You are on page 1of 57

Differential Geometry I Script

AUTHORS

04.10.2023
2
Contents

I Preliminaries 5
1 Curves 9
1.1 Definition and some Restrictions . . . . . . . . . . . . . . . . . . 9
1.2 Arc-length . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
1.3 Geometric quantities . . . . . . . . . . . . . . . . . . . . . . . . . 13
1.3.1 The tangent vector . . . . . . . . . . . . . . . . . . . . . . 13
1.3.2 The curvature vector . . . . . . . . . . . . . . . . . . . . . 15
1.4 Curves in R2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
1.4.1 The main interpretations of the curvature scalar . . . . . 22
1.4.2 Curvature determines curve up to rigid motion . . . . . . 27
1.4.3 The interaction between global and local . . . . . . . . . . 29
1.5 Curves in R3 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
1.5.1 First remarks . . . . . . . . . . . . . . . . . . . . . . . . . 31
1.5.2 The curvature scalar in three dimensions . . . . . . . . . 33
1.5.3 Torsion . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34
1.5.4 How Curvature and Torsion determine curve . . . . . . . 37
1.5.5 The Fréchet-Frame . . . . . . . . . . . . . . . . . . . . . . 38
1.5.6 Global theorems for curves in R3 . . . . . . . . . . . . . . 40

2 Surfaces 43
2.1 Some definitions and basic quantities . . . . . . . . . . . . . . . . 43
2.2 The curvature of surfaces in R3 . . . . . . . . . . . . . . . . . . . 48
2.3 The Geometric Definition of Curvature on Surfaces . . . . . . . . 49
2.3.1 A bit about Qp . . . . . . . . . . . . . . . . . . . . . . . . 50
2.3.2 The independence of Qp (v) from the curve we choose . . 52
2.3.3 Proof of the theorem . . . . . . . . . . . . . . . . . . . . . 54
2.4 The second fundamental form . . . . . . . . . . . . . . . . . . . . 57

3
4 CONTENTS
Part I

Preliminaries

5
7

We start with some of the most intuitive examples of the type of manifolds
we will be working with, that is, with curves and surfaces embedded in some
form of RN .
8
Chapter 1

Curves

In this chapter, we will deal with curves. We first define what we mean by
a curve, and impose some restrictions on the kind of curve we want to deal
with. We won’t prove all the things we claim in this chapter, as some of these
things you should have already seen in a Calculus class and this is only a quick
overview.

1.1 Definition and some Restrictions


Given that this is Differential Geometry, we do not want to work with discon-
tinuous curves. We therefore choose to work with smooth curves.

Definition 1.1.1 (Smooth curve). We define a smooth curve in N dimensions


to be a function from some interval I to RN , which is smooth. Mathematically:

A smooth curve is a function γ : I → RN so that it is in C ∞ (I)

The interval can be any sort of interval you want, open, closed, half-open,
etc. We also allow things like (∞, 0]. You can see an example in Figure 1.1.
The thing we are interested the most in in Differential Geometry, is not actually
the parametrization of the curve. What we usually mean by the curve is the
actual line in RN that you can draw on a piece of paper. That is what we mean
by a curve. The actual real geometric line, not the function that assigns it a
parameter value. That is why the parametrization is not the main player in
Differential Geometry. The curve exists independent of parametrization. It is
(mathematically) the image of the function γ, We will use names like γ for the
image of the function, not just for the function, as that is what we care about
the most.
Smoothness is not the only property we will (usually) want a curve we work
with to have, because smoothness in the sense above does not guarantee that
the image of the curve is a smooth object. (It only requires the parametrization
to be smooth.) We can see this with the example below.

9
10 CHAPTER 1. CURVES

Figure 1.1: An example of a smooth curve. The Interval I gets mapped onto a
curve in R3 with the function γ

Figure 1.2: The curve γ : t ∈ R → (t2 , t3 ) ∈ R2 . At t = 0 we see that the image


of the curve is not a smooth object
1.2. ARC-LENGTH 11

Example 1.1.1 (Smooth parametrization doesn’t imply smooth image). Take


the curve γ : t ∈ R → (t2 , t3 ) ∈ R2 . It is clear that this is a smooth curve. (The
coefficients are polynomials in t, all the derivatives exist and are continuous.)
But look at Figure 1.2. The image of the curve is obviously not smooth in R2
at t = 0 or equivalently x = (0, 0). What is happening over there? Well, it
resembles the absolute value function a bit. It also had a sort of sharp bend
at a point. The problem back then was with the derivative. It simply did not
exist, which made the curve have a weird behavior (the sharp bend).Similarly,
here the problem is also with the derivative. It exists, obviously, since this is a
smooth curve. But it becomes 0 at the problem point (t = 0 or the origin). A
curve with such a bend is not something we want to really work with, therefore
we put another restriction on the curves we work with. We eliminate curves like
the one from this example simply by saying we don’t work with curves whose
derivative becomes 0 anywhere.

As we saw in the previous example, we get into problem situations if the


derivative of the curve with respect to the parametrization parameter is zero.
We therefore define regular curves as those for which this doesn’t happen, or in
other words, where the velocity never vanishes.

Definition 1.1.2 (regular curves). A smooth curve is called a regular curve, if:


̸= 0 for all t ∈ I (1.1)
dt
where γ is the smooth curve and I is the interval it is defined on.

Note 1. We will use various notations for the derivative of a curve. These
include:

= γt = γ̇ (1.2)
dt

1.2 Arc-length
Now that we have said what we mean by a curve and restricted it so as to not
run into problems like the one in the example above, we can start with the
geometry. Undeniably, one of the most important quantities in geometry is the
length. If you know the lengths of a problem, you already know quite a bit of
the geometry. What is the length of a (piece of a) curve?
Well, we already restricted ourselves to work with regular curves, so our
motivation will be more on the intuitive side.
Imagine you have any curve, like the one in the figure 1.3. The idea is that
we divide the curve into very small almost-straight parts, calculate the length
of these parts by approximating that part as a straight line and then summing
up all of those back together. We can do it for reasonable (i.e regular) curves.
Of course, in reality what we do is go infinitesimal, at which point this becomes
an integral.
12 CHAPTER 1. CURVES

For the small piece as seen in the figure, we have ∆s = dγ


dt ∆t, where by ’s’
we mean the length and by ∆s the very small length of that very small part.
Afterwards we add all of these up, and in the continuum limit we get an integral:
Z Z t

s(t) = ds = dt (1.3)
γ t0 dt

Figure 1.3: A curve and our intuitive way to understand the definition of the
arc-length. We zoom in on a very small part of the curve, between t′ and t′ +∆t.
There, if ∆t is small enough, the line will be approximately straight and we can
use the velocity vector to calculate the length of that piece approximately. Note
that the velocity vector is drawn in way smaller than it would actually be for
any reasonable ∆t, just so that the picture is clearer.

Definition 1.2.1 (Arc-length). We define the arc-length of a curve γ : I → Rn ,


by first choosing a specify point t0 ∈ I and it’s image γ(t0 ) as a reference point.
Then the arc length s(t) between t and t0 is:
Z t

s(t) = dt (1.4)
t0 dt Rn

Note that in the definition we did not assume that t > t0 , a negative arc-
length is possible, simply by going into the opposite direction of the parametriza-
tion of the curve.
We already mentioned that the geometrically interesting object is the image
of γ, not the function γ (i.e the parametrization) itself. We will care mostly
for things we can define on the image of γ that are not dependant on that
1.3. GEOMETRIC QUANTITIES 13

parametrization. The arc-length is something independent of the parametriza-


tion1
In line with this philosophy, we can define a very convenient, but also more
geometrically ”real” parametrization. The idea is that the arc-length is a geo-
metric object independent of parametrization, and that, for regular curves, we
can use the arc-length to parameterize the curve.
Lemma 1.2.1 (Reparametrization of a regular curve). Let γ : t ∈ I → RN be
a regular curve. Then we can re-parameterize it to a new (regular) curve β(s),
so that

=1 (1.5)
ds
In other words, we parameterize it by the arc-length s,
Proof. We will only sketch the proof, as this is something rather simple and you
very likely already saw the proof in a calculus class2
The first step is to take the arc-length and see it as a function of t:
Z t

s = f (t) = dt (1.6)
t0 dt

We can then take the inverse of this function, call it g(s) = f −1 and express
t as a function of s. If we take β = γ ◦ g = γ(g(s)), we found the right
parametrization. The only thing left that you need to convince yourself is that
the velocity is really of unit length. (You can do this using the chain rule.)
Of course, the curve is still regular, and all the properties like smoothness
are still obeyed by the curve. The image of β and γ is of course exactly the
same, i.e, you can’t change the curve simply by re-parameterizing. You might
find Figure 1.4 helpful in visualizing this.

1.3 Geometric quantities


We continue our search for geometric quantities (other than arc-length), that
we can find in connection to curves. We will find that the tangent vector and
the curvature vector (see definition below) are both independent of the specific
parametrization of our curve and that they tell us a lot of geometric information.

1.3.1 The tangent vector


You have, throughout your studies, seen many functions, and many curves.
Your intuition from Calculus about the tangent vector will probably make you
very quickly say that the the tangent vector (here called τ ) should look like this:
1 Of course, we can always choose to parameterize the curve in the other direction, which

changes the arc-length by a minus. We can also choose a different reference point other than
γ(t0 ). But these are choices that are rather trivial and we won’t really mention them from
now on.
2 If not, try it yourself as an exercise.
14 CHAPTER 1. CURVES

Figure 1.4: Different Parametrizations of the same curve. The curve is drawn in
green, the ticks are the points on the curve with the parameter-values written
next to them. In (a) you see a typical non-special parametrization (i.e the ”t”
, in (b) you find the curve parameterized by the arc-length (s). It is intuitively
clear, why the parametrization is not something geometrically interesting. The
real curve (green line) exists, independent of the ticks. In (c) you find another
parametrization by the arc-length, except with a different choice of reference
point on the curve.
1.3. GEOMETRIC QUANTITIES 15

dγ ?
τ (t) = (1.7)
dt
A bit of thought however, reveals that this cannot be true. Why? Well, it
is not independent of parametrization. Imagine, for example, you were to go
twice as fast along the curve. Then your velocity vector (=tangent vector in
this example) would be, twice as big at every point. But if we want the tangent
vector to be something fundamentally independent of the parametrization, then
equation 1.7 cannot be the correct definition of the tangent vector.
How can we fix this? Well, look at figure 1.5. It shows the same curve,
parameterized in three different ways, with the ”fake” tangent vectors (from
equation 1.7 drawn in. The thing that should jump at you is that, while the
length of the vectors does change, the direction does not 3 . The tangent vector
how we defined it is not the geometrically real thing, rather the unit tangent
vector is, which is exactly how we choose to define it below.

Definition 1.3.1 (tangent vector of a curve). We define the tangent vector τ


of a curve γ : t ∈ I → RN to be the vector:

dγ/dt
τ (t) = (1.8)
|dγ/dt|

If we parameterize by the arc-length, then the formula for the tangent vector
becomes:

τ (s) = (1.9)
ds

since ds = 1.

In this sense, we get something geometric, independent of the parametriza-


tion. You can also think of this as us choosing the tangent vector to be the one
from figure1.5(c)

Note 2. We abused notation a bit. We write γ(s) instead of β(s), since, as we


already mentioned, we only really care about the image of those, and they are
the same, and it avoids cluttered notation. We also write dγ dβ
ds for ds for the same
reason. By the chain rule, we get ds = dt ds , which we also write as dγ
dβ dγ dg dt
dt ds , to
avoid introducing another symbol,g, since it is just the function that expresses
the parameter t in terms of s.

1.3.2 The curvature vector


We now come to the curvature vector. It is the fundamental object that de-
scribes how much the tangent vector of the curve changes. Look at, for example,
figure 1.5(c). It depicts the tangent vector for the curve, which is parameterized
by its arc-length. This vector does not change it’s length (it is per definition of
3 At least to the precision of my drawing skills
16 CHAPTER 1. CURVES

Figure 1.5: The ”fake” tangent vectors from equation 1.7 for different
parametrizations of the same curve (green). In (a) you see a random
parametrization (black ticks) and its ”fake” tangent vector (blue) from equation
1.7. In (b) you have the same situation, only that this time you go twice as fast
along the curve. Notice that the vectors (the physically drawn arrows) change.
In (c) you have the same thing, but this time parameterized with arc-length.
1.3. GEOMETRIC QUANTITIES 17

unit length), but it does rotate as you follow along the curve. Notice also, that
the more the tangent vector rotates, the more ”curved”4 the curve is. That’s
where the name comes from.

Definition 1.3.2 (The Curvature Vector of a Curve). Let γ : s ∈ I → RN be


a regular curve, parameterized by the arc-length. Then we define the curvature
vector κ to be:
dτ d2 γ
κs = = 2 (1.10)
ds ds
It is clear, that, because s is independent of any sort of parametrization, dτ
ds
is too.
Figure 1.6 shows a curve with the curvature vector drawn in. Notice that
the curvature vector seems to be orthogonal to the tangent vector5 . This turns
out to be true, universally, as we will now prove.

Lemma 1.3.1 (κ ⊥ τ ). The curvature vector κ is orthogonal to the tangent


vector τ .

Proof. The proof is strikingly simple. We know that the length of τ is set to
d
one. Therefore ⟨τ, τ ⟩ = 1 and ds ⟨τ, τ ⟩ = 0 since the length (and therefore the
scalar product) doesn’t change along the trajectory. We can use the product
rule:
d dτ
0= ⟨τ, τ ⟩ = 2⟨ , τ ⟩ = 2⟨κ, τ ⟩ (1.11)
ds ds
Therefore, the the scalar product of the two vectors is 0, i.e they are orthogonal,
as claimed.

This is something physics students are very familiar with. The situation is
very analogous to the trajectory of a particle. The speed of the particle doesn’t
change, so the only direction the acceleration (= curvature vector) can have is
perpendicular to the curve.

Note 3. We want to make a quick check on the units of all the quantities that
we described so far. Let’s assume that our RN holds some sort of length unit,
like the cm, which we will write as [L]. Let’s also assume the parameter of
our parametrization has the units of time, like sec, which we denote [T ]. Then
both γ and s have units [L], so the tangent vector dγ ds has units of [L]/[L] = 1
and is unit-less. This is something we want explicitly, as the geometric object
should not be dependant on the parametrization, which means it should also be
independent of the unit of the parametrization [T ]. The ”fake” tangent vector
we defined before has, on the other hand. units of [L]/[T ].
2
The curvature vector ddsγ2 has units of [L]/[L]2 = 1/[L].
4 At least in the intuitive sense, for well-behaving curves.
5 The physics students among you might find this very similar to how in physics we some-
times separate acceleration into a parallel and perpendicular part, the former changing the
speed, the latter curving the trajectory. Since here we don’t change the speed, only the curving
part is left.
18 CHAPTER 1. CURVES

Figure 1.6: A curve with τ and κ drawn in. Notice that κ is orthogonal to τ .

Until now, we have only given a formula for the curvature vector in the arc-
length-parametrization. We will now write down the formula for the curvature
vector with any parametrization.

Lemma 1.3.2 (Curvature Vector in arbitrary Parametrization). Let γ : t ∈


dγ/dt
I → RN be any curve and τ (t) = γt (t) = |dγ/dt| the tangent vector. Then the
6
curvature vector κ(t) can be written as :
 
1 γt γt
κ= 2 γ tt − ⟨γ tt , ⟩ (1.12)
|γt | |γt | |γt |

d2 γ
where γtt is dt2 .

Before we go on to prove this, we first want to talk about what each part of
the equation means.
2
We know that κ = ddsγ2 and therefore expect it to have something to do with
d2 γ
dt2 . This turns out to be the case, the first term is indeed γtt . But there is a
correction term of −⟨γt , |γγtt | ⟩ |γγtt | , which has a nice geometric explanation.
It projects γtt onto the normal plane of the tangent vector. See Figure 1.7
for a visual example. After we have projected γtt onto the normal plane, we
2
still divide it by |γt | . You can see it as just a factor that makes sure that the
units work out.
We can see this simply by comparing units. The part that projects γtt onto
the Normal plane has the same unit as γtt , so we can just look at γtt . (Because
6 If you already have some experience of Differential Geometry or you are rereading this after

already learning further chapters, you might notice how this is the the covariant derivative of
the tangent vector
1.3. GEOMETRIC QUANTITIES 19

2
we add them. That doesn’t change the units.) The unit of γtt = ddt2γ are clearly
[L]/[T ]2 , while the unit of κ is 1/[L], as we saw above. Therefore, to get a
2
consistent formula, we need something that has units of [T ]2 /[L]2 . 1/ |γt | is
exactly a factor like that.

Note 4. We call the normal plane a plane, even though that is technically only
correct if we have a curve in R3 . In R2 it is a line, in R4 a hyperplane and in
general an (N − 1)-dimensional vector-space.

Figure 1.7: A curve with τ and κ drawn in, as well as γtt . Because we move
along the curve faster and faster (the ticks are more spread out), γtt has a
component in the ”forward” direction, which we cancel out in equation 1.12.
2
The vector is still too long though, which is why we need to divide by |γt | .

Proof. We now prove equation 1.12. The proof consist in its most basic form just
of taking the definition of κ in the arc-length-parametrization and switching to
the t-parametrization, using the normal rules of derivatives (chain rule / product
rule). We start with the chain rule.

dτ dt dτ
κ= = (1.13)
ds ds dt
dg
Rt dγ
where by dt
ds we of course mean ds where g = f −1 and f (t) = t0 dt dt.
20 CHAPTER 1. CURVES

Therefore:
 −1  −1
dt dg df dγ 1
= = = = = 1/ |γt | (1.14)
ds ds dt dt |dγ/dt|
If we insert the definition of τ in the t-parametrization we get:
dt dτ 1 d γt
κ= = (1.15)
ds dt |γt | dt |γt |
d|γt |
Now we need to use the quotient rule and the fact7 that dt = ⟨γtt , |γγtt | ⟩.

1 d γt
κ= (1.16)
|γt | dt |γt |
d|γt |
1 (γtt |γt | − γt ( dt )
= 2 (1.17)
|γt | |γt |
1 (γtt |γt | − γt ⟨γtt , γt / |γt |⟩
= 2 (1.18)
|γt | |γt |
 
1 γt γt
= 2 γ tt − ⟨γ tt , ⟩ (1.19)
|γt | |γt | |γt |

And we get the result as promised.

1.4 Curves in R2
We have, by now, defined exactly what we mean by a curve, seen the concept of
what sort of object is geometric, and defined a few of these, like the arc-length,
tangent and curvature vectors. We will now use all of these concepts to describe
curves in the two dimensional plane.
The main idea that makes this a lot simpler, is that the curvature vector κ
reduces to a number. This is because the direction of the curvature is always
predetermined in two dimensions by the direction of the tangent vector.
To see this, we note that, as we showed before, the curvature vector κ lies in
the normal ”plane” of the tangent vector, which in two dimensions means that
it lies on a straight line perpendicular to τ . Therefore, we only need to specify
one number8 to determine the curvature vector.
Let’s say we are at a point on a curve, like the one drawn in figure 1.8. We
can construct a right handed basis of R2 at that point by taking τ as our first
basis vector, and the vector that one gets if one rotates τ by 90 deg (in the
positive sense.), which we will call N . Since τ is of unit length and we get N
by rotating τ by 90 deg, this is an orthonormal basis. (One that is right hand
sided.) Notice that it immediately follows that:
7 You should recognize this from Calculus II, given maybe in a different notation: dr/dt =

x/dt = ⃗xr d⃗
(∇r) ∗ d⃗ x
dt
8 On each point of the curve
1.4. CURVES IN R2 21

Figure 1.8: A curve with its tangent, curvature and normal vectors drawn in
at a point on the curve. As you see, the curvature vector is just some number
times the normal vector. Note however, that k is not just the absolute value of
κ, since it can also be the negative of its length, if it points in the other direction

κ = kN (1.20)
for some k ∈ R, because we know that κ and τ are orthogonal. We call this k the
curvature scalar. It is an important quantity in differential geometry, and we
will find its equivalents for different geometric objects throughout the subject.

Example 1.4.1. Our first example is the simplest curve that is not a straight
line (because a straight line, of course, has no curvature9 ),which is a circle of
radius R.
The curvature of that circle is

k = 1/R (1.21)

if the circle is parameterized in the counter-clockwise direction. Before you


dive into algebra, let’s consider why this result makes a lot of sense. Take the
few circles in figure 1.9. It is intuitively clear, that the bigger the radius of
the circle, the less curved it gets. As you get progressively bigger radii, the
circles look more and more ”flat” at the top, or in other words, less curved.
The biggest circle (only drawn in partially) is almost flat, and if you were to
draw something like R = 100 you could probably not see the difference anymore
between a straight line and the circle. So the result, that the curvature scalar
is the inverse of this radius makes a lot of sense.
9 If you don’t immediately believe this, convince yourself of it.
22 CHAPTER 1. CURVES

The proof of this claim is a very good exercise in converting parametrizations


and getting geometric information out of coordinates, and we will therefore leave
it as an exercise.

Figure 1.9: A few circles with different radii, with their respective τ, κ, N drawn
at the point (0, R) of each curve. The bigger the circle, the less curved it is, as
reflected by the formula k = 1/R.

1.4.1 The main interpretations of the curvature scalar


The curvature scalar has a lot of interpretations. Let’s first state them, then
discuss their consequences.
Proposition 1.4.1 (Interpretations of the curvature scalar). As always, let γ
be a two dimensional curve, with all the properties we already discussed. Then
the curvature scalar has the following interpretations:

1. The curvature scalar is the rate of change of the angle the tangent vector
makes with the x axis. Mathematically, let θ = arctan ττ12 (s)
(s)
be exactly
that angle Then:

k= (1.22)
ds
2. The absolute value of the curvature scalar k tells us the radius of the
osculating circle, which is the distinct circle, that agrees with the curve
1.4. CURVES IN R2 23

up to order two.
1
|k(s)| = (1.23)
R(s)

where R(s) is the radius of that circle at the point of the curve whose
parameter-value is s.

The first interpretation of the curvature scalar should make a lot of sense
intuitively. We know that the tangent vector cannot change its length, since per
construction it is of unit length. Therefore the only thing that can really change
is the direction, i.e the angle it makes with the x-axis. This, along with the fact
that the curvature vector describes how the tangent vector changes, makes the
first part of the proposition rather intuitive. See figure 1.10 for a visualisation.

Figure 1.10: A curve, with its tangent vectors drawn in, and a table that shows
how the tangent vector rotates.

The proof is not to complicated, you just need to derive θ(s) and remember
1
that (1) the derivative of arctan(x) is 1+x 2 and (2) the normal vector in terms

of the components of τ is (−τ2 , τ1 ).

Proof of the first interpretation. As we said, you only need to derive θ. Let’s
24 CHAPTER 1. CURVES

start:
 
dθ d τ2
= arctan (1.24)
ds ds τ1
 
d(arctan(x)) d τ2
= (1.25)
dx ds τ1
1 τ˙2 τ1 − τ2 τ˙1
= (1.26)
1 + x2 τ12
1 τ˙2 τ1 − τ2 τ˙1
= (1.27)
τ2
1 + 22 τ12
τ1
1
= 2 ⟨(τ˙1 , τ˙2 ), (−τ2 , τ1 )⟩ (1.28)
τ1 + τ22
1
= ⟨κ, N ⟩ (1.29)
1
=k (1.30)

The factor in the fraction is one because it’s the square of the length of τ , which
is one.
Now, what about the second interpretation? Well, you can imagine a circle,
going along the curve, that locally looks like the curve. (The curve tries to be as
similar to the circle as possible, but because the radius of the osculating circle
changes with s, it doesn’t become a circle.)
Figure 1.11 gives a picture of a curve and it’s osculating circles at different
points of the curve. As you hopefully agree with, the bigger the radius of the
circle, the more straight the curve will be at that point (as both of them agree
to order two so they locally behave quite similarly.) Therefore we expect that
the second interpretation is correct, that is, the curvature scalar is inverse to
the radius of the osculating circle.
Proof of the second interpretation. We will not prove this, as it is quite simple,
but we will sketch a proof. The osculating circle agrees with γ up to order
two. Therefore, we can expect that the second derivatives (i.e the k’s) agree
for the curve and the osculating circle (which we can see as a second curve.) at
that point. We know that at that point, the circle has kcircle = 1/Rcircle , and
therefore this should also be true for the first curve. The only missing parts of
the proof are (1) the proof that an osculating circle exists, which it does10 , and
a more rigorous way of presenting the above argument.

There is actually also a third interpretation of the curvature scalar, for a


special kind of curve. Let’s say, that the curve is the graph of a function y = u(x)
10 A straight line is a circle of infinite radius in many aspects of geometry, this is also

true here, if the curve is locally straight at a point, the radius of it’s osculating circle will
blow up and the circle will become as straight line, but the theorem will still hold. For the
mathematicians: 1/∞ = 0 in this case.
1.4. CURVES IN R2 25

Figure 1.11: A curve with its osculating circle drawn in at a few places along the
curve. (The biggest one only partially drawn in) It is clear that the bigger the
osculating circle is, the straighter the curve will be, which gives the connection
to the curvature scalar.

Figure 1.12: A graph of a function as a curve.


26 CHAPTER 1. CURVES

that assigns a y-value to every x-value, like the one in figure 1.12. Consider the
second derivative of u. Can we connect it to k, which is also, a second derivative?
Yes. In-fact, this is a very common theme that will accompany you throughout
differential geometry. Curvature is a second derivative and a second derivative
is curvature in some sense11 . The relationship between k and uxx is not trivial
however. k ̸= uxx ! The actual relationship is:
uxx
k= 3/2
(1.31)
(1 + u2x )

We have to compensate, because x is not s. The proof of this is quite simple,


if quite long. The basic strategy is, as with many of these proofs, differentiate
until you get to where you want to be.

Proof. Let’s start, by collecting different terms that might be useful. Firstly,
γ(x) = (x, u(x)) and therefore:

γx = (1, ux ) (1.32)
1/2
|γx | = 1 + u2x (1.33)
γx (1, ux )
τ= = 1/2
(1.34)
|γx | (1 + u2x )
(−ux , 1)
N= 1/2
(1.35)
(1 + u2x )

The first three should be rather clear, coming straight from the definition. The
last one comes from the fact that N is just τ , but rotated by 90 deg, which means
we switch the two entries of the vector and put a minus in-front of the first one12
We can now just use the definition of κ and the chain rule and calculate until
we get there.

dx dτ
κ= (1.36)
ds dx
 −1
dx ds −1
= = |γx | (1.37)
ds dx
1 dτ
→κ= (1.38)
|γx | dx
1 d γx
= (1.39)
|γx | dx |γx |

Before we continue, there is something to note about what we already found.


Inside the derivative, we already normalize once, and then again outside of the
11 Conditions apply, as always.
12 Ifthis is not clear to you, try it out with the rotation matrix of positive 90 deg. You’ll
see that this is correct.
1.4. CURVES IN R2 27

derivative. In this sense, κ is a normalized version of a second derivative.


1 d (1, ux )
... = (1.40)
|γx | dx |γx |
 
1 (0, uxx ) d 1
= − (1, ux ) (1.41)
|γx | |γx | dx |γx |
(We used the product rule.) Now, we know that k = ⟨κ, N ⟩. The last term in
the equation above for k is proportional to (1, ux ) which is proportional to τ ,
which means that when we form the scalar product to get k, it drops out, since
τ is orthogonal (per construction) to N . We get:
k = ⟨κ, N ⟩ (1.42)
1 (0, uxx ) (−ux , 1)
=⟨ , ⟩ (1.43)
|γx | |γx | |γx |
uxx uxx
= 3 = 3/2
(1.44)
|γx | (1 + u2x )
Here, we used that that aforementioned second term is orthogonal to N and left
it out. At the end we just collected terms.
Now, after seeing how much manual computation this took, you might be
a bit astounded as to why. The reason is the same reason why anytime you
actually want to compute something in differential geometry it usually turns
into a mess of derivatives. We are turning something fundamentally coordinate-
based13 (uxx ) into something geometric (k). Coordinate-based objects usually
have, as you might imagine, a lot of information in them that is only related to
the choice of our coordinates and we have to filter that information out when
we do the conversion. This is the reason why there is so much to compute, even
if the steps aren’t too complicated.

1.4.2 Curvature determines curve up to rigid motion


There is one more thing we want to discuss about the curvature scalar before
we move on. We want to talk about how much the curvature (scalar) actually
tells us about a curve, or to what degree it determines the curve, in the sense
that you have a function k(s) which you say is the curvature scalar of the curve,
and ask yourself how much freedom you still have left. It will turn out that the
curvature determines the curve, up to it’s position at s = 0 and the angle of the
tangent vector at that point. This mirrors Newton’s law a lot, the reason being
that both are differential equations of second order14 . You can also see this as
having the freedom to preform any rigid motion (a rotation or translation, but
no mirroring or stretching) and still getting a curve with the same curvature
scalar. You can look at figure 1.13 for an example.
13 To make this discussion more general, we write coordinate-based, even though right now

it’s just a parametrization. You can see a parametrization as coordinates on the curve.
14 The only difference to Newton’s law is that the speed (with respect to s) can’t change for

a curve, that is why we don’t get to pick any tangent vector.


28 CHAPTER 1. CURVES

Figure 1.13: You can preform a rigid motion and not change anything about
the curvature of the curve.

Figure 1.14: A picture of the theorem.


1.4. CURVES IN R2 29

Theorem 1.4.1 (Curvature determines curve up to a rigid motion). The cur-


vature k(s) of a curve determines that curve up to a rigid motion.
Proof. We will, again, not prove this rigorously. We will give you a sketch, from
which it should be clear that a proof can be constructed.
The basic idea is to integrate twice.
Firstly, integrate the equation:

k(s) = (1.45)
ds
to get: Z s
θ(s) = θ(0) + k(s) ds′ (1.46)
s0
where θ(0) is a constant we can choose freely. The next step is to integrate the
equation:

(s) = τ (s) = eiθ(s) (1.47)
ds
and get: Z s

γ(s) = γ0 + eiθ(s ) ds′ (1.48)
s0
giving us yet another constant γ0 , which we can choose freely. To get a more
rigorous proof, you would need to show that these are the solutions (just dif-
ferentiate them) and that these are the only solutions (use a theorem from
calculus.)

1.4.3 The interaction between global and local


Before we close the subject of curves in R2 , we want to talk about a general
theme of differential geometry, that shows up throughout the subject, and apply
it to curves in R2 . The theme is the interaction between global and local prop-
erties of geometric objects. The idea is that local properties (like curvature),
which only feel a tiny piece of the object (around any point), give constrains on
(or even determine) global properties. Local things are things that only need
a small surrounding of a point to be defined at that point, like the curvature
vector, or later the metric. Global things are typically integral quantities, which
often are related to topology. We will give an example (without proof) of a the-
orem that follows along the line of this idea, but before we do that, we need to
define two further restrictions, that we will need to make (and will from now on
assume that the curves we usually work with will usually obey.)
Definition 1.4.1 (more restrictions, simple curves).

1. An N -dimensional smooth curve γ is called simple, if it has no self-


intersections, i.e if γ(s) = γ(t) then s = t. The one exception we make
are the edges, as we don’t want to call a closed curve, like a circle, self-
intersecting, just because it returns back to it’s beginning.
30 CHAPTER 1. CURVES

Figure 1.15: A problem case of a closed curve, for which the tangent vector at
the beginning is not the same as the tangent vector at the end. We want to avoid
this, so we just take these kinds of curves (and ones where higher derivatives
don’t match up) out of the set of curves we consider
1.5. CURVES IN R3 31

2. Similarly, a curve is closed if it is defined on an interval [a, b] and γ(a) =


γ(b)

3. If we are working with closed curves, we will want them to have the (nice)
property that, if we extend them periodically to a curve from R → RN ,
they are smooth. This is to avoid annoying situations like the one in figure
1.15

With these restrictions, we can state the theorem.

Theorem 1.4.2. Let γ be a two dimensional regular closed curve that obeys
the above restriction. Then: Z
kds = 2πn (1.49)
γ

for some n ∈ Z

The integral quantity is the global quantity we mentioned before, while k


Rb
is the curvature scalar, which is a local quantity. Since γ k ds = a dθ
R
ds ds, the
global quantity is just the total angle (with signs) by which the tangent vector
rotated, a profoundly global thing.
It makes sense that this would be so. If the curved is closed (and is smooth
on the edge, if made periodic), then the angle τ has to rotate by must be some
multiple of a whole rotation, since it ends up where it started.

Proposition 1.4.2. If the curve is simple, n = ±1

This proposition tells us that a non-intersecting curve’s τ can only turn once
in total. See the examples in figure 1.16
With this we conclude the topic of two-dimensional curves, and move on to
three-dimensional curves.

1.5 Curves in R3
1.5.1 First remarks
Very early in our discussion of two-dimensional curves we figured out that in
two dimension, the curvature vector is not really necessary for the description
of how the curve curves. Better said, the curvature scalar (the signed length of
the curvature vector) held all the information about curvature, the direction of
the curvature vector was always predetermined. This is very different for curves
in three dimensions. Here, the vector character of the curvature vector really
stands out.
We saw at the beginning of this chapter that the curvature vector lies in
the normal plane of the tangent vector. In two dimensions this helped us, by
letting us forget about the direction of the curvature vector and only consider
the curvature scalar. This time, we cannot do this, as we have an entire plane
that the curvature vector could lie in.
32 CHAPTER 1. CURVES

Figure 1.16: Two simple curves and the graph that shows by how much the
tangent vector rotated.

In the two dimensional case, we defined a moving frame composed of τ and


N , which was a right handed orthogonal basis. This is the idea we will use to
get further in three dimensions.
Definition 1.5.1 (The moving frame). Let, as always, γ be a curve, this time in
R3 with all the properties we already mentioned before (smoothness, regularity
etc.) We define three vectors at each point, that will compose the moving frame
we will usually use.
1. The first is the tangent vector τ . Remember that it is normalized.
2. The second one is the normalized curvature vector, which we will call N ,
κ
N = |κ|

3. The third vector will be defined by the equation β = N × τ . Because both


τ and N are of unit length, β is too, and all three together form a right
handed orthonormal basis.
The second vector will be called the normal vector15 , while the third vector is
called the bi-normal vector.
Together, N and β span the normal plane of τ . For a visualisation of the
moving frame, look to figure 1.17.
15 Even though it is not the only normal vector to τ , but it is a very special normal vector,
1.5. CURVES IN R3 33

Figure 1.17: A three dimensional curve, with the moving frame drawn in. N is
κ, but normalized, β is the cross product of N and τ .

1.5.2 The curvature scalar in three dimensions


Because the curvature in three dimensions is really a vector, unlike in two di-
mensions, we will not find that we can describe the entire curve just with the
curvature scalar. We still introduce it, as |κ|. But this time, it cannot change
signs, which means the two dimensional version is not exactly the same thing
as the three dimensional.

Definition 1.5.2 (Curvature scalar for curves in three dimensions). The cur-
vature scalar is simply the absolute value of the curvature vector, defined as
k = |κ|

Now, you might have noted that to define N , we divided by the curvature
scalar and this becomes a problem, if k is zero. This is an actual problem and
happens for any curve that is (to second order) straight at some point. We will
exclude this, simply by adding another restriction on our definition of a curve.

Definition 1.5.3 (ordinary). A curve γ is called ordinary, if the curvature


vector never vanishes, which means that for all points on the curve |κ| =
̸ 0.

With this aside, let’s go back to the curvature vector. We saw that the
curvature scalar will simply not provide enough information about our curve
that we can paint a complete picture. We will need another agent, which will
be called the torsion.

since it goes in the direction of κ. Therefore we call it the normal vector


34 CHAPTER 1. CURVES

1.5.3 Torsion
Here we will define what we mean by the new agent we said we needed in the
previous section, called torsion. Torsion will also be another geometric object16 .
It will tell us, in a sense, how much the curvature vector changes.
Definition 1.5.4. Let γ be an ordinary three dimensional curve. We define
the torsion vector λ to be the projection of the derivative of the normal vector
onto the bi-normal:
dN
λ=⟨ , β⟩β (1.50)
ds
and the torsion scalar l to be:
dN
l=⟨ , β⟩ (1.51)
ds
Why do we do this? Why don’t we just define the torsion vector to be dNds ?
Well, it turns out, that this has a lot of unnecessary information.
Consider it’s τ and N components. For the τ component we get (product
rule):
dN d dτ d
⟨ , τ⟩ = ⟨N, τ ⟩ − ⟨N, ⟩ = 0 − ⟨N, κ⟩ = −k (1.52)
ds ds ds ds
which is just (minus) the curvature scalar, which we already know, and for the
other component we get (Product rule again):
dN d⟨N, N ⟩ d
2⟨ , N⟩ = = 1=0 (1.53)
ds ds ds
since a unit vector can’t change in it’s own direction (otherwise it’s length would
change.).
That is why we take the projection.It provides us with the only new infor-
mation. The information we want is how the normal plane changes, but only in
the direction of the bi-normal-vector.
We can form a table with all the objects we have so-far introduced and a
few things to note on them. See table 1.1.

object formula name/interpretation derivatives units


γ - Position 0 [L]

τ ds tangent vector / velocity 1 []
κ κ = kN curvature vector / acceleration 2 1/[L]
λ λ = lβ torsion / ”jerk” 3 1/[L]

Table 1.1: The main geometric objects we have defined up-til now.

One thing that you might find surprising at first is that the units of λ are
not 1/[L]2 . The reason is because we normalize κ before differentiating, which
means we multiplied the units by [L].
16 As a reminder, something is a geometric object or geometrically invariant when it does

not change if we change the parametrization, or rotate RN .


1.5. CURVES IN R3 35

Figure 1.18: A curve, and it’s normal vector and plane changing along the curve.

We can see what we are doing as a Taylor-expansion, which we stop after 4


terms (3 derivatives.)
We can draw a parallel between k and l. k measures how much γ deviates
from being a straight line. Look at figure 1.19(a). k is the value that measures
how γ1 deviates from a straight line. In the same figure, in (b), you see a curve
that lives entirely in a plane (even though it might have been defined in three
dimensions.) It’s l value is zero17 . In (c), you see a a curve γ3 , which lives
almost in a plane, close to the vicinity of a particular point on the curve, but
which deviates from that plane. For it l is not zero18 .
Imagine the red plane in the figure is the xy-plane. Then the one curve
would look like γ2 (s) = (x2 (s), y2 (s), 0) and the other curve would look like
γ3 (s) = (x3 (s), y3 (s), z3 (s)), i.e it would have a non-zero third component.
We can expand z3 (s) (which we will call z(s)) around s = 0, which we can
assume to be the parameter of the thick point in the picture. Then, since the
xy-plane is the one that the curve ”almost” lives in, z(0) = 0, of course, but
also z ′ (0) = 0, z ′′ (0) = 0. This means that both the tangent and curvature
vector lie in the xy-plane. The first non-zero derivative will be the third, and if
we Taylor-expand z around 0, we will get something like:

z(s) = 0 + 0s + 0s2 + cls3 + ... (1.54)

where l is the torsion scalar, and c is some universal constant. This is why the
torsion scalar measures how much the curve deviates from living in a plane. As
an exercise, you should prove all the claims we did not prove in this discussion
and find c. You can take the curve γ = (s, as2 , bs3 ), as a first example and see
where in the Taylor-series around 0 you find k and l)
17 Think about why.
18 Itmight be zero on some particular points of an arbitrary curve, where to high order the
curve locally really almost lives in a plane, and only deviates a bit ”further”.
36 CHAPTER 1. CURVES

Figure 1.19: The parallel between k and l. k measures how lose a curve is to
a straight line (Part (a)). In part (b) you see a curve that lives entirely in a
plane, while in (c) you see a curve that deviates from the plane it is almost in,
at least in the direct vicinity of the thick point with the moving frame drawn
in. l is the thing that measures this.
1.5. CURVES IN R3 37

You can also prove that if k = 0 for all s, then the curve is a straight line,
and if l = 0 for all s, the curve lies in a plane. (You can do this as an exercise,
or decide that the above discussion convinced you of this.)
In summary, torsion measures how much a curve twists away from a plane
and into the third dimension.
Example 1.5.1. What is the curve with constant curvature and torsion? We
won’t show it here, but it can be shown that it is a helix. A helix, by the way,
can be parameterized by (R cos(t), R sin(t), mt) where m is some constant.

Figure 1.20: A helix

Why does it make sense that it’s a helix? Well, we want constant curvature,
which means a circle is involved, but we also want it to twist out of the plane
it lives in, at a constant rate, which is why we get a helix.

1.5.4 How Curvature and Torsion determine curve


We have gotten a good intuition about curvature and torsion by now. Are these
all the things we need to know about a curve in R3 ? It turns out that, pretty
much, yes, will be the answer, very similarly to how in two dimensions we only
really need the curvature scalar.
This is stated in the following theorem, which can be proven similarly to its
equivalent in two dimensions.
Theorem 1.5.1 (k and l determine curve up to a rigid motion). Let k(s) ≥ 0
and l(s) be any smooth functions. If we set the curvature scalar and torsion
scalar to these functions, respectively, we determine the curve uniquely in R3
up to a rigid motion in R3 .
As we said before, we won’t prove this. But we hope you see how interesting
the result it. We need only two real functions to describe a curve in three
38 CHAPTER 1. CURVES

dimensions, and only one real function in two dimensions. In both of these
cases, we reduced our description by an entire function.

1.5.5 The Fréchet-Frame


We want to talk a bit more about τ, N and β. We already said that these form a
right-hand-sided orthonormal moving frame, basically a coordinate system that
moves with the curve. Now, you can define arbitrary moving frames, but this one
is somewhat special, not least because both of the components are normalized
versions of geometrically important vectors (κ and λ). That is why this choice
holds a special name, it’s called the Fréchet-frame. (Sometimes Frenet–Serret
frame).
Before we talk about what makes this choice of moving frame special, we
want to talk a bit about general moving frames.
Let’s say we have some curve γ in three-dimensional space, and some moving
frame (not necessarily the Fréchet-frame) attached to it, that is orthonormal.
See figure 1.21.

Figure 1.21: A curve with its Fréchet-frame and a second picture of the same
curve with a random moving frame.
1.5. CURVES IN R3 39

We will abuse notation a bit and write the three basis vectors of the frame
e1 (s), e2 (s), e3 (s) as a vector like this:
 
e1 (s)
e2 (s) (1.55)
e3 (s)
Then, as we will show in a second, for a general moving frame we can write:
   
e (s) e1 (s)
d  1 
e2 (s) = A(s) e2 (s) (1.56)
ds
e3 (s) e3 (s)
where A(s) is a 3 × 3 matrix.
This matrix has a special property, generally, for any moving frame.

Proposition 1.5.1. Let γ be any curve and (e1 (s), e2 (s), e3 (s)) an orthonormal
moving frame. Then the matrix A(s) from equation 1.56 is anti-symmetric.

Where does the anti-symmetry come from? Well, there are two parts of the
anti-symmetry. Firstly, the diagonal is zero, which comes simply from the fact
that a vector of unit length cannot change in its own direction, otherwise the
length would change. Then there is the anti-symmetry of the other components,
which stems from the orthogonality. If one vector changes, the others have to
change in a way that they all stay orthogonal to each other.

Proof. We know that ⟨ei (s), ej (s)⟩ = δij , in-dependant of s. Therefore:

d dei (s) dej (s)


0= ⟨ei (s), ej (s)⟩ = ⟨ , ej (s)⟩ + ⟨ei (s), ⟩ = Aij + Aji (1.57)
ds ds ds
Aha, this is exactly the condition for anti-symmetry.

Now we will be able to see why the Fréchet-frame is special.

Theorem 1.5.2 (Fréchet-Frame-derivatives). For a Fréchet-Frame, we get:

    
τ 0 k 0 τ
d   
N = −k 0 l  N  (1.58)
ds
β 0 −l 0 β
where k = k(s), l = l(s) are the curvature and torsion scalars, respectively,
This is a particularly simple matrix. The special thing about a Fréchet-
frame is that it reduces the three independent matrix-components of A to two,
specifically two components we already know.

Proof. Many of these are definitions (for example, the first equation is just the
definition of N ), the rest aren’t too hard to check and we leave them to you as
an exercise.
40 CHAPTER 1. CURVES

1.5.6 Global theorems for curves in R3


We already introduced the principle that local and global quantities influence
each other, when we discussed the topic for curves in R3 . The same principle,
of course, applies to curves in R3 .
We will not fully prove any of the below, just state them. The first one will
be familiar from curves in two dimensions.
Theorem 1.5.3 (Fenchel’s theorem). Let γ be a closed curve in R3 (RN works
as well.) Then: Z
|k| ds ≥ 2π (1.59)
γ

The main steps in proving Fenchel’s theorem are the following:


(i) The curve τ has image in S 2 not contained in any open hemisphere. It is in
a closed hemisphere iff γ is a plane curve.
(ii) Any curve of length ≤ 2π in S 2 is contained in a closed hemisphere, and
any curve of lenght < 2π is strictly contained in an open hemisphere.

Proof. (see exercise sheet 1, nr 1c)


(i) Suppose the image of τ in contained in an open hemisphere. By rotating
R3 we can assume that the last coordinate of τ (s) is > 0. Then (as τ is the
derivative of γ) the last coordinate of γ(t) in R3 is strictly increasing in t, so
γ(0) = γ(L) is not possible, which contradicts γ being a closed curve. If the
last coordinate of τ (s) is only ≥ 0 for all s, then it actually must be = 0for all
s. Because if the last coordinate of τ (s) is < 0 for some s, there needs to be an
s̃ such that the last coordinate of τ (s) is < 0 to get a closed curve (we need to
lose height again). But if the last coordinate of τ (s) is 0 everywhere the curve
stays in the plane with last coordinate 0.
(ii) is left as an exercise to the reader.
The next theorem says something about the same quantity as above, for
knotted curves. Knotted curves are curves that cannot smoothly be transformed
into a circle, without crossing themselves. You can see a few examples in figure
1.22
Theorem 1.5.4. Millner’s theorem Let γ be any three-dimensional curve, that
is knotted, closed and simple. Then:
Z
|k| ds ≥ 4π (1.60)
γ

The bound in Millner’s theorem is sharp: for example the trefoil knot has
total absolute curvature of exactly 4π.

With these two examples, we conclude our discussion of curves and move
on to the second type of object we want to discuss before talking of differential
geometry in a general matter, those objects being surfaces in R3 .
1.5. CURVES IN R3 41

Figure 1.22: The trefoil, figure-8-knot and unknot. The first two are knotted,
the last one is not, even though it might look like it at first.
42 CHAPTER 1. CURVES
Chapter 2

Surfaces

In this chapter, we deal with surfaces, which are the obvious next step after
curves in our discussion. We will not treat surfaces in the more general RN , but
just surfaces in R3 , as they are more intuitive and are enough for the purposes
of this lecture.
We start with the the definition and then define a few basic properties, before
moving further to a discussion of the geometry and curvature.

2.1 Some definitions and basic quantities


The easiest type of surface in R3 is simply a graph of a function of x and y (Like
the one in figure 2.1(a)) We have a function f (x, y) and we set this to be the
z-coordinate. You can have a look at the specific example below.

Example 2.1.1 (The surface z = f (x, y) = (x2 + y 2 )). In figure 2.1(b), you
can see the graph of the function z = f (x, y) = (x2 + y 2 ). It is, of course, a
surface (in every intuitive way, but also in the more general definition we will
give below.)

It is clear that the surface in the example should be a surface. The idea that
a surface should always be the graph of a function is however, not a good one.
There are two simple examples that should definitely be surfaces, but wouldn’t
be, if that was our definition. The first is the xz-plane. Obviously, it should be
a surface. If anything should be a surface, the xz-plane should be. And yet, it
is quite easy to see that you can’t write it as a function of x and y. Well, you
might say, that there is nothing special about x and y in R3 and that we should
be allowed to choose any plane to describe our surface. For example, we could
take y = f (x, z) = 0 to describe the xz-plane. Yes, that is a possibility, but
still we find a problem. Take the sphere. It should definitely be included in our
definition of a surface. But I dare you to find any plane for which you can write
the whole sphere as a graph of a function. It should be very clear, that this is
not possible.

43
44 CHAPTER 2. SURFACES

Figure 2.1: A picture of many surfaces, that explain our definition of a surface.
In (a) you can a graph of some random function of x and y. The picture in
(b) is simply the graph of f (x, y) = x2 + y 2 . In (c) you see the first problem
with the naive definition, because we cannot write the xz-plane as the graph of
a function of x and y. In (d) you see the further problem, that even if we allow
for the function to be defined on an arbitrary plane, the sphere can simply never
be of sure form globally, but it can be made such locally (e).
2.1. SOME DEFINITIONS AND BASIC QUANTITIES 45

Can we repair this situation? Yes, if we recognize that, while we cannot


write surfaces like the sphere as a graph of a function, at every point on the
surface, in some (possibly very small) region of the surface, we can write that
region as the graph of a function (on some plane in R3 ). That is we can write the
surfaces locally as the graph of a function1 . This leads us to our full definition.
Definition 2.1.1 (Smooth surface in R3 ). A smooth surface M in R3 is a subset
of R3 , so that for every point, there is a surrounding of that point which can be
written as a smooth graph of some function in some direction.
Now that we have settled on what we mean by surfaces mathematically, we
need to be able to describe derivatives, since we are, after all, doing differential
geometry. You can imagine that you are a small ant, living on this surface, and
you, as an ant, can only move on the surface, never out of it2 . Then everything
you can see and feel is two dimensional, but the usual calculus we know in R3
is obviously three dimensional.We somehow need to reduce the dimensions. An
intuitive idea, is that smooth surfaces, locally (that might be very locally) look
like a plane, the same way that a smooth curve locally looks like a straight line.
We can take this plane, called the tangent plane, as the space of directions, and
do our differentiation there, because in the limit (of a very small area around
the point of interest), the surface and plane will converge to the same thing.
See figure 2.2.
We define the tangent plane a bit differently however, through curves that
live in M . This is mostly because it will be this definition that we will use to
define the tangent space when talking of manifolds generally, in later chapters.
Definition 2.1.2 (Tangent plane). The tangent plane to M at p is the set of
all vectors based at p and tangent to M at point p. We can formally define it
as:

Tp M = v ∈ R3 and v = γt (0) for γ smooth curve on M and γ(0) = p (2.1)




Apart from tangent vectors, we also have normal vectors. These are vectors
that are normal to M , which, of course, means that they are also normal to
Tp M
Definition 2.1.3 (Unit normal vector). The unit normal vector to M at p,
denoted N (p), is a normalised vector perpendicular to the tangent plane at p.
You can look at figure 2.3 for an example.
We have, of course, at every point, two choices of normal vectors. We can
either choose N to point in one, or the other direction. When we were discussing
curves, we also had this choice, and simply solved it by saying that we would
always take N in a way so that we get a right-handed coordinate system with
the other important vectors3 . Here, without any further important objects, we
1A function that depends on the point!
2 You are an ant. You cannot fly or jump or anything like that.
3 In 2d: τ , In 3d τ, β
46 CHAPTER 2. SURFACES

Figure 2.2: A surface M , with a point p and the tangent plane Tp M drawn in,
at different levels of ”zoom”. Globally, M and Tp M are very different, but the
more we zoom in, the more do the coincide, until they become ”almost” the
same.

Figure 2.3: A surface, it’s tangent plane, and a normal vector.


2.1. SOME DEFINITIONS AND BASIC QUANTITIES 47

cannot really do this. There isn’t an obvious choice of preferred directions in the
tangent plane, so that we could just choose N ’s orientation to get a right-hand
coordinate system. With closed surfaces (like a sphere), we can at least talk of
N pointing ”inwards” or ”outwards”, but for something that isn’t closed (like a
plane), we don’t even have this luxury. What we will do, is simply choose some
direction for N whenever we need it, and use that one consistently.
But we don’t need to only talk of a normal vector at a point, we can talk of
the normal vectors at all the points of the surface, see figure 2.4

Figure 2.4: The normal field N of some manifold M

Definition 2.1.4 (Unit normal field). The unit normal field of M is the function
N : M → R3 which maps points p in M onto the unit normal vector at p. We
require N to be continuous (or sometimes smooth).

Some remarks to the unit normal field:

• We have two possible choices for a unit normal vector at a point (”inwards”
or ”outwards”). Since we require N to be continuous it must be in the
same direction on a connected set, so overall we get a total of 2k possible
unit normal fields, where k is the number of connected components in M .

• In some cases we can only define N locally. A classical example being the
Möbius-strip4 , which you can see in figure 2.5.

• If M is the boundary of an open set in R3 , then N can be defined globally.

Now that we have developed the most basic tools we could use to do dif-
ferential geometry on surfaces (The analogues of τ and N for curves) we can
proceed to talk about curvature.
4 You can easily build a Möbius-strip, from a strip of paper. You just need to tape the ends

of the strip of paper together, but in a way that the strip turns once. If you never did this,
you should try.
48 CHAPTER 2. SURFACES

Figure 2.5: The Möbius strip. You can try to draw a normal field by starting
at some point and drawing the next normal vector and then the next until you
get back to where you started, but you will find that when you come back, your
normal will point into the other direction than the one you started with (= not
smooth).

2.2 The curvature of surfaces in R3


We will find that the the curvature of a surface at a point will be described by
an object called the second fundamental form5 . There are three equivalent ways
to define the curvature of surface, which are summarized below.

A Through the curvature of curves through p (geometric definition, QP )

B By using the Hessian of M , regarded as a graph over its own tangent plane
Tp M (2nd fundamental form, Ap (x, y))

C By defining the Weingarten map Wp

Furthermore, we will prove two useful formulas for explicitly computing curva-
ture:

D Formula via appropriate parametrisation

E Formula for a graph of a function

We shall start with the geometric definition of curvature, since it is the most
intuitive one.
5 You might wonder whether there is a first fundamental form. Yes there is, but for peda-

gogical reasons, we will introduce it a bit later.


2.3. THE GEOMETRIC DEFINITION OF CURVATURE ON SURFACES49

2.3 The Geometric Definition of Curvature on


Surfaces
The first way to define curvature on a surface is using curves that lie on the
surface. If the surface is curved, then any curve passing through the point of
interest p has to be curved as well. We cannot just take the curvature of some
curve on M going through p, simply because this will give us different values for
different curves. (There is an infinite amount of curves going through p that live
on M , all curved differently, see figure 2.6) There is however a specific amount
of curving that a curve has to curve at point p to still stay on M , otherwise it
would leave. This amount of curvature, will be in the direction of N . Why?
Well, in the tangent direction, we can have pretty much any curvature we want,
the curve can be as curved as it want on M . That pretty much leaves on the
direction of N . From this simple thought follows the next definition.

Figure 2.6: A surface M and a point p. There is a lot of curves going through
p, some less, some more curved. (They all have to live on M though.)

Definition 2.3.1 (Geometric definition of curvature). Let M be a surface and


N the unit normal vector field. Let p ∈ M , v ∈ Tp M with |v| = 1, γ(t) any
curve in M with some parametrisation t, such that γ(0) = p, γt (0) = v. Then
define:
QN
p (v) = Qp (v) = ⟨κγ (0), N (p)⟩

What we basically do is define a function that gives us the normal component


of the curvature of a curve going through p on M , with its tangent vector
matching with v. See figure 2.7
50 CHAPTER 2. SURFACES

Figure 2.7: The idea of the previous definition. The function Qp takes a (nor-
malized) v from the tangent plane, takes a curve through p with a matching
tangent vector, and puts out the normal component of κγ .

There are two questions you might have already asked yourself. Firstly,
why do we only define this when |v| = 1? That is an easy question to answer.
Mostly convenience. It will simplify the proof and calculations, while not leaving
out any real information. Let’s say you want to know Qp (v) for some vector
with length 2. Then you can do this entire procedure for the same vector, but
normalized, and simply parameterize the curve γ in such a way that you move
twice as fast along it.
The second question is whether this definition makes any sense, whether it
is well-defined, because we have, after all, many curves going through p with
their tangent-vector equaling v and it is not obvious that we always get the
same number for Qp (v) for any choice of appropriate curve. This will turn out
to be true however, as we will prove in a short time. For now, we will quickly
assume that this is true and talk about the object Qp a bit.

2.3.1 A bit about Qp


We want to take a few notes and intuitive ideas about what Qp is and describes
at this point.

• Firstly, Qp (v) is not standard notation, because it will turn out that Qp
will simply be the second fundamental form in a slightly different manner,
which will be denoted as Ap (X, Y ), and which we will introduce shortly.

• Qp (v) will turn out to be a quadratic form in the components of v, i.e it


2.3. THE GEOMETRIC DEFINITION OF CURVATURE ON SURFACES51

P2
will be of the form Qp (v) = i,j=1 aij v i v j , where aij are some numbers
that come from the surface’s geometry and v i is the i-the component6 of
v.

• Qp (v) = Qp (−v). This follows from the fact that for an appropriate curve
which we use to calculate Qp (v), we can simply take the same curve, but
reverse it’s parametrization. Then the tangent-vector changes direction (or
sign), but the curvature vector stays the same, and therefore by definition
Qp (v) does too7 .

• Qp (v) can be either positive or negative, there are no general restrictions


on it’s sign. The signs hold geometric information however. Also, if we
use −N as our normal field, the sign of Qp (v) flips, simply because N is
in the scalar product.

• The largest and smallest values of Qp belong to two normalized vectors


e1 , e2 that are orthogonal to each other. To know these two directions
and their Qp values is also enough to specify Qp (v) for all v in Tp M with
unit length. The values of these are most often denoted kmax = k1 and
kmin = k2 .The quantities k1 and k2 are called the principle curvatures at p,
while the two directions e1 and e2 are called the principle directions/axes
of curvature.

We will prove many of these claims soon, for now we just wanted to familiarize
you with these properties.

Example 2.3.1 (A sphere of radius R). As always, a sphere is a good way


to find an intuitive understanding of what an object like Qp (v) describes. Let
us assume that for all v in Tp M of unit length it really is enough to calculate
Qp (v) for a single curve and that this number will be the same for all other
appropriate curves. Then we can calculate Qp (v) very easily. Because the
sphere is very symmetric, we can get away with just calculating Qp for the
north pole, and expect it to be the same (geometrically) at all other points8 .
The coordinates (Cartesian) of the north-pole are (0, 0, R) and the normal vector
(we choose the inwards normal vector) is N = (0, 0, −1). Let’s say we are given
a v = (v 1 , v 2 , 0) to calculate Qp (v) of. For this v, we can choose the curve to be
the appropriately parameterized greater circle lying in the plane spanned by N
and v. The curvature vector of this greater circle at p must, of course, be R1 N ,
so that
1 1
Qp (v) = ⟨κγ , N ⟩ = ⟨ N, N ⟩ = (2.2)
R R
6 Notice the index is upstairs. There is a reason for this sort of notation, where upstairs and

downstairs indexes have different meanings. For now, however, there really is no difference,
but we will use this notation to get you accustomed.
7 Prove that the tangent-vector changes sign, but the curvature doesn’t.
8 You can alternatively see this as us using special (Cartesian) coordinates, for which the

point on the sphere that is of interest p has coordinates (0, 0, R). Then, since Qp is defined by
a scalar product of two fundamentally geometric vectors, it should transform geometrically.
52 CHAPTER 2. SURFACES

Figure 2.8: A sphere, the north pole, a random unit tangent vector and a great
circle.

Also, this is independent of v, which really just says something about the (local)
symmetry of the sphere.

Example 2.3.2 (Independence of Qp from the curve on S 2 ). We have calculated


Qp for the sphere. We will do the same thing again, but with a different curve
and observe how it is that we get the same result. We have the same situation
as in the last example, but this time we pick a different curve, that is not a
greater circle. Instead, let’s take a circle on the sphere that passes through p
and has the tangent vector v, but is smaller than a greater circle. See figure 2.9.
Because the circle is smaller, κβ is a lot larger than κγ from the last example.
But κβ also points in a different direction, not in the direction of the normal
vector of the sphere at the north pole. When we project9 κβ back onto N , we
get the same result (prove this). In this way, the projection onto N ensures that
we get the same result, no matter the curve.

2.3.2 The independence of Qp (v) from the curve we choose


Now that we have a bit more intuition/information about Qp (v), we want to
prove that it is independent of our choice of curve, so long as the curve lies on
M and has the tangent vector v.

Theorem 2.3.1 (Qp is well-defined). Let M be a surface, p some point on


the surface and v some unit vector in Tp M . Then Qp (v) = ⟨κγ (0), N (p)⟩ is
independent of our choice of the appropriate curve γ. (Appropriate means that
γ lies on M , γ(0) = p and that the tangent vector γt = v.)
9 Because we are calculating a scalar product
2.3. THE GEOMETRIC DEFINITION OF CURVATURE ON SURFACES53

Figure 2.9: Here we have the same situation as in the previous figure, but instead
of a greater circle, we take a smaller circle to compute Qp (v). We have however
rotated the sphere (v is pointing towards you), so that it would be simpler to
see what is going on, namely that the curvature vector of the smaller curve is
far larger, but that the projection onto N is still the same.

To prove this theorem, we will need the lemma below.


Lemma 2.3.1. Let everything be the same as in the theorem above. Then

⟨κγ (0), N (p)⟩ = ⟨γtt (0), N (p)⟩ (2.3)

What the lemma says, is basically that instead of the geometric curvature κ
of the curve, we can instead use the second derivative of γ.
The proof is quite simple.
Proof. Let γ be any appropriate
 curve. We simply compute ⟨κγ (0), N (p)⟩, using
1 γt
the fact that κγ = |γ |2 γtt − ⟨γtt , γt ⟩ |γ |2 . We can use the fact that |γt | =
t t
|v| = 1, which simplifies the formula for κ

⟨κγ (0), N (p)⟩ = ⟨γtt − ⟨γtt , γt ⟩γt , N (p)⟩ = ⟨γtt (0), N (p)⟩ (2.4)

where in the last step we have used the geometric fact that γt = v is in the
tangent plane, and therefore normal to N .

Curvature is a second derivative


Before we go on to prove the theorem, we will discuss a big idea the proof
relies on, which is an overarching theme in differential geometry. It’s the idea
that curvature is a second derivative. This is something that is fully universal
throughout differential geometry. This fact however, shows up only if you either
54 CHAPTER 2. SURFACES

use some smart ”natural” coordinates or differentiate with respect to a geometric


quantity. In general coordinates, it is usually an equation containing a second
derivative but with some other non-geometric ”decorations”.
For a curve, we defined the curvature vector to be:

d2 γ
κ= (2.5)
ds2
It is manifestly a second derivative, and because we are differentiating after
a geometric quantity (s), it is simply just the second derivative. Later, while
discussing curves, we saw that, if you can write the curve as a graph of a function
y = u(x), then the curvature becomes:
uxx
k= (2.6)
(1 + u2x )3/2

At any point whose tangent is parallel to the x-axis, ux = 0 and:

k = uxx (2.7)

It is clear that we can also (locally) write the curve as a graph over it’s tangent,
as seen in figure 2.10, where we also find that the curvature scalar becomes a
second derivative.
We will see that the curvature of a surface will also become a second deriva-
tive, when you write the surface as the graph of a function over its tangent plane
(locally).

2.3.3 Proof of the theorem


We will now, finally, get to the proof.

Proof.

Step 1 We start, by choosing an orthonormal basis of R3 , so that p = (0, 0, 0)


and Tp M is the xy-plane, and N = (0, 0, 1). This is mostly a matter
of convenience, and certainly doesn’t change ⟨κγ (0), N (p)⟩. Then we can
write the surface (locally) as the graph of a function where z = f (x, y) for
a function given by f . For a picture see figure 2.11

Step 2 Now let v ∈ Tp M . Then it will be of the form v = (v 1 , v 2 , 0) ∼ (v 1 , v 2 )


(We associate the xy-plane in R3 with R2 in the usual way)

Step 3 We now let γ be any curve on M with γ(0) = p and γt 0 = v and compute
⟨κγ (0), N (p)⟩. To do this, we notice that γ(t) = (x(t), y(t), z(t)), and
γtt (t) = (x′′ (t), y ′′ (t), z ′′ (t)), obviously. Then:

⟨κγ (0), N (p)⟩ = ⟨γtt (0), N ⟩ = ⟨(x′′ (0), y ′′ (0), z ′′ (0)), (0, 0, 1)⟩ = z ′′ (0)
(2.8)
2.3. THE GEOMETRIC DEFINITION OF CURVATURE ON SURFACES55

Figure 2.10: A curve that is just the graph of the function u(x). When written
as the graph of a function over its tangent, k becomes the second derivative.
56 CHAPTER 2. SURFACES

Figure 2.11: The coordinates we are using for the proof of the theorem. The
point p is at the origin and the tangent plane is the xy-plane

where we used the lemma from above. We found the theme of the last
subsection. Curvature is a second derivative. What we have basically
done is write (a local part of) the surface as a graph over its tangent plane
at p, and found that, similarly to curves, curvature became the second
derivative in those coordinates.
Step 4 Somehow, we need to use the fact that γ lies on the surface, which we have
not yet used. Of course, because γ does lie on the surface, its coordinates
have to be: γ(t) = (x(t), y(t), f (x(t), y(t))), because the z component (for
a local part, again) is not independent of x and y. If we now evaluate
z ′′ (0) we get:

zt t = fx (x(t), y(t))xt (t) + fy (x(t), y(t))yt (t) (2.9)


ztt (t) = fxx xt xt + fxy xt yt + fx xtt + fyx xt yt + fyy yt yT + fy ytt (2.10)

But, we only need it at t = 0, where it simplifies a lot. Firstly, xt =


v 1 , yt = v 2 because we are using a curve whose tangent vector is v. But
also fx (0) = fx (x(0), y(0) has to be 0, because the xy-plane is Tp M is
tangent to the surface, which we describe with the function f . So what
we get is:

ztt (0) = fxx v 1 v 1 + 2fxy v 1 v 2 + fyy v 2 v 2 (2.11)

Firstly, we have just proven that Qp (v) is a quadratic form, but we have
also proven that Qp (v) is independent of the curve we choose. The quan-
tities fxx , fxy and fyy have nothing to do with our choice of curve, only
with the underlying surface.
2.4. THE SECOND FUNDAMENTAL FORM 57

We can rewrite our result quite a bit and see another way of looking at it.

Qp (v) = fxx v 1 v 1 + 2fxy v 1 v 2 + fyy v 2 v 2 (2.12)


  
 fxx fxy v1
= v1 v2 = v T Hv (2.13)
fyx fyy v2

where H is the Hessian-matrix of f evaluated at the point p. If we write x =


x1 , y = y 1 , z = z 1 we can write it in one sum:
2
X ∂f
Qp (v) = vi vj (2.14)
i,j
∂xi ∂xj

which makes the quadratic-form-ness of Qp a bit more visible. (The partials are
of course evaluated at p)

2.4 The second fundamental form


Now that we have seen the geometric definition, we can manipulate it a tiny bit
and define the second fundamental form.
Definition 2.4.1. Let M and p be, as always, a surface and a point on it.
Let us also use coordinates where locally we write the surface as the graph of a
function f over its tangent Tp M at p. The the second fundamental is defined
as:
2
X ∂f
Ap (X, Y )) = i ∂xj
X iY j (2.15)
i,j
∂x

which is a function Ap : Tp M × Tp M → R, or in other words the vectors X, Y


are both in the tangent space.
We can very quickly note that the second fundamental form is symmet-
ric, in the sense that Ap (X, Y ) = Ap (Y, X) simply because the hessian is a
symmetric matrix. You can also see the second fundamental form as the first
Taylor-coefficient of f , which actually tells us anything geometric. The zero-th
coefficient just tells us where the surface is, the first only how the tangent plane
is oriented, both of which are not geometric things. Only the second one tells
us anything geometric, that is, the curvature at the point p.
Note 5. The formula from above can only work if we use the coordinates
where p = (0, 0, 0) and the xy plane is the tangent-space, because there fx
and fy vanish. We need to carefully pick this coordinate system to extract the
geometric information. There is a general formula in any coordinates, but it
long and tedious, which is the reason why we won’t do it here.

You might also like