DG
DG
Lecture Notes
Gilbert Weinstein
Contents
Preface 5
Chapter 1. Curves 7
1. Preliminaries 7
2. Local Theory for Curves in R3 8
3. Plane Curves 10
4. Fenchel’s Theorem 14
Exercises 16
Chapter 2. Local Surface Theory 19
1. Surfaces 19
2. The First Fundamental Form 21
3. The Second Fundamental Form 23
4. Examples 25
5. Lines of Curvature 28
6. More Examples 30
7. Surface Area 33
8. Bernstein’s Theorem 37
9. Theorema Egregium 39
Exercises 41
Chapter 3. Local Intrinsic Geometry of Surfaces 45
1. Riemannian Surfaces 45
2. Lie Derivative 47
3. Covariant Differentiation 48
4. Geodesics 50
5. The Riemann Curvature Tensor 53
6. The Second Variation of Arclength 56
Exercises 59
Index 61
3
Preface
These notes are for a beginning graduate level course in differential geometry.
It is assumed that this is the students’ first course in the subject. Thus the choice
of subjects and presentation has been made to facilitate as much as possible a
concrete picture. For those interested in a deeper study, a second course would take
a more abstract point of view, and in particular, could go further into Riemannian
geometry.
Much of the material is borrowed from the following sources, but has been
adapted according to my own taste:
[1] M. P. Do Carmo, Differential geometry of curves and surfaces, Prentice-Hall.
[2] L. P. Eisenhart An introduction to differential geometry with use of the ten-
sor calculus, Princeton University Press.
[3] W. Klingenberg, A course in differential geometry, Springer-Verlag.
[4] B. O’Neill Elementary differential geometry, Academic Press.
[5] M. Spivak, A comprehensive introduction to Differential Geometry, Publish
or Perish.
[6] J. J. Stoker, Differential Geometry, Wiley & Sons.
The prerequisites for this course are: linear algebra, preferably with some ex-
posure to multilinear algebra; calculus up to and including the inverse and implicit
function theorem; the fundamental theorem of ordinary differential equations con-
cerning existence of solutions, uniqueness, and continuous dependence on parame-
ters, and some knowledge of linear systems of ordinary differential equations; linear
first order partial differential equations; complex analysis including Liouville’s the-
orem; and some elementary topology.
It is highly recommended for the students to complete all the exercises included
in these notes.
Gilbert Weinstein
Birmingham, Alabama
April 2000
5
CHAPTER 1
Curves
1. Preliminaries
Definition 1.1. A parametrized curve is a smooth (C ∞ ) function γ : I → Rn .
A curve is regular if γ 0 6= 0.
When the interval I is closed, we say that γ is C ∞ on I if there is an interval
J and a C ∞ function β on J which agrees with γ on I.
Definition 1.2. Let γ : I → Rn be a parametrized curve, and let β : J → Rn
be another parametrized curve. We say that β is a reparametrization (orientation-
preserving reparametrization) of γ if there is a smooth map τ : J → I with τ 0 > 0
such that β = γ ◦ τ .
Note that the relation β is a reparametrization of γ is an equivalence relation.
A curve is an equivalence class of parametrized curves. Furthermore, if γ is regular
then every reparametrization of γ is also regular, so we may speak of regular curves.
Definition 1.3. Let γ : I → Rn be a regular curve. For any compact interval
[a, b] ⊂ I, the arclength of γ over [a, b] is given by:
Z b
Lγ ([a, b]) = |γ 0 | dt.
a
By the uniqueness of solutions of the initial value problem, it follows that QX = X̃.
In particular, (R ◦ γ)0 = γ̃ 0 , and since R ◦ γ(0) = γ̃(0) we conclude R ◦ γ ≡ γ̃.
3. Plane Curves
3.1. Local Theory. Let γ : [a, b] → R2 be a regular plane curve parametrized
by arclength, and let κ be its curvature. Note that κ is signed, and in fact changes
sign (but not magnitude) when the orientation of γ is reversed. The Frenet frame
equations are:
e01 = κ e2 , e02 = −κ e1
Proposition 1.6. Let γ : [a, b] → R2 be a regular curve with |γ 0 | = 1. Then
there exists a differentiable function θ : [a, b] → R such that
(1.3) e1 = (cos θ, sin θ).
Moreover, θ is unique up to a constant integer multiple of 2π, and in particular
θ(b) − θ(a) is independent of the choice of θ. The derivative of θ is the curvature:
θ0 = κ.
Proof. Let a = t0 < t1 < · · · < tn = b be a partition of [a, b] so that the
diameter of e1 ([ti−1 , ti ]) is less than 2, i.e., e1 restricted to each subinterval maps
into a semi-circle. Such a partition exists since e1 is uniformly continuous on [a, b].
Choose θ(a) so that (1.3) holds at a, and proceed by induction on i: if θ is defined
at ti then there is a unique continuous extension so that (1.3) holds. If ψ is any
other continuous function satisfying (1.3), then k = (1/2π)(θ − ψ) is a continuous
integer-valued function, hence is constant. Finally, e2 = (− sin θ, cos θ) hence
e01 = κ e2 = θ0 (− sin θ, cos θ),
and we obtain θ0 = κ.
3.2. Global Theory.
Definition 1.7. A curve γ : [a, b] → Rn is closed if γ (k) (a) = γ (k) (b). A closed
curve γ : [a, b] → Rn is simple if γ|(a,b) is one-to-one. The rotation number of a
smooth closed curve is:
1
(1.4) nγ = θ(a) − θ(b) ,
2π
where θ is the function defined in Proposition 1.6.
We note that the rotation number is always an integer. For reference, we also
note that the rotation number of a curve is the winding number of the map e1 .
Finally, in view of the last statement in Proposition 1.6, we have:
Z
1
nγ = κ ds.
2π [0,L]
Theorem 1.7 (Rotation Theorem). Let γ : [0, L] → R2 be a smooth, regular,
simple, closed curve. Then nγ = ±1. In particular
Z
1
κ ds = ±1.
2π [0,L]
3. PLANE CURVES 11
For the proof we will need the following technical lemma. We say that a set
∆ ⊂ Rn is star-shaped with respect to x0 ∈ ∆ if for every y ∈ ∆ the line segment
x0 y lies in ∆.
Lemma 1.8. Let ∆ ⊂ Rn be star-shaped with respect to x0 ∈ ∆, and let e : ∆ →
1
S be a continuous function. Then there exists a continuous function θ : ∆ → R
such that:
(1.5) e = (cos θ, sin θ).
Moreover, if ψ is another continuous function satisfying (1.5), then θ − ψ = 2πk
where k is a constant integer.
In fact, it is sufficient to assume that ∆ is simply connected, but we will not
prove this more general result here.
Proof. Define θ(x0 ) so that (1.5) holds at x0 . For each x ∈ ∆ define θ
continuously along the line segment x0 x as in the proof of Proposition 1.6. Since
∆ is star-shaped with respect to x0 , this defines θ everywhere in ∆. It remains
to show that θ is continuous. Let y0 ∈ ∆. Since x0 y0 is compact, it is possible
to choose δ small enough that the following holds: y 0 ∈ x0 y0 and |y − y 0 | < δ
implies |e(y) − e(y 0 )| < 2 or equivalently e(y) and e(y 0 ) are not antipodal. Let
0 < < π. Then there exists a neighborhood U ⊂ Bδ (y0 ) of y0 such that y ∈ U
implies θ(y) − θ(y0 ) = 2πk(y) + 0 (y) where |0 (y)| < and k(y) is integer-valued.
It remain to prove that k ≡ 0. Let y ∈ U and consider the continuous function:
φ(s) = θ x0 + s(y − x0 ) − θ x0 + s(y0 − x0 ) , 0 6 s 6 1.
Since
x0 + s(y − x0 ) − x0 + s(y0 − x0 ) = |s(y − y0 )| < δ,
it follows from our choice of δ that e x0 + s(y − x0 ) and e x0 + s(y0 − x0 ) are not
antipodal. Thus, φ(s) 6= π for all 0 6 s 6 1, and since φ(0) = 0 we conclude that
|φ| < π. In particular
|2πk(y) + 0 (y)| = |θ(y) − θ(y0 )| = |φ(1)| < π,
and it follows that
|2πk(y)| 6 |2πk(y) + 0 (y)| + |0 (y)| < 2π.
Since k(y) is integer-valued this implies k(y) = 0.
Proof of the Rotation Theorem. Pick a line which intersects the curve
γ and pick a last point p on this line, i.e., a point with the property that one ray
of the line from p has no other intersection points with γ. Let h be the unit vector
pointing in the direction of that ray. We assume without loss of generality that γ
is parametrized by arclength, γ(0) = γ(L) = 0. Now, let ∆ = {(t1 , t2 ) ∈ R2 : 0 6
t1 6 t2 6 L}, and note that ∆ is star-shaped. Define the S1 -valued function:
0
γ (t1 ) if t1 = t2 ;
0
e(t1 , t2 ) = −γ (0) if (t1 , t2 ) = (0, L);
γ(t 2 ) − γ(t 1 )
otherwise.
|γ(t2 ) − γ(t1 )|
It is straightforward to check that e is continuous on ∆. By the Lemma, there
is a continuous function θ : ∆ → R such that e = (cos θ, sin θ). We claim that
12 1. CURVES
θ(L, L) − θ(0, 0) = ±2π which proves the theorem, since θ(t, t) is a continuous
function satisfying (1.3) in Proposition 1.6, and thus can be used on the right-hand
side of (1.4) to compute the rotation number.
To prove this claim, note that, for any 0 < t < L, the unit vector
γ(t) − γ(0)
e(0, t) =
|γ(t) − γ(0)|
is never equal to h. Hence, there is some value α such that θ(0, t)−θ(0, 0) 6= α+2πk
for any integer k. Thus, |θ(0, t) − θ(0, 0)| < 2π, and since e(0, L) = −e(0, 0) it
follows that θ(0, L) − θ(0, 0) = ±π.
Since the curves e(0, t) and e(t, L) are related via a rigid motion, i.e., e(t, L)
=
Re(0, t) where R is rotation by π, it follows that ψ(t) = θ(t, L) − θ(0, L) −
θ(0, t) − θ(0, 0) is a constant. Since clearly ψ(0) = 0, we get θ(0, L) − θ(0, 0) =
θ(L, L) − θ(0, L), and we conclude:
θ(L, L) − θ(0, 0) = θ(t, L) − θ(0, L) + θ(0, t) − θ(0, 0) = ±2π.
Definition 1.8. A piecewise smooth curve is a continuous function γ : [a, b] →
Rn such that there is a partition of [a, b]:
a = a0 < a1 < · · · < bn = b
such that for each 1 6 j 6 n the curve segment γj = γ|[aj−1 ,aj ] is smooth.
The points γ(aj ) are called the corners of γ. The directed angle −π < ψj 6 π
from γ 0 (aj −) to γ 0 (aj +) is called the exterior angle at the j-th corner. Define
θj : [aj−1 , aj ] → R as in Proposition 1.6, i.e., so that γj0 = (cos θj , sin θj ). The
rotation number of γ is given by:
n n
1 X 1 X
nγ = θj (aj ) − θj (aj−1 ) + ψj .
2π j=1 2π j=1
Note that the three tangents at t− , t+ and t0 are parallel but distinct. Since
φ0 (t− ) = φ0 (t+ ) = 0, we have that e1 (t− ) and e1 (t+ ) are both equal to ±e1 (t0 ).
Thus, at least two of these vectors are equal. We may assume, after reparametriza-
tion, that there exists 0 < s < L such that e1 (0) = e1 (s). This implies that
θ(s) − θ(0) = 2πk, θ(L) − θ(s) = 2πk 0
with k, k 0 ∈ Z. By the Rotation Theorem, nγ = k + k 0 = ±1. Since γ(0) and γ(s)
do not lie on a line parallel to e1 (t0 ), it follows that θ is not constant on either
[0, s] or [0, L]. If k = 0 then θ0 changes sign on [0, s], and similarly if k 0 = 0 then θ0
changes sign on [s, L]. If kk 0 6= 0, then since k + k 0 = ±1, it follows that kk 0 < 0
and θ0 changes sign on [0, L].
Definition 1.10. Let γ : [0, L] → R2 be a regular plane curve. A vertex of γ
is a critical point of the curvature κ.
Theorem 1.11 (The Four Vertex Theorem). A regular simple convex closed
curve has at least four vertices.
14 1. CURVES
Proof. Clearly, κ has a maximum and minimum on [0, L], hence γ has at least
two vertices. We will assume, without loss of generality, that γ is parametrized by
arclength, has its minimum at t = 0, its maximum at t = t0 where 0 < t0 < L,
that γ(0) and γ(t0 ) lie on the x-axis, and that γ enters the upper-half plane in
the interval [0, t0 ]. All these properties can be achieved by reparametrizing and
rotating γ.
We now claim that p = γ(0) and q = γ(t0 ) are the only points of γ on the
x-axis. Indeed, suppose that there is another point r = γ(t1 ) on the x-axis, then
one of these points lies between the other two, and the tangent at that point must,
by convexity, contain the other two. Thus, by the argument used in the proof of
Theorem 1.10 the segment between the outer two is contained in γ, and in particular
pq is contained in γ. If follows that κ = 0 at p and q where κ has its minimum
and maximum, hence κ ≡ 0, a contradiction since then γ is a line and cannot be
closed. We conclude that γ remains in the upper half-plane in the interval [0, t0 ]
and remains in the lower half-plane in the interval [t0 , L].
Suppose now by contradiction that γ(0) and γ(t0 ) are the only vertices of γ.
Then it follows that:
κ0 > 0 on [0, t0 ], κ0 6 0 on [t0 , L].
Thus, if we write γ = (x, y), then we have κ0 y > 0 on [0, L], and x00 = −κy 0 , hence:
Z L Z L Z L
00 0
0= x ds = − −κy ds = κ0 y ds.
0 0 0
Since the integrand in the last integral is non-negative, we conclude that κ0 y ≡ 0,
hence y ≡ 0, again a contradiction.
It follows that κ has another point where κ0 changes sign, i.e., an extremum.
Since extrema come in pairs, κ has at least four extrema.
4. Fenchel’s Theorem
We will use without proof the fact that the shortest path between two points
on a sphere is always an arc of a great circle. We also use the notation γ1 + γ2 to
denote the curve γ1 followed by the curve γ2 .
Definition 1.11. Let γ : [0, L] → Rn be a regular curve parametrized by ar-
clength. The spherical image of γ is the curve γ 0 : [0, L] → Sn−1 . The total curvature
of γ : [0, L] → Rn is: Z
Kγ = |γ 00 | ds.
I
We note that the total curvature is simply the length of the spherical image.
Theorem 1.12. Let γ be a regular simple closed curve in Rn parametrized by
arclength. Then the total curvature of γ is at least 2π:
Kγ > 2π,
with equality if and only if γ is planar and convex.
The proof will follow from two lemmata which are interesting in their own right.
Lemma 1.13. Let γ : [0, L] → Rn be a regular closed curve parametrized by
arclength. Then the spherical image of γ cannot map into an open hemisphere. If
γ 0 maps into a closed hemisphere, then γ maps into an equator.
4. FENCHEL’S THEOREM 15
Lemma 1.14. Let n > 3, and let γ : [0, L] → Sn−1 be a regular closed curve on
the unit sphere parametrized by arclength.
(1) If the arclength of γ is less than 2π then γ is contained in an open hemi-
sphere.
(2) If the arclength of γ is equal to 2π then γ is contained in a closed hemi-
sphere.
Proof. (1) First observe that no piecewise smooth curve of arclength less
than 2π contains two antipodal points. Otherwise the two segments of of the curve
between p and q would each have length at least π, and hence the length of the
curve would have to be at least 2π. Now pick a point p on γ and let q on γ be
chosen so that the two segments γ1 and γ2 from p to q along γ have equal length.
Note that p and q cannot be antipodal. Let v be the midpoint along the shorter of
the two segments of the great circle between p and q. Suppose that γ1 intersects
the equator, the great circle v · x = 0. Let γ̃1 be the reflection of γ with respect
to v, then the length of γ1 + γ̃1 is the same as the length of γ hence is less than
2π. But γ1 + γ̃1 contains two antipodal points, a contradiction. Thus, γ1 cannot
intersect the equator. Similarly, γ2 cannot intersect the equator, and we conclude
γ stays in the open hemisphere v · x > 0.
(2) If the arclength of γ is 2π, we refine the above argument. If p and q are
antipodal, then both γ1 and γ2 are great semi-circle, thus, γ stays in a closed
hemisphere.1 So we can assume that p and q are not antipodal and proceed as
before, defining v to be the midpoint on the shorter arc of the great circle between
p and q. Now, if γ1 crosses the equator, then γ1 + γ̃1 contains two antipodal points
on the equator, and the two segments joining these points enter both hemispheres.
Thus, these segments are not semi-circle, and consequently both have arclength
strictly greater than π. Thus the arclength of γ1 + γ̃1 is strictly larger than 2π
a contradiction. Similarly, γ2 does not cross the equator, and we conclude that γ
stays in the closed hemisphere v · x > 0.
Proof of Fenchel’s Theorem. Note that the total curvature is simply the
arclength of the spherical image of γ. By Lemma 1.13 γ 0 is not contained in an
open hemisphere, so by Lemma 1.14
Z
Kγ = |γ 00 | ds > 2π.
I
sphere, and by Lemma 1.13, γ maps into an equator. If n > 3, we may proceed
by induction until we obtain that γ is planar. Once we have that γ is planar, the
1In fact, since γ is smooth, γ and γ are contained in the same great circle, and hence γ is
1 1
itself a great circle.
16 1. CURVES
Rotation Theorem gives nγ = ±1. Without loss of generality,2 we may assume that
nγ = 1. Hence Z
06 |κ| − κ ds = Kγ − 2π = 0,
I
and it follows that κ = |κ| > 0, which by Theorem 1.10 implies that γ is convex.
Exercises
Exercise 1.1. A regular space curve γ : [a, b] → R3 is a helix if there is a fixed
unit vector u ∈ R3 such that e1 · u is constant. Let κ and τ be the curvature and
torsion of a regular space curve γ, and suppose that κ 6= 0. Prove that γ is a helix
if and only if τ = cκ for some constant c.
Exercise 1.3. Prove the Fundamental Theorem for curves in R4 : Given func-
tions κ1 , κ2 , κ3 on I with κ1 , κ2 > 0, there is a smooth curve γ parameterized by
arclength on I such that κ1 , κ2 , κ3 are the curvatures of γ. Furthermore, γ is unique
up to a rigid motion of R4 .
Exercise 1.4. Let γ : [a, b] → R2 be a regular plane curve with non-zero cur-
vature κ 6= 0, and let β = γ + κ−1 N be the locus of the centers of curvature of
γ.
(1) Prove that β is regular provided that κ0 6= 0.
(2) Prove that each tangent ` of β intersects γ at a right angle.
A curve satisfying (1) and (2) is called an evolute of γ.
(3) Prove that each regular plane curve γ : [a, b] → R2 has at most one evolute.
Exercise 1.6. Let γ : [0, L] → R2 be a strictly convex simple closed curve. The
width w(t) of γ at t ∈ [0, L] is the distance between the tangent line at γ(t) and the
tangent line at the unique point γ(t0 ) satisfying e1 (t0 ) = −e1 (t) (see Exercise 1.5).
A curve has constant width if w is independent of t. Prove that if γ has constant
width then:
(1) The line between γ(t) and γ(t0 ) is perpendicular to the tangent lines at
those points.
(2) The curve γ has length L = πw.
2Reversing the orientation of γ if necessary.
EXERCISES 17
Exercise 1.8. Prove that if a convex simple closed curve has four vertices,
then it cannot meet any circle in more than four points.
CHAPTER 2
1. Surfaces
Definition 2.1. A parametric surface patch is a smooth mapping:
X : U → R3 ,
where U ⊂ R2 is open, and the Jacobian dX is non-singular.
Write X = (x1 , x2 , x3 ), and each xi = xi (u1 , u2 ), then the Jacobian has the
matrix representation: 1
x1 x12
dX = x21 x22
x31 x32
where we have used the notation fi = fui = ∂f /∂ui . According to the definition,
we are requiring that this matrix has rank 2, or equivalently that the vectors X1 =
(x11 , x21 , x31 ) and X2 = (x12 , x22 , x32 ) are linearly independent. Another equivalent
requirement is that dX : R2 → R3 is injective.
Example 2.1. Let U ⊂ R2 be open, and suppose that f : U → R is smooth.
Define the graph of f as the parametric surface X(u1 , u2 ) = (u1 , u2 , f (u1 , u2 )). To
verify that X is indeed a parametric surface, note that:
1 0
dX = 0 1
f1 f2
so that clearly X is non-singular.
A diffeomorphism between open sets U, V ⊂ R2 is a map φ : U → V which is
smooth, one-to-one, and whose inverse is also smooth. If det(dφ) > 0, then we say
that φ is an orientation-preserving diffeomorphism.
Definition 2.2. Let X : U → R3 , and X̃ : Ũ → R3 be parametric surfaces.
We say that X̃ is reparametrization of X if X̃ = X ◦ φ, where φ : Ũ → U is a
diffeomorphism. If φ is an orientation-preserving diffeomorphism, then X̃ is an
orientation-preserving reparametrization.
Clearly, the inverse of a diffeomorphism is a diffeomorphism. Thus, if X̃ is a
reparametrization of X, then X is a reparametrization of X̃.
Definition 2.3. The tangent space Tu X of the parametric surface X : U → R3
at u ∈ U is the 2-dimensional linear subspace of R3 spanned by the two vectors X1
and X2 .1
1Note that the tangent plane to the surface X(U ) at u is actually the affine subspace X(u) +
Tu X. However, it will be very convenient to have the tangent space as a linear subspace of R3 .
19
20 2. LOCAL SURFACE THEORY
of g is at each point u0 ∈ U an
Thus, the so-called coordinate representation
instance of Example 2.4. In fact, if A = gij , and B(ξ, η) = ξ · Aη for ξ, η ∈ R2 as
in Example 2.4, then B is the pull-back by dXu : R2 → Tu X of the restriction of
the Euclidean inner product on Tu X.
22 2. LOCAL SURFACE THEORY
The classical (Gauss) notation for the first fundamental form is g11 = E, g12 =
g21 = F , and G = g22 , i.e.,
E F
gij =
F G
Clearly, F 2 < EG, and another condition equivalent to the condition that X1 and
X2 are linearly independent is that det gij = EG − F 2 > 0. The first fundamental
form is also sometimes written:
Note that the gij ’s are functions of u. The reason for the notation ds2 is that the
square root of the first fundamental form can be used to compute length of curves
on X. Indeed, if γ : [a, b] → R3 is a curve on X, then γ = X ◦ β, where β is a curve
in U . Let β(t) = β 1 (t), β 2 (t) , and denote time derivatives by a dot, then
Z b Z b q
Lγ ([a, b]) = |γ̇| dt gij β̇ i β̇ j dt.
a a
and the angle between two curves β and γ on X is the angle between their tangents
β̇ and γ̇. Intrinsic geometry is all the information which can be obtained from the
three functions gij and their derivatives.
Clearly, the first fundamental form is invariant under reparametrization. The
next proposition shows how the gij ’s change under reparametrization.
∂uk ∂ul
(2.4) g̃ij = gkl ,
∂ ũi ∂ ũj
where dφ = (∂ui /∂ ũj ).
To prove (2), extend Y and Z to be vector fields in a neighborhood of u, and use (1):
∂Y N · Z − ∂Z N · Y = −N · ∂Y Z − ∂Z Y = 0.
Note that while proving the proposition, we have established the following
formula for the commutator:
[Y, Z] = y i ∂i z j − z i ∂i y j Xj
(2.5)
Definition 2.8. Let X : U → R3 be a surface, and let N : U → S2 be its
unit normal. The second fundamental form of X is the symmetric bilinear form k
defined on each tangent space Tu X by:
(2.6) k(Y, Z) = −∂Y N · Z.
We remark that since N · N = 1, we have ∂Y N · N = 0, hence ∂Y N is tangen-
tial. Thus, according to (2.6), the second fundamental form is minus the tangential
directional derivative of the unit normal, and hence measures the turning of the tan-
gent plane as one moves about on the surface. Note that part (2) of the proposition
guarantees that k is indeed a symmetric bilinear form. Note that it is not neces-
sarily positive definite. Furthermore, if we set kij = k(Xi , Xj ) to be the coordinate
representation of the second fundamental form, then we have:
(2.7) kij = Xij · N.
24 2. LOCAL SURFACE THEORY
`ji = kim g mj .
It is customary to say that g raises the index of k and to write the new object
ki j = kim g mj . Here since kij is symmetric, it is not necessary to keep track of the
position of the indices, and hence we write: `ji = kij . In particular, we have:
1 i det kij
(2.10) H = ki , K= .
2 det gij
Now, k ij = g im g jl klm , and we have
2
|k| = kij k ij = tr `2 = k12 + k22 = 4H 2 − 2K.
Hence, we conclude
1 2
(2.11) K = 2H 2 − |k|
2
4. Examples
In this section, we use u = u, and u2 = v in order to simplify the notation.
1
4.2. Spheres. Let U = (0, π) × (0, 2π) ⊂ R2 , and let X : U → R2 be given by:
X(u, v) = (sin u cos v, sin u sin v, cos u).
The surface X is a parametric representation of the unit sphere. A straightforward
calculation shows that the first fundamental form is:
ds2 = du2 + sin2 u dv 2 ,
and the unit normal is N = X. Thus, Ni = Xi , and consequently kij = −Ni · Xj =
−Xi · Xj = −gij , i.e., k = −g. In particular, the principal curvatures are both
equal to −1 and all the points are umbilical. We have for the mean and Gauss
curvatures:
H = −1, K=1
Proposition 2.7. Let X : U → R3 be a parametric surface and suppose that
all the points of X are umbilical. Then, X(U ) is either contained in a plane or a
sphere.
Proof. By hypothesis, we have
(2.12) Ni = λXi .
We first show that λ is a constant. Differentiating (2.12), we get Nij = λj Xi +λXij .
Interchanging i and j, subtracting these two equations, and taking into account
Nij − Nji = Xij − Xji = 0, we obtain λi Xj − λj Xi = 0, e.g.,
λ1 X2 − λ2 X1 = 0.
Since X1 and X2 are linearly independent, we conclude that λ1 = λ2 = 0 and
it follows that λ is constant. Now, if λ = 0 then all points are planar, and by
Proposition 2.6, X is contained in a plane. Otherwise, let A = X − λ−1 N , then A
is constant:
Ai = Xi − λ−1 Ni = 0,
−1
and |X − A| = |λ| is also constant, hence X is contained in a sphere.
4. EXAMPLES 27
is the directrix , and the lines γ(u) + tY (u) for u fixed are the generators of X. We
may assume that Y is a unit vector field. Provided Ẏ 6= 0. We will also assume that
Ẏ 6= 0. In this case, it is possible to arrange by reparametrization that γ̇ · Ẏ = 0,
in which case γ is said to be a line of striction. Indeed, if this is not the case, then
we can set φ = γ̇ · Ẏ /|Ẏ |2 , and note that the curve
α = γ + φY
lies on the surface X, and satisfies α̇ · Ẏ = 0. Consequently, the surface:
X̃(s, t) = α(s) + tY (s)
is a reparametrization of X. Furthermore, there is only one line of striction on X.
Indeed, if β and γ are two lines of striction, then since both β is a curve on X we
may write β = γ + φY for some function φ and consequently:
β̇ = γ̇ + φ̇Y + φẎ .
Taking
2 inner product with Ẏ and using the fact that Y is a unit vector, we obtain
φẎ = 0 which implies that φ = 0 and thus, β = γ.
We have Xu = γ̇ + v Ẏ , Xv = Y , and Xvv = 0. Thus, the first fundamental is:
!
1 + v 2 |Ẏ |2 γ̇ · Y
gij =
γ̇ · Y 1
and 2
det gij = 1 + v 2 |Ẏ |2 + γ̇ · Y > v 2 |Ẏ |2 .
and
2 2
|Nv | = .
1 + 4v 2 + 4v 4
It follows from Proposition 2.8 that X has Gauss curvature K < 0.
5. Lines of Curvature
Definition 2.10. A curve γ on a parametric surface X is called a line of
curvature if γ̇ is a principal direction.
The following proposition, due to Rodriguez, characterizes lines of curvature
as those curves whose tangents are parallel to the tangent of their spherical image
under the Gauss map.
Proposition 2.9. Let γ be a curve on a parametric surface X with unit normal
N , and let β = N ◦ γ be its spherical image under the Gauss map. Then γ is a line
of curvature if and only if
(2.13) β̇ + λγ̇ = 0.
Proof. Suppose that (2.13) holds, then we have:
∂γ̇ N + λγ̇ = 0.
Let ` be the linear transformation on Tu X associated with k as defined by (2.9).
Then, we have for every Y ∈ Tu X:
g `(γ̇), Y ) = k γ̇, Y = −∂γ̇ N · Y = λg λγ̇, Y .
Thus, ` γ̇ = λγ̇, and γ̇ is a principal direction. The proof of the converse is
similar.
5. LINES OF CURVATURE 29
It is clear from the proof that λ in (2.13) is the associated principal curvature.
The coordinate curves of a parametric surface X are the two family of curves
γc (t) = X(t, c) and βc (t) = X(c, t). A surface is parametrized by lines of curvature
if the coordinate curves of X are lines of curvature. We will now show that any non-
umbilical point has a neighborhood in which the surface can be reparametrized by
lines of curvature. We first prove the following lemma which is also of independent
interest.
Proof. Suppose that (1) holds. Then Equation (2.5) shows that [X̃1 , X̃2 ] = 0.
However, since the commutator is invariant under reparametrization, it follows that
[Y1 , Y2 ] = 0.
Conversely, suppose that [Y1 , Y2 ] = 0. Express Xi = aji Yj and Yi = bji Xj , and
note that bji is the inverse of aji . We now calculate:
0 = [Xi , Xj ]
= aki Yk , alj Yl
= ali bm k l m k
l ∂m aj − aj bl ∂m ai Yk
= ∂i akj − ∂j aki Yk .
Proof. By Lemma 2.10 is suffices to show that there are function fi such that
f1 Y1 and f2 Y2 commute. Write [Y1 , Y2 ] = a1 Y1 − a2 Y2 , and compute:
[f1 Y1 , f2 Y2 ] = f1 f2 a1 Y1 − a2 Y2 + f1 ∂Y1 f2 Y2 − f2 ∂Y2 f1 Y1 .
Thus, the commutator [f1 Y1 , f2 Y2 ] vanishes if and only if the following two equa-
tions are satisfied:
∂Y2 f1 − a1 f1 = 0
∂Y1 f2 − a2 f2 = 0.
We can rewrite those as:
∂Y2 log f1 = a1
∂Y1 log f2 = a2 .
Each of those equation is a linear first-order partial differential equation, and can
be solved for a positive solution in a neighborhood of u0 .
In a neighborhood of a non-umbilical point, the principal directions define two
orthogonal unit vector fields. Thus, we obtain the following Theorem as a corollary
to the above proposition.
Theorem 2.12. Let X : U → R3 be a parametric surface, and let u0 be a
non-umbilical point. Then there is neighborhood U0 of u0 and a diffeomorphism
φ : Ũ0 → U0 such that X̃ = X ◦ φ is parametrized by lines of curvature.
If X is parametrized by lines of curvature, then the second fundamental form
has the coordinate representation:
!
k1 g11 0
kij =
0 k2 g22
Definition 2.11. A curve γ on a parametric surface X is called an asymptotic
line if it has zero normal curvature, i.e., k(γ̇, γ̇) = 0.
The term asymptotic stems from the fact that those curve have their tangent γ̇
along the asymptotes of the Dupin indicatrix , the conic section kij ξ i ξ j = 1 in the
tangent space. Since the Dupin indicatrix has no asymptotes when K > 0, we see
that the Gauss curvature must be non-positive along any asymptotic line.
The following Theorem can be proved by the same method as used above to
obtain Theorem 2.12.
Theorem 2.13. Let X : U → R3 be a parametric surface, and let u0 be a hyper-
bolic point. Then there is neighborhood U0 of u0 and a diffeomorphism φ : Ũ0 → U0
such that X̃ = X ◦ φ is parametrized by asymptotic lines.
6. More Examples
A surface of revolution is a parametric surface of the form:
X(u, v) = f (u) cos(v), f (u) sin(v), g(u) ,
where f (t), g(t) is a regular curve, called the generator , which satisfies f (t) 6= 0
. Without loss of generality, we may assume that f (t) > 0. The curves
γv (t) = f (t) cos(v), f (t) sin(v), g(t) , v fixed.
6. MORE EXAMPLES 31
In order to construct parametric surfaces which are both conformal and har-
monic,
√ we will use complex analysis in the domain U . Let ζ = u+iv where i denotes
−1, and let f (ζ) and h(ζ) be two complex analytic functions on U . Define
F1 = f 2 − h2 , F 2 = i f 2 + h2 ,
F3 = 2f h.
We have:
2 2 2
F1 + F2 + F3 = 0.
If we write Fj = ξj + iηj , then this can be written as:
3 h 3
X 2 2 i2 X
ξj − ηj + 2i ξj ηj = 0.
j=1 j=1
and
3
X
Xu · Xv = − ξj ηj = 0,
j=1
and hence, X is conformal.3 Since X is real analytic, the zeroes of det Xi · Xj are
7. Surface Area
In this section we will give interpretations of the Gauss curvature and the mean
curvature. Both of these involve the concept of surface area. Before introducing
the definition, we first prove a proposition which will show that the definition is
reparametrization invariant.
Proposition 2.16. Let X : U → R3 be a parametric surface with first funda-
mental form gij , and V ⊂ U . Let X̃ : Ũ → R3 be a reparametrization of X, let
Ṽ = φ−1 (V ), and let g̃ij be the coordinate representation of the first fundamental
where φij = ∂ui /∂ ũj . Thus, for any open subset V ⊂ U , and Ṽ = φ−1 (V ), we have:
Z q Z q Z q
det g̃ij dũ1 dũ2 = det gij det φij dũ1 dũ2 = det gij du1 du2
Ṽ Ṽ V
Thus, the integral on the right-hand side of (2.15) is reparametrization invari-
ant. This justifies the following definition.
Definition 2.14. Let X : U → R3 be a parametric surface and let gij be its
first fundamental form. The surface area element of X is:
q
dA = det gij du1 du2 .
It is easy to show, as in the proof of Proposition 2.16 that the total curvature
of X over V is invariant under reparametrization. We now introduce the signed
surface area, a variant of Definition 2.14 which allows for smooth maps Y into a
surface X, with Jacobian dY not necessarily everywhere non-singular, and which
also accounts for multiplicity.
Definition 2.16. Let X : U → R3 be a parametric surface, and let Y : U →
X(U ) be a smooth map. Define σ(u) to be 1, −1, or 0, according to whether the
pair Y1 (u), Y2 (u) has the same orientation as the pair X1 (u), X2 (u), the opposite
34 2. LOCAL SURFACE THEORY
In particular,
2
det kij
det hij =
det gij
Note also that Equation (2.17) implies
that the pair N1 , N2 has the same orientation
as X1 , X2 if and only if det kij > 0. Furthermore, since N (u) is also the outward
normal to the unit sphere at N (u), and since X1 , X2 , N is positively oriented in
R3 , it follows that X1 (u), X2 (u) also gives the positive orientation
on the tangent
space to the S2 at N (u). Thus, we deduce that sign det kij = σ. Consequently, in
view of Equation (2.10), we obtain:
q sign det kij det hij q
σ det hij = q = K det gij
det gij
The proposition follows by integrating over V .
We now turn to an interpretation of the mean curvature. Let X : U → be a
parametric surface. A variation of X is a smooth family F (u; t) : U × (−ε, ε) → R3
such that F (u; 0) = X. Note that since dF (u; 0) is non-singular, the same is true
of dF (u; t0 ) for any fixed u0 , perhaps after shrinking the interval (−ε, ε). Thus, all
the maps F (u; t0 ) for t0 close enough to 0 are parametric surfaces. The generator
of the variation is the vector field dF/dt(u; 0). The variation is compactly supported
if F (u; t) = X(u) outside a compact subset of U . The smallest such compact set is
called the support of the variation F . Clearly, if a variation is compactly supported,
then the support of its generator is compact in U . We say that a variation is
tangential if the generator is tangential; we say it is normal if the generator is
normal. Suppose now that the closure V is compact in U . We consider the area
AF (V ) of F (u; t) as a function of t. The next proposition shows that the derivative
of this function depends only on the generator, and in fact is a linear functional in
the generator.
7. SURFACE AREA 35
We also have that det B = det D. Thus taking into the account that (2.19) holds
for for D:
0 0
log det B = log det D = tr D−1 D0 = tr B −1 B 0 .
In order to prove the general case, it is more convenient to look at the equivalent
identity:
0
det B = tr (det B)B −1 B 0 .
(2.20)
Note that by Kramer’s rule, the matrix (det B)B −1 is the matrix of co-factors
of B, hence its components being determinants of minors of B, are multivariate
polynomials in the components of B. Thus, both sides of the identity (2.20) are
linear polynomials
n
X n
X
p(B 0 ; B) = pij (B)b0ij , q(B 0 ; B) = qij (B)b0ij ,
i,j=1 i,j=1
Proof of Proposition 2.18. Differentiating the area (2.16) under the inte-
gral sign, and using (2.21), we get:
Z Z
dAF (V ) 1 dgij 1 dgij
q
g ij det gij du1 du2 = g ij
= dA.
dt 2 V dt 2 V dt
Since Y is smooth, we have at t = 0 that dFi /dt = (dF/dt)i = Yi , and thus
dgij
g ij = g ij Yi · Xj + Xi · Yj = 2g ij Xi · Yj .
dt
This completes the proof of the proposition.
Since the variation of the area dAF (V )/dt is a linear functional in the generator
dF/dt of the variation, it is possible to decompose any variation into tangential and
normal components. We begin by showing that the area doesn’t change under a
tangential variation. This is simply the infinitesimal version of Proposition (2.16).
Proposition 2.20. Let X : U → R3 be a parametric surface, and let F (u; t) be
a compactly supported tangential variation. If V ⊂ U is open with V compact in
U , and the support of F contained in V , then dAF (V )/dt = 0.
Proof. Let Y be the generator of F (u; t). We will show that there is a smooth
family of diffeomorphisms φ : U × (−δ, δ) → U such that Y is also the generator of
the variation G = X ◦ φ. This proves the proposition since Proposition 2.16 gives
that AG (U ) is constant. Since Y is tangential, we can write Y = y i Xi . Consider
the initial value problem:
dv i
= y i (v), v i (0) = ui .
dt
Since the y i ’s are compactly supported, a solution v = v(u; t) exists for all t.
Defining φ(u; t) = v(u; t), then an application of the inverse function theorem shows
that φ(u; t) is a diffeomorphism for t in some small interval (−δ, δ). Finally, we see
that:
dX ◦ φ dv i
= Xi = Xi y i = Y.
dt dt
Our next theorem gives an interpretation of the mean curvature as a measure
of surface area variation under normal perturbations.
Theorem 2.21. Let X : U → R3 be a parametric surface, and let F (u; t) be a
compactly supported variation with generator Y . If V ⊂ U is open with V compact
in U , and the support of F contained in V , then
Z
dAF (V )
(2.22) = −2 (Y · N ) H dA.
dt V
Proof. By Propositions 2.18 and 2.20, it suffices to consider normal variations
with generator Y = f N . In that case, we find that Yj = fj N + f Nj , so that
g ij Xi · Yj = f g ij Xi · Nj = −f kii = −2f H. The theorem follows by substituting
into (2.18).
Definition 2.17. A parametric surface X is area minimizing if AX (U ) 6
AX̃(U ) for any parametric surface X̃ such that X̃ = X on the boundary of U .
A parametric surface X : U → R3 is locally area minimizing if for any compactly
supported variation F (u; t), the area AF (U ) has a local minimum at t = 0.
8. BERNSTEIN’S THEOREM 37
8. Bernstein’s Theorem
In this section, we prove Bernstein’s Theorem: A minimal surface which is a
graph over an entire plane must itself be a plane. We say that a surface X is a
graph over a plane Y : R2 → R3 , where Y is linear, if there is a function f : R2 → R
such that X = Y + f N where N is the unit normal of Y .
Theorem 2.23 (Bernstein’s Theorem). Let X be a minimal surface which is a
graph over an entire plane. Then X is a plane.
We may without loss of generality assume that X is a graph over the plane
Y (u, v) = (u, v, 0), i.e. X(u, v) = u, v, f (u, v) as in example 2.1. It is then
straightforward to check that X is a minimal surface if and only if f satisfies the
non-parametric minimal surface equation:
(2.23) (1 + q 2 )pu − 2pqpv + (1 + p2 )qv = 0,
where we have used the classical notation: p = fu , q = fv . We say that a solution
of a partial differential equation defined on the whole (u, v)-plane is entire. Thus,
to prove Bernstein’s Theorem, it suffices to prove that any entire solution of (2.23)
is linear.
Proposition 2.24. Let f be an entire solution of (2.23). Then f is a linear
function.
By Exercise 2.7, if f satisfies (2.23), then p and q satisfy the following equations:
! !
∂ 1 + q2 ∂ pq
(2.24) p = p ,
∂u 1 + p2 + q 2 ∂v 1 + p2 + q 2
! !
∂ pq ∂ 1 + p2
(2.25) p = p .
∂u 1 + p2 + q 2 ∂v 1 + p2 + q 2
Since the entire plane is simply connected, Equation (2.25) implies that there exists
a function ξ satisfying:
1 + p2 pq
ξu = p , ξv = p ,
1 + p2 + q 2 1 + p2 + q 2
and Equation (2.24) implies that there exists a function η satisfying:
pq 1 + q2
ηu = p , ηv = p .
1 + p2 + q 2 1 + p2 + q 2
Furthermore, ξv = ηu , hence there is a function h so that hu = ξ, hv = η. The
Hessian of the function h is:
! !
huu huv ξu ξv
hij = = ,
hvu hvv ηu ηv
38 2. LOCAL SURFACE THEORY
9. Theorema Egregium
In this section, we prove that the Gauss curvature can be computed in terms
of the first fundamental form and its derivatives. We then prove the Fundamental
Theorem for surfaces in R3 , analogous to Theorem 1.2 for curves, which states that a
parametric surface is uniquely determined by its first and second fundamental form.
Partial derivatives with respect to ui will be denoted by a subscript i following a
comma, unless there is no ambiguity in which case the comma may be omitted.
Proposition 2.26. Let X : U → R3 be a parametric surface. Then the follow-
ing equations hold:
(2.27) Xij = Γm
ij Xm + kij N,
where,
1 mn
Γm
(2.28) ij = g gni,j + gnj,i − gij,n ,
2
and gij and kij are the coordinate representations of its first and second fun-
damental form.
Proof. Clearly, Xij can be expanded in the basis X1 , X2 , N of R3 . We al-
ready saw in Equation (2.7), that the component of Xij along N is kij , hence
Equation (2.27) holds with the coefficients Γm
ij given by
Substituting Xml from (2.27) and Nl from (2.17), and decomposing into tangential
and normal components, we obtain:
Xijl = Am
ijl Xm + Bijl N,
where:
Am m n m
ijl = Γij,l + Γij Γnl − g
mn
kij kln ,
Bijl = kij,l + Γm
ij klm .
Taking note of the fact that Xijl = Xilj , we now interchange j and l and subtract
to obtain (2.30) and (2.31).
Equation (2.30) is called the Gauss Equation, and Equation (2.31) is called
the Codazzi Equation. The Gauss Equation has the following corollary which has
been coined Theorema Egregium. It’s discovery marked the beginning of intrinsic
geometry, the geometry of the first fundamental form.
Corollary 2.28. Let X : U → R3 be a parametric surface. Then the Gauss
curvature K of X can be computed in terms of only its first fundamental form gij
and its derivatives up to second order:
1
K = g ij Γm m n m n m
ij,m − Γim,j + Γij Γnm − Γim Γnj ,
2
where Γmij are the Christoffel symbols of the first kind.
(2.33) Ni = −kij g jm Xm ,
where Γm ij is defined in terms of gij by (2.28). The integrability conditions for
this system are:
Γm m
(2.34) ij Xm + kij N l = Γil Xm + kil N j
kij g jm Xm l = klj g jm Xm i .
(2.35)
6Here X is not to be understood as the derivative of X with respect to ui until later in the
i
proof.
EXERCISES 41
The proof of Theorem 2.27 also shows that the Gauss-Codazzi Equations (2.30)–
(2.31) imply (2.34) if Xi and N satisfy (2.32) and (2.33). We now check that (2.31)
also implies (2.35). First note that since Γm
ij is defined by (2.28), we have
1
Γm
ij gmn = gni,j + gnj,i − gij,n .
2
Interchanging n and i and adding, we get (2.29). Now, differentiate (2.33), and
taking into account that g,lij = −g ia gab,l g bj , substitute (2.29) to get:
Note that the last term is symmetric in i and l so that interchanging i and l, and
subtracting, we get:
Ni,l − Nl,i = −kij,l + kil,j − Γnij kln + Γnil kjn g jm Xm
which vanishes by (2.31). Thus, it follows that (2.35) is satisfied. We conclude that
given values for X1 , X2 , N at a point u0 ∈ U there is a unique solution of (2.32)–
(2.33) in U . We can choose the initial values to that Xi · Xj = gij , N · Xi = 0, and
N · N = 1 at u0 . Using (2.32) and (2.33), it is straightforward to check that the
functions hij = Xi · Xj , pi = N · Xi and q = N · N , satisfy the differential equations:
hij,l = Γnil hnj + Γnjl hni + kil pj + kjl pi ,
qi = −2kij g jm pm .
However, the functions hij = gij , pi = 0 and q = 1 also satisfy these equations, as
well as the same initial conditions as hij = Xi · Xj , pi = N · Xi and q = N · N at u0 .
Thus, by the uniqueness statement mentioned above, it follows that Xi · Xj = gij ,
N · Xi = 0, and N · N = 1. Clearly, in view of (2.32) we have Xi,j = Xj,i , hence
3
there is a function X : U → R whose partial derivatives are Xi , cf. foonote 6. Since
gij is positive definite we have that X1 , X2 are linearly independent, hence X is a
parametric surface with first fundamental form gij . Furthermore, it is easy to see
that the unit normal of X is N , and Ni · Xj = −N · Xij = −kij , hence the second
fundamental form of X is kij . This completes the proof of the existence statement.
Assume now that X̃ is another surface with the same first and second fun-
damental forms. Since X and X̃ have the same first fundamental form, it fol-
lows that there is a rigid motion R(x) = Qx + y with Q ∈ SO(n; R) such that
R X(u0 ) = X̃(u0 ), QXi (u0 ) = X̃i (u0 ), QN (u0 ) = Ñ (u0 ). Let X̂ = R ◦ X. Since
the two triples (X̃1 , X̃2 , Ñ ) and (X̂1 , X̂2 , N̂ ) both satisfy the same partial differen-
tial equations (2.32) and (2.33), it follows follows that they are equal everywhere,
and consequently X̃ = X̂ = R ◦ X.
Exercises
Exercise 2.1. Let X : U → R3 and X̃ : Ũ → R3 be two parametric surfaces.
The angle θ between them is the angle between their unit normals: cos θ = N · Ñ .
Let γ be a regular curve which lies on both X and X̃, and suppose that the angle
42 2. LOCAL SURFACE THEORY
Exercise 2.4. Let M n×n be the space of all n×n matrices, and let B : (a, b) →
n×n
M be continuously differentiable. Prove that:
0
det B = tr(B ∗ B 0 ),
where B ∗ is the matrix of co-factors of B.
Exercise 2.6. Prove that setting f (ζ) = 1, g(ζ) = 1/ζ in the Weierstrass
representation, we get the catenoid. Find the conjugate harmonic surface of the
catenoid.
In this chapter, we change our point of view, and study intrinsic geometry, in
which the starting point is the first fundamental form. Thus, given a parametric
surface, we will ignore all information which cannot be recovered from the first
fundamental form and its derivatives only. In particular, we will ignore the Gauss
map and the second fundamental form. Thanks to Gauss’ Theorema Egregium, we
will still be able to take the Gauss curvature into account.
1. Riemannian Surfaces
Definition 3.1. Let U ⊂ R2 be open. A Riemannian metric on U is a smooth
function g : U → S2×2
+ . A Riemannian surface patch is an open set U equipped
with a Riemannian metric.
The tangent space of U at u ∈ U is R2 . The Riemannian metric g defines an
inner-product on each tangent space by:
g(Y, Z) = gij y i z j ,
where y i and z j are the components of Y and Z with respect to the standard
2
basis of R2 . We will write |Y |g = g(Y, Y ), and omit the subscript g when it is not
ambiguous.
Two Riemannian surface patches (U, g) and (Ũ , g̃) are isometric if there is a
diffeomorphism φ : Ũ → U such that
(3.1) g̃ij = glm φli φm
j ,
Example 3.1. Let U ⊂ R2 be open, and let δij be the identity matrix,
then (U, δ) is a Riemannian surface. The Riemannian metric δ will be called the
Euclidean metric.
Example 3.2. Let X : U → R3 be a parametric surface, and let g be the
coordinate representation of its first fundamental form, then (U, g) is a Riemannian
surface patch. We say that the metric g is induced by the parametric surface X. If
X̃ = X ◦ φ : Ũ → R3 is a reparametrization of X and g̃ the coordinate representation
of its first fundamental form, then (Ũ , g̃) is isometric to (U, g).
Example 3.3 (The Poincaré Disk). Let D = {(u, v) : u2 + v 2 < 1} be the unit
disk in R2 , and let
4
gij = δij
(1 − r2 )2
√
where r = u2 + v 2 is the Euclidean distance to the origin. We can write this line
element also as
du2 + dv 2
(3.2) ds2 = 4 .
(1 − u2 − v 2 )2
The Riemannian surface (D, g) is called the Poincaré Disk. Let U = {(x, y) : y > 0}
be the upper half-plane, and let
1
hij = 2 δij .
y
Then it is not difficult to see that D, gij and U, hij are isometric with the
isometry given by:
1 − u2 − v 2
2v
φ : (u, v) 7→ (x, y) = , .
(1 + u)2 + v 2 (1 + u2 ) + v 2
In fact, a good bookkeeping technique to check this type of identity is to compute
the differentials :
v(1 + u) (1 + u)2 − v 2
dx = −4 2 du + 2 2 dv
(1 + u)2 + v 2 (1 + u)2 + v 2
(1 + u)2 − v 2 v(1 + u)
dy = −2 2 du + 4 2 dv,
2
(1 + u) + v 2 (1 + u)2 + v 2
substitute into
dx2 + dy 2
,
y2
and then simplify using du dv = dv du to obtain (3.2). It is not difficult to see that
this is equivalent to checking (3.1).
Definition 3.2. Let (U, g) be a Riemannian surface. The Christoffel symbols
of the second kind of g are defined by:
1 mn
Γm
(3.3) ij = g gni,j + gnj,i − gij,n .
2
The Gauss curvature of g is defined by:
1
K = g ij Γm m n m n m
(3.4) ij,m − Γim,j + Γij Γnm − Γim Γnj .
2
2. LIE DERIVATIVE 47
2. Lie Derivative
In this section, we study the Lie derivative. We denote the standard basis on
R2 by ∂1 , ∂2 . Let f be a smooth function on U , and let Y = y i ∂i ∈ Tu U be a vector
at u ∈ U . The directional derivative of f along Y is:
(3.5) ∂Y f = y i ∂i f = y i fi .
Since y i = ∂Y ui where (u1 , u2 ) are the coordinates on U , we see that Y = Z
follows from ∂Y = ∂Z as operators. The next proposition shows that the directional
derivative of a function is reparametrization invariant.
Proposition 3.1. Let φ : Ũ → U be a diffeomorphism, and let Ỹ be a vector
at ũ ∈ Ũ . Then for any smooth function f on U , we have:
∂dφ(Ỹ ) f ◦ φ = ∂Ỹ (f ◦ φ).
i i
We define the commutator of two tangent vector fields Y = y ∂i and Z = z ∂i ,
as in Section (3), Equation (2.5):
[Y, Z] = y i ∂i z j − z i ∂i y j ∂j .
(3.6)
Note that
(3.7) ∂[Y,Z] f = ∂Y ∂Z f − ∂Z ∂Y f.
This observation together with Proposition 3.1 are now used to show that the
commutator is reparametrization invariant.
Proposition 3.2. Let Ỹ and Z̃ be vector fields on Ũ , and let φ : Ũ → U be a
diffeomorphism, then
dφ [Ỹ , Z̃] = dφ(Ỹ ), dφ(Z̃) .
Proof. For any smooth function f on U , we have:
(3.8) ∂ [Ỹ ,Z̃] (f ◦ φ) = ∂Ỹ ∂Z̃ (f ◦ φ) − ∂Z̃ ∂Ỹ (f ◦ φ)
f = ∂
dφ [Ỹ ,Z̃]
= ∂Ỹ ∂dφ(Z̃) f ◦ φ − ∂Z̃ ∂dφ(Ỹ ) f ◦ φ = ∂dφ(Ỹ ) ∂dφ(Z̃) f − ∂dφ(Z̃) ∂dφ(Ỹ ) f
= ∂ f,
dφ(Ỹ ),dφ(Z̃)
3. Covariant Differentiation
Definition 3.3. Let (U, g) be a Riemannian metric, and let Z be a vector field
on U . The covariant derivative of Z along ∂i is:
∇i Z = ∂i z j + Γjik z k ∂j .
(3.9)
Let Y ∈ Tu U , the covariant derivative of Z along Y is:
∇Y Z = y i Z;i .
We write the components of ∇i Z as:
(3.10) z j ;i = z j ,i + Γjik z k ,
so that ∇Y Z = y i z j ;i ∂j . Furthermore, note that
(3.11) ∇i ∂j = Γkij ∂k .
Our first task is to show that covariant differentiation is reparametrization
invariant. However, since the metric g was used in the definition of the covari-
ant derivative, it stands to reason that it would be invariant only under those
reparametrization which preserve the metric, i.e., under isometries.
Proposition 3.3. Let φ : (Ũ , g̃) → (U, g) be an isometry. Let Ỹ ∈ Tũ Ũ , and
let Z̃ be a vector field on Ũ . Then
(3.12) ˜ Z̃) = ∇
dφ(∇ Ỹ dφ(Ỹ ) dφ(Z̃).
1Indeed ∂ Z as defined in Chapter 2 does depend only on the value of Y at a single point
Y
and satisfies ∂f Y Z = f ∂Y Z.
3. COVARIANT DIFFERENTIATION 49
∂i g(Y, Z) = ∂i gjk y j z k = Γm j k m j k j k j k
ji gkm y z + Γki gmj y z + gjk y ,i z + gjk y z ,i
4. Geodesics
Definition 3.6. Let (U, g) be a Riemannian surface, and let γ : I → U be
a curve. A vector field along γ is a smooth function Y : I → R2 . The covariant
derivative of Y = y i ∂i along γ is the vector field:
∇γ̇ Y = ẏ i + Γijk y j γ̇ k ∂i .
Given initial conditions y i (a) = y0i , the existence and uniqueness of a solution on
[a, b] follows from the theory of ordinary differential equations.
The proposition together with the comment preceding it shows that parallel
translation along a curve γ is an isometry between inner-product spaces Pγ : Ta U →
Tb U .
Definition 3.8. A curve γ is a geodesic if its tangent γ̇ is parallel along γ:
∇γ̇ γ̇ = 0.
If γ is a geodesic, then |γ̇| is constant and hence, every geodesic is parametrized
proportionally to arclength. In particular, if β = γ ◦ φ is a reparametrization of γ,
then β is not a geodesic unless φ is a linear map.
Proposition 3.8. Let (U, g) be a Riemannian surface, let u0 ∈ U and let
0 6= Y0 ∈ Tu0 U . Then there is and ε > 0, and a unique geodesic γ : (−ε, ε) → U ,
such that γ(0) = u0 , and γ̇(0) = Y0 .
Proof. We have:
∇γ̇ γ̇ = γ̈ i + Γijk γ̇ j γ̇ k ∂i .
Thus, the condition that γ is a geodesic can written as a pair of non-linear second-
order ordinary differential equations:
γ̈ i = −Γijk (γ(t))γ̇ j γ̇ k .
Given initial conditions γ i (0) = ui0 , γ̇ i (0) = y0i , there is a unique solution on defined
on a small enough interval (−ε, ε).
4. GEODESICS 51
Furthermore, we have:
R1212
(3.26) K= ,
det(g)
or equivalently:
Γjik,l + Γnik Γjnl = g jm g(∇l ∇k ∂i , ∂m ).
Proof. Denote the right-hand side of (3.27) by Sijkl , and note that it satis-
fies (3.17). Thus, the same comment which follows Proposition 3.14 applies and
the only non-zero components of Sijkl are determined by S1212 :
In view of (3.27), we have R1212 = S1212 , thus it follows that Rijkl = Sijkl
which is more in spirit with our derivation of the second variation formula. First
note that if σ is a fixed-endpoint variation of γ with generator σ 0 = Y , and with
σ̇ = X, then [X, Y ] = 0. Here Y denotes the vector field σ 0 along σ rather than
just along γ. Indeed, since X = dσ(d/dt) and Y = dσ(d/ds), it follows, as in
Propositions 3.1 and 3.2, that for any smooth function f on U , we have
d d
∂[X,Y ] f = , f ◦ σ = 0.
dt ds
In view of the symmetry Γijk = Γikj , this implies:
∇Y X − ∇X Y = [X, Y ] = 0.
We can now calculate:
Z Z b Z b
1
Eσ0 (s) = ∂Y g(X, X) dt = g(∇Y X, X) dt = g(∇X Y, X) dt
2 a a
Z b Z b Z b
d
= g(Y, X) dt − g(Y, ∇X X) dt = g(Y, X)|ba − g(Y, ∇X X) dt
a dt a a
where as above X = σ̇, and Y = σ 0 . Now, the first term integrates to g(∇Y Y, X)|ba =
0, and when we set s = 0, the second term also vanishes since ∇X X = ∇γ̇ γ̇ = 0.
Furthermore, the last term becomes g(∇γ̇ Y, ∇γ̇ Y ). Hence, we conclude:
Z b
Eσ00 (0) = ∇γ̇ Y 2 − R(X, Y, X, Y ) dt.
(3.31)
a
The proposition now follows from (3.28).
Thus, Eσ00 (0) can be viewed as a quadratic form in the generator Y . The
corresponding symmetric bilinear form is called the index form of γ:
Z b
I(Y, Z) = g(∇γ̇ Y, ∇γ̇ Z) − K ◦ γ g(Y, Z) − g(γ̇, Y ) g(γ̇, Z) dt.
a
It is the Hessian of the functional E, and if E has a local minimum, I is positive
semi-definite. We will also write I(Y ) = I(Y, Y ).
Definition 3.11. Let γ be a geodesic parametrized by arclength on the Rie-
mannian surface (U, g). A vector field Y along γ is called a Jacobi field , if it satisfies
the following differential equation:
∇γ̇ ∇γ̇ Y + K (Y − g(γ̇, Y )γ̇) = 0.
Two points γ(a) and γ(b) along a geodesic γ are called conjugate along γ if there
is a non-zero Jacobi field along γ which vanishes at those two points.
The Jacobi field equation is a linear system of second-order differential equa-
tions. Hence given initial data specifying the initial value and initial derivative of
Y , a unique solution exists along the entire geodesic γ.
Proposition 3.18. Let γ be a geodesic on the Riemannian surface (U, g). Then
given two vectors Z1 , Z2 ∈ Tγ(a) U , there is a unique Jacobi field Y along γ such
that Y (a) = Z1 , and ∇γ̇ Y (a) = Z2 .
In particular, any Jacobi field which is tangent to γ is a linear combination of
γ̇ and tγ̇. The significance of Jacobi fields is seen in the following two propositions.
We say that σ is a variation of γ through geodesics if the curves t 7→ σ(t; s) are
geodesics for all s.
Proposition 3.19. Let γ be a geodesic, and let σ be a variation of γ through
geodesics. Then the generator Y = σ 0 of σ is a Jacobi field.
Proof. As before, denote X = σ̇ and Y = σ 0 . We first prove the following
identity:
∇Y , ∇X X = −K Y − g(X, Y )X .
Indeed, in the proof of Lemma 3.13, it was seen that the left-hand side above is
a tensor, i.e., is linear over functions, and hence depends only on the values of
the vector fields X and Y at one point. Fix that point. If X and Y are linearly
dependent, then both sides of the equation above are zero. Otherwise, X and Y
are linearly independent, and it suffices to check the inner product of the identity
against X and Y . Taking inner product with X, both sides are zero, and equa-
tion (3.27) implies that the inner products with Y are equal. Since ∇X X = 0, we
get:
0 = ∇Y ∇X X = ∇X ∇Y X + ∇Y , ∇X X = ∇X ∇X Y − K Y − g(X, Y )X .
Thus, Y is a Jacobi field.
58 3. LOCAL INTRINSIC GEOMETRY OF SURFACES
If I(Z, Z) = 0, then f˙ = 0 and ḣY = 0 on [a, b]. Since Y 6= 0 on (a, b], we conclude
that ḣ = 0 on (a, b], and in view of h(b) = f (b) = 0, we get that Z = 0. Thus, I is
positive definite.
Exercises
Exercise 3.1. Two Riemannian metrics g and g̃ on an open set U ⊂ R2 are
conformal if g̃ = e2λ g for some smooth function λ.
(1) Prove that a parametric surface X : U → R3 is conformal if and only if its
first fundamental form g is conformal to the Euclidean metric δ on U .
(2) Let g̃ = e2λ g be conformal metrics on U , and let Γkij and Γ̃kij be their
Christoffel symbols. Prove that:
Γ̃kij = Γkij + δik λj + δjk λi + gij g km λm
60 3. LOCAL INTRINSIC GEOMETRY OF SURFACES
(3) Let g̃ and g be two conformal metrics on U , g̃ = e2λ g, and let K and K̃
be their Gauss curvatures. Prove that:
K̃ = e−2λ (K − ∆λ).
Index
angle, 22 Lie, 47
between surfaces, 41 developable, 28
exterior, 12 diffeomorphism, 19
arclength, 7, 45 differentials, 46
area minimizing, 36 directrix, 27
asymptotic line, 30, 42 distance, 22
divergence, 49
Bernstein’s Theorem, 37 Dupin indicatrix, 30
binormal, 8
Einstein summation convention, 20
catenoid, 31, 42 entire, 37
Cauchy-Riemann equations, 32, 38, 42 Euclidean metric, 46
Christoffel Symbols, 39 Euler, 24
Christoffel symbols, 46 evolute, 16
Codazzi Equation, 40 expanding map, 38, 43
commutator, 23, 47
conformal, 31, 59 Fenchel’s Theorem, 14
conjugate, 42 form
conjugate points, 56, 57 first fundamental, 21
convex, 12, 38 quadratic, 21
strictly, 16 second fundamental, 23
corner, 12 symmetric bilinear, 21
curvature Four Vertex Theorem, 13
center of, 16 Frenet frame, 8
Gauss, 25 Frenet frame equation, 8
line of, 29, 42 Fundamental Theorem
mean, 25 for curves in R3 , 8
normal, 24 for surfaces, 39, 40
of a curve, 8
Gauss curvature, 46
principal, see also principal, curvature
Gauss Equation, 40
Riemann tensor, 53
Gauss map, 20
total, see also total curvature generator, 27
curvature, line of, 28 geodesic, 50
curve geometry
closed, 10 intrinsic, 22
parametrized, 7 gradient, 43, 49
piecewise smooth, 12 graph, 19, 37
regular, 7
simple, 10 harmonic, 31
cylinder, 27 helix, 16
Hessian, 37
derivative hyperboloid, 28
covariant, 48
directional, 23, 47 index
61
62 INDEX
contravariant, 20 of a curve, 14
covariant, 20 under Gauss map, 28
raise, 25 star-shaped, 11
index form, 57 striction, line of, 27
induced metric, 46 surface
intrinsic geometry, 45 minimal, see also minimal surface
isometry, 21, 45 of revolution, 30
isoperimetric inequality, 17 generator, 30
parametric, 19
Jacobi field, 57 ruled, 27
Jacobi fields, 56 tangent, 28
surface area, 33
Laplacian, 49
element, 33
Leibniz, 47
signed, 34
length-minimizing, 51
locally, 51 tangent plane, 19
line element, 22, 45 tangent space, 19
tensor, 53
meridian, 31 Theorema Egregium, 39, 40
minimal surface, 31 torsion, 8
non-parametric, 37, 42 total curvature
Monge Ampère equation, 38 of a curve, 14
of a surface, 33
Nitsche, 38
normal section, 24 unit normal, 20
unit tangent, 8
orientation, 7, 19, 20
upper half-plane, 46
osculating paraboloid, 24
variation, 34, 51
parallel, 31 fixed-endpoint, 51
plane, 25 vector field, 20
normal, 9 vertex, 13
osculating, 9
rectifying, 9 Weierstrass representation, 31, 42
Poincaré Disk, 46 width, 16
point
ellitpic, 24
parabolic, 24
planar, 24, 26
umbilical, 25, 26
point,hyperbolic, 24
principal
curvature, 25
direction, 25
principal normal, 8
pull-back, 21
reparametrization
of curves, 7
of surfaces, 19
representation
coordinate, 21
Riemannian metric, 45
Riemannian surface, 45
Rodriguez, 28
rotation number, 10, 12
Rotation Theorem, 10, 12
sphere, 26
spherical image