You are on page 1of 54

DIFFERENTIAL GEOMETRY

IAN MCINTOSH
Figure 1. The Wente torus: an immersed torus with constant
mean curvature.
Contents
1. The Geometry of Space Curves. 3
1.1. Smooth Paths and Regular Paths. 3
1.2. Smooth Curves and Arc Length Parametrisation. 6
1.3. Curvature and torsion. 8
1.4. Congruence and Frenet formulas. 13
2. Smooth Surfaces and their Calculus. 17
2.1. Fundamental Concepts. 18
2.2. Charts, Atlases and Surfaces. 21
2.3. Tangent Spaces. 24
2.4. The Regular Value Theorem. 25
2.5. Maps between Surfaces. 27
Date: October 3, 2013.
1
2 IAN MCINTOSH
3. The Geometry of Smooth Surfaces. 30
3.1. The Riemannian Metric. 30
3.2. Lengths and Areas. 32
3.3. Isometries and Local Isometries. 33
3.4. The Shape Operator. 35
3.5. The Geometry of Curves on a Surface. 38
3.6. Normal Curvature and Principal Curvatures. 41
3.7. Gaussian and Mean Curvatures. 43
3.8. Geometric classication of points on a surface. 44
3.9. The Second Fundamental Form. 45
3.10. Gausss Theorema Egregium. 50
Appendix A. Brioschis intrinsic formulae for the Gaussian curvature. 52
References 54
DIFFERENTIAL GEOMETRY 3
1. The Geometry of Space Curves.
1.1. Smooth Paths and Regular Paths.
Denition 1.1. A smooth path in space is a smooth vector-valued function
p: I R
3
, where I R is an interval of positive length. We will call its
image in space its track.
Remark 1.1.
(i) By a smooth function we mean one whose r-th derivative exists for all
r N; synonyms are innitely dierentiable, or C

-smooth. (Note that


smooth sometimes just means that the rst derivative exists, and is con-
tinuous. However this is not sucient for our purposes.)
(ii) The interval I need not be open. If not, then the understanding is that
p(t) is dened and smooth on some open interval containing I. The
interval I need not be nite.
Example 1.1. Let p(t) = (t,

t, 0). This parameterises the track (x, y, z) : y =

x, z = 0, x 0. The path p : (0, ) R


3
is smooth, but the path
p : [0, ) R
3
is not, because
lim
t0
+
p

(t) = lim
t0
+
(1,
1
2

t
, 0) = (1, , 0).
This means it is impossible to nd a smooth extension of p(t) to any open interval
containing [0, ). Notice however that the path q(t) = (t
2
, t, 0) is smooth on
[0, ) (it is dened for all t R), with the same track as p.
If the path has components:
p(t) = (x(t), y(t), z(t))
then the derivative will be written as
dp
dt
=
_
dx
dt
,
dy
dt
,
dz
dt
_
, or p

(t) =
_
x

(t), y

(t), z

(t)
_
.
This is usually visualised, and referred to, as the tangent vector. It is also known
as the velocity vector, thinking of p(t) as the position of a moving particle at time
t.
0
p(t)
p

(t)
Figure 2. Position vector and tangent/velocity vector.
4 IAN MCINTOSH
Example 1.2 (Helix). Dene
p(t) = (a cos t, a sin t, bt), where a, b > 0. (1.1)
This path describes a right-handed helix on the cylinder x
2
+y
2
= a
2
. The number
2b is called the pitch of the helix. The tangent vector is:
x
y
z
t
2b = pitch
Figure 3. The right-handed helix.
p

(t) = (a sin t, a cos t, b),


which has constant length

a
2
+b
2
. Notice that the tangents describe a circle
lying in a horizontal plane.
1
It is very important to realise that smoothness is a property of the parametri-
sation, and not necessarily a property of its track. In particular, the tangent line
to the track need not be well-dened for a smooth path: this can only happen
when p

(t) vanishes at the corresponding point. Also, the track of a smooth path
can be self-intersecting. The following examples illustrate these points.
Example 1.3. For any k R dene a smooth path p
k
(t) as follows
2
:
p
k
(t) = (t
3
+kt, t
2
+k, 0), t R. (1.2)
Since x(t) = ty(t) the track of p
k
(t) can be described as
(x, y, 0) : x
2
= y
2
(y k).
The behaviour within this family demonstrates that our denition allows cusp
points, where the slope of the tangent line to the track is discontinuous, and
nodal points, where the track is self-intersecting. This behaviour is governed by
the sign of k.
Each p
k
(t) is clearly a smooth planar path (lying in the plane z = 0), with
tangent vectors
p

k
(t) = (3t
2
+k, 2t, 0).
1
In general the path p

(t) is called the hodograph of p(t).


2
These are historically known, perhaps confusingly for our modern terminology, as Newtons
diverging parabolas. Newton classied cubic equations of the form y
2
= ax
3
+ bx
2
+ cx + d
by distinguishing them into four classes, allowing linear transformations of x: the diverging
parabolas form the third class.
DIFFERENTIAL GEOMETRY 5
Since this is a planar curve we can assigne to each point p
k
(t) the slope of the
tangent line, which is 2t/(3t
2
+k).
When k > 0 then p
k
(t) parameterises an embedded cubic. Its path is one-to-
one (injective) and the slope of its tangent line at p
k
(t) is a smooth function.
When k = 0 the path is still one-to-one but the slope of its tangent line is 2/3t,
which has a discontinuity at p
0
(0) = (0, 0): this gives the curve a cusp, and it is
known as a cuspidal cubic. When k < 0 there is a point of self-intersection, since
p
k
(

k) = p(

k) = (0, 0). At this point the tangent line is not well-dened,


since the two tangent vectors (2k, 2

k, 0) are not parallel. This family of


curves are known as nodal cubics: each has a node or ordinary double point at the
origin. But the tangent vector of the path is well-dened, since these tangents
occur at dierent times t =

k.
x
y
x
y
x
y
Figure 4. From left to right, embedded, cuspidal and nodal cubics.
A more extreme example of a path whose track has no tangent line at one point,
but which can be smoothly parameterised, is given by the following example.
Example 1.4. Let p(t) = (t, [t[, 0). This is of course a parametrisation of the
right-angle curve. The path p(t) is not smooth at t = 0, but it is possible to
x
y
Figure 5. Right-angle curve.
reparameterise the track to obtain a smooth path. We take
q(t) =
_
(e
1/t
2
, e
1/t
2
, 0), if t 0,
(e
1/t
2
, e
1/t
2
, 0), if t 0.
This has q
(n)
(0) = 0 for all n = 1, 2, 3, . . . , which is the only way to smoothly
negotiate the right-angle bend.
6 IAN MCINTOSH
These examples make clear the need to distinguish those smooth paths whose
tangent vectors do not vanish at any point.
Denition 1.2. Let p : I R
3
be a smooth path. We say a parameter value
t
0
I is:
(i) singular or critical if p

(t
0
) = 0, in which case p = p(t
0
) is a singular or
critical point, and;
(ii) regular if it is not singular, in which case p is a regular point.
A smooth path is regular if all its points are regular.
In particular, by this denition whenever p(t) has a point p of self-intersection,
this is only a regular point if each of the parameter values t
0
for which p(t
0
) = p
is regular. It is still possible for regular curves to have multiple tangent lines at
one point, but regularity only allows this to happen at point of self-intersection.
For example, it can be shown that the cuspidal cubic in Example 1.3 does not
admit any regular parametrisation about its cusp, but the nodal cubic is a regular
path.
1.2. Smooth Curves and Arc Length Parametrisation. Any smooth path
p(t) has many dierent reparametrisations, dened as follows.
Denition 1.3. If p : I R
3
and q : J R
3
are smooth paths for which
q(u) = p(t(u)), u J, for some smooth change of variable function t(u) : J
I with t

(u) ,= 0 then we say that q(u) is a reparametrisation of p(t). When


t

(u) > 0 (resp. t

(u) < 0) we say u is an orientation preserving (resp. orientation


reversing) reparametrisation.
Remark 1.2. We want the reparametrisation function u(t) to be a smooth bijec-
tion u : I J, with smooth inverse. This will follow if its derivative is not allowed
to vanish, for then it is either a strictly increasing function (when u

(t) > 0) or
strictly decreasing function (when u

(t) < 0). In particular, a reparametrisation


cannot double back on itself.
Both smooth paths p, q will have the same track. By the chain rule
dq
du
=
dt
du
dp
dt
,
so that the condition t

(u) ,= 0 is necessary and sucient to allow the parametri-


sation to be inverted, i.e., we can write p(t) = q(u(t)). We can see from this
that an orientation preserving reparametrisation moves us in the same direction
along the curve, whereas an orientation reversing reparametrisation reverses the
direction of travel. Notice that p(t) will be a regular path if and only if q(u) is a
regular path, since [q

(u)[ = [t

(u)[[p

(t)[.
Our denition for a smooth curve intends to get at the geometric object which
remains unchanged by reparametrisation: this is slightly more subtle than than
the idea of the track.
DIFFERENTIAL GEOMETRY 7
Denition 1.4. A smooth curve (resp. oriented smooth curve) C is an equiv-
alence class of smooth paths any two of which are equal after reparametrisation
(resp. orientation preserving reparametrisation). Paths in the equivalence class
are said to parameterise C. A smooth curve is regular if one, and hence all, its
parametrisations are regular. If C is an oriented curve with parametrisation p(t)
then its opposite is the curve C with parametrisation p(t).
Remark 1.3. What is the dierence between a curve and its track? Clearly every
curve determines a single track, but there are many examples of tracks which
come from two dierent curves. One of the simplest examples is as follows. Let
p : [0, 2) R
3
, p(t) = (cos(t), sin(t), 0),
q : [0, 2) R
3
, q(t) = (cos(2t), sin(2t), 0).
Both curves have the unit circle in the plane z = 0 as their track. The curve p
winds once around this circle, whereas the curve q winds twice around. There
is no reparametrisation u : [0, 2) [0, 2) for which q(t) = p(u(t)), because u
is invertible (by the previous remark) and p is invertible, but q is not injective.
[You might think that u(t) = 2t should work, but it doesnt map [0, 2) to itself.]
There are a particular collection of (re)parametrisations which are geometri-
cally meaningful, corresponding to the arc length along the curve from a chosen
initial point and in a chosen direction along the path. Recall from Vector Calculus
the denition of arc length.
Denition 1.5. Let p : I R
3
be a smooth path. The (oriented) arc length
from p(a) to p(b) is dened to be
s(a, b) =
_
b
a
[p

(t)[ dt, a, b I. (1.3)


Note that s(a, b) < 0 if b < a. If we x an initial point p(t
0
) and measure the
oriented arc length to any other point p(t), then s(t) = s(t
0
, t) is a dierentiable
function with derivative [p

(t)[, by the Fundamental Theorem of Calculus.


Denition 1.6. We say p(t) is arc length parameterised (and t is an arc length
parametrisation) whenever s(t, t
0
) = t t
0
, or equivalently, when [p

(t)[ = 1. For
the latter reason, this is also referred to as a unit speed parametrisation, and
T(t) = p

(t) is called the unit tangent vector eld along p(t).


It seems geometrically obvious that the track of any smooth curve should admit
an arc length parametrisation. However, we can only ensure that this is smooth
when the arc length function is smooth, and therefore we must avoid paths which
have [p

(t)[ = 0 at some point (because f(x) = [x[ is smooth everywhere except


at x = 0, where it is not even dierentiable).
8 IAN MCINTOSH
Lemma 1.7. Let p(t) be a regular path. Then p can be reparameterised by arc
length s(t) to obtain a regular path p(s) for which
d p
ds
=
p

(t)
[p

(t)[
.
Proof. The expression above is a consequence of the chain rule mentioned above,
since
dt
ds
= 1
_
ds
dt
=
1
[p

(t)[
,
and we have just seen that this is smooth whenever [p

(t)[ does not vanish, i.e.,


at all regular points.
Example 1.5. Consider the helix p(t) = (a cos t, a sin t, bt) with a, b R. Then
p

(t) = (a sin t, a cos t, b), [p

(t)[
2
= a
2
+b
2
.
The arc length from p(0) to p(t) is therefore
s(0, t) =
_
t
0

a
2
+b
2
du = ct, where c =

a
2
+b
2
.
It follows that t(s) = s/c and an arc length reparametrisation is therefore
p(s) =
_
a cos(s/c), a sin(s/c), bs/c
_
.
We will reserve the symbol s to denote arc length parametrisations. Such a
parametrisation is not unique for a given regular curve C, but can only be changed
by translation of s or a change of orientation.
Theorem 1.8. Suppose p(s) and p() are both arc length parametrisations of the
same curve C. Then
= s +c, from some c R.
Proof. Since we require [ p

((s))[ = [p

(s)[ the chain rule forces [

(s)[ = 1. Thus

(s) = 1, whence (s) = 1 +c for some constant of integration c.


It follows that a smooth regular oriented curve has precisely one unit speed
tangent vector eld along it.
1.3. Curvature and torsion. Intuitively, the curvature of a space curve C is
its rate of change of direction. Since rate of change is measured by the derivative,
and direction is measure by the unit tangent vector, we measure curvature by
taking the derivative of the unit tangent vector, i.e., the second derivative of the
unit speed parameterised path.
Denition 1.9. Let C be a regular smooth curve, and p(s) an arc length parametri-
sation with unit speed tangent vector eld T(s) = p

(s). We call
k(s) = T

(s) = p

(s)
DIFFERENTIAL GEOMETRY 9
the curvature vector of p(s) and
(s) = [k(s)[ = [p

(s)[ 0
the curvature.
Note that
k

T = p

=
1
2
(p

= 0.
Since the length of T is constant, as s varies only the direction of T changes.
T(s)
k(s)
Figure 6. The curvature and unit tangent vectors.
Thus the magnitude of k = T

does indeed measure the rate of change of the


curves direction.
Example 1.6. Let a, b R
3
, with b of unit length. Then the path p(s) = a +sb
is an arc length parametrisation of the straight line through a in direction b.
Clearly all straight lines may be parameterised in this way. Then:
p

(s) = b, p

(s) = 0,
so (s) = 0 for all s, as we would expect. Conversely, if the curvature is identically
zero then p

(s) = 0 for all s, and two integrations yield p(s) = a + sb. Hence a
curve has zero curvature if and only if it is a straight line.
Provided (s) ,= 0 the two vectors T(s), k(s) span a plane, called the osculating
plane at the point p(s) on the curve. The torsion measures the extent to which
the curve twists out of this plane. To dene torsion we proceed as follows.
Denition 1.10. Let p(s) be an arc length parameterised curve. We say a point
p(s
0
) is a point of inection when (s
0
) = 0. When p(s) has no points of infection
we dene:
the principal normal vector
N(s) =
k(s)
[k(s)[
=
k(s)
(s)
=
T

(s)
(s)
=
p

(s)
[p

(s)[
,
the binormal vector
B(s) = T(s) N(s). (1.4)
At each point p(s) on the curve the vectors T(s), N(s), B(s) form a postively
oriented orthonormal basis for R
3
, called the Frenet frame.
10 IAN MCINTOSH
T(s)
N(s)
B(s)
Figure 7. The Frenet frame.
Dierentiating B

B = 1 and (1.4) gives two identites:


B

B = 0, B

= (T

N) +T N

) = T N

, (1.5)
where we have used that T

and N are parallel. Hence B

is orthogonal to T and
B, which means it is parallel to N.
Denition 1.11. The torsion (s) is the function for which B

(s) = (s)N(s).
Equally, = B

N.
Recall that (s) = [p

(s)[. Combining the denition of with (1.5), we have


= B

N = (T N

N = [T, N

, N], the triple scalar product,


= [T, N, N

],
using the antisymmetric property of the triple scalar product. Now
T = p

, N =
1

=
1

, N

=
1

2
(p

).
Therefore in arc length parametrisation
(s) =
1
(s)
2
[p

(s), p

(s), p

(s)] =
[p

(s), p

(s), p

(s)]
[p

(s)[
2
. (1.6)
Example 1.7 (Curvature and torsion of helices). From our previous work with
helices, a unit speed parametrisation is
p(s) = (a cos(s/c), a sin(s/c), bs/c), where c =

a
2
+b
2
Hence
p

(s) = (
a
c
sin(s/c),
a
c
cos(s/c),
b
c
),
p

(s) =
a
c
2
(cos(s/c), sin(a/c), 0),
p

(s) =
a
c
3
(sin(s/c), cos(a/c), 0),
and
[p

, p

, p

](s) =
ba
2
c
6
(cos
2
(s/c) + sin
2
(s/c)) =
ba
2
c
6
.
DIFFERENTIAL GEOMETRY 11
Hence
k(s) = p

(s) =
_

a
c
2
cos
_
s
c
_
,
a
c
2
sin
_
s
c
_
, 0
_
,
(s) =
[a[
c
2
=
[a[
a
2
+b
2
,
(s) =
c
4
a
2
a
2
b
c
6
=
b
c
2
.
So a helix has constant curvature and constant torsion (in fact we will see later
that a curve with such properties must be a helix). Notice that the torsion is
negative since this helix is right-handed. Taking b = 0 gives a circle of radius
a > 0, with (s) = 1/a and torsion = 0. Thus the curvature of a circle is the
reciprocal of its radius.
1.3.1. Planar curves. We say a curve p(s) is planar if it lies in a xed plane in
R
3
. Thus p(s) is planar when there is a point p R
3
a non-zero vector n R
3
for which
(p(s) p)

n = 0,
i.e., p(s) lies in the plane through p with normal n. By dierentiating this
equation twice we see that for a planar curve without points of inection
p

(s)

n = 0 = p

(s)

n, i.e., T

n = 0 = N

n.
Thus p(s) is planar if and only if the osculating plane spanned by T(s) and
N(s) is constant. This is equivalent to saying that the binormal B is constant.
Conversely, suppose the binormal B is constant, then for any choice of point
p = p(s
0
) on the curve
((p(s) p)

B)

= p

(s)

B = T

B = 0,
thus (p(s) p)

B) is a constant function of s. But it vanishes at s = s


0
, so
(p(s) p)

B = 0, i.e., p(s) is a planar curve. We have therefore proved:


Theorem 1.12. A curve without inections is planar if and only if it has constant
binormal vector, that is, if and only if its torsion is identically zero all along
the curve.
A plane R
3
is oriented by making a choice of unit normal vector . For a
smooth curve in the oriented plane (, ) we can talk about the signed curvature.
Given T(s) = p

(s) and , the vector = T(s) is a unit vector normal to


both T(s) and , chosen so that (T, , ) is positively oriented. It follows that
N = . The signed curvature is dened to be

s
= k

. (1.7)
Thus
s
= , with the sign determined by N

. Notice that (T, N, B) =
(T, , ) precisely when N = .
12 IAN MCINTOSH
1.3.2. Curvature and torsion in any regular parametrisation. Since it is not usu-
ally practical to nd the arc length parametrisation of a curve, we now derive the
expressions for curvature and torsion in any regular parametrisation.
Theorem 1.13. If p(t) is a regular parametrisation of a space curve then the
curvature vector and curvature are given by:
k =
[p

[
2
p

(p

)p

[p

[
4
=
[p

[
[p

[
3
(1.8)
Further, when p(t) has no inection points it has torsion
=
1

2
[p

, p

, p

]
[p

[
6
=
[p

, p

, p

]
[p

[
2
(1.9)
Proof. Dene p(s) = p(t(s)). Then
d p
ds
=
p

(t(s))
[p

(t(s))[
Note rst that:
d
dt
[p

(t)[ =
d
dt
_
p

(t)

(t)
_
1/2
=
p

(t)

(t)
[p

(t)[
Recalling that ds/dt = [p

[, and hence dt/ds = 1/[p

[, we obtain:
k =
d
2
p
ds
2
=
_
[p

[p

dt
ds

d
dt
[p

(t)[
dt
ds
p

__
[p

[
2
=
_
p

[p

[
2
p

__
[p

[
2
=
[p

[
2
p

(p

)p

[p

[
4
Then

2
= [k[
2
=
_
[p

[
2
p

(p

)p

[p

[
4
_

_
[p

[
2
p

(p

)p

[p

[
4
_
=
[p

[
4
[p

[
2
2[p

[
2
(p

)
2
+ (p

)
2
[p

[
2
[p

[
8
=
[p

[
2
[p

[
2
(p

)
2
[p

[
6
=
[p

[
2
[p

[
6
,
by Lagranges formula (1.10) below. Now from (1.6)
(t(s)) =
1
(s)
2
_
p

(s), p

(s), p

(s)

=
1

2
_
p

[p

[
,
p

[p

[
2
, p

_
,
DIFFERENTIAL GEOMETRY 13
where we have used the expression for k = p

above. Now p

is of the form
ap

+ bp

+ cp

, and since the triple scalar product is antisymmetric only the


term cp

is required:
cp

=
1
[p

[
2
d
ds
p

=
1
[p

[
2
_
p

dt
ds
_
=
p

[p

[
3
.

Remark 1.4. For any a, b R


3
we have:
[a b[
2
= [a[
2
[b[
2
sin
2
= [a[
2
[b[
2
(1 cos
2
) = [a[
2
[b[
2
(a

b)
2
. (1.10)
This is sometimes called Lagranges formula, and can be used to evaluate
without having to compute the vector product.
1.4. Congruence and Frenet formulas. Our aim now is to show that the
curvature and torsion of a curve determine it completely up to rotations and
translations. These transformations generate the group of orientation preserving
Euclidean motions. We start with a short discussion of Euclidean motions.
Denition 1.14. A transformation T : R
3
R
3
is called a Euclidean motion
(or isometry) when it preserves distances and angles, i.e.,
(Tp
2
Tp
1
)

(Tq
2
Tq
1
) = (p
2
p
1
)

(q
2
q
1
), p
1
, p
2
, q
1
, q
2
R
3
. (1.11)
It can be shown that all Euclidean motions must have the form Tv = Lv +c,
where L : R
3
R
3
is an invertible linear map, and c is a constant. It follows that
the set of Euclidean motions is a group (under composition of transformations).
The condition (1.11) amounts to the condition
Lv

Lw = v

w, v, w R
3
.
This implies that L must represented by an orthogonal matrix, i.e, LL
t
= I
3
, the
3 3 identity matrix. It follows from this that det(L)
2
= 1, since det(L
t
) =
det(L), and therefore det(L) = 1. L is a rotation about the origin when
det(L) = 1, and is a reection (through some plane containing the origin) when
det(L) = 1. Thus every Euclidean motion is the composition of a rotation or
reection with a translation.
Denition 1.15. We say the Euclidean motion Tv = Lv + c is orientation
preserving when det(L) = 1, equally, when
[Le
1
, Le
2
, Le
3
] = [e
1
, e
2
, e
3
]
where e
1
, e
2
, e
3
is the standard orthonormal basis for R
3
. Otherwise it is called
orientation reversing.
Denition 1.16. We will say two space curves p, q : [a, b] R
3
are congruent
when there is a Euclidean motion T for which (T p)(t) = q(t) for all t [a, b],
and we will say they are properly congruent when T is also orientation preserving.
14 IAN MCINTOSH
Remark 1.5. The set of all 3 3 orthogonal matrices is denoted by O(3). This
forms a group under matrix multiplication, which is called the orthogonal group.
The subset of those matrices which also have determinant 1 is denoted by SO(3).
It is a subgroup, called the special orthogonal (sub)group. Thus SO(3) represents
the group of all rotations of R
3
. Notice that the set of reections does not form
a subgroup of O(3) (since the identity transformation is not a refelection), but
that the product of two reections is a rotation.
The Frenet (sometimes called Frenet-Serret) formulas tell us how to rebuild a
curve from its curvature and torsion, up to rotations and translations in space.
They show how the Frenet frame can be obtained from a system of linear dier-
ential equations whose data is the curvature and torsion of the curve. These are
derived as follows.
Let p(s) be a unit speed path. By denition of the principal normal and
binormal
T

= N, B

= N.
Since [N[ = 1 we have N

N = 0, so
N

= T +B,
for some scalar functions , . Since N

T = 0 = N

B it follows that
= N

T = N

T

= , = N

B = N

B

= ,
and we thereby obtain the Frenet formulas.
T

= N, N

= T B B

= N. (1.12)
It is useful to write these in matrix form
_
T

_
=
_
T N B
_
_
_
0 0
0
0 0
_
_
. (1.13)
We can simplify this notationally into F

= FA where F is the matrix represent-


ing the Frenet frame and A is the skew-symmetric matrix (i.e., A
t
= A) contain-
ing the curvature and torsion. Notice that F is a special orthogonal matrix, since
its columns come from an orthonormal basis with right-handed orientation. We
should think of (1.12), equally (1.13), as a linear system of ordinary dierential
equations for the vector-valued functions T(s), N(t), B(t) given the functions (t)
and (t). Using this perspective we can state and prove the following theorem.
Fundamental Theorem of Space Curves.
(i) Let p, p : I R
3
be unit speed paths with curvatures (s) = (s) ,= 0 and
torsions (s) = (s) for all s I. Then p and p are properly congruent.
(ii) Let , : I R be smooth functions, with > 0. There these exists a
smooth unit speed path p : I R
3
with curvature and torsion , and p
is determined uniquely up to proper congruence.
DIFFERENTIAL GEOMETRY 15
The proof rests on the standard existence and uniqueness theorem for linear
ordinary dierential equations, which we state here without proof.
Theorem 1.17. Let A(t) be a smooth 3 3 matrix-valued function on some
interval I R, and t
0
I. Then for each x
0
R
3
the linear o.d.e. x

= xA for
the vector-valued function
x(t) =
_
x(t) y(t) z(t)
_
has a unique solution x(t) over t I satisfying the initial condition x(t
0
) = x
0
.
We will make use of this in the form of the following corollary. Let M
3
3
(R)
denote the set (indeed, vector space) of all 3 3 matrices with real entries.
Corollary 1.18. Let A : I M
3
3
(R) be a smooth matrix-valued function. For
each invertible matrix F
0
M
3
3
(R) there exists a unique matrix-valued solution
F : I M
3
3
(R) to the o.d.e. F

= FA satisfying the initial condition F(t


0
) = F
0
.
This solution F(t) is invertible for all t I. If

F(t) is the unique solution to the
same equation with initial condition

F(t
0
) =

F
0
then

F = LF where L =

F
0
F
1
0
is constant.
Further, when A
t
= A and F
0
O(3) then F(t) is an orthogonal matrix for
all t I (and F(t) is special orthogonal if F
0
SO(3)).
The relationship between the Theorem and its Corollary is that the rows of F
0
give three linearly independent initial conditions x
1
(t
0
), x
2
(t
0
), x
3
(t
0
) which span
R
3
, and the corresponding solutions x
1
(t), x
2
(t), x
3
(t) must therefore span R
3
at
each t I. Since x

j
= x
j
A for each of these, the matrix F with rows x
1
, x
2
, x
3
satises F

= FA and is invertible for all time. It is called a fundamental matrix


solution, since every solution to the o.d.e. can be obtained from its rows by
linear combination. Now if

F satises the same equation with a dierent initial
condition

F(t
0
) =

F
0
, then both F
1
0
F and

F
1
0

F satisfy the equation with intial


condition I
3
, so

F
1
0

F = F
1
0
F, i.e.,

F =

F
0
F
1
0
F.
To see the last part of the corollary, when A
t
= A it follows that
(FF
t
)

= F

F
t
+F(F

)
t
, since (F
t
)

= (F

)
t
,
= FAF
t
+FA
t
F
t
, since (FA)
t
= A
t
F
t
,
= F(A +A
t
)F
t
= 0.
So FF
t
is constant. Thus F(t
0
)F(t
0
)
t
= I
3
implies FF
t
= I
3
for all time. Hence
F is an orthogonal matrix. Now F(t) is a continuous function of t, therefore so
is det(F(t)). But det(F(t)) = 1, so either det(F) = 1 or det(F) = 1 for all
time. The sign is therefore determined by the initial condition det(F
0
).
Proof of the Fundamental Theorem.
(i) Suppose we have two arc length parameterised space curves p, p : I R
3
with the same curvature and torsion, and . Let T, N, B and

T,

N,

B be their
16 IAN MCINTOSH
respective Frenet frames, with corresponding matrices F and

F. These are both
solutions to the o.d.e. F

= FA where
A =
_
_
0 0
0
0 0
_
_
. (1.14)
By Corollary 1.18 there is a constant matrix L for which

F = LF. Since F,

F are
special orthogonal, so is L. This means in particular that

T = LT. Now T = p

and

T = p

, so
p(t) p(t
0
) =
_
t
t
0

T()d =
_
t
t
0
LT()d = L
_
t
t
0
T()d = L(p(t) p(t
0
)),
since L is constant. Therefore
p = Lp +c, c = p(t
0
) Lp(t
0
).
i.e., the two paths are properly congruent.
(ii) Given smooth functions , : I R, with > 0, we solve F

= FA for the
matrix A in (1.14). Use the columns of F to dene T, N, B, then these satisfy
the Frenet equations, by construction. Now dene
p(t) =
_
t
t
0
T()d.
Then p : I R
3
is a smooth path with p

= T, parameterised by arc length since


[T[ = 1, and with normal N and binormal B, and therefore with curvature and
torsion . By part (i), any other curve with the same curvature and torsion is
properly congruent to it.
Remark 1.6.
(i) Notice in the construction of a path p from , , there are constants of
integration chosen at two steps: rst, the solution to F

= FA requires
an initial condition F(t
0
) SO(3), and second, the integral to obtain p
sets a value for p(t
0
). These choices will, respectively, rotate and translate
the path.
(ii) Let us consider the eect of a reection L O(3), det(L) = 1, on a
path p. Let p = Lp. Then p has Frenet frame

T = LT,

N = LN, but

B = LB, since
[LT, LN, LB] = det(L)[T, N, B] = det(L) = 1.
Using the Frenet formulas this means = but = . By the Fun-
damental Theorem, we deduce that two paths with the same curvature
but opposite torsion are still congruent, but not properly congruent.
DIFFERENTIAL GEOMETRY 17
Example 1.8. Let us now show that a curve is a helix if and only if it has constant
curvature > 0 and torsion . Example 1.7 gave an arc length parameterised
helix with curvature and torsion
(s) =
[a[
a
2
+b
2
, (s) =
b
a
2
+b
2
,
for arbitrary constants a, b R. Conversely, given constants > 0 and we can
nd a, b R satisfying these equations. To be precise,
[a[ =

2
+
2
, b =

2
+
2
.
Thus every choice of constants > 0 and has a helix with those values for cur-
vature and torsion. By the Fundamental Theorem any path with that curvature
and torsion is a helix (congruent to the standard form given in Example 1.7).
2. Smooth Surfaces and their Calculus.
Our aim is to adapt vector calculus to work on smooth surfaces, but before
we can do that we need to dene the notion of a smooth surface itself. From a
vector calculus course we expect that this idea should include two common types
of construction.
Level surfaces. For a dierentiable function f : R
3
R for each k R the set
S
k
= (x, y, z) : f(x, y, z) = k
is usually called the level set at level k. For example, the sphere of radius r > 0
is the level surface
S
2
(r) = (x, y, z) : x
2
+y
2
+z
2
= r
2
.
For simplicity the unit sphere S
2
(1) is usually denoted by S
2
. This denition
also gives a non-empty set when r = 0, but that set is 0, the single point set
containing the origin. It would be absurd to call this a surface, so not every level
set S
k
gives a surface. Notice that the problem with S
2
(0) is that it has the
wrong dimension: a surface ought to be two-dimensional.
A second family of examples turns up another reason for caution. Consider
the level sets
S
k
= (x, y, z) : x
2
+y
2
z
2
= k.
For k < 0 these are the hyperboloids of one sheet, while for k > 0 we get the
hyperboloids of two sheets. In between these cases lies the cone S
0
. This looks
two dimensional at almost every point except the origin, where its tangent plane
fails to be well-dened. Our denition will exclude S
0
from being a smooth surface
for this reason.
Later, the Regular Value Theorem will give us sucient conditions for a level
set to be a surface.
18 IAN MCINTOSH
Parametric Surfaces. Another common construction of surfaces is to describe
them using the two dimensional version of the parametrisation of a curve, in the
form
S = p(u, v) = (x(u, v), y(u, v), z(u, v)) : (u, v) U R
2

For example, the map p : R


2
R
3
given by
p(u, v) = (cos ucos v, sin ucos v, sin v),
has image S
2
, i.e., this is a parametrisation of S
2
. It is clearly not injective, since
it is made from periodic functions. By restricting the domain it can be made
injective, but to make it simultaneously surjective we have to carefully choose
the domain. For example, the domain
D = (0, v) : /2 v /2 (u, v) : 0 < u < 2, /2 < v < /2,
will make p : D S
2
bijective.
A good parametrisation ought not the have redundant variables. For example,
the map p(u, v) = (u, 0, 0) doesnt really depend upon v, and its image is clearly
not a surface. We will see later that what makes for a good parametrisation of a
surface is the independence of the vectors of partial derivatives p
u
= p/u and
p
v
= p/v at each point. This also indicates the two dimensional nature of a
proper surface: these two vectors should span the tangent plane.
2.1. Fundamental Concepts.
Denition 2.1. If a R
n
, r > 0, then:
B
r
(a) = x R
n
: [x a[ < r,
is the open ball of radius r and centre a. Of course:
[x a[ =
_
(x
1
a
1
)
2
+ + (x
n
a
n
)
2
.
A subset U R
n
is open if for each a U there exists r > 0 such that B
r
(a) U.
A subset E R
n
is closed if its complement E

is open.
An open set containing a R
n
is called a neighbourhood of a.
Remark 2.1. These ideas come from topology (the so-called standard topology of
R
n
), but we do not need to study topology to understand them in our context.
An important observation is that when we use standard inclusions like R R
2
,
x (x, y), open sets are never mapped to open sets, since every open interval
(a, b) R is mapped to the subset
(x, y) : a < x < b, y = 0 R
2
,
which is neither open nor closed
3
. Thus the notion of open is dimension de-
pendent, as the denition indicates.
3
Use the denition to convince yourself that this set is not open and its complement is also
not open.
DIFFERENTIAL GEOMETRY 19
A function f : U R
m
on an open set U R
n
will be called smooth at a U
if for every k N all k-th order partial derivatives of f exist in a neighbourhood
of a and are continuous at a, and f is smooth on U if f is smooth at all a U.
Denition 2.2. Suppose U
1
R
n
and U
2
R
m
are open subsets. A smooth
function f : U
1
U
2
is said to be a dieomorphism if f is invertible and the
inverse function f
1
: U
2
U
1
is smooth. In this case U
1
and U
2
are said to be
dieomorphic open sets.
One of our main tasks will be to adapt the idea of a smooth map to non-open
subsets of R
3
(because surfaces, being two dimensional, are never open subsets
of R
3
). We achieve this using the following idea.
Denition 2.3. Suppose f : D R
m
with D R
n
and a D. If V is a
neighbourhood of a and

f : V R
m
satises:

f(x) = f(x), for all x V D,


then

f is called a local extension of f at a.
The idea now is that a function on a not-necessarily-open subset D R
n
will
be smooth if it admits about each point a local extension which is smooth. From
this we can extend the notion of dieomorphism.
Denition 2.4. Function f : D R
m
is smooth at a D if there exists a local
extension of f at a which is smooth at a. If f is smooth at all a D then f is
smooth on D.
Let D
1
R
n
and D
2
R
m
be subsets which are not necessarily open. A
function f : D
1
D
2
is smooth if it is smooth as a function into R
m
. Given
this, a smooth function f : D
1
D
2
is a dieomorphism if f is invertible and
f
1
: D
2
D
1
is smooth. Then D
1
is said to be dieomorphic to D
2
, written
D
1

= D
2
.
It is easy to see that dieomorphism is an equivalence relation.
The notion of dieomorphism is an important one in surface theory, so its
worth spending some time understanding. We start with some one dimensional
examples.
Example 2.1. Let D
1
= [0, 1) R, D
2
= (x, y) R
2
: y =
1
1x
, x [0, 1) and
S
1
be the unit circle (x, y) R
2
: x
2
+ y
2
= 1. We claim that D
1

= D
2
, but
D
1
,

= S
1
(and therefore D
2
,

= S
1
).
Let f : D
1
D
2
, f(x) = (x, 1/(1 x)). Clearly f is smooth on (0, 1), and
a bijection. To show that it is smooth at 0 we need a local smooth extension.
We can take

f : (, 0) R
2
using the same rule

f(x) = (x, 1/(1 x)). The
inverse of f is g : D
2
D
1
, g(x, 1/(1 x)) = x. To show this is smooth we need
to show there is a smooth extension. We can take g : (, 1) R R with
g(x, y) = x. Its domain is open and it is clearly smooth. Hence f is smooth with
smooth inverse.
20 IAN MCINTOSH
Now consider the claim that [0, 1) ,

= S
1
. The problem is not the lack of a
smooth function with an inverse, for the map
f : [0, 1) S
1
; f(t) = (cos(2t), sin(2t)),
is easily seen to be a smooth bijection (it has local smooth extension given by the
same rule extended to any open interval containing [0, 1)). The problem is that no
such function can have a continuous (let alone smooth) inverse. The (informal)
argument is straightfoward. A continuous injection g : S
1
R must be either
strictly increasing or strictly decreasing as we go around the circle clockwise.
Therefore it cannot return to the value at which it started after one circuit: the
function must break (be discontinuous) at some point.
Example 2.2. It is very easy to adapt the previous examples to two dimensions:
let D
1
= [0, 1) R,
D
2
= (x, y, z) R
3
: z = 1/(1 x), D
3
= (x, y, z) : x
2
+y
2
= 1.
Notice that D
3
is a cylinder. By adapting the functions and the arguments above
one can show that D
1
D
2
but D
1
,

= D
3
.
To proceed further with dierential geometry, and develop our understanding
of when two sets are dieomorphic, we have to recall the notion of the total
derivative from vector calculus.
Denition 2.5. For a smooth map f : U R
m
dened on an open set U R
n
its dierential or total derivative at a U is the linear map
df(a) : R
n
R
m
; df(a)[h] =
d
dt
f(a +th)

t=0
.
The matrix which represents this linear map (using the canonical bases of R
n
and
R
m
) is the Jacobian matrix J
f
(a), the m n matrix whose ij-th entry is the
partial derivative:
D
j
f
i
(a) =
f
i
x
j
(a).
Remark 2.2. Recall that the vector df(a)[h] gives the directional derivative of f
in the direction h. When f : R
n
R is a scalar function then the 1 3 matrix
J
f
is also denoted by f, the gradient vector of f:
f(p) =
_
f
x

p
,
f
y

p
,
f
z

p
_
= (f
x
(p), f
y
(p), f
z
(p)).
Then, if h = (h
1
, h
2
, h
3
) we have:
df(p)[h] = f
x
(p)h
1
+f
y
(p)h
2
+f
z
(p)h
3
= f(p)

h.
We recall also the vector calculus version of the chain rule.
DIFFERENTIAL GEOMETRY 21
Denition 2.6. Suppose we have a chain of maps:
R
n
f
R
m
g
R
p
,
satisfying the following conditions:
f is smooth at a R
n
(thus df(a): R
n
R
m
exists).
g is smooth at b = f(a) R
m
(thus dg(b): R
m
R
p
exists).
Then g f : R
n
R
p
is smooth at a, and:
d(g f)(a) = dg(b) df(a)
From the chain rule we can immediately derive a necessary condition for a
smooth map f : U V , between open subset U R
n
and V R
m
, to be a
dieomorphism. For this necessarily requires
f
1
f = id
U
, f f
1
= id
V
,
the identity maps on U and V respectively. Since f and f
1
are both smooth
the chain rule implies that
df
1
(b) df(a) = id
R
n, df(a) df
1
(b) = id
R
m .
This means the linear map df(a) has both a left and right inverse. Thus n = m
and df(a) must be invertible; equally, a linear isomorphism (in particular, dieo-
morphic sets must have the same dimension). It turns out that this condition
is also sucient locally. That is the content of the Inverse Function Theorem,
whose proof belongs in a course on calculus of several variables (see, for example,
[Bartle]). So we will simply state it and make use of it without further comment.
Inverse Function Theorem. Let U R
n
be open, and let f : U R
n
be
smooth at a U. Suppose that df(a): R
n
R
n
is an isomorphism. Then there
exist neighbourhoods A of a and B of b = f(a) such that f : A B is a smooth
dieomorphism. Furthermore:
df
1
(b) = (df(a))
1
2.2. Charts, Atlases and Surfaces. Our denition of a surface will essentially
say that about each point the surface has a patch which is dieomorphic to an
open subset of R
2
, together with a requirement that this patch sits well in R
3
.
The next two denitions make this precise.
Denition 2.7. A (smooth) surface patch is a subset D R
3
together with a
dieomorphism : D U to an open subset U R
2
.
Denition 2.8 (Smooth surface). Let S R
3
. A chart on S is a surface patch
(D, ) where D = S V for some open subset V R
3
. The chart map : D
U R
2
is also referred to as local coordinates on D, and its inverse
p =
1
: U D S,
is called a local parametrisation.
22 IAN MCINTOSH
If each point of S lies in a chart then S is called a smooth surface. A collection
of charts whose union is S is called an atlas for S.
As the example of the cylinder (Example 2.7) above shows, it would be too
restrictive to insist that the whole of a surface can be covered by one chart, since
many surfaces are not dieomorphic to an open subset of R
2
.
Example 2.3 (Graphs). Let f : U R be a smooth function, where U R
2
is
open, and let S
f
R
3
to be its graph:
S
f
= (u, v, f(u, v)) : (u, v) U.
Figure 8 gives a visualisation of this. We claim that S
f
is a surface patch, and
U
S
f
u
v
Figure 8. Graph of a function of two variables.
therefore a surface. For, if we dene (u, v, f(u, v)) = (u, v) then : S
f
R
2
is smooth: it has smooth extension : R
3
R
2
given by the projection onto
the plane (x, y, z) = (x, y). Its inverse is p : R
2
S
f
, p(u, v) = (u, v, f(u, v)),
which is also smooth since each component is a smooth function of u, v. Thus
(S
f
, ) is a global chart for S
f
.
Example 2.4 (Sphere: hemispherical charts). We can use graphs to put an atlas
on the unit sphere S
2
. Take D = (x, y, z) S
2
: z > 0. Then D = S
f
, where
Figure 9. Northern hemisphere of the unit sphere.
f(u, v) =
_
1 (u
2
+v
2
), dened on the open disc U = (u, v) R
2
: u
2
+ v
2
<
1. So D is a surface patch by the previous example. Furthermore D = S
2
V
where V = (x, y, z) : z > 0 is the open upper half space; so D is a chart of S.
We can use 6 hemispherical charts like this (or 4, if we want to be economical)
to make an atlas for S
2
. Thus S
2
is a smooth surface.
DIFFERENTIAL GEOMETRY 23
Example 2.5 (Sphere: spherical polar charts). Every point on S
2
has equatorial
spherical polar coordinates, which are obtained from the inverse of the parametri-
sation seen earlier:
p(u, v) = (cos ucos v, sin ucos v, sin v).
If we restrict this to the open subset
U = (u, v) : 0 < u < 2, /2 < v < /2 = (0, 2) (/2, /2) R
2
,
we obtain a smooth bijection p : U S
2
C, where C S
2
C = (x, 0, z) S
2
: x 0,
is the Greenwich meridian. See Figure 10. Let D = S
2
C. The inverse of p
is the map : D U which gives coordinates (u, v) to p(u, v). These can be
interpreted geometrically as the longitude and latitude relative to C. To show
x
z
u
v
C
y
Figure 10. Longitude and latitude on the sphere.
that (D, ) is a chart it suces to show that is smooth (since we know it has
smooth inverse p). Its smooth local extension is given by
: V R
3
; (x, y, z)
_
(x, y), arcsin(z)
_
,
where V = (x, y, z) : [z[ < 1 (x, 0, z) : x 0 and (x, y) is the polar angle of
the point (x, y) R
2
. Therefore D is a surface patch. Finally, D = S
2
V , so D
is a chart. It may be combined with one other similar chart to obtain an atlas.
Example 2.6 (Sphere: stereographic charts). Let D be the subset of S
2
obtained
by deleting the North Pole n = (0, 0, 1):
D = (x, y, z) S
2
: z ,= 1,
and for each p D let (p) denote the unique point in the xy-plane lying
on the straight line through n and p (Figure 11). The map : D R
2
is
called stereographic projection from n. If p = (x, y, z) then an easy exercise in
elementary geometry shows that
(p) =
_
x
1 z
,
y
1 z
_
.
24 IAN MCINTOSH
n
p
(p)
Figure 11. Stereographic projection from the north pole.
This extends to a smooth map of the open subset V = (x, y, z) : z ,= 1 R
3
,
using the same formula. Hence is smooth. The inverse map is
p(u, v) =
1
u
2
+v
2
+ 1
(2u, 2v, u
2
+v
2
1).
This map is clearly smooth. Since D = S
2
V our surface patch is a chart. In
order to obtain an atlas we simply combine this chart with the chart obtained by
stereographic projection from the south pole.
2.3. Tangent Spaces.
Denition 2.9. For a smooth surface S, p S, the tangent space to S at p is
the set
T
p
S = c

(0) R
3
: c(t) is a smooth curve on S with c(0) = p.
We will call the elements of T
p
S tangent vectors, although this denition puts
them at the origin 0 R
3
. We will use the phrase tangent plane to refer to the
parallel plane parallel to T
p
S through p, i.e., p +T
p
S.
Theorem 2.10. T
p
S R
3
is a two dimensional vector space.
Proof. Choose a chart (D, ) about p, with local parametrisation p : U D
about a = (p) U. We claim dp(a) : R
2
R
3
is injective with image T
p
S.
That it is injective follows at once from the existence of a local smooth extension
of about p, since p = id
U
, the identity map on U, therefore by the chain
rule
d (p) dp(a) = id
R
2,
so that dp(a) must have trivial kernel. To see that the image is T
p
S we argue as
follows.
(i) First, imdp(a) T
p
S. For if h R
2
then (t) = a + th lies in U for t
small, so p lies on S. Thus
dp(a)[h] = (p )

(0) T
p
S.
DIFFERENTIAL GEOMETRY 25
(ii) Next, T
p
S imdp(a), for if X T
p
S then X = c

(0) for some smooth


curve c(t) in S with c(0) = p. Now ( c)(t) is a smooth curve in R
2
.
Dene h = ( c)

(0), and observe that


X = (p c)

(0) = dp(a)[( c)

(0)] = dp(a)[h] imdp(a).

Remark 2.3. When S has local parametrisation p : U S, for which p(a) = p


for some a U, we can write p(u, v) = (x(u, v), y(u, v), z(u, v)). The coordinate
lines through a in U are given by (t) = a+te
1
and (t) = a+te
2
, for e
1
= (1, 0),
e
2
= (0, 1). Since

(0) = e
1
and

(0) = e
2
we have
p
u
(a) = dp(a)[e
1
] = (p )

(0),
p
v
(a) = dp(a)[e
2
] = (p )

(0),
Since dp(a) is injective these span T
p
S. Thus
T
p
S = Spp
u
(a), p
v
(a).
2.4. The Regular Value Theorem. Recall that for any smooth function f :
R
3
R its level sets are the subsets of R
3
of the form
S
k
= (x, y, z) : f(x, y, z) = k, k R.
Denition 2.11. Let V R
3
be an open set and f : V R a smooth function.
We say that:
(i) p V is a critical point of f if df(p) = 0, the zero linear map R
3
R,
(ii) p is a regular point of f if p is not a critical point,
(iii) k is a critical value of f if k = f(p) for some critical point p V ,
(iv) k is a regular value of f if k is not a critical value.
Remark 2.4.
(i) Since df(p)[h] = f(p)

h, p is a critical point if and only if f(p) = 0,


and a regular point if and only if f(p) ,= 0.
(ii) If S
k
= then k is a regular value of f. For, if not then S
k
contains a
critical point. Thus, for example, 1 is a regular value of f(x, y, z) =
x
2
+y
2
+z
2
.
(iii) If k is a regular value of f then all points of S
k
are regular points; however
if k is a critical value then at least one point of S
k
is a critical point.
Regular Value Theorem. Suppose V R
3
is open, and f : V R is a smooth
function. If k R is a regular value of f, and the level set S
k
= p V : f(p) =
k is non-empty, then S
k
is a smooth surface. Furthermore T
p
S
k
= ker df(p) for
all p S
k
; equivalently T
p
S
k
= f(p)

.
The proof of this rests largely on the following lemma.
26 IAN MCINTOSH
Lemma 2.12 (Regular Point Lemma). Suppose V R
3
is open and f : V R
is smooth at p V . If p is a regular point of f then there exist:
a neighbourhood A V of p,
an open subset B R
3
,
a smooth dieomorphism : B A,
such that f((u, v, w)) = w for all (u, v, w) B.
Proof. For convenience dene:

1
(x, y, z) = x,
2
(x, y, z) = y,
3
(x, y, z) = z,
and maps F
i
: R
3
R
3
, i = 1, 2, 3:
F
1
(x, y, z) =
_
f(x, y, z), y, z
_
,
F
2
(x, y, z) =
_
x, f(x, y, z), z
_
,
F
3
(x, y, z) =
_
x, y, f(x, y, z)
_
.
Thus
i
F
i
= f. We claim that at least one F
i
is invertible on a neighbourhood
of p. First note that:
J
f
(p) =
_
f
x
(p) f
y
(p) f
z
(p)
_
,
a 1 3 matrix. So, since p is a regular point, at least one of these partial
derivatives is non-zero; say f
x
(p). Now:
J
F
1
(p) =
_
_
f
x
(p) f
y
(p) f
z
(p)
0 1 0
0 0 1
_
_
hence:
det J
F
1
(p) = f
x
(p) ,= 0.
So dF
1
(p) is a linear isomorphism, and by the Inverse Function Theorem F
1
is
locally smoothly invertible about p, with inverse F
1
1
:

B A for some open
sets A V ,

B R
3
. Now let : R
3
R
3
be the dieomorphism which
permutes coordinates in the order (u, v, w) = (w, u, v), and let B =
1
(

B).
Dene : B A by = F
1
1
, then
(f )(u, v, w) = (
1
F
1
F
1
1
)(u, v, w) =
1
(w, u, v) = w.
It is easy to see how to adapt this argument to the case where only f
y
(p) or f
z
(p)
is non-zero, by replacing F
1
by F
2
or F
3
, and by
2
or
3
= id.
Proof of the Regular Value Theorem. Given f : V R with regular value k,
every p S
k
is regular. By the Regular Point Lemma each p has an open
neighbourhood A V in which f = w in the language of that lemma. Thus
D = S
k
A = (u, v, w) : w = k.
is the domain for a chart : D R
2
, ((u, v, w)) = (u, v). Since every point
has a chart of this type, S
k
is a smooth surface.
DIFFERENTIAL GEOMETRY 27
It remains to show that T
p
S = ker df(p). If X T
p
S then X = c

(0) for some


smooth path c(t) in S with c(0) = p. Now
df(p)[X] = (f c)

(0) = 0,
since (f c)(t) = k for all t. Thus T
p
S ker df(p). Since p is a regular point, the
linear map df(p) : R
3
R has rank 1, and its kernel is therefore 2-dimensional.
But T
p
S is also 2-dimensional. So T
p
S = ker df(p). By Remark 2.4(i) this equals
f(p)

.
2.5. Maps between Surfaces. Let f : S
1
S
2
be a smooth map between two
smooth surfaces. For any p S
1
and X T
p
S
1
we have X = c

(0) for some


smooth curve c(t) on S
1
, and (f c)(t) is a smooth curve on S
2
. Therefore
df(p)[X] = (f c)

(0) T
f(p)
S
2
.
Therefore the following denition makes sense.
Denition 2.13. For a smooth map f : S
1
S
2
between two smooth surfaces
the dierential of f at p is the linear map
df(p) : T
p
S
1
T
f(p)
S
2
; df(p)[X] = (f c)

(0),
where c(t) is any smooth curve on S
1
with c(0) = p and c

(0) = X.
Remark 2.5.
(i) Since every smooth map f : S
1
S
2
is, by denition, smooth as a map
f : S
1
R
3
, it follows immediately that this dierential is a linear map.
(ii) Recall that if p(u, v), p : U S
1
, is a local parametrisation about p
(p(a) = p for some a U R
2
) then T
p
S
1
= Spp
u
(a), p
v
(a). Now
for a smooth map f : S
1
S
2
between two smooth surfaces, and with
(t) = a +te
1
, (t) = a +te
2
on U, we have
df(p)[p
u
(a)] = (f p )

(0) = (f p)
u
(a),
df(p)[p
v
(a)] = (f p )

(0) = (f p)
v
(a).
This gives us a way of calculating df in the local coordinate expression
f(p(u, v)). For if X T
p
S
1
then X = X
u
p
u
(a) +X
v
p
v
(a) for coecients
X
u
, X
v
R, and then by linearity
df(p)[X] = X
u
df(p)[p
u
(a)] +X
v
df(p)[p
v
(a)] = X
u
(f p)
u
(a) +X
v
(f p)
u
(a).
(2.1)
Removing the clutter of notation labelling points, we can simplify this:
df[X] = X
u
(f p)
u
+X
v
(f p)
v
. (2.2)
(iii) When

f is a local smooth extension for f about p and X T
p
S
1
is
tangent to the smooth curve c(t) on S
1
, we have
d

f(p)[X] = (

f c)

(0) = (f c)

(0) = df(p)[X].
28 IAN MCINTOSH
This gives us a way of calculating df given a local smooth extension for
f.
Lemma 2.14 (Chain Rule for Surface Maps). Let f : S
1
S
2
and g : S
2
S
3
be two smooth maps between smooth surfaces, and let p S
1
with q = f(p).
Then g f : S
1
S
3
is smooth with dierential at p given by
d(g f)(p) = dg(q) df(p) : T
p
S
1
T
g(q)
S
3
.
The proof is a trivial application of the usual chain rule for maps in R
3
, given
that f and g both have local smooth extensions into R
3
.
Corollary 2.15. Suppose f : S
1
S
2
is a dieomorphism of surfaces. Then
at every p S
1
the dierential df(p) : T
p
S
1
T
f(p)
S
2
is a linear isomorphism,
with inverse df
1
(f(p)).
This follows immediately from the chain rule applied to f
1
f = id
S
1
. The
more interesting fact is that the Inverse Function Theorem can also be adapted
to maps between surfaces. For the statement of this we need to adapt the notion
of a local dieomorphism.
Denition 2.16. A smooth map f : S
1
S
2
is a local dieomorphism at p S
1
if the dierential df(p) : T
p
S
1
T
f(p)
S
2
is a linear isomorphism.
Theorem 2.17 (Inverse Function Theorem for Surface Maps.). A smooth map
f : S
1
S
2
is a local dieomorphism at p if and only if there exist chart domains
D
1
S
1
containing p and D
2
S
2
containing f(p) such that f : D
1
D
2
is a
dieomorphism.
Proof. This is simply a consequence of the standard Inverse Function Theorem
2.1, applied to a local representative F of f at p. This means the following. Let
(
1
, D
1
) be a chart on S
1
about p, and (
2
, D
2
) be a chart about f(p) on S
2
. We
will denote the corresponding inverses (local parametrisations) by p
j
: U
j
D
j
.
Now dene F : U
1
U
2
to be the composition F =
2
f p
1
;
U
1
p
1
D
1
f
D
2

2
U
2
.
Then if
1
(p) = a U
1
and q = f(p) it follows from the Chain Rule (for
Surfaces) that:
dF(a) = d
2
(q) df(p) dp
1
(a).
Now all three linear maps on the right hand side are isomorphisms, so dF(a) :
R
2
R
2
must also be one. Therefore by the Inverse Function Theorem there
exists a neighbourhood W
1
U
1
of a such that the restriction of F to W
1
is a
dieomorphism. By reducing the size of the chart domains D
1
, D
2
if necessary
(so that D
1
= p
1
(W
1
), D
2
= f p
1
(W
1
), we have a dieomorphism F : W
1
W
2
.
It follows that the restrcition of f to D
1
is
f = p
2
F
1
: D
1
D
2
,
which is a composite of dieomorphisms, and therefore a dieomorphism.
DIFFERENTIAL GEOMETRY 29
Notice what this theorem says: if the dierential df(p) is invertible at each p
then locally f possesses an inverse, but this need not mean f itself is invertible.
The following example demonstrates this point.
Example 2.7. Let S
1
R
3
be the plane z = 0 and S
2
be the cylinder (x, y, z) :
x
2
+z
2
= 1. The map
f : S
1
S
2
; f(x, y, 0) = (cos(x), y, sin(x)),
wraps the plane S
1
around the cylinder S
2
innitely many times, so it is not
injective and cannot have an inverse. It is plainly smooth and at each p = (x, y, 0)
clearly T
p
S
1
is the subspace of R
3
with equation z = 0. S
1
can be globally
parameterised by p(u, v) = (u, v, 0), for which
(f p)
u
= (sin(u), 0, cos(u)), (f p)
v
= (0, 1, 0).
These are linearly independent at every point p(u, v), therefore
df(p)[X] = X
u
(sin(x), 0, cos(x)) +X
v
(0, 1, 0) = (X
u
sin(x), X
2
, X
1
cos(x)),
has trivial kernel. So by Theorem 2.17 f is locally invertible.
A smooth map f : S
1
S
2
which is a local dieomorphism about each point
is simply called a local dieomorphism.
30 IAN MCINTOSH
3. The Geometry of Smooth Surfaces.
Throughout this chapter S will be a smooth surface in R
3
.
3.1. The Riemannian Metric. Euclidean space R
3
comes equipped with a
canonical inner product: the dot product. Since each tangent space T
p
S R
3
is
a vector subspace each inherits this inner product:
X, Y
p
= X

Y, X, Y T
p
S R
3
. (3.1)
Recall that the dening properties of an inner product are
X, Y = Y, X
aX +bY, Z = aX, Z +bY, Z, a, b, R
X, X > 0, X ,= 0.
The rst two say that , is a symmetric bilinear form, and the third one says
it is positive denite.
Denition 3.1. The rst fundamental form of a smooth surface S R
3
is the
inner product (3.1) induced on each tangent space T
p
S by the dot product in R
3
.
This is also called the induced Riemannian metric.
We should keep in mind that the rst fundamental form is a family of metrics,
one for each tangent space. In modern language it is an example of a tensor on
S. When we want to emphasise that we are making a calculation at a particular
point p we can write ,
p
.
We can express the information carried by the Riemannian metric in coor-
dinates. Let p : U S be a local parametrisation corresponding to a co-
ordinate chart : D U on S. Recall that at each p = p(a) we have
T
p
S = Spp
u
(a), p
v
(a). Thus X, Y T
p
S can be expressed in this basis as
X = X
u
p
u
+X
v
p
v
, Y = Y
u
p
u
+Y
v
p
v
,
where X
u
, X
v
, Y
u
, Y
v
R are the components of X, Y in these coordinates. Then
X, Y = X
u
p
u
+X
v
p
v
, Y
u
p
u
+Y
v
p
v

= X
u
Y
u
[p
u
[
2
+ (X
u
Y
v
+X
v
Y
u
) p
u
, p
v
+X
v
Y
v
[p
v
[
2
=
_
X
u
X
v
_
_
[p
u
[
2
p
u
, p
v

p
v
, p
u
[p
v
[
2
__
Y
u
Y
v
_
.
Notice that this is nothing other than the usual process for expressing an inner
product as a matrix using a basis for the vector space. The matrix will, of course,
be symmetric and positive denite.
Following the terminology due to Gauss himself we dene the components of
this matrix to be
E = [p
u
[
2
= p
u

p
u
, F = p
u
, p
v
= p
u

p
v
, G = [p
v
[
2
= p
v

p
v
.
DIFFERENTIAL GEOMETRY 31
These are called the components of the Riemannian metric in the given chart.
With this notation
_
X
u
X
v
_
_
E F
F G
__
Y
u
Y
v
_
. (3.2)
N.B. Notice that as we move though the chart domain D = p(U), E, F, G will be
functions of (u, v). If we choose another parametrisation then we should expect
all the components E, F, G (and X
u
, X
v
, Y
u
, Y
v
) to be dierent, but the quantity
X, Y is geometric and will not change.
Since p
u
, p
v
are the tangent vectors along the coordinate curves through the
chart domain D = p(U) we make the following denitions.
Denition 3.2. We say a coordinate parametrisation p(u, v) is orthogonal when
F(u, v) = p
u
, p
v
= 0 throughout U. We say these coordinates are isothermal (or
conformal) when additionally E(u, v) = G(u, v) throughout U, and orthonormal
when we also have E = G = 1.
These are properties of the coordinates, not the Riemannian metric.
Example 3.1 (Planes). Let S be an arbitrary plane in R
3
, containing the point
b R
3
and with normal n R
3
. For any two linearly independent X, Y R
3
with X

n = 0 = Y

n we can parameterise S by
p(u, v) = b +uX +vY, u, v R
2
.
Clearly p
u
= X and p
v
= Y so in this parametrisation the Riemannian metric
has components
E = [X[
2
, F = X

Y, G = [Y [
2
.
The coordinate system (u, v) will therefore be orthogonal precisely when X, Y
are orthogonal, isothermal when X, Y are orthogonal and have the same length,
and orthonormal when X, Y are orthonormal.
Example 3.2 (Cylinder). We can parameterise part of the cylinder S = (x, y, z) :
x
2
+z
2
= 1 by
p(u, v) = (cos(u), v, sin(u)); < u < , v R.
In this case
p
u
= (sin(u), 0, cos(u)), p
v
= (0, 1, 0),
so that the Riemannian metric components are
E = [p
u
[
2
= sin
2
(u) + cos
2
(u) = 1, F = p
u

p
v
= 0, G = [p
v
[
2
= 1.
So these coordinates on the cylinder are also orthonormal.
Example 3.3 (Helicoid). The helicoid may be dened parametrically as follows:
S = p(u, v) = (ucos v, u sin v, av) : u, v R,
where a ,= 0 is constant. In this parametrisation
p
u
= (cos v, sin v, 0), p
v
= (usin v, ucos v, a),
32 IAN MCINTOSH
so that
E = cos
2
v + sin
2
v = 1, F = 0, G = u
2
(sin
2
v + cos
2
v) +a
2
= u
2
+a
2
.
So this is an orthogonal coordinate system on the helicoid which is not isothermal:
see Figure 12.
Figure 12. Orthogonal coordinate curves on a helicoid.
3.2. Lengths and Areas. The Riemannian metric provides all the information
required to calculate the arc length of curves on the surface S, or the area of
bounded regions on S.
Lengths of curves. Let c : [a, b] S be a smooth curve, and suppose it lies entirely
in a chart domain (D, ). In this chart c(t) has coordinates (u(t), v(t)) = (c(t)),
equally, c(t) = p(u(t), v(t)). Therefore
dc
dt
=
p
u
du
dt
+
p
v
dv
dt
= u

p
u
+v

p
v
.
Therefore
[c

[
2
= E.(u

)
2
+ 2Fu

+G.(v

)
2
= c

, c

.
Of course, this is just the expression (3.2) for the squared length of the tangent
vector c

and only involves the Riemannian metric. It follows that the arc length
of c is
s(a, b) =
_
b
a
[c

(t)[dt =
_
b
a
_
E.(u

)
2
+ 2Fu

+G.(v

)
2
dt. (3.3)
In case c(t) does not lie inside one coordinate domain, we can break c into a nite
number of segments each of which lies in some coordinate domain, and add the
arc lengths of each segment.
Remark 3.1. This is the origin of the classical (and still quite commonly used)
expression for the Riemannian metric:
ds
2
= Edu
2
+ 2Fdudv +Gdv
2
,
sometimes referred to as the element of arc length. This has the virtue that
it makes clear the role of E, F, G as coordinate dependent components of a sym-
metric bilinear form.
DIFFERENTIAL GEOMETRY 33
Areas. We can compute the area of any region R S which is the image of
a region Q U R
2
, by some local parametrisation p : U S, over which
integration in the plane is well-dened
4
. For in that case standard vector calculus
tells us that
Area(R) =
__
Q
[p
u
p
v
[du dv =
__
Q

EGF
2
du dv. (3.4)
using Lagranges formula [p
u
p
v
[
2
= [p
u
[
2
[p
v
[
2
(p
u

p
v
)
2
. As with arc length,
the element of surface area

EGF
2
du dv is a coordinate invariant quantity
even though the function

EGF
2
may change.
3.3. Isometries and Local Isometries. Recall that a smooth map f : S

S
between smooth surfaces S and

S is said to be a local dieomorphism at p S
if df(p): T
p
S T
f(p)

S is a linear isomorphism.
Denition 3.3. A smooth map f : S

S is said to be a local isometry at p S
if df(p): T
p
S T
f(p)

S is a linear isometry of vector spaces:

df(p)[X], df(p)[Y ]
_
= X, Y , X, Y T
p
S. (3.5)
Alternatively, we say that f preserves the rst fundamental forms of S and

S.
Remark 3.2. It suces to check equation (3.5) on all pairs of vectors in a basis of
T
p
S. In particular, if df(p) maps an orthonormal basis of T
p
S to an orthonormal
basis of T
f(p)

S then df(p) is a linear isometry, and conversely. It follows that a
linear isometry is a linear isomorphism. Hence if f is a local isometry at p then
f is a local dieomorphism at p.
Denition 3.4. We say that f is a local isometry if f is a local isometry at each
point. If in addition f is a dieomorphism then f is called an isometry and S,

S are said to be isometric.


The relation of being isometric is an equivalence relation on the set of all
smooth surfaces. The relationship between these dierent types of map may be
summarised by the following diagram:
Isometry Dieomorphism

Local isometry Local dieomorphism
Example 3.4. In example 2.7 we showed that the plane S with equation z = 0
and the cylinder

S with equation x
2
+ z
2
= 1 are locally dieomorphic, but not
dieomorphic. Recall we used the map
f : S

S; f(x, y, 0) = (cos x, y, sin x).
4
For technical reasons, we stick to regions which are the closure of bounded open sets whose
boundary is a nite union of piecewise continuous curves.
34 IAN MCINTOSH
We will show that this is a local isometry. In terms of the global parametrisation
p(u, v) = (u, v, 0) on S we have:
p
u
= (1, 0, 0), p
v
= (0, 1, 0),
So this is an orthonormal basis for T
p(u,v)
S. Now
df[p
u
] =

u
f(p(u, v)) = (sin u, 0, cos u), df[p
v
] =

v
f(p(u, v)) = (0, 1, 0),
which is an orthonormal basis of T
f(p(u,v))

S. Thus df maps an orthonormal basis
to an orthonormal basis, hence f is a local isometry. However, f is cannot be an
isometry because it is not a dieomorphism.
We can check whether a map is a local isometry by looking at how it relates
the components E, F, G on S with the components

E,

F,

G on

S, but since these
components are calculated in charts we need to rst ensure that we are making
the calculation in compatible charts. The correct notion for compatibility is given
by the next denition.
Denition 3.5. Suppose f : S

S is smooth. Charts (D, ) on S and (

D, )
on

S are said to be f-adapted if (D) = (

D) and f = .
This means the following diagram commutes
D
f
-
D
U

This can be equally well expressed using the corresponding local parametrisations
(p, U) and ( p,

U) by saying U =

U and f p = p.
Notice that since both and are dieomorphisms the only way they can be
f-adapted is if f : D

D is a dieomorphism.
With this notion of f-adapted charts we have the following useful result.
Lemma 3.6 (E, F, G Lemma). Suppose f : S

S is smooth. Then f is a local
isometry at p S if and only if there exist f-adapted charts (D, ) about p and
(

D, ) about f(p) such that
(

E,

F,

G) = (E, F, G), at (p).
Proof. For f-adapted charts we have
p
u
= (f p)
u
= df[p
u
], p
v
= (f p)
v
= df[p
v
].
Hence

E = [ p
u
[
2
= [df[p
u
][
2
,

F = p
u
, p
v
= df[p
u
], df[p
v
],

G = [ p
v
[
2
= [df[p
v
][
2
.
DIFFERENTIAL GEOMETRY 35
Write X
1
= p
u
(a) and X
2
= p
v
(a), where a = (p). Then
(

E,

F,

G) = (E, F, G) at (p)

df(p)[X
i
], df(p)[X
j
]
_
= X
i
, X
j
, i, j = 1, 2
df(p) is a linear isometry
f is a local isometry at p.
It remains to show that if f is a local isometry at p then f-adapted charts exist.
Since f is a local dieomorphism at p we can nd charts D about p and

D about
f(p) such that f : D

D is a dieomorphism, by the Inverse Function Theorem
for surfaces. If the chart map for D is : D U, redene the chart map for

D
by :

D U; = f
1
. Then is smooth (by the Chain Rule for surfaces),
and invertible with inverse p = f p also smooth; hence is indeed a chart map
for

D.
3.4. The Shape Operator. At each point on a smooth surface S we have a
choice of two unit normal vectors. In any coordinate chart it is always possible
to smoothly assign a unit normal to each point, using the local parametrisation
p : U S as follows:
(p(a)) =
p
u
(a) p
v
(a)
[p
u
(a) p
v
(a)[
, a U.
We say that the coordinates give an orientation to S in D = p(U). For some
surfaces it is not possible to extend such an orientation smoothly to the whole
surface: the Mobius band is one example.
Denition 3.7. A smooth surface S is said to be orientable if there exists a
smooth unit normal eld on S, i.e., a smooth function : S R
3
satisfying for
all p S:
(p) T
p
S and [(p)[ = 1.
If S is orientable then there are precisely two smooth unit normal elds, choice
of either of which constitutes an orientation of S. An oriented surface is an
orientable surface together with an orientation.
Example 3.5. Every smooth level surface S
k
= (x, y, z) : F(x, y, z) = k, for a
regular value k, is orientable by the unit normal obtained from F:
(p) =
F(p)
[F(p)[
.
Denition 3.8. For a smooth oriented surface S the chosen unit normal eld,
thought of as a surface map : S S
2
, is called the Gauss map.
The dierential d(p) of the unit normal eld measures the rate of turning of
the tangent plane T
p
S as p moves along S. It is a linear map d(p) : T
p
S
T
(p)
S
2
, but we know that
T
(p)
S
2
= (p)

= T
p
S. (3.6)
36 IAN MCINTOSH
S

S
2
p
(p)
(p)
Figure 13. The Gauss map
Using this identication we regard this dierential as a linear map from T
p
S to
itself:
d(p) : T
p
S T
p
S.
Denition 3.9. The shape operator (or Weingarten map) of S at p is the linear
map/operator:
A
p
: T
p
S T
p
S; A
p
= d(p).
We often abbreviate A
p
= A when the point p S is understood, or to indicate
the family of maps, one at each point.
Notice that choosing the opposite orientation, , changes the sign of the shape
operator: A A.
Example 3.6 (Plane). Whichever orientation is selected, is constant and hence
d(p) = 0 (the zero map T
p
S R
3
) for all p S. Thus A(X) = 0 for all
X T
p
S.
Example 3.7 (Sphere). Let S be the sphere of radius R centred at the origin, and
choose (p) = p/R for all p S, which is the outward-pointing unit normal.
Then:
d(p)[X] = (X) = X/R, for all X T
p
S,
because is the restriction of the linear map v v/R on R
3
. Thus A(X) =
(1/R)X for all tangent vectors X.
Example 3.8 (Paraboloids). For each r R we dene a smooth function f
r
: R
2

R by:
f
r
(x, y) = x
2
+ry
2
.
Let S
r
be the graph of f
r
. Then S
r
is a smooth surface, with a global chart. The
intersection of S
r
with any vertical plane is a parabola. If r > 0 (resp. r < 0)
then the level curves of f
r
are ellipses (resp. hyperbolas). Consequently S
r
is
called an elliptic (resp. hyperbolic) paraboloid. The hyperbolic paraboloid is the
archetypal saddle surface. When r = 0 the surface is a parabolic cylinder: The
paraboloid S
r
may also be viewed as the level set of the smooth function:
F
r
(x, y, z) = z x
2
ry
2
,
DIFFERENTIAL GEOMETRY 37
r > 0 r < 0
Figure 14. Elliptic and hyperbolic paraboloids; parabolic cylinder.
corresponding to the regular value 0 R. We give S
r
the induced orientation:
(x, y, z) =
F
r
(x, y, z)
[F
r
(x, y, z)[
=
(2x, 2ry, 1)
_
4x
2
+ 4r
2
y
2
+ 1
.
Let p = (0, 0, 0), the unique point in common to all the S
r
. Since F
r
(p) =
(0, 0, 1) the tangent space T
p
S
r
is the xy-plane. We will compute the shape
operator of S
r
at p, acting on unit vectors in the form
T

= (cos , sin , 0),


where [0, 2). A smooth curve c

(t) in S
r
with c

(0) = T

is
c

(t) =
_
t cos , t sin , t
2
(cos
2
+r sin
2
)
_
.
Therefore, by denition of the shape operator:
A(T

) = d(p)[T

] = ( c

(0) =
d
dt
[
t=0
(2t cos , 2rt sin , 1)

4t
2
cos
2
+ 4r
2
t
2
sin
2
+ 1
= (2 cos , 2r sin , 0).
Notice that for all , [0, 2) we have:
A(T

), T

= 2 cos cos + 2r sin sin = T

, AT

.
It follows from the linearity of A that:
A(X), Y = X, A(Y ), for all X, Y T
p
S
r
.
If = +/2 then (T

, T

) is an orthonormal basis of T
p
S
r
and
A(T

), T

= 2 cos
2
+ 2r sin
2
,
A(T

), T

= 2 sin
2
+ 2r cos
2
,
A(T

), T

= (r 1) sin(2) = T

, A(T

).
It follows that the matrix of A
p
with respect to this basis is
_
2 cos
2
+ 2r sin
2
(r 1) sin(2)
(r 1) sin(2) 2 sin
2
+ 2r cos
2

_
38 IAN MCINTOSH
In particular, if = 0 then this symmetric matrix becomes diagonal
_
2 0
0 2r
_
from which it follows that T
0
and T
/2
are eigenvectors of A
p
, with corresponding
eigenvalues 2 and 2r respectively.
Now recall, from linear algebra, that a linear operator L on a nite-dimensional
inner product space (V, , ) is said to be self-adjoint, or symmetric if:
L(v), w = v, L(w), for all v, w V .
An operator is symmetric if and only if its matrix with respect to any orthonormal
basis of V is symmetric (hence the name). The previous example illustrates a
general property of the shape operator.
Lemma 3.10 (Shape Lemma). For any smooth surface S, the shape operator A
p
is a self-adjoint operator on T
p
S, for all p S.
Proof. Since A is a linear operator, it suces to establish its symmetry on a basis
(X
1
, X
2
) of T
p
S. Clearly A(X
i
), X
i
= X
i
, A(X
i
), so it suces to show:
A(X
1
), X
2
= X
1
, A(X
2
).
For this, we choose a chart about p and take the basis X
1
= p
u
and X
2
= p
v
.
Then
A(X
1
) = d(p)[ p
u
] = ( p)
u
, A(X
2
) = d(p)[ p
v
] = ( p)
v
.
Thus
A(X
1
), X
2
= ( p)
u

p
v
= ( p)

p
vu
,
where p
vu
=
2
p/uv. The last equality follows from ( p)

p
v
= 0, by dier-
entiation with respect to u. Now swapping u with v in the previous calculation
gives
A(X
2
), X
1
= ( p)
v

p
u
= ( p)

p
uv
.
Since p
uv
= p
vu
we have shown that A(X
1
), X
2
= A(X
2
), X
1
.
3.5. The Geometry of Curves on a Surface. Let c(t) be a regular smooth
path in a smooth surface S, with unit tangent vector T(t) = c

(t)/[c

(t)[ T
c(t)
S.
Denition 3.11 (Darboux Frame). Let be the unit normal eld (Gauss map)
on S. Along the curve c(t) in S we dene
V (t) = (c(t)), U(t) = V (t) T(t). (3.7)
Notice that U(t) T
c(t)
S. It is called the intrinsic normal to c(t). The triple of
vectors (T(t), U(t), V (t)) is a positively oriented (i.e. right-handed) orthonormal
basis of R
3
, called the Darboux frame along c(t).
DIFFERENTIAL GEOMETRY 39
The intrinsic normal may or may not agree with the principal normal N(t) of
c(t), and so one does not expect the Darboux frame to agree with the Frenet
frame. In fact the Darboux frame is dened at every point on c(t), including
points of inection (where N(t) is not dened).
Recall that c(t) has curvature vector k(t) orthogonal to T(t). Therefore
k = (k

U)U + (k

V )V. (3.8)
Denition 3.12. In the previous decomposition, the component (k

U)U is called
the geodesic curvature vector, and
g
= k

U is called the geodesic curvature. If

g
vanishes identically then c(t) is called a geodesic (curve) of S.
The component (k

V )V is called the normal curvature vector, and


n
= k

V
is called the normal curvature. If
n
vanishes identically then c(t) is said to be
an asymptotic curve of S.
If the orientation of S is reversed then both U and V , and hence
g
and
n
,
change sign.
Remark 3.3. Geodesics are a very important family of curves on a surface, but
sadly we dont have time in this course to study them properly. The condition

g
= 0 says that the geodesic is doing the least amount of turning possible to stay
on the surface. It turns out that this makes geodesics locally distance minimising,
i.e., given two points on a geodesic which are close enough together, that segment
of the geodesic has shortest length over all curves on the surface with the same
end points. This is the reason why they are of such interest: they generalise the
idea of a straight line to curved spaces.
The geodesic and normal curvatures can be explicitly calculated from the ex-
pression for the curvature vector.
Lemma 3.13. For a curve c(t) on a surface S, not necessarily unit speed, the
geodesic curvature and normal curvature are given by

g
=
[c

, c

, V ]
[c

[
3
,
n
=
c

V
[c

[
2
. (3.9)
Further, the curvature = [k[ of c(t) satises
2
=
2
g
+
2
n
.
Proof. The equation
2
=
2
g
+
2
n
follows immediately from (3.8) since U, V are
unit vectors. Now we recall from earlier the expression
k =
1
[c

[
2
c

[c

[
4
c

.
Since U = V T and c

T = 0 we notice that
c

U = [c

, V, T] = [V, T, c

] = 0,
Thus

g
= k

U =
1
[c

[
2
[c

, V, T] =
1
[c

[
3
[c

, c

, V ],
40 IAN MCINTOSH
since T = c

/[c

[. Similarly, c

V = 0 so that

n
= k

V =
1
[c

[
2
c

V.

The geodesic curvature is the amount of curvature perceivable in S, and


geodesics are therefore the analogues in S of straight lines in the Euclidean plane.
They play an essential r ole in any deeper investigation of the Riemannian geom-
etry of S.
Example 3.9. Let S be the helicoid from Example 3.3, with the standard parametri-
sation p(u, v) = (ucos v, u sin v, av). The u coordinate lines are the straight lines
with equation, for each v R,
c
v
(t) = (t cos(v), t sin(v), av).
Every point of the helicoid lies on one of these lines, so it is called a ruled surface
(and these lines are called the rulings). Since every straight line has = 0 it
follows immediately from
2
=
2
g
+
2
n
that
g
= 0 =
n
for these rulings.
Therefore they are simultaneously geodesics and asymptotic curves.
Now x u R and c(t) = (ucos t, usin t, at) be the corresponding v-coordinate
line. This is a helix. We choose the orientation of S induced by the parametrisa-
tion, that is
=
p
u
p
v
[ p
u
p
v
[
.
Since
p
u
= (cos v, sin v, 0), p
v
= (usin v, u cos v, a),
we have
p
u
p
v
= (a sin v, a cos v, u),
and hence:
=
_
a sin v

a
2
+u
2
,
a cos v

a
2
+u
2
,
u

a
2
+u
2
_
. (3.10)
Therefore
V (t) = (c(t)) =
_
a sin t

a
2
+u
2
,
a cos t

a
2
+u
2
,
u

a
2
+u
2
_
.
Since also
c

(t) = (usin t, ucos t, a), c

(t) = (ucos t, usin t, 0),


we nally compute

n
=
c

V
[c

[
2
= 0,
DIFFERENTIAL GEOMETRY 41
and

g
=
[c

, c

, V ]
[c

[
3
=
u
(a
2
+u
2
)
2

usin t cos t a sin t


ucos t sin t a cos t
a 0 u

=
u
a
2
+u
2
.
Thus, all the v-coordinate lines are asymptotic curves. Notice also that each one
has constant geodesic curvature, and precisely one (the axis u = 0 of the helicoid)
is a geodesic of the helicoid (of course, since it is a Euclidean straight line).
3.6. Normal Curvature and Principal Curvatures. For a smooth surface S
with unit normal (p) at p S, we say a plane
p
R
3
is a normal plane at p
if
p
contains the normal line p + t(p). Each normal plane intersects S along
a normal section S S, which is an unparameterised smooth curve on S. It
is not hard to see that for every unit tangent vector T T
p
S there is a normal
plane
p
whose normal section is tangent to T: simply take

p
= p +uT +v(p) : u, v R.
As an oriented plane (with orientation T (p)),
p
is uniquely determined by
the unit tangent vector T.
It is important to realise that although
p
is a normal plane to S at p,
p
is not necessarily normal to S at other points of
p
S. So, in general, normal
sections at p will not be normal sections at other points.
S
(p)
p
Figure 15. Normal sections
Remark 3.4. Once
p
has been oriented by the choice of T, there is a unique unit
speed curve c(t) with c(0) = p and c

(0) = T. Its curvature vector k is unchanged


by orientation preserving reparametrisation, and its normal curvature k(p)

(p)
may therefore be considered the normal curvature of the oriented normal section

p
S determined by T. We will write this as
n
(T). In fact, it is easy to show
that
n
(T) equals the signed curvature of c(t), as a curve in the oriented plane

p
, since the choice of orientation makes (p) the intrinsic normal for c(t).
42 IAN MCINTOSH
Lemma 3.14. Given a unit vector T T
p
S, the normal curvature of the normal
section through p which it determines satises

n
(T) = T, A
p
(T), (3.11)
where A
p
is the shape operator on T
p
S.
The proof follows from a more general calculation. If c(t) is any smooth regular
path in S (not necessarily a normal section), then from (3.9)
[c

[
2

n
= c

V = c

, since c

is orthogonal to V ,
= c

( c)

= c

d[c

],
= c

, A(c

).
Therefore, dividing through by [c

[
2
gives (3.11) at every point along c(t). Hence
the normal curvature of any curve depends only on its unit tangent vector at
each point.
The equation (3.11) gives a geometric interpretation of the shape operator as
a quadratic form. Further, since A
p
is a symmetric operator it is diagonalisable,
with real eigenvalues
1
,
2
: there exists an orthonormal basis (Z
1
, Z
2
) of T
p
S for
which Z
1
and Z
2
are eigenvectors of A
p
:
A
p
(Z
1
) =
1
(p)Z
1
, A
p
(Z
2
) =
2
(p)Z
2
.
Then
i
= Z
i
, A(Z
i
) =
n
(Z
i
).
Denition 3.15. The eigenvalues
i
(p) are called the principal curvatures (of
S at p), and the unit eigenvectors Z
i
are called the principal directions (of S
at p).
Remark 3.5. If the orientation of S is reversed (i.e. we choose instead of )
then the shape operator changes sign and hence so do
1
and
2
.
Theorem 3.16 (Euler). The principal curvatures (of S at p) are the maximum
and minimum normal curvatures (of S at p).
We should keep in mind that it is possible for
1
=
2
, in which case
n
(T) =
1
for all T and every direction is a principal direction.
Proof. Any unit vector T T
p
S may be written
T = (cos )Z
1
+ (sin )Z
2
,
for some [0, 2). Then
A(T) = (cos )
1
Z
1
+ (sin )
2
Z
2
,
so

n
(T) = T, A(T) = (cos
2
)
1
+ (sin
2
)
2
,
from which it follows that
n
always lies between
1
and
2
.
DIFFERENTIAL GEOMETRY 43
Remark 3.6. Eulers Theorem is nothing other than an application of the more
general statement that any real quadratic form q(x, y) = ax
2
+2bxy +cy
2
, when
restricted to the unit circle x
2
+y
2
= 1, takes its maximum and minimum values
on the eigenvectors of the symmetric matrix
A =
_
a b
b c
_
which corresponds to it. It is also a general fact that when A has distinct eigen-
values the eigenvectors are orthogonal, hence for
1
,=
2
the principal directions
are orthogonal.
3.7. Gaussian and Mean Curvatures. Given any basis E
1
, E
2
for the tangent
space T
p
S the shape operator A
p
is represented by a matrix A dened by the
property that, for any X = X
1
E
1
+X
2
E
2
T
p
S,
A
_
X
1
X
2
_
=
_
U
1
U
2
_
for AX = U
1
E
1
+U
2
E
2
.
By standard linear algebra, a change of this basis E
1
, E
2
transforms the ma-
trix A into P
1
AP, where P is an invertible 2 2 matrix determined by the
expressions for the new basis in terms of the old basis. The invariants of this
tranformation will be geometric information, and not basis dependent. Recall
that
tr(P
1
AP) = tr(A), and det(P
1
AP) = det(A),
so these are geometric information, and depend only on the shape operator itself.
Therefore we may write these as tr(A) and det(A). Of course, we may assume A
is the diagonal form, whose diagonal entries are the principal curvatures (in any
order).
Denition 3.17. The Gaussian curvature of S at p is dened to be
K(p) = det(A
p
) =
1
(p)
2
(p),
and the mean curvature of S at p is dened to be
H(p) =
1
2
tr(A
p
) =
1
2
(
1
(p) +
2
(p)).
The Gaussian and mean curvatures together carry all the information carried
by the principal curvature. This is because the characteristic polynomial of A
(whose roots are
1
,
2
) is

A
() =
2
2H +K,
and therefore

1
,
2
= H

H
2
K. (3.12)
Notice that in particular H
2
K 0.
44 IAN MCINTOSH
3.8. Geometric classication of points on a surface. The points of S are
classied according to the relative signs of the principal curvatures. There are
four mutually exclusive classes.
Denition 3.18. We say that p S is:
an elliptic point if
1
,
2
are non-zero and have the same sign at p.
a hyperbolic point if
1
,
2
are non-zero and have opposite signs at p.
a parabolic point if precisely one of
1
,
2
vanishes at p.
a planar point if
1
=
2
= 0 at p.
This point classication is independent of the orientation of S.
The terminology arises from the special case when S is a paraboloid. From
Example 3.8, the principal curvatures of the paraboloid S = S
r
at p = (0, 0, 0)
are 2 and 2r; so p is an elliptic (resp. hyperbolic) point precisely when S is an
elliptic (resp. hyperbolic) paraboloid, and p is a parabolic point if and only if
S is a parabolic cylinder. The paradigm surface for locating planar points is of
course any plane, all of whose points are planar.
The geometric point classication of S may be achieved by inspecting the
Gaussian and mean curvatures:
p elliptic K(p) > 0;
p hyperbolic K(p) < 0;
p parabolic K(p) = 0, H(p) ,= 0;
p planar K(p) = 0 = H(p).
There are some additional geometric conditions which may apply at some (or all)
points p of S.
Denition 3.19. We say that p S is:
an umbilic point if
1
(p) =
2
(p); equivalently, if H(p)
2
K(p) = 0;
a minimal point if
1
(p) =
2
(p); equivalently, if H(p) = 0;
a at point if
1
(p) = 0 or
2
(p) = 0; equivalently, if K(p) = 0.
We say that S is an umbilic (resp. minimal, resp. at) surface if all points of S
are umbilic (resp. minimal, resp. at).
Remark 3.7. A minimal surface gets its name from the fact that on any such
surface a small enough bounded patch of the surface minimises area amongst all
surfaces with the same boundary.
This geometric point classication can be summarised in the following table.
DIFFERENTIAL GEOMETRY 45
Point type Principal curvatures Gaussian & mean curvatures
elliptic
1
,
2
,= 0 have same sign K > 0
hyperbolic
1
,
2
,= 0 have opposite signs K < 0
parabolic precisely one
i
vanishes K = 0, H ,= 0
planar
1
= 0 =
2
K = 0 = H
at at least one
i
vanishes K = 0
umbilic
1
=
2
H
2
K = 0
minimal
1
=
2
H = 0
Example 3.10 (Circular Cylinder). Let S be the circular cylinder with equation
x
2
+y
2
= R
2
, where R > 0. Orient this by outward normal
(x, y, z) =
1
R
(x, y, 0).
This is the restriction of a linear map on R
3
, so
A
p
(X) = d(p)[X] =
1
R
(X
1
, X
2
, 0), for X = (X
1
, X
2
, X
3
).
Notice that Z
1
= (0, 0, 1) T
p
S for all p S, and A(Z
1
) = 0, so 0 is an
eigenvalue. Thus Z
1
is a principal direction, with principal curvature
1
= 0.
Since Z
2
Z
1
it follows that Z
2
= (X
1
, X
2
, 0) where

Z
2
= xX
1
+ yX
2
= 0.
Then:
A(Z
2
) =
1
R
Z
2
.
Hence
2
= 1/R, so:
K = 0 and H =
1
2R
.
Thus S is at, but not planar.
3.9. The Second Fundamental Form. Given an oriented surface S with shape
operator A we have seen, in the Shape Lemma 3.10, that X, A(Y ) is a symmetric
bilinear form on each tangent space.
Denition 3.20. The second fundamental form of S at p is the symmetric
bilinear form

p
: T
p
S T
p
S R;
p
(X, Y ) = X, A
p
(Y ).
46 IAN MCINTOSH
Just like the rst fundamental form we can represent this by a symmetric ma-
trix in any coordinate chart (D, ). We choose this so that its local parametri-
sation p =
1
is compatible with the orientation, i.e.
(p(u, v)) =
p
u
p
v
[p
u
p
v
[
.
Dene:
e = (p
u
, p
u
), f = (p
u
, p
v
) = (p
v
, p
u
), g = (p
v
, p
v
),
which are called the components of the second fundamental form in the given
chart. They are analogous to the components E, F, G of the Riemannian metric.
For example, if X = X
u
p
u
+X
v
p
v
and Y = Y
u
p
u
+Y
v
p
v
then
(X, Y ) =
_
X
u
X
v
_
_
e f
f g
__
Y
u
Y
v
_
Furthermore, e, f, g may be readily computed.
Lemma 3.21. In an orientation preserving parametrisation p : U S the
components of the second fundamental form are given by
e =
[ p
u
, p
v
, p
uu
]

EGF
2
, f =
[ p
u
, p
v
, p
uv
]

EGF
2
, g =
[ p
u
, p
v
, p
vv
]

EGF
2
. (3.13)
[Recall that [u, v, w] = (u v)

w, the scalar triple product.]


Proof. The proof is by straightforward computation, exploiting the identities
( p)

p
u
= 0 = ( p)

p
v
.
By taking partial derivatives of these we see that
e = p
u
, A(p
u
) = p
u

( p)
u
= ( p)

p
uu
,
f = p
u
, A(p
v
) = p
u

( p)
v
= ( p)

p
uv
,
g = p
v
, A(p
v
) = p
v

( p)
v
= ( p)

p
vv
.
Now (3.13) follow from the identity
p =
p
u
p
v
[p
u
p
v
[
=
p
u
p
v

EGF
2
,
in which we have used
[p
u
p
v
[
2
= [p
u
[
2
[p
v
[
2
(p
u

p
v
)
2
= EGF
2
.

The Gaussian and mean curvatures can now be directly computed from the
components of the rst and second fundamental forms.
DIFFERENTIAL GEOMETRY 47
Theorem 3.22. Let S be an oriented surface whose rst and second fundamental
forms have components (E, F, G) and (e, f, g) respectively in some orientation
preserving coordinate chart. Then the Gaussian and mean curvatures in that
chart are given by
K =
eg f
2
EGF
2
, H =
eG2fF +gE
2(EGF
2
)
. (3.14)
Proof. To compute K and H, we need the determinant and trace of A. Suppose
the matrix of A with respect to the basis (p
u
, p
v
) is
_
a b
c d
_
where
A(p
u
) = ap
u
+cp
v
,
A(p
v
) = bp
u
+dp
v
.
(3.15)
Then det A = ad bc and tr A = a + d. In order to express a, b, c, d in terms of
e, f, g we note that
_
e f
f g
_
=
_
p
u
, A(p
u
) p
u
, A(p
v
)
p
v
, A(p
u
) p
v
, A(p
v
)
_
=
_
a[ p
u
[
2
+c p
v
, p
u
b[ p
u
[
2
+d p
v
, p
u

a p
u
, p
v
+c[ p
v
[
2
b p
u
, p
v
+d[ p
v
[
2
_
by (3.15)
=
_
aE +cF bE +dF
aF +cG bF +dG
_
=
_
E F
F G
__
a b
c d
_
Therefore:
_
a b
c d
_
=
_
E F
F G
_
1
_
e f
f g
_
(3.16)
Note that
det
_
E F
F G
_
= EGF
2
= [p
u
p
v
[
2
,= 0,
so the inverse matrix exists. Hence
K = det A = det
_
a b
c d
_
= det
_
e f
f g
__
det
_
E F
F G
_
=
eg f
2
EGF
2
Notice that the multiplicative property of determinants allowed us to achieve this
without having to explicitly perform the matrix inversion in (3.16). However the
trace is not multiplicative, so we now need to develop (3.16)
_
a b
c d
_
=
1
EGF
2
_
G F
F E
__
e f
f g
_
=
1
EGF
2
_
eGfF
gE fF
_
Hence
H =
1
2
tr A =
a +d
2
=
eG2fF +gE
2(EGF
2
)

48 IAN MCINTOSH
Example 3.11 (Circular Torus.). We will nd the Gaussian and mean curvature
of the circular torus of revolution S = T
2
(a, b) where 0 < b < a. This surface is
obtained by rotating around the z-axis a circle in the y, z-plane, of radius b and
centre y = a, z = 0. The whole surface can be described using cylindrical polar
coordinates (r, , z), where r
2
= x
2
+y
2
, as
T(a, b) = (r, , z) : (r a)
2
+z
2
= b
2
.
We will make our calculations using the following local parametrisation:
p(u, v) =
_
cos u(a +b cos v), sin u(a +b cos v), b sin v
_
, 0 < u, v < 2.
From this we will determine the geometric point classication of S, locate special
points, and compute the principal curvatures.
We have:
p
u
= (a +b cos v)(sin u, cos u, 0), p
v
= b(cos usin v, sin usin v, cos v),
and
p
uu
= (a +b cos v)(cos u, sin u, 0),
p
uv
= b sin v(sin u, cos u, 0) = p
vu
,
p
vv
= b(cos ucos v, sin ucos v, sin v).
Therefore
E = [p
u
[
2
= (a +b cos v)
2
, F = p
u

p
v
= 0, G = [p
v
[
2
= b
2
,
so that

EGF
2
= b(a +b cos v). Hence by (3.13)
e =
[ p
u
, p
v
, p
uu
]

EGF
2
=
b(a +b cos v)
2
b(a +b cos v)

sin u cos u
cos u sin u
0 cos v 0

= cos v(a +b cos v),


while
f =
[ p
u
, p
v
, p
uv
]

EGF
2
= 0,
since p
uv
and p
u
are linearly dependent, and
g =
[ p
u
, p
v
, p
vv
]

EGF
2
=
b
2
(a +b cos v)
b(a +b cos v)

sin u cos usin v cos ucos v


cos u sin usin v sin ucos v
0 cos v sin v

= b
_
cos
2
v

sin u cos u
cos u sin u

+ sin
2
v

sin u cos u
cos u sin u

_
= b.
So
eg f
2
= b cos v(a +b cos v),
DIFFERENTIAL GEOMETRY 49
Hence by (3.14)
K =
eg f
2
EGF
2
=
cos v
b(a +b cos v)
and
H =
eG2fF +gE
2(EGF
2
)
=
(a + 2b cos v)
2b(a +b cos v)
Notice that K and H only depend on the latitude (v), as might be expected on
a surface of revolution. Notice also that a + b cos v > 0. Since K and H are
smooth functions on the whole of S, we can obtain their values at the edges of
the parametrisation p(u, v) by continuity (taking limits).
The geometric point classication of S is primarily determined by the sign of
K
K > 0, if 0 v < /2, or 3/2 < v 2;
K < 0, if /2 < v < 3/2.
Thus the outside of S is elliptic, whereas the inside is hyperbolic. Further-
more S has at points (K = 0) on the top and bottom circles of latitude
v = /2 and v = 3/2. Since H = 1/2b ,= 0 at these latitudes, points on these
circles are parabolic. Of the possible special geometric point types, minimal
elliptic
hyperbolic
parabolic
Figure 16. Geometric point classication of the torus.
points occur when cos v = a/2b, which is only possible if b a/2. If b = a/2
the minimal points constitute the inner equator (v = ) of S; otherwise they
constitute the pair of latitudes v = arccos(a/2b), which lie in the hyperbolic
region of S. To investigate the possibility of umbilic points, we compute:
H
2
K =
a
2
4b
2
(a +b cos v)
2
> 0.
Thus there are no umbilic points. Finally, the principal curvatures may be de-
termined using (3.12)

1
,
2
= H

H
2
K =
cos v
a +b cos v
,
1
b
.
50 IAN MCINTOSH
3.10. Gausss Theorema Egregium. The classical approach to the denition
of the Gaussian curvature, given above, is via the shape operator, which measures
normal curvatures. The normal curvatures measure extrinisic geometry of the
surface: they rely on the way the surface sits in the ambient Euclidean space.
Gausss Theorema Egregium (Remarkable Theorem) states that despite this
the Gaussian curvature K = det(A) =
1

2
relies only on the intrinsic geometry
determined by the surfaces metric.
Theorema Egregium. Let S,

S be smooth surfaces with Gaussian curvatures
K,

K respectively. If f : S

S is a local isometry then

K(f(p)) = K(p) for all
p S.
Proof. Our aim is to show that K can be expressed entirely in terms of E, F, G,
the components of the metric for S, hence

K can be expressed entirely in terms of

E,

F,

G, the components of the metric for

S. In that case, since by the (E, F, G)
Lemma 3.6 a local isometry f : S

S has the property that (E, F, G) =
(

E,

F,

G) in f-adapted charts, it follows that K f =

K.
We start from the expression (3.14) for K, which we rewrite as
W
4
K = [ p
u
, p
v
, p
uu
] [ p
u
, p
v
, p
vv
] [p
u
, p
v
, p
uv
]
2
,
where W =

EGF
2
and we have used (3.13) for e, f, g. For any vectors
a, b, c R
3
we write (a b c) for the 3 3 matrix with columns a, b, c and note
that the matrix
_
_
a
b
c
_
_
,
with rows a, b, c, is its transpose and therefore has the same determinant. Then
we can write:
W
4
K = det
_
p
u
p
v
p
uu
_
det
_
p
u
p
v
p
vv
_
det
_
p
u
p
v
p
uv
_
2
= det
_
_
p
u
p
v
p
uu
_
_
det
_
p
u
p
v
p
vv
_
det
_
_
p
u
p
v
p
uv
_
_
det
_
p
u
p
v
p
uv
_
= det
_
_
_
_
p
u
p
v
p
uu
_
_
_
p
u
p
v
p
vv
_
_
_
det
_
_
_
_
p
u
p
v
p
uv
_
_
_
p
u
p
v
p
uv
_
_
_
=

E F p
u

p
vv
F G p
v

p
vv
p
u

p
uu
p
v

p
uu
p
uu

p
vv

E F p
u

p
uv
F G p
v

p
uv
p
u

p
uv
p
v

p
uv
[p
uv
[
2

(3.17)
=

E F p
u

p
vv
F G p
v

p
vv
p
u

p
uu
p
v

p
uu
p
uu

p
vv
[p
uv
[
2

E F p
u

p
uv
F G p
v

p
uv
p
u

p
uv
p
v

p
uv
0

,
DIFFERENTIAL GEOMETRY 51
where to obtain the last line we have used the observation that both determinants
in (3.17) have the same top left hand 2 2 block and therefore it is possible to
move the bottom right hand corner element (viz. [p
uv
[
2
) of the second determinant
across to the rst determinant. (To see this, simply expand the determinants
along their nal column, or row.)
Now we claim that all the entries involving second order derivatives of p(u, v)
can be expressed in terms of E, F, G are their derivatives. The computations are:
p
uu

p
u
= =
1
2
(p
u

p
u
)
u
=
1
2
E
u
,
p
uu

p
v
= (p
u

p
v
)
u
p
u

p
vu
= F
u

1
2
(p
u

p
u
)
v
= F
u

1
2
E
v
,
p
uv

p
u
= p
vu

p
u
=
1
2
(p
u

p
u
)
v
=
1
2
E
v
,
p
uv

p
v
= p
vu

p
v
=
1
2
(p
v

p
v
)
u
=
1
2
G
u
,
p
vv

p
u
= = (p
v

p
u
)
v
p
v

p
uv
= F
v

1
2
(p
v

p
v
)
u
= F
v

1
2
G
u
,
p
vv

p
v
= =
1
2
(p
v

p
v
)
v
=
1
2
G
v
.
Finally, from these it follows that
p
uu

p
vv
= (p
u

p
vv
)
u
p
u

p
vvu
= (F
v

1
2
G
u
)
u
p
u

p
vvu
,
[p
uv
[
2
= (p
u

p
uv
)
v
p
u

p
uvv
=
1
2
E
vv
p
u

p
vvu
.
The equality of the third order partial derivatives gives
p
uu

p
vv
[p
uv
[
2
= F
uv

1
2
G
uu

1
2
E
vv
.

Remark 3.8.
(i) This proof shows that one can write down an explicit expression for the
Gaussian curvature purely in terms of the metric components E, F, G.
Such expressions were rst published by Brioschi, in 1852, 24 years after
Gauss gave the rst proof of his Remarkable Theorem. Our proof follows
Brioschis: Gausss is not so easy to follow.
(ii) By contrast, the mean curvature H is not an intrinsic quantity. This
follows at once from the fact that the plane and the cylinder in Example
3.4 are local isometric but have dierent mean curvatures.
For completeness we give Brioschis formulae for the Gaussian curvature.
52 IAN MCINTOSH
Theorem 3.23 (Brioschi). The Gauss curvature of a smooth surface may be
calculated locally using any one of the following three intrinsic ways:
2WK =

u
_
2GF
v
FG
v
GG
u
GW
_
+

v
_
FG
u
GE
v
GW
_
(3.18)
2WK =

u
_
FE
v
EG
u
EW
_
+

v
_
2EF
u
EE
v
FE
u
EW
_
(3.19)
2WK =

u
_
F
v
G
u
+F
_
ln
_
E/G
_
v
W
_
+

v
_
F
u
E
v
+F
_
ln
_
G/E
_
u
W
_
(3.20)
where W =

EGF
2
. If the coordinates are orthogonal (F = 0) then these
reduce to the following formula:
K =
1
2

EG
_

u
_
G
u

EG
_
+

v
_
E
v

EG
_
_
. (3.21)
Appendix A. Brioschis intrinsic formulae for the Gaussian
curvature.
The aim here is to derive the formulae for the Gaussian curvature given in
Theorem 3.23. Expanding the right hand side and making some initial cursory
groupings of terms produces the following:
W
4
K =
1
2
_
(E
vv
+G
uu
2F
uv
)W
2
+E
u
F
v
G+E
v
FF
v
+EF
u
G
v
+FF
u
G
u
2FF
u
F
v

1
2
_
E
u
FG
v
+E
u
GG
u
+EE
v
G
v
+E
v
FG
u
E
v
FG
u
E
v
FG
u
+ (E
v
)
2
G+E(G
u
)
2
_
_
We now note:
2WW
u
= (W
2
)
u
= E
u
G+EG
u
2FF
u
, (A.1)
and a similar formula for WW
v
, allowing us to regroup as follows:
2W
4
K = (E
vv
+G
uu
2F
uv
)W
2

1
2
(E
u
G+EG
u
2FF
u
)G
u

1
2
(E
v
G+EG
v
2FF
v
)E
v
+E
u
F
v
G+EF
u
G
v
2FF
u
F
v

1
2
E
u
FG
v
+
1
2
E
v
FG
u
= (E
vv
+G
uu
2F
uv
)W
2
WW
u
G
u
WW
v
E
v
+R,
where R denotes the following residual terms:
R = E
u
F
v
G+EF
u
G
v
2FF
u
F
v

1
2
(E
u
G
v
E
v
G
u
)F.
DIFFERENTIAL GEOMETRY 53
These may be written in two ways. First
R = (E
u
G+EG
u
2FF
u
)F
v
EF
v
G
u
+EF
u
G
v

1
2
(E
u
G
v
E
v
G
u
)F
= 2WW
u
F
v
+E(F
u
G
v
F
v
G
u
)
1
2
F(E
u
G
v
E
v
G
u
),
yielding
2W
4
K = W
3
_

v
_
E
v
W
_
+

u
_
G
u
W
_
2

u
_
F
v
W
__
+E(F
u
G
v
F
v
G
u
)
1
2
F(E
u
G
v
E
v
G
u
). (A.2)
On the other hand
R = (E
v
G+EG
v
2FF
v
)F
u
+E
u
F
v
GE
v
F
u
G
1
2
(E
u
G
v
E
v
G
u
)F
= 2WW
v
F
u
+G(E
u
F
v
E
v
F
u
)
1
2
F(E
u
G
v
E
v
G
u
),
yielding
2W
4
K = W
3
_

v
_
E
v
W
_
+

u
_
G
u
W
_
2

v
_
F
u
W
__
+G(E
u
F
v
E
v
F
u
)
1
2
F(E
u
G
v
E
v
G
u
). (A.3)
We now require a slightly more complicated calculation.
Lemma A.1.

v
_
FE
u
EW
_


u
_
FE
v
EW
_
=
1
W
3
_
G(E
u
F
v
E
v
F
u
)
1
2
F(E
u
G
v
E
v
G
u
)
_
. (A.4)

v
_
FG
u
GW
_


u
_
FG
v
GW
_
=
1
W
3
_
1
2
F(E
u
G
v
E
v
G
u
) E(F
u
G
v
F
v
G
u
)
_
. (A.5)
Proof. It suces to prove (A.4). We have:
E
2
W
2
_

v
_
FE
u
EW
_


u
_
FE
v
EW
_
_
= EW(E
uv
F +E
u
F
v
) E
u
F(E
v
W +EW
v
)
EW(E
vu
F +E
v
F
u
) +E
v
F(E
u
W +EW
u
)
= EW(E
u
F
v
E
v
F
u
) EF(E
u
W
v
E
v
W
u
).
Bearing in mind (A.1) we write:
2EW
3
_

v
_
FE
u
EW
_


u
_
FE
v
EW
_
_
= 2W
2
(E
u
F
v
E
v
F
u
) F
_
E
u
(2WW
v
) E
v
(2WW
u
)
_
= 2W
2
(E
u
F
v
E
v
F
u
) E
u
F(E
v
G+EG
v
2FF
v
)
+E
v
F(E
u
G+EG
u
2FF
u
)
= (2W
2
2F
2
)(E
u
F
v
E
v
F
u
) EF(E
u
G
v
E
v
G
u
)
= 2EG(E
u
F
v
E
v
F
u
) EF(E
u
G
v
E
v
G
u
),
54 IAN MCINTOSH
and the result follows on division by 2EW
3
.
Now Brioschis formula (3.18) follows from (A.2) and (A.5), whereas (3.19)
follows from (A.3) and (A.4). Formula (3.20) is simply the average of (3.18) and
(3.19).
References
[Bartle] R Bartle, The elements of real analysis, 2nd ed. Wiley, 1976.
[Spivak] M Spivak, A comprehensive introduction to dierential geometry, Vol. 2, Publish or
Perish, 1999.
Dept. of Mathematics, University of York

You might also like