THE GEOMETRY OF QUASI-SASAKI MANIFOLDS by ADAM P. WELLY A DISSERTATION Presented to the Department of Mathematics and the Graduate School of the University of Oregon in partial fulfillment of the requirements for the degree of Doctor of Philosophy June 2016 DISSERTATION APPROVAL PAGE Student: Adam P. Welly Title: The Geometry of quasi-Sasaki Manifolds This dissertation has been accepted and approved in partial fulfillment of the requirements for the Doctor of Philosophy degree in the Department of Mathematics by: Dr. Weiyong He Chair Dr. Peter Gilkey Core Member Dr. Peng Lu Core Member Dr. Christopher Sinclair Core Member Dr. Dietrich Belitz Institutional Representative and Scott Pratt Dean of the Graduate School Original approval signatures are on file with the University of Oregon Graduate School. Degree awarded June 2016 ii c© 2016 Adam P. Welly iii DISSERTATION ABSTRACT Adam P. Welly Doctor of Philosophy Department of Mathematics June 2016 Title: The Geometry of quasi-Sasaki Manifolds Let (M, g) be a quasi-Sasaki manifold with Reeb vector field ξ. Our goal is to understand the structure of M when g is an Einstein metric. Assuming that the S1 action induced by ξ is locally free or assuming a certain non-negativity condition on the transverse curvature, we prove some rigidity results on the structure of (M, g). Naturally associated to a quasi-Sasaki metric g is a transverse Ka¨hler metric gT . The transverse Ka¨hler-Ricci flow of gT is the normalized Ricci flow of the transverse metric. Exploiting the transverse Ka¨hler geometry of (M, g), we can extend results in Ka¨hler-Ricci flow to our transverse version. In particular, we show that a deep and beautiful theorem due to Perleman has its counterpart in the quasi-Sasaki setting. We also consider evolving a Sasaki metric g by Ricci flow. Unfortunately, if g(0) is Sasaki then g(t) is not Sasaki for t > 0. However, in some instances g(t) is quasi-Sasaki. We examine this and give some qualitative results and examples in the special case that the initial metric is η-Einstein. iv CURRICULUM VITAE NAME OF AUTHOR: Adam P. Welly GRADUATE AND UNDERGRADUATE SCHOOLS ATTENDED: University of Oregon, Eugene, OR Western Washington University, Bellingham, WA DEGREES AWARDED: Doctor of Philosophy, Mathematics, 2016, University of Oregon Bachelor of Science, Mathematics, 2010, Western Washington University AREAS OF SPECIAL INTEREST: Differential Geometry Geometric Analysis PROFESSIONAL EXPERIENCE: Graduate Teaching Fellow, Department of Mathematics, University of Oregon, Eugene OR, 2010–2016 Peer Tutor, Department of Mathematics, Western Washington University, Bellingham Wa, 2008–2010 v ACKNOWLEDGEMENTS I would like to thank my advisor Weiyong He for his encouragement and support, as well as many pieces of valuable advice. I would also like to thank Joey Iverson for providing the LATEX template for this document. vi Dedicated to my wife Allison and my son Owen. vii TABLE OF CONTENTS Chapter Page I. INTRODUCTION . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1 1.1. Overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1 1.2. Preliminaries . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4 1.3. Transverse Ka¨hler Geometry . . . . . . . . . . . . . . . . . . . 12 1.4. Transverse Distance and Preferred Coordinates . . . . . . . . . 20 II. RIGIDITY OF QUASI-SASAKI-EINSTEIN METRICS . . . . . . . . . 25 2.1. The Ricci tensor and Einstein equation . . . . . . . . . . . . . 25 2.2. The Regular Case . . . . . . . . . . . . . . . . . . . . . . . . . 31 2.3. The quasi-Regular Case . . . . . . . . . . . . . . . . . . . . . . 39 2.4. Transverse Curvature Conditions . . . . . . . . . . . . . . . . . 42 2.5. Splitting the Contact Bundle . . . . . . . . . . . . . . . . . . . 45 2.6. The General Case with a Curvature Condition . . . . . . . . . 48 III. TRANSVERSE KA¨HLER-RICCI FLOW . . . . . . . . . . . . . . . . . 54 3.1. Transverse Ka¨hler-Ricci Flow . . . . . . . . . . . . . . . . . . 56 3.2. Perelman’s entropy functional on quasi-Sasaki manifolds . . . . 64 3.3. Bounds on Scalar Curvature and the Transverse Ricci Potential 74 viii Chapter Page 3.4. Upper Bound on Diameter . . . . . . . . . . . . . . . . . . . . 85 3.5. Transverse Holomorphic Bisectional Curvature . . . . . . . . . 90 IV. THE RICCI FLOW OF A SASAKI METRIC . . . . . . . . . . . . . . 96 4.1. η-Einstein metrics . . . . . . . . . . . . . . . . . . . . . . . . . 97 4.2. Transverse Calabi-Yau Structure . . . . . . . . . . . . . . . . . 100 4.3. Transverse Fano Structure . . . . . . . . . . . . . . . . . . . . 102 4.4. Transverse Canonical Structure . . . . . . . . . . . . . . . . . 105 4.5. Rigidity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 108 4.6. Questions and Directions for Future Research . . . . . . . . . . 108 V. APPENDIX . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 110 REFERENCES CITED . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 119 ix CHAPTER I INTRODUCTION 1.1. Overview This dissertation is an exploration into the geometry of quasi-Sasaki manifolds. As the name suggests, a quasi-Sasaki structure is similar to a Sasaki one, but is more flexible than the latter. Indeed, Sasaki manifolds form a subset of the quasi- Sasaki manifolds, but in general the two can have quite different features. For example, cosymplectic manifolds are also quasi-Sasaki. A cosymplectic manifold has an integrable contact bundle whereas the contact bundle of a Sasaki manifold is as far from being integrable as possible. But like a Sasaki manifold, a quasi-Sasaki manifold (M, g) comes equipped with a one-dimensional Riemannian foliation, Fξ, determined by a nowhere vanishing vector field ξ. The space transverse to this foliation carries a Ka¨hler structure. Dual to the vector field ξ (via the metric g) is a 1-form η. The kernel of η is called the contact bundle. Perhaps the most significant difference between a quasi-Sasaki and a true Sasaki structure has to do with the rank of the 2-form dη. For a Sasaki structure, η ∧ (dη)n is a volume form and the transverse Ka¨hler metric ω = dη. If a quasi-Sasaki manifold has (dη)n = 0, then it lacks a contact structure and, even when (dη)n 6= 0, there may be no clear relationship between the transverse Ka¨hler metric and the 2-form dη. In the next sections we will recall all of the necessary definitions and elementary facts concerning quasi-Sasaki manifolds that will be used throughout the remainder of this dissertation. The pioneering work on quasi-Sasaki manifolds was iniated by Blair in [3] during the late 1960’s. However, his work was not without errors. These errors were later 1 recognized by Tanno who then gave complete and correct revisions of Blair’s results in [37]. The main results of these papers describe a splitting of quasi-Sasaki manifolds by assuming the integrability of a certain almost product structure TM = P ⊕ Q and positive definiteness of dη(·,Φ·) when restricted to P . These assumptions are somewhat unnatural and unmotivated and we do not make them here. Rather, we want to prove a splitting result like their’s, but under the assumption that the metric g is Einstein. In fact, one of our motivating reasons for studying quasi-Sasaki manifolds is the search for “new” Einstein metrics. A Riemannian metric g is called Einstein if its Ricci tensor is proportional to g. That is, if there is a constant λ ∈ R such that Ricg = λg. Differential geometers are interested in Einstein metrics for a wide variety of reasons. Entire books have been written on the subject; the book by Besse is a classic. Often times Einstein metrics are somehow the “best” or most desirable metrics. They appear as critical points of certain geometrically defined functionals. Physicists are interested in Einstein metrics for their role in the theory of general relativity. Thus we are curious about the rigidity of quasi-Sasaki-Einstein metrics. For example, we can create quasi-Sasaki-Einstein metrics by taking products of Sasaki-Einstein and Ka¨hler-Einstein metrics, but we don’t consider these to be “new” Einstein metrics. So the question is, are all quasi-Sasaki-Einstein metrics of this form or are there more “interesting” examples possible? We investigate this in chapter II. By assuming some regularity on the orbits of the Reeb vector field ξ, we can give a precise description of the possible Einstein metrics. We define a transverse version of non-negative quadratic orthogonal bisectional curvature (NQOBC) and show that a quasi-Sasaki-Einstein metric satisfying this curvature conditions is rigid. We show that M is locally the product of a Ka¨hler-Einstein manifold M0 and a full rank quasi-Sasaki-Einstein manifold M1. Furthermore, the transverse Ka¨hler metric of M1 2 is a product of Ka¨hler-Einstein metrics. If it happens that the second basic Betti number bB2 (M1) = 1, then the metric on M1 is simply a scaling of a Sasaki-Einstein metric. Note that Sasaki metrics are not preserved by homotheties, but quasi-Sasaki metrics are. Under the stronger assumption of positive or non-negative transverse holomorphic bisectional curvature, we conjecture that something like the Frankel and the generalized Frankel conjectures (which are actually no longer conjectures) in Sasaki geometry ought to be true. See [20] and [21]. As we will see later in this chapter, a quasi-Sasaki manifold (M, g) has naturally associated to it a transverse Ka¨hler metric gT . It is known by the work of Cao in [7] that if the first Chern class of a Ka¨hler manifold is negative or null, then the Ka¨hler- Ricci flow will converge to a Ka¨hler-Einstein metric. This gives us some motivation to consider the transverse Ka¨hler-Ricci flow (TKRF). By exploiting the transverse Ka¨hler geometry of (M, g), we can extend deep results in Ka¨hler-Ricci flow to our transverse version. In particular, if the basic first Chern class is negative or null, we can reproduce the results of Cao in [7] in our transverse setting. If M is compact and the basic first Chern class is positive, then defining functionals which are monotonic along the TKRF, we can mimic the work of Perleman in [34] and prove uniform bounds on the diameter and scalar curvature along the flow. We also show that, analogous to [2] and [29], positivity of transverse holomorphic bisectional curvature is preserved along the TKRF. All of this and more is carried out in chapter III. In chapter IV we start to explore Hamilton’s Ricci flow on a Sasaki manifold (M, g). That is, we let the Sasaki metric g itself evolve by the Ricci flow: ∂g ∂t = −2Ricg. 3 It is well-known that evolving a Ka¨hler metric by Ricci flow preserves the Ka¨hler condition, but the Ricci flow does not preserve the Sasaki condition. However, we will show that in the special case where the initial Sasaki metric is η-Einstein, that there is a quasi-Sasaki structure associated to the evolving metric. The geometry of an η-Einstein manifold is special enough that the Ricci flow equation can be reduced to a system of two ordinary differential equations. The qualitative behavior of the flow is discussed in each of the three cases determined by the transverse Ka¨hler- Einstein structure. We then exhibit concrete examples of η-Einstein metrics where the transverse metric gT satisfies RicTg = λg T for λ = −1, 0, 4. We mention in passing the classification theorem of three-dimensional Sasaki manifolds due to Geiges and the uniformization theorem due to Belgun and how our examples tie in. We raise some questions about the general case of evolving an arbitrary Sasaki metric by the Ricci flow. 1.2. Preliminaries In this section we will review the (Riemannian) basics of quasi-Sasaki geometry. Throughout, M = M2n+1 will denote a smooth, oriented, connected, Riemannian manifold of real dimension 2n + 1 for n ≥ 1. The metric will be denoted by g and ∇ will denote the Levi-Civita connection of g. We begin by defining a quasi-Sasaki structure as a certain type of almost contact structure. This is the definition that Blair gave in [3] and the one that we shall work with here. Definition 1.2.1. An almost contact structure on M2n+1 is a triple (ξ, η,Φ) where ξ is a nowhere vanishing vector field (called the Reeb vector field), η is a 1-form with η(ξ) = 1 and Φ is a (1,1)-tensor satisfying Φ2 = −I + ξ ⊗ η. 4 We let Lξ ⊂ TM denote the line bundle generated by ξ. The contact bundle D ⊂ TM is defined to be the kernel of η. We have the direct sum decomposition TM = Lξ ⊕D. The following theorem will be used frequently. Theorem 1.2.2. Suppose M2n+1 has an almost contact structure (ξ, η,Φ). Then Φξ = 0, the endomorphism Φ has rank 2n and η ◦ Φ = 0. Proof. Since Φ2 = −I on D, we see that the rank of Φ is at least 2n. Now Φ2ξ = 0 implies that Φξ 6∈ D \ {0}. Thus Φξ = fξ for some f ∈ C∞(M). But we have that 0 = Φ2ξ = fΦξ = f 2ξ. Since ξ is nowhere vanishing, f = 0. Thus Φξ = 0 and this implies that the rank of Φ is exactly 2n. For any vector field X ∈ TM , we now have η(ΦX)ξ = Φ2ΦX + ΦX = Φ(Φ2X +X) = η(X)Φξ = 0. Therefore η ◦ Φ = 0. Definition 1.2.3. An almost contact metric structure on M2n+1 is a quadruple (ξ, η,Φ, g) where (ξ, η,Φ) is an almost contact structure and g is a Riemannian metric on M that is compatible with Φ in the sense that Φ∗g = g − η ⊗ η. If (ξ, η,Φ, g) is an almost contact metric structure, the metric compatibility with Φ and Φξ = 0 implies that η = g(ξ, ·). Hence η and ξ are dual to each other via the metric g. Given an almost contact structure, a compatible metric always exists. 5 Proposition 1.2.4. Given an almost contact structure (ξ, η,Φ) on M , there is a Riemannian metric g such that (ξ, η,Φ, g) is an almost contact metric structure. Proof. Let g˜ be any Riemannian metric on M . First define a metric g′ by g′(X, Y ) := g˜ ( Φ2X,Φ2Y ) + η(X)η(Y ). Then we have g′(X, ξ) = η(X) for all X ∈ TM . Now define a metric g by g(X, Y ) := 1 2 (g′(X, Y ) + g′(ΦX,ΦY ) + η(X)η(Y )) and check that g(ΦX,ΦY ) = g(X, Y )− η(X)η(Y ) for all X, Y ∈ TM. Let M2n+1 be equipped with an almost contact structure. Consider the manifold M × R with the almost complex structure J : T (M × R)→ T (M × R), defined by J ( f1X + f2 d dt ) = f1ΦX − f2ξ + f1η(X) d dt . (1.1) Here X ∈ TM , f1, f2 ∈ C∞(M × R) and d dt spans the tangent space to R. It is straightforward to verify that J2 = −I. Definition 1.2.5. An almost contact structure on M is called normal if the almost complex structure J , defined by (1.1), is integrable. Recall that an almost complex structure is called integrable if it is induced by an actual complex structure. The celebrated theorem of Newlander and Nirenberg asserts that an almost complex structure J is integrable if and only if the Nijenhuis torsion tensor NJ ≡ 0. For a (1,1)-tensor T , the Nijenhuis torsion tensor NT is the 6 (1,2)-tensor defined by NT (X, Y ) := T 2[X, Y ] + [TX, TY ]− T [TX, Y ]− T [X,TY ]. (1.2) So if our almost product structure is normal, then NJ ≡ 0 and this implies that for any vector fields X, Y ∈ TM , N (1)(X, Y ) := NΦ(X, Y ) + 2dη(X, Y )ξ = 0, N (2)(X, Y ) := (LΦXη)(Y )− (LΦY η)(X) = 0, N (3)(X) := (LξΦ)(X) = 0, N (4)(X) := (Lξη)(X) = 0. Conversely, the vanishing of these four tensor fields implies that NJ ≡ 0. In fact, the vanishing of NJ is equivalent to the vanishing of N (1). A proof of these facts and the following proposition can be found in chapter 6 of [4]. Proposition 1.2.6. If N (1) vanishes then so do N (2), N (3) and N (4). Therefore, an almost contact structure is normal if and only if N (1) ≡ 0. An equation that we will use frequently, and one that holds for any 1-form η is 2dη(X, Y ) = Xη(Y )− Y η(X)− η([X, Y ]). (1.3) The normality condition has some interesting consequences. We record a few of these in the following lemma. 7 Lemma 1.2.7. Let (ξ, η,Φ) be a normal, almost contact structure on M . Then dη(ξ, ·) = 0 and Φ∗dη = dη. Moreover, if g is a compatible metric then the integral curves of ξ are geodesics. Proof. Since N (4)(X) = (Lξη)(X) = 0 for any X ∈ TM , we have 2dη(ξ,X) = ξ(η(X))−X(η(ξ))− η([ξ,X]) = ξ(η(X))− η([ξ,X]) = (Lξη)(X) = 0. Thus dη(ξ, ·) = 0. Cartan’s magic formula says that LΦXη = dιΦXη + ιΦXdη. Since η ◦ Φ = 0, we get (LΦXη)(Y ) = dη(ΦX, Y ). From N (2)(X, Y ) = 0 and the above, we have dη(ΦX, Y ) = dη(ΦY,X). Now replacing Y with ΦY and using dη(ξ, ·) = 0, we see that dη(ΦX,ΦY ) = −dη(Y,X) = dη(X, Y ). For the last claim, notice that g(∇Xξ, ξ) = 12Xg(ξ, ξ) = 0. Again using N (4)(X) = 0, we compute that for any X ∈ TM g(X,∇ξξ) = ξg(X, ξ)− g(∇ξX, ξ) + g(∇Xξ, ξ) = ξ(η(X))− η([ξ,X]) = (Lξη)(X) = 0. Thus ∇ξξ = 0. Hence the integral curves of ξ are geodesics. 8 The fundamental 2-form associated to an almost contact metric structure (ξ, η,Φ, g) is the 2-form ω defined for all X, Y ∈ TM by ω(X, Y ) := g(X,ΦY ). (1.4) From Φ∗g = g − η ⊗ η and η ◦ Φ = 0, it follows that Φ∗ω = ω. Note also that ω(ξ, ·) = 0. Now we are ready to give the definition of a quasi-Sasaki structure. Definition 1.2.8. A quasi-Sasaki structure is a normal, almost contact metric structure whose fundamental 2-form is closed (i.e. dω = 0). A smooth manifold equipped with a quasi-Sasaki structure is called a quasi- Sasaki manifold. A fundamental difference between a Sasaki and a quasi-Sasaki structure has to do with whether or not dη is non-degenerate on D. We can categorize quasi-Sasaki structures according to their rank, which indicates how degenerate dη is. The following definition was introduced in [3]. Definition 1.2.9. The rank of a quasi-Sasaki structure is 2p when (dη)p 6= 0 and η ∧ (dη)p = 0. The rank is 2p+ 1 when η ∧ (dη)p 6= 0 and (dη)p+1 = 0. A quasi-Sasaki structure of rank 1 is commonly called a cosymplectic structure. In this case we have dη = 0. A quasi-Sasaki structure of rank 2n + 1 with ω = dη is Sasaki. The next proposition follows easily from lemma 1.2.7. Proposition 1.2.10. There are no quasi-Sasaki structures of even rank. Proof. Since dη(ξ, ·) = 0, if (dη)p 6= 0 there are vector fieldsX1, . . . , X2p ∈ D such that (dη)p(X1, . . . , X2p) 6= 0. Thus η∧(dη)p(ξ,X1, . . . , X2p) = (dη)p(X1, . . . , X2p) 6= 0. 9 Some facts about the geometry of a quasi-Sasaki manifold are contained in the next lemma. Important is the fact that ξ is a Killing vector field (i.e. Lξg = 0). This means that moving along the integral curves of ξ is an isometry of the metric. Lemma 1.2.11. Let (ξ, η,Φ, g) be a quasi-Sasaki structure on M . Then Lξg = 0, g(∇Xξ, Y ) = dη(X, Y ) and ∇ξΦ = 0. Proof. Since N (3)(Y ) = (LξΦ)(Y ) = 0, we have Φ[ξ, Y ] = [ξ,ΦY ]. Thus we see that (Lξω)(X, Y ) = (Lξg)(X,ΦY ). The Lie derivative Lξω = dιξω + ιξdω = 0. Hence we have (Lξg)(X,ΦY ) = 0. Next, using N (4)(X) = (Lξη)(X) = 0 and ∇ξξ = 0, one computes that (Lξg)(X, η(Y )ξ) = 0. The above shows that (Lξg)(X, (Φ + ξ ⊗ η)Y ) = 0. Since the map Φ + ξ ⊗ η is non-singular, it follows that Lξg = 0. Therefore, ξ is a Klling vector field. For a vector field Z, let θZ be the 1-form defined by θZ(Y ) := g(Y, Z). It is then straightforward to verify that (LZg)(X, Y ) + 2dθZ(X, Y ) = 2g(∇XZ, Y ). Taking Z = ξ and using that ξ is a Killing vector field gives g(∇Xξ, Y ) = dη(X, Y ). With this and the vanishing of tensors N (2) and N (3), a straightforward computation, using that ∇ is Riemannian and torsion free, reveals that ∇ξΦ = 0. Remark 1.2.12. The condition that ξ is a Killing vector field is equivalent to the condition g(∇Xξ, Y ) + g(X,∇Y ξ) = 0 for all X, Y ∈ TM. Also, one can actually 10 show more than just ∇ξΦ = 0. In particular, lemma 6.1 of [4] shows that g((∇XΦ)Y, Z) = dη(ΦY,X)η(Z)− dη(ΦZ,X)η(Y ). (1.5) We end this section with some results which characterize cosymplectic manifolds (those with dη = 0) among quasi-Sasak manifolds. Lemma 1.2.13. A quasi-Sasaki manifold is cosymplectic if and only if ∇Φ = 0 if and only if ∇ω = 0. Proof. If ∇Φ = 0 then NΦ ≡ 0. By normality, 2dη(X, Y )ξ = −NΦ(X, Y ) = 0 for all X, Y ∈ TM . Hence dη = 0. Conversely, if dη = 0 then from equation (1.5) we conclude that ∇Φ = 0. The second if and only if statement follows easily from ω = g(·,Φ·) and ∇g = 0. Lemma 1.2.14. A quasi-Sasaki manifold is cosymplectic if and only if ∇η = 0. Proof. Since ξ is a Killing vector field and η = g(ξ, ·), we compute that (∇Xη)(Y ) = g(∇Xξ, Y ) = −g(∇Y ξ,X) = −(∇Y η)(X). We can write equation (1.3) as 2dη(X, Y ) = (∇Xη)(Y )− (∇Y η)(X). Combining these equations we have dη(X, Y ) = (∇Xη)(Y ) and the result follows. Remark 1.2.15. The cosymplectic case is the only one where the contact subbundle D is integrable. Indeed, it follows from equation (1.3) that a subbundle defined as the kernel of a closed 1-form is involutive, hence integrable (see definition 5.0.2 and 11 theorem 5.0.3). Furthermore, from ∇η = 0 we see that D is a parallel distribution. As TM = Lξ ⊕D and both Lξ and D are integrable in this case, it follows that M is locally the product of a 1-dimensional manifold (line or circle) and a Ka¨hler manifold. 1.3. Transverse Ka¨hler Geometry In this section we discuss the basics of Riemannian foliations and the associated transverse geometry, as it applies to our study of quasi-Sasaki manifolds. For a more thorough treatment of foliations, the interested reader should see [30] and/or [39]. Let (M, ξ, η,Φ, g) be a quasi-Sasaki manifold. Recall that Lξ is the line bundle generated by ξ and D = ker(η). From Φ∗g = g − η ⊗ η and Φξ = 0 we see that the decomposition TM = Lξ ⊕ D is orthogonal with respect to g. Hence D = L⊥ξ . The integral submanifolds of Lξ, which are the integral curves of ξ, define the leaves of a 1-dimensional foliation of M , which we denote by Fξ. We let ν(Fξ) := TM/Lξ denote the normal bundle of the foliation. Then we have the short exact sequence 0→ Lξ → TM → ν(Fξ)→ 0. The metric provides a splitting σ : ν(Fξ)→ D of this short exact sequence. Explicitly, for X¯ ∈ ν(Fξ) we have X¯ = X + Lξ for some X ∈ TM . Then σ(X¯) = X − g(X, ξ)ξ. Clearly σ(X¯) ∈ D and it is straightforward to check that σ is a well-defined bijection. Thus we may identify D and ν(Fξ). A transverse metric is a Riemannian metric on the quotient bundle ν(Fξ). Since D ' ν(Fξ), we define the transverse metric gT by gT := g|D. 12 The orthogonal projection piD : TM → D is given by piD(X) = X − η(X)ξ. Observe: gT (X, Y ) = g(piD(X), piD(Y )) = g(X, Y )− η(X)η(Y ) = Φ∗g(X, Y ). Thus the transverse metric is related to the quasi-Sasaki metric via g = η⊗η+gT . We want to know, under what conditions does the transverse metric reflect the geometry of the leaf space M/Fξ? To answer this question, we need the next definition. Definition 1.3.1. A vector field X ∈ TM is called foliate with respect to Fξ if [X, ξ] ∈ Lξ. A Riemannian metric g is called bundle-like with respect to Fξ if for any foliate vector fields X and Y which are orthogonal to Lξ, we have ξg(X, Y ) = 0. Bundle-like metrics are important in the theory of foliations because they provide globally defined transverse metrics which are locally the pull-back of a metric on the local Riemannian quotient. The following is proposition 2.5.7 in [5]. Proposition 1.3.2. If g is bundle-like with respect to Fξ, then Fξ is a Riemannian foliation. Conversely, if Fξ is a Riemannian foliation with transverse metric gT , then there is a bundle-like metric g whose associated transverse metric is gT . Thus we see that the transverse metric is locally the pull-back of a metric on the local Riemannian quotient if and only if g is bundle-like. In this way, bundle-like metrics reflect the geometry of the leaf space M/Fξ. The next lemma is immediate. Lemma 1.3.3. Quasi-Sasaki metrics are bundle-like with respect to Fξ. Proof. This follows easily from definition 1.3.1 and the fact that Lξg = 0. 13 The following definition is used to describe the behavior of the orbits of ξ. Definition 1.3.4. The foliation Fξ is quasi-regular if each x ∈M has a neighborhood Ux such that any leaf of Fξ passes through Ux at most mx ∈ Z+ times. If mx = 1 for all x ∈ M , then Fξ is called regular. If for some point, no such m exists, then Fξ is called irregular. If all of the orbits of ξ are compact then Fξ is quasi-regular. In this case, ξ generates an S1 action on M . If this action is free, then Fξ is regular. If the S1 action is only locally free, then Fξ is strictly quasi-regular. Assuming M is compact, quasi-regularity implies that the orbits of ξ are all compact. Thus, if ξ has a non- compact orbit, then Fξ is irregular. We will often say that M is regular (quasi-regular, irregular) to mean that Fξ is regular (quasi-regular, irregular). Example 1.3.5. A good example to keep in mind is the weighted Sasaki structure on S3 = {(z0, z1) ∈ C2 : |z0|2 + |z1|2 = 1}. Pick real numbers a0, a1 > 0 and consider the vector field ξ = √−1 ∑ j aj ( zj ∂ ∂zj − z¯j ∂ ∂z¯j ) . Restricting to S3, ξ is everywhere tangent to S3 and nowhere vanishing. The 1-form η = − √−1 2 ∑ j 1 aj ( z¯jdz j − zjdz¯j ) is a contact form with η(ξ) = 1. On D ⊂ TS3, we let Φ be the restriction of the standard complex structure on C2 and extend it trivially in the ξ direction. Then g = η ⊗ η + dη(Φ·, ·) defines a Sasaki-metric on S3. The integral curve of ξ through the point (z0, z1) ∈ S3 is given by (z0, z1) 7→ (e2piia0tz0, e2piia1tz1). 14 Choosing a0 = a1 = 1 gives the standard Sasaki structure on S 3. If a0 and a1 are chosen to be relatively prime integers, then Fξ is regular. If a0 and a1 are not coprime or a0, a1 ∈ Q \ Z, then Fξ is quasi-regular. Suppose that a0a1 is irrational. Orbits through points of the form (z0, 0) are circles as they return to (z0, 0) at time t = a−10 . Similarly, orbits through (0, z1) are circles that return in time t = a −1 1 . However, the orbits through points (z0, z1) where z0z1 6= 0 are not compact. Since a0 a1 is irrational, there is no time t such that both a0t and a1t are integers. Thus the integral curve never returns to (z0, z1). The closure of these orbits is the torus T 2. This example easily generalizes to higher odd dimensional spheres. Adapting theorem 6.3.8 of [5] to the special case of a quasi-Sasaki manifold, we get the following structure theorem. This important theorem will be used frequently in the sequel. See chapter 4.3 of [5] for the definition of the orbifold cohomology. If the orbifold is in fact a manifold, then the orbifold cohomology is the usual cohomology. Theorem 1.3.6. Let (M, ξ, η,Φ, g) be a quasi-Sasaki manifold of rank 2p + 1 such that the leaves of Fξ are all compact. Then 1. The leaf space M/Fξ is a Ka¨hler orbifold such that the canonical projection pi : M → M/Fξ is an orbifold Riemannian submersion. The fibers are totally geodesic submanifolds of M diffeomorphic to S1. 2. M is the total space of a principal S1-orbibundle over M/Fξ with connection 1-form η. The curvature form dη = pi∗Ω where Ω is a closed (1,1)-form of rank p on M/Fξ such that 12pi [Ω] ∈ H2orb(M/Fξ,Z). 3. If Fξ is regular, then M is the total space of a principal S1-bundle over the Ka¨hler manifold M/Fξ. 15 Now we will show that the transverse geometry of a quasi-Sasaki manifold has a Ka¨hler structure. From Φ2 = −I + ξ ⊗ η, we have (Φ|D)2 = −I. Hence Φ|D is an almost complex structure on D. We consider the complexified contact bundle DC := D ⊗R C. Then Φ|D extends C-linearly to an almost complex structure on DC, which we continue to call Φ. Let D1,0 and D0,1 denote the √−1 and −√−1 eigenspaces of Φ. Then DC = D1,0 ⊕ D0,1. Complex conjugation is an R-linear isomorphism between D1,0 and D0,1. As the transverse metric is Φ-invariant (i.e. Φ∗gT = gT ), it defines a Hermitian metric on D1,0. Explicitly, we C-linearly extend gT in both entries to a symmetric, complex bilinear form on DC, which we continue to call gT . Then for any U, V ∈ D1,0, we define the Hermitian metric h by h(U, V ) := gT (U, V¯ ). The map σ : D → D1,0 that sends X 7→ 1 2 ( X −√−1ΦX) is an isomorphism of complex vector bundles (D,Φ) ' (D1,0,√−1). Under this isomorphism, a straightforward computation reveals that σ∗h = 1 2 ( gT + √−1ω) . It is well-known that the holomorphic tangent bundle of an almost complex manifold is involutive if and only if the almost complex structure is integrable. In the almost contact setting, we have a similar theorem. Theorem 1.3.7. An almost contact structure is normal if and only if [D1,0, Lξ] ⊂ D1,0 and [D1,0,D1,0] ⊂ D1,0. 16 Proof. By proposition 1.2.6, normality is equivalent to NΦ + 2dη ⊗ ξ ≡ 0. Now let X, Y ∈ D1,0. Using equation (1.3) and Φ2 = −I + ξ ⊗ η, we compute that NΦ(X, Y ) + 2dη(X, Y )ξ = −2[X, Y ]− 2 √−1Φ[X, Y ], NΦ(X, Y¯ ) + 2dη(X, Y¯ )ξ = 0, NΦ(X, ξ) + 2dη(X, ξ)ξ = −[X, ξ]− √−1Φ[X, ξ]. So NΦ+2dη⊗ξ ≡ 0 if and only if Φ[X, Y ] = √−1[X, Y ] and Φ[X, ξ] = √−1[X, ξ]. Corollary 1.3.8. An almost contact structure is normal if and only if the Nijenhuis torsion tensor NΦ vanishes on D1,0 ⊕ Lξ. Proof. Since Φ∗dη = dη, we have dη(X, Y ) = 0 for X, Y ∈ D1,0. By lemma 1.2.7, dη(X, ξ) = 0. The corollary now follows from the computation in theorem 1.3.7. So for a quasi-Sasaki structure, the contact bundle D is endowed with a complex structure Φ and a Hermitian metric whose Ka¨hler form is closed. Thus the data (D,Φ, ω) gives M a transverse Ka¨hler structure. We define a connection ∇T on D by ∇TXY := piD (∇XY ) , X ∈ D, ∇Tξ Y := piD [ξ, Y ] . One can check that ∇T is the unique, torsion-free connection on D with ∇TgT = 0. For this reason we call ∇T the transverse Levi-Civita connection. We define the transverse curvature operator for X, Y ∈ D by RT (X, Y ) := ∇TX∇TY −∇TY∇TX −∇T[X,Y ]. 17 With this we can then define the transverse Riemannian curvature tensor, the transverse Ricci curvature and the transverse scalar curvature in the obvious way. When computing transverse curvatures, we have all of the familiar formulas from Ka¨hler geometry. Next we briefly mention the transverse de Rham cohomology and its associated Hodge theory. For a detailed discussion, see [15] and [24]. Definition 1.3.9. A function f ∈ C∞(M) is called basic if df(ξ) = 0. A p-form θ ∈ Ωp(M) is called basic if ιξθ = 0 and Lξθ = 0. We let ΛpB(Fξ) be the sheaf of germs of basic p-forms and ΩpB(Fξ) the set of global sections of ΛpB(Fξ). Observe that if θ is basic then so is dθ. This follows from Lξθ = ιξdθ + dιξθ and Lξdθ = dLξθ. We set dB := d|ΩpB . Then dB : Ω p B → Ωp+1B is well-defined, dB ◦ dB = 0 and we have a subcomplex of the de Rham complex: 0→ C∞B → Ω1B → · · · → Ω2nB → 0. We denote the cohomology ring by H∗B(Fξ) and call it the basic cohomology ring. There is a transverse Hodge theory for the basic cohomology ring and in many ways it resembles the de Rham-Hodge theory of a Ka¨hler manifold. The transverse Hodge star, ∗B, is defined in terms of the usual Hodge star by ∗Bα := ∗(η ∧ α). The adjoint operator δB : Ω p B → Ωp−1B of dB is defined by δB := − ∗B ◦dB ◦ ∗B. We define the basic Hodge Laplacian (with respect to dB) by 4B := dBδB + δBdB. We say that θ ∈ ΩpB is harmonic if 4Bθ = 0. The transverse Hodge theorem of [14] states that every basic cohomology class has a unique harmonic representative. We define the transverse tensor Laplacian by ∆T := ∇T∇T . For a basic function f , one 18 can check that 4Bf = −∆Tf = −gij¯∂i∂j¯f. The transverse complex structure Φ induces a natural decomposition of the basic p-forms into basic (i, j)-forms where i+ j = p: ΩpB ⊗R C = ⊕i+j=pΩi,jB and the exterior derivative splits as dB = ∂B + ∂¯B where ∂B : Ω i,j B → Ωi+1,jB and ∂¯B : Ωi,jB → Ωi,j+1B . From d2B = 0 we get ∂ 2 B = ∂¯ 2 B = 0 and ∂B∂¯B + ∂¯B∂B = 0. As we did above, we can define the basic Hodge Laplacians with respect to ∂B and ∂¯B. We can also define a transverse Lefschetz operator L : θ 7→ θ ∧ ω. One can then recover the expected transverse Ka¨hler identities. We set dcB := √−1 2 (∂¯B − ∂B). Then we have dBd c B = √−1∂B∂¯B and dcB ◦ dcB = 0. The transverse Ricci form ρT is the basic 2-form defined by ρT := RicT (·,Φ·). Just as in the Ka¨hler setting, one can show that ρT = −√−1∂B∂¯B log(det(gT )). 19 Thus ρT is a dB closed (1,1)-form and so defines a basic cohomology class. The class [ρT ] ∈ H2B(Fξ) is independent of the choice of transverse Ka¨hler metric and, as in the Ka¨hler setting, 2pic1B = [ρ T ], where c1B is the basic first Chern class. 1.4. Transverse Distance and Preferred Coordinates First we introduce the transverse distance function that will be important for us in chapter III. Let (M, g) be quasi-Sasaki and assume that M is compact and quasi-regular. For a point x ∈M , let γx denote the orbit of ξ through x. Since γx is compact for all x ∈M , we can make the following definition: Definition 1.4.1. The transverse distance function dT is defined on M by dT (x, y) := d(γx, γy) where d is the geodesic distance function on M , determined by g. The transverse ball of radius r > 0 about a point x ∈M is defined as BT (x, r) := {y ∈M : dT (x, y) < r}. The transverse diameter of M is the number ΘT defined as ΘT := max x,y∈M dT (x, y). The transverse distance function defines a metric on the leaf space M/Fξ. It turns out to be the same as the geodesic distance function on M/Fξ, determined by gT . In order to prove that dT is a metric, we need to show that it satisfies the triangle inequality. To do so, we make use of the following proposition: 20 Proposition 1.4.2. For x, y ∈ M and p ∈ γx, there is q ∈ γy such that dT (x, y) = d(p, q). Proof. Since the curves γx and γy are compact, there are points x1 ∈ γx and y1 ∈ γy such that dT (x, y) = d(x1, y1). Now let σ : [0, 1] → M be the length minimizing geodesic with σ(0) = x1 and σ(1) = y1. Flowing along γx is an isometry of M . As γx is compact, it is diffeomorphic to S1. Thus there is ζ ∈ S1 such that ζ(x1) = p. Let q := ζ(y1) and σ˜ = ζ ◦ σ. Then σ˜ is a geodesic connecting p and q whose length is the same as σ. Therefore, dT (x, y) = d(x1, y1) = d(p, q). Corollary 1.4.3. The transverse distance function satisfies the triangle inequality. For fixed x0 ∈M , r(x) := dT (x, x0) is a basic Lipschitz function. Proof. Pick points x, y, z ∈M . By the previous proposition, there are points p ∈ γx, q ∈ γy and s ∈ γz such that dT (x, y) = d(p, q) and dT (y, z) = d(q, s). Thus dT (x, z) ≤ d(p, s) ≤ d(p, q) + d(q, s) = dT (x, y) + dT (y, z). Let r be as above. By the triangle inequality, |r(x)−r(y)| ≤ dT (x, y) ≤ d(x, y). Thus r is Lipschitz continuous with Lipschitz constant 1. Hence r is differentiable almost everywhere and |∇r| ≤ 1 at any point of differentiability. Since r is constant along the orbits of ξ, dr(ξ) = 0. Therefore, r is basic. We conclude this chapter by introducing the local coordinates and local frame that we will use in our computations in chapter II, III and IV. By normality, the almost complex structure J , defined by (1.1), on M × R is integrable; hence it is induced by a complex structure. This means that there are (real) local coordinates (x0, y0, . . . , xn, yn) on M × R such that J∂xj = ∂yj and J∂yj = −∂xj . 21 We will identify M as a submanifold of M × R via M ' M × {0}. Let t be the coordinate on R. Then the tangent bundle T (M×R) = TM⊕R[ d dt ] = Lξ⊕D⊕R[ ddt ]. Observe that Jξ = d dt and J d dt = −ξ. Thus, Lξ ⊕ R[ ddt ] forms an involutive, complex subbundle of T (M ×R). By the complex version of the Frobenius theorem, there are local coordinates as above with ∂x0 = ξ and ∂y0 = d dt . Thus y0 = t. Now consider the complexification T (M × R) ⊗R C and C-linearly extend J . For j = 0, . . . , n, put zj = 1 2 (xj − √−1yj). Then (z0, . . . , zn, z¯0, . . . , z¯n) are local coordinates on M × R with J∂zj = √−1∂zj and J∂z¯j = − √−1∂z¯j . We know that ∂z0 = 1 2 (ξ − √−1 d dt ). For j ≥ 1, we can write ∂zj = Zj + fj ddt for some vector field Zj ∈ TM ⊗R C and some complex-valued function fj. By the definition of J , J∂zj = ΦZj − fjξ + η(Zj) d dt . But from J∂zj = √−1∂zj , we have J∂zj = √−1 ( Zj + fj d dt ) . Equating the two expression for J∂zj , we find that fj = − √−1η(Zj) and ΦZj = √−1(Zj − η(Zj)ξ). Setting t = 0, the local coordinates (z0, . . . , zn, z¯0, . . . , z¯n) on M ×R induce local coordinates on M . Since t = y0, we get the local coordinates (x, z1, . . . , zn, z¯1, . . . , z¯n) where x = x0. The coordinate vectors satisfy ∂x = ξ and Φ∂zj = √−1(∂zj − η(∂zj)ξ). We call these preferred local coordinates. 22 Since η is a real 1-form and dη is a basic (1,1)-form, in a small coordinate neighborhood U we can write η = dx−√−1(hjdzj − hj¯dz¯j) where h : U → R is a (local) real basic function (i.e. ξh = ∂xh = 0). We use the notation hj¯ = ∂z¯jh = ∂h ∂z¯j = ∂h ∂zj = ∂zjh = hj. Then, defining dηij¯ := hij¯, we have dη = √−1dηij¯dzi ∧ dz¯j. Now, for j = 1, . . . , n, define vector fields Xj := ∂zj + √−1hjξ and Xj¯ := X¯j = ∂z¯j − √−1hj¯ξ. Notice that η(Xj) = η(X¯j) = 0, while ΦXj = √−1Xj and ΦX¯j = − √−1X¯j. Thus, D1,0 = span{Xj} andD0,1 = span{X¯j}. We call the frame (ξ,X1, . . . , Xn, X¯1, . . . , X¯n) a preferred local frame. Since dη(·, ξ) = 0, we see that dη(Xi, X¯j) = dη(∂zi , ∂z¯j) = √−1dηij¯. Since the hj are basic functions, we compute that [Xi, Xj] = [X¯i, X¯j] = [ξ,Xi] = [ξ, X¯j] = 0 and [Xi, X¯j] = −2 √−1dηij¯ξ. 23 From the compatibility of Φ with g, in preferred local coordinates we can write g = η ⊗ η + gTij¯dzi ⊗ dz¯j where gTij¯ = g T ji¯ = gTj¯i. Since Lξg = Lξη = 0, the functions gTij¯ are basic. The transverse metric and Ka¨hler form are, respectively, gT = gTij¯dz i ⊗ dz¯j and ω = √−1gTij¯dzi ∧ dz¯j. Hence gT (Xi, X¯j) = g T (∂zi , ∂z¯j) = g T ij¯ and ω(Xi, X¯j) = ω(∂zi , ∂z¯j) = √−1gTij¯. The transverse metric defines a pointwise norm on basic forms. If α = αidz i and β = βj¯dz¯ j are basic (1,0) and (0,1)-forms, respectively, then we define |α|2gT = gij¯αiαj and |β|2gT = gij¯βj¯βi¯. We can then extend this to basic forms of type (p, q). For example, the norm of a (1, 1)-form θ = θij¯dz i ∧ dz¯j is given by |θ|2gT = gij¯gkl¯θil¯θjk¯. At a point where gij¯ = δij, we see that |θ|2gT = ∑ i,k θik¯θik¯ = ∑ i,k |θik¯|2, as we would expect. For any basic form θ, the norm with respect to the quasi-Sasaki metric g is twice that of the norm with respect to the transverse metric gT ; |θ|2g = 2|θ|2gT . 24 CHAPTER II RIGIDITY OF QUASI-SASAKI-EINSTEIN METRICS 2.1. The Ricci tensor and Einstein equation A Sasaki metric g and dη are intimately related by g = η ⊗ η + dη(Φ·, ·). It is not so clear, a priori, how a quasi-Sasaki metric and dη are related. In this section we compute the Ricci tensor of a quasi-Sasaki metric. To better understand the relationship between the metric and dη, we want to know how the Ricci tensor depends on dη. Given that a quasi-Sasaki structure has a Riemannian foliation associated to it, one could compute the Ricci tensor of g using curvature formulas involving the O’Neill tensors A and T (see theorems 2.5.16 and 2.5.18 of [5]). However, these curvature formulas will not give us what we want directly and to use them requires a sufficiently large amount of computation anyway. Thus, we will present a barehanded computation of the Ricci tensor and we will see the role that dη plays in the curvature. We shall compute the Ricci tensor with respect to a preferred local frame {ξ,Xi, X¯i}. Recall that {Xi} is a local frame for D1,0, gT (Xi, X¯j) = gTij¯ and [ξ,Xi] = [ξ, X¯i] = 0. In lemma 1.2.7 we showed that ∇ξξ = 0. Now we observe that ∇Xiξ = √−1gkj¯dηij¯Xk and ∇X¯iξ = ∇Xiξ = − √−1gjk¯dηji¯X¯k. (2.1) To see this, write∇Xiξ = αξ+βjXj+γkX¯k and use g(∇Xξ, Y ) = dη(X, Y ), keeping in mind that dη is a basic (1,1)-form. Let piLξ : TM → Lξ be the orthogonal projection onto Lξ. Proposition 2.5.14 in [5] asserts that for X, Y ∈ D, piLξ (∇XY ) = 1 2 piLξ ([X, Y ]) . (2.2) 25 Writing [X, Y ] = αξ+βiXi +γ iX¯i and applying η to both sides yields α = η([X, Y ]). Hence piLξ([X, Y ]) = η([X, Y ])ξ. By equation (1.3), for X, Y ∈ D, we have η([X, Y ]) = −2dη(X, Y ). Therefore, by (2.2) and the above, piLξ (∇XY ) = −dη(X, Y )ξ. (2.3) Now, as ∇XY = ∇TXY + piLξ (∇XY ) for X, Y ∈ D, equation (2.3) yields ∇XY = ∇TXY − dη(X, Y )ξ. (2.4) Since the transverse metric is Ka¨hler and dη is a (1,1)-form, it follows from (2.4) that ∇XiXj = ∇TXiXj and ∇XiX¯j = − √−1dηij¯ξ. (2.5) Equations (2.1) and (2.5) along with ∇ξξ = 0 allow us to compute the curvature tensor of a quasi-Sasaki structure with respect to our preferred frame. For notational convenience, we will drop the superscript T from the transverse metric gT . First we compute R(Xi, ξ)ξ = ∇Xi∇ξξ −∇ξ∇Xiξ −∇[Xi,ξ]ξ = 0−√−1gkj¯dηij¯∇ξXk − 0 = gkj¯gmp¯dηij¯dηkp¯Xm. Hence we have g(R(Xi, ξ)ξ,Xl¯) = g kj¯dηij¯dηkl¯ = g kj¯dηij¯dηlk¯ and therefore, Ric(ξ, ξ) = 2gil¯g(R(Xi, ξ)ξ,Xl¯) = 2g il¯gkj¯dηij¯dηlk¯ = |dη|2g ≥ 0. 26 Next we will compute Ric(Xj, ξ). We start with R(X¯i, ξ)Xj = ∇X¯i∇ξXj −∇ξ∇X¯iXj −∇[X¯i,ξ]Xj = √−1gkp¯∇X¯idηjp¯Xk − √−1∇ξ(dηji¯ξ)− 0 = √−1gkp¯ (X¯idηjp¯Xk + dηjp¯∇X¯iXk)− 0 = √−1gkp¯ (X¯idηjp¯Xk +√−1dηjp¯dηki¯ξ) . From equation (2.5) we have [Xi, Xj¯] = −2 √−1dηij¯ξ. Using this we compute that R(X¯i, Xj)ξ = ∇X¯i∇Xjξ −∇Xj∇X¯iξ −∇[X¯i,Xj ]ξ = √−1gkp¯ (∇X¯idηjp¯Xk)+√−1gmq¯ (∇Xjdηmi¯Xq¯)− 2√−1dηji¯∇ξξ = √−1gkp¯ (X¯idηjp¯Xk + dηjp¯∇X¯iXk)+√−1gmq¯ (Xjdηmi¯Xq¯ + dηmi¯∇XjXq¯)− 0 = √−1gkp¯ (X¯idηjp¯Xk +√−1dηjp¯dηki¯ξ)+√−1gmq¯ (Xjdηmi¯Xq¯ −√−1dηmi¯dηjq¯ξ) . Now we can compute the Ricci curvature Ric(Xj, ξ) = g l¯i ( g(R(Xl, Xj)ξ, X¯i) + g(R(X¯i, Xj)ξ,Xl) ) = g l¯i ( g(R(X¯i, ξ)Xj, Xl) + g(R(X¯i, Xj)ξ,Xl) ) = 0 + √−1g l¯i(∇Xjgmq¯dηmi¯)glq¯ = √−1∇Xjgmi¯dηmi¯ = ∇XjtrgT (dη) where trgT denotes the trace by the transverse metric g T . For basic forms, trgT = 1 2 trg. 27 Lastly we want to compute Ric(X¯j, Xk). We start by computing R(Xi, X¯j)Xk. Since we will eventually pair this with Xl¯, we will neglect any terms in the ξ direction. R(Xi, X¯j)Xk = ∇Xi∇X¯jXk −∇X¯j∇XiXk −∇[Xi,X¯j ]Xk = ∇Xi( √−1dηkj¯ξ +∇TX¯jXk)−∇X¯j(∇TXiXk) + 2 √−1dηij¯∇ξXk = √−1dηkj¯∇Xiξ +∇Xi∇TX¯jXk −∇X¯j∇TXiXk − 2gmp¯dηij¯dηkp¯Xm = −gmp¯dηkj¯dηip¯Xm +∇TXi∇TX¯jXk −∇TX¯j∇TXiXk − 2gmp¯dηij¯dηkp¯Xm. As [ξ,Xk] = 0, ∇TξXk = piD([ξ,Xk]) = 0. Since [Xi, X¯j] = −2 √−1dηij¯ξ, it follows that ∇T [Xi,X¯j ] Xk = 0. Thus the above can be written as R(Xi, X¯j)Xk = R T (Xi, X¯j)Xk − gmp¯ ( dηkj¯dηip¯ + 2dηij¯dηkp¯ ) Xm. By a computation similar to the one above, modulo the terms in the ξ direction, R(X¯l, X¯j)Xk = R T (X¯l, X¯j)Xk + g mp¯ ( dηkj¯dηml¯ − dηkl¯dηmj¯ ) X¯p. Now we are ready to compute the Ricci curvature. Using the above we find Ric(X¯j, Xk) = g il¯ ( g(R(Xi, X¯j)Xk, Xl¯) + g(R(X¯l, X¯j)Xk, Xi) ) + g(R(ξ, X¯j)Xk, ξ) = RicT (X¯j, Xk)− 2gil¯dηij¯dηkl¯. The above computations prove our next proposition. 28 Proposition 2.1.1. With respect to a preferred frame, the Ricci tensor of a quasi- Sasaki structure (ξ, η,Φ, g) satisfies Ric(ξ, ξ) = 2gil¯gkj¯dηij¯dηkl¯ = |dη|2g, (2.6) Ric(Xk, ξ) = √−1∇Xkgij¯dηij¯ = ∇XktrgT (dη), (2.7) Ric(Xi, X¯l) = Ric T (Xi, X¯l)− 2gkj¯dηij¯dηkl¯. (2.8) If our quasi-Sasaki structure is actually Sasaki, then the above equations should yield the well-known formulae for the Ricci tensor of a Sasaki metric. In the Sasaki setting, one has dηij¯ = gij¯. Substituting this into the above equations we get Ric(ξ, ξ) = 2n, Ric(Xk, ξ) = 0, Ric(Xi, X¯l) = Ric T (Xi, X¯l)− 2g(Xi, X¯l), which is indeed correct. Recall that an Einstein metric g is one where Ricg = λg for some constant λ ∈ R. Proposition 2.1.1 gives us more information about the nature of a quasi- Sasaki-Einstein metric. Since ξ has unit length, Lξ ⊥ D and g(Xi, X¯j) = gT (Xi, X¯j), proposition 2.1.1 gives us our next result. Proposition 2.1.2. A quasi-Sasaki metric is Einstein with Ricg = λg if and only if |dη|2g = λ, (2.9) ∇XktrgT (dη) = 0, (2.10) RicT (Xi, X¯l)− 2gkj¯dηij¯dηkl¯ = λgT (Xi, X¯l). (2.11) 29 Remark 2.1.3. The above proposition lends us some useful information. First we see that λ ≥ 0 and that the norm of dη is constant over M . If λ > 0 (i.e. dη 6= 0), then by Myers’s theorem M is compact. Since dη is basic, the second equation tells us that the trace of dη is constant. Since dη is an exact (1,1)-form, this implies that δ(dη) = 0 and therefore dη is harmonic. Finally, the trace of the third equation by the transverse metric yields RT = λ(n+ 1) ≥ 0. Thus the transverse scalar curvature is constant. This implies that the transverse Ricci form ρT is harmonic. Proposition 2.1.4. If a Ka¨hler metric has constant scalar curvature then the Ricci form is harmonic. Proof. Let g = gαβ¯dz α ⊗ dzβ¯ be a Ka¨hler metric and ρ = √−1Rαβ¯dzα ∧ dzβ¯ the Ricci form. Recall that ρ is a closed (1,1)-form and Rαβ¯ = −∂α∂β¯ log det(g). Using δ = −gklι∂k∇∂l we compute that δρ = √−1gλµ¯ ( Rαµ¯/λdz α −Rλβ¯/µ¯dzβ¯ ) = √−1gλµ¯ ( Rλµ¯/αdz α −Rλµ¯/β¯dzβ¯ ) = √−1 ( ∂R ∂zα dzα − ∂R ∂zβ¯ dzβ¯ ) . So if the scalar curvature R is constant, then δρ = dρ = 0. Hence ρ is harmonic. Remark 2.1.5. If the manifold is compact, then the converse to the above is true. That is, a compact Ka¨hler metric with harmonic Ricci form has constant scalar curvature. To see this, let Λ denote the dual Lefschetz operator and recall that Λρ = R and Λ commutes with the Hodge Laplacian 4 = dδ + δd. So if ρ is harmonic then so is R. Harmonic functions on compact manifolds are constant. Hence R is constant. 30 2.2. The Regular Case Let (M, ξ, η,Φ, g) be a quasi-Sasaki manifold. In this section we assume that the orbits of ξ are compact and the induced S1 action is free. Hence Fξ is regular. By theorem 1.3.6, M is the total space of a principal circle bundle over the Ka¨hler manifold B := M/Fξ. The transverse geometry of M is the Ka¨hler geometry of (B, gT ). For a manifold B (not necessarily Ka¨hler), we let P(B, S1) denote the collection of all principal circle bundles pi : P → B. We will quickly review some facts about P(B, S1). For a reference to these facts, see chapter two of [4]. Theorem 2.2.1. There is a binary operation on P(B, S1) giving it the structure of a group isomorphic to H2(B,Z). For any integer m 6= 0, the cyclic group Zm := Z/mZ is a subgroup of S1. Thus, for any P ∈ P(B, S1) there is a Zm action on P and P/Zm is a principal S1/Zm bundle over B. Since S1/Zm ' S1, we can consider P/Zm ∈ P(B, S1). Theorem 2.2.2. For P ∈ P(B, S1) and an integer m 6= 0, P/Zm ' mP. This implies that if P is simply connected and |m| > 1, then P 6= mP ′ for any P ′ ∈ P(B, S1). We let V := ker(pi∗) ⊂ TP and call it the vertical distribution. For θ ∈ S1, we define θˆ : P → P by p 7→ θ · p. A connection on P ∈ P(B, S1) is a smooth distribution H ⊂ TP such that TP = H ⊕ V and θˆ∗(H) = H for all θ ∈ S1. We call H the horizontal distribution. A point p ∈ P induces a map pˆ : S1 → P by θ 7→ θ · p. The image of pˆ is the orbit through p. It is diffeomorphic to pi−1(pi(p)) ' S1. At 1 ∈ S1, the differential of pˆ is a map pˆ∗ : R → TpP. The image of pˆ∗ is Vp. The connection form of H is the S1 invariant 1-form η on P such that for X ∈ TpP , η(X) = t where pˆ∗(t) is the vertical part of X. Note that H = ker(η). For a principal S1 bundle, the curvature form of the connection H is dη. 31 Lemma 2.2.3. Given a connection form η on P ∈ P(B, S1), there exists a unique, closed 2-form Ω on B such that 1 2pi [Ω] ∈ H2(B,Z) and dη = pi∗Ω. The cohomology class of Ω is independent of the choice of connection form. The cohomology class 1 2pi [Ω] is called the characteristic class of the principal circle bundle P . Theorem 2.2.1 shows that P(B, S1) is in bijection with H2(B,Z). The above lemma and the following theorem due to Kobayashi in [28] show that specifying a connection form η on P corresponds to choosing a representative of the characteristic class. Principal circle bundles are classified by their characteristic class. Theorem 2.2.4. Let Ω be a closed 2-form on B with 1 2pi [Ω] ∈ H2(B,Z). Then there is P ∈ P(B, S1) and a connection form η on P such that dη = pi∗Ω. If g is Einstein with Ricg = λg, the first equation in proposition 2.1.2 shows that λ ≥ 0. If λ = 0 then dη = 0 and by (2.11), RicgT = 0. Therefore, M is cosymplectic and B is Calabi-Yau. We know already that M is locally the product of a circle and a Ka¨hler manifold (see remark 1.2.15), but in this case we can say more. Since dη = 0, the characteristic class 1 2pi [Ω] = 0. This means that M is the trivial circle bundle over B. Therefore, M = S1 ×B. This proves our next theorem. Theorem 2.2.5. If (M, g) is a quasi-Sasaki manifold with Ricg = 0 and Fξ is regular with compact leaves, then M = S1 ×B where B is a Calabi-Yau manifold and g is a product metric. Now we assume that g is Einstein with Ricg = λg for λ > 0. Then by Myers’s theorem M is compact, in which case the regularity of Fξ is equivalent to the assumption that the orbits of ξ are compact and the S1 action is free. The (1,1)- tensor θ associated to dη is defined implicitly by g(θ(X), Y ) = dη(X, Y ). 32 If we write θ(Xi) = θ k iXk, then θ k i = √−1gkj¯dηij¯. With this and proposition 2.1.2, the condition that g is Einstein with Ricg = λg is equivalent to |dη|2g = λ, 4Bdη = 0, RicgT (X, Y ) = λg T (X, Y ) + 2gT (θ(X), θ(Y )). From the third equation above we see that RicgT > 0. By theorem 1 of [27], B is simply connected. Up to a factor of 2pi, the Ricci form of gT is a representative of the first Chern class of B. Thus c1(B) > 0. Before we give the main theorem of this section, we will prove the existence of quasi-Sasaki-Einstein metrics on certain principal circle bundles. These examples illustrate the general case. Let (Bnii , gi), i = 1, . . . , k, be Ka¨hler-Einstein manifolds with complex dimension ni and c 1(Bi) > 0. We can write c 1(Bi) = qiαi where qi ∈ Z+ and αi ∈ H2(Bi,Z) is an indivisible class. By scaling the metrics appropriately, we can assume that Ricgi = qigi. Then we have 2pic 1(Bi) = [ρi] = qi[ωi] and, hence, 1 2pi [ωi] = αi is an indivisible, integral class. Let B = B1 × · · · × Bk and pii : B → Bi the projections. Pick integers bi, not all zero, and set Ω = b1pi ∗ 1ω1 + · · ·+ bkpi∗kωk. Then 0 6= 1 2pi [Ω] ∈ H2(B,Z), so by theorem 2.2.4 there is a nontrivial principal circle bundle pi : M → B and a connection form η on M such that dη = pi∗Ω. Theorem 2.2.6. There exists a quasi-Sasaki-Einstein metric of positive scalar curvature on the principal circle bundle described above. 33 The proof of the theorem is mostly the same as the proof of theorem 1.4 in [41]. We use the construction due to Wang and Ziller in [41] to produce the Einstein metrics and then we show that they are in fact quasi-Sasaki. For the sake of completeness, we include the construction here. Proof. We consider a metric on M of the form g = a2η ⊗ η + pi∗h where h = a1pi ∗ 1g1 + · · ·+ akpi∗kgk for constants a 6= 0 and ai > 0. The choice of a is inconsequential. We will show that we can choose the ai so that g is Einstein, but first let us verify that this indeed defines a quasi-Sasaki metric. The 1-form ηa = aη determines the contact bundle D = ker(ηa) = ker(η). Since it is the horizontal distribution of the connection, it is invariant under the S1 action and TM = D ⊕ Lξ where Lξ = ker(pi∗) is the vertical distribution and ξ is a vector field on M with η(ξ) = 1. Let ξa = a −1ξ. Let J be the complex structure on B. Define an endomorphism Φ : TM → TM by Φ := p˜i ◦ J ◦ pi∗, where p˜i denotes the unique horizontal (with respect to η) lift of a vector field on B. Then Φξ = 0 and for horizontal vector fields X ∈ D, we have Φ2X = −X. Therefore Φ2 = −I + ξa ⊗ ηa. As h is invariant under J , it is easy to show that pi∗h(ΦX,ΦY ) = pi∗h(X, Y ). Hence g(ΦX,ΦY ) = g(X, Y )− ηa(X)ηa(Y ). Thus (ξa, ηa,Φ, g) is an almost contact metric structure on M . Since the base manifold B is complex, a computation shows that the structure is normal. The fundamental 2-form ω := g(·,Φ·) = pi∗h(·,Φ·) = pi∗(a1pi∗1ω1 + · · ·+ akpi∗kωk), where ωi is the Ka¨hler 34 form of gi. Since each ωi is closed, it follows that dω = 0. Therefore, (ξa, ηa,Φ, g) is quasi-Sasaki. Note that it is Sasaki precisely when ai = abi > 0 for all i = 1, . . . , k. If the metric g is to satisfy Ricg = λg, then by proposition 2.1.2, we must have λ = |dη|2g = 2a2 k∑ i=1 ni ( bi ai )2 > 0. (2.12) Since the Ka¨hler forms ωi are harmonic, we have that dη is harmonic. Thus equation (2.10) is satisfied. The contact bundle has a natural splitting D = D1 ⊕ · · · ⊕ Dk where pi∗Di = TBi. On Di we have dη = bipi∗i ω1. So to satisfy equation (2.11), for i = 1, . . . , k, we need qi ai = λ+ 2a2 ( bi ai )2 . (2.13) Combining equations (2.12) and (2.13), for each j = 1, . . . , k, we have ( k∑ i=1 ni ( bi ai )2)( qj aj − λ ) = λ ( bj aj )2 . (2.14) Introducing the new variables sj := qj λaj , equation (2.14) becomes ( k∑ i=1 ni ( bi qi )2 s2i ) (sj − 1) = ( bj qj )2 s2j . (2.15) Since not all bj = 0, this shows that sj ≥ 1 with equality if and only if bj = 0. We can rearrange (2.15) to get ( 1 + 1 nj − sj ) k∑ i=1 ni ( bi qi )2 s2i = 1 nj ∑ i 6=j ni ( bi qi )2 s2i . (2.16) 35 This shows that sj ≤ 1 + 1nj with equality if and only if bi = 0 for all i 6= j. Let us assume that all bj 6= 0; we will comment on the general case later. Then 1 < sj < 1 + 1 nj ≤ 2. If we multiply equation (2.15) by nj and then sum on j, we get k∑ j=1 nj(sj − 1) = 1. (2.17) Thus, solving the system (2.15) is equivalent to solving ( bj qj )2 s2j sj − 1 = c, (2.18) for each j = 1, . . . , k and some constant c > 0, subject to the constraint (2.17). On the interval (1, 2], the function f(s) = s 2 s−1 decreases monotonically with f(s) → ∞ as s → 1+ and f(2) = 4. So we start by choosing c large enough that c > 4 ( bj qj )2 for all j. Then there is a unique solution to (2.18) with 1 < sj < 2. If c is extremely large, then all sj must be very close to 1. But then we will have∑k j=1 nj(sj − 1) < 1. As c decreases, the sj increase monotonically. So we decrease c until the largest sj = 2. Then we will have ∑k j=1 nj(sj − 1) > 1 since nj ≥ 1. Thus there is a unique value of c for which (2.18) has a solution with 1 < sj < 2 and∑k j=1 nj(sj − 1) = 1. With the values of the sj in hand, we return to equation (2.12) and solve for λ. Once we know sj and λ, we have aj and we are done. Remark 2.2.7. If one of the bi, say b1 = 0, then on D1 we have dη = 0 and the bundle over B1 is trivial so M splits as M ′×B1. Here pi′ : M ′ → B2× · · · ×Bk is a principal S1 bundle with connection form η′ such that dη′ = pi′∗Ω. See section 2.5. We should also note that if any of the bi = 0, then the quasi-Sasaki structure has rank < 2n+ 1. 36 Example 2.2.8. Suppose the base B = Bn1 consists of a single Ka¨hler-Einstein manifold and consider the principal circle bundle pi : M → B with dη = bpi∗ω1, b 6= 0. By theorem 2.2.6, g = η ⊗ η + 2b 2(n+ 1) q pi∗g1 is quasi-Sasaki-Einstein with Ricg = nq2 2b2(n+ 1)2 g. The quasi-Sasaki structure has full rank 2n + 1. If we set a = q 2b(n+1) and ηa = aη, then a2g = ηa ⊗ ηa + abpi∗1g. Since dηa = abpi∗ω1, we see that a2g is Sasaki-Einstein. Example 2.2.9. Recall that CP n with the Fubini-Study metric ωFS is Ka¨hler- Einstein with ρ = (n + 1)ωFS. It is well known that H 2(CP n,Z) ' Z and that α = 1 2pi [ωFS] is a generator. Thus c 1(CP n) = (n + 1)α and every principal circle bundle pi : P → CP n has characteristic class 1 2pi [Ω] = bα for some integer b. A classic example is the Hopf fibration pi : S2n+1 → CP n. Since the total space S2n+1 is simply connected, by theorem 2.2.2, b = ±1. We choose an orientation so that b = 1. Let Bi = CP ni , i = 1, . . . , k, and B = B1 × · · · × Bk. Then H2(B,Z) ' ⊕ki=1Z with generators αi ∈ H2(Bi,Z). Thus every principal circle bundle pi : P → B has characteristic class 1 2pi [Ω] = ∑k i=1 biαi for some integers bi. The total space P is simply connected if and only if the characteristic class is indivisible if and only if gcd(b1, . . . , bk) = 1. For a concrete example, take ni = 1 for i = 1, 2, 3, b1 = 1, b2 = −1 and b3 = 0. Then qi = 2 and it is not too hard to solve the equations from theorem 2.2.6. Working through them, we find that a1 = a2 = 3 and a3 = 9 2 . Thus the fundamental 2-form 37 ω = pi∗(3pi∗1ωFS + 3pi ∗ 2ωFS + 9 2 pi∗3ωFS) and the metric g = η ⊗ η + ω(Φ·, ·) is quasi- Sasaki-Einstein with Ricg = 4 9 g. This example illustrates some interesting features of quasi-Sasaki-Einstein metrics. First, the rank of this quasi-Sasaki structure is 5, whereas full rank is 7. Thus this quasi-Sasaki structure is not a deformation of a Sasaki structure. However, since b3 = 0, the total space P splits as P ′×B3 where pi′ : P ′ → B1×B2 is a principal circle bundle with connection form η′ such that dη′ = pi′∗Ω. Thus g′ = η⊗η+ω′(Φ·, ·), where ω′ = pi′∗(3pi∗1ωFS + 3pi ∗ 2ωFS), is a quasi-Sasaki-Einstein metric on P ′ with Ric′g = 4 9 g′. This quasi-Sasaki structure does indeed have full rank, but still it cannot be a deformation of a Sasaki structure because dη′(Φ·, ·) is not positive definite; it is negative definite over B2. Now we will present the main theorem of this section. It is a consequence of theorem 1 in [40]. Theorem 2.2.10. If (M, ξ, η,Φ, g) is a regular, quasi-Sasaki-Einstein manifold with positive scalar curvature, then M is a principal circle bundle over a compact Ka¨hler manifold B, where B is a product of Ka¨hler-Einstein manifolds and the metric g is among those constructed in theorem 2.2.6. Proof. Since g is Einstein with positive scalar curvature, M is compact. Then the regularity of Fξ implies that the orbits of ξ are compact, so by theorem 1.3.6 M is a principal circle bundle over a compact Ka¨hler manifold B with η as the connection form. By lemma 1.2.7, dη(Φ·,Φ·) = dη. Thus the curvature form of the circle bundle is of type (1,1). Now by theorem 1 in [40], B is a product of Ka¨hler-Einstein manifolds and (M, g) is among those constructed in theorem 2.2.6. 38 Theorems 2.2.5 and 2.2.10 give us a clear picture of the possible quasi-Sasaki- Einstein metrics in the regular case. As we have seen, they are very rigid and do not offer us “new” Einstein metrics that we didn’t already know about. 2.3. The quasi-Regular Case In this section we assume, more generally, that the orbits of ξ are compact and the induced S1 action is locally free. Then Fξ is quasi-regular. By theorem 1.3.6, M is the total space of a principal S1-orbibundle over the Ka¨hler orbifold B := M/Fξ with connection form η and curvature form dη = pi∗Ω. Here Ω is a closed (1,1)-form with 1 2pi [Ω] ∈ H2orb(B,Z). The treatment here closely follows that of the previous section. We will not dive too deeply into the general theory of orbifolds and orbibundles, but only reference pertinent information. For definitions and fundamental results, see [5] and [26] and the references therein. Many facts about principal circle bundles have an orbibundle counterpart. For instance, theorem 4.3.15 of [5] says that the (isomorphism classes of) principal S1- orbibundles over an orbifold B are in one-to-one correspondence with the elements of H2orb(B,Z) and the bijection is given by the orbifold first Chern class. In particular, the trivial orbibundle corresponds to the trivial orbifold cohomology class. Also, it is not hard to produce orbibundle versions of lemma 2.2.3 and theorem 2.2.4. The same argument as in the previous section gives us the orbifold version of theorem 2.2.5. Theorem 2.3.1. If (M, g) is a quasi-Sasaki manifold with Ricg = 0 and Fξ has compact leaves, then M = S1 × |B| where |B| is the underlying topological space of a Calabi-Yau orbifold B. Remark 2.3.2. Since M is a manifold, |B| must be a manifold too. 39 The orbifolds given by theorem 1.3.6 (i.e. those coming from the leaf space of a quasi-regular quasi-Sasaki structure) are locally cyclic, meaning that all of the local uniformizing groups are cyclic groups. Furthermore, given an orbifold B, theorem 4.3.15 of [5] also states that the total space of an S1-orbibundle over B is a smooth manifold when all of the local uniformizing groups of B inject into S1. For these reasons, we will only concern ourselves with locally cyclic orbifolds. As a sort of converse to theorem 1.3.6, we have the following: Proposition 2.3.3. Given a locally cyclic Ka¨hler orbifold B with 1 2pi [Ω] ∈ H2orb(B,Z), there is a quasi-Sasaki structure on the principal S1-orbibundle pi : M → B corresponding to 1 2pi [Ω] with connection form η such that dη = pi∗Ω. Proof. Let ξ be the unit vector field along the fibers and consider the metric on M given by g = η ⊗ η + pi∗gB where gB is a Ka¨hler metric on B. Define Φ = p˜i ◦ J ◦ pi∗ where J is the complex structure on B and p˜i is the horizontal lift with respect to η. Then Φ2 = −I + ξ ⊗ η and Φ∗g = g − η ⊗ η. It follows that M is quasi-Sasaki. We also have the orbi-analogue of theorem 2.2.6. Since the equations we must solve are the same, the proof follows mutatis mutandis, and so we omit it. Theorem 2.3.4. Let (Bnii , ωi), i = 1, . . . , k, be locally cyclic Ka¨hler-Einstein orbifolds with c1orb(Bi) > 0. We can assume that the metrics have been normalized so that 1 2pi [ωi] is an indivisible, integral class and c 1 orb(Bi) = qi 1 2pi [ωi] for some qi ∈ Z+. Let B = B1 × · · · × Bk and pii : B → Bi the projections. Pick integers bi, not all zero, and set Ω = b1pi ∗ 1ω1 + · · ·+ bkpi∗kωk. 40 Then there is a quasi-Sasaki-Einstein metric with positive scalar curvature on the principal S1-orbibundle pi : M → B corresponding to 1 2pi [Ω] with connection form η such that dη = pi∗Ω. We conclude this section with the orbi-analogue of the rigidity theorem 2.2.10. We will show that theorem 1 in [40] can be carried over to the orbifold setting and, as in the previous section, it implies our theorem. Theorem 2.3.5. If (M, g) is a quasi-regular, quasi-Sasaki-Einstein manifold with positive scalar curvature, then M is a principal circle orbibundle over a compact Ka¨hler orbifold B, where B is a product of Ka¨hler-Einstein orbifolds and the metric g is among those constructed in theorem 2.3.4. Proof. Since (M, g) is compact, quasi-regularity of Fξ implies that the orbits of ξ are compact. By the structure theorem 1.3.6, M is a principal S1-orbibundle over a compact Ka¨hler orbifold (B, gT ) with connection form η. By lemma 1.2.7, the curvature form dη is of type (1,1). The conditions on dη and ρT imposed by the Einstein equations are the same as the conditions on the curvature form and Ricci form in [40]. Proposition 1 in [40] is proved by local computations, exploiting these conditions. Thus we would have the same proposition in orbifold setting. Proposition 2.3.6. Under the assumptions of the theorem, there is a global, orthogonal decomposition TB = E1 ⊕ · · · ⊕ Er where the Ei are the eigenspaces of the Ricci curvature. The eigenvalues are constant and any sum of the eigenspaces is integrable and invariant under the complex structure. The leaves of the foliation TB = E1⊕· · ·⊕Er have an induced Ka¨hler structure and the Einstein equations for g show that they have positive Ricci curvature. Thus the leaves are compact and simply connected by Myers’s theorem and Kobayashi’s 41 theorem in orbifold setting, respectively. By the same argument as in step 3 of proposition 2 of [40] (using the local stability theorem of foliations with compact leaves), we can find a local frame for E1 consisting of vector fields which are foliate with respect to E⊥1 . Then we can perform the same local computation as in [40] to show that both E1 and E ⊥ 1 are totally geodesic. Then it follows that E1 and E ⊥ 1 are parallel. Iterating this argument we find that each Ei is parallel. Then, since B is compact (hence complete) and simply connected, by the de Rham decomposition theorem for orbifolds (see [26]) B splits isometrically as a product of Ka¨hler orbifolds B = B1 × · · · × Br according to the decomposition of TB. Hence B is a product of Ka¨hler-Einstein orbifolds with c1orb(Bi) > 0. Then since the curvature form is of type (1,1), it follows that the characteristic class of the S1-orbibundle must be of the form in theorem 2.3.4. 2.4. Transverse Curvature Conditions In the previous two sections we assumed that all of the orbits of ξ were compact. We want to make no assumptions on the orbits of the Reeb field, but the irregular case, where M is compact and ξ has noncompact orbits, is inherently more difficult to deal with. The difficulty comes from the potential unruliness of the leaf space M/Fξ. It may not have a tractable structure; for instance, it may not even be a Hausdorff space! Since we are unable to work with the irregular case directly, to get a handle on it we will assume that the transverse curvature satisfies a certain non- negativity condition. In this section, we introduce the curvature conditions that we will work with and mention a few of their topological implications. We first consider a curvature condition known as non-negative quadratic orthogonal bisectional curvature (NQOBC). 42 Definition 2.4.1. A Ka¨hler manifold of complex dimension n has NQOBC at a point x ∈M if for any unitary frame {e1, . . . , en} of T 1,0x M and any real numbers α1, . . . , αn, n∑ i,j=1 Ri¯ijj¯(αi − αj)2 ≥ 0. We say that the manifold has NQOBC if it does so at every point x ∈M . For more information about Ka¨hler manifolds with NQOBC, see the paper [8]. The primary reason why we are interested in this curvature condition is for the following theorem: Theorem 2.4.2. If a closed (compact with no boundary) Ka¨hler manifold M has NQOBC, then every harmonic (1,1)-form is parallel. Proof. The Bochner formula for (1,1)-forms on a Ka¨hler manifold is 4αij¯ = −∆αij¯ − 2Rij¯kl¯αlk¯ +Rik¯αkj¯ +Rkj¯αik¯. Here 4 = dδ + δd is the Hodge laplacian and ∆ = tr(∇2) is the tensor laplacian. If 4α = 0, then the Bochner formula yields αij¯∆αij¯ = −2Rij¯kl¯αlk¯αij¯ +Rik¯αkj¯αij¯ +Rkj¯αik¯αij¯. (2.19) Since α is a (1,1)-form, A := α(J ·, ·) is a symmetric bilinear form that is invariant under the complex structure J . Hence A defines a Hermitian form and thus there is a local unitary frame {ei, ei} that diagonalizes A. So with respect to this frame, 43 αi¯i := α(ei, ei) are the only nonzero components of α and equation (2.19) becomes αi¯i∆αi¯i = −2Ri¯ikk¯αkk¯αi¯i +Ri¯iα2i¯i +Rkk¯α2kk¯ = Ri¯ikk¯(αi¯i − αkk¯)2. (2.20) Integrating equation (2.20) and using integration by parts, we get ∫ M −|∇α|2 dV = ∫ M Ri¯ikk¯(αi¯i − αkk¯)2 dV. From this, the assumption of NQOBC implies that ∇α = 0. If we assume that the transverse Ka¨hler metric of a quasi-Sasaki structure has non-negative quadratic orthogonal transverse bisectional curvature, then the following corollary is immediate. Corollary 2.4.3. If the transverse Ka¨hler metric of a closed quasi-Sasaki manifold has NQOBC, then every basic, harmonic (1,1)-form is transversely parallel. That is, if α is a basic (1,1)-form and 4Bα = 0, then ∇Tα = 0. The other curvature condition that we will consider is transverse holomorphic bisectional curvature (THBC). This is just the holomorphic bisectional curvature of the transverse Ka¨hler metric. Definition 2.4.4. Let σ be a plane in Dp ⊂ TpM . We say that σ is Φ-invariant if Φσ = σ. It follows that σ = span{Xp,ΦXp} for some 0 6= Xp ∈ Dp. Definition 2.4.5. Let σ1 and σ2 be Φ-invariant planes in Dp ⊂ TpM . The transverse holomorphic bisectional curvature (THBC) HT (σ1, σ2) is defined by HT (σ1, σ2) := 〈RT (X,ΦX)ΦY, Y 〉, 44 where X ∈ σ1, Y ∈ σ2 and |X| = |Y | = 1. It is a straightforward exercise to check that HT (σ1, σ2) is well-defined. That is, HT (σ1, σ2) depends only on σ1 and σ2, not on the choice of X and Y . We say that HT ≥ 0 at a point p ∈ M if HT (σ1, σ2) ≥ 0 for any two Φ-invariant planes σ1, σ2 ⊂ Dp. We say that M has non-negative transverse holomorphic bisectional curvature if HT ≥ 0 for all p ∈M . Let σu = span{u,Φu} and σv = span{v,Φv} for some unit vectors u, v ∈ D. Writing u = U+U¯ and v = V +V¯ where U = 1 2 (u−√−1Φu) and V = 1 2 (v−√−1Φv), we have HT (σu, σv) = 4〈RT (U, U¯)V, V¯ 〉. So HT ≥ 0 if and only if 〈RT (U, U¯)V, V¯ 〉 ≥ 0 for all U, V ∈ D1,0. In the Ka¨hler setting, positive and non-negative holomorphic bisectional curvature has been studied extensively. For example, see any of [2], [18], [22], [29], [35] or [42]. Positive and non-negative THBC has been studied in the Sasaki setting in [20] and [21]. The following is a corollary of theorem 4 in [18]. Since the proof would follow in exactly the same way, we shall omit it. Corollary 2.4.6. If M is a compact, connected quasi-Sasaki manifold with positive transverse bisectional curvature, then the second basic Betti number b2B = 1. Note that NQOBC is a strictly weaker condition than non-negative holomorphic bisectional curvature. Clearly NHBC implies NQOBC and indeed there are examples of Ka¨hler manifolds with NQOBC that do not admit any Ka¨hler metrics with NHBC. 2.5. Splitting the Contact Bundle In this section we show that the presence of transversely parallel (1,1)-forms induces a splitting of the contact bundle of a quasi-Sasaki manifold. We begin with a definition. 45 Definition 2.5.1. A subbundle D1 ⊂ D is called invariant or transversely parallel if for all Y ∈ D1, ∇TXY ∈ D1 for any X ∈ TM . The contact bundle D is reducible (with respect to gT ) if there are invariant subbundles D1 and D2 and an orthogonal decomposition D = D1 ⊕ D2. We say that gT is (irreducible) reducible if D is (not) reducible with respect to gT . Let (M, g) be a quasi-Sasaki manifold. The fundamental 2-form ω (i.e. the Ka¨hler form of the transverse metric gT ) is transversely parallel (i.e. ∇Tω = 0). Suppose that α is another basic, transversely parallel (1,1)-form. Then for X, Y ∈ D, define the (1,1)-tensor S by the equation α(X, Y ) = ω(S(X), Y ). (2.21) Notice that S is self-adjoint with respect to ω. Indeed, ω(S(X), Y ) = α(X, Y ) = −α(Y,X) = −ω(S(Y ), X) = ω(X,S(Y )). Hence S is a diagonalizable linear map on each contact space Dp ⊂ TpM . Since both α and ω are transversely parallel, it follows from (2.21) that S is transversely parallel. Therefore, the linear map Sp : Dp → Dp has the same set of eigenvalues for every point p ∈ M . Let {c0, . . . , cr} be the distinct eigenvalues of S. As S is self-adjoint, the eigenvalues ci ∈ R. If α is not a multiple of ω, then S is not a multiple of the identity and thus r ≥ 1. Let Di ⊂ D be the eigenbundle corresponding to the eigenvalue ci of S. Then D = D0 ⊕ · · · ⊕ Dr (2.22) 46 and the distribution p 7→ Di(p) is transversely parallel. Indeed, let Y ∈ Di(p) and X ∈ TpM . Since ∇TS = 0, we have S(∇TXY ) = ∇TXS(Y ) = ci∇TXY. Therefore ∇TXY ∈ Di(p) and Di is invariant in the sense of definition 2.5.1. The splitting (2.22) is orthogonal with respect to ω. Indeed, if X ∈ Di and Y ∈ Dj and i 6= j, then ciω(X, Y ) = ω(S(X), Y ) = ω(X,S(Y )) = cjω(X, Y ). Since ci 6= cj, this shows that ω(X, Y ) = 0. Since ci is real, Di is closed under complex conjugation (i.e. Di = Di). Since both α and ω are (1,1)-forms, they are Φ-invariant. As S is self-adjoint, we get that ω(X,S(ΦY )) = ω(S(X),ΦY ) = α(X,ΦY ) = −α(ΦX, Y ) = −ω(S(ΦX), Y ) = −ω(ΦX,S(Y )) = ω(X,ΦS(Y )). Since ω is non-degenerate on D, this implies that S ◦Φ = Φ◦S. Hence, Di is invariant under Φ (i.e. ΦDi = Di). With this we see that the splitting of D in (2.22) is orthogonal with respect to gT and therefore, if α and ω are not proportional to each other, D is reducible. Corresponding to this splitting, we write ω = ω0 ⊕ · · · ⊕ ωr, α = α0 ⊕ · · · ⊕ αr and Φ = Φ0 ⊕ · · · ⊕ Φr, where Φi(X) =  Φ(X) if X ∈ Di, 0 otherwise . On Di we have αi = ciωi. Since Di has a Ka¨her structure, Di has even real dimension. Therefore, Di has a local frame {X1, . . . , Xpi , X¯1, . . . , X¯pi} where Xj ∈ D1,0. 47 Lemma 2.5.2. Let M be a compact quasi-Sasaki manifold with transverse NQOBC. Then D is reducible if and only if b1,1B > 1. Proof. If b1,1B > 1 then by the transverse Hodge theorem, we have at least two distinct, basic, harmonic (1,1)-forms. With the assumption of transverse NQOBC, these harmonic forms are transversely parallel (corollary 2.4.3) and as we just saw above, this implies that D is reducible. Conversely, suppose D = D1 ⊕ D2 and the Di are invariant with respect to gT . Let ωi = ω|Di . Then ω = ω1 ⊕ ω2 and 0 6= [ωi]B ∈ H1,1B (M,Fξ). Furthermore, since ωn = (ω1 ⊕ ω2)n 6= 0, [ω1]B and [ω2]B are independent. Thus b1,1B > 1. 2.6. The General Case with a Curvature Condition In this section we assume that the quasi-Sasaki manifold (M, g) is Einstein with Ricg = λg. Then dη is harmonic and, as we saw in remark 2.1.3, the transverse scalar curvature RT = λ(n + 1) is constant. Proposition 2.1.4 then implies that the transverse Ricci form ρT is harmonic. We also assume further that (M, g) satisfies a transverse curvature condition at least as strong as NQOBC. Then by corollary 2.4.3, dη and ρT are transversely parallel. Thus we can split the contact bundle as in the previous section. Let’s split the contact bundle with respect to dη. Let c0, . . . , cr, be the distinct eigenvalues of S (the (1,1)-tensor form of dη). Then we have an orthogonal splitting of D into transversely parallel distributions, D = D0⊕ · · · ⊕Dr. On each Di we have dηi = ciωi. If we write ρ T = ⊕ρTi according to this decomposition, then equation (2.11) shows that ρTi = (λ + 2c 2 i )ωi ≥ 0. This implies that the splitting of D with respect to ρT is the same as the splitting of D with respect to dη. 48 If our quasi-Sasaki structure has rank 2p + 1 for p < n, then ci = 0 for some i. Reindexing if neccessary, we may assume that c0 = 0 and hence dη0 = 0. That Di is transversely parallel implies that for any X, Y ∈ Di we have piD([X, Y ]) ∈ Di. Since piLξ([X, Y ]) = −2dη(X, Y )ξ, we see that D0 and Lξ ⊕ Di are involutive, hence integrable. Again using that the Di are transversely parallel, it is not hard to see that Lξ ⊕ D1 ⊕ · · · ⊕ Dr is also integrable and thus M is locally a product manifold. In fact, (M, g) is locally a Riemannian product, as we will see next. Since gT and dη are both transversely parallel and we have the relations cig T (X,ΦiY ) = dηi(X, Y ), it follows that ∇TΦi = 0. We will show that Φ0, in particular, is parallel with respect to the metric g. That is, ∇Φ0 = 0. Note that this need not be the case for the other Φi. Remember that by lemma 1.2.13, ∇Φ = 0 if and only if M is cosymplectic. First we prove the following useful proposition. Proposition 2.6.1. If D = D0 ⊕ · · · ⊕ Dr admits a splitting such that dηi = ciωi, then for Y ∈ D, ∇Y ξ = − ∑ i ciΦiY . In particular, if Y ∈ Di then ∇Y ξ = −ciΦY . Proof. By lemma 1.2.11 we have dη(X, Y ) = −g(X,∇Y ξ). Since dη = ∑ i dηi and dηi(X, Y ) = cig(X,ΦiY ), the above equation gives ∑ i g(X, ciΦiY ) = −g(X,∇Y ξ). Since X was arbitrary, the result follows. Now we will show that ∇Φ0 = 0. Recall that the transverse Levi-Civita connection ∇T is defined by ∇TXY := piD (∇XY ) , X ∈ D, ∇Tξ Y := piD [ξ, Y ] . 49 Lemma 2.6.2. The (1,1)-tensor Φ0 is parallel with respect to the metric g. Proof. As Φ : TM → D and Φξ = 0, for X, Y ∈ D we have 0 = (∇TXΦ0)(Y ) = piD ((∇XΦ0)(Y )) = piD(∇XΦ0(Y )− Φ0(∇XY )). This says that (∇XΦ0)(Y ) has no part in D. Does it have any part in Lξ? Well, Φ0 : D0 → D0 and the term ∇XΦ0(Y ) ∈ D because the ξ part is dη(X,Φ0(Y ))ξ = 0 since dη = 0 on D0. Thus, for X, Y ∈ D, we have (∇XΦ0)(Y ) = 0. By proposition 2.6.1 and Φi ◦ Φj = 0 for i 6= j we compute that (∇XΦ0)(ξ) = ∇XΦ0(ξ)− Φ0(∇Xξ) = ∑ i ciΦ0(ΦiX) = 0. Thus ∇XΦ0 = 0 for X ∈ D. By lemma 1.2.11, ∇ξΦ = 0 which implies that ∇ξΦ0 = 0 (we could also show this by direct computation). Therefore ∇Φ0 = 0. Now let Φ˜ := ∑r i=1 Φi and D˜ := ∑r i=1Di. The map P := −Φ˜2 + ξ ⊗ η is projection onto Lξ ⊕ D˜ and Q := −Φ20 is projection onto D0. Thus P 2 = P , Q2 = Q and PQ = QP = 0. As in section ??, F := P − Q defines an almost product structure. Since Φ = Φ0 + Φ˜ and Φ 2 = −I + ξ ⊗ η, we have F = I + 2Φ20. By proposition 2.6.2, ∇Φ0 = 0 and this clearly implies that ∇Φ20 = 0. Hence ∇F = 0. So by proposition 5.0.7, (M, g) is a locally decomposable Riemannian manifold. Thus, locally, (M, g) is the Riemannian product of manifolds (M0, g0) and (M1, g1) whose tangent spaces are (isomorphic to) D0 and Lξ ⊕ D˜ respectively. We will show that (M0, g0) is Ka¨hler-Einstein and (M1, g1) is quasi-Sasaki-Einstein. 50 Pick a point x ∈ M and let M0 be a maximal integral submanifold of D0 containing the point x. Then dimR(M0) = 2(n − p). Define 2-forms ω0 and ω˜ by ω0(X, Y ) = g(X,Φ0Y ) and ω˜(X, Y ) = g(X, Φ˜Y ). Then ω0 = ω− ω˜. Since ω is closed and ω˜ = ∑r i=1 c −1 i dηi is proportional to dη, we have dω0 = 0. Because Φ0 has rank 2(n − p), ω0|D0 is non-degenerate. We have (Φ0|D0)2 = −I, ∇Φ0 = 0 and g|D0 is compatible with Φ0. Therefore, (M0,Φ0|D0 , g|D0) is Ka¨hler. To see that the metric on M0 is Einstein we refer back to equations (2.8) and (2.11). Since dη = 0 on D0, these equations say precisely that g|D0 is an Einstein metric on M0 Let M1 be a maximal integral submanifold of Lξ ⊕ D˜ containing x. Then dimR(M1) = 2p + 1. Since η ∧ (dη)p 6= 0 on Lξ ⊕ D˜, restricting η to Lξ ⊕ D˜ gives M1 a contact structure. As Φ and Φ˜ agree on D˜ and both are zero on Lξ, it is easily verified that the restriction of (ξ, η, Φ˜, g) to Lξ⊕D˜ gives M1 an almost contact metric structure. Now we check the normality condition. Recall that NΦ + ξ ⊗ 2dη = 0 and since ∇Φ0 = 0, we have NΦ0 = 0. Take X, Y ∈ Lξ ⊕ D˜. Since Φ˜ = Φ− Φ0 and Φ0X = Φ0Y = 0, we compute that NΦ˜(X, Y ) + 2dη(X, Y )ξ = Φ0[X,ΦY ] + Φ0[ΦX, Y ]−ΦΦ0[X, Y ]−Φ0Φ[X, Y ]. (2.23) As the distribution Lξ ⊕ D˜ is integrable, it is involutive. Hence, each of the bracket terms in (2.23) are in Lξ ⊕ D˜. Since Φ0 annihilates Lξ ⊕ D˜, it follows that NΦ˜(X, Y ) + 2dη(X, Y )ξ = 0. Therefore, M1 has a normal almost contact metric structure. Clearly, the fundamental 2-form ω˜ associated to this almost contact metric structure is closed. Hence (ξ, η, Φ˜, g) restricted to Lξ ⊕ D˜ gives M1 a full rank quasi- Sasaki structure. Moreover, the metric is Einstein. Summarizing, we have proved: 51 Theorem 2.6.3. Let M2n+1 be a quasi-Sasaki-Einstein manifold of rank 2p + 1 for p < n with transverse NQOBC. Then (M, g) is locally the Riemannian product of a Ka¨hler-Einstein manifold (M0, g0) with dimR(M0) = 2(n − p) and quasi-Sasaki- Einstein manifold (M1, g1) of rank 2p+ 1. Moreover, the transverse Ka¨hler metric of g1 is a product of transverse Ka¨hler-Einstein metrics. Remark 2.6.4. That M is locally a product manifold does not require the transverse curvature assumption. The Einstein equations alone are enough to get this and the proof can go along the same lines as the proof of proposition 1 in [40]. However, to get a local Riemannian product, the almost product structure must be parallel (proposition 5.0.7). At this time, we don’t know if this can be proved solely on the strength of the Einstein equations in the irregular case. For the remainder of this section we may assume that our quasi-Sasaki structures have full rank and if D is reducible, then none of the Di are integrable on their own. If D is irreducible, then the transversely parallel forms ω, dη and ρT must all be proportional to each other. That is, dη = cω and ρT = µω for some constants c and µ. Since c 6= 0, our quasi-Sasaki structure is simply a scaling of a Sasaki-Einstein structure. By replacing η with cη, replacing ξ with c−1ξ and scaling the transverse metric by c2 we get a Sasaki-Einstein structure. Combining corollaries 2.4.3, 2.4.6 and lemma 2.5.2 we get: Theorem 2.6.5. A quasi-Sasaki-Einstein manifold with positive transverse holomorphic bisectional curvature is, up to homothey, a Sasaki-Einstein manifold. Now by [21] we get that 52 Corollary 2.6.6. A quasi-Sasaki-Einstein manifold with positive transverse holomorphic bisectional curvature is diffeomorphic to S2n+1 and the metric can be deformed to the round metric. Lemma 2.6.7. A reducible quasi-Sasaki manifold with positive Ricci curvature is quasi-regular. Proof. Suppose that M is quasi-Sasaki with positive Ricci curvature and D = D1⊕D2 is reducible. Then the Di are transversely parallel and this implies that Lξ ⊕ Di is an integrable subbundle of TM . Since the Di are Φ-invariant, using proposition 2.6.1 and equation (2.4), it is not hard to show that the leaves of Lξ ⊕ Di are totally geodesic submanifolds of M . Hence the induced metric on them is complete. By assumption, D1 and D2 have positive Ricci curvature. So by Myers’s theorem, the leaves of Lξ ⊕ Di are compact. An integral curve of ξ is a transversal intersection of a leaf from Lξ ⊕ D1 and a leaf from Lξ ⊕ D2. Thus the ξ orbits are compact and therefore Fξ is quasi-regular. From this lemma we see that an irregular quasi-Sasaki-Einstein manifold with positive scalar curvature is irreducible. Hence we have the following corollary: Corollary 2.6.8. An irregular quasi-Sasaki-Einstein manifold with positive scalar curvature and transverse NQOBC is a scaling of a Sasaki-Einstein manifold. 53 CHAPTER III TRANSVERSE KA¨HLER-RICCI FLOW The aim of this chapter is to extend results in Ka¨hler-Ricci flow to quasi-Sasaki manifolds. The transverse Ka¨hler-Ricci flow is the normalized Ka¨hler-Ricci flow of the quasi-Sasaki manifold’s transverse Ka¨hler metric. We give a proof of its short-time existence. The long-time existence follows from the work in [36]. In comparison with [7], we can show that when the basic first Chern class c1B ≤ 0, the flow converges to a transverse Ka¨hler-Einstein metric. This can be done just as in [36], but we choose not to pursue it here. Rather, we are interested in the case that c1B > 0. In [31], Perelman introduced hisW functional. TheW functional is monotone increasing along the Ricci flow and this property has many important applications. In particular, for a Ka¨hler manifold with positive first Chern class, the monotonicity of the W functional can be used to prove uniform bounds on the scalar curvature and the diameter of the manifold along the flow. This was first presented in [34]. We have two primary goals in this chapter. The first is to establish uniform bounds on the transverse scalar curvature and the transverse diameter along the transverse Ka¨hler-Ricci flow when c1B > 0. This will imply uniform bounds on the scalar curvature and the diameter of the manifold along the flow. We first define the WT functional, which is the analogue of the W functional for the quasi-Sasaki setting. We show that it is monotonic along the flow and then we proceed in a way similar to [34]. A subtle difference in our situation is that the WT functional is only involved with basic functions. That is, functions which are constant in the ξ direction. This poses some problems as the distance function on the manifold is not basic. To deal with this, we utilize the transverse distance function from chapter I. 54 This approach was taken in [12] and [19], where these results were, independently, established for Sasaki manifolds. The essential ingredient is really the transverse Ka¨hler structure of the manifold. The additional contact structure afforded by Sasaki manifolds is not required. Thus we can prove these results in the quasi-Sasaki setting as well. On a compact Ka¨hler manifold, it is known that if the initial metric has nonnegative bisectional curvature, then this property is preserved along the Ka¨hler- Ricci flow. The second goal of this chapter is to show that nonnegativity of the transverse (holomorphic) bisectional curvature, as defined in 2.4.5, is preserved along the transverse Ka¨hler-Ricci flow. This chapter is organized as follows: First we define the transverse Ka¨hler- Ricci flow and prove its short-time existence. Next we give an evolving quasi- Sasaki structure, compatible with the evolving transverse metric, and provide some commentary on how the rank of this structure may change along the flow. Next we define the WT and µT functionals and prove their monotonicity. After that we show that the transverse scalar curvature and Ricci potential are bounded along the flow by the diameter of the manifold. Then we prove a uniform upper bound on the transverse diameter of the manifold along the flow. This implies a uniform bound on the diameter of the manifold, and hence on the scalar curvature and Ricci potential. The presentation in sections 3 and 4 is much like that in [34], but here we provide more details for the arguments and computations. The last section is devoted to studying the transverse Ka¨hler-Ricci flow when the initial metric has non-negative (positive) transverse bisectional curvature. Some of the technical lemmas that get used in this chapter are collected in the appendix. 55 3.1. Transverse Ka¨hler-Ricci Flow Following the introduction of Hamilton’s Ricci flow, similar evolution equations were considered on Ka¨hler manifolds. Since the 1980’s, Ka¨hler-Ricci flow (KRF) has been a very fruitful area of mathematics. Many deep and beautiful results are known and KRF is still an active area of research today. It is well-known in Ka¨hler geometry via the Chern-Weil theory, that the first Chern class, c1, of a Ka¨hler manifold is represented by a multiple of the Ricci form and that the de Rham cohomology class of the Ricci form is independent of the Ka¨hler metric. Thus a necessary condition for the existence of a Ka¨hler-Einstein metric ω is that c1 = λ[ω]. Hence c1 must be positive definite, negative definite or null depending on whether λ is positive, negative or zero respectively. The existence of Ka¨hler-Einstein metrics in the cases c1 < 0 and c1 = 0 was proved independently by Aubin in [1] and Yau in [43]. In [7], Cao showed that the normalized KRF converges to a Ka¨hler-Einstein metric when c1 ≤ 0. When c1 > 0, things are more complicated and there are obstructions to the existence of Ka¨hler-Einstein metrics, see [16]. However, in certain cases, it is known that the KRF converges to a Ka¨hler-Ricci soliton when c1 > 0. See [33] and [38]. In [36], Smoczyk, Wang and Zhang defined the Sasaki-Ricci flow (SRF) as the KRF of the underlying transverse Ka¨hler metric associated with a Sasaki structure. They proved results analogous to those in [7]. The work in [36] on the SRF does not rely heavily on the contact structure of a Sasaki manifold and many of their results can be carried over to the quasi-Sasaki setting. Collins in [12] and He in [19], independently, extended a remarkable result of Perelman in KRF with c1 > 0 (see [34]) to the SRF with c1B > 0. In this chapter, we will do the same in the quasi-Sasaki setting. In this section we define the flow and then prove its short time existence. 56 Definition 3.1.1. Let (ξ, η0,Φ0, g0) be a quasi-Sasaki structure with 2pic 1 B = κ[ω0] for some constant κ. The transverse Ka¨hler-Ricci flow (TKRF) is given by ∂ ∂t gT = κgT − RicTg , gT (0) = gT0 . (3.1) Let ρT := RicT (·,Φ·) be the transverse Ricci form. Recall that ρT is independent of the transverse Ka¨hler metric and that the class [ρT ] = 2pic1B. As ω = g T (·,Φ·), equation (3.1) is equivalent to ∂ ∂t ω = κω − ρT , ω(0) = ω0. (3.2) A transverse homothety, also called a D-homothety, is a deformation of a quasi- Sasaki structure that scales the transverse metric. Given a quasi-Sasaki structure (ξ, η,Φ, g) and λ > 0, we set ξλ := λ −1ξ, ηλ := λη and gλ := λ(λ− 1)η ⊗ η + λg. Then (ξλ, ηλ,Φ, gλ) is a quasi-Sasaki structure and the transverse metric g T λ = λg T . So if 2pic1B = κ[ω], then by a D-homothety we may assume that κ = −1, 0 or 1, depending on the sign of κ. 3.11. Short-Time Existence To prove the short-time existence of a solution to equation (3.2), we reduce it to a parabolic Monge-Ampe´re equation. Having done this, the standard theory of parabolic equations guarantees the existence of a solution for at least a short time. We will prove the case c1B > 0 since we shall assume this in the sequel. The cases 57 c1B = 0 and c 1 B < 0 can be proved similarly. We also note, although we don’t prove it here, that equation (3.1) has a short-time solution even with no assumption on c1B. We assume that [ω0] = 2pic 1 B and we know that 2pic 1 B = [ρ T 0 ]. Thus, by the transverse ∂∂¯-lemma of [15], there is a basic function f such that ω0 = ρ T 0 + √−1∂∂¯f . Recall that ρT0 = − √−1∂∂¯ log(det(ω0)) Thus ω0 = − √−1∂∂¯ log(det(ω0)) + √−1∂∂¯f = −√−1∂∂¯ log(e−f det(ω0)) = −√−1∂∂¯ log(det(e−f/nω0)). We let Ω := det(e−f/nω0) so that we can write ω0 = − √−1∂∂¯ log(Ω). Next we consider the equation ∂φ ∂t = log det(ω0 + √−1∂∂¯φ) Ω + φ+ c(t), φ(0) ≡ 1, (3.3) defined for all t ≥ 0 such that ω0 + √−1∂∂¯φ > 0. Here c(t) is a time-dependent constant. If φ solves (3.3), then it is easy to check that the metric ω := ω0 + √−1∂∂¯φ solves (3.2). Conversely, given a solution to (3.2), its basic cohomology class must evolve as ∂ ∂t [ω] = [ω]− 2pic1B, [ω](0) = [ω0]. Solving this ODE, we find that [ω] = 2pic1B+e t([ω0]−2pic1B). Since we assume that [ω0] = 2pic 1 B, we have [ω(t)] = [ω0]. By the transverse ∂∂¯-lemma, there is φ ∈ C∞B (M) such that ω = ω0 + √−1∂∂¯φ. (3.4) 58 Since ω(0) = ω0, we have that ∂∂¯φ(0) = 0. Differentiating the above, we get ω − ρT = √−1∂∂¯ ∂φ ∂t . (3.5) Putting (3.4) into (3.5) and writing ρT = −√−1∂∂¯ log(det(ω)), we then have √−1∂∂¯ ∂φ ∂t = √−1∂∂¯ ( log det(ω) Ω + φ ) . (3.6) Thus there is a time-dependent constant c(t) such that ∂φ ∂t = log det(ω) Ω + φ+ c(t). Hence φ solves (3.3). Therefore, equations (3.2) and (3.3) are equivalent. Now consider the parabolic Monge-Ampe´re equation, given by ∂φ ∂t = log det(ω0ij¯ + φ,ij¯) Ω + ξ2φ+ φ. (3.7) Here φ is not assumed to be basic and the indices after the comma denote covariant derivatives. Lemma 3.1.2. Equation (3.7) preserves the property ξφ = 0. Proof. The proof of this lemma is by a maximum principal argument and is the same as the proof of Lemma 5.2 in [36]. See [36] for the details. With this we can prove the short time existence of our flow. Proposition 3.1.3. For some T > 0, equation (3.2) has a solution for t ∈ [0, T ). 59 Proof. Equation (3.7) is parabolic whenever we have det(ω0ij¯ + φ,ij¯) > 0. (3.8) By the standard parabolic theory, for any initial function φ(0) satisfying (3.8), there is T > 0 and a solution, φ : M × [0, T )→ R, to (3.7) with det(ω0ij¯ + φ,ij¯) > 0 for all t ∈ [0, T ). Take φ(0) ≡ 1. Then (3.8) is satisfied and since ξφ(0) = 0, the solution φ is a basic function by lemma 3.1.2. Recall that φ,ij¯ := ∇2φ(Xi, Xj¯) = ∇dφ(Xi, Xj¯) = XiXj¯φ− dφ(∇XiXj¯). By (2.5) we have ∇XiXj¯ = −dη(Xi, Xj¯)ξ. So when φ is basic we have φ,ij¯ = φij¯. Thus our solution to (3.7) also solves (3.3) (with c(t) = 0). Hence, we have a solution to (3.2) on [0, T ). 3.12. Long-Time Existence The main theorem in [7] implies that for a compact Ka¨hler manifold (M, g0) with c1 = κ[ω0], the Ka¨hler-Ricci flow ∂ ∂t g = κg − Ricg, g(0) = g0, has a unique solution defined for all t ≥ 0 and when κ ≤ 0 the flow g(t) converges to the unique Ka¨hler-Einstein metric in c1. Using this result, it is shown in [36] that for 60 a compact Sasaki manifold (M, η0, g0) with with c 1 B = κ[dη0], the Sasaki-Ricci flow ∂ ∂t gT = κgT − RicTg , gT (0) = gT0 , has a unique solution for t ∈ [0,∞) and when κ ≤ 0 the flow gT (t) converges to a transverse Ka¨hler-Einstein metric. In the Sasaki setting, dη is non-degenerate on D and it is the transverse Ka¨hler form of the Sasaki metric. So the Sasaki-Ricci flow is really just our TKRF where the initial structure is Sasaki (rank 2n + 1 quasi-Sasaki with ω = dη). To prove the long-time existence of the flow, one needs uniform Ck-estimates for the metric potential function φ. The estimates obtained in [36] do not rely on the fact that the initial structure is Sasaki and thus can be carried over to the setting of our TKRF. Therefore, the long-time existence of the TKRF follows from the work in [36], which, in turn, follows from the work in [7]. We record this as a corollary. Corollary 3.1.4. Let (M, ξ, η0,Φ, g0) be a compact quasi-Sasaki manifold with c 1 B = κ[ω0]. Then the transverse Ka¨hler-Ricci flow ∂ ∂t gT = κgT − RicTg , gT (0) = gT0 , has a unique solution that exists for all t ≥ 0. When κ ≤ 0, the flow gT (t) converges to a transverse Ka¨hler-Einstein metric. 3.13. The Evolving quasi-Sasaki Structure Now that we have an evolving transverse Ka¨hler metric ωt = ω0 + √−1∂∂¯φ(t) that satisfies (3.2), we need an associated evolving quasi-Sasaki structure. We define 61 the evolving structure (ξ, ηt,Φt, gt) by ηt := η0 + d c Bφ, Φt := Φ0 − ξ ⊗ dcBφ ◦ Φ0, gt := ηt ⊗ ηt + ωt(Φt·, ·). As dcBφ is basic, it is clear that ηt(ξ) = η0(ξ) = 1. It is not hard to check that the relations Φ2t = −I + ξ ⊗ ηt and gt(Φt·,Φt·) = gt − ηt ⊗ ηt hold for this evolving structure. Hence, for each t ≥ 0, we have a quasi-Sasaki structure (ξ, ηt,Φt, gt) where the transverse metric gTt evolves by TKRF. It is natural to ask what happens to the rank of the quasi-Sasaki structure as it evolves under the TKRF. If the initial structure is Sasaki (rank 2n + 1), then dηt = ωt and thus the evolving structure remains Sasaki. However, the rank need not be constant along the flow. If the initial structure is co-symplectic (rank 1), then dη0 = 0. If the rank stays constant then dηt = √−1∂∂¯φ = 0 for all t. This implies that ω0 is a constant solution to (3.2). Hence ω0 is a transverse Ka¨hler- Einstein metric. So if the initial structure is rank 1 and ω0 is not Einstein, then the rank cannot remain constant. To construct such a manifold, simply take a compact Ka¨hler manifold (M˜, J, g˜) with c1 > 0 which is not Einstein (these exist) and let M = S1 × M˜ and g = dθ ⊗ dθ + g˜. Then {M,∂θ, dθ, J, g} is co-symplectic and ω is not transverse Ka¨hler-Einstein. If the initial structure has rank 2p + 1 with p < n, we would like to know what happens to the rank of the evolving structure. If the initial structure is rank 1, then the rank can only increase along the flow, but is this what happens in general? 62 We end this section with a few words on the volume form naturally associated to a quasi-Sasaki manifold. Given a quasi-Sasaki structure on M , the top form η ∧ (ω)n defines a volume form. If ωt = ω0 + √−1∂∂¯φ evolves by transverse Ka¨hler-Ricci flow, then as we noted above, ηt = η0 + d c Bφ is the evolving (almost) contact form. Since dcBφ is basic, dcBφ ∧ (ωt)n = 0, because it is a basic form of degree 2n+ 1. In a preferred coordinate chart we have gt = ηt ⊗ ηt + (gTij¯ + φij¯)dzi ⊗ dz¯j. It follows that (ωt) n = c(n) det(gTij¯ + φij¯)dz 1 ∧ dz¯1 ∧ · · · ∧ dzn ∧ dz¯n, where the constant c(n) = n! ( √−1)n 2n . Therefore, the evolving volume form ηt ∧ (ωt)n = η0 ∧ (ωt)n = c(n) det(gTij¯ + φij¯)dx ∧ dz1 ∧ dz¯1 ∧ · · · ∧ dzn ∧ dz¯n = n!dVgt . where dVgt is the Riemmanian volume form associated to gt. 63 3.2. Perelman’s entropy functional on quasi-Sasaki manifolds In this section we develop the analog of Perelman’s W functional in the quasi- Sasaki setting. Our transverse entropy functional WT can be viewed as the entropy functionalW for the transverse Ka¨hler structure. We will show thatWT is monotonic along the transverse Ka¨hler-Ricci flow. This will allow us, in the next section, to prove a non-collapsing result which can then be used to prove bounds on the scalar curvature and the diameter of the manifold along the TKRF, as in [34]. 3.21. The transverse entropy functional Let M be a quasi-Sasaki manifold. The WT functional is defined in terms of the quasi-Sasaki metric g, but it really only depends on the induced transverse metric. We make the following definition: Definition 3.2.1. For a quasi-Sasaki manifold (M, ξ, η,Φ, g), a basic function f ∈ C∞B (M) and a real number τ > 0, the transverse entropy functionalWT is defined by WT (g, f, τ) := ∫ M (τ(RT + |∇f |2) + f)e−fτ−n dV, (3.9) where dV = η ∧ (ω)n. Even though the WT functional is defined in terms of the metric g, it really only depends on the induced transverse metric gT . For this reason we may write WT (gT , ·, ·) to mean WT (g¯, ·, ·) where g¯ is any bundle-like metric inducing gT . Since the function f in (3.9) is basic, the integrand of the WT functional only involves the transverse Ka¨hler structure. Thus, when we compute the first variation of WT we are essentially just computing as we would in the Ka¨hler setting, with all of the familiar integration by parts formulas from Ka¨hler geometry at our disposal. 64 Proposition 5.0.10 and corollary 5.0.11 help make this precise. As noted in [17], corollary 5.0.11 implies a general principle that will aid us greatly in our computations. The principle is that if a result for a compact Ka¨hler manifold can be proved using only Stokes theorem and integration by parts, then the same result holds for compact quasi-Sasaki manifolds and the proof can go as in the Ka¨hler setting by inserting a “∧η” in each line of the proof. See also proposition 4.6 in [19]. We consider a variation of the quasi-Sasaki structure that fixes ξ and the transverse holomorphic structure. Let vij¯ = δg T ij¯, v = (g T )ij¯vij¯, δf = h ∈ C∞B (M) and δτ = σ. We assume that, at least locally, there is a basic function ψ such that vij¯ = ∂i∂j¯ψ. Then we have vij¯,i = v,i. Proposition 3.2.2. The first variation of WT on compact quasi-Sasaki manifolds is given by δWT (vij¯, h, σ) = ∫ M ( σ(RT + ∆Tf)− τ〈vij¯, RTij¯ + fij¯〉+ h ) τ−ne−f dV + ∫ M (v − h− nστ−1)(τ(2∆Tf − |∇Tf |2 +RT ) + f)τ−ne−f dV. Before we prove the proposition, a few comments and observations are in order. First, the Christoffel symbols of the transverse Levi-Civita connection in the ξ direction, as well as those with mixed barred and unbarred indices, are all zero. The others are given by the usual formula from Ka¨hler geometry, Γkij = (g T )kl¯ ∂gT il¯ ∂zj and Γk¯i¯j¯ = Γ k ij. 65 For a function f on M , we will use the notation fi := ∂f ∂zi = ∂if and fij¯ := ∂2f ∂zi∂z¯j = ∂i∂j¯f . Next, if f is a basic function, then |∇f |2 = |∇Tf |2 = (gT )ij¯fifj¯ and ∆f = ∆Tf = (gT )ij¯∇i∇j¯f = (gT )ij¯fij¯. Finally, for symmetric (0,2)-tensors Sij¯ and Tij¯, we have an inner product given by 〈Sij¯, Tij¯〉 := (gT )il¯(gT )kj¯Sij¯Tkl¯, and we write |Sij¯|2 = 〈Sij¯, Sij¯〉. To ease notation, we will drop the superscript T from the transverse connection, Laplacian and metric in what follows. As all of the functions involved are basic, this should not cause any confusion. The presence of one barred index and one unbarred index in the metric indicates that we are considering the transverse Ka¨hler metric. Now for the proof: Proof. From the relation gil¯gkl¯ = δ k i , we find that δg ij¯ = −gil¯vkl¯gkj¯. Now the variation of the transverse scalar curvature is δRT = δ(gij¯RTij¯) = −δ(gij¯∂i∂j¯ log(det(gTij¯))) = −〈vij¯, RTij¯〉 −∆v. 66 Next we compute δ|∇f |2 = δ(gij¯fifj¯) = −〈vij¯, fifj¯〉+ gij¯(hifj¯ + fihj¯) = −〈vij¯, fifj¯〉+ 2gT (∇f,∇h). Recall that dV = η ∧ (ω)n = c(n) det(gTij¯)dx ∧ dz1 ∧ dz¯1 ∧ · · · ∧ dzn ∧ dz¯n. Hence δdV = vdV and it follows that δ(τ−ne−fdV ) = (v − h− nστ−1)τ−ne−fdV. Putting these all together, we have δWT (vij, h, σ) = ∫ M (σ(RT + |∇f |2) + h)τ−ne−fdV + ∫ M τ(−〈vij¯, RTij¯ + fifj¯〉 −∆v + 2gT (∇f,∇h))τ−ne−fdV + ∫ M (τ(RT + |∇f |2) + f)(v − h− nστ−1)τ−ne−fdV. Observe that ∆e−f = (|∇f |2−∆f)e−f . So, since M is closed, by Stokes’ theorem we have ∫ M (|∇f |2 −∆f)e−fdV = 0. (3.10) 67 Since f, h, v ∈ C∞B (M), we have the following integration by parts formulas: ∫ M e−f∆v dV = ∫ M v(|∇f |2 −∆f)e−f dV, (3.11)∫ M 〈vij¯, fifj¯〉e−fdV = ∫ M (〈vij¯, fij¯〉 − v(∆f − |∇f |2))e−f dV, (3.12)∫ M gT (∇f,∇h)e−fdV = ∫ M h(|∇f |2 −∆f)e−f dV. (3.13) Equations (3.11) and (3.13) are straightforward to verify. For a proof of (3.12), see proposition 4.3 in [19]. Now, using equations (3.10)-(3.13) in our above expression for δWT we get δWT (vij¯, h, σ) = ∫ M ( σ(RT + ∆f)− τ〈vij¯, RTij¯ + fij¯〉+ h ) τ−ne−f dV + ∫ M (v − h− nστ−1)(τ(2∆f − |∇f |2 +RT ) + f)τ−ne−f dV. If we choose h = v − nστ−1, then the second integral in the expression for δWT vanishes. Now let vij¯ = gij¯ − RTij¯ − fij¯. Then v = n − RT − ∆f and h = n− RT −∆f − nστ−1. Putting these choices of vij¯ and h into the above and noting that 〈gij¯, RTij¯ + fij¯〉 = RT + ∆f , we find that δWT (vij¯, h, σ) = ∫ M ( τ |Rij¯ + fij¯|2 − (τ − σ + 1)(R + ∆f) + n− nστ−1 ) τ−ne−f dV. 68 Now setting σ = τ − 1, this becomes δWT (vij¯, h, σ) = ∫ M ( τ |Rij¯ + fij¯|2 − 2(R + ∆f) + nτ−1 ) τ−ne−f dV = ∫ M |Rij¯ + fij¯ − τ−1gij¯|2τ−n+1e−f dV ≥ 0. (3.14) The next goal is to show that WT is monotonically increasing along the TKRF. Let us outline one way to accomplish this. (i) First show that for a diffeomorphism ϕ : M →M we have WT (g, f, τ) =WT (ϕ∗g, ϕ∗f, τ). (3.15) This is relatively straightforward. (ii) Next find a smooth basic function f = f(t) satisfying ∂f ∂t = nτ−1 −RT −∆f + |∇f |2. This is a backward heat equation and for any time T > 0, if f(T ) ∈ C∞B (M), then there is a unique basic solution on [0, T ]. To prove this, one can argue as follows: Proof (sketch). Pick T > 0 and consider the equation ∂w ∂s = ∆w − (RT − nτ−1)w. (3.16) This is a linear parabolic equation as long as gT remains positive definite. Since gT is positive definite for all time along the TKRF, by the standard parabolic 69 theory, given an initial condition, a unique solution exists for s ∈ [0, T ]. If w(0) is basic, then the solution remains basic. Indeed, as RT is basic, ξw satisfies ∂ ∂s (ξw) = ξ ∂w ∂s = ξ∆w − ξ(RT + nτ−1)w = ∆(ξw)− (RT + nτ−1)(ξw). So if ξw(0) = 0, then by the uniqueness of the solution, we must have ξw ≡ 0. If w(0) ≥ 0, then w(s) ≥ 0. Furthermore, if w(0) is positive at some point p ∈ M , then w(s) > 0. By the standard regularity theory for linear parabolic equations, if we have w(0) ∈ W 1,2(M) then w(s) ∈ C∞(M). See proposition 4.9 in [19]. We can solve equation (3.16) with w(0) > 0 a basic function. Then we can write w(s) = e−f(s) for some smooth function f . As w is basic for all s, so is f . Let t = T − s. Then for t ∈ [0, T ], it’s not hard to check that f(t) solves the backward heat equation ∂f ∂t = nτ−1 −RT −∆Bf + |∇f |2. (iii) Now consider the time-dependent diffeomorphisms ϕ(t, ·) : M →M having ∂ ∂t ϕ(t, p) = −∇f(t)|ϕ(t,p), ϕ(0, ·) = I. That these diffeomorphisms exist is somewhat standard and a proof can be found, for example, in lemma 3.15 of [10]. Next, direct computation shows that if g evolves by TKRF, then g˜ := ϕ∗g induces a transverse metric which evolves 70 by ∂ ∂t g˜Tij = g˜ij¯ − R˜Tij¯ − f˜ij¯, where R˜Tij¯ is the transverse Ricci tensor of g˜ and f˜ = ϕ ∗f is a basic function satisfing ∂f˜ ∂t = nτ−1 − R˜T −∆f˜ . (iv) Finally, letting τ evolve by ∂τ ∂t = τ − 1, the computations in (3.14) and (3.15) then imply that ∂W T ∂t ≥ 0 along the TKRF. This outline is similar to the approach in [25] for the Ricci flow and the one taken in [12] for the Sasaki-Ricci flow. We can also make this outline work for us, but there are some technical difficulties involved. One difficulty encountered with this approach is that the diffeomorphisms ϕ may not preserve the transverse holomorphic structure. The pulled back metrics ϕ∗g may not be quasi-Sasaki as the induced transverse metric may not even by Hermitian. This is not a terrible difficulty to overcome though. One must simply define WT on a larger class of metrics. In particular, the class M of Riemannian metrics on M for which ξ is a unit-length Killing vector field will suffice. The metrics in M are bundle-like with respect to Fξ and thus induce transverse metrics. But now one must check that WT has the same variational formula when the metrics vary inM rather than as quasi-Sasaki metrics. This can be done and the proof is the same as in the Riemannian setting. Then one must show that if g ∈ M then ϕ∗g ∈ M, that ϕ∗ξ = ξ and that if f is basic then ϕ∗f is basic too. All of this can be done but we will choose not to pursue it any further. 71 Another approach is to show by direct computation that ∂W T ∂t ≥ 0 under the coupled transverse Ka¨hler-Ricci flow  ∂gT ∂t = gT − RicTg ∂f ∂t = nτ−1 −RT −∆f + |∇f |2 ∂τ ∂t = τ − 1 (3.17) This approach takes advantage of the transverse Ka¨hler structure and we can remain in the category of quasi-Sasaki metrics. We can prove the following proposition, which is the same as proposition 4.4 in [19]: Proposition 3.2.3. Under the coupled transverse Ka¨hler-Ricci flow (3.17), the WT functional on quasi-Sasaki manifolds satisfies ∂WT ∂t = ∫ M |Rij¯ + fij¯ − τ−1gij¯|2τ−n+1e−f dV + ∫ M |fij|2τ−n+1e−f dV ≥ 0. (3.18) Moreover, WT is strictly increasing unless RTij¯ + fij¯ − τ−1gTij¯ = 0 and fij = 0. In this case, ∇f is a real holomorphic vector field and gT is a transverse (gradient-shrinking) Ka¨hler-Ricci soliton. The proof involves a careful computation with several uses of integration by parts. Some of the integration by parts formulas are not obvious. As the proof is the same as in [19], we do not reproduce it here. Next we discuss the analog of Perelman’s µ functional. 72 3.22. The transverse µ-functional In the last subsection, we saw that if g evolves as ∂tgij¯ = gij¯ − RTij¯ − fij¯, f evolves as ∂tf = nτ −1 −RT −∆Bf and τ evolves as ∂tτ = τ − 1, then WT (g, f, τ) is non-decreasing with t and gij¯∂tgij¯ − ∂tf − nτ ∂tτ = 0. Hence ∂ ∂t (τ−ne−fdV ) = 0. This implies that ∫ M τ−ne−fdV is constant and motivates our next definition. Definition 3.2.4. The transverse µ-functional, µT , is defined by µT (g, τ) := inf f∈C∞B (M) {WT (g, f, τ) : ∫ M τ−ne−fdV = 1}. As it is the case with the WT -functional, it is not surprising that the µT - functional is also monotonically non-decreasing along the TKRF when ∂tτ = τ − 1. We prove this next. Proposition 3.2.5. If g(t) evolves by TKRF and τ satisfies ∂tτ = τ − 1, then µT (g(t), τ(t)) is monotone non-decreasing in t. In particular, µT (g(t), 1) is non- decreasing along the transverse Ka¨hler-Ricci flow. Proof. Fix a time t1 > 0 and let ft1 ∈ C∞B (M) be normalized by ∫ M τ−ne−ft1dV = 1. Then solve the backward heat equation ∂f ∂t = nτ−1 −RT −∆Bf + |∇f |2 73 for t ∈ [0, t1] with f(t1) = ft1 . We know that the solution f(t) remains basic. When g evolves by TKRF and ∂tτ = τ − 1, we can compute that ∂ ∂t (τ−ne−fdV ) = (∆f − |∇f |2)τ−ne−fdV = −τ−n∆e−fdV. Thus, ∂ ∂t ∫ M τ−ne−f dV = − ∫ M τ−n∆e−f dV = 0. This implies that ∫ M τ−ne−f(t) dV = 1 for all t ∈ [0, t1]. Now chose 0 ≤ t0 < t1. By the definition of µT and the monotonicity of WT , we get µT (g(t0), τ(t0)) ≤ WT (g(t0), f(t0), τ(t0)) ≤ WT (g(t1), f(t1), τ(t1)). Since ft1 = f(t1) was chosen arbitrarily, we conclude that µT (g(t0), τ(t0)) ≤ inf f(t1) WT (g(t1), f(t1), τ(t1)) = µT (g(t1), τ(t1)). 3.3. Bounds on Scalar Curvature and the Transverse Ricci Potential We assume there is a transverse Ka¨hler metric ω such that [ω] = 2pic1B. Then by the transverse ∂∂¯-lemma, there is smooth function u(x, t) such that gTij¯−RTij¯ = ∂i∂j¯u. We call u a transverse Ricci potential. We may assume that u is normalized so that∫ M e−u dV = 1. If φ is a solution to (3.3), then ∂i∂j¯ ∂φ ∂t = ∂gTij¯ ∂t = gTij¯ −RTij¯ = ∂i∂j¯u. 74 So we can take u = ∂tφ. As R T ij¯ = −∂i∂j¯ log(det(gT )), we compute that ∂i∂j¯ ∂u ∂t = ∂ ∂t (gTij¯ −RTij¯) = gTij¯ −RTij¯ + ∂i∂j¯gij¯ ∂gTij¯ ∂t = ∂i∂j¯(u+ g ij¯∂i∂j¯u) = ∂i∂j¯(u+ ∆ Tu). Thus, there is time-dependent constant A(t) such that ∂u ∂t = u+ ∆Tu+ A(t). (3.19) As ∂tdV = ∆ Tu dV , differentiating ∫ M e−u dV = 1 with respect to t, we find that A(t) = − ∫ M ue−u dV. We can use the monotonicity of the µT -functional to show that A is uniformly bounded. Recall from proposition 3.2.5 that µT (g(t), 1) is non-decreasing in t. Lemma 3.3.1. The function A(t) is uniformly bounded. Proof. It is a routine calculus exercise to show that f(t) = te−t is bounded above by e−1. From ∂tdV = ∆Tu dV , we see that the volume of the manifold is constant along the TKRF. For simplicity, let us assume that Vol(M) = ∫ M dV = 1. Then A(t) = − ∫ M ue−u dV ≥ −e−1 ∫ M dV = −e−1. 75 Tracing gTij¯ −RTij¯ = ∂i∂j¯u, we have ∆Tu = n−RT . Thus WT (g(t), u(t), 1) = ∫ M (n−∆Tu+ |∇u|2 + u)e−u dV = ∫ M ∆T e−u dV + ∫ M (n+ u)e−u dV = n− A(t). Therefore, A(t) = n−WT (g(t), u(t), 1) ≤ n− µT (g(t), 1) ≤ n− µT (g(0), 1). Next we will see how the transverse scalar curvature evolves along the TKRF, but first we recall a useful lemma. The proof is a standard calculus argument. Lemma 3.3.2. If f = f(x, t) is a smooth function on M × [0, T ] for some T > 0 and f achieves its minimum (maximum) at (x0, t0), then either t0 = 0 or at the point (x0, t0) ∂f ∂t ≤ 0 (≥ 0), ∇f = 0 and ∆f ≥ 0 (≤ 0). Lemma 3.3.3. Along the TKRF, the transverse scalar curvature evolves by ∂RT ∂t = −RT + |RicT |2 + ∆TRT and is uniformly bounded from below. Proof. Because of the transverse Ka¨hler structure, the Ricci tensor RTij¯ = −∂i∂j¯ log(det(gij¯)). Thus we can write the scalar curvature as RT = −gij¯∂i∂j¯ log(det(gij¯)). 76 From gil¯gkl¯ = δ i k, we compute ∂tg ij¯ = −gil¯(∂tgkl¯)gkj¯. Therefore, ∂RT ∂t = gil¯(∂tgkl¯)g kj¯∂i∂j¯ log(det(gij¯))− gij¯∂i∂j¯gij¯∂tgij¯ = −gil¯(gkl¯ −RTkl¯)gkj¯RTij¯ − gij¯∂i∂j¯gij¯(gij¯ −RTij¯) = −δikgkj¯RTij¯ + gil¯gkj¯RTkl¯RTij¯ − gij¯∂i∂j¯(n) + gij¯∂i∂j¯gij¯RTij¯ = −RT + |RicT |2 + ∆TRT . Now fix a time T > 0 and suppose that RT (x, t) on M × [0, T ] takes its minimum at (x0, t0). If t0 = 0, then we have inf M×[0,T ] RT (x, t) ≥ inf M RT (x, 0). Since RT is basic, ∆TRT = ∆RT . So if t0 6= 0, then by the above lemma, at (x0, t0), 0 ≥ ∂R T ∂t −∆TRT = −RT + |RicT |2 ≥ −RT . So either way, as T > 0 was arbitrary, we have RT ≥ min{0, infM RT (x, 0)}. Lemma 3.3.4. The function u(x, t) is uniformly bounded from below. Proof. Fix T > 0. By lemmas 3.3.1 and 3.3.3, A(t)−RT is uniformly bounded above. Set C = max{0, sup x∈M,t∈[0,T ] n−RT + A(t)} <∞. Suppose that u(x, t) is not uniformly bounded below. Then there is a point (x0, t0) where u(x0, t0) +C < 0. Since u is smooth in x, there is a neighborhood U of x0 such that u(x, t0) + C < 0 77 for all x ∈ U . Recall that ∂tu = u+ ∆Tu+A(t) and ∆Tu = n−RT . Thus, for x ∈ U , ∂u ∂t |(x,t0) = u(x, t0) + n−RT (x, t0) + A(t0) < 0. Since u is smooth in t, u(x, t) ≤ u(x, t0) for x ∈ U and t ≥ t0. From ∂tu < u+ C we get u(x, t) < (C + u(x, t0))e t−t0 ≤ −C1et (3.20) for x ∈ U , t ≥ t0 and some constant C1 > 0 that depends on t0. Recall that ∂tφ = u. Integrating equation (3.20) from t0 to t gives φ(x, t) ≤ φ(x, t0)− C1(et − et0) ≤ −C2et (3.21) for x ∈ U , t ≥ t0 sufficiently large and C2 > 0. As u is normalized so that ∫ M e−u dV = 1, −u cannot be very large. Thus there is a constant C3 > 0 such that u(xt, t) := max x∈M u(x, t) ≥ −C3. From ∂t(u− φ) = n−RT + A(t) ≤ C we get u(xt, t)− φ(xt, t) ≤ max x∈M (u(x, 0)− φ(x, 0)) + Ct. This implies there is a constant C4 so that φ(x′t, t) := max x∈M φ(x, t) ≥ −C4 − Ct. (3.22) 78 Tracing gTij¯(t) = gij¯(0) + ∂i∂j¯φ, with the metric g(0) gives n + ∆g(0)φ = gij¯(0)gTij¯(t) > 0. Recall that the volume of the manifold is constant along the TKRF and we are assuming that vol(M) = 1. Let G0 be the Green’s function for ∆g(0). The function φ−∫ M φ dV0 integrates to zero with respect to dV0, so we may apply Green’s formula to write φ(x′t, t) = ∫ M φ(y, t) dV0 − ∫ M ∆g(0)φ(y, t)G0(x ′ t, y) dV0. (3.23) By (3.21), φ(x, t) ≤ −C2et on U × [t0,∞). Since −∆g(0)φ < n and G0(x, y) is integrable, from (3.23) we get that for t ≥ t0 φ(x′t, t) ≤ vol(M \ U)φ(x′t, t)− vol(U)C2et + C˜. Note that vol(M \ U) < 1. So for large enough t ≥ t0, φ(x′t, t) ≤ −C5et + C6. (3.24) Combining (3.22) and (3.24), we have −C4 − Ct ≤ φ(x′t, t) ≤ −C5et + C6. This is a contradiction for sufficiently large t as all of the constants involved are independent of t > t0. Therefore, u(x, t) is uniformly bounded from below. Lemma 3.3.5. There is a uniform constant C > 0 such that |∇u|2 ≤ C(u+ C). Before we begin the proof, we need one new bit of notation. For a Ka¨hler metric g, we define ∇f to be the vector field dual to ∂¯f . We define ∇¯f to be the 79 vector field dual to ∂f . In local coordinates, ∇f = gij¯fj¯dzi, ∇¯f = gij¯fidz¯j and g(∇f, ∇¯f) = gij¯fifj¯ = |∇f |2. We make the analogous definition for a basic function and the transverse Ka¨hler metric of a quasi-Sasaki structure. Proof. Define the parabolic operator  := ∂t − ∆ and denote g(·, ·) = 〈·, ·〉. One computes  ( f h ) = f h − fh h2 + 2<〈∇f, ∇¯h〉 h2 − 2f |∇h| 2 h3 , where < denotes the real part of a complex number. From (3.19) we have u = u+ A(t). A local coordinate computation gives ∂t|∇u|2 = |∇u|2 + 〈Rij¯, uiuj¯〉+ 2<〈∇u, ∇¯∆u〉, and the Bochner-Kodaira formula gives ∆|∇u|2 = |∇∇u|2 + |∇∇¯u|2 + 〈Rij¯, uiuj¯〉+ 2<〈∇u, ∇¯∆u〉. Therefore |∇u|2 = |∇u|2 − |∇∇u|2 − |∇∇¯u|2. 80 By lemma 3.3.4, there is a constant B > 0 such that u(x, t) + B > 0 for all x ∈M and all t ∈ [0,∞). Now let H := |∇u|2/(u+ 2B). Then H = |∇u| 2 u+ 2B − |∇u| 2u (u+ 2B)2 + 2<〈∇|∇u|2, ∇¯u〉 (u+ 2B)2 − 2|∇u| 2|∇u|2 (u+ 2B)3 = |∇u|2 − |∇∇u|2 − |∇∇¯u|2 u+ 2B − |∇u| 2(u+ A) (u+ 2B)2 + 2<〈∇|∇u|2, ∇¯u〉 (u+ 2B)2 − 2|∇u| 4 (u+ 2B)3 = −|∇∇u|2 − |∇∇¯u|2 u+ 2B + |∇u|2(2B − A) (u+ 2B)2 + 2<〈∇|∇u|2, ∇¯u〉 (u+ 2B)2 − 2|∇u| 4 (u+ 2B)3 . (3.25) By the definition of H, ∇H = ∇|∇u| 2 u+ 2B − |∇u| 2∇u (u+ 2B)2 . Hence <〈∇H, ∇¯u〉 u+ 2B = <〈∇|∇u|2, ∇¯u〉 (u+ 2B)2 − |∇u| 4 (u+ 2B)3 . (3.26) Next we observe that |〈∇|∇u|2, ∇¯u〉| (u+ 2B)2 ≤ |∇u| 2(|∇∇u|+ |∇∇¯u|) (u+ 2B)3/2(u+ 2B)1/2 ≤ |∇u| 4 2(u+ 2B)3 + |∇∇u|2 + |∇∇¯u|2 u+ 2B . (3.27) Choosing 0 < ε < 1/4 and combining (3.25), (3.26) and (3.27), yields H ≤ |∇u| 2(2B − A) (u+ 2B)2 + (2− ε)<〈∇H, ∇¯u〉 u+ 2B − ε|∇u| 4 2(u+ 2B)3 . (3.28) 81 If H is unbounded from above, then there is a time t0 such that max M×[0,t0] H > 2B − A ε/2 . At the point where H achieves its maximum, ∇H = 0 and H ≥ 0. At this point, (3.28) gives 0 ≤ H ≤ |∇u| 2 (u+ 2B)2 ( 2B − A− ε 2 H ) < 0. This is a contradiction and therefore H is bounded from above. Thus there is a constant C > 0 such that |∇u|2 ≤ C(u+ C). Lemma 3.3.6. There is a uniform constant C > 0 such that RT ≤ C(u+ C). Proof. From ∂i∂j¯u = g T ij¯ −RTij¯, ∆u = n−RT and ∂tRT = −RT + |RicT |2 + ∆TRT , we compute (∆u) = −RT = RT − |RicT |2 = n−∆u− 〈gTij¯ − uij¯, gTij¯ − uij¯〉 = n−∆u− (n− 2∆u+ |uij¯|2) = ∆u− |uij¯|2. Let K := −∆u/(u+ 2B) where B > 0 is as in the previous lemma. We compute K = −∆u u+ 2B + ∆uu (u+ 2B)2 − 2<〈∇∆u, ∇¯u〉 (u+ 2B)2 + 2∆u|∇u|2 (u+ 2B)3 = −∆u+ |uij¯|2 u+ 2B + ∆u(u+ A) (u+ 2B)2 − 2<〈∇∆u, ∇¯u〉 (u+ 2B)2 + 2∆u|∇u|2 (u+ 2B)3 = |uij¯|2 u+ 2B + ∆u(A− 2B) (u+ 2B)2 − 2<〈∇∆u, ∇¯u〉 (u+ 2B)2 + 2∆u|∇u|2 (u+ 2B)3 . (3.29) 82 By the definition of K, ∇K = −∇∆u u+ 2B + ∆u∇u (u+ 2B)2 . Hence <〈∇K, ∇¯u〉 u+ 2B = −<〈∇∆u, ∇¯u〉 (u+ 2B)2 + ∆u|∇u|2 (u+ 2B)3 . Combining this with (3.29) yields K = |uij¯| 2 u+ 2B + ∆u(A− 2B) (u+ 2B)2 + 2<〈∇K, ∇¯u〉 u+ 2B . (3.30) Let H be as in the previous lemma. There we computed that H = −|uij| 2 − |uij¯|2 u+ 2B + |∇u|2(2B − A) (u+ 2B)2 + 2<〈∇H, ∇¯u〉 u+ 2B . Now let G := 2H +K. Then by (3.30) and the above, we have G = −2|uij| 2 − |uij¯|2 u+ 2B + (2|∇u|2 −∆u)(2B − A) (u+ 2B)2 + 2<〈∇G, ∇¯u〉 u+ 2B . (3.31) At the point where G achieves its maximum, ∇G = 0 and G ≥ 0. So at this point, equation (3.31) yields 0 ≤ G ≤ −|uij¯| 2 u+ 2B + 2|∇u|2(2B − A) (u+ 2B)2 − ∆u(2B − A) (u+ 2B)2 . (3.32) In normal coordinates at the point where G is at a maximum, by Cauchy-Schwarz (∆u)2 = (∑ i ui¯i )2 ≤ n ∑ i u2i¯i = n|uij¯|2. 83 So by (3.32) we get 0 ≤ G ≤ 2|∇u| 2(2B − A) (u+ 2B)2 − ∆u u+ 2B ( ∆u n + 2B − A u+ 2B ) . (3.33) Since inf u+ 2B > 0 and A(t) is uniformly bounded, (2B−A)/(u+ 2B) is uniformly bounded from above. By the bound on |∇u|2 from the previous lemma, the first term on the right hand side of (3.33) is uniformly bounded from above. So if −∆u/(u+2B) becomes arbitrarily large, then ∆u −(u+2B) < 0 and the right hand side of (3.33) goes negative. This is a contradiction. Thus −∆u/(u+ 2B) is bounded from above. Hence there is a constant C > 0 such that −∆u ≤ C(u+C). Since RT = n−∆u, the proof is complete. Proposition 3.3.7. Let xt ∈M be such that u(xt, t) = miny∈M u(y, t). Then there is a uniform constant C > 0 such that u(y, t) ≤ C(d2t (xt, y) + 1), |∇u| ≤ C(dt(xt, y) + 1), RT (y, t) ≤ C(d2t (xt, y) + 1), where dt is the distance function for the metric g(t). We prove the bound on u. The bounds on |∇u| and RT then follow from lemmas 3.3.5 and 3.3.6 respectively. By lemma 3.3.4, u is uniformly bounded below. Thus we may assume that u > 0. If not, then we just consider the function u˜ := u + C where C > 0 is a constant. Then u˜ is a transverse Ricci potential and choosing C large enough makes u˜ > 0. 84 Proof. The bound on |∇u|2 from lemma 3.3.5 implies that |∇√u| ≤ C0. So all of the first derivatives of √ u are uniformly bounded (recall that u is basic). Thus √ u is uniformly Lipschitz. Hence | √ u(y, t)− √ u(z, t)| ≤ C0dt(y, z). Now if u(xt, t) ≥ C(t) and C(t)→∞ as t→∞, then by the normalization of u and the fact that the volume of the manifold is constant along the flow, we get 1 = ∫ M e−u dV ≤ vol(M)e−C(t) → 0. Contradiction. Thus u(xt, t) ≤ C1 for some constant C1 > 0. Therefore, √ u(y, t) ≤ C0dt(y, xt) + √ u(xt, t) ≤ C0dt(y, xt) + √ C1. This implies that u(y, t) ≤ Cd2t (y, xt) + C for some constant C > 0. We see from this proposition that if we can bound the diameter of the manifold along the TKRF, then we will have uniform bounds on the transverse Ricci potential u and the transverse scalar curvature RT . We take this up in the next section. 3.4. Upper Bound on Diameter In this section we will prove a uniform upper bound on the diameter of a quasi- Sasaki manifold along the TKRF. This along with proposition 3.3.7 will provide uniform bounds on u, ∇u and RT . We take an approach analogous to that of Sesum and Tian in [34], but we must make some adjustments. The W T functional only 85 involves basic functions, but the distance function of a quasi-Sasaki manifold is not basic. To deal with this, we work with the transverse distance function from definition 1.4.1. We begin with a non-collapsing theorem. Since RT is bounded from below along the TKRF (lemma 3.3.3), there is r0 > 0 small enough so that R T (x, t) ≥ −r−20 for all x and t. Theorem 3.4.1. Let M be a quasi-Sasaki manifold whose ξ orbits are compact. Let g(t) be a solution to the TKRF. Then there is a constant C > 0 such that for every x ∈M , if RT ≤ Cr−2 on BTg(t)(x, r) for r ∈ (0, r0], then Vol(BTg(t)(x, r)) ≥ Cr2n. Proof. We give a proof by contradiction. Suppose that the theorem is false. Then there is a sequence of points and times (pk, tk) ∈M × [0,∞) with tk →∞ such that RT ≤ Cr−2k on Bk := BTg(tk)(pk, rk) but r−2nk Vol(Bk)→ 0 as k →∞. Let ψ : [0,∞)→ R be a plateau function such that ψ = 1 on [0, 1 2 ], is decreasing on [1 2 , 1] with bounded derivative and ψ = 0 on [1,∞). Let τk := r2k and define wk(x) := e Ckψ ( r−1k d T g(tk) (pk, x) ) =: eCkψk(x), where the constant Ck is chosen so that ∫ M w2kτ −n k dVk = 1. Here dVk = ηtk ∧ (ωtk)n is the volume form at time tk. Note that the support of wk is contained in Bk. Thus 1 = ∫ M w2kτ −n k dVk = e 2Ckr−2nk ∫ M ψ2 ( r−1k d T k (pk, x) ) dVk ≤ e2Ckr−2nk vol(Bk). Since 1 ≤ e2Ckr−2nk Vol(Bk) and r−2nk Vol(Bk)→ 0, it must be that Ck →∞ as k →∞. Recall, WT (g, f, τ) = τ−n ∫ M τ(RTw2 + 4|∇w|2) − w2 log(w2) dV, where w = e−f/2. 86 Now WT (g(tk), wk, τk) = τ−nk ∫ M r2k(R Tw2k + 4|∇wk|2)− w2k log(w2k) dVk ≤ e2Ckτ−nk ∫ M 4|ψ′k|2 − ψ2k log(ψ2k) dVk + r2k max Bk RT − 2Ck, Here we have used that dTg(tk) is Lipschitz continuous, hence differentiable almost everywhere, with Lipschitz constant 1. Let Vk(r) := Vol(Bk(r)). Notice that for t ∈ R, limt→0 t2 log(t2) = 0 and for t ∈ (0, 1], t2 log(t2) ≥ −1. Since ψk takes its values in [0, 1] and ψ ′ k is bounded with support in [ 1 2 , 1], there is some constant C ′ > 0 such that ∫ M 4|ψ′k|2 − ψ2k log(ψ2k) dVk ≤ C ′(Vk(rk)− Vk(rk/2)). By lemma 5.0.13, we may assume that Vk(rk) ≤ 5nVk(rk/2). Hence, ∫ M 4|ψ′k|2 − ψ2k log(ψ2k) dV gk ≤ C˜Vk(rk/2) ≤ C˜ ∫ M ψ2k dVk. Combining this with the above, we find that µT (g(tk), τk) ≤ WT (g(tk), wk, τk) ≤ C˜ ∫ M w2kτ −n k dVk + r 2 k max Bk RT − 2Ck ≤ C˜ + C − 2Ck → −∞. By proposition 3.2.5, µ(g(t), 1 + (τ0 − 1)et) is non-decreasing in t. Now for each k, set τ k0 := 1− (1− τk)e−tk . Then µ(g(0), τ k0 ) ≤ µ(g(tk), τk)→ −∞ as k → −∞. This 87 is a contradiction because µ(g(0), τ) is a continuous function of τ and τ k0 → 1 as k →∞. Let u be a transverse Ricci potential. Recall that u is a basic function. By lemma 3.3.4, u is uniformly bounded from below. Let xt ∈M be such that u(xt, t) = minx∈M u(x, t) and let dTt (y) = d T g(t)(xt, y). For k1 ≤ k2, we define the transverse annulus Bξ(k1, k2) := {y ∈M : 2k1 ≤ dTt (y) ≤ 2k2}. Now we can prove the diameter bound. Theorem 3.4.2. Let M be a quasi-Sasaki manifold whose ξ orbits are compact. Let g(t) be a solution to the TKRF. Then there is a uniform constant C such that the transverse diameter ΘTg(t) < C. This implies the diameter of M is uniformly bounded along the flow. Proof. Suppose that the transverse diameters are not uniformly bounded. Then there is a sequence of times ti → ∞ such that ΘTg(ti) → ∞. Let εi > 0 be a sequence such that εi → 0. By lemma 5.0.14, there are sequences ki1, ki2, ri1, ri2 and a uniform constant C such that 1. Vi(k i 1, k i 2) := Volg(ti)(Bξ(k i 1, k i 2)) < εi 2. Vi(k i 1, k i 2) ≤ 210nVi(ki1 + 2, ki2 − 2) 3. ri1 ∈ [ki1, ki1 + 1], ri2 ∈ [ki2 − 1, ki2] and ∫ Bξ(r i 1,r i 2) RT dVti ≤ CVi(ki1, ki2). Now let ψi be a cut-off function such that ψi(t) = 1 for t ∈ [2ki1+2, 2ki2−2], ψi(t) = 0 for t ∈ (−∞, 2ri1 ] ∪ [2ri2 ,∞) and |ψ′(t)| < 2. Define functions wi(y) := eCiψi(dTti(y)) 88 where the constant Ci is chosen so that ∫ M w2i dVti = 1. Then we have 1 = e2Ci ∫ M ψ2i dVti ≤ e2CiV1(ki1, ki2) < e2Ciεi. Since εi → 0, it must be that Ci →∞. Now we compute WT (g(ti), wi, 1) = ∫ M RTw2i + 4|∇wi|2 − w2i log(w2i ) dVti = e2Ci ∫ M RTψ2i + 4|ψ′i|2 − ψ2i log(ψi)2 − 2Ciψ2i dVti = e2Ci ∫ M RTψ2i + 4|ψ′i|2 − ψ2i log(ψ2i ) dVti − 2Ci. (3.34) By lemma 5.0.14, we can estimate the first term in the integrand of (3.34) as e2Ci ∫ M RTψ2i ≤ e2Ci ∫ Bξ(r i 1,r i 2) RT dVti ≤ e2CiCVi(ki1, ki2) ≤ e2CiC210nVi(ki1 + 2, ki2 − 2) ≤ C210n ∫ m w2i dVti = C210n Since 0 ≤ ψi ≤ 1 and ψ′i is uniformly bounded, there is a uniform constant C such that 4|∇ψi|2 − ψ2i log(ψ2i ) ≤ C. Using lemma 5.0.14 again we can estimate the remaining 89 terms in (3.34). Note that Bξ(r i 1, r i 2) ⊆ Bξ(ki1, ki2). e2Ci ∫ M 4|ψ′i|2 − ψ2i log(ψ2i ) dVg(ti) ≤ e2CiCVi(ki1, ki2) ≤ e2CiC210nVi(ki1 + 2, ki2 − 2) ≤ C210n ∫ M w2i dVti = C210n Therefore µT (g(0), 1) ≤ µT (g(ti), 1) ≤ WT (g(ti), wi, 1) ≤ C210n − 2Ci → −∞. This contradicts lemma 5.0.12. Thus the transverse diameter is uniformly bounded along the TKRF. 3.5. Transverse Holomorphic Bisectional Curvature Recall the transverse holomorphic bisectional curvature defined in section 2.4. In this section we will prove Theorem 3.5.1. Let gT (t) be a solution to the TKRF. If the initial metric gT (0) has nonnegative transverse holomorphic bisectional curvature then so does gT (t) for t ≥ 0. If the initial metric has positive transverse holomorphic bisectional curvature at some point x ∈M , then gT (t) has positive transverse holomorphic bisectional curvature for t > 0. To prove theorem 3.5.1, we need to compute the evolution of the transverse holomorphic bisectional curvature along the TKRF. By standard computations in 90 Ka¨hler-Ricci flow, ∂ ∂t Ri¯ijj¯ = ∆Ri¯ijj¯ + F (Rm)i¯ijj¯ +Ri¯ijj¯, where, for a tensor S with the same type as the curvature tensor, F (S)i¯ijj¯ = ∑ k,l (|Sik¯lj¯|2 − |Sik¯jl¯|2 + Si¯ilk¯Skl¯jj¯)−∑ k < (Sik¯Ski¯jj¯ + Sjk¯Si¯ikj¯) . See proposition 2.79 and corollary 2.82 in [9] for the computations in the Ka¨hler setting. We can easily mimic these computations and derive a similar formula in the quasi-Sasaki setting. We let HT = HTij := R T i¯ijj¯. Then we have ∂ ∂t HT = ∆THT + F (HT ) +HT . Using Hamilton’s maximum principle for tensors, Bando (n ≤ 3) in [2] and Mok (n ≥ 4) in [29] proved the analog of theorem 3.5.1 for the Ka¨hler-Ricci flow. To apply this maximum principle, one must show that F has the null vector property. That is, if there are nonzero X, Y ∈ TpM such that Sp(X, X¯, Y, Y¯ ) = 0, then F (S)p(X, X¯, Y, Y¯ ) ≥ 0. We have the following: Proposition 3.5.2. If S ≥ 0 and there exists X, Y ∈ D1,0, both non-zero, such that Sp(X, X¯, Y, Y¯ ) = 0, then F (S)p(X, X¯, Y, Y¯ ) ≥ 0. Proof. This is a local problem, so we choose a coordinate patch pi : U ⊂M → V ⊂ Cn with p ∈ U . Then the transverse metric gT on U induces a Ka¨hler metric gV on V which is evolving by Ka¨hler-Ricci flow. So F (RT ) restricted to V is the same as in the Ka¨hler case. The result now follows from the analogous result in the Ka¨hler setting. See [2], [29] or page 107 of [9]. Now we can prove theorem 3.5.1: 91 Proof. It suffices to prove the theorem for short time, say for t ∈ [0, t0] for some t0 > 0. Then all metrics involved have bounded geometry. We define an auxiliary tensor S in local coordinates by Sij¯kl¯ := 1 2 ( gTij¯g T kl¯ + g T il¯g T kj¯ ) . Note that S has the same type as the curvature tensor, S > 0 at all points of M and S is parallel, hence ∆S = 0. Then there is a constant C1 > 0 such that −C1S ≤ ∂S ∂t ≤ C1S. Since F is smooth and S > 0, there is a constant C2 > 0 such that F (HT )− F (HT + fS) ≥ −C2|f |S for f ∈ R with |f | ≤ 1. When f is a basic function, fS has the same type as the transverse curvature tensor. Let f = f(t, x) be a basic function for all t and choose ε > 0 small enough so that ε|f(t, x)| ≤ 1 for all (t, x). Then we compute ∂ ∂t (HT + εfS) = ∆HT + F (HT ) +HT + ε ∂f ∂t S + εf ∂S ∂t = F (HT )− F (HT + εfS) + F (HT + εfS) +HT + εfS + ∆(HT + εfS) + εS ( ∂f ∂t −∆f ) + εf ( ∂S ∂t − S ) ≥ F (HT + εfS) + ∆(HT + εfS) +HT + εfS + εS ( ∂f ∂t −∆f − (C1 + 1 + C2)|f | ) . 92 Now let f satisfy ∂f ∂t −∆f − (C1 + 1 + C2)f = 1 with initial condition f(·, 0) ≡ 1. In local coordinates, ∆f = ξ2f + gij¯∂i∂j¯f . Since ξgij¯ = 0, we have ξ(∆f) = ∆(ξf). Hence ∂(ξf) ∂t = ∆(ξf) + (C1 + 1 + C2)ξf. Since ξf(0) = 0, it follows from the uniqueness of the solution that ξf ≡ 0. Thus f(·, t) is basic. By the maximum principle, f > 0 for t > 0. Therefore ∂ ∂t (HT + εfS) ≥ F (HT + εfS) + ∆(HT + εfS) +HT + εfS + εS. (3.35) Since HT ≥ 0 at t = 0, we have HT + εfS > 0 at t = 0. In fact, HT + εfS > 0 for all t. We give a proof by contradiction. If not, then there is a first time t0 > 0, a point p ∈M and unit vectorsX, Y ∈ D1,0p such that (HT + εfS)(X, X¯, Y, Y¯ ) = 0 at (t0, p). By proposition 3.5.2, at (t0, p), F (HT + εfS)(X, X¯, Y, Y¯ ) ≥ 0. Now choose a coordinate patch U about p ∈ M . Then pi : U → V ⊂ Cn is a Riemannian submersion. Let XV = pi∗X, YV = pi∗Y . Extend XV and YV to a normal neighborhood of (pi(p), t0) by parallel translation along radial geodesics such that ∇XV = ∇YV = 0 at (pi(p), t0). Then extend XV and YV around t0 such that 93 ∂tXV = ∂tYV = 0 at (pi(p), t0). Using (3.35) at (p, t0) we have 0 ≥ ∂ ∂t ( HT + εfS ) (XV , X¯V , YV , Y¯V ) = ( ∂ ∂t (HT + εfS) ) (XV , X¯V , YV , Y¯V ) ≥ (∆(HT + εfS)) (XV , X¯V , YV , Y¯V ) + εS(XV , X¯V , YV , Y¯V ) = ∆ ( (HT + εfS)(XV , X¯V , YV , Y¯V ) ) + εS(XV , X¯V , YV , Y¯V ) > 0. Contradiction. Thus HT + εfS > 0 for all t. Letting ε → 0, we have HT ≥ 0 for all t. Now suppose that HT > 0 at some point in M when t = 0. Define f0(p) := minH T p (X, X¯, Y, Y¯ ) for X, Y ∈ D1,0p with |X| = |Y | = 1. Then f0 ≥ 0 everywhere and f0 > 0 somewhere. As ξ acts on g isometrically, f0 is constant along the orbits of ξ. Hence ξf0 = 0. As before, we compute that ∂ ∂t (HT − fS) ≥ F (HT − fS) + ∆(HT − fS) + S ( −∂f ∂t + ∆f − (C1 + C2)|f | ) . Now let f satisfy ∂f ∂t = ∆f − (C1 + C2)f with initial condition f(·, 0) = f0. Again, since ξf0 = 0 we have ξf ≡ 0. If we let f˜ = e(C1+C2)tf , then f˜ solves the heat equation ∂f˜ ∂t = ∆f˜ . 94 It is well known that a function f˜ satisfying the heat equation with nonnegative initial condition is positive for t > 0, see [32]. Thus f > 0 for t > 0 and ∂ ∂t (HT − fS) ≥ F (HT − fS) + ∆(HT − fS). By the definition of f0, we have H T − fS ≥ 0 when t = 0. Arguing as above, we can show that HT −fS ≥ 0 for t > 0. Since S > 0 and f > 0, it follows that HT > 0. 95 CHAPTER IV THE RICCI FLOW OF A SASAKI METRIC In this chapter we want to investigate the behavior of the Ricci flow when the initial metric is Sasaki. We are not talking about the Sasaki-Ricci flow as in [12], [19] or [36], which is just the transverse Ka¨hler-Ricci flow of a Sasaki manifold, but rather the Ricci flow of the actual Sasaki metric. The Ricci flow of a Ka¨hler metric preserves the Ka¨hler condition and this has been an extremely fruitful area of mathematical research since the 1980s. Since Sasaki geometry is closely related to Ka¨hler geometry, it seems reasonable to suspect that the Ricci flow of a Sasaki metric should be a tractable subject, but one quickly runs into trouble; the Ricci flow need not preserve the Sasaki condition. For example, consider a Sasaki-Einstein metric g0 with Einstein constant λ. The Ricci flow ∂g ∂t = −2Ricg with g(0) = g0 is given by g(t) = (1−2λt)g0. Since Sasaki metrics are not preserved by homothety, g(t) is not Sasaki for t > 0. However, a scaling of a Sasaki metric is a quasi-Sasaki metric. This gives us hope that the Ricci flow does preserve some sort of structure related to the Sasaki geometry. We begin with the following useful observation: Given a Sasaki structure (ξ, η,Φ, g), we can create a two-parameter family of quasi-Sasaki structures. Given A,B > 0, let ξB = 1√ B ξ, ηB = √ Bη and gA,B = ηB ⊗ ηB + AgT . Using the same (1,1)-tensor Φ, it is easy to see that the (almost) contact structure is preserved. A straighforward computation shows that the metric compatibility and normality conditions are also preserved. However, gA,B(X,ΦY ) 6= dηB(X, Y ). Rather we have gA,B(X,ΦY ) = A√ B dηB(X, Y ). Since Ka¨hler metrics are preserved by homothety, the transverse metric gTA,B = Ag T is Ka¨hler and thus (ξB, ηB,Φ, gA,B) defines a quasi- Sasaki structure. 96 Definition 4.0.3. A quasi-Sasaki structure obtained via the above construction will be called an (A,B)-deformation of a Sasaki structure. By choosing A = √ B, the deformed structure remains Sasaki. These type of deformations are known in the literature as transverse or D-homotheties. Note also that an (A,A)-deformation results in an honest scaling of the metric, but does not preserve the Sasaki structure unless A = 1. From proposition 2.1.1 we get the following corollary. Corollary 4.0.4. If (ξ, η,Φ, g) is an (A,B)-deformation of a Sasaki structure, then for X, Y ∈ D the Ricci tensor of g satisfies Ric(ξ, ξ) = 2n B A2 Ric(X, ξ) = 0 Ric(X, Y ) = RicT (X, Y )− 2 B A2 gT (X, Y ). 4.1. η-Einstein metrics It is difficult (as far as the author can tell) to say much about the Ricci flow of an arbitrary Sasaki metric. However, for a special class of Sasaki metrics known as η-Einstein, we can actually say quite a lot. It is these metrics whose Ricci flow we shall investigate. Definition 4.1.1. A Sasaki metric g is η-Einstein if there are real numbers µ and ν such that Ricci tensor of g satisfies Ricg = µg + νη ⊗ η. (4.1) 97 Recall that for a Sasaki structure (ξ, η,Φ, g), η ∧ (dη)n is a volume form, g(X,ΦY ) = dη(X, Y ) for all X, Y ∈ TM and for X, Y ∈ D the Ricci tensor satisfies Ric(ξ, ξ) = 2n, (4.2) Ric(X, ξ) = 0, (4.3) Ric(X, Y ) = RicT (X, Y )− 2gT (X, Y ), (4.4) From equations (4.1) and (4.2) we see that µ+ν = 2n. From equations (4.3) and (4.4) we see that a Sasaki manifold is η-Einstein if and only if the transverse metric is Ka¨hler-Einstein. Indeed, if RicT = λgT then Ric = (λ − 2)g + (2n + 2 − λ)η ⊗ η. Conversely, if Ric = µg + νη ⊗ η then RicT = (µ + 2)gT . We should also note that an η-Einstein manifold has constant scalar curvature. Tracing equation (4.1) we see that the scalar curvature R = (2n + 1)µ + ν = 2n(µ + 1). An excellent reference for the geometry of η-Einstein manifolds is the paper [6]. Let (ξ0, η0,Φ, g0) be η-Einstein. Then we can write g0 = η0 ⊗ η0 + gT0 where gT0 is transverse Ka¨hler-Einstein with RicT0 = λg T 0 . Now consider an (A,B)-deformation of this Sasaki structure. By corollary 4.0.4, for X, Y ∈ D, the Ricci tensor of the deformed metric satisfies Ric(ξ0, ξ0) = 2n B2 A2 , Ric(X, ξ0) = 0, Ric(X, Y ) = ( λ− 2B A ) gT0 (X, Y ). 98 Therefore, by the above equations, a metric of the form g(t) := A(t)gT0 +B(t)η0 ⊗ η0 (4.5) is the solution to the Ricci flow equation ∂g ∂t = −2Ricg, g(0) = g0, (4.6) with g0 an η-Einstein metric, if and only if A(t), B(t) > 0 satisfy dA dt = 4 B A − 2λ, A(0) = 1, (4.7) dB dt = −4nB 2 A2 , B(0) = 1. (4.8) By standard existence and uniqueness theory for ODEs (e.g. the Picard-Lindelo¨f theorem), the system (4.7), (4.8) has a unique solution that exists for at least a short time. Note that if we try to impose the additional restriction that A = √ B, then the system has no solution. Hence the metric cannot evolve by Sasaki metrics in this way. But, as we mentioned before, the metrics in (4.5) are quasi-Sasaki. For our qualitative analysis of these differential equations, we will make use of the conserved quantity Λ := ( λ 2(n+ 1) − B A ) B− n+1 n . (4.9) Indeed, a straightforward computation reveals that d dt Λ = 0. From the requirement that A(0) = B(0) = 1, we find that Λ = λ 2(n+1) − 1. 99 Since a Sasaki metric is η-Einstein if and only if the transverse metric is Ka¨hler- Einstein, we naturally have three cases to consider. In the following sections we investigate the Ricci flow equations (4.7) and (4.8) when the transverse Einstein constant λ is zero, positive and negative, respectively. We want to understand the behavior of the functions A and B in each case. Knowing their behavior, we can determine how the curvature is changing along the Ricci flow. Computing as in chapter II, the sectional curvature of a plane spanned by ξ and X ∈ D is κ(X, ξ) = B A2 κ0(X, ξ), (4.10) where κ0 denotes the sectional curvature with respect to the initial metric g0. The sectional curvature of a plane spanned by X, Y ∈ D is κ(X, Y ) = 1 A κT0 (X, Y ) +O ( B A2 ) . (4.11) 4.2. Transverse Calabi-Yau Structure When λ = 0 we have c1B = 0 and the transverse structure is Calabi-Yau. In this case we can solve the Ricci flow equations explicitly. We compute d dt ( B A2 ) = −4(n+ 2) ( B A2 )2 . (4.12) We solve this to get B A2 = 1 1 + 4(n+ 2)t . (4.13) Now we have dB dt = −4n ( B A2 ) B = −4n 1 + 4(n+ 2)t B. 100 This yields B(t) = (1 + 4(n+ 2)t) −n n+2 . Then from equation (4.13) we get A(t) = (1 + 4(n+ 2)t) 1 n+2 . Therefore, the flow exists for all t ∈ [0,∞) and A(t)→∞ while B(t)→ 0 as t→∞. In the limit, all sectional curvatures tend to zero and the metric degenerates to a flat metric on the transverse space. Example 4.2.1. For an example of an η-Einstein metric with transverse Calabi-Yau structure, we consider the Heisenberg group H = H(3,R). It is a nilpotent Lie group that can be identified algebraically as H =   1 x z 0 1 y 0 0 1  : x, y, z ∈ R  . We identify this Lie group topologically with R3 in the obvious way. We consider the group action to be from the right. Hence, the action of right multiplication by (a, b, c) is (x, y, z) 7→ (x + a, y + b, z + c + bx). The Lie algebra is generated by the right invariant vector fields ξ = 2 ∂ ∂z , Y = 2 ∂ ∂y and X = 2( ∂ ∂x + y ∂ ∂z ) which satisfy the bracket relations [ξ, Y ] = [ξ,X] = 0 and [Y,X] = 2ξ. Define a 1-form η := 1 2 (dz − ydx). Note that η is invariant under the group action. As η ∧ dη = 1 4 dx ∧ dy ∧ dz, we see that η is a contact form. We have D := ker(η) = {X, Y }. The complex structure J : D → D defined by JX = Y and 101 JY = −X gives (D, dη, J) a Ka¨hler structure. This is just the standard, flat, Ka¨hler structure on R2 = C. Define a metric g0 := η ⊗ η + 14(dx⊗ dx + dy ⊗ dy). Then g0 is invariant under the group action, {ξ, Y,X} is an orthonormal frame and (ξ, η,Φ, g0) gives H a Sasaki structure. Here Φ is just the extension of J such that Φξ = 0. Explicitly we have Φ = ∂ ∂y ⊗ dx − ( ∂ ∂x + y ∂ ∂z ) ⊗ dy. The nonzero components of the Ricci tensor of g0 are Ric(ξ, ξ) = 2 and Ric(Y, Y ) = Ric(X,X) = −2. Therefore Ricg0 = −2g0 + 4η⊗ η. Hence g0 is an η-Einstein metric. Under the Ricci flow, the curvature tends to zero and the metric degenerates to a flat metric on R2. Now define the compact quotient manifold N := H(3,R)/H(3,Z) where H(3,Z) denotes matrices of the above form with x, y, z ∈ Z. Then N is a non-trivial circle bundle over a complex torus and the Sasaki structure on H descends to a Sasaki structure on N . The S1 fiber corresponds to the integral curves of the Reeb vector field ξ = 2 ∂ ∂z . The Ricci flow on N behaves the same as the Ricci flow on H. Under the flow, the S1 fiber shrinks to a point and we have convergence to a flat torus. Remark 4.2.2. The behavior of the Ricci flow seen in this example is precisely the behavior of the Ricci flow on any Nil 3-manifold with an invariant metric. See, for example, chapter 1.6 of [11]. 4.3. Transverse Fano Structure When λ > 0 we have c1B > 0 and the transverse structure is Fano. In general, the flow only exists for finite time t ∈ [0, T ), T < ∞, and both A(t), B(t) → 0 as t→ T . We will consider three cases: 102 1. If λ > 2(n+ 1) then, from equation (4.9), Λ > 0. In this case we have B A = λ 2(n+ 1) − ΛB n+1n < λ 2(n+ 1) . Then dA dt = 4 B A − 2λ < 2λ n+ 1 − 2λ < 0. This shows that A(t)→ 0 in finite time. Since 0 < B < λ 2(n+1) A, it follows that B(t)→ 0 in finite time as well. 2. If λ = 2(n+ 1), then Λ = 0 and the solution is given by A(t) = B(t) = 1− 4nt, as is easily checked. In this case the initial metric is Sasaki-Einstein. Notice that the flow only exists for t ∈ [0, 1 4n ) and both A(t), B(t)→ 0 as t→ 1 4n . 3. If 0 < λ < 2(n+ 1), then Λ < 0. In this case we have B A = λ 2(n+ 1) − ΛB n+1n > λ 2(n+ 1) . Then dB dt = −4n ( B A )2 < −nλ2 (n+ 1)2 < 0. This implies that B(t) → 0 in finite time. Since 0 < A < 2(n+1) λ B, it follows that A(t)→ 0 in finite time as well. In this transverse Fano case, we see that all the sectionals curvatures tend to infinity as we approach the terminal time. This is to be expected (at least in one direction) of a Ricci flow with a finite-time singularity. It would be interesting to determine the terminal time explicitly in terms of λ and n in cases 1 and 3 above. 103 Example 4.3.1. The canonical example of an η-Einstein manifold with transverse Fano structure is the classical Hopf fibration. This gives S3 the structure of a principle circle bundle over CP 1 and the standard Sasaki structure on S3 (compare with example 1.3.5). Writing S3 = {(w, z) ∈ C2 : |w|2 + |z|2 = 1}, the Hopf fibration S1 ↪→ S3 → CP 1 is induced by the map pi : S3 → CP 1, where (w, z) 7→ [w : z]. Here [w : z] denotes the homogenous coordinates on CP 1. As [w : z] = [λw : λz] for λ ∈ C∗, it is clear that the fibers of pi are circles S1 = {λ ∈ C : |λ| = 1}. As we saw in chapter II, the collection of principal circle bundles over CP 1 has a group structure isomorphic to H2(CP 1,Z) ' Z with generator 1 2pi [ωFS]. The Fubini- Study metric has constant holomorphic sectional curvature equal to 4. As in the proof of theorem 2.2.6, there is a contact form η on S3 such that 1 2 dη = pi∗ωFS and g := η ⊗ η + pi∗ωFS(J ·, ·) defines a Sasaki metric. The metric g is in fact the round metric of constant sectional curvature 1. So g is an Einstein (hence η-Einstein) metric with Ric = 2g. The Ricci flow on S3 with initial metric g is g(t) := (1− 4t)g. The flow exists for t ∈ [0, 1 4 ) and S3 shrinks to a point at the terminal time T = 1 4 . As t→ T the sectional curvatures blows up. Now fix ε ∈ (0, 1) and define a Sasaki metric gε via a D-homothetic deformation of g. That is, gε := εg + (ε 2 − ε)η ⊗ η. Then g¯ := ε−1gε = εη ⊗ η + pi∗ωFS(J ·, ·) and (S3, g¯) is the classical ε-collapsed Berger sphere. Let g¯(t) be the Ricci flow with initial metric g¯. This flow is well-understood (see chapter 1.5 of [11]). The flow exists on a maximal finite time interval [0, T ) and the metric becomes asymptotically round as t → T . Thus, gε(ε−1t) = εg¯(ε−1t) is the Ricci flow on [0, εT ) with intitial metric gε and it behaves the same as the flow g¯(t). 104 4.4. Transverse Canonical Structure When λ < 0 we have c1B < 0 and we say that the transverse structure is canonical. In this case the flow exists for all t ≥ 0, A(t) → ∞ as t → ∞ and B(t) tends to a constant c(n, λ) ∈ (0, 1) as t→∞. Since dB dt = −4nB2 A2 < 0 and B(0) = 1, we have 0 < B(t) ≤ 1. From (4.8) and (4.9), dB dt = −4n ( λ 2(n+ 1) − ΛB n+1n )2 . (4.14) Thus dB dt is uniformly bounded. Hence, the right hand side of (4.14) and its first derivative are uniformly bounded. By a standard result in ODEs (see [23]), this implies that (4.14) with the initial condition B(0) = 1 has a unique solution defined for all t ≥ 0. Then it follows from (4.9) that equation (4.7) has a unique solution defined for all t ≥ 0. Since we must have A,B > 0, right away we see that dA dt = 4B A − 2λ ≥ −2λ > 0. Thus A(t) ≥ 1− 2λt. Hence A(t)→∞ as t→∞. Now, from equation (4.9) we have B A = λ 2(n+ 1) − ΛB n+1n , and it follows that lim t→∞ B(t) = ( λ 2Λ(n+ 1) ) n n+1 = ( λ λ− 2(n+ 1) ) n n+1 . As in the λ = 0 case, as t → ∞ all of the sectional curvatures tend to zero and the metric degenerates to a flat metric on the transverse space. 105 Example 4.4.1. For an example of an η-Einstein metric with transvere canonical structure, we consider M = B × R where B = {z ∈ C : |z| < 1}. Let ω be the Bergman metric on B. We take ω = 2i∂∂¯ log(1− |z|2) = 2i dz ∧ dz¯ (1− |z|2)2 . Then (B,ω, J) is Ka¨hler with constant, negative, holomorphic sectional curvature. Here J is the usual complex structure on C. Since ω is a real, closed 2-form on B, there is some real 1-form α such that ω = dα. In particular, α = i zdz¯ − z¯dz (1− |z|2) . Let pi : M → B be the projection. Let t be the coordinate on R and define 1 2 η := pi∗α + dt. Then 1 2 dη = pi∗ω and we see that η is a contact form. The Reeb vector field ξ = 1 2 ∂ ∂t . We take Φ = p˜i ◦ J ◦ pi∗. Now define a metric g0 := pi ∗ω(·, J ·) + η ⊗ η = 1 2 dη(·,Φ·) + η ⊗ η. Then (ξ, η,Φ, g0) is a Sasaki structure on M . In real coordinates (x, y), we have α = 2 xdy − ydx 1− x2 − y2 and ω = 4 dx ∧ dy (1− x2 − y2)2 . Define vector fields X := 1− x2 − y2 2 ( ∂ ∂x − η ( ∂ ∂x ) ξ ) and Y := 1− x2 − y2 2 ( ∂ ∂y − η ( ∂ ∂y ) ξ ) . 106 Then D = span{X, Y } and {ξ,X, Y } is an orthonormal frame for g. Working with this frame, we compute the nonzero components of the Ricci tensor to be Ric(ξ, ξ) = 2 and Ric(X,X) = Ric(Y, Y ) = −3. Thus Ric = −3g + 5η ⊗ η. Hence, g is η-Einstein. If we consider a metric of the form g = Api∗ω(·, J ·) + Bη ⊗ η, then we compute that, with respect to the frame {ξ,X, Y }, the nonzero components of the Ricci tensor of g are Ric(ξ, ξ) = 2B 2 A2 and Ric(X,X) = Ric(Y, Y ) = −1 − 2B A . Thus, we will have the Ricci flow with inital metric g0 if we can solve dA dt = 2 + 4 B A , A(0) = 1, dB dt = −4B 2 A2 , B(0) = 1. From the above we know that the solution exists for all t ≥ 0 and as t → ∞, A(t)→∞, B(t)→ √ 5 5 and the sectional curvatures all tend to zero. Remark 4.4.2. There is a classification of 3-dimensional Sasaki manifolds due to Geiges. The three examples we have presented in this chapter are the universal covers of the geometries given in the classification. Our examples easily generalize to higher dimensions, but in higher dimensions there is no such classification. In dimension 3, the basic first Chern class is always signed or null. As one might expect, there is a correspondence between the type of Sasaki structure (positive, negative or null c1B) and the three geometries given in Geiges’ classification. A well-known theorem of Belgun makes this precise. See [6] for more information about this. 107 4.5. Rigidity To have a Ricci flow of the form (4.5), the initial Sasaki metric g0 must have very special geometry. Indeed, we can prove the following: Proposition 4.5.1. Let (ξ, η0,Φ, g0) be a Sasaki structure. Suppose that the Ricci flow with initial metric g0 is given by g(t) = A(t)g T 0 + B(t)η0 ⊗ η0. Then g0 is an η-Einstein metric. Proof. As we saw before, the metric g(t) provides a quasi-Sasaki structure and we have Rict(ξ, ξ) = 2n B2 A2 , Rict|D = RicTt − 2BAgT0 and Rict(X, ξ) = 0 for all X ∈ D. Since gT (t) = A(t)gT0 is just a scaling of g T 0 , we have Ric T t = Ric T 0 . Hence, A satisfies dA dt gT0 (X, Y ) = 4 B A gT0 (X, Y )− 2RicT0 (X, Y ) for all X, Y ∈ D. Rewrite the above as 2RicT0 (X, Y ) = ( 4 B A − dA dt ) gT0 (X, Y ). Since the left-hand side is independent of time, 4B A − dA dt must be a constant, call it 2δ. Then RicT0 = δg T 0 . Thus, the initial transverse metric g T 0 is Ka¨hler-Einstein. Therefore, the initial Sasaki metric is η-Einstein. 4.6. Questions and Directions for Future Research We have merely scratched the surface of the investigation into the Ricci flow of a Sasaki metric and there are many questions still to be answered. We list just a few of them below. 108 1. The Ricci flow with an initial Sasaki metric is not Sasaki for t > 0, but is it quasi-Sasaki in general? 2. The author can show that, in general, the length of ξ with respect to the evolving metric is decreasing. If the Ricci flow has a finite time singularity, does the length of ξ going to zero characterize the singular time? 3. Is there a relationship between the Ricci flow of a Sasaki metric and the transverse Ka¨hler-Ricci flow of a Sasaki metric? 109 CHAPTER V APPENDIX In chapter II we proved that under certain conditions a quasi-Sasaki-Einstein manifold has a local Riemannian product structure. The proof of theorem 2.6.3 made use of an integrable, in fact parallel, almost product structure. We decided to include this short section here to review the necessary terminology and some criteria for integrability. Definition 5.0.1. An almost product structure on a smooth manifold M is a (1,1)-tensor F (i.e. an endomorphism of the tangent bundle) such that F 2 = I. If we let P = 1 2 (I + F ) and Q = 1 2 (I − F ), then P 2 = P, Q2 = Q and PQ = QP = 0. (5.1) Conversely, given (1,1)-tensors P and Q satisfying (5.1), setting F = P − Q gives an almost product structure. The images of P and Q define complementary distributions P and Q of TM . We will assume that P and Q have constant rank p and q, respectively. That is, the dimension of the corresponding distribution is constant for all x ∈ M . On the other hand, given two distributions P and Q such that TM = P ⊕Q, setting P and Q to be their respective projections, we see that P and Q satisfy (5.1). Definition 5.0.2. Let D be a k-dimensional distribution on a smooth m-dimensional manifold M . We say that D is involutive if for any X, Y ∈ D, the Lie bracket [X, Y ] ∈ D. 110 We say that D is integrable if for each point x ∈ M there is a k-dimensional submanifold N containing x such that TxN = Dx. We call N an integral submanifold of D (through x). We say that D is completely integrable if for each point x ∈ M , there is a coordinate neighborhood (U, x1, . . . , xm) of x such that {∂x1 , . . . , ∂xk} is a local frame for D. Setting xi = ci for i = k + 1, . . . ,m, and some constants ci gives an integral submanifold of D. The famous theorem of Frobenius says that these three notions are equivalent. Theorem 5.0.3 (Frobenius). A distribution is involutive if and only if it is integrable if and only if it is completely integrable. Definition 5.0.4. An almost product structure F on M is integrable if the distributions induced by P and Q are integrable in the sense of Frobenius. In this case we say that M is a locally product manifold. So if F = P − Q is an integrable almost product structure on M , then by the Frobenius theorem, for each point x ∈ M , there is a coordinate neighborhood (U, x1, . . . , xm) such that {∂x1 , . . . , ∂xp} is a local frame for P and {∂xp+1 , . . . , ∂xm} is a local frame for Q. Setting xp+1, . . . , xm constant gives an integral submanifold of P and setting x1, . . . , xp constant gives an integral submanifold of Q. Thus M is locally the product of two manifolds whose tangent spaces are (isomorphic to) P and Q. By proposition 3.1.4 in [13], the integrability of the almost product structure F is equivalent to the vanishing of the Nijenhuis tensor NF defined by (1.2). Furthermore, since NI = 0, one can check that NF = 2NP = −2NQ. Therefore: Proposition 5.0.5. An almost product structure F = P −Q is integrable if and only if NF = 0 if and only if NP = 0 if and only if NQ = 0. 111 Now let g be a Riemannian metric and ∇ the Levi-Civita connection of g. Recall that ∇g = 0 and ∇XY −∇YX = [X, Y ]. If T is a parallel (1,1)-tensor (i.e. ∇T = 0), then it is straightforward to check that the Nijenhuis tensor NT = 0. It is also easy to check that the image of a parallel (1,1)-tensor is an involutive distribution. Since ∇F = 2∇P = −2∇Q, we see that if any of F , P or Q is parallel then the almost product structure is integrable, but in this case even more is true. Definition 5.0.6. Let (M, g) be a locally product manifold with TM = P ⊕ Q an orthogonal decomposition. We say that (M, g) is a locally decomposable Riemannian manifold if for any U, V ∈ P and X, Y ∈ Q we have Xg(U, V ) = 0 and Ug(X, Y ) = 0. In this case the metric splits as an honest product metric corresponding to the locally product structure. It is known by work of Yano that a necessary and sufficient condition for a locally product manifold to be a locally decomposable Riemannian manifold is ∇F = 0. Therefore we have the following: Proposition 5.0.7. Let F = P −Q be an almost product structure on (M, g). Then (M, g) is a locally decomposable Riemannian manifold if and only if ∇F = 0 if and only if ∇P = 0 if and only if ∇Q = 0 if and only if P and Q are parallel distributions. Remark 5.0.8. We remark that if F is an integrable almost product structure, then there exists a torsion-free, affine connection (not necessarily the Levi-Civita connection) with respect to which F is parallel. See the book [13] for reference. In this section we have collected the technical lemmas and propositions used in chapter III. We begin with 112 Lemma 5.0.9. Let x ∈M and V (r) := Vol(BT (x, r)). If Fξ is quasi-regular, then lim r→0 V (r) V (r/2) = 4n. Proof. Recall that pi : M → Z is a principle S1-orbibundle over the orbifold Z = M/Fξ. Pick a point x ∈ M and choose r > 0 small enough so that BT (x, r) is a trivial S1 bundle over the geodesic ball pi(BT (x, r)) ⊂ Z. Since the set of orbifold singularities in Z has measure zero, it does not contribute anything when we compute volumes. Thus we may assume that pi(BT (x, r)) contains only smooth points. For smooth points in Z, all S1 fibers have the same length, call it `. On BT (x, r) the metric and volume form can be written as g = η ⊗ η + pi∗h, dV = η ∧ (pi∗ωh)n, where h is a Ka¨hler metric on pi(BT (x, r)) and ωh is the Ka¨hler form. Hence V (r) = ∫ BTg (x,r) dV = ∫ S1×pi(BTg (x,r)) η ∧ (pi∗ωh)n = ` · Vol(pi(BTg (x, r)). The result follows from this and the fact that for any geodesic ball B(r) ⊂ Z of radius r, lim r→0 Vol(B(r)) Vol(B(r/2)) = 22n = 4n. 113 Proposition 5.0.10. Let M2n+1 be a compact quasi-Sasaki manifold without boundary. Let α and β be basic forms with deg(α) + deg(β) = 2n− 1. Then ∫ M dα ∧ β ∧ η + (−1)deg(α) ∫ M α ∧ dβ ∧ η = 0. Proof. We compute that d(α ∧ β ∧ η) = dα ∧ β ∧ η + (−1)deg(α)α ∧ dβ ∧ η − α ∧ β ∧ dη. The form α ∧ β ∧ dη is basic and of degree 2n + 1, thus α ∧ β ∧ dη = 0. The result now follows from Stokes’ theorem. Corollary 5.0.11. Let α be a basic (2n− 1)-form on a closed quasi-Sasaki manifold M2n+1. Then ∫ M dBα ∧ η = 0. Proof. Take β = 1 in the above proposition. Lemma 5.0.12. For any bundle-like metric g and τ > 0, we have µT (g, τ) > −∞. Proof. Letting w := τ−n/2e−f/2, we can write the WT -functional as WT (g, f, τ) = ∫ M τ(RTw2 + 4|∇w|2)− 2(log(w) + n log(τ))w2 dV. For a constant c > 0, it is not hard to see that WT (cgT , f, cτ) =WT (gT , f, τ). Hence we may assume without loss of generality that τ = 1. Thus we need to show that there is a constant C = C(g) such that ∫ M RTw2 + 4|∇w|2 − 2w2 log(w) dV ≥ C. 114 Since infx∈M RT > −∞ and ∫ M w2 dV = 1, we only have to show that ∫ M 4|∇w|2 − 2w2 log(w) dV ≥ C. This follows from the log Sobolev inequality (see, for example, lemma 6.36 in [9]). Lemma 5.0.13. Suppose there is a sequence rk > 0 and a sequence of points and times (pk, tk) ∈M × [0,∞) with tk →∞ such that r−2nk Vk(rk)→ 0 as k →∞, where Vk(rk) = Vol(B T g(tk) (pk, rk)). Then we may assume that Vk(rk) ≤ 5nVk(rk/2) Proof. If it is not the case that Vk(rk) ≤ 5nVk(rk/2), then let rik = 2−irk for integers i > 0. By lemma 5.0.9, we can choose i0 to be the least integer such that Vk(r i0 k ) ≤ 5nVk(r i0 k /2). So for i < i0 we have Vk(r i k) > 5 nVk(r i k/2) = 5 nVk(r i+1 k ). Iterating this inequality, we get Vk(r i0 k ) < 5 −i0nVk(rk). Therefore, as k →∞, (ri0k ) −2nVk(r i0 k ) < ( 4 5 )i0n r−2nk Vk(rk)→ 0. Thus we can replace rk with r i0 k and we have Vk(r i0 k ) ≤ 5nVk(ri0k /2). Lemma 5.0.14. For any ε > 0 there are k1 < k2 such that if the transverse diameter is sufficiently large then 1. Vol(Bξ(k1, k2)) < ε 2. Vol(Bξ(k1, k2)) ≤ 210nVol(Bξ(k1 + 2, k2 − 2)) 115 3. There exists r1 ∈ [k1, k1 + 1], r2 ∈ [k2 − 1, k2] and a uniform constant C such that ∫ Bξ(r1,r2) RT dV ≤ CVol(Bξ(k1, k2)). Proof. Let ε > 0 be given. As the manifold M is closed and ∂tdVt = ∆ Tu dVt = ∆u dVt, the volume of M is constant along the flow. So if the transverse diameter becomes sufficiently large, then there is K big enough so that for all k2 > k1 ≥ K we have Vol(Bξ(k1, k2)) < ε. Take k2 > k1 ≥ K. If (2) holds then we are done. Suppose that (2) does not hold. We then consider the annulus Bξ(k1 + 2, k2 − 2) and check to see if (2) holds now. If it does then we are done, otherwise we repeat this step. Suppose that after repeating this step m times, we are still unable to find k1 and k2 so that (1) and (2) hold. Then we would have Vol(Bξ(k1, k2)) > 2 10nmVol(Bξ(k1 + 2m, k2 − 2m)). We can assume that k2−k1−4m is very close to 1. Then 2m is close to 12(k2−k1−1). If we choose k  1 and set k1 = k/2 and k2 = 3k/2, then m is close to k/4, k1 + 2m is close to k and k2 − 2m is close to k + 1. The length of ξ is constant along the flow. Hence, the lengths of the integral curves of ξ (which are closed geodesics) are constant along the flow. Thus they are uniformly bounded above by some ` > 0. So for any p, q ∈ M we have dg(t)(p, q) ≤ dTg(t)(p, q) + `. This, along with proposition 3.3.7, gives R T ≤ C22k on Bξ(k, k + 1) for some uniform constant C. The annulus Bξ(k, k + 1) contains at least 2 2k disjoint 116 transverse balls of radius 2−k. Then by theorem 3.4.1 we have Vol(Bξ(k, k + 1)) ≥ 22k∑ i=1 Vol(BTg(t)(2 −k)) ≥ C22k2−2kn. Combining this with the above we find that ε > Vol(Bξ(k1, k2)) > 2 10nmVol(Bξ(k, k + 1)) ≥ C210nm22k2−2kn. If k is large enough, then 10m > 2k and the above is a contradiction. Therefore (1) and (2) hold. To prove (3) we first define the transverse sphere STg (x, r) := {y ∈M : dTg (x, y) = r}. Then d dr Vol(BTg (x, r)) = Vol(S T g (x, r)). Given k1  k2 such that (1) and (2) hold, there is r1 ∈ [k1, k1 + 1] such that Vol(STg(t)(x, 2 r1)) ≤ Vol(Bξ(k1, k2)) 2k1−1 . If not, then we would have Vol(Bξ(k1, k1 + 1)) = ∫ 2k1+1 2k1 Vol(ST (x, r)) dr > 2Vol(Bξ(k1, k2)). This is a contradiction since k1  k2. Similarly we can prove there is r2 ∈ [k2− 1, k2] such that Vol(STg(t)(x, 2 r2)) ≤ Vol(Bξ(k1, k2)) 2k2−1 . 117 Using these volume estimates for the spheres and proposition 3.3.7, we have ∫ Bξ(r1,r2) RT − n dV = ∫ Bξ(r1,r2) −∆u dV = ∫ ST (x,2r1 ) |∇u| dV + ∫ ST (x,2r2 ) |∇u| dV ≤ Vol(Bξ(k1, k2)) 2k1−1 C2k1+1 + Vol(Bξ(k1, k2)) 2k2−1 C2k2 = CVol(Bξ(k1, k2)). 118 REFERENCES CITED [1] T. Aubin, E´quations du type Monge-Ampe`re sur les varie´te´s ka¨hle´riennes compactes, Bull. Sci. Math. (2) 102 (1) (1978) 63–95. [2] S. Bando, On the classification of three-dimensional compact Kaehler manifolds of nonnegative bisectional curvature, J. Differential Geom. 19 (2) (1984) 283–297. [3] D. E. Blair, The theory of quasi-Sasakian structures, J. Differential Geometry 1 (1967) 331–345. [4] D. E. Blair, Riemannian geometry of contact and symplectic manifolds, Vol. 203 of Progress in Mathematics, Birkha¨user Boston Inc., Boston, MA, 2002. [5] C. P. Boyer, K. Galicki, Sasakian geometry, Oxford Mathematical Monographs, Oxford University Press, Oxford, 2008. [6] C. P. Boyer, K. Galicki, P. Matzeu, On eta-Einstein Sasakian geometry, Comm. Math. Phys. 262 (1) (2006) 177–208. [7] H. D. Cao, Deformation of Ka¨hler metrics to Ka¨hler-Einstein metrics on compact Ka¨hler manifolds, Invent. Math. 81 (2) (1985) 359–372. [8] A. Chau, L.-F. Tam, On quadratic orthogonal bisectional curvature, J. Differential Geom. 92 (2) (2012) 187–200. [9] B. Chow, S.-C. Chu, D. Glickenstein, C. Guenther, J. Isenberg, T. Ivey, D. Knopf, P. Lu, F. Luo, L. Ni, The Ricci flow: techniques and applications. Part I, Vol. 135 of Mathematical Surveys and Monographs, American Mathematical Society, Providence, RI, 2007, geometric aspects. [10] B. Chow, D. Knopf, The Ricci flow: an introduction, Vol. 110 of Mathematical Surveys and Monographs, American Mathematical Society, Providence, RI, 2004. [11] B. Chow, D. Knopf, The Ricci flow: an introduction, Vol. 110 of Mathematical Surveys and Monographs, American Mathematical Society, Providence, RI, 2004. [12] T. C. Collins, The transverse entropy functional and the Sasaki-Ricci flow, Trans. Amer. Math. Soc. 365 (3) (2013) 1277–1303. [13] M. de Leo´n, P. R. Rodrigues, Methods of differential geometry in analytical mechanics, Vol. 158 of North-Holland Mathematics Studies, North-Holland Publishing Co., Amsterdam, 1989. 119 [14] A. El Kacimi-Alaoui, G. Hector, De´composition de Hodge basique pour un feuilletage riemannien, Ann. Inst. Fourier (Grenoble) 36 (3) (1986) 207–227. [15] A. El Kacimi-Alaoui, Ope´rateurs transversalement elliptiques sur un feuilletage riemannien et applications, Compositio Math. 73 (1) (1990) 57–106. [16] A. Futaki, An obstruction to the existence of Einstein Ka¨hler metrics, Invent. Math. 73 (3) (1983) 437–443. [17] A. Futaki, H. Ono, G. Wang, Transverse Ka¨hler geometry of Sasaki manifolds and toric Sasaki-Einstein manifolds, J. Differential Geom. 83 (3) (2009) 585–635. [18] S. I. Goldberg, S. Kobayashi, Holomorphic bisectional curvature, J. Differential Geometry 1 (1967) 225–233. [19] W. He, The Sasaki-Ricci flow and compact Sasaki manifolds of positive transverse holomorphic bisectional curvature, J. Geom. Anal. 23 (4) (2013) 1876–1931. [20] W. He, S. Sun, The generalized Frankel conjecture in Sasaki geometry, Int. Math. Res. Not. IMRN (1) (2015) 99–118. [21] W. He, S. Sun, Frankel conjecture and Sasaki geometry, Adv. Math. 291 (2016) 912–960. [22] A. Howard, B. Smyth, H. Wu, On compact Ka¨hler manifolds of nonnegative bisectional curvature. I, Acta Math. 147 (1-2) (1981) 51–56. [23] W. Hurewicz, Lectures on Ordinary Differential Equations, Dover Phoenix editions. [24] F. W. Kamber, P. Tondeur, de Rham-Hodge theory for Riemannian foliations, Math. Ann. 277 (3) (1987) 415–431. [25] B. Kleiner, J. Lott, Notes on Perelman’s papers, Geom. Topol. 12 (5) (2008) 2587–2855. [26] B. Kleiner, J. Lott, Geometrization of three-dimensional orbifolds via Ricci flow, Aste´risque (365) (2014) 101–177. [27] S. Kobayashi, On compact Ka¨hler manifolds with positive definite Ricci tensor, Ann. of Math. (2) 74 (1961) 570–574. [28] S. Kobayashi, Topology of positively pinched Kaehler manifolds, Toˆhoku Math. J. (2) 15 (1963) 121–139. 120 [29] N. Mok, The uniformization theorem for compact Ka¨hler manifolds of nonnegative holomorphic bisectional curvature, J. Differential Geom. 27 (2) (1988) 179–214. [30] P. Molino, Riemannian foliations, Vol. 73 of Progress in Mathematics, Birkha¨user Boston, Inc., Boston, MA, 1988, translated from the French by Grant Cairns, With appendices by Cairns, Y. Carrie`re, E´. Ghys, E. Salem and V. Sergiescu. [31] G. Perelman, The entropy formula for the Ricci flow and its geometric applications, ArXiv Mathematics e-printsarXiv:math/0211159. [32] M. H. Protter, H. F. Weinberger, Maximum principles in differential equations, Springer-Verlag, New York, 1984, corrected reprint of the 1967 original. [33] N. Sesum, Convergence of a Ka¨hler-Ricci flow, Math. Res. Lett. 12 (5-6) (2005) 623–632. [34] N. Sesum, G. Tian, Bounding scalar curvature and diameter along the Ka¨hler Ricci flow (after Perelman), J. Inst. Math. Jussieu 7 (3) (2008) 575–587. [35] Y. T. Siu, S. T. Yau, Compact Ka¨hler manifolds of positive bisectional curvature, Invent. Math. 59 (2) (1980) 189–204. [36] K. Smoczyk, G. Wang, Y. Zhang, The Sasaki-Ricci flow, Internat. J. Math. 21 (7) (2010) 951–969. [37] S. Tanno, Quasi-Sasakian structures of rank 2p+ 1, J. Differential Geometry 5 (1971) 317–324. [38] G. Tian, X. Zhu, Convergence of Ka¨hler-Ricci flow, J. Amer. Math. Soc. 20 (3) (2007) 675–699. [39] P. Tondeur, Geometry of foliations, Vol. 90 of Monographs in Mathematics, Birkha¨user Verlag, Basel, 1997. [40] J. Wang, Einstein metrics on principal circle bundles, Differential Geom. Appl. 7 (4) (1997) 377–387. [41] M. Y. Wang, W. Ziller, Einstein metrics on principal torus bundles, J. Differential Geom. 31 (1) (1990) 215–248. [42] H. Wu, On compact Ka¨hler manifolds of nonnegative bisectional curvature. II, Acta Math. 147 (1-2) (1981) 57–70. [43] S. T. Yau, On the Ricci curvature of a compact Ka¨hler manifold and the complex Monge-Ampe`re equation. I, Comm. Pure Appl. Math. 31 (3) (1978) 339–411. 121