Six-Dimensional Solvmanifolds with Holomorphically Trivial Canonical Bundle

Similar documents
REVISTA DE LA REAL ACADEMIA DE CIENCIAS. Exactas Físicas Químicas y Naturales DE ZARAGOZA. Serie 2.ª Volumen 69

arxiv: v2 [math.dg] 23 Dec 2016

joint work with A. Fino, M. Macrì, G. Ovando, M. Subils Rosario (Argentina) July 2012 Sergio Console Dipartimento di Matematica Università di Torino

Dolbeault cohomology and. stability of abelian complex structures. on nilmanifolds

LECTURE 2: SYMPLECTIC VECTOR BUNDLES

Geometria Simplettica e metriche Hermitiane speciali

Flat complex connections with torsion of type (1, 1)

MATRIX LIE GROUPS AND LIE GROUPS

Multi-moment maps. CP 3 Journal Club. Thomas Bruun Madsen. 20th November 2009

Siegel Moduli Space of Principally Polarized Abelian Manifolds

arxiv: v2 [math.dg] 16 Sep 2015

B 1 = {B(x, r) x = (x 1, x 2 ) H, 0 < r < x 2 }. (a) Show that B = B 1 B 2 is a basis for a topology on X.

Hyperkähler geometry lecture 3

arxiv: v2 [math.dg] 10 May 2015

Infinitesimal Einstein Deformations. Kähler Manifolds

Hard Lefschetz Theorem for Vaisman manifolds

LIE ALGEBRAS: LECTURE 7 11 May 2010

ON k-abelian p-filiform LIE ALGEBRAS I. 1. Generalities

INSTANTON MODULI AND COMPACTIFICATION MATTHEW MAHOWALD

Homogeneous para-kähler Einstein manifolds. Dmitri V. Alekseevsky

NilBott Tower of Aspherical Manifolds and Torus Actions

arxiv:math/ v2 [math.dg] 24 Nov 2007 ANNA FINO AND ADRIANO TOMASSINI

Math 868 Final Exam. Part 1. Complete 5 of the following 7 sentences to make a precise definition (5 points each). Y (φ t ) Y lim

Math 121 Homework 4: Notes on Selected Problems

Formality of Kähler manifolds

A Joint Adventure in Sasakian and Kähler Geometry

LIE ALGEBRA PREDERIVATIONS AND STRONGLY NILPOTENT LIE ALGEBRAS

INDECOMPOSABLE LIE ALGEBRAS WITH NONTRIVIAL LEVI DECOMPOSITION CANNOT HAVE FILIFORM RADICAL

Some remarks on symmetric correspondences

Quantizations and classical non-commutative non-associative algebras

Generalized complex geometry and topological sigma-models

Spin(10,1)-metrics with a parallel null spinor and maximal holonomy

DIFFERENTIAL FORMS AND COHOMOLOGY

Metrics and Holonomy

A MARSDEN WEINSTEIN REDUCTION THEOREM FOR PRESYMPLECTIC MANIFOLDS

Geometry of moduli spaces

THE THEOREM OF THE HIGHEST WEIGHT

Introduction to Group Theory

ALGEBRA QUALIFYING EXAM PROBLEMS LINEAR ALGEBRA

Citation Osaka Journal of Mathematics. 43(1)

L(C G (x) 0 ) c g (x). Proof. Recall C G (x) = {g G xgx 1 = g} and c g (x) = {X g Ad xx = X}. In general, it is obvious that

CONTROLLABILITY OF NILPOTENT SYSTEMS

CALIBRATED FIBRATIONS ON NONCOMPACT MANIFOLDS VIA GROUP ACTIONS

Formal power series rings, inverse limits, and I-adic completions of rings

CONSIDERATION OF COMPACT MINIMAL SURFACES IN 4-DIMENSIONAL FLAT TORI IN TERMS OF DEGENERATE GAUSS MAP

Math 249B. Nilpotence of connected solvable groups

ON NEARLY SEMIFREE CIRCLE ACTIONS

arxiv: v1 [math.dg] 30 Aug 2018

A PROOF OF BOREL-WEIL-BOTT THEOREM

SYMMETRIES OF SECTIONAL CURVATURE ON 3 MANIFOLDS. May 27, 1992

Solvable Lie groups and the shear construction

HYPERKÄHLER MANIFOLDS

TRANSITIVE HOLONOMY GROUP AND RIGIDITY IN NONNEGATIVE CURVATURE. Luis Guijarro and Gerard Walschap

Stable bundles with small c 2 over 2-dimensional complex tori

Classification of solvable Lie algebras - new approaches and developments

Math 210C. The representation ring

Groups of Prime Power Order with Derived Subgroup of Prime Order

INTRODUCTION TO LIE ALGEBRAS. LECTURE 7.

Abstract Algebra II Groups ( )

Kac-Moody Algebras. Ana Ros Camacho June 28, 2010

Math 210C. A non-closed commutator subgroup

LECTURE 3: TENSORING WITH FINITE DIMENSIONAL MODULES IN CATEGORY O

Chern forms and the Fredholm determinant

ON COSTELLO S CONSTRUCTION OF THE WITTEN GENUS: L SPACES AND DG-MANIFOLDS

Representations of semisimple Lie algebras

THE GROUP OF UNITS OF SOME FINITE LOCAL RINGS I

Techniques of computations of Dolbeault cohomology of solvmanifolds

Riemann surfaces with extra automorphisms and endomorphism rings of their Jacobians

An introduction to Birkhoff normal form

Mirror Symmetry: Introduction to the B Model

Math 121 Homework 5: Notes on Selected Problems

Seminario de Geometría y Topología Facultad de Ciencias Matemáticas, Universidad Complutense de Madrid

Intertwining integrals on completely solvable Lie groups

Algebraic v.s. Analytic Point of View

Trace Class Operators and Lidskii s Theorem

Mathematical Research Letters 2, (1995) A VANISHING THEOREM FOR SEIBERG-WITTEN INVARIANTS. Shuguang Wang

FORMAL GROUPS OF CERTAIN Q-CURVES OVER QUADRATIC FIELDS

These notes are incomplete they will be updated regularly.

LECTURE 25-26: CARTAN S THEOREM OF MAXIMAL TORI. 1. Maximal Tori

Real Analysis Prelim Questions Day 1 August 27, 2013

Two-Step Nilpotent Lie Algebras Attached to Graphs

GROUP THEORY PRIMER. New terms: so(2n), so(2n+1), symplectic algebra sp(2n)

Honors Algebra 4, MATH 371 Winter 2010 Assignment 4 Due Wednesday, February 17 at 08:35

Holomorphic line bundles

Decay to zero of matrix coefficients at Adjoint infinity by Scot Adams

1 Hermitian symmetric spaces: examples and basic properties

THE HODGE DECOMPOSITION

RIEMANN SURFACES: TALK V: DOLBEAULT COHOMOLOGY

SYMPLECTIC GEOMETRY: LECTURE 5

Classification of discretely decomposable A q (λ) with respect to reductive symmetric pairs UNIVERSITY OF TOKYO

Atiyah classes and homotopy algebras

Surjectivity in Honda-Tate

TOTALLY REAL SURFACES IN THE COMPLEX 2-SPACE

Algebraic Number Theory

MILNOR SEMINAR: DIFFERENTIAL FORMS AND CHERN CLASSES

REPRESENTATION THEORY, LECTURE 0. BASICS

THE EULER CHARACTERISTIC OF A LIE GROUP

Lie algebra cohomology

Cohomology jump loci of local systems

Lecture III: Neighbourhoods

Transcription:

International Mathematics Research Notices Advance Access published May 7, 2015 A. Fino et al. (2015) Six-Dimensional Solvmanifolds with Holomorphically Trivial Canonical Bundle, International Mathematics Research Notices, rnv112, 43 pages. doi:10.1093/imrn/rnv112 Six-Dimensional Solvmanifolds with Holomorphically Trivial Canonical Bundle Anna Fino 1, Antonio Otal 2, and Luis Ugarte 3 1 Dipartimento di Matematica, Università di Torino, Via Carlo Alberto 10, 10123 Torino, Italy, 2 Centro Universitario de la Defensa-I.U.M.A., Academia General Militar, Ctra. de Huesca s/n. 50090 Zaragoza, Spain, and 3 Departamento de Matemáticas-I.U.M.A., Universidad de Zaragoza, Campus Plaza San Francisco, 50009 Zaragoza, Spain In memory of Sergio Console Correspondence to be sent to: e-mail: annamaria.fino@unito.it We determine the 6D solvmanifolds admitting an invariant complex structure with holomorphically trivial canonical bundle. Such complex structures are classified up to isomorphism, and the existence of strong Kähler with torsion, generalized Gauduchon, balanced and strongly Gauduchon metrics is studied. As an application, we construct a holomorphic family (M, J a ) of compact complex manifolds such that (M, J a ) satisfies the -lemma and admits a balanced metric for any a =0, but the central limit neither satisfies the -lemma nor admits balanced metrics. 1 Introduction Any compact complex surface with holomorphically trivial canonical bundle is isomorphic to a K3 surface, a torus, or a Kodaira surface; the first two are Kähler, and the latter is a nilmanifold, that is, a compact quotient of a nilpotent Lie group by a lattice. It is well known that in any real dimension 2n the canonical bundle of a nilmanifold Γ \G endowed with an invariant complex structure is holomorphically trivial, where by Received February 7, 2014; Revised February 4, 2015; Accepted March 30, 2015 c The Author(s) 2015. Published by Oxford University Press. All rights reserved. For permissions, please e-mail: journals.permissions@oup.com.

2 A. Fino et al. invariant complex structure we mean one induced by a complex structure on the Lie algebra of the nilpotent Lie group G. In fact, Salamon proved in [27] the existence of a closed nonzero invariant (n, 0)-form (see also [5] for some applications of this fact to hypercomplex nilmanifolds). For 6D nilmanifolds, the classification of invariant complex structures J as well as the existence of some special Hermitian metrics (as strong Kähler with torsion (SKT), generalized first Gauduchon, balanced and strongly Gauduchon metrics) with respect to such Js have been studied in [7, 12, 13]. In this paper, we are interested in the Hermitian geometry of 6D solvmanifolds admitting an invariant complex structure with holomorphically trivial canonical bundle, that is, with an invariant nonzero closed (3,0)-form (see Proposition 2.1 for details). Throughout the paper, by a solvmanifold, we mean a compact quotient Γ \G of a non-nilpotent solvable Lie group G by a lattice Γ. Since the complex structures we consider are invariant, they come from complex structures on the Lie algebra g of G. Given an almost complex structure J : g g on a 6D Lie algebra g, the existence of a nonzero closed form of bidegree (3,0) implies that J is integrable, that is, that J is a complex structure. If the Lie algebra is nilpotent, then both conditions are equivalent [27]. The 6D nilpotent Lie algebras admitting a complex structure were classified in [27] and recently, these complex structures have been classified up to equivalence [7]. For solvable Lie algebras admitting a complex structure, there exists a general classification only in real dimension four [23], but no general result is known in higher dimension. In this paper, we consider the bigger class of 6D solvable Lie algebras g and look for almost complex structures admitting a nonzero closed (3,0)-form. Note that the latter condition implies that b 3 (g) 2. We will consider both the indecomposable and the decomposable solvable Lie algebras which are unimodular as we aim to find invariant complex structures with nonzero closed (3,0)-form on solvmanifolds [20]. We recall that the 6D indecomposable solvable Lie algebras have been classified by Turkowski [31], Mubarakzyanov [21], and Shabanskaya, who recovered and corrected in [29] the work of Mubarakzyanov (see the Appendix for more details). In Section 2, we consider the formalism of stable forms [17], together with ideas in [9, 14, 15, 28], and we explain in detail the method that we follow to classify unimodular solvable Lie algebras admitting a complex structure with nonzero closed (3,0)- form. The classification is obtained in Propositions 2.6 and 2.7 for the decomposable and indecomposable cases, respectively. As a consequence, we have that if a solvmanifold admits a complex structure arising from an invariant nonvanishing holomorphic (3,0)-form, then its underlying Lie algebra g must be isomorphic to one in the list of

Six-Dimensional Solvmanifolds with Holomorphically Trivial Canonical Bundle 3 Theorem 2.8, that is, g = g 1, g α 2 (α 0), g 3,...,g 8 or g 9. The Lie algebras g 1, g α 2 (α 0) and g 3 are decomposable, whereas g 4,...,g 8, and g 9 are indecomposable. The Lie algebra g 8 is precisely the real Lie algebra underlying Nakamura manifold [22]. Moreover, using some results in [6], we ensure in Proposition 2.10 the existence of a lattice for the simply connected Lie groups associated with the Lie algebras in the list, although for g α 2 we are able to find a lattice only for a countable number of different values of α (note that one cannot expect a lattice to exist for any real α>0 according to [32, Proposition 8.7]). In Section 3, we consider the whole space of complex structures having closed (3,0)-form and classify them up to equivalence (see Theorem 3.10). It turns out that the classification is finite, except for g 3 (Proposition 3.4) and g 8 (Proposition 3.7). As a consequence of this complex classification, we study in Section 4 the existence of several special Hermitian metrics on the corresponding complex solvmanifolds. The SKT geometry is studied in Theorem 4.1 and provides a new example of compact SKT manifold based on the Lie algebra g 4. In Theorem 4.3, we investigate the existence of invariant first Gauduchon metrics, in the sense of [16], and find that a solvmanifold corresponding to g 6 has invariant first Gauduchon metrics which are not SKT, moreover, it does not admit any SKT metric (note that any invariant first Gauduchon metric on a 6-nilmanifold is necessarily SKT as proved in [13, Proposition 3.3]). Finally, Theorems 4.5 and 4.6 deal with the existence of balanced metrics [19] and strongly Gauduchon metrics [24 26], respectively, and provide new compact examples of such Hermitian geometries. Our goal in Section 5 is to construct a holomorphic family of compact complex manifolds (M, J a ) a Δ, Δ ={a C a < 1}, such that (M, J a ) satisfies the -lemma and admits balanced metric for any a =0, but (M, J 0 ) neither satisfies the -lemma nor admits balanced metric. The construction is based on the balanced Hermitian geometry studied in Theorem 4.5 for g 8, the Lie algebra underlying Nakamura manifold, together with a recent result by Angella and Kasuya [3] on deformations of the holomorphically parallelizable Nakamura manifold. Note that the central limit (M, J 0 ) admits strongly Gauduchon metric by Theorem 4.6, as it must happen according to Popovici s result [24, Proposition 4.1]. We recall that there exists [7] a holomorphic family (N, J a ), N being a 6D nilmanifold, such that (N, J a ) has balanced metrics for any a =0, but the central limit (N, J 0 ) does not admit strongly Gauduchon metric. 2 The Classification We first show that the existence of a holomorphic form of bidegree (n, 0) with respect to an invariant complex structure on a 2n-dimensional solvmanifold implies the existence of an invariant nonzero closed (n, 0)-form. Furthermore:

4 A. Fino et al. Proposition 2.1. Let M = Γ \G be a 2n-dimensional solvmanifold endowed with an invariant complex structure J. IfΩ is a nowhere vanishing holomorphic (n, 0)-form on (M, J), thenω is necessarily invariant. Proof. Since J is invariant, we consider a global basis {ω 1,...,ω n } of invariant (1, 0)- forms on (M, J). Then, there is a nowhere vanishing complex-valued function f : M C such that Ω = f ω 1 ω n.sinceω is holomorphic, we have Ω = f ω 1 ω n + f (ω 1 ω n ) = 0, that is, (ω 1 ω n ) = (log f) ω 1 ω n. The latter form is an invariant (n, 1)-form on (M, J), so there is an invariant form α of bidegree (0, 1) on (M, J) such that (log f) = α. (1) Now, we apply the well-known symmetrization process, which assigns an invariant k-form γ to any k-form γ on M by integration with respect to a volume element on M induced by a bi-invariant one on the Lie group G (its existence is guaranteed by [20]). By extending the symmetrization to the space of complex forms, since J is invariant, we have that symmetrization preserves the bidegree and commutes with, because it commutes with the differential d (see [7, 11] for more details). Applying this to (1), we get ( log f) = α = α, because α is invariant. But log f is the symmetrization of a function, so it is a constant and, then ( log f) = 0. Therefore, α = 0andby(1), we get (log f) = 0. This means that log f is a holomorphic function on a compact complex manifold, which implies that log f = c, where c is a constant. In conclusion, f = exp(c) is a constant function, and Ω is necessarily invariant. As a consequence of this result, in order to describe the solvmanifolds M = Γ \G admitting an invariant complex structure with holomorphically trivial canonical bundle, we are led to study the unimodular solvable Lie algebras g admitting a complex structure J with nonzero closed (n, 0)-form. Next, we classify such Lie algebras in dimension 6. We will recall first the formalism of stable forms which enable us to construct the space of almost complex structures on a given Lie algebra [9, 17]. Let (V,ν) be an oriented 6D vector space, a 3-form ρ is stable if its orbit under the action of the group GL(V) is open. Let κ : Λ 5 V V be the isomorphism defined by κ(η)= X, where X is such that ι X ν = η, andletk ρ : V V be the endomorphism given by K ρ (X) = κ(ι X ρ ρ).

Six-Dimensional Solvmanifolds with Holomorphically Trivial Canonical Bundle 5 The square of this endomorphism is proportional to the identity map so it is an automorphism of the vector space. Let λ(ρ) be the constant of proportionality, that is, λ(ρ) = 1 6 tr(k2 ρ ). Then, the stability of ρ is equivalent to the open condition λ(ρ) =0. When λ(ρ) < 0, the endomorphism J ρ := 1 λ(ρ) K ρ gives rise to an almost complex structure on V. Moreover, λ(ρ) enables to construct a specific volume form φ(ρ):= λ(ρ) ν Λ 6 V such that the action of the dual endomorphism Jρ on 1-forms is given by the formula for any α g and X g. As a consequence, there is a natural mapping ((J ρ α)(x))φ(ρ) = α ι Xρ ρ, (2) {ρ Λ 3 V λ(ρ) < 0} {J : V V J 2 = I } (3) assigning each ρ to J = J ρ through relation (2). Although this map is not injective, as for example J ρ = J ρ when ρ is proportional to ρ, it is onto and therefore, it covers the space of almost complex structures on V. We are interested in the complex structures on a Lie algebra g admitting nonzero closed (3,0)-form. Let Z 3 (g) ={ρ Λ 3 g dρ = 0}. Themap(3) restricts to the surjective mapping {ρ Z 3 (g) λ(ρ) < 0, d(j ρ ρ) = 0} {J : g g J2 = I, Ψ Λ 3,0 closed}. The closed (3,0)-form is given by Ψ = ρ + ijρ ρ. The next result provides an equivalent condition to determine the existence of such complex structures on g. Lemma 2.2. Let g be a Lie algebra and ν a volume form on g. Then, g admits an almost complex structure with a nonzero closed (3, 0)-form if and only if there exists ρ Z 3 (g) such that the endomorphism J ρ : g g defined by (( J ρ α)(x))ν = α ι Xρ ρ, (4) for any α g and X g, satisfies that J ρ ρ is closed and tr( J 2 ρ )<0. Proof. Let J : g g be an almost complex structure admitting a nonzero (3,0)-form Ψ = Ψ + + iψ which is closed. Let ρ = Ψ +. Then, λ(ρ) < 0, J = J ρ is determined by (2), and

6 A. Fino et al. the form Jρ ρ = Ψ is closed. Since the associated form φ(ρ) is a volume form on g, we have that ν = s φ(ρ) for some s =0. Now, for the endomorphism J ρ : g g given by (4), we get s(( J ρ α)(x))φ(ρ) = (( J ρ α)(x))ν = α ι Xρ ρ = ((Jρ α)(x))φ(ρ), for any α g and X g. ThisimpliesthatJρ = s J ρ. Therefore, tr( J ρ 2)<0ifandonlyif tr(jρ 2)<0, and moreover, d( J ρ ρ) = 0ifandonlyifd(J ρ ρ) = 0. For the computation of the endomorphism J ρ, we will use the simplest volume form ν = e 123456, where {e 1,...,e 6 } is the basis of g in which the Lie algebra is expressed. Note that we will follow the notation given in [14, 15, 28, 31] to name the Lie algebras; for instance, the notation e(2) e(1, 1) = (0, e 13, e 12, 0, e 46, e 45 ) means that e(2) e(1, 1) is the (decomposable) Lie algebra determined by a basis {e i } 6 i=1 such that de1 = 0, de 2 = e 13, de 3 = e 12, de 4 = 0, de 5 = e 46, de 6 = e 45. The next two concrete examples show how we will proceed in general in the proofs of Propositions 2.6 and 2.7 in order to exclude candidates. Some of the computations in these examples, as well as in Propositions 2.6 and 2.7 and in Section 3, were carried out using Mathematica and the differential forms package scalaredc by S. Bonanos available at www.inp.demokritos.gr/sbonano/. Example 2.3. Let us consider the indecomposable solvable Lie algebra g = A 0, 1 6,25 = (e 23, e 26, e 36, 0, e 46, 0). Anyρ Z 3 (g) is given by ρ = a 1 e 123 + a 2 e 126 + a 3 e 136 + a 4 e 234 + a 5 (e 235 e 146 ) + a 6 e 236 + a 7 e 246 + a 8 e 256 + a 9 e 346 + a 10 e 356 + a 11 e 456, for a 1,...,a 11 R. Fixν = e 123456,andlet J ρ calculation shows that be the endomorphism given by (4). A direct tr( J 2 ρ ) = 6(a2 5 a 1a 11 ) 2 0. In this case, it is not worth evaluating the closedness of J ρ ρ, because by Lemma 2.2, there is no almost complex structure Jρ coming from a closed 3-form ρ Z 3 (g) and in particular, g does not admit a closed complex volume form.

Six-Dimensional Solvmanifolds with Holomorphically Trivial Canonical Bundle 7 Example 2.4. Let us consider the 5 1 decomposable solvable Lie algebra g = A 1 5,15 R = (e 15 + e 25, e 25, e 35 + e 45, e 45, 0, 0). Anyρ Z 3 (g) is given by ρ = a 1 e 125 + a 2 e 135 + a 3 e 145 + a 4 e 156 + a 5 e 235 + a 6 (e 236 e 146 ) + a 7 e 245 + a 8 e 246 + a 9 e 256 + a 10 e 345 + a 11 e 356 + a 12 e 456, for a 1,...,a 12 R. Take ν = e 123456,andlet J ρ be the endomorphism given by (4). Then, we have and 1 6 tr( J ρ 2 ) = (a 3 + a 5 ) 2 a6 2 + 4(a 1a 10 a 2 a 7 )a6 2 2(a 3 a 5 )a 2 a 6 a 8 + a2 2 a2 8 d( J ρ ρ) = 2a2 6 (2a 1 e 1256 + a 2 (e 1456 + e 2356 ) + (a 3 + a 5 ) e 2456 2a 10 e 3456 ). Since the form J ρ ρ must be closed, we distinguish two cases depending on the vanishing of the coefficient a 6.Ifa 6 = 0, then tr( J ρ 2) = 6(a 2a 8 ) 2 0, and if a 6 0, then a 1 = a 2 = a 3 + a 5 = a 10 = 0andsotr( J ρ 2 ) = 0. Consequently, Lemma 2.2 assures that there is no almost complex structure Jρ admitting a nonzero closed (3,0)-form. From now on, we shall denote by b k (g) the dimension of the kth Chevalley Eilenberg cohomology group H k (g) of the Lie algebra. Lemma 2.5 shows a simple but useful obstruction to the existence of complex structures with nonzero closed (3, 0)- volume forms in the unimodular case involving b 3. We remind that the unimodularity of g is equivalent to b 6 (g) = 1. Lemma 2.5. If g is an unimodular Lie algebra admitting a complex structure with a nonzero closed (3, 0)-form Ψ,thenb 3 (g) 2. Proof. Let Ψ +,Ψ Λ 3 (g ) be the real and imaginary parts of Ψ,thatis,Ψ = Ψ + + i Ψ. Since Ψ is closed we have that d(ψ + ) = d(ψ ) = 0 and therefore, [Ψ + ], [Ψ ] H 3 (g). Itis sufficient to see that both classes are nonzero and, moreover, that they are not cohomologous. Suppose that there exist a, b R with a 2 + b 2 0 such that aψ + + bψ = dα for some α Λ 2 (g ).Since0 = i 2 Ψ Ψ = Ψ + Ψ Λ 6 (g ), we get d(α ( bψ + + aψ )) = (aψ + + bψ ) ( bψ + + aψ ) = (a 2 + b 2 )Ψ + Ψ 0. But this is in contradiction to the unimodularity of g.

8 A. Fino et al. As a consequence of Lemma 2.5, we will concentrate on unimodular (nonnilpotent) solvable Lie algebras g with b 3 (g) 2. We will tackle the classification problem first in the decomposable case. Let g = b c. The unimodularity and solvability of g and Lemma 2.5 imply restrictions on the factors. In fact, g is unimodular, resp. solvable, if and only if b and c are unimodular, resp. solvable. Moreover, by Lemma 2.5 and the well-known formula relating the cohomology of g with the cohomologies of the factors, we have Proposition 2.6. b 3 (b)b 0 (c) + b 2 (b)b 1 (c) + b 1 (b)b 2 (c) + b 0 (b)b 3 (c) = b 3 (g) 2. (5) Let g = b c be a 6D decomposable unimodular (non-nilpotent) solvable Lie algebra admitting a complex structure with a nonzero closed (3, 0)-form. Then, g is isomorphic to e(2) e(1, 1), A 1, 1,1 5,7 R or A α, α,1 5,17 R with α 0. Proof. Since g is decomposable, we divide the proof into the three cases 3 3, 4 2, and 5 1. In the 3 3 case, the inequality (5) is always satisfied. The 3 3 decomposable unimodular (non-nilpotent) solvable Lie algebras are e(2) e(2), e(2) e(1, 1), e(2) h 3, e(2) R 3, e(1, 1) e(1, 1), e(1, 1) h 3, and e(1, 1) R 3 (see Table A1 in the Appendix for a description of the Lie algebras). An explicit computation shows that there is no ρ Z 3 satisfying the conditions λ(ρ) < 0andd(Jρ ρ) = 0, except for g = e(2) e(1, 1). Wegivean example of a closed complex volume form for e(2) e(1, 1) in Table A1. Since R 2 is the only 2D unimodular Lie algebra, the 4 2 case is reduced to the study of g = b R 2 for any 4D unimodular (non-nilpotent) solvable Lie algebra b satisfying (5), that is, b 3 (b) + 2b 2 (b) + b 1 (b) 2. The resulting Lie algebras are: A 2 4,2 R2, A α, 1 α 4,5 R 2 with 1 <α 1 2, α Aα, 2 4,6 R 2, A 4,8 R 2, and A 4,10 R 2 (see Table A1). However, all of them satisfy λ(ρ) 0 for any ρ Z 3 (g). Finally, the 5 1 case consists of Lie algebras of the form g = b R for any 5D unimodular (non-nilpotent) solvable Lie algebra b such that (b 2 (b), b 3 (b)) = (0, 0), (1, 0), (0, 1). Therefore, the Lie algebras are: A 1, 1,1 5,7 R, A 1,β, β 5,7 R with 0 <β<1, A 1 5,8 R, A 1, 1 5,9 R, A 1,0,γ 5,13 R with γ>0, A 0 5,14 with 0 <γ <1, A α, α,1 5,17 with α 0, A 0 5,18 A 1, 1 5,33 R, and A 0, 2 5,35 the following three situations: R, A 1 5,15 R, A0,0,γ 5,17 R R, A0,±1 R, A 1,2 5,19 R, A1, 2 5,19 R, A0 5,20 5,26 R, R. The explicit computation of each case allows us to distinguish If g = A 1, 1 5,9 R or A 0,±1 5,26 R, thenλ(ρ) 0 for all ρ Z 3 (g). The Lie algebras A 1, 1,1 5,7 R and A α, α,1 5,17 R with α 0 admit a closed complex volume form (see Table A1 for a concrete example).

Six-Dimensional Solvmanifolds with Holomorphically Trivial Canonical Bundle 9 For the rest of Lie algebras, there is no ρ Z 3 (g) satisfying d(j ρ ρ) = 0and λ(ρ) < 0 simultaneously. In conclusion, in the 5 1 case, the only possibilities are A 1, 1,1 5,7 R and the family A α, α,1 5,17 R with α 0. Next, we obtain the classification in the indecomposable case. Proposition 2.7. Let g be a 6D indecomposable unimodular (non-nilpotent) solvable Lie algebra admitting a complex structure with a nonzero closed (3, 0)-form. Then, g is isomorphic to N 0, 1, 1 6,18, A 0,0,1 6,37, A0,1,1 6,82, A0,0,1 6,88, B1 6,4,orB1 6,6. Proof. The Lie algebras g such that b 3 (g) 2 are listed in Table A2 in the Appendix. The indecomposable case is long to analyze because of the amount of Lie algebras, but after performing the computations, we distinguish the following three situations: Let g be one of the following Lie algebras: A a, 2a,2a 1 6,13 (a R { 1, 0, 1 3, 1 2 }), A a, a, 1 6,13 (a > 0, a =1), A 1 3, 2 3 6,14, Aa,b 6,18 with (a, b) {( 1, 2), ( 2, 1)}, Aa,b 2 6,25 with (a, b) {(0, 1), ( 1 2, 1 )}, A0,b, b 2 6,32 (b > 0), A 0,0,ɛ 6,34 (ɛ = 0, 1), Aa,b,c 6,35 with a > 0and (b, c) {( 2a, a), ( a, 0)}, and A 0,0,c 6,37 (c > 0, c =1). Then, λ(ρ) 0 for any ρ Z 3 (g). The Lie algebras N 0, 1, 1 6,18, A 0,0,1 6,37, A0,1,1 6,82, A0,0,1 6,88, B1 6,4,andB1 6,6 admit a nonzero closed (3,0)-form (see Table A2 for a concrete example). For the rest of Lie algebras, there is no ρ Z 3 (g) such that d(j ρ ρ) = 0and λ(ρ) < 0. From Propositions 2.6 and 2.7, it follows the following classification: Theorem 2.8. Let g be a unimodular (non-nilpotent) solvable Lie algebra of dimension 6. Then, g admits a complex structure with a nonzero closed (3, 0)-form if and only if it is isomorphic to one in the following list: g 1 = A 1, 1,1 5,7 R = (e 15, e 25, e 35, e 45, 0, 0), g α 2 = Aα, α,1 5,17 R = (α e 15 + e 25, e 15 + α e 25, α e 35 + e 45, e 35 α e 45, 0, 0), α 0, g 3 = e(2) e(1, 1) = (0, e 13, e 12, 0, e 46, e 45 ), g 4 = A 0,0,1 6,37 = (e23, e 36, e 26, e 56, e 46, 0),

10 A. Fino et al. g 5 = A 0,1,1 6,82 = (e24 + e 35, e 26, e 36, e 46, e 56, 0), g 6 = A 0,0,1 6,88 = (e24 + e 35, e 36, e 26, e 56, e 46, 0), g 7 = B 1 6,6 = (e24 + e 35, e 46, e 56, e 26, e 36, 0), g 8 = N 0, 1, 1 6,18 = (e 16 e 25, e 15 + e 26, e 36 + e 45, e 35 e 46, 0, 0), g 9 = B 1 6,4 = (e45, e 15 + e 36, e 14 e 26 + e 56, e 56, e 46, 0). From now on, we will use only the simplified notation g 1, g α 2, g 3,...,g 9 when referring to the Lie algebras listed in the previous theorem. 2.1 Existence of lattices In this section, we show that the simply connected solvable Lie groups G k corresponding to the Lie algebras g k in Theorem 2.8 admit lattices Γ k of maximal rank. Therefore, we get compact complex solvmanifolds Γ k \G k with holomorphically trivial canonical bundle. Let H be a n-dimensional nilpotent Lie group. We remind that a connected and simply connected Lie group G with nilradical H is called almost nilpotent (resp. almost abelian) if it can be written as G = R μ H (resp. G = R μ R n,thatis,h = R n ). If we denote by e the identity element of H, then following [6, 8] wehavethatd e (μ(t)) = exp GL(n,R) (tϕ), ϕ being a certain derivation of the Lie algebra h of H. Although it is not easy in general to know whether a solvable Lie group G admits a lattice, the next result allows to construct one in the case that G is almost nilpotent. Lemma 2.9 ([6]). Let G = R μ H be a (n+ 1)-dimensional almost nilpotent Lie group with nilradical H and h the Lie algebra of H. If there exists 0 =t 1 R and a rational basis {X 1,...,X n } of h such that the coordinate matrix of d e (μ(t 1 )) in such basis is integer, then Γ = t 1 Z μ exp H (Z X 1,...,X n ) isalatticeing. Now, we will use the former lemma to prove the following result: Proposition 2.10. For any k =2, the connected and simply connected Lie group G k with underlying Lie algebra g k admits a lattice. For k = 2, there exists a countable number of distinct αs, including α = 0, for which the connected and simply connected Lie group with underlying Lie algebra g α 2 admits a lattice.

Six-Dimensional Solvmanifolds with Holomorphically Trivial Canonical Bundle 11 Proof. The Lie algebra g 8 is not almost nilpotent, but its corresponding connected and simply connected Lie group G 8 admits a lattice by [33]. It is not hard to see that for k =8 the Lie algebra g k of Theorem 2.8 is either almost nilpotent or a product of almost nilpotent Lie algebras. In fact, we find the following correspondence with some of the Lie algebras studied in [6] (we use the notation in that paper in order to compare directly with the Lie algebras therein): g 1 = g 1, 1,1 5,7 R, g 0 2 = g 0,0,1 5,17 R, g 3 = g 0 3,5 g 1 3,4, g 4 = g 0,0, 1 6,37, g 5 = g 0, 1,0 6,88, g 6 = g 0, 1, 1 6,92, and g 7 = g 6,92. The simply connected Lie group G 3 admits a lattice, since it is product of two 3D Lie groups and every 3D completely solvable simply connected Lie group has a lattice. For g 1, g 0 2, g 4, g 5, g 6, and g 7, the existence of lattices in the corresponding Lie groups is already proved in [6]. In fact, a lattice is respectively given by Γ 1 = (t 1 Z μ exp H (Z e 1,...,e 4 )) Z, H = R 4, t 1 = log Γ 0 2 = (t 1Z μ exp H (Z e 1,...,e 4 )) Z, H = R 4, t 1 = π; Γ 4 = t 1 Z μ exp H (Z e 1,...,e 5 ), H = Heis 3 R 2, t 1 = π; Γ 5 = t 1 Z μ exp (Z H 1 e 1, 5 1 (e 3 e 5 ), αe 3 1 5 5α e 5, ( 5(3 + 5) α =, H = Heis 5, t 1 = log 10 3 + 5 2 Γ 6 = t 1 Z μ exp H (Z e 1,...,e 5 ), H = Heis 5, t 1 = π; ) ; ( 3 + ) 5 ; 2 1 (e 2 e 4 ), αe 2 1 ) 5 5α e 4, Γ 7 = t 1 Z μ exp H (Z 2e 1, e 2 e 3, e 4 + e 5, e 2 e 3, e 4 e 5 ), H = Heis 5, t 1 = π, where Heis 3 and Heis 5 are the real Heisenberg group of dimension 3 and 5. So, it remains to study g α 2 with α>0, and g 9. We will show first that there exists a countable subfamily of g α 2 with α>0 whose corresponding Lie group G α 2 admits lattice. The 5D factor Aα, α,1 5,17 in the decomposable Lie algebra g α 2 = Aα, α,1 5,17 R is given by [e 1, e 5 ] = αe 1 + e 2, [e 2, e 5 ] = e 1 αe 2, [e 3, e 5 ] = αe 3 + e 4, [e 4, e 5 ] = e 3 + αe 4, which is almost abelian since A α, α,1 5,17 = R ade5 R 4. If we denote by B α the coordinate matrix of the derivation ad e5 : R 4 R 4 in the basis {e 1, e 2, e 3, e 4 }, then the coordinate

12 A. Fino et al. matrix of d e (μ(t)) is the exponential e αt cos(t) e αt sin(t) 0 0 e tb α = e αt sin(t) e αt cos(t) 0 0 0 0 e αt cos(t) e αt sin(t). 0 0 e αt sin(t) e αt cos(t) If t l = lπ with l Z and l > 0, then the characteristic polynomial of the matrix e t l B α p(λ) = (1 ( 1) l (e αt l + e αt l )λ + λ 2 ) 2, which is integer if α l,m = 1 lπ log( m+ m 2 4 2 ) with m Z and m > 2. Moreover, e t l B α = P 1 C l,m P, where 0 0 ɛ β + 0 1 0 0 P 1 = ɛ β + 0 0 0 0 ɛ β, C l,m = 1 m( 1) l 0 0 0 0 0 1 ɛ β 0 0 0 0 1 m( 1) l with ɛ = 1 and m 2 4 β± = m2 4±( 1) l m 2 4. Taking the basis X 2(m 2 4) 1 = ɛ(e 2 e 4 ), X 2 = β + e 2 + β e 4, X 3 = ɛ(e 1 + e 3 ), and X 4 = β + e 1 + β e 3 of R 4 and using Lemma 2.9, we have that Γ = lπz μ Z X 1,...,X 4 is a lattice of the simply connected Lie group associated with A α, α,1 5,17 with α = α l,m. Hence, Γ = Γ Z isalatticeing α l,m 2. The Lie algebra g 9 can be seen as an almost nilpotent Lie algebra g = R ade6 h, where h = e 1,...,e 5 [e 1, e 4 ] = e 3, [e 1, e 5 ] = e 2, [e 4, e 5 ] = e 1 is a 5D nilpotent Lie algebra. Proceeding in a similar manner as for g α 2, we compute the characteristic polynomial of d e (μ(t)) getting that p(λ) = (λ 2 2λcos(t) + 1) 2.Ift 1 = π, then p(λ) Z[λ] and the coordinate matrix of d e (μ(t 1 )) on the basis X 1 = π 2 e 1, X 2 = π 2 e 4, X 3 = π 2 e 5, X 4 = ( π 2 )3/2 e 2 + π 2 e 4 and X 5 = ( π 2 )3/2 e 3 + π 2 e 5 of h is 1 0 0 0 0 0 2 0 1 0 C = 0 0 2 0 1. 0 1 0 0 0 0 0 1 0 0, is Moreover, {X 1,...,X 5 } is a rational basis of h because [X 1, X 2 ] = [X 1, X 4 ] = X 3 + X 5, [X 1, X 3 ] = [X 1, X 5 ] = X 2 X 4,[X 2, X 3 ] = [X 2, X 5 ] = [X 3, X 4 ] = X 1. Hence, if we denote by

Six-Dimensional Solvmanifolds with Holomorphically Trivial Canonical Bundle 13 H the simply connected Lie group corresponding to h, then using Lemma 2.9 we have that Γ = πz μ exp H (Z X 1,...,X 5 ) is a lattice in the Lie group G 9. Remark 2.11. Bock found a lattice for the Lie group associated with A α, α,1 2 with α = α 1,3 = 1 π log 3+ 5 2,thatis,forl = 1andm = 3. Note that our result for k = 2 is consistent with the result obtained by Witte [32, Proposition 8.7], where it is showed that only countably many nonisomorphic simply connected Lie groups admit a lattice, so that one cannot expect a lattice to exist for any real α>0. The Lie algebra g 9 does not appear in [6] because its nilradical is the 5D Lie algebra h, which is isomorphic to g 5,3 (in the notation of [6]), but there are only two solvable and unimodular Lie algebras with nilradical g 5,3 considered in that paper (namely g 1 6,76 and g 6,78 ) which are both completely solvable, but g 9 is not. 3 Moduli of Complex Structures In this section, we classify, up to equivalence, the complex structures having closed (3,0)-form on the Lie algebras of Theorem 2.8. Recall that two complex structures J and J on g are said to be equivalent if there exists an automorphism F : g g such that F J = J F. 3.1 The decomposable case We consider first the 5 1 decomposable Lie algebras. Lemma 3.1. Let J be any complex structure on g 1 or g α 2, α 0, with closed volume (3, 0)- form, then there is a nonzero closed (1, 0)-form. More concretely, the (1, 0)-form e 5 ij e 5 is closed. Proof. Let us consider first g = g 1 with structure equations given as in Theorem 2.8. Any ρ Z 3 (g) is given by ρ = a 1 e 125 + a 2 e 126 + a 3 e 135 + a 4 e 136 + a 6 e 156 + a 8 e 245 + a 9 e 246 + a 10 e 256 + a 11 e 345 + a 12 e 346 + a 13 e 356 + a 14 e 456, where a i R. We use Equation (2) to compute the space of almost complex structures corresponding to ρ Z 3 (g). When we compute the images of e 5, e 6 by Jρ, we find that the

14 A. Fino et al. subspace spanned by e 5, e 6 is Jρ -invariant, because J ρ e5 = J ρ e6 = 1 λ(ρ) ((a 1 a 12 + a 11 a 2 a 4 a 8 a 3 a 9 ) e 5 + 2(a 12 a 2 a 4 a 9 ) e 6 ), 1 λ(ρ) ( 2(a 1 a 11 a 3 a 8 ) e 5 (a 1 a 12 + a 11 a 2 a 4 a 8 a 3 a 9 ) e 6 ). Therefore, the (1,0)-form η = e 5 ij ρ e5 is closed for every ρ Z 3 (g). The same situation appears for g = g α 2, α 0, because again the subspace spanned by e 5, e 6 is found to be Jρ -invariant for all ρ Z 3 (g). Lemma 3.2. Let J be any complex structure on g 1 or g α 2, α 0, with closed volume (3, 0)- form. Then, there is a (1, 0)-basis {ω 1,ω 2,ω 3 } satisfying the following reduced equations where A= cos θ + i sin θ, θ [0,π). dω 1 = Aω 1 (ω 3 + ω 3 ), dω 2 = Aω 2 (ω 3 + ω 3 ), dω 3 = 0, Proof. By Lemma 3.1, we can consider a basis of (1, 0)-forms {η 1,η 2,η 3 } such that η 3 = e 5 ij e 5 is closed. The structure equations of g 1 and g α 2 with α 0 force the differential of any 1-form to be a multiple of e 5 = 1 2 (η3 + η 3 ), so there exist A, B, C, D, E, F C such that dη 1 = (Aη 1 + Bη 2 + Eη 3 ) (η 3 + η 3 ), dη 2 = (C η 1 + Dη 2 + F η 3 ) (η 3 + η 3 ), dη 3 = 0. Moreover, since d(η 123 ) = 0 necessarily D = A. Let us consider the nonzero 1-form τ 1 = Aη 1 + Bη 2 + Eη 3.Notethat (6) dτ 1 = ((A 2 + BC)η 1 + (AE + BF)η 3 ) (η 3 + η 3 ), which implies that A 2 + BC =0, because, otherwise, dτ 1 would be a multiple of e 56. Then, with respect to the new (1,0)-basis {τ 1,τ 2,τ 3 } given by τ 1 = Aη 1 + Bη 2 + Eη 3, τ 1 = C η 1 Aη 2 + F η 3, τ 3 = η 3,

Six-Dimensional Solvmanifolds with Holomorphically Trivial Canonical Bundle 15 the complex structure equations are dτ 1 = (Aτ 1 + Bτ 2 ) (τ 3 + τ 3 ), dτ 2 = (C τ 1 Aτ 2 ) (τ 3 + τ 3 ), dτ 3 = 0. (7) Now, we distinguish two cases: If B =0, then we consider the new basis {ω 1,ω 2,ω 3 } given by ω 1 = (A + A 2 + BC)τ 1 + B τ 2, ω 2 = (A A 2 + BC)τ 1 + Bτ 2, ω 3 = A 2 + BC τ 3. With respect to this basis, Equations (7) reduce to dω 1 A2 + BC = A 2 + BC ω1 (ω 3 + ω 3 ), dω 2 A2 + BC = A 2 + BC ω2 (ω 3 + ω 3 ), dω 3 = 0, that is, the equations are of the form (6) where the new complex coefficient has modulus equal to 1. If C =0, then with respect to the basis {ω 1,ω 2,ω 3 } given by ω 1 = C τ 1 (A A 2 + BC)τ 2, ω 2 = C τ 1 (A + A 2 + BC)τ 2, ω 3 = A 2 + BC τ 3, Equations (7) again reduce to equations of the form (6) where the new complex coefficient has modulus 1. Finally, note that in Equations (6) one can change the sign of A by changing the sign of ω 3, so we can suppose that A= cos θ + i sin θ with angle θ [0,π). Proposition 3.3. Up to isomorphism, there is only one complex structure with closed (3, 0)-form on the Lie algebras g 1 and g 0 2, whereas gα 2 has two such complex structures

16 A. Fino et al. for any α>0. More concretely, the complex structures are (g 1, J): dω 1 = ω 1 (ω 3 + ω 3 ), dω 2 = ω 2 (ω 3 + ω 3 ), dω 3 = 0; (8) (g 0 2, J): dω1 = iω 1 (ω 3 + ω 3 ), dω 2 = iω 2 (ω 3 + ω 3 ), dω 3 = 0; (9) α= cos θ sin θ (g2, J ± ): dω 1 = (± cos θ + i sin θ) ω 1 (ω 3 + ω 3 ), dω 2 = (± cos θ + i sin θ) ω 2 (ω 3 + ω 3 ), dω 3 = 0, where θ (0,π/2). (10) Proof. A real Lie algebra underlying Equations (6) is isomorphic to g 1 or g α 2 for some α 0. In fact, in terms of the real basis β 1,...,β 6 given by ω 1 = β 1 + iβ 2, ω 2 = β 3 + iβ 4, and ω 3 = 1 2 (β5 + iβ 6 ),wehave In particular: dβ 1 = cos θβ 15 sin θβ 25, dβ 3 = cos θβ 35 + sin θβ 45, dβ 5 = 0, dβ 2 = sin θβ 15 + cos θβ 25, dβ 4 = sin θβ 35 cos θβ 45, dβ 6 = 0. If θ = 0, then taking e 1 = β 1,e 2 = β 4,e 3 = β 3,e 4 = β 2,e 5 = β 5, and e 6 = β 6 the resulting structure equations are precisely those of the Lie algebra g 1. If θ (0,π), then sin θ =0 and taking e 1 = β 1,e 2 = β 2,e 3 = β 3,e 4 = β 4,e 5 = sin θβ 5, and e 6 = β 6 we get the structure equations of g α 2 with α = cos θ sin θ. Note that α takes any real value when θ varies in (0,π),andifθ = π 2, then θ and π θ correspond to two complex structures on the same Lie algebra. By a standard argument, one can prove that these two complex structures are nonequivalent. Let us consider now the 3 3 decomposable Lie algebra g 3. With respect to the structure equations given in Theorem 2.8, any closed 3-form ρ Z 3 (g 3 ) is given by ρ = a 1 e 123 + a 2 e 124 + a 3 e 134 + a 4 e 145 + a 5 e 146 + a 6 e 156 + a 7 e 234 + a 8 (e 136 e 245 ) + a 9 (e 135 e 246 ) + a 10 (e 126 + e 345 ) + a 11 (e 125 + e 346 ) + a 12 e 456,

Six-Dimensional Solvmanifolds with Holomorphically Trivial Canonical Bundle 17 ρ where a 1,...,a 12 R. By imposing the closedness of Jρ ρ together with the condition tr(j 2 )<0, one arrives by a long computation to an explicit description of the complex structure J ρ, which allows us to prove that {e 1, e 2, e 3, J ρ e1, J ρ e2, J ρ e3 } are always linearly independent. Therefore, the forms ω 1 = e 1 ij ρ e1, ω 2 = e 2 ij ρ e2, ω 3 = e 3 ij ρ e3, constitute a (1,0)-basis for the complex structure J ρ. Moreover, one can show that with respect to this basis the complex structure equations have the form dω 1 = 0, dω 2 = 1 2 ω13 + bω 1 1 + fiω 1 2 fiω 2 1 ( 1 2 + gi) ω 1 3 + giω 3 1, dω 3 = 1 2 ω12 + cω 1 1 + ( 1 2 + hi) ω 1 2 hiω 2 1 fiω 1 3 + fiω 3 1, where the coefficients b, c, f, g, h are real and satisfy 4gh = 4 f 2 1. (Explicit expression of each coefficient in terms of a 1,...,a 12 can be given, but this information is not relevant and so we omit it.) Note that the condition 4gh = 4 f 2 1 is equivalent to the Jacobi identity d 2 = 0. Furthermore, these equations can be reduced as follows: Proposition 3.4. Up to isomorphism, the complex structures with closed (3, 0)-form on the Lie algebra g 3 are where x R +. Proof. (g 3, J x ): dω 1 = 0, dω 2 = 1 2 ω13 ( 1 2 + xi) ω 1 3 + xiω 3 1, dω 3 = 1 2 ω12 + ( 1 2 ) i 4x ω 1 2 + i ω2 1, 4x Observe first that with respect to the (1,0)-basis {ω 1,ω 2 + 2cω 1,ω 3 2bω 1 },the complex structure equations express again as in (11) butwithb = c = 0, that is to say, one can suppose that the coefficients b and c both vanish. Let us prove next that we can also take the coefficient f to be zero. To see this, let {ω 1,ω 2,ω 3 } be a (1,0)-basis satisfying (11)withb = c = 0and f =0, and let us consider the (1,0)-basis {η 1,η 2,η 3 } given by (11) (12) η 1 = ω 1, η 2 = ω 2 g h 1 + (g + h) 2 2 f ω 3, η 3 = g h 1 + (g + h) 2 ω 2 + ω 3. 2 f

18 A. Fino et al. A direct calculation shows that with respect to {η 1,η 2,η 3 } the corresponding coefficient f vanishes. Therefore, since 4gh = 1, we are led to the reduced Equations (12), where we have written x instead of g. Finally, let J x and J x be two complex structures corresponding to x, x R. Itis easy to see that the structures are equivalent if and only if xx = 1. This represents an 4 hyperbola in the (x, x )-plane, so the equivalence class is given by one of the branches of the hyperbola, that is, we can take x > 0. 3.2 The indecomposable case Next, we classify the complex structures on the indecomposable Lie algebras g 4,...,g 9. Lemma 3.5. Let J be any complex structure on g k (4 k 7) with closed (3, 0)-form. Then, there is a (1, 0)-basis {ω 1,ω 2,ω 3 } such that dω = Aω 1 (ω 3 + ω 3 ), dω 2 = Aω 2 (ω 3 + ω 3 ), (13) dω 3 = G 11 ω 1 1 + G 12 ω 1 2 + Ḡ 12 ω 2 1 + G 22 ω 2 2, where A, G 12 C and G 11, G 22 R, with(g 11, G 12, G 22 ) =(0, 0, 0), satisfy A =1, (A + Ā)G 11 = 0, (A + Ā)G 22 = 0, (A Ā)G 12 = 0. (14) Proof. Let us consider first the Lie algebra g 4 with structure equations given as in Theorem 2.8. Any element ρ Z 3 (g 4 ) is given by ρ = a 1 e 123 + a 2 e 126 + a 3 (e 125 e 134 ) + a 4 (e 124 + e 135 ) + a 5 e 136 + a 6 (e 156 + e 234 ) + a 7 (e 146 e 235 ) + a 8 e 236 + a 9 e 246 + a 10 e 256 + a 11 e 346 + a 12 e 356 + a 13 e 456, where a 1,...,a 13 R. A direct calculation shows that if a3 2 + a2 4 = 0, there do not exist closed 3-forms ρ satisfying the conditions d(j ρ ρ) = 0andλ(ρ) < 0. Suppose that a3 2 + a2 4 0. Then, an element ρ Z 3 (g 4 ) satisfies the condition d(j ρ ρ) = 0 if and only if a 10 = a 3(a6 2 a2 7 )+2a 4a 6 a 7 a 11 (a3 2+a2 4 ), a a3 2 12 = 2a 3a 6 a 7 a 4 (a6 2 a2 7 )+a 9(a3 2+a2 4 ), and +a2 4 a3 2+a2 4 a 13 = 0. Moreover, under these relations, one has that λ(ρ) = 4(a 3 a 9 a 4 a 11 + a 6 a 7 ) 2 0.

Six-Dimensional Solvmanifolds with Holomorphically Trivial Canonical Bundle 19 Let ρ Z 3 (g 4 ) be such that λ(ρ) < 0andd(Jρ ρ) = 0. A direct calculation shows that J ρ e6 is given by J ρ e6 = 2(a 2 3 + a2 4 ) e1 + 2(a 3 a 6 + a 4 a 7 ) e 2 + 2(a 3 a 7 a 4 a 6 ) e 3 + 2(a 3 a 11 + a 4 a 9 + a 2 7 ) e6. Therefore, the coefficient of J ρ e6 in e 1 is nonzero for any ρ. A similar computation for the Lie algebras g 5, g 6, and g 7 shows that for any complex structure J ρ with closed (3, 0)-form,wealsohavethat where the coefficient c 1 is nonzero. J ρ e6 = c 1 e 1 + c 2 e 2 + c 3 e 3 + c 4 e 4 + c 5 e 5 + c 6 e 6, Let us consider the (1, 0)-form η 3 = e 6 ij ρ e6. From the structure equations of g k (4 k 7) in Theorem 2.8, it follows that dη 3 = ic 1 e 23 iα e 6, if g = g 4, dη 3 = ic 1 (e 24 + e 35 ) iα e 6, if g = g 5, g 6, g 7, where α is a 1-form. Since c 1 0, we can write the 2-forms e 23 and e 24 + e 35 as e 23 = i c 1 dη 3 + 1 c 1 α e 6, if g = g 4, e 24 + e 35 = i c 1 dη 3 + 1 c 1 α e 6, if g = g 5, g 6, g 7. Now, let η 1,η 2 be such that {η 1,η 2,η 3 } isabasisof(1, 0)-forms. Since e 6 is closed and η 3 + η 3 = 2e 6, the integrability of the complex structure implies that dη 3 has no component of type (2, 0) and dη 3 = G 11 η 1 1 + G 12 η 1 2 + G 13 η 1 3 + Ḡ 12 η 2 1 + G 22 η 2 2 + G 23 η 2 3 + Ḡ 13 η 3 1 + Ḡ 23 η 3 2 + G 33 η 3 3, for some G 11, G 22, G 33 R and G 12, G 13, G 23 C. From the structure of the Lie algebras g k (4 k 7), the relation (15) and taking into account that dη 3 is of type (1,1), it follows that there exist λ, μ C such that dη 1 = λdη 3 + (Aη 1 + Bη 2 + Eη 3 ) (η 3 + η 3 ), dη 2 = μdη 3 + (C η 1 + Dη 2 + F η 3 ) (η 3 + η 3 ), dη 3 = G 11 η 1 1 + G 12 η 1 2 + G 13 η 1 3 +Ḡ 12 η 2 1 + G 22 η 2 2 + G 23 η 2 3 +Ḡ 13 η 3 1 + Ḡ 23 η 3 2 + G 33 η 3 3, (15) (16)

20 A. Fino et al. for some A, B, C, D, E, F C. We will see next that these complex equations can be reduced to equations of the form (13). Note first that with respect to the (1,0)-basis {η 1 λη 3,η 2 μη 3,η 3 } we get complex equations of the form (16) withλ = μ = 0. So, without loss of generality, we can suppose λ = μ = 0. Moreover, the coefficients E and F also vanish. In fact, suppose for example that E =0 (the case F =0 is similar). Using (16) withλ = μ = 0, the condition d(dη 1 ) = 0 is equivalent to EG 11 = EG 12 = EG 13 = EG 22 = EG 23 = 0, so E =0 implies dη 3 = G 33 η 3 3 = G 33 η 3 (η 3 + η 3 ). But this is a contradiction with the structure of the Lie algebras g k (4 k 7), because d(g k ) would be annihilated by the real 1-form η 3 + η 3. From now on, we suppose that λ = μ = E = F = 0 in Equations (16). A direct calculation shows that dη 123 = Ḡ 13 η 123 1 + Ḡ 23 η 123 2 + (A + D + G 33 )η 123 3, so η 123 is closed if and only if G 13 = G 23 = 0andD = A G 33. Moreover, the unimodularity of the Lie algebras g k (4 k 7) implies that G 33 = 0. In fact, taking the real basis { f 1,..., f 6 } of g k given by η 1 = f 2 + if 3, η 2 = f 4 + if 5, η 3 = f 6 + if 1, we get that the trace of ad f6 is zero if and only if G 33 = 2Re A 2Re D, which implies, using that G 33 = A D, that the coefficient G 33 = 0. Summing up, we have proved the existence of a (1, 0)-basis {η 1,η 2,η 3 } satisfying the reduced complex equations dη 1 = (Aη 1 + Bη 2 ) (η 3 + η 3 ), dη 2 = (C η 1 Aη 2 ) (η 3 + η 3 ), dη 3 = G 11 η 1 1 + G 12 η 1 2 + Ḡ 12 η 2 1 + G 22 η 2 2, (17) where A, B, C, G 12 C and G 11, G 22 R. Note that A 2 + BC =0 because otherwise the (1,0)-form Aη 1 + Bη 2 would be closed, but this is a contradiction to b 1 (g k ) = 1, for 4 k 7. Therefore, arguing as in the proof of Lemma 3.2 we can suppose that B = C = 0 and A = 1 in (17). Finally, the condition d(dη 3 ) = 0 is satisfied if and only if (A + Ā)G 11 = (A + Ā)G 22 = (A Ā)G 12 = 0.

Six-Dimensional Solvmanifolds with Holomorphically Trivial Canonical Bundle 21 As a consequence of the previous lemma, we have the following classification of complex structures on g k,for4 k 7. Proposition 3.6. Up to isomorphism, there is only one complex structure J with closed (3, 0)-form on the Lie algebras g 5 and g 6, and two such complex structures on the Lie algebras g 4 and g 7. More concretely, the complex structures are (g 4, J ± ): dω 1 = iω 1 (ω 3 + ω 3 ), dω 2 = iω 2 (ω 3 + ω 3 ), dω 3 =±ω 1 1 ; (18) (g 5, J): dω 1 = ω 1 (ω 3 + ω 3 ), dω 2 = ω 2 (ω 3 + ω 3 ), dω 3 = ω 1 2 + ω 2 1 ; (19) (g 6, J): dω 1 = iω 1 (ω 3 + ω 3 ), dω 2 = iω 2 (ω 3 + ω 3 ), dω 3 = ω 1 1 + ω 2 2 ; (20) (g 7, J ± ): dω 1 = iω 1 (ω 3 + ω 3 ), dω 2 = iω 2 (ω 3 + ω 3 ), dω 3 =±(ω 1 1 ω 2 2 ). (21) Proof. First note that in Equations (13), after changing the sign of ω 3 if necessary, we can always suppose that A= cos θ + i sin θ with angle θ [0,π).Wehavethefollowing cases: If cos θ =0, then (14) implies G 11 = G 22 = 0, and sin θg 12 = 0, so sin θ = 0 because (G 11, G 12, G 22 ) =(0, 0, 0) is satisfied if and only if G 12 0. Therefore, in this case A= 1 and, moreover, we can normalize the coefficient G 12 (it suffices to consider G 12 ω 1 instead of ω 1 ). So the complex structure equations take the form (19), and in terms of the real basis {e 1,...,e 6 } defined by ω 1 = e 2 i e 3, ω 2 = e 5 + i e 4 and ω 3 = 1 2 e6 2i e 1, one has de 1 = e 24 + e 35, de 2 = e 26, de 3 = e 36, de 4 = e 46, de 5 = e 56, de 6 = 0, that is, the underlying Lie algebra is g 5. If cos θ = 0, then (14) implies that A= i and G 12 = 0. Therefore, the complex structure equations become dω 1 = iω 1 (ω 3 + ω 3 ), dω 2 = iω 2 (ω 3 + ω 3 ), dω 3 = G 11 ω 1 1 + G 22 ω 2 2, where (G 11, G 22 ) =(0, 0). We have the following possibilities: When G 22 = 0, we can suppose that G 11 =±1 (it suffices to consider G11 ω 1 instead of ω 1 ), and then the complex structure equations reduce to (18). In terms of the real basis {e 1,...,e 6 } given by ω 1 =

22 A. Fino et al. e 2 i e 3, ω 2 = e 4 + i e 5, and ω 3 = 1 2 e6 ± 2i e 1, we arrive at de 1 = e 23, de 2 = e 36, de 3 = e 26, de 4 = e 56, de 5 = e 46, de 6 = 0, that is, the underlying Lie algebra is g 4. A standard argument allows to conclude that the two complex structures in (18) are nonisomorphic. The case G 11 = 0 easily reduces to the previous case, so it does not produce any nonisomorphic complex structure. Finally, if G 11 0andG 22 0, then we can suppose G 11 =±1and G 22 =±1 (it suffices to consider G kk ω k instead of ω k for k = 1, 2). It is clear that the case G 11 = G 22 = 1 is equivalent to G 11 = G 22 = 1, so it remains to study the following three cases: (G 11, G 22 ) = (1, 1), (1, 1), ( 1, 1). In terms of the real basis {β 1,...,β 6 } defined by ω 1 = β 2 + iβ 4, ω 2 = β 3 + iβ 5, and ω 3 = 1 2 β6 + 2iβ 1, one has dβ 1 = G 11 β 24 G 22 β 35, dβ 2 = β 46, dβ 3 = β 56, dβ 4 = β 26, dβ 5 = β 36, dβ 6 = 0. When (G 11, G 22 ) = (1, 1), taking the basis e 1 = 2β 1,e 2 = β 2 + β 3,e 3 = β 4 + β 5,e 4 = β 4 + β 5,e 5 = β 2 β 3, and e 6 = β 6, the real structure equations are de 1 = e 24 + e 35, de 2 = e 36, de 3 = e 26, de 4 = e 56, de 5 = e 46, de 6 = 0, so the underlying Lie algebra is g 6 and the complex structure is given by (20). The cases (G 11, G 22 ) = (1, 1) and (G 11, G 22 ) = ( 1, 1) both correspond to the same Lie algebra (in fact, a change in the sign of β 1 gives an isomorphism), so we suppose next that (G 11, G 22 ) = (1, 1), thatis, dβ 1 = β 24 + β 35, dβ 2 = β 46, dβ 3 = β 56, dβ 4 = β 26, dβ 5 = β 36, dβ 6 = 0. Taking e 1 = β 1,e 3 = β 3 and e 6 = β 6, we conclude that g 7 is the underlying Lie algebra. Therefore, the complex structures on g 7 are given by (18), and it can be proved that they are nonisomorphic. Next, we find that there are infinitely many nonisomorphic complex structures on the Lie algebra g 8.

Six-Dimensional Solvmanifolds with Holomorphically Trivial Canonical Bundle 23 Proposition 3.7. Let J be any complex structure on g 8 with closed volume (3, 0)-form. Then, there is a (1, 0)-basis {ω 1,ω 2,ω 3 } satisfying one of the following reduced equations: (g 8, J): dω 1 = 2iω 13 + ω 3 3, dω 2 = 2iω 23, dω 3 = 0; (22) Proof. (g 8, J ): dω 1 = 2iω 13 + ω 3 3, dω 2 = 2i ω 23 + ω 3 3, dω 3 = 0; (23) (g 8, J A ): dω 1 = (A i)ω 13 (A + i)ω 1 3, dω 2 = (A i)ω 23 + (A + i)ω 2 3, dω 3 = 0, where A C with Im A =0. (24) Moreover, the complex structures above are nonisomorphic. With respect to the structure equations of g 8 given in Theorem 2.8, any closed 3-form ρ Z 3 (g 8 ) is given by ρ = a 1 e 126 + a 2 e 135 + a 3 e 145 + a 4 e 156 + a 5 e 235 + a 6 (e 146 + e 236 ) + a 7 e 245 + a 8 (e 136 e 246 ) + a 9 e 256 + a 10 e 346 + a 11 e 356 + a 12 e 456, where a 1,...,a 12 R. A direct calculation shows that such a ρ satisfies the conditions d(j ρ ρ) = 0andλ(ρ) < 0ifandonlyifa 1 = 0, a 2 = a 7, a 3 = a 5, a 10 = 0, and a 6 a 7 a 5 a 8 0. Moreover, in this case λ(ρ) = 4(a 6 a 7 a 5 a 8 ) 2. {e 1,...,e 6 } as The associated complex structures J ρ J ρ e1 = e 2 + a 5a 12 a 7 a 11 a 6 a 7 a 5 a 8 e 5 + a 6a 12 a 8 a 11 a 6 a 7 a 5 a 8 e 6, J ρ e2 = e 1 + a 5a 11 + a 7 a 12 a 6 a 7 a 5 a 8 e 5 + a 6a 11 + a 8 a 12 a 6 a 7 a 5 a 8 e 6, J ρ e3 = e 4 + a 4a 7 a 5 a 9 a 6 a 7 a 5 a 8 e 5 + a 4a 8 a 6 a 9 a 6 a 7 a 5 a 8 e 6, express in terms of the real basis J ρ e4 = e 3 a 4a 5 + a 7 a 9 a 6 a 7 a 5 a 8 e 5 a 4a 6 + a 8 a 9 a 6 a 7 a 5 a 8 e 6, J ρ e5 = a 5a 6 + a 7 a 8 a 6 a 7 a 5 a 8 e 5 + a2 6 + a2 8 a 6 a 7 a 5 a 8 e 6, J ρ e6 = a2 5 + a2 7 a 6 a 7 a 5 a 8 e 5 a 5a 6 + a 7 a 8 a 6 a 7 a 5 a 8 e 6.

24 A. Fino et al. Let us consider the basis of (1,0)-forms {ω 1,ω 2,ω 3 } given by ω 1 = e 1 ij ρ e1 = e 1 i(e 2 + k 1 e 5 + k 2 e 6 ), ω 2 = e 3 ijρ e3 = e 3 i(e 4 + k 3 e 5 + k 4 e 6 ), ω 3 = 1 2c (e5 ijρ e5 ) = 1 ( b 2c e5 i 2c e5 + 1 ) 2 e6, where k 1 = a 5a 12 a 7 a 11 a 6 a 7 a 5 a 8, k 2 = a 6a 12 a 8 a 11 a 6 a 7 a 5 a 8, k 3 = a 4a 7 a 5 a 9 a 6 a 7 a 5 a 8, k 4 = a 4a 8 a 6 a 9 a 6 a 7 a 5 a 8, b = a 5a 6 +a 7 a 8 a 6 a 7 a 5 a 8, and c = a6 2+a2 8 a 6 a 7 a 5 a 8.Notethatc =0, and 2(a 6 + ia 8 )ω 123 = ρ + ijρ ρ. With respect to this basis, the complex structure equations are dω 1 = (A i)ω 13 (A + i)ω 1 3 + B ω 3 3, dω 2 = (A i)ω 23 + (A + i)ω 2 3 + C ω 3 3, (25) dω 3 = 0, where A= b + ic, B = 2c(k 1 + ik 2 ) and C = 2c(k 3 + ik 4 ).NotethatIm A= c =0. Now, we will reduce the complex Equations (25) as follows: If A = i, then with respect to the (1,0)-basis {η 1,η 2,η 3 } given by η 1 = (A + i)ω 1 + Bω 3, η 2 = (A + i)ω 2 + C ω 3, η 3 = ω 3, the complex structure equations are of the form (24). If A= i, Equations (25) reduce to J (B,C ) : dω 1 = 2iω 13 + Bω 3 3, dω 2 = 2iω 23 + C ω 3 3, dω 3 = 0. Note that the structures J (B,C ) and J (C,B) are equivalent, since it suffices to consider the change of basis η 1 = ω 2, η 2 = ω 1, η 3 = ω 3.Now: if B = C = 0, then the complex equations are of the form (24) with A= i; if only one of the coefficients B, C is nonzero, for instance B, then taking 1 B ω1 instead of ω 1, we arrive at the complex Equations (22) and finally, if B, C =0, then we can normalize both coefficients and the corresponding complex equations are (23). It is straightforward to check that the complex structures given in Equations (22) (24) are nonisomorphic.

Six-Dimensional Solvmanifolds with Holomorphically Trivial Canonical Bundle 25 Remark 3.8. Note that on g 8 there exists a unique complex structure J that is abelian [2], that is, satisfying [JX, JY] = [X, Y], which corresponds to the value A= i in Equations (24), and a unique bi-invariant complex structure [22], corresponding to A = i in (24). Finally, let us consider now the Lie algebra g 9. With respect to the structure equations given in Theorem 2.8, any closed 3-form ρ Z 3 (g 9 ) is given by ρ = a 1 (e 124 e 135 ) + a 2 e 145 + a 3 e 146 + a 4 e 156 + a 5 (e 136 e 245 ) + a 6 (e 125 + e 134 e 246 ) + a 7 e 256 + a 8 (e 126 + e 345 ) + a 9 e 346 + a 10 (e 125 + e 134 + e 356 ) + a 11 e 456. By imposing the closedness of Jρ ρ together with the condition tr(j 2 ρ )<0, one can arrive after a long computation to an explicit description of the complex structure J ρ,which allows us to prove that {e 2, e 4, e 6, J ρ e2, J ρ e4, J ρ e6 } are always linearly independent. Therefore, the forms ω 1 = e 6 ij ρ e6, ω 2 = e 2 ij ρ e2, ω 3 = e 4 ij ρ e4, constitute a (1,0)-basis for the complex structure J ρ, and one can show that with respect to this basis the complex structure equations have the form dω 1 = c 2 ω 1 1 cω 3 1 cω 1 3 ω 3 3, dω 2 = ( c 2 + ce i 2 G) ω 1 1 i ω2 1 + Eω 3 1 + ( 1 2 2 + cg) ω 1 3 +Gω 3 3 + i 2 ω12 + (cg E)ω 13, dω 3 = c ( c 2 + 2) i ω 1 1 + ( c 2 + 2) i ω 3 1 + c 2 ω 1 3 + cω 3 3 i 2 ω13, where c is real and E, G C. (Explicit expression of each coefficient in terms of a 1,...,a 11 can be given, but this information will not be relevant in what follows and so we omit it.) In the following result, we prove that all the complex structures are equivalent. Proposition 3.9. Up to isomorphism, there is only one complex structure with closed (3, 0)-form on the Lie algebra g 9, whose complex equations are (26) (g 9, J): dω 1 = ω 3 3, dω 2 = i 2 ω12 + 1 ω1 3 i ω2 1, dω 3 = i 2 2 2 ω13 + i ω3 1. (27) 2 Proof. First, note that one can suppose that G = 0bytakingω 2 + Gω 1 instead of ω 2 in Equations (26). Now, let {ω 1,ω 2,ω 3 } be a (1,0)-basis satisfying (26) withg = 0, and