This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Masur’s Divergence for Higher Dimensional Tori and Kummer Surfaces

Zhijing Wang Yau Mathematical Sciences Centre
Tsinghua University
Beijing, China
[email protected]
Abstract.

Masur’s divergence states that the horizontal foliation of translation surfaces is uniquely ergodic if the geodesic flow is recurrent on the moduli space. This established a relationship between geometrical properties of foliations and the dynamics on the moduli space. In this paper, we extend this theorem to complex torus and Kummer surfaces. We define and calculate horizontal foliations and the corresponding geodesic flow in the moduli space of Kähler metrics and prove that the horizontal foliation is uniquely ergodic if the geodesic flow is recurrent. We also find the a necessary algebraic condition on the geodesic flow for the horizontal foliation to be uniquely ergodic.

1. Introduction

1.1. Motivations

Given a surface SgS_{g} with genus g>1g>1, its Teichmüller space Teich(Sg)\mathrm{Teich}(S_{g}) is defined as complex structures on SgS_{g} up to isotopy, which is equivalent to the space of hyperbolic metrics on SgS_{g} up to isotopy. The Teichmüller distance on the Teichmüller space is given by the logarithmic of the infimum dilatation among maps between the two complex structures: dTeich(X,Y)=inff:XYlogKfd_{Teich}(X,Y)=\inf_{f:X\to Y}\log K_{f} where the dilatation KfK_{f} of a map ff is given by Kf=suppX|fz|+|fz¯||fz||fz¯|K_{f}=\sup_{p\in X}\frac{|f_{z}|+|f_{\bar{z}}|}{|f_{z}|-|f_{\bar{z}}|} which at every point describes the ratio of the major and minor axes of the ellipse given by the image of infinitesimal circles under ff. The Teichmüller distance roughly measures how much two conformal structures are different.

Teichmüller showed that for every diffeomorphism on SgS_{g}, there is a unique map (differentiable outside finitely many points) in its isotopy class, called the Teichmüller map, minimizing the dilatation, thus achieving the Teichmüller distance.

The Teichmüller map is described explicitly by holomorphic quadratic differentials. For a Teichmüller map f:XYf:X\to Y, there are two holomorphic quadratic differentials qq on XX and pp on YY, such that under the natural coordinates of qq and pp that writes q=dz2q=dz^{2} and p=dw2p=dw^{2}, ff is of the form w=Kfz,w=1Kfz\Re w=\sqrt{K_{f}}\Re z,\Im w=\frac{1}{\sqrt{K_{f}}}\Im z. Given a quadratic differential qq on XX, taking the horizontal foliations and vertical foliations to be parallel to the xx and yy axes in the natural coordinates of qq and pp, we see that the Teichmüller map simply contracts the vertical foliations and expands the horizontal foliations.

The Teichmüller flow is given as the geodesic flow on the Teichmüller space under the Teichmüller distance. We also call it the geodesic flow on Teichmüller space and denote it by Rat(X,q)Ra_{t}(X,q) where the quadratic differential qq is identified with the cotangent space of Teich(Sg)\mathrm{Teich}(S_{g}). The Teichmüller flow gives a set of complete geodesics on the Teichmüller space, whose end points can be identified with the horizontal and vertical foliations of the quadratic differentials.

The moduli space of SgS_{g} is defined as the complex structures on SgS_{g} up to diffeomorphisms, so it is a quotient of the Teichmüller space by the mapping class group Γ=Diff(Sg)/Diff0(Sg)\Gamma=\mathrm{Diff}(S_{g})/\mathrm{Diff}_{0}(S_{g}) where Diff0(Sg)\mathrm{Diff}_{0}(S_{g}) is the group of diffeomorphisms on SgS_{g} and Diff0(Sg)\mathrm{Diff}_{0}(S_{g}) is the identity component of the diffeomorphism group, or diffeomorhpisms of SgS_{g} isotopic to identity. The mapping class group acts properly discontinuously on the Teichmüller space, so the geodesic flows descends to the moduli space.

Masur’s divergence established a relationship between unique ergodicity of foliations and the dynamics of the geodesic flow on the moduli space.

Theorem 1.1 ([Mas92]).

Let XX be a Riemann surface with quadratic differential qq. Suppose the Teichmüller flow Rat(X,q)Ra_{t}(X,q) is recurrent in the moduli space, then the horizontal foliation of (X,q)(X,q) is uniquely ergodic.

Recently, there have been many interesting results making analogies between Riemann surfaces and higher dimensional objects. For example, Cantat [Can14] established an analogy of stable and unstable foliations for pseudo-Anosov mappings, proving that that there exists stable and unstable currents contracted and expanded by the hyperbolic automorphisms of K3 surface. Filip [Fil20] made an analogy between periodic billiard trajectories on translation surfaces and special Lagrangian tori on K3 surfaces and counted the growth of the number of special Lagrangian tori in a twistor family.

1.2. Main results

In this article, we are interested in extending the definitions and properties of foliations and the Teichmüller flow to higher dimensional subjects. We start with the case of complex tori.

Unlike the case of Riemann surfaces, for higher dimensional objects, the space of complex structures up to isotopy and the space of flat (hyperbolic) metrics up to isotopy are different objects. For the complex torus, and hyperkähler manifolds (including K3 surfaces), the moduli space of complex structures is even highly non-Hausdorff. In fact, almost all complex structures are ergodic in the moduli space [Ver15]. Thus to study dynamics on the moduli space and to achieve geometric results related to the dynamics, the non-Hausdorff complex moduli space may not be the appropriate analogy.

In this article, we consider instead the moduli space of Kähler metrics on a differentiable manifold MM which is the set of equivalent Kähler structures of MM up to isotopy. We define this as

Kah(M)={(X,I,ω)}/\mathrm{Kah}(M)=\{(X,I,\omega)\}/\sim

where (X,I,ω)(X,I,\omega) are taken over Kähler manifolds XX equipped with complex structure II and Kähler form ω\omega such that XX is diffeomorphic to MM; the equivalence relation is defined as (X,I,ω)(Y,J,ν)(X,I,\omega)\sim(Y,J,\nu) if there exists a homeomorphism f:XYf:X\to Y such that fJ=If^{*}J=I and [fν]=[ω]H2(X,)[f^{*}\nu]=[\omega]\in H^{2}(X,\mathbb{R}).

For a complex torus, up to an isometry of Kähler structures, we may view the complex torus as n/Λ\mathbb{C}^{n}/\Lambda for a lattice Λ\Lambda of n\mathbb{C}^{n} with the standard Kähler metric ω0=i=1ndxidyi\omega_{0}=\sum_{i=1}^{n}dx_{i}\wedge dy_{i} on n\mathbb{C}^{n}. The moduli space of Kähler metrics of a 2n2n-torus 𝕋2n\mathbb{T}^{2n} is Kah(𝕋2n)=SL(2n,)\SL(2n,)/U(n)\mathrm{Kah}(\mathbb{T}^{2n})=\mathrm{SL}(2n,\mathbb{Z})\backslash\mathrm{SL}(2n,\mathbb{R})/\mathrm{U}(n).

The group SL(2n,)\mathrm{SL}(2n,\mathbb{R}) acts on SL(2n,)\SL(2n,)\mathrm{SL}(2n,\mathbb{Z})\backslash\mathrm{SL}(2n,\mathbb{R}) by right multiplication. This descends to the moduli space of Kähler structures of complex torus. More concretely, for a torus X=n/Λ(g)X=\mathbb{C}^{n}/\Lambda(g), we identify the lattice Λ(g)\Lambda(g) with a matrix gSL(2n,)g\in\mathrm{SL}(2n,\mathbb{R}) by taking a basis of Λ(g)\Lambda(g) to be the row vectors of gg and identifying n\mathbb{C}^{n} with 2n\mathbb{R}^{2n} by (z1=x1+iy2,,zn=xn+iyn)(x1,y1,,xn,yn)(z_{1}=x_{1}+iy_{2},...,z_{n}=x_{n}+iy_{n})\mapsto(x_{1},y_{1},...,x_{n},y_{n}).

We define the geodesic flow RatRa_{t} on the moduli space to be image of the action of at=diag(et,et,,et,et)a_{t}=\mathrm{diag}(e^{-t},e^{t},...,e^{-t},e^{t}) on SL(2n,)\SL(2n,)\mathrm{SL}(2n,\mathbb{Z})\backslash\mathrm{SL}(2n,\mathbb{R}) in the moduli space Kah(𝕋2n)\mathrm{Kah}(\mathbb{T}^{2n}), given as Rat(n/Λ(g))=n/Λ(gat)Ra_{t}(\mathbb{C}^{n}/\Lambda(g))=\mathbb{C}^{n}/\Lambda(ga_{t}). The geodesic flow is said to be recurrent if it visits a compact subset of the moduli space infinitely many times.

We define the horizontal and vertical foliations on n/Λ\mathbb{C}^{n}/\Lambda to be nn-dimensional foliations on the torus whose leaves are the subspaces in n\mathbb{C}^{n} parallel to the xx-axis and yy-axis respectively. A foliation is said to be uniquely ergodic if it only has one transverse measure up to a non-zero scalar multiplication. This is roughly saying that the foliation is equally distributed on the manifold. In this way, the geodesic flow acts on the torus n/Λ\mathbb{C}^{n}/\Lambda by contracting the horizontal foliations and expanding the vertical foliations.

Then we have the following analogy for Masur’s criterion.

Theorem 1.2.

Let X=n/Λ(g)X=\mathbb{C}^{n}/\Lambda(g) be a 2n2n-torus for gGL(2,)g\in GL(2,\mathbb{R}). Assume that the geodesic flow {RatX:t0}\{Ra_{t}X:t\leq 0\} is recurrent in the moduli space of Kähler metrics Kah(𝕋2n)\mathrm{Kah}(\mathbb{T}^{2n}). Then the horizontal foliation on XX must be uniquely ergodic.

This illustrates an interesting relationship between dynamical property of the geodesic flow on the geometric property of foliations on XX.

We remark that there are, of course, many ways to define geodesic flows on SL(2n,)\mathrm{SL}(2n,\mathbb{R}), and we choose at=diag(et,et,,et,et)a_{t}=\mathrm{diag}(e^{-t},e^{t},...,e^{-t},e^{t}) so that it acts on foliations similar to the way Teichmüller flow does for complex dimension 1. In fact, any action with exactly nn eigenvalues larger than 1 and nn eigenvalues smaller than 1 would have similar properties, one simply needs to define the horizontal and vertical foliations to be the respective eigen-spaces.

For real dimension 22 where X=𝕋2X=\mathbb{T}^{2}, Masur’s divergence is simply given by the fact that non-rational foliations are uniquely ergodic. (If a foliation is not uniquely ergodic, then a geodesic curve of XX must be a leaf of the foliation, so the length of the geodesic tends to zero along the geodesic flow RatXRa_{t}X which implies that RatXRa_{t}X must be divergent. ) For higher dimensions, the failure of unique ergodicity of the horizontal foliation does not imply the existence of a closed leaf of the foliation, so Theorem 1.2 cannot be simply deduced by the fact (See Theorem 4.1) that non-rational straight line flows on torus are uniquely ergodic.

Furthermore, the converse of Theorem 1.2 is not true for higher dimensional complex tori, as is the case for Riemann manifolds. For example, on the torus 2/Λ(g)\mathbb{C}^{2}/\Lambda(g) where Λ(g)\Lambda(g) is a lattice in 2\mathbb{C}^{2} spanned by the row vectors of g=(10i03+5i1+2i0i)g=\begin{pmatrix}1&0\\ i&0\\ \sqrt{3}+\sqrt{5}i&1+\sqrt{2}i\\ 0&i\end{pmatrix}, where i=1i=\sqrt{-1}, the horizontal foliation is uniquely ergodic but the geodesic flow starting at gg is divergent. For proof see Proposition 4.4.

We next extend the result to other complex manifolds. The simplest extension is to Kummer surfaces. A Kummer surface is defined as the minimal resolution of a complex torus XX mod involution. More explicitly, taking τ\tau to be the involution on 2/Λ(g)\mathbb{C}^{2}/\Lambda(g) descending from the inversion map (z1,z2)(z1,z2)(z_{1},z_{2})\to(-z_{1},-z_{2}) on 2\mathbb{C}^{2}, the Kummer surface Kum(g)\mathrm{Kum}(g) is given by the blow-up of (2/Λ(g))/τ(\mathbb{C}^{2}/\Lambda(g))/\tau at the 16 fixed points of τ\tau. We denote by Cj,j=1,,16C_{j},j=1,...,16 the 1616 exceptional curves given by the blow-ups of the fixed points. The orbifold (2/Λ(g))/τ(\mathbb{C}^{2}/\Lambda(g))/\tau before blow-up is called the exceptional Kummer surface. Kummer surfaces are examples of K3 surfaces.

For Kummer surfaces, we consider the moduli space of degenerate Kähler forms, i.e. Kähler forms defined outside some exceptional curves, denoted by Kah¯(M0)\overline{\mathrm{Kah}}(M_{0}) (see Section 5 for more details on this) where M0M_{0} is the differentiable manifold underlying Kummer surfaces. Moreover, we define the geodesic flow in the moduli space of Kummer surfaces to be given by that of the torus, i.e. RatKum(g)=Kum(gat)Ra_{t}\mathrm{Kum}(g)=\mathrm{Kum}(ga_{t}). We define the horizontal foliation and vertical foliation of a Kummer surface to be the image of the horizontal and vertical foliations of the torus.

Thus using our result for torus, we also establish the analogy of Masur’s divergence to Kummer surfaces.

Theorem 1.3.

Let XX be a Kummer surface. Assume that the geodesic flow {atX:t0}\{a_{t}X:t\leq 0\} is recurrent in the moduli space of degenerate Kähler metrics Kah¯(X)\overline{\mathrm{Kah}}(X). Then the horizontal foliation on XX must be uniquely ergodic.

The moduli space of degenerate Kähler metrics of a Kummer surface, which is the same as the moduli space of degenerate Kähler metrics of K3 surfaces, can be explicitly calculated via the period map, using the Torelli theorem.

We also have the following necessary condition for a horizontal foliation on the Kummer surface to be uniquely ergodic.

Theorem 1.4.

For a Kummer surface XX with flat Kähler form ω\omega, let h\mathcal{F}^{h} be its horizontal foliation. Then there exists cohomology classes η1,η2,η3H2(X,)\eta_{1},\eta_{2},\eta_{3}\in H^{2}(X,\mathbb{R}) such that the geodesic flow Rat(X)Ra_{t}(X) is of the form [e2tη1+e2tη2+1η3,ω]Kah¯(M0)[e^{-2t}\eta_{1}+e^{2t}\eta_{2}+\sqrt{-1}\eta_{3},\omega]\in\mathrm{\overline{Kah}}(M_{0}). Moreover, h\mathcal{F}^{h} is uniquely ergodic only if rank(η2H2(X,))<19rank(\eta_{2}^{\perp}\cap H^{2}(X,\mathbb{Z}))<19.

We remark that the cohomology classes η1\eta_{1}, η2\eta_{2} are on the boundary of the period domain and our result suggests that they can be thought of as representing the vertical and horizontal foliations. This gives an analogy of the identification of the boundary of Teichmüller space with horizontal and vertical foliations.

The form of the geodesic flow seems to suggest a generalization to general K3 surfaces. However, in [McM02] McMullen constructed examples of hyperbolic automorphisms of K3 surfaces that admits Siegel disks. The work of Cantat and Dupont [CD20] (for the projective case) and Filip and Tosatti [FT19] (for the general case) on Kummer rigidity also suggests that foliation preserved by automorphisms of K3 surfaces cannot have full Lebesgue measure except for the case of Kummer surfaces we discussed in this article. Thus, it may be too optimistic to hope for horizontal and vertical foliations on general K3, even outside divisors. Nevertheless, these results do not rule out the possibility of geodesic flows not containing automorphisms to have full measure horizontal and vertical foliations. Moreover, if we hope the geodesic flows to contain automorphism as in translation surfaces, there is still hope of developing the theory for laminations instead of foliations to be preserved by “geodesic flows”.

1.3. Organization

In Section 2 we recall the preliminaries on moduli spaces. In Section 3 we define the geodesic flow and foliations for complex torus. In Section 4 we prove Theorem 1.2 for torus and discuss the converse. In Section 5, we briefly recall the definitions and properties of K3 surfaces and Kummer surfaces, especially the period map and the generalized moduli spaces. In Section 6, we apply the results on torus to Kummer surfaces and calculate explicit form and properties of the geodesic flow.

Acknowledgements

I would like to thank my instructor Jinxin Xue and Yitwah Cheung for introducing me to the topic, for useful discussions and for comments on this article.

2. Moduli Spaces

First we give the definitions for the moduli spaces we shall be interested in, namely, the moduli space of complex structures and the moduli space of Kähler structures.

Definition 2.1.

For a differentiable manifold MM, we denote by Comp(M)\mathrm{Comp}(M) the space of complex structures on M, i.e.

Comp(M)={JEnd(TM),J2=1,[TJ1,0M,TJ1,0M]TJ1,0M}.\mathrm{Comp}(M)=\{J\in\mathrm{End}(TM),J^{2}=-1,[T^{1,0}_{J}M,T^{1,0}_{J}M]\subset T^{1,0}_{J}M\}.

We define its Teichmüller space to be the space of complex structures up to isotopy:

Teich(M)=Comp(M)/Diff0(M).\mathrm{Teich}(M)=\mathrm{Comp}(M)/\mathrm{Diff}_{0}(M).

The mapping class group of MM is defined as the diffeomorphisms of MM up to isotopy:

Γ(M)=Diff(M)/Diff0(M).\Gamma(M)=\mathrm{Diff}(M)/\mathrm{Diff}_{0}(M).

The moduli space of complex structures of MM is the space of complex structures up to diffeomorphisms

Mod(M)=Comp(M)/Diff(M)=Teich(M)/Γ(M).\mathrm{Mod}(M)=\mathrm{Comp}(M)/\mathrm{Diff}(M)=\mathrm{Teich}(M)/\Gamma(M).

The moduli space of Kähler structures of MM is defined as

Kah(M)={(X,I,ω)}/\mathrm{Kah}(M)=\{(X,I,\omega)\}/\sim

where (X,I,ω)(X,I,\omega) are taken over a Kähler manifold XX equipped with complex structure II and Kähler form ω\omega which is diffeomorphic to MM and the equivalence is defined as (X,I,ω)(Y,J,ν)(X,I,\omega)\sim(Y,J,\nu) if there exists a homeomorphism f:XYf:X\to Y such that fJ=If^{*}J=I and [fν]=[ω]PH2(X,)[f^{*}\nu]=[\omega]\in\mathrm{P}H^{2}(X,\mathbb{R}).

Remark 2.1.

For complex tori, the moduli space of Kähler structures is exactly the set of flat Kähler structures of volume 1 on MM. For Calabi-Yau manifolds, this is the set of Ricci-flat Kähler structures of volume 1 on MM as proved by Yau’s celebrated theorem that every Kähler class of a Calabi-Yau manifold admits a unique Ricci-flat Kähler metric [Yau77].

For the complex tori, the calculation of Teichmüller space, the mapping class group and the moduli spaces is a classical result.

We calculate that the Teichmüller space of an 2n2n-torus is Teich(𝕋2n)=GL(2n,)/GL(n,)\mathrm{Teich}(\mathbb{T}^{2n})=\mathrm{GL}(2n,\mathbb{R})/\mathrm{GL}(n,\mathbb{C}). Indeed, for every complex structure JJ on X𝕋2nX\simeq\mathbb{T}^{2n}, we may lift the complex structure to a complex structure J~\tilde{J} on 2n\mathbb{R}^{2n} which is invariant under translations by the fundamental group π1(X)\pi_{1}(X). Thus there exists an isotopy on 2n\mathbb{R}^{2n} commuting with π1(X)\pi_{1}(X) (so it descends to an isotopy on 𝕋2n\mathbb{T}^{2n}) that takes J~\tilde{J} to the standard complex structure J0J_{0} on n\mathbb{C}^{n}. This identifies X,JX,J with (n,J0)/π1(X)(\mathbb{C}^{n},J_{0})/\pi_{1}(X) in the Teichmüller space Teich(𝕋2n)\mathrm{Teich}(\mathbb{T}^{2n}). The lifting of the 2n2n generators of π1(X)\pi_{1}(X) in n\mathbb{C}^{n} may be expressed as an element of GL(2n,)\mathrm{GL}(2n,\mathbb{R}). These generators determine an element of the Teichmüller space of a 2n2n-torus up to complex linear transformation. Thus we have Teich(𝕋2n)=GL(2n,)/GL(n,)\mathrm{Teich}(\mathbb{T}^{2n})=\mathrm{GL}(2n,\mathbb{R})/\mathrm{GL}(n,\mathbb{C}).

Diffeomorphisms of the torus up to isotopy are given exactly by automoprhism of the fundamental group, so the mapping class group of an 2n2n-torus is Γ(𝕋n)=SL(2n,)\Gamma(\mathbb{T}^{n})=\mathrm{SL}(2n,\mathbb{Z}).

Thus the moduli space of complex structures is given by Mod(𝕋2n)=Teich(𝕋2n)/Γ=SL(2n,)\GL(2n,)/GL(n,)\mathrm{Mod}(\mathbb{T}^{2n})=\mathrm{Teich}(\mathbb{T}^{2n})/\Gamma=\mathrm{SL}(2n,\mathbb{Z})\backslash\mathrm{GL}(2n,\mathbb{R})/\mathrm{GL}(n,\mathbb{C}).

By similar arguments, the Kähler structures are always equivalent to the standard Kähler structure on (n/Λ,J0,ω0)(\mathbb{C}^{n}/\Lambda,J_{0},\omega_{0}) where Λ\Lambda is a lattice in n\mathbb{C}^{n}. And two lattices induce the same Kähler structure if and only if they differ by a unitary transformation and a rescaling. Thus the moduli space of flat Kähler structures is

Kah(𝕋2n)=SL(2n,)\GL(2n,)/(U(n)×)SL(2n,)\SL(2n,)/U(n).\mathrm{Kah}(\mathbb{T}^{2n})=\mathrm{SL}(2n,\mathbb{Z})\backslash\mathrm{GL}(2n,\mathbb{R})/(\mathrm{U}(n)\times\mathbb{R})\simeq\mathrm{SL}(2n,\mathbb{Z})\backslash\mathrm{SL}(2n,\mathbb{R})/\mathrm{U}(n).

We note that the complex moduli space of 𝕋2n\mathbb{T}^{2n} is non-Hausdorff for n>1n>1 since GL(n,)\mathrm{GL}(n,\mathbb{C}) acts ergodically on SL(2n,)\GL(2n,)\mathrm{SL}(2n,\mathbb{Z})\backslash\mathrm{GL}(2n,\mathbb{R}). The flat Kähler moduli space is Hausdorff since U(n)\mathrm{U}(n) is compact.

3. Geodesic Flow and Foliations on Complex Torus

As for the case of Riemann surfaces where we considered the natural coordinates, for higher dimensional object we may also define the geodesic flow and the horizontal and vertical foliations using the flat structures. In the case of complex tori, the flat structure readily comes from the universal covering n\mathbb{C}^{n}.

3.1. Foliations

First we briefly introduce the definition of measured foliations.

Definition 3.1.

A nn dimensional foliation \mathcal{F} on a 2n2n-dimensional differentiable manifold MM is defined as a decomposition of MM into a collection of disjoint union of subsets of MM, called the leaves of \mathcal{F}, such that for each pMp\in M there is a smooth chart from a neighborhood of pp to n\mathbb{C}^{n} that takes the leaves to the horizontal planes (n(n-planes parallel to the xx-axes)) with transition maps of the form (x1,y1,,xn,yn)(f1(x1,y1,..,xn,yn),g1(y1,,yn),.,(x_{1},y_{1},...,x_{n},y_{n})\to(f_{1}(x_{1},y_{1},..,x_{n},y_{n}),g_{1}(y_{1},...,y_{n}),....,fn(x1,y1,..,xn,yn),gn(y1,,yn))f_{n}(x_{1},y_{1},..,x_{n},y_{n}),g_{n}(y_{1},...,y_{n})).

A singular foliation is a foliation defined outside finitely many points.

A transverse measure μ\mu of a ((singular)) foliation \mathcal{F} is a map that assigns a non-negative real number to every smooth nn-submanifold of MM such that μ\mu is invariant under isotopies of MM that preserves each leaf of \mathcal{F} and μ\mu is absolutely continuous with respect to the nn-dimensional Lebesgue measure.

A measured (nn-)foliation (,μ)(\mathcal{F},\mu) on a 2n2n-manifold MM is a foliation \mathcal{F} equipped with a transverse measure μ\mu.

A foliation is said to be uniquely ergodic if it has only one transverse measure up to scale.

We shall be interested in very nice foliations, namely the horizontal and vertical foliations on 𝕋2n\mathbb{T}^{2n}, they are Lagrangian, or even special Lagrangian foliations. First we briefly introduce the definitions (See section 7, part 1 of [GHJ03] for detailed definition and examples on this).

Definition 3.2.

A foliation \mathcal{F} on a Kähler manifold (X,ω)(X,\omega) is called Lagrangian if the Kähler form vanishes when restricted to each leaf FF, i.e. ω|F=0\omega|_{F}=0.

In the case that XX is equipped with an holomorphic nn-form Ω\Omega, ((e.g. XX is a Calabi-Yau or hyperkähler manifold)), we say that a foliation \mathcal{F} is special Lagrangian if it is Lagrangian, the imaginary part of Ω\Omega vanishes and the real part of Ω\Omega is positive when restricted to each leaf FF, i.e. ω|F=0,(Ω)|F=0,(Ω)|F>0\omega|_{F}=0,\Im(\Omega)|_{F}=0,\Re(\Omega)|_{F}>0 for each leaf FF.

Now we define the horizontal and vertical foliations on complex tori in the following way. This is motivated by the definition of horizontal and vertical foliations on Riemann surfaces which is given by the natural coordinates.

We consider the universal covering of the torus. For g=(gij)1i,j2nSL(2n,)g=(g_{ij})_{1\leq i,j\leq 2n}\in\mathrm{SL}(2n,\mathbb{R}) we take Λ(g)\Lambda(g) to be the lattice in n\mathbb{C}^{n} given by the row vectors of

(1) (g11+g12ig13+g14ig1,2n1+g1,2nig21+g22ig23+g24ig2,2n1+g2,2nig2n,1+g2n,2ig2n,3+g2n,4ig2n,2n1+g2n,2ni)\begin{pmatrix}g_{11}+g_{12}i&g_{13}+g_{14}i&...&g_{1,2n-1}+g_{1,2n}i\\ g_{21}+g_{22}i&g_{23}+g_{24}i&...&g_{2,2n-1}+g_{2,2n}i\\ \vdots&\vdots&\ddots&\vdots\\ g_{2n,1}+g_{2n,2}i&g_{2n,3}+g_{2n,4}i&...&g_{2n,2n-1}+g_{2n,2n}i\end{pmatrix}

where i=1i=\sqrt{-1}. We take nn/Λ(g)=X\mathbb{C}^{n}\to\mathbb{C}^{n}/\Lambda(g)=X the universal covering of the complex torus XX given by gg. The leaves of the horizontal foliation h\mathcal{F}^{h} on XX given by gg are the images of {(x1+iy1,,xn+iyn)2:y1=c1,,yn=cn}n\{(x_{1}+iy_{1},...,x_{n}+iy_{n})\in\mathbb{C}^{2}:y_{1}=c_{1},...,y_{n}=c_{n}\}\subset\mathbb{C}^{n} in XX and the transverse measure is given by dμh=dy1dynd\mu_{h}=dy_{1}\wedge...\wedge dy_{n}. Respectively, the vertical foliation v\mathcal{F}^{v} is given by {(x1+iy1,,xn+iyn)2:x1=c1,,xn=cn}\{(x_{1}+iy_{1},...,x_{n}+iy_{n})\in\mathbb{C}^{2}:x_{1}=c_{1},...,x_{n}=c_{n}\} in XX with transverse measure dμv=dx1dxnd\mu_{v}=dx_{1}\wedge...\wedge dx_{n}. The horizontal and vertical foliations are transverse Lagrangian foliations on XX, where the Kähler form is given by ω0\omega_{0} where ω0=dxidyi\omega_{0}=\sum dx_{i}\wedge dy_{i} is the standard one on n\mathbb{C}^{n}.

We note that the horizontal and vertical foliations are taken with respect to gSL(2n,)g\in\mathrm{SL}(2n,\mathbb{R}). If we have a complex torus XX with a flat Kähler metric ω\omega to begin with and require that (X,ω)=(n/Λ(g),ω0)(X,\omega)=(\mathbb{C}^{n}/\Lambda(g),\omega_{0}), then there is a U(n)\mathrm{U}(n) set of possible choices for gg and a U(n)/O(n)\mathrm{U}(n)/\mathrm{O}(n) set of choices for the horizontal and vertical foliations. In fact, in the case n=2n=2, every linear special Lagrangian foliation in a flat Kähler torus can be given as the horizontal foliation for some choice of gSL(2n,)g\in\mathrm{SL}(2n,\mathbb{R}) (c.f. Proposition 6.1).

3.2. Geodesic flow

The geodesic flow is given as expansion of the vertical foliations and contraction of the horizontal foliations. This is motivated by the action of Teichmüller flow on natural coordinates. We may also write this more explicitly in the language of a homogeneous flow.

Definition 3.3.

We denote by RatRa_{t} the geodesic flow on the homogeneous space SL(2n,)\SL(2n,)\mathrm{SL}(2n,\mathbb{Z})\backslash\mathrm{SL}(2n,\mathbb{R}) starting at ySL(2n,)\SL(2n,)y\in\mathrm{SL}(2n,\mathbb{Z})\backslash\mathrm{SL}(2n,\mathbb{R}) given by Rat(y)=yatRa_{t}(y)=ya_{t} where

at=diag(et,et,,et,et)a_{t}=diag(e^{-t},e^{t},...,e^{-t},e^{t})

is the diagonal matrix of entries ete^{-t} and ete^{t} on the real and imaginary parts respectively.

For a complex torus XX expressed as n/Λ(g)\mathbb{C}^{n}/\Lambda(g) for gSL(2n,)g\in\mathrm{SL}(2n,\mathbb{R}), the geodesic flow (depending on gg) is defined as

Rat(X,ω)=(n/Λ(Ratg),ω0)=(n/Λ(gat),ω0)Ra_{t}(X,\omega)=(\mathbb{C}^{n}/\Lambda(Ra_{t}g),\omega_{0})=(\mathbb{C}^{n}/\Lambda(ga_{t}),\omega_{0})

Take X=n/Λ(g)X=\mathbb{C}^{n}/\Lambda(g) and consider the horizontal and vertical foliations of XX given by nX\mathbb{C}^{n}\to X, we see that the geodesic flow RatR_{a_{t}} preserves the horizontal and vertical foliations and while contracting and expanding the transverse measures.

We may project the geodesic flow down to the moduli spaces of the torus by projecting SL(2n,)\SL(2n,)\mathrm{SL}(2n,\mathbb{Z})\backslash\mathrm{SL}(2n,\mathbb{R}) down to the Kähler moduli space Kah(𝕋2n)SL(2n,)\SL(2n,)/U(n)\mathrm{Kah}(\mathbb{T}^{2n})\simeq\mathrm{SL}(2n,\mathbb{Z})\backslash\mathrm{SL}(2n,\mathbb{R})/\mathrm{U}(n), the Kähler moduli space of complex torus. More explicitly, this is given by the right multiplication Rat((n,ω0)/Λ(g))=(n,ω0)/Λ(gat)Ra_{t}((\mathbb{C}^{n},\omega_{0})/\Lambda(g))=(\mathbb{C}^{n},\omega_{0})/\Lambda(ga_{t}), where ω0\omega_{0} is the standard Kähler form on n\mathbb{C}^{n}.

4. Unique Ergodicity versus Recurrence

Now we are ready to prove our analogy of Masur’s divergence on the case of complex torus. The rough idea is to approximate the transverse measure of a transversal sub-manifold with the number of intersections of the transversal sub-manifold with a higher-dimensional “segment” of the foliation.

By the translation structure of torus given by n/Λ(g)\mathbb{C}^{n}/\Lambda(g) we may also define the straight line flow on the torus to be the image of the straight line flow on n\mathbb{C}^{n} under the projection. More precisely, the straight line flow from xX=n/Λ(g)x\in X=\mathbb{C}^{n}/\Lambda(g) in the direction vnv\in\mathbb{C}^{n} is defined as ftv(x+Λ(g))=(x+tv)+Λ(g)f^{v}_{t}(x+\Lambda(g))=(x+tv)+\Lambda(g).

Proof of Theorem 1.2.

Suppose RatXRa_{t}X is recurrent in the moduli space of Kähler structures, we hope to prove unique ergodicity of the horizontal foliation h\mathcal{F}^{h}.

Suppose a subsequence of RatXRa_{t}X converges to XX_{\infty} in Kah(Tn)\mathrm{Kah}(T^{n}) and take htm:gtmXXh_{t_{m}}:g_{t_{m}}X\to X_{\infty} to be diffeomorphisms with local differentials of norm close to 11 for tnt_{n} large enough.

The horizontal foliation is uniquely ergodic if and only if the straight line flows ftv,tf^{v}_{t},t\in\mathbb{R} are uniquely ergodic, where vv is taken over all possible directions in the tangent space TpFT_{p}F of a leaf FF\in\mathcal{F}. Note that since the foliation is linear, the tangent space does not depend on the points of the leaves.

We prove by contradiction. Suppose μ\mu and ν\nu are different ergodic measures of the straight line flow that give different transverse measures on a sub-manifold Q0Q_{0} transverse to h\mathcal{F}^{h}. Take a generic point xx of μ\mu and a generic point yy of ν\nu. For any basis vi,1inv_{i},1\leq i\leq n of the space given by h\mathcal{F}^{h}, and for a flow box Q={i=1nftivix:0tiδ,xQ0}Q=\{\prod_{i=1}^{n}f^{v_{i}}_{t_{i}}x:0\leq t_{i}\leq\delta,x\in Q_{0}\}, we have by the Birkhoff ergodic theorem

1Ti0T10Tn1Q((ftivi)(x))𝑑t1𝑑tnμ(Q)\frac{1}{\prod T^{i}}\int_{0}^{T^{1}}...\int_{0}^{T^{n}}1_{Q}(\prod(f^{v_{i}}_{t_{i}})(x))dt_{1}...dt_{n}\to\mu(Q)

and

1Ti0T10Tn1Q((ftivi)(y))𝑑t1𝑑tnν(Q)\frac{1}{\prod T^{i}}\int_{0}^{-T^{1}}...\int_{0}^{-T^{n}}1_{Q}(\prod(f^{v_{i}}_{t_{i}})(y))dt_{1}...dt_{n}\to\nu(Q)

as TiT^{i}\to\infty for any QXQ\subset X.

Take RR to be the 2n2n-cube with xx_{\infty} and yy_{\infty} as diagonal vertices in XX_{\infty} whose edges are parallel to viv_{i} in h\mathcal{F}^{h} or parallel to wiw_{i} in v\mathcal{F}^{v}.

Suppose the lengths of the edges in h\mathcal{F}^{h} are lil^{i} in the direction viv_{i}, and the edges in v\mathcal{F}^{v} are of length at most ll^{-} in the direction wiw_{i}. Then htm1Rh_{t_{m}}^{-1}R is a cube in gtmXg_{t_{m}}X. We take RmR_{m} to be the cube in gtmXg_{t_{m}}X with diagonal vertices gtmxg_{t_{m}}x and gtmyg_{t_{m}}y whose edges are of lengths lmil^{i}_{m} and lml^{-}_{m} and directions viv_{i} and vv_{-}. By our assumption, RmR_{m} is close to htm1Rh_{t_{m}}^{-1}R and we have lmilil^{i}_{m}\to l^{i}, lmll^{-}_{m}\to l^{-}, as nn\to\infty. Now consider gtm1RmXg_{t_{m}}^{-1}R_{m}\subset X, it will become a very thin cube with xx and yy as diagonal vertices and edges of lengths Tni=etmlmiT_{n}^{i}=e^{-t_{m}}l^{i}_{m}\to\infty and etmlm0e^{t_{m}}l^{-}_{m}\to 0 as tmt_{m}\to-\infty. And we see that for a transverse sub-manifold Q0Q_{0} of h\mathcal{F}^{h} the nn-face in h\mathcal{F}^{h} starting from xx and the nn-face in h\mathcal{F}^{h} ending with yy intersects QQ approximately the same amount of times, unless xx is very close to the boundary Q0\partial Q_{0}. Take viv_{i} so that the nn-face of the cube in h\mathcal{F}^{h} is non-degenerate.

Now take QQ to be Q={i=1nftivix:0tiδ,xQ0}Q=\{\prod_{i=1}^{n}f^{v_{i}}_{t_{i}}x:0\leq t_{i}\leq\delta,x\in Q_{0}\}. Then we have 1Q((ftivi)(x))=1Q((fsivi)(y))1_{Q}(\prod(f^{v_{i}}_{t_{i}})(x))=1_{Q}(\prod(f^{v_{i}}_{s_{i}})(y)) for tisi=Tmit_{i}-s_{i}=T^{i}_{m} unless (ftivi)(x)Qϵ\prod(f^{v_{i}}_{t_{i}})(x)\in Q_{\epsilon} where Qϵ={i=1nftivix:0tiδ,xQ0,dist(x,Q0)<ϵ}Q_{\epsilon}=\{\prod_{i=1}^{n}f^{v_{i}}_{t_{i}}x:0\leq t_{i}\leq\delta,x\in Q_{0},dist(x,Q_{0})<\epsilon\} and ϵcetm\epsilon\geq ce^{-t_{m}}. Here cc depends only on vv_{-} and yy. Thus we have

1Tmi0Tm10Tmn(1Q((ftivi)(x))1Q((ftiTmivi)(y)))𝑑t1𝑑tn\mid\frac{1}{\prod T^{i}_{m}}\int_{0}^{T^{1}_{m}}...\int_{0}^{T^{n}_{m}}(1_{Q}(\prod(f^{v_{i}}_{t_{i}})(x))-1_{Q}(\prod(f^{v_{i}}_{t_{i}-T^{i}_{m}})(y)))dt_{1}...dt_{n}\mid
1Tmi0Tm10Tmn1Qϵ((ftivi)(x))𝑑t1𝑑tnμ(Qϵ)\leq\frac{1}{\prod T^{i}_{m}}\int_{0}^{T^{1}_{m}}...\int_{0}^{T^{n}_{m}}1_{Q_{\epsilon}}(\prod(f^{v_{i}}_{t_{i}})(x))dt_{1}...dt_{n}\to\mu(Q_{\epsilon})

as nn\to\infty. Thus we get |μ(Q)ν(Q)||μ(Qϵ)||\mu(Q)-\nu(Q)|\leq|\mu(Q_{\epsilon})| for any distinct ergodic probability measures μ\mu and ν\nu. Since μ\mu supports the Lebesgue measure, we have μ(Qϵ)0\mu(Q_{\epsilon})\to 0 as ϵ0\epsilon\to 0. This contradicts to our assumption. Thus the horizontal foliation h\mathcal{F}^{h} can only have a unique transverse measure up to scale. This proves our theorem. ∎

Now we shall consider in greater details when exactly are foliations uniquely ergodic and when are the corresponding geodesic flow recurrent.

For complex tori, the unique ergodicity of irrational straight line flow is a classical result.

Theorem 4.1.

The straight line flow on a complex tori is uniquely ergodic if and only if it is totally irrational, i.e. the direction of the straight line flow is not in the span of any lower-dimensional sub-lattice of Λ(g)\Lambda(g).

Thus for linear foliation we have the following corollary.

Corollary 4.2.

The horizontal foliation is either uniquely ergodic and thus minimal, or non-minimal where each leaf is rational in the sense that it is contained in the span of any lower-dimensional sub-lattice of Λ(g)\Lambda(g).

Proof.

If the horizontal foliation is contained in a subspace V2nV\subset\mathbb{R}^{2n} which is the span of a lower-dimensional sub-lattice of Λ(g)\Lambda(g) then it cannot be uniquely ergodic: take VV^{\prime} is a subspace of 2n\mathbb{R}^{2n} spanned by another sub-lattice of Λ(g)\Lambda(g) so that VV=2nV^{\prime}\oplus V=\mathbb{R}^{2n}, then any measure on VV^{\prime} defines a transverse measure of the horizontal foliation.

However if the horizontal foliation is contained any rational subspace then there must be a direction in the foliation that is totally irrational, so by the unique ergodicity of totally irrational straight line flows the foliation is also uniquely ergodic.

On the moduli space of flat Kähler metrics, however, there doe not exist such a clear distinction between divergent and non-divergent geodesic flows. However, we do know that geodesic flows generated by horizontal foliations are“almost always” ergodic by the following theorem.

Theorem 4.3 (See [Moo66]).

The right multiplication RatRa_{t} is ergodic on SL(2n,)\SL(2n,)\mathrm{SL}(2n,\mathbb{Z})\backslash\mathrm{SL}(2n,\mathbb{R}) with respect to the Haar measure.

This implies that almost every geodesic flow is dense in the Kähler moduli space, hence recurrent.

The converse of Theorem 1.2 is not true: there exists divergent geodesic flows that are generated by uniquely ergodic horizontal and vertical foliations. For example, if a leaf of the horizontal foliation contains a closed curve in the torus with non-trivial topology then the geodesic flow must be divergent in the moduli space of flat metrics. However, the foliation itself may still be uniquely ergodic as long as one direction in the foliation is totally irrational. One example is given as follows.

Proposition 4.4.

Take

g=(1000010035120001)g=\begin{pmatrix}1&0&0&0\\ 0&1&0&0\\ \sqrt{3}&\sqrt{5}&1&\sqrt{2}\\ 0&0&0&1\end{pmatrix}

and X=2/Λ(g)X=\mathbb{C}^{2}/\Lambda(g). Then the horizontal foliation of X=2/Λ(g)X=\mathbb{C}^{2}/\Lambda(g) is uniquely ergodic, but the geodesic flow Rat(X,ω0)Kah(𝕋4)Ra_{t}(X,\omega_{0})\in\mathrm{Kah}(\mathbb{T}^{4}) is divergent.

Proof.

The horizontal foliation is unique ergodic since (0,0,1,0)(0,0,1,0) is a totally irrational linear sum of the row vectors of gg. Thus the straight line flow in this direction is uniquely ergodic.

However, the geodesic flow starting at gg is divergent since the length of the closed curve given by the vector (1,0,0,0)(1,0,0,0) tends to zero as tt\to\infty. ∎

5. Preliminaries on Kummer Surface

Next, we further exploit the form and property of the geodesic flow for Kummer surfaces in complex dimension 2. In this case, we may, by the Torelli theorem, explicitly calculate the geodesic flow and deduce some properties. First we introduce some preliminaries on Kummer surfaces and K3 surfaces, period map and the complex and generalized Kähler moduli spaces.

5.1. K3 surfaces and its moduli spaces

Here we introduce the definition and basic properties and Kummer surface and K3 surface, including the Torelli theorem and the denseness of Kummer surface in the complex moduli space of K3.

Definition 5.1.

A K3 surface is a complex surface XX which is simply connected and has a unique ((up to scale)) non-vanishing holomorphic 22-form ΩX\Omega_{X}.

Theorem 5.1 (See Theorem 7.1.1 of [Huy16]).

All K3 surfaces are diffeomorphic.

From now on we denote the differentiable manifold underlying a K3 surface by M0M_{0}. For any K3 surface XX, we define a marking ff of XX to be a diffeomorphism f:M0Xf:M_{0}\to X. A marking ff identifies the homology space H2(X,)H_{2}(X,\mathbb{Z}) with H2(M0,)H_{2}(M_{0},\mathbb{Z}) through ff_{*}.

Definition 5.2 (Period Domain).

Let M0M_{0} be the differentiable manifold underlying K3 surfaces. The period domain of K3 surfaces is the set of Hodge structure on H2(M0,)H^{2}(M_{0},\mathbb{C}), denoted as Per(M0)={f(H2,0(X,))H2(M0,):f:M0X a diffeomorphism}H2(M0,)\mathrm{Per}(M_{0})=\{f^{*}(H^{2,0}(X,\mathbb{C}))\subset H^{2}(M_{0},\mathbb{C}):f:M_{0}\to X\text{ a diffeomorphism}\}\subset\mathbb{P}H^{2}(M_{0},\mathbb{C}).

The period map from the Teichmüller space of M to the period domain is given by Per:Teich(M0)Per(M0):[(f,X)]f(H2,0(X,))\mathrm{Per}:\mathrm{Teich}(M_{0})\to\mathrm{Per}(M_{0}):[(f,X)]\to f^{*}(H^{2,0}(X,\mathbb{C})).

The relation between the period domain, the Teichmüller space and the moduli space of complex structures of K3 surfaces is described by the Torelli theorem.

Theorem 5.2 (Torelli Theorem).

The period domain of K3 surfaces is given by Per(K3)={αH2(M0,):αα=0,αα¯>0}.\mathrm{Per}(K3)=\{\alpha\in\mathbb{P}H^{2}(M_{0},\mathbb{C}):\alpha\cdot\alpha=0,\alpha\cdot\bar{\alpha}>0\}. The period map Per\mathrm{Per} is a local isomorphism. Two complex K3 surfaces XX and XX^{\prime} are bi-holomorphic if and only if there exists a Hodge isometry H2(X,)H2(X,)H^{2}(X,\mathbb{Z})\to H^{2}(X^{\prime},\mathbb{Z}).

Noting that mapping class group is generated by change of markings on the Teichmüller space, the Torelli theorem give us the following description of the moduli space.

Corollary 5.3.

The mapping class group of K3 surfaces is given by Γ=O(H2(M0,))\Gamma=O(H^{2}(M_{0},\mathbb{Z})) which is a lattice in O(H2(M0,))O(H^{2}(M_{0},\mathbb{R})). The moduli space of complex structures of K3 surfaces is Mod(K3)=Γ\Per(K3)\mathrm{Mod}(K3)=\Gamma\backslash\mathrm{Per}(K3).

Remark 5.4.

We have a bijection between Hodge structures and oriented positive 2-planes in the real cohomology space of M0M_{0}, given by ΩH2,0(X,)span(Ω,Ω)H2(X,)\Omega\in H^{2,0}(X,\mathbb{C})\to span(\Re\Omega,\Im\Omega)\subset H^{2}(X,\mathbb{R}). Indeed, the condition ΩΩ=0,ΩΩ¯>0\Omega\cdot\Omega=0,\Omega\cdot\bar{\Omega}>0 is equivalent to ΩΩ=ΩΩ>0,ΩΩ=0\Re\Omega\cdot\Re\Omega=\Im\Omega\cdot\Im\Omega>0,\Re\Omega\cdot\Im\Omega=0 which is equivalent to the two plane being positive and a complex scalar for Ω\Omega is simply rotation and real rescale of basis in the 2-plane. So we may identify the period domain with the Grassmannian of oriented positive two-planes in H2(M0,)H^{2}(M_{0},\mathbb{R}). More precisely, this gives an isomorphism Per(K3)SO(3,19,)/(SO(2)×SO(1,19))\mathrm{Per}(K3)\simeq SO(3,19,\mathbb{R})/(SO(2)\times SO(1,19)). The complex moduli space Γ\Per(K3)\Gamma\backslash\mathrm{Per}(K3) is highly non-Hausdorff, see [Ver15] for a more detailed discussion on this.

If we add the Kähler structure into consideration, the Kähler class and the Hodge structure together determine a positive 3-plane in the real cohomology space together with a direction in it specifying the Kähler class. So we have a map from the moduli space of Kähler structures to the space of positive 3-Grassmannian with a specified unit vector up to change of marking. We write this as Kah(K3)Γ\SO(3,19,)/(SO(2)×SO(19))\mathrm{Kah}(K3)\to\Gamma\backslash SO(3,19,\mathbb{R})/(SO(2)\times SO(19)).

However, this map is not surjective since a Kähler class cannot have zero intersection with curves. One way to solve this is to allow for degenerate Kähler forms, i.e. non-degenerate symplectic 2-forms that are in the positive cone but on the boundaries of the Kähler cone.

Definition 5.3 (Degenerate Kähler Form [Kob90]).

We define a degenerate Kähler form on a K3 surface XX to be a Kähler-Einstein orbifold metric on XX which is given by a Kähler-Einstein form on YY where XYX\to Y is a minimal resolution of YY and YY is a compact complex surface with at most rational double points. We call such YY a generalized K3 surface.

We define the period domain of degenerate Kähler forms of K3 to be KPer={(α,ω)H2(M0,)×H2(M0,):αPer(K3),αω=0,ωω>0}\mathrm{KPer}=\{(\alpha,\omega)\in\mathbb{P}H^{2}(M_{0},\mathbb{C})\times\mathbb{P}H^{2}(M_{0},\mathbb{R}):\alpha\in\mathrm{Per}(K3),\alpha\cdot\omega=0,\omega\cdot\omega>0\} this is equivalent to pairs {(V,ω)}\{(V,\omega)\} where VV is taken over positive 3-Grassmannians in H2(M0,)H^{2}(M_{0},\mathbb{R}) and ω\omega are taken over vectors in VV. So we have the identification KPerSO(3,19)/(SO(2)×SO(19))\mathrm{KPer}\simeq SO(3,19)/(SO(2)\times SO(19)). And we take the period map as (X,ω)(Per(X),[ω])(X,\omega)\mapsto(\mathrm{Per}(X),[\omega]) from the space of marked K3 surfaces with degenerate Kähler forms to the period domain.

The following proposition shows that after adding in the degenerate Kähler forms, every positive 3-plane in the real cohomology space can indeed be realized by some K3 surface.

Proposition 5.5 (Torelli Thoerem for Generalized Kähler Polarized K3 [Kob90]).

The period map (X,ω)(Per(X),[ω])(X,\omega)\to(\mathrm{Per}(X),[\omega]) from marked K3 surfaces with degenerate Kähler forms to its period domain KPer\mathrm{KPer} is surjective. Moreover, if F:H2(X,)H2(X,)F:H^{2}(X,\mathbb{Z})\to H^{2}(X^{\prime},\mathbb{Z}) is an isometry ((with respect to inner product given by intersection)) whose extension to H2(X,)H2(X,)H^{2}(X,\mathbb{C})\to H^{2}(X^{\prime},\mathbb{C}) preserves the Hodge structure and the Kähler class, then there is a unique isomoprhism f:XXf:X\to X^{\prime} such that F=fF=f^{*}.

Adding the change of markings into consideration, we have the following.

Corollary 5.6.

The moduli space of generalized Kähler forms of K3 surfaces is given by

Kah(M0)¯Γ\KPer\mathrm{\overline{Kah(M_{0})}}\simeq\Gamma\backslash\mathrm{KPer}
Γ\SO(3,19)/(SO(2)×SO(19)).\simeq\Gamma\backslash SO(3,19)/(SO(2)\times SO(19)).

This gives a homogeneous description of the moduli space of generalized Kähler froms of K3 surfaces. Note that this space is Hausdorff while the complex moduli space of K3 surfaces is not. Elements in the generalized period domain KPer\mathrm{KPer} leaves the compact sets exactly when the positive 33-plane approaches the light cone in the real cohomology space.

5.2. Kummer Surfaces

Kummer surfaces is a special class of K3 surfaces with a very concrete construction.

Definition 5.4 (Kummer surface).

Given a complex torus A=2/Λ(g)A=\mathbb{C}^{2}/\Lambda(g) where Λ(g)\Lambda(g) is the lattice in 2\mathbb{C}^{2} generated by the row vectors of gGL(4,)g\in\mathrm{GL}(4,\mathbb{R}), we may consider the involution τ:(z,w)(z,w)\tau:(z,w)\to(-z,-w) on n\mathbb{C}^{n} which has 1616 fixed points in AA. The Kummer minifold Kum(g)\mathrm{Kum}(g) is defined as the minimal resolution of A/τA/\tau where A=2/Λ(g)A=\mathbb{C}^{2}/\Lambda(g) is a complex torus given by gGL(4,)g\in\mathrm{GL}(4,\mathbb{R}). The minimal resolution is given by blowing up the 1616 fixed points of the involution τ\tau.

We call A/τA/\tau the exceptional Kummer surface and we call the blow-up of the 1616 fixed points the exceptional curves of the Kummer surface Kum(g)\mathrm{Kum}(g). We denote the exceptional curves by Ci,1i16C_{i},1\leq i\leq 16.

Indeed, we may check that the image of the holomorphic 2-form dzdwdz\wedge dw extends to a non-vanishing holomorphic 2-form on the Kummer surface and that the Kummer surface is simply connected. Therefore it is a K3 surface.

Next we fix a marking on Kummer surfaces for simplicity for calculation.

For any gSL(4,)g\in\mathrm{SL}(4,\mathbb{R}), we identify its row vectors with complex vectors in 2\mathbb{C}^{2} to obtain a lattice Λ(g)\Lambda(g), and take a marked Kummer surface (ϕg,Kum(g))(\phi_{g},\mathrm{Kum}(g)) in the following way. For the identity element ISL(4,)I\in\mathrm{SL}(4,\mathbb{R}) we fix a marking ϕI:Kum(I)M0\phi_{I}:\mathrm{Kum}(I)\to M_{0}, and for any gSL(4,)g\in\mathrm{SL}(4,\mathbb{R}), we take ϕg=gϕI\phi_{g}=g_{*}\phi_{I} where we consider gg acting on the left on 2\mathbb{C}^{2} as a homeomorphism 2/Λ(I)2/Λ(g)\mathbb{C}^{2}/\Lambda(I)\to\mathbb{C}^{2}/\Lambda(g) and gg_{*} is the corresponding homeomorphism between the Kummer surfaces. We denote by CiC_{i} the image of the 16 exceptional curves of Kum(I)\mathrm{Kum}(I) in M0M_{0} under ϕI\phi_{I}, then by the way we defined ϕg\phi_{g} they are also the image of the 16 exceptional curves of Kum(g)\mathrm{Kum}(g) under ϕg\phi_{g}. We denote by ωg\omega_{g} the degenerate Kähler form on the Kummer surface Kum(g)\mathrm{Kum}(g) given by the standard Kähler form ω0\omega_{0} on 2\mathbb{C}^{2}. Then under this special marking (ϕg)(ωg)(\phi_{g})^{*}(\omega_{g}) is constant on M0M_{0}.

We define the Kummer lattice KK to be the smallest primitive sublattice of H2(X,)H^{2}(X,\mathbb{Z}) that contains Ci,1i16C_{i},1\leq i\leq 16.

The map from the Teichmüller space of complex-2-tori to Per(M0)\mathrm{Per}(M_{0}) given by 2/gϕg(H2,0(Kum(g))\mathbb{C}^{2}/g\to\phi_{g}^{*}(H^{2,0}(\mathrm{Kum}(g)) has image P(Kum)={αPer(M0):αCi=0,1i16}\mathrm{P(Kum)}=\{\alpha\in\mathrm{Per}(M_{0}):\alpha\cdot C_{i}=0,1\leq i\leq 16\} by the following proposition and the Torelli thoerem.

Proposition 5.7 ([Nik75], see also Proposition 14.3.17 of [Huy16] ).

A K3 surface is Kummer if and only if there exists a primitive embedding of the Kummer lattice KK into NS(X,)={cH2(X,):cΩ=0}NS(X,\mathbb{Z})=\{c\in H^{2}(X,\mathbb{Z}):c\cdot\Omega=0\}.

So we see that Kummer surfaces with a certain marking make up a 5 dimensional closed subspace of the period domain. We call it the specially marked Kummer locus and denote it by P(Kum)\mathrm{P(Kum)}. However, considering the different markings, or equivalently the action of the mapping class group, the orbit of this closed subspace is in fact dense in the period domain.

Proposition 5.8 (Density theorem).

Kummer surfaces are dense in the period domain.

Sketch of proof, See [Huy16], [Kob90] for more detail.

By Proposition 5.7 a K3 surface of maximal Picard rank is a Kummer surface if T(X)=NS(X,)T(X)=NS(X,\mathbb{R})^{\perp} is a positive definite even oriented rank 2 lattice in H2(X,)H^{2}(X,\mathbb{Z}). This happens if the 2-plane given by the period [ΩX][\Omega_{X}] of XX is a rationally defined 2-plane in H2(X,)H^{2}(X,\mathbb{R}) such that xx4x\cdot x\in 4\mathbb{Z} if xspan(Ω,Ω)H2(X,)x\in span(\Re\Omega,\Im\Omega)\cap H^{2}(X,\mathbb{Z}). Such 2-planes are dense in the Grassmannian of positive 2-planes in H2(X,)H^{2}(X,\mathbb{R}). ∎

If we add the Kähler from into consideration, the flat metrics on tori maps to degenerate Kähler forms on the K3 surface, in fact, they give exactly the degenerate Kähler forms that have zero volume on the 16 exceptional curves.

Proposition 5.9.

The image of (ϕg,Kum(g),ϕg(ω0))(\phi_{g},\mathrm{Kum}(g),\phi_{g}^{*}(\omega_{0})) in KPer\mathrm{KPer} is

K(Kum)={(α,ω)KPer:αCi=0,ωCi=0,1i16}KPer.\mathrm{K(Kum)}=\{(\alpha,\omega)\in\mathrm{KPer}:\alpha\cdot C_{i}=0,\omega\cdot C_{i}=0,1\leq i\leq 16\}\subset\mathrm{KPer}.

We call this the degenerate flat Kummer locus. In particular, the image of the Kähler moduli space of complex torus into generalized Kähler moduli spcae of K3 surfaces is Γ\Γ(K(Kum))\Gamma\backslash\Gamma(\mathrm{K(Kum)}).

Proof.

Fix an ΩP(Kum)\Omega\in\mathrm{P(Kum)} and the complex K3 surface X=Kum(g)X=\mathrm{Kum}(g) with period Ω\Omega, we only need to prove that every Kähler class in the Kähler cone that is orthogonal to CiC_{i} can be realized as the image of the standard Kähler form ω0\omega_{0} on 2\mathbb{C}^{2} for some hgGL(n,)h\in g\mathrm{GL}(n,\mathbb{C}). Firstly, the image (Ω,ϕg(ω0))(\Omega,\phi_{g}^{*}(\omega_{0})) is in K(Kum)\mathrm{K(Kum)} since the degenerate Kähler form has zero volume on the 16 exceptional curves. And we note that the fiber of K(Kum)\mathrm{K(Kum)} over Ω\Omega, denoted Kah(Ω)\mathrm{Kah}(\Omega) is one component (determined by orientation) of {ωH2(X,):ωCi=ωΩ=0,ωω>0}\{\omega\in H^{2}(X,\mathbb{R}):\omega\cdot C_{i}=\omega\cdot\Omega=0,\omega\cdot\omega>0\} which is a connected open subset of a 44 dimensional subspace in H2(X,)H^{2}(X,\mathbb{R}). The right multiplication of GL(n,)\mathrm{GL}(n,\mathbb{C}) on GL(2n,)\mathrm{GL}(2n,\mathbb{R}) preserves the complex structure. Hence we fix the marking ϕg\phi_{g} to descend this action to Kah(Ω)\mathrm{Kah}(\Omega). The stablizer of ωg\omega_{g} is a conjugate of U(n)\mathrm{U}(n). So we have a differentiable map GL(n,)/U(n)Kah(Ω):hωgh\mathrm{GL}(n,\mathbb{C})/\mathrm{U}(n)\to\mathrm{Kah}(\Omega):h\to\omega_{gh} which is locally injective. Thus the image must be the whole set. ∎

6. Special Lagrangian Foliation and Geodesic Flow for Kummer Surfaces

We define the linear foliations and the geodesic flow for Kummer surfaces to be the image of the linear foliations and the geodesic flow on the tori generating the Kummer surfaces.

The horizontal and vertical foliations of a Kummer surface X=Kum(g)X=\mathrm{Kum}(g) are defined for a specified gGL(4,)g\in\mathrm{GL}(4,\mathbb{R}) as the image of the horizontal and vertical foliation of the torus X=2/Λ(g)X=\mathbb{C}^{2}/\Lambda(g) under the map XX/{±1}X\to X/\{\pm 1\}. These are singular special Lagrangian foliations on the corresponding exceptional Kummer surfaces and they are defined outside the exceptional curves on the Kummer surfaces.

Proposition 6.1.

Suppose (X,ω)K(Kum)(X,\omega)\in\mathrm{K(Kum)} is a Kummer surface given by X=Kum(g)X=\mathrm{Kum}(g) with degenerate Kähler form ω\omega, then the horizontal foliations of Kum(gh),ω,hU(2)\mathrm{Kum}(gh),\omega,h\in U(2) are in one to one correspondence to linear special Lagrangian foliations of (X,ω)(X,\omega) outside the 16 exceptional curves.

Proof.

This is done by direct calculation. For a linear subspace V={y1=ax1+bx2,y2=cx1+dx2}V=\{y_{1}=ax_{1}+bx_{2},y_{2}=cx_{1}+dx_{2}\}, ω0\omega_{0} and (dz1dz2)\Im(dz_{1}\wedge dz_{2}) to vanish on VV if and only if a=d,b=ca=-d,b=c, in which case (dz1dz2)V>0\Re(dz_{1}\wedge dz_{2})\mid V>0 for a suitable orientation. On the other hand the leaves of possible horizontal foliations on XX are the images {(x1,c1,x2,c2):x1,x2}h\{(x_{1},c_{1},x_{2},c_{2}):x_{1},x_{2}\in\mathbb{R}\}h, where hU(n)h\in\mathrm{U}(n), for det((hij)i,j=1,3)0\det((h^{ij})_{i,j=1,3})\neq 0, we may write this as a subspace where (abcd)\begin{pmatrix}a&b\\ c&d\end{pmatrix} is given by h(h)1\Im h(\Re h)^{-1} so we have a=d,b=ca=-d,b=c. Furthermore, any (abba)\begin{pmatrix}a&b\\ b&-a\end{pmatrix} can always be written as h(h)1\Im h(\Re h)^{-1} for some hU(n)h\in\mathrm{U}(n).

For VV that cannot be written in the form {y1=ax1+bx2,y2=cx1+dx2}\{y_{1}=ax_{1}+bx_{2},y_{2}=cx_{1}+dx_{2}\}, we simply observe that we may apply a unitary transformation to bring it to this form. ∎

A geodesic flow of Kummer surfaces starting at Kum(g)\mathrm{Kum}(g) (with direction given by gg) is defined as Rat(Kum(g),ωg)=(Kum(gat),ωgat)Ra_{t}(\mathrm{Kum}(g),\omega_{g})=(\mathrm{Kum}(ga_{t}),\omega_{ga_{t}}), i.e. bringing the Kummer surface generated by gGL(4,)g\in\mathrm{GL}(4,\mathbb{R}) to the one generated by gatga_{t}. This is obtained by projecting the geodesic flow in the moduli space of complex torus down to the Kummer locus K(Kum)\mathrm{K(Kum)}.

We may apply our results on complex tori to Kummer surfaces to get similar connection between geodesic flow and ergodicity of foliations for Kummer surfaces.

Proof of Theorem 1.3 for Kummer surfaces.

If the vertical or horizontal foliations on a torus is uniquely ergodic, then its image on the exceptional Kummer surface must be uniquely ergodic. If the geodesic flow in generalized Kähler moduli space of K3 surfaces is recurrent in compact subsets, then the pre-image in the Kähler moduli space of tori must also be recurrent. Thus by Theorem 1.2, assume the image of {RatX:t0}\{Ra_{t}X:t\leq 0\} in Kah¯(M0)\overline{\mathrm{Kah}}(M_{0}) is recurrent. Then the horizontal foliation must be uniquely ergodic. ∎

By the Torelli theorem and the period map, we may explicitly calculate the form of the geodesic flow.

Proposition 6.2.

The geodesic flow RatXRa_{t}X of a Kummer surface XX in the period domain of degenerate Kähler forms KPer\mathrm{KPer} is of the form (e2tη1+e2tη2+1η3,ω)(e^{-2t}\eta_{1}+e^{2t}\eta_{2}+\sqrt{-1}\eta_{3},\omega) where ηi,ωH2(X,)\eta_{i},\omega\in H^{2}(X,\mathbb{R}) satisfy ηiH2(X,),ηiCj=0,ωCj=0,1i3,1j16\eta_{i}\in H^{2}(X,\mathbb{R}),\eta_{i}\cdot C_{j}=0,\omega\cdot C_{j}=0,1\leq i\leq 3,1\leq j\leq 16 and (ηi)2=0,ηiη3=0(\eta_{i})^{2}=0,\eta_{i}\cdot\eta_{3}=0 for i=1,2i=1,2 and 2η1η2=(η3)2>02\eta_{1}\cdot\eta_{2}=(\eta_{3})^{2}>0, ωηi=0,i=1,2,3\omega\cdot\eta_{i}=0,i=1,2,3.

Thus in the Kähler moduli space Kah¯(M0)\mathrm{\overline{Kah}}(M_{0}), the geodesic flows are of the form [e2tη1+e2tη2+1η3,ω]=Γ(span(e2tη1+e2tη2,η3,ω))[e^{-2t}\eta_{1}+e^{2t}\eta_{2}+\sqrt{-1}\eta_{3},\omega]=\Gamma(span(e^{-2t}\eta_{1}+e^{2t}\eta_{2},\eta_{3},\omega))Kah¯(M0)Γ\SO(3,19,)/(SO(2)×SO(19))\in\mathrm{\overline{Kah}}(M_{0})\simeq\Gamma\backslash SO(3,19,\mathbb{R})/(SO(2)\times SO(19)).

Proof.

Suppose X=Kum(g)X=\mathrm{Kum}(g), first we calculate the periods of RatXRa_{t}X using the markings ϕgat\phi_{ga_{t}} and the basis given as follows.

Suppose g=(λ1λ2λ3λ4)=(λ11λ21λ12λ22λ13λ23λ14λ24)g=\begin{pmatrix}\lambda^{1}\\ \lambda^{2}\\ \lambda^{3}\\ \lambda^{4}\end{pmatrix}=\begin{pmatrix}\lambda^{1}_{1}&\lambda^{1}_{2}\\ \lambda^{2}_{1}&\lambda^{2}_{2}\\ \lambda_{1}^{3}&\lambda_{2}^{3}\\ \lambda_{1}^{4}&\lambda_{2}^{4}\end{pmatrix} where λji\lambda^{i}_{j}\in\mathbb{C}. Take TijT_{ij} to be the torus in the Kummer surface given by the image of the torus in 2/Λ(g)\mathbb{C}^{2}/\Lambda(g) spanned by λi\lambda^{i} and λj\lambda^{j}. Then Tij,CkT_{ij},C_{k} makes up a basis for H2(X,)H_{2}(X,\mathbb{Q}) and TijT_{ij} is a basis for ({Cj})H2(X,)(\{C_{j}\})^{\perp}\subset H_{2}(X,\mathbb{Z}).

Under this basis and the corresponding marking, we have Per(Kum(g))=[Tij𝑑z1dz2,0,,016 zeros ]\mathrm{Per}(\mathrm{Kum}(g))=[\int_{T_{ij}}dz^{1}\wedge dz^{2},\overbrace{0,...,0}^{16\text{ zeros }}]. So by direct calculation, we have

Per(Rat(X))=Per(Kum(etλ11etλ11etλ21etλ21etλ12etλ12etλ22etλ22etλ13etλ13etλ23etλ23etλ14etλ14etλ24etλ24))\mathrm{Per}(Ra_{t}(X))=\mathrm{Per}(\mathrm{Kum}\begin{pmatrix}e^{-t}\Re\lambda^{1}_{1}&e^{t}\Im\lambda^{1}_{1}&e^{-t}\Re\lambda^{1}_{2}&e^{t}\Im\lambda^{1}_{2}\\ e^{-t}\Re\lambda^{2}_{1}&e^{t}\Im\lambda^{2}_{1}&e^{-t}\Re\lambda^{2}_{2}&e^{t}\Im\lambda^{2}_{2}\\ e^{-t}\Re\lambda_{1}^{3}&e^{t}\Im\lambda_{1}^{3}&e^{-t}\Re\lambda_{2}^{3}&e^{t}\Im\lambda_{2}^{3}\\ e^{-t}\Re\lambda_{1}^{4}&e^{t}\Im\lambda_{1}^{4}&e^{-t}\Re\lambda_{2}^{4}&e^{t}\Im\lambda_{2}^{4}\end{pmatrix})
=[(etλ1i+1etλ1i)(etλ2j+1etλ2j)=[(e^{-t}\Re\lambda^{i}_{1}+\sqrt{-1}e^{t}\Im\lambda^{i}_{1})(e^{-t}\Re\lambda^{j}_{2}+\sqrt{-1}e^{t}\Im\lambda^{j}_{2})
(etλ2i+1etλ2i)(etλ1j+1etλ1j),0,,0]-(e^{-t}\Re\lambda^{i}_{2}+\sqrt{-1}e^{t}\Im\lambda^{i}_{2})(e^{-t}\Re\lambda^{j}_{1}+\sqrt{-1}e^{t}\Im\lambda^{j}_{1}),0,...,0]
=[e2tη1ij+e2tη2ij+1η3ij,0,,0]=[e^{-2t}\eta_{1}^{ij}+e^{2t}\eta_{2}^{ij}+\sqrt{-1}\eta_{3}^{ij},0,...,0]

where

η1ij=λ1iλ2jλ2iλ1j,\eta_{1}^{ij}=\Re\lambda^{i}_{1}\Re\lambda^{j}_{2}-\Re\lambda^{i}_{2}\Re\lambda^{j}_{1},
η2ij=λ1iλ2jλ2iλ1j,\eta_{2}^{ij}=\Im\lambda^{i}_{1}\Im\lambda^{j}_{2}-\Im\lambda^{i}_{2}\Im\lambda^{j}_{1},
η3ij=λ1iλ2jλ2iλ1j+λ1iλ2jλ2iλ1j.\eta_{3}^{ij}=\Re\lambda^{i}_{1}\Im\lambda^{j}_{2}-\Re\lambda^{i}_{2}\Im\lambda^{j}_{1}+\Im\lambda^{i}_{1}\Re\lambda^{j}_{2}-\Im\lambda^{i}_{2}\Re\lambda^{j}_{1}.

For the Kähler form ωg\omega_{g}, we have under these coordinates Per(ϕgat(ωg))=[Tij𝑑x1dy1+dx2dy2,0,,016 zeroes]\mathrm{Per}(\phi^{*}_{ga_{t}}(\omega_{g}))=[\int_{T_{ij}}dx_{1}\wedge dy_{1}+dx_{2}\wedge dy_{2},\overbrace{0,...,0}^{16\text{ zeroes}}] where Tij𝑑x1dy1+dx2dy2\int_{T_{ij}}dx_{1}\wedge dy_{1}+dx_{2}\wedge dy_{2} on Kum(gat)\mathrm{Kum}(ga_{t}) is Tij𝑑x1dy1+dx2dy2=etλ1ietλ1jetλ1jetλ1i+etλ2ietλ2jetλ2jetλ2i=λ1iλ1jλ1jλ1i+λ2iλ2jλ2jλ2i\int_{T_{ij}}dx_{1}\wedge dy_{1}+dx_{2}\wedge dy_{2}=e^{-t}\Re\lambda^{i}_{1}e^{t}\Im\lambda^{j}_{1}-e^{-t}\Re\lambda^{j}_{1}e^{t}\Im\lambda^{i}_{1}+e^{-t}\Re\lambda^{i}_{2}e^{t}\Im\lambda^{j}_{2}-e^{-t}\Re\lambda^{j}_{2}e^{t}\Im\lambda^{i}_{2}=\Re\lambda^{i}_{1}\Im\lambda^{j}_{1}-\Re\lambda^{j}_{1}\Im\lambda^{i}_{1}+\Re\lambda^{i}_{2}\Im\lambda^{j}_{2}-\Re\lambda^{j}_{2}\Im\lambda^{i}_{2} which is invariant under RatRa_{t}. Denote this 2-form as ω\omega (which is equal to [ωg][\omega_{g}]).

So take ηk=ηkijTij\eta_{k}=\sum\eta_{k}^{ij}T_{ij}^{*}, we have Per(Rat(X))=e2tη1+e2tη2+1η3\mathrm{Per}(Ra_{t}(X))=e^{-2t}\eta_{1}+e^{2t}\eta_{2}+\sqrt{-1}\eta_{3}.

Next, we check the conditions they satisfy. We may check that (ηi)2=0,ηiη3=0(\eta_{i})^{2}=0,\eta_{i}\cdot\eta_{3}=0 for i=1,2i=1,2 and 2η1η2=(η3)2>02\eta_{1}\cdot\eta_{2}=(\eta_{3})^{2}>0, ωηi=0,i=1,2,3\omega\cdot\eta_{i}=0,i=1,2,3 either by direct calculation, or by the fact that (e2tη1+e2tη2+1η3,ω)K(Kum)(e^{2t}\eta_{1}+e^{-2t}\eta_{2}+\sqrt{-1}\eta_{3},\omega)\in\mathrm{K(Kum)} for any tt.

Descending from the case of tori, for Kummer surfaces, the geodesic flow we defined acts on exceptional Kummer surfaces exactly by expanding the horizontal foliation and contracting the vertical foliations. Although we defined the geodesic flow as given by a “direction” from gSL(2n,)g\in\mathrm{SL}(2n,\mathbb{R}), we see in the calculation that the geodesic flow itself only depends on the directions of v\mathcal{F}^{v} and h\mathcal{F}^{h}, and we may view η1\eta_{1} and η2\eta_{2} as invariants representing those foliations.

We may also apply Corollary 4.2 to get a necessary condition for horizontal foliations to be uniquely ergodic: it is uniquely ergodic only if rank(η2H2(X,))<19rank(\eta_{2}^{\perp}\cap H^{2}(X,\mathbb{Z}))<19.

Proof of Theorem 1.3.

The horizontal foliation h\mathcal{F}^{h} in the Kummer surface XX outside the exceptional curves is unique ergodic if and only if its pre-image in the torus Y=2/Λ(g)Y=\mathbb{C}^{2}/\Lambda(g) is uniquely ergodic, which happens if and only if its leaves are not contained in the real span of any sub-lattice of g=(λ1λ2λ3λ4)g=\begin{pmatrix}\lambda^{1}\\ \lambda^{2}\\ \lambda^{3}\\ \lambda^{4}\end{pmatrix}. We next prove by contradiction. We show that if a leaf of the horizontal foliation is in the real span of a sub-lattice of Λ\Lambda, then rank(η2H2(X,))19.rank(\eta_{2}^{\perp}\cap H^{2}(X,\mathbb{Z}))\geq 19.

Assume that a leaf of the horizontal foliation is in the real span of a sub-lattice of Λ\Lambda.

We first assume that it is in the real span of λ1,λ2,λ3\lambda^{1},\lambda^{2},\lambda^{3}. We claim that this happens if and only if λ1,λ2,λ3\Im\lambda^{1},\Im\lambda^{2},\Im\lambda^{3} are collinear over \mathbb{R}. Indeed, if λ1,λ2,λ3\Im\lambda^{1},\Im\lambda^{2},\Im\lambda^{3} are collinear, then λ1,λ2,λ3\Re\lambda^{1},\Re\lambda^{2},\Re\lambda^{3} cannot be collinear, so the 2-plane {y1=y2=0}\{y_{1}=y_{2}=0\} is in the real span of λ1,λ2\lambda^{1},\lambda^{2} and λ3\lambda^{3}. Conversely, if λ1,λ2,λ3\Im\lambda^{1},\Im\lambda^{2},\Im\lambda^{3} are not collinear, then span(λ1,λ2,λ3){y1=y2=0}span(\lambda^{1},\lambda^{2},\lambda^{3})\cap\{y_{1}=y_{2}=0\} must be 1-dimensional, which contradicts to a leaf of the horizontal foliation being in the real span span(λ1,λ2,λ3)span(\lambda^{1},\lambda^{2},\lambda^{3}).

Now if we only have that a leaf of the horizontal foliation is in the real span of a sub-lattice of Λ\Lambda, then we have λ1,λ2,λ3\Im\lambda^{1},\Im\lambda^{2},\Im\lambda^{3} are mutually linear after a left multiplication of SL(4,)\mathrm{SL}(4,\mathbb{Z}) on Λ\Lambda. So using the basis Tik,1i<k4,Cj,1j16T_{ik},1\leq i<k\leq 4,C_{j},1\leq j\leq 16 for H2(X,)H^{2}(X,\mathbb{Q}), η2\eta_{2} has 19 zero entries after a left multiplication of SO(3,3,)×I16SO(3,3,\mathbb{Z})\times I_{16} on η2\eta_{2}. Thus we have rank(η2H2(X,))19rank(\eta_{2}^{\perp}\cap H^{2}(X,\mathbb{Z}))\geq 19. ∎

We remark that the converse is not true. For example consider

g=(1100120003100011) and h=(1000110000110312).g=\begin{pmatrix}1&1&0&0\\ 1&\sqrt{2}&0&0\\ 0&-\sqrt{3}&1&0\\ 0&0&1&1\end{pmatrix}\text{ and }h=\begin{pmatrix}1&0&0&0\\ 1&1&0&0\\ 0&0&1&1\\ 0&\sqrt{3}&1&\sqrt{2}\end{pmatrix}.

Then the horizontal foliations on Kum(g)\mathrm{Kum}(g) is uniquely ergodic while the horizontal foliations on Kum(h)\mathrm{Kum}(h) is not. However, we may calculate that η2(g)\eta_{2}(g) and η2(h)\eta_{2}(h) differ only by an element in O(H2(M0,))O(H^{2}(M_{0},\mathbb{Z})), so they are the same elements on the boundary of the complex moduli space.

References

  • [Can14] Serge Cantat. Dynamics of automorphisms of compact complex surfaces. In Frontiers in complex dynamics, volume 51 of Princeton Math. Ser., pages 463–514. Princeton Univ. Press, Princeton, NJ, 2014.
  • [CD20] Serge Cantat and Christophe Dupont. Automorphisms of surfaces: Kummer rigidity and measure of maximal entropy. J. Eur. Math. Soc. (JEMS), 22(4):1289–1351, 2020.
  • [Fil20] Simion Filip. Counting special Lagrangian fibrations in twistor families of K3 surfaces. Ann. Sci. Éc. Norm. Supér. (4), 53(3):713–750, 2020. With an appendix by Nicolas Bergeron and Carlos Matheus.
  • [FT19] Simion Filip and Valentino Tosatti. Kummer rigidity for k3 surface automorphisms via ricci-flat metrics, 2019.
  • [GHJ03] M. Gross, D. Huybrechts, and D. Joyce. Calabi-Yau manifolds and related geometries. Universitext. Springer-Verlag, Berlin, 2003. Lectures from the Summer School held in Nordfjordeid, June 2001.
  • [Huy16] Daniel Huybrechts. Lectures on K3 surfaces, volume 158 of Cambridge Studies in Advanced Mathematics. Cambridge University Press, Cambridge, 2016.
  • [Kob90] Ryoichi Kobayashi. Moduli of Einstein metrics on a K3K3 surface and degeneration of type I{\rm I}. In Kähler metric and moduli spaces, volume 18 of Adv. Stud. Pure Math., pages 257–311. Academic Press, Boston, MA, 1990.
  • [Mas92] Howard Masur. Hausdorff dimension of the set of nonergodic foliations of a quadratic differential. Duke Math. J., 66(3):387–442, 1992.
  • [McM02] Curtis T. McMullen. Dynamics on K3K3 surfaces: Salem numbers and Siegel disks. J. Reine Angew. Math., 545:201–233, 2002.
  • [Moo66] Calvin C. Moore. Ergodicity of flows on homogeneous spaces. Amer. J. Math., 88:154–178, 1966.
  • [Nik75] V. V. Nikulin. Kummer surfaces. Izv. Akad. Nauk SSSR Ser. Mat., 39(2):278–293, 471, 1975.
  • [Ver15] Misha Verbitsky. Ergodic complex structures on hyperkähler manifolds. Acta Math., 215(1):161–182, 2015.
  • [Wri15] Alex Wright. Translation surfaces and their orbit closures: an introduction for a broad audience. EMS Surv. Math. Sci., 2(1):63–108, 2015.
  • [Yau77] Shing Tung Yau. Calabi’s conjecture and some new results in algebraic geometry. Proc. Nat. Acad. Sci. U.S.A., 74(5):1798–1799, 1977.