B1 , . . . , Bl and convex
arcs C1 , . . . , Ch with h ≥ 1.
Step 5. Existence when ∂D contains only geodesic arcs A1 , . . . , Ak , B1 , . . . , Bl .
Step 6. Uniqueness.
Proof of Step 1. Let un : D−→R be the minimal solution with boundary values

un |A = +n, un |C = min(n, f ).
Let us prove that un exists. Define Γ(n) to be the union of the following geodesic
arcs: the geodesic arc A raised to height n, the graph of the function min(n, f ) and
the vertical geodesic arcs joining the endpoints of the curves just described. Let Σ(n)
be the solution of the Plateau problem for the curve Γ(n). By Rado’s theorem, Σ(n)
is the graph of a function un defined in D, with the desired boundary values. By the
maximum principle, {un } is an increasing sequence.
We need to prove that the sequence {un } is uniformly bounded on compact subsets of
D.
Given a complete geodesic α in H2 , E one of the components of H2 \ α, there exists
a minimal graph h defined on D, asymptotic to +∞ on α and to zero on ∂∞ (E) (see
Figure 7 in the disk model of H2 ). This function was found independently by U.
Abresch and R. Sa Earp (see [9]).
In the halfspace model of H2 with E = {x > 0, y > 0} :
Ãp !
x2 + y 2 + y
h(x, y) = ln x > 0, y > 0
x
Let K be a compact subset of D. Let α be a complete geodesic, disjoint from the
geodesic containing A, intersecting C in two points, such that the region of D bounded
by α and A is disjoint from K. The geodesic α separates the circle at infinity in two
arcs; let B denote the arc at infinity such that the disk E, bounded by α ∪ B, contains

13
0

Figure 7

K. Let h be the minimal graph on E which is +∞ on α and on B it is the maximum


of f on K ∩ C. By the maximum principle, each un is bounded by h on K. Then, the
sequence {un } converges to a minimal solution u on D.
The existence of Scherk’s type surface in a triangle, guarantees that u takes the right
boundary values, as in [3]
u
t

Remark 6. Let C be a strictly convex arc and denote by C(C) the (open) convex hull
of C. Let u be a minimal solution in C(C) with bounded values on C. As a result of
the previous proof, u is bounded on every compact set of of C(C) depending only on
the values of u on C and on the distance of the compact set to the boundary of C(C).

Assertion: if u is a minimal solution that is unbounded on C, then u is unbounded in


C(C).
This assertion implies that for solving the Dirichlet problem one can not assign infinite
data on a strictly convex arc of the boundary of the domain.
For the proof of the assertion we use a modification of the argument of step 1. Let
A be the geodesic arc in the boundary of C(C). For each n ∈ IN, define the function
un = min{n, u}. Let ∆ be a horizontal geodesic triangle containing C(C) with sides a,
b, c, such that the side a contains A in its interior. Let ϕ+ be the Scherk type solution
equal to +∞, on a, and zero on b and c.
Let K be a compact set in C(C) ∪ C. On ∂K we have

min un ≤ un ≤ max un + ϕ+
K∩C K∩C

By the maximum principle, the previous inequality holds in K. Letting n−→∞, one
sees that u is unbounded on C(C).

14
For the proof of Step 2, we need several preliminary results.
The first depends on the fact that the tangent plane to the graph of a minimal solution
is almost vertical at points near to a geodesic arc of the boundary where the solution
diverges to infinity. Let us be more precise.
Denote by S the graph of a minimal solution u : D−→R and let

(ν)u = ((ν1 )u , (ν2 )u , (ν3 )u )

be the inward unit conormal to the boundary of S.


Let (x1 (s), x2 (s), x3 (s)) be an arc length parametrization of the boundary of S. A
straightforward computation yields:
∂x1 u2 ∂x2 u1
(ν3 )u = − + .
∂s τ ∂s τ
Then |(ν3 )u | < 1 and (ν3 )u is integrable on arcs of ∂D regardless of the boundary
behaviour of u on such arcs. The behaviour of the flux of (ν3 )u on geodesic arcs of the
boundary is established in the following Lemma.

Lemma 1. Let D be a domain and let A be a geodesic arc of the boundary of D.


(i) Let u : D−→R be a minimal solution such that u|A = ∞. Then
Z
(ν3 )u ds = ||A||.
A

(ii) Let {un } be a sequence of minimal solutions in D continuous in D ∪ A. If {un }


diverges uniformly to infinity on compact subsets of A and remains uniformly bounded
in compact subsets of D, then
Z
lim (ν3 )n ds = ||A||,
n−→∞ A

where (ν3 )n is the third component of the unit conormal to the boundary of the graph
of un .
If {un } diverges uniformly to infinity in compact subsets of D and remains uniformly
bounded on compact subsets of A, then
Z
lim (ν3 )n ds = −||A||.
n−→∞ A

Proof. (i) First we prove that the tangent plane to S at points (z, u(z)) with z near
to A is almost
p vertical. Let N = (N1 , N2 , N3 ) be the upward unit normal to S, then
|(ν3 )u | = 1 − N32 at boundary points. We extend ν3 to the interior points of D

15
p
by setting |(ν3 )u | = 1 − N32 and choosing the sign that makes ν3 continuous at the
boundary (where it is already defined).
At points where the tangent plane is almost vertical, N3 approaches zero, hence (ν3 )u
approaches one. In other words the tangent plane at points (z, u(z)) with z next to A
is almost vertical if and only for any ε > 0 there is a neighborhood of A in D such that

(10) |(ν3 )u | > 1 − ε

at each point of the neighborhood.


A minimal graph is stable, so one has Schoen’s curvature estimates for the surface S :
let p be a point of S and let D(p, R) be a disk contained in S centered at p of intrinsic
radius R, then

µ ¶
R
(11) |A(q)| ≤ κ ∀q ∈ D p,
2
where A is the second fundamental form of S and κ is an absolute constant (see [Sc]).
Now, assume by contradiction that there is a sequence of points {zm } in D approaching
A ( i.e. un (zm ) → ∞ as m → ∞) such that (10) does not hold. Then, there is a radius
R independent on m such that D(pm , R) ⊂ S, where pm = (zm , u(zm )). Hence, by the
curvature estimate (11), around each pm the surface S is a graph over a disk D(pm , r)
of the tangent plane at pm , and the graph has bounded distance from the disk D(pm , r).
The radius of the disk depends only on R, hence it is independent of m. It is clear that,
if zm is close enough to A, then the horizontal projection of D(pm , r) and thus of the
surface S is not contained in D. Contradiction. Hence (10) holds in a neighborhood of
A.
Now, fix ε > 0 and let δ ≤ ε. Let q1 , q2 be the points of A at distance δ from
the endpoints of A and call Aδ the subarc of A bounded by q1 , q2 . We construct a
neighborhood of Aδ in D (see Figure 8). Let τ be a horizontal geodesic orthogonal
to A passing through the mid-point of A. We can assume that τ passes through the
origin. For i = 1, 2, let αi be a horizontal geodesic through qi forming an angle of π2
with A. Finally, let β be the horizontal geodesic orthogonal to τ, at distance δ from
A and call q3 = β ∩ α1 , q4 = β ∩ α2 . Denote by Qδ the geodesic quadrilateral having
vertices at points q1 , q2 , q3 , q4 .
The form (ν3 )u ds is exact, hence:
Z Z
0= (ν3 )u ds + (ν3 )u ds.
Aδ Qδ \Aδ

Then, if ε is small enough, using (10) we obtain:


Z
(ν3 )u ds ≥ −2ε + (1 − ε)||Aδ ||.

16
q 01
0α 1
3
1
τ β q1

q Aδ
Qδ 02 α2
1
q4 0
1

Figure 8

Letting ε (and so δ) tend to zero yields:


Z
(ν3 )u ds ≥ ||A||.
A
The opposite inequality is obvious, so (i) follows.
For the proof of (ii) one makes the obvious modifications of the arguments in (i).
u
t
Let us prove another useful result.

Lemma 2. Let u : D−→R be a minimal solution continuous on a convex arc C of the


boundary of D. Then: Z
(ν3 )u ds < ||C||.
C

Proof. It is enough to prove the result for a closed subarc of C, say C̃. Then we can
assume that u is defined in a convex set D̃ with continuous boundary data. Denote by
v the solution of the Dirichlet problem in D̃ such that:

v|∂ D̃\C̃ = u, vC̃ = u + a


where a is a constant to be fixed later. Let w = v − u, then

w|∂ D̃\C̃ = 0, w|C̃ = a


By the maximum principle w is uniformly bounded in D̃.
Using Stokes’ theorem (together with a standard approximation argument at the points
of discontinuity) we obtain:
Z Z Z · µ ¶ µ ¶¸
v 1 u1 v 2 u2
w[(ν3 )u − (ν3 )v ]ds = w1 − + w2 − dx1 dx2 .
∂ D̃ D̃ τv τu τv τu

17
The argument of the last integral is equal to the following expression:
µ ¶ "µ ¶2 µ ¶2 µ ¶2 #
τu + τv u1 v 1 u2 v 2 1 1 1
− + − + − .
2 τu τv τu τv F τu τv

Hence, it is non negative and not identically zero in D̃. Then we have:
Z
a [(ν3 )u − (ν3 )v ]ds > 0.

choosing alternatively a = ±1 we obtain the result. u
t

Remark 7. We point out that the results of Lemma 1 and 2 hold for non convex
domains as well.

Proof of Step 2. We prove that the first condition in (9) is sufficient and necessary for
existence. We start by sufficiency.
Let un : D−→R be the minimal solution with the following boundary values

un |Ai = +n, un |Cs = min(n, f s ).


By Remark 6, {un } is uniformly bounded in compact sets contained in each of the
convex hulls C(Cs ), s = 1, . . . , h. Hence, passing to a subsequence, {un } converges on
compact subsets of
∪hs=1 C(Cs )
to a minimal solution u defined in an open set U containing ∪hs=1 C(Cs ). Furthermore
{un } diverges uniformly on compact subsets of D \ U and u is a countinuous function
with values in R ∪ ∞.
Let V = D \ U . We claim that V = ∅. We start by showing that ∂V has a very special
structure, when V is not empty.

Lemma 3. With the notation above, one has:


(i) ∂V consists only of geodesic chords of D and parts of the boundary of D;
(ii) two chords of ∂V cannot have a common endpoint;
(iii) the endpoints of chords of ∂V are among the vertices of the geodesic arcs Ai ;
(iv) a component of V cannot consist only of an interior chord of D.

Proof. It is clear by Remark 6 that each arc of ∂V must be geodesic and that no vertex
of ∂V lies in D, then (i) follows. Now assume by contradiction that (ii) does not hold.
Let K1 , K2 be two arcs of ∂V having a common endpoint q ∈ ∂D. Choose two points
q1 ∈ K1 and q2 ∈ K2 such that the triangle T with vertices q, q1 , q2 lies in D. We have:

18
Z
(ν3 )n ds = 0
∂T
where (ν3 )n is defined as in Lemma 1 at interior points of D. The triangle T may be
either in U or in V. Assume the former is true, then, by the first equality in (ii) of
Lemma 1, choosing correctly the orientation, we have:

Z Z
(12) lim (ν3 )n ds = ||qq1 ||, lim (ν3 )n ds = ||qq2 ||.
n→∞ qq1 n→∞ q2 q

Here ∗ indicates the geodesic arc between two points and (ν3 )n is defined as in Lemma
1 at interior points of D.
On the other hand:
¯Z ¯
¯ ¯
(13) ¯ (ν3 )n ds¯¯ ≤ ||q1 q2 ||.
¯
q1 q2

(12) and (13) together with the triangle inequality give a contradiction.
If T ⊂ V, we make the same reasoning using the second equality in (ii) of Lemma 1.
(iii) and (iv) are proved with analogous arguments , using Lemma 1. We leave this to
the reader.
u
t
Now, we come back to the proof of Step 2. Assume by contradiction that V is not
empty. The convex hull of each Ci is contained in U, and each component of V is
bounded by a geodesic polygon P, whose vertices are among the endpoints of the Ai .
Denote by ÂP i those edges of Ai that are contained in P. In the notation of Theorem
3, ||P|| = γ, ||Âi || = α.
For each un we have:
Z Z Z
0 = (ν3 )n ds = (ν3 )n ds + (ν3 )n ds.
P ∪Âi P\∪Âi

By (ii) of Lemma 1, we infer:


Z
lim (ν3 )n ds = −(γ − α).
n→∞ P\∪Âi

For every n, |(ν3 )n | < 1, so


¯Z ¯
¯ ¯ X
¯ (ν3 )n ds¯¯ ≤ ||Âi || = α.
¯
∪Âi
Hence α ≥ γ − α, that contradicts the assumed conditions.
We are left with the proof of the necessity of the condition 2α < γ. Let u be the
minimal solution with the given boundary values and let P be a polygon as in the
hypothesis of Theorem 3 . We have:

19
Z Z
(ν3 )u ds + (ν3 )u ds = 0.
∪Âi P\∪Âi

Furthermore |(ν3 )u | < 1 on P \ ∪Âi , hence


¯Z ¯
¯ ¯
¯ (ν ) ds¯<γ−α
¯ 3 u ¯
P\∪Âi

and by (i) of Lemma 1, we have:


Z
(ν3 )u ds = α
∪Âi
Hence 2α < γ.
ut
Proof of Step 3. Let {un } be defined as in Step 2. First we prove that {un } is bounded
at some point of D. Assume that this is not the case, then V = D and we have:
Z Z
0= (ν3 )n ds + (ν3 )n ds.
∪Ai ∪Ci

(ii) of Lemma 1 implies


Z X
lim (ν3 )n ds = − ||Ci || ≤ −(γ − α).
n→∞ ∪Ci
P
Here γ = ||∂D|| and α = ||Ai ||.
On the other hand, as |(ν3 )n | < 1 for every n, we have
¯Z ¯ X
¯ ¯
¯ (ν ) ds¯< ||Ai || = α.
¯ 3 n ¯
∪Ai
Then α ≥ γ − α, a contradiction.
Hence the sequence {un } is bounded at some point of D. In fact we will prove that
there is a disk in D of radius independent on n where each un is uniformly bounded.
Up to an isometry, we can assume that {un } is bounded at the origin σ ∈ H2 . We
remark that by the maximum principle each un is positive in the domain of definition.
Let mn = un (σ). We assert that the gradient of un at σ is bounded depending only on
the constant mn . In order to prove it, we will compare the gradient of un with that of
a Scherk type surface.
Up to a rotation of x1 , x2 coordinates, we can assume that
∂un ∂un
(σ) > 0, (σ) = 0.
∂x1 ∂x2
Let ∆(n) be a geodesic triangle contained in D with edges a, b, c such that the x1 axis
bisects the edge a orthogonally and ∆(n) is symmetric with respect to the x1 axis (see
Figure 9).

20
D
1
0
b

σ a
1
0

c
1
0

Figure 9

Let ϕ∆(n) denote the Scherk type surface over ∆(n) with value +∞ on a, value 0 on b,
c and ϕ∆(n) (σ) = mn (we allow translations of ∆(n) along the x1 axis in order to find
such a Scherk type surface). Define C(mn ) = |∇ϕ∆(n) (σ)|. We claim that:

(14) |∇un (σ)| ≤ C(mn ).

In fact, assume by contradiction that (14) does not hold. Then, the symmetries of
ϕ∆(n) imply:

∂un ∂ϕ∆(n) ∂un ∂ϕ∆(n)


(σ) > (σ), (σ) = (σ) = 0.
∂x1 ∂x1 ∂x2 ∂x2
Now, we move ∆(n) by hyperbolic translations along the x1 axis, pushing the edge a
∂ϕ
towards σ. As ϕ∆(n) and ∂x∆(n)
1
diverge as one approaches the side a, there is a position
of ∆(n) such that:

un (σ) < ϕ∆(n) (σ)


and

∂un ∂ϕ∆(n) ∂un ∂ϕ∆(n)


(σ) = (σ), (σ) = (σ) = 0.
∂x1 ∂x1 ∂x2 ∂x2
Define w = ϕ∆(n) − un . We have:

w(σ) = χ > 0, ∇w(σ) = 0.


Then, there are at least four level lines of w = χ through σ ([CM],[Se]). These level lines
divide every small neighborhood of σ in at least four domains in which w is alternately

21
C−

w= χ

C+

Figure 10

greater than and less than χ. We prove that this yields a contradiction (our argument
is analogous to [Se], we give it for the sake of completeness).
Let G be the subset of ∆(n) whose points are at distance less than ε from the boudary
of ∆(n). The function un has bounded continuous gradient in ∆(n), hence, using the
form of the graph of ϕ∆(n) , one has that the set G is divided into two components by
the conditions

w > χ, w < χ,
for suitably small ε.
The first component is adjacent to edge a, while the second is adjacent to b and c and
the components themselves are separated by two level lines w = χ exiting from the
vertices of a. By the maximum principle, each component of the set w > χ must extend
to the boundary of ∆(n). It follows that the set w > χ consists of one component. Then,
any two regions near σ where w > χ can be joined by a simple Jordan arc C + along
which w > χ. Analogously any two regions near σ where w < χ can be joined by a
simple Jordan arc C − along which w < χ. Then, the curves C+ and C− must intersect
(see Figure 10). This is a contradiction.
Now let M = supn∈IN un (σ). As mn ≤ M, with the same reasoning used for proving
(14), one has that C(mn ) ≤ C(M ). Hence the sequence {un } has uniformly bounded
gradient at σ.
Now, by Schoen’s curvature estimates (see (11)), one has that in a neighborhood of
the point pn = (σ, un (σ)) the surface is a graph of bounded height and slope over a
disk D(pn , R) of the tangent plane to the surface at pn , of radius R independent of n.
As |∇un (σ)| is uniformly bounded, the projection of each D(pn , R) on the horizontal
plane contains a disk of fixed radius and {un } is uniformly bounded there. Then, there
exists an open set U in which {un } converges uniformly. Now we can apply the same
reasoning as in Step 2, in order to prove that U = D.
u
t
+
Proof of Step 4. By the previous arguments we can find a minimal solution u : D−→R
such that

u+ |Ai = ∞, u+ |Bj = 0, u+ |Cs = max{0, f s }.

22
Furthermore we can find a minimal solution u− : D−→R such that

u− |Ai = 0, u− |Bj = −∞, u− |Cs = min{0, f s }.


½
2x − y = 0
y + kz = 0
Then , define for each s :

 −n if f s < −n
s
(f )n = f s if |f s | ≤ n

n if f s > n
and let un : D−→R be the minimal solution such that

un |Ai = n, un |Bj = −n, un |Cs = (f s )n .


By the maximum principle:

u− ≤ un ≤ u+ in D.
Hence, the sequence {un } is uniformly bounded on compact subsets of D and there ex-
ists a subsequence converging to a minimal solution that takes the prescribed boundary
values.
This proves existence under condition (9). The necessity of condition (9) is proved as
in Step 2.
ut
Proof of Step 5. We remark that in this case the number of edges Ai is equal to the
number of edges Bj , say k. We need to construct some auxiliary sets and minimal
solutions.
Let vn : D−→R be the minimal solution such that

vn|Ai = n, vn|Bi = 0.
For c ∈]0, n[, we introduce the following subsets of D :

Ec = {vn > c} ∩ D, Fc = {vn < c} ∩ D.


Let Eci be the component of Ec whose closure contains the edge Ai and let Fcj be
the component of Fc whose closure contains the edge Bj . By the maximum principle
Ec = ∪ki=1 Eci and Fc = ∪ki=1 Fci . We choose c close enough to n such that the Eci are
disjoint and we define:

µ(n) = lim sup{c ∈]0, n[ | Eci ∩ Ecj = ∅ i 6= j}.

23
Of course there is at least one pair i, j such that

E i µ(n) ∩ E j µ(n) 6= ∅
i j
and this implies that for any given Fµ(n) , the set Fµ(n) is disjoint from it.
For each n, we define the following minimal solution in D :

un = vn − µ(n).

In order to prove that the sequence {un } is uniformly bounded on compact subsets of
D, let us define two auxiliary minimal solutions in D.
Let ui + and ui − be the minimal solutions in D with the following boundary values:

ui + |Ai = ∞, ui + |∂D\Ai = 0

ui − |Bj = −∞, j 6= i, ui − |∂D\∪j6=i Bj = 0


For every i = 1, . . . , k, ui + and ui − exist by previous steps.
Finally, for any z ∈ D we define:

u+ (z) = max {ui + (z)}, u− (z) = min {ui − (z)}.


1≤i≤k 1≤i≤k

We claim that at any point of D :

(15) u− ≤ un ≤ u+ .
i i
Let p ∈ D such that un (p) > 0, then p belongs to Eµ(n) for some i. On ∂Eµ(n) one has
+ i
un ≤ ui , then this inequality holds in Eµ(n) , and

un (p) ≤ u+ +
i (p) ≤ u (p).

Since u− is non positive, the left inequality in (15) is obvious at the point p.
i
The proof of (15) at points where un is negative is analogous, using the set Fµ(n) .
Hence {un } has a subsequence converging to a minimal solution u : D−→R. Let us
prove that u takes the right boundary values.
Recall that:

un |Ai = n − µ(n), un |Bi = −µ(n),


so we must prove that the sequences {µ(n)} and {n − µ(n)} both diverge to infinity.
We prove it for the sequence {µ(n)}; the proof will be analogous for the latter sequence.
The assumption that {µ(n)} does not diverge will give a contradiction to the hypothesis
α = β. By contradiction, take a subsequence (still denoted by {un }) such that µ(n)
tends to a finite limit µ0 . Then:

24
un −→∞ on Ai , un −→ − µ0 on Bi .
Then, for the limit function u we have:

u|Ai = ∞, u|Bi = −µ0 .


Let (ν3 )u be the unit inward conormal to the boundary of the graph of u. We finally
obtain:
Z Z
α= (ν3 )u ds = − (ν3 )u ds > −β,
∪Ai ∪Bi

where the first equality is given by (i) of Lemma 1. This is a contradiction.


The necessity of the condition α = β is proved as in Step 2. u
t
Proof of Step 6 (Uniqueness). Let u and v be two minimal solutions assuming values
+∞ on each Ai , −∞ on each Bj and the same data on each geodesic arc Cs .
Let M be a large constant and define:

 −M if u − v < −M
ψ= u − v if |u − v| < M

M if u − v > M
Let 0 < δ < ε and denote by Dδε the subset of D whose distance from ∂D is greater
than δ and whose distance from the vertex of each Ai and Bj is greater than ². It is
clear that the boundary Γ of Dδε consists of bounded arcs Ãi , B̃j , C̃s adjacent to the
Ai , Bj , Cs and circular arcs adjacent to the vertices of Ai , Bj . Let (ν3 )u and (ν3 )v be
defined as in Lemma 1 for the functions u and v respectively. Consider the following
integral:
Z
ψ[(ν3 )u − (ν3 )v ]ds.
Γ
For δ small enough, we have:
Z Z X
ψ[(ν3 )u − (ν3 )v ]ds ≤ 2 |ψ| ≤ 2ε ||Cs ||.
∪C̃s ∪C̃s

We recall that next to boundary arcs where the solution is infinity, |ν3 | is almost one,
hence:

Z Z Z X
ψ[(ν3 )u − (ν3 )v ]ds = ψ[(ν3 )u − 1]ds − ψ[(ν3 )v − 1]ds ≤ 2εM ||Ãi ||
∪Ãi ∪Ãi ∪Ãi

In the same way we obtain:


Z X
ψ[(ν3 )u − (ν3 )v ]ds ≤ 2εM ||B̃j ||
∪B̃j

25
By summing the previous inequalities, we infer:

Z X
(16) ψ[(ν3 )u − (ν3 )v ]ds ≤ 2ε (||Cs || + M ||Ãi || + M ||B̃j ||) + 4πsinh(ε)M (k + l)
Γ

where the last term is the contribution of the circular arcs next to vertices.
On the other hand, integrating by parts, we obtain:
Z Z Z · µ ¶ µ ¶¸
v 1 u1 v2 u2
ψ[(ν3 )u − (ν3 )v ]ds = ψ1 − + ψ2 − dx1 dx2
Γ τv τu τv τu

where the double integral is taken over the set Dδε ∩ {|u − v| < M }. The argument of
the last integral is non-negative and it is zero only at points where ∇u = ∇v (see the
proof of Lemma 2).
Then, letting ε−→0 in (16), we obtain:
µ ¶ µ ¶
v1 u 1 v 2 u2
ψ1 − + ψ2 − =0
τv τu τv τu
at points where |u − v| < M. Hence ∇u = ∇v in D, since M can be taken arbitrarily
large.
It follows that u = v + const in D. If the family of boundary convex arcs {Cs } is empty,
this proves the result. If not, the constant must be zero by the boundary condition on
Cs . u
t

6 Existence of Complete Minimal Graphs

Theorem 4. Let Γ be a continuous rectifiable Jordan curve in ∂∞ H2 × R, that is


a vertical graph. Then, there exists a minimal vertical graph on H2 having Γ as
asymptotic boundary. The graph is unique.

Proof. We claim that, if we prove the Theorem for a C ∞ curve Γ, then, the result is
true for a continuous rectifiable Γ. Let us prove it.
Consider two families of differentiable curves approximating Γ, constructed as follows.
For any ε > 0, let Γ+ 2
ε ⊂ ∂∞ H × R be a differentiable vertical graph above Γ, such that
the vertical distance between Γ and Γ+ − 2
ε is at most ε. Then, let Γε ⊂ ∂∞ H × R be a
differentiable vertical graph below Γ, such that the vertical distance between Γ and Γ−
ε
is at most ε.
For any ε > 0 there exist a minimal graph Mε+ ( Mε− ) with asymptotic boundary Γ+ ε
(resp. Γ−ε ).

26
By the maximum principle, each surface Mn in the proof of Theorem 4 in [?] is above
Mε− and below Mε+ , hence also the limit surface M. Now, let ε go to zero: as both Γ+ ε
and Γ− ε converge to the curve Γ, the boundary of M must be Γ.
Then, assume that Γ is C ∞ .
In the model D = {0 ≤ x21 + x22 < 1} for H2 , the curve Γ is a graph over the circle
x21 + x22 = 1. Consider an exhaustion of D by disks Dn centered at the origin, of
Euclidean radius 1 − n1 . For each n, let Γn be a vertical C 2 graph over ∂Dn converging
to Γ as n−→∞. We choose the curves Γn contained in the convex hull of Γ. The
curves Γn may be taken as the trace on ∂Dn × R of the function whose graph is a C 2
extension of Γ inside D. Let Mn be the Plateau solution with boundary Γn ; by Rado’s
theorem Mn is a vertical graph of a C 2 function vn : Dn −→R. The sequence {vn } is
uniformly bounded on compact subsets of D, hence there is a subsequence converging
to a minimal solution v : D−→R, uniformly on compact subsets of D. Let M be the
graph of the function v. We have only to prove that the asymptotic boundary of M is
Γ. By definition of M, one has that Γ ⊂ ∂∞ M. In order to prove the converse, we show
that any point p 6∈ Γ is not contained in ∂∞ M. Let p such a point and assume that
it lies above Γ (the reasoning is analogous, when p lies below Γ). We use a suitable
Abresch-Sa Earp graph as barrier. One can choose α such that a suitable translation
of the relative Abresch-Sa Earp graph has boundary above the boundary of Mn for n
large, and its asymptotic boundary is below p. Hence, by the maximum principle, such
Abresch-Sa Earp graph lies above Mn . Hence the asymptotic boundary of M (that is
the limit of the sequence {Mn }) does not contain the point p.
Uniqueness follows from the maximum principle. Theorem 4 is proved.
ut

7 Isolated Singurarities
The original idea of the proof of the following Theorem is from [11]. A removable
singularities Theorem analogous to Theorem 5, for prescribed mean curvature graphs
in Euclidean space and hyperbolic space is proved in [8] and [4] respectively.

Theorem 5. Let Dρ∗ be a punctured disk of radius ρ in H2 . Let u : Dρ∗ −→R be a C 2


function such that the graph of u is minimal in H2 × R. Then u extends C 2 to the
puncture.

Proof. First we prove that u is bounded on Dρ∗ . Up to isometry, we can take Dρ∗ centered
at the origin of H2 . Let ε < ρ and consider the annulus Aε = Dρ \ Dε . Let Cε the half
catenoid that is a graph over H2 \ D² with waist on ∂Dε . The function u is bounded
on Aε , hence there is a vertical translation of Cε that does not touch the graph of u.
Now translate vertically Cε towards the graph of u. By the maximum principle the first
contact point is on ∂Dρ × R.
Letting ²−→0, Cε tends to a plane, hence:

27
min u ≤ u ≤ max u
∂Dρ ∂Dρ

Now, let v : Dρ −→R be the minimal solution with boundary values u|∂Dρ . We will show
that u ≡ v in Dρ∗ .
Consider the form θ defined in Dρ∗ by:
µ ¶ µ ¶
u1 v 1 u 2 v2
θ = (u − v) − dx2 − − dx1
τu τv τu τv

We have: Z Z Z
θ= θ= dθ
∂Dε ∂Aε Aε

where the first equality depends on the fact that u ≡ v on ∂Dρ and the last equality is
by Stokes’ theorem.
The form θ is bounded on Dρ∗ because u, |ui |τu −1 , |vi |τv −1 are bounded for i = 1, 2.
Then:

Z
(9) θ−→0 as ε−→0
∂Aε

As in Lemma 2, we obtain that dθ is non negative and it is 0 if and only if ui = vi , for


i = 1, 2. Letting ε−→0 we obtain ∇u ≡ ∇v and so u ≡ v on Dρ∗ . u
t

References
[1] T. H. Colding, W. P. Minicozzi: Minimal surfaces, Courant Lecture Notes
in Mathematics, 4, New York University (1999).

[2] M. D. Duong , V. H. Nguyen: Graphs with Prescribed Mean Curvature on


Poincaré Disk, Bull. London Math. Soc. 27 (1995) 353-358.

[3] H. Jenkins, J. Serrin: Variational Problems of Minimal Surface Type II.


Boundary Value Problems for the Minimal Surface Equation, Arch. Rational Mech.
Anal. 21 (1966), 321-342.

[4] B. Nelli, R. Sa Earp: Some Properties of surfaces of prescribed mean curvature


in Hn+1 , Bul. Sc. Math. 120 (1996) 537-553.

[5] B. Nelli, R. Sa Earp, W. Santos, E. Toubiana: Existence and Uniqueness


of H-surfaces with one or two parallel convex curves as boundary, in H2 × R, Ann.
of Gl. Analysis and Geometry 33 (4) (2008) 307-321.

28
[6] R. Pedrosa, M. Ritoré: Isoperimetric Domains in the Riemannian product of
a Circle with a simply connected Space Form and Applications to Free Boundary
Value Problems, Indiana Univ. Math. Jour. 48 (1999), 1357-1349.

[7] H. Rosenberg: Minimal Surfaces in M 2 × R, Ill. Jour. of Math. 46 (4) (2002)


1177-1195.

[8] H. Rosenberg, R. Sa Earp: Some Remarks on Surfaces of Prescribed Mean


Curvature, Pitman Monographs and Surveys in Pure and Applied Mathematics
52,(1991), 123-148.

[9] R. Sa Earp: Parabolic and hyperbolic screw motion surfaces in H2 × R, Journal


of the Australian Mathematical Society, 85 (1) (2008) 113-143.

[10] R. Schoen: Estimates for Stable Minimal Surfaces in Three Dimensional Man-
ifolds, Seminar on Minimal Submanifolds, Princeton University Press (1983), 111-
126.

[11] J. Serrin: A Priori Estimates for Solutions of Minimal surface Equation, Arch.
Rational Mech. Anal. 14 (1963), 376-383.

[12] R. Sa Earp, E. Toubiana: Screw Motion Surfaces in H2 × R and S2 × R,


Illinois Jour. of Math. 49 n.4 (2005) 1323-1362.

Barbara Nelli
Universitá di L’Aquila
nelli@univaq.it
Harold Rosenberg
IMPA- Rio de Janeiro
rosenbergharold@gmail.com

29

You might also like