The theory of complex hyperbolic manifolds and complex-hyperbolic Kleinian groups (i.e., discrete holomorphic isometry groups of complex hyperbolic spaces \(\mathbb {H}^n_\mathbb {C}\)) is a rich mixture of Riemannian and complex geometry, topology, dynamics, symplectic geometry and complex analysis. The purpose of this note is to discuss interactions of the theory of complex-hyperbolic Kleinian groups and the function theory of complex-hyperbolic manifolds. Let \(\varGamma \) be a discrete group of isometries acting on the complex-hyperbolic n-space, \(\mathbb {H}^n_\mathbb {C}\), the unit ball \(\mathbf {B}^n\subset \mathbb {C}^n\) equipped with the Bergman metric. A fundamental numerical invariant associated with \(\varGamma \) is the critical exponent \(\delta (\varGamma )\) of \(\varGamma \), defined by

$$\begin{aligned} \delta (\varGamma ) = \inf \left\{ s : \sum _{\gamma \in \varGamma } e^{-s\cdot d(x,\gamma x)} <\infty \right\} , \end{aligned}$$

where \(x\in \mathbb {H}^n_\mathbb {C}\) is anyFootnote 1 point. The critical exponent measures the rate of exponential growth the \(\varGamma \)-orbit \(\varGamma x\subset \mathbb {H}^n_\mathbb {C}\); it also equals the Hausdorff dimension of the conical limit set of \(\varGamma \), see [7, 8].

Our main result is:

FormalPara Theorem 1

Suppose that \(\varGamma < {\text {Aut}}(\mathbf {B}^n)\) is a convex-cocompact, torsion-free discrete subgroup satisfying \(\delta (\varGamma )<2\). Then, \(M_\varGamma = \mathbf {B}^n/\varGamma \) is Stein.

The condition on the critical exponent in the above theorem is sharp, since for a complex Fuchsian subgroup \(\varGamma < {\text {Aut}}(\mathbf {B}^n)\), \(\delta (\varGamma ) = 2\), but the quotient \(M_\varGamma = \mathbf {B}^n/\varGamma \) is non-Stein, because the convex core of \(M_\varGamma \) is a complex curve, see Example 4. On the other hand, if \(\varGamma \) is a torsion-free real Fuchsian subgroup or a small deformation of such (see Example 3), then \(\varGamma \) satisfies the condition of the above theorem.

The main ingredients in the proof of Theorem 1 are Proposition 11 and Theorem 15. The condition “convex-cocompact” is only used in Proposition 11, whereas Theorem 15 holds for any torsion-free discrete subgroup \(\varGamma <{\text {Aut}}(\mathbf {B}^n)\) satisfying \(\delta (\varGamma )<2\).

FormalPara Conjecture 2

Theorem 1 holds if we omit the “convex-cocompact” assumption on \(\varGamma \).

In Sect. 4, we discuss other conjectural generalizations of Theorem 1 and supporting results.

1 Preliminaries

In this section, we recall some definitions and basic facts about the n-dimensional complex hyperbolic space, we refer to [9, 11] for details.

Consider the n-dimensional complex vector space \(\mathbb {C}^{n+1}\) equipped with the pseudo-hermitian bilinear form:

$$\begin{aligned} \langle z, w\rangle =-z_0 \bar{w}_0 + \sum _{k=1}^n z_k \bar{w}_k \end{aligned}$$
(1)

and define the quadratic form q(z) of signature (n, 1) by \(q(z):= \langle z, z\rangle \). Then, q defines the negative light cone \(V_-:= \{z: q(z)<0\} \subset \mathbb {C}^{n+1}\). The projection of \(V_-\) in the projectivization of \(\mathbb {C}^{n+1}\), \(\mathbb {P}^{n}\), is an open ball which we denote by \(\mathbf {B}^n\).

The tangent space \(T_{[z]} \mathbb {P}^n\) is naturally identified with \(z^\perp \), the orthogonal complement of \(\mathbb {C}z\) in V, taken with respect to \(\langle \cdot , \cdot \rangle \). If \(z\in V_-\), then the restriction of q to \(z^\perp \) is positive-definite, hence, \(\langle \cdot , \cdot \rangle \) project to a hermitian metric h (also denoted \(\langle \cdot , \cdot \rangle _h\)) on \(\mathbf {B}^n\). The complex hyperbolic n-space \(\mathbb {H}^n_{\mathbb {C}}\) is \(\mathbf {B}^n\) equipped with the hermitian metric h. The boundary \(\partial \mathbf {B}^n\) of \(\mathbf {B}^n\) in \(\mathbb {P}^n\) gives a natural compactification of \(\mathbf {B}^n\).

In this note, we usually denote the complex hyperbolic n-space by \(\mathbf {B}^n\). The real part of the hermitian metric h defines a Riemannian metric g on \(\mathbf {B}^n\). The sectional curvature of g varies between \(-4\) and \(-1\). We denote the distance function on \(\mathbf {B}^n\) by d. The distance function satisfies

$$\begin{aligned} \cosh ^2(d(0, z))= (1-|z|^2)^{-1}. \end{aligned}$$
(2)

A real linear subspace \(W\subset \mathbb {C}^{n+1}\) is said to be totally real with respect to the form (1) if for any two vectors \(z, w\in W\), \(\langle z, w\rangle \in \mathbb {R}\). Such a subspace is automatically totally real in the usual sense: \(J W\cap W=\{0\}\), where J is the almost complex structure on V. (Real) geodesics in \(\mathbf {B}^n\) are projections of totally real indefinite (with respect to q) 2-planes in \(\mathbb {C}^{n+1}\) (intersected with \(V_-\)). For instance, geodesics through the origin \(0\in \mathbf {B}^n\) are Euclidean line segments in \(\mathbf {B}^n\). More generally, totally-geodesic real subspaces in \(\mathbf {B}^n\) are projections of totally real indefinite subspaces in \(\mathbb {C}^{n+1}\) (intersected with \(V_-\)). They are isometric to the real hyperbolic space \(\mathbb {H}^n_\mathbb {R}\) of constant sectional curvature \(-1\).

Complex geodesics in \(\mathbf {B}^n\) are projections of indefinite complex 2-planes. Complex geodesics are isometric to the unit disk with the hermitian metric:

$$\begin{aligned} \frac{dz d{\bar{z}}}{(1-|z|^2)^2}, \end{aligned}$$

which has constant sectional curvature \(-4\). More generally, k-dimensional complex hyperbolic subspaces \(\mathbb {H}^k_\mathbb {C}\) in \(\mathbf {B}^n\) are projections of indefinite complex \((k+1)\)-dimensional subspaces (intersected with \(V_-\)).

All complete totally-geodesic submanifolds in \(\mathbb {H}^n_\mathbb {C}\) are either real or complex hyperbolic subspaces.

The group \({\mathrm {U}}(n,1)\cong {\mathrm {U}}(q)\) of (complex) automorphisms of the form q projects to the group \({\text {Aut}}(\mathbf {B}^n) \cong {\mathrm {PU}}(n,1)\) of complex (biholomorphic, isometric) automorphisms of \(\mathbf {B}^n\). The group \({\text {Aut}}(\mathbf {B}^n)\) is linear, its matrix representation is given, for instance, by the adjoint representation, which is faithful since \({\text {Aut}}(\mathbf {B}^n)\) has trivial center.

A discrete subgroup \(\varGamma \) of \({\text {Aut}}(\mathbf {B}^n)\) is called a complex-hyperbolic Kleinian group. The accumulation set of an(y) orbit \(\varGamma x\) in \(\partial \mathbf {B}^n\) is called the limit set of \(\varGamma \) and denoted by \(\varLambda (\varGamma )\). The complement of \(\varLambda (\varGamma )\) in \(\partial \mathbf {B}^n\) is called the domain of discontinuity of \(\varGamma \) and denoted by \(\varOmega (\varGamma )\). The group \(\varGamma \) acts properly discontinuously on \(\mathbf {B}^n \cup \varOmega (\varGamma )\).

For a (torsion-free) complex-hyperbolic Kleinian group \(\varGamma \), the quotient \(\mathbf {B}^n/\varGamma \) is a Riemannian orbifold (manifold) equipped with push-forward of the Riemannian metric of \(\mathbf {B}^n\). We reserve the notation \(M_\varGamma \) to denote this quotient. The convex core of \(M_\varGamma \) is the the projection of the closed convex hull of \(\varLambda (\varGamma )\) in \(\mathbf {B}^n\). The subgroup \(\varGamma \) is called convex-cocompact if the convex core of \(M_\varGamma \) is a nonempty compact subset. Equivalently (see [3]), \(\overline{M}_\varGamma =(\mathbf {B}^n \cup \varOmega (\varGamma ))/\varGamma \) is compact.

Below are two interesting examples of convex-cocompact complex-hyperbolic Kleinian groups which will also serve as illustrations our results.

Example 3

(Real Fuchsian subgroups) Let \(\mathbb {H}^2_\mathbb {R}\subset \mathbf {B}^n\) be a totally real hyperbolic plane. This inclusion is induced by an embedding \(\rho :{{\,\mathrm{Isom}\,}}(\mathbb {H}^2_\mathbb {R}) = \mathrm {PSL}(2,\mathbb {R}) \rightarrow {\text {Aut}}(\mathbf {B}^n)\), whose image preserves \(\mathbb {H}^2_\mathbb {R}\). Let \(\varGamma '< {{\,\mathrm{Isom}\,}}(\mathbb {H}^2_\mathbb {R})\) be a uniform lattice. Then \(\varGamma = \rho (\varGamma ')\) preserves \(\mathbb {H}^2_\mathbb {R}\) and acts on it cocompactly. Such subgroups \(\varGamma <{\text {Aut}}(\mathbf {B}^n)\) will be called real Fuchsian subgroups. The compact surface-orbifold \(\varSigma = \mathbb {H}^2_\mathbb {R}/{\varGamma }\) is the convex core of \(M_{\varGamma }\). The critical exponent \(\delta (\varGamma )\) is 1.

Let \(\varGamma _t\), \(t\ge 0\), be a continuous family of deformations of \(\varGamma _0 = \varGamma \) in \({\text {Aut}}(\mathbf {B}^n)\) such that \(\varGamma _t\)’s, for \(t>0\), are convex-cocompact but not real Fuchsian. Such deformation exist as long as \(\varGamma _t\) is, say, torsion-free, see, e.g., [15]. The groups \(\varGamma _t\), \(t>0\), are called real quasi-Fuchsian subgroups. The critical exponents of such subgroups are strictly greater than 1.

Example 4

(Complex Fuchsian subgroups) Let \(\varGamma '\) be a cocompact subgroup of \({\mathrm {SU}}(1,1)\), the identity component isometry group of the real-hyperbolic plane (modulo \({\mathbb {Z}}_2\)) and let \({\mathrm {SU}}(1,1) \rightarrow {\mathrm {SU}}(n,1)\) be any embedding. Note that \({\mathrm {SU}}(n,1)\) modulo center (isomorphic to \({\mathbb {Z}}_{n+1}\)) is isomorphic to \({\mathrm {PU}}(n,1)\). By taking compositions, we get a representation \(\rho :\varGamma '\rightarrow {\mathrm {PU}}(n,1)\). Then \(\varGamma := \rho (\varGamma ')\) leaves a complex geodesic invariant in \(\mathbf {B}^n\). Such subgroups \(\varGamma \) will be called complex Fuchsian subgroups. In this case, \(\mathrm {core}(M_{\varGamma }) = \mathbb {H}^1_\mathbb {C}/\varGamma \) is a compact complex curve in \(M_{\varGamma }\) where \(\mathbb {H}^1_\mathbb {C}\) is the \(\varGamma \)-invariant complex geodesic. The critical exponent \(\delta (\varGamma )\) is 2.

2 Generalities on Complex Manifolds

By a complex manifold with boundary M, we mean a smooth manifold with (possibly empty) boundary \(\partial M\) such that \({\text {int}}(M)\) is equipped with a complex structure and that there exists a smooth embedding \(f: M\rightarrow X\) to an equidimensional complex manifold X, biholomorphic on \({\text {int}}(M)\). A holomorphic function on M is a smooth function which admits a holomorphic extension to a neighborhood of M in X.

Let X be a complex manifold and \(Y\subset X\) is a codimension 0 smooth submanifold with boundary in X. The submanifold Y is said to be strictly Levi-convex if every boundary point of Y admits a neighborhood U in X such that the submanifold with boundary \(Y\cap U\) can be written as

$$\begin{aligned} \{\phi \le 0\}, \end{aligned}$$

for some smooth submersion \(\phi : U\rightarrow \mathbb {R}\) satisfying \( {{\,\mathrm{Hess}\,}}(\phi ) >0, \) where \({{\,\mathrm{Hess}\,}}(\phi )\) is the holomorphic Hessian:

$$\begin{aligned} \left( \frac{\partial ^2 \phi }{\partial \bar{z}_i\partial z_j}\right) . \end{aligned}$$

Definition 5

A strongly pseudoconvex manifold M is a complex manifold with boundary which admits a strictly Levi-convex holomorphic embedding in an equidimensional complex manifold.

Definition 6

An open complex manifold Z is called holomorphically convex if for every discrete closed subset \(A\subset Z\) there exists a holomorphic function \(Z\rightarrow \mathbb {C}\) which is proper on A.

Alternatively,Footnote 2 one can define holomorphically convex manifolds as follows: For a compact K in a complex manifold M, the holomorphic convex hull \(\hat{K}_M\) of K in M is

$$\begin{aligned} \hat{K}_M= \{z\in M: |f(z)|\le \sup _{w\in K} |f(w)|, \forall f\in {\mathcal {O}}_M\}. \end{aligned}$$

In the above, \({\mathcal {O}}_M\) denotes the ring of holomorphic functions on M. Then M is holomorphically convex iff for every compact \(K\subset M\), the hull \(\hat{K}_M\) is also compact.

Theorem 7

(Grauert [10]) The interior of every compact strongly pseudoconvex manifold M is holomorphically convex.

Definition 8

A complex manifold M is called Stein if it admits a proper holomorphic embedding in \(\mathbb {C}^n\) for some n.

Equivalently, M is Stein iff it is holomorphically convex and holomorphically separable: That is, for every distinct points \(x,y\in M\), there exists a holomorphic function \(f: M\rightarrow \mathbb {C}\) such that \(f(x) \ne f(y)\). We will use:

Theorem 9

(Rossi [13], Corollary on page 20) If a compact complex manifold M is strongly pseudoconvex and contains no compact complex subvarieties of positive dimension, then \({\text {int}}(M)\) is Stein.

We now discuss strong quasiconvexity and Stein property in the context of complex-hyperbolic manifolds. A classical example of a complex submanifold with Levi-convex boundary is a closed round ball \(\overline{\mathbf {B}}^n\) in \(\mathbb {C}^n\). Suppose that \(\varGamma < {\text {Aut}}(\mathbf {B}^n)\) is a discrete torsion-free subgroup of the group of holomorphic automorphisms of \(\mathbf {B}^n\) with (nonempty) domain of discontinuity \(\varOmega = \varOmega (\varGamma )\subset \partial \mathbf {B}^n\). The quotient

$$\begin{aligned} \overline{M}_\varGamma = (\mathbf {B}^n \cup \varOmega )/\varGamma \end{aligned}$$

is a smooth manifold with boundary.

Lemma 10

\(\overline{M}_\varGamma \) is strongly pseudoconvex.

Proof

We let \(T_\varLambda \) denote the union of all projective hyperplanes in \(P_\mathbb {C}^n\) tangent to \(\partial \mathbf {B}^n\) at points of \(\varLambda \), the limit set of \(\varGamma \). Let \(\widehat{\varOmega }\) denote the connected component of \(P_\mathbb {C}^n{\setminus }T_\varLambda \) containing \(\mathbf {B}^n\). It is clear that \(\mathbf {B}^n\cup \varOmega \subset {\widehat{\varOmega }}\) is strictly Levi-convex. By the construction, \(\varGamma \) preserves \(\widehat{\varOmega }\). It is proven in [5, Thm. 7.5.3] that the action of \(\varGamma \) on \({\widehat{\varOmega }}\) is properly discontinuous. Hence, \(X:={\widehat{\varOmega }}/\varGamma \) is a complex manifold containing \(\overline{M}_\varGamma \) as a strictly Levi-convex submanifold with boundary. \(\square \)

Specializing to the case when \(\overline{M}_\varGamma \) is compact, i.e. \(\varGamma \) is convex-cocompact, we obtain:

Proposition 11

Suppose that \(\varGamma \) is torsion-free, convex-cocompact and \(n>1\). Then:

  1. 1.

    \(\partial \overline{M}_\varGamma \) is connected.

  2. 2.

    If \({\text {int}}(\overline{M}_\varGamma ) = M_\varGamma \) contains no compact complex subvarieties of positive dimension, then \(M_\varGamma \) is Stein.

For example, as it was observed in [4], the quotient-manifold \(\mathbf {B}^2/\varGamma \) of a real-Fuchsian subgroup \(\varGamma < {\text {Aut}}(\mathbf {B}^2)\) is Stein while the quotient-manifold of a complex-Fuchsian subgroup \(\varGamma < {\text {Aut}}(\mathbf {B}^2)\) is non-Stein.

3 Proof of Theorem 1

In this section, we construct certain plurisubharmonic functions on \(M_\varGamma \), for each finitely generated, discrete subgroup \(\varGamma < {\text {Aut}}(\mathbf {B}^n)\) satisfying \(\delta (\varGamma ) < 2\). We use these functions to show that \(M_\varGamma \) has no compact subvarieties of positive dimension. At the end of this section, we prove the main result of this paper.

Let X be a complex manifold. Recall that a continuous function \(f:X\rightarrow \mathbb {R}\) is called plurisubharmonicFootnote 3 if for any homomorphic map \(\phi : V( \subset \mathbb {C}) \rightarrow X\), the composition \(f\circ \phi \) is subharmonic. Plurisubharmonic functions f satisfy the maximum principle; in particular, if f restricts to a nonconstant function on a connected complex subvariety \(Y\subset X\), then Y is noncompact.

Now, we turn to our construction of plurisubharmonic functions. Let \(\varGamma < {\text {Aut}}(\mathbf {B}^n)\) be a discrete subgroup. Consider the Poincaré series

$$\begin{aligned} \sum _{\gamma \in \varGamma } ({1 - |\gamma (z)|^2}), \quad z\in \mathbf {B}^n. \end{aligned}$$
(3)

Lemma 12

Suppose that \(\delta (\varGamma )<2\). Then (3) uniformly converges on compact sets.

Proof

Since \(\delta (\varGamma ) < 2\), the Poincaré series

$$\begin{aligned} \sum _{\gamma \in \varGamma } e^{ -2 d(0,\gamma (z))} \end{aligned}$$

uniformly converges on compact subsets in \(\mathbf {B}^n\). By (2), we get

$$\begin{aligned} e^{-2d(0,\gamma (z))} \le (1 - |\gamma (z)|^2) \le 4 e^{-2d(0,\gamma (z))}. \end{aligned}$$
(4)

Then, the result follows from the upper inequality. \(\square \)

Remark 13

Note that when \(\delta (\varGamma )>2\), or when \(\varGamma \) is of divergent type (e.g., convex-cocompact) and \(\delta (\varGamma ) = 2\), then (3) does not converge. This follows from the lower inequality of (4).

Assume that \(\delta (\varGamma )<2\). Define \(F: \mathbf {B}^n\rightarrow {\mathbb {R}}\),

$$\begin{aligned} F(z) = \sum _{\gamma \in \varGamma } ({ |\gamma (z)|^2 - 1}). \end{aligned}$$

Since F is \(\varGamma \)-invariant, i.e., \(F(\gamma z) = F(z)\), for all \(\gamma \in \varGamma \) and all \(z\in \mathbf {B}^n\), F descends to a function

$$\begin{aligned} f : M_\varGamma \rightarrow {\mathbb {R}}. \end{aligned}$$

Lemma 14

The function \(f : M_\varGamma \rightarrow {\mathbb {R}}\) is plurisubharmonic.

Proof

Enumerate \(\varGamma \) as \(\varGamma = \{\gamma _1,\gamma _2,\ldots \}\). Consider the sequence of partial sums of the series F:

$$\begin{aligned} S_k(z) = \sum _{j\le k} ({ |\gamma _j(z)|^2 - 1}). \end{aligned}$$

Since each summand in the above is plurisubharmonicFootnote 4, \(S_k\) is plurisubharmonic for each \(k\ge 1\). Moreover, the sequence of functions \(S_k\) is monotonically decreasing. Thus, the limit \(F = \lim _{k\rightarrow \infty } S_k\) is also plurisubharmonic, and hence so is f. \(\square \)

Note, however, that at this point, we do not yet know that the function f is nonconstant.

Now, we prove the main result of this section.

Theorem 15

Let \(\varGamma \) be a torsion-free discrete subgroup of \({\text {Aut}}(\mathbf {B}^n)\). If \(\delta (\varGamma ) < 2\), then \(M_\varGamma \) contains no compact complex subvarieties of positive dimension.

Proof

Suppose that Y is a compact connected subvariety of positive dimension in \(M_\varGamma \). Since \(\pi _1(Y)\) is finitely generated, so is its image \(\varGamma '\) in \(\varGamma =\pi _1(M_\varGamma )\). Since \(\delta (\varGamma ')\le \delta (\varGamma )\), by passing to the subgroup \(\varGamma '\) we can (and will) assume that the group \(\varGamma \) is finitely generated.

We construct a sequence of functions \(F_k: \mathbf {B}^n\rightarrow \mathbb {R}\) as follows. For \(k \in \mathbb {N}\), let \(\varSigma _k\subset \varGamma -\{1\}\) denote the subset consisting of \(\gamma \in \varGamma \) satisfying \(d(0,\gamma (0)) \le k\). Since \(\varGamma \) is a finitely generated linear group, it is residually finite and, hence, there exists a finite index subgroup \(\varGamma _k<\varGamma \) disjoint from \(\varSigma _k\). For each \(k\in \mathbb {N}\), define \(F_k: \mathbf {B}^n\rightarrow \mathbb {R}\) as the sum:

$$\begin{aligned} F_k(z) = \sum _{\gamma \in \varGamma _k} ({ |\gamma (z)|^2 - 1}). \end{aligned}$$

Since

$$\begin{aligned} \bigcap _{k\in \mathbb {N}} \varGamma _k=\{1\}, \end{aligned}$$

the sequence of functions \(F_k\) converges to \((|z|^2-1)\) uniformly on compact subsets of \(\mathbf {B}^n\). As before, each \(F_k\) is plurisubharmonic (cf. Lemmata 12, 14).

Let \({\widetilde{Y}}\) be a connected component of the preimage of Y under the projection map \(\mathbf {B}^n \rightarrow M_\varGamma \). Since \({\widetilde{Y}}\) is a closed, noncompact subset of \(\mathbf {B}^n\), the function \((|z|^2-1)\) is nonconstant on \({\widetilde{Y}}\). As the sequence \((F_k)\) converges to \((|z|^2-1)\) uniformly on compacts, there exists \(k\in \mathbb {N}\), such that \(F_k\) is nonconstant on \({\widetilde{Y}}\). Let \(f_k : M_k = M_{\varGamma _k} \rightarrow \mathbb {R}\) denote the function obtained by projecting \(F_k\) to \(M_k\), and \(Y_k\) be the image of \({\widetilde{Y}}\) under the projection map \(\mathbf {B}^n\rightarrow M_k\). Since \(M_k\) is a finite covering of \(M_\varGamma \), the subvariety \(Y_k\subset M_k\) is compact. Moreover, \(f_k\) is a nonconstant plurisubharmonic function on \(Y_k\) since \(F_k\) is such a function on \({\widetilde{Y}}\). This contradicts the maximum principle. \(\square \)

Remark 16

Regarding Remark 13: The failure of convergence of the series (3) as pointed out in Remark 13 is not so surprising. In fact, if \(\varGamma \) is a complex Fuchsian group, then \(\delta (\varGamma ) = 2\) and the convex core of \(M_\varGamma \) is a compact Riemann surface, see Example 4. Thus, our construction of F must fail in this case.

We conclude this section with a proof of the main result of this paper.

Proof of Theorem 1

By Theorem 15, \(M_\varGamma \) does not have compact complex subvarieties of positive dimensions. Then, by the second part of Proposition 11, \(M_\varGamma \) is Stein. \(\square \)

4 Further Remarks

In relation to Theorem 1, it is also interesting to understand the case when \(\delta (\varGamma ) = 2\), that is: For which convex-cocompact, torsion-free subgroups \(\varGamma \) of \({\text {Aut}}(\mathbf {B}^n)\) satisfying \(\delta (\varGamma )=2\), is the manifold \(M_\varGamma \) Stein? It has been pointed out before that a complex Fuchsian subgroup \(\varGamma < {\text {Aut}}(\mathbf {B}^n)\) satisfies \(\delta (\varGamma ) = 2\), but the manifold \(M_\varGamma \) is not Stein. In fact, the convex core of \(M_\varGamma \) is a complex curve, see Remark 16. We conjecture that complex Fuchsian subgroups are the only such non-Stein examples.

Conjecture 17

Let \(\varGamma <{\text {Aut}}(\mathbf {B}^n)\) be a convex-cocompact, torsion-free subgroup such that \(\delta (\varGamma ) = 2\). Then, \(M_\varGamma \) is non-Stein if and only if \(\varGamma \) is a complex Fuchsian subgroup.

We illustrate this conjecture in the following very special case: Let \(\phi : \pi _1(\varSigma ) \rightarrow {\text {Aut}}(\mathbf {B}^n)\) be a faithful convex-cocompact representation where \(\varSigma \) is a compact Riemann surface of genus \(g \ge 2\). Then \(\phi \) induces a (unique) equivariant harmonic map:

$$\begin{aligned} F: {\widetilde{\varSigma }} \rightarrow \mathbf {B}^n. \end{aligned}$$

which descends to a harmonic map \(f: \varSigma \rightarrow M_\varGamma \).

Proposition 18

Suppose that F is a holomorphic immersion. Then \(\varGamma = \phi (\pi _1(\varSigma ))\) satisfies \(\delta (\varGamma )\ge 2\). Moreover, if \(\delta (\varGamma )= 2\), then \(\varGamma \) preserves a complex line. In particular, \(\varGamma \) is a complex Fuchsian subgroup of \({\text {Aut}}(\mathbf {B}^n)\).

Proof

Noting that \(M_\varGamma \) contains a compact complex curve, namely, \(f(\varSigma )\), the first part follows directly from Theorem 1.

For the second part, we let Y denote the surface \({{\tilde{\varSigma }}}\) equipped with the Riemannian metric obtained via pull-back of the Riemannian metric g on \(\mathbf {B}^n\). The entropyFootnote 5h(Y) of Y is bounded above by \(\delta (\varGamma )\), that is

$$\begin{aligned} h(Y) \le 2. \end{aligned}$$
(5)

This can be seen as follows: The distance function \(d_Y\) on Y satisfies

$$\begin{aligned} d_Y(y_1,y_2) \ge d(F(y_1), F(y_2)). \end{aligned}$$

Therefore, the exponential growth-rate \(\delta _Y\) of \(\pi _1(\varSigma )\)-orbits in Y satisfies \(\delta _Y\le \delta (\varGamma )\). On the other hand, the quantity \(\delta _Y=h(Y)\) since \(\pi _1(\varSigma )\) acts cocompactly on Y.

Assume that \({\widetilde{\varSigma }}\) is endowed with a conformal Riemannian metric of constant \(-4\) sectional curvature. Since \({{\tilde{\varSigma }}}\) is a symmetric space, we have

$$\begin{aligned} h^2(Y)\mathrm {Area}(Y/\varGamma ) \ge h^2({\widetilde{\varSigma }})\mathrm {Area}(\varSigma ), \end{aligned}$$

see [1, p. 624]. The inequality (5) together with the above implies that

$$\begin{aligned} \mathrm {Area}(Y/\varGamma ) \ge \mathrm {Area}(\varSigma ). \end{aligned}$$

On the other hand, since \(f: Y/\varGamma \rightarrow M_\varGamma \) is holomorphic, \(4\cdot \mathrm {Area}(Y/\varGamma )\) equals to the Toledo invariant \(c(\phi )\) (see [14]) of the representation \(\phi \). Since \(c(\phi ) \le 4\pi (g-1)\), the inequality \(\mathrm {Area}(Y/\varGamma ) \ge \mathrm {Area}(\varSigma ) = \pi (g-1)\) shows that \(\mathrm {Area}(Y/\varGamma ) = \pi (g-1)\) or, equivalently, \(c(\phi ) = 4\pi (g-1)\). By the main result of [14], \(\varGamma \) preserves a complex-hyperbolic line in \(\mathbf {B}^n\). \(\square \)

Remark 19

The assumption that F is an immersion can be eliminated: Instead of working with a Riemannian metric, one can work with a Riemannian metric with finitely many singularities.

Motivated by Theorem 15, we also make the following conjecture.

Conjecture 20

If \(\varGamma < {\text {Aut}}(\mathbf {B}^n)\) is discrete, torsion-free, and \(\delta (\varGamma )<2k\), then \(M_\varGamma \) does not contain compact complex subvarieties of dimension \(\ge k\).

We conclude this section with a verification of this conjecture under a stronger hypothesis.

Proposition 21

If \(\varGamma < {\text {Aut}}(\mathbf {B}^n)\) is discrete, torsion-free, and \(\delta (\varGamma )<2k-1\), then \(M_\varGamma \) does not contain compact complex subvarieties of dimension \(\ge k\).

Proof

Note that if \(\varGamma \) is elementary (i.e., virtually abelian), then \(\delta (\varGamma ) = 0\). In this case, the result follows from Theorem 15. For the rest, we assume that \(\varGamma \) is nonelementary.

By [2, Sec. 4], there is a natural map \(f: M_\varGamma \rightarrow M_\varGamma \) homotopic to the identity map \({\mathrm {id}}_{M_\varGamma }: M_\varGamma \rightarrow M_\varGamma \) and satisfying

$$\begin{aligned} |\mathrm {Jac}_p(f)| \le \left( \frac{\delta (\varGamma )+1}{p}\right) ^p,\quad 2\le p\le 2n, \end{aligned}$$

where \(\mathrm {Jac}_p(f)\) denotes the p-Jacobian of f. When \(\delta (\varGamma )<2k-1\), we have \(|\mathrm {Jac}_p(f)|<1\), for \(p\in [2k,2n]\). This means that f strictly contracts the volume form on each p-dimensional tangent space at every point \(x\in M_\varGamma \), for \(p\in [2k,2n]\).

Let \(Y\subset M_\varGamma \) be a compact complex subvariety of dimension \(\ge k\) (real dimension \(\ge 2k\)). Then, Y is also a volume minimizer in its homology class. Since f strictly contracts volume on Y, f(Y) has volume strictly lesser than that of Y. However, f being homotopic to \({\mathrm {id}}_{M_\varGamma }\), f(Y) belongs to the homology class of Y. This is a contradiction to the fact that Y minimizes volume its homology class. \(\square \)

Remark 22

Note that Proposition 21 gives an alternative proof of Theorem 15 (hence Theorem 1) under a stronger hypothesis, namely \(\delta (\varGamma )\in (0,1)\). However, this method fails to verify Theorem 15 in the case when \(\delta (\varGamma )\in [1,2)\).

Finally, we note that the papers [6, 16] contain other interesting results and conjectures on Stein properties of complex-hyperbolic manifolds.