Tobias H. Colding and William P. Minicozzi Ii
Tobias H. Colding and William P. Minicozzi Ii
Contents
1. Introduction 2
The authors were partially supported by NSF Grants DMS 0104453 and DMS 0405695.
1
2 TOBIAS H. COLDING AND WILLIAM P. MINICOZZI II
1. Introduction
In this article, we survey some old and new results about minimal surfaces and
submanifolds. The field of minimal surfaces has its origin in the mid eighteenth
century with the work of Euler and Lagrange but it has very recently seen major
advances that have solved many long standing open conjectures in the field. In what
follows, we give a quick tour through the field, starting with the definition and the
classical results and ending up with current areas of research. Many references are
given for further reading.
Soap films, soap bubbles, and surface tension were extensively studied by the
Belgian physicist and inventor (the inventor of the stroboscope) Joseph Plateau in
the first half of the nineteenth century. At least since his studies, it has been known
that the right mathematical model for soap films are minimal surfaces – the soap
film is in a state of minimum energy when it is covering the least possible amount
of area. Minimal surfaces and equations like the minimal surface equation have
served as mathematical models for many physical problems.
The field of minimal surfaces dates back to the publication in 1762 of Lagrange’s
famous memoir “Essai d’une nouvelle méthode pour déterminer les maxima et les
minima des formules intégrales indéfinies”. Euler had already in a paper published
in 1744 discussed minimizing properties of the surface now known as the catenoid,
but he only considered variations within a certain class of surfaces. In the almost
one quarter of a millennium that has past since Lagrange’s memoir minimal surfaces
has remained a vibrant area of research and there are many reasons why. The study
of minimal surfaces was the birthplace of regularity theory. It lies on the intersection
of nonlinear elliptic PDE, geometry, topology and general relativity.
where H is the mean curvature (vector) of Σ. Here, and throughout this paper,
integration is with respect to dvol. (When Σ is noncompact, then Σt,Φ in (1.2) is
MINIMAL SUBMANIFOLDS 3
replaced by Γt,Φ , where Γ is any compact set containing the support of Φ.) The
submanifold Σ is said to be a minimal submanifold (or just minimal) if
d
(1.3) Vol(Σt,Φ ) = 0 for all Φ ∈ C0∞ (N Σ)
dt t=0
d2
Z
(1.4) Vol(Σt,φnΣ ) = − φ LΣ φ ,
dt2 t=0 Σ
where
(1.5) LΣ φ = ∆Σ φ + |A|2 φ
1.1. The Gauss map. Let Σ2 ⊂ R3 be a surface (not necessarily minimal). The
Gauss map is a continuous choice of a unit normal n : Σ → S2 ⊂ R3 . Observe that
there are two choices n and −n corresponding to a choice of orientation of Σ. If Σ
is minimal, then the Gauss map is an (anti) conformal map since the eigenvalues
of the Weingarten map are κ1 and κ2 = −κ1 . Moreover, for a minimal surface
where KΣ is the Gauss curvature. It follows that the area of the image of the Gauss
map is a multiple of the total curvature.
4 TOBIAS H. COLDING AND WILLIAM P. MINICOZZI II
This shows that Graphu is area-minimizing among all surfaces in the cylinder and
with the same boundary.
If the domain Ω is convex, the minimal graph is absolutely area-minimizing.
To see this, observe first that if Ω is convex, then so is Ω × R, and hence the
nearest point projection P : R3 → Ω × R is a distance nonincreasing Lipschitz
map that is equal to the identity on Ω × R. If Σ ⊂ R3 is any other surface with
∂Σ = ∂ Graphu , then Σ′ = P (Σ) has Area(Σ′ ) ≤ Area(Σ). Applying (1.20) to Σ′ ,
we see that Area(Graphu ) ≤ Area(Σ′ ) and the claim follows.
If Ω ⊂ R2 contains a ball of radius r, then, since ∂Br ∩ Graphu divides ∂Br
into two components at least one of which has area at most (Area(S2 )/2) r2 , we get
from (1.20) the crude estimate
Area(S2 ) 2
(1.21) Area(Br ∩ Graphu ) ≤ r .
2
Very similar calculations to the ones above show that if Ω ⊂ Rn−1 and u : Ω → R
is a C 2 function, then the graph of u is a critical point for the area functional if
and only if u satisfies (1.14). Moreover, as in (1.20), the graph of u is actually
area-minimizing. Consequently, as in (1.21), if Ω contains a ball of radius r, then
Vol(Sn−1 ) n−1
(1.22) Vol(Br ∩ Graphu ) ≤ r .
2
1.3. The maximum principle. The first variation formula, (1.2), showed that a
smooth submanifold is a critical point for area if and only if the mean curvature
vanishes. We will next derive the weak form of the first variation formula which
is the basic tool for working with “weak solutions” (typically, stationary varifolds).
Let X be a vector field on Rn . We can write the divergence div Σ X of X on Σ as
(1.23) div Σ X = div Σ X T + div Σ X N = div Σ X T + hX, Hi ,
where X T and X N are the tangential and normal projections of X. In particular,
we get that, for a minimal submanifold,
(1.24) div Σ X = div Σ XT .
Moreover, from (1.23) and Stokes’ theorem, we see that Σ is minimal if and only if
for all vector fields X with compact support and vanishing on the boundary of Σ,
Z
(1.25) div Σ X = 0 .
Σ
The key point is that (1.25) makes sense as long as we can define the divergence of
a vector field on Σ.
The two most common notions of weakly minimal submanifolds are minimal
currents and stationary varifolds. A k-varifold Σ on Rn is a Radon measure on
Rn × G(k, n), where G(k, n) is the space of (unoriented) k-planes through the
6 TOBIAS H. COLDING AND WILLIAM P. MINICOZZI II
For example, L. Simon, [97], proved that the strong maximum holds when both
hypersurfaces are area-minimizing even in the presence of singularities. B. Solomon
and B. White, [100], showed that it holds when at least one of the hypersurfaces
is smooth (the other may be just a stationary varifold). Finally, T. Ilmanen, [48],
showed that it holds for stationary hypersurfaces as long as each singular set has
zero codimension two measure.
Thus far, the examples of minimal submanifolds have all been smooth. The
simplest non-smooth example is given by a pair of planes intersecting transversely
along a line. To get an example that is not even immersed, one can take three
half–planes meeting along a line with an angle of 2π/3 between each adjacent pair.
When we apply this formula below, h will be Euclidean distance to a fixed point
x0 .
Proposition 2.1. Suppose that Σk ⊂ Rn is a minimal submanifold and x0 ∈ Rn ;
then for all 0 < s < t
|(x − x0 )N |2
Z
(2.2) t−k Vol(Bt (x0 )∩Σ)−s−k Vol(Bs (x0 )∩Σ) = k+2
.
(Bt (x0 )\Bs (x0 ))∩Σ |x − x0 |
Using this and the coarea formula (i.e., (2.1)), an easy calculation gives
d −k |x − x0 |
Z
−k−1 −k
s Vol(Bs ∩ Σ) = −k s Vol(Bs ∩ Σ) + s
ds ∂Bs ∩Σ |(x − x0 )T |
2
|x − x0 |
Z
−k−1 T
(2.5) =s − |(x − x0 ) |
∂Bs ∩Σ |(x − x0 )T |
|(x − x0 )N |2
Z
= s−k−1 T
.
∂Bs ∩Σ |(x − x0 ) |
Integrating and applying the coarea formula once more gives the claim.
Notice that (x − x0 )N vanishes precisely when Σ is conical about x0 , i.e., when
Σ is invariant under dilations about x0 . As a corollary, we get the following:
8 TOBIAS H. COLDING AND WILLIAM P. MINICOZZI II
t
|(x − x0 )N |2 1
Z Z Z
= f + τ −k−1 (τ 2 − |x − x0 |2 ) ∆Σ f dτ .
(Bt (x0 )\Bs (x0 ))∩Σ |x − x0 |k+2 2 s Bτ (x0 )∩Σ
We get immediately the following mean value inequality for the special case of
non–negative subharmonic functions:
Corollary 2.4. Suppose that Σk ⊂ Rn is a minimal submanifold, x0 ∈ Rn , and f
is a non–negative subharmonic function on Σ; then
Z
(2.8) s−k f
Bs (x0 )∩Σ
3. Rado’s theorem
One of the most basic questions is what does the boundary ∂Σ tell us about
a compact minimal submanifold Σ? We have already seen that Σ must lie in the
convex hull of ∂Σ, but there are many other theorems of this nature. One of the
first is a beautiful result of T. Rado which says that if ∂Σ is a graph over the
boundary of a convex set in R2 , then Σ is also graph (and hence embedded). The
proof of this uses basic properties of nodal lines for harmonic functions.
MINIMAL SUBMANIFOLDS 9
5. Simons inequality
In this section, we recall a very useful differential inequality for the Laplacian of
the norm squared of the second fundamental form A of a minimal hypersurface Σ
in Rn and illustrate its role in a priori estimates. This inequality, originally due to
J. Simons (see [14] for a proof and further discussion), is:
MINIMAL SUBMANIFOLDS 11
then
(5.3) |A|2 (0) ≤ r−2 .
Proof. (Sketch.) Observe first that it suffices to prove the estimate for σ = r0 , i.e.,
to show that
(6.2) |A|2 (0, u(0)) ≤ C r0−2 .
Recall that minimal graphs are automatically stable. As in the proof of Theorem
4.1, the area estimate for graphs (1.21) allows us to use a logarithmic cutoff function
in the the stability inequality (1.7) to get that
C
Z
(6.3) |A|2 ≤ .
Br1 ∩Graphu log(r0 /r1 )
Taking r0 /r1 sufficiently large, we can then apply Theorem 5.2 to get (6.2).
Heinz’s estimate gives an effective version of the Bernstein’s theorem; namely,
letting the radius r0 go to infinity in (6.1) implies that |A| vanishes, thus giving
Bernstein’s theorem.
then
(7.6) sup |A|2 ≤ CP s−2 .
Bs
We will mention two such estimates. The first is R. Schoen’s curvature estimate
for stable surfaces:
Theorem 8.1. [88] There exists a constant C so that if Σ ⊂ R3 is an immersed
stable minimal surface with trivial normal bundle and Br0 ⊂ Σ \ ∂Σ, then
(8.2) sup |A|2 ≤ C σ −2 .
Br0 −σ
The second is an estimate for the area and total curvature of a stable surface;
for simplicity, we will state only the area estimate:
Theorem 8.2. [15] If Σ ⊂ R3 is an immersed stable minimal surface with trivial
normal bundle and Br0 ⊂ Σ \ ∂Σ, then
(8.3) Area (Br0 ) ≤ 4π r02 /3 .
As mentioned, we can use (8.3) to bound the energy of a cutoff function in the
stability inequality and, thus, bound the total curvature of sub-balls. Combining
this with the curvature estimate of Theorem 5.2 gives Theorem 8.1; see [15]. Note
that the bound (8.3) is surprisingly sharp; even when Σ is a plane, the area is πr02 .
We will explain next why Theorem 8.2 holds. The stability inequality can be
used to get upper bounds for the total curvature in terms of the area of a minimal
surface. On the other hand, we can use either the Gauss–Bonnet theorem or the
Jacobi equation to get the opposite bound. Combining these two bounds will give
the a priori bound on the area of intrinsic balls in a stable surface. More precisely,
integrating the Jacobi equation and using the Gauss equation KΣ = −|A|2 /2 gives
Z R Z tZ Z
2 2
(8.4) 4 (Area (BR ) − π R ) = 2 |A| = |A|2 (R − r)2 .
0 0 Bs BR
RR
The second equality uses two integrations by parts (i.e., 0 f (t) g ′′ (t) dt with f (t) =
RtR 2 2
0 Bs |A| and g(t) = (R − t) ). See corollary 1.7 of [17] for further details.
Using φ = R − r (where r(x) = distΣ (0, x)) in the stability inequality gives
Z Z
2 2 2
(8.5) 4 (Area (BR ) − π R ) = |A| (R − r) ≤ |∇(R − r)|2 = Area (BR ) .
BR BR
2
Consequently, Area (BR ) ≤ 4π R /3.
9. Regularity theory
In this section, we survey some of the key ideas in classical regularity theory,
such as the role of monotonicity, scaling, ǫ-regularity theorems (such as W. Allard’s
theorem) and tangent cone analysis (such as F. Almgren’s refinement of H. Federer’s
dimension reducing).
The starting point for all of this is the monotonicity of volume for a minimal
k–dimensional submanifold Σ. Namely, Corollary 2.2 gives that the density
Vol(Bs (x0 ) ∩ Σ)
(9.1) Θx0 (s) =
Vol(Bs ⊂ Rk )
is a monotone non-decreasing function of s. Consequently, we can define the density
Θx0 at the point x0 to be the limit as s → 0 of Θx0 (s). It also follows easily from
monotonicity that the density is semi–continuous as a function of x0 .
14 TOBIAS H. COLDING AND WILLIAM P. MINICOZZI II
9.2. Tangent cone analysis. It is not hard to see that scaling preserves the space
of minimal submanifolds of Rn . Namely, if Σ is minimal, then so is
(9.4) Σy,λ = {y + λ−1 (x − y) | x ∈ Σ} .
(To see this, simply note that this scaling multiplies the principal curvatures by
λ.) Suppose now that we fix the point y and take a sequence λj → 0. The mono-
tonicity formula bounds the density of the rescaled solution, allowing us to extract
a convergent subsequence and limit. This limit, which is called a tangent cone at
y, achieves equality in the monotonicity formula and, hence, must be homogeneous
(i.e., invariant under dilations about y).
The usefulness of tangent cone analysis in regularity theory is based on two
key facts. For simplicity, we illustrate these when Σ ⊂ Rn is an area minimizing
hypersurface. First, if any tangent cone at y is a hyperplane Rn−1 , then Σ is
smooth in a neighborhood of y. This follows easily from the Allard regularity
theorem since the density at y of the tangent cone is the same as the density at y
of Σ. The second key fact, known as “dimension reducing,” is due to F. Almgren,
[3], and is a refinement of an argument of H. Federer. To state this, we first stratify
the singular set S of Σ into subsets
(9.5) S0 ⊂ S1 ⊂ · · · ⊂ Sn−2 ,
where we define Si to be the set of points y ∈ S so that any linear space contained
in any tangent cone at y has dimension at most i. (Note that Sn−1 = ∅ by Allard’s
Theorem.) The dimension reducing argument then gives that
(9.6) dim (Si ) ≤ i ,
MINIMAL SUBMANIFOLDS 15
where dimension means the Hausdorff dimension. In particular, the solution of the
Bernstein problem then gives codimension 7 regularity of Σ, i.e., dim (S) ≤ n − 8.
See lecture 2 in [96] for a proof of (9.6).
Tangent cones produced by scalings as above may very well depend on the par-
ticular convergent subsequence. In some cases, one can prove uniqueness of the
tangent cone and this is often quite useful (see, for instance, section 3.4 in [96] for
one such application).
x3 -axis
x3-axis
u(ρ, θ + 2π)
w
u(ρ, θ)
10.2. The sublinear growth of the separation. As we have seen, the separation
is constant for the multi–valued graphs coming from each half of the helicoid. This
can be viewed as a type of Liouville Theorem reflecting the conformal properties of
an infinite–valued graph. In Proposition II.2.12 of [16], we proved a corresponding
gradient estimate (i.e., a pointwise estimate for |∇ log |w||) for a general multi-
valued graph. Integrating this gradient estimate gives that the separation grows
sublinearly (see figure 2):
Proposition 10.1. [16] Given α > 0, there exists N so that if u satisfies the
minimal surface equation on
(10.4) {e−N r1 ≤ ρ ≤ eN r2 , −N ≤ θ ≤ 2π + N } ,
|∇u| ≤ 1, and has separation w 6= 0, then
α
r2
(10.5) |w|(r2 , 0) ≤ |w|(r1 , 0) .
r1
Proof. (Sketch.) As mentioned above, the inequality (10.5) follows from integrating
the gradient estimate
α
(10.6) |∇ log |w||(r, 0) ≤
r
in the same way that the Harnack inequality for positive harmonic functions follows
from integrating the gradient estimate.
MINIMAL SUBMANIFOLDS 17
To see why a gradient estimate like (10.6) holds, observe that u(·, ·) and its
2π–rotation u(·, · + 2π) are both solutions of the minimal surface equation and,
thus, the difference w is (almost) a positive solution of the linearized equation. The
linearized equation is itself a perturbation of the Laplace equation and it is not
difficult to get an estimate
C
(10.7) |∇ log |w||(r, 0) ≤ ,
r
for some constant C.
The point now is to show that if N is large, then we can make the constant C in
(10.7) small. For simplicity, suppose that w is actually a positive harmonic func-
tion on the region (10.4). Since harmonic functions are invariant under conformal
transformations, w ◦ ez is a positive harmonic function on the “rectangle”
(10.8) {(x + iy) | − N + log r1 < x < N + log r2 and |y| < N } .
In the extreme case when N = ∞, the positive harmonic function w ◦ ez is defined
on the entire plane and, hence, is constant by the Liouville theorem. It is not hard
to see that applying the gradient estimate on this rectangle when N is large, and
then translating this back to the original function w, gives (10.6).
Using the notion of multi–valued graphs, the lamination theorem (the main the-
orem of [19]), can now be stated:
Figure 3. Theorem 11.1 – the singular set, S, and the two multi–valued
graphs. First published in the Notices of the American Mathematical
Society in 2003, published by the American Mathematical Society.
Theorem 11.1. (Theorem 0.1 in [19]). See figure 3. Let Σi ⊂ BRi = BRi (0) ⊂ R3
be a sequence of embedded minimal disks with ∂Σi ⊂ ∂BRi where Ri → ∞. If
(11.1) sup |A|2 → ∞ ,
B1 ∩Σi
Bǫr0 x3 = 0
Br0
B2r0
x3
Rescaled catenoid.
x2
x1 x3 = 0
Figure 5. The catenoid given Figure 6. Rescaling the
by revolving x1 = cosh x3 catenoid shows that simply
around the x3 –axis. First pub- connected (and embedded) is
lished in the Notices of the needed in the one–sided curva-
American Mathematical Society ture estimate. First published
in 2003, published by the Amer- in the Notices of the American
ican Mathematical Society. Mathematical Society in 2003,
published by the American
Mathematical Society.
Note that the assumption in Theorem 11.2 that Σ is simply connected is crucial
as can be seen from the example of a rescaled catenoid. Recall that the catenoid is
the minimal surface in R3 given by
(11.3) (cosh s cos t, cosh s sin t, s)
20 TOBIAS H. COLDING AND WILLIAM P. MINICOZZI II
where s, t ∈ R; see figure 5. Under rescalings this converges (with multiplicity two)
to the flat plane; see figure 6. Likewise, by considering the universal cover of the
catenoid, one sees that embedded, and not just immersed, is needed in Theorem
11.2.
As an almost immediate consequence of Theorem 11.2 and a simple barrier ar-
gument we get that if in a ball two embedded minimal disks come close to each
other near the center of the ball then each of the disks are graphs. Precisely, this
is the following:
graph
graph
Corollary 11.3. (Corollary 0.4 in [19]). See figure 7. There exist c > 1, ǫ > 0 so
that the following holds:
Let Σ1 and Σ2 ⊂ Bcr0 ⊂ R3 be disjoint embedded minimal surfaces with ∂Σi ⊂
∂Bcr0 and Bǫ r0 ∩ Σi 6= ∅. If Σ1 is a disk, then for all components Σ′1 of Br0 ∩ Σ1
which intersect Bǫ r0
(11.4) sup |A|2 ≤ r0−2 .
Σ′1
Σ2i ⊂ R3 is uniformly locally simply connected (or ULSC) if for each compact subset
K of R3 , there exists a constant r0 > 0 (depending on K) so that for every x ∈ K,
all r ≤ r0 , and every surface Σi
(12.1) each connected component of Br (x) ∩ Σi is a disk.
For instance, a sequence of rescaled catenoids where the necks shrink to zero is not
ULSC, whereas a sequence of rescaled helicoids is.
Remark 12.1. If each component of the intersection of a minimal surface with a
ball of radius r0 is a disk, then so are the intersections with all sub–balls by the
convex hull property (see, e.g., lemma C.1 in [19]). Therefore, it would be enough
that (12.1) holds for r = r0 .
We will next describe briefly the case of ULSC sequences. See [20] for more on
this and for the general case of fixed genus.
We will assume here that the surfaces are not disks (the case of disks was dealt
with in the previous subsection). In particular, we will assume that for each i, there
exists some yi ∈ R3 and si > 0 so that
(12.2) some component of Bsi (yi ) ∩ Σi is not a disk.
Loosely speaking, the next result shows that when the sequence is ULSC (but
not simply connected), a subsequence converges to a foliation by parallel planes
away from two lines S1 and S2 ; see figure 10. The lines S1 and S2 are disjoint and
orthogonal to the leaves of the foliation and the two lines are precisely the points
where the curvature is blowing up. This is similar to the case of disks, except that
we get two singular curves for non-disks as opposed to just one singular curve for
disks.
Theorem 12.2. [20] Let Σi ⊂ BRi = BRi (0) ⊂ R3 be a sequence of compact
embedded minimal surfaces with fixed genus and with ∂Σi ⊂ ∂BRi where Ri → ∞.
Suppose that each Σi is ULSC and satisfies (12.2) with si = R > 1 and yi = 0. If
(12.3) sup |A|2 → ∞ ,
B1 ∩Σi
then there exists a subsequence Σj , two disjoint parallel lines S1 and S2 , and a
rotation of R3 so that:
22 TOBIAS H. COLDING AND WILLIAM P. MINICOZZI II
(A) For each 1 > α > 0, Σj \ (S1 ∪ S2 ) converges in the C α -topology to the
foliation {x3 = t} by parallel planes.
(B) supBr (x)∩Σj |A|2 → ∞ as j → ∞ for all r > 0 and x ∈ S1 ∪ S2 . (The
curvatures blow up along S1 and S2 .)
(Culsc ) Away from S1 ∪ S2 , each Σj consists of exactly two multi-valued graphs
spiraling together. Near S1 and S2 , the pair of multi-valued graphs form
double spiral staircases with opposite orientations at S1 and S2 . Thus,
circling only S1 or only S2 results in going either up or down, while a path
circling both S1 and S2 closes up (see figure 11).
(Dulsc ) S1 and S2 are orthogonal to the leaves of the foliation.
We should point out that the multi-valued graphs in (Culsc ) are defined over
the doubly-punctured plane; see figure 11. The multi-valued graphs considered
previously were defined over a plane punctured at just one point.
Despite the similarity of Theorem 12.2 to the case of disks, it is worth noting
that the results for disks do not alone give this theorem. Namely, even though the
ULSC sequence consists locally of disks, the compactness result for disks was in the
global case where the radii go to infinity. One might wrongly think that Theorem
12.2 could be proven using the results for disks and a blow up argument. However,
local examples constructed in [24] show the difficulty with such an argument.
For embedded minimal surfaces, S consists of two types of points. The first type
is roughly modelled on rescaled helicoids and the second on rescaled catenoids:
• A point y in R3 is in Sulsc if the curvature for the sequence Σi blows up at
y and the sequence is ULSC in a neighborhood of y.
• A point y in R3 is in Sneck if the sequence is not ULSC in any neighborhood
of y. In this case, a sequence of closed non-contractible curves γi ⊂ Σi
converges to y.
The sets Sneck and Sulsc are obviously disjoint and the curvature blows up at both,
so Sneck ∪ Sulsc ⊂ S. An easy argument shows that, after passing to a subsequence,
we can assume that
(12.5) S = Sneck ∪ Sulsc .
Note that Sneck = ∅ is equivalent to that the sequence is ULSC as is the case
for sequences of rescaled helicoids. On the other hand, Sulsc = ∅ for sequences of
rescaled catenoids.
We will show that every sequence Σi has a subsequence that is either ULSC or
for which Sulsc is empty. This is the next “no mixing” theorem. These two different
cases give two very different structures.
Theorem 12.3. [20] If Σi ⊂ BRi = BRi (0) ⊂ R3 is a sequence of compact em-
bedded minimal planar domains with ∂Σi ⊂ ∂BRi where Ri → ∞, then there is a
subsequence with either Sulsc = ∅ or Sneck = ∅.
The case where Sneck = ∅ was dealt with in the previous section.
Common for both the ULSC case and the case where Sulsc is empty is that the
limits are always laminations by flat parallel planes and the singular sets are always
closed subsets contained in the union of the planes. This is the content of the next
theorem:
Theorem 12.4. [20] Let Σi ⊂ BRi = BRi (0) ⊂ R3 be a sequence of compact
embedded minimal surfaces with fixed genus and with ∂Σi ⊂ ∂BRi where Ri → ∞.
24 TOBIAS H. COLDING AND WILLIAM P. MINICOZZI II
If
(12.6) sup |A|2 → ∞ ,
B1 ∩Σi
The simplest examples are the plane (where the total curvature is zero) and the
catenoid. In the case of the catenoid, the Gauss map gives a conformal diffeo-
morphism to the sphere punctured at the north and south poles. Since A is the
differential of the Gauss map, it follows that the catenoid has total curvature 8π.
The fundamental result for minimal surfaces with ftc is that they are all conformally
diffeomorphic to compact Riemann surfaces with a finite number of points removed.
Namely, we have the following result of R. Osserman (see [79]):
Theorem 13.1. Let Σ ⊂ R3 be a complete minimal immersion with finite total
curvature. Then:
(1) Σ is conformally diffeomorphic to a compact Riemann surface with a finite
set of points removed. Each point corresponds to an end of the surface.
(2) Σ is proper.
(3) The Weierstrass data extends across the punctures meromorphically. (See
Section 17 for the definition of the Weierstrass representation.)
(4) The total curvature is an integral multiple of 8π.
One application of this theorem is a classification of the ends of an embedded
minimal surface with finite total curvature. Namely, one can show that any such
end is asymptotic to either a plane or to half of a catenoid.
13.2. The uniqueness of the catenoid. Given the structure result, Theorem
13.1, it is natural to try to understand the space of minimal surfaces with finite
total curvature in terms of the genus and the number of ends. Two such results were
proven by R. Schoen and F. Lopez–A. Ros, respectively. The theorem of Schoen
says that the catenoid is the unique embedded minimal surface with finite total
curvature and exactly two ends.
Theorem 13.2. [89] If Σ ⊂ R3 is a connected embedded minimal surface with
finite total curvature and exactly two ends, then Σ is a catenoid.
The theorem of Lopez and Ros says that the catenoid and the plane are the
only embedded minimal surfaces with finite total curvature and genus zero. Here,
“genus zero” means conformal to the sphere with a finite set of points removed.
Theorem 13.3. [59] If Σ ⊂ R3 is an embedded minimal surface with finite total
curvature and genus zero, then Σ is a catenoid or a plane.
properly embedded minimal surface in R3 (recall that properness here means that
the intersection of Σ with any compact subset of R3 is compact).
We say that Σ has finite topology if it is homeomorphic to a closed Riemann
surface with a finite number of punctures; the genus of Σ is then the genus of this
Riemann surface and the number of punctures is the number of ends. It follows
that a neighborhood of each puncture corresponds to a properly embedded annular
end of Σ. Perhaps surprisingly at first, the more restrictive case is when Σ has
more than one end. The reason for this is that a barrier argument gives a stable
minimal surface between any pair of ends. Such a stable surface is then asymptotic
to a plane (or catenoid), essentially forcing each end to live in a half–space. Using
this restriction, P. Collin proved:
Theorem 14.1. [30] Each end of a complete properly embedded minimal surface
with finite topology and at least two ends is asymptotic to a plane or catenoid.
In particular, outside some compact set, Σ is given by a finite collection of disjoint
graphs over a common plane (and has finite total curvature).
As mentioned above, Collin essentially proved Theorem 14.1 by showing that an
embedded annular end that lives in a half–space must have finite total curvature.
[27] used the one–sided curvature estimate to strengthen this from a half–space
to a strictly larger cone, and in the process give a very different proof of Collin’s
theorem.
Theorem 14.2. [27] There exists ǫ > 0 so that any complete properly embedded
minimal annular end contained in the cone
(14.1) {x3 ≥ −ǫ (x21 + x22 + x23 )1/2 }
is asymptotic to a plane or catenoid.
When Σ has only one end (e.g., for the helicoid), it need not have finite total
curvature so the situation is more delicate. However, the regularity results of the
previous section can be applied. For example, if Σ is a (non–planar) embedded mini-
mal disk, then we get a multi–valued graph structure away from a “one–dimensional
singular set.” Using Theorems 11.1 and 11.2, W. Meeks and H. Rosenberg proved
the uniqueness of the helicoid:
Theorem 14.3. [70] The plane and helicoid are the only complete properly embed-
ded simply–connected minimal surfaces in R3 .
This uniqueness has many applications. Recall that if we take a sequence of
rescalings of the helicoid, then the singular set S for the convergence is the vertical
axis perpendicular to the leaves of the foliation. In [64], W. Meeks used this fact
together with the uniqueness of the helicoid to prove that the singular set S in
Theorem 11.1 is always a straight line perpendicular to the foliation. Recently, W.
Meeks and M. Weber have constructed a local example (i.e., a sequence of embedded
minimal disks in a unit ball) where S is a circle.
We have not even touched on the case where Σ has infinite topology (e.g., when
Σ is one of the Riemann examples). This is an area of much current research, see
[18], the work of Meeks, J. Perez and A. Ros, [67], [68], and [69], the survey [66]
and references therein.
MINIMAL SUBMANIFOLDS 27
We close this section with a local analog of the two–ended case. Namely, in [22],
we proved that any embedded minimal annulus in a ball (with boundary in the
boundary of the ball and) with a small neck can be decomposed by a simple closed
geodesic into two graphical sub–annuli. Moreover, we gave a sharp bound for the
length of this closed geodesic in terms of the separation (or height) between the
graphical sub–annuli. This serves to illustrate our “pair of pants” decomposition
from [18] in the special case where the embedded minimal planar domain is an
annulus (we will not touch on this further here). The catenoid
(14.2) {x21 + x22 = cosh2 x3 }
is the prime example of an embedded minimal annulus.
The precise statement of this decomposition for annuli is:
Theorem 14.4. [22] There exist ǫ > 0, C1 , C2 , C3 > 1 so: If Σ ⊂ BR ⊂ R3 is an
embedded minimal annulus with ∂Σ ⊂ ∂BR and π1 (BǫR ∩ Σ) 6= 0, then there exists
a simple closed geodesic γ ⊂ Σ of length ℓ so that:
R component of BR/C1 ∩ Σ containing it into
• The curve γ splits the connected
two annuli Σ+ , Σ− each with |A|2 ≤ 5 π.
• Furthermore, Σ± \ TC2 ℓ (γ) are graphs with gradient ≤ 1.
• Finally, ℓ log(R/ℓ) ≤ C3 h where the separation h is given by
(14.3) h= min |x+ − x− | .
x± ∈∂BR/C1 ∩Σ±
As a corollary of this intrinsic one-sided curvature estimate we get that the sec-
ond, and more ambitious, of Calabi’s conjectures is also true for embedded minimal
disks.
MINIMAL SUBMANIFOLDS 29
In fact, [29] proved both of Calabi’s conjectures and properness also for embedded
surfaces with finite topology. Recall that a surface Σ is said to have finite topology
if it is homeomorphic to a closed Riemann surface with a finite set of points removed
or “punctures”. Each puncture corresponds to an end of Σ.
The following generalization of the halfspace theorem gives Calabi’s second, more
ambitious, conjecture for embedded surfaces with finite topology:
Theorem 15.4. The plane is the only complete embedded minimal surface with
finite topology in R3 in a halfspace.
Likewise, we get the properness of embedded surfaces with finite topology:
Theorem 15.5. A complete embedded minimal surface with finite topology in R3
must be proper.
Compare also W. Meeks and H. Rosenberg, [71].
There has been extensive work on both properness and the halfspace property
assuming various curvature bounds. Jorge and Xavier, [49] and [50], showed that
there cannot exist a complete immersed minimal surface with bounded curvature
in ∩i {xi > 0}; later Xavier proved that the plane is the only such surface in a
halfspace, [107]. Recently, G.P. Bessa, Jorge and G. Oliveira-Filho, [7], and H.
Rosenberg, [86], have shown that if such a surface is embedded, then it must be
proper. This properness was extended to embedded minimal surfaces with locally
bounded curvature and finite topology by Meeks and Rosenberg in [70]; finite topol-
ogy was subsequently replaced by finite genus in [67] by Meeks, J. Perez and A.
Ros.
Inspired by Nadirashvili’s examples, F. Martin and S. Morales constructed in
[61] a complete bounded minimal immersion which is proper in the (open) unit
ball. That is, the preimages of compact subsets of the (open) unit ball are compact
in the surface and the image of the surface accumulates on the boundary of the unit
ball. They extended this in [62] to show that any convex, possibly noncompact or
nonsmooth, region of R3 admits a proper complete minimal immersion of the unit
disk; cf. [78].
Note that some restriction on the boundary curve γ is certainly necessary. For
instance, if the boundary curve was knotted (e.g., the trefoil), then it could not
be spanned by any embedded disk (minimal or otherwise). Prior to the work of
Meeks and Yau, embeddedness was known for extremal boundary curves in R3 with
small total curvature by the work of R. Gulliver and J. Spruck [38]; see chapter 4
in [14] for other results and further discussion. Recently, in [33], T. Ekholm, B.
White, and D. Wienholtz proved that minimal surfaces whose boundary has total
curvature less than 4π also must be embedded.
If we instead fix a homotopy class of maps, then the two fundamental exis-
tence results are due to J. Sacks–K. Uhlenbeck and R. Schoen– S.T. Yau (with
embeddedness proven by W. Meeks–S.T. Yau and M. Freedman–J. Hass–P. Scott,
respectively):
Theorem 18.2. [87], [75] Given M 3 , there exist conformal (stable) minimal im-
mersions u1 , . . . , um : S2 → M which generate π2 (M ) as a Z[π1 (M )] module.
Furthermore,
• If u : S2 → M and [u]π2 6= 0, then Area(u) ≥ mini Area(ui ).
• Each ui is either an embedding or a 2–1 map onto an embedded 2–sided
RP 2 .
Theorem 18.3. [93], [36] If Σ2 is a closed surface with genus g > 0 and i0 : Σ →
M 3 is an embedding which induces an injective map on π1 , then there is a least
area embedding with the same action on π1 .
In [72], W. Meeks, L. Simon, and S.T. Yau find an embedded sphere minimizing
area in an isotopy class in a closed 3–manifold.
We end this section by mentioning two applications of Theorem 18.3. First, in
[26], we showed that any topological 3–manifold M had an open set of metrics so
that, for each such metric, there was a sequence of embedded minimal tori whose
area went to infinity. In [32], B. Dean showed that this was true for every genus
g ≥ 1.
do satisfy Condition C and then obtain minimal surfaces as limits of critical points
for the perturbed problems:
Theorem 19.1. [87] If πk (M ) 6= 0 for some k > 1, then there exists a branched
immersed minimal 2–sphere in M (for any metric).
The main new contribution of Smith was to control the topological type of the
resulting minimal surface while keeping it embedded; see also J. Pitts and J.H.
Rubinstein, [83], for some generalizations.
20.1. The positive mass theorem. The (Riemannian version of the) positive
mass theorem of R. Schoen and S.T. Yau states that an asymptotically flat 3-
manifold M with non-negative scalar curvature must have positive mass. The Rie-
mannian manifold M here arises as a maximal space-like slice in a 3+1-dimensional
space-time solution of Einstein’s equations.
The asymptotic flatness of M comes from that the space-time models an isolated
gravitational system and hence is a perturbation of the vacuum solution outside a
large compact set. To make this precise, suppose for simplicity that M has only
one end; M is then said to be asymptotically flat if there is a compact set Ω ⊂ M so
that M \ Ω is diffeomorphic to R3 \ BR (0) and the metric on M \ Ω can be written
as
4
M
(20.1) gij = 1 + δij + pij ,
2 |x|
where
(20.2) |x|2 |pij | + |x|3 |D pij | + |x|4 |D2 pij | ≤ C .
The constant M is the so called mass of M . Observe that the metric gij is a
perturbation of the metric on a constant-time slice in the Schwarzschild space-time
of mass M; that is to say, the Schwarzschild metric has pij ≡ 0.
A tensor h is said to be O(|x|−p ) if |x|p |h| + |x|p+1 |D h| ≤ C. For example, an
easy calculation shows that
(20.3) gij = (1 + 2 M/|x|) δij + O(|x|−2 ) ,
√
g ≡ det gij = 1 + 3 M |x|−1 + O(|x|−2 ) .
p
The positive mass theorem states that the mass M of such an M must be non-
negative:
Theorem 20.1. [92] With M as above, M ≥ 0.
There is a rigidity theorem as well which states that the mass vanishes only when
M is isometric to R3 :
Theorem 20.2. [92] If |∇3 pij | = O(|x|−5 ) and M = 0 in Theorem 20.1, then M
is isometric to R3 .
We will give a very brief overview of the proof of Theorem 20.1, showing in the
process where minimal surfaces appear.
Proof. (Sketch.) The argument will by contradiction, so suppose that the mass is
negative. Note first that a “rounding off” argument shows that the metric on M
can be perturbed to have positive scalar curvature outside of a compact set and
still have negative mass.
It is not hard to prove that the slab between two parallel planes is mean-convex.
That is, we have the following:
Lemma 20.3. If M < 0 and M is asymptotically flat, then there exist R0 , h > 0
so that for r > R0 the sets
(20.4) Cr = {|x|2 ≤ r2 , −h ≤ x3 ≤ h}
have strictly mean convex boundary.
MINIMAL SUBMANIFOLDS 35
Since the compact set Cr is mean convex, we can solve the Plateau problem (as
in Section 16) to get an area minimizing (and hence stable) surface Γr ⊂ Cr with
boundary
(20.5) ∂Γr = {|x|2 = r2 , x3 = h} .
Using the disk {|x|2 ≤ r2 , x3 = h} as a comparison surface, we get uniform local
area bounds for any such Γr . Combining these local area bounds with the a priori
curvature estimates for minimizing surfaces, we can take a sequence of r’s going to
infinity and find a subsequence of Γr ’s that converge to a complete area-minimizing
surface
(20.6) Γ ⊂ {−h ≤ x3 ≤ h} .
Since Γ is pinched between the planes {x3 = ±h}, the estimates for minimizing
surfaces implies that (outside a large compact set) Γ is a graph over the plane
{x3 = 0} and hence has quadratic area growth and finite total curvature. Moreover,
using the form of the metric gij , we see that |∇u| decays like |x|−1 and
Z
(20.7) kg = (2 π s + O(1)) (s−1 + O(s−2 )) = 2 π + O(s−1 ) ,
σs
20.2. Black holes. Another way that minimal surfaces enter into relativity is
through black holes. Suppose that we have a three-dimensional time-slice M in
a 3 + 1-dimensional space–time. For simplicity, assume that M is totally geodesic
and hence has non-negative scalar curvature. A closed surface Σ in M is said to be
trapped if its mean curvature is everywhere negative with respect to its outward
normal. Physically, this means that the surface emits an outward shell of light
whose surface area is decreasing everywhere on the surface. The existence of a
closed trapped surface implies the existence of a black hole in the space-time.
Given a trapped surface, we can look for the outermost trapped surface con-
taining it; this outermost surface is called an apparent horizon. It is not hard to
see that an apparent horizon must be a minimal surface and, moreover, a barrier
argument shows that it must be stable. Since M has non-negative scalar curvature,
36 TOBIAS H. COLDING AND WILLIAM P. MINICOZZI II
stability in turn implies that it must be diffeomorphic to a sphere. See, for instance,
[9] and [47] for some results on black holes, horizons, etc.
20.3. Constant mean curvature surfaces. At least since the time of Plateau,
minimal surfaces have been used to model soap films. This is because the mean
curvature of the surface models the surface tension and this is essentially the only
force acting on a soap film. Soap bubbles, on other hand, enclose a volume and
thus the pressure gives a second counterbalancing force. It follows easily that these
two forces are in equilibrium when the surface has constant mean curvature.
For the same reason, constant mean curvature surfaces arise in the isoperimetric
problem. Namely, a surface that minimizes surface area while enclosing a fixed
volume must have constant mean curvature (or “cmc”). It is not hard to see that
such an isoperimetric surface in Rn must be a round sphere. There are two inter-
esting partial converses to this. First, by a theorem of H. Hopf, any cmc 2-sphere in
R3 must be round. Second, using the maximum principle (“the method of moving
planes”) A.D. Alexandrov showed that any closed embedded cmc hypersurface in
Rn must be a round sphere. It turned out, however, that not every closed immersed
cmc surface is round. The first examples were immersed cmc tori constructed by
H. Wente, [105]. N. Kapouleas constructed many new examples, including closed
higher genus cmc surfaces, [52], [53]. See also [63] and [55] for other results on the
space of such cmc surfaces.
Many of the techniques developed for studying minimal surfaces generalize to
general constant mean curvature surfaces.
20.4. Finite extinction for Ricci flow. We close this survey with indicating how
minimal surfaces can be used to show that on a homotopy 3-sphere the Ricci flow
become extinct in finite time (see [25], [80] for details).
Let M 3 be a smooth closed orientable 3–manifold and let g(t) be a one–parameter
family of metrics on M evolving by the Ricci flow, so
(20.10) ∂t g = −2 RicMt .
In an earlier section, we saw that there is a natural way of constructing minimal
surfaces on many 3-manifolds and that comes from the min–max argument where
the minimal of all maximal slices of sweep–outs is a minimal surface. The idea
is then to look at how the area of this min–max surface changes under the flow.
Geometrically the area measures a kind of width of the 3–manifold and as we will
see for certain 3–manifolds (those, like the 3–sphere, whose prime decomposition
contains no aspherical factors) the area becomes zero in finite time corresponding
to that the solution becomes extinct in finite time.
Fix a continuous map β : [0, 1] → C 0 ∩ L21 (S2 , M ) where β(0) and β(1) are
constant maps so that β is in the nontrivial homotopy class [β] (such β exists when
M is a homotopy 3-sphere). We define the width W = W (g, [β]) by
(20.11) W (g) = min max Energy(γ(s)) .
γ∈[β] s∈[0,1]
The next theorem gives an upper bound for the derivative of W (g(t)) under the
Ricci flow which forces the solution g(t) to become extinct in finite.
MINIMAL SUBMANIFOLDS 37
References
[1] W. Allard, On the first variation of a varifold, Ann. of Math. (2) 95 (1972) 417–491.
[2] F. J. Almgren, Jr., Some interior regularity theorems for minimal surfaces and an extension
of Bernstein’s theorem, Ann. of Math. (2) 84 (1966) 277–292.
[3] F. J. Almgren, Jr., Almgren’s big regularity paper. Q-valued functions minimizing Dirichlet’s
integral and the regularity of area-minimizing rectifiable currents up to codimension 2, With
a preface by Jean E. Taylor and Vladimir Scheffer. World Scientific Monograph Series in
Mathematics, 1. World Scientific Publishing Co., Inc., River Edge, NJ, 2000.
[4] F. J. Almgren, Jr., Minimal surface forms, Math. Intell. 4 (1982) 164–172.
38 TOBIAS H. COLDING AND WILLIAM P. MINICOZZI II
[5] S. Bernstein, Über ein geometrisches Theorem und seine Anwendung auf die partiellen Dif-
ferentialgleichungen vom ellipschen Typos. Math. Zeit. 26 (1927) 551–558 (translation of the
original version in Comm. Soc. Math. Kharkov 2-ème sér. 15 (1915–17) 38–45).
[6] L. Bers, Isolated singularities of minimal surfaces, Ann. of Math. (2) 53 (1951) 364–386.
[7] G.P. Bessa, L. Jorge and G. Oliveira-Filho, Half-space theorems for minimal surfaces with
bounded curvature, J. Diff. Geom. 57 (2001) 493–508.
[8] E. Bombieri, E. De Giorgi, and E. Giusti, Minimal cones and the Bernstein problem, Invent.
Math. 7 (1969) 243–268.
[9] H. Bray, Proof of the Riemannian Penrose inequality using the positive mass theorem, J.
Differential Geom. 59 (2001), no. 2, 177–267.
[10] E. Calabi, Problems in differential geometry, Ed. S. Kobayashi and J. Eells, Jr., Proceedings
of the United States-Japan Seminar in Differential Geometry, Kyoto, Japan, 1965. Nippon
Hyoronsha Co., Ltd., Tokyo (1966) 170.
[11] S.S. Chern, The geometry of G-structures, Bull. Amer. Math. Soc. 72 (1966) 167–219.
[12] H.I. Choi and R. Schoen, The space of minimal embeddings of a surface into a three-
dimensional manifold of positive Ricci curvature, Invent. Math. 81 (1985) 387–394.
[13] T.H. Colding and C. De Lellis, The min–max construction of minimal surfaces, Surveys in
differential geometry, Vol. 8, Lectures on Geometry and Topology held in honor of Calabi,
Lawson, Siu, and Uhlenbeck at Harvard University, May 3–5, 2002, Sponsored by the Journal
of Differential Geometry, (2003) 75–107, math.AP/0303305.
[14] T.H. Colding and W.P. Minicozzi II, Minimal surfaces, Courant Lecture Notes in Math., v.
4, 1999.
[15] T.H. Colding and W.P. Minicozzi II, Estimates for parametric elliptic integrands, Interna-
tional Mathematics Research Notices, no. 6 (2002) 291-297.
[16] T.H. Colding and W.P. Minicozzi II, The space of embedded minimal surfaces of fixed genus
in a 3-manifold I; Estimates off the axis for disks, Annals of Math., 160 (2004) 27–68,
math.AP/0210106.
[17] T.H. Colding and W.P. Minicozzi II, The space of embedded minimal surfaces of fixed
genus in a 3-manifold II; Multi-valued graphs in disks, Annals of Math., 160 (2004) 69–92,
math.AP/0210086.
[18] T.H. Colding and W.P. Minicozzi II, The space of embedded minimal surfaces of fixed genus in
a 3-manifold III; Planar domains, Annals of Math., 160 (2004) 523–572, math.AP/0210141.
[19] T.H. Colding and W.P. Minicozzi II, The space of embedded minimal surfaces of fixed
genus in a 3-manifold IV; Locally simply connected, Annals of Math., 160 (2004) 573–615,
math.AP/0210119.
[20] T.H. Colding and W.P. Minicozzi II, The space of embedded minimal surfaces of fixed genus
in a 3-manifold V; Fixed genus, math.DG/0509647.
[21] T.H. Colding and W.P. Minicozzi II, Multi-valued minimal graphs and properness of disks,
International Mathematics Research Notices, no. 21 (2002) 1111-1127.
[22] T.H. Colding and W.P. Minicozzi II, On the structure of embedded minimal annuli, Interna-
tional Mathematics Research Notices, no. 29 (2002) 1539–1552.
[23] T.H. Colding and W.P. Minicozzi II, Disks that are double spiral staircases, Notices of the
AMS, Vol. 50, no. 3, March (2003) 327–339.
[24] T.H. Colding and W.P. Minicozzi II, Embedded minimal disks: Proper versus nonproper -
global versus local, Transactions of the AMS, 356 (2004) 283-289, math.DG/0210328.
[25] T.H. Colding and W.P. Minicozzi II, Estimates for the extinction time for the Ricci flow
on certain 3–manifolds and a question of Perelman, JAMS, 18 (2005), no. 3, 561–569,
math.AP/0308090.
[26] T.H. Colding and W.P. Minicozzi II, Examples of embedded minimal tori without area
bounds, International Mathematics Research Notices, 99 no. 20 (1999) 1097–1100.
[27] T.H. Colding and W.P. Minicozzi II, Complete properly embedded minimal surfaces in R3 ,
Duke Math. J. 107 (2001) 421–426.
[28] T.H. Colding and W.P. Minicozzi II, Embedded minimal disks, Global theory of min-
imal surfaces, 405–438, Clay Math. Proc., 2, Amer. Math. Soc., Providence, RI, 2005,
math.DG/0206146.
[29] T.H. Colding and W.P. Minicozzi II, The Calabi–Yau conjectures for embedded surfaces,
preprint, math.DG/0404197.
MINIMAL SUBMANIFOLDS 39
[30] P. Collin, Topologie et courbure des surfaces minimales proprement plonges de R3 , Ann. of
Math. (2) 145 (1997) 1–31.
[31] E. De Giorgi, Frontiere orientate di misura minima, Sem. Mat. Scuola Norm. Sup. Pisa
(1961) 1–56.
[32] B. Dean, Compact Embedded Minimal Surfaces of Positive Genus Without Area Bounds,
Geom. Ded. 102 (2003), 45–52.
[33] T. Ekholm, B. White, and D. Wienholtz, Embeddedness of minimal surfaces with total bound-
ary curvature at most 4π, Ann. of Math. (2) 155 (2002), no. 1, 209–234.
[34] H. Federer, Geometric measure theory, Springer-Verlag, Berlin–Heidelberg–New York, 1969.
[35] R. Finn, Capillary surface interfaces, Notices Amer. Math. Soc. 46 (1999), no. 7, 770–781.
[36] M.H. Freedman, J. Hass, and P. Scott, Least area incompressible surfaces in 3-manifolds,
Invent. Math. 71 (1983), no. 3, 609–642.
[37] A. Fraser, On the free boundary variational problem for minimal disks, Comm. Pure Appl.
Math. 53 (2000) 931–971.
[38] R. Gulliver and J. Spruck, On embedded minimal surfaces, Ann. of Math. (2) 103 (1976)
331–347; Ann. of Math. (2) 109 (1979) 407–412.
[39] R. Hardt, D. Kinderlehrer, and F. H. Lin, The variety of configurations of static liquid
crystals, Variational methods (Paris, 1988), 115–131, Progr. Nonlinear Differential Equations
Appl., 4, Birkhuser Boston, Boston, MA, 1990.
[40] R. Hardt and L. Simon, Boundary regularity and embedded solutions for the oriented Plateau
problem, Ann. of Math. (2) 110 (1979), no. 3, 439–486.
[41] E. Heinz, Über die Lösungen der Minimalflächengleichung, Nachr. Akad. Wiss. Göttingen
Math.–Phys. Kl, II (1952) 51–56.
[42] D. Hoffman, Mixing materials and mathematics, Nature 384 (1996) 28 - 29.
[43] D. Hoffman and H. Karcher, Complete embedded minimal surfaces with finite total curvature,
Geometry V (R. Osserman, ed.) Encyclopaedia Math. Sci. 90, Springer-Verlag, New York
(1997) 5–93.
[44] D. Hoffman and W. Meeks III, A complete embedded minimal surface in R3 with genus one
and three ends, J. Diff. Geom. 21 (1985) 109–127.
[45] D. Hoffman, M. Weber, and M. Wolf, An embedded genus-one helicoid, math.DG/0401080.
[46] D. Hoffman, M. Weber, and M. Wolf, An embedded genus-one helicoid, PNAS, November 15,
2005, vol. 102, no. 46.
[47] G. Huisken and T. Ilmanen, The inverse mean curvature flow and the Riemannian Penrose
inequality, J. Differential Geom. 59 (2001), no. 3, 353–437.
[48] T. Ilmanen, A strong maximum principle for singular minimal hypersurfaces, Calc. Var.
Partial Differential Equations 4 (1996), no. 5, 443–467.
[49] L. Jorge and F. Xavier, On the existence of complete bounded minimal surfaces in Rn , Bol.
Soc. Brasil. Mat. 10 (1979), no. 2, 171–173.
[50] L. Jorge and F. Xavier, A complete minimal surface in R3 between two parallel planes,
Annals of Math. (2) 112 (1980) 203–206.
[51] L. Jorge and F. Xavier, An inequality between the exterior diameter and the mean curvature
of bounded immersions, Math. Zeit. 178 (1981), no. 1, 77–82.
[52] N. Kapouleas, Compact constant mean curvature surfaces in Euclidean three-space, J. Dif-
ferential Geom. 33 (1991), no. 3, 683–715.
[53] N. Kapouleas, Complete constant mean curvature surfaces in Euclidean three-space, Ann. of
Math. (2) 131 (1990), no. 2, 239–330.
[54] N. Kapouleas, Complete embedded minimal surfaces of finite total curvature, J. Diff. Geom.,
47 (1997) 95–169.
[55] N. Korevaar and R. Kusner, The global structure of constant mean curvature surfaces, Invent.
Math. 114 (1993), no. 2, 311–332.
[56] H. B. Lawson, Lectures on minimal submanifolds, vol. I, Publish or Perish, Inc, Berkeley,
1980.
[57] F. Lopez, F. Martin, and S. Morales, Adding handles to Nadirashvili’s surfaces, J. Diff. Geom.
60 (2002), no. 1, 155–175.
[58] F. Lopez, F. Martin, and S. Morales, Complete nonorientable minimal surfaces in a ball of
R3 , preprint.
[59] F. Lopez and A. Ros, On embedded complete minimal surfaces of genus zero, J. Differential
Geom. 33 (1991), no. 1, 293–300.
40 TOBIAS H. COLDING AND WILLIAM P. MINICOZZI II
[60] F. Martin and S. Morales, A complete bounded minimal cylinder in R3 , Michigan Math. J.
47 (2000), no. 3, 499–514.
[61] F. Martin and S. Morales, On the asymptotic behavior of a complete bounded minimal surface
in R3 , Trans. Amer. Math. Soc. 356 (2004), 3985–3994.
[62] F. Martin and S. Morales, Complete proper minimal surfaces in convex bodies of R3 , Duke
Math. J. 128 (2005), no. 3, 559–593.
[63] R. Mazzeo, F. Pacard, and D. Pollack,Connected sums of constant mean curvature surfaces
in Euclidean 3 space, J. Reine Angew. Math. 536 (2001), 115–165.
[64] W. Meeks III, The regularity of the singular set in the Colding and Minicozzi lamination
theorem, Duke Math. Jour., 123 (2004), no. 2, 329–334.
[65] W. Meeks III, The lamination metric for the Colding-Minicozzi minimal lamination, Illinois
J. Math. 49 (2005), no. 2, 645–658.
[66] W. Meeks III and J. Perez, Conformal properties in classical minimal surface theory, Sur-
veys in Diff. Geom. IX: Eigenvalues of Laplacians and other geometric operators, Ed. by A.
Grigor’yan and S.T. Yau, International Press (2004), 275–335.
[67] W. Meeks III, J. Perez, and A. Ros, The geometry of minimal surfaces of finite genus I;
Curvature estimates and quasiperiodicity, J. Differential Geom. 66 (2004), 1–45.
[68] W. Meeks III, J. Perez, and A. Ros, The geometry of minimal surfaces of finite genus II;
Nonexistence of one limit end examples, Invent. Math. 158 (2004), no. 2, 323–341.
[69] W. Meeks III, J. Perez, and A. Ros, The geometry of minimal surfaces of finite genus III;
bounds on the topology and index of classical minimal surfaces, preprint.
[70] W. Meeks III and H. Rosenberg, The uniqueness of the helicoid and the asymptotic geometry
of properly embedded minimal surfaces with finite topology, Ann. of Math., 161 (2005), no. 2,
727–758.
[71] W. Meeks III and H. Rosenberg, The minimal lamination closure theorem, preprint.
[72] W. Meeks III, L. Simon, and S.T. Yau, Embedded minimal surfaces, exotic spheres and
manifolds with positive Ricci curvature, Ann. of Math. (2) 116 (1982) 621–659.
[73] W. Meeks III and M. Weber, Existence of bent helicoids and the regularity of the singular
set in the Colding-Minicozzi lamination theorem, in preparation.
[74] W. Meeks III and S.T. Yau, The classical Plateau problem and the topology of three dimen-
sional manifolds, Topology 21 (1982) 409–442.
[75] W.H. Meeks and S.T. Yau, Topology of three–dimensional manifolds and the embedding
problems in minimal surface theory, Ann. of Math. (2) 112 (1980), no. 3, 441–484.
[76] M.J. Micallef and J.D. Moore, Minimal two–spheres and the topology of manifolds with
positive curvature on totally isotropic two–planes, Ann. of Math. (2) 127 (1988) no. 1 199–
227.
[77] N. Nadirashvili, Hadamard’s and Calabi-Yau’s conjectures on negatively curved and minimal
surfaces, Invent. Math. 126 (1996) 457–465.
[78] N. Nadirashvili, An application of potential analysis to minimal surfaces, Mosc. Math. J. 1
(2001), no. 4, 601–604, 645.
[79] R. Osserman, A survey of minimal surfaces, Dover, 2nd. edition (1986).
[80] G. Perelman, Finite extinction time for the solutions to the Ricci flow on certain three–
manifolds, math.DG/0307245.
[81] J. Perez, Limits by rescalings of minimal surfaces: Minimal laminations, curvature decay
and local pictures, notes for the workshop ”Moduli Spaces of Properly Embedded Minimal
Surfaces”, American Institute of Mathematics, Palo Alto, California (2005).
[82] J.T. Pitts, Existence and regularity of minimal surfaces on Riemannian manifolds, Princeton
University Press, Princeton, N.J.; University of Tokyo Press, Tokyo (1981).
[83] J.T. Pitts and J.H. Rubinstein, Applications of minmax to minimal surfaces and the topology
of 3–manifolds, Miniconference on geometry and partial differential equations, Proceedings of
the CMA, Australia National University (1986).
[84] H. Rosenberg, Some recent developments in the theory of properly embedded minimal surfaces
in R3 , Seminare Bourbaki 1991/92, Asterisque No. 206 (1992) 463–535.
[85] H. Rosenberg, Some recent developments in the theory of minimal surfaces in 3-manifolds,
IMPA Mathematical Publications. 24th Brazilian Mathematics Colloquium, Instituto de
Matemtica Pura e Aplicada (IMPA), Rio de Janeiro, 2003.
[86] H. Rosenberg, A complete embedded minimal surface in R3 of bounded curvature is proper,
preprint.
MINIMAL SUBMANIFOLDS 41
[87] J. Sacks and K. Uhlenbeck, The existence of minimal immersions of 2–spheres, Ann. of Math.
(2) 113 (1981) no. 1, 1–24.
[88] R. Schoen, Estimates for stable minimal surfaces in three–dimensional manifolds, In Semi-
nar on Minimal Submanifolds, Ann. of Math. Studies, vol. 103, Princeton University Press,
Princeton, N.J., (1983) 111–126.
[89] R. Schoen, Uniqueness, symmetry, and embeddedness of minimal surfaces, J. Differential
Geom. 18 (1983), no. 4, 791–809 (1984).
[90] R. Schoen and L. Simon, Regularity of simply connected surfaces with quasi-conformal Gauss
map, In Seminar on Minimal Submanifolds, Annals of Math. Studies, vol. 103, Princeton
University Press, Princeton, N.J., (1983) 127–145.
[91] R. Schoen, L. Simon, and S.T. Yau, Curvature estimates for minimal hypersurfaces, Acta
Math. 134 (1975) 275–288.
[92] Schoen, R. and Yau, S. T., On the proof of the positive mass conjecture in general relativity,
Comm. Math. Phys. 65 (1979), no. 1, 45–76.
[93] R. Schoen and S.T. Yau, Existence of incompressible minimal surfaces and the topology of
three dimensional manifolds with nonnegative scalar curvature, Ann. of Math. (2) 110 (1979)
127–142.
[94] L. Simon, Asymptotic behaviour of minimal graphs over exterior domains, Ann. Inst. H.
Poincar Anal. Non Linaire 4 (1987) 231–242.
[95] L. Simon, Remarks on curvature estimates for minimal hypersurfaces, Duke Math. J. 43
(1976) 545–553.
[96] L. Simon, Singularities of Geometric Variational Problems, In Nonlinear Partial Differential
Equations in Differential Geometry (R. Hardt and M. Wolf, Ed.), American Mathematical
Society, Providence (1996) 185–223.
[97] L. Simon, A strict maximum principle for area minimizing hypersurfaces, J. Differential
Geom. 26 (1987), no. 2, 327–335.
[98] J. Simons, Minimal varieties in Riemannian manifolds, Ann. of Math. (2) 88 (1968) 62–105.
[99] F. Smith, On the existence of embedded minimal 2–spheres in the 3–sphere, endowed with
an arbitrary Riemannian metric, supervisor L. Simon, University of Melbourne (1982).
[100] B. Solomon and B. White, A strong maximum principle for varifolds that are stationary
with respect to even parametric elliptic functionals, Indiana Univ. Math. J. 38 (1989), no. 3,
683–691.
[101] J. Taylor, Some mathematical challenges in materials science, Mathematical challenges of
the 21st century (Los Angeles, CA, 2000). Bull. Amer. Math. Soc. (N.S.) 40 (2003), no. 1,
69–87.
[102] D.W. Thompson, On growth and form, New ed., Cambridge Univ. Press, Cambridge (1942).
[103] M. Traizet, Adding handles to Riemann’s minimal surfaces, J. Inst. Math. Jussieu 1 (2002)
145–174.
[104] M. Weber and M. Wolf, Teichmüller theory and handle addition for minimal surfaces, Ann.
of Math. (2), 156 (2002) 713–795.
[105] H. Wente, Counterexample to a conjecture of H. Hopf, Pacific J. Math. 121 (1986), no. 1,
193–243.
[106] B. White, Evolution of curves and surfaces by mean curvature, Proceedings of the ICM
(2002).
[107] F. Xavier, Convex hulls of complete minimal surfaces, Math. Ann. 269 (1984) 179–182.
[108] S.D. Yang, A connected sum construction for complete minimal surfaces of finite total
curvature, Comm. Anal. Geom. 9 (2001), no. 1, 115–167.
[109] S.T. Yau, Nonlinear analysis in geometry, L’Eseignement Mathematique (2) 33 (1987) 109–
158.