Diff Geometry
Diff Geometry
f (x, y) = c,
1
f
2 Dr. V. V. Acharya
4.
2.
3.
1.
2.
−2. −1. 0 1. 2. c
1.
−1.
y − 2x = 1 y − x2 = 0
2.
e
1.
1.
d −3. −2. −1. 0 1. 2. 3.
−1.
−1. 0 1.
−1. −2.
x2 y 2
x2 + y 2 = 1 + = 1.
9 4
f1 (x, y, z) = c1 , f2 (x, y, z) = c2 .
Curves of this kind are called level curves, the idea being that the curve in
Differential Geometry 3
Eq. 1.1, for example, is the set of points (x, y) in the plane at which the
quantity f (x, y) reaches the level c.
But there is another way to think about curves which turns out to be more
useful in many situations. For this, a curve is viewed as the path traced out
by a moving point. Thus, if γ(t) is the position of the point at time t, the
curve is described by a function γ of a scalar parameter t with vector values
(in R2 for a plane curve, in R3 for a curve in space). We use this idea to give
our first formal definition of a curve in Rn (we shall be interested only in the
cases n = 2 or 3, but it is convenient to treat both cases simultaneously).
for all values of t in the interval (α, β) where γ is defined (yet to be decided),
and ideally every point on the parabola should be equal to (γ 1 (t), γ 2 (t)) for
some value of t ∈ (α, β).
Of course, there is an obvious solution to Eq. 1.2: take γ 1 (t) = t, γ 2 (t) =
t2 . To get every point on the parabola we must allow t to take every real
number value (since the x-coordinate of γ(t) is just t, and the x-coordinate
of a point on the parabola can be any real number), so we must take (α, β)
to be (−∞, ∞). Thus, the desired parametrization is
But this is not the only parametrization of the parabola. Another choice is
γ(t) = (t3 , t6 ) (with (α, β) = (−∞, ∞). ). Yet another is (2t, 4t2 ), and of
course there are (infinitely many) others. So the parametrization of a given
level curve is not unique.
4 Dr. V. V. Acharya
Example 1.4 A cycloid is the plane curve traced out by a point on the
circumference of a circle as it rolls without slipping along a straight line.
Show that, if the straight line is the x-axis and the circle has radius a > 0,
the cycloid can be parametrised as γ(t) = a(t − sin t, 1 − cos t).
Solution. When the circle has rotated through an angle t, its centre has
moved to (at, a), so the point on the circle initially at the origin is now at
the point (a(t − sin t), a(1 − cos t)).
Generalise this by finding parametrisations of an epicycloid (resp.
hypocycloid), the curve traced out by a point on the circumference
of a circle as it rolls without slipping around the outside (resp.
inside) of a fixed circle.
Differential Geometry 5
Let the fixed circle have radius a, and the moving circle radius b (so that
b < a in the case of the hypocycloid), and let the point P of the moving
circle be initially in contact with the fixed circle at (a, 0). When the moving
circle has rotated through an angle ϕ the line joining the origin to the point
of contact of the circles makes an angle θ with the positive x-axis, where
aθ = bϕ. The point P is then at the point
and g(t) are smooth functions, it follows from standard results of calculus
that the sum f (t) + g(t), product f (t)g(t), quotient f (t)/g(t), and composite
f (g(t)) are smooth functions, where they are defined.
To differentiate a vector-valued function such as γ(t), we differentiate
componentwise: if γ(t) = (γ 1 (t), γ 2 (t), . . . , γ n (t)), then
2 2
d γ 1 d2 γ 2 d2 γ n
dγ dγ 1 dγ 2 dγ n dγ
= , ,..., , 2 = , ,..., , etc.
dt dt dt dt dt dt2 dt2 dt2
dγ d2 γ
To save space, we often denote by γ̇(t), 2 by γ̈(t), etc. We say that
dt dt
dn γ
γ is smooth if the derivatives n exist for all n ≥ 1 and all t ∈ (α, β); this
dt
is equivalent to requiring that each of the components γ 1 , γ 2 , . . . , γ n of γ is
smooth. From now on, all parametrized curves studied will be assumed to
be smooth.
Definition 1.2 If γ is a parametrized curve, its first derivative γ̇(t) is called
the tangent vector of γ at the point γ(t).2
γ(t + δt) − γ(t)
To see the reason for this terminology, note that the vector
δt
is parallel to the chord joining the points γ(t) and γ(t + δt) of the image
C of γ. As δt tends to zero the length of the chord also tends to zero, but
we expect that the direction of the chord becomes parallel to that of the
tangent to C at γ(t). But the direction of the chord is the same as that of the
γ(t + δt) − γ(t)
vector which tends to dγ/dt as δt tends to zero. Of course,
δt
this only determines a well defined direction tangent to the curve if dγ/dt is
non-zero. If that condition holds, we define the tangent line to C at a point
p of C to be the straight line passing through p and parallel to the vector
dγ/dt.
Theorem 1.1 If the tangent vector of a parametrized curve is constant, the
image of the curve is (part of) a straight line.
Proof. If γ̇(t) = a for all t, where a is a constant vector, we have, integrating
component-wise,
Z Z
dγ
γ(t) = dt = adt = ta + b,
dt
2
at the parameter value t.
Differential Geometry 7
Example 1.6 For a logarithmic spiral γ(t) = (et cos t, et sin t), show that
the angle between γ(t) and the tangent vector at γ(t) is independent of t.
Solution: Note that γ̇(t) = (et (cos t − sin t), et (sin t + cos t)), so the angle
between γ(t) and γ̇(t) is given by
Suppose the point O is the origin, and that M is on the positive y-axis.
Suppose the radius of the circle is a. Then the curve has Cartesian equation
8a3
y= .
x2 + 4a2
Note that if a = 1/2, then this equation becomes rather simple:
1
y= .
x2 +1
Parametrically, if θ is the angle between OM and OA, measured clock-
wise, then the curve is defined by the equations
x = 2a tan θ, y = 2a cos2 θ.
Another parametrization, with being the angle between OA and the x-axis,
increasing anti-clockwise is
x = 2a cot θ, y = 2a sin2 θ.
Differential Geometry 9
1.5
0.5
-0.5
-1
-1.5
(see the diagram below). Note that γ has a self-intersection at the origin in
the sense that γ(t) = 0 for t = 2π/3 and for t = 4π/3. The tangent vector is
This example shows that we must be careful while talking about a ‘point’
of a parametrized curve γ : strictly speaking, this should be the same thing as
a value of the curve parameter t,and not the corresponding geometric point
γ(t) ∈ Rn .
1.2 Arc-length
We recall that, if v = (v1 , . . . , vn ) is a vector in Rn , its length is
q
kvk = v12 + · · · + vn2 .
10 Dr. V. V. Acharya
Example 1.10 For a logarithmic spiral γ(t) = (et cos t, et sin t) we have
Hence,
kγ̇(t)k2 = e2t (cos t − sin t)2 + (sin t − cos t)2 = 2e2t .
d t
Z
ds
= kγ̇(u)k du = kγ̇(t)k (1.5)
dt dt t0
ds
Thinking of γ(t) as the position of a moving point at time t, is the
dt
speed of the point (rate of change of distance along the curve). For this
reason, we make
n(t) · ṅ(t) = 0
d db da
(a(t) · b(t)) = a(t) · + · b(t).
dt dt dt
Using this to differentiate both sides of the equation n(t) · n(t) = 1 with
respect to t, gives
1.3 Reparametrisation
Definition 1.5 A parametrised curve γ̃ : (α̃, β̃) → Rn is a reparametrisation
of a parametrised curve γ : (α, β) → Rn if there is a smooth bijective map
φ : (α̃, β̃) → (α, β) (the reparametrisation map) such that the inverse map
φ−1 : (α, β) → (α̃, β̃) is also smooth and
Example 1.11 Note that γ(t) = (cos t, sin t) and γ̃(t) = (sin t, cos t) are
parametrizations of the unit circle x2 + y 2 = 1.
To see that γ is a reparametrisation of γ̃, we have to find a reparametri-
sation map φ such that (cos φ(t), sin φ(t)) = (sin t, cos t). One solution is
φ(t) = π2 − t.
ds
= f (u̇2 + v̇ 2 ). (1.6)
dt
The crucial point is that f is smooth on R2 \ {(0, 0}, which means that all
the partial derivatives of f of all orders exist and are continuous functions
except at the origin (0, 0). For example,
∂f u ∂f v
=√ , =√ ,
∂u u + v ∂v
2 2 u + v2
2
are well defined and continuous except where u = v = 0, and similarly for
higher derivatives. Since γ is regular, u̇ and v̇ are never both zero, so the
chain rule and Eq. (1.6) shows that ds/dt is smooth.
dγ̃ dγ
Since γ̃ is unit-speed, = 1, so clearly cannot be zero.
dt̃ dt
dγ
Conversely, suppose that the tangent vector is never zero. Note that
dt
ds dγ
= > 0 for all t, where s is the arc·length of γ starting at any point of
dt dt
the curve, and by Proposition 1.4 s is a smooth function of t. It follows from
the inverse function theorem of multivariable calculus that s : (α, β) → R
is injective, that its image is an open interval (α̃, β̃), and that the inverse
map s−1 : (α̃, β̃) → (α, β) is smooth. We take φ = s−1 and let γ̃ be the
corresponding reparametrization of γ, so that γ̃(s) = γ(t). Then,
dγ̃ ds dγ dγ̃ ds dγ ds
= . Hence, = = .
ds dt dt ds dt dt dt
dγ̃
Thus, = 1.
ds
Example 1.12 Show that for the parabola y = x2 , γ 1 (t) = (t, t2 ) and
γ 2 (t) = (t3 , t6 ) are parametrisations. Further, show that γ 2 (t) is not a
reparametrisation of γ 1 (t).
Differential Geometry 15
(a) γ(t) = (t3 , t2 ), t ∈ R, has a weak tangent but not a strong tangent
at t = 0.
(b) If γ : I → R3 is of class C 1 and regular at t = t0 , then it has a
strong tangent at t = t0 .
(c) The curve given by γ(t) = (t2 , t|t|), t ∈ R is of class C 1 but not of
class C 2 . Draw a sketch of the curve and its tangent vectors.
8. Let γ : [a, b] → R3 be a differentiable curve, where [a, b] is a closed and
bounded interval. For every partition
a = t0 < t1 < · · · < tn = b
of [a, b], consider the sum ni=1 kγ(ti ) − γ(ti−1 )k = l(γ, P ), where P
P
stands for the given partition. The norm kP k ( of a partition P is
defined as
kP k = max(ti − ti−1 ), i = 1, . . . , n.
Geometrically, l(γ, P ) is the length of a polygon inscribed in γ([a, b])
with vertices in γ(ti) (see Fig. Ex1).
The point of the exercise is to show that the arc length of γ([a, b]) is,
in some sense, a limit of lengths of inscribed poIygons.
Figure Ex 1
Prove that given >0 there exists δ > 0 such that if kP k < δ then
Z b
kγ̇(t)k − l(γ, P ) < .
a
Chapter 2
How Much Does a Curve Curve?
In this chapter, we associate to any curve in R3 two scalar functions,
called its curvature and torsion. The curvature measures the extent to which
a curve is not contained in a straight line (so that straight lines have zero
curvature), and the torsion measures the extent to which a curve is not
contained in a plane (so that plane curves have zero torsion). It turns out
that the curvature and torsion together determine the shape of a curve.
2.1 Curvature
We want to find a measure of how ‘curved’ a curve is. Since this ‘curvature’
should depend only on the ‘shape’ of the curve:
(ii) the curvature of a straight line should be zero, and large circles should
have smaller curvature than small circles.
17
18 Dr. V. V. Acharya
The first part of condition (ii) will certainly be satisfied. As to the second
part, consider the circle centred at (x0 , y0 ) and of radius R. This has a unit-
speed parametrisation
s s
γ(s) = (x0 + R cos , y0 + R sin ).
R R
We have γ̇(s) = (− sin Rs , cos Rs ), Hence,
kγ̇(s)k
s = 1, showing s that γ
is indeed unit-speed, and hence γ̈(s) = − R1 cos , − R1 sin . so the
R R
curvature of the circle is inversely proportional to its radius. As to condition
(i), recall from Corollary 1.1 that, if u = ±s + c, and c is a constant. Then,
by the chain rule,
dγ dγ du dγ
= =± .
ds du ds du
Hence,
d2 γ d2 γ
d dγ
= ± = .
ds2 du du du2
This shows that the curvature of the curve computed using the unit-speed
parameter s is the same as that computed using the unit-speed parameter u.
But what if we are given a curve γ(t) that is not unit-speed? If γ is regular
then it has a unit-speed reparametrisation γ̃. We define the curvature of γ
to be that of the unit-speed curve γ̃. But since it is not always possible to
find the unit-speed reparametrisation explicitly, we really need a formula for
the curvature in terms of γ and t only.
kγ̈ × γ̇k
κ= (2.1)
kγ̇k3
where the × indicates the vector (or cross) product and the dot denotes d/dt.
Proof. Let γ̃ (with parameter s) be a unit-speed reparametrisation of γ,
and let us denote d/ds by a dash. Then, by the chain rule,
dγ̃ ds dγ
= = γ̇, so
ds dt dt
! 2
d2 γ̃ d γ̇ γ̈ ds
dt
− γ̇ ddt2s
= = .
ds2 ds ds ds 3
dt dt
Differential Geometry 19
2
ds
Now, = kγ̇k2 = γ̇ · γ̇, and differentiating with respect to t gives
dt
ds d2 s
· = γ̇ · γ̈. Thus,
dt dt2
ds 2 d2 s
− γ̇ ds
d2 γ̃ γ̈ dt dt dt2 γ̈(γ̇ · γ̇) − γ̇(γ̇ · γ̈)
= = .
ds2 ds 4 kγ̇k4
dt
(where a, b, c ∈ R3 ), we get
Proposition 2.2 Let γ(s) be a unit-speed plane curve, and let ϕ(s)1 be the
angle through which a fixed unit vector must be rotated anti-clockwise to
bring it into coincidence with the unit tangent vector t of γ.2 Then,
dϕ
κs = . (2.4)
ds
Proof. Let a be the fixed unit vector and let b be the unit vector obtained
by rotating a anti-clockwise by π/2. Then, t = a cos ϕ + b sin ϕ.
dϕ dϕ
Hence, ṫ = (−a sin ϕ + b cos ϕ) . Hence, ṫ · a = − sin ϕ . Note that
ds ds
ṫ = κs ns . Hence,
dϕ
ṫ · a = κs (ns · a) = − sin ϕ . (2.5)
ds
But the angle between ns and a is ϕ + π/2, since t must be rotated anti-
clockwise by π/2 to bring it into coincidence with ns
Hence, ns · a = cos(ϕ + π/2) = − sin ϕ. Inserting this into (2.5) gives the
required result.
The next result shows that a unit-speed plane curve is essentially deter-
mined once we know its signed curvature at each point of the curve. The
meaning of ‘essentially’ here is ‘up to a rigid motion of R2 . Recall that a rigid
motion of R2 is a map M : R2 → R2 of the form
M = Ta ◦ Rθ ,
1
Note that, even though the angle ϕ is only determined up to adding an integer multiple
of 2π, the derivative dϕ
ds is well defined.
2
Thus, the signed curvature is the rate at which the tangent vector of the curve rotates.
Differential Geometry 21
Proof. For the first part, fix s0 ∈ (α, β) and define, for any s ∈ (α, β),
Z s
ϕ(s) = κ(u)du, (cf.P roposition2.2),
s0
Z s Z s
γ(s) = cos ϕ(t)dt, sin ϕ(t)dt .
s0 s0
Then, the tangent vector of γ(s) is γ̇(s) = (cos ϕ(s), sin ϕ(s)) which is a
unit vector making an angle ϕ(s) with the x-axis. Thus, γ is unit-speed and,
by Proposition 2.2, its signed curvature is
d s
Z
dϕ
= κ(u)du = κ(s).
ds ds s0
For the second part, let ϕ̃(s) be the angle between the x·axis and the unit
tangent vector γ̃˙ of γ̃. Thus,
˙
γ̃(s) = (cos ϕ̃(s), sin ϕ̃(s)),
Z s Z s
γ̃(s) = cos ϕ̃(t)dt, sin ϕ̃(t)dt + γ̃(s0 ). (2.6)
s0 s0
By Proposition 2.2,
dϕ̃
= κ(s).
ds
22 Dr. V. V. Acharya
Rs
ϕ̃(s) = s0 κ(u)du + ϕ̃(s0 ). Inserting this into Eq. (6), and writing a for the
constant vector γ̃(s0 ) and θ for the constant scalar ϕ̃(s0 ), we get
Z s Z s
γ̃(s) = Ta ( cos(ϕ̃(t) + θ)dt, sin(ϕ̃(t) + θ)dt)
s0 s0
Z s Z s
= Ta (cos θ cos ϕ̃(t)dt − sin θ sin ϕ̃(t)dt,
s0 s0
Z s Z s
sin θ cos ϕ̃(t)dt + cos θ sin ϕ̃(t)dt))
s0 s0
Z s Z s
= Ta Rθ ( cos ϕ̃(t)dt, sin ϕ̃(t)dt)
s0 s0
= Ta Rθ (γ(s)) .
Proposition 2.3 Any regular plane curve whose curvature is a positive con-
stant is part of a circle.
Proof. Let κ be the (constant) curvature of the curve γ, and let κs be its
signed curvature. Then, κs = ±κ.
A priori, we could have κs = κ at some points of the curve and κs = −κ.
at others, but in fact this cannot happen since κs is a continuous function of
s, so the Intermediate Value Theorem tells us that, if κs takes both the value
κ and the value −κ, it must take all values between. Thus, either κs = κ at
all points of the curve, or κs = −κ at all points of the curve. In particular,
κs is constant.
The idea now is to show that, whatever the value of κs , we can find a
parametrised circle whose signed curvature is κs . The theorem then tells us
that any curve whose signed curvature is κs can be obtained by applying a
rigid motion to this circle. Since rotations and translations obviously take
circles to circles, it follows that any curve whose signed curvature is constant
is (part of) a circle.
2.3 Curves in R3
Let γ be a unit speed curve. Then t = γ̇. We assume that ṫ = γ̈ 6= 0 for all
1 1
t. Then we define κ = ṫ = kγ̈k and n = ṫ = γ̈. Thus,
κ κ
dt
= κn. (2.7)
ds
Differential Geometry 23
b = t × n, n = b × t, t=n × b.
db dt dn dn
Hence, = ×n+t× = t× . Note that b is a unit vector.
ds ds ds ds
db db
Hence, is orthogonal to b as well as t. Hence, is parallel to n. We
ds ds
define
db
= −τ n. (2.8)
ds
Note that n = b × t. Differentiating w.r.t. s, we get
dn db dt
= ×t+b× = −τ n × t + b × κn = −κt + τ b. (2.9)
ds ds ds
Theorem 2.2 Let γ(t) be a regular curve in R3 with nowhere vanishing
d
curvature. Then, denoting by a dot, its torsion is given by
dt
...
(γ̇ × γ̈) · γ
τ= (2.10)
kγ̇ × γ̈k2
In the general case, let s be arc-length along γ and denote d/ds by a dash.
Then
ds 0
γ̇ = γ
dt
2
ds 00 d2 s 0
γ̈ = γ + 2γ
dt dt
3
... ds ds d2 s 00 d3 s 0
γ = γ 000 + 3 γ + 3γ .
dt dt dt2 dt
3
ds
Hence, γ 0 × γ 00 = (γ̇ × γ̈)
dt
6
... 0 ds
γ (γ × γ 00 ) = γ 000 × (γ̇ × γ̈)
dt
and so ...
(γ̇ × γ̈) · γ (γ 0 × γ 00 ) · γ 000
= .
kγ̇ × γ̈k2 kγ 0 × γ 00 k2
Theorem 2.3 Let γ(s) and γ̃(s) be two unit-speed curves in R3 with the
same curvature κ(s) > 0 and the same torsion τ (s) for all s. Then, there is
a rigid motion M of R3 such that
γ̃(s) = M (γ(s)) for all s.
Further, if κ and τ are smooth functions with κ > 0 everywhere, there is a
unit-speed curve in R3 whose curvature is κ and whose torsion is τ.
Proof. Let t, n and b be the tangent vector, principal normal and bi-
normal of γ and let t̃, ñ and b̃ be those of γ̃. Let s0 be a fixed value of
the parameter s. Since {t(s0 ), n(s0 ), b(s0 )} and {t̃(s0 ), ñ(s0 ), b̃(s0 )} are both
right-handed orthonormal bases of R3 , there is a rotation about the origin
of R3 that takes t(s0 ), n(s0 ) and b(s0 ) to t̃(s0 ), ñ(s0 ) and b̃(s0 ) respectively.
Further, there is a translation that takes γ(s0 ) to γ̃(s0 ) (and this has no
effect on t, n and b). By applying the rotation followed by the translation,
we can therefore assume that
γ(s0 ) = γ̃(s0 ), t(s0 ) = t̃(s0 ), n(s0 ) = ñ(s0 ), b(s0 ) = b̃(s0 ) (2.11)
Consider the expression
A(s) = t(s)t̃(s) + n(s)ñ(s) + b(s)b̃(s).
Differential Geometry 25
Using (2.11), we get A(s0 ) = 3. On the other hand, since t and t̃ are unit
vectors, t · t̃ ≤ 1, with equality holding if and only if t = t̃ and similarly
n · ñ ≤ 1, and b · b̃ ≤ 1. It follows that A(s) ≤ 3, with equality holding if and
only if t(s) = t̃(s), n(s) = ñ(s) and b(s) = b̃(s). Thus, if we can prove that
A is constant, it will follow in particular that t(s) = t̃(s), i.e. that γ̇ = γ̃,˙
and hence that γ(s) − γ̃(s) is a constant. But by (2.11) again, this constant
vector must be zero, so γ = γ̃.
We are therefore reduced to proving that A is constant. But, using the
dA
Frenet-Serret equations, it follows that = 0.
ds
26 Dr. V. V. Acharya
Chapter 3
Global properties of curves
All the properties of curves that we have discussed so far are ‘local’: they
depend only on the behaviour of a curve near a given point and not on the
‘global’ shape of the curve. Proving global results about curves often requires
concepts from topology, in addition to the calculus techniques we have used
in the first two chapters of this book. Since we are not assuming that readers
of this book have extensive familiarity with topological ideas, we will not be
able to give complete proofs of some of the global results about curves that
we discuss in this chapter.
27
28 Dr. V. V. Acharya
Example 3.1 The ellipse γ(t) = (p cos t, q sin t), where p and q are non-zero
constants, is a simple closed curve with period 2π. The interior and exterior
of γ are, of course, given by
2
y2 2
y2
2 x 2 x
(x, y) ∈ R | 2 + 2 < 1 and (x, y) ∈ R | 2 + 2 > 1 .
p q p q
respectively.
Not all examples of simple closed curves have such an obvious interior
and exterior, however. Is the point p in the interior or the exterior of the
simple closed curve shown below?
Example 3.2 The limacon in Example 1.1.7 is closed but is not a simple
closed curve as it has a self-intersection . See Exercise 3.1.1.
The fact that a simple closed curve has an interior and an exterior enables
us to distinguish between the two possible orientations of γ. We shall say
that γ is positively-oriented if the signed unit normal ns of γ. (see Section
2.2) points into int(γ) at every point of γ. This can always be achieved by
replacing the parameter t of γ by −t, if necessary. In the diagrams below, the
arrow indicates the direction of increasing parameter. Is the simple closed
curve shown above positively-oriented?
1 1
Proof. Taking f (x, y) = − y, g(x, y) = x in Green’s theorem, we get
2 2
Z Z ZZ
1 1 1
f (x, y)dx + g(x, y)dy = xdy − ydx = − − dxdy
2 2 2
γ γ int γ
Z a ZZ
1
Thus, (xẏ − ẋy)dt = dxdy = A(int (γ)).
2
0 int γ
Further, equality holds if and only if F (t) = A sin t for all t ∈ [0, π], where A
is a constant.
Proof. Let G(t) = F (t)/ sin t. Then, denoting d/dt by a dot as usual,
Z π
dF
( )2 dt
dt
Z0 π
= (Ġ sin t + G cos t)2 dt
Z0 π Z π Z π
2 2
= Ġ sin t + 2 GĠ sin t cos t + G2 cos2 tdt
0 0 0
30 Dr. V. V. Acharya
Integrating by parts:
Z π Z π
2 π
2 GĠ sin t cos tdt = G sin t cos t|0 − G2 (cos2 t − sin2 t)dt
0 0
Z π
= G2 (sin2 t − cos2 t)dt
0
Hence,
Z π 2 Z π Z π Z π
dF 2 2 2 2 2
dt = Ġ sin t + G (sin t − cos t)dt + G2 (cos2 t)dt
0 dt
Z0 π 0 0
and so Z π Z π Z π
dF
( )2 dt − 2
F dt = Ġ2 sin2 tdt
0 dt 0 0
Theorem 3.4 (Four Vertex Theorem) Every convex simple closed curve
in R2 has at least four vertices. 1
Proof. Let γ be a parametrization of a convex simple closed curve in R2 ,
and let ` be its length. Assume for a contradiction that γ has fewer than
four vertices. We show first that there is a straight line L that divides γ into
two segments, in one of which κ̇s > 0 and in the other κ̇s ≤ 0 (or possibly
κ̇s ≥ 0 on one and κ̇s < 0 on the other). Indeed, κs attains all of its values on
1
The conclusion of this theorem actually remains true without the assumption of con-
vexity, but the proof is then more difficult than the one we are about to give.
Differential Geometry 31
the closed interval [0, l], so κs must attain its maximum and minimum values
at some points P and Q of γ. We can assume that P 6= Q, since otherwise
κ̇s would be constant, γ would be a circle,and every point of γ would be a
vertex.
If P and Q were the only vertices of γ, we would have κ̇s > 0 on one of
the segments into which the line through P and Q divides γ and κ̇s < 0 on
the other.
Let a be a unit vector perpendicular to L, so that γ · a > 0 on one side
of L and γ · a < 0 on the other. Then, the quantity κ̇s (γ · a) is either always
> 0 or always < 0, except at the two points in which L intersects the curve.
It follows that Suppose now that there is just one more vertex, say R. Then,
P, Q and R divide γ into three segments, on each of which either κ̇s > 0 or
κ̇s < 0. It follows that there are two adjacent segments on which κ̇s > 0 or
two on which κ̇s < 0 (except at the point at which the two segments meet).
This proves our assertion.
Let a be a unit vector perpendicular to L, so that γ · a > 0 on one side
of L and γ · a < 0 on the other. Then, the quantity κ̇s (γ · a) is either always
> 0 or always < 0, except at the two points in which L intersects the curve.
It follows that Z a
κ̇s (γ · a) ≥ 0, (3.4)
0
as this integral is definitely > 0 in the first case and < 0 in the second. But,
using the equation ṅs = −κs t, we get
d d
κ̇s γ = (κs γ) − κs γ̇ = (κs γ + ns ),
dt dt
and hence κs (t + `) = κs (t). It follows that λ(t + `) = λ(t) for all t, so the
integral in (3.4) is equal to
Z `
λ̇(t)dt = λ(`) − λ(0) = 0. (3.5)
0
33
34 Dr. V. V. Acharya
4.2 Homeomorphism
distance of a is also in U :
a ∈ U and ku − ak < =⇒ u ∈ U.
the open ball with centre a and radius r > 0. 2 (If n = 1, an open ball is
called an open interval; if n = 2 it is called an open disc.) However,
Dr (a) = {u ∈ Rn | ku − ak ≤ r},
is not open because however small the positive number is, there is a point
within a distance of the point (a1 + r, a2 , ..., an ) ∈ Dr (a) (say) that is not in
Dr (a) (for example, the point (a1 + r + 2 , a2 , ..., an )).
Next, if X and Y are subsets of Rm and Rn , respectively, a map f : X →
Y is said to be continuous at a point a ∈ X if points in X near a are mapped
by f to points in Y near f (a). More precisely, f is continuous at a if, given
any number > 0, there is a number δ > 0 such that
Example 4.2 A circular cylinder is the set of points of R3 that are at a fixed
distance (the radius of the cylinder) from a fixed straight line (its axis). For
example, the circular cylinder of radius 1 and axis the z-axis, which we shall
call the unit cylinder, is
and the restriction σ|U of σ to U is a surface patch. However, σ|U does not
cover the whole of S, but only the open subset obtained by removing the line
x = 1, y = 0 from S.
To get an atlas for S we therefore need at least one more surface patch.
We can take σ|U1 , where U1 = {(u, v) ∈ R2 | − π < u < π}; this covers the
open subset of S obtained by removing the line x = 1, y = 0. Every point of
S is in the image of at least one of the surface patches σ|U , σ|U1 , so σ|U , σ|U1
is an atlas for S, and S is a surface.
Example 4.3 A sphere is the set of points of R3 that are at a fixed distance
(the radius of the sphere) from a fixed point (its centre). For example, the
sphere of radius 1 and centre the origin, called the unit sphere, is
1
0.5
0
-0.5
-1
1
0.5
-0.5
-1
-1
-0.5
0
0.5
1
Denote the right-hand side of this equation by σ(θ, ϕ); this is the latitude-
longitude parametrization of S 2 . See Fig. 4.3.
(i) (ii)
However, the set of points (θ, ϕ) satisfying these inequalities is not an open
subset of R2 . The largest open set consistent with the above inequalities is
π π
U = {(θ, ϕ)| < θ < , 0 < ϕ < 2π}.
2 2
however, the image of σ|U is not the whole of the sphere, but the open subset
obtained by removing the semicircle C consisting of the points of the sphere
of the form (x, 0, z) with x ≥ 0. (See Fig. 4.4 (i))
To show that the sphere is a surface, we must therefore produce at least
one more surface patch covering the part of the sphere omitted by σ. One
possibility is the patch σ
b obtained by first rotating σ by π about the z-axis
π
and then by about the x-axis. Explicitly, σ b : U → R3 is given by
2
b (θ, ϕ) = (− cos θ cos ϕ, − sin θ, − cos θ sin ϕ).
σ
(the open set U is the same as for σ). The image of σ b is the open subset
of S 2 obtained by removing the semicircle C̃ consisting of the points of the
sphere of the form (x, y, 0) with x ≤ 0. (Fig. 4.4 (ii))
It is clear that C and C̃ do not intersect, so the union of the images of
σ|U and σ b |U is the whole sphere.
Example 4.4 Show that the unit cylinder can be covered by a single surface
patch, but that the unit sphere cannot.
Solution.
√ For the first part, let U = {(u, v) ∈ R2 |0 < u2 + v 2 < π 2 }, let
r = u2 + v 2 , and define σ : U → R3 by σ(u, v) = ( ur , vr , tan(r − π2 )). It
40 Dr. V. V. Acharya
Example 4.5 The circular cone with vertex a point v, axis a straight line l
passing through v, and angle α, where 0 < α < π2 , is the set of points p in
R3 such that the straight line through v and p makes an angle α with the
line l. For example, if v is the origin, l is the z-axis and α = π4 , the circular
cone is
S = {(x, y, z) ∈ R3 |x2 + y 2 = z 2 }.
To see that this is not a surface, suppose that σ : U → S ∩ W is a surface
patch containing the vertex (0, 0, 0) of the cone, and let a ∈ U correspond to
the vertex. We can assume that U is an open ball with centre a, since any
open set U containing a must contain such an open ball. The open set W
must obviously contain a point p in the lower half S− of S where z < 0 and
a point q in the upper half S+ where z > 0; let b and c be the corresponding
points in U.
entirely in S, passing through p and q, and not passing through the vertex.
(It is true that γ will, in general, only be continuous, and not smooth, but this
does not affect the argument.) This is clearly impossible for if we take the
projection map Π : R3 → R defined by Π(x, y, z) = z then Π is a continuous
map and Π(p) < 0 and Π(q) > 0. Hence, there must exist a point r on γ
such that Π(r) = 0. But the only possible such point on the double cone is
(0, 0, 0) which does not lie on γ.
If we remove the vertex, however, we do get a surface S− ∪ S+ . It has
an atlas consisting of the two surface patches σ ± : U → R3 , where U =
R2 − {(0, 0)}, given by the inverse of projection onto the xy plane:
√
σ ± (u, v) = (u, v, ± u2 + v 2 ).
As the example of the sphere shows, a point a of a surface S will generally
lie in the image of more than one surface patch. In general, suppose then that
σ : U → S ∩W and σ e :Ue → S ∩W f are two patches such that a ∈ S ∩W ∩ W f.
e are homeomorphisms, σ −1 (S ∩ W ∩ W
Since σ and σ e −1 (S ∩ W ∩ W
f ) and σ f)
are open sets V ⊆ U and Ve ⊆ U e , respectively. The composite homeomor-
−1
phism σ ◦ σ e : Ve → V is called the transition map from σ to σ e . If we
denote this map by Φ, we have
σ
e (e
u, ve) = σ(F (e u, ve) ∈ Ve .
u, ve)) for all (e
Example 4.6 Define surface patches σ x± : U → R3 for S 2 by solving the
equation x2 + y 2 + z 2 = 1 for x in terms of y and z:
√
σ x± (u, v) = (± 1 − u2 − v 2 , u, v),
42 Dr. V. V. Acharya
defined on the open set U = {(u, v) ∈ R2 |u2 + v 2 < 1}. Define σ y± and σ z± .
similarly (with the same U ) by solving for y and z, respectively. Show that
these six patches give S 2 the structure of a surface.
open subset of S. A smooth surface is a surface S such that, for any point
p ∈ S, there is an allowable surface patch σ as above such that p ∈ σ(U ).
A collection A of allowable surface patches for a surface S such that every
point of S is in the image of at least one patch in A is called an atlas for the
smooth surface S.
Example 4.7 The plane in Example 1 is a smooth surface. For
σ(u, v) = a + up + vq
is clearly smooth and σ u = p and σ v = q are linearly independent because
p and q were chosen to be perpendicular unit vectors.
Example 4.8 The unit cylinder in Example 2 is a smooth surface. Indeed,
σ(u, v) = (cos u, sin u, v)
is clearly smooth and σ u = (− sin u, cos u, 0), σ v = (0, 0, 1) are mutually
orthonormal vectors and hence are linearly independent for all (u, v), so σ|U
and σ̃|Ũ are regular surface patches.
Example 4.9 For the unit sphere S 2 in Example 3, it is again clear that σ
and σ̃ are smooth. As for regularity,
σ θ = (− sin θ cos ϕ, sin θ sin ϕ, cos θ), σ ϕ = (− cos θ sin ϕ, cos θ cos ϕ, 0),
σ θ × σ ϕ = (− cos2 θ cos ϕ, − cos2 θ sin ϕ, − sin θ cos θ) and hence
kσ θ × σ ϕ k = | cos θ|.
But if (θ, ϕ) ∈ U, then − π2 < θ < π
2
so cos θ 6= 0. Similarly, one checks that
σ̃ is regular.
In Example 4.6 we gave another family of allowable surface patches cov-
ering the unit sphere S 2 (it is easy to check that they are regular.) An
obvious question is: which of these two atlases should we use to study the
sphere? The answer is that we can use either, or both. The eight patches
in Example 3 and Example 6 together form a third atlas. In most situations
(although not in all.), one might as well use the maximal atlas for a given
surface S consisting of all of its allowable surface patches. The maximal atlas
is independent of any arbitrary choices.
Although not at first sight very interesting, the next two results are very
important for what is to follow.
44 Dr. V. V. Acharya
The proof of this will be given later. The next result is a kind of converse.
We recall from calculus that, if Ψ and Ψ̃ are two smooth maps between
open sets of R2 ,
J(Ψ̃ ◦ Ψ) = J(Ψ̃)J(Ψ).
(In fact, this is equivalent to the chain rule that expresses the first partial
derivatives of Ψ̃ ◦ Ψ in terms of those of Ψ̃ and Ψ.) Taking Ψ = Φ and
Ψ̃ = Φ−1 , we see that J(Φ−1 ) = J(Φ)−1 . In particular, J(Φ) is invertible, so
its determinant is non-zero and Eq. (4.1) shows that σ̃ is regular.
provided we can define it for any regular surface patch in such a way that it
is unchanged when the patch is reparametrized. We shall give an important
application of this principle in the next section.
For the rest of this course, by a surface we shall mean a smooth
surface S, and by a surface patch for S we shall mean an allowable
surface patch for S.
Unless we indicate otherwise, we shall also assume that all surfaces
we consider are connected, which means that any two points of S can
be joined by a curve lying entirely in S. This is not a serious restriction,
for it is not difficult to prove that any surface S is a disjoint union
of connected surfaces, each of which is an open subset of S, so S can
be studied by studying each of its connected parts separately. All the
surfaces we have encountered so far are connected except the double
cone of Example 5, which breaks into the union of two disjoint half
cones S± when the vertex is removed, as it must be to have a surface.
σ̃ −1 −1 −1 −1
2 ◦ f ◦ σ̃ 1 = σ̃ 2 ◦ (σ 2 ◦ σ 2 ) ◦ f ◦ (σ 1 ◦ σ 1 )σ̃ 1
= (σ̃ −1 −1 −1
2 ◦ σ 2 ) ◦ σ 2 ) ◦ f ◦ σ 1 ◦ (σ 1 σ̃ 1 )
= Φ−1 −1
2 ◦ (σ 2 ◦ f ◦ σ 1 ) ◦ Φ1
Theorem 4.1 Let S be a subset of R3 with the following property: for each
point p ∈ S, there is an open subset W of R3 containing p and a smooth
function f : W → R such that
Differential Geometry 47
where the sign is that of the determinant of J(Φ). This leads to the following
definition.
all points in the image of σ, otherwise discard it. The patches that remain
form an atlas a satisfying the condition in Definition 6.
From now onwards, whenever we are dealing with an oriented surface S,
we shall only use surface patches for S whose standard unit normal is the
same as the chosen normal of S.
∂f ∂f
and are not both zero. If P is a point of C, with coordinates (x0 , y0 ),
∂x ∂y
say, there is a regular parametrized curve γ(t), defined on an open interval
containing 0, such that γ passes through P when t = 0 and γ(t) is contained
in C for all t.
∂f ∂f
Proof. We note that n = , is non-zero at every point of C. We
∂x ∂y
∂f
may assume that 6= 0 at P. Then, n is not parallel to the x-axis, so the
∂y
tangent to C at P is not parallel to the y-axis.
This implies that vertical lines x = constant near x = x0 all intersect C
in a unique point (x, y) near P. In other words, the equation
f (x, y) = 0 (4.4)
has a unique solution y near y0 for every x near x0 . 6 Thus, there is a function
g(x), defined for x near x0 , such that y = g(x) is the unique solution of Eq.
4.4 near y0 . We can now define a parametrization γ of the part of C near P
by
γ(t) = (t, g(t)).
Note that g is smooth (which follows from the inverse function theorem),
then γ is regular since ġ = (1, ġ) is obviously never zero.
Theorem 4.4 Let γ be a regular parametrized plane curve, and let γ(t0 ) =
(x0 , y0 ) be a point in the image of γ. Then, there is a smooth real-valued
function f (x, y), defined for x and y in open intervals containing x0 and y0 ,
respectively, and satisfying the conditions in Theorem 4.3, such that γ(t)
is contained in the level curve f (x, y) = 0 for all values of t in some open
interval containing t0 .
Proof. Let γ(t) = (u(t), v(t)), where u and v are smooth functions. Since γ
is regular, at least one of u0 (t0 ) and v 0 (t0 ) is non-zero, say u0 (t0 ). This means
that the graph of u as a function of t is not parallel to the t-axis at t0 .
This implies that any line parallel to the t-axis close to u = x0 intersects
the graph of u at a unique point u(t) with t close to t0 . This gives a function
h(x), defined for x in an open interval containing x0 , such that t = h(x) is
6
Note that this may fail to be the case if the tangent to C at P is parallel to the y-axis
∂f
(i.e., if = 0.)
∂y
Differential Geometry 53
f (x, y) = y − v(h(x))
2 ×
of σ has rank 2 everywhere. Hence, at least one of its three 2 submatrices
f f
is invertible at each point. Suppose that the submatrix u v is invertible
gu gv
at P. (The proof is similar in the other two cases.) By the inverse function
theorem applied to the map F : U → R2 given by
Theorem 4.6 Let S be a subset of R3 with the following property: for each
point P ∈ S, there is an open subset W of R3 containing P and a smooth
function f : W → R such that
(i) S ∩ W = {(x, y, z) ∈ W |f (x, y, z) = 0};
(ii) The gradient ∇f = (fx , fy , fz ) of f does not vanish at P.
Then, S is a smooth surface.
Proof. Suppose that P = (x0 , y0 , z0 ) and that fz 6= 0 at P. (The proof is
similar in the other two cases.) Consider the map F : W → R3 defined by
is called the first fundamental form of the surface patch σ(u, v). Note that
the coefficients E, F, G and the linear maps du, dv depend on the choice of
surface patch for S, but the first fundamental form itself depends only on S
and p. Let
L = σ uu · N, M = σ uv · N, N = σ vv · N. (5.3)
The expression
Ldu2 + 2M dudv + N dv 2 (5.4)
is called the second fundamental form of the surface patch σ.
Example 5.1 For the plane σ(u, v) = a + up + vq with p and q being
perpendicular unit vectors, we have σ u = p, σ v = q, so E = kσ u k2 =
kpk2 = 1, F = σ u · σ v = p · q = 0, G = kσ v k2 = kqk2 = 1, and the first
fundamental form is simply du2 + dv 2 . Since σ u = p and σ v = q are constant
vectors, we have σ uu = σ uv = σ vv = 0. Hence, the second fundamental form
of a plane is zero.1
55
56 Dr. V. V. Acharya
of sinh u, cosh u are cosh u and sinh u respectively. Hence, the derivative of
tanh u and sechu are sech2 u and −sechu tanh u resp.
Hence,
Note that the first and the second fundamental form of S 2 are same.
called its meridians. (This agrees with the use of these terms in geography,
if we think of the earth as the surface obtained by rotating a circle passing
through the poles about the polar axis and we take u and v to be latitude
and longitude, respectively.)
Let us take the axis of rotation to be the z-axis and the plane to be the
xz-plane. Any point p of the surface is obtained by rotating some point q of
the profile curve through an angle v (say) around the z-axis. If
γ(u) = (f (u), 0, g(u))
is a parametrization of the profile curve containing q, p is of the form
σ(u, v) = (f (u) cos v, f (u) sin v, g(u)).
To check regularity, we compute (with a dot denoting d/du):
σ u = (f˙ cos v, f˙ sin v, ġ), σ v = (−f sin v, f cos v, 0),
σ u × σ v = (f ġ cos v, −f ġ sin v, f f˙),
kσ u × σ v k2 = f 2 (f˙2 + ġ 2 ).
Thus, σ u × σ v will be non-vanishing if f (u) is never zero, i.e., if γ does not
intersect the z-axis, and if f˙ and ġ are never zero simultaneously, i.e., if γ is
regular. In this case, we might as well assume that f (u) > 0, so that f (u) is
the distance of σ(u, v) from the axis of rotation. Then, σ is injective provided
that γ does not self-intersect and the angle of rotation v is restricted to lie
in an open interval of length ≤ 2π. Under these conditions, surface patches
of the form σ give the surface of revolution the structure of a surface.
We now calculate the first fundamental form for surface of revolution.
Note that f (u) > 0 for all values of u and that the profile curve u 7→
(f (u), 0, g(u)) is unit-speed, i.e., f˙2 + ġ 2 = 1 (a dot denoting d/du). Then:
σ u = (f˙ cos v, f˙ sin v, ġ), σ v = (−f sin v, f cos v, 0). Hence, E = kσ u k2 =
f˙2 + ġ 2 = 1, F = σ u · σ v = 0, G = kσ v k2 = f 2 . So the first fundamental form
is du2 + f (u)2 dv 2 .
Note that
σu × σv
N= = (ġ cos v, −ġ sin v, f˙),
kσ u × σ v k
σ uu = (f¨ cos v, f¨ sin v, g̈), σ uv = (−f˙ sin v, f˙ cos v, 0), and
σ vv = (−f cos v, −f sin v, 0). Hence,
L = σ uu · N = f˙g̈ − f¨f˙, M = σ uv · N = 0, N = σ vv · N = f ġ,
58 Dr. V. V. Acharya
σ u = ġ + v δ̇, σ v = δ.
σ u = γ̇, σ v = a,
so E = kσ u k2 = kġk2 = 1, F = σ u · σ v = ġ · a = 0, G = kσ v k2 = kak2 = 1,
and the first fundamental form of σ is du2 + dv 2 . Note that this is the same
as the first fundamental form of the plane.
Further, N = γ̇ × a and σ uu = γ̈, σ uv = 0 and σ vv = 0. Hence, L =
σ uu · N = γ̈ · (γ̇ × a), M = 0, N = 0 and the second fundamental form is
σ uu · N = γ̈ · (γ̇ × a)du2 .
In particular, if the surface is the unit cylinder, γ(u) = (cos u, sin u, 0) and
a = (0, 0, 1). Then the first fundamental form of σ is du2 + dv 2 , where
60 Dr. V. V. Acharya
Example 5.5 The second special case we shall consider is that in which
the rulings all pass through a certain fixed point, say v; then S is called a
generalized cone with vertex v.
We can take δ(u) = γ(u) − v, giving σ(u, v) = (1 + v)γ(u) − vv. Now,
σ(u, v) = σ(u0 , v 0 ) ⇐⇒ (1 + v)γ(u) − (1 + v 0 )γ 0 (u0 ) + (v 0 − v)v = 0; since
(1 + v) − (1 + v 0 ) + (v 0 − v) = 0, the equation on the right-hand side means
that the points v, σ(u) and σ(u0 ) are collinear. So, for σ to be a surface
patch, no straight line passing through v should pass through more than one
point of γ (in particular, γ should not pass through v). Finally, we have
σ u = (1 + v)γ̇, σ v = γ − v, so σ is regular provided v 6= −1, i.e., the vertex
of the cone is omitted (cf. Example 4.1.5), and none of the straight lines
forming the cone is tangent to γ.
The parametrization is simplest when γ lies in a plane. If this plane
contains v, the cone is simply part of that plane. Otherwise, we can take
v to be the origin and the plane to be z = 1. Then, σ(u) = (f (u), g(u), 1)
for some smooth functions f and g, and the parametrization takes the form
σ(u, v) = v(f (u), g(u), 1), after making the reparametrization v 7→ v − 1.
Before computing its first fundamental form, we make some simplifica-
tions to σ.
First, translating the surface by v (which does not change its first funda-
mental form), we get the surface patch σ 1 = σ − v = (1 + v)(γ − v), so if
we replace γ − v by γ 1 we get σ 1 = (1 + v)γ 1 . This means that we might as
well assume that v = 0 to begin with.
Next, we have already seen that for σ to be a regular surface patch, γ
must not pass through the origin, so we can define a new curve γ̃ by γ̃(u) =
γ(u)/ kγ(u)k . Setting ũ = u, ṽ = (1+v)/ kγ(u)k , we get a reparametrization
σ̃(ũ, ṽ) = ṽ γ̃(ũ) of σ with kγ̃k = 1. We can therefore assume to begin with
that σ(u, v) = vγ(u) with kγ(u)k = 1 for all values of u (geometrically,
this means that we can replace γ by the intersection of the cone with S 2 ).
Finally, reparametrizing again, we can assume that γ is unit-speed, for we
have already seen that for σ to be regular, γ must be regular.
Differential Geometry 61
It is clear that γ 1 and γ 2 have the same length, since both lengths are found
by integrating the expression (E u̇2 +2F u̇v̇ +Gv̇ 2 )1/2 , where Edu2 +2F dudv +
Gdv 2 is the (common) first fundamental form of σ 1 and σ 2 .
Conversely, suppose that f is an isometry. If t 7→ (u(t), v(t)) is any
curve in U, defined for t ∈ (α, β), say, the curves γ 1 (t) = σ 1 (u(t), v(t)) and
γ 2 (t) = σ 2 (u(t), v(t)) have the same length. Hence,
Zt1 Zt1
(E1 u̇2 + 2F1 u̇v̇ + G1 v̇ 2 )1/2 dt = (E2 u̇2 + 2F2 u̇v̇ + G2 v̇ 2 )1/2 dt,
t0 t0
for all t0 , t1 ∈ (α, β), where E1 , F1 and G1 are the coefficients of the first
fundamental form of σ 1 , and E2 , F2 and G2 those of σ 2 . This implies that
the two integrands are the same, and hence that
E1 u̇2 + 2F1 u̇v̇ + G1 v̇ 2 = E2 u̇2 + 2F2 u̇v̇ + G2 v̇ 2 (5.5)
Fix t0 ∈ (α, β), and let u0 = u(t0 ), v0 = v(t0 ). We now apply Eq. (5.5)
for the following three choices of the curvet 7→ (u(t), v(t)) in U : (i) u =
u0 + t − t0 , v = v0 : this gives E1 = E2 ; (ii) u = u0 , v = v0 + t − t0 : this gives
G1 = G2 ; (iii) u = u0 + t − t0 , v = v0 + t − t0 : this gives
E1 + 2F1 + G1 = E2 + 2F2 + G2 ,
and hence (in view of (i) and (ii)) F1 = F2 .
F
cos θ = √ ,
EG
Proof. We can assume that S1 and S2 are covered by single surface patches.
Moreover, since f is a diffeomorphism, Proposition 4.3 shows that we can
assume that these patches are of the form σ 1 : U → R3 (for S1 ) and f ◦ σ 1 =
σ 2 (for S2 ). We have to show that f is a conformal map if and only if the
first fundamental form σ 1 and σ 2 are proportional.
Suppose that their first fundamental forms E1 du2 + 2F1 dudv + G1 dv 2 and
E2 du2 + 2F2 dudv + G2 dv 2 are proportional, say
for some smooth function λ(u, v), where (u, v) are coordinates on U. Note
that λ > 0 everywhere, since (for example) E1 and E2 are both > 0.
Suppose that γ and γ̃ lie in a surface patch σ of S1 , so that γ(t) =
σ 1 (u(t), v(t)) and γ̃(t) = σ 1 (ũ(t), ṽ(t)) for some smooth functions u, v, ũ and
ṽ, then f takes γ and γ̃ to the curves σ 2 (u(t), v(t)) and σ 2 (ũ(t), ṽ(t)) in S2
resp. Using (5.7), the angle θ) of intersection of the latter curves on S2 is
given by
since the λ’s cancel. But, using (5.7) again, we see that the right-hand side
is the cosine of the angle of intersection of the curves γ and on γ̃. Hence, f
is conformal.
For the converse, we must show that if
for all pairs of intersecting curves γ(t) = σ 1 (u(t), v(t)) and γ̃(t) = σ 1 (ũ(t), ṽ(t))
in S1 , then the first fundamental forms of σ 1 and σ 2 are proportional. Fix
Differential Geometry 65
E1 cos φ + F1 sin φ
(E1 )1/2 (E1 φ + 2F1 cos φ sin φ + G1 sin2 φ)1/2
cos2
E2 cos φ + F2 sin φ
= . (5.9)
(E2 )1/2 (E2 cos2 φ + 2F2 cos φ sin φ + G2 sin2 φ)1/2
− E1 G1 − F12 sin2 φ
(5.10)
= E2 G2 − F22 (E1 )(E1 cos2 φ + 2F1 cos φ sin φ + G1 sin2 φ). (5.11)
E2 cos2 φ+2F2 cos φ sin φ+G2 sin2 φ = λ(E1 cos2 φ+2F1 cos φ sin φ+G1 sin2 φ).
(E2 − λE1 ) cos2 φ + 2 (F2 − λF1 ) cos φ sin φ + (G2 − λG1 ) sin2 φ = 0. (5.12)
π π π
Taking φ = we get G2 = λG1 and then φ = and φ = gives E2 = λE1 ,
2 4 6
and F2 = λF1 .
Let P = (u, v, 0), Q = (x, y, z). Since P, Q, N lie on a straight line, there
is a scalar ρ such that
Q − N = ρ(P − N ),
and hence
u2 + v 2 − 1
2u 2v
Q= , , .
u2 + v 2 + 1 u2 + v 2 + 1 u2 + v 2 + 1
|w|2 − 1
2w
σ 1 (w) = , .
|w|2 + 1 |w|2 + 1
The inconvenience of having to exclude the north pole from the domain of
definition of Π can be overcome by introducing a ‘point at infinity’ ∞ and
defining the ‘extended complex plane’ C∞ = C ∪ {∞}. If we agree that Π
maps the north pole to ∞, it defines a bijection Π : S 2 → C∞ .
Definition 5.3 The area Aσ (R) of the part σ(R) of a surface patch σ :
U → R3 corresponding to a region R ⊆ U is
ZZ
Aσ (R) = kσ u × σ v k dudv.
R
68 Dr. V. V. Acharya
kσ u × σ v k2 = (σ u · σ u )(σ v · σ v ) − (σ u · σ v )2 = EG − F 2 .
Hence,
ZZ ZZ
∂(u, v)
kσ̃ ũ × σ̃ ṽ k dũdṽ = det kσ ũ × σ ṽ k dũdṽ
∂(ũ, ṽ)
R̃ R̃
By the change of variables formula for double integrals, the right-hand side
of this equation is exactly
ZZ
kσ u × σ v k dudv.
R
taken over the disc u2 + v 2 < 1. To calculate the area, we use change of
variables. Let u = r sin θ, v = r cos θ then the area is
Z 1√ Z 5√
2π π
2π 1 + 4r2 rdr = tdt = (53/2 − 1).
0 8 0 6
This is less than the area 2π of the hemisphere.3
Example 5.9 Determine the surface area of the portion of the plane y+2z =
2 within the cylinder x2 + y 2 = 3.
v
Solution. Note that σ(u, v) = (u, v, 1 − ). Hence, σ u = (1, 0, 0) , σ v =
√ 2 √
1 5 3 5
0, 1, − . Hence, kσ u × σ v k = . Thus, A = π.
2 2 2
xy
Example 5.10 Determine the area of the surface z = intercepted by the
3
cylinder x2 + y 2 ≤ 16.
3
Put 1 + 4r2 = t. Hence, 8rdr = dt.
70 Dr. V. V. Acharya
uv v
Solution. Note that σ(u, v) = (u, v, ). Hence, σ u = 1, 0, , σv =
3 3
2
u v 2 uv u2 2
0, 1, . Thus, the first fundamental form is du + 2 dudv + dv .
3 9 9 9
1 RR √
Hence, the required area is u + v + 9dudv, taken over the disc u2 +
2 2
3
v 2 < 16. 4 To calculate the area, we use change of variables. Let u =
r sin θ, v = r cos θ then the area is
π
Z 4 √ π 25 √
Z
π 2 25 2π 196π
2
9 + r 2rdr = tdt = × t3/2 9 = [125 − 27] = .
3 0 3 9 3 3 9 9
Now that we have a good definition of area, we can ask which maps between
surfaces are area-preserving.
for all regions R ⊆ U. This holds if and only if the two integrands are equal
everywhere, i.e., if and only if E1 G1 − F12 = E2 G2 − F22 .
v u
4
Alternatively, we can find σ u × σ v = − , − , 1 . Hence, kσ u × σ v k =
3 3
1√ 2
u + v 2 + 9.
3
Differential Geometry 71
where κ is the curvature of γ and ψ is the angle between N and the principal
normal n of γ.
Proof. Equations (5.17) and (5.18) follow from (5.16) and the fact that N
and N× γ̇ are perpendicular unit vectors. The first equation in (5.19) follows
from γ̈ = κn, and the second then follows from Eq. (5.18)
N · σ u = N · σ v = 0.
Nu · σ u = −N · σ uu = −L, Nv · σ u = −N · σ uv = −M,
Nu · σ v = −N · σ uv = −M, Nv · σ v = −N · σ vv = −N.
κn = Lu̇2 + 2M u̇v̇ + N v̇ 2 ,
Differential Geometry 73
This result means that two curves which touch each other at a point p of
a surface (i.e., which intersect at p and have parallel tangent vectors at p)
have the same normal curvature at p.
κn = ±κ, κg = 0.
t1 · t2 = (ξ1 σ u + η1 σ v ) · (ξ2 σ u + η2 σ v )
= Eξ1 ξ2 + F (ξ1 η2 + ξ2 η1 ) + Gη1 η2
E f ξ2
= T1t FI T2 ,
= ξ1 η1
F G η2
ξ1 ξ2 u̇
where, T1 = and T2 = . Similarly, if γ̇ = σ u u̇ + σ v v̇ and T =
η1 η2 v̇
t
then using proposition 5.7, we see that κn = T FII T as
L M u̇
= Lu̇2 + 2M u̇v̇ + N v̇ 2 .
u̇ v̇
M N v̇
Definition 5.6 The principal curvatures of a surface patch are the roots of
the equation
det(FII − κFI ) = 0, (5.20)
i.e
L − κE M − κF
= 0. (5.21)
M − κF N − κG
Since, (5.21) is a quadratic equation for κ, there are two roots. It turns out
that both the roots are real! If κ is one of the principal curvatures, equation
(5.20) says that FII − κFI is not invertible, so, assuming that κ is real, there
is a non-zero 2 × 1 column matrix T having real number entries such that
Example 5.12 Consider the unit sphere S 2 and use the latitude longitude
parametrization as usual.We found that E = 1, F = 0, G = cos2 θ and that
L = 1, M = 0, N = cos2 θ. So the principal curvatures are the roots of
1−κ 0
= 0,
0 cos2 θ − κ cos2 θ
Example 5.13 Consider the unit cylinder parametrized in the usual way:
0−κ 0
= 0,
0 1−κ
(ii) if κ1 = κ2 = κ, say, then FII = κFI and (hence) every tangent vector
to σ at P is a principal vector;
K = κ1 κ2 (5.23)
Proposition 5.10 Let σ(u, v) be a surface patch with first and second fun-
damental forms
respectively. Then,
LN − M 2
(i) K = ,
EG − F 2
LG − 2M F + N E
(ii) M = ,
2 (EG − F 2 )
√
(iii) the principal curvatures are H ± H 2 − K.
Proof. Note that by definition, the principal curvatures are the roots of the
equation
L − κE M − κF
= 0.
M − κF N − κG
Hence, (L − κE) (N − κG) − (M − κF )2 = 0, i.e.
2
EG − F 2 κ2 − (LG − 2M F + N E) + LN − M 2 = 0.
78 Dr. V. V. Acharya
Using the relation between roots and coefficients of the quadratic equation,
we get (i) and (ii). By the definition of H and K, κ1 and κ2 are the roots of
κ2 − 2Hκ + K = 0,
√
Hence, κ = H ± H 2 − K.
Example 5.16 Find the Gaussian and mean curvature of the unit sphere
and right circular cylider of radius 1.