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

Levi-Civita Ricci-flat metrics on non-Kähler Calabi-Yau manifolds

Eder M. Correa UFMG, Avenida Antônio Carlos, 6627, 31270-901 Belo Horizonte - MG, Brazil E-mail: [email protected].
Abstract.

In this paper, we provide new examples of Levi-Civita Ricci-flat Hermitian metrics on certain compact non-Kähler Calabi-Yau manifolds, including every compact Hermitian Weyl-Einstein manifold, every compact locally conformal hyperKähler manifold, certain suspensions of Brieskorn manifolds, and every generalized Hopf manifold provided by suspensions of exotic spheres. These examples generalize previous constructions on Hopf manifolds. Additionally, we also construct new examples of compact Hermitian manifolds with nonnegative first Chern class that admit constant strictly negative Riemannian scalar curvature. Further, we remark some applications of our main results in the study of the Chern-Ricci flow on compact Hermitian Weyl-Einstein manifolds. In particular, we describe the Gromov-Hausdorff limit for certain explicit finite-time collapsing solutions which generalize previous constructions on Hopf manifolds.

1. Introduction

Given a compact Hermitian manifold (M,g,J)(M,g,J), with fundamental 22-form Ω\Omega, as shown in [28, Theorem 1.2], the first Levi-Civita Ricci curvature (1)(Ω)\mathfrak{R}^{(1)}(\Omega) of (M,g,J)(M,g,J) represents its first Aeppli-Chern class111See [28, Definition 1.1]. For more details on Aeppli cohomology, see [2] and references therein. c1AC(M)HA1,1(M)c_{1}^{AC}(M)\in H^{1,1}_{A}(M). More precisely,

(1)(Ω)=Ric(1)(Ω)12(Ω+¯¯Ω),\mathfrak{R}^{(1)}(\Omega)={\rm{Ric}}^{(1)}(\Omega)-\frac{1}{2}\big{(}\partial\partial^{\ast}\Omega+\bar{\partial}\bar{\partial}^{\ast}\Omega\big{)}, (1.1)

where Ric(1)(Ω)=loc12ddclog(det(Ω)){\rm{Ric}}^{(1)}(\Omega)\mathrel{\stackrel{{\scriptstyle\makebox[0.0pt]{\mbox{\tiny loc}}}}{{=}}}-\frac{1}{2}{\rm{d}}{\rm{d}}^{c}\log(\det(\Omega)) is the associated (first) Chern Ricci curvature. By the celebrated Calabi-Yau theorem [40] a compact Kähler manifold has c1(M)=0c_{1}(M)=0 if and only if it has a Ricci-flat Kähler metric, i.e. Ric(g)=0{\rm{Ric}}(g)=0. In the compact non-Kähler setting, since (1)(Ω)=0\mathfrak{R}^{(1)}(\Omega)=0 implies c1AC(M)=0c_{1}^{AC}(M)=0, a natural question to ask inspired by the Calabi-Yau theorem is the following:

Problem 1.1 ([28]).

On a compact complex manifold MM, if c1AC(M)=0c_{1}^{AC}(M)=0, does there exist a smooth Levi-Civita Ricci-flat Hermitian metric Ω\Omega, i.e. such that (1)(Ω)=0\mathfrak{R}^{(1)}(\Omega)=0?

From Eq. (1.1), the Levi-Civita Ricci-Flat condition (1)(Ω)=0\mathfrak{R}^{(1)}(\Omega)=0 is equivalent to

Ric(1)(Ω)=12(Ω+¯¯Ω).{\rm{Ric}}^{(1)}(\Omega)=\frac{1}{2}\big{(}\partial\partial^{\ast}\Omega+\bar{\partial}\bar{\partial}^{\ast}\Omega\big{)}. (1.2)

Since there are non-elliptic terms on the right-hand side of Eq. (1.2), it is not of Monge-Ampère type. Thus, it is particularly challenging to solve such equations. Since c1AC(M)=0c_{1}^{AC}(M)=0, if c1(M)=0c_{1}(M)=0, it is very natural to study the Problem 1.1 on non-Kähler Calabi-Yau manifolds, i.e. compact non-Kähler manifolds satisfying c1(M)=0c_{1}(M)=0. It is known that the Hopf manifold M=S2n+1×S1M=S^{2n+1}\times S^{1} is non-Kähler Calabi-Yau. Moreover, since the first Bott-Chern class of MM does not vanish, i.e. c1BC(M)0c_{1}^{BC}(M)\neq 0, there does not exist a Hermitian metric on MM, such that Ric(1)(Ω)=0{\rm{Ric}}^{(1)}(\Omega)=0. On the other hand, as it was shown in [28], every Hopf manifold M=S2n+1×S1M=S^{2n+1}\times S^{1} admits a Hermitian metric Ω\Omega satisfying (1)(Ω)=0\mathfrak{R}^{(1)}(\Omega)=0. Inspired by this example and by Problem 1.1, in this paper we generalize the construction provided in [28] on Hopf manifolds to a more general setting of locally conformal Kähler manifolds obtained as certain suspensions of Sasaki-Einstein manifolds. More precisely, we prove the following:

Theorem A.

Let (Q,gSE)(Q,g_{SE}) be a compact Sasaki-Einstein manifold, ϕ:QQ\phi\colon Q\to Q a Sasaki automorphism, and c>0c>0. Then, the suspension Σϕ,c(Q)\Sigma_{\phi,c}(Q) by (ϕ,c)(\phi,c) of QQ admits a Levi-Civita Ricci-flat Hermitian metric.

Let us observe that by a suspension of QQ by (ϕ,c)(\phi,c) we mean

Σϕ,c(Q):=Q×[0,log(c)](ϕ(x),0)(x,log(c)).\Sigma_{\phi,c}(Q):=\frac{Q\times[0,\log(c)]}{(\phi(x),0)\sim(x,\log(c))}. (1.3)

In particular, since b1(Σϕ,c(Q))=1b_{1}(\Sigma_{\phi,c}(Q))=1 (e.g. [38]), it follows that Σϕ,c(Q)\Sigma_{\phi,c}(Q) cannot be Kähler. From Theorem A, one can recover the Levi-Civita Ricci-flat Hermitian metric on Hopf manifolds provided in [28] just by taking Q=S2n+1Q=S^{2n+1} and ϕ=id\phi={\rm{id}}. The main idea to prove Theorem A is the following: Considering the locally conformal Kähler metric Ω\Omega on Σϕ,c(Q)\Sigma_{\phi,c}(Q) induced from the Calabi-Yau metric of the metric cone Q×+Q\times\mathbbm{R}_{+}, e.g. [23], one can write explicitly the Levi-Civita Ricci-Flat condition for the perturbed Hermitian metric

Ωζ:=Ω+2ζ(1)(Ω),\Omega_{\zeta}:=\Omega+2\zeta\mathfrak{R}^{(1)}(\Omega), (1.4)

such that ζ>1\zeta>-1. Then, we show that one can always find ζ\zeta, such that (1)(Ωζ)=0\mathfrak{R}^{(1)}(\Omega_{\zeta})=0. The key point which allows us to solve Eq. (1.2) is that (Σϕ,c(Q),Ω)(\Sigma_{\phi,c}(Q),\Omega) is a compact Hermitian Weyl-Einstein manifold. This fact allows us to perform all computations needed using differential forms and to reduce the problem to a simple equation in terms of ζ\zeta just like in the case Hopf manifolds (cf. [28, Theorem 6.2]). The framework on Sasaki-Einstein geometry of Theorem A allows us to obtain several examples of Levi-Civita Ricci-flat manifolds. In fact, since every 3-Sasakian manifold is, in particular, Sasaki-Einstein, see for instance [8], [26], we obtain the following result.

Corollary A.

Let (Q,gQ)(Q,g_{Q}) be a compact 33-Sasakian manifold, ϕ:QQ\phi\colon Q\to Q a Sasaki automorphism, and c>0c>0. Then, the suspension Σϕ,c(Q)\Sigma_{\phi,c}(Q) by (ϕ,c)(\phi,c) of QQ admits a Levi-Civita Ricci-flat Hermitian metric.

Also, from [38] and Theorem A, we have a positive answer for the question proposed in Problem 1.1 in the case that MM is a compact Hermitian Weyl-Einstein manifold. In fact, we obtain the following.

Corollary B.

Every compact Hermitian Weyl-Einstein manifold admits a Levi-Civita Ricci-flat Hermitian metric. In particular, every compact locally conformal hyperKähler manifold admits a Levi-Civita Ricci-flat Hermitian metric.

Our next result is an application of Theorem A in the setting of quasi-regular Sasaki-Einstein manifolds provided by links of hypersurfaces singularities. More precisely, from the result provided in [29, Theorem 1.4], see also [7, Conjecture 4], and Theorem A, we have the following corollary.

Corollary C.

Let L(𝐚):=Y(𝐚)S2n+1{\rm{L}}({\bf{a}}):=Y({\bf{a}})\cap S^{2n+1} be the link of a Brieskorn-Pham singularity

Y(𝐚):=(z0a0++znan=0)n+1,Y({\bf{a}}):=\Big{(}z_{0}^{a_{0}}+\cdots+z_{n}^{a_{n}}=0\Big{)}\subset\mathbbm{C}^{n+1}, (1.5)

such that n3n\geq 3. Assume that a0ana_{0}\leq\cdots\leq a_{n}. Then L(𝐚)×S1{\rm{L}}({\bf{a}})\times S^{1} admits a Levi-Civita Ricci-flat Hermitian metric if

1<j=0n1aj<1+nan.1<\sum_{j=0}^{n}\frac{1}{a_{j}}<1+\frac{n}{a_{n}}. (1.6)

In particular, from Theorem A and [29], one can obtain Levi-Civita Ricci-flat Hermitian metrics on generalized Hopf manifolds [11]. In fact, we obtain the following corollary which generalizes [29, Theorem 1.5].

Corollary D.

Let 𝚺{\bf{\Sigma}} be an odd dimensional homotopy sphere which bounds a parallelizable manifold. Then 𝚺×S1{\bf{\Sigma}}\times S^{1} admits a Levi-Civita Ricci-flat Hermitian metric.

Given two homotopy spheres 𝚺{\bf{\Sigma}} and 𝚺{\bf{\Sigma}}^{\prime} of dimension 5\geq 5, then we have that 𝚺×S1{\bf{\Sigma}}\times S^{1} is diffeomorphic to 𝚺×S1{\bf{\Sigma}}^{\prime}\times S^{1} if and only if 𝚺{\bf{\Sigma}} is diffeomorphic to 𝚺{\bf{\Sigma}}^{\prime}, see for instance [11, Proposition 3]. Therefore, in the setting of Corollary D the underlying differentiable manifolds 𝚺×S1{\bf{\Sigma}}\times S^{1} are exotic. Additionally, inspired by [29, Theorem 6.4], we prove the following result.

Theorem B.

Let (Q,gSE)(Q,g_{SE}) be a compact Sasaki-Einstein manifold, ϕ:QQ\phi\colon Q\to Q a Sasaki automorphism, and c>0c>0. Then, the suspension Σϕ,c(Q)\Sigma_{\phi,c}(Q) by (ϕ,c)(\phi,c) of QQ admits three different Hermitian metrics Ωi\Omega_{i}, i=1,2,3i=1,2,3, satisfying the following properties:

  1. (1)

    Ric(1)(Ω1)=Ric(1)(Ω2)=Ric(1)(Ω3)0{\rm{Ric}}^{(1)}(\Omega_{1})={\rm{Ric}}^{(1)}(\Omega_{2})={\rm{Ric}}^{(1)}(\Omega_{3})\geq 0;

  2. (2)

    Ω1\Omega_{1} has strictly positive Riemannian scalar curvature;

  3. (3)

    Ω2\Omega_{2} has zero Riemannian scalar curvature;

  4. (4)

    Ω3\Omega_{3} has strictly negative Riemannian scalar curvature.

In particular, all compact Hermitian manifolds of the previous corollaries admit three different Hermitian metrics satisfying the above properties.

As it can be seen, Theorem B generalizes some ideas introduced in [29]. In particular, it provides a huge class of examples of compact Hermitian manifolds with nonnegative first Chern class which admit Hermitian metrics with strictly negative Riemannian scalar curvature. This fact contrasts with the setting of compact Kähler manifolds with nonnegative first Chern class. Further, we also make some remarks related to applications of the ideas used in the proof of Theorem A and Theorem B in order to construct explicit solutions of the Chern-Ricci flow (e.g. [21], [34]) on compact Hermitian Weyl-Einstein manifolds. More precisely, we prove the following.

Theorem C.

Let (M,g,J)(M,g,J) be a compact Hermitian Weyl-Einstein manifold, then there exists an explicit solution g(t)g(t) of the Chern-Ricci flow on MM for t[0,2n)t\in[0,\frac{2}{n}), starting at gg, satisfying the following properties:

  1. (1)

    Vol(M,g(t))0{\rm{Vol}}(M,g(t))\to 0 as t2nt\to\frac{2}{n} (i.e. g(t)g(t) is finite-time collapsing);

  2. (2)

    limt2ng(t)=hT\lim_{t\to\frac{2}{n}}g(t)=h_{T}, where hTh_{T} is a nonnegative symmetric tensor on MM;

  3. (3)

    The Chern scalar curvature of g(t)g(t) blows up like (n1)/(2nt)(n-1)/(\frac{2}{n}-t);

  4. (4)

    scal(g(t)){\rm{scal}}(g(t))\to-\infty as t2nt\to\frac{2}{n};

  5. (5)

    limt2ndGH((M,dt),(S1,dS1))=0\lim_{t\to\frac{2}{n}}d_{GH}\big{(}(M,d_{t}),(S^{1},d_{S^{1}})\big{)}=0,

where dtd_{t} is the distance induced by g(t)g(t) on MM and dS1d_{S^{1}} is the distance on the unit circle S1S^{1} induced by a suitable scalar multiple of the standard Riemannian metric.

As it can be seen, the results of Theorem C generalize some previous results provided in [34], [22], and [35], for Hopf manifolds S2n+1×S1S^{2n+1}\times S^{1}. The proof which we present for Theorem C takes into account the framework on Sasaki-Einstein geometry which underlies compact Hermitian Weyl-Einstein manifolds. Therefore, the results of Theorem C hold for any suspension of a Sasaki-Einstein manifold as in the previous corollaries and theorems.

Outline of the paper

This paper is organized as follows. In Section 2, we present some generalities on Hermitian geometry, Sasaki geometry, and Hermitian Weyl-Einstein geometry. In Section 3, we prove Theorem A, Theorem B, and we provide a huge class of examples which illustrate our results. In Appendix A, we prove Theorem C and we illustrate the result by means of an explicit example.

2. Preliminary results

In this section, we review some basic definitions and results related to Hermitian geometry [17], [18], [39], [28], Calabi-Yau cones [32], [6], and Hermitian Weyl-Einstein geometry [36], [15], [18], [19], [38].

2.1. Generalities on Hermitian manifolds

Let (M,g,J)(M,g,J) be a Hermitian manifold of complex dimension nn. Denoting by Ω:=g(J𝟙)\Omega:=g(J\otimes\mathbbm{1}) the associated fundamental 22-form, consider

(dΩ)0:=dΩ1n1L(Λ(dΩ)),({\rm{d}}\Omega)_{0}:={\rm{d}}\Omega-\frac{1}{n-1}{\rm{L}}(\Lambda({\rm{d}}\Omega)), (2.1)

where L=Ω(){\rm{L}}=\Omega\wedge(-) is the Lefschetz operator and Λ\Lambda is its adjoint [39]. Since Λ(dΩ)\Lambda({\rm{d}}\Omega) is primitive, it follows that Λ(L(Λ(dΩ)))=(n1)Λ(dΩ)\Lambda({\rm{L}}(\Lambda({\rm{d}}\Omega)))=(n-1)\Lambda({\rm{d}}\Omega), thus Λ((dΩ)0)=0\Lambda(({\rm{d}}\Omega)_{0})=0, i.e., the primitive part of dΩ{\rm{d}}\Omega is given by Eq. (2.1). In the above setting, we have the following definition.

Definition 2.1.

The Lee 11-form of a Hermitian manifold (M,Ω,J)(M,\Omega,J) of complex dimension nn is defined by

θ:=1n1Λ(dΩ).\theta:=\frac{1}{n-1}\Lambda({\rm{d}}\Omega). (2.2)
Remark 2.2.

The Lee form θ\theta associated to a Hermitian manifold plays an important role in the study of the classification of Hermitian structures, e.g. [24].

Now we consider the following result.

Lemma 2.3.

Let (M,Ω,J)(M,\Omega,J) be a Hermitian manifold of complex dimension nn. Then

d(Ωn1)=(n1)θΩn1.{\rm{d}}(\Omega^{n-1})=(n-1)\theta\wedge\Omega^{n-1}. (2.3)
Remark 2.4.

The above lemma follows directly from the fact that

Λ((dΩ)0)=0(dΩ)0Ωn2=0,\Lambda(({\rm{d}}\Omega)_{0})=0\iff({\rm{d}}\Omega)_{0}\wedge\Omega^{n-2}=0, (2.4)

see for instance [39, Corollary 3.13].

Let Ch\nabla^{Ch} be the Chern connection of a Hermitian manifold (M,g,J)(M,g,J), and let 𝒯\mathcal{T}_{\nabla} be its torsion. Considering the 11-form222Notice that tr(𝒯)=Λ(dΩ){\rm{tr}}(\mathcal{\mathcal{T}}_{\nabla})=\Lambda(\rm{d}\Omega), see for instance [17], [18].

tr(𝒯)(X):=trace(Y𝒯(X,Y)),X𝔛(M),{\rm{tr}}(\mathcal{\mathcal{T}}_{\nabla})(X):={\text{trace}}\big{(}Y\mapsto\mathcal{\mathcal{T}}_{\nabla}(X,Y)\big{)},\ \ \ \forall X\in\mathfrak{X}(M), (2.5)

we define the torsion (1,0)(1,0)-form associated to Ch\nabla^{Ch} as being

τ:=tr(𝒯)1,0=Λ(Ω).\tau:={\rm{tr}}(\mathcal{T}_{\nabla})^{1,0}=\Lambda(\partial\Omega). (2.6)

From Definition 2.1, it follows that θ=1n1(τ+τ¯)\theta=\frac{1}{n-1}(\tau+\overline{\tau}). By considering the Hodge \ast-operator defined by gg and the associated codifferential δ:=d\delta:=-\ast{\rm{d}}\ast, from the decomposition d=+¯{\rm{d}}=\partial+\bar{\partial}, we have δ:=+¯\delta:=\partial^{\ast}+\bar{\partial}^{\ast}, such that

:=¯,¯:=.\partial^{\ast}:=-\ast\bar{\partial}\ast,\ \ \ \bar{\partial}^{\ast}:=-\ast\partial\ast. (2.7)

In the above context, since Ωk=k!(n1)!Ωnk\ast\Omega^{k}=\frac{k!}{(n-1)!}\Omega^{n-k}, we have from Lemma 2.3 that

δΩ=1(n1)!d(Ωn1)=1(n2)!(θΩn1)=1(n2)!Ln1(θ).\displaystyle{\delta\Omega=-\frac{1}{(n-1)!}\ast{\rm{d}}(\Omega^{n-1})=-\frac{1}{(n-2)!}}\ast\big{(}\theta\wedge\Omega^{n-1}\big{)}=-\frac{1}{(n-2)!}\ast{\rm{L}}^{n-1}(\theta).

Considering J(θ)=θJ=1n1(τ¯τ)J(\theta)=-\theta\circ J=\frac{\sqrt{-1}}{n-1}(\overline{\tau}-\tau), since Λ(θ)=0\Lambda(\theta)=0, it follows that Ln1(θ)=(n1)!J(θ)\ast{\rm{L}}^{n-1}(\theta)=(n-1)!J(\theta), e.g. [39, Theorem 3.16]. Therefore, we conclude that

θ=1n1J(δΩ).\theta=\frac{1}{n-1}J(\delta\Omega). (2.8)

From above, we obtain the following description for the torsion (1,0)(1,0)-form associated to the Chern connection Ch\nabla^{Ch}.

Lemma 2.5.

Let (M,Ω,J)(M,\Omega,J) be a compact Hermitian manifold and τ\tau the torsion (1,0)(1,0)-form of the associated to the Chern connection Ch\nabla^{Ch}. Then

τ=1¯Ω.\tau=-\sqrt{-1}\bar{\partial}^{\ast}\Omega. (2.9)

2.2. Levi-Civita Ricci curvature

Given a Hermitian manifold (M,Ω,J)(M,\Omega,J), consider its first Chern Ricci curvature

Ric(1)(Ω)=loc12ddclog(det(Ω)),{\rm{Ric}}^{(1)}(\Omega)\mathrel{\stackrel{{\scriptstyle\makebox[0.0pt]{\mbox{\tiny loc}}}}{{=}}}-\frac{1}{2}{\rm{d}}{\rm{d}}^{c}\log(\det(\Omega)), (2.10)

such that dc=Jd{\rm{d}}^{c}=J\circ{\rm{d}}. It is well-known that Ric(1)(Ω)2πc1(M){\rm{Ric}}^{(1)}(\Omega)\in 2\pi c_{1}(M) and, if dΩ=0{\rm{d}}\Omega=0, it follows that

Ric(1)(Ω)=Ric(g)(J𝟙),{\rm{Ric}}^{(1)}(\Omega)={\rm{Ric}}(g)(J\otimes\mathbbm{1}), (2.11)

where Ric(g){\rm{Ric}}(g) is the Ricci tensor associated to Riemannian metric g=Ω(𝟙J)g=\Omega(\mathbbm{1}\otimes J). In the case that dΩ0{\rm{d}}\Omega\neq 0, we have several different types of Ricci curvatures, e.g. [28]. Besides the first Chern Ricci curvature, we also shall consider in this paper the following notion of Ricci curvature.

Definition 2.6 (Theorem 1.12, [28]).

Let (M,Ω,J)(M,\Omega,J) be a compact Hermitian manifold. The first Levi-Civita Ricci curvature of (M,Ω,J)(M,\Omega,J) is defined by

(1)(Ω):=Ric(1)(Ω)12(Ω+¯¯Ω).\mathfrak{R}^{(1)}(\Omega):={\rm{Ric}}^{(1)}(\Omega)-\frac{1}{2}\big{(}\partial\partial^{\ast}\Omega+\bar{\partial}\bar{\partial}^{\ast}\Omega\big{)}. (2.12)

Given a compact Hermitian manifold (M,Ω,J)(M,\Omega,J), we have the Bott-Chern cohomology and the Aeppli cohomology of (M,Ω,J)(M,\Omega,J) defined, respectively, by

HBCp,q(M):=ker(d)Ωp,q(M)im(¯)Ωp,q(M),HAp,q(M):=ker(¯)Ωp,q(M)im()Ωp,q(M)+im(¯)Ωp,q(M)\displaystyle{H^{p,q}_{BC}(M):=\frac{\ker({\rm{d}})\cap\Omega^{p,q}(M)}{\operatorname{im}(\partial\bar{\partial})\cap\Omega^{p,q}(M)},\ \ H^{p,q}_{A}(M):=\frac{\ker(\partial\bar{\partial})\cap\Omega^{p,q}(M)}{\operatorname{im}(\partial)\cap\Omega^{p,q}(M)+\operatorname{im}(\bar{\partial})\cap\Omega^{p,q}(M)}}.

For more details, see for instance [2] and references therein.

Definition 2.7 (Definition 1.1, [28]).

Let 𝐋M{\bf{L}}\to M be a holomorphic line bundle over MM. The first Aeppli-Chern class of 𝐋{\bf{L}} is defined by

c1AC(𝐋):=[12π𝚯()]AHA1,1(M),c_{1}^{AC}({\bf{L}}):=\bigg{[}\frac{\sqrt{-1}}{2\pi}{\bf{\Theta}}(\nabla)\bigg{]}_{A}\in H^{1,1}_{A}(M), (2.13)

where 𝚯(){\bf{\Theta}}(\nabla) is the curvature of the Chern connection =d+dlog(𝐇)\nabla={\rm{d}}+{\rm{d}}\log({\bf{H}}), for some Hermitian metric 𝐇{\bf{H}} on 𝐋{\bf{L}}. In particular, the first Aeppli-Chern class of MM is defined by c1AC(M):=c1AC(𝐊M1)c_{1}^{AC}(M):=c_{1}^{AC}({\bf{K}}_{M}^{-1}).

Remark 2.8.

In the setting of the above definition, similarly, we define the first Bott-Chern class of 𝐋M{\bf{L}}\to M by

c1BC(𝐋):=[12π𝚯()]BCHBC1,1(M).c_{1}^{BC}({\bf{L}}):=\bigg{[}\frac{\sqrt{-1}}{2\pi}{\bf{\Theta}}(\nabla)\bigg{]}_{BC}\in H^{1,1}_{BC}(M). (2.14)

The first Bott-Chern class of MM is defined by c1BC(M):=c1BC(𝐊M1)c_{1}^{BC}(M):=c_{1}^{BC}({\bf{K}}_{M}^{-1}).

From Definition 2.6, given a compact Hermitian manifold (M,Ω,J)(M,\Omega,J), it follows that

c1AC(M)=[Ric(1)(Ω)2π]A=[(1)(Ω)2π]A\displaystyle c_{1}^{AC}(M)=\bigg{[}\frac{{\rm{Ric}}^{(1)}(\Omega)}{2\pi}\bigg{]}_{A}=\bigg{[}\frac{\mathfrak{R}^{(1)}(\Omega)}{2\pi}\bigg{]}_{A}.

The vanishing of the cohomology classes described above are related by the following result.

Proposition 2.9 (Corollary 1.3, [28]).

Let MM be a complex manifold. Then

c1BC(M)=0c1(M)=0c1AC(M)=0.c_{1}^{BC}(M)=0\Longrightarrow c_{1}(M)=0\Longrightarrow c_{1}^{AC}(M)=0. (2.15)

Moreover, on a complex manifold satisfying the ¯\partial\bar{\partial}-lemma, we have

c1BC(M)=0c1(M)=0c1AC(M)=0.c_{1}^{BC}(M)=0\iff c_{1}(M)=0\iff c_{1}^{AC}(M)=0. (2.16)

2.3. Scalar curvatures

Given a Hermitian manifold (M,g,J)(M,g,J), we can define its Chern scalar curvature as sC(g):=trΩ(Ric(1)(Ω))s_{C}(g):={\rm{tr}}_{\Omega}({\rm{Ric}}^{(1)}(\Omega)), such that

trΩ(Ric(1)(Ω))nΩn=Ric(1)(Ω)Ωn1.\frac{{\rm{tr}}_{\Omega}({\rm{Ric}}^{(1)}(\Omega))}{n}\Omega^{n}={\rm{Ric}}^{(1)}(\Omega)\wedge\Omega^{n-1}. (2.17)

By considering the underlying Riemannian structure of (M,g,J)(M,g,J), we also have the notion of Riemannian scalar curvature scal(g):=trg(Ric(g)){\rm{scal}}(g):={\rm{tr}}_{g}({\rm{Ric}}(g)). In the particular case that dΩ=0{\rm{d}}\Omega=0, i.e. when (M,g,J)(M,g,J) is Kähler, these two notions of scalar curvature are related by

scal(g)=2sC(g).{\rm{scal}}(g)=2s_{C}(g). (2.18)

It is worth pointing out that the converse of the above fact is not true in general, see for instance [14]. However, if MM is compact, then we have that scal(g)=2sC(g){\rm{scal}}(g)=2s_{C}(g) if and only if dΩ=0{\rm{d}\Omega=0}. Given a compact Hermitian manifold (M,g,J)(M,g,J), it will be useful for us to consider the following characterization of scal(g){\rm{scal}}(g):

scal(g)=2sC(g)+(Ω+¯¯Ω,Ω2|Ω|2)12|𝒯|2,{\rm{scal}}(g)=2s_{C}(g)+\Big{(}\big{\langle}\partial\partial^{\ast}\Omega+\bar{\partial}\bar{\partial}^{\ast}\Omega,\Omega\big{\rangle}-2|\partial^{\ast}\Omega|^{2}\Big{)}-\frac{1}{2}|\mathcal{T}_{\nabla}|^{2}, (2.19)

where333 α,βΩnn!=αβ¯\langle\alpha,\beta\rangle\frac{\Omega^{n}}{n!}=\alpha\wedge\ast\bar{\beta}, α,βΩp,q(M)\forall\alpha,\beta\in\Omega^{p,q}(M). |𝒯|2=1Λ(¯Ωζ),Ωζ+Ωζ+¯¯Ωζ,Ωζ|\mathcal{T}_{\nabla}|^{2}=\big{\langle}\sqrt{-1}\Lambda(\partial\bar{\partial}\Omega_{\zeta}),\Omega_{\zeta}\big{\rangle}+\big{\langle}\partial\partial^{\ast}\Omega_{\zeta}+\bar{\partial}\bar{\partial}^{\ast}\Omega_{\zeta},\Omega_{\zeta}\big{\rangle}, see [28, Corollary 4.2].

2.4. Calabi-Yau cones

Regarding r(0,+)r\in(0,+\infty) as a coordinate on the positive real line +\mathbbm{R}_{+}, we consider the following definition.

Definition 2.10.

A Riemannian manifold (Q,gQ)(Q,g_{Q}) is Sasakian if and only if its metric cone (𝒞(Q):=Q×+,g𝒞:=r2g+drdr)(\mathcal{C}(Q):=Q\times\mathbbm{R}_{+},g_{\mathcal{C}}:=r^{2}g+{\rm{d}}r\otimes{\rm{d}}r) is a Kähler cone.

Given a Sasakian manifold (Q,g)(Q,g), it follows that dim(Q)=2n+1\dim_{\mathbbm{R}}(Q)=2n+1. Denoting by 𝒥\mathcal{J} the associated complex structure on (𝒞(Q),g𝒞)(\mathcal{C}(Q),g_{\mathcal{C}}), we have the following facts:

  1. (1)

    the vector field rrr\partial_{r} is real holomorphic, i.e. rr𝒥=0;\mathscr{L}_{r\partial_{r}}\mathcal{J}=0;

  2. (2)

    the (Reeb) vector field ξ=𝒥(rr)\xi=\mathcal{J}(r\partial_{r}) is real holomorphic and Killing, i.e.

    ξ𝒥=0\mathscr{L}_{\xi}\mathcal{J}=0   and   ξg𝒞=0;\mathscr{L}_{\xi}g_{\mathcal{C}}=0;

  3. (3)

    the 11-form η=dclog(r)\eta={\rm{d}}^{c}\log(r), such that dc=𝒥d{\rm{d}}^{c}=\mathcal{J}\circ{\rm{d}}, satisfies the following

    η(ξ)=1\eta(\xi)=1   and   ξdη=0\xi\lrcorner{\rm{d}}\eta=0.

From above, one can describe the Kähler form ω𝒞=g𝒞(𝒥𝟙)\omega_{\mathcal{C}}=g_{\mathcal{C}}(\mathcal{J}\otimes\mathbbm{1}) as follows

ω𝒞=14ddcr2=d(r2η2).\omega_{\mathcal{C}}=\frac{1}{4}{\rm{d}}{\rm{d}}^{c}r^{2}={\rm{d}}\Big{(}\frac{r^{2}\eta}{2}\Big{)}. (2.20)

By considering the natural inclusion Q𝒞(Q)Q\hookrightarrow\mathcal{C}(Q), such that Q={r=1}Q=\{r=1\}, we can consider η\eta and ξ\xi as tensor fields on QQ. From this, since ω𝒞\omega_{\mathcal{C}} is a symplectic structure on 𝒞(Q)\mathcal{C}(Q), it follows that (dη)nη0({\rm{d}}\eta)^{n}\wedge\eta\neq 0 on QQ. Therefore, we have that (Q,η,ξ)(Q,\eta,\xi) is a contact manifold. Further, denoting 𝒟=ker(η)\mathcal{D}=\ker(\eta) and ξ=ξ\mathcal{F}_{\xi}=\big{\langle}\xi\big{\rangle}, we can define ΦEnd(TQ)\Phi\in{\rm{End}}(TQ), such that Φ=𝒥|𝒟\Phi=\mathcal{J}|_{\mathcal{D}} and Φ|ξ=0\Phi|_{\mathcal{F}_{\xi}}=0. From this, we have

  1. (1)

    𝒥2=𝟙Φ2=𝟙+ηξ\mathcal{J}^{2}=-\mathbbm{1}\Longrightarrow\Phi^{2}=-\mathbbm{1}+\eta\otimes\xi;

  2. (2)

    g𝒞(𝒥𝟙)=g𝒞(𝟙𝒥)gQ(ΦΦ)=gQηηg_{\mathcal{C}}(\mathcal{J}\otimes\mathbbm{1})=-g_{\mathcal{C}}(\mathbbm{1}\otimes\mathcal{J})\Longrightarrow g_{Q}(\Phi\otimes\Phi)=g_{Q}-\eta\otimes\eta;

  3. (3)

    [𝒥,𝒥]=0[Φ,Φ]+dηξ=0.[\mathcal{J},\mathcal{J}]=0\Longrightarrow[\Phi,\Phi]+{\rm{d}}\eta\otimes\xi=0.

By using the above relations one can show that

g=12dη(𝟙Φ)+ηη.g=\frac{1}{2}{\rm{d}}\eta(\mathbbm{1}\otimes\Phi)+\eta\otimes\eta. (2.21)

Moreover, we have that g|𝒟=12dη(𝟙Φ)g|_{\mathcal{D}}=\frac{1}{2}{\rm{d}}\eta(\mathbbm{1}\otimes\Phi) defines a Hermitian metric on 𝒟\mathcal{D}. From this, denoting g|𝒟=gTg|_{\mathcal{D}}=g^{T}, we have a (transverse) Kähler foliation (gT,𝒥|𝒟,𝒟)(g^{T},\mathcal{J}|_{\mathcal{D}},\mathcal{D}) on QQ.

Remark 2.11.

In the above setting, we shall refer to 𝒮:=(ξ,η,Φ,g)\mathcal{S}:=(\xi,\eta,\Phi,g) as being a Sasakian structure on QQ.

Remark 2.12.

Denoting also by ξ\mathcal{F}_{\xi} the Reeb foliation on QQ defined by ξ\xi, unless otherwise stated, we will assume that all orbits of ξ\xi are all compact. Under this assumption, ξ\xi integrates to an isometric U(1){\rm{U}}(1) action on (Q,g)(Q,g). We call the Sasakian structure 𝒮\mathcal{S} quasi-regular if the U(1){\rm{U}}(1) action is locally free. If the U(1){\rm{U}}(1) action is free, we call the Sasakian structure 𝒮\mathcal{S} regular. In the regular or quasi-regular case, the leaf space X:=Q/ξX:=Q/\mathcal{F}_{\xi} has the structure of a manifold or orbifold, respectively. In both cases the transverse Kähler structure (gT,𝒥|𝒟,𝒟)(g^{T},\mathcal{J}|_{\mathcal{D}},\mathcal{D}) pushes down to a Kähler structure on XX. Moreover, we have the following facts:

  1. (1)

    QQ is the total space of a principal U(1){\rm{U}}(1)-(orbi)bundle over XX;

  2. (2)

    π:QX=Q/ξ\pi\colon Q\to X=Q/\mathcal{F}_{\xi} is a (an orbifold) Riemannian submersion;

  3. (3)

    dη=πωX{\rm{d}}\eta=\pi^{\ast}\omega_{X}, where ωX\omega_{X} is a non-trivial integral (orbifold) cohomology class;

  4. (4)

    (X,ωX)(X,\omega_{X}) is a Kähler (orbifold) manifold.

Proposition 2.13.

Let (Q,g)(Q,g) be a Sasakian manifold of dimension 2n+12n+1. Then the following are equivalent

  1. (1)

    (Q,g)(Q,g) is Sasaki-Einstein with Ric(g)=2ng{\rm{Ric}}(g)=2ng;

  2. (2)

    The Kähler cone (𝒞(Q),g𝒞)({\mathcal{C}}(Q),g_{\mathcal{C}}) is Ricci-flat, Ric(g𝒞)=0{\rm{Ric}}(g_{\mathcal{C}})=0;

  3. (3)

    The transverse Kähler structure to the Reeb foliation ξ\mathcal{F}_{\xi} is Kähler-Einstein with Ric(gT)=2(n+1)gT{\rm{Ric}}(g^{T})=2(n+1)g^{T}.

By a slight abuse of notation, we have the following

Ric(g𝒞)=Ric(g)2ng=Ric(gT)2(n+1)gT.{\rm{Ric}}(g_{\mathcal{C}})={\rm{Ric}}(g)-2ng={\rm{Ric}}(g^{T})-2(n+1)g^{T}. (2.22)

It follows from Remark 2.12 that every (quasi-)regular Sasaki-Einstein manifold is a principal U(1){\rm{U}}(1)-(orbi)bundle over a Kähler-Einstein (orbifold) manifold (X,ωX)(X,\omega_{X}). One can also obtain Sasaki-Einstein manifolds from Kähler-Einstein (orbifold) manifolds, we shall explore later this construction in the regular case (Example 4.1).

2.5. Hermitian Weyl-Einstein manifolds

Let (M,g,J)(M,g,J) be a connected complex Hermitian manifold, such that dim(M)2\dim_{\mathbbm{C}}(M)\geq 2.

Definition 2.14.

A Hermitian manifold (M,g,J)(M,g,J) is called locally conformally Kähler (l.c.K.) if it satisfies one of the following equivalent conditions:

  1. (1)

    There exists an open cover 𝒰\mathscr{U} of MM and a family of smooth functions fU:Uf_{U}\colon U\to\mathbbm{R}, U𝒰U\in\mathscr{U}, such that gU:=efUg|Ug_{U}:={\mathrm{e}}^{-f_{U}}g|_{U}, is Kählerian, U𝒰\forall U\in\mathscr{U}.

  2. (2)

    There exists a globally defined closed 11-form θΩ1(M)\theta\in\Omega^{1}(M), such that

    dΩ=θΩ.{\rm{d}}\Omega=\theta\wedge\Omega. (2.23)
Remark 2.15.

In the setting above, it follows from Definition 2.1 that θ\theta satisfying Eq. (2.23) is the Lee form associated to the Hermitian structure of a l.c.K. manifold (M,g,J)(M,g,J). Throughout this paper, unless otherwise stated, we shall assume for all l.c.K. manifold (M,g,J)(M,g,J) that θ\theta is not exact and θ0\theta\not\equiv 0.

An important subclass of l.c.K. manifolds is defined by the parallelism of the Lee form with respect to the Levi-Civita connection of gg. Being more precise, we have the following definition.

Definition 2.16.

A l.c.K. manifold (M,g,J)(M,g,J) is called a Vaisman manifold if θ=0\nabla\theta=0, where \nabla is the Levi-Civita connection of gg.

Remark 2.17 ([36], [15]).

Since the Lee form of a Vaisman manifold is parallel, it has constant norm. Thus, the underlying Hermitian metric can be rescaled such that θg=1||\theta||_{g}=1. The fundamental form Ω\Omega of the Vaisman metric with unit length Lee form satisfies the equality:

Ω=d(Jθ)+θJθ.\Omega=-{\rm{d}}(J\theta)+\theta\wedge J\theta. (2.24)

In the above description, we have that d(Jθ)0-{\rm{d}}(J\theta)\geq 0, e.g. [15, Theorem 5.1].

Example 2.18.

Let (𝒞(Q),g𝒞,𝒥)(\mathcal{C}(Q),g_{\mathcal{C}},\mathcal{J}) be the Kähler cone of a Sasaki manifold (Q,gQ)(Q,g_{Q}). By considering the identification 𝒞(Q)Q×\mathcal{C}(Q)\cong Q\times\mathbbm{R} defined by φ=log(r)\varphi=\log(r), it follows that g𝒞=e2φ(gQ+dφdφ)g_{\mathcal{C}}={\rm{e}}^{2\varphi}\big{(}g_{Q}+{\rm{d}}\varphi\otimes{\rm{d}}\varphi\big{)}. From above, we consider the Hermitian manifold (𝒞(Q),e2φg𝒞,𝒥)(\mathcal{C}(Q),{\rm{e}}^{-2\varphi}g_{\mathcal{C}},\mathcal{J}). Given a Sasaki automorphism ϕ:QQ\phi\colon Q\to Q and c>0c>0, let Γϕ,c\Gamma_{\phi,c} be the cyclic group defined by

Γϕ,c:=(x,φ)(ϕ(x),φ+log(c))\Gamma_{\phi,c}:=\Big{\langle}(x,\varphi)\mapsto(\phi(x),\varphi+\log(c))\Big{\rangle}.

Since Γϕ,c\Gamma_{\phi,c} acts by holomorphic isometries on (𝒞(Q),e2φg𝒞,𝒥)(\mathcal{C}(Q),{\rm{e}}^{-2\varphi}g_{\mathcal{C}},\mathcal{J}), the Hermitian structure (e2φg𝒞,𝒥)({\rm{e}}^{-2\varphi}g_{\mathcal{C}},\mathcal{J}) descends to a Hermitian structure (g,J)(g,J) on Σϕ,c(Q)=𝒞(Q)/Γϕ,c\Sigma_{\phi,c}(Q)=\mathcal{C}(Q)/\Gamma_{\phi,c}. By construction, since g=loce2φg𝒞g\mathrel{\stackrel{{\scriptstyle\makebox[0.0pt]{\mbox{\tiny loc}}}}{{=}}}{\rm{e}}^{-2\varphi}g_{\mathcal{C}}, it follows that (Σϕ,c(Q),g,J)(\Sigma_{\phi,c}(Q),g,J) defines a l.c.K. manifold with Lee form described by θ=loc2dφ\theta\mathrel{\stackrel{{\scriptstyle\makebox[0.0pt]{\mbox{\tiny loc}}}}{{=}}}-2{\rm{d}}\varphi, see for instance [23]. It is straightforward to show that θ\theta is also parallel with respect to the Levi-Civita connection of gg. Thus, we have that (Σϕ,c(Q),g,J)(\Sigma_{\phi,c}(Q),g,J) is a Vaisman manifold.

Remark 2.19.

In the previous example we have the following identification

Σϕ,c(Q)Q×[0,log(c)](ϕ(x),0)(x,log(c)),\Sigma_{\phi,c}(Q)\cong\frac{Q\times[0,\log(c)]}{(\phi(x),0)\sim(x,\log(c))}, (2.25)

i.e. Σϕ,c(Q)\Sigma_{\phi,c}(Q) can be seen as the suspension of QQ by ϕ\phi over the circle of length 2πlog(c)2\pi\log(c), e.g. [3]. We shall refer to Σϕ,c(Q)\Sigma_{\phi,c}(Q) as the suspension by (ϕ,c)(\phi,c) of QQ.

Given a conformal manifold (M,[g])(M,[g]), we have the following notion of compatible connection with the conformal class [g][g].

Definition 2.20.

A Weyl connection DD on a conformal manifold (M,[g])(M,[g]) is a torsion-free connection which preserves the conformal class [g][g]. In this last setting, we say that DD defines a Weyl structure on (M,[g])(M,[g]) and that (M,[g],D)(M,[g],D) is a Weyl manifold.

In the above definition by preserving the conformal class it means that for each g[g]g^{\prime}\in[g], we have a 11-form θg\theta_{g^{\prime}} (Higgs field), such that

Dg=θgg.Dg^{\prime}=\theta_{g^{\prime}}\otimes g^{\prime}. (2.26)

Let (M,[g],D)(M,[g],D) be a Weyl manifold, in what follows we shall fix a representative gg for the underlying conformal class and consider the 11-form θg\theta_{g} which defines its Higgs field.

Definition 2.21.

We say that a Weyl manifold (M,[g],D)(M,[g],D) is a Hermitian-Weyl manifold if it admits an almost complex structure JEnd(TM)J\in{\text{End}}(TM), which satisfies:

  1. (1)

    g(JX,JY)=g(X,Y)g(JX,JY)=g(X,Y), X,Y𝔛(M);\forall X,Y\in\mathfrak{X}(M);

  2. (2)

    DJ=0.DJ=0.

An important result to be considered in the setting of Hermitian-Weyl manifolds is the following proposition.

Proposition 2.22 (Vaisman).

Any Hermitian-Weyl manifold of (real) dimension at least 66 is l.c.K.. Conversely, a l.c.K. manifold of (real) dimension at least 44 admits a compatible Hermitian-Weyl structure.

For a compact Hermitian-Weyl manifold (M,[g],D,J)(M,[g],D,J) Gauduchon [18] showed that, up to homothety, there is a unique choice of metric g0g_{0} in the conformal class [g][g] such that the corresponding 11-form θg0\theta_{g_{0}} is co-closed.

Definition 2.23.

The unique (up to homothety) l.c.K. metric g0g_{0} in the conformal class [g][g] of (M,[g],D,J)(M,[g],D,J) with harmonic associated Lee form is called the Gauduchon metric.

It follows from Proposition 2.22 that any compact Vaisman manifold admits a Hermitian-Weyl structure uniquely determined by the Gauduchon metric.

Definition 2.24.

A Hermitian-Weyl manifold is Hermitian Weyl-Einstein if the symmetric part of the Ricci tensor of the Weyl connection is proportional to the metric.

In the setting of the above definition, it can be shown that the Hermitian Weyl-Einstein condition is equivalent to

Ric(g0)=(n2)(θg02g0θg0θg0),{\rm{Ric}}(g_{0})=(n-2)\big{(}||\theta_{g_{0}}||^{2}g_{0}-\theta_{g_{0}}\otimes\theta_{g_{0}}\big{)}, (2.27)

see for instance [19]. From a deep result of Gauduchon in [19], it follows that:

Theorem 2.25.

Let (M,[g],D,J)(M,[g],D,J) be a compact Hermitian Weyl-Einstein manifold. Then the Ricci tensor of the Weyl connection vanishes identically and the Lee form is parallel. In particular, (M,g0,J)(M,g_{0},J) is Vaisman.

In the above setting, we have that θg0\theta_{g_{0}} is harmonic and Ric(g0){\rm{Ric}}(g_{0}) is non-negative, using the Weitzenböck formula, one can show that b1(M)=1b_{1}(M)=1. Now we consider the following structure theorem [38].

Theorem 2.26.

Every compact Vaisman manifold with b1(M)=1b_{1}(M)=1 is isomorphic to (Σϕ,c(Q),g,J)(\Sigma_{\phi,c}(Q),g,J), where QQ is some compact Sasakian manifold.

Combining the result above with the previous comments, one can show that every compact Hermitian Weyl-Einstein manifold is isomorphic to (Σϕ,c(Q),g0,J)(\Sigma_{\phi,c}(Q),g_{0},J) as a Vaisman manifold, where QQ is a Sasaki-Einstein manifold. Thus, in the setting of Theorem 2.25 we have g0=loce2φgCYg_{0}\mathrel{\stackrel{{\scriptstyle\makebox[0.0pt]{\mbox{\tiny loc}}}}{{=}}}{\rm{e}}^{-2\varphi}g_{CY}, where gCYg_{CY} is a Calabi-Yau metric on the metric cone 𝒞(Q)\mathcal{C}(Q) of a Sasaki-Einstein manifold QQ.

3. Proof of main results

Lemma 3.1.

Let (M,Ω,J)(M,\Omega,J) be a compact l.c.K. manifold. Then

(1)(Ω)=Υ12d(Jθ),\mathfrak{R}^{(1)}(\Omega)=\Upsilon-\frac{1}{2}{\rm{d}}(J\theta), (3.1)

where θ\theta is the Lee form of (M,Ω,J)(M,\Omega,J) and Υ2πc1(M)\Upsilon\in 2\pi c_{1}(M). In particular, if (1)(Ω)=0\mathfrak{R}^{(1)}(\Omega)=0, then c1(M)=0c_{1}(M)=0.

Proof.

Given a l.c.K. manifold (M,Ω,J)(M,\Omega,J), it follows that Ω=locefUωU\Omega\mathrel{\stackrel{{\scriptstyle\makebox[0.0pt]{\mbox{\tiny loc}}}}{{=}}}{\rm{e}}^{f_{U}}\omega_{U}, thus

Ric(1)(Ω)=loc12ddclog(det(ωU))n2ddcfU.{\rm{Ric}}^{(1)}(\Omega)\mathrel{\stackrel{{\scriptstyle\makebox[0.0pt]{\mbox{\tiny loc}}}}{{=}}}-\frac{1}{2}{\rm{d}}{\rm{d}}^{c}\log(\det(\omega_{U}))-\frac{n}{2}{\rm{d}}{\rm{d}}^{c}f_{U}. (3.2)

Since, θ=locdfU\theta\mathrel{\stackrel{{\scriptstyle\makebox[0.0pt]{\mbox{\tiny loc}}}}{{=}}}{\rm{d}}f_{U}, it follows that Ric(1)(Ω)=loc12ddclog(det(ωU))n2d(Jθ){\rm{Ric}}^{(1)}(\Omega)\mathrel{\stackrel{{\scriptstyle\makebox[0.0pt]{\mbox{\tiny loc}}}}{{=}}}-\frac{1}{2}{\rm{d}}{\rm{d}}^{c}\log(\det(\omega_{U}))-\frac{n}{2}{\rm{d}}(J\theta). Thus, gluing 12ddclog(det(ωU))-\frac{1}{2}{\rm{d}}{\rm{d}}^{c}\log(\det(\omega_{U})), we obtain a globally defined (1,1)(1,1)-form Υ\Upsilon, such that

Ric(1)(Ω)=Υn2d(Jθ).{\rm{Ric}}^{(1)}(\Omega)=\Upsilon-\frac{n}{2}{\rm{d}}(J\theta). (3.3)

In particular, we have Υ2πc1(M)\Upsilon\in 2\pi c_{1}(M). Thus, if (M,Ω,J)(M,\Omega,J) is a compact l.c.K. manifold, it follows that

(1)(Ω)=Υn2d(Jθ)12(Ω+¯¯Ω).\mathfrak{R}^{(1)}(\Omega)=\Upsilon-\frac{n}{2}{\rm{d}}(J\theta)-\frac{1}{2}\big{(}\partial\partial^{\ast}\Omega+\bar{\partial}\bar{\partial}^{\ast}\Omega\big{)}. (3.4)

Considering the Hodge decomposition θ=θ1,0+θ0,1\theta=\theta^{1,0}+\theta^{0,1}, we observe the following:

  1. (1)

    dθ=0¯θ1,0=θ0,1,θ1,0=¯θ0,1=0{\rm{d}}\theta=0\Rightarrow\bar{\partial}\theta^{1,0}=-\partial\theta^{0,1},\ \ \partial\theta^{1,0}=\bar{\partial}\theta^{0,1}=0;

  2. (2)

    ¯θ1,0=θ0,1¯τ=τ¯d(Jθ)=2n11¯τ\bar{\partial}\theta^{1,0}=-\partial\theta^{0,1}\iff\bar{\partial}\tau=-\partial\bar{\tau}\Rightarrow{\rm{d}}(J\theta)=-\frac{2}{n-1}\sqrt{-1}\bar{\partial}\tau;

  3. (3)

    Ω+¯¯Ω=1τ¯+1¯τ=21¯τ=(n1)d(Jθ)\partial\partial^{\ast}\Omega+\bar{\partial}\bar{\partial}^{\ast}\Omega=-\sqrt{-1}\partial\bar{\tau}+\sqrt{-1}\bar{\partial}\tau=2\sqrt{-1}\bar{\partial}\tau=-(n-1){\rm{d}}(J\theta).

By using the above relations in Eq. (3.4), we conclude that

(1)(Ω)=Υ12d(Jθ).\mathfrak{R}^{(1)}(\Omega)=\Upsilon-\frac{1}{2}{\rm{d}}(J\theta). (3.5)

From this, it follows that [(1)(Ω)]2πc1(M)[\mathfrak{R}^{(1)}(\Omega)]\in 2\pi c_{1}(M). In particular, if (1)(Ω)=0\mathfrak{R}^{(1)}(\Omega)=0, we have c1(M)=0c_{1}(M)=0. In this last case, Eq. (3.1) holds for Υ=12d(Jθ)\Upsilon=\frac{1}{2}{\rm{d}}(J\theta). ∎

Remark 3.2.

In the setting of the above lemma, the fact that [(1)(Ω)]2πc1(M)[\mathfrak{R}^{(1)}(\Omega)]\in 2\pi c_{1}(M) also can be seen as a consequence of [28, Theorem 3.14, item (2)]. In fact, since ¯θ0,1=0\bar{\partial}\theta^{0,1}=0, it follows that ¯Ω=1¯τ¯=1(n1)¯θ0,1=0.\bar{\partial}\partial^{\ast}\Omega=-\sqrt{-1}\bar{\partial}\bar{\tau}=-\sqrt{-1}(n-1)\bar{\partial}\theta^{0,1}=0.

Theorem 3.3.

Let (Q,gSE)(Q,g_{SE}) be a compact Sasaki-Einstein manifold, ϕ:QQ\phi\colon Q\to Q a Sasaki automorphism, and c>0c>0. Then, the suspension Σϕ,c(Q)\Sigma_{\phi,c}(Q) by (ϕ,c)(\phi,c) of QQ admits a Levi-Civita Ricci-flat Hermitian metric.

Proof.

Given a Sasaki-Einstein manifold (Q,gSE)(Q,g_{SE}), it follows that the Kähler cone (𝒞(Q),ωCY:=14ddcr2,𝒥)(\mathcal{C}(Q),\omega_{CY}:=\frac{1}{4}{\rm{d}}{\rm{d}}^{c}r^{2},\mathcal{J}) is Calabi-Yau. Given a Sasaki automorphism ϕ:QQ\phi\colon Q\to Q and c>0c>0, we have that the Hermitian structure (𝒞(Q),e2φωCY,𝒥)(\mathcal{C}(Q),{\rm{e}}^{-2\varphi}\omega_{CY},\mathcal{J}), such that φ=log(r)\varphi=\log(r), descends to a Vaisman structure (Ω,J)(\Omega,J) on the suspension Σϕ,c(Q)\Sigma_{\phi,c}(Q). From Lemma 3.1, it follows that

(1)(Ω)=Υ12d(Jθ).\mathfrak{R}^{(1)}(\Omega)=\Upsilon-\frac{1}{2}{\rm{d}}(J\theta). (3.6)

By considering the projection :𝒞(Q)Σϕ,c(Q)\wp\colon\mathcal{C}(Q)\to\Sigma_{\phi,c}(Q), since Ω=loce2φωCY\Omega\mathrel{\stackrel{{\scriptstyle\makebox[0.0pt]{\mbox{\tiny loc}}}}{{=}}}{\rm{e}}^{-2\varphi}\omega_{CY}, it follows that Υ=Ric(1)(ωCY)=0\wp^{\ast}\Upsilon={\rm{Ric}}^{(1)}(\omega_{CY})=0. Thus, we have (1)(Ω)=12d(Jθ)\mathfrak{R}^{(1)}(\Omega)=-\frac{1}{2}{\rm{d}}(J\theta). Given ζ>1\zeta>-1, we consider the perturbed Hermitian metric Ωζ\Omega_{\zeta} on Σϕ,c(Q)\Sigma_{\phi,c}(Q) given by

Ωζ:=Ω+2ζ(1)(Ω)=Ωζd(Jθ).\Omega_{\zeta}:=\Omega+2\zeta\mathfrak{R}^{(1)}(\Omega)=\Omega-\zeta{\rm{d}}(J\theta). (3.7)

Since Ω\Omega is Vaisman, it follows from Eq. (2.24) that

Ωζ=(1+ζ)d(Jθ)+θJθdΩζ=θζΩζ,\Omega_{\zeta}=-(1+\zeta){\rm{d}}(J\theta)+\theta\wedge J\theta\Rightarrow{\rm{d}}\Omega_{\zeta}=\theta_{\zeta}\wedge\Omega_{\zeta}, (3.8)

such that θζ=11+ζθ\theta_{\zeta}=\frac{1}{1+\zeta}\theta. By construction, we have Ωζn=(1+ζ)n1Ωn\Omega_{\zeta}^{n}=(1+\zeta)^{n-1}\Omega^{n}, thus

Ric(1)(Ωζ)=Ric(1)(Ω)=Υn2d(Jθ)=n2d(Jθ).{\rm{Ric}}^{(1)}(\Omega_{\zeta})={\rm{Ric}}^{(1)}(\Omega)=\Upsilon-\frac{n}{2}{\rm{d}}(J\theta)=-\frac{n}{2}{\rm{d}}(J\theta). (3.9)

By considering the Hodge \ast-operator induced by Ωζ\Omega_{\zeta}, it follows that

Ωζ+¯¯Ωζ=(n1)d(Jθζ)=(n11+ζ)d(Jθ).\partial\partial^{\ast}\Omega_{\zeta}+\bar{\partial}\bar{\partial}^{\ast}\Omega_{\zeta}=-(n-1){\rm{d}}(J\theta_{\zeta})=-\bigg{(}\frac{n-1}{1+\zeta}\bigg{)}{\rm{d}}(J\theta). (3.10)

Combining the above expressions, we obtain

(1)(Ωζ):=Ric(1)(Ωζ)12(Ωζ+¯¯Ωζ)=(n+n11+ζ)d(Jθ).\mathfrak{R}^{(1)}(\Omega_{\zeta}):={\rm{Ric}}^{(1)}(\Omega_{\zeta})-\frac{1}{2}\big{(}\partial\partial^{\ast}\Omega_{\zeta}+\bar{\partial}\bar{\partial}^{\ast}\Omega_{\zeta}\big{)}=\bigg{(}-n+\frac{n-1}{1+\zeta}\bigg{)}{\rm{d}}(J\theta). (3.11)

Therefore, for ζ=1n\zeta=-\frac{1}{n}, we have that (Σϕ,c(Q),Ωζ,J)(\Sigma_{\phi,c}(Q),\Omega_{\zeta},J) is Levi-Civita Ricci-flat. ∎

Since every 3-Sasakian manifold is Sasaki-Einstein (e.g. [8], [26]) we obtain the following corollary.

Corollary 3.4.

Let (Q,gQ)(Q,g_{Q}) be a compact 33-Sasakian manifold, ϕ:QQ\phi\colon Q\to Q a Sasaki automorphism, and c>0c>0. Then, the suspension Σϕ,c(Q)\Sigma_{\phi,c}(Q) by (ϕ,c)(\phi,c) of QQ admits a Levi-Civita Ricci-flat Hermitian metric.

From [38] and Theorem 3.3, we have the following result.

Corollary 3.5.

Every compact Hermitian Weyl-Einstein manifold admits a Levi-Civita Ricci-flat Hermitian metric. In particular, every compact locally conformal hyperKähler manifold admits a Levi-Civita Ricci-flat Hermitian metric.

Combining the result of Theorem 3.3 with [29], see also [7, Conjecture 4], we have the following corollaries.

Corollary 3.6.

Let L(𝐚):=Y(𝐚)S2n+1{\rm{L}}({\bf{a}}):=Y({\bf{a}})\cap S^{2n+1} be the link of a Brieskorn-Pham singularity

Y(𝐚):=(z0a0++znan=0)n+1,Y({\bf{a}}):=\Big{(}z_{0}^{a_{0}}+\cdots+z_{n}^{a_{n}}=0\Big{)}\subset\mathbbm{C}^{n+1}, (3.12)

such that n3n\geq 3. Assume that a0ana_{0}\leq\cdots\leq a_{n}. Then L(𝐚)×S1{\rm{L}}({\bf{a}})\times S^{1} admits a Levi-Civita Ricci-flat Hermitian metric if

1<j=0n1aj<1+nan.1<\sum_{j=0}^{n}\frac{1}{a_{j}}<1+\frac{n}{a_{n}}. (3.13)
Corollary 3.7.

Let 𝚺{\bf{\Sigma}} be an odd dimensional homotopy sphere which bounds a parallelizable manifold. Then 𝚺×S1{\bf{\Sigma}}\times S^{1} admits a Levi-Civita Ricci-flat Hermitian metric.

Our next result generalizes some ideas introduced in [29, Theorem 6.4].

Theorem 3.8.

Let (Q,gSE)(Q,g_{SE}) be a compact Sasaki-Einstein manifold, ϕ:QQ\phi\colon Q\to Q a Sasaki automorphism, and c>0c>0. Then, the suspension Σϕ,c(Q)\Sigma_{\phi,c}(Q) by (ϕ,c)(\phi,c) of QQ admits three different Hermitian metrics Ωi\Omega_{i}, i=1,2,3i=1,2,3, satisfying the following properties:

  1. (1)

    Ric(1)(Ω1)=Ric(1)(Ω2)=Ric(1)(Ω3)0{\rm{Ric}}^{(1)}(\Omega_{1})={\rm{Ric}}^{(1)}(\Omega_{2})={\rm{Ric}}^{(1)}(\Omega_{3})\geq 0;

  2. (2)

    Ω1\Omega_{1} has strictly positive Riemannian scalar curvature;

  3. (3)

    Ω2\Omega_{2} has zero Riemannian scalar curvature;

  4. (4)

    Ω3\Omega_{3} has strictly negative Riemannian scalar curvature.

In particular, all compact Hermitian manifolds of the previous corollaries admit three different Hermitian metrics satisfying the above properties.

Proof.

For every ζ>1\zeta>-1, let (Σϕ,c(Q),Ωζ,J)(\Sigma_{\phi,c}(Q),\Omega_{\zeta},J) be the Hermitian compact manifold constructed as in the proof of Theorem 3.3. It follows from Eq. (2.19) that the Riemannian scalar curvature of the underlying Riemannian metric gζ=Ωζ(𝟙J)g_{\zeta}=\Omega_{\zeta}(\mathbbm{1}\otimes J) is given by

scal(gζ)=2sC(gζ)+(Ωζ+¯¯Ωζ,Ωζ2|Ωζ|2)12|𝒯|2,{\rm{scal}}(g_{\zeta})=2s_{C}(g_{\zeta})+\Big{(}\big{\langle}\partial\partial^{\ast}\Omega_{\zeta}+\bar{\partial}\bar{\partial}^{\ast}\Omega_{\zeta},\Omega_{\zeta}\big{\rangle}-2|\partial^{\ast}\Omega_{\zeta}|^{2}\Big{)}-\frac{1}{2}|\mathcal{T}_{\nabla}|^{2}, (3.14)

where |𝒯|2=1Λ(¯Ωζ),Ωζ+Ωζ+¯¯Ωζ,Ωζ|\mathcal{T}_{\nabla}|^{2}=\big{\langle}\sqrt{-1}\Lambda(\partial\bar{\partial}\Omega_{\zeta}),\Omega_{\zeta}\big{\rangle}+\big{\langle}\partial\partial^{\ast}\Omega_{\zeta}+\bar{\partial}\bar{\partial}^{\ast}\Omega_{\zeta},\Omega_{\zeta}\big{\rangle}. Since we have

Ωζ=(1+ζ)d(Jθ)+θJθ,dΩζ=θζΩζ,\Omega_{\zeta}=-(1+\zeta){\rm{d}}(J\theta)+\theta\wedge J\theta,\ \ \ {\rm{d}}\Omega_{\zeta}=\theta_{\zeta}\wedge\Omega_{\zeta}, (3.15)

such that θζ=11+ζθ\theta_{\zeta}=\frac{1}{1+\zeta}\theta, in order to compute scal(gζ){\rm{scal}}(g_{\zeta}), it will be useful to consider the following identities:

  1. (A)

    vol:=1n!Ωζn=(1)n1(ζ+1)n1(n1)!(d(Jθ))n1θJθ\displaystyle{\rm{vol}}:=\frac{1}{n!}\Omega_{\zeta}^{n}=\frac{(-1)^{n-1}(\zeta+1)^{n-1}}{(n-1)!}({\rm{d}}(J\theta))^{n-1}\wedge\theta\wedge J\theta;

  2. (B)

    d(Jθ)Ωζn1=(1)n2(n1)(ζ+1)n2(d(Jθ))n1θJθ{\rm{d}}(J\theta)\wedge\Omega_{\zeta}^{n-1}=(-1)^{n-2}(n-1)(\zeta+1)^{n-2}({\rm{d}}(J\theta))^{n-1}\wedge\theta\wedge J\theta;

  3. (C)

    θJθΩζn1=(1)n1(ζ+1)n1(d(Jθ))n1θJθ.\theta\wedge J\theta\wedge\Omega_{\zeta}^{n-1}=(-1)^{n-1}(\zeta+1)^{n-1}({\rm{d}}(J\theta))^{n-1}\wedge\theta\wedge J\theta.

Notice that, from (A), (B), and (C), it follows that

d(Jθ)Ωζn1=(n1)!(n1ζ+1)volandθJθΩζn1=(n1)!vol.{\rm{d}}(J\theta)\wedge\Omega_{\zeta}^{n-1}=-(n-1)!\bigg{(}\frac{n-1}{\zeta+1}\bigg{)}{\rm{vol}}\ \ \ {\text{and}}\ \ \ \theta\wedge J\theta\wedge\Omega_{\zeta}^{n-1}=(n-1)!{\rm{vol}}. (3.16)

Since Ωζ+¯¯Ωζ=(n1)d(Jθζ)\partial\partial^{\ast}\Omega_{\zeta}+\bar{\partial}\bar{\partial}^{\ast}\Omega_{\zeta}=-(n-1){\rm{d}}(J\theta_{\zeta}), it follows that

Ωζ+¯¯Ωζ,Ωζvol=(n1)d(Jθζ)Ωζ=(n1ζ+1)d(Jθ)Ωζ.\big{\langle}\partial\partial^{\ast}\Omega_{\zeta}+\bar{\partial}\bar{\partial}^{\ast}\Omega_{\zeta},\Omega_{\zeta}\big{\rangle}{\rm{vol}}=-(n-1){\rm{d}}(J\theta_{\zeta})\wedge\ast\Omega_{\zeta}=-\bigg{(}\frac{n-1}{\zeta+1}\bigg{)}{\rm{d}}(J\theta)\wedge\ast\Omega_{\zeta}. (3.17)

As Ωζ=Ωζn1(n1)!\ast\Omega_{\zeta}=\frac{\Omega_{\zeta}^{n-1}}{(n-1)!}, it follow from Eq. (3.16) that

(n1ζ+1)d(Jθ)Ωζ=(n1ζ+1)2vol.-\bigg{(}\frac{n-1}{\zeta+1}\bigg{)}{\rm{d}}(J\theta)\wedge\ast\Omega_{\zeta}=\bigg{(}\frac{n-1}{\zeta+1}\bigg{)}^{2}{\rm{vol}}. (3.18)

Therefore, we conclude that

Ωζ+¯¯Ωζ,Ωζ=(n1ζ+1)2.\big{\langle}\partial\partial^{\ast}\Omega_{\zeta}+\bar{\partial}\bar{\partial}^{\ast}\Omega_{\zeta},\Omega_{\zeta}\big{\rangle}=\bigg{(}\frac{n-1}{\zeta+1}\bigg{)}^{2}. (3.19)

Now we observe that |Ωζ|2=Ωζ,Ωζ=Ωζ,Ωζ|\partial^{\ast}\Omega_{\zeta}|^{2}=\big{\langle}\partial^{\ast}\Omega_{\zeta},\partial^{\ast}\Omega_{\zeta}\big{\rangle}=\big{\langle}\partial\partial^{\ast}\Omega_{\zeta},\Omega_{\zeta}\big{\rangle}. Since

Ωζ=(n1)2d(Jθζ)=12(n1ζ+1)d(Jθ),\partial\partial^{\ast}\Omega_{\zeta}=-\frac{(n-1)}{2}{\rm{d}}(J\theta_{\zeta})=-\frac{1}{2}\bigg{(}\frac{n-1}{\zeta+1}\bigg{)}{\rm{d}}(J\theta), (3.20)

from a similar argument as in Eq. (3.18), we have

Ωζ,Ωζvol=12(n1ζ+1)d(Jθ)Ωζ=12(n1ζ+1)2vol.\big{\langle}\partial^{\ast}\Omega_{\zeta},\partial^{\ast}\Omega_{\zeta}\big{\rangle}{\rm{vol}}=-\frac{1}{2}\bigg{(}\frac{n-1}{\zeta+1}\bigg{)}{\rm{d}}(J\theta)\wedge\ast\Omega_{\zeta}=\frac{1}{2}\bigg{(}\frac{n-1}{\zeta+1}\bigg{)}^{2}{\rm{vol}}. (3.21)

Hence, we obtain

|Ωζ|2=12(n1ζ+1)2.|\partial^{\ast}\Omega_{\zeta}|^{2}=\frac{1}{2}\bigg{(}\frac{n-1}{\zeta+1}\bigg{)}^{2}. (3.22)

Replacing Eq. (3.19) and Eq. (3.22) in Eq. (3.14), we obtain

scal(gζ)=2sC(gζ)12|𝒯|2.{\rm{scal}}(g_{\zeta})=2s_{C}(g_{\zeta})-\frac{1}{2}|\mathcal{T}_{\nabla}|^{2}. (3.23)

Since Ric(1)(Ωζ)=n2d(Jθ){\rm{Ric}}^{(1)}(\Omega_{\zeta})=-\frac{n}{2}{\rm{d}}(J\theta), it follows that

sC(gζ)nΩζn=Ric(1)(Ωζ)Ωζn1=n2d(Jθ)Ωζn1.\frac{s_{C}(g_{\zeta})}{n}\Omega_{\zeta}^{n}={\rm{Ric}}^{(1)}(\Omega_{\zeta})\wedge\Omega_{\zeta}^{n-1}=-\frac{n}{2}{\rm{d}}(J\theta)\wedge\Omega_{\zeta}^{n-1}. (3.24)

From Eq. (3.16), we obtain

sC(gζ)nΩζn=n!2(n1ζ+1)volsC(gζ)=n(n1)2(ζ+1).\frac{s_{C}(g_{\zeta})}{n}\Omega_{\zeta}^{n}=\frac{n!}{2}\bigg{(}\frac{n-1}{\zeta+1}\bigg{)}{\rm{vol}}\iff s_{C}(g_{\zeta})=\frac{n(n-1)}{2(\zeta+1)}. (3.25)

In order to describe scal(gζ){\rm{scal}}(g_{\zeta}), it remains to compute 12|𝒯|2\frac{1}{2}|\mathcal{T}_{\nabla}|^{2}. From Eq. (3.19), we have

|𝒯|2=1Λ(¯Ωζ),Ωζ+(n1ζ+1)2.|\mathcal{T}_{\nabla}|^{2}=\big{\langle}\sqrt{-1}\Lambda(\partial\bar{\partial}\Omega_{\zeta}),\Omega_{\zeta}\big{\rangle}+\bigg{(}\frac{n-1}{\zeta+1}\bigg{)}^{2}. (3.26)

We notice that

1Λ(¯Ωζ),Ωζ=1¯Ωζ,L(Ωζ)=1¯Ωζ,Ωζ2.\big{\langle}\sqrt{-1}\Lambda(\partial\bar{\partial}\Omega_{\zeta}),\Omega_{\zeta}\big{\rangle}=\big{\langle}\sqrt{-1}\partial\bar{\partial}\Omega_{\zeta},{\rm{L}}(\Omega_{\zeta})\big{\rangle}=\big{\langle}\sqrt{-1}\partial\bar{\partial}\Omega_{\zeta},\Omega_{\zeta}^{2}\big{\rangle}. (3.27)

By using that dΩζ=θζΩζ{\rm{d}}\Omega_{\zeta}=\theta_{\zeta}\wedge\Omega_{\zeta} and dθζ=0{\rm{d}}\theta_{\zeta}=0, we obtain

1¯Ωζ=12d(Jθζ)Ωζ+12θζJθζΩζ.\sqrt{-1}\partial\bar{\partial}\Omega_{\zeta}=\frac{1}{2}{\rm{d}}(J\theta_{\zeta})\wedge\Omega_{\zeta}+\frac{1}{2}\theta_{\zeta}\wedge J\theta_{\zeta}\wedge\Omega_{\zeta}. (3.28)

From above, by using that Ωζ2=2Ωζn2(n2)!\ast\Omega_{\zeta}^{2}=\frac{2\Omega_{\zeta}^{n-2}}{(n-2)!}, it follows that:

  1. (i)

    12d(Jθζ)ΩζΩζ2=d(Jθ)Ωζn1(n2)!(ζ+1)=(n1ζ+1)2vol\displaystyle\frac{1}{2}{\rm{d}}(J\theta_{\zeta})\wedge\Omega_{\zeta}\wedge\ast\Omega_{\zeta}^{2}=\frac{{\rm{d}}(J\theta)\wedge\Omega_{\zeta}^{n-1}}{(n-2)!(\zeta+1)}=-\bigg{(}\frac{n-1}{\zeta+1}\bigg{)}^{2}{\rm{vol}};

  2. (ii)

    12θζJθζΩζΩζ2=θJθΩζn1(n2)!(ζ+1)2=(n1)(ζ+1)2vol\displaystyle\frac{1}{2}\theta_{\zeta}\wedge J\theta_{\zeta}\wedge\Omega_{\zeta}\wedge\ast\Omega_{\zeta}^{2}=\frac{\theta\wedge J\theta\wedge\Omega_{\zeta}^{n-1}}{(n-2)!(\zeta+1)^{2}}=\frac{(n-1)}{(\zeta+1)^{2}}{\rm{vol}}.

Here we have used Eq. (3.16) to obtain the above expressions. Hence, we conclude that

1Λ(¯Ωζ),Ωζ=1¯Ωζ,Ωζ2=(n1ζ+1)2+(n1)(ζ+1)2.\big{\langle}\sqrt{-1}\Lambda(\partial\bar{\partial}\Omega_{\zeta}),\Omega_{\zeta}\big{\rangle}=\big{\langle}\sqrt{-1}\partial\bar{\partial}\Omega_{\zeta},\Omega_{\zeta}^{2}\big{\rangle}=-\bigg{(}\frac{n-1}{\zeta+1}\bigg{)}^{2}+\frac{(n-1)}{(\zeta+1)^{2}}. (3.29)

Therefore, we have

|𝒯|2=(n1)(ζ+1)2.|\mathcal{T}_{\nabla}|^{2}=\frac{(n-1)}{(\zeta+1)^{2}}. (3.30)

Combining the above expression with Eq. (3.25), we obtain

scal(gζ)=n(n1)(ζ+1)(n1)2(ζ+1)2=n(n1)(ζ+1)2[ζ(12n)2n].{\rm{scal}}(g_{\zeta})=\frac{n(n-1)}{(\zeta+1)}-\frac{(n-1)}{2(\zeta+1)^{2}}=\frac{n(n-1)}{(\zeta+1)^{2}}\bigg{[}\zeta-\frac{(1-2n)}{2n}\bigg{]}. (3.31)

From this, we have the following:

  1. (1)

    ζ>(12n)2nΩζ\zeta>\frac{(1-2n)}{2n}\Rightarrow\Omega_{\zeta} has strictly positive constant Riemannian scalar curvature;

  2. (2)

    ζ=(12n)2nΩζ\zeta=\frac{(1-2n)}{2n}\Rightarrow\Omega_{\zeta} has constant zero Riemannian scalar curvature;

  3. (3)

    ζ<(12n)2nΩζ\zeta<\frac{(1-2n)}{2n}\Rightarrow\Omega_{\zeta} has strictly negative constant Riemannian scalar curvature.

Hence, since Ric(1)(Ωζ)=n2d(Jθ)0{\rm{Ric}}^{(1)}(\Omega_{\zeta})=-\frac{n}{2}{\rm{d}}(J\theta)\geq 0, ζ>1\forall\zeta>-1, see Remark 2.17, we can always find three different Hermitian metrics Ωi\Omega_{i}, i=1,2,3i=1,2,3, satisfying the desired properties. The last statement of the theorem follows immediately from the above construction, and from the fact that all compact Hermitian manifolds mentioned can be obtained as a suspension Σϕ,c(Q)\Sigma_{\phi,c}(Q) of some suitable Sasaki-Einstein manifold (Q,gQ)(Q,g_{Q}). ∎

4. Examples

In this section, in order to illustrate the main results, we provide a general method to construct explicit examples of Levi-Civita Ricci-flat Hermitian metric from Kähler-Einstein Fano (orbifolds) manifolds. Also, we illustrate the results of Theorem 3.8 in the case that QQ is an exotic 77-sphere.

Example 4.1.

Let (X,ωX,J)(X,\omega_{X},J) be a Kähler-Einstein Fano manifold of complex dimension nn and Fano index 𝐈(X){\bf{I}}(X)\in\mathbbm{Z}. Suppose that Ric(ωX)=λωX{\rm{Ric}}(\omega_{X})=\lambda\omega_{X}, for some λ>0\lambda>0. Considering 𝐋=𝐊X𝐈(X){\bf{L}}={\bf{K}}_{X}^{\otimes\frac{\ell}{{\bf{I}}(X)}}, for some >0\ell\in\mathbbm{Z}_{>0}, and let 𝐇{\bf{H}} be a Hermitian structure on 𝐋{\bf{L}}, such that

12π𝚯()=λωX2π𝐈(X),\frac{\sqrt{-1}}{2\pi}{\bf{\Theta}}(\nabla)=-\frac{\ell\lambda\omega_{X}}{2\pi{\bf{I}}(X)}, (4.1)

where 𝚯(){\bf{\Theta}}(\nabla) is the curvature of the associated Chern connection =d+dlog(𝐇)\nabla={\rm{d}}+{\rm{d}}\log({\bf{H}}). Considering the complex manifold Tot(𝐋×){\rm{Tot}}({\bf{L}}^{\times}) underlying the total space of the principal ×\mathbbm{C}^{\times}-bundle 𝐋×:=𝐋{0-section}{\bf{L}}^{\times}:={\bf{L}}-\{0{\text{-section}}\}, let r:Tot(𝐋×)r\colon{\rm{Tot}}({\bf{L}}^{\times})\to\mathbbm{R}, such that r2=𝐇(u,u)r^{2}={\bf{H}}(u,u), uTot(𝐋×)\forall u\in{\rm{Tot}}(\bf{L}^{\times}). Denoting by 𝒥\mathscr{J} the canonical complex structure on Tot(𝐋×){\rm{Tot}}({\bf{L}}^{\times}), we have

ddclog(r)=p(1𝚯())=p(λωX𝐈(X)),{\rm{d}}{\rm{d}}^{c}\log(r)=-p^{\ast}(\sqrt{-1}{\bf{\Theta}}(\nabla))=p^{\ast}\bigg{(}\frac{\ell\lambda\omega_{X}}{{\bf{I}}(X)}\bigg{)}, (4.2)

where p:Tot(𝐋×)Xp\colon{\rm{Tot}}({\bf{L}}^{\times})\to X is the associated bundle projection and dc=𝒥d{\rm{d}}^{c}=\mathscr{J}\circ{\rm{d}}. By considering the sphere bundle Q(𝐋)={u𝐋|𝐇(u,u)=1}Q({\bf{L}})=\big{\{}u\in{\bf{L}}\ \big{|}\sqrt{{\bf{H}}(u,u)}=1\big{\}} we have an identification Tot(𝐋×)Q(𝐋)×+{\rm{Tot}}({\bf{L}}^{\times})\cong Q({\bf{L}})\times\mathbbm{R}_{+} provide by the map

u(u𝐇(u,u),𝐇(u,u)).u\mapsto\Bigg{(}\frac{u}{\sqrt{{\bf{H}}(u,u)}},\sqrt{{\bf{H}}(u,u)}\Bigg{)}. (4.3)

Under the above identification, by considering the rescaled potential ϱ=r𝐈(X)(n+1)\varrho=r^{\frac{{\bf{I}}(X)}{\ell(n+1)}} we have a Kähler structure ωCY\omega_{CY} on 𝒞(Q(𝐋))\mathcal{C}(Q({\bf{L}})) defined by

ωCY:=14ddcϱ2=𝐈(X)(n+1)d(ϱ2η2),\omega_{CY}:=\frac{1}{4}{\rm{d}}{\rm{d}}^{c}\varrho^{2}=\frac{{\bf{I}}(X)}{\ell(n+1)}{\rm{d}}\Big{(}\frac{\varrho^{2}\eta}{2}\Big{)}, (4.4)

such that η=dclog(r)\eta={\rm{d}}^{c}\log(r). We notice that

dlog(ϱ)=𝐈(X)(n+1)dlog(r)𝒥(dϱϱ)=𝐈(X)(n+1)η\displaystyle{{\rm{d}}\log(\varrho)=\frac{{\bf{I}}(X)}{\ell(n+1)}{\rm{d}}\log(r)\Rightarrow\mathscr{J}\Big{(}\frac{{\rm{d}}\varrho}{\varrho}\Big{)}=\frac{{\bf{I}}(X)}{\ell(n+1)}\eta}.

Thus, a straightforward computation shows that

gCY=ωCY(𝟙𝒥)=ϱ2gSE+dϱdϱ,g_{CY}=\omega_{CY}(\mathbbm{1}\otimes\mathscr{J})=\varrho^{2}g_{SE}+{\rm{d}}\varrho\otimes{\rm{d}}\varrho, (4.5)

such that gSEg_{SE} is a Riemannian metric on Q(𝐋)Q({\bf{L}}) defined by

gSE=π(gKE)+𝐈(X)22(n+1)2ηη,g_{SE}=\pi^{\ast}(g_{KE})+\frac{{\bf{I}}(X)^{2}}{\ell^{2}(n+1)^{2}}\eta\otimes\eta, (4.6)

where gKE=λ2(n+1)ωX(𝟙J)g_{KE}=\frac{\lambda}{2(n+1)}\omega_{X}(\mathbbm{1}\otimes J) and π:Q(𝐋)X\pi\colon Q({\bf{L}})\to X is the associated projection. By construction, we have

Ric(gKE)=2(n+1)gKERic(gSE)=2ngSERic(gCY)=0{\rm{Ric}}(g_{KE})=2(n+1)g_{KE}\iff{\rm{Ric}}(g_{SE})=2ng_{SE}\iff{\rm{Ric}}(g_{CY})=0.

Therefore, we have that (Q(𝐋),gSE)(Q({\bf{L}}),g_{SE}) is a regular Sasaki-Einstein manifold and its metric cone (𝒞(Q(𝐋)),ωCY)(\mathcal{C}(Q({\bf{L}})),\omega_{CY}) is Calabi-Yau. From this, given a Sasaki automorphism ϕ:Q(𝐋)Q(𝐋)\phi\colon Q({\bf{L}})\to Q({\bf{L}}), and c>0c>0, it follows from Theorem 3.3 that the suspension Σϕ,c(Q(𝐋))\Sigma_{\phi,c}(Q({\bf{L}})) by (ϕ,c)(\phi,c) of Q(𝐋)Q({\bf{L}}) admits a Levi-Civita Ricci-flat Hermitian metric Ω\Omega given by

Ω=(11n)d(Jθ)+θJθ\Omega=-\big{(}1-\frac{1}{n}\big{)}{\rm{d}}(J\theta)+\theta\wedge J\theta,

such that

θ=locdlog(ϱ2)=𝐈(X)(m+1)dlog(r2)=𝐈(X)(m+1)dlog(𝐇).\theta\mathrel{\stackrel{{\scriptstyle\makebox[0.0pt]{\mbox{\tiny loc}}}}{{=}}}-{\rm{d}}\log(\varrho^{2})=-\frac{{\bf{I}}(X)}{\ell(m+1)}{\rm{d}}\log(r^{2})=-\frac{{\bf{I}}(X)}{\ell(m+1)}{\rm{d}}\log({\bf{H}}). (4.7)

Locally, we have 𝐇=h|w|2{\bf{H}}=h|w|^{2}, such that hC(U)h\in C^{\infty}(U), for some open set UXU\subset X which trivializes 𝐋{\bf{L}}, satisfying 12ddclog(h)=locλωX𝐈(X)\frac{1}{2}{\rm{d}}{\rm{d}}^{c}\log(h)\mathrel{\stackrel{{\scriptstyle\makebox[0.0pt]{\mbox{\tiny loc}}}}{{=}}}\frac{\ell\lambda\omega_{X}}{{\bf{I}}(X)}, see Eq. (4.2). Therefore, we conclude that

θ=loc𝐈(X)(m+1)[dlog(h)+w¯dw+wdw¯|w|2].\theta\mathrel{\stackrel{{\scriptstyle\makebox[0.0pt]{\mbox{\tiny loc}}}}{{=}}}-\frac{{\bf{I}}(X)}{\ell(m+1)}\bigg{[}{\rm{d}}\log(h)+\frac{\bar{w}{\rm{d}}w+w{\rm{d}}\bar{w}}{|w|^{2}}\bigg{]}. (4.8)

Summarizing, the Levi-Civita Ricci-flat metric Ω\Omega can be explicitly described in terms of the local potentials of the Chern connection \nabla on 𝐋{\bf{L}} which satisfies Eq. (4.1).

Remark 4.2.

An alternative way to equip 𝒞(Q(𝐋))\mathcal{C}(Q({\bf{L}})) with a Calabi-Yau metric is the following. In the setting of the previous example, consider (Q(𝐋),g)(Q({\bf{L}}),g), such that

g=π(λgKE2𝐈(X))+ηη,g=\pi^{\ast}\bigg{(}\frac{\ell\lambda g_{KE}}{2{\bf{I}}(X)}\bigg{)}+\eta\otimes\eta, (4.9)

where gKE=ωX(𝟙J)g_{KE}=\omega_{X}(\mathbbm{1}\otimes J). In this case, we have that (Q(𝐋),g)(Q({\bf{L}}),g) is a Sasakian manifold, and its Sasakian structure 𝒮=(ξ,η,Φ,g)\mathcal{S}=(\xi,\eta,\Phi,g) can be obtained from the Kähler cone (𝒞(Q(𝐋)),ω𝒞=d(r2η2),𝒥)(\mathcal{C}(Q({\bf{L}})),\omega_{\mathcal{C}}={\rm{d}}\big{(}\frac{r^{2}\eta}{2}\big{)},\mathscr{J}). Given a>0a>0, consider 𝒮a=(1aξ,aη,Φ,ga)\mathcal{S}_{a}=(\frac{1}{a}\xi,a\eta,\Phi,g_{a}), such that

ga=ag+(a2a)ηηg_{a}=ag+(a^{2}-a)\eta\otimes\eta.

By construction, we have that (Q(𝐋),ga)(Q({\bf{L}}),g_{a}) is a Riemannian manifold. From 𝒮a\mathcal{S}_{a} we can construct a complex structure 𝒥a\mathcal{J}_{a} on 𝒞(Q(𝐋)\mathcal{C}(Q({\bf{L}}) by setting

𝒥a(Y)=ϕ(Y)aη(Y)rr,𝒥a(rr)=1aξ,\mathcal{J}_{a}(Y)=\phi(Y)-a\eta(Y)r\partial_{r},\ \ \ \ \ \mathcal{J}_{a}(r\partial_{r})=\frac{1}{a}\xi,

Y𝔛(Q(𝐋))\forall Y\in\mathfrak{X}(Q({\bf{L}})). It follows that (𝒞(Q(𝐋)),g𝒞,a:=r2ga+drdr,𝒥a)(\mathcal{C}(Q({\bf{L}})),g_{\mathcal{C},a}:=r^{2}g_{a}+{\rm{d}}r\otimes{\rm{d}}r,\mathcal{J}_{a}) is a Kähler manifold. Notice that ω𝒞,a=g𝒞,a(𝒥a𝟙)\omega_{\mathcal{C},a}=g_{\mathcal{C},a}(\mathcal{J}_{a}\otimes\mathbbm{1}) is given by

ω𝒞,a=14ddcr2=ad(r2η2),\omega_{\mathcal{C},a}=\frac{1}{4}{\rm{d}}{\rm{d}}^{c}r^{2}=a{\rm{d}}\Big{(}\frac{r^{2}\eta}{2}\Big{)}, (4.10)

such that dc:=𝒥ad{\rm{d}}^{c}:=\mathcal{J}_{a}\circ{\rm{d}}. Thus, we have that (Q(𝐋),ga)(Q({\bf{L}}),g_{a}) is also Sasakian. In particular, we have the following:

a=𝐈(X)(n+1)ga=gSERic(g𝒞,a)=0,a=\frac{{\bf{I}}(X)}{\ell(n+1)}\Longrightarrow g_{a}=g_{SE}\Longrightarrow{\rm{Ric}}(g_{\mathcal{C},a})=0, (4.11)

i.e., a=𝐈(X)(n+1)(𝒞(Q(𝐋)),g𝒞,a:=r2ga+drdr,𝒥a)a=\frac{{\bf{I}}(X)}{\ell(n+1)}\Longrightarrow(\mathcal{C}(Q({\bf{L}})),g_{\mathcal{C},a}:=r^{2}g_{a}+{\rm{d}}r\otimes{\rm{d}}r,\mathcal{J}_{a}) is Calabi-Yau. In this case, from Theorem 3.3, we obtain a Levi-Civita Ricci-flat Hermitian manifold (Σϕ,c(Q(𝐋)),Ω,Ja)(\Sigma_{\phi,c}(Q({\bf{L}})),\Omega,J_{a}), here the complex structure JaJ_{a} is obtained from 𝒥a\mathcal{J}_{a}, for a=𝐈(X)(n+1)a=\frac{{\bf{I}}(X)}{\ell(n+1)}. In particular, in this case we have the underlying Lee form given by θ=locdlog(r2)\theta\mathrel{\stackrel{{\scriptstyle\makebox[0.0pt]{\mbox{\tiny loc}}}}{{=}}}-{\rm{d}}\log(r^{2}).

Remark 4.3.

By following the results given in [13] on principal elliptic bundles over flag varieties, and the construction provided by Example 4.1 (or by Remark 4.2), one can describe a huge class of explicit examples which illustrate the results of Theorem 3.3 and Theorem 3.8 by using tools of Lie theory.

Example 4.4.

Given a (n+1)(n+1)-tuple 𝐚=(a0,,an){\bf{a}}=(a_{0},\ldots,a_{n}), such that aj>0a_{j}\in\mathbb{Z}_{>0}, aj>1a_{j}>1, and n3n\geq 3, consider the Brieskorn-Pham singularity (e.g. [10], [16]) given by

Y(𝐚):=(z0a0++znan=0)n+1.Y({\bf{a}}):=\Big{(}z_{0}^{a_{0}}+\cdots+z_{n}^{a_{n}}=0\Big{)}\subset\mathbbm{C}^{n+1}. (4.12)

Let us assume that a0ana_{0}\leq\cdots\leq a_{n}. Set C=lcm(aj:j=0,,n)C={\rm{lcm}}(a_{j}:j=0,\ldots,n) and consider the ×\mathbbm{C}^{\times}-action on Y(𝐚)Y({\bf{a}}) defined by

λ(z0,,zn)=(λCa0z0,,λCanzn),λ×.\lambda\cdot(z_{0},\ldots,z_{n})=(\lambda^{\frac{C}{a_{0}}}z_{0},\ldots,\lambda^{\frac{C}{a_{n}}}z_{n}),\ \ \forall\lambda\in\mathbbm{C}^{\times}. (4.13)

From above, we have the following facts [7]:

  1. (1)

    Xorb(𝐚):=Y(𝐚)/×X^{{\rm{orb}}}({\bf{a}}):=Y({\bf{a}})/\mathbbm{C}^{\times} is a Kähler orbifold;

  2. (2)

    Xorb(𝐚)X^{{\rm{orb}}}({\bf{a}}) is a Fano orbifold if and only if 1<j=0n1aj1<\sum_{j=0}^{n}\frac{1}{a_{j}}.

Consider now the Brieskorn manifold (Figure 1)

L(𝐚):=Y(𝐚)S2n+1.{\rm{L}}({\bf{a}}):=Y({\bf{a}})\cap S^{2n+1}. (4.14)

The Brieskorn manifold L(𝐚){\rm{L}}({\bf{a}}) is a smooth (2n1)(2n-1)-dimensional compact manifold. Moreover, by following [1], [33], and [5], we have a well-known quasi-regular Sasakian structure on L(𝐚){\rm{L}}({\bf{a}}) naturally obtained from a weighted Sasakian structure of S2n+1S^{2n+1}.

Refer to caption
Figure 1. Link of isolated singularity which represents a Brieskorn manifold L(𝐚){\rm{L}}({\bf{a}}).

In this setting, we have the following commutative diagram:

L(𝐚){{\rm{L}}({\bf{a}})}S2n+1{S^{2n+1}}Xorb(𝐚){X^{{\rm{orb}}}({\bf{a}})}(𝐰){\mathbbm{P}({\bf{w}})}

where (𝐰)=(n+1{0})/×\mathbbm{P}({\bf{w}})=(\mathbbm{C}^{n+1}-\{0\})/\mathbbm{C}^{\times} is the weighted projective space defined by the action described in Eq. 4.13, here we denote 𝐰=(Ca0,,Can){\bf{w}}=(\frac{C}{a_{0}},\ldots,\frac{C}{a_{n}}). In the above diagram the horizontal arrows are Sasakian and Kählerian embeddings, respectively, and the vertical arrows are principal S1S^{1} V-bundles and orbifold Riemannian submersions. Moreover, L(𝐚){\rm{L}}({\bf{a}}) is the total space of principal S1S^{1} V-bundle over the orbifold Xorb(𝐚)X^{{\rm{orb}}}({\bf{a}}) whose the first Chern class in Horb2(Xorb(𝐚),)H^{2}_{orb}(X^{{\rm{orb}}}({\bf{a}}),\mathbbm{Z}) is 1Ic1(Xorb(𝐚))\frac{1}{I}c_{1}(X^{{\rm{orb}}}({\bf{a}})), where II is the Fano index of Xorb(𝐚)X^{{\rm{orb}}}({\bf{a}}), see for instance [5]. Thus, in a similar way as in the previous example, we have that L(𝐚){\rm{L}}({\bf{a}}) admits a Sasaki-Einstein metric if and only if the orbifold Xorb(𝐚)X^{{\rm{orb}}}({\bf{a}}) admits a Kähler-Einstein orbifold metric of scalar curvature 4n(n+1)4n(n+1). From the Lichnerowicz obstruction [20, Eq. (3.23)] and the recent result provided in [29], see also [7, Conjecture 4], it follows that L(𝐚){\rm{L}}({\bf{a}}) admits a Sasaki-Einstein metric if

1<j=0n1aj<1+nan.1<\sum_{j=0}^{n}\frac{1}{a_{j}}<1+\frac{n}{a_{n}}. (4.15)

In this case, we have from Theorem 3.3 (or Corollary 3.6) that L(𝐚)×S1{\rm{L}}({\bf{a}})\times S^{1} admits a Levi-Civita Ricci-Flat Hermitian metric. Also, from Theorem 3.8, we obtain a huge class of new examples of complex manifolds with nonnegative first Chern class that admit constant strictly negative Riemannian scalar curvature.

Example 4.5.

The construction presented in the last example above plays an important role in the study homotopy spheres and exotic spheres. Being more precise, in [25] Hirzebruch showed that links as in Eq. (4.14) can sometimes be homeomorphic, but not diffeomorphic to standard spheres. As shown by Egbert V. Brieskorn in [9], links of the form

L(2,2,2,3,6k1)=(z02+z12+z22+z33+z46k1=0)S9,{\rm{L}}(2,2,2,3,6k-1)=\Big{(}z_{0}^{2}+z_{1}^{2}+z_{2}^{2}+z_{3}^{3}+z_{4}^{6k-1}=0\Big{)}\cap S^{9}, (4.16)

with k=1,,28k=1,\ldots,28, realize explicitly all the distinct smooth structures on the 77-sphere S7S^{7} classified by Kervaire and Milnor in [27], see also [30]. Denoting by 𝚺7{\bf{\Sigma}}^{7} any one of the 28 homotopy 77-sphere, it follows from Theorem 3.3 (or Corollary 3.7) that 𝚺7×S1{\bf{\Sigma}}^{7}\times S^{1} admits a Levi-Civita Ricci-Flat Hermitian metric. Moreover, from Theorem 3.8, we have a family of Hermitian metrics gζg_{\zeta} on 𝚺7×S1{\bf{\Sigma}}^{7}\times S^{1}, such that ζ>1\zeta>-1 (see Eq. 3.7). In this case, it follows from Eq. (3.31) that

scal(gζ)=12(ζ+1)2(ζ+78),ζ>1.{\rm{scal}}(g_{\zeta})=\frac{12}{(\zeta+1)^{2}}\bigg{(}\zeta+\frac{7}{8}\bigg{)},\ \ \ \ \ \ \forall\zeta>-1. (4.17)

Thus, we have scal(gζ)(,24]{\rm{scal}}(g_{\zeta})\in(-\infty,24]. In particular, we obtain the following:

  1. (1)

    ζ>78(𝚺7×S1,gζ,J)\zeta>-\frac{7}{8}\Longrightarrow({\bf{\Sigma}}^{7}\times S^{1},g_{\zeta},J) has constant strictly positive Riemannian scalar curvature;

  2. (2)

    ζ=78(𝚺7×S1,gζ,J)\zeta=-\frac{7}{8}\Longrightarrow({\bf{\Sigma}}^{7}\times S^{1},g_{\zeta},J) has constant zero Riemannian scalar curvature;

  3. (3)

    1<ζ<78(𝚺7×S1,gζ,J)-1<\zeta<-\frac{7}{8}\Longrightarrow({\bf{\Sigma}}^{7}\times S^{1},g_{\zeta},J) has constant strictly negative Riemannian scalar curvature.

Notice that scal(gζ){\rm{scal}}(g_{\zeta})\to-\infty as ζ1\zeta\to-1. According to [11], the existence of an exotic 88-sphere 𝚺8{\bf{\Sigma}}^{8} implies the existence of 30 different differentiable structures on S7×S1S^{7}\times S^{1}, i.e., besides those 28 obtained from 𝚺7×S1{\bf{\Sigma}}^{7}\times S^{1}, we also have an additional one obtained from (𝚺7×S1)#𝚺8({\bf{\Sigma}}^{7}\times S^{1})\#{\bf{\Sigma}}^{8}, where 𝚺7{\bf{\Sigma}}^{7} is any one of the 28 homotopy 77-spheres. It would be interesting to know whether the above results also hold for (𝚺7×S1)#𝚺8({\bf{\Sigma}}^{7}\times S^{1})\#{\bf{\Sigma}}^{8}.

Remark 4.6.

Another source of examples which illustrate the results of Theorem 3.3 and Theorem 3.8 is provided by manifolds of the form k×S1\mathcal{M}_{k}\times S^{1}, such that k=k#(S2×S3)\mathcal{M}_{k}=k\#(S^{2}\times S^{3}), k1k\geq 1. Actually, it is known that k\mathcal{M}_{k}, k1k\geq 1, admits infinitely many Sasaki-Einstein structures, see for instance [6], [37]. In particular, from Theorem 3.3 one can always solve the equation (1)(Ω)=0\mathfrak{R}^{(1)}(\Omega)=0 on k×S1\mathcal{M}_{k}\times S^{1}, k1k\geq 1.

Appendix A Remarks on Chern-Ricci flow

On a Hermitian manifold (M,Ω0,J)(M,\Omega_{0},J), a solution of the Chern-Ricci flow starting at Ω0\Omega_{0} is given by a smooth family of Hermitian metrics Ω=Ω(t)\Omega=\Omega(t) satisfying

{tΩ=Ric(1)(Ω), 0t<T,Ω(0)=Ω0,\begin{cases}\displaystyle\frac{\partial}{\partial t}\Omega=-{\rm{Ric}}^{(1)}(\Omega),\ \ 0\leq t<T,\\ \Omega(0)=\Omega_{0},\end{cases} (A.1)

for T(0,]T\in(0,\infty], see for instance [21] and [34]. Inspired by the results on Hopf manifolds provided in [34], [22], and [35], we observe that the ideas involved in the proof of Theorem 3.3 can be used to obtain explicit solutions of the Chern-Ricci flow on compact Hermitian Weyl-Einstein manifolds. More precisely, we have the following result:

Theorem A.1.

Let (M,g,J)(M,g,J) be a compact Hermitian Weyl-Einstein manifold, then there exists an explicit solution g(t)g(t) of the Chern-Ricci flow on MM for t[0,2n)t\in[0,\frac{2}{n}), starting at gg, satisfying the following properties:

  1. (1)

    Vol(M,g(t))0{\rm{Vol}}(M,g(t))\to 0 as t2nt\to\frac{2}{n} (i.e. g(t)g(t) is finite-time collapsing);

  2. (2)

    limt2ng(t)=hT\lim_{t\to\frac{2}{n}}g(t)=h_{T}, where hTh_{T} is a nonnegative symmetric tensor on MM;

  3. (3)

    The Chern scalar curvature of g(t)g(t) blows up like (n1)/(2nt)(n-1)/(\frac{2}{n}-t);

  4. (4)

    scal(g(t)){\rm{scal}}(g(t))\to-\infty as t2nt\to\frac{2}{n};

  5. (5)

    limt2ndGH((M,dt),(S1,dS1))=0\lim_{t\to\frac{2}{n}}d_{GH}\big{(}(M,d_{t}),(S^{1},d_{S^{1}})\big{)}=0,

where dtd_{t} is the distance induced by g(t)g(t) on MM and dS1d_{S^{1}} is the distance on the unit circle S1S^{1} induced by a suitable scalar multiple of the standard Riemannian metric.

Remark A.2.

In order to prove Theorem A.1, we proceed highlighting the background on Sasaki-Einstein geometry underlying all our main results. In fact, in what follows we shall prove the results of Theorem A.1 for compact Hermitian manifolds as in the proof of Theorem 3.3, i.e., for compact Hermitian manifolds of the form (Σϕ,c(Q),g,J)(\Sigma_{\phi,c}(Q),g,J), such that gg is a Hermitian Weyl-Einstein metric induced by the Calabi-Yau structure of 𝒞(Q)\mathcal{C}(Q). We also include in the proof some comments relating our results with previous known results on Hopf manifolds.

Before we start the proof, we recall that the Gromov-Hausdorff distance of metric spaces can be defined as follows (see for instance [12]). Given two metric spaces (X,dX)(X,d_{X}) and (Y,dY)(Y,d_{Y}), a correspondence between the underlying sets XX and YY is a subset RX×YR\subseteq X\times Y satisfying the following property: for every xXx\in X there exists at leas one yYy\in Y, such that (x,y)X×Y(x,y)\in X\times Y, and similarly for every yYy\in Y there exists an xXx\in X, such that (x,y)X×Y(x,y)\in X\times Y. Let us denote by (X,Y)\mathcal{R}(X,Y) the set of all correspondences between XX and YY. Now we consider the following definition

Definition A.3.

Let R(X,Y)R\in\mathcal{R}(X,Y) be a correspondence between two metric spaces (X,dX)(X,d_{X}) and (Y,dY)(Y,d_{Y}). The distortion of RR is defined by

dis(R)=sup{|dX(x,x)dY(y,y)||(x,y),(x,y)R}.{\rm{dis}}(R)=\sup\Big{\{}|d_{X}(x,x^{\prime})-d_{Y}(y,y^{\prime})|\ \ \big{|}\ \ (x,y),(x^{\prime},y^{\prime})\in R\ \Big{\}}. (A.2)

From above, we can define the Gromov-Hausdorff distance between two metric spaces as follows.

Definition A.4.

We define the Gromov-Hausdorff distance of any two metric spaces (X,dX)(X,d_{X}) and (Y,dY)(Y,d_{Y}) as being

dGH((X,dX),(Y,dY))=12inf{dis(R)|R(X,Y)}.d_{GH}\big{(}(X,d_{X}),(Y,d_{Y})\big{)}=\frac{1}{2}\inf\Big{\{}{\rm{dis}}(R)\ \ \Big{|}\ \ R\in\mathcal{R}(X,Y)\Big{\}}. (A.3)

Now we can prove Theorem A.1.

Proof.

(Theorem A.1) As we have seen, given a compact Hermitian Weyl-Einstein manifold of the form (Σϕ,c(Q),g,J)(\Sigma_{\phi,c}(Q),g,J), it follows that

Ω=d(Jθ)+θJθandRic(1)(Ω)=n2d(Jθ).\Omega=-{\rm{d}}(J\theta)+\theta\wedge J\theta\ \ \ {\text{and}}\ \ \ {\rm{Ric}}^{(1)}(\Omega)=-\frac{n}{2}{\rm{d}}(J\theta). (A.4)

Therefore, we set Ω0:=Ω\Omega_{0}:=\Omega and

Ω(t):=Ω0tRic(1)(Ω0)=(1n2t)d(Jθ)+θJθ.\Omega(t):=\Omega_{0}-t{\rm{Ric}}^{(1)}(\Omega_{0})=-\Big{(}1-\frac{n}{2}t\Big{)}{\rm{d}}(J\theta)+\theta\wedge J\theta. (A.5)

Since d(Jθ)0-{\rm{d}}(J\theta)\geq 0, we have that Ω(t)\Omega(t) is a Hermitian metric for all t[0,2n)t\in[0,\frac{2}{n}). Moreover, a straightforward computation shows that

Ω(t)n=(1n2t)n1Ω0n.\Omega(t)^{n}=\Big{(}1-\frac{n}{2}t\Big{)}^{n-1}\Omega_{0}^{n}. (A.6)

Thus, we have Ric(1)(Ω(t))=Ric(1)(Ω0){\rm{Ric}}^{(1)}(\Omega(t))={\rm{Ric}}^{(1)}(\Omega_{0}), t[0,2n)\forall t\in[0,\frac{2}{n}). From this, we conclude that

tΩ(t)=Ric(1)(Ω0)=Ric(1)(Ω(t)),\frac{\partial}{\partial t}\Omega(t)=-{\rm{Ric}}^{(1)}(\Omega_{0})=-{\rm{Ric}}^{(1)}(\Omega(t)), (A.7)

for all t[0,2n)t\in[0,\frac{2}{n}). Therefore, the family of Hermitian metrics Ω(t)=Ω0tRic(1)(Ω0)\Omega(t)=\Omega_{0}-t{\rm{Ric}}^{(1)}(\Omega_{0}) on Σϕ,c(Q)\Sigma_{\phi,c}(Q) gives a solution of the Chern-Ricci flow on the maximal existence interval [0,T)[0,T), such that T=2nT=\frac{2}{n}. Let us observe that

dΩ(t)=θ(t)Ω(t),such that θ(t)=TTtθ,{\rm{d}}\Omega(t)=\theta(t)\wedge\Omega(t),\ \ {\text{such that }}\ \ \theta(t)=\frac{T}{T-t}\theta, (A.8)

for all t[0,T)t\in[0,T). Moreover, one can easily verify that θ(t)\theta(t) is parallel with respect to the Levi-Civita connection induced by Ω(t)\Omega(t). Therefore, we have that Ω(t)\Omega(t) is Vaisman for all t[0,T)t\in[0,T). Also, from Eq. (A.6) and Eq. (A.5), it follows that

limtTVol(Σϕ,c(Q),Ω(t))=0andlimtTΩ(t)=θJθ,\lim_{t\to T}{\rm{Vol}}\big{(}\Sigma_{\phi,c}(Q),\Omega(t)\big{)}=0\ \ \ {\text{and}}\ \ \ \lim_{t\to T}\Omega(t)=\theta\wedge J\theta, (A.9)

i.e., Ω(t)\Omega(t) is finite-time collapsing [35]. Observing that ΩT:=θJθ\Omega_{T}:=\theta\wedge J\theta is a nonnegative (1,1)(1,1)-form, we conclude that the behavior of the Chern-Ricci flow Ω(t)\Omega(t) is quite similar to the behavior of the Chern-Ricci flow on Hopf manifolds provided in [34, Porposition 1.8]. Further, as it was shown in [22], finite-time singularities are characterized by the blow-up of the Chern scalar curvature. This last fact can be easily verified for the Chern-Ricci flow describe in Eq. (A.5). In fact, by a straightforward computation one can show that

d(Jθ)Ω(t)n1=(n1)!(n11n2t)Ω(t)nn!.{\rm{d}}(J\theta)\wedge\Omega(t)^{n-1}=-(n-1)!\bigg{(}\frac{n-1}{1-\frac{n}{2}t}\bigg{)}\frac{\Omega(t)^{n}}{n!}. (A.10)

Thus, since Ric(1)(Ω(t))=Ric(1)(Ω0){\rm{Ric}}^{(1)}(\Omega(t))={\rm{Ric}}^{(1)}(\Omega_{0}), t[0,T)\forall t\in[0,T), we obtain the following

sC(Ω(t))nΩ(t)n=n2d(Jθ)Ω(t)n1=12(n11n2t)Ω(t)n.\frac{s_{C}(\Omega(t))}{n}\Omega(t)^{n}=-\frac{n}{2}{\rm{d}}(J\theta)\wedge\Omega(t)^{n-1}=\frac{1}{2}\bigg{(}\frac{n-1}{1-\frac{n}{2}t}\bigg{)}\Omega(t)^{n}. (A.11)

From above we have

sC(Ω(t))=n2(n11n2t)=n1Tt,s_{C}(\Omega(t))=\frac{n}{2}\bigg{(}\frac{n-1}{1-\frac{n}{2}t}\bigg{)}=\frac{n-1}{T-t}, (A.12)

for all 0t<T0\leq t<T. It follows that sC(Ω(t))+s_{C}(\Omega(t))\to+\infty as tTt\to T (cf. [22]). Also, denoting by g(t)g(t) the underlying Riemannian metric associated to Ω(t)\Omega(t), it follows from Eq. (3.31) that

scal(g(t))=n(n1)(1n2t)2[2n12nn2t]=n(n1)(Tt)2[T1n2t],{\rm{scal}}(g(t))=\frac{n(n-1)}{(1-\frac{n}{2}t)^{2}}\bigg{[}\frac{2n-1}{2n}-\frac{n}{2}t\bigg{]}=\frac{n(n-1)}{(T-t)^{2}}\bigg{[}T-\frac{1}{n^{2}}-t\bigg{]}, (A.13)

for all 0t<T0\leq t<T. Hence, we obtain scal(g(t)){\rm{scal}}(g(t))\to-\infty as tTt\to T. In particular, we have:

  1. (a)

    0t<T1n2scal(g(t))>00\leq t<T-\frac{1}{n^{2}}\Longrightarrow{\rm{scal}}(g(t))>0;

  2. (b)

    t=T1n2scal(g(t))=0t=T-\frac{1}{n^{2}}\Longrightarrow{\rm{scal}}(g(t))=0;

  3. (c)

    T1n2<t<Tscal(g(t))<0T-\frac{1}{n^{2}}<t<T\Longrightarrow{\rm{scal}}(g(t))<0.

From above, we obtain a complete picture of the behavior of scal(g(t)){\rm{scal}}(g(t)), t[0,T)t\in[0,T). Further, by means of a suitable change in the argument presented in [35, §\S 4] one can show that

(Σϕ,c(Q),dt)G.H.(S1,dS1),astT,\big{(}\Sigma_{\phi,c}(Q),d_{t}\big{)}\xrightarrow{\text{G.H.}}\big{(}S^{1},d_{S^{1}}\big{)},\ \ {\text{as}}\ \ t\to T, (A.14)

where dtd_{t} is the distance induced by g(t)g(t) and dS1d_{S^{1}} is the distance on the unit circle S1S^{1} induced by a suitable scalar multiple of the standard Riemannian metric. From Definition A.3, we say that (Σϕ,c(Q),dt)G.H.(S1,dS1)\big{(}\Sigma_{\phi,c}(Q),d_{t}\big{)}\xrightarrow{\text{G.H.}}\big{(}S^{1},d_{S^{1}}\big{)}, as tTt\to T, if

limtTdGH((Σϕ,c(Q),dt),(S1,dS1))=0.\lim_{t\to T}d_{GH}\big{(}(\Sigma_{\phi,c}(Q),d_{t}),(S^{1},d_{S^{1}})\big{)}=0. (A.15)

In order to conclude the proof, consider F:Q×F\colon Q\times\mathbbm{R}\to\mathbbm{R}, such that F(x,φ)=2φF(x,\varphi)=-2\varphi, (x,φ)Q×\forall(x,\varphi)\in Q\times\mathbbm{R}. From this, we set Fc:Σϕ,c(Q)S1F_{c}\colon\Sigma_{\phi,c}(Q)\to S^{1}, such that

Fc([x,φ]):=exp(2π1F(x,φ)log(c2)),F_{c}([x,\varphi]):=\exp\bigg{(}\frac{2\pi\sqrt{-1}F(x,\varphi)}{\log(c^{2})}\bigg{)}, (A.16)

for all [x,φ]Σϕ,c(Q)[x,\varphi]\in\Sigma_{\phi,c}(Q). Since

F(ϕ(x),φ+nlog(c))=F(x,φ)nlog(c2),F(\phi(x),\varphi+n\log(c))=F(x,\varphi)-n\log(c^{2}), (A.17)

it follows that FcF_{c} is a well-defined smooth map. Considering the canonical angular 11-form dσ{\rm{d}}\sigma on S1S^{1}, a straightforward computation shows us that

Fc(dσ)=θlog(c2).F_{c}^{\ast}({\rm{d}}\sigma)=\frac{\theta}{\log(c^{2})}. (A.18)

From above, since θ\theta is a non-vanishing 11-form, it follows that FcF_{c} is a submersion. Also, we notice that ker((Fc))=ker(θ(t))\ker((F_{c})_{\ast})=\ker(\theta(t)), 0t<T\forall 0\leq t<T. Since

g(t)=(1n2t)d(Jθ)(𝟙J)h(t)+θθ+JθJθhT,g(t)=\underbrace{-\Big{(}1-\frac{n}{2}t\Big{)}{\rm{d}}(J\theta)(\mathbbm{1}\otimes J)}_{h(t)}+\underbrace{\theta\otimes\theta+J\theta\otimes J\theta}_{h_{T}}, (A.19)

see Eq. (A.5), by considering the g(t)g(t)-orthogonal complement ker((Fc))\ker((F_{c})_{\ast})^{\perp}, it follows that

g(t)|ker((Fc))=θθ=λ2Fc(dσdσ),g(t)|_{\ker((F_{c})_{\ast})^{\perp}}=\theta\otimes\theta=\lambda^{2}F_{c}^{\ast}\big{(}{\rm{d}}\sigma\otimes{\rm{d}}\sigma\big{)}, (A.20)

where λ=log(c2)\lambda=\log(c^{2}), i.e., Fc:(Σϕ,c(Q),g(t))(S1,λ2dσdσ)F_{c}\colon(\Sigma_{\phi,c}(Q),g(t))\to(S^{1},\lambda^{2}{\rm{d}}\sigma\otimes{\rm{d}}\sigma) is a Riemannian submersion. Since Σϕ,c(Q)\Sigma_{\phi,c}(Q) is compact, we have that Fc:Σϕ,c(Q)S1F_{c}\colon\Sigma_{\phi,c}(Q)\to S^{1} is in fact a locally trivial fiber bundle with typical fiber diffeomorphic to QQ (e.g. [4]). From hTh_{T} given in Eq. (A.19), we set

𝒟:={XTΣϕ,c(Q)|hT(X,Y)=0,Y}.\mathcal{D}:=\big{\{}X\in T\Sigma_{\phi,c}(Q)\ \big{|}\ h_{T}(X,Y)=0,\ \forall Y\big{\}}. (A.21)

Let us denote by SΣϕ,c(Q)S\subset\Sigma_{\phi,c}(Q) a generic fiber of FcF_{c}. Denote also by 𝒟S\mathcal{D}_{S} the distribution 𝒟\mathcal{D} restricted to SS. By construction, since ker((Fc))=ker(θ)\ker((F_{c})_{\ast})=\ker(\theta), we have that (S,ηS:=i(Jθ))(S,\eta_{S}:=i^{\ast}(J\theta)) is a contact manifold, where i:SΣϕ,c(Q)i\colon S\hookrightarrow\Sigma_{\phi,c}(Q) is the natural inclusion, see for instance [15]. Hence, we obtain the following description

𝒟S={XTS|ηS(X)=0},\mathcal{D}_{S}=\big{\{}X\in TS\ \big{|}\ \eta_{S}(X)=0\big{\}}, (A.22)

i.e. 𝒟S\mathcal{D}_{S} is the contact distribution on SS induced by ηS\eta_{S}. Since every contact distribution is a bracket-generating distribution, it follows from Chow’s theorem [31] that any two points of SS can be connected by a smooth path tangent to 𝒟S\mathcal{D}_{S}. It is worth observing that g(t)|𝒟S=h(t)|𝒟Sg(t)|_{\mathcal{D}_{S}}=h(t)|_{\mathcal{D}_{S}}, where h(t)h(t) is given as in Eq. (A.19). Given p,qΣϕ,c(Q)p,q\in\Sigma_{\phi,c}(Q), let α\alpha be a path connecting Fc(p)F_{c}(p) and Fc(q)F_{c}(q) in S1S^{1}. Since (Σϕ,c(Q),g(t))(\Sigma_{\phi,c}(Q),g(t)) is complete, there exists a lift α~\widetilde{\alpha} of α\alpha starting at pp and tangent to ker((Fc))\ker((F_{c})_{\ast})^{\perp}. Let us denote by zΣϕ,c(Q)z\in\Sigma_{\phi,c}(Q) the endpoint of the path α~\widetilde{\alpha} (Figure 2). From this, we have dt(p,z)dS1(Fc(p),Fc(z))=dS1(Fc(p),Fc(q))d_{t}(p,z)\leq d_{S^{1}}(F_{c}(p),F_{c}(z))=d_{S^{1}}(F_{c}(p),F_{c}(q)), where dS1d_{S^{1}} denotes the distance induced by λ2dσdσ\lambda^{2}{\rm{d}}\sigma\otimes{\rm{d}}\sigma on S1S^{1}. Hence, we obtain

dt(p,q)dt(p,z)+dt(z,q)dS1(Fc(p),Fc(q))+dt(z,q).d_{t}(p,q)\leq d_{t}(p,z)+d_{t}(z,q)\leq d_{S^{1}}(F_{c}(p),F_{c}(q))+d_{t}(z,q). (A.23)

Since z,qFc1(Fc(q))=Sz,q\in F_{c}^{-1}(F_{c}(q))=S, we have a smooth path γ:[0,1]Σϕ,c(Q)\gamma\colon[0,1]\to\Sigma_{\phi,c}(Q) tangent to 𝒟S\mathcal{D}_{S} connecting zz and qq (Figure 2).

Refer to caption
Figure 2. Representation of horizontal and vertical paths used to estimate the distance between pp and qq.

Therefore, since g(t)|𝒟S=h(t)|𝒟Sg(t)|_{\mathcal{D}_{S}}=h(t)|_{\mathcal{D}_{S}} and Σϕ,c(Q)\Sigma_{\phi,c}(Q) is compact, we obtain

dt(z,q)01γ(s)g(t)𝑑s=(1n2t)01γ(s)g(0)𝑑sC(1n2t),d_{t}(z,q)\leq\int_{0}^{1}||\gamma^{\prime}(s)||_{g(t)}ds=\Big{(}1-\frac{n}{2}t\Big{)}\int_{0}^{1}||\gamma^{\prime}(s)||_{g(0)}ds\leq C\Big{(}1-\frac{n}{2}t\Big{)}, (A.24)

where C:=diam(Σϕ,c(Q),g(0))C:={\rm{diam}}\big{(}\Sigma_{\phi,c}(Q),g(0)\big{)}. Hence, it follows that

dt(p,q)dS1(Fc(p),Fc(q))C(1n2t),d_{t}(p,q)-d_{S^{1}}(F_{c}(p),F_{c}(q))\leq C\Big{(}1-\frac{n}{2}t\Big{)}, (A.25)

for all t[0,T)t\in[0,T) and for all p,qΣϕ,c(Q)p,q\in\Sigma_{\phi,c}(Q). From Fc:Σϕ,c(Q)S1F_{c}\colon\Sigma_{\phi,c}(Q)\to S^{1}, we set

R(Fc):={(p,Fc(p))Σϕ,c(Q)×S1|pΣϕ,c(Q)}.R(F_{c}):=\big{\{}(p,F_{c}(p))\in\Sigma_{\phi,c}(Q)\times S^{1}\ \big{|}\ p\in\Sigma_{\phi,c}(Q)\big{\}}. (A.26)

Since FcF_{c} is a surjective map, it follows that R(Fc)(Σϕ,c(Q),S1)R(F_{c})\in\mathcal{R}(\Sigma_{\phi,c}(Q),S^{1}). From Eq. (A.25), we conclude that

dGH((Σϕ,c(Q),dt),(S1,dS1))12dis(R(Fc))<nC2(Tt).d_{GH}\big{(}(\Sigma_{\phi,c}(Q),d_{t}),(S^{1},d_{S^{1}})\big{)}\leq\frac{1}{2}{\rm{dis}}(R(F_{c}))<\frac{nC}{2}\Big{(}T-t\Big{)}. (A.27)

Therefore, it follows that

limtTdGH((Σϕ,c(Q),dt),(S1,dS1))=0.\lim_{t\to T}d_{GH}\big{(}(\Sigma_{\phi,c}(Q),d_{t}),(S^{1},d_{S^{1}})\big{)}=0. (A.28)

As it can be seen, the arguments provided above generalize certain ideas introduced in [35, §4] for Hopf manifolds. Following Theorem 2.26 and the above constructions, we conclude the proof of Theorem A.1. ∎

Example A.5.

Consider 𝚺7×S1{\bf{\Sigma}}^{7}\times S^{1}, where 𝚺7{\bf{\Sigma}}^{7} is any one of the 28 homotopy 77-spheres. Fixed a Hermitian Weyl-Einstein metric ΩWE\Omega_{WE} on 𝚺7×S1{\bf{\Sigma}}^{7}\times S^{1}, it follows from Theorem A.1 that there exists a family of Hermitian metrics Ω(t)\Omega(t), t[0,12)t\in[0,\frac{1}{2}), satisfying

{tΩ=Ric(1)(Ω), 0t<12,Ω(0)=ΩWE.\begin{cases}\displaystyle\frac{\partial}{\partial t}\Omega=-{\rm{Ric}}^{(1)}(\Omega),\ \ 0\leq t<\frac{1}{2},\\ \Omega(0)=\Omega_{WE}.\end{cases} (A.29)

Moreover, it follows form Eq. A.12 and from Eq. A.13 that

sC(Ω(t))=612t,andscal(g(t))=48(12t)2[716t].s_{C}(\Omega(t))=\frac{6}{1-2t},\ \ \ \ {\text{and}}\ \ \ \ {\rm{scal}}(g(t))=\frac{48}{(1-2t)^{2}}\bigg{[}\frac{7}{16}-t\bigg{]}. (A.30)

From above, we obtain the following:

  1. (a)

    0t<716scal(g(t))>00\leq t<\frac{7}{16}\Longrightarrow{\rm{scal}}(g(t))>0;

  2. (b)

    t=716scal(g(t))=0t=\frac{7}{16}\Longrightarrow{\rm{scal}}(g(t))=0;

  3. (c)

    716<t<12scal(g(t))<0\frac{7}{16}<t<\frac{1}{2}\Longrightarrow{\rm{scal}}(g(t))<0.

Regarding 𝚺7×S1{\bf{\Sigma}}^{7}\times S^{1} as a suspension by (id,e)({\rm{id}},\sqrt{{\rm{e}}}) of 𝚺7{\bf{\Sigma}}^{7}, from Theorem A.1 we conclude that

limt12dGH((𝚺7×S1,dt),(S1,dS1))=0,\lim_{t\to\frac{1}{2}}d_{GH}\big{(}({\bf{\Sigma}}^{7}\times S^{1},d_{t}),(S^{1},d_{S^{1}})\big{)}=0, (A.31)

where dtd_{t} is the distance induced by g(t)g(t) on 𝚺7×S1{\bf{\Sigma}}^{7}\times S^{1} and dS1d_{S^{1}} is the distance on the unit circle S1S^{1} induced by the standard Riemannian metric.

Conflict of interest statement

The author declares that there is no conflict of interest.

References

  • [1] Abe, K.; On a generalization of the Hopf fibration, I. Contact structures on the generalized Brieskorn manifolds. Tohoku Math. J. (2) 29 (3) 335-374, 1977.
  • [2] Angella, D.; Cohomological Aspects in Complex Non-Kähler Geometry. Lecture Notes in Mathematics, 2095. Springer, Cham, 2014.
  • [3] Bazzoni, G.; Marrero, J. C.; On locally conformal symplectic manifolds of the first kind, Bull. Sci. Math. 143, 1-57 (2018).
  • [4] Besse, A.; Einstein manifolds, Springer 1987.
  • [5] Boyer, C. P.; Galicki, K.; New Einstein metrics in dimension five. J. Differential Geom., 57 (3): 443-463, 2001.
  • [6] Boyer, C. P.; Galicki, K.; Sasakian Geometry, Oxford Mathematical Monographs, Oxford University Press; 1 edition (2008).
  • [7] Boyer, C. P.; Galicki, K.; Kollár. J.; Einstein metrics on spheres. Ann. of Math. (2), 162 (1): 557-580, 2005.
  • [8] Boyer, C. P.; Galicki, K.; 3-Sasakian manifolds, Surveys in differential geometry: essays on Einstein manifolds, 123-184. Surv. Differ Geom., VI, Int. Press, Boston (1999).
  • [9] Brieskorn, E. V.; Beispiele zur Differentialtopologie von Singularitäten, Inventiones Mathematicae, 2 (1): 1-14 (1966).
  • [10] Brieskorn, E. V.; Examples of singular normal complex spaces which are topological manifolds, Proceedings of the National Academy of Sciences, 55 (6): 1395-1397 (1966).
  • [11] Brieskorn, E. V.; van de Ven, A.; Some complex structures on products of homotopy spheres, Topology 7 (1968), 389-393.
  • [12] Burago, D.; Burago, Y.; Ivanov, S.; A Course in Metric Geometry, Graduate Studies in Mathematics, vol. 33. American Mathematical Society, Providence (2001).
  • [13] Correa, E. M.; Principal elliptic bundles and compact homogeneous l.c.K. manifolds, arXiv:1904.10099v2 (2022).
  • [14] Dabkowski, M. G.; Lock, M. T.; An Equivalence of Scalar Curvatures on Hermitian Manifolds. Jour. Geom. Anal. 27 (1), 239-270 (2017).
  • [15] Dragomir, S.; Ornea, L.; Locally Conformal Kähler Geometry; Progress in Mathematics, Birkhäuser; 1998 edition (1997).
  • [16] Frédéric, P.; Formules de Picard-Lefschetz généralisées et ramification des intégrales, Bulletin de la Société Mathématique de France, 93: 333-367 (1965).
  • [17] Gauduchon P., Hermitian connections and Dirac operators, Boll. Un. Mat. Ital. B 11 (1997), 257-288.
  • [18] Gauduchon, P.; La 1-forme de torsion d’une variété e hermitienne compacte, Math. Ann. 267 (1984), 495-518.
  • [19] Gauduchon, P.; Structures de Weyl-Einstein, espaces de twisteurs et variété de type S1×S3S^{1}\times S^{3}, J. Reine Angew. Math. 469 (1995), 1-50.
  • [20] Gauntlett, J. P.; Martelli, D.; Sparks, J.; Yau, S.-T.; Obstructions to the existence of Sasaki-Einstein metrics, Comm. Math. Phys. 273 (2007), no. 3, 803-827.
  • [21] Gill, M.; Convergence of the parabolic complex Monge-Ampère equation on compact Hermitian manifolds, Comm. Anal. Geom. 19 (2011), no. 2, 277-303.
  • [22] Gill, M.; Smith, D.; The behavior of Chern scalar curvature under Chern-Ricci flow. Proc. Amer. Math. Soc. 143, no. 11 (2015): 4875-83.
  • [23] Gini, R.; Ornea, L.; Parton, M.; Locally conformal Kähler reduction. J. Reine Angew. Math., 581:1-21, 2005.
  • [24] Gray, A.; Hervella, L. M.; The sixteen classes of almost Hermitian manifolds and their linear invariants, Ann. Mat. Pura Appl. (4) 123 (1980), 35-58.
  • [25] Hirzebruch, F.; Singularities and exotic spheres. Séminaire Bourbaki, Vol. 10, Soc. Math. France, Paris, 1995, pp. 13-32.
  • [26] Kashiwada, T.; A note on a Riemannian space with Sasakian 3-structure, Nat. Sci. Rep. Ochanomizu Univ. , 22 (1971) pp. 1-2.
  • [27] Kervaire, M.; Milnor, J.; Groups of Homotopy Spheres I, Ann. Math. 77 (1963) 504.
  • [28] Liu, K.; Yang, X.; Ricci curvatures on Hermitian manifolds, Trans. Amer. Math. Soc. 369 (2017), no. 7, 5157-5196.
  • [29] Liu, Y.; Sano, T.; Tasin. L.; Infinitely many families of Sasaki-Einstein metrics on spheres, arXiv:2203.08468 (2022).
  • [30] Milnor, J.; On manifolds homeomorphic to the seven sphere, Ann. of Math. vol. 64 (1956) pp. 399-405.
  • [31] Montgomery, R.; A tour of subriemannian geometries, their geodesics and applications, vol. 91 of Mathematical Surveys and Monographs (American Mathematical Society, Providence, Rhode Island, 2002).
  • [32] Sparks, J.; Sasaki-Einstein manifolds, Surveys Diff. Geom. 16 (2011) 265.
  • [33] Takahashi, T.; Deformations of Sasakian structures and its application to the Brieskorn manifolds. Tohoku Math. J. (2) 30 (1) 37-43, 1978.
  • [34] Tosatti, V.; Weinkove, B.; On the evolution of a Hermitian metric by its Chern-Ricci form, J. Differential Geom. 99 (2015), no.1, 125-163.
  • [35] Tosatti, V.; Weinkove, B.; The Chern-Ricci flow on complex surfaces, Compos. Math. 149 (2013), no. 12, 2101-2138.
  • [36] Vaisman, I.; Generalized Hopf manifolds, Geom. Dedicata 13 (1982), 231-255.
  • [37] van Coevering, C.; Sasaki-Einstein 5-manifolds associated to toric 3-Sasaki manifolds, New York J. Math., 18 (2012), 555-608.
  • [38] Verbitsky, M.; Vanishing theorems for locally conformal hyperkähler manifolds, Proc. of Steklov Institute, 246 (2004), 54-79.
  • [39] Wells, R.; Differential Analysis on Complex Manifolds, GTM 65, Springer, New York, 2008.
  • [40] Yau, S.-T.; On the Ricci curvature of a compact Kähler manifold and the complex Monge-Ampère equation. I, Comm. Pure Appl. Math. 31 (1978), no. 3, 339–411. MR480350.