1 Introduction

The quaternionic Monge–Ampère equation on compact HKT manifolds was introduced by Alesker and Verbitsky in [AV10]. On a general hyperhermitian manifold (MIJKg) of quaternionic dimension n it takes the form

$$\begin{aligned} \begin{aligned}&(\Omega + \partial \partial _J\phi )^n = e^f \Omega ^n, \\&\Omega + \partial \partial _J\phi > 0. \end{aligned} \end{aligned}$$
(1.1)

The goal of this paper is to prove the higher order estimates for the quaternionic Monge–Ampère equation (1.1) on compact hyperKähler manifolds. The highlight of our result is that we do not assume any flatness or additional integrability of the underlying hypercomplex structure as was always the case for the higher order estimates known so far. The hyperKähler condition, as will be seen during the derivation of the estimates, may in turn be interpreted as a curvature condition which we hope can be removed. As a result we solve this equation on any compact hyperKähler manifold, cf. Remark 6.1.

Our main result is stated as follows:

Theorem 1.1

Let (MIJKg) be a compact, connected hyperKähler manifold. For any real function \(f \in C^\infty (M)\) there exists a unique, up to the addition of a constant, smooth solution \(\phi \) to the quaternionic Monge–Ampère equation (1.1) provided the necessary normalization condition

$$\begin{aligned} \int _M e^f \Omega ^n \wedge \overline{\Omega }^n = \int _M \Omega ^n \wedge \overline{\Omega }^n \end{aligned}$$
(1.2)

is satisfied.

Equation (1.1) was motivated by the prior research in the local case, cf. [A03], and the attempt to prove the analog of the Calabi conjecture in quaternionic geometry, cf. [AV10, V09], as we explain in the next section. It was conjectured in [AV10] that Eq. (1.1) can always be solved at least in the case when the canonical bundle of the HKT manifold is trivial holomorphically. Given the recent progress in solving Calabi–Yau type equations for non Kähler metrics, cf. [GL10, TW10b, SzTW17, TW17, Sz18, TW19], this is expected to hold even without the latter assumption, cf. [AS17, Sr19]. To sum up Theorem 1.1 constitutes the ansatz for proving the following conjecture.

Conjecture 1.2

[AV10] Let (MIJKg) be a compact, connected HKT manifold. Suppose that there exists a non vanishing I-holomorphic (2n, 0) form \(\Theta \) on M. For every real smooth function f the quaternionic Monge–Ampère equation (1.1) admits a unique, up to the constant, smooth solution provided the function f satisfies the necessary condition

$$\begin{aligned} \int _M (e^f-1) \Omega ^n \wedge \overline{\Theta }=0.\end{aligned}$$
(1.3)

Let us just mention that the name for the Eq. (1.1) is justified as follows. Consider the quaternionic variables

$$\begin{aligned} q_i=x_{4i}+x_{4i+1} \mathfrak {i}+ x_{4i+2} \mathfrak {j}+ x_{4i+3} \mathfrak {k}, \end{aligned}$$
(1.4)

in the flat quaternionic space \(\mathbb {H}^n\) of quaternionic dimension n. There are the so called Cauchy–Riemann–Fueter derivatives

$$\begin{aligned}{} & {} \frac{\partial \phi }{ \partial \bar{q}_{\alpha }} = \frac{\partial \phi }{\partial x_{4\alpha }} + \mathfrak {i}\frac{\partial \phi }{\partial x_{4\alpha +1}} + \mathfrak {j}\frac{\partial \phi }{\partial x_{4\alpha +2}} + \mathfrak {k}\frac{\partial \phi }{\partial x_{4\alpha +3}}, \end{aligned}$$
(1.5)
$$\begin{aligned}{} & {} \frac{\partial \phi }{ \partial q_{\alpha }}= \frac{\partial \phi }{\partial x_{4 \alpha }} - \frac{\partial \phi }{\partial x_{4 \alpha +1}} \mathfrak {i}- \frac{\partial \phi }{\partial x_{4 \alpha +2}} \mathfrak {j}- \frac{\partial \phi }{\partial x_{4 \alpha +3}} \mathfrak {k}. \end{aligned}$$
(1.6)

It may be checked, cf. [AV06] or [Sr18] for an elementary calculation, that in this case

$$\begin{aligned} \big (\partial \partial _J\phi \big )^n = \det \Bigg [ \frac{\partial ^2 \phi }{ \partial q_\alpha \partial \bar{q_{\beta }}} \Bigg ]_{\alpha , \beta } \Theta , \end{aligned}$$
(1.7)

where \(\Theta \) is the canonical trivialization of \(K_I(\mathbb {H}^n)\) and the determinant has to be understood in a proper way—as the Moore determinant, cf. [M22], of the hyperhermitian matrix. The Dirichlet problem for the operator (1.7) was first considered by Alesker in [A03] where continuous solutions were found for continuous right hand sides. After that, the problem was solved in the smooth category by Zhu [Z17]. Later, thanks to the form of (1.7), the pluripotential approach was taken up resulting in providing continuous solutions even for right hand sides in \(L^p\) spaces. For \(p \ge 4\) it is due to Wan, cf. [W20], and for \(p>2\) due to the second named author, cf. [Sr18]. In the latter case the exponent was proven to be optimal. This equation is also covered by the very general approach taken up in the last 2 decades by Harvey and Lawson. They provide viscosity solutions for the Dirichlet problem even for domains in quaternionic manifolds, cf. [HL09, HL11, HL20].

Coming back to the advances towards proving Conjecture 1.2. Generally speaking the difficulties with obtaining a priori estimates for the Eq. (1.1) are caused, as we explain in depth in the next section, by the fact that a generic hypercomplex structure locally is not the pull back of the flat structure from \(\mathbb {H}^n\). The lack of quaternionic coordinates forces one to work in the general case, at best, with holomorphic coordinates for the reference complex structure I. But this in turn results in Eq. (1.1) depending not only on the coefficients of the metric tensor but also of the endomorphism field J since, in the holomorphic coordinates for I,

$$\begin{aligned} \Omega + \partial \partial _J\phi := \Omega ^\phi _{ij} dz_i \otimes dz_j= \Big ( -(g_{i \bar{k}} + \phi _{i \bar{k}})J^{\bar{k}}_j + ( g_{j \bar{k}} + \phi _{j \bar{k}})J^{\bar{k}}_i \Big ) dz_i \otimes dz_j.\end{aligned}$$
(1.8)

Proving estimates in such coordinates is similar to solving the complex Monge–Ampère equation by performing the calculations in generic real coordinates. Another drawback is that Eq. (1.1), instead of being of the form (1.7), is an equation on the Pfaffian of the coefficients of the two form \(\Omega + \partial \partial _J\phi \) in the complex coordinates

$$\begin{aligned} {\text {Pf }}\big [\Omega ^\phi _{ij}\big ]_{i,j} = e^f \cdot {\text {Pf }}\big [\Omega _{ij}\big ]_{i,j}. \end{aligned}$$
(1.9)

As will be seen in the calculations, the more times the Eq. (1.9) is differentiated the more the lack of quaternionic derivatives becomes an issue.

In the paper [AV10], the \(C^0\) estimates for the solutions to the Eq. (1.1) were derived by the classical Moser iteration method in the setting exactly as in Conjecture 1.2. There the existence of the holomorphic trivialization is crucial for the argument to work. Later it was shown that the \(C^0\) estimate holds for (1.1) even without the assumption on the holomorphic triviality of \(K_I(M)\), cf. [AS17, Sr19], but the methods are much more involved. As for the higher order estimates, as we mentioned, they have been derived only under the assumption that the hypercomplex structure is integrable in the strong sense, i.e. is locally flat. In addition, either the initial metric should be very special or the manifold has to admit in addition a hyperKähler metric compatible with the flat hypercomplex structure. Precisely, in [A13] it was shown that in the case the manifold is a torus or its quotient endowed with the flat hyperKähler metric (this implies in particular the flatness of the hypercomplex structure), the Laplacian bound on the solution of (1.1) holds. Alesker proved also that the analogue of the Evans–Krylov theorem, cf. [E82], holds under the assumption of flatness of the hypercomplex structure. In [GV21] the authors show the Laplacian bound, unlike in Alesker’s result depending on the gradient bound, for Eq. (1.1) on certain eight dimensional nilmanifolds endowed with the flat hypercomplex structures and torus action invariant initial HKT metrics. Shortly before this preprint was written down the preprint [BGV21] appeared on arXiv. The authors take up there the parabolic approach for the Eq. (1.1) but the assumption under which they are able to prove the convergence of the flow to the solution are exactly as in Alesker’s paper [A13].

In the current note we carry out the computations for all the higher order estimates in geodesic coordinates for the Obata connection. This seems to be the main technical input which allows us to overcome the issues coming from the dependence of the Eq. (1.1) on the second complex structure J. This is still not enough though to deal with the dependence on the metric. Because of that the hyperKähler assumption appears.

Our strategy in this paper is as follows. First of all, motivated by an influential idea of Błocki from [B09] (see also [Gu]), we show, cf. Theorem 3.1, the gradient estimate in the setting of Theorem 1.1. Here, the hyperKähler condition plays similar role as non negativity of the holomorphic bisectional curvature of the background Kähler metric in the case of the complex Monge–Ampère equation. This is somewhat surprising since the hyperKähler metric is not necessarily of such a curvature. In the general hypercomplex case the gradient estimate does not seem to follow from arguments as in [B09]. The troubles in this case are caused partially by the failure of the Leibniz rule for the operators (1.5) and (1.6). Let us mention that (direct) gradient estimate is not known for general complex Hessian equations. We use Theorem 3.1 in Sect. 5 to derive, with the aid of Theorem 4.1, the full \(C^2\) estimate. In this case the proof is standard and relies on an idea of Błocki from [B11]. The crucial result is Theorem 4.1 proven in Sect. 4 which gives the bound on the Laplacian, or equivalently the quaternionic Hessian \(\partial \partial _J\phi \), for the solutions of (1.1). Let us remark that for general HKT metric a major issue is how to handle third order terms. Roughly speaking the positive term appearing, being formed by the squares of sums of certain third order derivatives, compensates only half of the negative term. At the moment we do not know how to deal with this difficulty and the classical methods of [Y78, B09, GL10] does not seem, cf. Remark 4.2, to work. The hyperKähler assumption allows us to get rid of some terms coming from differentiating the metric coefficients and in turn, by considering a more general perturbation than the classical one in the Pogorelov approach for the Laplacian bound, cf. [Y78, B09, GL10], we are able to ignore the negative term.

As we explain in details in the next section, Theorem 1.1 allows one to draw some conclusions concerning the HKT geometry of the underlying hyperKähler manifold. First of all having a hypercomplex manifold (MIJK) any hyperhermitian metric g provides a canonically associated smooth section of the canonical bundle \(K_I(M)\) via the map

$$\begin{aligned} g \longmapsto \Omega ^n.\end{aligned}$$
(1.10)

Sections obtained in (1.10) satisfy the properties, defined rigorously in the next section, which we call positivity and J-realness.

Proposition 1.3

Let (MIJKg) be a compact, connected hyperKähler manifold. Given any positive and J-real trivialization \(\Theta \), i.e. of the form \(e^f\Omega ^n\) for some smooth function f, of the canonical bundle \(K_I(M)\) there is an HKT metric \(\hat{g}\) such that the associated HKT form \(\hat{\Omega }\) satisfies

$$\begin{aligned} \hat{\Omega }^n = \Theta . \end{aligned}$$
(1.11)

What is more the metric \(\hat{g}\) may be chosen so that the associated HKT form is of the form

$$\begin{aligned} a \big (\Omega + \partial \partial _J\phi \big ) \end{aligned}$$

for a smooth function \(\phi \) and a positive constant a.

From Proposition 1.3 it is easy to obtain the so called Calabi–Yau type theorem for HKT metrics, yet only on hyperKähler manifolds.

Remark 1.4

Let (MIJKg) be a compact, connected hyperKähler manifold. Given any positively oriented volume form \(\sigma \) on (MI) there is an HKT metric \(g_\phi \), such that the associated HKT form is \(a \big ( \Omega + \partial \partial _J\phi \big )\), for a smooth function \(\phi \) and a positive constant a, such that

$$\begin{aligned} \omega _{I,g_\phi }^{2n} = \sigma . \end{aligned}$$
(1.12)

A classical calculation allows us to obtain the following as a result of Remark 1.4.

Proposition 1.5

Let (MIJKg) be a compact, connected hyperKähler manifold. Given any representative

$$\begin{aligned} \rho \in c_1^{BC}(M,I) \end{aligned}$$

of the first Bott–Chern class of (MI), see [T15], there is an HKT metric \(g_\phi \), whose associated HKT form is \(\Omega + \partial \partial _J\phi \), such that

$$\begin{aligned} Ricc\big (\nabla ^{Ch}_{I,g_\phi }\big ) = \rho .\end{aligned}$$
(1.13)

In (1.13) the symbol \(Ricc(\nabla ^{Ch}_{I,g_\phi })\) denotes the well known Chern–Ricci form associated to the Chern connection of the hermitian structure \((I, g_\phi )\) on M.

As the non constant conformal deformation of an HKT metric is never an HKT metric the last three results stated above are non trivial, even under the hyperKähler assumption. Provided one can remove the extra assumption on the initial metric the Calabi–Yau theorem for yet another class of metrics would be settled. Proving the Calabi–Yau type theorem for different class of hermitian (non Kähler) metrics was a subject of an intense study in the last decade. In the classical case of Kähler metrics it is known due to Yau [Y78]. For the class of Gauduchon metrics it was settled only recently in [SzTW17], building on [Sz18, TW19], confirming in turn an old conjecture of Gauduchon from ’80 s. Actually the same result was proven there for strongly Gauduchon metrics as well. In [TW17] a Calabi–Yau type theorem, Corollary 1.3 in there, for balanced metrics was proven, yet like in our case, under an extra assumption that the manifold admits a Kähler metric. In this case the result follows of course from the Calabi–Yau theorem for Kähler metrics as well. The assumption of admitting Kähler metric was later relaxed to admitting merely an Astheno–Kähler one, cf. [SzTW17], in which case the classical Calabi–Yau theorem can not be applied. It is still an open problem whether the Calabi–Yau type theorem for the class of balanced metric holds in general, cf. [TW19, SzTW17].

Let us finish this introduction by the remark in the spirit of the mentioned Harvey and Lawson theory. Like in the local case, in the global situation the Eq. (1.1) is a companion of the real and complex Monge–Ampère equations which received great attention in the last century and, in certain forms, are related to the fundamental differential geometric problems.

The complex Monge–Ampère equation on complex n dimensional hermitian manifold (MIg) taking the form

$$\begin{aligned} (\omega + \mathfrak {i}\partial \overline{\partial }\phi )^n = e^f \omega ^n\end{aligned}$$
(1.14)

was solved, as we mentioned, on Kähler manifolds by Yau [Y78] and on general hermitian manifold by Tosatti and Weinkove [TW10a, TW10b], cf. also [GL10]. It was proven that it can be solved even on almost complex manifolds, cf. [ChTW19].

As for the real Monge–Ampère equation on a Riemannian manifold (Mg) of dimension n, its rough version does not carry a substantial geometric meaning. It is nevertheless interesting from an analytic point of view. In this mentioned rough form it can be written as

$$\begin{aligned} \det \big ( g + (\nabla ^{LC})^2 \phi \big ) = e^f \det g.\end{aligned}$$
(1.15)

In (1.15) the meaning of taking the determinant is that the symmetric bilinear forms \(g + (\nabla ^{LC})^2 \phi \) and g are treated, via the metric, as the endomorphisms of the tangent bundle and we take the determinant of this endomorphisms. These equations were treated for example in Li [L90], where they were solved under the non negative curvature assumption on g. This assumption was later removed by Urbas in [U02]. Let us just mention that in case of this equation it is hard to obtain an easy normalization of f’s for which the equation can be solved.

The very special modification of this equation was treated earlier by Cheng and Yau in [ChY82]. They considered affine manifolds, i.e. smooth real manifolds endowed with a flat torsion free connection, admitting a Riemannian metric, which they called affine Kähler metric, being locally given as the Hessian of a potential function in affine coordinates. The equation they considered is obtained by taking the affine Kähler metric in (1.15) and by exchanging the Levi–Civita connection there for the affine connection.

2 Preliminaries

In this section we introduce the notation and collect basic facts concerning hyperhermitian manifolds. We also prove technical or computational in nature results needed for the a priori estimates of Sects. 35 in order to make the presentation there more straightforward.

2.1 Hypercomplex geometry.

We denote by

$$\begin{aligned} \mathbb {H}= \{ x_0+x_1\mathfrak {i}+x_2\mathfrak {j}+x_3\mathfrak {k}\, | \, x_0,x_1,x_2,x_3 \in \mathbb {R}\} \end{aligned}$$

where \(\mathfrak {i}^2= \mathfrak {j}^2= \mathfrak {k}^2=-1\) and \(\mathfrak {i}\mathfrak {j}\mathfrak {k}= -1\) the field of quaternions with addition and multiplication being defined in the standard way. We consider \(\mathbb {H}^n\) as a right \(\mathbb {H}\) vector space. Let us recall what have become the standard definition.

Definition 2.1

For a manifold M of the real dimension 4n endowed with a triple of complex structures I, J, K satisfying the quaternion relation

$$\begin{aligned} I \circ J \circ K = -id_{TM} \end{aligned}$$

the tuple (MIJK) is called a hypercomplex structure.

A hypercomplex manifold admits many complex structures in particular the ones given

$$\begin{aligned} S_M= \{aI+bJ+cK \, | \, a^2+b^2+c^2=1 \} \end{aligned}$$

(the so called twistor sphere).

Remark 2.2

We warn the reader that for us, in the whole text, endomorphisms act from the right on the tangent space. This convention is compatible with the one usually taken up in papers on hypercomplex geometry.

In that case each tangent space \(T_x M\), for \(x \in M\), becomes a right \(\mathbb {H}\)-vector space where multiplication by \(\mathfrak {i}\), \(\mathfrak {j}\) and \(\mathfrak {k}\) is given by \(I_x\), \(J_x\) and \(K_x\) respectively.

Clearly the structure group of the hypercomplex manifold is reduced to \(Gl_n(\mathbb {H})\) and vice versa each such reduction induces the almost complex structures I, J and K as in the definition above. The condition of the almost complex structures being integrable is though not equivalent to the induced \(Gl_n(\mathbb {H})\) structure being integrable in the strong sense of differential geometry, i.e locally I, J and K are not pull backs of the standard hypercomplex structure induced by \(\mathfrak {i}\), \(\mathfrak {j}\) and \(\mathfrak {k}\) in \(\mathbb {H}^n\). In case the latter condition is satisfied such structures were studied in [S75]. The latter is also equivalent to the existence of an atlas whose transition functions are affine maps with the endomorphism parts belonging to \(Gl_n(\mathbb {H})\). On the bright side the integrability of the almost complex structures I, J and K implies the 0-integrability of the induced \(Gl_n(\mathbb {H})\) structures (the reverse implication holds due to the Newlander–Nirenberg theorem), i.e. the existence of the \(Gl_n(\mathbb {H})\) compatible torsion free connection. This is a non obvious result of Obata. Strong integrability of the \(Gl_n(\mathbb {H})\) structure is equivalent to the Obata connection being in addition flat.

Theorem 2.3

[Ob56] For a hypercomplex manifold (MIJK) there exists a unique torsion free connection, denoted by \(\nabla ^{Ob}\), such that

$$\begin{aligned} \nabla ^{Ob}I=\nabla ^{Ob}J=\nabla ^{Ob}K=0.\end{aligned}$$

The coordinate expression for this connection can be found in [Ob56]. An invariant global formula can be found for example in Gauduchon’s paper [G97b]. The existence of this connection will be very important for the technical results stated below and for the computations involved in deriving the a priori estimates for the Eq. (1.1).

Remark 2.4

For a more detailed discussion on quaternionic geometry one can refer to the two excellent papers [AM96] and [S86].

When considering a hypercomplex manifold (MIJK) we obtain that

$$\begin{aligned} \mathbb {H}\cong \{a id_{TM} + b I +cJ +dK \, | \, a,b,c,d \in \mathbb {R}\} \subset End(M)\end{aligned}$$

acts from the right on TM and from the left on \(T^*M\) or more generally from the left on differential forms. The convention we use for the latter action is against the commonly used. Namely, given any field of endomorphisms L on TM, acting according to our convention from the right, we define its left action on the space of complex valued smooth differential forms by

$$\begin{aligned} L: \Lambda ^k_\mathbb {C}(M) \ni \alpha \longmapsto \alpha (\cdot L,\ldots , \cdot L) \in \Lambda ^k_\mathbb {C}(M).\end{aligned}$$

Remark 2.5

From now on, whenever it happens that on a hypercomplex manifold (MIJK) we do not specify with respect to which complex structure the Hodge bidegree is taken, it is taken with respect to I.

Let (MIJK) be a hypercomplex manifold. Let us remind that we have the Dolbeault operators

$$\begin{aligned} \partial :=\partial _I \text { and } \overline{\partial }:=\overline{\partial _I}\end{aligned}$$

associated to the complex structure I on M. We are going to introduce the quaternionic analogue of the \(\overline{\partial }\) operator, or rather \(d^c_I:= I^{-1} \circ d \circ I\). In this we follow Verbitsky, cf. [V02], who defined the differential operator \(\partial _J\) by

$$\begin{aligned} \partial _J:=J^{-1} \circ \overline{\partial } \circ J. \end{aligned}$$
(2.1)

Since the operator J acts on the complex forms by exchanging the bidegree components

$$\begin{aligned} J: \Lambda ^{p,q}_I(M) \longrightarrow \Lambda ^{q,p}_I(M) \end{aligned}$$
(2.2)

the operator \(\partial _J\) acts on this forms by

$$\begin{aligned} \partial _J: \Lambda ^{p,q}_I(M) \rightarrow \Lambda ^{p+1,q}_I(M).\end{aligned}$$
(2.3)

We also introduce the operator \(\overline{\partial _J}\) defined formally again by twisting

$$\begin{aligned} \overline{\partial _J}:= J^{-1} \circ \partial \circ J, \end{aligned}$$

but as the operator J is real it is equal to \(\overline{(\partial _J)}\) as well.

It was observed by Verbitsky, [V02, V07a], that the bicomplex

$$\begin{aligned} \Big (A^{p,q}:=\Lambda ^{p+q,0}_I(M),\partial , \partial _J\Big ),\end{aligned}$$

called by him the quaternionic Dolbeault bicomplex, not only resembles the Dolbeault bicomplex

$$\begin{aligned} \Big (\Lambda ^{p,q}, \partial , \overline{\partial }\Big ) \end{aligned}$$

but it is also isomorphic to the so called Salamon complex, cf. [S86], introduced by Salamon in the broader context of quaternionic manifolds (see [V07a] for details).

For the further reference we would like to introduce the notion of J-realness. This is done as follows. The composition of the operator (2.2) with the bar operator is an involution on \(\Lambda ^{p,q}_I(M)\) if \(p+q\) is even. The bundle of fixed points for this endomorphism in \(\Lambda ^{2k,0}_I (M)\) was denoted in [V10] by \(\Lambda ^{2k,0}_{I,\mathbb {R}} (M)\). In short, \(\alpha \in \Lambda ^{2k,0}_{I,\mathbb {R}} (M) \) if and only if

$$\begin{aligned} J \alpha = \overline{\alpha } \end{aligned}$$

and such a from is called J-real.

2.2 Hyperhermitian metrics.

Definition 2.6

A Riemannian metric g on a hypercomplex manifold (MIJK) is called hyperhermitian if it is hermitian with respect to I, J and K. For a hyperhermitian manifold (MIJKg) and any \(L \in S_M\) we denote the associated hermitian form by

$$\begin{aligned} \omega _L(X,Y)=g(X L,Y) \end{aligned}$$
(2.4)

for \(X,Y \in \Gamma (TM)\). We define also the associated hyperhermitian form

$$\begin{aligned} \Omega = \omega _J - \mathfrak {i}\omega _K. \end{aligned}$$
(2.5)

It is elementary to check that \(\Omega \in \Lambda ^{2,0}_{I,\mathbb {R}}(M)\). Let us elaborate on the role of \(\Omega \) in encoding the hyperhermitian metric g, cf. [H90] Chapter 2, Lemma 2.72. Suppose (Vg) is a right \(\mathbb {H}\)-vector space with a hyperhermitian inner product g. Such inner products correspond bijectively to the hyperhermitian sesquilinear forms

$$\begin{aligned} H = g + \mathfrak {i}\omega _I + \mathfrak {j}\omega _J + \mathfrak {k}\omega _K. \end{aligned}$$
(2.6)

By introducing

$$\begin{aligned} \overline{h} = g + \mathfrak {i}\omega _I \end{aligned}$$
(2.7)

and \(\Omega \) as above we obtain

$$\begin{aligned} H = \bar{h} + \mathfrak {j}\Omega .\end{aligned}$$
(2.8)

What is more

$$\begin{aligned} \Omega = \overline{h}(\cdot \mathfrak {j}, \cdot ).\end{aligned}$$
(2.9)

Consequently \(\bar{h}\), and in turn also H, is completely determined by \(\Omega \) satisfying for any \(v,w \in V\)

$$\begin{aligned} \Omega (v \mathfrak {j},w \mathfrak {j}) = \overline{\Omega (v,w)},\end{aligned}$$

or after taking the complexification of V,

$$\begin{aligned} \Omega (\cdot \mathfrak {j}, \cdot \mathfrak {j}) = \overline{\Omega }\end{aligned}$$
(2.10)

and

$$\begin{aligned} \Omega (v,v \mathfrak {j}) = \overline{h}( v \mathfrak {j}, v \mathfrak {j}) \ge 0,\end{aligned}$$

or after taking the complexification,

$$\begin{aligned} \Omega (z,\bar{z} \mathfrak {j}) = \overline{h}( z \mathfrak {j}, \bar{z} \mathfrak {j}) \ge 0 \end{aligned}$$
(2.11)

for any \(z \in V^{1,0}_\mathfrak {i}\). The conditions (2.10) and (2.11) give the meaning of the inequality in (1.1).

Remark 2.7

In calculations it will be customary to assume that at the point of interest the hyperhermitian structure \((T_x M, I_x, J_x,K_x,g_x)\) is isomorphic to the standard model below, cf. (2.13) and (2.14). This can trivially be seen to be possible by taking the orthonormal basis of the form \(e_0\), \(e_0I\), \(e_0J\), \(e_0K\),..., \(e_{n-1}\), \(e_{n-1}I\), \(e_{n-1}J\), \(e_{n-1}K\) for \((T_x M, I_x, J_x,K_x,g_x)\). As was discussed in [Sr19] in the presence of a second hyperhermitian metric the first one may still be assumed to be standard while the second one being diagonal.

The right multiplications by \(\mathfrak {i}\), \(\mathfrak {j}\), and \(\mathfrak {k}\) act on \(\mathbb {H}^n\) defining the almost complex structures I, J and K respectively. Let us introduce, next to the real coordinates (1.4), the holomorphic coordinates, for the complex structure I, by decomposing

$$\begin{aligned} q_i=z_{2i}+\mathfrak {j}z_{2i+1} \end{aligned}$$
(2.12)

for \(i \in \{0,\ldots ,n-1\}\). As an easy calculation shows the action of J in this holomorphic coordinates is

$$\begin{aligned} \begin{aligned}&(\partial _{z_{j}})J=(-1)^{j}\partial _{\overline{z_{j+(-1)^j}}}, \\&J^{-1}(d z_{j})=(-1)^{j}d\overline{z_{j+(-1)^j}} \end{aligned} \end{aligned}$$
(2.13)

for \(j \in \{0,\ldots ,2n-1\}\). Take a standard inner product on \(\mathbb {H}^n\), in coordinates from (1.4),

$$\begin{aligned} g = dx_{4i} \otimes dx_{4i} + dx_{4i+1} \otimes dx_{4i+1}+ dx_{4i+2} \otimes dx_{4i+2} +dx_{4i+3} \otimes dx_{4i+3}. \end{aligned}$$

We easily get the following expressions for the quantities associated with this hyperhermitian structure \((\mathbb {H}^n, I,J,K,g)\)

$$\begin{aligned} \begin{aligned}&\omega _I=-dx_{4i+1} \otimes dx_{4i } + dx_{4i} \otimes dx_{4i+1} + dx_{4i+3} \otimes dx_{4i+2} - dx_{4i+2} \otimes dx_{4i+3} \\&\quad = dx_{4i} \wedge dx_{4i+1} + dx_{4i+3} \wedge dx_{4i+2}=\frac{\mathfrak {i}}{2}(dz_{2i} \wedge d\overline{z_{2i}} + dz_{2i+1} \wedge d \overline{z_{2i+1}}),\\&\omega _J=dx_{4i} \wedge dx_{4i+2} + dx_{4i+1} \wedge dx_{4i+3},\\&\omega _K=dx_{4i+2} \wedge dx_{4i+1} + dx_{4i} \wedge dx_{4i+3},\\&\Omega =\omega _J - \mathfrak {i}\omega _K =dx_{4i} \wedge dx_{4i+2} + dx_{4i+1} \wedge dx_{4i+3} - \mathfrak {i}dx_{4i+2} \wedge dx_{4i+1} -\mathfrak {i}dx_{4i} \wedge dx_{4i+3}\\&\quad = (dx_{4i} + \mathfrak {i}dx_{4i+1})\wedge (dx_{4i+2}-\mathfrak {i}dx_{4i+3})=dz_{2i} \wedge dz_{2i+1}. \end{aligned}\end{aligned}$$
(2.14)

Remark 2.8

One should note that the real coordinates introduced in (1.4) are not given by taking the real and imaginary part decomposition of the complex coordinates (2.12). The relation between these coordinates is

$$\begin{aligned} \begin{aligned}&z_{2j}=x_{4j}+ x_{4j+1} \mathfrak {i}, \\&z_{2j+1}= x_{4j+2} - x_{4j+3} \mathfrak {i}, \end{aligned} \end{aligned}$$

for \(j=0,\ldots ,n-1\).

Before introducing certain classes of hyperhermitian metric let us recall that in the paper [G97a] Gauduchon has distinguished in the affine space of all the hermitian connections, i.e. those satisfying

$$\begin{aligned} \nabla I = \nabla g = 0\end{aligned}$$

for a hermitian manifold (MIg), the affine line of the canonical connections. Among them two classical ones will be important for our presentation.

Proposition 2.9

[G97a] Let (MIg) be a hermitian manifold. There exists a unique hermitian connection, denoted by \(\nabla ^{Ch}_{I,g}\), which will be called the Chern connection characterized by the fact that

$$\begin{aligned} \big (\nabla ^{Ch}_{I,g} \big )^{0,1} = \overline{\partial }. \end{aligned}$$

There exists as well a unique hermitian connection, denoted by \(\nabla ^{B}_{I,g}\), which will be called the Bismut connection characterized by the fact that its torsion tensor, after lowering the upper index by g, is a three form.

Coming back to the hyperhermitian metric on hypercomplex manifolds, when considering the above connections associated to the hermitian structure (MLg) for any \(L \in S_M\) we add a subscript L, eg. \(\nabla ^{Ch}_{L,g}\). Let us recall the following definition:

Definition 2.10

A hyperhermitian metric g on (MIJK) is called hyperKähler, HK for short, if any of the equivalent conditions are satisfied

  • \(d\omega _I=d \omega _J = d \omega _K=0\),

  • \(d \Omega = 0\),

  • \(\nabla ^{Ob} = \nabla ^{LC}\),

  • \(\nabla ^{B}_{I,g}= \nabla ^{B}_{J,g}=\nabla ^{B}_{K,g}= \nabla ^{LC}\).

If moreover M is simply connected the conditions above are equivalent to

$$\begin{aligned} Hol(g) \subset Sp(m)\end{aligned}$$

and IJK being induced by this holonomy group.

This class of metrics is standard to consider from the point of view of Berger’s Riemannian holonomy theorem. Indeed it implies that Sp(n), and \(Sp(n)\cdot Sp(1)\), corresponding respectively to the hyperKähler and quaternionic Kähler metrics, are the only infinite families occurring, cf. [Be87], which correspond to the hypercomplex, respectively quaternionic, geometry. There are only two known deformation classes in each dimension and two isolated examples, due to O’Grady, of hyperKähler manifolds. Partially because of that one may be tempted to look for a natural generalizations of those. A possible attempt is as follows.

Definition 2.11

A hyperhermitian manifold (MIJKg) is called HKT, which stands for hyperKähler with torsion, if any of the equivalent conditions is satisfied

  • \(\partial \Omega =0\),

  • \(\nabla ^{B}_{I,g} = \nabla ^{B}_{J,g}=\nabla ^{B}_{K,g}\).

From the second condition it follows that HKT structures are natural differential geometric generalizations of HK structures where the torsion just vanishes. These metrics emerged originally from mathematical physics. More exactly connections with special holonomy and skew torsion occur naturally while studying the target space of sigma models in quantum theory. In the presence of the so called Wess-Zumino term and supersymmetry the HKT metrics appear, cf. [HP96].

An established mathematical treatment of basic properties of HKT manifolds is [GP00]. One should note though that despite the name HKT manifolds do not admit Kähler metrics in general. In fact a result of Verbitsky [V05] shows they can be Kähler only in the case when the manifold already admits HK metric.

2.3 Quaternionic Monge–Ampère equation.

The quaternionic Monge–Ampère equation (1.1) on HKT manifolds proposed by Alesker and Verbitsky, cf. [AV10], naturally solves the prescribed trivialization problem (1.10)–(1.11). The authors suggested to look for an HKT metric whose associated HKT form is

$$\begin{aligned} \Omega _\phi := \Omega + \partial \partial _J \phi \end{aligned}$$

for some smooth real function \(\phi \) and for which \(\Omega _\phi ^n\) is the section we want to obtain because the \(\partial \partial _J\phi \) perturbation preserves the HKT condition. If such a \(\phi \) exists then this new HKT metric \(g_\phi \) can be obtained from \(\Omega _\phi \) by applying the reasoning from the section above. We denote the associated hermitian forms by adding a subscript \(g_\phi \), eg. \(\omega _{I,g_\phi }\).

As we noted the canonical bundle \(K_I(M)\) of a given HKT manifold is trivial topologically, with \(\Omega ^n\) providing a smooth trivialization. Unlike in the Kähler case where, up to the finite covering, topological triviality gives the holomorphic one, here the canonical bundle is not holomorphically trivial in many cases, Hopf surfaces being one example. Arguably, cf. [V09], HKT metrics for which \(\Omega ^n\) is holomorphic trivialization constitute a hypercomplex analogue of a Calabi–Yau manifold. These metrics have in particular the property of being balanced with respect to any complex structure from \(S_M\). They perfectly fit into the recently active stream of research on generalizations of Calabi–Yau spaces, the so called torsion Calabi–Yau manifolds, cf. [T15, Pi19].

Let us now turn to other consequences of Theorem 1.1 advocated in the introduction. In doing so let us also keep in mind that actually an expectation is that Conjecture 1.2 should hold even after dropping the assumption on the holomorphic triviality of the canonical bundle \(K_I(M)\).

For a hyperhermitian manifold (MIJKg) and the constant \(c_n\) depending only on the dimension, as can be seen from (2.14),

$$\begin{aligned} \Omega ^n \wedge \overline{\Omega }^n=c_n \omega _I^{2n}. \end{aligned}$$

Consequently we see that the quaternionic Monge–Ampère equation (1.1) is solvable for any f if and only if

$$\begin{aligned} (\omega _{I, g_\phi })^{2n} = \frac{1}{c}_n (\Omega + \partial \partial _J \phi )^n \wedge (\overline{\Omega + \partial \partial _J \phi })^n = \frac{1}{c}_n e^{(2f+2b)}\Omega ^n \wedge \overline{\Omega }^n = e^{F+C}\omega _I^{2n} \end{aligned}$$

is satisfied with suitable \(\phi \) and C for any F. This shows that Remark 1.4 follows from Theorem 1.1.

The above means that any representative of \(c_1^{BC}(M,I)\) can be obtained as the Chern–Ricci curvature of an HKT metric \(g_\phi \) as the standard calculation shows. This was noted already in [Ma11] for what Madsen calls the projected Chern form, cf. Section 7.1.3 in [Ma11]. Indeed, as is well known prescribing the Chern–Ricci curvature is equivalent to solving the Monge–Ampère equation

$$\begin{aligned} \left( \det ({g_{\phi }}_{i \overline{j}})_{i,j}\right) = e^{F+b}\left( \det (g_{i\bar{j}})_{i,j}\right) \end{aligned}$$

for some \(b \in \mathbb {R}\). This in turn means, by going from the chart expression to the global one, that

$$\begin{aligned} \omega _{I,g_\phi }^n = e^{F+b} \omega ^n_I \end{aligned}$$

justifying that Proposition 1.5 follows from Remark 1.4.

2.4 Technical results.

From now on we assume that the components of all tensors are taken with respect to a holomorphic coordinates \(z_i\) for I. Another important convention we use is that whenever an unknown, eg. i, appears as an index its range is in \(\{0,\ldots ,2n-1\}\). When an expressions 2i or \(2i+1\) involving an unknown appears as an index the range for i is in \(\{0,\ldots ,n-1\}\). We often omit the summation symbols when it is clear that the summation takes place even when the Einstein summation convention does not apply directly.

First of all let us note the following properties of the operators \(\partial \), \(\partial _J\), \(\overline{\partial }\), \(\overline{\partial _J}\) introduced in this section. This result is elementary and based only on the integrability and anti commutativity of I and J but we do not know a reference containing the proof.

Lemma 2.12

For a hypercomplex manifold (MIJK) the following holds

$$\begin{aligned}{} & {} \partial ^2 = \overline{\partial }^2=\partial _J^2=\overline{\partial _J}^2=0, \end{aligned}$$
(2.15)
$$\begin{aligned}{} & {} \partial \overline{\partial }+ \overline{\partial }\partial = \partial _J\overline{\partial _J}+ \overline{\partial _J}\partial _J= \partial \partial _J+ \partial _J\partial = \overline{\partial }\overline{\partial _J}+ \overline{\partial _J}\overline{\partial }= \partial _J\overline{\partial }+ \overline{\partial }\partial _J= \overline{\partial _J}\partial + \partial \overline{\partial _J}= 0. \nonumber \\ \end{aligned}$$
(2.16)

Proof

One can simply use the facts that

$$\begin{aligned}{} & {} dd^c_J+d^c_Jd=0, \end{aligned}$$
(2.17)
$$\begin{aligned}{} & {} dd^c_K + d^c_K d = 0. \end{aligned}$$
(2.18)

This follows from the integrability of J and K. Then by rewriting d as \(\partial + \overline{\partial }\) we obtain

$$\begin{aligned} d= & {} \partial + \overline{\partial }, \\ d^c_I= & {} \mathfrak {i}(\overline{\partial }- \partial ), \\ d^c_J= & {} J^{-1} \circ (\partial + \overline{\partial }) \circ J = \overline{\partial _J}+ \partial _J, \\ d^c_K= & {} (IJ)^{-1} \circ d \circ (IJ) = J^{-1} \circ (I^{-1} \circ d \circ I) \circ J = J^{-1} \circ d^c_I \circ J = \mathfrak {i}( \partial _J- \overline{\partial _J}). \end{aligned}$$

Comparing both sides in (2.17) and (2.18) and taking into an account the Hodge bidegrees with respect to I gives the claim. More precisely (2.17) gives

$$\begin{aligned} \partial \overline{\partial _J}+ \partial \partial _J+ \overline{\partial }\overline{\partial _J}+ \overline{\partial }\partial _J+ \overline{\partial _J}\partial + \partial _J\partial + \overline{\partial _J}\overline{\partial }+ \partial _J\overline{\partial }=0, \end{aligned}$$
(2.19)

while (2.18) gives

$$\begin{aligned} \partial \partial _J- \partial \overline{\partial _J}+ \overline{\partial }\partial _J- \overline{\partial }\overline{\partial _J}+ \partial _J\partial - \overline{\partial _J}\partial + \partial _J\overline{\partial }- \overline{\partial _J}\overline{\partial }=0.\end{aligned}$$
(2.20)

It turns out that

$$\begin{aligned} \partial \partial _J+ \partial _J\partial = 0, \end{aligned}$$

as it is the component of bidegree (2, 0) of the left hand side of (2.19),

$$\begin{aligned} \overline{\partial }\overline{\partial _J}+ \overline{\partial _J}\overline{\partial }= 0, \end{aligned}$$

as it is of bidegree (0, 2) of (2.19),

$$\begin{aligned} \partial \overline{\partial _J}+ \overline{\partial }\partial _J+ \overline{\partial _J}\partial + \partial _J\overline{\partial }= 0, \end{aligned}$$
(2.21)

as it is of bidegree (1, 1) of (2.19),

$$\begin{aligned} - \partial \overline{\partial _J}+ \overline{\partial }\partial _J- \overline{\partial _J}\partial + \partial _J\overline{\partial }= 0, \end{aligned}$$
(2.22)

as it is of bidegree (1, 1) of (2.20). Adding and subtracting (2.21) and (2.22) we obtain

$$\begin{aligned} \overline{\partial }\partial _J+ \partial _J\overline{\partial }= 0 \end{aligned}$$

and

$$\begin{aligned} \partial \overline{\partial _J}+ \overline{\partial _J}\partial = 0. \end{aligned}$$

The only two remaining identities

$$\begin{aligned}\partial \overline{\partial }+ \overline{\partial }\partial = 0, \end{aligned}$$
$$\begin{aligned} \partial _J\overline{\partial _J}+ \overline{\partial _J}\partial _J= J^{-1} \circ ( \overline{\partial }\partial + \partial \overline{\partial }) \circ J = 0\end{aligned}$$

of course do hold. \(\square \)

The first conclusion we may draw from this is, what we will use constantly during the computations of the sections to follow, that certain identities involving derivatives of the components of J vanish locally and not only at a fixed point.

Remark 2.13

For any holomorphic coordinates for I we have

$$\begin{aligned} 0 = \big (\partial \partial _J+ \partial _J\partial \big ) (z_i) = \partial J^{-1} \overline{\partial }z_i + J^{-1} \overline{\partial }J dz_i,\end{aligned}$$

consequently

$$\begin{aligned} \overline{\partial }J \big (dz_i \big ) = 0. \end{aligned}$$

In coordinates this reads

$$\begin{aligned} \overline{\partial }\big ( J_{\overline{k}}^i d\overline{z_k} \big ) = J_{\bar{k},\bar{l}}^i d\overline{z_l} \wedge d\overline{z_k} = \sum _{l<k} \big ( J_{\bar{k},\bar{l}}^i - J_{\bar{l},\bar{k}}^i \big ) d\overline{z_l} \wedge d\overline{z_k} =0. \end{aligned}$$
(2.23)

This in turn provides

$$\begin{aligned} J_{\bar{k},\bar{l}}^i = J_{\bar{l},\bar{k}}^i \end{aligned}$$

and by conjugation

$$\begin{aligned} J_{k,l}^{\bar{i}} = J_{l,k}^{\bar{i}} \end{aligned}$$

for all i, j and k.

In order to have even better control of the derivatives of J we have to stick to the point. Let \(\nabla ^{Ob}\) be the Obata connection for (MIJK). Since it is the complex, in particular for I, torsion free connection we have, in any holomorphic chart for I,

$$\begin{aligned}{} & {} \nabla ^{Ob}_{\partial _{z_i}} \partial _{\overline{z_j}}=\nabla ^{Ob}_{\partial _{\overline{z_i}}} \partial _{{z_j}}=0, \end{aligned}$$
(2.24)
$$\begin{aligned}{} & {} \begin{aligned}&\nabla ^{Ob}_{ \partial _{z_i}} \partial _{z_j} = \Gamma _{ij}^k \partial _{z_k}, \\&\nabla ^{Ob}_{ \partial _{\overline{z_i}} } \partial _{\overline{z_j}} = \Gamma _{\overline{i} \overline{j}}^{\overline{k}} \partial _{\overline{z_k}},\end{aligned} \end{aligned}$$
(2.25)
$$\begin{aligned}{} & {} \begin{aligned}&\nabla ^{Ob}_{ \partial _{z_i}} {dz_j} = -\Gamma _{im}^j d{z_m}, \\&\nabla ^{Ob}_{ \partial _{\overline{z_i}}} {\overline{dz_j}} = -\Gamma _{\bar{i}\bar{m}}^{\bar{j}} d{\overline{z_m}}. \end{aligned} \end{aligned}$$
(2.26)

The condition

$$\begin{aligned}\nabla ^{Ob}J = 0 \end{aligned}$$

gives

$$\begin{aligned} 0= & {} \nabla ^{Ob}_{\partial _{z_i}} \big ( J_{\overline{k}}^l d \overline{z_k} \otimes \partial _{z_l} + J_{k}^{\overline{l}} d z_k \otimes \partial _{\overline{z_l}} \big )\\= & {} J_{\overline{k},i}^l d \overline{z_k} \otimes \partial _{z_l} + J_{\bar{k}}^l d \overline{z_k} \otimes \big ( \Gamma _{il}^m \partial _{z_m}\big ) + J_{k,i}^{\overline{l}} d z_k \otimes \partial _{\overline{z_l}} + J_{k}^{\overline{l}} \big ( -\Gamma _{im}^k d{z_m} \big ) \otimes \partial _{\overline{z_l}}. \end{aligned}$$

This in turn allows us to obtain the expressions

$$\begin{aligned} \begin{aligned}&J_{\overline{k},i}^l = -J_{\bar{k}}^m \Gamma _{im}^l, \\&J_{k,i}^{\overline{l}} = J_{m}^{\overline{l}} \Gamma _{ik}^m,\\&J_{k,\bar{i}}^{\bar{l}} = -J_{k}^{\bar{m}} \Gamma _{\overline{i} \overline{m}}^{\bar{l}}, \\&J_{\bar{k},\bar{i}}^{l} = J_{\bar{m}}^{l} \Gamma _{\overline{i} \overline{k}}^{\bar{m}}.\end{aligned} \end{aligned}$$
(2.27)

Since the Obata connection is torsion free and I-complex, for any chosen \(p \in M\) we can choose I-holomorphic, geodesic coordinates which, from (2.27), gives at p the equalities

$$\begin{aligned} J_{\overline{k},i}^l = J_{k,i}^{\overline{l}} = J_{k,\bar{i}}^{\bar{l}}=J_{\bar{k},\bar{i}}^{l}=0.\end{aligned}$$
(2.28)

Finally, let us find the expression for the \(\partial \partial _J\phi \) perturbation at the point, in local coordinates.

Lemma 2.14

At the point on (MIJK), in any coordinates satisfying (2.13), for any \(\phi \) we have:

$$\begin{aligned} \begin{aligned} \partial _J \phi&=(J^{-1} \overline{\partial }J) \phi =J^{-1} \big ( \overline{\partial }\phi \big ) = J^{-1} \left( \sum \limits _{j=0}^{2n-1} \phi _{\overline{j}} d\overline{z_j} \right) \\&=\sum \limits _{j=0}^{2n-1} \phi _{\overline{j}} J^{-1}(d \overline{z_j})=\sum \limits _{j=0}^{2n-1} \phi _{\overline{j}} (-1)^j d {z_{j+(-1)^j}}= \sum \limits _{j} \big ( \phi _{\overline{2j}} d {{z_{2j+1}}} - \phi _{\overline{2j+1}} d {{z_{2j}}} \big ), \\ \partial \partial _J\phi&= \partial \big ( \phi _{\overline{j}} J^{-1} d \overline{z_j} \big ) = \phi _{i \overline{j}} dz_i \wedge J^{-1} d \overline{z_j}=\sum \limits _{i,j} \left( (-1)^{j+1} \phi _{i \overline{j+(-1)^j}} \right) dz_i \wedge dz_j \\&= \sum \limits _{i,j} \big ( \phi _{2i \overline{2j}} dz_{2i} \wedge d z_{2j+1} + \phi _{2i+1 \overline{2j}} dz_{2i+1} \wedge d z_{2j+1} \\&\quad - \phi _{2i \overline{2j+1}} dz_{2i} \wedge d z_{2j} - \phi _{2i+1 \overline{2j+1}} dz_{2i+1} \wedge d z_{2j} \big ). \end{aligned} \end{aligned}$$
(2.29)

The last one, after rearrangement, gives

$$\begin{aligned}{} & {} \begin{aligned} \partial \partial _J\phi&= \sum \limits _{i,j} \big ( \phi _{2i \overline{2j}} + \phi _{2j+1 \overline{2i+1}} \big ) dz_{2i} \wedge d z_{2j+1} \\&\quad + \sum \limits _{i<j} \big ( \phi _{2i+1 \overline{2j}} - \phi _{2j+1 \overline{2i}} \big ) dz_{2i+1} \wedge d z_{2j+1} \\&\quad + \sum \limits _{i<j} \big ( \phi _{2j \overline{2i+1}} - \phi _{2i \overline{2j+1}} \big ) dz_{2i} \wedge d z_{2j}. \end{aligned} \end{aligned}$$
(2.30)

In the coordinates in which \(\partial \partial _J\phi \) is diagonal, formula (2.30) reads

$$\begin{aligned} \phi _{2i \overline{2j}} =- \phi _{2j+1 \overline{2i+1}}\end{aligned}$$
(2.31)

for \(i \not = j\) and

$$\begin{aligned} \begin{aligned}&\phi _{2i+1 \overline{2j}} = \phi _{2j+1 \overline{2i}} \\&\phi _{2j \overline{2i+1}} = \phi _{2i \overline{2j+1}} \end{aligned} \end{aligned}$$
(2.32)

for any ij.

Using the above discussion we provide, for later reference, the expression for the Chern laplacian involving the quaternionic Hessian.

Proposition 2.15

Let (MIJKg) be a hyperhermitian manifold and \(\phi \!\in \! C^\infty (M)\), then

$$\begin{aligned} 2n \frac{\partial \partial _J\phi \wedge \Omega ^{n-1}}{\Omega ^n} = \Delta ^{Ch}_{I,g} \phi .\end{aligned}$$

Proof

It is well known that the Chern laplacian can be expressed as

$$\begin{aligned}2n \frac{\partial \overline{\partial }\phi \wedge \omega _I^{2n-1}}{\omega _I^{2n}}.\end{aligned}$$

Let us choose any holomorphic coordinates such that at the point \(x \in M\) the hyperhermitian structure is standard, in the sense that (2.12) and (2.13) are satisfied. In those coordinates we see, since \(\omega _I=\frac{\mathfrak {i}}{2}(dz_{2i} \wedge d\overline{z_{2i}} + dz_{2i+1} \wedge d \overline{z_{2i+1}})\),

$$\begin{aligned}2n \frac{\partial \overline{\partial }\phi \wedge \omega _I^{2n-1}}{\omega _I^{2n}} = 2(\phi _{2i\overline{2i}} + \phi _{2i+1\overline{2i+1}}).\end{aligned}$$

On the other hand, because of (2.30) and \(\Omega = dz_{2i} \wedge dz_{2i+1}\), we see that also

$$\begin{aligned}2n \frac{\partial \partial _J\phi \wedge \Omega ^{n-1}}{\Omega ^n} = 2(\phi _{2i\overline{2i}} + \phi _{2i+1\overline{2i+1}})\end{aligned}$$

as required.

We recall the basic facts concerning the Pfaffian. The following proposition is probably well known but we do not know the reference. The proof reduces to defining Pfaffian as below and checking the claimed equality (2.33) of polynomials on sufficiently many skew-symmetric matrices.

Proposition 2.16

There exists a polynomial which we denote by \({\text {Pf }}\), with real coefficients, of degree n on the space of skew-symmetric complex matrices of size 2n, i.e. those satisfying \(A^T=-A\), such that

$$\begin{aligned} \det = {{\text {Pf }}}^2 \end{aligned}$$
(2.33)

as polynomials on this space.

The polynomial \({\text {Pf }}\) from Proposition 2.16, is defined only up to the sign. We make the following choice.

Definition 2.17

For a skew-symmetric complex matrix

$$\begin{aligned} M = \left( m_{ij} \right) _{i,j =1,\ldots ,2n} \end{aligned}$$

we define the Pfaffian of M as

$$\begin{aligned} {\text {Pf }}(M) e_1 \wedge \cdots \wedge e_{2n} = \frac{1}{n!}\left( \sum \limits _{i<j}m_{ij}e_i \wedge e_j \right) ^n\end{aligned}$$
(2.34)

where \(e_i\) is the canonical basis of \(\mathbb {C}^{2n}\).

With this definition, the mentioned convention for tensors components and notation (1.8), it follows immediately that writing the Eq. (1.1) in holomorphic coordinates gives the Eq. (1.9). Note that the original Eq. (1.1) could have been rewritten in that way only because the associated hyperhermitian forms are of the Hodge type (2, 0).

In deriving the estimates we will differentiate Eq. (1.9) and in order to do that we would like to know the formula for the derivative of the Pfaffian. As we will be concerned only with the matrices with positive Pfaffian this can be derived from the more familiar formula for determinant derivatives coupled with Proposition 2.16. The result is as follows.

Lemma 2.18

Suppose \(A = \big [ A_{ij}\big ]_{i,j}\) is a complex skew-symmetric \(2n \times 2n\) matrix with positive Pfaffian depending on variables t and s. Its derivatives are given by

$$\begin{aligned}{} & {} \frac{\partial }{\partial t} \Big ( \log {\text {Pf }}(A) \Big ) = \frac{1}{2} \text {tr}\Big ( A^{-1} \frac{\partial }{\partial t} A \Big ), \end{aligned}$$
(2.35)
$$\begin{aligned}{} & {} \frac{\partial }{\partial s} \frac{\partial }{\partial t} \Big ( \log {\text {Pf }}(A) \Big ) = \frac{1}{2} \bigg [ \text {tr}\Big ( A^{-1} \Big ( \frac{\partial }{\partial s} \frac{\partial }{\partial t} A\Big ) \Big ) - \text {tr}\Big ( A^{-1} \Big ( \frac{\partial }{\partial s} A \Big ) A^{-1} \Big ( \frac{\partial }{\partial t} A\Big ) \Big ) \bigg ].\nonumber \\ \end{aligned}$$
(2.36)

Remark 2.19

From now on we assume that at the point of interest the holomorphic coordinates are chosen so that the hyperhermitian structure (IJKg) is standard, i.e. (2.13) and (2.14) hold, the form \(\Omega _\phi := \Omega + \partial \partial _J\phi \) is diagonal, or equivalently the metric \(g_\phi \) is and that (2.28) holds. Such an arrangement is possible because after choosing the geodesic coordinates we may make a linear change of those gaining the simultaneous diagonalization of both metrics. The vanishing of the Christoffel symbols is preserved under linear change of coordinates. From time to time we may refer to such a coordinates as canonical.

3 \(C^1\) estimate

The goal of this section is to prove the following \(C^1\) a priori estimate:

Theorem 3.1

Let (MIJKg) be a compact, connected hyperKähler manifold. There exists a constant C depending on f, \(\sup _M |\phi |\) and the hyperhermitian structure (IJKg) such that for any solution \(\phi \) of the Eq. (1.1) the estimate

$$\begin{aligned} |d \phi |_g \le C \end{aligned}$$
(3.1)

holds.

Proof

Let us define \(\beta \) by

$$\begin{aligned} \beta \Omega ^n = n \partial \phi \wedge \partial _J\phi \wedge \Omega ^{n-1}.\end{aligned}$$
(3.2)

It is easy to see, for example by rewriting this in canonical coordinates, that

$$\begin{aligned} \beta = \frac{1}{4} |d\phi |_g^2.\end{aligned}$$

Since, after taking the logarithm, the linearization of the Eq. (1.1) is, up to the constant, the Chern Laplacian with respect to the hermitian structure \((I,g_\phi )\) on M, we note the useful form-type formula for this, cf. Proposition 2.15, in our setting

$$\begin{aligned} \partial \partial _Jf \wedge \Omega ^{n-1}_\phi = \Big ( \frac{1}{2n} \Delta ^{Ch}_{I,g_\phi } f \Big ) \Omega ^n_\phi .\end{aligned}$$
(3.3)

Following Błocki, cf. [B09], we consider the quantity

$$\begin{aligned} \alpha = \log \beta - \gamma \circ \phi \end{aligned}$$
(3.4)

for a function \(\gamma : \mathbb {R}\rightarrow \mathbb {R}\) to be specified below.

All the computations from now on will be curried out at a maximum point of \(\alpha \). As the operators \(\partial \) and \(\partial _J\) are of pure first order we note that

$$\begin{aligned}{} & {} \partial \alpha = \frac{\partial \beta }{\beta } - \gamma ' \partial \phi = 0, \end{aligned}$$
(3.5)
$$\begin{aligned}{} & {} \partial _J \alpha = \frac{\partial _J \beta }{\beta } - \gamma ' \partial _J \phi = 0. \end{aligned}$$
(3.6)

Furthermore

$$\begin{aligned} \begin{aligned} \partial \partial _J\alpha =&\frac{\partial \partial _J\beta }{\beta } - \frac{\partial \beta \wedge \partial _J\beta }{\beta ^2} - \gamma '' \partial \phi \wedge \partial _J\phi - \gamma '\partial \partial _J\phi \\&= \frac{\partial \partial _J\beta }{\beta } - \frac{\partial \beta \wedge \partial _J\beta }{\beta ^2} - \gamma '' \partial \phi \wedge \partial _J\phi - \gamma ' \big (\Omega + \partial \partial _J\phi \big ) + \gamma ' \Omega \\&= \frac{\partial \partial _J\beta }{\beta } - \big ( (\gamma ')^2 + \gamma '' \big ) \partial \phi \wedge \partial _J\phi - \gamma ' \big (\Omega + \partial \partial _J\phi \big ) + \gamma ' \Omega . \end{aligned} \end{aligned}$$
(3.7)

Taking the bar of (3.2) results in

$$\begin{aligned} \beta \overline{\Omega }^n = n \overline{\partial }\phi \wedge \overline{\partial _J}\phi \wedge \overline{\Omega }^{n-1}.\end{aligned}$$

Next, taking \(\partial _J\) of both sides, because of the hyperKähler assumption, we get

$$\begin{aligned}\partial _J\beta \wedge \overline{\Omega }^n = n \partial _J\overline{\partial }\phi \wedge \overline{\partial _J}\phi \wedge \overline{\Omega }^{n-1} - n \overline{\partial }\phi \wedge \partial _J\overline{\partial _J}\phi \wedge \overline{\Omega }^{n-1}\end{aligned}$$

and by taking \(\partial \), from the same reason as above, we end up with

$$\begin{aligned}{} & {} \partial \partial _J\beta \wedge \overline{\Omega }^n \nonumber \\{} & {} \quad = n \partial \partial _J\overline{\partial }\phi \wedge \overline{\partial _J}\phi \wedge \overline{\Omega }^{n-1} + n \partial _J\overline{\partial }\phi \wedge \partial \overline{\partial _J}\phi \wedge \overline{\Omega }^{n-1} \nonumber \\{} & {} \qquad - n \partial \overline{\partial }\phi \wedge \partial _J\overline{\partial _J}\phi \wedge \overline{\Omega }^{n-1} + n \overline{\partial }\phi \wedge \partial \partial _J\overline{\partial _J}\phi \wedge \overline{\Omega }^{n-1}.\end{aligned}$$
(3.8)

From the Eq. (1.1)

$$\begin{aligned}\big ( \Omega + \partial \partial _J\phi \big )^n = e^f \Omega ^n\end{aligned}$$

by taking \(\overline{\partial }\) and applying Lemma 2.12 we obtain

$$\begin{aligned}n \partial \partial _J\overline{\partial }\phi \wedge \Omega _\phi ^{n-1} = \overline{\partial }e^f \wedge \Omega ^n\end{aligned}$$

while by taking \(\overline{\partial _J}\)

$$\begin{aligned}n \partial \partial _J\overline{\partial _J}\phi \wedge \Omega _\phi ^{n-1} = \overline{\partial _J}e^f \wedge \Omega ^n.\end{aligned}$$

From this we obtain

$$\begin{aligned}{} & {} \partial \partial _J\beta \wedge \Omega _\phi ^{n-1}\wedge \overline{\Omega }^n \nonumber \\{} & {} = - \overline{\partial _J}\phi \wedge \overline{\partial }e^f \wedge \Omega ^n \wedge \overline{\Omega }^{n-1} + n \partial _J\overline{\partial }\phi \wedge \partial \overline{\partial _J}\phi \wedge \Omega _\phi ^{n-1} \wedge \overline{\Omega }^{n-1}\nonumber \\{} & {} \quad - n \partial \overline{\partial }\phi \wedge \partial _J\overline{\partial _J}\phi \wedge \Omega _\phi ^{n-1} \wedge \overline{\Omega }^{n-1} + \overline{\partial }\phi \wedge \overline{\partial _J}e^f \wedge \Omega ^n \wedge \overline{\Omega }^{n-1}. \end{aligned}$$
(3.9)

We now turn to evaluating the required quantities of second order present in the expression from (3.9). They are equal to

$$\begin{aligned}{} & {} \partial _J\overline{\partial }\phi = \partial _J\left( \phi _{\bar{j}} d\overline{z_j}\right) = J^{-1} \overline{\partial }\left( \phi _{\bar{j}} J d\overline{z_j}\right) = J^{-1} \left( \phi _{\bar{j} \bar{k}} d\overline{z_k} \wedge J d\overline{z_j}\right) = \phi _{\bar{j}\bar{k}} J^{-1}d\overline{z_k} \wedge d\overline{z_j}, \nonumber \\ \end{aligned}$$
(3.10)
$$\begin{aligned}{} & {} \partial \overline{\partial _J}\phi = \partial J^{-1} \partial \phi = \partial J^{-1} \left( \phi _i d z_i \right) = \partial \left( \phi _i J^{-1} dz_i \right) = \phi _{ij} dz_j \wedge J^{-1} d z_i, \end{aligned}$$
(3.11)
$$\begin{aligned}{} & {} \partial \overline{\partial }\phi = \phi _{i \bar{j}} dz_i \wedge d\overline{z_j}, \end{aligned}$$
(3.12)
$$\begin{aligned}{} & {} \partial _J\overline{\partial _J}\phi = J^{-1} \overline{\partial }J J^{-1} \partial \phi = J^{-1} \overline{\partial }\partial \phi = \phi _{i \bar{j}} J^{-1}(d\overline{z_j}) \wedge J^{-1} dz_i. \end{aligned}$$
(3.13)

At a maximum point of \(\alpha \) we have

$$\begin{aligned} \begin{aligned} 0&\ge \frac{1}{2n} \Delta ^{Ch}_{I,g_\phi } \alpha = \frac{\partial \partial _J\alpha \wedge \Omega _\phi ^{n-1} \wedge \overline{\Omega }^n}{\Omega _\phi ^n\wedge \overline{\Omega }^n} \\&= \frac{\partial \partial _J\beta \wedge \Omega _\phi ^{n-1} \wedge \overline{\Omega }^n}{ \beta \Omega _\phi ^n\wedge \overline{\Omega }^n} - \big ( (\gamma ')^2 + \gamma '' \big ) \frac{ \partial \phi \wedge \partial _J\phi \wedge \Omega _\phi ^{n-1} \wedge \overline{\Omega }^n}{ \Omega _\phi ^n\wedge \overline{\Omega }^n} \\&\quad - \gamma '\frac{ \Omega _\phi ^{n} \wedge \overline{\Omega }^n}{ \Omega _\phi ^n\wedge \overline{\Omega }^n} + \gamma ' \frac{\Omega \wedge \Omega _\phi ^{n-1} \wedge \overline{\Omega }^n}{ \Omega _\phi ^n\wedge \overline{\Omega }^n}. \end{aligned} \end{aligned}$$
(3.14)

Applying (3.10)–(3.13) in (3.9) and rewriting in coordinates gives us that those quantities are equal to

$$\begin{aligned}{} & {} \begin{aligned}&\frac{\partial \partial _J\beta \wedge \Omega _\phi ^{n-1} \wedge \overline{\Omega }^n}{ \beta \Omega _\phi ^n \wedge \overline{\Omega }^n} \\&\quad = \frac{1}{n} \frac{1}{ \beta } \left( \frac{\phi _{\bar{i}} (e^f)_i}{e^f} + \frac{ \phi _{i} (e^f)_{\bar{i}}}{e^f} + \frac{\phi _{\bar{j}\bar{k}} \phi _{jk}}{|\Omega ^\phi _{k k+(-1)^k}|} + \frac{\phi _{i\bar{j}} \phi _{ \bar{i}j}}{|\Omega ^\phi _{i i+(-1)^i}|} \right) , \end{aligned} \end{aligned}$$
(3.15)
$$\begin{aligned}{} & {} \quad \quad - \big ( (\gamma ')^2 + \gamma '' \big ) \frac{ \partial \phi \wedge \partial _J\phi \wedge \Omega _\phi ^{n-1} \wedge \overline{\Omega }^n}{ \Omega _\phi ^n\wedge \overline{\Omega }^n}= - \frac{1}{n} \big ( (\gamma ')^2 + \gamma '' \big ) \frac{\phi _i \phi _{\bar{i}}}{|\Omega ^\phi _{i i+(-1)^i}|}, \end{aligned}$$
(3.16)
$$\begin{aligned}{} & {} \quad \quad - \gamma '\frac{ \Omega _\phi ^{n} \wedge \overline{\Omega }^n}{ \Omega _\phi ^n\wedge \overline{\Omega }^n} = -\gamma ', \end{aligned}$$
(3.17)
$$\begin{aligned}{} & {} \quad \quad \gamma ' \frac{\Omega \wedge \Omega _\phi ^{n-1} \wedge \overline{\Omega }^n}{ \Omega _\phi ^n\wedge \overline{\Omega }^n} = \frac{1}{n} \gamma ' \frac{1}{\Omega _{2i 2i+1}^{\phi }}.\end{aligned}$$
(3.18)

We have thus obtained from (3.14)

$$\begin{aligned}{} & {} \begin{aligned}&0 \ge \frac{\phi _{\bar{i}} (e^f)_i}{\beta e^f} + \frac{ \phi _{i} (e^f)_{\bar{i}}}{\beta e^f} + \frac{|\phi _{2i j}|^2 + |\phi _{2i+1 j}|^2}{ \beta \Omega ^\phi _{2i 2i+1}} + \frac{|\phi _{2i \bar{j}}|^2 + |\phi _{2i+1 \bar{j}}|^2}{\beta \Omega ^\phi _{2i 2i+1}}\\&\quad - \big ( {\gamma '}^2 + \gamma '' \big ) \frac{|\phi _{2i}|^2+ |\phi _{2i+1}|^2}{\Omega ^\phi _{2i 2i+1}} - n \gamma ' + \gamma ' \frac{1}{\Omega ^\phi _{2i 2i+1}}. \end{aligned} \end{aligned}$$
(3.19)

Note that we may assume \(\beta > 1\), otherwise we are finished. Under this assumption the first two terms in (3.19) are bounded from below by a quantity not depending on \(\phi \). The next two terms are positive and after fixing \(\gamma \) the penultimate one is bounded from below as well. All of this allows us to rewrite the inequality (3.19) as

$$\begin{aligned} C(\gamma ) \ge - \big ( {\gamma '}^2 + \gamma '' \big ) \frac{|\phi _{2i}|^2+ |\phi _{2i+1}|^2}{\Omega ^\phi _{2i 2i+1}} + \gamma ' \frac{1}{\Omega ^\phi _{2i 2i+1}}. \end{aligned}$$
(3.20)

Now we can take, as in for example [B09],

$$\begin{aligned} \gamma (t) = \frac{\log (2t+1)}{2}. \end{aligned}$$
(3.21)

Under this choice and the \(C^0\) bound we have

$$\begin{aligned} C \ge C_1 \frac{|\phi _{2i}|^2+ |\phi _{2i+1}|^2}{\Omega ^\phi _{2i 2i+1}} + C_2 \frac{1}{\Omega ^\phi _{2i 2i+1}}. \end{aligned}$$
(3.22)

From (3.22) we obtain for any fixed j

$$\begin{aligned} \Omega ^\phi _{2j2j+1} \ge \frac{C_2}{C} \end{aligned}$$

which coupled with the Eq. (1.1) written in canonical coordinates as

$$\begin{aligned} \prod \limits _{i} \Omega _{2i2i+1}^\phi =e^f\end{aligned}$$
(3.23)

gives

$$\begin{aligned} \frac{1}{\Omega ^\phi _{2j2j+1}} \ge \frac{C^{n-1}}{\sup _M e^f}.\end{aligned}$$
(3.24)

Having the bound on \(\Omega ^\phi _{2i2i+1}\)’s from (3.24) we obtain the bound for \(\beta \), from (3.22), at a maximum point of \(\alpha \). This results in a uniform bound for \(\beta \).

Remark 3.2

The argument above corresponds to a gradient bound for the complex Monge–Ampère equation with a background metric of non negative holomorphic bisectional curvature. Let us briefly discuss the problems in the argument above for general HKT metrics. Due to non vanishing of \(d\Omega \) extra terms controlled by

$$\begin{aligned} -C\frac{1}{\Omega ^\phi _{2i 2i+1}} \end{aligned}$$

appear. Hence \(\gamma \) would have to satisfy both

$$\begin{aligned} (\gamma ')^2+\gamma ''\ge 0 \text { and } \gamma '>C \end{aligned}$$

which is possible only if the oscillation of \(\phi \) is small compared to C. Furthermore the idea from [B09] to exploit the terms containing squares of the pure second order derivatives does not seem to work. This is partially explained by the fact that, even in the flat case, though the gradient can be written as

$$\begin{aligned}\beta = \phi _{q_i} \phi _{\bar{q_i}}, \end{aligned}$$

its \(q_j\)’th derivative is not given by the Leibniz rule as this fails for the Cauchy-Riemann-Fueter operators (1.5) and (1.6). What one obtains instead of \(\phi _{q_i q_j} \phi _{\bar{q_i}}+\phi _{q_i } \phi _{\bar{q_i}q_j}\) are \(\phi _{q_i } \phi _{\bar{q_i}q_j}\) and the conjugations of \(\phi _{q_i q_j} \phi _{\bar{q_i}}\). Lack of orthogonality between these conjugates results in insufficient positivity to beat

$$\begin{aligned} -{\gamma '}^2 \frac{|\phi _{2i}|^2+ |\phi _{2i+1}|^2}{\Omega ^\phi _{2i 2i+1}}\end{aligned}$$

- the main negative term.

4 Bound on \(\partial \partial _J\phi \)

In this section we bound partially the Hessian of \(\phi \). More specifically we prove the following a priori estimate:

Theorem 4.1

Let (MIJKg) be a compact, connected hyperKähler manifold. There exists a constant C depending on f, \(\sup _M |\phi |\) and the hyperhermitian structure (IJKg) such that for any solution \(\phi \) of the Eq. (1.1) the estimate

$$\begin{aligned} |\partial \partial _J\phi |_g \le C \end{aligned}$$
(4.1)

holds.

Proof

Let us define this time

$$\begin{aligned} \eta \Omega ^n = \Omega _\phi \wedge \Omega ^{n-1} \end{aligned}$$
(4.2)

and consider the quantity

$$\begin{aligned} \alpha = \log \eta - \gamma \circ \phi , \end{aligned}$$

where the function \(\gamma \) is as in the previous section, cf. (3.21). We note that in order to obtain (4.1) it is sufficient to bound \(\eta \) from above, at a maximal point of \(\alpha \), as it is a positive quantity due to

$$\begin{aligned} \Omega _\phi > 0. \end{aligned}$$
(4.3)

This truly implies (4.1) as in the canonical coordinates

$$\begin{aligned} \eta = \frac{1}{n} \big ( \phi _{i \bar{i}} + n \big ) \end{aligned}$$

is the constant plus the sum of the coefficients of \(\partial \partial _J\phi \) which in light of (4.3) are bounded from below.

We note that at a maximum point of \(\alpha \)

$$\begin{aligned} \begin{aligned} \partial \partial _J\alpha&= \frac{\partial \partial _J\eta }{\eta } - \frac{ \partial \eta \wedge \partial _J\eta }{\eta ^2} - \gamma '' \partial \phi \wedge \partial _J\phi - \gamma ' \partial \partial _J\phi \\&= \frac{\partial \partial _J\eta }{\eta } - \Big ( \big ( \gamma ' \big )^2 + \gamma '' \Big ) \partial \phi \wedge \partial _J\phi - \gamma ' \Omega _\phi + \gamma ' \Omega \end{aligned} \end{aligned}$$
(4.4)

Here we have used the fact that at the extremal point

$$\begin{aligned}{} & {} \frac{\partial \eta }{\eta } = \gamma ' \partial \phi , \\{} & {} \frac{\partial _J\eta }{\eta } = \gamma ' \partial _J\phi .\end{aligned}$$

Let us focus for the moment on the term \(\partial \partial _J\eta \) appearing in (4.4). Differentiating twice the conjugation of the relation (4.2) (recall that we work under the assumption \(d \Omega = 0\)) we obtain

$$\begin{aligned} \partial \partial _J\eta \wedge \Omega _\phi ^{n-1} \wedge \overline{\Omega }^n = \partial \partial _J\overline{\partial }\overline{\partial _J}\phi \wedge \Omega _\phi ^{n-1} \wedge \overline{\Omega }^{n-1}.\end{aligned}$$
(4.5)

An easy calculation in the canonical coordinates shows that

$$\begin{aligned} \begin{aligned} \partial \partial _J\overline{\partial }\overline{\partial _J}\phi&= \partial J^{-1} \overline{\partial }J \overline{\partial }J^{-1} \partial J \phi = \partial J^{-1} \overline{\partial }J \overline{\partial }J^{-1} \phi _i dz_i = \partial J^{-1} \overline{\partial }\phi _{i \bar{j}} J d \overline{z_j} \wedge dz_i \\&= \partial \phi _{i \overline{j} \overline{k}} J^{-1} d \overline{z_k} \wedge d \overline{z_j} \wedge J^{-1} dz_i \\&= \phi _{i \overline{j} \overline{k} l} dz_l \wedge J^{-1} d \overline{z_k} \wedge d \overline{z_j} \wedge J^{-1} dz_i \\&= \Big (\phi _{i \overline{j} \overline{k} l} d \overline{z_j} \wedge J^{-1} dz_i \Big ) \wedge dz_l \wedge J^{-1} d \overline{z_k}. \end{aligned} \end{aligned}$$
(4.6)

Here we have used the relation (2.23) and its conjugation

$$\begin{aligned} \begin{aligned}&\overline{\partial }J^{-1} dz_i = \overline{\partial }J^{-1} \partial J z_i = \overline{\partial }\overline{\partial _J}z_i = - \overline{\partial _J}\overline{\partial }z_i = 0,\\&\partial J^{-1} \overline{\partial }J d \overline{z_j} = \partial \partial _J\overline{\partial }\overline{z_j} = \partial _J\overline{\partial }\partial \overline{z_j} = 0 \end{aligned} \end{aligned}$$

and the fact that the first derivatives of the components of J vanish at the point, cf. (2.28).

Formulas (4.5) and (4.6) allows us to conclude that

$$\begin{aligned} \frac{\partial \partial _J\eta \wedge \Omega _\phi ^{n-1}\wedge \overline{\Omega }^n}{\eta \Omega _\phi ^n \wedge \overline{\Omega }^n}=\frac{\partial \partial _J\overline{\partial }\overline{\partial _J}\phi \wedge \Omega _\phi ^{n-1} \wedge \overline{\Omega }^{n-1}}{ \eta \Omega _\phi ^n \wedge \overline{\Omega }^n} = \frac{1}{\eta n^2}\sum _{l=0}^{n-1} \frac{\sum _{i=0}^{2n-1} \big ( \phi _{i \overline{i} \overline{2l} 2l} + \phi _{i \overline{i} \overline{2l+1} 2l+1} \big )}{\Omega ^{\phi }_{2l 2l+1}}. \nonumber \\ \end{aligned}$$
(4.7)

Now we find another expression for the last quantity in (4.7). Recall that the Eq. (1.1) can be written in the form (1.9)

$$\begin{aligned}Pf\big (\Omega _{ij}^\phi \big ) = e^f Pf(\Omega _{ij}). \end{aligned}$$

After taking the logarithm this reads

$$\begin{aligned} \log Pf(\Omega _{ij}^\phi ) = f + \log Pf(\Omega _{ij}).\end{aligned}$$
(4.8)

Differentiating (4.8) once provides, due to (2.35),

$$\begin{aligned} \frac{1}{2} tr \Big ( \big (\Omega ^{ij}_\phi \big )\big (\Omega _{ij, \bar{p}}^\phi \big ) \Big ) = f_{\bar{p}}. \end{aligned}$$
(4.9)

because, due to the hyperKähler assumption,

$$\begin{aligned} 0= \overline{\partial }\Omega ^n = n! Pf(\Omega _{ij})_{\bar{p}} d\overline{z_p} \wedge dz_0 \wedge \cdots \wedge dz_{2n-1}. \end{aligned}$$
(4.10)

In particular the first barred derivatives vanish locally and not only at the fixed point. Differentiating (4.9) once more, due to the formula (2.36), yields

$$\begin{aligned} \frac{1}{2} tr \Big ( \big (\Omega ^{ij}_\phi \big )\big (\Omega _{ij, \bar{p}p}^\phi \big ) \Big ) - \frac{1}{2} tr \Big ( \big (\Omega ^{ij}_\phi \big )\big (\Omega _{ij, \bar{p}}^\phi \big ) \big (\Omega ^{ij}_\phi \big )\big (\Omega _{ij,p}^\phi \big )\Big ) = f_{\bar{p}p}.\end{aligned}$$
(4.11)

Summing, over p, the formulas (4.11) give us (recall that \(\big [\Omega _{ij}^\phi \big ]_{i,j}\) is block diagonal)

$$\begin{aligned} \begin{aligned}&\frac{\sum _p \Omega ^\phi _{2i2i+1,p \bar{p}}}{\Omega _{2i2i+1}^\phi } = \frac{1}{2} \Delta ^{Ch}_{I,g} f + \frac{1}{2} \Omega _\phi ^{ka}\Omega ^\phi _{al,p}\Omega _\phi ^{lb}\Omega ^\phi _{bk,\bar{p}} = \frac{1}{2} \Delta ^{Ch}_{I,g} f \\&\quad +\frac{1}{2} \frac{ \Omega _{2k+1 2l, p}^\phi \Omega _{2l+1 2k, \bar{p}}^\phi + \Omega _{2k2l+1, p}^\phi \Omega _{2l2k+1, \bar{p}}^\phi - \Omega _{2k+12l+1, p}^\phi \Omega _{2l2k, \bar{p}}^\phi - \Omega _{2k2l, p}^\phi \Omega _{2l+1 2k+1, \bar{p}}^\phi }{\Omega _{2k2k+1}^\phi \Omega _{2l2l+1}^\phi }. \end{aligned} \end{aligned}$$
(4.12)

Using the hyperKähler assumption and (2.23) as well as (2.28) we obtain the formula

$$\begin{aligned} \begin{aligned}&\sum \limits _{k<l} \Omega ^\phi _{kl,\bar{p}} d \overline{z_p} \wedge dz_k \wedge dz_l = \overline{\partial }\sum \limits _{k <l} \Omega ^\phi _{kl} dz_k \wedge dz_l = \overline{\partial }\Omega _\phi = \overline{\partial }\partial \partial _J\phi \\&\quad = \overline{\partial }\sum \limits _{k,l} \phi _{k \bar{l}} dz_k \wedge J^{-1} d \overline{z_l} = \sum \limits _{k,l} \phi _{k \bar{l} \bar{p}} d \overline{z_p} \wedge dz_k \wedge J^{-1} d \overline{z_l} \end{aligned} \end{aligned}$$
(4.13)

and similarly

$$\begin{aligned} \begin{aligned}&\sum \limits _{k< l} \Omega ^\phi _{kl, p} J^{-1} d z_p \wedge dz_k \wedge dz_l = \overline{\partial _J}\sum \limits _{k <l} \Omega ^\phi _{kl} dz_k \wedge dz_l = \overline{\partial _J}\Omega _\phi = \overline{\partial _J}\partial \partial _J\phi \\&\quad = \overline{\partial _J}\sum \limits _{k,l} \phi _{k \bar{l}} dz_k \wedge J^{-1} d \overline{z_l} = \sum \limits _{k,l} \phi _{k \bar{l} p} J^{-1} d z_p \wedge dz_k \wedge J^{-1} d \overline{z_l}. \end{aligned} \end{aligned}$$
(4.14)

From (4.13) and (4.14) we obtain that, for any k, l and p,

$$\begin{aligned} \begin{aligned}&\Omega ^\phi _{2k+1 2l, p} = - \phi _{2k+1 \overline{2l+1} p} - \phi _{2l \overline{2k} p}, \\&\Omega ^\phi _{2l+1 2k, \bar{p}} = - \phi _{2l+1 \overline{2k+1} \bar{p}} - \phi _{2k \overline{2l} \bar{p}}, \\&\Omega ^\phi _{2k+1 2l+1, p} = \phi _{2k+1 \overline{2l} p} - \phi _{2l+1 \overline{2k} p}, \\&\Omega ^\phi _{2l+1 2k+1, \bar{p}} = \phi _{2k+1 \overline{2l} \bar{p}} - \phi _{2l+1 \overline{2k} \bar{p}}, \\&\Omega ^\phi _{2k 2l, p} = \phi _{2k \overline{2l+1} p} - \phi _{2l \overline{2k+1} p}, \\&\Omega ^\phi _{2l 2k, \bar{p}} = \phi _{2k \overline{2l+1} \bar{p}} - \phi _{2l \overline{2k+1} \bar{p}}. \end{aligned}\end{aligned}$$
(4.15)

This gives the expression in terms of derivatives of \(\phi \) for the fourth order component obtained in (4.12)

$$\begin{aligned} \begin{aligned}&\frac{1}{2} \frac{ \Omega _{2k+1 2l, p}^\phi \Omega _{2l+1 2k, \bar{p}}^\phi + \Omega _{2k2l+1, p}^\phi \Omega _{2l2k+1, \bar{p}}^\phi - \Omega _{2k+12l+1, p}^\phi \Omega _{2l2k, \bar{p}}^\phi - \Omega _{2k2l, p}^\phi \Omega _{2l+1 2k+1, \bar{p}}^\phi }{\Omega _{2k2k+1}^\phi \Omega _{2l2l+1}^\phi } \\&\quad = \frac{1}{2} \frac{ |\phi _{2k+1 \overline{2l+1} p} + \phi _{2l \overline{2k} p}|^2 + |\phi _{2l+1 \overline{2k+1} p} + \phi _{2k \overline{2l} p}|^2}{\Omega _{2k2k+1}^\phi \Omega _{2l2l+1}^\phi } \\&\qquad +\frac{1}{2} \frac{ |\phi _{2k+1 \overline{2l} p} - \phi _{2l+1 \overline{2k} p}|^2 + |\phi _{2k \overline{2l+1} p} - \phi _{2l \overline{2k+1} p}|^2}{\Omega _{2k2k+1}^\phi \Omega _{2l2l+1}^\phi }. \end{aligned}\end{aligned}$$
(4.16)

From (4.16) we see that the quantity from (4.12) satisfies

$$\begin{aligned} \frac{\sum _p \Omega ^\phi _{2i2i+1,p \bar{p}}}{\Omega _{2i2i+1}^\phi } \ge \frac{1}{2} \Delta ^{Ch}_{I,g} f \ge - C(f). \end{aligned}$$
(4.17)

Finally observe that from (4.6), much like in (4.13) and (4.14), it is easy to see that

$$\begin{aligned} \phi _{i \overline{i} \overline{2l} 2l} + \phi _{i \overline{i} \overline{2l+1} 2l+1} = \Omega ^\phi _{2l2l+1,i \bar{i}}\end{aligned}$$
(4.18)

for any il.

Having this we return to the estimation of \(\Omega ^\phi _{ij}\)’s. At a maximum point of \(\alpha \) we have

$$\begin{aligned} \begin{aligned}&0 \ge \frac{1}{2n} \Delta ^{Ch}_{I,g_\phi } \alpha = \frac{\partial \partial _J\alpha \wedge \Omega _\phi ^{n-1} \wedge \overline{\Omega }^n}{\Omega _\phi ^n\wedge \overline{\Omega }^n} \\&\quad = \frac{\partial \partial _J\eta \wedge \Omega _\phi ^{n-1} \wedge \overline{\Omega }^n}{ \eta \Omega _\phi ^n\wedge \overline{\Omega }^n} - \big ( (\gamma ')^2 + \gamma '' \big ) \frac{ \partial \phi \wedge \partial _J\phi \wedge \Omega _\phi ^{n-1} \wedge \overline{\Omega }^n}{ \Omega _\phi ^n\wedge \overline{\Omega }^n} \\&\qquad - \gamma '\frac{ \Omega _\phi ^{n} \wedge \overline{\Omega }^n}{ \Omega _\phi ^n\wedge \overline{\Omega }^n} + \gamma ' \frac{\Omega \wedge \Omega _\phi ^{n-1} \wedge \overline{\Omega }^n}{ \Omega _\phi ^n\wedge \overline{\Omega }^n}. \end{aligned} \end{aligned}$$
(4.19)

We may assume \(\eta > 1\), otherwise we are done. By (4.7), (4.18) and (4.17) the first term on the right hand side of (4.19) is bounded from below. The same holds for the third term of (4.19) as \(\gamma \) is chosen as in (3.21). This choice of \(\gamma \) ensures also the positivity of the coefficients of the two remaining terms on the right hand side of (4.19). This means we can rewrite the inequality (4.19) as

$$\begin{aligned} C \ge C_1 \frac{|\phi _{2i}|^2+ |\phi _{2i+1}|^2}{\Omega ^\phi _{2i 2i+1}} + C_2 \frac{1}{\Omega ^\phi _{2i 2i+1}}\end{aligned}$$
(4.20)

for positive constants \(C_1\), \(C_2\). This allows us, as in the previous section, to obtain the bounds on \(\Omega ^\phi _{2i 2i+1}\)’s at a maximum point of \(\alpha \). This in turn gives us a bound on \(\eta \), which is a multiple of the sum of \(\Omega ^\phi _{2i 2i+1}\)’s, at a maximum point of \(\alpha \) which in turn yields the uniform bound on \(\eta \) itself.

Remark 4.2

Let us note that for the general HKT metric the presence of terms coming from differentiating \(\Omega \) in (4.4) significantly complicates the computations. Instead, one has to bound the term

$$\begin{aligned} \frac{ \partial \eta \wedge \partial _J\eta }{\eta ^2}\end{aligned}$$
(4.21)

in (4.4). Unfortunately as one can easily see the quantity

$$\begin{aligned} \frac{ \partial \eta \wedge \partial _J\eta \wedge \Omega _\phi ^{n-1} \wedge \overline{\Omega }^n}{\eta ^2 \Omega ^n \wedge \overline{\Omega }^n} = \frac{1}{\eta ^2 n^3} \sum _i \frac{|\phi _{j \bar{j} i}|^2}{|\Omega _{ii+(-1)^i}^\phi |}\end{aligned}$$
(4.22)

is only bounded by twice the quantity (4.7) as can bee seen from what we have obtained in (4.16).

5 Full \(C^2\) estimate

This section fully exploits the fact that under the assumptions of Theorem 1.1

$$\begin{aligned} \nabla := \nabla ^{Ob}= \nabla ^{LC} = \nabla ^{Ch}_{I,g}. \end{aligned}$$
(5.1)

As we shall see this coupled with the previous a priori bounds suffices to bound the full Hessian of \(\phi \). The \(C^2\) a priori estimate reads as follows:

Theorem 5.1

Let (MIJKg) be a compact, connected hyperKähler manifold. There exists a constant C depending on f, \(\sup _M |\phi |\) and the hyperhermitian structure (IJKg) such that for any solution \(\phi \) of the Eq. (1.1) the estimate

$$\begin{aligned} | \nabla ^2 \phi |_g \le C \end{aligned}$$
(5.2)

holds.

Proof

We wish to estimate the quantity \(\theta \) being, this time, defined for any \(x \in M\) as

$$\begin{aligned} \theta (x) = \lambda _{max}(x):= \sup \limits _{X \in T_x M, \, |X|_g = 1} g(\nabla _X \nabla \phi ,X) = \sup \limits _{X \in T_x M, \, |X|_g = 1} \big (\nabla ^2 \phi \big ) (X,X)\end{aligned}$$
(5.3)

as was done originally in [B11] in the case of the complex Monge–Ampère equation on Kähler manifolds. This is sufficient for the bound on the full Hessian (5.2) because \(\theta \) is the maximum eigenvalue of the Hessian at the point x. More precisely, the sum of all the eigenvalues, being the Laplacian, is bounded from below by

$$\begin{aligned} \frac{1}{2n} \Delta ^{Ch}_{I,g} \phi \ge - \frac{\Omega \wedge \Omega ^{n-1}}{\Omega ^n} = -1\end{aligned}$$
(5.4)

as \(\Omega _\phi > 0 \). Once the sum is under control from below and \(\lambda _{max}\) is bounded from above we obtain the lower bound for the smallest eigenvalue, \(\lambda _{min}\), and consequently we get both sided bounds for all the entries of the matrix of \(\nabla ^2 \phi \).

Consider the quantity \(\alpha \) this time given by

$$\begin{aligned} \alpha = \theta + \frac{1}{4} |d\phi |_g^2. \end{aligned}$$
(5.5)

Since we obtained the gradient bound in Theorem 3.1 it is enough to estimate \(\alpha \) at a maximum point \(p \in M\).

In this section it will be customary to introduce also the real coordinates

$$\begin{aligned} z_i = t_i + \mathfrak {i}t_{2n+i} \end{aligned}$$
(5.6)

for \(i=0,\ldots ,2n-1\), different from the one introduced in (1.4) as can be seen from Remark 2.8.

As the quantity (5.5) is in general non smooth, due to (5.3) not being smooth, we extend a fixed vector

$$\begin{aligned} X = X^j \partial _j(p) \in T_p M,\end{aligned}$$
(5.7)

realizing the supremum in the definition of \(\theta (p)\), to a constant coefficient local vector field

$$\begin{aligned} X = X^j \partial _j.\end{aligned}$$
(5.8)

This X is fixed for the rest of the proof. Consider instead of (5.5) the quantity

$$\begin{aligned} \tilde{\alpha } = \frac{\tilde{\theta }}{|X|_g^2} + \frac{1}{4} |d\phi |_g^2,\end{aligned}$$
(5.9)

where

$$\begin{aligned} \tilde{\theta } = \big ( \nabla ^2 \phi \big ) (X,X). \end{aligned}$$
(5.10)

Observe that

$$\begin{aligned} \frac{\tilde{\theta }}{|X|_g^2} \le \theta \end{aligned}$$
(5.11)

and

$$\begin{aligned} \tilde{\theta }(p) = \theta (p)\end{aligned}$$
(5.12)

which means that also the quantity (5.9) attains a maximum at p. We may assume that

$$\begin{aligned} \theta (p) \ge 0\end{aligned}$$
(5.13)

since otherwise we are done.

We have the following expression in the introduced coordinates (5.6)

$$\begin{aligned} \nabla ^2 \phi = \nabla \phi _{t_j} dt_j = \phi _{t_i t_j} dt_i \otimes dt_j - \Gamma _{ik}^j \phi _{t_j} dt_i \otimes dt_k,\end{aligned}$$
(5.14)

where \(\Gamma _{ji}^k\) are Christoffel symbols in the real frame \(\partial _{t_i}\). From (5.14) we find that

$$\begin{aligned} \tilde{\theta } = D^2_X \phi - \Gamma _{ik}^j \phi _{t_j} X^i X^k, \end{aligned}$$
(5.15)

where D denotes the flat connection in coordinates (5.6). Our goal is to exploit the estimate

$$\begin{aligned} \begin{aligned}&0 \ge \frac{1}{2n} \Delta ^{Ch}_{I,g_\phi } \tilde{\alpha } \\&\quad = \frac{\partial \partial _J\tilde{\theta } \wedge \Omega _\phi ^{n-1} \wedge \overline{\Omega }^n}{\Omega _\phi ^n \wedge \overline{\Omega }^n} - \tilde{\theta }\frac{\partial \partial _J\left( {|X|_g^2} \right) \wedge \Omega _\phi ^{n-1} \wedge \overline{\Omega }^n}{\Omega _\phi ^n \wedge \overline{\Omega }^n} + \frac{\partial \partial _J\frac{1}{4} | d \phi |_g^2 \wedge \Omega _\phi ^{n-1} \wedge \overline{\Omega }^n}{\Omega _\phi ^n \wedge \overline{\Omega }^n}, \end{aligned}\nonumber \\ \end{aligned}$$
(5.16)

where we have used the fact that at the point p

$$\begin{aligned}\partial |X|_g^2 = \partial _J|X|_g^2 = 0. \end{aligned}$$

As we already noticed in the previous section, in the canonical coordinates, we have at p

$$\begin{aligned} \frac{\partial \partial _J\tilde{\theta } \wedge \Omega _\phi ^{n-1} \wedge \overline{\Omega }^n}{\Omega _\phi ^n \wedge \overline{\Omega }^n} = \frac{1}{n} \frac{\tilde{\theta }_{2p\overline{2p}} + \tilde{\theta }_{2p+1\overline{2p+1}}}{\Omega _{2p2p+1}^\phi }. \end{aligned}$$
(5.17)

Differentiating the expression (5.15) for \(\tilde{\theta }\) we obtain

$$\begin{aligned} \tilde{\theta }_{p \bar{p}} = D^2_X \phi _{p \bar{p}} - \Gamma _{ik,p \bar{p}}^j \phi _{t_j} X^i X^k - \Gamma _{ik,p}^J \phi _{t_j \bar{p}} X^i X^k - \Gamma _{ik, \bar{p}}^j \phi _{t_j p} X^i X^k \ge D^2_X \phi _{p \bar{p}} - C(\tilde{\theta } +1 ).\nonumber \\ \end{aligned}$$
(5.18)

In the estimation (5.18) we have used the facts that \(\Gamma _{ij}^k\) vanish at the point p (recall (2.27) and (2.28)), \(\Gamma _{ij}^k\)’s derivatives depend only on the derivatives of the initial metric g, the gradient of \(\phi \) is bounded and

$$\begin{aligned} |\phi _{t_it_j}|<C(1 + \tilde{\theta }). \end{aligned}$$
(5.19)

Since we know from Sect. refSec4, (4.1), that

$$\begin{aligned} \frac{1}{C} \le \Omega ^\phi _{2i2i+1} \le C \end{aligned}$$
(5.20)

we can estimate the quantity (5.17) by applying (5.18) and (5.20)

$$\begin{aligned} \frac{1}{n} \frac{\tilde{\theta }_{2p\overline{2p}} + \tilde{\theta }_{2p+1\overline{2p+1}}}{\Omega _{2p2p+1}^\phi } \ge \frac{1}{n} \frac{D^2_X \phi _{2p \overline{2p}} + D^2_X \phi _{2p+1 \overline{2p+1}}}{\Omega ^\phi _{2p2p+1}} - C(\tilde{\theta } + 1). \end{aligned}$$
(5.21)

In order to deal with the last remaining terms involving derivatives of \(\phi \) in (5.21) let us differentiate the equation (4.8) twice in the direction of X obtaining

$$\begin{aligned} \frac{1}{2} tr \Big ( (\Omega ^{ij}_\phi )(D^2_X \Omega _{ij}^\phi ) \Big ) - \frac{1}{2} tr \Big ( (\Omega ^{ij}_\phi )(D_X \Omega _{ij}^\phi ) (\Omega ^{ij}_\phi )(D_X \Omega _{ij}^\phi )\Big ) = D^2_X f + D^2_X \log Pf(\Omega _{ij}). \nonumber \\ \end{aligned}$$
(5.22)

Rewriting the quantity in (5.22) explicitly gives

$$\begin{aligned} \begin{aligned} \frac{ D^2_X \Omega ^\phi _{2i2i+1}}{\Omega _{2i2i+1}^\phi }&= D^2_X f + D^2_X \log Pf(\Omega _{ij}) + \frac{1}{2} \Omega _\phi ^{ka}D_X \Omega ^\phi _{al}\Omega _\phi ^{lb} D_X \Omega ^\phi _{bk} \\&= D^2_X f + D^2_X \log Pf(\Omega _{ij}) \\&\quad + \frac{1}{2} \frac{ D_X \Omega _{2k+1 2l}^\phi D_X \Omega _{2l+1 2k}^\phi + D_X \Omega _{2k2l+1}^\phi D_X \Omega _{2l2k+1}^\phi }{\Omega _{2k2k+1}^\phi \Omega _{2l2l+1}^\phi } \\&\quad - \frac{1}{2} \frac{D_X \Omega _{2k+12l+1}^\phi D_X \Omega _{2l2k}^\phi + D_X \Omega _{2k2l}^\phi D_X \Omega _{2l+1 2k+1}^\phi }{\Omega _{2k2k+1}^\phi \Omega _{2l2l+1}^\phi }. \end{aligned}\end{aligned}$$
(5.23)

From the formulas we obtained in (4.15) we have the expression for (5.23) in terms of the derivatives of \(\phi \) as follows

$$\begin{aligned} \begin{aligned} \frac{1}{2}&\frac{ D_X \Omega _{2k+1 2l}^\phi D_X \Omega _{2l+1 2k}^\phi + D_X \Omega _{2k2l+1}^\phi D_X \Omega _{2l2k+1}^\phi }{\Omega _{2k2k+1}^\phi \Omega _{2l2l+1}^\phi } \\&\qquad - \frac{1}{2} \frac{D_X \Omega _{2k+12l+1}^\phi D_X \Omega _{2l2k}^\phi + D_X \Omega _{2k2l}^\phi D_X \Omega _{2l+1 2k+1}^\phi }{\Omega _{2k2k+1}^\phi \Omega _{2l2l+1}^\phi } \\&\quad = \frac{1}{2} \frac{ |D_X \phi _{2k+1 \overline{2l+1}} + D_X\phi _{2l \overline{2k}}|^2 + |D_X \phi _{2l+1 \overline{2k+1}} + D_X \phi _{2k \overline{2l} }|^2}{\Omega _{2k2k+1}^\phi \Omega _{2l2l+1}^\phi } \\&\qquad + \frac{1}{2} \frac{|D_X \phi _{2k+1 \overline{2l} } - D_X \phi _{2l+1 \overline{2k} }|^2 + |D_X \phi _{2k \overline{2l+1} } - D_X \phi _{2l \overline{2k+1} }|^2}{\Omega _{2k2k+1}^\phi \Omega _{2l2l+1}^\phi } \end{aligned}\end{aligned}$$
(5.24)

which is seen to be non negative. We also note that

$$\begin{aligned} \Omega _{ij}^\phi = \Omega _{ij} + \left( - \phi _{i \bar{k}}J^{\bar{k}}_j + \phi _{j \bar{k}}J^{\bar{k}}_i\right) \end{aligned}$$
(5.25)

which gives (recall (2.28)),

$$\begin{aligned} D^2_X \Omega ^\phi _{2i2i+1} = D_X^2 \Omega _{2i2i+1} + D^2_X \phi _{2i \overline{2i}} + D^2_X \phi _{2i+1 \overline{2i+1}} + \big (- \phi _{2i \bar{k}} D^2_X J^{\bar{k}}_{2i+1} + \phi _{2i+1 \bar{k}} D^2_X J^{\bar{k}}_{2i}\big ).\nonumber \\ \end{aligned}$$
(5.26)

Applying (5.26) in (5.23) coupled with (5.19) and (5.24) provides

$$\begin{aligned} \frac{D^2_X \phi _{2p \overline{2p}} + D^2_X \phi _{2p+1 \overline{2p+1}}}{\Omega ^\phi _{2p2p+1}} \ge - C(\tilde{\theta } + 1).\end{aligned}$$
(5.27)

Finally, applying (5.27) in (5.21) gives

$$\begin{aligned} \frac{1}{n} \frac{\tilde{\theta }_{2p\overline{2p}} + \tilde{\theta }_{2p+1\overline{2p+1}}}{\Omega _{2p2p+1}^\phi } \ge - C(\tilde{\theta } + 1) \end{aligned}$$
(5.28)

providing the lower bound for the first quantity in (5.16).

As for the third factor of (5.16) we recall from the computations for the gradient estimate, cf. (3.15), that

$$\begin{aligned} \begin{aligned}&\frac{\partial \partial _J\frac{1}{4} | d \phi |_g^2 \wedge \Omega _\phi ^{n-1} \wedge \overline{\Omega }^n}{\Omega _\phi ^n \wedge \overline{\Omega }^n} \\&\quad = \frac{1}{n} \left( \frac{\phi _{\bar{i}} e^f_i}{e^f} + \frac{ \phi _{i} e^f_{\bar{i}}}{e^f} + \frac{\phi _{\bar{j}\bar{k}} \phi _{jk}}{|\Omega ^\phi _{k k+(-1)^k}|} + \frac{\phi _{i\bar{j}} \phi _{ \bar{i}j}}{|\Omega ^\phi _{i i+(-1)^i}|} \right) . \end{aligned} \end{aligned}$$
(5.29)

By (5.20) and (3.1) we obtain from (5.29) the bound

$$\begin{aligned} \frac{\partial \partial _J\frac{1}{4} | d \phi |_g^2 \wedge \Omega _\phi ^{n-1} \wedge \overline{\Omega }^n}{\Omega _\phi ^n \wedge \overline{\Omega }^n} \ge -C' + C(|\phi _{ij}|^2+ |\phi _{i\bar{j}}|^2).\end{aligned}$$
(5.30)

We note that

$$\begin{aligned} |\phi _{ij}|^2+ |\phi _{i\bar{j}}|^2 \ge C \tilde{\theta }^2.\end{aligned}$$
(5.31)

Applying (5.31) in (5.30) gives us

$$\begin{aligned} \frac{\partial \partial _J\frac{1}{4} | d \phi |_g^2 \wedge \Omega _\phi ^{n-1} \wedge \overline{\Omega }^n}{\Omega _\phi ^n \wedge \overline{\Omega }^n} \ge -C' + C \tilde{\theta }^2.\end{aligned}$$
(5.32)

The second term in (5.16) can be easily seen to satisfy

$$\begin{aligned} -\tilde{\theta }\frac{\partial \partial _J\left( {|X|_g^2} \right) \wedge \Omega _\phi ^{n-1} \wedge \overline{\Omega }^n}{\Omega _\phi ^n \wedge \overline{\Omega }^n} \ge -C \tilde{\theta }.\end{aligned}$$
(5.33)

This is because after rewriting this term in coordinates as in (5.17) and applying (5.20) we observe it depends only on \(\tilde{\theta }\) and second derivatives of the metric g alone (as the coefficients of X are constant and we are computing in normal coordinates).

The estimations (5.28), (5.32) and (5.33) applied in (5.16) deliver

$$\begin{aligned} 0 \ge \frac{1}{2n} \Delta ^{Ch}_{I,g_\phi } \tilde{\alpha } \ge C \tilde{\theta }^2 - C' \tilde{\theta } - C''.\end{aligned}$$
(5.34)

Inequality (5.34) provides the desired estimate on \(\tilde{\theta }\) in terms of C, \(C'\) and \(C''\). The desired bound on \(\theta \) follows from that.

6 Proof of Theorem 1.1

As we advocated in the introduction having Theorem 3.1, Theorem 5.1 and using Corollary 5.7 in [AV10] (or Theorem 1.1.13 of [AS17] or Theorem A in [Sr19]), one obtains the \(C^{2,\alpha }\) a priori estimate for the solutions of (1.1) for some \(\alpha \in (0,1)\). The rest of the proof are completely standard. For the convenience of the reader we sketch them and we refer to [A13] Sect. 5 for a the detailed discussion.

Turning to the proof of Theorem 1.1. For the given f satisfying (1.2) we set up the continuity path

$$\begin{aligned} (\Omega + \partial \partial _J\phi )^n = \big ( te^f +(1-t) \big ) \Omega ^n\end{aligned}$$
(6.1)

for \(t \in [0,1]\). In order to prove Theorem 1.1 it is enough to show that the set S of \(t \in [0,1]\), for any fixed \(k \ge 1\) and \(\alpha \in (0,1)\), such that there exists \(\phi \in C^{k+2,\alpha }\) solving (6.1) is both open and closed.

Openness in our setting is completely standard. One has to prove that for a fixed \(\phi \) the operator

$$\begin{aligned} U^{k+2,\alpha } \ni \psi \longmapsto \frac{(\Omega + \partial \partial _J\phi + \partial \partial _J\psi )^n}{\Omega ^n} \in V^{k,\alpha }\end{aligned}$$
(6.2)

has an open image. In (6.2) the Banach manifolds, with the induced Hölder norms, are given by

$$\begin{aligned} \begin{aligned} U^{k+2,\alpha }&:= \{ \phi \in C^{k+2,\alpha }(M) \, | \, \int _M \phi \Omega ^n \wedge \overline{\Omega }^n = 0 \} \\ V^{k,\alpha }&:= \{ F \in C^{k,\alpha }(M) \, | \, \int _M F \Omega ^n \wedge \overline{\Omega }^n = \int _M \Omega ^n \wedge \overline{\Omega }^n \}. \end{aligned}\end{aligned}$$
(6.3)

The fact that (6.2) has an open image was proven in Proposition 5.1 of [A13] and relies on the theory of linear elliptic operators.

As for the closedness it is enough to know that once \(t_i \in S\) are such that \(t_i \rightarrow t\) then \(t \in S\) as well. For any \(t_i\) let us take \(\phi _{t_i}\) solving (6.1) normalized by

$$\begin{aligned} \sup _M \phi _{t_i}=0.\end{aligned}$$
(6.4)

Once we know that the sequence \(\phi _{t_i}\) is bounded in \(C^{k+3,\alpha }\) the Kondrakov theorem yields a subsequence converging in a \(C^{k+3}\) norm (and hence in \(C^{k+2, \alpha }\) norm) to the solution \(\phi \) of (6.1). All we need then is to have a priori estimates for the solutions of (1.1) normalized by (6.4) up to the order \(C^{k+3,\alpha }\). From Theorem 3.1, Theorem 5.1 and Corollary 5.7 in [AV10] we have the \(C^2\) estimate. Applying now the Evans–Krylov theorem, cf. [E82], to the operator

$$\begin{aligned} \log \bigg ( \frac{(\Omega + \partial \partial _J\phi )^n}{\Omega ^n} \bigg )\end{aligned}$$
(6.5)

defined on those functions for which the hyperhermitian matrix associated to \(\Omega + \partial \partial _J\phi \) is positive we obtain a \(C^{2,\alpha }\) estimate for some fixed \(\alpha \). From this the standard procedure of bootstrapping provides the bounds of any higher order as in Section 17.5 of [GT01].

Remark 6.1

For simplicity of presentation (and calculations for obtaining a priori estimates) we stated Theorem 1.1 in the setting when the initial metric g is already HK. Actually, our method works equally well just under the assumption that the hypercomplex manifold (MIJK) admits some compatible hyperKähler metric \(g'\) and the initial hyperhermitian metric g is arbitrary (in particular HKT). In that setting no new, essential, complications arise while performing a priori estimates, provided we still define the test quantities using \(g'\). This is because all the new terms are estimable from below by

$$\begin{aligned} - \frac{C}{\delta ^\sigma } \cdot \frac{1}{\Omega ^\phi _{2i 2i+1}},\end{aligned}$$

where C is a constant depending on the curvature of the initial metric g, \(\sigma \in \{\frac{1}{2},1\}\) and \(\delta \) is the test quantity we are trying to estimate at the moment. Thus, we can always assume \(\delta \) makes the above term arbitrarily small in comparison with

$$\begin{aligned} \gamma ' \cdot \frac{1}{\Omega ^\phi _{2i 2i+1}},\end{aligned}$$

for \(\gamma \) which is chosen in the sections above, since otherwise we already obtain a bound on the test quantity.