Geodesic Equations on asymptotically locally Euclidean Kähler manifolds
Abstract.
In this paper, We solve the geodesic equation in the space of Kähler metrics under the setting of asymptotically locally Euclidean (ALE) Kähler manifolds and establish global regularity of the solution. The solution of the geodesic equation is then related to the uniqueness of scalar-flat ALE metrics. To this end, we study the asymptotic behavior of -geodesics at spatial infinity. We will prove the convexity of Mabuchi energy along -geodesics under the assumption that the Ricci curvature of a reference ALE Kähler metric is non-positive. However, by testing the Ricci curvature of ALE Kähler metrics, we find that on the line bundle over with and , all ALE Kähler metrics cannot have non-positive (or non-negative) Ricci curvature.
1. Introduction
In the paper, we study the geodesic equation in the setting of ALE Kähler cases, assuming relatively weak fall-off conditions. Let be a complete non-compact Kähler manifold of complex dimension (), we say is ALE if there is a compact subset such that is a diffeomorphism, where is a closed ball in with radius and is a finite subgroup of (any ALE Kähler manifold has only one end according to [19, Proposition 1.5, 3.2]) and the metric satisfies the following condition on the end :
-
The metric is asymptotic to the Euclidean metric at the end with decay rate for some , i.e., for ,
(1.1)
The fall-off condition is the weakest decay rate to make the ADM mass coordinate-invariant in general, referring to Bartnik [4] and Chrus̀ciel [11].
One of the difficulties in building up a general theory of scalar-flat Kähler metrics in the ALE setting is that the decay rate of such metrics to their asymptotic models is not good enough compared to the Ricci-flat case. For instance, consider the family of scalar-flat Kähler metric constructed on by LeBrun [23],
where , are constants, , , are three invariant vector fields on -sphere and is a radial function on . It can be checked that , where denotes the geodesic distance from a fixed basepoint, indicating that the Kähler potential function should be of growth. In Arezzo-Pacard [2, Lemma 7.2], an expansion theorem is proved for scalar-flat Kähler metrics in the complement of in , where is a finite subgroup of , assuming that the -lemma holds in this situation. In [30], the author proved a lemma and an expansion theorem under the setting of asymptotically conical (AC) Kähler manifolds. Here, we only need a theorem of weaker version under the setting of ALE Kähler manifolds.
Theorem 1.1.
(Yao 2022) Let be an ALE Kähler manifold asymptotic to . Let , be Kähler forms in the same Kähler class of with the corresponding metrics satisfying (1.1) and such that the scalar curvatures of and are equal, . Then
(1.2) |
for some depending on .
Let be the corresponding Kähler form of . According to Theorem 1.1, given two Kähler forms , if the corresponding ALE Kähler metrics satisfy the decay condition (1.1) and that the scalar curvatures of and are identically equal, , then and decays at infinity with higher rate , with , for some . Hence, for the prescribed scalar curvature problem, we consider the following restricted weighted Kähler potential space,
where the class of functions, , is defined as follows
Define , , for any two boundary data . Also introduce the linear reference path in . Another path in with the same endpoints is called a geodesic in if
(1.3) |
As observed by Donaldson [15] and Semmes [26], the geodesic equation is equivalent to a homogeneous complex Monge-Ampère equation in the product space , where can be embedded as an annulus in . Notice that any path of functions on can be viewed as a function on via . Let , where is the projection from to and is computed on . Then the equation (1.3) can be rewritten as follows:
(1.4) | |||
(1.5) | |||
(1.6) |
In [15], Donaldson proposed a program to attack the existence and uniqueness problems regarding canonical metrics by studying the geometric structure of the potential space , where the geodesic equation plays a central role. In the cases of compact Kähler manifolds, Chen [9] showed that for any , , the geodesic equation has a unique solution up to -regularity. Blocki [6] and He [18] built up direct calculations to prove the gradient estimate and Laplacian estimate. The full estimate was proved by Chu-Tosatti-Weinkove in [12]. In the other direction, Lempert-Vivas [24] and Darvas-Lempert [14] constructed counter-examples to assert that is not continuous in general, hence the regularity is optimal in general. In [3], Auvray generalized the -regularity to singular cases (precisely, there exist cusp singularities along simple normal crossings). The main theorem of sections 2-5 is to generalize the full estimates to ALE Kähler manifolds.
Theorem A.
Let be an ALE Kähler manifold and . Then and can be connected by a geodesic solving (1.4), (1.5), (1.6). Moreover, there is a uniform constant depending only on , and on the geometry of such that
(1.7) |
Here, is a Kähler form on given by with the linear path introduced above, and with , where is fixed depending only on , such that .
Then, we relate the solution of the geodesic equation to the uniqueness of scalar-flat ALE Kähler metrics in each Kähler class. The main idea is to follow the framework of Chen [9] in the compact case, under the assumption that the Ricci curvature of the reference metric is non-positive. This was extended to the non-compact case with Poincaré cusp ends by Auvray [3]. In the ALE case, it is first necessary to prove sufficient decay at infinity of solutions to the -geodesic equation.
In Section 6, we discuss the asymptotic behavior of -geodesics. Given any two functions
we set and let denote the corresponding function on . We fix large depending on , such that is positive on , where with the projection. Then, we introduce the following -geodesic equations
where is a smooth nonnegative function defined in satisfies the following conditions
(1.8) |
where is an real number with . In particular, we are interested in the following case. By taking
(1.9) |
where is a smooth increasing function in equal to (resp. ) in a neighborhood of (resp. ) and is defined as follows
(1.10) |
and in this case, , . By taking to be small enough, can be written as
Due to the positivity of the right hand side of , it is well known that for every there exists a solution . We now prove:
Theorem B.
Let be the -geodesic constructed above. Then, there exists a constant depending on and on an upper bound for such that
where denotes the Levi-Civita connection of the ALE Kähler metric on , acting as a differential operator in the directions on . And
where is a function only depending on . Hence, for any two potentials , in , there exist -geodesics in connecting and .
In section 6, we prove a stronger statement. Let , then due to the fact that was chosen to be linear in (see section 6 for details).
Hence, while we still cannot define the Mabuchi -energy along geodesics, the Mabuchi -energy is now actually well-defined along -geodesics assuming .
In Section 7, the second derivative of the Mabuchi -energy will be calculated. Throughout section 7, we assume (Here it turns out that if , are only in , there would be boundary terms at infinity breaking the positivity of the second derivative. This is a new phenomenon compared to Chen [9] and Auvray [3]). However, under the assumption that the Ricci curvature of some reference ALE Kähler metric, , is non-positive, we can then prove the convexity of Mabuchi -energy:
Theorem C.
Assume that is an ALE Kähler metric on such that the Ricci curvature of is non-positive, . Then, along each -geodesic in with , , the Mabuchi -energy is convex.
A quick corollary of Theorem C is that assuming , the scalar-flat Kähler metric, if it exists, is unique in . However, if there exists a scalar-flat Kähler metric in , the condition, implies . Hence, the uniqueness of scalar-flat ALE metric can be reduced to the uniqueness result of Ricci-flat ALE Kähle metric, which can be found in reference [20, 28, 13]. The point is that implies by definition that the ADM masses of and are equal, . According to mass formula by Hein-LeBrun [19], it follows that . The assumption that implies that (see Remark 7.4 for details). In fact, in Section 8, we will prove that many ALE Kähler manifolds do not admit any ALE Kähler metrics with (or ) at all:
Theorem D.
Let be the standard negative line bundle over with , , and let be an ALE Kähler metric on with decay rate , . Then, the Ricci form of , is of mixed type, i.e., neither nor is true.
In Riemannian geometry, AE metrics of negative Ricci curvature are well-known to exist in by explicit construction in Lohkamp [25]. Theorem D gives a negative answer to this question in the setting of ALE Kähler metrics.
An interesting question in this context is to ask whether some version of the Nonexistence Theorem D holds in general ALE Kähler manifolds or even AC Kähler manifolds.
Question.
Is it true in any ALE Kähler manifold that the Ricci curvature form of an ALE Kähler metric can only be identically zero or of mixed type?
This paper is a part of the Ph.D. thesis of the author. The author would like to express his gratitude to Professor Hans-Joachim Hein and Professor Bianca Santoro for suggesting the problem, and for constant support, many helpful comments, as well as much enlightening conversation. The author is also thankful to professor Gustav Holzegel for providing financial support via his Alexander von Humboldt Professorship during the last semester at University of Münster. The whole project is Funded by the DFG under Germany’s Excellence Strategy EXC 2044-390685587, Mathematics Münster: Dynamics-Geometry-Structure, and by the CRC 1442, Geometry: Deformations and Rigidity, of the DFG.
2. -geodesic equations and openness
Recall that geodesic equations can be written as follows,
where and are constants depending on . The family of equations is called the -geodesic equations. The idea to solve the equation is the following. Firstly, we apply the continuity method to show that there exists a solution of in . In particular, consider the family of equations , . There is a trivial solution at . Then, we shall prove the openness and closedness of in certain regularity. In the current section, we deal with the openness of .
Assuming that there exists a solution of in for some , we will show in this subsection that can be solved for all in a small open neighborhood of . For simplicity, we write as in Theorem A and . Then, the equation can be written as, in , with the boundary condition on . Then, the Monge-Ampère operator is defined to be
Let be a solution of for some . By assumption, is -plurisubharmonic satisfying . Then, the linearization of Monge-Ampère operator at is uniformly elliptic, and is given by
where represents the Laplacian with respect to . Let be the functions in vanishing on the boundary . Then, we have the following property of , from which the desired openness is clear by the implicit function theorem.
Proposition 2.1.
Let be the solution of , then the linearized operator is an isomorphism for all integers and .
Proof.
Let us first prove the surjectivity. Fixing , the exhaustion argument will be applied to solve the equation . Take an exhaustive sequence of pre-compact sets, , with smooth boundary. In particular, by taking a sequence of subsets, where , and smoothing the corners, we can obtain the exhaustive sequence . Then, we can solve the following Dirichlet problems,
where . The existence of the solution of is a classic result of the Dirichlet problem on compact Riemannian manifolds with boundary. The key to complete the proof is to give the uniform estimates of . The main idea to show the uniform estimates is to construct barrier functions. Consider the function . The fact that implies . If we suppose that and take , then we have . Combining with the fact that on the boundary , the maximum principle implies that,
(2.1) |
The uniform estimates follows directly from the standard Schauder estimates. Precisely, for interior points away from the boundary, we pick a pair of balls centered at , . Then, the interior Schauder estimates implies that . If is close to the boundary, we can apply the boundary Schauder estimate. After straightening the boundary in case the boundary portion on is not flat, we can pick half balls, for some . Together with the interior estimates, we have
(2.2) |
where depends only on . After passing to a subsequence, we conclude that the limit function, , satisfies in and on . The uniqueness directly follows from the following maximum principle, Lemma 2.3. ∎
The following lemma comes from Yau’s generalized maximum principle, referring to [10, 29]. To describe the model metric on , we introduce the asymptotic coordinates of . Let be asymptotic coordinates of the end of and let be the complex coordinate of . Real asymptotic coordinates are given by , where the complex coordinates are written as . The asymptotic coordinate system will be applied to describe the asymptotic behavior of prescribed Kähler metrics on .
Lemma 2.2.
Let be the noncompact Kähler manifold as above with the Kähler metric associated with satisfying, for some uniform constant ,
in the asymptotic coordinates of . Let be a function bounded from above on . Suppose that , then there exists a sequence in such that
(2.3) |
Proof.
Let be the radial function inherited from the asymptotic chart of , for instance, . The radial function can be extended to a non-negative smooth function in the whole space satisfying the estimate
(2.4) |
for some uniform constant . Consider the function . Since tends to negative infinity as goes to infinity, achieves its maximum at some point . And must be an interior point in based on the assumption that . At , the function satisfies
and
Choosing to be points achieving the maximum of , then combining with (2.4) and letting go to infinity, we complete the proof of (2.3). ∎
The following lemma is a strengthened version of the above maximum principle, based on solving the Dirichlet problem in .
Lemma 2.3.
Let be the same as in Lemma 2.2. Suppose that is a function in and bounded from above. Suppose that satisfies in and on . Then in .
3. A priori estimate up to
From section 3 to 5, we complete the proof of Theorem A. The key ingredient is to prove uniform a priori estimates up to order for the solution of the -geodesic equation (). These estimates will be uniform with respect to and with respect to the distance from a fixed point in . (In section 6, we will also see that for a fixed it can be proved that is decaying at spatial infinity. However, we are currently unable to make these decay estimates uniform with respect to .)
These uniform estimates are then used in two ways:
-
First, they allow us to solve for any fixed via the continuity method in for any . Recall this is done by considering the family of equations with , where openness in follows from Proposition 2.1. The uniform estimates that we will prove, together with the general regularity theory of the Monge-Ampère equation, then imply closedness. Here, it is not yet important that the estimates are uniform in , and the higher estimates will depend on because the ellipticity of the equation does. Also, note that these higher-order estimates follow from standard local regularity in the interior and from [8, Section 2.1–2.2] near the boundary because we already have a true bound.
We omit these standard arguments and instead focus on the proof of the uniform a priori estimates of the -geodesic solution . For this we follow the outline of [7] in the compact case. However, we provide all the necessary details that are required to generalize this theory to the ALE case. In addition, we also make use of the recent advance [12] to obtain a estimate which is uniform in .
In this section, we only deal with the uniform estimate. We begin with a standard comparison principle [5, Proposition 3.1].
Lemma 3.1.
Let be a bounded connected domain in with smooth boundary and , plurisubharmonic functions in . If on and , then we have
Then we can prove the following maximum principle for Monge-Ampère operators.
Theorem 3.2.
Let be a fixed reference Kähler form and , the pull-back of a semipositive -form in . Assume that are bounded functions with , . If for some positive constants , , we have the following properties:
(3.1) | ||||
(3.2) | ||||
(3.3) |
then in .
Proof.
Assume at some point . Let . Then, we can modify to be as follows:
(3.4) |
It can be checked that , are bounded functions satisfying that on and . By Wu-Yau’s generalized maximum principle, there exists a sequence in such that
For a sufficiently small constant , there exist a point , and . Fix a local holomorphic chart around , with . Without loss of generality, we assume contains the unit disk in and for any local vector field ,
where the constant only depends on the geometry of and the reference metric . Let and . To derive the contradiction, we construct the following local functions in ,
(3.5) |
If we denote the unit ball contained in the coordinate chart of by , we have and on . Consider the following subset of ,
Let be the local potential of in , . According to Lemma 3.1,
(3.6) |
Taking ,
Together with the construction of and in (3.4), (3.5),
(3.7) |
By picking smaller, , and combining with (3.1), we have,
(3.8) |
Since the second term of (3.8) is strictly positive, which leads to a contradiction, we complete the proof. ∎
Let be the solution of after subtracting . According to Theorem 3.2, we have a uniform lower bound ; hence, . The upper bound is easy to construct. Consider the function defined in , . By restricting to each section , we have
Hence, in and on its boundary . The maximum principle on compact manifolds with boundary implies that on each section. Hence, we get the desired uniform estimate,
4. A priori estimate up to
For the bound, Blocki gives an explicit estimate in the compact setting in [6]. We generalize this estimate to the non-compact case. The boundary estimate follows directly from the fact that in and , , agree along . Let be the Levi-Civita connection of on . Then we have
Hence, , where is a uniform constant.
Proposition 4.1.
Let be a solution of and let be the Levi-Civita connection of the Kähler metric on . Assume that lies in the space . Then,
where is a positive constant depending only on upper bounds for , on lower bounds for the bisectional curvature of , and on , but not on .
Proof.
Suppose that and . Consider the following function,
where and is a smooth function to be determined later. According to the assumption that lies in the space , Yau’s maximum principle can be applied here. In particular, there exists a sequence in in such that,
where . Then, for a sufficiently small to be determined later and all , we have
(4.1) |
Fixing satisfying (4.1), we can pick the normal coordinates around . Let and denote the metric tensors corresponding to and . Then there exist local holomorphic coordinates near such that,
By taking derivative of ,
Combining with condition (4.1), . Then, at the point , we have
(4.2) |
If we write the local potential of as near , then the -geodesic equation is locally given by . The direct derivative of the equation at gives,
(4.3) |
Also, notice that,
where is the negative lower bound of bisectional curvature of . Recall that we have the assumption and , where is the constant from our assumption at the beginning of this section and we will get rid of this constant in the end. Together with (4.2) and (4.3), we have,
(4.4) |
According to Blocki’s key observation in [6], after modified in our case, at the point , we have
and assuming that , we have
where is a uniform constant satisfying
Combining with (4.4),
(4.5) |
Now, we choose the function and the small number in (4.5) as follows. Let and , then we have
Then, it is straightforward to conclude that . Noting that , hence, is controlled by some uniform constant only depending on , , and . ∎
5. A priori estimate up to
First, we deal with the uniform boundary estimate on . The technique is to construct local barrier functions near the boundary, which is completely parallel to [8, 9, 17]. The statement is the following:
Lemma 5.1.
Let the data be the same as in Proposition 4.1. Let denote the Levi-Civita connection of on . Then
where the constant only depends on and on .
Proof.
Fixing a point , we pick the local holomorphic coordinates around the point such that the coordinates system is normal in and direction, we still pick the standard coordinate function of the annulus, denoted by and the corresponding holomorphic coordinates, . Throughout the proof, we assume the metric tensor associated with satisfies . In general, we need to prove the boundary estimate at in tangential-tangential, tangential-normal and normal-normal directions respectively. However, the tangential-tangential is trivial in our case and the normal-normal estimate follows directly from the tangential-normal estimate. Here, we briefly summarize the proof of tangential-normal estimate by explicitly constructing the barrier functions.
Consider a small neighborhood near , , where the small constant will be determined later. Firstly, we construct the following auxiliary function in ,
(5.1) |
where is a large constant to be determined. Then, it can be easily checked that
where again denotes the metric tensor associated with and denotes the corresponding Laplacian. Notice that
By taking , we have . Noting that , we have on . Then, the barrier functions can be constructed as follows:
By differentiating the Monge-Ampère equation in the local coordinates,
where and are large positive constants to be determined. According to the estimate of , we assume that . By picking a very large constant such that, on , , we have on . Then, we choose a large constant such that in . Then, by maximum principle, in . Together with the fact that , we have at , which implies the tangential-normal estimate on the boundary. ∎
Lemma 5.1 together with Yau’s standard calculation on Laplacian estimate implies the following interior Laplacian estimate, referring to [31].
Lemma 5.2.
Let be the solution of and , , the Laplacian operators of and respectively. Then, for any constant ,
where denotes the curvature tensor of . From this, we can deduce the estimate
where only depends on and on a negative lower bound of .
Lemma 5.2, together with Lemma 5.1, implies that there exists a uniform constant only depending on such that . This is already enough to apply the standard local regularity theory of the Monge-Ampère equation to prove estimates for any that depend on a positive lower bound for . In this way the equation can be solved using the continuity path , . However, in order to construct an honest geodesic by letting , we require a full estimate which is uniform in . In [12], regularity is proved in the compact case. The method can also be applied in the ALE Kähler setting.
Proposition 5.3.
Let the data be the same as in Proposition 4.1. If lies in the space , then there exists a constant such that
where again denotes the Levi-Civita connection of the metric and depends only on and on , , , .
Proof.
We again write for the metric tensor associated with . Let be the largest eigenvalue of the real Hessian . By observing that there exists a uniform constant such that , it suffices to prove that has a uniform upper bound. Consider the following quantity,
where is defined to be and is a uniform large positive constant to be determined later. We can further modify this quantity to , where is a small positive constant to be determined later. According to the assumption that is bounded and hence so is , the modified quantity attains its maximum at some point . The same argument as in Lemma 2.2 implies that . In the following, we assume is small enough such that and always write . Since might not be smooth at if the eigenspace of has dimension greater than one, a perturbation argument used in [12] can be applied to the quantity here.
Fix normal coordinates with respect to at such that is diagonal at . Define the corresponding real coordinates by . Let be the eigenvalues of at and , the corresponding unit eigenvectors at . The eigenvectors can be extended to vector fields with constant coefficients in a small neighborhood of , also denoted by , and can be represented by in the local coordinates. The perturbation argument is to perturb locally around and to ensure that near . Precisely, consider the following locally defined tensor field,
Let . Then, one can easily check that and for . Hence, there exists a neighborhood of such that and . Consider the following perturbed quantities,
Therefore, is a smooth quantity with a local maximum at . Then, we have,
The following inequality follows directly from the calculation in [12, Lemma 2.1]. The only information we need in the calculation is the second derivative of the Monge-Ampère equation at . We will not repeat the details here. By assuming at , and again writing for the metric tensor associated with , we have
(5.2) |
where the constant only depends on and . To cancel the annoying terms, we deal with the third term in (5.2), . To estimate the term, we split it into the following two parts,
where is to be determined later. For , referring to [12, Lemma 2.2], by assuming that , where only depends on and , we have
(5.3) |
To estimate , recall the fact that and apply the derivative of eigenvalues referring to [12, Lemma 5.2]. Then, we have
(5.4) |
6. The asymptotic behavior of -geodesics
In this section, we prove Theorem B on the asymptotic behavior of -geodesics for a fixed . We use the notation introduced before Theorem B and we assume . We are really interested in the case when due to theorem 1.1. In -geodesic equation , the derivatives of function decays at infinity with order , with for . Without loss of generality, we assume , otherwise theorem B can be proved more easily without iteration (step 3).
We also write , so that the solution is given by with on . In Aleyasin [1], a rough idea is given to prove the asymptotic behavior of -geodesics by constructing barrier functions in the (strictly easier) special case where the asymptotic coordinates are -holomorphic and the decay rate of the ALE Kähler metric to the Euclidean metric is high enough. However, even in this special case, the details are more involved than what is suggested in [1]. Here we give a complete proof in the general setting.
Step 1: Differentiating the Monge-Ampère equation.
The Monge-Ampère equation can be written explicitly in the asymptotic coordinates of . As the complex structure of does not coincide with the Euclidean complex structure of the asymptotic coordinates in general, we will use real coordinates for clarity. By passing to the universal covering of the end, we are able to work with the global coordinates. Precisely, let be the asymptotic complex coordinates of and the complex coordinate of . The corresponding real coordinates are , where for . From now on:
-
Latin indices will denote the real coordinates from to .
-
Greek indices will denote the real coordinates from to .
-
The bold Greek indices will denote the real coordinates from to .
In these coordinates, we write the Riemannian metric tensors corresponding to and as and , respectively.
Throughout this section, we work in the asymptotic chart of . This allows us to use the Euclidean metric on as a reference metric to measure derivatives. This is helpful because it enables us to write down equations with a good structure. Let denote the Euclidean length, the Euclidean Levi-Civita connection and () the component of acting only in the space (time) directions on .
Then, the equation can be written as
(6.1) |
Recall that satisfies conditions in (1.8). By differentiating the log of both sides by , we have
(6.2) |
The first goal is to rewrite the equation (6.2) to be an elliptic equation in terms of . Let represent the real coordinate vector fields of . Notice that . We compute of the second term:
(6.3) |
Observe that is completely horizontal because preserves the product structure of the tangent bundle and is constant. Thus,
(6.4) |
where the coefficients depend only on and not on . In the same way, we can also see that
(6.5) |
Moreover, it is obvious that
(6.6) |
(6.7) |
where denotes a contraction and denotes the following behavior of a tensor :
Then, abbreviating the estimates
by and
by the equation (6.2) can be rewritten as
(6.8) |
Step 2: Barrier estimate of the first derivatives.
The next target is to construct the upper barrier and lower barrier functions to control . Consider a smooth cutoff function satisfying for , for and , for . The function can be extended smoothly to by defining
(6.10) |
where is a large positive constant to be determined later such that is contained in the asymptotic chart of and . From (6.9),
(6.11) |
Then, we can pick a barrier function as follows:
(6.12) |
where the constant is to be determined later. The barrier function is defined in with on . We also have
Using the estimate , we obtain that
(6.13) |
By taking
(6.14) |
and comparing with the inequality (6.11), we have . Together with the fact that on , Lemma 2.2 implies that in . The same method shows the upper bound , which, together with the lower bound, implies that for each spatial index ,
(6.15) |
Step 3: Barrier estimate of the second derivatives.
Now, it comes to deal with the asymptotic behavior of the second derivative.
For a preliminary estimate, we go back to the full formula (6.8) for . For every , the Euclidean norm of the right-hand side on a restricted unit ball with is still bounded by thanks to the Evans-Krylov estimates applied to in the interior and the estimates of [8, Sections 2.1–2.2] at the boundary. (The precise dependence of this constant on the ellipticity, and hence on , is not clear but also not needed.) Likewise, the norm of the coefficient tensor of the PDE, , is bounded by . Applying the classic interior and boundary Schauder estimates to (6.8), we thus obtain from (6.15) that
(6.16) |
These estimates will now be used to start a bootstrap to obtain some decay for using the same barrier method as in Step 2. Differentiate the equation (6.8) again by for . This yields
(6.17) |
As before, we have that , and we also have
(6.18) |
thanks to the preliminary estimate (6.16). Similarly, all derivatives of on the right-hand side of (6.17) are at worst of order , with at least one purely spatial derivative, and hence can be bounded by thanks to (6.16). In this way, we obtain that
(6.19) |
The majority of terms on the right-hand side actually decay faster than , and the only term that might decay more slowly is . So far, we can only bound this by . However, by applying the same method as in the weighted estimate of the first derivative in Step 2, we can then construct the following barrier function for :
(6.20) |
where is the same constant as in (6.12) and is another uniform constant depending on , , , , and on the constant of (6.16). Hence, we get the weighted estimate for :
(6.21) |
Step 4: Iterative improvement of the barrier estimates.
In this step, we improve the decay order of the estimates we obtain in (6.16) and (6.22) by an iteration argument. Recall that from Steps 2–3 we have the following weighted estimates to start the iteration process (see (6.16) and (6.22)):
(6.23) |
To complete the iteration argument, we need to improve the decay of the term . More precisely, this term occurs in a combination in the first derivative estimate (Step 2), and in combinations and (to get optimal decay rate of , we need to analyze this term) in the second derivative estimate (Step 3). We will now analyze these combinations more carefully. All constants in this step may depend on . Let be a continuous function defined in with at most polynomial growth rate at infinity, for simplicity, we introduce the notation to denote the decay rate of and , to denote the decay rate of , respectively.
The metric tensor and its inverse can be written as -matrices
where , are -matrices, , are -matrices and , are -matrices. By direct calculation, we have
(6.24) |
The fact that implies that . The weighted estimate (6.23), together with the fall-off condition of the metric , implies that . Then, from (6.24), we have
(6.25) |
Similarly, let denote the matrix of in asymptotic coordinates. If we write
then we have
(6.26) |
where . According to the estimate (6.23), and hence as well because are uniformly bounded. Moreover, are all uniformly equivalent to but there is no reason for to decay. Then (6.25) and (6.26) imply that
(6.27) |
Then, by calculating blockwise and using that , we have
(6.28) |
By inserting (6.27), (6.28) into (6.8), we have
(6.29) |
For the last but one term of (6.17),
(6.30) |
For the first term of right-hand side of (6.30), by using , we obtain that the decay rate of the first term is given by . For the second and third terms, we need to analyze . Similar to (6.28), we have
(6.31) |
By inserting (6.27) into (6.31), we have
(6.32) |
Then, inserting (6.27), (6.28) into (6.17), we have
We can go one step further by applying Schauder estimates to (6.8) and (6.17) and to obtain estimates for and in . Indeed, those terms on the right-hand side of the PDEs (6.8), (6.17) that were known to decay pointwise with rate already after Step 3 are actually also decaying at rate in norm. This is clear from (6.23). So we just need to find the decay rates of the most difficult terms, in (6.8) and , in (6.17) in norm as well. For this, we need to go back and also estimate the -norm of and in , as follows. By using (6.23), we have that
Then, based on (6.25), we have that
(6.33) |
Then we can proceed as in (6.28) and (6.32), obtaining that the decay rates of and are and respectively. According to the classic interior and boundary Schauder estimates, we improve (6.29) to norm,
(6.34) |
Inserting (6.23) into (6.34), and using (6.34) again to improve (6.23), we finally obtain the following estimates:
(6.35) |
Note that according to (6.35), because was chosen to be linear in , the decay rate of is faster than the decay rate of the boundary data .
Step 5: Proof of Theorem B
In Step 4, we have obtained the optimal decay rates in the cases of (even though it is not required in the proof of Theorem B). In this step, we give optimal estimates for and complete the proof of Theorem B.
For the higher order derivatives, by differentiating the Monge-Ampère equation (6.2) times, similar to (6.8) and (6.17) and writing (, for ), instead of giving a full formula as (6.8) and (6.17), we write a simplified formula of :
(6.36) |
Applying induction on , according to iteration process (step 4), we can assume for
(6.37) |
To find the optimal decay rates, the most difficult term is . Notice that by (6.31) and (6.34), we have
where , are -, -multi-indices respectively. Then, we apply induction on to find the decay rate of , where is a -multi-index. Applying one derivative to , by (6.30), we can prove that
(6.38) |
Then, by (6.30) and (6.38), we have
Combining with (6.37), we have that the right-hand side of (6.36) is . Using the construction of barrier functions in Step 2–3, we obtain that . To apply Schauder estimates to the -th derivative of Monge-Ampère equation, we also need to know the decay rate of :
Hence, we have , for . To prove that is in , by integrating in the radial direction from infinity to , we obtain a function defined in with decay rate . Then,
(6.39) |
where is a function in independent of radius and be viewed as a variable on the link. It suffices to prove that is independent of . By taking derivative of (6.39), we have . In the case that is not constant with respect to , , which contradicts to the fact that . Hence we proved that . We conclude that, for ,
In conclusion, we have proved Theorem B.
7. Convexity of the Mabuchi -energy
According to Theorem 1.1 (assuming ), we can restrict ourselves to the space
and the function is constructed by (1.9) and (1.10). In the previous section, we proved that for any two given boundary data , there exists a solution of the -geodesic equation in the same space .
The derivative of the Mabuchi -energy can be defined as follows: for ,
The integral converges because , equivalently, . In the following proposition, the second derivative of Mabuchi -energy will be calculated in containing boundary terms, and it will be clear that these boundary terms go to zero as . Precisely, we consider Mabuchi -energy restricted in ,
(7.1) |
The calculation of the second variation of is due to my advisor Bianca Santoro in one of her unpublished notes, several years before I started this project. The limiting case was previously stated by Aleyasin [1] without details concerning the vanishing of boundary terms.
To simplify the notation, in the following proposition, we write , , , , and , where is a covariant derivative of with respect to . Recall that is called the Lichnerowicz operator, and if and only if is a holomorphic type vector field.
Proposition 7.1 (Santoro).
Proof.
By taking the second derivative of Mabuchi -energy in , we have
(7.4) | ||||
The second term of (7.4) needs one integration by parts, and we get
Now, to the term . For simplicity, ,
where is the traceless part of Ricci. If is any primitive -form, then
Hence,
Note that
Thus, we get that
(7.5) |
Lemma 7.2.
Let be a smooth function defined on . Then we have that
Hence,
(7.6) |
Proof.
Notice that
Then, we have
Here . Then, we have
The Stokes’ theorem can be applied to the first term in the above formula by observing that if we write , then . Hence,
Similarly,
We have
Notice that
and integrating by parts,
Hence, we proved that
which completes the proof of the lemma. ∎
Theorem 7.3.
Assume that is an ALE Kähler metric on such that the Ricci curvature of is non-positive, . Then, along each -geodesic in , , the Mabuchi -energy is convex.
Proof.
The proof is parallel to Chen [9]. Here, we just do the calculation in the ALE setting. Define . Then the -geodesic equation can be written as
According to (7.3), together with the observation, , we have
We have the last equality because and , so that the relevant boundary integral vanishes. Hence, we have proved the convexity of the Mabuchi -energy. ∎
Remark 7.4.
A quick corollary of Theorem 7.3 is that assuming , the scalar-flat Kähler metric if it exists, is unique in . The proof is also parallel to Chen [9]. However, if there exists a scalar-flat Kähler metric in , the condition, , implies . Hence, the uniqueness of scalar-flat ALE Kähler metric can be reduced to the uniqueness result of Ricci-flat ALE Kähler metric (which can be found in many reference [20, 28, 13]). A short proof is given as follows. Let be a scalar-flat Kähler metric in . The fact, , implies that the ADM mass of and are equal, . According to the mass formula by Hein-LeBrun [19],
where is a constant only determined by the dimension , the first Chern class of and the cohomology class of and only depends on dimension . The fact, , together with the mass formula, implies that
The assumption that implies that . Then, by a simple argument, we can prove that all scalar-flat ALE Kähler metrics in is actually Ricci-flat. The expansion of scalar-flat Kähler metrics (Theorem 1.1) implies that the Ricci form, , decays to zero at infinite with decay rate faster than . The ddbar lemma implies that there exist such that
Taking trace with respect to , we have that . By solving the Laplacian equation (for instance, see [30, Propsition 2.3]), there is a unique solution in the space (for ). Hence, , which implies that is Ricci-flat.
8. Nonexistence of non-positive (or non-negative) Ricci curvature
Consider the standard family of negative line bundles, , over together with their natural projections . The total spaces of are fundamental examples of ALE Kähler manifolds by viewing as a resolution space of . Let be any ALE Kähler metric on asymptotic to the Euclidean metric with decay rate (). In the following, we shall prove the nonexistence of a sign of the Ricci curvature of in the case . When , there always exists a Ricci-flat ALE Kähler metric in each compactly supported ALE Kähler class, see [21, 22, 27].
Theorem 8.1.
Let be the standard negative line bundle over with and . Let be an ALE Kähler metric on with decay rate . Then, the Ricci form of , , is of mixed type, i.e., neither nor is true.
Proof.
Notice that for each integer , there is a compactification of by adding a divisor at infinity, . We denote the compactified manifold as and the natural embedding is holomorphic. is a -bundle over . Denote as the divisor corresponding to the base manifold, . Then, the normal line bundles of and are given by
(8.1) |
The following facts on the geometry of can be checked by viewing as a smooth toric variety. can be described by coordinate charts with coordinates , (), where the coordinates are related by
The divisor classes of are generated by the class of , the zero section of , and the class of , the total space of the restriction of the -bundle to a linear subspace of . Restricting and to , we can write
The divisor at infinity, , can be represented by and can be represented in terms of and as follows,
(8.2) |
By viewing , and as smooth complex hypersurfaces of , the Poincaré duals of , and have natural explicit representatives denoted by , , respectively. For instance, in ,
(8.3) | ||||
(8.4) | ||||
(8.5) |
Step 1: Extension of the ALE Ricci form to .
Recall that the diffeomorphism gives a holomorphic asymptotic chart of . The diffeomorphism can be explicitly written as
In the coordinate chart , we have . By the asymptotic condition of , in an asymptotic chart of , can be viewed as a function of decay order , where is the standard Euclidean metric on the asymptotic chart. Thus, the Ricci form satisfies
The adjunction formula tells us that as line bundles over ,
Since is the curvature form of a Hermitian metric on and is the curvature form of a Hermitian metric on , it follows that
By restricting in (8.3) to the asymptotic chart of , we have
Hence, by Theorem 1.1, can be written as
Since cannot be extended smoothly to , we define a smooth cut-off function ,
and we define . Applying the cutoff function, we can extend to be
Step 2: Integral argument for .
Recall that the intersection numbers between , and are given by
(8.6) |
In particular, if we integrate over , then
(8.7) |
On the other hand, we have pointwise and uniformly as . Hence, by the dominated convergence theorem,
(8.8) |
Now assume that is seminegative (or semi-positive). Then the left-hand sides of both (8.7) and (8.8) are non-positive (or non-negative). However, the right-hand sides have opposite signs because . This is a contradiction.
Step 3: Integral argument for .
In higher dimension, the difficulty is to calculate the intersection numbers of divisors. However, in the case of , we can apply the formula of intersection numbers on toric varieties [16, Chapter VII.6], or, more explicitly, take integral of formulas of Poincaré dual (8.3)–(8.5). Notice that
(8.9) |
Then, we have
On the other hand, we have pointwise and uniformly as . Hence, by the dominated convergence theorem,
where the last equality can be observed from (8.2) and (8.9):
because . By the same argument as in dimension , we complete the proof. ∎
References
- [1] S. Ali Aleyasin, The space of Kähler potentials on an asymptotically locally euclidean Kähler manifold, preprint, arXiv:1311.0419 (2013).
- [2] C. Arezzo and F. Pacard, Blowing up and desingularizing constant scalar curvature Kähler manifolds, Acta Math. 196 (2006), 179–228.
- [3] H. Auvray, The space of Poincaré type Kähler metrics on the complement of a divisor, J. reine angew. Math. 722 (2017), 1–64.
- [4] R. Bartnik, The mass of an asymptotically flat manifold, Comm. Pure Appl. Math. 39 (1986), 661–693.
- [5] E. Bedford and B. A. Taylor, The Dirichlet problem for a complex Monge-Ampère equation, Invent. math. 37 (1976), 1–44.
- [6] Z. Błocki, A gradient estimate in the Calabi-Yau theorem, Math. Ann. 344 (2009), 317–327.
- [7] S. Boucksom, Monge-Ampère equations on complex manifolds with boundary, Lecture Notes in Mathematics, vol. 2038, Springer, 2012, pp. 257–282.
- [8] L. Caffarelli, J. Kohn, L. Nirenberg, and J. Spruck, The Dirichlet problem for nonlinear second order elliptic equations, II: Complex Monge-Ampère, and uniformly elliptic equations, Comm. Pure Appl. Math. 38 (1985), 209–252.
- [9] X. Chen, The space of Kähler metrics, J. Differ. Geom. 56 (2000), 189–234.
- [10] S.-Y. Cheng and S.-T. Yau, On the existence of a complete Kähler metric on non-compact complex manifolds and the regularity of Fefferman’s equation, Comm. Pure Appl. Math. 33 (1980), 507–544.
- [11] P. Chruściel, Boundary conditions at spatial infinity from a Hamiltonian point of view, Topological Properties and Global Structure of Space-Time (Erice, 1985), Plenum Press, 1986, pp. 49–59.
- [12] J. Chu, V. Tosatti, and B. Weinkove, On the regularity of geodesics in the space of Kähler metrics, Ann. PDE 3 (2017), 1–12.
- [13] R. J. Conlon and H.-J. Hein, Asymptotically conical Calabi–Yau manifolds, I, Duke Math. J. 162 (2013), 2855–2902.
- [14] T. Darvas and L. Lempert, Weak geodesics in the space of Kähler metrics, Math. Res. Lett. 19 (2012), 1127–1135.
- [15] S. K. Donaldson, Symmetric spaces, Kahler geometry and Hamiltonian dynamics, Translations of the AMS, vol. 196, American Mathematical Society, 1999, pp. 13–34.
- [16] G. Ewald, Combinatorial convexity and algebraic geometry, Graduate Texts in Mathematics, vol. 168, Springer, 1996.
- [17] B. Guan, The Dirichlet problem for complex Monge-Ampère equations and regularity of the pluri-complex Green function, Comm. Anal. Geom. 6 (1998), 687–703.
- [18] W. He, On the space of Kähler potentials, Comm. Pure Appl. Math. 68 (2015), 332–343.
- [19] H.-J. Hein and C. LeBrun, Mass in Kähler geometry, Comm. Math. Phys. 347 (2016), 183–221.
- [20] D. D. Joyce, Compact manifolds with special holonomy, Oxford University Press, 2000.
- [21] P. B. Kronheimer, The construction of ALE spaces as hyper-Kähler quotients, J. Differ. Geom. 29 (1989), 665–683.
- [22] P. B. Kronheimer, A Torelli-type theorem for gravitational instantons, J. Differ. Geom. 29 (1989), 685–697.
- [23] C. LeBrun, Counter-examples to the generalized positive action conjecture, Comm. Math. Phys. 118 (1988), 591–596.
- [24] L. Lempert and L. Vivas, Geodesics in the space of Kähler metrics, Duke Math. J. 162 (2013), 1369–1381.
- [25] J. Lohkamp, Metrics of negative ricci curvature, Annals of Mathematics 140 (1994), no. 3, 655–683.
- [26] S. Semmes, Complex Monge-Ampère and symplectic manifolds, Amer. J. Math. 114 (1992), 495–550.
- [27] C. van Coevering, Ricci-flat Kähler metrics on crepant resolutions of Kähler cones, Math. Ann. 347 (2010), 581–611.
- [28] Craig van Coevering, Regularity of asymptotically conical ricci-flat kähler metrics, 2010.
- [29] D. Wu, Kähler-Einstein metrics of negative Ricci curvature on general quasi-projective manifolds, Comm. Anal. Geom. 16 (2008), 395–435.
- [30] Q. Yao, Mass and expansion of asymptotically conical Kähler metrics, preprint, arXiv:2206.06505 (2022).
- [31] S.-T. Yau, On the Ricci curvature of a compact Kähler manifold and the complex Monge-Ampére equation, I, Comm. Pure Appl. Math. 31 (1978), 339–411.