602 Notes1
602 Notes1
Sébastien Picard
Contents
1 Complex Geometry 2
1.1 Complex manifolds . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
1.2 Holomorphic vector bundles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
1.2.1 Definitions and notation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
1.2.2 Bundle constructions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
1.2.3 Constructions from the tangent bundle . . . . . . . . . . . . . . . . . . . . . . 14
1.3 Geometry of bundles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
1.3.1 Chern connection . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
1.3.2 Curvature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
1.3.3 Hermitian geometry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
2 Kähler Manifolds 26
2.1 Hodge theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
2.2 Kodaira vanishing theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34
2.3 Sheaves and the Lefschetz hyperplane theorem . . . . . . . . . . . . . . . . . . . . . 38
4 Calabi-Yau Threefolds 56
4.1 Parameters of threefolds . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56
4.2 Ricci flat metrics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57
4.3 Deformations of complex structure . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59
4.4 Quintic threefolds . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62
4.4.1 Holomorphic volume form . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62
4.4.2 Hodge numbers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63
4.4.3 Nodal singularities . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64
4.4.4 Examples of conifold transitions . . . . . . . . . . . . . . . . . . . . . . . . . 66
1
5.1 Blowing-up a nodal singularity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69
5.1.1 Blow-up review . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69
5.1.2 ODP in C3 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 70
5.1.3 ODP in C4 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71
5.2 Smoothing a nodal singularity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73
5.3 Candelas-de la Ossa metrics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75
5.3.1 Metrics on the small resolution . . . . . . . . . . . . . . . . . . . . . . . . . . 75
5.3.2 Metrics on the smoothings . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78
5.3.3 The cone metric . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 82
5.4 Special Lagrangian cycles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 85
5.4.1 Definition of special Lagrangian cycles . . . . . . . . . . . . . . . . . . . . . . 85
5.4.2 Examples on the smoothing . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89
References 96
1 Complex Geometry
This section is an introduction to complex geometry. For other references in the style of these notes,
see Kodaira’s book [19], Chapter 1 of Siu’s notes [24], Chapter 1 of Song-Weinkove’s notes [25], or
Chapter 1 of Szekelyhidi’s book [26].
z “ pz 1 , . . . , z n q,
so that
z k “ xk ` iy k , z̄ k “ xk ´ iy k , k P t1, . . . , nu,
2n 1 1 n n
and the real variables on Ω Ď R are px , y , . . . , x , y q. The inverse transformation is
1 k 1 k
xk “ pz ` z̄ k q, yk “ pz ´ z̄ k q. (1.1)
2 2i
By the chain rule, for a smooth function f : Ω Ñ C, we have
ˆ ˙ ˆ ˙
Bf 1 B B Bf 1 B B
“ ´i f, “ `i f. (1.2)
Bz 2 Bx By Bz̄ 2 Bx By
f “ pf 1 ppq, . . . , f k ppqq.
2
We say f is holomorphic if
Bf i
“0
Bz̄ k
for all i, k.
Let M be a compact complex manifold.
Ť This means that M is a smooth compact manifold admiting
a finite cover by open sets M “ i Ui with homeomorphisms
zU : U Ñ U Ď Cn , 1
zU ppq “ pzU n
ppq, . . . , zU ppqq
with the following property. For any pair pU, zU q, pV, zV q with U XV ‰ H, then we can write
zVp “ fV U p pzU q, ´1
fV U “ zV ˝ zU
p “ rZ0 : Z1 s
and rZ0 : Z1 s „ rX0 : X1 s if and only if pZ0 , Z1 q “ λpX0 , X1 q for λ P C˚ . Holomorphic charts are:
‚ U0 “ tZ0 ‰ 0u, with coordinate
Z1
z“ .
Z0
Example 1.3. Complex projective space in higher dimensions Pn “ pCn`1 zt0uq{ „ is defined
similarly. Points are denoted
p “ rZ0 : Z1 : ¨ ¨ ¨ : Zn s
and rZ0 : Z1 : ¨ ¨ ¨ : Zn s „ rX0 : X1 : ¨ ¨ ¨ : Xn s if and only if pZ0 , Z1 , . . . , Zn q “ λpX0 , X1 , . . . , Xn q
for λ P C˚ . Holomorphic charts are of the form Uk “ tZk ‰ 0u. For example, pU0 , zq has coordinate
ˆ ˙
Z1 Zn
z “ pz 1 , . . . , z n q “ ,..., ,
Z0 Z0
3
and the change of coordinates function is z̃ “ f10 pzq with
1 zk
z̃ 1 “ , z̃ k “ , k P t2, . . . , nu.
z1 z1
The coordinates on the other open sets Uk are defined similarly.
Example 1.4. Let P pZ0 , . . . , Zn q be a homogeneous polynomial of degree r, meaning P pλZ0 , . . . , λZn q “
λr P pZ0 , . . . , Zn q. Suppose that P has the property that only the point Z0 “ ¨ ¨ ¨ “ Zn “ 0 solves
BP
BZi “ 0 for all i. Then
X “ tx P Pn : P pxq “ 0u
defines a complex manifold of dimension n ´ 1. To see this, we look at tP “ 0u inside the local
charts Uk Ď Pn . For example, in coordinates pU0 , zq, the equation defining X is
0 “ f pz 1 , . . . , z n q “ P p1, z 1 , . . . , z n q.
Bf
At a point p P tf “ 0u, there must be a coordinate z i such that Bz i ‰ 0, in other words one
BP BP BP
of BZ1 , . . . BZn must be nonzero. The assumption on P is that we cannot have BZ i
“ 0 for all
BP
i P t0, . . . , nu, so we only need to rule out BZ0 ‰ 0 with all other partials zero. This is ruled out by
Euler’s identity
n
ÿ BP
Zi “ rP,
i“0
BZ i
BP BP
which shows that on tP “ 0u X U0 if BZ 0
‰ 0 then one of BZi for i ě 1 must be non-vanishing.
dˇ
ˇ dˇ
ˇ
Euler’s identity follows from dt t“1 P ptZq “ dt t“1
tr P pZq.
Bf
Without loss of generality, Bz n ‰ 0. By the holomorphic implicit function theorem, there exists a
4
Bf
or (implicit function on Bz 1 ‰ 0)
pψ̃pw̃q, w̃1 , . . . , w̃n´1 q.
The change of coordinates is
which is holomorphic.
‚ Suppose p P X X pU0 X U1 q. Then we can use coordinates coming from either U0 or U1 . As before
we denote pU0 , zq and pU1 , z̃q with z̃ 1 “ 1{z 1 and z̃ k “ z k {z 1 for k ě 2. The submanifold appears
as the equation
0 “ P p1, z 1 , z 2 , . . . , z n q
on U0 , and
0 “ P pz̃ 1 , 1, z̃ 2 , . . . , z̃ n q
BP
on U1 . Suppose BZ n
‰ 0. As before, the implicit function theorem gives coordinates wi and w̃i , so
that the equations become
0 “ P p1, w1 , w2 , . . . , ψpwqq
on U0 , and
0 “ P pw̃1 , 1, w̃2 , . . . , ψ̃pw̃qq
on U1 . The change of coordinates is
5
1.2 Holomorphic vector bundles
1.2.1 Definitions and notation
Ť
We recall the cocycle definition of a rank r complex vector bundle. Let M “ i Ui be a finite cover-
ing of open coordinate charts, together with matrix-valued functions on the nonzero overlaps
tU V : U X V Ñ GLpr, Cq
satisfying
tU V ppq “ tU W ppqtW V ppq, p P U X V X W.
We call the tU V transition functions. They satisfy tU U “ Irˆr and t´1
U V “ tV U . We define a complex
vector bundle E by ˆď ˙
r
E“ Ui ˆ C { „,
i
where the relation is as follows. For pp, uq P U ˆ Cr and pp, vq P V ˆ Cr , we identify pp, uq „ pp, vq
if
u “ tU V ppqv.
This is written using matrix notation. In terms of components, this is written
uk “ rtU V ppqsk ` v ` ,
where repeated indices are summed. Here we write v “ pv 1 , . . . , v r q and the components of the
matrix tU V are denoted tU V i j , e.g. for 2x2,
» fi » fi » fi
u1 tU V 1 1 tU V 1 2 v1
– fl “ – fl – fl .
u2 tU V 2 1 tU V 2 2 v2
The Ui ˆ Cr are the trivializations of the bundle. The projection map π : E Ñ M is given by
πpp, uq “ p.
6
We denote E|p “ π ´1 ppq to be the fiber over p, and note that E|p is a vector space of dimension r.
For two points pp, uq, pp, vq in the same trivialization Ui ˆ Cr , the vector space structure is
7
A section s P ΓpEq defines a well-defined map s : M Ñ E such that
sppq P E|p .
Indeed, in this formalism we set sppq “ pp, sU ppqq when p P U , and the condition sU “ tU V sV
ensures that if p P U X V then pp, sU ppqq „ pp, sV ppqq.
Remark 1.8. If E Ñ M is a holomorphic bundle and the sU are holomorphic functions, then we
say s is a holomorphic section and write s P H 0 pM, Eq.
Remark 1.9. In components, the transformation law is
on U X V .
There is another viewpoint on this from the perspective of basis vectors rather than vector compo-
nents. For a trivialization U ˆ Cr , let
eU r
a ppq “ pp, p0, . . . , 0, 1, 0, . . . , 0qq P U ˆ C ,
eU V b
a “ eb tV U a . (1.4)
eU V b
1 “ tU V eb tV U 1 .
which is » fi ¨» fi » fi ˛ ¨ » fi˛
1 1 0 1 ‹
˝– fl tV U 1 1 ` – fl tV U 2 1 ‚ “ rtU V s ˝rtV U s – fl‚
– fl “ rtU V s ˚ ‹ ˚
0 0 1 0
s “ sa ea .
8
In the 2 ˆ 2 case, this notation means
» fi » fi » fi
s1 ppq 1 0
– fl “ s1 ppq – fl ` s2 ppq – fl .
s2 ppq 0 1
so we simply write s “ sa ea . Indeed, substituting the transformation laws (1.3), (1.4) gives
saU eU a b V c c b V b V
a “ rtU V b sV srec tV U a s “ δ b sV ec “ sV eb .
In terms of linear algebra, this is just the statement that the same vector v will appear in different
components v a using different bases ea .
Ť
Example 1.10. Let M “ i Ui be a complex manifold with holomorphic coordinate charts pU, zq.
The holomorphic tangent bundle T 1,0 M Ñ M is the holomorphic bundle defined by transition
matrices
Bz k
tU V k i “ U .
BzVi
Sections X P ΓpT 1,0 M q are denoted
B
X “ X p pzq .
Bz p
On an overlap of coordinate charts pU, zq, pŨ , z̃q, components transform as
Bz̃ p `
X̃ p “ X
Bz `
while the basis transforms as
B Bz p B
k
“ k p.
Bz̃ Bz̃ Bz
It follows that
B B
Xp p
“ X̃ p p
Bz Bz̃
on overlaps.
Let E Ñ M , F Ñ M be two holomorphic
Ť vector bundles with transition functions tU V , t̃U V with
respect to a trivialization M “ i Ui . An isomorphism of holomorphic bundles h : E Ñ F is given
by a collection thU : U Ñ GLpr, Cqu of holomorphic invertible matrices satisfying
hU “ t̃U V hV t´1
UV . (1.5)
hppq : E|p Ñ F |p
9
Example 1.11. We return to the example M “ P1 “ U0 YU1 . There are two charts pU0 , zq, pU1 , z̃q
and z̃ “ z ´1 . Therefore
B Bz̃ B 1 B B
“ “´ 2 “ ´z̃ 2 .
Bz Bz Bz̃ z Bz̃ Bz̃
Said otherwise, a section of T 1,0 M may be written on U0 X U1 as vpzq Bz
B B
or ṽpz̃q Bz̃ with
ṽ “ ´z̃ 2 v.
so that the transition function is t10 “ ´z̃ 2 .
For example, defining BzB
over U0 extends to a global vector field V over M by setting ´z̃ 2 Bz̃
B
over
B
U1 . However, even though Bz is nowhere vanishing over U0 , this vector field must acquire a zero at
z̃ “ 0 in U1 . In component notation, this vector field V is given by the data
V “ tpU0 , vpzqq, pU1 , ṽpz̃qqu P H 0 pM, T 1,0 M q
with
vpzq “ 1, ṽpz̃q “ ´z̃ 2 .
Example 1.12. Let k P Z. Define the bundle Opkq Ñ P1 with trivializations pU0 , zq, pU1 , z̃q by
setting t10 “ z̃ k , so that sections transform as
s̃ “ z̃ k s.
The previous example, combined with (1.5) and suitable choice of hU0 , hU1 , shows that T 1,0 P1 –
Op2q. Let k ą 0.
‚ There are no holomorphic sections of Op´kq. Suppose such a section appears as a holomorphic
function spzq over the trivialization U0 . Then over U1 , that same section takes the form s̃ “ z̃ ´k s,
which in the z̃ coordinates belonging to U1 is
s̃pz̃q “ z̃ ´k spz̃ ´1 q.
ř8
Writing spzq “ i“0 ai z i , we see that s̃pz̃q must have a pole and cannot be holomorphic.
‚ Holomorphic sections of Opkq correspond to homogeneous polynomials P pZ0 , Z1 q of degree k: any
section σ P H 0 pP1 , Opkqq is σ “ tpU0 , sq, pU1 , s̃qu locally the form s “ P pZ0 , Z1 q{Z0k over U0 , and
of the form s̃ “ P pZ0 , Z1 q{Z1k over U1 . Indeed, let σ P H 0 pP1 , Opkqq be an arbitrary holomorphic
section. Then spzq, s̃pz̃q are both holomorphic and
s̃pz̃q “ z̃ k spz̃ ´1 q. (1.6)
ř8 ř8
After writing s̃ “ k“0 bk z̃ k , s “ k“0 ak z k and comparing coefficients, we see that spzq “
a0 ` a1 z ` ¨ ¨ ¨ ` ak z k . It follows that
„
1 k k´1 k
s “ k a0 Z0 ` a1 Z0 Z1 ` ¨ ¨ ¨ ` ak Z1
Z0
since on U0 “ tZ0 ‰ 0u the coordinate is z “ Z1 {Z0 . The transformation (1.6) implies
„
1 k k´1 k
s̃ “ k a0 Z0 ` a1 Z0 Z1 ` ¨ ¨ ¨ ` ak Z1
Z1
since on U1 “ tZ1 ‰ 0u the coordinate is z̃ “ Z0 {Z1 . Therefore the homogeneous polynomial
corresponding to this section is P “ a0 Z0k ` a1 Z0k´1 Z1 ` ¨ ¨ ¨ ` ak Z1k .
10
Example 1.13. In higher dimensional projective space, define Opkq Ñ Pn by pUi , tij q where
Ui “ tZi ‰ 0u and tij : Ui X Uj Ñ C˚ is
ˆ ˙k
Zj
tij “
Zi
For example on Op1q Ñ P2 , with coordinates pU0 , zq, pU1 , z̃q with z “ pZ1 {Z0 , Z2 {Z0 q and z̃ “
pZ0 {Z1 , Z2 {Z1 q, then t10 “ Z0 {Z1 “ z̃ 1 .
siU “ tU V i k skV , ϕU V k
i “ ϕk tV U i .
This is the dual bundle because sections s P ΓpM, Eq and ϕ P ΓpM, E ˚ q can be paired together to
form a function
ϕpsq :“ pϕi si q P C 8 pM, Rq.
This is because the transformation laws imply
ϕU i V i
i sU “ ϕi sV
and so pϕi si qppq is independent of the choice of trivialization. In matrix notation Q “ rtU V s, the
transformation laws for s P ΓpEq and ϕ P ΓpE ˚ q are
In terms of local frames, if tei u is a local frame for E, we denote the corresponding dual frame on
E ˚ by tei u. This is defined as ei pej q “ δ i j , and a section ϕ P ΓpE ˚ q is written as
ϕ “ ϕi ei .
The pairing ϕpsq can then be seen by the formula for the dual frame: ϕpsq “ pϕi ei qpsk ek q “
ϕj sj .
‚ Determinant bundle. The line bundle det E Ñ M is defined by the trivializations det tU V .
‚ Tensor product. If E Ñ M , Ẽ Ñ M are vector bundles, then the bundle E b Ẽ Ñ M has
trivializations tU V b t̃U V . In components, if indices i, j denotes indices on E and indices α, β
indices on Ẽ, then
siα i α
U “ tU V j t̃U V β sV
jβ
.
11
‚ Endomorphism bundle. We will later encounter sections of E ˚ b E “ End E ˚ , and our convention
for h P ΓpEnd E ˚ q will be
h “ hα β eα b eβ
so that the transformation law for components reads
rhU sα β “ tV U µ α rhV sµ ν tU V β ν ,
Note that h defines a map h|p : E ˚ |p Ñ E ˚ |p by hα β ϕβ . Verifying that this map is well-defined is
a similar calculation as (1.5).
In fact, h P ΓpEnd E ˚ q also defines an endomorphism of E by acting on the right as uT h, or in
index notation uα hα β . That uT h transforms like a section follows from
Thus uT h P ΓpEq. Thus, we will sometimes view h P ΓpEnd E ˚ q with h “ hα β as h P ΓpEnd Eq.
‚ Divisor bundle. Let Y Ă X be an analytic hypersurface. This means that near each p P Y , there
is neighborhood U such that U X Y is locally given by the vanishing set of a holomorphic function.
The theory of holomorphic functions (see e.g. [13]) implies that there exists the notion of a local
defining function: this means that f is holomorphic with
U X Y “ tf “ 0u
and any other local holomorphic function g vanishing on Y factors as gpzq “ hpzqf pzq with h a local
holomorphic function. The notion of local defining function is not unique: if f1 and f2 are local
defining functions, then f1 “ hf2 where h is a holomorphic function non-vanishing on U .
We can associate a line bundle OpY q Ñ X in the following way. In a coordinate chart U , the
submanifold Y appears as Y X U “ tfU pzq “ 0u where fU pzq is a local holomorphic function. The
transition function of OpY q is given by tU V “ fU {fV on U X V . If Y X U “ H, we can take
fU “ 1.
‚ Note: if another choice of local defining function is taken, by (1.5) it follows that this defines an
isomorphic bundle.
‚ Note: there is a global section s P H 0 pOpY qq given by the local data pU, sU q with sU “ fU , since
sU “ tU V sV is tautology.
Example 1.14. Let P pZ0 , . . . , Zn q be a homogeneous polynomial of degree k, and let Y “ tP “
0u Ă Pn . Then OpY q “ Opkq. To see this, in the local chart U0 Ă Pn the equation in coordinates
pU0 , zq is 0 “ P pZ0 , . . . , Zn q{Z0k “ s0 and in the local chart U1 the equation in coordinates pU1 , z̃q
is 0 “ P pZ0 , . . . , Zn q{Z1k “ s1 . The transition function t10 is then
„ k
P pZ0 , . . . , Zn q{Z1k Z0
t10 “ “
P pZ0 , . . . , Zn q{Z0k Z1
which matches with the transition functions of Opkq.
12
Let Y Ď X be a smooth analytic hypersurface. This means that at p P U , the local defining function
U X Y “ tf pzq “ 0u has the property that Bi f ppq ‰ 0 for some coordinate direction Bi . In this case,
there exists new holomorphic local coordinates tz̃ i u such that (after possibly shrinking U )
U X Y “ tz̃ n “ 0u.
To see this, let tz i u be the original holomorphic coordinates and suppose after relabeling that
Bf
Bz n ppq ‰ 0. By the holomorphic implicit function theorem, after possibly shrinking U we have
13
1.2.3 Constructions from the tangent bundle
We now list some bundle constructions which come from the holomorphic tangent bundle T 1,0 M .
‚ Complexified tangent bundle. The conjugate of T 1,0 M is denoted T 0,1 M :“ T 1,0 M . A local
frame is given by " *
B B
, . . . , ,
Bz̄ 1 Bz̄ n
and sections denoted V ī Bz̄B i . We can write the px, yq coordinate basis in terms of the pz, z̄q coordinate
basis by the change of variables z k “ xk ` iy k , z̄ k “ xk ´ iy k and the chain rule:
„
B B B B B B
“ ` , “ i ´ .
Bxk Bz k Bz̄ k By k Bz k Bz̄ k
It follows that the complexified tangent bundle TC M can be written as a direct sum
TC M “ T 1,0 M ‘ T 0,1 M.
Example 1.16. We will sometimes use the notion of a complex structure J : TC M Ñ TC M . Given
a complex manifold, the complex structure J is defined by setting
J|T 1,0 M “ `iId, J|T 0,1 M “ ´iId,
or in other words, J BzBk “ i BzBk and J Bz̄Bk “ ´i Bz̄Bk . In components,
J p q “ iδ p q , J p̄ q̄ “ ´iδ p q , J p q̄ “ J p̄ q “ 0,
in complex coordinates.
‚ Complexified cotangent bundle. We denote smooth sections of the dual of the holomorphic cotan-
gent bundle pT 1,0 M q˚ by Λ1,0 pM q, and holomorphic sections of this bundle by H 0 pM, pT 1,0 M q˚ q.
A local frame is given by
tdz 1 , . . . , dz n u
meaning dz k pBzi q “ δ k j , and so that a section α P Λ1,0 pM q is written α “ αi dz i and αpV q “ αi V i
for V “ V i Bzi P ΓpM, T 1,0 M q. Transformation laws for components and frames are
Bz p Bz̃ k i
α̃i “ αp , dz̃ k “ dz .
Bz̃ i Bz i
Denote Λ0,1 pM q “ Λ1,0 pM q. Complexified 1-forms Λ1C M can be decomposed as
Λ1C “ Λ1,0 pM q ‘ Λ0,1 pM q,
from the decompositions (1.1)
1 1
dxk “ pdz k ` dz̄ k q, dy k “ pdz k ´ dz̄ k q. (1.10)
2 2i
‚ Differential forms. Let z k “ xk `iy k be local complex coordinates, and write w “ px1 , . . . , xn , y 1 , . . . , y n q P
R2n . A differential form on M appears in w coordinates as
1
η“ ηi ¨¨¨i dwi1 ^ ¨ ¨ ¨ ^ dwik .
k! 1 k
14
From (1.10), we see that this can be written in the complex basis of dz, dz̄. We will use the following
convention for complex components: ÿ
η“ η p,q
p`q“k
with
1
η p,q “ η dz i1 ^ ¨ ¨ ¨ ^ dz ip ^ dz̄ j1 ^ ¨ ¨ ¨ ^ dz̄ jq .
p!q! i1 ¨¨¨ip j̄1 ¨¨¨j̄q
We call η p,q a pp, qq-form, denoted Λp,q pM q.
‚ Exterior derivative. The exterior derivative acting on a function is
Bf i Bf
df “ i
dx ` i dy i
Bx By
which in complex coordinates becomes
Bf i Bf
df “ dz ` i dz̄ i .
Bz i Bz̄
We write this as df “ Bf ` B̄f , with
Bf i Bf i
Bf “ dz , B̄f “ dz̄ .
Bz i Bz̄ i
Similarly, the exterior derivative d : Λk Ñ Λk`1 on higher differential forms decomposes into
types.
d “ B ` B̄.
Acting on χ P Λp,q pM q, we have
1 B
Bχ “ χ dz ` ^ dz i1 ^ ¨ ¨ ¨ ^ dz ip ^ dz̄ j1 ^ ¨ ¨ ¨ ^ dz̄ jq
p!q! Bz ` i1 ¨¨¨ip j̄1 ¨¨¨j̄q
and
1 B
B̄χ “ χ dz̄ ` ^ dz i1 ^ ¨ ¨ ¨ ^ dz ip ^ dz̄ j1 ^ ¨ ¨ ¨ ^ dz̄ jq
p!q! Bz̄ ` i1 ¨¨¨ip j̄1 ¨¨¨j̄q
so that B : Λp,q Ñ Λp`1,q and B̄ : Λp,q Ñ Λp,q`1 .
Example 1.17. For α P Λ1,1 , we write α “ αj k̄ dz j ^ dz̄ k and
Bα “ B` αj k̄ dz ` ^ dz j ^ dz̄ k
1
“ pB` αj k̄ ´ Bj α`k̄ qdz ` ^ dz j ^ dz̄ k
2
1
“ pBαq`j k̄ dz ` ^ dz j ^ dz̄ k . (1.11)
2
and the components formula is
pBαq`j k̄ “ B` αj k̄ ´ Bj α`k̄ .
15
1.3 Geometry of bundles
1.3.1 Chern connection
Let E Ñ M be a holomorphic vector bundle of rank r over a complex manifold. A hermitian metric
on E is H P ΓpE ˚ b Ē ˚ q which is represented in a local frame teα u of E by
H “ Hαβ̄ eα b eβ
with Hαβ̄ ppq a positive-define r ˆ r hermitian matrix at all points p. The hermitian condition is
Hαβ̄ “ Hβ ᾱ . Our conventions for the inner product on E given by H is
so that xu, λvy “ λ̄xu, vy. The hermitian condition is xu, vy “ xv, uy. The norm of a section is
|u|2 “ xu, uy.
We note that xu, vy does not depend on the choice of trivialization. Let U, Ũ be two trivializations
of E with transition matrix rQs “ Qα β and denote components on Ũ with tildes, so for u P ΓpEq
and ϕ P ΓpE ˚ q we have
ũα “ Qα β uβ , ϕ̃α “ ϕα pQ´1 qα β
and the transformation law on H P ΓpE ˚ b Ē ˚ q is
From here we can verify ũα H̃αβ̄ ṽ β “ uα Hαβ̄ v β . This can also be written using matrix nota-
tion:
xu, vy “ uT H v̄,
ũ “ Qu, H̃ “ pQ´1 qT HQ´1 , (1.12)
and it is straightforward to verify that ũT H̃ ṽ “ uT H v̄. In other words, though the direction of uα
as a column vector is not a well-defined quantity (depends on the choice of trivialization), its norm
|u|H is a measurable number.
The inverse of H is denoted in components as H ᾱβ , so that HH ´1 “ I becomes in components
Hαβ̄ H β̄γ “ δα γ . The inverse H ´1 produces a metric on E ‹ .
Similarly as above, it can be verified that xψU , ϕU y “ xψŨ , ϕŨ y, so that xψ, ϕy takes two sections
and produces a global function on X.
The metric can be used to raise and lower indices. From uα P ΓpEq, we will write
uβ̄ “ uα Hαβ̄ ,
and uβ̄ defines a section of Ē ˚ . This is because if u ÞÑ Qu and H ÞÑ pQ´1 qT HQ´1 , then uT H ÞÑ
uT HQ´1 “ pQ´1 qT puT Hq. Said another way, given u P ΓpEq and a metric H, we obtain a dual
element u˚ P ΓpĒ ˚ q defined by
u˚ pv̄q “ xu, vyH .
16
Similarly, from uα P ΓpĒ ˚ q, then uα “ H αβ̄ uβ̄ is a section of ΓpEq. We note that
uα v α “ uᾱ vᾱ .
∇k̄ uα “ Bk̄ uα ,
∇k uα “ Bk uα ` uβ Akβ α , Akβ α “ Bk Hβ ν̄ H ν̄α , (1.14)
or without indices as
∇ “ pB ` BHH ´1 q ` B̄.
For ∇k̄ sα to be a section, we need to verify that if pŨ , s̃α q and pU, sα q are two overlapping trivial-
izations of E with s̃α “ Qα β sβ , then
This is true because B̄Qα β “ 0. It can also be checked directly that ∇k sα is a section, namely
∇k s̃α “ Qα β ∇k sβ ,
Proof. Let ∇ be a connection satisfying (1.15) with ∇0,1 “ B̄. We will solve for ∇1,0 . Our notation
for the unknown connection is
∇Bk eα “ Akα β eβ
where A are unknown coefficients to be solved. In other words
∇k uα “ Bk uα ` Akβ α uβ .
17
which simplies to
Bk Hij̄ ui v j “ Hrj̄ Aki r ui v j
If this is true for all sections u, v, then
Bk Hij̄ “ Aki r Hrj̄
and solving for A gives
Aki ` “ Bk Hij̄ H j̄`
or A “ BHH ´1 .
There is a formula for how the Chern connection changes when changing the metric. Let Ĥ and H
be two metrics on E. Let A “ BHH ´1 and h “ H Ĥ ´1 . Then
A “ BHH ´1
“ BphĤqĤ ´1 h´1
“ Bhh´1 ` hÂh´1
“ Bhh´1 ` hÂh´1 ` p ´ Âhh´1 q
“ ˆ ´1
 ` ∇hh (1.16)
ˆ “ Bh ` h ´ Âh in matrix notation, or using index notation for the component of
where ∇h
hα β “ Hαµ̄ Ĥ µ̄β , then
∇i hα β “ Bi hα β ` hα γ Aiγ β ´ Aiα γ hγ β . (1.17)
Here is the reason for the term with a minus sign. Let ∇ be a connection on E acting on sections
u P ΓpEq by
∇i uα “ Bi uα ` uβ Aiβ α , Aiβ α “ Bi Hβ ν̄ H ν̄α .
The induced dual connection acting on ϕ P ΓpE ˚ q is defined with a minus sign:
∇i ϕα “ Bi ϕα ´ Aiα β ϕβ .
This minus sign is introduced so that we can differentiate contracted indices using the product
rule
Bi puα ϕα q “ p∇i uα qϕα ` uα p∇i ϕα q.
The formula (1.17) follows from the rule for covariant differentiation where each upper index receives
a `A term and each lower index receives a ´A term. As another example,
∇i T α βγ “ Bi T α βγ ` T µ βγ Aiµ α ´ T α µγ Aiβ µ ´ T α βµ Aiγ µ ,
for T P ΓpE b E ˚ b E ˚ q. Sections of Ē follow the rules
∇i uᾱ “ Bi uᾱ , ∇ī uᾱ “ Bī uᾱ ` uν̄ Aiν α
¯ for u P ΓpEq. For example,
so that ∇ū “ ∇u
∇i Gαβ̄ “ Bi Gαβ̄ ´ Aiα ν Gν β̄
for G P ΓpE ˚ b Ē ˚ q. From this formula and Ai “ pBi HqH ´1 , we see that
∇i Hαβ̄ “ 0, ∇ī Hαβ̄ “ 0
when ∇ is the Chern connection of H.
18
1.3.2 Curvature
The curvature of the Chern connection defines a notion of Hessian of a metric tensor H. The
local combinations Bj Bk̄ Hαβ̄ do not transform as the section of any bundle, and taking covariant
derivatives gives zero: ∇i Hαβ̄ “ 0. The curvature tensor is a way to encode second derivatives of
the metric.
Definition 1.20. Let pE, Hq be a holomorphic vector bundle with metric. The curvature of the
Chern connection F P ΓpΛ1,1 b End Eq is given by F “ B̄pBHH ´1 q. In components
F “ Fβ α j k̄ eα b eβ dz j ^ dz̄ k ,
the definition is
Fα β j k̄ “ ´Bk̄ pBj Hαµ̄ H µ̄β q,
or
Fj k̄ “ ´Bk̄ pBj HH ´1 q,
without showing the endomorphism indices.
The action of Fj k̄ P ΓpEnd Eq on u P ΓpEq is
uα ÞÑ uβ Fβ α j k̄ .
We now verify that the formula for Fj k̄ gives a well-defined section of End E. The transformation
law H̃ “ pQ´1 qT HQ´1 implies
„
F̃j k̄ “ ´Bk̄ Bj ppQ´1 qT HQ´1 qpQH ´1 QT q
„ „
“ ´Bk̄ p´QT Bj QT q ` ´Bk̄ ppQ´1 qT Bj HH ´1 QT q
„
“ pQ q ´ Bk̄ pBj HH q QT
´1 T ´1
(1.18)
using B̄Q “ 0, B Q̄ “ 0. This matches with (1.7), and so Fj k̄ P ΓpEnd Eq acting on sections of E on
the right.
Remark 1.21. Let L Ñ M be a holomorphic line bundle. Then H is a 1 ˆ 1 matrix, and the
transformation law for a metric reads
1
H̃ “ H. (1.19)
|tŨ U |2
19
Remark 1.22. The formula for change of curvature is
ˆ ´1 q.
F “ F̂ ` B̄p∇hh
Remark 1.23. The formula for Chern curvature is consistent with the general formula for the
curvature of a connection. In general, the curvature of a connection ∇ on E is F “ dA ´ A ^ A,
F “ 12 Fj i µν dxµ ^ dxν b ej b ei , with
For the Chern connection, ∇0,1 “ B̄, so Aīα β “ 0. Therefore Fα β k̄j̄ “ 0, and since the Chern
connection is unitary then Fα β kj “ 0 which can also be checked directly. Therefore F only has
mixed p1, 1q form indices, and
Fj i mn̄ “ ´Bn̄ Amj i
since mixed connection terms are zero.
Example 1.24. Fubini-Study metric on Op1q Ñ P1 . Recall that P1 is covered by two trivializations
pU0 , zq, pU1 , z̃q, with change of coordinates z̃ “ z ´1 and sections s P ΓpOp1qq transform as s̃ “ z̃s.
The Fubini-Study metric is defined as
This transforms correctly as (1.19): h̃ “ p1{|z̃|2 qh. In other words, the norm
|s|2h “ ss̄h
gives the same result in either trivialization. The curvature is a 2-form iF P Λ1,1 with components
We compute
z̄ 1 |z|2
Fzz̄ “ Bk̄ “ ´ “ p1 ` |z|2 q´2 ą 0.
1 ` |z|2 1 ` |z|2 p1 ` |z|2 q2
Therefore iF is a closed positive p1, 1q form; this is a Kähler metric on P1 . The Fubini-Study Kähler
metric is sometimes denoted
ωF S “ iB B̄ logp1 ` |z|2 q.
g “ gj k̄ dz j b dz̄ k .
20
The collection of local matrices pU, gj k̄ q are related by
Bz ` Bz m
g̃j k̄ “ j
g`m̄ k .
Bz̃ Bz̃
Locally we view g as a matrix, e.g. in 2 dimensions
» fi
g11̄ g12̄
g“– fl .
g21̄ g22̄
where X “ X i BzB i ` X ī Bz̄B i . From this perspective, the p1, 1q-form ω may be defined as ωpX, Y q “
gR pJX, Y q since ωpBj , Bk̄ q “ igj k̄ .
We can also write the metric g (we now stop using the notation gR ) in terms of q “ px1 , y 1 , . . . , xn , y n q
coordinates where z k “ xk ` iy k . We give the details in complex dimension 1. Let z “ x ` iy be
complex coordinates, and we would like to write
» fi
gxx gxy
g“– fl
gyx gyy
21
A hermitian metric is given by a mixed type gzz̄ ą 0, and we declare gzz “ 0 and gz̄z̄ “ 0. Changing
coordinates
Note: there is a way to go from a Riemannian metric on T M to a hermitian metric on T 1,0 M , but
this requires that g is compatible with the complex structure so that gpJX, JXq “ gpX, Xq and we
omit the details.
We will use a normalization factor of p!q! for the inner product defined on pp, qq forms. For ϕ, ψ P
Λp,q with
1
ϕ“ “ ϕα1 ¨¨¨αp β̄1 ¨¨¨β̄q dz α1 ^ ¨ ¨ ¨ ^ dz αp ^ dz̄ β1 ^ ¨ ¨ ¨ ^ dz̄ βq ,
p!q!
and
ϕ̄α1 ¨¨¨αp β̄1 ...β̄q “ g ā1 α1 ¨ ¨ ¨ g āp αp g β̄1 b1 ¨ ¨ ¨ g β̄q bq ϕa1 ¨¨¨ap b̄1 ¨¨¨b̄q
then we will use the convention
1
xϕ, ψy “ ϕ ψ̄ α1 ¨¨¨αp β̄1 ...β̄q . (1.22)
p!q! α1 ¨¨¨αp β̄1 ¨¨¨β̄q
22
using the variation formula for the inverse of a matrix δA´1 “ ´A´1 pδAqA´1 . The conjugate of
this is
Rj p̄mk̄ “ ´Bm̄ Bk gpj̄ ` pBk gpā qg āq pBm̄ gqj̄ q
so
Rj p̄mk̄ “ Rpj̄km̄ , Rj i mk̄ “ g īp Rpj̄km̄ “ Rī j̄km̄ .
The curvature appears when exchanging covariant derivatives. We will write
where ∇m ∇k̄ V i means: let Wk̄ i “ ∇k̄ V i be a tensor and compute the components ∇m Wk̄ i . The
commutator formula is
r∇m , ∇k̄ sV i “ V p Rp i mk̄ . (1.23)
Here is the check:
Remark 1.26. A similar calculation gives that the commutator r∇m , ∇k sV i involves another
tensor, the torsion tensor, but we will not need this formula.
Taking the conjugate of (1.23) gives
Similarly
r∇k , ∇m̄ sVā “ Rp̄ āmk̄ Vp̄ .
Let
Rj k̄ “ Rp p j k̄ “ g q̄p Rpq̄j k̄
and define the Chern-Ricci form by
Viewing R P Λ1,1 pEnd T 1,0 Xq, we have iRicω “ iTr R P Λ1,1 pXq where we trace out the endomor-
phism indices and retain the 2-form indices. Using the general formula for the derivative of the
determinant of an invertible hermitian matrix A,
ˇ
d ˇˇ
det Aptq “ det Ap0q Tr rAp0q´1 Ap0qs
9
dt ˇt“0
23
and
iRicω “ ´iB B̄ log det g.
This expression can also be connected to Riemannian geometry: in the case when gij̄ is a Kähler
metric, then it turns out that Rj k̄ is the Levi-Civita Ricci tensor of the Riemannian metric g on T X.
This calculation can be found in Kähler’s original paper (see p.178 in [17], where (1.25) is described
as “very elegant”), and it is one of the main motivations for the field of Kähler geometry.
We note that when g is a general hermitian metric on T 1,0 X, the Chern-Ricci curvature Rj k̄ (1.25)
is different than the Riemannian Levi-Civita Ricci tensor.
Remark 1.27. From the point of view of Riemannian geometry, it looks like Rj k̄ “ Rp p j k̄ is
tracing the wrong index in the definition of the Ricci curvature and should be zero. This is not the
case because we are only tracing over holomorphic indices. Tracing over all real indices does indeed
give zero. Let α, β represents real coordinates xα “ pz 1 , . . . , z n , z̄ 1 , . . . , z̄ n q and a, b, j, k represents
holomorphic coordinates z i , and let Rαβγµ be the Riemannian curvature tensor. Then
since the Riemannian curvature tensor satisfies Rαβγµ “ ´Rβαγµ . But the trace
g ab̄ Rab̄j k̄
is not summing over all coordinate indices α, β, and need not be zero.
Example 1.28. We start with the line bundle Op1q Ñ Pn with trivializations Ui “ tZi ‰ 0u.
Recall that the transition functions pUi X Uj , tij q are tij “ Zj {Zi .
‚ Show that the collection pUi , hi q with
|Zi |2
hi “ ř 2
k |Zk |
as the Fubini-Study metric on Pn . We see that dωF S “ 0, and to show ωF S is a Kähler metric, we
need to verify that gj k̄ is a positive-definite matrix. A computation gives
24
For ξ P Cn , we have
25
2 Kähler Manifolds
This section will follow the textbook by Kodaira-Morrow [20].
26
Here I “ i1 ¨ ¨ ¨ ip , J “ j1 ¨ ¨ ¨ jq are multi-indices, and we raise indices using the metric: e.g. if
1
ϕ “ q! ϕJ¯dz̄ J is p0, qq form, then ϕJ “ g j1 ā1 ¨ ¨ ¨ g jq āq ϕā1 ¨¨¨āq , or e.g. if ϕ “ ϕrs̄ dz r ^ dz̄ s is a
p1, 1q-form, then ϕīj “ g īr g s̄j ϕrs̄ .
Let us verify formula (2.1) in the special case of p0, 2q forms. Let θ P Λ0,2 pXq and ϕ P Λ0,1 pXq.
Then
ϕ “ ϕī dz̄ i , pB̄ϕqj̄ k̄ “ Bj̄ ϕk̄ ´ Bk̄ ϕj̄ .
Using the inner product (1.22) we have
1 j̄i k̄`
xB̄ϕ, θyg “ g g pBj̄ ϕk̄ ´ Bk̄ ϕj̄ qθī`¯ “ Bj̄ ϕk̄ θ̄j̄ k̄
2
and
Bj̄ ϕk̄ θ̄j̄ k̄ pdet gq “ ´ϕk̄ Bj̄ pθ̄j̄ k̄ pdet gqq ` Bj̄ rϕk̄ pθ̄j̄ k̄ pdet gqqs.
Let W k̄ “ pdet gq´1 Bj̄ ppdet gqθ̄j̄ k̄ q. Multiplying by idz 1 ^ dz̄ 1 ¨ ¨ ¨ ^ idz n ^ dz̄ n and using (1.20)
gives
ωn ωn ωn
xB̄ϕ, θy “ ´ϕk̄ W k̄ ` dιV .
n! n! n!
The last term involves V j̄ “ ϕk̄ θj̄ k̄ and will be explained below. Integrating this identity over X
and applying Stokes’s theorem gives
ωn
ż
pB̄ϕ, θqL2 “ ´ pϕk̄ W k̄ q .
X n!
We compare this with the definition of the adjoint:
k̄ ω n
ż
pB̄ϕ, θqL2 “ pϕk̄ pB̄ : θq q “ pϕ, B̄ : θqL2 .
X n!
Thus pB̄ : θqk “ ´W̄ k , and since
W̄ k “ pdet gq´1 Bp ppdet gqθpk q
this is the formula (2.1). We now explain why
ωn
Bj̄ rV j̄ pdet gqs idz 1 ^ dz̄ 1 ¨ ¨ ¨ ^ idz n ^ dz̄ n “ dιV .
n!
First, we recall the definition of the interior product: if V is a vector field, then ιV : Λk Ñ Λk´1
via
pιV ηqpW1 , . . . , Wk´1 q “ V i ηpBi , W1 , . . . , Wk´1 q,
and it satisfies ιV pη1 ^ η2 q “ ιV η1 ^ η2 ` p´1qk η1 ^ ιV η2 if η1 P Λk . Therefore applying ιV to (1.20)
gives
ωn
„
1̄ 1 2 2 2̄ 1 1 2
´ιV “ det g V pidz q ^ pidz ^ dz̄ q ^ p¨ ¨ ¨ q ` V pidz ^ dz̄ q ^ pidz q ^ p¨ ¨ ¨ q ` . . .
n!
and d “ B ` B̄ becomes
ωn
dιV “ Bk̄ rpdet gqV k̄ s pidz 1 ^ dz̄ 1 q ^ ¨ ¨ ¨ ^ pidz n ^ dz̄ n q,
n!
27
as claimed.
The B̄-Laplacian can be studied using techniques from the theory of elliptic PDE. We now give the
general definition of an elliptic operator.
Definition 2.1. Let E, F Ñ X be vector bundles trivialized by a finite cover X “ YN
i“1 Ui . An
elliptic operator of order k is a map L : ΓpEq Ñ ΓpF q such that:
‚ In each trivialization Ui Ă X, L appears as
ÿ ÿ
pLuqα “ AIα β BI uβ ` B Iα β BI uβ .
|I|“k 0ď|I|ăk
σpL, ξqppq : Ep Ñ Fp
and therefore
σp∆, ξq “ p´g j̄i ξi ξj̄ q id “ ´|ξ|2g id.
Here we write ξ “ ξi dz i ` ξī dz̄ i for ξ P Λ1 pM, Rq, and since ξ is real then ξ¯ “ ξ and ξj̄ “ ξj .
We verify (2.2) for p0, 1q forms ϕ “ ϕī dz̄ i . By using the expression for the adjoint (2.1) and
pB̄ϕqj̄ k̄ “ Bj̄ ϕk̄ ´ Bk̄ ϕj̄ , we compute
and
„
: ´1 pβ
pB̄ B̄ϕqᾱ “ ´pdet gq Bp pdet gqpB̄ϕq gβ ᾱ
„
´1 q̄p ν̄β
“ ´pdet gq Bp pg g det gqpBq̄ ϕν̄ ´ Bν̄ ϕq̄ q gβ ᾱ . (2.4)
28
We say that an elliptic operator L is self-adjoint if
ΓpEq “ ker L ‘ Im L.
Applying this to ∆B̄ , we see that we can write any η P Λp,q pXq as
η “ h ` pB̄ B̄ : ` B̄ : B̄qβ,
η “ h ` B̄β1 ` B̄ : β2 .
η “ h ` B̄β1 .
This is because pB̄ : β2 , B̄ : β2 q “ pβ2 , B̄ηq “ 0. It follows that rηs “ rhs for h P ker ∆B̄ . For uniqueness,
suppose η “ h1 ` B̄β1 “ h̃1 ` B̄ β̃1 . Then
and so
0 “ ph1 ´ h̃1 , h1 ´ h˜1 qL2 ` pB̄pβ1 ´ β̃1 q, h1 ´ h˜1 qL2
and }h1 ´ h̃1 }2L2 “ 0 since B̄ : ph1 ´ h̃1 q “ 0.
29
There is a similar theory for vector bundle valued pp, qq forms, and in general
Theorem 2.5. (Serre duality) Let E Ñ X be a holomorphic vector bundle over a compact complex
manifold. Then
dim HB̄p,q pEq “ dim HB̄n´p,n´q pE ˚ q,
which implies that
dim H q pX, Eq “ dim H n´q pX, KX b E ˚ q. (2.6)
p,q n´p,n´q
and h pXq “ h pXq.
Proof. We will use the Hodge star operator. This is a linear map
‹ : Λp,q Ñ Λn´q,n´p
30
This can be verified by Stokes’s theorem and substitution of (2.9) into the defining relation pB̄ϕ, ψq “
pϕ, B̄ : ψq. Indeed, for ϕ P Λp,q pEq and ψ P Λp,q`1 pEq, then
ż
pϕ, B̄ : ψq “ ϕα ^ ‹pB̄ : ψqβ Hαβ̄
ż ˆ ˙
α 2 β̄µ ν̄
“ ´ ϕ ^ ‹ H B̄pHµν̄ ‹ ψ̄ q Hαβ̄
ż
n´p n´q
“ p´1qp´1q p´1q ϕα ^ B̄pHαν̄ ‹ ψ̄ ν̄ q
ż
“ p´1qp´1qn´p p´1qn´q p´1qp´1qp`q B̄ϕα ^ ‹ψ̄ ν̄ Hαν̄
Therefore
HB̄p,q pEq “ tη P Λp,q pEq : B̄η “ 0, B̄ : η “ 0u
can be written in this notation as
It follows that
η ÞÑ #η
is a map from H p,q pEq to H n´p,n´q pE ˚ q, and this is an isomorphism.
Note that here we used # because this would not have worked using ‹ : Λp,q pXq Ñ Λn´q,n´p pXq,
since ∆B̄ η “ 0 if and only if B̄η “ 0 and B ‹ η “ 0, and so if ∆B̄ η “ 0 then it is not necessarily true
that ∆B̄ ‹ η “ 0.
On a Kähler manifold, there are the following symmetries for the Hodge numbers.
Theorem 2.6. Let X be a compact Kähler manifold. Then
à
H k pX, Cq “ H p,q pXq, hp,q “ hq,p , hn´p,n´q “ hp,q .
p`q“k
31
For example,
Here
kerpd : Λq pXq Ñ Λq`1 pXqq
H q pX, Cq “
im pd : Λq´1 pXq Ñ Λq pXqq
are the deRham cohomology groups, and Λk pXq denotes differential k-forms with coefficients in C.
The Laplacians are
∆d “ dd: ` d: d, ∆B “ ∆B̄ “ BB : ` B : B,
where d: , B : are the L2 adjoints of d, B. These Laplacians are elliptic operators and satisfy the
Hodge decomposition Theorem 2.3, and it follows from elliptic PDE theory that each de Rham
class rηs P H k pX, Cq admits a unique representative h P rηs with ∆d h “ 0.
H k pX, Cq “ tη P Λk : ∆d η “ 0u
p,q
H pXq “ tη P Λp,q : ∆B̄ η “ 0u, (2.14)
we can decompose η P Λk as η “ p`q“k η p,q . Since ∆B̄ preserves type, we have ∆d η “ 2∆B̄ η p,q .
ř ř
p`q“k
Therefore if η P Λk with ∆d η “ 0, then
ÿ
η ÞÑ η p,q
p`q“k
Therefore η ÞÑ η̄ is a map from H p,q pXq to H q,p pXq and this is an isomorphism.
hk,k pPn q “ 1
32
We now show ∆B̄ “ ∆B “ 12 ∆d . First, we note the following Kähler identities:
33
and
∆d “ pdd: ` d: dq
“ pB ` B̄qpB : ` B̄ : q ` pB : ` B̄ : qpB ` B̄q
“ ∆B ` ∆B̄ ` pB B̄ : ` B̄ : Bq ` pB̄B : ` B : B̄q
“ 2∆B̄ . (2.20)
Proof. We prove the first statement, and the second statement has a similar proof. Using the Hodge
decomposition (Theorem 2.3) for ∆B ,
α “ Bα1 ` B : α2 ` α1 , α1 P ker ∆B .
Substituting gives
α “ B B̄β1 ` B B̄ : β2 .
To remove the last term, we must use the Kähler identity (2.18) which reads B B̄ : “ ´B̄ : B. Using
this and B̄α “ 0, we see that
0 “ B̄ B̄ : Bβ2 .
It follows that pB̄ : Bβ2 , B̄ : Bβ2 q “ 0.
Fj k̄ v j v k ě εgj k̄ v j v k
for all v P Cn .
34
Example 2.9. The main example of a positive line bundle is Op1q Ñ Pn , as
|Zi |2
phF S qi “ ř 2
, over Ui “ tZi ‰ 0u
p |Zp |
|si |2
hi “ ϕ˚ hF S “ ř 2
.
k |sk |
Then ´iB B̄ log h “ ϕ˚ p´iB B̄ log hF S q ą 0. Therefore ample line bundles are positive. Kodaira’s
embedding theorem (e.g. [20]) states that positive line bundles are ample.
Theorem 2.12. Let L Ñ pM, ωq be a positive holomorphic line bundle over a compact Kähler
manifold. Then
H q pX, L b KX q “ 0
for all q ě 1.
We will show that dim H q pX, L b KX q “ dim ker ∆B̄ |Λ0,q pLbKX q “ 0. We will give the proof for
q “ 1 for simplicity. For the general calculation, see e.g. [20].
Let h be a metric on L, so that the inner product on sections u, v P ΓpLq is xu, vyh “ uvh. Let
ϕ P Λ0,1 pLq, which we write as
ϕ “ ϕk̄ dz̄ k ,
where ϕk̄ is a local section of L. The L2 inner product for u, v P ΓpLq is
ωn
ż
pu, vq “ puv̄hq
X n!
35
We will verify this for ϕ P Λ0,1 pLq. The definition of the adjoint is
We start with
xϕ, B̄uyg,h “ g īk pϕī Bk̄ uhq “ ϕk Bk uh
which implies
xϕ, B̄uyg,h pdet gq “ Bk pϕk updet gqhq ´ Bk phpdet gqϕk qu
The first term integrates to zero by Stokes’s theorem (see earlier notes for more justification on
this), and so wedging by dz 1 ^ ¨ ¨ ¨ ^ dz̄ n and integrating gives
ωn
ż
pϕ, B̄uq “ ´ Bk phpdet gqϕk qpdet gq´1 h´1 hu “ pB̄ : ϕ, uq
X n!
where
B̄ : ϕ “ ´h´1 pdet gq´1 Bk phpdet gqg kī ϕī q. (2.21)
On a Kähler manifold, this is in fact
B̄ : ϕ “ ´g kī Bk ϕī ´ h´1 Bk hg kī ϕī ´ pdet gq´1 Bk ppdet gqg kī qϕī
by the Kähler condition Bk gab̄ “ Ba gkb̄ . This proves (2.22). With this formula for the adjoint, we
now compute the Laplacian.
Lemma 2.14. Let pL, hq Ñ pX, ωq be a holomorphic line bundle with metric over a compact Kähler
manifold. For any ϕ P Λ0,1 pLq, we have
36
since ∇k̄ g āb “ 0. Next,
This is because Γij k “ Γji k for a Kähler manifold since Γij k “ Bi gj p̄ g p̄k . Therefore
In Kähler geometry, one can see directly from Rj i mk̄ “ ´Bk̄ pBm gj p̄ g p̄i q the symmetry
Ra b j k̄ “ Rj b ak̄ .
We now let ϕ P Λ0,1 pLbKX q. We can apply the previous formula with ϕ P Λ0,1 pL̃q, and L̃ “ LbKX
equipped with the product metric h̃ “ h b pdet gq´1 . Using the formula Rik̄ “ ´Bi Bk̄ log det g and
F̃j k̄ “ ´Bj Bk̄ log h, we see that the curvature is
F̃j k̄ “ Fj k̄ ´ Rj k̄
which is
0 ě p∇ϕ, ∇ϕqL2 ` εpϕ, ϕqL2 .
It follows that ϕ “ 0. This proves that ker ∆ “ t0u. Therefore dim H 1 pX, L b KX q “ 0.
Here we used integration by parts on a Kähler manifold, which follows from the divergence theo-
rem ż
∇a V a ω n “ 0, V P ΓpT 1,0 Xq. (2.25)
X
37
In the above computation, this was used with V a “ ∇a ϕk̄ ϕ̄k̄ h P ΓpT 1,0 Xq and
ωn n
ωn
ż ż ż
a k̄ a k̄ ω
0“ ∇a p∇ ϕk̄ ϕ̄ hq “ ∇a ∇ ϕk̄ ϕ̄ h ` ∇a ϕk̄ ∇a pϕ̄k̄ hq
X n! X n! X n!
and we also then used metric compatibility ∇a h “ 0. The divergence theorem (2.25) comes
from
ωn
ż
dιV “ 0,
X n!
and the integrand is
ωn
Bi pV i det gqpidz 1 dz̄ 1 q . . . pidz n dz̄ n q “ pdet gq´1 Bi pV i det gq .
n!
We compute
Therefore ż
pBi V i ` Γib b V i q ω n “ 0.
X
On the other hand
∇i V i “ Bi V i ` Γib i V b .
Since Γib “ Γbi , these are equal.
This vanishing theorem can be generalized, and more generally there holds (see e.g. [20]
Theorem 2.15. Let L Ñ pX, ωq be a positive holomorphic line bundle over a compact Kähler
manifold. Then
H q pX, Ωp b Lq “ 0
for all integers p, q with p ` q ą n.
38
ř
so that pBxi Bxi ` Byi Byi qRe f “ 0. By the maximum principle for harmonic functions on B1 p0q Ď
R2n , since pRe f qp0q “ M then Re f ” M in all of B1 p0q. Hence S is open, and S “ X. Since
Re f “ M and |f | ď M , then Im f “ 0 and f is a constant.
We will define sheaves of OpU q modules (there is also a notion of sheaves of groups, vector spaces,
etc).
Definition 2.16. A presheaf (of OX modules) F on a complex manifold X is defined by the
following information. For any non-empty open set U Ď X, we associate a nonempty OpU q-module
FpU q, and a collection of restriction maps ρU,V : FpV q Ñ FpU q defined when U Ď V satisfying
for U Ď V Ď W . The set FpU q is called the set of sections of F over U . We also use the notation
s|U “ ρU V psq for s P FpV q.
Definition
Ť 2.17. A sheaf F on X is a presheaf satisfying the following glueing property. Suppose
Ω “ Uµ are open sets in X. If sµ P FpUµ q are such that
then there exists s P FpΩq such that ρUµ ,Ω psq “ sµ . Also, if s, t P FpΩq and ρUµ ,Ω psq “ ρUµ ,Ω ptq
for all µ, then s “ t.
In other words, local sections of a sheaf can be uniquely glued together.
Example 2.18. We write OX for the sheaf of holomorphic functions on a complex manifold X.
Example 2.19. Let E Ñ M be a holomorphic bundle. We write E for the sheaf of holomorphic
sections: EpU q are holomorphic sections over U .
Example 2.20. Sheaf I0 described by holomorphic functions in C2 vanishing at the origin. If U
does not contain the origin, then this is generated by 1 so I0 pU q – OpU q. In a neighbourhood V
of the origin, this is generated by x and y: any local holomorphic function f with f p0q “ 0 can be
written as f pxq “ gpx, yqx ` hpx, yqy. Thus I0 pV q is a module of rank 2. Thus the rank jumps up
to 2. Also, at the origin, the module is not free. For example, we have the relation ´y ¨ x ` x ¨ y “ 0.
In this sense, sheaves are sometimes viewed as a generalization of vector bundles where the rank
may jump.
Example 2.21. Another example to note are the constant sheaves Z, R, C. These are sheaves of
groups, meaning that FpU q attaches a group for every open set U (rather than a module). So for
example, ZpU q are locally constant Z-valued functions on U .
Ů Stalk of a sheaf. Let x P X. The stalk Fx is the set of equivalence classes in the
Definition 2.22.
disjoint union xPU FpU q with s1 P FpU1 q and s2 P F pU2 q satisfying s1 „ s2 if s1 |V “ s2 |V for
some V Ă U1 X U2 . The stalk Fx is an Ox,X -module.
A map of sheaves ϕ : F Ñ E is a collection of homomorphisms ϕU : FpU q Ñ EpU q such that ϕU ,
ϕV commute with the restriction maps. We say
0ÑE ÑF ÑGÑ0
39
is an exact sequence of sheaves if, denoting the arrows by fi , we have that all fi are maps of sheaves
with fi`1 ˝ fi “ 0 and the associated complex of stalks
0 Ñ Ex Ñ Fx Ñ Gx Ñ 0
is exact for all x P X. Recall that exact means that the kernel of one arrow is the image of the
previous arrow.
We will use the following two results from Cech cohomology [13]. Rather than define the Cech
cohomology groups Ȟ q pX, Eq, we will just directly use the following two facts:
‚ Given an exact sequence of sheaves 0 Ñ E Ñ F Ñ G Ñ 0, there exists a long exact sequence in
cohomology
¨ ¨ ¨ Ñ Ȟ p pX, Eq Ñ Ȟ p pX, Fq Ñ Ȟ p pX, Gq Ñ Ȟ p`1 pX, Eq Ñ ¨ ¨ ¨ .
‚ Dolbeault theorem:
kerpB̄ : Λp,q pEq Ñ Λp,q`1 pEqq
Ȟ q pX, Ωp b Eq – .
im pB̄ : Λp,q´1 pEq Ñ Λp,q pEqq
H q pX, Cq Ñ H q pY, Cq
is an isomorphism for q ď n ´ 2.
Example 2.24. Let Y “ tP “ 0u Ă Pn be a smooth complex manifold cut out by a homogeneous
polynomial P of degree m ě 1. We showed earlier that OpY q “ Opmq, which is positive. Therefore
H k pY, Cq is isomorphic to H k pPn q for k ď n ´ 2.
0 Ñ Op´Y q Ñ OX Ñ OY Ñ 0.
Here Op´Y q is the dual bundle of OpY q. This means that sections s P ΓpOp´Y qq transform as
sU “ tU V sV with tU V “ fV {fU , where Y has defining function fU “ 0 over an open set U . The
transformation relation shows that the combination sU fU is a well-defined function on the manifold.
To explain the exact sequence, over an open set U where Y has defining function fU “ 0, a local
holomorphic section sU P Op´Y qU gets sent to sU fU P OU which is a holomorphic function on U
vanishing along Y .
40
Tensoring with bΩpX implies the exact sequence
H q pX, ΩpX p´Y qq Ñ H q pX, ΩpX q Ñ H q pY, ΩpX |Y q Ñ H q`1 pX, ΩpX p´Y qq
We note that
by Serre duality and the vanishing theorem (Theorem 2.15) when pn ´ qq ` pn ´ pq ą n. Also
considering this with q replaced by q ` 1, we see that when p ` q ă n ´ 1 we have
Here y P Y and coordinates are chosen over an open set U Ă X such that U X tz n “ 0u “ U X Y ,
and
and we may multiply the generators by local holomorphic functions on Y . The sequence (2.30)
implies
0 Ñ Op´Y q|Y b Ωp´1
Y Ñ ΩpX |Y Ñ ΩpY Ñ 0,
as sheaves over Y . This is because
pN |Y q˚ “ Op´Y q
which can be seen as follows: if there are two sets of coordinates z, z̃ where both z n “ 0 and z̃ n “ 0
locally cut out Y , then z̃ n pzq “ z n f pzq, where f pzq is the transition function for OpY q. Note that
f pzq is non-vanishing; this is because Y “ tz̃ n pzq “ 0u smooth means that Bzn z̃ n pyq ‰ 0. Next,
n
dz̃ n “ Bz̃ i n n
Bz i dz implies dz̃ |Y “ f pzqdz |Y . This is the transformation law for the local frame dz ,
n
˚
so components of pN |Y q transform by the inverse 1{f which is why the dual Op´Y q appears.
Thus
H q pY, Ωp´1 q p q p
Y p´Y qq Ñ H pY, ΩX |Y q Ñ H pY, ΩY q Ñ H
q`1
pY, Ωp´1
Y p´Y qq.
41
Here we used that OpY q|Y Ñ Y is positive if OpY q Ñ X is positive. The dimension of Y is n ´ 1,
so the vanishing theorem applies if
pn ´ q ´ 1q ` pn ´ p ` 1q ą pn ´ 1q
so H q pY, ΩpX |Y q Ñ H q pY, ΩpY q is an isomorphism, which combined with (2.29) gives that
is an isomorphism.
42
3 Deformations of Complex Manifolds
3.1 Families of complex manifolds
From a complex manifold given by
X “ tP “ 0u Ă Pn ,
we can consider a 1-parameter family Xt by inserting a parameter t P C in front of one of the
polynomial coefficients. A famous example [3] is
"ÿ 4
ź *
Xt “ Zk5 ´ t Zk “ 0 Ă P4 .
k“0 k“0
X “ tpp, tq P Xt ˆ Cu.
We can also consider families where t P Cr , and we use the notation ∆ Ă Cr for a ball of radius 1.
The formal definition of a family of complex manifolds is:
Definition 3.1. Let X0 be a compact complex manifold. A family of deformations of X0 over ∆ Ă
Cr is given by π : X Ñ ∆ where X is a complex manifold, π is a holomorphic map, π ´1 p0q “ X0 ,
and the Jacobian of π has maximal rank.
πpz 1 , . . . , z n , tq “ t.
Change of coordinates on an overlap pz, tq, pz̃, t̃q are of the form
z̃ k “ f k pz 1 , . . . , z n , tq, t̃ “ t. (3.1)
We will now show that in a family of complex manifolds, all underlying smooth manifolds are
diffeomorphic.
43
Lemma 3.2. Let π : X Ñ ∆ be a family of compact complex manifolds.
Let t1 P ∆. From any path γ : r0, 1s Ñ ∆ with γp0q “ 0 and γp1q “ t1 , we can construct a
1-parameter family of diffeomorphisms Θs : X0 Ñ Xγpsq such that Θ0 “ id.
In particular Xt1 is diffeomorphic to X0 for all t1 P ∆.
Proof. Let γpsq be a path on ∆ from 0 to t0 . Extend the vector field γ9 arbitrarily to all of ∆. Let
tbi u be real coordinates on ∆, and write
B
γpbq
9 “ γ9 i pbq .
Bbi
We will lift up the vector field on the base γ9 P ΓpT B, Bq using a partition of unity. Cover X with
finitely many coordinates charts
ř pz 1 , . . . , z n , b1 , . . . , bn q as described earlier where πpz, bq “ b. Let
ρα be a partition of unity ( ρα “ 1, supppρα q Ă Uα ) subordinate to this cover. Define on X the
vector field
ÿ „
i B
V “ ρα γ9 .
α
BbiUα
It may be clearer to write V using fixed coordinates on say U0 Ă X , in which case
B ÿ Bz k B
V “ γ9 i i
` ρα γ9 i iU0 k .
BbU0 α
Bb Uα BzU0
44
9 Next, we solve the ODE system on X
So we get a lifted vector field V P ΓpT X q with π˚ V “ γ.
d
Θε “ V, Θ0 “ id,
dε
which is well-known to produce a 1-parameter family of diffeomorphisms Θε : X Ñ X together with
d
inverses Θ´ε : X Ñ X satisfying dε Θ´ε “ ´V and Θε ˝ Θ´ε “ id.
The last thing to check is that this construction produces diffeomorphisms from fiber to fiber
Θε : X0 Ñ Xγpεq . Fix x P X0 and consider the function f pεq “ π ˝ Θε pxq ´ γpεq. Then
f p0q “ 0, f 1 pεq “ π˚ V ´ γ9 “ 0.
Looking relative to the moving family of diffeomorphisms, we may regard Xt as the fixed differen-
tiable manifold X0 , and let the complex structure tensor Jt vary in t. More precisely: let X Ñ ∆
be a family of complex manifolds, so that pXt , Jˇt q is a complex manifold for each parameter t P ∆.
Take a path γ : p´ε, εq Ñ ∆ with γp0q “ 0, γp0q
9 ‰ 0, and so from Lemma 3.2 we obtain a family
of diffeomorphisms Θt : X0 Ñ Xγptq . Then we may consider
45
Let X Ñ ∆ be a family of complex manifolds with central fiber pX, Jq, together with a path
γ : p´ε, εq Ñ ∆ with γp0q “ 0 and γp0q
9 ‰ 0. From this data, we would like to produce an
element
rηs P H 1 pX, T 1,0 Xq.
Given our discussion so far, from this information we can create a path of complex structures
Jt P ΓpEndT Xq on the fixed differentiable manifold X with J0 “ J satisfying the constraints
Jt2 “ ´id and N pJt q “ 0. We will now show:
9
‚ Differentiation η “ Jp0q produces an element η P Λ0,1 pT 1,0 Xq which satisfies B̄η “ 0 and hence
defines a Dolbeault cohomology class rηs P H 1 pX, T 1,0 Xq.
‚ Different choices of diffeomorphisms Θt : X Ñ Xγptq in Lemma 3.2 produce the same class
rηs P H 1 pX, T 1,0 Xq.
d
We start by differentiating Jt2 “ ´id. Take dt |t“0 and obtain
η α β̄ :“ J9α β̄ .
d
Next, we differentiate N pJt q “ 0. Taking dt |t“0 of (3.2) gives
which becomes „
9 α 9 α
0 “ 2 ´ iBβ̄ J γ̄ ` iBγ̄ J β̄ .
46
From the complex manifold pXt , Jˇt q, we produce a family of complex structures on the fixed X0 as
before by Jt “ pΘt q´1 ˇ ˜ ´1 ˇ
˚ Jγptq pΘt q˚ and Jt “ pΨt q˚ Jγptq pΨt q˚ . These are related by
d
with ft “ Ψ´1
t ˝ Θt . Then ft be a 1-parameter family of diffeomorphisms with dt |t“0 ft “ V (for
some vector field V ) and f0 “ id.
We now compute dt d
|t“0 of J˜t . Let y α “ ftα pxq be a change of coordinates by the diffeomorphism.
The formulas for the pushfoward f˚ : Tp M Ñ Tf ppq M are
By a Bxi
pf˚ V qa pf ppqq “ ppqV i ppq, pf˚´1 V qa ppq “ pf ppqqV a pf ppqq,
Bxi By a
and so acting Jt on V P Tf ppq M gives
By a k Bx
i
rJ˜t a b V b spf ppqq “ ppqJt ppq i pf ppqqV a pf ppqq
Bxk By a
The components of Jt are then
By a Bxi
J˜ta b pyt pxqq “ tk pxqJt pxqk i b pyt pxqq.
Bx Byt
Differentiating this in time at t “ 0 along a path with y0 pxq “ x and yp0q 9 “ V gives
ˇ
d ˇ Bxi
J9̃a b ` Bk J a b V k “ Bk V a J k b ` J9a b ` J a i ˇˇ pyt pxqq
dt t“0 By b
In complex coordinates, the second term on the left is zero since the components J a b are constant.
For the last term, we differentiate in time the chain rule identity
Bxi By k B
k
pypxqq j
pxq “ j xi “ δ i j
By Bx Bx
to obtain ˇ
d ˇˇ Bxi
pyt pxqq “ ´Bj V i .
dt ˇt“0 By j
Therefore
J9̃a b “ J9a b ` Bk V a J k b ´ J a i Bb V i .
We showed earlier that considering J 2 “ ´id implies that the non-zero contributions are η α β̄ “
9 α β̄ and its conjugate. So we let a “ α, b “ β̄ and obtain
Jp0q
η̃ α β̄ “ η α β̄ ´ 2iBβ̄ V α .
It follows that
η̃ “ η ´ 2iB̄V 1,0
and so rηs “ rη̃s P H 1 pT 1,0 Xq.
47
Remark 3.4. Let pX, Jq be a complex manifold. This calculation shows that deformations
J˜t “ pft q˚ Jpft q´1
˚ created by 1-parameter families of diffeomorphisms ft produce the zero class
rηs “ 0 P H 1 pX, T Xq (since J9 “ 0). Deformations of complex structure coming from families of
diffeomorphisms are not counted by rηs P H 1 pX, T Xq.
A central question in deformation theory is the inverse problem: given η P H 1 pT 1,0 Xq, does it
9
come from a family pXptq, Jptqq of complex manifolds with Jp0q “ η? This statement is not true
in general, but it is true for Kähler Calabi-Yau manifolds (this is the Bogomolov-Tian-Todorov
theorem). For more references on this topic, see for example [20, 16, 14].
Returning to complex geometry, let pX, Jt q be a family of complex structures on a compact manifold
X with J0 “ J. A differential form α P Λk pXq has different decompositions into pp, qq types for
each parameter t. Write Λ1,0 0 for p1, 0q forms with respect to the initial structure J, and Λt1,0
for p1, 0q forms with respect to Jt . We can decompose α P Λ1 pXq into pp, qq components via the
formula
1 1
α “ pα ´ iJt αq ` pα ` iJt αq
2 2
:“ pαq1,0
t ` pαqt
0,1
(3.4)
Here we define JαpXq “ αpJXq. With this notation, we have that α P Λ1,0
0 if and only if Jα “ `iα.
The map ϕt : Λ1,0
0 Ñ Λ 1,0
t given by
α ÞÑ pαq1,0
t
is an isomorphism for small t, since ϕ0 “ id and ϕt varies continuously as seen from the explicit
expression above. Similarly, we obtain isomorphisms
ϕt : Λp,q p,q
0 Ñ Λt
To view all operators on the same space, instead of ∆B̄ we will use
Lt : Λp,q p,q
0 Ñ Λ0 , Lt “ ϕ´1
t ˝ ∆B̄t ˝ ϕt ,
48
and show
dim ker Lt ď dim ker L0 , |t| ă ε.
To prove this, we will need some PDE estimates.
Let EŤÑ X be a vector bundle over a compact manifold. Cover X by finitely many trivializations
N
X “ i“1 Ui where Bi Ă Ui Ă Cn are balls of radius 1 still covering X. Let ψ P ΓpX, Eq and let ψUi
denote the vector valued function of the components of ψ in the trivialization Ui . Let 0 ă α ă 1.
Define
}ψ}C k,α :“ sup }ψUi }C k,α pBi q
i
where }f }C k pBq “ sup|I|“k supB |DI f |. We write ψ P C k,α pX, Eq if ψ is a k-times differentiable
section of E with finite } ¨ }C k,α norm. The main features of C k,α spaces for our purposes are:
‚ C k,α pX, Eq is a Banach space.
‚ Compactness: suppose tψn u P C k,α pX, Eq is a sequence of sections such that
}ψn }C k,α ď C
for uniform constant C ą 0. Let 0 ă α1 ă α. Then there exists a limiting section ψ8 P C k,α and a
1
subsequence tψnk u such that ψnk Ñ ψ8 in C k,α .
‚ The Schauder estimates. (Theorem 3.5 below)
Let us prove that C 0,α pX, Eq :“ C α is a Banach space and its compactness property. It is routine
to check that } ¨ }C α is a norm. To show completeness, we must show that if tψn u is a Cauchy
sequence, then ψn converges to ψ8 P C α . By Arzela-Ascoli applied on each coordinate ball Bi , a
subsequence ψnk converges in C 0 to a continuous limit ψ8 , and since tψn u is Cauchy then the full
sequence converges
ψn Ñ ψ8 in C 0 .
The limit section ψ8 is in C α , since for x, y P Bi with |x ´ y| “ δ, then
|ψ8 pxq ´ ψ8 pyq| |ψ8 pxq ´ ψk pxq| |ψk pxq ´ ψk pyq| |ψk pyq ´ ψ8 pyq|
ď ` `
|x ´ y|α |x ´ y|α |x ´ y|α |x ´ y|α
ď C (3.5)
for k large enough such that }ψ8 ´ ψk }C 0 ă δ. To show the sequence converges in C α , let ε ą 0.
There exists N such that if k, ` ě N then for all x ‰ y there holds
|rψk ´ ψ` spxq ´ rψk ´ ψ` spyq|
ď ε.
|x ´ y|α
Fix x, y and send ` Ñ 8 to conclude
}ψk ´ ψ8 }C α ă ε.
49
We now show compactness. Suppose tψn u P C α is a sequence with }ψn }C α ď C. By the Arzela-
Ascoli theorem applied to coordinate balls Bi , we obtain a subsequence ψnk converging in C 0 to a
continuous limit ψ8 . The limiting section is Hölder continuous ψ8 P C α by estimate (3.5). To see
convergence in 0 ă α1 ă α, we let vk “ ψnk ´ ψ8 and write
ˆ ˙α1 {α
|vk pxq ´ vk pyq| |vk pxq ´ vk pyq| 1
“ |vk pxq ´ vk pyq|pα{α q´1 .
|x ´ y|α1 |x ´ y|α
Here C only depends on the constant of ellipticity and the C α norms of the coefficients of L.
For a reference, see [19], remark after Theorem 4.3 in the appendix.
We can upgrade this estimate for sections s P pker LqK .
Theorem 3.6. (Schauder estimates II) Let E, F Ñ X be complex vector bundles over a compact
complex manifold. Let L : ΓpX, Eq Ñ ΓpX, F q be an elliptic operator of order k. Let H be a metric
on E and ω a hermitian metric on X. There exists C ą 0 such that for all s P C k,α pX, Eq with
s P pker LqK
then
}s}C k,α pXq ď C}Ls}C 0,α .
Here s P pker LqK means: ps, ϕqL2 “ş 0 for all ϕ P ΓpX, Eq with Lϕ “ 0, and the L2 inner product
on ΓpX, Eq is as before: ps, ϕqL2 “ X xs, ϕyH ω n .
Proof. Suppose this estimate is false. Then there exists a sequence tsi u P pker LqK X C k,α such that
u8 “ 0.
50
We want to obtain a contradiction, but we cannot take a limit of }un }C k,α “ 1 since α ą α1 .
However, by the usual Schauder estimates, for n " 1 large enough we have
1
1 “ }un }C k,α ď Cp}un }L8 ` }Lun }C 0,α q ď C}un }L8 `
2
and so }un }L8 ě 1{2C for all n large. Taking a limit gives }u8 }L8 ą 0 which is a contradiction.
It follows that
pker L0 qK X pker Lt q “ t0u.
From
pker Lt q Ă ΓpX, Eq “ pker L0 q ‘ pker L0 qK ,
we have ker Lt Ď ker L0 .
p,q p,q
To conclude semi-continuity of Hodge numbers, we let Lt “ ϕ´1 t ˝ ∆B̄ ˝ ϕt : Λ0 Ñ Λ0 as before.
Then
hp,q pXt q ď hp,q pX0 q, |t| ă ε
and we see that t ÞÑ dim H q pXt , Ωp q is upper semi-continuous in a family of complex mani-
folds.
51
3.3 Stability of Kähler metrics
Kodaira-Spencer’s stability theorem states that if X0 is a compact Kähler manifold, then any
deformation Xt admits a Kähler metric for small enough t [21].
After this section, we will be restricting our attention from general complex manifolds to simply
connected Calabi-Yau threefolds. We will in this section give the proof of Kodaira-Spencer’s stability
theorem assuming that X0 is a simply connected Calabi-Yau threefold. Explicitly, the only extra
hypothesis that we will use is H 0,2 pX0 q “ 0, so the proof here applies to any complex manifold
X0 satisfying this property. Our proof under this hypothesis is simpler compared compared to
Kodaira-Spencer [21], and we will follow the argument and exposition in Fu-Li-Yau [11]. The setup
in Fu-Li-Yau [11] is different as it concerns balanced metrics rather than Kähler metrics and the
central fiber X0 has nodal singularities, but the outline of the argument is readily adapted.
Let ω0 be the Kähler metric on X0 , and Θt be a family of diffeomorphisms Θt : X0 Ñ Xt .
Then
Θ˚t ω0 “ pΘ˚t ω0 q2,0 ` pΘ˚t ω0 q1,1 ` pΘ˚t ω0 q0,2
is a closed 2-form on Xt . To obtain a Kähler metric on Xt , it must be of type p1, 1q, so we let
χt “ pΘ˚t ω0 q1,1 .
0 “ dpΘ˚t ω0 q
„ „ „ „
“ BpΘ˚t ω0 q2,0 ` B̄pΘ˚t ω0 q2,0 ` BpΘ˚t ω0 q1,1 ` B̄pΘ˚t ω0 q1,1 ` BpΘ˚t ω0 q0,2 ` B̄pΘ˚t ω0 q0,2
Taking the conjugate, we conclude dωγ “ 0. To show that ωγ is a Kähler metric, we will show its
positivity ωγ ą 0. For this, we will show that the correction γ is small.
Roughly speaking, the strategy is to correct ω̃t “ χt ` Bα ` B̄ ᾱ so that α solves B̄Bα “ ´B̄ ω̃t and
so B̄ ω̃t “ 0. If B̄B were an elliptic operator, one could try to use elliptic PDE theory to estimate
}α} ď C}B B̄α} “ C}B̄χt } and show smallness of α since B̄χt Ñ 0 as t Ñ 0. To make this strategy
work, we do not use α but rather γ with α “ B̄ : B : γ. Furthermore (3.9) is not quite an elliptic PDE
for γ, but this equation can be modified to make this strategy work.
52
To make (3.9) elliptic, we consider the Kodaira-Spencer operator
Et “ B B̄ B̄ : B : ` B̄ : B : B B̄ ` B̄ : BB : B̄ ` B : B̄ B̄ : B ` B̄ : B̄ ` B : B
and solve
Et pγt q “ B̄χt .
In the definition of Et , the B, B̄ are with respect to the complex structure at t, and the adjoints
B : , B̄ : are with respect to the non-Kähler metrics χt . The operator Et is a 4th order elliptic operator
as proved by Kodaira-Spencer. The first term in Et matches with (3.9), so we will need to look
for solutions in a space where all the other terms vanish; we see that if we can find a solution to
Epγq “ B̄ω with dγ “ 0, we will solve (3.9).
By the Fredholm alternative, we can find a solution γt P pker Et qK if B̄χt K ker Et . Note that
considering pEt ϕ, ϕq and integrating by parts shows that
ker Et “ tϕ : dϕ “ 0, B̄ : B : ϕ “ 0u.
Here we use the assumption of vanishing cohomology H 0,2 pX0 q “ 0. By the semi-continuity theo-
rem,
h0,2 pXt q ď h0,2 pX0 q “ 0
and by (3.7), we have rpΘ˚t ω0 q0,2 s P H 0,2 pXt q “ 0 and so we can write pΘ˚t ω0 q0,2 “ B̄ν for ν P Λ0,1 .
By (3.7), we have B̄χt “ ´B B̄ν.
It follows from (3.10) that for β P ker Et then
xB̄χt , βy “ xµt , B̄ : B : βy “ 0.
Therefore, we can find a solution to Et pγt q “ B̄χt “ iB B̄µt . Next, we prove that such a solution is
closed.
Lemma 3.8. If Epγq “ iB B̄µ, then dγ “ 0.
Proof. We compute
0 “ Et pγq ´ iB B̄µt
„ „ „
: : : : : : :
“ B B̄ B̄ B γt ´ iµt ` B B̄ B̄ Bγt ` Bγt ` B̄ B B B̄γt ` BB B̄γt ` B̄γt
:“ B B̄α1 ` B : α2 ` B̄ : α3 . (3.11)
Setting σ “ 0, we obtain
„ „
: : : : :
0 “ B B̄ B̄ Bγt ` Bγt ` B̄ B B B̄γt ` BB B̄γt ` B̄γt .
53
and taking an inner product with γt implies
In summary, by the Fredholm alternative we have γt P pker Et qK with Et γt “ B̄χt and dγt “ 0.
Therefore (3.9) holds and the corrected ωγ (3.8) satisfies dωγ “ 0. To show ωγ is a Kähler metric,
we prove the estimate
}γt }C 3 pX,g0 q ď C|t|. (3.12)
Then since χt ą p1{2qω0 for small t, we see that ωγ ą 0 for small t.
Lemma 3.9. There exists C ą 1 independent of t such that for all t small, we can estimate
Proof. Suppose by contradiction that the estimate fails, and there is a sequence ti Ñ 0 such that
The Eti Ñ E0 smoothly, and after relabeling a subsequence we have that γti Ñ γ0 in C 4,α{2 with
E0 γ0 “ 0 and dγ0 “ 0. This limit is non-trivial due to the Schauder estimates
Bγ0 “ 0, B : γ0 “ 0.
dγ0 “ 0, d: γ0 “ 0.
We will now use γt P pker Et qK , dγt “ 0, to show that γ0 “ 0, which is a contradiction. Since
Λ1,2 “ Im Et ‘ ker Et , we can write γt “ Et pβt q, and so
γt “ B B̄β1 ` B̄ : β2 ` B : β3 .
54
Since dγt “ 0, we have
0 “ pγt , B̄ : β2 ` B : β3 q “ }B̄ : β2 ` B : β3 }2 .
So γt “ dαt . We note
pγt , γ0 qL2 pX0 q “ pαt , d:0 γ0 qL2 pX0 q “ 0.
Let t Ñ 0, we conclude
pγ0 , γ0 qL2 pX0 q “ 0
which is a contradiction.
Proof. By the Kodaira-Spencer theorem, Xt is a Kähler manifold for all t small enough. By the
semi-continuity theorem, hp,q pXt q ď hp,q pX0 q for small t, so we suppose by contradiction that there
exists t and p, q such that hp,q pXt q ă hp,q pX0 q. Let p ` q “ k, and
ÿ ÿ ÿ
hi,j pXt q ă hi,j pX0 q “ bk “ hi,j pXt q
i`j“k i`j“k i`j“k
55
4 Calabi-Yau Threefolds
4.1 Parameters of threefolds
There are various inequivalent definitions of Calabi-Yau manifolds used in the literature. The
Wikipedia page for Calabi-Yau manifolds gives some of the commonly used definitions. In these
notes, we will use the following definition:
Definition 4.1. Our definition of a Kähler Calabi-Yau manifold is a simply-connected compact
complex manifold of dimension n admitting a Kähler metric ω and a nowhere vanishing holomorphic
pn, 0q form Ω.
The section Ω is a nowhere vanishing holomorphic section of the canonical bundle KX “ pdet T 1,0 Xq˚ ,
and so KX “ OX is the trivial holomorphic bundle.
Let pY, Ω, ω̂q be a Kähler Calabi-Yau threefold. Since we assume that Y is simply connected, we
have b1 pXq “ 0 and so by the Hodge decomposition (2.11) then
h1,0 “ h0,1 “ 0.
h1,1 , h2,1 .
By the Hodge decomposition (2.11), we have b2 “ h1,1 and b3 “ 2 ` 2h2,1 . The Euler characteristic
is defined by
ÿ6
χpY q “ p´1qi bi ,
i“1
56
and a given Kähler class rωs is parametrized by functions.
‚ h2,1 encodes the infinitesimal complex structure deformations of X. We discussed earlier how
a 1-parameter family of complex structures pX, Jptqq produces an element rηs P H 1 pX, T 1,0 Xq by
9
η “ Jp0q. Note that by Serre duality
ω “ igj k̄ dz j ^ dz̄ k , Ω “ f dz 1 ^ . . . dz n
where f pzq is a local nowhere vanishing holomorphic function. The norm of Ω induced on pdet T 1,0 Xq˚
is
f pzqf pzq
|Ω|2ω “ .
det gj k̄
The Chern-Ricci curvature Rj k̄ “ ´Bk̄ Bj log det g can also be written
Indeed,
Bk̄ Bj log det |Ω|2ω “ Bk̄ Bj log |f |2 ´ Bk̄ Bj log det g,
and Bk̄ Bj log |f |2 “ 0 for any such function f . This is because
f Bk̄ f¯ Bk f
Bj Bk̄ log |f |2 “ Bj ¯ “ Bj̄ “ 0.
ff f
The maximum principle on compact manifolds implies that |Ω|ω is constant. Here is a quick proof
if ω is Kähler. (Note: this PDE result holds without the Kähler assumption. The general proof
uses the Hopf maximum principle instead of integration by parts. The general statement is: if
aij Bi Bj f ` bi Bi f “ 0 on a compact manifold with aij positive-definite, then f is constant.)
57
First, note the identities for a real function u : X Ñ R:
1 j k̄
iB B̄u ^ ω n´1 “ pg uj k̄ qω n , uj k̄ “ Bj Bk̄ u,
n
and
1
iBu ^ B̄u ^ ω n´1 “
|Bu|2g ω n .
n
These can be checked at a point p P X with ω|p “ k idz k ^ dz̄ k . By Stokes’s theorem and
ř
dω “ 0, ż ż
plog |Ω|2ω qidB̄ log |Ω|2ω ^ ω n´1 “ ´ dplog |Ω|2ω q ^ iB̄ log |Ω|2ω ^ ω n´1 .
X X
which implies ż ż
0“ plog |Ω|2ω qpg j k̄ Bj Bk̄ log |Ω|2ω q ω n “´ |B log |Ω|2ω |2g ω n
X X
and so |B log |Ω|2ω |2g “ 0 and log |Ω|2ω is a constant.
In summary, on a compact complex manifold with trivial canonical bundle, then Rj k̄ “ 0 is equiv-
alent to |Ω|2ω “ const. To find such metrics, we fix ω an arbitrary hermitian metric, and look for
solutions to this equation of the form
g̃j k̄ “ gj k̄ ` Bj Bk̄ ϕ ą 0
or in differential form notation
ω̃ “ ω ` iB B̄ϕ ą 0.
The equation |Ω|2ω̃ pxq b
“ e , b P R, can be written
|Ω|2ω̃ log |Ω|2ω
„
b det g log |Ω|2ω
e “ e “ e ,
|Ω|2ω det g̃
which leads to the complex Monge-Ampère equation
„
detpgj k̄ ` ϕj k̄ q 2
“ elog |Ω|ω ´b .
det gj k̄
When dω “ 0, the constant b can be identified from the initial data pω, Ωq, since the equation can
also be written
pω ` iB B̄ϕqn “ e´b |Ω|2ω ω n
and integration of both sides and Stokes’s theorem gives
ż ż
n
ω “ e´b |Ω|2ω ω n .
X X
Indeed:
ż ż n
ÿ ż
pω ` iB B̄ϕqn “ ωn ` ck ω n´k ^ piB B̄ϕqk
X X k“1 X
ż ÿn ż
“ ωn ` ck dpω n´k ^ piB̄ϕq ^ piB B̄ϕqk´1 q
X k“1 X
ż
“ ωn . (4.1)
X
58
Another way to write the Kähler-Ricci flat metric equation is
2
pω ` iB B̄ϕqn “ e´b in Ω ^ Ω̄,
RicpωCY q “ 0, dωCY “ 0.
Furthermore, each Kähler class rωs P H 1,1 pX, Rq contains a unique Calabi-Yau representative ωCY P
rωs by solving the Monge-Ampère equation with ansatz ω ` iB B̄ϕ.
Proof. The existence of the family of Kähler metrics ωptq “ ωt with ωt Ñ ω is Kodaira-Spencer’s
stability theorem [21] which we discussed earlier. We now describe how to construct Ωptq “ Ωt . The
3-form Ω is defined on pX, Jptqq, however it does not necessarily have type p3, 0q. So we take the
p3, 0q part pΩq3,0
t , and this is nowhere vanishing for small t by continuity of the complex structures
Jptq. We now need to correct pΩq3,0t to make it holomorphic. Write
pΩq3,0
t “ σt dzt1 ^ dzt2 ^ dzt3 .
59
Note α “ σt´1 B̄t σt is a well-defined p0, 1q-form on Xt . This is because of the transformation law
σ ÞÑ tU V σ with tU V holomorphic, which implies αk̄ ÞÑ αk̄ . We also note that B̄α “ 0, since
B̄α “ ´σ ´2 B̄σ ^ B̄σ ` σ ´1 B̄ 2 σ “ 0.
We claim that we can find a smooth function ut such that
ż
B̄t ut “ ´αt , u ωtn “ 0
Xt
Hence |Ωt ´ Ω|g ď C|t|. To prove (4.2), we will use the complex Laplacian
∆gt : C 8 pX, Rq Ñ C 8 pX, Rq,
given by
∆gt f “ pgt qj k̄ Bj Bk̄ f.
ş
We proved earlier that ker ∆gt “ R ¨ 1 are constants. Therefore since Xt ut ωtn “ 0, we have that
ut P kerp∆gt qK with respect to the gt inner product for all t. By elliptic estimates,
}ut }C 2,α pX,gq ď C}∆gt ut }C α pX,gq .
One should verify that the constant C is uniform in t. We omit the proof, but it follows an outline
similar to (3.13): the usual Schauder estimates }u}C 2,α ď Cp}u}L8 ` }∆u}C α q are uniform in t, and
to remove the extra }u}L8 we assume the estimate fails as t Ñ 0 and derive a contradiction.
Therefore since Bk̄ u “ σt´1 Bk̄ σt
60
9
Let pX, Jptq, ωptq, Ωptqq be a family of Kähler Calabi-Yau threefolds with Jp0q “ η. We use the
d
notation δ “ dt |t“0 . To end this section, we will compute the variation
χ :“ δΩ
and
´ iδΩᾱβγ “ ´δJ r ᾱ Ωrβγ ` δJ s j Ωᾱsγ ` δJ t k Ωᾱβt ` p´iqi2 δΩᾱβγ (4.7)
which implies „
i µ
χᾱβγ :“ δΩᾱβγ “ ´ δJ ᾱ Ωµβγ . (4.8)
2
We can also invert the formula for χ to get δJ from δΩ. At a point where gij “ δij , then
and „
i
Ω̄βγν ´ Ωµβγ δJ µ ᾱ “ ´i|Ω|2ω δJ ν ᾱ . (4.10)
2
Thus
i
δJ ν ᾱ “ Ω̄βγν δΩᾱβγ (4.11)
|Ω|2ω
or
i
η ν ᾱ “ Ω̄βµν χᾱβµ . (4.12)
|Ω|2ω
61
4.4 Quintic threefolds
4.4.1 Holomorphic volume form
Consider X Ă P4 given by tP “ 0u, where
Since
Df “ r5z14 , 5z24 , 5z34 , 5z44 s
we see that Df has maximal rank except at z “ 0, which is not included in the set tf “ 0u.
Therefore X is a smooth complex manifold of complex dimension three. This manifold admits a
holomorphic volume form Ω. This can be seen by the adjunction formula, but also by the explicit
expression
dz1 ^ dz2 ^ dz3
Ω|U0 XV4 “ (4.13)
Bf {Bz4
over the set U0 “ tx0 ‰ 0u intersected with V4 “ tBf {Bz4 ‰ 0u. We now verify that Ω extends
from U0 X V to a holomorphic volume form on all of X.
B3 f B4 g ` B4 f “ 0.
Hence
dz1 ^ dz2 ^ dz4
Ω|U0 XV3 “ ´
Bf {Bz3
This is holomorphic and nowhere vanishing on V3 , therefore Ω extends from V4 to V3 . Similar
arguments show that Ω defines a holomorphic volume form on all of U0 . Next, we need to extend
Ω beyond U0 Ă X.
62
since Bf {Bz4 “ 5z44 , and therefore
As before, Ω extends from U1 X Ṽ4 to all of U1 , and similarly Ω extends to a nowhere vanishing
holomorphic form on U2 , U3 , U4 .
Putting everything together, we see that the local expression (4.13) defines a holomorphic volume
form on all of X.
‚ h1,1 “ 1. This follows from the Lefschetz hyperplane theorem, which we recall states: let Y Ď X
be a complex hypersurface such that its associated line bundle OpY q Ñ X is positive. Then
The number of parameters are: number of ways of placing 5 objects (the powers) in 5 bins (the Zi ),
which is p5 ` 4q!{5!4! “ 126. But some of these 126 coefficients do not give genuine deformations
63
of complex structure. There are 52 degrees of freedom coming from matrices A P GLp5q which
produces a biholomorphism
A : tP pZq “ 0u Ď P4 Ñ tP pAZq “ 0u Ď P4 .
So we are left with 126 ´ 25 “ 101 parameters, which matches up with h1,2 “ 101. This is not a
proof because a priori there could be ways to deform the complex structure of a quintic which is
not by (4.14).
We note in passing that this family was used by [3] to construct one of the first examples of mirror
symmetry. We first notice that at t “ 1, this is no longer a smooth manifold as it contains singular
points. To find the singular points, we set all derivatives of Qt to zero.
ź
5Zk4 “ 5t Zi
i‰k
and so ˙4 ˙4
4
ˆź 4
ˆź
5
Zi “t Zi
i“0 i“0
If one of the Zi “ 0, then they all are, which is not a point in projective space. So we conclude
that singular points occur when t5 “ 1.
We let t “ 1 and investigate the singular quintic which we denote X. There are 125
ś singular points:
these occur when Z05 “ Z15 “ ¨ ¨ ¨ “ Z45 “ Zi . Dividing the singular points by Zi (in projective
ś
space), singular points are given by roots of unity Zk5 “ 1.
q “ rζ α0 , . . . , ζ α4 s, ζ 5 “ 1
ř
with αi P t0, 1, 2, 3, 4u. Since q P X, we must have i αi “ 0 mod 5. We can always represent q in
projective spaceřwith the first entry equal to 1, so that leaves 3 free parameters α1 , α2 , α3 with α4
determined by i αi “ 0, so there are p5qp5qp5q “ 125 singular points.
We now look locally near the point q “ r1, 1, 1, 1, 1s. In the coordinate chart pU0 , zq, q “ p1, 1, 1, 1q
and the equation of the singular quintic is
4
ÿ
f pzq “ 1 ` zi5 ´ 5z1 z2 z3 z4 “ 0.
i“1
The holomorphic function f satisfies f pqq “ 0 and Df pqq “ 0. Its holomorphic Hessian matrix
64
B2 f
Bz i Bz j at q is » fi
20 ´5 ´5 ´5
— ffi
— ffi
—´5 20 ´5 ´5ffi
— ffi
— ffi
—´5 ´5 20 ´5ffi
– fl
´5 ´5 ´5 20
This is a non-singular matrix. Thus we have a holomorphic function f pz1 , . . . , z4 q with f pqq “ 0,
df pqq “ 0, but D2 f pqq is non-degenerate. There is a holomorphic Morse lemma (e.g. Lemma
2.11 in the book [30]) which gives the existence of holomorphic coordinates w with qpwq “ 0 such
that
ÿ 4
f pwq “ wi2 .
i“1
We give the proof in complex dimension n “ 2 following (Lemma 42 p.242 in [8]). First, shift
coordinates so that q “ 0 and f pz1 , z2 q is a holomorphic function with f p0q “ 0, Df p0q “ 0, and
Bi Bj f p0q a non-degenerate symmetric matrix. For a symmetric complex matrix A, there is a unitary
matrix U such that U AU T “ D where D is diagonal (Autonne–Takagi factorization). Since A is
non-degenerate the diagonal elements are non-zero, so we may multiply on both sides by a diagonal
complex matrix to form SAS T “ I.
So we can write S T rD2 f p0qsS “ I for a complex matrix S, and let z k “ S i k y k . By the chain
rule
Bf Bf Bz p p B2 f
“ “ S k f p , “ S p k fpq S q j .
By k Bz p By k By j By k
Written in matrix notation, this is
Bf
Let f2 “ Bz 2
. Since B2 f2 p0q ‰ 0, by the holomorphic implicit function theorem there is a function
apz1 q with ap0q “ 0 such that f2 pz1 , apz1 qq “ 0. Differentiating this gives a1 p0q “ 0. Define new
coordinates by
z̃1 “ z1 , z̃2 “ z2 ´ apz1 q
and let f˜pz̃1 , z̃2 q “ f pz1 , z2 ´ apz1 qq. We will compute the power series of f˜. We start with 1st
derivatives:
B1 f˜ “ B1 f ´ a1 B2 f, B2 f˜ “ B2 f.
The key observation is that B2 f˜ ” 0 when z̃2 “ 0, and so
65
B1 B1 f˜ “ B1 B1 f ´ 2a1 B1 B2 f ` pa1 q2 B2 B2 f ´ a2 B2 f
B2 B2 f˜ “ B2 B2 f.
Evaluating all these at zero, and using that (4.15) implies B1k Bf f p0q “ 0, we get the expansion
ÿ ÿ
f “ z̃12 ` bk z̃1k ` z̃22 ` aij z̃1i z̃2j
kě3 i`jě3, jě2
ˆ ÿ ˙ ˆ ÿ ˙
“ z̃12 1` bk z̃1k ` z̃22 1` aij z̃1i z̃2j (4.16)
kě1 i`jě3, jě0
using that a holomorphic function ψ : C Ñ C with ψp0q ‰ 0 has a local square root defined in a
neighborhood of the origin. In these new coordinates then, f “ w12 ` w22 .
In summary, the singular quintic X has holomorphic coordinates around each singular point where
the singularity appears as
"ÿ 4 *
zi “ 0 Ď C4 .
2
i“1
Such singularities are called nodes, or nodal singularities, or ordinary double points (ODP).
We discussed earlier how for t ‰ 0, the space Xt no longer has singular points. In the chart pU0 , zq,
we see the zero set
ÿ4
ft pzq “ 1 ` zi5 ´ 5z1 z2 z3 z4 ´ tz1 z2 z3 z4 “ 0.
i“1
66
wi2 (here w “ 0 corresponds to the point p1, 1, 1, 1q) so that the zero set
ř
and write gpwq “ i
tft “ 0u becomes
4
ÿ
t wi2 “ tu Ă C4 . (4.18)
i“1
‚ The second way to desingularize X is by small resolution. We will discuss how this is done in full
detail in the next section, but it results in a map
σ:XÑX
where X is a complex manifold, σ ´1 ppq “ P1 for each singular point p, and σ is a biholomorphism
outside the singular set on X. In other words, each singular point of X is replaced by P1 .
The holomorphic volume form Ω on X defines a holomorphic volume form Ω̂ “ σ ˚ Ω defined on
XzE with E “ YP1 . By Hartog’s theorem, Ω̂ extends to all of X and so X has trivial canonical
bundle. We will show later that χpXq ´ 2N “ χpXt q where N is the number of nodes. Therefore
the two threefolds have different topologies.
Note Hartog’s theorem (e.g. p.46 in [7]) states: Let X be a complex manifold. Let S Ă X be a
closed complex submanifold of complex codimension ě 2: this means there are local coordinates
where U X S is given by z1 “ ¨ ¨ ¨ “ zp “ 0 for p ě 2. Then every holomorphic function f on XzS
extends holomorphically to X. This is major difference with complex analysis in C: f pzq “ 1{z
does not extend on Czt0u, but any holomorphic f pz1 , z2 q extends on C2 zt0u.
Example 4.6. Here is another example from Greene-Morrison-Strominger [12] (see also the expo-
sition in [23]). Define a singular quintic X Ă P4 by the polynomial
where G “ Z34 ` Z24 ´ Z04 and H “ ´Z44 ´ Z14 ´ Z04 . The singular points are where DP “ 0, which
happens when
Z3 “ 0, Z4 “ 0, G “ 0, H “ 0. (4.20)
There are 16 singular points. (This is also expected by Bezout’s theorem: n homogeneous poly-
nomials of degrees di in projective space of dimension n has d1 ¨ dn intersection points generically.
16 “ p1qp1qp4qp4q singular points.) We now look at the local model for these singularities. Suppose
p P X is a singular point with p P U0 “ tZ0 ‰ 0u. In coordinates zi “ Zi {Z0 the equation for X is
Since gpzq has surjective derivative, by the implicit function theorem there is a holomorphic change
of coordinates
w1 “ gpz1 , z2 , z3 , z4 q, wi`1 “ zi`1 (4.22)
with inverse z1 “ ϕpw1 , w2 , w3 , w4 q. We can repeat this for hpzq and obtain coordinates w̃i with
w̃2 “ hpw1 , w2 , w3 , w4 q. Then in these coordinates, the equation for X is a neighborhood of the
origin in
tw̃3 w̃1 ` w̃4 w̃2 “ 0u Ă C4 . (4.23)
67
This is the model for a nodal singularity. There is another change of coordinates
w̃1 ` iw̃2 w̃1 ´ iw̃2 w̃3 ` iw̃4 w3 ´ iw4
z1 “ , z2 “ , z3 “ , z4 “ . (4.24)
2 2 2 2
such that this singularity is represented by
4
ÿ
t zi2 “ 0u Ă C4 . (4.25)
i“1
68
5 Conifold Transitions: Local Model
5.1 Blowing-up a nodal singularity
5.1.1 Blow-up review
We start this section by recalling the blow-up construction. The blow-up of Cn at 0 is
where x P rus means x “ λu for some λ P C. The exceptional divisor E Ă X is the set of points
p0, rusq, so that x “ 0 and E is a copy of Pn´1 . The projection
π : X Ñ Cn , πpx, rusq “ x
such that
x y
“ .
u v
To obtain coordinate charts on X, we use the coordinate charts on P1 . So in this case, there are
two charts:
x “ x
y “ xv
u “ 1
v “ v,
so that we only keep the two coordinate px, vq on this patch on X. The exceptional divisor (where
px, yq “ p0, 0q) appears as E X U “ tx “ 0u, and v is a free coordinate.
x “ uy
y “ y
u “ u
v “ 1,
and we only keep the coordinates py, uq on this set. The exceptional divisor is E X V “ ty “ 0u
with free coordinate u.
x “ uy, v “ u´1 .
69
Indeed, in chart U , we have v “ y{x which substituting in chart V givs v “ y{uy “ u´1 . The
change of coordinates v “ u´1 is the coordinate change on P1 . The change x “ uy is the coordinate
change on a line bundle over P1 . Recall in general a section over a line bundle over P1 transforms
as
σ̃pvq “ τ puqσpuq
where τ puq is a local function on P1 which is the transition function of the bundle. In the case
above, we have τ puq “ u. This transition function defines a line bundle
Op´1q Ñ P1 .
So X is the total space of this line bundle. The other way to define Op´1q is to cover P1 “ tru1 : u2 su
by U1 “ tu1 ‰ 0u and U2 “ tu2 ‰ 0u. Then declare the transition function on the overlap U1 X U2
to be
u1
τ12 “ ,
u2
so that local sections transform as σ1 “ τ12 σ2 . Coordinates over U1 are ζ “ u2 {u1 and coordinates
over U2 are ξ “ u1 {u2 . Then ξ “ ζ ´1 and over U2 , we do have τ12 pξq “ ξ. Thus this is the same
space as above.
5.1.2 ODP in C3
We now illustrate how the blow-up procedure can be used to resolve singularities. Consider the
space
X “ txz ´ y 2 “ 0u Ă C3 .
This space is of the form F px, y, zq “ 0 with DF “ pz, ´2y, xq, and there is a singularity preventing
it from being a submanifold at p0, 0, 0q. We can resolve this singularity
σ : X̃ Ñ X
x “ x, y “ vx, z “ wx.
xpwxq ´ pvxq2 “ 0
w ´ v2 “ 0
70
which is smooth, and called the proper transform. Therefore X̃ has two coordinates px, vq, and the
exceptional divisor in X is the curve x “ 0 with v a free coordinate.
x “ uy, y “ y, z “ wy,
x “ uz, y “ vz, z “ z.
u ´ v 2 “ 0.
Therefore X̃ has two coordinates pz, vq, and the exceptional divisor is the curve z “ 0 with v a free
coordinate.
Next, we note that X̃ is the total space of the line bundle Op´2q Ñ P1 . To see the P1 coordinate,
convert from U to V on U X V to obtain
5.1.3 ODP in C4
Before discussing singularities, we recall that the blow-up of Cn along a subspace Z replaces Z by
the projectivization of its normal bundle PpN Zq. We give the concrete example of the blow-up
of
Z “ tz2 “ z4 “ 0u Ă C4 .
The blow-up along Z is
71
in other words
z2 z4
“ .
u v
The P1 provides two charts for Z.
z1 “ z1 , z2 “ z2 , z3 “ z3 , z4 “ vz2 , u “ 1, v“v
z1 “ z1 , z2 “ uz4 , z3 “ z3 , z4 “ z4 , u “ u, v“1
V “ tz1 z2 ´ z3 z4 “ 0u Ă C4 .
Note that this singulariy is the same as t i zi2 “ 0u by a change of coordinates (4.24). We can
ř
desingularize this space by blowing-up C4 along Z “ tz2 “ z4 “ 0u and taking the proper transform
Ṽ .
‚ Chart U “ tu “ 1u. Using the relations above, the equation for V becomes
z1 z2 ´ z3 pvz2 q “ 0.
z1 ´ vz3 “ 0,
which is now smooth. Therefore here Ṽ has three coordinates pz2 , z3 , vq, and the relations are
z1 “ vz3 , z2 “ z2 , z3 “ z3 , z4 “ vz2
on U .
uz1 ´ z3 “ 0,
and here Ṽ has three coordinates pz1 , z4 , uq and the relations are
z1 “ z1 , z2 “ uz4 , z3 “ uz1 , z4 “ z4 .
on W .
We note that Ṽ can be identified with the total space of Op´1q ‘ Op´1q Ñ P1 . For this, we
compute the change of coordinates on the overlap U X W . Relabel
72
ps̃, σ̃, ξq “ pz1 , z4 , uq on W.
Then chasing relations gives for example,
s̃ “ z1 “ vz3 “ ζs,
and altogether
s̃ “ ζs, σ̃ “ ζσ, ξ “ ζ ´1 .
This is the change of coordinates on the total space of Op´1q ‘ Op´1q Ñ P1 with fiber coordiantes
s, σ and P1 coordinate ζ. Thus we have a resolution of singularities
σ : Ṽ Ñ V
where the origin in V is replaced by σ ´1 p0q “ P1 . This can be seen since in local coordinates on U
then σ ´1 p0q is freely parametrized by coordinate v (and coordinates z2 “ z3 “ 0), and on V then
σ ´1 is freely parametrized by coordinate u (and coordinates z1 “ z4 “ 0), and on the overlap we
noted u “ v ´1 . We note that the modulus
4
ÿ
}z}2 “ |zi |2
i“1
We recognize this as
}z}2 “ |pσ, ζq|2hF S
where hF S is the Fubini-Study metric on the Op´1q fibers.
In summary, the total space of the bundle Op´1q ‘ Op´1q Ñ P1 is given by two trivializations
tU, pσ, s, ζqu and tV, pσ̃, s̃, ξqu with ξ “ 1{ζ and s̃ “ ζs, σ̃ “ ζσ. The set }z} “ 0 is a P1 .
Definition 5.1. A p´1, ´1q curve C in a compact threefold X is defined by C – P1 and there
exists a neighborhood of C which is biholomorphic to a neighborhood of t}z} “ 0u in the total space
Op´1q ‘ Op´1q Ñ P1 .
We see that V is a cone, since if z P V then so is λz for λ P C. We also have the radius function
ÿ
}z}2 “ |zi |2 .
i
73
On this cone, we have the holomorphic volume form
1
Ω“ dz2 ^ dz3 ^ dz4 , on tz1 ‰ 0u
z1
and the corresponding formula on the other open sets tzi ‰ 0u. These local expressions glue on V
to give a global holomorphic volume form. Next, we write
zk “ xk ` iyk .
}z}2 }z}2
|x|2 “ , |y|2 “ , xx, yy “ 0.
2 2
For each fixed r ą 0, the cross-section where }z} “ r is seen from here to be a S 2 bundle over S 3 .
These are topologically trivial, so for each r ą 0, the cross section where }z} “ r is S 3 ˆ S 2 .
The smoothing of V is given by
4
ÿ
Vt “ t pzi q2 “ tu.
i“1
ř4 2
This is smooth because the only point where the derivative of i“1 pzi q is not surjective is at the
origin, which is not included in Vt .
After rotating coordinates z, we may suppose t ą 0. The defining equation becomes
|x|2 “ t, y “ 0.
The space of x describes an S 3 of vanishing radius as t Ñ 0. This is sometimes called the vanishing
sphere.
Lastly, we note that Vt also admits a holomorphic volume form, given by local expressions such
as
1
Ωt “ dz2 ^ dz3 ^ dz4 , on tz1 ‰ 0u.
z1
74
5.3 Candelas-de la Ossa metrics
To summarize our discussion so far, we have described two ways to resolve the singular space
ÿ
V “ t zi2 “ 0u Ă C4
The first is by small blow-up, which replaced the origin by P1 “ S 2 . We call this the small
resolution
σ : Ṽ Ñ V.
The second is by smoothing, which replaced the origin by S 3 .
ÿ
Vt “ t zi2 “ tu Ă C4 .
Candelas-de la Ossa [2] constructed a sequence of metrics on both sides. We will discuss how one
side of the sequence sends the area of a holomorphic curve P1 to zero, and the other side sends the
area of a special Lagrangian 3-sphere to zero.
which measures the distance to the zero section P1 (coordinate λ) along the Op´1q fibers (coordi-
nates u, v) using the Fubini-Study metric. For a ą 0, the Candelas-de la Ossa ansatz is
We want to solve for fa such that these metrics are Ricci-flat. We expand the ansatz as
„
iB f τ B̄ log τ ` 4a2 ωF S
1
75
3
To compute the Ricci curvature, we first compute ωco,a . We note that these 3 terms each square
to zero. For example
„ 2
2 2
iB B̄ logp|u| ` |v| q “ 0. (5.2)
Indeed,
"„ „ *
1
B B̄ logp|u|2 `|v|2 q “ p|u|2 `|v|2 qpdu^dū`dv ^dv̄q ´ pūdu` v̄dvq^pudū`vdv̄q .
p|u|2 ` |v|2 q2
Squaring this is of the form pα ´ βq2 where β 2 “ 0. A direct computation then shows α2 ´ 2αβ “
2p|u|2 ` |v|2 q2 ´ 2p|u|2 ` |v|2 q2 “ 0. This verifies (5.2).
3
Going back to (5.1), computing ωco,a is of the form pa ` b ` cq3 “ 6abc since a2 “ b2 “ c2 “ 0. We
have
3
ωco,a “ 6pf 2 ` τ ´1 f 1 qpf 1 τ qpf 1 τ ` 4a2 qiBτ ^ B̄τ ^ ωF S ^ iB B̄ logp|u|2 ` |v|2 q.
Next, we have
ωF S “ p1 ` |λ|2 q´2 idλ ^ dλ̄,
and
ˆ ˙ˆ ˙
2 2 2 2 2 2
Bτ ^ B̄τ “ p|u| ` |v| qλ̄dλ ` p1 ` |λ| qpūdu ` v̄dvq p|u| ` |v| qλdλ̄ ` p1 ` |λ| qpudū ` vdv̄q
and
ˆ ˙
1
B B̄ logp|u|2 ` |v|2 q “ p|u| 2
` |v|2
qpdu ^ dū ` dv ^ dv̄ ´ pūdu ` v̄dvq ^ pudū ` vdv̄q
p|u|2 ` |v|2 q2
Using this, direct calculation gives
iBτ ^ B̄τ ^ ωF S ^ iB B̄ logp|u|2 ` |v|2 q “ cpidλ ^ dλ̄q ^ pidu ^ dūq ^ pidv ^ dv̄q.
To find a Ricci-flat metric, from Rj k̄ “ ´Bj Bk̄ log det g we look for solutions to det gco,a “ const,
which is the equation (with convenient choice of constant)
2
γ 1 γpγ ` 4a2 q “ τ.
3
76
This has solution
γ 3 ` 6a2 γ 2 “ τ 2 .
This is
τ pfa1 q3 ` 6a2 pfa1 q2 “ 1.
From here, we note that if f1 is a solution for a “ 1, then fa “ a2 f1 pτ {a3 q is a solution for arbitrary
a. We discuss the solution f for a “ 1 and obtain the other fa by rescaling.
3 2 1
We look
? for a positive solution of τ y ` 6y “ 1 with y “ f pτ q. At τ “ 0, we choose the solution
y “ 1{ 6. It turns out that there is an explicit solution ypτ q ą 0 on τ ě 0 given by
„
1 4
y“ z ` ´ 2 , z “ 2´1{3 pτ 2 ` pτ 4 ´ 32τ 2 q1{2 ´ 16q1{3 . (5.3)
τ z
By construction, it solves
pf 2 f 1 τ ` pf 1 q2 qpf 1 τ ` 4q “ 1.
The form ωco,1 is ą 0, since it is given by (5.1) with f 1 ą 0 and f 2 τ ` f 1 ą 0.
Next, we show that fa Ñ f0 smoothly on compact sets away from u “ v “ 0. We study the
asymptotic behavior of f1 as τ Ñ 8. For τ ě K, then the Puiseux series of (5.3) turns out to
be
2 16
y “ τ ´1{3 ´ ` 4τ ´5{3 ´ τ ´7{3 ` . . .
τ 3
Integrating gives
f1 pτ q “ c0 τ 2{3 ` c1 log τ ` c2 τ ´2{3 ` c3 τ ´4{3 ` . . .
Therefore when τ ě Ka3 , then
on tr ě K0 au, where r “ τ 1{3 . Therefore ωco,a Ñ ωco,0 as a Ñ 0 uniformly on compact sets which
are disjoint from the exceptional curve P1 “ tτ “ 0u.
Remark 5.2. Let us rewrite things slightly differently. Rescale the metrics so that c0 “ 1, let
r “ τ 1{3 so that r “ p1 ` |λ|2 q1{3 p|u|2 ` |v|2 q1{3 and
77
where
SR pλ, u, vq “ pλ, R3{2 u, R3{2 vq
˚
satisfies SR ˚
r “ Rr and SR ωco,0 “ R2 ωco,0 . Let AR “ tR ď r ď 2Ru and note SR : A1 Ñ AR . We
pullback the compact set estimate
to obtain
|iB B̄Sρ˚ pf1 ´ f0 q|Sρ˚ ω0 ď C, tρ´1 ď r ď 2ρ´1 u
and
|ωco,1 ´ ω0 |ωco,0 ď Cr´2 .
We now pullback this estimate via Sa´1 .
Therefore
|Sa˚´1 ωco,1 ´ a´2 ωco,0 |ωco,0 ď Cr´2 ,
and
|ωco,a ´ ωco,0 |ωco,0 ď Cr´2 a2
which shows that ωco,a Ñ ωco,0 as a Ñ 0 on sets tr ą εu.
ϕpzq “ f pτ q,
We compute
B̄ϕ “ f 1 B̄τ,
and
iB B̄ϕ “ f 1 iB B̄τ ` f 2 iBτ ^ B̄τ.
78
Its Monge-Ampère mass is
piB B̄ϕq3 “ pf 1 q3 piB B̄τ q3 ` 3pf 1 q2 f 2 piB B̄τ q2 ^ iBτ ^ B̄τ. (5.4)
We have ÿ
Bτ ^ B̄τ “ wj w̄i dwi ^ dw̄j , B B̄τ “ dwk ^ dw̄k .
k
We now compute in coordinates pw1 , w2 , w3 q on Vt over the chart tw4 ‰ 0u. On tw4 ‰ 0u, we
have
1
Ωt “ dw1 ^ dw2 ^ dw3 .
w4
Therefore
1
iΩt ^ Ω̄t “ pidw1 ^ dw̄1 q ^ pidw2 ^ dw̄2 q ^ pidw3 ^ dw̄3 q.
|w4 |2
We write
1
dµ “ pidw1 ^ dw̄1 q ^ pidw2 ^ dw̄2 q ^ pidw3 ^ dw̄3 q
|w4 |2
for simplicity. We need to equate this to piB B̄ϕq3 (5.4). For this we will use the defining relation of
Vt , which gives
4
ÿ 1
wk dwk “ 0, dw4 “ ´ pw1 dw1 ` w2 dw2 ` w3 dw3 q.
k“1
w 4
We also have
λ11̄ ^ λ44̄ ^ iBτ ^ B̄τ
“ λ11̄ ^ λ44̄ ^ p|w2 |2 λ22̄ ` |w3 |2 λ33̄ ` w3 w̄2 λ23̄ ` w2 w̄3 λ32̄ q (5.8)
79
which by the defining relation becomes
By symmetry,
pf 1 q3 τ ` pf 1 q2 f 2 pτ 2 ´ |t|2 q “ 1{6,
for a function f pτ q, and where t is a fixed parameter. Here is the result. When t “ 0, the solution
is proportional to
f0 pτ q “ τ 2{3
and for given t ‰ 0, the solution is proportional to
ż cosh´1 pτ {|t|q
ft pτ q “ 2´1{3 |t|2{3 psinhp2λq ´ 2λq1{3 dλ.
0
Remark 5.3. Let’s work out how to find this solution when t “ 0. The ODE is
6τ rpf 1 q3 ` τ pf 1 q2 f 2 s “ 1.
80
which admits the solution γ 3 “ 14 s4 . Therefore
1 4
s6 rf 1 ps2 qs3 “ s
4
which for f 1 ą 0 is the ODE
f 1 pτ q “ c0 τ ´1{3 .
The solution is f pτ q “ c1 τ 2{3 .
Remark 5.4. Let’s compute the asymptotics of f1 . Its derivative is
ˆ ˙1{3
1
f11 pτ q “ sinhp2 cosh´1 pτ qq ´ 2 cosh´1 τ ? .
τ2 ´ 1
?
We have sinhp2 cosh´1 pτ qq “ 2τ pτ ´ 1q1{2 pτ ` 1q1{2 and cosh´1 τ “ logpτ ` τ 2 ´ 1q. Then as
τ Ñ 8, the series expansion turns out to be
Integrating gives
f pτ q “ c0 τ 2{3 ` a1 τ ´4{3 log τ ` a2 τ ´4{3 ` Opτ ´7{3 q.
Therefore, if we let r “ τ 1{3 , then
Remark 5.5. The metrics ωco,1 are asymptotically conical. We now describe what this means (see
[6] for more on asymptotically conical Kähler-Ricci flat metrics). Let the Kähler-Ricci flat metrics
be denoted
ωco,0 “ iB B̄ϕ0 , ωco,1 “ iB B̄ϕ1 .
We consider the map Φ : V X t|z| ą Ru Ñ V1 given by
x̄
Φpxq “ x ` .
2|x|2
We will estimate the order of Φ˚ ωco,1 ´ ωco,0 . By the asymptotics of ϕ1 and r “ }x}2{3 ,
Therefore
Φ˚ ϕ1 ´ ϕ0 “ Opr´1 q.
Next, we start from
sup |Φ˚ ωco,1 ´ ωco,0 |ωco,0 ď C,
1ďrď2
81
and pull this back via Sλ : V Ñ V , Sλ pzq “ λ3{2 z to get
The definition of Sλ is such that Sλ˚ r “ λr and Sλ˚ ωco,0 “ λ2 ωco,0 . Since
we conclude
|Φ˚ ωco,1 ´ ωco,0 |ωco,0 ď Cr´3 . (5.11)
Metrics satisfying estimates of this form are said to be asymptotically conical (with rate 3).
Remark 5.6. Next, we notice that Sλ pzq “ λ3{2 z implies
St1{3 : V1 Ñ Vt .
|t´2{3 Φ˚t ωco,t ´ t´2{3 ωco,0 |t´2{3 ωco,0 “ |Φ˚t ωco,t ´ ωco,0 |ωco,0
gives the estimate (5.12). This estimate can be interpreted as ωco,t Ñ ωco,0 on compact sets away
from the cone singularity as t Ñ 0.
82
identification. So we will work on V , and the Candelas-de la Ossa cone metric derived in the two
previous sections is, up to a constant,
ωco,0 “ iB B̄r2
where r “ }z}2{3 .
3
Here is the reason why the power 2/3 is natural. We are looking to solve ωco,0 “ iΩ ^ Ω̄. We can
rescale z ÞÑ λz on both sides, which gives
which is consistent.
Since the cone radius is r, the natural scaling on the cone is
g “ dr b dr ` r2 gΣ . (5.13)
This type of metric is very useful in Riemann geometry. To see why, convert the Euclidean metric
gEuc on Rn to polar coordinates pr, θi q. It will be of the form (5.13) with gΣ the metric on the
sphere S n´1 . Riemannian geometry with metrics of the form (5.13) behaves like polar coordinates:
for example, the kernel of the Laplacian ∆g can be decomposed by separation of variables.
Here is why ωco,0 can be put in this form. We insert a factor of p1{2q in the definition of ωco “ 12 iB B̄r2
for convenience. Compute
83
We can write this as ˆ ˙
2 i
ωco “ ´dr ^ Jdr ` r ´ dJd log r
2
where pJβqpXq “ βpJXq. This can be seen by writing dr “ Br ` B̄r and Jdr “ iBr ´ iB̄r.
Let η “ ´Jd log r. To obtain the metric, we use gco pX, Y q “ ωco pX, JY q.
with rBr “ 32 xi BxB i ` 23 yi ByB i . This vector field generates the t¨ action on V in the sense
ˇ
d ˇˇ
ξpf q “ 2 f pt ¨ z, z̄q.
dt ˇt“1
In R2 with polar coordinates reiθ , we can think of JprBr q “ rBθ . This discussion implies ηpBr q “
´JpBr q log r “ 0 and so
η “ ηi pr, θqdθi .
The next step is to show that ηi does not depend on r, and for this we will use the Lie deriva-
tive.
‚ Next, we note
Lξ J “ 0.
The Lie derivative is
Lξ pJXq “ pLξ JqpXq ` JLξ pXq
and Lξ pXq “ rξ, Xs for vector fields X. So this amounts to showing
84
‚ We now show that
LrBr η “ 0.
By Cartan’s formula, LrBr “ dιrBr η ` ιrBr dη. Since ιrBr η “ 0, we need to show
dηprBr , ¨q “ 0. (5.16)
Indeed, the invariant formula for dη gives
dηprBr , Xq “ rBr ηpXq ´ XηprBr q ´ ηprrBr , Xsq
“ ´rBr JX log r ` JrrBr , Xs log r
“ ´rBr JX log r ` rrBr , JXs log r
“ ´JXrBr log r “ 0. (5.17)
Here we used the real part of rξ, JXs “ Jrξ, Xs (5.15).
‚ Altogether, since ˇ
d ˇˇ
0 “ LrBr η “ ηi ptr, θqdθi
dt ˇt“1
it follows that in a local chart we can write η “ ηi pθqdθi and ηi does not depend on r. In other
words, η “ p˚ ηΣ where p : V Ñ Σ is ppr, θq “ θ. Similarly, (5.16) implies
dηp¨, J¨q “ αij pr, θqdθi b dθj ,
and LrBr dη “ 0 and LrBr J “ 0 implies that αij pθq does not dependent on r. We have proved (5.14),
and obtain the cone formula g “ dr2 ` r2 gΣ .
pointwise. The norm is defined as: write ϕ|L “ f pxqdvolL and then take |f pxq|. Here dvolL is
the volume form of the induced metric g|L . A submanifold L is calibrated with respect to ϕ if
ϕ|L “ dvolL .
Let rLs P Hk pM, Zq be a fixed homology class. Any other representative can be written as L1 “
L ´ BΣ. We consider the functional on rLs defined by
ż ż
EpL1 q “ dvolL1 ` dϕ, L1 “ L ´ BΣ. (5.18)
L1 Σ
In the standard definition of a calibration, we require dϕ “ 0 and the bulk term does not appear so
that this is the area functional. The more general definition is here for potential future application
to non-Kähler geometry.
85
Proposition 5.7. Let L be a calibrated submanifold with respect to ϕ. Then L minimizes the
functional EpLq in a given homology class.
When dϕ “ 0, this is the topological number rϕs ¨ rLs. We will show that this is the lower bound
of the functional. By Stokes’s theorem,
ż „ż ż
1
EpL q “ dvolL `
1 ϕ´ ϕ .
L1 L L1
Let Ω be a holomorphic volume form. Let ω be a hermitian metric, conformally rescaled so that
|Ω|ω “ 23{2 . A special Lagrangian cycle is a submanifold calibrated with respect to Re Ω. The
calibration argument implies that special Lagrangian cycles minimize the area functional
ż
EpLq “ dvolL
L
Proof. We closely follow Harvey-Lawson’s proof [15]. Let L Ď M . Suppose v1 , v2 , v3 are orthonor-
mal vectors spanning Tp L. Harvey-Lawson’s identity is
In fact, equality in Hadamard’s inequality is achieved if and only if the vectors are orthogonal. This
implies xvi , Jvk yg “ 0, which translates to ωpvi , vk q “ 0 and so ω|L “ 0. The inequality above is
Therefore, since the vi are orthonormal, then |Re Ω|L |g ď 1 and Re Ω is a calibration. If equality
Re Ω|L “ dvolL holds, then ω|L “ 0 and Im Ω|L “ 0.
86
We now prove Harvey-Lawson’s identity (5.19). We start with the left-hand side. Let e1 , e2 , e3 , Je1 , Je2 , Je3
be an orthonormal basis for Tp M . Since Ω|p is an element of the 1 dimensional vector space Λ3,0 p
it can be written Ω|p “ f ppqε1 ^ ε2 ^ ε3 with
1
εk “ pek ´ iJek q.
2
To find f ppq P C, we take the norm and use that |Ω|ω “ c. For suitable normalization (c “ 23{2 as
will be computed below), we conclude Ω|p “ eiθppq ε1 ^ ε2 ^ ε3 .
Next, we expand vi in the basis εk , εk
¯
vk “ A` k ε` ` A` k ε` .
¯
Since vk “ vk , using uniqueness of expansion of a basis we obtain that A` k “ A` k . Since Ω|p “
eiθ ε1 ^ ε2 ^ ε3 , we have
Next, we compute the right-hand side of (5.19). Let e4 “ Je1 , e5 “ Je2 , e6 “ Je3 so that tei u is
an oriented basis. Define a linear map M by its action on this basis
Then
|v1 ^ Jv1 ^ . . . v3 ^ Jv3 | “ |M pe1 q ^ M pe2 q ^ ¨ ¨ ¨ ^ M pe6 q| “ | det M |
To compute det M , we compute it in the basis εi , εi . We compute
1 1 i
M pεk q “ pM pek q ´ iM pJek qq “ vk ´ Jvk .
2 2 2
Using our earlier matrix Ai k and Jεk “ iεk , Jεk “ ´iεk , this is
1 n i
M pεk q “ pA k εn ` An̄ k εn q ´ piAn k εn ´ iAn̄ k εn q “ An k εn .
2 2
Similarly M pεk q “ Āεk , and so in this basis
» fi
A 0
M “– fl .
0 Ā
A useful formula for calibrated cycles are the special Lagrangian equations. The submanifold L is
special Lagrangian if it solves the equations
87
Proposition 5.9. If L Ď pM, ω, Ωq solves the special Lagrangian equations (5.20), then L is a
calibrated cycle. This means we can orient L such that Re Ω|L “ dvolL and L minimizes the area
functional in its homology class.
eiθ |Ω|ω
Ω|L “ dvol. (5.21)
23{2
Therefore the conditions ω|L “ 0, |Ω|ω “ 23{2 and ImΩ|L “ 0 imply Re Ω|L “ ˘dvol|L .
To prove (5.21), fix a point p P Tp L and an orthonormal basis e1 , e2 , e3 of Tp L with dual basis ek .
Since ωpei , ej q “ gpei , Jej q, the condition ω|L “ 0 implies that
εk “ ek ` iJek . (5.22)
ω3
iΩ ^ Ω “ |Ω|2ω .
3!
We see that
|Ω|2ω 11̄ 22̄ 33̄
i|f |2 ε1231̄2̄3̄ “ iε iε iε .
23
Therefore |f |2 “ |Ω|2 {23 and
eiθppq |Ω|ω 1
Ω|p “ ε ^ ε2 ^ ε3
23{2
Thus
eiθ |Ω|ω 1
Ω|L “ e ^ e2 ^ e3 ,
23{2
which is (5.21).
88
5.4.2 Examples on the smoothing
We give two examples of special Lagrangian cycles on
If zk “ xk ` iyk , the constraint for Vt implies t “ |x|2 ´ |y|2 , xx, yy “ 0, and the constraint for L is
|x|2 ` |y|2 “ t. Therefore y “ 0 and |x|2 “ t, so L is topologically a 3-sphere. Since y “ 0, we see
that pIm Ωt q|L “ 0. Since τ is constant, we see that ωco,t |L “ 0.
‚ Special Lagrangian discs. Here we will use the metric ωEuc “ k idz k ^ dz̄ k restricted to Vt . It
ř
is not Calabi-Yau, so this example involves a more general setup. Consider the two discs L` , L´
given by
L˘ “ tx21 ` x22 ` x23 ď tu Ď Vt X ty “ 0u.
with b
z4 “ ˘ t ´ x21 ´ x22 ´ x23 .
These share the same boundary BL which is a 2-cycle S 2 which lies on the holomorphic surface
z4 “ 0. Since y “ 0, we have pIm Ωt q|L “ 0 and ωEuc |L “ 0.
89
6 Conifold Transitions: Global Geometry
6.1 Overview
Let X̂ be a compact Calabi-Yau threefold. A conifold transition X̂ Ñ X0 ù Xt is defined as
follows:
Step 1: Find holomorphic curves Ci Ă X̂ with a neighborhood biholomorphic to the open set
t}z} ă 1u in the total space Op´1q ‘ Op´1q Ñ P1 , with Ci “ t}z} “ 0u. Such curves Ci are called
p´1, ´1q curves.
Step 2: Contract each Ci to a point pi and obtain singular
ř4 space X0 . A change of variables shows
that each pi admits a neighborhood biholomorphic to t i“1 zi2 “ 0u Ď C4 (§5.1.2).
Step 3: Realize X0 as the central fiber of a deformationřπ : X Ñ ∆ such that Xt is smooth for
t ‰ 0 and locally near pi we see the local smoothing t zi2 “ tu. For this to exist, we need to
satisfy Friedman’s condition on the initial curves, and the global smoothing result of Friedman-
Tian-Kawamata is stated below. The smoothing has the effect of replacing each pi P X0 with
3-sphere S 3 (§5.2).
Theorem 6.1. [9, 27, 18] Let X̂ be a Calabi-Yau threefold with p´1, ´1q curves Ci Ă X̂. Let
µ : X̂ Ñ X0 be the blow-down map sending Ci to nodal singularities pi P X0 . If the curves are
linearly dependent in homology, so that
ÿ
λi rCi s “ 0 P H 4 pX̂, Cq
i
with λi ‰ 0, then X0 admits a smoothing Xt . This means there is a complex space X with a proper
flat map π : X Ñ ∆ such that π ´1 p0q “ X0 and π ´1 ptq “ Xt is a smooth complex manifold for
t ‰ 0.
ř
Let Ci be an initial configuration of p´1, ´1q curves on X̂ satisfying Friedman’s condition λi rCi s “
0. The theorem above guarantees the existence of a conifold transition X̂ Ñ X0 ù Xt . Here is a
summary of known results on the geometry and topology of such a transition:
‚ Xt is a compact complex manifold with trivial canonical bundle. [9]
‚ It is conjectured [1] that all Calabi-Yau threefolds can be connected to each other by conifold
transitions.
‚ However, a conifold transition may produce an Xt which is non-Kähler. (Examples at the end
of §6.2) This suggests that these limiting non-Kähler objects should be included in the web of
Calabi-Yau threefolds.
‚ Though non-Kähler in general, Xt is expected to satisfy the ddbar lemma, but as far as I can
tell, this is still unknown in full generality [10].
‚ The topological change is
N “ k ` c,
where N is the number of nodes, k is the decrease of b2 , and 2c is the increase of b3 . (§6.2)
‚ [11] There is a sequence of metrics pX̂, ga q with dωa2 “ 0 such that ga Ñ g0 uniformly on compact
sets for a limiting metric g0 on X0 locally modeled near the singularities by a scaling of gco,0 ,
90
and
VolpCi , ga q Ñ 0, a Ñ 0.
On the smoothing side, for small t there is a sequence of metrics pXt , gt q with dωt2 “ 0 and gt Ñ g0
uniformly on compact sets with
VolpLi,t , gt q Ñ 0, t Ñ 0,
where Lt,i “ t}z}2 “ |t|u Ď Vt Ď Xt are the vanishing 3-spheres. This result is due to Fu-Li-Yau
[11]. Near the singularities, the FLY metrics are modeled by the Candelas-de la Ossa metrics:
Here r : X Ñ p0, 8q is a function which agrees with r “ }z}2{3 near the singular points and
r´1 p0q “ SingpX0 q. Each ith component of Xt X tr ă 1u is biholomorphic to Vt X tr ă 1u, and the
metric gt is close to a λi -scaled version of gco,t for λi ą 0. For higher order derivatives, the estimate
is
|∇kgco,t pgt ´ λgco,t q|gco,t ď Ck |t|2{3 r´k , on X X tr ă 1u.
Let us give a few more details of the Fu-Li-Yau construction. There are two steps:
The correction θ comes from solving Et pγt q “ B̄ ω̂t2 (Et is the Kodaira-Spencer operator) and
θt “ B B̄ : B : γt , ´B̄θt “ Et pγt q. Fu-Li-Yau show that the correction is θt is small: it satisfies |θt |ω̂t ď
C|t|2{3 .
‚ [5] The vanishing cycles Lt,i Ď Xt can be represented by special Lagrangian 3-spheres with respect
to the global geometry gt . Thus from the point of view of submanifold geometry, the transition
exchanges holomorphic 2-cycles with special Lagrangian 3-cycles.
‚ [4] The compact manifolds Xt for small t admits a pair of metrics pgt , ht q solving
g j k̄ Hpj k̄ “ 0, g j k̄ F p qj k̄ “ 0
91
where H is the 3-form field strength H “ ipB ´ B̄qω and F is the 2-form field strength F “
B̄ph´1 Bhq. In particular, Xt admits balanced metrics and has stable tangent bundle. (The equation
g j k̄ Hpj k̄ “ 0 is equivalent to dω 2 “ 0.) Note that Calabi-Yau metrics gCY solve these equations for
h “ g “ gCY , so these are generalizations of Calabi-Yau metrics on the non-Kähler spaces reached
by conifold transitions.
‚ It is conjectured by S.-T. Yau that the pair pg, hq can be further deformed to solve the Strominger
system:
dp|Ω|ω ω 2 q “ 0, g j k̄ F p qj k̄ “ 0
iB B̄ω “ α1 pTr Rω ^ Rω ´ Tr Fh ^ Fh q. (6.1)
2
The equation dp|Ω|ω ω q “ 0 can be solved by conformally changing the Fu-Li-Yau metric on Xt ,
but whether (6.1) is solvable through conifold transitions is unknown.
92
Let Û “ YN 2 4
i“1 Ûi Ď X̂ be an open set containing all curves Ci with Û “ \S ˆ B , and Ut “
Yi“1 Ut,i Ď Xt be an open set containing all vanishing cycles with Ut “ \S ˆ B 3 . We now
N 3
while on X, we have
and on Xt we have
To connect these diagrams, we use the contraction maps X̂ Ñ X and Xt Ñ X, and that X̂zÛ and
Xt zUt are both diffeomorphic to XzU (as they are a smooth family of complex structures). The
5-lemma implies that H1 pX̂q “ H1 pXq “ H1 pXt q. Recall the 5-lemma: given a diagram between
two exact sequence
A /B /C /D /E
f g h i j
A1 / B1 / C1 / D1 / E1
If g, i are isomorphisms, f surjective and j injective, then h is an isomorphism.
Since we are assuming that X̂ is a simply connected Calabi-Yau threefold, it follows that b1 pXt q “ 0
on the other side of a conifold transition.
Lemma 6.3. b2 pX̂q “ b2 pXt q ´ κ, where κ is the dimension of the subspace in H2 pX̂, Rq spanned
by the rCi s.
Proof. Let C “ YCi Ă X̂ be degenerating 2-cycles and L “ YLi Ă Xt be the vanishing 3-cycles.
The long exact sequence for relative homology gives
ι˚ ι˚
¨ ¨ ¨ Ñ H2 pCqÑH2 pX̂q Ñ H2 pX̂, Cq Ñ H1 pCqÑ ¨ ¨ ¨ (6.4)
Recall that relative homology HpX̂, Cq means: homology of space where we identify all points in
C to be a single point (or cycles α P Zn pX̂q such that Bα P Zn´1 pCq). But we will not need to have
intuition for this definition here; the only properties we will use is the exact sequence above, and
the Lefschetz duality
Hi pX̂, Cq – Hcn´i pX̂zCq
93
which holds for topological manifolds and implies in our setup that
since X̂zC – Xt zL. The long exact sequence for relative homology gives
N
ğ ι˚ ϕ
rCi sÑH2 pX̂qÑH2 pX̂, Cq Ñ 0.
i“1
which for L “ \N 3
i“1 S implies H2 pXt q – H2 pXt , Lq.
Lemma 6.4. N “ k ` c
ř6 k
Proof. Recall the Euler characteristic χ “ k“1 p´1q bk . The excision property for the Euler
characteristic is: for U Ď X open,
χpX̂q ´ 2N “ χpXt q.
94
Remark 6.5. If the initial manifold has b2 pX̂q “ 1 (if for example X̂ is a quintic threefold in
P4 ), then two homologically linearly dependent curves C1 , C2 produce a conifold transition where
the resulting manifold has b2 pXt q “ 0. A compact Kähler manifold cannot have b2 “ 0, since ω
represents a non-zero cohomology class in H 2 . We see that conifold transitions possibly take us out
of Kähler geometry.
Example 6.6. Here is another example of a non-Kähler transition from [2]. Start with a small
resolution X̂ of a singular quintic P “ Z3 GpZ0 , . . . , Z4 q ` Z4 HpZ0 , . . . , Z4 q “ 0. This has h1,1 “ 2,
and there are 16 p´1, ´1q curves Ci Ď X̂ from the small resolution, but they are all linearly
dependent. Suppose they are generated by C1 . Take 2 curves C1 , C2 ; they satisfy Friedman’s
condition so we can produce a transition X̂ Ñ X ù Xt . Now the other 14 curves are trivial in
homology rCi s “ rBDi s. Furthermore by [Remark 3.2.8, Lemma 3.3.1] in McDuff-Salamon [22], the
p´1, ´1q curves deform along the family so that Ci,t are holomorphic curves. If Xt admits a Kähler
metric ωt , then ż ż
0ă ωt “ dωt “ 0,
Ci,t Di,t
95
References
[1] P. Candelas, P. Green, and T. Hubsch, Rolling among Calabi-Yau vacua, Nuclear Physics B
330(1) (1990), 49-102.
[2] P. Candelas and X. de la Ossa, Comments on conifolds, Nuclear Physics B 342, no. 1 (1990),
246-268.
[3] P. Candelas, X. de la Ossa, P. Green, and L. Parkes, A pair of Calabi-Yau manifolds as an
exactly soluble superconformal theory, Nuclear Physics B, Vol 359, Issue 1 (1991), 21-74.
[4] T.C. Collins, S. Picard and S.-T. Yau, Stability of the tangent bundle through conifold transi-
tions, arXiv:2102.11170.
[5] T.C. Collins, S. Gukov, S. Picard and S.-T. Yau, Special Lagrangian cycles and Calabi-Yau
transitions, arXiv:2111.10355.
[6] R. Conlon and H.-J. Hein, Asymptotically conical Calabi-Yau manifolds, I, Duke Math. J. 162
(2013), 2855-2902.
[7] J.P. Demailly, Complex Analytic and Differential Geometry. Book available on the author’s
website.
[8] S.K. Donaldson, Riemann Surfaces, Oxford University Press 2011.
[9] R. Friedman, On threefolds with trivial canonical bundles, Complex geometry and Lie theory,
Sundance, UT, 1989, Proc. Sympos. Pure Math. 53, Amer. Math. Soc. (1991), 103– 134.
[10] R. Friedman, The B B̄-lemma for general Clemens manifolds, Pure and Applied Mathematics
Quarterly 15(4) (2019), 1001-1028.
[11] J.Fu, J. Li, and S.-T. Yau, Balanced metrics on non-Kahler Calabi-Yau threefolds, J. Differen-
tial Geom. 90(1), 81-129, (2012).
[12] B.R. Greene, D.R. Morrison, and A. Strominger. Black hole condensation and the unification
of string vacua. Nuclear Physics B 451, no. 1-2 (1995), 109-120.
[13] P. Griffiths and J. Harris, Principles of algebraic geometry, John Wiley and Sons, 2014.
[14] M. Gross, D. Joyce and D Huybrechts, Calabi-Yau manifolds and related geometries, Springer
2001.
[15] R. Harvey and H.B. Lawson, Calibrated geometries, Acta Mathematica (1982), 148, pp.47-157.
[16] D. Huybrechts, Complex geometry: an introduction, Springer 2005.
[17] E. Kahler, Uber eine bemerkenswerte Hermitesche Metrik, Abh. Math. Sem. Univ. Hamburg,
9 (1933), 173–186. DOI:10.1007/BF02940642
[18] Y. Kawamata, Unobstructed deformations – a remark on a paper of Z. Ran, J. Alg. Geom. 1
(1992), 183–190.
[19] K. Kodaira, Complex manifolds and deformation of complex structures, Grundlehren der math-
matischen Wissenschaften 1986.
[20] K. Kodaira and J. Morrow, Complex manifolds, Holt, Rinehart and Winston Inc, 1971.
96
[21] K. Kodaira and D.C. Spencer, On Deformations of Complex Analytic Structures, III. Stability
Theorems for Complex Structures, Annals of Mathematics 71(1) (1960), p.43-76.
[22] D. McDuff and D. Salamon, J-holomorphic Curves and Symplectic Topology, American Math-
ematical Society 2012.
[23] M. Rossi, Geometric transitions, J.Geom.Phys. 56 (2006), 1940-1983.
[24] Y.-T. Siu, Lectures on Hermitian-Einstein metrics for stable bundles and Kähler-Einstein met-
rics, DMV Seminar, Band 8 Birkhauser 1987. Notes available on the author’s website.
[25] J. Song and B. Weinkove, An introduction to the Kahler-Ricci flow, in An introduction to the
Kahler-Ricci flow, 89-188, Lecture Notes in Math., 2086, Springer 2013.
[26] G. Szekelyhidi, An Introduction to Extremal Kahler Metrics, AMS Graduate Studies in Math-
ematics 152 (2014).
[27] G. Tian, Smoothing threefold with trivial canonical bundle and ordinary double points, Essays
on Mirror Manifolds Internat. Press, Hong Kong (1992), 458–479.
[28] V. Tosatti, KAWA lecture notes on the Kahler-Ricci flow, Annales de la Faculte des sciences
de Toulouse: Mathematiques 27 (2018), no. 2, 285-376.
[29] V. Tosatti and B. Weinkove, The complex Monge–Ampere equation on compact Hermitian
manifolds, J. Amer. Math. Soc., 23 (2010), 1187-1195.
[30] C. Voisin, Hodge Theory and Complex Algebraic Geometry II, Cambridge University Press
2003.
[31] S.-T. Yau, On the Ricci curvature of a compact Kahler manifold and the complex
Monge–Ampere equation. I, Comm. Pure Appl. Math. 31 (1978) 339-411.
97