1  

THE MONGE-AMPÈRE OPERATOR AND GEODESICS IN THE SPACE OF KÄHLER POTENTIALS

D.H. Phong * and Jacob Sturm †

November 27, 2006

1  

1 Introduction

Let X   be an n   -dimensional compact complex manifold, L X   a positive holomorphic line bundle, and   the space of positively curved hermitian metrics on L   . The purpose of this article is to prove that geodesics in the infinite-dimensional symmetric space   can be uniformly approximated by geodesics in the finite-dimensional symmetric spaces k = G L ( N k + 1 , C ) / U ( N k + 1 , C )   , where N k + 1 = dim ( H 0 ( L k ) )   . Thus the k   are becoming flat as k   .
The motivation for this work comes from Donaldson's far reaching program [9, 10relating the geometry of   to the existence and uniqueness of constant scalar curvature Kähler metrics. As advocated by S.T. Yau over the years,   should be approximated by k   , and the properties of this approximation should be closely reflected in many basic questions of Kähler geometry. In particular, the condition of “stability” is one which concerns the growth of energy functionals along the geodesics of k   . On the other hand, the existence and uniqueness of metrics of constant scalar curvature concerns the growth of energy functionals along the geodesics of   (see [11, 12, 24, 25, 26). Thus a good understanding of the relationship between these different types of geodesics is desirable.
In order to state the precise theorem, we need some notation: Let h : L [ 0 , )   be a smooth hermitian metric. If s L   we write h ( s ) = | s | h   and for k > 0   , we denote by h k   the induced metric on L k   . The curvature of h   is the ( 1 , 1 )   form on X   defined locally by the formula R ( h ) = 1 2 ¯ log | s ( z ) | h 2   , where s ( z )   is a local, nowhere vanishing holomorphic section. In particular, R ( h k ) = k R ( h )   . Let = { h : L [ 0 , ) : h i s s m o o t h m e t r i c o n L s u c h t h a t R ( h ) > 0 } .   If we fix h 0   and let ω 0 = R ( h 0 )   we have a natural isomorphism = { φ C ( M ) : ω φ = ω 0 + 1 2 ¯ φ > 0 } ,   where φ   is identified with h = h 0 e φ   so that R ( h ) = ω φ   . Then   is an infinite dimensional manifold whose tangent space T φ   at φ   is naturally identified with C ( M )   .
Now let φ   and ψ C ( M )   and define a metric on   by
| | ψ | | φ 2 = X | ψ | 2 ω φ n . (1.1)
Donaldson [10, Mabuchi [18and Semmes [23have shown that ( 1.1 ) defines a Riemannian metric which makes   into an infinite dimensional negatively curved symmetric space.
Furthermore, the geodesics of   in this metric are the paths φ t   which satisfy the partial differential equation
φ ¨ | φ | ω φ 2 = 0 . (1.2)
The space   contains a canonical family of finite-dimensional negatively curved symmetric spaces k   which are defined as follows: For k > > 0   and for s ̲ = ( s 0 , . . . . , s N k )   an ordered basis of H 0 ( L k )   , let ι s ̲ : X P N   be the Kodaira embedding given by z ( s 0 ( z ) , . . . , s N k ( z ) )   . Then we have a canonical isomorphism L k = ι s ̲ * O ( 1 )   . Let h F S   is the Fubini-Study metric on O ( 1 ) P N   and let
h s ̲ = ( ι s ̲ * h F S ) 1 / k = h 0 ( j = 0 N k | s j | h 0 k 2 ) 1 / k (1.3)
Note that the right side of ( 1.3 ) is independent of the choice of h 0   . In particular
j = 0 N k | s j | h s ̲ k 2 = 1 (1.4)
Let k = { h s ̲ : s ̲ a b a s i s o f H 0 ( L k ) }   Then k = G L ( N k + 1 ) / U ( N k + 1 )   is a finite-dimensional negatively curved symmetric space sitting inside of   . It is well known that the k   are topologically dense in   :
If h   then there exists h ( k ) k   such that h ( k ) h   in the C   topology. This follows from the Tian-Yau-Zelditch theorem on the density of states (see also Catlin [4for corresponding results for the Bergman kernel). In fact, if h   , then there is a canonical choice of the approximating sequence h ( k )   : Let s ̲   be a basis of H 0 ( L k )   which is orthonormal with respect to the metrics h   . In other words,
X ( s i , s j ) h k ω n = δ i j w h e r e ω = R ( h ) . (1.5)
Note that s ̲   is not unique: if s ̲   is orthonormal, so is u s ̲   , for any u U ( N k + 1 )   . Define ρ k ( h ) = ρ k ( ω ) = j | s j | h k 2   . Then Theorem 1 of [27says that for h   fixed, we have a C   asymptotic expansion as k   :
ρ k ( ω ) k n + A 1 ( ω ) k n 1 + A 2 ( ω ) k n 1 + (1.6)
where the A j ( ω )   are smooth functions on X   defined locally by ω   . Let s ̲ ^ = k n / 2 s ̲   and h ( k ) = h s ̲ ^   . In particular, ( 1.3 ) and ( 1.6 ) imply that for each r > 0  
| | h ( k ) h 1 | | = O ( 1 k 2 ) , | | ω ( k ) ω | | = O ( 1 k 2 ) , | | φ ( k ) φ | | = O ( 1 k 2 ) (1.7)
where the norms are all taken with respect to C r ( ω 0 )   . Here, as before, ω = R ( h )   , ω ( k ) = R ( h ( k ) )   , h = h 0 e φ   and h ( k ) = h 0 e φ ( k )   .
Now let h 0 , h 1   . It is known by the work of Chen [5(see also more recent progress in Donaldson [13and Chen-Tian [6) that there is a unique C 1 , 1   geodesic h : [ 0 , 1 ]   joining h 0   to h 1   which, according to a conjecture of Donaldson [10, should be smooth.
Our main theorem says that h t = h 0 e φ ( t )   is a uniform limit of the smooth geodesics in k = G L ( N k + 1 ) / U ( N k + 1 )   which join h 0 ( k ) = h s ̲ ^ ( 0 )   to h 1 ( k ) = h s ̲ ^ ( 1 )   .
To be precise, let σ G L ( N k + 1 )   be the change of basis matrix defined by σ s ̲ ^ ( 0 ) = s ̲ ^ ( 1 )   .
Without loss of generality, we may assume that σ   is diagonal with entries e λ 0 , . . . , e λ N   for some λ j R   . Let s ̲ ^ ( t ) = σ t s ̲ ^ ( 0 )   where σ t   is diagonal with entries e λ j t   . Define
h ( t ; k ) = h s ̲ ^ ( t ) = h 0 e φ ( t ; k ) . (1.8)
Then h ( t ; k )   is the smooth geodesic in G L ( N k + 1 ) / U ( N k + 1 )   joining h 0 ( k )   to h 1 ( k )   .
Explicitly, using ( 1.3 ), we can also write
φ ( t ; k ) = 1 k log ( j = 0 N e 2 λ j t | s ^ j ( 0 ) | h 0 k 2 ) . (1.9)
Theorem 1 Let h 0 , h 1   and h t = h 0 e φ t   the unique C 1 , 1   geodesic joining h 0   to h 1   .
Then
φ t = lim l [ sup k l φ ( t ; k ) ] * u n i f o r m l y a s l , (1.10)
where, for any bounded function u : X × [ 0 , 1 ] R   , we define the upper envelope u *   of u   by u * ( ζ 0 ) = lim ε 0 sup | ζ ζ 0 | < ε u ( ζ )   .
Remark 1. The proof will show that
φ t = lim l [ sup k l φ ( t ; k ) ] a l m o s t e v e r y w h e r e . (1.11)
Remark 2. We note that the upper envelope u *   is independent of the choice of coordinate systems defining the balls | ζ ζ 0 | < ε   . It is the smallest upper semicontinuous function which is greater than or equal to u   . If u k   is a sequence of plurisubharmonic functions which are locally uniformly bounded, then [ s u p u k ] *   is plurisubharmonic and equal to s u p u k   almost everywhere. Similarly, we also define for later use the lower envelope u *   of a bounded function u   by u * ( ζ 0 ) = lim ε 0 inf | ζ ζ 0 | < ε u ( ζ )   . The function u *   is the largest lower semi-continuous function which is less than or equal to u   .
To prove the theorem we first apply an observation of Donaldson [10which shows that solving the geodesic equation on   is equivalent to solving the degenerate Monge-Ampère equation
Ω Φ n + 1 = 0 (1.12)
on the manifold M ¯ = X × A   , where A   is the annulus A = { w C : 1 | w | e }   , the values of Φ   are prescribed on the boundary of M ¯   by smooth rotationally symmetric data, and Ω Φ | X × { t }   is positive for every t   . Here we are writing Ω Φ = Ω 0 + 1 2 ¯ Φ   where Ω 0 = π 1 * ω 0   and Φ   is a smooth on M ¯   . We then use the Tian-Yau-Zelditch theorem to prove that the k   geodesics are, in a certain sense, approximate solutions to ( 1.12 ). A key step is then proving that the limit of the sequence of approximate solutions is in fact a weak solution of the Monge-Ampère equation and that the weak solution thus obtained is unique. This is accomplished using the methods of pluripotential theory which were introduced and systematically developed by Bedford-Taylor in their fundamental work [1and [2. In recent years, this subject has been extended by Demailly [8, Klimek [15, Blocki [3, and others. The uniqueness implies that the limit coincides with the C 1 , 1   solution φ t   . Now the uniqueness theorem for bounded open sets in C n   , which is due to Bedford-Taylor, requires smoothing techniques for rough solutions which do not generalize in a simple way to manifolds. It is quite possible that the regularization methods of Demailly, which have recently been successfully applied by Guedj and Zeriahi [14in the setting of compact manifolds, will also apply in our setting. However, we use a different approach which exploits the particular structure of the weak solutions in our case. A particularly important ingredient for us is the existence of a vector field Y   , transversal to the boundary of M ¯   , with
| Y ( Φ k ) | C , (1.13)
where Φ k   is a sequence of approximate solutions to the Dirichlet problem for the Monge-Ampère equation (c.f. Theorem 3 and the proof of Theorem 6 below). For the application to Theorem 1, this hypothesis is a consequence of the a priori estimate
| φ ˙ ( t , k ) | C , (1.14)
which is the one global estimate for the derivatives of φ ( t ; k )   that we can actually obtain.
One approach to Donaldson's conjecture on the existence of smooth geodesics is the establishment of a priori estimates on the smooth approximate solutions φ ( t ; k )   . As we just noted, the uniform C 0   estimates for φ ( t ; k )   and for φ ˙ ( t ; k )   do hold. But extending these to higher derivative estimates seems rather hard, and at the same time, quite intriguing:
For example, the estimate which is needed for φ ¨ ( t ; k )   can be carried out in certain special cases, and appears to be related to the central limit theorem in probability theory. In the last section, we shall make some remarks along these lines.

2 The volume estimate

In this section, we establish the basic properties of the functions φ ( t ; k )   introduced in ( 1.9 ).
The properties of the functions φ ( t ; k )   themselves are summarized in Lemma 1 below, while the properties of their corresponding Monge-Ampere measures are given in Theorem 2.   A first key ingredient is the following very general volume formula. Let ( X , ω 0 )   be a compact Kähler manifold of dimension n   , A = { w C : 1 | w | e }   , M ¯ = X × A   , π 1 : M ¯ X   be the projection on the first factor, and Ω 0 = π 1 * ω 0   . Then Ω 0   is a closed positive ( 1 , 1 )   -form on M ¯   such that Ω 0 n + 1 = 0   .
Let φ : [ 0 , 1 ] φ ( t ) = { φ C ( X ) : ω 0 + 1 2 ¯ φ }   be a smooth path joining φ ( 0 )   to φ ( 1 )   and define Φ : M ¯ R   by
Φ ( z , w ) = φ ( t ) ( z ) w h e r e t = log | w | (2.1)
and the corresponding ( 1 , 1 )   -form Ω Φ   by
Ω Φ = Ω 0 + 1 2 ¯ Φ . (2.2)
In local coordinates z i   for X   and v = log w   for A   , if we identify a Kähler form ω 0 = 1 2 i j g j ¯ i 0 d z i d z ¯ j   with the hermitian matrix { g j ¯ i 0 }   , then we can write Ω Φ = ( g j ¯ i 0 + i j ¯ φ 1 2 B 1 2 B * 1 4 φ ¨ )   where B   is the column vector B = ( φ ˙ z ¯ 1 φ ˙ z ¯ 2 φ ˙ z ¯ n )   and B *   is the conjugate transpose of B   . It follows that
Ω Φ n + 1 = 1 4 ( φ ¨ | φ ˙ | ω φ 2 ) ω φ n ( 1 2 d v d v ¯ ) , (2.3)
and the condition Ω Φ n + 1 = 0   is equivalent to the geodesic equation φ ¨ | φ ˙ | ω φ 2 = 0   . Now
X × A Ω Φ n + 1 = 0 1 X ( φ ¨ | φ ˙ | 2 ) ω φ n d t = 0 1 d d t ( X φ ˙ ω φ n ) d t . (2.4)
and we obtain the desired volume formula:
X × A Ω Φ n + 1 = X φ ˙ ( 1 ) ω φ ( 1 ) n X φ ˙ ( 0 ) ω φ ( 0 ) n . (2.5)
Note that ( 2.5 ) is only a true volume if Ω Φ 0   , but that ( 2.5 ) is valid even in the absence of this hypothesis.   It is perhaps noteworthy that 1 V X φ ˙ ω φ n = F ˙ ω φ 0 0   , where V = X ω 0 n   , and F ω φ 0 0 ( φ )   is precisely the functional whose critical points (with the constraint 1 V X e h 0 φ ω 0 n = 1   , and h   is a fixed function satisfying R i c ( ω 0 ) μ ω 0 = 1 2 ¯ h   ), give Kähler-Einstein metrics.
Thus the preceding relation can be rewritten as
1 V X × A Ω Φ n + 1 = 0 1 F ¨ ω φ 0 0 d t = F ˙ ω φ 0 0 ( φ ( 1 ) ) F ˙ ω φ 0 0 ( φ ( 1 ) ) . (2.6)
The asymptotic behavior of F ω φ 0 0   along geodesics in k   is known to be closely related to Chow-Mumford stability (see [28, and also [19, [20,[22). The simple formula ( 2.6 ) raises the intriguing possibility that the behavior of the Monge-Ampère operator Ω Φ n + 1   and geodesics in   can be linked even more directly to stability and constant scalar curvature Kähler metrics.
  Fix now two potentials h 0   and h 1   in   , and let s ̲ ( 0 )   , s ̲ ( 1 )   be bases for H 0 ( L k )   which are orthonormal with respect to the metrics induced by h 0   and h 1   . As in the Introduction, let σ G L ( N k + 1 )   be the corresponding change of bases, σ s ̲ ( 0 ) = s ̲ ( 1 )   . We may assume that σ   is diagonal with eigenvalues e λ 0 , , e λ N k   . We consider the functions φ ( t ; k )   defined as in ( 1.9 ).
Lemma 1 Let Ω 0 = π 1 * ω 0   where ω 0 = R ( h 0 )   .
(a) For each k   , let Φ ( k )   be the extension of φ ( t ; k )   to M ¯ = X × A   as in ( 2.1 ), and let Ω Φ ( k )   is the corresponding ( 1 , 1 )   -form as in ( 2.2 ). Then Ω Φ ( k )   is a smooth positive ( 1 , 1 )   -form,
Ω Φ ( k ) 0 . (2.7)
In particular, the ( n + 1 , n + 1 )   -form Ω Φ ( k ) n + 1   form is a positive smooth measure on M ¯   ; (b) There is a constant C > 0   which does not depend on k   such that
C 1 k m a x 0 j N | λ j | C k ; (2.8)
(c) With the same constant C > 0   , we also have
| φ ( t ; k ) | + | φ ˙ ( t ; k ) | 4 C . (2.9)
Proof. Locally on M ¯   , we have e 2 λ j t = | w λ j | 2   , where w λ j = e λ j ( t + i s ) A   is a holomorphic function. Since the logarithms of sums of squares of absolute values of holomorphic functions are plurisubharmonic in the usual sense on C n + 1   , and since we can write
Ω Φ ( k ) = Ω 0 + 1 2 ¯ ( 1 k log j = 0 N k e 2 λ j t | s j ( z ) | h 0 k 2 )
= 1 2 k ¯ log j = 0 N k | w λ j s j ( z ) | 2 , (2.10)
where | s j ( z ) |   is just the absolute value of s ^ j ( z )   in a local trivialization, the positivity of Ω Φ ( k )   follows. Here we have simplified the notation by denoting s j ( 0 )   just by s j   .
We turn next to the proof of (b). Let φ = log h 1 h 0   , and order the eigenvalues λ j   so that λ 0 λ 1 λ N   . Clearly,
2 k λ N 1 k log j = 0 N k e 2 λ j | s j ( z ) | 2 j = 0 N k | s j ( z ) | 2 2 k λ 0 . (2.11)
By the Tian-Yau-Zelditch theorem, the expression in the middle tends to φ   as k   .
Thus, if we set C 1 = s u p φ   and C 2 = i n f φ   , we have λ 0 C 1 k / 4   and λ N C 2 k / 4   , for k   large enough.
To get inequalities in the opposite direction, we note that for each j   ,
| s j | h 0 k 2 ω 0 n n ! = 1 , 1 N k l = 0 N | s l | h 0 k 2 ω 0 n n ! = 1 . (2.12)
This implies that for some x X   , we have | s j | 2 1 N k l = 0 N | s l | 2   . Choosing j = 0   , we obtain for k   large,
1 k log e λ 0 N k 1 k log e λ 0 | s 0 ( x ) | 2 | s ̲ ( x ) | 2 1 k log e λ l | s l ( x ) | 2 | s l ( x ) | 2 2 C 1 . (2.13)
This shows that λ 0 2 C 1 k + log N k   . But N k k n   . Thus we conclude
C 1 2 k λ 0 3 C 1 k , k > > 1 . (2.14)
To get the bound on λ N k   , interchange the roles of h 0   and h 1   , which changes φ   to φ   and σ   to σ 1   . Thus we have
C 2 2 k λ N k 3 C 2 k , (2.15)
and (b) is proved.
The first inequality in (c), namely that | φ ( t ; k ) |   is uniformly bounded, follows immediately from the bounds in (b) for | λ j |   . To get the second inequality, we write
φ ˙ ( t ; k ) = 1 k j = 0 N k 2 λ j e 2 λ j t | s j ( z ) | 2 j = 0 N k e 2 λ j t | s j ( z ) | 2 (2.16)
and hence
| φ ˙ ( t ; k ) | 2 m a x | λ j | k . (2.17)
Applying (b) again, we find that | φ ˙ |   is uniformly bounded. Q.E.D.   We can now state and prove the following theorem, which is a key step in the construction of a generalized solution of the Monge-Ampère equation Ω Φ n + 1 = 0   :
Theorem 2 The volume of X × A   approaches zero as k   tends to infinity. More precisely,
0 X × A Ω Φ ( k ) n + 1 C k (2.18)
where C   is a constant which is independent of k   . In particular, the positive measures Ω Φ ( k ) n + 1   tend weakly to 0.
Proof. Recall that φ ˙ ( t ; k )   has been evaluated and is given by ( 2.16 ). In particular, it is independent of the choice of metric on L k   . We now apply ( 2.5 ) to the path of Fubini-Study metrics φ ( t ; k )   defined by ( 1.9 ) and obtain
X × A Ω Φ ( k ) n + 1 = 2 k X j = 0 N k λ j | s j ( 1 ) | 2 j = 0 N k | s j ( 1 ) | 2 ω φ ( 1 ) n 2 k X j = 0 N k λ j | s j ( 0 ) | 2 j = 0 N k | s j ( 0 ) | 2 ω φ ( 0 ) n (2.19)
Here Φ ( k )   is defined by ( 2.1 ) with φ ( t )   replaced by φ ( t ; k )   . Since h 1 ( k ) = h s ̲ ^ ( 1 )   the formula ( 1.4 ) implies k n j = 0 N k | s j ( 1 ) | h 1 k ( k ) 2 = j = 0 N k | s ^ j ( 1 ) | h 1 k ( k ) 2 = 1   and, observing that ω φ ( 1 ) = ω 1 ( k )   and ω φ ( 0 ) = ω 0 ( k )   , we can rewrite ( 2.19 ) as
X × A Ω Φ ( k ) n + 1 = 2 k n + 1 X j = 0 N k λ j | s j ( 1 ) | h 1 k ( k ) 2 ω 1 n ( k ) 2 k n + 1 X j = 0 N k λ j | s j ( 0 ) | h 0 k ( k ) 2 ω 0 n ( k ) . (2.20)
Now observe that X | s j ( 1 ) | h 1 k ( k ) 2 ω 1 n ( k ) = X | s j ( 1 ) | h 1 k 2 ω 1 n h 1 k ( k ) h 1 k ω 1 n ( k ) ω 1 n .   On the other hand, ( 1.7 ) implies h 1 k ( k ) h 1 k = 1 + O ( 1 k )   and ω 1 n ( k ) ω 1 n = 1 + O ( 1 k 2 )   . Moreover, X | s j ( 1 ) | h 1 k 2 ω 1 n = 1   since the s j ( 1 )   are orthonormal with respect to h 1 k   . Thus X | s j ( 1 ) | h 1 k ( k ) 2 ω 1 n ( k ) = 1 + O ( 1 k )   and hence
2 k n + 1 X j = 0 N k λ j | s j ( 1 ) | h 1 k ( k ) 2 ω 1 n ( k ) 2 k n + 1 j = 0 N k λ j = O ( 1 k n + 2 ) N k max 0 j N k | λ j | (2.21)
Now the Riemann-Roch theorem implies N k = O ( k n )   and thus we obtain from Lemma  1 that the right side of ( 2.21 ) is of the size O ( 1 k )   . Theorem  2 follows now from ( 2.20 ) and ( 2.21 ). Q.E.D.

3 Generalized solutions of the Monge-Ampère equation

In the previous section, we have seen that the functions Φ ( k )   form a uniformly bounded sequence of functions whose Monge-Ampère operators Ω Φ ( k ) n + 1   are positive measures on M ¯   which tend to 0. The Chern-Levine-Nirenberg inequality [7implies that if any subsequence of the Φ ( k )   's converges uniformly, then its limit Φ   would satisfy the Monge-Ampère equation Ω Φ n + 1 = 0   in the generalized sense. A major problem is the fact that the bounds available to us at the present time (c.f. Lemma 1) are not strong enough to guarantee the existence of a uniformly convergent subsequence of the Φ ( k )   's. Of course, weakly convergent subsequences can always be found. However, it is well-known that the Monge-Ampère operator P S H C 0 Φ Ω Φ n + 1   is not lower semi-continuous under weak limits [16.
To circumvent these difficulties, we shall formulate and establish extensions of the classical convergence and uniqueness theorems of Bedford-Taylor for the Monge-Ampère operator for domains in C n   to the case of Kähler manifolds with boundary.

3.1 Convergence of approximate solutions

The first of these extensions is the following convergence theorem, where the key hypothesis is the existence of uniform C 0   bounds for some transversal derivative of the Φ ( k )   's at the boundary of the manifold M ¯   .Let M ¯   be a compact complex manifold with smooth boundary and M M ¯   be the interior of M ¯   . Let M ¯ = α = 1 N U α   be a covering of M ¯   by a finite number of coordinate charts U α   .
Fix a smooth closed ( 1 , 1 )   -form Ω 0   on M   , and let Ψ α   be smooth potentials for Ω 0   on U α   , that is, Ω 0 = 1 2 ¯ Ψ α   on U α   . We define the class of Ω 0   -plurisubharmonic functions on M   to be the class of functions Φ   on M ¯   by
P S H ( M , Ω 0 ) = { Φ ; Ψ α + Φ i s p l u r i s u b h a r m o n i c o n U α , 1 α N } . (3.1)
Note that this condition means that Ψ α + Φ   is upper semi-continuous and satisfies the sub-mean value property. However, it is a stronger condition than the condition Ω 0 + 1 2 ¯ Φ 0   by itself, since it is a pointwise condition, while the condition Ω 0 + 1 2 ¯ Φ 0   depends only on the values of Φ   almost everywhere.
Next, let φ   be a continuous function on M ¯   , and Ω 0   a real, smooth closed ( 1 , 1 )   form on M   . Set d = + ¯   , d c = 1 4 ( ¯ )   , so that d d c = 1 2 ¯   .
Definition 1 Let Φ : M ¯ R   be an upper-semicontinuous function. We say Φ   is a solution of the Dirichlet problem with boundary values φ   if
  • 1. Φ P S H ( M , Ω 0 )  
  • 2. Φ   is continuous at p   for all p M ¯   and Φ | M ¯ = φ   .
  • 3. ( Ω 0 + d d c Φ ) m = 0   on M ¯   where m = dim ( M )   .
We wish to obtain a solution of the Dirichlet problem on M ¯   from a sequence of approximate solutions. Let Y   be a smooth real nowhere vanishing vector field on a neighborhood U M ¯   of M ¯   which is transversal to M ¯   , in the sense that for p M ¯   , the vector Y ( p )   is not tangent to M ¯   .
Theorem 3 Assume Ω 0 m = 0   . Let Φ k P S H ( M , Ω 0 ) C ( M ¯ )   have the properties:
  • 1. | Φ k |   is uniformly bounded on M ¯   .
  • 2. | Y ( Φ k ) |   is uniformly bounded on U   ;
  • 3. Φ k | M ¯ φ   uniformly;
  • 4. There is a sequence a k > sup x M ¯ | Φ k ( x ) φ ( x ) |   with the property a k 0   (i.e. a k   is decreasing and approaching zero) and k a k <   .
  • 5. lim k M ( Ω 0 + d d c Φ k ) m = 0  
Let Φ = lim k [ sup l k Φ l ] *   Then Φ   is a solution of the Dirichlet problem with boundary values φ   .
Remark: Assumption 4. implies Assumption 3. On the other hand, after passing to a subsequence, Assumption 3. implies Assumption 4.Proof. We want to choose a sequence c k 0   with the property in such a way that Φ k + c k   is monotonically decreasing on M   . To do this, we just define c k = 2 j k a j   . Then c k c k + 1 = 2 a k   . Moreover, if x M   , then
Φ k ( x ) Φ k + 1 ( x ) > ( φ a k ) ( φ + a k + 1 ) 2 a k = c k + 1 c k (3.2)
Thus, replacing Φ k   by Φ k + c k   , we may assume that Φ k | Ω > Φ k + 1 | Ω   .
Let W k = [ sup l k Φ l ] *   Then W k P S H ( M , Ω 0 )   by Theorem 5.7 of [8. Moreover, ( 3.2 ) implies that W k = Φ k   in an open neighborhood of M ¯   . Thus
M ( Ω 0 + d d c W k ) m = M ( Ω 0 + d d c Φ k ) m (3.3)
This follows from the following simple lemma:
Lemma 2 Let W , Φ P S H ( M , Ω 0 )   with Ω 0 m = 0   and assume that W = Φ   on some neighborhood of M ¯   . Then M ( Ω 0 + d d c W ) m = M ( Ω 0 + d d c Φ ) m  
Proof of Lemma  2 . To see this, let K   be any compact subset such that W = Φ   on M \ K   , and let Ψ C 0 ( M )   be a function with Ψ = 1   on a neighborhood of K   . Then, expanding ( Ω 0 + d d c W ) m   , we obtain M Ψ ( Ω 0 + d d c W ) m = M Ψ Ω 0 m + M Ψ d ( Θ W ) = M \ K d Ψ Θ W   Here Θ W = k = 0 m 1 ( m k ) Ω 0 k d c W ( d d c W ) m 1 k   . Since Θ W = Θ Φ   on M \ K   , the lemma is proved.
Now assumption 5. of the theorem together with ( 3.3 ) implies that ( Ω 0 + d d c W k ) m 0   weakly. On the other hand, since W k Φ   monotonically, the Bedford-Taylor monotonicity theorem (Theorem 2.1 of [2; this is stated for domains in C n   , but generalizes in a straightforward fashion to manifolds) we have ( Ω 0 + d d c W k ) m ( Ω 0 + d d c Φ ) m   weakly. Thus we have ( Ω 0 + d d c Φ ) m = 0   . To finish the proof of the theorem, we must show that Φ   is continuous at the boundary and has the right boundary values.
Let ε > 0   . Choose k 0   such that k k 0 sup x M ¯ | Φ k ( x ) φ ( x ) | < ε   . Extend φ   to a continuous function on a neighborhood U M ¯   of M ¯   in such a way that φ   is constant on the flow lines of Y   . Then assumption 2. implies that if U   is sufficiently small, then
sup x U | Φ k ( x ) φ ( x ) | < 2 ε (3.4)
and thus sup x U | Φ ( x ) φ ( x ) | 2 ε   . In particular, Φ | Ω = φ   and Φ   is continuous at all points p M ¯   . This proves the theorem. Q.E.D.

3.2 The domination principle for the Monge-Ampère operator

The proof of the extension of the Bedford-Taylor uniqueness theorem in this section follows closely the original arguments of [1, 2, and especially the exposition of Blocki [3.
When we consider generalized solutions of a partial differential equation, a particularly desirable property is their uniqueness. For bounded domains in C n   , the uniqueness of the generalized solution of the Dirichlet problem for the Monge-Ampère equation in the class P S H L   has been established by Bedford and Taylor [1. It seems that this uniqueness theorem should extend as well to bounded domains in Kähler manifolds, at least if good smooth approximations of Ω Φ ( 0 )   -plurisubharmonic functions exist. Although there are now many powerful approximation theorems (see [8, 14and references therein), we found it more convenient to extend the Bedford-Taylor uniqueness theorem to a situation adapted to the problem at hand, in the spirit of the earlier extension. The key hypothesis which we will exploit is a capacity zero condition.Recall the following notion of capacity of a set introduced by Bedford and Taylor [2: If E U   is a Borel subset of a bounded domain U C n   then
c ( E , U ) = s u p { E ( d d c v ) n ; v P S H ( U ) , 0 v 1 } . (3.5)
For our purposes, we can adapt this notion to Kähler manifolds as follows: Let M = α = 1 N U α   be a finite cover of M   by coordinate neighborhoods. Then we say c ( E , M ) < ε   if we can write E = α E α   with E α U α   a Borel subset and
α c ( E α , U α ) < ε . (3.6)
We say that c ( E , M ) = 0   if c ( E , M ) < ε   for every ε > 0   .
Lemma 3 There is a constant C > 0   with the following property: If E M   is a Borel subset, ε > 0   and Φ P S H ( Ω 0 , M )   , then
E ( Ω 0 + d d c Φ ) m C ε ( 1 + sup | Φ | ) m (3.7)
if c ( E , M ) < ε   . In particular, if c ( E , M ) = 0   , then for all functions Φ P S H ( Ω 0 , M )   , we have
E ( Ω 0 + d d c Φ ) m = 0 . (3.8)
Proof. Fix a smooth potential Ψ α   on U α   such that Ω 0 = d d c Ψ α   . Then
E ( Ω 0 + d d c Φ ) m α E α ( Ω 0 + d d c Φ ) m = α E α ( d d c ( Ψ α + Φ ) ) m
α ( sup | Ψ α + Φ | ) m c ( E α , U α ) . (3.9)
Since the Ψ α   are fixed, we have | Ψ α + Φ | C ( 1 + | Φ | )   and the lemma follows.
The following lemma follows immediately from the quasi-continuity theorem of Bedford-Taylor [2:
Lemma 4 Let Φ P S H ( M , Ω 0 )   . Then for every ε > 0   , there is an open set G M   such that c ( G , M ) < ε   and Φ   is continuous on M \ G   .
Proof. Since Ψ α + Φ   is a plurisubharmonic function on U α   , the quasi-continuity theorem of Bedford-Taylor ([2, Theorem 3.5) implies that there is an open set G α U α   such that Ψ α + Φ   is continuous on U α \ G α   and c ( G α , U α ) < ε   . Let G = α G α   . Then Φ   is continuous on M \ G   and, by definition, c ( G , M ) < N ε   .
We also require a notion of “nearly continuous” functions:
Definition 2 We say that a bounded function v : M ¯ R   is “nearly continuous” if
  • 1. There exists a lower semi-continuous function v 0   on M ¯   such that v = v 0 *   ;
  • 2. { v 0 < v }   has capacity zero, that is, c ( { v 0 < v } , M ) = 0   ;
  • 3. v = v 0   on M   .
With this notion, we shall prove the following:
Theorem 4 Assume that M ¯   is a complex manifold of dimension m   with smooth boundary, let M = M ¯ \ M   and let Ω 0   be a real closed smooth ( 1 , 1 )   form on M   satisfying Ω 0 m = 0   .
Let u , v P S H ( M , Ω 0 ) L   be such that ( u v ) * 0   on M ¯   . Assume as well:
  • 1. u   is continuous;
  • 2. There is a decreasing sequence v k   of nearly continuous functions in P S H ( M , Ω 0 )   such that v k v   ;
  • 3. For every δ > 0   there is a compact set K M   such that v k | M \ K < v | M \ K + δ   for k   sufficiently large.
Then
u < v ( Ω 0 + d d c v ) m u < v ( Ω 0 + d d c u ) m . (3.10)
Proof. We divide the proof into several steps.
Step 1. If we replace u   by u + δ   , then we have { u + δ < v } { u < v }   as δ 0   . Since for any positive measure μ   we have μ ( E j ) μ ( E j )   whenever E j   is an increasing family of measurable sets, we may replace u   by u + δ   . Thus we may assume that
( u v ) * δ a n d t h a t M = { u < v } i s r e l a t i v e l y c o m p a c t i n M . (3.11)
Step 2. We prove the theorem under the assumption that v   is continuous (in which case the hypotheses 2. and 3. are automatic, since we can take v k = v   for all k   ).
For ε > 0   let u ε = max ( u + ε , v )   . Then u ε = u + ε   on a neighborhood of M   . Since d d c ( u + ε ) = d d c u   , we can invoke Lemma 2 and conclude that
M ( Ω 0 + d d c u ε ) m = M ( Ω 0 + d d c u ) m . (3.12)
On the other hand, u ε v   on M   and, by the Bedford-Taylor monotonicity theorem, ( Ω 0 + d d c u ε ) m ( Ω 0 + d d c v ) m o n M ( w e a k c o n v e r g e n c e o f m e a s u r e ) .   Since { u < v }   is open, we obtain { u < v } ( Ω 0 + d d c v ) m liminf ε 0 { u < v } ( Ω 0 + d d c u ε ) m .   This completes step 2.
Step 3. Now we treat the case where v   itself is nearly continuous (in which case the hypotheses 2. and 3. are again automatic, since we can take v k = v   for all k   ). This step will be parallel to the argument in Step 2:
For ε > 0   let u ε = max ( u + ε , v )   . Then u ε v   on the open set { u < v 0 }   (this set is open since v 0   is lower semi-continuous). In particular, the Bedford-Taylor monotonicity theorem implies ( Ω 0 + d d c u ε ) m ( Ω 0 + d d c v ) m   weakly on { u < v 0 }   (as measures or as currents the two notions of weak convergence are equivalent). Thus we have { u < v } ( Ω 0 + d d c v ) m = { u < v 0 } ( Ω 0 + d d c v ) m liminf ε 0 { u < v 0 } ( Ω 0 + d d c u ε ) m   = liminf ε 0 { u < v } ( Ω 0 + d d c u ε ) m   The first equality follows from the assumption that { v 0 < v }   has capacity zero and the inequality from the fact that ( Ω 0 + d d c u ε ) m ( Ω 0 + d d c v ) m   weakly. Here we are making strong use of the fact that { u < v 0 }   is open.
Next we claim that for all ε > 0   :
{ u < v } ( Ω 0 + d d c u ε ) m = { u < v } ( Ω 0 + d d c u ) m . (3.13)
To see this, let A   be an open set containing { u < v }   , which is relatively compact in M   .
Then, p A   implies that p / A   , and we have u ( p ) v ( p )   , so u ε ( p ) = u ( p ) + ε > v ( p )   .
Since the set { u + ε > v }   is open ( u   is continuous and v   is upper semi-continuous), we see that u + ε = u ε   in a neighborhood of A   . Thus Lemma  2 implies A ( Ω 0 + d d c u ε ) m = A ( Ω 0 + d d c ( u + ε ) ) m = A ( Ω 0 + d d c u ) m   Since ( Ω 0 + d d c u ε ) n   and ( Ω 0 + d d c u ) n   are positive Borel measures which are finite on compact subsets of M   , they are both regular. In particular
{ u < v } ( Ω 0 + d d c u ε ) m = inf A A ( Ω 0 + d d c u ε ) m
{ u < v } ( Ω 0 + d d c u ) m = inf A A ( Ω 0 + d d c u ) m (3.14)
where the inf is taken over all open sets A   which contain { u < v }   . This proves ( 3.13 ).
Step 4. We treat the general case. From step 1. we can assume that ( u v ) * δ > 0   .
Thus u > v k   on a neighborhood of M   for k   sufficiently large. By Lemma  4 , there is an open set G M   such that c ( G , M ) < ε   (the capacity of G   in M   ) and such that v   is continuous on F = M \ G   . Choose φ   which is continuous M   such that φ = v   on F   . Then { u < v } { u < φ } G   implies
{ u < v } ( Ω 0 + d d c v ) m { u < φ } ( Ω 0 + d d c v ) m + G ( Ω 0 + d d c v ) m
{ u < φ } ( Ω 0 + d d c v ) m + C ε (3.15)
where the last inequality follows from Lemma  3 , the fact that c ( G , M ) < ε   and that the function v   is bounded. Here the constant C   depends only on the sup norm of v   . Šince { u < φ }   is open and v k v   , the Bedford-Taylor monotonicity theorem implies { u < φ } ( Ω 0 + d d c v ) m liminf k { u < φ } ( Ω 0 + d d c v k ) m   Now { u < φ } { u < v } G { u < v k } G   Thus { u < v } ( Ω 0 + d d c v ) m liminf k { u < v k } ( Ω 0 + d d c v k ) m + 2 C ε ,   since our assumptions imply that the functions v k   are uniformly bounded. Next, the assumption 3. implies that the sets { u < v k }   are all contained in a relatively compact subset of M   . Using Step 3, { u < v k } ( Ω 0 + d d c v k ) m { u < v k } ( Ω 0 + d d c u ) m   and since v k v   , k { u < v k } = { u v }   , we can conclude that { u < v } ( Ω 0 + d d c v ) m { u v } ( Ω 0 + d d c u ) m + 2 C ε   Since ε   is arbitrary:
{ u < v } ( Ω 0 + d d c v ) m { u v } ( Ω 0 + d d c u ) m   Applying this last inequality to u + η   and v   , for some η > 0   :
{ u + η < v } ( Ω 0 + d d c v ) m { u + η v } ( Ω 0 + d d c u ) m   Finally, taking the limit as η 0   and noting that η > 0 { u + η < v } = η > 0 { u + η v } = { u < v }   , we obtain ( 3.16 ).
Next we prove another version of the domination theorem, but this time with the roles of u , v   reversed:
Theorem 5 Assume that M ¯   is a complex manifold of dimension m   with smooth boundary, let M = M ¯ \ M   and let Ω 0   be a real closed smooth ( 1 , 1 )   form on M   with the property:
Ω 0 m = 0   . Let u , v P S H ( M ¯ , Ω 0 ) L   satisfy ( u v ) * 0   on M ¯   . Assume as well:
  • 1. v   is continuous;
  • 2. There is a decreasing sequence u k   of nearly continuous PSH functions on M   such that u k u   .
Then
u < v ( Ω 0 + d d c v ) m u < v ( Ω 0 + d d c u ) m . (3.16)
Proof. The proof is parallel to that of Theorem  4 , although there are some important differences. We again divide it into several steps. Štep 1. As before, we may assume that ( u v ) * δ a n d t h a t M = { u < v } i s r e l a t i v e l y c o m p a c t i n M   Moreover, we have { u 0 < v } i s r e l a t i v e l y c o m p a c t i n M   To see this, observe first that u v + δ / 2   in an open neighborhood of M   . Also, { u 0 u > δ / 4 }   is open since u 0   is lower semi-continuous and u   is upper semi-continuous.
Since u 0 = u   on M   , the set { u 0 u > δ / 4 }   is an open neighborhood of M   . Thus u 0 > v + δ / 4   in an open neighborhood of M   .
Step 2. Now we treat the case where u   itself is nearly continuous (in which case hypotheses 2. and 3. are automatic, since we can take u k = u   for all k   ).
For ε > 0   let u ε = max ( u + ε , v )   . Then u ε v   on the open set { u < v }   (this set is open since u   is upper semi-continuous). As before, the Bedford-Taylor monotonicity theorem implies ( Ω 0 + d d c u ε ) m ( Ω 0 + d d c v ) m   weakly and, making use of the fact that { u < v }   is open, we get
{ u < v } ( Ω 0 + d d c v ) m liminf ε 0 { u < v } ( Ω 0 + d d c u ε ) m
= liminf ε 0 { u 0 < v } ( Ω 0 + d d c u ε ) m , (3.17)
where u 0   is the function which appears in the definition of the near continuity of u   . Next we claim that for all ε > 0   :
{ u 0 < v } ( Ω 0 + d d c u ε ) m = { u 0 < v } ( Ω 0 + d d c u ) m (3.18)
The proof is similar to that of ( 3.13 ), using A   an open set containing { u 0 < v }   , u 0 ( p ) v ( p )   for p A   so that u 0 ( p ) + ε > v ( p )   . We use now the continuity of v   and the lower semi-continuity of u 0   to deduce that the set { u 0 + ε > v }   is open, and u + ε = u ε   in a neighborhood of A   . As before, Lemma 2 implies
A ( Ω 0 + d d c u ε ) m = A ( Ω 0 + d d c ( u + ε ) ) m = A ( Ω 0 + d d c u ) m , (3.19)
and, using the fact that ( Ω 0 + d d c u ε ) m   and ( Ω 0 + d d c u ) m   are positive Borel measures which are finite on compact subsets of M   ,
{ u 0 < v } ( Ω 0 + d d c u ε ) m = inf A A ( Ω 0 + d d c u ε ) m
{ u 0 < v } ( Ω 0 + d d c u ) m = inf A A ( Ω 0 + d d c u ) m (3.20)
where the inf is taken over all open sets A   which contain { u 0 < v }   . This proves ( 3.18 ).
Since u   and u 0   differ only on a set of capacity 0, we obtain the desired inequality.
Step 4. We treat the general case: From step 1. we can assume that ( u v ) * δ > 0   . Thus u j > v   on a neighborhood of M   for j   sufficiently large. Choose an open set G M   such that c ( G , M ) < ε   (the capacity of G   in M   ) and such that u   is continuous on F = M \ G   .
Choose φ   which is continuous M   such that φ = u   on F   . Now
{ u < v } ( Ω 0 + d d c v ) m = lim j { u j < v } ( Ω 0 + d d c v ) m lim j { u j < v } ( Ω 0 + d d c u j ) m
liminf j { u < v } ( Ω 0 + d d c u j ) m (3.21)
where we have made use of step 3. and the fact that u j > v   on M   to prove the first inequality. Now we have { u < v } { φ < v } G ( K { φ v } G   where K M   is a compact set such that { u < v } K   so
{ u < v } ( Ω 0 + d d c u j ) m { φ v } K ( Ω 0 + d d c u j ) m + G ( Ω 0 + d d c u j ) m
{ φ v } K ( Ω 0 + d d c u j ) m + C ε (3.22)
where the last inequality follows from the fact that c ( G , M ) < ε   and that the function v   is bounded. Here the constant C   depends only on the sup norm of v   .
Since { φ v } K   is compact, the Bedford-Taylor monotonicity theorem implies limsup k { φ v } K ( Ω 0 + d d c u j ) m { φ v } K ( Ω 0 + d d c u ) m   Now { φ v } K { φ v } { u v } G   Thus { u < v } ( Ω 0 + d d c v ) m { u v } ( Ω 0 + d d c u ) m + 2 C ε   Since ε   is arbitrary:
{ u < v } ( Ω 0 + d d c v ) m { u v } ( Ω 0 + d d c u ) m   Applying this last inequality to u + η   and v   , for some η > 0   :
{ u + η < v } ( Ω 0 + d d c v ) m { u + η v } ( Ω 0 + d d c u ) m   Finally, taking the limit as η 0   , we obtain ( 3.16 ).
We can now state and prove the uniqueness theorem which we need for the proof of Theorem 1:
Theorem 6 Let u , v   be as in Theorem  4 , that is,
  • 1. u   is continuous;
  • 2. There is a decreasing sequence v k   of nearly continuous PSH functions on M   such that v k v   ;
  • 3. For every δ > 0   there is a compact set K M   such that v k | M \ K < v | M \ K + δ   for k   sufficiently large.
Assume that ( u v ) * = ( u v ) * = 0   on M   and that ( Ω 0 + d d c u ) m = ( Ω 0 + d d c v ) m = 0   .
Then u = v   .
Proof. We may assume, after replacing u   and v   by u + C   and v + C   for some constant C   , that u   and v   are positive. We wish to show u = v   . Assume not: Let ψ C ( M )   be such that Ω 0 + d d c ψ > 0   . Replacing ψ   by ψ C   , we may assume that ψ < 0   on M   .
Case 1. { u < v }   . This implies { u < ( 1 ε ) v + ε ψ }   for some ε > 0   . Let p { u < ( 1 ε ) v + ε ψ }   and let D   be a disk in some coordinate neighborhood of p   . Then D { u < ( 1 ε ) v + ε ψ }   has non-zero Lebesgue measure (in general, if u , v   are psh functions such that u = v   almost everywhere in a disk D   , then u = v   everywhere in D   ; this follows from local regularization). Now we have, using Theorem  4 
0 = { u < ( 1 ε ) v + ε ψ } ( Ω 0 + d d c u ) m { u < ( 1 ε ) v + ε ψ } ( Ω 0 + d d c [ ( 1 ε ) v + ε ψ ] ) m
ε m { u < ( 1 ε ) v + ε ψ } ( Ω 0 + d d c ψ ) m > 0 (3.23)
Case 2. { v < u }   . This is treated exactly in the same was as in case 1 except that we use Theorem  5 instead of Theorem  4 . Q.E.D.

4 Proof of Theorem 1

We can give now the proof of Theorem 1.First, we apply Theorem 3 to construct a generalized solution of the Dirichlet problem for the Monge-Ampère equation
Ω Φ n + 1 = 0 o n M = X × A , Φ | M = φ , (4.1)
where φ : M R   is defined by φ | | w | = 1 = 0   and φ | | w | = e = log h 0 h 1   . Define Φ ( k ) ( z , w ) = φ ( t ; k )   , where t = log | w |   and φ ( t ; k )   is defined as in ( 1.9 ). Let Y = t   . Then | Φ ( k ) | C   and | Y ( Φ ( k ) ) | C   by Lemma 1. We also have | | Φ ( k ) | M ¯ φ | | L C 1 k 2   by ( 1.7 ), so that Φ ( k ) M ¯ φ   uniformly. Furthermore, Theorem 2 implies that M Ω Φ ( k ) n + 1 0   as k   .
Thus, since k = 1 1 k 2 <   , we can apply Theorem 3, and conclude that
Φ = l i m k [ s u p l k Φ ( l ) ] * (4.2)
is a generalized solution of the desired Dirichlet problem ( 4.1 ).
Consider next the C 1 , 1   geodesic φ t   joining φ 0   to φ 1   in the space   of Kähler potentials.
Let Φ ~ ( z , w ) = φ t ( z )   with t = log | w |   as before. We shall show that Φ = Φ ~   . To do this, we would like to apply Theorem 6 with
u = Φ ~ , ( v k ) 0 = s u p l k Φ ( l ) , v k = ( v k ) 0 * , v = Φ = lim k v k (4.3)
First, we show that v k   is nearly continuous. Since Φ ( l )   is smooth, ( v k ) 0   is lower semi-continuous.
Moreover, { ( v k ) 0 < v k }   has capacity 0, by Proposition 5.1 of [2. It remains to show that v k = ( v k ) 0   on M ¯   . The equation ( 1.7 ) implies | D Φ ( l ) ( z , w ) | C   , for some constant C   independent of l   , if ( z , w ) M ¯   and D   is any derivative tangent to M ¯   . Thus, if δ > 0   , there exists ε > 0   such that if ( z 0 , w 0 ) M ¯   then
Φ ( l ) ( z 1 , w 0 ) + δ > Φ ( l ) ( z 0 , w 0 ) > Φ ( l ) ( z 1 , w 0 ) δ , (4.4)
for all l   if | z 1 z 0 | < ε   . Also, | Y ( Φ ( l ) ) | C   implies that
Φ ( l ) ( z 1 , w 1 ) + δ > Φ ( l ) ( z 1 , w 0 ) > Φ ( l ) ( z 1 , w 1 ) δ i f | w 0 w 1 | < ε . (4.5)
We have then Φ ( l ) ( z 1 , w 1 ) + 2 δ > Φ ( l ) ( z 0 , w 0 ) > Φ ( l ) ( z 1 , w 1 ) 2 δ   , which implies that
( v k ) 0 ( z 1 , w 1 ) + 2 δ > ( v k ) 0 ( z 0 , w 0 ) > ( v k ) 0 ( z 1 , w 1 ) 2 δ , (4.6)
and hence
( v k ) 0 ( z 1 , w 1 ) + 2 δ > ( v k ) 0 ( z 0 , w 0 ) > v k ( z 1 , w 1 ) 2 δ , (4.7)
if | z 1 z 0 | < ε   , | w 1 w 0 | < ε   . In particular, ( v k ) 0 ( z 0 , w 0 ) > v k ( z 0 , w 0 ) 2 δ   for all δ   so ( v k ) 0 = v k   on M   so v k   is indeed nearly continuous.
Thus the first two assumptions of Theorem  6 are satisfied.
In the next step, we will need the bound
v ( z 1 , w 1 ) + 2 δ > v ( z 0 , w 0 ) > v ( z 1 , w 1 ) 2 δ , (4.8)
which follows by taking the limit of ( 4.7 ) as k   .
Next, we verify assumption 3: Since v k = ( v k ) 0   on M   we see that v k   is both upper and lower semi-continuous on M   and thus v k   is continuous on M   . Moreover, by Theorem 3, v = φ   on M   so v   is also continuous on M   . Since v k v   we see, by Dini's theorem, that v k v   uniformly on M   . Thus, for every δ > 0   ( 4.7 ) and ( 4.8 ) imply that for k > > 0   v k ( z 1 , w 1 ) 2 δ < v k ( z 0 , w 0 ) < v ( z 0 , w 0 ) + δ < v ( z 1 , w 1 ) + 3 δ   for all | z 1 z 0 | < ε   and | w 1 w 0 | < ε   .
All the conditions of Theorem  6 are satisfied. We can thus conclude that Φ = Φ ~   .
Finally, the uniform convergence of the functions [ s u p k l φ ( t , k ) ] *   follows from their upper semi-continuity and the compactness of X   . This is essentially Dini's theorem, and can be proven as follows. Assume that u n   is a sequence of upper semi-continuous functions, decreasing to a continuous limit u   . For each ε > 0   , the sets { x X ; u n ( x ) u ( x ) < ε }   form an open covering of X   . Since X   is compact, it admits a finite subcover, and since the sets are increasing as n   increases, we must have X = n N ε { x X ; u n ( x ) u ( x ) < ε }   for some N ε   . Q.E.D.

5 Remarks

We conclude with a few remarks.   Donaldson's conjecture says any two Kähler metrics can be connected by a smooth geodesic. One way to prove this conjecture is to establish a priori bounds on the derivatives of φ ( t ; k )   . This was done in Lemma  1 for φ   and φ ˙   . Let us now consider φ ¨   :
φ ¨ ( 0 ) = 1 k ( λ α λ β ) 2 | s i | h 0 k ( k ) 2 | s j | h 0 k ( k ) 2 (5.1)
Note that Lemma  1 implies | φ ¨ ( 0 ) | C k   , but this is not strong enough, since we need a bound which is independent of k   .
Define a random variable Z   whose probability distribution is given by P ( Z = λ α ) = | s α ( z ) | h 0 k ( k ) 2   This is indeed a distribution since the total probability is α | s α ( z ) | h 0 k ( k ) 2 = 1   . Moreover, ( λ α λ β ) 2 | s i | h 0 k ( k ) 2 | s j | h 0 k ( k ) 2   is just the variance of Z   . Thus we need to prove that the variance of Z   is bounded by k   . In the simplest case where X = P 1   and the line bundle L = O ( 1 )   and the metric h   is the Fubini-Study metric, then an easy computation shows that Z   is just the binomial distribution with k   trials, where the probability p   of flipping heads is a function of z P 1   . As is well known, the variance of the binomial distribution is k p   , which is the bound we need.
In the case where ω   is the Fubini-Study metric on P 1   , the eigenvalues of the change of basis matrix are just 0 , 1 , . . . , k   . More generally, we can show that if φ   is a radially symmetric Kähler potential on P 1   , and if λ 0 λ 1 λ k   are the eigenvalues of the change of basis matrix, then | λ j λ j + 1 | C   for some constant C   , independent of k   . From this one can show without difficulty that φ ¨ ( 0 )   is uniformly bounded.
  The function l i m l [ s u p k l φ ( t ; k ) ] * ( x )   is equal almost everywhere to the l i m s u p φ ( t ; k )   .
The convergence can also be guaranteed to take place in a Sobolev norm of positive order. Indeed, quite generally, the L 2   -norm of φ   can be bounded by | | φ | | C 0   if φ   is P S H ( X , ω 0 )   -plurisubharmonic. Indeed,
| | φ | | ω 0 2 = X φ ¯ φ ω 0 n 1 = X φ ¯ φ ω 0 n 1 = X φ ω φ ω 0 n 1 X φ ω 0 n (5.2)
and the right hand side can be bounded in turn by
| X φ ω φ ω 0 n 1 X φ ω 0 n | | | φ | | C 0 ( X ω φ ω 0 n 1 + X ω 0 n ) = 2 | | φ | | C 0 X ω 0 n . (5.3)
(In fact, the same argument gives the following useful inequality
J ω 0 ( φ ) 2 n | | φ | | C 0 , (5.4)
where J ω 0 ( φ ) = V 1 1 i = 0 n 1 n i n + 1 X φ ¯ φ ω φ i ω 0 n i 1   , V = X ω 0 n   , is the familiar Aubin-Yau functional.) Returning to the problem at hand, we deduce that the H ( 1 ) ( X × [ 0 , 1 ] )   Sobolev norms of the functions φ ( t ; k )   are uniformly bounded. The same is true for the H ( 1 ) ( X )   Sobolev norms of φ ( t ; k )   for each t [ 0 , 1 ]   .   The functions φ ˙ ( t ; k )   also satisfy an interesting Harnack inequality of Li-Yau type. Let 0 τ < T 1   and let ξ , X M   . Then we claim
φ ˙ ( ξ , τ ) φ ˙ ( X , T ) + 1 8 Δ ( ξ , τ , X , T ) (5.5)
where
Δ ( ξ , τ , X , T ) = inf γ τ T ( d s d t ) 2 d t (5.6)
where d s d t   is the velocity in space at time t   and the infimum is taken over all paths from ( ξ , τ )   to ( X , T )   parametrized by τ t T   . To see this, let L = 2 φ ˙   . Then φ ¨ | φ ˙ | ω φ 2 = | π N V | 2 0   , where π N V   is the normal component of the holomorphic vector field on C P N k   generated by σ t   [21, and so
L t | D L | φ 2 . (5.7)
As in [17, we can now choose a path ( t , s ( t ) )   joining ( ξ , τ )   to ( X , T )   where τ < T   and ξ , X M   . Then
L ( X , T ) L ( ξ , τ ) = τ T d L d t d t = τ T { L t + L s d s d t } d t
τ T { | D L | 2 + L s d s d t } d t . (5.8)
Completing the square, we obtain
L ( X , T ) L ( ξ , τ ) 1 4 ( d s d t ) 2 d t . (5.9)
References

  1. Bedford, E. and B.A. Taylor, “The Dirichlet problem for a complex Monge-Ampre equation”, Invent. Math. 37 (1976), 1-44.
  2. Bedford, E. and B.A. Taylor, “A new capacity for plurisubharmonic functions”, Acta Math. 149 (1982), 1-40.
  3. Blocki, Z., “The complex Monge-Ampère operator and pluripotential theory”, lecture notes available from the author's website.
  4. Catlin, D., “The Bergman kernel and a theorem of Tian”, Analysis and geometry in several complex variables (Katata, 1997), 1-23, Trends Math., Birkhäuser Boston, Boston, MA, 1999.
  5. Chen, X.X., “The space of Kähler metrics”, J. Differential Geom. 56 (2000), 189-234.
  6. Chen, X.X. and G. Tian, “Geometry of Kähler metrics and foliations by discs”, arXiv: math.DG / 0409433.
  7. Chern, S.S., H. Levine, and L. Nirenberg, “Intrinsic norms on a complex manifold”, Global Analysis, Papers in honor of K. Kodaira, University of Tokyo Press (1969) 119-139.
  8. Demailly, J.P., “Complex analytic and differential geometry”, book available from the author's website.
  9. Donaldson, S.K., “Remarks on gauge theory, complex geometry and 4   -manifold topology”, Fields Medallists' lectures, 384-403, World Sci. Ser. 20th Century Math., 5, World Sci. Publishing, River Edge, NJ, 1997.
  10. Donaldson, S.K., “Symmetric spaces, Kähler geometry, and Hamiltonian dynamics”, Amer. Math. Soc. Transl. 196 (1999) 13-33.
  11. Donaldson, S.K., “Scalar curvature and projective imbeddings I”, J. Differential Geom. 59 (2001) 479-522.
  12. Donaldson, S.K., “Scalar curvature and stability of toric varieties”, J. Differential Geom. 62 (2002), 289-349.
  13. Donaldson, S.K., “Scalar curvature and projective imbeddings II”, arXiv: math.DG / 0407534.
  14. Guedj, V. and A. Zeriahi, “Intrinsic capacities on compact Kähler manifolds”, arXiv: math.CV / 0401302.
  15. Klimek, M., “Pluripotential theory”, London Mathematical Society monographs, New Series 6 (1991) Oxford University Press, New York.
  16. Lelong, P., “Fonctions plurisousharmoniques et formes différentielles positives”, Gordon & Breach, Paris-London-New York (1968).
  17. Li, P. and S.T. Yau, “On the parabolic kernel of the Schrödinger operator”, Acta Math. 156 (1986), 153-201.
  18. Mabuchi, T., “Some symplectic geometry on compact Kähler manifolds”, Osaka J. Math. 24 (1987) 227-252.
  19. Paul, S., “Geometric analysis of Chow Mumford stability”, Adv. Math. 182 (2004), 333-356.
  20. Phong, D.H. and J. Sturm, “Stability, energy functionals, and Kähler-Einstein metrics”, Comm. Anal. Geometry 11 (2003) 563-597, arXiv: math.DG / 0203254.
  21. Phong, D.H. and J. Sturm, “Scalar curvature, moment maps, and the Deligne pairing”, Amer. J. Math. 126 (2004) 693-712, arXiv: math.DG / 0209098.
  22. Phong, D.H. and J. Sturm, “The Futaki invariant and the Mabuchi energy of a complete intersection”, Comm. Anal. Geometry 12 (2004) 321-343, arXiv: math.DG / 0312529.
  23. Semmes, S., “Complex Monge-Ampère and symplectic manifolds”, Amer. J. Math. 114 (1992) 495-550.
  24. Tian, G., “Kähler-Einstein metrics with positive scalar curvature”, Invent. Math. 130 (1997) 1-37.
  25. Yau, S.T., “On the Ricci curvature of a compact Kähler manifold and the complex Monge-Ampère equation I”, Comm. Pure Appl. Math. 31 (1978) 339-411.
  26. Yau, S.T., “Open problems in geometry”, Proc. Symp. Pure Math. 54, AMS Providence, RI (1993) 1-28.
  27. Zelditch, S., “The Szegö kernel and a theorem of Tian”, Int. Math. Res. Notices 6 (1998) 317-331.
  28. Zhang, S., “Heights and reductions of semi-stable varieties”, Compositio Math. 104 (1996) 77-105.

1   Research supported in part by National Science Foundation grants DMS-02-45371 and DMS-01-00410

1   *   Department of Mathematics Columbia University, New York, NY 10027   Department of Mathematics Rutgers University, Newark, NJ 07102