Mathematical term
In mathematics , in particular in algebraic geometry and differential geometry , Dolbeault cohomology (named after Pierre Dolbeault ) is an analog of de Rham cohomology for complex manifolds . Let M be a complex manifold. Then the Dolbeault cohomology groups
H
p
,
q
(
M
,
C
)
{\displaystyle H^{p,q}(M,\mathbb {C} )}
depend on a pair of integers p and q and are realized as a subquotient of the space of complex differential forms of degree (p ,q ).
Construction of the cohomology groups [ edit ]
Let
Ω
p
,
q
{\displaystyle \Omega ^{p,q}}
be the vector bundle of complex differential forms of degree
(
p
,
q
)
{\displaystyle (p,q)}
. In the article on complex forms , the Dolbeault operator is defined as a differential operator on smooth sections
∂
¯
:
Ω
p
,
q
→
Ω
p
,
q
+
1
.
{\displaystyle {\bar {\partial }}:\Omega ^{p,q}\to \Omega ^{p,q+1}.}
Since
∂
¯
2
=
0
{\displaystyle {\bar {\partial }}^{2}=0}
, this operator has some associated cohomology . Specifically, define the cohomology to be the quotient space
H
p
,
q
(
M
,
C
)
=
ker
(
∂
¯
:
Ω
p
,
q
→
Ω
p
,
q
+
1
)
i
m
(
∂
¯
:
Ω
p
,
q
−
1
→
Ω
p
,
q
)
.
{\displaystyle H^{p,q}(M,\mathbb {C} )={\frac {\ker \,({\bar {\partial }}:\Omega ^{p,q}\to \Omega ^{p,q+1})}{\mathrm {im} \,({\bar {\partial }}:\Omega ^{p,q-1}\to \Omega ^{p,q})}}.}
Dolbeault cohomology of vector bundles [ edit ]
If E is a holomorphic vector bundle on a complex manifold X , then one can define likewise a fine resolution of the sheaf
O
(
E
)
{\displaystyle {\mathcal {O}}(E)}
of holomorphic sections of E , using the Dolbeault operator of E . This is therefore a resolution of the sheaf cohomology of
O
(
E
)
{\displaystyle {\mathcal {O}}(E)}
.
In particular associated to the holomorphic structure of
E
{\displaystyle E}
is a Dolbeault operator
∂
¯
E
:
Γ
(
E
)
→
Ω
0
,
1
(
E
)
{\displaystyle {\bar {\partial }}_{E}:\Gamma (E)\to \Omega ^{0,1}(E)}
taking sections of
E
{\displaystyle E}
to
(
0
,
1
)
{\displaystyle (0,1)}
-forms with values in
E
{\displaystyle E}
. This satisfies the characteristic Leibniz rule with respect to the Dolbeault operator
∂
¯
{\displaystyle {\bar {\partial }}}
on differential forms, and is therefore sometimes known as a
(
0
,
1
)
{\displaystyle (0,1)}
-connection on
E
{\displaystyle E}
, Therefore, in the same way that a connection on a vector bundle can be extended to the exterior covariant derivative , the Dolbeault operator of
E
{\displaystyle E}
can be extended to an operator
∂
¯
E
:
Ω
p
,
q
(
E
)
→
Ω
p
,
q
+
1
(
E
)
{\displaystyle {\bar {\partial }}_{E}:\Omega ^{p,q}(E)\to \Omega ^{p,q+1}(E)}
which acts on a section
α
⊗
s
∈
Ω
p
,
q
(
E
)
{\displaystyle \alpha \otimes s\in \Omega ^{p,q}(E)}
by
∂
¯
E
(
α
⊗
s
)
=
(
∂
¯
α
)
⊗
s
+
(
−
1
)
p
+
q
α
∧
∂
¯
E
s
{\displaystyle {\bar {\partial }}_{E}(\alpha \otimes s)=({\bar {\partial }}\alpha )\otimes s+(-1)^{p+q}\alpha \wedge {\bar {\partial }}_{E}s}
and is extended linearly to any section in
Ω
p
,
q
(
E
)
{\displaystyle \Omega ^{p,q}(E)}
. The Dolbeault operator satisfies the integrability condition
∂
¯
E
2
=
0
{\displaystyle {\bar {\partial }}_{E}^{2}=0}
and so Dolbeault cohomology with coefficients in
E
{\displaystyle E}
can be defined as above:
H
p
,
q
(
X
,
(
E
,
∂
¯
E
)
)
=
ker
(
∂
¯
E
:
Ω
p
,
q
(
E
)
→
Ω
p
,
q
+
1
(
E
)
)
i
m
(
∂
¯
E
:
Ω
p
,
q
−
1
(
E
)
→
Ω
p
,
q
(
E
)
)
.
{\displaystyle H^{p,q}(X,(E,{\bar {\partial }}_{E}))={\frac {\ker \,({\bar {\partial }}_{E}:\Omega ^{p,q}(E)\to \Omega ^{p,q+1}(E))}{\mathrm {im} \,({\bar {\partial }}_{E}:\Omega ^{p,q-1}(E)\to \Omega ^{p,q}(E))}}.}
The Dolbeault cohomology groups do not depend on the choice of Dolbeault operator
∂
¯
E
{\displaystyle {\bar {\partial }}_{E}}
compatible with the holomorphic structure of
E
{\displaystyle E}
, so are typically denoted by
H
p
,
q
(
X
,
E
)
{\displaystyle H^{p,q}(X,E)}
dropping the dependence on
∂
¯
E
{\displaystyle {\bar {\partial }}_{E}}
.
Dolbeault–Grothendieck lemma[ edit ]
In order to establish the Dolbeault isomorphism we need to prove the Dolbeault–Grothendieck lemma (or
∂
¯
{\displaystyle {\bar {\partial }}}
-Poincaré lemma ). First we prove a one-dimensional version of the
∂
¯
{\displaystyle {\bar {\partial }}}
-Poincaré lemma; we shall use the following generalised form of the Cauchy integral representation for smooth functions :
Proposition : Let
B
ε
(
0
)
:=
{
z
∈
C
∣
|
z
|
<
ε
}
{\displaystyle B_{\varepsilon }(0):=\lbrace z\in \mathbb {C} \mid |z|<\varepsilon \rbrace }
the open ball centered in
0
{\displaystyle 0}
of radius
ε
∈
R
>
0
,
{\displaystyle \varepsilon \in \mathbb {R} _{>0},}
B
ε
(
0
)
¯
⊆
U
{\displaystyle {\overline {B_{\varepsilon }(0)}}\subseteq U}
open and
f
∈
C
∞
(
U
)
{\displaystyle f\in {\mathcal {C}}^{\infty }(U)}
, then
∀
z
∈
B
ε
(
0
)
:
f
(
z
)
=
1
2
π
i
∫
∂
B
ε
(
0
)
f
(
ξ
)
ξ
−
z
d
ξ
+
1
2
π
i
∬
B
ε
(
0
)
∂
f
∂
ξ
¯
d
ξ
∧
d
ξ
¯
ξ
−
z
.
{\displaystyle \forall z\in B_{\varepsilon }(0):\quad f(z)={\frac {1}{2\pi i}}\int _{\partial B_{\varepsilon }(0)}{\frac {f(\xi )}{\xi -z}}d\xi +{\frac {1}{2\pi i}}\iint _{B_{\varepsilon }(0)}{\frac {\partial f}{\partial {\bar {\xi }}}}{\frac {d\xi \wedge d{\bar {\xi }}}{\xi -z}}.}
Lemma (
∂
¯
{\displaystyle {\bar {\partial }}}
-Poincaré lemma on the complex plane ): Let
B
ε
(
0
)
,
U
{\displaystyle B_{\varepsilon }(0),U}
be as before and
α
=
f
d
z
¯
∈
A
C
0
,
1
(
U
)
{\displaystyle \alpha =fd{\bar {z}}\in {\mathcal {A}}_{\mathbb {C} }^{0,1}(U)}
a smooth form, then
C
∞
(
U
)
∋
g
(
z
)
:=
1
2
π
i
∫
B
ε
(
0
)
f
(
ξ
)
ξ
−
z
d
ξ
∧
d
ξ
¯
{\displaystyle {\mathcal {C}}^{\infty }(U)\ni g(z):={\frac {1}{2\pi i}}\int _{B_{\varepsilon }(0)}{\frac {f(\xi )}{\xi -z}}d\xi \wedge d{\bar {\xi }}}
satisfies
α
=
∂
¯
g
{\displaystyle \alpha ={\bar {\partial }}g}
on
B
ε
(
0
)
.
{\displaystyle B_{\varepsilon }(0).}
Proof. Our claim is that
g
{\displaystyle g}
defined above is a well-defined smooth function and
α
=
f
d
z
¯
=
∂
¯
g
{\displaystyle \alpha =f\,d{\bar {z}}={\bar {\partial }}g}
. To show this we choose a point
z
∈
B
ε
(
0
)
{\displaystyle z\in B_{\varepsilon }(0)}
and an open neighbourhood
z
∈
V
⊆
B
ε
(
0
)
{\displaystyle z\in V\subseteq B_{\varepsilon }(0)}
, then we can find a smooth function
ρ
:
B
ε
(
0
)
→
R
{\displaystyle \rho :B_{\varepsilon }(0)\to \mathbb {R} }
whose support is compact and lies in
B
ε
(
0
)
{\displaystyle B_{\varepsilon }(0)}
and
ρ
|
V
≡
1.
{\displaystyle \rho |_{V}\equiv 1.}
Then we can write
f
=
f
1
+
f
2
:=
ρ
f
+
(
1
−
ρ
)
f
{\displaystyle f=f_{1}+f_{2}:=\rho f+(1-\rho )f}
and define
g
i
:=
1
2
π
i
∫
B
ε
(
0
)
f
i
(
ξ
)
ξ
−
z
d
ξ
∧
d
ξ
¯
.
{\displaystyle g_{i}:={\frac {1}{2\pi i}}\int _{B_{\varepsilon }(0)}{\frac {f_{i}(\xi )}{\xi -z}}d\xi \wedge d{\bar {\xi }}.}
Since
f
2
≡
0
{\displaystyle f_{2}\equiv 0}
in
V
{\displaystyle V}
then
g
2
{\displaystyle g_{2}}
is clearly well-defined and smooth; we note that
g
1
=
1
2
π
i
∫
B
ε
(
0
)
f
1
(
ξ
)
ξ
−
z
d
ξ
∧
d
ξ
¯
=
1
2
π
i
∫
C
f
1
(
ξ
)
ξ
−
z
d
ξ
∧
d
ξ
¯
=
π
−
1
∫
0
∞
∫
0
2
π
f
1
(
z
+
r
e
i
θ
)
e
−
i
θ
d
θ
d
r
,
{\displaystyle {\begin{aligned}g_{1}&={\frac {1}{2\pi i}}\int _{B_{\varepsilon }(0)}{\frac {f_{1}(\xi )}{\xi -z}}d\xi \wedge d{\bar {\xi }}\\&={\frac {1}{2\pi i}}\int _{\mathbb {C} }{\frac {f_{1}(\xi )}{\xi -z}}d\xi \wedge d{\bar {\xi }}\\&=\pi ^{-1}\int _{0}^{\infty }\int _{0}^{2\pi }f_{1}(z+re^{i\theta })e^{-i\theta }d\theta dr,\end{aligned}}}
which is indeed well-defined and smooth, therefore the same is true for
g
{\displaystyle g}
. Now we show that
∂
¯
g
=
α
{\displaystyle {\bar {\partial }}g=\alpha }
on
B
ε
(
0
)
{\displaystyle B_{\varepsilon }(0)}
.
∂
g
2
∂
z
¯
=
1
2
π
i
∫
B
ε
(
0
)
f
2
(
ξ
)
∂
∂
z
¯
(
1
ξ
−
z
)
d
ξ
∧
d
ξ
¯
=
0
{\displaystyle {\frac {\partial g_{2}}{\partial {\bar {z}}}}={\frac {1}{2\pi i}}\int _{B_{\varepsilon }(0)}f_{2}(\xi ){\frac {\partial }{\partial {\bar {z}}}}{\Big (}{\frac {1}{\xi -z}}{\Big )}d\xi \wedge d{\bar {\xi }}=0}
since
(
ξ
−
z
)
−
1
{\displaystyle (\xi -z)^{-1}}
is holomorphic in
B
ε
(
0
)
∖
V
{\displaystyle B_{\varepsilon }(0)\setminus V}
.
∂
g
1
∂
z
¯
=
π
−
1
∫
C
∂
f
1
(
z
+
r
e
i
θ
)
∂
z
¯
e
−
i
θ
d
θ
∧
d
r
=
π
−
1
∫
C
(
∂
f
1
∂
z
¯
)
(
z
+
r
e
i
θ
)
e
−
i
θ
d
θ
∧
d
r
=
1
2
π
i
∬
B
ε
(
0
)
∂
f
1
∂
ξ
¯
d
ξ
∧
d
ξ
¯
ξ
−
z
{\displaystyle {\begin{aligned}{\frac {\partial g_{1}}{\partial {\bar {z}}}}=&\pi ^{-1}\int _{\mathbb {C} }{\frac {\partial f_{1}(z+re^{i\theta })}{\partial {\bar {z}}}}e^{-i\theta }d\theta \wedge dr\\=&\pi ^{-1}\int _{\mathbb {C} }{\Big (}{\frac {\partial f_{1}}{\partial {\bar {z}}}}{\Big )}(z+re^{i\theta })e^{-i\theta }d\theta \wedge dr\\=&{\frac {1}{2\pi i}}\iint _{B_{\varepsilon }(0)}{\frac {\partial f_{1}}{\partial {\bar {\xi }}}}{\frac {d\xi \wedge d{\bar {\xi }}}{\xi -z}}\end{aligned}}}
applying the generalised Cauchy formula to
f
1
{\displaystyle f_{1}}
we find
f
1
(
z
)
=
1
2
π
i
∫
∂
B
ε
(
0
)
f
1
(
ξ
)
ξ
−
z
d
ξ
+
1
2
π
i
∬
B
ε
(
0
)
∂
f
1
∂
ξ
¯
d
ξ
∧
d
ξ
¯
ξ
−
z
=
1
2
π
i
∬
B
ε
(
0
)
∂
f
1
∂
ξ
¯
d
ξ
∧
d
ξ
¯
ξ
−
z
{\displaystyle f_{1}(z)={\frac {1}{2\pi i}}\int _{\partial B_{\varepsilon }(0)}{\frac {f_{1}(\xi )}{\xi -z}}d\xi +{\frac {1}{2\pi i}}\iint _{B_{\varepsilon }(0)}{\frac {\partial f_{1}}{\partial {\bar {\xi }}}}{\frac {d\xi \wedge d{\bar {\xi }}}{\xi -z}}={\frac {1}{2\pi i}}\iint _{B_{\varepsilon }(0)}{\frac {\partial f_{1}}{\partial {\bar {\xi }}}}{\frac {d\xi \wedge d{\bar {\xi }}}{\xi -z}}}
since
f
1
|
∂
B
ε
(
0
)
=
0
{\displaystyle f_{1}|_{\partial B_{\varepsilon }(0)}=0}
, but then
f
=
f
1
=
∂
g
1
∂
z
¯
=
∂
g
∂
z
¯
{\displaystyle f=f_{1}={\frac {\partial g_{1}}{\partial {\bar {z}}}}={\frac {\partial g}{\partial {\bar {z}}}}}
on
V
{\displaystyle V}
. Since
z
{\displaystyle z}
was arbitrary, the lemma is now proved.
Proof of Dolbeault–Grothendieck lemma[ edit ]
Now are ready to prove the Dolbeault–Grothendieck lemma; the proof presented here is due to Grothendieck .[ 1] [ 2] We denote with
Δ
ε
n
(
0
)
{\displaystyle \Delta _{\varepsilon }^{n}(0)}
the open polydisc centered in
0
∈
C
n
{\displaystyle 0\in \mathbb {C} ^{n}}
with radius
ε
∈
R
>
0
{\displaystyle \varepsilon \in \mathbb {R} _{>0}}
.
Lemma (Dolbeault–Grothendieck): Let
α
∈
A
C
n
p
,
q
(
U
)
{\displaystyle \alpha \in {\mathcal {A}}_{\mathbb {C} ^{n}}^{p,q}(U)}
where
Δ
ε
n
(
0
)
¯
⊆
U
{\displaystyle {\overline {\Delta _{\varepsilon }^{n}(0)}}\subseteq U}
open and
q
>
0
{\displaystyle q>0}
such that
∂
¯
α
=
0
{\displaystyle {\bar {\partial }}\alpha =0}
, then there exists
β
∈
A
C
n
p
,
q
−
1
(
U
)
{\displaystyle \beta \in {\mathcal {A}}_{\mathbb {C} ^{n}}^{p,q-1}(U)}
which satisfies:
α
=
∂
¯
β
{\displaystyle \alpha ={\bar {\partial }}\beta }
on
Δ
ε
n
(
0
)
.
{\displaystyle \Delta _{\varepsilon }^{n}(0).}
Before starting the proof we note that any
(
p
,
q
)
{\displaystyle (p,q)}
-form can be written as
α
=
∑
I
J
α
I
J
d
z
I
∧
d
z
¯
J
=
∑
J
(
∑
I
α
I
J
d
z
I
)
J
∧
d
z
¯
J
{\displaystyle \alpha =\sum _{IJ}\alpha _{IJ}dz_{I}\wedge d{\bar {z}}_{J}=\sum _{J}\left(\sum _{I}\alpha _{IJ}dz_{I}\right)_{J}\wedge d{\bar {z}}_{J}}
for multi-indices
I
,
J
,
|
I
|
=
p
,
|
J
|
=
q
{\displaystyle I,J,|I|=p,|J|=q}
, therefore we can reduce the proof to the case
α
∈
A
C
n
0
,
q
(
U
)
{\displaystyle \alpha \in {\mathcal {A}}_{\mathbb {C} ^{n}}^{0,q}(U)}
.
Proof. Let
k
>
0
{\displaystyle k>0}
be the smallest index such that
α
∈
(
d
z
¯
1
,
…
,
d
z
¯
k
)
{\displaystyle \alpha \in (d{\bar {z}}_{1},\dots ,d{\bar {z}}_{k})}
in the sheaf of
C
∞
{\displaystyle {\mathcal {C}}^{\infty }}
-modules, we proceed by induction on
k
{\displaystyle k}
. For
k
=
0
{\displaystyle k=0}
we have
α
≡
0
{\displaystyle \alpha \equiv 0}
since
q
>
0
{\displaystyle q>0}
; next we suppose that if
α
∈
(
d
z
¯
1
,
…
,
d
z
¯
k
)
{\displaystyle \alpha \in (d{\bar {z}}_{1},\dots ,d{\bar {z}}_{k})}
then there exists
β
∈
A
C
n
0
,
q
−
1
(
U
)
{\displaystyle \beta \in {\mathcal {A}}_{\mathbb {C} ^{n}}^{0,q-1}(U)}
such that
α
=
∂
¯
β
{\displaystyle \alpha ={\bar {\partial }}\beta }
on
Δ
ε
n
(
0
)
{\displaystyle \Delta _{\varepsilon }^{n}(0)}
. Then suppose
ω
∈
(
d
z
¯
1
,
…
,
d
z
¯
k
+
1
)
{\displaystyle \omega \in (d{\bar {z}}_{1},\dots ,d{\bar {z}}_{k+1})}
and observe that we can write
ω
=
d
z
¯
k
+
1
∧
ψ
+
μ
,
ψ
,
μ
∈
(
d
z
¯
1
,
…
,
d
z
¯
k
)
.
{\displaystyle \omega =d{\bar {z}}_{k+1}\wedge \psi +\mu ,\qquad \psi ,\mu \in (d{\bar {z}}_{1},\dots ,d{\bar {z}}_{k}).}
Since
ω
{\displaystyle \omega }
is
∂
¯
{\displaystyle {\bar {\partial }}}
-closed it follows that
ψ
,
μ
{\displaystyle \psi ,\mu }
are holomorphic in variables
z
k
+
2
,
…
,
z
n
{\displaystyle z_{k+2},\dots ,z_{n}}
and smooth in the remaining ones on the polydisc
Δ
ε
n
(
0
)
{\displaystyle \Delta _{\varepsilon }^{n}(0)}
. Moreover we can apply the
∂
¯
{\displaystyle {\bar {\partial }}}
-Poincaré lemma to the smooth functions
z
k
+
1
↦
ψ
J
(
z
1
,
…
,
z
k
+
1
,
…
,
z
n
)
{\displaystyle z_{k+1}\mapsto \psi _{J}(z_{1},\dots ,z_{k+1},\dots ,z_{n})}
on the open ball
B
ε
k
+
1
(
0
)
{\displaystyle B_{\varepsilon _{k+1}}(0)}
, hence there exist a family of smooth functions
g
J
{\displaystyle g_{J}}
which satisfy
ψ
J
=
∂
g
J
∂
z
¯
k
+
1
on
B
ε
k
+
1
(
0
)
.
{\displaystyle \psi _{J}={\frac {\partial g_{J}}{\partial {\bar {z}}_{k+1}}}\quad {\text{on}}\quad B_{\varepsilon _{k+1}}(0).}
g
J
{\displaystyle g_{J}}
are also holomorphic in
z
k
+
2
,
…
,
z
n
{\displaystyle z_{k+2},\dots ,z_{n}}
. Define
ψ
~
:=
∑
J
g
J
d
z
¯
J
{\displaystyle {\tilde {\psi }}:=\sum _{J}g_{J}d{\bar {z}}_{J}}
then
ω
−
∂
¯
ψ
~
=
d
z
¯
k
+
1
∧
ψ
+
μ
−
∑
J
∂
g
J
∂
z
¯
k
+
1
d
z
¯
k
+
1
∧
d
z
¯
J
+
∑
j
=
1
k
∑
J
∂
g
J
∂
z
¯
j
d
z
¯
j
∧
d
z
¯
J
∖
{
j
}
=
d
z
¯
k
+
1
∧
ψ
+
μ
−
d
z
¯
k
+
1
∧
ψ
+
∑
j
=
1
k
∑
J
∂
g
J
∂
z
¯
j
d
z
¯
j
∧
d
z
¯
J
∖
{
j
}
=
μ
+
∑
j
=
1
k
∑
J
∂
g
J
∂
z
¯
j
d
z
¯
j
∧
d
z
¯
J
∖
{
j
}
∈
(
d
z
¯
1
,
…
,
d
z
¯
k
)
,
{\displaystyle {\begin{aligned}\omega -{\bar {\partial }}{\tilde {\psi }}&=d{\bar {z}}_{k+1}\wedge \psi +\mu -\sum _{J}{\frac {\partial g_{J}}{\partial {\bar {z}}_{k+1}}}d{\bar {z}}_{k+1}\wedge d{\bar {z}}_{J}+\sum _{j=1}^{k}\sum _{J}{\frac {\partial g_{J}}{\partial {\bar {z}}_{j}}}d{\bar {z}}_{j}\wedge d{\bar {z}}_{J\setminus \lbrace j\rbrace }\\&=d{\bar {z}}_{k+1}\wedge \psi +\mu -d{\bar {z}}_{k+1}\wedge \psi +\sum _{j=1}^{k}\sum _{J}{\frac {\partial g_{J}}{\partial {\bar {z}}_{j}}}d{\bar {z}}_{j}\wedge d{\bar {z}}_{J\setminus \lbrace j\rbrace }\\&=\mu +\sum _{j=1}^{k}\sum _{J}{\frac {\partial g_{J}}{\partial {\bar {z}}_{j}}}d{\bar {z}}_{j}\wedge d{\bar {z}}_{J\setminus \lbrace j\rbrace }\in (d{\bar {z}}_{1},\dots ,d{\bar {z}}_{k}),\end{aligned}}}
therefore we can apply the induction hypothesis to it, there exists
η
∈
A
C
n
0
,
q
−
1
(
U
)
{\displaystyle \eta \in {\mathcal {A}}_{\mathbb {C} ^{n}}^{0,q-1}(U)}
such that
ω
−
∂
¯
ψ
~
=
∂
¯
η
on
Δ
ε
n
(
0
)
{\displaystyle \omega -{\bar {\partial }}{\tilde {\psi }}={\bar {\partial }}\eta \quad {\text{on}}\quad \Delta _{\varepsilon }^{n}(0)}
and
ζ
:=
η
+
ψ
~
{\displaystyle \zeta :=\eta +{\tilde {\psi }}}
ends the induction step. QED
The previous lemma can be generalised by admitting polydiscs with
ε
k
=
+
∞
{\displaystyle \varepsilon _{k}=+\infty }
for some of the components of the polyradius.
Lemma (extended Dolbeault-Grothendieck). If
Δ
ε
n
(
0
)
{\displaystyle \Delta _{\varepsilon }^{n}(0)}
is an open polydisc with
ε
k
∈
R
∪
{
+
∞
}
{\displaystyle \varepsilon _{k}\in \mathbb {R} \cup \lbrace +\infty \rbrace }
and
q
>
0
{\displaystyle q>0}
, then
H
∂
¯
p
,
q
(
Δ
ε
n
(
0
)
)
=
0.
{\displaystyle H_{\bar {\partial }}^{p,q}(\Delta _{\varepsilon }^{n}(0))=0.}
Proof. We consider two cases:
α
∈
A
C
n
p
,
q
+
1
(
U
)
,
q
>
0
{\displaystyle \alpha \in {\mathcal {A}}_{\mathbb {C} ^{n}}^{p,q+1}(U),q>0}
and
α
∈
A
C
n
p
,
1
(
U
)
{\displaystyle \alpha \in {\mathcal {A}}_{\mathbb {C} ^{n}}^{p,1}(U)}
.
Case 1. Let
α
∈
A
C
n
p
,
q
+
1
(
U
)
,
q
>
0
{\displaystyle \alpha \in {\mathcal {A}}_{\mathbb {C} ^{n}}^{p,q+1}(U),q>0}
, and we cover
Δ
ε
n
(
0
)
{\displaystyle \Delta _{\varepsilon }^{n}(0)}
with polydiscs
Δ
i
¯
⊂
Δ
i
+
1
{\displaystyle {\overline {\Delta _{i}}}\subset \Delta _{i+1}}
, then by the Dolbeault–Grothendieck lemma we can find forms
β
i
{\displaystyle \beta _{i}}
of bidegree
(
p
,
q
−
1
)
{\displaystyle (p,q-1)}
on
Δ
i
¯
⊆
U
i
{\displaystyle {\overline {\Delta _{i}}}\subseteq U_{i}}
open such that
α
|
Δ
i
=
∂
¯
β
i
{\displaystyle \alpha |_{\Delta _{i}}={\bar {\partial }}\beta _{i}}
; we want to show that
β
i
+
1
|
Δ
i
=
β
i
.
{\displaystyle \beta _{i+1}|_{\Delta _{i}}=\beta _{i}.}
We proceed by induction on
i
{\displaystyle i}
: the case when
i
=
1
{\displaystyle i=1}
holds by the previous lemma. Let the claim be true for
k
>
1
{\displaystyle k>1}
and take
Δ
k
+
1
{\displaystyle \Delta _{k+1}}
with
Δ
ε
n
(
0
)
=
⋃
i
=
1
k
+
1
Δ
i
and
Δ
k
¯
⊂
Δ
k
+
1
.
{\displaystyle \Delta _{\varepsilon }^{n}(0)=\bigcup _{i=1}^{k+1}\Delta _{i}\quad {\text{and}}\quad {\overline {\Delta _{k}}}\subset \Delta _{k+1}.}
Then we find a
(
p
,
q
−
1
)
{\displaystyle (p,q-1)}
-form
β
k
+
1
′
{\displaystyle \beta '_{k+1}}
defined in an open neighbourhood of
Δ
k
+
1
¯
{\displaystyle {\overline {\Delta _{k+1}}}}
such that
α
|
Δ
k
+
1
=
∂
¯
β
k
+
1
{\displaystyle \alpha |_{\Delta _{k+1}}={\bar {\partial }}\beta _{k+1}}
. Let
U
k
{\displaystyle U_{k}}
be an open neighbourhood of
Δ
k
¯
{\displaystyle {\overline {\Delta _{k}}}}
then
∂
¯
(
β
k
−
β
k
+
1
′
)
=
0
{\displaystyle {\bar {\partial }}(\beta _{k}-\beta '_{k+1})=0}
on
U
k
{\displaystyle U_{k}}
and we can apply again the Dolbeault-Grothendieck lemma to find a
(
p
,
q
−
2
)
{\displaystyle (p,q-2)}
-form
γ
k
{\displaystyle \gamma _{k}}
such that
β
k
−
β
k
+
1
′
=
∂
¯
γ
k
{\displaystyle \beta _{k}-\beta '_{k+1}={\bar {\partial }}\gamma _{k}}
on
Δ
k
{\displaystyle \Delta _{k}}
. Now, let
V
k
{\displaystyle V_{k}}
be an open set with
Δ
k
¯
⊂
V
k
⊊
U
k
{\displaystyle {\overline {\Delta _{k}}}\subset V_{k}\subsetneq U_{k}}
and
ρ
k
:
Δ
ε
n
(
0
)
→
R
{\displaystyle \rho _{k}:\Delta _{\varepsilon }^{n}(0)\to \mathbb {R} }
a smooth function such that:
supp
(
ρ
k
)
⊂
U
k
,
ρ
|
V
k
=
1
,
ρ
k
|
Δ
ε
n
(
0
)
∖
U
k
=
0.
{\displaystyle \operatorname {supp} (\rho _{k})\subset U_{k},\qquad \rho |_{V_{k}}=1,\qquad \rho _{k}|_{\Delta _{\varepsilon }^{n}(0)\setminus U_{k}}=0.}
Then
ρ
k
γ
k
{\displaystyle \rho _{k}\gamma _{k}}
is a well-defined smooth form on
Δ
ε
n
(
0
)
{\displaystyle \Delta _{\varepsilon }^{n}(0)}
which satisfies
β
k
=
β
k
+
1
′
+
∂
¯
(
γ
k
ρ
k
)
on
Δ
k
,
{\displaystyle \beta _{k}=\beta '_{k+1}+{\bar {\partial }}(\gamma _{k}\rho _{k})\quad {\text{on}}\quad \Delta _{k},}
hence the form
β
k
+
1
:=
β
k
+
1
′
+
∂
¯
(
γ
k
ρ
k
)
{\displaystyle \beta _{k+1}:=\beta '_{k+1}+{\bar {\partial }}(\gamma _{k}\rho _{k})}
satisfies
β
k
+
1
|
Δ
k
=
β
k
+
1
′
+
∂
¯
γ
k
=
β
k
∂
¯
β
k
+
1
=
∂
¯
β
k
+
1
′
=
α
|
Δ
k
+
1
{\displaystyle {\begin{aligned}\beta _{k+1}|_{\Delta _{k}}&=\beta '_{k+1}+{\bar {\partial }}\gamma _{k}=\beta _{k}\\{\bar {\partial }}\beta _{k+1}&={\bar {\partial }}\beta '_{k+1}=\alpha |_{\Delta _{k+1}}\end{aligned}}}
Case 2. If instead
α
∈
A
C
n
p
,
1
(
U
)
,
{\displaystyle \alpha \in {\mathcal {A}}_{\mathbb {C} ^{n}}^{p,1}(U),}
we cannot apply the Dolbeault-Grothendieck lemma twice; we take
β
i
{\displaystyle \beta _{i}}
and
Δ
i
{\displaystyle \Delta _{i}}
as before, we want to show that
‖
(
β
i
I
−
β
i
+
1
I
)
|
Δ
k
−
1
‖
∞
<
2
−
i
.
{\displaystyle \left\|\left.\left({\beta _{i}}_{I}-{\beta _{i+1}}_{I}\right)\right|_{\Delta _{k-1}}\right\|_{\infty }<2^{-i}.}
Again, we proceed by induction on
i
{\displaystyle i}
: for
i
=
1
{\displaystyle i=1}
the answer is given by the Dolbeault-Grothendieck lemma. Next we suppose that the claim is true for
k
>
1
{\displaystyle k>1}
. We take
Δ
k
+
1
⊃
Δ
k
¯
{\displaystyle \Delta _{k+1}\supset {\overline {\Delta _{k}}}}
such that
Δ
k
+
1
∪
{
Δ
i
}
i
=
1
k
{\displaystyle \Delta _{k+1}\cup \lbrace \Delta _{i}\rbrace _{i=1}^{k}}
covers
Δ
ε
n
(
0
)
{\displaystyle \Delta _{\varepsilon }^{n}(0)}
, then we can find a
(
p
,
0
)
{\displaystyle (p,0)}
-form
β
k
+
1
′
{\displaystyle \beta '_{k+1}}
such that
α
|
Δ
k
+
1
=
∂
¯
β
k
+
1
′
,
{\displaystyle \alpha |_{\Delta _{k+1}}={\bar {\partial }}\beta '_{k+1},}
which also satisfies
∂
¯
(
β
k
−
β
k
+
1
′
)
=
0
{\displaystyle {\bar {\partial }}(\beta _{k}-\beta '_{k+1})=0}
on
Δ
k
{\displaystyle \Delta _{k}}
, i.e.
β
k
−
β
k
+
1
′
{\displaystyle \beta _{k}-\beta '_{k+1}}
is a holomorphic
(
p
,
0
)
{\displaystyle (p,0)}
-form wherever defined, hence by the Stone–Weierstrass theorem we can write it as
β
k
−
β
k
+
1
′
=
∑
|
I
|
=
p
(
P
I
+
r
I
)
d
z
I
{\displaystyle \beta _{k}-\beta '_{k+1}=\sum _{|I|=p}(P_{I}+r_{I})dz_{I}}
where
P
I
{\displaystyle P_{I}}
are polynomials and
‖
r
I
|
Δ
k
−
1
‖
∞
<
2
−
k
,
{\displaystyle \left\|r_{I}|_{\Delta _{k-1}}\right\|_{\infty }<2^{-k},}
but then the form
β
k
+
1
:=
β
k
+
1
′
+
∑
|
I
|
=
p
P
I
d
z
I
{\displaystyle \beta _{k+1}:=\beta '_{k+1}+\sum _{|I|=p}P_{I}dz_{I}}
satisfies
∂
¯
β
k
+
1
=
∂
¯
β
k
+
1
′
=
α
|
Δ
k
+
1
‖
(
β
k
I
−
β
k
+
1
I
)
|
Δ
k
−
1
‖
∞
=
‖
r
I
‖
∞
<
2
−
k
{\displaystyle {\begin{aligned}{\bar {\partial }}\beta _{k+1}&={\bar {\partial }}\beta '_{k+1}=\alpha |_{\Delta _{k+1}}\\\left\|({\beta _{k}}_{I}-{\beta _{k+1}}_{I})|_{\Delta _{k-1}}\right\|_{\infty }&=\|r_{I}\|_{\infty }<2^{-k}\end{aligned}}}
which completes the induction step; therefore we have built a sequence
{
β
i
}
i
∈
N
{\displaystyle \lbrace \beta _{i}\rbrace _{i\in \mathbb {N} }}
which uniformly converges to some
(
p
,
0
)
{\displaystyle (p,0)}
-form
β
{\displaystyle \beta }
such that
α
|
Δ
ε
n
(
0
)
=
∂
¯
β
{\displaystyle \alpha |_{\Delta _{\varepsilon }^{n}(0)}={\bar {\partial }}\beta }
. QED
Dolbeault's theorem[ edit ]
Dolbeault's theorem is a complex analog[ 3] of de Rham's theorem . It asserts that the Dolbeault cohomology is isomorphic to the sheaf cohomology of the sheaf of holomorphic differential forms. Specifically,
H
p
,
q
(
M
)
≅
H
q
(
M
,
Ω
p
)
{\displaystyle H^{p,q}(M)\cong H^{q}(M,\Omega ^{p})}
where
Ω
p
{\displaystyle \Omega ^{p}}
is the sheaf of holomorphic p forms on M .
A version of the Dolbeault theorem also holds for Dolbeault cohomology with coefficients in a holomorphic vector bundle
E
{\displaystyle E}
. Namely one has an isomorphism
H
p
,
q
(
M
,
E
)
≅
H
q
(
M
,
Ω
p
⊗
E
)
.
{\displaystyle H^{p,q}(M,E)\cong H^{q}(M,\Omega ^{p}\otimes E).}
A version for logarithmic forms has also been established.[ 4]
Let
F
p
,
q
{\displaystyle {\mathcal {F}}^{p,q}}
be the fine sheaf of
C
∞
{\displaystyle C^{\infty }}
forms of type
(
p
,
q
)
{\displaystyle (p,q)}
. Then the
∂
¯
{\displaystyle {\overline {\partial }}}
-Poincaré lemma says that the sequence
Ω
p
,
q
→
∂
¯
F
p
,
q
+
1
→
∂
¯
F
p
,
q
+
2
→
∂
¯
⋯
{\displaystyle \Omega ^{p,q}{\xrightarrow {\overline {\partial }}}{\mathcal {F}}^{p,q+1}{\xrightarrow {\overline {\partial }}}{\mathcal {F}}^{p,q+2}{\xrightarrow {\overline {\partial }}}\cdots }
is exact. Like any long exact sequence , this sequence breaks up into short exact sequences. The long exact sequences of cohomology corresponding to these give the result, once one uses that the higher cohomologies of a fine sheaf vanish.
Explicit example of calculation [ edit ]
The Dolbeault cohomology of the
n
{\displaystyle n}
-dimensional complex projective space is
H
∂
¯
p
,
q
(
P
C
n
)
=
{
C
p
=
q
0
otherwise
{\displaystyle H_{\bar {\partial }}^{p,q}(P_{\mathbb {C} }^{n})={\begin{cases}\mathbb {C} &p=q\\0&{\text{otherwise}}\end{cases}}}
We apply the following well-known fact from Hodge theory :
H
d
R
k
(
P
C
n
,
C
)
=
⨁
p
+
q
=
k
H
∂
¯
p
,
q
(
P
C
n
)
{\displaystyle H_{\rm {dR}}^{k}\left(P_{\mathbb {C} }^{n},\mathbb {C} \right)=\bigoplus _{p+q=k}H_{\bar {\partial }}^{p,q}(P_{\mathbb {C} }^{n})}
because
P
C
n
{\displaystyle P_{\mathbb {C} }^{n}}
is a compact Kähler complex manifold . Then
b
2
k
+
1
=
0
{\displaystyle b_{2k+1}=0}
and
b
2
k
=
h
k
,
k
+
∑
p
+
q
=
2
k
,
p
≠
q
h
p
,
q
=
1.
{\displaystyle b_{2k}=h^{k,k}+\sum _{p+q=2k,p\neq q}h^{p,q}=1.}
Furthermore we know that
P
C
n
{\displaystyle P_{\mathbb {C} }^{n}}
is Kähler, and
0
≠
[
ω
k
]
∈
H
∂
¯
k
,
k
(
P
C
n
)
,
{\displaystyle 0\neq [\omega ^{k}]\in H_{\bar {\partial }}^{k,k}(P_{\mathbb {C} }^{n}),}
where
ω
{\displaystyle \omega }
is the fundamental form associated to the Fubini–Study metric (which is indeed Kähler), therefore
h
k
,
k
=
1
{\displaystyle h^{k,k}=1}
and
h
p
,
q
=
0
{\displaystyle h^{p,q}=0}
whenever
p
≠
q
,
{\displaystyle p\neq q,}
which yields the result.
^ Serre, Jean-Pierre (1953–1954), "Faisceaux analytiques sur l'espace projectif" , Séminaire Henri Cartan , 6 (Talk no. 18): 1– 10
^ "Calculus on Complex Manifolds". Several Complex Variables and Complex Manifolds II . 1982. pp. 1– 64. doi :10.1017/CBO9780511629327.002 . ISBN 9780521288880 .
^ In contrast to de Rham cohomology, Dolbeault cohomology is no longer a topological invariant because it depends closely on complex structure.
^ Navarro Aznar, Vicente (1987), "Sur la théorie de Hodge–Deligne", Inventiones Mathematicae , 90 (1): 11– 76, Bibcode :1987InMat..90...11A , doi :10.1007/bf01389031 , S2CID 122772976 , Section 8