Mfds Merged
Mfds Merged
The line and surface integrals studied in vector calculus require a concept of a curve and a
surface in Euclidean 3-space. These latter objects are introduced via a parameterization:
a smooth map of, respectively, an interval on the real line R or a domain in the plane R2
into R3 . In fact, one often requires a so-called regular parameterization. For a curve r(t),
this means non-vanishing of the ‘velocity vector’ at any point, ṙ(t) 6= 0. On a surface a
point depends on two parameters r = r(u, v) and regular parameterization means that the
two vectors of partial derivatives in these parameters are linearly independent at every
point ru (u, v) × rv (u, v) 6= 0.
Differential Geometry develops a more general concept of a smooth n-dimensional
differentiable manifold.1 and a systematic way to do differential and integral calculus (and
more) on manifolds. Curves and surfaces are examples of manifolds of dimension d = 1
and d = 2 respectively. However, in general a manifold need not be given (or considered)
as lying in some ambient Euclidean space.
The last two parts of the above definition will be needed to avoid some pathological
examples (see below).
1
More precisely, it is often useful to also consider appropriate ‘structures’ on manifolds (e.g. Riemannian
metrics), as we shall see in due course.
1
2 alexei kovalev
The knowledge of open subsets enables one to speak of continuous maps: a map between
topological spaces is continuous if the inverse image of any open set is open. Exercise: check
that for maps between open subsets (in the usual sense) of Euclidean spaces this definition
is equivalent to other definitions of a continuous map.
A homeomorphism is a bijective continuous map with continuous inverse. More ex-
plicitly, to say that ‘a bijective mapping ϕ of U onto V is a homeomorphism’ means that
‘D ⊂ U is open if and only if ϕ(D) ⊂ V is open’.
Let M be a topological space. A homeomorphism ϕ : U → V of an open set U ⊆ M
onto an open set V ⊆ Rd will be called a local coordinate chart (or just ‘a chart’) and
U is then a coordinate neighbourhood (or ‘a coordinate patch’) in M .
Definition. A C ∞ differentiable structure, or smooth structure, on M is a collection
of coordinate charts ϕα : Uα → Vα ⊆ Rd (same d for all α’s) such that
(i) M = ∪α∈A Uα ;
(ii) any two charts are ‘compatible’: for every α, β the change of local coordinates ϕβ ◦ϕ−1
α
is a smooth (C ∞ ) map on its domain of definition, i.e. on ϕα (Uβ ∩ Uα ) ⊆ Rd .
(iii) the collection of charts ϕα is maximal with respect to the property (ii): if a chart ϕ
of M is compatible with all ϕα then ϕ is included in the collection.
A bijective smooth map with a smooth inverse is called a diffeomorphism. Notice
that clause (ii) in the above definition implies that any change of local coordinates is a
diffeomorphism between open sets ϕα (Uβ ∩ Uα ) and ϕβ (Uβ ∩ Uα ) of Rd .
In practice, one only needs to worry about the first two conditions in the above defini-
tion. Given a collection of compatible charts covering M , i.e. satisfying (i) and (ii), there
is a unique way to extend it to a maximal collection to satisfy (iii). I leave this last claim
without proof but refer to Warner, p. 6.
Definition . A topological space equipped with a C ∞ differential structure is called a
smooth manifold. Then d is called the dimension of M , d = dim M .
Sometimes in the practical examples one starts with a differential structure on a set
of points M (with charts being bijective maps onto open sets in Rd ) and then defines
the open sets in M to be precisely those making the charts into homeomorphisms. More
explicitly, one then says that D ⊂ M is open if and only if for every chart ϕ : U → V ⊆ Rd ,
ϕ(D ∩ U ) is open in Rd . (For this to be well-defined, every finite intersection of coordinate
neighbourhoods must have an open image in Rd under some chart.) We shall refer to this
as the topology induced by a C ∞ structure.
Remarks . 1. Some variations of the definition of the differentiable structure are possible.
Much of the material in these lectures could be adapted to C k rather than C ∞ differentiable
manifolds, for any integer k > 0. 2
On the other hand, replacing Rd with the complex coordinate space Cn and smooth
maps with holomorphic (complex analytic) maps between domains in C, leads of an im-
portant special class of complex manifolds—but that is another story.
2
if k = 0 then the definition of differentiable structure has no content
differential geometry 3
By a manifold I will always mean in these lectures a smooth (real) manifold, unless
explicitly stated otherwise.
2. Here is an example of what can happen if one omits a Hausdorff property. Consider
the following ‘line with a double point’
M = (−∞, 0) ∪ {00 , 000 } ∪ (0, ∞),
with two charts being the obvious ‘identity’ maps (the induced topology is assumed)
ϕ1 : U1 = (−∞, 0) ∪ {00 } ∪ (0, ∞) → R, ϕ2 : U2 = (−∞, 0) ∪ {000 } ∪ (0, ∞) → R, so
ϕ1 (00 ) = ϕ1 (000 ) = 0. It is not difficult to check that M satisfies all the conditions of a
smooth manifold, except for the Hausdorff property (00 and 000 cannot be separated).
Omitting the 2nd countable property would allow, e.g. an uncountable collection (dis-
joint union) of lines
t0<α<1 Rα ,
each line Rα equipped with the usual topology and charts being the identity maps Rα → R.
3
Examples. The Euclidean Rd is made into a manifold using the identity chart. The
complex coordinate space Cn becomes a 2n-dimensional manifold via the chart Cn → R2n
replacing every complex coordinate zjP by a pair of real coordinates Re zj , Im zj .
n
The sphere S n = {x ∈ Rn+1 : 2
i=0 xi = 1} is made into a smooth manifold of
dimension n, by means of the two stereographic projections onto Rn ∼ = {x ∈ Rn+1 :
x0 = 0}, from the North and the South poles (±1, 0, . . . , 0). The corresponding change of
coordinates is given by (x1 , . . . , xn ) 7→ (x1 /|x|2 , . . . , xn /|x|2 ).
The real projective space RP n is the set of all lines in Rn+1 passing through 0. Elements
of RP n are denoted by x0 : x1 : . . . : xn , where not all xi are zero. Charts can be given by
ϕi (x0 : x1 : . . . : xn ) = (x0 /xi , . . . î . . . , xn /xi ) ∈ Rn with changes of coordinates given by
ϕj ◦ ϕ−1 n
i : (y1 , . . . , yn ) 7→ y1 : . . . : (1 in ith place) : . . . : yn ∈ RP 7→
y1 yi−1 1 yi yj−1 yj+1 yn
,..., , , ,..., , ... ,
yj yj yj yj yj yj yj
smooth functions on their domains of definition (i.e. for yj 6= 0). Thus RP n is a smooth
n-dimensional manifold.
Definition . Let M, N be smooth manifolds. A continuous map f : M → N is called
smooth (C ∞ ) if for each p ∈ M , for some (hence for every) charts ϕ and ψ, of M and
N respectively, with p in the domain of ϕ and f (p) in the domain of ψ, the composition
ψ ◦ f ◦ ϕ−1 (which is a map between open sets in Rn , Rk , where n = dim M , k = dim N )
is smooth on its domain of definition.
Exercise: write out the domain of definition for ψ ◦ f ◦ ϕ−1 .
Two manifolds M and N are called diffeomorphic is there exists a smooth bijective
map M → N having smooth inverse. Informally, diffeomorphic manifolds can be thought
of as ‘the same’.
3
For a more interesting example on a ‘non 2nd countable manifold’ see Example Sheet Q1.12.
4 alexei kovalev
Definition . A group G is called a Lie group if it is a smooth manifold and the map
(σ, τ ) ∈ G × G → στ −1 ∈ G is smooth.
Let A be an n × n complex matrix. The norm given by |A| = n maxij |aij | has a useful
property that |AB| ≤ |A||B| for any A, B. The exponential map on the matrices is defined
by
exp(A) = I + A + A2 /2! + . . . + An /n! + . . . .
The series converge absolutely and uniformly on any set {|A| ≤ µ}, by the Weierstrass
M-test. It follows that e.g. exp(At ) = (exp(A))t and exp(C −1 AC) = C −1 exp(A)C,
for any invertible matrix C. Furthermore, the term-by-term differentiated series also con-
verge uniformly and so exp(A) is C ∞ -smooth in A. (This means smooth as a function of
2n2 real variables, the entries of A.)
The logarithmic series
log(I + A) = A − A2 /2 + . . . + (−1)n+1 An /n + . . .
converge absolutely for |A| < 1 and uniformly on any closed subset {|A| ≤ ε}, for ε < 1,
and log(A) is smooth in A.
One has
exp(log(A)) = A, when |A − I| < 1. (1.1)
This is true in the formal sense of composing the two series in the left-hand side. The
formal computations are valid in this case as the double-indexed series in the left-hand
side is absolutely convergent.
For the other composition, one has
Proposition 1.3. The orthogonal group O(n) = {A ∈ GL(n, R) : AAt = I} has a smooth
structure making it into a manifold of dimension n(n − 1)/2.
n(n−1)
The charts take values in the 2
-dimensional linear space of skew-symmetric n × n
real matrices. E.g.
As can be seen from (1.4), the standard basis vectors of Tp M given by the local coor-
dinates xi and x0i are related by
0
∂xj
∂ ∂
= . (1.40 )
∂xi p ∂xi p ∂x0j p
This is what one would expect in view of the chain rule from the calculus. The formula (1.40 )
tells us that every tangent vector ai ( ∂x∂ i )p gives a well-defined first-order ‘derivation’
∂ ∂f
ai ( )p : f ∈ C ∞ (M ) → ai (p) ∈ R. (1.5)
∂xi ∂xi
In fact the following converse statement is true although I shall not prove it here5 . Given
p ∈ M , every linear map a : C ∞ (M ) → R satisfying Leibniz rule a(f g) = a(f )g(p) +
f (p)a(g) arises from a tangent vector as in (1.5).
Example 1.6. Let r = r(u, v), (u, v) ∈ U ⊆ R2 be a smooth regular-parameterized surface
in R3 . Examples of tangent vectors are the partial derivatives ru , rv —these correspond
∂ ∂
to just ∂u , ∂v in the above notation, as the parameterization by u, v is an instance of a
coordinate chart.
Definition. A vector space with a multiplication [·, ·], bilinear in its arguments (thus satis-
fying the distributive law), is called a Lie algebra if the multiplication is anti-commutative
[a, b] = −[b, a] and satisfies the Jacobi identity [[a, b], c] + [[b, c], a] + [[c, a], b] = 0.
Theorem 1.7 (The Lie algebra of a Lie group). Let G be a Lie group of matrices and
suppose that log defines a coordinate chart near the identity element of G. Identify the
tangent space g = TI G at the identity element with a linear subspace of matrices, via the
log chart, and then g is a Lie algebra with [B1 , B2 ] = B1 B2 − B2 B1 .
The space g is called the Lie algebra of G.
Proof. It suffices to show that for every two matrices B1 , B2 ∈ g, the [B1 , B2 ] is also an
element of g. As [B1 , B2 ] is clearly anticommutative and the Jacobi identity holds for
matrices, g will then be a Lie algebra.
The expression
defines, for |t| < ε with sufficiently small ε, a path A(t) in G such that A(0) = I. Using
for each factor the local formula exp(Bt) = I + Bt + 21 B 2 t2 + o(t2 ), as t → 0, 6 we obtain
Hence
B(t) = log A(t) = [B1 , B2 ]t2 + o(t2 ) and exp(B(t) = A(t)
hold for any sufficiently small |t| and so B(t) ∈ g (as B(t) is in the image of the log
chart). Hence B(t)/t2 ∈ g for every small t 6= 0 (as g is a vector space). But then
also limt→0 B(t)/t2 = limt→0 ([B1 , B2 ] + o(1)) = [B1 , B2 ] ∈ g (as g is a closed subset of
matrices).
Notice that the idea behind the above proof is that the Lie bracket [B1 , B2 ] on a Lie
algebra g is an ‘infinitesimal version’ of the commutator g1 g2 g1−1 g2−1 in the corresponding
Lie group G.
Theorem 1.8 (The ‘manifold of tangent vectors’). The tangent bundle T M has a
canonical differentiable structure making it into a smooth 2n-dimensional manifold, where
n = dim M .
Theorem 1.9. Suppose that on a smooth manifold M of dimension n there exist n vector
fields X (1) , X (2) , . . . , X (n) , such that X (1) (p), X (2) (p), . . . , X (n) (p) form a basis of Tp M at
every point p of M . Then T M is isomorphic to M × Rn .
Φ : ~a ∈ T M → (π(~a); a1 , . . . , an ) ∈ M × Rn .
It is clear from the construction and the hypotheses of the theorem that Φ is a bijection
and Φ converts every tangent space into a copy of Rn . It remains to show that Φ and Φ−1
are smooth.
7
The topology on T M is induced from the smooth structure.
8 alexei kovalev
(i)
where xi are local coordinates on U , and ~a = bj ∂x∂ j . Writing X (i) = Xj (x) ∂x∂ j , we obtain
(i) (i)
bj = ai Xj (x), which shows that Φ−1 is smooth. The matrix X(x) = (Xj (x)) expresses a
change of basis of Tp M , from ( ∂x∂ j )p to X (j) (p), and is smooth in x, so the inverse matrix
C(x) is smooth in x too. Therefore ai = bj Cij (x) verifies that Φ is smooth.
Remark . The hypothesis of Theorem 1.9 is rather restrictive. In general, a manifold need
not admit any non-vanishing smooth (or even continuous) vector fields at all (as we shall
see, this is the case for any even-dimensional sphere S 2n ) and the tangent bundle T M will
not be a product M × Rn .
It follows, by direct calculation in local coordinates using (1.40 ), that the differential is
independent of the choice of charts and that the chain rule d(F2 ◦ F1 )p = (dF2 )F1 (p) ◦ (dF1 )p
holds.
Every (smooth) vector field, say X on M , defines a linear differential operator of first
order X : C ∞ (M ) → C ∞ (M ), according to (1.5) (with p allowed to vary in M ). Suppose
that F : M → N is a diffeomorphism. Then for each vector field X on M , (dF )X is a
well-defined vector field on N . For any f ∈ C ∞ (N ), the chain rule in local coordinates
∂yj
∂xi
(p) ∂y∂ j F (p) f = ∂x∂ i p (f ◦ F ) (xi on M and yj on N ) yields a coordinate-free relation
((dF )X)f ◦ F = X(f ◦ F ), (1.10)
and using the left-invariant property (1.14) of Xξ and Xη for the next step
= [Xξ ◦ Lg , Xη ◦ Lg ]f ◦ Lg = ([Xξ , Xη ] ◦ Lg )f ◦ Lg .
Thus (dLg )[Xξ , Xη ]f ◦ Lg = ([Xξ , Xη ] ◦ Lg )f ◦ Lg , for each g ∈ G, f ∈ C ∞ (G), so the vector
field [Xξ , Xη ] satisfies (1.14) and is left-invariant.
10 alexei kovalev
It follows that [Xξ , Xη ] = Xζ for some ζ ∈ g, thus the Lie bracket on l(G) induces one
on g. We can identify this Lie bracket more explicitly for the matrix Lie groups.
Theorem 1.16. If G is a matrix Lie group, then the map ξ ∈ g → Xξ ∈ l(G) is an
isomorphism of the Lie algebras, where the Lie bracket on g is as defined in Theorem 1.7.
We shall prove Theorem 1.16 in the next section.
1.4 Submanifolds
Suppose that M is a manifold, N ⊂ M , and N is itself a manifold, denote by ι : N → M
the inclusion map.
Definition. N is said to be an embedded submanifold8 of M if
(i) the map ι is smooth;
(ii) the differential (dι)p at any point of p ∈ N is an injective linear map;
(iii) ι is a homeomorphism onto its image, i.e. a D ⊆ N is open in the topology of
manifold N if and only if D is open in the topology induced on N from M (i.e. the open
subsets in N are precisely the intersections with N of the open subsets in M ).
Remark . Often the manifold N is not given as a subset of M but can be identified with a
subset of M by means of an injective map ψ : N → M . In this situation, the conditions in
the above definition make sense for ψ(N ) (regarded as a manifold diffeomorphic to N ). If
these conditions hold for ψ(N ) then one says that the map ψ embeds N in M and writes
ψ : N ,→ M .
Example. A basic example of embedded submanifold is a (parameterized) curve or surface
in R3 . Then the condition (i) means that the parameterization is smooth and (ii) means
that the parameterization is regular (cf. introductory remarks on p.1).
Remark . A map ι satisfying conditions (i) and (ii) is called an immersion and respectively
N is said to be an immersed submanifold. The condition (iii) eliminates e.g. the irrational
twist flow t ∈ R → [(t, αt)] ∈ R2 /Z2 = S 1 × S 1 , α ∈ R \ Q, on the torus.
A surface or curve in R3 (or more generally in Rn ) is often defined by an equation or
a system of equations, i.e. as the zero locus of a smooth map on R3 (respectively on Rn ).
E.g. x2 + y 2 − 1 = 0 (a circle) or (x2 + y 2 + b2 + z 2 − a2 )2 − 4b2 (x2 + y 2 ) = 0, b > a > 0
(a torus). However a smooth (even polynomial) map may in general have ‘bad’ points in
its zero locus (cf. Example Sheet 1, Q.8). When does a system of equations on a manifold
define a submanifold?
Definition. A value q ∈ N of a smooth map f between manifolds M and N is called a
regular value if for any p ∈ M such that f (p) = q the differential of f at p is surjective,
(df )p (Tp M ) = Tq N .
Theorem 1.17. Let f : M → N be a smooth map between manifolds and q ∈ N a regular
value of f . The inverse image of a regular value P = f −1 (q) = {p ∈ M : f (p) = q}
(if it is non-empty) is an embedded submanifold of M , of dimension dim M − dim N .
8
In these notes I will sometimes write ‘submanifold’ meaning ‘embedded submanifold’.
differential geometry 11
so we must have det(∂x0i0 /∂xj )(0) 6= 0 for the n × n Jacobian matrix, and the change from
xi ’s to x0i0 ’s is a diffeomorphism near 0.
Remarks. 1. It is not true that every submanifold of M is obtainable as the inverse image
of a regular value for some smooth map on M . One counterexample is RP 1 ,→ RP 2
(an exercise: check this!).
2. Sometimes in the literature one encounters a statement ‘a subset P is (or is not) a
submanifold of M ’. Every subset P of a manifold M has a topology induced from M . It
turns out that there is at most one smooth structure on the topological space P such that
P is an embedded submanifold of M , but I shall not prove it here.
12 alexei kovalev
Theorem 1.18 (Whitney embedding theorem). Every smooth n-dimensional manifold can
be embedded in R2n (i.e. is diffeomorphic to a submanifold of R2n ).
We shall assume the Whitney embedding theorem without proof here (note however,
Examples Sheet 1 Q.9). A proof of embedding in R2n+1 is e.g. in Guillemin and Pollack,
Ch.1 §9.
It is worth to remark that the possibility of embedding any manifold in some RN , with
N possibly very large, does not particularly simplify the study of manifolds in practical
terms (but is relatively easier to prove). The essence of the Whitney embedding theorem is
the minimum possible dimension of the ambient Euclidean space as way of measuring the
‘topological complexity’ of the manifold. The result is sharp, in that the dimension of the
ambient Euclidean space could not in general be lowered (as can be checked by considering
e.g. the Klein bottle).
We can now give, as promised, a proof of Theorem 1.16
Theorem 1.16. Suppose that G ⊂ GL(n, R) is a subgroup and an embedded submani-
fold of GL(n, R), and smooth structure on G is defined by the log-charts. Then the map
ξ ∈ g → Xξ ∈ l(G) is an isomorphism of the Lie algebras, where the Lie bracket on g is
the Lie bracket of matrices, as in Theorem 1.7.
Proof of Theorem 1.16. We want to show [Xξ , Xη ] = X[ξ,η] for a matrix Lie group where
the LHS is the Lie bracket of left-invariant vector fields and the RHS is defined using
Theorem 1.7. Note first that for G = GL(n, R), all the calculations can be done on an
2
open subset of Rn = Matr(n, R) with coordinates xij , i.j = 1. . . . , n. (The coordinate
chart on GL(n, R) given by the identity map is compatible, i.e. belongs to the same C ∞
structure, with the log-charts.) The map Lg , hence also dLg , is a linear map given by the
left multiplication by a fixed matrix g = (xij ). We obtain that the left-invariant vector
fields are Xξ (g) = xik ξjk ∂x∂ i . The coefficients of Xξ depend linearly on the coordinates xik
j
and [Xξ , Xη ] = X[ξ,η] follows by an elementary calculation.
In the above, we considered the coefficients of the left-invariant vector fields at I ∈
GL(n) using the basis of ‘coordinate vector fields’ corresponding to the identity chart.
On the other hand, the Lie bracket of matrices in Theorem 1.7 is defined in terms of the
basis corresponding to the log-chart. Recall from §1.2 that the log-chart around I is a
diffeomorphism
log : UI ⊂ GL(n) → V0 ⊂ Matr(n) (1.19)
between an open neighbourhood UI of I and an open neighbourhood V0 of the zero matrix.
For GL(n, R), the map (1.19) may also be considered as a change of local coordinates.
The derivative (the Jacobi matrix) of log A at A = I ∈ UI is the identity matrix and so
the bases of coordinate vector fields corresponding to the two choices of local coordinates
coincide at I.
Theorem 1.16 therefore holds for GL(n, R) (the argument also applies with some change
of notation to GL(n, C)).
Now consider a general case G ⊂ GL(n, R). Denote, as before, the inclusion map by ι.
Any left-invariant vector field Xξ on G (ξ ∈ g) may be identified by means of dι with a
differential geometry 13
vector field defined on a subset G ⊂ GL(n, R). Further, the left translation Lg on G is
the restriction of the left translation on GL(n, R) (if g ∈ G). We find that the vector
fields (dι)Xξ (ξ ∈ g) correspond bijectively to the restrictions to G ⊂ GL(n, R) of the
left-invariant vector fields Xξ on GL(n, R), such that ξ ∈ g. Let Xξ , Xη ∈ l(GL(n, R))
with ξ, η ∈ g, where g is understood as the image of log-chart for G near I. We have
[Xξ , Xη ] = X[ξ,η] , by the above calculations on GL(n, R), and the Lie bracket of matrices
[ξ, η] ∈ g by Theorem 1.7. Therefore, X[ξ,η] restricts to give a well-defined left-invariant
vector field on G, and [Xξ , Xη ] = X[ξ,η] holds for any Xξ , Xη ∈ l(G) as claimed.
Remark. Notice that the differential dι considered in the above proof identifies g with a
2
linear subspace of n × n matrices, by considering G as a hypersurface in GL(n, R) ⊂ Rn .
In the example of G = O(n), for any path A(t) or orthogonal matrices with A(0) = I, we
may compute 0 = dtd t=0 A(t)A∗ (t) = Ȧ(0)A∗ (0) + A(0)Ȧ∗ (0) = Ȧ(0) + Ȧ∗ (0), i.e. Ȧ(0) is
skew-symmetric. Thus the log chart at the identity actually maps onto a neighbourhood
of zero in the tangent space TI G —which explains why the log-chart construction in §1.2
worked.
Like (1.40 ) the eq.(1.20) is a priori merely a notation which resembles familiar results from
the calculus. See however further justification in the remark on the next page.
A disjoint union of all the cotangent spaces T ∗ M = tp∈M Tp∗ M of a given manifold
M is called the cotangent bundle of M . The cotangent bundle can be given a smooth
structure making it into a manifold of dimension 2 dim M by an argument very similar
to one for the tangent bundle (but with a change of notation, replacing any occurrence
of (1.40 ) with (1.20)).
14 alexei kovalev
A smooth field of linear functionals is called a (smooth) differential 1-form (or just
1-form). More precisely, a differential 1-form is a map α : M → T ∗ M such that αp ∈ Tp∗ M
for every p ∈ M and α is expressed in any local coordinates x = (x1 , . . . , xn ) by α =
P
i ai (x)dxi where ai (x) are some smooth functions of x.
Remark . The 1-forms are of course the dual objects to the vector fields. In particular,
ai (x(p)) is obtained as the value of αp on the tangent vector ( ∂x∂ i )p . Consequently, α is
smooth on M if and only if α(X) is a smooth function for every vector field X on M .
Notice that the latter condition does not use local coordinates.
One can similarly consider a space Λr Tp∗ M of alternating multilinear functions on Tp M ×
. . . × Tp M (r factors), for any r = 0, 1, 2, . . . n, and proceed to define the r-th exterior power
Λr T ∗ M of the cotangent bundle of M and the (smooth) differential r-forms on M .
Details are left as an exercise.
The space of all the smooth differential r-forms on M is denoted by Ωr (M ) and r is
referred to as the degree of a differential form. If r = 0 then Λ0 T ∗ M = M × R and
Ω0 (M ) = C ∞ (M ). The other extreme case r = dim M is more interesting.
Theorem 1.21 (Orientation of a manifold). Let M be an n-dimensional manifold. The
following are equivalent:
(a) there exists a nowhere vanishing smooth differential n-form on M ;
(b) there exists a family of charts in the differentiable structure on M such that the re-
spective coordinate domains cover M and the Jacobian matrices have positive determinants
on every overlap of the coordinate domains;
(c) the bundle of n-forms Λn T ∗ M is isomorphic to M × R.
Proof (gist). That (a)⇔(c) is proved similarly to the proof of Theorem 1.9.
Using linear algebra, we find that the transformation of the differential forms of top
degree under a change of coordinates is given by
∂xj 0
dx1 ∧ . . . ∧ dxn = det dx1 ∧ . . . ∧ dx0n .
∂x0i
Now (a)⇒(b) is easy to see.
To obtain, (b)⇒(a) we assume the following.
Theorem 1.22 (Partition of unity). For any open cover M ⊂ ∪α∈A Uα , there exists a
countable collection of functions ρi ∈ C ∞ (M ), i = 1, 2, ..., such that the following holds:
(i) for any i, the closure of supp(ρi ) = {x ∈ M : ρi (x) 6= 0} is compact and contained in
Uα for some α = αi (i.e. depending on i);
(ii) the collection is locally finite: each x ∈ M has a neighbourhood Wx such that ρi (x) 6= 0
on Wx for only finitely many i;P and
(iii) ρi ≥ 0 on M for all i and i ρi (x) = 1 for all x ∈ M .
The collection {ρi } satisfying the above is called a partition of unity subordinate to {Uα }.
Choose a partition of unity {ρi } subordinate to the given family of coordinate neigh-
(α)
bourhoods covering M . For each i, choose local coordinates xi valid on the support
differential geometry 15
(α) (α)
of ρi . Define ωα = dx1 ∧ . . . ∧ dxn in these local coordinates, then ρi ωαi is a well-
defined (smooth)
P n-form on all of M (extended by zero outside the coordinate domain)
and ω = i ρi ωαi is the required n-form.
A manifold M satisfying any of the conditions (a),(b),(c) of the above theorem is
called orientable. A choice of the differential form in (a), or family of charts in (b),
or diffeomorphism in (c) defines an orientation of M and a manifold endowed with an
orientation is said to be oriented.
Exterior derivative
Recall that the differential of a smooth map f : M → N between manifolds is a linear map
between respective tangent spaces. In the special case N = R, the (df )p at each p ∈ M is
a linear functional on Tp M , i.e. an element of the dual space Tp∗ M . In local coordinates
∂f
xi defined near p we have df (x) = ∂x i
(x)dxi , thus df is a well-defined differential 1-form,
whose coefficients are those of the gradient of f .
Remark . Observe that any local coordinate xi on an open domain U ⊂ M is a smooth
function on U . Then the formal symbols dxi actually make sense as the differentials of
these smooth functions (which justifies the previously introduced notation, cf.(1.20)).
Theorem 1.23 (exterior differentiation). There exists unique linear operator
k k+1
d : Ω (M ) → Ω (M ), k ≥ 0, such that
(i) if f ∈ Ω0 (M ) then df coincides with the differential of a smooth function f ;
(ii) d(ω ∧ η) = dω ∧ η + (−1)deg ω ω ∧ dη for any two differential forms ω,η;
(iii) ddω = 0 for every differential form ω.
Proof (gist). On an open set U ⊆ Rn , or in the local coordinates on a coordinate domain
on a manifold, application of conditions (ii), then (iii) and (i), yields
∂f
d(f dxi1 ∧ . . . ∧ dxir ) = df ∧ dxi1 ∧ . . . ∧ dxir = dxi ∧ dxi1 ∧ . . . ∧ dxir , (1.24)
∂xi
for any smooth function f . Extend this to arbitrary differential forms by linearity. The
conditions (i),(ii),(iii) then follow by direct calculation, in particular the last of these holds
by independence of the order of differentiation in second partial derivatives. This proves
the uniqueness, i.e. that if d exists then it must be expressed by (1.24) in local coordinates.
Observe another important consequence of (1.24): the operator d is necessarily local,
which means that the value (dω)p at a point p is determined by the values of differential
form ω on a neighbourhood of p.
To establish the existence of d one now needs to show that the defining formula (1.24)
is consistent, i.e. the result of calculation does not depend on the system of local coordi-
nates in which it is performed. So let d0 denote the exterior differentiation constructed as
in (1.24), but using different choice of local coordinates. Then, by (ii), we must have
r
X
d0 (f dxi1 ∧ . . . ∧ dxir ) = d0 f ∧ dxi1 ∧ . . . ∧ dxir + (−1)j+1 f dxi1 ∧ . . . d0 (dxij ) . . . ∧ dxir .
j=1
16 alexei kovalev
But d0f = df and d0 (dxk ) = d0 (d0 xk ) = 0, by (i) and (iii) and because we know that
the differential of a smooth function (0-form) is independent of the coordinates. Hence the
∂f
right-hand side of the above equality becomes ∂x j
dxj ∧dxi1 ∧. . .∧dxir = d(f xi1 ∧. . .∧dxir ),
and so the exterior differentiation is well-defined.
De Rham cohomology
A differential form α is said to be closed when dα = 0 and exact when α = dβ for some
differential form β. Thus exact forms are necessarily closed (but the converse is not in
general true, e.g. Example Sheet 2, Q4).
Definition. The quotient space
closed k-forms on M
k
HdR (M ) =
exact k-forms on M
is called the k-th de Rham cohomology group of the manifold M .
Any smooth map between manifolds, say f : M → N , induces a pull-back map
between exterior powers of cotangent spaces f ∗ : Λr Tf∗(p) N → Λr Tp∗ M (r = 0, 1, 2, . . .),
which is a linear map defined, for any differential r-form α on N , by
(f ∗ α)p (v1 , . . . , vr ) = αf (p) ((df )p v1 , . . . , (df )p vr ),
using the differential of f . The chain rule for differentials of smooth maps immediately
gives
(f ◦ g)∗ = g ∗ f ∗ . (1.25)
It is also straightforward to check that f ∗ preserves the ∧-product f ∗ (α∧β) = (f ∗ α)∧(f ∗ β)
and f ∗ commutes with the exterior differentiation, f ∗ (dα) = d(f ∗ α), hence f ∗ preserves the
subspaces of closed and exact differential forms. Therefore, every smooth map f : M → N
induces a linear map on the de Rham cohomology
f ∗ : H r (N ) → H r (M )
A consequence of the chain rule (1.25) is that if f is a diffeomorphism then f ∗ is a linear
isomorphism. Thus the de Rham cohomology is a diffeomorphism invariant, i.e. diffeo-
morphic manifolds have isomorphic de Rham cohomology.9
Poincaré lemma. H k (D) = 0 for any k > 0, where D denotes the open unit ball in Rn .
The proof goes by working out a way to invert the exterior derivative. More precisely,
one constructs linear maps hk : Ωk (U ) → Ωk−1 (U ) such that
hk+1 ◦ d + d ◦ hk = idΩk (U ) .
Remark . In the degree 0, one has H 0 (M ) = R for any connected manifold.
9
In fact, more is true. It can be shown, using topology, that the de Rham cohomology depends only on
the topological space underlying a smooth manifold and that homeomorphic manifolds have isomorphic de
Rham cohomology. The converse in not true: there are manifolds with isomorphic de Rham cohomology
but e.g. with different fundamental groups.
differential geometry 17
Note that the sum in the right-hand side may be assumed to have only finitelyR many
non-zero terms (by the compact support assumption). It can be checked that M ω does
not depend on the choice of the open cover {Uα } or a partition of unity on M , and is
therefore well-defined.
Stokes’ Theorem (for manifolds R without boundary)10 . Suppose that η ∈ Ωn−1 (M )
has a compact support. Then M dη = 0.
Let ρα be a partition of unity subordinate to some oriented system of coordinate neigh-
bourhoods covering M (as above). Then dη is a finite sum of the forms ρα η. Now the proof
of Stokes’ Theorem can be completed by considering compactly supported exact forms on
Rn and using calculus.
Corollary 1.26 (Integration by parts). Suppose that α and β are compactly supported
differential deg αR+ deg β = dim M − 1.
R forms on M and1+deg α
Then M α ∧ dβ = (−1) M
(dα) ∧ β.
One rather elegant application of the results discussed above is the following.
Proof of Theorem 1.27. Notation: recall that we write S n ⊂ Rn+1 for the unit sphere about
the origin. For r > 0, let ι(r) : x ∈ S n ,→ rx ∈ Rn+1 denote the embedding of S n in the
Euclidean space as the sphere of radius r about the origin and write S n (r) = ι(r)(S n )
(thus, in particular, S n (1) = S n ).
Suppose, for a contradiction, that X(x) is a nowhere-zero vector field on S n . We
may assume, without loss of generality, that |X| = 1, identically on S n . For any real
parameter ε, define a map
f : x ∈ Rn+1 \ {0} → x + ε|x|X(x/|x|) ∈ Rn+1 \ {0}.
Here we used the inclusion S n ⊂ Rn+1 (and hence Tp S n ⊂ Tp Rn+1 ∼
= Rn+1 ) to define a
(smooth) map x ∈ Rn+1 \ {0} → X(x/|x|) ∈ Rn+1 \ {0}.
Step 1.
We claim that f is a diffeomorphism, whenever |ε| is sufficiently small. Firstly, for any
x0 6= 0,
(df )x0 = idRn+1 +ε d(|x|X(x/|x|)) x0
and straightforward
calculus
shows that the norm of the linear map defined by the Jacobi
matrix d(|x|X(x/|x|)) x0 is bounded independent of x0 6= 0. Hence there is ε0 > 0, such
that (df )x0 is a linear isomorphism of Rn+1 onto itself for any x0 6= 0 and any |ε| < ε0 . But
then, by the Inverse Mapping Theorem (page 10), for any x 6= 0, f maps some open ball
B(x, δx ) of radius δx > 0 about x diffeomorphically onto its image.
Furthermore, it can be checked, by inspection of the proof of the Inverse Mapping The-
orem, that (1) δx can be taken to be continuous in x and (2) δx can be chosen independent
of ε if |ε| < ε0 . We shall assume these two latter claims without proof. Consequently, δx
can be taken to depend only on |x| (as S n (|x|) is compact).
Taking a smaller ε0 > 0√if necessary, we ensure that f is one-to-one if |ε| < ε0 . For the
latter, note that |f (x)| = 1 + ε2 |x| and so it suffices to check that that f is one-to-one
on each S n (|x|). But two points on S n (|x|) far away from each other cannot be mapped
to one because |f (x) − x| < ε|x| and two distinct points at a distance less than say 21 δ|x|
cannot be mapped to one because f restricts to a diffeomorphism (hence a bijection) on a
δ|x| -ball about each point.
A similar reasoning shows that f is surjective (onto) if ε in sufficiently small. Indeed,
f (B(x, δx )) is an open set homeomorphic to a ball and the boundary of f (B(x, δx )) is
within small distance ε(1 + δx )|x| from the boundary of B(x, δx ). Therefore, x must be
inside the boundary of f (B(x, δx )) and thus in the image of f . In all of the above, ‘small
ε0 ’ can be chosen independent of x because δx depends only on |x| and f is homogeneous
of degree 1, f (λx) = λf (x) for each positive λ.
Step 2. Now, as f is a diffeomorphism f maps the embedded submanifold S n (1)
√
diffeomorphically onto the embedded submanifold f (S n (1)) = S n ( 1 + ε2 ). Consider a
differential n-form on Rn+1
n
X
ω= (−1)i xi dx0 ∧ . . . (omit dxi ) . . . ∧ dxn ,
i=0
differential geometry 19
We have f (S n (1)) ω = S n (1) f ∗ ω (change of variable formula) and this integral depends
R R
where a constant cn+1 is n + 1 times the √ volume of (n + 1)-dimensional ball (the value
of cn+1 does not matter here). Put r = 1 + ε2 and then the right-hand side is not a
polynomial in ε if n + 1 is odd (i.e. when n is even). A contradiction.
2 Vector bundles.
Definition. Let B be a smooth manifold. A manifold E together with a smooth submer-
sion1 π : E → B, onto B, is called a vector bundle of rank k over B if the following
holds:
(i) there is a k-dimensional vector space V , called typical fibre of E, such that for any
point p ∈ B the fibre Ep = π −1 (p) of π over p is a vector space isomorphic to V ;
20
differential geometry 21
p ∗
more
generally, the bundle of differential p-forms Λ T M are real vector bundles of rank
n
p
with sections being the differential 1-forms, respectively p-forms. Exercise: verify that
the vector bundles Λp T ∗ M (1 ≤ p ≤ dim M ) will be trivial if T M is so.
2. ‘Tautological vector bundles’ may be defined over projective spaces RP n , CP n (and,
more generally, over the Grassmannians). Let B = CP n say. Then let E be the disjoint
union of complex lines through the origin in Cn+1 , with π assigning to a point in p ∈ E
the line ` containing that point, so π(p) = ` ∈ CP n . We shall take a closer look at one
example (Hopf bundle) below and show that the tautological construction indeed gives a
well-defined (and non-trivial) complex vector bundle of rank 1 over CP 1 .
Φβ ◦ Φ−1
α (b, v) = (b, ψβα (b)v),
(b, v) ∈ (Uβ ∩ Uα ) × Rk . For every fixed b the above composition is a linear isomorphism
of Rk depending smoothly on b. The maps ψβα : Uβ ∩ Uα → GL(k, R). are called the
transition functions of E.
It is not difficult to see that transition functions ψαβ satisfy the following relations,
called ‘cocycle conditions’
ψαα = idRk ,
ψαβ ψβα = idRk , (2.1)
ψαβ ψβγ ψγα = idRk .
The left-hand side is defined on the intersection Uα ∩ Uβ , for the second of the above
equalities, and on Uα ∩ Uβ ∩ Uγ for the third. (Sometimes the name ‘cocycle condition’
refers to just the last of the equalities (2.1); the first two may be viewed as notation.)
Now it may happen that a vector bundle π : E → B is endowed with a system of trivi-
alizing neighbourhoods Uα covering the base and such that all the corresponding transition
functions ψβα take values in a subgroup G ⊆ GL(k, R), ψβα (b) ∈ G for all b ∈ Uβ ∩ Uα , for
all α, β, where k is the rank of E. Then this latter system {(Uα , Φα )} of local trivializations
over Uα ’s is said to define a G-structure on vector bundle E.
Examples. 0. If G consists of just one element (the identity) then E has to be a trivial
bundle E = B × Rk .
1. Let G = GL+ (k, R) be the subgroup of matrices with positive determinant. If the
typical fibre Rk is considered as an oriented vector space then the transition functions ψβα
preserve the orientation. The vector bundle E is then said to be orientable.
A basic example arises from a system of coordinate charts giving an orientation of a
manifold M . The transition functions of T M are just the Jacobians and so M is orientable
precisely when its tangent bundle is so.
2. A more interesting situation occurs when G = O(k), the subgroup of all the non-
singular linear maps in GL(k, R) which preserve the Euclidean inner product on Rk . It
22 alexei kovalev
follows that the existence of an O(k) structure on a rank k vector bundle E is equivalent
to a well-defined positive-definite inner product on the fibres Ep . This inner product
is expressed in any trivialization over U ⊆ B as a symmetric positive-definite matrix
depending smoothly on a point in U .
Conversely, one can define a vector bundle with inner product by modifying the defi-
nition on page 20: replace every occurrence of ‘vector space’ by ‘inner product space’ and
(linear) ‘isomorphism’ by (linear) ‘isometry’. This will force all the transition functions to
take value in O(k) (why?).
A variation on the theme: an orientable vector bundle with an inner product is the
same as vector bundle with an SO(k)-structure.
3. Another variant of the above: one can play the same game with rank k complex
vector bundles and consider the U (k)-structures (U (k) ⊂ GL(k, C)). Equivalently, consider
complex vector bundles with Hermitian inner product ‘varying smoothly with the fibre’.
Furthermore, complex vector bundles themselves may be regarded as rank 2k real vector
bundles with a GL(k, C)-structure (the latter is usually called a complex structure on a
vector bundle).
Principal bundles.
Let G be a Lie group. A smooth free right action of G on a manifold P is a smooth
map P × G → P , (p, h) 7→ ph, such that (1) for any p ∈ P , ph = p if and only if h is the
identity element of G; and (2) (ph1 )h2 = p(h1 h2 ) for any p ∈ P , any h1 , h2 ∈ G. (It follows
that for each h ∈ G, P × {h} → P is a diffeomorphism.)
Φ
π −1 (U ) −−−
U
→ U ×G
πy
pr
y 1 (2.2)
U U
and ΦU commutes with the action of G, i.e. for each h ∈ G, ΦU (ph) = (b, gh), where
(b, g) = ΦU (p), π(p) = b ∈ U .
A local section of the principal bundle P is a smooth map s : U → P defined on a
neighbourhood U ⊂ B and such that π ◦ s = idU .
differential geometry 23
Φβ ◦ Φ−1
α (b, g) = (b, ψβα (b, g)),
where for each b ∈ Uα ∩ Uβ , the ψβα (b, ·) is a map G → G. Then, for each b in the domain
of ψβα , we must have ψβα (b, g)h = ψβα (b, gh) for all g, h ∈ G, in view of (2.2). It follows (by
taking g to be the unit element 1G ) that the map ψβα (b, ·) is just the multiplication on the
left by ψβα (b, 1G ) ∈ G. It is sensible to slightly simplify the notation and write g 7→ ψβα (b)g
for this left multiplication. We find that, just like vector bundles, the principal G-bundles
have transition functions ψβα : Uα ∩ Uβ → G between local trivializations. In particular,
ψβα for a principal bundle satisfy the same cocycle conditions (2.1).
A principal G-bundle over B may be obtained from a system of ψβα , corresponding
to an open cover of B and satisfying (2.1), — via the following ‘Steenrod construction’.
For each trivializing neighbourhood Uα ⊂ B for E consider Uα × G. Define an equivalence
relation between elements (b, h) ∈ Uα ×G, (b0 , h0 ) ∈ Uβ ×G, so that (b, h) ∼ (b0 , h0 ) precisely
if b0 = b and h0 = ψβα (b)h. Now let
P = tα Uα × G / ∼ (2.3)
the disjoint union of all Uα × G’s glued together according to the equivalence relation.
Theorem 2.4. P defined by (2.3) is a principal G-bundle.
Remark . The ψβα ’s can be taken from some vector bundle E over B, then P will be
‘constructed from E’. The construction can be reversed, so as to start from a principal
G-bundle P over a base manifold B and obtain the vector bundle E over B. Then E will
be automatically given a G-structure.
In either case the data of transition functions is the same for the principal G-bundle P
and the vector bundle P . The difference is in the action of the structure group G on the
typical fibre. G acts on itself by left translations in the case of the principal bundle and G
acts as a subgroup of GL(k, R) on Rk in the case of vector bundle E. 3 The vector bundle
E is then said to be associated to P via the action of G on Rk .
CP 1 = U1 ∪ U2 , Ui = {z1 : z2 ∈ CP 1 , zi 6= 0}, i = 1, 2,
3
G need not be explicitly a subgroup of GL(k, R), it suffices to have a representation of G on Rk .
24 alexei kovalev
with the local complex coordinate z = z2 /z1 on U1 , and ζ = z1 /z2 on U2 , and ζ = 1/z when
z 6= 0. We shall denote points in the total space E as (wz1 , wz2 ), with |z1 |2 + |z2 |2 6= 0,
w ∈ C, so as to present each point as a vector with coordinate w relative to a basis (z1 , z2 )
of a fibre. (This clarification is needed in the case when (wz1 , wz2 ) = (0, 0) ∈ C2 .) An
‘obvious’ local trivialization over Ui may be given, say over U1 , by (w, wz) ∈ π 1 (U1 ) →
(1 : z, w) ∈ U1 × C, —but in fact this is not a very good choice. Instead we define
p
Φ1 : (w, wz) ∈ π −1 (U1 ) → (1 : z, w 1 + |z|2 ) ∈ U1 × C,
p
Φ2 : (wζ, w) ∈ π −1 (U2 ) → (ζ : 1, w 1 + |ζ|2 ) ∈ U2 × C.
giving the transition function τ2,1 (1 : z) = (z/|z|), for 1 : z ∈ U1 ∩ U2 (i.e. z 6= 0). The
τ2,1 takes values in the unitary group U (1) = S 1 = {z ∈ C : |z| = 1}, a subgroup of
GL(1, C) = C \ {0} (it is for this reason the square root factor was useful in the local
trivialization). Theorem 2.4 now ensures that Hopf bundle E is a well-defined vector
bundle, moreover a vector bundle with a U (1)-structure. Hence there is a invariantly
defined notion of length of any vector in any fibre of E. The length of (wz1 , wz2 ) may
be calculated in the local trivializations Φ1 or Φ2 by taking the modulus of the second
component
p of Φi (in C). For each i = 1, 2, this coincides with the familiar Euclidean
length |wz1 |2 + |wz2 |2 of (wz1 , wz2 ) in C2 .
We can now use τ2,1 to construct the principal S 1 -bundle (i.e. U (1)-bundle) P → CP 1
associated to Hopf vector bundle E, cf. Theorem 2.4. If U (1) is identified with a unit
circle S 1 ⊂ C , any fibre of P may be considered as the unit circle in the respective
fibre of E. Thus P is identified as the space of all vectors in E of length 1, so P =
{(w1 , w2 ) ∈ C2 | w1 w̄1 +w2 w̄2 = 1} is the 3-dimensional sphere and the bundle projection is
π : (w1 , w2 ) ∈ S 3 → w1 : w2 ∈ S 2 , π −1 (p) ∼
= S 1,
where we used the diffeomorphism S 2 ∼ = CP 1 for the target space. (Examples 1, Q3(ii).)
1 3 2
This principal S -bundle S over S is also called Hopf bundle. It is certainly not trivial,
as S 3 is not diffeomorphic to S 2 × S 1 . (The latter claim is not difficult to verify, e.g.
by showing that the de Rham cohomology H 1 (S 3 ) is trivial, whereas H 1 (S 2 × S 1 ) is not.
Cf. Examples 2 Q5.)
differential geometry 25
f
B −−−→ B 0
is a commutative diagram, π 0 ◦ F = f ◦ π.
An isomorphism from a vector bundle E to itself covering the identity map idB is
called a bundle automorphism of E. The set Aut E of all the bundle automorphisms
of E forms a group (by composition of maps). If E = B × V is a trivial bundle then any
automorphism of E is defined by a smooth maps B → GL(V ), so Aut E = C ∞ (B, GL(V )).
If a vector bundle E has a G-structure (G ⊆ GL(V )) then it is natural to consider the
group of G-bundle automorphisms of E, denoted AutG E and defined as follows. Recall
that a G-structure means that there is a system of local trivializations over neighbourhoods
covering the base B and with the transition functions of E taking values in G. Now a bundle
automorphism F ∈ Aut E of E → B is determined in any local trivialization over open
differential geometry 27
4
...and informally the ‘group of gauge transformations’ is often abbreviated as the ‘gauge group’ of E,
although the ‘gauge group’ is really a different object! (It is the structure group G of vector bundle.) Alas,
there is a danger of confusion.
28 alexei kovalev
2.2 Connections.
Sections of a vector bundle generalize vector-valued functions on open domains in Rn . Is
there a suitable version of derivative for sections, corresponding to the differential in multi-
variate calculus? In order to propose such a derivative, it is necessary at least to understand
which sections are to have zero derivative, corresponding to the constant functions on Rn .
(Note that a section which is expressed as a constant in one local trivialization need not
be constant in another.)
Definition. The vector space Ker (dπ)p is called the vertical subspace of Tp E, denoted
T vp E. A subspace Sp of Tp E is called a horizontal subspace if Sp ∩ T vp E = {0} and
Sp ⊕ T vp E = Tp E.
Thus any horizontal subspace at p is isomorphic to the quotient Tp E/T vp E and has
the dimension dim(Tp E) − dim(T vp E) = dim B. Notice that, unlike the vertical tangent
space, a horizontal space can be chosen in many different ways (e.g. because there are
many choices of local trivialization near a given point in B).
It is convenient to specify a choice of a horizontal subspace at every point of E as the
kernel of a system of differential 1-forms on E, using the following
Fact from linear algebra: if θ1 , . . . , θm ∈ (Rn+m )∗ are linear functionals then one
will have dim(∩m i 1 m
i=1 Ker θ ) = n if and only if θ , . . . , θ are linearly independent
in (Rn+m )∗ .
Now let θp1 , . . . , θpm be linearly independent ‘covectors’ in Tp∗ E, p ∈ π −1 (U ), and define
and any tangent vector in Tp E as v = B k ( ∂x∂ k )p + C i ( ∂a∂ i )p , B k , C i ∈ R. The θpj cannot all
vanish on a vertical vector, i.e. on a vector having B k = 0 for all k. That is,
Therefore the m × m matrix g = (gji ) must be invertible. Denote the inverse matrix by
c = g −1 , c = (cil ). Replace θi by θ̃i = cil θl = dai + eik dxk , this does not change the space Sp .
The above arrangement can be made for every p ∈ π −1 (U ), with fki (p) and gji (p)
in (2.7) becoming functions of p. Call a map p 7→ Sp a field of horizontal subspaces if
the functions fki (p) and gji (p) are smooth. To summarize,
and so
As we shall see below, the linearity condition in ai ensures that the horizontal sections
(which are to become the analogues of constant vector-functions) form a linear subspace of
the vector space of all sections of E. (A very reasonable thing to ask for.)
I will sometimes use an abbreviated notation
0 0
to U 0 are written as ψii and from U 0 to U as ψii0 , thus the matrix (ψii0 ) is inverse to (ψii ).
0 0
Likewise, (∂xk /∂xk ) denotes the inverse matrix to (∂xk /∂xk ).
Recall that on U 0 ∩ U we have
0 0 0 0
xk = xk (x), ai = ψii (x)ai . (2.11)
Then
0
∂xk k
k0 0 0 0
dx = dx , dai = dψii ai + ψii dai
∂xk
0
∂ψii i k 0
= k
a dx + ψii dai ,
∂x
so 0
i0 i0 0 0 0 0 0 ∂xk i0 j 0 k
θ = da + Γij 0 k0 aj dxk = dψii ai + ψii dai + Γ 0 0 a dx ,
∂xk j k
and 0 0
0 ∂ψji ∂xk i0 j0 j
ψii0 θi
= da + · k
+ ψi
i
i
0 ·
k
(ψii0
· Γj
k
0 k 0 · ψj )a dx .
∂x ∂x
i i0 i
But then ψi0 θ = θ and we find, by comparing with (2.10b), that
0
0 ∂x
k0 ∂ψji
0
Γijk = Γij 0 k0 ψii0 ψjj k
+ ψii0 (2.12a)
∂x ∂xk
and, using Aij = Γijk dxk ,
0 0 0
Aij 0 = ψii Aij ψjj0 + ψii dψji 0 , (2.12b)
0
Writing AΦ and AΦ for the matrix-valued 1-forms expressing the connection A in the local
trivializations respectively Φ and Φ0 and abbreviating (2.11) to Φ0 = ψΦ for the transition
function ψ we obtain from the above that
θ will determine a connection on P if and only if θ is (1) horizontal and (2) G-equivariant
with respect to the smooth free right action on P .
We may determine a choice of a field of horizontal subspaces by using the kernels of local
matrices of 1-forms θp = (θji )p defined around each p ∈ P , where (θji )p ∈ Ω1 (π −1 (U )) for
i, j = 1, . . . , m and U ⊂ B a trivializing neighbourhood (and a coordinate neighbourhood)
U ⊂ B. Denote by cij the ‘obvious’ coordinates on the space Matr(m, R) and require that
(θji )p = dcij + Γisk (x)csj dxk for some functions Γisk ∈ C ∞ (U ) satisfying the transformation
law (2.12). In a more abbreviated notation, θ = dC + AC, where C = (cij ) and A = (Aij ) =
(Γijk dxk ). One can verify, noting (2.12) and C 0 = ψC (cf. (2.11) and page 23), that the
forms C −1 θ = C −1 dC + C −1 AC agree on the intersections π −1 (U ) ∩ π −1 (U 0 ) and patch
together to define a global 1-form θ on P . Cf. Sheet 3 Q8.
Note the construction of θ does not extend to the vector bundles.
Covariant derivatives
Definition . A covariant derivative on a vector bundle E is a R-linear operator
∇E : Γ(E) → Γ(T ∗ B ⊗ E) satisfying a ‘Leibniz rule’
∇E (f s) = df ⊗ s + f ∇E s (2.14)
dA s = ds + As, s ∈ Γ(E).
More explicitly, one can write s = (s1 , . . . , sm ) with the help of a local trivialization, where
sj are smooth functions on the trivializing neighbourhood, and then
∂s1 ∂sm
dA (s1 , . . . , sm ) = ( 1 j k m j k
+ Γ jk s )dx , . . . , ( + Γ jk s )dx .
∂xk ∂xk
The operator dA is well-defined as making a transition to another trivialization we have
s = ψs0 and A = ψA0 ψ −1 − (dψ)ψ −1 , which yields the correct transformation law for
dA s = ds + As = d(ψs0 ) + (ψA0 ψ −1 − (dψ)ψ −1 )s0 = ψ(ds0 + A0 s0 ) = ψ(dA s)0 .
Theorem 2.15. Any covariant derivative ∇E arises as dA from some connection A.
Proof (gist). Firstly, any covariant derivative ∇E is a local operation, which means that
is s1 , s2 are two sections which are equal over an open neighbourhood U of b ∈ B then
(∇E s1 )|b = (∇E s2 )|b . Indeed, let U0 be a smaller neighbourhood of b with the closure
U0 ⊂ U and consider a cut-off function α ∈ C ∞ (B), so that 0 ≤ α ≤ 1, α|U0 = 1, α|B\U = 0.
Then 0 = d(α(s1 − s2 )) = (s1 − s2 ) ⊗ dα + α∇E (s1 − s2 ), whence (∇E s1 )|b = (∇E s2 )|b as
required. So, it suffices to consider ∇E in some local trivialization of E.
differential geometry 33
The proof now simply produces the coefficients Γijk of the desired A in an arbitrary
local trivialization of E, over U say. Any local section of E defined over U may be written,
with respect to the local trivialization, as a vector valued function U → Rm (m being
the rank of E). Let ej , j = 1, . . . m denote the sections corresponding in this way to the
constant vector-valued functions on U equal to the j-th standard basis vector of Rm . Then
the coefficients of the connection A are (uniquely) determined by the formula
∂ i
Γijk = (∇E ej )( ) , (2.16)
∂xk
where ∇E (ej ) is a vector of differential 1-forms which takes the argument a vector field. We
used a local coordinate vector field ∂x∂ k (which is well-defined provided that a trivializing
neighbourhood U for E is also a coordinate neighbourhood for B) and obtained a local
section of E, expressed as a smooth map U → V . Then Γijk ∈ C ∞ (U ) is the i-th component
of this map in the basis ej of V .
It follows, from the R-linearity and Leibniz rule for ∇E , that for an arbitrary local
section we must have
∇E s = ∇E (sj ej ) = (dsi + sj Γijk dxk )ei = dA s
where as usual A = (Aij ) = (Γijk dxk ), so we recover the dA defined above. It remains to ver-
ify that Γijk ’s actually transform according to (2.12) in any change of local trivialization, so
we get a well-defined connection. The latter calculation is straightforward (and practically
equivalent to verifying that dA is well-defined independent of local trivialization.)
The definition of covariant derivative further extends to differential forms with values
in E by requiring the Leibniz rule, as follows
dA (σ ∧ ω) = (dA σ) ∧ ω + (−1)q σ ∧ (dω), σ ∈ ΩqB (E), ω ∈ Ωr (B),
with ∧ above extended in an obvious way to multiply vector-valued differential forms and
usual differential forms. It is straightforward to verify, considering local bases of sections
ej and differential forms dxk , that in any local trivialization one has dA σ = dσ + A ∧ σ,
where σ ∈ ΩqB (E).
2.3 Curvature.
Let A be a connection on vector bundle E and consider the repeated covariant differentia-
tion of an arbitrary section (or r-form) s ∈ ΩrB (E) (assume r = 0 though). Calculation in
a local trivialization gives
dA dA s = d(ds+As)+A∧(ds+As) = (dA)s−A∧(ds)+A∧(ds)+A∧(As) = (dA+A∧A)s.
Thus dA dA is a linear algebraic operator, i.e. unlike the differential operator but dA , the
dA dA commutes with the multiplication by smooth functions.
dA dA (f s) = f dA dA s, for any f ∈ C ∞ (B). (2.17)
Notice that the formula (2.17) does not make explicit reference to any local trivialization.
We find that (dA dA s) at any point b ∈ B is determined by the value s(b) at that point. It
follows that the operator dA dA is a multiplication by an endomorphism-valued differential
2-form. (This 2-form can be recovered explicitly in coordinates similarly to (2.16), using
a basis ei say of local sections and a basis of differential 1-forms dxk in local coordinates
on B.)
Definition. The form
F (A) = dA + A ∧ A ∈ Ω2 (B; End E).
is called the curvature form of a connection A.
Definition. A connection A is said to be flat is its curvature form vanishes F (A) = 0.
Example. Consider a trivial bundle B × Rm , so the space of sections is just the vector-
functions C ∞ (B; Rm ). Then exterior derivative applied to each component of a vector-
function is a well-defined linear operator satisfying Leibniz rule (2.14). The corresponding
connection is called trivial, or product connection. It is clearly a flat connection.
The converse is only true with an additional topological condition that the base B is
simply-connected; then any flat connection on E induces a (global) trivialization E ∼ =B×
Rm (Examples 3, Q7(ii)) and will be a product connection with respect to this trivialisation.
differential geometry 35
Bianchi identity.
Covariant derivative on a vector bundle E with respect to a connection A can be extended,
in a natural way, to any section of B of End E by requiring the following formula to hold,
for every section s of E. Notice that this is just a suitable form of Leibniz rule.
The definition further extends to differential forms with values End E, by setting for
every µ ∈ Ωp (B; End E) and σ ∈ Ωq (B; E),
dA µ = dµ + A ∧ µ − µ ∧ A.
Now, for any section s of E, we can write dA (dA dA )s = (dA dA )dA s, i.e. dA (F (A)s) =
F (A)dA s and comparing with the Leibniz rule above we obtain.
Proposition 2.18 (Bianchi identity). Every connection A satisfies dA F (A) = 0.
for any two sections s1 , s2 of E, where h·, ·i denotes the inner product on the fibres of E.
36 alexei kovalev
Proof.
0 = hdA (si1 ei ), sj2 ej i + hsi1 ei , dA (sj2 ej )i − dhsi1 ei , sj2 ej i = (Aij + Aji )si1 sj2 for any si1 , sj2 ,
Corollary 2.20. The curvature form F (A) of an orthogonal (resp. unitary) connection A
is skew-symmetric (resp. skew-Hermitian) in any orthogonal (unitary) trivialization.
Proof. It suffices to show that there exists a well-defined covariant derivative ∇E on sections
of E. We shall construct a example of ∇E using a partition of unity.
Let Wα be an open covering of B by trivializing neighbourhoods for E and Φα the
corresponding local trivializations. Then on each restriction E|Wα we may consider a
trivial product connection d(α) defined using Φα . Of course, the expression d(α) s will only
make sense over all of B if a section s ∈ Γ(E) is equal to zero away from Wα . Now consider
a partition of unity ρi subordinate to Wα . The expressions ρi s, ρi d(i) s make sense over all
of B as we may extend by zero away from Wi . Now define
∞
X ∞
X
E
∇ s := d(i) (ρi s) = ρi d(i) s, (2.22)
i=1 i=1
3 Riemannian geometry
3.1 Riemannian metrics and the Levi–Civita connection
Let M be a smooth manifold.
Definition. A bilinear symmetric positive-definite form
gp : Tp M × Tp M → R
The bilinear form (metric) g will be smooth if and only if the local coefficients gij = gij (x)
are smooth functions of local coordinates xi on each coordinate neighbourhood.
Example 3.1. Recall (from Chapter 1) that any smooth regularly parameterized surface S
in R3 ,
r : (u, v) ∈ U ⊂ R2 → r(u, v) ∈ R3 .
is a 2-dimensional manifold (more precisely, we assume here that S satisfies all the defining
conditions of an embedded submanifold). The first fundamental form1 Edu2 + 2F dudv +
Gdv 2 is a Riemannian metric on S.
The following formulae are proved in multivariate calculus.
• A curve on S may be given as γ(t) = r(u(t), v(t)), a ≤ t ≤ b. The length of γ is then
Rb Rb√
computed as a |γ̇(t)|dt = a E u̇2 + 2F u̇v̇ + Gv̇ 2 dt.
RR √
• The area of S is U EG − F 2 du dv.
1
E = (ru , ru ), F = (ru , rv ), G = (rv , rv ) using the Euclidean inner product
37
38 alexei kovalev
Proof. Uniqueness. The conditions (1) and (2) determine the coefficients of Levi–Civita in
local coordinates as follows. A ‘coordinate vector field’ ∂x∂ i with constant coefficients has
covariant derivative D ∂x∂ i = Γpik ∂x∂ p dxk . The condition (1) with X = ∂x∂ i , Y = ∂x∂ j , Z = ∂x∂ k
becomes
∂
gij = Γpik gpj + Γpjk gip . (3.7a)
∂xk
Cycling i, j, k in the above formula, one can write two more relations
∂
j
gki = Γpkj gpi + Γpij gkp , (3.7b)
∂x
∂
i
gjk = Γpji gpk + Γpki gjp . (3.7c)
∂x
Let (g iq ) denote the inverse matrix to (giq ), so Γpjk gqp g iq = Γijk . Adding the first two
equations of (3.7) and subtracting the third, dividing by 2, and multiplying both sides of
the resulting equation by (g iq ), one obtains the formula
i 1 iq ∂gqj ∂gkq ∂gjk
Γjk = g + − (3.8)
2 ∂xk ∂xj ∂xq
(also taking account of the symmetry condition (2)). Thus if the Levi–Civita connection
exists then its coefficients in local coordinates are expressed in terms of the metric by (3.8).
Exericise. By adapting the above method to arbitrary vector fields X, Y, Z on M , using
the symmetry condition (2) in the form DX Y − DY X = [X, Y ], show that the Levi–Civita
connection is uniquely determined by the identity
1
g(DX Y, Z) = Xg(Y, Z)+Y g(Z, X)−Zg(X, Y )−g(Y, [X, Z])−g(Z, [Y, X])+g(X, [Z, Y ]) .
2
(3.9)
Verify that D defined by (3.9) satisfies the conditions (1) and (2) in Theorem 3.6 (this
might be argued by essentially following your calculation of (3.9) backwards).
40 alexei kovalev
Existence. Proof 1. One way of proving the existence is to check that the Γijk computed
by the formula (3.8) are indeed the coefficients of a well-defined connection on M . This
can be done by verifying that the Γijk ’s transform in the right way, i.e. as in (3.3), under
∂xi ∂xj
a change of local coordinates. The transformation law for gij is gi0 j 0 = 0 g ij , by the
∂xi ∂xj 0
usual linear algebra. Differentiating this latter formula and using the respective formula
for the induced inner product on the dual spaces, i.e. on the cotangent spaces to M , we
can verify that the coefficients given by (3.8) indeed transform according to (3.3) and so
the Levi–Civita connection of the metric g on M is well-defined.
Proof 2. Alternatively, assuming the exercise above, we shall be done if we show that
Dx Y defined by the formula (3.9) is C ∞ (M )-linear in X and satisfies the Leibniz rule in Y .
For the first property, we note that [f X, Z] = f XZ − Z(f X) = f [X, Z] − (Zf )X, for every
f ∈ C ∞ (M ). Then 2g(Df X Y, Z) becomes
f Xg(Y, Z) + Y g(Z, f X) − Zg(f X, Y ) − g(Y, [f X, Z]) − g(Z, [Y, f X]) + g(f X, [Z, Y ])
= f Xg(Y, Z) + (Y f )g(Z, X) + f Y g(Z, X) − (Zf )g(X, Y ) − f Zg(X, Y )
−g(Y, f [X, Z] − (Zf )X) − g(Z, (Y f )X + f [Y, X]) + f g(X, [Z, Y ]),
using the Leibniz rule for vector fields. It follows that g(Df X Y, Z) = g(f DX Y, Z), thus D
is C ∞ (M )-linear in X.
For the Leibniz rule we calculate, with h a smooth function,
2g(DX (hY ), Z) = X(hg(Y, Z)) + hY g(Z, X)
− Z(hg(X, Y )) − hg(Y, [X, Z]) − g(Z, [hY, X]) + g(X, [Z, hY ])
= (Xh)g(Y, Z) + hXg(Y, Z) + hY g(Z, X) − (Zh)g(X, Y ) − hZg(X, Y )
−hg(Y, [X, Z]) − hg(Z, [Y, X]) + (Xh)g(Z, Y ) + hg(X, [Z, Y ]) + (Zh)g(X, Y )
= 2(Xh)g(Y, Z) + 2hg(DX Y, Z)
which gives (3.4) as required. Since DX Y is clearly R-linear in Y we have proved that D
is a connection on M .
a non-linear second-order ordinary differential equation for a path x(t) = (xi (t)) (here
i, j, k = 1, . . . , dim M ). By the basic existence and uniqueness theorem from the theory
of ordinary differential equations, it follows that for any choice of the initial conditions
x(0) = p, ẋ(0) = a there is a unique solution path x(t) defined for |t| < ε for some
positive ε. Thus for any p ∈ M and a ∈ Tp M there is a uniquely determined (at least
for any small |t|) geodesic with this initial data (i.e. ‘coming out of p in the direction a’).
Denote this geodesic by γp (t, a) (or γ(t, a) if this is not likely to cause confusion).
Proposition 3.12. If γ(t) is a geodesic on (M, g) then |γ̇(t)|g = const.
Proof. We shall first make a rigorous sense of the equation
Dγ̇ γ̇ = 0 (3.13)
and show that (3.13) is satisfied at each γ(t) if and only if γ is a geodesic curve. The
problem with (3.13) at the moment is that γ̇ is not a vector field defined on any open set
in M , but only along a curve γ. We define an extension, still denoted by γ̇, on a coordinate
neighbourhood U of γ(0) as follows. It may be assumed, without loss, that γ̇(0) = (ẋi (0))
has ẋ1 (0) 6= 0. We may further assume, taking a smaller U if necessary, that γ ∩ U , is
a graph of a smooth function x1 7→ (x2 (x1 ), . . . , xn (x1 )). In particular, ẋ1 (t) 6= 0 for any
small |t| and also any hyperplane x1 = x10 , such that |x10 − x1 (γ(0))| is small, meets the
curve γ ∩ U in exactly one point. Denote by π the projection along hyperplanes x1 = const
onto γ ∩ U . Define, for every p ∈ U , γ̇(p) = γ̇(π(p)) and then γ̇ is a smooth vector field
on U , such that (γ̇)p = γ̇(t) whenever p = γ(t).
Now let Γijk be the coefficients of the Levi–Civita in the coordinates on U . So DZ Y =
(Z l ∂l Y i + Γijk Y j Z k )∂i for any vector fields Z = Z l ∂l , Y = Y i ∂i on U . Let Y = Z = γ̇.
∂ ẋi
Then at any point p = γ(t) we have Z l ∂l Y i = ẋl l = ẍi by the chain rule. It follows
∂x
that the equation (3.11) is equivalent to (3.13) if the latter if restricted to the points of
the curve γ. It can also be seen, by inspection of the above construction, that Dγ̇ γ̇ at the
points of γ is independent of the choice of extension of γ̇(t) to a vector field on U .
We have γ̇(γ̇, γ̇)g = (D γ̇ γ̇, γ̇)g + (γ̇, Dγ̇ γ̇)g on U from the defining properties of the
2
Levi–Civita. Hence γ̇ |γ̇|g = 0 at each γ(t) ∈ U , by (3.13). From the construction of the
extension γ̇ on U , we find that the directional partial derivative γ̇ |γ̇|2g at the points γ(t)
∂ d
is expressed as ẋl l |γ̇|2g = |γ̇(t)|2g by the chain rule again, whence |γ̇|g = const as we
∂x dt
had to prove.
42 alexei kovalev
Examples. 1. On Rn with the Euclidean metric (dxi )2 we have Γiik = 0, so the Levi–
P
Civita is just the exterior derivative D = d. The geodesics ẍi = 0 are straight lines
γp (t, a) = p + at parameterized with constant velocity.
2. Consider the sphere S n with the round metric (i.e. the restriction of the Euclidean
metric to S n ⊂ Rn+1 ). Then p ∈ S n and a ∈ Tp S n may be regarded as the vectors
in Rn+1 . Suppose a 6= 0, then the orthogonal reflection L in the 2-dimensional subspace
P = span{p, p + a} is an isometry of S n . Now L preserves the metric and p and a, the data
which determines the geodesic γp (·, a). As γp (·, a) is moreover uniquely determined it must
be contained in the fixed point set of L. But the fixed point set is a curve, the great circle
P ∩ S n . We find that great circles, parameterized with velocity of constant length—and
only these—are the geodesics on S n .
Observe that for any geodesic γp (t, a) and any real constant λ the path γp (λt, a) is also
a geodesic and γp (λt, a) = γp (t, λa).
By application of a general result in the theory of ordinary differential equations, a
geodesic γp (t, a) must depend smoothly on its initial conditions p, a. Furthermore, there
exist ε1 > 0 and ε2 > 0 independent of a and such that if |a| < ε1 then γp (t, a) exists for
all −2ε2 < t < 2ε2 . It follows that γp (1, a) is defined whenever |a| < ε = ε1 ε2 .
Corollary 3.15. The exponential map expm defines a diffeomorphism from a neighbour-
hood of zero in Tm M to a neighbourhood of m in M .
Here ]0, ε[×S n−1 is regarded as a subset in Tp M ∼ = Rn via the inner product g(p). If
0 < r < ε then the image Σr = f ({r} × S n−1 ⊂ Tp M ) of the metric sphere of radius r
is well-defined on M and is called a geodesic sphere about p. (So Σr is an embedded
submanifold of M .) The following remarkable result asserts that ‘the geodesic spheres are
orthogonal to their radii’.
Gauss Lemma. The geodesic γp (t, a) is orthogonal to Σr . Thus the metric g in geodesic
polar coordinates has local expression g = dr2 + h(r, v), where for any 0 < r < ε, h(r, v) is
the metric on Σr induced by restriction of g.
Proof. Let X be an arbitrary smooth vector field on the unit sphere S n−1 ⊂ Tp M . Use
polar coordinates to make sense of X as a vector field (independent of r) on the punctured
unit ball B \ {0} ⊂ Tp M . Define a vector field X̃(r, v) = rX(v) on B \ {0}. The map
expp induces a vector field Y (f (r, v)) = (d expp )rv X̃(r, v) on the punctured geodesic ball
B 0 \ {p} = expp (B \ {0}) in M .
We shall be done if we show that Y is everywhere orthogonal to the radial vector
∂
field ∂r . Note that, by construction, any geodesic from p is given in normal coordinates
∂
by γp (t, a) = at, so γ̇p (t, a)/|a| = ∂r . Here |a| means the norm in the inner product gp
on the vector space Tp M . By application of Corollary 3.15, the family γ̇p (t, a), where
|a| = 1 and 0 < |t| < ε, defines a smooth vector field on B 0 \ {p}. Recall from (3.13) that
Dγ̇ γ̇ = 0 for any geodesic γ, where D denotes the Levi–Civita covariant derivative. Also
d ∂ ∂
dt
g( ∂r , ∂r ) = dtd g(γ̇, γ̇) = 0 by Proposition 3.12, so g( ∂r
∂ ∂
, ∂r ) = 1. It remains to show that
g(Y, γ̇) = 0.
Using the diffeomorphism f in (3.16) to go to polar geodesic coordinates, we obtain
∂
d
Dγ̇ Y − DY γ̇ = (df ) D ∂ X̃ − DX̃ ∂r = (df ) X̃ = (df )(X̃/r) = Y /r,
∂r dr
with the help of Proposition 3.5. Therefore, we find
d 1 1
g(Y, γ̇) = g(Dγ̇ Y, γ̇) + g(Y, Dγ̇ γ̇) = g(Dγ̇ Y, γ̇) = g(DY γ̇ + Y, γ̇) = g(Y, γ̇).
dr r r
d
as 2g(Dγ̇, γ̇) = d g(γ̇, γ̇) = 0 by Proposition 3.13. Thus dr G = G/r, where G = g(Y, γ̇).
d d ∂
Hence G is linear in r and dr G independent of r. But limr→0 dr G = limr→0 g(X, ∂r ) = 0, as
(d expp )0 is an isometry by Proposition 3.14, and so g(Y, γ̇) = 0 and the result follows.
∂ 2 gkl ∂ ∂ ∂ ∂ ∂ ∂ ∂ ∂
j i
= g(Dj Di k , l ) + g(Di k , Dj l ) + g(Dj k , Di l ) + g( k , Dj Di l ).
∂x ∂x ∂x ∂x ∂x ∂x ∂x ∂x ∂x ∂x
2
∂ gkl ∂ gkl2
The right-hand side of the above expression is symmetric in i, j as ∂x i ∂xj = ∂xj ∂xi . The
anti-symmetric part of the right-hand side (which has to be zero) equals Rij,kl + Rji,kl .
(ii) Firstly, (Dk ∂x∂ j )i = Γijk = (Dj ∂x∂ k )i , by the symmetric property of the Levi–Civita.
The claim now follows by straightforward computation using (3.17).
We note for use in the proof of (iii) that multiplying (ii) by giq gives Rij,kl + Rik,lj +
Ril,jk = 0.
(iii) We organize the argument using the vertices and faces of an octahedron (see the
next page).
2
also known as the algebraic Bianchi identity, not to be confused with the differential Bianchi identity
in Chapter 2.
differential geometry 45
Remark . Notice that the proof of (ii) shows the first Bianchi identity is valid for every
symmetric connection on M .
Corollary 3.20. The Riemann curvature tensor (Rij,kl )p defines, at any point p ∈ M a
symmetric bilinear form on the fibres of Λ2 Tp M .
There are natural ways to extract “simpler” quantities (i.e. with less components) from
the Riemann curvature tensor.
If local coordinates
P are chosen soPthat gij (p) = δij at a point, then the latter definition
means that s(p) = Ricii (p) = i,j Rij,ji (p). For a general gij , the formula may be
P ij i ij
written as s = i g Ricij , where g is the induced inner product on the cotangent space
with respect to the dual basis, algebraically (g ij ) is the inverse matrix of (gij ).
3
I learned this argument from the lectures of M.M. Postnikov.
46 alexei kovalev
Ric = λg (3.21)
for some constant λ ∈ R, as both the metric and its Ricci curvature are symmetric bilinear
forms on the tangent spaces to M . When the condition (3.21) is satisfied, the Riemannian
manifold (M, g) is called Einstein manifold. In particular, if (3.21) holds with λ = 0 then
M is said to be Ricci-flat.
(2) Recall that if Σ is a surface in R3 (smooth, regularly parameterized by (u, v) in an
open set in R2 ) then there is a metric induced on Σ, expressed as the first fundamental
form Edu2 + 2F dudv + Gdv 2 . The second fundamental form Ldu2 + 2M dudv + N dv 2 is
defined by taking the inner products L = (ruu , n), M = (ruv , n), N = (rvv , n) with the
unit normal vector to Σ, n = ru × rv /|ru × rv | (the subscripts u and v denote respective
partial derivatives). The quantity
LN − M 2
K=
EG − F 2
is called the gaussian curvature of Σ. A celebrated theorema egregium, proved by Gauss,
asserts that K is determined by the coefficients of first fundamental form, i.e. by the metric
on Σ (and so K is independent of the choice of an isometric embedding of Σ in R3 ).
Taking up a general view on Σ as a 2-dimensional Riemannian manifold, one can check
that 2(EG − F 2 )−1 R12,21 = s, the scalar curvature of Σ. From the results of the next
section, we shall see (among other things) that the scalar curvature of a surface Σ is twice
its gaussian curvature s = 2K.
D̃X Y = DX Y + II(X, Y ),
∆ = δd + dδ.
Ωp (M ) = ∆Ωp (M ) ⊕ Hp
= dδΩp (M ) ⊕ δdΩp (M ) ⊕ Hp
= dΩp−1 (M ) ⊕ δΩp+1 (M ) ⊕ Hp
Short summary of the proof. We need to introduce the concept of a weak solution of
∆ω = α. (3.30)
A weak solution of (3.30) is by definition, a linear functional l : Ωp (M ) → R which is
(i) bounded, |l(β)| ≤ Ckβk, for some C > 0 independent of β, and
(ii) satisfies l(∆ϕ) = hα, ϕiL2 .
Any solution ω of (3.30) defines a weak solution by putting lω (β) = hω, βiL2 .
The proof of Hodge Decomposition Theorem requires some results from Functional
Analysis.
Regularity Theorem. Any weak solution l of (3.30) is of the form l(β) = hω, βiL2 , for
some ω ∈ Ωp (M ) (and hence defines a solution of (3.30)).
Compactness Theorem. Assume that a sequence αn ∈ Ωp (M ) satisfies kαn k < C and
k∆αn k < C, for some C independent of n. Then αn contains a Cauchy subsequence.
We shall assume the above two theorems (and the Hahn–Banach theorem below)
without proof.
Compactness Theorem implies at once that Hp must be finite-dimensional (for,
otherwise, there would exist an infinite orthonormal sequence of forms). As Hp is finite-
dimensional, we can write an L2 -orthogonal decomposition Ωp (M ) = Hp ⊕ (Hp )⊥ .
It is easy to see that ∆Ωp (M ) ⊆ (Hp )⊥ (use Proposition 3.28). For the reverse inclusion,
suppose that α ∈ (Hp )⊥ . We want to show that the equation (3.30) has a solution.
Assuming the Regularity Theorem, we shall be done if we obtain a weak solution l :
Ωp (M ) → R of (3.30).
Define l first on a subspace ∆Ωp (M ), by putting l(∆η) = hη, αiL2 . It is not hard to
check that l is well-defined. Further, (ii) is automatically satisfied (on this subspace); we
claim that (i) holds too. To verify the latter claim, we show that l is bounded below on
∆Ωp (M ) using, once again, the Compactness Theorem.
In order to extend l to all of Ωp (M ), we appeal to
Hahn–Banach Theorem. Suppose that L is a normed vector space, and L0 a subspace
of L, and l : L0 → R a linear functional satisfying l(x0 ) < kx0 k, for all x0 ∈ L0 . Then l
extends to a linear functional on L with l(x) < kxk for all x ∈ L.
Thus we obtain a weak solution of (3.30) and deduce that Ωp (M ) = ∆Ωp (M ) ⊕ Hp
as desired. The two other versions of the L2 -orthogonal decomposition of Ωp (M ) follow
readily by application of Proposition 3.28.
Corollary 3.31 is a surprising result: an analytical object (harmonic forms) turns out to
be equivalent to a topological object (de Rham cohomology) via some differential geometry.
Here is a way to see ‘why such a result can be true’.
A de Rham cohomology class, a ∈ H r (M ) say, can be represented by many differential
forms; consider the (infinite-dimensional) affine space
Ba = {ξ ∈ Ωr (M ) | dξ = 0, [ξ] = a ∈ H r (M )}
= {ξ ∈ Ωr (M )|ξ = α + dβ, for some β ∈ Ωr−1 (M )}.
When does a closed form α have the smallest L2 -norm amongst all the closed forms in a
given de Rham cohomology class Ba ?
Such a form must be a critical point of the function F (α + dβ) = kα + dβ)k2 on Ba , so
the partial derivatives of F in any direction should vanish. That is, we must have
d
0= hα + t dβ, α + t dβiL2 = 2hα, dβiL2 .
dt t=0
R
Integrating by parts, we find that M hδα, βig = 0 must hold for every β ∈ Ωr−1 (M ). This
forces δα = 0, and so the extremal points of F are precisely the harmonic forms α.
Let V be a vector space over R. Assume, for simplicity, that V has finite dimension n say.
1. If W is another real finite-dimensional vector space then the tensor product V ⊗ W may
be defined as the real vector space consisting of all formal linear combinations of elements
v ⊗ w (for v ∈ V , w ∈ W ), with the relations
vi ⊗ wj ∈ V ⊗ W 7→ v ∗ (vi ) w∗ (wj ) ∈ R
v ⊗ w ∈ V ⊗ W → ψ(v, w) ∈ R
defines a natural isomorphism between the vector space of all bilinear forms on V × W and
the dual space (V ⊗ W )∗ . Recall from linear algebra that the space of bilinear forms on V × U
may be naturally identified with the space L(V, U ∗ ) of linear maps V → U ∗ . Putting U ∗ = W
and noting the above relations, one obtains a natural linear isomorphism
L(V, W ) ∼
= V ∗ ⊗ W.
In the special case W = R this recovers the definition of dual vector space V ∗ .
λj (vj ) = δij .
Here δij is the ‘Kronecker delta’, δij = 1 if i = j and is 0 otherwise. The p-th exterior power
Λp V ∗ of V ∗ (p ≥ 0) is the vector space of all the functions h : V × . . . × V → R, such that h is
(1) multilinear: h(u1 , . . . , aui + bu0i , . . . , up ) = ah(u1 , . . . , ui , . . . , up ) + bh(u1 , . . . , u0i , . . . , up );
(i.e. linear in each argument) and
(2) antisymmetric: ∀i < j ∈ {1, 2, . . . , n}, swapping the ith and j th vector changes the sign
It follows that for 1 ≤ p ≤ n, dim Λp V ∗ = np , a basis may be given by {λi1 ∧ . . . ∧ λip : λij ∈
V ∗ , 1 ≤ i1 < . . . < ip ≤ n}. Also Λp V ∗ = {0} when p > n. One formally defines Λ0 V ∗ = R.
The exterior product (or wedge product) is a bilinear map
(λ, µ) ∈ Λp V ∗ × Λq V ∗ → λ ∧ µ ∈ Λp+q V ∗ .
(λi1 ∧. . .∧λip )∧(λip+1 ∧. . .∧λip+q )(u1 , . . . , up+q ) = det(λij (uk )), (uk ∈ V, j, k = 1, . . . , p+q).