Complex Manifolds 2018; 5: 103–110

Complex Manifolds

Research Article

Hiroshi Sawai* Examples of without LCK structures https://doi.org/10.1515/coma-2018-0005 Received September 28, 2017; accepted March 6, 2018. Abstract: The purpose in this paper is to construct solvmanifolds without LCK structures such that the complex structure is left-invariant.

Keywords: , Locally conformal Kähler structure, Vaisman structure

MSC: 53C55, 17B30

Introduction

Let G be a simply-connected solvable . A discrete co-compact Γ of G is said to be a in G. We call the compact manifold Γ ƒG a solvmanifold. If G is a nilpotent Lie group, we call the compact manifold Γ ƒG a nilmanifold. It is well known that a Kähler solvmanifold is a nite quotient of a complex torus which has a structure of a complex torus bundle over a complex torus ([2], [8]). Let (M, g, J) be a 2n-dimensional compact Hermitian manifold. We denote by Ω the fundamental 2- form, that is, the 2-form dened by Ω(X, Y) = g(X, JY). A Hermitian manifold (M, g, J) is said to be locally conformal Kähler (LCK) if there exists a closed 1-form ω such that dΩ = ω ∧ Ω. The closed 1-form ω is called the Lee form. Note that if ω = df, then (M, e−f g, J) is Kähler. A LCK manifold (M, g, J) is said to be a Vaisman manifold if the Lee form ω is parallel with respect to the Levi-Civita connection ∇ of the metric g. In this paper, we consider LCK structures and Vaisman structures on solvmanifolds with a left-invariant complex structure. The main non-Kähler examples of LCK manifolds are Hopf manifold [24], Inoue surfaces [22], Kodaira- Thurston manifold [5] and Oeljeklaus-Toma manifold [15] (cf. [6]). Note that Hopf manifold and Kodaira- Thurston manifold are Vaisman manifolds, and Inoue surfaces, Kodaira-Thurston manifold and Oeljeklaus- Toma manifold have a structure of a solvmanifold. The author of [18] proved that, in the case of a nilmanifold Γ ƒG with a left-invariant complex structure J, (Γ ƒG, J) has a LCK structure if and only if G = R × H(n), where H(n) is a (2n + 1)-dimensional Heisenberg Lie group. Thus, a LCK structure on a nilmanifold with a left-invariant complex structure is a Vaisman structure. Bazzoni [3] proved that a Vaisman nilmanifold is a quotient of G = R × H(n) even if J is not left-invariant. Kasuya [12] proved that Oeljeklaus-Toma manifold has no Vaisman structures. On the other hand, not much is known about LCK structures and Vaisman structures on general solvmanifolds.

In this paper, we consider a (2n + 2)-dimensional solvable Lie group Gn given by

xi Gn = œŒt, Œ ‘ , z‘ ∶ t, xi , yi , z ∈ R, i = 1, ⋯, n¡ , yi

*Corresponding Author: Hiroshi Sawai: National Institute of Technology, Numazu College, 3600 Ooka, Numazu-shi, Shizuoka 410-8501, Japan, E-mail: [email protected]

Open Access. © 2018 Sawai, published by De Gruyter. This work is licensed under the Creative Commons Attribution- NonCommercial-NoDerivs 4.0 License. 104 Ë H. Sawai

where the structure group on Gn is dened by

x x′ eai t x′ + x 1 n Œt, Œ i‘ , z‘ ⋅ Œt′, Œ i ‘ , z′‘ = Œt + t′, Œ i i ‘ , z′ + Q(−eai t y x′ + e−ai t x y′) + z‘ ′ −ai t ′ i i i i yi yi e yi + yi 2 i=1 and ai ∈ Z − {0}. In section 1, we construct a lattice Γn in Gn (cf. [17]). In the case of n = 1, the 4-dimensional solvmanifold Γ1ƒG1 has a LCK structure with non-parallel Lee form [1], and it is biholomorphic to Inoue surface S+ [10]. Belgun [4] considered the universal covering space of a Vaisman surface and proved that Inoue surface S+ has no Vaisman structures. In section 3, we reconsider LCK structures and Vaisman structures on Inoue surface S+ of the form Γ ƒG, and have the same result by a dierent proof:

Main Theorem 1. The solvmanifold Γ1ƒG1 has no Vaisman structures.

Moreover, we determine a LCK structure on the solvmanifold Γ1ƒG1. In the case of n ≥ 2, we prove

Main Theorem 2. In the case of n ≥ 2, let Jn be a left-invariant complex structure on the solvmanifold ΓnƒGn. Then the solvmanifold (ΓnƒGn , Jn) has no LCK structures.

A non-degenerate 2-form Ω is said to be locally conformal symplectic (LCS) if there exists a closed 1-form ω such that dΩ = ω ∧ Ω. Note that the solvmanifold ΓnƒGn has LCS structures.

1 Construction of solvmanifolds

In this section, we construct a lattice Γn in Gn (cf. [17]). Let B be a unimodular matrix with distinct positive eigenvalues λ, λ−1 given by

0 −1 B = Œ ‘ ∈ SL(2, ), 1 k Z

2 1 λ where k > 2. The characteristic polynomial of B is given by f (x) = x − kx + 1. Let P = Œ −1‘ ∈ GL(2, R). 1 λ λ 0 Then we see that SPS = (λ−1 − λ) and PBP−1 = Œ ‘. 0 λ−1 2n+2 We dene a dieomorphism ϕ ∶ R → Gn by

x ϕ(A) = Œ(log λ)t, 2P Œ i‘ , SPS z‘ yi

2n+2 2n+2 for A = (t, xi , yi , z) ∈ R . Dene Γn = ϕ(Z ). We prove that Γn is a lattice in Gn. It is obvious that Γn is discrete and ϕ(U) is compact, where U = {(t, xi , yi , z) ∶ 0 ≤ t, xi , yi , z ≤ 1}. Thus, it is sucient to prove that Γn is a subgroup of Gn. ′ ′ ′ ′ ′ ′ Let γ, γ ∈ Γn such that γ = ϕ((s, ui , vi , w)), γ = ϕ((s , ui , vi , w )). Put

u (γ) u s(γ) = (log λ)s, Œ i ‘ = 2P Œ i‘ , w(γ) = SPS w. vi(γ) vi

Then we get s(γγ′) = log λ(s + s′). For each i, we see that

u (γγ′) eai (log λ)s u (γ′) + u (γ) Œ i ‘ = Œ i i ‘ ′ −ai (log λ)s ′ vi(γγ ) e vi(γ ) + vi(γ) Examples of solvmanifolds without LCK structures Ë 105

λai s 0 u (γ′) u (γ) = Œ ‘ Œ i ‘ + Œ i ‘ −ai s ′ 0 λ vi(γ ) vi(γ) ′ ′ ai s −1 ui ui ai s ui ui = PB P 2P Œ ′ ‘ + 2P Œ ‘ = 2P œB Œ ′ ‘ + Œ ‘¡ . vi vi vi vi

Moreover, since ai (log λ)s ′ −ai (log λ)s ′ −e vi(γ)ui(γ ) + e ui(γ)vi(γ ) ai s ′ −ai s ′ = −λ vi(γ)ui(γ ) + λ ui(γ)vi(γ ) λai s u (γ′) = (−v (γ), u (γ)) Œ i ‘ i i −ai s ′ λ vi(γ ) λai s 0 u (γ′) = (−v (γ), u (γ)) Œ ‘ Œ i ‘ i i −ai s ′ 0 λ vi(γ ) ′ −1 ai s −1 ui = 2(−ui − λ vi , ui + λvi)PB P 2P Œ ′ ‘ vi λ ′ −1 1 ai s ui = 4{ui(−1, 1) + vi(−λ , λ)} Œ −1‘ B Œ ′ ‘ 1 λ vi ′ ai s ui = 4{ui(0, SPS) + vi(−SPS, 0)}B Œ ′ ‘ vi ′ ai s ui = 4SPS(−vi , ui)B Œ ′ ‘ , vi we have

n ′ ′ 1 ai (log λ)s ′ −ai (log λ)s ′ ′ w(γγ ) = w(γ ) + Q(−e vi(γ)ui(γ ) + e ui(γ)vi(γ )) + w(γ ) 2 i=1 n ′ ′ ai s ui = SPS w + Q 2SPS(−vi , ui)B Œ ′ ‘ + SPS w i=1 vi n ′ ′ ai s ui = SPS œw + Q 2(−vi , ui)B Œ ′ ‘ + w¡ . i=1 vi

Then Γn is a subgroup of Gn. Thus we have

Proposition 1.1. Γn is a lattice in Gn.

2 LCK structures on solvmanifolds

In this section, we determine a form of the fundamental 2-form on a Vaisman manifold, and obtain a left- invariant LCK metric provided one begins with a left-invariant complex structure. Let M be a manifold and A⋆(M) be the de Rham complex of M with the exterior dierential operator d. p p+1 For a closed 1-form θ on M, we dene the new dierential operator dθ from A (M) to A (M) by

dθα = θ ∧ α + dα.

2 ∗ Since θ is closed, we see that dθ = 0. Then, we can dene the new cohomology group Hθ (M) in a natural manner. A p-form α is called θ-closed if dθα = 0. It is called θ-exact if there exists a (p − 1)-form β such that α = dθβ. Similarly, we can dene a new dierential operator and the new cohomology group on a Lie algebra. Let Ω be the fundamental 2-form of a LCK structure (g, J). There exists a closed 1-form ω such that dΩ = ω ∧ Ω, which implies that Ω is −ω-closed. Moreover, we have

Proposition 2.1 ([13]). Let (M, g) be a compact Riemannian manifold. If a closed 1-form θ is parallel with respect to the Levi-Civita connection ∇ of the metric g, then any θ-closed form is θ-exact. 106 Ë H. Sawai

From Proposition 2.1, the fundamental 2-from Ω of a Vaisman manifold is −ω-exact, that is, it is given by

Ω = d−ωη = −ω ∧ η + dη, where η is a 1-form. A solvable Lie group G is said to be completely solvable if a linear map ad(X) from g to g has only real eigenvalues for each X ∈ g, where g is the Lie algebra of G. We have

Theorem 2.2 ([9]). If a completely solvable Lie group G admits a lattice Γ , then

p p HDR(Γ ƒG) ≅ H (g) for each p, where g is the Lie algebra of G.

Let (M = Γ ƒG, g, J) be a LCK completely solvable solvmanifold with a left-invariant complex structure J, and g be the Lie algebra of G. From Theorem 2.2, there exists a left-invariant closed 1-form ω0 such that ω−ω0 = df. Then, for left-invariant vector elds X, Y, we dene an inner product ⟨ , ⟩ by

−f ⟨X, Y⟩ = S e g(X, Y)dµ, M where dµ is the volume element induced by a bi-invariant volume element on G. Since the complex structure

J is left-invariant, we get a left-invariant Hermitian structure (⟨ , ⟩, J). Let Ω0 be the fundamental 2-form of (⟨ , ⟩, J). We have

Proposition 2.3 ([4]). dΩ0 = ω0 ∧ Ω0.

Thus a LCK structure (g, J) with left-invariant J induces a left-invariant LCK structure (⟨ , ⟩, J) on g. We say that (g, ⟨ , ⟩, J) is a locally conformal Kähler (LCK) solvable Lie algebra. Moreover, we have

Proposition 2.4 ([19]). If Ω is −ω-exact, then Ω0 is also −ω0-exact, that is, it is given by

Ω0 = d−ω0 η0 = −ω0 ∧ η0 + dη0, where η0 is a 1-form on g.

Since the solvable Lie group Gn is completely solvable, we see that a LCK structure (g, Jn) on ΓnƒGn induces a left-invariant LCK structure (⟨ , ⟩, Jn), where Jn is a left-invariant complex structure.

3 Proof of Main Theorem 1

In this section, we consider LCK structures in the case of n = 1 and prove Main Theorem 1.

In the case of n = 1, the Lie algebra g1 of G1 is given by

g1 = span{A, X, Y, Z ∶ [A, X] = a1X, [A, Y] = −a1Y, [X, Y] = Z},

̂ 1 ′ where a1 ≠ 0. Put A = A. Then the Lie algebra g1 is isomorphic to g1 given by a1 ′ ̂ ̂ ̂ g1 = span{A, X, Y, Z ∶ [A, X] = X, [A, Y] = −Y, [X, Y] = Z}. ′ ̂ From now on, we write g1 as g1 and A as A, respectively. Let {θ, α, β, δ} be the dual basis of {A, X, Y, Z}:

dθ = 0, dα = −θ ∧ α, dβ = θ ∧ β, dδ = −α ∧ β.

1 Then we see that H (g1) = span{[θ]}. The 4-dimensional solvmanifold Γ1ƒG1 has a LCK structure with non-parallel Lee form as follows: Examples of solvmanifolds without LCK structures Ë 107

Example 3.1 ([1]). Let ⟨ , ⟩ be a left-invariant metric such that {A, X, Y, Z} is an orthonormal frame, and J be a left-invariant complex structure dened by

JA = Y, JY = −A, JZ = X, JX = −Z.

Then (Γ1ƒG1, ⟨ , ⟩, J) is a LCK manifold with the fundamental 2-form Ω given by

Ω = −θ ∧ β − δ ∧ α, and Lee form −θ. We easily see that

∇Y θ(Y) = −θ(∇Y Y) = −⟨A, ∇Y Y⟩ = −⟨[A, Y], Y⟩ = ⟨Y, Y⟩ ≠ 0.

Then Lee form −θ is not parallel.

Two complex structures J1 and J2 on a Lie algebra g are equivalent if there exists F ∈ Aut(g) such that F ○ J2 = J1 ○ F. If J1 and J2 are equivalent, then a LCK metric g2 with a complex structure J2 is one-to-one ∗ correspondence with a LCK metric g1 with a complex structure J1, by g2 = F g1 [23]. Proof of Main Theorem 1. A complex structure J on a 4-dimensional solvmanifold is left-invariant [7]. Then, a LCK structure (g, J) on Γ1ƒG1 induces a left-invariant LCK structure (⟨ , ⟩, J) as in section 2. Particularly, we assume that the LCK structure (g, J) is a Vaisman structure. From Proposition 2.1 and Proposition 2.4, the fundamental 2-form Ω of (⟨ , ⟩, J) is −kθ-exact (k ∈ R). A complex structure J on g1 is equivalent to a complex structure Jq such that

Jq A = Y + qZ, Jq Y = −A − qX, Jq Z = X, Jq X = −Z, where q = 0 or 1 [16]. Thus, the LCK structure (⟨ , ⟩, J) induces the LCK structure (⟨ , ⟩q , Jq) with the fundamental 2-form Ωq given by

Ωq = d−kθη = −kθ ∧ η + dη, ∗ because there exists F ∈ Aut(g1) such that Ωq = F Ω. However, we easily see that

⟨Z, Z⟩ = d−kθη(Jq Z, Z) = (−kθ ∧ η + dη)(X, Z) = 0 because X, Z ∈ [g1, g1] and Z is in the center of g1. This is a contradiction. We give some remarks:

n m Remark 3.2. Kasuya [12] proved that, in the case of G = R ⋉ϕ R , a solvmanifold Γ ƒG with some conditions has no Vaisman structures. Thus, Inoue surface S0 and Oeljeklaus-Toma manifold have no Vaisman structures. On the other hand, the solvable Lie group G1 is the form of R⋉ϕ H(1), where H(1) is a 3-dimensional Heisenberg Lie group.

Remark 3.3. It is known that the rst Betti number of a Vaisman manifold is odd ([11], [24]). From Theorem 2.2 1 1 [9], we see that dim HDR(Γ1ƒG1) = dim H (g1) = 1.

Next, we consider LCK structures on the solvable Lie algebra g1. From the argument in [14], the new 2 cohomology group Hkθ(g1) of g1 is given by

2 H−θ(g1) = span{[δ ∧ β]−θ , [θ ∧ β]−θ}, 2 Hθ(g1) = span{[δ ∧ α]θ , [θ ∧ α]θ}, 2 Hkθ(g1) = {0} for k ≠ ±1,

2 where []±θ is a cohomology class of a new cohomology group H±θ(g1) dened by d±θ, respectively. From the above argument, we see that g1 has no LCK structures such that the fundamental 2-form is exact. 108 Ë H. Sawai

A −θ-closed 2-form Ω− is given by

− − − − Ω = a δ ∧ β + b θ ∧ β + d−θη .

Then we see that − − − − Ω (Jq Z, Z) = Ω (X, Z) = (−θ ∧ η + dη )(X, Z) = 0

− because X, Z ∈ [g1, g1] and Z is in the center of g1. Thus, (Ω , Jq) is not a LCK structure for any q = 0 or 1. A θ-closed 2-form Ω+ is given by

+ + + + Ω = a δ ∧ α + b θ ∧ α + dθη .

+ If (Ω , Jq) is a LCK structure, then we see that

+ + + + Ω (Jq Z, Z) = Ω (X, Z) = (a δ ∧ α)(X, Z) = −a ≠ 0.

Moreover, we get

+ + + Ω (X, Y) = Ω (Jq X, Jq Y) = Ω (−Z, −A − qX) dη+(X, Y) = (a+δ ∧ α + θ ∧ η+)(−Z, −A − qX) −η+([X, Y]) = −η+(Z) = −η+(Z) + a+q 0 = a+q, that is, q = 0. + Let (Ω , J0) be a LCK structure with Lee form θ as above. Then we see that

+ + + Ω (A, X) = Ω (J0A, J0X) = Ω (Y, −Z) + + (b θ ∧ α + dθη )(A, X) = 0 b+ + η+(X) − η+(X) = 0,

+ + that is, b = 0. Then, we see that [Ω ]θ ∈ span{[δ ∧ α]θ}. This is Example 3.1. Thus, we have

+ Proposition 3.4. (Γ1ƒG1, Jq) has a LCK structure Ω if and only if q = 0, that is, Inoue surface S . Moreover, a LCK structure (Ω, J0) has a Lee form θ and [Ω]θ ∈ span{[δ ∧ α]θ}.

4 Proof of Main Theorem 2

In this section, we consider LCK structures in the case of n ≥ 2 and prove Main Theorem 2.

In the case of n ≥ 2, the Lie algebra gn of Gn is given by

gn = span{A, Xi , Yi , Z ∶ [A, Xi] = ai Xi , [A, Yi] = −ai Yi , [Xi , Yi] = Z}, where ai ≠ 0. Let {θ, αi , βi , δ} be the dual basis of {A, Xi , Yi , Z}:

n dθ = 0, dαi = −aiθ ∧ αi , dβi = aiθ ∧ βi , dδ = − Q αi ∧ βi . i=1

1 Then we see that H (gn) = span{[θ]}. We have

Lemma 4.1. Let Ω0 be a −kθ-closed 2-form on gn (k ∈ R). Then Ω0(Z, Xi) = Ω0(Z, Yi) = 0 for each i. Examples of solvmanifolds without LCK structures Ë 109

Proof. For each i, since n ≥ 2, we can take j such that j ≠ i. Then we get

Ω0(Z, Xi) = Ω0([Xj , Yj], Xi)

= −dΩ0(Xj , Yj , Xi) + Ω0([Xj , Xi], Yj) − Ω0([Yj , Xi], Xj)

= −kθ ∧ Ω0(Xj , Yj , Xi) = 0.

Similarly, we get Ω0(Z, Yi) = 0.

Proof of Main Theorem 2. Let Jn be a left-invariant complex structure on ΓnƒGn. We assume that the solvman- ifold (ΓnƒGn , Jn) has a LCK metric g. From section 2, a LCK structure (g, Jn) induces a left-invariant LCK structure (⟨ , ⟩, Jn). Let Ω0 be the fundamental 2-form of (⟨ , ⟩, Jn) and −kθ be Lee form. ∗ Let γ0 be the isomorphism from gn to gn induced by ⟨ , ⟩. Since [gn , gn] = span{Xi , Yi , Z} and θ is closed, we ⊥ see that [gn , gn] = span{γ0(θ)}. From Lemma 4.1, we get

Corollary 4.2. J ○ γ0(θ) ∈ span{Z}.

Proof. From Lemma 4.1, we see that

⟨JZ, Xi⟩ = −Ω0(Z, Xi) = 0 and ⟨JZ, Yi⟩ = −Ω0(Z, Yi) = 0 ⊥ for each i. Since [gn , gn] = span{Xi , Yi , Z}, we have JZ ∈ [gn , gn] = span{γ0(θ)}, that is, J ○γ0(θ) ∈ span{Z}.

Let (g, ⟨ , ⟩, J) be a LCK solvable Lie algebra as in section 2, and ω0 be the Lee form. In the previous papers [20], [21], we have

Theorem 4.3 ([20]). If ⟨[γ0(ω0), J ○ γ0(ω0)], J ○ γ0(ω0)⟩ = 0, then (⟨ , ⟩, J) is a Vaisman structure.

Theorem 4.4 ([21]). If g is completely solvable and (⟨ , ⟩, J) is a Vaisman structure, then g = R × h(n), where h(n) is a (2n + 1)-dimensional Heisenberg Lie algebra.

Since Z is in the center of gn, from Corollary 4.2, we see that (gn , ⟨ , ⟩, Jn) satises the assumption on Theorem 4.3. Then a LCK structure (⟨ , ⟩, Jn) on gn is a Vaisman structure. However, gn is completely solvable and it is the form of R ⋉ϕ h(n), and the action ϕ is non-trivial. This is contradiction to Theorem 4.4. Thus, (gn , Jn) has no LCK structures. We give some remarks:

Remark 4.5. A non-degenerate −kθ-closed 2-form is said to be locally conformal symplectic (LCS). The solvmanifold ΓnƒGn has LCS structures Ω1, Ω2 dened as follows: n Ω1 = d−θδ = −θ ∧ δ − Q αi ∧ βi , i=1 n = ∧ + = ∧ + ( + ) ∧ − ( + ) ∧ Ω2 αk αl d(ak +al )θδ αk αl ak al θ δ ak al Q αi βi , i=1 where ak + al ≠ 0. Note that Ω1 is exact and Ω2 is non-exact.

1 Remark 4.6. If ai = 0 for each i, then the solvmanifold ΓnƒGn is a nilmanifold S × Γ ƒH(n), where H(n) is a (2n + 1)-dimensional Heisenberg Lie group and Γ is its lattice. A nilmanifold S1 × Γ ƒH(n) has a Vaisman structure [5].

Remark 4.7. From Theorem 4.4, we can easily prove Main Theorem 1.

Acknowledgement: The author would like to express his deep appreciation to Professor Takumi Yamada for their thoughtful guidance and encouragement during the completion of this paper. The author also thanks the referee for many suggestions. This work was supported by JSPS KAKEN Grant number JP17K05235. 110 Ë H. Sawai

References

[1] L. C. de Andrés, L. A. Cordero, M. Fernández and J. J. Mencía : Examples of four dimensional locally conformal Kähler solvmanifolds, Geom. Dedicata 29 (1989), 227-232. [2] D. Arapura and M. Nori: Solvable fundamental groups of algebraic varieties and Kähler manifolds, Compositio Math, 116 (1999), 173-188. [3] G. Bazzoni: Vaisman nilmanifolds, Bull. Lond. Math. Soc. 49 (2017), no. 5, 824-830. [4] F. A. Belgun: On the metric structure of non-Kähler complex surfaces, Math. Ann. 317 (2000), 1-40. [5] L. A. Cordero, M. Fernández and M. de Léon: Compact locally conformal Kähler nilmanifolds, Geom. Dedicata 21 (1986), 187-192. [6] S. Dragomir and L. Ornea: Locally conformal Kähler geometry, Birkhäuser (1998). [7] K. Hasegawa: Complex and Kähler structures on compact solvmanifolds, J. Symplectic Geom. 3 (2005), 749-767. [8] K. Hasegawa: A note on compact solvmanifolds with Kähler structures, Osaka J. Math. 43 (2006), no. 1, 131-135. [9] A. Hattori : Spectral sequence in the de Rham cohomology of bre bundles, J. Fac. Sci. Univ. Tokyo Sect. I . 8 (1960), 289-331. [10] Y. Kamishima: Note on locally conformal Kähler surfaces, Geom. Dedicata 84 (2001), no. 1-3, 115-124. [11] T. Kashiwada and S. Sato: On harmonic forms in compact locally conformal Kähler manifolds with the parallel Lee form, Ann. Fac. Sci. Univ. Nat. Zaire (Kinshasa) Sect. Math.-Phys. 6 (1980), 17-29 [12] H. Kasuya : Vaisman metrics on solvmanifolds and Oeljeklaus-Toma manifolds, Bull. Lond. Math. Soc. 45 (2013), no. 1, 15-26. [13] M. de León, B. López, J. C. Marrero and E. Padrón: On the computation of the Lichnerowicz-Jacobi cohomology, J. Geom. Phys. 44 (2003), no. 4, 507-522. [14] D. V. Millionschikov: Cohomology of solvable Lie algebras and solvmanifolds, Mat. Zametki 77 (2005), no. 1, 67-79, [15] K. Oeljeklaus and M. Toma: Non-Kähler compact complex manifolds associated to number elds, Ann. Inst. Fourier (Grenoble) 55 (2005) 161-171. [16] G. Ovando: Complex, symplectic and Kähler structures on four dimensional Lie groups, Rev. Un. Mat. Argentina 45 (2004), no. 2, 55-67. [17] H. Sawai: A construction of lattices on certain solvable Lie groups, Topology Appl. 154 (2007), no. 18, 3125-3134. [18] H. Sawai: Locally conformal Kähler structures on compact nilmanifolds with left-invariant complex structures, Geom. Dedicata 125 (2007), 93-101. [19] H. Sawai: Locally conformal Kähler structures on compact solvmanifolds, Osaka J. Math. 49 (2012), no. 4, 1087-1102. [20] H. Sawai: Vaisman structures on compact solvmanifolds, Geom. Dedicata 178 (2015), 389-404. [21] H. Sawai: Structure theorem for Vaisman completely solvable solvmanifolds, J. Geom. Phys. 114 (2017), 581-586. [22] F. Tricerri: Some examples of locally conformal Kähler manifolds, Rend. Sem. Math. Univ. Politec. Torino 40 (1982), no. 1, 81-92. [23] L. Ugarte: Hermitian structures on six-dimensional nilmanifolds, Transform. Groups 12 (2007), no. 1, 175-202. [24] I. Vaisman: Generalized Hopf manifolds, Geom. Dedicata 13 (1982), no. 3, 231-255.