<<

arXiv:1905.10222v1 [math.DG] 24 May 2019 -qaini endas defined is J-equation nti ae,ormi oli opoeteeuvlneo h solvab the K¨ahler of metrics equivalence Given the stability. prove of to notion is a goal and main J-equation our paper, this In Introduction 1 oetmpitrrtto 2] twstefis perneo the of appearance first the w literature. was the It direction in [21]. J-equat this of interpretation conn in study map Hermitian-Yang-Mills the moment results of including questions study first many Herm the proposed the on by Donaldson 39] Inspired of [20, One connections. Donaldson-Uhlenbeck-Yau Mills of work geometry. celebrated in the common very is -tblt srdcdt osToa’ nfr lp -tblt [31]. K-stability slope uniform Ross-Thomas’s cones to normal reduced to is r “degeneration When K-stability called conjec substitution. configurations folklore correct test a a is be special evidenc There may K-stability is [1]. of case there version cscK However, uniform Chen-Do in wrong by K¨ahler-Einstein case. be proved Fano may gene been conjecture in was has 11] K¨ahler It (cscK conjecture 10, curvature This [9, [37]. scalar K¨ah condition [22]. constant Fano case K-stability the in projective the to precise called it Donaldson made was by Tian it ized and [41]. stability case of Einstein kind some to equivalent for ngnrl h qiaec ftesaiiyadteslaiiyo neq an of solvability the and stability the of equivalence the general, In a ojcue htteeitneo aoK¨ahler-Einstein met Fano of existence the that conjectured Yau aemto,w lopoeasmlrrsl o h eomdH ( deformed in the is for angle result the similar when a equation prove Yang-Mills also we method, same aycntn clrcraueKalrmtiswith K¨ahler metrics curvature scalar constant many n nyi ( if only and hr exists there nti ae,w rv htfrayK¨ahler metrics any for that prove we paper, this In ω M, ϕ = [ ω 0 ω ] 0 , [ + χ )i nfrl -tbe sacrlay ecnfind can we corollary, a As J-stable. uniformly is ]) √ ω nJ-equation On − ϕ 1 = ∂ a 7 2019 27, May ϕ> ∂ϕ ¯ ω a Chen Gao 0 tr Abstract + ω aifigteJeuto tr J-equation the satisfying 0 ϕ √ χ 1 − = 1 ∂ c ϕ> ∂ϕ ¯ nπ 2 − 0 π 4 . , nπ 2 ). c 1 ω < ω 0 0 and .Uigthe Using 0. and ω ,teuniform the ”, ermitian- ϕ χ rbe in problem ) ueta the that ture o sn the using ion χ χ naldson-Sun on itian-Yang- srce to estricted htthis that e = J-equation lt fthe of ility on i salso is ric M c ections, M if uation , More the , ral- ler- as recently, the projective assumption was removed by the work of Dervan-Ross [19] and independently by Sj¨ostr¨om Dyrefelt [33]. It is easy to see that cscK metrics are critical points of the K-energy func- tional [5] n n ωϕ ωϕ K(ϕ)= log( ) + Ric(ω )(ϕ). ωn n! J− 0 ZM 0 The functional for any (1,1)-form χ is defined by Jχ n 1 n − 1 k n 1 k 1 k n k χ(ϕ)= ϕ χ ω0 ωϕ− − c0ϕ ω0 ωϕ− , J n! M ∧ ∧ − (n + 1)! M ∧ Z Xk=0 Z Xk=0 where c0 is the constant given by

n 1 n ω − ω χ 0 c 0 =0. ∧ (n 1)! − 0 n! ZM − When χ is a K¨ahler form, it is well known that the critical point of the χ func- tional is exactly the solution to the J-equation. It was the second appearanceJ of the J-equation in the literature. Following this formula, using the interpola- tion of the K-energy and the χ functional, Chen-Cheng [6, 7, 8] proved that the existence of cscK metric isJ equivalent to the geodesic stability of K-energy. However, the relationship between the existence of cscK metrics and the uniform K-stability is still open. When we replace the K-energy by the χ functional for a K¨ahler form χ, the analogy of the K-stability and the slopeJ stability conditions were proposed by Lejmi and Sz´ekelyhidi [28]. See also Section 6 of [19] for the extension to non-projective case. The main theorem of this paper proves the equivalence between the existence of the critical point of functional, the solvability of Jχ J-equation, the coerciveness of χ functional, and the uniform J-stability as well as the uniform slope J-stability.J Theorem 1.1. (Main Theorem) Fix a K¨ahler manifold M n with K¨ahler metrics χ and ω0. Let c0 > 0 be the constant such that

n 1 n ω − ω χ 0 = c 0 , ∧ (n 1)! 0 n! ZM − ZM then the followings are equivalent: (1) There exists a unique smooth function ϕ up to a constant such that ω = ω + √ 1∂∂ϕ¯ > 0 satisfies the J-equation ϕ 0 −

trωϕ χ = c0; (2) There exists a unique smooth function ϕ up to a constant such that ω = ω + √ 1∂∂ϕ¯ > 0 satisfies the J-equation ϕ 0 − n 1 n ω − ω χ ϕ = c ϕ ; ∧ (n 1)! 0 n! −

2 (3) There exists a unique smooth function ϕ up to a constant such that ϕ is the critical point of the functional; Jχ (4) The χ functional is coercive, in other words, there exist ǫ1.1 > 0 and another constantJ C such that (ϕ) ǫ (ϕ) C ; 1.2 Jχ ≥ 1.1Jω0 − 1.2 (5) (M, [ω0], [χ]) is uniformly J-stable, in other words, there exists ǫ1.1 > 0 such that for all K¨ahler test configurations ( , Ω) defined as Definition 2.10 of X [19], the numerical invariant J[χ]( , Ω) defined as Definition 6.3 of [19] satisfies J ( , Ω) ǫ J ( , Ω); X [χ] X ≥ 1.1 [ω0] X (6) (M, [ω0], [χ]) is uniformly slope J-stable, in other words, there exists ǫ1.1 > 0 such that for any subvariety V of M, the degeneration to normal cone ( , Ω) defined as Example 2.11 (ii) of [19] satisfies J ( , Ω) ǫ J ( , Ω); X [χ] X ≥ 1.1 [ω0] X (7) There exists ǫ1.1 > 0 such that

p p 1 (c (n p)ǫ )ω pχ ω − 0 0 − − 1.1 0 − ∧ 0 ≥ ZV for all p-dimensional subvarieties V with p =1, 2, ..., n. Remark 1.2. Lejmi and Sz´ekelyhidi’s original conjecture is that the solvability of

trωϕ χ = c0 is equivalent to p p 1 c ω pχ ω − > 0 0 0 − ∧ 0 ZV for all p-dimensional subvarieties V with p = 1, 2, ..., n 1 [28]. However, it seems that our uniform version is more natural from geometri− c point of view.

Remark 1.3. It is well known that there exists a constant C1.3 depending on n such that the functional Jω0 1 n 1 n 1 n 1 ω − ω tω − ( ϕ(ω tϕ n tϕ ))dt = (√ 1 ∂ϕ ∂ϕ¯ tϕ )dt 0 ∧ (n 1)! − n! − ∧ ∧ (n 1)! Z0 ZM − Z0 ZM − and Aubin’s I-functional

n n n ¯ k n k ϕ(ω0 ωϕ)= √ 1 ∂ϕ ∂ϕ ω0 ωϕ− M − − M ∧ ∧ ∧ Z Z kX=0 satisfy 1 n n n n C− ϕ(ω ω ) (ϕ) C ϕ(ω ω ). 1.3 0 − ϕ ≤ Jω0 ≤ 1.3 0 − ϕ ZM ZM For example, Collins and Sz´ekelyhidi used this fact and their Definition 20 in n n [13] replaced ω0 (ϕ) by M ϕ(ω0 ωϕ) in the definition of the coerciveness which was called “properness”J in [13].− By (3) of [2], Aubin’s I-functional can also be replaced by Aubin’s J-functionalR in the definition of coerciveness. Accordingly, in the definition of uniform stability, the numerical invariant J ( , Ω) can be [ω0] X

3 replaced by the minimum norm of ( , Ω) defined as Definition 2.18 of [19]. By X (62) of [16], Aubin’s J-functional can be further replaced by the d1 distance in the definition of the coerciveness when ϕ is normalized such that the Aubin-Mabuchi energy of ϕ is 0. Remark 1.4. By Proposition 2 of [5], if the solution to the J-equation exists, it is unique up to a constant. It is easy to see that (1) and (2) are equivalent. The equivalence between (2) and (3) follows from the formula n 1 n d ∂ϕ ω − ω Jχ = (χ ϕ c ϕ ). dt ∂t ∧ (n 1)! − 0 n! ZM − By Proposition 21 and Proposition 22 of [13] and Remark 1.3, (1) and (4) are equivalent. By Corollary 6.5 of [19], (4) implies (5). It is trivial that (5) implies (6). By [28], (6) implies (7) in the projective case if ǫ1.1 is replaced by 0. However, it is easy to see that it is also true in non-projective case and for positive ǫ1.1. Thus, we only need to prove that (7) implies (1) in Theorem 1.1. Remark that there is a simpler proof that (1) implies (7). Let χ = δij and ωϕ = λiδij , then for any c> 0, the condition p p 1 cω pχ ω − 0 ϕ − ∧ ϕ ≥ is equivalent to p 1 c λ ≤ j=1 ij X for all distinct p numbers i ,i , ..., i 1, 2, ..., n . So tr χ = c as well as { 1 2 p}⊂{ } ωϕ 0 the upper bounds of λi imply that for small enough ǫ1.1 > 0, p p 1 (c (n p)ǫ )ω pχ ω − 0 0 − − 1.1 ϕ − ∧ ϕ ≥ for all p =1, 2, ..., n. (4) follows from the fact that

p p 1 p p 1 (c (n p)ǫ )ω pχ ω − = (c (n p)ǫ )ω pχ ω − . 0 − − 1.1 0 − ∧ 0 0 − − 1.1 ϕ − ∧ ϕ ZV ZV

When c1(M) < 0, we can choose χ as a K¨ahler form in c1(M). Since x log x −n n R ωϕ ωϕ is bounded from below for any x , the entropy M log( ωn ) n! is also bounded ∈ 0 from below. So the coerciveness of functional implies the coerciveness of K- χ R energy. This observation appearedJ as Remark 2 of [5]. Using this observation, as a corollary of Theorem 1.3 of [7] and Theorem 1.1, we can find many cscK metrics with c1(M) < 0.

Corollary 1.5. If c1(M) < 0, and ǫ1.1 > 0, then for any K¨ahler class [ω0] such that n 1 n[c1(M)] [ω0] − p p 1 ((− · (n p)ǫ )ω pω − ( c (M)) 0 [ω ]n − − 1.1 0 − 0 ∧ − 1 ≥ ZV 0 for all p-dimensional subvarieties V with p = 1, 2, ..., n, there exists a cscK metric in [ω0].

4 Remark 1.6. If there exists ωϕ [ω0] such that Ric(ωϕ) < 0 and ωϕ has constant scalar curvature, then the∈ condition above is also necessary. Besides the appearances in the moment map picture and the study of the cscK problem, J-equation also arises from the study of mirror symmetry. In fact, using the following observation of Collins-Jacob-Yau [12]

n n π 1 lim k( arctan(kλi)) = , k 2 − λ →∞ i=1 i=1 i X X the J-equation is exactly the limit of the deformed Hermitian-Yang-Mills equa- tion n arctan λi = constant, i=1 X where λi are the eigenvalues of ωϕ with respect to χ. It plays an important role in the study of mirror symmetry [35, 29]. Motivated by the J-equation, Collins-Jacob-Yau [12] conjectured that the solvability of the deformed Hermitian-Yang-Mills equation is also equivalent to a notion of stability. In this paper, we prove the uniform version of their conjecture when the angle is in ( nπ π , nπ ): 2 − 4 2 n Theorem 1.7. Fix a K¨ahler manifold M with K¨ahler metrics χ and ω0. Let ˆ nπ π nπ θ ( 2 4 , 2 ) be a constant. Then there exists a unique smooth function ϕ up∈ to a constant− such that n arctan λi = θˆ i=1 X for eigenvalues λ of ω = ω + √ 1∂∂ϕ¯ > 0 with respect to χ if and only if i ϕ 0 − there exists a constant ǫ1.1 > 0 and for all p-dimensional subvarieties V with p =1, 2, ..., n (V can be chosen as M), there exist smooth functions θV (t) from (n p)π pπ [1, ) to [θˆ − + (n p)ǫ , ) such that for all t [1, ), ∞ − 2 − 1.1 2 ∈ ∞

p p pπ (χ + √ 1tω0) =0,θV (t) = arg( (χ + √ 1tω0) ), lim θV (t)= . − 6 − t 2 ZV ZV →∞

Moreover, when V = M, it is required that θM (1) = θˆ.

Remark 1.8. We only study the case when θˆ ( nπ π , nπ ) in this paper. So it ∈ 2 − 4 2 is natural to assume that [ω0] is a K¨ahler class. However, usually we need extra conditions in addition to [ω ] being K¨ahler to make sure (χ + √ 1tω )p is 0 V − 0 not 0 so that θV (t) is well defined. When p =1, 2, θV (t) is always well defined and increasing for t ( , ) without any extra assumption.R In addition, ∈ −∞ ∞ θV (0) = 0. When p =3,

(χ + √ 1tω )3 = χ3 3t2 χ ω2 + √ 1t(3 χ2 ω t2 ω3). − 0 − ∧ 0 − ∧ 0 − 0 ZV ZV ZV ZV ZV

5 So if the inequality

( χ3)( ω3) < 9( χ ω2)( χ2 ω ) 0 ∧ 0 ∧ 0 ZV ZV ZV ZV in Proposition 3.3 of [14] holds, then θV (t) is well defined for t ( , ). Moreover, if θ (1) > π, then θ (t) is increasing for t [1, ).∈ In−∞ addition,∞ V V ∈ ∞ θV (0) = 0. So the choice of θV (1) in this paper is the same as the choice of θV (1) in Proposition 8.4 of [15]. In higher dimensions, more inequalities are involved. ˆ (n p)π Remark 1.9. Collins-Jacob-Yau conjectured that θV (1) > θ −2 for all p = 1, 2, ..., n 1 is equivalent to the solvability of the deformed− Hermitian- Yang-Mills equation− [12]. However, it seems that our uniform version is more natural because Definition 8.10 (2) of [15] also assumed the uniform positive lower bound. Remark 1.10. By Theorem 1.1 of [26], the solution to the deformed Hermitian- Yang-Mills equation is unique up to a constant if it exists. The “only if” part of Theorem 1.7 is a combination of Proposition 3.1 of [12] and Remark 1.4. So we only need to prove the “if” part of Theorem 1.7. Theorem 1.7 will be proved in Section 5 using the same method of the proof of Theorem 1.1. Instead of Theorem 1.1, we will prove the following stronger statement by induction:

n Theorem 1.11. Fix a K¨ahler manifold M with K¨ahler metrics χ and ω0. Let 1 1 n 1 c> 0 be a constant and f > ( ) − be a smooth function satisfying − 2n c n n n 1 χ ω ω − f = c 0 χ 0 0, n! n! − ∧ (n 1)! ≥ ZM ZM ZM − then there exists ω = ω + √ 1∂∂ϕ¯ > 0 satisfying the equation ϕ 0 − χn χ f c trωϕ + n = ωϕ and the inequality n 1 n 2 cω − (n 1)χ ω − > 0 ϕ − − ∧ ϕ if there exists ǫ1.1 > 0 such that

p p 1 (c (n p)ǫ )ω pω − χ 0 − − 1.1 0 − 0 ∧ ≥ ZV for all p-dimensional subvarieties V with p =1, 2, ..., n. Remark 1.12. By Remark 1.4, Theorem 1.1 is a corollary of Theorem 1.11 by choosing f =0.

6 Remark 1.13. When n = 1, Theorem 1.11 is trivial. When n = 2, Theorem 1.11 is the statement that the Demailly-Paun’s characterization [18] for [cω0 χ] being K¨ahler implies the solvability of the

(cω χ)2 = (cf + 1)χ2 ϕ − by Yau [40]. In the toric case when f is a non-negative constant, Theorem 1.11 was proved by Collins and Sz´ekelyhidi [13]. There are several steps to prove Theorem 1.11. Step 1: Prove the following:

n Theorem 1.14. Fix a K¨ahler manifold M with K¨ahler metrics χ and ω0. Let 1 1 n 1 c> 0 be a constant and f > ( ) − be a smooth function satisfying − 2n c n n n 1 χ ω ω − f = c 0 χ 0 0, n! n! − ∧ (n 1)! ≥ ZM ZM ZM − then there exists ω = ω + √ 1∂∂ϕ¯ satisfying the equation ϕ 0 − χn χ f c trωϕ + n = ωϕ and the inequality n 1 n 2 cω − (n 1)χ ω − > 0 ϕ − − ∧ ϕ if n 1 n 2 cω − (n 1)χ ω − > 0. 0 − − ∧ 0

We will use the continuity method to prove Theorem 1.14. The details will be provided in Section 2.

Remark 1.15. Let χ = δij and ωϕ = λiδij , then the equation χn χ f c trωϕ + n = ωϕ is equivalent to n 1 f + = c. λ n λ i=1 i i=1 i X 1 Q Remark 1.16. Suppose i=k λ c for all k =1, 2, ..., n and 6 i ≤ P n 1 f + = c, λ n λ i=1 i i=1 i X 1 1 n 1 Q 1 then as long as f > 2n ( c ) − , it is easy to see that i=k λ

n Theorem 1.18. Fix a K¨ahler manifold M with K¨ahler metrics χ and ω0. Suppose that for all t > 0, there exist a constant ct > 0 and a smooth K¨ahler form ω [(1 + t)ω ] satisfying t ∈ 0 n 1 n 2 cω − (n 1)χ ω − > 0 t − − ∧ t and χn trωt χ + ct n = c. ωt Then there exist a constant ǫ > 0 and a current ω [ω ǫ χ] such that 1.4 1.5 ∈ 0 − 1.4 n 1 n 2 cω − (n 1)χ ω − 0 1.5 − − ∧ 1.5 ≥ in the sense of Definition 3.3.

Remark 1.19. In general we can only take the wedge product of ωϕ when ϕ is 2 in C . Bedford-Taylor [3] proved that it can also be defined when ϕ is in L∞. In our case, ϕ is unbounded, so we have to figure out the correct definition of

p p 1 cω pχ ω − 0 ϕ − ∧ ϕ ≥ for unbounded ϕ and p =1, 2, ..., n. This will be done in Definition 3.3. Remark 1.20. When n =2, it is same as Theorem 2.12 of [18]. Now let us sketch the proof here. It is analogous to the proof of Theorem 2.12 of Demailly-Paun’s paper [18]. Consider the diagonal ∆ inside the product manifold M M. Cover it by finitely many open coordinate balls B . Since ∆ × j is non-singular, we can assume that on Bj , gj,k, k =1, 2, ..., 2n are coordinates and ∆ = gj,k =0, 1 k n . Assume that θj are smooth functions supported { 2≤ ≤ } in Bj such that θj = 1 in a neighborhood of ∆. For ǫ1.6 > 0, define P n ψ = log( θ2 g 2 + ǫ2 ). ǫ1.6 j | j,k| 1.6 j X Xk=1 Define χM M = π1∗χ + π2∗χ, × and ¯ χM M,ǫ . ,ǫ . = χM M + ǫ1.7√ 1∂∂ψǫ . . × 1 6 1 7 × − 1 6 Let χ2n χ2n M M,ǫ . ,ǫ . ct M M,ǫ . ,ǫ . f × 1 6 1 7 > × 1 6 1 7 , t,ǫ1.6,ǫ1.7 = 2n 1+ n 2n 1 χM M − c χM M − × ×

8 then by Lemma 2.1 (ii) of [18], there exists ǫ1.7 > 0 such that for ǫ1.6 small enough, χ2n M M,ǫ1.6,ǫ1.7 1 1 2n 1 × 2n 1 > ( ) − . χM M − −4n (n + 1)c × 1 Now we consider ω0,M M,t = π1∗ωt + π2∗χ. By Theorem 1.14, there exists × c ωt,ǫ . ,ǫ . [ω0,M M,t] such that 1 6 1 7 ∈ × 2n χM M trωt,ǫ ,ǫ χM M + ft,ǫ . ,ǫ . × = (n + 1)c. 1.6 1.7 × 1 6 1 7 ω2n t,ǫ1.6,ǫ1.7

Now define ω1,t,ǫ1.6,ǫ1.7 by

n 1 c − n ω1,t,ǫ1.6,ǫ1.7 = (π1) (ωt,ǫ ,ǫ π2∗χ). n ∗ 1.6 1.7 M nχ ∧

Fix ǫ1.7 and let t and ǫ1.6 convergeR to 0. For small enough ǫ1.4, let ω1.5 be the weak limit of ω ǫ χ. Then we shall expect 1,t,ǫ1.6,ǫ1.7 − 1.4 n 1 n 2 cω − (n 1)χ ω − 0 1.5 − − ∧ 1.5 ≥ in the sense of Definition 3.3. The details will be provided in Section 3. Step 3: Consider the set I of t 0 such that there exist a constant ct 0 and a smooth K¨ahler form ω [(1≥ + t)ω ] satisfying ≥ t ∈ 0 n 2 (cω (n 1)χ) ω − > 0 t − − ∧ t and χn trωt χ + ct n = c. ωt By Theorem 1.14, it suffices to show that 0 I. When t is large enough, the condition of Theorem 1.14 is satisfied. So t ∈I. It is easy to see that if t I, then for nearby t, the condition of Theorem∈ 1.14 is also satisfied. So I is open.∈ Still by Theorem 1.14, as long as t I, then for all t′ t, t′ I. Thus, in ∈ ≥ ∈ order to prove the closedness of I, it suffices to show that if t I for all t>t0, then t I. After replacing (1 + t )ω by ω , we can without∈ loss of generality 0 ∈ 0 0 0 assume that t0 = 0. In particular we can apply Theorem 1.18 to get ω1.5. Let ν(x) be the Lelong number of ω1.5 at x. For ǫ1.8 > 0 to be determined, let Y be the set Y = x : ν(x) ǫ . { ≥ 1.8} By the result of Siu [32], Y is a subvariety with dimension p

2 ω = ω + √ 1∂∂¯(Proj∗ ϕ + C d (., Y ) ) 1.11 0 − Y 1.9 1.10 χ

9 satisfies n p n 1 n 2 (c − ǫ )ω − (n 1)χ ω − > 0 − 2 1.1 1.11 − − ∧ 1.11 on a tubular neighborhood of Y , where ProjY means the projection to Y . Bya generalization of the result of Blocki and Kolodziej [4], we can glue the smooth- ing of ω outside Y and ω near Y into ω = ω + √ 1∂∂ϕ¯ satisfying 1.5 1.11 1.12 0 − 1.12 n 1 n 2 cω − (n 1)χ ω − > 0 1.12 − − ∧ 1.12 on M. Then we are done by Theorem 1.14. In general, Y is singular and we need to use Hironaka’s desingularization theorem to resolve it. The details will be provided in Section 4.

Acknowledgement

The author wishes to thank for suggesting him this problem and providing valuable comments. The author is also grateful to Jingrui Cheng for pointing out a gap in the first version of this paper, Helmut Hofer for a discussion about symplectic geometry as well as Simone Calamai and Long Li for minor suggestions. This material is based upon work supported by the National Science Foundation under Grant No. 1638352, as well as support from the S. S. Chern Foundation for Mathematics Research Fund.

2 The analysis part

In this section, we use the continuity method twice to prove Theorem 1.14. First of all, for t [0, 1], define χ by ∈ t c χ = tχ + (1 t) ω t − n 0 and define f 0 as the constant such that t ≥ n n n 1 n n 1 χ ω ω − ω ω − f t = c 0 χ 0 = t(c 0 χ 0 ) 0. t n! n! − t ∧ (n 1)! n! − ∧ (n 1)! ≥ ZM ZM ZM − ZM ZM − Now we consider the set I consisting of all t [0, 1] such that there exists ω = ω + √ 1∂∂ϕ¯ > 0 for smooth ϕ satisfying∈ t 0 − t t n χt trωt χt + ft n = c ωt and n 1 n 2 cω − (n 1)χ ω − > 0. t − − t ∧ t Then it is easy to see that 0 I. Remark that the equation is the same as ∈ n n 1 n ω ω − χ c t χ t = f t . n! − t ∧ (n 1)! t n! −

10 The linearization is

n n 1 1 n 1 n 2 ∂ϕt ∂ χt ∂χt ωt − (cω − (n 1)χ ω − ) √ 1∂∂¯ = (f )+ . (n 1)! t − − t ∧ t ∧ − ∂t ∂t t n! ∂t ∧ (n 1)! − − Assume that t I, then the left hand side is a second order elliptic equation on ∂ϕt ∈ ∂t . On the other hands, the integrability condition implies that the integral of the right hand side is 0. By standard elliptic theory and the implicit function theorem, I is open when we replace the smoothness assumption of ϕ by C100,α. However, standard elliptic regularity theory implies that any C100,α solution is automatically smooth. So I is in fact open. Assume that we are able to show the closedness of I, then we have proved Theorem 1.14 for f replaced by f1. We can use another continuity path by fixing χ and ω0 but choosing fˆt = tf1 + (1 t)f. However, it is the same as before ˆ 1 1 n 1 − except that ft > 2n ( c ) − is a function instead of a constant. Thus, we only − 1 1 n 1 need to prove the a priori estimate of ωt by assuming that ft > 2n ( c ) − is a function. We start from the following proposition which is analogous− to Lemma 3.1 of Song-Weinkove’s paper [34]: Proposition 2.1. Assume that t I and ω = ω + √ 1∂∂ϕ¯ is the corre- ∈ t 0 − t sponding solution, then there exist constants C2.1 and C2.2 depending only on c, 2 ω , the C∞-norm of χ with respect to ω , the C -norm of f with respect to 0 t 0 || t|| ω0 such that C . (ϕt inf ϕt) tr ω C e 2 1 − . χt t ≤ 2.2

i ¯j i ¯j Proof. In local coordinates, χ = √ 1χ ¯dz dz and ω = √ 1g ¯dz dz . t − ij ∧ t − ij ∧ Fix any point x, choose a χt-normal coordinate such that χi¯j = δi¯j , χi¯j,k = χi¯j,k¯ = 0 and gi¯j = λiδi¯j at x, where the derivatives are all ordinary derivatives. Then the equation is i¯j det χαβ¯ g χi¯j + ft = c. det g ¯ i,j αβ X Define an operator ∆˜ by

i¯l g det χ ¯ ˜ i¯j kl¯ αβ ∆u = ( g g χk¯j + ft )u,i¯l, det gαβ¯ Xi,l Xj,k then it is easy to see that ∆˜ is independent of the choice of local coordinates. At x, 1 1 1 ∆˜ u = ( + f )u ¯. λ2 t λ n λ ,ii i i i α=1 α X 1 1 1 n 1 Since ( ) − , it is easyQ to see that λα − 2n c 1 1 1 2 + ft n > 0 λi λi α=1 λα Q 11 for all i. So ∆˜ is a second order elliptic operator. ˜ ˜ i¯j Now we compute ∆(log trχt ωt)= ∆(log( i,j χ gi¯j )). It equals to

i¯i P2 (g ¯ ¯ + (χ ) ¯λi) g ¯ 1 1 1 i ii,kk ,kk | i ii,k| f ( 2 )( 2 + t n )) λi − ( λi) λ λk λα k P i P i k α=1 X at x. P P Q Now we differentiate the equation

i¯j det χαβ¯ g χi¯j + ft = c, det g ¯ i,j αβ X then

i¯j i¯b a¯j det χαβ¯ g χi¯j,k g ga¯b,kg χi¯j + (ft,k − det g ¯ i,j αβ X i,j,a,bX i¯j i¯j +f (χ χ ¯ g g ¯ ))=0. t ij,k − ij,k i,j X So

1 1 1 1 1 2 2 ( χ ¯ ¯ g ¯ ¯ + ( g ¯ + g ¯ ¯ ) λ λ ii,kk − λ2 ii,kk λ2 λ | ij,k| | ij,k| i i i i i i i,j i j Xk X X X 1P 1 2 1 2 1 + (f ( ( g ¯ ) + χ ¯ ¯ + g ¯ g ¯ ¯) λ t | λ ii,k | ii,kk λ λ | ij,k| − λ ii,kk α α i i i i,j i j i i X X X X Q 1 1 +f ¯ f ¯ ( g ¯ ) f ( g ¯ ¯))) = 0 t,kk − t,k λ ii,k − t,k λ ii,k i i i i X X at x. By K¨ahler condition, gi¯i,kk¯ = gkk,i¯ ¯i, gi¯j,k = gk¯j,i and gi¯j,k¯ = gik,¯ ¯j . Using the i¯i 1 bounds that χ ¯ ¯ + (χ ) ¯ + f + f ¯ + f ¯ + f + < C for all ii,kk ,kk t,k t,k t,kk t λi 2.3 i, k, it is easy| to see| that| by taking| | the| sum| of| the| previous| | | two equations,

2 g ¯ 1 1 1 ˜ ω C | i ii,k| f ∆(log trχ t) 2.4 ( 2 )( 2 + t n ) ≥− − ( λi) λ λk λα k P i k α=1 X 1 1 1 2 P2 1 Q1 2 + ( ( g ¯ + g ¯ ¯ )+ (f ( ( g ¯ ) λ λ2 λ | ij,k| | ij,k| λ t | λ ii,k | i i i,j i j α α i i Xk X X P 1 2 1 Q 1 + g ¯ ) f ¯ ( g ¯ ) f ( g ¯ ¯))). λ λ | ij,k| − t,k λ ii,k − t,k λ ii,k i,j i j i i i i X X X

12 Remark that 1 1 1 1 f ¯ ( g ¯ ) = f ( g ¯ ¯) | λ t,k λ ii,k | | λ t,k λ ii,k | α α i i α α i i X X Q 1 1 Q 1 = f ¯ ( g ¯ ) C g ¯ | t,k|| λ λ ii,k |≤ 2.5 |λ2 ii,k| i α α i i i X X 1 1 2 Q 1 1 1 2 g ¯ + C g ¯ ¯ + C ≤ 4 λ3 | ii,k| 2.6 λ ≤ 4 λ3 | ii,k| 2.7 i i i i i i X X X and

ft 1 2 nft 1 2 1 1 2 ( g ¯ ) g ¯ g ¯ ¯ . − λ | λ ii,k | ≤− λ λ2 | ii,k| ≤ 2 λ3 | ii,k| α α i i α α i i i i X X X So Q Q 2 g ¯ 1 1 1 ˜ ω C | i ii,k| f ∆(log trχt t) 2.8 ( 2 )( 2 + t n ) ≥− − ( λi) λ λk λα k P i k α=1 X 1 1 1 P 2 ft 1Q 2 + ( g ¯ + g ¯ ). λ λ2 λ | ij,k| λ λ λ | ij,k| i i i,j i j α α i,j i j Xk X X We have used P Q 1 1 2 1 2 g ¯ ¯ g ¯ ¯ λ2 λ | ij,k| ≥ λ3 | ii,k| i,j i j i i X X here. By Cauchy-Schwarz inequality and the fact that gi¯i,k = gk¯i,i,

2 1 1 1 ( g ¯ )( + f ) | ii,k| λ2 t λ n λ i k k α=1 α Xk X 1 1Q 1 gi¯i,k gj¯j,k ( 2 + ft n ) ≤ | || | λk λk α=1 λα i,j,kX Q 2 1 1 1 2 1 1 1 gi¯i,k ( 2 + ft n ) gj¯j,k ( 2 + ft n ) ≤ | | λ λk λα | | λ λk λα i,j s k k α=1 s k k α=1 X X X Q Q 2 1 1 1 2 = ( gi¯i,k ( 2 + ft n )) | | λ λk λα i s k k α=1 X X 2 gi¯i,k 1 Q 1 1 ( λi) | | ( 2 + ft n ) ≤ λi λ λk λα i i k k α=1 X X X 2 g ¯ 1 1 Q1 = ( λ ) | ik,k| ( + f ) i λ λ2 t λ n λ i k i i α=1 α X Xi,k 2 g ¯ 1 1 Q 1 ( λ ) | ij,k| ( + f ), ≤ i λ λ2 t λ n λ i j i i α=1 α X i,j,kX Q 13 ˜ ˜ so ∆(log trχt ωt) C2.9 at x. However, since x is arbitrary and ∆ is indepen- ≥− ˜ dent of the local coordinates, we see that ∆(log trχt ωt) C2.9 on M. c ≥− Choose ǫ2.10 < 2n as a small constant such that

n 1 n 2 n 1 cω − (n 1)χ ω − > 2ǫ ω − , 0 − − ∧ 0 2.10 0 then n 1 n 2 n 1 cω − (n 1)χ ω − > 2ǫ ω − 0 − − t ∧ 0 2.10 0 2C2.9 by the definition of χt. Choose C2.1 as , then at the maximal point of ǫ2.10 log tr ω C ϕ , χt t − 2.1 t i¯l g det χ ¯ ˜ i¯j kl¯ αβ 0 ǫ2.10 ∆ϕt = ( g g χk¯j + ft )(gi¯l gi¯l) < . − − det gαβ¯ − 2 Xi,l Xj,k If i¯j kl¯ 0 c g g χ ¯(2g ¯ g ) <ǫ , − kj il − i¯l 2.10 Xi,l Xj,k by the proof of Lemma 3.1 of [34], tr ω C . If χt t ≤ 2.11 i¯j kl¯ 0 c g g χ ¯(2g ¯ g ) ǫ , − kj il − i¯l ≥ 2.10 Xi,l Xj,k then

i¯l g det χαβ¯ 0 ǫ2.10 i¯j ǫ2.10 det χαβ¯ ft (gi¯l gi¯l) < + c g χi¯j = + ft , − det g ¯ − − 2 − − 2 det g ¯ αβ i,j αβ Xi,l X

det gαβ¯ 1 so i λi = < C2.12. Using the fact that λi > , trχt ωt = i λi C2.13 det χαβ¯ c ≤ is also true. This completes the proof of the proposition. Q P By adding a constant if necessary, we can without loss of generality assume 0 that supM ϕt = 0. Then we have the following C estimate: Proposition 2.2.

ϕ 0 C . || t||C ≤ 2.14 1 Moreover, C− χ ω C χ . 2.15 t ≤ t ≤ 2.15 t Proof. Lemma 3.3 and Lemma 3.4 and Proposition 3.5 of [34] only used the inequality in Proposition 2.1. So they are still true in our case. Proposition 2.3. I is closed. Proof. First of all, we want to check the uniform ellipticity and concavity for the Evans-Krylov estimate. The equation is

i¯j det χαβ¯ g χi¯j ft = c. − − det gαβ¯ −

14 View it as a function in terms of gi¯j , χi¯j and ft, then the partial derivative in the ga¯b direction is i¯b a¯j det χαβ¯ a¯b g g χi¯j + ft g . det gαβ¯ At x, it equals to 1 1 f t δ . ( 2 + ) a¯b λa λa i λi It has uniform upper bound and lower bound.Q The second order derivative in ga¯b and gcd¯ direction is

id¯ c¯b a¯j i¯b ad¯ c¯j det χαβ¯ a¯b cd¯ det χαβ¯ ad¯ c¯b g g g χi¯j g g g χi¯j ft g g ft g g . − − − det gαβ¯ − det gαβ¯

At x, when taking the product with wa¯bwcd¯ and summing a,b,c,d for any matrix wi¯j , we get

1 2 1 2 ft waa¯ 2 ft 1 2 w ¯ w ¯ ( ) w ¯ . − λ2 λ | ab| − λ2λ | ab| − λ λ − λ λ λ | ab| a b b a i i a a i i a b Xa,b Xa,b X Xa,b Q Q It is easy to see that it is non-positive. Thus, if we replace the complex second derivatives by real second derivatives, the uniform ellipticity and concavity for the Evans-Krylov estimate [23, 24, 27, 38] are satisfied. By checking Evans-Krylov’s estimate carefully, it is easy to see that in our complex case, the estimate

[(ϕ ) ¯] α C t ij C ≤ 2.16 is still true. By standard elliptic estimate, ϕ 101,α is bounded. By Arzela-Ascoli the- || t||C orem, if ti t and ti I, then a subsequence of ϕt converges to ϕt∞ in C100,α-norm.→ By∞ Remark∈ 1.16,

n 1 n 2 cω − (n 1)χ ω − > 0. t∞ − − t ∧ t∞

So by standard elliptic regularity, ϕt∞ is smooth. In other words, t I. ∞ ∈ 3 Concentration of mass and its application

In this section, we prove Theorem 1.18. However, before that, we need to figure out the correct definition of

p p 1 cω pχ ω − 0 − ∧ ≥ when ω is only a current. Recall the following definition of the smoothing:

15 Definition 3.1. Fix a smooth non-negative function ρ supported in [0,1] such that 1 2n 1 ρ(t)t − Vol(∂B1(0))dt =1 Z0 and ρ is a positive constant near 0. For any δ > 0, the smoothing ϕδ is defined by 2n y ϕδ(x)= ϕ(x y)δ− ρ( )dVoly. Cn − | δ | Z We can define the smoothing of a current using similar formula. It is easy to see that the smoothing commutes with derivatives. So (√ 1∂∂ϕ¯ ) = √ 1∂∂¯(ϕ ). − δ − δ

Recall that √ 1∂∂ϕ¯ 0 if and only if √ 1∂∂ϕ¯ δ 0 for all δ > 0. As an analogy, we can define− ≥ − ≥ p p 1 cω pχ ω − 0 − ∧ ≥ for a closed positive (1,1) current ω using smoothing. Remark that any closed positive (1,1) current can be written as √ 1∂∂¯ acting on a real function locally. − Definition 3.2. Suppose that χ is a K¨ahler form with constant coefficients on an open set O Cn. Then we say that ⊂ p p 1 c(√ 1∂∂ϕ¯ ) pχ (√ 1∂∂ϕ¯ ) − 0 − − ∧ − ≥ on O if for any δ > 0, the smoothing ϕδ satisfies

p p 1 c(√ 1∂∂ϕ¯ ) pχ (√ 1∂∂ϕ¯ ) − 0 − δ − ∧ − δ ≥ on the set O = x : B (x) O . δ { δ ⊂ } We can also define it without the constant coefficients assumption. Definition 3.3. We say that

p p 1 cω pχ ω − 0 − ∧ ≥ if on any coordinate chart, for any open subset O, as long as χ χ on O for a ≥ 0 K¨ahler form χ0 with constant coefficients, then

p p 1 cω pχ ω − 0. − 0 ∧ ≥

Remark 3.4. p p 1 c(√ 1∂∂ϕ¯ ) pχ (√ 1∂∂ϕ¯ ) − 0 − − 0 ∧ − ≥ is a convex property for ϕ. So if ω is smooth, then

p p 1 cω pχ ω − 0 − ∧ ≥ on O pointwise if and only if it is true on O in the sense of Definition 3.3.

16 For simplicity, for any positive definite n n matrix A, we define P (A) by × I 1 1 P (A) = max( )= max (tr(A )− ), I − V k λj V n 1 Cn | j=k ⊂ X6 where λj are the eigenvalues of A. Then

n 1 n 2 cω − (n 1)χ ω − 0 − − ∧ ≥ is equivalent to Pχ(ω) c. Now we need a lemma:≤ Lemma 3.5.

1 T 1 A C P (A CB− C¯ ) + tr(B− ) P ( ) I − ≤ I C¯T B  

Proof. By restricting on the codimension 1 subspaces, it suffices to prove that

1 1 T 1 1 A C − tr((A CB− C¯ )− ) + tr(B− ) tr( ). − ≤ C¯T B   It is easy to see that

1 1 T I CB− A C I O A CB− C¯ O − T 1 T = − . O I C¯ B B− C¯ I O B   −    So

1 T 1 1 1 I O A CB− C¯ O − I CB− A C − 1 T − − = T . B− C¯ I O B O I C¯ B −       After taking traces, the left hand side equals to

1 T 1 1 1 T 1 T 1 1 tr((A CB− C¯ )− ) + tr(B− ) + tr(B− C¯ (A CB− C¯ )− CB− ). − −

Now we start the proof of Theorem 1.18. By assumption, for any t > 0, there exist c > 0 and ω [(1 + t)ω ] satisfying t t ∈ 0 n 1 n 2 cω − (n 1)χ ω − > 0 t − − ∧ t and χn trωt χ + ct n = c. ωt 1 Consider ω0,M M,t = π1∗ωt + π2∗χ and χM M = π1∗χ + π2∗χ. At each point, × c × diagonalizing them so that χi¯j = δi¯j and (ωt)i¯j = λiδi¯j . Then the eigenvalues on the product manifold are λ , ...λ , 1 , ..., 1 . Their inverses are 1 , ..., 1 , c, ..., c. 1 n c c λ1 λj

17 So the sum of them is at most (n + 1)c because ct > 0. In particular, the sum of (2n-1) distinct elements among them is also at most (n + 1)c. Define ft,ǫ1.6,ǫ1.7 as in Section 1, then there exists ǫ1.7 > 0 such that for ǫ1.6 small 1 1 2n 1 enough, f > ( ) − . So we can apply Theorem 1.14 to get t,ǫ1.6,ǫ1.7 − 4n (n+1)c ωt,ǫ . ,ǫ . [ω0,M M,t] such that PχM×M (ωt,ǫ . ,ǫ . ) < (n + 1)c and 1 6 1 7 ∈ × 1 6 1 7 2n χM M trωt,ǫ ,ǫ χM M + ft,ǫ . ,ǫ . × = (n + 1)c. 1.6 1.7 × 1 6 1 7 2n ωt,ǫ1.6,ǫ1.7

(1) (1) For each point (x1, x2), we assume that z1 , ..., zn are the local coordinates (2) (2) on M x , and z , ..., zn are the local coordinates on x M. Then we ×{ 2} 1 { 1}× can express ωt,ǫ1.6,ǫ1.7 as

(1) (2) (1,2) (2,1) ωt,ǫ1.6,ǫ1.7 = ωt,ǫ1.6,ǫ1.7 + ωt,ǫ1.6,ǫ1.7 + ωt,ǫ1.6,ǫ1.7 + ωt,ǫ1.6,ǫ1.7 , where n (1) (1) (1) (1) ωt,ǫ ,ǫ = √ 1ω ¯dz dz¯ , 1.6 1.7 − t,ǫ1.6,ǫ1.7,ij i ∧ j i,j=1 X n (2) (2) (2) (2) ωt,ǫ ,ǫ = √ 1ω ¯dzi dz¯j , 1.6 1.7 − t,ǫ1.6,ǫ1.7,ij ∧ i,j=1 X n (1,2) (1,2) (1) (2) ωt,ǫ ,ǫ = √ 1ω ¯dz dz¯ , 1.6 1.7 − t,ǫ1.6,ǫ1.7,ij i ∧ j i,j=1 X ω(2,1) ω(1,2) . t,ǫ1.6,ǫ1.7 = t,ǫ1.6,ǫ1.7 (2) After changing the definition of zi if necessary, we can assume that n (2) (2) π∗χ = √ 1 dz dz¯ 2 − i ∧ i i=1 X and n (2) (2) (2) ω = √ 1 λidz dz¯ t,ǫ1.6,ǫ1.7 − i ∧ i i=1 X at (x1, x2).

Now consider ω1,t,ǫ1.6,ǫ1.7 defined as

ω1,t,ǫ1.6,ǫ1.7 n 1 c − n = (π1) (ωt,ǫ ,ǫ π2∗χ) n ∗ 1.6 1.7 M nχ ∧ n 1 R c − (1) (2) n 1 = (π1) (nωt,ǫ ,ǫ (ωt,ǫ ,ǫ ) − π2∗χ) n ∗ 1.6 1.7 1.6 1.7 M nχ ∧ ∧ n 1 Rc − (1,2) (2,1) (2) n 2 + (π1) (n(n 1)ωt,ǫ ,ǫ ωt,ǫ ,ǫ (ωt,ǫ ,ǫ ) − π2∗χ). n ∗ 1.6 1.7 1.6 1.7 1.6 1.7 M nχ − ∧ ∧ ∧ R 18 At (x1, x2),

(1) (2) n 1 nω (ω ) − π∗χ t,ǫ1.6,ǫ1.7 ∧ t,ǫ1.6,ǫ1.7 ∧ 2 n n (1) (1) (1) 1 (2) n = √ 1ω ¯dz dz¯ ( )(ωt,ǫ ,ǫ ) , − t,ǫ1.6,ǫ1.7,ij i ∧ j ∧ λ 1.6 1.7 i,j=1 α=1 α X X and

(1,2) (2,1) (2) n 2 n(n 1)ω ω (ω ) − π∗χ − t,ǫ1.6,ǫ1.7 ∧ t,ǫ1.6,ǫ1.7 ∧ t,ǫ1.6,ǫ1.7 ∧ 2 n 1 1 = √ 1ω(1,2) ω(1,2) dz(1) dz¯(1) ( )(ω(2) )n t,ǫ ,ǫ ,ik¯ t,ǫ ,ǫ ,jk¯ i j t,ǫ1.6,ǫ1.7 − − 1.6 1.7 1.6 1.7 ∧ ∧ λk λα i,j,k=1 α=k X X6 n n (1,2) (1,2) (1) (1) 1 1 (2) n √ 1ω ¯ω ¯dzi dz¯j ( )(ωt,ǫ ,ǫ ) . ≥− − t,ǫ1.6,ǫ1.7,ik t,ǫ1.6,ǫ1.7,jk ∧ ∧ λ λ 1.6 1.7 k α=1 α i,j,kX=1 X By Lemma 3.5,

n n (1) 1 (1,2) (1,2) (1) (1) ∗ Pπ χ(√ 1 ((ω ¯ ω ¯ω ¯)dzi dz¯j )) 1 − t,ǫ1.6,ǫ1.7,ij − λ t,ǫ1.6,ǫ1.7,ik t,ǫ1.6,ǫ1.7,jk ∧ i,j=1 k X Xk=1 P ω (2) π χ χM×M ( t,ǫ1.6,ǫ1.7 ) trω ( 2∗ ) ≤ − t,ǫ1.6,ǫ1.7

n c (2) π χ . ( + 1) trω ( 2∗ ) ≤ − t,ǫ1.6,ǫ1.7 Now we view n 1 c − (2) n (2) (tr π2∗χ)(ωt,ǫ . ,ǫ . ) n ωt,ǫ ,ǫ 1 6 1 7 M nχ 1.6 1.7 as a measure on x1 R M, then it is easy to see that { }× n 1 c − (2) n (tr (2) π∗χ)(ω ) n ω 2 t,ǫ1.6,ǫ1.7 nχ x M t,ǫ1.6 ,ǫ1.7 M Z{ 1}× n 1 R c − (2) n 1 = (π∗χ) (ω ) − χn 2 t,ǫ1.6,ǫ1.7 M x1 M ∧ Z{ }× n 1 Rc − χ n 1 = χ ( ) − χn ∧ c M ZM =1R.

19 By the monotonicity and convexity of Pχ,

n 1 c − 2 (2) n P (ω ) (n + 1)c (tr (2) π∗χ) (ω ) χ 1,t,ǫ1.6,ǫ1.7 n ω 2 t,ǫ1.6,ǫ1.7 nχ t,ǫ1.6 ,ǫ1.7 ≤ − M x1 M Z{ }× (2) n 2 R n 1 ( (tr (2) π∗χ)(ω ) ) c x1 M ω 2 t,ǫ1.6,ǫ1.7 (n + 1)c − { }× t,ǫ1.6,ǫ1.7 ≤ − nχn R (ω(2) )n M x M t,ǫ1.6,ǫ1.7 { 1}× n 1 χ n 1 2 Rnc − ( ( ) − R χ) = (n + 1)c M c ∧ − χn ( χ )n M R M c = c. R R Up to here, we have not used the equation

2n χM M trωt,ǫ ,ǫ χM M + ft,ǫ . ,ǫ . × = (n + 1)c. 1.6 1.7 × 1 6 1 7 ω2n t,ǫ1.6,ǫ1.7

f ω2n t,ǫ1.6 ,ǫ1.7 χ2n By the equation, t,ǫ1.6,ǫ1.7 (n+1)c M M . So as in Proposition 2.6 of [18], ≥ × ωn t ǫ it is easy to see that for any weak limit Θ of t,ǫ1.6,ǫ1.7 when and 1.6 converging to 0, 1∆Θ= ǫ3.1[∆] for a constant ǫ3.1 > 0. Let ∆ǫ3.2 be the ǫ3.2 neighborhood of ∆ with respect to χM M . Then for any δ > 0, for any small enough ǫ3.2 and × ǫ3.3, the smoothing of

n 1 n 1 c − n c − ω π∗χ ǫ χ n t,ǫ1.6,ǫ1.7 2 n 3.1 nχ ( x M) ∆ǫ ∧ − nχ M Z { 1}× ∩ 3.2 M R R is at least ǫ3.3χ for small enough t and ǫ1.6. Similarly,− locally for any n + 1 dimensional subvariety V containing ∆, for n 1 any weak limit Θ′ of ω − ,1 Θ′ = ǫ [V ] for ǫ 0. Since the dimension t,ǫ1.6,ǫ1.7 V 3.4 3.4 ≥ of ∆ is strictly smaller then V , for any fixed smoothing function, for any ǫ3.3 > 0, there exists ǫ3.2 > 0 such that the smoothing of

n 1 c − (2) n 1 π∗χ((ω ) − π∗χ) n 1 t,ǫ1.6,ǫ1.7 2 nχ ( x M) ∆ǫ ∧ M Z { 1}× ∩ 3.2 R is at most ǫ3.3χ for small enough t and ǫ1.6. Now let ǫ3.5 be an arbitrary small positive number. Then we can choose ǫ3.3 n−1 ǫ3.3 1 c such that + ǫ3.3 < n ǫ3.1. Then we choose the number ǫ3.2 depending ǫ3.5 2 RM nχ on ǫ . For any K¨ahler form ω restricted to the first n coordinates of M M, 3.3 × after choosing a good coordinate, assume that π1∗χ = δi¯j and ωi¯j = λiδi¯j . We ∗ ∗ 1 define the truncation T π χ (ω) by (T π χ (ω)) ¯ = min λi, δ ¯. Now consider 1 1 ij ǫ3.5 ij ǫ . ǫ . { } ∗ 3 5 3 5 π χ ( 1 ) ω ǫ3.5 the truncation 1,t,ǫ1.6,ǫ1.7 defined as

n 1 c − (1) (2) ∗ n 1 (π1) (T π χ (˜ωt,ǫ ,ǫ ) (ωt,ǫ ,ǫ ) − π2∗χ), n ∗ 1 1.6 1.7 1.6 1.7 M nχ ǫ3.5 ∧ ∧ R 20 (1) where the (1,1)-formω ˜t,ǫ1.6,ǫ1.7 is defined by

(1) (2) n 1 nω (ω ) − π∗χ t,ǫ1.6,ǫ1.7 ∧ t,ǫ1.6,ǫ1.7 ∧ 2 (1,2) (2,1) (2) n 2 + n(n 1)ω ω (ω ) − π∗χ − t,ǫ1.6,ǫ1.7 ∧ t,ǫ1.6,ǫ1.7 ∧ t,ǫ1.6,ǫ1.7 ∧ 2 (1) (2) n 1 =ω ˜ (ω ) − π∗χ. t,ǫ1.6,ǫ1.7 ∧ t,ǫ1.6,ǫ1.7 ∧ 2 The smoothing of

∗ π1 χ n 1 ( ǫ ) c − ω ω 3.5 ǫ χ 1,t,ǫ1.6,ǫ1.7 1,t,ǫ1.6,ǫ1.7 n 3.1 − − M nχ

ǫ3.3 is at least χ ǫ3.3χ for small enough t andR ǫ1.6. In fact, the sum of the − ǫ3.5 − first two terms is non-negative outside ∆3.2, the sum of the first and third term ǫ3.3 inside ∆3.2 is at least ǫ3.3χ and the second term inside ∆3.2 is at least χ. − − ǫ3.5 By the choice of ǫ3.3, the smoothing of ∗ π χ ( 1 ) 1 cn 1 ω ω ǫ3.5 − ǫ χ 1,t,ǫ1.6,ǫ1.7 1,t,ǫ1.6,ǫ1.7 n 3.1 − − 2 M nχ is nonnegative for small enough t and ǫ1.6. On theR other hands,

∗ P ∗ T π χ ω P ∗ ω n ǫ π1 χ( 1 ( )) π1 χ( ) ( 1) 3.5 ǫ3.5 − ≤ − for any (1,1)-form ω on the first n coordinates of M M. So using the estimate × of Pχ(ω1,t,ǫ1.6,ǫ1.7 ), it is easy to see that

∗ π χ ( 1 ) ǫ3.5 Pχ(ω ) c + (n 1)ǫ3.5. 1,t,ǫ1.6,ǫ1.7 ≤ − So if χ χ on the support of the smoothing function for a K¨ahler form χ with 0 ∗ 0 ≥ π χ ( 1 ) P ω ǫ3.5 constant coefficients, then χ0 acting on the smoothing of 1,t,ǫ1.6,ǫ1.7 is at most n−1 c n ǫ P ω 1 c ǫ χ + ( 1) 3.5. So χ0 acting on the smoothing of 1,t,ǫ1.6,ǫ1.7 2 R nχn 3.1 is − − M 1 cn−1 also at most c+(n 1)ǫ3.5. Let ω1.5 be a weak limit of ω1,t,ǫ1.6,ǫ1.7 n ǫ3.1χ, 2 RM nχ − n−1 − ω ω ǫ χ ǫ 1 c ǫ P then 1.5 [ 0 1.4 ], where 1.4 = 2 R nχn 3.1. Moreover, χ0 acting on the ∈ − M smoothing of ω1.5 is at most c + (n 1)ǫ3.5. Since ǫ3.5 is arbitrary, it is at most c. This completes the proof of Theorem− 1.18.

4 Regularization

In this section, we prove Theorem 1.11. By Remark 1.13, the n = 1 and n =2 cases have been proved. By induction, we can assume that Theorem 1.11 has been proved in dimension 1, 2, ..., n 1. By Section 1, we can in addition assume that the condition of Theorem− 1.18 are satisfied. So by Theorem 1.18, there exist a constant ǫ > 0 and a current ω [ω ǫ χ] such that 1.4 1.5 ∈ 0 − 1.4 n 1 n 2 cω − (n 1)χ ω − 0 1.5 − − ∧ 1.5 ≥

21 in the sense of Definition 3.3. 1 Pick small enough ǫ4.1 < 10000 such that ω 100ǫ ω (1 + ǫ )2(ω ǫ χ). 0 − 4.1 0 ≥ 4.1 0 − 1.4 Then there exists a current ω = ω 100ǫ ω + √ 1∂∂ϕ¯ such that 4.2 0 − 4.1 0 − 4.2 c ωn 1 n χ ωn 2 2 4.−2 ( 1) 4.−2 0 (1 + ǫ4.1) − − ∧ ≥ in the sense of Definition 3.3. Now we pick a finite number of coordinate balls B2r(xi) such that Br(xi) is a cover of M. Moreover, we require that

i i χ0 <χ< (1 + ǫ4.1)χ0

i on B2r(xi) for K¨ahler forms χ0 with constant coefficients. We also assume that (1 ǫ )√ 1∂∂¯ z 2 ω (1 + ǫ )√ 1∂∂¯ z 2 − 4.1 − | | ≤ 0 ≤ 4.1 − | | B x ϕi √ ∂∂ϕ¯ i ω B x on 2r( i). Let ω0 be potential such that 1 ω0 = 0 on 2r( i). Then we also assume that − i 2 2 ϕ z ǫ4.1r . | ω0 − | | |≤ Let ϕi be the smoothing of ϕ + (1 100ǫ )ϕi . When δ < r , it is well δ 4.2 − 4.1 ω0 5 defined on B 9 (xi). By assumption, it is easy to see that 5 r c √ ∂∂ϕ¯ i n 1 n χi √ ∂∂ϕ¯ i n 2 . 2 ( 1 δ) − ( 1) 0 ( 1 δ) − 0 (1 + ǫ4.1) − − − ∧ − ≥ So c ¯ i n 1 ¯ i n 2 (√ 1∂∂ϕδ) − (n 1)χ (√ 1∂∂ϕδ) − > 0. 1+ ǫ4.1 − − − ∧ − i i i Now define the function ϕ from B 9 (xi) to R as ϕ ϕ , our goal is to show 4.3 5 r δ ω0 that for any x M, − ∈ 2 i i ǫ4.1r + max ϕ4.3(x) < max ϕ4.3(x). i:x B 9 r(xi) B 8 r (xi) i:x Br(xi) { ∈ 5 \ 5 } { ∈ }

i If this is true, then for the smooth function ϕ4.4 = ˜maxϕ4.3 on M, where ˜max means the regularized maximum by choosing the parameters “ηi” in Lemma 2 ǫ4.1r ¯ I.5.18 of [17] to be smaller than 3 , the K¨ahler form ω4.4 = ω0 + √ 1∂∂ϕ4.4 n 1 n 2 − will satisfy cω4.−4 (n 1)χ ω4.−4 > 0. In general, this is not true. However, using the proof of− the results− ∧ of Blocki and Kolodziej [4], it is in fact true if the Lelong number is small enough. The details consist the rest of this section. r i j It is easy to see that if x B 9 r(xi) B 9 r(xj ) and δ < 10 , B δ (x) Bδ (x). ∈ 5 ∩ 5 2 ⊂ r i For any δ < and x B 9 r(xi), we defineϕ ˆδ by 20 ∈ 5 i i ϕˆδ(x) = sup (ϕ4.2 + (1 100ǫ4.1)ϕω0 ) i − Bδ (x)

22 and define νi(x, δ) by i i ϕˆ r (x) ϕˆδ(x) νi(x, δ)= 16 − . log( r ) log δ 16 − Then νi(x, δ) is monotonically non-decreasing in δ. Recall that the Lelong number is defined by νi(x) = lim νi(x, δ). δ 0 → It is independent of i and can be denoted as ν(x) instead. Recall the definition of ρ in Definition 3.1. Let

2 ǫ4.1r ǫ4.5 = , 1 1 2n 1 32n−1 5( 0 log( t )Vol(∂B1(0))t − ρ(t)dt + log 2 + 22n−3 log 2) R then by the result of Siu [32], the set Y = x : ν(x) ǫ4.5 is a subvariety. For simplicity, we assume that Y is smooth.{ The≥ singular} case will be done at the end of this section. Since Y is smooth by our assumption, as in Section 1, there exists a smooth function ϕ1.11 in a neighborhood O of Y such that

n p n 1 n 2 (c − ǫ )ω − (n 1)χ ω − > 0 − 2 1.1 1.11 − − ∧ 1.11 on O. Now we pick smaller neighborhoods O′ and O′′ such that O O and ′ ⊂ O O′. We need to prove the following proposition: ′′ ⊂ r Proposition 4.1. (1) For small enough δ < 20 , if

max νi(x, δ) 2ǫ , ≤ 4.5 i:x B 9 r(xi) { ∈ 5 } then

ϕ ǫ δ ǫ r2 < ϕi x ϕi x . sup 1.11 +3 4.5 log + 4.1 max ( δ( ) ω0 ( )) ′ − O i:x B 9 r(xi) { ∈ 5 }

r (2) For small enough δ < 20 , if

ϕ ǫ δ ǫ r2 ϕi x ϕi x , inf 1.11 +3 4.5 log 4.1 max ( δ( ) ω0 ( )) O′ − ≤ − i:x B 9 r(xi) { ∈ 5 } then i max ν (x, δ) < 4ǫ4.5. i:x B 9 r(xi) { ∈ 5 } r (3) For small enough δ < 20 , if

max νi(x, δ) 4ǫ . ≤ 4.5 i:x B 9 r(xi) { ∈ 5 }

23 then

i i 2 i i max (ϕ (x) ϕ (x)) + ǫ4.1r < max (ϕ (x) ϕ (x)). δ − ω0 δ − ω0 i:x B 9 r (xi) B 8 r(xi) i:x Br(xi) { ∈ 5 \ 5 } { ∈ }

If Proposition 4.1 is true, for small enough δ, we can define ϕ1.12 as the i i regularized maximum of ϕ1.11(x)+3ǫ4.5 log δ on O and ϕ ϕ on B 9 (xi). ′ δ ω0 5 r − i Since ν(x) <ǫ4.5 for x Y , for small enough δ, max ν (x, δ) 2ǫ4.5 i:x B 9 r(xi) 6∈ { ∈ 5 } ≤ for all x O′′. So by Proposition 4.1 (1), we do not need to worry about the 6∈ discontiuty near the boundary of O′. By Proposition 4.1 (2) and (3), there is also no need to worry about the discontinuity near the boundary of B 9 (xi). In 5 r conclusion, ϕ1.12 will be smooth and satisfy

n 1 n 2 cω − (n 1)χ ω − > 0 1.12 − − ∧ 1.12 on M as long as Y is smooth and Proposition 4.1 is true. In order to prove Proposition 4.1, we need the following lemma of Bloc ki and Kolodziej [4].

r Lemma 4.2. For any δ < and x B 9 (xi), the following estimates hold: 20 5 r i i i ∈ (1) 0 ϕˆδ ϕˆ δ ν (x, δ) log a for all a 1, a ≤ − ≤ ≥ − i i i 1 1 2n 1 32n 1 (2) 0 ϕˆ ϕ ν (x, δ)( log( )Vol(∂B (0))t − ρ(t)dt + 2n−3 log 2). ≤ δ − δ ≤ 0 t 1 2 Proof. For readers’ convenience,R we almost line by line copy the paper [4] here: i (1) It follows from the logarithmical convexity ofϕ ˆδ and the definition of νi(x, δ). i (2) Define another regularizationϕ ˜δ by

i 1 i ϕ˜ (x)= (ϕ4.2 + (1 100ǫ4.1)ϕ )dVol. δ Vol(∂B (x)) − ω0 δ Z∂Bδ(x) Then by the Poisson kernel for subharmonic functions [4] and the estimate in (1), 32n 1 32n 1 ϕi x ϕi x − ϕi ϕi − νi x,tδ ˆtδ( ) ˜tδ( ) 2n 2 (ˆtδ ˆtδ/2) ( 2n 2 log 2) ( ) − ≤ 2 − − ≤ 2 − for all t (0, 1]. By monotonicity, ∈ 32n 1 32n 1 ϕi x ϕi x − νi x,tδ − νi x, δ . ˆtδ( ) ˜tδ( ) ( 2n 2 log 2) ( ) ( 2n 2 log 2) ( ) − ≤ 2 − ≤ 2 − Define 2n 1 ρ˜(t) = Vol(∂B1(0))t − ρ(t), 1 then 0 ρ˜(t)=1. So R 1 1 32n 1 ϕ˜i ϕi = (˜ϕi ϕ˜i )˜ρ(t)dt (ˆϕi ϕˆi )˜ρ(t)dt + ( − log 2)νi(x, δ). δ − δ δ − tδ ≤ δ − tδ 22n 2 Z0 Z0 −

24 By the estimate in (1) again, 1 ϕˆi ϕˆi νi(x, δ) log( ). δ − tδ ≤ t The other side of inequality 0 ϕˆi ϕi is trivial. ≤ δ − δ r It is easy to see that there exists a constant C4.6 such that for any δ < 20 i and x B 9 r(xi), ν (x, δ) < C4.6. Now we are ready to prove Proposition 4.1. ∈ 5 r (1) Suppose δ < , x B 9 r(xi) and 20 ∈ 5 i i ϕˆ r (x) ϕˆδ(x) νi(x, δ)= 16 − 2ǫ , log( r ) log δ ≤ 4.5 16 − then

i i r ϕˆδ(x) ϕˆ r (x)+2ǫ4.5(log δ log( )) C4.7 +2ǫ4.5 log δ. ≥ 16 − 16 ≥− By Lemma 4.2 (2), ϕi (x) C +2ǫ log δ. δ ≥− 4.8 4.5 It is easy to see that for δ small enough,

2 i i sup ϕ1.11 +3ǫ4.5 log δ + ǫ4.1r < ϕδ(x) ϕω0 (x) O′ −

i because ϕ is uniformly bounded on B 9 (xi). ω0 5 r r (2) Suppose δ < , x B 9 r(xi) and 20 ∈ 5 ϕ ǫ δ ǫ r2 ϕi x ϕi x , inf 1.11 +3 4.5 log 4.1 δ( ) ω0 ( ) O′ − ≤ − then as before ϕˆi (x) ϕi (x) C +3ǫ log δ. δ ≥ δ ≥− 4.9 4.5 i By Lemma 4.2 (1) and the definition ofϕ ˆ δ (x), 2

sup ϕ4.2 C4.10 +3ǫ4.5 log δ. i ≥− B δ (x) 2

i j If x B 9 r(xj ), then B δ (x) Bδ (x) and therefore ∈ 5 2 ⊂

sup ϕ4.2 sup ϕ4.2 C4.10 +3ǫ4.5 log δ. j ≥ Bi (x) ≥− Bδ (x) δ 2

j j j By the definition ofϕ ˆδ(x) and ν (x, δ), it is easy to see that ν (x, δ) < 4ǫ4.5 if δ is small enough.

25 r (3) Suppose δ < , x (B 9 r(xi) B 8 r(xi)) Br(xj ) and 20 ∈ 5 \ 5 ∩ max νi(x, δ) 4ǫ , ≤ 4.5 i:x B 9 r(xi) { ∈ 5 } then

i i 2 2 2 7 2 ϕ δ x ϕ x ϕ ǫ r r δ r ǫ r , ˆ ( ) ω0 ( ) sup 4.2 +2 4.1 + (2 + ) (2 ) 100 4.1( ) 2 − ≤ i − − 5 B δ (x) 2 and 6 ϕ ϕj x ϕj x ǫ r2 r δ 2 r 2 ǫ r 2. sup 4.2 ˆδ( ) ω0 ( )+2 4.1 + (2 + ) (2 ) + 100 4.1( ) j ≤ − − 5 Bδ (x)

By Lemma 4.2 (1),

i i i ϕˆδ ϕˆ δ ν (x, δ) log 2 4ǫ4.5 log 2. − 2 ≤ ≤ By Lemma 4.2 (2), ϕi ϕˆi and δ ≤ δ 1 2n 1 j j 1 2n 1 3 − ϕˆ ϕ 4ǫ ( log( )Vol(∂B (0))t − ρ(t)dt + log 2). δ − δ ≤ 4.5 t 1 22n 3 Z0 −

Since supBi (x) ϕ4.2 sup j ϕ4.2, by summing everything together, for δ small δ Bδ (x) 2 ≤ ϕi x ϕi x ǫ r2 < ϕj x ϕj x Y enough, δ( ) ω0 ( )+ 4.1 δ( ) ω0 ( ). We are done if is smooth. In general −Y is singular. By Hironaka’s− desingularization theorem, there exists a blow-up M˜ of M obtained by a sequence of blow-ups with smooth centers such that the proper transform Y˜ of Y is smooth. Without loss of generality, assume that we only need to blow up once. Let π be the projection of M˜ to M. Let E be the exceptional divisor. Let s be the defining section of E. Let h √ 1 ¯ 2 be any smooth metric on the line bundle [E], then 2−π ∂∂ log s h = [E]+ ω4.11 by the Poincar´e-Lelong equation. Then it is well known that the| | smooth (1,1)- form ω [E] on M˜ and ω > C π∗ω . For example, see Lemma 3.5 4.11 ∈− 4.11 − 4.12 0 of [18] for the explanation. Define ω4.13 = C4.12π∗ω0 + ω4.11. Then ω4.13 is a K¨ahler form on M˜ .

6n Lemma 4.3. Let C4.14 = . Then for all small enough t and q-dimensional ǫ1.1 subvarieties V of M˜ , as long as q

n q 2 q (c − ǫ )((1 + C t)π∗ω + C t ω ) − 3n 1.1 4.14 0 4.14 4.13 ZV 2 q 1 2 q((1 + C t)π∗ω + C t ω ) − (π∗χ + t ω ). ≥ 4.14 0 4.14 4.13 ∧ 4.13 ZV

26 Proof. By assumption,

ǫ1.1 q q 1 ǫ1.1 q q 1 (c )π∗ω qπ∗ω − π∗χ = (c )ω qω − χ 0. − 3 0 − 0 ∧ − 3 0 − 0 ∧ ≥ ZV Zπ(V ) So

ǫ1.1 q q 1 (c )((1+C t)π∗ω ) q((1+C t)π∗ω ) − ((1+C t)π∗χ) 0. − 3 4.14 0 − 4.14 0 ∧ 4.14 ≥ ZV It suffices to show that

ǫ1.1 2 q (c )((1 + C4.14t)π∗ω0 + C4.14t ω4.13) V − 3 Z 2 q 1 2 q((1 + C t)π∗ω + C t ω ) − (π∗χ + t ω ) − 4.14 0 4.14 4.13 ∧ 4.13 ǫ1.1 q (c )((1 + C4.14t)π∗ω0) ≥ V − 3 Z q 1 q((1 + C t)π∗ω ) − ((1 + C t)π∗χ). − 4.14 0 ∧ 4.14

Since it only depends on the cohomology classes, we want to replace ω0 by a better representative in its cohomology class. Remark that π(E) is smooth by assumption. So we can apply Theorem 1.11 to π(E). As in Section 1, there exists a smooth function ϕ4.15 on a neighborhood O4.16 of π(E) in M such that ω = ω + √ 1∂∂ϕ¯ satisfies 4.15 0 − 4.15 ǫ1.1 n 1 n 2 (c )ω − (n 1)χ ω − > 0 − 2 4.15 − − ∧ 4.15

2 log s h on O4.16. Define ϕ4.17 = π | | on M π(E). Recall the definition of the ∗ 4πC4.12 \ regularized maximum in Lemma I.5.18 of [17]. For large enough C4.18, let ϕ4.19 be the regularized maximum of ϕ4.17 + C4.18 and ϕ4.15. Then ϕ4.19 is smooth on M and ω4.19 = ω0 +√ 1∂∂ϕ¯ 4.19 > 0 on M. Moreover, there exists a smaller neighborhood O of π(−E) such that ϕ = ϕ on O O . 4.20 4.19 4.15 4.20 ⊂ 4.16 After replacing ω0 by ω4.19, it suffices to show that

q ǫ1.1 q! q i 2 i (c ) ((1 + C t)π∗ω ) − (C t ω ) − 3 i!(q i)! 4.14 4.19 ∧ 4.14 4.13 i=1 − q X (q 1)! q i i 1 2 i q − ((1 + C t)π∗ω ) − C − (t ω ) − (q i)!(i 1)! 4.14 4.19 ∧ 4.14 4.13 i=1 X − − q 1 − (q 1)! q 1 i q − ((1 + C t)π∗ω ) − − − i!(q 1 i)! 4.14 4.19 i=1 − − X 2 i (C t ω ) (1 + C t)π∗χ ∧ 4.14 4.13 ∧ 4.14 2 q 1 +q((1 + C t)π∗ω + C t ω ) − C tπ∗χ 4.14 4.19 4.14 4.13 ∧ 4.14 0. ≥

27 By definition of C4.14, (q 1)! ǫ q! q − < 1.1 C (q i)!(i 1)! 6 i!(q i)! 4.14 − − − for all i = 1, 2, ..., q. So we can combine the first term and the second term. If 1 the point is inside π− (O ), then for all i =1, 2, ..., q 1, 4.20 − ǫ1.1 q i q 1 i (c )(π∗ω ) − (q i)(π∗ω ) − − π∗χ − 2 4.19 ≥ − 4.19 ∧ because ǫ1.1 n 1 n 2 (c )ω − (n 1)ω − χ − 2 4.19 ≥ − 4.19 ∧ on O4.20. So the sum of the first three terms is non-negative if i =1, 2, ..., q 1. So we are done because the i = q term and the fourth term are non-negative.− If 1 the point is outside π− (O4.20), then there exists C4.21 such that

1 C4.21π∗χ>ω4.13 > C4−.21π∗χ and 1 C4.21π∗χ > π∗ω4.19 > C4−.21π∗χ on M˜ π 1(O ). The only first order term in t is \ − 4.20 q 1 qπ∗ω − C tπ∗χ. 4.19 ∧ 4.14 Since it is positive, for small enough t, we also get the required inequality. Now we pick t> 0 such that t satisfies Lemma 4.3 and c ǫ c > c 1.1 , . 2 max 1+ C4.14t + C4.12C4.14t { − 4n 1+ ǫ4.1 }

We apply Theorem 1.11 to the lower dimensional smooth manifold Y˜ with the 2 2 K¨ahler forms (1 + C4.14t)π∗ω0 + C4.14t ω4.13 and π∗χ + t ω4.13. As in Section 1, there exists a smooth function ϕ4.22 on a neighborhood of Y˜ in M˜ such that

2 ω = (1+ C t)π∗ω + C t ω + √ 1∂∂ϕ¯ 4.22 4.14 0 4.14 4.13 − 4.22 satisfies ǫ1.1 n 1 2 n 2 (c )ω − (n 1)(π∗χ + t ω ) ω − > 0 − 4n 4.22 − − 4.13 ∧ 4.22 near Y˜ . Similarly, let ϕ4.23 be the potential near E. For large enough constant C4.24, define

1 ϕ = ˜max ϕ , ϕ + C− π∗ϕ + C 4.25 { 4.23 4.22 4.24 4.17 4.24} and 2 ω = (1+ C t)π∗ω + C t ω + √ 1∂∂ϕ¯ . 4.25 4.14 0 4.14 4.13 − 4.25

28 Then ǫ1.1 n 1 2 n 2 (c )ω − (n 1)(π∗χ + t ω ) ω − > 0 − 4n 4.25 − − 4.13 ∧ 4.25 on a neighborhood O of Y˜ E in M˜ . Since t2ω > 0, it is easy to see that ∪ 4.13 ǫ1.1 n 1 n 2 (c )(π ω4.25) − (n 1)χ (π ω4.25) − > 0 − 4n ∗ − − ∧ ∗ on π(O E). Now we choose neighborhoods O′ and O′′ of Y π(E) in M such \ ∪ that O π(O) and O O′. Then as before, for small enough δ, we can ′ ⊂ ′′ ⊂ define ϕ4.26 as the regularized maximum of π ϕ4.25 +3ǫ4.5 log δ on O′ π(E) i i ∗ \ and ϕ ϕ on B 9 (xi). Then ϕ4.26 is smooth and bounded on M π(E). δ ω0 5 r Moreover,− for \

2 ω4.26 =(1+ C4.14t)ω0 + C4.14t π ω4.13 + √ 1∂∂ϕ¯ 4.26 ∗ − 2 2 =(1+ C4.14t + C4.12C4.14t )ω0 + C4.14t π ω4.11 + √ 1∂∂ϕ¯ 4.26, ∗ − it is easy to see that

ǫ1.1 c n 1 n 2 (max c , )ω4.−26 (n 1)χ ω4.−26 > 0 { − 4n 1+ ǫ4.1 } − − ∧

2 on M π(E) because C4.14tω0 + C4.14t π ω4.13 > 0. Now we define \ ∗ 2 ω4.26 C4.14t π ω4.11 + √ 1∂∂ϕ¯ 4.26 ω ω ∗ , 4.27 = 2 = 0 + − 2 1+ C4.14t + C4.12C4.14t 1+ C4.14t + C4.12C4.14t then by the choice of t,

n 1 n 2 cω − (n 1)χ ω − > 0 4.27 − − ∧ 4.27 on M π(E). For large enough constant C , define \ 4.28 2 C4.14t 2 π log s + ϕ4.26 ϕ 2π ∗ h C , ϕ , 4.29 = ˜max | | 2 + 4.28 4.15 {1+ C4.14t + C4.12C4.14t } then ϕ is smooth on M and ω = ω + √ 1∂∂ϕ¯ satisfies 4.29 4.29 0 − 4.29 n 1 n 2 cω − (n 1)χ ω − > 0 4.29 − − ∧ 4.29 on M. We are done.

5 Deformed Hermitian-Yang-Mills Equation

In this section, we prove Theorem 1.7. The equation

n arctan λi = θˆ i=1 X

29 for eigenvalues λi of ωϕ = ω0 + √ 1∂∂ϕ¯ > 0 with respect to χ is the same as the equation − n 1 nπ arctan( )= θ.ˆ λ 2 − i=1 i X nπ ˆ To simplify the notations, define θ0 = 2 θ. Then the equation is equivalent to the inequality − 1 arctan( ) <θ0 λi i=j X6 for j =1, 2, ..., n and the equation

Im(ω + √ 1χ)n = tan(θ )Re(ω + √ 1χ)n. ϕ − 0 ϕ − Inspired by the work of Pingali in the toric case [30], the analogy of Theorem 1.11 is the following:

n Theorem 5.1. Fix a K¨ahler manifold M with K¨ahler metrics χ and ω0. Let θ (0, π ) be a constant and f > 1 be a smooth function satisfying 0 ∈ 4 − 100n χn tan(θ )Re(ω + √ 1χ)n Im(ω + √ 1χ)n f = ( 0 ϕ − ϕ − ) 0, n! n! − n! ≥ ZM ZM then there exists a smooth function ϕ satisfying the equation

Im(ω + √ 1χ)n + fχn = tan(θ )Re(ω + √ 1χ)n ϕ − 0 ϕ − and the inequality 1 arctan( ) <θ0 λi i=j X6 for j = 1, 2, ..., n and eigenvalues λ of ω = ω + √ 1∂∂ϕ¯ > 0 with respect i ϕ 0 − to χ if there exists a constant ǫ1.1 > 0 and for all p-dimensional subvarieties V with p =1, 2, ..., n (V can be chosen as M), there exist smooth functions θV (t) from [1, ) to [ pπ θ + (n p)ǫ , pπ ) such that for all t [1, ), ∞ 2 − 0 − 1.1 2 ∈ ∞

p p pπ (χ + √ 1tω0) =0,θV (t) = arg( (χ + √ 1tω0) ), lim θV (t)= . − 6 − t 2 ZV ZV →∞

When n = 1, it is trivial. In higher dimensions, we need to prove it by induction. Inspired by the work of Collins-Jacob-Yau [12], the analogy of Theorem 1.14 is the following:

n Theorem 5.2. Fix a K¨ahler manifold M with K¨ahler metrics χ and ω0. Let θ (0, π ) be a constant and f > 1 be a smooth function satisfying 0 ∈ 4 − 100n χn tan(θ )Re(ω + √ 1χ)n Im(ω + √ 1χ)n f = ( 0 ϕ − ϕ − ) 0, n! n! − n! ≥ ZM ZM

30 then there exists a smooth function ϕ satisfying the equation

Im(ω + √ 1χ)n + fχn = tan(θ )Re(ω + √ 1χ)n ϕ − 0 ϕ − and the inequality 1 arctan( ) <θ0 λi i=j X6 for j =1, 2, ..., n and eigenvalues λi of ωϕ = ω0 + √ 1∂∂ϕ¯ > 0 with respect to χ if − 1 arctan( ) <θ0 λi,0 i=j X6 for j =1, 2, ..., n and eigenvalues λi,0 of ω0 > 0 with respect to χ.

We will use the continuity method three times to prove Theorem 5.2. Letω ˜0 be the form ω0 and f˜ be the function f in Theorem 5.2. There exists a constant θ0 C5.1 such thatω ˜0 C5.1χ. We start from f = 0 and ω0 = cot( n )χ. In this case, ≥ θ0 1 θ0 ϕ = 0 is the solution. Then we let ω0 = t cot( n )χ + (1 t)(C5−.1 cot( n )+1)˜ω0 and f be the non-negative constant satisfying the integrability− condition as the 1 θ0 first path. It will imply the result for ω0 = (C5−.1 cot( n )+1)˜ω0. Then we let ω0 = tω˜0 and f be the constant satisfying the integrability condition as the second path. f must be non-negative because d tan(θ )Re(tω + √ 1χ)n Im(tω + √ 1χ)n ( 0 0 − 0 − ) dt n! − n! ZM n 1 n 1 tan(θ )Re(tω + √ 1χ) − Im(tω + √ 1χ) − = ( 0 0 − 0 − ) ω > 0 (n 1)! − (n 1)! ∧ 0 ZM − − by the assumption on ω0 and Lemma 8.2 of [12]. The continuity method will imply the result forω ˜0 when f is the non-negative constant f0 satisfying the integrability condition. Finally, we let ω0 =ω ˜0 and f = tf˜+ (1 t)f0 be the third continuity path. It will imply Theorem 5.2. − It is easy to see the openness along the paths. Thus, we only need to prove the a priori estimate along the paths. It will be achieved by Sz´ekelyhidi’s estimates in [36]. First of all, we need to rewrite the equation. The equation

Im(ω + √ 1χ)n + fχn = tan(θ )Re(ω + √ 1χ)n ϕ − 0 ϕ − can be written as n n Im (λ + √ 1) + f = tan(θ )Re (λ + √ 1). i − 0 i − i=1 i=1 Y Y It is equivalent to

n 1 f n 1 sin( arctan( )) + = tan(θ ) cos( arctan( )), λ n 2 0 λ i=1 i i=1 λi +1 i=1 i X X Q p 31 which is the same as

n 1 f cos(θ ) sin(θ arctan( )) = 0 . 0 − λ n 2 i=1 i i=1 λi +1 X Let Γ be the region consisting of (λ , ..., λQ) Rpn such that λ > 0 and 1 n ∈ i 1 arctan( ) <θ0 λi i=j X6 for all j =1, 2, ..., n, then we want to study the function

n 1 f cos(θ ) F (f, λ , ..., λ ) = sin(θ arctan( )) 0 1 n 0 − λ − n 2 i=1 i i=1 λi +1 X 1 ¯ ¯ Q p on ( 100n , ) Γ, where Γ is the closure of Γ. For any K¨ahler form ω, we say ω −Γ if the∞ eigenvalues× of ω with respect to χ is in Γ. Similarly, we define ∈ χ Fχ(f,ω) as F (f, λi) for eigenvalues λi of ω with respect to χ. In order to apply Sz´ekelyhidi’s estimates in [36], we claim the following: Proposition 5.3. Assume that n 2. If f > 1 , then ≥ − 100n (1) ∂ F (f, λ) > 0 if λ Γ; ∂λi 2 ∈ ∂ λiδij (2) ∂λ ∂λ F (f, λ) cos(θ0) 2(λ2+1)2 if λ Γ and F (f, λ)=0; i j ≤− i ∈ ∂ ∂ (3) F (f, λ) F (f, λ) if λ Γ and λi λj ; ∂λi ≤ ∂λj ∈ ≥ (4) If λ ∂Γ, then F (f, λ) < 0; (5) For any∈ λ Γ, the set ∈

λ′ Γ: F (f, λ′)=0, λ′ λ { ∈ i ≥ i} is bounded. Proof. (1)

n 1 ∂ cos(θ0 arctan( )) f cos(θ ) λ F (f, λ)= − k=1 λk + 0 i . ∂λ λ2 +1 n 2 λ2 +1 i P i k=1 λk +1 i Using the definition of Γ, it is easy to see that Q p

n 1 nθ 0 < arctan( ) < 0 λk n 1 Xk=1 − on Γ. So n 1 1 cos(θ0 arctan( )) > cos(θ0) > . − λk √2 kX=1 It is easy to see that ∂ F (f, λ) > 0 if λ Γ. ∂λi ∈

32 (2)

n ∂2 1 2λ δ F f, λ θ i ij ( )= cos( 0 arctan( )) 2 2 ∂λi∂λj − − λk (λ + 1) k=1 i n X 1 1 sin(θ0 arctan( )) 2 2 − − λk (λi + 1)(λj + 1) kX=1 f cos(θ ) λ λ f cos(θ ) 1 λ2 0 i j + 0 − i δ . n 2 λ2 +1 λ2 +1 n 2 (λ2 + 1)2 ij − k=1 λk +1 i j k=1 λk +1 i

Q p 2λiδij Q p The first term is at most cos(θ0) (λ2+1)2 . When F (f, λ) = 0, the second term − i f cos(θ0) 1 f equals to n 2 (λ2+1)(λ2+1) . When 0, it is a non-negative definite − Qk=1 √λk+1 i j ≥ matrix. When 0 >f > 1 , for any ξ , ...ξ R, − 100n 1 n ∈ n f cos(θ0) ξiξj − n 2 (λ2 + 1)(λ2 + 1) i,j=1 k=1 λk +1 i j X n Q 1 p ξiξj cos(θ0) | 5 | 5 ≤ 100n 2 4 2 4 i,j=1 (λi + 1) (λj + 1) X n 1 ξi 2 = cos(θ0)( | | 5 ) 100n 2 4 i=1 (λi + 1) X n 2 1 ξi cos(θ0) 5 ≤ 100 2 2 i=1 (λi + 1) X 1 n n ξ ξ δ cos(θ ) i j ij . ≤ 100 0 λ (λ2 + 1)2 i=1 j=1 i i X X

Since λi > cot(θ0) > 1 on Γ, it is easy to see that the second term is at most 1 4 times the first term. Similarly the third term and the fourth term are also − 1 at most 4 times the first term. (3) It− suffices to show that

d 1 f d x 2x f 1 x2 + = − + − dx x2 +1 n 2 dx x2 +1 (x2 + 1)2 n 2 (x2 + 1)2 k=1 λk +1 k=1 λk +1 p p is non-positiveQ for all x [λj , λi]. When f 0, it isQ trivial. When 0 >f > 1 , it is at most ∈ ≥ − 100n 2x 1 x2 1 2x 1 x2 1 . 2− 2 + 2 − 2 2− 2 + 2 − 2 (x + 1) 100nλi (x + 1) ≤ (x + 1) x (x + 1) This is indeed non-positive because x λ > 1. ≥ j

33 θ0 (4)The point λi = cot( n 1 ) belongs to ∂Γ and F at this point is negative because −

n 2 θ0 2 1 (cot ( n 1 )+1) θ0 1 − < sin( )= n 1 θ . 100n cos(θ ) n 1 2 0 0 cos(θ0) sin − ( n 1 ) − − Thus, it suffices to prove that if F (f, λ) = 0, then λ ∂Γ. If f 0, it is obvious. n 6∈ ≥ If 0 >f > 1 , then 0 θ arctan( 1 ) θ0 π . So λ ∂Γ 100n 0 i=1 λi n 1 4 i because − ≥ − ≥ − − ≥ − 6∈ 1 P 1 sin x < ( inf ( x2 + 1 arctan( )))( inf ). π 100n x (1, ) x x [ 4 ,0] x ∈ ∞ p ∈ − (5) is obvious. Compared to Sz´ekelyhidi’s conditions in [36], there are three major differ- ences. First of all, F also depends on f. Second of all, Γ does not contain the positive orthant. Finally, even if we fix the f variable, F is only concave when F = 0. However, we will show that his works still survive without much changes. Proposition 5 of [36] only requires the concavity of F when F = 0. So it still holds. Sz´ekelyhidi’s C0 estimate relies on the variant of Alexandroff-Bakelman- Pucci maximum principle similar to Lemma 9.2 of [25]. Clearly it does not take derivatives of f. So Sz´ekelyhidi’s C0 estimate is still true. The next step is to prove that

¯ 2 √ 1∂∂ϕ χ C5.2(1 + sup ϕ χ). | − | ≤ M |∇ |

We will use the same notations as in [36] except that letter f in [36] is replaced by F , the letter F is replaced by Fχ and the letter u is replaced by ϕ. It is easy to see that (78) of [36] still holds. Now we differentiate the equation Fχ(f,ωϕ) = 0. We see that

ij f Fχ gi¯j1 + Fχ f1 =0, and

pq,rs kk kk,f kk,f f Fχ gpq¯1grs¯1¯ + Fχ gkk¯11¯ + Fχ gkk¯1f1¯ + Fχ gkk¯1¯f1 + Fχ f11¯ =0

ff f cos(θ0) f because Fχ = 0. Since Fχ = n 2 1, the term Fχ f11¯ is bounded. | | |− Qi=1 √λi +1 |≤ 1 kk,f So the only additional term in (85) of [36] is C0λ1− Fχ gkk¯1 on the right hand side. Instead of (94) of [36], we get − | |

kk f Fχ gkkp¯ + Fχ fp =0.

f Since Fχ fp is bounded, the estimate in (95) still holds. So the only additional | | 1 kk,f term in (99) and (104) of [36] is C λ− F g ¯ on the right hand side. − 0 1 | χ kk1|

34 The case 1 in [36] will not happen. The additional term in (120) of [36] is also 1 kk,f C λ− F g ¯ . However, recall that (67) of [36] is that − 0 1 | χ kk1|

ij,rs F1 Fi F g ¯ g ¯ F g ¯ g ¯¯ − g ¯ . − χ ij1 rs¯1 ≥− ij ii1 jj1 − λ λ | i11| i>1 1 i X − (Remark that the letter f in [36] is replaced by F and the letter F is replaced by Fχ.) The term Fij gi¯i1gj¯j1¯ was thrown away. However, this term is at least λi 2 − 2(λ2+1)2 gi¯i1 . The term i | |

kk,f f cos(θ0) λi λi C0 Fχ gkk¯1 = C0 n 2 gi¯i1 C0 f 3 gi¯i1 2 λ +1 2 2 − | | − | k=1 λk +1 i |≥− | |(λi + 1) | |

λi 2Q p is at least 2(λ2+1)2 gi¯i1 C5.3. So Sz´ekelyhidi’s estimate − i | | − ¯ 2 √ 1∂∂ϕ χ C5.2(1 + sup ϕ χ) | − | ≤ M |∇ | still holds. Sz´ekelyhidi used the property that Γ contains the positive orthant to prove the C2 estimate [36]. We do not have this property. However, we can use Proposition 5.1 of [12] to achieve this. The Evans-Krylov estimate requires the uniform ellipticity and concavity of Fχ(f, .). Its relationship with the function F was cited as (63) and (64) of [36]. By Proposition 5.3, the conditions for the Evans-Krylov estimate are indeed true. The higher order estimate follows from standard elliptic theories. Finally, ωϕ will stay in the region Γχ along the continuity paths by Proposition 5.3 (4). The analogy of Theorem 1.18 is the following:

n Theorem 5.4. Fix a K¨ahler manifold M with K¨ahler metrics χ and ω0. Sup- pose that for all t> 0, there exist a constant ct > 0 and a smooth K¨ahler form ω [(1 + t)ω ] satisfying ω Γ , and t ∈ 0 t ∈ χ Im(ω + √ 1χ)n + c χn = tan(θ )Re(ω + √ 1χ)n. ϕ − t 0 ϕ −

Then there exist a constant ǫ5.4 > 0 and a current ω5.5 [ω0 ǫ5.4χ] such that ω Γ¯ in the sense of current. ∈ − 5.5 ∈ χ The definition of a current being in Γ¯χ is similar to Definition 3.3 except that we replace the condition

n 1 n 2 cω − (n 1)χ ω − 0 − − ∧ ≥ by ω Γ¯ for K¨ahler form ω. To simplify notations, for a K¨ahler form ω, we ∈ χ define Pχ(ω) and Qχ(ω) by 1 Pχ(ω) = max( arctan( )), j λi i=j X6

35 and 1 Q (ω)= arctan( ), χ λ i i X where λ are the eigenvalues of ω with respect to χ. Then ω Γ¯ is equivalent i ∈ χ to Pχ(ω) θ0. The analogy≤ of Lemma 3.5 is the following: Lemma 5.5. Suppose that A C I O > . C¯T B O I     Then 1 T A C P (A CB− C¯ )+ Q (B) P ( ). I − I ≤ I C¯T B  

Proof. It is easy to see that B > I. Moreover, for any ξ =0 Cn, 6 ∈ ¯T 1 ¯T ¯T ¯T 1 A C ξ ξ (A CB− C )ξ = ξ ξ CB− T 1 T − − C¯ B B− C¯ ξ  −  2 1 T 2 > ξ + B− C¯ ξ , | | | | 1 T so A CB− C¯ > I. Therefore, both hand sides of the inequality are well defined.− By restricting on the codimension 1 subspaces, it suffices to prove that

1 T A C Q (A CB− C¯ )+ Q (B) Q ( ). I − I ≤ I C¯T B   For any s [0, 1], it is easy to see that ∈ A + √ 1I sC det − sC¯ T B + √ 1I  −  2 1 T A s C(B + √ 1I)− C¯ + √ 1I O = det − − − . O B + √ 1I  −  2 1 T So we need to compute det(A s C(B + √ 1I)− C¯ + √ 1I). We already know that B >− I, so − −

1 1 1 1 (B + √ 1I)− = B− (I + √ 1B− )− − − ∞ k 2k 1 ∞ k 2k 2 = ( 1) B− − √ 1 ( 1) B− − . − − − − kX=0 Xk=0 Therefore 2 1 T A s C(B + √ 1I)− C¯ + √ 1I − − − 2 ∞ k 2k 1 T 2 ∞ k 2k 2 T = A s ( ( 1) CB− − C¯ )+ √ 1(I + s ( 1) CB− − C¯ ). − − − − Xk=0 Xk=0

36 2 1 T The real part is at least A s CB− C¯ > I and the imaginary part is also at least I. So − A + √ 1I sC arg det − sC¯ T B + √ 1I  −  2 1 T = arg det(A s C(B + √ 1I)− C¯ + √ 1I) + arg det(B + √ 1I). − − − − if we define arg det(X + √ 1Y ) as QY (X) for X,Y > 0. In fact, this is true up to an integer times 2π. However,− both hand sides are continuous with respect to s and this equation holds for s = 0. So it holds for all s [0, 1]. Now it suffices to show that ∈

1 T 1 T arg det(A C(B + √ 1I)− C¯ + √ 1I) arg det(A CB− C¯ + √ 1I). − − − ≥ − − It follows from the facts that

∞ k 2k 2 T ∞ k+1 2k 1 T ( 1) CB− − C¯ ( 1) CB− − C¯ 0 − ≥ − ≥ kX=0 Xk=1 and 1 T A CB− C¯ > I. −

Choose C large enough such that θ + n arctan( 1 ) < π . The definitions 5.6 0 C5.6 4 of χM M , χM M,ǫ5.7,ǫ5.8 and ft,ǫ5.7,ǫ5.8 are still the same as in Section 1. As before,× there exists× ǫ > 0 such that for ǫ small enough, f > 1 . 5.8 5.7 t,ǫ5.7,ǫ5.8 − 200n Now we consider ω0,M M,t = π1∗ωt + C5.6π2∗χ. By Theorem 5.2, there exists × 1 ωt,ǫ . ,ǫ . [ω0,M M,t] such that PχM×M (ωt,ǫ . ,ǫ . ) <θ0 + n arctan( ) and 5 7 5 8 ∈ × 5 7 5 8 C5.6 √ 2n 2n Im(ωt,ǫ5.7,ǫ5.8 + 1χM M ) + ft,ǫ5.7,ǫ5.8 χM M − × × 1 2n = tan(θ0 + n arctan( ))Re(ωt,ǫ5.7,ǫ5.8 + √ 1χM M ) . C5.6 − ×

Define ω1,t,ǫ5.7,ǫ5.8 by

n−1 ⌊ 2 ⌋ k n! n 2k 2k+1 ( 1) (π1) (ωt,ǫ− ,ǫ π2∗χ ) k=0 (n 2k)!(2k+1)! ∗ 5.7 5.8 ω1,t,ǫ ,ǫ = − − ∧ . 5.7 5.8 Im(C χ + √ 1χ)n P M 5.6 −

Fix ǫ5.8 and let t and ǫ5.7 convergeR to 0. For small enough ǫ5.4, we shall expect ω to be the weak limit of ω ǫ χ. 5.5 1,t,ǫ5.7,ǫ5.8 − 5.4 As before, we write ωt,ǫ5.7,ǫ5.8 as

ω ω(1) ω(2) ω(1,2) ω(2,1) , t,ǫ5.7,ǫ5.8 = t,ǫ5.7,ǫ5.8 + t,ǫ5.7,ǫ5.8 + t,ǫ5.7,ǫ5.8 + t,ǫ5.7,ǫ5.8 and assume that n (2) (2) π∗χ = √ 1 dz dz¯ 2 − i ∧ i i=1 X

37 and n (2) (2) (2) ω = √ 1 λidz dz¯ t,ǫ5.7,ǫ5.8 − i ∧ i i=1 X at (x1, x2). To simplify notations, we omit t,ǫ5.7,ǫ5.8, then

n−1 ⌊ 2 ⌋ k k=0 ( 1) (π1) ωˆk ω1 = − ∗ , Im(C χ + √ 1χ)n MP 5.6 − whereω ˆk equals to R

n! (1) (2) n 2k 1 2k+1 ω (ω ) − − π∗χ (n 2k 1)!(2k + 1)! ∧ ∧ 2 − − n! (1,2) (2,1) (2) n 2k 2 2k+1 + ω ω (ω ) − − π∗χ (n 2k 2)!(2k + 1)! ∧ ∧ ∧ 2 − − n (1) (1) (1) √ 1ω ¯ dzi dz¯j 1 = − ij ∧ ( )(ω(2))n (2k + 1)! ∧ λ ...λ i,j=1 α1 α2k+1 X α1,...α2Xk+1distinct n (1,2) (1,2) (1) (1) √ 1ω ¯ ω ¯ dzi dz¯j 1 − il jl ∧ (ω(2))n. − (2k + 1)! ∧ λlλα1 ...λα2k+1 i,j,lX=1 α1,...α2kX+1,ldistinct Remark that

n−1 ⌊ 2 ⌋ ( 1)k 1 1 − ( ) (2k + 1)! λα1 ...λα2k+1 − λα1 ...λα2k+1 Xk=0 α1,...α2kX+1,ldistinct α1,...α2Xk+1distinct 1 1 1 1 1 = Re(1 + √ 1 )...(1 + √ 1 )(1 + √ 1 )...(1 + √ 1 ) −λl − λ1 − λl 1 − λl+1 − λn − 0, ≤ and

n−1 ⌊ 2 ⌋ ( 1)k 1 − (ω(2))n (2k + 1)! λα1 ...λα2k+1 Xk=0 α1,...α2Xk+1distinct 1 1 = Im(1 + √ 1 )...(1 + √ 1 )(ω(2))n − λ1 − λn (2) n = Im(ω + √ 1π∗χ) . − 2 By Lemma 5.5, n n (1) 1 (1,2) (1,2) (1) (1) Pπ∗χ(√ 1 ((ω ω ω )dz dz¯ )) 1 − i¯j − λ ik¯ jk¯ i ∧ j i,j=1 k X kX=1 ∗ (2) Pχ × (ω) Qπ χ(ω ) ≤ M M − 2 1 (2) <θ n Q ∗ ω . 0 + arctan( ) π2 χ( ) C5.6 −

38 So by the monotonicity and convexity of Pχ,

Pχ(ω1) 1 (2) (2) n (θ + n arctan( ) Q ∗ (ω ))Im(ω + √ 1π∗χ) x M 0 C5.6 π2 χ 2 < { 1}× − − Im(C χ + √ 1χ)n R M 5.6 − (Q ∗ (ω(2)))Im(ω(2) + √ 1π χ)n 1 R x M π2 χ 2∗ = θ + n arctan( ) { 1}× − . 0 n C5.6 − Im(C χ + √ 1χ) R M 5.6 − Using the convexity of x arctan x and theR fact that

(2) n (2) Im(ω + √ 1π2∗χ) Qπ∗χ(ω ) = arctan( − ), 2 Re(ω(2) + √ 1π χ)n − 2∗ it is easy to see that the minimum of

(2) (2) n Q ∗ ω ω √ π χ ( π2 χ( ))Im( + 1 2∗ ) x M − Z{ 1}× (2) 1 is achieved if Qπ∗χ(ω ) is a constant, which must be n arctan( ). Thus 2 C5.6 Pχ(ω1) <θ0. Back to our original notations, it means that Pχ(ω1,t,ǫ5.7,ǫ5.8 ) <θ0. ωn k The rest part of Section 3 still holds because t,ǫ−5.7,ǫ5.8 has no concentration of mass on the diagonal if k 1. Most part of Section 4 also holds because Γ¯χ is convex. We only need to prove≥ the following analogy of Lemma 4.3:

ǫ1.1 Lemma 5.6. Let C5.9 = cot( 6n ). Then for all small enough t and all q- dimensional subvarieties V of M˜ , as long as q

√ 1( ǫ1.1 θ ) 2 2 q Im e − 3 − 0 ((1 + C t)π∗ω + C t ω + √ 1(π∗χ + t ω )) 0. 5.9 0 5.9 4.13 − 4.13 ≤ ZV

Proof. We already know that

√ 1( ǫ1.1 θ ) q Im e − 3 − 0 ((1 + C t)π∗ω + √ 1(1 + C t)π∗χ) 0. 5.9 0 − 5.9 ≤ ZV So it suffices to show that

√ 1( ǫ1.1 θ ) 2 2 q Im e − 3 − 0 ((1 + C t)π∗ω + C t ω + √ 1(π∗χ + t ω )) 5.9 0 5.9 4.13 − 4.13 ZV ǫ √ 1( 1.1 θ ) q Im e − 3 − 0 ((1 + C t)π∗ω + √ 1(1 + C t)π∗χ) . ≤ 5.9 0 − 5.9 ZV As before, there exists a smooth function ϕ5.10 on a neighborhood O5.11 of π(E) in M such that ω = ω +√ 1∂∂ϕ¯ > 0 satisfies P (ω ) <θ ǫ1.1 5.10 0 − 5.10 χ 5.10 0 − 2 on O5.11. We define ϕ4.17 as before on M π(E). For any s > 0, by our assumption, there exists ϕ such that ω \ = (1+ s)ω + √ 1∂∂ϕ¯ > 0 5.12 5.12 0 − 5.12

39 π ω5.12 π satisfies Pχ(ω5.12) < θ0 < 4 . Choose s small enough such that Pχ( 1+s ) < 4 . Choose a large enough constant C5.13. Let ϕ5.14 be the regularized maximum ω5.12 1 of ϕ5.10 and 1+s + C5−.13ϕ4.17 + C5.13. Then it is easy to see that ϕ5.14 is smooth and P (ω ) < π on M for ω = ω + √ 1∂∂ϕ¯ > 0. Moreover, χ 5.14 4 5.14 0 − 5.14 there exists a smaller neighborhood O5.15 of π(E) such that ϕ5.14 = ϕ5.10 on O O . 5.15 ⊂ 5.11 After replacing ω0 by ω5.14, it suffices to show that

√ 1( ǫ1.1 θ ) 2 2 q Im(e − 3 − 0 ((1 + C t)π∗ω + C t ω + √ 1(π∗χ + t ω )) ) 5.9 5.14 5.9 4.13 − 4.13 √ 1( ǫ1.1 θ ) q Im(e − 3 − 0 ((1 + C t)π∗ω + √ 1(1 + C t)π∗χ) ). ≤ 5.9 5.14 − 5.9 First of all,

ǫ √ 1( 1.1 θ ) q Im(e − 3 − 0 ((1 + C t)π∗ω + √ 1(1 + C t)π∗χ) ) 5.9 5.14 − 5.9 √ 1( ǫ1.1 θ ) q Im(e − 3 − 0 ((1 + C t)π∗ω + √ 1π∗χ) ) − 5.9 5.14 − 1+C t 5.9 ǫ √ 1( 1.1 θ ) q 1 = Re(qe − 3 − 0 ((1 + C t)π∗ω + √ 1τπ∗χ) − π∗χ)dτ 5.9 5.14 − ∧ Z1 0 ≥ π using a calculation similar to Lemma 8.2 of [12] because Pχ(ω5.14) < 4 on M. 1 If the point is outside π− (O5.15), then as before, we get the required inequality 1 if t is small enough. If the point is inside π− (O5.15), then

ǫ √ 1( 1.1 θ ) 2 2 q Im(e − 3 − 0 ((1 + C t)π∗ω + C t ω + √ 1(π∗χ + t ω )) ) 5.9 5.14 5.9 4.13 − 4.13 √ 1( ǫ1.1 θ ) q Im(e − 3 − 0 ((1 + C t)π∗ω + √ 1π∗χ) ) − 5.9 5.14 − q 1 ǫ − √ 1( 1.1 θ ) q! k =Im(e − 3 − 0 ((1 + C t)π∗ω + √ 1π∗χ) k!(q k)! 5.9 5.14 − kX=0 − 2 2 q k (C t ω + √ 1t ω ) − ) ∧ 5.9 4.13 − 4.13 0 ≤ using a calculation similar to Lemma 8.2 of [12] because P (ω ) < θ ǫ1.1 χ 5.14 0 − 2 on O5.15. We are done.

References

[1] Vestislav Apostolov, David M. J. Calderbank, Paul Gauduchon, and Christina W. Tønnesen-Friedman, Hamiltonian 2-forms in K¨ahler geom- etry III: extremal metrics and stability, Invent. Math., 173 (2008), no. 3, 547–601. [2] Thierry Aubin, R´eduction du cas positif de l’´equation de Monge-Amp`ere sur les vari´et´es k¨ahl´eriennes compactes `ala d´emonstration d’une in´egalit´e, J. Funct. Anal., 57 (1984), no. 2, 143–153.

40 [3] Eric Bedford and B. A. Taylor, The Dirichlet problem for a complex Monge-Amp`ere equation, Invent. Math., 37 (1976), no. 1, 1–44. [4] Zbigniew B locki and Slawomir Kolodziej , Regularization of plurisubhar- monic functions on manifolds, Proc. Amer. Math. Soc., 135 (2007), no. 7, 2089–2093 . [5] Xiuxiong Chen, On the Lower Bound of the Mabuchi Energy and Its Ap- plication, Internat. Math. Res. Notices, 2000, no. 12, 607–623. [6] Xiuxiong Chen and Jingrui Cheng, On the constant scalar curvature K¨ahler metrics, apriori estimates, preprint, https://arxiv.org/abs/1712.06697 [7] Xiuxiong Chen and Jingrui Cheng, On the constant scalar curvature K¨ahler metrics, existence results, preprint, https://arxiv.org/abs/1801.00656 [8] Xiuxiong Chen and Jingrui Cheng, On the constant scalar cur- vature K¨ahler metrics, general automorphism group, preprint, https://arxiv.org/abs/1801.05907

[9] Xiuxiong Chen, and , K¨ahler-Einstein metrics on Fano manifolds. I: Approximation of metrics with cone singularities,J. Amer. Math. Soc., 28 (2015), no.1, 183–197. [10] Xiuxiong Chen, Simon Donaldson and Song Sun, K¨ahler-Einstein metrics on Fano manifolds. II: Limits with cone angle less than 2π, J. Amer. Math. Soc., 28 (2015), no.1, 199–234. [11] Xiuxiong Chen, Simon Donaldson and Song Sun, K¨ahler-Einstein metrics on Fano manifolds. III: Limits as cone angle approaches 2π and completion of the main proof, J. Amer. Math. Soc., 28 (2015), no.1, 235–278. [12] Tristan C. Collins, Adam Jacob and Shing-Tung Yau, (1,1) forms with specified Lagrangian phase: a priori estimates and algebraic obstructions, preprint, https://arxiv.org/abs/1508.01934

[13] Tristan C. Collins and G´abor Sz´ekelyhidi, Convergence of the J-flow on toric manifolds, J. Differential Geom., 107 (2017), no. 1, 47–81. [14] Tristan C. Collins, Dan Xie, and Shing-Tung Yau, The deformed Hermitian-Yang-Mills equation in geometry and physics, preprint, https://arxiv.org/abs/1712.00893

[15] Tristan C. Collins and Shing-Tung Yau, Moment maps, non- linear PDE, and stability in mirror symmetry, preprint, https://arxiv.org/abs/1811.04824

[16] Tam´as Darvas, The Mabuchi geometry of finite energy classes, Adv. Math., 285 (2015), 182–219.

41 [17] Jean-Pierre Demailly, Complex Analytic and Differential Geometry, preprint, https://www-fourier.ujf-grenoble.fr/~demailly/manuscripts/agbook.pdf [18] Jean-Pierre Demailly and Mihai Paun, Numerical characterization of the K¨ahler cone of a compact K¨ahler manifold, Ann. of Math., (2) 159 (2004), no. 3, 1247–1274. [19] Ruadha´ıDervan and Julius Ross, K-stability for K¨ahler manifolds, Math. Res. Lett., 24 (2017), no. 3, 689–739. [20] Simon K. Donaldson, Anti self-dual Yang-Mills connections over complex algebraic surfaces and stable vector bundles, Proceedings of the London Mathematical Society, (3) 50 (1985), 1–26. [21] S. K. Donaldson, Moment maps and diffeomorphisms, Sir Michael Atiyah: a great mathematician of the twentieth century, Asian J. Math., 3 (1999), no. 1, 1–15. [22] S. K. Donaldson, Scalar Curvature and Stability of Toric Varieties, J. Dif- ferential Geom., 62 (2002), no. 2, 289–349. [23] Lawrence C. Evans, Classical solutions of fully nonlinear, convex, second order elliptic equations, Comm. Pure Appl. Math., 25 (1982), 333–363. [24] Lawrence C. Evans, Classical solutions of the Hamilton-Jacobi Bellman equation for uniformly elliptic operators, Trans. Amer. Math. Soc., 275 (1983), 245–255. [25] David Gilbarg and Neil S. Trudinger, Elliptic partial differential equations of second order. Reprint of the 1998 edition, Springer-Verlag, Berlin, 2001. [26] Adam Jacob and Shing-Tung Yau, A special Lagrangian type equation for holomorphic line bundles, Math. Ann., 369 (2017), no. 1–2, 869–898. [27] N. V. Krylov, Boundedly nonhomogeneous elliptic and parabolic equations, Izvestia Akad. Nauk. SSSR, 46 (1982), 487–523. English translation in Math. USSR Izv., 20 (1983), No. 3, 459–492. [28] Mehdi Lejmi and Gbor Sz´ekelyhidi, The J-flow and stability, Adv. Math., 274 (2015), 404–431. [29] Naichung Conan Leung, Shing-Tung Yau, and Eric Zaslow, From special Lagrangian to Hermitian-Yang-Mills via Fourier-Mukai, Adv. Theor. Math. Phys., 4 (2000), no. 6, 1319–1341. [30] Vamsi Pingali, A note on the deformed Hermitian Yang-Mills PDE, Com- plex Var. Elliptic Equ., 64 (2019), no. 3, 503–518. [31] Julius Ross and Richard Thomas, An obstruction to the existence of con- stant scalar curvature K¨ahler metrics, J. Differential Geom., 72 (2006), no. 3, 429–466.

42 [32] Yum-Tong Siu, Analyticity of sets associated to Lelong numbers and the extension of closed positive currents, Invent. Math., 27 (1974), 53–156. [33] Zakaria Sj¨ostr¨om Dyrefelt, K-semistability of cscK manifolds with tran- scendental cohomology class, J. Geom. Anal., 28 (2018), no. 4, 2927–2960. [34] Jian Song and Ben Weinkove, On the convergence and singularities of the J-flow with applications to the Mabuchi energy, Comm. Pure Appl. Math., 61 (2008), no. 2, 210–229. [35] Andrew Strominger, Shing-Tung Yau, and Eric Zaslow, Mirror Symmetry is T-duality, Nucl. Phys., B479 (1996), 243–259. [36] G´abor Sz´ekelyhidi, Fully non-linear elliptic equations on compact Hermitian manifolds, J. Differential Geom., 109 (2018), no. 2, 337–378. [37] Gang Tian, K¨ahler-Einstein metrics with positive scalar curvature, Inv. Math., 130 (1997) 1–37. [38] Neil S. Trudinger, Fully nonlinear, uniformly elliptic equations under nat- ural structure conditions, Trans. Amer. Math. Soc., 278 (1983), 751–769. [39] Karen Uhlenbeck and Shing-Tung Yau, On the existence of HermitianYang- Mills connections in stable vector bundles. Frontiers of the mathematical sciences: 1985 (New York, 1985). Communications on Pure and Applied Mathematics 39 (1986), no. S, suppl., S257S293. [40] Shing-Tung Yau, On the Ricci curvature of a compact K¨ahler manifold and the complex Monge-Amp`ere equation. I, Comm. Pure Appl. Math., 31 (1978), no. 3, 339–411. [41] Shing-Tung Yau, Review of K¨ahler-Einstein metrics in algebraic geometry, Israel Math. Conference Proceedings, Bar Ilan Univ., 433–443, 1996. [42] Kai Zheng, I-properness of Mabuchi’s K-energy, Calc. Var. Partial Differ- ential Equations, 54 (2015), no. 3, 2807–2830.

43