Blowing up Kähler manifolds with constant scalar curvature II
Claudio Arezzo
Frank Pacard
claudio.arezzo@unipr.it, Dipartimento di Matematica, Universita' di Parma, Via M. D'Azeglio, 85, 43100 Parma, Italy pacard@univ-paris12.fr, University Paris 12 and Institut Universitaire de France, France
-
Abstract.
In this paper we prove the existence of Kähler metrics of constant scalar curvature on the blow up at finitely many points of a compact manifold which already carries a Kähler constant scalar curvature metric. Necessary conditions of the number and locations of the blow up points are given.
1991 Math. Subject Classification: 58E11, 32C17.
1 Introduction
Assume that
is a constant scalar curvature compact Kähler manifold. Given
distinct points
, we define
to be the blow up of
at the points
. The question we would like to address in the paper is whether
can be endowed with a constant scalar curvature Kähler form. In [2] , we have already given a positive answer to this question in the case where the manifold has no nontrivial vanishing holomorphic vector field (i.e. has no non zero holomorphic vector field which vanish somewhere). This condition is for example fulfilled when the group of automorphisms of
is discrete. Under this condition, we have obtained the following :
Theorem 1.1.
[
2]
Assume that
is a constant scalar curvature compact Kähler manifold without nontrivial vanishing holomorphic vector field. Then,
, the blow up at finitely many points has a constant scalar curvature Kähler form
. In addition, if the scalar curvature of
is not zero then the scalar curvatures of
and of
have the same signs.
In the case of
-scalar curvature metrics, we also have the :
Theorem 1.2.
[
2]
Assume that
is a
-scalar curvature compact Kähler manifold without nontrivial vanishing holomorphic vector field. Then, the blow up of
at finitely many points has a
-scalar curvature Kähler form provided the first Chern class of
is non zero.
This last result compliments, in any dimension, previous constructions which have been obtained in complex dimension
and for
-scalar curvature metrics, by Kim-LeBrunPontecorvo [6] , LeBrun-Singer [13] and Rollin-Singer [15] . Indeed, using twistor theory, Kim-LeBrunPontecorvo and LeBrun-Singer have been able to construct such metrics by desingularizing some quotients of minimal ruled surfaces and, more recently, Rollin-Singer [16] have shown that one can desingularize compact manifolds of zero scalar curvature with cyclic orbifold groups, keeping the scalar curvature zero by solving on the desingularization the hermitian anti-selfdual equation which implies the Kähler scalar flat equation.
Theorem 1.1 and Theorem 1.2 are obtained using a connected sum of the Kähler form
at each
with a
-scalar curvature Kähler form
which is defined on
, the blow up of
at the origin. When
this model
has been known for long time and it is usually referred to in the literature as Burns metric [9] . In higher dimensions a similar model has been produced by Simanca [17] .
In the present paper, we are interested in the case where
has nontrivial vanishing holomorphic vector fields
, with
(this condition in particular implies that
has a nontrivial continuous automorphisms group) since this is precisely the case which is not covered by our previous result. Our main result states that the blow up of
at sufficiently many points can be endowed with a constant scalar curvature Kähler metric provided these points are carefully chosen.
Theorem 1.3.
Assume that
is a constant scalar curvature compact Kähler manifold. Then, there exists
and, for all
there exist a nonempty open subset
such that, for all
there exist constant scalar curvature Kähler forms on the blow up of
.
In fact, we can be more precise and show that there exists
, for
, a one parameter family of Kähler forms on
which converges in
norm as
tends to
, to the Kähler form
away from the points
. Moreover, the sequence of Kähler forms
converges in
norm to the Kähler form
on
the blow up at the origin of
. Finally, if
had positive (resp. negative) scalar curvature then the Kähler forms
have positive (resp. negative) scalar curvature. Let us emphasize that, in contrast with the case where the manifold
does not carry any nontrivial vanishing holomorphic vector field, this time the number and position of the blow up points is not arbitrary.
Remark 1.1.
We will see that, in some explicit cases, one can make use of the symmetries of the manifold
, and construct constant scalar curvature Kähler forms on the blow up of
at
with
. However this construction will not hold anymore for the choice of the points in some open set of
.
Let us mention that, for the
-dimensional complex manifolds, Kähler metrics with constant scalar curvature have been obtained by Rollin-Singer in [15] , [16] on blow ups of
,
or
using a different approach based on both algebraic tools and a connect sum result, while for zero scalar curvature metrics on blow ups of
,
, the problem was solved by LeBrun-Singer [13] using twistor theory. We will recall these results in §
.
In the statement of Theorem 1.3, the value of
and the choice of the blow up points are very delicate issues. To describe this, we still assume that
is a constant scalar curvature compact Kähler manifold and we define (see §
) the elliptic fourth order operator
where all operators are computed with respect to the metric induced by
. We assume that the kernel of
is not trivial and we denote by
the linearly independent functions which span this kernel. We can assume, without loss of generality that the functions
have mean
. Observe that
either denotes the dimension of the space of vanishing holomorphic vector fields or the dimension of the nontrivial kernel of
since, by a result of LeBrun-Simanca [12] these two dimensions are the same. This being understood, we define the matrix
|
(1)
|
as well as the integer valued functions
where
is the cone of vectors with positive entries in
.
Our main result is a consequence of the following sequence of results. The first one gives a sufficient condition on the functions
and
for Theorem 1.3 to hold :
Proposition 1.1.
Assume that
is a constant scalar curvature compact Kähler manifold. Let us assume that
and
are chosen so that
and that
. Then there exists
and for all
, a constant scalar curvature Kähler form
on
.
Furthermore, the Kähler form
satisfy the properties described after the statement of Theorem 1.3.
As will be explained in §
, the first condition
is easily seen to be generic (and open) in the sense that :
Lemma 1.1.
With the above notations, assume that
then the set of the points
such that
is open and dense in
.
A result of LeBrun-Simanca [12] (see also [4] and [13] ) implies that nontrivial vanishing holomorphic vector fields are in one to one correspondence with nonconstant elementsof the kernel of
. Furthermore, it is well known that, for a choice of blow up points
in some open and dense subset of
, the group of automorphisms of
is trivial (observe that, by virtue of the above mentioned result of LeBrun and Simanca
is also equal to the dimension of the identity component of the automorphisms group of
). In view of all these results, one is tempted to conjecture that our condition
is equivalent to the fact that the group of automorphisms of
is trivial.
However, these two conditions turn out to be of a different nature. Indeed, for example when
the kernel of
is spanned by
and
, and the set of nontrivial vanishing holomorphic vector fields is spanned by
. Then it is enough to choose
not in the zero set of
to get
, while the group of automorphisms of
is trivial if and only if
is chosen away from the
set of
, but, according to LeBrun and Simanca, the
set of
corresponds to the set of critical points of
.
The second sufficient condition on the existence of points for which
is more subtle. We prove, in §
, that this condition is always fulfilled for some careful choice of the points, provided their number
is chosen larger than some value
.
Lemma 1.2.
With the above notations, assume that
, then the set of points
for which
and
is open. Moreover, there exists
such that, for all
the set of points
for which
and
is nonempty.
The proof of the Lemma 1.2 is due to E. Sandier, and we are very grateful to him for allowing us to present it here. By opposition to the first condition, it is easy to convince oneself that this condition does not hold for generic choice of the points.
Remark 1.2.
We believe that
, the minimal number of points for which the second condition is fulfilled, should be equal to
. We have only been able to prove this fact for some special manifolds where all elements of the kernel of
are explicitly known.
It is interesting to observe that our sufficient conditions depends only on the complex structure of the base manifold and not on riemannian properties of its Kähler metric.
This is shared with all known necessary conditions, from Matsushima-Lichnerovicz's one, Futaki's character and the known stability conditions implied by the existence of canonical metrics thanks to the work of Donaldson, Mabuchi and Paul-Tian. We now give a number of explicit examples for which our result can be applied.
As already mentioned in Remark 1.1, by definition
is larger than the dimension of vanishing holomorphic vector fields on
and, using some symmetries of the manifolds, one can still construct constant scalar curvature Kähler forms on the blow up of
at fewer points. To illustrate this fact, we have, when
, we have :
Lemma 1.3.
When
, then
and
.
Applying Theorem 1.3, we get the existence of constant scalar curvature Kähler forms on the blow up of
at
points for
which belong to some nonempty open set of
. One can also work equivariantly using the symmetries of
to get the :
Corollary 1.1.
For any
there exist points
such that the blow up of
at
carries a constant scalar curvature Kähler metric.
The above corollary is optimal in the number of points because, for
,
does not satisfy the Matsushima-Lichnerowicz obstruction. Observe also that
still has vanishing holomorphic vector fields.
On the other hand it is well known that on
,
points forming a projective frame are enough to kill all holomorphic vector fields when blowing up, and indeed wecan prove that this condition also guarantees the existence of Kähler constant scalar curvature metric.
Corollary 1.2.
Given
points in
such that
form a projective frame,
has a Kähler constant scalar curvature metric and no holomorphic vector fields. Moreover
can vary in a dense open subset of
and
are arbitrary.
When
where
is any manifold with no nontrivial vanishing holomorphic vector field, all the above result carry to
. In particular :
Corollary 1.3.
Assume that
is a Kähler manifold with constant scalar curvature and without nontrivial vanishing holomorphic vector fields and define
, then
and
. Moreover, the blow up of
at
points has a constant scalar curvature Kähler metric.
As already mentioned, when
and
is a Riemann surface of genus
greater than
, the existence of zero scalar curvature Kähler metrics on blow ups of
is due to LeBrun and Singer [12] .
Finally, we have also considered the case where
, for which we have obtained the :
Corollary 1.4.
Suppose
. The blow up of
at
special points has a constant scalar curvature Kähler metric. Moreover, if the projections of these points to the two factors contain projective frames, these points can move in a dense open subset of
and the other ones are arbitrary.
The above results for
translate directly in the following :
Corollary 1.5.
For any
there exist points
such that
has a constant scalar curvature Kähler metric.
Rollin-Singer in [16] have already found such constant scalar curvature Kähler metrics on the blow up of
at
points. It is worth mentioning that, for
,
still has nontrivial vanishing holomorphic vector fields. When
the construction is performed using the symmetries of the manifold, working equivariantly.
2 Weighted spaces
In this section, we describe weighted spaces on the noncompact manifold
with
points removed, as well as weighted spaces on the noncompact manifolds
. We also introduce notations which will be used in the next sections.
For all
, we agree that
|
(2)
|
Assume that
and
are fixed. Given
and a function
, we define
and, for any function
, we define
Assume that we are given
, a Kähler manifold with Kähler form
and that we are also given
distinct points
. By definition, near
, the manifold
is biholomorphic to a neighborhood of
in
. In particular, we can choose complex coordinates
in a neighborhood of
in
, to parameterize a neighborhood of
in
. In order to distinguish between the different neighborhoods and coordinate systems, we agree that, for all
small enough, say
,
(resp.
) denotes the ball (resp. the punctured ball) of radius
in the above defined coordinates
parameterizing a fixed neighborhood
of
. Furthermore, it follows from [5] , that there is no loss of generality in assuming that the Kähler form
can be expended as
|
(3)
|
where the function
. This in particular implies that, in these coordinates, the standard metric on
and the metric induced by
agree up to order
.
For all
, we define
|
(4)
|
The weighted space for functions defined on the noncompact manifold
|
(5)
|
is then defined as the set of functions whose decay or blow up near any
is controlled by a power of the distance to
. More precisely, we have the :
Definition 2.1.
Given
,
and
, we define the weighted space
to be the space of functions
for which the following norm is finite
We now turn to the description of weighted space on
, the blow up at the origin of
endowed with the Burns-Simanca metric. Away from the exceptional divisor, the Kähler form
is given by
where
are complex coordinates in
and where the function
is a solution of the ordinary differential equation
which satisfies
near
. It turns out that, when
, the function
is explicitly given by
where
and
, while in dimension
, even though there is no explicit formula for
we have the following :
Lemma 2.1.
Assume that
. Then the function
can be expanded as
for
, where
and
.
Proof : This result follows from the analysis done in [17] even though this is not explicitly stated. In fact, we define the function
by
. Direct computation shows that
solves
If in addition we take
, then
remains positive and one can check that
is well defined for all time and converges to some positive constant
, as
tends to
.
This immediately implies that
at infinity. The expansion then follows easily.
Changing variables
, we see from the previous Lemma that the Kähler form
can be expanded near
as
|
(6)
|
in dimension
and as
|
(7)
|
in dimension
, where
for all
. The property which will be crucial for our construction is that, in these two expansions, the coefficient in from of
, in dimension
and the coefficients in front of
are not zero.
For all
, we define
We are now in a position to define weighted spaces on the noncompact complete manifold
. This time, we are interested in functions which decay (or blow up) near infinity at a rate which is controlled by a power of the distance to a fixed point in
.
More precisely, we have the :
Definition 2.2.
Given
,
and
, we define the weighted space
to be the space of functions
for which the following norm is finite
3 The geometry of the equation
Recall that
is a compact Kähler manifold of complex dimension
. We will indicate by
the Riemannian metric induced by
,
its Ricci tensor,
the corresponding Ricci form, and
its scalar curvature.
Following LeBrun-Simanca [12] , we want to understand the behavior of the scalar curvature of the metric under deformations of the form
where
a function defined on
. The following result is proven, for example in [12] :
Proposition 3.1.
The scalar curvature of
can be expanded in terms of
as
where
collects all the nonlinear terms and where all operators in the right hand side of this identity are computed with respect to the Kähler form
.
The operator
|
(9)
|
which acts on functions is the one which has been defined in the introduction.
For a general Kähler metric it can be very difficult to analyze these operators. Nevertheless geometry comes to the rescue at a constant scalar curvature metric. Indeed, in this case we have
|
(10)
|
where
denotes the
-part of the
-gradient of
. Using this result, the key observation of LeBrun-Simanca is that to any element
of
one can associate a holomorphic vector field, namely
, which must vanish somewhere on
. This is the crucial relation between constant scalar curvature (or extremal) metrics and the space of holomorphic vector fields
(and, in turn, the automorphism group
).
We recall the following important result, which has been mentioned in the introduction and which helps to understand the kernel of the operator
:
Theorem 3.1.
[
12]
Assume that
is a compact Kähler manifold of complex dimension
, then the space of the elements of the kernel of
whose mean over
is
, is in one to one correspondence with the space of holomorphic vector fields which vanish on
.
The previous considerations extend to
and this implies the following important result of [8] which states that there are no elements in the kernel of the operator
which decay at infinity.
Proposition 3.2.
[
8]
There are no nontrivial solution to
, which belong to
, for some
.
Proof : Assume that
and that
, for some
. Then
is an holomorphic vector field which tends to
at infinity. Using Hartogs theorem, the restriction of
to
can be extended to a holomorphic vector field on
. Since this vector field decays at infinity, it has to be identically equal to
. This implies that
is identically equal to
on
. However
being a real valued function, this implies that
in
. Hence the function
is constant in
and decays at infinity. This implies that
is identically equal to
in
and satisfies
in
. Now, use the unique continuation theorem to conclude that
is identically equal to
in
.
4 Mapping properties
4.1 Analysis of the operator defined on
The results we want to obtain are based on the fact that, near each
and in the above defined coordinates, the metric on
is asymptotic to the Euclidean metric. This implies that, in each
the operator
is close to the operator
|
(11)
|
where
denotes the Laplacian in
when endowed with the standard Kähler form.
In particular these two operators have the same indicial roots at
. This set of indicial roots is given by
when
and is given by
when
. The mapping properties of
when defined between weighed spaces follows from the general theory explained in [14] and this will allow us to construct some right inverse for
.
We define in
the function
by
Observe that, unless the metric
is the Euclidean metric, these functions are not solutions of the homogeneous equation associated to
, however they can be perturbed into
solutions of the homogeneous problem
. Indeed, reducing
if this is necessary, we have the following :
Lemma 4.1.
There exist functions
which are solutions of
in
and which are asymptotic to
in the sense that
when
and
for any
when
.
Proof : Observe that
and, thanks to the expansion given in ( 3 ), we conclude that
. The result then follows from the analysis of §6 in [2] . Observe that, as long as
is not an indicial root of
we conclude that
while, if
is an indicial root of
, we only conclude that
for any
.
Observe that there is no uniqueness of
. With the functions
at hand, we define the following deficiency spaces
where
is a cutoff function which is identically equal to
in
and identically equal to
in
.
When
, we fix
and define the operator
| |
Whereas, when
, we fix
and define the operator
| |
To keep notations short, it will be convenient to set
when
and
when
. The main result of this section reads :
Proposition 4.1.
Assume that the points
are chosen so that
. Then, the operator
defined above is surjective (and has a
-dimensional kernel).
Proof : The proof of this result follows from the general theorem described in [14] , however, we choose here to describe an almost self contained proof. Recall that the kernel of
is spanned by the functions
and, by assumption
, for
have mean
. We use the fact that
is self adjoint and hence, for
, the problem
is solvable if and only if
satisfies
for
.
Observe that
when
. Now, given
, we choose
and, since
, we also choose
solution of the system
Then, the problem
is solvable in
for all
and uniqueness follows if we impose that
is orthogonal to
. To complete the proof, we invoke regularity theory which implies that
when
and
when
. The estimate of the dimension of the kernel is left to the reader since it will not be used in the paper.
Observe that, when we will solve
, the constant
is determined by
4.2 Analysis of the operator defined on
As above, we use the fact that the metric on
is asymptotic to Euclidean metrics, as explained in ( 6 ) and ( 7 ). This implies that, in
, the operator
is close to the operator
and they have the same set of indicial roots which is given by
when
and is given by
when
. Given
, we define the operator
| |
and recall the following result from [2]
Proposition 4.2.
Assume that
. Then the operator
defined above is surjective and has a one dimensional kernel spanned by the constant function.
Proof : The result of Proposition 3.2 precisely states that the operator
is injective when
. This implies that the operator
is surjective when
. When
, this also implies that the operator
has a one dimensional kernel, spanned by the constant function.
4.3 Bi-harmonic extensions
We end up this section by the following simple result which follows at once from the application of the maximum principle and whose proof can be found in [2] .
Proposition 4.3.
Assume that
. Given
,
there exists a function
such that
with
Moreover,
We will also need the following result which differs slightly from what we have already used in [2] . The rational being that there exists a bi-harmonic extension of the data
which is defined on the complement of the unit ball and decays at infinity. Moreover, this function is bounded by a constant times the distance to the origin to the power
(when
). In the case where the function
is assumed to have mean
, then the rate of decay can be improved and estimated as the distance to the origin to the power
:
Proposition 4.4.
Assume that
. Given
,
such that
there exists a function
such that
with
Moreover,
4.4 Perturbation of
We consider the Kähler form
|
(12)
|
The scalar curvature of
is given by
where the operator
has been defined in ( 9 ) and where
collects all the nonlinear terms. The structure of
is quite complicated however, it follows from the explicit computation of the Ricci curvature that, near
, the nonlinear operator
can be decomposed as
|
(13)
|
where the
and
are bilinear forms with bounded coefficients and the sums are finite. In addition, identifying the Hessian of
with a symmetric matrix, one can check that the nonlinear operators
and
are uniformly bounded and uniformly Lipschitz over the set of symmetric matrices whose entries are bounded by some fixed constant
. More precisely, they satisfy
|
(14)
|
and
|
(15)
|
for all symmetric matrices
and
satisfying
and
.
As usual, we denote by
, independent functions which span the kernel of
. We agree that
is the constant function and that all functions
are normalized to have mean
. Now, assume that we are given
such that there exists a solution of
where the constant
is defined by
Observe that such a function
exists if and only if
is given by
and the coefficients
are solutions of the system
for
. This amounts to ask that the vector
has positive entries and is in the kernel of the matrix
. This is precisely the origin of the condition
. It is not hard to check that :
Lemma 4.2.
Near each
, the function
satisfies
for some constant
.
We fix
|
(16)
|
Assume that we are given
and
,
, satisfying
|
(17)
|
where
will be fixed later on. Further assume that
We define
|
(19)
|
where
and
stand for
and
and where the cutoff functions
are identically equal to
in
and identically equal to
in
.
We fix
. The space
(resp.
) is just defined as the space of restrictions of functions in
(resp.
) to
, endowed with the induced norm.
We would like to solve the equation
where
and
have to be determined. This amounts to solve the equation
|
(20)
|
in
. We consider an extension (linear) operator
such that
in
,
is supported in
and
.
The equation we would like to solve can be rewritten as
|
(21)
|
where
is determined by
Granted these notations, we prove the :
Lemma 4.3.
We fix
. There exists
,
and there exists
such that, for all
and
where the norms are the restrictions of the weighted norms in Definition 2.1 to functions defined in
. In addition, we have and
Proof : The proof of these estimates can be obtained as in [2] . We briefly recall it here.
First, we use the result of Proposition 4.4 to estimate
|
(22)
|
Now observe that
in
and
in each
, hence
in this set. Now, we use the expansion ( 3 ) which reflects the fact that the metric induced by
is in each
asymptotic to the Euclidean metric. This implies that there exists a constant
such that
Collecting these, and taking into account the effect of the cutoff functions, we conclude that
The first estimate follows at once from the fact that
.
The second estimate follows from the structure of the nonlinear operator
described in ( 13 ) together with the estimate
and also ( 22 ). Roughly speaking this estimate reflects the fact the most important nonlinear terms in the expression of
are of the form
and
provided the second derivatives of
remain bounded (which is precisely our case). The proof of the second estimate follows easily from these considerations. The third estimate is easy to obtain and left to the reader.
We will also need the following result which will be useful for the contraction mapping argument we will apply :
Lemma 4.4.
We fix
and
. There exists a constant
such that, for all
satisfying
we have
| |
| |
and we also have
| |
| |
The important observation is that
Using the results of §5.2, we reduce the solvability of the equation ( 21 ) to a fixed point problem in
where the constant
is fixed large enough. And in fact, using Lemma 4.3 and Lemma 4.4 it is easy to check that there exists a fixed point solution of ( 21 ) provided
is chosen small enough, say
by applying a fixed point theorem for contraction mapping.
Hence we have obtained the :
Proposition 4.5.
Assume that
and choose
. Given
, there exists
and
such that, for all
, for all
and all
satisfying ( 17 ), there exists a Kähler form
defined on
, which has constant scalar curvature equal. Moreover,
| |
| |
and
Recall that
is the function which appears in ( 3 ), the expansion of the Kähler potential
at
so that
in each
. We define,
when
and
when
. Observe that, locally, the Kähler potential is not uniquely defined and addition of constant does not modify the corresponding Kähler form.
Thanks to the previous analysis, the following expansions are valid in each
.
Lemma 4.5.
The following expansions holds
|
(23)
|
when
and
|
(24)
|
when
.
Similar expansions are valid for the partial derivatives of these functions. The subscript
in
is meant to recall that, as
tends to
, this quantity is bounded by a constant, depending on
, times
, while
is bounded by a constant (independent of
) times
.
The key point is that
as
tends to
(here we have to assume that
, when
) and this implies that, in the expansion of the boundary data of the potential
, the part corresponding to the solution of the nonlinear problem ( 20 ) is much smaller than
, the characteristic size of the boundary data
and
.
It should be clear that the Kähler form provided by Proposition 4.5 depends continuously on the data (such as
and
).
4.5 Perturbation of
Given
, we now consider on
, the Kähler form
which, after a change of variable, can be expanded in
as
|
(25)
|
when
and
|
(26)
|
when
, where the function
enjoys properties similar to the one enjoyed by
. Everything being uniform in
as long as this parameter remains both bounded from above and bounded away from
. In fact we will need to choose
so that
where the coefficients
are the one which have been defined at the beginning of the previous section.
As in the previous section, we consider the perturbed Kähler form
The scalar curvature of
is given by
since the scalar curvature of
is identically equal to
. Even though this is not explicit in the notations, the operators
and
both depend on
. Again, the structure of the nonlinear operator
is quite complicated but, in
, it enjoys a decomposition similar to the ones described in ( 13 ) and the estimates ( 14 ) and ( 15 ) hold uniformly in each
and in
, as long as
remains bounded from above and bounded away from
.
We define
Given
and
satisfying
|
(28)
|
where
will be fixed later on, we define
|
(29)
|
where
are cutoff functions which are identically equal to
in
and identically equal to
in
.
We fix
. The space
(resp.
) is defined as the space of restrictions of functions in
(resp.
) to
, endowed with the induced norm.
We would like to solve the equation
where
satisfies
.
We consider an extension (linear) operator
such that
in
,
is supported in
and
. The equation we would like to solve can be rewritten as
|
(30)
|
We prove the :
Lemma 4.6.
Assume that
. There exists
such that, for all
and
Proof : Again, the proof of these estimates can be obtained as in [2] . First, we use the result of Proposition 4.3 to estimate
|
(31)
|
Now observe that
in each
, hence
in this set.
Now, we use the expansion ( 6 ) or the expansion ( 7 ) which reflect the fact that the metric induced by
is, in each
asymptotic to the Euclidean metric. This implies that there exists a constant
such that
Collecting these, and taking into account the effect of the cutoff functions, we conclude that
Observe that the main contribution comes from the effect of the cutoff function
.
The second estimate is easy to derive. The third estimate follows from the structure of the nonlinear operator
, which is similar to the one described in ( 13 ), together with ( 31 ). Again, the most important nonlinear terms in the expression of
are of the form
and
provided the second derivatives of
remain bounded (which is precisely our case). The proof of the second estimate follows easily from these considerations and, again, the main contribution comes from the effect of the cutoff function
.
We will also need the :
Lemma 4.7.
We fix
and
. There exists
such that, for all
, satisfying
we have
Using the result of §5.3, we reduce the solvability of the equation ( 30 ) to a fixed point problem in
where the constant
is fixed large enough. In fact, using Lemma 4.6 and Lemma 4.7 it is easy to check that there exists a fixed point solution of ( 30 ) provided
is chosensmall enough, say
, which is obtained by applying a fixed point theorem for contraction mapping.
We have obtained the :
Proposition 4.6.
Given
,
and
, there exists
such that, for all
, for all
and
, there exists a Kähler form
defined on
, which has constant scalar curvature. Moreover
for some constant
independent of
. In addition,
.
We now apply this result when
is the scalar curvature of the metric defined in the previous section. The corresponding Kähler form will be denoted by
even though it also depends on
and
. Recall that
is the function which appears in ( 25 ) and ( 26 ), the expansion of the Kähler potential
at
. Hence, the Kähler form
has constant scalar curvature equal to
and, near
, it can be expanded as
when
and
when
.
As in the previous section, we define,
when
and
when
. Observe that, locally, the Kähler potential is not uniquely defined and addition of constant does not modify the corresponding Kähler form.
Thanks to the previous analysis, the following expansions are valid in each
.
Lemma 4.8.
The following expansion holds
|
(32)
|
when
and
|
(33)
|
when
.
Similar expansions are valid for the partial derivatives of these functions.
The key point is that the quantity
is bounded by a constant, independent of
and of
, times
, as
tends to
. Again, it should be clear that the Kähler form provided by Proposition 4.6 depends continuously on the data (such as
and
).
5 Gluing the pieces together
We are now in a position to describe the connected sum construction. For all
small enough, we define a complex manifold
be removing small balls centered at the points
,
and replacing them by properly rescaled versions of the
. We define
|
(34)
|
By construction
is obtained by performing a connected sum of
(which has
boundaries) with the truncated ALE spaces
. This connected sum is obtained by identifying
with
using
where
are coordinates in
and
are coordinates in
.
Next, a Kähler form is defined on
by gluing together the Kähler form
define in §4.4 and the rescalled Kähler forms
defined in §4.5 on each copy of
.
It remains to explain to choose the data
,
satisfying ( 17 ) and ( 18 ),
,
satisfying ( 28 ), and
in such a way that, for each
, the function
defined in
as in §4.4 on the one hand and the function
defined in
as in §4.5 with data
and
on the other hand, have their derivatives up to order
which coincide on
.
Now, assume that the functions
and
are
functions which satisfy
|
(35)
|
on
where
. Consider polar coordinates and decompose
Using the decomposition of
in polar coordinates, it is clear that these identities guaranty that
and
have their partial derivatives up to order
which coincide on
.
Assuming for the moment that we have already found the construction data in such a way that ( 35 ) are satisfied, then the Kähler form
defined in
and the Kähler form
defined in
have constant scalar curvature which are equal. This implies that the potential function defined by
in
and
in
is a weak solution of the nonlinear elliptic partial differential equation
in
. In addition
is
and smooth away from
. It then follows from elliptic regularity theory together with a bootstrap argument that this function is in fact smooth. Hence, by gluing the Kähler metrics
and
on the different pieces constituting
, we have produced a Kähler metric on
which has constant scalar curvature. This will end the proof of the main Theorem. It remains to explain how to find the functions
,
,
,
and the parameters
. To this aim, let us assume that
since only notational changes are needed to handle the general case
. It follows from the result of Lemma 4.5 and Lemma 4.8 that the following expansions hold
| |
| |
We change parameters and define the functions
and
by
| |
| |
and also
Recall that the functions
are assumed to satisfy ( 18 ) while the function
is not assumed to satisfy such a constraint. The role of the scalar
is precisely to recover this lost degree of freedom in the assignment of the boundary data. With these new variables, the expansions for both
and
can now be written as
| |
| |
The data functions
,
are assumed to be bounded by a constant
times
in
and the data functions
and
are assumed to be bounded by a constant
times
in
. The terms
are nonlinear terms of
which are bounded by a constant (independent of
times
provided
is chosen small enough.
We recall the following result :
Lemma 5.1.
[
2]
The mapping
| |
| |
is an isomorphism.
Using this Lemma, it is straightforward to check that the solvability of ( 35 ) reduces to a fixed point problem which can be written as
where the nonlinear operator
satisfies
for some constant
which does not depend on
, provided
is small enough. We finally choose
and
, where
is fixed small enough. We have therefore proved that
is a map from
into itself. This mapping is clearly continuous and if in addition it were compact, the application of Leray-Schauder's fixed point theorem would directly guaranty the existence of a fixed point. However, the nonlinear equations we have solved being fully nonlinear, there is no gain of regularity and hence To overcome this last difficulty, we define a family of smoothing mappings
[1] such that
|
(36)
|
and replace the nonlinearities
by
. This time the corresponding operator
is compact and for
sufficiently small, maps the ball of radius
to itself.
Hence it has a fixed point
when
is small enough. Finally, the fixed points are bounded uniformly in
, so for any fixed
we may extract a sequence
such that
converges in
to a fixed point of
. This completes our proof of the existence the Kähler metric on
which has constant scalar curvature.
Observe that the use of smoothing operator is not strictly necessary and the fixed point for
could have been obtained through the application of a fixed point theorem for contraction mappings, but this would have required more work in §4.4 and §4.5 to prove that
is a contraction for
small enough.
6 Understanding the constraints
In this section, we give the proof of Lemma 1.1 and Lemma 1.2.
As usual, let us denote by
the set of independent functions which span the kernel of
. We assume that, for
, the functions
are normalized to have mean
and to be mutually
-orthogonal. We keep the notations of the introduction and define the matrix
as in ( 1 ) and the integer values functions
and
as in ( 1 ).
Lemma 6.1.
Assume that
. Then the set of points
such that
is open and dense in
.
Proof : Observe that it is sufficient to consider the case where
, since increasing the number of columns of the matrix can only increase its rank !
The proof is by induction on
. When
there is nothing to prove. Now, assume that the result if true for
functions. We write the function
as a linear combination of functions depending on
, the coefficients of which are functions which depend on
. In other words, we expand the determinant
with respect to the last row. We obtain
where
is (up to factor
) equal to the determinant of the matrix
where
and
. By assumption, the set of
for which
is open and dense in
. We fix
such that
. The functions
being independent, the set of
for which
is open and non empty. Furthermore, unique continuation theorem for the solutions of
implies that the the set of
for which
is dense. This completes the proof of the result.
The second condition for our construction to work asks for the existence
, with
, such that there exists a solution
to the system
This amounts to ask for the existence of points
such that the kernel of
contains a vector whose entries are all positive, i.e. an element of
, the positive cone in
. Equivalently, we have to prove that it is possible to find points
such that the image of
is included in a hyperplane of
whose normal belongs to
.
Observe that, according to Lemma 6.1 , for all
, there exists an open and dense set
such that for all
, the kernel of
is
-dimensional and in fact varies continuously as the points change in
. Equivalently, for any choice of the points in
, the image of
is a
-dimensional subspace of
which also varies continuously as the points change in
. This clearly shows that the set of points
such that
and
is an open (probably empty !) set. Also, it should now be clear that once we have found
satisfying both conditions, then the conditions remain fulfilled after any adjunction of points to this list.
Now, the condition that
is included in a hyperplane of
whose normal belongs to
is equivalent to the requirement that
does not contain any vector of
, i.e. that all nonzero elements of
have entries which change sign.
This being understood, for all
, we define on
the function
The previous discussion can be summarized as follows : We have to prove that, it is possible to find
such that, for all
, the vector
has entries which change sign.
As stated in the introduction, except in special cases, we have not been able to prove that the minimal number of points for which the above condition is satisfied is
even though we suspect that this is the case. Nevertheless, we have the general result :
Lemma 6.2.
There exists
and, for all
, there exists an nonempty open set
such that
, for all
.
Proof : We keep the above notations. Pick
. Then the function
has mean
(since it is a linear combination of the functions
which are are assumed to have mean
). Therefore it is possible to find
such that
Now, by continuity, we also have
for all
in some open neighborhood
of
in
, all
in some open neighborhood
of
in
and all
in some open neighborhood
of
in
. The sets
constitute an open cover of
, and by compactness one can find a finite sub-cover
Given any
and given any
, it belongs to some
and hence the
-th and the
-th entries of the vector
do not have the same sign. Therefore, we have found
points satisfying the required conditions.
One can then define
to be the least number of points for which
and
.
7 Geometric interpretations of the constraints
Let us go back to a well known interpretation of the kernel of the operator
in terms of holomorphic vector fields. We recall some important results (which go back to Matsushima and Lichnerowicz) and emphasize only the part relevant to the geometric interpretation of the analysis we have carried out in the previous sections. We refer the readers to the papers [13] and [12] and the book of Futaki [4] for a complete introduction and other applications. Given a real valued function
satisfying
and having called
the
part of the gradient of
, one can use ( 10 ) to conclude that
, i.e.
is a holomorphic vector field which vanishes somewhere on the manifold (since
certainly has critical points !). Conversely, every Killing vector field which vanishes somewhere is the imaginary part of a holomorphic vector field of the form
for some
(see for example [4] , [13] ,[12] ). If in addition
has constant scalar curvature, then, modulo the space of parallel holomorphic vector fields, every holomorphic vector field vanishing somewhere on
arises as
for some
[4] , [13] , [12] . Of course the parallel part of the Lie algebra of holomorphic vector fields does not come in since the requirement of vanishing somewhere makes it trivial.
We will therefore disregard (and in fact we have already done it in the statements of our results) these vector fields.
The correspondence just described should not shadow some important differences in dealing with holomorphic vector fields and with function annihilating the operator
. This appears clearly when comparing our sufficient conditions with the existence of holomorphic vector fields on the blow up manifold. The following is standard, but we include it for reader's convenience.
Proposition 7.1.
Let
be a holomorphic vector field on
and
any point in
. Then
lifts to a holomorphic vector field
on
if and only if
. Moreover any holomorphic vector field on
projects to a holomorphic vector field on
vanishing at
.
Proof : If
lifts to
on
, then the (holomorphic) flow associated to
preserves
, where
is the exceptional divisor, and
is fixed. Therefore
. Moreover being the blow up map an isomorphism away from the exceptional divisor, the same observation proves the last part of the proposition, since extending the projected vector field to zero at
gives a continuous vector field and hence a holomorphic one.
Conversely, we can look directly on a local chart in
, with
identified with the origin, since away from the origin the problem is trivial. Then
We consider the chart given by
for which the defining equations become
. In this chart,
Hence a holomorphic vector field on
, which can be written as
lifts to
as the vector field
In order to extend
to the exceptional divisor we have to verify that the functions
are bounded for any
near
. This clearly forces
for any
, and therefore the vanishing of the vector field
at the blow up point. Once this is verified the extension is continuous and hence holomorphic.
Whether a holomorphic vector field (vanishing somewhere) lifts to a blow up depends only on its value at the point one is blowing up. Therefore, if the space of holomorphic vector fields with zeros has complex dimension
, blowing up
points in generic position (generic in the sense described by the above proposition) one gets a manifold without such vector fields. In terms of the potentials used throughout this paper this translates in looking for
points such that
Our first condition (
) is sensitive to the zeros of the bounded functions in
with mean zero instead of their critical values.
The second condition is certainly encoding a more subtle phenomenon. In fact, this second condition should be related to some suitable stability property of the blown up manifold. The fact that some positivity condition must hold is present in all known examples in different veins, and has been deeply investigated in the case of complex surfaces with zero scalar curvature by LeBrun-Singer [13] , Rollin-Singer [15] , and for Del Pezzo surfaces by Rollin-Singer [16] . In our construction it appears naturally from the analytical approach we developed, but a geometric interpretation is certainly worth seeking.
8 Examples
8.1 The case of
A convenient way to study our problems on the blow up at points of the projective spaces is to look at
as the quotient of the unit sphere in
with complex coordinates
via the standard
-action given by the restriction of complex scalar multiplication.
It is well known that the automorphism group of
is given by the projectivization of
, whose complex dimension is
. We therefore seek for
real functions whose
-part of the gradient generate the Lie algebra of the automorphism group as explained in the previous subsection. This can be done in two equivalent ways: either by explicit computation on the automorphism group, or by relying on the equivalence described in the previous section between this and the study of the kernel of the operator
which for
with its Fubini-Study metric induced by the Hopf fibration becomes
Our problem reduces to seeking a basis of functions with mean zero of the eigenspace of the Laplacian
associated to the eigenvalue
(i.e. the eigenspace of the Laplacian
which are associated to the eigenvalue
and are invariant under the
action), and this is clearly given by the
functions
for
and
for
. Recall that we should add
to this list.
Obviously any explicit calculation will be rather troublesome. It is hence very convenient (and giving best results) to introduce symmetries acting on the projective space in order to reduce as much as possible the elements of the kernel of
which are invariant under these symmetries.
Example 1: Let us consider the group
acting on
generated by the transformations
Of course, the action of any element of the group on
maps the unit sphere into itself. The space of elements of the kernel of
which are invariant under the action of the elements of
is generated by
.
Let us fix the following set of blow up points
We find the matrix
whose rank is clearly equal to
and has a one dimensional kernel spanned by the vector
which has positive entries ! This proves that, working equivariantly with respect to the action of the group
, the blow up of
at the above
points carries a constant scalar curvature Kähler form.
We have then proved
Corollary 8.1.
The blow up of
at
has a constant scalar curvature Kähler metric.
It is worth remarking that this result is optimal in the number of points to be blown up, since for fewer points the manifold would have nonreductive automorphisms group, and hence no Kähler metrics of constant scalar curvature by the Mathushima-Lichnerovicz obstruction. Another interesting aspect of this example is that the manifold obtained still has non trivial (in fact
dimensional) automorphism group. The point is that the surviving automorphisms are precisely those which are not
-invariant.
Exemple 2 : Now, we still work equivariantly with respect to the action of the group
defined above but we fix the following set of blow up points
and
where
. This time, we find the matrix
whose rank is again equal to
and has a two dimensional kernel containing the vector
which has positive entries, provided
! This proves that, working equivariantly with respect to the action of the group
, the blow up of
at the above
points carries a constant scalar curvature Kähler form.
It is an easy general observation that the addition of points to a list of points satisfying our sufficient conditions preserves these conditions satisfied. In particular we can add one point to the above lists, so, for example, the
-orbit of any point
(which does not initially belong to the list) and keep the two conditions fulfilled. Observe that, for generic choice of the point
the
orbit of
has
points, so this substantially increases the number of points one can blow.
However, one can add to one of the above lists
points of the form
where
,
, so that the list of points remains invariant under the action of
. This clearly increases the number of blow up points by
. Using this idea and starting from the list of points given in example 1, one shows that the blow up of
at
points carries a constant scalar curvature Kähler form and starting from the second list of points, one shows that the blow up of
at
points carries a constant scalar curvature Kähler form. Therefore, we have obtained :
Corollary 8.2.
The blow up of
at
(special) points has a constant scalar curvature Kähler metric.
Example 3 : The important observation is that, so far we have worked equivariantly and the
blow up points cannot be chosen into some open set of
. We now give a upper bound for the number
which corresponds to the least number of points (largerthan
) for which the conditions
and
are fulfilled. Given
satisfying
and
, we consider the following set of points :
| |
| |
| |
| |
where
correspond to the indices of the nonzero entries. There are exactly
such points and it is easy to check that both conditions are fulfilled.
Indeed, the first condition
is easy to check and left to the reader. Concerning the second condition, we show that
is in the kernel of
. Observe that this condition can be translated into the fact that for each
or
where summation over
is understood. Observe that it is enough to check that the formula holds for the functions
and the functions
. It is now easy too check that
.
Therefore, we have obtained the :
Corollary 8.3.
When
, we have the estimate
.
Example 4 : For general Kähler manifolds with space of holomorphic vector fields of dimension
, one needs to blow up
points in general position to have a manifold without holomorphic vector fields. In this respect
is very special since it is easy to observe that
points suffice provided they form a so called projective frame, namely any choice of
of them are linearly independent in
(such sets of points are also often called in generic position with respect to hyperplanes). The freedom of choices of projective frames ranges clearly in an open and dense subset of
.
Suppose that we find a projective frame
for which we can prove that
has a Kähler constant scalar curvature metric necessarely, being
, using some equivariant construction. The first simple observation is that
is biholomorphic to
for any other projective frame
. The manifold obtained is also with at most discrete automorphisms and we can then apply to it the results of [2] to blow up any other set of points and still get Kähler constant scalar curvature metrics.
For all these reasons we now seek for a projective frame for which some equivariant construction works. To this aim, consider the group
of permutations on the
affine coordinates and the points
with
. The equivariant kernel is spanned by
and therefore the matrix we obtain is
Our sufficient conditions are then fulfilled for
. It is easy to see that the points given by the orbits of
(which is fixed), and
(which consists of
elements) form a projective frame and hence the resulting manifold is without holomorphic vector fields.
Corollary 8.4.
Given
points in
which contain a projective frame, then the blow up of
at these
points has a constant scalar curvature Kähler metric. Moreover the
of these points which form a projective frame can move in an open set of
and the remaining
are arbitrary.
This example naturally rises an important comment. For
Tian's solution of the Calabi Conjecture tells us that
admits Kähler-Einstein metrics (of course in the classes
) as long as no three collinear points are blown up, no
of them lie on a quadric and
on a cubic. These conditions of course trivially prevent the blown up manifold to have a Kähler-Einstein metric. Moreover it is well known that, under these assumptions, for
,
has only discrete automorphisms.
Therefore, as noted in [2] , we can use these metrics for successive blow ups and apply the main result in [2] to represent the classes
with metrics of constant scalar curvature. What we have gained with the present construction is firstly to represent also the classes
with canonical positively curved metrics, therefore getting another open set (by perturbation arguments) in the Kähler cone. Applying their technique on parabolically stable bundles over riemann surfaces Rollin-Singer proved the same result as ours for
.
Moreover for
the open set of points for which our construction works is larger than the above mentioned conditions for the existence of a Kähler -Einstein metric. For example once four points forming a projective frame are blown up, we can add a fifth aligned to two of the previous ones still getting a canonical metric.
Example 5 : The case of
. The type of manifolds we now treat are of particular interest since they have been a central object of study in the papers of LeBrun-Singer, Kim-LeBrun-Pontecorvo up to the recent works of Rollin-Singer, when
is taken to be a Riemann surface. The point is that a complete understanding of these examples leads via algebraic geometric techniques to the relation with stability of rank two vector bundles over Riemann surfaces. By understanding via a different approach these models in our more general setting, we hope to give a tool to the study of similar approach to higher rank vector bundles over any Kähler constant scalar curvature manifold.
We assume throughout this example that
is a Kähler manifold of any dimension with no nonvanishing holomorphic vector fields. Moreover on
we consider the product metric, where
is endowed with the Fubini-Study metric normalized as in the previous subsection though for any positive value of the scalar curvature the results we obtain would hold verbatim.
With these conventions the bounded kernel of the operator
is naturally identified with
,
and
the functions on
which have been described above.
We then look at the group generated by
which reduces the invariant kernel to be generated by
. Let us choose the points
and
. The first condition then requires the matrix
Obviously the two conditions are fulfilled. Since the
-orbits of the points chosen are projectively the points themselves, we have proved :
Corollary 8.5.
Given any two points
and
in
(possibly coinciding), the blow up of
at
,
has a constant scalar curvature Kähler metric.
A stronger version of the above Corollary was proved by LeBrun-Singer when
is a Riemann surface of genus at least
in [13] (in the case
is a torus one must recall that all holomorphic vector fields are parallel so they do not interfere in the analysis of the operator
). In fact they proved that the metric on the blow up can be chosen to have zero scalar curvature. For
we have gained the freedom of assigning any sign to the constant scalar curvature abtained.
Example 6 : The case of
. The results for this case easily follow from the corresponding results for the previous examples.
The case of
falls directly in the previous discussion since the blow up of this manifold at
points is biholomorphic to the blow up at
suitably choosen points of
. By the previous calculation we already know that for
our problem has a positive solution, and that this is the least number of points for which this can happen. Moreover we have already seen that for
points onwards (suitably chosen in an open set) the manifold has only discrete automorphisms, therefore iteration can start.
Concerning the higher dimensional examples, for sake of simplicity we illustrate the case of
, but everything can be easily extended to any of these products. Let us consider affine coordinates
on
and
on
and let us denote by
the functions elements of the kernel on
as described above and by
the functions elements of the kernel on
. To get the optimal result we have to work equivariantly and we consider the diagonal action used in the first example. The equivariant bounded kernel is then spanned by the union of the generators of each single factor, hence by
,
and
.
Let us choose the points
The matrix is then
which is clearly of rank
and whose kernel contains elements with positive entries.
Therefore the blow up has a Kähler constant scalar curvature metric. It is moreover not difficult to see that the blow up of
has no holomorphic vector fields and therefore can be used both for moving these points in some open set and for starting iterations of further blow ups. Note that this examples generalizes immediately to any dimension in the following
Corollary 8.6.
Suppose
. The blow up of
at
special points has a constant scalar curvature Kähler metric. Moreover, if the projections of these points to the two factors contain projective frames, these points can move in an open set of
and the other ones are arbitrary.
References
-
S. Alinhac and P. Gérard, Opérateurs pseudo-différentiels et théorème de Nash-Moser, InterEditions/Editions du CNRS (1991).
-
C. Arezzo and F. Pacard F., Blowing up and desingularizing Kähler orbifolds with constant scalar curvature, math.DG/041240522.
-
E. Calabi, Métriques kählériennes et fibrés holomorphes, Ann. Sci. Icole Norm. Sup. 4 12 (1979), no. 2, 269-294.
-
A. Futaki, Kähler-Einstein metrics and Integral Invariants, Lecture Notes in Math. 1314 (1988), Springer-Verlag.
-
P. Griffiths and J. Harris, Principles of algebraic geometry, Wiley-Interscience Publ., 1978.
-
J. Kim, C. LeBrun and M. Pontecorvo, Scalar-flat Kähler surfaces of all genera, J. Reine Angew. Math. 486 (1997), 69-95.
-
J. Kim and M. Pontecorvo, A new method of constructing scalar-flat Kähler surfaces, J. Differential Geom. 41 (1995), no. 2, 449-477.
-
A. Kovalev and M. Singer, Gluing theorems for anti-self-dual spaces, GAFA 11 (2001), 1229-1281.
-
C. LeBrun, Counter-examples to the generalized positive action conjecture Comm. Math. Phys. 118 (1988), no. 4, 591-596.
-
C. LeBrun, Scalar-flat Kähler metrics on blown-up ruled surfaces J. Reine Angew. Math. 420 (1991), 161-177.
-
C. LeBrun, On the scalar curvature of complex surfaces, Geom. Funct. Anal. 5 (1995), 619-628.
-
C. LeBrun and S. Simanca, Extremal Kähler metrics and complex deformation theory, Geom. Funct. Anal. 4 (1994), 298-336.
-
C. LeBrun and M. Singer, Existence and deformation theory for scalar flat Kähler metrics on compact complex surfaces Invent. Math. 112 (1993), 273-313.
-
R. Melrose, The Atiyah-Patodi-singer index theorem, Research notes in Math 4 (1993).
-
Y. Rollin and M. Singer, Non-minimal scalar-flat Kaehler surfaces and parabolic stability, math.DG/0404423, to appear in Invent. Math.
-
Y. Rollin and M. Singer, Construction of Kaehler surfaces with constant scalar curvature, math.DG/0412405.
-
S. Simanca, Kähler metrics of constant scalar curvature on bundles over
, Math. Ann. 291 (1991), no. 2, 239-246.
claudio.arezzo@unipr.it, Dipartimento di Matematica, Universita' di Parma, Via M. D'Azeglio, 85, 43100 Parma, Italy pacard@univ-paris12.fr, University Paris 12 and Institut Universitaire de France, France