MX Cartan For Beginners Differential Geometry Via Mov
MX Cartan For Beginners Differential Geometry Via Mov
I N M AT H E M AT I C S 175
Thomas A. Ivey
Joseph M. Landsberg
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
https://doi.org/10.1090//gsm/175
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
GRADUATE STUDIES
I N M AT H E M AT I C S 175
Thomas A. Ivey
Joseph M. Landsberg
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
EDITORIAL COMMITTEE
Dan Abramovich
Daniel S. Freed (Chair)
Gigliola Staffilani
Jeff A. Viaclovsky
2010 Mathematics Subject Classification. Primary 35N10, 37K35, 53A05, 53A20, 53A30,
53A55, 53C10, 53C15, 58A15, 58A17.
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 2016 by the American Mathematical Society. All rights reserved.
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 21 20 19 18 17 16
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
Contents
Introduction xi
Preface to the Second Edition xi
Preface to the First Edition xiv
Chapter 1. Moving Frames and Exterior Differential Systems 1
§1.1. Geometry of surfaces in E3 in coordinates 2
§1.2. Differential equations in coordinates 5
§1.3. Introduction to differential equations without coordinates 8
§1.4. Introduction to geometry without coordinates:
curves in E2 13
§1.5. Submanifolds of homogeneous spaces 16
§1.6. The Maurer-Cartan form 18
§1.7. Plane curves in other geometries 22
§1.8. Curves in E3 25
§1.9. Grassmannians 28
§1.10. Exterior differential systems and jet spaces 31
Chapter 2. Euclidean Geometry 39
§2.1. Gauss and mean curvature via frames 40
§2.2. Calculation of H and K for some examples 43
§2.3. Darboux frames and applications 46
§2.4. What do H and K tell us? 47
§2.5. Invariants for n-dimensional submanifolds of En+s 48
§2.6. Intrinsic and extrinsic geometry 52
v
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
vi Contents
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
Introduction
xi
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
xii Introduction
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
Preface to the Second Edition xiii
Dependence of Chapters
1
2 5 7.1 4
3 6
7.2–7.5 12
8
9
10
11
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
xiv Introduction
We define exterior differential systems and jet spaces, and explain how
to rephrase any system of partial differential equations as an EDS using jets.
We state and prove the Frobenius system, leading up to it via an elementary
example of an overdetermined system of PDE.
There are four appendices, covering background material for the main
part of the book: linear algebra and rudiments of representation theory,
differential forms and vector fields, complex and almost complex manifolds,
and a brief discussion of initial value problems and the Cauchy-Kowalevski
Theorem, of which the Cartan-Kähler Theorem is a generalization.
5 Now Chapter 8
6 Now Chapter 7
7 Now Chapter 9
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
Preface to the First Edition xvii
Further reading on EDS. To our knowledge, there are only a small num-
ber of textbooks on exterior differential systems. The first is Cartan’s clas-
sic text [40], which has an extraordinarily beautiful collection of examples,
some of which are reproduced here. We learned the subject from our teacher
Bryant and the book by Bryant, Chern, Griffiths, Gardner and Goldschmidt
[27], which is an elaboration of an earlier monograph [26], and is at a more
advanced level than this book. One text at a comparable level to this book,
but more formal in approach, is [190]. The monograph [86], which is cen-
tered around the isometric embedding problem, is similar in spirit but covers
less material. The memoir [189] is dedicated to extending the Cartan-Kähler
Theorem to the C ∞ setting for hyperbolic systems, but contains an exposi-
tion of the general theory. There is also a monograph by Kähler [109] and
lectures by Kuranishi [119], as well the survey articles [82, 110]. Some dis-
cussion of the theory may be found in the differential geometry texts [174]
and [177].
We give references for other topics discussed in the book in the text.
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
https://doi.org/10.1090//gsm/175/01
Chapter 1
1
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
2 1. Moving Frames and Exterior Differential Systems
Once one has found invariants for a given submanifold geometry, one
may ask questions about submanifolds with special invariants. For surfaces
in E3 , one might ask which surfaces have K constant or H constant. These
can be treated as questions about solutions to certain partial differential
equations (PDE). For example, from (1.3) we see that surfaces with K = 1
are locally given by solutions to the PDE
(1.4) zxx zyy − zxy
2
= (1 + zx2 + zy2 )2 .
We will soon free ourselves of coordinates and use moving frames and dif-
ferential forms. As a provisional definition, a moving frame is a smoothly
varying basis of the tangent space to E3 defined at each point of our sur-
face. In general, using moving frames one can obtain formulas valid at every
point, analogous to coordinate formulas valid at just one preferred point. In
the present context, the Gauss and mean curvatures will be described at all
points by expressions like (1.2) rather than (1.3); see §2.1.
Another reason to use moving frames is that the method gives a uni-
form procedure for dealing with diverse geometric settings. Even if one is
originally only interested in Euclidean geometry, other geometries arise nat-
urally. For example, consider the warp of a surface, which is defined to be
(k1 − k2 )2 , where the kj are the eigenvalues of (1.1). It turns out that this
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
1.2. Differential equations in coordinates 5
arise because mixed partials must commute, i.e., (ux )y = (uy )x . In our
example,
∂ ∂u
(ux )y = A(x, y, u) = Ay (x, y, u) + Au (x, y, u) = Ay + BAu ,
∂y ∂y
(uy )x = Bx + ABu ,
so setting (ux )y = (uy )x reveals a “hidden equation”, the compatibility
condition
(1.9) Ay + BAu = Bx + ABu .
We will prove in §1.10 that the commuting of second-order partials in this
case implies that all higher-order mixed partials commute as well, so that
there are no further hidden equations. In other words, if (1.9) is an identity
in x, y, u, then solving the ODE’s (1.7) and (1.8) in succession gives a solution
to (1.6), and solutions depend on one constant.
Exercise 1.2.4: Show that, if (1.9) is an identity, then one gets the same
solution by first solving for ũ(y) = u(0, y).
2. Show that the conditions for the span of differential forms being Fro-
benius is equivalent to the Frobenius condition for the corresponding dis-
tribution, i.e., Γ(Δ) ⊂ Γ(T Σ) satisfies [Γ(Δ), Γ(Δ)] ⊆ Γ(Δ) if and only if
Δ⊥ ⊂ T ∗ Σ satisfies dθ ≡ 0 mod Δ⊥ for all θ ∈ Γ(Δ⊥ ).
3. On R3 let θ = Adx + Bdy + Cdz, where A = A(x, y, z), etc. Assume
the differential ideal generated by θ is Frobenius, and explain how to find a
function f (x, y, z) such that the differential systems {θ}diff and {df }diff are
equivalent.
C : R → ASO(2),
1 0
t → .
c(t) (e1 (t), e2 (t))
because |e1 (t)| ≡ 1 (see Exercise 1.4.2.1 below). Thus λ(t) is a differential
invariant, but it again depends on the parametrization of the curve. To
determine an invariant of the image alone, we let c̃(t) = c(φ(t)) be another
parametrization of the same curve for some regular φ : R → R. We calculate
ṽ(t)
that λ̃(t) = v(φ(t)) λ(φ(t)), so the quantity κ(t) = λ(t)
v(t) is unchanged under
reparametrization. This κ(t), called the curvature of the curve, measures
how much c is infinitesimally moving away from its tangent line at c(t).
The above discussion implies that a necessary condition for two curves
c, c̃ to have congruent images (i.e., differing by a Euclidean motion) is that
there exists a diffeomorphism ψ : R → R such that κ(t) = κ̃(ψ(t)). It will
follow from Corollary 1.6.15 that the images of curves are locally classified
up to congruence by their curvature functions, and that parametrized curves
are locally classified by κ, v.
Exercises 1.4.2:
1. Let V be a vector space with a nondegenerate inner product , . Let
v(t) be a curve in V such that F (t) := v(t), v(t) is constant. Show that
v (t), v(t) = 0 for all t. Show the converse is also true.
t
2. Suppose that c is regular. Let s(t) = 0 |c (τ )|dτ and consider c paramet-
rized by s instead of t. Since s gives the length of the image of c : [0, s] → E2 ,
s is called an arclength parameter. Show that in this preferred parametriza-
de1
tion κ(s) = e2 , ds .
3. Show that κ(t) is constant if and only if the curve is an open subset of
a line (if κ = 0) or circle of radius κ1 .
4. Let c(t) = (x(t), y(t)) be given in coordinates. Calculate κ(t) in terms
of x(t), y(t) and their derivatives.
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
16 1. Moving Frames and Exterior Differential Systems
5. Calculate the function κ(t) for an ellipse. Characterize the points on the
ellipse where the maximum and minimum values of κ(t) occur.
6. Can κ(t) be unbounded if c(t) is the graph of a polynomial?
A homogeneous space G/H has a natural left G-action on it; the sub-
group stabilizing the coset [e] of the identity element is H, and the stabilizer
of any point is conjugate to H. Conversely, a manifold X is a homogeneous
space if it admits a smooth transitive action by a Lie group G. If H is the
isotropy group of a point x0 ∈ X, then X G/H, and x0 corresponds to
[e] ∈ G/H. (See [94, 174] for additional facts about homogeneous spaces.)
In the spirit of Klein’s Erlanger Programm (see [93, 112] for historical
accounts), we will consider G as the group of motions of G/H. We will
study the geometry of submanifolds M ⊂ G/H, where two submanifolds
M, M ⊂ G/H will be considered equivalent if there exists a g ∈ G such that
g(M ) = M .
To determine necessary conditions for equivalence we will find differential
invariants as we did in §1.1 and §1.4. (Note that we need to specify whether
we are interested in invariants of a mapping or just of the image.) After
finding invariants, we will then interpret them as we did in the exercises in
§1.4.
We begin to derive invariants for maps f : M → G/H by constructing
lifts F : M → G as we did for curves in E2 .
Definition 1.5.3. A lift of a mapping f : M → G/H is defined to be a map
F : M → G such that the following diagram commutes:
G
F
?
M -
f
G/H
By associating the value of the lift F (x) with the action of F (x)−1 on
G/H, we may think of choosing a lift to G as analogous to putting a point
p on a surface in a normalized position, as we did in §1.1.
Given f : M → G/H, we will choose lifts adapted to the infinitesimal
geometry. To explain what this statement means, we first remark that
in the situations we will be dealing with, the fiber at x ∈ G/H of the
fibration π : G → G/H admits the interpretation of being a subset of the
space of frames, i.e., bases of Tx G/H. Since H fixes the point [e] ∈ G/H,
its infinitesimal action on tangent vectors gives a representation ρ : H →
GL(T[e] G/H), called the isotropy representation [187]. Now fix a reference
basis (v1 , . . . , vn ) of T[e] G/H. We identify the H-orbit of this basis (induced
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
18 1. Moving Frames and Exterior Differential Systems
Remark 1.6.4. The above proof shows more generally that a left-invariant
vector-valued form α ∈ Ωk (G, V ) is uniquely determined by αg for any
g ∈ G. In this way, the set of left-invariant k-forms may be identified with
Λk g∗ .
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
1.6. The Maurer-Cartan form 19
Example 1.6.5. Let G ⊆ GL(N, R) be a matrix Lie group with Lie algebra
g ⊂ gl(N, R), and let g = (gji ) be the matrix-valued function which embeds
G into the vector space MN ×N of N × N matrices with real entries, with
differential dga : Ta G → Tg(a) MN ×N MN ×N . The Maurer-Cartan form of
G is ωa = L∗a−1 (ωe ) = g(a)−1 dga , which is often written as
ω = g −1 dg.
In more detail, writing g = (gji ), we consider each gji : G → R as a func-
tion on G, and thus have the function (gji ) : G → MN ×N , so (ωa )ij =
(g(a)−1 )ik (dga )kj .
Definition 1.6.8. Let ω = (ωki ) and η = (ηki ) be matrices whose entries are
elements of a vector space V , so that ω, η ∈ V ⊗Mn×n . Define their matrix
wedge product ω ∧ η ∈ Λ2 V ⊗Mn×n by
(ω ∧ η)ij := ωki ∧ ηjk .
by the forms θji . Submanifolds of dimension n to which these forms pull back
to be zero are graphs of maps f : M → G such that φ = f ∗ ω. We check the
conditions given in the Frobenius Theorem. Calculating derivatives (and
omitting the pullback notation), we have
dθ = −φ ∧ φ + ω ∧ ω
= −φ ∧ φ + (θ − φ) ∧ (θ − φ)
≡ 0 mod I.
Thus, the system is Frobenius and by Theorem 1.3.7 there is a unique con-
nected maximal n-dimensional integral manifold through any (x, g) ∈ Σ.
Suppose f1 , f2 are two different solutions. Say f1 (x) = g. Let a =
gf2 (x)−1 . Then the graph of f = La ◦ f2 passes through (x, g) and f ∗ ω = φ.
By uniqueness, it follows that f1 = La ◦ f2 .
Remark 1.6.14. If we assume in Theorem 1.6.13 that M is connected and
simply connected, then the desired map f may be extended to all of M
[187].
The group of FLT’s is P SL(2, C) := SL(2, C)/{± Id}, which acts tran-
sitively on CP1 . Thus, CP1 is a homogeneous space P SL(2, C)/P , where
a b
P = ∈ P SL(2, C)
0 a−1
is the stabilizer of t[1, 0]. Although P SL(2, C), as presented, is not a matrix
Lie group, we may avoid problems by working locally as follows:
If Δ ⊂ C ⊂ CP1 is a domain, then P SL(2, C) acts on maps f : Δ → C
by
af + b
. f →
cf + d
(Since we will be working locally, there is no harm in considering f as a map
to C; then to think of f as a map to CP1 , write it as t[f, 1].)
Suppose f, g : Δ → C are two holomorphic maps with nonzero first
derivatives. When are these locally equivalent via a fractional linear trans-
formation, i.e., when does g = A ◦ f for some FLT A? (One can ask the
same question in the real category for analytic maps f, g : (0, 1) → RP1 ,
and the answer will be the same.) Note that in this example the target is of
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
1.7. Plane curves in other geometries 23
Exercises 1.7.2:
1. Show that f is an FLT if and only if Sf ≡ 0. So, just as the curvature
of a curve in R2 measures the failure of a curve to be a line, Sf (z) is an
infinitesimal measure of the failure of a holomorphic map to be an FLT.
Since FLT’s map circles to circles, Sf may be thought of as measuring how
much circles are being distorted under f .
2. Calculate Sf for f (z) = aebz , and f (z) = z n . How do these compare
asymptotically? What does this say about how circles are distorted as one
goes out to infinity?
acting on R2 in the same way as ASO(2) acts in §1.4. Since the origin is
fixed by the subgroup SL(2, R), in this context we relabel R2 as the special
affine plane SA2 = ASL(2, R)/SL(2, R).
(a) Find differential invariants for curves in SA2 . (As with the Euclidean
case, one can consider invariants of a parametrized curve or invariants of
just the image curve.)
(b) What are the image curves with invariants zero? The image curves with
constant invariants?
(c) Let κA denote the differential invariant that distinguishes image curves.
Interpret κA (t) in terms of an osculating curve, as we did with the osculating
circles to a curve in §1.4.
(d) The preferred frame will lead to a unique choice of e2 . Give a geometric
interpretation of e2 .
2. (Curves in the projective plane) Carry out the analogous exercise for
curves in the projective plane P2 = GL(3)/P , where P is the subgroup pre-
serving a line. Show that the curves with zero invariants are the projective
lines and plane conics. Derive the Monge equation (y )−2/3 = 0, charac-
terizing plane conics among graphs y = f (x), by working in a local adapted
coordinate system. Note that one may do this exercise over R or C.
3. Carry out the analogous exercise for curves in the conformal plane
ACO(2)/CO(2), where equivalence is up to translations, rotations and di-
lations.
4. Carry out the analogous exercise for curves in L2 = ASO(1, 1)/SO(1, 1),
where SO(1, 1) is the subgroup of GL(2, R) preserving the quadratic form
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
1.8. Curves in E3 25
Q(v, w) = t v −1 0
0 1 w. Note that there will be three distinct types of
curves: spacelike curves, where Q(c (t), c (t)) > 0; timelike curves, where
Q(c (t), c (t)) < 0; and lightlike curves, where Q(c (t), c (t)) = 0. What are
the curves with constant invariants?
1.8. Curves in E3
The group ASO(3) and its Maurer-Cartan form. The group ASO(3)
is the set of transformations of E3 of the form x → t + Rx, i.e.,
⎛ 1⎞ ⎛ 1⎞ ⎛ 1⎞
x t x
⎝x2 ⎠ → ⎝t2 ⎠ + R ⎝x2 ⎠ ,
x3 t3 x3
where R ∈ SO(3) is a rotation matrix. Like ASO(2), it may be represented
as a matrix Lie group by writing
1 0
(1.23) ASO(3) = M ∈ GL(4, R) M = , t ∈ R3 , R ∈ SO(3) .
t R
The action on E3 is given by x̂ → M x̂, where we represent points in E3 by
x = t (1, x1 , x2 , x3 ).
Having expressed ASO(3) as in (1.23), we may express an arbitrary
element of its Lie algebra aso(3) as
⎛ ⎞
0 0 0 0
⎜x1 0 −x2 −x3 ⎟
⎜ 1 1⎟ , ai , aij ∈ R.
⎝x2 x21 0 −x32 ⎠
x3 x31 x32 0
In this presentation, the Maurer-Cartan form of ASO(3) is
⎛ ⎞
0 0 0 0
⎜ω 1 0 −ω12 −ω13 ⎟
(1.24) ω=⎜⎝ω 2 ω12
⎟,
0 −ω23 ⎠
ω 3 ω13 ω23 0
where ω i , ωji ∈ Ω1 (ASO(3)). Recall from §1.6 that the forms ω i , ωji are left-
invariant and are a basis for the space of left-invariant 1-forms on ASO(3).
We identify ASO(3) with the space of oriented orthonormal frames of
E3 , as follows. Given a point x ∈ E3 and an oriented orthonormal basis
(e1 , e2 , e3 ) of Tx E3 , let
1 0
(1.25) g= ∈ ASO(3),
x R
where R = (e1 , e2 , e3 ) is a rotation matrix, and x = (x1 , x2 , x3 ) corresponds
to translation.
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
26 1. Moving Frames and Exterior Differential Systems
−ω 12 e2
ω 21
−e1 e1
ω 12
−e2 −ω 12
1.9. Grassmannians
Perhaps the most important manifold is the Grassmannian G(k, V ) which
parametrizes k-dimensional subspaces of a vector space V . It plays a cen-
tral role in differential geometry, algebraic geometry, algebraic topology and
representation theory and helps bring these subjects together.
Fix index ranges 1 ≤ i, j ≤ k, and k + 1 ≤ s, t, u ≤ n.
Let V be a vector space over R or C and let G(k, V ) denote the Grass-
mannian of k-planes that pass through the origin in V . To specify a k-plane
E, it is sufficient to specify a basis v1 , . . . , vk of E. We continue our nota-
tional convention that {v1 , . . . , vk } denotes the span of the vectors v1 , . . . , vk .
After fixing a reference basis, we identify GL(V ) with the set of bases for V
by associating to g ∈ GL(V ) the columns of the matrix representing it, and
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
1.9. Grassmannians 29
define a map
π : GL(V ) → G(k, V ),
(e1 , . . . , en ) → {e1 , . . . , ek }.
If we let ẽ1 , . . . , ẽn denote a reference basis of V , the fiber of this mapping
over π(Id) = {ẽ1 , . . . , ẽk } is the subgroup
i
g gi
Pk = g = j st | det(g) = 0 ⊂ GL(V ).
0 gs
More generally, for g ∈ GL(V ), π −1 (π(g)) = gPk , so the Grassmannian has
the structure of a homogeneous space.
Of particular importance is projective space PV = G(1, V ), the space of
all lines through the origin in V . When V = CN we write PN −1 = P(CN ).
We define a line in PV to be the projectivization of a two-dimensional linear
subspace of V and we define a (k − 1)-plane in PV as the projectivization
of a k-dimensional linear subspace in V . We may thus consider G(k, V ) as
the space of all Pk−1 ’s in PV .
Notation. If Y ⊂ PV , we let Ŷ ⊂ V denote its pre-image under the
projection V \{0} → PV , called the cone over Y . If Z ⊂ V , we let [Z] =
π(Z\{0}).
Exercises 1.9.1:
1. Show there is a canonical isomorphism G(k, V ) G(n − k, V ∗ ). In
particular, PV ∗ is the space of hyperplanes (i.e., Pn−1 ’s) in PV .
2. Show that the following map, called the Plücker embedding of the Grass-
mannian, is well-defined:
G(k, V ) → P(Λk V ),
{e1 , . . . , ek } → [e1 ∧ . . . ∧ ek ].
Proposition 1.9.2. Let E ∈ G(k, V ). There is a canonical identification
TE G(k, V ) E ∗ ⊗(V /E).
We consider the term corresponding to a fixed l: If el (0) ∈ {e1 , . . . , el−1 , el+1 ,
. . . , ek }, the term is zero. If el (0) is a multiple of el , then the term in E
(0)
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
30 1. Moving Frames and Exterior Differential Systems
which descends to an element of TE G(k, V ). Now one just checks that this
correspondence is independent of our choices of bases and representative
curves el (t) with el (0) = f (el ).
Proof. A curve c(t) = (e1 (t), . . . , en (t)) ⊂ GL(V ) is vertical for the projec-
tion π : GL(V ) → G(k, V ) if and only if
ej (t) ⊂ {e1 (t), . . . , ek (t)}
for all 1 ≤ j ≤ k. But dej = ei ωji + es ωjs , and thus on a curve c(t),
dej
= ei ωji (c (t)) + es ωjs (c (t)).
dt
So, c(t) is vertical if and only if ωjs (c (t)) = 0 for all s, j.
Second proof of 1.9.2. Let (ej , es ) be the dual basis to (ei , et ). Write
E = {ej }. For f = (e1 , . . . , en ) ∈ GL(V ), consider the tensor
Lf := ωis ⊗ei ⊗(es mod E) ∈ Tf∗ GL(V )⊗(E ∗ ⊗V /E).
In Exercise 1.9.4 you will show that L is basic for the projection π, so L
descends to a well-defined linear map TE G(k, V ) → E ∗ ⊗V /E, which is an
isomorphism because the ωis are semi-basic and independent, and the two
vector spaces have the same dimension.
Exercise 1.9.4: Show that L is indeed basic, by showing that the actions
of Pk on the three terms cancel each other.
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
1.10. Exterior differential systems and jet spaces 31
Third proof of 1.9.2. Recall from Exercise 1.6.1 that Tπ(Id) G/P g/p.
In the case of G = GL(V ), we have g = V ∗ ⊗ V . Let E0 = π(Id) ∈
G(k, V ); then p (E0∗ ⊗ E0 ) ⊕ ((V /E0 )∗ ⊗ E0 ) ⊕ ((V /E0 )∗ ⊗ (V /E0 )). (If
this confuses you, look at the block form of Pk above.) Writing V ∗ ⊗V =
(E0∗ + (V /E0 )∗ )⊗(E0 + V /E0 ), we obtain g/p E0∗⊗(V /E0 ). In fact we may
drop the subscript 0, because the tangent space at g ∈ GL(V ) of the fiber
over E = π(g) is Lg∗ p, and the conjugate of g = gl(V ) is still isomorphic to
V ∗ ⊗V .
Jets.
Definition 1.10.5. Let t be a coordinate on R and let k ≥ 0. Two differ-
entiable maps f, g : R → R with f (0) = g(0) = 0 are said to have the same
k-jet at 0 if
df dg d2 f d2 g dk f dk g
(0) = (0), 2
(0) = 2 (0), . . . , k (0) = k (0).
dt dt dt dt dt dt
Let M, N be differentiable manifolds and f, g : M → N be two maps.
Then f and g are said to have the same k-jet at p ∈ M if
i. f (p) = g(p) = q, and
ii. for all maps u : R → M and v : N → R with u(0) = p and v(q) = 0,
the differentiable maps v ◦ f ◦ u and v ◦ g ◦ u have the same k-jet at 0.
Exercise 1.10.6: Show that to determine if f, g : M → N have the same
k-jet at p, it is sufficient to check derivatives up to order k with respect to
coordinate directions in any pair of local coordinate systems around p and
q.
which we will call contact forms. (Note the summation on ik .) We will use
multi-index notation to abbreviate the forms in (1.29) as
Proof. We follow the proof in [49], as that proof will get us accustomed to
calculations with differential forms.
We proceed by induction on n, the rank of the distribution annihilated
by I. If n = 1, then the distribution defines a line field and we are done by
Theorem 1.3.1. Assume that the theorem is true up to n − 1, and we will
show that it is true for n.
Let x : M → R be a smooth function such that θ1 ∧ . . . ∧ θm−n ∧ dx = 0
on a neighborhood U of p, and consider the ideal I = {θ1 , . . . , θm−n , dx}diff .
Since I = {θ1 , . . . , θm−n }diff is Frobenius, I is also Frobenius. By our
induction hypothesis, there exist local coordinates (y 1 , . . . , y m ) such that
I = {dy 1 , . . . , dy m−n+1 }alg .
At this point we have an (n − 1)-dimensional integral manifold of I
(hence, also of I) passing through p, obtained by setting y 1 , . . . , y m−n+1
equal to the appropriate constants. We want to enlarge it to an n-dimension-
al integral manifold.
Let 1 ≤ i, j ≤ m − n. We may write
dx = ai dy i + am−n+1 dy m−n+1 ,
θi = cij dy j + cim−n+1 dy m−n+1 ,
where ai , am−n+1 , cij , cim−n+1 are smooth functions. Without loss of gener-
ality, we may assume an−m+1 = ∂x/∂y m−n+1 = 0, so we may rewrite the
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
1.10. Exterior differential systems and jet spaces 35
second line as
θi = c̃ij dy j + f i dx
for some smooth functions c̃ij , f i . The matrix of functions c̃ij is invertible at
each point in a (possibly smaller) neighborhood Ũ of p, so locally we may
take a new set of generators for I, of the form
θ̃i = dy i + ei dx
for some smooth functions ei , and with θi = c̃ij θ̃j . Then dθ̃i = dei ∧ dx and,
since I is Frobenius,
dei ∧ dx ≡ 0 mod {θ̃j }.
Hence
dei = adx + bij dy j
for some functions a, bij . In particular, the ei are functions of the y j and
x only, and it follows that the θ̃i are defined in terms of the variables
y 1 , . . . , y m−n+1 only.
Let V ⊂ Ũ be the submanifold through p obtained by setting y m−n+2
through y m constant. Then I|V is a codimension-one Frobenius system.
Hence there are coordinates (ỹ 1 , . . . , ỹ m−n+1 ) on V that are functions of
the y 1 , . . . , y m−n+1 , such that I|V is generated by dỹ 1 , . . . , dỹ m−n . These
relationships extend to Ũ , so that
(ỹ 1 , . . . , ỹ m−n+1 , y m−n+2 , . . . , y m )
is the desired coordinate system.
For a proof of the maximality assertion in the first version of the Frobe-
nius Theorem (1.3.7), see [187, Thm. 1.6.4].
Contact manifolds.
Exercise 1.10.15: Consider the contact system on J 1 (Rn , R) = R2n+1 with
coordinates (z, x1 , . . . , xn , y 1 , . . . , y n ). Here θ = dz − Σi y i dxi generates the
contact system I = {θ}diff .
(a) Show that at any point ∂
∂y 1
∧ ...∧ ∂
∂y n is an integral element.
(b) At which points is ∂
∂x1
∧ . . . ∧ ∂x∂n an integral element?
(c) Show that any graph z = h(x1 , . . . , xn ), y j = f j (x1 , . . . , xn )
such that
f j = ∂h/∂xj is an integral manifold. (In fact, all n-dimensional integral
manifolds are locally of this form.)
(d) Show that there are no (n + 1)-dimensional integral elements for I.
y 0 dy 1 + y 2 dy 3 + . . . + y 2r dy 2r+1 if (dθ)r+1 = 0,
θ=
dy 1 + y 2 dy 3 + . . . + y 2r dy 2r−1 if (dθ)r+1 = 0,
on U .
The integer r is sometimes referred to as the rank of θ [27, II.3]; to avoid
confusion with the rank of a Pfaffian system, we will call r the Pfaff rank.
If n = 2r, then Pfaff’s Theorem implies that M is locally diffeomorphic to
the jet bundle J 1 (Rr , R) and θ̃ is the pullback of the standard contact form.
Thus, r-dimensional integral manifolds of I are given in the coordinates
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
38 1. Moving Frames and Exterior Differential Systems
w0 , . . . , wn by
w0 = f (w1 , . . . , wr ),
∂f
wr+1 = ,
∂w1
..
.
∂f
. w2r =
∂wr
A 1-form θ on a (2n + 1)-dimensional manifold Σ is called a contact form if
it is as nondegenerate as possible, i.e., if θ ∧ (dθ)n is nonvanishing.
Since we use the 1-form θ to define an EDS, we really only care about
it up to multiplication by a nonvanishing function. A contact manifold is
defined to be a manifold with a contact form, defined up to scale—more
precisely, a rank-one subbundle I ⊂ T ∗ M any nonvanishing section of which
is a contact form. This generalizes the contact structure on J 1 (M, R), our
first example of a contact manifold.
Example 1.10.17. The projectivized tangent bundle PT M may be given
the structure of a contact manifold by taking the distribution α⊥ ⊂ T (T M )
annihilated by the tautological form and projecting to PT M .
Exercise 1.10.18 (Normal form for degenerate contact forms): On R3 , con-
sider a 1-form θ such that θ ∧ dθ = f Ω, where Ω is a volume form and f
is a function such that df |p = 0 whenever f (p) = 0. Show that there are
coordinates (x, y, z), in a neighborhood of any such point where θp = 0, such
that θ = dz − y 2 dx.
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
https://doi.org/10.1090//gsm/175/02
Chapter 2
Euclidean Geometry
39
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
40 2. Euclidean Geometry
order, and (2.2) implies that to first order x may move independently along
the span of e1 and e2 .
Let π : F 1 → U denote the bundle whose fiber over x ∈ U is the set of
oriented orthonormal bases (e1 , e2 , e3 ) of Tx E3 such that e3 ⊥ Tf (x) M . The
first-order adapted lifts are exactly the sections of F 1 . By fixing a reference
frame at the origin in E3 , ASO(3) may be identified as the bundle of all
oriented orthonormal frames of E3 , and F 1 is a subbundle of f ∗ (ASO(3)).
Throughout this chapter we will not distinguish between U and M when
the distinction is unimportant. In particular, we will usually consider F 1 as
a bundle over M .
Consequences of our adaptation. Thanks to the Maurer-Cartan equation
(1.19), we may calculate the derivatives of the left-invariant forms on
ASO(3) algebraically:
⎛ ⎞
0 0 0 0
⎜ω 1 0 −ω12 −ω13 ⎟
d⎜
⎝ω 2 ω12
⎟
0 −ω23 ⎠
ω 3 ω13 ω23 0
⎛ ⎞ ⎛ ⎞
0 0 0 0 0 0 0 0
⎜ω 1 0 −ω 2 −ω 3 ⎟ ⎜ω 1 0 −ω 2 −ω 3 ⎟
= −⎜⎝ω 2 ω12
1 1⎟ ∧ ⎜ 1 1⎟ .
0 −ω23 ⎠ ⎝ω 2 ω12 0 −ω23 ⎠
ω 3 ω13 ω23 0 ω 3 ω13 ω23 0
Write i : F 1 → ASO(3) for the inclusion map. By our definition of F 1 ,
i∗ ω 3 = 0, and hence
(2.3) 0 = i∗ (dω 3 ) = −i∗ (ω13 ∧ ω 1 + ω23 ∧ ω 2 ).
By (2.2), i∗ ω 1 and i∗ ω 2 are independent, and we can apply the Cartan
Lemma A.1.9 to the right hand side of (2.3). We obtain
3
∗ ω1 h11 h12 ∗ ω 1
(2.4) i = i ,
ω23 h21 h22 ω2
where hij = hji are some functions defined on F 1 . This h = (hij ) is analo-
gous to the Hessian at the origin in (1.1), but it has the advantage of being
defined on all of F 1 .
Given an adapted lift F : U → F 1 , we have
3 1
∗ ω1 ∗ ω
F = hF F ,
ω23 ω2
where hF = F ∗ (h). We now determine the invariance of hF . Since F
is uniquely defined up to a rotation in the tangent plane to M , all other
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
42 2. Euclidean Geometry
Exercises 2.1.2:
cos θ sin θ
1. Let F̃ be as in (2.5) and let R = . Calculate F̃ ∗ ω12 in
− sin θ cos θ
terms of θ and compare with F ∗ ω12 .
2. Show that H, K are invariants of the image of f .
3. Express k1 , k2 in terms of H and K.
Surfaces with H identically zero are called minimal surfaces and are
discussed in more detail in §3.2 and §7.4.
Developable surfaces. A surface M 2 ⊂ E3 is said to be a tangential devel-
opable if it is describable as (a subset of) the union of tangent rays to a
curve. (These surfaces are also called tangential surfaces.)
Let c : R → E3 be a regular parametrized curve, and consider the surface
f : R2 → E3 defined by (u, v) → c(u) + vc (u). Since
df = (c (u) + vc (u))du + c (u)dv,
we see that f is regular, i.e., df is of maximal rank, when c (u) is linearly
independent from c (u) and v = 0. (We will assume v > 0.) Note that the
tangent space, spanned by c (u) and c (u), is independent of v.
Assume that c is parametrized by arclength. Then c (u), c (u) = 0,
and we can take
e1 (u, v) = c (u),
e2 (u, v) = c (u)/||c (u)||.
Using df = ω 1 e1 + ω 2 e2 , we calculate
ω 1 = du + dv,
ω 2 = vκ(u)du,
where κ(u) is the curvature of c.
Note that our frame is the same as if we were to take an adapted framing
of c (as in §1.8), so we have
de3 = (−τ (u)e2 )du.
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
2.2. Calculation of H and K for some examples 45
Thus,
τ (u) 2
ω13 = 0, ω23 = ω ,
κ(u)v
τ (u)
showing that H(u, v) = 2κ(u)v and K ≡ 0.
Surfaces with K identically zero are called flat, and we study their ge-
ometry more in §2.4.
Developable surfaces are also examples of ruled surfaces (as is the heli-
coid). A surface is ruled if through any point of the surface there passes a
straight line (or line segment) contained in the surface.
Surfaces of revolution. Let U ⊂ R2 be an open set with coordinates u, v and
let f : U → E3 be a map of the form
x(u, v) = r(v) cos(u),
y(u, v) = r(v) sin(u),
z(u, v) = t(v),
where r, t are smooth functions. The resulting surface is called a surface of
revolution because it is constructed by rotating a generating curve (e.g., in
the xz-plane) about the z-axis. Call the image M .
Assuming that the generating curve is regular, we can choose v to be an
arclength parameter, so that (r (v))2 + (t (v))2 = 1. Let
e1 = (− sin u, cos u, 0),
(2.9)
e2 = (r (v) cos u, r (v) sin u, t (v)).
Note that ej ∈ Γ(U, f ∗ (T E3 )).
Exercises 2.2.2:
1. (a) Show that e1 , e2 in (2.9) is an oriented orthonormal basis of Tf (u,v) M .
(b) Calculate e3 such that e1 , e2 , e3 is an orthonormal basis of Tf (u,v) E3 .
2. Considering this frame as an adapted lift F : U → F 1 , calculate the
pullback of the Maurer-Cartan forms in terms of du, dv.
3. Calculate the Gauss and mean curvature functions of M .
4. Consider the surfaces of revolution generated by the following data. In
each case, describe the surface geometrically. (Take time out to draw some
pictures and have fun!) Calculate H, K and describe their asymptotic be-
havior.
(a) r(v) = constant, t(v) = v.
(b) r(v) = av, t(v) = bv, where a, b are constants such that a2 + b2 = 1.
(c) r(v) = cos v, t(v) = sin v.
(d) The generating curve in the xz-plane is a parabola, e.g., x − bz 2 = c.
(e) The generating curve is a hyperbola, e.g., x2 − bz 2 = c.
2 2
(f) The generating curve is an ellipse, e.g., xa2 + zb2 = 1.
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
46 2. Euclidean Geometry
Although we have the same Gauss and mean curvature functions, the
helicoid is ruled, and it is not hard to check that the catenoid contains no
line segments. Since a Euclidean transformation takes lines to lines, we see it
is impossible for the helicoid to be equivalent to the catenoid via a Euclidean
motion.
We will see in Example 7.4.24 that, given a nonumbilic surface with
constant mean curvature, there are a circle’s worth of noncongruent surfaces
with the same Gauss curvature function and mean curvature as the given
surface. On the other hand, the functions H and K are usually sufficient
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
48 2. Euclidean Geometry
Remark 2.4.6. It turns out the property of being flat is invariant under a
larger group than ASO(3), and flat surfaces are best studied by exploiting
this larger group. This topic will be taken up in Chapter 4, where we classify
all flat surfaces: in the analytic category they are either open subsets of
cones, cylinders, or tangential surfaces to a curve. Even in the C ∞ category,
the only complete flat surfaces are cylinders; see [174], where there are also
extensive comments about the local characterization of flat C ∞ -surfaces.
Let (x, e1 , . . . , en+s ) denote an element of ASO(n + s). Define the pro-
jection
π : ASO(n + s) → En+s ,
(x, e1 , . . . , en+s ) → x.
ωia = haij ω j ,
In this situation, if we were to look for scalar functions that are constant
on the fibers, we would get a mess, but there are simple vector bundle valued
functions that are constant on the fibers. Recall our general formula (1.21)
for how the Maurer-Cartan form changes under a change of lift. Under a
motion (2.11) we have
To view our calculations more invariantly, let N M denote the normal bundle
of M , the bundle with fiber Nx M = (Tx M )⊥ ⊂ Tx En+s . Define
: = ω a ω j ⊗ ea
II j
The Gauss equation. There is another way to calculate the Gauss curva-
ture of a surface, namely by differentiating ω12 . Using the Maurer-Cartan
equation, we obtain
(2.14) dω12 = −Kω 1 ∧ ω 2 ,
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
52 2. Euclidean Geometry
Figure 1. The shaded area of the surface maps to the shaded area
of the sphere
Exercise 2.5.9: Let M n ⊂ En+s . Show that dγ = II, the Euclidean second
fundamental form.
dη 1 = aη 1 ∧ η 2 ,
dη 2 = bη 1 ∧ η 2 .
dη 1 = −α ∧ η 2 ,
dη 2 = α ∧ η 1 .
Curves such that ∇c c = 0 are called geodesics. Define the geodesic
curvature κg (c(t)) := |∇c (t) c (t)|, a measure of the failure of c(t) to be a
geodesic.
The Codazzi equation. Given two functions H and K on an open subset U ⊂
R2 , does there exist (locally) a map f : U → E3 such that H and K are the
mean and Gauss curvature functions of M = f (U )? There are inequalities
on admissible pairs of functions because H, K are symmetric functions of
the principal curvatures (so, e.g., H = 0 implies K ≤ 0). However, as we
have already seen in (2.10), stronger restrictions are uncovered when one
differentiates and checks that mixed partials commute.
To derive these restrictions without assuming a Darboux framing, set
up an EDS on ASO(3) × R3 for lifts of surfaces equipped with second fun-
damental forms. Give R3 coordinates hij = hji and define
0 0 τ 0
In Exercise 2.8.11, you will show that geodesics are locally the curves
that are the shortest distance between two points on a Riemannian manifold.
Thus the notion of a small geodesic disk about a point, which we will use in
the proof of the Gauss-Bonnet Theorem, makes sense. Notice that if κg ≡ 0,
then all of the curvature of the curve lies in the normal direction, and is
parallel to the binormal B of the curve.
Next, (2.16) shows that κn measures the curving of the surface in the
direction of T (by means of measuring how the surface normal e3 is bending);
it is called the normal curvature of the surface along c. Since it only depends
on the pointwise value of T , it is really an invariant of the surface. We say
c ⊂ M is an asymptotic line on M if κn ≡ 0.
Finally, Equation (2.17) shows τg measures what the torsion (as a curve
in R3 ) of a geodesic having tangent vector T would be; it is called the
geodesic torsion of c.
Exercises 2.7.2:
1. Show that
κn = −II(T, T ), e3 ,
(2.17) τg = II(, T ), e3 ,
(2.18) κg = −(dθ + ω12 )(T ).
2. Show that κg ≡ 0 if and only if the osculating plane to c is perpendicular
to the surface at each point.
3. Show that ∇c c = κg (c).
4. Find formulas for κn , τg in terms of the principal curvatures k1 , k2 when
our surface is given a principal (Darboux) framing.
5. Find formulas for κn , τg when our surface is given a framing such that
e1 = T .
6. Calculate τg of a curve c such that c is a principal direction (i.e., a
direction where κn is a principal curvature) at each point along c. Such
curves are called lines of curvature.
Exercise 2.7.3: Prove the local Gauss-Bonnet Theorem: Let (M 2 , g) be
a Riemannian manifold and let U ⊂ M be a connected, simply connected
oriented open subset with smooth boundary ∂U . Show that
KdA = 2π − κg ds,
U ∂U
Notice that each vertex of the triangulation becomes a zero of index +1,
and each edge and face contains a zero of index −1 or +1, respectively.
To prove that χmetric (M, g) = χ(M ), we follow ([174], vol. III): Divide
M into two pieces, a subset U ⊂ M where X is complicated but the topology
of U is trivial, and M \U where X is simple but we know nothing about the
topology.
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
2.8. The Gauss-Bonnet and Poincaré-Hopf theorems 59
KdA = − dη12 = − η12 = η12 .
N( ) N( ) ∂N ( ) i ∂Di ( )
Let ẽj1 , ẽj2 be an oriented orthonormal framing in Dj (), and from now
on we suppress the j index. Let θ denote the angle between e1 and ẽ1 . Then
η̃12 = η12 − dθ wherever both framings are defined. Thus
C2 δ1
δ2
C1
R
C3
δ4
δ3 C4
Then, use this formula to obtain a second proof of the Gauss-Bonnet The-
orem using the triangulation definition of the Euler characteristic.
you’ve made inside the circle? Zero! Next, take a round sphere, sit on it,
twist it, and fold it so it acquires regions of negative curvature. What’s the
average curvature of your distorted sphere? 4π, no matter how strong you
are!
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
2.8. The Gauss-Bonnet and Poincaré-Hopf theorems 61
Exercises 2.8.4:
1. Let M 2 ⊂ E3 . Show that the degree of the Gauss map is χ(M )/2.
2. What is ∂R κg ds, where R is the region enclosed by the lower dashed
curve (only half of which is pictured) in Figure 3?
Computing de1 and de2 gives ω13 = d(fx ) and ω23 = d(fy ), so that h = (hij )
is just the Hessian of f .
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
2.8. The Gauss-Bonnet and Poincaré-Hopf theorems 63
Exercise 2.8.6: Show that, relative to our framing for the graph z =
f (x, y),
1
2 2 − 21 1 2 fxx fxy ω
Q = (1 + fx + fy ) (ω , ω ) ,
fyx fyy ω2
2 −1 1 + fx −fx fy
2
−1 2
(gij ) = (1 + fx + fy ) .
−fx fy 1 + fy2
Then confirm that H, K are given by (1.3).
Now we wave our hands a little and ask you to trust that calculus in
infinite dimensions behaves the same way as in finite dimensions. That is,
a function has a critical point at a point where the derivative vanishes, and
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
64 2. Euclidean Geometry
in our case its easy to see we are at a minimum. (If you don’t believe us,
consult a rigorous book on the calculus of variations, such as [72].)
Exercise 2.8.10: More generally, for M n ⊂ En+s , define H ∈ Γ(N M ), the
mean curvature vector, to be traceg (II). Show that M is minimal if and
≡ 0. In particular, show that straight lines are locally the shortest
only if H
curves between two points in the plane.
Exercise 2.8.11: More generally, let X n+s be a Riemannian manifold and
M n ⊂ X a submanifold, and the mean curvature vector define H ∈ Γ(N M ),
to be traceg (II). Show that M is minimal if and only if H ≡ 0. In particular,
show that geodesics are locally the shortest curves between two points of X.
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
https://doi.org/10.1090//gsm/175/03
Chapter 3
Riemannian Geometry
65
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
66 3. Riemannian Geometry
(3.1) dη i = −αji ∧ η j .
Forms αji satisfying (3.1) are called connection forms and we also write
α = (ei ⊗ η j ) ⊗ αji ∈ Ω1 (F , End(T M )). Such forms always exist; however,
unlike in the case of the Maurer-Cartan form, the αji will not be uniquely
defined unless we specify additional conditions beyond (3.1).
Exercise 3.1.1: Suppose η i is a local coframing defined on U ⊂ M . Define
a local trivialization
t : GL(V ) × U F (M )|U
by (g, x) → g −1 η x . Show the existence of a desired coframing (η, α) on
F (M )|U as follows:
(a) Show that t∗ η = g −1 η.
(b) Show that there exist αji ∈ Ω1 (F (M )|U ) satisfying (3.1) such that
t∗ αji ≡ (g −1 dg)ij modulo {η i }.
(c) Show that (η i , αji ) is a coframing of F (M )|U .
(d) Show that any other coframing satisfying (3.1) must be of the form
α̃ji = αji + Cjki η k for some functions C i = C i .
jk kj
FGL (M ) to distinguish it from Fon (M ).) The orthogonal group O(V ) acts
simply transitively on the fibers of Fon (M ).
Definition 3.1.2. For a bundle π : B → M , define the vertical tangent
space at u ∈ B to be ker π∗u ⊂ Tu B.
Exercise 3.1.3: Show that the vertical tangent space of Fon (M ) at each
point is isomorphic to so(V ).
To prove uniqueness, assume that ηji , η̃ji both satisfy (i) and (ii). The
Cartan Lemma applied to (η̃ji − ηji ) ∧ η j = 0 shows η̃ji − ηji = Cjk
i η k for some
i = C i . Moreover C i = −C j .
functions Cjk kj jk ik
i = 0.
Exercise 3.1.6: Show that Cjk
Proof. First assume that dηji + ηki ∧ ηjk = 0. By Cartan’s Theorem 1.6.13,
around any point x ∈ M there exist an open set U and a map a : U → SO(n)
such that ηji = (a−1 da)ij on U . We have
dη j = −(a−1 da)jk ∧ η k .
Taking a new frame ηi = aij η j , where aij are the entries of a, we obtain
η i = daij ∧ η j − aij dη j
d
= daij ∧ η j − aij (a−1 da)jk ∧ η k = 0.
Thus ηi = dxi for some functions xi defined on a possibly smaller open set
U . The other direction is straightforward
A wish list for our differential operator ∇. Here are properties we might like
∇ to satisfy for all X, Y ∈ Γ(T M ) and f ∈ C ∞ (M ):
(1) ∇(X + Y ) = ∇X + ∇Y (additivity).
(2) ∇X+Y = ∇X + ∇Y (linearity).
(3) ∇(f X) = X ⊗df + f ∇X (Leibniz rule).
(4) ∇f X Y = f ∇X Y , i.e., (∇X Y )p only depends on Xp and not on its
extension to a vector field.
(5) Mixed partials commute, in the sense that in coordinates ∇ ∂ ∇ ∂
∂xi ∂xj
−∇ ∂ ∇ ∂ .
∂xj ∂xi
Exercise 3.1.20: Show that the last property is equivalent to having ∇X ∇Y
−∇Y ∇X − ∇[X,Y ] = 0.
The first two properties are immediate from the definition. The next
two follow from the first two exercises below. The last property fails by
Exercise 3.1.21.4 below; however the weaker commutation of mixed partials
∂ ∂
∇ ∂ −∇ ∂ =0
∂xi ∂xj ∂xj ∂xi
holds. More generally,
(3.5) ∇X Y − ∇Y X = [X, Y ]
holds by Exercise 3.1.21.3 below. Note that the right-hand side of (3.5),
which is the Lie bracket of X and Y , is independent of the Riemannian
metric. A connection α whose associated differential operator satisfies (3.5)
is said to be torsion free.
Exercises 3.1.21:
1. Show that ∇Y (f X) = f ∇Y X + Y (f )X for f ∈ C ∞ (M ).
2. Show that ∇f Y (X) = f ∇Y X for f ∈ C ∞ (M ).
3. Show that (3.5) indeed holds.
4. For X, Y ∈ Γ(T M ), define
Θ̃(X, Y ) = ∇X ∇Y − ∇Y ∇X − ∇[X,Y ] .
least n2 . In this section we will deal with the especially restrictive case of
realizing M as a hypersurface.
We are now finally in a position to answer our original question regarding
the equivalence of surfaces in E3 (the case n = 2), and at the same time see
the generalization to hypersurfaces. On an oriented hypersurface in En+1 ,
the second fundamental form can be made scalar valued. Denote this by
h ∈ Γ(S 2 T ∗ M ).
Exercise 3.2.6: Show that when n = 2 the Codazzi equation (2.15) is
equivalent to ∇h ∈ Γ(S 3 T ∗ M ).
Exercise 3.2.7: More generally, for a hypersurface M n ⊂ En+1 , show that
h ∈ Γ(S 2 T ∗ M ) satisfies ∇h ∈ Γ(S 3 T ∗ M ).
Theorem 3.2.8 (The fundamental theorem for hypersurfaces in En+1 ). Let
(M n , g) be a Riemannian manifold with curvature tensor R ∈
Γ(Λ2 T ∗ M ⊗Λ2 T ∗ M ), and let h ∈ Γ(S 2 T ∗ M ). Assume that
i. R = G(h) (Gauss)
and
ii. ∇h ∈ Γ(S 3 T ∗ M ) (Codazzi).
Then for every x ∈ M there exist an open neighborhood U containing x,
and an embedding f : U → En+1 as oriented hypersurface, such that f ∗ (I) =
g and f ∗ (II, en+1 ) = h, where en+1 is the outward unit normal vector
determined by the orientation. Moreover, f is unique up to a Euclidean
motion.
n
Corollary 3.2.9. Let M n , M ⊂ En+1 be two orientable hypersurfaces with
fundamental forms I, II and I, II. Suppose there exist a diffeomorphism
φ : M → M and unit vector fields en+1 , en+1 such that
i. φ∗ (I) = I,
and
ii. φ∗ II, ēn+1 = II, en+1 .
Then there exists a ∈ AO(n + 1) (the group ASO(n + 1) plus reflections)
such that φ = a |M .
Integral manifolds of the EDS (I, Ω) are graphs of immersions i : Fon (M )|U →
AO(n + 1) that are adapted frame bundles of immersions j : U → En+1 , for
U ⊂ M , satisfying j ∗ (I) = g and j ∗ (II, en+1 ) = h. We obtain existence
by:
Exercise 3.2.10: Show that (I, Ω) is Frobenius.
S 1 V = S 2 V = Λ2 V .
2 1
Let S 2 1 V denote the kernel of the skew-symmetrization map S 1 V ⊗
3
S 2 V → S 2 V . Elements of S 2 1 V are tensors of the form u⊗v⊗w − u⊗w⊗
3 3 3
1
v + v⊗u⊗w − v⊗w⊗u. The first Bianchi identity implies Θx ∈ T ⊗S 2 1 T ∗ .
3
Exercise 3.3.2: Show that R ∈ Γ(S 2 (Λ2 T ∗ M )), i.e., Rijkl = Rklij .
Just as we can “subtract” the scalar curvature from Ricci to get a tensor
in a smaller space, we can subtract the Ricci from the Riemann curvature
tensor to get a tensor, called the Weyl curvature tensor and denoted W ,
which is the “totally traceless” part of the Riemann curvature tensor. The
formula for W is given in Equation (3.11) below. We will return to these
decompositions in §3.5.
Remark 3.3.5. Myers’ Theorem states that if for all x ∈ M and for all
v ∈ Tx M with gx (v, v) = 1, Ric(v, v) > C > 0, then M is compact; see, e.g.,
[149, Thm. 19.5]. The geometry and topology of manifolds with positive
scalar curvature is an active area of research; see e.g. [138].
Thus, H n may be considered as (one half of) the “sphere of radius −1” in
Ln+1 .
Exercise 3.4.1: Show that Q restricts to be positive definite on vectors
tangent to H n .
Exercises 3.4.2:
1. Show that the sectional curvature of X() is constant for all 2-planes.
(In particular, it is zero for En , positive for S n , and negative for H n .)
2. Compute Ric, s and Ric0 for the X()’s.
3. Compute ∇R for S n and H n .
4. Show that the above construction generalizes to space forms X(r) with
r ∈ R, i.e., replace by an arbitrary real constant, and that the sectional
curvature function K : G(2, T M ) → R of a Riemannian manifold is constant
if and only if M = X(r) for some r, or an open subset thereof.
5. What are the geodesics in each of the X()’s?
6. Let M n−1 ⊂ X n () be a hypersurface. Define its second fundamental
form and describe the hypersurfaces with II ≡ 0.
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
82 3. Riemannian Geometry
Questions. In Exercise 3.4.2.2 we saw that space forms had zero traceless
Ricci curvature and constant scalar curvature. We ask more generally:
What are the Riemannian manifolds with s = 0?
What are the Riemannian manifolds with Ric0 = 0?
What are the Riemannian manifolds with W = 0?
What are the Riemannian manifolds with ∇R = 0?
These questions will lead us to study other geometries. For example, W is
related to conformal geometry, in fact metrics with W = 0 are conformally
equivalent to a flat metric; see §11.2. Metrics with Ric0 = 0 are solutions
of the following variational problem, which in the (1, n − 1) signature case
correspond to Einstein’s field equations for gravity, and so are called Einstein
(in any signature). For simplicity, only
consider metrics on M with volume
equal to one, i.e., metrics g such that M dvolg = 1. Let S(g) := M sg dvolg ,
the integral of the scalar curvature. Then Einstein metrics are the critical
points for this functional.
Intuitively, metrics with ∇R = 0 are metrics where the curvature is in
some sense “constant”, and one would expect these to be homogeneous. In
Theorem 9.7.19 we will see something even stronger is true. We could also
ask for part of the curvature to be “constant”. For example, one can study
metrics with constant scalar curvature: this is related to the famous Yamabe
problem, to find a metric with constant scalar curvature that is conformally
equivalent to a given one (see [21] for a discussion). It turns out that there
is always such a metric, and moreover the metric is unique if the scalar
curvature is negative.
Exercises 3.4.4:
1. Assume n > 2. Show that if (M, g) is Einstein, then the scalar curvature
is constant. Conclude that the only Riemannian manifolds with W = 0
and Ric0 = 0 are (open subsets of) spheres, planes, and hyperbolic space.
2. Calculate the Riemann curvature tensor, how it decomposes, and com-
pute ∇R for the product of two spheres S n (r1 ) × S N (r2 ) endowed with
the product metric. What are the geodesics? For which n, r1 , N, r2 is the
product metric Einstein?
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
3.5. Representation theory for Riemannian geometry 83
(3.10) (∇R)x ∈ S 1 3 5 T ∗.
2 4
The Pieri rule (see, e.g., [65, §6]) implies that STπ V ⊗V ⊂ V ⊗d+1 decom-
poses into a direct sum all the ways of adding a box filled with the integer
d + 1 to Tπ and to still have a Young tableau.
Example 3.5.4.
S 1 3 V ⊗V = S 1 3 4 V ⊕ S 1 3 V ⊕ S 1 3 V.
2 2 2 4 2
4
we could view the contraction with the quadratic form as first identifying
one copy of V with V ∗ via the quadratic form and then taking the trace.
The contraction on S 2 V is just R.
As an O(T )-module T is isomorphic to T ∗ and we have the decomposition
S 1 3 T = (S 1 3 T )0 ⊕ (S 1 2 T )0 ⊕ R.
2 4 2 4
These are respectively the homes of Weyl, traceless Ricci and scalar curva-
ture.
To compute the Weyl tensor and determine the contribution of the Ricci
to the Riemann curvature tensor, we need to explicitly compute the inclusion
S 2 V ∗ ⊗S 2 V ∗ → S 2 (Λ2 V ∗ ). This is given by the Kulkarni-Nomizu product:
for h, k ∈ S 2 V ∗ and a basis v 1 , . . . , v n of V ∗ , define h k ∈ S 2 (Λ2 V ∗ ) by
(h k)ijkl = hil kjk − hik kjl + hjk kil − hjl kik. If h = k this is the map from
symmetric n × n matrices to symmetric n2 × n2 matrices given in bases
by taking the entries of the new matrix to be twice the 2 × 2 minors of the
original matrix. Then (when n > 2)
s 1
(3.11) R= gg+ Ric0 g + W.
2n(n − 1) n−2
When n = 2 the formula is R = 4s g g.
Exercise 3.5.5: Verify that R − s
2n(n−1) g g− 1
n−2 Ric0 g ∈ (S 1 3 T ∗ )0 ,
2 4
which implies (3.11).
Exercise 3.5.6: What are the Weyl curvature tensors of the space forms
X()?
where LX is the Lie derivative along X. The flow of X will provide the de-
sired one-parameter group and X = dφ dt |t=0 . We first derive the differential
t
This looks a lot like ∇X = ei ⊗(xj η ij + dxi ) from (3.3). In fact, recalling the
isomorphism : Tx M → Tx∗ M induced by gx and Exercise 3.1.10, we have
(3.15) LX g = c 1 2 ◦ (⊗IdT ∗ M )(∇X).
Thus LX g = 0 if and only if (⊗IdT ∗ M )∇X is skew-symmetric. In summary:
Proposition 3.6.3. LX g = 0 if and only if ∇X ∈ Γ(so(T M, g)).
Exercises 3.6.4:
1. Show that the vector space of Killing vector fields is a Lie subalgebra of
Γ(T M ).
2. Show that LX = ∇X − ∇X, e.g., LX Y = ∇X Y − ∇Y X for Y ∈ Γ(T M )
and LX α = ∇X α − α ◦ ∇X for α ∈ Ω1 (M ). In this formula ∇X is a
differential operator on sections of a tensor bundle and ∇X is the endomor-
phism of the tensor bundle induced from an element of gl(Tx M ). (Note:
this explains why, although in general to take a Lie derivative one must look
at a tensor along a curve, for LX g we do not, as ∇X g = 0 for all X, i.e.,
LX g = (∇X).g.) This observation is due to Kostant [116].
3. Use the previous exercise to obtain a second proof that LX g = 0 if and
only if ∇X ∈ Γ(so(T M )).
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
3.6. Infinitesimal symmetries: Killing vector fields 87
3.6.2. An EDS for Killing vector fields. Given (M, g), we now set up an
EDS for Killing vector fields X ∈ Γ(T M ). We will solve for the horizontal lift
X̃ of a Killing vector field X ∈ Γ(T M ) to the orthonormal coframe bundle
F . Since the condition involves X̃ and its derivatives, we introduce new
2
variables to account for them: we work on a submanifold of F × Rn × Rn ×
n
Rn( 2 ) where the last three factors have coordinates (xi , xij , xi j ) respectively,
(k)
with all indices running from 1 to n and xi j = −xi k . On any integral
(k ) (j )
manifold, the xi will be functions such that X̃ = xi ei , and the remaining
variables will be such that dxi = xij η j + 12 xijk ηkj . Thus, our EDS will include
differential forms dxi − xij η j − 12 xijk ηkj whose vanishing ensures this.
Note that our initial value problem will be specifying the value of X at
a point and the value of its first derivatives at that point. The questions
will then be: is there an extension of X to a Killing vector field, and if there
is, is the extension unique?
Now we introduce the Killing condition, i.e., we restrict to Σ ⊂ F ×Rn ×
n
R × Rn( 2 ) defined by (3.14) assuming dxi = xij η j + 12 xijk ηkj , i.e.,
n2
1
(3.16) 0= ((δki xj + xi k )ηjk + xik η k ) η i .
2 (j )
i
More precisely, since the forms ηjk , η i are all linearly independent, for (3.16)
to hold we must have
(3.17) xi k = −δki xj + δji xk ,
(j )
(3.18) xik + xki = 0,
n
and these are the defining equations for Σ. Thus Σ = F × Rn × R( 2 ) , where
n
R( 2 ) has coordinates xij = −xji . Set θi := dxi − xij η j + xj ηji . Our Pfaffian
EDS on Σ is generated by
(3.19) I = {θi }.
Since π ∗ (∇X) = ei ⊗(dxi + xj ηji ) ≡ ei ⊗xij η j mod I, where π : F → M is the
projection, we see that we are indeed requiring that ∇X be so(V )-valued.
Recalling dη j = −ηkj ∧ η k and dηji = −ηki ∧ ηjk + 12 Rjkl
i η k ∧ η l , we compute
1 i k
dθi = −dxij ∧ η j + xij ηlj ∧ η l + dxj ∧ ηji + xj (−ηki ∧ ηjk + Rjkl η ∧ ηl )
2
1
(3.20) ≡ [−dxil + xij ηlj − xjl ηji + xj Rjkli
η k ] ∧ η l mod I.
2
Now if we add
xj (− 12 Rjki − 12 Rjk i )η k
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
88 3. Riemannian Geometry
to the term in the brackets, it does not change the expression because
[xj (− 12 Rjki − 12 Rjk i )η k ] ∧ η l = 0. But now
1
−dxil + xij ηlj − xjl ηji + xj Rjkl
i
η k + xj (− 12 Rjki − 12 Rjk i )η k
2
is skew in i and l. Applying the Bianchi identity, this becomes
1
−dxil + xij ηlj − xjl ηji − xj Rljk
i
ηk ,
2
and these forms must vanish on any integral manifold, so we add the forms
1
θli := −dxil + xij ηlj − xjl ηji − xj Rljk i
ηk
2
to our ideal and start over. Let I˜ = {θi , θli }. We have arranged such that
dθi ≡ 0 mod I,˜ and it remains to compute the derivative of θi :
l
1 i
dθli = 0 + dxij ∧ ηlj + xij dηlj − dxjl ∧ ηji − xjl dηji + Rljk dxj ∧ η k
2
1 1
+ xj dRljk i
∧ η k + xj Rljk i
dη k
2 2
1
(3.21) ≡ − xj [(∇R)iljku η u ] ∧ η k
2
+ [xit Rlsj
t
− xtl Rtsj
i
− xts Rltj
i
− xtj Rlst
i
]η s ∧ η j mod I.
3.6.3. Exercises.
(1) Give geometric interpretations in S n and H n of the Killing vector
fields at a point x that fix x.
(2) Let X ∈ K and show LX ◦ ∇ = 0.
we may do the same, however we further need the inner product on g/k to
be K-invariant in order to left translate unambiguously, as in Proposition
3.7.5.
Notation. For a K-module W , W K denotes the K-invariant vectors in
W , the isotypic component of the trivial representation in W .
Exercise 3.7.10: Compute the Riemann curvature tensor for each metric
in this one-parameter family as well as its decomposition, and ∇R. Which
metrics are Einstein?
1
Example 3.7.11. Let G = SU (n + 1) and K = | B ∈ SU (n)}
B
SU (n). Take the map G → G/K, (e0 , . . . , en ) → e0 so G/K = S 2n+1 . Write
ix −z T
g= | x ∈ R, z ∈ Cn , X ∈ su(n) .
z − ix
n Id n +X
Then (g/k)∗ R ⊕ Cn , so there is a two-parameter family of SU (n + 1)-
invariant metrics on S 2n+1 .
Remark 3.7.12. The maps of Examples 3.7.9 and 3.7.11 are compatible.
Factoring we obtain the Hopf fibration S 1 → S 2n+1 → CPn .
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
92 3. Riemannian Geometry
Exercise 3.7.13: Compute the Riemann curvature tensor for each metric
in this two-parameter family as well as its decomposition and ∇R. Which
metrics are Einstein?
and the operator agrees with the classical Laplacian (up to a sign which is
inserted to make the operator have nonnegative eigenvalues).
Since Δg : Ωk (M ) → Ωk (M ) is a map from a vector space to itself, we
can talk about eigenvalues and eigenvectors (eigenfunctions).
What is the geometric meaning of these eigenfunctions?
In general, they
are critical for the “energy” functional E(α) = M (α, α)g dvolM .
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
3.8. The Laplacian 93
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
https://doi.org/10.1090//gsm/175/04
Chapter 4
Projective Geometry I:
Basic Definitions and
Examples
Until the middle of the twentieth century, there was not such a distinc-
tion between algebraic and differential geometry. In this chapter we discuss
work of Darboux, Cartan, Segre, Terracini and others who approached the
geometry of subvarieties (and submanifolds) of projective space via local
differential geometry. This approach was revisited in the paper of Griffiths
and Harris [85], which began a synthesis of modern algebraic geometry and
moving frames techniques. An earlier version of this chapter, together with
Chapter 12, containing more algebraic results than presented here, consti-
tuted the monograph [129].
We study the local geometry of submanifolds of projective space and
applications to algebraic geometry. We describe moving frames for subman-
ifolds of projective space and define the projective second fundamental form
in §4.1. In §4.2 we give basic definitions from algebraic geometry. We also
give examples of homogeneous algebraic varieties and explain several con-
structions of auxiliary varieties from a given variety X ⊂ PV : the secant
variety σ(X), the tangential variety τ (X) and the dual variety X ∨ . In §4.3
we describe the basic properties of varieties with degenerate Gauss maps
and classify surfaces with this property.
In this chapter, unless we say otherwise, when we work over the complex
numbers, all tangent, cotangent, etc., spaces are the holomorphic tangent,
cotangent, etc., spaces (see Appendix C). We will generally use X to denote
95
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
96 4. Projective Geometry I: Basic Definitions and Examples
Note that Tv M̂ = Tw M̂ for all nonzero v, w ∈ x̂, so T̂x M, T̃x M are well-
defined. For any submanifold of V , the affine tangent space is the naı̈ve
tangent space obtained by translating tangent vectors to the origin using the
identification Tv V V . The affine and embedded tangent spaces should not
be confused with the (intrinsic) tangent space Tx M defined in Appendix B.
Recall from §1.9 that the intrinsic tangent space to PV has the description
Tx PV x̂∗ ⊗(V /x̂).
Exercise 4.1.1: Let M ⊂ PV be a submanifold. Show that Tx M x̂∗ ⊗
(T̂x M/x̂). Find the corresponding description of Tx∗ M .
Since de0 = e0 ω00 +e1 ω01 +. . .+en+a ω0n+a , and de0 ≡ 0 mod{e0 , . . . , en } when
pulled back to F 1 , we must have ω0μ = 0 ∀μ.
Notation. Given f = (e0 , . . . , en+a ) ∈ F 1 , we let (e0 , . . . , en+a ) denote
the dual basis of V ∗ and eα = e0 ⊗(eα mod e0 ) ∈ T[e0 ] M denote the tangent
vector corresponding to eα (see Exercise 4.1.1). Note that for any section
s : M → F 1 , we have s∗ (ω0α )(eβ ) = δβα . Also note that eα depends on f and
not just eα . Similarly, for normal vectors, we let eμ = e0⊗(eμ mod T̂[e0 ] M ) ∈
N[e0 ] M .
F 1 → G(n + 1, V ),
f → [e0 ∧ · · · ∧ en ].
ωβμ = qαβ
μ
ω0α
μ μ
for some functions qαβ = qβα defined on F 1 .
Exercise 4.1.8: Prove the “moreover” assertion in Proposition 4.1.4.
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
4.1. Frames and the projective second fundamental form 99
ẽ0 = g00 e0 ,
ẽα = gα0 e0 + gαβ eβ ,
ẽμ = gμ0 e0 + gμβ eβ + gμν eν .
We first determine how ω0α , ωαμ vary. Let “g·” denote the action of G both
on V and on the Maurer-Cartan form, and observe that
g · de0 = g · (e0 ω00 + eα ω0α )
= (g00 e0 )(g · ω00 ) + (gα0 e0 + gαβ eβ )(g · ω0α ).
On the other hand, since d commutes with the g-action,
g · de0 = d(e0 g00 )
= g00 (e0 ω00 + eα ω0α ).
Let h denote the inverse matrix to g. The analogous calculation for deα
combined with our calculation above yields
ω̃0α = hαβ g00 ω0β ,
ω̃αμ = gαβ hμν ωβν .
Substituting into the equation ω̃βμ = q̃βα
μ
ω̃0α , we obtain
μ
q̃αβ = h00 hγβ hδα gνμ qγδ
ν
.
Thus the quantity
= q μ ω α ω β ⊗eμ ∈ π ∗ (S 2 T ∗ M ⊗N[e ] M )
II αβ 0 0 [e0 ] 0
of a quadratic form over C is its rank. For real projective hypersurfaces the
only numerical information is rank and signature.
Remark 4.1.10. If we are in the holomorphic category, we may compare
the projective second fundamental form with its Hermitian counterpart. In
this case the projective second fundamental form is simply the holomorphic
component of the Hermitian second fundamental form.
Remark 4.1.11. The projective second fundamental form admits the fol-
lowing interpretation. As remarked in Exercise 4.1.3, the conormal space
Nx∗ M may be identified (after tensoring with a line bundle) with the space
of linear forms on V annihilating T̂x M , so we may think of H ∈ N ∗ M as
x
determining a hyperplane H containing the tangent projective space T̃x M .
As such, as long as M H, M ∩ H must be singular at x because it is
defined by the equations of M and the linear form defining H, but this
linear form has differential in the span of the differentials of the defining
equations of M . With this perspective II(H) ∈ S 2 T ∗ M may be identified
x
as the quadratic part of the singularity.
M
M
Tx M Tx M
IP T x M IP T x M
Baseloc|II| Baseloc|II|
Exercises 4.1.12:
1. Show that Singloc |IIM,x | = {v ∈ Tx M | (s∗ ωαμ )(v) = 0, ∀α, μ}, where
s : M → F 1 is any section.
2. Show that if IIM,x ≡ 0 for all x ∈ M , then M is (an open subset of) a
linear Pn ⊂ PV . In particular, if M is analytic, then M is an open subset of
a linear space if and only if IIM,x = 0 at a general point x ∈ M as defined
below.
Example 4.2.1. If one considers the plane curve y = x3 , then all points
on the curve are general except the origin which is a flex point (also called
a point of inflection).
Proof. By Exercise 4.2.2 any line having contact to order two with X at x
is contained in X.
Remark 4.2.6. Borel and Remmert proved that any compact homogeneous
Kähler manifold is of the form G/P × T , where T is the quotient of Cg by
a lattice; see [2].
v2 : PV → PS 2 V,
[v] → [v 2 ].
F = {(E1 , . . . , Ep ) ∈ S | E1 ⊂ E2 , E2 ⊂ E3 , . . . , Ep−1 ⊂ Ep },
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
106 4. Projective Geometry I: Basic Definitions and Examples
and observe that F = Fa1 ,...,ap (V ). The incidence relations are described by
polynomials, namely various contractions being zero.
Joins, cones and secant varieties. Let Y, Z ⊂ PV be varieties (we allow the
possibility that Y = Z). For x, y ∈ PV , let P1xy denote the projective line
containing x and y. Define the join of Y and Z to be
J(Y, Z) = x∈Y,y∈Z,x=y P1xy .
If Z = Pk is a k-plane, J(Y, Pk ) is called the cone over Y with vertex Pk . If
Y = Z, σ(Y ) = J(Y, Y ) is called the secant variety of Y . We may similarly
define the join of k varieties to be the union of the corresponding Pk−1 ’s,
or by induction as J(Y1 , . . . , Yk ) = J(Y1 , J(Y2 , . . . , Yk )). In particular, write
σk (Y ) = J(Y, . . . , Y ) for the join of k copies of Y , called the k-th secant
variety of Y .
Aside 4.2.11. Given Y ⊂ PV , if Y is not contained in a hyperplane we
obtain a stratification of PV by the secant varieties of Y . Given x ∈ PV ,
define the Y -border rank of x to be the minimum k such that x ∈ σk (Y )
and the Y -rank of x to be the minimum k such that x is in the span of k
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
4.2. Algebraic varieties 107
Dual varieties.
Here, and it what follows, we identify Fs+1,t+1 with the space of Ps ’s con-
tained in Pt ’s contained in PV . Similarly consider
with picture
I
S
/
w
S
X X ∨ ⊂ PV ∗ .
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
4.3. Varieties with degenerate Gauss mappings 109
FX
1
Sρ
π
γ S
/ w
X - G(n + 1, V )
where π(e0 , . . . , en+a ) = [e0 ] and ρ(e0 , . . . , en+a ) = [e0 ∧ . . . ∧ en ]. Here the
geometric object of interest is the incidence correspondence Γ(X) defined in
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
110 4. Projective Geometry I: Basic Definitions and Examples
Recall that the ωαμ are semi-basic for π; thus assertion 1 and the “conversely”
part of Proposition 4.3.1 follow from Exercise 4.1.12.2.
We now implement a key idea in the method of the moving frame: now
that we have more geometric information, we adapt frames further to reflect
this information.
Let F γ ⊂ F 1 be the subbundle of the first-order adapted frame bundle
consisting of frames preserving the flag
Exercise 4.3.3: Show that our adaptations have the effect that the pull-
backs of the forms ωsμ to F γ are zero and that Singloc |II| = {ω0j }⊥ = {es }.
(Recall that S ∗ (ω0α )(eβ ) = δβα for any section S : X → F 1 .)
We now show that the fibers are indeed linear spaces. The Maurer-
Cartan equation implies
spaces, it will suffice to show that IIF,[e0 ] = 0. Because ωsμ = 0 when pulled
back to F γ and we may restrict F γ to be a frame bundle over F , we have
But ωsj ≡ 0 mod{ω0k } and the ω0k pulled back to F are zero; hence IIF = 0,
and the proposition is proved.
Thus, p is a smooth point of X if and only if the matrix (u0 δjk + us Csj
k ) is
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
https://doi.org/10.1090//gsm/175/05
Chapter 5
Cartan-Kähler I:
Linear Algebra and
Constant-Coefficient
Homogeneous Systems
We have seen that differentiating the forms that generate an exterior dif-
ferential system often reveals additional conditions that integral manifolds
must satisfy (e.g., the Gauss and Codazzi equations for a surface in Eucli-
dean space). The conditions are consequences of the fact that mixed partials
must commute. What we did not see was a way of telling when one has dif-
ferentiated enough to find all hidden conditions. We do know the answer
in two cases: If a system is in Cauchy-Kowalevski form there are no extra
conditions. In the case of the Frobenius Theorem, if the system passes a
first-order test, then there are no extra conditions.
What will emerge over the next few chapters is a test, called Cartan’s
Test, that will tell us when we have differentiated enough.
The general version of Cartan’s Test is described in Chapter 8. For a
given integral element E ∈ Vn (I)x of an exterior differential system I on
a manifold Σ, it guarantees existence of an integral manifold to the system
with tangent plane E if E passes the test.
In Chapter 6, we present a version of Cartan’s Test valid for a class
of exterior differential systems with independence condition called linear
Pfaffian systems. These are systems that are generated by 1-forms and have
the additional property that the variety of integral elements through a point
115
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
116 5. Constant-Coefficient Homogeneous Systems
5.1. Tableaux
Let x = xi vi , u = ua wa denote elements of V and W respectively. We will
consider (x1 , . . . , xn ), respectively (u1 , . . . , un ), as coordinate functions on
V and W respectively. Any first-order, constant-coefficient, homogeneous
system of PDE for maps f : V → W is given in coordinates by equations
∂ua
(5.1) Bari = 0, 1 ≤ r ≤ R,
∂xi
where the Bari are constants. For example, the Cauchy-Riemann system
u1x1 − u2x2 = 0, u1x2 + u2x1 = 0 has B111 = 1, B212 = −1, B112 = 0, B211 = 0,
B221 = 1, B122 = 1, B121 = 0 and B222 = 0.
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
5.1. Tableaux 117
Example 5.1.3. The Cauchy-Riemann system u1x1 = u2x2 , u1x2 = −u2x1 has
tableau
A ={a(w1 ⊗ v 1 + w2 ⊗ v 2 ) + b(−w2 ⊗ v 1 + w1 ⊗ v 2 )|a, b ∈ R}
a −b
|a, b ∈ R .
b a
Example 5.1.4. The tableau A = (0) corresponds to a Frobenius system.
The equations are uaxi = 0, ∀i, a. The only solutions to this system are
constant maps. We will say solutions depend on s constants.
Example 5.1.5. When A = W ⊗ V ∗ , there are no equations and any map
is a solution. Here, solutions depend on s functions of n variables.
Example 5.1.6. Let L∗ ⊂ V ∗ be a k-dimensional linear subspace and let
A = W ⊗ L∗ . If L∗ = {v 1 , . . . , v k }, the equations are uaxρ = 0, k + 1 ≤ ρ ≤ n,
and the solutions are ua (x1 , . . . , xn ) = f a (x1 , . . . , xk ), where the f a are
arbitrary functions. Here, solutions depend on s functions of k variables.
With respect to adapted bases we may write A in block form as (∗, 0),
where the left block is free and the right block is zero.
Example 5.1.7. Let A = Y ⊗ V ∗ , where Y is a p-dimensional linear sub-
space of W . If Y is spanned by uλ , 1 ≤ λ ≤ p, the equations are uξxi = 0,
where p + 1 ≤ ξ ≤ s. Solutions are
uλ (x1 , . . . , xn ) = f λ (x1 , . . . , xn ),
uξ (x1 , . . . , xn ) = f0ξ ,
where the f λ (x1 , . . . , xn ) are arbitrary functions and the f0ξ are constants;
so, solutions depend on p functions of n variables and s − p constants. One
may think of this tableau as a combination of two disjoint systems of types
5.1.5 and 5.1.6. With respect to adapted bases we may write A in block
form as ( ∗0 ), where the top block is free and the bottom block is zero.
In the rest of this chapter, we will develop a test for a tableau such that
if it passes, we will know exactly what initial data one should specify to
determine an analytic solution to (5.1). What to do if the tableau fails the
test is taken up in Chapter 6.
Given a tableau A, suppose we try to specify the solution in terms of a
power series
(5.3) ua (x) = pa + pai xi + paij xi xj + paijk xi xj xk + . . . .
Since the system is homogeneous with constant coefficients, u is a solution
if and only if each term satisfies the system and the series converges. As we
have seen, there is no restriction on the constant terms, and the coefficients
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
5.1. Tableaux 119
paij xi wa ⊗v j ∈ A ∀x ∈ V,
so that
(paij wa ⊗ v j ) ⊗ v i ∈ (A ⊗ V ∗ ).
Since mixed partials commute, (paij wa ⊗ v i ) ⊗ v j is also an element of W ⊗
S 2 V ∗ ; in indices this means paij = paji . Combining these two conditions gives
paij wa ⊗v i ⊗v j ∈ (A ⊗ V ∗ ) ∩ (W ⊗ S 2 V ∗ ) =: A(1) ,
where, at the risk of being redundant, the first factor is due to the equations
and the second to the commuting of derivatives. The space A(1) is called
the (first) prolongation of A.
Similarly, the condition on the next term of (5.3) is
(5.4) paijk wa ⊗v i ⊗v j ⊗v k ∈ (A ⊗ V ∗ ⊗ V ∗ ) ∩ (W ⊗ S 3 V ∗ ) =: A(2) ,
and so on for all l. In general, we define
A(l) := (A ⊗ V ∗⊗l ) ∩ (W ⊗ S (l+1) V ∗ )
to be the l-th prolongation of A. A map f : V → W , given in terms of a
convergent Taylor series, is a solution to the system defined by A if and only
if for all k, the tensor formed from the k-th term lies in A(k−1) .
Exercises 5.1.8:
1. Explicitly calculate A(1) when A is the tableau for the Cauchy-Riemann
equations.
2. Show that A(l) is naturally identified with the set of solutions to the
system defined by A that are homogeneous polynomials of degree l + 1.
3. Show that if A(l) = 0 for some l, then solutions depend on a finite number
of constants.
Definition 5.1.9. A tableau of order p is a linear subspace A ⊂ W ⊗S p V ∗ .
It determines a homogeneous constant-coefficient system of PDE of order p
for W -valued functions on V . In particular, the (p − 1)-st prolongation of a
tableau (of order one) is a tableau of order p. For a tableau A of order p,
we define its prolongations by A(l) = (A⊗V ∗⊗l ) ∩ (W ⊗S p+l V ∗ ).
Example 5.1.10. The equation uxy = 0 may be encoded as a tableau of
order two as
A = {p11 v 1 v 1 + p22 v 2 v 2 | p11 , p22 ∈ R}.
Here V = R2 , W = R.
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
120 5. Constant-Coefficient Homogeneous Systems
Exercises 5.1.11:
1. Express the system uxx + vyy = 0 of one equation for two functions of
two variables as a tableau of order two. What is its dimension?
2. Show that A(k+l) = (A(k) )(l) .
A = {(pa1 v 1 + C2b
a b 2
p1 v ) ⊗ wa |pa1 ∈ R, 1 ≤ a ≤ s}
a b n
p1 v + . . . + Cnb
(5.5)
= {(p1 , C2 p1 , . . . , Cn p1 )},
where the Cρ are fixed s × s matrices and p1 = (pa1 ) is any column vector.
The symbol relations corresponding to (5.5) are
∂ua a ∂u
b
(5.6) − C ρb = 0, ρ = 2, . . . , n.
∂xρ ∂x1
We saw that when Cρ ≡ 0 ∀ρ, any convergent series with coefficients
pa1 , pa11 , pa111 , . . . determines a solution u. The hope is that under some con-
ditions the solutions of the system (5.6) are describable in a similar way. In
fact, we will determine when solutions to (5.6) can be given in terms of s
arbitrary functions of one variable, and will see that this is the largest space
of solutions one could hope for.
Assume we are given constants pa1 , pa11 , pa111 , · · · . Then we may determine
the remaining paI for any multi-index I = (i1 , . . . , ik ) as follows: To have
(pai wa ⊗ v i ) ∈ A, the terms paρ must be given by paρ = Cρb a pb . To have
1
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
5.2. First example 121
paρ1 = Cρb
a b
p11 ,
(5.7)
paρσ = Cρb
a b a b c
pσ1 = Cρb Cσc p11 , 2 ≤ σ ≤ n.
These equations imply dim A(1) ≤ s. They may lead to conflicting equations
because it is also necessary that paρσ = paσρ , i.e., that Cρb
a C b pc = C a C b pc .
σc 11 σb ρc 11
In other words, to ensure that no choice of pa11 ’s leads to a conflict, it is
necessary that
(5.8) [Cρ , Cσ ]ac = Cρb
a b
Cσc − Cσb
a b
Cρc =0 ∀ a, c, ρ, σ.
If (5.8) holds, we are free to specify not only pa11 but in fact any convergent
power series for ua (x1 , 0, . . . , 0). More precisely,
Proposition 5.2.1. If the Cρ in tableau (5.5) are such that [Cρ , Cσ ] =
0 ∀ρ, σ, then there exists a unique solution to the initial value problem
γu (x) ∈ A = (p1 , C2 p1 , . . . , Cn p1 )
with initial condition
ua (x1 , 0, . . . , 0) = f a (x1 ),
where the f a are analytic.
Geometrically, the f a (x1 ) determine a curve in V × W , and we would
like to enlarge this curve to an n-dimensional integral manifold. By (5.7) we
see that there is at most one such, and the proposition states that if (5.8)
holds, this candidate is in fact an integral manifold. If (5.8) fails to hold,
there may still be solutions to (5.5), but there will be additional conditions
on the functions f a .
Proof of 5.2.1. We have seen that in this case the paρi are exactly deter-
mined by the pa11 , and the same computation shows the paρij are exactly
determined by the pa111 , etc. (The commutation relations take care of all
possible conflicts.) It remains to show that the formal power series for
u(x1 , . . . , xn ) determined by these coefficients converges.
Let c =max{1, maximum eigenvalue of any Cρ }. Consider the system
∂U a ∂U a
= c , ρ = 2, . . . , n.
∂xρ ∂x1
Its general solution is given by
U a (x1 , . . . , xn ) = F a (x1 + c(x2 + . . . + xn )),
where F a (t) are analytic functions of one variable. If
F a (t) = pa + pa1 t + pa11 t2 + pa111 t3 + . . . ,
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
122 5. Constant-Coefficient Homogeneous Systems
In summary, the largest space of solutions one could hope for in a system
of the form (5.5) is that solutions depend on s functions of 1 variable, and
whether or not this is the case can be determined just by computing A(1) .
One always has dim A(1) ≤ s, and solutions depend on s functions of one
variable if and only if dim A(1) = s.
Proof. The only thing to check is the claim about the determination of A(l)
and the convergence of the power series. These are left to the reader.
For any tableau of the form (5.11), we have dim A(1) ≤ s + 2k. Proposi-
tion 5.3.1 says that, when equality holds, solutions depend on k functions of
2 variables and (s − k) functions of 1 variable, which is the largest possible
space of solutions.
Remark 5.3.2 (another view of 5.3.1). One could attempt to solve the
system (5.11) by solving a sequence of Cauchy problems. For example,
suppose
uξ (x1 , 0, 0) = f ξ (x1 ),
uλ (x1 , x2 , 0) = f λ (x1 , x2 )
are the functions we specify to determine a putative solution to the system
(5.11). We could first solve the Cauchy problem
(5.13)
uξx2 (x1 , x2 , 0) = Cλξ uλx1 (x1 , x2 , 0) + Dηξ uηx1 (x1 , x2 , 0) + Eλξ uλx2 (x1 , x2 , 0),
uξ (x1 , 0, 0) = f ξ (x1 )
to determine uξ (x1 , x2 , 0), then solve the Cauchy problem
(5.14)
uλx3 (x1 , x2 , x3 ) = Fμλ uμx1 (x1 , x2 , x3 ) + Gλξ uξx1 (x1 , x2 , x3 ) + Eμλ uμx2 (x1 , x2 , x3 ),
uλ (x1 , x2 , 0) = f λ (x1 , x2 )
to determine uλ (x1 , x2 , x3 ), and finally obtain a map u : V → W by solving
(5.15)
uξx3 (x1 , x2 , x3 ) = Iλξ uλx1 (x1 , x2 , x3 ) + Jηξ uηx1 (x1 , x2 , x3 ) + Kλξ uλx2 (x1 , x2 , x3 ),
uξ (x1 , x2 , 0) = solution to (5.13).
Solving (5.13) determines the constants q2 , q12 , q22 , q112 , . . . for which all sub-
scripts are 1 or 2, with at least one being 2. Solving (5.14) determines
p3 , p13 , . . . and solving (5.15) determines p23 , p33 , . . . .
Just as with the similar procedure we tried in Example 1.2.3, this pro-
cedure yields a function which may not solve our system. As with Example
1.2.3, there are other sequences of Cauchy problems—e.g., we could have
solved first for uξ (x1 , 0, x3 )—to obtain a function. However, condition (5.12)
guarantees that the function we obtain is actually a solution to our system
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
5.4. Third example 125
Then dim A(1) = s1 + 2s2 + 3s3 + . . . + ksk , since an element of A(1) is deter-
mined by a choice of s1 constants p111 , . . . , ps111 , then 2s2 constants p112 , . . . , ps122 ,
p122 , . . . , ps222 , then 3s3 constants p113 , . . . , ps133 , p123 , . . . , ps323 , p133 , . . . , ps333 , and so
on, up to ksk constants p11k , . . . , pskkk .
The general solution of the corresponding system of differential equations
is
ua = f a (x1 , . . . , xk ), 1 ≤ a ≤ sk ,
ua = f a (x1 , . . . , xk−1 ), 1 + sk ≤ a ≤ sk−1 ,
..
.
ua = f a (x1 ), 1 + s2 ≤ a ≤ s1 ,
u =a
ua0 (constant), 1 + s1 ≤ a ≤ s.
with equality if and only if any choice of the s1 + 2s2 + . . . + ksk constants
paij for a point in A(1) determines the other paij ’s without conflict. Moreover,
in this case A(k) has its maximum possible dimension for all k as well,
i.e., commutation of all second-order derivatives implies commutation of all
higher-order derivatives.
Note that the sk (b) do not depend on the choice of basis for W , but only
on the flag F = (F0 , F1 , . . . , Fn ) of subspaces in V ∗ induced by b, which we
denote by
Fj = {v1 , . . . , vj }⊥ = {v j+1 , . . . , v n },
with F0 = V ∗ and Fn = (0). Hence, we write sk (F ) instead of sk (b).
We define
Ak (F ) = (W ⊗Fk ) ∩ A
and observe that
(5.17) dim Ak (F ) = sk+1 (F ) + . . . + sn (F ).
One can visualize Ak (F ) as the subspace of matrices in A for which the first
k columns are zero, when we use the basis b for V . In our first example,
A1 = (0); in our second example, A1 is the k-dimensional subspace of A
obtained by setting p1 , q1 = 0.
Definition 5.5.2. Let A ⊂ W ⊗V ∗ be a tableau. Define
s1 (A) = max{s1 (F )| all flags},
s2 (A) = max{s2 (F )| flags with s1 (F ) = s1 (A)},
..
.
sn (A) = max{sn (F )| flags with s1 (F ) = s1 (A), . . . , sn−1 (F ) = sn−1 (A)}.
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
5.5. The general case 127
For the “model” involutive tableau (5.16), solutions are uniquely deter-
mined by specifying
uσ (x1 , . . . , xk , 0, . . . , 0) for σ ≤ sk .
This motivates the following:
Definition 5.5.5. For an integer σ between 1 and s, define the level of σ to
be the largest k such that σ ≤ sk . (If σ > s1 , its level is defined to be zero.)
Example 5.5.6. Tableau (5.10) has characters s1 = s, s2 = s − k, s3 = 0,
so
level(1) = level(2) = . . . = level(s − k) = 2
while
level(s − k + 1) = . . . = level(s) = 1.
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
128 5. Constant-Coefficient Homogeneous Systems
For any tableau, the most general set of initial data one could hope to
specify is
(5.19) uσ (x1 , . . . , xk , 0, . . . , 0), where level(σ) = k,
with an integral manifold obtained by then solving a sequence of Cauchy
problems. In fact, an argument that is conceptually the same as in the
examples of the previous sections yields
Theorem 5.5.7 (Cartan-Kähler for tableaux). Let A ⊂ W⊗V ∗ be a tableau.
Choose A-generic bases of V, W which induce coordinates xi on V and ua on
W . If A is involutive, then any choice of analytic functions (5.19) uniquely
determines an integral manifold of the differential system represented by A
in some neighborhood of the origin.
Definition 5.5.8. If A is an involutive tableau such that sl = 0 but sl+1 = 0,
then sl is called the character of the system and the number l is called the
Cartan integer.
Seg(PV ∗ × PW )
pV ∗ S
/ SS
w
PV ∗ PW.
Proof. We have already seen the first inequality. Equality occurs if and
only if PFk−1 ∩ ΞA = ∅, i.e., iff PAk−1 ∩ Seg(PFk−1 × PW ) = ∅. But as
a tableau, Ak−1 has characters s1 (Ak−1 ) = sk (A), s2 (Ak−1 ) = 0, so we are
reduced to the special case. Since PFk−1 is a generic linear space, the same
calculation as above is also valid for the degree.
of the characteristic variety may be easier to
The dimension and deg
compute than carrying out Cartan’s Test. More importantly, this theorem
is an important step towards proving Theorem 5.6.12, stated above.
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
https://doi.org/10.1090//gsm/175/06
Chapter 6
We now generalize the test from Chapter 5 to a test valid for a large class of
exterior differential systems called linear Pfaffian systems, which are defined
in §6.1. In §§6.2–6.4 we present three examples of linear Pfaffian systems that
lead us to Cartan’s algorithm and the definitions of torsion and prolongation,
all of which are given in §6.5. For easy reference, we give a summary and
flowchart of the algorithm in §6.6. Additional aspects of the theory, includ-
ing characteristic hyperplanes, Spencer cohomology and the Goldschmidt
version of the Cartan-Kähler Theorem, are given in §6.7. In the remainder
of the chapter we give numerous examples, beginning with elementary prob-
lems coming mostly from surface theory in §6.8, then an example motivated
by variation of Hodge structure in §6.9, then the Cartan-Janet Isometric Im-
mersion Theorem in §6.10, followed by a discussion of isometric embeddings
of space forms in §6.11 and concluding with a discussion of calibrations and
calibrated submanifolds in §6.12.
135
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
136 6. The Cartan Algorithm for Linear Pfaffian Systems
Exercise 6.1.2: Let (I, J) be a linear Pfaffian system as above. Let π , for
1 ≤ ≤ dim Σ − n − s, be a collection of 1-forms such that T ∗ Σ is locally
spanned by θa , ω i , π . Show that there exist functions Aai , Tija defined locally
on Σ such that
and the same calculation shows that the pullback of this system to any
submanifold Σ ⊂ J 2 (R2 , R2 ) is linear Pfaffian. More generally, we have
Example 6.1.4. Any system of PDE expressed as the pullback of the con-
tact system on J k (M, N ) to a subset Σ is a linear Pfaffian system. If M
has local coordinates (x1 , . . . , xn ) and N has local coordinates (u1 , . . . , us ),
then J k = J k (M, N ) has local coordinates (xi , ua , paL ), where L = (l1 , . . . , lp )
ranges over symmetric multi-indices with 1 ≤ |L| ≤ k. In these coordinates,
the contact system I is {θL̃a = dpaL̃ − paL̃j dxj }, where L̃ is a symmetric multi-
index with 1 ≤ |L̃| ≤ k − 1, θLa = dpa − pa dxj }, and J = {θ a , θ a }. On
L Lj L̃
J k (M, N ),
)
dθa = −dpaj ∧ dxj
≡ 0 mod J,
dθL̃a = −dpaL̃j ∧ dxj
dθ = dy 1 ∧ dy 2 + dy 3 ∧ dy 4 + dy 5 ∧ dx
y3 1
≡ (dy 3 − dy ) ∧ dy 4 mod{θ, dx}.
y1
In this case, (I, J) is not a linear Pfaffian system.
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
6.2. First example 137
Exercises 6.1.6:
1. On ASO(n + s), with the notation from §2.5, let I = {ω a } and J =
{ω a , ω j }. The integral manifolds of (I, J) are the first-order adapted lifts of
immersed submanifolds M n ⊂ En+s . Show that (I, J) is a linear Pfaffian
system.
n+1
2. On ASO(n + s) × Rs( 2 ) , let the second factor have coordinates hajk =
hakj . Let I = {ω a , ωja − hajk ω k } and J = {ω a , ωja − hajk ω k , ω j }. Show that
the integral manifolds of (I, J) are the first-order adapted lifts of immersed
submanifolds M n ⊂ En+s together with their second fundamental forms,
and that (I, J) is a linear Pfaffian system.
Input:
linear Pfaffian system
Rename Σ as Σ - (I, J) on Σ;
calculate dI mod I
6
?
N Q
Q Q Y
Q Q
QQ Q Is [T ] = 0? Q
- ?
Q Is Σ empty? Q
Q k
Q QQ
Q
Q Q N
Y Q
Q ?
? Q Restrict to Σ ⊂ Σ
Q
Done: defined by [T ] = 0
there are no and Ω |Σ = 0
integral manifolds
Remark 6.2.2. The set defined by setting the torsion to zero may have
many components and strata (as an analytic space). To find all integral
manifolds to a system, one would have to restrict to each smooth stratum
in turn.
denoted
I = {θa },
J = {θa , ω i },
and we have
(6.2) dθa ≡ −dpai ∧ ω i mod I.
However, pulled back to Σ, the 1-forms dpai are not all independent. Say
we choose forms π , where dim J + 1 ≤ ≤ dim Σ, such that {θa , ω i , π }
gives a local coframing of Σ. Since Σ is defined by homogeneous equations
Bari pai = 0, we may choose the π so that
dpai = −Aai π
for some constants Aai such that Bari Aai = 0. Thus we may rewrite (6.2) as
(6.3) dθa ≡ Aai π ∧ ω i mod I.
Observe that (J/I)x V ∗ and Ix W ∗ . When we work with general
linear Pfaffian systems, these will be the definitions for V and W . The
tableau is recovered by taking
A := {Aai v i ⊗wa ⊆ V ∗ ⊗W | dim J + 1 ≤ ≤ dim Σ}.
We saw in Chapter 5 that if the tableau A is involutive, we have local exis-
tence of integral manifolds, roughly depending on sl functions of l variables,
where l is the Cartan integer and sl is the character of the tableau. For an
arbitrary linear Pfaffian system with no torsion, we will define a tableau at
a general point x ∈ Σ, and if this tableau is involutive, we will again have
local existence of integral manifolds.
Our informal algorithm now looks like:
Input:
linear Pfaffian system
Rename Σ as Σ - (I, J) on Σ;
calculate dI mod I
?
6
6
? N
Q Q
Q
N
Q Y Q
Q Q Q
Q Is [T ] = 0? Q - Is tableau Q
Is Σ empty?Q Q Q involutive?
Q Q Q
Q k
Q QQ Q
QQ Q N Q
Y Q Y
Q ?
? Q Restrict to Σ ⊂ Σ ?
Q
Done: defined by [T ] = 0 Done:
there are no and Ω |Σ = 0 local existence of
integral manifolds integral manifolds
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
140 6. The Cartan Algorithm for Linear Pfaffian Systems
Remark 6.3.1. We can consider the more general case of a linear constant-
coefficient but not necessarily homogeneous system. Say there are constants
Cir such that the equations are of the form
Bari pai = Cir xi .
Then we must have
dpai = −Aai π + Tija ω i ,
where the Aai are as before, and Tija are constants which satisfy Cir = Barj Tija .
Now we have apparent torsion:
dθa ≡ Aai π ∧ ω i + Tija ω i ∧ ω j mod I.
As seen in Exercise 6.2.1, it might be possible to modify the π , by adding
multiples of the ω j ’s, so that the Tija all zero, i.e., so that the apparent torsion
is absorbed. We examine the absorption of apparent torsion systematically
in §6.5.
Input:
“Prolong”, i.e., start over
linear Pfaffian system
- on a larger space Σ̃;
Rename Σ as Σ (I, J) on Σ;
rename Σ̃ as Σ
calculate dI mod I
and new system as (I, J)
6
6
? N
Q Q
Q
N
Q Y Q
Q Q Q
Q Is [T ] = 0? Q - Is tableau QQ
Is Σ empty?Q Q Q involutive?
Q Q Q
Q QQ Q
QQQ kQ N Q
Y Q Y
Q ?
? Q Restrict to Σ ⊂ Σ ?
Q
Done: defined by [T ] = 0 Done:
there are no and Ω |Σ = 0 local existence of
integral manifolds integral manifolds
We will say that a manifold admits a local isometric embedding into Eucli-
dean space if an isometric embedding exists on some neighborhood of any
given point.
The above problem is stated vaguely—we will study it in the following
form:
Determine a function r(n) with values in Z+ ∪ ∞ such
that every analytic Riemannian manifold of dimension n
admits a local isometric embedding into En+r .
While the answer to this question is known—in fact we will find r(n) in
§6.10—the corresponding question in the C ∞ category is still open.
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
142 6. The Cartan Algorithm for Linear Pfaffian Systems
In coordinates, choosing a metric corresponds to choosing n+1
2 func-
tions of n variables subject to some open (nondegeneracy) conditions, and
choosing a map Rn → Rn+r is the choice of n + r functions of n variables.
So the determined case (the
case where there are as many equations as un-
n
knowns) is when r = 2 . We will specialize to this case when we discuss
the Cartan-Janet Theorem in §6.10.
As usual, instead of working on M × En+r , we will work on a larger
space that takes into account the group actions. Let FM and FEn+r be
the respective orthonormal frame bundles. Fix index ranges 1 ≤ i, j ≤ n,
n + 1 ≤ μ, ν ≤ n + r. We use η to denote forms on FM and ω to denote
forms on FEn+r . Recall from §2.6 the structure equations on FM :
dη j = −ηkj ∧ η k ,
dηji = −ηli ∧ ηjl + 12 Rjkl
i
ηk ∧ ηl ,
i
where the functions Rjkl : FM → R are the coefficients of the Riemann
curvature tensor. We set up an EDS whose integral manifolds are graphs
Γ ⊂ FEn+r ×FM of isometric embeddings M n → En+r . As usual, we commit
the standard abuse of notation, omitting pullbacks.
To simplify our calculations, we restrict to graphs Γ on which ω μ = 0
and ω i = η i , so Tx M is spanned by e1 , . . . , en , and the first n basis vectors
of Tx M and Tx En+r are aligned. On FM × FEn+r , define the Pfaffian system
I = {ω μ , ω j − η j }diff
with independence condition Ω = η 1 ∧ . . . ∧ η n , so that
I = {ω μ , ω j − η j },
J = {ω μ , ω j , η j }.
We calculate
dω μ = −ωjμ ∧ ω j − ωνμ ∧ ω ν ,
d(ω i − η i ) = −ωji ∧ ω j − ηji ∧ η j − ωνi ∧ ω ν .
Thus,
)
dω μ ≡ 0
mod J,
d(ω i − η i ) ≡ 0
and our system is a linear Pfaffian system, so we can begin the algorithm.
We calculate
)
dω μ ≡ −ωjμ ∧ η j
(6.4) mod I.
d(ω i − η i ) ≡ −(ωji − ηji ) ∧ η j
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
6.4. The local isometric embedding problem 143
There is no torsion, since the forms ωjμ , (ωji − ηji ) are linearly independent.
At a point x ∈ Σ, let V ∗ = (J/I)x , W ∗ = Ix . We can use the structure
equations (6.4), just as we used (6.3), to define the tableau
μ
aj
A= ,
aij
where the entries are subject only to the relations aij + aji = 0. Here the π ’s
are the ωjμ and the (ωji − ηji ) for i < j. The Cartan characters (see §3.1) are
given by sj = r + (n − j).
Write an element of A(1) as (aijk wi + aμjk wμ )⊗v j v k . We see that aijk =
−ajik = aikj , so aijk = 0 for all i, j, k. The only relations on the aμjk are
aμjk = aμkj . Thus dim A(1) = r n+1
2 < s1 + 2s2 + . . . + nsn , and the tableau
is not involutive.
We continue our algorithm, starting over on a larger space where the
second derivatives are included as independent variables. In this case, we
will call these new variables hμij . (They will turn out to be the coefficients
of the second fundamental form.) We start over on the manifold
n+1
Σ̃ = FM × FEn+r × Rr( 2 ).
˜ J)
The system (I, ˜ defined by
Here we are presented with the same situation as in §6.2: there is torsion,
and so there are no integral elements except at points of Σ̃ where
μ μ
(6.6) i
Rjkl − (hik hjl − hμil hμjk ) = 0.
μ
holds. Assume for the moment that the pullback to Σ of our independence
condition is nonvanishing, so that we may go on to calculate our new tableau:
(6.7) d(ω μ − hμ ω k ) ≡ (−dhμ − hν ω μ + hμ ω l + hμ ω l ) ∧ ω k mod I.
j jk jk
˜
jk ν jl k kl j
μ
To simplify notation, let = − + πjk −dhμjk
+ hμkl ωjl . On Σ̃, hνjk ωνμ hμjl ωkl
μ
the forms πjk are all independent, but when we restrict to Σ , they satisfy
relations obtained by differentiating (6.6).
The next step is to check the torsion. First, we restrict to a simple,
special case. Other cases will be treated below and in §6.10.
Case n = 2, r = 1. The only independent curvature component is R1212 =
R, and equation (6.6) becomes
(6.8) R = h11 h22 − (h12 )2 ,
where we have suppressed the μ = 3 index on h.
Let θj = ωj3 − hjk ω k . Then (6.7) specializes to
1
θ1 −dh11 + 2h12 η12 −dh12 − (h11 − h22 )η12 η ˜
d ≡ ∧ 2 mod I.
θ2 −dh12 − (h11 − h22 )η12 −dh22 − 2h12 η12 η
Since R : FM → R is constant on the fibers of the projection to M , we
may write dR = R1 η 1 + R2 η 2 . Differentiating (6.8) gives
(6.9) R1 η 1 + R2 η 2 − h11 dh22 − h22 dh11 + 2h12 dh12 = 0.
Upon setting
π1 = −dh11 + 2h12 η12 ,
π2 = −dh12 − (h11 − h22 )η12 ,
π3 = −dh22 − 2h12 η12 ,
the relation (6.9) becomes
(6.10) R1 η 1 + R2 η 2 + h11 π3 + h22 π1 − 2h12 π2 = 0.
Aside 6.4.2. The seeming miracle that the η12 terms all cancel is due to
the fact that the form η12 is dual to a Cauchy characteristic vector field (see
Chapter 7). Intuitively, integral curves of this vector field correspond to
spinning the frame within the tangent space of the surface, which doesn’t
change the embedding.
In fact, Theorem 7.1.21 implies that our EDS actually lives on the quo-
tient of Σ by this foliation. Since integral manifolds of the EDS exist if and
only if torsion vanishes on the quotient manifold, we expect that the torsion
upstairs will not involve η12 .
In n dimensions, the forms ηji are dual to Cauchy characteristics for
the isometric embedding system. Rather than ignoring them, we could have
incorporated them into the independence condition. Then integral manifolds
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
6.4. The local isometric embedding problem 145
Aside 6.4.5 (Another easy case). Suppose that the codimension r is small
enough that n (n12−1) > n+1
2 2
2 r. To see if (6.6) is solvable, let V = R ,
n
γ : S 2 V ∗ ⊗W → K(V ),
haij v i v j ⊗wa → (haik hajl − hail hajk )v i ⊗v j ⊗v k ⊗v l .
a
Torsion. Recall that the terms Tija ω i ∧ ω j in (6.1) are called apparent tor-
sion. If it is possible to choose new π̃ ’s such that T̃ija = 0, we say the
apparent torsion is absorbable. If it is not possible, we will say there is tor-
sion. In other words, a particular choice of complement to Jx ⊂ Tx∗ Σ may
yield apparent torsion at x, while we say there is torsion at x if all choices
yield apparent torsion. In that case there are no integral elements at x. We
now discuss these statements without reference to bases.
The change in the Tija that a change of complement π̃ = π +ei ω i effects
is
Tija = Tija + 12 (Aaj ei − Aai ej ).
a = Aa e − Aa e and let
Let Eij j i i j
δ : W ⊗V ∗ ⊗V ∗ → W ⊗Λ2 V ∗
be the skew-symmetrization map. The space of possible modifications to
apparent torsion is δ(A⊗V ∗ ), because E := Eij
a w ⊗v i ⊗v j lies in δ(A⊗V ∗ ).
a
Then the torsion of (I, J) at x is defined to be
[T ] := [Tija wa ⊗v i ∧ v j ] ∈ W ⊗Λ2 V ∗ /δ(Ax ⊗V ∗ ).
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
148 6. The Cartan Algorithm for Linear Pfaffian Systems
We will use the notation H 0,2 (A) = W ⊗Λ2 V ∗ /δ(Ax ⊗V ∗ ). (This is one of
the Spencer cohomology groups, which will be discussed in §6.7.)
Exercise 6.5.4: Compute the Cjr of Remark 6.5.2 in terms of the coefficients
Aai , Tija .
Remark 6.6.2. Each time one prolongs, there may be many different com-
ponents of Vn (I) on which to restrict. To find all possible integral manifolds,
one must carry out the algorithm on each component. The Cartan-Kuranishi
Prolongation Theorem [27] says in effect that this process terminates even-
tually, but gives no hint of how long it will take.
Proof. Let π̃ia ≡ πia +paij ω j mod I. The π̃ia must satisfy the symbol relations
Bari π̃ia ≡ 0 mod I. But Bari π̃ia ≡ Bari (πia + paij ω j ) ≡ Bari (paij ω j ) mod I, since
πia also satisfies the symbol relations. This implies Bari paij = 0 for all j,
since the ω j are independent, so paij wa ⊗v i ⊗v j ∈ A⊗V ∗ . But we also need
π̃ia ∧ ω i ≡ πia ∧ ω j + paij ω i ∧ ω j ≡ 0 mod I. We already have πia ∧ ω i ≡ 0 mod I,
so this forces paij ω i ∧ ω j = 0, which implies paij = paji .
(1)
Remark 6.7.2. If there is no torsion, Ax is the space of integral elements
at x. After fixing an integral element, it becomes a linear space. This is the
reason for the adjective “linear” in the name linear Pfaffian systems.
be defined by
δj (f ⊗ξ) = df ∧ ξ,
The Kuranishi Prolongation Theorem (see [27]) implies that after a finite
number of prolongations a tableau will become involutive (including the
possibility that it becomes empty). This implies
Theorem 6.7.5 (Goldschmidt version of Cartan-Kähler [74]). Let (I, J)
be a linear Pfaffian system on an analytic manifold Σ and let x ∈ Σ be a
general point. If H p,2 (Ax ) = 0 for all p, then there exist integral manifolds
passing through x.
This theorem is useful because sometimes one can show that the groups
H p,2 (Ax ) are zero without actually calculating the prolongations (see [70]).
More generally, it is sufficient to show that the corresponding vector
bundles with induced torsion sections are such that the sections are zero;
see Chapter 9 for the construction of the associated vector bundles.
Now recall the symbol mapping from Definition 5.6.2, defining the char-
acteristic variety ΞA for A = Ax . If ξH ∈ PE ∗ corresponds to H ⊂ E, then
ξH ∈ ΞA if and only if
Barn wa = 0
for some nonzero vector w ∈ W . In other words, ξH ∈ ΞA if and only if
the tableau contains a matrix with the first n − 1 columns zero and the
last column nonzero, or equivalently, some linear combination of the πna is
linearly independent from the forms in the first n − 1 columns, so that v is
not uniquely determined by (6.18). We have now proved
Theorem 6.7.7. A hyperplane H ⊂ E ⊂ Tx Σ is characteristic if and only
if the corresponding element ξH ∈ PE ∗ is such that ξH ∈ ΞA .
Example (6.5.3 continued). Let Σ ⊂ J 2 (R2 , R) be the submanifold defined
by a regular second-order PDE and let A be the tableau (6.16) at a point
u ∈ Σ. Then a nonzero covector ξ = ξ1 dx + ξ2 dy ∈ V ∗ belongs to ΞA if and
only if the map σξ (w) := w ⊗ ξ mod A has a nonzero kernel. Write
⎛ ⎞
w1 ξ1 w1 ξ2
w ⊗ ξ = ⎝w2 ξ1 w2 ξ2 ⎠ .
w3 ξ1 w3 ξ2
Then w ⊗ ξ ∈ A if and only if w1 = 0, w2 ξ2 = w3 ξ1 and Fr w2 ξ1 + Fs w2 ξ2 +
Ft w3 ξ2 = 0. This homogeneous system of linear equations for the wa has a
nontrivial solution if and only if
(6.19) Fr ξ12 + Fs ξ1 ξ2 + Ft ξ22 = 0.
6.8. Examples
The examples with Roman numerals are adapted from Cartan’s 1945 treatise
[40].
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
6.8. Examples 155
where a, b, c, e are arbitrary. Thus dim A(1) = 4 = s1 + 2s2 , and the system
is involutive. The Cartan-Kähler Theorem implies that integral manifolds
depend on one function of two variables. This should come as no surprise,
since locally any surface can be written as a graph z = f (x, y).
Example 6.8.2 (Cartan III: Linear Weingarten surfaces). Let A, B, C be
constants, with A = 0, and let AK + 2BH + C = 0 be a linear relation
between functions H and K. We set up and describe the space of general
integral manifolds for first-order adapted lifts of surfaces in R3 , together
with the components of the second fundamental form, having Gauss and
mean curvatures satisfying this relation. (Note that this class of examples
includes constant mean curvature and constant Gauss curvature surfaces.)
This is the system above restricted to a codimension one submanifold
Σ ⊂ ASO(3) × R3 hij . The submanifold is defined by the equation
A(h11 h22 − h212 ) + B(h11 + h22 ) + C = 0.
The differential of this equation implies that on Σ we may write
dh12 = αdh11 + βdh22
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
156 6. The Cartan Algorithm for Linear Pfaffian Systems
may write
a −b 0
A= a, b ∈ R .
b a 0
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
6.8. Examples 157
Matrices in the Lie algebra u(3) of U (3) = SO(6) ∩ Sp(R6 , ω) have the
form
⎛ ⎞
0 −α12 −α13 −ρ1 −β12 −β13
⎜α12 0 −α23 −β12 −ρ2 −β23 ⎟
⎜ 3 ⎟
⎜α1 α23 0 −β13 −β23 −ρ3 ⎟
⎜ ⎟.
⎜ ρ1 β 2 β 3 0 α12 α13 ⎟
⎜ 1 1 ⎟
⎝β 2 ρ2 β23 −α12 0 α23 ⎠
1
3
β1 β2 3 ρ3 −α13 −α23 0
Let FU (3) be the unitary frame bundle for C3 , and let I = {ω 4 , ω 5 , ω 6 } be
defined on FU (3) , with independence condition Ω = ω 1 ∧ ω 2 ∧ ω 3 . We have
structure equations
⎛ ⎞
0 0 ... 0
⎜ω 1 0 −α12 −α13 −ρ1 −β12 −β13 ⎟
⎜ 2 ⎟
⎜ω α12 0 −α23 −β12 −ρ2 −β23 ⎟
⎜ 3 ⎟
d(x, e1 , . . . , e6 ) = (x, e1 , . . . , e6 ) ⎜ 3
⎜ω α1 α2
3 0 −β13 −β23 −ρ3 ⎟
⎟.
⎜ω 4 ρ1 β 2 3
β1 0 α12 α1 ⎟
3
⎜ 1 ⎟
⎝ω 5 β 2 ρ2 β2 −α1
3 2 0 α23 ⎠
1
ω 6 β13 β23 ρ3 −α13 −α23 0
Let I = {ω 3 , θ13 , θ23 } and let J = {ω 3 , θ13 , θ23 , ω 1 , ω 2 }, where we set θ13 =
ω13 − kω 1 and θ23 = ω23 + kω 2 . The structure equations are
⎛ 3⎞ ⎛ ⎞
ω 0 0 1
ω
d ⎝ θ13 ⎠ ≡ ⎝ −dk −2kω12 ⎠ ∧ 2 .
ω
θ2 3 −2kω12 dk
Write dk = k1 ω 1 + k2 ω 2 . The symbol relations are
⎫
π10 ≡ π20 ≡ 0 ⎪
⎪
⎪
⎪
⎬
π1 − π2 ≡ 0
2 1
mod I.
π11 + π22 ≡ 0 ⎪
⎪
⎪
⎪
⎭
π11 ≡ k1 ω 1 + k2 ω 2
Again we can change bases to attempt to get rid of the apparent torsion. If
we rewrite our equations as
⎛ 3⎞ ⎛ ⎞
ω 0 0 1
ω
d ⎝ θ1 ⎠ ≡ ⎝
3 0 −2kω1 + k2 ω ⎠ ∧
2 1
2 ,
ω
3
θ2 −2kω1 − k1 ω
2 2 0
the symbol relations become
⎫
π10 ≡ π20 ≡ 0 ⎪
⎪
⎬
π12 − π21 ≡ −k1 ω 1 + k2 ω 2 mod I.
⎪
⎪
π11 ≡ π22 ≡ 0 ⎭
Here, the forms ω13 − k1 ω 1 , ω23 − k2 ω 2 are in the ideal to ensure that (ω 1 )⊥ ,
(ω 2 )⊥ correspond to principal directions on the surface, and k1 , k2 will be
the principal curvatures. If we can multiply ω 1 , ω 2 by a common function u
to make them both closed, then the fact that closed forms are locally exact
implies local existence of the isothermal coordinates.
Note that I is not a Pfaffian system. But if we adjoin new variables
u1 , u2 such that du = u1 ω 1 + u2 ω 2 , then the generators of I of degree two
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
6.9. Functions whose Hessians commute 161
are
d(uω 1 ) = (−u2 ω 1 + uω12 ) ∧ ω 2 ,
d(uω 2 ) = (−u1 ω 2 − uω12 ) ∧ ω 1 .
Thus, an equivalent system is (I, Ω) defined on ASO(3)×R2k1 ,k2 ×R1u ×R2u1 ,u2
with
I := {ω 3 , ω13 − k1 ω 1 , ω23 − k2 ω 2 , du − u1 ω 1 − u2 ω 2 , uω12 − u2 ω 1 + u1 ω 2 },
which is a linear Pfaffian system.
Exercise 6.8.15: Determine the conditions on k1 , k2 such that there exist
integral manifolds of the above system. (The answer will be in the form of a
PDE that must be satisfied in addition to the Codazzi equation.) Determine
the space of integral manifolds for generic functions k1 , k2 that satisfy these
conditions.
Exercises 6.8.16:
1. (Cartan VI) Set up an EDS for pairs of surfaces S, S ⊂ E3 together
with a local diffeomorphism φ : S → S that is an isometry preserving a
family of asymptotic directions. What does the general solution depend on?
Geometrically characterize the surfaces S admitting a noncongruent mate
S.
2. (Cartan VII) Set up an EDS for pairs of surfaces S, S ⊂ E3 together
with a local diffeomorphism φ : S → S that is an isometry preserving a
family of lines of curvature. What does the general solution depend on?
Geometrically characterize the surfaces S admitting a noncongruent mate
S.
3. (Cartan XII) Set up an EDS for pairs of surfaces S, S ⊂ E3 such
that there exists a diffeomorphism φ : S → S that is a conformal map
preserving both families of asymptotic directions. What does the general
solution depend on? Show that, if S is not minimal, there exists at most
one such S , up to congruence. In the special case where S is minimal, what
are the surfaces S such that there exists such a φ?
where the aij ’s are from (6.20) and the bijk ’s are additional constants which
must satisfy bijk = bikj = bjik .
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
6.9. Functions whose Hessians commute 163
Let 2 ≤ ρ, σ ≤ s and let mρij = (liρ − ljρ )/(li1 − lj1 ). Then ψ ρij = mρij ψ 1ij for
each i = j, and consequently mρij bijk = mρik bikj . Since generically the values
of the mρij are all distinct, all the bijk must be zero. The only remaining
freedom in choosing an integral element are the “diagonal” elements, i.e.,
we may set
n
dim A (1)
= n + (s − 1)n = 2
2
+ s1 .
2
Thus if we are in the case where n2 ≤ sn, i.e., s ≥ n−1
2 , then the system
n
is involutive and general integral manifolds depend on 2 functions of two
variables and sn functions of one variable.
There are many other families of integral manifolds of this system. For
example: a naı̈ve way to construct s functions of n variables whose Hessians
commute is to take the first function to be arbitrary, and the rest of the
form f a (x1 , . . . , xn ) = ua (x1 + . . . + xn ), so that the Hessians of f a are all
multiples of the identity matrix. Such integral manifolds are called singular
solutions.
From the perspective of EDS, these integral manifolds are obtained by
restricting to the stratum Σ ⊂ Σ where paij = C a δij for a > 1. Note that
they depend roughly on one function of n variables, and thus the space of
these integral manifolds is larger than for integral manifolds with generic
initial data.
Singular solutions. So far we have discussed integral manifolds of a system
(I, J) on Σ obtained by solving a series of Cauchy problems starting at a
generic point x ∈ Σ and a generic flag in Tx Σ among flags leading to n-
dimensional integral elements. A singular solution to an exterior differential
system is an integral manifold obtained by either starting at a nongeneric
point y ∈ Σ, or taking a nongeneric flag for the Cauchy problems.
If one is at a nongeneric point y ∈ Σ and restricts to the submanifold
Σ ⊂ Σ of points near y having the same nongenericity properties, one can
again use the Cartan algorithm restricted to Σ to find integral manifolds.
For singular solutions corresponding to a nongeneric flag, one can often
still apply Cartan’s algorithm to a modified system. In small dimensions,
singular solutions can often be treated on a case by case basis; see the
numerous examples in [40].
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
164 6. The Cartan Algorithm for Linear Pfaffian Systems
s1 + 2s2 + . . . + nsn
n−1
p
n−p+1
≤ p r(n − p + 1) − (n − p + 1) − (p − 1) .
2 2
p=1
The summation on the right-hand side equals (6.23), so we obtain the desired
inequality.
Consider
0 = (v ∧ v0 ) γ(h) = (v hξ ) ∧ v ξ .
ξ
By the Cartan Lemma A.1.9 we have
(6.26) {v hξ | v ∈ V, 1 ≤ ξ ≤ dim U0 } ⊆ U0 .
This implies Singloc{hξ } = {v hξ }⊥ ⊇ U0 ⊥ and thus is nonempty. Let
v1 ∈ Singloc{hξ }. There is some index φ and v2 ∈ V such that hφ (v1 , v2 ) =
0, because Singloc(h) = 0. For all v ∈ V , v0 + v will be generic if is
sufficiently small, because v0 is generic. Consider
U1 = {(v0 + v2 ) hs } = {(v0 + v2 ) hξ , v2 hφ }.
By (6.26), {(v0 +v2 ) hξ } ⊆ U0 , and in fact we have equality because v0 +v2
is generic. But {v2 hφ } ⊂ U0 , which implies dim U1 > dim U0 , giving a
contradiction.
To prove the normal form, consider a basis v1 , . . . , vn of V where v1 is
h-generic. Let 2 ≤ ρ, σ ≤ n. Again, since
(vρ ∧ v1 ) γ(h) = 0,
the Cartan Lemma implies that vρ hs = Cts (ρ)(v1 ht ) for some symmetric
matrices C(ρ). Using (vρ ∧ vσ ) γ(h) = 0, we see that the matrices commute
and are therefore simultaneously diagonalizable by an orthogonal matrix,
say B ∈ O(W ). Let q s = Bts ht . We have arranged things so that for each
index s, dim{vj q s | 1 ≤ j ≤ n} = 1, so rank q s = 1 for all s.
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
168 6. The Cartan Algorithm for Linear Pfaffian Systems
It is easy to verify that this determines an integral element, so dim A(1) ≥
n2 + 2 n2 . But we already know that dim A(1) ≤ n2 + 2 n2 , so we have
equality, and the tableau is involutive.
In summary, we have shown:
Theorem 6.11.3 (Cartan [37]). Local isometricembeddings En → E2n
with immersive Gauss maps exist and depend on n2 functions of 2 variables.
R = −γ 0 (g0 , g0 ).
Let M (t) denote the parametrized line segment from a to b and let M (t)
denote any rectifiable path from a to b. Without loss of generality, use
coordinates (x1 , x2 ) such that the line segment is along the x1 -axis.
M
~
a M b
We have
dx1 (M (t)) = |M (t)|,
dx1 (M̃ (t)) ≤ |M̃ (t)|.
Letting φ = dx1 , we have
i
ω = (dz 1 ∧ dz 1 + . . . + dz m ∧ dz m ).
2
1 ∧k
Proposition 6.12.5. Let (X, ω) be Kähler and let φ = k! ω . Then φ is
a calibration and Facex (φ) = G(C , C ) ⊂ G(2k, Tx X). In other words,
k m
|dz(E)| = |E ∧ J(E)|.
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
https://doi.org/10.1090//gsm/175/07
Chapter 7
Applications to PDE
175
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
176 7. Applications to PDE
Cauchy characteristics.
Definition 7.1.5. Let I be an EDS on Σ. A vector field v ∈ Γ(T Σ) is a
Cauchy characteristic vector field of I if v ψ ∈ I for all ψ ∈ I.
The flow lines of v are called Cauchy characteristic curves (or, simply,
Cauchy characteristics) of I.
Exercises 7.1.6:
1. If v is a Cauchy characteristic vector field for I, then it is also a symme-
try.
2. The Lie bracket of any two Cauchy characteristic vector fields is a Cauchy
characteristic vector field.
3. Show that v is a Cauchy characteristic vector field if and only if v ψ ∈ I
for a set of forms ψ that generate I algebraically.
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
178 7. Applications to PDE
Exercise 7.1.9: Generalize this argument to the case where N0 has arbi-
trary dimension.
Example 7.1.10 (Surfaces in E3 ). Recall from §2.1 that the integral man-
ifolds of I = {ω 3 }diff on F = Fon (E3 ) are the first-order adapted lifts of
surfaces Σ2 ⊂ E3 . This system admits a Cauchy characteristic vector field
as follows:
Let (ω j )∗ , (ωji )∗ be a basis of vector fields on F dual to the entries of the
Maurer-Cartan form. Write v = aj (ω j )∗ + aji (ωji )∗ . Then v ω 3 = 0 implies
that a3 = 0, and
v dω 3 = v (−ω13 ∧ ω 1 − ω23 ∧ ω 2 )
= a1 ω13 − a13 ω 1 + a2 ω23 − a23 ω 2
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
7.1. Symmetries and Cauchy characteristics 179
large values of t, the surface defined by (7.7) may not be a graph over the
xt-plane.
Exercises 7.1.15:
1. Consider the PDE u2x + u2y = 2u.
(a) Show that, when x, y, p, q are used as coordinates, the Cauchy charac-
teristic line field is generated by v = p(∂p + ∂x ) + q(∂q + ∂y ).
(b) Find all regular solutions satisfying initial condition u(x, 0) = 25 x2 .
2. Solve 12 (u2x + u2y ) + xux + yuy = u subject to u(x, 0) = (1 − x2 )/2.
3. [105, Ch. 1] Solve the PDE uy = u3x with the initial condition u(x, 0) =
2x3/2 . Show that the only solutions which are regular over all of R2 are
linear functions of x and y.
Exercise 7.1.17: Show that the only quasilinear systems for u, v which have
nontrivial Cauchy characteristics are those that uncouple into two first-order
equations with the same characteristics.
Proof. Let J be a Frobenius system such that there exists a finite collection
of forms ψ1 , . . . , ψm ∈ ΛJ which generate I algebraically. In terms of these
generators, any form Φ ∈ I can be written as
m
Φ= βk ∧ ψk .
k=1
Let v be any vector field annihilated by J. Then v Φ = mk=1 (v βk ) ∧ ψk ,
which belongs in I. Hence v ∈ Γ(A(I)). Since v and Φ are arbitrary, we
conclude that C(I) ⊂ J.
Furthermore, the smallest such J cannot be larger than C(I). For, sup-
pose C(I) has codimension l > 0 within J, and let π1 , . . . , πl be linearly
independent forms in J with πk ∈ / C(I). Partition the generators for I into
ϕ1 , . . . , ϕm ∈ ΛC(I), plus generators ψ1 , . . . , ψn involving the πk . If v is any
vector in A(I), and ψ1 has lowest degree among the ψ’s, then v ψ1 must be
expressible in terms of the ϕ’s. In particular, any term in ψ1 involving one
of the πk may be omitted since that term already is generated by the ϕ’s.
It follows that we may take all the generators to be in Λ C(I).
The above proposition implies that we may calculate the retracting space
simply by obtaining a minimal set of algebraic generators for the ideal.
Exercises 7.1.19:
1. Consider an EDS generated by a single 1-form θ ∈ Ω1 (Σ). Show that
the dimension of C(I) will be 2k + 1, where k is the largest integer such that
θ ∧ (dθ)k |Σ = 0.
2. [76, §126] Consider the system of two second-order PDE given by uxx +
3 uxy = 0 and uxy uyy = 1. Let I be the pullback of the standard contact
1 3
about t = 0.) Then (7.8) implies that φ∗t ψ ∈ Γ(I), and thus (φt ◦σ)∗ ψ ∈ Γ(I)
for any t.
Proof. Let J = C(I), and let IJp be the bundle spanned by semi-basic p-
forms in I. Then Proposition 7.1.20 implies that IJp is the pullback of the
degree p piece of a differential ideal I on Q (see the exercises below).
alg ⊂ I. Moreover, since I has
Because Γ(IJp ) ⊂ I p , we have {π ∗ I}
a set of algebraic generators that are semi-basic (see Proposition 7.1.18),
alg .
I ⊂ {π ∗ I}
In practice, one exploits the existence of the quotient manifold and its
EDS to obtain a better geometric understanding of the system. However, it
may be easier to continue to calculate using the coordinates or coframing on
the original space. For example, if the original space is a submanifold of a
jet bundle, or a Lie group, then these spaces come equipped with coframes
for which the structure equations are particularly simple.
Exercises 7.1.22:
1. Show that I, as constructed degree by degree in the proof of Theorem
7.1.21, is actually a differential ideal on Q.
2. Let Σ, π, Q and J be as above. Suppose Q has dimension 2m, and Ω is
a 2-form in J such that Ωm = 0. Under what conditions is Ω the pullback
of a well-defined symplectic form on Q?
For the rest of this discussion we will assume the determinant is nonzero
at every point of Σ. In fact, we will assume the PDE is hyperbolic; however,
everything we do may be carried out in the elliptic case, by using complex-
valued forms (see §7.4).
Hyperbolic PDE and characteristic systems. Note that (7.11) implies that a
linear combination of the system 2-forms is congruent, modulo I, to a de-
composable 2-form with ξ as one of the factors. Therefore, when the PDE is
hyperbolic, there are two linearly independent decomposable 2-forms in the
ideal that are independent of the θ’s. We may choose linearly independent
forms π1 , π2 , ω1 , ω2 that are not in I 1 , such that π1 ∧ω1 and π2 ∧ω2 are these
decomposable forms in I 2 , and ω1 ∧ ω2 = 0 is the independence condition.
In order to simplify the tableau, we may also choose new forms θ̃1 , θ̃2 in I 1
so that
(7.12) dθ̃1 ≡ −π1 ∧ ω1 , dθ̃2 ≡ −π2 ∧ ω2 mod I.
(However, this is not immediately necessary.)
At each point of an integral surface for I, the tangent vectors to the
surface that are annihilated by ω1 or ω2 are characteristic in the sense defined
above. Moreover, since each πi pulls back to the surface to be a multiple of
ωi , each such vector is annihilated by all of the forms of one of the following
two characteristic systems:
M1 := {θ0 , θ1 , θ2 , π1 , ω1 }, M2 := {θ0 , θ1 , θ2 , π2 , ω2 }.
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
186 7. Applications to PDE
Exercises 7.2.4:
1. Assume that Fr = 0. Using (7.11), show that ξ = dy − m dx belongs to
the characteristic variety if and only if m is a root of the quadratic equation
(7.13) Fr m2 − Fs m + Ft = 0.
Assuming that this equation has distinct roots m1 , m2 , show that the de-
composable 2-forms in I are
Remark 7.2.5. Note that if the PDE (6.13) is quasilinear, then the equa-
tion (7.13) defining the characteristics only involves the variables x, y, z, p, q.
More generally, we say that a second-order PDE for one function of two
variables has first-order characteristics if there exists a rank three Pfaffian
system on J 1 (R2 , R) whose pullback to Σ is contained in one of the char-
acteristic systems of I. This is the case when (6.13) is a Monge-Ampère
equation (see §7.4); for other examples, see [76, Ch. IV]. The existence of
first-order characteristics is intrinsic to the system, since J 1 (R2 , R) is the
quotient of Σ by the foliation dual to the retracting space of I’s first derived
system (see §7.3).
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
7.2. Second-order PDE and Monge characteristics 187
Exercise 7.2.8: Show that a system of the form ux = f (v), vy = g(u), when
converted to an EDS on a codimension-two submanifold of J 1 (R2 , R2 ), gives
a hyperbolic EDS of class two. Show that it may also be encoded by a
hyperbolic EDS of class zero (i.e., generated by a pair of decomposable 2-
forms) defined on J 0 (R2 , R2 ).
r0 ≤ 12 m(m − 1),
r1 ≤ mr0 + 12 r0 (r0 − 1),
r2 ≤ r1 (r0 + m) + 12 r1 (r1 − 1),
....
For generic systems these inequalities are equalities; see [66] for further
inequalities and invariants of Pfaffian systems.
Because dω = 2dx ∧ θ2 , I (1) is not Frobenius. Hence the derived flag termi-
nates in I (2) = 0.
The two extra equations (7.14) which we impose on the solution of Li-
ouville’s equation are compatible with it, in the sense that the restriction of
I to the submanifold defined by them is involutive—in fact, it is Frobenius.
(In classical terminology, any extra equation that defines a submanifold to
which the system restricts to be involutive is called an “integral” for the
given PDE. If the extra equation is of lower order than the given PDE, then
it is an “intermediate integral”, and otherwise it is a “general integral”.)
The Frobenius condition means that the solution is uniquely determined,
up to a choice of constants, by the two arbitrary functions in (7.14). Thus,
we have a concrete realization of what Cartan’s Test predicts for this system,
namely that the solution depends on two functions of one variable.
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
192 7. Applications to PDE
Exercises 7.3.15:
1. Show that r − qs + pt = 0 is Darboux-integrable.
2. Determine for which functions f (v) and g(u) the system of Exercise 7.2.8
is Darboux-integrable.
3. Let m1 = m2 be the roots of the characteristic equation (7.13) for a
PDE of the form r + f (s, t) = 0.
(a) Show that each characteristic system contains one of the 1-forms ds +
mi dt, and that both of these forms are integrable.
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
7.3. Derived systems and the method of Darboux 193
Exercises 7.3.17:
1. Show that s = pq/(x − y)n is Darboux-integrable when n = 0, and
Darboux-integrable after one prolongation when n = 1.
2. Show that s+n(n−1)z/(x−y)2 , where n is a positive integer, is Darboux-
integrable only after n prolongations [54, IV, Ch. 3].
Remark 7.3.18. The problem of classifying which second-order PDE be-
come Darboux-integrable after a finite number of prolongations has a long
history and is still a subject of current research [7, 108].
One of the earliest results in this area is Lie’s proof that the only equa-
tions of the form s = f (u) that are Darboux-integrable after a finite number
of prolongations are those with f (u) = exp(au + b); see [76, §170] for a
proof. Another significant early result was Goursat’s classification [77] of
quasilinear equations s = f (x, y, u, p, q) that are Darboux-integrable with-
out prolongation. This classification was later extended, using symmetry
methods, by E. Vessiot; see [182] for more details.
Recently, some progress has been made on the related problem of de-
termining which systems of even class are Darboux-integrable [30]. As in
Exercise 7.2.8, such systems can arise from a system of first-order PDE for
two functions u(x, y), v(x, y) and its prolongations.
Remark 7.3.19 (Semi-integrability). If a rank-two Frobenius system is
present in only one of the two characteristic systems of a hyperbolic EDS,
it is said to be semi-integrable by the method of Darboux. In this case,
one can still construct solutions for the system using ODE techniques alone.
For simplicity, we will explain how this is done for a single second-order
hyperbolic PDE (cf. [182, Prop. 3.3]):
For dW, dX ∈ M1 , the imposed equation W = f (X) defines a six-
dimensional submanifold N ⊂ Σ on which the pullback of I has a Cauchy
characteristic vector field annihilated by M2 . Thus, we can construct an
integral surface of I by starting with an integral curve K transverse to
these Cauchy characteristics, and taking the union of the characteristics
through K. Constructing these characteristics involves solving a system
of first-order ODE for four unknowns which, along with the transversality
condition, depend on the choice of f .
Remark 7.3.20 (Darboux for parabolic PDE). Suppose that, for a para-
bolic second-order equation, the single characteristic system M contains a
rank two Frobenius system. Then M itself is Frobenius [175, §16.1]. The
contact system contains a well-defined rank two Pfaffian subsystem J, not
containing θ0 , such that M = C(J). By Theorem 7.1.21, J is well-defined
on the quotient manifold Q by the leaves of M. One can construct integral
surfaces of I by starting with an integral curve N ⊂ Q of J, and then writing
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
7.4. Monge-Ampère systems and Weingarten surfaces 195
dθ ≡ ω1 ∧ π1 + ω2 ∧ π2 mod θ.
dθ = −β0 ∧ θ + ω1 ∧ π1 + ω2 ∧ π2 ,
d(ω1 ∧ π1 ) = −β1 ∧ ω1 ∧ π1 ,
d(ω2 ∧ π2 ) = −β2 ∧ ω2 ∧ π2 .
(c) By differentiating the first structure equation, show that we can assume
β1 = β0 − aθ, β2 = β0 + aθ for some function a.
(d) By differentiating the other equations, show that a = 0 identically, and
conclude that dβ0 = 0.
(e) On a possibly smaller neighborhood, there exist functions λ, p, q, x, y, z
such that β0 = dλ, eλ ω1 ∧ π1 = dx ∧ dp, eλ ω2 ∧ π2 = dy ∧ dq, and eλ θ =
dz − p dx − q dy.
Thus, we recover the theorem of Bonnet, to the effect that each surface
of constant positive Gauss curvature is parallel to two surfaces of constant
mean curvature.
It follows that solutions of the PDE (1.4) for surfaces of constant positive
Gauss curvature may be obtained by a contact transformation from solutions
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
7.4. Monge-Ampère systems and Weingarten surfaces 199
We finish this section with three extended examples that show how char-
acteristics may be interpreted geometrically, and how to work with them
when the system is elliptic.
Example 7.4.17 (Pseudospherical surfaces and the sine-Gordon equation).
A pseudospherical surface is a surface S ⊂ E3 with Gauss curvature K ≡ −1.
Setting A = C = 1 and B = 0 in the above EDS shows that, in this case,
the 2-forms of I are spanned by the decomposable forms
Ψ + Θ = (ω13 − ω 2 ) ∧ (ω23 + ω 1 ), Ψ − Θ = (ω13 + ω 2 ) ∧ (ω23 − ω 1 ),
and the system is hyperbolic. Moreover, we may use the factors of the
decomposable forms to express the second fundamental form as
II = [ 12 (ω13 + ω 2 ) (ω23 + ω 1 ) − 12 (ω13 − ω 2 ) (ω23 − ω 1 )] ⊗ e3 .
This shows that each characteristic curve projects to be an asymptotic line
on S (see §2.7).
On an open subset of a pseudospherical surface that is free of umbilic
points, we may choose a Darboux framing (see §2.3). This breaks the Cauchy
characteristic symmetry of the EDS, but allows us to see how pseudospher-
ical surfaces are connected to solutions of the sine-Gordon equation (7.20).
any integral surface. Hence there exist local coordinates t1 , t2 such that
ω 1 = cos u dt1 and ω 2 = sin u dt2 .
Exercise 7.4.18: Verify that ω 1 / cos u and ω 2 / sin u are closed, and show
that x = (t1 − t2 )/2 and y = (t1 + t2 )/2 are arclength coordinates along the
asymptotic lines. This implies that the asymptotic lines form what is called
a Chebyshev net [55] on the surface.
gives the standard pseudosphere when a2 = 1 and gives Dini’s surface when
a2 = 1 (see [146] for pictures).
Hilbert’s Theorem states that pseudospherical surfaces in Euclidean
space cannot be geodesically complete. Hence, even if the sine-Gordon
solution is defined for all x and y, the corresponding surface will become
singular. In fact, ω 1 ∧ ω 2 = sin 2u dx ∧ dy shows that the surface will be sin-
gular whenever u is a multiple of π/2, i.e., the asymptotic lines are tangent
to each other.
Note that each of x + y and x − y is constant along one of the two sets
of lines of curvature. This observation is useful if we suppose that S is a
pseudospherical surface of revolution.
Exercise 7.4.19: Show that if S is a surface of revolution, the corresponding
sine-Gordon solution can be assumed to be of the form u = f (x + y). Then,
determine all sine-Gordon solutions of this form.
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
7.4. Monge-Ampère systems and Weingarten surfaces 201
Exercise 7.4.21: Verify that, under our identification, the point (7.25) in N
corresponds to the point in the unit sphere given by standard stereographic
projection
(x, y) → 2x, 2y, 1 − x2 − y 2 /(1 + x2 + y 2 )
when w = x + iy. Thus, the complex structure on S 2 defined by τ is the
same as that given by the usual stereographic projection from C.
The following exercises show how some well-known minimal surfaces can
be obtained from (7.26) for simple choices of f and g. (More generally, g(z)
can be a meromorphic function whose domain has nontrivial topology, and
this allows the resulting surface to have nontrivial topology itself.)
Exercises 7.4.23:
1. Identify the surfaces obtained from the Weierstrass representation using
f (z) = 1 and g(z) = 0, g(z) = z, g(z) = 1/z and g(z) = i/z, respectively.
2. Show that the parametrization (7.26) is regular provided the poles of g
coincide with (and are at most half the order of) the zeros of f .
3. Conclude that the surface obtained using g(z) = tan(z/2) and f (z) =
cos2 (z/2) is regular, and identify it with one of the surfaces in question #1.
(Thus, the Weierstrass representation of a minimal surface is not unique.)
Proof. To see that the differential σ is well-defined, let v be the vector field
tangent to the fibers of Fon that is dual to η21 . Then it is easy to check
that Lv σ = 0. To see that it is holomorphic, let σ = P (dz)2 , where z is a
local complex coordinate such that dz is a multiple of η 1 + iη 2 . Then (7.28)
implies that ∂P/∂ z̄ = 0.
As explained in §6.4, any isometric immersion of M into R3 will induce
an immersion from Fon (M ) to FE3 , and the graph of this map will be a
three-dimensional integral manifold of the Pfaffian system on Fon (M ) × FE3
generated by
{ω 3 , ω 1 − η 1 , ω 2 − η 2 , ω21 − η21 }.
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
204 7. Applications to PDE
To this system we adjoin the form θ (defined by (7.27)) and its conjugate,
where Q is now given. Our hypothesis about Q implies that this enlarged
system is Frobenius. Thus, there exists a unique integral 3-fold through each
point. Since the left action of FE3 ∼
= ASO(3) on itself covers rigid motion
in E , any two immersions will differ only by rigid motion.
3
Corollary 7.4.26 (Bonnet). Any constant mean curvature (CMC) surface
has a circle’s worth of isometric deformations as a CMC surface.
Example 7.5.9.
ux − ūx = λ sin(u + ū),
(7.36) 1
uy + ūy = sin(u − ū).
λ
If u(x, y), ū(x, y) are smooth functions that satisfy this system for some
λ = 0, then differentiating and equating mixed partials implies that u and
ū must be solutions of sine-Gordon. Conversely, given a solution u(x, y) of
(7.20), we can determine ūx and ūy from (7.36), and integrate to get the new
solution ū(x, y) (since we are guaranteed that mixed partials commute). For
example, the traveling wave solution (7.23) can be produced in this way, by
starting with the identically zero solution of (7.20).
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
7.5. Integrable extensions and Bäcklund transformations 209
p − p̄ = λ sin(u + ū),
1
q + q̄ = sin(u − ū).
λ
Let θ be the standard contact form on J 1 (R2 , R). On Bλ , let J = {π1∗ θ, π2∗ θ}.
Then it is easy to check that (7.37) is satisfied.
x̄ = x,
ȳ = y,
ū = 2(μ − y 2 ) − u,
p̄ = 4y(μ − u − y 2 ) − p,
r̄ = −4(u + y 2 − μ)(u + 3y 2 − μ) − 4yp − r.
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
210 7. Applications to PDE
family is tangent. Leaving degeneracies aside, the lines locally give a 1-to-1
correspondence between points on the focal surfaces.
Theorem (Bäcklund). If the distance λ between corresponding points on
the focal surfaces and the angle ψ between the surface normal at corre-
sponding points are both constant, then the two surfaces have constant
Gauss curvature equal to − sin2 ψ/λ2 .
(We will take λ = sin ψ, so that K = −1.)
The starting point for a proof of Bäcklund’s Theorem (cf. [48]) is adapt-
ing frames (X, e1 , e2 , e3 ) and (X̄, ē1 , ē2 , ē3 ) along the two surfaces so that
e1 = ē1 is tangent to the line connecting corresponding points X and X̄.
For any fixed λ, the graph of the Bäcklund transformation is a 6-dimensional
submanifold of F × F̄ defined by
X̄ = X + λe1 ,
ē1 = e1 ,
(7.38)
ē2 = e2 cos ψ + e3 sin ψ,
ē3 = e3 cos ψ − e2 sin ψ.
However, we will regard (7.38) as defining a 7-dimensional submanifold B ⊂
F × F̄ × R, with λ as the extra coordinate. On B, let J = {ω 3 , ω̄ 3 , dλ}.
Then (7.37) is satisfied, and Bäcklund’s Theorem follows from the fact that
{ω13 ∧ ω23 + ω 1 ∧ ω 2 , ω̄13 ∧ ω̄23 + ω̄ 1 ∧ ω̄ 2 } ≡ {dω 3 , dω̄ 3 } mod J.
Exercises 7.5.13:
1. Differentiate (7.38) to obtain the following relationships between the
pullbacks to B of the canonical forms:
ω̄ 1 = ω 1 + dλ, ω̄32 = ω32 − dψ,
2 2
ω̄ cos ψ sin ψ ω − λω21
= ,
ω̄ 3 − sin ψ cos ψ ω 3 − λω31
1 1
ω̄2 cos ψ sin ψ ω2
= .
ω̄31 − sin ψ cos ψ ω31
Then, use these relationships to verify (7.37).
2. Show that asymptotic lines are taken to asymptotic lines under the
Bäcklund transformation.
3. Show that if two pseudospherical surfaces M1 , M2 are related by a Bäck-
lund transformation and both are generated by solutions ui (x, y) of the sine-
Gordon equation, then u1 and u2 are related by a Bäcklund transformation
for the same λ-value.
4. ([192]) Consider the system defined by
u √ u
(7.39) ux = + 2wx , uy = + 2wy .
Licensed to AMS.
x+y x+y
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
212 7. Applications to PDE
Show that this defines a Bäcklund transformation between the wave equation
uxy = 0 and
√ √
wx wy
(7.40) wxy + 2 = 0.
x+y
5. Explain why the Cole-Hopf transformation (Example 7.5.2) is not a
Bäcklund transformation.
6. Show that existence of a Bäcklund transformation between two systems
is an equivalence relation on exterior differential systems.
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
https://doi.org/10.1090//gsm/175/08
Chapter 8
215
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
216 8. Cartan-Kähler III: The General Case
the study of linear Weingarten surfaces (see §7.4); a more complex example
is Bryant’s proof of the existence of Riemannian manifolds with holonomy
G2 [23]. See §9.8 for a discussion.
We begin this chapter with a detailed study of the space of integral
elements of an EDS. In §8.2 we give an example that shows how one can
use the Cauchy-Kowalevski Theorem to construct triply orthogonal systems
of surfaces, and this serves as a model for the proof of the full Cartan-
Kähler Theorem in §8.3. In §8.4 we discuss Cartan’s Test, a procedure by
which one can test for involution, and which has already been described in
Chapter 6 for the special case of linear Pfaffian systems. Then in §8.5 we
give a few more examples that illustrate how one applies Cartan’s Test in
the non-Pfaffian case.
Ann(ϕ)p := {v ∈ Tp Σ | v ϕ = 0}
ẽ (ω 1 ∧ π1 + ω 2 ∧ π2 ) = π2 ,
ẽ (ω 1 ∧ π2 ) = 0
Certainly, Vn (I) must lie inside the common zero locus of the
codimE Vn−1 + codim H(E) functions Gν and H α near E + . We will show
that they have linearly independent differentials at E + .
Note that F ν is a polynomial in the qja and uj , with coefficients de-
pending on xi and y a , and Gν is a polynomial in the pai . In fact Gν (pai ) =
F ν (paj , 0), i.e., Gν is obtained from F ν by setting qja = paj and uj = 0.
(In particular, Gν does not involve the pan .) Furthermore, since any other
codimension one subspace of E + is also in Vn−1 (I), dF ν |E (∂/∂uj ) = 0. It
follows that the Gν have codimE Vn−1 (I) linearly independent differentials
at E + .
Let Ψα = Φα − dy α ∧ dx1 ∧ · · · ∧ dxn−1 . Then (8.4) shows that Ψα is
a sum of terms that either vanish at p, or are wedge products of degree at
least two in (dxn , dy 1 , . . . , dy s ). Thus, Ψα (X1 , . . . , Xn ) will be a polynomial
in the pai consisting of terms that either vanish at p, or are of degree two in
the pai , or, when they are of degree one, do not involve the pan . It follows
that
dH α |E + ≡ dpαn mod {dpaj , j < n}.
Thus, Gν and H α have linearly independent differentials at E + .
We would like to find conditions under which the ability to solve the
infinitesimal Cauchy problem (i.e., extending E to E + ) implies the ability
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
222 8. Cartan-Kähler III: The General Case
If φν ∧ θ|E+ = 0, then
for i = 1, 2, 3. Since
dω 1 ∧ ω 1 = −ω21 ∧ ω 2 ∧ ω 1 − ω31 ∧ ω 3 ∧ ω 1 ,
(8.6) dω 2 ∧ ω 2 = −ω32 ∧ ω 3 ∧ ω 2 − ω12 ∧ ω 1 ∧ ω 2 ,
dω 3 ∧ ω 3 = −ω13 ∧ ω 1 ∧ ω 3 − ω23 ∧ ω 2 ∧ ω 3 ,
∂ ∂F α ∂ ∂ ∂F α ∂
+ and + will span Tπ(p) N . (Here and in what
∂x0 ∂x0 ∂y α ∂xi ∂xi ∂y α
follows, Greek indices run from 1 to 3 and Roman indices from 1 to 2.)
We will now show that the F α satisfy a system of PDE of Cauchy-
Kowalevski type. For variables q β and pβj , let
α ∂ β ∂ ∂ β ∂ ∂ β ∂
Ω +q , + p1 β , 2 + p2 β = Aαβ q β + B α ,
∂x0 ∂y β ∂x1 ∂y ∂x ∂y
where Aαβ and B α are polynomial in the pβj with analytic functions on U as
coefficients. Since at p, the equations
α ∂ β ∂ ∂ ∂
Ω +q , , =0
∂x0 ∂y β ∂x1 ∂x2
coincide with the polar equations of E, they are of rank 3 in the q β . We
may assume that Aαβ is invertible on some (possibly smaller) U ⊂ F and for
%pβj % sufficiently small. Then
−1 γ α ∂ β ∂ ∂ β ∂ ∂ β ∂
(A )α Ω +q , + p1 β , 2 + p2 β = q γ − C γ ,
∂x0 ∂y β ∂x1 ∂y ∂x ∂y
where C γ is an analytic function of the x’s, y’s and the pβj . So, we must
have
∂F α α 0 1 2 1 2 3 ∂F
β
=C x ,x ,x ,F ,F ,F ;
∂x0 ∂xj
with F α (0, x1 , x2 ) = 0. Since this is now a system of PDE in Cauchy-
Kowalevski form, we conclude that a unique solution exists, on a possibly
smaller U .
Proof. As in §8.2, we will reduce our problem to solving a system using the
Cauchy-Kowalevski Theorem.
Choose analytic coordinates x0 , x1 , . . . , xn , y 1 , . . . , y s , centered at p, such
∂ ∂ ∂
that 1
, . . . , n span Tp P and 0
∈ H(Tp P ). As in Lemma 8.1.13, let
∂x ∂x ∂x
{φ } be a basis for I near p, and let the (n + 1)-forms Φα , 1 ≤ α ≤ s,
ν n
As in §8.2, the Aαβ are the entries of a matrix which is invertible when
evaluated at the origin; so, on a possibly smaller open set (which we continue
to denote by U ) and for sufficiently small %pβj %, we can define forms Φ α =
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
8.3. Statement and proof of Cartan-Kähler 227
Proof. Let ck = codim H(Ek ) for 0 ≤ k < n; note that the ck are nonde-
creasing. Let x1 , . . . , xn , y 1 , . . . , y s be local coordinates centered at p, defined
∂ ∂
on an open set U ⊂ Σ, and chosen so that Ek is spanned by ,..., k
∂x1 ∂x
for k ≤ n and H(Ek ) is annihilated by dy 1 , . . . , dy ck for k < n.
Let R1 ⊂ U be the submanifold given by setting x2 = · · · = xn = 0
and y a = f1a (x1 ) for a > c0 , where f1a are some analytic functions such
that f1a (0) = 0. By Theorem 8.3.2 with R = R1 , there exists a unique
1-dimensional integral manifold N1 ⊂ R1 tangent to E1 at p.
Next, let R2 ⊂ U be the manifold given by setting x3 = · · · = xn = 0
and y a = f2a (x1 , x2 ) for a > c1 . To arrange that R2 contains N1 , we require
that f2a (x1 , 0) = f1a (x1 ). By Theorem 8.3.2 with R = R2 , there exists a
unique 2-dimensional integral manifold N2 ⊂ R2 tangent to E2 at p. (Note,
however, that we may have to shrink the open set U to ensure that H(Tq N1 )
is transverse to Tq R2 at every point q ∈ N1 .)
Proceeding in this way, we eventually obtain an n-dimensional integral
manifold tangent to En at p. (At the last step, Rn is defined by y a =
fna (x1 , . . . , xn ) for a > cn−1 .)
dimension at most s − cn−1 . (Here we again set s = dim Σ − n.) Since the
dimension of Gn (T Σ) is n + s + ns, Vn (I) has dimension
d = n + s + ns − (c0 + · · · + cn−1 ).
So, we expect the image of this map to have dimension at least
d + n − 1 − (s − cn−1 − 1) = n + s + (n − 1)(s + 1) − (c0 + · · · + cn−2 ),
which is exactly the largest dimension it can be. We will now make this
argument rigorous.
Let x1 , . . . , xn , y 1 , . . . , y s be local coordinates such that Ek is spanned by
∂/∂x1 , . . . , ∂/∂xk and H(Ek ) is annihilated by dy 1 , . . . , dy ck . As in the proof
of Lemma 8.1.10, we define functions pai completing a coordinate system on
U ⊂ Gn (T Σ) centered at En , and also define Xi , qja , uj and Zj as before.
Let F : U × Rn−1 → Gn−1 (T Σ) take (En , (v1 , . . . , vn−1 )) to the subspace
n spanned by the vectors Xj + vj Xn , 1 ≤ j ≤ n − 1. In coordinates,
of E
this map is given by uj = vj and qja = paj + vj pan . Differentiating these
equations shows that the kernel of F∗ at (En , (0, . . . , 0)) is spanned by the
vectors ∂/∂pan .
Let f denote the restriction of F to (Vn (I) ∩ U ) × Rn−1 . We want to
show that the rank of f is at least d − (s − cn−1 ) at (En , (0, . . . , 0)), or
equivalently that the intersection of the kernel of F∗ with TEn Vn (I) has
dimension at most s − cn−1 . Once this is known, the image of f contains
a smooth submanifold through En−1 with maximum codimension, and it
follows that En−1 is Kähler-ordinary; in fact, the image of f fills out Vn−1 (I)
near En−1 .
We will need to look at the equations that cut out Vn (I) near En in
local coordinates. By the above inequality, codimEn−1 Vn−1 (I) ≥ c0 + · · · +
cn−2 . So, let φν , 1 ≤ ν ≤ c0 + · · · + cn−2 , and Φα , 1 ≤ α ≤ cn−1 , be
defined as in the proof of Lemma 8.1.10. We saw then that the n-forms
φν ∧ dxn and Φα give rise to functions Gν and H α on Gn (T Σ) with linearly
independent differentials at En . Since Vn (I) is smooth at En , and of exactly
this codimension, the equations Gν = 0 and H α = 0 define Vn (I) near En .
As before, Gν is a polynomial in the pai which does not involve the pan ,
and
(8.8) dH α |En ≡ dpαn
mod {dpai , i < n}, 1 ≤ α ≤ cn−1 .
s
Now, if some linear combination a=1 ta ∂/∂pan is tangent to Vn (I) at En ,
(8.8) shows that ta = 0 for 1 ≤ a ≤ cn−1 . It follows that the kernel of f∗ at
En is as small as desired, and the rank of f is as large as desired.
Now it is easy to show that En−1 is Kähler-regular. For, suppose a
nearby E n−1 ) > codim H(En−1 ). This E
n−1 ∈ Vn−1 (I) has codim H(E n−1
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
8.4. Cartan’s Test 231
n ∈ Vn (I) near En .
will be in the image of f , and so is contained in an E
Now (8.3) implies that the codimension of Vn (I) at E n is larger than it
really is, so we reach a contradiction.
To conclude the proof of the theorem, we apply the above argument
inductively to show that every Ek is Kähler-regular.
Recall from §3.1 that the components of the curvature tensor are deter-
mined by the curvature 2-forms:
(8.9) Θij = 12 Ri jkl ω k ∧ ω l .
Now, if we set Φij = Θik g kj , then from (8.9) and the above identities we
obtain
ω(pij) ∧ Φij = 2Rjj ω(p) − 4Rpj ω(j) ,
where Rij are the components of the Ricci tensor, and Rji = g ik Rkj . Finally,
this equation yields an expression for the scalar curvature R if we wedge with
ω p and sum on the index p:
ω(pij) ∧ Φij ∧ ω p = 2(n − 2)R ω 1 ∧ . . . ∧ ω n .
In particular, on a three-manifold this says Φ12 ∧ ω 3 + Φ23 ∧ ω 1 + Φ31 ∧
ω 2 = R ω 1 ∧ ω 2 ∧ ω 3 . We will use this identity to set up an EDS whose
3-dimensional integral manifolds correspond to hypersurfaces of constant
scalar curvature R0 in S 4 .
We will work on F = FSon4 O(5). Let (e0 , e1 , e2 , e3 , e4 ) represent a
frame with basepoint e0 . Recall that deα = eβ ωαβ , 0 ≤ α, β ≤ 4, ωβα = −ωαβ ,
and dωβα = −ωγα ∧ ωβγ on F . The forms ω i = ω0i furnish a basis of the semi-
basic forms on F , and ωji , 1 ≤ i, j ≤ 4, are the Levi-Civita connection forms
for the standard metric on S 4 .
Along a hypersurface in S 4 we can choose a framing so that e4 is always
normal to the hypersurface; this will give a section of F which is a three-
dimensional integral manifold of the 1-form ω 4 , and to which Ω = ω 1 ∧ω 2 ∧ω 3
restricts to be nonzero. (This will be our independence condition.) From
now on, let i, j, k be indices that run from 1 to 3. Since dω i ≡ −ωji ∧
ω j mod ω 4 , and the ωji are skew-symmetric, the connection forms φij for the
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
236 8. Cartan-Kähler III: The General Case
by
(8.11) (tr h)2 − tr(h2 ) = R0 − 3.
Differentiation shows that V3 (I, Ω) is a smooth submanifold at points where
the entries of h are not all zero. Its fiber dimension is five.
Let E be a smooth point of V3 (I, Ω). Let v1 , v2 , v3 be the basis of E dual
to the restrictions of ω 1 , ω 2 , ω 3 to E, let Ei ⊂ E be spanned by v1 , . . . , vi ,
and of course let E0 be the zero subspace. Recall that H(E0 ) is just the
kernel of the 1-forms in I; H(E1 ) is cut out by these 1-forms and by those
obtained by feeding the vector v1 into generator 2-forms of I (cf. Exercise
8.1.7.5). Since
v1 (ωi4 ∧ ω i ) ≡ −ω14 mod ω 1 , ω 2 , ω 3 ,
we get one additional 1-form, and s1 = codim H(E1 ) − codim H(E0 ) = 1.
Similarly, v2 (ωi4 ∧ ω i ) ≡ −ω24 mod ω i , so s2 ≥ 1. But
(e1 , e2 ) Υ ≡ (h11 + h22 )ω34 mod ω 1 , ω 2 , ω 3 , ω14 , ω24 ,
so if h11 + h22 = 0, we get s2 = 2 for this flag. Following Remark 8.4.5, we
ignore Cauchy characteristics and take s3 = 0. Then s1 + 2s2 + 3s3 = 5,
and by the Cartan-Kähler Theorem we can construct an integral manifold
through E.
We now address the case when h11 +h22 = 0. If we can obtain an integral
manifold that has second fundamental form matrix h with respect to a frame
f ∈ F , then by rotation we can obtain an integral manifold through any
other point in the Cauchy characteristic leaf that contains f , and the second
fundamental form matrix there will be ghg −1 , g ∈ SO(3). So, it is enough
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
8.5. More examples of Cartan’s Test 237
where [x, y, z] denotes the associator (see Exercises A.5.3), we see that
Face(φ) consists of 3-planes in R7 to which the octonionic multiplication
restricts to be associative.
We define an EDS I for associative submanifolds by taking the com-
ponents of the ImO-valued 3-form ψ as generators. (Since ψ is constant-
coefficient, all of these generators are closed.) However, we can avoid a
lengthy character calculation by using the fact that G2 acts transitively on
Face(φ), as follows:
Choose the basis e1 , . . . , e7 for R7 to coincide with the basis 1 , . . . , 7
for ImO in §A.5; then g2 has the form (6.30). Without loss of generality
we may assume that E = {e1 , e2 , e3 } ∈ V3 (I). Then the stabilizer of E
in G2 is six-dimensional. Thus, V3 (I) is smooth of codimension four in
G(3, R7 ). On the other hand, for any flag in E, c0 = c1 = 0 and c2 = 4 (two
independent vectors in E determine the third one by multiplication). Thus,
by Theorem 8.4.1 we have I involutive at E (hence involutive everywhere,
by homogeneity). Integral manifolds depend on s2 = 4 functions of two
variables. (Note that this system has the same character and Cartan integer
as the Pfaffian system for associative submanifolds discussed in §6.12.)
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
8.5. More examples of Cartan’s Test 239
Exercises 8.5.9:
1. Recall that, if x is a point in Euclidean space, lying on a surface S with
principal curvatures k1 , k2 , then x + (1/ki )e3 , i = 1, 2, is a point on one of
the two focal surfaces of S. Show that, for the surfaces in Example 8.5.1,
one of the focal surfaces degenerates to a curve. Show that the surfaces are,
in fact, canal surfaces (see [178]) of radius 1/k0 about this curve, and use
this to interpret the “two functions of one variable” in another way.
2. Give an example of a surface whose principal curvatures k1 , k2 are con-
stant along their respective lines of curvature. Set up and investigate an
EDS for these surfaces. (Hint: You should introduce k1 and k2 as new
variables, and define your EDS on F × R2 .)
3. Show that, if we don’t ignore Cauchy characteristics in applying Cartan’s
Test to Example 8.5.3, the system is still involutive but now s2 = 1. Explain
how the integral surfaces depend on one function of two variables. (Hint:
How does the frame change as we move along the Cauchy characteristics?)
4. Set up and investigate an EDS for hypersurfaces in S 4 that have con-
stant scalar curvature and are minimal—that is, the trace of the second
fundamental form is zero. Can this be done without introducing the hij as
new variables?
5. Set up and investigate a non-Pfaffian EDS for hypersurfaces in S 5 that
are Einstein manifolds [15].
6. Set up and investigate a non-Pfaffian EDS for coassociative submanifolds
of E7 .
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
https://doi.org/10.1090//gsm/175/09
Chapter 9
Geometric Structures
and Connections
9.1. G-structures
In this section we present two examples of G-structures and then give a
formal definition.
241
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
242 9. Geometric Structures and Connections
0 = d2 ω 1 = −dθ ∧ ω 1 − θ ∧ (θ ∧ ω 1 ) = −dθ ∧ ω 1 ,
0 = d2 ω 2 = −dθ ∧ ω 2 − θ ∧ (θ ∧ ω 2 ) = −dθ ∧ ω 2 .
Remark 9.1.13 (Hexagonality). Recall from Remark 2.5.6 that one can
interpret the Gauss curvature of a surface in E3 in terms of a limit of ratios
of areas. There is a similar geometric interpretation of the Blaschke-Chern
curvature, in terms of the following construction (see Figure 2):
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
246 9. Geometric Structures and Connections
L
3 L
2
L3 L3
L2
L2
c
L
b 1
d
x L
1
g a
e L1
f
Let V != Rn with its standard basis, and GL(V ) = GL(n, R). Let
T(V ) := ⊗k ⊗ V ∗⊗l . Let ψ ∈ T(V ) be a tensor (e.g. ψ ∈ S 2 V ∗ or
k,l V
∗
ψ ∈ Λ V ) and let G = Gψ ⊂ GL(V ) be the subgroup preserving ψ. (In
2
Recall from §2.6 that the coframe bundle FGL comes equipped with—
and FG inherits—a tautological V -valued 1-form ω defined by, for w ∈
Tu FGL (M ),
ωu (w) = u(π∗ (w)).
If one fixes a basis of V , ω furnishes a basis {ω i } for the semi-basic forms
on FG .
Exercise 9.2.1: Prove that FG and FG are equivalent if and only if there
We are less interested in Tθ than in the part of the torsion that is in-
dependent of our choice of θ. Recall from §6.5 that the torsion of a linear
Pfaffian system with tableau A ⊆ W ⊗V ∗ lies in
H 0,2 (A) = (W ⊗Λ2 V ∗ )/δ(A⊗V ∗ ),
where δ : (W ⊗V ∗ )⊗V ∗ → W ⊗Λ2 V ∗ is the skew symmetrization map. In
our case W = V and A = g.
Definition 9.2.7. Let FG be a G-structure on M , θ a connection form and
Tθ ∈ Γ(FG , V ⊗Λ2 V ∗ ) the torsion of θ. Let π : V ⊗Λ2 V ∗ → H 0,2 (g) denote
the quotient map and write [Tθ ] for π ◦ Tθ . Then [Tθ ] : FG → H 0,2 (g) is
called the torsion of the G-structure, or the intrinsic torsion of FG . A G-
structure whose intrinsic torsion is zero is called 1-flat, because it resembles
a flat G-structure to first order.
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
9.2. Connections on FG and differential invariants of G-structures 251
Exercises 9.2.8:
1. If g ⊆ so(V ), then H 0,2 (g) = (so(V )/g)⊗V ∗ .
2. Verify that [Tθ ] is independent of θ, i.e., if θ is another connection form,
then [Tθ ] = [Tθ ].
3. Show that if FG → M is any G-structure, then for any fixed x ∈ M ,
there exist local coordinates (xi ) such that, (dx1 , . . . , dxn )|x ∈ (FG )x , i.e.,
there exists a section ρ of FG such that ρx = (dx1 , . . . , dxn )|x . If FG is 1-flat,
then moreover there exists a connection form θ such that, ρ∗ (θ)x = 0. (We
cannot have dθx = 0 unless FG is 2-flat at x; see Remark 9.2.14 below.)
dω i = θji ∧ ω j .
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
252 9. Geometric Structures and Connections
In general
dω i = Aijk ω j ∧ ω k + Bjk
i
ω j ∧ ω k + Cjk
i
ωj ∧ ωk
for some functions Aijk , Bjk i , C i . Taking θ i = Ai ω j + B i ω j eliminates
jk k jk jk
as much torsion as permissible. Thus the nonabsorbable torsion lives in
V (1,0) ⊗Λ2 V ∗(0,1) , so that H 0,2 (g) = V (1,0) ⊗Λ2 V ∗(0,1) .
We will see in Exercise 9.5.7 that if g(1) = 0, then Θ is basic and descends
to give a well-defined vector bundle valued 2-form on M .
If g(1) = 0 and θ is a torsion-free connection, then we may still define
the semi-basic form Θθ ∈ Ω2 (FG , g), but now it is just the curvature of θ,
rather than being intrinsic to the G-structure. We obtain the curvature of
the G-structure below in a manner similar to our passage from apparent to
intrinsic torsion.
Exercise 9.2.12: Show that the curvature of a connection Θθ,u : Λ2 Tu F →
gl(V ) on F = FGL (M ) → M satisfies the following equivariance: for g ∈
GL(V )
(9.4) Ad(g −1 )Θu (v, w) = ΘRg u (Rg∗ v, Rg∗ w),
where Rg : Fx → Fx is the right action on the fiber and Rg∗ : Tu F → TRg u F
is the induced action on the tangent spaces. Note that there is a g −1 in the
adjoint action so that the GL(V )-action on both sides of the equality is a
right action.
Exercise 9.2.13 (The first Bianchi identity): Let δ2 : (V ⊗V ∗ )⊗Λ2 V ∗ →
V ⊗Λ3 V ∗ denote the skew-symmetrization map. Show that
Θθ,u ∈ ker δ2 |g⊗Λ2 V ∗ = (g⊗Λ2 V ∗ ) ∩ (V ⊗S 2 1 V ∗ ),
3
Here H 1,2 (g) is the Spencer cohomology group (see Definition 6.7.3) where
we take A = g and W = V . As defined, we are viewing [Θθ ] as a function
on FG taking values in a fixed vector space, but we could equally view it as
a vector bundle valued 2-form on FG . Here (Θθ )u ∈ ker δ2 and the brackets
denote its equivalence class modulo the image of δ1 .
Remark 9.2.14. If s is a flat G-structure on M , then there exists a con-
nection θ with Tθ = 0 and Θθ = 0. A G-structure such that there exists a
connection θ such that Tθ = 0 and Θθ = 0 is called 2-flat, because at each
point the structure appears flat to order two.
1. Show that WH• is an open dense subset of H 0,2 (g)H andWH• − > H 0,2 (g)H• .
2. Show that FH is smooth.
Example 9.4.3. Let G ⊂ GLn+1 (R) be a matrix Lie group, and let M n =
G/H be a homogeneous space, with quotient map π : G → M . Assume
that if we write g ∈ G as g = (x, e1 , . . . , en ), where x and e1 , . . . , en are
the columns of the matrix representing g, the projection is π(g) = x and
we may identify e1 , . . . , en with a basis of Tx M . (This was the situation in
many of our examples in Chapters 1–3.) Assume that there exists a splitting
g = h ⊕ m such that m is H-invariant. (This will be the case, e.g., if H is
reductive.) Then we may write dej = ei ⊗ωji , where the ωji are components
of the Maurer-Cartan form of H.
Let s : M → G be a section and for X = xi s∗ (ei ) ∈ Γ(T M ) define
Note that here the Christoffel symbols Γijk for the connection are defined
by s∗ (ωji ) = Γijk s∗ (ω k ). To relate this example to our first attempt, the
correction term we are adding on to the naı̈ve derivative dxi ⊗ ei is the
pullback of the Maurer-Cartan form of h. Moreover, the choice of different
sections is analogous to different local trivializations, and we see that we do
indeed get the desired cancellation.
P = B is a Borel subgroup. (In the case of SL(n, C), the subgroup B con-
sists of the upper triangular matrices with determinant equal to one and
SL(n, C)/B is the space of complete flags in Cn . In general G/B is a space
of complete flags; see §4.2.) Then, taking one-dimensional representations
of B, one obtains holomorphic line bundles L → G/B.
Let Γ(L) denote the vector space of holomorphic sections of L. It is finite
dimensional because G/B is a compact complex manifold. Moreover, Γ(L)
is naturally a G-module. The Bott-Borel-Weil theorem gives an explicit one-
to-one correspondence between the irreducible finite dimensional G-modules
and the vector spaces Γ(L).
The correction term should come as no surprise, since in our example that
worked, Example 9.4.3, we had ∇X = ei ⊗(dxi + xj s∗ (θji )).
Proposition 9.5.6. Let FG → M be a G-structure with connection form
(θji ), and ∇ the induced connection on E = T M . Let s : M → FG be any
section, let ei be the corresponding frame field, and let X = xj ej ∈ Γ(T M ).
Then
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
260 9. Geometric Structures and Connections
∂ ∂
c∗ (∇)c = (xi ) dt⊗ + (xi ) (xk ) Γjik dt⊗ j .
∂xi ∂x
So
(9.7) c∗ (∇)c c = (xi ) + (xj ) (xk ) Γijk ,
and we see that c∗ (∇)c c = 0 is a system of n second-order ODEs.
Now specialize to the case dim M = 2 with coordinates (x, y).
Exercise 9.5.8: Eliminate t from (9.7) to obtain a single ODE:
3 2
d2 y dy dy dy
= Γ 1
22 + (2Γ 1
12 − Γ 2
22 ) + (Γ111 − 2Γ212 ) − Γ211 .
dx2 dx dx dx
Proposition 9.5.9 (Cartan [38]). A projective structure on a surface de-
termines a second-order ordinary differential equation on the surface of the
form
d2 y dy dy dy
2
= A( )3 + B( )2 + C( ) + D
dx dx dx dx
for some functions A, B, C, D on the surface. The solutions of this ODE are
the geodesics of the projective structure.
In Proposition 9.8.26 we will see a converse to Proposition 9.5.9.
9.7. Holonomy
In this section we only discuss the local theory, so throughout this section
(and exercises) assume M is connected and simply connected.
Given a G-structure FG → M , is there some proper subgroup H ⊂ G
that induces the “same geometry” as FG ? (This question will be made
precise below.)
One idea to make this question precise would be to ask if the curva-
ture Θ ∈ Ω2 (M, H1,2 (g)) takes values in some Ω2 (M, H1,2 (h)) for some
Lie subalgebra h ⊂ g. We will soon see that this is the correct idea (see
the Ambrose-Singer Theorem 9.7.14 below), but our definition of the corre-
sponding subgroup H will be different to facilitate proofs.
Our definition will be provided by the holonomy of a connection. Recall
that in a vector space, we may identify the tangent spaces at all points and
thus move vectors around from one tangent space to another. We can’t do
this on a manifold, even one equipped with a connection, but part of this
can be salvaged with a connection: we can identify tangent spaces to points
using curves that connect those points.
Parallel transport. Let π : FG → M be a G-structure with torsion-free
connection θ. (Here and below, this includes the case of G = GL(V ), i.e.,
when FG is the full frame bundle.) Let α : [a, b] → M be a smooth (or even
piecewise smooth) curve, let x0 = α(a) and let u0 ∈ π −1 (x0 ).
Exercise 9.7.1: Show that there exists a unique curve α̃ : [a, b] → FG such
that α̃(a) = u0 , π ◦ α̃ = α and α̃∗ (θ) = 0.
We call α̃ the horizontal lift of α through u0 , and curves that are hori-
zontal lifts of some curve in the base will be referred to as horizontal curves.
Exercise 9.7.2: Show that if α̃(t) is a horizontal curve in FG , then β(t) =
Rg α̃(t) is also horizontal for any g ∈ G, and all horizontal lifts are of this
form.
For those of you wondering about the origin of the terminology “connec-
tion”, parallel transport allows us to “connect” tangent spaces at different
points of M as follows: Consider the linear map
τα : Tα(a) M → Tα(b) M
~
α(a) ~(b)
α
Tα(a)M Tα(b) M
α(a) α(b)
Define
Holθx = {τα | ∃ α : [a, b] → M, α(a) = α(b) = x} ⊆ GL(Tx M ).
Let u ∈ π −1 (x) ⊂ FG . Then Holθx Holθu under the isomorphism between
GL(Tx M ) and G induced by u : Tx M → V . Since we are working locally,
each of the groups Holθx is isomorphic, so we sometimes drop reference to x
in the notation and write Holθ .
We may understand Holθ as measuring the failure of parallel translation
along closed curves to take a vector to itself.
We now explore relations between holonomy and infinitesimal geometry:
Exercise 9.7.6: Let E = Eρ → M be an induced vector bundle and let
s : M → E be a section. Let X ∈ Tx M and let α be a curve in M with
α(0) = x and α (0) = X. Let τ t be parallel translation from x to α(t) along
α with respect to connection ∇ on E induced from the connection on FG .
Show that
1
∇X s = lim [(τ t )−1 (s(xt )) − s(x)].
t→0 t
Then
∇φ =ω j1 ⊗. . .⊗ω jl ⊗ei1 ⊗. . .⊗eik
i ,...,i ,m,is+1 ,...,ik ∗ ∗ m
⊗[dpij11,...,i k 1 s−1
,...,jl + pj1 ,...,jl
is
s (θm ) − pij11,...,i k
,...,js−1 ,m,js+1 ,...,jl s (θjs )].
Now let α : [a, b] → M be a curve with α(a) = α(b) = x and α̃ its horizontal
lift. By definition, α̃∗ (θ) = 0, so if φ is G-parallel, then dpij11,...,i k
,...,jl (α ) = 0, i.e.,
the functions pij11,...,i k ∗
,...,jl are constant on lifts of horizontal curves; so τ (φ) = φ,
and thus Holθx ⊆ Gφ,x .
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
9.7. Holonomy 265
Proof. First, note that W ⊆ hol because θ pulled back to Pu0 is hol-valued
(the fibers of Pu0 are Holθu0 ), and thus Θ = dθ + θ ∧ θ is as well. Next, we
show that W is an ideal in g. In general, U ⊆ g is an ideal (i.e., [U, g] ⊆ U )
if and only if AdG (U ) ⊆ U . Let w = Θ(v1 , v2 ). By (9.4)
Adg (w) = Θ(Rg−1 ∗ v1 , Rg−1 ∗ v2 ) ∈ W,
so W is indeed an ideal.
Now fix a splitting hol = W ⊕ W c and write θ = θW + θc , where θW is
W -valued and θc is W c -valued. We need to show that W c = 0. Let I be the
Pfaffian system on FG spanned by the components of θc . Then
Θ = (dθW + θW ∧ θW ) + (dθc + θc ∧ θc ) + (θW ∧ θc + θc ∧ θW ).
The left hand side, and the first and third terms in parentheses, all take
values in W (the third because W is an ideal), while dθc + θc ∧ θc is W c -
valued. Thus 0 = dθc +θc ∧θc ; in particular 0 ≡ dθc mod I, so I is Frobenius.
Let Q ⊆ Pu be the maximal integral manifold of I through u. Since any
horizontal curve in FG is an integral of I (as all of θ vanishes on it), we have
Q = Pu . Hence W c = 0.
We conclude:
Theorem 9.7.18. Let M be an n-dimensional manifold equipped with a
torsion-free connection θ and curvature form Θ = Θθ such that ∇Θ =
0. Let H = Holθ = exp(hol) be the associated holonomy group. Then
locally M is isomorphic, as a manifold with H-structure, to G/H where
G = exp(hol ⊕ V ).
The homogeneous space G/H has extra structure: Define a group auto-
morphism σ : G → G with σ 2 = e by exponentiating the map σ∗ : Te G →
Te G given by σ∗ = Idh ⊕(− IdV ). This descends to give an H-structure au-
tomorphism σ[e] : G/H → G/H that squares to be the identity and with [e]
an isolated fixed point. For all x ∈ G/H we may obtain such an automor-
phism σx : G/H → G/H, with x an isolated fixed point, by conjugating by
elements of g. In particular:
Theorem 9.7.19. Let (M, g) be a Riemannian manifold with curvature
tensor Θ. Then the following are equivalent:
(1) For all x ∈ M , there exists an open neighborhood U ⊂ M of x and
an isometry σx : U → U , such that σx2 = Id, with x an isolated
fixed point.
(2) ∇Θ = 0.
(3) Given x ∈ M , there exists a neighborhood U ⊂ M and an isometry
from U to an open subset U ⊂ G/H, where H = exp(hol), hol is
the holonomy algebra of M , the group G is constructed as above,
and the Riemannian metric on G/H is the one induced from H.
A manifold satisfying the conditions in Theorem 9.7.19 is called locally
symmetric, and it is called symmetric, or a symmetric space, if U = M .
sp(V ) S 2 V ∗ ,
δ(S 2 V ∗ ⊗V ∗ ) = S21 V ∗ ,
We will see below that every path geometry is locally equivalent to the
path geometry induced by the solutions to a second-order ODE.
Example 9.8.4. Let X have a Riemannian, or more generally a reversible
Finsler metric, and take the paths to be the geodesics.
Example 9.8.5. Let X have a projective structure and take the paths to
be the geodesics.
Example 9.8.6. Let X = R2 with the origin O omitted, and let the paths
through a point x be circles and straight lines through x and O.
Definition 9.8.7. We will call a set of paths in X flat if for any x ∈ X
there is an open set U ⊂ X, x ∈ U , and a local diffeomorphism from U to an
open set in R2 that carries the paths in U to straight line segments. In other
words, the flat case locally corresponds to solutions of the ODE y = 0.
Problem 9.8.8. Which Riemannian metrics on a surface induce a flat set
of paths?
It is clear that the Euclidean metric on R2 induces the flat set of paths,
but are there others?
Proposition 9.8.9. S 2 with its standard metric induces the flat set of paths.
π
?
X
gives another foliation of P by the fibers of π. Call the leaves of this foliation
L2 and let L2 ⊂ T P also denote the associated distribution.
Suppose that the quotient of P by the leaves of L1 is a smooth surface
Z. Then Z is the space of paths in X, and we have the following double
fibration:
P
S
/
w
S
P/L1 = Z X = P/L2
(If the path geometry results from a second-order ODE in the plane, we may
think of Z as the space of solutions to, or initial conditions for, the ODE.)
While a point in Z corresponds to a path in X, a point x ∈ X also gives
a curve in Z (the one-parameter family of paths passing through x). The
resulting set of paths in Z is called the dual path geometry.
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
274 9. Geometric Structures and Connections
Example 9.8.12. Let X = RP2 (or P2 for short) have the standard set
of straight-line paths. Then Z = P2∗ , the set of lines in P2 , and P is the
flag manifold F12 := {(x, H)|x ∈ H} ⊂ P2 × P2∗ , a homogeneous space of
SL3 R. The paths in P2∗ are the linear P1∗ ’s, so the picture is completely
symmetric. The map from SL(3, R) to pairs (x, H) is given in bases by
(e1 , e2 , e3 ) → ({e3 }, {e2 , e3 }).
(1)
(4) On B1 , we normalize torsion and reduce to a subbundle with
smaller structure group several times, until we obtain a subbundle
(1)
B2 ⊂ B1 where the structure group G2 ⊂ g1 (1) is such that the
connection is unique.
(5) The torsion of the connection on B2 gives two functionally inde-
pendent relative differential invariants H1 , H2 . From these we con-
struct tensors on M that are diffeomorphism invariants of the path
geometry.
On a first reading of this section, the reader may now wish to skip ahead
to page 281, where we interpret the invariants geometrically. In particular,
we show that if H1 = H2 = 0, the path geometry is locally diffeomorphic to
the flat case, Example 9.8.12. Note: the flat path geometry does not induce
a flat G-structure, since it turns out that B1 always has torsion.
⎛ −1 1 ⎞
a ω
Rg∗ ω = ⎝ b−1 ω 2 ⎠ ,
c−1 ω 3
2 = b T 2 . Let B ⊂ B be the subbundle where
and it follows that g · T13 ac 13 1
T13 ≡ 1. Its structure group is
2
⎧⎛ ⎞ ⎫
⎨ a e 0 ⎬
G1 = ⎝0 ac 0⎠ ac = 0 .
⎩ ⎭
0 f c
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
9.8. Extended example: path geometry 277
In our usual abuse of notation, we now let θ = (θji ) denote the pullback of
the connection form to B1 . Since θ must be g1 -valued when restricted to
the fiber of B1 , we have θ22 ≡ θ11 + θ33 mod{ω j }. New apparent torsion may
arise in trying to make θ g1 -valued. Suppose that
θ22 = θ11 + θ33 + a1 ω 1 + a2 ω 2 + a3 ω 3 .
Then if we let φ1 = θ11 + a1 ω 1 and φ2 = θ33 + a3 ω 3 , our new structure
equations are
⎛ 1⎞ ⎛ 1 ⎞ ⎛ 1⎞ ⎛ ⎞
ω φ μ1 0 ω 0
(9.11) d ⎝ω 2 ⎠ = − ⎝ 0 φ1 + φ2 0 ⎠ ∧ ⎝ω 2 ⎠ + ⎝ω 1 ∧ ω 3 ⎠ ,
ω3 0 μ2 φ2 ω3 0
where μ1 = θ21 and μ2 = θ23 . Thus the apparent torsion is absorbable.
Remark 9.8.18. Our change of notation for the connection forms has an
ulterior motive. We want to distinguish the torus t ⊂ g1 which is spanned
by vectors dual to φ1 , φ2 . Recall that for a semi-simple Lie algebra, its
action on a vector space is entirely determined up to isomorphism by the
action of a maximal torus. In particular, in the case of semi-simple g, to
test if a quantity is invariant it is sufficient to check if it is annihilated by
the maximal torus t ⊂ g (or equivalently, acted upon trivially by the torus
T ⊂ G). Recall from Appendix A that such quantities are said to have weight
zero. In the case of a G-structure with G an arbitrary matrix Lie group,
it is still easy to check if a form is invariant under the action of a maximal
torus, and thus gives a necessary condition for the form to be invariant. We
will still say quantities invariant under the action of a maximal torus have
weight zero.
Example (9.8.16, continued). For the ODE y = F (x, y, y ), we obtain a
section of the G1 -structure by taking
ω 1 = dx,
ω 2 = dy − pdx,
ω 3 = dp − F (x, y, p)dx.
In this case, a set of connection forms which satisfy (9.11) is
(9.12) φ1 = Fp dx, φ2 = −φ1 , μ1 = 0, μ2 = −Fy dx.
Proposition 9.8.19. The nonuniqueness in the choice of connection satis-
fying (9.11) is
1 1 1
φ̃ φ 0 s1 ω
= 1 + ,
μ̃ 1 μ s1 t1 ω2
(9.13) 2 2 2
μ̃ μ t2 s2 ω
= 2 +
φ̃2 φ s2 0 ω3
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
278 9. Geometric Structures and Connections
(1)
Prolongation. Let B1 be the prolongation of B1 , as described in §9.3.
(1)
The forms (ω j , φi , μi ) pulled back to B1 become a basis of the tautological
(1)
forms on this bundle. We would like to fix a connection on B1 and de-
termine the torsion. In order to deduce the specific form of the torsion, we
(1)
differentiate the semi-basic forms on B1 . We already know the derivatives
of the ω j , so we need the derivatives of the φi and μi , which we calculate by
computing 0 = d2 ω j :
⎛ 1⎞ ⎛ 1 ⎞ ⎛ 1⎞
ω dφ dμ1 0 ω
2 ⎝ 2⎠ ⎝ ⎠
0=d ω =− 1
0 dφ + dφ 2 0 ∧ ω2⎠
⎝
ω 3 0 dμ 2 dφ 2 ω3
⎛ 1 ⎞ ⎧ ⎛ ⎞ ⎛ 1⎞ ⎛ ⎞⎫
φ μ1 0 ⎨ φ1 μ1 0 ω 0 ⎬
+ ⎝ 0 φ1 + φ2 0 ⎠ ∧ − ⎝ 0 φ1 + φ2 0 ⎠ ∧ ⎝ω 2 ⎠ + ⎝ω 1 ∧ ω 3 ⎠
⎩ ⎭
0 μ2 φ2 0 μ2 φ2 ω3 0
⎛ ⎞
0
+ ⎝−(φ1 ∧ ω 1 + μ1 ∧ ω 2 ) ∧ ω 3 + ω 1 ∧ (μ2 ∧ ω 2 + φ2 ∧ ω 3 )⎠ .
0
Separating the rows, we have
(9.14) 0 = −d2 ω 1 = (dφ1 + μ1 ∧ ω 3 ) ∧ ω 1 + (dμ1 + μ1 ∧ φ2 ) ∧ ω 2 ,
(9.15) 0 = −d2 ω 2 = −(dφ1 + dφ2 ) ∧ ω 2 − μ1 ∧ ω 2 ∧ ω 3 − μ2 ∧ ω 1 ∧ ω 2 ,
(9.16) 0 = −d2 ω 3 = (dμ2 + μ2 ∧ φ1 ) ∧ ω 2 + (dφ2 − μ1 ∧ ω 1 ) ∧ ω 3 .
where
rI1
rμ 2 rσ rH1
rω 1 uφ1,φ2 rμ 1
rω 2 rω 3
In this diagram, weights are assigned according to how objects scale with
respect to directions in the maximal torus dual to φ1 and φ2 respectively;
the large dot represents the origin.
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
282 9. Geometric Structures and Connections
Proof. For the first part, notice that for the ODE y = 0, we obtain H1 =
H2 = 0 from the formulas (9.21). For the second part, let
⎛ 1 ⎞
− 3 (2φ1 + φ2 ) −μ2 σ
ψ=⎝ ω1 3 (φ − φ )
1 1 2 μ1 ⎠.
2 3 1 1 2
ω ω 3 (φ + 2φ )
Note that ψ takes values in the Lie algebra sl(3, R). In general,
⎛ ⎞
0 −H1 ω 1 ∧ ω 2 (I1 ω 1 + I2 ω 3 ) ∧ ω 2
dψ + ψ ∧ ψ = ⎝0 0 H2 ω 3 ∧ ω 2 ⎠.
0 0 0
If H1 and H2 are identically zero, then I1 and I2 are also zero by (9.24).
Since dψ + ψ ∧ ψ = 0, by Theorem 1.6.13 there exists, in a neighborhood
U of any point on B2 , a local diffeomorphism g : U → SL(3, R) such that
g −1 dg = ψ. Moreover, the fibers of submersion to M , which are annihilated
by the ω’s, are carried to right cosets of the subgroup P of upper triangular
matrices in SL(3, R). This means that M is locally diffeomorphic to the flag
variety F12 from Example 9.8.12.
Proof. It is easy to check that the coframe in Exercise 9.8.20 gives a section
of B2 . Differentiating the structure equations of Fon (X) gives dK ≡ 0
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
9.8. Extended example: path geometry 283
modulo ω 1 , ω 2 . Let
(9.25) dK = K1 ω 1 + K2 ω 2 .
Using σ = −Kω 2 , we compute that H1 = K2 and H2 = 0. Furthermore,
differentiating (9.25) shows that dK2 ≡ −K1 ω12 modulo ω 1 , ω 2 . Hence, H1
is identically zero if and only if K is constant.
Remark 9.8.25. In fact, the same result is true for path geometries of
arbitrary dimensions: the only Riemannian metrics inducing the flat path
geometries are those of constant sectional curvature.
Proposition 9.8.26. A path geometry is locally equivalent to a projective
structure if and only if H2 ≡ 0.
Proof. Since every path geometry is locally equivalent to one arising from
a second-order ODE, we can use the formulas (9.21) in that case to give
interpretations to each of H1 and H2 . For example, H2 = 0 implies that our
system of paths is locally equivalent to the integral curves of an ODE of the
form
y = A(y )3 + B(y )2 + C(y ) + D,
where A, B, C, D are functions of x and y. Now apply Proposition 9.5.9.
Remark 9.8.27. If one really wanted to find the explicit integral curves of
an ODE giving rise to a flat set of paths, it would be rather difficult to solve,
because although one knows abstractly that one has the flat case, actually
finding the transformation of integral curves can be quite difficult.
If H1 ≡ 0 and H2 = 0, then up to a few exceptions the equation
y = F (x, y, p) is solvable by quadrature. (In other words, the differential
equation has a solvable symmetry group.) The idea is that differentiating
H2 and I2 yields scalar invariants that are constant along the paths, yielding
first integrals for the ODE. We explain this in more detail, following [31]:
Assume that H1 ≡ 0. Then I1 ≡ 0 and J ≡ 0 from (9.24), and we have
1
I2 2φ + 3φ2 μ2 I2
(9.26) d ≡ mod ω 2 , ω 3 .
H2 −ω 1 φ1 + 3φ2 H2
This implies that P = (I2 ω 2 + H2 ω 3 ) ⊗ (ω 2 ∧ ω 3 ) is a well-defined tensor on
the path space Z = M/L1 . (Note that the weight diagram indicates that P
is invariant under the maximal torus.) To see this, suppose we set P2 = I2 ,
P3 = H2 . Then we may write (9.26) as
(9.27) dPa = (ψab + (tr Ψ)δab ) Pb + Pab ω b ,
for some functions Pab , where we use index ranges 2 ≤ a, b ≤ 3 and we have
set 1
φ + φ2 μ2
Ψ = (ψab ) = .
−ω 1 φ2
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
284 9. Geometric Structures and Connections
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
https://doi.org/10.1090//gsm/175/10
Chapter 10
Superposition for
Darboux-Integrable
Systems
285
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
286 10. Superposition for Darboux-Integrable Systems
10.1. Decomposability
Throughout this chapter I will be an EDS generated algebraically by 1-
forms and 2-forms on manifold Σ; we let I ⊂ T ∗ Σ be the bundle spanned
by the 1-forms.
Recall from §7.2 that I is hyperbolic of class k if it is generated by k
1-forms and a pair of decomposable 2-forms. In other words, at each point
of Σ there is a coframe (θ1 , . . . , θk , π11 , π12 , π21 , π22 ) such that
I = {θ1 , . . . , θk , π11 ∧ π12 , π21 ∧ π22 }alg .
In §7.2 we saw that hyperbolic systems with k = 3 occur when one makes
an EDS encoding a second-order hyperbolic PDE for one function of two
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
10.1. Decomposability 287
variables in the standard way (and systems with k = 1 can encode second-
order hyperbolic Monge-Ampère equations), while systems with k = 2 arise
when one encodes a quasilinear hyperbolic first-order system for a pair of
functions of two variables.
Recall that the Monge characteristic systems V1 , V2 of a hyperbolic sys-
tem are defined by
Vi = {θ1 , . . . , θk , πi1 , πi2 }.
These systems play a key role in the method of Darboux, since I is Darboux-
integrable if each Vi possesses a pair of characteristic invariants that are
independent of each other and of the 1-forms of I. (A characteristic invariant
is a first integral of Vi , i.e., a function f whose differential df is a section of
Vi ; it follows that it also must be a section of the terminal derived system
(∞)
Vi .) In seeking to extend integrability to a wider class of systems, we will
focus on the role played by V1 , V2 in the algebraic structure of the ideal.
Because I is closed under exterior differentiation,
dθj ≡ Aj π11 ∧ π12 + B j π21 ∧ π22 mod θ1 , . . . , θk
for some functions Aj , B j . A natural extension of this structure, which turns
out to encompass a variety of other kinds of systems, is based on assuming
a direct sum decomposition
T ∗ Σ = I ⊕ J1 ⊕ J2
such that I is generated algebraically by Γ(I) and 2-forms that either lie in
Γ(Λ2 J1 ) or Γ(Λ2 J2 ). This condition is unchanged if J1 or J2 is modified by
adding 1-forms belonging to I, so it makes sense to formulate it in terms of
I + J1 and I + J2 .
Definition 10.1.1. Let I be an EDS generated algebraically by 1-forms and
2-forms on Σ. Then I is decomposable if there exist subbundles V̂ , V̌ ⊂ T ∗ Σ
such that
• T ∗ Σ = V̂ + V̌ and I = V̂ ∩ V̌ ;
• I is algebraically generated by 1-forms in I and 2-forms that either
are sections of Λ2 V̂ or in Λ2 V̌ .
Some examples.
Example 10.1.2. Consider a system of f -Gordon equations for s functions
of variables x1 , x2 :
(10.1) ua12 = F a (x, u, u1 , u2 ), a = 1, . . . , s.
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
288 10. Superposition for Darboux-Integrable Systems
(On the right we group the independent and dependent variables into vectors
x and u respectively, and ui denotes the partial derivative of u with respect
to xi .) This system is encoded by an EDS generated by 1-forms θa =
dua − ua1 dx1 − ua2 dx2 and 2-forms π1a ∧ dx1 , π2a ∧ dx2 with
Here, π i indicates the s-vector of 1-forms π1a , and θ indicates the vector of
forms θa .
where Cba is the Cartan matrix (see, e.g., [19]) of a simple Lie algebra. (For
example, the A2 Toda lattice system appears in Exercise 10.5.11.6 below.)
These systems are Darboux-integrable after sufficiently many prolongations;
see [154] for a systematic way of writing down the characteristic invariants.
If the matrix in (10.2) is replaced by certain other choices (called gener-
alized Cartan matrices), the resulting system is completely integrable, and
solutions can be used to construct harmonic maps into symmetric spaces
[87].
Remark 10.1.4. Systems of the form (10.1) are related to the defining
equations for wave maps from the Minkowski plane R1,1 into a Riemannian
manifold M , where x, y are null coordinates in the domain and the ua are
local coordinates in M . (Wave maps are exactly analogous to harmonic
maps, but with the domain being a semi-Riemannian manifold.) Two-
dimensional Riemannian manifolds M for which the wave map system is
Darboux-integrable after at most one prolongation have been classified [52];
one example appears in Exercise 10.5.11.5 below.
10.2. Integrability
A key geometric feature of classical Darboux integrability is the double
foliation of an integral surface S by characteristic curves: since the first
integrals of each characteristic system are constant along the corresponding
curves, each set of invariants restricts to S to be functionally related, and
these relations can be used to recover S. For more general decomposable
systems, the tangent bundle of an integral submanifold S still splits as a sum
of characteristic distributions (annihilated by the pullbacks to S of V̌ and V̂ ,
respectively), but the integrability of these distributions is not automatic:
what’s necessary is that the pullbacks to S of V̂ and V̌ be Frobenius. A
condition that guarantees this, and is independent of S, is that V̂ (∞) and
V̌ (∞) are large enough that T ∗ Σ = I + V̂ (∞) + V̌ (∞) as a nondirect sum.
This is the gist of the next definition.
Definition 10.2.1. A decomposable system is Darboux-integrable if
(i) V̂ (∞) + V̌ = V̂ + V̌ (∞) = T ∗ Σ;
(ii) V̂ (∞) ∩ V̌ (∞) = 0.
Again, note that in the first condition the sum need not be direct. The
second condition is technical, ensuring that I contains no closed 1-forms.
This is not a serious restriction, for if I did contain some closed 1-forms,
these would locally be exact 1-forms dfi , and any integral submanifold would
necessarily lie in a joint level set of the fi ; but then the pullback of I to the
level set would contain no exact 1-forms.
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
290 10. Superposition for Darboux-Integrable Systems
σ̂ η̂ θ η̌ σ̌
Since the σ̂ a must span a complement to {η̂} within V̂ (∞) , they can be
chosen to be closed 1-forms, and similarly we can choose the σ̌ α to be closed.
But we can also simplify the structure equations satisfied by η̂, η̌:
Proposition 10.2.7 (Theorem 2.9 in [6]). Near any point, there exists a
0-adapted coframe satisfying
dσ̂ = 0, dη̂ = (Ê η̂ + F̂ σ̂) ∧ σ̂, dσ̌ = 0, dη̌ = (Ě η̌ + F̌ σ̌) ∧ σ̌.
(Following [6], here we begin to use boldface capitals to indicate the ap-
propriately indexed arrays of coefficients. For example, the last structure
equation reads in full
μ α
dη̌ μ = Ěνα
μ ν
η̌ ∧ σ̌ α + 12 F̌αβ σ̌ ∧ σ̌ β ,
Proof. By the Frobenius Theorem, the system V̂ (∞) locally has a basis of
exact differentials. We can group these differentials into two vectors dIˆ1 and
dIˆ2 (of length p and r respectively) such that the 1-forms
η̂ = dIˆ2 + R̂ dIˆ1
form a basis of V̂ (∞) ∩ I, for some r × p matrix R̂. We will take σ̂ = dIˆ1 ,
so that dσ̂ = 0. Because dη̂ involves no σ̌ ∧ σ̂ terms, dR̂ ∈ V̂ , and hence
dR̂ ∈ V̂ (∞) . So we have dR̂ = Ê η̂ + F̂ σ̂ and the desired form for dη̂ follows.
We obtain σ̌ and η̂ similarly.
Exercise 10.2.8: Show that the components of Ê and F̂ are expressible in
terms of the entries of R̂ and their partial derivatives with respect to the
functions in Iˆ1 and Iˆ2 . (This establishes the last assertion of the proposi-
tion.)
The last assertion of the proposition (which follows from the exercise)
is not part of the definition of 1-adapted given in [6], but it is an easy
consequence of the proof, and is necessary for subsequent adaptations.
Example (10.2.4 continued). To construct a 0-adapted coframe, we first
choose a basis for the intersection of V̂ (∞) with I = {θ0 , θ1 , θ2 }. This inter-
section has rank one, and is spanned by η̂ = (θ1 −wθ0 )/u = dw+(w2 −R)dx.
We complete a basis for V̂ (∞) with exact 1-forms σ̂ 1 = dx and σ̂ 2 = dw.
Similarly, we span V̌ (∞) using σ̌ 1 = dy and σ̌ 2 = dT . The intersection of
V̌ (∞) with I is zero, so there are no 1-forms η̌ in this example.
We complete the 0-adapted coframe with
θ1 = −θ0 = −du + p dx + q dy, θ2 = −θ2 = −dq + q(w dx + T dy).
Finally, we note that
dη̂ = σ̂ 1 ∧ σ̂ 2 + 2wη̂ ∧ σ̂ 1 ,
indicating that this coframe is also 1-adapted.
that
I = {θ, η̌, η̂, Ω̂, Ω̌}alg .
Moreover, with respect to a 1-adapted coframe, we can assume that the Ω̂α
and Ω̌β are expressible in terms of pure wedge products of the σ̂ a and σ̌ b
respectively. By adjoining one or the other of these groups of 1-forms to the
generators of I, we obtain the singular systems
(10.5) V̂ = {θ, σ̂, η̂, η̌, Ω̌}alg , V̌ = {θ, σ̌, η̌, η̂, Ω̂}alg .
Exercise 10.2.9: Show that these are differential ideals.
The next result will be needed in §10.5 to prove that the superposition
formula takes solutions to solutions. (In what follows, let 1 ≤ i, j, k ≤ s,
where s = dim Σ − rkV̂ (∞) − rkV̌ (∞) .)
Proposition 10.2.10. Let (θ, η̂, σ̂, η̌, σ̌) be a 1-adapted coframe, and let
(10.6) dθi ≡ 12 Aiab σ̂ a ∧ σ̂ b + 12 Bαβ
i
σ̌ α ∧ σ̌ β mod θ, η̂, η̌.
(Note that the decomposability condition implies that these are the only
kinds of wedge products that can occur in dθi mod I.) Then
(10.7) I = {θ, η̂, dη̂, η̌, dη̌, Aσ̂ ∧ σ̂, B σ̌ ∧ σ̌}alg ,
where Aσ̂ ∧ σ̂ and B σ̌ ∧ σ̌ denote vectors of 2-forms whose members are
the individual terms on the right-hand side of (10.6).
Proof. Let (∂θ , ∂σ̂ , ∂η̂ , ∂σ̌ , ∂σ̂ ) be the dual frame to a 1-adapted coframe.
Using iterated Lie brackets define the vector fields ŜA and ŠB (where A and
B are multi-indices whose entries have ranges 1 . . . p and 1 . . . q respectively)
by
Ŝa = ∂σ̂a , Šα = ∂σ̌α , ŜAa = [ŜA , Ŝa ], ŠBα = [ŠB , Šα ].
Note from the structure equations that [∂σ̂ , ∂σ̌ ] = 0, and that ŜA ∈ {∂θ , ∂η̂ },
ŜB ∈ {∂θ , ∂η̌ } when |A| > 1 and |B| > 1. Furthermore, it follows by the
Jacobi identity that
(10.8) [ŜA , ŠB ] = 0.
Since the ∂σ̂ span Ĥ = (V̌ )⊥ , then a finite number of the ŜA will span
Ĥ(∞) = (V̌ (∞) )⊥ . Among these we can choose vectors ŜAm , where 1 ≤ m ≤
rk(V̂ ∞ ∩ I), so that (∂θ , ∂σ̂ , ŜAm ) are a basis for Ĥ(∞) . Similarly, there are
vectors ŠBμ , where 1 ≤ μ ≤ rk(V̌ ∞ ∩ I), such that (∂θ , ∂σ̌ , ŠBμ ) is a basis
for Ȟ(∞) .
Then (∂θ , ∂σ̂ , ŜAm , ∂σ̌ , ŠBμ ) is a local frame, and our new coframe is its dual
(θ , σ̂, η̂ , σ̌, η̌ ), where θi ≡ θi mod η̂, η̌ and
η̂ = Ĥ η̂, η̌ = Ȟ η̌
for some nonsingular matrices Ĥ and Ȟ. Thus, the new coframe is 0-
adapted.
Exercise 10.3.2: Show that the components of Ĥ and Ȟ are first integrals
of V̂ (∞) and V̌ (∞) respectively. Conclude that the new coframe is 1-adapted.
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
10.3. Coframe adaptations 295
Moreover, from (10.8) it follows that the structure equations of the new
coframe contain no terms of the form σ̂ ∧ η̌ , σ̌ ∧ η̂ or η̂ ∧ η̌ .
Example (10.2.4 continued). The remaining structure equations for our
1-adapted coframe are
dθ1 = uη̂ ∧ σ̂ 1 − wθ1 ∧ σ̂ 1 − θ2 ∧ σ̌ 1 ,
dθ2 = q η̂ ∧ σ̂ 1 − wθ2 ∧ σ̂ 1 − T θ2 ∧ σ̌ 1 − qσ̌ 1 ∧ σ̌ 2 .
Since these contain no η̂ ∧ σ̌ terms, this coframe is also 2-adapted.
Proof. Given a 2-adapted coframe (θ, π̂, π̌), we replace this with a tempo-
rary coframe (θ, dJˆ, dJ),
ˇ where the exact 1-forms dJˆ and dJˇ form a basis
for the Frobenius systems V̂ (∞) and V̌ (∞) respectively. We let (∂θ , Ûa , Ǔα )
be the dual frame to the temporary coframe, where indices a, b range from
1 to rkV̂ (∞) while α, β range from 1 to rkV̌ (∞) . As in the previous proof
we form iterated Lie brackets ÛAa = [ÛA , Ûa ] and ǓBα = [ǓB , Ǔα ], where
A, B are multi-indices. Note that because no π̂ ∧ π̌ terms occur in the
structure equations of the 2-adapted frame, [Ûa , Ǔα ] = 0 and consequently
[ÛA , ǓB ] = 0.
Since
Ĥ = V̌ ⊥ = {θ, η̂, π̌}⊥ ⊂ {θ, dJ}
ˇ ⊥ = {Ûa },
then a finite number of the ÛA will span Ĥ(∞) . But because [Ûa , Ûb ] ∈
{∂θ } = D and [Ûa , ∂θ ] ∈ D, then the ÛA for |A| > 1 all lie in D =
Ĥ(∞) ∩ Ȟ(∞) . Therefore, there are a finite number of these ÛA which form
a basis for D; let ÛAk denote these, where k = 1, . . . , s and |Ak | > 1. Thus,
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
296 10. Superposition for Darboux-Integrable Systems
we started with, the coframe (θ X , π̂, π̌) is still 2-adapted. Moreover, because
[Ǔα , ÛAi ] = 0 the structure equations of the new coframe contain no π̌ ∧ θ
terms.
The 1-forms θYi are similarly derived from the ǓB for |B| > 1.
Let G denote the simply connected Lie group whose algebra is the Vessiot
algebra g, and let {vi } be a fixed basis for Te G such that [vjL , vkL ] = −Cjk
i vL .
i
By the existence theory for group actions (see §10.4 below), the vector fields
Xi and Yj generate local right and left G-actions respectively on Σ. In other
words, there is an open neighborhood V of {e} × Σ in G × Σ, and mappings
λ : V → Σ, ρ : V → Σ,
which are left and right actions respectively, such that ρ∗ : vi → Xi and
λ∗ : vj → Yj . The actions are free, and the orbits are open subsets of
maximal integral submanifolds of D. Moreover, because [Xi , Yj ] = 0 these
actions commute. However, for reasons we now explain, we will end up using
a different set of vector fields to generate the action of G in §10.5.
The superposition formula is constructed by defining simpler Pfaffian
systems on integral manifolds of V̂ (∞) and V̌ (∞) , and using the G-action
to define a surjective mapping from the product of these manifolds to a
G-invariant open set U ⊂ Σ, in a way that carries solutions to solutions.
The mapping is based on choosing a slice S for the group action—i.e., a
submanifold in Σ which intersects each orbit in U transversely in a single
point. Given a slice, we can associate to each point x ∈ U a unique point
x ∈ S in the same orbit, and a group element γ(x) that takes x to x. (Thus,
S = γ −1 ({e}).) This enables us to identify U with the product G×S, but we
would like the identification to be equivariant with respect to the G-action.
For this to happen, we need the foliation of U by level sets of γ to be
G-invariant; in other words, the level sets must be integrals of a G-invariant
Frobenius system. One might take the 1-forms θ X as spanning a Frobenius
system whose integrals are transverse to the G-orbits; however, LXj θX i does
(where the components of matrices R̂, Ŝ are first integrals of V̂ (∞) , the
components of Ř, Š are first integrals of V̌ (∞) , and R̂, Ř are nonsingular)
such that
dθ̂i = 12 Cjk
i j
θ̂ ∧ θ̂k + 12 Giαβ π̌ α ∧ π̌ β ,
(10.9)
dθ̌i = − 12 Cjk
i j
θ̌ ∧ θ̌k + 12 Hab
i a
π̂ ∧ π̂ b .
The coframes (θ̂, π̂, π̌) and (θ̌, π̂, π̌) are called 5-adapted.
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
298 10. Superposition for Darboux-Integrable Systems
i = Qi θ j for a nonsingular s × s
Corollary 10.3.6 (Lemma 5.8 in [6]). If θX j Y
matrix Q, and we define the matrix Λ = R̂Q(Ř)−1 , then the 1-forms
satisfy
The spans {ω̂ i } and {ω̌ i } coincide, and (by making constant-coefficient ad-
justments to R̂, Ř, Ŝ, Š) we can arrange that ω̂ i = ω̌ i at a given point x0 .
Exercise 10.3.7: Deduce the corollary from the proposition.
Let X̂i and X̌j denote the vector fields annihilated by the π̂ and π̌
forms, and dual to θ̂i and θ̌j respectively. Note that, unlike previous steps,
5-adapted coframes are not a special case of the previous adaptation; in fact,
the 1-forms θ̂i and θ̌j are not necessarily in I. Nevertheless, both sets of
vector fields span D at each point, and they still satisfy bracket relations
as well as
(modulo π̌) to a multiple of θY2 . So, it’s easier to solve for θ̂2 and θ̌2 first.
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
10.4. Some results on group actions 299
γ ∗ ωLi = ω̂ i and γ ∗ ωR
j
= ω̌ j .
Moreover, if we define Ψ : U → S × G by Ψ(x) = (x, γ(x)), then Ψ is
bi-equivariant, where G acts trivially on the factor S in the codomain.
Exercise 10.5.2: Show that γ(ρ(g, x)) = γ(x) · g and γ(λ(g, x)) = g · γ(x),
and use these to show the bi-equivariance of Ψ.
The flow generated by X̂1 is (u, q) → (u − tewx , q) and the flow generated
by X̂2 is (u, q) → (e−t u, e−t q), so that the right action is
ρ(a, b, u, q) = (e−a u − bewx , e−a q).
Exercise 10.5.3: Verify that the left and right actions commute.
The first integrals of the Frobenius system {ω̂ 1 , ω̂ 2 } are u/q and qe−xw .
Therefore, the slice S through x0 is defined by u = 0 and q = exw . Equation
(10.15), which implicitly defines the mapping γ of Proposition 10.5.1, now
reads
(u, q) = ρ(a, b, 0, exw )
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
302 10. Superposition for Darboux-Integrable Systems
where a, b are the coordinates of γ(x). Solving this gives ea = exw /q, b =
−ue−xw , and thus
1 ue−xw
(10.16) γ(x) = .
0 qe−xw
Exercise 10.5.4: Compute the pullbacks via γ of0 the left-invariant 1-forms
ωLi that are dual to the basis v1 = 00 −1
0 , v2 = 0
0 −1 for Te G, and confirm
that these pullbacks coincide with the ω̂ .i
Let S1 denote the maximal integral manifold of V̂ (∞) |S and S2 the max-
imal integral manifold of V̌ (∞) |S passing through x0 . By the Frobenius
Theorem, on S there are functions pa such that V̂ (∞) |S = {dpa }, and func-
tions q α such that V̌ (∞) |S = {dq α }, and because T ∗ S is the direct sum of
these systems, the pa and q α form a local coordinate system on S. We can
arrange that S1 is the zero locus of the pa and S2 is the zero locus of the q α ,
and by concatenating the coordinates of a point on S1 with the coordinates
of a point on S2 , we have a diffeomorphism
χ : S1 × S2 → S
whose image is an open set S0 ⊂ S.
Let M1 and M2 be the intersections with U of maximal integral manifolds
through x0 of V̂ (∞) and V̌ (∞) respectively. Since these systems are G-
invariant, the actions λ and ρ restrict to M1 and M2 . Furthermore, if x1 ∈
M1 and x2 ∈ M2 , then xi ∈ Si = Mi ∩ S for i = 1, 2. We define the
superposition map Γ : M1 × M2 → M by
(10.17) Γ(x1 , x2 ) = ρ (γ(x1 )γ(x2 ), χ(x1 , x2 )) .
Using the results of Exercise 10.5.2, it is easy to show that Γ is invariant
under the diagonal action of G on M1 × M2 , i.e.,
Γ(ρ(g −1 , x1 ), λ(g, x2 )) = Γ(x1 , x2 ) ∀g ∈ G.
Thus, Γ covers a well-defined mapping from the quotient (M1 × M2 )/G to
U , which turns out to be a local diffeomorphism. To see this, we need to
examine how Γ relates the 5-adapted coframe on U to coframes on M1 and
M2 .
A coframe on M1 is given by the restrictions of θ̂, σ̌ and η̌, while on M2
we use the restrictions of θ̌, σ̂ and π̂. We will use a subscript i = 1 or i = 2
to denote the restriction of a 1-form to Mi , and we pull these 1-forms back
to the product M1 × M2 . Then we have:
Lemma 10.5.5 (Lemma 5.9 in [6]).
(10.18a) Γ∗ (σ̂) = σ̂ 2 , Γ∗ (η̂) = η̂ 2 , Γ∗ (σ̌) = σ̌ 1 , Γ∗ (η̌) = η̌ 1 ,
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
10.5. The superposition formula 303
and
(10.18b) Γ∗ (R̂ji θX
j
) = Λij (x2 )(θ̂1j + θ̌2j ).
j
θ X , η̌, η̂ are all in W, hence so are their derivatives. On M1 , θ̂i = R̂ji θX , so
that
dθ̂ 1 ≡ R̂B σ̌ 1 ∧ σ̌ 1 mod θ̂ 1 , η̌ 1 .
Since the left-hand side belongs in W, the 2-forms B σ̌ 1 ∧ σ̌ 1 are in W, and
similarly for Aσ̂ 2 ∧ σ̂ 2 . Thus, the pullbacks of the generators of I given by
(10.7) all lie in W.
dθ̂ 1 ≡ 0 mod θ̂ 1 , Γ∗ I.
It follows that W = {θ̂1 , Γ∗ I}alg , and so is an integrable extension of I.
Remark 10.5.10. We have just proved that any Darboux-integrable system
I arises (locally) as the quotient EDS of the product of two systems W1 and
W2 which have a common symmetry group G. The converse is also true:
given two such systems (generated by 1-forms and 2-forms on manifolds
M1 and M2 ) with a common symmetry group, the quotient of the product
system W1 + W2 is Darboux-integrable, under certain technical assumptions
(see Theorem 6.1 in [5]). Note that there is actually a larger group G × G
acting on M1 × M2 as symmetries of W1 + W2 , and we recover I as the
quotient by the action of the diagonal subgroup Gdiag ⊂ G × G. If we
form the quotient by a different subgroup, then under suitable regularity
assumptions the resulting quotient differential system is related to I by a
Bäcklund transformation (see Theorem 9.1 in [5]).
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
10.5. The superposition formula 305
Example (10.2.4 continued). From Exercise 10.5.7 (and noting that there
are no 1-forms η̌),
W1 = {θ̂11 , θ̂12 }diff = {q1−1 dq1 − T dy, du1 − q1 dy}diff
and similarly
W2 = {θ̌21 , θ̌22 , η̂2 }diff = {dw+(w2 −R)dx, q2−1 dq2 −w dx, du2 +(x−u2 w)dx}diff .
(For the sake of convenience, we have omitted the subscripts on coordinates
x2 , y1 , w2 , R2 and T1 .) To finish the example, we will generate integral curves
for each of these systems, and use the superposition formula to produce
solutions for the PDE we started with.
Integral curves of W1 are easily obtained, by letting u1 = f (y), q1 = f (y)
and T = f (y)/f (y) for an arbitrary monotone function f . For W2 , observe
that along integral curves
d u2 x
=− .
dx q2 q2
Thus, integral curves of W2 can be generated by letting u2 /q2 = g(x) for an
arbitrary monotone function g, and setting
x xg(x)
q2 = − , u2 = − .
g (x) g (x)
Formulas for w and T in terms of g and its derivatives can then be obtained
by substituting these into the system W2 . Finally, the superposition formula
gives
x
(10.20) u(x, y) = q2 u1 + u2 = − (f (y) + g(x)).
g (x)
Exercises 10.5.11:
1. Check that (10.20) satisfies the PDE (10.4).
2. Carry out the coframe adaptations and construct the superposition for-
mula for the Darboux-integrable PDE uxy = ux uy /(u − x). In particular,
show that the Vessiot algebra is again 2-dimensional and non-Abelian (cf.
Example 6.2 in [6]).
3. Do the same for Liouville’s equation uxy = eu , and show that its Vessiot
algebra is sl(2). In particular, re-derive the solution formula (7.16) due to
Goursat (cf. Example 6.1 in [6]).
4. Show that the Born-Infeld system uy = vux , vy = uvx is Darboux-
integrable with a 1-dimensional Vessiot algebra.
5. The system
vx vy − ux uy ux vy + uy vx
uxy = u
, vxy = −
2(1 + e ) 2(1 + eu )
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
306 10. Superposition for Darboux-Integrable Systems
defines wave maps from the Minkowski plane into R2 with the confor-
mally flat metric (du2 + dv 2 )/(1 + e−u ). Show that the system is Darboux-
integrable, with Vessiot algebra R ⊕ sl(2).
6. Show that the system uxy = 2eu − ev , vxy = 2ev − eu is Darboux-
integrable after one prolongation, and its Vessiot algebra is sl(3).
7. Show that the system
(uxy + zuxx )uz (uyy + zuxy − z)uz
uxz = , uyz =
uy + zux − yz uy + zux − yz
is decomposable of type [5, 2], and is Darboux-integrable with a 3-dimension-
al nilpotent Vessiot algebra.
8. Show that the system
2u + 1 ux uz uy uz
uxy = ux uy , uxz = , uyz =
u(u + 1) u+1 u
is decomposable of type [4, 2], and is Darboux-integrable with a 2-dimension-
al non-Abelian Vessiot algebra.
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
https://doi.org/10.1090//gsm/175/11
Chapter 11
Conformal Differential
Geometry
307
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
308 11. Conformal Differential Geometry
Proof. The totally traceless part of Θ̂ equals the totally traceless part of
Θ.
We set up the EDS for metrics in a conformal class that are of constant
scalar curvature: Let F denote the orthonormal frame bundle n+1
of the Rie-
mannian manifold (M, g). Work on Σs ⊂ F × R+ × R × R n ( 2 ) where the
second, third and fourth factors have coordinates (σ, σj , sij ) and Σs is the
hypersurface given by setting (11.4) to zero. Consider the differential forms
θ := dσ − σj η j ,
θj := dσj − σm ηjm − sjm η m
and the ideal they generate. Integral manifolds of this ideal correspond to
metrics in the conformal class of constant scalar curvature.
On Σs we have the identity
. σ n /
0 = d σ2s + traceg (∇g dσ) − |dσ|2g
2(n − 1) 2(n − 1)
which will be needed in the following exercise.
Exercise 11.1.8: Show that the system on Σs is involutive, so that locally
there are many metrics in a conformal class with constant scalar curvature.
metric there are no local integral manifolds. What are the local integral
manifolds when (M, g) is Euclidean space?
M.
(b) Show that A = r−2 B, b = tv defines a group isomorphism from P1,n+1
to CO(V ) V ∗ . Thus, we conclude that the fiber of FCO(V ) is isomorphic
(1)
to P1,n+1 .
To further the comparison with the flat case, define the ray bundle R+ ⊂
S 2 T+∗ M whose fiber corresponds to the choices of g ∈ [g], which generalizes
the choices of e0 ∈ [e0 ] that fixed a pointwise choice of metric in the flat
case. Write θ = 2ρ IdV +α where α is so(V )-valued and ρ is a scalar-valued
1-form. Then, since ρ corresponds to infinitesimal dilations, the forms semi-
(1) (1)
basic to the projection FCO(V ) → R+ , where R+ = FCO(V ) /(SO(V ) V ),
are ω i , ρ.
Following the discussion in §9.3, we want to obtain a V ∗ -valued 1-form
(1)
β to give a connection form for the bundle FCO(V ) → FCO(V ) . In particular,
(1)
we will need ω, β, ρ, α to give a coframing of FCO(V ) . Differentiating both
sides of
(11.8) dω i = 2ρ ∧ ω i − αji ∧ ω j
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
314 11. Conformal Differential Geometry
shows
(−2δij dρ + dαji + αki ∧ αjk ) ∧ ω j = 0,
which, by the generalized Cartan Lemma, implies −2δij dρ + dαji + αki ∧
αjk ≡ 0 mod{ω 1 , . . . , ω n }. In particular, setting i = j we see that dρ ≡
(1)
0 mod{ω 1 , . . . , ω n }. Thus there exist forms βi ∈ Ω1 (FCO(V ) ) such that
The choice of βi is not unique, as setting β̃i = sij ω j + βi with sij = sji ,
one gets the same equation with β̃i , as was predicted by Remark 11.2.5.
Writing dαji + αki ∧ αjk − βj ∧ ω i + βi ∧ ω j = 12 Rjkl
i ω k ∧ ω l , we see that under
such a change,
i
R̃jkl i
= Rjkl + (−δli sjk + δlj sik − δkj sil + δki sjl ),
which is similar to, but simpler than the expression (11.2). In particular,
since the change here is by s g where s is an arbitrary quadratic form, it
is clear that one can choose s to make R̃ traceless. To see this explicitly, set
l = i and sum:
i
R̃jki i
= Rjki + (−nsjk + sjk − δkj sii + sjk )
i
= i
Rjki − (n − 2)sjk − δjk sii .
i
Thus if we take
1 1
(11.10) sjk = i
(Rjki − i
δjk Rlli ),
n−2 2n − 2
l
then i
R̃jki = 0.
Remark 11.2.6. The above normalization is another example of the utility
of working on the proper bundle. With a fixed metric in a conformal class
we can have Ricci vanishing at a point by Exercise 11.1.5. Here we obtain
the analog of Ricci vanishing at all points.
Equation (11.10) is the most natural (but not canonical) choice to nor-
malize the curvature. We now have our coframe.
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
11.2. Conformal differential geometry as a G-structure 315
(1)
Proposition 11.2.7. On FCO(V ) there is a unique coframing ω, α, ρ, β where
the forms are V -, so(V )-, R-, and V ∗ -valued respectively, satisfying
dω = (2ρ IdV −α) ∧ ω,
dρ = − 12 β ∧ ω,
dα = −α ∧ α + ω ∧ β − t (ω ∧ β) + W,
where W = 12 Wjkl i (e ⊗ ω j ) ⊗ ω k ∧ ω l satisfies the symmetries of the Weyl
i
curvature tensor. Here we write β as a row vector and ω as a column
vector, so, e.g., β ∧ ω is the scalar-valued 2-form βi ∧ ω i . With indices, these
equations read as
dω i = (2ρδji − αji ) ∧ ω j ,
1
dρ = − βi ∧ ω i ,
2
1 i k
dαji = −αki ∧ αjk + βj ∧ ω i − βi ∧ ω j + Wjkl ω ∧ ωl .
2
It remains to study dβi , which may or may not contain new differen-
tial invariants. When n = 3 we expect it to do so, because otherwise all
conformal structures in three dimensions would be flat. Differentiating the
structure equations gives
(11.11) 0 = 2d2 ρ = −dβj ∧ ω j + βi ∧ (2δji ρ − αji ) ∧ ω j ,
(11.12) 0 = d2 αji = 12 (DW )ijkl ∧ ω k ∧ ω l + dβj ∧ ω i − dβi ∧ ω j
− βk ∧ (2δki ρ − αjk ) ∧ ω i + βk ∧ (2δkj ρ − αik ) ∧ ω j .
Here
(11.13) (DW )ijkl := dWjkl
i m i
+ Wjkl θm − Wmkl
k
θjm − Wjml
k
θkm − Wjkm
k
θlm .
By the generalized Cartan Lemma A.1.9, (11.11) implies
dβj + 2ρ ∧ βj + βi ∧ αji = Cjk ∧ ω k
(1)
for some Cjk ∈ Ω1 (FCO(V ) ). Substitute this into the second equation to
obtain
[ 12 (DW )ijkl − Cik δlj + Cjk δli + Cil δkj − Cjl δki ] ∧ ω k ∧ ω l = 0,
so the term in the brackets is semi-basic. Setting i = l and summing we
obtain
(n − 2)Cjk + δkj Cii ≡ 0 mod{ω 1 , . . . , ω n }.
i
Cjk ≡ 0 mod{ω 1 , . . . , ω n }. Examining
Examining the case j = k we see that
the term j = k gives (n − 1)Cjj + i=j Cii ≡ 0 mod{ω 1 , . . . , ω n }. Since
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
316 11. Conformal Differential Geometry
n ≥ 3, this implies the Cjj are semi-basic as well. Thus we may write
In summary:
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
11.2. Conformal differential geometry as a G-structure 317
(1)
Proposition 11.2.9. On FCO(V ) there is a unique coframing ω, α, ρ, β where
the forms are respectively V -, so(V )-, R-, and V ∗ -valued, satisfying
We will show that when n > 3, the Cotton-York tensor lives precisely
in the copy of (S21 V ∗ )0 that lives in S221 V ∗ , so in particular there must
be a relationship between it and the component of ∇W in (S21 V ∗ )0 . That
relation is:
Proposition 11.2.12. For all i, j, k
(∇W )ijkmm + (n − 3)Cijk = 0.
m
so
(∇P)ijm − (∇P)imj
1 sm 1 sj
= (∇ Ric0 )ijm + δij − (∇ Ric0 )imj − δim .
n−2 2n(n − 1) n−2 2n(n − 1)
Setting i = j and summing over i gives the desired identity.
Proof. While a direct proof is possible, the following proof is also instruc-
tive. The idea is to change coordinates via the antipodal map α of the
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
11.3. Conformal Killing vector fields 321
x
α(φX (x, t)) = φα∗ X (α(x), t) = y + 12 tv = − + 1 tv.
|x|2 2
Applying the antipodal map to both sides gives
x
x
|x|2
− 2t v
1
φX (x, t) = α(− 2 + 2 tv) = x .
|x| | |x|2 − 2t v|2
So
xj − 2t v j |x|2
xj (t) = .
1 − tv, x + 14 t2 |v|2 |x|2
Remark 11.3.3. The above proof, and the proof of Liouville’s theorem
below were communicated to us by M. Eastwood.
We know from §9.6 that the largest possible space (even locally) of con-
formal Killing vector fields on anyconformal manifold (when n ≥ 3) is given
by V ⊕ co(V ) ⊕ V ∗ (i.e., a n+22 -dimensional space), and this occurs for
conformally flat manifolds. In the Riemannian case there were examples of
nonflat manifolds (namely, space forms) with the maximal space of (Rie-
mannian) Killing vector fields. This does not occur in conformal geometry:
Proposition 11.3.5. Let (M, [g]) be a manifold of dimension n ≥ 3 with
conformal structure. If the
dimension of the space of conformal Killing
n+2
vector fields on M is 2 , then (M, [g]) is conformally flat.
(1)
Proof. The hypothesis implies that any element of FCO(V ) over a fixed
(1)
point x gives rise to a conformal Killing vector field by embedding FCO(V )
into the bundle of 2-jets of vector fields. (Recall that FGL(V ) (M ) may be
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
322 11. Conformal Differential Geometry
(1)
embedded in J 1 (M, V ) and similarly FCO(V ) may be identified with a sub-
bundle of the bundle of 2-jets sitting inside J 1 (J 1 (M, V )).) This in turn
(1)
implies, via parallel transport in FCO(V ) , that the coefficients of W, C must
be constant. In particular, (DW )ijkl , as defined in (11.13), is zero (by the
same argument as the proof of Proposition 9.7.7), but this implies C = 0 by
(11.16) (assuming n > 3).
l β
Exercise 11.3.6: Compute dCijk and show it has a term of the form Wijki l
appearing in it.
Let E[τ ] := Γ(M, R[τ ]) denote the space of smooth sections and E[τ ]+ the
space of sections where t > 0.
τ
Remark 11.4.1. In [29] the bundle R[τ ] is denoted D n . For those familiar
with the notation from algebraic geometry, the line bundles R[τ ], when τ
is an integer and M is the sphere, are real analogs of OCPN (τ ) and, for
arbitrary M , of OX (τ ) for projective manifolds X.
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
11.4. Conformal densities and the Laplacian 323
Differentiate (11.20) to obtain Rh∗ (dfuh ) = −2τ f (u)r−1 dr + r−2τ dfu , and
take h = Idn+2 to see that the derivative of f in the fiber direction is −2τ f ρ.
Therefore
(11.21) df = −2τ f ρ + fi ω i
(1)
for some functions fi : FCO(V ) → R.
Remark 11.4.3. If one prefers to work with sections, let f˜ ∈ E[τ ] denote the
section corresponding to f and let ∇R[τ ] be the connection on R[τ ] induced
(1) (1)
from the connection on FCO(V ) . Let S : M → FCO(V ) be a local section.
Then (11.21) may be written
∇R[τ ] f˜ = S ∗ (fi ω i ).
is basic.
Exercise 11.4.4: Verify that gf is indeed basic.
The following two identities will be useful when discussing the conformal
Laplacian.
Differentiating (11.21) gives
thus
(11.26) P = Pf = τ −1 f jk η j η k .
dh = hi η i ,
dhi = hj ηij + (hij + hf ij )η j .
Writing ∇ = ∇gf ,
i.e., on Ff ,
n−2
(11.29) Δ[g] h̃ = Δgf h + sg h,
4(n − 1) f
where s = sgf is the scalar curvature.
In summary:
Proposition 11.4.6. Fix a metric g ∈ [g]. Relative to the trivializations of
R[ n2 − 1] and R[ n2 + 1], the conformal Laplacian becomes
n−2
Δ[g] = Δg + sg ,
4(n − 1)
where sg denotes the scalar curvature.
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
326 11. Conformal Differential Geometry
(1)
The tractor bundle. The most natural vector bundle associated to FCO(V )
is the one corresponding to the standard representation of P1,n+1 on Rn+2
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
11.5. Einstein manifolds in a conformal class and the tractor bundle 327
2 ∗
have ∇S T M [2] g = 0 since ∇T h = 0.
• The Lorentzian metric h allows us to identify T and T ∗ , and the
conformal metric g allows us to identify T M [−1] T ∗ M [1].
• A choice of g ∈ [g], equivalently a choice of any positive section of
any R[s] with s = 0, determines a splitting of T into its graded
components because there is a reduction of structure group P1,n+2
to SO(n) ⊂ SO(n + 1, 1).
Define a second order differential operator
(11.31) D : E[1] → Γ(T ∗ ),
⎛ ⎞
h
h → ⎝ dh mod T1∗ ⎠ .
Δgf h mod T2∗
This operator is well defined because modulo T2∗ , Δgf h is independent of
the choice of f .
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
328 11. Conformal Differential Geometry
The tractor bundle and the Einstein equations. Following the nota-
tion of [79], write a section of T ∗ as I = t (h, hi ω i mod T1∗ , k mod T2∗ ). Then
⎛ ⎞ ⎛ ⎞
h dh − 2hρ − hj ω j
∗
(11.32) ∇T ⎝hi ω i ⎠ = ⎝(dhi − hj αij − hβi + ω i k)⊗ω i mod T1∗ ⎠ .
k dk + 2kρ − j hj βj mod T2∗
First note that I is covariant constant if and only if I = D(h) for some
h ∈ E[1]. Comparing the EDS of Exercise 11.5.2 and (11.32), we see that a
section of T ∗ that is covariant constant is nearly the same thing as a choice
of an Einstein metric in a conformal class! The only thing that could go
wrong is if h = 0 at points. (If h < 0, simply use −h instead of h.) R.
Gover [78, 79] was the first to observe this and recognize its utility. Let
∗
I denote a section of T ∗ . Call a manifold (M, g, I) with ∇T I = 0 almost
Einstein. Note that if h is not identically zero, then we obtain an honest
Einstein metric on an open subset of M . We may use sgh to get more precise
information about the locus where the metric gh is well-defined.
Theorem 11.5.7 ([79]). Let (M, g, I = D(h)) be almost Einstein and let
Zh ⊂ M denote the zero set of h.
(1) If the constant sh is positive, then Zh = ∅ and (M, gh ) is an Einstein
manifold with positive scalar curvature.
(2) If sh = 0, then either Zh = ∅ and (M, gh ) is a Ricci flat Einstein
manifold, or Zh consists of isolated points and (M \Zh , gh ) is a Ricci
flat Einstein manifold.
(3) If sh < 0, then Zh is either empty or a smooth hypersurface and
(M \Zh , gh ) is an Einstein manifold with negative scalar curvature.
So σ(x) = 0 (i.e., h(x) = 0) implies sgσ ≤ 0, which proves the first case.
For the second case, if σ(x) = 0, we see dσx = 0 as well. Since a parallel
section of a vector bundle is nonvanishing, ∇gf (dσ)x must be nonzero to
have lx nonzero. This implies x is an isolated zero of σ, which is what we
were trying to show.
For the third case, since |dσx | = 0, the inverse function theorem guar-
antees the zero locus is smooth in the neighborhood of any x ∈ Zh , that is,
smooth everywhere.
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
https://doi.org/10.1090//gsm/175/12
Chapter 12
Projective Geometry
II: Moving Frames and
Subvarieties of
Projective Space
331
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
332 12. Projective geometry II
The Fubini cubic form. Unlike Euclidean geometry, the projective funda-
mental forms do not give enough information to determine local equivalence
of complex submanifolds up to projective transformations (e.g., the case of
hypersurfaces discussed in Remark 4.1.9). Thus we differentiate further to
find higher-order differential invariants:
μ
0 = d(ωαμ − qαβ ω0β )
μ μ μ δ μ δ
= (−dqαβ − qαβ ω00 − qαβ
ν
ωνμ + qαδ ωβ + qβδ ωα ) ∧ ω0β .
μ
The Cartan Lemma implies there exist functions rαβγ , defined on F 1 , sym-
metric in their lower indices, satisfying
μ
(12.1) rαβγ ω0γ = −dqαβ
μ μ
− qαβ ω00 − qαβ
ν μ δ
ωνμ + qαδ μ δ
ωβ + qβδ ωα .
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
12.1. The Fubini cubic and higher order differential invariants 333
Moving in the fiber by the block diagonal matrices g00 , gβα , gνμ does not
change F3 . However, F3 does not descend to be a well-defined section of
S 3 T ∗ M ⊗N M , because if (ẽ0 , ẽα , ẽμ ) is a new frame with
ẽμ = eμ + gμ0 e0 + gμα eα ,
(12.3) ẽα = eα + gα0 e0 ,
then
μ μ μ ν μ
(12.4) r̃αβγ = rαβγ + Sαβγ gα0 qβγ + Sαβγ gνδ qαβ qγδ ,
where Sαβγ is cyclic summation in the fixed indices α, β, γ.
As mentioned above, F3 ∈ Γ(F 1 , π ∗ (S 3 T ∗ M ⊗ N M )) is an example of
μ
a relative differential invariant. Adopt the notation Δrαβγ to denote the
μ
change in rαβγ by a fiber motion of the type in (12.3). By (12.4),
μ μ ν μ
Δrαβγ = Sαβγ (gα0 qβγ + gνδ qαβ qγδ ).
Exercises 12.1.8:
1. Prove Theorem 12.1.6.
2. Show that the prolongation property is equivalent to the quadratic form
v P being in II(Nx∗ M ) for every v ∈ Tx M and every P ∈ III(N2,x
∗ M ).
The k-th fundamental form FFk . The exercises below give two definitions
of the k-th fundamental form FFkx,M ∈ S k Tx∗ M ⊗N(k−1),x M and the k-th
osculating space
T̂ (k) (k−1) 1k (S k T M ).
M + FF
x M = T̂ x x
Below the exercises we give yet another definition of the k-th fundamental
(k)
form. Let Nk,x M = Tx PV /T x M .
Exercises 12.1.9:
1. Define FFk by defining a higher-order Gauss map and taking its deriva-
tive.
2. Define FFk by refining the flag and differentiating, as we did with III =
FF3 .
3. Show the two procedures above give the same differential invariant,
called the k-th fundamental form, which measures to first order how M
is leaving its (k − 1)-st osculating space.
∗
Notation. Let |FFk |x,M = FFk (N(k−1),x M ) ⊂ S k Tx∗ M .
Exercise 12.1.10: Show that at general points x ∈ M , |FFk |M,x ⊆ (Tx∗ M ⊗
|FFk−1 |x,M ) ∩ S k Tx∗ M , i.e., that the prolongation property persists to higher
fundamental forms.
the choice of extension to vector fields is standard (see, e.g., [174]). Then
dk e0 ⊗e0 = FFk .
For example, keeping the index ranges on F 2 that we used above to
study the third fundamental form, we have
d1 e0 ≡ ω0α eα mod{e0 },
d2 e0 = ω0α deα mod{e0 , eα }
= ω0α ωαξ eξ mod{e0 , eα },
d3 e0 = ω0α ωαξ deξ mod{e0 , eα , eξ }
= ω0α ωαξ ωξφ eφ mod{e0 , eα , eξ }.
proving that the quadrics in |II| are spanned by the 2 × 2 minors of the
matrices in TE G(k, W ).
Exercise 12.2.1: Determine FFlG(k,W ) and its base locus.
Since the Grassmannian is cut out by polynomials of degree two, the set
Baseloc |IIG(k,W ),E | consists of tangent directions to lines on G(k, W ) that
pass through E. To see these lines explicitly, a point [x ⊗ y] ∈
Baseloc |IIG(k,W ),E | ⊂ P(TE G(k, W )) determines a (k − 1)-plane Fx ⊂ E ⊂
V and a (k+1)-plane Ly ⊂ V containing E, where Fx = ker x and Ly = E+y.
The corresponding line in G(k, W ) is {E | Fx ⊂ E ⊂ Ly }.
Fundamental forms of Segre varieties. Let U = Ca+1 , W = Cb+1 , and con-
sider the Segre embedding Seg(PU × PW ) ⊂ P(U ⊗ W ). Here we use
G = GL(U )×GL(W ) ⊂ GL(U⊗W )-frames. Let (e0 , . . . , ea ) and (f0 , . . . , fb )
denote bases of U, W , and let the respective Maurer-Cartan forms be de-
noted (ω) and (η). Use index ranges 1 ≤ α ≤ a, 1 ≤ j ≤ b. Let
[e0 ⊗ f0 ] ∈ Seg(PU × PW ). Note that e0 ⊗ f0 , eα ⊗ f0 , e0 ⊗ fj , eα ⊗ fj form
a basis of U ⊗W . We compute
(12.8) d(e0 ⊗f0 ) ≡ ω0α eα ⊗f0 + η0j e0 ⊗fj mod e0 ⊗f0 ,
(12.9) d2 (e0 ⊗f0 ) ≡ ω0α η0j eα ⊗fj mod T̂[e0⊗f0 ] Seg(PU × PW ).
Equation (12.8) shows that
T̂[e0⊗f0 ] Seg(PU × PW ) = e0 ⊗W + U ⊗f0 .
Let U = (U⊗f0 mod e0⊗f0 )⊗(e0⊗f 0 ) and W = (e0⊗W mod e0⊗f0 )⊗(e0⊗f 0 ),
so
T[e0⊗f0 ] Seg(PU × PW ) = U ⊕ W .
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
12.3. Ruled and uniruled varieties 339
IIC,[e0 ] = ω12 ω01 ⊗(e2 mod T̂[e0 ] C)⊗e0 + ω1μ ω01 ⊗(eμ mod T̂[e0 ] C)⊗e0 .
0 = deA , eB = ωA
C
eC , eB + eA , deB ,
which implies
deB , eA = −ωA
B
,
i.e.,
deB = −ωC
B C
e .
Since ρ is submersive, we may calculate dim X ∨ by determining the
number of independent forms in den+a :
(12.10) den+a ≡ −ωαn+a eα − ωn+λ
n+a n+λ
e mod en+a .
Here we use the index range 1 ≤ λ, κ ≤ a − 1. The forms ωn+λ n+a
are inde-
pendent and independent of the semi-basic forms for π , and dim{ωαn+a } =
rank q n+a . We obtain an explicit description of the tangent space:
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
342 12. Projective geometry II
∗ X ∨.
is a linear Pδ∗ ⊂ X and may be identified with PNH
Reversing the roles of X and X ∨ in Corollary 12.4.4, we obtain:
Theorem 12.4.8. Let X n ⊂ Pn+a be a smooth variety, let H ∈ X ∨
be a general point, let v ∈ NH ∗ X ∨ be any nonzero vector, and let r =
∨
rank IIX ∨ ,H (v). Then dim(X ) = r + a − 1. In particular, |IIX ∨ ,H | is a
system of quadrics of projective dimension equal to the dual defect δ∗ , and
of constant rank n − δ∗ .
The first assertion of Theorem 12.4.8 is due to B. Segre [168]. It was
rediscovered by Griffiths and Harris in [85]. The second appeared in [103].
Corollary 12.4.9. Let X n ⊂ Pn+a be a smooth variety, and let H ∈ X ∨
be a general point. Then IIIX ∨ ,H = 0.
It is difficult to have a system of quadrics of constant rank. By a classical
result in linear algebra, it is impossible if r is odd, and, by [103], impossible
if the dimension of the system is one greater than the dimension of the vector
space minus the rank. This unifies and recovers two important results:
Theorem 12.4.10 (Landman parity theorem [57]). Let X n ⊂ Pn+a be a
smooth variety. Then n − δ∗ is even.
Theorem 12.4.11 (Zak [191]). Let X n ⊂ Pn+a be a smooth variety not
contained in a hyperplane. Then δ∗ < a − 1, i.e., dim X ∨ ≥ dim X.
One consequence of Zak’s theorem is the following:
Theorem 12.4.12 ([122]). Let X n ⊂ Pn+a be a smooth variety, let x ∈
X be a general point and let q ∈ |IIX,x | be a generic quadric. Then
dim(Singloc(q)) ≤ a − 1.
We will generalize Theorem 12.4.12 in §12.11.
2. Determine |IIG(2,5) |.
3. One gets the second fundamental form of what variety when one uses
Exercise 1 in the case B = M2×q ?
We conclude:
Proposition 12.5.3 ([85]). Let x ∈ X be a general point and let v ∈ Tx X
be a generic tangent vector. Then dim τ (X) = n + rank IIv .
This agrees with our expectation that, generally, τ (X) has dimension
min{2n, n + a}. If this expectation fails, then we say τ (X) is degenerate.
For X with τ (X) degenerate, define the tangential defect to be
δτ (X) = 2n − dim τ (X).
Exercise 12.5.4: Calculate the tangential defects of the following vari-
eties:
(a) a curve X 1 ⊂ Pa+1 that is not a line;
(b) the Veronese vd (Pn );
(c) the Segre Seg(Pa × Pb );
(d) the Grassmannian G(2, 6).
the secant defect of X, and we say σ(X) is degenerate if δσ (X) > 0. While
Terracini’s Lemma 4.2.12 enables us to easily compute dim σ(X) using one
derivative at two general points of X, computing dim σ(X) at one general
point of X is more subtle. We begin our calculation by first observing that
for varieties with degenerate secant varieties, part of the cubic form descends
to be a well-defined differential invariant.
For a vector space T and a subspace A ⊂ S 2 T ∗ , recall that
Singloc(A) = {v ∈ T | v q = 0, ∀q ∈ A}.
For v ∈ T , let
(12.11)
ωμν ωβμ ω0β |π∗ (S 3 (Singloc(Ann(v)))) ⊗(eν mod{T̂ , II v (T )})⊗e
0
by (12.11).
Exercises 12.5.6:
1. Show that III may be recovered from the forms III v for all v ∈ T .
2. Show that if v ∈ T is a II-generic vector, then ker IIv ⊂ Singloc(Ann(v)).
General points of σ(X) are of the form [e0 + sf0 ], where [e0 ], [f0 ] are
points of X and s ∈ C. Label the corresponding frames for X above [e0 ], [f0 ]
with the same letters. Terracini’s Lemma 4.2.12 implies that the dimension
of σ(X) is one less than the number of linearly independent vectors among
e 0 , e α , f0 , fβ .
Now we use the assumption that X is smooth and connected. Set
v = e1 ∈ T[e0 ] X. Let (f0 (t), . . . , fn+a (t)) be a coframing over an arc f0 (t)
such that (f0 (0), . . . , fn+a (0)) = (e0 , . . . , en+a ) and f0 (0) = e1 . Expand
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
12.5. Secant and tangential varieties 349
t2 t3
f0 (t) = e0 + te1 + f0 (0) + f0 (0) + O(t4 ),
2 3!
t 2 t3
fα (t) = eα + tfα (0) + fα (0) + fα (0) + O(t4 ).
2 3!
Now (ignoring twists),
)
f0 (0) ≡ II(v, v)
mod T̂[e0 ] X
fα (0) ≡ II(v, eα )
and
)
f0 (0) ≡ III v (v, v, v)
mod{T̂[e0 ] X, IIv (T )}.
fα (0) ≡ III v (v, v, eα )
We will use the Fulton-Hansen Theorem when proving the rank restric-
tion theorems in §12.11, and again when giving a proof of Zak’s theorem on
Severi varieties using moving frame techniques in §12.12.
The tangential variety can be generalized as follows: if Y ⊆ X ⊆ Pn+a ,
and y ∈ Y , define Ty (Y, X) to be the union of P1∗ ’s, where P1∗ is a limit
of P1xy ’s when x ∈ X and y ∈ Y , and x, y → y0 ∈ Y . Define τ (Y, X) =
y∈Y Ty (Y, X), the variety of relative tangent stars. As observed by Zak,
the Fulton-Hansen Theorem 12.5.7 generalizes to relative tangent stars and
joins as follows:
Theorem 12.5.10 ([191]). Let X n , Y y ⊂ PV be varieties, respectively of
dimensions n, y. Assume Y ⊆ X. Then either
i. dim J(Y, X) = n + y + 1 and dim τ (Y, X) = n + y,
or
ii. J(Y, X) = τ (Y, X).
Using Theorem 12.5.10, one obtains:
Theorem 12.5.11 (Zak’s theorem on tangencies [191]). Let X n ⊂ Pn+a
be a variety not contained in a hyperplane. Let dim Xsing = b, with the
convention that b = −1 if X is smooth. Let L be any Pn+k ⊂ Pn+a ; then
dim{x ∈ X | T̃x X ⊆ L} ≤ k + (b + 1).
The cubic form detA tells us which elements of H are of less than full
rank. One can also unambiguously define a notion of being rank one, either
by taking 2 × 2 minors or by noting that under the G action each x ∈ H
is diagonalizable and one can take as the rank of x the number of nonzero
elements in the diagonalization of x. Let
X := P{rank one elements of H} = P{G-orbit of any rank one matrix}.
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
352 12. Projective geometry II
Exercise 12.6.5: What are the varieties analogous to S5 for the other di-
vision algebras?
Cominuscule varieties.
Definition 12.6.6. Let X = G/P ⊂ PV be a homogeneous variety where
G is a complex semi-simple Lie group acting linearly on V and P is the
subgroup stabilizing a point of X (such a P is called a parabolic subgroup).
We say X is a generalized cominuscule variety if Tx X contains no proper
irreducible P -submodule. X is a cominuscule variety if moreover G is simple
and the embedding is minimal, which is equivalent to the embedding not
being a Veronese re-embedding.
Remark 12.6.8. For readers familiar with Dynkin diagrams, H and the H-
module Tx X can be determined pictorially: H is a semi-simple group with
Lie algebra whose Dynkin diagram is that of G with the node corresponding
to the root defining the maximal parabolic P removed. Moreover, if the
diagram is simply laced, the H-module Tx X is the irreducible module with
highest weight the sum of the fundamental weights corresponding to the
nodes of the Dynkin diagram adjacent to the deleted node.
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
354 12. Projective geometry II
Exercise 12.2.1 shows that equality in (12.6) holds for the examples of
Grassmannians. In fact, equality holds for all cominuscule varieties:
Theorem 12.6.9 ([136]). Let X = G/P ⊂ PV be a cominuscule variety
and let x ∈ X. Then for k ≥ 2,
X,x | = |FFX,x |
|FFk+1 2 (k−1)
.
Spinor varieties and the spin representation. For sl(W ), the funda-
mental representations are just the exterior powers of the standard repre-
sentation (see §A.6). For so(W ), assuming for simplicity that dim W = 2m,
fundamental representations are furnished by the exterior powers of the
standard representation, up to Λm−2 W . There are two other fundamental
representations of so(W ), dual to one another, one of which we now describe.
These representations are called the spin representations. In terms of Lie
groups, the spin representations are representations of the simply connected
double cover of SO(W ), called Spin(W ), but not of SO(W ).
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
12.6. Cominuscule varieties and their differential invariants 355
We will use Corollary 12.6.10 to construct a model for the spin repre-
sentation. Readers familiar with Dynkin diagrams may skip the first step,
where the tangent space is calculated, thanks to Remark 12.6.8.
Before starting, we examine the Grassmannian G(k, W ) from the per-
spective we will use to construct the spin representation. Say we knew
Corollary 12.6.10 but were unaware of the Plücker embedding. We would
then construct V = Λk W as follows: fix E ∈ G(k, W ); then V is the sum
of the successive quotient spaces Λk E (which we identify with the point
E), E ∗ ⊗W/E = Λk−1 E ⊗W/E, Λ2 E ∗ ⊗Λ2 (W/E) = Λk−2 E ⊗Λ2 (W/E),...,
Λk E ∗ ⊗ Λk (W/E) = Λ0 E ⊗ Λk (W/E). Note that only the first space is
well-defined as a subspace of Λk W , but the sum of the first p spaces is a
well-defined subspace of Λk W for all p, so we obtain a Λk W equipped with
a filtration.
We now apply the same method to recover the spin representation. The
variety GQ−null (m, C2m ) has two isomorphic components; we define Sm to
be one of them.
Choose a basis of V = C2m so that Q = 0I I0 . With respect to this
basis SO(V, Q) has Maurer-Cartan form
i i
ωj ωm+k
m+k ,
ωjn+l ωm+l
working with its simply connected double cover Spin(W ). Since our calcu-
lation was on the level of Lie algebras, we got the correct answer.
Exercises 12.6.12:
1. Verify that Baseloc |IISm ,E | = G(2, E) by showing that II(v, v) = 0 if
and only if v = w1 ∧ w2 .
2. Calculate the higher fundamental forms of Sm via moving frames to
verify that |FFk | = Λ2k E ∗ = Ik (σk−1 (G(2, E))).
l−3
− (p − 2 + l)rαμ1 ,...,αp rαν p+1 ,...,αl−1 ων0 ].
p=2
This formula corrects the typographical errors in (2.20) of [124] (where the
proof is given without typographical errors).
1 defined by e = ∂ + f μ ∂ , e = ∂ . In
Consider the section of FM α ∂xα xα ∂xμ μ ∂xμ
this framing ω0 = dx , or, more precisely, s∗ (ω0α ) = dxα . We compute
α α
∂
. deα ≡ fxμα xβ dxβ ⊗
∂xμ
In terms of the Maurer-Cartan form, we have
deα ≡ ωαμ eμ mod{e0 , eβ },
compute
FF1v2 (X),v2 (x) = 2F1 F0 |x̂2⊥ ,
FF2v2 (X),v2 (x) = 2(F2 F0 + F1 F1 )|(x̂T̂ )⊥ ,
FF3v2 (X),v2 (x) = 2(F3 F0 + 3F2 F1 )|ker FF2 ,
v2 (X)
(12.13)
FF4v2 (X),v2 (x) = 2(F4 F0 + 4F3 F1 + 3F2 F2 )|ker FF3 ,
v2 (X)
..
.
Here Fj = Fj,X,x are the relative differential invariants of X ⊂ PV .
We make several observations. First, note that |FF1v2 (X),v2 (x) | Tx∗ X,
|FF2v2 (X),v2 (x) | S 2 Tx∗ X, and thus the dimensions of their kernels are re-
spectively n+a+1 2 − (n + 1) and n+a+12 − (n + 1) − n+1
2 . This says:
Assuming ker FF3v2 (X),v2 (x) is as small as possible, ker FF4v2 (X),v2 (x) will
be assured to be as small as possible if the map |II|⊗|II| → S 4 T ∗ has no
kernel.
Exercise 12.7.4: Show that the map |II| ⊗ |II| → S 4 T ∗ has no kernel if
there are no linear syzygies among the quadrics in |II|.
Proof. In each case, the extension to linear spaces will hold by polarizing
the forms. We prove only the first assertion in each case, as an analogous
equation at each order proves the next higher order.
μ
1. Assume v = e1 and q = qαβ ω0α ω0β . Our hypotheses imply q1β
μ
= 0 for
all β. Formula (12.1) reduces to
μ
r11β ω0β = −q11
ν μ
ων .
and again, both sides are zero. This and the corresponding higher-order
equations also prove both the classical Bertini Theorem and part 1 of the
theorem.
2. If h = en+a is not generic, we have to reduce to a subbundle F ⊂ F 1
where the ωνn+a are not necessarily independent. However, the hypotheses
ν = 0 for all ν, and the required vanishing still holds.
of part 2 state that q11
n+a δ ν ω n+a + r n+a ω + q n+a q ν ω and the
For part 3, note that r111δ ω0 = r111 ν 11 1 1 11 ν
right hand side is zero under our hypotheses. Part 4 is proven similarly.
and for all h, which will imply that there is a p-dimensional space of lines
passing through X at x.
We proceed by induction on h, using our lower-order equations to solve
for the connection forms ω1j in terms of the semi-basic forms ω0l , and then
plug into the higher Fh to show that Fh (e1 , . . . , e1 ) and Fh (e1 , . . . , e1 , es )
both vanish. Using (12.7) we obtain that, for λ ≤ k,
λ
r1,...,1,i ω0i = −r1,...,1,j
λ−1
ω1j .
These are k − 2 equations for the n − p 1-forms ω1j in terms of the n − p
semi-basic forms ω0j . Recall that n − p ≤ k − 2. The system is solvable,
because were the (k − 2) × (n − p) matrix (r1,...,1,j
λ ) not of maximal rank,
there would be additional directions in the tangent space to Σk0 . Thus
ω1j ≡ 0 mod{ω0p+1 , . . . , ω0n },
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
12.8. Varieties with vanishing Fubini cubic 363
so the equation
k+1
r1,...,1,β ω0β = −r1,...,1,j
k
ω1j
k+1 k+1 k+1
implies r1,...,1 = r1,...,1,2 = . . . = r1,...,1,p = 0. Thus the line through [e1 ] has
contact to order k + 1, and moreover T̃[e1 ] Σk+1 0 = T̃[e1 ] Σk0 . Now one can use
these equations iteratively to show the same holds to order k + 2 and all
orders, i.e., the component of Σk containing [e1 ] at a general point equals
the component of Σ∞ containing [e1 ].
use Q to denote both the quadratic form and the quadric hypersurface it
determines.
Exercise 12.8.3: Show that, with this choice of basis,
⎧⎛ 0 ⎞ ⎫
⎨ a0 a0β 0 0 β ⎬
a + an+1
= 0, a n+1
− a = 0,
so(V, Q) = ⎝aα0 aαn+1 ⎠
0 n+1 β 0
aαβ .
⎩ n+1 n+1 a α
+ aβ
= 0, aα
− a0
= 0 ⎭
0 aβ an+1 β α n+1 α
i∗ (F3 ) = 0.
The second equation implies that ωαα = ωββ for all α, β, and the condition
det(f ) = 1 implies trace(ω) = 0, so
n+1
(n + 2)ωαα = ω00 + ωn+1 + ω11 + . . . + ωnn
= 0.
Thus
ωαα = 0, ∀α,
n+1
ω00 + ωn+1 = 0.
To reduce the Maurer-Cartan form to the Maurer-Cartan form for
α
SO(V ), we need to show that ωn+1 − ωα0 = 0 and ωn+1
0 = 0. We consider
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
12.9. Associated varieties 365
the coefficients of F4 :
rαβγ ω0 = 0, α, β, γ distinct,
β
rααβ ω0 = (ωβ0 − ωn+1 ), α = β,
rααα ω0 = 3(ωα0 − α
ωn+1 ).
We see that for all α = β, we have rαααβ = 0, rααββ = r and rαααα = 3r for
some function r.
Exercise 12.8.5: Show that we may reduce our frame bundle further to a
subbundle where r ≡ 0.
Now that all coefficients of F3 and F4 are zero, we see that F5 and all
higher Fk must be zero as well, as, for example, the coefficients of F5 are
polynomials in the coefficients of II, F3 , F4 , but the coefficients of II only
occur multiplied by the coefficients of F3 .
Exercise 12.8.6: Show that Fubini’s Theorem still holds for manifolds with
degenerate Gauss maps as long as the rank of the Gauss map is at least two.
In this case the closure of M is a singular quadric hypersurface.
Exercise 12.8.7: Show that, for G(k, W ) ⊂ P(Λk W ) and Seg(Pk × Pl ),
there are natural framings with F3 = 0.
Zr,s (X)
= {E ∈ G(r + 1, V ) | ∃x ∈ Xsmooth such that dim(T̂x X ∩ E) ≥ s + 1}.
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
366 12. Projective geometry II
Note that Zn,n (X) = γ(X) and Zn+a−1,n = X ∨ , and for r ≥ a we have
Zr,0 (X) = G(r + 1, V ). We will see below that Za−1,0 (X) is always a hy-
persurface, and we expect Zs+a−1,s (X) to be a hypersurface. Za−1,0 (X) is
called the associated hypersurface to X in [73] and the varieties Zs+a−1,s (X)
are called higher associated hypersurfaces, although they are not always
hypersurfaces—see below. Let δs (X) = codim Zs+a−1,s (X) − 1 and recall
the notation δ∗ (X) = δn (X).
Proposition 12.9.1. δs (X) = max{0, s + δ∗ (X) − n}.
Exercise 12.9.5: Prove the proposition by imitating our proof for Gauss
images. Namely, take the natural variety X formed from Z and show that
Z is an associated variety to X.
μ
where W = {es }. But rsik = 0 implies ωsj = 0, as for each j there exists a μ
μ α
with qjα ω0 = 0. We now have
Discussion of the generic case. We now consider the generic case, where all
n+1 n+1
the eigenvalues of (r1jk ) are distinct. Write r1jk = λj δjk and consider
Proof. Our calculation above shows that all quadrics in |IIX,x | can be si-
multaneously diagonalized; thus the dimension of the system of quadrics is
at most n − f . A short calculation shows that the only way the prolongation
of such a system can be nonempty is for there to be a quadric of rank one
in the system.
IΞ = {(y, H) | T̃y X ⊂ H, H ∈ Ξy }.
S
/ w
S
X Z
Observe that
dim πΞ (ρΞ −1 (H)) = dim ρΞ −1 (H),
dim IΞ = dim X + dim Ξx ,
dim Z = rank dρΞ,(x,H) = dim IΞ − dim ρΞ −1 (H).
Putting these together, we conclude that
rank dρΞ,(x,H) = dim X + dim Ξx − dim πΞ ρΞ −1 (H).
We now refine the rank restriction theorem in the case σ(X) = Pn+a .
For simplicity, assume σ(X) is a hypersurface. (See [123] for the general
case.) In this case we may smoothly project X to Pn+a−1 from a point
not on σ(X) to obtain a variety of codimension a − 1, which gives us an
immediate improvement of the rank restriction theorem, replacing a − 1 by
a − 2 in the inequality (12.16).
We get a further improvement as follows: take Ξx to be the set of h ∈
Nx∗ X such that (q h )sing contains a generic vector. By our assumptions,
Ξx ⊂ PNx∗ X is a hypersurface.
Exercise 12.11.5: Show that Ξx is indeed a hypersurface.
We obtain
r ≥ n − dim{ωn+λ
n+a
mod ωαn+a },
so at least r ≥ n − (a − 1). In §12.12 we will define a subbundle of F 1 on
which dim{ωn+λn+a
mod ωαn+a } ≤ a − 2, which will prove
Theorem 12.11.6 (Rank restriction [123]). Let X n ⊂ Pn+a be a smooth
variety with degenerate tangential variety that is a hypersurface. Let x ∈ X
be a general point and let
Ξx = P{h ∈ Nx∗ X | II(h)sing contains a II-generic vector}.
Let r be the rank of a generic quadric in Ξx . Then r ≥ n − a + 2.
Proof. It is clear that if Y = γ(X), then (i) and (ii) are satisfied, as then
F (E) is just a general fiber of γ.
Now assume (i) and (ii) are satisfied and let X = E∈Y F (E). Since
X is the transform of the variety Y under an incidence correspondence (the
image under an algebraic map of the inverse image of an algebraic map), X
is a variety. Let 1 ≤ s, t, u ≤ n − r, n − r + 1 ≤ j, k, l ≤ n, and adapt frames
over each x = [e0 ] ∈ Xgeneral such that F = P{e0 , es }, E = {e0 , es , ej }.
The set of pairs {(F (E), PE) | E ∈ Y } gives a subvariety of the flag
variety Fdim F (E),n (V ), and in particular, dF (E) ⊆ E by Exercise 4.2.9.2.
Expressed in frames,
de0 ≡ 0 mod{e0 , es , ej }.
Thus T̂x X ⊆ E. Condition ii implies dim T̂x X = dim E, so we have equality.
Since we are at a general point and X is a variety, we must have equality at
all smooth points of X, and therefore Y = γ(X).
We have
n = dim(ker IIv ) + rank(Ann(v)) + 1 + S = δτ + r + 1 + S,
where S = dim Singloc(Ann(v))−(1+dim ker IIv ). Combined with the rank
restriction Theorem 12.11.6, which gives r ≥ n − a + 2, we have
n S
+2+ . a≥
2 2
This implies a ≥ n
2 + 2, proving Zak’s theorem 12.5.1 on linear normality.
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
374 12. Projective geometry II
which imply
qkn+j + qjn+k = 0 ∀j, k, .
Phrased invariantly, we have shown:
Lemma 12.12.1. Let II ∈ S 2 T ∗ ⊗N be a critical tangential defect. Fix a
generic v ∈ T to obtain a quadratic form on T / Singloc(Ann(v)) which is
well-defined up to scale, and fix its scale. Then the mapping
ker IIv → End(T / Singloc(Ann(v)))
takes image in so(T / Singloc(Ann(v))).
The quadratic form of the lemma is q n+a |T / Singloc(Ann(v)) .
Observe that q n+1 |ker IIv is well-defined because
ker IIv ⊆ Baseloc{q n+2 , . . . , q n+a }.
Consider
n+j β n+j
r11β ω0 = −ωn+1 + 2ω1j ,
r1n+j β j n+j k
β ω0 = ω + q k ω1 ,
which imply
n+j
ωn+1 ≡ 2ω1j mod{ω0α },
ω j ≡ −q n+j
k ω1 mod{ω0 }.
k α
Computing
rn+j β n+1 n+j n+j k n+j k
δβ ω0 = −q δ ωn+1 + q k ωδ + qδk ω ,
and mod-ing out by the semi-basic forms and using (12.12), we obtain
(12.22) q n+j n+k
k qδi
n+j n+k
+ qδk q i = −2q n+1 i
δ δj ∀, δ, j, k, i.
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
12.12. Smooth varieties with degenerate tangential varieties 375
We now outline the proof that the second fundamental form must be
|II| = {uu, vv, uv}, i.e., the fundamental form of a Severi variety.
frame, it holds in any choice of frame (with different constants aμνγ ), so the
expression (12.24) has intrinsic meaning. If (12.24) holds, then
μ α β
ker FF3v2 (X) = {xμ x0 − qαβ x x − aμνβ xν xβ , xμ xν }.
Continuing in the same fashion, we uncover the following conditions:
F3μ = 3aμνγ ω0γ II ν ,
(12.25) F4μ = 4aμνα ω0α F3ν + 3bμντ II ν II τ ,
F5μ = 5aμνγ ω0γ F4ν + 10bμντ F3ν II τ ,
where aμνα , bμντ = bμτν ∈ C. Moreover, if there are no linear syzygies among
the quadrics in |II|, as explained in §12.7, then F6 = 0; thus, Nx∗ is spanned
by the differentials of quadrics, and these quadrics are smooth along X, so
they generate I(X). In this case, we will call (12.25) the generalized Monge
system for quadrics.
In summary:
Theorem 12.13.3 ([124]). Let X ⊂ PV = Pn+a be a variety and x ∈ X a
general point. Assume IIIXx = 0 and that there are no linear syzygies in
|II|x . Then
a+1
dim{quadrics osculating to order three at x} ≤ a + − 1,
(12.26) 2
dim{quadrics osculating to order four at x} ≤ a − 1.
If the generalized Monge system (12.25) holds, then
I2 (X) = ker FF4v2 (X)x .
Equality occurs in the first (respectively second) line of (12.26) if and only
if the first (resp. second) line of (12.25) holds at x, in which case I(X) is
generated by I2 (X). If the generalized Monge system does not hold, then
I(X) is not generated by quadrics.
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
378 12. Projective geometry II
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
https://doi.org/10.1090//gsm/175/13
Appendix A
In this appendix we briefly discuss tensors (§A.1), matrix Lie groups (§A.2)
and Lie algebras (§A.4), complex vector spaces and complex structures
(§A.3), the octonions and the exceptional group G2 (§A.5), Clifford alge-
bras and spin groups (§A.7), and outline some rudiments of representation
theory (§A.6).
Unless otherwise noted, V, W are real vector spaces of dimensions n
and m, with bases v1 , . . . , vn and w1 , . . . , wm . We use the index ranges
1 ≤ i, j ≤ n and 1 ≤ s, t ≤ m.
Exercises A.1.2:
1. Let αi ∈ V ∗ be defined by αi (vj ) = δji . Show that α1 , . . . , αn is a basis
for V ∗ , called the dual basis to v1 , . . . , vn . In particular, dim V ∗ = n.
2. Define, in a coordinate-free way, an isomorphism V → (V ∗ )∗ . (Note that
these spaces would not necessarily be isomorphic if V were replaced by an
infinite-dimensional vector space.)
381
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
382 A. Linear Algebra and Representation Theory
Geometric Aside. Given V , one can form the associated projective space
PV , which is the space of all lines thorough the origin in V . A nonzero
vector α ∈ V ∗ determines a codimension-one linear subspace ker α ⊂ V , so
that PV ∗ may be interpreted as the space of hyperplanes through the origin
in V .
Definition A.1.3. Let Hom(V, W ) denote the space of linear maps from V
to W . Like V ∗ , this is a vector space under the addition of maps. The space
of linear maps from V to V , called endomorphisms, is denoted by End(V ).
We may think of Hom(V, W ) as the space of W -valued linear functions
on V , and when doing so we denote it by V ∗ ⊗ W , the tensor product of
V ∗ and W . (Taking the tensor product with W = R is trivial because
V ∗ ⊗ R = Hom(V, R) = V ∗ .) Similarly, we define V ⊗ W as the space of
W -valued linear functions on V ∗ .
and
One may induce inner products on all tensor spaces constructed from V
and V ∗ . For example, the inner product on V ⊗V is given on decomposable
vectors by a⊗b, c ⊗d = a, cb, d and extended bilinearly. However, since
our definition (A.3) for the wedge product produces k! terms, we normalize
the inner product on Λk V by dividing the inner product inherited from
V ⊗k by k!. With this normalization, the inner product satisfies Hadamard’s
inequality
(A.5) |a ∧ b| ≤ |a||b|.
When V = Rn with the standard inner product, then we write SO(n) for
SO(V ). When V = Rn or Cn , we respectively write SL(n, R) or SL(n, C)
(or SLn when we wish to remain ambiguous) for SL(V ).
When n = 2m, a 2-form ω ∈ Λ2 V ∗ is nondegenerate if ω ∧ · · · ∧ ω = 0,
where ω is wedged with itself m times (or, equivalently, if the map v → v ω
is an isomorphism from V to V ∗ ). In this case, ω is called a symplectic form
on V , and we may define the symplectic group
Sp(V, ω) := {g ∈ V ⊗V ∗ | ω(v ∧ w) = ω(gv ∧ gw) ∀v, w ∈ V }.
Since all nondegenerate 2-forms in Λ2 (R2m ) are linearly equivalent, we often
denote this group by Sp(m, R) (see Exercise A.4.9.8).
Exercises A.2.4:
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
388 A. Linear Algebra and Representation Theory
1. Show that all of the above are matrix Lie groups, i.e., that they are
smooth submanifolds of GL(V ), and they are closed under multiplication
and inverses.
2. Show that if we write the matrices of O(V ) with respect to an orthonor-
mal basis, then
O(V ) ∼
= O(n) = {A ∈ Mn×n | tAA = Id}.
CO(V, Q) ∼
= O(V, Q) × {λ Id | λ > 0}.
Exercises A.3.3:
1. Show that J preserves Q if and only if J ∈ so(Q), so that J ∈ SO(Q) ∩
so(Q).
2. Show that ω as defined above is also preserved by J.
One can make any vector space into a Lie algebra by using a bracket
which is identically zero. Such Lie algebras are called abelian Lie algebras.
Inside a Lie algebra, any subspace that is closed under the bracket is called
a subalgebra. It is a Lie algebra in its own right, since the Jacobi identity
holds by restriction.
A representation of a Lie algebra g is a linear transformation ρ : g →
gl(V ) which preserves the brackets. We say such a V is a g-module. Define
a representation ad : g → End(g) by ad(X)Y := [X, Y ], called the adjoint
representation of g. In particular, any finite-dimensional Lie algebra may
be realized as a matrix Lie algebra.
Exercise A.4.4: Show that ad is indeed a representation.
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
A.4. Lie algebras 391
End(V ) composed with itself j times. The image is a Lie subgroup of GL(V )
(see, e.g., [174, Chap. 10] or [65, §8.3]). When V = g we drop V from the
notation, and denote the corresponding group G, the adjoint form of the Lie
group associated to g. Different representations give rise to different groups,
but all the groups will have the same universal covering group (see, e.g.
[65, §7.3, §23.1]). One can also go in the other direction, as the following
example shows:
called the adjoint representation. For matrix Lie groups, Ad(g)X = gXg −1 .
Given any representation of a Lie group G, one obtains a representation
of the corresponding Lie algebra g by differentiation. In particular Ad : G →
GL(g) gives rise to a representation g → End(g).
Exercise A.4.6: Let g(t) ∈ G be a curve with g(0) = Id and g (0) = Y ∈
d
g. Show that dt t=0
Ad(g(t))X = [Y, X]. In particular the representation
obtained by differentiating Ad is ad.
Remark A.4.7. Where does the Jacobi identity (A.7) come from? One can
interpret it as a Leibniz rule. For, if A is an algebra with product indicated
by a dot, then a map D : A → A is a derivation if D(a·b) = D(a)·b+a·D(b)
for all a, b ∈ A. Now take A = g with the product given by the bracket and
D = DX by the bracket with X ∈ g.
where Idk denotes the k × k identity matrix. We write SO(p, q) and so(p, q)
for the corresponding orthogonal groups and Lie algebras.
Show that if Q has signature (n, 0), and we choose bases such that
Q = Idn , then so(V, Q) ∼
= {X | X + X t = 0}.
6. If Q has signature (3, 1), we could take bases such that
⎛ ⎞ ⎛ ⎞
−1 0 0 0 1
⎜ 1 ⎟ ⎜0 1 0 0⎟
Q=⎜ ⎝
⎟
⎠ or Q=⎜ ⎝0 0 1 0⎠ .
⎟
1
1 1 0 0 0
If we use the first basis, we have
⎧⎛ ⎞ ⎫
⎪
⎪ 0 x21 x31 x41 ⎪
⎪
⎨⎜ 2 3 −x4 ⎟ ⎬
⎜ x 0 −x ⎟
so(3, 1) ⎝ 3 1 2 2 | x 2 3
, x , ... ∈ R .
⎪
⎪ x1 x32 0 −x43 ⎠ 1 1 ⎪
⎪
⎩ ⎭
x41 x42 x43 0
Find the matrix presentation of so(3, 1) using the second choice for Q.
7. Similarly, given a complex structure on R4 , two choices of bases yield
⎛ ⎞ ⎛ ⎞
0 −1 0 0 0 0 −1 0
⎜1 0 0 0 ⎟ ⎜0 0 0 −1⎟
J =⎜ ⎝0 0 0 −1⎠
⎟ and J = ⎜
⎝1 0 0
⎟.
0⎠
0 0 1 0 0 1 0 0
Find the matrix presentation of gl(2, C) ⊂ gl(4, R) using these two choices
for J.
8. Two choices of symplectic form on V R2n are
ω = dx1 ∧ dxn+1 + · · · + dxn ∧ dx2n ,
ω = dx1 ∧ dx2 + · · · + dx2n−1 ∧ dx2n .
Example A.4.10. Let su(2) denote the Lie algebra of SU (2). Taking
e1 , e2 ∈ C2 as a unitary basis gives
1 ix −y + iz
su(2) =
2 y + iz −ix x, y, z ∈ R ⊂ gl(2, C).
√
Let S 4 (C2 ) have basis v1 = e41 , v2 = 2e31 e2 , v3 = 6e21 e22 , v4 = 2e1 e32 , v5 =
e42 . Then the induced representation of gl(2, C) on S 4 (C2 ) restricts to give
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
394 A. Linear Algebra and Representation Theory
Over the reals, two types of signature are possible, one of which is def-
inite. Say we are in the case where the signature is definite; then, by com-
paring homogeneity, one can see that Q is in fact well-defined. A little more
calculation (see [91, Thm. 6.80]) shows that one obtains a new simple Lie
group of dimension 14, which is named G2 .
e1 e2
e1 e2 e3 e3
Figure 1. Products are positive if one multiplies with the arrow (e.g.,
e1 e2 = e3 ), negative against (e.g. e3 e2 = −e1 )
e3
e2 e4
e7
e1 e5
e6
is skew-symmetric in a, b, c.
4. Calculate φ in terms of coordinates x1 , . . . , x7 on R7 = ImO.
5. Show that the Lie algebra g2 = der(φ) has a matrix presentation
κ −tA
(A.13) g2 = κ, τ ∈ so(3) ,
A ρ(κ ⊕ τ )
where ρ is the isomorphism so(3) ⊕ so(3) ∼ = so(4) and a13 = a31 + a42 ,
a23 = a41 − a32 , a33 = a22 − a11 , and a43 = −a21 − a12 .
6. Determine the classical group(s) preserving a generic φ ∈ Λ3 R6 . (Over
C there is a unique generic 3-form up to GL(6, C)-equivalence, but over R
there can be several.)
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
A.6. A smidgen of representation theory 397
Example A.6.2. Consider the rational normal curve Cd ⊂ Pd = P(S d (C2 )),
which is the image of CP1 under [x, y] → [xd , xd−1 y, xd−2 y 2 , . . . , y d ], or,
without coordinates, [v] → [v d ] for v ∈ C2 \{0} (see Chapter 4). What are
the linear changes of coordinates in Cd+1 that leave Cd invariant? The
set of all such changes forms a subgroup of GL(d + 1, C) which is iso-
morphic to GL(2, C). The action may be seen explicitly by g.[v1 · · · vd ] =
[(gv1 ) · · · (gvd )].
semi-simple and an abelian Lie algebra. (The Killing form of a Lie algebra
is nondegenerate if and only if it is semi-simple; see, e.g., [98, §5.1].)
We discuss irreducible representations of simple Lie algebras. The irre-
ducible representations of a semi-simple Lie algebra are just the tensor prod-
ucts of the irreducible representations of its simple components. Thanks to
the action of g on itself, without loss of generality, we may assume that g is
a matrix Lie algebra, i.e., g ⊆ gl(V ).
Simple Lie algebras are best studied via a certain kind of abelian sub-
algebra they contain. We decompose a given g-module V first with respect
to commuting matrices A1 , . . . , Ar that span such a subalgebra.
Exercise A.6.4: Let A1 , . . . , Ar be n × n matrices that commute. Show
that if each Aj is diagonalizable, then A1 , . . . , Ar are simultaneously diago-
nalizable.
Now, back to our simple Lie algebra g. There always exists a maximal
simultaneously diagonalizable (and thus abelian) subalgebra t ⊂ g, unique
up to conjugation by G, called a maximal torus. We define the rank of g to be
the dimension of a maximal torus. Amazingly, irreducible representations
of g are completely determined up to equivalence by how t acts. More
precisely, suppose V is an irreducible g-module. Then as a t-module V
admits a weight space decomposition. If two irreducible g modules V, W
have the same weights (with the same multiplicities) as t-modules, then
they are isomorphic as g-modules.
Let g act on itself by the adjoint action. Then as a t-module, we have
the weight space decomposition
4
g= gα ,
α∈R
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
A.6. A smidgen of representation theory 399
where R ⊂ t∗ is some finite subset whose nonzero members are called the
roots of g, and gα is the eigenspace associated to each root. (Of course,
g0 = t since t is maximal.) Another amazing fact is that the eigenspaces gα
for α = 0 are all one-dimensional.
Remark A.6.6 (Why the word “root”?). Consider the adjoint representa-
tion ad : g → End(g) restricted to t. The roots of the characteristic polyno-
mial pλ (X) = det(ad(X) − λ Idg ) are the eigenvalues of X. By varying X
one obtains linear functions on t which are the roots of g.
Exercise A.6.7: In fact, weights were originally called “generalized roots”
when the theory was being developed. They too are roots of a characteristic
polynomial—which one?
−xn 0
Exercise A.6.10: Determine the roots and the root space decomposition
for so(2n).
Remark A.6.11. The classification of complex simple Lie algebras g (due
to Killing and Cartan) is based on classifying the possible root systems, the
collection of roots for g. The rules for such systems are rather strict and
concise. For the record, we need a subset R ⊂ t∗ such that
(1) R must span t∗ , and moreover if g∗ is complex, R lies in an R-linear
(in fact Q-linear) subspace;
(2) for each α ∈ R, reflection in the hyperplane perpendicular to α
must map R to R;
(3) for all α, β ∈ R, the quantity 2β, α/α, α must be an integer;
and
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
400 A. Linear Algebra and Representation Theory
L
v2 v
v1
refl (v)
L
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
https://doi.org/10.1090//gsm/175/14
Appendix B
Differential Forms
Exercises B.1.2:
1. Show that Tx∗ M is a vector space. Note that if x1 , . . . , xn are local coor-
dinates on M , then the dxi |x provide a basis of Tx∗ M , and thus dim Tx∗ M =
dim M .
2. Show that the cotangent bundle T ∗ M = x∈M Tx∗ M is a vector bundle
over M . Thus, given f : M → R, we obtain a section df ∈ Γ(T ∗ M ).
3. We define Tx M , the tangent space of M at x, as the space of equivalence
classes of differentiable mappings f : R → M with f (0) = x, where f ∼ g
if their first derivatives agree in some coordinate system near x. Show that
if f ∼ g, then their derivatives agree in every such coordinate system. The
vector bundle T M = x∈M Tx M is called the tangent bundle, and sections
X ∈ Γ(T M ) are called vector fields.
4. Show that Tx M and Tx∗ M are naturally dual vector spaces.
405
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
406 B. Differential Forms
Exercises B.2.2:
1. Show that, according to the second definition, dα is C ∞ (M )-linear in
each Xi . Hence dα(X1 , . . . , Xk+1 )|x depends only on the values of the Xi at
x.
2. Show that the two definitions of d agree.
3. Using the second definition, show that φ∗ dα = d(φ∗ α).
4. Show that d2 = 0 (i.e., d(dα) = 0) under both definitions. In particular,
d2 = 0 is equivalent to the fact that mixed partials commute. What property
of vector fields is it equivalent to?
We extend the Lie derivative to Γ((T ∗ M )⊗k ⊗(T M )⊗l ) by the Leibniz rule.
For example, for α, β ∈ Ω1 (M ),
(B.3) LX (α ⊗ β) = (LX α) ⊗ β + α ⊗ LX β.
Similarly, LX (α ∧ β) = (LX α) ∧ β + α ∧ LX β.
Exercises B.2.5:
1. Show that LX Y = [X, Y ].
2. Show that
(B.4) LX α = X dα + d(X α),
where the interior product is as defined in §A.1.
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
https://doi.org/10.1090//gsm/175/15
Appendix C
Complex Structures
and Complex
Manifolds
411
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
412 C. Complex Structures and Complex Manifolds
Do these two ideas agree? If not, which is the “better” one? The second
idea is a local definition, while the first is an infinitesimal one. It’s clear
that the second idea implies the first, since one can make identifications
using the pushforward φ∗ : Tx M → Tφ(x) Cn Cn , and the requirement
that composition maps on the overlaps are holomorphic ensures that the
identifications are well-defined up to multiplication by a complex matrix. So
the question is whether or not the infinitesimal condition is strong enough
to imply the local one.
To answer the question, we first need to describe the infinitesimal idea
more precisely. Recall from Appendix A that a complex structure on a real
vector space V is a map J ∈ End(V ) with J ◦ J = − Id. Recall also that,
when we extend J to VC , that space splits as a sum of +i and −i eigenspaces
of J, denoted by V (1,0) and V (0,1) respectively, and that J is uniquely deter-
mined by this splitting. So, specifying a complex structure on V is equivalent
to specifying the n-dimensional complex subspace V (1,0) ⊂ VC (assuming its
intersection with its complex conjugate, which will be V (0,1) , is zero). In
∗(1,0) ∗(0,1)
turn, this is equivalent to specifying a dual splitting VC∗ = VC ⊕ VC ,
∗(1,0)
where the forms in VC vanish on the −i-eigenspace and vice versa.
Definition C.1.1. An almost complex structure on a real 2n-dimensional
manifold M is an n-dimensional smooth distribution T (1,0) M ⊂ TC M . It
determines a unique compatible J ∈ Γ(End(T M )) having T (1,0) M as its
+i-eigenspace. A manifold equipped with an almost complex structure will
be called an almost complex manifold.
A complex structure on M is an atlas of charts φi : Ui → Cn such that the
compositions φj ◦ φ−1
i , where defined, are invertible holomorphic functions
from domains in Cn to domains in Cn . A complex manifold is a real manifold
equipped with a complex structure.
Our question above may be rephrased as, are all almost complex mani-
folds complex? We will see below that if dim M = 2, the answer is yes.
Proposition C.1.2. An almost complex structure is complex if and only
if T ∗(1,0) M ⊂ Ω1C (M ) is a Frobenius ideal (equivalently, if T (1,0) M is closed
under the Lie bracket). When this is the case, the almost complex structure
is said to be integrable.
Here is the idea of the proof:
∗(1,0)
At a point p one has a subspace Tp ⊂ TC∗ M with basis, say,
αp1 , . . . , αpn . We want to find local coordinates (x1 , . . . , xn , y 1 , . . . , y n ), such
that if we define z j = xj + iy j , then dz j |p = αpj (which we can always do)
and moreover we can extend the αpj ’s to sections αj ∈ Γ(T ∗(1,0) M ) such
that in some neighborhood of p we have dz j = αj . In the real case, this
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
C.1. Complex manifolds 413
We refer to smooth sections of Λ(p,q) TC∗ M as (p, q)-forms and use the notation
Ω(p,q) (M ) = Γ(Λ(p,q) TC∗ M ).
dα ∈ Ω(p+1,0) (M ),
J(e1 ⊗ v + e2 ⊗ w) = e1 ⊗ w − e2 ⊗ v.
Almost complex manifolds are far more abundant than complex mani-
folds and have been used extensively in symplectic geometry—see [9]. On
the other hand, interesting explicit examples of almost complex structures
that are not complex are rare (just as interesting explicit ‘generic’ Riemann-
ian metrics are rare). Here is one:
Example C.1.11. Let R7 = ImO with its standard flat metric (see Ap-
pendix A for a description of the octonions O). Consider the unit sphere
S 6 ⊂ ImO. Define Ju (v) = uv, where we identify Tu S 6 with the linear
subspace u⊥ ⊂ R7 and uv is octonionic multiplication.
Exercise C.1.12: Show that this defines an almost complex structure on
S 6 . Show that the structure is not integrable.
calculate that
1 1
θ π1 π2 dx
(C.3) d 2 =− ∧ ,
θ −π2 π1 dx2
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
https://doi.org/10.1090//gsm/175/16
Appendix D
Example D.1.1. The wave equation utt − c2 uxx = 0 describes the propa-
gation of a wave in a one-dimensional medium with velocity c, where u(x, t)
419
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
420 D. Initial Value Problems
represents the height of the wave at point x and time t and damping is ne-
glected. Intuitively, since wave motion is the aggregate of the dynamics of
many small particles, we expect the solution to be determined by the initial
position u(x, 0) = h(x) and the velocity ut (x, 0) = v(x). In fact, the general
solution of this wave equation is
and f, g may be determined from h(x) and v(x). So, the solution is defined
for all x and t, and depends on two arbitrary functions of one variable,
specified along the hypersurface t = 0.
Example D.1.2. The heat equation ut = uxx models heat flow in a one-
dimensional medium, with u(x, t) representing temperature at point x and
time t. Intuitively, one expects that the temperature distribution along t = 0
exactly determines the temperature in the domain for all later times, so we
are led to pose the problem of solving the heat equation subject to u = f (x)
along the hypersurface t = 0. In fact, under reasonable assumptions on
f (x), there is a unique solution defined on R × [0, ∞) given by a convolution
formula [61].
be written as
∂ua ∂ub
(D.2) = F a (xi , t, ub , i ),
∂t ∂x
(D.3) u |t=t0 = g (x , . . . , x ),
a a 1 n
∂t ∂x
with the corresponding initial data u = h(x), u2 = h (x) and u3 = v(x).
1
Exercises D.2.2:
1. Give an explicit expression for the solution of the wave equation in terms
of initial data h(x) and v(x). Conclude that real-analyticity is not necessary.
2. Show that the initial value problem of Example D.1.2 cannot be ex-
pressed in Cauchy-Kowalevski form. (However, the heat equation by itself
can be put in the form (D.2) if we exchange x and t.)
Theorem D.2.3 (Cauchy-Kowalevski). Suppose that the system (D.2) is
quasilinear, i.e.,
∂ua ∂ub ∂ub
= F a (xi , t, ub , i ) = Aabi (x, t, u) i + B a (x, t, u)
∂t ∂x ∂x
b,i
and that the functions Aabi and B a are real-analytic on an open set U ⊂
Rn+1 × Rs . Suppose there is an open set V ⊂ Rn on which the initial data
(D.3) is real-analytic, and such that the graph of g a |V lies in the intersection
of U with the hyperplane t = t0 . Then for any point (x10 , . . . , xn0 ) ∈ V , there
is an open set W ⊂ Rn+1 containing (x10 , . . . , xn0 , t0 ) and a solution U a (x, t)
of the initial value problem which is real-analytic on W . Moreover, if W
is an open subset of W with a non-empty intersection with V , then any
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
422 D. Initial Value Problems
solution of the initial value problem defined on W must coincide with the
solution ua = U a (x, t).
Remark D.2.4. The restriction of quasilinearity is not essential, since dif-
ferentiating (D.2) with respect to the xi produces equations which are linear
in the highest-order derivatives. Then adding more variables leads us to a
quasilinear system. When generalized to exterior differential systems (EDS),
this shows that the prolongation of any EDS is a linear Pfaffian system (see
Chapter 6).
Exercise D.2.5: Add more variables to the nonlinear first-order equation
u∂u/∂y = (∂u/∂x)2 to obtain a quasilinear system.
where the csjl are linear combinations of powers of the cjl . Then (D.4) implies
that
lcjl xj tl−1 = iajs cskl cim xi+j+k−1 tl+m + bjs cskl xj+k tl .
j,l i,j,k,l,m,s j,k,l,s
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
D.3. Generalizations 423
In order for this equality to hold, it must hold for each power of t.
Equating the coefficients of the t0 terms on each side, we get
cj1 xj = iajs csk0 ci0 xi+j+k−1 + bjs csk0 xj+k .
i,j,k,s j,k,s
Observe that all terms on the right hand side are known, so if we now equate
coefficients of like powers of x, we can solve for the cj1 ’s. When we equate the
coefficients of the t1 terms, the right hand side will only involve previously
known quantities and the cj1 ’s, while the left hand side will involve the cj2 ’s.
Continuing, we can solve for all the cjl ’s.
We conclude that if there is a solution, it must be unique and given by
this procedure. To show that a solution in fact exists, it remains to show
this series converges in some neighborhood of the origin. For a proof of
Theorem D.2.3 using the method of majorants, see [105]; a special case is
given in Chapter 5.
D.3. Generalizations
When trying to discover solution formulas for PDE like the d’Alembert
formula (D.1), or when studying PDE systems that arise in geometry, we will
rewrite the differential equations as exterior differential systems—in other
words, replacing coordinates by moving frames, and equations by differential
ideals. By appropriately generalizing the notion of characteristic to these
systems, we can detect and construct explicit solution formulas for PDE
when they exist (see Chapter 7). As well, the Cauchy-Kowalevski Theorem
has a generalization which constructs solutions to initial value problems
for EDS. This theorem, called the Cartan-Kähler Theorem, is discussed in
Chapter 6 and Chapter 8.
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
Hints and Answers to
Selected Exercises
Chapter 1
1.2.4 Use Stokes’ Theorem applied to integrating about the boundary
of a rectangle.
1.4.3(f) Let A(t) be a family of circles with radius r(t) and centers c(t),
dt > | dt |, then the circles are nested.
and show that if dr dc
1.7.3.1(d) Consider the area of the region bounded by the curve and a line
parallel to the curve, and take the limit as the line approaches
the tangent line (see [83]).
1.8.4.3 For (b): Write c(t) = c(t)+r(t)e2 (t), and show that r is constant.
1.8.4.5 Consider the point c(t) + ρN + σB.
1.9.1.2 Use the skew-symmetry of the wedge product, defined in Appen-
dix A.
1.10.7 For part (c), n + m kj=0 n+j−1
j .
Chapter 2
2.1.2.2 Compare the speed of a curve to its curvature.
2.1.2.3 What is the characteristic polynomial of h?
2.2.2.5 Take a line in the x − z plane and rotate it about the z-axis.
v
2.2.2.6 Let r(v) = C cos(v), t(v) = 0 1 − C 2 sin2 v dv. The only com-
plete ones are the spheres.
425
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
426 Hints and Answers to Selected Exercises
Chapter 3
3.1.1 For (b) differentiate the result of part (a). For (d) use the Cartan
Lemma A.1.9.
3.1.3 The fibers are isomorphic to SO(V ) without the identity labeled.
If we consider the point we are taking the differential at as the
identity, then that fiber becomes SO(V ) and thus its tangent
space at that point becomes so(V ).
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
Hints and Answers to Selected Exercises 427
3.1.8 Show that Θ̃ is unchanged under the action of O(n) on the fiber
of Fon (M ).
3.1.16 dβ is minus the skew-symmetrization of ∇β.
3.1.19 For arbitrary α = ai η i , we have ∇(α(X)) = d(ai xi ) = ai dxi +
xi dai and X ∇α = xi (dai − aj ηij ) from (3.3).
3.1.21.3 Use dα(X, Y ) = X(α(Y )) − Y (α(X)) − α(X, Y ) for α = η i , and
combine with (3.4).
3.1.21.4 Differentiating (3.4), with Z in place of X, gives a relationship
between Θ and the second derivatives of Z.
3.2.3 To do this exercise, you need to use the higher-dimensional ver-
sion of the Poincaré-Hopf Theorem; see p. 450 in Volume 1 of
[173] for further details.
3.2.7 On one hand ∇h = ω i ω j ⊗(dhij − hik ωjk − hjk ωik ) by the defini-
tion of the covariant derivative. On the other hand taking 0 =
d(ωjn+1 − hjk ω k shows (dhij − hik ωjk − hjk ωik ) ∧ ω j = 0. Now ap-
ply the Cartan Lemma to conclude (dhij − hik ωjk − hjk ωik ) = hijk
with hijk = hikj , but we also have hijk = hjik .
3.2.10 The differentials of η i − ω i are clearly zero modulo I, those
of ηji − ωji are zero thanks to the Gauss equation, and that of
ω n+1 , ωjn+1 − hjk ω k is zero thanks to the Codazzi equation.
Chapter 4
4.2.2 Use the fundamental theorem of algebra.
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
428 Hints and Answers to Selected Exercises
Chapter 5
5.1.11.1 A = {p111 w1 ⊗v 1 ◦ v 1 + p112 w1 ⊗v 1 ◦ v 2 + p122 w1 ⊗v 2 ◦ v 2 + p211 w2 ⊗
v 1 ◦ v 1 + p212 w2 ⊗ v 1 ◦ v 2 + p222 w2 ⊗ v 2 ◦ v 2 | p111 + p222 = 0} has
dimension 5 = 6 − 1.
Chapter 7
7.1.2.1 Use the fact that L[v,w] = Lv ◦ Lw − Lw ◦ Lv . This identity is
easily verified on 0-forms and exact 1-forms, and then follows for
all forms from the Leibniz rule (B.3) for Lie derivatives.
7.1.2.2 Suppose Lv ψ ∈ I for ψ ∈ I. Then use the Leibniz rule to show
that Lv (α ∧ ψ) ∈ I for any α, and use (B.4) to get Lv (dψ) ∈ I.
7.1.2.3 The most general v is given by (7.2) for f = −hz and g = h+zhz ,
where h is an arbitrary function of x, y, z.
7.1.3 Note that if ψ ∈ I, then φ∗t (ψ) ∈ I also.
7.1.15.2 u(x, y) = 12 (1 − x2 ) ± y.
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
Hints and Answers to Selected Exercises 429
√
7.1.15.3 u(x, y) = 2x3/2 / 1 − 27y. In general, the solution of u(s, 0) =
f (s) is u = f (s) − 2yf (s)3 , where s is implicitly defined as a
function of x, y by s = x + 3yf (s)2 . If f (s) isn’t linear, then
there exists an s such that f (s) = 0, and hence there exists a y
such that ∂s/∂x is undefined. Then ∂ 2 u/∂x2 is undefined.
7.1.19.1 Use the Pfaff Theorem 1.10.16.
7.4.15 For part (a), note that ω 3 is the result of taking the dot product
of dx with e3 . For part (b), note that Fr is covered by the
transformation on F given by (x; e1 , e2 , e3 ) → (x+re3 ; e1 , e2 , e3 ).
Then use (1.26) and (1.27) to compute pullbacks of the ω’s.
Chapter 8
8.1.5 Consider the bundle map i : T Σ → T ∗ Σ given by v → v (ω 1 ∧ω 2 ),
which is well-defined up to multiple. Then the singular points of
V2 (I) are those 2-planes to which this map restricts to be zero.
8.1.7.3 If ψ ∈ I i for k ≤ n, then v ψ|E = 0 implies that there exists
ψ ∧ α ∈ I n+1 such that v ψ|E = 0.
8.2.1.1 Express the right-hand side of (8.6) in terms of the Ωi and solve
for the Ωi in terms of the dω i ∧ ω i ’s.
8.2.1.3 Consider the intersection of a hyperboloid and an ellipsoid in a
system of confocal quadrics.
8.5.2 Let q = κeiθ and differentiate N = e2 sin θ + e3 cos θ.
8.5.9.1 In fact, d(x + (1/k0 )e3 ) ∧ ω 1 = 0, so each line of curvature in the
e2 -direction lies on a sphere of radius 1/k0 . The surface is the
envelope of these spheres, and the two functions may be taken
to be the curvature and torsion of the curve traced out by the
centers of the spheres.
8.5.9.2 These surfaces may also be characterized as those for which both
of the focal surfaces are degenerate. One such surface is the torus
swept out by revolving a circle about an axis. The EDS for these
surfaces becomes Frobenius after two prolongations, and the so-
lutions are the cyclides of Dupin. These are the surfaces that
are obtained by acting on circular tori of revolution by the group
O(4, 1) of conformal transformations of E3 . See [104] for more
information, including an interesting application of conformal
moving frames.
8.5.9.4 Add the 3-form ω14 ∧ ω 2 ∧ ω 3 + ω24 ∧ ω 3 ∧ ω 1 + ω34 ∧ ω 1 ∧ ω 3 to the
ideal.
Chapter 9
9.1.3 All are equivalent.
1 ∂2 Hx
9.1.12.5 (log )dx ∧ dy.
2 ∂x∂y Hy
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
432 Hints and Answers to Selected Exercises
F Fxy − Fx Fy
9.1.12.6 dx ∧ dy.
F2
9.7.1 Show that the components of θ span a codimension-one Pfaffian
system on α∗ FG .
9.7.2 Use the equivariance of θ.
Chapter 10
10.2.6 Given α ∈ Γ(V̂ ), let α = α1 + α0 where α1 ∈ Γ(V̌ ) and α0 ∈
Γ(V̂ (∞) ). Conclude that α1 is a section of V̂ ∩ V̌ = I. Thus,
V̂ = {σ̂, η̂, θ, η̌}, and then use T ∗ M = V̂ + V̌ (∞) .
10.2.9 To show, e.g., that V̂ is a differential ideal, note that modulo the
1-forms in V̂ , the Ω̌ are spanned by the derivatives of θ and η̌.
Chapter 11
11.1.6 Under a change of metric
Θ̂ij = [ + σ −2 ( σk2 )]η i ∧ η j + σ −1 (sjm η i ∧ η m − sim η j ∧ η m ).
k
1
11.5.5 Use (11.27) and (11.3), keeping in mind that s = 2(n−1) P.
Chapter 12
12.1.8.1 Use equation (12.5).
12.1.8.3 Consider a cubic form in two variables and differentiate.
12.2.1 Use the prolongation property.
12.2.2.3 TE Gω (k, W ) E ∗ ⊗ (E ⊥ /E) ⊕ S 2 E ∗ . Here E ⊥ = {w ∈ W |
ω(v, w) = 0 ∀v ∈ E}.
12.2.5 Note that Mr is a join.
12.4.3 The cases (a), (b), (e) are always hypersurfaces, (c), (d) are hy-
persurfaces respectively when a = b and m is even. The smooth
quadric surface in P3 , v2 (P1 ), G(2, 5) and P1 × Pn are all self-
dual. The self-duality can be proven directly; or, once one knows
the dimension is correct, since the dual variety of a homogeneous
variety inherits a group action and they are the dimension of the
minimal orbit, they must be the minimal orbit.
12.4.14.1 Given A ∈ Mp×q , consider
0 A
tA 0 .
12.5.5
Since we are working with eμ mod{T̂ , II v (T )}, adapt frames
further such that {eξ } IIv (T ) and {eφ } gives a comple-
ment. You will need to show that if w ∈ Singloc(Ann(v)), then
IIw (T ) ⊆ IIv (T ).
12.5.6.1 Note that III can be recovered from III(v, v, v) for all v ∈ T .
12.6.12.1 Note that II(w1 ∧ w2 , u1 ∧ u2 ) = w1 ∧ w2 ∧ u1 ∧ u2 .
12.8.9 Adapt frames so that II = ω i ω j ⊗eij and calculate F4 , F5 . The
calculation is a little involved; see [127] for the details.
12.9.5 Use the calculation (12.14).
12.11.4 Use Corollary 12.11.2. What is ker II?
12.13.2 If x ∈ X is a general point and IIIX,x = 0, then a hyperplane
osculating to order two at x contains X.
Appendix A
A.1.6.3 Use the preceding exercise.
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
Hints and Answers to Selected Exercises 435
Appendix B
∂bj ∂ i ∂a
j ∂
B.1.3.2 [X, Y ] = ai − b .
∂xi ∂xj ∂xi ∂xj
B.2.1 Let y jbe a second set of coordinates and similarly let α =
(1/k!) J bJ dy j1 ∧ . . . ∧ dy jk . Using the chain rule, show that
dα is the same computed in either coordinate system.
Appendix C
C.1.10(b) Recall from Chapter 4 that the forms ωik for 3 ≤ k ≤ n and
i = 1, 2 are semi-basic for the projection from GL(V ) to G. Show
that the system spanned by the forms ω1k − iω2k is Frobenius, and
drops to G and forms a basis for T ∗(1,0) .
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
Bibliography
437
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
438 Bibliography
16. R. Bishop, There is more than one way to frame a curve, Am. Math. Monthly 82 (1975),
246–251.
17. A. Bobenko, Exploring surfaces through methods from the theory of integrable systems:
Lectures on the Bonnet Problem, preprint (1999), arXiv:math.DG/9909003
18. R. Bott, L. Tu, Differential forms in algebraic topology, Springer, 1982.
19. N. Bourbaki, Groupes et algèbres de Lie, Chap. 4–6, Hermann, 1968.
20. Simon Brendle, Embedded minimal tori in S 3 and the Lawson conjecture, Acta Math.
211 (2013), no. 2, 177–190.
21. Simon Brendle and Fernando C. Marques, Recent progress on the Yamabe problem,
pp. 29–47 in Surveys in geometric analysis and relativity, International Press, 2011.
22. R. Bryant, Conformal and Minimal Immersions of Compact Surfaces into the 4-sphere,
J. Diff. Geom. 17 (1982), 455–473.
23. —, Metrics with exceptional holonomy, Ann. of Math. 126 (1987), 525–576.
24. —, An introduction to Lie groups and symplectic geometry, pp. 5–181 in Geometry and
quantum field theory, Amer. Math. Soc., Providence, 1995.
25. —, Rigidity and quasi-rigidity of extremal cycles in Hermitian symmetric spaces,
Annals of Mathematics Studies vol. 153, Princeton, 2002.
26. R. Bryant, S.-S. Chern, P. Griffiths, Exterior Differential Systems, pp. 219–338 in
Proceedings of the 1980 Beijing Symposium on Differential Geometry and Differential
Equations, Science Press, Beijing, 1982.
27. R. Bryant, S.-S. Chern, R.B. Gardner, H. Goldschmidt, P. Griffiths, Exterior Differ-
ential Systems, MSRI Publications, Springer, 1990.
28. R. Bryant, P. Griffiths, Characteristic Cohomology of Differential Systems (II): Conser-
vation Laws for a Class of Parabolic Equations, Duke Math. J. 78 (1995), 531–676.
29. R. Bryant, P. Griffiths, D. Grossman, Exterior differential systems and Euler-
Lagrange partial differential equations, University of Chicago Press, 2003.
30. R. Bryant, P. Griffiths, L. Hsu, Hyperbolic exterior differential systems and their con-
servation laws (I), Selecta Mathematica (N.S.) 1 (1995), 21–112.
31. —, Toward a geometry of differential equations, pp. 1–76 in Geometry, topology, and
physics, International Press, 1995.
32. R. Bryant, S. Salamon, On the construction of some complete metrics with exceptional
holonomy, Duke Math. J. 58 (1989), 829–850.
33. Andreas Čap and Jan Slovák, Parabolic geometries I: Background and general theory,
American Mathematical Society, 2009.
34. J. Carlson, D. Toledo, Generic Integral Manifolds for weight two period domains, Trans.
AMS 356 (2004), 2241-2249.
35. E. Cartan, Sur la structure des groupes infinis de transformations, Ann. Sci. Ecole
Norm. Sup. 26 (1909), 93–161.
36. —, Les systèmes de Pfaff à cinq variables et les équations aux dérivées partielles du
second ordre, Ann. Ecole Norm. Sup. 27 (1910), 109–192.
37. —, Sur les variétés de courbure constante d’un espace euclidien ou non euclidien, Bull.
Soc. Math France 47 (1919) 125–160 and 48 (1920), 132–208; see also pp. 321–432 in
Oeuvres Complètes Part 3, Gauthier-Villars, 1955.
38. —, Sur les variétés à connexion projective, Bull. Soc. Math. Fr. 52 (1924), 205–241.
39. —, Sur la théorie des systèmes en involution et ses applications à la relativité, Bull. Soc.
Math. Fr. 59 (1931), 88–118; see also pp. 1199–1230 in Oeuvres Complètes, Part 2.
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
Bibliography 439
69. R.B. Gardner, W.F. Shadwick, The GS algorithm for exact linearization, IEEE Trans.
Autom. Control 37 (1992), 224–230.
70. J. Gasqui, Sur la résolubilité locale des équations d’Einstein, Compositio Math. 47
(1982), 43–69.
71. —, Formal integrability of systems of partial differential equations, pp. 21–36 in Non-
linear equations in classical and quantum field theory, Lecture Notes in Phys. 226,
Springer, 1985.
72. I.M. Gel’fand, S. Fomin, Calculus of Variations, Prentice-Hall, 1963.
73. I.M. Gel’fand, M. Kapranov, A. Zelevinsky, Discriminants, resultants, and multidi-
mensional determinants, Birkhäuser, 1994.
74. H. Goldschmidt, Existence theorems for analytic linear partial differential equations,
Ann. of Math. 86 (1967), 246–270.
75. A.B. Goncharov, Generalized Conformal Structures on Manifolds, Selecta Math. Sov. 6
(1987), 307–340.
76. E. Goursat, Leçons sur l’intégration des équations aux dérivées partielles du second
ordre, Gauthier-Villars, 1890.
77. —, Recherches sur quelques équations aux dérivées partielles du second ordre, Annales
de la Faculté de Toulouse, deuxième serie 1 (1899), 31–78.
78. A. Rod Gover, Almost conformally Einstein manifolds and obstructions, pp. 247–260 in
Differential geometry and its applications, Matfyzpress, Prague, 2005.
79. —, Almost Einstein, Poincaré-Einstein manifolds in Riemannian signature, J. Geom.
Phys. 60 (2010), no. 2, 182–204.
80. C. Robin Graham and Kengo Hirachi, The ambient obstruction tensor and Q-curvature,
pp. 59–71 in AdS/CFT correspondence: Einstein metrics and their conformal bound-
aries, Eur. Math. Soc., Zürich, 2005.
81. Mark L. Green, Generic initial ideals, pp. 119–186 in Six lectures on commutative
algebra (J. Elias et al, eds.), Progr. Math. 166, Birkhäuser, 1998.
82. P. Griffiths, Some aspects of exterior differential systems, pp. 151–173 in Complex
geometry and Lie theory, Proc. Sympos. Pure Math. 53 (1991), AMS.
83. —, Exterior Differential Systems and the Calculus of Variations, Birkhäuser, 1983.
84. P. Griffiths, J. Harris, Principles of Algebraic Geometry, Wiley, 1978.
85. —, Algebraic geometry and local differential geometry, Ann. Sci. Ecole Norm. Sup. 12
(1979), 355–432.
86. P. Griffiths, G. Jensen, Differential systems and isometric embeddings, Princeton
University Press, 1987.
87. M. Guest, Harmonic Maps, Loop Groups and Integrable Systems, Cambridge Uni-
versity Press, 1997.
88. V. Guillemin, The integral geometry of line complexes and a theorem of Gel’fand-Graev,
pp. 135–149 in The mathematical heritage of Elie Cartan, Astérisque, Numero Hors
Serie (1985).
89. J. Harris, Algebraic geometry, a first course, Springer, 1995.
90. Robin Hartshorne, Varieties of small codimension in projective space, Bull. Amer. Math.
Soc. 80 (1974), 1017–1032.
91. F.R. Harvey, Spinors and Calibrations, Academic Press, 1990.
92. F.R. Harvey, H.B. Lawson, Calibrated geometries, Acta Math. 148 (1982), 47–157.
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
Bibliography 441
93. T. Hawkins, Emergence of the theory of Lie groups. An essay in the history of math-
ematics 1869–1926, Springer, 2000.
94. S. Helgason, Differential Geometry, Lie Groups and Symmetric Spaces, Academic
Press, 1978.
95. N. Hitchin, Kählerian twistor spaces, Proc. London Math. Soc. (3) 43 (1981), 133–150.
96. —, Complex manifolds and Einstein’s equations, pp. 73–99 in Twistor geometry and
nonlinear systems, Lecture Notes in Math. 970, Springer, 1982.
97. W.V.D. Hodge, D. Pedoe, Methods of Algebraic Geometry Vol. 2, Cambridge Uni-
versity Press, 1954.
98. J. Humphreys, Introduction to Lie algebras and representation theory, Springer, 1972.
99. J.M. Hwang, N. Mok, Uniruled projective manifolds with irreducible reductive G-
structures, J. reine angew. Math. 490 (1997), 55–64.
100. J.M. Hwang, N. Mok, Rigidity of irreduicible Hermitian symmetric spaces of the compact
type under Kahler deformation, Invent. Math. 131 (1998), 393–418.
101. J.M. Hwang, K. Yamaguchi, Characterization of Hermitian symmetric spaces by funda-
mental forms, Duke Math. J. 120 (2003), 621–634.
102. J. Jost, Riemannian geometry and geometric analysis, Springer, 2011.
103. Bo Ilic, J. M. Landsberg, On symmetric degeneracy loci, spaces of symmetric matrices
of constant rank and dual varieties, Math. Ann. 314 (1999), no. 1, 159–174.
104. T. Ivey, Surfaces with orthogonal families of circles, Proc. AMS 123 (1995), 865–872.
105. F. John, Partial Differential Equations (4th ed.), Springer, 1982.
106. D. Joyce, Compact Riemannian 7-manifolds with holonomy G2 , I, II, J. Diff. Geom. 43
(1996), 291–375.
107. D. Joyce, Compact 8-manifolds with holonomy Spin(7), Invent. Math. 123 (1996),
507–552.
108. M. Juráš, I. Anderson, Generalized Laplace invariants and the method of Darboux, Duke
Math. J. 89 (1997), 351–375.
109. E. Kähler, Einfürhung in die Theorie der Systeme von Differentialgleichungen, Teub-
ner, 1934.
110. N. Kamran, An elementary introduction to exterior differential systems, pp. 151–173 in
Geometric approaches to differential equations, Cambridge University Press, 2000.
111. S. Kleiman, Tangency and duality, pp. 163–225 in Proceedings of the 1984 Vancouver
conference in algebraic geometry, CMS Conf. Proc., vol. 6, Amer. Math. Soc., 1986.
112. M. Kline, Mathematical Thought from Ancient to Modern Times, Oxford University
Press, 1972.
113. A. Knapp, Lie groups beyond an introduction, Birkhäuser, 1996.
114. S. Kobayashi, K. Nomizu, Foundations of Differential Geometry, Vols. 1,2, Wiley,
1963/1969.
115. S. Kobayashi, T. Ochiai, Holomorphic structures modeled after compact Hermitian
symmetric spaces, pp. 207–222 in Manifolds and Lie groups, Birkhäuser, 1981.
116. Bertram Kostant, Holonomy and the Lie algebra of infinitesimal motions of a Riemannian
manifold, Trans. Amer. Math. Soc. 80 (1955), 528–542.
117. J.Krasil’shchick, A.M. Vinogradov (eds.), Symmetries and conservation laws for dif-
ferential equations of mathematical physics, American Math. Society, 1999.
118. Wolfgang Kühnel and Hans-Bert Rademacher, Einstein spaces with a conformal group,
Results Math. 56 (2009), no. 1-4, 421–444.
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
442 Bibliography
119. M. Kuranishi, Lectures on exterior differential systems, Tata Institute, Bombay, 1962.
120. —, CR geometry and Cartan geometry, Forum Math. 7 (1995), 147–205.
121. J.M. Landsberg, Minimal submanifolds defined by first-order systems of PDE, J. Diff.
Geom. 36 (1992), 369–415.
122. —, On second fundamental forms of projective varieties, Inventiones Math. 117 (1994),
303–315.
123. —, On degenerate secant and tangential varieties and local differential geometry, Duke
Mathematical Journal 85 (1996), 605–634.
124. —, Differential-geometric characterizations of complete intersections, J. Diff. Geom. 44
(1996), 32–73.
125. —, Introduction to G-structures via three examples, pp. 133–149 in Web theory and
related topics, World Scientific, 2001,
126. —, Exterior differential systems: a geometric approach to pde, pp. 77–101 in Topology
and Geometry, Proc. Workshop Pure Math. vol. 17, Korean Academic Council, 1998.
127. —, On the infinitesimal rigidity of homogeneous varieties, Compositio Math. 118 (1999),
189–201.
128. —, Is a linear space contained in a submanifold? - On the number of derivatives needed
to tell, J. reine angew. Math. 508 (1999), 53–60.
129. —, Algebraic geometry and projective differential geometry, Lecture Notes Series, vol.
45, Seoul National University, Research Institute of Mathematics, Global Analysis
Research Center, Seoul, 1999.
130. —, Lines on projective varieties, J. Reine Angew. Math. 562 (2003), 1–3.
131. —, Griffiths-Harris rigidity of compact Hermitian symmetric spaces, J. Diff. Geom. 74
(2006), 395–405.
132. —, The border rank of the multiplication of 2 × 2 matrices is seven, J. Amer. Math.
Soc. 19 (2006), no. 2, 447–459.
133. —, Tensors: geometry and applications, American Mathematical Society, Providence,
RI, 2012.
134. J.M. Landsberg, L. Manivel, The projective geometry of Freudenthal’s magic square,
J. Algebra 239 (2001), 477–512.
135. —, Construction and classification of complex simple Lie algebras via projective geometry,
Selecta Math. (N.S.) 8 (2002), 137–159.
136. —, On the projective geometry of homogeneous varieties, Commentari Math. Helv. 78
(2003), 65–100.
137. J.M. Landsberg, C. Robles, Fubini-Griffiths-Harris rigidity and Lie algebra cohomology,
Asian J. Math. 16 (2012), 561–586.
138. H.B. Lawson, M. Michelsohn, Spin geometry, Princeton University Press, 1989.
139. R. Lazarsfeld, Positivity in algebraic geometry, Volume I, Springer, 2004.
140. Claude R. LeBrun, Ambi-twistors and Einstein’s equations, Classical and Quantum
Gravity 2 (1985), no. 4, 555–563.
141. Claude LeBrun and Simon Salamon, Strong rigidity of positive quaternion-Kähler man-
ifolds, Invent. Math. 118 (1994), no. 1, 109–132.
142. M. van Leeuwen, LiE, a computer algebra package,
available at http://young.sp2mi.univ-poitiers.fr/∼marc/LiE/.
143. S. L’vovsky, On Landsberg’s criterion for complete intersections, Manuscripta Math. 88
(1995), 185–189.
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
Bibliography 443
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
Index
445
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms
446 Index
175 Thomas A. Ivey and Joseph M. Landsberg, Cartan for Beginners: Differential
Geometry via Moving Frames and Exterior Differential Systems, Second Edition, 2016
174 Alexander Kirillov Jr., Quiver Representations and Quiver Varieties, 2016
173 Lan Wen, Differentiable Dynamical Systems, 2016
172 Jinho Baik, Percy Deift, and Toufic Suidan, Combinatorics and Random Matrix
Theory, 2016
171 Qing Han, Nonlinear Elliptic Equations of the Second Order, 2016
170 Donald Yau, Colored Operads, 2016
169 András Vasy, Partial Differential Equations, 2015
168 Michael Aizenman and Simone Warzel, Random Operators, 2015
167 John C. Neu, Singular Perturbation in the Physical Sciences, 2015
166 Alberto Torchinsky, Problems in Real and Functional Analysis, 2015
165 Joseph J. Rotman, Advanced Modern Algebra: Third Edition, Part 1, 2015
164 Terence Tao, Expansion in Finite Simple Groups of Lie Type, 2015
163 Gérald Tenenbaum, Introduction to Analytic and Probabilistic Number Theory, Third
Edition, 2015
162 Firas Rassoul-Agha and Timo Seppäläinen, A Course on Large Deviations with an
Introduction to Gibbs Measures, 2015
161 Diane Maclagan and Bernd Sturmfels, Introduction to Tropical Geometry, 2015
160 Marius Overholt, A Course in Analytic Number Theory, 2014
159 John R. Faulkner, The Role of Nonassociative Algebra in Projective Geometry, 2014
158 Fritz Colonius and Wolfgang Kliemann, Dynamical Systems and Linear Algebra,
2014
157 Gerald Teschl, Mathematical Methods in Quantum Mechanics: With Applications to
Schrödinger Operators, Second Edition, 2014
156 Markus Haase, Functional Analysis, 2014
155 Emmanuel Kowalski, An Introduction to the Representation Theory of Groups, 2014
154 Wilhelm Schlag, A Course in Complex Analysis and Riemann Surfaces, 2014
153 Terence Tao, Hilbert’s Fifth Problem and Related Topics, 2014
152 Gábor Székelyhidi, An Introduction to Extremal Kähler Metrics, 2014
151 Jennifer Schultens, Introduction to 3-Manifolds, 2014
150 Joe Diestel and Angela Spalsbury, The Joys of Haar Measure, 2013
149 Daniel W. Stroock, Mathematics of Probability, 2013
148 Luis Barreira and Yakov Pesin, Introduction to Smooth Ergodic Theory, 2013
147 Xingzhi Zhan, Matrix Theory, 2013
146 Aaron N. Siegel, Combinatorial Game Theory, 2013
145 Charles A. Weibel, The K-book, 2013
144 Shun-Jen Cheng and Weiqiang Wang, Dualities and Representations of Lie
Superalgebras, 2012
143 Alberto Bressan, Lecture Notes on Functional Analysis, 2013
142 Terence Tao, Higher Order Fourier Analysis, 2012
141 John B. Conway, A Course in Abstract Analysis, 2012
140 Gerald Teschl, Ordinary Differential Equations and Dynamical Systems, 2012
139 John B. Walsh, Knowing the Odds, 2012
138 Maciej Zworski, Semiclassical Analysis, 2012
137 Luis Barreira and Claudia Valls, Ordinary Differential Equations, 2012
www.ams.org
GSM/175
Licensed to AMS.
License or copyright restrictions may apply to redistribution; see https://www.ams.org/publications/ebooks/terms