760 Notes
760 Notes
January 2015
Later, the ideas of geometrization leached into other areas of manifold theory; of special
note is the case of 3-manifolds. As opposed to the 3 homogeneous geometries of 2-manifolds,
Thurston was noticed that 3-manifolds could admit just 8 homogeneous geometries (with
compact stabilizers), called the model geometries. He noticed that, although incompressible
spheres and tori obstructed the existence of a model geometry, incompressible 2-manifolds
of other kinds did not. This, along with other evidence, led Thurston conjecture that any
3-manifold can be cut along incompressible tori (in an essentially unique way) such that the
resulting connected components admit a complete metric with one of the 8 geometries.
Since the topological structure of any manifold with a model geometry is completely
understood, if any compact 3-manifold admits such a decomposition, then the topology of
all compact 3-manifolds would be understood.
1
Starting in the early 1980’s, Richard Hamilton began a program of using the Ricci
Flow to prove the geometrization conjecture for 3-manifolds. The Ricci flow is an evolution
equation for the metric:
dgij
= −2Ricij . (1)
dt
This is a non-linear 2nd order PDE, which resembles a heat equation, or more accurately
a reaction-diffusion equation. Hamilton proved that if singularities do not develop, the
resulting manifold in the limit carries a model geometry. Singularities proved difficult to
control, but in 2003 Perelman classified the singularities, proved that there are at most
finitely many of them, and thereby completed the geometrization program.
Many of the problems we’ll explore are involved in the search for canonical metrics.
Our list of topics (we’ll surely not get to all of them) is
• The Yamabe Problem: does a conformal class contain a constant-scalr curvature met-
ric? Is it unique?
• (prescribed scalar curvature) Given a function f : M n → R, is there a metric on M n
so that scal = f ?
• The Ricci flow: can we canonically smooth-out a metric using a heat-like flow in
curvature, like the heat flow “smooths out” any distribution of heat on a compact
body? (Possibly related topic: the Yamabe flow.)
• Complex geometry, especially Kähler geometry: if we add additional structure, do our
problems become more solvable? (Answer: Yes, but they’re still hard.)
2
• The Calabi conjectures, and Yau’s and Aubin’s partial solutions (possible related topic:
the Calabi flow)
• The Yang-Mills equation and the Yang-Mills flow: vector bundles are of decisive sig-
nificance in Riemannian geometry. Do bundles have canonical geometries? Can we
establish a heat-type flow to attempt to find them? (Yes and yes.)
The primary nexus of analysis and differential topology is Stokes Theorem: if ω is a k-form
on M n and N k+1 ⊂ M n is a submanifold, then
Z Z
dω = ω. (2)
N ∂N
The primary nexus of analysis and differential geometry is the cluster of results including and
surrounding the Sobolev embedding theorems. These are familiar in the Euclidean setting,
where they are usually proven (and understood) as consequences of the Poincare inequality,
which is normally proven via iterated applications of FTCII (which is Stokes Theorem in
dimension 1). But careful study of the proof shows that it is dependent on the Euclidean
structure of Rn ; the proof cannot be imitated on manifolds. What’s more, the statement of
the Poincare inequality is very distinctly not diffeomorphism invariant.
The proof of the Sobolev inequalities of Federer-Fleming, dating to the late 1950’s,
showed us their fundamental connections with geometry; it is their proof that we’ll give
here. Before giving the proof, we introduce two constants. If (Ω, g) is a domain with metric
g, then the ν-isometric inequality is
|∂Ωo |
Iν (Ω) = inf ν−1 (3)
Ωo ⊂⊂Ω |Ωo | ν
where |∂Ωo | indicates the (n−1) Hausdorff measure of ∂Ωo and |Ωo | indicates the n-Hausdorff
measure of Ωo . The ν-Sobolev constant is
R
|∇f |
Cν (Ω) = inf ν−1 (4)
f ∈C ∞ (Ω) R ν
Supp(f )⊂Ω f ν−1 ν
3
Theorem 1.1 (Federer-Fleming)
Iν (Ω) = Cν (Ω)
This is the more interesting case. Assume f is smooth, with isolated critical points.
On any non-critical point p, let dA indicate the area measure on the level-set of f passing
through p. The co-area formula is
1
dV olp = df ∧ dA (7)
|∇f |
Then using Ωt = {|f | ≥ t} to denote the superlevel sets, we have
Z Z Z sup |f | Z Z ∞
|∇f | dV ol = df ∧ dA = dA df = |∂Ωt | dt
inf |f | {f =0} 0
Z ∞
(8)
ν−1
≥ Iν (Ω) |Ωt | ν
and using the “layer-cake” representation and then a change of variables (a lá multivariable
calculus), we have
Z Z Z |f (p)| Z ∞Z
ν ν 1 ν 1
f ν−1 = t ν−1 dt dV ol(p) = t ν−1 dV ol dt
ν−1 Ω 0 ν−1 0 Ωt
Z ∞ (9)
ν 1
= t ν−1 |Ωt | dt
ν−1 0
Lastly we have to connect these. We have
4
Proof. First taking a derivative
! 1s ! 1−s
s
Z T Z T
d s−1 s−1 s−1
s t g(t) dt = T g(T ) s t g(t) dt
dT 0 0
! 1−s (11)
Z T s
1 1
≤ T s−1 g(T ) s s ts−1 = g(T ) s
0
ν
Applying this lemma with s = ν−1 we get
Z ν
ν−1 Z ∞ ν−1 Z ∞ Z
ν−1 ν 1
ν ν−1 1 (13)
f ν = t ν−1 |Ωt | dt ≤ |Ωt | ν ≤ |∇f |
ν−1 0 0 Iν (Ω)
pν
Theorem 1.3 For any p < ν we obtain embeddings L ν−p ⊂ W01,p .
If f has more than one weak derivative, then the embedding theorem also improves.
pν
Theorem 1.4 If 1 ≤ p < ∞, k ∈ N, and kp < ν then we obtain embeddings L ν−pk ⊂ W0k,p .
pν
So what happens if pk ≥ ν? As pk % ν, then ν−pk % ∞, so perhaps the embedding
∞
is into L . This naive supposition fails because the constant in the Sobolev inequality also
degenerates (see the exercises).
n
Theorem 1.5 If l + α = k − p and Cn (Ω) > 0, then
5
Theorem 1.6 The embeddings above are all compact embeddings.
|∂Ωo |
Iν (Ω) = inf ν−1
Ωo ⊂⊂Ω
|Ωo |≤ 21 |Ω|
|Ωo | ν
R
|∇f | (15)
Cν (Ω) = inf ν−1
f ∈C ∞ (Ω)
R ν
ν
Supp(f )⊂Ω f ν−1
| supp(f )|≤ 12 |Ω|
1.4 Laplacians
If T is a tensor, say
O O
p q ∗
T : TM ⊗ T M → R
T (X, Xi1 , . . . , Xip , η j1 , . . . , η jq ) ∈ R (16)
∂ ∂
T = Ti1 ...ip j1 ...jq dxi1 ⊗ · · · ⊗ dxip ⊗ ⊗ · · · ⊗ jq
∂xji ∂x
then we can define its covariant derivative, defined via a Leibniz rule:
6
The tensor ∇T has an expression in coordinates, where, by convention, the derivative is
indicated with a comma:
∂ ∂
∇T = Ti1 ...ip j1 ...jq ,k dxi1 ⊗ . . . dxip ⊗ ⊗ · · · ⊗ j1 ⊗ dxk . (18)
∂xj1 ∂x
One can compute the following expression for the component functions:
d
Ti1 ...ip j1 ...jq ,k = Ti ...i j1 ...jq + Γski1 Ts...ip j1 ...jq + . . . + Γskip Ti1 ...s j1 ...jq
dxk 1 p (19)
j
+ Γjks1 Ti1 ...s s...jq + . . . + Γksp Ti1 ...s j1 ...s .
∂2f df
4f = g ij − Γk k , Γk , g ij Γkij . (21)
∂xi ∂xj dx
1.4.2 Forms, the Hodge Duality Operator, and the Hodge Laplacian
V
k
We recall a few definitions. We have the sections of k-forms, denoted Γ M n . By abuse
Vk n Vk
of notation, we denote this by simply M or just . The Hodge duality operator or just
Hodge star is term given to the tensoral forms of the volume form:
√
dV ol = gij dx1 ∧ · · · ∧ dxn ,
^
k
^
n−k
(22)
∗: −→
Vk Vk+1
Now let’s look at the exterior derivative d : → . We would like to find its L2
adjoint. Now d is a linear operator, but is only densely defined. If we restrict the L2 space to
just the C ∞ k-forms, then d is defined everywhere, but it is not a bounded (or continuous)
7
linear operator, and the inner peoduct space is not a Hilbert space (due to non-completeness
of C ∞ under the L2 norm). Despite all this, we can find an adjoint for d, by making use of
Stoke’s theorem. We find that d∗ , defined implicitly by
The first order operators d, d∗ can be combined give a second order operator, the Hodge
Laplacian, by
^ ^
k k
4H : −→ ,
(26)
∗ ∗
4H η = d dη + dd η.
This is not the same as the rough Laplacian, although at the first and second order levels
they are essentially the same. Notice that 4 is negative definite and 4H is positive definite:
(4η, η) = −|∇η|2
(27)
(4H η, η) = |dη|2 + |d∗ η|2 .
[d, 4H ] = d4H − 4H d = 0
[d∗ , 4H ] = d∗ 4H − 4H d∗ = 0 (29)
[∗, 4H ] = ∗4H − 4H ∗ = 0.
There are no such simple commutation relation for the rough Laplacian, although later we
shall talk about the important commutator [4, ∇].
8
10) Prove that a rough-harmonic form is constant, while a Hodge-harmonic form need not
be constant, but must be both closed and co-closed.
11) In addition to the inner product (24), there is another inner product
Show that the two are proportional, and find the constant of proportionality.
Let η : M n → R≥0 be a test function, meaning η ∈ Cc∞ (M n ), and let f be any weakly
differentiable function that weakly satisfies
4f + uf ≥ 0 (30)
(more general elliptic equations can be considered, but for simplicity we’ll stick to (30) as
the method in the general case is essentially the same.) A standing assumption will be
f ≥ 0.
n
1,p
In addition to (30) assume f ∈ Wloc for some p > 1 and assume u ∈ Lloc 2
. Then an
application of the divergence theorem along with the Cauchy-Schwartz inequality (and some
propitious re-arrangement) gives
Z Z
(p − 1) η 2 f p−2 |∇f |2 = η 2 ∇f p−1 , ∇f
Z
Div η 2 ∇f p−1 ∇f
=
Z p Z (31)
p
−1 2 p
2 p−2
− p − 1ηf 2 ∇f, √ f ∇η −
2 η f 4f
p−1
p−1
Z Z Z
2
≤ η 2 f p−2 |∇f |2 + |∇η|2 f p + η2 f p u
2 p−1
which gives us an equation that could be regarded as a form of reverse Sobolev inequality:
Z 2 Z Z
2 p−2 2 2 2 p 2
η f |∇f | ≤ |∇η| f + η2 f p u (32)
p−1 p−1
n
Set γ = n−2 and notice γ > 1 (this is convenient for applications of Hölder’s inequality as
1 2
γ + n = 1). One form of the Sobolev inequality is
Z 1
2γ Z 12
2γ 2
Cn (Ω) g ≤ |∇g| (33)
9
where g ∈ W01,2 (Ω). Using our “reverse Sobolev inequality” along with the actual Sobolev
inequality gives
Z γ1 Z
p
2 2γ pγ
Cn (supp η) η f ≤ |∇(ηf 2 )|2
p2
Z Z
≤ 2 |∇η|2 f p + η 2 f p−2 |∇f |2 (34)
2
2 Z!
p2
Z
p 2 p
≤ 2 1+ |∇η| f + η2 f p u
p−1 p−1
" n2 # Z γ1 2 ! Z
p2
Z
2 n
2γ pγ p
Cn (supp η) − u2 η f ≤ 2 1+ |∇η|2 f p
p−1 supp η p−1
R n
Thus, provided supp η U 2 is sufficiently small with respect to both p and the Sobolev
constant, then the local Lγp -norm of f is controlled in terms of the local Lp -norm of f .
Lifting apriori from lower Lp -norms to higher Lp -norms is called bootstrapping.
Iterating this estimate as many times as necessary, and choosing appropriate test func-
tions at each stage (see below), gives us the following theorem
10
implies
Z ! q1 Z p1
q p
f ≤ C f (39)
Br/2 Br
n
In other words, f ∈ Lploc , p > 1 and u ∈ Lloc2
implies f ∈ Lqloc for all q ∈ (p, ∞). However
both = (p, q, Cn ) and C = C(p, q, Cn , r) deteriorate as q → ∞ (see the exercises), so we
do not obtain apriori L∞ bounds.
It is worthwhile to formalize the choice of test functions necessary to the proof of Theorem
1.8. The test functions we construct below are called cutoff functions.
Let m ∈ M n be any point, and let r = dist(m, ·) be the distance function from m.
Distance functions are not smooth, but are Lipschitz and therefore weakly differentiable.
Since we encounter at worst gradients of η in our integral estimates, the fact that our “test
functions” are just C 0,1 and not C ∞ is no bother to us.
Given some radius r0 ∈ (0, ∞), we may wish to have a test function that is zero
outside Br0 (m), is 1 inside Br0 /2 (m) and has bounded gradient in the intermediate annulus
Br0 (m) \ Br0 /2 (m). Then define
1
, r ∈ [0, r0 /2]
2
η(r) = 2 − r0 r , r ∈ (r0 /2, r0 ) (40)
0 , r ∈ [r0 , ∞)
It is also sometimes necessary to have a variety of cutoff functions, with the collection
of cutoff functions operating on progressively smaller nested balls, say. To this end consider
the functions
1
, r ∈ [0, (2−1 + 2−i−2 )r0 ]
2i+1 −1 −i−2
, r ∈ (2−1 + 2−i−2 )r0 , (2−1 + 2−i−1 )r0 (41)
ηi (r) = 1 − r0 r − (2 − 2 )r0
, r ∈ [(2−1 + 2−i−1 )r0 , ∞).
0
11
1.6.3 Nash-Moser Iteration and the L∞ estimates
In addition to f ∈ Lp for some p > 1, now assume u is in a strictly larger Lp space than
p = n2 . We shall assume that u ∈ Lpγ . (Aside: u ∈ Lp+δ for any δ > 0 is sufficient to obtain
all our conclusions; we refer you to the sources (eg. Gilbarg-Trudinger) for the details.)
1 2 2 1
Beginning with (34), we observe γ2 + n + nγ = 1, which allows us to use Hölder’s
inequality with greater subtlety:
Z γ1
2 2γ pγ
Cn (supp η) η f
2 ! Z
p2
Z
p 2 1
4 2
≤ 2 1+ 2 p
|∇η| f + η γ fpγ · η n fpn · u (43)
p−1 p−1
2 ! Z 1 n2 Z n2 γ1
p2
Z Z
p γ2 n
≤ 2 1+ |∇η|2 f p + η 2γ f pγ η2 f p u2γ
p−1 p−1 supp η
n
Notice that p 2p is uniformly bounded. This gives us the possibility of iteration without the
deterioration of the estimate that we saw above.
Let ηi , i ∈ {0, 1, 2, . . . } be the ith cutoff function constructed above, and set pi = pγ i .
Also, abbreviate Cn = Cn (Br0 ). Then Lemma 1.9 gives us
!p 1 ! p1
Z i+1
1 2 h n
i p1 Z i
f pi+1 (2i r0 )2 + pi f pi
pi i
≤ C(n) Cn
pi 2
. (46)
B(2−1 +2−i−2 )r B(2−1 +2−i−1 )r
0 0
12
This can be iterated all the way down, to obtain
!p 1 " i # Z ! p1
Z i+1 Y 1 2 h n
i p1
pi+1 pk k 2 k p
f ≤ C(n) Cn (2 r0 ) + pk
pk 2
f . (47)
B(2−1 +2−i−2 )r k=0 B r0
0
then
Z ! 12
sup |f | ≤ C f2 (49)
Br0 /2 (m) Br0 (m)
This method of finding apriori L∞ control after assuming just L2 control is called Moser
iteration or Nash-Moser iteration, depending on whether you’re an analyst or a geometric
analyst.
13
1.8 Variational Problems and Euler-Lagrange Equations
One is sometimes interested in finding minimizers (or more generally extremizers) of such
a functional. To do this, we consider variations Tt = T + tH of the tensor T and compute
the first order effect on F (when M n is non-compact it is normal to put a restriction on H,
namely that H ∈ W01,2 or H ∈ Cc∞ even, in order that we can use Stokes’ theorem). If first
derivatives of F(Tt ) vanish regardless of what variation H is chosen, then the tensor T is
said to be a critical point for F. Then if T is an extremizer, then we compute
Z
d d
0 = F(Tt ) = |∇Tt |2
dt t=0 dt
Z Z (52)
= 2 h∇T, ∇Hi = −2 h4T, Hi .
Since H is arbitrary, we can choose H = η4T , where η ≥ 0 is any test function, to obtain the
requirement that 4T = 0. One easily sees that indeed 4T = 0 if and only if T extremizes F.
In general, the pointwise equations one obtains as a requirement for a quantity to extremize
a functional are called the Euler-Lagrange equations of that functional.
There is a more abstract way to look at this process. Let T be the vector space
of all W 1,2 sections of tensors on M n ; this is a Banach space. Then F : T −→ R is a
map from one Banach space to another, and more specifically, it is just a function on the
(infinite dimensional) space T . As such, we should be able to linearize it (that is, find its
“gradient”). This linearization at T ∈ T is called its Frechet derivative of F at T , and is
sometimes denoted
DF T
: W01,2 −→ R. (53)
We are looking for those points T ∈ T at which the linearized operator DF|T vanishes.
Unfortunately we won’t have the time to explore this abstract point of view much further
than this.
14
1.8.2 Spaces of Riemannian Metrics, and the Riemannian Moduli Space
We will be applying functionals to the Riemannian structure itself, so we will have to consider
the appropriate space of Riemannian structures. So consider the space of Riemannian
metrics
n O o
2 ∗
M et(M n ) = g ∈ T M g is symmetric and positive definite (55)
Obviously if ϕ : M n → M n is any diffeomorphism the metric pulls back. Thus the diffeomor-
phism group Dif f (M n ) acts on M et(M n ), giving the so-called moduli space of Riemannian
metrics
The space M et(M n ) is a cone, not a vector space, and is infinte dimensional. The group
Dif f (M n ) is an extremely complicated infinite dimensional Lie group. It is non-compact,
and otherwise very difficult to understand. Thus M od(M n ) is also difficult to directly
understand. Yet is is our fundamental object, as the curvature functionals are invariant
under diffeomorphism and therefore pass from maps on M et(M n ) to maps on M od(M n ).
J2 V2 V2
Indeed we may regard Rm as an operator Rm : → that is diffeomorphism-
invariant:
This is scale invariant only in dimension n = 2, in which case it is related to the Euler
constant. To fix this we can normalize by volume
Z
n 1
H(M , g) = n−2 R dV ol (59)
V ol(M n ) n M n
15
Obviously functionals for other curvature quantities also exists:
Z
n
n
RM(M , g) = | Rm | 2 dV ol
n
ZM
n
n
RIC(M , g) = | Ric | 2 dV ol (61)
n
Z M
n
n
W(M , g) = |W | 2 dV ol.
Mn
gt = g + th (63)
dg
where h = hij dxi ⊗ dxj is any
symmetric 2-tensor. We have dt = h. With the usual formula
1 ∂gis ∂gjs ∂gij
k
Γij = 2 ∂xj + ∂xi − ∂xs g sk we compute
dΓkij
1 ∂his ∂hjs ∂hij 1 ∂gis ∂gjs ∂gij
= j
+ i
− g sk − + − g su g vk hsv . (64)
dt 2 ∂x ∂x ∂xs 2 ∂x j ∂x i ∂xs
t=0
For theoretical reasons we know that Γ̇kij is tensorial, so we apply the following standard
trick: express Γ̇kij in a coordinate system in which all Christoffel symbols vanish, then convert
all partial derivatives to covariant derivatives. In geodesic normal coordinates, at any single
∂g
chosen point p we have gij = δij + O(2) so the ∂x terms vanish. Thus we have
dΓkij
1 ∂his ∂hjs ∂hij
= + − g sk
dt 2 ∂xj ∂xi ∂xs
t=0
(65)
dΓkij 1
= (his,j + hjs,i + hij,s ) g sk
dt 2
t=0
16
Using the standard formulas for Rmijk l in terms of the Christoffel symbols, we obtain
d 1
Rm ijk l (gt ) = (hjs,ki + hks,ji − hjk,si − his,kj + hks,ij − hik,sj ) g sl
dt t=0 2
d 1 s
Ric ij (gt ) = (h i,js + hs j,is − hjk,s s − (T r h),ij ) (66)
dt t=0 2
d
R(gt ) = hst ,ts − 4(T r h) − Ricij hij
dt t=0
d
From this, along with ol = 12 (T r h) dV ol we can easily compute
dt dV
Z Z
d 1
R dV ol = Ric − R g , h dV ol (67)
dt t=0 2
so that a stable point for the Hilbert functional occurs when the gravitational tensor van-
ishes:
1
Gij , Ric ij − Rgij
2 (68)
Gij = 0.
16) (The p-Laplacian) Determine the Euler-Lagrange equations for the non-quadratic func-
tionals
Z
F(u) = |∇u|p (70)
This pulls back to S2 , under standard streographic projection, to the metric of constant
curvature +1. Given any constant Λ > 0, we have diffeomorphisms R2 → R2 given
17
by (x, y) 7→ (Λx, Λy). Prove that this lifts to a diffeomorphism on S2 (obviously this
map is differentiable away from the north pole; you have to verify it is differentiable
at the north pole as well). The metric on S2 therefore transforms via diffeomorphism
by following this proceedure: pushforward under standard stereographic projection,
pullback under coordinate multiplication by Λ, and then pullback under standard
stereographic projection. Gaussian curvature is a diffeomorphism invariant, so the
sectional curvatures of these metrics remains +1 (this can also be verified by direct
computation). By letting Λ → 0 or Λ → ∞, show that the set of round metrics of
constant curvature +1 is not compact the C ∞ (or C 0 ) topology. Conclude that the
diffeomorphism group of S2 is non-compact.
18) Show that any linear fractional transformation on R2 ∪ {∗} lifts to a diffeomorphism
on S2 , so that P SL(2, R) ⊂ Dif f (S2 ). With this, find a sequence of diffeomorphisms
ϕi : S2 → S2 such that the sequence of pullbacks ϕ∗i (g) of a round metric g converges
pointwise to zero at every point of S2 .
19) Show that the functionals Rk dV ol are diffeomorphism invariant, and therefore in-
R
23) Find the Euler-Lagrange equations for the quadratic scalar curvature functional. In
dimension other than 4, normalize this functional, and re-compute the Euler-Lagrange
equations.
24) (Einstein’s Cosmological Constant) If RΛ : M n → R is a function, we have the Hilbert
functional with a cosmological term: (R − Λ)dV ol. Find the Euler-Lagrange equa-
tions. Show that there exist non-Ricci flat solutions, and that for these solutions Λ
must be constant. Such manifolds are called Einstein manifolds.
We have seen that the half-dimension Lp norms play two important roles that appear, at
first to be independent. First, the apriori estimates for 4f ≥ uf require u ∈ Lploc for any
p > n2 , with p = n2 being the critical case where a weak conclusion holds (in the case p < n2
n
no conclusions at all are possible). Second, the L 2 Riemannian norms are scale invariant,
and therefore their values carry intrinsic geometric meaning, as opposed to Lp norms for
other values of p, which can be made to take any value whatever simply by scaling the
metric.
Here we shall see how these two roles combine to produce a remarkable regularity
theory for critical metrics (for simplicity we restrict ourselves to Einstein metrics).
18
1.10.1 Derivative Commutator Formulas
From basic Riemannian geometry we have the definition of the Riemann tensor as a com-
mutator: if X is any vector field then
X l ,ij − X l ,ji = Rmijs l X s . (72)
By computation this can be extended to other tensors:
T l1 ...ls ,ij − T l1 ...ls ,ji = Rmijs k1 T sl2 ...ln + Rmijs l2 T l1 s...ln + . . . + Rmijs ln T l1 ...ln−1 s (73)
and to covectors η = ηi dxi
ηk,ij − ηk,ji = −Rmijk s ηs
(74)
Tk1 ...kn ,ij − ηk,ji = − Rmijk1 s Tsk2 ...kn − Rmijk2 s Tk1 s...kn − . . . − Rmijkn s Tk1 ...kn−1 s .
This can be extended to tensors of any type. For instance
Ta bc ,ij − Ta bc ,ji = Rmija s Ts bc − Rmijs b Ta sc − Rmijs c Ta bs (75)
Both the first and second Bianchi identities can be derived from the diffeomorphism invari-
ance of the Riemann tensor (Reference: J. Kazdan, 1981). If we regard Rm = Rm(g) as a
second order differential operator on the metric g, then diffeomorphism invariance takes the
form
Rm(ϕ∗ g) = ϕ∗ (Rm(g)) (76)
n n ∗
where ϕ : M → M is any diffeomorphism and ϕ is the pullback. Letting ϕt be a time-
dependent family of diffeomorphisms with a well-chosen variational field, then taking a Lie
derivative on both sides of (76) gives the second Bianchi identity (see exercises). A similar
process gives the first Bianchi identity. In this light the Bianchi identities are understood
as Leibniz rules.
Alternatively, the second Bianchi identity identity can be computed directly from the
definition of Rm, where it ultimately follows from the Jacobi idenity, which itself is both an
expression of diffeomorphism invariance (of the bracket), and also a Leibniz rule.
The 2nd Bianchi identity can be expressed in either of the following two ways:
Rmijkl,m + Rmijlm,k + Rmijmk,l = 0
(77)
Rmijkl,m + Rmjmkl,i + Rmmikl,j = 0.
Tracing once gives
Ricij,k − Ricik,j = Rmsijk ,s (78)
Tracing again gives
R,k = 2Ricsk ,s . (79)
19
1.10.3 The Laplacian of the Riemann Tensor
Combining the 2nd Bianchi identity with the commutator formula, we can obtain and ex-
pression for the Laplacian of the full Riemann tensor. First using the 2nd Bianchi idenitity
we get
(4 Rm)ijkl , g st Rmijkl,st
(80)
= −g st Rmijls,kt − g st Rmijsk,lt
Using the commutator formula on both terms gives
(4 Rm)ijkl = −g st Rmijls,tk − g st Rmijsk,tl
(81)
+ g st Rmkti u Rmujls + . . .
where “. . . ” indicates a further seven quadratic monomials in the Riemann tensor. Whenever
the exact expression doesn’t matter, we shall abbreviate any linear combination of tensor
products and traces of a tensor T with a tensor S simply by S ∗ T . To (81), apply the 2nd
Bianchi identity again to get
(4 Rm)ijkl = g st Rmjtls,ik + g st Rmtils,jk + g st Rmjtsk,il + g st Rmtisk,jl + Rm ∗ Rm
(82)
= −Ricjl,ik + Ricil,jk + Ricjk,il − Ricik,jl + Rm ∗ Rm
The second derivative terms are now entirely on various copies of the Ricci tensor. Schemat-
ically expressed, we have
4 Rm = ∇2 Ric + Rm ∗ Rm . (83)
In the n ≥ 3 Einstein case we have Ric = λg where λ = const, and so we obtain the
tensoral elliptic equation 4 Rm = Rm ∗ Rm. This is difficult to deal with using the above
elliptic theory, however, because it was developed for functional Laplace equations of the
sort 4f ≥ uf . We therefore resort to the so-called Kato inequality: for any tensor T we
have
|∇|T || ≤ |∇T |. (84)
Then the two identities
1
4|T |2 = h4T, T i + |∇T |2
2 (85)
1
4|T |2 = |T |4|T | + |∇|T ||2
2
along with the Kato inequality give
|T |4|T | ≥ h4T, T i . (86)
Using T = Rm and | hRm ∗ Rm, Rmi | ≥ −C| Rm |3 (from Cauchy-Schwartz) we get the
equation
4| Rm | ≥ C(n)| Rm |2 (87)
which holds strongly away from | Rm | = 0 and holds weakly everywhere.
20
1.10.4 Consequences of the Einstein Condition
We have seen that the Einstein condition gives the weak elliptic inequality
4u ≥ −u2 (88)
where u = C| Rm |. Using the Lp theory above, where f = u, gives the statement that there
are constants = (n, Cs ), = C(n, Cs , r) so that if
Z ! n2
n
| Rm | 2 ≤ (89)
Br (m)
then
n2 γ1 ! n2
Z Z
n n
γ
| Rm | 2 ≤ C | Rm | 2 (90)
B 3 r (m) Br (m)
4
4
(this is from Theorem 1.8). It can be shown that the r–dependency of C is r− n (see
n
exercises). But then with a uniform local estimate of the L 2 γ -norm of u we can actually
use the more powerful -regularity theorem (Theorem 1.10) to obtain
Z ! n2 Z ! n2
n n
| Rm | 2 ≤ =⇒ sup | Rm | ≤ C | Rm | 2 (91)
Br (m) B 1 r (m) Br (m)
2
26) (Harder, but not terrible) Use the process from Problem 25 and the variation formula
for the Riemann tensor to derive the traced 2nd Bianci and finally the full 2nd Bianchi
identity.
21
27) Derive the exact expression for Rm ∗ Rm in (83). Then derive an explicit estimate for
C = C(n) in the weak inequality 4| Rm | ≥ −C| Rm |2 .
28) Prove the Kato inequality.
29) Run through the “First Bootstrapping Inequality” (Lemma 1.7) with a proper choice
of η to obtain
n2 γ1 ! n2
Z Z
γn 4
−n n
| Rm | 2 ≤ Cr | Rm | 2 (92)
B 3 r (m) Br (m)
4
R n
n2
for C = C(n, Cn ), under the condition that Br (m)
| Rm | 2 ≤ .
30) Use your result from Problem 29, along with a slightly more careful Moser iteration
process from Theorem 1.10 to obtain
Z ! n2
n
sup | Rm | ≤ Cr−2 | Rm | 2 (93)
B 1 r (m) Br (m)
2
R n
n2
for C = C(n, Cn ), under the condition that Br (m)
| Rm | 2 ≤ .
One interested in what local coordinates are “best.” Obviously what “best” means will
vary from situation to situation , and geodesic normal coordinates are particularly useful in
certain cases, as we already saw when we computed variation formulas.
However, for investigating regularity issues, geodesic normal coordinates are not good.
If a metric is only C k,α , for instance, then the exponential map causes a derivative loss
when constructing the coordinate frame, so, in fact, a C k,α metric will not appear even to
be C k,α in its own exponential normal coordinate chart (see the DeTurck-Kazdan paper for
details). So if the metric is C k,α , what sort of coordinate chart will show this?
4 xi = 0. (94)
22
2
Using 4 f = g ij ∂x∂i ∂x
f k ∂f 1 n
j − Γ ∂xk , we easily see that {x , . . . , x } is an harmonic chart if
and only if each of the traced Christoffel symbols Γk = g ij Γkij vanishes. A theorem of
DeTurck-Kazdan states that, for regularity issues, this coordinate frame is the best:
In particular, a C k,α metric will appear to be C k,α in its own harmonic coordinate chart.
The Ricci curvature tensor can be considered a non-linear differential operator that takes
the metric g as its input and produces a symmetric 2-tensor as its output. In harmonic
coordinates this transformation is explicitly a semilinear elliptic differential operator, as we
shall see.
We have
∂Γsij ∂Γssj
Ricij = − + Γstj Γtis − Γsij Γtts . (95)
∂xs ∂xi
We shall be interested exclusively in the second order part, so we shall denote all lower order
terms with the expression Q = Q(∂g, g), which will be quadratic separately in ∂g and in
g, and a polynomial of 4th degree in both together. Expanding out the Christoffel symbols
and making a cancellation we obtain
1 st ∂ 2 gij
2
∂ 2 gjs ∂ 2 gst
1 ts ∂ git
Ricij = − g + g + − + Q. (96)
2 ∂xs ∂xt 2 ∂xs ∂xj ∂xi ∂xt ∂xi ∂xj
j
∂Γ
Next consider the expression g ui ∂xu . We have
r
2
∂ 2 gvs ∂ 2 guv
∂Γ 1 ∂ gus
gri j = gri g uv g sr + − + Q
∂x 2 ∂xj ∂xv ∂xj ∂xu ∂xj ∂xs
(97)
∂ 2 gui
2
1 ∂ guv
= g uv v j − g uv + Q.
∂x ∂x 2 ∂xi ∂xj
Comparing this with (98) we see that, in general,
∂ 2 gij ∂Γr ∂Γr
1 1
Ricij = − g st s t + gri j + grj i + Q. (98)
2 ∂x ∂x 2 ∂x ∂x
Therefore in an harmonic coordinate system this simplifies to
1
Ric ij = − 4c gij + Q (99)
2
2
where 4c indicates the coordinate Laplacian 4c , g st ∂x∂s ∂xt .
23
1.12.3 Bootstrapping Regularity for Einstein Manifolds
Recall that in the case of Einstein metrics in dimension n ≥ 3, we derived apriori local
estimates on the Riemann curvature tensor, obtaining local L∞ bounds. Via the Jacobi
equation, this puts apriori C 2 control on the metric tensor. Now consider, in an harmonic
coordinate system, the equation
1
λgij = − 4c gij + Q, (100)
2
where λ = const, which comes from (99) and the Einstein condition. Because g ∈ C 2
we have Q ∈ C 1 , and that the coefficients on 4c are C 2 . The Schauder apriori estimates
therefore imply that the components gij must be C 2,α functions. But then Q ∈ C 1,α , so
the Schauder theory gives gij ∈ C 3,α . Now Q ∈ C 2,α , so Schauder implies gij ∈ C 4,α .
Continuing in this way, we obtain apriori C ∞ estimates for gij .
Many approaches to complex geometry exist; in these notes we’ll choose an approach that
should give differential geometers the gentlest possible introduction. We will regard com-
plex geometry and/or topology to simply be differential geometry and/or topology with an
additional differentiable structure, the so-called almost complex structure J, which is simply
any tensor
24
Note that J can be dualized: if η is any covector then we define J(η) to be the covector
η ◦ J (beware: sometimes the convention J(η) = −η ◦ J is used instead).
√
Because J squares to −1 its eigenvalues are ± −1, and its eigenvectors come in com-
plex conjugate pairs. If we define the complexified tangent space to be
TC M , C ⊗ T M (102)
Because T(0,1) M is a complex vector bundle, it does not make sense to talk about its
“integrability” in the sense that in spans any kind of submanifold. Yet it makes complete
sense that we can talk about the Frobenius criterion, so we say that T(1,0) M is integrable
√ √
iff [T(1,0) M, T(1,0) M ] ⊆ T(1,0) M . Letting Y − −1JY and Z − −1JZ be sections of the
holomorphic vector bundle, linearity of the bracket gives that
√ √
[Y − −1JY, Z − −1JZ] ∈ T(1,0) M (105)
if and only if
25
It is convenient to define the Nijenhuis tensor NJ to be
NJ (Y, Z) = [Y, Z] − [JY, JZ] + J ([Y, JZ] − [JY, Z]) . (107)
Although it may not be obvious, the Nijenhuis tensor is actually linear with respect to
multiplying Y or Z by a function (see exercises). Thus the condition for integrability is that
NJ ≡ 0. An almost complex structure that is integrable is called a complex structure.
Notice that in dimension 2 (and only dimension 2) the Laplace equation 4f = 0 is also
conformally invariant, meaning that if a function f is harmonic with respect to some metric,
it remains harmonic in any conformally related metric. This can be verifies as such: 4f =-
∗d(∗df ) so therefore 4f = 0 is the same as d(∗df ) = 0. But the 1-form ∗df is conformally
invariant (because d is independent of the metric and ∗ is conformally invariant), and so the
equation d(∗df ) = 0 is conformally invariant.
Theorem 1.12 If f is√any harmonic function and g is its conjugate in the sense that
dg = ∗df , then z = g + −1f is a holomorphic function.
26
so that the 2n many functions x1 , y 1 , . . . , xn , y n form a coordinate system, and so that the
function {z i } are all holomorphic. When this happens, we say that {z 1 , . . . , z n } form a
holomorphic coordinate system.
One final word. If {z 1 , . . . , z n } and {w1 , . . . , wn } are two different holomorphic coordi-
∂z i
nates systems, then the transition functions ∂w j are in fact automatically holomorphic. If
NJ ≡ 0, the Newlander-Nirenber theorem states that we can cover M n with domains that
each have holomorphic coordinate systems. Since the transitions on the overlaps are auto-
matically holomorphic, this means we have an equivalence amongst three different notions
of a complex manifold. Namely (M 2n , J) is a complex manifold iff the holomorphic tangent
bundle T(1,0) M is integrable iff the Nijenhuis tensor NJ vanishes iff M 2n can be covered by
J-holomorphic coordinate charts with holomorphic coordinate transitions.
27
1.16 The Bigraded Exterior Algebra, and the ∂ and ∂¯ Operators
The decomposition of TC∗ M into eigenspaces of the J operator implies a refinement of the
usual grading of the exterior algebra.
Define operators
^k,l ^k+1,l
∂ : −→ , ∂η = π (k+1,l) (dη)
^k,l ^k,l+1 (111)
∂¯ : −→ , ∂η = π (k,l+1) (dη) .
V0 V1 ¯ However, this does not hold in
The operator d : → obviously decomposes d = ∂ + ∂.
general. All we can say is that
^k,l ^k+l+1
d : → , (112)
28
and therefore dz i = ∂z i . Newlander-Nirenberg says that T(0,1) M is integrable, and a Frobe-
V1,0 V2,0 V1,1
nius theorem says therefore d : → ⊕ (see exercises) and, but taking complex
V0,1 V0,2 V1,1
conjugates, also d : → ⊕ . This means, in particular, that for any function f
we have
¯ )
0 = d2 f = d(∂f ) + d(∂f (114)
Vk,l
Theorem 1.14 If J is an integrable complex structure, there is a unique operator ∂ : →
Vk+1,l
that satisfies
Vk,l
Theorem 1.15 If J is an integrable complex structure, there is a unique operator ∂¯ : →
Vk,l+1
that satisfies
¯
1) (Constant linearity) ∂(nη ¯ + m∂γ
+ mγ) = n∂η ¯
¯ ∧ γ) = ∂η
2) (Leibniz Rule) ∂(η ¯ ∧ γ + (−1)deg η η ∧ ∂γ
¯
V0,0 V2,0
3) (Flatness) The operator ∂¯∂¯ : → is precisely zero
¯ = π 0,1 df .
4) (Action on functions) ∂f
29
Proof. See exercises.
A rank k real vector bundle E over M n can be described via charts {Uα }α that cover
M 2n (countably many with finite intersections) and “trivializations” Uα × Rk , along with a
collection of “transition functions”
that satisfy the cocycle conditions that ραβ = ρ−1βα and that on triple intersection ργα =
ργβ ρβα . Then we can form equivalence classes: given (x, v) ∈ Uα × Ck and (y, w) ∈ Uβ × Ck
we say
A rank k complex vector bundle is the same thing, except now the transitions are
A complex vector bundle is a holomorphic vector bundle provided the base space M n
is itself a complex manifold and provided, after choosing a basis for Ck and therefore being
able to describe the ρβ,α as matrix-values maps ρβα : Uα ∩ Uβ → Mn×n C
, we have that the
transitions are themselves holomorphic:
∂¯ ρβα = 0. (120)
In the real case, the exterior derivative exists exclusively on the base manifold. By contrast,
in the case of a holomorphic vector bundle E, there does exist a uniquely-defined exterior
derivative
30
(however it will be impossible to iterate this operator, so for instance ∂¯∂¯ will not be defined).
If {e1 , . . . , ek } is a basis for Ck , then any section e can be expressed over trivializations
Uα × Ck by e = ai ei and over Uβ × Ck by e = bi fi where the transition is fi = (ρβα )ji ej .
However the transition component functions (ρβα )ij are holomorphic, so by defining
¯ = ∂(a
∂e ¯ i ) ⊗ ei (122)
over a trivialization, we see that this definition is invariant over which trivialization is chosen.
To wit:
¯ = ∂(b
∂e ¯ i ) ⊗ fi
¯ j (ρβα )i ) ⊗ fi
= ∂(a j
(123)
¯
= ∂(a ) ⊗ (ρβα )i fi
j
j
¯ j ) ⊗ ei .
= ∂(a
Thus we see that the definition of ∂¯ is trivialization-independent, so ∂¯ exists as an operator
on E. A section e of E that has ∂e¯ = 0 is called a holomorphic section.
The holomorphic tangent bundle T(1,0) M of course has a ∂¯ operator. A vector field X
¯ = 0.
on M that is a section of T(1,0) M is called an holomorphic vector field provided ∂X
31
1.19 Covariant Exterior Differentiation and Curvature of Vector
Bundles
For a moment we retreat from complex geometry as such and develop some ideas in dif-
ferential geometry that will be of use to us, even in the complex case. A connection on a
vector bundle E (whether real or complex) is a map
∇ : E −→ T ∗ M ⊗ E (124)
that obeys a Leibniz rule: if σ is any section of E and f is any function, then
∇(f σ) = df ⊗ E + f ∇σ. (125)
We can combine any connection ∇ on E with the differential operator d on the base manifold
to create covariant exterior differentiation, which is often denoted D. We do this by simply
requiring that a Leibnz rule hold:
^k ^k+1
D : ⊗ E −→ ⊗E
(126)
D(η ⊗ σ) = dη ⊗ σ + (−1)deg(η) η ∧ Dσ.
Suppose {σ1 , . . . , σk } is some local frame for E. Then we define the 1-forms θij by
Of course iterating the exterior derivative d gives zero: d2 = 0. What happens when
we iterate the covariant exterior derivative D?
DD(η ⊗ σ) = D dη ⊗ σ + (−1)deg(η) η ∧ Dσ
= ddη ⊗ σ + (−1)deg(η)+1 dη ∧ Dσ + dη ⊗ σ + (−1)2deg(η) η ∧ DDσ (128)
= η ∧ DDσ
V0
Therefore D2 , while not necessarily zero, is tensorial. In particular, if η ∈ is any function
f , then the computation above shows D2 (f σ) = f D2 (σ). It is usual to denote this tensor
by Θ, and call it the curvature tensor:
^k ^k+2
Θ : ⊗ E −→ ⊗E
(129)
Θ(η ⊗ σ) = D (η ⊗ σ) = η ∧ D2 (σ).
2
This is identical to the more familiar notion of the curvature of line bundles in Riemnnian
geometry (see exercises). We can express this tensor in terms of the connection 1-forms by
D2 (σi ) = D θij ⊗ σj
= dθij ⊗ σj − θij ∧ θjk ⊗ σk (130)
= dθij + θis ∧ θsj σj
32
Therefore
The theory of covariant exterior differentiation dovetails nicely with the holomorphic vector
bundle theory, in the sense that holomorphic vector bundles automatically have a natural
connection, the ∂¯ operator, as we have already discussed. A connection ∇ : E → T M ⊗ E
is said to be compatible with the holomorphic structure of E provided ∇0,1 = ∂. ¯ Then if E
has an hermitian inner product h·, ·i, we obtain the following uniqueness theorem:
• ∇ : E −→ T ∗ M ⊗ E
• (Leibniz rule) ∇(f σ) = df ⊗ σ + f ∇σ
• (Compatibility with the Hermitian inner product) d hσ1 , σ2 i = h∇σ1 , σ2 i + hσ1 , ∇σ2 i
¯
• (Compatibility with the holomorphic structure) ∇0,1 σ = ∂σ.
This can be seen in two steps. Let {σi } be a holomorphic frame, meaning ∂σ ¯ = 0.
1,0 i
V2,0 V1,1 V2,0 V2,0 i V1,1
Therefore ∇σi = ∇ σi implies dθj ∈ ⊕ and θ ∧ θ ∈ so that Θ ∈ ⊕ .
Then we note
0 = d2 hσi , σj i
= d hDσi , σj i + d hσi , Dσj i
(132)
= hΘ(σi ), σj i − hDσi , Dσj i + hDσi , Dσj i + hσi , Θ(σj )i
= hΘ(σi ), σj i + hσi , Θ(σj )i
Now we specialize to the case that (M, J) is a complex manifold, so T1,0 M is a holo-
morphic vector bundle and has a ∂¯ operator. Suppose (M, J, h) is a complex manifold
with Hermitian inner product h. Then M has a natural connection, ∇C , called the Chern
connection—in general this is different from the Levi-Civita connection.
33
We would like to establish the conditions under which the ∇C = ∇L−C . That is, the
conditions under which the holomprhic theory of curvature and the real (Riemnnian) theory
coincide. This question leads us to the topic of Kähler geometry.
connection 1-forms are just the Christoffel symbols: θjk = Γkij dxi .
46) In the Riemannian setting where E = T M , show that Θlk is indeed the Riemannian
curvature operator.
47) Consider the 2-sphere, with metric given by gij = (1 + r2 )−2 δij . Determine the
connection 1-forms θji and the curvature tensor Θij .
48) In the real setting, what can you say about the traced curvature 2-form ρ = Θii ?
49) Prove the second Bianchi identity: dΘ + θ ∧ Θ + Θ ∧ θ = 0. Hint: start by showing
D (Θ(σ)) = Θ(D(σ)), and work out what this means in terms of the θ-symbols.
50) Starting with dhī = d hei , ej i = θis hs̄ + θjs hs̄ and pairing off by bi-degree, show that
θji = his̄ ∂hj s̄ and that Θij = ∂¯ his̄ ∂hj s̄
We shall not attempt to develop the theory of characteristic classes or even Chern classes
of complex vector bundles, but shall content ourselves with describing the first Chern class
of a holomorphic vector bundle.
34
Recall the Jacobi formula for matrices: if A = (Aij ) is a matrix dependent on some variable
d dA
x, then dx log A = Aji dxij . Using this, we obtain
√ √
−1∂¯ his̄ ∂his̄ = − −1∂ ∂¯ log det(h),
ρ = (136)
which is clearly a (1, 1)-form. We shall show below that it is a real (1,1)-form, but for now
¯
we will take this for granted. Because ∂∂(f ) = d∂(f ) (as a result of ∂ 2 = 0, we see that
ρ is locally exact, and therefore it is globally closed. Therefore [ρ] represents a deRham
cohomology class.
However, it is not yet clear that this class represents anything fundamental, because
within the definition of ρ is a choice of Hermitian form h on E. So suppose h and k are two
Hermitian forms on E. Then we can connect them through a path:
Note that ht is an Hermitian metric for all t ∈ [0, 1], and that h0 = h and h1 = k. For each
ht we have a new ρt , but we see
d √ d √
ρ = − −1∂ ∂¯ log det ht = − −1∂ ∂¯ (ht )ī (kī − hī )
dt √ dt (138)
= d − −1∂¯ T rht (k − h)
d
But T rht (k − h) is a globally defined function. Therefore dt ρt is an exact 2-form, which
means that the class [ρt ] is actually invariant. The class [ρ] is called the first Chern class of
¯ and denoted c1 (E). This gives the following remarkable theorem:
(E, ∂),
Theorem 1.18 (The First Chern Class) Assume (E, ∂) ¯ is a holomorphic vector bun-
dle. Given any Hermitian form h on E, the resulting Ricci 2-form ρ is real and closed. The
2
resulting deRham class [ρ] ∈ HdR (M n , R) is invariant under choice of Hermitian form h,
and is therefore a constant of the holomorphic structure ∂¯ of the bundle E.
The theory of Chern curvature obviously exists when we choose E = TC0 M , the holomorphic
tangent bundle. The resulting first Chern class c1 (T M ) = c1 (M ) is referred to as the first
Chern class of the manifold itself.
h = hi dz i ⊗ dz̄ j
1 1 (139)
= hi dz i z̄ j + hi dz i ∧ dz̄ j
2 2
35
We make the definitions
1
g = hī dz i z̄ j
2
√ (140)
−1
ω = hī dz i z̄ j
2
√
so that h = g − −1ω. Notice that, since hī = hjı̄ , we have
1 1
g =hī dz i dz̄ j = hjı̄ dz̄ i dz j
2 2 (141)
1 j i
= hjı̄ dz dz̄ = g.
2
Likewise we can compute ω = ω, so ω is also real. The real symmetric tensor g is called
the Riemannian metric associated to h, and the real antisymmetric 2-tensor ω is called the
Kähler form associated to h. We can prove that ω(·, ·) = g(J·, ·) (see exercises).
Every complex manifold has at least one Hermitian inner product; this can be obtained, for
instance, by using creating locally defined Hermitian forms, and gluing them together with
a partition of unity. But this is too “flexible” to be of further use.
This is actually a restrictive condition, and not every complex manifold admits a Kähler
metric. For instance, if M n is compact, note that
Z
n 1
Vol(M ) = ω ∧ · · · ∧ ω. (145)
n! M n
However,
R if ω = dη is actually exact, then ω∧· · ·∧ω = d(η∧ω∧· · ·∧ω) is also exact, meaning
ω ∧ · · · ∧ ω = 0, an impossibility. Therefore ω is closed but not exact, and therefore defines
a non-zero cohomology class [ω], called the Kähler class of the manifold.
36
1.22.4 Various Forms of the Kähler Condition, and the Equivalence of Real
and Complex Geometry on Kähler manifolds
√
Setting ω = ωī dz i ∧ dz̄ j where of course ωij̄ = −1hī , the Kähler condition is
dωī k dωī k
0 = dω = dz ∧ dz i ∧ dz̄ j + dz̄ ∧ dz i ∧ dz̄ j . (146)
dz k dz̄ k
for this to be zero, it is necessary and sufficient that
dωī dωk̄
k
= (147)
dz dz i
for all i, j, k. But this has an expression in terms of Christoffel symbols: because ωī =
√
−1hī , we have
d d d d d
hī = ∇ d , + , ∇ d/dz k dz j
dz k dz k dz i dz j dz i
= Γski hs̄ (148)
d
hk̄ = Γsik hs̄ .
dz i
Therefore the Kähler condition is equivalent to Γtik = Γtki . However this is precisely the
torsion-free condition of Riemannian geometry.
Theorem 1.19 The Chern connection on an Hermitian complex manifold is the Levi-Civita
connection if and only if the metric is Kähler.
Proof. Exercise.
ω(·, ·) =
54) Prove the converse of (52) and (53), namely if g is J-invariant, then setting √
g(·, ·), show that ω is a J-invariant, antisymmetric 2-form, and that h = g − −1ω is
an Hermitian metric.
37
55) If (M n , J, g) is a complex manifold with metric g, show that this is a Kähler manifold
provided the two compatibility conditions g(J·, J·) = g(·, ·) and dg(J·, ·) = 0 both
hold.
dωī dωk̄
56) Verify the assertion before (147). In particular, show that dz k
= dz k
is equivalent
dω
to dz̄ī
k = dω ik̄
dz̄ j
1.23.1 The Hodge Stars ∗ and ¯∗, and the Bidegree Decomposition
1.23.2 ¯
The ∂ ∂-operator
38