(Graduate Studies in Mathematics 178.) Clelland, Jeanne N. - From Frenet To Cartan - The Method of Moving Frames-American Mathematical Society (2017)
(Graduate Studies in Mathematics 178.) Clelland, Jeanne N. - From Frenet To Cartan - The Method of Moving Frames-American Mathematical Society (2017)
I N M AT H E M AT I C S 178
From Frenet
to Cartan:
The Method of
Moving Frames
Jeanne N. Clelland
From Frenet
to Cartan:
The Method of
Moving Frames
Jeanne N. Clelland
2010 Mathematics Subject Classification. Primary 22F30, 53A04, 53A05, 53A15, 53A20,
53A55, 53B25, 53B30, 58A10, 58A15.
Copying and reprinting. Individual readers of this publication, and nonprofit libraries acting
for them, are permitted to make fair use of the material, such as to copy select pages for use
in teaching or research. Permission is granted to quote brief passages from this publication in
reviews, provided the customary acknowledgment of the source is given.
Republication, systematic copying, or multiple reproduction of any material in this publication
is permitted only under license from the American Mathematical Society. Permissions to reuse
portions of AMS publication content are handled by Copyright Clearance Center’s RightsLink
service. For more information, please visit: http://www.ams.org/rightslink.
Send requests for translation rights and licensed reprints to [email protected].
Excluded from these provisions is material for which the author holds copyright. In such cases,
requests for permission to reuse or reprint material should be addressed directly to the author(s).
Copyright ownership is indicated on the copyright page, or on the lower right-hand corner of the
first page of each article within proceedings volumes.
c 2017 by the American Mathematical Society. All rights reserved.
The American Mathematical Society retains all rights
except those granted to the United States Government.
Printed in the United States of America.
∞ The paper used in this book is acid-free and falls within the guidelines
established to ensure permanence and durability.
Visit the AMS home page at http://www.ams.org/
10 9 8 7 6 5 4 3 2 1 22 21 20 19 18 17
To Rick, Kevin, and Valerie, who make everything worthwhile
Contents
Preface xi
Acknowledgments xv
vii
viii Contents
Chapter 11. Curves and surfaces in elliptic and hyperbolic spaces 339
§11.1. Introduction 339
§11.2. The homogeneous spaces Sn and Hn 340
§11.3. A more intrinsic view of Sn and Hn 345
§11.4. Moving frames for curves in S3 and H3 348
§11.5. Moving frames for surfaces in S3 and H3 351
§11.6. Maple computations 357
x Contents
Perhaps the earliest example of a moving frame is the Frenet frame along
a nondegenerate curve in the Euclidean space R3 , consisting of a triple
of orthonormal vectors (T, N, B) based at each point of the curve. First
introduced by Bartels in the early nineteenth century [Sen31] and later
described by Frenet in his thesis [Fre47] and Serret in [Ser51], the frame
at each point is chosen based on properties of the geometry of the curve
near that point, and the fundamental geometric invariants of the curve—
curvature and torsion—appear when the derivatives of the frame vectors are
expressed in terms of the frame vectors themselves.
In the late nineteenth century, Darboux studied the problem of construct-
ing moving frames on surfaces in Euclidean space [Dar72a], [Dar72b],
[Dar72c], [Dar72d]. In the early twentieth century, Élie Cartan general-
ized the notion of moving frames to other geometries (for example, affine and
projective geometry) and developed the theory of moving frames extensively.
A very nice introduction to Cartan’s ideas may be found in Guggenheimer’s
text [Gug77].
More recently, Fels and Olver [FO98], [FO99] have introduced the notion
of an “equivariant moving frame”, which expands on Cartan’s construction
and provides new algorithmic tools for computing invariants. This approach
has generated substantial interest and spawned a wide variety of applications
in the last several years. This material will not be treated here, but several
surveys of recent results are available; for example, see [Man10], [Olv10],
and [Olv11a].
xi
xii Preface
A special feature of this book is that it includes guidance on how to use the
mathematical software package Maple to perform many of the computa-
tions involved in the exercises. (If you do not have access to Maple, rest
assured that, with very few exceptions, the exercises can be done perfectly
well by hand.) The computations here make use of the custom Maple pack-
age Cartan, which was written by myself and Yunliang Yu of Duke Univer-
sity. The Cartan package can be downloaded either from the AMS webpage
www.ams.org/bookpages/gsm-178
or from my webpage at
http://euclid.colorado.edu/~jnc/Maple.html.
(Installation instructions are included with the package.) The last section of
Chapter 2 contains an introduction to the Cartan package, and beginning
with Chapter 3, each chapter includes a section at the end describing how
to use Maple and the Cartan package for some of the exercises in that
chapter. Additional exercises are worked out in Maple worksheets for each
chapter that are available on the AMS webpage.
Remark. As of Maple 16 and above, much of Cartan’s functionality is now
available as part of the DifferentialGeometry package, which is included
in the standard Maple installation and covers a wide range of applications.
The two packages have very different syntax, and no attempt will be made
here to translate—but interested readers are encouraged to do so!
Acknowledgments
xv
xvi Acknowledgments
Thanks to all the other students who have worked through subsequent ver-
sions of the manuscript over the last several years: Brian Carlsen, Michael
Schmidt, Edward Estrada, Molly May, Jonah Miller, Sean Peneyra, Duff
Baker-Jarvis, Akaxia Cruz, Rachel Helm, Peter Joeris, Joshua Karpel, An-
drew Jensen, and Michael Mahoney. These independent study courses—and
the research projects that followed—have been, hands down, the most re-
warding experiences of my teaching career. I hope you all enjoyed them
half as much as I did! And thanks to Sunita Vatuk and George Wilkens for
sitting in on some of these courses, contributing many valuable insights to
our discussions, and making great suggestions for the manuscript.
I am grateful to the Mathematical Sciences Research Institute for sponsoring
the 1999 Summer Graduate Workshop where I gave the lectures that were
the genesis for this book; videos of the original lectures are available on
MSRI’s webpage at [Cle99]. I am also grateful to the National Science
Foundation for research support; portions of this book were written while I
was supported by NSF grants DMS-0908456 and DMS-1206272.
Finally, profound thanks to my husband, Rick; his love and support have
been constant and unwavering, and I count myself fortunate beyond all
measure to have him as my best friend and partner in life.
Part 1
Background material
Chapter 1
This chapter contains some useful background material from differential ge-
ometry. Much of this material would typically be covered in courses on
multivariable analysis, algebra, and graduate-level differential geometry and
linear algebra. This chapter is not intended to be a comprehensive intro-
duction to any of these topics; the focus will be on ideas that will be used
in the remainder of the book, and details will be kept to a minimum.
1.1. Manifolds
Just about all the objects that we discuss in this book will be defined on a
manifold of some sort. The simplest manifolds are things that you already
know about: regular curves (1-dimensional manifolds) and regular surfaces
(2-dimensional manifolds). Before making a more general definition, let’s
remind ourselves of some of the fundamental properties of regular surfaces.
(1) x is differentiable.
(2) x is a homeomorphism.
(3) For each u ∈ U , the differential dxu : R2 → R2 is one-to-one.
3
4 1. Assorted notions from differential geometry
Remark 1.2. Throughout this book, the word “differentiable” will be taken
to mean “infinitely differentiable” (i.e., C ∞ ). We will also use the word
“smooth” as a synonym.
(1.1) x−1 −1 −1
j ◦ xi : xi (V ) → xj (V ), x−1 −1 −1
i ◦ xj : xj (V ) → xi (V )
.................................................................................
................................................ .........................
......................... ....................
.................... ..................
.................... .............................................. .......................................................
... ....... ....
....
.................. .......... ..
... ..........
. .
........... ..... ........ ...
..
... ..
. ..
.
..... . .
..... ...... Σ .
..
..
... .
.. .
... .
... .
... ...
x (U ) ... x (U ) ..
.. .. i .. V ...
.. j ...
... ..
..
..
.. .... ..... . . ..
.. ... ... .... ...
. ...
.. ... ... .
.. .... .... ... ...
. ...
.. .... .... .... .... ...
.. ..... ...... .... ..... ...
. ........ ......... ............. .......
.... ........... ....
.................................................. ...................................................... ......... ..
..
.. ..
.. ..
..
.. ...............................................................................................................
..
.
.... .................................................................. . .......................... ...
.
................. ...............
@
I
@ xj
xi
@
@
v @ v̄
6 6
.............................................. ............................ ..... ...............
.............. .... ............. .........
......... .............. ........... .......
....... ...... ......... ....... .......... .....
..... .... ..... .... ....
x−1
. .... .
j ◦ xi
. .... . .... ..
... U . ... ... U j ......
-
. i . . ...
... ... −1 ... ... −1 .... ...
.... x.... (V ) ... x
.... j (V )
... ... i -u ...
. ...
...
.. - ū...
.
..
..
..
.. .. ..
... ... .. ... ... ..
... ... .... ... .
..
...
.
x−1
... ... . ... ... ...
i ◦ xj
.
...
.... ....
...
. ...
.... .... ....
..... ....
..... ........ ..... ... ...
........ .............. ................... .
.........
.......... .......... .
.................... ............................. .................... .......................... ....
........... ...........
sic objects, not living in any particular Euclidean space. In this case, the
differentiability of (1.1) becomes part of the definition of a regular surface:
(2) For each pair i, j with V = xi (Ui ) ∩ xj (Uj ) = ∅, the sets x−1
i (V )
−1 −1
and xj (V ) are open sets in R , and the mappings xj ◦ xi and
2
x−1
i ◦ xj are differentiable.
x−1 −1 −1
j ◦ xi : xi (V ) → xj (V )
x−1
j ◦ xi (u, v) = (ū(u, v), v̄(u, v)).
(2) For any pair i, j with V = xi (Ui ) ∩ xj (Uj ) = ∅, the sets x−1
i (V ) and
xj (V ) are open sets in R , and the mappings xj ◦xi and x−1
−1 n −1
i ◦xj
are differentiable. The mappings xi are called parametrizations of
M , or systems of local coordinates on M .
Remark 1.5. When studying regular surfaces in R3 , it is traditional to use
(u, v) as local coordinates on R2 and to write parametrizations x : U → R3
as ⎡ ⎤
x(u, v)
x(u, v) = ⎣y(u, v)⎦ .
z(u, v)
When we graduate to manifolds of arbitrary dimension, we need to use
variables with indices so that we don’t run out of letters. So, depending on
the context, we will generally use either u = (u1 , . . . , un ) or x = (x1 , . . . , xn )
as local coordinates on an open set U ⊂ Rn .
1.1.3. Examples.
Example 1.6 (The unit sphere). Let Sn ⊂ Rn+1 be the set
Sn = {t[x1 , . . . , xn+1 ] ∈ Rn+1 | (x1 )2 + · · · + (xn+1 )2 = 1};
i.e., Sn is the set of all vectors in Rn+1 of Euclidean length 1.
Remark 1.7. Vectors in Rk are assumed to be column vectors unless
otherwise specified. The notation t[x1 , . . . , xn+1 ] in Example 1.6 denotes
the transpose of the row vector [x1 , . . . , xn+1 ], which is the column vector
1.1. Manifolds 7
⎡ ⎤
x1
⎢ .. ⎥
⎣ . ⎦. We will often write column vectors in this way in order to save
xn+1
space, using the transpose notation in order to maintain the distinction
between column vectors and row vectors.
Let
U = {t[u1 , . . . , un ] ∈ Rn | (u1 )2 + · · · + (un )2 < 1};
i.e., U is the open unit ball in Rn . Define maps
i : U → Vi ,
x+ +
x−
i : U → Vi
−
by
t
x+ 1 n
i (u , . . . , u ) = u1 , . . . , ui−1 , 1 − (u1 )2 − · · · − (un )2 , ui , . . . , un ,
t
x− 1 n
i (u , . . . , u ) = u1 , . . . , ui−1 , − 1 − (u1 )2 − · · · − (un )2 , ui , . . . , un .
−
(a) What portion of Sn is covered by the images of x+ i and xi ? Show
that every point in Sn is contained in the image of at least one of these
parametrizations.
Example 1.9 (Real projective space of dimension n). Let Pn denote the
set of lines through the origin in the vector space Rn+1 . Since any such line
is determined by any point on the line other than the origin, we can think
of Pn as the quotient space
Pn = (Rn+1 \ {0})/ ∼,
the same line through the origin. It is customary to use local coordinates
(x0 , x1 , . . . , xn ) on Rn+1 ; then for any nonzero real number λ, we have
[x0 , . . . , xn ] ∼ t[λx0 , . . . , λxn ].
t
P2
..................................
........... .......
....... .....
S1 ..... ....
...
P1 ..
.. ...
...
............. .
..... .......
...
........... ....
...r .r -
.
..
.... .... r...
......
..... .
.... ..
... .
.... .
..
. .... ... .. ... ..... .. ... ...r
. .. ..
... ..
..... .... ..
...
.
r
........
.........r..... ....... ..
r
............ . ........
. .
..
... .
...........................................................
. .
O
... . . .
.... .....
...
.. ....
....
.......
K
............ ................. .................
.....................................
(a) What portion of Pn is covered by the image of [xi ]? Show that every point
in Pn is contained in the image of at least one of these parametrizations.
(b) For i = j, identify the set Vij = [xi (U )] ∩ [xj (U )] in Pn and the open sets
[xj ]−1 (Vij ) and [xi ]−1 (Vij ) in U , and compute the local coordinate transfor-
mation [xj ]−1 ◦ [xi ] between the latter two open sets.
In equation (1.2), the variables are indexed with superscripts rather than
subscripts. You may be wondering—why on earth would we do that? So,
before we go any farther, let’s discuss how (and why!) indices are used
throughout this book. Some objects will be indexed with superscripts (“up-
per indices” or, less formally, “up indices”), some by subscripts (“lower
indices” or “down indices”), and some by a combination of both. Most of
these objects will be tensors or tensor fields. Without getting too specific,
a tensor is just an element of a certain type of vector space, and a tensor
field on a manifold is defined by assigning a tensor to each point of the
manifold. (A precise definition will be given in Chapter 2.) For example,
tensors include objects such as vectors and linear transformations, while
tensor fields include objects such as vector fields, metrics, and differential
forms on manifolds.
A key feature of tensors is the way in which their coordinate expressions
change when the basis for the underlying vector space is changed. The
following examples demonstrate some typical behavior.
v = a1 e1 + · · · + an en
where ⎡ ⎤
r11 . . . rn1
⎢ .. .. ⎥ .
R=⎣. .⎦
r1n . . . rnn
Then we have
e1 . . . en = ē1 . . . ēn R−1 ,
and so
v = a1 e1 + · · · + an en
⎡ 1⎤
a
⎢ . ⎥
= e1 . . . en ⎣ .. ⎦
an
⎤⎡
a1
⎢ ⎥
= ē1 . . . ēn R−1 ⎣ ... ⎦
an
= ā1 ē1 + · · · + ān ēn ,
where ⎡ ⎤ ⎡ 1⎤
ā1 a
⎢ .. ⎥ −1 ⎢ .. ⎥
⎣ . ⎦ = R ⎣ . ⎦.
ān an
Let’s think carefully about what this means. We usually write a vector
v ∈ V as a column vector
⎡ 1⎤
a
⎢ .. ⎥
a=⎣ . ⎦
am
(by which we really mean that v = [e1 . . . em ] · a), and we write w = T (v)
as the matrix product
⎡ 1⎤ ⎡ 1 ⎤⎡ 1⎤
b c1 . . . c1m a
⎢ .. ⎥ ⎢ .. .. ⎦ ⎣ ... ⎥
. ⎥ ⎢
b=⎣.⎦=⎣. ⎦
bn n
c1 . . . cm n am
(by which we really mean that w = [f1 . . . fn ] · b). The use of different
notations here for the vectors v and w, which are well-defined independently
of a choice of basis, and their expressions a and b in terms of specific bases,
is deliberate: The equation
b = AT a
is a basis-dependent expression of the basis-independent equation
w = T (v).
Again, we ask what happens if we change bases for V and W . Suppose that
we set
(1.4) ē1 . . . ēm = e1 . . . em R,
f̄1 . . . f̄n = f1 . . . fn S,
AT = S −1 AT R.
Remark 1.13. These were fairly simple examples of tensors, dealing with
a single vector and a single linear transformation. We will mostly be inter-
ested in tensor fields, which consist of an underlying manifold with a tensor
defined at each point. It may sound complicated, but you already know
several examples—e.g., vector fields and metrics on surfaces. The only real
differences are that:
(1) The components are functions on the base manifold rather than
constants.
(2) The vector spaces in question are usually the tangent or cotangent
spaces to the manifold at each point.
(3) Basis changes to the vector spaces usually arise as derivatives (i.e.,
Jacobian matrices) of local coordinate transformations on the man-
ifold.
gq : Tq M × Tq M → R
The assumption that g is symmetric implies that gji (q) = gij (q), and the
bilinearity of g implies that for any vectors
n n
v= ai ei , w= bj ej
i=1 j=1
in Tq M , we have
n n
gq (v, w) = gij (q)ai bj .
i=1 j=1
For instance, when M is a regular surface Σ ⊂ R3 with a local parametriza-
tion x : U → Σ, we typically use the basis
e1 = xu , e2 = xv
for Tq M , and the bilinear form is given by the dot product in R3 :
gq (v, w) = v · w.
Then we have
g11 = E = xu · xu , g12 = g21 = F = xu · xv , g22 = G = xv · xv .
A metric g is often represented by the symmetric matrix
⎡ ⎤
g11 . . . g1n
⎢ .. ⎥ .
Ag = ⎣ ... . ⎦
gn1 . . . gnn
etc. So, in order to reduce notational clutter, we omit the explicit sum
notation and simply write
v = ai ei ,
g(v, w) = gij ai bj .
The Einstein summation convention says that whenever the same index ap-
pears twice in an expression, once up and once down, it indicates a sum over
the entire range of that index. This takes a bit of getting used to, but it’s
actually extremely handy. For instance, it makes the multivariable chain
rule look exactly like the single-variable version, where you can “cancel” the
intermediate variables:
∂f ∂f ∂y j
= .
∂xi ∂y j ∂xi
It also gives a quick method for error checking, much like dimensional analy-
sis in chemistry or physics: Once repeated indices are “canceled”, the indices
on both sides of an equation should match.
also called the tangent vector to α at the point α(t). But the expression
(1.5) is actually an incomplete description of the tangent vector because the
base point of a vector is a crucial part of its definition. A more accurate
definition would be
⎛⎡ 1 ⎤ ⎡ dy1 ⎤⎞
y (t) dt (t)
⎜⎢ .. ⎥ ⎢ .. ⎥⎟
(1.6) α (t) = ⎝⎣ . ⎦ , ⎣ . ⎦⎠ ,
y n (t) dy n
dt (t)
including both the base point and its derivative. While the abbreviated
notation (1.5) is common, it is important to remember that it represents a
vector based at the specific point α(t) in Rn .
Given a point y = t[y 1 , . . . , y n ] ∈ Rn , the tangent space to Rn at y, denoted
Ty Rn , is the set of all tangent vectors to all curves in Rn passing through
y; i.e.,
v=α (0)
....................
............. F◦α
.........
r..
.......
..... rP dF (v)=(F ◦ α) (0)
............................................................
............
PP
....
F .........
....
....
....
x=α(0) - F (x) q
P
........
......
.....
x
.... ....
....
... ...
α .. ...
.. .
..
....
..
..
Now let’s get a little more abstract and define the analogous notions for
manifolds.
Definition 1.17. Suppose that M and N are m- and n-dimensional man-
ifolds, respectively. A map F : M → N is called differentiable if for any
parametrizations
x : U ⊂ Rm → M, y : V ⊂ Rn → N
on M and N such that F (x(U )) ⊂ y(V ), the composite map
y−1 ◦ F ◦ x : U → V
is differentiable. (Definition 1.4 implies that this condition is independent
of the choice of parametrizations x and y.)
Remark 1.18. Keeping track of all these different maps can be challenging,
and so they are often described using diagrams. For example, the maps
1.3. Differentiable maps, tangent spaces, and vector fields 19
M
F - N
x 6 6y
y−1 ◦F ◦x
Rm ⊇ U - V ⊂ Rn .
xi = xi (x̄1 , . . . , x̄m ), i = 1, . . . , m,
on M . Use the definition of tangent vectors and the multivariable chain rule
to show that
∂
∂ x̄1 · · · ∂
∂ x̄m = ∂
∂x 1 · · · ∂
∂xm J,
Moreover, the vector spaces associated to each point all have the same
dimension—2, in this case. T Σ is called the total space of the tangent bundle,
and it can be given the structure of a manifold of dimension 4, as follows. If
x = (x1 , x2 ) is a system of local coordinates on an open set V ⊂ Σ, it can be
canonicallyextended to a system of local coordinates (x, y) = (x1 , x2 , y 1 , y 2 )
on T V = q∈V Tq V ⊂ T Σ by associating to any tangent vector v ∈ Tq V
its unique representation as
∂ ∂
(1.7) v = y1 1 + y2 2 .
∂x ∂x
*Exercise 1.24. Let T Σ be the tangent bundle of a regular surface, and let
x = (x1 , x2 ) and x̄ = (x̄1 , x̄2 ) be two overlapping systems of local coordinates
on Σ, related by a coordinate transformation of the form
(1.8) x1 = x1 (x̄1 , x̄2 ), x2 = x2 (x̄1 , x̄2 ).
Show that the local coordinate transformation between the canonically asso-
ciated local coordinates (x, y) = (x1 , x2 , y 1 , y 2 ) and (x̄, ȳ) = (x̄1 , x̄2 , ȳ 1 , ȳ 2 )
on T Σ is given by equations (1.8) together with the equations
y1 ȳ 1
= J ,
y2 ȳ 2
where J is the Jacobian matrix of the transformation (1.8). (Hint: The
result of Exercise 1.22 should be helpful here.)
Tangent bundles for manifolds of arbitrary dimension work exactly the same
way. If M is a manifold of dimension m and x : U ⊂ Rm → M is a
parametrization of M , then the associated canonical parametrization of T M
is the map (x, y) : U × Rm → T M given by
∂
(x, y)(x1 , . . . , xm , y 1 , . . . , y m ) = y i i ∈ Tx(x1 ,...,xm ) M.
∂x
In other words, for each point q ∈ x(U ), the tangent space Tq M isidentified
with the vector space Rm by identifying the basis ∂x∂ 1 , . . ., ∂x∂m for Tq M
with the standard basis for Rm . The same calculation as in Exercise 1.24
shows that the transition map between any two canonical parametrizations
(x, y) and (x̄, ȳ) is differentiable and, in fact, has the form
∂xi j
xi = xi (x̄1 , . . . , x̄m ), yi = ȳ ;
∂ x̄j
1.3. Differentiable maps, tangent spaces, and vector fields 23
x : U ⊂ Rm → M, y : V ⊂ Rn → N
H = y−1 ◦ F ◦ x : U → V.
We can write
in other words,
⎛ ⎡ ⎤⎞ ⎡ ⎤
a1 a1
⎜ ⎢ . ⎥⎟ ⎢ ⎥
dFq ⎝ ∂
∂x1
··· ∂
∂xm ⎣ . ⎦⎠ =
. ∂
∂y 1
··· ∂
∂y n J ⎣ ... ⎦ ,
am am
where J is the n × m Jacobian matrix of H.
BONUS: What happens to this formula if you perform local coordinate
transformations on M and N ?
(b) The differential of F can be extended to a map
dF : T M → T N
in the obvious way: For v ∈ Tq M , define
dF (v) = dFq (v).
Show that dF is a differentiable map from T M to T N .
(b) Give an explicit interpretation of the chain rule in terms of local coor-
dinates and Jacobian matrices. (What are the sizes of each of the matrices
involved, in terms of the dimensions m, n, p of M, N , and P ?) Write out the
(i, j)th component of this matrix equation using the Einstein summation
convention, and make sure that the indices on both sides of the equation
match up correctly.
This is a concise way of stating that group multiplication and group inverse
are both differentiable operations on G, as the following exercise shows.
Exercise 1.32. Let G be a Lie group. Show that the multiplication map
G × G → G defined by
(g, h) → gh
and the inverse map G → G defined by
g → g −1
are both differentiable. Conversely, show that differentiability of both of
these maps implies that the map μ in Definition 1.31 is differentiable. (Hint:
Consider the inverse map first.)
To any Lie group G is associated a Lie algebra g. The Lie algebra g is simply
the tangent space to G at the identity element e ∈ G; thus g is a vector space
of the same dimension as the manifold G. There is also a product structure
on g known as the Lie bracket. Defining this structure will require a bit of
effort.
For each element g ∈ G, the left translation or left multiplication map Lg :
G → G is defined by
Lg (h) = gh.
(Similarly, the right translation map Rg : G → G is defined by
Rg (h) = hg.)
For any element h ∈ G, the differential (dLg )h of Lg at h is a linear map from
Th G to Tgh G. In particular, the differential of Lg at the identity element
e is a linear map from the Lie algebra g to Tg G. We can use this map to
associate to any element v ∈ g a left-invariant vector field ṽ on G defined
at each point g ∈ G by
(1.11) ṽ(g) = (dLg )e (v).
1.4. Lie groups and matrix groups 27
Finally, we use the Lie bracket of the left-invariant vector fields ṽ, w̃ to
define the Lie bracket operation on the Lie algebra g:
Definition 1.35. Let G be a Lie group with associated Lie algebra g. For
any two elements v, w ∈ g, the Lie bracket [v, w] of v and w is the unique
element z ∈ g such that
[ṽ, w̃] = z̃,
where tildes denote the extensions to left-invariant vector fields on G, as
defined above.
In order for this definition to make sense, we need to know that the Lie
bracket of two left-invariant vector fields on G is also left-invariant.
*Exercise 1.36. Let ṽ, w̃ be left-invariant vector fields on a Lie group G.
Show that the Lie bracket [ṽ, w̃] satisfies the condition (1.12) and that it
is therefore equal to z̃ for some z ∈ g. Conclude that the Lie bracket is a
well-defined operation on g.
While this all sounds fairly abstract, the most common examples of Lie
groups—and the only ones that we will encounter in this book—are sub-
groups of the group GL(n) of invertible n × n matrices. The differentiable
structure on GL(n) is simply that inherited by regarding GL(n) as an open
2
set in Rn . The Lie algebra of any such Lie group is a subspace of the vector
space of n × n matrices.
Remark 1.37. If it seems a little strange to think of a matrix as a vector,
let G be any subgroup of GL(n), and let α(t) be a curve in G with α(0) = I.
Then we can write
α(t) = g(t),
where g(t) is an n × n matrix. The tangent vector to this curve at t = 0 is,
of course,
α (0) = g (0),
2
which, while it could be thought of as a vector in Rn , is most naturally
regarded as an n × n matrix.
The following exercise shows that the Lie bracket on the Lie algebra of any
subgroup of GL(n) is simply the matrix commutator:
[A, B] = AB − BA.
*Exercise 1.38. Let G be a subgroup of GL(n).
(a) Let g ∈ G, and let Lg : G → G be the left multiplication map. Show
that the differential dLg : g → Tg G is given explicitly by
dLg (A) = g · A, A ∈ g,
1.4. Lie groups and matrix groups 29
where the right-hand side is just the matrix product of the two n×n matrices
g and A. Therefore, the vector field à on G defined by
Ã(g) = g · A
is left-invariant. (Hint: Consider a curve h(t) in G with h(0) = I, h (0) = A.)
(b) Let’s get explicit about exactly what differential operator the vector field
à represents on G: If we use local coordinates (xij ) on the manifold of n × n
matrices, then a general element g ∈ G is represented by a matrix [xij ]. If
we write A = [aij ], then the tangent vector A ∈ g represents the differential
operator
∂
A = aij i ,
∂xj
and the matrix product à = g · A, whose entries are
Ãij = xir arj ,
represents the differential operator
∂
à = xir arj .
∂xij
Now, let A = [aij ], B = [bkl ] ∈ g. Then we can write
∂ ∂
à = xir arj , B̃ = xks bsl .
∂xij ∂xkl
Use the formula that you computed in Exercise 1.34 (and some very careful
index manipulation!) to show that
! ∂
[Ã, B̃] = xir arl blj − brl alj .
∂xij
Thus, its Lie algebra gl(n) is the tangent space to the space of all n × n
matrices at the identity In . This tangent space is naturally isomorphic to
Mn×n .
is a curve of matrices in SL(n) with A(0) = In . Since A(t) ∈ SL(n) for all
t, we have
where the sum is over all permutations σ in the symmetric group Sn and
sgn(σ) = ±1 is the sign of σ. In order to determine what conditions this
imposes on the tangent vector A (0), we need to differentiate this equation
and evaluate the result at t = 0. Fortunately, this computation is not as
bad as it looks: The fact that A(0) = In means that aij (0) = 0 unless i = j,
in which case aij (0) = 1. So most of the terms in the derivative will drop
out at t = 0, leaving the equation
(aii ) (0) = 0.
(Don’t forget the summation convention!) In other words, A (0) must have
trace equal to zero.
Aside from the trace condition, there are no other restrictions on A (0).
This is most easily verified by comparing dimensions: SL(n) has dimension
n2 − 1, and the set of n × n matrices with trace equal to zero is a vector
space of dimension n2 − 1. Since sl(n) is contained in this vector space and
has the same dimension as the entire space, it must be equal to the entire
space. Therefore,
sl(n) = {B ∈ Mn×n | tr(B) = 0}.
Example 1.42. The orthogonal group O(n). O(n) is defined as
O(n) = {A ∈ GL(n) | tA A = In },
where tA denotes the transpose of the matrix A. A standard computation
using the implicit function theorem (which we omit here) shows that O(n)
is a manifold of dimension 12 n(n − 1). In order to compute its Lie algebra
o(n), suppose that A(t) is a curve of matrices in O(n) with A(0) = In . Since
A(t) ∈ O(n) for all t, we have
t
A(t) A(t) = In .
Differentiating this equation and setting t = 0 yields
t
A (0) + A (0) = 0;
in other words, A (0) must be a skew-symmetric matrix. Comparing dimen-
sions as in the previous example shows that there are no other restrictions
on A (0), and so
o(n) = {B ∈ Mn×n | tB + B = 0}.
Example 1.43. The special orthogonal group SO(n). SO(n) is defined as
SO(n) = {A ∈ GL(n) | tA A = In and det(A) = 1}.
Clearly, SO(n) is a subgroup of O(n); it consists of those matrices in O(n)
having determinant equal to 1. Now, it follows from the defining equation
for O(n) that every element of O(n) has determinant equal to either 1 or
−1. Moreover, the function
det : O(n) → R
is continuous, with image equal to the set {1, −1}. Therefore, the subsets
SO(n) = {A ∈ O(n) | det(A) = 1},
O(n) \ SO(n) = {A ∈ O(n) | det(A) = −1}
are disconnected. In fact, O(n) consists of two connected components, one
of which is SO(n) and the other of which is the coset of SO(n) consisting of
those elements with determinant −1. It follows that the Lie algebra so(n)
is, in fact, equal to the Lie algebra o(n). This Lie algebra is usually denoted
by so(n).
32 1. Assorted notions from differential geometry
For a more thorough discussion of Lie groups and Lie algebras, see, e.g.,
[Lee13].
It is important to note that, while vector bundles always have local trivial-
izations, they may not have global trivializations. That is to say, the local
trivializations may patch together in such a way that their union becomes
“twisted” in a way that cannot be straightened out. So, a vector bundle
π : E → B is not necessarily diffeomorphic to B × Rk . For example, the
tangent bundle of a compact surface has no global trivialization unless the
surface has Euler characteristic equal to zero.
Next, we introduce the notion of a section of a vector bundle.
Definition 1.45. A (smooth) section of a vector bundle π : E → B is
a differentiable map σ : B → E with the property that the composition
π ◦ σ : B → B is the identity map on B.
G - P
π
?
B.
34 1. Assorted notions from differential geometry
The arrow G → P doesn’t really represent a map here; it just indicates that
each fiber of P is diffeomorphic to G and that G acts freely and transitively
on the fibers of P .
We will see some examples of principal bundles beginning in Chapter 3. In
many cases, the total space P will itself be a Lie group, and the fibers will
all be diffeomorphic to some subgroup G of P .
Exercise 1.46. Let Σ be the unit sphere S2 ⊂ R3 . Let F (S2 ) denote the
orthonormal frame bundle of S2 ; this is the set of all orthonormal bases
(a.k.a. “frames”) e = (e1 , e2 ) for the tangent space Tq S2 at each point
q ∈ S2 . Given any orthonormal frame e at a point q ∈ S2 , any other frame
at q can be obtained from e by composition of a rotation and (possibly) a
reflection, i.e., an element of the Lie group O(2). Conversely, every element
of O(2) acts on e to produce a new orthonormal frame at q, and distinct
elements of O(2) will produce different frames. Thus, the fiber of F (S2 ) over
any point q ∈ S2 is diffeomorphic to O(2), and F (S2 ) is a principal bundle
over S2 with fiber group O(2).
Prove that F (S2 ) has no global sections. (Hint: Use the Poincaré-Hopf
theorem.)
Chapter 2
Differential forms
2.1. Introduction
∂f ∂f ∂f
df = dx + dy + dz.
∂x ∂y ∂z
The expression df is called a 1-form and you probably learned various ways
of manipulating such things—for instance, how to integrate a 1-form along
a parametrized curve in R3 . But what is a 1-form, really?
In this chapter, we will give formal definitions for 1-forms and then more
generally for p-forms with p ≥ 0 on Rn , and we will explore a bit about how
they work. Once we understand how differential forms behave in Rn , we will
see how to define them more generally on manifolds. But before we launch
into all that, it will be helpful to go into a bit more detail about certain
types of tensors.
35
36 2. Differential forms
dimension n:
*Exercise 2.2. Let (e1 , . . . , en ) be a basis for V . Define linear functions
e∗i : V → R, 1 ≤ i ≤ n, by the property that
∗i i 1, i = j,
e (ej ) = δj =
0, i = j.
(Note that requiring e∗i to be linear and specifying the value of e∗i on all of
the basis vectors ej completely determines the function e∗i on V .)
(a) Show that the functions (e∗1 , . . . , e∗n ) are linearly independent. (Hint:
The hypothesis “ci e∗i = 0” means that for every v ∈ V , ci e∗i (v) = 0.)
(b) Show that the functions (e∗1 , . . . , e∗n ) span V ∗ . (Hint: Let α : V → R be
a linear mapping, and let ci = α(ei ). Consider the function ci e∗i : V → R.)
(c) Show that (V ∗ )∗ ∼
= V . (Hint: Associate to any vector v ∈ V the function
∗
v̂ : V → R defined by
v̂(α) = α(v).)
The basis (e∗1 , . . . , e∗n ) for V ∗ constructed in Exercise 2.2 is called the
dual basis to the basis (e1 , . . . , en ) for V . Also, note that the isomorphism
(V ∗ )∗ ∼
= V is canonical: It is completely independent of any choice of basis
for V . Thus it is accurate—and common—to regard (V ∗ )∗ as being equal
to V .
The most common example of a dual space that we will encounter is the
following:
Example 2.3. Let M be a manifold of dimension m, let q ∈ M , and let
V = Tq M . The dual space V ∗ = Tq∗ M is called the cotangent space to M
at q. The cotangent bundle of M is the union of all these cotangent spaces:
T ∗M = Tq∗ M.
q∈M
Most of the tensor fields that commonly appear in geometry are sections of
vector bundles that are constructed from the tangent and cotangent bundles
of a manifold. The construction used to build more complicated bundles
from these two is the tensor product. The official definition starts by defining
tensor products for dual spaces:
Definition 2.5. Let V be a vector space of dimension m and W a vector
space of dimension n. Let α ∈ V ∗ , β ∈ W ∗ . The tensor product of α and β,
denoted α ⊗ β, is the bilinear function
α⊗β :V ×W →R
defined by
(α ⊗ β)(v, w) = α(v)β(w).
The tensor product of V ∗ and W ∗ , denoted V ∗ ⊗ W ∗ , is the vector space
consisting of all linear combinations of such tensor products; i.e.,
V ∗ ⊗ W ∗ = span{α ⊗ β | α ∈ V ∗ , β ∈ W ∗ }.
*Exercise 2.6. (a) Show that α ⊗ β is indeed a bilinear function on V × W ,
i.e., that for any vectors v, v1 , v2 ∈ V , w, w1 , w2 ∈ W , and real numbers
a, b,
(α ⊗ β)(av1 + bv2 , w) = a(α ⊗ β)(v1 , w) + b(α ⊗ β)(v2 , w),
(α ⊗ β)(v, aw1 + bw2 ) = a(α ⊗ β)(v, w1 ) + b(α ⊗ β)(v, w2 ).
v1 ∧ · · · ∧ vk = sgn(σ)vσ(1) ⊗ · · · ⊗ vσ(k) ,
σ
where the sum is over all permutations σ in the symmetric group Sk and
sgn(σ) = ±1 is the sign of σ. The spaces of symmetric and skew-symmetric
k-tensors of V are the subspaces S k V , Λk V of V ⊗k defined by
S k V = span{v1 ◦ · · · ◦ vk | v1 , . . . , vk ∈ V },
Λk V = span{v1 ∧ · · · ∧ vk | v1 , . . . , vk ∈ V }.
Exercise 2.14. (a) Show that V ⊗ V = S 2 V ⊕ Λ2 V ; i.e., every element of
V ⊗ V can be uniquely expressed as the sum of a symmetric product and a
skew-symmetric product.
(b) Show that the analogous decomposition does not hold for V ⊗3 . (Hint:
Compute the dimensions of V ⊗3 , S 3 V , and Λ3 V in terms of the dimension
n of V .)
All these tensor products can be used to form tensor bundles on a manifold
M . A tensor bundle on M is simply a vector bundle on M where each
fiber is isomorphic to a fixed tensor product of vector spaces. In the most
common examples, the fiber over each point q ∈ M is a tensor product of
40 2. Differential forms
some numbers of copies of the tangent space Tq M and the cotangent space
Tq∗ M , or some natural subspace of such a tensor product. Finally, this allows
us to give a formal definition of tensor fields:
Definition 2.15. A rank k tensor field on a manifold M is a section of a
vector bundle π : E → M , where the fiber Eq of E over each point q ∈ M
has the form
Eq = V1 ⊗ · · · ⊗ Vk ,
where V1 , . . . , Vk are vector spaces. (In most cases, each of the vector spaces
V1 , . . . , Vk is equal to either Tq M or Tq∗ M .)
Remark 2.16. In practice, tensor fields on a manifold M are almost always
referred to as “tensors on M ”. For example, a Riemannian manifold M has
a “metric tensor” g and a “Riemann curvature tensor” R. Strictly speaking,
these are both tensor fields on M , but you will probably never encounter
the phrase “Riemann curvature tensor field”.
Example 2.17 (Cf. Exercise 1.15). Let M be a manifold of dimension n,
and let
S 2 (T ∗ M ) = S 2 (Tq∗ M ).
q∈M
S 2 (T ∗ M ) is a subbundle of the tensor bundle T ∗ M ⊗ T ∗ M , and it is a
bundle of symmetric 2-tensors on M . (Note that I didn’t say “the” bundle
of symmetric 2-tensors on M , because there are other bundles that would
satisfy the same description, e.g., S 2 (T M ).) A metric g on M is a section
of S 2 (T ∗ M ) because, at every point q ∈ M , g defines a symmetric, bilinear
function
gq : Tq M × Tq M → R.
2.3. 1-forms on Rn
In particular, the 1-forms (dx1 , . . . , dxn ) are defined by the property that
for any vector v = t[v 1 , . . . , v n ] ∈ Tx Rn ,
(2.1) dxi (v) = v i .
*Exercise 2.19 (Cf. Exercise 2.2). In this exercise, we will show directly
that for any point x ∈ Rn , the 1-forms (dx1 , . . . , dxn ) defined by (2.1) form
a basis for the cotangent space Tx∗ Rn .
(a) In order to show that the 1-forms (dx1 , . . . , dxn ) are linearly independent,
suppose that
ci dxi = 0.
This means that ci dxi (v) = 0 for every vector v ∈ Tx Rn . Show that this
implies that c1 = · · · = cn = 0.
(b) In order to show that the 1-forms (dx1 , . . . , dxn ) span Tx∗ Rn , let φx ∈
Tx∗ Rn be arbitrary. Let (e1 , . . . , en ) be the standard basis for Tx Rn , and let
ci = φx (ei ), i = 1, . . . , n.
Show that φx = ci dxi . (It suffices to show that φx (v) = ci dxi (v) for every
v ∈ Tx Rn .)
It follows from Exercise 2.19 that the 1-forms (dx1 , . . . , dxn ) form a basis for
the 1-forms on Rn as a module over C ∞ (Rn ); this means that any 1-form φ
on Rn can be expressed uniquely as
(2.2) φ = fi (x) dxi
for some smooth, real-valued functions f1 , . . . , fn : Rn → R. If a vector field
v on Rn has the form
v(x) = t[v 1 (x), . . . , v n (x)],
then we have
φ(v(x)) = fi (x) v i (x).
2.4. p-forms on Rn
The 1-forms on Rn are the building blocks of an algebra, called the algebra
of differential forms on Rn . The multiplication in this algebra is the wedge
product of Definition 2.13. The wedge product is multilinear and skew-
42 2. Differential forms
φ ∧ ψ = −ψ ∧ φ.
∧ : Λp (T ∗ Rn ) × Λq (T ∗ Rn ) → Λp+q (T ∗ Rn ).
A basis for the p-forms on Rn (as a module over C ∞ (Rn )) is given by the
set
{dxi1 ∧ · · · ∧ dxip | 1 ≤ i1 < i2 < · · · < ip ≤ n};
this simply means that any p-form Φ on Rn can be expressed uniquely as
where the sum is over all permutations σ in the symmetric group Sp and
sgn(σ) = ±1 is the sign of σ.
Exercise 2.22. Show that the wedge product is related to the determinant
as follows: If φ1 , . . . , φp are 1-forms and v1 , . . . , vp are vector fields, then
⎡ ⎤
φ1 (v1 ) · · · φ1 (vp )
⎢ ⎥
⎢ .. ⎥ .
(φ1 ∧ · · · ∧ φp )(v1 , . . . , vp ) = det ⎢ ... . ⎥
⎣ ⎦
φp (v1 ) · · · φp (vp )
If Φ is a p-form and Ψ is a q-form, then the Leibniz rule takes the form
(2.7) d(Φ ∧ Ψ) = dΦ ∧ Ψ + (−1)p Φ ∧ dΨ.
The following theorem is possibly the most important one in this entire
book!
Next, note that by Exercise 2.24 and the argument above, d(dxi ) = 0. Now
suppose that
Φ= fI (x) dxi1 ∧ · · · ∧ dxip .
|I|=p
Then by the Leibniz rule,
⎛ ⎞
∂
(2) the curl of a vector field v(x) = v i (x) i :
∂x
" 3 2
# " 1 3
# " 2 #
∂v ∂v ∂ ∂v ∂v ∂ ∂v ∂v 1 ∂
curl(v) = 2
− 3 1
+ 3
− 1 2
+ 1
− 2 ;
∂x ∂x ∂x ∂x ∂x ∂x ∂x ∂x ∂x3
∂
(3) the divergence of a vector field v(x) = v i (x) :
∂xi
3
∂v i
div(v) = .
∂xi
i=1
These operations can all be expressed in terms of differential forms and their
exterior derivatives via the following correspondences between forms on the
one hand and vector fields and functions on the other:
3
∂
f1 (x)dx1 + f2 (x)dx2 + f3 (x)dx3 ←→ fi (x) ,
∂xi
i=1
3
∂
f1 (x)dx ∧ dx + f2 (x)dx ∧ dx + f3 (x)dx ∧ dx ←→
2 3 3 1 1 2
fi (x) ,
∂xi
i=1
f (x) dx ∧ dx ∧ dx ←→ f (x).
1 2 3
46 2. Differential forms
By Theorem 2.28, every exact form is closed. The converse is only partially
true: Every closed form is locally exact. Precisely, we have
Closely related to Corollary 2.32 is the following special case of the Frobenius
theorem:
then dΦ is the (p + 1)-form defined by the property that for any p + 1 vector
fields v1 , . . . , vp+1 on M ,
p+1
(2.8) dΦ(v1 , . . . , vp+1 ) = (−1)i+1 vi [Φ(v1 , . . . , v̂i , . . . , vp+1 )]
i=1
Here a hat over a symbol indicates that that element is omitted from the list;
vi [Φ(v1 , . . . , v̂i , . . . , vp+1 )] denotes the directional derivative of the scalar-
valued function Φ(v1 , . . . , v̂i , . . . , vp+1 ) in the direction of vi , and [vi , vj ]
denotes the Lie bracket of the vector fields vi and vj .
2.8. Pullbacks
Remark 2.39. This is why the same notation is used both for the differen-
tial dF of a map F : M → N and for the exterior derivative df of a function
f : M → R: The exterior derivative of a function really is just the differen-
tial of the function regarded as a map between manifolds. (This equivalence
relies on the canonical isomorphism Tf (q) R ∼
= R.)
*Exercise 2.40. Let M be a manifold of dimension m, N a manifold of
dimension n, and F : M → N a differentiable map with differential dF :
T M → T N . Let (f1 , . . . , fn ) be vector fields on N that form a basis for Tq N
at each point q ∈ N . Show that there exist ordinary, scalar-valued 1-forms
(φ1 , . . . , φn ) on M defined by the condition that
dF = fi φi .
(Recall that this means that for any v ∈ Tq M ,
dFq (v) = fi φi (v) ∈ TF (q) N.)
If the basis (f1 , . . . , fn ) is understood, then we may identify dF with the
Rn -valued 1-form ⎡ 1⎤
φ
⎢ .. ⎥
φ=⎣ . ⎦
φn
on M .
F∗ -
v ∈ TM T N F∗ v
? F ?
M - N
and
F∗
F ∗φ ∈ T ∗M T ∗N φ
? F ?
M - N.
The pullback map behaves as nicely as one could hope with respect to the
various operations on differential forms, as described in the following theo-
rem.
2.9. Integration and Stokes’s theorem 53
(1) F ∗ (Φ + Ψ) = F ∗ Φ + F ∗ Ψ.
(2) F ∗ (Φ ∧ Ψ) = (F ∗ Φ) ∧ (F ∗ Ψ).
(3) F ∗ (dΦ) = d(F ∗ Φ).
*Exercise 2.45. Prove Theorem 2.44 in the case where Φ, Ψ are 1-forms.
*Exercise 2.46. Suppose that m = n = 2 and that F : M → N is described
in local coordinates by
F (x1 , x2 ) = y 1 (x1 , x2 ), y 2 (x1 , x2 ) .
Show that 1
∂y ∂y 1
1 ∂x1 ∂x2
F dy ∧ dy = 2
∗ 2
dx1 ∧ dx2 .
∂y 1 ∂y 2
∂x ∂x2
(You may recognize this expression from the change of variables formula for
double integrals.)
(2.10) F ∗ ⎝ fI (y) dy i1 ∧ · · · ∧ dy ip ⎠
|I|=p
⎛ ⎛ ⎞⎞
m
∂y i1 ∂y ip ⎠⎠ j1
=⎝ ⎝ fI (y(x)) j1 · · · jp dx ∧ · · · dxjp .
∂x ∂x
|I|=p j1 ,...,jp =1
While we won’t need it for the topics in this book, no discussion of differential
forms is complete without a mention of Stokes’s theorem, which has to do
with the integration of differential forms over manifolds. Generally speaking,
a p-form Φ on a manifold M can be integrated over a p-dimensional sub-
manifold P of M , and pullbacks allow us to define such integrals in terms of
ordinary integrals over regions in Rp . This construction requires the notion
of a “manifold with boundary”, which means more or less exactly what it
sounds like; common examples might include a curve segment in Rn (includ-
ing its endpoints) or a hemisphere in R3 (including the boundary circle). It
also requires that the submanifold P be orientable, which means that P can
54 2. Differential forms
*Exercise 2.48. Show that the following theorems are all special cases of
Stokes’s theorem:
(a) The Fundamental Theorem of Calculus: If f : [a, b] → R is a differen-
tiable function, then
& b
f (x) dx = f (b) − f (a).
a
2.10. Cartan’s lemma 55
(Hint: Let M ⊂ R be the closed interval [a, b]. The oriented boundary of M
is the formal sum (b) − (a), and the “integral” of a function f : [a, b] → R
over ∂M is f (b) − f (a).)
(b) Green’s theorem: Let D ⊂ R2 be a region bounded by a simple closed
curve C (oriented counterclockwise), and let P, Q : D → R be differentiable
functions. Then
' &&
(P dx + Q dy) = (Qx − Py ) dx ∧ dy.
C D
Remark 2.51. Technically, this proof only works if the tangent bundle
T M has a global trivialization, because otherwise there might not exist
complementary 1-forms η n+1 , . . . , η s that are globally defined on M . But
such 1-forms always exist on some neighborhood of any given point in M ,
so this argument can be used to prove the lemma in general by showing that
equation (2.13) holds in a neighborhood of any point in M and hence on all
of M .
The final operation that we will define on differential forms is the Lie deriv-
ative. This is a generalization of the notion of the directional derivative of
a function.
Suppose that v is a vector field on a manifold M , and let ϕ be the flow of
v. This is, by definition, the unique map ϕ : V → M , defined on some open
neighborhood V of the set M × {0} in M × R, that satisfies the conditions
∂ϕ
(2.14) (q, t) = v(ϕ(q, t)),
∂t
ϕ(q, 0) = q.
2.11. The Lie derivative 57
ϕ∗t Φ − Φ
(2.15) Lv Φ = lim ,
t→0 t
where ϕ∗t Φ denotes the pullback of the p-form Φ via the map ϕt .
With the interior product in hand, the following theorem tells us how to
compute the Lie derivative:
Theorem 2.55. Given a p-form Φ and a vector field v on a manifold M ,
(2.16) Lv φ = v dφ + d(v φ).
The formula (2.16) is called Cartan’s formula for the Lie derivative.
Exercise 2.56. Use the following steps to prove Cartan’s formula for the
Lie derivative:
(a) Prove that the Lie derivative (Definition 2.52) is a derivation of order
zero; i.e., it takes p-forms to p-forms and satisfies the Leibniz rule
Lv (Φ ∧ Ψ) = (Lv Φ) ∧ Ψ + Φ ∧ (Lv Ψ).
[(v ) ◦ d + d ◦ (v )] (Φ ∧ Ψ)
= [(v ) ◦ d + d ◦ (v )] (Φ) ∧ Ψ + Φ ∧ [(v ) ◦ d + d ◦ (v )] (Ψ).
(d) It is shown in [Wil61] that any two derivations of the same degree that
agree on 0-forms and 1-forms must be equal. Use this fact to conclude that
Cartan’s formula holds.
declares omega and theta to be 1-forms and Omega to be a 2-form. You can
declare functions (i.e., 0-forms); e.g.,
> Form(f=0);
but it isn’t necessary; anything which hasn’t been declared with a Form
command is automatically assumed to be a 0-form. You can also declare
constants by assigning them to have degree −1 (this is purely a Maple
convention and has nothing to do with their degree as differential forms!);
e.g., the command
> Form(a=-1, b=-1);
declares a and b to be constants. Finally, if you want to declare an object
to be a 1-form (certainly the most common use of this command), the “=1”
can be omitted; e.g., the command
> Form(omega, theta, eta);
declares the objects omega, theta, and eta to be 1-forms.
The &ˆ command. This is the command for wedge product. For in-
stance, the commands
> omega:= x*d(y);
> theta:= y*d(z);
> omega &ˆ theta;
can be used to compute the wedge product of the 1-forms x dy and y dz.
Maple will choose a default ordering of forms for wedge products; e.g., the
command
> d(y) &ˆ d(x);
returns
If you don’t like the order that Maple chooses, you can change it with the
Forder command; see Maple’s help page for the Cartan package for more
details. Furthermore, if you don’t like this symbol for the wedge product,
you can change it with the WedgeProduct command.
The d command. This is the all-purpose exterior derivative command.
(WARNING: When you’re using the Cartan package, don’t use the letter d
as a variable or Maple will get hopelessly confused!) It can be used both
for computation and for assignment, and it knows how to use the chain rule
2.12. Introduction to the Cartan package for Maple 61
If you like, you can clean this up a little bit by using the PDETools[declare]
command to tell Maple that f is a function of the variables (x, y, z):
> PDETools[declare](f(x,y,z));
> d(f(x,y,z));
returns
fx d(x) + fy d(y) + fz d(z)
Note that for coordinate 1-forms, you must type, e.g., d(x) rather than dx;
Maple will regard dx as a completely different variable having nothing to
do with the variable x or its exterior derivative.
An important feature of the Cartan package is that exterior derivatives
can be assigned as well as computed, and no explicit local coordinates are
required. (The value of this feature for the method of moving frames will
become apparent in later chapters!) For instance, the commands
> Form(omega, theta, eta);
> d(omega):= theta &ˆ eta;
declare that omega, theta, and eta are 1-forms and that the exterior de-
rivative of omega is the 2-form (theta &ˆ eta).
The Simf command. This is the all-purpose simplification command
for differential forms, and you should use it liberally. (Maple fails to make
some fairly obvious simplifications without it.) For instance, the command
> d(xˆ2*d(y) + yˆ2*d(x));
returns
−2y (d(x) &ˆ d(y)) + 2x (d(x) &ˆ d(y))
62 2. Differential forms
where form2 contains no terms involving omega and returns form1. For
instance, the command
> pick(x*d(y) &ˆ d(z) + y*d(z) &ˆ d(x) + z*d(x) &ˆ d(y),
d(x));
returns
−z dy + y dz
for the 1-forms on R5 . (Don’t worry about why you might want to do such
a thing; it has to do with an exterior differential system representing the
partial differential equation
zxy = ez ,
but that’s well beyond the scope of this book!) You could simply make
assignments such as
> theta:= d(z) - p*d(x) - q*d(y);
etc., but then everything that follows would be expressed in terms of the
coordinate basis. If you want to do something like computing dθ and ex-
pressing it in terms of the basis (θ, ω 1 , ω 2 , π1 , π2 ), it’s more effective to set
up substitutions to go back and forth between the two bases. You can do
this as follows. First, you need to declare the 1-forms in your new basis:
> Form(theta, omega1, omega2, pi1, pi2);
(You don’t have to declare the coordinate basis; the functions x, y, z, p, q
are automatically assumed to be 0-forms, so their exterior derivatives are
1-forms.) Then define the substitution:
> sub1:= [theta = d(z) - p*d(x) - q*d(y), omega1 = d(x),
omega2 = d(y), pi1 = d(p) - exp(z)*d(y),
pi2 = d(q) - exp(z)*d(x)];
You can now use the substitution sub1 to go from the basis (θ, ω 1 , ω 2 , π1 , π2 )
to the coordinate basis; e.g., the command
> Simf(subs(sub1, theta));
yields
d(z) − p d(x) − q d(y)
Where the makebacksub command comes in is when you want to go the
other way. The command
> backsub1:= makebacksub(sub1);
produces a substitution backsub1 that is the inverse of the substitution
sub1; thus, it will go from the coordinate basis to the basis (θ, ω 1 , ω 2 , π1 , π2 ).
So, e.g., the command
> Simf(subs(backsub1, d(z)));
yields
θ + p ω1 + q ω2
One warning about the makebacksub command: It only works properly
when the substitution sub1 is a complete list of the elements of one basis of
2.12. Introduction to the Cartan package for Maple 65
1-forms expressed in terms of another basis. So, for instance, if you define
> sub2:= [theta = d(z) - p*d(x) - q*d(y),
pi1 = d(p) - exp(z)*d(y), pi2 = d(q) - exp(z)*d(x)];
without including omega1 and omega2 in the list, the command
> makebacksub(sub2);
returns a less than helpful backwards substitution, and it might not even
do it consistently if you execute it more than once. (Try it and see what
happens!)
You can use these two substitutions to express dθ in terms of the basis
(θ, ω 1 , ω 2 , π1 , π2 ) via the following sequence of commands:
> Simf(subs(sub1, theta));
> Simf(d(%));
(d(x)) &ˆ (d(p)) + (d(y)) &ˆ (d(q))
ω1 &ˆ π1 + ω2 &ˆ π2
Homogeneous spaces
3.1. Introduction
In the late nineteenth century, there were several different types of geometry
under investigation: There was the classical Euclidean geometry with its
standard notions of lengths and angles, various non-Euclidean geometries in
which the parallel postulate was replaced by alternative versions and lengths
were measured differently than in Euclidean geometry, and even affine and
projective geometries, where lengths and angles weren’t well-defined notions.
In 1872, Felix Klein published a revolutionary treatise on geometry
([Kle93b]; an English translation is available in [Kle93a]), in which he pro-
posed that the most useful way to study a geometric structure is to study its
group of symmetries, i.e., the group of transformations that preserve the key
features of the structure. This approach revolutionized the study of geome-
try, and it continues to influence the development of the subject today. For
instance, when studying curves in R3 (with the standard Euclidean metric
on R3 ), you probably learned the following theorem:
69
70 3. Homogeneous spaces
smooth curves in R3 : They are exactly the properties of smooth curves that
remain unchanged when a curve is transformed by a rigid motion.
Remark 3.2. Technically, translations and rotations are the orientation-
preserving symmetries of Euclidean space. Reflections are also symmetries
of Euclidean space, but they reverse orientation. (They also reverse certain
other quantities, such as the sign of the torsion of a curve.) It is generally
advantageous to restrict consideration to orientation-preserving symmetries,
mainly because doing so typically allows us to work with connected Lie
groups. (Think SO(n) vs. O(n).)
Definition 3.4 makes it sound as though there might be many different Eu-
clidean spaces corresponding to different inner products, but in fact they
are all equivalent, as the following exercise shows.
Exercise 3.6. Let ·, · be any inner product on Rn , and construct a basis
(e1 , . . . , en ) for Rn inductively as follows:
(Initial step) Choose any nonzero vector v1 ∈ Rn , and set
v1
e1 = .
v1 , v1
Let (e1 )⊥ denote the orthogonal complement of e1 with respect to ·, ·; i.e.,
(e1 )⊥ = {v ∈ Rn | v, e1 = 0}.
72 3. Homogeneous spaces
the n×n identity matrix), so that tx represents the translation tx (y) = y+x.
Then for any element h ∈ H0 , define
φ(h) = tx ht−1
x ∈ E(n).
Exercise 3.9. (a) Show that for any h ∈ H0 , we have φ(h) ∈ Hx ; i.e.,
φ(h)(x) = x. Therefore, φ is a map from H0 to Hx .
(b) Show that φ : H0 → Hx is a homomorphism; i.e., show that for any two
elements h1 , h2 ∈ H0 , we have
φ(h1 h2 ) = φ(h1 )φ(h2 ).
(c) Show that φ : H0 → Hx is injective; i.e., show that if φ(h) is the identity
element in Hx , then h must be the identity element in H0 .
(d) Show that φ : H0 → Hx is surjective; i.e., show that for any h̃ ∈ Hx ,
there exists h ∈ H0 with φ(h) = h̃.
and we say that all the isotropy groups Hx ⊂ E(n) are conjugate in E(n).
Moreover, they are all isomorphic to SO(n).
3.3.1. The orthonormal frame bundle. Another way to look at all this
is in terms of orthonormal frames on En .
Definition 3.12. An (oriented) orthonormal frame f on En is a list of
vectors f = (x; e1 , . . . , en ), where x ∈ En and (e1 , . . . , en ) is an oriented,
orthonormal basis for the tangent space Tx En . Alternatively, we may say
that (e1 , . . . , en ) is a orthonormal frame based at x.
SO(n) - E(n)
π
?
En ∼
= E(n)/SO(n).
In this context, E(n) is also called the (oriented) orthonormal frame bundle
of En , and it is denoted F (En ).
Remark 3.13. In fact, this requires some abuse of notation. The function x
is legitimately En -valued, but the functions (e1 , . . . , en ) actually take values
in the tangent bundle T En , and their derivatives—which we will get to
shortly—take values in the bundle T (T En ). This distinction is sometimes
important, but for the most part we will ignore it by making use of the
canonical isomorphisms Tx En ∼ = En , which allow us to regard (e1 , . . . , en )
as E -valued functions.
n
Now, consider the exterior derivatives dx, dei of the functions x, ei on F (En ).
(Recall from Remark 2.39 that these are simply the differentials of the maps
x, ei : F (En ) → En .) For any point f = (x; e1 , . . . , en ) ∈ F (En ), these are
maps
dx : Tf F (En ) → Tx En ,
dei : Tf F (En ) → Tei En ∼= Tei (Tx En ) ∼
= Tx En .
dx = ei ω i ,
(3.1)
dei = ej ωij ,
where 1 ≤ i, j, ≤ n.
Remark 3.14. While it may look strange to write the scalar-valued 1-forms
after the vectors in these equations, order is important here: For example,
the first equation may be written as the matrix product
⎡ 1⎤
ω
⎢ . ⎥
(3.2) dx = e1 · · · en ⎣ .. ⎦ ,
ωn
and it wouldn’t make sense in the other order without interchanging the
roles of row and column vectors.
3.3. Orthonormal frames on Euclidean space 77
In other words, the 1-forms (ω i , ωji ) on F (En ) are defined by the property
that for any f = (x; e1 , . . . , en ) ∈ F (En ), v ∈ Tf F (En ),
dx(v) = ei ω i (v) ∈ Tx En ,
dei (v) = ej ωij (v) ∈ Tei (Tx En ) ∼
= Tx En .
Remark 3.15. Note that the 1-forms (ω i , ωji ) are not well-defined on the
base space En since they are defined relative to a particular choice of frame
(e1 , . . . , en ) for the tangent space Tx En .
A = e1 · · · en .
⎡ ⎤
ω1
⎢ ⎥
dx = A ⎣ ... ⎦ .
ωn
Therefore,
⎡ ⎤ ⎡ 1⎤
ω1 dx
⎢ .. ⎥ −1 −1 ⎢ .. ⎥
(3.3) ⎣ . ⎦ = A dx = A ⎣ . ⎦ .
ωn dxn
Similarly, the equations for dei in (3.1) can be combined into the matrix
equation
⎡ 1 ⎤
ω1 · · · ωn1
⎢ .. ⎥
dA = de1 · · · den = e1 · · · en ⎣ ... . ⎦
ω1 · · · ωnn
n
⎡ 1 ⎤
ω1 · · · ωn1
⎢ .. ⎥ .
= A ⎣ ... . ⎦
ω1n · · · ωnn
Therefore,
⎡ ⎤
ω11 · · · ωn1
⎢ .. .. ⎥ = A−1 dA.
(3.4) ⎣ . . ⎦
ω1n · · · ωnn
where θ is the angle between e1 and the standard basis vector e1 = ∂x∂ 1 .
Thus, θ may be regarded as a local coordinate on the 1-dimensional fibers
of the orthonormal frame bundle F (E2 ). Now, write
cos(θ) − sin(θ)
A = e1 e2 = .
sin(θ) cos(θ)
The 1-forms (ω 1 , . . . , ω n ) are often called the dual forms on the orthonormal
frame bundle F (En ) because they have the property that at any point f ∈
F (En ),
i i 1, i = j,
(3.7) ω (ej ) = δj =
0, i = j.
Exercise 3.18. Verify equations (3.7) by direct computation.
The dual forms also have the property that ω i (v) = 0 for any vector v ∈
T F (En ) that is tangent to the fibers of the projection π : F (En ) → En . The
technical way to say this is that the pullback of ω i to any fiber π −1 (x0 ) of π
via the inclusion map ι : π −1 (x0 ) → F (En ) vanishes.
Exercise 3.19. Prove this statement. (Hint: Let α : I → F (En ) be a curve
tangent to v. Since v is tangent to some fiber π −1 (x0 ) ⊂ F (En ), you can
assume that α(I) ⊂ π −1 (x
0 ). So, what is the value of x at any point α(t)?
d
Now compute dx(v) = dt t=0 x(α(t)).)
Forms with this property (i.e., their pullbacks to each fiber of π vanish) are
called semi-basic for the projection π; for this reason, the dual forms are also
sometimes called the semi-basic forms on F (En ). The pullbacks of (ωji ), on
the other hand, form a basis for the 1-forms on each fiber of π. They are
called the connection forms on F (En ).
Now for the fun part: Start differentiating! In order to compute the exterior
derivatives of the dual forms and connection forms, we need to differentiate
equations (3.1).
*Exercise 3.20. Differentiate equations (3.1) (taking the second equation
into account!) and do some careful index-juggling to obtain
0 = ei (ωji ∧ ω j + dω i ),
0 = ei (ωki ∧ ωjk + dωji ).
80 3. Homogeneous spaces
3.3.3. The Maurer-Cartan form. This all fits into a larger structure,
which is easier to see if we go back to regarding F (En ) as the Lie group
E(n).
Exercise 3.24. Prove that the Lie algebra of E(n) is
( )
0 t0
e(n) = : B ∈ so(n), b ∈ E .
n
b B
(Recall that B ∈ so(n) if and only if B is a skew-symmetric n × n matrix.)
While this definition is fairly simple in theory, it is not so easy to work with
computationally. The Maurer-Cartan form is more commonly written as
(3.10) ω = g −1 dg.
This notation requires some explanation. The variable g here essentially
denotes the identity map g : E(n) → E(n), but it really should be thought of
as a coordinate function on E(n), which realizes any element of the abstract
Lie group E(n) as its (n + 1) × (n + 1) matrix representation. Specifically,
for any f = (x; e1 , . . . , en ) ∈ E(n),
1 0 ··· 0
(3.11) g(f ) = .
x e1 · · · en
Now, let f0 be any element of E(n), and let g0 = g(f0 ). (So f0 and g0
represent the same element of E(n); the difference in notation is meant to
suggest that f0 is an element of the abstract group and g0 is its matrix
representation.) Differentiating (3.11) shows that the mapping
dg : Tf0 E(n) → Tg0 E(n)
can be written as the matrix-valued 1-form
0 0 ··· 0
(3.12) dg = .
dx de1 · · · den
Since g is essentially the identity map, the same is true for dg: It simply
identifies any tangent vector v ∈ Tf0 E(n) with its matrix representation as
a tangent vector to the matrix group E(n) at g0 . So
dg(v) = v ∈ Tg0 E(n).
The left multiplication by g −1 , applied to a vector v ∈ Tg0 E(n), then means
to multiply the matrix representation for v by the matrix representation
for g0−1 .
82 3. Homogeneous spaces
As you might suspect, it is no accident that we have used the letter ω both
for the scalar-valued dual and connection forms and for the matrix-valued
Maurer-Cartan form. The following exercise shows how they are related:
*Exercise 3.27. (a) Show that the Maurer-Cartan form on E(n) is given
by
⎡ ⎤
0 0 ··· 0
⎢ω1 ω1 · · · ω1 ⎥
⎢ 1 n⎥
ω=⎢ . .. .. ⎥ .
⎣ .. . . ⎦
ω n ω1n · · · ωnn
(Hint: Use equations (3.1), (3.11), (3.12) and the fact that ω = g −1 dg is
equivalent to dg = gω.) Write out this matrix explicitly in the case n = 3.
(b) Use the skew-symmetry of the forms (ωji ) to confirm that for any v ∈
T E(n), we have ω(v) ∈ e(n). (This is what it means to say that ω is
“e(n)-valued”.)
0 0 1
84 3. Homogeneous spaces
to compute the dual forms (ω̄ i ) and the connection forms (ω̄ji ) for this frame
field. Show by direct computation that these forms satisfy the structure
equations (3.8).
Exercise 3.32. Repeat Exercise 3.31 for the spherical frame field
⎡ ⎤ ⎡ ⎤ ⎡ ⎤
cos(ϕ) cos(θ) − sin(θ) − sin(ϕ) cos(θ)
⎢ ⎥ ⎢ ⎥ ⎢ ⎥
e1 = ⎢ ⎥
⎣ cos(ϕ) sin(θ) ⎦ , e2 = ⎢ ⎥
⎣ cos(θ) ⎦ , e3 = ⎢ ⎥
⎣ − sin(ϕ) sin(θ) ⎦ ,
sin(ϕ) 0 cos(ϕ)
where (ρ, ϕ, θ) are spherical coordinates on E3 (minus the z-axis), related to
the usual Euclidean coordinates (x, y, z) by
x = ρ cos(ϕ) cos(θ), y = ρ cos(ϕ) sin(θ), z = ρ sin(ϕ).
Let’s take a moment to summarize our work so far: We started with Eu-
clidean space En , i.e., the vector space Rn endowed with an inner product
·, ·. By considering the group of symmetries of this structure, we arrived at
a description of En as the set of left cosets of the closed subgroup SO(n) of
the Lie group E(n). These particular groups arose as the symmetry group of
the structure (E(n)) and the isotropy group of a particular point (SO(n)).
This description of En as the group quotient E(n)/SO(n) realizes En as a
homogeneous space; it is a special case of the following general construction:
Definition 3.33. A homogeneous space is the set G/H of left cosets of a
closed subgroup H of a Lie group G, endowed with the unique manifold
structure with respect to which the natural action of G on G/H is smooth.
(The existence and uniqueness of this manifold structure on G/H is a con-
sequence of a theorem of Cartan [Car30] which states that every closed
subgroup of a Lie group is a Lie subgroup.) The Lie group G is called the
symmetry group of the homogeneous space G/H.
defined by
π(g) = gH
describes G as a principal bundle over G/H with fiber H:
H - G
π
?
G/H.
Just as we identified the group E(n) with the set of orthonormal frames on
Euclidean space, there is often some natural way to think of the group G
as the set of “frames” on the space G/H. The Maurer-Cartan form on G is
defined by
ω(v) = (Lg−1 )∗ (v)
for any v ∈ Tg G, and if G is a matrix group, then we can write
ω = g −1 dg.
Either way, ω takes values in the Lie algebra g of G, and it satisfies the
Maurer-Cartan equation:
(3.13) dω = −ω ∧ ω.
This equation will turn out to play a crucial role in the geometry of sub-
manifolds of the space G/H.
What distinguishes Minkowski space from Euclidean space is its inner prod-
uct structure: Instead of a positive definite inner product, Minkowski space
is equipped with an indefinite inner product.
Given any symmetric bilinear form Q on the vector space Rn , there exists
a basis (e1 , . . . , en ) for Rn and integers p, q ≥ 0 with the property that for
any v = v i ei ∈ Rn ,
Q(v, v) = (v 1 )2 + · · · + (v p )2 − (v p+1 )2 − · · · − (v p+q )2 .
The integers p, q are invariants of Q, and the ordered pair (p, q) is called the
signature of Q. The form Q is nondegenerate if and only if p + q = n. A
positive definite form Q has signature (n, 0), whereas Q is called indefinite
if p and q are both greater than zero.
Definition 3.35. The (1 + n)-dimensional Minkowski space M1,n is the
vector space Rn+1 endowed with a nondegenerate, symmetric bilinear form
·, · with signature (1, n). This bilinear form is called the Minkowski inner
product on M1,n .
Equivalently, for any basis (e0 , . . . , en ) for Rn+1 , the symmetric matrix
g = [gαβ ] = [eα , eβ ]
that describes the Minkowski inner product ·, · in terms of this basis has
one positive eigenvalue and n negative eigenvalues.
Remark 3.36. Even when n has a fixed value, say n = 3, the Minkowski
space M1,3 is still often referred to as “(1+3)-dimensional Minkowski space”
to emphasize the distinction between the time and space dimensions.
Remark 3.37. The Minkowski inner product is sometimes taken to have
signature (n, 1) rather than (1, n). Our convention is the same as that used
in [Cal00]; it is chosen so that tangent vectors v to curves corresponding
to the world lines of particles in spacetime will have v, v > 0.
Remark 3.38. For clarity, we will generally use Roman letters (i, j, etc.)
for indices that range from 1 to n and Greek letters (α, β, etc.) for indices
that range from 0 to n.
Just as in the Euclidean case, all Minkowski inner products on Rn+1 are
equivalent, as the following exercise shows. (Cf. Exercise 3.6.)
Exercise 3.39. Let ·, · be any Minkowski inner product on Rn+1 . Con-
struct a basis (e0 , . . . , en ) for Rn+1 as follows: Choose any nonzero vector
v0 ∈ Rn+1 with v0 , v0 > 0 , and set
v0
e0 = .
v0 , v0
3.5. Minkowski space 87
Let (e0 )⊥ denote the orthogonal complement of e0 with respect to ·, ·; i.e.,
(e0 )⊥ = {v ∈ Rn+1 | v, e0 = 0}.
Henceforth, we will let (e0 , . . . , en ) denote the standard basis for M1,n .
Nonzero vectors in M1,n are divided into three types:
Definition 3.40. A nonzero vector v ∈ M1,n is called
(See Figure 3.1; note that the e0 -axis is drawn as the vertial axis.)
3.5.2. Symmetries and isotropy groups. Let M1,n denote the Minkow-
ski space of dimension 1 + n with the standard inner product (3.14). Just
as we did for Euclidean space, we ask: What are the symmetries of M1,n ?
3.5. Minkowski space 89
O(1, n) =
⎧ ⎡ ⎤ ⎡ ⎤⎫
⎪
⎪ 1 0 ··· 0 1 0 ··· 0 ⎪
⎪
⎪
⎨ ⎢0 −1 · · · ⎥ ⎢0 −1 · · · ⎥⎬⎪
⎢ 0 ⎥ ⎢ 0 ⎥
A ∈ GL(n + 1) : tA ⎢ . .. ..⎥ A = ⎢ .. .. .. ⎥ .
⎪
⎪ ⎣ .. . .⎦ ⎣. . . ⎦⎪⎪
⎪
⎩ ⎪
⎭
0 0 ··· −1 0 0 ··· −1
Furthermore, show that O(1, n) consists of all matrices of the form
A = e 0 · · · en ,
Remark 3.45. Unlike O(n), which has two connected components, O(1, n)
has four connected components. The connected component that contains
the identity matrix is denoted SO + (1, n), and it consists of those transfor-
mations in O(1, n) that have determinant equal to 1 and map the vector e0 to
a future-pointing vector. These transformations are called proper (because
they preserve the orientation of the spatial dimensions) and orthochronous
(because they preserve the orientation of the time direction). SO+ (1, n) is
therefore referred to as the “proper, orthochronous Lorentz group” or, more
succinctly, the “restricted Lorentz group”. We will restrict our attention to
the proper, orthochronous symmetries of M1,n , and for simplicity we will
refer to elements of this group as Lorentz transformations.
(c) Experiment with the action of SO+ (1, 1) on various shapes in the plane.
For instance, consider the unit circle, parametrized by
x(t) = t[x0 (t), x1 (t)] = t[cos(t), sin(t)].
Use Maple to plot the curve Ax(t), where A ∈ SO+ (1, 1), for various values
of the group parameter θ.
Exercise 3.47. Show that the Lie algebra so(1, n) of O(1, n) is given by
so(1, n) =
⎧ ⎡ ⎤ ⎡ ⎤ ⎫
⎪
⎪ 1 0 ··· 0 1 0 ··· 0 ⎪
⎪
⎪
⎨ ⎢0 −1 · · · ⎥ ⎢0 −1 · · · ⎥ ⎪
⎬
⎢ 0 ⎥ ⎢ 0 ⎥
B ∈ M(n+1)×(n+1) : tB ⎢ . .. ..⎥ + ⎢ .. .. .. ⎥B = 0 .
⎪
⎪ ⎣ .. . .⎦ ⎣. . . ⎦ ⎪
⎪
⎪
⎩ ⎪
⎭
0 0 ··· −1 0 0 ··· −1
In particular, this implies that the entries (bαβ ) of B satisfy the relations
b00 = 0,
bi0 = b0i , i = 1, . . . , n,
bji = −bij , i, j = 1, . . . , n.
Since the Minkowski inner product is also preserved by translation, the full
symmetry group of M1,n consists of all transformations of the form
ϕ(x) = Ax + b,
where A ∈ SO + (1, n) and b ∈ M1,n . These transformations form a Lie
group, called the Poincaré group M (1, n), which can be represented as
( )
1 t0
M (1, n) = : A ∈ SO (1, n), b ∈ M
+ 1,n
.
b A
1
dimensional vector , then elements of M (1, n) act by matrix multiplica-
x
tion:
1 t0 1 1
= .
b A x Ax + b
Once again, we ask: Given a vector x ∈ M1,n , what is the isotropy group of
x?
*Exercise 3.48. Show that:
(a) The isotropy group H0 of 0 ∈ M1,n is
( t )
1 0
H0 = : A ∈ SO (1, n) ∼
+
= SO + (1, n).
0 A
3.5. Minkowski space 91
The same reasoning as in the Euclidean case shows that the set of orthonor-
mal frames may be regarded as the Poincaré group M (1, n) via the one-to-
one correspondence
1 0 ··· 0
g(x; e0 , . . . , en ) = .
x e0 · · · en
We can define a projection map π : M (1, n) → M1,n by
π(x; e0 , . . . , en ) = x;
the fiber of this map is the set of all orthonormal frames based at x.
SO + (1, n) acts freely and transitively on each fiber, and so this map gives
an explicit description of M (1, n) as a principal bundle over M1,n with fiber
group SO+ (1, n):
SO + (1, n) -M (1, n)
π
?
M1,n ∼
= M (1, n)/SO+ (1, n).
In this context, M (1, n) is also called the orthonormal frame bundle of M1,n ,
and it is denoted F (M1,n ).
The Maurer-Cartan forms on M (1, n) are defined exactly as in the Euclidean
case by equations (3.1). The structure equations (3.8) are also the same as
in the Euclidean case; the only difference is that with the indefinite inner
product, the connection forms (ωβα ) are no longer quite skew-symmetric.
92 3. Homogeneous spaces
(1) ω00 = 0;
(2) ω0i = ωi0 for i = 1, . . . , n;
3.6.2. Symmetries and isotropy groups. Now we ask: What are the
symmetries of equi-affine space? (It turns out that “orientation-preserving”
is an unneccesary stipulation here, because any volume-preserving transfor-
mation necessarily preserves orientation.) Since equi-affine space has less
structure that must be preserved by a symmetry than Euclidean space does,
we might expect that the symmetry group would be larger.
*Exercise 3.57. Let A ∈ GL(n). Show that for any basis (e1 , . . . , en ) for
An , we have
dV (Ae1 , . . . , Aen ) = (det A) dV (e1 , . . . , en ).
Therefore, dV (Ae1 , . . . , Aen ) = dV (e1 , . . . , en ) if and only if A ∈ SL(n).
Since the volume form is also preserved by translation, the full symmetry
group of An consists of all transformations ϕ : An → An of the form
ϕ(x) = Ax + b,
where A ∈ SL(n) and b ∈ An . These transformations form a Lie group,
called the equi-affine group or special affine group A(n), which can be rep-
resented as ( )
1 t0
A(n) = : A ∈ SL(n), b ∈ A .
n
b A
space Tx An that spans a parallelepiped of volume 1. (We may also say that
(e1 , . . . , en ) is a unimodular frame based at x.)
The same reasoning as in the Euclidean case shows that the set of unimodu-
lar frames may be regarded as the equi-affine group A(n) via the one-to-one
correspondence
1 0 ··· 0
g(x; e1 , . . . , en ) = .
x e1 · · · en
We can define a projection map π : A(n) → An by
π(x; e1 , . . . , en ) = x;
the fiber of this map is the set of all unimodular frames based at x. SL(n)
acts freely and transitively on each fiber, and so this map gives an explicit
description of A(n) as a principal bundle over An with fiber group SL(n):
SL(n) - A(n)
π
?
An ∼
= A(n)/SL(n).
In this context, A(n) is also called the unimodular frame bundle of An , and
it is denoted F (An ).
Just as for Euclidean and Minkowski spaces, the Maurer-Cartan forms on
A(n) are defined by equations (3.1). The structure equations (3.8) are also
the same as before, but without any sort of inner product structure, there
is no symmetry or skew-symmetry condition on the connection forms (ωji ).
There is, however, one relation among the (ωji ) that comes from the uni-
modular condition
det e1 · · · en = 1.
We can differentiate this condition more easily if we express it as
(3.15) e1 ∧ · · · ∧ en = e1 ∧ · · · ∧ en ,
where (e1 , . . . , en ) denotes the standard basis for An . (Recall that wedge
product is simply the skew-symmetrization of the tensor product.)
Exercise 3.60. Show that equation (3.15) holds for any unimodular basis
by showing that for any basis (e1 , . . . , en ) of An , unimodular or not, we have
e1 ∧ · · · ∧ en = det e1 · · · en e1 ∧ · · · ∧ en .
*Exercise 3.61. Differentiate equation (3.15) and conclude that the con-
nection forms (ωji ) satisfy the trace condition ωii = 0. (Hint: Don’t let the
96 3. Homogeneous spaces
In order to see that this captures the idea of “Rn plus points at infinity”,
we will use homogeneous coordinates [x0 : · · · : xn ] (cf. Example 1.9) to
represent a point in Pn . Consider the dense, open subset V0 of Pn given by
V0 = {[x0 : · · · : xn ] ∈ Pn | x0 = 0}.
Any point [x] ∈ V0 can be written as
x1 xn
[x] = [x : · · · : x ] = 1 : 0 : · · · : 0
0 n
x x
3.7. Projective space 97
!
x1 n
and so may be represented by the point x0
, . . . , xx0 ∈ Rn . We will denote
this point by x̄ = (x̄1 , . . . , x̄n ); these are referred to as affine coordinates on
the open set V0 . We can visualize this by observing that any line in Rn+1
that passes through the origin and another point x with x0 = 0 intersects
the plane x0 = 1 in exactly one point. (See Figure 3.2.)
Figure 3.2. V0 ⊂ Pn
A ∈ GL(n + 1) fixes every line in Rn+1 —and therefore acts as the identity
transformation on Pn —if and only if A = λI for some λ = 0. Moreover, any
two matrices A, B ∈ GL(n + 1) induce the same transformation on Pn if and
only if A = λB for some λ = 0. Therefore, the most natural choice for the
symmetry group of Pn is the quotient group GL(n + 1)/ ∼, where A ∼ B if
and only if A = λB for some λ = 0.
Remark 3.67. The group GL(m)/ ∼ is called the projective general linear
group, denoted P GL(m); similarly, the group SL(m)/ ∼ is called the pro-
jective special linear group, denoted P SL(m). When m is odd, these two
groups are identical and isomorphic to SL(m); when m is even, P GL(m)
is the group (SL(m)/{±I}) {±1} of Exercise 3.66, while P SL(m) is the
proper subgroup SL(m)/{±I} of P GL(m).
*Exercise 3.68. Consider the plane R2 with coordinates (x̄, ȳ), and let the
x̄-axis represent the open set
V0 = {[x0 : x1 ] ∈ P1 | x0 = 0}
1
via the identification x̄ = xx0 . Suppose that an object lies along the line L
with equation ȳ = mx̄+b and that two viewers located at points p = (p1 , p2 )
3.7. Projective space 99
sq
p s
Q
Q
```QQ
``Q
`s``
Q `
r Q ``` L
Q
Q
Q
Q
s Qs
T (x̄) x̄
lying at (x̄, 0), and observer q sees r as lying at (T (x̄), 0). Use the following
steps to compute the transformation T (x̄):
(a) Show that the point r of intersection between L and the line joining the
point (x̄, 0) to p has coordinates
" #
(p2 − b)x̄ + bp1 mp2 x̄ + bp2
r= , .
mx̄ + p2 − mp1 mx̄ + p2 − mp1
(b) Show that the line joining q to r intersects the x̄-axis at the point
(T (x̄), 0), where
αx̄ + β
(3.16) T (x̄) =
γ x̄ + δ
(a) Let
⎡ 0 0 0⎤
a0 a1 a2
⎢ 1 1 1⎥
A = ⎣a0 a1 a2 ⎦ ∈ SL(3).
a20 a21 a22
Show that the transformation
T ([x]) = [Ax]
corresponds to the map
" #
1 2 a10 + a11 x̄1 + a12 x̄2 a20 + a21 x̄1 + a22 x̄2
T (x̄ , x̄ ) = ,
a00 + a01 x̄1 + a02 x̄2 a00 + a01 x̄1 + a02 x̄2
on V0 .
(b) Show that if A has the form
⎡ ⎤
1 0 0
⎢ ⎥
A = ⎣b1 a11 a12 ⎦
b2 a21 a22
with a11 a22 − a12 a21 = 1, then T : V0 → V0 is the equi-affine transformation
T (x̄) = Āx̄ + b̄,
a11 a12 b1
where Ā = 2 2 , b̄ = 2 . Therefore, the equi-affine group A(2) is a
a1 a2 b
proper subgroup of the group of projective transformations.
3.7. Projective space 101
where t[x] is any matrix in SL(n + 1) whose first column is [x]. (t[x] will
then have the property that t[x] ([x0 ]) = [x].)
(c) There is a natural correspondence between Pn and SL(n + 1)/H[x0 ] , the
set of left cosets of H[x0 ] in SL(n + 1).
the tangent space T[e0 ] Pn . (We may also say that (e0 , . . . , en ) is a projective
frame based at [e0 ].) Note that e0 needs to be part of the frame since it is
not uniquely determined by its equivalence class [e0 ] and choosing a different
representative for [e0 ] would affect the scaling of the vectors (e1 , . . . , en ).
The same reasoning as in the Euclidean case shows that the set of projec-
tive frames may be regarded as the group SL(n + 1) via the one-to-one
correspondence
g(e0 , . . . , en ) = e0 · · · en .
We can define a projection map π : SL(n + 1) → Pn by
π(e0 , . . . , en ) = [e0 ];
the fiber of this map is the set of all projective frames based at [e0 ]. H[x0 ]
acts freely and transitively on each fiber, and so this map gives an explicit
description of SL(n+1) as a principal bundle over Pn with fiber group H[x0 ] :
H[x0 ] - SL(n + 1)
π
?
Pn ∼
= SL(n + 1)/H[x0 ] .
where 0 ≤ α, β ≤ n.
The structure equations are
and so the semi-basic forms for π are (ω01 , . . . , ω0n ). The remaining (ωαβ ) are
the connection forms.
In this section, we will explore how the Cartan package may be used to
facilitate some of the computations involved in the exercises in this chapter.
While most of these computations are simple enough to be done by hand,
seeing how to do them in Maple will help you to familiarize yourself with
how the package works.
You should begin by loading the Cartan package into Maple, and the
LinearAlgebra package will be useful as well:
> with(Cartan);
> with(LinearAlgebra);
Exercise 3.32: Define A to be the matrix whose columns are the frame
field vectors:
> A:= Matrix([
[cos(phi)*cos(theta), -sin(theta), -sin(phi)*cos(theta)],
[cos(phi)*sin(theta), cos(theta), -sin(phi)*sin(theta)],
[sin(phi),0,cos(phi)]
]);
Let dx be the vector of the Cartesian coordinate 1-forms:
> dx:= Vector([d(x), d(y), d(z)]);
We know that the dual forms (ω̄ 1 , ω̄ 2 , ω̄ 3 ) are the components of the following
vector:
> dualforms:= simplify(MatrixInverse(A).dx);
We can assign them as follows:
> for i from 1 to 3 do
omega[i]:= dualforms[i];
end do;
104 3. Homogeneous spaces
Similarly, the connection forms (ω̄ji ) are given by the following matrix:
> connectionforms:= simplify(MatrixInverse(A).map(d, A));
Assign these as follows:
> for i from 1 to 3 do
for j from 1 to 3 do
omega[i,j]:= connectionforms[i,j];
end do;
end do;
In order to verify the structure equations for the (dω̄ i ), check that the fol-
lowing quantities are all equal to zero:
> for i from 1 to 3 do
Simf(d(omega[i]) + add(omega[i,j] &ˆ omega[j], j=1..3));
end do;
In order to verify the structure equations for the (dω̄ji ), check that the follow-
ing quantities are all equal to zero. (The print command is to force Maple
to print the output; normally output is suppressed for computations nested
this deeply.)
> for i from 1 to 3 do
for j from 1 to 3 do
print(Simf(d(omega[i,j]) + add(omega[i,k] &ˆ omega[k,j],
k=1..3)));
end do;
end do;
The following two exercises don’t require the Cartan package, but they nicely
illustrate some of Maple’s graphic capabilities, so we will include them here
anyway. First, we need to load the plots package:
> with(plots);
Exercise 3.46, part (c): The general element of SO+ (1, 1) is represented
by the following matrix:
> A:= Matrix([[cosh(theta), sinh(theta)], [sinh(theta),
cosh(theta)]]);
Consider the unit circle, parametrized as follows:
> unitcircle:= Vector([cos(t), sin(t)]);
3.8. Maple computations 105
The image of this curve under the action of the matrix A (for a fixed value
of θ) is given by
> newcurve:= A.unitcircle;
We can now define plots of this curve for various values of θ; by using the
display command, we can view them all on one graph:
> g1:= plot([subs([theta=0], newcurve[1]), subs([theta=0],
newcurve[2]), t=0..2*Pi], scaling=constrained):
g2:= plot([subs([theta=0.5], newcurve[1]),
subs([theta=0.5], newcurve[2]), t=0..2*Pi],
scaling=constrained):
g3:= plot([subs([theta=1], newcurve[1]),
subs([theta=1], newcurve[2]), t=0..2*Pi],
scaling=constrained):
display(g1, g2, g3);
We can animate the resulting curve, using ϕ as the time parameter, to see
how the unit circle transforms through a family of quadrics and back again
as ϕ varies from 0 to π:
> animate(plot, [ [Tx[1], Tx[2], t=0..2*Pi],
view = [-5..5, -5..5], scaling=constrained], phi = 0..Pi,
frames=100);
In order to view the animation, click on the plot, and then from the “Plot”
menu choose “Animation → Play”. (Alternatively, right-click on the plot
and select “Animation → Play” from the pop-up menu.)
Chapter 4
4.1. Introduction
(1) How can we tell when we have found a complete set of invariants?
This is a question about uniqueness: Given two submanifolds of a
homogeneous space, when is it possible to transform one into the
other via an element of the symmetry group? This is also known
as the equivalence problem: When are two submanifolds equivalent
under the action of the symmetry group?
107
108 4. Curves and surfaces in Euclidean space
(2) Once we know what a complete set of invariants should look like,
can they be prescribed arbitrarily? This is a question about exis-
tence: Given prescribed values for the invariants, does there neces-
sarily exist a submanifold whose invariants coincide with the given
values?
In §4.2, we will address the theory underlying the first question, and in §4.3
we will show how it applies to curves in E3 . Then in §4.4 and §4.5, we will
take up the second question and show how the theory as a whole applies to
surfaces in E3 .
G
>
f˜
π
f ?
U - G/H.
It follows that f˜1∗ ω = f˜2∗ ω if and only if dg = 0, i.e., if and only if g(u) is
constant.
So, our approach from here on will go something like this: Given an immer-
sion f : U → G/H, we will look for an equivariant method of constructing
a canonical adapted frame field f˜ : U → G. Then we will examine the
pulled-back Maurer-Cartan form f˜∗ ω, which will contain a complete set of
geometric invariants for the original immersion f . This is known as the
method of moving frames, and we will start by demonstrating how to carry
it out for curves in E3 .
Recall that α is regular if α (t) = 0 for every t ∈ I. The first condition that
we will require in order for α to be “nice enough” is that α must be a regular
curve. With this assumption, we can make our first frame adaptation by
setting
α (t)
e1 (t) = ;
|α (t)|
i.e., we require that e1 (t) be the unit tangent vector to the curve at α(t).
Exercise 4.5. Show that this choice of e1 (t) is equivariant under the action
of E(3): If we replace α by g · α for some g ∈ E(3), then e1 (t) ∈ Tα(t) E3 will
be replaced by (Lg )∗ (e1 (t)) ∈ Tg·α(t) E3 .
The vector e1 (t) is now uniquely determined, but we still have the freedom
to vary the pair (e2 (t), e3 (t)) by an arbitrary rotation in SO(2). We will
112 4. Curves and surfaces in Euclidean space
need to delve deeper into the geometry of the curve α in order to determine
how to choose the remainder of the adapted frame field.
Here we make an observation that will simplify the remainder of our com-
putations. Fix t0 ∈ I and define the arc length function along α to be
& t
s(t) = |α (u)| du.
t0
Exercise 4.6. Show that s(t) is invariant under the action of E(3); that is,
for any g ∈ E(3), the curves α and g · α have the same arc length function.
Since α (t) = 0 for all t ∈ I, the inverse function theorem implies that s(t)
has a differentiable inverse function t(s). By setting α(s) = α(t(s)), we
may assume that α is parametrized by arc length, so that |α (s)| = 1 and
e1 (s) = α (s).
In order to make the next adaptation, we need to make another assumption
about the curve. We will say that α is nondegenerate if α is regular and, in
addition, e1 (s) = 0 for all s ∈ I. In this case, differentiating the equation
e1 (s), e1 (s) = 1
with respect to s yields
e1 (s), e1 (s) = 0.
Thus, e1 (s) is orthogonal to e1 (s), and we can make our second adaptation
by setting
e (s)
e2 (s) = 1 .
|e1 (s)|
This vector is called the unit normal vector to the curve at α(s).
Exercise 4.7. Show that e2 (s) is equivariant under the action of E(3): If
we replace α by g ·α for some g ∈ E(3), then e2 (s) ∈ Tα(s) E3 will be replaced
by (Lg )∗ (e2 (s)) ∈ Tg·α(s) E3 .
The adapted frame field is now uniquely determined: Because the frame
must be oriented and orthonormal, e3 (s) is uniquely determined by the
condition that
e3 (s) = e1 (s) × e2 (s).
The vector e3 (s) is called the binormal vector to the curve at α(s). The
adapted frame field (e1 (s), e2 (s), e3 (s)) is called the Frenet frame of the
curve α(s); it determines a canonical, left-invariant lifting α̃ : I → E(3)
given by
α̃(s) = (α(s); e1 (s), e2 (s), e3 (s))
for any nondegenerate curve α parametrized by arc length. (See Figure 4.1.)
4.3. Moving frames for curves in E3 113
As in Chapter 3, write (ω̄ i , ω̄ji ) for the pulled-back forms (α̃∗ ω i , α̃∗ ωji ). Note
that, since these are all 1-forms on I, they must all be multiples of ds.
We can write the pullbacks of equations (3.1) as
α (s)ds = ei (s) ω̄ i ,
(4.3)
ei (s)ds = ej (s) ω̄ij .
Now recall how we constructed our adapted frame field. First, we chose
e1 (s) so that α (s) = e1 (s); therefore, the first equation in (4.3) implies that
ω̄ 1 = ds, ω̄ 2 = ω̄ 3 = 0.
Then we chose e2 (s) so that e1 (s) is a multiple of e2 (s), say e1 (s) =
κ(s)e2 (s). The function κ(s) is called the curvature of α at s; note that
α is nondegenerate if and only if κ(s) > 0 for all s ∈ I. So the equation for
e1 (s) in (4.3) implies that
ω̄12 = κ(s)ds, ω̄13 = 0.
(Recall that ω̄11 = 0 by the skew-symmetry of the (ωji ).) The only remaining
Maurer-Cartan form is ω̄32 ; it must be equal to some multiple of ds, so define
a function τ (s) by the condition that
ω̄32 = −τ (s)ds.
The function τ (s) is called the torsion of α at s.
114 4. Curves and surfaces in Euclidean space
Remark 4.8. The minus sign in the definition of τ (s) is a convention pre-
ferred by some authors, but it is not universal. This choice of sign has the
feature that it results in a positive value of τ for the standard right-handed
helix
α(s) = t [a cos(s), a sin(s), b] ,
where a, b > 0 and a2 + b2 = 1.
Using the skew-symmetry of the (ω̄ji ), the remaining two equations in (4.3)
become
e2 (s)ds = e1 (s) ω̄21 + e3 (s) ω̄23 = (−e1 (s)κ(s) + e3 (s)τ (s))ds,
e3 (s)ds = e1 (s) ω̄31 + e2 (s) ω̄32 = −e2 (s)τ (s)ds.
Thus, we have the familiar Frenet equations:
⎡ ⎤
0 0 0 0
⎢ ⎥
⎢1 0 −κ(s) 0 ⎥
α (s) e1 (s) e2 (s) e3 (s) = α(s) e1 (s) e2 (s) e3 (s) ⎢⎢ ⎥.
−τ ⎥
⎣0 κ(s) 0 (s) ⎦
0 0 τ (s) 0
Note that if we regard α̃(s) as the matrix
α̃(s) = α(s) e1 (s) e2 (s) e3 (s) ,
then the matrix on the right multiplied by the 1-form ds is equal to
α̃(s)−1 d(α̃(s)),
and so it is exactly the pullback of the Maurer-Cartan form ω = g −1 dg on
E(3) via α̃.
Applying Lemma 4.2 yields the following theorem:
Theorem 4.9. Two nondegenerate curves α1 , α2 : I → E3 parametrized by
arc length differ by a rigid motion if and only if they have the same curvature
κ(s) and torsion τ (s).
under the action of E(4). (Hint: The tricky part is how to choose
e3 (s) so that it is orthogonal to both e1 (s) and e2 (s). For guidance,
use the Frenet equations to convince yourself that for curves in E3 ,
The answer to this existence question is yes, but this result is particular to
1-dimensional submanifolds of homogeneous spaces G/H. It follows from
116 4. Curves and surfaces in Euclidean space
Outline of Proof. The full proof of this lemma requires the Frobenius
theorem and is beyond the scope of this book. (If you’re curious, the proof
may be found in [Gri74].) However, the main idea goes something like
this: f˜∗ ω contains quantities involving derivatives of the unknown function
f˜ : U → G, and for any given g-valued 1-form ω̄ on U , the equation f˜∗ ω = ω̄
may be regarded as a system of partial differential equations for f˜. In
general, this system is overdetermined and may have no solutions. However,
equation (4.4) is precisely the compatibility condition that must be satisfied
in order to guarantee that solutions exist, at least locally. (In this case,
it turns out that a solution is uniquely determined by specifying an initial
condition f˜(u0 ) for any u0 ∈ U .) Once we know that local solutions exist,
a patching argument can be used to construct a solution f˜ on the entire
domain U .
Remark 4.13. Even without the hypothesis that U is simply connected,
the result of Lemma 4.12 holds in some neighborhood of any point u ∈ U ;
simple connectivity is only necessary to ensure that these local solutions
can be patched together to form a single solution that is globally defined
on U . For simplicity of exposition, we will not explicitly state topological
hypotheses on the domain U every time we introduce an immersion f : U →
G/H. But keep in mind that if U is topologically nontrivial, then many
of our constructions may be possible only locally and not globally on U .
For example, because a frame bundle over a topologically nontrivial base
space may have no global sections, it might not be possible to construct an
adapted frame field globally on U . Because of these limitations, the method
of moving frames is a tool best suited to the study of the local geometry of
submanifolds of homogeneous spaces; it has very little to say about global
properties.
Assuming that the conditions of Lemma 4.12 are satisfied, composing the
map f˜ with the natural projection π : G → G/H gives a smooth map
f : U → G/H that, in most cases of interest, realizes M = f (U ) as a
submanifold of the homogeneous space G/H. According to Lemma 4.2,
specifying a g-valued 1-form ω̄ on U is equivalent to prescribing the values
of a complete set of invariants for an unknown submanifold M ⊂ G/H;
4.5. Moving frames for surfaces in E3 117
Lemma 4.12 then gives a necessary and sufficient condition for the existence
of a smooth map f : U → G/H whose image has the prescribed invariants.
Moreover, Lemma 4.2 implies that any such f is unique up to left action by
an element g ∈ G.
*Exercise 4.16. Show how Corollary 4.15 follows from Lemma 4.12. (Hint:
Observe that both sides of equation (4.4) are 2-forms on I.)
Having chosen e3 (u) in this way, e1 (u) and e2 (u) must form a basis for
Tx(u) Σ no matter how we choose them. We will explore how we might refine
our choices later, but for now, we allow (e1 (u), e2 (u)) to be an arbitrary
orthonormal basis of Tx(u) Σ.
Now consider the pullbacks of equations (3.1) to U via x̃. (As in Chapter
3, write (ω̄ i , ω̄ji ) for the pulled-back forms (x̃∗ ω i , x̃∗ ωji ) on U .) Our first
observation about these forms is the following:
Proposition 4.18. Let U ⊂ R2 be an open set, and let x : U → E3 be an
immersion. For any adapted orthonormal frame field (e1 (u), e2 (u), e3 (u))
along Σ = x(U ), the associated dual and connection forms (ω̄ i , ω̄ji ) have the
property that ω̄ 3 = 0.
You may recall that the metric properties of a regular surface in E3 are
encapsulated in the first fundamental form of the surface.
Definition 4.20. Let U ⊂ R2 be an open set, and let x : U → E3 be an
immersion. The first fundamental form of Σ = x(U ) is the quadratic form I
on T U defined by
I(v) = dx(v), dx(v)
for v ∈ Tu U .
4.5. Moving frames for surfaces in E3 119
(b) Convince yourself that ·, ·u is a section of the symmetric tensor bundle
S 2 (T ∗ U ). Any section of this bundle defines a symmetric bilinear form
B : T U × T U → R,
which in turn defines a quadratic form
Q : TU → R
by setting Q(v) = B(v, v).
*Exercise 4.24. If you’ve seen the first fundamental form before, you prob-
ably saw it written as
I = E du2 + 2F du dv + G dv 2 ,
where (u, v) are local coordinates on U and
E = xu , xu , F = xu , xv , G = xv , xv .
Suppose that x : U → E3 is an immersion with F = 0. (Such a parametriza-
tion for a given surface Σ always exists, at least locally; the proof is beyond
the scope of this book but can be found in [dC76]. This assumption isn’t
necessary, but it keeps the calculations simpler.)
120 4. Curves and surfaces in Euclidean space
(c) Show that if (e1 (u), e2 (u), e3 (u)) is replaced by another adapted frame
field (ẽ1 (u), ẽ2 (u), ẽ3 (u)) of the form
ẽ1 = cos(θ) e1 + sin(θ) e2 ,
ẽ2 = − sin(θ) e1 + cos(θ) e2 ,
ẽ3 = e3 ,
˜ 1 , ω̄
then the dual forms (ω̄ ˜ 2 ) of the new adapted frame field are
˜ 1 = cos(θ) ω̄ 1 + sin(θ) ω̄ 2 ,
ω̄
˜ 2 = − sin(θ) ω̄ 1 + cos(θ) ω̄ 2 .
ω̄
Moreover,
I = (ω̄ 1 )2 + (ω̄ 2 )2 = (ω̄
˜ 1 )2 + (ω̄
˜ 2 )2 .
II = e du2 + 2f du dv + g dv 2 ,
where
e = e3 , xuu = −(e3 )u , xu = −de3 ∂u , dx ∂u ,
∂ ∂
∂ ∂
f = e3 , xuv = −(e3 )v , xu = −de3 ∂v , dx ∂u
∂ ∂
= −(e3 )u , xv = −de3 ∂u , dx ∂v ,
∂ ∂
g = e3 , xvv = −(e3 )v , xv = −de3 ∂v , dx ∂v .
Now, we still haven’t figured out how we should choose the vectors (e1 (u),
e2 (u)), except that they should form an orthonormal basis for Tx(u) Σ at each
point. In order to refine our adapted frame field further, we will examine how
the matrix [hij ] changes if we vary the frame. So, let (e1 (u), e2 (u), e3 (u))
be any adapted frame field, with associated Maurer-Cartan forms (ω̄ i , ω̄ji ).
Any other adapted frame field (ẽ1 (u), ẽ2 (u), ẽ3 (u)) has the form (up to sign)
⎡ ⎤
cos(θ) − sin(θ) 0
⎢ ⎥
ẽ1 (u) ẽ2 (u) ẽ3 (u) = e1 (u) e2 (u) e3 (u) ⎣ sin(θ) cos(θ) 0⎦
0 0 1
˜ i , ω̄
for some function θ on U . Let (ω̄ ˜ ji ) be the Maurer-Cartan forms associ-
cos(θ) − sin(θ)
ated to the new frame field, and set B = .
sin(θ) cos(θ)
*Exercise 4.28. (a) Show that the result in part (c) of Exercise 4.24 can
be expressed as
1 1
˜
ω̄ ω̄
(4.5) = B −1 .
˜2
ω̄ ω̄ 2
4.5. Moving frames for surfaces in E3 123
Recall from linear algebra that any symmetric matrix can be transformed
to a diagonal matrix by just such an orthogonal change of basis. Therefore,
for each u ∈ U there exists an adapted frame (e1 (u), e2 (u), e3 (u)) at the
point x(u) ∈ Σ with the property that
h11 (u) h12 (u) κ1 (u) 0
(4.8) =
h12 (u) h22 (u) 0 κ2 (u)
for some real numbers κ1 (u), κ2 (u).
*Exercise 4.29. Let (e1 (u), e2 (u), e3 (u)) be an adapted frame satisfying
(4.8), and let v1 , v2 ∈ Tu U be vectors with the property that dx(vi ) = ei (u)
for i = 1, 2. Show that
d(e3 )u (vi ) = dNx(u) (ei (u)) = −κi (u)ei (u), i = 1, 2.
This implies that e1 (u) and e2 (u) are eigenvectors for the linear transfor-
mation dNx(u) , the differential of the Gauss map N : Σ → S2 at the point
x(u) ∈ Σ, with eigenvalues −κ1 (u), −κ2 (u), respectively.
Therefore, there exists an adapted frame (e1 (u), e2 (u), e3 (u)) at each point
x(u) ∈ Σ with the property that e1 (u) and e2 (u) are principal vectors at
x(u). Such a frame will be called a principal adapted frame at the point
x(u) ∈ Σ, and an adapted frame field on Σ which has this property at every
point x(u) ∈ Σ will be called a principal adapted frame field on Σ.
124 4. Curves and surfaces in Euclidean space
Definition 4.31. If κ1 (u) = κ2 (u) for some point u ∈ U , then the corre-
sponding point x(u) of Σ is called an umbilic point of Σ.
Suppose that Σ = x(U ) has no umbilic points. Now that we (finally!) have
a way of defining a canonical adapted frame field along Σ, we can apply
Lemma 4.2 to find a complete set of invariants for the surface.
Theorem 4.35 (Bonnet). Let U ⊂ R2 be an open set. Two immersions
x1 , x2 : U → E3 without umbilic points differ by a rigid motion if and only
if they have the same first and second fundamental forms.
4.5. Moving frames for surfaces in E3 125
Remark 4.36. This theorem is true even for surfaces with umbilic points,
but the proof is slightly more involved due to the issue of how to choose a
canonical adapted frame field near umbilic points.
Proof. One direction is clear: Since all our constructions are equivariant
under the action of E(3), any two surfaces that differ by a rigid motion must
have the same first and second fundamental forms.
Conversely, suppose that x1 , x2 have the same first and second fundamental
forms. Let x̃1 , x̃2 : U → E(3) be principal adapted frame fields for x1 , x2 ,
respectively; let (ω̄ i , ω̄ji ) denote the pulled-back dual and connection forms
for x1 and let (Ω̄i , Ω̄ij ) denote those for x2 . By hypothesis,
Ix1 = (ω̄ 1 )2 + (ω̄ 2 )2 = (Ω̄1 )2 + (Ω̄2 )2 = Ix2 ,
IIx1 = (κ1 )x1 (ω̄ 1 )2 + (κ2 )x1 (ω̄ 2 )2 = (κ1 )x2 (Ω̄1 )2 + (κ2 )x2 (Ω̄2 )2 = IIx2 .
Equality of the first fundamental forms implies that
1
Ω̄ cos(θ) − sin(θ) ω̄ 1
=
Ω̄2 ± sin(θ) ± cos(θ) ω̄ 2
for some function θ on U , where the signs on the bottom row of the matrix
are the same. Substituting this relation into the equation for the second
fundamental forms yields
(κ1 )x1 (ω̄ 1 )2 + (κ2 )x1 (ω̄ 2 )2 = (κ1 )x2 cos2 (θ) + (κ2 )x2 sin2 (θ) (ω̄ 1 )2
+ 2 (((κ2 )x2 − (κ1 )x2 ) sin(θ) cos(θ)) ω̄ 1 ω̄ 2
+ (κ1 )x2 sin2 (θ) + (κ2 )x2 cos2 (θ) (ω̄ 2 )2 .
Since (κ1 )x2 > (κ2 )x2 , the vanishing of the middle term on the right-hand
side implies that θ is a multiple of π2 . Then the remaining terms, together
with the inequality κ1 > κ2 on both sides, imply that θ is a multiple of π.
Therefore, (κ1 )x2 = (κ1 )x1 , (κ2 )x2 = (κ2 )x1 , and
Ω̄1 = ±ω̄ 1 , Ω̄2 = ±ω̄ 2 .
By making one of the permissible frame changes described in Remark 4.33
on one side or the other if necessary, we can arrange that both signs above
are positive.
Since we now have Ω̄1 = ω̄ 1 , we must have dΩ̄1 = dω̄ 1 . According to the
Cartan structure equations (3.8), this implies that
(Ω̄12 − ω̄21 ) ∧ ω̄ 2 = 0.
By Cartan’s lemma, Ω̄12 − ω̄21 must be a multiple of ω̄ 2 . But the same
reasoning applied to the equation dΩ̄2 = dω̄ 2 implies that Ω̄12 − ω̄21 must also
126 4. Curves and surfaces in Euclidean space
Finally, since x̃1 and x̃2 are both principal adapted frame fields, we have
Ω̄31 = (κ1 )x2 Ω̄1 = (κ1 )x1 ω̄ 1 = ω̄13 ,
Ω̄32 = (κ2 )x2 Ω̄2 = (κ2 )x1 ω̄ 2 = ω̄23 .
The theorem now follows from Lemma 4.2.
Now we consider the question discussed in §4.4; namely, can the first and
second fundamental forms be prescribed arbitrarily? We must require that
I be a positive definite quadratic form (i.e., that I(v) > 0 for every v = 0 ∈
T U ) in order to define a metric on the surface. And in order to avoid the
issue of umbilic points, we will assume that I and II are prescribed in such
a way that IIu is not a scalar multiple of Iu at any point u ∈ U .
Exercise 4.37. Why is this the right condition to impose on I and II in
order to avoid umbilic points?
The form ω̄21 determined by the first two equations in (4.9) is called the
Levi-Civita connection form of the metric defined by the first fundamental
form I = (ω̄ 1 )2 + (ω̄ 2 )2 . The third equation just says that (ω̄13 , ω̄23 ) must be
symmetric linear combinations of (ω̄ 1 , ω̄ 2 ), which will automatically be true
under our assumptions.
The remaining structure equations may be written as
dω̄21 = ω̄13 ∧ ω̄23 ,
(4.10) dω̄13 = ω̄23 ∧ ω̄21 ,
dω̄23 = −ω̄13 ∧ ω̄21 .
The first of these equations is called the Gauss equation, and the last two
are called the Codazzi-Mainardi equations, or simply the Codazzi equations.
By Lemma 4.12, we have the following theorem:
Theorem 4.39 (Bonnet). Let (ω̄ 1 , ω̄ 2 , ω̄13 , ω̄23 ) be 1-forms on a connected and
simply connected open set U ⊂ R2 satisfying the conditions that (ω̄ 1 , ω̄ 2 ) are
linearly independent at each point of U and that ω̄i3 is a scalar multiple of
ω̄ i for i = 1, 2. Suppose that, together with the Levi-Civita connection form
ω̄21 determined by ω̄ 1 and ω̄ 2 , these forms satisfy the Gauss and Codazzi
equations (4.10). Then there exists an immersed surface x : U → E3 , unique
up to rigid motion, whose first and second fundamental forms are
I = (ω̄ 1 )2 + (ω̄ 2 )2 ,
II = ω̄13 ω̄ 1 + ω̄23 ω̄ 2 .
Because of this result, the Gauss and Codazzi equations are also referred to
as the compatibility equations of the theory of surfaces in E3 .
(a) Show that the 1-forms (ω̄ 1 , ω̄ 2 , ω̄13 , ω̄23 ) determined by I and II according
to the conditions of Theorem 4.39 are
(c) Prove that λ is constant. (Hint: Use the structure equations to differen-
tiate equations (4.13), taking into account the fact that we must have
dλ = λ1 ω̄ 1 + λ2 ω̄ 2
for some functions λ1 , λ2 on U . Then use Cartan’s lemma.)
(d) Show that if λ = 0, then de3 = 0. Conclude that the normal vector field
of Σ is constant and that Σ is therefore contained in a plane.
(e) Show that if λ = 0, then d(x+ λ1 e3 ) = 0. Conclude that the vector-valued
function x + λ1 e3 : U → E3 is equal to some constant point q ∈ E3 and that
1
Σ is therefore contained in the sphere of radius |λ| centered at q.
Thus, the only totally umbilic surfaces in E3 are (open subsets of) planes
and spheres.
Exercise 4.43. Suppose that the surface x : U → E3 has the property that
both principal curvatures κ1 , κ2 are constants. We know from Exercise 4.42
that if κ1 = κ2 , then Σ = x(U ) is contained in either a plane or a sphere,
so assume that κ1 = κ2 . Let (e1 (u), e2 (u), e3 (u)) be a principal adapted
frame field on Σ, and let x̃ : U → E(3) denote the corresponding lifting of
x : U → E3 , with associated Maurer-Cartan forms (ω̄ i , ω̄ji ). Then we have
(d) Integrate equations (4.15) (beginning with the equations for de1 , de2 , de3
and working backwards) to show that there exist constant vectors ē1 , ē2 , ē3 ,
x̄ ∈ E3 such that
e1 (u, v) = ē1 ,
e2 (u, v) = cos(κ2 v) ē2 + sin(κ2 v) ē3 ,
(4.16) e3 (u, v) = − sin(κ2 v) ē2 + cos(κ2 v) ē3 ,
1 1
x(u, v) = x̄ + u ē1 + sin(κ2 v) ē2 − cos(κ2 v) ē3 .
κ2 κ2
(e) Use equations (4.16) and the fact that (e1 (u), e2 (u), e3 (u)) is an or-
thonormal frame field to show that (ē1 , ē2 , ē3 ) is an orthonormal frame.
4.5. Moving frames for surfaces in E3 131
Together, Exercises 4.42 and 4.43 prove the following classification theorem:
Theorem 4.44. Let Σ be a connected, regular surface in E3 whose principal
curvatures are constant. Then Σ is contained in either a plane, sphere, or
cylinder.
several steps. (For simplicity, we will assume that the surface is oriented,
meaning that a choice of e3 has been specified.)
dA = ω̄ 1 ∧ ω̄ 2
(Note: The notation dA for the area form is traditional, but the d does not
signify that dA is the exterior derivative of some 1-form.)
(b) Show that if ω̄21 is the Levi-Civita connection form corresponding to
(ω̄ 1 , ω̄ 2 ), then
˜ 21 = ω̄21 − dθ.
ω̄
Conclude that dω̄21 is an intrinsic quantity.
(c) Show that
dω̄21 = ω̄13 ∧ ω̄23 = K ω̄ 1 ∧ ω̄ 2 = K dA.
Conclude that K must be an intrinsic quantity. (Note that this is simply
another version of equation (4.11), which expresses the Gauss curvature
eg
K = EG as a function of E, G, and their derivatives.)
Surfaces for which the mean curvature H is identically zero are called min-
imal surfaces; these surfaces are of considerable interest and will be treated
in detail in Chapter 8. Surfaces for which the Gauss curvature K is iden-
tically zero are called flat, and we will conclude this chapter with a brief
exploration of their local theory.
Because the Gauss curvature of a regular surface Σ is an intrinsic quantity, it
is not changed by any deformation of Σ that preserves the first fundamental
form of Σ. Intuitively, this means that the surface may be smoothly bent
and/or twisted, but not stretched or contracted. So for instance, any surface
that can be obtained by smoothly bending a sheet of paper must be flat.
4.5. Moving frames for surfaces in E3 133
(a) Differentiate the equation ω̄13 = 0 and use Cartan’s lemma to conclude
that
(4.19) ω̄21 = μ ω̄ 2
(e) Use the structure equation for dω̄21 to conclude that λuu = 0, and there-
fore
We have now proved the following theorem, keeping in mind that with the
notation of Exercise 4.48, the mean curvature H of Σ is given by H = 12 κ2
with κ2 as in equation (4.22):
Theorem 4.49. Let Σ be a flat surface whose mean curvature H is nonzero
everywhere. Then for each point x ∈ Σ, there exists a unique straight line
x in E3 such that x ∈ x and x ∩ Σ is an open neighborhood of x in x .
Moreover, the restriction of the function H1 to the open interval x ∩ Σ is an
affine linear function of the arc length parameter along this interval.
In order to get set up to use Maple for some of the exercises in this chapter,
begin by loading the Cartan and LinearAlgebra packages into Maple:
> with(Cartan);
> with(LinearAlgebra);
Next, introduce the Maurer-Cartan forms on the frame bundle F (E3 ); these
need to be declared so that Maple will recognize them as 1-forms. It suffices
to declare a linearly independent subset; we’ll define the others in terms of
these shortly.
4.6. Maple computations 135
Note that we first computed dω̄ 3 and then applied the substitution to tell
Maple that ω̄ 3 = 0. In this case the knowledge that ω̄ 3 = 0 didn’t affect
the computation of dω̄ 3 , but it’s a good idea to get in the habit of applying
such substitutions when you intend for them to be in effect.
Applying Cartan’s lemma tells us that (ω̄13 , ω̄23 ) must be symmetric linear
combinations of (ω̄ 1 , ω̄ 2 ), so we add this information to our substitution:
> adaptedsub1:= [op(adaptedsub1),
omega[3,1] = h[1,1]*omega[1] + h[1,2]*omega[2],
omega[3,2] = h[1,2]*omega[1] + h[2,2]*omega[2]];
Exercise 4.28: In order to keep up with both the original Maurer-Cartan
forms (ω̄ i , ω̄ji ) and the transformed forms (ω̄
˜ i , ω̄
˜ ji ), introduce new 1-forms to
represent the transformed forms, with the same symmetry conditions as the
original forms:
> Form(Omega[1], Omega[2], Omega[3]);
Form(Omega[1,2], Omega[3,1], Omega[3,2]);
Omega[1,1]:= 0;
Omega[2,2]:= 0;
Omega[3,3]:= 0;
Omega[2,1]:= -Omega[1,2];
Omega[1,3]:= -Omega[3,1];
Omega[2,3]:= -Omega[3,2];
(It won’t be necessary to assign their exterior derivatives because these will
be computed in terms of the exterior derivatives of the original forms when
needed.)
We can introduce the relations (4.5), (4.6) via the following substitution:
> framechangesub:= [
Omega[1] = cos(theta)*omega[1] + sin(theta)*omega[2],
Omega[2] = -sin(theta)*omega[1] + cos(theta)*omega[2],
Omega[3,1] = cos(theta)*omega[3,1] + sin(theta)*omega[3,2],
Omega[3,2] = -sin(theta)*omega[3,1] + cos(theta)*omega[3,2]];
We’ll also need the reverse substitution so that we can go back and forth
between the two sets of Maurer-Cartan forms:
> framechangebacksub:= makebacksub(framechangesub);
In order to compare the functions (h̃ij ) associated to the transformed forms
to the functions (hij ) associated to the original forms, introduce another
4.6. Maple computations 137
˜ 1 , ω̄
Finally, convert this to an expression in terms of (ω̄ ˜ 2 ):
sinh(v)
a :=
cosh(v)
omega[3,1] = (e(u,v)/sqrt(E(u,v)))*d(u),
omega[3,2] = (g(u,v)/sqrt(G(u,v)))*d(v),
omega[1,2] = a*d(u) + b*d(v)];
Compute the coefficients in ω̄21 as we did in the previous exercise:
> Simf(d(Simf(subs(coordsub, omega[1])))
- subs(coordsub, Simf(d(omega[1]))));
> a:= solve(%, a);
> Simf(d(Simf(subs(coordsub, omega[2])))
- subs(coordsub, Simf(d(omega[2]))));
> b:= solve(%, b);
The Gauss equation comes from comparing the two expressions for dω̄21 :
> Simf(d(Simf(subs(coordsub, omega[1,2])))
- subs(coordsub, Simf(d(omega[1,2]))));
> Gausseq1:= pick(%, d(u), d(v));
Now, you’ll probably notice that this expression doesn’t look quite like the
one in part (e) of the exercise. But we can ask Maple to compare the two
expressions to confirm that they are, in fact, equivalent. First, give a name
to the expression that results from moving all the terms in equation (4.11)
to the left-hand side:
> Gausseq2:= (e(u,v)*g(u,v))/(E(u,v)*G(u,v))
+ (1/(2*sqrt(E(u,v)*G(u,v))))*
(diff(diff(E(u,v), v)/sqrt(E(u,v)*G(u,v)), v)
+ diff(diff(G(u,v), u)/sqrt(E(u,v)*G(u,v)), u));
Solve this equation for one of the variables (say, g), and then substitute this
expression into the first version of the Gauss equation. If the two equations
are equivalent, then the result should be zero.
> solve(Gausseq2, {g(u,v)});
> Simf(subs(%, Gausseq1));
Similar manipulations involving dω̄13 and dω̄23 will confirm that their struc-
ture equations are equivalent to the Codazzi equations (4.12).
Now, in fact, there’s nothing special about assuming that F = f = 0,
except that it makes the computations simpler. For a challenge, you might
try redoing this exercise without this assumption. You’ll need to start by
4.6. Maple computations 141
5.1. Introduction
143
144 5. Curves and surfaces in Minkowski space
where for each t ∈ I, (e0 (t), e1 (t), e2 (t)) is an oriented, orthonormal basis
for the tangent space Tα(t) M1,2 . (Recall that, by convention, we require that
e0 (t) be timelike and that e1 (t) and e2 (t) be spacelike; cf. Exercise 3.39.)
As in the Euclidean case, such an adapted frame field is usually called an
orthonormal frame field along α.
As in the Euclidean case, we say that α is regular if α (t) = 0 for every
t ∈ I; henceforth, we will only consider regular curves. The construction of
an orthonormal frame field along α is very similar to that for a curve in E3 ;
we begin by setting
α (t)
e0 (t) = ,
|α (t)|
where |α (t)| = α (t), α (t) is computed using the Minkowski inner prod-
uct; i.e., we take e0 (t) to be the unit tangent vector to the curve at α(t).
(Note that, since α is a timelike curve, e0 (t) is a timelike vector, as desired.)
The Minkowski analog of arc length is the following:
Definition 5.1. Given a fixed point t0 ∈ I, the proper time function along
α is
& t
τ (t) = |α (u)| du.
t0
This terminology arises from the key fact that in relativity, time is not
an absolute quantity; different observers may measure the passage of time
differently, and the proper time along a world line α(t) represents how time
passes for an observer traveling along α.
Remark 5.2. The use of the letter τ to denote proper time is traditional
in relativity; however, it should not be confused with the torsion of a non-
degenerate curve in E3 , which is also denoted by τ ! Hopefully it will be
clear from the context which quantity is intended. This convention will also
necessitate a departure from the traditional geometric notation for curves
in E3 ; thus, we will denote the analogs of curvature and torsion for timelike
curves in M1,2 by κ1 , κ2 .
5.2. Moving frames for timelike curves in M1,2 145
Exercise 5.3. Show that τ (t) is invariant under the action of M (1, 2); that
is, for any g ∈ M (1, 2), the curves α and g · α have the same proper time
function.
The following exercise shows that the Minkowski cross product has proper-
ties analogous to those of the Euclidean cross product.
5.2. Moving frames for timelike curves in M1,2 147
Exercise 5.6. (a) Show that with respect to the Minkowski inner product,
v × w is orthogonal to both v and w.
(b) Show that
v × w, v × w = v, vw, w − v, w2 .
Conclude that if v, w are orthogonal vectors of Minkowski norm 1, then
v × w also has Minkowski norm equal to 1.
(c) Show that for any three vectors u, v, w ∈ M1,2 ,
u, v × w = det u v w .
(This is an orientation condition that dictates how the sign of v × w should
be chosen.)
(Recall from Exercise 3.50 that ω̄00 = 0.) The only remaining Maurer-Cartan
form is ω̄21 ; it must be equal to some multiple of dτ , so define a function κ2 (τ )
by the condition that
ω̄21 = κ2 (τ )dτ.
The function κ2 (τ ) is the analog of the torsion τ (s) for curves in E3 . We
will call κ2 (τ ) the Minkowski torsion of α.
Using the symmetry relations between the (ω̄βα ) from Exercise 3.50, we have
the Minkowski analog of the Frenet equations for timelike curves:
⎡ ⎤
0 0 0 0
⎢ ⎥
⎢1 0 κ1 (τ ) 0 ⎥
α (τ ) e0 (τ ) e1 (τ ) e2 (τ ) = α(τ ) e0 (τ ) e1 (τ ) e2 (τ ) ⎢
⎥
⎢0 κ (τ ) 0 κ (τ )⎥ .
⎣ 1 2 ⎦
0 0 −κ2 (τ ) 0
(b) Compute the Frenet frame (e0 (τ ), e1 (τ ), e2 (τ )) for α and show that the
Minkowski curvature and torsion of α are
k2r k
κ1 (τ ) = , κ2 (τ ) = .
1−k r
2 2 1 − k2 r2
Thus, α is the Minkowski analog of a helix in E3 , i.e., a nondegenerate
curve with constant nonzero curvature and torsion. (This shouldn’t be too
surprising since the world line of α is a helix in M1,2 !)
The vectors (e0 (u), e1 (u)) must now form a basis for the tangent space
Tx(u) Σ. We will assume that e0 (u) is timelike and e1 (u) is spacelike, but
otherwise (e0 (u), e1 (u)) is allowed to be an arbitrary orthonormal basis of
Tx(u) Σ. An orthonormal frame field satisfying this condition will be called
adapted.
Let (ω̄ α , ω̄βα ) represent the pulled-back forms (x̃∗ ω α , x̃∗ ωβα ) on U . Precisely
the same reasoning as in the Euclidean case can be used to prove the fol-
lowing:
Proposition 5.12. Let U ⊂ R2 be an open set, and let x : U → M1,2
be a timelike immersion. For any adapted orthonormal frame field (e0 (u),
e1 (u), e2 (u)) along Σ = x(U ), the associated dual and connection forms
(ω̄ α , ω̄βα ) have the property that ω̄ 2 = 0. Moreover, (ω̄ 0 , ω̄ 1 ) form a basis for
the 1-forms on U .
The metric properties of the surface Σ are once again contained in the first
fundamental form of the surface:
Definition 5.13. Let U ⊂ R2 be an open set, and let x : U → M1,2 be a
timelike immersion. The first fundamental form of Σ = x(U ) is the quadratic
form I on T U defined by
I(v) = dx(v), dx(v)
for v ∈ Tu U , where ·, · is the Minkowski inner product.
*Exercise 5.14. Show that for any v ∈ Tu U ,
2 2
I(v) = ω̄ 0 (v) − ω̄ 1 (v) .
This is often written more concisely as
I = (ω̄ 0 )2 − (ω̄ 1 )2 .
Note that I is a quadratic form of signature (1, 1) rather than a positive
definite quadratic form. This reflects the fact that Σ is a timelike surface;
if Σ were spacelike, then its first fundamental form would have signature
(0, 2).
(The minus sign is included here in order to minimize minus signs in the
second fundamental form below.)
Once again, the functions (hαβ ) are related to the differential of the Gauss
map. The Gauss map and the second fundamental form are defined essen-
tially as before:
Definition 5.15. Let U ⊂ R2 be an open set, and let x : U → M1,2 be a
timelike immersion with image Σ = x(U ). The Gauss map of Σ = x(U ) is
the map N : Σ → M1,2 defined by
N (x(u)) = e2 (u),
where (e0 (u), e1 (u), e2 (u)) is any adapted frame field on Σ = x(U ). (Note
that N takes values in the “sphere” S−1 .)
Definition 5.16. Let U ⊂ R2 be an open set, and let x : U → M1,2 be
a timelike immersion. The second fundamental form of Σ = x(U ) is the
quadratic form II on T U defined by
II(v) = −de2 (v), dx(v)
for v ∈ Tu U , where (e0 (u), e1 (u), e2 (u)) is any adapted frame field on Σ =
x(U ).
*Exercise 5.17. Show that for any v ∈ Tu U ,
II(v) = − ω̄02 (v) ω̄ 0 (v) + ω̄12 (v) ω̄ 1 (v)
= h00 (ω̄ 0 (v))2 + 2h01 ω̄ 0 (v) ω̄ 1 (v) + h11 (ω̄ 1 (v))2 .
This is often written more concisely as
II = −(ω̄02 ω̄ 0 + ω̄12 ω̄ 1 ) = h00 (ω̄ 0 )2 + 2h01 ω̄ 0 ω̄ 1 + h11 (ω̄ 1 )2 .
(Hint: Although the result looks the same as in the Euclidean case, there are
some sign differences in the details of the computation. Note that, because
of the slightly different symmetries in the (ω̄βα ), we have
de2 = e0 ω20 + e1 ω21 = e0 ω02 − e1 ω12 .
And don’t forget to use the Minkowski inner product in the definition of II!)
Now we will examine how the matrix [hαβ ] changes if we vary the frame.
Let (e0 (u), e1 (u), e2 (u)) be any adapted frame field along Σ, with associated
Maurer-Cartan forms (ω̄ α , ω̄βα ). Any other adapted frame field (ẽ0 (u), ẽ1 (u),
ẽ2 (u)) has the form (up to sign)
(5.2) ⎡ ⎤
cosh(θ) sinh(θ) 0
⎢ ⎥
ẽ0 (u) ẽ1 (u) ẽ2 (u) = e0 (u) e1 (u) e2 (u) ⎣ sinh(θ) cosh(θ) 0⎦
0 0 1
152 5. Curves and surfaces in Minkowski space
˜ α , ω̄
for some function θ on U . Let (ω̄ ˜ βα ) be the Maurer-Cartan forms associ-
cosh(θ) sinh(θ)
ated to the new frame field, and set B = .
sinh(θ) cosh(θ)
Recall that at this point in the Euclidean case, we used the fact that any
symmetric matrix has an orthogonal basis of eigenvectors to conclude that
we could choose B so as to diagonalize the matrix [hij ]. The key fact is not
so much that the matrix [hij ] is symmetric (which depends on the fact that
it is expressed relative to an orthonormal basis (e1 , e2 )), but rather that
the linear operator de3 : Tx(u) Σ → Tx(u) Σ is self-adjoint; i.e., for any two
vectors v, w ∈ Tx(u) Σ, we have
de3 (v), w = v, de3 (w).
The formal statement of the linear algebra theorem is that a self-adjoint op-
erator on En has all real eigenvalues and an orthogonal basis of eigenvectors.
5.3. Moving frames for timelike surfaces in M1,2 153
(b) Show that the linear transformation de2 : Tx(u) Σ → Tx(u) Σ is self-
adjoint with respect to the Minkowski metric; that is, for any two vectors
v, w ∈ Tx(u) Σ, we have
de2 (v), w = v, de2 (w).
Conclude that the matrix S that expresses a self-adjoint operator on a
(1+1)-dimensional Minkowski vector space V relative to an orthonormal
basis (e0 , e1 ) for V has the property that the matrix S + tS is diagonal.
(Contrast this with the Euclidean case, where any self-adjoint operator is
expressed by a symmetric matrix S relative to an orthonormal basis.)
(c) Show that:
• If |h00 + h11 | > 2|h01 |, then de2 has distinct, real eigenvalues and
an orthogonal basis of eigenvectors, one timelike and one spacelike.
• If |h00 + h11 | < 2|h01 |, then de2 has complex eigenvalues and no
real eigenvectors.
• If |h00 + h11 | = 2|h01 | = 0, then de2 is a multiple of the identity
transformation.
• If |h00 + h11 | = 2|h01 | = 0, then de2 has a repeated real eigenvalue
and a 1-dimensional, lightlike eigenspace.
√
Definition 5.22. The function H = H 2 − K on Σ is called the skew
curvature of Σ. (By convention, if H 2 − K < 0, then H is chosen so that
iH < 0.)
*Exercise 5.23. (a) Show that for a regular surface Σ in Euclidean space
E3 , the skew curvature H is always real and H = 0 precisely at umbilic
points.
(b) Show that for a timelike, regular surface in Minkowski space M1,2 ,
9
H = 12 (h00 + h11 )2 − 4h201 .
(c) Conclude from the result of Exercise 5.19 that at any point x(u) ∈ Σ:
• If H (u) is real and H (u) > 0, then d(e2 )x(u) has distinct, real
eigenvalues and an orthogonal basis of eigenvectors, one timelike
and one spacelike.
In light of this result, we will need to divide into cases based on the skew
curvature in order to make further refinements to our adapted frame field.
5.3.1. Case 1: H (u) is real and d(e2 )x(u) has an orthogonal ba-
sis of eigenvectors for all u ∈ U . This assumption covers the first
and third cases in Exercise 5.23, and the frame adaptation proceeds much
as it did for surfaces in E3 . In this case, there exists an adapted frame
(e0 (u), e1 (u), e2 (u)) at each point x(u) ∈ Σ with the property that
−h00 (u) −h01 (u) −κ0 (u) 0
=
h01 (u) h11 (u) 0 −κ1 (u)
for some real numbers κ0 (u), κ1 (u). As in the Euclidean case, e0 (u) and
e1 (u) are eigenvectors for d(e2 )u , with
(5.4) d(e2 )u (e0 (u)) = −κ0 e0 (u), d(e2 )u (e1 (u)) = −κ1 e1 (u).
5.3. Moving frames for timelike surfaces in M1,2 155
The proof of this theorem (at least for surfaces with no umbilic points) is
very similar to that in the Euclidean case.
*Exercise 5.30. Prove Theorem 5.29 for surfaces without umbilic points
as follows: Let x1 , x2 : U → M1,2 be immersions as in Theorem 5.29, with
the same first and second fundamental forms and κ0 (u) = κ1 (u) for every
u ∈ U.
(a) Let x̃1 , x̃2 : U → M (1, 2) be principal adapted frame fields for x1 , x2 .
Let (ω̄ α , ω̄βα ) denote the pulled-back dual and connection forms for x1 and
156 5. Curves and surfaces in Minkowski space
(c) Prove that λ is constant. (Hint: Use the structure equations to dif-
ferentiate the equations above, taking into account the fact that we must
5.3. Moving frames for timelike surfaces in M1,2 157
have
dλ = λ0 ω̄ 0 + λ1 ω̄ 1
for some functions λ0 , λ1 on U . Then use Cartan’s lemma.)
(d) Show that if λ = 0, then de2 = 0. Conclude that the normal vector field
of Σ is constant and that Σ is therefore contained in a plane.
(e) Show that if λ = 0, then d(x+ λ1 e2 ) = 0. Conclude that the vector-valued
function x + λ1 e2 : U → M1,2 is equal to some constant point q ∈ M1,2 and
that Σ is therefore contained in the hyperboloid of one sheet defined by the
equation
1
x − q, x − q = − 2 .
|λ|
(Note that the minus sign on the right-hand side is due to the fact that e2 is
a spacelike vector and that this hyperboloid is actually the “sphere” S− 1 .)
λ2
Thus, the only totally umbilic timelike surfaces are (open subsets of) planes
and hyperboloids of one sheet.
Now consider the possible world lines for particles traveling in the de Sitter
space dS2 . Since dS2 is spatially finite, we might think that a moving particle
should be able to “circumnavigate” the space and return to its original
position. But in fact this is not possible, as the following exercise shows.
*Exercise 5.35. Consider the path β of a photon emitted at the point
x(0, 0) and traveling in the positive v direction. We can parametrize the
photon’s world line as
β(t) = x(t, v(t))
for some function v(t) which is determined by the conditions that v(0) = 0,
v (t) > 0, and β (t) is a lightlike vector.
(a) Show that the lightlike tangent directions to dS2 at the point x(u0 , v0 )
are spanned by the vectors v± = cosh(u0 ) xu ± xv .
(b) Show that the function v(t) above must satisfy the differential equation
v (t) = sech(t)
and that the solution to this equation with v(0) = 0 and v (t) > 0 is
π
v(t) = 2 tan−1 et − .
2
and therefore,
h00 h01 λ0 λ1
=
h01 h11 λ1 −λ0
for some functions λ0 , λ1 on U with |λ1 | > 0 (cf. Exercise 5.18).
(b) Show that the Maurer-Cartan forms associated to such a frame field
satisfy
ω̄02 = −λ0 ω̄ 0 − λ1 ω̄ 1 ,
ω̄12 = −λ1 ω̄ 0 + λ0 ω̄ 1 .
(c) Show that for such a frame field, the second fundamental form is
II = λ0 (ω̄ 0 )2 − (ω̄ 1 )2 + 2λ1 ω̄ 0 ω̄ 1 = λ0 I + 2λ1 ω̄ 0 ω̄ 1 .
(d) What goes wrong with part (a) when H = 0 and h01 = 0?
In order to get a feel for what such a surface Σ might look like, consider the
case where Σ is a graph of the form
x2 = f (x0 , x1 ).
(Any timelike surface can locally be described in this way, possibly after a
rotation in the (x1 , x2 )-plane.) Consider a parametrization x : U → M1,2 of
the form
x(u, v) = t[u, v, f (u, v)].
Exercise 5.37. Show that Σ is timelike if and only if |fu | < 1.
Now choose a point (u0 , v0 ) ∈ U , and “rotate” the surface (via a Lorentzian
transformation) if necessary so that
fu (u0 , v0 ) = fv (u0 , v0 ) = 0,
so that the tangent plane to Σ at the point x0 = x(u0 , v0 ) is parallel to the
(x0 , x1 )-plane.
Exercise 5.38. (a) Show that the normal vector field along Σ is
1
e2 (u, v) = [−fu , fv , −1].
t
1 + fv2 − fu2
In particular,
e2 (u0 , v0 ) = t[0, 0, −1].
and let (ω̄ α , ω̄βα ) be the associated Maurer-Cartan forms. Show that at the
point x0 ,
ω̄ 0 = du, ω̄ 1 = dv,
ω̄02 = −fuu du − fuv dv, ω̄12 = −fuv du − fvv dv.
Therefore,
h00 h01 fuu (u0 , v0 ) fuv (u0 , v0 )
= .
h01 h11 fuv (u0 , v0 ) fvv (u0 , v0 )
(u0 ,v0 )
5.3.3. Case 3: H (u) = 0 and h01 (u) = 0 for all u ∈ U . Note that this
is a degenerate condition, and generically it will only hold along a closed
subset of Σ, similar to the set of umbilic points for a surface in Case 1.
Definition 5.39. A point x(u) in a timelike surface Σ = x(U ) in M1,2 will
be called a quasi-umbilic point if H (u) = 0 and h01 (u) = 0. If every point
of Σ is quasi-umbilic, then Σ is called totally quasi-umbilic.
Exercise 5.40. Suppose that x : U → M1,2 is a totally quasi-umbilic time-
like surface.
(a) Show that
|h01 | = 12 |h00 + h11 | = 0.
5.4. An alternate construction for timelike surfaces 161
(d) Show that for such a frame field, the second fundamental form is
II = λ (ω̄ 0 )2 − (ω̄ 1 )2 + ε1 (ω̄ 0 )2 + (ω̄ 1 )2 + 2ε2 ω̄ 0 ω̄ 1
= λI + ε1 (ω̄ 0 )2 + (ω̄ 1 )2 + 2ε2 ω̄ 0 ω̄ 1 .
By analogy with the totally umbilic surfaces, one might expect that the
totally quasi-umbilic surfaces would form a fairly limited set. But in fact,
the analysis of totally quasi-umbilic surfaces is considerably more involved
than that for totally umbilic surfaces, and it turns out that there is an
infinite-dimensional family of such surfaces! We will explore some examples
in Exercise 5.49.
we will construct frame fields (f1 (u), f2 (u), f3 (u)) along Σ = x(U ) with the
property that f1 (u) and f2 (u) are null vectors.
Exercise 5.41. Suppose that (e0 (u), e1 (u), e2 (u)) is an orthonormal frame
field along Σ = x(U ). Show that the vectors
f1 (u) = √1 (e0 (u)
2
+ e1 (u)), f2 (u) = √1 (e0 (u)
2
− e1 (u))
Definition 5.42. A null adapted frame field (f1 (u), f2 (u), f3 (u)) along a
timelike surface Σ = x(U ) is a basis for the tangent space Tx(u) M1,2 with
the properties that
(1) (f1 (u), f2 (u)) are null vectors that span the tangent space Tx(u) Σ
at each point x(u) ∈ Σ;
(2) f1 (u), f2 (u) = 1;
(3) f3 (u) is the normal vector f3 (u) = f1 (u) × f2 (u) to Tx(u) Σ at each
point x(u) ∈ Σ.
Now, let (η̄ i , η̄ji ) be the Maurer-Cartan forms associated to a null adapted
frame field along Σ. These forms are defined just as in (3.1):
dx = fi η̄ i ,
dfi = fj η̄ij ,
where 1 ≤ i, j ≤ 3, and they still satisfy the structure equations
dη̄ i = −η̄ji ∧ η̄ j ,
dη̄ji = −η̄ki ∧ η̄jk .
The main difference is that the vectors (f1 , f2 , f3 ) satisfy the somewhat un-
usual inner product relations
f1 , f1 = 0, f2 , f2 = 0, f3 , f3 = −1,
(5.6)
f1 , f2 = 1, f1 , f3 = 0, f2 , f3 = 0.
*Exercise 5.45. Show that the first fundamental form of Σ = x(U ) (cf.
Definition 5.13) is
I = 2η̄ 1 η̄ 2 .
(b) The Gauss map (cf. Definition 5.15) of Σ = x(U ) is the map N : Σ →
M1,2 defined by
N (x(u)) = f3 (u).
Show that
df3 = f1 η̄31 + f2 η̄32 = −f1 (k12 η̄ 1 + k22 η̄ 2 ) − f2 (k11 η̄ 1 + k12 η̄ 2 ).
Conclude that with respect to the basis (f1 (u), f2 (u)) for Tx(u) Σ, the matrix
of the linear transformation df3 : Tx(u) Σ → Tx(u) Σ is
k12 k22
− .
k11 k12
(c) The second fundamental form (cf. Definition 5.16) of Σ = x(U ) is defined
by
II(v) = −df3 (v), dx(v)
for v ∈ Tu U . Show that
II = k11 (η̄ 1 )2 + 2k12 η̄ 1 η̄ 2 + k22 (η̄ 2 )2 .
(d) Show that the Gauss and mean curvatures (cf. Definition 5.20) of Σ are
2
K = k12 − k11 k22 , H = k12
and that the skew curvature (cf. Definition 5.22) is
H = k11 k22 .
164 5. Curves and surfaces in Minkowski space
Now we will examine how the matrix [kij ] changes if we vary the frame.
*Exercise 5.47. Let (f1 (u), f2 (u), f3 (u)) be any null adapted frame field
along Σ, with associated Maurer-Cartan forms (η̄ i , η̄ji ).
(a) Show that any other null adapted frame field (f̃1 (u), f̃2 (u), f̃3 (u)) has the
form (up to sign)
⎡ θ ⎤
e 0 0
⎢ ⎥
(5.9) f̃1 (u) f̃2 (u) f̃3 (u) = f1 (u) f2 (u) f3 (u) ⎣ 0 e−θ 0⎦
0 0 1
for some function θ on U . (Hint: Consider the transformation (5.2) for the
adapted orthonormal frame field
e0 (u) = √1 (f1 (u)
2
+ f2 (u)), e1 (u) = √1 (f1 (u)
2
− f2 (u)), e2 (u) = f3 (u).)
(b) Let (η̄˜i , η̄˜ji ) be the Maurer-Cartan forms associated to the new frame
eθ 0
field, and set B = . Show that
0 e−θ
1 1 −θ 1 1 1 −θ 1
η̄˜ η̄ e η̄ η̄˜3 η̄3 e η̄3
−1 −1
(5.10) =B = , =B = .
η̄˜2 η̄ 2 eθ η̄ 2 η̄˜32 η̄32 eθ η̄32
(c) Cartan’s lemma implies that there exist functions k̃11 , k̃12 , k̃22 on U such
that
3
η̄˜1 k̃11 k̃12 η̄˜1
=− .
η̄˜23 k̃12 k̃22 η̄˜2
Show that
k̃12 k̃22 k12 k22
= B −1 B
k̃11 k̃12 k11 k12
5.4. An alternate construction for timelike surfaces 165
and that
k̃11 k̃12 k11 k12 e2θ k11 k12
(5.11) = tB B= .
k̃12 k̃22 k12 k22 k12 e−2θ k22
(1) If H = 0, then there exists a null adapted frame field (f1 (u),
f2 (u), f3 (u)) (unique up to sign) along Σ with the property that
k11 = ±k22 . The sign is positive if H is real and negative if H is
imaginary.
(2) It may not be possible to choose such a frame field continuously
near umbilic or quasi-umbilic points. But in a totally quasi-umbilic
neighborhood where, without loss of generality, k22 = 0 and k11 = 0,
there exists a null adapted frame field (f1 (u), f2 (u), f3 (u)) (unique
up to sign) along Σ with the property that k11 = ±1.
I = 2F du dv
for some function F > 0 on U . We can write the second fundamental form
as
II = e du2 + 2f du dv + g dv 2
for some functions e, f, g on U . Follow the outline of Exercises 4.24 and
4.27 to derive the analogs of the Gauss and Codazzi equations for timelike
surfaces in null coordinates.
x(u, v) = α(u) + vf 0 .
α (u), f 0 = 1.
166 5. Curves and surfaces in Minkowski space
η̄ 1 = du, η̄ 2 = dv,
η̄31 = η̄23 = 0, η̄32 = η̄13 = λ du
K ≡ 0, H ≡ 0.
This exercise shows that, unlike in the Euclidean case, there exist non-planar
timelike surfaces in M1,2 whose Gauss and mean curvatures are identically
zero! Conversely, it can be shown that if Σ ⊂ M1,2 is a timelike surface with
K ≡ H ≡ 0, then every point of Σ is either umbilic or quasi-umbilic; see
[Cle12] for details.
The Maple setup for the orthonormal frame approach in this chapter is
much the same as in Chapter 4, except that indices now range from 0 to 2
and the symmetries of the connection forms are slightly different. Likewise,
Exercise 5.18 is very similar to Exercise 4.28; aside from the adjustments
in the index ranges, the only change is in the matrix B. We leave these
modifications as an exercise for the reader. (More details are given in the
Maple worksheet for this chapter on the AMS webpage.
Here, we will explore how to do analogous computations for a null adapted
frame field along a timelike surface in M1,2 . After loading the Cartan and
LinearAlgebra packages into Maple, begin by declaring the necessary 1-
forms:
> Form(eta[1], eta[2], eta[3]);
Form(eta[1,1], eta[3,1], eta[3,2]);
5.5. Maple computations 167
and in order for the second fundamental form to have the desired form, we
must have
" # " #
e f f g
η̄1 = − √ du + √ dv ,
3
η̄2 = − √ du + √ dv .
3
F F F F
+ (g(u,v)/sqrt(F(u,v)))*d(v)),
eta[1,1] = a*d(u) + b*d(v)];
Use the structure equations for dη̄ 1 and dη̄ 2 to compute the coefficients in
η11 :
> Simf(d(Simf(subs(coordsub, eta[1])))
- subs(coordsub, Simf(d(eta[1]))));
> b:= solve(%, b);
> Simf(d(Simf(subs(coordsub, eta[2])))
- subs(coordsub, Simf(d(eta[2]))));
> a:= solve(%, a);
The Gauss equation comes from comparing the two expressions for dη̄11 :
> Simf(d(Simf(subs(coordsub, eta[1,1])))
- subs(coordsub, Simf(d(eta[1,1]))));
> Gausseq:= pick(%, d(u), d(v));
Similar manipulations involving dη̄13 and dη̄23 show that the Codazzi equa-
tions take the form
" # " #
F fu − f Fu f F fv − f Fv f
ev = =F , gu = =F .
F F u F F v
Chapter 6
6.1. Introduction
Now we will apply the method of moving frames to study the geometry
of curves and surfaces in equi-affine space. The lack of a metric structure
will lead to results that may seem less intuitive than those of the last two
chapters, but the general procedure for constructing invariants remains the
same. And rather than rediscovering the familiar invariants for curves and
surfaces in Euclidean space, we will see how this method naturally leads to
an alternative set of invariants that are preserved under the action of the
entire equi-affine group.
T (x1 , x2 , x3 ) = t 2x1 , 12 x2 , x3 .)
171
172 6. Curves and surfaces in equi-affine space
where for each t ∈ I, (e1 (t), e2 (t), e3 (t)) is a unimodular basis for the tangent
space Tα(t) A3 . Such an adapted frame field is called a unimodular frame
field along α. The situation is quite different from that of either Euclidean
or Minkowski space; for instance, there is no obvious notion of arc length
for a curve that is invariant under the action of A(3). Moreover, we have
much greater freedom in choosing our frame; the only requirement is that
det[e1 (t) e2 (t) e3 (t)] = 1. So, how should we proceed?
In the Euclidean case, we used the first derivative of α to choose e1 (t)
and the second derivative of α to choose e2 (t), pausing along the way to
normalize according to arc length so that the frame would be orthonormal.
These choices determined e3 (t) uniquely, but it is clear from the structure
equations that e3 (t) is related to the third derivative of α. In order for this
procedure to work, we had to assume that α was “nondegenerate”, i.e., that
the vectors (α (t), α (t)) were linearly independent for each t ∈ I.
So, how might we use similar reasoning to construct a canonical adapted
frame field in the equi-affine case, and what would be the right notion of
“nondegenerate” for equi-affine curves? Since orthonormality is no longer
required, our first guess towards constructing an adapted frame field might
be to take
e1 (t) = α (t),
In order for this to work, we must assume that the vectors (α (t), α (t),
α (t)) are linearly independent for each t ∈ I; such a curve will be called
nondegenerate. (Note that this word means something different for curves
in equi-affine space than for curves in Euclidean space!) For nondegenerate
curves, the only problem with this choice of frame field is that it is not
6.2. Moving frames for curves in A3 173
Now, wouldn’t it be nice to get rid of that ugly denominator? In the Eu-
clidean case, we were able to get rid of the denominator for e1 (t) by repa-
rametrizing according to arc length. So let’s see if we can find a suitable
reparametrization to do the trick here; specifically, we would like to find a
reparametrization α(s) of α for which
det[α (s) α (s) α (s)] = 1.
This will be a bit more complicated than in the Euclidean case since the
denominators in (6.1) involve the first three derivatives of α instead of just
the first derivative.
Suppose that we reparametrize the curve by setting α(s) = α(t(s)) for some
invertible function t(s). Then
dα dα
= t (s) ,
ds dt
d2 α
2
2 d α dα
(6.2) ≡ t (s) mod ,
ds2 dt2 ds
d3 α d3 α dα d2 α
3
≡ t (s)3 3 mod , .
ds dt ds ds2
2
Note that it doesn’t matter what multiple of dα d α
ds we are ignoring in ds2 —
2 3
and similarly for the dα d α d α
ds , ds terms in ds2 —because they won’t affect the
determinant det[α (s) α (s) α (s)]. Therefore,
det[α (s) α (s) α (s)] = t (s)6 det[α (t(s)) α (t(s)) α (t(s))].
Note that the sign of det[α (s) α (s) α (s)] is fixed, so the best that we can
hope for is to arrange that det[α (s) α (s) α (s)] = ±1. This suggests that
174 6. Curves and surfaces in equi-affine space
Equi-affine arc length is a very different notion from Euclidean arc length.
Some of the most notable differences are:
(1) Unlike Euclidean arc length, which depends only on the first de-
rivative of α, the equi-affine arc length depends on the first three
derivatives of α. In fact, this number is dependent on the dimen-
sion of the ambient equi-affine space: The equi-affine arc length of
a curve α : I → An depends on the first n derivatives of α.
(2) The equi-affine arc length is only nonzero for nondegenerate curves;
so, for instance, any curve contained in a plane in A3 has equi-affine
arc length zero according to this definition. It could, however, have
nonzero equi-affine arc length when regarded as a curve in A2 .
Remark 6.8. This exercise implies that, while the curvature κ and torsion
τ of a unit-speed curve in E3 are not preserved by the action of the equi-
affine group, the 1-form 6 |κ2 τ | ds̄ (where s̄ is the Euclidean arc length of
the curve) is invariant under the action of this group!
Remark 6.10. The vectors (e1 (s), e2 (s), e3 (s)) of the equi-affine Frenet
frame (6.4) may not be exactly equal to those of the frame field (6.1); they
will differ by precisely the terms that we ignored in the chain rule compu-
tations (6.2).
(d) Use the pullbacks of equations (3.1) to find a complete set of invariants
for nondegenerate equi-affine curves α : I → A2 parametrized by equi-affine
arc length.
*Exercise 6.14. Hopefully you discovered a single invariant κ(s), called
the equi-affine curvature of α, in Exercise 6.13. Suppose that α : I → A2 is
nondegenerate and that its equi-affine curvature κ(s) is equal to a constant
κ. Show that:
(a) If κ = 0, then α is a parabola.
(b) If κ > 0, then α is a hyperbola.
(c) If κ < 0, then α is a circle or an ellipse.
(Hint: In each case, you should be able to show that α(s) satisfies a dif-
ferential equation whose general solution is not difficult to find. Since each
of these conditions is preserved under equi-affine transformations, you can
perform an equi-affine transformation to eliminate most of the arbitrary con-
stants in the general solution. Finally, eliminate the parameter s to show
that α lies on the appropriate conic section.)
*Exercise 6.15. Let α : I → A2 be a curve, parametrized by its Euclidean
arc length s̄, and let κ̄(s̄) denote the Euclidean curvature of α.
(a) Show that the equi-affine arc length of α is given by
&
3
(6.6) s(s̄) = κ̄(u)2 du.
s̄0
(b) Let (ē1 (s̄), ē2 (s̄)) denote the Euclidean Frenet frame field along α. Show
that the equi-affine Frenet frame field along α is given by
e1 (s) = κ̄(s̄(s))−1/3 ē1 (s̄(s)),
(6.7) κ̄ (s̄(s))
e2 (s) = − ē1 (s̄(s)) + κ̄(s̄(s))1/3 ē2 (s̄(s)),
3κ̄(s̄(s))5/3
where s̄(s) is the inverse function of the equi-affine arc length function (6.6)
and primes denote derivatives of κ̄(s̄) with respect to s̄. (Hint: Use the
chain rule very carefully!)
(c) Differentiate the expression (6.7) for e2 (s) (again, being very careful with
the chain rule) and conclude that the equi-affine curvature of α is given by
1
(6.8) κ(s) = − 8/3
3κ̄(s̄(s))κ̄ (s̄(s)) − 5(κ̄ (s̄(s)))2 + 9κ̄(s̄(s))4 .
9κ̄(s̄(s))
This shows that, while Euclidean curvature is a second-order invariant of α
(i.e., it may be expressed in terms of α and its derivatives up to order 2), the
178 6. Curves and surfaces in equi-affine space
Exercise 6.16. Now that you know how things work for curves in A2 and
A3 , what sorts of invariants would you expect to appear for curves in An ?
*Exercise 6.17. Let α : I → A3 be a nondegenerate curve parametrized
by equi-affine arc length s, and suppose that its equi-affine curvatures κ1 (s),
κ2 (s) are both identically equal to zero.
(a) Show that there exist vectors v0 , v1 , v2 , v3 ∈ A3 , with det[v1 v2 v3 ] = 1,
such that
α(s) = v0 + sv1 + 12 s2 v2 + 16 s3 v3 .
(b) Show that there exists an equi-affine transformation g ∈ A(3) such that
g · α(s) = t s, 12 s2 , 16 s3 .
This curve is called the rational normal curve of degree 3; we will encounter
it again in a slightly different form in Chapter 7 (cf. Exercise 7.31).
Let (ω̄ i , ω̄ji ) represent the pulled-back forms (x̃∗ ω i , x̃∗ ωji ) on U . Precisely the
same reasoning as in the Euclidean case can be used to prove the following:
Once again, with an eye towards making further adaptations to our frame
field, we will investigate how the matrix [hij ] changes if we vary the frame.
At this point things begin to look different from the Euclidean case because
we have a larger symmetry group to make use of.
Let (e1 (u), e2 (u), e3 (u)) be any 0-adapted frame field, with associated Mau-
rer-Cartan forms (ω̄ i , ω̄ji ). Any other 0-adapted frame field (ẽ1 (u), ẽ2 (u),
ẽ3 (u)) must have the properties that
span(ẽ1 (u), ẽ2 (u)) = span(e1 (u), e2 (u))
and
det ẽ1 (u) ẽ2 (u) ẽ3 (u) = det e1 (u) e2 (u) e3 (u) .
This will be true if and only if
⎡ ⎤
r1
⎢ B ⎥
(6.10) ẽ1 (u) ẽ2 (u) ẽ3 (u) = e1 (u) e2 (u) e3 (u) ⎣ r2 ⎦
0 0 (det B)−1
for some GL(2)-valued function B and real-valued functions r1 , r2 on U . Let
˜ i , ω̄
(ω̄ ˜ ji ) be the Maurer-Cartan forms associated to the new frame field.
(c) Cartan’s lemma implies that there exist functions h̃11 , h̃12 , h̃22 on U such
that 3 1
˜1
ω̄ h̃11 h̃12 ω̄ ˜
= .
˜ 23
ω̄ h̃12 h̃22 ω̄ ˜2
Show that
h̃11 h̃12 h11 h12
(6.13) = (det B) tB B.
h̃12 h̃22 h12 h22
For the remainder of this section, we will assume that Σ is elliptic; the
hyperbolic case will be treated in Exercise 6.42.
The transformation (6.13) acts transitively on the set of 2 × 2 matrices with
positive determinant; therefore, there exists a choice of 0-adapted frame field
(e1 (u), e2 (u), e3 (u)) for which
h11 h12 1 0
= .
h12 h22 0 1
Such a frame field will be called 1-adapted.
6.3. Moving frames for surfaces in A3 181
*Exercise 6.22. Let (e1 (u), e2 (u), e3 (u)) be any 1-adapted frame field for
an elliptic equi-affine surface Σ = x(U ) ⊂ A3 .
(a) Show that any other 1-adapted frame field (ẽ1 (u), ẽ2 (u), ẽ3 (u)) for Σ
must have the form
⎡ ⎤
r1
⎢ B ⎥
(6.14) ẽ1 (u) ẽ2 (u) ẽ3 (u) = e1 (u) e2 (u) e3 (u) ⎣ r2 ⎦
0 0 (det B) −1
The result of Exercise 6.22 implies that the equi-affine first fundamental
form is well-defined, independent of the choice of a particular 1-adapted
frame field on Σ.
Exercise 6.24. Show that the equi-affine first fundamental form is invariant
under the action of A(3).
come from the restriction of the same ambient metric on the entire space.
But there is no reason why they should be the same for curves in equi-affine
space, especially considering the fact that the equi-affine arc length of a
curve α : I → A3 depends on the first three derivatives of α, whereas the
equi-affine first fundamental form of a surface x : U → A3 depends only on
the first two derivatives of x. In fact, as the following exercise will show,
these two notions of arc length do not necessarily agree!
Exercise 6.26. Let U = (0, 2π) × (− π2 , π2 ) ⊂ R2 , and let x : U → A3 be the
standard parametrization of the unit sphere S2 :
x(u, v) = t[cos(u) cos(v), sin(u) cos(v), sin(v)] .
Let α : (− π2 , π2 ) → A3 be the curve
α(t) = x(t, t) = t cos2 (t), sin(t) cos(t), sin(t) .
(d) (Maple recommended) Plot both arc length functions for α, and com-
pute the arc length of α with respect to both arc length functions. What
happens if you change the pitch of α, i.e., set α(t) = x(ct, t) for various
choices of c?
This raises the question: Given an elliptic surface Σ ⊂ A3 , are there any
curves in Σ for which these two notions of arc length do agree? This question
is explored in [CEM+ 14].
6.3. Moving frames for surfaces in A3 183
Now, we still don’t have a well-defined normal vector field e3 (u) on Σ because
we still allow transformations between 1-adapted frame fields with
ẽ3 (u) = e3 (u) + r1 (u)e1 (u) + r2 (u)e2 (u)
for arbitrary functions r1 , r2 . In order to address this issue, consider the
connection form ω̄33 .
*Exercise 6.27. (a) Show that under a transformation of the form (6.14),
we have
˜ 33 = ω̄33 + r1 ω̄ 1 + r2 ω̄ 2 .
ω̄
Since (ω̄ 1 , ω̄ 2 ) form a basis for the 1-forms on U , there is a unique choice
of r1 , r2 for which ω̄33 = 0. A 1-adapted frame field satisfying the condition
ω̄33 = 0 will be called 2-adapted.
(b) Give a geometric interpretation of the condition ω̄33 = 0.
*Exercise 6.28. Let (e1 (u), e2 (u), e3 (u)) be any 2-adapted frame field for
an elliptic equi-affine surface Σ = x(U ) ⊂ A3 .
(a) Show that any other 2-adapted frame field (ẽ1 (u), ẽ2 (u), ẽ3 (u)) for Σ
must have the form
⎡ ⎤
0
⎢ B ⎥
(6.15) ẽ1 (u) ẽ2 (u) ẽ3 (u) = e1 (u) e2 (u) e3 (u) ⎣ 0⎦
0 0 1
for some SO(2)-valued function B on U .
(b) Conclude that ẽ3 (u) = e3 (u) and that the vector field e3 (u) on Σ is
therefore now well-defined.
(c) Show that this choice of e3 (u) is equivariant under the action of A(3).
Definition 6.29. Let U ⊂ R2 be an open set, and let x : U → A3 be
an immersion. Let (e1 (u), e2 (u), e3 (u)) be any 2-adapted frame field along
Σ = x(U ). The vector field e3 (u) along Σ is called the equi-affine normal
vector field of Σ.
Remark 6.30. The equi-affine normal direction at a point x0 ∈ Σ has the
following geometric interpretation: Consider the family of planes in A3 that
are parallel to the tangent plane Tx0 Σ. Because Σ is elliptic, planes in this
family sufficiently close to Tx0 Σ each intersect Σ in a convex plane curve.
Each of these plane curves bounds a region that has a center of mass. These
centers of mass trace out a smooth curve in A3 , and the limiting tangent line
to this curve as the curve approaches the original point x0 is parallel to the
equi-affine normal vector to Σ at x0 . And since the notions of parallelism
184 6. Curves and surfaces in equi-affine space
and center of mass are equivariant under the action of A(3), this construction
is equivariant as well.
where (ω̄ i , ω̄ji ) are the Maurer-Cartan forms associated to any 2-adapted
frame field (e1 (u), e2 (u), e3 (u)) on Σ = x(U ).
*Exercise 6.32. Let (e1 (u), e2 (u), e3 (u)) be any 2-adapted frame field
along Σ with associated Maurer-Cartan forms (ω̄ i , ω̄ji ), and let (ẽ1 (u), ẽ2 (u),
ẽ3 (u)) be any other 2-adapted frame field of the form (6.15), with associated
Maurer-Cartan forms (ω̄ ˜ i , ω̄
˜ ji ).
Remark 6.33. The quantity L = 12 (11 + 22 ) is called the equi-affine mean
curvature of Σ. It has the property (similar to that of the mean curvature
H in the Euclidean case) that it vanishes identically if and only if Σ is a
critical point of the equi-affine area functional.We will explore this further
in Chapter 8.
(b) Use the fact that ω̄11 + ω̄22 = 0 (Why is this true for a 2-adapted frame
field? Hint: It is not true for a 1-adapted frame field.) to show that
(6.20) h122 = −h111 , h112 = −h222 .
Remark 6.35. If we were to define functions (hijk ) with i, j, k = 1, 2 by
the equations
2ω̄11 = h111 ω̄ 1 + h112 ω̄ 2 ,
ω̄21 + ω̄12 = h121 ω̄ 1 + h122 ω̄ 2 = h211 ω̄ 1 + h212 ω̄ 2 ,
2ω̄22 = h221 ω̄ 1 + h222 ω̄ 2
(where the second line makes sense because the expression on the left is
symmetric in the first two indices), then the result of Exercise 6.34 says
that the (hijk ) are symmetric in all their indices.
Definition 6.36. Let U ⊂ R2 be an open set, and let x : U → A3 be
an immersion. Let (e1 (u), e2 (u), e3 (u)) be any 2-adapted frame field along
Σ = x(U ), with associated Maurer-Cartan forms (ω̄ i , ω̄ji ). The cubic form
ẽ3 (u)) be any other 2-adapted frame field of the form (6.15), with associated
Maurer-Cartan forms (ω̄ ˜ i , ω̄
˜ ji ). We can write
cos(θ) − sin(θ)
B=
sin(θ) cos(θ)
for some function θ on U .
(a) Show that
˜ 11 ω̄
ω̄ ˜ 21 ω̄11 ω̄21 0 −dθ
−1
(6.21) =B B+ .
˜ 12 ω̄
ω̄ ˜ 22 ω̄12 ω̄22 dθ 0
More explicitly:
˜ 11 = cos2 (θ) ω̄11 + sin(θ) cos(θ) (ω̄21 + ω̄12 ) + sin2 (θ) ω̄22 ,
ω̄
˜ 21 = cos2 (θ) ω̄21 + sin(θ) cos(θ) (ω̄22 − ω̄11 ) − sin2 (θ) ω̄12 − dθ,
ω̄
(6.22)
˜ 12 = cos2 (θ) ω̄12 + sin(θ) cos(θ) (ω̄22 − ω̄11 ) − sin2 (θ) ω̄21 + dθ,
ω̄
˜ 22 = cos2 (θ) ω̄22 − sin(θ) cos(θ) (ω̄21 + ω̄12 ) + (sin2 θ)ω̄11 .
ω̄
(c) Conclude that any two elliptic surfaces Σ1 , Σ2 ⊂ A3 with the same first
and second equi-affine fundamental forms and Fubini-Pick form differ by an
equi-affine transformation. In other words, the first and second equi-affine
fundamental forms, together with the Fubini-Pick form, form a complete set
of local invariants for elliptic equi-affine surfaces with no umbilic points.
I = M du2 + N dv 2 ,
(6.23) II = m du2 + n dv 2 ,
P = p111 du3 + 3p112 du2 dv + 3p122 du dv 2 + p222 dv 3
(b) Show that the Maurer-Cartan forms associated to this equi-affine prin-
cipal adapted frame field are given by
√ √
ω̄ 1 = ω̄13 = M du, ω̄ 2 = ω̄23 = N dv,
m n
ω̄31 = √ du, ω̄32 = √ dv,
M N
1 p111 p112 2 p122 p222
ω̄1 = du + dv, ω̄2 = du + dv,
2M
" 2M # " 2N 2N
#
p p
ω̄21 = √112 + s1 du + √122 + s2 dv,
2 MN 2 MN
" # " #
p112 p122
2
ω̄1 = √ − s1 du + √ − s2 dv
2 MN 2 MN
(c) Use the Cartan structure equations for dω̄ 1 and dω̄ 2 to find the functions
s1 , s2 .
(d) Suppose that the functions M, N, m, n, pijk are prescribed arbitrarily,
subject to the conditions that M, N > 0 and that equation (6.24) is satisfied.
Use the Cartan structure equations (3.8) to determine the PDE system that
must be satisfied by these functions in order to guarantee the local existence
of an elliptic equi-affine surface with equi-affine first and second fundamental
forms and Fubini-Pick form given by (6.23). This PDE system is the analog
of the Gauss-Codazzi system for elliptic surfaces in A3 .
dx = e1 ω̄ 1 + e2 ω̄ 2 = ẽ1 ω̄
˜ 1 + ẽ2 ω̄
˜ 2,
ω̄31 = λ ω̄ 1 , ω̄32 = λ ω̄ 2
Exercise 6.42. In this exercise, we will explore the frame adaptation pro-
cess for hyperbolic equi-affine surfaces. Suppose that x : U → A3 is a
parametrization for a hyperbolic equi-affine surface Σ = x(U ) ⊂ A3 . Then
the matrix [hij ] in (6.9) has det[hij ] < 0.
(a) Show that there exists a 0-adapted frame field (f1 (u), f2 (u), f3 (u)) for
which
h11 h12 0 1
= .
h12 h22 1 0
190 6. Curves and surfaces in equi-affine space
Such a frame field will be called a 1-adapted null frame field on Σ = x(U ).
The associated Maurer-Cartan forms (η̄ i , η̄ji ) satisfy
η̄13 = η̄ 2 , η̄23 = η̄ 1 ,
and the equi-affine first fundamental form becomes
I = 2 η̄ 1 η̄ 2 .
Thus, it defines an indefinite metric on Σ. (Based on this observation, we
might expect that the geometry of hyperbolic equi-affine surfaces will be
somewhat reminiscent of the geometry of timelike surfaces in Minkowski
space!)
(b) Let (f1 (u), f2 (u), f3 (u)) be any 1-adapted null frame field for a hyperbolic
equi-affine surface Σ = x(U ) ⊂ A3 . Show that any other 1-adapted null
frame field (f̃1 (u), f̃2 (u), f̃3 (u)) for Σ must have the form
⎡ θ ⎤
e 0 r1
⎢ ⎥
(6.25) f̃1 (u) f̃2 (u) f̃3 (u) = f1 (u) f2 (u) f3 (u) ⎣ 0 e−θ r2 ⎦
0 0 1
for some functions θ, r1 , r2 on U .
(c) Show that under a transformation of the form (6.25), we have
η̄˜33 = η̄33 + r2 η̄ 1 + r1 η̄ 2 .
(No, that’s not a typo in the indices!) Conclude that there is a unique
choice of r1 , r2 for which η̄33 = 0. A 1-adapted null frame field satisfying the
condition η̄33 = 0 will be called a 2-adapted null frame field.
For the remainder of this exercise, assume that (f1 (u), f2 (u), f3 (u)) is a 2-
adapted null frame field along Σ.
(d) Differentiate the equation η̄33 = 0 and use Cartan’s lemma to conclude
that there exist functions , 12 , 21 on U such that
1
η̄3 12 η̄ 1
= .
η̄32 21 η̄ 2
The equi-affine second fundamental form of Σ is given by
II = η̄31 η̄13 + η̄32 η̄23
= η̄31 η̄ 2 + η̄32 η̄ 2
= 21 (η̄ 1 )2 + 2η̄ 1 η̄ 2 + 12 (η̄ 2 )2 .
6.4. Maple computations 191
Once again, the Maple setup is similar to that of Chapter 4, but since
there are fewer relations among the connection forms, we need to declare
more of them. (In fact, we’ll go ahead and declare them all; it turns out
to be convenient to wait until after we define the structure equations to tell
Maple about the single relation among the connection forms because it
allows us to use a for loop to define the structure equations without creating
redundant assignments.) So, after loading the Cartan and LinearAlgebra
packages into Maple, begin by declaring the necessary 1-forms:
> Form(omega[1], omega[2], omega[3]);
Form(omega[1,1], omega[1,2], omega[1,3],
omega[2,1], omega[2,2], omega[2,3],
omega[3,1], omega[3,2], omega[3,3]);
192 6. Curves and surfaces in equi-affine space
Tell Maple how to differentiate these forms according to the Cartan struc-
ture equations (3.8):
> for i from 1 to 3 do
d(omega[i]):= -add(’omega[i,j] &ˆ omega[j]’, j=1..3);
end do;
for i from 1 to 3 do
for j from 1 to 3 do
d(omega[i,j]):= -add(’omega[i,k] &ˆ omega[k,j]’,
k=1..3);
end do;
end do;
Now tell Maple about the relation between the connection forms:
> omega[3,3]:= -(omega[1,1] + omega[2,2]);
Set up a substitution for the Maurer-Cartan forms associated to a 0-adapted
frame field, taking into account the relations (6.9) that result from comput-
ing dω̄ 3 = 0:
> adaptedsub1:= [omega[3]=0,
omega[3,1] = h[1,1]*omega[1] + h[1,2]*omega[2],
omega[3,2] = h[1,2]*omega[1] + h[2,2]*omega[2]];
Exercise 6.20: Introduce new 1-forms to represent the transformed forms,
with the same relation as the original forms:
> Form(Omega[1], Omega[2], Omega[3]);
Form(Omega[1,1], Omega[1,2], Omega[1,3],
Omega[2,1], Omega[2,2], Omega[2,3],
Omega[3,1], Omega[3,2], Omega[3,3]);
Omega[3,3]:= -(Omega[1,1] + Omega[2,2]);
Under a transformation of the form (6.10), we have the relations (6.11),
(6.12). Since B is now an arbitrary matrix in GL(2), the corresponding
substitution requires a bit more typing than in previous chapters:
> framechangesub:= [
Omega[1] = (1/(b[1,1]*b[2,2] - b[1,2]*b[2,1]))*
(b[2,2]*omega[1] - b[1,2]*omega[2]),
Omega[2] = (1/(b[1,1]*b[2,2] - b[1,2]*b[2,1]))*
(-b[2,1]*omega[1] + b[1,1]*omega[2]),
Omega[3,1] = (b[1,1]*b[2,2] - b[1,2]*b[2,1])*
(b[1,1]*omega[3,1] + b[2,1]*omega[3,2]),
Omega[3,2] = (b[1,1]*b[2,2] - b[1,2]*b[2,1])*
6.4. Maple computations 193
(b[1,2]*omega[3,1] + b[2,2]*omega[3,2])];
> framechangebacksub:= makebacksub(framechangesub);
In order to compare the functions (h̃ij ) to the functions (hij ), introduce
another substitution describing the adaptations for the transformed frame:
> adaptedsub2:= [Omega[3]=0,
Omega[3,1] = H[1,1]*Omega[1] + H[1,2]*Omega[2],
Omega[3,2] = H[1,2]*Omega[1] + H[2,2]*Omega[2]];
Now combine all these substitutions to see how the (h̃ij ) are expressed in
terms of the (hij ):
> zero2:= Simf(subs(adaptedsub2, Omega[3,1])
- Simf(subs(framechangebacksub, Simf(subs(adaptedsub1,
Simf(subs(framechangesub, Omega[3,1])))))));
> eqns:= {op(ScalarForm(zero2))};
> zero3:= Simf(subs(adaptedsub2, Omega[3,2])
- Simf(subs(framechangebacksub, Simf(subs(adaptedsub1,
Simf(subs(framechangesub, Omega[3,2])))))));
> eqns:= eqns union {op(ScalarForm(zero3))};
> solve(eqns, {H[1,1], H[1,2], H[2,2]});
> assign(%);
Check that these expressions for the (h̃ij ) agree with those in equation (6.13):
> hmatrix:= Matrix([[h[1,1], h[1,2]], [h[1,2], h[2,2]]]);
Hmatrix:= Matrix([[H[1,1], H[1,2]], [H[1,2], H[2,2]]]);
B:= Matrix([[b[1,1], b[1,2]], [b[2,1], b[2,2]]]);
> Hmatrix - simplify(Determinant(B)*Transpose(B).hmatrix.B);
b[1,1]:= cos(theta);
b[1,2]:= -sin(theta);
b[2,1]:= sin(theta);
b[2,2]:= cos(theta);
Note that our adapted frame substitution is now as it should be:
> Simf(adaptedsub1);
[ω3 = 0, ω3,1 = ω1 , ω3,2 = ω2 ]
At this point, we know that ω̄33 = −(ω̄11 + ω̄22 ) = 0, and from equation (6.16)
we have expressions for ω̄31 and ω̄32 . Add these relations to our substitution
(we will use ell for the letter in order to avoid confusing it with the
number 1):
> adaptedsub1:= [op(adaptedsub1), omega[2,2] = -omega[1,1],
omega[1,3] = ell[1,1]*omega[1] + ell[1,2]*omega[2],
omega[2,3] = ell[1,2]*omega[1] + ell[2,2]*omega[2]];
Add similar relations to the substitution for the transformed forms (we can’t
use L because it is a command in the Cartan package, so use LL instead):
> adaptedsub2:= [op(adaptedsub2), Omega[2,2] = -Omega[1,1],
Omega[1,3] = LL[1,1]*Omega[1] + LL[1,2]*Omega[2],
Omega[2,3] = LL[1,2]*Omega[1] + LL[2,2]*Omega[2]];
Next, add the relations (6.17) to our frame change substitution (keeping in
mind that we now know that B ∈ SO(2)):
> framechangesub:= Simf([op(framechangesub),
Omega[1,3] = cos(theta)*omega[1,3] + sin(theta)*omega[2,3],
Omega[2,3] = -sin(theta)*omega[1,3]
+ cos(theta)*omega[2,3]]);
> framechangebacksub:= makebacksub(framechangesub);
Now combine all these substitutions to see how the (˜ij ) are expressed in
terms of the (ij ):
> zero4:= Simf(subs(adaptedsub2, Omega[1,3])
- Simf(subs(framechangebacksub, Simf(subs(adaptedsub1,
6.4. Maple computations 195
Simf(subs(framechangesub, Omega[1,3])))))));
> eqns:= {op(ScalarForm(zero4))};
> zero5:= Simf(subs(adaptedsub2, Omega[2,3])
- Simf(subs(framechangebacksub, Simf(subs(adaptedsub1,
Simf(subs(framechangesub, Omega[2,3])))))));
> eqns:= eqns union {op(ScalarForm(zero5))};
> solve(eqns, {LL[1,1], LL[1,2], LL[2,2]});
> assign(%);
Check that these expressions for the (˜ij ) agree with those in equation (6.18):
> ellmatrix:= Matrix([[ell[1,1], ell[1,2]],
[ell[1,2], ell[2,2]]]);
LLmatrix:= Matrix([[LL[1,1], LL[1,2]], [LL[1,2], LL[2,2]]]);
> LLmatrix - simplify(Transpose(B).ellmatrix.B);
d(ω̄13 − ω̄ 1 ), d(ω̄23 − ω̄ 2 ),
(See the Maple worksheet for this chapter on the AMS webpage for an
illustration of why the inner Simf command is a good idea—or experiment
for yourself and see what happens without it!)
We can recognize this as an expression of the form
φ1 ∧ ω̄ 1 + φ2 ∧ ω̄ 2 ,
to which we should apply Cartan’s lemma. If you want Maple to help you
identify φ1 and φ2 , you can use the pick command:
> pick(zero6, omega[1]);
2 ω1,1
> pick(zero6, omega[2]);
ω1,2 + ω2,1
196 6. Curves and surfaces in equi-affine space
Exercise 6.40: First declare the functions associated to the Euclidean first
and second fundamental forms of Σ. Since we want to express everything
in terms of κ1 , κ2 , declare these functions and then express e, g in terms of
them:
> PDETools[declare](E(u,v), G(u,v), kappa1(u,v), kappa2(u,v));
> e(u,v):= E(u,v)*kappa1(u,v);
g(u,v):= G(u,v)*kappa2(u,v);
198 6. Curves and surfaces in equi-affine space
For Maple purposes, use (η̄ i , η̄ji ) to denote the Maurer-Cartan forms as-
sociated to the Euclidean adapted frame field (e1 (u), e2 (u), e3 (u)) and use
(ω̄ i , ω̄ji ) to denote those associated to the 2-adapted equi-affine frame field
(ẽ1 (u), ẽ2 (u), ẽ3 (u)). We have computed (η̄ i , η̄ji ) previously, and we can
simply assign them:
> eta[1]:= E(u,v)ˆ(1/2)*d(u);
eta[2]:= G(u,v)ˆ(1/2)*d(v);
eta[1,2]:= (diff(E(u,v), v)*d(u) - diff(G(u,v), u)*d(v))/
(2*E(u,v)ˆ(1/2)*G(u,v)ˆ(1/2));
eta[3,1]:= (e(u,v)/E(u,v)ˆ(1/2))*d(u);
eta[3,2]:= (g(u,v)/G(u,v)ˆ(1/2))*d(v);
eta[2,1]:= -eta[1,2];
eta[1,3]:= -eta[3,1];
eta[2,3]:= -eta[3,2];
Next, we need to introduce variables for the vector fields (ei (u)) and (ẽi (u))
so that we can use their structure equations to compute the associated
Maurer-Cartan forms. There’s no good way to tell Maple that these func-
tions are vector-valued, but it won’t really matter.
First we tell Maple the structure equations for the exterior derivatives of
the Euclidean frame field, the components of which we call (e01, e02,
e03):
> d(e01):= e02*eta[2,1] + e03*eta[3,1];
d(e02):= e01*eta[1,2] + e03*eta[3,2];
d(e03):= e01*eta[1,3] + e02*eta[2,3];
We’ll need the Gauss and Codazzi equations later, and these can be com-
puted directly from the equations d(dei ) = 0. Start with d(de3 ):
zero9:= Simf(d(d(e03)));
Pick off the scalar coefficient of du ∧ dv, and collect terms in (e1 (u), e2 (u)):
> zero9a:= collect(pick(zero9, d(u), d(v)), {e01, e02});
Since (e1 (u), e2 (u)) are linearly independent, both coefficients must be zero;
indeed, these are the Codazzi equations. We can collect these into a conve-
nient substitution:
> Codazzisub:= [
op(solve({coeff(zero9a, e01), coeff(zero9a, e02)},
{diff(kappa1(u,v), v), diff(kappa2(u,v), u)}))];
6.4. Maple computations 199
Rather than just computing the Maurer-Cartan forms for the frame field
given in Exercise 6.40(a), let’s explore where these expressions came from.
First, the fact that the coordinate curves of Σ are principal curves means
that the Maurer-Cartan forms associated to this Euclidean frame field have
the property that η̄13 is a multiple of η̄ 1 and η̄23 is a multiple of η̄ 2 . This means
that this Euclidean frame field is actually a 0-adapted equi-affine frame field
for which the matrix [hij ] is diagonal. Based on the transformation rule
(6.13), this suggests that we should be able to create a 2-adapted equi-affine
frame field for which (ẽ1 (u), ẽ2 (u)) are scalar multiples of (e1 (u), e2 (u)),
respectively. So, set
for some (unknown) functions λ1 , λ2 . Then in order to keep the frame uni-
modular, we must have
1
ẽ3 (u) = r1 e1 (u) + r2 e2 (u) + e3
λ1 λ2
200 6. Curves and surfaces in equi-affine space
for some functions r1 , r2 . For Maple purposes, we will call this frame field
(e1, e2, e3):
> e1:= lambda1(u,v)*e01;
e2:= lambda2(u,v)*e02;
e3:= r1(u,v)*e01 + r2(u,v)*e02
+ e03/(lambda1(u,v)*lambda2(u,v));
In order for this frame field to be 2-adapted, the corresponding Maurer-
Cartan forms must satisfy the conditions
solve for λ1 , λ2 :
> solve({pick(coeff(zero11, e03), d(u)),
pick(coeff(zero12, e03), d(v))},
{lambda1(u,v), lambda2(u,v)});
At this point, Maple returns the rather obtuse expression
(
1
λ1 (u, v) = ,
RootOf (−κ1 + Z 8 κ23 )3 κ2
3
λ2 (u, v) = RootOf (−κ1 + Z 8 κ23 )
and we will need to intervene by hand. The expression
RootOf (−κ1 + Z 8 κ23 )
means any one of the values obtained by setting the expression in parentheses
equal to zero and solving for Z. Since we know that we want everything to
be positive-valued, we can easily see by inspection that the desired solution
is
(1/8)
κ
Z = 1(3/8) ,
κ2
which leads to the solution
(1/8) (1/8)
κ2 κ1
λ1 = (3/8)
, λ2 = (3/8)
.
κ1 κ2
So make these assignments:
> lambda1(u,v):= kappa2(u,v)ˆ(1/8)/kappa1(u,v)ˆ(3/8);
lambda2(u,v):= kappa1(u,v)ˆ(1/8)/kappa2(u,v)ˆ(3/8);
Now we can use the e3 (u) term in dẽ3 to solve for r1 , r2 :
> solve({op(ScalarForm(Simf(coeff(zero13, e03))))},
{r1(u,v), r2(u,v)});
> assign(%);
You can now inspect the following expressions to see that they agree with
those in Exercise 6.40(a):
> Simf(e1);
Simf(e2);
Simf(e3);
Moreover, the remaining Maurer-Cartan forms can now be computed from
the e1 (u) and e2 (u) terms in zero11, zero12, zero13. This is left as an
exercise for the reader; details may be found in the Maple worksheet for
this chapter on the AMS webpage.
Chapter 7
7.1. Introduction
203
204 7. Curves and surfaces in projective space
and
det e0 (t) e1 (t) e2 (t) = 1.
Such an adapted frame field is called a projective frame field along [α]. How
should we choose such a lifting in some geometrically natural way?
As in the equi-affine case, it seems natural to choose the vectors (e0 (t), e1 (t),
e2 (t)) so that
This suggests that we look for a lifting α : I → R3 of [α] with the property
that
(7.1) det α(t) α (t) α (t) = 1.
In particular, the vectors (α(t), α (t), α (t)) should be linearly independent
for all t ∈ I.
for all t ∈ I. Show that the vectors (α1 (t), α1 (t), α1 (t)) are linearly indepen-
dent if and only if the vectors (α2 (t), α2 (t), α2 (t)) are linearly independent.
(Hint: α2 (t) = λ(t)α1 (t) for some nonvanishing, real-valued function λ(t).)
Exercise 7.1 implies that the linear dependence or independence of the vec-
tors (α(t), α (t), α (t)) does not depend on the choice of lifting α : I → R3 .
7.2. Moving frames for curves in P2 205
Exercise 7.5. Prove that the canonical projective frame field (e0 (t), e1 (t),
e2 (t)) associated to [α] is equivariant under the action of SL(3) on P2 : If
we replace [α] by g · [α] for some g ∈ SL(3), then eγ (t) ∈ Tα(t) R3 will be
replaced by g · eγ (t) ∈ Tg·α(t) R3 for γ = 0, 1, 2.
Now things start to get interesting. The canonical projective frame field de-
pends on the parametrization of [α]. How will it change if we reparametrize
[α]? Is there some particular parametrization of [α]—something akin to
the arc length parametrization for curves in Euclidean space—that is some-
how geometrically natural? In order to answer this question, we will first
compute invariants for parametrized nondegenerate curves. Then we will in-
vestigate how these invariants transform under a change of parametrization
for [α] and look for special parametrizations that normalize the invariants
in some natural way.
The pullbacks of equations (3.1) to I via α can be written as
for some functions κ0 (t), κ1 (t). These functions are called the Wilczynski in-
variants of the parametrized curve [α] : I → P2 . They were first introduced
by Wilczynski in the context of studying invariants for linear differential
operators; see [Wil62] for details.
*Exercise 7.6. Why is there no e2 (t) term in the equation for e2 (t)? (Cf.
Exercise 6.11.)
1
β(s) = α(t(s)),
t (s)
1
ẽ0 (s) = e0 (t(s)),
t (s)
t (s)
ẽ1 (s) = − e0 (t(s)) + e1 (t(s)),
(t (s))2
" #
t (s) (t (s))2 t (s)
ẽ2 (s) = − − 2 e 0 (t(s)) − e1 (t(s)) + t (s)e2 (t(s)),
(t (s))2 (t (s))3 t (s)
where (e0 (t), e1 (t), e2 (t)) is the canonical projective frame field associated
to [α].
7.2. Moving frames for curves in P2 207
(c) Show that the Wilczynski invariants κ̃0 (s), κ̃1 (s) associated to [β] are
given by
" #3
t(4) (s) t (s)t (s) t (s)
(7.3) κ̃0 (s) = − +4 −3
t (s) (t (s))2 t (s)
+ (t (s))3 κ0 (t(s)) + t (s)t (s)κ1 (t(s)),
" #2
t (s) t (s)
(7.4) κ̃1 (s) = −2 +3 + (t (s))2 κ1 (t(s)),
t (s) t (s)
The transformation rule (7.4) for κ̃1 is simpler than the transformation rule
(7.3) for κ̃0 , in that (7.4) involves only κ1 , whereas (7.3) involves both κ0
and κ1 . So, first consider equation (7.4). Local existence theory for ordinary
differential equations guarantees that, given any smooth function κ1 (t) on
an interval I ⊂ R, any point t0 ∈ I, and any point s0 ∈ R, there exists an
interval J ⊂ R containing the point s0 and a function t : J → I with the
properties that
(1) t(s0 ) = t0 ;
S(g ◦ f ) = S(f ).
z − 12 z 2 = 0
f
for the function z = f .)
Remark 7.10. Lest this seem too bizarre, note that a curve in Euclidean
space does not actually have a unique arc length parameter s. Given an arc
length parameter s along a curve α : I → En , any reparametrization of the
form
(7.8) s̃ = s + c,
as + b
(7.9) t(s) =
cs + d
with ad − bc = 1. Let κ0 (t), κ̃0 (s) be the invariants associated to [α], [β],
respectively. Use equation (7.3) to show that
" #3
κ0 (t(s)) dt
κ̃0 (s) = = κ0 (t(s)) .
(cs + d)6 ds
The cubic form (7.10) is called the projective curvature form of [α]. In the
following exercise, we show that the projective curvature form is a well-
defined rank 3 symmetric tensor along [α].
point amounts to choosing two arbitrary functions along the curve. Here we
have only two constant parameters’ worth of choices for a projective frame
field along the curve, and choosing a specific projective frame at any point
of the curve uniquely determines the canonical projective frame field along
the entire curve.
For any canonical projective frame field (e0 (t), e1 (t), e2 (t)) associated to a
nondegenerate, projectively parametrized curve [α] : I → P2 , we have the
following analog of the Frenet equations:
⎡ ⎤
0 0 κ0 (t)
⎢ ⎥
e0 (t) e1 (t) e2 (t) = e0 (t) e1 (t) e2 (t) ⎣1 0 0 ⎦ .
01 0
Exercise 7.17. This exercise demonstrates the result of Exercise 7.16 ex-
plicitly in the case where [α] is a circle. Let [α] : R → P2 be the circle
parametrized in homogeneous coordinates as
(b) Show that the invariants κ0 (t), κ1 (t) associated to [α] are
2s 0 1 0
Therefore, β is a parabola in the plane spanned by the vectors t[1, −1, 0],
t[0, 0, 1] and passing through the point t[ 1 , 1 , 0].
2 2
(b) Show that the invariants κ0 (s̄), κ1 (s̄) associated to [β] have the property
that κ0 (s̄) ≡ 0.
(c) Conclude that [β(s̄)] is not a projective parametrization unless [α] is a
conic section. It follows that for non-conic curves, the canonical lifting of a
projective parametrization is never contained in a plane in R3 .
(d) Show that, while [β(s̄)] is not a projective parametrization, it has the
property that the curve β̄ : I → A2 defined by
β̄(s̄) = t[x̄1 (t(s̄)), x̄2 (t(s̄))]
is parametrized by its equi-affine arc length, and the Wilczynski invariant
κ1 (s̄) of [β] is equal to the equi-affine curvature κ(s̄) of β̄. This is related to
214 7. Curves and surfaces in projective space
the fact that the equi-affine transformation group A(2) can be realized as
the subgroup of the projective transformation group SL(3) that preserves
the plane {t[x0 , x1 , x2 ] ∈ R3 | x0 = 1} in R3 .
(e) Let [γ(s)] = [β(s̄(s))] be a projective reparametrization of [β(s̄)], so that
the Wilczynski invariants κ̃0 (s), κ̃1 (s) associated to [γ] satisfy κ̃1 (s) = 0.
Use the formulas (7.3), (7.4) and the fact that κ0 (s̄) = 0 to show that
where κ(s̄) is the equi-affine curvature of β̄ and κ (s̄(s)) denotes the deriva-
tive of κ(s̄) with respect to s̄, evaluated at s̄ = s̄(s).
(f) Conclude that the projective curvature form of the original curve [α] may
be expressed as
κ̂0 = κ̃0 (s)(ds)3 = − 12 κ (s̄)(ds̄)3 ,
where s̄ is the equi-affine arc length of the curve β̄ : I → A2 and κ(s̄) is the
equi-affine curvature of β̄. Comparing with the result of Exercise 6.15, we
see that the projective curvature form is a fifth-order invariant of [α].
and
(7.13) det e0 (t) e1 (t) e2 (t) e3 (t) = 1.
In particular, the vectors (α(t), α (t), α (t), α (t)) should be linearly inde-
pendent for all t ∈ I.
7.3. Moving frames for curves in P3 215
*Exercise 7.24. Show that the canonical projective frame field associated
to a parametrized nondegenerate curve [α] : I → P3 satisfies the structure
equations
e0 (t) = e1 (t), e1 (t) = e2 (t), e2 (t) = e3 (t),
e3 (t) = κ0 (t) e0 (t) + κ1 (t) e1 (t) + κ2 (t) e2 (t)
for some functions κ0 (t), κ1 (t), κ2 (t).
As for curves in P2 , the functions κ0 (t), κ1 (t), κ2 (t) are called the Wilczynski
invariants of the parametrized curve [α] : I → P3 . Next, we investigate how
the Wilczynski invariants transform under a reparametrization of [α].
*Exercise 7.25 (Maple recommended; cf. Exercise 7.7). Let J ⊂ R, and
let [β] : J → P3 be a reparametrization of [α] given by [β(s)] = [α(t(s))],
with t (s) > 0. (The sign of t (s) is not particularly important, but choosing
it to be positive will simplify the following computations.)
(a) Show that the canonical lifting of [β] is given by
1
β(s) = α(t(s)),
t (s)
where α(t) is the canonical lifting of [α].
(b) Show that the canonical projective frame field (ẽ0 (s), ẽ1 (s), ẽ2 (s), ẽ3 (s))
associated to [β] is given by
1
ẽ0 (s) = e0 (t(s)),
(t (s))3/2
3t (s) 1
ẽ1 (s) = − e 0 (t(s)) + e1 (t(s)),
(2t (s))5/2 (t (s))1/2
" #
3t (s) 15(t (s))2 t (s)
ẽ2 (s) = − − e 0 (t(s)) − 2 e1 (t(s))
2(t (s))5/2 4(t (s))7/2 (t (s))3/2
where (e0 (t), e1 (t), e2 (t), e3 (t)) is the canonical projective frame field asso-
ciated to [α].
(c) Show that the Wilczynski invariants κ̃0 (s), κ̃1 (s), κ̃2 (s) associated to [β]
are given by
" #2
1 t(5) (s) t(4) (s)t (s) t (s)
κ̃0 (s) = −24 + 120
+ 60
16 t (s) (t (s)) 2 t (s)
" #4
t (s)(t (s))2 t (s)
−300 + 135
(t (s))3 t (s)
(7.15)
3
+ (t (s))4 κ0 (t(s)) + t (s)(t (s))2 κ1 (t(s))
2
" #
3 3
+ t (s)t (s) − (t (s)) κ2 (t(s)),
2
2 4
" #3
t(4) (s) t (s)t (s) t (s)
κ̃1 (s) = 5 − +4 2
−3
(7.16) t (s) (t (s)) t (s)
+ (t (s))3 κ1 (t(s)) + 2t (s)t (s)κ2 (t(s)),
" #2
5 t (s) t (s)
(7.17) κ̃2 (s) = −2 +3 + (t (s))2 κ2 (t(s)),
2 t (s) t (s)
where κ0 (t), κ1 (t), κ2 (t) are the Wilczynski invariants associated to [α].
Observe that equation (7.17) is very similar to equation (7.4). Precisely the
same argument that we used in Section 7.2 can be applied to equation (7.17)
to show that there exists a reparametrization [β(s)] for which κ̃2 (s) ≡ 0.
Thus, we have the analog of Definition 7.8:
Definition 7.26. Let I ⊂ R, and let [α] : I → P3 be a nondegenerate
curve. [α] is called a projective parametrization if the Wilczynski invariants
κ0 (t), κ1 (t), κ2 (t) associated to [α] have the property that κ2 (t) ≡ 0. In this
case, the parameter t ∈ I is called a projective parameter for [α].
κ0 (t(s)) 3c κ1 (t(s))
κ̃0 (s) = 8
− .
(cs + d) (cs + d)7
Any choice for the function e0 (u) defines a lifting x : U → R4 of [x], given
by
x(u) = e0 (u),
whose image is a surface Σ ⊂ R4 .
7.4. Moving frames for surfaces in P3 221
We might like to proceed as in the equi-affine case, by choosing (e1 (u), e2 (u))
so that they span the tangent space Tx(u) Σ. However, this choice is compli-
cated by the fact that this space is not well-defined: If we replace the lifting
x by an alternate lifting
x̂(u) = λ(u)x(u)
for some nonvanishing function λ : U → R, then the tangent vectors xu , xv ,
which span Tx(u) Σ, are replaced by the tangent vectors
x̂u = λxu + λu x, x̂v = λxv + λv x.
Since the function λ and its partial derivatives at any point are arbitrary
(aside from the requirement that λ = 0), the tangent space Tx(u) Σ is only
well-defined modulo the vector x(u) = e0 (u). Consequently, we can only
require that (e1 (u), e2 (u)) span Tx(u) Σ modulo e0 (u). More concretely, this
means that we will require that the three vectors (e0 (u), e1 (u), e2 (u)) span
the 3-dimensional subspace of R4 determined by the position vector x(u) and
the tangent plane Tx(u) Σ. In particular, we will require that this space be 3-
dimensional, which means that the position vector of Σ cannot be contained
in its tangent plane at any point. (In fact, this condition is precisely what it
means for the smooth map [x] : U → P3 to be an immersion.) A projective
frame field (e0 (u), e1 (u), e2 (u), e3 (u)) along [Σ] satisfying these conditions
will be called 0-adapted.
Exercise 7.32. Show that, given a smooth map [x] : U → P3 , the condition
that span(x, xu , xv ) be 3-dimensional is independent of the choice of lifting
x : U → R4 of [x].
and Cartan’s lemma (cf. Lemma 2.49) implies that there exist functions
h11 , h12 , h22 on U such that
3
ω̄1 h11 h12 ω̄01
(7.21) = .
ω̄23 h12 h22 ω̄02
The next step is to investigate how the matrix [hij ] changes if we vary
the frame. First, we must identify the group of transformations between
0-adapted frames.
Let (e0 (u), e1 (u), e2 (u), e3 (u)) be any 0-adapted frame field, with associated
Maurer-Cartan forms (ω̄βα ). Any other 0-adapted frame field (ẽ0 (u), ẽ1 (u),
ẽ2 (u), ẽ3 (u)) must have the properties that
span(ẽ0 (u)) = span(e0 (u)),
span(ẽ0 (u), ẽ1 (u), ẽ2 (u)) = span(e0 (u), e1 (u), e2 (u)),
and
det ẽ0 (u) ẽ1 (u) ẽ2 (u) ẽ3 (u) = det e0 (u) e1 (u) e2 (u) e3 (u) = 1.
This will be true if and only if
ẽ0 (u) ẽ1 (u) ẽ2 (u) ẽ3 (u)
⎡ ⎤
λ r1 r2 s0
(7.22) ⎢
⎢0 s1 ⎥
⎥
= e0 (u) e1 (u) e2 (u) e3 (u) ⎢ B ⎥
⎣0 s2 ⎦
0 0 0 (λ det B)−1
for some GL(2)-valued function B and real-valued functions λ, r1 , r2 , s0 , s1 ,
s2 on U , with λ = 0. Let (ω̄
˜ βα ) be the Maurer-Cartan forms associated to
the new frame field.
The computations in the following exercise should feel familiar from the
equi-affine case:
*Exercise 7.34 (Cf. Exercise 6.20). (a) Show that
1 1
˜0
ω̄ ω̄0
= λB −1 .
˜ 02
ω̄ ω̄02
(c) Cartan’s lemma implies that there exist functions h̃11 , h̃12 , h̃22 on U such
that 3 1
˜1
ω̄ h̃11 h̃12 ω̄ ˜0
= .
˜ 23
ω̄ h̃12 h̃22 ω̄ ˜ 02
Show that
h̃11 h̃12 h11 h12
(7.23) = (det B) tB B.
h̃12 h̃22 h12 h22
For the remainder of this section, we will assume that [Σ] is elliptic. The
hyperbolic case was treated in detail by Cartan in [Car20]; we will explore
this case in Exercise 7.50.
The transformation (7.23) acts transitively on the set of 2 × 2 matrices with
positive determinant; therefore, there exists a choice of 0-adapted frame field
(e0 (u), e1 (u), e2 (u), e3 (u)) for which
h11 h12 1 0
= .
h12 h22 0 1
Such a frame field will be called 1-adapted.
*Exercise 7.36. Let (e0 (u), e1 (u), e2 (u), e3 (u)) be any 1-adapted frame
field for an elliptic projective surface [Σ] = [x(U )] ⊂ P3 .
(a) Show that any other 1-adapted frame field (ẽ0 (u), ẽ1 (u), ẽ2 (u), ẽ3 (u))
for [Σ] must have the form
ẽ0 (u) ẽ1 (u) ẽ2 (u) ẽ3 (u)
⎡ ⎤
λ r1 r2 s0
(7.24) ⎢
⎢0 s1 ⎥
⎥
= e0 (u) e1 (u) e2 (u) e3 (u) ⎢ B ⎥
⎣0 s2 ⎦
0 0 0 (λ det B) −1
Note that the result of Exercise 7.36(b) is slightly different from the equi-
affine case. Now the quadratic form
(7.25) I = ω̄13 ω̄01 + ω̄23 ω̄02 = (ω̄01 )2 + (ω̄02 )2
is well-defined only up to a scalar multiple. Rather than defining a metric,
it defines a conformal structure on [Σ], in which angles between tangent
vectors are well-defined, but lengths of tangent vectors are not.
Exercise 7.37. Show that the conformal structure I is invariant (up to a
scalar multiple) under the action of SL(4).
Next, we need to compute how the functions (hijk ) vary under a transfor-
mation of the form (7.24). This computation gets rather complicated, but
we can make it simpler by breaking it down into two steps.
*Exercise 7.39. Consider transformations of the form (7.24) with B equal
to the identity matrix and λ = 1, so that
⎡ ⎤
1 r1 r2 s0
⎢0 1 0 s1 ⎥
ẽ0 (u) ẽ1 (u) ẽ2 (u) ẽ3 (u) = e0 (u) e1 (u) e2 (u) e3 (u) ⎢ ⎥
⎣0 0 1 s2 ⎦ .
0 0 0 1
7.4. Moving frames for surfaces in P3 225
Therefore,
so there exists a choice of 1-adapted frame field (e0 (u), e1 (u), e2 (u), e3 (u))
for which
Any 1-adapted frame field satisfying the condition (7.28) will be called 2-
adapted.
Remark 7.40. This is certainly not the only way that one might choose to
normalize these functions! But our experience with the equi-affine case—
where a similar relation held for any 2-adapted frame field—suggests that
this might turn out to be a nice choice.
*Exercise 7.41. Let (e0 (u), e1 (u), e2 (u), e3 (u)) be any 2-adapted frame
field for an elliptic projective surface [Σ] = [x(U )] ⊂ P3 .
(a) Show that any other 2-adapted frame field (ẽ0 (u), ẽ1 (u), ẽ2 (u), ẽ3 (u))
for [Σ] must have the form (7.24), where
r1 s1
(7.29) = λ tB .
r2 s2
(b) Show that under a transformation of the form (7.24) satisfying equation
(7.29), we have
h̃111 h111
−1 3
(7.30) =λ B .
h̃222 h222
(c) Let (ω̄βα ) be the Maurer-Cartan forms associated to the 2-adapted frame
˜ βα ) be the Maurer-Cartan forms
field (e0 (u), e1 (u), e2 (u), e3 (u)), and let (ω̄
associated to the 2-adapted frame field (ẽ0 (u), ẽ1 (u), ẽ2 (u), ẽ3 (u)). Show
226 7. Curves and surfaces in projective space
that
1 3 2 3
˜ 0 ) − 3ω̄
h̃111 (ω̄ ˜ 01 (ω̄ ˜ 0 ) − 3(ω̄
˜ 02 )2 + h̃222 (ω̄ ˜ 01 )2 ω̄
˜ 02
= λ2 h111 (ω̄01 )3 − 3ω̄01 (ω̄02 )2 + h222 (ω̄02 )3 − 3(ω̄01 )2 ω̄02 .
I = (ω̄01 )2 + (ω̄02 )2
Remark 7.44. For a 3-adapted frame field, the cubic form (7.31) becomes
P = 2 (ω̄01 )3 − 3 ω̄01 (ω̄02 )2 .
The 3-adapted frames have the property that the tangent vector e2 (u) is one
of the three null directions for P at the point [x(u)] ∈ [Σ]. These directions
are called the Darboux tangents at the point [x(u)] ∈ [Σ]; for more details,
see [Su83].
*Exercise 7.45. (a) Show that the Maurer-Cartan forms associated to any
3-adapted frame field satisfy the equations
(Hint: Apply the result of Exercise 7.38, keeping in mind condition (7.28),
and recall the defining condition for the Lie algebra sl(4).)
228 7. Curves and surfaces in projective space
ω̄00 + ω̄33 = 0
and use Cartan’s lemma to conclude that there exist functions 11 , 12 , 22
on U such that
1
ω̄10 − ω̄31 11 12 ω̄0
(7.33) = .
ω̄20 − ω̄32 12 22 ω̄02
*Exercise 7.46. (a) Show that under a transformation of the form (7.32),
we have
(b) Conclude that there exists a choice of 3-adapted frame field for which
where (λ, B) are one of the pairs from Exercise 7.42. Therefore, the 4-
adapted frames at each point form a discrete set, and a continuous 4-adapted
frame field along [Σ] is uniquely determined by its value at any point u ∈ U .
(b) Differentiate the equations in part (a) and use Cartan’s lemma to con-
clude that there exist functions p1 , p2 , q11 , q12 , q21 , q22 such that
ω̄21 = p1 ω̄01 + (p2 − 1)ω̄02 ,
ω̄12 = −p1 ω̄01 − (p2 + 1)ω̄02 ,
ω̄31 = q11 ω̄01 + q12 ω̄02 ,
(7.35)
ω̄32 = q21 ω̄01 + q22 ω̄02 ,
ω̄00 = −ω̄33 = 3p2 ω̄01 − 3p1 ω̄02 ,
ω̄30 = (q11 − q22 )ω̄01 − (q12 + q21 )ω̄02 .
The functions p1 , p2 , q11 , q12 , q21 , q22 are the fundamental invariants for el-
liptic surfaces in P3 with no umbilic points.
Conclude that there exists a 2-adapted frame field with σ = 0 and that for
such a frame field we have
(7.37) ω̄31 = ω̄10 , ω̄32 = ω̄20 .
We will call such a frame field fully adapted.
(d) Show that any two fully adapted frame fields differ by a transformation
of the form
ẽ0 (u) ẽ1 (u) ẽ2 (u) ẽ3 (u)
⎡ ⎤
λ r1 r2 s0
(7.38) ⎢
⎢ 0 s1 ⎥⎥
= e0 (u) e1 (u) e2 (u) e3 (u) ⎢ B ⎥,
⎣0 s2 ⎦
0 0 0 λ−1
where
1
B ∈ SO(2), [r1 r2 ] = λ[s1 s2 ]B, s0 = λ(s21 + s22 ).
2
−1 0 0 0
Q represents the quadratic form
Q(x) = (x1 )2 + (x2 )2 − 2x0 x3 ,
which is an indefinite quadratic form of signature (3, 1) on R4 . The Lie
group SO(Q) is defined to be the group of matrices of determinant 1 that
preserve this quadratic form; i.e.,
SO(Q) = {A ∈ SL(4) | tAQA = Q}.
The group SO(Q) is isomorphic to the Lie group SO(3, 1).
(a) Show that the Lie algebra so(Q) is defined by
so(Q) = {B ∈ sl(4) | tBQ + QB = 0}.
(Hint: Let A(t) be a curve in SO(Q) with A(0) = I. Differentiate the
equation
t
A(t)QA(t) = Q
and evaluate at t = 0.)
(b) Show that so(Q) consists of all matrices of the form
⎡ 0 0 0 ⎤
a0 a1 a2 0
⎢ 1 ⎥
⎢a0 0 a12 a01 ⎥
B=⎢ ⎢a2 −a1 0 a0 ⎥ .
⎥
⎣ 0 2 2 ⎦
0 a10 a20 −a00
Conclude that the reduced Maurer-Cartan form on the bundle of fully a-
dapted frames in Exercise 7.48 takes values in so(Q).
(c) Recall that for any fully adapted frame field
x̃(u) = (e0 (u), e1 (u), e2 (u), e3 (u))
as in Exercise 7.48, the pullback of the Maurer-Cartan form ω̄ to U via x̃
is given by x̃∗ (ω̄) = x̃−1 dx̃. Use this fact to show that any fully adapted
frame field has the form
x̃(u) = A0 A(u),
where A0 ∈ SL(4) is a constant matrix and A(u) ∈ SO(Q) for all u ∈ U .
232 7. Curves and surfaces in projective space
[A−1 −1 −1
0 Σ] = [A0 x(U )] = [A0 e0 (U )] ⊂ P .)
3
(b) Let (f0 (u), f1 (u), f2 (u), f3 (u)) be any 1-adapted null frame field for a
hyperbolic projective surface [Σ] = [x(U )] ⊂ P3 . Show that any other 1-
adapted null frame field (f̃0 (u), f̃1 (u), f̃2 (u), f̃3 (u)) for Σ must have the form
(7.40) ⎡ ⎤
λ r1 r2 s0
⎢ ⎥
⎢ 0 eθ 0 s1 ⎥
f̃0 (u) f̃1 (u) f̃2 (u) f̃3 (u) = f0 (u) f1 (u) f2 (u) f3 (u) ⎢ ⎢ 0 0 e−θ s ⎥
⎥
⎣ 2 ⎦
0 0 0 λ−1
for some real-valued functions λ, θ, r1 , r2 , s0 , s1 , s2 on U , with λ = 0.
(c) Differentiate equations (7.39) and use Cartan’s lemma to conclude that
there exist functions h111 , h112 , h122 , h222 on U such that
⎡ ⎤ ⎡ ⎤
2η̄12 h111 h112
⎢ 1 ⎥ ⎢ ⎥ η̄ 1
⎢(η̄ + η̄ 2 ) − (η̄ 0 + η̄ 3 )⎥ = ⎢h112 h122 ⎥ 0 .
⎣ 1 2 0 3 ⎦ ⎣ ⎦ 2
η̄0
2η̄21 h122 h222
Equations (7.41) imply that the functions h111 , h222 are relative invariants;
for the remainder of this exercise, we will assume that both h111 , h222 are
nonzero. (Cartan showed in [Car20] that if either of these functions is
identically zero, then [Σ] is a ruled surface in P3 .) Equations (7.41) then
imply that there exists a choice of 2-adapted null frame field for which
h111 = h222 = 2.
234 7. Curves and surfaces in projective space
0 0 0 1
(g) Show that the Maurer-Cartan forms associated to any 3-adapted null
frame field satisfy the equations
η̄00 + η̄33 = η̄11 + η̄22 = 0.
Differentiate the equation
η̄00 + η̄33 = 0
and use Cartan’s lemma to conclude that there exist functions 11 , 12 , 22
on U such that 0
η̄1 − η̄32 11 12 η̄01
= .
η̄20 − η̄31 12 22 η̄02
Differentiate these equations and use Cartan’s lemma to conclude that there
exist functions p1 , p2 , q11 , q12 , q21 , q22 such that
η̄11 = −η̄22 = p1 η̄01 + p2 η̄02 ,
η̄00 = −η̄33 = −3p1 η̄01 + 3p2 η̄02 ,
η̄31 = q11 η̄01 + q12 η̄02 ,
η̄32 = q21 η̄01 + q22 η̄02 ,
η̄30 = q12 η̄01 + q21 η̄02 .
The functions p1 , p2 , q11 , q12 , q21 , q22 are the fundamental invariants for non-
ruled, hyperbolic surfaces in P3 .
In this chapter, the computations for curves in P2 and P3 are already suffi-
ciently complicated that some assistance from Maple may be appreciated.
Here we will explore some of the exercises for curves in P3 ; the case of curves
in P2 is left as an exercise for the reader. As usual, begin by loading the
Cartan and LinearAlgebra packages into Maple.
Exercise 7.25: The canonical lifting ẽ0 (s) of [β(s)] must be equal to a
scalar multiple of e0 (t(s)); say
ẽ0 (s) = λ(s) e0 (t(s)).
For Maple purposes, we will denote the canonical projective frame field for
[β(s)] by (f0(s), f1(s), f2(s), f3(s)).
> f0(s):= lambda(s)*e0(t(s));
Compute the first three derivatives of ẽ0 (s):
> f1(s):= diff(f0(s), s);
f2(s):= diff(f1(s), s);
f3(s):= diff(f2(s), s);
Maple expresses the output from these computations in terms of the deriva-
tives of e0 (t(s)); for instance, it returns
f 1(s) := λs e0(t(s)) + λ(s) D(e0)(t(s)) ts
d
(The expression D(e0)(t(s)) refers to the derivative dt (e0 (t)) evaluated at
t = t(s).)
We don’t have any good way of telling Maple that these derivatives are all
linearly independent, but we can still collect terms and extract the coeffi-
cients needed to compute the necessary determinant.
236 7. Curves and surfaces in projective space
1
λ(s) = .
(t (s))(3/2)
Now computing the rest of the canonical projective frame field for [β] is
simply a matter of differentiating the expression
1
ẽ0 (s) = e0 (t(s)).
(t (s))(3/2)
But keep in mind that we still need to compute the Wilczynski invariants,
which requires expressing ẽ3 (s) as a linear combination of (ẽ0 (s), ẽ1 (s),
ẽ2 (s)). This can get a bit messy, but we can keep things better organized if,
instead of explicitly expressing everything as functions of t and s, we instead
work with appropriately chosen differential forms. In this context, we assign
the derivatives (and the Wilczynski invariants) of the original frame field as
follows:
> d(e0):= e1*d(t);
d(e1):= e2*d(t);
d(e2):= e3*d(t);
d(e3):= e0*kappa0*d(t) + e1*kappa1*d(t) + e2*kappa2*d(t);
Next, observe that the function t(s) never appears explicitly; all that matters
is its derivative t (s). So, let r(s) denote t (s), and set up substitutions that
relate dt to ds and ẽ0 to e0 :
> PDETools[declare](r(s));
tsub:= [d(t) = r(s)*d(s)];
fsub:= [f0 = e0/r(s)ˆ(3/2)];
7.5. Maple computations 237
Compute the derivative of ẽ0 with respect to s by computing dẽ0 and then
substituting for dt in terms of ds:
> Simf(subs(tsub, Simf(d(Simf(subs(fsub, f0))))));
The coefficient of ds in this expression should be ẽ1 (s), so add it to fsub:
> fsub:= [op(fsub), f1 = pick(%, d(s))];
Similarly for ẽ2 (s) and ẽ3 (s):
> Simf(subs(tsub, Simf(d(Simf(subs(fsub, f1))))));
fsub:= [op(fsub), f2 = pick(%, d(s))];
Simf(subs(tsub, Simf(d(Simf(subs(fsub, f2))))));
fsub:= [op(fsub), f3 = pick(%, d(s))];
In order to compute the Wilczinksi invariants for the reparametrized curve,
we need to express dẽ3 in terms of (ẽ0 , ẽ1 , ẽ2 ). The makebacksub command
won’t work here because (e0 , e1 , e2 , e3 ) are not 1-forms, but we can create
the reverse substitution as follows:
> esub:= [op(solve({op(fsub)}, {e0, e1, e2, e3}))];
Now we can compute dẽ3 and express the result in terms of the canonical
projective frame field for [β]:
> Simf(subs(esub, Simf(subs(tsub,
Simf(d(Simf(subs(fsub, f3))))))));
df3:= collect(pick(%, d(s)), {f0, f1, f2, f3});
As expected, we see that the output contains no ẽ3 term, and the invariants
κ̃0 (s), κ̃1 (s), κ̃2 (s) associated to [β] are given by
> Kappa0:= coeff(df3, f0);
Kappa1:= coeff(df3, f1);
Kappa2:= coeff(df3, f2);
Exercise 7.28: First, it will be helpful to introduce variables to represent
the derivatives of κ0 (t), κ1 (t), κ2 (t):
> d(kappa0):= kappa0 t*d(t);
d(kappa1):= kappa1 t*d(t);
d(kappa2):= kappa2 t*d(t);
d(kappa0 t):= kappa0 tt*d(t);
d(kappa1 t):= kappa1 tt*d(t);
d(kappa2 t):= kappa2 tt*d(t);
238 7. Curves and surfaces in projective space
Next, we will need to express some of the derivatives of κ̃0 (s), κ̃1 (s), κ̃2 (s)
(with respect to s) in terms of κ0 (t), κ1 (t), κ2 (t) and their derivatives (with
respect to t). As it turns out, these are the ones that we will need:
> Kappa1 s:= Simf(pick(Simf(subs(tsub, d(Kappa1))), d(s)));
Kappa2 s:= Simf(pick(Simf(subs(tsub, d(Kappa2))), d(s)));
Kappa2 ss:= Simf(pick(Simf(subs(tsub, d(Kappa2 s))), d(s)));
In order to identify invariant forms, the trick is to look for combinations of
κ̃0 (s), κ̃1 (s), κ̃2 (s) and their derivatives that don’t contain any derivatives
of r(s), so that they transform tensorially under the change of variables
t → t(s). The first one is relatively easy to spot: The computations above
show that
and hence
κ̃1 (s) − κ̃2 (s) ds3 = κ1 (t) − κ2 (t) (r(s) ds)3 = κ1 (t) − κ2 (t) dt3 .
The quartic form requires a bit more finesse to identify; for details, see the
Maple worksheet for this chapter on the AMS webpage. But once we know
the answer, verifying it is straightforward:
> Simf(Kappa0 - (1/2)*Kappa1 s + (1/5)*Kappa2 ss
+ (9/100)*Kappa2ˆ2);
1 4
r 100 κ0 + 9 κ22 + 20 kappa2 tt − 50 kappa1 t
100
at this point because doing so would require solving for one of the forms in
terms of the others, and it’s not yet clear which form would be best to solve
for. So for now, we’ll simply keep the relation in mind, and we’ll tell Maple
about it when it becomes convenient to do so.
Next, set up the initial substitution for the Maurer-Cartan forms of a 0-
adapted frame field:
> adaptedsub1:= [omega[3,0]=0,
omega[3,1] = h[1,1]*omega[1,0] + h[1,2]*omega[2,0],
omega[3,2] = h[1,2]*omega[1,0] + h[2,2]*omega[2,0]];
Exercise 7.34: Introduce new 1-forms to represent the transformed forms:
> Form(Omega[0,0], Omega[0,1], Omega[0,2], Omega[0,3],
Omega[1,0], Omega[1,1], Omega[1,2], Omega[1,3],
Omega[2,0], Omega[2,1], Omega[2,2], Omega[2,3],
Omega[3,0], Omega[3,1], Omega[3,2], Omega[3,3]);
Ultimately, we’re going to need to know how all of the Maurer-Cartan forms
transform as we gradually reduce the transformation group. And while it
would be quite a mess to write everything out explicitly, we know that under
a transformation of the form (7.22), we have
(7.44) ˜ βα = A−1 dA + A−1 ω̄βα A,
ω̄
where A is the matrix in equation (7.22). We can go ahead and set up this
substitution for all the Maurer-Cartan forms at once—although you will
probably want to suppress the output so as not to have Maple spew out
pages and pages of nasty expressions that we don’t really need to see!
240 7. Curves and surfaces in projective space
b[2,1]:= sin(theta);
b[2,2]:= cos(theta);
Differentiate the first equation in (7.26):
> zero4:= Simf(subs(adaptedsub1,
Simf(d(omega[3,1] - omega[1,0]))));
Use the pick command to write this expression in the form
φ1 ∧ ω̄01 + φ2 ∧ ω̄02
so that we can apply Cartan’s lemma:
pick(zero4, omega[1,0]);
−ω0,0 + 2 ω1,1 − ω3,3
pick(zero4, omega[2,0]);
ω1,2 + ω2,1
Equation (7.27) then follows from Cartan’s lemma. Now add these condi-
tions to adaptedsub1:
> adaptedsub1:= [op(adaptedsub1),
omega[1,1] = (1/2)*(omega[0,0] + omega[3,3]
+ h[1,1,1]*omega[1,0] + h[1,1,2]*omega[2,0]),
omega[2,1] = -omega[1,2] + h[1,1,2]*omega[1,0]
+ h[1,2,2]*omega[2,0],
omega[2,2] = (1/2)*(omega[0,0] + omega[3,3]
+ h[1,2,2]*omega[1,0] + h[2,2,2]*omega[2,0])];
Exercise 7.39: We may as well go ahead and determine how the (hijk )
transform under the full 1-adapted group action. We can compute the func-
tions (h̃ijk ) as follows:
> newform1:= Simf(subs(adaptedsub2,
Simf(subs(framechangebacksub, Simf(subs(adaptedsub1,
7.5. Maple computations 243
Simf(subs(framechangesub,
2*Omega[1,1] - Omega[0,0] - Omega[3,3]))))))));
> H[1,1,1]:= pick(newform1, Omega[1,0]);
H[1,1,2]:= pick(newform1, Omega[2,0]);
> newform2:= Simf(subs(adaptedsub2,
Simf(subs(framechangebacksub, Simf(subs(adaptedsub1,
Simf(subs(framechangesub,
2*Omega[2,2] - Omega[0,0] - Omega[3,3]))))))));
> H[1,2,2]:= pick(newform2, Omega[1,0]);
H[2,2,2]:= pick(newform2, Omega[2,0]);
The resulting expressions look fairly complicated, but they simplify consid-
erably if we take θ = 0, λ = 1:
> Simf(subs([theta=0, lambda=1], H[1,1,1]));
h1,1,1 − 3 s1 + 3 r1
h1,1,2 − s2 + r2
h1,2,2 − s1 + r1
h2,2,2 − 3 s2 + 3 r2
Exercise 7.41: Now we assume that both frame fields are 2-adapted, so that
the condition (7.28) holds for both sets of Maurer-Cartan forms. So, we tell
Maple that it holds for the (hijk ) and then determine what conditions must
be satisfied by the group parameters in order for it to hold for the (h̃ijk ) as
well:
> h[1,2,2]:= -h[1,1,1];
h[1,1,2]:= -h[2,2,2];
> zero6:= Simf(H[1,1,1] + H[1,2,2]);
zero7:= Simf(H[1,1,2] + H[2,2,2]);
> solve({zero6, zero7}, {r[1], r[2]});
> assign(%);
244 7. Curves and surfaces in projective space
Now compute how the functions h111 , h222 transform under this restricted
group action:
> collect(Simf(H[1,1,1]), {h[1,1,1], h[2,2,2]});
> collect(Simf(H[2,2,2]), {h[1,1,1], h[2,2,2]});
If the output looks too complicated to recognize immediately, we can check
that this does, in fact, yield the result in equation (7.30):
> Simf(Vector([H[1,1,1], H[2,2,2]])
- (1/lambda)*B.B.B.Vector([h[1,1,1], h[2,2,2]]));
0
0
ω0,0 + ω3,3
It follows from Cartan’s lemma that equations (7.33) must hold, and so we
add them to our substitution:
> adaptedsub13:= [op(adaptedsub13),
omega[1,3] = omega[0,1] + ell[1,1]*omega[1,0]
+ ell[1,2]*omega[2,0],
omega[2,3] = omega[0,2] + ell[1,2]*omega[1,0]
+ ell[2,2]*omega[2,0]];
Exercise 7.46: Now we compute how the (ij ) transform under the trans-
formation (7.32) between 3-adapted frame fields.
> newform4:= Simf(subs(threeadaptedsub, Simf(subs(adaptedsub2,
Simf(subs(framechangebacksub, Simf(subs(adaptedsub13,
Simf(subs(framechangesub, Omega[1,3] - Omega[0,1]))))))))));
> newform5:= Simf(subs(threeadaptedsub, Simf(subs(adaptedsub2,
Simf(subs(framechangebacksub, Simf(subs(adaptedsub13,
Simf(subs(framechangesub, Omega[2,3] - Omega[0,2]))))))))));
> LL[1,1]:= pick(newform4, Omega[1,0]);
LL[1,2]:= pick(newform4, Omega[2,0]);
LL[2,2]:= pick(newform5, Omega[2,0]);
In order to show that we can always find a 3-adapted frame for which the
(ij ) are all equal to zero, it suffices to show that we can solve the equations
˜11 = ˜12 = ˜22 = 0
for s0 , s1 , s2 :
> solve({LL[1,1], LL[1,2], LL[2,2]}, {s[0], s[1], s[2]});
1 1 1 2 1 2 1 1 2
{s0 = − ell1,1 − ell2,2 + ell1,2 + ell2,2 − ell2,2 ell1,1 + ell1,1 ,
4 4 8 32 16 32
1 1 1
s1 = ell2,2 − ell1,1 , s2 = ell1,2 }
4 4 2
246 7. Curves and surfaces in projective space
Exercise 7.47: For a 4-adapted frame field, we have the following condi-
tions:
> fouradaptedsub:= [h[1,1,1] = 2, h[2,2,2] = 0,
ell[1,1] = 0, ell[1,2] = 0, ell[2,2] = 0];
We can refine our 3-adapted frame adaptation by incorporating these con-
ditions as follows:
> adaptedsub14:= Simf(subs(fouradaptedsub, adaptedsub13));
We can now read off the relations (7.34) directly from adaptedsub14.
At this point, all the remaining Maurer-Cartan forms must be linear com-
binations of (ω̄01 , ω̄02 ). It follows from equations (7.34) that the first four
equations in (7.35) must hold for some functions (pi , qij ). Before we con-
tinue, add these expressions to adaptedsub14 in two steps:
> adaptedsub14:= Simf(subs([
omega[1,2] = p[1]*omega[1,0] + (p[2] - 1)*omega[2,0],
omega[0,1] = q[1,1]*omega[1,0] + q[1,2]*omega[2,0],
omega[0,2] = q[2,1]*omega[1,0] + q[2,2]*omega[2,0]],
adaptedsub14));
> adaptedsub14:= [op(adaptedsub14),
omega[1,2] = p[1]*omega[1,0] + (p[2] - 1)*omega[2,0],
omega[0,1] = q[1,1]*omega[1,0] + q[1,2]*omega[2,0],
omega[0,2] = q[2,1]*omega[1,0] + q[2,2]*omega[2,0]];
Now differentiate the first two equations in (7.34):
> Simf(subs(adaptedsub14, Simf(d(omega[1,1] - omega[1,0]))));
> pick(%, omega[1,0]);
−ω0,0 − 3 p1 ω2,0
2 ω0,0 − 6 p2 ω1,0
By Cartan’s lemma, (2ω̄00 − 6p2 ω̄01 ) must be equal to a multiple of ω̄02 . To-
gether, these conditions imply that
Applications of
moving frames
Chapter 8
Minimal surfaces
in E3 and A3
8.1. Introduction
The study of minimal surfaces goes back to 1760, when Lagrange posed the
following problem: Given a closed curve in E3 , can we find a surface of min-
imum area among all surfaces that have the given curve as their boundary?
This question is called the Plateau problem; it is named after the physicist
Joseph Plateau, who in the nineteenth century performed experiments with
wire frames and soap films in order to study the properties of such surfaces.
Despite the long history, rigorous existence results for area-minimizing sur-
faces with general boundary curves were only proved beginning in the 1930s
[Dou31].
251
252 8. Minimal surfaces in E3 and A3
i.e., A(Σ) is simply the area of Σ. The key idea is that minimal surfaces
should be critical points of this functional.
Unfortunately, since the space S is infinite-dimensional (with a rather com-
plicated topology!), it isn’t entirely obvious how to go about finding critical
points of the functional A. The solution to this problem lies in the idea of
a variation. We start with the observation that if
f :M →R
(1) For each t ∈ (−ε, ε), the set Σt = X(U, t) is a regular surface in E3 .
(2) For each u ∈ U , X(u, 0) = x(u). (In particular, Σ0 = Σ.)
(c) Let nt (u) denote the unit normal vector to the surface Σt at the point
xt (u). Then we can decompose ∂X ∂t as
∂X ∂X ∂X
= a(u, t) 1 + b(u, t) 2 + c(u, t)nt
∂t ∂u ∂u
∂ X̄
for some functions a(u, t), b(u, t), c(u, t). Show that the condition that ∂t
is parallel to nt (u) is equivalent to the system of differential equations
∂u1 ∂u2
(8.2) = −a(u1 , u2 , t), = −b(u1 , u2 , t)
∂t ∂t
for the functions u1 (ū, t), u2 (ū, t). (These equations should be regarded as
ordinary differential equations, with t as the independent variable and ū =
(ū1 , ū2 ) as parameters.) Moreover, the condition u(ū, 0) = ū is equivalent
to the initial conditions
Remark 8.5. You might worry that, because the value of ε in the previous
exercise depends on ū, it could happen that there is no single ε > 0 that
suffices for all ū ∈ U . Fortunately, the hypothesis that the variation is com-
pactly supported saves the day: Any positive-valued function on a compact
set must have a positive lower bound; therefore there exists a real number
ε > 0 such that the desired variation X̄ is well-defined on U × (−ε, ε).
8.2. Minimal surfaces in E3 255
Now we can give a rigorous definition that corresponds to our intuitive idea
of a critical point for the area functional A on S:
Remark 8.8. Despite the name, minimal surfaces are not necessarily global
minimizers for the area functional; for instance, there may be multiple
minimal surfaces with different areas spanning a given boundary curve or
curves. However, it can be shown that any minimal surface Σ is locally
area-minimizing in the sense that, given any point q ∈ Σ, there exists a
neighborhood V ⊂ Σ of q that is area-minimizing among all surfaces with
the same boundary curve as V .
Now, it still isn’t entirely obvious how to use this definition to find minimal
surfaces. In order to find critical points q0 for a function f : M → R on a
finite-dimensional manifold M , it suffices to identify those points q0 ∈ M
where all the partial derivatives of f vanish; however, for the functional
A on the infinite-dimensional space S, there is no finite set of compactly
supported variations that plays a role analogous to the partial derivatives of
f . Instead, we will need to find a way to work directly with Definition 8.6.
Let x : U → E3 be an immersion whose image is a surface Σ ⊂ E3 . We
begin by choosing an adapted orthonormal frame field (e1 (u), e2 (u), e3 (u))
along Σ, so that (e1 (u), e2 (u)) span the tangent plane Tx(u) Σ and e3 (u) is
orthogonal to Tx(u) Σ. Recall that choosing an orthonormal frame field along
Σ is equivalent to choosing a lifting x̃ : U → E(3) defined by
The pullbacks (ω̄ i , ω̄ji ) to U of the Maurer-Cartan forms (ω i , ωji ) on E(3) via
the lifting x̃ satisfy the conditions that ω̄ 3 = 0 and
3
ω̄1 h11 h12 ω̄ 1
=
ω̄23 h12 h22 ω̄ 2
256 8. Minimal surfaces in E3 and A3
for some functions h11 , h12 , h22 on U , and the Gauss and mean curvature
functions of Σ (cf. Definition 4.45) are given by
K = h11 h22 − h212 , H = 12 (h11 + h22 ).
8
via X.
We can define such a frame field as follows: For each (u, t) ∈ U × (−ε, ε),
let (e1 (u, t), e2 (u, t), e3 (u, t)) be an orthonormal frame for the surface Σt at
the point xt (u), with e3 (u, t) normal to the tangent plane Txt (u) Σt . Then
the pullbacks (ω̄ i ) of the (ω i ) to U via X 8 must satisfy the equation
dX = ei ω̄ i .
*Exercise 8.9. (a) For each t ∈ (−ε, ε), let ıt : U → U × (−ε, ε) denote
the inclusion map defined by ıt (u) = (u, t). Show that the pullbacks of the
1-forms
ω̄ 1 = dX, e1 , ω̄ 2 = dX, e2
to U via ıt are the usual dual forms on the surface Σt = xt (U ). (Hint: This
isn’t as complicated as it sounds. It is an immediate consequence of the fact
that the immersion xt can be written as the composition
xt = X ◦ ıt
and the fact that pullbacks behave well with respect to composition—spe-
cifically,
(xt )∗ = (ıt )∗ ◦ (X)∗ .)
For convenience, let χ(u, t) = ± ∂X
∂t (u, t) , with the sign chosen so that
3
ω̄ = χ(u, t) dt.
*Exercise 8.10. (a) Differentiate the equation ω̄ 3 = χ dt to obtain
ω̄13 ∧ ω̄ 1 + ω̄23 ∧ ω̄ 2 + dχ ∧ dt = 0.
Now, recall that for any adapted orthonormal frame field x̃ : U → E(3) on
the surface Σ = x(U ), the area of Σ is given by
& &
A(Σ) = ω̄ ∧ ω̄ =
1 2
x̃∗ (ω 1 ∧ ω 2 )
U U
(cf. Exercise 4.47). Applying this formula to each surface Σt in the variation
yields &
A(Σt ) = x̃∗t (ω 1 ∧ ω 2 ).
U
The surface Σ = Σ0 is minimal if and only if for every compactly supported
normal variation, we have
d
0 = A(Σt )
dt t=0
&
d
(8.6) = x̃∗ (ω 1 ∧ ω 2 )
dt t=0 U t
&
d ∗ 1
= x̃t (ω ∧ ω 2
) .
U dt t=0
In order to evaluate this integral, recall from §2.11 that the Lie derivative
of a p-form Φ along a vector field v is the p-form
ϕ∗t Φ − Φ
Lv Φ = lim ,
t→0 t
where ϕt is the flow of the vector field v. In other words,
d
Lv Φ = ϕ∗t Φ.
dt t=0
*Exercise 8.12. (a) Show that the inclusion map ıt : U → U × (−ε, ε) may
∂
be regarded as the flow of the vector field ∂t on U × (−ε, ε), restricted to
the set U × {0} ⊂ U × (−ε, ε).
(b) Use part (a) to show that
d ∗ 1
ıt (ω̄ ∧ ω̄ 2 ) = L∂/∂t (ω̄ 1 ∧ ω̄ 2 ).
dt t=0
(See §2.11 for the relevant definitions.)
(c) Use equation (8.7) to conclude that if the immersion x : U → E3 is
minimal, then for any compactly supported, normal variation of x, we must
have
&
(8.8) L∂/∂t (ω̄ 1 ∧ ω̄ 2 ) = 0.
U
(b) Use the Cartan structure equations, the equation ω̄ 3 = χ dt, and equa-
tion (8.5) to show that
d(ω̄ 1 ∧ ω̄ 2 ) = −(h11 + h22 )χ ω̄ 1 ∧ ω̄ 2 ∧ dt = −2Hχ ω̄ 1 ∧ ω̄ 2 ∧ dt.
(d) Use the results of parts (b) and (c) to compute the matrix [hij ], and
show that the mean curvature of Σ is H = 0. Therefore, the catenoid is a
minimal surface.
(e) (Maple recommended) Repeat the computations of parts (a)–(d) for an
arbitrary non-planar surface of revolution, parametrized by
Show that the surface is minimal if and only if the function ρ(v) satisfies
the differential equation
(d) Use the results of parts (b) and (c) to compute the matrix [hij ], and
show that the mean curvature of Σ is H = 0. Therefore, the helicoid is a
minimal surface.
previous subsection, the associated Maurer-Cartan forms (ω̄ i , ω̄ji ) have the
property that
3
ω̄1 h11 h12 ω̄ 1
=
ω̄23 h12 h22 ω̄ 2
for some functions (hij ) on U , and Σ is minimal if and only if h11 + h22 = 0.
Consider the complex, vector-valued 1-form
Remark 8.20. In keeping with the notation used throughout this book,
we will continue to use (ω̄ i , ω̄ji ) to denote the (real-valued!) Maurer-Cartan
forms associated to the orthonormal frame field (e1 (u), e2 (u), e3 (u)). In
order to avoid confusion, we will use the notation z ∗ rather than z̄ to denote
complex conjugation.
Now suppose that Σ is minimal. Let (e1 (u), e2 (u), e3 (u)) be any adapted or-
thonormal frame field along Σ, with associated Maurer-Cartan forms (ω̄ i , ω̄ji ),
and consider the complex, scalar-valued 1-form ω̄ 1 + iω̄ 2 on U .
*Exercise 8.22. (a) Use the Frobenius theorem (cf. Theorem 2.33) to show
that every point u ∈ U has a neighborhood V ⊂ U on which there exist
complex-valued functions z, ϕ : V → C such that
(b) Show that the function ϕ must be nonzero at every point of V . (Hint:
The 2-form
is nonvanishing on Σ.)
(c) Write the function z as
z = u + iv,
The next exercise introduces some basic results from complex analysis that
will be needed for the remainder of this section.
1 1
u = (z + z ∗ ), v= (z − z ∗ ).
2 2i
∂w
(8.11) = 0,
∂z ∗
i.e., if w is a function of z alone.
(b) Show that equation (8.11) is equivalent to
(8.12) wu + iwv = 0.
264 8. Minimal surfaces in E3 and A3
(c) Write
w(u, v) = x(u, v) + iy(u, v),
where x, y : V → R are real-valued, differentiable functions on V . Show
that equation (8.12) is equivalent to the pair of equations
(8.13) xu = yv , xv = −yu .
Equations (8.13) are called the Cauchy-Riemann equations.
(d) Show that any differentiable functions x, y : V → R satisfying equations
(8.13) must also satisfy the equations
(8.14) xuu + xvv = yuu + yvv = 0.
In other words, both x and y must be harmonic functions.
(c) Apply the Poincaré lemma (cf. Theorem 2.31) to the closed 1-form ξ =
f (z) dz to conclude that there exists a vector-valued, holomorphic function
8.2. Minimal surfaces in E3 265
z : V → C3 such that
ξ = dz,
and therefore, f (z) = z (z).
(d) Show that
Re(dz) = Re(ξ) = e1 ω̄ 1 + e2 ω̄ 2 = dx,
where x : U → Σ is the original parametrization of Σ. (Here Re denotes
the “real part”; i.e., Re(a + ib) = a.) Conclude that, up to a translation, we
must have
x(z) = Re (z(z)) .
Therefore, we can write z(z) as
z(z) = x(z) + iy(z)
for some function y : V → R3 . Moreover, from part (b), we have
z (z), z (z) = 0.
(Note, however, that the requirement that x be an immersion implies that
the vector z (z) is never zero for any z ∈ V .)
*Exercise 8.27. (a) Show that under the given assumptions on f and g,
the integrands in (8.15) are holomorphic and therefore the function z(z) is
holomorphic.
(b) Show that z (z) = 0 for all z ∈ U and that z (z), z (z) = 0. Conclude
that x = Re(z) : U → R3 is a parametrization for an immersed minimal
surface in E3 .
Conversely, the following exercise shows that every minimal surface has a
local parametrization of this form.
(b) Define
ξ3 (z)
f (z) = ξ1 (z) − iξ2 (z), g(z) = .
ξ1 (z) − iξ2 (z)
Show that f is holomorphic and that if g has a pole of order k at z0 ∈ U ,
then f has a zero of order 2k at z0 . (Hint: Clearly, zeros of f and poles of
g both occur at points z0 ∈ U where ξ1 (z0 ) − iξ2 (z0 ) = 0. Use the relation
in part (a) to relate the order of the pole of g at z0 to the order of the zero
of f at z0 .)
(c) Show that with f, g as in part (b), z has the form (8.15).
Recall that in the process of constructing adapted frame fields for elliptic
surfaces Σ ⊂ A3 , we found two invariant quadratic forms that we referred to
as the first and second equi-affine fundamental forms. These were defined
in terms of the Maurer-Cartan forms (ω̄ i , ω̄ji ) associated to any 2-adapted
frame field (e1 (u), e2 (u), e3 (u)) along Σ by
I = (ω̄ 1 )2 + (ω̄ 2 )2 ,
II = ω̄31 ω̄ 1 + ω̄32 ω̄ 2 .
These quadratic forms are equi-affine analogs of the Euclidean first and sec-
ond fundamental forms for an elliptic surface Σ in E3 , but despite the appar-
ent similarities, they are quite different from their Euclidean counterparts.
For instance, if x : U → A3 is a parametrization of Σ, then the coefficients of
8.3. Minimal surfaces in A3 269
the equi-affine first fundamental form are defined in terms of second deriva-
tives of x, whereas the Euclidean first fundamental form is defined in terms
of first derivatives of x. (In fact, the equi-affine first fundamental form of a
surface Σ is a scalar multiple of the Euclidean second fundamental form of
Σ; cf. Exercise 6.40(c).) Nevertheless, we will see that an equi-affine ana-
log of Theorem 8.16 is true: If we define an area functional for equi-affine
surfaces based on the equi-affine first fundamental form I, then the critical
points of this functional are precisely those surfaces for which the trace of
II with respect to I is identically zero.
dA = ω̄ 1 ∧ ω̄ 2
By analogy with the Euclidean case, we define the equi-affine area functional
A on the set S of closed and bounded elliptic surfaces Σ ⊂ A3 to be
&
A(Σ) = dA.
Σ
In order to look for critical points of the equi-affine area functional, we apply
the same ideas that we developed in the Euclidean case. The notion of a
compactly supported normal variation is precisely the same for surfaces in
A3 as for surfaces in E3 ; the only difference is that “normal” refers to the
equi-affine normal vector, which is well-defined for any 2-adapted frame field
on Σ. Thus, the equi-affine analog of Definition 8.6 is as follows:
*Exercise 8.34. Convince yourself that the results of Exercise 8.4 are
equally valid in the equi-affine case; the only change is that the unit normal
vector nt (u) to the surface Σt should be replaced with the equi-affine nor-
mal vector to Σt at the point xt (u). Therefore, as in the Euclidean case, it
suffices to consider normal variations of x.
*Exercise 8.35 (Cf. Exercise 8.9). (a) For each t ∈ (−ε, ε), let ıt : U →
U × (−ε, ε) denote the inclusion map defined by ıt (u) = (u, t). Show that
the pullbacks of the 1-forms (ω̄ 1 , ω̄ 2 ) to U via ıt are the usual dual forms on
the surface Σt = xt (U ).
(b) Show that ω̄ 3 = χ dt, where the function χ(u, t) is determined by the
condition that
∂X
e3 (u, t) ω̄ 3 = dt.
∂t
(c) Show that the pullbacks of the connection forms (ω̄ji ) to U via ıt are
the usual connection forms on the surface Σt = xt (U ). In particular, the
relations
1
ω̄3 11 12 ω̄ 1
(8.17) = ,
ω̄32 12 22 ω̄ 2
which hold for a 2-adapted frame field on an elliptic equi-affine surface, still
hold modulo dt for the corresponding forms on U × (−ε, ε). (This means,
e.g., that the 1-form
ω̄31 − (11 ω̄ 1 + 12 ω̄ 2 )
on U × (−ε, ε) is equal to a multiple of dt.)
8.3. Minimal surfaces in A3 271
*Exercise 8.36. Convince yourself that the results of Exercises 8.11 and
8.12 hold in the equi-affine case.
(b) Use the Cartan structure equations for a 2-adapted coframing, the equa-
tion ω̄ 3 = χ dt, and equation (8.17) to show that
d(ω̄ 1 ∧ ω̄ 2 ) = 2Lχω̄ 1 ∧ ω̄ 2 ∧ dt,
where L = 12 (11 + 22 ) is the equi-affine mean curvature of Σ. (Note that
you only need to know that equation (8.17) holds modulo dt for this com-
putation.)
(c) Use parts (a) and (b) to show that
& &
L∂/∂t (ω̄ ∧ ω̄ ) =
1 2
2Lχ ω̄ 1 ∧ ω̄ 2 .
U U
Together with Proposition 8.38, this proves the following equi-affine analog
of Theorem 8.16:
Exercise 8.41. (a) Show that for any values of a, b, c with ac − b2 > 0, the
elliptic paraboloid
z = ax2 + bxy + cy 2
is equi-affinely equivalent to the paraboloid Σ ⊂ A3 defined by the equation
(8.18) z = 12 (x2 + y 2 ).
(a) Let (e1 (u), e2 (u), e3 (u)) be the 0-adapted equi-affine frame field on Σ
given by
e1 (u, v) = xu = t [1, 0, fu ] ,
e2 (u, v) = xv = t[0, 1, fv ] ,
e3 (u, v) = t [0, 0, 1] .
Show that the dual forms associated to this frame field are
ω̄ 1 = du, ω̄ 2 = dv
and that the only nonzero connection forms are
ω̄13 = fuu du + fuv dv,
ω̄23 = fuv du + fvv dv.
Thus, we have
h11 h12 fuu fuv
= .
h12 h22 fuv fvv
Assume that fuu fvv −fuv 2 > 0, so that Σ is elliptic, and for simplicity assume
that fuu , fvv > 0. In order to compute the equi-affine mean curvature of
Σ, we need to construct a 2-adapted frame field on Σ and compute the
associated Maurer-Cartan forms. Recall from Chapter 6 that any other
0-adapted frame field (ẽ1 (u), ẽ2 (u), ẽ3 (u)) on Σ has the form
⎡ ⎤
r1
⎢ B ⎥
(8.20) ẽ1 (u) ẽ2 (u) ẽ3 (u) = e1 (u) e2 (u) e3 (u) ⎣ r2 ⎦
0 0 (det B)−1
for some GL(2)-valued function B and real-valued functions r1 , r2 on U and
that the Maurer-Cartan forms (ω̄ ˜ i , ω̄
˜ ji ) associated to this frame field satisfy
the relations
1 1 3 3
˜
ω̄ ω̄ ˜1
ω̄ ω̄1
(8.21) = B −1 , = (det B) tB .
˜2
ω̄ ω̄ 2 ˜ 23
ω̄ ω̄23
then
h̃11 h̃12 10
= ,
h̃12 h̃22 01
˜ 13 = ω̄
and therefore ω̄ ˜ 1 , ω̄
˜ 23 = ω̄
˜ 2.
(c) Show that for this frame field (with r1 , r2 still arbitrary), we have
" #
(fuu r1 + fuv r2 ) (fuu fuvv − 2fuv fuuv + fvv fuuu )
˜ 33 =
ω̄ + du
(fuu fvv − fuv
2 )1/4 4(fuu fvv − fuv
2 )
" #
(fuv r1 + fvv r2 ) (fuu fvvv − 2fuv fuvv + fvv fuuv )
+ + dv.
(fuu fvv − fuv )
2 1/4 4(fuu fvv − fuv
2 )
we can arrange that ω̄˜ 33 = 0 and hence (together with the result of part (b))
that the resulting frame field (ẽ1 (u), ẽ2 (u), ẽ3 (u)) is 2-adapted.
(d) Show that the dual forms associated to this frame field are
√
fuu fuv
˜ 1
ω̄ = du + √ dv,
(fuu fvv − fuv )
2 1/8 fuu (fuu fvv − fuv
2 )1/8
Now, suppose that Σ is equi-affine minimal. Let (e1 (u), e2 (u), e3 (u)) be any
2-adapted frame field on Σ, with associated Maurer-Cartan forms (ω̄ i , ω̄ji ),
and consider the scalar-valued, complex 1-form ω̄ 1 + iω̄ 2 on U .
*Exercise 8.44. Convince yourself that the results of Exercise 8.22 hold in
the equi-affine case. Thus, every point u ∈ U has a neighborhood V ⊂ U
on which there exist complex-valued functions z, ϕ : V → C such that
ω̄ 1 + iω̄ 2 = ϕ dz.
276 8. Minimal surfaces in E3 and A3
*Exercise 8.48. (a) Use the results of Exercise 8.45 to show that
for any choice of base point z0 ∈ V . This formula is called the Blaschke
representation for Σ.
(8.26) Xu = Yv , Xv = −Yu ,
Yuu = h0 Y + h1 Yu + h2 Yv ,
Yuv = k0 Y + k1 Yu + k2 Yv ,
Yvv = −h0 Y − h1 Yu − h2 Yv
Conclude that the frame field (8.28) is 2-adapted. (Hint: For the last con-
dition, first show that
Then use the assumption that the frame field is unimodular to conclude that
ω̄33 = 0.)
280 8. Minimal surfaces in E3 and A3
symbolic);
e2:= Xv/simplify(Norm(Xv, Euclidean, conjugate=false),
symbolic);
e3:= simplify(CrossProduct(e1, e2));
We can use the following substitution to go back and forth between the
(du, dv) basis and the (ω̄ 1 , ω̄ 2 ) basis as necessary (but first we need to declare
the 1-forms (ω̄ 1 , ω̄ 2 )):
> Form(omega[1], omega[2]);
> framesub:= [
omega[1] = simplify(Norm(Xu, Euclidean, conjugate=false),
symbolic)*d(u),
omega[2] = simplify(Norm(Xv, Euclidean, conjugate=false),
symbolic)*d(v)];
> framebacksub:= makebacksub(framesub);
In order to compute the (hij ), we need to express de3 as a linear combination
of (e1 (u), e2 (u)). We know from the Cartan structure equations (3.1) that
the coefficients of (e1 (u), e2 (u)) will be the 1-forms ω̄31 = −ω̄13 , ω̄32 = −ω̄23 ,
respectively.
> de3:= map(Simf, subs(framebacksub, map(d, e3)));
> zero1:= Simf(de3 + (omega[3,1]*e1 + omega[3,2]*e2));
> Simf(solve({zero1[1], zero1[2], zero1[3]},
{omega[3,1], omega[3,2]}));
1
ω1 ρv,v ω2 1 + ρ2v
ω3,1 = − , ω3,2 =
1 + ρ2v ρ 1 + 2 ρ2v + ρ4v
> assign(%);
We can read off the (hij ) directly from ω̄13 and ω̄23 :
> h:= Matrix([
[pick(omega[3,1], omega[1]), pick(omega[3,1], omega[2])],
[pick(omega[3,2], omega[1]), pick(omega[3,2], omega[2])]]);
Finally, Σ is minimal if and only if H = 12 (h11 + h22 ) = 0:
> minsurfeq:= numer(simplify(Trace(h)));
Now, Maple can solve this equation with the dsolve command. Unfor-
tunately, Maple can be rather clumsy about simplifying exponentials and
282 8. Minimal surfaces in E3 and A3
" !2 !2 #
v C2
C1 e C1 e C1 +1
1
f= v C2
2 e C1 e C1
Maple apparently doesn’t even realize that these two solutions are, in fact,
the same! But we can verify that our eyes are not deceiving us on this point:
> Simf(subs(soln[1], rho(v)) - subs(soln[2], rho(v)));
0
So, we can choose either solution and perform some gymnastics to force
Maple to put it into a nicer form. (Try unpacking this command and
applying these operations one at a time to see the intermediate steps.)
> Simf(convert(combine(expand(Simf(subs(soln[1], rho(v)))),
exp), trig));
" #
v + C2
C1 cosh
C1
The plot3d command can be used to plot the surface over various parameter
ranges; e.g.,
> plot3d(X0, u=-2..2, v=-2..2, axes = normal,
scaling=constrained);
Since we have e1 (u) = xu (u), e2 (u) = xv (u), it follows that the associated
Maurer-Cartan forms ω̄ 1 , ω̄ 2 are equal to
ω̄ 1 = du, ω̄ 2 = dv.
Therefore, we have
ω̄13 = ω̄ 1 , ω̄23 = ω̄ 2 ,
and the remaining connection forms (ω̄ji ) are equal to zero. It follows that
this frame field is 2-adapted. Moreover, since ω̄31 = ω̄32 = 0, Σ is an equi-
affine minimal surface.
Exercise 8.42: Define the parametrization for Σ and the 0-adapted frame
field of part (a) (we’ll call this frame field (e10, e20, e30) because it will
be refined later):
> PDETools[declare](f(u,v));
> X:= Vector([u, v, f(u,v)]);
> e10:= map(diff, X, u);
e20:= map(diff, X, v);
e30:= Vector([0,0,1]);
As in the previous exercise, for this frame field we have ω̄ 1 = du, ω̄ 2 = dv:
> framesub0:= [omega[1] = d(u), omega[2] = d(v)];
> framebacksub0:= makebacksub(framesub0);
In order to compute the connection forms (ω̄ji ), differentiate the frame fields:
> de10:= map(Simf, subs(framebacksub0, map(d, e10)));
de20:= map(Simf, subs(framebacksub0, map(d, e20)));
de30:= map(Simf, subs(framebacksub0, map(d, e30)));
8.4. Maple computations 285
- diff(f(u,v),u,v)ˆ2)ˆ(3/8));
b[2,2]:= sqrt(diff(f(u,v),u,u))/
(diff(f(u,v),u,u)*diff(f(u,v),v,v)
- diff(f(u,v),u,v)ˆ2)ˆ(3/8);
Set up a substitution for the Maurer-Cartan forms for this new frame field:
> framesub:= [omega[1] = Simf(dualformsvec[1]),
omega[2] = Simf(dualformsvec[2])];
> framebacksub:= makebacksub(framesub);
We still need to solve for r1 , r2 . This requires computing dẽ3 so that we can
˜ 33 and set it equal to zero. We can do this fairly compactly as follows:
find ω̄
If we let A be the matrix
A = ẽ1 (u) ẽ2 (u) ẽ3 (u) ,
then we have ⎡ ⎤
˜ 31
ω̄
⎢ ˜ 2⎥
dẽ3 = A ⎣ω̄ 3⎦ .
˜ 33
ω̄
Therefore, ω̄˜ 33 is the last entry of the vector A−1 dẽ3 , and we can solve for
r1 , r2 by setting the scalar coefficients of ω̄˜ 33 equal to zero.
Pseudospherical
surfaces and
Bäcklund’s theorem
9.1. Introduction
In this chapter, we will show how moving frames may be used to prove
Bäcklund’s theorem. Bäcklund’s theorem concerns surfaces of constant neg-
ative Gauss curvature, also known as pseudospherical surfaces. The best-
known pseudospherical surface is, of course, the pseudosphere, which is the
surface of revolution obtained by revolving the tractrix
about the x-axis. But there are infinitely many other pseudospherical sur-
faces as well.
Bäcklund’s theorem is based on a beautiful geometric construction that
starts with a given pseudospherical surface and produces from it a 2-pa-
rameter family of new pseudospherical surfaces. (See [RS82] for a discus-
sion of Bäcklund’s original construction.) The construction can be iterated,
thereby producing an arbitrary number of increasingly complicated fami-
lies of pseudospherical surfaces from a single starting surface. For example,
we might take the pseudosphere as our initial surface and use Bäcklund’s
construction to generate new families of surfaces.
287
288 9. Pseudospherical surfaces and Bäcklund’s theorem
Line congruences were the object of much study in the nineteenth century;
for a thorough treatment, see [Eis60].
Example 9.3. Let U be an open set in R2 , and let x : U → E3 be an
immersion whose image is a regular surface Σ ⊂ E3 . For each u ∈ U , let
(u) denote the line in E3 passing through the point x(u) and parallel to
the normal vector e3 (u). Then the collection
{(u) | u ∈ U }
is a line congruence in E3 . A line congruence of this type is called a normal
congruence.
The construction in part (2) of Theorem 9.5 is due to Bianchi and Bäcklund
([B8̈3], [Bia79]) and is called a Bäcklund transformation; the terminology
8 is a “transformation” of the original
refers to the idea that the new surface Σ
surface Σ.
Bäcklund’s theorem can be proved using local coordinates on the surfaces
8 but the computations are rather ugly. The proof can be greatly sim-
Σ, Σ,
plified by using the method of moving frames because frame fields can be
adapted to the geometry of the problem in a way that local coordinates can-
not. Whereas in previous chapters we have adapted our frames according to
the geometry of a single surface, here we have to consider two surfaces and
the geometric conditions relating them. We will use these considerations to
guide our choices of orthonormal frame fields on the surfaces Σ, Σ.8 (This
proof is taken from [CT80].)
8 respectively,
Proof. (1) Let x, x̃ : U → E3 be the parametrizations of Σ, Σ,
induced from the line congruence : U → G1 (E ) as above. We can choose
3
orthonormal frame fields (e1 (u), e2 (u), e3 (u)) along Σ and (ẽ1 (u), ẽ2 (u),
8 such that:
ẽ3 (u)) along Σ
(1) e3 (u) is the unit normal vector to Σ at x(u) and ẽ3 (u) is the unit
normal vector to Σ 8 at x̃(u); therefore, (e (u), e (u)) span the tan-
1 2
8
gent space Tx(u) Σ and (ẽ1 (u), ẽ2 (u)) span the tangent space Tx̃(u) Σ.
9.3. Bäcklund’s theorem 291
(2) e1 (u) = ẽ1 (u) is the common unit tangent vector to both surfaces
in the direction of x̃(u) − x(u).
Remark 9.6. The notation ei (u) is intended to distinguish these orthonor-
mal frame vectors from the principal orthonormal frame vectors ei (u) that
will be introduced in §9.4. This notational distinction will become impor-
tant in §9.5, when we need to consider both of these orthonormal frame
fields simultaneously! Similarly, we will denote the Maurer-Cartan forms
associated to the frame field (e1 (u), e2 (u), e3 (u)) by (ω̄ i , ω̄ ij ), and those
˜ i , ω̄
associated to the frame field (ẽ1 (u), ẽ2 (u), ẽ3 (u)) by (ω̄ ˜ ji ).
*Exercise 9.7. Show that it follows from these conditions (and the defini-
tion of α as the angle between e3 (u), ẽ3 (u)) that
ẽ1 (u) = e1 (u),
(9.2) ẽ2 (u) = cos(α)e2 (u) + sin(α)e3 (u),
ẽ3 (u) = − sin(α)e2 (u) + cos(α)e3 (u)
and that
(9.3) x̃(u) = x(u) + re1 (u).
*Exercise 9.8. (a) Show that taking the exterior derivative of equation
(9.3) and applying the Cartan structure equations (3.1) yields
(9.4) ˜ 1 + ẽ2 ω̄
ẽ1 ω̄ ˜ 2 = e1 ω̄ 1 + e2 ω̄ 2 + r(e2 ω̄ 21 + e3 ω̄ 31 ).
Then substitute the expressions (9.2) for (ẽ1 (u), ẽ2 (u)) into equation (9.4)
to obtain
˜ 1 + e2 cos(α) ω̄
e1 ω̄ ˜ 2 + e3 sin(α) ω̄
˜ 2 = e1 ω̄ 1 + e2 (ω̄ 2 + r ω̄ 21 ) + e3 (r ω̄ 31 ).
Conclude that we have the following relationships between the Maurer-
8
Cartan forms on Σ and those on Σ:
˜ 1 = ω̄ 1 ,
ω̄
(9.5) ˜ 2 = ω̄ 2 + r ω̄ 21 ,
cos(α) ω̄
˜ 2 = r ω̄ 31 .
sin(α) ω̄
(b) Show that the last two equations in (9.5) imply that
(9.6) ω̄ 2 + r ω̄ 21 = r cot(α) ω̄ 31 .
292 9. Pseudospherical surfaces and Bäcklund’s theorem
Next, recall that the coefficients (hij ) of the second fundamental form of Σ
are defined by the equations
ω̄ 31 = h11 ω̄ 1 + h12 ω̄ 2 ,
(9.8)
ω̄ 32 = h12 ω̄ 1 + h22 ω̄ 2 .
Because we have not made any attempt to arrange for e1 (u) and e2 (u) to
be principal vector fields, we should not expect to have h12 = 0; in fact, the
following exercise shows that h12 cannot be equal to zero.
*Exercise 9.9. Use the first and third equations in (9.5), the first equation
˜ 1 , ω̄
in (9.8), and the fact that (ω̄ ˜ 2 ) are linearly independent 1-forms on U to
conclude that h12 = 0.
8 Recall that
Finally, consider the Gauss curvature K̃ of Σ.
(9.9) ˜ 13 ∧ ω̄
ω̄ ˜ 1 ∧ ω̄
˜ 23 = K̃ ω̄ ˜2
(cf. Exericse 4.47).
*Exercise 9.10. (a) Use equations (9.7) and (9.8) to show that
sin(α)
˜ 13 ∧ ω̄
ω̄ ˜ 23 = − h12 ω̄ 1 ∧ ω̄ 2 .
r
(b) Use equation (9.9), the first and third equations of (9.5), and (9.8) to
show that
r
˜ 13 ∧ ω̄
ω̄ ˜ 23 = K̃ h12 ω̄ 1 ∧ ω̄ 2 .
sin(α)
(c) Conclude from parts (a) and (b) and the fact that h12 = 0 that
sin2 (α)
K̃ = − .
r2
2
An analogous argument shows that K = − sinr2(α) as well.
The proof of part (2) of Theorem 9.5 involves concepts from the theory of
exterior differential systems similar to those needed for the proof of Lemma
4.12. The key step involves constructing an adapted orthonormal frame
field (e1 (u), e2 (u), e3 (u)) along Σ with the property that a parametrization
x̃ : U → E3 for the desired surface Σ 8 will be given by
(9.10) x̃(u) = x(u) + re1 (u).
9.4. Pseudospherical surfaces and the sine-Gordon equation 293
The Maurer-Cartan forms for such a frame field must satisfy equation (9.6);
in fact (as we will see in §9.5), equation (9.6) is equivalent to an overde-
termined system of partial differential equations for the frame field (e1 (u),
e2 (u), e3 (u)), and the compatibility condition for this system is precisely the
2
condition that Σ has Gauss curvature K = − sinr2(α) . The initial condition
e1 (u0 ) = e0 ∈ Tx0 Σ then determines the desired frame field uniquely, and
the remainder of the proof consists of showing that the surface Σ 8 defined by
(9.10) satisfies all the desired conditions.
The proof of part (2) of Theorem 9.5 shows how, given a surface of constant
negative Gauss curvature K, one can construct new pseudospherical surfaces
8 The 2-parameter family of such surfaces alluded to earlier arises as
Σ.
follows: One parameter comes from the choice of constants r, α such that
2
K = − sinr2(α) , and one comes from the choice of a non-principal unit vector
e0 ∈ Tx0 Σ. The choice of r, α determines the coefficients of the PDE system
(9.6), while the choice of e0 determines the initial conditions that give rise
to a particular solution of this PDE system.
Let Σ be a pseudospherical surface, and for simplicity assume that its Gauss
curvature is K = −1. Since the Gauss curvature of Σ is negative, Σ cannot
have any umbilic points; consequently, it can be shown (for a proof, see
[dC76]) that every point x ∈ Σ has a neighborhood for which there exists a
local parametrization x : U → E3 of Σ whose coordinate curves are principal
curves in Σ (cf. Exercises 4.24 and 4.27). As in Exercise 4.24, we can then
choose the adapted orthonormal frame field
1 1
e1 (u) = √ xu , e2 (u) = √ xv , e3 (u) = e1 (u) × e2 (u)
E G
along Σ. (Note that, unlike in §9.3, we are only considering a single pseu-
dospherical surface, so there is no line congruence to take into consideration
when choosing an adapted frame field.) Then, as we saw in Exercises 4.24
and 4.27, the coefficients of the first and second fundamental forms satisfy
F = f = 0, and the associated Maurer-Cartan forms (ω̄ i , ω̄ji ) are given by
√ √
ω̄ 1 = E du, ω̄ 2 = G dv,
e √ g √
ω̄13 = √ du = κ1 E du, ω̄23 = √ dv = κ2 G dv,
(9.11) E G
1
ω̄21 = √ (Ev du − Gu dv),
2 EG
where κ1 = Ee , κ2 = Gg are the principal curvatures of Σ.
294 9. Pseudospherical surfaces and Bäcklund’s theorem
The following exercises will show how the surface Σ = x(U ) gives rise to a
solution of the sine-Gordon equation (9.1). First, we investigate the Gauss
and Codazzi equations for Σ and show that the parametrization x : U → E3
can be fine-tuned to arrange that the first and second fundamental forms of
Σ can be expressed nicely in terms of a single function ψ : U → R.
*Exercise 9.11. (a) Show that the Codazzi equations of Exercise 4.41(f)
can be written in the form
∂κ1 ∂ √ ! ∂κ2 ∂ √ !
(9.12) = (κ2 − κ1 ) ln( E) , = (κ1 − κ2 ) ln( G) .
∂v ∂v ∂u ∂u
(b) Divide equations (9.12) by (κ1 − κ2 ), multiply the left-hand sides by κκ11
and κκ22 , respectively, and use the Gauss equation κ1 κ2 = −1 to show that
∂ ∂ √ !
ln(κ21 + 1) = −2 ln( E) ,
∂v ∂v
(9.13)
∂ ∂ √ !
ln(κ22 + 1) = −2 ln( G) .
∂u ∂u
(e) Now, assume that the coordinate functions (u, v) (without the tildes)
have been chosen so that c1 (u) = c2 (v) = 1. Show that there exists a
function ψ(u, v) such that
κ1 = tan(ψ), κ2 = − cot(ψ),
2
E = cos (ψ), G = sin2 (ψ).
9.4. Pseudospherical surfaces and the sine-Gordon equation 295
(d) Let φ = 2ψ, so that φ(u, v) is the angle between the asymptotic directions
at the point x(u, v). Show that equation (9.16) is equivalent to the PDE
(9.17) φuu − φvv = sin(φ)
for the function φ.
(e) Consider the change of coordinates
1 1
x = (u + v), y = (u − v).
2 2
Show that in terms of the (x, y)-coordinates, the first and second fundamen-
tal forms of Σ are given by
I = dx2 + 2 cos(2ψ) dx dy + dy 2 ,
(9.18)
II = 2 sin(2ψ) dx dy.
Note that the x- and y-coordinate directions are now the asymptotic direc-
tions at each point of Σ; for this reason, (x, y) are called asymptotic coordi-
nates on Σ, and the corresponding parametrization is called an asymptotic
parametrization of Σ.
(f) Show that equation (9.17) is equivalent to equation (9.1).
296 9. Pseudospherical surfaces and Bäcklund’s theorem
Remark 9.13. Equations (9.17) and (9.1) are both referred to as the sine-
Gordon equation. The local coordinates (u, v) of equation (9.17) are called
space-time coordinates because the left-hand side has the same form as the
wave equation
φuu − φvv = 0,
where u is often thought of as a time coordinate and v as a spatial coordinate.
The local coordinates (x, y) of equation (9.1), on the other hand, are called
null or characteristic coordinates. The term “characteristic” comes from the
fact that the x- and y-coordinate curves are the characteristic curves for the
PDE (9.1), while the term “null” arises from thinking of the (u, v)-plane as
the Minkowski space M1,1 with its standard metric, for which the x- and
y-coordinate curves are the null lines.
Remark 9.15. In recent years, many other integrable systems have been
shown to be connected with pseudospherical geometry, and this connection
9.5. Bäcklund transformation for the sine-Gordon equation 297
provides an important tool for studying the space of solutions to these equa-
tions. This connection was first considered by Chern and Tenenblat [CT86]
and further explored by Reyes [Rey98] and others.
In this section, we will see how the geometric Bäcklund transformation be-
tween pseudospherical surfaces gives rise to a corresponding analytic trans-
formation between solutions of the sine-Gordon equation (9.1).
Suppose that we have a Bäcklund transformation between two pseudospher-
ical surfaces Σ, Σ8 of Gauss curvature K = −1. (Note that the condition
K = −1 implies that r = sin(α).) Let (e1 (u), e2 (u), e3 (u)) be the orthonor-
mal frame field on Σ adapted to the Bäcklund transformation as in §9.3,
and let (e1 (u), e2 (u), e3 (u)) be the principal adapted frame field on Σ as in
§9.4.
Let η(u) denote the angle between e1 (u) and e1 (u). The following exercise
shows how the function η is related to the function ψ of §9.4.
*Exercise 9.17. (a) Suppose that the pair of functions (ψ(x, y), η(x, y))
satisfies the PDE system (9.22), where λ is any nonzero constant. Show
that the functions 2ψ, 2η must each be solutions of the sine-Gordon equa-
tion (9.1). (Hint: Differentiate the first equation in (9.22) with respect to y
and the second equation with respect to x; then add and subtract the result-
ing equations and apply trigonometric identities to simplify the right-hand
sides.)
(b) Now, suppose that 2ψ is any known solution of (9.1). Then the system
(9.22) can be regarded as the overdetermined PDE system
ηx = −ψx + λ sin(ψ − η),
1
ηy = ψy − sin(ψ + η)
λ
for the unknown function η(x, y). Show that this system is compatible, i.e.,
that (ηx )y = (ηy )x —precisely because 2ψ satisfies (9.1). It follows that this
system has a 1-parameter family of solutions η(x, y) and that these solutions
can be constructed using only techniques of ordinary differential equations.
The PDE system (9.22) is called a Bäcklund transformation for the sine-
Gordon equation (9.1); the construction in Exercise 9.17 is the analog of
part (2) of Theorem 9.5. It shows how, given one solution 2ψ of the sine-
Gordon equation (9.1), the PDE system (9.22) can be used to construct
a 2-parameter family of new solutions 2η: One parameter comes from the
choice of the constant λ = 0, and one comes from the 1-parameter family of
solutions η to the system (9.22).
At this point, we have established several relationships involving pseudo-
spherical surfaces and solutions of the sine-Gordon equation:
8 are
(b) Show that the first and second fundamental forms of Σ
˜ 1 )2 + (ω̄
I = (ω̄ ˜ 2 )2 = dx2 + 2 cos(2η) dx dy + dy 2 ,
(9.24)
˜ 13 ω̄
II = ω̄ ˜ 1 + ω̄ ˜ 2 = −2 sin(2η) dx dy.
˜ 23 ω̄
(c) Conclude that the angle between the asymptotic directions of Σ 8 = x̃(U )
at the point x̃(x, y) is equal to 2η(x, y). (The sign of the second fundamental
form is unimportant, as it can be reversed by reversing the orientation of
8
Σ.)
The result of Exercise 9.18 completes the picture: The analytic Bäcklund
transformation (9.22) between solutions of the sine-Gordon equation (9.1)
is the precise analog of the geometric Bäcklund transformation between
pseudospherical surfaces of Gauss curvature K = −1.
Exercise 9.19. Let 2ψ(x, y) ≡ 0 be the trivial solution of the sine-Gordon
equation (9.1). (This solution corresponds to the degenerate “surface” con-
sisting of a straight line in E3 .) Show that the corresponding solutions
2η(x, y) generated by the Bäcklund transformation (9.22) are given by
η(x, y) = 2 tan−1 (Ce−(λx+ λ y) ),
1
300 9. Pseudospherical surfaces and Bäcklund’s theorem
where C = 0 is constant. (Hint: You may find the trig identity csc(η) +
cot(η) = cot( 12 η) useful.) The functions
2η = 4 tan−1 (Ce−(λx+ λ y) )
1
(9.25)
are called the 1-soliton solutions of the sine-Gordon equation. Iterating this
procedure gives the 2-solitons, etc.
(a) Show that the first and second fundamental forms of Σ are given by
2(cosh2 (x + y) − 2)
I = dx2 + 2 dx dy + dy 2 ,
cosh (x + y)
4 sinh(x + y)
II = dx dy.
cosh2 (x + y)
cosh2 (x + y) − 2 2 sinh(x + y)
(9.26) cos(2ψ(x, y)) = 2 , sin(2ψ(x, y)) = .
cosh (x + y) cosh2 (x + y)
Use the sine-Gordon equation (9.1) and the Fundamental Theorem of Cal-
culus to show that the area of x(R) ⊂ Σ is given by
&
A(x(R)) = ω̄ 1 ∧ ω̄ 2
R
= 2 (ψ(b, d) − ψ(b, c) − ψ(a, d) + ψ(a, c)) .
Conclude from part (b) that A(x(R)) < 2π.
(d) Observe that, because H2 is complete and has infinite area, the area of R
can be made arbitrarily large by choosing a, b, c, d appropriately. Therefore,
since x is an isometric immersion, the area of x(R) can be made arbitrarily
large as well. This contradicts the result of part (c); thus, no such isometric
immersion can exist.
Exercise 9.22. While the Bäcklund transformation (9.22) relates two dif-
ferent solutions of the same PDE (9.1), it is also possible for a Bäcklund
transformation to relate solutions of two different PDEs. For example, con-
sider the first-order system of partial differential equations
(ψ−η)
ψx + ηx = 2λe 2 ,
(9.27) 1 (ψ+η)
ψy − ηy = e 2 .
λ
(a) Suppose that the pair of functions (ψ(x, y), η(x, y)) satisfies the PDE
system (9.27), where λ is any nonzero constant. Show that η(x, y) must be
a solution of the wave equation (in characteristic coordinates)
(9.28) ηxy = 0,
while ψ(x, y) must be a solution of Liouville’s equation
(9.29) ψxy = eψ .
(b) Show by integration that the general solution of the wave equation (9.28)
is
(9.30) η(x, y) = ρ(x) + σ(y),
where ρ(x), σ(y) are arbitrary functions of a single variable.
(c) Substitute equation (9.30) into the system (9.27) to obtain a system of
two first-order PDEs for the function ψ(x, y). By treating each of these
equations as a separable ODE in the appropriate variable, show that
" & & #
− 21 ψ 1
−ρ(x) 1
(9.31) e = −e 2 (ρ(x)−σ(y))
λ e dx + e σ(y)
dy .
2λ
9.6. Maple computations 303
theta1[1,1]:= 0;
theta1[2,2]:= 0;
theta1[3,3]:= 0;
theta1[2,1]:= -theta1[1,2];
theta1[1,3]:= -theta1[3,1];
theta1[2,3]:= -theta1[3,2];
theta2[1,1]:= 0;
theta2[2,2]:= 0;
theta2[3,3]:= 0;
theta2[2,1]:= -theta2[1,2];
theta2[1,3]:= -theta2[3,1];
theta2[2,3]:= -theta2[3,2];
> for i from 1 to 3 do
d(omega[i]):= -add(’omega[i,j] &ˆ omega[j]’, j=1..3);
end do;
d(omega[1,2]):= -add(’omega[1,k] &ˆ omega[k,2]’, k=1..3);
d(omega[3,1]):= -add(’omega[3,k] &ˆ omega[k,1]’, k=1..3);
d(omega[3,2]):= -add(’omega[3,k] &ˆ omega[k,2]’, k=1..3);
for i from 1 to 3 do
d(theta1[i]):= -add(’theta1[i,j] &ˆ theta1[j]’, j=1..3);
end do;
d(theta1[1,2]):= -add(’theta1[1,k] &ˆ theta1[k,2]’, k=1..3);
d(theta1[3,1]):= -add(’theta1[3,k] &ˆ theta1[k,1]’, k=1..3);
d(theta1[3,2]):= -add(’theta1[3,k] &ˆ theta1[k,2]’, k=1..3);
for i from 1 to 3 do
d(theta2[i]):= -add(’theta2[i,j] &ˆ theta2[j]’, j=1..3);
end do;
d(theta2[1,2]):= -add(’theta2[1,k] &ˆ theta2[k,2]’, k=1..3);
d(theta2[3,1]):= -add(’theta2[3,k] &ˆ theta2[k,1]’, k=1..3);
d(theta2[3,2]):= -add(’theta2[3,k] &ˆ theta2[k,2]’, k=1..3);
We’ll also need the vector form of the structure equations; we’ll use x1 and
8 respectively, and (e11, e12, e13),
x2 for the parametrizations of Σ and Σ,
(e21, e22, e23) for the frame fields on the two surfaces that are adapted
to the Bäcklund transformation.
> d(x1):= e11*theta1[1] + e12*theta1[2];
d(e11):= e12*theta1[2,1] + e13*theta1[3,1];
d(e12):= e11*theta1[1,2] + e13*theta1[3,2];
d(e13):= e11*theta1[1,3] + e12*theta1[2,3];
9.6. Maple computations 305
˜ 13 = dẽ1 , ẽ3 ,
ω̄ ˜ 23 = dẽ2 , ẽ3 .
ω̄
Since we haven’t told Maple that (e11, e12, e13) are vectors, we’ll have
to define our own procedure to compute these inner products. We can do
this as follows:
> innprod1:= proc(exp1, exp2)
RETURN(coeff(exp1, e11)*coeff(exp2, e11)
+ coeff(exp1, e12)*coeff(exp2, e12)
306 9. Pseudospherical surfaces and Bäcklund’s theorem
In order to verify that the structure equations are satisfied, check that com-
putations such as the following all yield zero:
> Simf(d(Simf(subs(SGEformsub, omega[1]))))
- Simf(subs(SGEformsub, Simf(d(omega[1]))));
The only one that doesn’t immediately reduce to zero is the one for ω̄21 ,
which requires an application of SGEsub in order to see that it vanishes.
Exercise 9.16: Equation (9.20) is immediate from the definition of η(u),
and the results of Chapter 4 yield the expressions for (ω̄ 1 , ω̄ 2 , ω̄ 31 , ω̄ 32 ) in
equations (9.21). Set up a substitution to go back and forth between the
two sets of Maurer-Cartan forms:
> PDETools[declare](eta(x,y));
> rotationsub:= [
theta1[1] = cos(eta(x,y))*omega[1] + sin(eta(x,y))*omega[2],
theta1[2] = -sin(eta(x,y))*omega[1]
+ cos(eta(x,y))*omega[2],
theta1[3,1] = cos(eta(x,y))*omega[3,1]
+ sin(eta(x,y))*omega[3,2],
theta1[3,2] = -sin(eta(x,y))*omega[3,1]
+ cos(eta(x,y))*omega[3,2]];
> rotationbacksub:= makebacksub(rotationsub);
We can then compute ω̄ 12 in terms of ω̄21 as follows: Compute dω̄ 1 and dω̄ 2
by differentiating the expressions for ω̄ 1 and ω̄ 2 in equation (9.21), and then
use the reverse substitution to express the result in terms of (ω̄ 1 , ω̄ 2 ):
> dtheta1[1]:= Simf(subs([omega[3]=0],
Simf(subs(rotationbacksub, Simf(d(Simf(subs(rotationsub,
theta1[1]))))))));
> pick(dtheta1[1], theta1[2]);
−ω1,2 + ηx d(x) + ηy d(y)
> dtheta1[2]:= Simf(subs([omega[3]=0],
Simf(subs(rotationbacksub, Simf(d(Simf(subs(rotationsub,
theta1[2]))))))));
> pick(dtheta1[2], theta1[1]);
ω1,2 − ηx d(x) − ηy d(y)
It follows from the structure equations for dω̄ 1 and dω̄ 2 that ω̄ 12 = ω̄21 − dη.
Add this to our substitution:
> rotationsub:= [op(rotationsub),
theta1[1,2] = omega[1,2] - d(eta(x,y))];
308 9. Pseudospherical surfaces and Bäcklund’s theorem
We can combine this with the expressions for (ω̄ i , ω̄ji ) in SGEformsub in order
to write (ω̄ i , ω̄ ij ) in terms of (dx, dy):
> SGErotationsub:= Simf(subs(SGEformsub, rotationsub));
Since K = −1, we have r = sin(α):
> r:= sin(alpha);
Now consider the Bäcklund equation (9.6), which says that the following
expression is zero:
> Backlundeq:= theta1[2] + r*theta1[2,1]
- r*cot(alpha)*theta1[3,1];
We can express this equation in terms of (dx, dy) via SGErotationsub:
> Backlundzero:= Simf(subs(SGErotationsub, Backlundeq));
The coefficients of dx and dy in this expression are PDEs involving the
functions ψ and η:
> PDE1:= pick(Backlundzero, d(x));
PDE2:= pick(Backlundzero, d(y));
These PDEs contain some coefficients involving trigonometric functions of
α, and they will look nicer if we rename them. In the first equation, divide
by sin(α) and set λ = cot(α) − csc(α) via the substitution:
> lambdasub:= [cos(alpha) = lambda*sin(alpha) + 1];
> Simf(subs(lambdasub, PDE1/sin(alpha)));
λ sin(η − ψ) + ψx + ηx
Similarly, for the second equation divide by sin(α) and set μ = −(cot(α) +
csc(α)):
> musub:= [cos(alpha) = -mu*sin(alpha) - 1];
> Simf(subs(musub, PDE2/sin(alpha)));
μ sin(η + ψ) − ψy + ηy
9.6. Maple computations 309
In this chapter, we will see how moving frames may be used to prove two
classical results in differential geometry: the classification of doubly ruled
surfaces in R3 and the Cauchy-Crofton formula for the length of a curve in
the Euclidean plane E2 .
311
312 10. Two classical theorems
You may have noticed that so far in this chapter, we have been describing
surfaces as subsets of R3 without specifying any particular homogeneous
10.1. Doubly ruled surfaces in R3 313
Based on the result of Exercise 10.2, for the remainder of this section, we
will regard R3 as the equi-affine space A3 .
The goal of this section is to prove the following theorem [HCV52]:
Theorem 10.3. Any non-planar doubly ruled regular surface Σ ⊂ A3 is
equivalent via an equi-affine transformation to (an open subset of ) either a
hyperboloid of one sheet as in (10.1) or a hyperbolic paraboloid as in (10.2).
Remark 10.4. We will use the theory developed in Chapter 6 to prove
Theorem 10.3 as stated, but in fact more is true:
(1) The set of doubly ruled surfaces is invariant under the action of
the full affine group without the equi-affine restriction. Under this
group action, the hyperboloids of one sheet (10.1) for all positive
values of r are all equivalent, whereas different values of r give
surfaces that are inequivalent under the action of the equi-affine
group.
(2) If we regard R3 as an open subset of P3 via the identification
(x1 , x2 , x3 ) ↔ 1 : x1 : x2 : x3 ,
the set of doubly ruled surfaces is also invariant under the action
of the group of projective transformations on P3 . Under this group
action, the surfaces (10.1) and (10.2) are all equivalent to each
other.
(e1 (u), e2 (u), e3 (u)) along Σ will be called 0-adapted if for each u ∈ U ,
the vectors e1 (u) and e2 (u) are each tangent to one of the rulings passing
through the point x(u). (Note that this notion of “0-adapted” is more
restrictive than that in Chapter 6 because here we have additional geometric
information to guide our initial choice for an adapted frame field.)
and that the pullbacks (ω̄ i , ω̄ji ) to U of the Maurer-Cartan forms (ω i , ωji ) on
A(3) via the lifting x̃ satisfy the conditions that ω̄ 3 = 0 and
dx = e1 ω̄ 1 + e2 ω̄ 2 ,
(10.3)
dei = e1 ω̄i1 + e2 ω̄i2 + e3 ω̄i3 , i = 1, 2, 3.
*Exercise 10.6. (a) Show that the condition that the vector fields (e1 (u),
e2 (u)) are tangent to the rulings implies that
(Hint: Recall that de1 (e1 ) computes the directional derivative of the vector
field e1 (u) in the direction of e1 (u). What limitations are imposed on this
derivative by the condition that the vector field e1 (u) is tangent to a straight
line along this direction?)
(b) Show that the condition (10.4) implies that
In order to minimize notational clutter, we will rename the (akij ) and write
ω̄12 = a ω̄ 2 , ω̄13 = c ω̄ 2 ,
(10.7)
ω̄21 = b ω̄ 1 , ω̄23 = c ω̄ 1 .
0 0 λ11λ2
for some functions λ1 , λ2 , r1 , r2 on U , with λ1 , λ2 = 0.
(b) Show that under a transformation of the form (10.8), we have
1 1 1 3 2
˜
ω̄ λ1 ω̄
˜1
ω̄ λ1 λ2 ω̄13
(10.9) = , = .
˜2 1 2 ˜ 23
ω̄ λ2 ω̄ ω̄ λ1 λ22 ω̄23
Conclude that the transformed function c̃ defined by the conditions
˜ 13 = c̃ ω̄
ω̄ ˜ 2, ˜ 23 = c̃ ω̄
ω̄ ˜1
is given by
(10.10) c̃ = λ21 λ22 c.
0 0 ±1
for some functions λ, r1 , r2 on U , with λ = 0.
(b) Note that the condition c = 1 is equivalent to the relations
(10.12) ω̄13 = ω̄ 2 , ω̄23 = ω̄ 1
among the associated Maurer-Cartan forms. Compare with Exercise 6.42(a),
and conclude that any 1-adapted frame field along Σ satisfies the defining
conditions for what we called a “1-adapted null frame” there. In particular,
Σ must be a hyperbolic equi-affine surface; otherwise no such frame field
could exist along Σ.
(c) Apply the result of Exercise 6.42(c) to show that there exists a choice of
1-adapted frame field along Σ for which ω̄33 = 0. (This condition corresponds
10.1. Doubly ruled surfaces in R3 317
0 0 ±1
*Exercise 10.11. Let (e1 (u), e2 (u), e3 (u)) be any 2-adapted frame field
along Σ.
(a) Apply the results of Exercises 6.42(e) and 6.42(f) and equations (10.7)
to show that
a = b = 0.
df ∧ ω̄ 1 = df ∧ ω̄ 2 = 0.
Once again, we need to divide into cases based on whether or not C is zero.
Case 1: C = 0.
*Exercise 10.12. Suppose that C = 0.
(a) Show that de3 = 0, and conclude that the vector field e3 (u) is constant
on U . (Note that this implies that Σ is an improper equi-affine sphere; cf.
Exercise 6.41.)
(b) Recall that the Maurer-Cartan forms associated to a 2-adapted frame
field on Σ satisfy
ω̄11 + ω̄22 = 0.
Show that dω̄11 = dω̄22 = 0, and use the Poincaré lemma (cf. Theorem 2.31)
to conclude that there exists a function μ on U such that
(10.16) ω̄11 = dμ = −ω̄22 .
Remark 10.13. Technically, this result only holds if U is homeomorphic to
an open disk in R2 . If this is not the case, then U can be covered by such
open sets, and the result of Theorem 10.3 can be applied to the restriction
of the parametrization x : U → A3 to each of these open subsets of U . The
theorem can then be obtained for the entire surface Σ = x(U ) via a patching
argument.
The following exercise shows how these equations can be integrated in order
to determine the surface Σ.
*Exercise 10.14. (a) Apply the Poincaré lemma to the first equations in
(10.20) to show that there exist functions u, v on U such that
(10.21) ω̄ 1 = du, ω̄ 2 = dv
(cf. Remark 10.13). Note that the condition ω̄ 1 ∧ ω̄ 2 = 0 implies that the
functions (u, v) form a local coordinate system on U .
(b) From equations (10.21) and the remainder of equations (10.20), conclude
that the structure equations (10.3) now take the form
dx = e1 du + e2 dv,
de1 = e3 dv,
(10.22)
de2 = e3 du,
de3 = 0.
(c) Integrate equations (10.22) (beginning with the equation for de3 and
working backwards) to show that there exist constant vectors (x̄, ē1 , ē2 , ē3 )
such that
e3 (u, v) = ē3 ,
e1 (u, v) = ē1 + vē3 ,
(10.23)
e2 (u, v) = ē2 + uē3 ,
x(u, v) = x̄ + uē1 + vē2 + uvē3 .
(d) Use equations (10.23) and the fact that (e1 (u), e2 (u), e3 (u)) is a uni-
modular frame field to show that
det ē1 ē2 ē3 = 1.
and hence
x(u, v) = t [u, v, uv] .
Therefore, up to equi-affine equivalence, Σ must be an open subset of the
hyperbolic paraboloid (10.2).
320 10. Two classical theorems
Case 2: C = 0.
Suppose that C = 0. Without loss of generality, we may suppose that C > 0:
If instead we have C < 0, an equi-affine transformation of the form
x = t x1 , x2 , x3 → t −x1 , x2 , −x3
will reverse the sign of C.
*Exercise 10.15. (a) Suppose that C > 0. Show that
(10.24) dω̄ 1 = −ω̄11 ∧ ω̄ 1 , dω̄ 2 = −ω̄22 ∧ ω̄ 2 ,
and use the Frobenius theorem (cf. Theorem 2.33) to conclude that every
point u ∈ U has a neighborhood V ⊂ U on which there exist functions
u, v, g1 , g2 such that
(10.25) ω̄ 1 = eg1 du, ω̄ 2 = eg2 dv.
It suffices to prove Theorem 10.3 on each of these restricted neighborhoods
V (cf. Remark 10.13); for simplicity, we will assume that V = U . Note
that the condition ω̄ 1 ∧ ω̄ 2 = 0 implies that the functions (u, v) form a local
coordinate system on U .
(b) Show that by a transformation of the form (10.13) with λ = e 2 (g1 −g2 ) ,
1
where U (u) and V (v) are arbitrary functions of a single variable with the
property that U (u) and V (v) are nonzero and have the same sign (cf.
Exercise 9.22). Show that by making the change of coordinates
ũ = U (u), ṽ = V (v),
(b) Use equations (10.12), (10.14), (10.15), (10.26), (10.27), and (10.31) to
show that
√ √
2 2
ω̄ = √
1
du, ω̄ = √
2
dv,
C(u + v) C(u + v)
ω̄21 = ω̄12 = ω̄33 = 0,
1
ω̄11 = −ω̄22 = (−du + dv),
(10.32) (u + v)
√ √
2 2
ω̄1 = √
3
dv, ω̄2 = √
3
du,
C(u + v) C(u + v)
√ √
1 2C 2 2C
ω̄3 = du, ω̄3 = dv.
(u + v) (u + v)
(c) From equations (10.32), conclude that the structure equations (10.3) now
take the form
√ √
2 2
dx = e1 √ du + e2 √ dv ,
C(u + v) C(u + v)
" # √
1 2
de1 = e1 (−du + dv) + e3 √ dv ,
(u + v) C(u + v)
(10.33) " # √
1 2
de2 = e2 (du − dv) + e3 √ du ,
(u + v) C(u + v)
√ √
2C 2C
de3 = e1 du + e2 dv .
(u + v) (u + v)
322 10. Two classical theorems
*Exercise 10.17. (a) Show that equations (10.33) are equivalent to the
PDE system
(10.34) √ √
2 2
xu = √ e1 , xv = √ e2 ,
C(u + v) C(u + v)
√
1 1 2
(e1 )u = − e1 , (e1 )v = e1 + √ e3 ,
(u + v) (u + v) C(u + v)
√
1 2 1
(e2 )u = e2 + √ e3 , (e2 )v = − e2 ,
(u + v) C(u + v) (u + v)
√ √
2C 2C
(e3 )u = e1 , (e3 )v = e2 .
(u + v) (u + v)
(b) Integrate the equations for (e1 )u and (e2 )v to show that there exist
A3 -valued functions f (v), g(u) such that
1 1
(10.35) e1 (u, v) = f (v), e2 (u, v) = g(u).
(u + v) (u + v)
(d) Differentiate equation (10.37) with respect to u and v, and use the result
to show that there exist constant vectors c10 , c20 , c11 , c21 , c2 such that
(f) Verify that the functions in equations (10.40), (10.41) satisfy the PDE
system (10.34).
(g) Use equations (10.40) to show that
√
C
det e1 (u, v) e2 (u, v) e3 (u, v) = √ det c2 c1 c0 .
2
Since (e1 , e2 , e3 ) must be a unimodular frame, conclude that
√
2
(10.42) det c2 c1 c0 = √ .
C
Finally, check that the coordinates of x(u, v) satisfy the defining equation
(10.1) for the hyperboloid, with r = C − 3 . Therefore, up to equi-affine
2
One familiar example is the area measure on the Euclidean plane E2 : The
area measure is the 2-form dA = dx ∧ dy, and the area of any (measurable)
subset Ω ⊂ E2 is given by
&
A(Ω) = dA.
Ω
An important feature of the area measure is that it is invariant under the
action of the Euclidean group E(2): For any element g ∈ E(2) and any
measurable subset Ω ⊂ E2 , we have
A(g · Ω) = A(Ω).
We can see how this invariance comes about as follows: Recall that we can
regard E2 as the homogeneous space E2 ∼= E(2)/SO(2); i.e., E2 is the set of
left cosets of the subgroup SO(2) ⊂ E(2):
SO(2) - E(2)
π
?
E2 ∼
= E(2)/SO(2).
*Exercise 10.18. (a) Show that the product E2 ×S1 is diffeomorphic to the
oriented, orthonormal frame bundle F (E2 ), which in turn is diffeomorphic
to the Lie group E(2). (Hint: Choosing one unit vector e1 ∈ Tx E2 uniquely
determines a second unit vector e2 ∈ Tx E2 such that (e1 , e2 ) is an oriented,
orthonormal frame for Tx E2 . This is a feature that is particular to E2 and
not true for En when n ≥ 3.) Consequently, we can write
G1 (E2 ) = E(2)/ ∼ .
for some t ∈ R, where the choice of sign is the same in both diagonal entries.
Therefore, G1 (E2 ) may be regarded as the set of left cosets of the subgroup
⎧⎡ ⎤ ⎫
⎨ 1 0 0 ⎬
H = ⎣ t ±1 0 ⎦ : t ∈ R ⊂ E(2),
⎩ ⎭
0 0 ±1
and E(2) may be regarded as a principal bundle over G1 (E2 ) with fiber
group H:
H - E(2)
π
?
G1 (E2 ) ∼
= E(2)/H.
(c) Show that the Maurer-Cartan forms (ω 2 , ω21 ) on E(2) are semi-basic for
the projection π : E(2) → G1 (E2 ).
G1 (E2 ) ∼
= R × S1
(where S1 represents R/πZ rather than the more typical R/2πZ), with co-
ordinates
{(ρ, θ) | ρ ∈ R, 0 ≤ θ < π} .
10.2. The Cauchy-Crofton formula 327
SO(2) - E(2) H
π1 S π2
/ SS
w
E2 G1 (E2 ).
Show that a point x ∈ E2 and a line ∈ G1 (E2 ) are incident if and only if
the left cosets π1−1 (x), π2−1 () ⊂ E(2) have nonempty intersection.
8 is a 2-dimensional submanifold
*Exercise 10.23. Convince yourself that Ω
of E(2).
Now, let
8 ⊂ G1 (E2 ),
Ω = π2 (Ω)
and observe that Ω consists of all lines incident with α. Moreover, for any
8 is precisely twice the number of
∈ Ω, the cardinality of the set π2−1 () ∩ Ω
distinct points in which intersects α. So the set Ω 8 represents all the lines
in E that are incident with α, counted with double multiplicity.
2
where the integral on the left must be counted with multiplicity. (We are
abusing notation slightly here by writing dμ for both the 2-form |ω 2 ∧ ω21 |
on E(2) and the well-defined 2-form dμ on G1 (E2 ) whose pullback to E(2)
via π2 is equal to |ω 2 ∧ ω21 |.)
defined by
λ12 λ22 c Ω2
λ12 λ22 c Ω1
Therefore, g = f .
> g:= f;
Finally, differentiate equations (10.15):
> Simf(subs(ruledsub, Simf(d(omega[1,3]))
- d(Simf(subs(ruledsub, omega[1,3])))));
Equation (10.27) then follows from Cartan’s lemma. Add this condition to
our substitution:
> ruledsubcase2:= Simf(subs([omega[1,1] = diff(h(u,v), u)*d(u)
- diff(h(u,v), v)*d(v)], ruledsubcase2));
> ruledsubcase2:= [op(ruledsubcase2),
omega[1,1] = diff(h(u,v), u)*d(u) - diff(h(u,v), v)*d(v)];
Finally, consider the structure equation for dω̄11 :
> Simf(subs(ruledsubcase2, Simf(d(omega[1,1]))
- d(Simf(subs(ruledsubcase2, omega[1,1])))));
11.1. Introduction
Until now, all the homogeneous spaces that we have encountered have been
modeled on the vector space Rn , and we have relied extensively on the fact
that all the various structures that we have defined (Euclidean, Minkowski,
equi-affine, projective) are flat. This property is encoded in the Cartan
structure equations (3.8): According to the second equation in (3.8), the
matrix of connection forms ωc = [ωji ] satisfies the structure equation
(11.1) dωc + ωc ∧ ωc = 0.
(Note that this is not quite the same thing as the Maurer-Cartan equa-
tion (3.13) because ωc is a submatrix of the matrix-valued Maurer-Cartan
form ω.)
This might not seem like such a big deal, but in fact flatness is a necessary
condition for the existence of canonical isomorphisms Tx Rn ∼ = Rn (cf. Re-
mark 3.13 and the discussion in §3.3.2). It is these canonical isomorphisms
that allow us to think of the components (e1 (x), . . . , en (x)) of a frame field
on Rn as functions from Rn to Rn rather than as sections of the tangent
bundle T Rn , which in turn allows us to define their exterior derivatives in a
straightforward way.
339
340 11. Curves and surfaces in elliptic and hyperbolic spaces
The same reasoning as in prior cases shows that the orthonormal frame
bundle F (Sn ) may be regarded as the group SO(n + 1) via the one-to-one
correspondence
g(e0 , . . . , en ) = e0 · · · en .
The projection map π : SO(n + 1) → Sn defined by
π([e0 . . . en ]) = e0
describes SO(n + 1) as a principal bundle over Sn with fiber group SO(n);
therefore, we have a natural correspondence Sn ∼
= SO(n + 1)/SO(n).
Because we have defined the components (e0 , . . . , en ) of a frame as elements
of the vector space En+1 , we can regard them as functions eα : F (Sn ) →
En+1 and define the Maurer-Cartan forms (ωβα ) on SO(n + 1) as usual by
the equations
(11.5) deα = eβ ωαβ ,
where 0 ≤ α, β ≤ n and ωαβ = −ωβα .
*Exercise 11.5. (a) Show that the forms (ω0i ) for 1 ≤ i ≤ n are semi-basic
for the projection π : SO(n + 1) → Sn and so may be regarded as the dual
forms of any orthonormal frame (e1 , . . . , en ) based at e0 ∈ Sn . The forms
(ωji ) for 1 ≤ i, j ≤ n may then be regarded as the connection forms.
(b) Set ω i = ω0i . Show that the dual forms (ω i ) and the connection forms
(ωji ) satisfy the structure equations
dω i = −ωji ∧ ω j ,
(11.6)
dωji = −ωki ∧ ωjk + ω i ∧ ω j .
342 11. Curves and surfaces in elliptic and hyperbolic spaces
The second equation in (11.6) implies that the curvature (11.2) of the con-
nection matrix ωc = [ωji ] is given by
Ω = [Ωij ] = [ω i ∧ ω j ].
The nonzero entries of Ω reflect the fact that all sectional curvatures of Sn
are identically equal to 1 (see, e.g., [dC92]).
(c) Show that the structure equations (11.6) are equivalent to the Maurer-
Cartan equation
dω = −ω ∧ ω,
where ⎡ ⎤
0 −ω 1 · · · −ω n
⎢ω1 ω1 · · · ω1 ⎥
⎢ 1 n ⎥
ω=⎢ . .. .. ⎥
⎣ .. . . ⎦
ω n ω1n · · · ωnn
is the so(n + 1)-valued Maurer-Cartan form on SO(n + 1).
In order to describe Hn as a homogeneous space of the Lie group SO+ (1, n),
we need to compute the isotropy group of a point x ∈ Hn .
*Exercise 11.8. Let x0 = t[1, 0, . . . , 0] ∈ Hn . Show that:
(a) The isotropy group Hx0 of x0 in SO+ (1, n) is
( t )
1 0
(11.8) Hx0 = : Ā ∈ SO(n) .
0 Ā
may be regarded as an orthonormal basis for the tangent space Te0 Hn . (We
may also say that (e1 , . . . , en ) is an orthonormal frame based at e0 .) The
collection of all orthonormal frames on Hn is called the orthonormal frame
bundle of Hn , denoted F (Hn ).
Remark 11.10. For an orthonormal frame f = (e0 , . . . , en ) on Sn , each of
the vectors eα ∈ En+1 satisfies eα , eα = 1, and so could, if desired, be
identified with a point of Sn . But for an orthonormal frame on Hn , only
the vector e0 satisfies the defining condition e0 , e0 = 1 for points in Hn ,
while the vectors (e1 , . . . , en ) each satisfy ei , ei = −1. This illustrates the
fact that for non-flat homogeneous spaces (and for Riemannian manifolds in
general), we will no longer be able to regard the frame vectors (e1 , . . . , en )
as taking values in the same space as the position vector e0 , as we did for
frames on flat homogeneous spaces. Instead, we must regard each of the
vectors (e1 , . . . , en ) as taking values in the tangent bundle of the underlying
manifold.
The same reasoning as in prior cases shows that the orthonormal frame
bundle F (Hn ) may be regarded as the group SO+ (1, n) via the one-to-one
correspondence
g(e0 , . . . , en ) = e0 · · · en .
The projection map π : SO+ (1, n) → Hn defined by
π([e0 . . . en ]) = e0
344 11. Curves and surfaces in elliptic and hyperbolic spaces
describes SO+ (1, n) as a principal bundle over Hn with fiber group SO(n);
therefore, we have a natural correspondence Hn ∼ = SO+ (1, n)/SO(n).
Because we have defined the components (e0 , . . . , en ) of a frame as elements
of the vector space M1,n , we can regard them as functions eα : F (Hn ) →
M1,n and define the Maurer-Cartan forms (ωβα ) on SO+ (1, n) as usual by
the equations
(11.9) deα = eβ ωαβ ,
where 0 ≤ α, β ≤ n. Recall from Exercise 3.50 that these forms satisfy the
relations ⎧
⎪
⎨0, α = β,
β α
ωα = ωβ , α = 0 or β = 0,
⎪
⎩ α
−ωβ , α, β ≥ 1.
*Exercise 11.11. (a) Show that the forms (ω0i ) for 1 ≤ i ≤ n are semi-basic
for the projection π : SO+ (1, n) → Hn and so may be regarded as the dual
forms of any orthonormal frame (e1 , . . . , en ) based at e0 ∈ Hn . The forms
(ωji ) for 1 ≤ i, j ≤ n may then be regarded as the connection forms.
(b) Set ω i = ω0i . Show that the dual forms (ω i ) and the connection forms
(ωji ) satisfy the structure equations
dω i = −ωji ∧ ω j ,
(11.10)
dωji = −ωki ∧ ωjk − ω i ∧ ω j .
(Compare with equations (11.6).) The second equation in (11.10) implies
that the curvature (11.2) of the connection matrix ωc = [ωji ] is given by
Ω = [Ωij ] = [−ω i ∧ ω j ].
The nonzero entries of Ω reflect the fact that all sectional curvatures of Hn
are identically equal to −1.
(c) Show that the structure equations (11.10) are equivalent to the Maurer-
Cartan equation
dω = −ω ∧ ω,
where ⎡ ⎤
0 ω1 · · · ωn
⎢ω1 ω11 · · · ωn1 ⎥
⎢ ⎥
ω=⎢ . .. .. ⎥
⎣ .. . . ⎦
ω n ω1n · · · ωnn
is the so+ (1, n)-valued Maurer-Cartan form on SO+ (1, n).
11.3. A more intrinsic view of Sn and Hn 345
The Maurer-Cartan form ω and its components (ω i , ωji ) are defined as usual
on G, and the structure equations (11.6) and (11.10) remain valid in the
intrinsic setting as consequences of the Maurer-Cartan equation on G. The
primary issue that must be addressed is how to make sense of the idea of
differentiating vector fields on Xn . We can no longer think of the compo-
nents (e0 , . . . , en ) of an orthonormal frame as functions from F (Xn ) to Vn+1 ;
rather, e0 is a function from F (Xn ) to Xn , while (e1 , . . . , en ) are functions
from F (Xn ) to the tangent bundle T Xn , with the property that for any
frame f = (e0 , . . . , en ) ∈ F (Xn ), we have
ei (f ) ∈ Te0 (f ) Xn
equations (11.5) and (11.9), as each of the exterior derivatives de1 , . . . , den
contains a nonzero e0 term.
It turns out that the way to solve this problem is simply to take the or-
thogonal projection of these extrinsically defined exterior derivatives onto
the tangent plane Te0 Xn . This idea leads to the notion of the covariant
derivative for vector fields on Xn :
The remarkable fact about equation (11.12) is that, even though we used
extrinsic objects to define the covariant derivative, the result is described
entirely in intrinsic terms: The tangent vector fields (ei (x)) and the pulled-
back Maurer-Cartan forms (ω̄ji ) are well-defined on Xn as a homogeneous
space of the Lie group G, without regard to its embedding as a submanifold
of Vn+1 . So, unlike the exterior derivative of equations (11.5) and (11.9),
the covariant derivative is intrinsically defined on Xn .
The following exercise shows how the covariant derivative may be thought
of as an analog to the exterior derivative for vector fields on Xn .
defined by
(11.13) ∇v(w) = ∇w v.
Use the explicit formula (11.12) to show that ∇v is a T Xn -valued 1-form on
Xn (cf. Definition 2.18). Specifically, if (e1 (x), . . . , en (x)) is an orthonormal
frame field on Xn with associated connection forms (ω̄ji ), then
∇v = ei dv i + v j ω̄ji .
In particular, if v(x) = ei (x), then we have
(11.14) ∇ei (x) = ej (x) ω̄ij .
Thus, we can think of ∇ as a generalization of the exterior derivative d that
appears in the structure equations (3.1).
Definition 11.16. Let Γ(T Xn ) denote the space of smooth local sections
of T Xn (i.e., smooth vector fields on open sets in Xn ). The map
∇ : Γ(T Xn ) × Γ(T Xn ) → Γ(T Xn )
defined by
∇(v, w)(x) = ∇w(x) v ∈ Tx Xn
is called the Levi-Civita connection on Xn . The 1-forms (ω̄ji ) determined by
the frame field (e1 (x), . . . , en (x)) and equation (11.14) are called the con-
nection forms associated to ∇ and the frame field (e1 (x), . . . , en (x)). (Note
that this terminology is consistent with their definition as the pullbacks of
the connection forms on the frame bundle G → Xn via the frame field.)
Remark 11.17. Although the Levi-Civita connection is defined as an op-
erator on vector fields on Xn , equation (11.14) suggests—correctly!—that
there should be a related operator (also denoted ∇) on an appropriate class
of smooth maps from F (Xn ) to T Xn , determined by the condition
(11.16) ∇ei = ej ωij
348 11. Curves and surfaces in elliptic and hyperbolic spaces
The next step is where things start to look a bit different from the Euclidean
case: Since e1 (s) is a vector field along α, we have to use the covariant
derivative to differentiate it. In particular, differentiating any vector field
along the curve α means taking its covariant derivative with respect to the
unit tangent vector field along α. So the natural analog for the Euclidean
derivative e1 (s) is the covariant derivative ∇e1 (s) e1 (s). We will say that α is
nondegenerate if α is regular and, in addition, ∇e1 (s) e1 (s) = 0 for all s ∈ I.
*Exercise 11.18. In the Euclidean case, any regular curve α : I → E3 with
e1 (s) = α (s) = 0 for all s ∈ I is contained in a straight line in E3 . Consider
the analogous condition for curves in X3 : Let α : I → X3 be a regular curve
parametrized by arc length; let e0 (s) = α(s), e1 (s) = α (s), and suppose
that ∇e1 (s) e1 (s) = 0 for all s ∈ I.
(a) Use the extrinsic definition (11.11) together with the structure equations
(11.5) and (11.9) to show that, when regarded as functions e0 , e1 : I → V4 ,
we have
(11.17) e0 (s) = e1 (s), e1 (s) = k(s)e0 (s),
where k(s) = ω10 (e1 (s)).
(b) Use the fact that ω10 = ±ω01 = ±ω 1 (with the sign depending on whether
X3 = S3 or H3 ) to show that
−1, X3 = S3 ,
k(s) =
1, X3 = H3 .
The differential equation ∇α (s) α (s) = 0 is called the geodesic equation, and
the curves α in equations (11.18) and (11.19) are the geodesics in S3 and
H3 , respectively.
350 11. Curves and surfaces in elliptic and hyperbolic spaces
*Exercise 11.19. Show that we have the following analog of the Frenet
equations for nondegenerate curves α : I → X3 parametrized by arc length:
(11.20) α (s) ∇e1 (s) e1 (s) ∇e1 (s) e2 (s) ∇e1 (s) e3 (s)
⎡ ⎤
0 0 0 0
⎢ ⎥
⎢1 0 −κ(s) 0 ⎥
= α(s) e1 (s) e2 (s) e3 (s) ⎢ ⎢ ⎥,
−τ ⎥
⎣ 0 κ(s) 0 (s) ⎦
0 0 τ (s) 0
where κ, τ : I → R are smooth functions along α with
κ(s) = |∇e1 (s) e1 (s)| > 0.
(Hint: Equation (11.14) might be helpful.) As in the Euclidean case, the
functions κ(s), τ (s) are called the curvature and torsion, respectively, of α.
11.5. Moving frames for surfaces in S3 and H3 351
Once again, the functions (hij ) are related to the derivative of the Gauss
map on Σ. However, there are two important differences:
As in the Euclidean case, the eigenvalues κ1 (u), κ2 (u) of the matrix [hij (u)]
are called the principal curvatures of Σ at the point x(u), and the eigen-
vectors of the self-adjoint map −dNx(u) : Tx(u) Σ → Tx(u) Σ are called the
principal vectors or principal directions of Σ at the point x(u).
11.5. Moving frames for surfaces in S3 and H3 353
(1) If X3 = S3 , then K = K̄ + 1.
(2) If X3 = H3 , then K = K̄ − 1.
So, unlike in the Euclidean case where the Gauss equation implies that
K = K̄, here these two notions of curvature differ by the sectional curvature
of the underlying homogeneous space.
As for surfaces in E3 , the first and second fundamental forms are invariants
of the surface. Consequently, if two surfaces have different first and second
fundamental forms, then they cannot be equivalent via an isometry of M .
Lemmas 4.2 and 4.12 imply that the converse is true as well, and we have
the following analog of Bonnet’s theorem:
Theorem 11.29. Let (ω̄ 1 , ω̄ 2 , ω̄13 , ω̄23 ) be 1-forms on a connected and simply
connected open set U ⊂ R2 satisfying the conditions that (ω̄ 1 , ω̄ 2 ) are lin-
early independent at each point of U and that ω̄i3 is a scalar multiple of ω̄ i
for i = 1, 2. Suppose that, together with the Levi-Civita connection form ω̄21
determined by ω̄ 1 and ω̄ 2 , these forms satisfy the Gauss and Codazzi equa-
tions (11.22). Then there exists an immersed surface x : U → X3 , unique up
to transformation by an element of G, whose first and second fundamental
forms are
I = (ω̄ 1 )2 + (ω̄ 2 )2 ,
II = ω̄13 ω̄ 1 + ω̄23 ω̄ 2 .
(b) Use the extrinsic definition (11.11) together with the structure equations
(11.5) and (11.9) to show that, when regarded as functions e0 , e1 , e2 , e3 :
U → V4 , equation (11.26) implies that
de3 = 0.
Conclude that there exist orthogonal unit vectors ē0 , ē1 , ē2 ∈ V4 such that
Σ = x(U ) is contained in the “great sphere” of S3 or the “great hyper-
boloid” of H3 determined by the intersection of X3 with the plane spanned
by (ē0 , ē1 , ē2 ) (cf. Exercise 11.20). This is the analog of the fact that any
totally geodesic surface in E3 is contained in a plane.
of the wave equation gives rise to a flat surface in S3 whose first and second
fundamental forms are given by (11.27).
(b) Let a, b > 0 with a2 +b2 = 1, and consider the map x : S1 ×S1 → S3 ⊂ E4
defined by
x(θ, ϕ) = t [a cos(θ), a sin(θ), −b cos(ϕ), b sin(ϕ)] ,
where θ and ϕ denote angle coordinates on the two copies of S1 . Show that
Σ = x(S1 × S1 ) is a flat torus in S3 , with angle function equal to a constant
ψ0 such that
cos(ψ0 ) = a, sin(ψ0 ) = b.
(Hint: Begin by considering the principal adapted frame field
e0 (θ, ϕ) = x(θ, ϕ) = t[a cos(θ), a sin(θ), −b cos(ϕ), b sin(ϕ)],
xθ (θ, ϕ)
e1 (θ, ϕ) = = t[− sin(θ), cos(θ), 0, 0],
|xθ (θ, ϕ)|
xϕ (θ, ϕ)
e2 (θ, ϕ) = = t[0, 0, sin(ϕ), cos(ϕ)],
|xϕ (θ, ϕ)|
e3 (θ, ϕ) = t[−b cos(θ), −b sin(θ), −a cos(ϕ), a sin(ϕ)]
and its associated Maurer-Cartan forms.)
cosh(ψ0 ) = a, sinh(ψ0 ) = b.
We’ll need to remove ω̄21 from the list in adaptedsub because Maple
won’t like this substitution in the expressions diff(psi0(u,v), u) and
11.6. Maple computations 359
diff(psi0(u,v), v). We’ll also introduce a new name for the function
ψ(x, y) = ψ0 (x + y, x − y).
> PDETools[declare](psi(x,y));
> adaptedsub asymp:= Simf(subs([u = x + y, v = x - y],
Simf(subs([psi0(u,v) = psi(x,y)],
[seq(adaptedsub[i], i=1..5)]))));
Maple doesn’t really know how to compute symmetric products of differ-
ential forms, but the following commands will work for computing the first
and second fundamental forms:
> collect(simplify(Simf(subs(adaptedsub asymp, omega[1]))ˆ2
+ Simf(subs(adaptedsub asymp, omega[2]))ˆ2), {d(x), d(y)});
> collect(simplify(Simf(subs(adaptedsub asymp, omega[3,1]))*
Simf(subs(adaptedsub asymp, omega[1]))
+ Simf(subs(adaptedsub asymp, omega[3,2]))*
Simf(subs(adaptedsub asymp, omega[2]))), {d(x), d(y)});
For Exercise 11.34(b), we can compute the Maurer-Cartan forms associated
to the given frame field, as follows. First, declare a, b to be constants and
define the frame vectors (e0 , e1 , e2 , e3 ):
> Form(a=-1, b=-1);
> e0:= Vector([a*cos(theta), a*sin(theta), -b*cos(phi),
b*sin(phi)]);
e1:= Vector([-sin(theta), cos(theta), 0,0]);
e2:= Vector([0,0, sin(phi), cos(phi)]);
e3:= Vector([-b*cos(theta), -b*sin(theta), -a*cos(phi),
a*sin(phi)]);
The fastest way to compute the Maurer-Cartan forms is to define the cor-
responding group element
g = e0 e1 e2 e3 ∈ SO(4)
ω̄ = g −1 dg,
as follows:
> g:= Matrix([e0, e1, e2, e3]);
connection matrix:= simplify(MatrixInverse(g).map(d, g));
360 11. Curves and surfaces in elliptic and hyperbolic spaces
⎡ ⎤
d(θ)a d(φ)
⎢ 0 − −b 0 ⎥
⎢ a2 + b2 a2 + b2 ⎥
⎢ d(θ)a 0 0 −b d(θ)⎥
⎢ ⎥
connection matrix = ⎢
⎢b d(φ)
⎥
⎢ 0 0 a d(φ) ⎥
⎥
⎢ ⎥
⎣ d(θ) d(φ)a ⎦
0 b 2 2
− 2 2
0
a +b a +b
Since a2 + b2 = 1, we see that we have the following Maurer-Cartan forms:
> examplesub:= [omega[1] = a*d(theta), omega[2] = b*d(phi),
omega[3] = 0, omega[1,2] = 0, omega[3,1] = b*d(theta),
omega[3,2] = -a*d(phi)];
These forms agree with the forms in adaptedsub, with cos(ψ0 ) = a, sin(ψ0 )
= b, as desired.
Chapter 12
The nonhomogeneous
case: Moving frames
on Riemannian
manifolds
12.1. Introduction
So far, we have been using moving frames to study the geometry of curves
and surfaces as submanifolds Σ of homogeneous spaces G/H. In this context,
the geometry of Σ is determined by the geometry of the ambient homoge-
neous space G/H and the particular way that Σ is embedded in G/H as a
submanifold. But there are many interesting geometric problems for which
this scenario is too restrictive. For instance, we may be interested in the geo-
metric structure of a nonhomogeneous manifold that is defined intrinsically
and not as a submanifold of some larger ambient homogeneous space. Or,
even in the study of submanifolds, we might be interested in submanifolds
Σ of some manifold M that is not homogeneous.
Recall that for a homogeneous space G/H, the natural projection map
π : G → G/H
361
362 12. Moving frames on Riemannian manifolds
v, w = g(v, w)
for v, w ∈ Tx M .
12.2. Orthonormal frames and connections 363
SO(n) - F (M )
π
?
M.
There are several important differences between the orthonormal frame bun-
dle of the homogeneous space En and that of a general Riemannian manifold
M . First, while each fiber of the frame bundle is acted on freely and tran-
sitively by SO(n), there is no larger group that acts transitively on the
entire frame bundle F (M ). But the most significant change is that, given
an orthonormal frame field (e1 (x), . . . , en (x)) on an open set in M , there
is no natural way of thinking of the frame vector fields (ei (x)) as functions
from M to a fixed vector space—not even by regarding M as a submanifold
of some larger Euclidean space, as we did for Sn and Hn in Chapter 11.
Rather, they are sections of the tangent bundle T M , which means that for
364 12. Moving frames on Riemannian manifolds
each point x ∈ M , the vectors (e1 (x), . . . , en (x)) take values in the vector
space Tx M .
In the case of En , we were able to differentiate vector fields by using the fact
that each tangent space Tx En is canonically isomorphic to En and regarding
the vector fields (e1 (x), . . . , en (x)) as functions into this fixed vector space
(cf. Remark 3.13). And even for Sn and Hn , we were able to make use of
these canonical isomorphisms for the ambient spaces En+1 and M1,n in order
to define the covariant derivatives of vector fields. But for a more general
Riemannian manifold M , there is no such canonical isomorphism between
each tangent space Tx M and a fixed vector space En , or even a canonical
embedding of Tx M into a larger fixed vector space; indeed, there are infin-
itely many ways of identifying each tangent space Tx M with a fixed vector
space En , all of which are equally valid. The following exercise illustrates
some of the complications that may arise as a result of this ambiguity.
*Exercise 12.4. To any local orthonormal frame field (e1 (x), . . . , en (x))
on an open set U ⊂ M , we can associate a local trivialization (cf. §1.5)
φ : T U → U × En of the tangent bundle T M by defining
(12.1) φ x, ai ei (x) = x, t[a1 , . . . , an ] .
(a) Show that the map φ defines an isometry between each tangent space
Tx M and the vector space En .
(b) We may regard a local trivialization φ as a local choice of basis vector
fields (e1 (x), . . ., en (x)) for sections of T M , and the identification (12.1)
makes it tempting to think that we might be able to regard these vector
fields as “constant” for purposes of differentiation. But what happens when
we choose a different basis? Let (ẽ1 (x), . . . , ẽn (x)) be any other orthonormal
frame field on U , related to the original frame field by
(12.2) ẽ1 (x) . . . ẽn (x) = e1 (x) . . . en (x) Ā(x),
where Ā(x) is an SO(n)-valued function on U . Show that under the analo-
gous local trivialization φ̃ associated to the orthonormal frame field (ẽ1 (x),
. . ., ẽn (x)), the vector fields (e1 (x), . . . , en (x)) are identified with the col-
−1
umns of the matrix Ā(x) . In particular, vector fields that appear “con-
stant” with respect to one trivialization do not necessarily remain “constant”
with respect to a different trivialization.
a vector bundle.) It turns out that, in order to make sense of the notion
of differentiation for vector fields on M , we need to introduce an additional
structure, called a connection, on the tangent bundle Tx M .
Definition 12.5. Let Γ(T M ) denote the space of smooth local sections of
T M (i.e., smooth vector fields on open sets in M ). An affine connection (or,
more succinctly, a connection) ∇ on T M is a map
∇ : Γ(T M ) × Γ(T M ) → Γ(T M ),
with ∇(w, v) denoted by ∇w v, such that for any vector fields v, v1 , v2 , w,
w1 , w2 ∈ Γ(T M ), any smooth, real-valued functions f, g on M , and any
real numbers a, b ∈ R, we have
Conditions (1) and (2) are linearity properties: They say that the map ∇ is
linear over smooth functions in its first input and linear over real numbers in
its second input. Condition (3) is an analog of the Leibniz rule that describes
how ∇ behaves when its second input is multiplied by a smooth function.
The vector field ∇w v should be regarded as defining a sort of “directional
derivative” of the vector field v in the direction of w, and conditions (1)–(3)
are precisely the conditions that such a directional derivative must satisfy.
Remark 12.7. This definition applies more generally to any smooth vector
bundle B over a manifold M : If π : B → M is a vector bundle with fibers
isomorphic to a fixed k-dimensional vector space V (cf. §1.5), then an affine
connection ∇ on B is a map
∇ : Γ(T M ) × Γ(B) → Γ(B)
satisfying the properties of Definition 12.5. This more general setting il-
lustrates the fact that the two copies of Γ(T M ) used for the inputs of ∇
in Definition 12.5 play significantly different roles: The first input is the
direction along which differentiation should occur (which must be a tangent
vector to M ), and the second input is the object to be differentiated. The
output is the resulting differentiated object, and it should live in the same
space as the second input.
366 12. Moving frames on Riemannian manifolds
(12.3) T En ∼
= En × En ,
where the first factor represents the base manifold En and the second factor
represents the fibers Tx En . (In other words, we have a canonical global
trivialization of the tangent bundle T En .) This, in turn, allows us to write
the tangent bundle T (T En ) as the product manifold
(12.4) T (T En ) ∼
= T En × T En
in the obvious way. The vector field v is a section of T En , and via the
identification (12.3), we can write it as a function σ : En → En × En defined
by
σ(x) = (x, v(x)) .
Show that ∇w v is given by the composition
(12.5) ∇w v = π2 ◦ dσ(w),
where dσ : T En → T (T En ) is the differential of the map σ (cf. §1.3) and
π2 : T En × T En → T En is the projection onto the second factor. This
connection is called the flat connection on En .
Remark 12.9. For a general Riemannian manifold M , there is generally
no global trivialization analogous to equation (12.3) for T M . However, for
each point (x, v) ∈ T M , the 2n-dimensional tangent space T(x,v) (T M ) can
be decomposed in a manner analogous to equation (12.4) in many different
ways. The n-dimensional subspace
V(x,v) = T(x,v) (Tx M ) ⊂ T(x,v) (T M )
(corresponding to the second factor in (12.4)) is canonically defined: It is
the tangent space to the fiber Tx M ⊂ T M and is called the vertical tangent
12.2. Orthonormal frames and connections 367
π2 : T(x,v) (T M ) → V(x,v)
(12.7) ∇v(w) = ∇w v.
(b) Show that if (e1 (x), . . . , en (x)) is an orthonormal frame field on an open
set U ⊂ M , then ∇ei is the T M -valued 1-form
(Note that, while the exterior derivative of a vector field on M is not well-
defined, it still makes perfect sense to compute the exterior derivative of
real-valued functions, such as v i , on M .) For this reason, a connection is
sometimes expressed with respect to a given trivialization of T M as
∇ = d + ω̄,
As you might have guessed from the notation, the dual forms and connection
forms (ω̄ i , ω̄ji ) on M associated to an orthonormal frame field (e1 (x), . . .,
en (x)) on M are the pullbacks to M of certain 1-forms (ω i , ωji ) on the frame
bundle F (M ) via the section
f (x) = (x; e1 (x), . . . , en (x))
of F (M ). Fortunately, as the following exercise will show, equations (12.10)
and (12.11) tell us precisely how these 1-forms should be defined. (It might
be helpful to review §3.3.2 at this point, particularly the derivations of equa-
tions (3.3) and (3.4).)
*Exercise 12.15. Let (e1 (x), . . ., en (x)) be a local orthonormal frame field
on an open set U ⊂ M , with associated dual and connection forms (ω̄ i , ω̄ji ).
Just as this orthonormal frame field determines a local trivialization of T M
(cf. Exercise 12.4), it can also be used to define a local trivialization of
the principal bundle F (M ): For any x ∈ U and any orthonormal frame
f = (x; e1 , . . . , en ) based at x, we can write
(12.12) e1 · · · en = e1 · · · en A
370 12. Moving frames on Riemannian manifolds
Note the distinction between these 1-forms and those in equations (12.10)
and (12.11): In those equations, Ā is an SO(n)-valued function on the open
set U ⊂ M , whereas in equations (12.14), the entries of A represent lo-
cal coordinates on SO(n), independent of the local coordinates on M . So,
while the 1-forms (ω̄ i , ω̄ji ) are 1-forms on M and the (ω̄ji ) are linear combi-
nations of the (ω̄ i ), the 1-forms (ω i , ωji ) are linearly independent 1-forms on
the orthonormal frame bundle F (M ). Use equations (12.10), (12.11), and
(12.13) to show that the 1-forms (12.14) are well-defined, independent of
the choice of orthonormal frame field (e1 (x), . . . , en (x)) used to define the
1-forms (ω̄ i , ω̄ji ) and the local trivialization φ of F (M ).
In the following exercises, we will show how to prove Theorem 12.17. The
strategy of the proof is to show that if a torsion-free, metric-compatible
connection exists, then it must be unique. In the process, an explicit formula
for this unique connection is derived, which serves to prove the existence
result as well.
To begin the proof, let (e1 (x), . . . , en (x)) be an orthonormal frame field
on an open set U ⊂ M , with associated dual forms (ω̄ 1 , . . . , ω̄ n ). (At this
point, we will drop the underscore notation on the vector fields ei (x) since
we no longer need to think of them as “basis” vector fields.) Let ∇ be an
affine connection on M , and let (ω̄ji ) be the corresponding connection forms
associated to the given frame field. Since the dual forms are a basis for the
372 12. Moving frames on Riemannian manifolds
*Exercise 12.20. Suppose that ∇ is torsion-free, and let ckij (x) = −ckji (x)
be the functions on U defined by the Lie bracket relations
[ei , ej ] = ckij ek .
*Exercise 12.21. Now, suppose that ∇ is both compatible with the metric
and torsion-free.
(a) Use equations (12.16) and (12.17) to show that
!
(12.18) akij = 12 cjki − cijk − ckij .
(Hint: This is an exercise in index juggling. Start with akij and apply equa-
tions (12.16) and (12.17) alternately until you come back to the index ar-
rangement that you started with. And keep in mind that ckij = −ckji !)
(b) Conclude that Theorem 12.17 is true and that the Levi-Civita connection
is defined by equation (12.15), with (akij ) as in equation (12.18).
The map dx is exactly analogous to the Euclidean case: The vectors (e1 ,
. . ., en ) form a basis for Tx M at each point, and the dual forms are defined
precisely so that
(12.19) dx = ei ω i .
The maps (dei ) are a bit more complicated. For each f ∈ F (M ), the image
of (dei )f takes values in the 2n-dimensional tangent space T(x,ei ) (T M ). The
Levi-Civita connection ∇ determines a linear projection operator
π2 : T(x,ei ) (T M ) → T(x,ei ) (Tx M ) ∼
= Tx M,
defined by the condition that
(12.20) (π2 ◦ d)(ei ) = ∇ei = ej ωij
(cf. Remark 12.9). This equation is the analog of the equation for dei in
equations (3.1).
In order to compute the structure equations for the forms (ω i , ωji ), we will
first need to differentiate equation (12.19). This requires some care: Since
both sides of the equation are 1-forms that take values in T M , we must use
the connection ∇ to differentiate them. And in general, it is not necessarily
true that ∇ ◦ d = 0, so we cannot directly apply any obvious analog of
the identity d ◦ d = 0. Fortunately, we can get around this problem by
considering two different expressions for the T M -valued 1-form dx, as the
following two exercises show.
374 12. Moving frames on Riemannian manifolds
*Exercise 12.22. Recall that with respect to any local coordinate system
x = (x1 , . . . , xn ) on an open set U ⊂ M , we can write
∂
(12.21) dx = dxi ,
∂xi
where ∂x∂ 1 , . . . , ∂x∂n are the coordinate vector fields on U .
(a) Show that the coordinate vector fields ∂x∂ 1 , . . . , ∂x∂n have pairwise Lie
∂ ∂
brackets equal to zero; i.e., ∂x i , ∂xj = 0 (cf. Exercise 1.34).
(b) Apply ∇ to equation (12.21) and use an argument similar to that of
Exercise 12.11, part (a), to show that
" #
∂ ∂
∇ (dx) = ∇ i
∧ dxi + i d(dxi ).
∂x ∂x
Conclude that
" #
∂
(12.22) ∇ (dx) = ∇ ∧ dxi .
∂xi
(c) Use the definition (12.7) for the T M -valued 1-form ∇ ∂
∂xi
to show that
" # " #
∂ ∂
(12.23) ∇ =∇ ∂ dxj .
∂xi ∂xj ∂xi
(e) Conclude from part (a) and the fact that the Levi-Civita connection is
torsion-free that ∇ (dx) = 0.
*Exercise 12.23. Now that we know that ∇ (dx) = 0, we can differentiate
equation (12.19) to obtain the structure equations for the dual forms.
(a) Apply ∇ to equation (12.19) to show that
(12.24) ∇ei ∧ ω i + ei dω i = 0.
(b) Apply equation (12.20) (and a bit of index juggling) to equation (12.24)
to obtain
ei dω i + ωji ∧ ω j = 0.
(c) Use the linear independence of the vectors (e1 , . . . , en ) in each tangent
space Tx M to conclude that the dual forms (ω 1 , . . . , ω n ) satisfy the structure
equations
(12.25) dω i = −ωji ∧ ω j .
12.4. The structure equations 375
*Exercise 12.24. In this exercise, we will show that the structure equations
(12.25) can be used to compute the connection forms directly from the dual
forms associated to any orthonormal frame field on M . To this end, let
(e1 (x), . . . , en (x)) be an orthonormal frame field on an open set U ⊂ M ,
with associated dual forms (ω̄ 1 , . . . , ω̄ n ).
(a) Show that if there exist 1-forms (ω̄ji ) on U that satisfy the metric com-
patibility condition ω̄ij = −ω̄ji and the equations
(c) In order to show that there exist such 1-forms (ω̄ji ), let ckij = −ckji be the
functions on U defined by the conditions
(12.27) dω̄ k = −ckij ω̄ i ∧ ω̄ j .
Suppose that the desired 1-forms (ω̄ji ) exist; then we must have
ω̄ji = aijk ω̄ k
for some functions (aijk ) on U . Since ω̄ji = −ω̄ij , the functions (aijk ) must
satisfy the conditions
aijk = −ajik .
Show that the structure equations (12.26) are equivalent to the equations
akji − akij = ckij .
Observe that these are precisely the same conditions as equations (12.16),
(12.17), and conclude that the functions (akij ) defined by equation (12.18)
produce the desired connection forms (ω̄ji ).
Our next task is to compute the structure equations for the connection
forms (ωji ). Differentiating equation (12.20) as we did for the analogous
equation in the Euclidean case turns out to be impractical: We would have
to use the connection ∇ to differentiate the equation, and there is no useful
identity for ∇ ◦ ∇ that would allow us to differentiate the left-hand side
376 12. Moving frames on Riemannian manifolds
(a) Use the fact that ωji = −ωij to show that Ωij = −Ωji .
(b) Show that we can write
(Hint: Ωij is a 2-form on F (M ), and the 1-forms (ω i , ωji ) are a basis for the
1-forms on F (M ).)
(c) Equation (12.28) can now be written as
!
1 i
2 Rjk ω
k
∧ ω + Sjm
ik
ωk ∧ ω m + Tjn ωk ∧ ωm
ikm n
∧ ω j = 0.
(d) Use equations (12.32) and (12.33) to show that Sjm ik = 0. (Hint: This
Remark 12.29. The reason for including the factor of 12 in equations (12.31)
and (12.35) is so that the curvature 2-forms (Ωij ) may be written as
(12.36) Ωij = 12 Rjk
i
ωk ∧ ω = i
Rjk ωk ∧ ω.
k<
In particular, when n = 2, the Riemann curvature tensor has a single non-
1 , which is equal to the Gauss curvature K.
trivial component R212
Exercise 12.30. The Nash embedding theorem [Nas56] states that any
n-dimensional Riemannian manifold M can be isometrically embedded into
a Euclidean space En+m of sufficiently large dimension. The structure equa-
tions for the Maurer-Cartan forms on M can then be derived from those on
En+m , as follows.
Suppose that M ⊂ En+m is a smooth submanifold of En+m . Choose an
orthonormal frame field (e1 (x), . . ., en+m (x)) for Tx En+m along M so
that for each x ∈ M , the tangent space Tx M is spanned by the vectors
(e1 (x), . . . , en (x)). Let (ω̄ α , ω̄βα ), where 1 ≤ α, β ≤ (n + m), denote the
Maurer-Cartan forms associated to this frame field on M .
(a) Show that
(12.37) ω̄ n+1 = · · · = ω̄ n+m = 0
(cf. Proposition 4.18).
(b) Differentiate equations (12.37) and use Cartan’s lemma to show that
there exist functions (haij ) on M , with 1 ≤ i ≤ n, (n + 1) ≤ a ≤ (n + m),
and haij = haji , such that
(12.38) ω̄ia = haij ω̄ j .
These are the coefficients of the second fundamental form
II = haij ea ⊗ ω̄ i ω̄ j
of M ⊂ En+m (cf. Exercise 4.27).
(c) We can define a connection ∇ on M in much the same way that we
did for Sn and Hn (cf. §11.3), as follows: For any vector field v on M
and any vector w ∈ Tx M , define the covariant derivative of v with respect
to w to be the vector ∇w v ∈ Tx M given by the orthogonal projection of
the Euclidean directional derivative dv(w) in En+m onto the tangent plane
Tx M ⊂ Tx En+m ∼ = En+m . Then let ∇v : T M → T M be the T M -valued
1-form on M defined by
∇v(w) = ∇w v,
as in Exercise 12.10. Show that for 1 ≤ i ≤ n,
∇ei (x) = ej (x) ω̄ij ,
12.5. Moving frames for curves in Riemannian manifolds 379
where the sum ranges over 1 ≤ j ≤ n. Conclude that the 1-forms (ω̄ji )
with 1 ≤ i, j ≤ n are the connection forms associated to the connection ∇
and the orthonormal frame field (e1 (x), . . . , en (x)) on M . Use the fact that
ω̄ji = −ω̄ij and the structure equations for the (dω̄ i ) on En+m to argue that
∇ is, in fact, the Levi-Civita connection on M .
(d) Substitute the result of part (b) into the structure equations
dω̄ji = −ω̄αi ∧ ω̄jα , 1 ≤ i, j ≤ n,
on En+m (where the sum ranges over 1 ≤ α ≤ (n + m)), and compare with
the structure equations (12.35) to obtain the Gauss equations
n+m
a a
(12.39) i
Rjk = hik hj − hai hajk .
a=n+1
These equations are the higher-dimensional analog of the Gauss equation
(4.11) for surfaces in E3 .
Similarly, the structure equations for (dω̄ia ) with 1 ≤ i ≤ n, (n + 1) ≤ a ≤
(n + m) are the higher-dimensional analog of the Codazzi equations (4.12),
while the structure equations for (dω̄ba ) with (n + 1) ≤ a, b ≤ (n + m)} are
called the Ricci equations. (The Ricci equations have no analog for surfaces
in E3 because they only appear for embeddings of codmension greater than
or equal to 2.)
The remainder of this chapter will be devoted to the study of curves and
surfaces in a 3-dimensional Riemannian manifold M .
of each tangent space T(x,v) (T M ) into horizontal and vertical tangent spaces,
where V(x,v) = T(x,v) (Tx M ) ⊂ T(x,v) (T M ) is canonically defined and the
complementary subspace H(x,v) is determined by ∇.
Let π : T M → M denote the standard projection operator from T M to
M , and let dπ : T(x,v) (T M ) → Tx M denote its differential at the point
(x, v) ∈ T M . Show that for each vector v ∈ Tx M , there exists a unique
vector v̂ ∈ H(x,v) such that dπ(v̂) = v.
(b) Any regular curve α : I → M has a unique lifting to a curve α :
I → T M defined by the condition that for each t ∈ I, α (t) ∈ Tα(t) is the
tangent vector to α at the point α(t). This curve, in turn, has a unique
lifting to a curve (α ) : I → T (T M ) defined analogously: For each t ∈ I,
(α ) (t) ∈ Tα (t) (T M ) is the tangent vector to α at the point α (t) ∈ T M .
Show that under the decomposition (12.40) of the tangent space Tα (t) (T M ),
we have
(α ) (t) = α (t) + ∇ α (t),
α (t)
where α (t) ∈ H
α (t) and ∇α (t) α (t) ∈ Vα (t) . (Hint: It follows from the
discussion in Remark 12.9 that ∇α (t) α (t) is the projection of (α ) (t) onto
the vertical tangent space Vα (t) , so you only need to show that the projection
of (α ) (t) onto the horizontal tangent space Hα (t) is equal to α (t).)
(c) Conclude from part (b) that α is a geodesic if and only if (α ) (t) ∈ Hα (t)
for all t ∈ I.
(d) A vector field w on T M is called horizontal if w(x, v) ∈ H(x,v) for all
(x, v) ∈ T M . The horizontal vector field
w(x, v) = v̂
12.6. Moving frames for surfaces in Riemannian manifolds 381
(b) Let (ω̄ i , ω̄ji ) denote the pullbacks of the Maurer-Cartan forms (ω i , ωji )
on F (M ) to U via x̃. Then we have ω̄ 3 = 0, and the 1-forms (ω̄ 1 , ω̄ 2 ) form
a basis for the 1-forms on U .
(c) The metric on M naturally induces a metric on Σ = x(U ) ⊂ M , given
by the first fundamental form
I = (ω̄ 1 )2 + (ω̄ 2 )2 .
As for surfaces in E3 , the first and second fundamental forms are invariants
of the surface. Consequently, if two surfaces have different first and second
fundamental forms, then they cannot be equivalent via an isometry of M .
However, the converse question is not so straightforward: In general, when
there is no group structure on F (M ), there are no analogs of Lemmas 4.2
and 4.12 that tell us when we have a complete set of invariants for a surface,
or when two surfaces are equivalent via a symmetry of M . In particular,
given arbitrary quadratic forms I and II on an open set U ⊂ R2 , checking
that the structure equations are satisfied is not sufficient to guarantee that
there exists an immersion x : U → M whose first and second fundamental
forms are I and II. The best that we can do is to arbitrarily specify one
of these quadratic forms, and even then the existence question may be very
difficult. For instance, we have the following local isometric embedding
theorem, which is due independently to Cartan [Car27] and Janet [Jan26].
The proof of this theorem uses exterior differential systems; it may be found
in [BCG+ 91]. The assumption of real analyticity is a crucial ingredient of
the proof; without this hypothesis, much less is known about the existence
of isometric embeddings, even locally. Even the local existence question for
embedding a surface with a C ∞ metric into E3 is not completely solved. (For
a comprehensive treatment of the current status of the isometric embedding
problem, see [HH06].)
For another example of how the geometry of surfaces in a general 3-dimen-
sional Riemannian manifold M differs from the geometry of surfaces in the
homogeneous spaces E3 , S3 , and H3 , consider the family of totally geodesic
surfaces in M , which are defined exactly as in Definition 11.30:
so that the vectors (a1 , a2 , a3 ) in R3 represent the columns of Ā. Show that
for any cyclic permutation (i, j, k) of the indices (1, 2, 3), we have
ai × aj = ak .
Similarly, if (a1 , a2 , a3 ) represent the rows of Ā, then
ai × aj = ak .
ω̄ 1 ∧ ω̄ 2 ˜ 1 ∧ ω̄
ω̄ ˜2
Use equation (12.11) and part (a) to show that
˜ † = Ā−1 ω̄ † .
ω̄
The symmetries in equation (12.43) imply that these matrices are symmetric,
and by definition, we have
(12.44) Ω̄† = R̄ ω̄ † , ˜ † = R̄
Ω̄ ˜ †.
˜ ω̄
Next, we will see how the curvature tensor can impose constraints on the
existence of totally geodesic surfaces.
*Exercise 12.41. Suppose that Σ = x(U ) ⊂ M is a totally geodesic surface
in M . Let (e1 (u), e2 (u), e3 (u)) be an adapted frame field along Σ, with
associated dual and connection forms (ω̄ i , ω̄ji ) on U .
(a) Show that the condition (12.42) is equivalent to the condition that
(12.46) ω̄13 = ω̄23 = 0.
386 12. Moving frames on Riemannian manifolds
dz
ω̄ 3 = .
(ax + by + cz 2 + 1)
2 2
12.6. Moving frames for surfaces in Riemannian manifolds 387
(b) Show that the Levi-Civita connection forms associated to the dual forms
(12.50) are
2
ω̄32 = (by dz − cz dy),
(ax2 + by 2 + cz 2 + 1)
2
(12.51) ω̄13 = (cz dx − ax dz),
(ax2 + by 2 + cz 2 + 1)
2
ω̄21 = (ax dy − by dx).
(ax + by + cz 2 + 1)
2 2
A similar argument shows that the only totally geodesic surfaces in M are
contained in the coordinate “planes”
Σ1 = {(0, y, z) | y, z ∈ R} ⊂ M,
Σ2 = {(x, 0, z) | x, z ∈ R} ⊂ M,
Σ3 = {(x, y, 0) | x, y ∈ R} ⊂ M.
Any open submanifold M0 ⊂ M that does not intersect these surfaces—for
instance, the first octant M0 = {(x, y, z) ∈ M | x, y, z > 0}—contains no
totally geodesic surfaces whatsoever.
omega[j,a]:= -omega[a,j];
end do;
end do;
Tell Maple how to differentiate these forms according to the Cartan struc-
ture equations (3.8), and be sure to define structure equations only for those
connection forms that have not been defined in terms of other forms:
> for i from 1 to 6 do
d(omega[i]):= sum(’-omega[i,j] &ˆ omega[j]’, ’j’=1..6);
end do;
d(omega[1,2]):= sum(’-omega[1,k] &ˆ omega[k,2]’, ’k’=1..6);
d(omega[2,3]):= sum(’-omega[2,k] &ˆ omega[k,3]’, ’k’=1..6);
d(omega[3,1]):= sum(’-omega[3,k] &ˆ omega[k,1]’, ’k’=1..6);
d(omega[4,5]):= sum(’-omega[4,k] &ˆ omega[k,5]’, ’k’=1..6);
d(omega[5,6]):= sum(’-omega[5,k] &ˆ omega[k,6]’, ’k’=1..6);
d(omega[6,4]):= sum(’-omega[6,k] &ˆ omega[k,4]’, ’k’=1..6);
for a from 4 to 6 do
for j from 1 to 3 do
d(omega[a,j]):= sum(’-omega[a,k] &ˆ omega[k,j]’,
’k’=1..6);
end do;
end do;
Now, consider an adapted orthonormal frame field along M for which the
frame vectors (e1 (u), e2 (u), e3 (u)) are tangent to M at each point. Accord-
ing to equation (12.37), for such a frame field, we must have ω̄ 4 = ω̄ 5 =
ω̄ 6 = 0. Set up a substitution for the Maurer-Cartan forms associated to an
adapted frame field:
> adaptedsub:= [omega[4]=0, omega[5]=0, omega[6]=0];
Now differentiate these equations:
> Simf(subs(adaptedsub, Simf(d(omega[4]))));
Simf(subs(adaptedsub, Simf(d(omega[5]))));
Simf(subs(adaptedsub, Simf(d(omega[6]))));
Applying Cartan’s lemma tells us that equations (12.38) hold, and we can
add these equations to our substitution as follows:
> for a from 4 to 6 do
h[a,2,1]:= h[a,1,2];
h[a,3,2]:= h[a,2,3];
390 12. Moving frames on Riemannian manifolds
h[a,1,3]:= h[a,3,1];
end do;
> unassign(’i’, ’j’, ’a’);
> adaptedsub:= [op(adaptedsub),
seq(seq(omega[a,i] = sum(h[a,i,j]*omega[j],
j=1..3), i=1..3), a=4..6)];
(The unassign command is needed because one of the previous for loops
assigned values to these indices that need to be removed before using them
in a seq command.)
Now look at the structure equations for dω̄ 1 , dω̄ 2 , dω̄ 3 :
> Simf(subs(adaptedsub, Simf(d(omega[1]))));
It follows from these equations that the 1-forms (ω̄ji ) with 1 ≤ i, j ≤ 3 are
the Levi-Civita connection forms associated to the metric
2 2 2
I = ω̄ 1 + ω̄ 2 + ω̄ 3
The scalar coefficients in these quantities are equivalent to the Gauss equa-
tions (12.39).
Next, look at the structure equations for the connection forms (ω̄ia ) with
1 ≤ i ≤ 3, 4 ≤ a ≤ 6:
> zero41:= Simf(subs(adaptedsub, Simf(d(omega[4,1]))
- d(Simf(subs(adaptedsub, omega[4,1])))));
The resulting expression has the form
φ1 ∧ ω̄ 1 + φ2 ∧ ω̄ 2 + φ3 ∧ ω̄ 3 ,
Applying Cartan’s lemma to all 9 structure equations for the (dω̄ia ) shows
that there exist functions (haijk ), symmetric in their lower indices, such that
these equations take the form
These are the Codazzi equations; there are 18 of them, one for each of the
components (haij ) of the second fundamental form.
Remark 12.43. From the PDE perspective, the symmetry of the (haijk ) is
equivalent to a system of PDEs of the form
∂haij ∂haik
− = (lower order terms).
∂xk ∂xj
From this point of view, there are 27 Codazzi equations, 24 of which are
independent.
392 12. Moving frames on Riemannian manifolds
Finally, look at the structure equations for the connection forms (ω̄ba ) with
4 ≤ a, b ≤ 6:
> Simf(subs(adaptedsub, Simf(d(omega[4,5]))));
> Simf(subs(adaptedsub, Simf(d(omega[5,6]))));
> Simf(subs(adaptedsub, Simf(d(omega[6,4]))));
These all have the form
In order to compute the connection forms, differentiate the dual forms and
compare the result to the structure equations. (It’s helpful to go ahead
and write the (ω̄ji ) as linear combinations of the (ω̄ i ) with undetermined
coefficients and then solve for these coefficients.)
> connectionformssub:= [
omega[1,2] = h[1,2,1]*omega[1] + h[1,2,2]*omega[2]
+ h[1,2,3]*omega[3],
omega[3,1] = h[3,1,1]*omega[1] + h[3,1,2]*omega[2]
+ h[3,1,3]*omega[3],
omega[3,2] = h[3,2,1]*omega[1] + h[3,2,2]*omega[2]
+ h[3,2,3]*omega[3]];
According to the structure equations, the following expressions should be
zero:
> zero1:= Simf(d(Simf(subs(dualformssub, omega[1])))
+ Simf(subs(dualformssub, Simf(subs(connectionformssub,
omega[1,2] &ˆ omega[2] + omega[1,3] &ˆ omega[3])))));
zero2:= Simf(d(Simf(subs(dualformssub, omega[2])))
+ Simf(subs(dualformssub, Simf(subs(connectionformssub,
omega[2,1] &ˆ omega[1] + omega[2,3] &ˆ omega[3])))));
zero3:= Simf(d(Simf(subs(dualformssub, omega[3])))
+ Simf(subs(dualformssub, Simf(subs(connectionformssub,
omega[3,1] &ˆ omega[1] + omega[3,2] &ˆ omega[2])))));
We can set the scalar coefficients of these forms equal to zero and solve for
the (hijk ) to determine the connection forms:
> solve({op(ScalarForm(zero1)), op(ScalarForm(zero2)),
op(ScalarForm(zero3))}, {h[1,2,1], h[1,2,2], h[1,2,3],
h[3,1,1], h[3,1,2], h[3,1,3], h[3,2,1], h[3,2,2],
h[3,2,3]});
{h1,2,1 = −2by, h1,2,2 = 2ax, h1,2,3 = 0, h3,1,1 = 2cz, h3,1,2 = 0, h3,1,3 = −2ax,
h3,2,1 = 0, h3,2,2 = 2cz, h3,2,3 = −2by}
> assign(%);
So here are the connection forms, expressed in terms of the coordinate 1-
forms:
> connectionformssub:= Simf(subs(dualformssub,
Simf(connectionformssub)));
connectionf ormssub :=
2cz d(y) 2by d(z)
ω3,2 = 2 −
ax + by 2 + cz 2 + 1 ax2 + by 2 + cz 2 + 1
Now compute the curvature 2-forms, expressed in terms of the dual forms:
> Omega[2,3]:= Simf(subs(dualformsbacksub,
Simf(d(Simf(subs(connectionformssub, omega[2,3])))
+ Simf(subs(connectionformssub,
omega[2,1] &ˆ omega[1,3])))));
Omega[3,1]:= Simf(subs(dualformsbacksub,
Simf(d(Simf(subs(connectionformssub, omega[3,1])))
+ Simf(subs(connectionformssub,
omega[3,2] &ˆ omega[2,1])))));
Omega[1,2]:= Simf(subs(dualformsbacksub,
Simf(d(Simf(subs(connectionformssub, omega[1,2])))
+ Simf(subs(connectionformssub,
12.7. Maple computations 395
It follows that any totally geodesic surface must be a level surface for one of
the coordinate functions (x, y, z). Suppose that Σ is contained in the level
set {(x, y, z) ∈ M | z = z0 } for some constant z0 and compute the pullbacks
of the connection forms to Σ:
> Form(z0=-1);
> Simf(subs([x=u, y=v, z=z0], connectionformssub));
2bv d(u) 2au d(v)
ω1,2 = − 2 + ,
au + bv 2 + cz02 + 1 au2 + bv 2 + cz02 + 1
2cz0 d(u) 2cz0 d(v)
ω3,1 = 2 2 2
, ω3,2 = 2
au + bv + cz0 + 1 au + bv 2 + cz02 + 1
Since a totally geodesic surface must have ω̄13 = ω̄23 = 0, Σ is not a totally
geodesic surface unless z0 = 0.
Bibliography
[AEG06] Juan A. Aledo, José M. Espinar, and José A. Gálvez, Timelike surfaces in the Lorentz-
Minkowski space with prescribed Gaussian curvature and Gauss map, J. Geom. Phys.
56 (2006), no. 8, 1357–1369.
[B8̈3] A. V. Bäcklund, Om ytor med konstant negativ krökning, Lunds Universitets Årsskrift
19 (1883), 1–48.
[BG80] Richard L. Bishop and Samuel I. Goldberg, Tensor Analysis on Manifolds, Dover
Publications Inc., New York, 1980, corrected reprint of the 1968 original.
[Bia79] Luigi Bianchi, Ricerche sulle superficie elicoidali e sulle superficie a curvatura
costante, Ann. Scuola Norm. Sup. Pisa Cl. Sci. 2 (1879), 285–341.
[Bla23] Wilhelm Blaschke, Vorlesungen über Differentialgeometrie, vol. II, Springer, Berlin,
1923.
[Cal00] James J. Callahan, The Geometry of Spacetime: An Introduction to Special and Gen-
eral Relativity, Undergraduate Texts in Mathematics, Springer-Verlag, New York,
2000.
397
398 Bibliography
[Car04] Élie Cartan, Sur la structure des groupes infinis de transformation, Ann. Sci. École
Norm. Sup. (3) 21 (1904), 153–206.
[Car20] , Sur la déformation projective des surfaces, Ann. Sci. École Norm. Sup. (3)
37 (1920), 259–356.
[Car30] , La théorie des groupes finis et continus et l’analysis situs, Mémorial des
sciences mathématiques, no. 42, Gauthier-Villars et Cie, 1930.
[Car35] , La Méthode du Repère Mobile, la Théorie des Groupes Continus, et les Es-
paces Généralisés, Exposés de Géométrie, no. 5, Hermann, Paris, 1935.
[Car46] , Leçons sur la Géométrie des Espaces de Riemann, 2nd ed., Gauthier-Villars,
Paris, 1946.
[Car92] , Leçons sur la Géométrie Projective Complexe. La Théorie des Groupes Finis
et Continus et la Géométrie Différentielle Traitées Par la Méthode du Repère Mobile.
Leçons sur la Théorie des Espaces à Connexion Projective, Les Grands Classiques
Gauthier-Villars. [Gauthier-Villars Great Classics], Éditions Jacques Gabay, Sceaux,
1992, reprint of the 1931, 1937, and 1937 editions.
[CEM+ 14] Jeanne Clelland, Edward Estrada, Molly May, Jonah Miller, and Sean Peneyra, A
Tale of Two Arc Lengths: Metric Notions for Curves in Surfaces in Equiaffine Space,
Proc. Amer. Math. Soc. 142 (2014), 2543–2558.
[Che42] Shiing-shen Chern, On integral geometry in Klein spaces, Ann. of Math. (2) 43 (1942),
178–189.
[Cle99] Jeanne N. Clelland, Lie groups and the method of moving frames, 1999, lecture
videos from Mathematical Sciences Research Institute Summer Graduate Workshop,
http://www.msri.org/publications/video/index2.html.
[CT80] Shiing Shen Chern and Chuu Lian Terng, An analogue of Bäcklund’s theorem in affine
geometry, Rocky Mountain J. Math. 10 (1980), no. 1, 105–124.
[CT86] Shiing-shen Chern and Keti Tenenblat, Pseudo-spherical surfaces and evolution equa-
tions, Stud. Appl. Math. 74 (1986), 55–83.
[Dac08] Bernard Dacorogna, Introduction to the Calculus of Variations, 2nd ed., Imperial
College Press, London, 2008.
[Dar72a] Gaston Darboux, Leçons sur la Théorie Générale des Surfaces et les Applications
Géométriques du Calcul Infinitésimal. Deuxième partie, Chelsea Publishing Co.,
Bronx, N.Y., 1972, Les Congruences et les Équations Linéaires aux Dérivées Par-
tielles. Les Lignes Tracées sur les Surfaces, réimpression de la deuxième édition de
1915.
[Dar72b] , Leçons sur la Théorie Générale des Surfaces et les Applications Géométriques
du Calcul Infinitésimal. Première partie, Chelsea Publishing Co., Bronx, N.Y., 1972,
Généralités. Coordonnées curvilignes. Surfaces minima, réimpression de la deuxième
édition de 1914.
[Dar72c] , Leçons sur la Théorie Générale des Surfaces et les Applications Géométriques
du Calcul Infinitésimal. Quatrième partie, Chelsea Publishing Co., Bronx, N.Y.,
Bibliography 399
[Dar72d] , Leçons sur la Théorie Générale des Surfaces et les Applications Géométriques
du Calcul Infinitésimal. Troisième partie, Chelsea Publishing Co., Bronx, N.Y., 1972,
Lignes Géodésiques et Courbure Géodésique. Paramètres Différentiels. Déformation
des Surfaces, réimpression de la première édition de 1894.
[Dou31] Jesse Douglas, Solution of the problem of Plateau, Trans. Amer. Math. Soc. 33 (1931),
no. 1, 263–321.
[Eis60] Luther Pfahler Eisenhart, A Treatise on the Differential Geometry of Curves and
Surfaces, Dover Publications Inc., New York, 1960.
[FO98] Mark Fels and Peter J. Olver, Moving coframes. I. A practical algorithm, Acta Appl.
Math. 51 (1998), no. 2, 161–213.
[FO99] , Moving coframes. II. Regularization and theoretical foundations, Acta Appl.
Math. 55 (1999), no. 2, 127–208.
[Fom90] A. T. Fomenko, The Plateau Problem. Part I, Studies in the Development of Mod-
ern Mathematics, vol. 1, Gordon and Breach Science Publishers, New York, 1990,
Historical survey, translated from the Russian.
[Fre47] J. F. Frenet, Sur les Courbes à Double Courbure, Ph.D. thesis, Toulouse, 1847.
[Gál09] José A. Gálvez, Surfaces of constant curvature in 3-dimensional space forms, Mat.
Contemp. 37 (2009), 1–42.
[Gar89] Robert B. Gardner, The Method of Equivalence and Its Applications, CBMS-NSF
Regional Conference Series in Applied Mathematics, vol. 58, Society for Industrial
and Applied Mathematics (SIAM), Philadelphia, PA, 1989.
[Gre78] Mark L. Green, The moving frame, differential invariants and rigidity theorems for
curves in homogeneous spaces, Duke Math. J. 45 (1978), no. 4, 735–779.
[Gri74] P. Griffiths, On Cartan’s method of Lie groups and moving frames as applied to
uniqueness and existence questions in differential geometry, Duke Math. J. 41 (1974),
775–814.
[HCV52] D. Hilbert and S. Cohn-Vossen, Geometry and the Imagination, Chelsea Publishing
Company, New York, N.Y., 1952, translated by P. Neményi.
400 Bibliography
[HH06] Qing Han and Jia-Xing Hong, Isometric Embedding of Riemannian Manifolds in Eu-
clidean Spaces, Mathematical Surveys and Monographs, vol. 130, American Mathe-
matical Society, Providence, RI, 2006.
[IL03] Thomas A. Ivey and J. M. Landsberg, Cartan for Beginners: Differential Geometry via
Moving Frames and Exterior Differential Systems, Graduate Studies in Mathematics,
vol. 61, American Mathematical Society, Providence, RI, 2003.
[Jan26] M. Janet, Sur la possibilité de plonger un espace Riemannien donné dans un espace
euclidien, Ann. Soc. Polon. Math. 5 (1926), 38–43.
[Kle93a] Felix Klein, A comparative review of recent researches in geometry, Bull. Amer. Math.
Soc. 2 (1893), no. 10, 215–249.
[Kog03] Irina Kogan, Two algorithms for a moving frame construction, Canad. J. Math. 55
(2003), 266–291.
[Lee13] John M. Lee, Introduction to Smooth Manifolds, 2nd ed., Graduate Texts in Mathe-
matics, vol. 218, Springer, New York, 2013.
[Man10] Elizabeth Louise Mansfield, A Practical Guide to the Invariant Calculus, Cambridge
Monographs on Applied and Computational Mathematics, vol. 26, Cambridge Univer-
sity Press, Cambridge, 2010.
[Mil83] Tilla Klotz Milnor, Harmonic maps and classical surface theory in Minkowski 3-space,
Trans. Amer. Math. Soc. 280 (1983), no. 1, 161–185.
[MR05] Sebastián Montiel and Antonio Ros, Curves and surfaces, Graduate Studies in Math-
ematics, vol. 69, American Mathematical Society, Providence, RI; Real Sociedad
Matemática Española, Madrid, 2005, translated and updated from the 1998 Span-
ish edition by the authors.
[Nas56] John Nash, The imbedding problem for Riemannian manifolds, Ann. of Math. (2) 63
(1956), 20–63.
[NMS08] Mehdi Nadjafikhah and Ali Mahdipour Sh., Affine classification of n-curves, Balkan
J. Geom. Appl. 13 (2008), no. 2, 66–73.
[NS94] Katsumi Nomizu and Takeshi Sasaki, Affine Differential Geometry, Geometry of affine
immersions, Cambridge Tracts in Mathematics, vol. 111, Cambridge University Press,
Cambridge, 1994.
[Olv00] Peter Olver, Applications of Lie Groups to Differential Equations, 2nd ed., Graduate
Texts in Mathematics, Springer-Verlag, New York, 2000.
Bibliography 401
[Olv10] , Recent advances in the theory and application of Lie pseudo-groups, XVIII
International Fall Workshop on Geometry and Physics, AIP Conf. Proc., vol. 1260,
Amer. Inst. Phys., Melville, NY, 2010, pp. 35–63.
[Opr07] John Oprea, Differential Geometry and Its Applications, 2nd ed., Classroom Resource
Materials Series, Mathematical Association of America, Washington, DC, 2007.
[OT05] V. Ovsienko and S. Tabachnikov, Projective Differential Geometry Old and New, Cam-
bridge Tracts in Mathematics, vol. 165, Cambridge University Press, Cambridge, 2005.
[Rad30] Tibor Radó, On Plateau’s problem, Ann. of Math. (2) 31 (1930), no. 3, 457–469.
[RS02] C. Rogers and W. K. Schief, Bäcklund and Darboux Transformations, Geometry and
modern applications in soliton theory, Cambridge Texts in Applied Mathematics, Cam-
bridge University Press, Cambridge, 2002.
[Ser51] J. Serret, Mémoire sur quelques formules relatives à la théorie des courbes à double
courbure, J. Math. Pures Appl. 16 (1851), 193–207.
[She99] Mary D. Shepherd, Line congruences as surfaces in the space of lines, Differential
Geom. Appl. 10 (1999), no. 1, 1–26.
[Ste07] James Stewart, Calculus: Early Transcendentals, 6th ed., Stewart’s Calculus Series,
Brooks Cole, Belmont, CA, 2007.
[Su83] Bu Chin Su, Affine Differential Geometry, Science Press, Beijing, 1983.
[Wei66] K. Weierstrass, Über die Flächen deren mittlere Krümmung überall gleich null ist.,
Monatsber. Berliner Akad. (1866), 612–625, 855–856.
402 Bibliography
[Wil61] T. J. Willmore, The definition of Lie derivative, Proc. Edinburgh Math. Soc. (2) 12
(1960/1961), 27–29.
403
404 Index
www.ams.org
GSM/178