Let \((M,J)\) be an almost complex manifold of a real dimension \(2n\) (the definitions are given in Sect. 1). Pali proved (in [7]) that, as it is in the case of complex geometry, for plurisubharmonic functions the \((1,1)\) current \(i\partial \bar{\partial }u\) is positive.Footnote 1 So for a smooth plurisubharmonic function \(u\) we have well defined Monge–Ampère operator \((i\partial \bar{\partial }u)^n\ge 0\) and we can study the complex Monge–Ampère equation

$$\begin{aligned} (i\partial \bar{\partial }u)^n=fdV, \end{aligned}$$
(1)

where \(f\ge 0\) and \(dV\) is a (smooth) volume form.

Let \(\Omega \Subset M\) be a strictly pseudoconvex domain of class \(\mathcal {C}^{\infty }\). In this article we study the following Dirichlet problem for the Monge–Ampère equation:

$$\begin{aligned} \left\{ \begin{array}{l} u\in \mathcal {PSH}(\Omega )\cap \mathcal {C}^\infty (\bar{\Omega })\\ (i\partial \bar{\partial }u)^n=dV \; \text{ in } \;\Omega \\ u=\varphi \; \text{ on } \;\partial \Omega , \end{array} \right. \end{aligned}$$
(2)

where \(\varphi \in \mathcal {C}^\infty (\bar{\Omega })\). The main theorem is the following:

FormalPara Theorem 1

There is a unique smooth plurisubharmonic solution \(u\) of the problem (2).

In [3] the theorem above was proved for \(\Omega \subset \mathbb {C}^n\) with \(J_{\mathrm{st}}\). Note that even in the integrable case it is not enough to assume that \({\partial \Omega }\) is strictly pseudoconvex.Footnote 2 Indeed, if \(\Omega \) is the blow-up of a strictly pseudoconvex domain in \(\mathbb {C}^n\) in one point, then \({\partial \Omega }\) is strictly pseudoconvex. But if \(u\in \mathcal {PSH}(\Omega )\cap \mathcal {C}^\infty (\bar{\Omega })\), then the form \((i\partial \bar{\partial }u)^n\) is not a volume form.

In case of \(J\) not integrable McDuff constructed a domain \(\Omega \) with a non connected strictly pseudoconvex boundary (see [6]).Footnote 3 One can prove the theorem above (in almost the same way) for \(\Omega \) not necessary strictly pseudoconvex but \(\partial \Omega \) strictly pseudoconvex and \(dV\le (i\partial \bar{\partial }\varphi )^n\). It is however not clear for the author, whether there is an example of such \(\varphi \) in McDuff’s example (or in any other not strictly pseudoconvex domain with a strictly pseudoconvex boundary).

In the last section we explain how Theorem 1 gives the theorem of Harvey and Lawson about existing a continuous solution of the Dirichlet Problem for maximal functions. We even improve their result by proving that the solution is Lipschitz (if the boundary condition is regular enough).

1 Notion

We say that \((M,J)\) is an almost complex manifold if \(M\) is a manifold and \(J\) is an (\(\mathcal {C}^\infty \) smooth) endomorphism of the tangent bundle \(TM\), such that \(J^2=-\mathrm {id}\). The real dimension of \(M\) is even in that case.

We have then a direct sum decomposition \(T_{\mathbb {C}}M=T^{1,0}M \oplus T^{0,1}M\), where \(T_{\mathbb {C}}M\) is a complexification of \(TM\),

$$\begin{aligned} T^{1,0}M=\left\{ X-iJX:X\in TM\right\} \end{aligned}$$

and

$$\begin{aligned} T^{0,1}M=\left\{ X+iJX:X\in TM\right\} \left( =\left\{ \zeta \in T_{\mathbb {C}}M:\bar{\zeta }\in T^{1,0}M\right\} \right) . \end{aligned}$$

Let \(\mathcal {A}^k\) be the set of \(k\)-forms, i.e. the set of sections of \(\bigwedge ^k(T_{\mathbb {C}}M)^\star \) and let \(\mathcal {A}^{p,q}\) be the set of \((p,q)\)-forms, i.e. the set of sections of \(\bigwedge ^p (T^{1,0}M)^\star \otimes _{(\mathbb {C})} ~\bigwedge ^q (T^{0,1}M)^\star \). Then we have a direct sum decomposition \(\mathcal {A}^k=\bigoplus _{p+q=k}\mathcal {A}^{p,q}\). We denote the projections \(\mathcal {A}^k\rightarrow \mathcal {A}^{p,q}\) by \(\Pi ^{p,q}\).

If \(d:\mathcal {A}^k\rightarrow \mathcal {A}^{k+1}\) is (the \(\mathbb {C}\)-linear extension of) the exterior differential, then we define \(\partial : \mathcal {A}^{p,q}\rightarrow \mathcal {A}^{p+1,q}\) as \(\Pi ^{p+1,q}\circ d\) and \(\bar{\partial }: \mathcal {A}^{p,q}\rightarrow \mathcal {A}^{p,q+1}\) as \(\Pi ^{p,q+1}\circ d\).

We say that an almost complex structure \(J\) is integrable, if any of the following (equivalent) conditions is satisfied:

  1. (i)

    \(d=\partial +\bar{\partial }\);

  2. (ii)

    \(\bar{\partial }^2=0\);

  3. (iii)

    \([\zeta ,\xi ]\in T^{0,1}M\) for vector fields \(\zeta ,\xi \in T^{0,1}M\).

By the Newlander–Nirenberg Theorem \(J\) is integrable if and only if it is induced by a complex structure.

In the paper \(\zeta _1,\ldots ,\zeta _n\) is always a (local) frame of \(T^{1,0}\). Let us put for a smooth function \(u\)

$$\begin{aligned} u_p=\zeta _p u,\;\;u_{p\bar{q}}=\zeta _p\bar{\zeta _q}u=u_{\bar{q}p}+\left[ \zeta _p,\bar{\zeta _q}\right] u,\;\;\mathrm{etc.} \end{aligned}$$

and

$$\begin{aligned} A_{p\bar{q}}=A_{p\bar{q}}(u)=u_{p\bar{q}}-\left[ \zeta _p,\bar{\zeta _q}\right] ^{0,1}u, \end{aligned}$$

where for any \(X\in T_{\mathbb {C}}M\) a vector \(X^{0,1}\in T^{0,1}M\) is such that \(X^{1,0}:=X-X^{0,1}\in T^{1,0}M\). Then for a smooth function \(u\) we have (see [7]):

$$\begin{aligned} i\partial \bar{\partial }u=i\sum A_{p\bar{q}}\zeta _p^\star \wedge \bar{\zeta }_q^\star , \end{aligned}$$

where \(\zeta _1^\star ,\ldots , \zeta _n^\star ,\bar{\zeta }_1^\star ,\ldots , \bar{\zeta }_n^\star \) is a base of \((T_{\mathbb {C}}M)^\star \) dual to the base \(\zeta _1,\ldots ,\zeta _n, \bar{\zeta _1}, \ldots ,\bar{\zeta _n}\) of \(T_{\mathbb {C}}M\).

Let \(\mathbb {D}=\{z\in \mathbb {C}:|z|<1\}\). We say that a (smooth) function \(\lambda :\mathbb {D}\rightarrow M\) is \(J\)-holomorphic or simpler holomorphic if \(\lambda '(\frac{\partial }{\partial \bar{z}})\in T^{0,1}M\). The following proposition from [5], where it is stated for \(C^{k',\alpha }\) class of \(J\), shows that there exists plenty of such disks:

Proposition 1.1

Let \(0\in M\subset \mathbb {R}^{2n}, k,k'\ge 1\). For \(v_0,v_1,\ldots ,v_k\in \mathbb {R}^{2n}\) close enough to 0, there is a holomorphic function \(\lambda :\mathbb {D}\rightarrow M\), such that \(\lambda (0)=v_0\) and \(\frac{\partial ^l\lambda }{\partial x^l}=v_l\), for \(l=1,\ldots ,k\). Moreover, we can choose \(\lambda \) with \(\mathcal {C}^1\) dependence on parameters \((v_0,\ldots ,v_k)\in (\mathbb {R}^{2n})^{k+1}\), where for holomorphic functions we consider \(\mathcal {C}^{k'}\) norm.

We can locally normalize coordinates with respect to a given holomorphic disc \(\lambda \), that is we can assume that \(\lambda (z)=(z,0)\in \mathbb {C}^n\) and \(J=J_\mathrm{st}\) on \(\mathbb {C}\times \{0\}\subset \mathbb {C}^n\), where \(J_\mathrm{st}\) is the standard almost complex structure in \(\mathbb {C}^n\) (see section 1.2 in [1]) and moreover we can assume that for every \(J\)-holomorphic \(\mu \) such that \(\mu (0)=0\) we have \(\triangle \mu (0)=0\) (see [8]).

An upper semi-continuous function \(u\) on an open subset of \(M\) is said to be \(J\)-plurisubharmonic or simpler plurisubharmonic, if a function \(u\circ \lambda \) is subharmonic for every holomorphic function \(\lambda \). We denote the set of plurisubharmonic functions on \(\Omega \subset M\) by \(\mathcal {PSH}(\Omega )\). For a smooth function \(u\) it means that a matrix \((A_{p\bar{q}})\) is nonnegative. Recently Harvey and Lawson proved that an upper semicontinuous locally integrable function \(u\) is plurisubharmonic iff a current \(i\partial \bar{\partial }u\) is positive. We say that a function \(u \in \mathcal {C}^{1,1}(\Omega )\) is strictly plurisubharmonic if for every \(K\Subset \Omega \) there is \(m>0\) such that \(\omega \le im\partial \bar{\partial }u\) a.e. in \(K\), where \(\omega \) is any hermitian metricFootnote 4 on \(\Omega \). If \(u\in \mathcal {C}^2(\Omega )\) then the following conditions are equivalent:

  1. (i)

    \(u\) is strictly plurisubharmonic;

  2. (ii)

    \(i\partial \bar{\partial }u>0\);

  3. (iii)

    \(u\) is plurisubharmonic and \((i\partial \bar{\partial }u)^n>0\).

We say that a domain \(\Omega \Subset M\) is strictly pseudoconvex of class \(\mathcal {C}^{\infty }\) (respectively of class \(\mathcal {C}^{1,1}\)) if there is a strictly plurisubharmonic function \(\rho \) of class \(\mathcal {C}^{\infty }\) (respectively of class \(\mathcal {C}^{1,1}\)) in a neighbourhood of \(\bar{\Omega }\), such that \(\Omega =\{\rho <0\}\) and \(\triangledown \rho \ne 0\) on \(\partial \Omega \). In that case we say that \(\rho \) is a defining function for \(\Omega \).

Let \(z_0\in M\). The basic example of a (strictly) plurisubharmonic function in a neighbourhood of \(z_0\) is \(u(z)=(\mathrm{dist}(z,z_0))^2\) (where \(\mathrm{dist}\) is a distance in some Rimannian metric). Domains \(\Omega _\varepsilon =\{u<\varepsilon \}\) are strictly pseudoconvex of class \(\mathcal {C}^{\infty }\) for \(\varepsilon >0\) small enough and they make a fundamental neighbourhood system for \(z_0\).

2 Comparison principle

In this section \(\Omega \Subset M\) is a domain not necessary strictly pseudoconvex but such that there is a bounded function \(\rho \in \mathcal {C}^2\cap \mathcal { PSH}(\Omega )\).

In the pluripotential theory in \(\mathbb {C}^n\), the comparison principle is a very effective tool. We give here the basic version for \(J\)-plurisubharmonic functions.

Proposition 2.1

(comparison principle) If \(u, v\in \mathcal {C}^2(\bar{\Omega })\) are such that \(u\) is a plurisubharmonic function, \((i\partial \bar{\partial } u)^n\ge (i\partial \bar{\partial } v)^n\) on the set \(\{i\partial \bar{\partial } v>0\}\) and \(u\le v\) on \({\partial \Omega }\), then \(u\le v\) in \(\bar{\Omega }\).

Proof

First, let us assume that \((i\partial \bar{\partial } u)^n>(i\partial \bar{\partial } v)^n\) on the set \(\{i\partial \bar{\partial } v\ge 0\}\) and a function \(u-v\) takes its maximum in \(z_0\in \Omega \). By Proposition 1.1 for small \(\zeta \in T_{z_0}^{1,0}M \) there is a holomorphic disk \(\lambda \) such that \(\lambda (0)=z_0\) and \(\frac{\partial \lambda }{\partial x}(0)-iJ\frac{\partial \lambda }{\partial x}(0)=\zeta \). Hence at \(z_0\)

$$\begin{aligned} \partial \bar{\partial }(v-u) (\zeta ,\bar{\zeta }) = \bigtriangleup \left( (v-u)\circ \lambda \right) (0) \ge 0 \end{aligned}$$

so we have \(i\partial \bar{\partial }u\le i\partial \bar{\partial }v\) and then we obtain \((i\partial \bar{\partial }u)^n\le (i\partial \bar{\partial }v)^n\) which is the contradiction with our first assumption.

In the general case we put \(u'=u+\varepsilon (\rho -\sup _{\bar{\Omega }}\rho )\) and the lemma follows from the above case (with \(u'\) instead of \(u\)). \(\square \)

In Sect. 4 we use a slight stronger version of the proposition above.

Proposition 2.2

Suppose that \(u, v\in \mathcal {C}^2(\bar{\Omega })\) are such that \(u\) is a plurisubharmonic function and \((i\partial \bar{\partial } u)^n\ge (i\partial \bar{\partial } v)^n\) on the set

$$\begin{aligned} \left\{ i\partial \bar{\partial } v>0\right\} . \end{aligned}$$

Then for any \(H\in \mathcal {PSH}\), an inequality

$$\begin{aligned} \varlimsup _{z\rightarrow z_0}(u+H-v)\le 0 \end{aligned}$$

for any \(z_0\in {\partial \Omega }\) implies \(u+H\le v\) on \(\Omega \).

Proof

Let \(z_0\in \Omega \) be a point where a function \(f=u+H-v\) attains a maximum and \(\lambda \) is a holomorphic disk such that \(\lambda (0)=z_0\). Because \(H\circ \lambda \) is a subharmonic function one can find a sequence \(t_k\) of nonzero complex numbers such that

$$\begin{aligned} \lim _{k\rightarrow \infty }t_k=0 \end{aligned}$$

and

$$\begin{aligned} 4H\circ \lambda (0)\le H\circ \lambda (t_k)+H\circ \lambda (it_k)+H\circ \lambda (-t_k)+H\circ \lambda (-it_k). \end{aligned}$$

Hence

$$\begin{aligned}&\bigtriangleup \left( (v-u) \circ \lambda \right) (0)\\&\quad \ge \varlimsup _{k\rightarrow \infty } \frac{4f\circ \lambda (0)-f \circ \lambda (t_k)-f \circ \lambda (it_k)-f \circ \lambda (-t_k)-f\circ \lambda (-it_k)}{|t_k|^2} \ge 0. \end{aligned}$$

Therefore we can obtain our result exactly as in the proof of the previous proposition. \(\square \)

3 A priori estimate

In this section we will prove a \(\mathcal {C}^{1,1}\) estimate for the smooth solution \(u\) of the problem (2). By the general theory of elliptic equations (see for example [3]) we obtain from this the \(\mathcal {C}^{k,\alpha }\) estimate and then the existence of a smooth solution. The uniqueness follows from the comparison principle.

Our proofs are close to these in [3] but more complicated because of the noncommutativity of some vector fields.

3.1 Some technical preparation

In this section we assume that \(\Omega \Subset M\) is strictly pseudoconvex of class \(\mathcal {C}^\infty \) with the defining function \(\rho \). Let us fix a hermitian metric \(\omega \) on \(M\). From now all norms, gradient and hessian are taken with respect to this metric or more precisely with respect to a Rimannian metric which is given by \(g(X,Y)=\omega (X,JY)\) for vector fields \(X, Y\).

Let \(f\in \mathcal {C}^\infty (\bar{\Omega })\) be such that \(dV=f\omega ^n\). Then locally our Monge–Ampère equation \((i\partial \bar{\partial }u)^n=dV\) has a form:

$$\begin{aligned} \det (A_{p\bar{q}})=\tilde{f}=gf, \end{aligned}$$

where \(g=\det (-i\omega (\zeta _p,\bar{\zeta }_q))\). So if vectors \(\zeta _1,\ldots ,\zeta _n\) are orthonormal (i.e. \(\omega (\zeta _p,\bar{\zeta }_q)=i\delta _{pq}\)), then \(g=1\).

The following elliptic operator is very useful

$$\begin{aligned} L=L_{\zeta }=A^{p\bar{q}}\left( \zeta _p\bar{\zeta _q}-\left[ \zeta _p,\bar{\zeta _q}\right] ^{0,1}\right) . \end{aligned}$$

Note that for \(X, Y\) vector fields we have

$$\begin{aligned} X(\log \tilde{f})&= A^{p\bar{q}}XA_{p\bar{q}},\\ XY(\log \tilde{f})&= A^{p\bar{q}}XYA_{p\bar{q}}-A^{p\bar{j}}A^{i\bar{q}}(YA_{i\bar{j}})(XA_{p\bar{q}}), \end{aligned}$$

where \((A^{p\bar{q}})\) is the inverse of the matrix \((\overline{A_{p\bar{q}}})\).

In the lemmas we specify exactly how a priori estimates depend on \(\rho , f\hbox { and }\varphi \). We should emphasize that they also depend strongly on \(M, J, \omega , M'\hbox { and }m(\rho )\), where \(M'\) is some fixed domain such that \(\Omega \Subset M'\Subset M\) and \(m(\rho )\) is defined as the smallest constant \(m>0\) such that \(\omega \le mi\partial \bar{\partial }\rho \) on \(\Omega \). The notion \(C=C(A)\) really means that \(C\) depends on an upper bound for \(A\).

In the proofs below \(C\) is a constant under control, but it can change from a line to a next line.

3.2 Uniform estimate

Lemma 3.1

We have \(\Vert u\Vert _{L^\infty (\Omega )}\le C\), where \(\;C = C(\Vert \rho \Vert _{L^\infty (\Omega )}, \Vert f\Vert _{L^\infty (\Omega )}, \Vert \varphi \Vert _{L^\infty (\Omega )})\).

Proof

From the comparison principle and the maximum principle we have

$$\begin{aligned} \Vert f^{1/n}\Vert _{L^\infty (\Omega )}m(\rho )\rho +\inf _{\partial \Omega }\varphi \le u\le \sup _{\partial \Omega }\varphi . \end{aligned}$$

\(\square \)

3.3 Gradient estimate

In the next two lemmas we shall prove a priori estimate for the first derivative.

Lemma 3.2

We have

$$\begin{aligned} \Vert u\Vert _{\mathcal {C}^{0,1}(\partial \Omega )}\le C, \end{aligned}$$

where \(\;C = C(\Vert \rho \Vert _{\mathcal {C}^{0,1}(\Omega )}, \Vert f\Vert _{L^\infty (\Omega )}, \Vert \varphi \Vert _{\mathcal {C}^{1,1}(\Omega )})\).

Proof

We can choose \(A>0\) such that \(Ai\partial \bar{\partial }\rho +i\partial \bar{\partial }\varphi \ge f^{1/n}\omega \) and \(Ai\partial \bar{\partial }\rho \ge i\partial \bar{\partial }\varphi \). Thus by the comparison principle and the maximum principle we have

$$\begin{aligned}\varphi +A\rho \le u\le \varphi -A\rho \end{aligned}$$

for \(A\) large enough. So on the boundary we have

$$\begin{aligned} |\nabla u|\le |\nabla A\rho |+|\nabla \varphi |. \end{aligned}$$

\(\square \)

Lemma 3.3

We have

$$\begin{aligned} \Vert u\Vert _{\mathcal {C}^{0,1}(\Omega )}\le C, \end{aligned}$$

where \(\;C = C(\Vert \rho \Vert _{\mathcal {C}^{0,1}(\Omega )}, \Vert f^{1/n}\Vert _{\mathcal {C}^{0,1}}, \Vert u\Vert _{\mathcal {C}^{0,1}(\partial \Omega )})\).

Proof

Let us consider the function \(v=\psi |\nabla u|^2\), where a smooth plurisubharmonic function \(\psi \) will be determined later. Let us assume that \(v\) takes its maximum in \(z_0\in \Omega \). We can choose \(\zeta _1,\ldots ,\zeta _n\), such that they are orthonormal in a neighbourhood of \(z_0\), and the matrix \(A_{p\bar{q}}\) is diagonal at \(z_0\). From now on all formulas are assumed to hold at \(z_0\).

We have \(Xv=0\), hence \(X(|\nabla |^2)=-|\nabla u|^2X\log \psi \). We can calculate

$$\begin{aligned} L(v)&= L(\psi )|\nabla u|^2+\psi L(|\nabla u|^2)+A^{p\bar{p}}\left( \psi _p(|\nabla u|^2)_{\bar{p}}+\psi _{\bar{p}}(|\nabla u|^2)_p\right) \\&= |\nabla u|^2A^{p\bar{p}}\left( \psi _{p\bar{p}}-\left[ \zeta _p,\bar{\zeta _p}\right] ^{0,1} \psi -2\frac{|\psi _p|^2}{\psi }\right) + \psi L(|\nabla u|^2),\\ L(|\nabla u|^2)&= A^{p\bar{p}}\left( (|\nabla u|^2)_{p\bar{p}} -[\zeta _p,\bar{\zeta _p}]^{0,1}|\nabla u|^2\right) \\&= A^{p\bar{p}}\sum _k\left( u_{p\bar{p}k}u_{\bar{k}}+u_{k}u_{p\bar{p}\bar{k}}+|u_{pk}|^2+|u_{\bar{p}k}|^2\right. \\&\left. \quad -\left[ \zeta _p,\bar{\zeta _p}\right] ^{0,1} u_ku_{\bar{k}}-u_k \left[ \zeta _p,\bar{\zeta _p}\right] ^{0,1}u_{\bar{k}}\right) ,\\&A^{p\bar{p}}(u_{p\bar{p}k}-\left[ \zeta _p,\bar{\zeta _p}\right] ^{0,1}u_k)\\&= A^{p\bar{p}}(u_{kp\bar{p}}-\zeta _k\left[ \zeta _p, \bar{\zeta _p}\right] ^{0,1}u+\zeta _p\left[ \bar{\zeta _p}, \zeta _k\right] u+\left[ \zeta _p,\zeta _k\right] \bar{\zeta _p}u)\\&= (\log f)_k+A^{p\bar{p}}\left( \zeta _p\left[ \bar{\zeta _p}, \zeta _k\right] u+\bar{\zeta _p}\left[ \zeta _p,\zeta _k\right] u+\left[ \left[ \zeta _p,\zeta _k\right] , \bar{\zeta _p}\right] u \right. \\&\left. \quad -\left[ \left[ \zeta _p,\bar{\zeta _p}\right] ^{0,1},\zeta _k\right] u\right) . \end{aligned}$$

Then we have

$$\begin{aligned}&|A^{p\bar{p}}(u_{p\bar{p}k}-[\zeta _p,\bar{\zeta _p}]^{0,1}u_k)|\\&\quad \le C\left( \frac{\Vert f^{1/n}\Vert _{\mathcal {C}^{0,1}}}{f^{1/n}}+A^{p\bar{p}}\left( \sum _s(|u_{ps}|+|u_{p\bar{s}}|)+|\nabla u|\right) \right) \end{aligned}$$

and similarly

$$\begin{aligned}&|A^{p\bar{p}}(u_{p\bar{p}\bar{k}}-[\zeta _p,\bar{\zeta _p}]^{0,1}u_{\bar{k}})|\\&\quad \le C\left( \frac{\Vert f^{1/n}\Vert _{\mathcal {C}^{0,1}}}{f^{1/n}}+A^{p\bar{p}}\left( \sum _s(|u_{ps}|+|u_{p\bar{s}}|)+|\nabla u|\right) \right) \end{aligned}$$

so for the proper choice of \(\psi \) (we can get \(\psi =e^{A\rho }+B\) for \(A, B\) large enough) we have \(L(v)(0)>0\) and this is a contradiction with the maximality of \(v\). \(\square \)

3.4 \(\mathcal {C}^{1,1}\) estimate

Let us fix a point \(P\in {\partial \Omega }\). Now we give the \(\mathcal {C}^{1,1}\) estimate in a point \(P\) (which does not depend on \(P\)). The estimate of \(XYu(P)\), where \(X,Y\) are tangent to \({\partial \Omega }\), follows from the gradient estimate.

Lemma 3.4

Let \(N\in T_PM\) be orthogonal to \({\partial \Omega }\) such that \(N\rho =-1\) and let \(X\) be a vector field on a neighbourhood of \(P\) tangent to \({\partial \Omega }\) on \({\partial \Omega }\). We have

$$\begin{aligned} |NXu(P)|\le C, \end{aligned}$$

where \(C = C(\Vert \rho \Vert _{\mathcal {C}^{0,1}(\Omega )}, \Vert f^{1/n}\Vert _{\mathcal {C}^{0,1}}, \Vert \varphi \Vert _{\mathcal {C}^{2,1}(\Omega )}, \Vert X\Vert _{\mathcal {C}^{0,1}}, \Vert u\Vert _{\mathcal {C}^{0,1}(\Omega )})\).

Proof

Let \(X_1,X_2,\ldots ,X_{n}\) be (real) vector fields on \(U\) a neighbourhood of \(P\), tangent at \(P\) to \({\partial \Omega }\), such that \(X_1,JX_1,\ldots ,X_n,JX_n\) is a frame. Consider the function

$$\begin{aligned} v=X(u-\varphi )+B\rho +\sum _{k=1}^n|X_k (u-\varphi )|^2-A(\mathrm{dist}(P,\cdot ))^2. \end{aligned}$$

Let \(V\Subset U\) be a neighbourhood of \(P\) and \(S=V\cap \Omega \). For \(A\) large enough \(v\le 0\) on \(\partial S\).

Our goal is to show that for \(B\) large enough we have \(v\le 0\) on \(\bar{S}\). Let \(z_0\in S\) be a point where \(v\) attains a maximum and let \(\zeta _1,\ldots ,\zeta _n\) be orthonormal and such that \((A_{p\bar{q}})\) is diagonal. From now on all formulas are assumed to hold at \(z_0\). Let us calculate:

$$\begin{aligned} m(\rho )L(\rho )\ge \sum A^{p\bar{p}} \end{aligned}$$

and

$$\begin{aligned} L(-X\varphi -A(\mathrm{dist}(P,\cdot ))^2)\ge -C\sum A^{p\bar{p}}, \end{aligned}$$

hence for \(B\) large enough

$$\begin{aligned} L(B\rho -X\varphi -A(\mathrm{dist}(P,\cdot ))^2)\ge \frac{B}{2m(\rho )}\sum A^{p\bar{p}}. \end{aligned}$$

To estimate \(L(Xu+\sum _{k=1}^{n}|X_k (u-\varphi )|^2)\) let us first consider \(Y\in \{X,X_1,\ldots X_{n}\}\) and calculate

$$\begin{aligned} L(Yu)&= A^{p\bar{q}}\left( \zeta _p\bar{\zeta }_q Yu-\left[ \zeta _p,\bar{\zeta }_q\right] ^{0,1}Yu\right) \\&= Y \log f +A^{p\bar{q}}\left( \zeta _p\left[ \bar{\zeta }_q, Y\right] u+[\zeta _p, Y]\bar{\zeta }_qu-\left[ \left[ \zeta _p,\bar{\zeta }_q\right] ^{0,1},Y\right] u\right) . \end{aligned}$$

There are \(\alpha _{q,k},\beta _{q,k}\in \mathbb {C}\) such that

$$\begin{aligned} \left[ \bar{\zeta }_q, Y\right] =\sum _{k=1}^{n}\alpha _{q,k}\bar{\zeta }_k+\beta _{q,k}X_k \end{aligned}$$

and so

$$\begin{aligned} A^{p\bar{q}}\zeta _p\left[ \bar{\zeta }_q, Y\right] u = \sum _{q}\alpha _{q,q}+\sum _{k=1}^{n}A^{p\bar{p}} \beta _{p,k}\zeta _pX_ku+A^{p\bar{p}}Z_p u, \end{aligned}$$

where \(Z_p\) are vector fields under control. This gives us

$$\begin{aligned} |A^{p\bar{q}}\zeta _p\left[ \bar{\zeta }_q, Y\right] u|\le C A^{p\bar{p}}\left( 1+\sum _{k}|\beta _{p,k}\zeta _pX_ku|\right) . \end{aligned}$$

In a similar way we can estimate \(A^{p\bar{q}}[\zeta _p, Y] \bar{\zeta }_qu\) and we obtain

$$\begin{aligned} |L(Yu)|\le CA^{p\bar{p}} \left( 1+\sum _k|\zeta _pX_ku|\right) . \end{aligned}$$

Therefore

$$\begin{aligned}&L(Xu+\sum _k|X_k (u-\varphi )|^2)\\&\quad \ge A^{p\bar{p}}\sum _{k=1}^{n}\left( \zeta _pX_k (u-\varphi )\right) \left( \bar{\zeta }_{ p}X_k (u-\varphi )\right) -CA^{p\bar{p}} \left( 1+\sum _k|\zeta _pX_ku|\right) \\&\quad \ge A^{p\bar{p}}\sum _k|\zeta _pX_k u|^2-CA^{p\bar{p}} (1+\sum _k|\zeta _pX_ku|). \end{aligned}$$

Now for \(B\) large enough, since \(L(v)(z_0)>0\), we have contradiction with maximality of \(v\). Hence \(v\le 0\) on \(S\) and so \(NXu(P)\le C\). \(\square \)

Lemma 3.5

Let \(X\) be a vector field orthogonal to \({\partial \Omega }\) at \(P\). We have

$$\begin{aligned} \Vert XXu(P)\Vert \le C, \end{aligned}$$

where

$$\begin{aligned} C = C(\Vert \rho \Vert _{\mathcal {C}^{2,1}(\Omega )}, \Vert f^{1/n}\Vert _{\mathcal {C}^{0,1}}, \Vert f^{-1}\Vert _{L^\infty ({\Omega })}, \Vert \varphi \Vert _{\mathcal {C}^{3,1}(\Omega )}, \Vert X\Vert _{\mathcal {C}^{0,1}},\Vert u\Vert _{\mathcal {C}^{0,1}(\Omega )}). \end{aligned}$$

Proof

By the previous Lemma it is enough to prove that

$$\begin{aligned} |\zeta |^2\le C\left( \zeta \bar{\zeta }-\left[ \zeta ,\bar{\zeta }\right] ^{0,1}\right) u(P) \end{aligned}$$

for every vector field \(\zeta \in T^{1,0}M\) tangent (at \(P\)) to \({\partial \Omega }\).

Because our argue are local we can assume that \(P=0\in \mathbb {C}^n\). Let \(\zeta _1,\zeta _2,\ldots \zeta _n\in T^{1,0}\) be an orthonormal frame in a neighbourhood of 0 such that \(\zeta _k\rho =-\delta _{kn}\). We can assume that \(\zeta _1=\zeta \). By the strictly pseudoconvexity we have \((\zeta \bar{\zeta }-[\zeta ,\bar{\zeta }]^{0,1})\rho (P)\ne 0,\) so we can also assume that \((\zeta \bar{\zeta }-[\zeta ,\bar{\zeta }]^{0,1})\varphi (P)=0\).

From the strictly pseudoconvexity and using the Proposition 1.1 (for \(k=2\)) we can choose \(J\)-holomorphic disk \(\lambda \) such that \(\lambda (0)=0, \frac{\partial \lambda }{\partial z}(0)=a\zeta \) and

$$\begin{aligned} \rho \circ \lambda (z)= b|z|^2+O(|z|^3) \end{aligned}$$
(3.1)

for some \(a,b>0\). In particular we have

$$\begin{aligned} |z|^2\le C\mathrm{dist}(\lambda (z),\bar{\Omega }). \end{aligned}$$
(3.2)

Indeed, for \(a>0\) small enough, by the proposition 1.1 (for \(k=1\)) there is a \(J\)-holomorphic disk \(\tilde{\lambda }\) such that \(\tilde{\lambda }(0)=0\) and \(\frac{\partial \tilde{\lambda }}{\partial z}(0)=a\zeta \). Then (by changing coordinates) we can assume \((\zeta J)(0)=0\), so for any \(J\)-holomorphic disk \(\lambda \) such that \(\lambda (0)= 0\) and \(\frac{\partial {\lambda }}{\partial z}(0)=a\zeta \) we have \(\frac{\partial ^2\lambda }{\partial x^2}(0)=-J\frac{\partial ^2\lambda }{\partial x\partial y}(0)=-\frac{\partial ^2\lambda }{\partial y^2}(0)\). Now if we put

$$\begin{aligned} \frac{\partial ^2\lambda }{\partial x^2}(0) = a^2\left( 0,4\frac{\partial ^2\rho }{\partial x_1^2}(0)-2\left( \zeta \bar{\zeta }-\left[ \zeta ,\bar{\zeta }\right] ^{0,1}\right) \rho (0),-4\frac{\partial ^2\rho }{\partial x_1\partial x_2}(0)\right) \end{aligned}$$

we obtain (3.1) with \(b=a^2(\zeta \bar{\zeta }-[\zeta ,\bar{\zeta }]^{0,1})\rho (0)\).

Once again changing coordinates we may assume \(\lambda (z_1)=(z_1,0), \zeta _k(0)=\frac{\partial }{\partial z_k}\) for \(k=1,\ldots ,n\) and for every \(J\)-holomorphic disk \(\mu \) such that \(\mu (0)=0\) we have

$$\begin{aligned} \frac{\partial ^2\mu }{\partial z\partial \bar{z}}(0)=0. \end{aligned}$$
(3.3)

We can find a holomorphic cubic polynomial \(p_1\) and a complex number \(\alpha \) such that

$$\begin{aligned} \varphi (z)&= \varphi (0)+\varphi '(0)(z)\\&\quad +\mathrm{Re}\left( \sum _{p=1 }^n\frac{\partial ^2\varphi }{\partial z_1\partial \bar{z}_p}z_1\bar{z}_p+p_1(z) + \alpha z_1|z_1|^2\right) +O(|z_1|^4+|z_2|^2+\ldots +|z_n|^2). \end{aligned}$$

By (3.2) we can choose another cubic polynomial \( p_2\) and numbers \(\beta _1,\beta _2,\ldots ,\beta _n\in \mathbb {C}, \beta _1>0\) such that

$$\begin{aligned} \mathrm{Re}z_n=\mathrm{Re}\left( \sum _{p=1}^n\beta _pz_1\bar{z}_p+ p_2(z)\right) +O(|z_1|^3+|z_2|^2+\ldots +|z_n|^2) \hbox { on } \partial \Omega . \end{aligned}$$

Then we obtain

$$\begin{aligned} \varphi (z)+\varphi (0) =\varphi '(0)(z)+\mathrm{Re}\left( \sum _{p=2}^na_pz_1\bar{z}_p+ p_3(z)\right) +O(|z_2|^2+\ldots +|z_n|^2) \end{aligned}$$

for some numbers \(a_2,\ldots ,a_n\in \mathbb {C}\) and a new cubic polynomial \(p_3\). Hence

$$\begin{aligned} u(z)-u(0)=\mathrm{Re}\left( \sum _{p=2}^na_pz_1\bar{z}_p+ p_4(z)\right) +O(|z_2|^2+\ldots +|z_n|^2) \end{aligned}$$
(3.4)

for \(z\in \partial \Omega \) and same polynomial \(p_4\).

Let \(B>1\) Footnote 5 and \(D=B^{-1}\max \{|a_2|,\ldots ,|a_n|\}\). By the Proposition 1.1 (again for \(k=2\)) there is a family of \(J\)-holomorphic disks \(g_w:\mathbb {D}\rightarrow \mathbb {C}^n,\,w\in \mathbb {C}^{n-1}, g_w=(g_w^1,\ldots ,g_w^n)\) such that

$$\begin{aligned} g_w(0)&= (0,w),\\ \frac{\partial g_w}{\partial z}(0)&= (1,-\frac{ a_2}{B},\ldots ,-\frac{ a_n}{B}),\\ \Vert g_w-\lambda \Vert _{\mathcal {C}^4}&\le C(|w|+D) \end{aligned}$$

and a function \(G:\mathbb {D}\times \mathbb {C}^{n-1}\rightarrow \mathbb {C}^n\) given by \(G(z,w)=g_w(z)\) is of class \(\mathcal {C}^4\). Then we have

$$\begin{aligned} |g_w(z)-(w_1+z,w_2-\frac{a_2z}{B},\ldots ,w_n-\frac{a_nz}{B})|<C|z|^2(|w|+D) \end{aligned}$$
(3.5)

for \(z\in \mathbb {D}\) and by (3.2)

$$\begin{aligned} |z|<C(\sqrt{|w|}+D) \end{aligned}$$
(3.6)

if \(g_w(z)\in \Omega \).

We can choose domains \( U\subset \mathbb {D}, V\subset \mathbb {C}^{n-1}, W\subset \mathbb {C}^{n}\) such that \(0\in W, G(\partial U\times V)\cap \bar{\Omega }=\emptyset \) and \(G|_{U\times V}\) is a diffeomorphism onto \(W\).

Let

$$\begin{aligned} h(g_w(z))=\mathrm{Re}p_w(z)+AB |w|^2+\varepsilon \rho , \end{aligned}$$

where \(A,\varepsilon >0\) and \(p_w\) is a holomorphic cubic polynomial in one variable such that

$$\begin{aligned} \mathrm{Re} p_4(g_w(z))=\mathrm{Re}p_w(z)+a_w|z|^2+\mathrm{Re} b_wz|z|^2+O(|z|^4) \end{aligned}$$

for some \(a_w\in \mathbb {R}, b_w\in \mathbb {C}\). Note that \(|b_w|< C(|w|+D)\) and by (3.3) \(|a_w|<C|w|\). Thus enlarging \(A\) (if necessary) and using also (3.6) we obtain

$$\begin{aligned} \mathrm{Re}p_4(g_w(z))\le h(g_w(z))+\frac{1}{2}D^2|z|^2 \end{aligned}$$
(3.7)

on \(\partial \Omega \).

By inequalities (3.5) and (3.6) we have

$$\begin{aligned}&2\sum _{k=2}^n\mathrm{Re}a_kg_w^1(z)g_w^k(z)\\&\quad =\sum _{k=2}^nB\left( |-\frac{a_kg_w^1(z)}{B}-g_w^k(z)|^2-| \frac{a_kg_w^1(z)}{B}|^2-|g_w^k(z)|^2\right) \\&\quad \le B(|w|^2-D^2|z|^2+C|z|^4(|w|^2+D^2))\le CB|w|^2-\frac{1}{2}D^2|z|^2 \end{aligned}$$

for \(B\) large enough. By an above estimate, (3.4), and (3.7) we obtain that if \(A\) is large enough then \(h\ge u-u(0)\) on \(\partial \Omega \cap W\). Again enlarging \(A\) we can assume \(h\ge u-u(0)\) on \(\partial S\) where \(S=\Omega \cap W\). Since \(i\partial \bar{\partial }h\) is under control for \(\varepsilon \) enough small we get an inequality

$$\begin{aligned} \left( i\partial \bar{\partial }h\right) ^n<\left( i\partial \bar{\partial }u\right) ^n \end{aligned}$$

on the set \(S\cap \{i\partial \bar{\partial }h>0\}\). This by the comparison principle gives us \( h\ge u-u(0)\) on \(S\). Note that \(h_N\ge u_N, \varphi _{1\bar{1}}=0\) and \(\varphi _N=h_N-\varepsilon \rho _N=h_N+\varepsilon \), so we can conclude that

$$\begin{aligned} u_{1\bar{1}}=u_{1\bar{1}}-\varphi _{1\bar{1}}=\left( \varphi _N-u_N\right) \rho _{1\bar{1}}\ge \varepsilon \rho _{1\bar{1}}. \end{aligned}$$

\(\square \)

Finally we will obtain the interior \(\mathcal {C}^{1,1}\) estimate, which together with previous lemmas gives us a full \(\mathcal {C}^{1,1}\) estimate. By a standard argumentation this ends the proof of Theorem 1.

Lemma 3.6

We have

$$\begin{aligned} \Vert Hu\Vert _{L^\infty (\Omega )}\le C, \end{aligned}$$
(3.8)

where \(Hu\) is a Hessian of \(u\) and

$$\begin{aligned} C = C\left( \Vert \rho \Vert _{\mathcal {C}^{0,1}(\Omega )}, \Vert f^{\frac{1}{2n}}\Vert _{\mathcal {C}^{1,1}}, \Vert u\Vert _{\mathcal {C}^{0,1}(\Omega )}, \Vert Hu\Vert _{L^\infty ({\partial \Omega })}\right) . \end{aligned}$$

Proof

Let us define \(M\) as the biggest eigenvalue of the Hessian \(Hu\). We will show that the function

$$\begin{aligned} \Lambda =\psi e^{K|\nabla u|^2}M, \end{aligned}$$

where a smooth plurisubharmonic function \(\psi >1\) and a small positive number \(K\) will be determined later, does not attain maximum in \(\Omega \). Because a function \(u\) is plurisubharmonic this will give (3.6).

Assume that a maximum of the function \(\Lambda \) is attained at a point \(z_0\in ~\Omega \) (otherwise we are done). There are \(\zeta _1,\ldots ,\zeta _n\in T_{z_0}^{1,0}M\) orthonormal at \(z_0\) such that the matrix \((A_{p\bar{q}})\) is diagonal at \(z_0\). Let \(X\in T_{z_0}M\) be such that \(\Vert X\Vert =1\) and \(M=H(X,X)\). We can normalize coordinates near \(z_0\) such that \(z_0=0\in \mathbb {C}^n,\,X=\frac{\partial }{\partial x_1}(0)\) and \(J(z,0)=J_{st}\) for small \(z\in \mathbb {C}\). Let us extend \(X\) as \(\frac{\partial }{\partial x_1}\) and then in a natural way we can extend \(\zeta _1,\ldots ,\zeta _n\) to some neighbourhood \(U\) of \(0\) such that \([\zeta _k,X](0)=0\) and \([\zeta _k,\bar{\zeta }_k](0)=0\) for \(k=1,\ldots ,n\). Indeed, on \(U\cap \mathbb {C}\times \{0\}\) we can put \(\zeta _k\) as the same linear combination of vectors \(\frac{\partial }{\partial z_1}\ldots ,\frac{\partial }{\partial z_n}\) as in \(0\). Then for some small \(a>0\) we can take (for \(\zeta _k\) not tangent to \(\mathbb {C}\times \{0\}\)) \(J\)-holomorphic disks \(d_k:\mathbb {D}\rightarrow U\) such that \(d_k(0)=0\) and \(\frac{\partial d_k}{\partial z}(0)=a\zeta _k(0)\), and on the image of \(d_k\) we can put \(\zeta _k(w)=a^{-1}\frac{\partial d_k}{\partial z}(d_k^{-1}(w))\). On the end we extend the vector fields on whole \(U\).

Let

$$\begin{aligned} v=\psi e^{K|\nabla u|^2}\frac{Hu\left( \frac{\partial }{\partial x_1},\frac{\partial }{\partial x_1}\right) }{|\frac{\partial }{\partial x_1}|^2}=\Psi e^{K|\nabla u|^2}(u_{x_1x_1}+Tu) \hbox {on} U, \end{aligned}$$

where \(\Psi =\frac{\psi }{|\frac{\partial }{\partial x_1}|^2}\) and \(T\) is a vector field (which is under control), then also a function \(v\) has a maximum at 0 (in particular \(L(v)\le 0\)). Let us put \(\mu = u_{x_1x_1}+Tu\) (then \(|\frac{\partial }{\partial x_1}(0)|^2\mu (0)=M(0)\)). Note that we have \(XYu\le C\mu \) for vector fields \(X, Y\) (which are under control). Assume \(\mu >1\) (otherwise we have \(\Lambda <C\), so we are done).

From now all formulas are assumed to hold at \(0\). We estimate \(L(v)\) from below:

$$\begin{aligned} L(v)= L\left( \Psi e^{K|\nabla u|^2}\right) \mu + \Psi e^{K|\nabla u|^2}L(\mu ) -2A^{p\bar{p}}\frac{(\Psi e^{K|\nabla u|^2})_p(\Psi e^{K|\nabla u|^2})_{\bar{p}}\mu }{\Psi e^{K|\nabla u|^2}} \end{aligned}$$

To estimate the first term let us calculate

$$\begin{aligned}&L(\Psi e^{K|\nabla u|^2})\\&\quad =e^{K|\nabla u|^2}A^{p\bar{p}}(\Psi _{p\bar{p}}+2K\mathrm{Re}(\Psi _p(|\nabla u|^2)_{\bar{p}})+K\Psi (|\nabla u|^2)_{p\bar{p}}\\&\qquad +K^2\Psi |(|\nabla u|^2)_p|^2),A^{p\bar{p}}(|\nabla u|^2)_{p\bar{p}}\\&\quad =A^{p\bar{p}}\sum _k((\zeta _p\bar{\zeta }_{ p }\eta _ku)u_{\bar{k}}+u_{k} (\zeta _p\bar{\zeta }_{ p }\bar{\eta }_ku) +(\bar{\zeta }_p\eta _ku) (\zeta _p\bar{\eta }_ku) +(\zeta _p\eta _ku) (\bar{\zeta }_p\bar{\eta }_ku))\\&\quad =\sum _k((\log \tilde{f})_ku_{\bar{k}}+(\log \tilde{f})_{\bar{k}}u_k)\\&\quad +A^{p\bar{p}}\sum _k((\zeta _p[\zeta _{\bar{p} },\eta _k]u)u_{\bar{k}}+([\zeta _p,\eta _k]u_{\bar{p} })u_{\bar{k}}+(\zeta _p[\bar{\zeta }_{p}, \bar{\eta }_k]u)u_{ k}+([\zeta _p,\bar{\eta }_k]u_{\bar{p} })u_{k})\\&\quad +A^{p\bar{p}}\sum _k\left( 2K\mathrm{Re} \left( \eta _k\left[ \zeta _p,\bar{\zeta }_p\right] ^{0,1}uu_{\bar{k}}\right) +\left( \bar{\zeta }_p\eta _k u\right) \left( \zeta _p\bar{\eta }_k u\right) +\left( \zeta _p\eta _k u\right) \left( \bar{\zeta }_p\bar{\eta }_k u\right) \right) , \end{aligned}$$

where \(\eta _1,\ldots \eta _n\) is an orthonormal frame such that \(\eta _k(0)=\zeta _k(0)\). Therefore we have

$$\begin{aligned} A^{p\bar{p}}(|\nabla u|^2)_{p\bar{p}}\ge -\,C+A^{p\bar{p}}\frac{1}{2}\sum _k\left( \left( \bar{\zeta }_p\zeta _k u\right) \left( \zeta _p\bar{\zeta }_k u\right) +\left( \zeta _p\zeta _k u\right) \left( \bar{\zeta }_p\bar{\zeta }_k u\right) -C\right) , \end{aligned}$$

hence

$$\begin{aligned} L(\Psi e^{K|\nabla u|^2})&\ge -\,Ce^{K|\nabla u|^2}\\&\quad +\,e^{K|\nabla u|^2}A^{p\bar{p}}\left( \Psi _{p\bar{p}}-CK|\Psi _p|\sum _k\left( |u_{p\bar{k}}|+|u_{pk}|+1\right) \right) \\&\quad +\,e^{K|\nabla u|^2}A^{p\bar{p}}\Psi (\frac{1}{2} K-CK^2)\sum _k\left( |u_{p\bar{k}}|^2+|u_{pk}|^2\right) . \end{aligned}$$

Let us start the calculation for the second term

$$\begin{aligned} L(\mu )&= L\left( u_{x_1x_1}\right) +L(Tu),\\ L(Tu)&\le T(\log \tilde{f})-C(\mu +1)\sum A^{p\bar{p}},\\ L(u_{x_1x_1})&= \left( \log \tilde{f}\right) _{x_1x_1}+A^{p\bar{p}}A^{q\bar{q}} |X(\zeta _p\bar{\zeta }_q-\left[ \zeta _p,\bar{\zeta }_q\right] ^{0,1}) u|^2\\&\quad +A^{p\bar{p}} \left( \zeta _p\left[ \bar{\zeta }_p,X\right] Xu +\left[ \zeta _p,X\right] \bar{\zeta }_pXu + X\zeta _p\left[ \bar{\zeta }_p,X\right] u + X\left[ \zeta _p,X\right] \bar{\zeta }_pu\right. \\&\left. \quad + XX\left[ \zeta _p,\bar{\zeta }_p\right] ^{0,1}u\right) \\&= (\log \tilde{f})_{x_1x_1}+A^{p\bar{p}}A^{q\bar{q}}|X\left( \zeta _p\bar{\zeta }_q-\left[ \zeta _p,\bar{\zeta }_q\right] ^{0,1}\right) u|^2\\&\quad +A^{p\bar{p}}\left( \left[ \zeta _p,\left[ \bar{\zeta }_p,X\right] \right] Xu+X[\zeta _p, [\bar{\zeta }_p,X]]u + \left[ X,\left[ \bar{\zeta }_p,X\right] \right] \zeta _pu\right. \\&\left. \quad + \left[ X,[\zeta _p,X]\right] \bar{\zeta }_pu\right) +A^{p\bar{p}}\left( X\left[ X,\left[ \zeta _p,\bar{\zeta }_p\right] ^{0,1}\right] u +[X,[\zeta _p,\bar{\zeta }_p]^{0,1}]Xu\right) \\&\ge -\,C(\mu +1)\sum A^{p\bar{p}} \end{aligned}$$

and we obtain

$$\begin{aligned} L(\mu )\ge -C\mu \sum A^{p\bar{p}}. \end{aligned}$$

Now we come to the last term

$$\begin{aligned}&-2A^{p\bar{p}}\frac{\left( \Psi e^{K|\nabla u|^2}\right) _p\left( \Psi e^{K|\nabla u|^2}\right) _{\bar{p}}}{\Psi e^{K|\nabla u|^2}}\\&\quad =-2A^{p\bar{p}}e^{K|\nabla u|^2}\left( \frac{|\Psi _p|^2}{\Psi }+2K\mathrm{Re}(\Psi _p(|\nabla u|^2)_{\bar{p}})+K^2\Psi (|\nabla u|^2)_{\bar{p}}(|\nabla u|^2)_{ p}\right) \\&\quad \ge -2e^{K|\nabla u|^2}A^{p\bar{p}} \left( \frac{|\Psi _p|^2}{\Psi }+CK|\Psi _p|\sum _k(|u_{p\bar{k}}|+|u_{pk}|)+K^2\sum _k(|u_{p\bar{k}}|^2+|u_{pk}|^2)\right) . \end{aligned}$$

Therefore there is a constant \(C_0>1\) which is under control such that

$$\begin{aligned} L(v)&\ge \mu e^{K|\nabla u|^2} A^{p\bar{p}} \left( \Psi _{p\bar{p}} + \Psi \left( \frac{1}{2} K-C_0K^2\right) \sum _k(|u_{p\bar{k}}|^2+|u_{pk}|^2)-2\frac{|\Psi _p|^2}{\Psi }\right) \\&\quad -\,C_0\mu e^{K|\nabla u|^2} A^{p\bar{p}} (1+K)(1+|\Psi _p|)\left( 1+\sum _k\left( |u_{p\bar{k}}|+|u_{pk}|\right) \right) . \end{aligned}$$

For a proper choice of \(K\) and \(\psi (K=\frac{1}{2C_0}, \psi =e^{A\phi }+4e^{A}\) where \(\frac{1}{2}<\phi <1\) is strictly plurisubharmonic in a neighborhood of \(\bar{\Omega }\) and \(A\) is large enough) we can conclude that \(L(v)>0\) and this is a contradiction with the maximality of \(v\). \(\square \)

4 Maximal plurisubharmonic functions

We say that a function \(u\in \mathcal {PSH}(\Omega )\) is maximal if for every function \(v\in \mathcal {PSH}(\Omega )\) such that \(v\le u\) outside a compact subset of \(\Omega \) we have \(v\le u\) in \(\Omega \).

Now we want to find the solution to the following Dirichlet problem:

$$\begin{aligned} \left\{ \begin{array}{l} u\in \mathcal {PSH}(\Omega )\cap \mathcal {C} (\bar{\Omega })\\ u \text{ is } \text{ maximal } \\ u=\varphi \; \text{ on } \;\partial \Omega , \end{array} \right. \end{aligned}$$
(4.1)

where \(\Omega \) is a strictly pseudoconvex domain of class \(\mathcal {C}^{1,1}\) and \(\varphi \) is a continuous function on \(\partial \Omega \).

Proposition 4.1

If \(\varphi \in \mathcal {C}^{1,1}(\bar{\Omega })\), then there is a unique solution \(u\in \mathcal {C}^{0,1}(\bar{\Omega })\) of the problem (4.1) and

$$\begin{aligned} \Vert u\Vert _{\mathcal {C}^{0,1}(\bar{\Omega })}\le C=C\left( \Vert \rho \Vert _{\mathcal {C}^{0,1}(\Omega )},m(\rho ),\Vert \varphi \Vert _{ \in \mathcal {C}^{1,1}}\right) . \end{aligned}$$

Proof

The uniqueness is a consequence of the definition.

To prove the existence assume that \(\rho \) is smooth. There are an increasing sequence \(\varphi _k\) of smooth functions such that \(\varphi _k\) tends to \(\varphi \) in \(\mathcal {C}^{1,1}\) norm. By Theorem 1 there is a solution \(u_k\) of the following Dirichlet Problem

$$\begin{aligned} \left\{ \begin{array}{l} u_k\in \mathcal {PSH}(\Omega )\cap \mathcal {C}^\infty (\bar{\Omega })\\ (i\partial \bar{\partial }u_k)^n=\frac{1}{k^n}(i\partial \bar{\partial }\rho )^n \; \text{ in } \;\Omega \\ u_k=\varphi _k\; \text{ on } \;\partial \Omega . \end{array} \right. \end{aligned}$$

By Lemma 3.3 \(\Vert u_k\Vert _{\mathcal {C}^{0,1}(\bar{\Omega })}\le C(\Vert \rho \Vert _{\mathcal {C}^{0,1}(\Omega )},m(\rho ),\Vert \varphi \Vert _{ \in \mathcal {C}^{1,1}})\). Now we can put

$$\begin{aligned} u:=\lim _{k\rightarrow \infty } u_k. \end{aligned}$$

It is enough to show that \(u\) is a maximal function. Let a function \(v\in \mathcal {PSH}(\Omega )\) be smaller than \(u\) outside a compact subset of \(\Omega \). From the comparison principle (Proposition 2.2) we obtain

$$\begin{aligned} v+\frac{\rho }{k}-\sup _{{\partial \Omega }}(\varphi -\varphi _k)\le u_p \end{aligned}$$

for \(p\ge k\). Taking the limit we conclude that \(v\le u\) in \(\Omega \).

In the general case we can assume that \(\varphi \) is a plurisubharmonic function on \(\Omega \) (by adding \(A\rho \) for \(A\) enough large). We can approximate \(\Omega \) by an increasing sequence of smooth strictly pseudoconvex domains \(\Omega _k\) such that \(\bigcup _k\Omega _k=\Omega \) and \(\Vert \rho \Vert _{\mathcal {C}^{0,1}(\Omega )},m(\rho )\) are under control, where \(\rho _k\) are strictly plurisubharmonic smooth defining functions for \(\Omega _k\). Let \(u_k\) be a solution of the following Dirichlet Problem

$$\begin{aligned} \left\{ \begin{array}{l} u_k\in \mathcal {PSH}(\Omega _k)\cap \mathcal {C} (\bar{\Omega }_k)\\ u_k \text{ is } \text{ maximal } \\ u_k=\varphi \; \text{ on } \;\partial \Omega _k. \end{array}\right. \end{aligned}$$

Then \(u_k\ge \varphi \), hence it is an increasing sequence and again we can put

$$\begin{aligned} u:=\lim _{k\rightarrow \infty } u_k. \end{aligned}$$

If \(v\) is as above, for every \(\varepsilon >0\) we have \(v-\varepsilon \le u_k\) outside a compact set for \(k\) large enough. So we obtain \(v\le u\) and conclude that \(u\) is a maximal function as in the statement. \(\square \)

Note that in the above proposition it is not enough to assume that \(\varphi \) is \(\mathcal {C}^{1,\alpha }\) regular for some \(\alpha <1\). Indeed, one can show that if \(\Omega \) is strictly pseudoconvex, \(P\in {\partial \Omega }\) and \(\varphi (z)\le \varphi (P)-(\mathrm{dist}(z,p))^{1+\alpha }\), then a solution of (4.1) is not Hölder continuous with the exponent greater than \(\frac{1+\alpha }{2}\).

Theorem 4.2

(Harvey, Lawson [4]) There is a unique solution \(u\) of the problem (4.1).

Proof

Let \(\varphi _k \) be an increasing sequence of smooth functions on \(\bar{\Omega }\) such that \(\lim _{k\rightarrow \infty }\varphi _k=\varphi \). By Proposition 4.1 there is a sequence \(u_k\) of solutions of (4.1) with boundary conditions \(\varphi _k\) (instead of \(\varphi \)). Because

$$\begin{aligned} u_k\le u_p\le u_k+\sup _{{\partial \Omega }}(\varphi -\varphi _k) \end{aligned}$$

for \(p\ge k\), the sequence \(u_k\) is a Cauchy sequence in \(\mathcal {C}(\bar{\Omega })\). Similar as in the previous proof we can conclude that its limit \(u\) is a solution of the problem (4.1). \(\square \)

Note that we can also prove the above theorem directly from Theorem 1.

The following proposition shows that being a continuous maximal plurisubharmonic function is a local property.

Proposition 4.3

Let \(\Omega \subset M\) and \(u\in \mathcal {PSH}(\Omega )\). Then

  1. (i)

    If \(u\) is maximal then \(u|_U\) is maximal for every \(U\subset \Omega \);

  2. (ii)

    If \(\Omega \) is such that there is a bounded strictly plurisubharmonic function \(\rho \in \mathcal {C}^2(\Omega ),\,u\) is continuous and every point in \(\Omega \) has a neighbourhood \(U\) such that \(u|_U\) is maximal, then \(u\) is maximal.

Proof

  1. (i)

    Suppose that \(v\in \mathcal {PSH}(U)\) is such that \(v\le u\) outside a compact subset of \(U\). Then \(\max \{u,v\}\in \mathcal {PSH}(\Omega )\) and we obtain \(v\le \max \{u,v\}\le u\) on \(U\).

  2. (ii)

    We can assume that \(\rho <0\). Let \(\varepsilon >0,\,v\in \mathcal {PSH}(\Omega )\) and let \(z_0\in \Omega \) be a point where a function \(v+\varepsilon \rho -u\) attains its maximum. By (i) there is a strictly pseudoconvex domain \(\tilde{\Omega }\subset \Omega \) with a smooth plurisubharmonic defining function \(\tilde{\rho }\) such that \(z_0\in \tilde{\Omega }\) and \(u|_{\tilde{\Omega }}\) is maximal. Note that there is \(\tilde{\varepsilon }>0\) such that a function \(\rho -\tilde{\varepsilon }\tilde{\rho }\) is plurisubharmonic in some neighbourhood of \(\mathrm{cl}(\tilde{\Omega })\). Hence a function \(\tilde{v}=\max \{v+\varepsilon \rho ,v+\varepsilon (\rho -\tilde{\rho } \tilde{\varepsilon })\}\) is also plurisubharmonic and \(\tilde{v}-u\) attains a maximum only in some compact subset of \(\tilde{\Omega }\), which is impossible because \(u|_{\tilde{\Omega }}\) is maximal. As \(\varepsilon \) and \(v\) were arbitrary we can conclude that the function \(u\) is maximal. \(\square \)

In [4] the authors consider problem (4.1) for \(\mathcal {F(J)}\)-harmonic functions which they define in a different way than we define maximal functions but we will see that these concepts agree.

Let \(\Omega \subset M\) and \(u\in \mathcal {C\cap PSH}(\Omega )\). We say that \(u\) is \(\mathcal {F(J)}\)-harmonic if for every \(U\subset \Omega \) and for every smooth strictly plurisubharmonic function \(\phi \le u\) on \(U\) we have \(\phi <u\) on \(U\). One can show (using the comparison principle) that \(\mathcal {C}^2\,\mathcal {F(J)}\)-harmonic functions are exactly \(\mathcal {C}^2\) solutions of (1) with \(f=0\).

Proposition 4.4

Let \(\Omega \) and \(u\) be as above. Then

  1. (i)

    If \(u\) is maximal then \(u\) is \(\mathcal {F(J)}\)-harmonic;

  2. (ii)

    If \(\Omega \) is such that there is a bounded strictly plurisubharmonic function \(\rho \in \mathcal {C}^2(\Omega )\) and \(u\) is \(\mathcal {F(J)}\)-harmonic, then \(u\) is maximal.

Proof

The first assertion follows from definitions. To proof (ii) we can assume, by Proposition 4.3, that \(\Omega \) is a smooth strictly pseudoconvex domain with defining function \(\rho \) such that \(u\in \mathcal {C}(\bar{\Omega })\). Let \(\varepsilon >0\). By Theorem 4.2 there is a continuous maximal plurisubharmonic function \(u_0\) equal to \(u\) on \({\partial \Omega }\). By Theorem 1 there is a smooth strictly plurisubharmonic function \(u_1\) such that \(u-\varepsilon <u_1<u\) on a boundary and

$$\begin{aligned} (i\partial \bar{\partial }u_1)^n = \frac{1}{2} \varepsilon ^n(i\partial \bar{\partial }\rho )^n. \end{aligned}$$

Then using the comparison principle (Proposition 2.2) we obtain

$$\begin{aligned} u_0+\varepsilon \rho -\varepsilon \le u_1\le u\le u_0 \end{aligned}$$

and thus we get \(u=u_0\). \(\square \)