Analysis of Manifolds IMPA
Analysis of Manifolds IMPA
Analysis of Manifolds IMPA
Version: 20201116.2250
Download the last version from here: http://luis.impa.br/aulas/anvar/aulas.pdf
Bibliography: [Tu], [Sp], [Le], [Ha], [Hi], [Hr]...
Clickable index
1
§1. Manifolds
We want to extend calculus: object needs to be locally a vector
space. Example: Sn.
Topological space, neighborhood, covering.
Countable basis.
Hausdorff (T2).
REM: Countable basis and Hausdorff are inherited by subspaces.
Locally Euclidean Topological space: charts and coordinates.
Dimension, notation: dim M n = n.
Topological manifold = Topological space + Locally Euclidean +
Countable basis + Hausdorff.
Examples: Rn, graph, cusp. Not a manifold: ‘ × ’ (⊂ R2).
Compatible (C ∞–)charts, transition functions, atlas (C ∞).
Example: Sn.
Differentiable structure = maximal atlas.
From now on, for us: Manifold = differentiable manifold = smooth
manifold = Topological manifold + maximal (C ∞) atlas.
Examples: Rn, Sn, U ⊂ M n, GL(n, R), graphs, products.
2
§3. The moduli space
2
As you know, Rn and the set of square matrices Rn×n are iso-
morphic as vector spaces. This means that, although they are
different as sets, they are indistinguishable as vector spaces:
every inherent property of vector spaces is satisfied by both.
In fact, the dimension is the only vector space property that
distinguishes between vector spaces (of finite dimension).
Now, regard M := R as a topological manifold, and N := R
as a smooth manifold. Consider the map τ : M → N given
by τ (t) = t3. Since τ is a homeomorphism, the topologies and
therefore the sets of continuous functions on M and N agree:
C 0(M ) = C 0(N ). On the other hand, since τ is a bijection, there
is a unique differentiable structure on M such that τ is a diffeo-
morphism, that is, the one induced by {τ } as an atlas. Let M̂ be
M with this differentiable structure. Now, although M̂ = N as
sets (and as topological manifolds), M̂ 6= N as smooth manifolds,
since τ is not even an immersion when we regard on M = R the
standard differentiable structure of R. In fact, F(M̂ ) 6= F(N ).
However, τ : M̂ → N is a diffeomorphism by definition (hence
F(M̂ ) = {g ◦ τ : g ∈ F(N )}), and thus, by the above discussion,
as smooth manifolds they should be indistinguishable! Huh????
Answer: As a general fact in math, when studying a mathemat-
ical structure as such, we should distinguish them only up to
the isomorphism of the category. That is, we should not really
study the set Mn of differentiable n-manifolds, but its moduli
space Mn/∼, where two manifolds are identified if they are dif-
2
feomorphic. So we finally obtain M̂ ∼ N , as we got Rn ∼ Rn×n.
3
In fact, every topological manifold of dimension n ≤ 3 has a dif-
ferentiable structure, which is also unique (in the above sense).
Yet, in 1956 John Milnor showed that the topological 7-sphere S7
has more than one differentiable structure! We now know exactly
how many smooth structures exist on Sn... except for n = 4
for which almost nothing is known. See here. (Don’t worry, you will
understand more of this Wiki article by the end of the course).
§4. Quotients
Exercise: Show that on any topological space quotient there is a unique minimal topo-
logical structure, called quotient topology, such that the projection π is continuous (i.e.,
the final topology of π). But the quotient of a manifold is not necessarily a manifold...
4
§5. The tangent space
Germs of functions: Fp(M ) = {f : U ⊂ M → R : p ∈ U }/ ∼
TpM , x : Up ⊂ M n → Rn chart ⇒ ∂x∂ i |p ∈ TpM , 1 ≤ i ≤ n.
Differential of functions ⇒ chain rule.
f local diffeomorphism ⇒ f∗p isomorphism ⇒ dimension is pre-
served by local diffeomorphisms.
Converse: Inverse function Theorem (it has to hold!).
Since every chart x is a diffeomorphism with its image and since
x∗p(∂/∂xi|p) = ∂/∂ui|x(p) ∀1 ≤ i ≤ n,
then { ∂x∂ 1 |p, . . . , ∂x∂ n |p} is a basis of TpM ⇒ dim TpM = dim M .
Local expression of the differential.
Curves: speed, local expression.
Differential using curves: every vector is the derivative of a curve.
REM: TpRn = Rn. Therefore, if f ∈ Fp(U ), v ∈ TpM , then
f∗p(v) = v(f ).
Differential of curves, and computation of differentials using curves.
Immersion, submersion, embedding. Rank.
Examples: projections and injections in product manifolds.
Identification of the tangent space of a product manifold:
Tp M × Tp 0 M 0 ∼
= T(p,p0)(M × M 0).
Definition 1. The point p ∈ M is a critical point of f : M → N
if f∗p is not surjective. Otherwise, p is a regular point. The point
q ∈ N is a critical value of f if it the image of some critical point.
Otherwise, q is a regular value of f (in particular, q ∈ N, q 6∈ Im (f ) ⇒ q is a
regular value of f ).
5
§6. Submanifolds
Regular submanifolds S ⊂ M , adapted charts ϕS .
Codimension. Topology.
Examples: sin(1/t) ∪ I; points and open sets.
The ϕS give an atlas of S.
Differentiable functions from and to regular submanifolds.
Level sets: f −1(q). Regular level sets.
Examples: Sn, SL(n, R): use the curve t 7→ det(tA) !!
Exercise: S ⊂ M is a submanifold ⇐⇒ ∃ covering C of S such that S ∩ U is a
submanifold of U , for all U ∈ C.
πk := (x1 , . . . , xm ) 7→ (x1 , . . . , xk , 0, . . . , 0) ∈ Rn .
Conclude from this the normal form of immersions and submersions as particular cases.
6
Exercise: Conclude for the previous exercise that, if f has constant rank = k in a
neighborhood U of f −1 (q) 6= ∅, then U ∩ f −1 (q) is a regular submanifold of M m with
dimension m − k.
§9. Orientation
Orientability: bundle! Example: T M is orientable as manifold.
Moebius strip: paper trick, knot: intrinsic vs extrinsic topology.
8
§10. Differential 1–forms
Ω1(M ) = Γ(T ∗M ) = {w : X (M ) → F(M )/w is F(M )−linear}:
Local operator ⇒ point-wise operator ⇒ F(M )-linear.
f ∈ F(M ) ⇒ df ∈ Ω1(M ), and df ∼ = f∗ .
∂ ∂
(x, U ) chart ⇒ { ∂x1 |p, . . . , ∂xn |p} is basis of TpM whose dual basis
is {dx1|p, . . . , dxn|p} (i.e., basis of Tp∗M ).
{dx1, . . . , dxn} are then a frame of T ∗U : local expression.
Example: Liouville form on T ∗M : λ(w) := w ◦ π∗w .
Pull-back: ϕ ∈ End(V , W ) ⇒ ϕ∗ ∈ End(W ∗, V ∗);
f : M → N ⇒ f ∗ : F(N ) → F(M ); f ∗ : Ω1(N ) → Ω1(M ).
Importance of pull-back!
Restriction of 1-forms to a submanifold i : S → M : w|S = i∗w.
F∗ ◦ d = d ◦ F∗
11
i.e., F ∗ : Ω•(N ) → Ω•(M ) is a morphism of differential graded
algebras (i.e., preserves degree and commutes with d).
REM: This also explains why dω|U = d(ω|U ) via inc∗.
Exercise: ∀ k, ∀ ω ∈ Ωk (M ), ∀ Y0 , . . . , Yk ∈ X (M ), it holds that dw(Y0 , . . . , Yk ) =
k
X k
X
(−1)i Yi ω(Y0 , . . . , Ŷi , . . . , Yk ) + (−1)i+j ω([Yi , Yj ], Y0 , . . . , Ŷi , . . . , Ŷj , . . . , Yk ).
i=0 0≤i<j≤k
12
Manifold with boundary: definition. (Rough idea of orbifold).
Interior points.
The boundary of M n = ∂M n is a manifold of dimension n − 1.
∂M versus topological boundary.
If p ∈ ∂M : Fp(M ), Tp∗M , v ∈ TpM (yet, it could be no curve
with α0(0) = v), T M , orientation: SAME as before.
If p ∈ ∂M : v ∈ TpM interior and exterior.
REM: In any manifold with boundary M there exists an ex-
terior vector field X along ∂M (i.e., considering the inclusion
inc : ∂M → M we have that X ∈ Xinc). Then, ∂M is ori-
entable if M is, with the induced orientation inc∗iX ω.
Examples: Hn, [a, b]; B n, B n.
Example: If j = inc : Sn−1 = ∂B n → B n, Z(p) = p ∈ Xinc is
exterior ⇒ orientation σ in Sn−1 ⊂ B n via B n ⊂ Rn and dvRn :
X
∗
σ = j (iZ dvRn ) = (−1)i−1 xi dx1 ∧ · · · ∧ dx
ci ∧ · · · ∧ dxn. (2)
i
13
Change of Variables Theorem (CVT) that
Z Z
ξ ∗ω = ξ ∗(f dx1 ∧ · · · ∧ dxn)
U ZU
= f ◦ ξ (ξ ∗dx1 ∧ · · · ∧ ξ ∗dxn)
ZU
= f ◦ ξ (dξ1 ∧ · · · ∧ dξn)
ZU Z
= f ◦ ξ det(Jξ ) dx1 ∧ · · · ∧ dxn = ω.
U V
Def.: If M n is oriented, ϕ : UR⊂ M n → Hn chart oriented, and
w ∈ Ωnc(U ), we define U ω = M ω := ϕ(U )(ϕ−1)∗w. Linear!
R R
n n n
R P R
Def.: RM oriented, w ∈ Ωc (M ) ⇒ M ω := α M ρα w.
CVT: N ϕ∗ω = M ω, ∀ ϕ ∈ Dif+(N, M ), ∀ w ∈RΩnc(M n).
R
Underlying idea: Sum integrals over small cubes, since the inte-
rior faces cancel down due to orientation
R (dim 1 and 2 pictures).
Cor.: M n compact oriented ⇒ M dω = 0, ∀ω ∈ Ωn−1(M ).
Exercise: The classical calculus theorems all follow from Stokes.
14
OBS (!!): i : N k ⊂ M , N k compact oriented regular sub-
k k k
manifold,
R and
R ∗ ω ∈ Ω (M ) (or N oriented and ω ∈ Ω c (M ))
⇒ N ω (= N i ω). R
If ρ ∈ Dif f+(N k ) ⇒ N ρ∗ω = N ω ⇒ we only care about the
R
15
fold M , w ∈ Ωk−1(M ), and c ∈ Ck (M ), we have that
Z Z
dω = ω.
c ∂c
R
In other words, ∂ (over R) is the dual (with respect to ) of d.
Everything works exactly the same considering k-simplex instead
of k-cubes.
16
§19. De Rham cohomology (Spivak, vol.1 chap.8)
Pull-back: F : M → N ⇒ F ∗(= F #) : H k (N ) → H k (M ).
(F ◦G)∗ = G∗ ◦F ∗ ⇒ H k (M ) is an invariant of the differentiable
structure (!), and invariant under diffeomorphisms.
∧ : H k (M ) × H r (M ) → H k+r (M ), [ω] ∧ [σ] := [ω ∧ σ] (well!).
H •(M ) := ⊕k∈ZH k (M ) is the de Rham cohomology ring of M .
17
In fact, H •(M ) is a anticommutative graded algebra, and F ∗ is
a homomorphism of graded algebras.
19
Definition 19. A deformation retract from M to S ⊂ M is a
function T : M × [0, 1] → M such that T0 = IdM , Im (T1) ⊆ S,
and T1|S = IdS (i.e., retract T1 ∼ T0 = IdM ⇒ T1∗ and inc∗S are
isomorphisms).
• gσ = r∗ (i∗ (gσ)) = d(r∗ λ) outside B1n , since (i ◦ r)∗p = kpk−1 Πp⊥ , (i ◦ r)∗ σ(p) = kpk−n σ(p),
and g(p) = kpk−n (g ◦ i ◦ r)(p), if kpk ≥ 1.
• If β := gσ − d(hr∗ λ) ⇒ w = d(gσ) = dβ, with sup(β) ⊆ B1n .
23
Let w = f (r)dx R ∧ dy = fR(r)rdr ∧ dθ with f with compact
support. Then, R2 h∗w = k R2 w ⇒ deg(g) = deg(h) = k > 0 ⇒
g is surjective.
Another proof: h is a local diffeo that preserves orientation on
C \ {0}, and ∀u ∈ C \ {0}, h−1(u) has k points ⇒ deg(h) = k.
24
§23. The birth of exact sequences
Let U, V ⊂ M open such that M = U ∪ V , k ∈ Z ⇒ iU : U ,→
M , jU : U ∩ V ,→ U ⇒ i∗U : Ωk (M ) → Ωk (U ), jU∗ : Ωk (U ) →
Ωk (U ∩ V ). Idem for iV , jV . We then have:
i = i∗U ⊕ i∗V : Ωk (M ) → Ωk (U ) ⊕ Ωk (V ),
j = jV∗ ◦ π2 − jU∗ ◦ π1 : Ωk (U ) ⊕ Ωk (V ) → Ωk (U ∩ V ),
i.e., i(ω) = (ω|U , ω|V ), j(σ, ω) = jV∗ ω − jU∗ σ = ω|U ∩V − σ|U ∩V .
Joining, we get
i j
0 → Ωk (M ) → Ωk (U ) ⊕ Ωk (V ) → Ωk (U ∩ V ) → 0, (3)
with each image contained in the kernel of the next. Now, the
fundamental point is that, in fact, they equal! (the only not
obvious is that j is surjective, but, if {ρU , ρV } is a partition of
unity subordinated to {U, V } and ω ∈ Ωk (U ∩ V ), then ωU :=
ρV ω ∈ Ωk (U ), ωV := ρU ω ∈ Ωk (V ), and j(−ωU , ωV ) = ω).
25
Proposition 30. (General linear algebra dimension Theorem)
α β
If 0 → V 1 → V 2 → · · · → V k → 0 is exact ⇒ i(−1)i dim V i = 0.
P
β[ ]
Proof: Induction on k, changing to 0 → V 2/Im α → V 3 → · · ·
Cochain complex: C = {C k }k∈Z + ‘differentials’ {dk }k∈Z:
d−1 d d
· · · C −1 → C 0 →0 C 1 →1 C 2 · · · , dk ◦ dk−1 = 0.
Direct sum of cochain complexes.
a ∈ C k is a k−cochain of C.
a ∈ Z k (C) := Ker dk ⊂ C k is a k−cocycle of C.
a ∈ B k (C) := Im dk−1 ⊂ C k is a k−coboundary of C.
The k-th cohomology of C is given by
H k (C) := Z k (C)/B k (C).
If a ∈ Z k (C) ⇒ [a] ∈ H k (C) is the cohomology class of a.
Um cochain map ϕ : A → B is a sequence {ϕk : Ak → B k }k∈Z
such that d ◦ ϕk = ϕk+1 ◦ d. This gives maps ϕ∗ : H •(A) →
i j
H •(B). The sequence 0 → A → B → C → 0 is said to be short
exact if at each level k is exact. In this situation,
i∗ j∗
H (A) → H (B) → H k (C)
k k
is exact for all k. But it is NOT exact with 0 at the left or at the
right... BUT:
26
i j
Theorem 31 (!!!!!!!). If 0 → A → B → C → 0 is short ex-
act, then there exist homomorphisms (explicit and natural)
δ ∗ : H k (C) → H k+1(A),
called connection homomorphisms, that induce the following
long exact sequence in cohomology:
27
coboundary and the b is a cocycle: a = da0. Thus db = ida0 =
dia0, i.e., d(b − ia0) = 0. So j ∗[b − ia0] = [jb − jia0] = [jb] = [c].
i∗ j∗
0 → H 1(M ) → H 1(U ) ⊕ H 1(V ) → · · ·
28
δ∗
are exact (since M is connected, and H (U ∩ V ) → H 1(M ) is
0
H •(T 2).
Mayer–Vietoris + Proposition 30 ⇒
χ(M ) = χ(U )+χ(V )−χ(U ∩ V ). (4)
Simplex ⇒ triangulations: always exist (by countable basis).
29
Proof: For each n-simplex σi of T , choose pi ∈ σio and pi ∈
Bpi ⊂ σio (think about pi as a small ball too). Let U1 be the
disjoint union of these αn balls, and Vn−1 = M \ {p1, . . . , pαn }.
Then, (4) ⇒ χ(M n) = χ(Vn−1) + (−1)nαn.
For each (n − 1)-face τj of T , pick the “long” ball Bτj joining
the two Bpi ’s of each n-simplex touching τj . Call U2 the union
of these disjoint αn−1 balls. Pick an arc (inside Bτj ) joining the
boundaries of the two Bpi ’s , and let Vn−2 be the complement of
these αn−1 arcs. Again, (4) ⇒ χ(Vn−1) = χ(Vn−2)+(−1)n−1αn−1.
Inductively we obtain Vn−3, · · · , V0, the last one being the union
of α0 contractible sets (each one a neighborhood of a vertex of T ),
so that χ(V0) = α0 and χ(Vk ) = χ(Vk−1) + (−1)k αk .
30
Platonic model of the solar system by Kepler; Circogonia icosahedra; Stones from 2000 AC
32
Now, since j ∗ : H k (Vi) → H k (N ) is an isomorphism for all i
and for all k, H k (N ) is isomorphic to H k (G) (exercise). Then,
Theorem 31 + Lemma 39 ⇒
33
That is, dim H n(M n) + 1 = b0(Rn+1 \ M n) ≥ 2 (Corolary 29).
Hence, by Theorem 22 and Theorem 23, H n(M n) ∼ = R, M n is
orientable, and #{connected components of Rn+1 \ M n} = 2.
By the same argument for winding numbers, each point of M n is
arbitrarily close to points in both connected components.
Corolary 44. Neither the Klein bottle nor the projective
plane can be embedded in R3.
35
Theorem 51 (Poincaré duality). If M n is connected and
orientable, the linear function P D: H k (M ) → (Hcn−k (M ))∗,
Z
P D([ω])([σ]) := ω∧σ
M
↓ PD ⊕ PD ↓ PD ↓ PD ↓ PD ⊕ PD ↓ PD
(Hcl+1 (U ) ⊕ Hcl+1 (V ))∗ → Hcl+1 (U ∩ V )∗ → Hcl (M )∗ → (Hcl (U ) ⊕ Hcl (V ))∗ → Hcl (U ∩ V )∗
where all vertical maps are isomorphisms, except maybe the middle
one. Moreover, all squares commute up to signs (exercise), and
hence up to some signs in the P D’s everything commutes. The
lemma follows now from the five Lemma (prove), which says
precisely that the middle one must also be an isomorphism.
36
30.1 The Poincaré sphere
Henri Poincaré conjectured that a 3-manifold with the homology of a sphere must be
homeomorphic to the 3-sphere S3 . Poincaré himself found a counterexample, essentially
creating the concept of fundamental group. Indeed, by Hurewicz theorem, it would be
enough to take S3 /G, with G ⊂ SO(4) a nontrivial perfect group (i.e., G = [G, G])
acting freely. The simplest such example that we can think of is G = A5 ⊂ SO(3) as
the order 60 icosahedral group since A5 is simple. This almost works, except that G
has to be extended to the binary icosahedral group G = 2A5 of order 120, which is still
perfect, though not simple (or work with A5 but on SO(3) ∼
= S3 /{±I} instead) . Then,
H1 (S3 /G, Z) = G/[G, G] = 0, and H2 (S3 /G) = H1 (S3 /G) = 0 by e.g. Poincaré duality.
Thus, H∗ (S3 /G) = H∗ (S3 ), yet S3 /G is not simply connected. It is remarkable that
this is the only example with finite fundamental group (there are plenty with infinite
fundamental group). After Poincaré found this counterexample to his own conjecture,
he made another one: the 3-sphere is the only simply connected homology 3-sphere.
This is of course the very famous Poincaré conjecture, proved (among other things!) by
G.Perelman in 2002. Notice that, by Perelman’s result, any homology 3-sphere with
finite fundamental group must be S3 /G, with G ⊂ SO(4) perfect, reducing the original
problem to a group one: find the finite perfect subgroups of SO(4) that act freely. It
turns out that 2A5 is the only one!
Proof: See here for a general argument, even for manifolds that
are not of finite type.
38
The End. :o)
References
[Ha] Hatcher, A.: Algebraic topology. Cambridge University
Press, 2002.
[Hi] Hitchin, N.: Differentiable manifolds. Lecture notes here.
[Hr] Hirsch, M.: Differential topology. Graduate text in Math-
ematics 33, Springer-Verlag, New York, 1972.
[Le] Lee, J.: Introduction to smooth manifolds. University of
Washington, Washington, 2000.
[Tu] Tu, L: An introduction to manifolds. Second edition. Uni-
versitext. Springer, New York, 2011.
[Sp] Spivak, M.: The comprehensive introduction to differen-
tial geometry.. Vol. III. Third edition. Publish or Perish,
Inc., Wilmington, Del., 1979.
[Zi] Zinger, A: Notes on vector bundles.. Lecture notes here.
39