Manifolds Work Material 1
Manifolds Work Material 1
CHRISTIAN BÖHNING
The good news is that now we are almost ready to define manifolds already!
Definition 1.3. Let M be a topological space. An atlas of M compatible with a
pseudogroup of transformations Γ is a family of pairs {(Ui , ϕi )}, called charts, such
that
S
(i) Each Ui is an open subset of M and M = i Ui , i.e. the Ui cover M .
(ii) Each ϕi is a homeomorphism of Ui onto an open set of S.
(iii) Whenever Ui ∩ Uj is non-empty, the mapping ϕj ◦ ϕ−1 i of ϕi (Ui ∩ Uj ) onto
ϕj (Ui ∩ Uj ) is an element of Γ.
Sometimes the ϕi are also called local Γ-coordinates, especially when S = Rn , or
Cn . For a point p ∈ Ui one then often writes
ϕi (p) = (x1 (p), . . . , xn (p))
and calls x1 (p), . . . , xn (p) the local coordinates of p, and the function x1 , . . . , xn
simply local coordinates. The atlas itself is sometimes also called a Γ-structure
on M . Basically, we would like to call such an M a Γ-manifold. However, to be
politically correct, we need to choose the atlas maximal in a certain sense:
Definition 1.4. A complete atlas on M compatible with Γ is an atlas of M com-
patible with Γ which is not contained in any other atlas of M compatible with
Γ.
Note: every atlas of M compatible with Γ is contained in a unique complete atlas
of M compatible with Γ. In fact, given an atlas A = {(Ui , ϕi )} of M compatible
with Γ, let à be the family of all pairs (U, ϕ) such that ϕ is a homeomorphism of
an open set U of M onto an open set of S and that
ϕi ◦ ϕ−1 : ϕ(U ∩ Ui ) → ϕi (U ∩ Ui )
is an element of Γ whenever U ∩ Ui is nonempty. Then à is the complete atlas
containing A.
Definition 1.5. A Γ-manifold is a Hausdorff topological space M that is second
countable, together with a fixed complete atlas compatible with Γ (sometimes also
called a Γ-structure on M ). Depending on the particular Γ, people have come up
with a plethora of special names; we keep the notation from Example 1.2.
a) If Γ = ΓC 0 , M is called a topological manifold.
b) If Γ = ΓC r , r = 1, 2, . . . , M is called a differentiable manifold of class C r . If
Γ = ΓC ∞ , M is called a smooth manifold or a differentiable manifold (without
further qualification).
c) If Γ = ΓC r ,0 , M is called an orientable differentiable manifold of class C r , r =
1, 2, . . . , and if Γ = ΓC ∞ ,0 , M is called an orientable differentiable manifold
or orientable smooth manifold.
d) If Γ = ΓC ω , then M is called a real analytic manifold.
e) If Γ = Γhol , M is called a complex manifold.
MA3H5 MANIFOLDS 3
For us the term manifold, without qualification, will always mean smooth manifold.
S 2 there exists only one, on S 6 (A. Adler) there exists none, but in general, there
can exist a continuous supply of complex structures on a given (even-dimensional)
smooth manifold, and these structures themselves form a space with rich geometry.
Benzecri (1959) showed that if a Γ+
af f -structure exists on a compact surface it must
be a torus. More recently, symplectic structures have gained greater importance
placing them on equal footing with complex structures in the framework of mirror
symmetry.
So far nothing really to knock your socks off, but there are tons of other more
interesting manifolds and constructions, which we will come to in a minute; here is
something a little more entertaining to begin with: take K = R or K = C; we will
construct a manifold PnK called (real or complex) projective space over K of (real
or complex) dimension n. As a set
PnK = (K n+1 − {0})/ ∼
where ∼ is the equivalence relation
(p0 , . . . , pn ) ∼ (q0 , . . . , qn ) ⇐⇒ ∃c ∈ K\{0} : c(p0 , . . . , pn ) = (q0 , . . . , qn ).
We give PnK the quotient topology of the usual Euclidean topology on K n+1 − {0}
(meaning a subset in PnK is declared open iff its preimage in K n+1 − {0} is open).
We denote the equivalence class of (p0 , . . . , pn ) in PnK by (p0 : · · · : pn ). We define
an atlas on PnK in the following way: For i = 0, . . . n we let
Ui = {(p0 : · · · : pn ) | pi 6= 0}
and
n p0 pi pn (i) (i) (i)
ϕi : Ui → K , ϕi ((p0 : · · · : pn )) = = (x0 , . . . , xi−1 , xi+1 , . . . , x(i)
c
,..., ,..., n ).
pi pi pi
MA3H5 MANIFOLDS 5
It is easy to see that G is isomorphic to Z and is without fixed points and properly
discontinuous (Exercise: Show this!), and the quotient manifold W/G is called a
Hopf manifold. It can be shown that it is diffeomorphic to S 1 × S 2n−1 (you can
think about this as a challenge).
Example 2.11. Let M = S 3 and view this as the unit sphere in R4 ' C2 (note
that if z1 , z2 are complex coordinates in C2 , this is the subset z1 z̄1 + z2 z̄2 = 1. Fix
coprime integers p, q > 1 and consider the group G generated by the diffeomorphism
of S 3 given by 2πi 2πiq
g(z1 , z2 ) = e p z1 , e p z2 .
Then G is isomorphic to Z/p and without fixed points (and properly discontinuous
since it is a finite group). The quotient manifold in this case is often denoted by
L(p, q) and called a lens space.
2.4. Surgeries. Another very useful method to produce new manifolds from old
ones is to modify a given manifold locally in a certain way we will now describe.
This is called surgery.
Theorem 2.12. Let M be a manifold, and let S ⊂ M be a compact submanifold.
Suppose we have an open neighbourhood W of S in M and another pair of manifolds
S ∗ ⊂ W ∗ with S ∗ compact and W ∗ an open neighbourhood of S ∗ . Suppose we are
given a diffeomorphism
f : W ∗ − S∗ → W − S
of W ∗ − S ∗ onto W − S. Then we can “replace” S by S ∗ and obtain a new manifold
M ∗ = (M − W ) ∪ W ∗ . More precisely, we consider
M ∗ = ((M − S) t W ∗ ) / ∼
where ∼ is the equivalence relation identifying each point z ∗ ∈ W ∗ − S ∗ with z =
f (z ∗ ). This, with the quotient topology, inherits a natural manifold structure from
M (and W ∗ ).
Proof. Exercise for the reader.
Example 2.13. We consider M = P1C × P1C . We let ζ = (ζ0 : ζ1 ) be (homogeneous)
coordinates on the first copy of P1 , and let z = ζζ01 be a local coordinate on the open
subset U ⊂ P1C , U ' C where ζ0 6= 0. Then M contains W = U ×P1 , and W contains
S = {0} × P1 . Let W ∗ = U ∗ × (P1 )∗ be a second copy of W (on which everything
is denoted by the same letters, but with a superscript ∗), and let S ∗ = {0} × (P1 )∗ .
Fix an integer m > 0 and define f : W ∗ − S ∗ → W − S by
ζ1∗
∗ ∗ ∗ ∗
(z, ζ) = f (z , ζ ) := z , ζ0 : ∗ m .
(z )
Then f is a diffeomorphism and we can define Mm ∗ = (M − S) ∪ W ∗ by surgery.
∗ are homeomorphic
This is called a Hirzebruch manifold. One can show that all M2n
∗
to each other, and so are all M2n+1 , but the ones for even and odd m are not
homeomorphic to each other. This requires more techniques from topology than we
have right now, however.
Example 2.14. Let M be a complex manifold with a point S := p ∈ M and a chart
(W, ϕ = (z1 , z2 )) with ϕ : W ' U where U = {(z1 , z2 ) ∈ C2 | ||z1 | < 1, |z2 | < 1} and
ϕ(p) = (0, 0). In particular, ϕ is a smooth chart of course. Define a submanifold
W ∗ of W × P1C as follows:
W ∗ = (z1 , z2 ; (ζ1 : ζ2 )) ∈ W × P1C | z1 ζ2 − z2 ζ1 = 0
10 CHRISTIAN BÖHNING
Then
X ∂xk
0= ξj = ξk , k = 1, . . . , n.
∂xj p
j
Thus we have shown that the set of tangent vectors to curves, i.e., derivations of
the form Xγ,p , form a vector subspace of dimension n of the vector space of all
derivations, where n = dim M .
Definition 3.1. We denote the n-dimensional vector space of all tangent vectors
to M at p by
Tp (M )
and call it the tangent space to M at p.
Remark 3.2. It is known that for a smooth (C ∞ -)manifold M , Tp (M ) coincides with
the space of all derivations of CM,p
∞ (not just a subspace).
(λ)
X ∂x(λ)
j (µ)
ξj = ξ
(µ) k
k ∂xk
on Uλ ∩ Uµ . Although computations with these local expressions can get messy
rather quickly, this description sometimes allows one to obtain useful information
about X(M ).
The set X(M ) is obviously an R-vector space and also a module over the algebra
C (M ) of smooth functions on M , but it has another useful structure: if X, Y ∈
∞
then
X ∂ηj
∂ξj ∂f
[X, Y ]f = ξk − ηk
∂xk ∂xk ∂xj
j,k
∂
so that [X, Y ] is a vector field whose components on U with respect to the ∂xj are
X ∂ηj ∂ξj
ξk − ηk .
∂xk ∂xk
k
The bracket
[·, ·] : X(M ) × X(M ) → X(M )
is R-bilinear and satisfies
[[X, Y ]Z] + [[Y, Z], X] + [[Z, X], Y ] = 0, ∀ X, Y, Z ∈ X(M )
as a direct calculation shows. The last identity is called the Jacobi identity, and
a vector space endowed with a bilinear bracket [·, ·] satisfying the Jacobi identity
MA3H5 MANIFOLDS 13
is called a Lie algebra. You will learn more about these in MA4E0 Lie Groups
and MA453 Lie Algebras. Suffice it to mention here that Lie algebras usually arise
as tangent spaces to the identity of some Lie group, which is a group endowed
with some kind of smooth or similar structure; indeed, X(M ) should be though
tof as the tangent space to the “manifold” Diff(M ), the diffeomorphism group of
M . Unfortunately, it is not a manifold in our sense here, but a “Fréchet manifold”,
locally modelled not on some Rd , but on some special infinite-dimensional topological
vector spaces called Fréchet spaces (sorry! The world out there is vast and cruel).
We will leave it at that here, but for cultural reasons we remark that “Lie” is
pronounced “L-ee”, after the Norwegian mathematician Sophus Lie.
If f, g are smooth functions on M and X, Y vector fields, a computation shows
that
[f X, gY ] = f g[X, Y ] + f (Xg)Y − g(Y f )X
(Exercise!)
Definition 3.4. Let
f : M → M0
be a smooth map of manifolds. The differential at p ∈ M of f is the linear map
f∗ : Tp (M ) → Tf (p) (M 0 )
defined as follows. For each tangent vector Xp ∈ Tp (M ), choose a curve γ(t) in M
such that Xp is the tangent vector to γ(t) at p = γ(t0 ). Then f∗ (Xp ) is the vector
tangent to the curve f (γ(t)) at f (p) = f (γ(t0 )).
It is easy to check that f∗ (Xp ) is independent of the choice of γ(t); in fact, if
(U, x1 , . . . , xn ) is a chart around p, and (V, y1 , . . . , ym ) is a chart around f (p) such
that f (U ) ⊂ V , it is immediate to see that the differential f∗ is the linear map given
by the Jacobian matrix
∂fi
(p)
∂xj i=1,...,m, j=1,...,n
if we use the bases
∂ ∂ ∂ ∂
,..., ; ,...,
∂x1 p ∂xn p ∂y1 f (p) ∂ym f (p)
for Tp (M ) and Tf (p) (M 0 ), respectively. You should check the preceding assertions as
an exercise and also verify that if g is a smooth function defined in a neighbourhood
of f (p), then
(f∗ (Xp ))g = Xp (g ◦ f ).
If we want to specify the point p, we write (f∗ )p instead of f∗ . Another common
notation used for (f∗ )p is dfp (or Dfp ).
The following is a useful consequence of the Inverse Function Theorem that you
encounter in MA259 Multivariable Calculus:
Proposition 3.5. Let f : M → M 0 be a smooth map and p ∈ M as before.
a) If (f∗ )p is injective, there exist local coordinates x1 , . . . , xn in a neighbourhood
U of p and local coordinates y1 , . . . , ym in a neighbourhood of f (p) such that
yi (f (q)) = xi (q) for q ∈ U and i = 1, . . . , n.
This can be paraphrased by saying that f locally around p “looks like (=is
diffeomorphic to) the inclusion of Rn into Rm ”.
In particular, f is a homeomorphism of U onto f (U ).
14 CHRISTIAN BÖHNING
are a solution to the initial system of differential equations with initial conditions
gi (0) = fi (s; x), and thus by the uniqueness of the solutions,
gi (t) = fi (t; Φs (x)).
Thus Φt (Φs (x)) = Φt+s (x); moreover, there exist δ > 0 and > 0 such that putting
U = {x | |xi | < δ ∀ i}, we have Φt (U ) ⊂ U1 for |t| < : indeed, this follows simply
because Φ0 is the identity on U1 . Since therefore
Φ−t (Φt (x)) = Φt (Φ−t (x)) = Φ0 (x) = x
for all x ∈ U and |t| < , all Φt are diffeomorphisms of U onto Φt (U ). Thus Φ is a
local 1-parameter group of local diffeomorphisms defined on I × U inducing X.
If under the hypotheses of Proposition 3.9, there exists a global 1-parameter group
of diffeomorphisms Φ : R × M → M (hence defined for all times t) inducing X, we
say X is complete. The following is a fact that adds something genuinely new to
facts known from the (local) theory of ordinary differential equations.
Proposition 3.10. On a compact manifold M , every vector field X is complete.
Proof. For every point p, let U (p) be a neighbourhood of p and (p) > 0 such that
there exists a local 1-parameter group of local diffeomorphisms Φ defined on I(p) ×
U (p) inducing X. Since M is compact, finitely many of the U (p), U (p1 ), . . . , U (pk )
say, cover M . Putting the minimum of the (pi ), i = 1, . . . , k, we see that Φt (p)
can be defined on I × M , hence on all of R × M . Another way to think of this is
to say that the integral curve γ of X with γ(0) = p exists for all time: after time ,
you can continue it for another time interval , and so on.
π is a submersion, π(U (p)) = V (p) ⊂ I is open (use Proposition 3.5, b)) and since
π∗ X = ∂/∂t, we have that
Φt (p) ∈ Mt1 +t
since π ◦ Φt (p) is an integral curve for the vector field ∂/∂t on R with initial value
for t = 0 equal to t1 . Now since π is proper, K := π −1 ([t1 , t2 ]) ⊂ M is compact. We
investigate the “maximum life-time” of the integral curve γp (t) = Φt (p) starting at
p ∈ Mt1 . More precisely, for any q ∈ M we look at the set of all integral curves γ of
X in a neighbourhood of q with initial condition γ(0) S = q. If Iγ ⊂ R is the domain
of definition of the map γ, then we can set J(q) = γ Iγ , the union being taken over
all such integral curves γ, and obtain an open interval J(q) together with an integral
curve γq : J(q) → M with γq (0) = q that is the maximal integral curve with initial
condition q (in the sense that J(q) is maximal among all such integral curves). Let
J(q) = (ω− , ω+ ), −∞ ≤ ω− < 0, 0 < ω+ ≤ +∞. Then a well-known fact from the
theory of ordinary differential equations, which carries over without change to the
context of manifolds, gives that (t, γq (t)) leaves every compact subset S of R × M
for t → ω− , t → ω+ : indeed, suppose that this was false. Then we can choose a
sequence of times tk → ω+ , k = 1, 2, . . . (same argument for ω− ) with xk = γq (tk ),
(tk , xk ) ∈ S, and passing to a convergent subsequence if necessary (which we can
do since S is compact), we can assume that (tk , xk ) converges to some (t0 , x0 ) ∈ S
(then clearly t0 = ω+ and this must be finite). Now picking a chart around (t0 , x0 ),
we can identify a neighbourhood of (t0 , x0 ) in M with an open subset Ω of Rn .
With our usual deliberate sloppiness of notation, we write (t0 , x0 ) ∈ Ω. By the
Picard-Lindelöf Theorem from MA254 Theory of ODEs it follows that there exist
real numbers r, > 0 such that the initial value problem corresponding to the given
vector field on M has a solution defined for all times t with
h r ri
t ∈ t0 − , t0 +
2 2
provided the initial point is inside
h r ri
G = t0 − , t0 + × B̄ x0 ,
2 2 2
(where B̄ x0 , 2 is the closed ball around x0 with radius /2). But for large k,
(tk , xk ) ∈ G, and is arbitrarily close to (t0 , x0 ). This is a contradiction since then
γq (t) cannot be maximal (i.e. the existence interval J(q) is not maximal). Hence
(t, γq (t)) has to leave every compact subset.
Returning to our geometric situation, the integral curve t 7→ Φt (p) cannot be
contained in K! In other words, it will exist at least on a small time interval
containing [t1 , t2 ]. The same is true for Φt (q) for any q ∈ Mt0 with t0 ∈ [t1 , t2 ].
Therefore, we can define a diffeomorphism
Φt2 −t1 : Mt1 → Mt2
using the flow of X. This is a diffeomorphism since its inverse is Φ−(t2 −t1 ) .
Remark 3.13. As a cultural digression and to emphasise how remarkable Theorem
3.12 is, we point out that it is completely false in the context of complex manifolds:
the complex structure can vary continuously with parameters in a family of complex
manifolds! Riemann called these parameters moduli. Theorem 3.12 on the contrary
shows that differentiable manifolds do not have moduli: the differentiable structure
remains constant in a family {Mt } (over any connected base manifold). Intuitively, a
complex structure on a given differentiable manifold can be thought of geometrically
as a conformal structure, i.e. a rule to measure angles in the manifold. Thus
MA3H5 MANIFOLDS 21
For the time being, we point out that 1-forms are natural objects occurring in
many applications in the sciences as well; for example, the work of a force field on
R3 is a 1-form acting on displacements (which are tangent vectors at points of R3 ).
It is a remarkable fact that the collection of all tangent vectors to a manifold M
itself forms a manifold (of dimension 2n if dim M = n), called the tangent bundle
T (M ), in a natural way, which we now explain. The same is true for covectors: they
lead to the so-called cotangent bundle T (M )∗ . As sets, as we have mentioned,
G G
T (M ) = Tp (M ), T (M )∗ = Tp (M )∗
p∈M p∈M
(we use the square union symbol to emphasise that these unions are disjoint). We
can then define the two projections
π : T (M ) → M, π(Xp ) = p for Xp ∈ Tp (M ),
∗ ∗
π : T (M ) → M, π(ωp ) = p for ωp ∈ Tp (M )∗ .
Notice that for now these are just maps of sets. M comes with a complete atlas
A = {(U, ϕ)} of charts (U, ϕ). Let ϕ = (x1 , . . . , xn ) with the xi local coordinates on
U . Then define
ϕ̃ : π −1 (U ) → R2n and ϕ̃∗ : (π ∗ )−1 (U ) → R2n
by
ϕ̃(v) = (x1 (π(v)), . . . , xn (π(v)), dx1 (v), . . . , dxn (v)) ,
∗ ∗ ∗ ∂ ∂
ϕ̃ (λ) = x1 (π (λ)), . . . , xn (π (λ)), λ ,...,λ
∂x1 ∂xn
for all v ∈ π −1 (U ) and λ ∈ (π ∗ )−1 (U ). Both ϕ̃ and ϕ̃∗ are one-to-one maps onto
open subsets of R2n . One can construct a topology and differentiable structure on
T (M ) (and T (M )∗ ) as follows. We do the case of T (M ) and leave the unproven
assertions as (easy, mechanical, hopefully at worst confusing) exercises:
(1) If (U, ϕ) and (V, ψ) are charts in A, then ψ̃ ◦ ϕ̃−1 is a smooth map.
(2) The collection of
−1
ϕ̃ (W ) : W open in Rn , (U, ϕ) ∈ A
is a basis for a topology on T (M ) which makes T (M ) into a Hausdorff second-
countable topological space.
(3) The set −1
(π (U ), ϕ̃) : (U, ϕ) ∈ A
is a smooth atlas on T (M ), and the complete atlas containing this makes
T (M ) into a differentiable manifold.
We will sometimes write the elements of T (M ) as pairs (p, Xp ) and those of T (M )∗
as (p, ωp ). If f : M → N is a smooth map of manifolds, then it is easy to check that
f∗ : T (M ) → T (N ), f∗ ((p, Xp )) = (f∗ )p (Xp )
is a smooth map of manifolds.
In fact, both T (M ) and T (M )∗ even have a little more structure to them than
just the smooth manifold structures, and this comes out automatically from the
above construction: they are vector bundles (of rank n) over M . We now define this
concept precisely. It is ubiquitous in most parts of geometry and shows up whenever
linearising certain structures locally. Intuitively, a rank r vector bundle over M is a
smoothly varying family of r-dimensional (real) vector spaces Vp , p ∈ M , on M .
MA3H5 MANIFOLDS 23
π πj
Uj
id / Uj
commutes, where πj (p, ζ (j) )) = p. (This morally just means that fj preserves
fibres).
(j) (j) (k) (k)
(b) If (p, ζ1 , . . . , ζr ) ∈ Uj × Rr and (p, ζ1 , . . . , ζr ) ∈ Uk × Rr , then
(k)
fj ◦ fk−1 (p, ζ1 , . . . , ζr(k) ) = (p, fjk (p) · (ζ (k) )t )
where (k)
ζ1
(k) t ..
(ζ ) = .
(k)
ζr
and fjk : Uj ∩ Uk → GL(r, R) are smooth maps (here we view GL(r, R) as an
2
open submanifold of Rr ).
We call Ep = π −1 (p) the fibre of E over p. By (a) and (b) we can endow Ep with
a well-defined structure of R-vector space of dimension r for which the fj become
fibrewise linear isomorphisms.
A morphism of vector bundles, sometimes also called a (smooth) map of vector
bundles, E and E 0 of ranks r, r0 over M is a smooth map of manifolds ϕ : E → E 0
that maps the fibre Ep into the fibre Ep0 and such that ϕ |Ep is a linear map of some
constant rank s that is independent of the point p ∈ M . One then defines the notion
of isomorphism in the usual way as a morphism with a two-sided inverse.
A (smooth) section in a vector bundle E over an open subset U ⊂ M is a smooth
map σ : U → E with π ◦ σ = idU .
There are variants of the above notion: for example, to get a smooth, complex
vector bundle of rank r, one would replace Rr by Cr in the above definition etc.
Note that (smooth) sections in T (M ) and T (M )∗ over an open subset U are
nothing but (smooth) vector fields on U and (smooth) 1-forms on U as defined
earlier.
Vector bundles on a manifold M can be (and are often) (re-)constructed starting
from an open cover {Uj }j∈J as in Definition 3.15 and smooth functions
fjk : Uj ∩ Uk → GL(r, R)
satisfying
(∗) fij ◦ fjk = fik on Ui ∩ Uj ∩ Uk for all i, j, k.
The condition (∗) is called the cocycle condition and an open cover {Uj }j∈J together
with a system of functions satisfying the cocyle condition is called a GL(r, R)-cocycle
24 CHRISTIAN BÖHNING
spaces by modules over a commutative ring). But this is not a course in algebra,
and we do not think we should strive for maximum generality here.
First, given two vector spaces U , V we define their tensor product U ⊗ V (some-
times also denoted by U ⊗K V if we want to recall the ground field) as follows. Let
F (U, V ) be the vector space which has the set U × V as a basis, i.e., the free vector
space (over K of course as always) generated by the pairs (u, v) where u ∈ U and
v ∈ V . Let R be the vector subspace of F (U, V ) spanned by all elements of the form
(u + u0 , v) − (u, v) − (u0 , v), (u, v + v 0 ) − (u, v) − (u, v 0 ),
(ru, v) − r(u, v), (u, rv) − r(u, v)
where u, u0 ∈ U , v, v 0 ∈ V , r ∈ K.
Definition 4.1. The quotient vector space
U ⊗ V := F (U, V )/R
is called the tensor product of U and V . The image of (u, v) ∈ F (U, V ) under the
projection F (U, V ) → U ⊗ V will be denoted by u ⊗ v. We define the canonical
bilinear mapping
β: U × V → U ⊗ V
by β(u, v) = u ⊗ v. Being very precise, one should refer to the pair (U ⊗ V, β) as the
tensor product of U and V , but usually people just use the term for U ⊗ V with β
tacitly understood.
Sometimes one does not need to know the construction of U ⊗ V when working
with it, but only has to use the following property it enjoys in proofs.
Proposition 4.2. Let W be a vector space with a bilinear mapping ψ : U × V → W .
We say that (W, ψ) has the universal factorisation property for U × V if for every
vector space S and every bilinear mapping f : U ×V → S there exists a unique linear
mapping g : W → S such that f = g ◦ ψ.
Then the couple (U ⊗ V, β) has the universal factorisation property for U × V .
If a couple (W, ψ) has the universal factorisation property for U × V , then (U ⊗
V, β) and (W, ψ) are canonically isomorphic in the sense that there exists a unique
isomorphism σ : U ⊗ V → W such that ψ = σ ◦ β.
Proof. Suppose we are given any bilinear mapping f : U × V → S. Since U × V is
a basis of F (U, V ) we can extend f to a unique linear mapping f 0 : F (U, V ) → S.
Now f 0 vanishes on R since f is bilinear so induces a linear mapping g : U ⊗ V → S
on the quotient. Clearly, f = g ◦ β by construction. The uniqueness of such a map
g follows from the fact that β(U × V ) spans U ⊗ V , so we have no other choice in
defining g.
Now if (W, ψ) is a couple having the universal factorisation property for U × V ,
then by the universal factorisation property of (U ⊗ V, β) (resp. of (W, ψ)), there
exists a unique linear mapping σ : U ⊗ V → W (resp. τ : W → U ⊗ V ) such that
ψ = σ ◦ β (resp. β = τ ◦ ψ). Hence
β = τ ◦ σ ◦ β, ψ = σ ◦ τ ◦ ψ.
Using the uniqueness of the g in the universal factorisation property, we conclude
that τ ◦ σ and σ ◦ τ are the identity on U × V and W respectively.
This universal property of the tensor product can be used to prove a great many
formal properties of the tensor product in a way that is almost mechanical once one
26 CHRISTIAN BÖHNING
gets practice with it. All these proofs are boring. So we give one, and you can easily
work out the rest for some practice with this.
Proposition 4.3. The tensor product has the following properties.
a) There is a unique isomorphism of U ⊗ V onto V ⊗ U sending u ⊗ v to v ⊗ u
for all u ∈ U , v ∈ V
b) There is a unique isomorphism of K ⊗ U with U sending r ⊗ u to ru for all
r ∈ K and u ∈ U ; similarly for U ⊗ K and U .
c) There is a unique isomorphism of (U ⊗ V ) ⊗ W onto U ⊗ (V ⊗ W ) sending
(u ⊗ v) ⊗ w to u ⊗ (v ⊗ w) for all u ∈ U , v ∈ V , w ∈ W .
d) Given linear mappings
fi : Ui → Vi , i = 1, 2,
there exists a unique linear mapping f : U1 ⊗ U2 → V1 ⊗ V2 such that
f (u1 ⊗ u2 ) = f1 (u1 ) ⊗ f2 (u2 )
for all u1 ∈ U1 , u2 ∈ U2 .
e) There is a unique isomorphism from (U1 ⊕ U1 ) ⊗ V onto (U1 ⊗ V ) ⊕ (U1 ⊗ V )
sending (u1 , u2 ) ⊗ v to (u1 ⊗ v, u2 ⊗ v) for all u1 ∈ U1 , u2 ∈ U2 , v ∈ V .
f ) If u1 , . . . , um is a basis for U and v1 , . . . , vn is a basis for V , then ui ⊗ vj ,
i = 1, . . . , m, j = 1, . . . , n, is a basis for U ⊗ V . In particular, dim U ⊗ V =
dim U dim V .
g) Let U ∗ be the dual vector space to U . Then there is a unique isomorphism g
from U ⊗ V onto Hom(U ∗ , V ) such that
(g(u ⊗ v))(u∗ ) = hu, u∗ iv for all u ∈ U , v ∈ V , u∗ ∈ U ∗ .
h) There is a unique isomorphism h of U ∗ ⊗ V ∗ onto (U ⊗ V )∗ such that
(h(u∗ ⊗ v ∗ ))(u ⊗ v) = hu, u∗ ihv, v ∗ i
for all u ∈ U , u∗ ∈ U ∗ , v ∈ V , v ∗ ∈ V ∗ .
Proof. We prove a) and g) just to illustrate the method, and leave the rest as easy
exercises.
For a) let f : U × V → V ⊗ U be the bilinear mapping with f (u, v) = v ⊗ u. By
the universal property of the tensor product, there is a unique linear mapping
g: U ⊗ V → V ⊗ U
such that g(u ⊗ v) = v ⊗ u. Similarly, there is a unique linear mapping g 0 : V ⊗ U →
U ⊗V with g 0 (v⊗u) = u⊗v. Clearly, g 0 ◦g and g◦g 0 are the identity transformations.
We now prove g) (using f)). Consider the bilinear mapping f : U ×V → Hom(U ∗ , V )
given by
(f (u, v))(u∗ ) = hu, u∗ iv
and apply the universal property of the tensor product in Proposition 4.2. Thus
there exists a unique linear mapping g : U ⊗ V → Hom(U ∗ , V ) such that (g(u ⊗
v))(u∗ ) = hu, u∗ iv. To prove that g is an isomorphism, let u1 , . . . , um be a basis for
U , u∗1 , . . . , u∗m ∈ U ∗ the dual basis, and v1 , . . . , vn a basis for V . We show that
{g(ui ⊗ vj ) : i = 1, . . . , m, j = 1, . . . , n}
is a linearly independent set of vectors. Indeed
X
aij g(ui ⊗ vj ) = 0, aij ∈ K
i,j
MA3H5 MANIFOLDS 27
gives
X X
0= aij g(ui ⊗ vj )(u∗k ) = akj vj
i,j j
hence all the aij vanish. Since dim U ⊗ V = dim Hom(U ∗ , V ) by f), g is an isomor-
phism.
We now turn to the properties of the exterior algebra we will need later. First of
all it is clear that for any v, w ∈ V we have
v ∧ v = 0, v ∧ w = −w ∧ v,
the first because v⊗v maps to zero under the quotient map T • (V ) → Λ• (V ), and the
second is implied by (v + w) ∧ (v + w) = 0. We say the wedge-product is alternating
or anti-symmetric. More generally, this implies that if ω ∈ Λr (V ) and ϕ ∈ Λs (V ),
then
ω ∧ ϕ = (−1)rs ϕ ∧ ω.
Proposition 4.5. The exterior powers and exterior algebra have the following prop-
erties.
a) If F : V ×· · ·×V → W (r copies of V ) is a multilinear alternating mapping of
vector spaces (which means F (v1 , . . . , vr ) is linear in each argument separately
and zero if two of the vi are equal), then there is a unique linear map
F̄ : Λr (V ) → W
with F̄ (v1 ∧ · · · ∧ vr ) = F (v1 , . . . , vr ).
b) If ϕ : V → W is a linear mapping, there is a unique linear mapping
Λr (ϕ) : Λr (V ) → Λr (W )
with the property
Λr (ϕ)(v1 ∧ · · · ∧ vr ) = ϕ(v1 ) ∧ · · · ∧ ϕ(vr ).
c) If e1 , . . . , en is a basis of V , then
{ei1 ∧ · · · ∧ eir : 1 ≤ i1 < · · · < ir ≤ n}
is a basis of Λr (V ). Consequently,
r n
dim Λ (V ) =
r
and Λi (V ) = 0 for i > n. Moreover, note that dim Λn (V ) = 1.
d) If f : V → V is an endomorphism, dim V = n, then the induced map
Λn (f ) : Λn (V ) → Λn (V )
is multiplication by det(f ).
e) For an n-dimensional vector space V we have a natural non-degenerate bilin-
ear pairing
Λr (V ∗ ) × Λr (V ) → K
mapping
(v1∗ ∧ · · · ∧ vr∗ , w1 ∧ · · · ∧ wr ) 7→ det(vi∗ (wj ))1≤i,j≤r
which induces an isomorphism
Λr (V ∗ ) ' (Λr (V ))∗ .
Proof. For a) notice that repeated application of the universal property of the tensor
product furnishes us with a linear map
F̃ : V ⊗r → W
with F̃ (v1 ⊗ · · · ⊗ vr ) = F (v1 , . . . , vr ); this factors over
Λr (V ) = V ⊗r / V ⊗r ∩ I
MA3H5 MANIFOLDS 29
since the ideal I is generated by elements v ⊗ v that get mapped to zero since F is
alternating.
To prove b) notice that inductive application of Proposition 4.3, d), gives an
induced mapping
⊗r ϕ : V ⊗r → W ⊗r
and this maps V ⊗r ∩ I into the corresponding graded piece of the ideal we divide
out by to get Λr W , so descends to give Λr (ϕ) as desired.
For c) we first show that Λn (V ) ' K via the map induced by the determinant.
Indeed, since the elements v1 ∧ · · · ∧ vr generate Λr (V ), it is clear that, if e1 , . . . , en
is a basis of V , then
{ei1 ∧ · · · ∧ eir : 1 ≤ i1 < · · · < ir ≤ n}
is at least a generating set for Λr (V ). In particular, Λn (V ) is at most one-dimensional,
and exactly one-dimensional, generated by e1 ∧ · · · ∧ en , if we can show it is nonzero.
But by a), the determinant gives a map det : Λn (V ) → K sending e1 ∧ · · · ∧ en to 1.
Now suppose there was a linear dependence relation between the ei1 ∧ · · · ∧ eir :
X
aI e I = 0
I
where we use multi-index notation I = (i1 , . . . , ir ), 1 ≤ i1 < · · · < ir ≤ n, aI ∈
K, eI = ei1 ∧ · · · ∧ eir . For a certain multi-index J = (j1 , . . . , jr ), let J¯ be the
complimentary indices to J in {1, . . . , n}, increasingly ordered. Then
X
( aI eI ) ∧ eJ¯ = ±aJ e1 ∧ · · · ∧ en = 0.
I
Hence all coefficients aJ are zero, proving c).
The endomorphism f : V → V gives a commutative diagram
Λn (f )
Λn (V ) / Λn (V )
det det
mult(c)
K /K
where the lower horizontal arrow is multiplication by some constant c. We want to
show that c = det(f ) and for this it suffices to consider what happens to det(e1 ∧
· · · ∧ en ) = 1: this gets mapped to the determinant of the matrix with columns
(f (e1 ), . . . , f (en )), which is det(f ). This proves d).
For e) first note that
det((vi∗ (wj ))1≤i,j≤r
is alternating in both the v1∗ , . . . , vn∗ and the w1 , . . . , wn , and multilinear in these
sets of variables; hence by an application of a), we get a well-defined map
β : Λr (V ∗ ) × Λr (V ) → K
of the type in e). All that remains to prove is that this pairing is nondegenerate,
i.e. that for any nonzero ω ∈ Λr (V ) there is a ψ ∈ Λr (V ∗ ) with β(ψ, ω) 6= 0, and
vice versa, for any nonzero ψ 0 ∈ Λr (V ∗ ) there is an ω 0 ∈ Λr (V ) with β(ψ 0 , ω 0 ) 6= 0.
We prove the first assertion since the second is then proven completely analogously.
If e1 , . . . , en is a basis of V , write in multi-index notation
X
ω= aI eI .
I
30 CHRISTIAN BÖHNING
Since ω 6= 0, there is an aJ 6= 0. Then let ψ = e∗J = e∗j1 ∧ · · · ∧ e∗jr where e∗1 , . . . , e∗n
is the dual basis to e1 , . . . , en . We have β(ψ, ω) = aJ 6= 0 then.
Let us work out what f ∗ does in local coordinates. Given a smooth map f : M →
N of manifolds, dim M = k, dim N = l, p ∈, we can choose a chart (V, y1 , . . . , yl )
around f (p) and a chart (U, x1 , . . . , xk ) around p with f (U ) ⊂ V ; in terms of these
coordinates we can write f = (f1 , . . . , fl ) with the fi smooth functions of x1 , . . . , xk
on U . Since (f∗ )q , for q ∈ U , is represented in the local coordinates by the Jacobian
matrix
∂fi
(q)
∂xj 1≤i≤l,1≤j≤k
and ((f∗ )q )∗ is given by the transpose matrix, we get
k
∗
X ∂fi
f dyi = dxj = dfi .
∂xj
j=1
Suppose now that M is an orientable manifold and that we have moreover chosen
an orientation. We can then define
Z
ω
M
for any n-form ω with compact support on M , in the following way: choose a
covering {Uα } satisfying the condition in Proposition 5.3 (transition functions have
positive Jacobian on intersections of charts) and coordinates yiα on Uα such that
dy1α ∧ · · · ∧ dynα is equivalent to the given orientation form. Then on Uα we can write
ω |Uα = f (y1α , . . . , ynα )dy1α ∧ · · · ∧ dynα .
Choosing a partition of unity {θi } subordinate to {Uα }, we have
θi ω |Uα(i) = gi (y1α , . . . , ynα )dy1α ∧ · · · ∧ dynα
MA3H5 MANIFOLDS 33
with gi a function which is smooth and has compact support in (an open subset of)
the Rn with coordinates x1 , . . . , xn . Then define
Z XZ XZ
ω := θi ω := gi (x1 , . . . , xn )dx1 dx2 . . . dxn .
M i M i Rn
The local finiteness of the family {θi }, property b) of Theorem 3.11, implies that
only finitely many θi are not identically zero on the compact support of ω, so the
above sum is finite. We leave it to you as an exercise to show this definition is
independent of the partition of unity.
We still owe you to exhibit some examples of orientable manifolds and orienta-
tions. This is rather easy. Clearly Rn is orientable; if x1 , . . . , xn are the standard
global coordinates, then the two orientations correspond to ±dx1 ∧ · · · ∧ dxn .
Suppose that M ⊂ Rn is a subset given by f (x) = c where f : Rn → Rm has
surjective derivative at all points of M , hence is a submersion in a neighbourhood
of M . Such an M is a submanifold either by Proposition 2.4 or Proposition 3.5, b).
If f = (f1 , . . . , fm ) the tangent space Ta (M ) can be identified with the subspace of
Ta (Rn ) of tangent vectors annihilated by the 1-forms df1 , . . . , dfm . Dually, Ta (M )∗
is the quotient of Ta (Rn )∗ by the subspace U spanned by df1 , . . . , dfm . Thus
Λm U ⊗ Λn−m Ta (M )∗ ' Λm U ⊗ Λn−m (Ta (Rn )∗ /U ) ' Λn (Ta (Rn )∗ )
(prove this as an exercise! It’s just linear algebra). Therefore, since df1 ∧ · · · ∧ dfm
gives a non-vanishing section in the Λm U on M , and dx1 ∧ · · · ∧ dxn gives a non-
vanishing section in T (Rn )∗ , the isomorphism furnishes us with a non-zero section
in Λn−m T (M )∗ .
In particular, it follows that all spheres S n are orientable, but obviously we get
tons of other examples as well. However, there are also compact manifolds that
are not orientable (we will give an example in a second). This gives an interesting
conclusion: it can be proved that any compact manifold M of dimension d can be
embedded into some RN for N sufficiently large (we will not do this in this course,
though); thus it follows that not every such M can be defined by N − d global
functions on RN with independent derivatives on M because such an M would
necessarily have to be orientable.
Here is a non-orientable (compact) example: consider the real projective space
PnR with the smooth quotient map
p : S n → PnR
mapping a unit vector in Rn+1 to the one-dimensional subspace it spans (this shows
that PnR is compact since S n is). We have the diffeomorphism
σ : Sn → Sn, σ(x) = −x.
Moreover, if x1 , . . . , xn+1 are global coordinates on Rn+1 and x1 6= 0 at a point of
S n , we can use x = (x2 , . . . , xn+1 ) as local coordinates on S n at that point and the
ratios
(x2 /x1 , . . . , xn+1 /x1 )
as local coordinates around the image point in PnR . With || · || the Euclidean norm
on Rn+1 one thus has in these coordinates
1
p(x) = p x
1 − ||x||2
34 CHRISTIAN BÖHNING
then
dγi
γ ∗ dxi = dγi = ,
dt
whence
Z 3 Z
X dγi
ω= fi (γ(t)) dt.
C I dt
i=1
If we define F to be the vector field (f1 , f2 , f3 ) you will recognise this as the line
integral of F over C (sometimes denoted by
I
F~ dγ
Here dA is usually called the area form of S (if you don’t recall from vector calculus,
try to figure out why!), and the integral on the right hand-side the flux of the vector
field F~ through the surface S.
We have already seen that for a 0-form on a manifold M , f ∈ A 0 (M ), which is
nothing but a smooth function on M , we can define a 1-form df , the differential of
f . This operation d can be extended in a canonical way to all r-forms and is then
called exterior differentiation. We will see that it provides a key link between the
geometry of a manifold and analysis on the manifold. The operation d is defined in
terms of the manifold structure alone without any additional choices.
Proposition 5.5. Let M be a smooth manifold, A (M ) = r A (M ). There is
r
L
a natural R-linear map d : A (M ) → A (M ), called exterior differentiation or the
exterior derivative, that has the following properties:
(i) d(A r (M )) ⊂ A r+1 (M );
(ii) For every function f ∈ A 0 (M ), df is the differential defined earlier.
(iii) For every ω ∈ A r (M ) and π ∈ A s (M ), we have
d(ω ∧ π) = dω ∧ π + (−1)r ω ∧ dπ.
(iv) d2 = d ◦ d = 0.
Moreover, if x1 , . . . , xn are local coordinates on some open subset U of M and we
write on U X
ω= fi1 ...ir dxi1 ∧ · · · ∧ dxir
i1 <···<ir
then X
dω = dfi1 ...ir ∧ dxi1 ∧ · · · ∧ dxir .
i1 <···<ir
Proof. We will define a map d having all the properties claimed by breaking up ω
as a sum of terms with support in some local coordinate system (using a partition
of unity), then define a d operator locally using a coordinate system, then showing
that the definition is independent of the choice of coordinate system. So we write
X
ω= fi1 ...ir dxi1 ∧ · · · ∧ dxir
i1 <···<ir
and define X
dω = dfi1 ...ir ∧ dxi1 ∧ · · · ∧ dxir .
i1 <···<ir
When r = 0, this is just the differential, so (ii) will be satisfied, and (i) is by
construction. Now
X ∂fi ...i
1 r
dω = dxj ∧ dxi1 ∧ · · · ∧ dxir .
∂xj
j, i1 <···<ir
whence
X ∂ 2 fi1 ...ir
d2 ω = dxk ∧ dxj ∧ dxi1 ∧ · · · ∧ dxir .
∂xk ∂xj
j,k, i1 <···<ir
Now
∂ 2 fi1 ...ir
∂xk ∂xj
is symmetric in k, j, but gets multiplied by dxk ∧ dxj , which is skew-symmetric, so
d2 ω = 0, in accordance with (iv).
MA3H5 MANIFOLDS 37
We now show that then d = d0 (on the intersection), using the properties (i)-(iv).
Now (ii) and (iii) give
X
dω = d( f˜i1 ...ir dyi1 ∧ · · · ∧ dyir )
i1 <···<ir
X X
= df˜i1 ...ir ∧ dyi1 ∧ · · · ∧ dyir + f˜i1 ...ir d(dyi1 ∧ · · · ∧ dyir )
i1 <···<ir i1 <···<ir
(you should recall the definition of the differential of a function dψp : h(dψ)p , Xp i =
Xp (ψ)). Thus
f ∗ (dϕ) = d(ϕ ◦ f ) = d(f ∗ ϕ)
38 CHRISTIAN BÖHNING
so our claim holds for 0-forms (functions) on N . Now in general in terms of local
coordinates y1 , . . . , yr on N
X
ω= gi1 ...ir dyi1 ∧ · · · ∧ dyir
i1 <···<ir
whence X
f ∗ω = f ∗ (gi1 ...ir )f ∗ dyi1 ∧ · · · ∧ f ∗ dyir
i1 <···<ir
and therefore
X
d(f ∗ ω) = d(f ∗ (gi1 ...ir )) ∧ f ∗ dyi1 ∧ · · · ∧ f ∗ dyir
i1 <···<ir
X
= f ∗ d(gi1 ...ir ) ∧ f ∗ dyi1 ∧ · · · ∧ f ∗ dyir
i1 <···<ir
∗
= f (dω).
Let us work out what d does on R3 ;
we will re-encounter some old friends (and if
not friends, at least acquaintances) in this way.
Example 5.7. If f is a function on R3 (with coordinates x1 , x2 , x3 ), then
df = g1 dx1 + g2 dx2 + g3 dx3
where
∂f ∂f ∂f
(g1 , g2 , g3 ) = , , = grad (f )
∂x1 ∂x2 ∂x3
the gradient vector field of f .
If
ω = f1 dx1 + f2 dx2 + f3 dx3
is a 1-form, then
dω = df1 ∧ dx1 + df2 ∧ dx2 + df3 ∧ dx3
= g1 dx2 ∧ dx3 + g2 dx3 ∧ dx1 + g3 dx1 ∧ dx2
where
∂f3 ∂f2 ∂f1 ∂f3 ∂f2 ∂f1
g1 =− , g2 = − , g3 = − .
∂x2 ∂x3 ∂x3 ∂x1 ∂x1 ∂x2
Thus if F and G are the vector fields (f1 , f2 , f3 ) and (g1 , g2 , g3 ), then G = curl F .
For a 2-form
ω = f1 dx2 ∧ dx3 + f2 dx3 ∧ dx1 + f3 dx1 ∧ dx2
we get
∂f1 ∂f2 ∂f3
dω = + + dx1 ∧ dx2 ∧ dx3
∂x1 ∂x2 ∂x3
= (div (F ))dx1 ∧ dx2 ∧ dx3 .
Of course d of any 3-form is zero, and all r-forms, r > 3 are zero on R3 .
From d2 = 0 we thus get
curl (grad f ) = 0, div(curl F ) = 0
which is well-known to you (but d2 = 0 is certainly easier to remember than those
two formulas!)
MA3H5 MANIFOLDS 39
It turns out that the connection between what you already know from vector
calculus and differential forms can be carried much further: the classical theorems
of Stokes and Green and Gauss in vector calculus are special cases of a single result
in the theory of differential forms, which is also called “Stokes’ theorem” (in the
present form it was formulated in 1945 by Élie Cartan following earlier work by
Volterra, Goursat and Poincaré on generalisations of the classical integral theorems
of vector calculus. Apparently this is an example of Arnold’s principle (named after
the famous Russian mathematician V.I. Arnold): If a famous result or notion bears
a personal name, then this name is not the name of the discoverer).
We first prove a simple version of the result and the full version afterwards.
Theorem 5.8. Let M be an oriented n-dimensional manifold (i.e., M is orientable
and we have chosen one of the two possible orientations). If ω ∈ A n−1 (M ) is a
differential form with compact support, then
Z
dω = 0.
M
Proof. Choose a coordinate covering of M (with positive Jacobian determinant of
the transition functions on intersections) and a partition of unity {θi } subordinate
to this covering to write X
ω= θi ω.
i
Then in a chart we can write
θi ω = f1 dx2 ∧ dx3 ∧ · · · ∧ dxn − f2 dx1 ∧ dx3 ∧ · · · ∧ dxn + . . .
whence
∂f1 ∂fn
d(θi ω) = + ··· + dx1 ∧ · · · ∧ dxn .
∂x1 ∂xn
From the definition of the integral over M , we simply need to sum all contributions
Z
∂f1 ∂fn
+ ··· + dx1 dx2 . . . dxn .
Rn ∂x1 ∂xn
Now consider Z
∂f1
dx1 dx2 . . . dxn .
Rn ∂x 1
Using Fubini’s theorem to evaluate this as a repeated integral we get that this equals
Z Z Z
∂f1
... dx1 dx2 dx3 . . . dxn .
R R R ∂x1
But since f1 has compact support, it vanishes for |x1 | ≥ N , N large, so
Z
∂f1
dx1 = [f1 ]N
−N = 0.
R ∂x 1
Continuing in the same way, we see that the entire integral is thus zero.
As you will recall from vector calculus, Green’s theorem relates an area integral
(in the plane) to a line integral (and so does the classical Stokes’ theorem for surfaces
in 3-space), and Gauss’s theorem relates a surface integral to a volume integral. To
extend such theorems to the manifold context, we thus have to enlarge the class of
objects under study to include spaces like a closed ball in Rn+1 (with boundary the
sphere S n ), and the cylinder S 1 × [0, 1] in R3 which has boundary two copies of the
circle. We will call these manifolds with boundary and we need to integrate over
them.
40 CHRISTIAN BÖHNING
y1 , . . . , yn overlap
yi (x1 , . . . , xn ), yn (x1 , . . . , xn−1 , 0) = 0,
so the Jacobian matrix at a point of the boundary has the form
∂y1 /∂x1 ∂y1 /∂x2 . . . ∂y1 /∂xn
.. .. .. ..
. . . .
.. .. .. ..
.
. . . .
.. .. .. ..
. . . .
0 0 . . . ∂yn /∂xn
Now ϕβ ϕ−1α maps xn > 0 to yn > 0, thus if xn = 0, then yn = 0, and if xn > 0,
yn > 0. Therefore
∂yn
|x =0 > 0
∂xn n
and since the determinant of the preceding matrix is positive, the determinant of the
upper left (n − 1) × (n − 1)-matrix is positive on xn = 0. Thus the (Uα |∂M , ϕα |∂M )
give an atlas for ∂M whose transition functions have everywhere positive Jacobian,
so ∂M is orientable.
Once we have chosen a specific orientation for M , we will always choose a certain
orientation of ∂M (one of the two possible ones) that is compatible with the orien-
tation of M in a way we now make precise. For surfaces in R3 , for example, one can
think of the convention we are about to make as the choice of an outward or inward
normal.
Definition 5.11. Suppose M is an orientable manifold with boundary for which
we have chosen a specific orientation, represented by a nowhere vanishing n-form ω.
If in local coordinates around a boundary point
ω ' dx1 ∧ · · · ∧ dxn
where = ±1 and with xn ≥ 0 defining M locally, then we define the induced
orientation ω 0 of ∂M locally by
(−1)n dx1 ∧ · · · ∧ dxn−1 .
Example 5.12. It is worthwhile to think about what this means, especially the
funny sign (−1)n ; for example, if we take the closed interval M = [0, 1] ⊂ R with
coordinate x, we see that ω = dx is a nowhere vanishing 1-form on [0, 1], giving it an
orientation. At 0, the local coordinate x1 = x is satisfied that M is locally given by
x1 ≥ 0 and ω ' dx1 , so the point 0 gets the induced orientation −1 (since n = 1).
Note that a choice of orientation for a point is simply the choice of +1 or −1 in R.
At the point 1, instead we can use as local coordinate y1 = 1 − x, whence locally
around 1, M is given by y1 ≥ 0, and now ω ' −dy1 whence the point 0 gets
the induced orientation +1. Note that this already indicates that our choice of
orientation of the boundary is well-suited for generalising the fundamental theorem
of differential and integral calculus: indeed,
Z 1
f 0 (x)dx = f (1) − f (0)
0
and this can be interpreted as saying that integrating the differential of a function
over M is the same as “integrating the function over the boundary”, which in this
case is just evaluating the function at the points of the boundary ∂M , with the
42 CHRISTIAN BÖHNING
boundary orientation taken into account: f (1) enters with a plus-sign, and f (0)
with a minus-sign.
More generally, if we consider H n ⊂ Rn with the standard basis e1 , . . . , en in Rn
and dual basis dx1 , . . . , dxn , we can give H n an orientation by taking the nowhere
vanishing n-form ω = dx1 ∧ · · · ∧ dxn . We call an ordered basis (v1 , . . . , vn ) in Rn
positively oriented (with respect to ω) if
ω(v1 ∧ · · · ∧ vn ) > 0.
The ω0 equivalent to ω are precisely those having positive values on a positively ori-
ented basis, and so the choice of a positively oriented basis in Rn is equivalent to the
choice of an orientation. The induced orientation on ∂H n = Rn−1 = he1 , . . . , en−1 i
in this case is given by ω∂H n = (−1)n dx1 ∧ · · · ∧ dxn−1 : this means that a basis
(w1 , . . . , wn−1 ) of ∂H n = Rn−1 is positively oriented with respect to ω∂H n if and
only if
(n, w1 , . . . , wn−1 )
is a positively oriented basis with respect to ω in Rn , where n is the outward normal
n = −en to H n at 0 ∈ Rn .
We can now prove the full version of Stokes’ theorem; in part its power
R lies in the
fact that it establishes a connection between the analytical operations and d and
the geometric operation ∂ of forming the (oriented) boundary of a manifold with
boundary.
Theorem 5.13 (Stokes’ Theorem, which, as you will recall, is not entirely due to
Stokes...). Let M be an n-dimensional oriented manifold with boundary ∂M carrying
the induced orientation. Let ω ∈ A n−1 (M ) be an (n − 1)-form of compact support.
Then Z Z
dω = ω |∂M .
M ∂M
R
Proof. We need to start with a word about how the integral M dω over the ori-
ented manifold M with boundary is defined; we completely mimic the procedure for
manifolds without boundary: if ϕ is the orientation form on M , as in the proof of
Proposition 5.3, we can write
ϕ = ϕ(x1 , . . . , xn )dx1 ∧ · · · ∧ dxn
in local charts, and provided ϕ is positive in every chart get a cover by charts of
this type such that the Jacobian of the transition functions is positive on overlaps.
We can then give a well-posed definition of the integral just as before: choose a
partition {θi } subordinate to such a cover and repeat the construction described
after the proof of Proposition 5.3. However, if you have been super vigilant, there is
one small catch here: to make ϕ positive in each chart, we may have to replace x1
by the local coordinate −x1 , and if n = 1, we cannot do that at a boundary point
because it would mess up the requirement that M be locally defined by xn ≥ 0 at a
boundary point! Thus if the dimension of M is 1 (and only then) are we seemingly
in trouble and cannot proceed in complete analogy with what we did before. It is
easy to sort out this small inconvenience and we leave it to the student to do this
as a nice and easy exercise to test his understanding of the material (maybe the
easiest fix is to allow local models for manifolds with boundary M that are either
open sets in the upper half space H n = H+ n = {x ∈ Rn | x ≥ 0} or the lower
n
n n
half-space H− = {x ∈ R | xn ≤ 0}; then one can always cover M by such charts
with positive Jacobian on overlaps and dx1 ∧ · · · ∧ dxn equivalent to the chosen
MA3H5 MANIFOLDS 43
so we proceed undeterred and give a proof now that works for all n ≥ 2 without the
need for modification.
P
As in the proof of Theorem 5.8, we write again ω = i θi ω whence
Z XZ
dω = d(θi ω)
M i M
and write
θi ω = f1 dx2 ∧dx3 ∧· · ·∧dxn −f2 dx1 ∧dx3 ∧· · ·∧dxn +· · ·+(−1)n−1 fn dx1 ∧· · ·∧dxn−1 .
There are now two types of open sets among the coordinate charts in use here: those
that do not intersect ∂M , and those which do intersect it. For those which do not
intersect it, the contribution to the integral is zero by what was shown in the proof
of Theorem 5.8. For those which do, we have
Z Z
∂f1 ∂fn
d(θi ω) = + ··· + dx1 dx2 . . . dxn
M xn ≥0 ∂x1 ∂xn
Z
= [fn ]∞
0 dx1 . . . dxn−1
Rn−1
Z
=− fn (x1 , . . . , xn−1 , 0)dx1 . . . dxn−1
n−1
Z R
= θi ω |∂M
∂M
If our endeavour here is likened to a mountain hike, you should not view Stokes’
Theorem as a peak or pinnacle that affords a beautiful view of the scenery and
everything that surrounds it, but rather as a solid pair of mountain boots. In other
words, we are after geometric applications. Here is a first one, we will give many
more in the next Section.
Theorem 5.14 (Brouwer’s fixed point theorem). Let B be the closed unit ball in
Rn and f : B → B a smooth map from B to itself. Then f has a fixed point.
Proof. Arguing by contradiction, suppose there is no fixed point, so f (x) 6= x for all
x ∈ B. Extend the segment f (x)x to a half ray starting from f (x) and intersecting
the boundary S n−1 of B in g(x). Then
g : B → ∂B
44 CHRISTIAN BÖHNING
is a smooth map with the property that g(x) = x for x ∈ ∂B. Now let ω be a
nowhere vanishing (n − 1)-form on ∂B = S n−1 (we have seen above that such an ω
exists), normalised so that Z
ω = 1.
∂B
Then Z Z
1= ω= g∗ω
∂B ∂B
since g is the identity on ∂B. By Stokes’ Theorem
Z Z Z
∗ ∗
g ω= d(g ω) = g ∗ (dω) = 0
∂B B B
using Proposition 5.6 and the fact that dω = 0 as ω is a form in dimension n − 1 on
S n−1 . This is a contradiction (1=0) and hence f must have a fixed point.
Finally, for (iii), one only needs to check that pullback of differential forms really
descends to cohomology groups; we need that the pull-back of a closed form α is
closed. This follows from
d(f ∗ α) = f ∗ (dα) = 0.
And we need that if we change α by an exact form, replacing it by α0 = α + dγ, the
pullback f ∗ α0 still represents the same cohomology class as f ∗ α: again, this follows
from the compatibility of f ∗ and d
f ∗ (α + dγ) = f ∗ α + d(f ∗ γ).
Since f ∗ (α ∧ β) = f ∗ (α) ∧ f ∗ (β), the operation f ∗ respects the product we defined.
We now come to a result that has more geometric content and says that the in-
duced maps ft∗ on cohomology do not vary if ft : M → N is a family of smooth maps,
smoothly varying with a parameter t. This is sometimes referred to as homotopy
invariance of de Rham cohomology. Precisely, it means the following.
Theorem 6.3. Let f : M × [0, 1] → N be a smooth map (by which of course we
mean it is the restriction of some smooth map M × (−, 1 + ) → N for some small
> 0). Put ft (x) = f (x, t). The for the induced maps ft∗ : H r (N ) → H r (M ) we
have
f0∗ = f1∗ .
Proof. We start with a class a ∈ H r (N ) and represent it by a closed r-form α; let
us consider f ∗ α, an r-form on M × [0, 1]. We can write uniquely
f ∗ α = β + dt ∧ γ
with β an r-form on M , depending smoothly on t, and γ an (r − 1)-form on M ,
depending smoothly on t. Here β is nothing but ft∗ α. Denote by dM the exterior
derivative on M . Then from the preceding displayed formula we get
∂β
0 = dM β + dt ∧ − dt ∧ dM γ
∂t
because α is closed. Therefore we must have
∂β
= dM γ.
∂t
Since also
∂ ∗ ∂β
ft α =
∂t ∂t
integration with respect to t yields
Z 1 Z 1
∗ ∗ ∂ ∗
f1 α − f0 α = ft α = dM γ dt.
0 ∂t 0
Thus the cohomology class of f1∗ α is the same as that of f0∗ α since their difference
is exact on M .
Computing de Rham cohomology groups is usually not very easy, but we are
now already in a position to calculate them for some “easy” manifolds, for example,
spheres.
Theorem 6.5. Let S n be the n-sphere in Rn+1 . Then
H 0 (S n ) = H n (S n ) = R and H r (S n ) = 0 otherwise.
Proof. We start with the case n = 1 and then make an induction.
Thus consider S 1 ' R1 /Z first. The 1-form dx on R is translation-invariant, hence
descends to a nowhere vanishing (closed) 1-form, also denoted dx, on R/Z. This dx
is not the differential of a function f on S 1 because such a function would have to
have a minimum on the compact S 1 and at such a point we would have df = 0, but
dx is nowhere vanishing. Hence
H 1 (R/Z) 6= 0.
Now suppose g(x)dx is any 1-form on S 1 (automatically closed); here g can be
viewed as a periodic function on R: g(x + 1) = g(x). If we want to write α = df on
S 1 we must have f 0 (x) = g(x) on R whence we can put
Z x
f (x) = g(s) ds.
0
But we need f periodic f (x + 1) = f (x) for it to descend to S 1 whence we must
have Z 1
g(s) ds = 0.
0
Thus if α = g(x) dx with g some smooth periodic function on R (not necessarily
integral 0 over [0, 1]) we can write
Z 1
α = g(x) dx = g(s) ds dx + df
0
48 CHRISTIAN BÖHNING
R1 Rx
because g̃ = g− 0 g(s) ds has integral zero over [0, 1], and we can put f = 0 g̃(s) ds.
Thus any 1-form on S 1 is of the form
c dx + df
with c ∈ R some constant and f a function on S 1 , whence H 1 (S 1 ) = R.
Now we assume by induction that the statement of the Theorem holds for all
spheres S j for j up to dimension n − 1 and we prove it for S n . As S n is connected,
H 0 (S n ) = R. We now prove first that H p (S n ) = 0 for all 1 < p < n. For this let α be
a closed p-form on S n representing some cohomology class in H p (S n ). The type of
argument we will employ now has been formalised in the form of the Mayer-Vietoris
exact sequence in topology, but we give a proof without the technical baggage. Let
U and V be the complements in S n of small closed balls in Rn+1 around the North
and South pole of S n . Then U and V are diffeomorphic to open balls in Rn via
stereographic projection, which are star-shaped with respect to the center, hence
they have trivial higher de Rham cohomology. Thus the restriction of α to U and
V is exact:
α = du on U and α = dv on V .
On U ∩ V , u − v is closed since d(u − v) = α − α = 0 on the intersection. Since
U ∩ V ' S n−1 × R
and H p−1 (S n−1 × R) = H p−1 (S n−1 ) = 0 by induction, u − v is exact on U ∩ V :
u − v = dw on U ∩ V .
Now U ∩ V is also diffeomorphic to S n−1 × (−2, 2) and we can choose a smooth
function ϕ that is equal to 1 on S n−1 × (−1, 1) and has support in S n−1 × (−2, 2).
Shrinking U and V slightly to get U 0 ⊂ U and V 0 ⊂ V such that U 0 ∩ V 0 =
S n−1 × (−1, 1), we can extend ϕw by 0 outside of its support to get a globally
defined (p − 2)-form on S n . Moreover we have the (p − 1)-form u on U 0 and the
(p − 1)-form
v + d(ϕw)
on V 0 that satisfy u = v + dw = v + d(ϕw) on U 0 ∩ V 0 . Thus we can patch these
together to get a (p − 1)-form β on all of S n with α = dβ globally (because this
holds on the open cover U 0 , V 0 be construction). Thus the cohomology class of α is
trivial and we have shown
H p (S n ) = 0 for 1 < p < n.
Now what about p = 1? Well, we can start the argument above as before and get
a function u − v on U ∩ V with d(u − v) = 0. Hence u − v is a constant c since U ∩ V
is connected for n > 1, which we can assume since the case n = 1 was the induction
base. Then d(v + c) = α on V as well, and v + c and u agree on U ∩ V , hence patch
together to give a globally defined function f on S n with df = α. Hence
H 1 (S n ) = 0; for n > 1
as well.
For p = n, the argument gives us an (n − 1)-form u − v on U ∩ V defining a class
in
H n−1 (S n−1 × R) ' H n−1 (S n−1 ) = R.
MA3H5 MANIFOLDS 49
β = (−1)n+1 kω + d(σ ∗ γ)
and we can write
1 + (−1)n+1 )
β=k ω + dδ
2
with
γ + σ∗γ
δ= .
2
Then δ is invariant under σ. If n is even, β = dδ, and δ descends to a form
δ ∈ A n−1 (PnR ) with α = dδ by construction. Thus H n (PnR ) = 0 in this case.
If n is odd, β = kω + dδ with σ ∗ δ = δ. But also
σ ∗ ω = ω.
Thus there exist σ̄ and δ̄ on PnR with
α = k ω̄ + dδ̄.
Thus dim H n (PnR )
≤ 1 in this case (every class is a multiple of [ω̄]), but since ω̄ is a
volume form inducing an orientation in this case,
H n (PnR ) = R.
The case of the complex projective spaces PnC is rather different: as a real manifold,
PnC has dimension 2n and it can be shown that H r (PnC ) is equal to R precisely if
0 ≤ r ≤ 2n and r is even, and otherwise these groups are zero. In fact, we mention
here as a cultural remark that this and many other calulations become much simpler
once one recognises that for most “nice” spaces M , such as manifolds, the groups
H r (M ) cannot only be calculated with the aid of differential forms, but via a number
52 CHRISTIAN BÖHNING
of different other methods! There is morally just one “cohomology” of the space M ,
and depending on how one initially defines it (e.g. via differential forms, or so-called
singular co-chains, or...) one speaks of de Rham cohomology, singular cohomology,
Alexander-Spanier cohomology, Čech cohomology, ... and the good news is they
all essentially give the same answer! This can be proven using sheaf theory and
sheaf cohomology. There is one further remark we would like to make since we
will see some “shadow” of this phenomenon in the next Section: the cohomology
H r (M ) actually “comes from” an object over Z (i.e., with an integral structure), in
the sense that one can define natural abelian groups (=Z-modules) H r (M, Z) with
H r (M ) = H r (M, Z) ⊗Z R. One can use other “coefficients” than Z as well, but Z is
in some sense a universal choice.
We finally compute the de Rham cohomology of tori.
Theorem 6.9. Let Tn = Rn /Zn (where (m1 , . . . , mn ) ∈ Zn acts on x = (x1 , . . . , xn ) ∈
Rn by x 7→ (x1 + m1 , . . . , xn + mn )). The translation-invariant 1-forms dx1 , . . . , dxn
on Rn descend to 1-forms on Tn which we denote by the same letters. Then
n
dim H r (Tn ) =
r
and H r (Tn ) has a basis consisting of the classes of the r-form
dxi1 ∧ · · · ∧ dxir , 1 ≤ i1 < · · · < ir ≤ n.
Proof. Let α ∈ F r (Tn ) be a closed r-form. We will show the following:
(1) There are n natural commuting one-parameter groups of diffeomorphisms
(i)
Gt , i = 1, . . . , n, acting on Tn = Rn /Zn : indeed,
(i)
Gt : [(x1 , . . . , xn )] 7→ [(x1 , . . . , xi + t, . . . , xn )],
Q (i)
so the action of i Gt is just the action of Tn on itself by translations. We
will define an average ᾱ of α under all of these translations, and show that this
average ᾱ is closed, defines the same cohomology class as α, and is invariant
under all translations.
(2) We show that r-forms invariant under all translations are combinations of the
dxi1 ∧ · · · ∧ dxir with constant coefficients. We denote these invariant forms
by Invr (Tn ). We have Invr (Tn ) ⊂ F r (Tn ).
(3) We know at that point that Invr (Tn ) and B r (Tn ) span F r (Tn ), so we show
Invr (Tn ) ∩ B r (Tn ) = 0
to conclude that Invr (Tn ) maps isomorphically onto H r (Tn ) under the pro-
jection F r (Tn ) → H r (Tn ).
Now let us implement this program in more detail. For (1), let us start looking at
(1)
Gt := Gt . We then define
Z 1
µ1 (α) = G∗θ α dθ.
0
Since G∗t is linear and commutes with the integral sign, we get that µ1 (α) is invariant
under Gt , i.e. G∗t (µ1 (α)) = µ1 (α) for
every t. Also µ1 (α) ∈ F r (Tn ) since
Z 1 Z 1 Z 1
∗ ∗
d(µ1 (α)) = d Gθ α dθ = d(Gθ α) dθ = G∗θ dα dθ = 0.
0 0 0
MA3H5 MANIFOLDS 53
Moreover, µ1 (α) and α define the same class in H r (Tn ) since, following the method
of proof for Theorem 6.3, we can write
G∗t α − G∗0 α = G∗t α − α = dβt
with β some (r − 1)-form depending smoothly on t. Hence
Z 1 Z 1 Z 1
∗
µ1 (α) = Gθ α dθ = (α + dβθ ) dθ = α + d βθ dθ
0 0 0
and thus µ1 (α) and α define the same class.
We can then employ these argument inductively to conclude that
ᾱ = µn (µn−1 (. . . (µ2 (µ1 (α))) . . . )),
(i)
where µi denotes the analogous average with respect to Gt , is invariant under all
translations, closed and of the same class as α.
Now for (2), we suppose we are given an r-form γ that is invariant under all
translations. We can write, using multi-index notation again,
X
γ= aI dxI
I
(thus dxI = dxi1 ∧ · · · ∧ dxir ), with the aI functions that, when we view them as
functions on Rn , are periodic for the lattice Zn . But since the dxI are invariant
under all translations, for γ to be invariant under all translations, it must be true
that the aI are invariant under all translations, i.e. they must be constants. In
particular, such a γ is closed.
Now Step (3) can be achieved as follows: assume α ∈ Invr (Tn ) and also α = dβ
for some β, so α ∈ B r (Tn ). We want to show that α = 0. Let (−) be the averaging
operator introduced above. Then
ᾱ = dβ = dβ̄.
But β̄ is invariant under all translations, hence β̄ is closed. Hence dβ̄ = 0, and thus
ᾱ = α = 0.
The idea used in the proof above can be used to compute the de Rham cohomology
of several other manifolds acted on by some groups that are so-called compact Lie
groups.
In particular, this shows that deg(f ) is always an integer. We will give several
applications of this afterwards.
We start with a crucial auxiliary result.
Lemma 7.1. Let Qn = (−1, 1)n be the open unit cube in Rn and let ω ∈ A n (Rn )
with support in Qn and Z
ω = 0.
Qn
Then there exists a β ∈ A n−1 (Rn ) with support in Qn and such that ω = dβ.
Remark that since Qn is star-shaped, Poincaré’s Lemma tells us that there exists
a β with ω = dβ (since dω = 0), but we do not know that such a β will have support
in Qn . In fact, just in case you are wondering why the Lemma is not just Poincaré’s
Lemma “+”, and why we have to R work quite a bit to get it at this point, note that
we do not have supp β ⊂ Qn if Qn ω 6= 0 (which explains the extra assumption):
indeed, by Stokes’ Theorem
Z Z Z Z
β= dβ = ω= ω
∂Qn Qn Qn Qn
(this is actually not quite the generality we proved Stokes’ Theorem in because
∂Qn has “corners”, but one can rather easily generalise it to this case), but of
supp β ⊂ Qn , the first integral would be zero.
Proof. We will prove the statement by induction on the dimension n, but we will
prove a slightly stronger assertion (which often simplifies proofs by induction because
the you can assume more in the inductive step): namely we will prove the statement
with ω and β depending smoothly on some additional parameter λ ∈ Rm , and will
also show that if ω vanishes identically for some value of λ, the same holds for β.
We start with n = 1. We can then write ω = f (x, λ)dx. Setting
Z x
β(x, λ) = f (t, λ)dt
−1
defines a function with dβ = ω; there is a δ > 0 such that f vanishes for x > 1 − δ
and x < −1 + δ since suppω ⊂ (−1, 1) and thus
Z x Z 1
f (t, λ)dt = f (t, λ)dt = 0
−1 −1
R
(by the hypothesis Qn ω = 0) for x > 1 − δ, and similarly for x < −1 + δ. So β
has support in (−1, 1) as well. Also if f (x, λ) = 0 for all x for some λ, then β(x, λ)
vanishes for all x, too, by its definition.
Now for the induction step, let us assume that the stronger statement we want
to prove is true for all dimensions less than n, and we will prove it for n. Write
ω = f (x1 , . . . , xn , λ)dx1 ∧ · · · ∧ dxn .
MA3H5 MANIFOLDS 55
and this is a finite sum. It suffices to show that each ηj α defines a cohomology class
proportional to [α0 ], so without loss of generality we can suppose that α has support
in one of the coordinate neighbourhoods Um , say. Since M is connected, we can
connect any p ∈ U1 and q ∈ Um by a path, and possibly by reordering the Ui and
allowing some duplications, we can assume that the path is covered by U1 , . . . , Um
such that Ui ∩ Ui+1 6= ∅. Now choose an n-form αi with support in Ui ∩ Ui+1 ,
1 ≤ i ≤ m − 1, and integral 1. Then on U1
Z
(α0 − α1 ) = 0
where sgn(det(f∗ )p is ±1 and defined as follows: the fact that M and N are assumed
to be oriented means that we have chosen orientations for M and N , i.e. equivalence
classes of nowhere vanishing
R globalRn-forms ωM and ωN , which we can also assume
to be normalised so that M ωM = N ωN = 1; the differential
(f∗ )p : Tp (M ) → Tf (p) (N )
induces a map
Λn ((f∗ )p )∗ : Λn Tf (p) (N )∗ → Λn Tp (M )∗
whence we have (Λn ((f∗ )p )∗ )(ωN,f (p) ) = λωM,p for some λ ∈ R. Then sgn(det(f∗ )p )
is the sign of λ. More concretely, if we choose local coordinates (U, x1 , . . . , xn )
around p and (V, y1 , . . . , yn ) around q = f (p) such that (ωM ) |U is equivalent to
dx1 ∧ · · · ∧ dxn and (ωN ) |V is equivalent to dy1 ∧ · · · ∧ dyn , and we write f =
(f1 (x1 , . . . , xn ), . . . , fn (x1 , . . . , xn )) in these local coordinates, then sgn(det(f∗ )p ) is
nothing but the sign of the determinant of the Jacobian matrix:
∂f
det ∂xji .
Definition 7.7. Let {f, g} be a pair of curves. The linking number of {f, g},
denoted by link(f, g), is the degree of the smooth map µ : S 1 × S 1 → S 2 given by
f (t) − g(s)
µ(t, s) =
||f (t) − g(s)||
where || · || is the Euclidean norm.
We now wish to relate this number link(f, g) to the geometric idea that the curves
can be “linked” or “unlinked”.
We will denote by {f0 , g0 } the pair of curves in R3 consisting of the standard
parametrisations of the circles
f0 : S 1 → (x, y, z) ∈ R3 : z = 0 and (x − 2)2 + y 2 = 1 ,
Note that the images of f0 and g0 are just translations of the standard copy of S 1
of radius 1 around the origin in the (x, y)-plane, and these translations give what
we call “standard parametrisations”.
Definition 7.8. We call a pair of curves {f, g} unlinked if it is isotopic to the pair
{f0 , g0 }. Otherwise we call it linked.
We then have the following criterion.
Theorem 7.9. If {f, g} and {f 0 , g 0 } are isotopic pairs of curves, we have
link(f, g) = link(f 0 , g 0 ).
Moreover, if link(f, g) 6= 0, then the pair {f, g} is linked.
Proof. Let {F, G} be an isotopy between the isotopic pairs of curves {f, g} and
{f 0 , g 0 }. For a fixed t, we associate to the pair {Ft , Gt } the map µt : S 1 × S 1 → S 2
given by
Ft (x) − Gt (y)
µt (x, y) = .
||Ft (x) − Gt (y)||
We can then define a smooth map
µ : [0, 1] × S 1 × S 1 → S 2
by µ(t, x, y) = µt (x, y). We can apply Theorem 6.3 to conclude that deg(µt ) is
constant in the family. Hence
link(f, g) = deg(µ0 ) = deg(µ1 ) = link(f 0 , g 0 ).
We now show that if f and g are unlinked, then their linking number is zero. By
definition, unlinkedness means that {f, g} is isotopic to {f0 , g0 }, so by the first part
we only have to show that the linking number of f0 and g0 is zero. But the images
of f0 and g0 are contained in the (x, y)-plane, so the image of µ is also contained
in the (x, y)-plane and cannot be all of S 2 . Thus µ is not surjective, which means
deg(µ) = link(f0 , g0 ) = 0 by Theorem 7.4.
Example 7.10. We should show that link(f, g) is not by chance always identically
zero. Consider the circles in Figure 1, top. The images of f and g lie in mutually
perpendicular planes, the (x, y)- and (x, z)-planes, and their centres lie one the x-
axis, the line of intersection of these two planes. We claim that link(f, g) = ±1 (the
sign depending on the chosen orientations). Consider the map
µ : S1 × S1 → S2
MA3H5 MANIFOLDS 61
from Definition 7.7. Let e1 = (1, 0, 0) ∈ S 2 . We want to know what µ−1 (e1 ) is. Now
if (a, b) ∈ µ−1 (e1 ), then f (a) − g(b) must be parallel to the x-axis and point in the
positive direction towards +∞ on that axis: hence f (a) and g(b) must be the points
shown in the figure. Therefore, µ−1 (e1 ) is a single point.
claim that the pair of curves on the right then also has to have linking number zero
(though it is intuitively clear that this pair of curves is linked). Indeed, the proof
of the first part of Theorem 7.9 shows a little more, namely that link(f, g) is not
only invariant under isotopy, bot under what we may call (smooth) homotopy: here,
in slight modification of Definition 7.6, we call pairs of curves {f, g} and {f 0 , g 0 }
(smoothly) homotopic if there are smooth maps
F : [0, 1] × S 1 → R3 , G : [0, 1] × S 1 → R3
such that Ft (x) := F (t, x) and Gt (x) := G(t, x) satisfy F0 = f, F1 = f 0 , G0 =
g, G1 = g 0 and
Ft (S 1 ) ∩ Gt (S 1 ) = ∅
for every t ∈ [0, 1]. In other words, we relax the assumption that
{Ft , Gt }
be a pair of curves for every t ∈ [0, 1], to allow more general maps Ft , Gt , t ∈ (0, 1),
that need not be smooth embeddings anymore: for example, the image of Ft (or Gt )
could be a curve with self-intersections (so that Ft is no longer injective), or even
wilder.
Then it is clear intuitively that the pairs of curves in Figure 1, bottom are
smoothly homotopic.
Example 7.12. As an example, that the linking number link(f, g) in some sense
really measures the number of times the curve f goes around the curve g, we en-
courage the reader to contemplate the homotopy depicted in Figure 2, where the red
“coil” winds around the green curve n times (n = 3 is shown). Then that homotopy
MA3H5 MANIFOLDS 63
shows that this situation is homotopic to one with a red circle being linked to the
green curve as shown, with the red circle being traversed n times (corresponding to
the map S 1 → S 1 , z 7→ z n , z ∈ C, |z| = 1), so link(f, g) = ±n.
We now leave linking numbers of curves and come to a final and very important
concept that can be accessed using degree theory: the index of an isolated singularity
of a vector field. As we will explain this has many wonderful applications and is
also intimately related to the global topological properties of a manifold.
We start with the situation in Rn , so let U ⊂ Rn be an open subset and let X
be a vector field on U . We may view X simply as a smooth map X : U → Rn . We
call a point p ∈ U a singularity of X if X(p) = 0. In addition, we call it an isolated
singularity if X(q) 6= 0 for every q 6= p in a neighbourhood V of p in U .
Proposition 7.13. Let p be an isolated singularity of the vector field X on the open
subset U in Rn , and denote by S n−1 (p, ) the sphere of radius > 0 and centre p in
Rn . Let S n−1 = S n−1 (0, 1). If V is a neighbourhood of p that contains no further
singularities of X than p and if is so small that S n−1 (p, ) ⊂ V , then the degree of
the map
fX : S n−1 → S n−1 ,
X(p + x)
x 7→
||X(p + x)||
does not depend on .
We call this degree the index of X at p and denote it by indp X.
Proof. Let and δ be small real numbers such that S(p, ) and S(p, δ) are both
contained in V . Define
F : [0, 1] × S n−1 → S n−1
by
X
F (t, x) = f(1−t)+tδ (x).
Since X does not vanish on V except in p, there exist an α > 0 such that for every
t ∈ (−α, 1 + α) the denominator of ftX does not vanish for any x. Thus, since
F0 = fX and F1 = fδX , we get the result from Theorem 6.3.
Example 7.14. You can easily calculate the following:
(1) If we put X = idRn , then the origin is an isolated singularity of index 1.
(2) For X = −idRn , the origin is an isolated singularity of index (−1)n .
(3) If X on R2 is given by X(x, y) = (−y, x), then the origin is an isolated
singularity of index one.
(4) If X(x, y) = (x2 − y 2 , 2xy) in R2 , then the origin is an isolated singularity of
index 2.
We just calculate (4) for fun (or punishment, whichever your view) and the rest is
seen similarly, but easier. We recognise (4) as the map z 7→ z 2 for z = x + iy, and
thus the preimage of any point p ∈ S 1 consists of two points ±p. Since the map
preserves orientation, ind0 X = 2.
You should look at the primitive sketches of these vector field that I made in
Figure 3, and use these to confirm that the index has the following geometric inter-
pretation: traverse a little circle around the origin in counterclockwise direction; at
each point of your way, the vector field furnishes you with an arrow that you can
view as the position of the hand of a clock. Then the index is the number of times
this hand turns on the clock as you traverse the circle and return to your starting
64 CHRISTIAN BÖHNING
point (where you count each full turn in counterclockwise direction as +1, and turns
in clockwise direction as −1).
We now give a fun application of the concept of the index of a vector field, called
the argument principle in complex analysis. Suppose we want to count the number
of zeros of a complex polynomial
p(z) = z m + a1 zm−1 + · · · + am
that lie in a smooth compact region Ω of the complex plane whose boundary we
assume contains no zero of p; we will assume for simplicity and concreteness that
Ω is just some some closed ball, i.e. a disc, in the plane, but other cases would be
possible, too. We can view this as a manifold with boundary, oriented in a natural
way using the standard orientation in the ambient R2 = C.
Proposition 7.15. The number of zeros of p(z) in Ω, counted with multiplicity,
equals the degree of the map
p
: ∂Ω → S 1 .
|p|
This number is equal to the sum of the indices of singularities of the vector field
X(z) = p(z) contained in Ω.
Here the expression “counted with multiplicity” means the following: if
N
Y
p(z) = (z − αi )mi
i=1
and α1 , . . . , αt are the zeros of p(z) lying inside Ω, the number in Proposition 7.15
is by definition
m1 + · · · + mt .
MA3H5 MANIFOLDS 65
Proof. We can find pairwise disjoint classed balls Di = B(αi , i ), i > 0 small real
numbers, i = 1, . . . , t, contained in the interior of Ω. We will now apply Stokes’
Theorem to the oriented manifold with boundary
t
[
D = Ω\ B(αi , i )
i=1
(here B(αi , i ) denotes the open disk). Note that the map
p(z)
Fp (z) :=
||p(z)||
is defined and smooth in a neighbourhood of D. It takes values in S 1 ⊂ C. Denoting
by ω the canonical volume form giving the orientation on S 1 (normalised so that its
integral over S 1 is 1), we thus have by Stokes’ Theorem
Z Z Xt Z
∗ ∗
0= d(Fp ω) = Fp ω − Fp∗ ω.
D ∂Ω i=1 S 1 (αi ,i )
Indeed, we can do somewhat better. The following result sometimes goes under
the name argument principle.
Theorem 7.16. The number of zeros (counted with multiplicity) of the complex
polynomial p(z) in the disk Ω (with zero-free boundary) is equal to
Z
1
d(arg p(z)).
2π ∂Ω
We first make some comments so that you can parse the formula in the statement
of the Theorem. Each nonzero complex number w can be written uniquely as
w = reiθ
where r is the norm of w, and θ is, by definition, the argument. However, note
that θ = arg (w) is not really a well-defined function: all the values θ + 2πn, n any
integer, qualify equally to be the argument of w. Luckily, if we take the exteriro
derivative d, this ambiguity disappears since any two choices of argument differ by
addition of a constant. Indeed, in a suitable neighbourhood of any point in C − {0},
we can always choose values of arg (w) so that this becomes a smooth function near
the point; and any other local such choice will differ from the preceding one by some
constant 2πn. Thus we get a well-defined 1-form
d arg
on all of C − {0}, but this is not the differential of a well-defined function as the
notation may erroneously suggest.
The integrand in Theorem 7.16 is the 1-form
z 7→ d arg p(z) = p∗ (d arg)
which is well-defined and smooth on the complex plane minus the zeros of p.
(Proof of Theorem 7.16). We already know by Proposition 7.15 that the number of
zeros counted with multiplicity inside Ω is equal to the degree of the map
p
f := : ∂Ω → S 1 .
|p|
Now
p(z)
= eiarg p(z) ,
|p(z)|
and all that remains to do is to identify the integral of d(arg p(z)) over ∂Ω with
2π deg(f ).
We apply Theorem 7.4 to the restriction of the smooth 1-form d arg to S 1 . If
w = f (z) = eiarg p(z)
then arg w = arg p(z) and since arg p(z) is locally a well-defined smooth function,
we get
d arg p(z) = d(f ∗ arg w) = f ∗ (d arg w).
Thus Z Z
d(arg p(z)) = deg(f ) d arg w.
∂Ω S1
But calculating Z
d arg w
S1
MA3H5 MANIFOLDS 67
We can use the argument in Lemma 7.18 to calculate the index of a vector field
in a much simpler way in certain special cases:
Proposition 7.19. Suppose X : U → Rn is a vector field on some open subset U of
Rn , and let p ∈ U be an isolated singularity of X. If det(dXp ) 6= 0, then we have
(
1 if det(dXp ) > 0,
indp X =
−1 if det(dXp ) < 0.
Proof. Indeed, we can assume p = 0, and then X is a diffeomorphism on a neigh-
bourhood of 0, which we can assume to be star-shaped. If det(dXp ) > 0, then X is
isotopic to the identity by Lemma 7.18, and each Ft in this isotopy can be consid-
ered as a vector field having an isolated singularity at p; by the invariance of degree
(which again follows from Theorem 6.3), we get
indp X = 1.
If X has det(dXp ) < 0, we reduce to the previous case by considering X ◦ σ where
σ is the reflection in a hyperplane.
We conclude by mentioning a celebrated result, the Poincaré-Hopf Index Theo-
rem, that relates the concept of indices of isolated singularities of vector fields on
compact manifolds with a basic topological invariant of the manifold, the Euler
characteristic (sometimes also called the Euler number ).
Theorem 7.20. If M is a compact manifold and X a vector field on M having only
isolated singularities p1 , . . . , pn , then
X n
indpi X = χ(M )
i=1
is the so-called Euler characteristic of M , which can be computed in the following
alternative way: all de Rham cohomology groups H i (M ) are finite-dimensional R-
vector spaces if M is compact, and
dim
XM
χ(M ) = (−1)i dimR H i (M ).
i=0
We do not have time in the course of this lecture to prove this result, but did
not want to pass it by in silence because it is so beautiful and important; a proof
can be found, for example, in the wonderful book by Raoul Bott and Loring W. Tu,
Differential Forms in Algebraic Topology, Springer (1982), p. 126 ff.
We just note that from our earlier computation of the de Rham cohomology of
spheres and tori, it follows that for M = S n , we get
χ(S n ) = 1 + (−1)n
and for M = Tn ,
χ(Tn ) = 0.
Moreover, for M = PnR we get
(
1 if n is even,
χ(PnR ) =
0 if n is odd.
It is easy in low dimensions to picture vector fields with isolated singularities on
these manifolds that allow us to confirm the equality stated in the Poincaré-Hopf
MA3H5 MANIFOLDS 69
Index Theorem explicitly for these examples. The reader should compare Figure 4
and check that in each case the sum of the indices of the vector fields depicted is
equal to the Euler characteristic.
References
[BeGo88] Berger, M., Gostiaux, B., Differential Geometry: Manifolds, Curves and Surfaces, GTM
115, Springer (1988)
[BoTu82] Bott, R., Tu, L.W., Differential Forms in Algebraic Topology, GTM 82, Springer (1982)
[GuPo74] Guillemin, V., Pollack, A., Differential Topology, reprint of the 1974 edition, AMS
Chelsea Publishing, American Mathematical Society, Providence, Rhode Island (2014)
[Hitchin] Hitchin, N., Differentiable Manifolds, online notes available at
http://people.maths.ox.ac.uk/ joyce/Nairobi2019/Hitchin-DifferentiableManifolds.pdf
[KoNo63] Kobayashi, S., Nomizu, K., Foundations of Differential Geometry, Volume I, John Wiley
& Sons Inc. (1963)
[MoKo71] Morrow, J., Kodaira, K., Complex Manifolds, reprint of the 1971 edition, AMS Chelsea
Publishing, American Mathematical Society, Providence, Rhode Island (2006)
[Tu11] Tu, L.W., An Introduction to Manifolds, 2nd Edition, Springer Universitext, Springer
(2011)
[Warner83] Warner, F.W., Foundations of Differentiable Manifolds and Lie Groups, GTM 94,
Springer (1983)