2017-curves
2017-curves
by 2409:40d2:a:8f73:10c2:8aff:fe61:88b7 on 03/10/25. Re-use and distribution is strictly not permitted, except for Open Access articles.
Chapter I
Curves
Differential Geometry of Curves and Surfaces Downloaded from www.worldscientific.com
Students at the college and university level are familiar with graphs of
functions and parametrized curves. In this chapter, we define the cur-
vature of plane curves and space curves, which measures the “amount
of bending” of the curve, and introduce its geometric meanings. As ap-
plications, the rotation index for closed plane curves and properties of
spirals are introduced. While space curves are closely related to surface
theory, which is treated in Chapter II, the rotation index for plane curves
is also used in the proof of the Gauss-Bonnet theorem for 2-dimensional
manifolds in Chapter III.
The first picture (i) in Fig. 1.1 is an example of a smooth curve. The
second picture (ii) is also of a smooth curve, but now without endpoints,
since traveling along the curve will bring you back to where you started.
This type of curve is called a closed curve. Like in the picture (iii), we
1
March 29, 2017 9:45 ws-book9x6 World Scientific Book - 9in x 6in 9901-main page 2
allow crossing points, which we can also call self-intersections, and we still
regard it as a smooth closed curve. The picture (iv) is a closed curve,
but as it has sharp angles at particular points, it is not smooth at those
by 2409:40d2:a:8f73:10c2:8aff:fe61:88b7 on 03/10/25. Re-use and distribution is strictly not permitted, except for Open Access articles.
points. This type of curve is called a piecewise smooth curve (cf. page 225
in Appendix B.3). In this case, we are dealing with a curve that is still
smooth almost everywhere. We now introduce mathematical methods for
representing curves in the plane R2 .
Differential Geometry of Curves and Surfaces Downloaded from www.worldscientific.com
1 Each point on the lemniscate satisfies that the product of its distances to the two
√
points (±a/ 2, 0) is a2 /2. This curve was studied by Jacob Bernoulli (1654–1705).
2 The astroid is a locus of one point of a circle with radius a/4 rotating around the
Curves 3
by 2409:40d2:a:8f73:10c2:8aff:fe61:88b7 on 03/10/25. Re-use and distribution is strictly not permitted, except for Open Access articles.
Differential Geometry of Curves and Surfaces Downloaded from www.worldscientific.com
An ellipse. A hyperbola.
Fig. 1.2
There are many varieties of shapes that singular points can appear as.
plane with respect to a time variable t, we denote by (x(t), y(t)) the point
depending on t. Then the trajectory of these points can be imagined as a
curve. Here, t is called a parameter of the curve. This way of representing
a curve is called a parametrization. Pairing up the two functions x(t), y(t)
and naming the pair γ(t), we have the curve
Differential Geometry of Curves and Surfaces Downloaded from www.worldscientific.com
For the curve obtained as the graph of y = f (x), we can choose the special
form
If we restrict s to the open interval (−π, π), we again get the same circle
as in (1.2) with one point removed. When (x(t), y(t)) and (x(s), y(s)) are
the same point, the parameters s and t are related by t = tan(s/2).
In general, a curve is assumed to be represented as a smooth parametr-
ization, with parameter t, by
Curves 5
where cosh t, sinh t are hyperbolic functions (see page 195 in Appendix A.1).
Given a parametrization γ(t) = (x(t), y(t)) for a curve, and viewing the
curve as the trajectory of the point γ(t) that is moving with respect to
t, and in turn regarding t as representing time, we then have the velocity
vector
dx dy
γ̇(t) := (ẋ(t), ẏ(t)) ẋ = , ẏ = ,
dt dt
where an overhead dot “ ˙ ” denotes differentiation with respect to t. If the
velocity vector γ̇(c) vanishes, the value t = c of the parameter or the image
γ(c) is called a singular point (cf. Fig. 1.3). The singular point appearing
in Fig. 1.3 is called a cusp. See Appendix B.9 for a criterion for cusps.
Fig. 1.3 The singular point of the map t 7→ (t2 , t3 ) and its image.
March 29, 2017 9:45 ws-book9x6 World Scientific Book - 9in x 6in 9901-main page 6
For a given curve, the word “singular point” may have different mean-
ings, depending on whether it is represented by an implicit function or as a
parametrized curve. For example, the origin (0, 0) is a singular point in the
by 2409:40d2:a:8f73:10c2:8aff:fe61:88b7 on 03/10/25. Re-use and distribution is strictly not permitted, except for Open Access articles.
Curves 7
Writing the change in the two coordinate functions x(t), y(t) of these two
points as
by 2409:40d2:a:8f73:10c2:8aff:fe61:88b7 on 03/10/25. Re-use and distribution is strictly not permitted, except for Open Access articles.
Adding up many such short distances and taking the limit of such sums as
those distances approach zero, we arrive at the length
s
Z b 2 2 Z bp
dx dy
(1.5) L(γ) = + dt = ẋ2 + ẏ 2 dt
a dt dt a
of the curve.4 This length does not depend on the choice of the parametr-
ization for the regular curve (Problem 2 at the end of this section asks for
a proof of this). p
The integrand in (1.5) is the norm |γ̇| := ẋ(t)2 + ẏ(t)2 of the velocity
vector γ̇, so we can write
Z b
(1.6) L(γ) = |γ̇(t)| dt.
a
For example, even for the graph y = 1/x, it is known that the length
cannot be represented in terms of elementary functions. But by knowing
the equation above, we can approximate the length by approximating this
integral (using a computer, for example). Next we give an example where
we can compute the length explicitly:
Differential Geometry of Curves and Surfaces Downloaded from www.worldscientific.com
give the fixed endpoints of the hanging string. By equation (1.8), we have
dy/dx = sinh ax, and so the length l of the curve is
Z cp Z c
2
(1.9) l= 1 + sinh2 ax dx = 2 cosh ax dx = sinh ac.
−c 0 a
Using this, one can observe that for a given length l (> 0) of the string and
a distance 2c (2c < l) between the two endpoints, there is only one positive
constant a that satisfies (1.9).
A closed curve without self-intersection points is called a simple closed
curve, and if it is regular, it is called a simple closed regular curve. It is
March 29, 2017 9:45 ws-book9x6 World Scientific Book - 9in x 6in 9901-main page 9
Curves 9
known as the Jordan closed curve theorem 5 that a simple closed curve on
the plane separates the plane into two connected open subsets, one bounded
and the other not.
by 2409:40d2:a:8f73:10c2:8aff:fe61:88b7 on 03/10/25. Re-use and distribution is strictly not permitted, except for Open Access articles.
Definition 1.5. For a given simple closed curve in R2 , the bounded (resp.
unbounded) region6 whose boundary is a given curve is called the interior
(resp. exterior ) of the curve.
The Jordan closed curve theorem can be proven in a rather simple way
Differential Geometry of Curves and Surfaces Downloaded from www.worldscientific.com
for polygons (cf. [4]). A proof for general continuous curves can be found
in the textbook [8, Sec. 18] on algebraic topology.
For a simple closed curve with length l, the area A of the interior of the
curve satisfies the inequality
(1.10) 4πA ≤ l2 ,
which is called the isoperimetric inequality, for which equality holds if and
only if the curve is a circle. See Appendix B.3 for the proof, noting that
this appendix is dependent on material in the next section.
Exercises 1
1 Find the length of the portion of the parabola y = x2 where |x| ≤ 1.
2 Show that the length of a regular curve as given in (1.5) does not depend on
the choice of parametrization.
3 (1) For 0 < b < a, show that the length of the ellipse (a cos t, b sin t), 0 ≤ t ≤
2π), is given by
r !
Z 2π p a2 − b2
a 1 − ε2 cos2 t dt ε := .
0 a2
4 Taking a circle of radius a (> 0) and rolling it without slippage along the
horizontal x-axis, any given point on the circle traces out an image curve
called a cycloid (see the next figure). We choose that given point so that it
starts at the origin (0, 0) in the x-axis, and we let t denote the oriented angle
between the spoke from the center of the circle to the given point and the
vertical direction. We choose t so that it is zero when the given point is at
(0, 0) and is continuously increasing as the circle rolls to the right. Then the
resulting cycloid can be parametrized as
As the circle rotates completely around once, t increases from 0 to 2π, giving
us one period of the cycloid. Show that the length of one period of the cycloid
is eight times the radius of the circle.
Curves 11
Z t
(2.1) s(t) = |γ̇(u)| du
a
is the length of the image under γ(t) of the interval [a, t]. Taking the
derivative of this equation and noting that γ̇ 6= 0, we have
ds
= |γ̇(t)| > 0.
dt
Hence, if l is the length of the regular curve γ(t) over the closed interval
[a, b], then s(t) is a strictly increasing function from the interval [a, b] to the
interval [0, l], and thus the inverse correspondence [0, l] 3 s 7→ t(s) ∈ [a, b]
exists. By the inverse function theorem (Theorem A.1.5 in Appendix A.1),
t(s) is also smooth. Then we can parametrize γ by s as well:
γ(s) := γ(t(s)) (0 ≤ s ≤ l).
We call this s an arc-length parameter of the regular curve.
Up to now we have been using the dot “ ˙ ” to represent the derivative
with respect to t, but when we have an arc-length parameter s, we will
denote the derivative with respect to s by using a “ 0”. The purpose of this
change of notation is to help clarify when we are using an arc-length pa-
rameter. By the chain rule for the derivative of a composition of functions,
we have
dγ dγ dt γ̇(t)
(2.2) γ 0 (s) = = = .
ds dt ds |γ̇(t)|
Therefore we have |γ 0 (s)| = 1. In other words, a curve parametrized by
arc-length always has unit speed.
Conversely, suppose we have a curve γ(t) whose speed is always 1, that
is, |γ̇| = 1 holds. Then, by (2.1), the distance along the curve from some
starting point differs from t only by an additive constant. Thus, an arc-
length parameter means a parameter of unit speed, that is, the norm of the
velocity vector is 1.
March 29, 2017 9:45 ws-book9x6 World Scientific Book - 9in x 6in 9901-main page 12
Thinking of the curve as a road, with a car moving along it with respect
to time t, the car is moving at constant speed 1 when t is an arc-length
parameter, and a general parametrization corresponds to travel where speed
by 2409:40d2:a:8f73:10c2:8aff:fe61:88b7 on 03/10/25. Re-use and distribution is strictly not permitted, except for Open Access articles.
can vary.
For a curve γ(s) = (x(s), y(s)) parametrized by the arc-length s,
(2.3) e(s) := γ 0 (s) = (x0 (s), y 0 (s))
is a unit tangent vector of γ(s). Furthermore,
(2.4) n(s) := (−y 0 (s), x0 (s))
Differential Geometry of Curves and Surfaces Downloaded from www.worldscientific.com
is a unit vector perpendicular to e(s), on the left side of e(s) (see Fig. 2.1).
This is called the leftward unit normal vector of γ(s), at each point of γ(s),
with respect to the direction of travel given by s.
Curves 13
For each given choice of s, the curvature becomes a real number, and so
the curvature κ(s) is a function of s; hence, we can call it the curvature
function. When κ(s) is positive (negative), the curve is turning to the left
by 2409:40d2:a:8f73:10c2:8aff:fe61:88b7 on 03/10/25. Re-use and distribution is strictly not permitted, except for Open Access articles.
(right) near γ(s). For example, while driving a car, if the steering is not
turned at all, the car is moving straight forward on a path with curvature 0,
and when the steering wheel is turned toward the left or right, the curvature
of the path is positive or negative, respectively (see Fig. 2.1).
Considering e and n as column vectors, together they form a 2 × 2
Differential Geometry of Curves and Surfaces Downloaded from www.worldscientific.com
orthogonal matrix (e, n). Since this matrix has determinant 1, we have
that (see Problem 2 at the end of this section)
and we have
00 1 s s 1
γ (s) = − cos , − sin = n(s).
a a a a
Thus a leftward-turning circle of radius a (> 0) has constant curvature
κ(s) = 1/a. When the circle is rightward-turning and of radius a (> 0),
the curvature would be −1/a.
March 29, 2017 9:45 ws-book9x6 World Scientific Book - 9in x 6in 9901-main page 14
ally, so we will also need a formula for the curvature that applies when the
parameter is not necessarily arc-length. If t is any parameter for a regular
curve γ(t) = (x(t), y(t)), the curvature function κ(t) is
ẋÿ − ẏẍ det(γ̇, γ̈)
(2.7) κ(t) = = .
2
(ẋ + ẏ )2 3/2 |γ̇|3
Differential Geometry of Curves and Surfaces Downloaded from www.worldscientific.com
Proof of equation (2.7). Using (2.2) and the chain rule for differentiating
compositions of functions, we have
0
0 γ̇ 00 γ̈ 1
γ = , γ = + γ̇.
|γ̇| |γ̇|2 |γ̇|
Substituting these into (2.6) and using the properties of determinants, we
arrive at the desired equation.
The osculating circle. At a point on a regular curve, the line that comes
closest to “approximating” the curve near that point is the tangent line at
that point. Now, instead of a line, let us consider what circle would best
“approximate” the curve near a given point on the curve.
Let us take a point γ(s) on the curve where the curvature κ(s) is not
zero. We can then place a circle of radius 1/|κ(s)| so that it is tangent to
the curve at γ(s), and we can give the circle a parametrization so that its
parameter is increasing in the same direction as the s along γ(s). When
κ(s) > 0, we place the circle on the left-hand side of the curve, and when
March 29, 2017 9:45 ws-book9x6 World Scientific Book - 9in x 6in 9901-main page 15
Curves 15
κ(s) < 0, we place it on the right-hand side. We call this circle the oscu-
lating circle of the curve γ at the point γ(s); see Fig. 2.2. The osculating
circle has radius 1/|κ(s)| and center
by 2409:40d2:a:8f73:10c2:8aff:fe61:88b7 on 03/10/25. Re-use and distribution is strictly not permitted, except for Open Access articles.
1
γ(s) + n(s).
κ(s)
The locus of the center of osculating circles is called the evolute (or focal
curve or caustic) of γ(s). See Appendix B.1 for the construction of the
evolutes and relationships with the original curves.
Differential Geometry of Curves and Surfaces Downloaded from www.worldscientific.com
Fig. 2.2 The osculating circles (lighter shaded) of curves (darker shaded).
At a point of γ(s) where the curvature is 0, we say that the tangent line
is the osculating circle (a circle with infinite radius).
We give the osculating circle a pa-
rametrization in the same direction as
the original curve. Then by Example
2.1, we see that the original curve and
the osculating circle will both have the
same curvature at the point where they
touch tangentially. If one is driving a
car, and fixes the steering wheel in the
position it is in at one moment, the car
will start driving along a circle. That
circle is the osculating circle of the path
of the car at the moment the steering Fig. 2.3
wheel was held fixed.
Theorem 2.4 on page 16 will show us that the osculating circle really
is the circle that best approximates the original curve at each point. But
March 29, 2017 9:45 ws-book9x6 World Scientific Book - 9in x 6in 9901-main page 16
us further suppose that γ(t0 ) is the origin and γ̇(t0 ) points in the direction
of the x-axis, Then the normal vector at t = t0 on the left-hand side of
the curve will point in the direction of the y-axis. We can thus write the
tangent vector as γ̇(t0 ) = (a, 0) (a := ẋ(t0 ) > 0), see Fig. 2.3. Therefore,
by what we stated in Section 1 (on page 6), for values of t close to t0 (i.e.
Differential Geometry of Curves and Surfaces Downloaded from www.worldscientific.com
for a portion of the curve near the origin), the curve can be written as a
graph y = f (x). We use these facts in the next definition.
Definition 2.3. If two curves γ1 (t) and γ2 (u) intersect at a point P so
that their derivatives point in the same direction, then we say that the two
curves have first order contact at P. In this case, taking P to be the origin,
and the curves to be moving in the direction of the positive x-axis, and
taking the normal vector to be on the left-hand side so that it points in the
direction of the positive y-axis at P, the two curves γ1 (t) and γ2 (u) given
above can each be written as graphs of functions y = f (x) and y = g(x) so
that
df dg
f (0) = g(0) = 0, (0) = (0) = 0.
dx dx
Then, if we further have
d2 f d2 g
(0) = (0),
dx2 dx2
we say that the two curves have second order contact (see the left-hand side
of Fig. 2.4).
We can similarly define the notion of higher order contact. For example,
if the f (x) and g(x) in Definition 2.3 have the same first, second and third
order derivatives at 0, we say that the two corresponding curves γ1 and
γ2 have third order contact. The parabola y = x2 /2 and the catenary
y = cosh x − 1 have third order contact at the origin. (See Fig. 2.4.)
Since a tangent line of a curve makes first order contact with that curve,
we say it is a first order approximation to the curve. Using the same
phrasing, the next theorem tells us that the osculating circle is a second
order approximation to the curve.
Theorem 2.4. The osculating circle at a point γ(t0 ) of a regular curve
γ(t) makes second order contact with the curve at that point. Conversely,
any circle that makes second order contact with γ(t) at γ(t0 ) must be the
osculating circle at that point.
March 29, 2017 9:45 ws-book9x6 World Scientific Book - 9in x 6in 9901-main page 17
Curves 17
by 2409:40d2:a:8f73:10c2:8aff:fe61:88b7 on 03/10/25. Re-use and distribution is strictly not permitted, except for Open Access articles.
Differential Geometry of Curves and Surfaces Downloaded from www.worldscientific.com
Fig. 2.4
To test for the contact order of regular curves from the definition, one
has to express the curves as graphs. To avoid this procedure, the following
criteria is useful:
March 29, 2017 9:45 ws-book9x6 World Scientific Book - 9in x 6in 9901-main page 18
Proposition 2.5. Two regular curves γ1 (t) and γ2 (u) will make first order
(resp. second order) contact at a point P if and only if one can choose a
parameter u for γ2 so that at t = t0 and u = u0 ,
by 2409:40d2:a:8f73:10c2:8aff:fe61:88b7 on 03/10/25. Re-use and distribution is strictly not permitted, except for Open Access articles.
dγ1 dγ2
(2.12) γ1 (t0 ) = γ2 (u0 ) = P, (t0 ) = (u0 )
dt du
d2 γ1 d2 γ2
resp. additionally (t 0 ) = (u0 ) .
dt2 du2
Differential Geometry of Curves and Surfaces Downloaded from www.worldscientific.com
Proof. The ‘only if’ part is evident because a graph is a special case of a
parametrization.
We shall now show the ‘if’ part: We denote by “ ˙ ” differentiation with
respect to the parameter t, and “ 0 ” with respect to u (though the parameter
u is not necessarily arc-length, we use this notation to distinguish the two
variables t and u). For the first order contact, the conclusion is evident by
Definition 2.3. We show when the two curves have second order contact.
Set γ1 (t) = (x(t), y(t)), γ2 (u) = (x̃(u), ỹ(u)) and express the two curves
as graphs y = f (x), ỹ = g(x̃), respectively, as in Definition 2.3. Then
y(t) = f (x(t)) and ỹ(u) = g(x̃(u)) hold, and thus we have
df ẏ dg ỹ 0 d2 f ÿ ẋ − ẏẍ d2 g ỹ 00 x̃0 − ỹ 0 x̃00
= , = 0, = , = ,
dx ẋ dx x̃ dx2 ẋ3 dx 2 x̃03
by the chain rule. Hence
d2 f d2 g
(0) = (0)
dx2 dx2
if γ̇1 (t0 ) = γ20 (u0 ) and γ̈1 (t0 ) = γ200 (u0 ), that is, the two curves have second
order contact.
Curves 19
For example, if we set x = r cos θ, y = r sin θ for r > 0 and θ, then (x, y)
by 2409:40d2:a:8f73:10c2:8aff:fe61:88b7 on 03/10/25. Re-use and distribution is strictly not permitted, except for Open Access articles.
is a point on the plane with polar coordinate (r, θ). This correspondence is
regarded as a map from the rθ-plane to the xy-plane as
ϕ : (r, θ) 7−→ (x, y) = (r cos θ, r sin θ).
If one takes a region D := {(r, θ) | r > 0, −π < θ < π} of this mapping, then
ϕ(D) is the xy-plane excluding the negative part of the x-axis, and ϕ gives
Differential Geometry of Curves and Surfaces Downloaded from www.worldscientific.com
Proof. Writing ϕ as ϕ(x, y) = (ξ(x, y), η(x, y)), for the curve γ(t) =
(x(t), y(t)) in the region D we have the corresponding curve
γ̃(t) = ϕ ◦ γ(t) = (ξ(x(t), y(t)), η(x(t), y(t))).
Applying the chain rule,
dξ dη
= ξx ẋ + ξy ẏ, = ηx ẋ + ηy ẏ,
dt dt
d2 ξ
(2.13) = ξx ẍ + ξy ÿ + ξxx ẋ2 + 2ξxy ẋẏ + ξyy ẏ 2 ,
dt2
d2 η
= ηx ẍ + ηy ÿ + ηxx ẋ2 + 2ηxy ẋẏ + ηyy ẏ 2 .
dt2
March 29, 2017 9:45 ws-book9x6 World Scientific Book - 9in x 6in 9901-main page 20
Fig. 2.5 Diffeomorphism and contact of curves: Two regular curves in the xy-plane
having second order contact are mapped to curves in the ξη-plane having second order
contact, under the diffeomorphism (x, y) 7→ ϕ(x, y) = (ξ, η). Here, the dotted lines in
the right-hand figure are the images under ϕ of dotted lines parallel to coordinate axes
in the left-hand figure.
Assume γ1 (t) := (x1 (t), y1 (t)) and γ2 (u) := (x2 (u), y2 (u)) have first order
contact at P = γ1 (t0 ) = γ2 (u0 ) so that the parameters t and u satisfy
(2.12) in Proposition 2.5. We write γ̃1 (t) = (ξ1 (t), η1 (t)) and γ̃2 (u) =
(ξ2 (u), η2 (u)), that is,
Then we have
γ̃2 have first order contact at ϕ(P). The case of second order contact can
be obtained in a similar way.
March 29, 2017 9:45 ws-book9x6 World Scientific Book - 9in x 6in 9901-main page 21
Curves 21
A similar result can be stated for third or higher order contact as well
(see (2) in Problem 9 at the end of this section).
by 2409:40d2:a:8f73:10c2:8aff:fe61:88b7 on 03/10/25. Re-use and distribution is strictly not permitted, except for Open Access articles.
The Frenet formula. Given a regular curve γ(s) = (x(s), y(s)) param-
etrized by an arc-length parameter s, equation (2.5) can be rewritten as:
(2.14) x00 (s) = −κ(s) y 0 (s), y 00 (s) = κ(s) x0 (s).
From this and (2.3) and (2.4), we know that n0 (s) = −κ(s)e(s), so, together
with (2.5), we have that
(2.15) e0 (s) = κ(s)n(s), n0 (s) = −κ(s)e(s).
These two equations are called the Frenet formula.1
It is natural to ask, for a given function, whether there always exists a
regular curve whose curvature function is equal to the given function, and
whether it is unique if it exists. The following theorem is the answer to this
question.
Theorem 2.8 (The fundamental theorem for plane curves). For a smooth
function κ(s) defined for s in the interval 0 ≤ s ≤ l, there exists a regular
curve γ(s) with arc-length parameter s ∈ [0, l] and curvature κ(s). Further-
more, this curve is uniquely determined up to rotations and translations of
the plane.
1 Frenet, Jean Frédéric (1816–1900).
March 29, 2017 9:45 ws-book9x6 World Scientific Book - 9in x 6in 9901-main page 22
Proof. Existence of the curve seems natural when one imagines driving a
car, steering in accordance with the curvature function, while driving at
unit speed.
by 2409:40d2:a:8f73:10c2:8aff:fe61:88b7 on 03/10/25. Re-use and distribution is strictly not permitted, except for Open Access articles.
0 −κ
(2.16) F 0 = FΩ, Ω := .
κ 0
This equation can be regarded as a linear ordinary differential equation
of the unknown matrix-valued function F(s), where κ(s) is a given data
(cf. Appendix A.2). Then there exists a unique solution F(s) of (2.16)
satisfying the initial condition
1 0
F(0) := (e(0), n(0)) = I, I := ,
0 1
because of Theorem A.2.2 in Appendix A.2. For each s, F(s) is an or-
thogonal matrix with determinant 1. In fact, since Ω is a skew-symmetric
matrix, Proposition A.3.9 in Appendix A.3 yields that
T
(2.17) (FF T )0 = F 0 F T + FF 0 = FΩF T + FΩ T F T = O,
where F T is the transposition of F, and O denotes the 2 × 2 zero matrix.
T
Hence F(s)F(s) does not depend on s, i.e., is a constant matrix:
T T
F(s)F(s) = F(0)F(0) = I.
This implies that F(s) is an orthogonal matrix for each s. Then the de-
terminant of F(s) is ±1 (cf. (A.3.11) in Appendix A.3), and by the fact
F(0) = I, we can conclude that the determinant of F(s) is 1 by continuity
of det F(s). Thus F(s) is an orthogonal matrix of determinant 1, and hence
n(s) points leftward with respect to e(s) (cf. (3) of Problem 2).
Let
Z s
γ(s) := e(t) dt.
0
Then γ(0) = 0 holds, and γ(s) is the desired curve. In fact, e(s) is the
unit tangent vector of γ(s) since γ 0 (s) = e(s). Moreover, since F(s) is
an orthogonal matrix of determinant 1, n(s) is the left-hand unit normal
vector. Then equation (2.16) for F(s) = (e(s), n(s)) is equivalent to the
Frenet formula for γ(s), and then κ(s) is a curvature function.
March 29, 2017 9:45 ws-book9x6 World Scientific Book - 9in x 6in 9901-main page 23
Curves 23
0 1 0
(2.18) γ(0) = Γ (0) = , e(0) = E(0) = , n(0) = N (0) =
0 0 1
by applying rotations and translations, without loss of generality. Here,
E(s) and N (s) are the unit tangent vector and the left-hand unit normal
Differential Geometry of Curves and Surfaces Downloaded from www.worldscientific.com
Remark 2.9. For a given curvature function κ(s), the desired regular curve
γ(s) can be obtained explicitly as follows:
Z s Z t Z t
(2.19) γ(s) = cos κ(u) du , sin κ(u) du dt.
0 0 0
The four-vertex theorem for convex curves. At the end of this sec-
tion we give an example of an application of the Frenet formula.
In Section 1, a curve was said to be closed when traveling along the
curve would bring you back to where you started. More precisely, closed
curves are defined as follows: If a curve γ(t) defined on the entire real line
R has a period l (> 0), that is,
(2.20) γ(t + l) = γ(t) (t ∈ R),
then the restriction of γ to the closed interval [0, l] is said to be a closed
curve. In other words, a curve γ : [a, b] → R2 is said to be a closed curve
March 29, 2017 9:45 ws-book9x6 World Scientific Book - 9in x 6in 9901-main page 24
at s = s0 if there exists an interval [a, b] such that s0 ∈ (a, b), and κ(s0 ) is the maximum
(resp. minimum) of κ(s) on [a, b], and κ(a), κ(b) < κ(s0 ) (resp. κ(a), κ(b) > κ(s0 )).
March 29, 2017 9:45 ws-book9x6 World Scientific Book - 9in x 6in 9901-main page 25
Curves 25
by 2409:40d2:a:8f73:10c2:8aff:fe61:88b7 on 03/10/25. Re-use and distribution is strictly not permitted, except for Open Access articles.
2 vertices 4 vertices
Differential Geometry of Curves and Surfaces Downloaded from www.worldscientific.com
Theorem 2.11. There exist at least four vertices on any strictly convex
curve that is not a round circle.
Then the closed curve would be a circle (see Corollary 2.10), contradicting
our assumption.
Let
(2.21) ax + by + c = 0 (a, b, c are constants)
be the equation for the straight line from γ(0) to γ(v). As the curve is con-
Differential Geometry of Curves and Surfaces Downloaded from www.worldscientific.com
vex, the line segment between these two points must lie in the interior region
of the curve (cf. Definition 1.5), and then the curve γ(s) = (x(s), y(s)) is
split into two parts by the line segment, one part where 0 ≤ s ≤ v lying to
one side, and the other part where v ≤ s ≤ l lying to the other side. The
function ax(s) + by(s) + c is non-negative on one of the intervals [0, v] and
[v, l], and is non-positive on the other. Hence κ0 (s)(ax(s) + by(s) + c) does
not change sign on 0 ≤ s ≤ l, and is not identically zero. Then the integral
Z l
J := κ0 (s) (ax(s) + by(s) + c) ds
0
will not be zero. On the other hand, integrating by parts and using (2.14)
and (2.20), we have
Z l
J =− κ(s) (ax0 (s) + by 0 (s)) ds
0
Z l
s=l
= (−ay 00 (s) + bx00 (s))ds = [−ay 0 (s) + bx0 (s)]s=0 = 0.
0
The last equality above holds because the curve is closed. This gives us a
contradiction, proving the result.
Exercises 2
1 An n × n matrix A is called an orthogonal matrix if AT is the inverse matrix
of A, where AT denotes the transposition of A (cf. Appendix A.3). When
setting n = 2, show that a 2 × 2 orthogonal matrix can be written in the form
cos α − sin α
(when the determinant is +1),
sin α cos α
cos α sin α
(when the determinant is −1),
sin α − cos α
Curves 27
2 (1) For a unit vector e in the plane, let n be the unit vector perpendicular
to e. Show that the matrix (e, n) (the 2 × 2 matrix whose first column
is e and second is n) is an orthogonal matrix (cf. Appendix A.3).
by 2409:40d2:a:8f73:10c2:8aff:fe61:88b7 on 03/10/25. Re-use and distribution is strictly not permitted, except for Open Access articles.
s2
γ(s) =γ(0) + se(0) + κ(0)n(0)
2
s3
+ (−κ(0)2 e(0) + κ0 (0)n(0)) + o(s3 ).
6
Here o is Landau’s symbol as in Appendix A.1.
∗
9 (1) Show that two regular curves γ1 (t), γ2 (u) have third order contact at P
if and only if one can choose a parametrization of γ2 such that γ1 (t0 ) =
γ2 (u0 ) = P and
dγ1 dγ2 d2 γ 1 d2 γ 2 d3 γ 1 d3 γ 2
(t0 ) = (u0 ), 2
(t0 ) = 2
(u0 ), 3
(t0 ) = (u0 )
dt du dt du dt du3
March 29, 2017 9:45 ws-book9x6 World Scientific Book - 9in x 6in 9901-main page 28
hold.
(2) Let ϕ : D → ϕ(D) be a diffeomorphism from a region D to ϕ(D) in the
plane. Show that for two curves γ1 and γ2 in D having third order contact
by 2409:40d2:a:8f73:10c2:8aff:fe61:88b7 on 03/10/25. Re-use and distribution is strictly not permitted, except for Open Access articles.
3. Closed curves*
Here we investigate the properties of closed regular curves that do not
change under continuous deformations.
Given two closed regular curves, if there exists a smooth deformation
from one to the other so that, at each moment of the deformation, the curve
at that moment has a nowhere vanishing velocity vector, then we say that
the two regular curves are equivalent or regular homotopic to each other.
More precisely, we define two closed regular curves to be regular homotopic
as follows: Changing parameter if necessary, two closed regular curves γ1 (s)
and γ2 (s) can be parametrized by the same parameter s (0 ≤ s ≤ l). Then,
γ1 and γ2 are regular homotopic if and only if there exists a family of closed
curves σt : [0, l] → R2 (0 ≤ t ≤ 1) so that:
(i) For each t, σt (s) (0 ≤ s ≤ l) is a smooth closed curve whose velocity
vector dσds (s) never vanishes.
t
(ii) σ0 (s) = γ1 (s) and σ1 (s) = γ2 (s). In other words, when t moves from
0 to 1, the curves σt move from γ1 to γ2 .
(iii) σt (s) is continuous with respect to both t and s.
Here we do not assume that s is an arc-length parameter.
This deformation gives an equivalence relation for closed regular curves,
and every closed regular curve lies in one of the classes shown in Fig. 3.1.
In order to check that this is so, we need to introduce the notion of rotation
index for closed regular curves.
Curves 29
by 2409:40d2:a:8f73:10c2:8aff:fe61:88b7 on 03/10/25. Re-use and distribution is strictly not permitted, except for Open Access articles.
iγ = 0 iγ = ±1 iγ = ±2 iγ = ±3 ...
where
Z s
(3.1) θ(s) := κ(u) du.
0
This gives γ 0 (s) = (cos θ(s), sin θ(s)), and so θ(s) gives us the angle between
γ 0 (s) and γ 0 (0) = (1, 0). Here, because we have the condition (2.20) for
closed curves, which implies γ 0 (l) = γ 0 (0), it follows that θ(l) − θ(0) is an
integer multiple of 2π. Therefore iγ is an integer.
Curves 31
by 2409:40d2:a:8f73:10c2:8aff:fe61:88b7 on 03/10/25. Re-use and distribution is strictly not permitted, except for Open Access articles.
Differential Geometry of Curves and Surfaces Downloaded from www.worldscientific.com
Fig. 3.3
γ 0 (s) (when s = t),
0
−γ (0) (when s = 0 and t = l),
w(s, t) :=
γ(t) − γ(s)
(otherwise),
|γ(t) − γ(s)|
this w is a continuous map from D to the unit circle S 1 . Then there exists
a unique continuous function Θ on D such that
(3.3) w(s, t) = (cos Θ(s, t), sin Θ(s, t)), Θ(0, 0) = 0.
The function Θ(s, t) represents the angle of the vector w(s, t) with (1, 0).
Since the angle has the ambiguity of integer multiples of 2π, it cannot
be determined uniquely, in general. But if we allow the angle Θ(s, t) to
take arbitrary real number values beyond the interval [0, 2π], and choose
values Θ(s, t) which vary continuously in (s, t), then the function Θ(s, t)
satisfies (3.3). (The existence of this Θ is usually shown using the notion
of covering spaces, and the fact that D is simply connected (cf. [8, 35]).
However, noticing that w(s, t) is of class C 1 , one can show the existence of
the required function Θ without using the notion of covering spaces. See
Problem 4.)
Since γ is a closed curve, γ(l) = γ(0) holds. So we have w(s, l) =
−w(0, s), and we have that Θ(s, l) − Θ(0, s) = π × (an odd integer). Since
the function Θ(s, l) − Θ(0, s) of s is continuous, there exists an integer n
independent of s so that we can write
(3.4) Θ(s, l) − Θ(0, s) = (2n + 1)π (0 ≤ s ≤ l).
March 29, 2017 9:45 ws-book9x6 World Scientific Book - 9in x 6in 9901-main page 32
Proof. The necessity of the equality of the rotation indices is clear. Con-
versely, suppose that the two closed regular curves γ1 and γ2 have the same
rotation index:
iγ1 = iγ2 = m.
We will construct a family σt (s) of curves satisfying the conditions (i)–(iii)
for regular homotopy. Because homotheties do not change the rotation
index, we may assume without loss of generality that both curves have
length 1 and have arc-length parametrizations γ1 (s), γ2 (s) (0 ≤ s ≤ 1).
Letting κ1 (s) and κ2 (s) be the respective curvatures of γ1 and γ2 , from
2 Whitney, Hassler (1907–1989).
March 29, 2017 9:45 ws-book9x6 World Scientific Book - 9in x 6in 9901-main page 33
Curves 33
equation (2.19) in the proof of the fundamental theorem for plane curves
(Theorem 2.8), we see that we may assume γj (0) = 0 and
Z s Z s
by 2409:40d2:a:8f73:10c2:8aff:fe61:88b7 on 03/10/25. Re-use and distribution is strictly not permitted, except for Open Access articles.
Here, we set
(3.8) ϕt (s) := (1 − t)θ1 (s) + tθ2 (s) (0 ≤ t ≤ 1).
By (3.7), it holds that ϕt (s + 1) = ϕt (s) + 2πm. Then defining
v t (s) := (cos ϕt (s), sin ϕt (s)),
Z 1
ṽ t (s) := v t (s) − v t (u) du,
0
Z s
σt (s) := ṽ t (u) du,
0
one can show that, for each t, v t (s), ṽ t (s) and σt (s) are periodic in s with
period 1, and then σt is a closed curve. Because γj (0) = 0 (j = 1, 2), we
have ṽ t = v t for t = 0, 1, and it follows that σ0 = γ1 , σ1 = γ2 . That is,
σt is a deformation from γ1 to γ2 . In other words, σt (s) satisfies the first
part of the condition (i) and the conditions (ii) and (iii) of the definition of
regular homotopy. So it is sufficient to show that σt (s) satisfies the second
part of the condition (i), that is, ṽ t = dσt /ds does not vanish anywhere in
the deformation. By definition, we have
Z 1
dσt (s)
= ṽ t (s) = v t (s) − v t (u) du,
ds 0
and the triangle inequality for integrals (cf. Theorem A.1.4 in Appendix
A.1) gives
Z 1 Z 1
(3.9) v t (u) du ≤ |v t (u)| du = 1.
0 0
In order for equality to hold, v t (s) (0 ≤ s ≤ 1) must be a constant vector.
(Here, we used the fact that |v t | = 1.)
When the rotation index m is not 0, (3.8) implies that ϕt moves from
0 to 2πm when s moves 0 to 1, and then v t cannot be a constant vector.
Consequently, equality does not hold in (3.9), implying
Z 1
dσt
≥ |v t | − v t (u) du > |v t | − 1 = 0,
ds 0
March 29, 2017 9:45 ws-book9x6 World Scientific Book - 9in x 6in 9901-main page 34
tθ2 (s) is zero for all s. In particular, two functions θ1 and θ2 differ by only
constant multiplication. If the curvatures of γ1 , γ2 are identically zero, they
cannot be closed curves. Hence θ1 and θ2 are not constant functions. If θ1
is proportional to θ2 , translating the parameter s of γ2 if necessary, one can
change θ2 with an additional constant, and then cause θ1 and θ2 not to be
Differential Geometry of Curves and Surfaces Downloaded from www.worldscientific.com
By Theorem 3.3, we can find the rotation index of a given curve without
computation. In particular, for the curves in Fig. 3.1, the leftmost curve
is regular homotopic to the lemniscate and so has rotation index 0, by
Problem 1. Also, the other curves in that figure are regular homotopic to
curves that wrap finitely many times about a circle, so their rotation indices
can be found as well. Also, for closed regular curves γ(s) (0 ≤ s ≤ l) that
are C 1 , we have existence of a continuous function θ : [0, l] → R so that
γ 0 (s) = (cos θ(s), sin θ(s)) holds, and so we can define the rotation index by
2πiγ := θ(l) − θ(0). Then, by adjusting the proofs of Theorems 3.2 and 3.3
a bit, we can reprove those results for the case of C 1 -regular curves.
Proof. A circle turning right (resp. left) has the rotation index 1 (resp. −1).
On the other hand, a given simple closed curve γ has the rotation index 1
or −1 by Theorem 3.2. Hence by Theorem 3.3, it is regular homotopic to
a circle.
Curves 35
by 2409:40d2:a:8f73:10c2:8aff:fe61:88b7 on 03/10/25. Re-use and distribution is strictly not permitted, except for Open Access articles.
Fig. 3.4
Differential Geometry of Curves and Surfaces Downloaded from www.worldscientific.com
Fig. 3.5
can always find such a point.) Then the rotation index iγ of γ is obtained
by
X
iγ = 1 + sgnP,γ (Q).
Q∈γ
Differential Geometry of Curves and Surfaces Downloaded from www.worldscientific.com
Proof. By the Jordan closed curve theorem (cf. page 9), the simple closed
curve σ divides the plane into two regions, the interior and the exterior. For
each adjacent pair of self-intersections, if the curve γ travels from the inte-
rior (resp. exterior) to the exterior (resp. interior) at the first intersection,
then at the next intersection, it travels from the exterior (resp. interior) to
the interior (resp. exterior). Noticing that the number of intersections is
even, the conclusion follows.
Curves 37
by 2409:40d2:a:8f73:10c2:8aff:fe61:88b7 on 03/10/25. Re-use and distribution is strictly not permitted, except for Open Access articles.
Differential Geometry of Curves and Surfaces Downloaded from www.worldscientific.com
Fig. 3.6
Q1 . In fact, if there exists such a self-intersection, the point has the same
property as Q1 , which contradicts the way we chose Q1 . Moreover, γ1 does
not pass through the base point P, because the curve passes through each
self-intersection twice before returning to P.
On the other hand, let γ2 be a closed regular curve which is on the
opposite side of γ1 where we rounded the vertical angle at Q1 . The integrals
of the curvature on the rounded angles are equal to the increment of the
angles at the rounded parts (cf. (3.1)). Since these angles at the rounded
parts are equal, these increments on γ1 and γ2 cancel, which implies iγ =
iγ1 + iγ2 . Since the rotation index of the simple closed regular curve γ1 is
the sign of the self-intersection Q1 , by the inductive assumption, we have
iγ = iγ1 + iγ2 = sgnP,γ (Q1 ) + iγ2
X
= sgnP,γ (Q1 ) + 1 + sgnP,γ2 (Q).
Q∈γ2
By Lemma 3.6, the sum of the signs of intersections of γ2 with γ1 vanishes.
So we have X
sgnγ2 ,γ1 (R) = 0.
R∈γ1 ∩γ2
Here, γ1 does not pass through self-intersections between P and Q1 . Thus
for each intersection R of γ2 with γ1 , it holds that sgnP,γ (Q) = sgnγ2 ,γ1 (R).
Hence
X X X
sgnP,γ (Q) = sgnP,γ (Q1 ) + sgnγ2 ,γ1 (R) + sgnP,γ2 (S)
Q∈γ R∈γ1 ∩γ2 S∈γ2
X
= sgnP,γ (Q1 ) + sgnP,γ2 (S) = iγ − 1,
S∈γ2
which proves the theorem.
March 29, 2017 9:45 ws-book9x6 World Scientific Book - 9in x 6in 9901-main page 38
Fig. 3.7
The left side of Fig. 3.7 has 3 self-intersections. Taking P as a base point,
the signs of these self-intersections are +, −, +, in order of appearance.
Thus by Theorem 3.5, the rotation index is 1 + (1 − 1 + 1) = 2. On the
other hand, the rotation index of the right side figure is 2, because the
sign of the self-intersection is +. Therefore these two curves are regular
homotopic, by Theorem 3.3. One can verify this fact by drawing pictures
of a smooth deformation.
Exercises 3
1 Compute the rotation indices of the ellipse and the lemniscate from the defi-
nition of the rotation index.
2∗ Assume a C ∞ -function ϕ(t) defined on R satisfies ϕ(0) = 0. If we set
(
ϕ(t)/t (when t 6= 0),
ψ(t) :=
ϕ̇(0) (when t = 0),
Curves 39
γ(t) − γ(s) |t − s|
(s − l < t < s + l, s 6= t),
by 2409:40d2:a:8f73:10c2:8aff:fe61:88b7 on 03/10/25. Re-use and distribution is strictly not permitted, except for Open Access articles.
t−s |γ(t) − γ(s)|
0
γ 0 (s) = γ (s)
(s = t),
w̃(s, t) = |γ 0 (s)|
γ(t) − γ(s) |t − (s + l)|
− (s + l < t < s + 2l, s 6= t),
t − (s + l) |γ(t) − γ(s)|
Differential Geometry of Curves and Surfaces Downloaded from www.worldscientific.com
0
−γ (s) (s + l = t),
show that w̃ is C ∞ on the domain {(s, t) | s − l < t < s + 2l}, and the
restriction of w̃ to D = {(s, t) | 0 ≤ s ≤ t ≤ l} coincides with the w in the
proof of Theorem 3.2.
4∗ For a positive constant l, let w : D → R2 be a C ∞ -map w : D → R2 defined
on a closed domain D = {(s, t) | 0 ≤ s ≤ t ≤ l} of the st-plane satisfying
|w(s, t)| = 1 and w(0, 0) = (1, 0). Show, in the following way, that there exists
a continuous function Θ on D satisfying w(s, t) = (cos Θ(s, t), sin Θ(s, t)) and
Θ(0, 0) = 0.
(3) Let
Z s Z t
Θ(s, t) = det(w(u, t), w1 (u, t)) du + det(w(0, v), w2 (0, v)) dv
0
0
∂w ∂w
w1 := , w2 :=
∂s ∂t
4. Geometry of spirals*
We first introduce two representative examples of spiral curves, the
by 2409:40d2:a:8f73:10c2:8aff:fe61:88b7 on 03/10/25. Re-use and distribution is strictly not permitted, except for Open Access articles.
The Archimedean spiral. Taking polar coordinates (r, θ) for the plane,
the curve satisfying
is called the Archimedean spiral (see Fig. 4.1, left side). If one were to
imagine a horizontal plane rotating about one fixed point at a constant
rate, and if one were to take a nail and press it gently onto that plane by
starting at the fixed point and moving at constant speed out in one un-
changing direction, one would scratch out such a spiral. In ancient Greece,
Archimedes1 investigated the properties of this spiral curve (see Problem 1
at the end of this section).
Fig. 4.1
The logarithmic spiral. In polar coordinates (r, θ), the curve given by
r = aθ (a > 1 is a constant, θ ∈ R)
is called the logarithmic spiral. When drawing a half-line out from the
origin, the half-line will always intersect the spiral at the same angle, re-
gardless of choice of direction of the half-line (see the right-hand side of
Fig. 4.1). One can find the logarithmic spiral in nature, for example, in
spiral shells or in the shape of galaxies. Similarity transformation of the
1 Archimedes (B.C. 287–212).
March 29, 2017 9:45 ws-book9x6 World Scientific Book - 9in x 6in 9901-main page 41
Curves 41
Fig. 4.2
the curvature function κ(s) is increasing. Here, consider the curve γ̃(s) :=
γ(l − s) (0 ≤ s ≤ l) obtained by changing the direction of γ(s). Then by
(2.6) on page 13, the curvature function κ̃(s) of γ̃(s) is
by 2409:40d2:a:8f73:10c2:8aff:fe61:88b7 on 03/10/25. Re-use and distribution is strictly not permitted, except for Open Access articles.
Theorem 4.1. Let γ(s) be a positive (resp. negative) spiral, and let the
osculating circle at each point of the spiral have the same orientation as the
spiral. Then γ(s) tangentially crosses the osculating circle from the circle’s
right side to its left (resp. left to right). In particular, if γ is positively curved
at s, then γ moves from outside (resp. inside) to inside (resp. outside) of
the circle.
Proof. A driver traveling along a positive (resp. negative) spiral keeps turn-
ing the wheel to the left (resp. right). If he would stop turning the wheel,
then the car follows the osculating circle at the point. Then one can see
intuitively that the trajectory of the car crosses the circle from right to left
(resp. left to right).
We give a rigorous proof: For a given positive spiral γ(s) and a point
γ(s0 ) of the spiral, choose an xy-coordinate system such that the point γ(s0 )
is the origin, the unit tangent vector e(s0 ) := γ 0 (s0 ) is the direction of the
x-axis, and the direction of the unit normal vector n(s0 ) is the direction
of the y-axis. Then the curve can be expressed as a graph y = f (x) (cf.
page 19). In particular, since f (0) = 0 and f˙(0) := dx df
(0) = 0 hold, (2.8)
March 29, 2017 9:45 ws-book9x6 World Scientific Book - 9in x 6in 9901-main page 43
Curves 43
implies that the curvature and the derivative of the curvature at the origin
are expressed as
by 2409:40d2:a:8f73:10c2:8aff:fe61:88b7 on 03/10/25. Re-use and distribution is strictly not permitted, except for Open Access articles.
f¨(0) ...
κ(0) = = f¨(0), κ̇(0) = f (0),
(1 + f˙(0)2 )3/2
where the dot means the derivative with respect to x. Then by Taylor’s
formula (Theorem A.1.1 in Appendix A.1),
1 1
κ(0)x2 + κ̇(0)x3 + o(x3 )
Differential Geometry of Curves and Surfaces Downloaded from www.worldscientific.com
(4.1) f (x) =
2 6
holds, where o(x3 ) is Landau’s symbol. On the other hand, the osculating
circle of the curve can be expressed as a graph y = g(x) near the origin,
and we have
1
g(x) = κ(0)x2 + o(x3 )
2
by replacing f in (4.1) with g, since the curvature of the circle y = g(x) is
constant. Hence we have
1
f (x) − g(x) = κ̇(0)x3 + o(x3 ).
6
Here, κ̇(0) > 0, because the graph f (x) is a positive spiral. Then, as x
increases, f (x) − g(x) changes sign from negative to positive at the origin.
This implies that γ(s) tangentially crosses the osculating circle from right
to left.
A translation z 7→ z + d (d ∈ C),
A rotation z 7→ eiθ z (θ ∈ R),
(4.3)
A similarity transformation z 7→ rz (r ∈ R \ {0}),
Conjugation composed with inversion z 7→ 1/z.
Here we explain inversion: Let C be a circle of radius ρ in the plane centered
Differential Geometry of Curves and Surfaces Downloaded from www.worldscientific.com
at the origin O. Take a map which maps each point P on the plane different
from O to the point Q on the half-line starting at O passing through P
such that OP · OQ = ρ2 . Such a map is called inversion with respect to the
circle C. The inversion on the complex plane C with respect to the unit
circle centered at the origin is the map z 7→ 1/z̄, where z̄ is the complex
conjugation of z.
Theorem 4.2. A Möbius transformation maps each positive spiral to a
positive spiral, and maps each negative spiral to a negative spiral. Further-
more, the osculating circles of a regular curve are mapped to the osculating
circles at the corresponding points of the image curve.
For example, by the transformation z 7→ 1/z, the logarithmic spiral is
mapped to a curve congruent to the original spiral. To prove Theorem 4.2,
we prepare the following lemma.
Lemma 4.3. A Möbius transformation maps circles to circles. Moreover,
the region to the left-hand side of a circle will be mapped to the left-hand
side of the image circle. Here the term “circles” includes straight lines,
which can be regarded as circles with infinite radius.
We used the terms “left-hand side” and “right-hand side” of the circle,
instead of the interior and exterior, since a Möbius transformation may
map the interior of the circle to the exterior of the image circle.
Proof. The conclusion is evident for the first three transformations of (4.3).
So we shall prove the lemma for the transformation z 7→ 1/z. A circle on
the xy-plane is expressed as
(4.4) a(x2 + y 2 ) + 2bx + 2cy + d = 0 (a, b, c, d ∈ R, b2 + c2 − ad > 0)
in general. In particular, when a = 0, (4.4) represents a straight line. It is
sufficient to prove that this circle is mapped to a circle by the transformation
z 7→ 1/z. Letting z = x + iy, (4.4) is rewritten as
a|z|2 + (b − ic)z + (b + ic)z̄ + d = 0 (z̄ := x − iy).
March 29, 2017 9:45 ws-book9x6 World Scientific Book - 9in x 6in 9901-main page 45
Curves 45
(A := d, B := b, C := −c, D := a).
In particular, since B 2 + C 2 − AD = b2 + c2 − ad > 0, the image is also
a circle. Note that the center of the original circle may not be mapped to
the center of the image circle.
Now we prove the latter conclusion. Assume a point P is mapped to a
Differential Geometry of Curves and Surfaces Downloaded from www.worldscientific.com
Fig. 4.4
Curves 47
the curvature function is constant (Corollary 2.10 in Section 2), this remains
so under Möbius transformations. By Theorem 4.2 and this fact, the prop-
erty of having a non-decreasing and non-increasing curvature function is
preserved by Möbius transformations. In particular, T (P) is the minimum
of the curvature function of the curve T ◦ γ. Therefore the curvature of
Differential Geometry of Curves and Surfaces Downloaded from www.worldscientific.com
Ds ⊂ Dt ,
where Ds is the closure of Ds , that is, the union of Ds and its boundary
∂Ds .
Curves 49
Exercises 4
1 Prove that the area bounded by the Archimedean spiral r = aθ and the half-
lines θ = α, θ = β (0 < β − α < 2π) is a2 (β 3 − α3 )/6. This fact was discovered
by Archimedes himself.
2 Show that a similarity expansion of the logarithmic spiral with respect to the
origin coincides with the original spiral up to the rotation about the origin.
Moreover, any half-line starting at the origin will make a single constant angle
with the logarithmic spiral, independent of choice of direction of the half-line
and also of choice of intersection point (see Fig. 4.1).
3 Explain, using the analogy of driving a car, that the positivity and negativity
of spirals do not depend on the orientation of curves.
4 Find the curvature functions of the Archimedean spirals and the logarithmic
spirals.
5 (1) Show that the number of vertices of the hyperbola x2 − y 2 = 1/a2 is 2.
Here a is a positive constant.
(2) Applying the Möbius transformation z 7→ 1/z (z = x + iy), show that the
lemniscate in Examples 1.1 and 1.3 is mapped to the hyperbola x2 − y 2 =
1/a2 . (In this case, the origin of the lemniscate is mapped to infinity.)
(3) Show that the lemniscate has exactly two vertices.
6 Show that any Möbius transformation can be expressed as a composition of
the four transformations in (4.3).
7∗ Show that the locus of the centers of the osculating circles of the logarithmic
spiral (i.e., the evolute, cf. Appendix B.1) is again the logarithmic spiral.
March 29, 2017 9:45 ws-book9x6 World Scientific Book - 9in x 6in 9901-main page 50
5. Space curves
In this section, we consider curves in 3-dimensional Euclidean space, or
by 2409:40d2:a:8f73:10c2:8aff:fe61:88b7 on 03/10/25. Re-use and distribution is strictly not permitted, except for Open Access articles.
more briefly “space curves”. To do this, we will need the vector product
(or cross product) of two vectors in space (cf. Appendix A.3).
Curves 51
is a unit vector perpendicular to both e(s) and n(s), and is called the
binormal vector. For each value of s,
{e(s), n(s), b(s)}
is a positively oriented orthonormal basis for R3 , that is, the determinant
of the 3 × 3-matrix (e(s), n(s), b(s)) is equal to 1 (cf. Appendix A.3).
For each s, the plane containing the
point γ(s) and spanned by e(s) and n(s)
is called the osculating plane of the reg-
ular curve γ at s (Fig. 5.1). In addi-
tion, the plane containing the point γ(s)
and spanned by n(s) and b(s) (resp.
e(s) and b(s)) is called the normal plane
(resp. the rectifying plane) of the curve
γ at s (Fig. 5.2).
Here, we define the torsion of the
curve as
Fig. 5.1 The osculating plane.
(5.1) τ (s) := −b0 (s) · n(s),
which measures how quickly the curve pulls away from the osculating plane
(cf. Problem 6 at the end of this section).
If two regular curves differ only by orientation preserving congruences
(cf. page 203 in Appendix A.3), then their curvatures and torsions coincide
(cf. Problem 1).
Example 5.1. The curve parametrized by
γ(t) = (a cos t, a sin t, bt)
is called a helix √
(Fig. 5.3). Here a and b are both non-zero constants.
Because |γ̇(t)| = a2 + b2 , this curve has arc-length parameter
p
s = t a 2 + b2 ,
March 29, 2017 9:45 ws-book9x6 World Scientific Book - 9in x 6in 9901-main page 52
and reparametrized as
s s bs p
γ(s) = a cos , a sin , c := a2 + b2
c c c
by the arc-length parameter s. Then we have
1 s s
e(s) = γ 0 (s) = −a sin , a cos , b ,
c c c
1 s s
e0 (s) = γ 00 (s) = 2 −a cos , −a sin , 0
c c c
which imply that the curvature κ and normal vector n are
e0 (s)
|a| |a| a s s
κ = |γ 00 (s)| = 2 = 2 , n = = − cos , − sin , 0 ,
c a + b2 |e0 (s)| |a| c c
March 29, 2017 9:45 ws-book9x6 World Scientific Book - 9in x 6in 9901-main page 53
Curves 53
respectively. We then have that the binormal vector b and torsion τ are
a 1 s s
b=e×n= b sin , −b cos , a ,
|a| c
by 2409:40d2:a:8f73:10c2:8aff:fe61:88b7 on 03/10/25. Re-use and distribution is strictly not permitted, except for Open Access articles.
c c
b b
τ = −b0 · n = 2 = 2 .
c a + b2
In particular, both the curvature and torsion of a helix are constant.
etrized by arc-length such that κ(s) > 0 for each s. Since the unit tangent
vector e(s), the principal normal vector n(s), and the binormal vector b(s)
form an orthonormal basis for R3 , any vector in R3 can be written as a
linear combination of them. Therefore n0 and b0 can be written as
dn
(5.2) n0 = = α1 e + α2 n + α3 b,
ds
db
(5.3) b0 = = β1 e + β2 n + β3 b,
ds
where αj , βj (j = 1, 2, 3) are real-valued functions depending on s.
Taking inner products of both sides of equation (5.2) with e, n and b,
we find that
α1 = n0 · e = (n · e)0 − n · e0 = −κn · n = −κ,
1
α2 = n0 · n = (n · n)0 = 0,
2
α3 = n0 · b = (n · b)0 − n · b0 = τ.
Again, taking inner products of both sides of equation (5.3) with e, n and
b, we have
β1 = b0 · e = (b · e)0 − b · e0 = −κb · n = 0,
β2 = b0 · n = −τ,
1
β3 = b0 · b = (b · b)0 = 0.
2
Substituting these into (5.2), (5.3) (together with e0 = κn), we have
0
e (s) = κ(s) n(s),
(5.4) n0 (s) = −κ(s) e(s) + τ (s) b(s),
0
b (s) = − τ (s) n(s).
We call these the Frenet-Serret 1 formula.
1 Serret, Joseph Alfred (1819–1885). See page 21 for Frenet.
March 29, 2017 9:45 ws-book9x6 World Scientific Book - 9in x 6in 9901-main page 54
Take the 3 × 3-matrix F(s) := (e(s), n(s), b(s)), where e(s), n(s) and
b(s) are considered as column vectors. Then F(s) is an orthogonal matrix
with determinant 1, since {e(s), n(s), b(s)} forms a positive orthonormal
by 2409:40d2:a:8f73:10c2:8aff:fe61:88b7 on 03/10/25. Re-use and distribution is strictly not permitted, except for Open Access articles.
basis (a right-handed system, see Appendix A.3) for each s. Using this, the
Frenet-Serret formula (5.4) is reformulated as
0 −κ 0
dF
(5.5) = FΩ, Ω := κ 0 −τ .
ds
0 τ 0
Differential Geometry of Curves and Surfaces Downloaded from www.worldscientific.com
Curves 55
differential equation
0 −κ 0
F 0 = FΩ,
by 2409:40d2:a:8f73:10c2:8aff:fe61:88b7 on 03/10/25. Re-use and distribution is strictly not permitted, except for Open Access articles.
(5.8) F(a) = I, Ω = κ 0 −τ
0 τ 0
whose unknown is the function F(s) taking values in the set of 3 × 3-
matrices, where I is the 3 × 3-identity matrix. Since this equation is linear
with respect to the nine components of F, by the fundamental theorem
Differential Geometry of Curves and Surfaces Downloaded from www.worldscientific.com
Then γ(s) is the desired curve. The uniqueness part and final part of the
proof are left as the exercises for readers.
Exercises 5
1 Show that the curvature and torsion of a space curve are invariant under ori-
entation preserving congruences. Namely, let Φ be an orientation preserving
congruence (cf. page 203 in Appendix A.3). Then the curvature and tor-
sion for a curve γ(s) and those for Φ ◦ γ(s) coincide. Note that whereas the
curvature is invariant under general congruence, the torsion may change sign.
2 Verify (5.6) and (5.7).
March 29, 2017 9:45 ws-book9x6 World Scientific Book - 9in x 6in 9901-main page 56
s2
γ(s) = γ(0) + se(0) + κ(0)n(0)
2
Differential Geometry of Curves and Surfaces Downloaded from www.worldscientific.com
s3
+ (−κ(0)2 e(0) + κ0 (0)n(0) + κ(0)τ (0)b(0)) + o(s3 ),
6
where o(s3 ) represents a higher order term (cf. Appendix A.1). This formula
is known as Bouquet’s formula 2 . Compare with Problem 8 in Section 2.
5 Show that a space regular curve whose curvature is a positive constant, and
whose torsion is a non-zero constant, coincides with a helix by an orientation
preserving congruence.
6∗ Show that a space regular curve whose torsion vanishes identically (and having
positive curvature) coincides with a curve obtained as in Problem 3 from a
planar curve, up to an orientation preserving congruence.
7∗ Show that a regular curve that lies on a sphere and has constant curvature is
a part of a circle.
8∗ Let γ1 (t) be a planar regular curve with curvature κ1 (t), and γ2 (t) a curve
obtained by reflecting γ1 (t) across a line. Show that the curvature of γ2 (t)
is −κ1 (t) (cf. Problem 4 in Section 2). Let γ̃1 (t) and γ̃2 (t) be space curves
obtained from γ1 and γ2 , respectively, as in Problem 3. Then show that the
curvature functions of these curves are |κ1 |, and the torsion functions vanish
identically.
Here, since γ1 and γ2 have different curvature functions as planar curves, they
cannot coincide by rotations and translations. In contrast, the space curves γ̃1
and γ̃2 have the same curvature and torsion, and hence by the fundamental
theorem for space curves (Theorem 5.2), they coincide by a rotation and
translation (an orientation preserving congruence). Explain the reason for
this difference between planar curves and space curves.