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

Diameter theorems on Kähler and quaternionic Kähler manifolds under a positive lower curvature bound

Maria Gordina Department of Mathematics
University of Connecticut
Storrs, CT 06269, U.S.A.
maria.gordina@uconn.edu
 and  Gunhee Cho Department of Mathematics
University of California, Santa Barbara
Santa Barbara, CA 93106
gunhee.cho@math.ucsb.edu
(Date:  File:Manuscript˙20210730.tex)
Abstract.

We define the orthogonal Bakry-Émery tensor as a generalization of the orthogonal Ricci curvature, and then study diameter theorems on Kähler and quaternionic Kähler manifolds under positivity assumption on the orthogonal Bakry-Émery tensor. Moreover, under such assumptions on the orthogonal Bakry-Émery tensor and the holomorphic or quaternionic sectional curvature on a Kähler manifold or a quaternionic Kähler manifold respectively, the Bonnet-Myers type diameter bounds are sharper than in the Riemannian case.

Key words and phrases:
Bakry-Émery Ricci tensor, Bonnet-Myers theorem, Kähler and quaternionic Kähler Laplacian comparison theorems
2010 Mathematics Subject Classification:
Primary:53C21, Secondary:58J65
11footnotemark: 1{\dagger} Research was supported in part by NSF Grant DMS-1954264.

1. Introduction

The Bonnet–Myers theorem in [Myers1941] is a fundamental theorem connecting compactness and upper bounds on the diameter of a complete Riemannian manifold. A typical result relies on the assumption of a positive lower bound of the Ricci curvature, or a positivity assumption of the Bakry-Émery Ricci tensor such as [AndersonM1990, Gromov1981a, Otsu1991, BakryQian2000, WeiWylie2009, Limoncu2012, Limoncu2010]. Such results connecting Ricci curvature or its generalizations and Bonnet-Myers type theorems have been studied for Riemannian manifolds, and on sub-Riemannian manifolds in [BaudoinGarofalo2017, BaudoinGrongKuwadaThalmaier2019]. Our focus is on the setting of Kähler geometry and quaternionic Kähler geometry for which such results have not been considered so far.

We will rely on the decomposition of Ricci curvature on Kähler manifolds into orthogonal Ricci curvature and holomorphic sectional curvature as follows.

Ric(X,X¯)=Ric(X,X¯)+R(X,X¯,X,X¯)/|X|4,\textup{Ric}(X,\overline{X})=\textup{Ric}^{\perp}(X,\overline{X})+R(X,\overline{X},X,\overline{X})/|X|^{4},

where XX is a (1,0)(1,0)-tangent vector of the holomorphic tangent bundle on a Kähler manifold MnM^{n}. A similar decomposition holds on a quaternionic Kähler geometry. Precise definitions are given by (2.1) and (2.3). Using such a decomposition, it is natural to consider a Bakry-Émery tensor in both Kähler geometry and quaternionic Kähler geometry similarly to how the Ricci curvature is replaced by a Bakry-Émery tensor in Riemannian geometry. The goal of this paper is to study diameter theorems for compact Kähler manifolds and quaternionic Kähler manifolds under various notions of positive lower bounds on the orthogonal Bakry-Émery type tensor corresponding to the orthogonal Ricci curvature.

First, we note that replacing the positivity of the orthogonal Ricci curvature with a weaker notion of positivity is justified by the following observation. Indeed, the class of complete Kähler manifolds with a positive orthogonal Ricci curvature is rather small although an orthogonal Ricci curvature bound is usually weaker than a Ricci curvature bound. For example, a complete Kähler manifold Mn,n2M^{n},n\geqslant 2 with a positive lower bound on the orthogonal Ricci curvature must be compact and always projective [NiZheng2018, Theorem 1.7]. Moreover, for n=2n=2, a compact M2M^{2} which admits a Kähler metric with Ric>0\textup{Ric}^{\perp}>0 must be biholomorphic to the two-dimensional complex projective space 2\mathbb{P}_{\mathbb{C}}^{2}, and for n=3n=3, a compact Kähler manifold under Ric>0\textup{Ric}^{\perp}>0 must be biholomorphic to either 3\mathbb{P}_{\mathbb{C}}^{3} or the smooth quadratic hypersurface in 4\mathbb{P}_{\mathbb{C}}^{4} as pointed out by [LeiQingsongFangyang2018, Theorem 1.8]. On the other hand, as there is an example of complete non-compact Riemannian manifold with a non-negative Ricci curvature lower bound, there are also complete non-compact examples of Kähler manifolds with Ric>0\textup{Ric}^{\perp}>0 [NiZheng2018, p151]. Therefore in order to consider the complete Kähler manifolds of the wide class rather than the complete Kähler manifolds of the limited classes possible under the positive orthogonal Ricci curvature, it is reasonable to make at least weaker the positivity condition than the orthogonal Ricci curvature. On the other hand, similarly to the Bakry-Émery tensor in the Riemannian case, a complete Kähler manifold need not be compact even under if the orthogonal Bakry-Émery tensor satisfies a positive lower bound as shown in Example 1.

As the Ricci curvature can be written as a sum of orthogonal Ricci curvature and holomorphic (respectively, quaternionic) sectional curvature for Kähler (respectively, quaternionic Kähler) manifolds, we consider two versions of an orthogonal Bakry-Émery tensor corresponding to the orthogonal Ricci curvature. In the first case we consider the orthogonal Bakry-Émery tensor Ric+Hess(ϕ)\textup{Ric}^{\perp}+\operatorname{Hess}(\phi), where ϕ\phi is a real-valued smooth function on MM and Hess(ϕ)\operatorname{Hess}(\phi) is the Riemannian Hessian. That is, we omit the holomorphic (respectively, quaternionic) sectional curvature. Note that here we only consider the Bakry-Émery tensor of a gradient type Hess(ϕ)\operatorname{Hess}(\phi). Similarly to how compactness and diameter have been treated on complete Riemannian manifolds in [Limoncu2012, Limoncu2010], here too, additional smoothness assumptions are needed to obtain the results.

Another approach to compactness and diameter bounds’ results using Ricci curvature or Bakry-Émery tensor assumptions on a Riemannian manifold is to rely on Bochner’s formula. We are exploring such an approach for complete Kähler manifolds with an orthogonal Ricci curvature bound or its generalization. For this purpose, we derive a new Bochner’s formula in Proposition 6 for the orthogonal Ricci curvature, and then establish such results under the assumptions which are compatible with this Bochner’s formula. The second case is to consider a non-gradient type Bakry-Émery type tensor Ricm,Z\textup{Ric}^{\perp}_{m,Z} defined by (4.1) with a vector field ZZ and an additional assumption on the holomorphic sectional curvature (quaternionic sectional curvature in the case of quaternionic Kähler manifolds). In the previous case, the second-order differential operator in Bochner’s formula for the orthogonal Ricci curvature is not hypoelliptic in general, whereas in the second case, as in the Riemannian case, we use the Laplace-Beltrami operator making it possible to use a weaker positivity assumption than in the first case when we replaced the orthogonal Ricci curvature by a Bakry-Émery tensor. In order to show the diameter upper bound, we follow Kuwada’s approach in [Kuwada2013a] and consider a stochastic process with this operator as a generator that might be non-symmetric. We then prove an upper bound on the diameter which is sharper than the diameter upper bound in the Riemannian case.

This paper is organized as follows. In Section 2 we introduce basic definitions and properties of Kähler manifolds and quaternionic Kähler manifolds, and in particular, how these structures are connected to their Riemannian structures. In Section 3, diameter theorems are covered under the Bakry-Émery orthogonal Ricci tensor of the gradient type. In Section 4 we prove diameter theorems for a non-gradient Bakry-Émery tensor under the additional assumption on the holomorphic (quaternionic) sectional curvature.

2. Preliminaries on Kähler and quaternionic Kähler manifolds

We start by reviewing basics of Kähler and quaternionic manifolds.

2.1. Kähler manifolds

Let MM be an nn-dimensional complex manifold equipped with a complex structure JJ and a Hermitian metric gg. The complex structure J:TMTMJ:T_{\mathbb{R}}M\rightarrow T_{\mathbb{R}}M is a real linear endomorphism that satisfies for every xMx\in M, and X,YT,xMX,Y\in T_{\mathbb{R},x}M, gx(JxX,Y)=gx(X,JxY)g_{x}(J_{x}X,Y)=-g_{x}(X,J_{x}Y), and for every xMx\in M, Jx2=𝐈𝐝TxMJ_{x}^{2}=-\mathbf{Id}_{T_{x}M}. We decompose the complexified tangent bundle TM=TMTM¯T_{\mathbb{R}}M\otimes_{\mathbb{R}}\mathbb{C}=T^{\prime}M\oplus\overline{T^{\prime}M}, where TMT^{\prime}M is the eigenspace of JJ with respect to the eigenvalue 1\sqrt{-1} and TM¯\overline{T^{\prime}M} is the eigenspace of JJ with respect to the eigenvalue 1-\sqrt{-1}. We can identify v,wv,w as real tangent vectors, and η,ξ\eta,\xi as corresponding holomorphic (1,0)(1,0) tangent vectors under the \mathbb{R}-linear isomorphism TMTMT_{\mathbb{R}}M\rightarrow T^{\prime}M, i.e. η=12(v1Jv),,ξ=12(w1Jw)\eta=\frac{1}{{\sqrt{2}}}(v-\sqrt{-1}Jv),,\xi=\frac{1}{{{\sqrt{2}}}}(w-\sqrt{-1}Jw).

A Hermitian metric on MM is a positive definite Hermitian inner product

gp:TpMTpM¯g_{p}:T^{\prime}_{p}M\otimes\overline{T^{\prime}_{p}M}\rightarrow\mathbb{C}

which varies smoothly for each pMp\in M. Here, varying smoothly means that if z=(z1,,zn)z=(z_{1},\cdots,z_{n}) are local coordinates around pp, and z1,,zn\frac{\partial}{\partial z_{1}},\cdots,\frac{\partial}{\partial z_{n}} is a standard basis for TpMT^{\prime}_{p}M, the functions

gij¯:U,pgp(zi,zj¯)g_{i\overline{j}}:U\rightarrow\mathbb{C},p\mapsto g_{p}(\frac{\partial}{\partial z_{i}},\frac{\partial}{\partial\overline{z_{j}}})

are smooth for all i,j=1,,ni,j=1,\cdots,n. Locally, a Hermitian metric can be written as

g=i,j=1ngij¯dzidzj¯,g=\sum_{i,j=1}^{n}g_{i\overline{j}}dz_{i}\otimes d\overline{z_{j}},

where (gij¯)(g_{i\overline{j}}) is an n×nn\times n positive definite Hermitian matrix of smooth functions and dz1,,dzndz_{1},\cdots,dz_{n} be the dual basis of z1,,zn\frac{\partial}{\partial z_{1}},\cdots,\frac{\partial}{\partial z_{n}}. The metric gg can be decomposed into the real part denoted by Re(g)\operatorname{Re}(g), and the imaginary part, denoted by Im(g)\operatorname{Im}(g). Re(g)\operatorname{Re}(g) induces an inner product called the induced Riemannian metric of gg, an alternating \mathbb{R}-differential 22-form. Define the (1,1)(1,1)-form ω:=12Im(g)\omega:=-\frac{1}{2}Im(g), which is called the fundamental (1,1)(1,1)-form of gg. In local coordinates this form can written as

ω=12i,j=1ngij¯dzidzj¯.\omega=\frac{\sqrt{-1}}{2}\sum_{i,j=1}^{n}g_{i\overline{j}}dz_{i}\wedge d\overline{z_{j}}.

In this setting we have two natural connections. The Chern connection c\nabla^{c} is compatible with the Hermitian metric gg and the complex structure JJ, and the Levi-Civita connection \nabla is a torsion free connection compatible with the induced Riemannian metric. The components of the curvature 44-tensor of the Chern connection associated with the Hermitian metric gg are given by

Rij¯kl¯:=R(zi,zi,zi,zi)\displaystyle R_{i\overline{j}k\overline{l}}:=R(\frac{\partial}{\partial z_{i}},\frac{\partial}{\partial z_{i}},\frac{\partial}{\partial z_{i}},\frac{\partial}{\partial z_{i}})
=g(ziczj¯czkzj¯cziczk[zi,zj¯]czk,zl¯)\displaystyle=g\left(\nabla^{c}_{\frac{\partial}{\partial z_{i}}}\nabla^{c}_{\frac{\partial}{\partial\overline{z_{j}}}}\frac{\partial}{\partial z_{k}}-\nabla^{c}_{\frac{\partial}{\partial\overline{z_{j}}}}\nabla^{c}_{\frac{\partial}{\partial z_{i}}}\frac{\partial}{\partial z_{k}}-\nabla^{c}_{[{\frac{\partial}{\partial z_{i}}},{\frac{\partial}{\partial\overline{z_{j}}}}]}{\frac{\partial}{\partial z_{k}}},{\frac{\partial}{\partial\overline{z_{l}}}}\right)
=2gij¯zkz¯l+p,q=1ngqp¯gip¯zkgqj¯z¯l,\displaystyle=-\frac{\partial^{2}g_{i\overline{j}}}{\partial z_{k}\partial\overline{z}_{l}}+\sum_{p,q=1}^{n}g^{q\overline{p}}\frac{\partial g_{i\overline{p}}}{\partial z_{k}}\frac{\partial g_{q\overline{j}}}{\partial\overline{z}_{l}},

where i,j,k,l{1.,n}i,j,k,l\in\left\{1.\cdots,n\right\}.

The Hermitian metric gg is called Kähler if dω=0d\omega=0, where dd is the exterior derivative d=+¯d=\partial+\overline{\partial}, and the Chern and Levi-Civita connections coincide precisely when the Hermitian metric is Kähler. There are several equivalent ways to show that a metric is Kähler, and one of them is that a metric gg is Kähler if and only if for any pMp\in M, there exist holomorphic coordinates (z1,,zn)(z_{1},\cdots,z_{n}) near pp such that gij¯(p)=δij¯g_{i\overline{j}}(p)=\delta_{i\overline{j}} and (dgij¯)(p)=0(dg_{i\overline{j}})(p)=0. Such coordinates are called holomorphic normal coordinates.

The holomorphic sectional curvature with the unit direction η\eta at xMx\in M (i.e., gω(η,η)=1g_{\omega}(\eta,\eta)=1) is defined by

H(g)(x,η)=R(η,η¯,η,η¯)=R(v,Jv,Jv,v),H(g)(x,\eta)=R(\eta,\overline{\eta},\eta,\overline{\eta})=R(v,Jv,Jv,v),

where vv is the real tangent vector corresponding to η\eta. We will often write H(g)(x,η)=H(g)(η)=H(η)H(g)(x,\eta)=H(g)(\eta)=H(\eta).

Following [NiZheng2019] we define the orthogonal Ricci curvature on a Kähler manifold (M,g,J)(M,g,J) by

Ric(v,v)=Ric(v,v)H(v),\operatorname{Ric}^{\perp}(v,v)=\operatorname{Ric}(v,v)-H(v), (2.1)

where vv is a real vector field and Ric\mathrm{Ric} is the Ricci 22-tensor of (M,g)(M,g). Unlike the Ricci tensor, Ric\operatorname{Ric}^{\perp} does not admit polarization, so we never consider Ric(u,v)\operatorname{Ric}^{\perp}(u,v) for uvu\not=v. For a real vector field vv, we can write

Ric(v,v)=R(v,Ei,Ei,v),{Ric}^{\perp}(v,v)=\sum R(v,E_{i},E_{i},v),

where {ei}\left\{e_{i}\right\} is any orthonormal frame of {v,Jv}\left\{v,Jv\right\}^{\perp}. We will assign index 1,21,2 to vv and JvJv in this summation expression for complex nn dimensional Kähler manifold MnM^{n}, and use indices from 33 to 2n2n for orthonormal frames {Ei}\left\{E_{i}\right\} of {v,Jv}\left\{v,Jv\right\}^{\perp}. Denote by Fi=12(Ei1J(Ei))F_{i}=\frac{1}{\sqrt{2}}(E_{i}-\sqrt{-1}J(E_{i})) a unitary frame such that E1=v/|v|=:v~E_{1}=v/|v|=:\widetilde{v} by following the convention En+i=J(Ei)E_{n+i}=J(E_{i}), then

1|v|2Ric(v,v)\displaystyle\frac{1}{|v|^{2}}\mathrm{Ric}^{\perp}(v,v) =Ric(v~,v~)=Ric(v~,v~)R(v~,Jv~,v~,Jv~)\displaystyle=\mathrm{Ric}^{\perp}(\widetilde{v},\widetilde{v})=\mathrm{Ric}(\widetilde{v},\widetilde{v})-R(\widetilde{v},J\widetilde{v},\widetilde{v},J\widetilde{v})
=Ric(F1,F1¯)R(F1,F1¯,F1,F1¯)=j=2nR(F1,F1¯,Fj,Fj¯).\displaystyle=\mathrm{Ric}(F_{1},\overline{F_{1}})-R(F_{1},\overline{F_{1}},F_{1},\overline{F_{1}})=\sum_{j=2}^{n}R(F_{1},\overline{F_{1}},F_{j},\overline{F_{j}}).

In particular, we have Ric(Fi,Fi¯)=Ric(Ei,Ei)\mathrm{Ric}(F_{i},\overline{F_{i}})=\mathrm{Ric}(E_{i},E_{i}), Ric(v~,v~)=Ric(F1,F1¯)R11¯11¯\mathrm{Ric}^{\perp}(\widetilde{v},\widetilde{v})=\mathrm{Ric}(F_{1},\overline{F_{1}})-R_{1\overline{1}1\overline{1}}.

2.2. Quaternionic Kähler manifolds

We start by recalling the following definition of quaternionic Kähler manifold following [BesseBook2008, Proposition 14.36].

Definition 1.

A Riemannian manifold (M,g)(M,g) is called a quaternionic Kähler manifold if there exists a covering of MM by open sets UiU_{i} and, for each ii there exist 3 smooth (1,1)(1,1) tensors I,J,KI,J,K on UiU_{i} such that

  • -

    For every xUix\in U_{i}, and X,YTxMX,Y\in T_{x}M, gx(IxX,Y)=gx(X,IxY)g_{x}(I_{x}X,Y)=-g_{x}(X,I_{x}Y), gx(JxX,Y)=gx(X,JxY)g_{x}(J_{x}X,Y)=-g_{x}(X,J_{x}Y), gx(KxX,Y)=gx(X,KxY)g_{x}(K_{x}X,Y)=-g_{x}(X,K_{x}Y) ;

  • -

    For every xUix\in U_{i}, Ix2=Jx2=Kx2=IxJxKx=𝐈𝐝TxMI_{x}^{2}=J_{x}^{2}=K_{x}^{2}=I_{x}J_{x}K_{x}=-\mathbf{Id}_{T_{x}M};

  • -

    For every xUix\in U_{i}, and XTxMX\in T_{x}M XI,XJ,XK𝐬𝐩𝐚𝐧{I,J,K}\nabla_{X}I,\nabla_{X}J,\nabla_{X}K\in\mathbf{span}\{I,J,K\};

  • -

    For every xUiUjx\in U_{i}\cap U_{j}, the vector space of endomorphisms of TxMT_{x}M generated by Ix,Jx,KxI_{x},J_{x},K_{x} is the same for ii and jj.

Note that in some cases such as the quaternionic projective spaces the tensors I,J,KI,J,K may not be defined globally for topological reasons. However, 𝐬𝐩𝐚𝐧{I,J,K}\mathbf{span}\{I,J,K\} may always be defined globally according to Definition 1.

On quaternionic Kähler manifolds, we will be considering the following curvature tensors. As above, let

R(X,Y,Z,W)=g((XYYX[X,Y])Z,W)R(X,Y,Z,W)=g((\nabla_{X}\nabla_{Y}-\nabla_{Y}\nabla_{X}-\nabla_{[X,Y]})Z,W) (2.2)

be the Riemannian curvature tensor of (M,g)(M,g). We define the quaternionic sectional curvature of the quaternionic Kähler manifold (M,g,J)(M,g,J) as

Q(X)=R(X,IX,IX,X)+R(X,JX,JX,X)+R(X,KX,KX,X)g(X,X)2.Q(X)=\frac{R(X,IX,IX,X)+R(X,JX,JX,X)+R(X,KX,KX,X)}{g(X,X)^{2}}.

Following [BaudoinYang2020, Section 2.1.2] we define the orthogonal Ricci curvature of the quaternionic Kähler manifold (M,g,I,J,K)(M,g,I,J,K) by

Ric(X,X)=Ric(X,X)Q(X),\mathrm{Ric}^{\perp}(X,X)=\mathrm{Ric}(X,X)-Q(X), (2.3)

where Ric\mathrm{Ric} is the usual Riemannian Ricci tensor of (M,g)(M,g) and XX is a vector field such that g(X,X)=1g(X,X)=1.

Lastly, given a vector field VV on a Riemannian manifold along a geodesic γ\gamma defined on [a,b][a,b], the index form \mathcal{I} associated to VV is defined as

(V,V)=ab(|V˙(s)|2R(V(s),γ˙(s),γ˙(s),V(s)))𝑑s,\mathcal{I}(V,V)=\int_{a}^{b}\left(|\dot{V}(s)|^{2}-R(V(s),\dot{\gamma}(s),\dot{\gamma}(s),V(s))\right)ds,

and using polarization the form \mathcal{I} can be extended to a bilinear form on the space of vector fields along the geodesic γ\gamma.

3. Bakry-Émery orthogonal Ricci tensor of the gradient type

Given a Riemannian manifold (M,g)(M,g) and a smooth function ϕ:M\phi:M\rightarrow\mathbb{R}, we denote the Hessian of ϕ\phi by Hess(ϕ)\operatorname{Hess}(\phi), i.e., Hess(ϕ)(X,Y)=g(Xϕ,Y)\operatorname{Hess}(\phi)(X,Y)=g(\nabla_{X}\nabla\phi,Y) for any real vector fields X,YX,Y. In this section, we define and consider the orthogonal Bakry-Émery tensor Ric+Hess(ϕ)\textup{Ric}^{\perp}+\operatorname{Hess}(\phi) with a smooth function ϕ\phi on either a Kähler manifold or a quaternionic Kähler manifold. On a Kähler manifold MnM^{n} with kk\in\mathbb{R}, we say Ric+Hess(ϕ)(2n2)k\textup{Ric}^{\perp}+\operatorname{Hess}(\phi)\geqslant(2n-2)k if for any unit vector vv

Ric(v,v)+Hess(ϕ)(v,v)(2n2)k,\textup{Ric}^{\perp}(v,v)+\mathrm{\operatorname{Hess}}(\phi)(v,v)\geqslant(2n-2)k,

and similarly, we assume Ric(v,v)+Hess(ϕ)(v,v)(4n4)k\textup{Ric}^{\perp}(v,v)+\mathrm{\operatorname{Hess}}(\phi)(v,v)\geqslant(4n-4)k for a quaternionic Kähler manifold. The Bakry-Émery tensor considered in this section is different from such a tensor in Section 4. Indeed, the modified Bochner’s formula in Proposition 6 shows the relationship between orthogonal Laplacian and orthogonal Ricci curvature, without any assumptions on holomorhpic (or quaternionic) sectional curvature. With the modified Bochner formula, assumptions on a smooth function ϕ\phi are important when trying to prove diameter theorems.

Previously Riemannian manifolds endowed with a weighted volume measure efdVge^{-f}dV_{g} satisfying a lower bound on the standard Bakry-Eḿery Ricci tensor has been studied in several settings. For example, a Riemannian manifold (M,g)(M,g) is called a gradient Ricci soliton if there exists a real-valued smooth function ff on MM such that the Ricci curvature and the Hessian of ff satisfy Ric+Hess(f)=λg\textup{Ric}+\operatorname{Hess}(f)=\lambda g for some λ\lambda\in\mathbb{R}. Gradient Ricci solitons play an important role in the theory of Ricci flow as in [Fernandez-LopezGarcia-Rio2008]. Bakry-Eḿery Ricci tensor plays fundamental role in [WeiWylie2009], and it has been extended to metric measure spaces using the Lott-Villani-Sturm theory initiated by [LottVillani2009, Sturm2006b]. In particular, if (M,g)(M,g) is a Kähler manifold and it is a gradient Ricci soliton with a real-valued smooth function ff, then f\nabla f is a real holomorphic vector field, i.e., its (1,0)(1,0)-part is a holomorphic vector field. Moreover, the weighted Hodge Laplacian from considered with respect to the weighted volume measure efdVge^{-f}dV_{g} maps the space of smooth (p,q)(p,q) forms to itself for 0<p+q<2n0<p+q<2n if and only if f\nabla f is a real holomorphic vector field by [OvidiuMunteanuJiapingWang15, Proposition 0.1]. Also, the real holomorphic vector field serves as a critical point of the Calabi functional which yields the Calabi’s extremal Kähler metric that can be used to study the existence of the Kähler-Einstein metric on Fano manifolds [EugenioCalabi82]. Lastly, on a compact Kähler manifold with Ricci curvature bounded below by a positive constant kk, if the first nonzero eigenvalue achieves its optimal lower bound 2k2k then the gradient vector field of the corresponding eigenfunction must be real holomorphic by [Udagawa1988].

In our case, we are interested in bounds on Ric+Hess(ϕ)\textup{Ric}^{\perp}+\operatorname{Hess}(\phi), where ϕ\phi is a real-valued smooth function, and the holomorphic sectional curvature is not controlled, hence it is different from the case where bounds on the Bakry-Émery tensor are used.

For compactness of a complete Riemannian manifold with an upper bound on the diameter, one needs additional assumptions on the function whose Hessian is used to define the Bakry-Émery Ricci tensor as pointed out by [WeiWylie2009]. The next example indicates that we need to address such an issue for the orthogonal Bakry-Émery tensor as well.

Examples 1.

Let M=nM=\mathbb{C}^{n} with the Euclidean metric gEg_{E}, and ϕ(x)=λ2|x|2\phi(x)=\frac{\lambda}{2}|x|^{2}. Then both the orthogonal Ricci curvature and the holomorphic sectional curvature are zero, and Hess(ϕ)=λgE\mathrm{\operatorname{Hess}}(\phi)=\lambda g_{E} and (Ric+Hess(ϕ))(v,v)=λgE(v,v)(\textup{Ric}^{\perp}+\mathrm{\operatorname{Hess}}(\phi))(v,v)=\lambda g_{E}(v,v) for any real vector field vv. This example shows that a complete Kähler manifold with positive lower bound of Ric+Hess(ϕ)\textup{Ric}^{\perp}+\mathrm{\operatorname{Hess}}(\phi) is not necessarily compact.

As we observed from the example, to prove that a complete Kähler (and also quaternionic Kähler) manifold with a positive lower bound on the orthogonal Bakry-Émery tensor is compact we need to have additional assumptions on ϕ\phi, and we will start with the case when ϕ\phi is bounded. In this case, we have the following elementary lemma to control such a function ϕ\phi. This fact has been used in the proof of [Limoncu2012, Theorem 1] for a Bakry-Émery Ricci tensor on Riemannian manifolds.

Lemma 2.

Let MM be a Riemannian manifold with a Riemannian metric gg and let ϕ:M\phi:M\rightarrow\mathbb{R} be a smooth function satisfying |ϕ|C|\phi|\leqslant C for some C0C\geqslant 0. Let γ\gamma be a minimizing unit speed geodesic segment from pp to qq of length ll. Then we have

0lf2Hess(ϕ)2Cl(0l(ddt(ff˙)2)𝑑t)1/2,\int_{0}^{l}f^{2}\mathrm{\operatorname{Hess}}(\phi)\leqslant 2C\sqrt{l}\left(\int_{0}^{l}(\frac{d}{dt}(f\dot{f})^{2})dt\right)^{1/2},

for any smooth function fC([0,l])f\in C^{\infty}([0,l]) such that f(0)=f(l)=0f(0)=f(l)=0. Here f˙(t)\dot{f}(t) means ddtf(γ(t))\frac{d}{dt}f(\gamma(t)).

Proof.

From

f(t)2Hess(ϕ)(γ˙,γ˙)(γ(t))=f(t)2ddt(g(ϕ,γ˙))(γ(t))\displaystyle f(t)^{2}\mathrm{\operatorname{Hess}}(\phi)({\dot{\gamma},\dot{\gamma}})(\gamma(t))=f(t)^{2}\frac{d}{dt}\left(g(\nabla\phi,\dot{\gamma})\right)(\gamma(t))
=2f(t)f(t)g(ϕ,γ˙)+ddt(f(t)2g(ϕ,γ˙))(γ(t))\displaystyle=-2f(t)f(t)^{\prime}g(\nabla\phi,\dot{\gamma})+\frac{d}{dt}\left(f(t)^{2}g(\nabla\phi,\dot{\gamma})\right)(\gamma(t))
=2ϕ(γ(t))ddt(f(t)f(t))(γ(t))2ddt(ϕf(t)f(t)(γ(t))+ddt(f(t)2g(ϕ,γ˙))(γ(t)),\displaystyle=2\phi(\gamma(t))\frac{d}{dt}(f(t)f(t^{\prime}))(\gamma(t))-2\frac{d}{dt}(\phi f(t)f(t^{\prime})(\gamma(t))+\frac{d}{dt}(f(t)^{2}g(\nabla\phi,\dot{\gamma}))(\gamma(t)),

and f(0)=f(l)=0f(0)=f(l)=0, we have

0lf2Hess(ϕ)(γ˙,γ˙))dt=20lϕddt(ff˙)dt.\displaystyle\int_{0}^{l}f^{2}\mathrm{\operatorname{Hess}}(\phi)({\dot{\gamma},\dot{\gamma}}))dt=2\int_{0}^{l}\phi\frac{d}{dt}(f\dot{f})dt.

Now the Cauchy-Schwarz inequality with the assumption |ϕ|C|\phi|\leqslant C implies

0lf2Hess(ϕ)=20lϕddt(ff˙)𝑑t2Cl(0l(ddt(ff˙)2)𝑑t)1/2.\int_{0}^{l}f^{2}\mathrm{\operatorname{Hess}}(\phi)=2\int_{0}^{l}\phi\frac{d}{dt}(f\dot{f})dt\leqslant 2C\sqrt{l}\left(\int_{0}^{l}(\frac{d}{dt}(f\dot{f})^{2})dt\right)^{1/2}.

Proposition 3.

Let (Mn,g)(M^{n},g) be a complete Kähler manifold with the complex dimension n2n\geqslant 2. Suppose that for some constant k>0k>0, Ric+Hess(ϕ)(2n2)k\textup{Ric}^{\perp}+\operatorname{Hess}(\phi)\geqslant(2n-2)k and |ϕ|C|\phi|\leqslant C for some C0C\geqslant 0. Then the diameter DD of MM has the upper bound

Dπk1+2Cn1.D\leqslant\frac{\pi}{\sqrt{k}}\sqrt{1+\frac{\sqrt{2}C}{n-1}}.
Proof.

Let p,qMp,q\in M and let γ\gamma be a minimizing unit speed geodesic segment from pp to qq of length ll. Consider a parallel orthonormal frame

{E1=γ˙,E2=JE1,,E2n}\left\{E_{1}=\dot{\gamma},E_{2}=JE_{1},\cdots,E_{2n}\right\}

along γ\gamma and a smooth function fC([0,l])f\in C^{\infty}([0,l]) such that f(0)=f(l)=0f(0)=f(l)=0. Here we used the Kähler condition J=0\nabla J=0 and parallel transport to have E2=JE1E_{2}=JE_{1}. From the definition of Ric\textup{Ric}^{\perp}, we have

i=32n(fEi,fEi)\displaystyle\sum_{i=3}^{2n}\mathcal{I}(fE_{i},fE_{i}) =0l((2n2)f˙2i=32nR(fEi,γ˙,γ˙,fEi)dt\displaystyle=\int_{0}^{l}((2n-2)\dot{f}^{2}-\sum_{i=3}^{2n}R(fE_{i},\dot{\gamma},\dot{\gamma},fE_{i})dt
=0l((2n2)f˙2f2Ric(γ˙,γ˙)dt.\displaystyle=\int_{0}^{l}((2n-2)\dot{f}^{2}-f^{2}\textup{Ric}^{\perp}(\dot{\gamma},\dot{\gamma})dt.

By the assumption on the orthogonal Bakry-Émery tensor,

i=32n(fEi,fEi)0l((2n2)(f˙2kf2)+f2Hess(ϕ)(γ˙,γ˙))𝑑t,\displaystyle\sum_{i=3}^{2n}\mathcal{I}(fE_{i},fE_{i})\leqslant\int_{0}^{l}((2n-2)(\dot{f}^{2}-kf^{2})+f^{2}\operatorname{Hess}(\phi)({\dot{\gamma},\dot{\gamma}}))dt,

where \mathcal{I} denotes the index form of γ\gamma. By Lemma 2, we have

i=32n(fEi,fEi)0l((2n2)(f˙2kf2)dt+2Cl(0l(ddt(ff˙)2)dt)1/2.\sum_{i=3}^{2n}\mathcal{I}(fE_{i},fE_{i})\leqslant\int_{0}^{l}((2n-2)(\dot{f}^{2}-kf^{2})dt+2C\sqrt{l}\left(\int_{0}^{l}(\frac{d}{dt}(f\dot{f})^{2})dt\right)^{1/2}.

Now, take ff to be f(t)=sin(πlt)f(t)=\sin(\frac{\pi}{l}t), then we get

i=32n(fEi,fEi)\displaystyle\sum_{i=3}^{2n}\mathcal{I}(fE_{i},fE_{i}) (2n2)0l(π2l2cos2(πlt)ksin2(πlt))𝑑t\displaystyle\leqslant(2n-2)\int_{0}^{l}\left(\frac{\pi^{2}}{l^{2}}\cos^{2}(\frac{\pi}{l}t)-k\sin^{2}(\frac{\pi}{l}t)\right)dt
+2Cπ2ll(0lcos2(2πlt)𝑑t)1/2,\displaystyle+\frac{2C\pi^{2}}{l\sqrt{l}}\left(\int_{0}^{l}\cos^{2}(\frac{2\pi}{l}t)dt\right)^{1/2},

and therefore

i=32n(fEi,fEi)1l((n1)kl22Cπ2(n1)π2).\sum_{i=3}^{2n}\mathcal{I}(fE_{i},fE_{i})\leqslant-\frac{1}{l}\left((n-1)kl^{2}-\sqrt{2}C\pi^{2}-(n-1)\pi^{2}\right).

Now if (n1)kl22Cπ2(n1)π2>0(n-1)kl^{2}-\sqrt{2}C\pi^{2}-(n-1)\pi^{2}>0, this forces (fEm,fEm)<0\mathcal{I}(fE_{m},fE_{m})<0 for some 3m2n3\leqslant m\leqslant 2n. On the other hand, since γ\gamma is a minimizing geodesic, the index form \mathcal{I} is positive semi-definite, which is a contradiction. Therefore,

lπk1+2Cn1.l\leqslant\frac{\pi}{\sqrt{k}}\sqrt{1+\frac{\sqrt{2}C}{n-1}}.

Remark 1.

By taking C=0C=0, we can conclude that a complete Kähler manifold MM with a positive lower bound on the orthogonal Ricci curvature implies compactness of MM and thereby implies that its fundamental group is finite by [NiZheng2018, Theorem 3.2].

The proof of Proposition 3 is easily modified to the quaternionic Kähler case.

Proposition 4.

Let (Mn,g,I,J,K)(M^{n},g,I,J,K) be a complete quaternionic Kähler manifold of the quaternionic dimension n2n\geqslant 2. Suppose that for some constant k>0k>0, Ric+Hess(ϕ)(4n4)k\textup{Ric}^{\perp}+\operatorname{Hess}(\phi)\geqslant(4n-4)k and |ϕ|C|\phi|\leqslant C for some C0C\geqslant 0. Then the diameter DD of MM satisfies the upper bound

Dπk1+2C2n2.D\leqslant\frac{\pi}{\sqrt{k}}\sqrt{1+\frac{\sqrt{2}C}{2n-2}}.
Proof.

We consider an orthonormal frame {X1(x),,X4m(x)}\{X_{1}(x),\cdots,X_{4m}(x)\} around xMx\in M such that

X1(x)=γ(0),X2(x)=Iγ(0),X3(x)=Jγ(0),X4(x)=Kγ(0)X_{1}(x)=\gamma^{\prime}(0),\,X_{2}(x)=I\gamma^{\prime}(0),\,X_{3}(x)=J\gamma^{\prime}(0),\,X_{4}(x)=K\gamma^{\prime}(0)

We introduce the function

𝔧(k,t)=coskt+1cosktsinktsinkt,\mathfrak{j}(k,t)=\cos\sqrt{k}t+\frac{1-\cos\sqrt{k}t}{\sin\sqrt{k}t}\sin\sqrt{k}t,

and we denote by X1,,X4mX_{1},\cdots,X_{4m} vector fields obtained by parallel transporting X1(x),,X4m(x)X_{1}(x),\cdots,X_{4m}(x) along γ\gamma and consider the vector fields defined along γ\gamma by

X~2(γ(t))=𝔧(4k,t)X2,X~3(γ(t))=𝔧(4k,t)X3,X~4(γ(t))=𝔧(4k,t)X4\tilde{X}_{2}(\gamma(t))=\mathfrak{j}(4k,t)X_{2},\,\tilde{X}_{3}(\gamma(t))=\mathfrak{j}(4k,t)X_{3},\tilde{X}_{4}(\gamma(t))=\mathfrak{j}(4k,t)X_{4}

and for i=5,,4mi=5,\cdots,4m by

X~i(γ(t))=𝔧(k,t)Xi.\tilde{X}_{i}(\gamma(t))=\mathfrak{j}(k,t)X_{i}.

Then the result follows by arguments similar to the previous proof. ∎

Using an argument similar to the case of a bounded function ϕ\phi in the summation of the index form in the proof of the Proposition 3, we can prove similar types of diameter theorems based on different assumptions on ϕ\phi. Let us just mention one setting that can easily replace Lemma 2. Given a vector field VV on a Riemannian manifold (M,g)(M,g), we denote the Lie derivative of VV by Vg\mathcal{L}_{V}g (see [Limoncu2010]).

Lemma 5.

Let MM be a Riemannian manifold with a Riemannian metric gg and let VV be the smooth vector field satisfying |V|C|V|\leqslant C for some C0C\geqslant 0. Let γ\gamma be a minimizing unit speed geodesic segment from pp to qq of length ll. Then we have

0lf2Vg(γ˙,γ˙)𝑑tCl(0lf2f˙2𝑑t)1/2,\int_{0}^{l}f^{2}\mathcal{L}_{V}g(\dot{\gamma},\dot{\gamma})dt\leqslant C\sqrt{l}\left(\int_{0}^{l}f^{2}\dot{f}^{2}dt\right)^{1/2},

for any smooth function fC([0,l])f\in C^{\infty}([0,l]) such that f(0)=f(l)=0f(0)=f(l)=0.

Now we will consider different assumptions on smooth functions ϕ\phi to apply the condition of Ric+Hess(ϕ)(2n2)k\textup{Ric}^{\perp}+\operatorname{Hess}(\phi)\geqslant(2n-2)k differently on both Kähler manifolds or quaternionic Kähler manifolds based on the modified Bochner type formula. The Bochner formula was used to control the Laplace-Beltrami operator under the Bakry-Émery tensor with the certain assumption on ϕ\phi [Limoncu2012], and we modify this approach. To do so, we first define the orthogonal Laplacian \triangle^{\perp} on Kähler manifolds (quaternion Kähler case would be similar) as follows: given any fixed point pp on a complete Kähler manifold (Mn,J)(M^{n},J) of complex dimension nn. For each real-valued smooth function ff, consider the holomorphic vector field Z=12(f1J(f))Z=\frac{1}{\sqrt{2}}(\nabla f-\sqrt{-1}J(\nabla f)) corresponding to the real vector field f\nabla f and define

f:=fHess(f)(Z,Z¯).\triangle^{\perp}f:=\triangle f-\operatorname{Hess}(f)(Z,\overline{Z}).

(also see [NiZheng2018, p.151]) Equivalently, with (Ei)i=12n(E_{i})^{2n}_{i=1} be an orthonormal frame with E1=fE_{1}=\nabla f and E2=JE1E_{2}=JE_{1},

f=fHess(f)(E1,E1)Hess(f)(JE1,JE1)=i=32nHess(f)(Ei,Ei).\triangle^{\perp}f=\triangle f-\operatorname{Hess}(f)(E_{1},E_{1})-\operatorname{Hess}(f)(JE_{1},JE_{1})=\sum_{i=3}^{2n}\operatorname{Hess}(f)(E_{i},E_{i}).

Similarly, the orthogonal Laplacian of a real-valued smooth function ff on a quaternionic Kähler manifold (M,I,J,K)(M,I,J,K) is defined by

f:\displaystyle\triangle^{\perp}f: =fHess(f)(E1,E1)Hess(f)(IE1,IE1)\displaystyle=\triangle f-\operatorname{Hess}(f)(E_{1},E_{1})-\operatorname{Hess}(f)(IE_{1},IE_{1})
Hess(f)(JE1,JE1)Hess(f)(KE1,KE1)\displaystyle-\operatorname{Hess}(f)(JE_{1},JE_{1})-\operatorname{Hess}(f)(KE_{1},KE_{1})
=i=52nHess(f)(Ei,Ei),\displaystyle=\sum_{i=5}^{2n}\operatorname{Hess}(f)(E_{i},E_{i}),

where (Ei)i=12n(E_{i})^{2n}_{i=1} is an orthonormal frame around pp with E2=IE1,E3=JE1,E4=KE1E_{2}=IE_{1},E_{3}=JE_{1},E_{4}=KE_{1}.

Since the orthogonal Laplacian is neither a self-adjoint operator nor a hypo-elliptic operator in general, it might be difficult to study this operator and its applications. Nevertheless, the modified Bochner type formula corresponding to the orthogonal Ricci curvature can be established. We hope this form may have several applications on the geometric analysis side in the future. In our paper, we use the following modified Bochner’s formula to use in Proposition 8. The idea of the proof is a Bochner’s formula modified to fit the orthogonal Ricci tensor.

Proposition 6 (Bochner’s formula: Kähler’s case).

Let (Mn,g,J)(M^{n},g,J) be a Kähler manifold of the complex dimension nn. Let ff be a real-valued smooth function on MM and \triangle^{\perp} be the orthogonal Laplacian. Then for any orthonormal frame (Ei)i=12n(E_{i})^{2n}_{i=1} around qMq\in M with E1=fE_{1}=\nabla{f}, E2=JE1E_{2}=JE_{1},

12i=32nEiEig(f,f)(q)=Ric(f,f)(q)+g(f,f)(q)\displaystyle\frac{1}{2}\sum^{2n}_{i=3}E_{i}E_{i}g(\nabla f,\nabla f)(q)=\textup{Ric}^{\perp}(\nabla f,\nabla f)(q)+g(\nabla f,\nabla\triangle^{\perp}f)(q)
+i=32ng(ff,EiEi))(q)+i=32n(g(Eif,Eif)2g(fEi,Eif))(q).\displaystyle+\sum^{2n}_{i=3}g(\nabla_{\nabla f}\nabla f,\nabla_{E_{i}}E_{i}))(q)+\sum_{i=3}^{2n}\left(g(\nabla_{E_{i}}\nabla f,\nabla_{E_{i}}\nabla f)-2g(\nabla_{\nabla f}{E_{i}},\nabla_{E_{i}}\nabla f)\right)(q).
Proof.

Let qMq\in M. Then at qq,

12i=32nEiEig(f,f)(q)=i=32nEig(Eif,f)(q)\displaystyle\frac{1}{2}\sum^{2n}_{i=3}E_{i}E_{i}g(\nabla f,\nabla f)(q)=\sum^{2n}_{i=3}E_{i}g(\nabla_{E_{i}}\nabla f,\nabla f)(q)
=i=32nEiHess(f)(Ei,f)(q)=i=32nEiHess(f)(f,Ei)(q)\displaystyle=\sum^{2n}_{i=3}E_{i}\operatorname{Hess}(f)(E_{i},\nabla f)(q)=\sum^{2n}_{i=3}E_{i}\operatorname{Hess}(f)(\nabla f,E_{i})(q)
=i=32nEig(ff,Ei)(q)=i=32n(g(Eiff,Ei)(q)+g(ff,EiEi))(q),\displaystyle=\sum^{2n}_{i=3}E_{i}g(\nabla_{\nabla f}\nabla f,E_{i})(q)=\sum^{2n}_{i=3}(g(\nabla_{E_{i}}\nabla_{\nabla f}\nabla f,E_{i})(q)+g(\nabla_{\nabla f}\nabla f,\nabla_{E_{i}}E_{i}))(q),

By using the Riemann curvature tensor defined by (2.2) we see that

i=32n(g(Eiff,Ei)(q)\displaystyle\sum^{2n}_{i=3}(g(\nabla_{E_{i}}\nabla_{\nabla f}\nabla f,E_{i})(q) (3.1)
=i=32n(g(R(Ei,f)f,Ei)+g(fEif,Ei)+g([Ei,f]f,Ei))(q)\displaystyle=\sum^{2n}_{i=3}\left(g(R(E_{i},\nabla f)\nabla f,E_{i})+g(\nabla_{\nabla f}\nabla_{E_{i}}\nabla f,E_{i})+g(\nabla_{[E_{i},\nabla f]}\nabla f,E_{i})\right)(q)
=Ric(f,f)(q)+i=32n(g(fEif,Ei)+g([Ei,f]f,Ei))(q).\displaystyle=\textup{Ric}^{\perp}(\nabla f,\nabla f)(q)+\sum^{2n}_{i=3}\left(g(\nabla_{\nabla f}\nabla_{E_{i}}\nabla f,E_{i})+g(\nabla_{[E_{i},\nabla f]}\nabla f,E_{i})\right)(q).

The second term i=32ng(fEif,Ei)\sum^{2n}_{i=3}g(\nabla_{\nabla f}\nabla_{E_{i}}\nabla f,E_{i}) in (3.1) can be written as

i=32n(fg(Eif,Ei)g(Eif,fEi))(q)\displaystyle\sum^{2n}_{i=3}\left(\nabla fg(\nabla_{E_{i}}\nabla f,E_{i})-g(\nabla_{E_{i}}\nabla f,\nabla_{\nabla f}E_{i})\right)(q)
=(fi=32ng(Eif,Ei))(q)i=32ng(Eif,fEi)\displaystyle=(\nabla f\sum^{2n}_{i=3}g(\nabla_{E_{i}}\nabla f,E_{i}))(q)-\sum^{2n}_{i=3}g(\nabla_{E_{i}}\nabla f,\nabla_{\nabla f}E_{i})
=f(f)i=32ng(Eif,fEi)=g(f,f)i=32ng(Eif,fEi).\displaystyle=\nabla f(\triangle^{\perp}f)-\sum^{2n}_{i=3}g(\nabla_{E_{i}}\nabla f,\nabla_{\nabla f}E_{i})=g(\nabla f,\nabla\triangle^{\perp}f)-\sum^{2n}_{i=3}g(\nabla_{E_{i}}\nabla f,\nabla_{\nabla f}E_{i}).

The last term g([Ei,f]f,Ei)(q)g(\nabla_{[E_{i},\nabla f]}\nabla f,E_{i})(q) in in (3.1) can be written as

i=32nHess(f)([Ei,f],Ei)(q)=i=32nHess(f)(EiffEi,Ei)(q)\displaystyle\sum^{2n}_{i=3}\operatorname{Hess}(f)([E_{i},\nabla f],E_{i})(q)=\sum^{2n}_{i=3}\operatorname{Hess}(f)(\nabla_{E_{i}}\nabla f-\nabla_{\nabla f}E_{i},E_{i})(q)
=i=32n(g(Eif,Eif)g(fEi,Eif))(q).\displaystyle=\sum_{i=3}^{2n}\left(g(\nabla_{E_{i}}\nabla f,\nabla_{E_{i}}\nabla f)-g(\nabla_{\nabla f}{E_{i}},\nabla_{E_{i}}\nabla f)\right)(q).

and we obtain the desired formula. ∎

Remark 2.

To compare with the usual Bochner’s formula, the formula in Proposition 6 uses the orthogonal Ricci tensor, the orthogonal Laplacian, and

i=32n(g(Eif,Eif)2g(fEi,Eif)+g(ff,EiEi))\sum^{2n}_{i=3}\left(g(\nabla_{E_{i}}\nabla f,\nabla_{E_{i}}\nabla f)-2g(\nabla_{\nabla f}{E_{i}},\nabla_{E_{i}}\nabla f)+g(\nabla_{\nabla f}\nabla f,\nabla_{E_{i}}E_{i})\right) (3.2)

instead of the Ricci tensor, the Laplace-Beltrami operator, and the Hessian norm squared of ff respectively. If we choose an orthonormal frame satisfying EiEj(q)=0\nabla_{E_{i}}E_{j}(q)=0 at some fixed point qq for any i,j=1,,2ni,j=1,\cdots,2n, then each of three terms in  (3.2) are zero at qq. However, this does not imply that the first two terms in  (3.2) vanish somewhere. For example, if we take ff to be the geodesic distance rr emanating from qMq\in M, then outside of the cut-locus of qq, one can see that limr0+rr=2n2\lim_{r\rightarrow 0+}r\triangle^{\perp}r=2n-2, but i=32ng(Eir,Eir)12n2(r)2\sum_{i=3}^{2n}g(\nabla_{E_{i}}\nabla r,\nabla_{E_{i}}\nabla r)\geqslant\frac{1}{2n-2}(\triangle^{\perp}r)^{2} (see the proof of Proposition 8), and by combining these two, i=32ng(Eir,Eir)\sum_{i=3}^{2n}g(\nabla_{E_{i}}\nabla r,\nabla_{E_{i}}\nabla r) cannot vanish in a small neighborhood of qq.

With similar computations, one can obtain the quaternionic version of the modified Bochner’s formula.

Proposition 7 (Bochner’s formula: quaternionic case).

Let (Mn,g,I,J,K)(M^{n},g,I,J,K) be a quaternionic Kähler manifold of the quaternionic dimension nn. Let ff be a real-valued smooth function on MM and \triangle^{\perp} be the orthogonal Laplacian. Then for any orthonormal frame (Ei)i=14n(E_{i})^{4n}_{i=1} around qq with E2=IE1,E3=JE1,E4=KE1E_{2}=IE_{1},E_{3}=JE_{1},E_{4}=KE_{1},

12|f|2(q)\displaystyle\frac{1}{2}\triangle^{\perp}|\nabla f|^{2}(q) =Ric(f,f)(q)+g(f,f)(q)+i=54ng(ff,EiEi))(q)\displaystyle=\textup{Ric}^{\perp}(\nabla f,\nabla f)(q)+g(\nabla f,\nabla\triangle^{\perp}f)(q)+\sum^{4n}_{i=5}g(\nabla_{\nabla f}\nabla f,\nabla_{E_{i}}E_{i}))(q)
+i=54n(g(Eif,Eif)g(fEi,Eif))(q).\displaystyle+\sum_{i=5}^{4n}\left(g(\nabla_{E_{i}}\nabla f,\nabla_{E_{i}}\nabla f)-g(\nabla_{\nabla f}{E_{i}},\nabla_{E_{i}}\nabla f)\right)(q).

By Bochner’s formula in Proposition 6 applied to the function ff being equal to the geodesic distance rr emanating from a point pMp\in M outside of the cut-locus of pp, although all terms g(rr,EiEi),i=3,,2ng(\nabla_{\nabla r}\nabla r,\nabla_{E_{i}}E_{i}),i=3,\cdots,2n vanish, we still need to control the term

2i=32ng(rEi,Eir)=2i=32nHess(r)(rEi,Ei).-2\sum_{i=3}^{2n}g(\nabla_{\nabla r}{E_{i}},\nabla_{E_{i}}\nabla r)=-2\sum_{i=3}^{2n}\operatorname{Hess}(r)(\nabla_{\nabla r}E_{i},E_{i}).

With this consideration, it would be natural to compensate the averaging effect i=32nHess(r)(rEi,Ei)\sum_{i=3}^{2n}\operatorname{Hess}(r)(\nabla_{\nabla r}E_{i},E_{i}) by the Hessian of some function ϕ\phi, for example,

Hess(ϕ)(r,r)2i=32nHess(r)(rEi,Ei).\operatorname{Hess}(\phi)(\nabla r,\nabla r)\geqslant 2\sum_{i=3}^{2n}\operatorname{Hess}(r)(\nabla_{\nabla r}E_{i},E_{i}).

By adding Hess(ϕ)\operatorname{Hess}(\phi) term to the orthogonal Ricci curvature with certain assumptions on ϕ\phi, we obtain the following diameter theorem. One can see that the assumptions on ϕ\phi in the Proposition below are analogous to the assumptions of Theorem 2 in [Limoncu2012] If we replace the complete Kähler manifold with Ric+Hess(ϕ)2i=32nHess(r)(rEi,Ei)\textup{Ric}^{\perp}+\mathrm{\operatorname{Hess}}(\phi)-2\sum_{i=3}^{2n}\operatorname{Hess}(r)(\nabla_{\nabla r}E_{i},E_{i}) by the Riemannian manifold with Ric+Hess(ϕ)\textup{Ric}+\operatorname{Hess}(\phi).

Proposition 8.

Let (Mn,g)(M^{n},g) be a complete Kähler manifold with the complex dimension n2n\geqslant 2. Take any pMp\in M and let rr be the geodesic distance function from pp. Suppose that for some constant k>0k>0, there exists a local orthonormal frame (Ei)i=12n(E_{i})^{2n}_{i=1} around pp with E1=rE_{1}=\nabla{r}, E2=JE1E_{2}=JE_{1} such that

Ric(r,r)+Hess(ϕ)(r,r)2i=32nHess(r)(rEi,Ei)(2n2)k\textup{Ric}^{\perp}(\nabla r,\nabla r)+\operatorname{Hess}(\phi)(\nabla r,\nabla r)-2\sum_{i=3}^{2n}\operatorname{Hess}(r)(\nabla_{\nabla r}E_{i},E_{i})\geqslant(2n-2)k

outside of the cut-locus of pp and |ϕ|2Cr(x)|\nabla\phi|^{2}\leqslant\frac{C}{r(x)} for some C0C\geqslant 0. Then the diameter DD of MM satisfies

Dπ(n1)k2C+n1.D\leqslant\frac{\pi}{\sqrt{(n-1)k}}\sqrt{2\sqrt{C}+n-1}.
Proof.

Define the modified orthogonal Laplacian

~f:=fg(ϕ,f)+F(f),\tilde{\triangle}^{\perp}f:=\triangle^{\perp}f-g(\nabla\phi,\nabla f)+F(f), (3.3)

where ff is a smooth function on MM and FF is a real-valued function taking values from \mathbb{R}. We will take f=rf=r, here rr is the distance function from the fixed point pMp\in M. Then outside of the cut-locus of pp,

g(r,~r)=g(r,~r)Hess(ϕ)(r,r)+F(r),F(r)=ddrF(r).g(\nabla r,\nabla\tilde{\triangle}^{\perp}r)=g(\nabla r,\nabla\tilde{\triangle}^{\perp}r)-\operatorname{Hess}(\phi)(\nabla r,\nabla r)+F^{\prime}(r),F^{\prime}(r)=\frac{d}{dr}F(r). (3.4)

On the other hand, from Proposition 6, since 12|r|20\frac{1}{2}\triangle^{\perp}|\nabla r|^{2}\equiv 0,

0=Ric(r,r)+g(r,r)+i=32n(g(Eir,Eir)+g(rEi,Eir)).0=\textup{Ric}^{\perp}(\nabla r,\nabla r)+g(\nabla r,\nabla\triangle^{\perp}r)+\sum_{i=3}^{2n}\left(g(\nabla_{E_{i}}\nabla r,\nabla_{E_{i}}\nabla r)+g(\nabla_{\nabla r}{E_{i}},\nabla_{E_{i}}\nabla r)\right).

From the Cauchy-Schwarz inequality, the term i=32ng(Eir,Eir)\sum_{i=3}^{2n}g(\nabla_{E_{i}}\nabla r,\nabla_{E_{i}}\nabla r) has the following lower bound:

i=32ng(Eir,Eir)=i=32nj=12ng(Eir,Ej)2i=32ng(Eir,Ei)2\displaystyle\sum_{i=3}^{2n}g(\nabla_{E_{i}}\nabla r,\nabla_{E_{i}}\nabla r)=\sum_{i=3}^{2n}\sum_{j=1}^{2n}g(\nabla_{E_{i}}\nabla r,E_{j})^{2}\geqslant\sum_{i=3}^{2n}g(\nabla_{E_{i}}\nabla r,E_{i})^{2}
=12n2i=32ng(Eir,Ei)2i=32n112n2(i=32ng(Eir,Ei))2=12n2(r)2,\displaystyle=\frac{1}{2n-2}\sum_{i=3}^{2n}g(\nabla_{E_{i}}\nabla r,E_{i})^{2}\sum_{i=3}^{2n}1\geqslant\frac{1}{2n-2}\left(\sum_{i=3}^{2n}g(\nabla_{E_{i}}\nabla r,E_{i})\right)^{2}=\frac{1}{2n-2}(\triangle^{\perp}r)^{2},

thus we have

0Ric(r,r)+g(r,r)+12n2(r)2,0\geqslant\textup{Ric}^{\perp}(\nabla r,\nabla r)+g(\nabla r,\nabla\triangle^{\perp}r)+\frac{1}{2n-2}(\triangle^{\perp}r)^{2}, (3.5)

outside of the cut-locus of pp.

From (3.5) and (3.4),

0Ric(r,r)+Hess(ϕ)(r,r)+g(r,r)+12n2(r)2,0\geqslant\textup{Ric}^{\perp}(\nabla r,\nabla r)+\operatorname{Hess}(\phi)(\nabla r,\nabla r)+g(\nabla r,\nabla{\triangle}^{\perp}r)+\frac{1}{2n-2}({\triangle}^{\perp}r)^{2}, (3.6)

Combining with (3.3), (3.6) becomes

0Ric(r,r)+Hess(ϕ)(r,r)+g(r,r)+12n2(~r+g(ϕ,r)F(r))2.0\geqslant\textup{Ric}^{\perp}(\nabla r,\nabla r)+\operatorname{Hess}(\phi)(\nabla r,\nabla r)+g(\nabla r,\nabla{\triangle}^{\perp}r)+\frac{1}{2n-2}(\tilde{\triangle}^{\perp}r+g(\nabla\phi,\nabla r)-F^{\prime}(r))^{2}. (3.7)

From an elementary inequality (ab)21γ+1a21γb2(a\mp b)^{2}\geqslant\frac{1}{\gamma+1}a^{2}-\frac{1}{\gamma}b^{2} for any real numbers a,ba,b and γ>0\gamma>0,

(~r+g(ϕ,r)F(r))21γ+1(~r+g(ϕ,r))21γ(F(r))2.(\tilde{\triangle}^{\perp}r+g(\nabla\phi,\nabla r)-F^{\prime}(r))^{2}\geqslant\frac{1}{\gamma+1}(\tilde{\triangle}^{\perp}r+g(\nabla\phi,\nabla r))^{2}-\frac{1}{\gamma}(F(r))^{2}.

By using the same inequality applied to (~r+g(ϕ,r))2(\tilde{\triangle}^{\perp}r+g(\nabla\phi,\nabla r))^{2}, any γ,η>0\gamma,\eta>0

(~r+g(ϕ,r))21(γ+1)η+γ+1(~r)21γ(F(r))21(γ+1)η(g(ϕ,r))2.(\tilde{\triangle}^{\perp}r+g(\nabla\phi,\nabla r))^{2}\geqslant\frac{1}{(\gamma+1)\eta+\gamma+1}(\tilde{\triangle}^{\perp}r)^{2}-\frac{1}{\gamma}(F(r))^{2}-\frac{1}{(\gamma+1)\eta}(g(\nabla\phi,\nabla r))^{2}. (3.8)

Inserting (3.8) into (3.7) with α=(2n2)γ>0,β=(2n2)(γ+1)η>0\alpha=(2n-2)\gamma>0,\beta=(2n-2)(\gamma+1)\eta>0,

0\displaystyle 0\geqslant Ric(r,r)+H(ϕ)(r,r)+g(r,r)\displaystyle\textup{Ric}^{\perp}(\nabla r,\nabla r)+H(\phi)(\nabla r,\nabla r)+g(\nabla r,\nabla{\triangle}^{\perp}r) (3.9)
+1α+β+2n2(~r)21α(F(r))21β(g(ϕ,r))2.\displaystyle+\frac{1}{\alpha+\beta+2n-2}(\tilde{\triangle}^{\perp}r)^{2}-\frac{1}{\alpha}(F(r))^{2}-\frac{1}{\beta}(g(\nabla\phi,\nabla r))^{2}. (3.10)

By the Cauchy-Schwarz inequality,

(g(ϕ,r))2g(ϕ,ϕ)g(r,r)=g(ϕ,ϕ),(g(\nabla\phi,\nabla r))^{2}\leqslant g(\nabla\phi,\nabla\phi)g(\nabla r,\nabla r)=g(\nabla\phi,\nabla\phi),

thus (3.9) becomes

0\displaystyle 0\geqslant Ric(r,r)+H(ϕ)(r,r)+g(r,r)\displaystyle\textup{Ric}^{\perp}(\nabla r,\nabla r)+H(\phi)(\nabla r,\nabla r)+g(\nabla r,\nabla{\triangle}^{\perp}r)
+1α+β+2n2(~r)21α(F(r))21β(g(ϕ,ϕ)).\displaystyle+\frac{1}{\alpha+\beta+2n-2}(\tilde{\triangle}^{\perp}r)^{2}-\frac{1}{\alpha}(F(r))^{2}-\frac{1}{\beta}(g(\nabla\phi,\nabla\phi)).

From the assumption on (g(ϕ,ϕ))(g(\nabla\phi,\nabla\phi)),

0\displaystyle 0\geqslant Ric(r,r)+H(ϕ)(r,r)+g(r,r)\displaystyle\textup{Ric}^{\perp}(\nabla r,\nabla r)+H(\phi)(\nabla r,\nabla r)+g(\nabla r,\nabla{\triangle}^{\perp}r)
+1α+β+2n2(~r)21α(F(r))2Cβr2.\displaystyle+\frac{1}{\alpha+\beta+2n-2}(\tilde{\triangle}^{\perp}r)^{2}-\frac{1}{\alpha}(F(r))^{2}-\frac{C}{\beta r^{2}}.

Now we take β=4Cα\beta=\frac{4C}{\alpha} and F(r)=αrF(r)=\frac{\alpha}{r}, then

0Ric(r,r)+H(ϕ)(r,r)+g(r,r)+αα2+(2n2)α+4C(~r)2.\displaystyle 0\geqslant\textup{Ric}^{\perp}(\nabla r,\nabla r)+H(\phi)(\nabla r,\nabla r)+g(\nabla r,\nabla{\triangle}^{\perp}r)+\frac{\alpha}{\alpha^{2}+(2n-2)\alpha+4C}(\tilde{\triangle}^{\perp}r)^{2}.

By the assumption on Ric+H(ϕ)\textup{Ric}^{\perp}+H(\phi),

0r(~r)+αα2+(2n2)α+4C(~r)2+(2n2)k.0\geqslant\partial_{r}({\tilde{\triangle}}^{\perp}r)+\frac{\alpha}{\alpha^{2}+(2n-2)\alpha+4C}(\tilde{\triangle}^{\perp}r)^{2}+(2n-2)k.

From (3.3),

limr0+r~r\displaystyle\lim_{r\rightarrow 0+}r{\tilde{\triangle}}^{\perp}r =limr0+(rrrg(ϕ,r)+α2)\displaystyle=\lim_{r\rightarrow 0+}\left(r\triangle r-rg(\nabla\phi,\nabla r)+\frac{\alpha}{2}\right)
=2n2+α2α2+(2n2)α+4Cα.\displaystyle=2n-2+\frac{\alpha}{2}\leqslant\frac{\alpha^{2}+(2n-2)\alpha+4C}{\alpha}.

Here, we used limr0+rr=limr0+rr=2n2\lim_{r\rightarrow 0+}r\triangle r=\lim_{r\rightarrow 0+}r\triangle^{\perp}r=2n-2. Thus by the Sturm-Liouville comparison argument,

~r(2n2)k(2n2+α+4Cα)cot(α(2n2)kα2+(2n2)α+4Cr).\tilde{\triangle}^{\perp}r\leqslant\sqrt{(2n-2)k\left(2n-2+\alpha+\frac{4C}{\alpha}\right)}\cot\left(\frac{\sqrt{\alpha(2n-2)k}}{\sqrt{\alpha^{2}+(2n-2)\alpha+4C}r}\right). (3.11)

Now we use the contradictory argument which was used in [Zhu1997, Limoncu2012] for ~r\tilde{\triangle}r. Let qMq\in M and let γ\gamma be a minimizing unit speed geodesic segment from pp to qq. Assume that

d(p,q)>π(2n2)k4C+2n2.d(p,q)>\frac{\pi}{\sqrt{(2n-2)k}}\sqrt{4\sqrt{C}+2n-2}.

Then γ(π(2n2)k4C+2n2)\gamma\left(\frac{\pi}{\sqrt{(2n-2)k}}\sqrt{4\sqrt{C}+2n-2}\right) must belong to MM outside of the cut-locus of pp. In particular, the distance function rr is smooth at this point. Now, the left-hand side of (3.11) is constant, whereas the right-hand side goes to -\infty, which yields the contradiction. Hence the diameter DD of MM must satisfy

Dα2+(2n2)α+4Crα(2n2)kπ.D\leqslant\frac{\sqrt{\alpha^{2}+(2n-2)\alpha+4C}r}{\sqrt{\alpha(2n-2)k}}\pi.

By taking α=2C\alpha=2\sqrt{C}, we have

Dπ(n1)k2C+n1.D\leqslant\frac{\pi}{\sqrt{(n-1)k}}\sqrt{2\sqrt{C}+n-1}.

After one obtains (3.5), Proposition (8) can be proven similar to the proof of [Limoncu2012, Theorem 2]. Also the proof can be also generalized to the quaternionic Kähler case.

Proposition 9.

Let (Mn,g,I,J,K)(M^{n},g,I,J,K) be a complete quaternion Kähler manifold with the quaternionic dimension n2n\geqslant 2. Take any pMp\in M and let rr be the geodesic distance function from pp. Suppose that for some constant k>0k>0, there exists a local orthonormal frame (Ei)i=14n(E_{i})^{4n}_{i=1} around pp with E1=rE_{1}=\nabla{r}, E2=IE1,E3=JE1,E4=KE1E_{2}=IE_{1},E_{3}=JE_{1},E_{4}=KE_{1} satisfying Ric(r,r)+Hess(ϕ)(r,r)2i=54nHess(r)(rEi,Ei)(4n4)k\textup{Ric}^{\perp}(\nabla r,\nabla r)+\operatorname{Hess}(\phi)(\nabla r,\nabla r)-2\sum_{i=5}^{4n}\operatorname{Hess}(r)(\nabla_{\nabla r}E_{i},E_{i})\geqslant(4n-4)k outside of the cut-locus of qq and |ϕ|2Cr(x)|\nabla\phi|^{2}\leqslant\frac{C}{r(x)} for some C0C\geqslant 0. Then the diameter DD of MM has the upper bound

Dπ(n1)kC+n1.D\leqslant\frac{\pi}{\sqrt{(n-1)k}}\sqrt{\sqrt{C}+n-1}.

4. Bakry-Émery orthogonal Ricci tensor associated with the possibly non-symmetric generator of a diffusion process

In this section we consider Bakry-Émery orthogonal Ricci tensor for the orthogonal Ricci curvature while imposing an additional assumption on the holomorphic sectional curvature on Kähler manifolds (quaternionic sectional curvature in the case of quaternion Kähler manifolds). Let MM be a real nn-dimensional connected smooth manifold with n2n\geqslant 2, either equipped with a Kähler structure or a quaternionic Kähler structure. Choose a complete Riemannian metric gg which is compatible with underlying Kähler or quaternionic Kähler structure and define :=+Z\mathcal{L}:=\triangle+Z, where \triangle is the Laplace-Beltrami operator associated with gg and ZZ a smooth vector field. We denote the Riemannian distance on MM associated with gg by rr. Let us define (0,2)(0,2)-symmetric tensor (Z)b(\nabla Z)^{{b}} by

(Z)b(X,Y):=12(XZ,Y+YZ,X).(\nabla Z)^{{b}}(X,Y):=\frac{1}{2}(\langle\nabla_{X}Z,Y\rangle+\langle\nabla_{Y}Z,X\rangle).

Given a constant mm\in\mathbb{R} with mnm\geqslant n, we define the Bakry-Émery type orthogonal Ricci tensor Ricm,Z\textup{Ric}^{\perp}_{m,Z} by

Ricm,Z:=Ric(Z)b1mnZZ.\textup{Ric}^{\perp}_{m,Z}:=\textup{Ric}^{\perp}-(\nabla Z)^{{b}}-\frac{1}{m-n}Z\otimes Z. (4.1)

In this definition, the convention is that for m=nm=n the vector field Z0Z\equiv 0.

Throughout this section, we will assume that with some constant m,k>0m,k>0, Ricm,Z(2m2)k\textup{Ric}^{\perp}_{m,Z}\geqslant(2m-2)k. This condition means that for a smooth curve γ(t)\gamma(t)

Ricm,Z(γ˙(t),γ˙(t))=Ric(γ˙(t),γ˙(t))(Z)b(γ˙(t),γ˙(t))1m2nZ(γ(t)),γ˙(t)\displaystyle\textup{Ric}^{\perp}_{m,Z}(\dot{\gamma}(t),\dot{\gamma}(t))=\textup{Ric}^{\perp}(\dot{\gamma}(t),\dot{\gamma}(t))-(\nabla Z)^{b}(\dot{\gamma}(t),\dot{\gamma}(t))-\frac{1}{m-2n}\langle Z(\gamma(t)),\dot{\gamma}(t)\rangle
(2m2)k.\displaystyle\geqslant(2m-2)k.
Proposition 10.

Suppose (Mn,g,J)(M^{n},g,J) is a Kähler manifold of the complex dimension n2n\geqslant 2. We denote by rr the Riemannian distance on MM from pp associated with the metric gg, and by CutpCut_{p} the cut-locus of pp. Suppose that for some constant k>0k>0, Ricm,Z(2m2)k\textup{Ric}^{\perp}_{m,Z}\geqslant(2m-2)k and H4kH\geqslant 4k. Let xM\Cutp{p}x\in M\backslash Cut_{p}\cup\left\{p\right\} with r(x)<π2kr(x)<\frac{\pi}{2\sqrt{k}}. Then

r(x)(m2)𝔰(k,r(x))𝔰(k,r(x))+𝔰(4k,r(x))𝔰(4k,r(x)),\mathcal{L}r(x)\leqslant(m-2)\frac{\mathfrak{s}^{\prime}(k,r(x))}{\mathfrak{s}(k,r(x))}+\frac{\mathfrak{s}^{\prime}(4k,r(x))}{\mathfrak{s}(4k,r(x))},

where 𝔰(k,t):=sinkt\mathfrak{s}(k,t):=\sin{\sqrt{k}t}.

Proof.

When m=2nm=2n, we have Z0Z\equiv 0 and we can use the same argument as we use below. Thus without loss of generality, we will assume m>2nm>2n.

Let pMp\in M and xpx\neq p which is not in the cut-locus of pp. Let γ:[0,r(x)]M\gamma:[0,r(x)]\rightarrow M be the unique arclength parameterized geodesic connecting pp to xx. At xx, we consider an orthonormal frame {X1(x),,X2n(x)}\left\{X_{1}(x),...,X_{2n}(x)\right\} such that

X1(x)=γ(r(x)),X2(x)=JX1(x).X_{1}(x)=\gamma^{\prime}(r(x)),X_{2}(x)=JX_{1}(x).

Then

r=r(x)+Zr(x)=i=12n2r(Xi(x),Xi(x))+Zr(x).\mathcal{L}r=\triangle r(x)+Zr(x)=\sum_{i=1}^{2n}\nabla^{2}r(X_{i}(x),X_{i}(x))+Zr(x).

Since X1(x)=γ(r(x))X_{1}(x)=\gamma^{\prime}(r(x)), 2r(X1(x),X1(x))\nabla^{2}r(X_{1}(x),X_{1}(x)) is zero. Now we divide the above sum into three parts: 2r(X2(x),X2(x))\nabla^{2}r(X_{2}(x),X_{2}(x)), i=32n2r(Xi(x),Xi(x))\sum_{i=3}^{2n}\nabla^{2}r(X_{i}(x),X_{i}(x)), and Zr(x)Zr(x).

For 2r(X2(x),X2(x))\nabla^{2}r(X_{2}(x),X_{2}(x)), since JJ is parellel and γ\gamma is a geodesic, the vector field defined along γ\gamma by JγJ\gamma^{\prime} is parellel. Define the vector field along γ\gamma by

X~(γ(t))=𝔰(4k,t)𝔰(4k,r(x))Jγ(t),\tilde{X}(\gamma(t))=\frac{\mathfrak{s}(4k,t)}{\mathfrak{s}(4k,r(x))}J\gamma^{\prime}(t),

where 𝔰(k,t):=sinkt\mathfrak{s}(k,t):=\sin{\sqrt{k}t}. From the index lemma,

2r(X2(x),X2(x))\displaystyle\nabla^{2}r(X_{2}(x),X_{2}(x)) 0r(x)(γX~,γX~R(γ,X~)X~,γ)𝑑t\displaystyle\leqslant\int_{0}^{r(x)}\left(\langle\nabla_{\gamma^{\prime}}\tilde{X},\nabla_{\gamma^{\prime}}\tilde{X}\rangle-\langle R(\gamma^{\prime},\tilde{X})\tilde{X},\gamma^{\prime}\rangle\right)dt
=1𝔰(4k,r(x))20r(x)(𝔰(4k,t)2𝔰(4k,t)2R(γ,Jγ)Jγ,γ)𝑑t\displaystyle=\frac{1}{\mathfrak{s}(4k,r(x))^{2}}\int_{0}^{r(x)}(\mathfrak{s}^{\prime}(4k,t)^{2}-\mathfrak{s}(4k,t)^{2}\langle R(\gamma^{\prime},J\gamma^{\prime})J\gamma^{\prime},\gamma^{\prime}\rangle)dt
1𝔰(4k,r(x))20r(x)(𝔰(4k,t)24k𝔰(4k,t)2)𝑑t\displaystyle\leqslant\frac{1}{\mathfrak{s}(4k,r(x))^{2}}\int_{0}^{r(x)}(\mathfrak{s}^{\prime}(4k,t)^{2}-4k\mathfrak{s}(4k,t)^{2})dt

Next, let us estimate i=32n2r(Xi(x),Xi(x))\sum_{i=3}^{2n}\nabla^{2}r(X_{i}(x),X_{i}(x)). We denote by {X3,,X2n}\left\{X_{3},\cdots,X_{2n}\right\} the vector fields along γ\gamma obtained by parallel transport of {X3(x),,X2n(x)}\left\{X_{3}(x),\cdots,X_{2n}(x)\right\}. Define the vector fields along γ\gamma by

X~i(γ(t))=𝔰(k,t)𝔰(k,r(x))Xi(γ(t)),i=3,,2m.\tilde{X}_{i}(\gamma(t))=\frac{\mathfrak{s}(k,t)}{\mathfrak{s}(k,r(x))}X_{i}(\gamma(t)),i=3,\cdots,2m.

By the index lemma,

i=32n2r(Xi(x),Xi(x))\displaystyle\sum_{i=3}^{2n}\nabla^{2}r(X_{i}(x),X_{i}(x)) i=32n0r(x)(γX~i,γX~iR(γ,X~)iX~i,γ)𝑑t\displaystyle\leqslant\sum_{i=3}^{2n}\int_{0}^{r(x)}\left(\langle\nabla_{\gamma^{\prime}}\tilde{X}_{i},\nabla_{\gamma^{\prime}}\tilde{X}_{i}\rangle-\langle R(\gamma^{\prime},\tilde{X})_{i}\tilde{X}_{i},\gamma^{\prime}\rangle\right)dt
=1𝔰(k,r(x))2i=32n0r(x)(𝔰(k,t)2𝔰(k,t)2R(γ,X~i)X~i,γ)𝑑t\displaystyle=\frac{1}{\mathfrak{s}(k,r(x))^{2}}\sum_{i=3}^{2n}\int_{0}^{r(x)}(\mathfrak{s}^{\prime}(k,t)^{2}-\mathfrak{s}(k,t)^{2}\langle R(\gamma^{\prime},\tilde{X}_{i})\tilde{X}_{i},\gamma^{\prime}\rangle)dt
1𝔰(k,r(x))20r(x)((2n2)𝔰(k,t)2𝔰(k,t)2Ric(γ,γ))𝑑t.\displaystyle\leqslant\frac{1}{\mathfrak{s}(k,r(x))^{2}}\int_{0}^{r(x)}((2n-2)\mathfrak{s}^{\prime}(k,t)^{2}-\mathfrak{s}(k,t)^{2}\textup{Ric}^{\perp}(\gamma^{\prime},\gamma^{\prime}))dt.

For the last term Zr(x)Zr(x), we have

Zr(x)\displaystyle Zr(x) =Z(x),γ˙(r(x))𝔰(k,r(x))2𝔰(k,r(x))2Z(p),γ˙(0)𝔰(k,0)2𝔰(k,r(x))2\displaystyle=\langle Z(x),\dot{\gamma}(r(x))\rangle\frac{\mathfrak{s}(k,r(x))^{2}}{\mathfrak{s}(k,r(x))^{2}}-\langle Z(p),\dot{\gamma}(0)\rangle\frac{\mathfrak{s}(k,0)^{2}}{\mathfrak{s}(k,r(x))^{2}}
=0r(x)t(Z(γ(t)),γ˙(t)𝔰(k,t)2𝔰(k,r(x))2dt\displaystyle=\int_{0}^{r(x)}\frac{\partial}{\partial t}(\langle Z(\gamma(t)),\dot{\gamma}(t)\rangle\frac{\mathfrak{s}(k,t)^{2}}{\mathfrak{s}(k,r(x))^{2}}dt
=0r(x)((Z)b(γ˙(t),γ˙(t))𝔰(k,t)2𝔰(k,r(x))2+2Z(γ(t)),γ˙(t)𝔰(k,t)𝔰(k,r(x)))𝔰(k,t)𝔰(k,r(x)))dt\displaystyle=\int_{0}^{r(x)}\left((\nabla Z)^{b}(\dot{\gamma}(t),\dot{\gamma}(t))\frac{\mathfrak{s}(k,t)^{2}}{\mathfrak{s}(k,r(x))^{2}}+2\langle Z(\gamma(t)),\dot{\gamma}(t)\rangle\frac{\mathfrak{s}^{\prime}(k,t)}{\mathfrak{s}(k,r(x))})\frac{\mathfrak{s}(k,t)}{\mathfrak{s}(k,r(x))}\right)dt
0r(x)((Z)b(γ˙(t),γ˙(t))𝔰(k,t)2𝔰(k,r(x))2+1m2n0r(x)Z(γ(t)),γ˙(t)𝔰(k,t)2𝔰(k,r(x))2)𝑑t\displaystyle\leqslant\int_{0}^{r(x)}\left((\nabla Z)^{b}(\dot{\gamma}(t),\dot{\gamma}(t))\frac{\mathfrak{s}(k,t)^{2}}{\mathfrak{s}(k,r(x))^{2}}+\frac{1}{m-2n}\int_{0}^{r(x)}\langle Z(\gamma(t)),\dot{\gamma}(t)\rangle\frac{\mathfrak{s}(k,t)^{2}}{\mathfrak{s}(k,r(x))^{2}}\right)dt
+(m2n)0r(x)(𝔰(k,t))2𝔰(k,r(x))2𝑑t.\displaystyle+(m-2n)\int_{0}^{r(x)}\frac{(\mathfrak{s}^{\prime}(k,t))^{2}}{\mathfrak{s}(k,r(x))^{2}}dt.

Here the last inequality follows from the arithmetic-geometric mean inequality. Since Ricm,Z(γ˙(t),γ˙(t))=Ric(γ˙(t),γ˙(t))(Z)b(γ˙(t),γ˙(t))1m2nZ(γ(t)),γ˙(t)(2m2)k\textup{Ric}^{\perp}_{m,Z}(\dot{\gamma}(t),\dot{\gamma}(t))=\textup{Ric}^{\perp}(\dot{\gamma}(t),\dot{\gamma}(t))-(\nabla Z)^{b}(\dot{\gamma}(t),\dot{\gamma}(t))-\frac{1}{m-2n}\langle Z(\gamma(t)),\dot{\gamma}(t)\rangle\geq(2m-2)k for all t[0,r(x)]t\in[0,r(x)], we obtain

r(x)\displaystyle\mathcal{L}r(x) 0r(x)(m2)(𝔰(k,t))2𝔰(k,r(x))2𝑑t0r(x)Ricm,Z(γ˙(t),γ˙(t))𝔰(k,t)2𝔰(k,r(x))2𝑑t\displaystyle\leqslant\int_{0}^{r(x)}(m-2)\frac{(\mathfrak{s}^{\prime}(k,t))^{2}}{\mathfrak{s}(k,r(x))^{2}}dt-\int_{0}^{r(x)}\textup{Ric}^{\perp}_{m,Z}(\dot{\gamma}(t),\dot{\gamma}(t))\frac{\mathfrak{s}(k,t)^{2}}{\mathfrak{s}(k,r(x))^{2}}dt
+1𝔰(4k,r(x))20r(x)(𝔰(4k,t)24k𝔰(4k,t)2)𝑑t\displaystyle+\frac{1}{\mathfrak{s}(4k,r(x))^{2}}\int_{0}^{r(x)}(\mathfrak{s}^{\prime}(4k,t)^{2}-4k\mathfrak{s}(4k,t)^{2})dt
(m2)0r(x)(𝔰(k,t)2𝔰(k,r(x))2k𝔰(k,t)2𝔰(k,r(x))2)𝑑u\displaystyle\leqslant(m-2)\int_{0}^{r(x)}\left(\frac{\mathfrak{s}^{\prime}(k,t)^{2}}{\mathfrak{s}(k,r(x))^{2}}-k\frac{\mathfrak{s}(k,t)^{2}}{\mathfrak{s}(k,r(x))^{2}}\right)du
+1𝔰(4k,r(x))20r(x)(𝔰(4k,t)24k𝔰(4k,t)2)𝑑t\displaystyle+\frac{1}{\mathfrak{s}(4k,r(x))^{2}}\int_{0}^{r(x)}(\mathfrak{s}^{\prime}(4k,t)^{2}-4k\mathfrak{s}(4k,t)^{2})dt
=(m2)[𝔰(k,r(x))𝔰(k,r(x))𝔰(k,r(x))𝔰(k,r(x))]t=0r(x)+[(𝔰(k,r(x)))𝔰(k,r(x))(𝔰(4k,r(x)))𝔰(4k,r(x))]t=0r(x)\displaystyle=(m-2)[\frac{\mathfrak{s}^{\prime}(k,r(x))}{\mathfrak{s}(k,r(x))}\frac{\mathfrak{s}^{\prime}(k,r(x))}{\mathfrak{s}(k,r(x))}]^{r(x)}_{t=0}+[\frac{(\mathfrak{s}^{\prime}(k,r(x)))}{\mathfrak{s}(k,r(x))}\frac{(\mathfrak{s}^{\prime}(4k,r(x)))}{\mathfrak{s}(4k,r(x))}]^{r(x)}_{t=0}
=(m2)𝔰(k,r(x))𝔰(k,r(x))+𝔰(4k,r(x))𝔰(4k,r(x)).\displaystyle=(m-2)\frac{\mathfrak{s}^{\prime}(k,r(x))}{\mathfrak{s}(k,r(x))}+\frac{\mathfrak{s}^{\prime}(4k,r(x))}{\mathfrak{s}(4k,r(x))}.

For the proposition below, let {Xtx}t0,xM\left\{X_{t}^{x}\right\}_{t\geqslant 0},x\in M be the diffusion process with the infinitesimal generator \mathcal{L}. Let us define a stopping time σp\sigma_{p} by

σp:=inf{t0:dp(Xt)=π2k}\sigma_{p}:=\inf\left\{t\geqslant 0:d_{p}(X_{t})=\frac{\pi}{2\sqrt{k}}\right\}

in the Kähler case and

σp:=inf{t0:dp(Xt)=π23k}\sigma_{p}:=\inf\left\{t\geqslant 0:d_{p}(X_{t})=\frac{\pi}{2\sqrt{3k}}\right\}

in the quaternionic case.

Proposition 11.

Given a Kähler manifold (Mn,g,J)(M^{n},g,J) of the complex dimension n2n\geqslant 2, suppose that for some constant k>0k>0, Ricm,Z(2m2)k\textup{Ric}^{\perp}_{m,Z}\geqslant(2m-2)k and H4kH\geqslant 4k hold on the open ball of raidus π2k\frac{\pi}{2\sqrt{k}} centered at pp. Then σp=\sigma_{p}=\infty holds q\mathbb{P}_{q}-almost surely for any qM\(Cutp{p})q\in M\backslash(Cut_{p}\cup\left\{p\right\}) with dp(q)<π2kd_{p}(q)<\frac{\pi}{2\sqrt{k}}.

Proof.

By Itô’s formula for the radial process dp(Xt)d_{p}(X_{t}) together with Proposition 10, we have

dp(Xt)dp(q)+2βt+0tdp(Xs)𝑑s\displaystyle d_{p}(X_{t})\leqslant d_{p}(q)+\sqrt{2}\beta_{t}+\int_{0}^{t}\mathcal{L}d_{p}(X_{s})ds
dp(q)+2βt+0t(m2)𝔰(k,r(Xs))𝔰(k,r(Xs))+𝔰(4k,r(Xs))𝔰(4k,r(Xs))ds\displaystyle\leqslant d_{p}(q)+\sqrt{2}\beta_{t}+\int_{0}^{t}(m-2)\frac{\mathfrak{s}^{\prime}(k,r(X_{s}))}{\mathfrak{s}(k,r(X_{s}))}+\frac{\mathfrak{s}^{\prime}(4k,r(X_{s}))}{\mathfrak{s}(4k,r(X_{s}))}ds

for t<σpt<\sigma_{p}, where βt\beta_{t} is a 11-dimensional standard Brownian motion. Let us define ρt\rho_{t} as the solution to the following stochastic differential equation

dρt=2dβt+((m2)𝔰(k,r(ρt))𝔰(k,r(ρt))+𝔰(4k,r(ρt))𝔰(4k,r(ρt)))dtd\rho_{t}=\sqrt{2}d\beta_{t}+\left((m-2)\frac{\mathfrak{s}^{\prime}(k,r(\rho_{t}))}{\mathfrak{s}(k,r(\rho_{t}))}+\frac{\mathfrak{s}^{\prime}(4k,r(\rho_{t}))}{\mathfrak{s}(4k,r(\rho_{t}))}\right)dt

with ρ0=dp(q)\rho_{0}=d_{p}(q). (see for example [HsuEltonBook, Theorem 3.5.3]). Thus it suffices to show that ρt\rho_{t} never hit π2k\frac{\pi}{2\sqrt{k}}. Since

𝔰(4k,r(ρt))𝔰(4k,r(ρt))=1uπ2k+o(1)\frac{\mathfrak{s}^{\prime}(4k,r(\rho_{t}))}{\mathfrak{s}(4k,r(\rho_{t}))}=\frac{1}{u-\frac{\pi}{2\sqrt{k}}}+o(1)

as uπ2ku\uparrow\frac{\pi}{2\sqrt{k}} and m2n2m\geqslant 2n\geqslant 2, a general theory of 11-dimensional diffusion processes yields that ρt\rho_{t} cannot hit π2k\frac{\pi}{2\sqrt{k}} (see e.g. [HsuEltonBook, Proposition 4.2.2]).

By using Proposition above, we can easily show the Bonnet-Myers theorem.

Corollary 12.

Given a Kähler manifold (Mn,g,J)(M^{n},g,J) of the complex dimension n2n\geqslant 2, suppose that for some constant k>0k>0, Ricm,Z(2m2)k\textup{Ric}^{\perp}_{m,Z}\geqslant(2m-2)k and H4kH\geqslant 4k hold on MM. Then the diameter of MM is less than equal to π2k\frac{\pi}{2\sqrt{k}}.

Proof.

Suppose that there are p,qMp,q\in M such that d(p,q)>π2kd(p,q)>\frac{\pi}{2\sqrt{k}}. We may assume that MM is compact and that Ricm,Z(2m2)k\textup{Ric}^{\perp}_{m,Z}\geqslant(2m-2)k, holds on the open ball of radius π2k\frac{\pi}{2\sqrt{k}} centered at pp by modifying outside of a ball of large radius. Then there is an open neighborhood GG of qq such that d(p,y)>π2kd(p,y)>\frac{\pi}{2\sqrt{k}} for all yGy\in G. Take pp^{\prime} from a small neighborhood of pp. Then Proposition yields that p[σp=]=1\mathbb{P}_{p^{\prime}}[\sigma_{p}=\infty]=1. It implies p[XtG]=0\mathbb{P}_{p^{\prime}}[X_{t}\in G]=0 for any t>0t>0. This is absurd since the law of XtX_{t} has a strictly positive density with respect to the Riemannian volume measure for t>0t>0. ∎

The similar propositions of the quaternion Kähler case can be obtained by repeating the proofs of the Kähler case. We follow the structure of the proof in [BaudoinYang2020, Theorem 3.2].

Proposition 13.

Given a quaternionic Kähler manifold (Mn,g,I,J,K)(M^{n},g,I,J,K) of the quaternionic dimension n2n\geqslant 2 and we denote the Riemannian distance on MM from pp associated with gg by rr and CutpCut_{p} the cut-locus of pp. Suppose that for some constant k>0k>0, Ricm,Z(4m4)k\textup{Ric}^{\perp}_{m,Z}\geqslant(4m-4)k and Q12kQ\geqslant 12k. Let xM\Cutp{p}x\in M\backslash Cut_{p}\cup\left\{p\right\} with r(x)<π23kr(x)<\frac{\pi}{2\sqrt{3k}}. Then

r(x)(m4)𝔰(k,r(x))𝔰(k,r(x))+𝔰(12k,r(x))𝔰(12k,r(x)),\mathcal{L}r(x)\leqslant(m-4)\frac{\mathfrak{s}^{\prime}(k,r(x))}{\mathfrak{s}(k,r(x))}+\frac{\mathfrak{s}^{\prime}(12k,r(x))}{\mathfrak{s}(12k,r(x))},

where 𝔰(k,t):=sinkt\mathfrak{s}(k,t):=\sin{\sqrt{k}t}.

Proof.

When m=4nm=4n, we just need to put Z0Z\equiv 0 and proceed the same argument that we provide below. Thus without loss of generality, we will assume m>4nm>4n.

Let pMp\in M and xpx\neq p which is not in the cut-locus of pp. Let γ:[0,r(x)]M\gamma:[0,r(x)]\rightarrow M be the unique arclength parameterized geodesic connecting pp to xx. At xx, we consider an orthonormal frame {X1(x),,X4n(x)}\left\{X_{1}(x),...,X_{4n}(x)\right\} such that

X1(x)=γ(r(x)),X2(x)=IX1(x),X3(x)=JX1(x),X2(x)=KX1(x).X_{1}(x)=\gamma^{\prime}(r(x)),X_{2}(x)=IX_{1}(x),X_{3}(x)=JX_{1}(x),X_{2}(x)=KX_{1}(x).

Then

r=r(x)+Zr(x)=i=14n2r(Xi(x),Xi(x))+Zr(x).\mathcal{L}r=\triangle r(x)+Zr(x)=\sum_{i=1}^{4n}\nabla^{2}r(X_{i}(x),X_{i}(x))+Zr(x).

Since X1(x)=γ(r(x))X_{1}(x)=\gamma^{\prime}(r(x)), 2r(X1(x),X1(x))\nabla^{2}r(X_{1}(x),X_{1}(x)) is zero. Now we divide the above sum into three parts: i=242r(Xi(x),Xi(x))\sum_{i=2}^{4}\nabla^{2}r(X_{i}(x),X_{i}(x)), i=34n2r(Xi(x),Xi(x))\sum_{i=3}^{4n}\nabla^{2}r(X_{i}(x),X_{i}(x)), and Zr(x)Zr(x).

For i=242r(Xi(x),Xi(x))\sum_{i=2}^{4}\nabla^{2}r(X_{i}(x),X_{i}(x)), note that vectors IX1,JX1,KX1IX_{1},JX_{1},KX_{1} might not be parallel along γ\gamma. Denote X2,X3X_{2},X_{3}, and X4X_{4} the vector fields along γ\gamma obtained by parallel transport along γ\gamma of X2(x),X3(x)X_{2}(x),X_{3}(x) and X4(x)X_{4}(x). Then one can deduce that

R(γ,X2,X2,γ)+R(γ,X3,X3,γ)+R(γ,X4,X4,γ)\displaystyle R(\gamma^{\prime},X_{2},X_{2},\gamma^{\prime})+R(\gamma^{\prime},X_{3},X_{3},\gamma^{\prime})+R(\gamma^{\prime},X_{4},X_{4},\gamma^{\prime})
=R(γ,Iγ,Iγ,γ)+R(γ,Jγ,Jγ,γ)+R(γ,Kγ,Kγ,γ)=Q(γ)\displaystyle=R(\gamma^{\prime},I\gamma^{\prime},I\gamma^{\prime},\gamma^{\prime})+R(\gamma^{\prime},J\gamma^{\prime},J\gamma^{\prime},\gamma^{\prime})+R(\gamma^{\prime},K\gamma^{\prime},K\gamma^{\prime},\gamma^{\prime})=Q(\gamma^{\prime})

(see [BaudoinYang2020, Theorem 3.2] for more details). Define the vector field along γ\gamma by

X~i(γ(t))=𝔰(12k,t)𝔰(12k,r(x))Jγ(t),i=2,3,4,\tilde{X}_{i}(\gamma(t))=\frac{\mathfrak{s}(12k,t)}{\mathfrak{s}(12k,r(x))}J\gamma^{\prime}(t),i=2,3,4,

where 𝔰(k,t):=sinkt\mathfrak{s}(k,t):=\sin{\sqrt{k}t}, we obtain by the same computation as in the Proposition 10,

i=242r(Xi(x),Xi(x))1𝔰(12k,r(x))20r(x)(𝔰(12k,t)212k𝔰(4k,t)2)𝑑t\displaystyle\sum_{i=2}^{4}\nabla^{2}r(X_{i}(x),X_{i}(x))\leqslant\frac{1}{\mathfrak{s}(12k,r(x))^{2}}\int_{0}^{r(x)}(\mathfrak{s}^{\prime}(12k,t)^{2}-12k\mathfrak{s}(4k,t)^{2})dt

The rest steps are similar as in the Proposition 10. Consequently,

r(x)\displaystyle\mathcal{L}r(x) =(m4)𝔰(k,r(x))𝔰(k,r(x))+𝔰(12k,r(x))𝔰(12k,r(x)).\displaystyle=(m-4)\frac{\mathfrak{s}^{\prime}(k,r(x))}{\mathfrak{s}(k,r(x))}+\frac{\mathfrak{s}^{\prime}(12k,r(x))}{\mathfrak{s}(12k,r(x))}.

Combining the proposition above with Proposition 13, we obtained:

Proposition 14.

Given a quaternionic Kähler manifold (Mn,g,I,J,K)(M^{n},g,I,J,K) of the quaternionic dimension n2n\geqslant 2, suppose that for some constant k>0k>0, Ricm,Z(4m4)k\textup{Ric}^{\perp}_{m,Z}\geqslant(4m-4)k and Q12kQ\geqslant 12k hold on the open ball of radius π2k\frac{\pi}{2\sqrt{k}} centered at pp. Then σp=\sigma_{p}=\infty holds q\mathbb{P}_{q}-almost surely for any qM\(Cutp{p})q\in M\backslash(Cut_{p}\cup\left\{p\right\}) with dp(q)<π23kd_{p}(q)<\frac{\pi}{2\sqrt{3k}}.

Corollary 15.

Given a quaternionic Kähler manifold (Mn,g,I,J,K)(M^{n},g,I,J,K) of the quaternionic dimension n2n\geqslant 2, suppose that for some constant k>0k>0, Ricm,Z(4m4)k\textup{Ric}^{\perp}_{m,Z}\geqslant(4m-4)k and Q12kQ\geqslant 12k hold on MM. Then the diameter of MM is less than equal to π23k\frac{\pi}{2\sqrt{3k}}.

Conflicts of interest

The corresponding author states that there is no conflict of interest.

References