Diffgeom Tsakanik Main
Diffgeom Tsakanik Main
2 Smooth Maps 9
2.1 Smooth Maps . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
2.2 Partitions of Unity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
5 Submanifolds 45
5.1 Embedded Submanifolds . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45
5.2 Immersed Submanifolds . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50
5.3 The Tangent Space to a Submanifold . . . . . . . . . . . . . . . . . . . . . 52
6 Vector Bundles 55
6.1 Vector Bundles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55
6.2 Sections of a Vector Bundle . . . . . . . . . . . . . . . . . . . . . . . . . . 59
6.3 Subbundles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62
iii
iv CONTENTS
7.3 Flows . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73
8 Differential Forms 79
8.1 Differential 1-Forms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79
8.2 Differential k-Forms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87
10 Orientations 99
10.1 Orientations of Vector Spaces . . . . . . . . . . . . . . . . . . . . . . . . . 99
10.2 Orientations of Smooth Manifolds . . . . . . . . . . . . . . . . . . . . . . . 101
Bibliography 141
Index 143
CHAPTER 1
TOPOLOGICAL AND SMOOTH MANIFOLDS
• M is a Hausdorff space: for each pair of distinct points p, q ∈ M there are disjoint
open sets U, V ⊆ M such that p ∈ U and q ∈ V .
Comments:
(1) Every topological manifold has, by definition, a specific, well-defined dimension. Thus,
we do not consider spaces of mixed dimension, such as the disjoint union of a plane
and a line, to be manifolds at all. It can be shown (using de Rham cohomology or
singular homology) that the dimension of a (non-empty) topological manifold is in fact a
topological invariant: A non-empty topological n-manifold cannot be homeomorphic to a
topological m-manifold unless n = m.
(2) The three conditions in Definition 1.1 ensure that topological manifolds behave in the
ways we expect from our experience with Euclidean spaces. For example, in a Hausdorff
1
2 Chapter 1. Topological and Smooth Manifolds
space, finite subsets are closed and limits of convergent sequences are unique. The mo-
tivation for second-countability is less evident, but stems from the existence of so-called
partitions of unity; see Section 2.2.
(3) There are also examples of topological spaces which are not topological manifolds;
see Exercise Sheet 1. For example:
The line with two origins is locally Euclidean and second-countable, but not Haus-
dorff.
A disjoint union of uncountably many copies of R is locally Euclidean and Haus-
dorff, but not second-countable.
Example 1.3.
with the subspace topology. Let π1 : Rn × Rk → Rn be the projection onto the first factor
and let φ : Γ(f ) → U be the restriction of π1 to Γ(f ):
and
Ui− := (x1 , . . . , xn+1 ) ∈ Rn+1 | xi < 0 .
Thus, each subset Ui± ∩ Sn is locally Euclidean of dimension n, see (1) above, and the
maps
φ± ± n
i : Ui ∩ S → B
n
f : R2 → R, f (x, y) 7→ x
√
is differentiable, but the composite map (f ◦ φ)(u, v) = 3
u is not differentiable
at (0, 0).
Definition 1.4. Let M be a topological manifold. If (U, φ) and (V, ψ) are two charts
such that U ∩ V ̸= ∅, then the composite map ψ ◦ φ−1 : φ(U ∩ V ) → ψ(U ∩ V ) is called
the transition map from φ to ψ (see Figure 1.3) and is clearly a homeomorphism. Two
charts (U, φ) and (V, ψ) are said to be smoothly compatible if either U ∩ V = ∅ or the
transition map ψ ◦ φ−1 is diffeomorphism (i.e., smooth and bijective with smooth inverse).
Since φ(U ∩ V ) and ψ(U ∩ V ) are open subsets of Rn , smoothness of this map is to be
interpreted in the ordinary sense of having continuous partial derivatives of all orders
(C ∞ ).
An atlas for M is a collection of charts whose domains cover M . An atlas A is called
a smooth atlas if any two charts in A are smoothly compatible. Finally, a smooth atlas A
on M is called maximal (or complete) if it is not properly contained in any larger smooth
atlas. This just means that any chart which is smoothly compatible with every chart in
A is already in A.
Remark 1.5. To show that an atlas is smooth, we need only verify that each transition
map ψ ◦ φ−1 is smooth whenever (U, φ) and (V, ψ) are charts in A; once we have proved
this, it follows that ψ ◦ φ−1 is a diffeomorphism because its inverse φ ◦ ψ −1 = (ψ ◦ φ−1 )−1
is one of the transition maps we have already shown to be smooth. Alternatively, given
Section 1.2. Smooth Manifolds 5
two particular charts (U, φ) and (V, ψ), it is often easier to show that they are smoothly
compatible by verifying that ψ ◦ φ−1 is smooth with non-singular Jacobian at each point
of its domain, since then the Inverse Function Theorem = [Lee13, Theorem C.34] implies
that ψ ◦ φ−1 is a diffeomorphism, see [Lee13, Corollary C.36].
M
V
U
ψ
φ
ψ ◦ φ−1
ψ(V )
φ(U )
Rn Rn
Remark 1.7. A smooth structure is an additional piece of data that must be added to
a topological manifold before we are entitled to talk about a “smooth manifold”. Note
that a given topological manifold may have many smooth structures (in fact, if it has one,
then it has infinitely many, see [Lee13, Problem 1.6]), but it may also admit no smooth
structures at all.
6 Chapter 1. Topological and Smooth Manifolds
(a) Every smooth atlas A for M is contained in a unique maximal smooth atlas, called
the smooth structure determined by A.
(b) Two smooth atlantes for M determine the same smooth structure if and only if their
union is a smooth atlas.
Proof.
(a) Given a smooth atlas A on M , set
A := (U, φ) chart for M | ∀(V, ψ) ∈ A : (U, φ) and (V, ψ) are smoothly compatible .
Definition 1.9. Let (M, A) be a smooth manifold. Any chart (U, φ) contained in the
maximal smooth atlas A is called a smooth coordinate chart. The corresponding coor-
dinate map φ is called a smooth coordinate map, and its domain U is called a smooth
coordinate domain, or smooth coordinate neighborhood of each of its points. A smooth
coordinate ball is a smooth coordinate domain whose image under a smooth coordinate
map is a ball in Euclidean space. A smooth coordinate cube is defined similarly.
Section 1.2. Smooth Manifolds 7
Here is how one usually thinks about (smooth) coordinate charts on a smooth manifold.
Once we choose a (smooth) coordinate chart (U, φ) on M n , the (smooth) coordinate map
φ: U → U b ⊆ Rn can be thought of as giving a temporary identification between U and U b.
Using this identification, while we work in this chart, we can think of U simultaneously
as an open subset of M and as an open subset of Rn . Under this identification, we can
represent a point p ∈ M by its coordinates (x1 , . . . , xn ) = φ(p), and think of this n-
tuple as being the point p. We typically express this by saying “(x1 , . . . , xn ) is the (local)
coordinate representation for p” or “p = (x1 , . . . , xn ) in local coordinates”.
Example 1.10.
(0) For each n ∈ N the Euclidean
n space Rn is a smooth n-manifold with smooth structure
determined by the atlas (R , IdRn ) . We call this the standard smooth structure on Rn
and the resulting coordinate map the standard coordinates on Rn . (Unless we explicitly
say otherwise, we always use this smooth structure on Rn .) With respect to this smooth
structure, the smooth coordinate charts for Rn are exactly those charts (U, φ) such that
φ is diffeomorphism (in the usual sense) from U ⊆ Rn to another open set U b ⊆ Rn .
and similar formulas hold in the other cases. When i = j, the domains of φ+ and φ−i are
± in ± n+1
disjoint, so there is nothing to check. Thus, the collection of charts (Ui ∩ S , φi ) i=1 is
a smooth atlas, so it defines a smooth structure on Sn , which we call its standard smooth
structure.
(3) Projective spaces: see Appendix A.
(4) Open submanifolds: If U is any open subset of Rn , then U is a topological n-manifold,
and the single chart (U, IdU ) determines a smooth structure on U .
More generally, let M be a smooth n-manifold and let U ⊆ M be an open subset.
Define an atlas on U by
Every point p ∈ U is contained in the domain of some chart (W, φ) for M . If we set
V = W ∩ U , then (V, φ|V ) is a chart in AU whose domain contains p. Therefore, U is
covered by the domains of the charts in AU , and it is easy to verify that AU is a smooth
atlas for U . In conclusion, any open subset of M is itself a smooth n-manifold in a natural
way. Endowed with this smooth structure, we call any open subset an open submanifold
of M .
We will encounter many more examples of smooth manifolds later in the course and
in the exercise sheets as well.
In the examples we have seen so far, we constructed a smooth manifold structure in
two stages: we started with a topological space and checked that it was a topological
manifold, and then we specified a smooth structure (by means of a smooth atlas due to
Proposition 1.8(a)). The following lemma shows how, given a set with suitable “charts”
that overlap smoothly, we can use these charts to define both a topology and a smooth
structure on the set.
Lemma 1.11 (Smooth manifold chart lemma). Let M be a set. Suppose that we are
given a collection {Uα } of subsets of M together with maps φα : Uα → Rn such that the
following properties are satisfied:
(ii) For each α and β, the sets φα (Uα ∩ Uβ ) and φβ (Uα ∩ Uβ ) are open in Rn .
φβ ◦ φ−1
α : φα (Uα ∩ Uβ ) → φβ (Uα ∩ Uβ )
is smooth.
(v) Whenever p, q ∈ M with p ̸= q, either there exists some Uα containing both p and q
or there exist disjoint sets Uα and Uβ with p ∈ Uα and q ∈ Uβ .
Then M has a unique smooth manifold structure such that each (Uα , φα ) is a smooth
chart.
Proof. For the details of the proof we refer to [Lee13, Lemma 1.35]. The basic idea is to
define a topology on M by taking all sets of the form φ−1 n
α (V ), where V ⊆ R is open, as
a basis.
CHAPTER 2
SMOOTH MAPS
The main reason for introducing smooth structures was to enable us to define smooth
functions on manifolds and smooth maps between manifolds. In Section 2.1 we carry out
this project. In Section 2.2 we introduce a powerful tool for blending together locally
defined smooth objects, called partitions of unity. They are used throughout smooth
manifold theory for building global smooth objects out of local ones. At the end of this
chapter we will give the first applications of partitions of unity.
Remark 2.2. Let M be a smooth manifold. The set C ∞ (M ) of all smooth real-valued
functions on M is an R-vector space: sums and constant multiples of smooth functions are
smooth. Note that C ∞ (M ) is infinite-dimensional, see Exercise 2.21. Moreover, pointwise
multiplication turns C ∞ (M ) into a commutative ring and an associative and commutative
R-algebra.
Definition 2.3. Let M be a topological manifold. Given a function f : M → Rk and
a chart (U, φ) for M , the function fb = f ◦ φ−1 : φ(U ) → Rk is called the coordinate
representation of f .
Let M be a smooth manifold and let f : M → Rk be a function on M . By definition,
f is smooth if and only if its coordinate representation is smooth in some smooth chart
around each point. According to [Exercise Sheet 3, Exercise 3], smooth functions have
smooth coordinate representations in every smooth chart; that is, f ◦ φ−1 : φ(U ) → Rk is
smooth for every smooth chart (U, φ) for M .
9
10 Chapter 2. Smooth Maps
F U
= ψ −1 ◦ (ψ ◦ F ◦ φ−1 ) ◦ φ : U → V
(a) For every p ∈ M there exist smooth charts (U, φ) containing p and (V, ψ) containing
F (p) such that U ∩F−1 (V ) is open in M and the composite map ψ◦F ◦φ−1 is smooth
from φ U ∩ F −1 (V ) to ψ(V ).
Section 2.1. Smooth Maps 11
(b) F is continuous and there exist smooth atlases (Uα , φα ) and (Vβ , ψβ ) for M
and N , respectively, such that for each α and β, ψβ ◦ F ◦ φ−1
α is a smooth map from
−1
φα Uα ∩ F (Vβ ) to ψβ (Vβ ).
Lemma 2.10 (Gluing lemma for smooth maps). Let M and N be smooth manifolds and
let {Uα }α∈A be an open cover of M . Suppose that for each α ∈ A we are given a smooth
map Fα : Uα → N such that the maps agree on overlaps: Fα |Uα ∩Uβ = Fβ |Uα ∩Uβ for all
α, β ∈ A. Then there exists a unique smooth map F : M → N such that F |Uα = Fα for
each α ∈ A.
(1) Write the map in smooth local coordinates and recognize its component functions
as compositions of smooth elementary functions.
(2) Exhibit the map as a composition of maps that are known to be smooth.
(3) Use some special-purpose theorem that applies to the particular case under consid-
eration.
We give below an example of a smooth map utilizing the first method above, and we will
encounter many more examples of smooth maps in the exercise sheets.
12 Chapter 2. Smooth Maps
Example 2.12. Consider the unit n-sphere Sn ⊆ Rn+1 with its standard smooth struc-
ture, see Example 1.10(2). The inclusion map ι : Sn ,→ Rn+1 is continuous (inclusion map
of topological spaces). It is a smooth map, because its coordinate representation with
respect to any of the graph coordinates of Example 1.3(2) is
ι(u1 , . . . , un ) = ι ◦ (φ± −1
1
b i ) (u , . . . , un )
p
= u1 , . . . , ui−1 , ± 1 − |u|2 , ui , . . . , un ,
↓
i-th
which is smooth on its domain (the set where |u|2 < 1).
Definition 2.13. Let M and N be smooth manifolds. A diffeomorphism from M to N
is a smooth bijective map M → N that has smooth inverse. We say that M and N are
diffeomorphic if there exists a diffeomorphism between them.
Example 2.14.
(1) Consider the maps
x
F : Bn → Rn , x 7→ p
1 − |x|2
and
y
G : Rn → Bn , y 7→ p .
1 + |y|2
These maps are smooth, and it is straightforward to check that they are inverses to each
other. Thus, they are both diffeomorphisms, so Bn is diffeomorphic to Rn .
(2) If M is any smooth manifold and if (U, φ) is any smooth coordinate chart on M , then
the coordinate map φ : U → φ(U ) ⊆ Rn is a diffeomorphism. Indeed, it has an identity
map as a coordinate representation.
Proposition 2.15 (Properties of diffeomorphisms).
(a) Every composition of diffeomorphisms is a diffeomorphism.
(b) Every finite product of diffeomorphisms between smooth manifolds is a diffeomor-
phism.
(c) Every diffeomorphism is a homeomorphism and an open map.
(d) The restriction of a diffeomorphism to an open submanifold is a diffeomorphism
onto its image.
(e) “Diffeomorphic” is an equivalence relation on the class of all smooth functions.
Proof. Exercise! (See also Proposition 4.9.)
Just as two topological spaces are considered to be “the same” if they are homeo-
morphic, two smooth manifolds are essentially indistinguishable if they are diffeomorphic.
The central concern of smooth manifold theory is the study of properties of smooth man-
ifolds that they preserved by diffeomorphisms. The dimension is one such property (cf.
p. 1):
Section 2.2. Partitions of Unity 13
Moreover,
Definition 2.18. Let M be a topological space and let X = (Xα )α∈A be an open cover
of M , indexed by a set A. A partition of unity subordinate to X is an indexed family
(ψα )α∈A of continuous functions ψα : M → R with the following properties:
(i) 0 ≤ ψα (x) ≤ 1, ∀ α ∈ A, ∀ x ∈ M .
(ii) supp ψα ⊆ Xα , ∀ α ∈ A.
(iii) The family of supports supp ψα α∈A is locally finite, i.e., every point p ∈ M has a
neighborhood Wp such that Wp ∩ supp ψα = ∅ for all but a finite number of α ∈ A.
X
(iv) ψα (x) = 1, ∀ x ∈ M .
α∈A
If now M is a smooth manifold, then a smooth partition of unity is one for which each
of the functions ψα is smooth.
Observe that, due to the local finiteness condition (iii), the sum in (iv) has only finitely
many non-zero terms in a neighborhood of each point, so there is no issue of convergence.
14 Chapter 2. Smooth Maps
Proof. For a detailed proof of the statement we refer to [Lee13, Theorem 2.23], see also
[Lee09, Theorem 1.73].
• 0 ≤ ψ(x) ≤ 1, ∀ x ∈ M ,
• ψ ≡ 1 on A, and
• supp ψ ⊆ U .
Proposition 2.20. Let M be a smooth manifold. For every closed subset A ⊆ M and any
open subset U ⊆ M containing A, there exists a smooth bump function for A supported
in U .
λ1 f1 + . . . + λk fk = 0 (2.1)
for some λi ∈ R. For each i ∈ {1, . . . , k}, pick x ∈ supp(fi ) such that fi (x) ̸= 0, and
note that fj (x) = 0 for every j ∈ {1, . . . , k} \ {i} by assumption. By evaluating (2.1)
Section 2.2. Partitions of Unity 15
at the chosen point x, we obtain λi fi (x) = 0, which implies λi = 0. This shows that
f1 , . . . , fk ∈ F are linearly independent, as claimed.
We will now show that there exists a countable collection of smooth functions on M
with non-empty disjoint supports, which in turn implies that the R-vector space C ∞ (M )
is infinite-dimensional, as desired. Fix a point p ∈ M and consider a smooth coordinate
chart (U, φ) containing p. In view of [Exercise Sheet 1, Exercise 1] and by further shrinking
U , we may assume that U is a smooth coordinate cube, i.e.,
are both smooth and agree at the point 0 where they overlap, but the continuous function
f : R → R, x 7→ |x| that they define is clearly not smooth at the origin. Our second
application of partitions of unity is an important result concerning the possibility of
extending smooth functions from closed sets.
Let M and N be smooth manifolds and let A ⊆ M be an arbitrary subset. We say
that a map F : A → N is smooth on A if it admits a smooth extension in a neighborhood
of each point; namely, if for every p ∈ A there exists an open subset W ⊆ M containing
p and a smooth map Fe : W → N whose restriction to W ∩ A agrees with F .
Lemma 2.22 (Extension lemma for smooth functions). Let M be a smooth manifold, let
A ⊆ M be a closed subset, and let f : A → Rk be a smooth function. For any open subset
U ⊆ M containing A, there exists a smooth function fe: M → Rk such that fe|A = f and
supp fe ⊆ U .
16 Chapter 2. Smooth Maps
Since the collection of supports {supp ψp }p∈A is locally finite, the sum actually has only
finitely many zero terms in a neighborhood of any point of M , and therefore defines a
smooth function. If x ∈ A, then ψ0 (x) = 0 by construction and fep (x) = f (x) for each p
such that ψp (x) ̸= 0 by (2.2), so
X X
fe(x) = ψp (x) fep (x) = ψ0 (x) + ψp (x) f (x) = f (x).
p∈A p∈A
where the equality in the middle is a property of locally finite collections, see [Lee13,
Lemma 1.13].
Comments:
(1) The conclusion of the extension lemma can be false if A is not closed; see [Lee13,
Exercise 2.27].
(2) The assumption in the extension lemma that the codomain is Rk , and not some other
manifold, is necessary: for other codomains, extensions can fail to exist for topological
reasons.
3 Closed subsets as level sets: The next result tells us that every closed subset of
a smooth manifold can be expressed as a level set of some smooth real-valued function.
This remarkable fact will not be used anywhere in these notes, so we omit its proof and
we refer to [Lee13, Theorem 2.29] for the details.
Theorem 2.23. Let M be a smooth manifold. If K is a closed subset of M , then there
exists a smooth non-negative function f : M → R such that f −1 (0) = K.
Exercise 2.24: Let A and B be disjoint closed subsets of a smooth manifold M . Show
that there exists f ∈ C ∞ (M ) such that 0 ≤ f (x) ≤ 1 for all x ∈ M , f −1 (0) = A and
f −1 (1) = B.
Section 2.2. Partitions of Unity 17
and observe that it is well-defined (that is, fA (x) + fB (x) ̸= 0 for all x ∈ M ) due to (2.3)
and since A ∩ B = ∅. Moreover, f is smooth as a quotient of smooth functions, and it
satisfies
0 ≤ f (x) ≤ 1 for all x ∈ M,
since fA and fB are non-negative. Finally, it follows from (2.3) that
va + wa := (v + w)a ,
λ va := (λ v)a .
The vectors ei |a , 1 ≤ i ≤ n, (where ei denotes the i-th standard basis vector of Rn ) are a
basis of Rna . In fact, Rna is essentially the same as Rn itself; the only reason why we add
the index a is so that the geometric tangent spaces Rna and Rnb at distinct points a and b
are disjoint sets.
Geometric tangent vectors provide a means of taking directional derivatives of func-
tions. For example, any geometric tangent vector v ∈ Rna yields a map
Dv a : C ∞ (Rn ) → R
d
f 7→ Dv a f = Dv f (a) = f (a + tv),
dt t=0
Dv a (f g) = f (a) Dv a g + g(a) Dv a f.
19
20 Chapter 3. The Tangent Bundle
If va = v i ei |a in terms of the standard basis, then by the chain rule Dv a f can be written
more concretely as
∂f
Dv a f = v i i (a). (3.1)
∂x
In particular, if va = ej |a , then
∂f
Dej a f = (a). (3.2)
∂xj
With this construction in mind, we make the following definition.
Definition 3.1. Given a ∈ Rn , a map w : C ∞ (Rn ) → R is called a derivation at a if it is
R-linear and satisfies the product rule:
(w1 + w2 )(f ) := w1 f + w2 f,
(λw)f := λ wf.
*1 *1
wf1 = w(f1 · f1 ) =
f1
(a) wf1 +
f1
(a) wf1 = 2wf1 ,
wf = w(cf1 ) = c wf1 = 0.
Dv a : C ∞ (Rn ) → R
d
f 7→ Dv a f = Dv f (a) = f (a + tv)
dt t=0
Φ : Rna → Ta Rn
v 7→ Dv a
is an R-linear isomorphism.
defined by
∂ ∂f
i
f := (a) , 1 ≤ i ≤ n,
∂x a ∂xi
form a basis of Ta Rn , and thus
dimR Ta Rn = n.
Proof.
(a) Easy to check (using calculus).
(b) Linearity: For every f ∈ C ∞ (Rn ) we have
= Df (a) · (λ1 v1 + λ2 v2 )
d d
= λ1 f (a + tv1 ) + λ2 f (a + tv2 )
dt t=0 dt t=0
(3.1) ∂
0 = Dv a xj === v i (xj ) = vj ,
∂xi x=a
∂xj ∂xi
where the last equality follows because ∂xi
= 0 for i ̸= j, and ∂xi
= 1. Hence, va = 0 ∈ Rna .
Surjectivity: Let w ∈ Ta Rn . Set v := v i ei |a ∈ Rna , where v i = w(xi ) ∈ R. We will
show that w = Φ(v) = Dv |a , To this end, let f ∈ C ∞ (Rn ). By Taylor’s theorem [Lee13,
22 Chapter 3. The Tangent Bundle
Note that each term in the last sum above is a product of two smooth functions of x that
vanish at x = a: one is (xi − ai ) and the other is (xj − aj ) · (integral). (The integral is a
smooth function of x by iterative application of [Lee13, Theorem C.14].) By Lemma 3.2(b)
the derivation w annihilates this entire sum. Thus, thanks to the R-linearity of w, we
obtain
n
0 X ∂f i i
(a) (x − a )
:
wf = wf (a) +
w i
i=1
∂x
n i
X ∂f *v
i *0
i
= (a) w(x ) − )
w(a
i=1
∂xi
n
X ∂f
= vi (a) = Dv |a f.
i=1
∂xi
∂
(c) By (3.2) we know that ∂xi
= Dei |a . Hence, (c) follows immediately from (b).
v
p Tp M
dFp : Tp M → TF (p) N,
v : C ∞ (U ) → R, f 7→ wfe,
where fe is any smooth function on M that agrees with f on V , see Lemma 2.22. By
Proposition 3.8, vf is independent of the choice of fe, so v is well-defined, and it is easy
to check that it is a derivation of C ∞ (U ) at p. For any g ∈ C ∞ (M ) we have
Proof. Fix p ∈ M and let (U, φ) be a smooth coordinate chart containing p. Since
φ: U → U b ⊆ Rn is a diffeomorphism by Example 2.14(2), dφp : Tp U → Tφ(p) U b is an
∼
isomorphism by Proposition 3.7(d). Since Proposition 3.9 guarantees that Tp U = Tp M
b∼
and Tφ(p) U = Tφ(p) Rn , it follows from Proposition 3.3(c) that
∂ ∂
,...,
∂x1 φ(p) ∂xn φ(p)
form a basis of Tφ(p) Rn . Therefore, the preimages of these vectors under the isomorphism
dφp , denoted by ∂x∂ i p , form a basis of Tp M . These vectors are characterized by
∂ −1 ∂ 3.7(d) −1 ∂
= (dφp ) === d(φ )φ(p) . (3.3)
∂xi p ∂xi φ(p) ∂xi φ(p)
∂ ∂ ∂ fb
i
f= (f ◦ φ−1 ) = (b
p),
∂x p ∂xi φ(p) ∂xi
The ordered basis ∂x∂ i p is called a coordinate basis for Tp M and the numbers (v i ) are
called the components of v with respect to the coordinate basis. If v is known, then its
26 Chapter 3. The Tangent Bundle
components can be easily computed from its action on the coordinate functions. For each
j ∈ {1, . . . , n}, the components of v are given by v j = v(xj ) (where we think of xj as a
smooth real-valued function on U ), because
j
j i ∂ j i ∂x
v(x ) = v (x ) = v (p) = v j .
∂xi p ∂xi
We now explore how differentials look in coordinates. We begin by considering the case
of a smooth map F : U ⊆ Rn → V ⊆ Rm between open subsets of Euclidean spaces. For
any p ∈ U we will determine the matrix of dFp : Tp Rn → TF (p) Rm in terms of the standard
coordinate bases. Denoting by (x1 , . . . , xn ) (respectively (y 1 , . . . , y m )) the coordinates in
the domain (respectively codomain), we use the chain rule to compute the action of dFp
on a typical basis vector as follows:
∂F j
∂ ∂ ∂f
dFp f= (f ◦ F ) = F (p) (p)
∂xi p ∂xi p ∂y j ∂xi
j
∂F ∂
= (p) j f.
∂xi ∂y F (p)
Thus,
∂F j
∂ ∂
dFp = (p) . (3.4)
∂xi p ∂xi ∂y j F (p)
that is, the Jacobian matrix of F at p, which is the matrix representation of the total
derivative DF (p) : Rn → Rm . Therefore, in this case, dFp : Tp Rn → TF (p) Rm corresponds
to the total derivative DF (p) : Rn → Rm , under the usual identification of Euclidean
spaces with their tangent spaces.
We now consider the more general case of a smooth map F : M → N between two
smooth manifolds. Choosing smooth coordinate charts (U, φ) for M containing p and
(V, ψ) for N containing F (p), we obtain the coordinate representation
Fb = ψ ◦ F ◦ φ−1 : φ U ∩ F −1 (V ) → ψ(V )
see Figure 3.3, and we also denote by pb = φ(p) the coordinate representation of p. By the
computation above, dFbpb is represented with respect to the standard coordinates bases by
Section 3.3. Computations in Local Coordinates 27
the Jacobian matrix of Fb at pb. Using the fact that F ◦ φ−1 = ψ −1 ◦ Fb, we compute
∂ dfn −1 ∂
dFp == dFp d(φ )pb
∂xi p ∂xi pb
Prop. −1 ∂
=== d(F ◦ φ )pb
3.7(b) | {z } ∂xi pb
ψ −1 ◦Fb
Prop. −1 ∂
=== d(ψ )Fb(bp) dFbpb
3.7(b) ∂xi pb
bj
(3.4) −1 ∂F ∂
== d(ψ )Fb(bp) (b
p) j
∂xi ∂y pb
Thus, dFp is represented in coordinate bases by the Jacobian matrix of (the coordinate
representation of) F . In fact, the definition of the differential was cooked up precisely in
order to give a coordinate-independent meaning to the Jacobian matrix.
Finally, suppose that U, φ = (xi ) and V, ψ = (e xi ) are two smooth charts on M
and that p ∈ U ∩ V . Any tangent vector at p can be represented with respect to either
coordinates basis ∂x∂ i p or ∂e∂xi p . How are the two representations related?
ψ ◦ φ−1 (x) = x
e1 (x), . . . , x
en (x) .
Here we are indulging in a typical abuse of notation: in the expression x ei (x) we think
ei as a coordinate function (whose domain is an open subset of M , identified with an
of x
open subset of Rn ), but we think of x as representing a point (in this case, in φ(U ∩ V )).
By (3.4) we have
xj
−1 ∂ ∂e ∂
d(ψ ◦ φ )φ(p) = φ(p) .
∂xi φ(p) ∂xi xj
∂e ψ(p)
28 Chapter 3. The Tangent Bundle
xj
−1 ∂e ∂
= d(ψ )ψ(p) φ(p)
∂xi xj
∂e ψ(p)
(3.3) xj
∂e ∂
===== i
φ(p) . (3.6)
linearity ∂x |{z} ∂e xj p
= pb
(This formula looks exactly the same as the chain rule for partial derivatives in Rn .)
Applying this to the components of a vector
∂ ∂
v = vi = vej ,
∂xi p xj
∂e p
We usually write an element of this disjoint union as an ordered pair (p, v) with p ∈ M
and v ∈ Tp M ; we sometimes also write vp for (p, v). The tangent bundle comes equipped
with a natural projection map π : TM → M , which sends each vector in Tp M to the
point p at which is tangent: (p, v) 7→ p.
For example, when M = Rn , using Proposition 3.3, we see that the tangent bundle of
n
R can be canonically identified with the disjoint union of its geometric tangent spaces,
which in turn is just the Cartesian product of Rn with itself:
Tp Rn ∼
G G G
T(Rn ) = = Rnp = {p} × Rn = Rn × Rn .
p∈Rn p∈Rn p∈Rn
An element of this Cartesian product can be thought of as representing either the geo-
metric tangent vector vp or the derivation Dv |p defined in Proposition 3.3(a). In general,
however, the tangent bundle of a smooth manifold cannot be identified in a natural way
with a Cartesian product, because there is no canonical way to identify tangent spaces at
distinct points with each other.
Section 3.4. The Tangent Bundle 29
The tangent bundle of a smooth manifold can be thought of simply as a disjoint union
of vector spaces, but it is much more than that. The next proposition shows that the
tangent bundle of a smooth manifold can be considered as a smooth manifold in its own
right. For its proof we will use Lemma 1.11.
Proposition 3.12. For any smooth n-manifold M , the tangent bundle TM has a natural
topology and smooth structure that make it into a smooth (2n)-manifold. With respect to
this structure, the projection π : TM → M is smooth.
Proof. We begin by defining the maps that will become our smooth charts. Given any
smooth chart (U, φ) for M , observe that π −1 (U ) is the set of all tangent vectors to M at
all points of U . Denote by (x1 , . . . , xn ) the coordinate functions of φ, and define a map
e : π −1 (U ) → R2n ,
φ
i ∂ 1 n 1 n
φ
e v = x (p), . . . , x (p), v , . . . , v . (3.8)
∂xi p
Its image is the set φ(U ) × Rn , which is an open subset of R2n . It is a bijection onto its
image, because its inverse can be explicitly written as
∂
e−1 x1 , . . . , xn , v 1 , . . . , v n = v i i
φ .
∂x φ−1 (x)
Now, suppose that we are given two smooth charts (U, φ) and (V, ψ) for M , and
consider the corresponding “charts” π −1 (U ), φ e and π −1 (V ), ψe for TM . The sets
e π −1 (U ) ∩ π −1 (V ) = φ(U ∩ V ) × Rn
φ
and
ψe π −1 (U ) ∩ π −1 (V ) = ψ(U ∩ V ) × Rn
xj ∂
(3.7) ∂e
== ψ v i i
e
∂x ∂e xj φ−1 (x)
x1
∂e xn
∂e
= xe1 , . . . , x
en , i v i , . . . , i v i ,
∂x ∂x
which is clearly smooth.
Choosing a countable cover {Ui } of M by smooth coordinate domains, we obtain a
countable cover of TM by coordinate domains {π −1 (Ui )} satisfying conditions (i)-(iv) of
Lemma 1.11. To check the Hausdorff condition (v), just note that any two points in the
30 Chapter 3. The Tangent Bundle
same fiber of π lie in one chart, while if (p, v) and (q, w) lie in different fibers, there exist
disjoint smooth coordinate domains U and V for M such that p ∈ U and q ∈ V , and then
π −1 (U ) and π −1 (V ) are disjoint coordinate neighborhoods containing (p, v) and (q, w),
respectively. This completes the proof of the first part of the statement.
π : TM → M is smooth, note that with respect
Finally, to check that to charts (U, φ)
for M and π −1 (U ), φ e for TM , its coordinate representation φ◦π ◦ φ e−1 is π
b(x, v) = x.
Proof. If (U, φ) is a global smooth chart for M , then φ is, in particular, a diffeomor-
phism from U = M to an open subset U b ⊆ Rn , see Example 2.14(2). The proof of
Proposition 3.12 showed that the natural coordinate chart φe is a bijection from TM to
n
U × R , and the smooth structure on TM is defined essentially by declaring φ
b e to be
diffeomorphism.
Comment: In general, the tangent bundle is not globally diffeomorphic (or ever homeo-
morphic) to a product of the manifold with Rn .
(a) d(G ◦ F ) = dG ◦ dF : TM → TP .
where d/dt|t0 is the standard coordinate basis vector in Tt0 R. Other common nota-
tions for the velocity vector are:
dγ
γ̇(t0 ) and (t0 )
dt
Assume that M , γ and t0 are as in Definition 3.15. The tangent vector γ ′ (t0 ) acts on
functions f ∈ C ∞ (M ) by
′ d d
γ (t0 )f = dγ f= (f ◦ γ) = (f ◦ γ)′ (t0 ).
dt t=t0 dt t=t0
In other words, γ ′ (t0 ) is the derivation at γ(t0 ) obtained by taking the derivative of a
function along γ. (If t0 is an endpoint of the interval J ⊆ R, this still holds, provided that
we interpret the derivative with respect to t as a one-sided derivative, or equivalently as
the derivative of any smooth extension of f ◦ γ to an open subset of R.)
Now, let (U, φ) be a smooth chart for M with coordinate functions (xi ). If γ(t0 ) ∈ U ,
then we can write the coordinate representation of γ as
at least for t ∈ J sufficiently close to t0 ∈ J, and then the coordinate formula for the
differential (3.5) yields
dγ i ∂
γ ′ (t0 ) = (t0 ) i .
dt ∂x γ(t0 )
This means that γ ′ (t0 ) is given by essentially the same formula as it would be in Euclidean
space: it is the tangent vector whose components in a coordinate basis are the derivatives
of the component functions of γ.
The next proposition shows that every tangent vector on a manifold is the velocity
vector of some curve. This gives a different and somewhat more geometric way to think
about the tangent bundle: it is just the set of all velocity vectors of smooth curves in M .
Proposition 3.16 (Tangent vectors as velocity vectors of smooth curves). Let M be
a smooth manifold. If p ∈ M , then for any v ∈ Tp M there exists a smooth curve
γ : (−ε, ε) → M such that γ(0) = p and γ ′ (0) = v.
Proof. See [Exercise Sheet 4, Exercise 5].
32 Chapter 3. The Tangent Bundle
Because the differential of a smooth map is supposed to represent the “best linear ap-
proximation” to the map near a given point, we can learn a great deal about a map by
studying linear-algebraic properties of its differential. The most essential property of the
differential is its rank (the dimension of its image). In this chapter we study the ways
in which geometric properties of smooth maps can be detected from their differentials.
The maps for which differentials give good local models turn out to be the ones whose
differentials have constant rank.
(b) a smooth submersion if its differential is surjective at each point or, equivalently, if
rk F = dim N ; and
33
34 Chapter 4. Maps of Constant Rank
and [Lee13, Chapter 4, Smooth Covering Maps]). We will see that smooth immersions
and submersions behave locally like injective and surjective linear maps, respectively (see
Theorem 4.11).
Proof. If we choose any smooth coordinates for M near p and for N near F (p), either
hypothesis means that the Jacobian matrix of F in coordinates has full rank at p ∈ M .
By [Exercise Sheet 2, Exercise 3] we know that the set of n × m matrices of full rank is
an open subset of M(n × m, R) (where m = dim M and n = dim N ), so by continuity, the
Jacobian of F (in coordinates) has full rank in some neighborhood of p ∈ M .
Example 4.4.
(2) If M is a smooth manifold and its tangent bundle TM is given the smooth manifold
structure described in Proposition 3.12, then the projection π : TM → M is a smooth
submersion. Indeed, we showed that with respect to any smooth local coordinates (xi ) on
an open subset U ⊆ M and the corresponding natural coordinates (xi , v i ) on π −1 (U ) ⊆
TM , the coordinate representation of π is π
b(x, v) = x, and thus
Jπb = Iddim M O ,
(3) If M is a smooth manifold and U ⊆ M is an open subset, then the inclusion map
U ,→ M is a smooth embedding, see Proposition 3.9.
Example 4.5.
Its image is a set that looks like a figure-eight in the plane, sometimes called a lemniscate,
see Figure 4.1. It is the locus of points (x, y) ∈ R2 such that x2 = 4y 2 (1 − y 2 ), as one can
easily check.
−1 0 1
−1
(a) an open map if for every open subset U of X, the image F (U ) is an open subset
of Y ;
(b) a closed map if for every closed subset C of X, the image F (C) is a closed subset
of Y ;
(c) a proper map if for every compact subset K ⊆ Y , the preimage F −1 (K) is a
compact subset of X.
36 Chapter 4. Maps of Constant Rank
The following proposition gives a few simple sufficient criteria for an injective immer-
sion to be an embedding.
(c) M is compact.
Proof. We first prove the following three claims, which will be then used crucially in the
proof of the statement.
Claim 1 : Let F : X → Y be a continuous map between topological spaces that is either
open or closed. If F is injective, then it is a topological embedding.
Proof : Assume that F is open and injective. Then F : X → F (X) is bijective, so
F −1 : F (X) → F exists. If U ⊆ X is open, then (F −1 )−1 (U ) = F (U ) is open in Y by
hypothesis, and therefore also open in F (X) by definition of the subspace topology on
F (X). Hence, F −1 is continuous, so that F is a topological embedding.
The proof of the assertion is similar when F is closed and injective.
Claim 2 (Closed map lemma): Let X be a compact space, let Y be a Hausdorff space,
and let F : X → Y be a continuous map. Then F is a closed map.
Proof : Let K ⊆ X be a closed subset. Since X is compact, K is also compact, and
since F is continuous, F (K) is also compact. Since Y is Hausdorff, F (K) ⊆ Y is a closed
subset. Thus, F is a closed map.
Claim 3 : Let X be a topological space and let Y be a locally compact Hausdorff space.
Then every proper continuous map F : X → Y is closed.
Proof : Let K ⊆ X be a closed subset. To show that F (K) ⊆ Y is closed, we will
show that its complement is open. Let y ∈ Y \ F (K). Since Y is locally compact, y
has an open neighborhood V with compact closure V , and since F is proper, F −1 ( V ) is
compact. Set E := K ∩ F −1 ( V ) and note that E is a compact set. Since F is continuous,
F (E) is also compact, and since Y is Hausdorff, F (E) is a closed subset of Y . Set
U := V \ F (E) = V ∩ Y \ F (E) and observe that U is open neighborhood of y, which
is disjoint from F (K). Hence, Y \ F (E) is open, which implies that F (K) is closed.
We are now ready to prove the statement.
(a) By assumption and by Claim 1, F is a topological embedding. Since it is also a smooth
immersion by assumption, we conclude that F is a smooth embedding.
(b) By assumption and by Claim 3, F is a closed map, so it is a smooth embedding by
(a).
(c) By assumption and by Claim 2, F is a closed map, so it is a smooth embedding by
(a).
Section 4.2. Local Diffeomorphisms 37
(b) Every finite product of local diffeomorphisms between smooth manifolds is a local
diffeomorphism.
(g) A map between smooth manifolds is a local diffeomorphism if and only if in a neigh-
borhood of each point of its domain, it has a coordinate representation that is a local
diffeomorphism.
Proof. Exercise! (See also Proposition 2.15.)
Proposition 4.10. Let M and N be smooth manifolds and let F : M → N be a map.
The following statements hold:
38 Chapter 4. Maps of Constant Rank
(a) F is a local diffeomorphism if and only if it is both a smooth immersion and a smooth
submersion.
Theorem 4.11 (Rank theorem). Let M and N be smooth manifolds of dimension m and
n, respectively, and let F : M → N be a smooth map of constant rank r. For each p ∈ M
there exist smooth charts (U, φ) for M centered at p and (V, ψ) for N centered at F (p)
such that F (U ) ⊆ V , in which F has a coordinate representation of the form
Proof. Since the theorem is local, after choosing smooth coordinates we can replace M
and N by open subsets U ⊆ Rm and V ⊆ Rn . The fact that DF (p) has rank r implies
that its matrix has some r × r submatrix with non-zero determinant. By reordering
i
the coordinates, we may assume that it is the upper left submatrix ∂F ∂xj
(p) for i, j ∈
{1, . . . , r}. We relabel the standard coordinates as
and
(v, w) = (v 1 , . . . , v r , w1 , . . . , wn−r ) in Rn .
By initial translation of the coordinates, without loss of generality we may assume that
p = (0, 0) and F (p) = (0, 0). If we write
F (x, y) = Q(x, y), R(x, y)
for some smooth maps Q : U → Rr and R : U → Rn−r , then our hypothesis is that the
∂Qi
matrix ∂xj is non-singular at (0, 0).
Section 4.3. The Rank Theorem 39
for some smooth functions A : Ue0 → Rr and B : Ue0 → Rn−r , we see that
(x, y) = φ ◦ φ−1 (x, y) = φ A(x, y), B(x, y) = Q A(x, y), B(x, y) , B(x, y) .
Comparing y components shows that B(x, y) = y, and therefore φ−1 has the form
φ−1 (x, y) = A(x, y), y . Comparing
now x components and taking this into account
also shows that Q A(x, y), y = x, and therefore F ◦ φ−1 has the form
δji
O
D(F ◦ φ−1 )(x, y) = ∂ R .
ei ei
∂R
(x, y) (x, y)
∂xj ∂y j
Since composing with a diffeomorphism does not change the rank of a map, the above
matrix has rank r everywhere in U e0 . The first r columns are obviously linearly indepen-
ei
dent, so the rank can be r only if ∂∂yRj vanish identically on U e0 , which implies that Re is
actually independent of (y 1 , . . . , y m−r ). (This is one reason why we arranged for U e0 to be
a cube.) Thus, if we let S(x) = R(x, e 0), then we have
and note that V0 is a neighborhood of (0, 0). Since Ue0 ∋ (0, 0) = φ(0, 0) is a cube and
−1 −1
F ◦ φ has the form (4.1), it follows that F ◦ φ (U e0 ) ⊆ V0 (because (v, w) ∈ U
e0 ⇒
−1
F ◦ φ (v, w) = v, S(v) ∈ V by construction and (v, 0) ∈ U e0 by the form of U
e0 ), and
hence F (U0 ) ⊆ V0 . Define the function
ψ : V0 → Rn , ψ(v, w) = v, w − S(v) .
This is an open map and a diffeomorphism onto its image, because its inverse is given
−1
explicitly by ψ (s, t) = s, t + S(s) . Thus, (V0 , ψ) is a smooth chart. It follows now
immediately from (4.1) that
(a) For each p ∈ M there exists smooth charts containing p and F (p) in which the
coordinate representation of F is linear.
Proof.
(b) ⇒ (a): Follows immediately from the rank theorem.
(a) ⇒ (b): Since every linear map has constant rank, it follows that the rank of F is
constant in a neighborhood of each point, and thus by connectedness it is constant on all
of M .
The rank theorem is a purely local statement. However, it has the following powerful
global consequence.
and thus F |U1 is injective. Now, consider a precompact neighborhood U of p such that
U ⊆ U1 . The restriction of F to U is an injective continuous map with compact domain
and Hausdorff codomain, so it is a topological embedding according to Claims 1 and 2
from the proof of Proposition 4.6. Since any restriction of a topological embedding is again
a topological embedding, F |U is both a topological embedding and a smooth immersion,
so it is a smooth embedding.
Cε := x |xi | < ε, 1 ≤ i ≤ m
Cε′ := y |y i | < ε, 1 ≤ i ≤ n .
σ(x1 , . . . , xn ) = (x1 , . . . , xn , 0, . . . , 0)
M
F ◦π
π
N F
P
⇝ [Exercise Sheet 7, Exercise 4] explains the sense in which the above property is
“characteristic”.
⇝ [Exercise Sheet 7, Exercise 5] shows that the converse of the Theorem 4.17 is false.
N P
Fe
M
π1 π2
N1 F
N2
Many familiar manifolds appear naturally as subsets of other manifolds. We have already
seen that open subsets of smooth manifolds can be viewed as smooth manifolds in their
own right. However, there are many interesting examples beyond the open ones. In this
chapter we explore smooth submanifolds, which are smooth manifolds that are subsets of
other smooth manifolds.
45
46 Chapter 5. Submanifolds
Proof. If we give S the subspace topology that it inherits from M , then the assumption
that F is an embedding means that F can be considered as a homeomorphism from N
onto S, and thus S is a topological manifold. We now give S a smooth structure by
taking the smooth charts to be those of the form F (U ), φ ◦ F −1 , where (U, φ) is a
smooth chart for N . Note that the smooth compatibility of these charts follows from the
smooth compatibility of the corresponding charts for N . With this smooth structure on
S, the map F is a diffeomorphism onto its image (essentially by definition), and this is
obviously the only smooth structure with this property. The inclusion map ι : S ,→ M is
equal to the composition of a diffeomorphism followed by a smooth embedding
F −1 F
S N M,
so it is a smooth embedding itself by [Exercise Sheet 6, Exercise 1(a)(iii)].
Since every embedded submanifold is the image of a smooth embedding (namely its
own inclusion map), Proposition 5.3 shows that embedded submanifolds are exactly the
images of smooth embeddings of smooth manifolds.
(a) Given an open subset U ⊆ Rn and an integer k ∈ {0, . . . , n}, a k-dimensional slice of
U (or simply a k-slice) is any subset of the form
(b) Let M be a smooth manifold and let (U, φ) be a smooth chart for M . If S is a subset
of U such that φ(S) is a k-slice of φ(U ) ⊆ Rn , then we say that S is a k-slice of U .
Theorem 5.6 (Local slice criterion for embedded submanifolds). Let M be a smooth
n-manifold. If S is an embedded k-dimensional submanifold of M , then S satisfies the
local k-slice condition. Conversely, if S ⊆ M is a subset that satisfies the local k-slice
condition, then with the subspace topology, S is a topological manifold of dimension k, and
it has a smooth structure making it into a k-dimensional embedded submanifold of M .
Proof.
“⇒”: Fix p ∈ S. Since the inclusion map ι : S ,→ M is in particular a smooth immersion,
by the rank theorem there are smooth charts (U, φ) for S (in its given smooth manifold
structure) and (V, ψ) for M , both centered at p, in which the inclusion map ιU : U ,→ V
has the coordinate representation
Now, choose 0 < ε ≪ 0 so that both U and V contain coordinate balls U0 ⊆ U and V0 ⊆ V
of radius ε > 0 centered at p. It follows that U0 ∼ = ι(U0 ) is exactly a single slice in V0
(using the above local description). Since S ⊆ M has the subspace topology and since U0
is open in S, there is an open subset W ⊆ M such that U0 = W ∩S. Setting V1 := W ∩V0 ,
we obtain a smooth chart (V1 , ψ|V1 ) for M containing p such that V1 ∩ S = U0 ∩ V0 = U0 ,
which is a single slice of V1 (as U0 is a single slice of V0 ).
“⇐”: With the subspace topology, S is Hausdorff and second-countable, because both
properties are inherited by subspaces. To show that S is locally Euclidean, we construct
an atlas. The basic idea of the construction is that if (x1 , . . . , xn ) are slice coordinates for
S in M , then we can use (x1 , . . . , xk ) as local coordinates for S.
48 Chapter 5. Submanifolds
Let π : Rn → Rk be the projection onto the first k-coordinates. Let (U, φ) be a slice chart
for S in M , and define
V = U ∩ S, Vb = (π ◦ φ)(V ), ψ = (π ◦ φ)|V : V → Vb .
By definition of slice charts, φ(V ) is the intersection of φ(U ) with a certain k-slice A ⊆ Rn
defined by setting xk+1 = ck+1 , . . . , xn = cn , and therefore φ(V ) is open in A. Since
π|A : A → Rk is a diffeomorphism, it follows that Vb is open in Rk . Moreover, ψ is a
homeomorphism, because it has a continuous inverse given by (φ−1 ◦ j)|Vb , where
ψ ′ ◦ ψ −1 = π ◦ φ′ ◦ φ−1 ◦ j,
which is smooth by Proposition 2.11(d) as a composite of four smooth maps. Hence, the
atlas we have constructed is actually a smooth atlas (see Remark 1.5), and it defines a
smooth structure on S by Proposition 1.8(a). In terms of a slice chart (U, φ) for S in M
and the corresponding chart (V, ψ) for S, the inclusion map ι : S ,→ M has a coordinate
representation of the form
which is a smooth immersion. Since the inclusion map is also a topological embedding,
we are done.
Notice that the local slice condition for S ⊆ M is a condition on the subset S only;
it does not presuppose any particular topology or smooth structure on S. According to
[Exercise Sheet 9, Exercise 1], the smooth manifold structure constructed in Theorem 5.6
is the unique one in which S can be considered as a submanifold, so a subset satisfying
the local slice condition is an embedded submanifold in only one way.
• A point c ∈ N is called a regular value of Φ if every point of the level set Φ−1 (c) is
a regular point; otherwise, we say that c is a critical value of Φ. (In particular, if
Φ−1 (c) = ∅, then c is a regular value.)
• A level set Φ−1 (c) is called a regular level set if c is a regular value of Φ.
(3) By Lemma 4.3, the set of regular points of Φ is an open subset of M (but may well
be empty).
Θ : R2 → R, (x, y) 7→ x2 − y,
Φ : R2 → R, (x, y) 7→ x2 − y 2 ,
Ψ : R2 → R, (x, y) 7→ x2 − y 3 .
Although the zero set Θ−1 (0) of Θ is an embedded submanifold of R2 , it will be shown in
[Exercise Sheet 8, Exercise 3(b)] and [Exercise Sheet 9, Exercise 5(c)], respectively, that
neither the zero set Φ−1 (0) of Φ nor the zero set Ψ−1 (0) of Ψ is an embedded submanifold
of R2 . Hence, it is fairly easy to find level sets of smooth functions that are not smooth
submanifolds. In fact, without further assumptions on the smooth function, the situation
is about as bad as could be imagined; namely, according to Theorem 2.23, every closed
subset of M can be expressed as the zero set of a smooth non-negative real-valued function.
However, using the rank theorem, we can prove the following result:
Proof. Set m = dim M , n = dim N and k = m − r. Pick c ∈ N and set S = Φ−1 (c). By
the rank theorem, for each p ∈ S there are smooth charts (U, φ) centered at p and (V, ψ)
centered at c = Φ(p) in which Φ has a coordinate representation of the form
e 1 , . . . , xm ) = (x1 , . . . , xr , 0, . . . , 0).
Φ(x
Corollary 5.10 (Regular level set theorem). Every regular level set of a smooth map
between smooth manifolds is a properly embedded submanifold of the domain whose codi-
mension is equal to the dimension of the codomain.
Proof. Let Φ : M → N be a smooth map and let c ∈ N be a regular value of Φ. By
Remark 5.8(3) the set
U = p ∈ M | rk(dΦp ) = dim N ⊆ M
Φ−1 (c) ,→ U ,→ M
Proof. The proof is very similar to that of Proposition 5.3, except that now we also have
to define the topology on S.
We give S a topology by declaring a subset U ⊆ S to be open if and only if F −1 (U ) ⊆ N
is open, and then we give it a smooth structure by taking the smooth charts to be those
of the form F (U ), φ ◦ F −1 , where (U, φ) is a smooth chart for N . (As in the proof
of Proposition 5.3, the smooth compatibility of these charts follows from the smooth
compatibility of the corresponding charts for N .) With this topology and smooth structure
on S, the map F is a diffeomorphism onto its image, and these are the only topology and
smooth structure on S with this property. The inclusion map ι : S ,→ M can be written
as the composition
F −1 F
S N M,
where the first map is a diffeomorphism and the second map is a smooth immersion, so ι
is itself a smooth immersion by [Exercise Sheet 6, Exercise 1(a)(ii)].
Example 5.14. The figure-eight curve (lemniscate) from Example 4.5(2) is the image of
the injective smooth immersion
Remark 5.15. In general, smooth (immersed) submanifolds can be closed without being
embedded (as is, for example, the figure-eight curve from Example 5.14) or embedded
without being closed (as is, for example, the open unit ball Bn in Rn ).
The following observation is sometimes useful when thinking about the topology of an
immersed submanifold.
Tp S = v ∈ Tp M | vf = 0 whenever f ∈ C ∞ (M ) with f |S = 0 .
U ∩ S = (x1 , . . . , xn ) ∈ U | xk+1 = . . . = xn = 0 ,
and (x1 , . . . , xk ) are coordinates for U ∩ S. Since the inclusion map ι : U ∩ S ,→ M has
the coordinate representation
ι(x1 , . . . , xk ) = (x1 , . . . , xk , 0, . . . , 0)
∂ ∂
,..., .
∂x1 p ∂xk p
Thus, v ∈ Tp S, as desired.
54 Chapter 5. Submanifolds
Here is an example of how this can be done; another one is given in [Exercise Sheet 9,
Exercise 5(c)].
S = (x, y) ∈ R2 | y = |x| ⊆ R2 .
In Section 3.4 we saw that the tangent bundle of a smooth manifold has a natural structure
as a smooth manifold in its own right. The natural coordinates we constructed on TM
make it look locally like the Cartesian product of an open subset of M n with Rn . This
kind of structure arises quite frequently – a collection of vector spaces, one for each point
in M , glued together in a way that looks locally like the Cartesian product of an open
subset of M n with Rn , but globally may be “twisted”. Such structures are called vector
bundles and will be briefly discussed in this chapter.
There is a deep and extensive body of theory about vector bundles on manifolds,
which we will not touch in this course. We introduce them primarily in order to have
a convenient language for talking about the tangent bundle and structures like it; see
Chapter 7 and Chapter 8.
(i) For each p ∈ M , the fiber Ep = π −1 (p) over p is endowed with the structure of a
k-dimensional R-vector space.
The space E is called the total space of the bundle, M is called its base, and π is called
its projection.
55
56 Chapter 6. Vector Bundles
If M and E are smooth manifolds, π is a smooth map, and the local trivializations
can be chosen to be diffeomorphisms, then E is called a smooth vector bundle over M .
In this case, any local trivialization that is a diffeomorphism onto its image is called a
smooth local trivialization.
Definition 6.2. If there exists a local trivialization of E over all of M , called a global
trivialization of E, then E is called a trivial bundle. In this case, E itself is homeomorphic
to the product space M × Rk .
If E → M is a smooth vector bundle that admits a smooth global trivialization, then
we say that E is smoothly trivial . In this case, E is diffeomorphic to M × Rk , not just
homeomorphic (as in previous case).
Example 6.3.
(1) Product bundles: Given any topological space M , the product space E = M × Rk
with the map π = πM : M × Rk → M as its projection is a rank-k vector bundle over
M . Any such bundle, called a product bundle, is clearly trivial (with the identity map
Φ = IdE : M × Rk → M × Rk as a global trivialization). If M is a smooth manifold, then
the (smooth) product bundle M × Rk is smoothly trivial.
Proposition 6.4 (The tangent bundle as a vector bundle). Let M be a smooth n-manifold
and let TM be its tangent bundle. With its standard projection map π : TM → M ,
its natural vector space structure on each fiber, and the topology and smooth structure
constructed in Proposition 3.12, π : TM → M is a smooth vector bundle of rank n over
M.
Proof. Given any smooth chart (U, φ) for M with coordinate functions (xi ), define a map
∂
Φ : π −1 (U ) → U × Rn , v i 7→ p, (v 1 , . . . , v n ) .
∂xi p
Any bundle that is not trivial requires more than one local trivialization. The next
lemma shows that the composition of two smooth local trivializations has a simple form
where they overlap.
Section 6.1. Vector Bundles 57
Lemma 6.5. Let π : E → M be a smooth vector bundle of rank k over M . Suppose that
Φ : π −1 (U ) → U × Rk and Ψ : π −1 (V ) → V × Rk
are two smooth local trivializations of E with U ∩ V ̸= ∅. Then there exists a smooth map
τ : U ∩ V → GL(k, R),
called the transition function between the smooth local trivializations Φ and Ψ, such that
the composition
Φ ◦ Ψ−1 : (U ∩ V ) × Rk → (U ∩ V ) × Rk
has the form
Φ ◦ Ψ−1 (p, v) = p, τ (p) · v .
π
π1 π1 =πU ∩V
U ∩V
and thus π1 ◦ (Φ ◦ Ψ−1 ) = π1 , which means that
Φα ◦ Φ−1 k
β : (Uα ∩ Uβ ) × R → (Uα ∩ Uβ ) × R
k
Then E has a unique topology and smooth structure making it into a smooth manifold
and a smooth vector bundle of rank k over M , with π as projection and (Uα , Φα ) α∈A as
smooth local trivializations.
Proof. For the details of the proof, which relies essentially on Lemma 1.11, we refer to
[Lee13, Lemma 10.6].
Here are some examples showing how the vector bundle chart lemma can be used to
construct new vector bundles from old ones.
Example 6.7 (Whitney sums). Let M be a smooth manifold. Let π ′ : E ′ → M and
π ′′ : E ′′ → M be two smooth vector bundles of ranks k ′ and k ′′ , respectively, over M . We
will construct a new smooth vector bundle π : E → M of rank k ′ + k ′′ over M , denoted
by E ′ ⊕ E ′′ and called the Whitney sum of E ′ and E ′′ , whose fiber over each p ∈ M is
the direct sum Ep := Ep′ ⊕ Ep′′ , which is a (k ′ + k ′′ )-dimensional R-vector space. For each
p ∈ M choose a small enough neighborhood U of p so that there exist local trivializations
(U, Φ′ ) of E ′ and (U, Φ′′ ) of E ′′ , and define the map
′ ′′
Φ : π −1 (U ) → U × Rk +k , Φ(v ′ , v ′′ ) := π ′ (v ′ ), πRk′ ◦ Φ′ (v ′ ), πRk′′ ◦ Φ′′ (v ′′ ) .
e ′ and U e ′′ .
Suppose that we are given another such pair of local trivializations U e, Φ e, Φ
Let τ ′ : U ∩ U e → GL(k ′ , R) and τ ′′ : U ∩ U e → GL(k ′′ , R) be the corresponding transition
functions. Then the transition function for E ′ ⊕ E ′′ has the form
e ◦ Φ−1 p, (v ′ , v ′′ ) = p, τ (p)(v ′ , v ′′ ) ,
Φ
Since this depends smoothly on p, it follows from Lemma 6.6 that E ′ ⊕ E ′′ is a smooth
vector bundle over M .
Example 6.8 (Restriction of a vector bundle). Let π : E → M be a rank-k vector bundle
S let S ⊆ M be any subset. We define the restriction of E to S to be
and
−1
the set E|S :=
E
p∈S p , with the projection E| S → S obtained by restricting π. If Φ : π (U ) → U × Rk
is a local trivialization of E over U ⊆ M , it restricts to a bijective map from (π|S )−1 (U ∩S)
to (U ∩ S) × Rk , and it is easy to check that these form local trivializations for a vector
bundle structure on E|S .
Section 6.2. Sections of a Vector Bundle 59
σ(U )
π
( σ
)
U M
If E → M is a smooth vector bundle, then the set of all smooth global sections of E
is an R-vector space under pointwise addition and scalar multiplication:
This vector space is usually denoted by Γ(E) (but for particular smooth vector bundles we
often introduce specialized notation for their spaces of global sections) and it is infinite-
dimensional, see [Exercise Sheet 10, Exercise 3] and Exercise 2.21. Moreover, smooth
sections of E → M can be multiplied by smooth real-valued functions: If f ∈ C ∞ (M )
and σ ∈ Γ(E), then we obtain a new smooth section f σ ∈ Γ(E) defined by
∈
R Ep
⇝ The various claims made above are proved in [Exercise Sheet 10, Exercise 3(b)].
⇝ The global sections of a product bundle are discussed in [Exercise Sheet 10, Exercise
3(c)].
Lemma 6.10 (Extension lemma for smooth vector bundles). Let π : E → M be a smooth
vector bundle. Let A ⊆ M be a closed subset and let σ : A → E be a section of E|A that
is smooth in the sense that σ extends to a smooth local section of E in a neighborhood of
each point. Then for each open subset U ⊆ M containing A, there exists a smooth global
e ∈ Γ(E) such that σ
section σ e = {p ∈ M | σ
e|A = σ and supp σ e(p) ̸= 0} ⊆ U .
⇝ For two applications of Lemma 6.10 we refer to [Exercise Sheet 10, Exercises 3(d)
and 4(c)].
eei : M → E, p 7→ (p, ei ).
⇝ For the correspondence between smooth local frames and smooth local trivializations
see [Exercise Sheet 10, Exercise 5] (which also settles the question of the existence
of smooth local frames).
⇝ For the completion of smooth local frames for smooth vector bundles see [Exercise
Sheet 10, Exercise 4].
We conclude this section with the important observation that smoothness of sections
of smooth vector bundles can be characterized in terms of local frames.
Assume that (σi ) is a smooth local frame for a smooth vector bundle E → M over
some open subset U ⊆ M . If τ : M → E is a rough section, then the value of τ at an
arbitrary point p ∈ U can be written as
Proof. We prove the statement in the smooth case; the other case can be treated similarly.
Let Φ : π −1 (U ) → U × Rk be the smooth local trivialization associated with the smooth
local frame (σi ), see [Exercise Sheet 10, Exercise 5(b)]. Since Φ is a diffeomorphism, τ is
smooth on U if and only if Φ ◦ τ is smooth on U . By the construction of Φ in [Exercise
Sheet 10, Exercise 5(b)] we know that
Φ ◦ τ (p) = Φ τ i (p)σi (p) = p, τ 1 (p), . . . , τ k (p) ,
Note that Proposition 6.14 applies equally well to local sections, since a local section
of E over an open subset V ⊂ M is a global section of the restricted bundle E|V .
62 Chapter 6. Vector Bundles
6.3 Subbundles
Definition 6.15. Given a vector bundle πE : E → M , a subbundle of E is a vector bundle
πD : D → M , in which D is a topological subspace of E and πD is the restriction of πE
to D, such that for each p ∈ M , the subset Dp = D ∩ Ep is a linear subspace of Ep , and
the vector space structure on Dp is the one inherited from Ep .
If E → M is a smooth vector bundle, then a subbundle of E is called a smooth
subbundle if it is a smooth vector bundle and an embedded submanifold of E.
Note that the condition that D be a vector bundle over M implies that all of the fibers
Dp must be non-empty and have the same dimension.
Thefollowing lemma gives a convenient condition for checking that a union of sub-
spaces Dp ⊆ Ep | p ∈ M is a smooth subbundle.
Lemma 6.16 (Local frame criterion for subbundles). Let π : E → M be a smooth vector
bundle of rank k. Suppose that S for each p ∈ M we are given an m-dimensional linear
subspace Dp ⊆ Ep . Then D = p∈M Dp ⊆ E is a smooth subbundle of E if and only if the
following condition is satisfied: “Each point of M has a neighborhood U on which there
exist smooth local sections σ1 , . . . , σm : U → E with the property that σ1 (q), . . . , σm (q)
form a basis for Dq at each q ∈ U .”
(1) Let M be a smooth manifold and let S ⊆ M be an immersed k-submanifold. Then the
tangent bundle TS is a smooth rank-k subbundle of the ambient tangent bundle TM |S ;
see [Lee13, Problem 10-14].
(3) Let E → M be any smoothly trivial vector bundle of rank k and let (E1 , . . . , Ek )
be a smooth global frame for E. If m ∈ {0, . . . , k}, then the subset D ⊆ E defined by
Dp = span E1 |p , . . . , Em |p for each p ∈ M is a smooth rank-m subbundle of E.
CHAPTER 7
VECTOR FIELDS AND FLOWS
⇝ The set X(M ) of all smooth (global) vector fields on a smooth manifold M is an
infinite-dimensional R-vector space and a module over the ring C ∞ (M ): this is a
special case of [Exercise Sheet 10, Exercise 3(b)].
⇝ Extension lemma for vector fields: this is a special case of Lemma 6.10; see also
[Exercise Sheet 10, Exercise 3(d)] for an application (any tangent vector at a point
can be extended to a smooth vector field on the entire manifold).
63
64 Chapter 7. Vector Fields and Flows
⇝ Completion of smooth local frames for M : this is a special case of [Exercise Sheet
10, Exercise 4].
Let M and X be as above. If U, (xi ) is a smooth coordinate chart for M , then we
can write the value of X at any point p ∈ U in terms of the coordinate basis vectors:
∂
Xp = X i (p) .
∂xi p
Proposition 7.2 (Smoothness criterion for vector fields). Let M be a smooth manifold
i
and let X : M → TM be a rough vector field on M . If U, (x ) is a smooth coordinate
chart for M , then the restriction of X to U is smooth if and only if its components
functions with respect to this chart are smooth.
Example 7.3.
(1) If U, (xi ) is any smooth chart on M , then the assignment p 7→ ∂x∂ i p determines a
vector field on U , called the i-th coordinate vector field and denoted by ∂x∂ i . It is smooth
by Proposition 7.2, because its component functions are constant.
In particular, the coordinate vector fields form a smooth local frame ∂x∂ i for TM ,
called a coordinate frame. Note that every point of M is in the domain of such a local
frame.
∂ ∂
V x = x1 + . . . + xn .
∂x1 x ∂xn x
We will encounter many more examples of vector fields (especially on Rn ) later in the
course and in the exercise sheets as well.
Section 7.1. Vector Fields 65
This construction yields another useful smoothness criterion for vector fields.
Proposition 7.4 (Smoothness criterion for vector fields). Let M be a smooth manifold
and let X : M → TM be a rough vector field on M . The following are equivalent:
(a) X is smooth.
Proof.
(a) ⇒ (b): Given p ∈ M , pick a smooth chart U, (xi ) for M containing p. For x ∈ U
we may write
i ∂ ∂ fb
f = X i (x) i x
Xf (x) = X (x) i b .
∂x x ∂x
Since the component functions X i of X are smooth on U by Proposition 7.2, it follows
that Xf is smooth on U . We conclude by [Exercise Sheet 3, Exercise 2(a)].
(b) ⇒ (c): Fix an open subset U ⊆ M and f ∈ C ∞ (U ). For any p ∈ U , let ψ be a
smooth bump function that is equal to 1 in a neighborhood of p and supported in U (see
Proposition 2.20), and define fe = ψf , extended to be zero on M \ supp ψ. Then X fe is
smooth by assumption, and equal to Xf in a neighborhood of p by construction (and by
the above discussion). We conclude by [Exercise Sheet 3, Exercise 2(a)].
(c) ⇒ (a): If (xi ) are smooth local coordinates on U ⊆ M , then we think of each coordinate
xi as a smooth function on U , and we have
∂x i
i
i
∂ i ∂x
j j = δj
X(x ) = X (x ) ==== Xi ,
∂xj
which is smooth by assumption. We conclude by Proposition 7.2 and [Exercise Sheet 3,
Exercise 2(a)].
66 Chapter 7. Vector Fields and Flows
One consequence of Proposition 7.4 is that a smooth vector field X ∈ X(M ) defines a
map
C ∞ (M ) → C ∞ (M ), f 7→ Xf,
which (as can be checked pointwise) is R-linear and satisfies the following product rule
for vector fields:
X(f g) = f Xg + g Xf ;
in other words, the map is a derivation of C ∞ (M ).
The next proposition shows that derivations of C ∞ (M ) can be identified with smooth
vector fields. Due to this result, we sometimes identify smooth vector fields on M with
derivations of C ∞ (M ), using the same letter for both the vector field (thought of as a
smooth map M → TM ) and the derivation (thought of as a linear map C ∞ (M ) →
C ∞ (M )).
Proof.
“⇒”: We just showed above that any smooth vector field induces a derivation of C ∞ (M ).
“⇐”: Let p ∈ M and consider the map
Xp : C ∞ (M ) → R, f 7→ (Df )(p).
Lemma 7.6. Let F : M → N be a smooth map. Let X ∈ X(M ) and Y ∈ X(N ). Then
X and Y are F -related if and only if for every smooth real-valued function f defined on
an open subset of N , we have
X(f ◦ F ) = (Y f ) ◦ F.
It is important to remember that for a given smooth map F : M → N and vector field
X ∈ X(M ), there may not be any vector field on N that is F -related to X. There is one
special case, however, in which there is always such a vector field, as the next proposition
shows.
on R2 . We compute
∂(x2 )
∂(xy)
(XY )(f g) = X x = X(x2 ) = = 2x
∂y ∂x
and
1
0
1
∂y
∂x
∂x
f XY g + gXY f = x X x + y X x = x = x,
∂y ∂y ∂x
so XY is not a derivation of C ∞ (R2 ).
However, we can also apply the same two vector fields in the opposite order, obtaining
a (usually different) smooth function Y Xf ∈ C ∞ (M ). Applying both of these operators
to f ∈ C ∞ (M ) and subtracting, we obtain the operator
[X, Y ] : C ∞ (M ) → C ∞ (M ), f 7→ XY f − Y Xf,
called the Lie bracket of X and Y . The key fact, following readily from Proposition 7.5,
is that this operator is a vector field.
Lemma 7.11. The Lie bracket of any pair of smooth vector fields on a smooth manifold
is a smooth vector field.
Proof. See [Exercise Sheet 11, Exercise 3].
We mention below the basic properties of the Lie bracket and we refer to Exercise
Sheet 11 for their proofs. The geometric interpretation of the Lie bracket will not be
covered in this course, but we refer to [Lee13, Chapter 9, Lie derivatives] for some details.
Proposition 7.12 (Coordinate formula for the Lie bracket). Let M be a smooth n-
manifold and let X, Y ∈ X(M ). Let
n n
X
i ∂ X ∂
X= X and Y = Yj
i=1
∂xi j=1
∂xj
Section 7.1. Vector Fields 69
be the coordinate expressions for X and Y , respectively, in terms of some smooth local
coordinates (xi ) for M . Then the Lie bracket [X, Y ] has the following coordinate expres-
sion:
n X n j j
i ∂Y i ∂X ∂
X
[X, Y ] = X i
−Y i j
.
j=1 i=1
∂x ∂x ∂x
Proposition 7.13 (Properties of the Lie bracket). Let M be a smooth manifold. The Lie
bracket satisfies the following identities for all X, Y, Z ∈ X(M ):
(b) Antisymmetry:
[X, Y ] = −[Y, X].
Corollary 7.16 (Lie brackets of smooth vector fields tangent to submanifolds). Let M be
a smooth manifold and let S be an immersed submanifold of M . If Y1 and Y2 are smooth
vector fields on M that are tangent to S, then their Lie bracket [Y1 , Y2 ] is tangent to S as
well.
γ ′ (t) = Vγ(t) , ∀ t ∈ J .
Finding integral curves of vector fields boils down to solving a system of ODEs in a
smooth chart. Suppose that V ∈ X(M ) and that γ : J ⊆ R → M is a smooth curve. On
a smooth coordinate domain U ⊆ M we can write γ in local coordinates as
Then the condition γ ′ (t) = Vγ(t) for γ to be an integral curve of V can be written as
∂ ∂
γ̇ i (t) = V i γ(t) ,
∂xi γ(t) ∂xi γ(t)
The fundamental fact about such systems is the following existence, uniqueness and
smoothness theorem. (This is the reason for the terminology “integral curves”, because
solving a system of ODEs is often referred to as “integrating” the system.)
Theorem 7.18 (Fundamental theorem for autonomous ODEs). Let V : U → Rn be a
smooth vector-valued function, where U ⊆ Rn is open. Consider the initial value problem
(a) Existence: For any t0 ∈ R and x0 ∈ U , there exists an open interval J0 ∋ t0 and an
open subset x0 ∈ U0 ⊆ U such that for each c ∈ U0 , there is a C 1 map y : Jo → U
that solves (1) - (2).
(b) Uniqueness: Any two differentiable solutions to (1) - (2) defined on intervals con-
taining t0 agree on their common domain.
Proposition 7.19. Let V be a smooth vector field on a smooth manifold M . For each
point p ∈ M , there exists ε > 0 and a smooth curve γ : (−ε, ε) → M that is an integral
curve of V starting at p ∈ M .
Proof. Follows immediately from the existence and smoothness part of Theorem 7.18.
The next two lemmas show how affine reparametrizations affect integral curves.
Lemma 7.20 (Rescaling lemma). Let V be a smooth vector field on a smooth manifold
M , let J ⊆ R be an interval, and let γ : J → M be an integral curve of V . For any a ∈ R,
the curve
e : Je → M, t 7→ γ(at)
γ
Lemma 7.21 (Translation lemma). Let V be a smooth vector field on a smooth manifold
M , let J ⊆ R be an interval, and let γ : J → M be an integral curve of V . For any b ∈ R,
the curve
b : Jb → M, t 7→ γ(t + b)
γ
(1) Let
∂
V = ∈ X(R2 )
∂x
be the first coordinate vector field. Note that the integral curves of V are precisely the
straight lines parallel to the x-axis (see Figure 7.4a), with parametrization of the form
γ(t) = (a + t, b) for constants a, b ∈ R. Thus, there is a unique integral curve starting at
each point of the plane, and the images of different integral curves are either identical or
disjoint.
(2) Let
∂ ∂
W = −y +x ∈ X(R2 ).
∂x ∂y
To determine the integral curves of W we proceed as follows (see p. 70):
(
γ̇1 (t) = −γ2 (t)
γ(t) = γ 1 (t), γ 2 (t)
=⇒ γ̇(t) = Wγ(t) =⇒
γ̇2 (t) = γ1 (t)
(
γ̈1 (t)+γ1 (t)=0 γ1 (t) = a cos t − b sin t
========⇒
γ2 (t) = a sin t + b cos t = −γ̇1 (t)
is an integral curve of W . When (a, b) = (0, 0), this is the constant curve γ(t) = (0, 0);
otherwise, it is a circle traversed clockwise (see Figure 7.4b). Since γ(0) = (a, b), we
see again that there is a unique integral curve staring at each point (a, b) ∈ R2 , and the
images of the various integral curves are either identical or disjoint.
7.3 Flows
Definition 7.24. Let M be a smooth manifold.
(a) A flow domain for M is an open subset D ⊆ R × M with the property that for each
p ∈ M , the set
D(p) := t ∈ R | (t, p) ∈ D
• ∀ p ∈ M : θ(0, p) = p.
• ∀ s ∈ D(p) ∀ t ∈ D(θ(s,p)) such that s + t ∈ D(p) , we have
θ t, θ(s, p) = θ(t + s, p).
(c) A maximal flow on M is a flow that admits no extension to a flow on a larger flow
domain.
Let θ : D → M be a flow on M .
Mt := p ∈ M | (t, p) ∈ D
and a map
θt : Mt → M, θt (p) = θ(t, p) = θ(p) (t) .
Note that
p ∈ Mt ⇐⇒ (t, p) ∈ D ⇐⇒ t ∈ D(p) .
74 Chapter 7. Vector Fields and Flows
′ d
V : M → TM, p 7→ Vp := θ(p) (0) = θ(p) (t) ,
dt t=0
is a smooth vector field on M , and each curve θ(p) is an integral curve of V starting at
p ∈ M.
Proof. If D = R × M , then this is shown in [Exercise Sheet 12, Exercise 4]. The proof of
the general case is essentially identical to the proof for global flows (after verifying that
all the expressions involved make sense).
The term “infinitesimal generator” comes from the following picture: in a smooth
chart, a good approximation to an integral curve can be obtained by composing many
small straight-line motions, with the direction and length of each motion determined by
the value of the vector field at the point arrived at in the previous step. Intuitively, one
can think of a flow as a sequence of infinitely many infinitesimally small linear steps.
(a) For each p ∈ M , the curve θ(p) : D(p) → M is the unique maximal integral curve of
V starting at p.
(c) For each t ∈ R, the set Mt is open in M , and the map θt : Mt → M−t is a diffeo-
morphism with inverse θ−t .
Proof.
(a) Proposition 7.19 shows that there exists an integral curve of V starting at each point
p ∈ M . Suppose that γ and γ e are two integral curves of V defined on the same open
e(t0 ) for some t0 ∈ J. Consider the set
interval J such that γ(t0 ) = γ
S := t ∈ J | γ(t) = γ
e(t)
θ(p) (t) = γ(t), where γ is any integral curve starting at p and defined on an open interval
containing 0 and t. Since all such integral curves agree at t by the argument above, θ(p)
is well defined, and it is obviously the unique maximal integral curve of V starting at p.
Next, for the verification that the set
D := (t, p) ∈ R × M | t ∈ D(p)
satisfies the claimed properties, as well as for the proof of (b), we refer to [Lee13, Theorem
9.12], which makes heavy use of Theorem 7.18.
(c) The fact that Mt is open in M is an immediate consequence of the fact that D is open.
We have
(b)
p ∈ Mt =⇒ t ∈ D(p) =⇒ D(θ(t,p)) = D(p) − t
dfn
=⇒ −t ∈ D(θ(t,p)) =⇒ θt (p) ∈ M−t ,
which shows that θt maps Mt to M−t for any (fixed) t ∈ R. Moreover, the group laws
then show that θ−t ◦ θt is equal to the identity on Mt . Reversing the roles of t and −t
shows that θt ◦ θ−t is equal to the identity on M−t . This completes the proof of (c).
The flow whose existence and uniqueness are asserted in Theorem 7.26 is called the
flow generated by V , or just the flow of V .
The naturality of integral curves (see Proposition 7.22) translates into the following
naturality statement for flows.
F
Mt Nt
θt ηt
F
M−t N−t
Corollary 7.33. Every smooth vector field on a compact smooth manifold is complete.
Exercise 7.34 (The escape lemma): Let M be a smooth manifold and let V be a smooth
vector field on M . Show that if γ : J → M is a maximal integral curve of V whose domain
J has a finite least upper bound b ∈ R, then for any t0 ∈ J the image γ [t0 , b) of the
interval [t0 , b) under γ is not contained in any compact subset of M .
CHAPTER 8
DIFFERENTIAL FORMS
T∗p M := (Tp M )∗ .
i ∂
ω = ωi λ p , where ωi = ω .
∂xi p
Given now another set of smooth local coordinates (exj ) whose domain contains p ∈ U ,
ej |p the basis for T∗ M dual to ∂ j . We can compute the components
denote by λ p ∂e
x p
∗
of the same covector ω ∈ Tp M with respect to the new coordinate system as follows.
According to (3.6), the coordinate vector fields transform as follows:
∂ xj
∂e ∂
= (p) (8.1)
∂xi p ∂xi xj
∂e p
79
80 Chapter 8. Differential Forms
ξi λi
p
7→ p, (ξ1 , . . . , ξn ) ,
where λi is the i-th coordinate covector field associated with (xi ). Suppose that (U e , φ)
e is
−1 e
another smooth chart for M with coordinate functions (e x ), and let Φ : π (U ) → U × Rn
j e e
−1
be defined analogously. On π (U ∩ U e ), it follows from (8.2) that
j
xj
−1
1 n
∂e
x ∂e
Φ◦Φ e p, (ξ , . . . , ξ ) = p,
e e (p) ξj , . . . , n (p) ξj
e e .
∂x1 ∂x
xj
The GL(n, R)-valued function ∂e
∂x i is smooth, so it follows from the vector bundle chart
lemma (= Lemma 6.6) that T∗ M has a smooth structure making it into a smooth vector
bundle for which the maps Φ are smooth local trivializations. Uniqueness follows as in
the proof of [Exercise Sheet 10, Exercise 6].
Section 8.1. Differential 1-Forms 81
As in the case of the tangent bundle (see the proof of Proposition 3.12), smooth local
coordinates for M yield smooth local coordinates for its cotangent bundle. If (xi ) are
smooth coordinates on an open subset U ⊆ M , then [Exercise Sheet 10, Exercise 5(d)]
shows that the map
is a smooth coordinate chart for T∗ M . We call (xi , ξi ) the natural coordinates for T∗ M
associated with (xi ).
⇝ Extension lemma for covector fields: this is a special case of Lemma 6.10; see also
[Exercise Sheet 10, Exercise 3(d)] for an application (any tangent covector at a point
can be extended to a smooth covector field on the entire manifold).
⇝ Completion of smooth local coframes for M : this is a special case of [Exercise Sheet
10, Exercise 4].
(c) Each point of M is contained in some coordinate chart in which ω has smooth
component functions.
(e) For every open subset U ⊆ M and every smooth vector field X on U , the function
ω(X) : U → R is smooth.
Of course, since any open subset of a smooth manifold is again a smooth manifold,
Proposition 8.5 applies equally well to covector fields defined only on some open subset
of M .
Example 8.6. For any smooth chart U, (xi ) , the coordinate covector fields (λi ) defined
above constitute a local coframe over U , called a coordinate coframe. By Proposition 8.5,
every coordinate coframe is smooth, because its component functions in the given chart
are constants.
More generally, if (Ei ) is a (rough) local frame for TM over an open subset U ⊆ M ,
then there is a uniquely determined (rough) local coframe (εi ) over U such that εi |p is
the dual basis to (Ei |p ) for each p ∈ U , or equivalently εi (Ej ) = δji . This coframe is called
the coframe dual to (Ei ). Conversely, if (εi ) is a (rough) local coframe over an open subset
U ⊆ M , then there is a uniquely determined (rough) local frame (Ei ) for TM over U ,
called the frame dual to (εi ) and determined by εi (Ej ) = δji . For example, in a smooth
chart, the coordinate frame ∂x∂ i and the coordinate coframe (λi ) are dual to each other.
Lemma 8.7. Let M be a smooth manifold. If (Ei ) is a rough local frame over an open
subset U ⊆ M and if (εi ) is its dual coframe, then (Ei ) is smooth if and only if (εi ) is
smooth.
∂
Ei = aki and εj = bjℓ λℓ
∂xk
for some matrices of real-valued functions aki and bjℓ defined on V . By virtue of Propo-
sitions 7.2 and 8.5, the vector fields Ei are smooth on V if and only if the functions aki
are smooth, and the covector fields εj are smooth on V if and only if the functions bℓj are
smooth. The fact that εj (Ei ) = δij implies that the matrices (aki ) and (bℓj ) are inverses to
each other. Since matrix inversion is a smooth map GL(n, R) → GL(n, R), we conclude
that either one of these matrix-valued functions is smooth if and only if the other one is
smooth.
Section 8.1. Differential 1-Forms 83
∂xj
dxj p
= (p) λi p
= δij λi p
= λj p ;
∂xi
in other words, the coordinate vector field λj is none other than the differential dxj .
Therefore, (8.3) can be rewritten as
∂f
dfp = (p) dxi p , p ∈ U,
∂xi
84 Chapter 8. Differential Forms
∂f
df = i
dxi . (8.4)
∂x
In particular, in the 1-dimensional case, this reduces to
df
df = dx .
dx
Thus, we have recovered the familiar classical expression for the differential of a function
f in coordinates. Henceforth, we abandon the notation λi for the coordinate coframe, and
use dxi instead.
Example 8.9. If
f : R2 → R, (x, y) 7→ x2 y cos x,
then
∂(x2 y cos x) ∂(x2 y cos x)
df = dx + dy
∂x ∂y
= (2xy cos x − x2 y sin x) dx + (x2 cos x) dy .
Proposition 8.10 (Properties of the differential). Let M be a smooth manifold and let
f, g ∈ C ∞ (M ). The following statements hold:
(b) d(f g) = f dg + g df .
Unlike vector fields, whose pushforwards are defined only in certain special cases (see,
e.g., Subsection 7.1.2), covector fields always pullback to covector fields.
Section 8.1. Differential 1-Forms 85
lin. (8.5)
== u(F (p)) dFp∗ ωF (p) == (u ◦ F )(p) (F ∗ ω)p
= (u ◦ F )(F ∗ ω) p ,
(8.5)
F ∗ (du) p (v) == dFp∗ (duF (p) ) (v)
dfn
== duF (p) dFp (v)
and the identific.
using ES4E1(b)
immediately
dfn of du
===== dFp (v) u
dfn of dFp
===== v(u ◦ F )
dfn of d(u◦F )
======= d(u ◦ F )p (v) ,
which yields the assertion.
Proposition 8.14. Let F : M → N be a smooth map and let ω be a (continuous) covector
field on N . Then F ∗ ω is a (continuous) covector field on M , and if ω is smooth, then so
is F ∗ ω.
Proof. Fix p ∈ M and choose smooth coordinates (y j ) for N in a neighborhood V of
F (p). Set U = F −1 (V ) and observe that U is a neighborhood of p in M . Writing ω in
coordinates as ω = ωj dy j for (continuous) functions on V and using Proposition 8.13
twice (for F |U ), we compute that
F ∗ ω = F ∗ (ωj dy j ) = (ωj ◦ F )F ∗ dy j = (ωj ◦ F ) d(y j ◦ F ) . (8.6)
In view of Proposition 8.8, this expression is continuous, and it is smooth when ω is
smooth, so we are done.
86 Chapter 8. Differential Forms
Formula (8.6) for the pullback of a covector field can also be written in the following
way:
F ∗ ω = (ωj ◦ F ) d(y j ◦ F ) = (ωj ◦ F ) dF j ,
where F j is the j-th component function of F in these coordinates. Using either of these
formulas, the computation of pullbacks in coordinates is quite simple.
ω = u dv + v du ∈ X∗ (R2 ) .
F ∗ ω = (u ◦ F ) d(v ◦ F ) + (v ◦ F ) d(u ◦ F )
= (x2 y) d(y sin z) + (y sin z) d(x2 y)
= (x2 y)(sin z dy + y cos z dz) + y sin z (2xy dx + x2 dy)
= (2xy 2 sin z) dx + (2x2 y sin z) dy + (x2 y 2 cos z) dz.
since dιp : Tp S → Tp M is just the inclusion map under our usual identification of Tp S
with the subspace dιp (Tp S) of Tp M . Thus, ι∗ ω is just the restriction of ω to vectors
tangent to S. For this reason, ι∗ ω is often called the restriction of ω to S. Note, however,
that ι∗ ω might equal zero at a given point of S, even though considered as a covector field
on M , ω might not vanish there. For example:
Example 8.17. Consider ω = dy ∈ X∗ (R2 ) and let S : (y = 0) be the x-axis, considered
as an embedded submanifold of R2 . As a covector field on R2 , ω is clearly nonzero
everywhere, because one of its components is always equal to 1. However, the restriction
ι∗ ω of ω to S is identically zero, because y vanishes identically on S:
ι∗ ω = ι∗ dy = d(y ◦ ι) = 0.
To distinguish the two ways in which we might interpret the statement “ω vanishes
on S”, one usually says that ω vanishes along S (or vanishes at points of S) if ωp = 0 for
every p ∈ S. The weaker condition that ι∗ ω = 0 is expressed by saying that the restriction
of ω to S vanishes (or the pullback of ω to S vanishes).
with the obvious projection map. It can be shown (exercise!) that it is a smooth
vector bundle of rank nk over M . Its (smooth) sections are called (smooth) covariant
k-tensor fields on M .
(b) The subset of Tk (T∗ M ) consisting of alternating k-tensors is defined as:
Vk ∗ G Vk
(T M ) := (T∗p M ).
p∈M
Ωk (M ) := Γ k (T∗ M ) .
V
(T∗p M ) ∼
V0 ∗ G V0 G
(T M ) = = R = M × R,
p∈M p∈M
(T∗p M ) ∼
V1 ∗ G V1 G
(T M ) = = T∗p M = T∗ M.
p∈M p∈M
(ω ∧ η)p = ωp ∧ ηp .
Comment: If we define n
M
∗
Ω (M ) = Ωk (M ),
k=0
where the coefficients ωI are continuous functions defined on the coordinate domain U ,
we use dxI as an abbreviation for dxi1 ∧ . . . ∧ dxik (where I = (i1 , . . . , in )) and the
primed summation sign denotes a sum over only increasing multi-indices. According to
Proposition 6.14, ω is smooth if and only if the component functions ωI are smooth. Since
i1 ik ∂ ∂
dx ∧ . . . ∧ dx j
,..., j = δJI
∂x 1 ∂x k
by Lemma C.20 and Proposition C.25(c), the component functions ωI of ω are determined
by
∂ ∂
ωI = ω ,..., i .
∂xi1 ∂x k
(b) F ∗ (ω ∧ η) = F ∗ ω ∧ F ∗ η.
Section 8.2. Differential k-Forms 89
(c) In any smooth chart V, (y i ) for N , we have
X′ X′
F∗ ωI dy i1 ∧ . . . ∧ dy ik = (ωI ◦ F ) d(y i1 ◦ F ) ∧ . . . ∧ d(y ik ◦ F ).
I I
= v du ∧ d(u2 − v 2 ) + u dv ∧ d(u2 − v 2 )
du∧du=0
= v du ∧ (2u du − 2v dv) + u dv ∧ (2u du − 2v dv) =====
dv∧dv=0
du∧dv=
= −2v 2 du ∧ dv + 2u2 dv ∧ du =====
−dv∧du
= −2(u2 + v 2 ) du ∧ dv.
Proposition 8.21 (Pullback formula for top forms). Let F : M → N be a smooth map
between smooth n-manifolds. If (xi ) and (y j ) are smooth coordinates on open subsets
U ⊆ M and V ⊆ N , respectively, and if u is a real-valued function on V , then the
following holds on U ∩ F −1 (V ):
F ∗ u dy 1 ∧ . . . ∧ dy n = (u ◦ F ) det DF dx1 ∧ . . . ∧ dxn ,
(8.7)
where DF represents the Jacobian matrix of F in these coordinates.
Proof. Since the fiber of n (T∗ M ) is spanned by dx1 ∧. . .∧dxn at each
V
∂ ∂
point, it suffices to
show that both sides of (8.7) agree when evaluated on ∂x1 , . . . , ∂xn . By Lemma 8.19(c)
we have
F ∗ u dy 1 ∧ . . . ∧ dy n = (u ◦ F ) d(y 1 ◦ F ) ∧ . . . ∧ d(y n ◦ F ),
| {z } | {z }
F1 Fn
so by Proposition C.25(c)(d) we obtain
∂ ∂ ∂ ∂
F ∗ u dy 1 ∧ . . . ∧ dy n , . . . , = (u ◦ F ) dF 1
∧ . . . ∧ dF n
, . . . ,
∂x1 ∂xn ∂x1 ∂xn
∂
= (u ◦ F ) det dF j
∂xi
1 n
∂ ∂
= (u ◦ F ) det DF dx ∧ . . . ∧ dx 1
,..., n ,
| {z ∂x ∂x }
=1
as desired.
90 Chapter 8. Differential Forms
Corollary 8.22. If U, (xi ) and U xj ) are overlapping smooth coordinate charts on
e , (e
a smooth manifold M , then the following identity holds on U ∩ U e:
xj
∂e
x1 ∧ . . . ∧ de
de xn = det dx1 ∧ . . . ∧ dxn .
∂xi
Proof. Apply Proposition 8.21 for F = IdU ∩Ue , but using coordinates (xi ) in the domain
xj ) in the codomain.
and (e
where dωJ is the differential of the smooth function ωJ , see Subsection 8.1.4. In somewhat
more detail, this is
X XX
′
j1 jk ∂ωJ i
d ωJ dx ∧ . . . ∧ dx = i
dx ∧ dxj1 ∧ . . . ∧ dxjk .
J J i
∂x
∂f
df = dxi ,
∂xi
which is just the differential of f , see (8.4), while for a smooth 1-form ω = ωj dxj we
compute that
X ∂ωj ∂ωi i
dω = i
− j dx ∧ dxj .
i<j
∂x ∂x
In order to transfer this definition to manifolds, we first need to check that it satisfies
the following properties.
(a) d is R-linear.
(c) d ◦ d ≡ 0.
Section 8.2. Differential k-Forms 91
Proof.
(b) Due to (a), it suffices to consider terms of the form ω = u dxI ∈ Ωk (U ) and η =
v dxJ ∈ Ωℓ (U ), where u, v ∈ C ∞ (U ).
Proof: If I has repeated indices, then clearly d(u dxI ) = 0 = du ∧ dxI . Otherwise, let
σ be a permutation sending I to an increasing multi-index J. Then
∂u
i
∂ 2u dxi ∧dxi =0
d(du) = d i
dx = i j
dxi ∧ dxj ======
∂x ∂x ∂x
X ∂ 2u ∂ 2u i
= i ∂xj
− j ∂xi
dx ∧ dxj
i<j
∂x ∂x
=0
92 Chapter 8. Differential Forms
X
′
j1 jk
d(du) = d dωJ ∧ dx ∧ . . . ∧ dx
J
0 by (b) and
by case k=0
X′ :
j1 jk
+ (−1) · dωJ ∧ d dx ∧
. . . ∧ dx
J
= 0.
1
(∗)
= d(u ◦ F ) ∧ d(xi1 ◦ F ) ∧ . . . ∧ d(xik ◦ F ) ==
Lemma 8.19(c)
= d (u ◦ F ) d(xi1 ◦ F ) ∧ . . . ∧ d(xik ◦ F ) =========
= d F ∗ (u dxi1 ∧ . . . ∧ dxik ) .
Example 8.24. Let us compute the exterior derivatives of arbitrary 1-forms and 2-forms
on R3 .
ω = P dx + Q dy + R dz
for some smooth functions P, Q, R on R3 . Using (8.8) and the fact that the wedge product
of any 1-form with itself is zero, we compute
dω = dP ∧ dx + dQ ∧ dy + dR ∧ dz
∂P ∂P ∂P
= dx + dy + dz ∧ dx +
∂x ∂y ∂z
∂Q ∂Q ∂Q ∂R ∂R ∂R
+ dx + dy + dz ∧ dy + dx + dy + dz ∧ dz
∂x ∂y ∂z ∂x ∂y ∂z
∂Q ∂P ∂R ∂P ∂R ∂Q
= − dx ∧ dy + − dx ∧ dz + − dy ∧ dz .
∂x ∂y ∂x ∂z ∂y ∂z
1
(∗): We have an expression of the form df ∧ η, where η = dg1 ∧ . . . ∧ dgk (with f = u ◦ F and
gℓ = xiℓ ◦ F ), so
p. 87 (b)
d(f η) === d(f ∧ η) == df ∧ η + (−1)0 f ∧ dη = df ∧ η,
since dη = 0 by (b) and (c).
Section 8.2. Differential k-Forms 93
η = u dx ∧ dy + v dx ∧ dz + w dy ∧ dz
d : Ωk (M ) → Ωk+1 (M ) ,
(c) d ◦ d ≡ 0.
This is well-defined, since for any other smooth chart (V, ψ) for M , the map φ ◦ ψ −1 is a
diffeomorphism between open subsets of Rn , so
∗ ∗
ψ ∗ d (ψ −1 ) ω = (φ−1 ◦ φ)∗ ψ ∗ d (ψ −1 ) ω
| {z }
Id
∗ (φ−1 )∗ ψ ∗ =(ψ◦φ−1 )∗
= φ∗ (φ−1 )∗ ψ ∗ d (ψ −1 ) ω
============
& P roposition 8.23(d)
∗ ∗
= φ∗ d (ψ ◦ φ−1 ) (ψ −1 ) ω
| {z }
(ψ −1 ◦ ψ ◦ φ−1 )∗ =(φ−1 )∗
∗
= φ∗ d (φ−1 ) ω .
Comment: The preceding theorem can be summarized by saying that the differential on
functions extends uniquely to an anti-derivation of Ω∗ (M ) of degree +1 whose square is
zero.
= d(F ∗ ω).
Remark 8.28. Every exact form is closed, since d ◦ d ≡ 0, but the converse does not hold
in general, see Example 11.27. However, it can be shown that closed forms are locally
exact (but not necessarily globally), so the question of whether a given closed form is
exact depends on global properties of the manifold.
CHAPTER 9
MANIFOLDS WITH BOUNDARY
We briefly discuss manifolds with boundary. They play a central role in the theory of
integration on manifolds, which will be developed in Chapter 11.
Hn = (x1 , . . . , xn ) ∈ Rn | xn ≥ 0 .
The interior and the boundary of Hn as a subset of Rn are denoted by Int Hn and ∂Hn ,
respectively.
If n > 0, then
∂Hn = (x1 , . . . , xn ) ∈ Rn xn = 0 ,
whereas if n = 0, then
H0 = R0 = {0} and ∂H0 = ∅ .
Definition 9.2. An n-dimensional topological manifold with boundary is a second-coun-
table, Hausdorff topological space M in which every point has a neighborhood homeo-
morphic either to an open subset of Rn or to a (relatively) open subset of Hn .
An open subset U ⊆ M together with a map φ : U → Rn that is a homeomorphism
onto an open subset of Rn or Hn is called a chart for M . When it is necessary to make the
distinction, we call (U, φ) an interior chart for M if φ(U ) is an open subset of Rn (which
includes the case of an open subset of Hn that does not intersect ∂Hn ), and a boundary
chart for M if φ(U ) is a open subset of Hn such that φ(U ) ∩ ∂Hn ̸= ∅.
A point p ∈ M is called an interior point of M if it is in the domain of some interior
chart, and a boundary point of M if it is in the domain of a boundary chart that sends p to
∂Hn . The set of all boundary points of M is denoted by ∂M and is called the boundary of
M , while the set of all interior points of M is denoted by Int M and is called the interior
of M .
95
96 Chapter 9. Manifolds with Boundary
Example 9.4.
(1) Every interval in R is a connected topological 1-manifold with boundary, whose man-
ifold boundary consists of its endpoints (if any).
n
(2) The closed unit ball B ⊆ Rn is a connected topological n-manifold with boundary,
whose (manifold) boundary is Sn−1 and whose interior is Bn ; see [Lee13, Problem 1.11].
(c) M is a topological manifold (in the sense of Definition 1.1) if and only if ∂M = ∅.
Proof. Exercise!
Definition 9.6. Let M be a topological manifold with boundary. A smooth structure for
M is defined to be a maximal smooth atlas (a collection of charts whose domains cover M
and whose transition maps (and their inverses) are smooth in the sense just described).
With such a structure, M is called a smooth manifold with boundary.
In the following lengthy remark we collect some basic definitions and facts about
smooth manifolds with boundary, referring to [Lee13] for further information.
Remark 9.7.
(2) Cf. Chapter 3: If M is a smooth n-manifold with boundary, then the tangent space
Tp M to M at p ∈ M is defined in the same way (see Definition 3.4), and it is an n-
dimensional R-vector space. For any smooth chart (U, (xi )) containing p, the coordinate
vectors
∂ ∂
1
,..., n
∂x p ∂x p
(where ∂x∂n p
should be interpreted as a one-sided derivative when p ∈ ∂M ) form a basis
for Tp M .
Let M be a smooth manifold with boundary and let p ∈ ∂M . It is intuitively evident
that the vectors in Tp M can be separated in three classes: those tangent to the boundary,
those pointing inward, and those pointing outward. Formally, we make the following
definition:
(3) Cf. Chapter 4: Submersions, immersions, embeddings and local diffeomorphisms are
defined in the same way (see Definitions 4.2 and 4.7(b)), and there is a version of the rank
theorem in this setting (see [Lee13, Theorem 4.15 and Problem 4.3]).
(4) Cf. Chapter 5: Immersed and embedded submanifolds of smooth manifolds with
boundary are defined in the same way (see 5.1 and 5.12) and are themselves smooth
manifolds with (possibly empty) boundary.
Theorem: If M is a smooth n-manifold with boundary, then with the subspace topology,
∂M is a topological (n − 1)-manifold (without boundary), and has a unique smooth
structure such that it is a properly embedded submanifold of M .
98 Chapter 9. Manifolds with Boundary
(5) Cf. Chapter 7: The tangent bundle of a smooth n-fold with boundary is defined in
the same way (see Definition 3.11) and it is a smooth vector bundle of rank n over the
given manifold (see Proposition 6.4). Vector fields are also defined in the same way (see
Definition 7.1), but flows in this setting need to be treated with extra care (see [Lee13,
Chapter 9, Flows and Flowouts on Manifolds with Boundary]).
(6)
Vk Cf. Chapter 8: The cotangent bundle T∗ M (respectively the k-th exterior power
(T∗ M ) of the cotangent bundle) of a smooth n-manifold M with boundary is defined in
the same way (see Definition 8.2, respectively Definition 8.18(b)), and it is a smooth vec-
tor bundle of rank n (respectively of rank nk ) over M (see Proposition 8.3, respectively
Definition 8.18(b)). Differential k-forms (0 ≤ k ≤ n) are also defined in the same way
(see Definition 8.18(b)), and so does their exterior derivative as well (see Theorem 8.25).
CHAPTER 10
ORIENTATIONS
The purpose of this chapter is to introduce a subtle but important property of smooth
manifolds, called orientation. An orientation of a line or a curve is simply a choice of direc-
tion along it. For 2-dimensional manifolds, an orientation is essentially a choice of which
rotational direction should be considered “clockwise” and which “counterclockwise”. For
3-dimensional ones, it is a choice between “left-handedness” and “right-handedness”. The
general definition of an orientation is an adaptation of these everyday concepts to arbitrary
dimensions.
Definition 10.1. Let V be a real vector space of dimension n ≥ 1. We say that two or-
dered bases (E1 , . . . , En ) and (E
e1 , . . . , E
en ) for V are consistently oriented if the transition
99
100 Chapter 10. Orientations
ω(E
e1 , . . . , E
en ) = ω(BE1 , . . . , BEn ) = (det B) ω(E1 , . . . , En ).
Section 10.2. Orientations of Smooth Manifolds 101
It follows that the basis (E ej ) is consistently oriented with (Ei ) if and only if ω(E
e1 , . . . , E
en )
and ω(E1 , . . . , En ) have the same sign, which is the same as saying that Oω is one equiv-
alence class. The last statement then follows easily (and is thus left as an exercise).
For example, the n-covector ε1...n = ε1 ∧ · · · ∧ εn is positively oriented for the standard
orientation on Rn ; see Lemma C.20(c).
Recall that if V is an n-dimensional real vector space, then the vector space Λn (V ∗ ) is
1-dimensional by Proposition C.21. Proposition 10.6 shows that choosing an orientation
for V is equivalent to choosing one of the two components of Λn (V ∗ ) \ {0}. This formu-
lation also works for 0-dimensional vector spaces, and explains why we have defined an
orientation of a 0-dimensional space in the way we did.
By itself, this is not a very useful concept, because the orientations at nearby points
may have no relation to each other. For example, a pointwise orientation on Rn might
switch randomly from point to point between the standard orientation and its opposite.
In order for pointwise orientations to have some relationship with the smooth structure,
we need an extra condition to ensure that the orientations of nearby tangent spaces are
consistent with each other.
Definition 10.9. Let M be a smooth manifold with or without boundary, endowed with
a pointwise orientation. If (Ei ) is a local frame for T M over an open subset U ⊆ M ,
then we say that (Ei ) is positively oriented (or simply oriented ) if (E1 |p , . . . , En |p ) is a
positively oriented ordered basis for Tp M at each point p ∈ U ; see Definition 10.4. A
negatively oriented frame for T M over U ⊆ M is defined analogously.
Definition 10.10. Let M be a smooth manifold with or without boundary (of dimension
n ≥ 1).
(c) We say that M is orientable if there exists an orientation for it; otherwise we say
that M is nonorientable.
102 Chapter 10. Orientations
Example 10.12. We give here some examples of orientable and nonorientable manifolds.
(2) For each n ∈ N, the unit n-sphere Sn ⊆ Rn+1 is orientable. Indeed, this follows
from Proposition 10.21, because Sn is a hypersurface in Rn+1 , to which the vector field
N = xi ∂/∂xi is nowhere tangent. We define the standard orientation of Sn to be the one
determined by N . (The standard orientation of S0 is the one that assigns the orientation
+1 to the point +1 ∈ S0 and −1 to the point −1 ∈ S0 .) Alternatively, this follows from
Proposition 10.23, because Sn is the boundary of the closed unit ball. (It can be checked
that the orientation thus induced on Sn is the standard one.)
(3) The so-called Möbius band is nonorientable; see [Lee13, Examples 10.3 and 15.38].
Proof.
“⇒”: Let ω be a nonvanishing n-form on M . By Proposition 10.6, ω defines a pointwise
orientation on M , so it remains to show that it is continuous. Since this is trivially true
for n = 0, we may assume that n ≥ 1. Given p ∈ M , let (Ei ) be any local frame for T M
over a connected open neighborhood U of p in M , and let (εi ) be the dual coframe. The
expression for ω in this frame over U is
ω = f ε 1 ∧ . . . ∧ εn
for some continuous function f on U . The fact that ω is nonvanishing means that f is
nonvanishing, and thus by Lemma C.20(c) we obtain
ωp (E1 |p , . . . , En |p ) = f (p) ̸= 0 for all p ∈ U.
Since U is connected, it follows that this expression is either always positive or always
negative on U , and therefore the given frame is either positively oriented or negatively
oriented. If the latter case holds, then we can replace E1 by −E1 to obtain a new frame
that is positively oriented. Hence, the pointwise orientation determined by ω is continuous.
“⇒”: We refer to [Lee13, Proposition 15.5] for the details.
Due to Proposition 10.14, we may now give the following definition.
Definition 10.15. Let M be a smooth n-manifold with or without boundary. Any
nonvanishing n-form on M is called an orientation form. If M is oriented and if ω is an
orientation form determining the given orientation, then we also say that ω is positively
oriented (or simply oriented ).
If M is zero-dimensional, then a nonvanishing 0-form (i.e., a nonvanishing smooth
real-valued function) on M assigns the orientation +1 to points where it is positive and
−1 to points where it is negative.
Remark 10.16. It is straightforward to check (see Proposition 10.6) that if ω and ω
e are
two positively oriented smooth n-forms on M , then ωe = f ω for some strictly positive
smooth real-valued function f on M .
Definition 10.17.
(a) A smooth coordinate chart U, (xi ) on an oriented smooth manifold with or without
boundary is said to be positively oriented (or simply oriented ) if the coordinate
frame (∂/∂xi ) is positively oriented, and negatively oriented if the coordinate frame
(∂/∂xi ) is negatively oriented; see Definition 10.9.
(b) A smooth atlas {(Uα , φα )} for a smooth manifold M with or without boundary is
said to be consistently oriented if for each α, β, the transition map φβ ◦ φ−1
α has
positive Jacobian determinant everywhere on φα (Uα ∩ Uβ ).
Proposition 10.18 (The orientation determined by a coordinate atlas). Let M be a
smooth manifold with or without boundary of dimension n ≥ 1. Given any consistently
oriented smooth atlas for M , there exists a unique orientation for M with the property
that each chart in the given atlas is positively oriented. Conversely, if M is oriented and
either ∂M = ∅ or n > 1, then the collection of all oriented smooth charts is a consistently
oriented atlas for M .
104 Chapter 10. Orientations
Proof. Assume first that M has a consistently oriented smooth atlas. Each chart in the
atlas determines a pointwise orientation at each point of its domain. Wherever two of
the charts overlap, the transition matrix between their respective coordinate frames is
the Jacobian matrix of the transition map (see the bottom of p. 27 and (3.6)), which has
positive determinant by assumption, so they determine the same pointwise orientation at
each point. The pointwise orientation on M thus determined is continuous, because each
point of M is in the domain of an oriented coordinate frame.
Conversely, assume that M is oriented and either ∂M = ∅ or n > 1. Each point is
in the domain of a smooth chart with connected domain, and if the chart is negatively
oriented (see Exercise 10.11), then we can replace x1 with −x1 to obtain a new chart that
is positively oriented. The fact that all these charts are positively oriented guarantees that
their transition maps have positive Jacobian determinants, so they form a consistently
oriented atlas.2
Definition 10.20. Let M be a smooth manifold with or without boundary and let S ⊆ M
be an immersed submanifold with or without boundary. A vector field along S is a section
of the ambient tangent bundle T M |S , i.e., a continuous map N : S → T M with the
property that Np ∈ Tp M for every p ∈ S. Such a vector field is said to be nowhere
tangent to S if Np ∈ Tp M \ Tp S for all p ∈ S; cf. Subsection 7.1.3.
Note that any vector field on M restricts to a vector field along S (not necessarily
tangent to S), but in general not every vector field along S is of this form, see Lemma 6.11.
Note that not every hypersurface admits a nowhere tangent vector field, see for instance
[Lee13, Problem 15.6]. However, the following result gives a sufficient condition that holds
in many cases.
Example 10.24. We determine the induced orientation on ∂Hn when Hn itself has the
standard orientation inherited from Rn . We can identify ∂Hn with Rn−1 under the corre-
spondence
(x1 , . . . , xn−1 , 0) ↔ (x1 , . . . , xn−1 ).
Since the vector field −∂/∂xn is outward-pointing along Hn , the standard coordinate
frame for Rn−1 is positively oriented for ∂Hn if and only if [−∂/∂xn , ∂/∂x1 , . . . , ∂/∂xn−1 ]
is the standard orientation for Rn ; see Proposition 10.21. This orientation satisfies
Thus, the induced orientation on ∂Hn is equal to the standard orientation on Rn−1 when
n is even, but it is opposite to the standard orientation when n is odd. In particular, the
standard coordinates on ∂Hn ≈ Rn−1 are positively oriented if and only if n is even.
106 Chapter 10. Orientations
(a) F is orientation-preserving.
(b) With respect to any positively oriented smooth charts for M and N , the Jacobian
matrix of F has positive determinant.
Proof. Exercise!
Here is another important method for constructing orientations.
Proposition 10.28 (The pullback orientation). Let M and N be smooth manifolds with
or without boundary. If F : M → N is a local diffeomorphism and if N is oriented, then
M has a unique orientation, called the pullback orientation induced by F , such that F is
orientation-preserving.
Proof. For each p ∈ M there is a unique orientation on Tp M that makes the isomorphism
dFp : Tp M → TF (p) N orientation-preserving. This defines a pointwise orientation on M ;
provided that it is continuous, it is the unique orientation on M with respect to which F
is orientation-preserving. To see that it is continuous, just choose a smooth orientation
form ω of N using Proposition 10.14 (so that ω is positively oriented) and note that F ∗ ω
is a smooth orientation form for M , determining by construction and by Proposition 10.14
the above pointwise orientation on M , which is thus continuous, as desired.
CHAPTER 11
INTEGRATION ON MANIFOLDS
1
Continuity of γ means that γ(t) approaches the same value as t approaches any of the points ai (other
than a0 or ak ) from the left or the right. Smoothness of γ in each subinterval means that γ has one-sided
velocity vectors at each such ai when approaching from the left or the right, but these one-sided velocities
need not be equal.
107
108 Chapter 11. Integration on Manifolds
Therefore,
Z k Z
X ai
ω= fi (t) dt.
γ i=1 ai−1
Proof.
(a) Follows immediately from the corresponding property of usual integrals.
(b) Since γ is constant, for any p ∈ [a, b] we have dγp = 0, and thus
Proof. Denote by s, resp. t, the standard coordinate on [c, d], resp. [a, b]. Then ω can
be written as ωt = f (t) dt for some smooth function f : [a, b] → R, and now (8.4) and
(8.6) show that φ∗ ω has the coordinate expression (φ∗ ω)s = f φ(s) φ′ (s) ds. Inserting
this into the definition of the line integral and using the change of variables formula for
ordinary integrals, we obtain
Z b
Z Z d
f (t) dt if φ is increasing,
a
φ∗ ω = f φ(s) φ′ (s) ds =
[c,d] c Z b
− f (t) dt if φ is decreasing,
a
Proof. Exercise! (First deal with the case when γ is smooth using Lemma 11.6 and
Remark 8.16, and then treat the general case using Proposition 11.3(c).)
Proposition 11.8. Let M be a smooth manifold with or without boundary and let ω ∈
X∗ (M ). If γ : [a, b] → M is a piecewise smooth curve segment in M , then the line integral
of ω over γ can also be expressed as the ordinary integral
Z Z b
ωγ(t) γ ′ (t) dt.
ω=
γ a
Proof. Suppose first that γ is smooth. By combining Proposition 11.8, [Exercise Sheet
13, Exercise 3(a)] and the fundamental theorem of calculus we obtain
Z Z b Z b
′
′
df = dfγ(t) γ (t) dt = f ◦ γ (t) = f ◦ γ (b) − f ◦ γ (a).
γ a a
Suppose now that γ is merely piecewise smooth and consider a finite partition a0 =
a < a1 < · · · < ak−1 < ak = b of [a, b] such that γ|[ai−1 ,ai ] is smooth for every 1 ≤ i ≤ k.
In view of Proposition 11.3(c), applying the above argument on each subinterval and
summing, we find that
Z k
X
df = f γ(ai ) − f γ(ai−1 ) = f γ(b) − f γ(a) ,
γ i=1
11.2.1 Integration in Rn
Definition 11.11. Let D ⊆ Rn be a domain of integration (i.e., a bounded subset of Rn
whose boundary has n-dimensional measure zero, such as a rectangle according to [Lee13,
Proposition C.18]), and let ω be a continuous n-form on D. Since ω can be written as
ω = f dx1 ∧ . . . ∧ dxn for some continuous function f : D → R, we define the integral of ω
over D to be the usual integral
Z Z Z Z
1 n 1 n
ω= f dx ∧ . . . ∧ dx := f dx . . . dx = f dV .
D D D D
(In simple terms, to compute the integral of a form such as f dx1 ∧ . . . ∧ dxn , just
“erase the wedges”.)
Proof. Follows from the (usual) change of variables formula ([Lee13, Theorem C.26])
and the pullback formula for n-forms (Proposition 8.21), taking also Lemma 10.27 into
account.
112 Chapter 11. Integration on Manifolds
As we cannot guarantee that arbitrary open or compact subsets are domains of inte-
gration, we need the following lemma in order to extend Proposition 11.13 to compactly
supported n-forms defined on open subsets.
Lemma 11.14. If U is an open subset of Rn or Hn and if K is a compact subset of U ,
then there is an open domain of integration D such that
K ⊆ D ⊆ D ⊆ U.
supp ω ⊆ E ⊆ E ⊆ V.
(See Figure 11.1.) Since diffeomorphisms take interiors to interiors, boundaries to bound-
aries, and sets of measure zero to sets of measure zero, we infer that D := G−1 (E) ⊆
U is an open domain of integration containing supp(G∗ ω). We conclude by Proposi-
tion 11.13.
Using the above proposition we can now make sense of the integral of a differential
n-form over an oriented n-manifold.
with the positive sign for a positively oriented chart, and the negative sign otherwise.
∗
(See Figure 11.2.) Since (φ−1 ) ω is a compactly supported n-form on the open subset
φ(U ) ⊆ Rn or Hn , its integral is defined as in Definition 11.12.
Section 11.2. Integration of Differential Forms 113
R
Proposition 11.17. If M and ω are as above, then M
ω does not depend on the choice
of smooth chart whose domain contains supp ω.
Proof. Let (U, φ) and (U e be two smooth charts such that supp ω ⊆ U ∩ U
e , φ) e . If both
e ◦ φ−1 : φ(U ∩ U
charts are similarly oriented, then φ e ) → φ(U
e ∩U e ) is an orientation-
preserving diffeomorphism (see the proof of Proposition 10.18 and Lemma 10.27), so
Z Z Z
−1 ∗ −1 ∗ P roposition 11.15 ∗ ∗
(φe )ω= (φ
e ) ω ========== (φe ◦ φ−1 ) (φe−1 ) ω
φ(
eU e) φ(
eU e ∩U ) φ(U ∩U
e)
Z Z
∗ ∗ ∗
= (φ−1 ) φ
e∗ (φ
e−1 ) ω = (φ−1 ) ω .
φ(U ∩U
e) | {z } φ(U )
= Id∗
If the charts are oppositely oriented, then the two definitions given by (11.1) have opposite
signs, but is compensated by the fact that φ e ◦ φ−1 is orientation-reversing, so Proposi-
tion 11.15 introduces an
R extra negative sign into the above computation. In either case,
the two definitions of M ω agree.
To integrate over an entire manifold, we combine this definition with a partition of
unity.
Definition 11.18. Let M be an oriented smooth n-manifold with or without boundary
and let ω be a compactly supported n-form on M , where n ≥ 1. Let {Ui } be a finite open
cover of supp ω by domains of positively or negatively oriented smooth charts2 , and let
{ψi } be a smooth partition of unity subordinate to this cover. We define the integral of
ω over M to be Z XZ
ω= ψi ω . (11.2)
M i M
Since for each i the n-form ψi ω is compactly supported in Ui , each of the terms in this
(finite) sum is well defined according to our previous discussion.
The following proposition shows that the integral is well defined.
Proposition 11.19. The definition (11.2) does not depend on the choice of open cover
or partition of unity.
Proof. Let {U
ej } be another open cover of supp ω by domains of positively or negatively
oriented smooth charts, and let {ψej } be a subordinate smooth partition of unity. Since
Z Z X XZ
ψi ω = ψj ψi ω =
e ψej ψi ω for every i,
M M j j M
we obtain XZ XZ
ψi ω = ψej ψi ω.
i M i,j M
2
The reason we allow for negatively oriented charts is that it may not be possible to find positively
oriented boundary charts on a 1-manifold with boundary, as noted in the proof of Proposition 10.18.
114 Chapter 11. Integration on Manifolds
Each term in this last sum is the integral of a form that is compactly supported in the
domain of a single smooth chart (e.g. in Ui ), so by Proposition 11.17 each term is well
defined, regardless
R of which coordinate map we use to compute it. The same argument,
starting with M ψej ω instead, shows that
XZ XZ
ψej ω = ψej ψi ω.
j M i,j M
R
Thus, both definitions yield the same value for M
ω.
We have a special definition in the zero-dimensional case. The integral of a compactly
supported 0-form (i.e., a real-valued function) f over an oriented 0-manifold M is defined
to be the sum Z X
f := ±f (p), (11.3)
M p∈M
where we take the positive sign at points where the orientation is positive and the negative
sign otherwise. The assumption that f is compactly supported implies that there are only
finitely many non-zero terms in this sum.
If S ⊆ M is an oriented immersed k-dimensional manifold (with or without boundary)
and
R if ω Ris a∗k-form on M whose restriction to S is compactly supported, then we interpret
S
ω as S ι ω, where ι : S ,→ M is the inclusion map. In particular, if M is a compact,
R n-manifold with boundary and if ω is an (n − 1)-form on M , then we
oriented, smooth
can interpret ∂M ω unambiguously as the integral of ι∗ ω over ∂M , where ∂M is always
understood to have the induced (Stokes) orientation; see Proposition 10.23.
Proposition 11.20 (Properties of integrals). Let M and N be nonempty oriented smooth
n-manifolds with or without boundary, and let ω and η be compactly supported n-forms
on M .
(a) Linearity: If a, b ∈ R, then
Z Z Z
aω + bη = a ω+b η.
M M M
Proof.
(a) Exercise.
(b) Exercise (follows from the usual change of variables formula).
(c) Since ω is a positively oriented orientation form on M , if (U, φ) is a positively oriented
∗
smooth chart, then (φ−1 ) ω is a positive function times dx1 ∧ . . . ∧ dxn (while if (U, φ) is
negatively oriented, then it is a negative function times the Rsame form); see the proof of
Proposition 10.14. Therefore, each term in (11.2) defining M ω is nonnegative, with at
least one strictly positive term, proving thus (c).
(d) It suffices to treat the case when ω is compactly supported in a single positively
or negatively oriented smooth chart. If (U, φ) is a positively oriented such chart and
if F is orientation-preserving, then it is easy to check that F −1 (U ), φ ◦ F is an ori-
ented smooth chart on N whose domain contains supp(F ∗ ω), so the result follows from
Proposition 11.15. The remaining cases follow from this result and (b).
Here, ∂M is understood to have the induced (Stokes) orientation, and the ω on the
right-hand side is to be interpreted as ι∗∂M ω. If ∂M = ∅, then the right-hand side is to be
interpreted as 0. When M is 1-dimensional, the right-hand integral is just a finite sum,
see (11.3).
Proof of the case M = R2 . We have to show that
Z
dω = 0, where ω = f dx + g dy ∈ Ω1c (R2 ).
R2
Since f and g have compact support, we may pick r > 0 such that both supp(f ) and
supp(g) are contained in the interior of the square [−r, r] × [−r, r]. Then
Z Z
∂f ∂g Fubini
dω = dy ∧ dx + dx ∧ dy ====
R2 R2 ∂y ∂x
Z rZ r Z rZ r
∂f ∂g
=− (x, y) dx dy + (x, y) dx dy
−r −r ∂y −r −r ∂x
Z r Z r
y=r x=r
=− f (x, y) y=−r dx + g(x, y) x=−r dy
−r | {z } −r | {z }
=0 =0
=0
116 Chapter 11. Integration on Manifolds
because the boundary orientation at γ(a) is −1, while at γ(b) is +1. Thus, Stokes’
theorem reduces to the fundamental theorem for line integrals (Theorem 11.9) in this
case. In particular, when γ : [a, b] ,→ R is the inclusion map, then Stokes’ theorem is just
the fundamental theorem of calculus.
Theorem 11.23 (Green’s theorem). Let D ⊆ R2 be a compact regular domain (i.e.,
properly embedded codimension-0 submanifold with boundary), and let P , Q be smooth
real-valued functions on D. Then
Z Z
∂Q ∂P
− dx dy = P dx + Q dy.
D ∂x ∂y ∂D
(b) S is not the boundary of an oriented, compact, smooth submanifold with boundary
in M .
Section 11.2. Integration of Differential Forms 117
Proof.
(a) If ω were exact on M , then ω = dη for some (k − 1)-form η on M , so we have
Z Z Z Z
dfn ∗ ∗ P roposition 8.26 Corollary 11.24
0 ̸= ω == ιS ω = ιS (dη) ========== d(ι∗S η) ========= 0,
S S S S
which is a contradiction.
(b) Argue again by contradiction and invoke Corollary 11.25.
• Setting
x y
f (x, y) = and g(x, y) = − ,
x2 + y2 x2 + y2
so that ω = f dy + g dx, we compute that
∂f ∂g
dω = df ∧ dy + dg ∧ dx = dx ∧ dy + dy ∧ dx
∂x ∂y
y 2 − x2 y 2 − x2
= dx ∧ dy + dy ∧ dx
(x2 + y 2 )2 (x2 + y 2 )2
= 0,
• If ω were exact, then there would exist a smooth function f : M → R such that ω = df ,
so by Stokes’ theorem and the fact that ∂γ = ∅ we would then obtain
Z Z Z
ω = df = f = 0,
γ γ ∂γ
Most of the smooth manifolds that we encountered in this course were intrinsically sub-
spaces of some Euclidean space Rn . However, the set-up of the general theory (that is,
endowing topological manifolds with a smooth structure) is designed precisely so as to
allow our objects of study to come along as abstract spaces, rather than requiring them
to be subsets of some Rn . Hence, it would be nice to see an example of a smooth manifold
which takes advantage of this abstract set-up. An elementary, yet important, example is
the real projective space RPn , which will be described in this appendix.
As with any group action, we can form the quotient set, whose points are the orbits of the
action. Concretely, we define the real projective space of dimension n, denoted by RPn ,
to be the quotient of the above action, i.e.,
In particular, notice that points of RPn are in one-to-one correspondence with one-
dimensional subspaces of Rn+1 : if [x] ∈ RPn , then [x] ∪ {0} = R · x is the one-dimensional
subspace of Rn+1 generated by x, while if L is any one-dimensional subspace of Rn+1 ,
then L \ {0} = [x] for any x ∈ L \ {0}. (This is the geometric picture you should have in
mind when thinking about RPn .) If
x = (x0 , . . . , xn )
119
120 Appendix A. The Real Projective Space
the corresponding point of RPn . Note that [x0 : . . . : xn ] = [y0 : . . . : yn ] if and only if
there exists λ ̸= 0 such that λxi = yi for all i.
is a topology on RPn . Moreover, if we endow RPn with this topology, then the quotient
map π : Rn+1 \ {0} → RPn is continuous, and a map f : RPn → X from RPn to some
topological space X is continuous if and only if so is the composite map f ◦ π. The same
is true for any subset A ⊆ RPn endowed with the subspace topology.
At this point, there are several things that need to be checked about the topological
space RPn .
Exercise A.1: Show that RPn is Hausdorff by going through the following steps:
(i) Show that the quotient map π : Rn+1 \ {0} → RPn is open.
U × V | U, V ∈ TRPn
is a basis for the topology of RPn × RPn by definition of the product topology.]
121
Solution:
(i) Note that we have
[
π −1 π(U ) =
λ · U.
λ∈R×
and notice that as π is open, this is a collection of open subsets of RPn . Let us show that
B′ is a basis for the topology of RPn . To this end, let U ⊆ RPn be open and [x] ∈ U a
point. Then x ∈ π −1 (U ), and thus there exists B ∈ B such that x ∈ B ⊆ π −1 (U ). But
then [x] ∈ π(B) ⊆ U . Hence, B′ is a countable basis for RPn .
φi : Ui → Rn
x0 xi−1 xi+1 xn
[x0 : . . . : xn ] 7→ ,..., , ,..., .
xi xi xi xi
Solution:
(ii) The ratio xj /xi is invariant under scaling x, and thus φi is well-defined. To check
that it is continuous, it suffices to check that φi ◦ π is a continuous map from
π −1 (Ui ) = Rn+1 n
xi ̸=0 to R , but this is straightforward by the defining formula. Finally,
to show that φi is a homeomorphism, we construct a continuous inverse. Consider
the map
Ψi : Rn → π −1 (Ui )
(y1 , . . . , yn ) 7→ (y1 , . . . , yi , 1, yi+1 , . . . , yn ),
Exercise A.4:
(ii) Show that the restriction of π to S n ⊆ Rn+1 \ {0} is still surjective. Conclude that
RPn is compact.
lines which do not intersect the plane Rn × {1}, which (as you may convince yourself) are
precisely the lines contained in Rn × {0}. Hence, we may somewhat suggestively write
Exercise A.5: Let 0 ≤ i < j ≤ n. Show that the transition map from (Ui , φi ) to (Uj , φj )
is a diffeomorphism by computing that
φj ◦ φ−1 n n
i : Rxj ̸=0 → Rxi+1 ̸=0
1
(x1 , . . . , xn ) 7→ (x1 , . . . , xi , 1, xi+1 , . . . , xj−1 , xj+1 , . . . , xn ) ,
xj
and
φi ◦ φ−1 n
j : Rxi+1 ̸=0 → Rxj ̸=0
n
1
(x1 , . . . , xn ) 7→ (x1 , . . . , xi , xi+2 , . . . , xj , 1, xj+1 , . . . , xn ) .
xi+1
is a smooth atlas for RPn , and the induced by Proposition 1.8(a) smooth structure on
RPn is referred to as the standard one. Thus, we now have a smooth manifold, namely
RPn , which is not intrinsically defined as a subset of Rn !
124 Appendix A. The Real Projective Space
(ii) A map F : RPn → M to a smooth manifold M is smooth if and only if the composite
map F ◦ π : Rn+1 \ {0} → M is smooth.
Exercise A.8: Let P : Rn+1 \ {0} → Rk+1 \ {0} be a smooth map, and suppose that for
some d ∈ Z we have P (λx) = λd P (x) for all λ ∈ R× and x ∈ Rn+1 \ {0}. Show that the
map Pe : RPn → RPk given by Pe [x] = [P (x)] is well-defined and smooth.
Exercise A.10: Show that the quotient map π : Rn+1 \ {0} → RPn is a smooth submer-
sion, and that the kernel of the differential dπp : Tp Rn+1 \ {0} → T[p] RPn is the subspace
generated by p.
F : R2 → RP2 , (x, y) 7→ [x : y : 1]
Show that there is a smooth vector field Y on RP2 that is F -related to X, and compute
its coordinate representation in terms of each of the charts defined in Exercise 3(ii).
APPENDIX B
SARD’S THEOREM AND WHITNEY’S THEOREMS
Theorem B.3 (Whitney’s immersion theorem). Every smooth n-manifold admits a smooth
immersion into R2n .
The above two theorems are sometimes referred to as the easy or weak Whitney
embedding and immersion theorems, because Whitney obtained later the following im-
provements.
For the proofs of all the above results, as well as a discussion of sets of measure zero
(in Rn or in smooth manifolds) we refer to [Lee13, Chapter 6 and Appendix C].
125
APPENDIX C
MULTILINEAR ALGEBRA
127
128 Appendix C. Multilinear Algebra
Definition C.3. Let V and W be real vector spaces and let A : V → W be a linear map.
The dual map of A is the linear map A∗ : W ∗ → V ∗ defined by
(A∗ ω)(v) := ω(Av), ω ∈ W ∗ , v ∈ V.
It is straightforward to check that the dual map satisfies the following properties:
(a) (A ◦ B)∗ = B ∗ ◦ A∗ .
(b) (IdV )∗ = IdV ∗ .
Proposition C.4. The assignment that sends a vector space to its dual space and a linear
map to its dual linear map is a contravariant functor from the category of real vector spaces
to itself.
Another important fact about the dual of a finite-dimensional vector space is the
following.
Proposition C.5. Let V be a finite-dimensional real vector space. For any given v ∈ V ,
define a linear functional ξ(v) by
ξ(v) : V ∗ → R
ω 7→ ξ(v)(ω) := ω(v).
Then ξ(v) ∈ (V ∗ )∗ ; that is, ξ(v) is a linear functional on V ∗ . Moreover, the map
ξ : V → (V ∗ )∗
v 7→ ξ(v)
is an R-linear isomorphism, which is canonical (it is defined without reference to any
basis).
Proof. The proof that both ξ(v) and ξ are R-linear maps are left as exercises. Since by
Proposition C.2 we have
dim V = dim V ∗ = dim(V ∗ )∗ ,
it suffices to prove that ξ is injective. To this end, let v ∈ V be non-zero, complete it to
a basis {v = E1 , E2 , . . . , En } of V , and let (εi ) be its dual basis. Then
ξ(v)(ε1 ) = ε1 (v) = ε1 (E1 ) = 1,
so ξ(v) ̸= 0. Therefore, ker ξ = 0; in other words, ξ is injective, as desired.
Due to Proposition C.5, the real number ω(v) obtained by applying a covector ω to a
vector v is sometimes denoted by either of the more symmetric-looking notations ⟨ω, v⟩
or ⟨v, ω⟩; both expressions can be thought of either as the action of the covector ω ∈ V ∗
on the vector v ∈ V , or as the action of the linear functional ξ(v) ∈ V ∗∗ on the element
ω ∈ V ∗ . There should be no cause for confusion with the use of the same angle bracket
notation for inner products: whenever one of the arguments is a vector and the other a
covector, the notation ⟨ω, v⟩ is always to be interpreted as the natural pairing between
vectors and covectors, not as an inner product.
There is also a symmetry between bases and dual bases for a finite-dimensional vector
space V : any basis for V determines a dual basis for V ∗ , and conversely, any basis for V ∗
determines a dual basis for V ∗∗ ∼= V . If (εi ) is the basis for V ∗ dual to a basis (Ej ) for V ,
then (Ej ) is the basis dual to (εi ), because both statements are equivalent to the relation
⟨εi , Ej ⟩ = δji .
Section 2. Multilinear Maps and Tensors 129
F ⊗ G : V 1 × · · · × V k × W1 × · · · × Wl → R
(v1 , . . . , vk , w1 , . . . , wl ) 7→ F (v1 , . . . , vk ) G(w1 , . . . , wl )
Exercise C.8:
(a) Show that, given F and G as above, the function F ⊗ G is multilinear, that is,
F ⊗ G ∈ L(V1 , . . . , Vk , W1 , . . . , Wl ; R).
is bilinear, i.e., multilinear with two variables, and associative, i.e., for any multilin-
ear real-valued functions F, G, H, we have F ⊗ (G ⊗ H) = (F ⊗ G) ⊗ H.
Proof. First, given F ∈ L(V1 , . . . , Vk ; R), define for each multi-index I = (i1 , . . . , ik ) with
1 ≤ ij ≤ nj for all 1 ≤ j ≤ k, a number FI ∈ R by
(1) (k)
FI := F Ei1 , . . . , Eik .
ε⊗I := εi(1)
1
⊗ · · · ⊗ εi(k)
k
.
where the sum is taken over all multi-indices as above, and thereby show that B spans
L(V1 , . . . , Vk ; R). To this end, take (v1 , . . . , vk ) ∈ V1 × · · · × Vk . For integers ij between 1
i (j) (j)
and nj , let vjj ∈ R be the coefficient of vj with respect to the basis E1 , . . . , Enj , i.e.,
i i
vjj = ε(j)
j
(vj ).
⊗I
agree at any k-tuple and thus are equal, so B indeed spans
P
Hence F and I FI ε
L(V1 , . . . , Vk ; R).
Finally, in order to see that B is linearly independent, suppose that we have
X
λI ε⊗I = 0
I
for
some real numbers λI ∈ R indexed by multi-indices I. Evaluating both sides at
(1) (k)
Ei1 , . . . , Eik for some fixed multi-index I = (i1 , . . . , ik ), we obtain by the same com-
putation as above that λI = 0. Hence, B is linearly independent.
Section 2. Multilinear Maps and Tensors 131
The proof of Proposition C.9 shows also that the components Fi1 ...ik of a multilinear
function F in terms of the basis elements in B are given by
(1) (k)
Fi1 ...ik = F Ei1 , . . . , Eik .
Thus, F is completely determined by its action on all possible sequences of basis vectors.
Remark C.10. You might have already encountered the abstract construction of the ten-
sor product of vector spaces. If so, then regarding the above discussion (which shows that
the real vector space L(V1 , . . . , Vk ; R) can be viewed as the set of all linear combinations
of objects of the form ω 1 ⊗ · · · ⊗ ω k , where ω i ∈ Vi∗ are covectors), one should remark
the following: given finite-dimensional real vector spaces V1 , . . . , Vk , there is a canonical
isomorphism
V1∗ ⊗ · · · ⊗ Vk∗ ∼= L(V1 , . . . , Vk ; R),
which is induced by the multilinear map
= ω 1 (v1 ) · · · ω k (vk ).
Under this canonical isomorphism, abstract tensors correspond to the concrete tensor
product of multilinear functions defined above. As it is a natural isomorphism, we may
use the expression V1∗ ⊗· · ·⊗Vk∗ as a notation for L(V1 , . . . , Vk ; R) (this is a typical example
of slight abuse of notation, where one identifies naturally isomorphic objects). Finally,
using Proposition C.5, we also obtain a canonical identification
V1 ⊗ · · · ⊗ Vk ∼
= L(V1∗ , . . . , Vk∗ ; R).
Therefore, we may view the above construction as a concrete construction of the abstract
tensor product.
Let us now turn our attention to various spaces of multilinear functions on a finite-
dimensional real vector space that naturally appear in (differential) geometry.
Definition C.11. Let V be a finite-dimensional real vector space. For any integer k ≥ 1,
we denote by T k (V ∗ ) the space of k-multilinear functions on V , i.e.,
T k (V ∗ ) := L(V, . . . , V ; R) ∼
= |V ∗ ⊗ .{z
.. ⊗ V ∗.
| {z } }
k times k copies
T 1 (V ∗ ) = V ∗ .
132 Appendix C. Multilinear Algebra
(see the proof of Proposition C.9) and for an integer m ∈ Z≥1 , denote by [m] the set
{1, . . . , m}. Then the set
⊗I
ε | I ∈ [n][k]
dimR T k (V ∗ ) = nk .
For example, T 2 (V ∗ ) is the space of bilinear forms on V – note that a covariant 2-tensor
on V is simply a real-valued bilinear function of two vectors – and every bilinear form on
V can be written as β = βij εi ⊗ εj for some uniquely determined n × n matrix (βij ).
σ
α: V × · · · × V → R
(v1 , . . . , vk ) 7→ α vσ(1) , . . . , vσ(k) .
3 Symmetric Tensors
In all probability, you have already encountered the concept of inner product on a finite-
dimensional real vector space V . It is a bilinear map ⟨·, ·⟩ : V ×V → R which is symmetric
and positive definite; in particular, ⟨·, ·⟩ is a covariant 2-tensor on V , having the additional
property that its value is unchanged when the two input arguments are exchanged; namely,
we have ⟨v1 , v2 ⟩ = ⟨v2 , v1 ⟩ for any v1 , v2 ∈ V . We now generalize this notion to any
covariant k-tensor on V .
Definition C.13. Let V be a finite-dimensional real vector space.
(a) A covariant k-tensor α ∈ T k (V ∗ ) on V is said to be symmetric if its value is un-
changed by interchanging any pair of its arguments; namely, for all v1 , . . . , vk ∈ V
and all 1 ≤ i < j ≤ k, we have
α(v1 , . . . , vi , . . . , vj , . . . , vk ) = α(v1 , . . . , vj , . . . , vi , . . . , vk ).
4 Alternating Tensors
Recall that the determinant may be regarded as a function det : Rn ×· · ·×Rn → R, taking
as input n column vectors with n entries each, and having as output the determinant of
the n × n matrix formed by these n column vectors. This map is multilinear, so det is
a covariant n-tensor on Rn . Moreover, it has the property that its value changes sign
whenever two of its input entries are interchanged; in other words, det is an alternating
n-tensor. We now generalize this notion to arbitrary covariant k-tensors.
Definition C.15. Let V be a finite-dimensional real vector space.
(a) A covariant k-tensor α ∈ T k (V ∗ ) on V is said to be alternating (or anti-symmetric
or skew-symmetric) if its value changes sign whenever any two of its arguments are
interchanged; namely, for all v1 , . . . , vk ∈ V and 1 ≤ i < j ≤ k, we have
α(v1 , . . . , vi , . . . , vj , . . . , vk ) = −α(v1 , . . . , vj , . . . , vi , . . . , vk ).
134 Appendix C. Multilinear Algebra
Note that every covariant 2-tensor β can be expressed as a sum of an alternating and
a symmetric tensor, because
1 1
β(v, w) = β(v, w) − β(w, v) + β(v, w) + β(w, v)
2 2
= α(v, w) + σ(v, w),
where
1
β(v, w) − β(w, v) ∈ Λ2 (V ∗ )
α(v, w) :=
2
is an alternating 2-tensor on V and
1
β(v, w) + β(w, v) ∈ Σ2 (V ∗ )
σ(v, w) :=
2
is a symmetric 2-tensor on V . However, this is not true for tensors of higher rank, as the
following exercise demonstrates.
Exercise C.16: Let (e1 , e2 , e3 ) be the standard dual basis for (R3 )∗ . Show that e1 ⊗e2 ⊗e3
is not equal to a sum of an alternating tensor and a symmetric tensor.
Recall that there is a group homomorphism sgn : Sk → {±1}, which maps a permuta-
tion σ ∈ Sk to 1 if it is a product of an even number of transpositions (even permutation),
and to −1 otherwise (odd permutation). We may use it to describe alternating tensors as
follows.
σ
where α was defined in Definition C.12. Show that Alt is well-defined and linear, and
that the following are equivalent:
(a) α is alternating,
(c) α = Alt(α),
Example C.18. Let us explicitly compute Alt for 1-, 2- and 3-tensors.
• If β is a 2-tensor, then
1
Alt(β)(u, v) = β(u, v) − β(v, u) .
2
• If γ is a 3-tensor, then
1
Alt(γ)(u, v, w) = γ(u, v, w) + γ(v, w, u) + γ(w, u, v)
6
−γ(v, u, w) − γ(u, w, v) − γ(w, v, u) .
where
ε⊗I = εi1 ⊗ · · · ⊗ εik ∈ T k (V ∗ )
is the elementary k-tensor. Therefore, if v1 , . . . , vk ∈ V , then the value of εI at the k-tuple
(v1 , . . . , vk ) is given by the formula
X
εI (v1 , . . . , vk ) = (sgn σ) ε⊗I vσ(1) , . . . , vσ(k)
σ∈Sk
X Y
εij vσ(j)
= (sgn σ)
σ∈Sk 1≤j≤k
εi1 (v1 ) · · · εi1 (vk )
= det ... ... .. .
.
ik ik
ε (v1 ) · · · ε (vk )
Example C.19. In terms of the standard dual basis (e1 , e2 , e3 ) for (R3 )∗ , we have
1
13 v w1
e (v, w) = det 3 = v 1 w3 − v 3 w1 ,
v w3
Proof. Exercise!
Lemma C.20 tells us that from the generating set εI | I ∈ [n][k] of Λk (V ∗ ), we may
discard all those εI ’s for which I has a repeated index, and for any I having no re-
peated index, we need only take one element from the set {εIσ | σ ∈ Sk } and dis-
card the rest. A nice choice is thus the following: notice that for any multi-index I
having no repeated indices, there exists a unique permutation σ ∈ Sk such that Iσ is
strictly increasing, i.e., iσ(1) < · · · < iσ(k) . Therefore, according to Lemma C.20, the set
ε | I ∈ [n][k] is strictly increasing still generates Λk (V ∗ ), and there is no obvious re-
I
dundancy in it. Essentially due to Lemma C.20(c), this set is linearily independent, and
thus we obtain the following result:
Proposition C.21. With the same notation as above, the set
I
ε | I ∈ [n][k] is a strictly increasing multi-index
Proof. Assume first that k > n. Since then every k-tuple of vectors is linearly dependent,
it follows from Exercise C.17(d) that Λk (V ∗ ) = {0}.
Assume now that k ≤ n. We need to show that
is linearly independent and spans Λk (V ∗ ). The fact that E generates Λk (V ∗ ) was already
discussed above. Suppose now that we have some linear relation
X
λ I εI = 0
I∈[n][k] strictly increasing
for some λI ∈ R. If we fix a strictly increasing multi-index J ∈ [n][k] , then evaluating the
above relation at (Ej1 , . . . , Ejk ) gives λJ = 0 according to Lemma C.20(c). Thus, E is
linearly independent. In conclusion, E is a basis of Λk (V ∗ ), as desired.
In particular, if V is a real vector space of dimension n, then the above proposition
implies that Λn (V ∗ ) is 1-dimensional, spanned by the elementary n-covector ε(1,...,n) . As
discussed in the beginning of this subsection, ε(1,...,n) sends an n-tuple (v1 , . . . , vn ) to the
determinant of the matrix (vji )1≤i,j,≤n , where vji = εi (vj ) is the i-th component of vj with
respect to the chosen basis of V . Note that when V = Rn with the standard basis, the
2
covector ε(1,...,n) (which by definition is a function from (Rn )n = Rn to R) is precisely the
usual determinant function.
One consequence of this observation is the following useful description of the behavior
of an n-covector on an n-dimensional vector space under linear maps. Recall that if
T : V → V is a linear map, then the determinant of T is defined to be the determinant of
the matrix representation of T with respect to any basis (recall that any two such matrix
representation are conjugations of each other and hence have the same determinant, so
this is well-defined).
ω(T E1 , . . . , T En ) = c ε(1,...,n) (T1 , . . . , Tn ) = c det (εj (Ti ))1≤i,j≤n = c det (Tij )1≤i,j≤n .
for any sequence of basis vectors (Ep1 , . . . , Epk+l ). We do this by considering several cases.
Case 1: The multi-index P = (p1 , . . . , pk+l ) has a repeated index. Then by part (e) of
Exercise C.17, both sides of (⋆) evaluate to 0.
Case 2: P contains an index that does not appear in either I or J. In this case, the
right-hand side of (⋆) is zero by part (c) of Lemma C.20. Similarly, each term in the
expansion of the left-hand side of (⋆) involves either I or J evaluated on a sequence of
basis vectors that is not a permutation of I or J, respectively, so the left-hand side is also
zero.
Case 3: P = I ⌢ J and P has no repeated indices. In this case, the right-hand side of
(⋆) is equal to 1, again by part (c) of Lemma C.20, so we need to show that the left-hand
side is also equal to 1. By definition,
εI ∧ εJ (Ep1 , . . . , Epk+l ) =
(k + l)!
= Alt(εI ⊗ εJ )
k!l!
1 X
= (sgn σ)εI (Epσ(1) , . . . , Epσ(k) )εJ (Epσ(k+1) , . . . , Epσ(k+l) ).
k!l! σ∈S
k+l
Section 4. Alternating Tensors 139
By Lemma C.20 again, the only terms in the sum above that give nonzero values are
those in which σ permutes the first k indices and the last l indices of P separately. In
other words, σ must be of the form σ = τ η, where τ ∈ Sk acts by permuting {1, . . . , k}
and η ∈ Sl acts by permuting {k + 1, . . . , k + l}. Since then sgn σ = (sgn τ )(sgn η), we
have
εI ∧ εJ (Ep1 , . . . , Epk+l ) =
1 X
= (sgn τ )(sgn η) εI (Epτ (1) , . . . , Epτ (k) ) εJ (Epk+η(1) , . . . , Epk+η(l) )
k!l! τ ∈S
k
η∈Sl
! !
1 X 1 X
= (sgn τ ) εI (Epτ (1) , . . . , Epτ (k) ) (sgn η) εJ (Epk+η(1) , . . . , Epk+η(l) )
k! τ ∈S l! η∈S
k l
I J
= Alt(ε )(Ep1 , . . . , Epk ) Alt(ε )(Epk+1 , . . . , Epk+l )
= εI (Ep1 , . . . , Epk ) εJ (Epk+1 , . . . , Epk+l )
=1
where we used that Alt fixes alternating tensors by Exercise C.17, and again used part
(c) of Lemma C.20 (recall that we are in the case P = I ⌢ J).
ω ∧ η = (−1)kl η ∧ ω.
εi 1 ∧ . . . ∧ εi k = εI .
Proof. Exercise!
140 Appendix C. Multilinear Algebra
Due to Proposition C.25(c), we generally use the notations εI and εi1 ∧ . . . ∧ εik
interchangably.
An element η ∈ Λk (V ∗ ) is said to be decomposable if it can be expressed in the form
η = ω 1 ∧ . . . ∧ ω k for some covectors ω 1 , . . . , ω k ∈ V ∗ . Note that not every k-covector
is decomposable when k > 1; however, it follows from Proposition C.21 and Proposi-
tion C.25(c) that every k-covector can be written as a linear combination of decomposable
ones.
BIBLIOGRAPHY
141
INDEX
143
144 Index