Matd67h3 2
Matd67h3 2
Elden Elmanto
Chapter 1. Introduction 5
Appendix. Bibliography 37
3
CHAPTER 1
Introduction
These are notes for MATD67H3, the UTSC version of a first class in differentiable manifolds.
5
CHAPTER 2
For our informal discussion, we assert that anytime we have a set Z ⊂ Rn defined by
some polynomial equation (in the above case Z is defined by x2 + y 2 = 1) then Z inherits a
standard differentiable structure. Roughly: for any x ∈ Z and any open subset U ⊂ Z
(these are of the form U ∩ Z where U ⊂ Rn is a union of open balls in Rn ), then we declare
differentiable functions to be those that come from infinitely differentiable functions f : U → R.
Of fundamental interest are the spheres:
Sn = {(x1 , x2 , · · · , xn+1 ) : x21 + x22 + · · · x2n+1 = 1} ⊂ Rn+1 ,
with its standard differentiable structure.
Now, when n = 1, we have the standard differentiable structure, and also one defined in the
previous paragraph using θ+ and θ− . It turns out that they are diffeomorphic. In topology,
we have the notion of a homeomorphism: two spaces X and Y are homeomorphic if there are
continuous functions
f :X→Y g:Y→X
such that g ◦ f, f ◦ g are the identities. A diffeomorphism is a homeomorphism that preserves
differentiable structure. For example, f, g could both be the identity map on underlying sets,
but one can imagine equipping X and Y with different differentiable structures and X and Y
would not be a diffeomorphic. In the case of the sphere, we say that a differentiable structure
on Sn is exotic if it is not diffeomorphic to the standard structure.
Question 1.0.1. Is there any exotic differentiable structure on the sphere?
It turns out that when n = 1, 2, 3, 5, 6 (Note that 4 is skipped) one cannot find any exotic
differentiable structure on Sn . Milnor shocked mathematics when he proved:
Theorem 1.0.2 (Milnor 1956, [Mil56]). There exists an exotic differentiable structure on
S7 .
The construction is not easy and even proving that his sphere is not the standard one
requires technology outside the scope of this class. Roughly, however, Milnor’s exotic 7 sphere
is constructed out of arranging S3 and S4 to “glued” in a strange way. He then proves (and in
fact develops an entire technology for proving) that it is not the boundary of any 8-dimensional
differentiable manifold: note that Sn is the boundary of
Dn+1 := {(x1 , x2 , · · · , xn+1 ) : x21 + x22 + · · · x2n+1 6 1}.
The concept of a manifold with boundary is something we will encounter in this class. In fact,
this result comes as a surprise to Milnor himself. At this point in history, we have understood
smooth structures better than ever before. As a sampler: there are 27 exotic smooth structures
on S7 , while there are 523263 on S19 and we know that S61 is the last odd-dimension where no
exotic smooth structure can exist (we only knew this in 2017). But a question remains:
Question 1.0.3. Is there an exotic 4-sphere?
Lemma 2.2.4 says that continuity is the preservation of a topological property of subsets,
namely of openness. It is more abstract than the − δ argument but also conceptually simpler.
As an application, we give a soft-analysis argument of the following result which might be
familiar in an analysis class.
2. SOME RECOLLECTIONS ON ANALYSIS 11
One of the first results that one proves about this definition is that Dfx0 is, in fact, unique and
so the notation is justified; we also note that differentiable functions are continuous.
In any case, the higher dimension generalization of being differentiable is as follows.
Definition 2.2.6. A function f : Rn → Rm is said to be differentiable at x0 ∈ Rn if
there exists a linear transformation
Dfx0 : Rn → Rm
such that
|f (x0 +h)−f (x0 )−Dfx0 (h)|
lim |h| = 0.
h→0
We call the transformation Dfx0 the derivative of f at x0 if it exists. This linear trans-
formation is unique.
Lemma 2.2.7. The derivative is unique.
Proof. Let T : Rn → Rm be a transformation that satisfies
|f (x0 +h)−f (x0 )−T(h)|
lim h = 0.
h→0
2.3. The Jacobian and partial derivatives. Since Dfx0 : Rn → Rm is a linear trans-
formation we can, up to choosing bases, represent Dfx0 as a matrix. This matrix, in fact,
encodes important properties about the derivative. Suppose that f : Rn → R is a func-
tion, then we can hold all variables constant in the domain except at the i-th spot, i.e., fix
a1 , · · · , ai−1 , ai+1 , · · · an ∈ R and consider the function
f :R→R x 7→ f (a1 , · · · , ai−1 , x, ai+1 · · · an ).
In this way, the i-th partial derivative at x0 = (a1 , · · · , an ) is the limit
f (a1 ,··· ,ai−1 ,ai +h,ai+1 ···an )−f (a1 ,··· ,ai−1 ,ai ,ai+1 ···an )
lim h =: Di (f )(x0 )
h→0
Remark 2.3.7 (Additional structure on C∞ (A)). There are two more important things to
note about smooth functions which are not necessarily used in this class, but are fundamental.
First, let f, g ∈ C∞ (A). Then note that we can define
The functions f + g and f · g are both smooth. Hence C∞ (A) forms the structure of a ring
with the constant functions at 1 and 0 as the multiplicative and additive identity respectively.
Furthermore, we have a compatible action of R on C∞ (A): if r ∈ R
(r · f )(x) := rf (x)
f 7→ f 0 : C∞ (A) → C∞ (A)
In essence, the main functions of interest in differential topology are smooth functions
f : Rn → R. In fact, at this point, we can define a manifold as follows: first, let X, Y be subsets
of Rn . We say that a function f : X → Y is a diffeomorphism if f is a bijection such that f
and f −1 are both smooth functions. Then a smooth manifold of dimension n is a subset
M ⊂ Rk such that any x ∈ M has a neighborhood W ⊂ M such that W is diffeomorphic to
W0 where W0 ⊂ Rn . This is a perfectly good definition except that it is rather unsatisfying:
for example what does k have to do with n and it is not clear if it depends on the inclusion
M ⊂ Rk . We want to get rid of these dependence and so we introduce the abstract concept of
a smooth manifold.
In the other class we will learn many interesting, but perhaps pathological, topological
spaces. For example, one can endow a finite set with a topology whose role seems to be
combinatorial in nature. We want to only consider “geometric” topological spaces in this class
and the axioms that let us do these is as follows.
Definition 3.1.6. A space X is said to be Hausdorff or separated if for any two distinct
points x, y there exists open neighborhoods x ∈ Ux , y ∈ Uy such that Ux ∩ Uy = ∅. It is called
second countable or completely separable if its topology has a countable basis.
Remark 3.1.7. Suppose that X is a Hausdorff space with the property that any x ∈ X
admits an open neighborhood homeomorphic to an open subset of Euclidean space. Then it is
second countable if and only if it is metrizable, that is, admits a metric space. We certainly want
differentiable manifolds to have a notion of distance and hence second countability is desired.
At this point, we have completed Definition 2.0.1 and understood what it means for a space
to have a smooth atlas. However, let us explain a quick warning.
Remark 3.1.8 (Invariance of domain). Technically, Definition 2.0.1 is a little abusive. We
have to specify that our atlas is made up of open’s of Rn for a specified n because there is no
reason for this particular n to be well-defined. What allows us to get away with this abuse is a
result called invariance of domain, proved by Brouwer using algebraic topology.
Theorem 3.1.9 (Brouwer). Let U ⊂ Rn be a open subset and f : U → Rn a continuous,
injective function. Then f (U) ⊂ Rn is open.
Using Brouwer’s theorem, we see that a space cannot have an atlas of different dimensions.
Indeed, if we have a chart which constitutes a homeomorphism f : U → V where U ⊂ Rk and
0
V ⊂ Rk where k > k 0 , then we can look at the composite
0
U → V ⊂ Rk ,→ Rk .
This produces a map which is a continuous injection and thus, by Brouwer’s theorem, an open
0
map. But no open subset of Rk can be exclusively contained in Rk , unless it is empty.
Before we go on to define what a differentiable manifold, let us extract an important piece
of data from a smooth atlas.
−1
Remark 3.1.10 (Atlases and cocycles). The maps ψαβ : ϕ−1 α (Vα ∩ Vβ ) → ϕβ (Vα ∩ Vβ ),
which are maps between open subsets of Eucildean space, are called transition functions and
they satisfy
ψαα = id ψαβ ◦ ψβγ = ψαγ .
These conditions make precise what we mean by “compatibility” and is often called the cocycle
condition.
Furthermore, the above two conditions imply that ψαβ ◦ ψβα = id. Hence ψαβ is a diffeo-
morphism: a smooth bijection with a smooth inverse.
In fact, there is another way to specify a differentiable manifolds by glueing together cocy-
cles; we fix n. Let us take as input a set I with elements denoted by i ∈ I. For each i ∈ I we
are given Xi ⊂ Rn an open and for each Xi we are given an open
Uij ⊂ Xi ∀j ∈ I.
For each pair i, j ∈ I we are given diffeomorphisms
ψij : Uij → Uji
−1
for which 1) Uii = Xi , 2) ψij (Uji ∩ Ujk ) = Uij ∩ Uik and the diagram
ψik
Uij ∩ Uik Uki ∩ Ukj
ψij ψjk
Uji ∩ Ujk .
16 2. WHAT IS DIFFERENTIAL TOPOLOGY?
commutes. Then there exists topological manifold X and a smooth atlas of dimension n,
{(Xi , Vi , ϕi )} such that
(1) ϕi : Xi → Vi is a homeomorphism;
(2) ϕi (Uij ) = Vi ∩ Vj
(3) ψij = ϕ−1 j |Vi ∩Vj ◦ ϕi |Uji .
This will be left as exercise in homework 2.
Now, two smooth atlases on X are said to be compatible if their union is an atlas and an
atlas is maximal if it is one with the property that any atlas compatible with it must be in it.
Definition 3.1.11. An n-dimensional smooth manifold or, simply, an n-manifold is
a second-countable, Hausdorff space X with a maximal atlas.
In practice, one finds a smooth atlas on X and then generate a maximal one by the following
lemma:
Lemma 3.1.12. Every smooth atlas is contained in a unique maximal atlas.
Proof. See [Kup, Lemma 2.2.2].
CHAPTER 3
We now discuss examples on smooth manifolds and how to get new smooth manifolds out of
old ones. From now on, by a manifold we really mean a smooth one unless otherwise stated. We
want to illustrate several methods for showing that something is a manifold. We will repeatedly
use the following fact without further comment.
Lemma 0.0.1. Let X be a space. If A ⊂ X is a subset, then the subspace topology on A
satisfies:
(1) if X is Hausdorff, then A is;
(2) if X is second countable, then A is.
1. Spheres
Let us now give Sn ⊂ Rn+1 the structure of a smooth manifold of dimension n. Evidently,
it is second countable and Hausdorff by Lemma 0.0.1 since it is a subspace of Rn+1 . Consider
U+ := Sn r {(1, 0, · · · , 0)} U− := Sn r {(−1, 0, · · · , 0)}.
We then define
ϕ−1
+ : U+ → R
n 1
1−x0 (x1 , · · · , xn ),
ϕ−1
− : U− → R
n 1
x0 +1 (x1 , · · · , xn ).
Indeed, ϕ−1 −1
+ and ϕ− are diffeomorphisms which generalizes the (inverse of) the example of S
1
Here’s our strategy: as Z := f −1 (0) is already second countable and Hausdorff, we just
need to produce an atlas and then take something maximal containing it. In particular, the
following is enough: for each x ∈ Z, find an open subset U ⊂ Z such that U is diffeomorphic
to a subset of Rn . In particular, this is a local check: just choose some U first and then find a
U0 ⊂ U (so, possibly after shrinking U) which is indeed diffeomorphic to a subset of Rn .
In multivariable calculus, we have already learned a method of showing that some function
is diffeomorphic, at least for a map f : Rn → Rn .
Theorem 1.0.3 is plainly very useful: it converts the task of showing that f is diffeomorphism
into one of linear algebra, at least if we are willing to shrink around x. The inverse function
theorem can be used to study maps into Rm when m 6 n. To state this result, let us introduce
a piece of terminology. Let U ⊂ Rn be an open subset. A map f : U → Rm is a submersion
at x ∈ U if the linear map Dfx : Rn → Rm is surjective.
Here is how we can use Theorem 1.0.4. Let y ∈ g −1 (g(x)); we want to find a diffeomorphism
between an open subset U1 ⊂ g −1 (g(x)) (in the subspace topology) and an open subset of
Rn−m ; this will then assemble (as y varies) into an atlas for g −1 (g(x)). The commutativity of
the diagram in part (2) of Theorem 1.0.4 then says that we have a diffeomorphism
g −1 (g(x)) ∩ g −1 (V) ∼
= g −1 (g(x)) ∩ U ∼
= π −1 (0) = {0} × Rn−m ,
which is the desired diffeomorphism. This provides an easy way to prove that spheres are
smooth.
Example 1.0.5. We have the function g : Rn+1 → R given by (x0 , · · · , xn ) 7→ x20 + · · · + x2n .
Its derivative is given by the matrix
2x0 2x1 · · · 2xn .
This matrix has full rank if and only if (x0 , · · · , xn ) 6= 0. Therefore Theorem 1.0.4 states that
g −1 (c) is a smooth manifold if and only if c 6= 0.
3. TORI 19
2. Projective spaces
Sn is just one of other possible ways to generalize S1 . An important class of spaces that are
ubiquitous throughout mathematics are the projective spaces. We first introduce real projec-
tive spaces. The real projective space RPn is the space of “lines in Rn+1 through the origin.”
To make this precise, we consider the following equivalence relation on Rn+1 r {0}:
(x0 , · · · , xn ) ∼ (rx0 , · · · , rxn ) r ∈ R r {0}.
The quotient is denoted by
RPn := Rn+1 r 0/ ∼ .
Points in RPn (in other words, equivalence classes) are denoted by [x0 : x1 : · · · : xn ]. We have
the quotient map q : Rn+1 r {0} → RPn and the inverse image over each point is a copy of
R r {0}. It is given the quotient topology: a subset U ⊂ RPn is open if and only if q −1 (U) is
open. One can check easily that it is second countable and Hausdorff.
The interesting part is to write down an smooth atlas for RPn . Consider the open subsets
Ui = {[x0 : x1 : · · · : xn ] : xi 6= 0};
n
then the maps ϕi : Ui → R defined by
ϕi ([x0 : x1 : · · · : xn ]) = (x1 /xi , · · · , x[
i /xi , · · · , xn /xi )
defines the (inverse) of an atlas. We also have an analog for complex space which is discussed
in the homework.
3. Tori
We now discuss tori. To begin with, we note:
Lemma 3.0.1 (Products of manifolds). Let M and N be manifolds of dimension m and n
respectively. Then M × N admits the structure of an m × n-manifold.
Proof. After noting that second countability and Hausdorfness are stable under products
(with the product topology on M × N, we need only equip M and N with an atlas. But now if
{(Uα , Vα , ϕα )} is an atlas for M and {(U0β , Vβ0 , ϕ0β )} is an atlas for N, then we can produce an
atlas for M × N given by {Uα × U0β , Vα × Vβ0 , ϕα × ϕ0β )}.
The n-torus is the n-manifold given by the n-fold product of S1 ’s.
Tn = S1 × · · · × S1
This is yet another way to specify a manifold — by producing them out of old ones. We
want to discuss a construction of the 2-torus which is somewhat “visual” and is useful in practice.
Consider the “box” I × I where I = [0, 1]; its boundary is defined as ∂(I × I) := {(x, y) : x ∈
{0, 1} or y ∈ {0, 1}}. Consider the equivalence relation given by
(0, y) ∼ (1, y) (x, 0) ∼ (x, 1)
0 0 0 0
and (x, y) ∼ (x , y ) if and only if (x, y) = (x , y ) if the points are not on the boundary. In
other words, we are trying to glue the edge of the square to its opposite edge. We can give this
the quotient topology: let us recall that this means, under the map:
q : I × I → I × I/ ∼
a set U ⊂ I × I/ ∼ is open if and only if q −1 (U) is open. Let us endow the quotient with a
smooth structure; tentatively we call this space T2q to indicate that we don’t quite know that it
is diffeomorphic to I × I and we will write elements in it as [(x, y)] to indicate the equivalence
class of (x, y).
There are several flavors of charts:
(Interior) if (x, y) 6∈ q(∂(I × I)) this is the easiest chart: just take a small ball around (x, y).
20 3. EXAMPLES OF SMOOTH MANIFOLDS
is a smooth map between open subsets of Euclidean spaces. A smooth map f : M → N is said
to be a diffeomorphism if it is a continuous bijection with a smooth inverse. In this class,
we want to say that two manifolds are the same whenever there is a diffeomorphism between
them. In the homeworks, you will be asked to show that S1 is diffeomorphic to RP1 and S2 is
diffeomorphic to CP1 .
In the next example, we exhibit ways in which we can prove that something is a smooth
map.
Example 1.0.1 (The diagonal). A smooth map that is always defined for any manifold is
the diagonal: as a continuous map of topological spaces we set
∆:M→M×M x 7→ (x, x).
To “officially” prove that ∆ is a smooth map, one needs to pick an atlas on both domain and tar-
get. Of course the atlas on the domain does induce an atlas on the target: if {(Uα , Vα , ϕα )}α∈A
is an atlas then {(Uα × Uβ , Vα × Vβ , ϕα × ϕβ }α,β∈A gives an atlas. However, it turns out that
we do not need to pick an atlas on the target by the next lemma which asserts that smoothness
is a local property.
To use the next lemma, we take a chart (Uα , Vα , ϕα ) containing x and take the chart
(Uα × Uα , Vα × Vα , ϕα ). We are then left to prove that the diagonal map between subsets of
Euclidean spaces:
Uα → Uα × Uα
is smooth.
The proof of the next lemma is completely analogous to the proof of the lemma in topology
that a function is continuous if and only if it is continuous at each point.
21
22 4. MORPHSIMS OF MANIFOLDS, TANGENT SPACES AND DERIVATIVES
Example 1.1.3 (Special linear groups). SLn (R) ⊂ GLn (R) is the subset of invertible ma-
trices with determinant 1. In other words, it is det−1 (1) where det : GLn (R) → R is the
determinant map. Using the submersion theorem, one can show that SLn (R) is smooth of
dimension n2 − 1.
The next example is very important and illustrates the kind of phenomena, in simplified
form, that we want to capture via the general notion of a derivative.
Example 1.1.4. [Orthogonal groups] Let (−, −) : Rn × Rn → R be the standard inner
product on n-space. Explicitly this is given by
(x = (x1 , · · · , xn ), y = (y1 , · · · , yn )) = x1 y1 + x2 y2 + · · · + xn yn .
In matrix multiplication, we can write this as
y1
t
x · y = x1 ··· xn · · · · = (x, y).
yn
The orthogonal group is the subset O(n) ⊂ Mn (R) given by those matrices A such that
(x, y) = (Ax, Ay)
n
for all x, y ∈ R . In other words, O(n) is the set of invertible matrices that preserve the inner
product. One should interpret this as the set of matrices that preserve angles and lengths.
Equivalently, this is the set of all matrices whose inverse is the transpose. The set O(n) itself
admits the structure of a manifold of dimension n(n − 1)/2.
Let’s prove this because it illustrates the submersion theorem quite nicely. Let Sn (R) ⊂
Mn (R) be the subset of symmetric matrices: those such that t A = A. Hopefully it is
clear why we call these symmetric matrices. The first claim is that Sn (R) is a vector space
of dimension (n2 − n)/2 + n = n(n + 1)/2 (hint: the first expression gives a hint for a proof).
Indeed: this follows from the standard properties of the transpose T (A + B) =T A +T B and
n(n+1)
T
(rA) = r(T A). Therefore, it is isomorphic to R 2 as a topological space and hence a
smooth manifold. We have the map
t : GLn (R) → Sn (R) A 7→T AA
and t−1 (In ) = O(n). We note that t is actually a map between Euclidean spaces!
2 n(n+1)
t : Rn → R 2 .
We want to prove that the set t is a submersion at all inverse elements in t−1 (In ) in order to
apply the submersion theorem. Hence fix A ∈ t−1 (In ) and we want to say that
2 n(n+1)
DtA : Rn → R 2
has full rank. The best way to compute that the Jacobian matrix has full rank is to straight
up compute it from first principles. We claim that
DtA (B) = BT A +T AB.
Indeed, we note that the following expression does make sense
T
(A+hB)(A+hB)−T AA T
AA+hT AB+hAT B+(h2 )T BB−T AA
lim ( t(A+hB)=t(A)
h = h = h )
h→0
which is equal to BT A +T AB and is the desired linear transformation solving the requirement
for the derivative as discussed in chapter 1. Now, to apply the submersion theorem, we need to
prove that for each C ∈ Sn (R) there indeed exists a B such that
DtA (B) = C.
Using that AA = In by assumption, we see that 12 CA works!
T
24 4. MORPHSIMS OF MANIFOLDS, TANGENT SPACES AND DERIVATIVES
Remark 2.1.2 (Germs and sheaves). There is a more “global” notion of germs: to each
chart (U, V, ϕ) we can associate the R-algebra
C∞ (U) =: C∞ (V),
and if (U0 , V0 , ϕ0 ) is another chart such that V ⊂ V0 then we have a map of R-algebras:
C∞ (V) → C∞ (V0 )
given by
ϕ0 ϕ−1 f
f : U → R 7→ U0 −→ V0 ⊂ V −−→ U −
→ R;
we write the latter function as f |V0 . In these terms, a function germ around a point x ∈ M is,
informally, an equivalence class of a function f ∈ C∞ (V) for V containing x and we identify f
with f |U0 whenever V0 ⊂ V as long as x ∈ V0 . The language of sheaves and stalks, which we
will not cover in this class, makes all of this precise.
The following is easy to verify:
Lemma 2.1.3. Let M, N, P be manifolds.
(1) Composition of germs: L let f : (M, x) → (N, f (x)) and g : (N, g(x)) → (P, (g ◦ f )(x))
be function germs. Then the composite, defined by
g ◦ f := g ◦ f
is well-defined.
(2) R-vector space structure: for f, g ∈ OM,x , the operations
f + g := f + g rf := λf
defines a R-vector space structure on OM,x .
(3) R-algebra structure: with the notation as above, multiplication
f · g := f g
defines a R-algebra structure on OM,x .
Construction 2.1.4 (Induced map on germs). Let f : M → N be a smooth map of
manifolds and x ∈ M, the induced map on germs is the map
f ∗ : ON,f (x) → OM,x g 7→ g ◦ f .
Concretely, f ∗ takes the equivalence class of (U, g) where g : U → R and f (x) ∈ U to the
f| 0
equivalence class of g ◦ f : U0 −−−
U
→ U → R where U0 is a neighborhood of x. It is easy to check
that this is, in fact, a R-algebra map.
Notice that the above elaboration on f ∗ only depends on open neighborhoods around x
and f (x). Hence, more generally, if we have a germ f : (M, x) → (N, f (x)) then we have the
induced map on function germs
f ∗ : ON,f (x) → OM,x g 7→ g ◦ f ,
given by the same formula.
We will not really use the R-algebra structure seriously, though we will use that this is a
map of R-vector spaces. The next proposition, albeit tautological, is key to the theory.
Lemma 2.1.5 (Local nature of germs). Let M be a manifold of dimension n and (Uα , Vα , ϕα )
be a chart around x such that ϕα (0) = x. Then there is an induced isomorphism of R-vector
spaces
ϕ∗
OM,x −−→
α
ORn ,0 .
26 4. MORPHSIMS OF MANIFOLDS, TANGENT SPACES AND DERIVATIVES
Proof. In fact, Construction 2.1.4 satisfies the following: suppose that we have germs
f : (M, x) → (N, f (x)) and (N, f (x)) → (P, g(f (x))) we have precompositions
f∗
ON,f (x) −→ OM,x
such that: 1) f ∗ is an R-linear map, 2) id∗ = id and 3) (g ◦ f )∗ = f ∗ ◦ g ∗ . Hence, we deduce
that (ϕ−1 ∗ ∗
α ) furnishes an inverse to ϕα .
We will use Lemma 2.1.5 repeatedly to reduce a lot of results about tangent spaces to results
about Rn . We now define the tangent space; we explain why this is a reasonable definition later.
Now, recall that we want a map f : M → N to induce a map Tx M → Tf (x) N of tangent spaces.
As explained in Construction 2.1.4, function germs compose in the other direction:
ON,f (x) → OM,x g : (N, f (x)) → R g ◦ f : (M, x) → R.
So we might guess that Tx M is the dual space of R-linear functionals O∨
M,x := HomR (OM,x , R).
This fixes the problem as we will now have a map
O∨
M,x → ON,f (x) .
∨
However, this is not quite what we want: the space O∨M,x is way too big and is not even finite
dimensional. Remember that we want this to just be a Rn where n = dim(M). The right thing
to do is to only consider “germs up to the first order.” This is captured by the notion of a
derivation.
Definition 2.1.6. A derivation (at x ∈ M) is a R-linear map
X : OM,x → R
satisfying the “Liebniz rule”
X(f g) = X(f )g(x) + f (x)X(g(x)).
The tangent space Tx M is the set {X : OM,x → R : X is a derivation}.
A derivation takes in a germ and spits out a number; hence we should think of as a certain
operation on functions which only depends on its germ. Furthermore, these operations only
really depend on “first order information” of f . This is what the Liebniz rule says but the best
way to see this is via the proof of Lemma 2.1.10.
Example 2.1.7. Let c be the germ of a constant funtion, then we claim that X(c) = 0. By
linearity, we may assume that c = 1, the constant function with value 1. Then:
X(1) = X(1 · 1) = 2X(1)
which means that X(1) = 0.
Example 2.1.8. Consider T0 Rn . Then, for each i = 1, · · · , n we have the derivation
∂
∂xi : ORn ,0 → R f 7→ Di f (0).
These are the most important kinds of derivation and they are linearly independent by evalu-
ating them on the germs of the projections: πi : Rn → R, (x1 , · · · , xn ) 7→ xi .
Example 2.1.9. Let f , g : (M, x) → R be function germs such that f (x) = g(x) = 0. Then
X(f g) = f (x)X(g) + g(x)X(f ) = 0 + 0 = 0. Hence X annihilates the product of functions that
vanish around x. We will use this result later.
The next lemma is a key result which verifies that Tx M is indeed what we want.
Lemma 2.1.10. The tangent space Tx M is a R-linear vector space of dimension dim(M).
More precisely: let (U, V, ϕ) be a chart around x such that ϕ(0) = x and thus induces an isomor-
ϕ∗
phism OM,x −−→ ORn ,0 as in Lemma 2.1.5. Then taking R-linear dual induces an isomorphism
∼
=
T 0 Rn −
→ Tx M
2. THE TANGENT SPACE AND THE DERIVATIVE 27
These linear functionals are derivations by the usual Liebniz rule and are linearly
independent. We prove that they span.
(3) Let us write xi : Rn → R for the i-th projection map and we think about it as the
germ at 0. Let X be derivation, our claim is that
n
X
∂
X= X(xi ) ∂x i
.
i=1
Let’s unpack what this means: this means that for all function germ f , X(f ) is the
sum of terms
∂f
X(xi ) ∂x i
(0) = X(xi )Di f (0).
where xi is the germ of the projection function to the i-variable and Di f (0) is a
number: it is the value of the derivation explained in Example 2.1.8.
(4) To do so, recall that we have Taylor’s theorem with remainder: we can write any
smooth function f locally around 0 (in other words, valid for x ∈ U0 a small enough
neighborhood of 0) as
n
X
∂f
f (x = (x1 , · · · , xn )) = f (0) + ∂xi (0)xi + R2 (x)
i=1
It almost doesn’t matter what R2 (x) is —- the only thing we ever need is to be able
to know that each summand can be written as a product of two functions vanishing
at 0 so as to invoke Example 2.1.9.
28 4. MORPHSIMS OF MANIFOLDS, TANGENT SPACES AND DERIVATIVES
The first equality comes from plugging in the Taylor expansion and the linearity of X,
the second equality comes from Examples 2.1.7 and 2.1.9 and the last equality comes
from linearity.
(4) Let (U, V, ϕ) and (U0 , V0 , ϕ0 ) be charts around x and f (x) respectively such that ϕ(0) =
x, ϕ0 (0) = f (x), then the composite
Dfx
Tx M Tf (x) N
Dϕ0 ,∼
= (Dϕ00 )−1 ,∼
=
T0 Rm T 0 Rn
by the formal properties of taking duals of linear maps. Thus, it holds for Dfx because f∗
preserves derivatives. Now we explicitly want to compute the composition
(Dϕ00 )−1 ◦ Dfx ◦ Dϕ0 = (ϕ0−1
0 ◦ f ◦ ϕ0 )∗ .
where the equality follows from (1)-(3). So let us rewrite
g := ϕ00−1 ◦ f ◦ ϕ0 : (Rm , 0) → (Rn , 0)
∂
as g = (g 1 , · · · , g n ) where each g i : Rm → R. Since we have already shown that ∂xj for
j = 1, · · · m forms a basis for T0 Rm we need only calculate
∂
g∗ ( ∂x j
)
∂
in terms of the bases ∂yi for i = i, · · · , n. In fact, to prove that g∗ is indeed the usual total
∂ ∂g i
derivative, we need to know the coefficient corresponding to ∂yi is exactly the number ∂xj (0);
∂ ∂
in the basis ( ∂x j
)j=1,··· ,m for the source and ( ∂y i
)i=1,···n we will have written out the Jacobian
of g.
So take a function germ h : (Rn , 0) → R. Then, for any fixed j ∈ {1, · · · m} we have
∂ ∂(h◦g)
g∗ ( ∂x j
)(h) = ∂xj
n
∂g j
X
∂h
= ∂yi (0) ∂xj (0),
i=1
where the first equality is by definition and the second equality comes from the multivariable
chain rule. Since this is true for all function germ h, we conclude that the coefficient the term
∂
∂yi is as desired.
Remark 2.1.13 (Tangent space as curves). We remark on another, equivalent way of defin-
ing the tangent space. It is somehow more “visual” and evokes the idea that a tangent vector
should be a point on the manifold and a “vector pointing parallel to it.” Let (−1, 1) := I be
the open interval and we think of it as a smooth manifold. A smooth curve around x ∈ M
(or simply a curve if the context is clear) is a smooth map γ : I → M such that γ(0) = x.
A smooth curve in M can be rather wild, but we are only interested in the first order
information derived from γ. To do so we associate to a curve γ a derivation given by:
Xγ : OM,x → R g 7→ (g ◦ γ)0 (0).
In other words, we look at g ◦ γ : I → R and just take its derivative at 0. By the usual Liebniz
and chain rules, this is a derivation. Set
Γx M := {γ : γ is smooth curve around x}
30 4. MORPHSIMS OF MANIFOLDS, TANGENT SPACES AND DERIVATIVES
f
M N
such that, for each x ∈ X the map
Tx M → Tf (x) N
is the map of Lemma 2.1.12.
Proof. Note that the commutativity of the diagram says that Df (x, v) = (f (x), Dfx (v))
so the map Df is constructed this way by defining it on the underlying set and checking that it
glues together to a map Df : TM → TN. The only thing we really need to explain is why Df
is smooth. The point here is that
∂f
Df (x, v = (v1 , · · · , vn )) = (f (x) = (f 1 (x), · · · , f n (x)), ∂x i
(x)vi ).
which is evidently smooth since f was.
It then follows that the following global version of Lemma 2.1.12 holds:
Corollary 2.2.7. The map
Df : TM → TN
satisfies the following properties:
(1) if f = id, then Df = id;
(2) if g : N → P is another smooth map then
D(g ◦ f )x = Dg ◦ Df
coincides with the total derivative.
32 4. MORPHSIMS OF MANIFOLDS, TANGENT SPACES AND DERIVATIVES
33
APPENDIX A
35
Bibliography
37