Differential Geometry Guide: Luis A. Florit (Luis@impa - BR, Office 404)
Differential Geometry Guide: Luis A. Florit (Luis@impa - BR, Office 404)
Differential Geometry Guide: Luis A. Florit (Luis@impa - BR, Office 404)
PART I
Prerequisites (part I): Analysis on Rn .
Analysis on manifolds, Stokes and de Rham recommended. Bibliography: [dC]
§1. Introduction
Differentiable manifolds: smooth world. Now we’re going to
measure in them. After all, geometry comes from the Greek:
“measurement of the Earth”:
Eratosthenes (Cirene, 276 AC - Alexandria, 194 AC)
Posidonius (135 AC - 51 AC) ⇒ Colombo
We will study in this first part curves, surfaces and hypersurfaces
of Euclidean space ⇒ Two aspects: intrinsic and extrinsic.
§2. Curves
Curve: intrinsically, nothing interesting: I ⊆ R.
Regular curves α in R2 and R3: arc length s, p.b.a.l, curvature
κα and torsion τ α . Frenet Trihedron: {t, n, b}.
FTC: explicit vs. ODE.
Curves in Rn: FTC.
p
Exercise. Prove that κα = kα′ k2 kα′′ k2 − hα′ , α′′ i2 /kα′ k3 , indepen-
dently of the parametrization of α.
1
§3. Surfaces in R3
Regular (Euclidean) surface: S = S 2 ⊂ R3 (embedding!)
Regular (Euclidean) hypersurface: M n ⊂ Rn+1 (embedding!)
Regular (Euclidean) submanifold: M n ⊂ Rn+p (embedding!)
It is enough to check that: ∀x ∈ M n, ∃ V ⊂ Rn+p open, x ∈ V ,
and a smooth map U ⊂ Rn 7→ M n ∩ V ⊂ Rn+p that is injective,
open and has rank n: coordinates (smooth = C r /C ∞/C w ).
Examples:
graf(f ) for f : U ⊂ Rn → R.
g −1(t0) for a regular value t0 (in the image) of g : W ⊂ Rn+1 → R:
Sphere Sn ⊂ Rn+1.
Ellipsoid g −1(r) for g(x, y, z) = x2/a2 + y 2/b2 + z 2/c2, r > 0.
Hyperboloid g −1(r) for g = x2 + y 2 − z 2 (Hyperboloid of two
sheets for r < 0, while the Cone g −1(0) is NOT a regular surface).
Hyperbolic paraboloid g −1(0) for g = x2/a2 − y 2/b2 − z.
Circular cylinder and Cylinders over conics.
2
Examples: F : U ⊂ R3 → R smooth ⇒ F |S is smooth ∀S ⊂ U
ϕ : U ⊂ R2 → R3 coordinates ⇒ U and ϕ(U ) are diffeomorphic
S symmetric ⇒ the symmetry restricted to S is smooth:
3
vector field of our surface:
∂ϕ
∂u ∧ ∂ϕ
∂v
N= ∂ϕ
(1)
k ∂u ∧ ∂ϕ
∂v k
ϕ : U → S ⊂ R3 coordinate system ⇒
∂ϕ 2 ∂ϕ ∂ϕ ∂ϕ
E=k k , F =h , i, G = k k2 ∈ C ∞(U )
∂u ∂u ∂v ∂v
coefficients of I in the coordinate
R t √ ϕ.
Im (α) ⊂ Im (ϕ) ⇒ s(t) = t0 Eu′2 + 2F u′v ′ + Gv ′2dr.
4
Same for arbitrary regular Euclidean submanifolds:
Example: Affine plane through p ∈ R3 or Cylinder over plane
curve ⇒ everywhere coordinate systems with E ≡ 1 ≡ G, F ≡ 0.
∂ϕ ∂ϕ
Z Z p
A(Ω) = k ∧ k= EG − F 2 dudv
ϕ−1 (Ω) ∂u ∂v ϕ−1 (Ω)
since kx ∧ yk2 + hx, yi2 = kxk2 kyk 2
5
§8. Gauss map and Second Fundamental Form
For any Euclidean hypersurface M n ⊂ Rn+1, locally we have a
unit normal vector field
N : U ⊂ M n → Sn ⊂ Rn+1 .
But TpM is parallel to TN (p)Sn, and hence dNp ∈ End(TpM ).
Moreover, dNp is self adjoint (w.r.t. I), so the quadratic form
IIp(w) := −hdNpw, wi
is called the second fundamental form of M n at p. We also give
the same name to the associated symmetric tensor,
Ap := −dNp.
Def.: Let α : I → M be a regular curve through p = α(0) ∈ M .
The normal curvature of α at p is given by
κn := κhN, ni,
where n is the normal vector of α at p and κ its curvature.
6
Examples: Second fundamental form of a graph; y = x4; Sn.
A self adjoint ⇒ principal curvatures {ki}, principal directions
{ei}, lines of curvature:
X X
v= viei ⇒ II(v) = kivi2.
i i
7
Notice that K does not depend on orientation, while H does.
Notice that both are smooth functions that determine k1 and k2:
p
ki = H ± H 2 − K.
8
§10. II and K in coordinates, part 1
For a coordinate system ϕ = ϕ(u, v) in a surface S ⊂ R3, let as
before
∂ϕ ∂ϕ
∂u ∧ ∂v
N = ∂ϕ ∂ϕ
.
k ∂u ∧ ∂v k
Denote by (aij ) the matrix of A in the coordinate basis,
Nu = a11ϕu + a12ϕv ,
Nv = a21ϕu + a22ϕv .
Define the functions
e := −hNu, ϕui = hN, ϕuui,
g := −hNv , ϕv i = hN, ϕvv i,
f := −hNu, ϕv i = −hNv , ϕui = hN, ϕuv i.
Hence, if v = v1ϕu + v2ϕv , II(v) = ev12 + 2f v2 v2 + gv22, and
a11 a12 −1 e f G −F
=
a21 a22 EG − F 2 f g −F E
These are known as the Weingarten equations. In particular,
for the Gaussian curvature we obtain
eg − f 2
K= .
EG − F 2
Example: The torus. For 0 < r < a, take the (almost global)
chart
ϕ(u, v) = ((a + r cos(u)) cos(v), (a + r cos(u)) sin(v), r sin(u)) .
9
Hence, N = (cos(u) cos(v), cos(u) sin(v), sin(u)) (geometrically!),
and hence E = r2, G = (a + r cos(u))2 , F = 0, and
cos(u)
K= .
r(a + r cos(u))
Make the computation, make a picture, sign K interpretation,
elliptic/hyperbolic points, move a and r and see how K varies,
independence of v... see everything geometrically!
R
Remark 9. Observe that K = 0, independently of a and r!
10
Proposition 11. Let p ∈ S ⊂ R3, and a sequence of compact
regions Bi ⊂ S such that Bi → p. Then,
Area(N (Bi))
|K(p)| = lim .
i→∞ Area(Bi )
Remark 20. Besides the plane, these are the only 2 doubly
ruled surfaces! How would you prove this??
Singularities of a ruled surface are contained in the striction curve:
Def.: For a noncylindrical ruled surface S, the striction curve
is given by σ(s) = α(s) + t(s)v(s) for which hσ ′, v ′i = 0 (i.e.,
t(s) = −hα′, v ′i/kv ′k2).
Notice that the striction curve does not depend on the directrix
α. In particular, we can assume that σ = α, that is, hα′, v ′i = 0.
13
niz persuaded Bernoulli to extend the six month limit to solve the
challenge for foreign mathematicians to be able to participate.
Five more mathematicians solved the problem: Tschirnhaus, Ja-
cob Bernoulli, Leibniz, de L’Hôpital, and... Isaac Newton, who
was teased by Bernoulli and Leibniz, and solved the problem in
one night. These solutions eventually lead to a general method by
Euler to solve these kind of problems: the calculus of variations.
When does a surface minimize the area for “close enough” sur-
faces?
14
§14. Intrinsic Geometry
Intrinsic objects of a Riemannian manifold are the ones that
only depend on the first fundamental form, i.e., invariant by
isometries: distance, angle, area, volume...
Cylinder ∼
= plane ∼
= cone (locally): they are intrinsically the same
thing, and hence the mean curvature H is not an intrinsic concept.
If ϕ : U → M , ϕ′ : U → M ′ are charts such that gij = gij′ ⇒
ϕ(U ) ⊂ M and ϕ′(U ) ⊂ M ′ are isometric.
Example: For a surface of revolution
ϕ(u, v) = (f (v) cos(u), f (v) sin(u), g(v))
we have E = f 2, F = 0, G = f ′2 + g ′2. In particular, the
catenoid, where f (v) = a cosh(v) and g(v) = av for a > 0, has
E = G = a2 cosh2(v), F = 0. Now, change variables on the
helicoid
ϕ(u, v) = v(cos(u), sin(u), 0) + aue3,
v = a sinh(v), u = u to get
ϕ(u, v) = a(sinh(v) cos(u), sinh(v) sin(u), u),
that also has E = G = a2 cosh2(v), F = 0. Therefore, the
catenoid and the helicoid are locally isometric (but not globally).
This is a general phenomenon for minimal surfaces in R3 : they
have a one parameter family of isometric deformations.
Def.: Isometries, conformal diffeomorphisms.
Existence of isothermal coordinates ⇒ any two surfaces are lo-
cally conformally equivalent.
15
§15. The Gaussian curvature in coordinates, part 2
Given a chart ϕ = ϕ(u1, u2) : U → S on a surface S ⊂ R3, we
have seen that K = (eg − f 2)/(EG − F 2). Let’s do this compu-
tation again in other way, by decomposing the second derivatives
on their tangent and normal components:
X
ϕij = Γkij ϕk + rij N.
k
16
Exercise. Show that for a surface of revolution around a curve α(v) =
(f (v), g(v)), it holds that Γ111 = Γ212 = Γ122 = 0, Γ211 = f f ′ /(f ′2 + g ′2 ), Γ112 =
f ′ /f, Γ222 = (f f ′′ + gg ′′ )/(f ′2 + g ′2 ).
Now, we get relations that come from taking the tangent and
normal components of ϕrij = ϕrji and Nij = Nji (3×3 equations
if n = 2) that have the form
n
X
ckijr ϕk + dkijr N = 0, ∀1 ≤ i, j, r ≤ n = 2. (2)
k=1
17
two more equations, called the Codazzi-Mainardi equations:
ev − fu = eΓ112 + f (Γ212 − Γ111) − gΓ211.
fv − gu = eΓ122 + f (Γ222 − Γ112) − gΓ212.
19
of this first order system of PDE: Pv − Qu = [P, Q] (Frobenius
Theorem). Integrating once more, we get ϕ, and it is easy to check
that it is a surface with the desired first and second fundamental
forms (for details, see R. Palais notes here).
Remark 30. Take a long time comparing this with the FTC.
22
P
we write η on V as η|V = i zi ξi ◦ f . Just by the definition of
a connection and its local nature, on V we get
X
∇ˆ η=
f ˆ f X ξi .
X(zi)ξi ◦ f + zi∇ (4)
X ∗
i
ˆ f . But we can define ∇
This implies the uniqueness of ∇ ˆ f locally
with (4): it is easy to check that ∇ ˆ f defined this way is well
defined, and a connection.
′
X
′
X ∂
Y = yk + αi′ yj Γkij ◦ α ◦α (6)
ij
∂u k
k
25
Beware: along α!! It does depend on α, not just on α(t0 ) and
α(t) (in contrast to Rm).
Examples: M = Rm: usual. Meridian in S2: Cylinder. Parallel
in S2: Cone ⇒ after a complete turn, the parallel transport does
not close ⇒ dependency on α.
§23. Geodesics
Lemma 38. Given ϕ : U ⊂ R2 → M ⇒ ∇∂u ϕv = ∇∂v ϕu.
Proof: Use coordinates and the symmetry of ∇ (it’s equivalent).
Proposition 39. A curve α parametrized by arc-length is a
critical point of the arc-length functional if and only if α′′ = 0.
Proof: Calculus of variations! :-)
29
Proof: Define σ ∈ Ω1(N ) by σ(X) = aX(b) − bX(a). Since
a2 + b2 = 1 we easily check that σ is closed, hence exact. Define
now ξ by dξ = σ, ξ(p0 ) = ξ0. The lemma follows simply by
differentiating (a−cos ξ)2 +(b−sin ξ)2 = 2−2(a cos ξ +b sin ξ).
31
Def.: We say that a compact region R ⊂ S is simple if R ∼ = D2
and ∂R is the trace of a closed simple piecewise regular curve α,
that we orient in such a way that α′ points to the interior of R.
Theorem 49. (Gauss-Bonnet; local) Let R ⊂ S be a simple
region contained in the image of an orthogonal oriented chart
ϕ:U ∼ = D2 → S, and let α : I → ∂R oriented p.b.a.l. with
vertices α(si) and external oriented angles θi. Then,
X Z si+1 X Z
καg + θi + K = 2π.
i si i R
32
Z Z
= − K + ξ(ℓ) − ξ(0) = ∆ξ − K,
R R
∆ξ
where ξ = ∢(w, ϕu). So, limR→p A(R) = K(p), and we conclude:
33
For compact connected surfaces it holds that:
χ(S) = χ(S ′) ⇔ S is (diffeo)homeomorphic to S ′ (!!)
If S is orientable ⇒ 4 − χ(S) ∈ 2N. Therefore, by Jordan’s
Theorem, the only compact regular surfaces in R3 are the n-tori
(up to diffeomorphism). Thus, the number g of “handles” of S,
g := 2−χ(S)
2 ∈ N0, is called the genus of S.
Def.: A region R ⊂ S is regular if it is compact and ∂R is a
disjoint union of simple closed piecewise differential curves.
Theorem 51. (Gauss-Bonnet; global) Let R be a regular re-
gion of an oriented surface S, and let ∂R = ∪ki=1Ci positively
oriented with external angles θ1, . . . , θm. Then,
X k Z Xm Z
καg + θj + K = 2πχ(R).
i=1 Ci j=1 R
34
where θrj are the 3 oriented external angles of Tr , since the
internal edges of each triangle get opposite orientations. Call
βrj := π − θrj the internal angles of the triangle Tr . Thus,
X X m
X
3πF − θrj = βrj = 2πVi + πVet + (π − θl ).
rj rj l=1
36
PART II
Prerequisites (part II): Analysis on manifolds, Stokes and de Rham cohomology.
Bibliography: [KN] Vol.II, Ch.12; [S] Vol.V Ch.13; [MS]
§30. Fiber and principal bundles ([KN], Vol. I, Ch. 1.4, 1.5)
Exercise. Show that the pull-back and Whitney sum of vector bundles is
a vector bundle.
Exercise. Show that the S1 and S3 Hopf bundles are principal bundles.
Exercise. Show that the lens spaces S3 /Zk for Zk ⊂ S1 are circle bundles
over S2 .
41
Proof: If X, Y ∈ H, follows from the definition of D. If
X = A∗ ∈ V is fundamental, Ω(A∗, ·) = 0 because it is verti-
cal. Now, if B ∗ ∈ V is fundamental, dw(A∗, B ∗) = A∗(w(B ∗)) −
B ∗(w(A∗)) − w([A∗ , B ∗]) = A∗(B) − B ∗(A) − [A, B] = −[A, B] =
−[w(A∗), w(B ∗)]. Now, if Y ∈ H, [w(A∗), w(Y )] = 0 and
dw(A∗, Y ) = −w([A∗, Y ]) = 0 by Lemma 64.
Proposition 68. (Bianchi’s identity) DΩ = 0.
Proof: We need to check that dΩ = d([w, w]) = 0 for 3 horizontal
vector fields, but this is immediate from H ⊂ Ker w.
Lemma 69. If σ is a horizontal 1-form of type adG,
Dσ = dσ + [σ, w] + [ω, σ].
Proof: The only nontrivial case is for v ∗ ∈ V fundamental and
Y ∗ ∈ H a lift. But Dσ(v ∗, Y ∗) = 0, and dσ(v ∗, Y ∗) = v ∗(σ(Y ∗))
since [v ∗, Y ∗] = 0 by Lemma 64. Since Y ∗ is G-invariant, by
Lemma 62,
v ∗(σ(Y ∗))(p) = σ(Y ∗(pβsv ))′ = σ((Rβsv )∗pY ∗(p))′
= ((Rβ∗sv σ)(Y ∗(p)))′ = (adβs−v (σ(Y ∗(p))))′
= [−v, σ(Y ∗(p))] = −[w(v ∗), σ(Y ∗)](p).
42
Now, let E be a principal G-bundle with principal connection
1-form w and curvature 2-form Ω. For f ∈ I k (G), we define the
2k-form f (Ω) = f (Ω, . . . , Ω) on E by
f (Ω)(X1 , . . . , X2k )
1 X
= sign(σ)f (Ω(Xσ1 , Xσ2 ), . . . , Ω(Xσ2k−1 , Xσ2k )).
(2k)! σ
Theorem 70. (A.Weil) For each f ∈ I k (G), the (2k)-form
f (Ω) ∈ Ω2k (E) projects to a unique closed (2k)-form f (Ω) ∈
Ω2k (B). Moreover, its cohomology class
ωf = [f (Ω)] ∈ H 2k (B)
is independent of the choice of the connection, and ω: I(G) →
H ∗(B) is an algebra homomorphism, called Weil homomorp..
Proof: Since Ω is horizontal by definition, so is f (Ω). Since
Ω is of type adG and f is adG-invariant, f (Ω) is G-invariant:
Rg∗(f (Ω)) = f (Ω). By Lemma 65 f (Ω) projects: f (Ω) = π ∗f (Ω).
Proposition 68 says that DΩ = 0, and hence D(f (Ω)) = 0. By
Lemma 66, f (Ω) is closed, and so is f (Ω) since π∗ is onto.
For the second part, take w1 and w2 two principal connections
on E, and define wt := w0 + t(w1 − w0). Obviously, wt and
α = w1 − w0 are also of type adG, and α(V) = 0. Let Dt and
Ωt be the exterior covariant differentiation and curvature form of
wt, respectively. By Proposition 67, Ωt = Dtwt = dwt + [wt, wt],
so, by Lemma 69, dtd Ωt = Dtα. Therefore, by Proposition 68,
d
f (Ωt) = kf (Dt α, Ωt, . . . , Ωt) = kDt(f (α, Ωt , . . . , Ωt))
dt
= kd(f (α, Ωt , . . . , Ωt)),
43
where the last equality follows from Lemma 65 and Lemma 66:
since both α and Ωt are horizontal and of type adG, and f is adG-
invariant, so βt := f (α, Ωt, . . . , Ωt) is horizontal and G-invariant.
R 1 then βt also projects to a (k −1)-form on B, and so does Φ =
But
k 0 βtdt. We conclude from the above that dΦ = f (Ω1 ) − f (Ω0 )
also projects, and thus f (Ω1) − f (Ω0) = dΦ is exact.
It’s easy to check that ω is an algebra homomorphism (exercise).
Remark 71. Notice that the homology class is in the base B,
not in the total space E!!
Let’s show a much more direct approach for vector bundles. No-
tice that, since G = Gl(n, R) ⊂ Rn×n is open, g = Rn×n.
π
Let Rn → P → M be a real rank n vector bundle over a manifold
M , and ∇ an affine connection on P . Given a local frame e =
{ξ1, . . . ξn} of π −1(U ) ∼
= U × Rn, U ⊂ M , write
X
∇X ξ j = Γij (e)(X)ξi.
i
Let’s see that the two constructions agree. Given P the vector
π̂
bundle above, its frame bundle of G → F(P ) → M is a principal
G-bundle, a trivializing neighborhood of which is F(π̂ −1(U )) ∼
=
U × G. If w is a principal connection on F(P ), and e : U ⊂
M → π̂ −1(U ) ⊂ F(P ) is a local section, the equation
ω(e) = e∗w (10)
relates w with the affine connection form ω. Then, check that
Ωω (e) = e∗Ωw , and so the forms f (Ωw ) project precisely to f (Ωω ):
π̂ ∗(f (Ωω )) = π̂ ∗(f (e∗Ωw )) = π̂ ∗e∗(f (Ωw )) = f (Ωw ), (11)
where for the last equality we used that f (Ωw ) projects.
Exercise. If ω and w are forms related by (10), then ω is well defined,
and it is a principal connection ⇔ w is an affine connection (form).
47
§36. Invariant polynomials ([KN], Vol. II, Ch. XII.2)
49
36.3 Special orthogonal group: SO(n) = {X ∈ O(n): det(X) = 1}.
Hence, we have:
Proposition 78. For n = 2m − 1 (resp. n = 2m) the
polynomial functions p1, . . . , pm−1 (resp. p1, . . . , pm−1, pf ) are
adSO(n)-invariant, algebraically independent, and generate (as
algebra) PSO(n)(so(n)).
Proof: See Theorem 2.7 in Kobayashi-Nomizu, Vol 2, Cap. XII.
Corollary 79. The characteristic classes pk (E) := ωpk ∈
H 4k (B), 1 ≤ k ≤ [ n−1
2 ], together with e(E) := (2π)
−n/2
ωpf ∈
H n(B) if n is even, generate all the characteristic classes of
an SO(n)-principal bundle as an algebra. The classes pi(E)
are called the Pontrjagin classes of the bundle, while, for n
even, e(E) is called the Euler class of the bundle.
50
by means of Section 35. By definition, the classes of a vector
bundle are the classes of its frame principal bundle.
Total Chern and Pontrjagin classes:
c(E) = 1 + c1(E) + c2(E) + · · · ∈ H ∗(B),
p(E) = 1 + p1(E) + p2(E) + · · · ∈ H ∗(B).
52
With these definitions, the Poincaré-Hopf Theorem 60 holds for
any compact oriented manifold and any vector field with isolated
singularities, and not just for surfaces. Indeed, by the proof of
the Gauss-Bonnet-Chern Theorem 83 below it follows that the
total index of a vector field is a topological invariant, i.e., does
not depend on the vector field. But it is easy to construct a
vector field whose total index is the Euler characteristic: for a
triangulation T , define VT as having precisely one singularity on
the ‘center’ of each simplex of T in a way so that the flow lines
of the vector field point from the centers of higher dimensional
simplexes towards the lower dimensional simplexes. Such a vector
field has total index equal to the Euler characteristic.
Remark 85. Again, notice that we not only proved the Gauss-
Bonnet-Chern Theorem, but also Poincaré-Hopf Theorem 60 for
any dimensions.
References
[C] S.S. Chern; A Simple Intrinsic Proof of the Gauss-
Bonnet Formula for Closed Riemannian Manifolds.
Ann. Math. 45 (4), 1944, 747-752.
[dC] M. do Carmo; Differential Geometry of Curves and Sur-
faces. Pearson Education Canada, 1976.
[H] A. Hubery; Notes on fibre bundles. Lecture notes here
with backup here.
[KN] S. Kobayashi, K. Nomizu; Foundations of Differential
Geometry. Wiley Classics Library, Volume I and II.
[L] Y. Li; The Gauss-Bonnet-Chern Theorem on Rieman-
nian Manifolds. ArXiV 1111.4972v4.
[MS] J. Milnor, J. Stasheff; Characteristic classes. Princeton
University Press, 1974.
[S] M. Spivak; A comprehensive introduction to differential
geometry. Publish or Perish Inc., Texas, 1975.
55