............ .....,. .....
MonoanPM Volume 135
The Ricci F·-low: Techniques and Applications Part I: Geometric Aspects Bennett Chow Sun-Chin Chu David Glickenstein Christine Guenther James Isenberg Tom Ivey Dan Knopf Peng Lu Feng Luo Lei Ni
•
American Mathematical Society
Mathematical Surveys and Monographs Volume 135
The Ricci Flow: Techniques and Applications Part I: Geometric Aspects
Bennett Chow Sun-Chin Chu David Glickenstein Christine Guenther James Isenberg Tom Ivey Dan Knopf Peng Lu Feng Luo Lei Ni
American Mathematical Society
EDITORIAL COMMITTEE Jerry L. Bona Michael G. Eastwood
Peter S. Landweber Michael P. Loss
J. T. Stafford, Chair 2000 Mathematics Subject Classification. Primary 53C44, 53C25, 58J35, 35K55, 35K05.
For additional information and updates on this book, visit www.ams.org/bookpages/surv-135
ISBN 978-0-8218-3946-1
Copying and reprinting. Individual readers of this publication, and nonprofit libraries acting for them, are permitted to make fair use of the material, such as to copy a chapter for use in teaching or research. Permission is granted to quote brief passages from this publication in reviews, provided the customary acknowledgment of the source is given. Republication, systematic copying, or multiple reproduction of any material in this publication is permitted only under license from the American Mathematical Society. Requests for such permission should be addressed to the Acquisitions Department, American Mathematical Society, 201 Charles Street, Providence, Rhode Island 02904-2294, USA. Requests can also be made by e-mail to reprint-permissionlilams. org.
© 2007 by Bennett
Chow. All rights reserved. Printed in the United States of America.
§
The paper used in this book is acid-free and falls within the guidelines established to ensure permanence and durability. Visit the AMS home page at http://www.ams.org/ 10 9 8 7 6 5 4 3 2 1
12 11 10 09 08 07
Contents Preface What this book is about Highlights of Part I Acknowledgments
xiii
Contents of Part I of Volume Two
xvii
IX
ix Xl
Chapter 1. Ricci Solitons 1. General solitons and their canonical forms 2. Differentiating the soliton equation - local and global analysis 3. Warped products and 2-dimensional solitons 4. Constructing the Bryant steady soliton 5. Rotationally symmetric expanding solitons 6. Homogeneous expanding solitons 7. When breathers and solitons are Einstein 8. Perelman's energy and entropy in relation to Ricci solitons 9. Buscher duality transformation of warped product solitons 10. Summary of results and open problems on Ricci solitons 11. Notes and commentary
1 2 6 11 17 26 32 41 44 46 50 52
Chapter 2. Kahler-Ricci Flow and Kahler-Ricci Solitons 1. Introduction to Kahler manifolds 2. Connection, curvature, and covariant differentiation 3. Existence of Kahler-Einstein metrics 4. Introduction to the Kahler-Ricci flow 5. Existence and convergence of the Kahler-Ricci flow 6. Survey of some results for the Kahler-Ricci flow 7. Examples of Kahler-Ricci solitons 8. Kahler-Ricci flow with nonnegative bisectional curvature 9. Matrix differential Harnack estimate for the Kahler-Ricci flow 10. Linear and interpolated differential Harnack estimates 11. Notes and commentary
55 55 62 70 74 80 95 97 103 109 118 124
Chapter 3. The Compactness Theorem for Ricci Flow 1. Introduction and statements of the compactness theorems 2. Convergence at all times from convergence at one time 3. Extensions of Hamilton's compactness theorem
127 127 132 138
v
CONTENTS
vi
4. 5.
Applications of Hamilton's compactness theorem Notes and commentary
142 148
Chapter 4. Proof of the Compactness Theorem 149 1. Outline of the proof 149 2. Approximate isometries, compactness of maps, and direct limits 150 3. Construction of good coverings by balls 158 4. The limit manifold (M~, 900) 165 5. Center of mass and nonlinear averages 175 6. Notes and commentary 187 Chapter 5. Energy, Monotonicity, and Breathers 1. Energy, its first variation, and the gradient flow 2. Monotonicity of energy for the Ricci flow 3. Steady and expanding breather solutions revisited 4. Classical entropy and Perelman's energy 5. Notes and commentary
189 190 197 203 214 219
Chapter 6. Entropy and No Local Collapsing 1. The entropy functional Wand its monotonicity 2. The functionals jJ, and 1/ 3. Shrinking breathers are shrinking gradient Ricci solitons 4. Logarithmic Sobolev inequality 5. No finite time local collapsing: A proof of Hamilton's little loop conjecture 6. Improved version of no local collapsing and diameter control 7. Some further calculations related to F and W 8. Notes and commentary
221 221 235 242 246
Chapter 7. The Reduced Distance 1. The C-Iength and distance for a static metric 2. The C-Iength and the L-distance 3. The first variation of C-Iength and existence of C-geodesics 4. The gradient and time-derivative of the L-distance function 5. The second variation formula for C and the Hessian of L 6. Equations and inequalities satisfied by Land f. 7. The f.-function on Einstein solutions and Ricci solitons 8. C-Jacobi fields and the C-exponential map 9. Weak solution formulation 10. Notes and commentary
285 286 288 296 306 312 322 335 345 363 379
251 264 273 284
Chapter 8. Applications of the Reduced Distance 381 1. Reduced volume of a static metric 381 2. Reduced volume for Ricci flow 386 3. A weakened no local collapsing theorem via the monotonicity of the reduced volume 399 4. Backward limit of ancient K-solution is a shrinker 406
CONTENTS
5.
vii
Perelman's Riemannian formalism in potentially infinite dimensions Notes and commentary
417 432
Chapter 9. Basic Topology of 3-Manifolds 1. Essential 2-spheres and irreducible 3-manifolds 2. Incompressible surfaces and the geometrization conjecture 3. Decomposition theorems and the Ricci flow 4. Notes and commentary
433 433 435 439 442
Appendix A. Basic Ricci Flow Theory 1. Riemannian geometry 2. Basic Ricci flow 3. Basic singularity theory for Ricci flow 4. More Ricci flow theory and ancient solutions 5. Classical singularity theory
445 445 456 465 470 474
Appendix B. Other Aspects of Ricci Flow and Related Flows 1. Convergence to Ricci solitons 2. The mean curvature flow 3. The cross curvature flow 4. Notes and commentary
477 477 482 490 500
Appendix C.
501
6.
Glossary
Bibliography
513
Index
531
Preface Trinity: It's the question that drives us, Neo. It's the question that brought you here. You know the question just as I did. Neo: What is the Matrix? ... Morpheus: Do you want to know what it is? The Matrix is everywhere .... Unfortunately, no one can be told what the Matrix is. You have to see it for yourself. - From the movie "The Matrix".
What this book is about This is the sequel to the book "The Ricci Flow: An Introduction" by two of the authors [108]. In the previous volume (henceforth referred to as Volume One) we laid some of the foundations for the study of Richard Hamilton's Ricci flow. The Ricci flow is an evolution equation which deforms Riemannian metrics by evolving them in the direction of minus the Ricci tensor. It is like a heat equation and tries to smooth out initial metrics. In some cases one can exhibit global existence and convergence of the Ricci flow. A striking example of this is the main result presented in Chapter 6 of Volume One: Hamilton's topological classification of closed 3-manifolds with positive Ricci curvature as spherical space forms. The idea of the proof is to show, for any initial metric with positive Ricci curvature, the normalized Ricci flow exists for all time and converges to a constant curvature metric as time approaches infinity. Note that on any closed 2-dimensional manifold, the normalized Ricci flow exists for all time and converges to a constant curvature metric. Many of the techniques used in Hamilton's original work in dimension 2 have influenced the study of the Ricci flow in higher dimensions. In this respect, of special note is Hamilton's 'meta-principle' of considering geometric quantities which either vanish or are constant on gmdient Ricci solitons. It is perhaps generally believed that the Ricci flow tries to make metrics more homogeneous and isotropic. However, for general initial metrics on closed manifolds, singularities may develop under the Ricci flow in dimensions as low as 3. 1 In Volume One we began to set up the study of singularities by discussing curvature and derivative of curvature estimates, looking at how generally dilations are done in all dimensions, and studying 1For noncompact manifolds, finite time singularities may even occur in dimension 2. ix
x
PREFACE
aspects of singularity formation in dimension 3. In this volume, we continue the study of the fundamental properties of the Ricci flow with particular emphasis on their application to the study of singularities. We pay particu-
lar attention to dimension 3, where we describe some aspects of Hamilton's and Perelman's nearly complete classification of the possible singularities. 2 As we saw in Volume One, Ricci solitons (Le., self-similar solutions), differential Harnack inequalities, derivative estimates, compactness theorems, maximum principles, and injectivity radius estimates play an important role in the study of the Ricci flow. The maximum principle was used extensively in the 3-dimensional results we presented. Some of the other techniques were presented only in the context of the Ricci flow on surfaces. In this volume we take a more detailed look at these general topics and also describe some of the fundamental new tools of Perelman which almost complete Hamilton's partial classification of singularities in dimension 3. In particular, we discuss Perelman's energy, entropy, reduced distance, and some applications. Much of Perelman's work is independent of dimension and leads to a new understanding of singularities. It is difficult to overemphasize the importance of the reduced distance function, which is a space-time distance-like function (not necessarily nonnegative!) which is intimately tied to the geometry of solutions of the Ricci flow and the understanding of forming singularities. We also discuss stability and the linearized Ricci flow. Here the emphasis is not just on one solution to the Ricci flow, but on the dependence of the solutions on their initial conditions. We hope that this direction of study may have applications to showing that certain singularity types are not generic. This volume is divided into two parts plus appendices. For the most part, the division is along the lines of whether the techniques are geometric or analytic. However, this distinction is rather arbitrary since the techniques in Ricci flow are often a synthesis of geometry and analysis. The first part is intended as an introduction to some basic geometric techniques used in the study of the singularity formation in general dimensions. Particular attention is paid to finite time singularities on closed manifolds, where the spatial maximum of the curvature tends to infinity in finite time. We also discuss some basic 3-manifold topology and reconcile this with some classification results for 3-dimensional finite time singularities. The partial classification of such singularities is used in defining Ricci flow with surgery. In particular, given a good enough understanding of the singularities which can occur in dimension 3, one can perform topological-geometric surgeries on solutions to the Ricci flow either right before or at the singularity time. One would then like to continue the solution to the Ricci flow until the next singularity and iterate this process. In the end one hopes to infer the existence of a geometric decomposition on the underlying 3-manifold. This is what Hamilton's program aims to accomplish and this is the same framework on which 2Not all singularity models have been classified, even for finite time solutions of the Ricci flow on closed 3-manifolds. Apparently this is independent of Hamilton's program for Thurston geometrization.
HIGHLIGHTS OF PART I
xi
Perelman's work is based. In view of the desired topological applications of Ricci flow, in Chapter 9 we give a more detailed review of 3-manifold topology than was presented in Volume One. We hope to discuss the topics of nonsingular solutions (and their variants), where one can infer the existence of a geometric decomposition, surgery techniques, and more advanced topics in the understanding of singularities elsewhere. 3 The second part of this volume emphasizes analytic and geometric techniques which are useful in the study of Ricci flow, again especially in regards to singularity analysis. We hope that the second part of this volume will not only be helpful for those wishing to understand analytic and geometric aspects of Ricci flow but that it will also provide tools for understanding certain technical aspects of Ricci flow. The appendices form an eclectic collection of topics which either further develop or support directions in this volume. We have endeavored to make each of the chapters as self-contained as possible. In this way it is hoped that this volume may be used not only as a text for self-study, but also as a reference for those who would like to learn any of the particular topics in Ricci flow. To aid the reader, we have included a detailed guide to the chapters and appendices of this volume and in the first appendix we have also collected the most relevant results from Volume One for handy reference. For the reader who would like to learn more about details of Perelman's work on Hamilton's program, we suggest the following excellent sources: Kleiner and Lott [231]' Sesum, Tian, and Wang [326], Morgan and Tian [273], Chen and Zhu [81], Cao and Zhu [56], and Topping [356]. For further expository accounts, please see Anderson [5], Ding [126], Milnor [267], and Morgan [272]. Part of the discussion of Perelman's work in this volume was derived from notes of four of the authors [102]. Finally a word about notation; if an unnumbered formula appears on p. <::? of Volume One, we refer to it as (V1-p. <::?); if the equation is numbered <).", then we refer to it as (V1-<)."). Highlights of Part I In Part I of this volume we continue to lay the foundations of Ricci flow and give more geometric applications. We also discuss some aspects of Perelman's work on the Ricci flow. 4 Some highlights of Part I of this volume are the following: (1) Proof of the existence of the Bryant steady soliton and rotationally symmetric expanding gradient Ricci solitons. Examples of homogeneous Ricci solitons. Triviality of breather solutions (no nontrivial steady or expanding breathers result). The Buscher duality transformation of gradient Ricci solitons of warped product type. An 3Some of these topics will appear in Part II of this volume. 4Further treatment of Perelman's work will appear in Part II and elsewhere.
PREFACE
xii
(2)
(3)
(4)
(5)
(6)
open problem list on the geometry and classification of Ricci solitons. Introduction to the Kahler-Ricci flow. Long-time existence of the Kahler-Ricci flow on Kahler manifolds with first Chern class having a sign. Convergence of the Kahler-Ricci flow on Kahler manifolds with negative first Chern class. Construction of the Koiso solitons and other U(n)-invariant solitons. Differential Harnack estimates and their applications under the assumption of nonnegative bisectional curvature. A survey of uniformization-type results for complete noncompact Kahler manifolds with positive curvature. Proof of the global version of Hamilton's Cheeger-Gromov-type compactness theorem for the Ricci flow. We take care to follow Hamilton and prove the compactness theorem for the Ricci flow in the category of pointed solutions with the convergence in Coo on compact sets. Outline of the proof of the local version of the aforementioned result. Application to the existence of singularity models. A unified approach to Perelman's monotonicity formulas for energy and entropy and the expander entropy monotonicity formula. Perelman's A-invariant and application to the second proof of the no nontrivial steady or expanding breathers result. Other entropy results due to Hamilton and Bakry-Emery. Proof of the no local collapsing theorem assuming only an upper bound on the scalar curvature. Relation of no local collapsing and Hamilton's little loop conjecture. Perelman's J.L- and v-invariants and application to the proof of the no shrinking breathers result. Discussion of Topping's diameter control result. Relation between the variation of the modified scalar curvature and the linear trace Harnack quadratic. Second variation of energy and entropy. Theory of the reduced length. Comparison between the reduced length for static metrics and solutions of the Ricci flow. The c,length, L-, L-, and f-distances and the first and second variation formulas for the c'-length. Existence of C,-geodesics and estimates for their speeds. Formulas for the gradient and time-derivative of the L-distance function and its local Lipschitz property. Formulas for the Laplacian and Hessian of L and differential inequalities for L, L, and f including a space-time Laplacian comparison theorem. Upper bound for the spatial minimum of f. Formulas for f on Einstein and gradient Ricci soliton solutions. c'-Jacobi fields, the c'-Jacobian, and the c'-exponential map, and their properties. Estimate for the time-derivative of the c'-Jacobian. Bounds for f, its space-derivative, and its time-derivative. Properties of Lipschitz functions applied to f and equivalence of notions of supersolutions in view of differential inequalities for f.
ACKNOWLEDGMENTS
xiii
(7) Applications of the reduced distance. Reduced volume of a static metric and its monotonicity. Monotonicity formula for the reduced volume and application to weakened no local collapsing for complete (possibly noncompact) solutions of the Ricci flow with bounded curvature. Certain backward limits of ancient ~-solutions are shrinking gradient Ricci solitons. (8) A survey of basic 3-manifold topology and a brief description of the role of Ricci flow as an approach to the geometrization conjecture. (9) Concise summary of the contents of Volume One including some main formulas and results. Formulas for the change in geometric quantities given a variation of the metric, evolution of geometric quantities under Ricci flow, maximum principles, curvature estimates, classical singularity theory including applications of classical monotonicity formulas, ancient 2-dimensional solutions, Hamilton's partial classification of 3-dimensional finite time singularities. (10) List of some results in the basic theory of Ricci flow and the background Riemannian geometry. Bishop-Gromov volume comparison theorem, Laplacian and Hessian comparison theorems, Calabi's trick, geometry at infinity of gradient Ricci solitons, dimension reduction, properties of ancient solutions, existence of necks using the combination of classical singularity theory in dimension 3 and no local collapsing. (11) Discussion of some results on the asymptotic behavior of complete solutions of the Ricci flow on noncom pact manifolds diffeomorphic to Euclidean space. A brief discussion of the mean curvature flow (MCF) of hypersurfaces in Riemannian manifolds. Huisken's monotonicity formula for MCF of hypersurfaces in Euclidean space, including a generalization by Hamilton to MCF of hypersurfaces in Riemannian manifolds. Short-time existence (Buckland) and monotonicity formulas (Hamilton) for the cross curvature flow of closed 3-manifolds with negative sectional curvature. Acknowledgments
We would like to thank the following mathematicians for helpful discussions and/or encouragement: Sigurd Angenent, Robert Bryant, Esther Cabezas-Rivas, Huai-Dong Cao, Xiaodong Cao, Jim Carlson, Albert Chau, Bing-Long Chen, Xiuxiong Chen, Li-Tien Cheng, Shiu-Yuen Cheng, Yuxin Dong, Klaus Ecker, David Ellwood, Bob Gulliver, Hongxin Guo, Emmanuel Hebey, Gerhard Huisken, Bruce Kleiner, Brett Kotschwar, Junfang Li, Peter Li, John Lott, Robert McCann, John Morgan, Andre Neves, Hugo Rossi, Rick Schoen, Natasa Sesum, Weimin Sheng, Luen-Fai Tam, Gang Tian, Peter Topping, Yuan-Long Xin, Nolan Wallach, Jiaping Wang, Guofang Wei, Neshan Wickramasekera, Jon Wolfson, Deane Yang, Rugang Ye, Yu Yuan, Qi Zhang, Yu Zheng, and Xi-Ping Zhu. We would like to thank Jiaping
xiv
PREFACE
Wang for advice on how to structure the book. We are especially grateful to Esther Cabezas-Rivas for a plethora of helpful suggestions; we have incorporated a number of her suggestions essentially verbatim in the first part of this volume. The authors would like to express special thanks to the following people. Helena Noronha, formerly a geometric analysis program director at NSF, for her encouragement and support during the early stages of this project. Ed Dunne and Sergei Gelfand for their help, encouragement and support to make the publication of these books possible with the AMS. We are especially grateful and indebted to Ed Dunne for his constant help and encouragement throughout the long process of writing this volume. We thank Arlene O'Sean for her expert and wonderful help as the copy and production editor. During the preparation of this volume, Bennett Chow was partially supported by NSF grants DMS-9971891, DMS-020392, DMS-0354540 and DMS0505507. Christine Guenther was partially supported by the Thomas J. and Joyce Holce Professorship in Science. Jim Isenberg was partially supported by NSF grant PHY-0354659. Dan Knopf was partially supported by NSF grants DMS-0511184, DMS-0505920, and DMS-0545984. Peng Lu was partially supported by NSF grant DMS-0405255. Feng Luo was partially supported by NSF grant DMS-0103843. Lei Ni was partially supported by NSF grants DMS-0354540 and DMS-0504792. Bennett Chow and Lei Ni were partially supported by NSF FRG grant DMS-0354540 (jointly with Gang Tian). Bennett Chow would like to thank Richard Hamilton and Shing-Tung Yau for making the study of the Ricci flow possible for him and for their help and encouragement over a period of many years. He would like to thank East China Normal University, Mathematical Sciences Research Institute in Berkeley, National Center for Theoretical Sciences in Hsinchu, Taiwan, and Universite de Cergy-Pontoise. Ben expresses extra special thanks to Classic Dimension for continued support, faith, guidance, encouragement, and inspiration. He would like to thank his parents, Yutze and Wanlin, for their encouragement and love throughout his life. Ben lovingly dedicates this book to his daughters Michelle and Isabelle. Sun-Chin Chu would like to thank Nai-Chung Leung and Wei-Ming Ni for their encouragement and help over the years. Sun-Chin would like to thank his parents for their love and support throughout his life and dedicates this book to his family. David Glickenstein would like to thank his wife, Tricia, and his parents, Helen and Harvey, for their love and support. Dave dedicates this book to his family. Christine Guenther would like to thank Jim Isenberg as a friend and colleague for his guidance and encouragement. She thanks her family, in
ACKNOWLEDGMENTS
xv
particular Manuel, for their constant support and dedicates this book to them. Jim Isenberg would like to thank Mauro Carfora for introducing him to Ricci flow. He thanks Richard Hamilton for showing him how much fun it can be. He dedicates this book to Paul and Ruth Isenberg. Tom Ivey would like to thank Robert Bryant and Andre Neves for helpful comments and suggestions. Dan Knopf thanks his colleagues and friends in mathematics, with whom he is privileged to work and study. He is especially grateful to Kevin McLeod, whose mentorship and guidance has been invaluable. On a personal level, he thanks his family and friends for their love, especially Dan and Penny, his parents, Frank and Mary Ann, and his wife, Stephanie. Peng Lu benefits from the notes on Perelman's work by Bruce Kleiner and John Lott. Peng thanks Gang Tian for encouragement and help over the years. Peng thanks his family for support. Peng dedicates this book to his grandparents and wishes them well in their world. Feng Luo would like to thank the NSF for partial support. Lei Ni would like to thank Jiaxing Hong and Yuanlong Xin for initiating his interests in geometry and PDE, Peter Li and Luen-Fai Tam for their teaching over the years and for collaborations. In particular, he would like to thank Richard Hamilton and Grisha Perelman, from whose papers he learned much of what he knows about Ricci flow. Sun-Chin Chu, David Glickenstein, Christine Guenther, Jim Isenberg, Tom Ivey, Dan Knopf, Peng Lu, Feng Luo, and Lei Ni would like to collectively thank Bennett Chow for organizing this project. 5
Bennett Chow, UC San Diego and East China Normal University Sun-Chin Chu, National Chung Cheng University David Glickenstein, University of Arizona Christine Guenther, Pacific University Jim Isenberg, University of Oregon Tom Ivey, College of Charleston Dan Knopf, University of Texas, Austin Peng Lu, University of Oregon Feng Luo, Rutgers University Lei Ni, UC San Diego
[email protected] August 23, 2006
5Bennett Chow would like to thank his coauthors for letting him ride on their coattails.
Contents of Part I of Volume Two "There must be some way out of here, ... " - From "All Along the Watchtower" by Bob Dylan
We describe the main topics considered in each of the chapters of Part I of this volume.
Chapter 1. We consider the self-similar solutions to the Ricci flow, called Ricci solitons. These special solutions evolve purely by homotheties and diffeomorphisms. When the diffeomorphisms are generated by gradient vector fields, the Ricci solitons are called gradient. We begin by presenting a canonical form for gradient Ricci solitons and by systematically differentiating the gradient Ricci soliton equations to obtain higher-order equations. These equations play an important role in the qualitative study of gradient Ricci solitons. Warped products provide elegant and important examples of solitons, such as the cigar metric, which is an explicit steady Ricci soliton defined on ~2 and which is conformal to the Euclidean metric, positively curved and asymptotic to a cylinder. We also discuss the construction of the higherdimensional generalization of the cigar, the rotationally symmetric Bryant soliton defined on ~n. Interestingly, the qualitative behavior of the Bryant soliton is quite different from the cigar. We also construct rotationally symmetric expanding gradient Ricci solitons. An interesting class of Ricci solitons is the homogenous Ricci solitons. It is notable that in dimensions as low as 3, there exist expanding homogeneous Ricci solitons which are not gradient. We especially discuss the 3-dimensional case of expanding Ricci solitons. The scarcity of Ricci solitons on closed manifolds is exhibited by the fact that the only steady or expanding Ricci solitons on such manifolds are Einstein metrics. Moreover, in dimensions 2 and 3, the only shrinking Ricci solitons on closed manifolds are Einstein. In dimension 2, this follows from the Kazdan-Warner identity and in dimension 3 this relies on a pinching estimate for the curvature due independently to Hamilton and Ivey. The consideration of gradient Ricci solitons, in particular those geometric quantities which either vanish or are constant on gradient Ricci solitons, has played an important role in the discovery of monotonicity formulas. We briefly introduce Perelman's energy and entropy functionals from this point xvii
xviii
CONTENTS OF PART I OF VOLUME TWO
of view. These functionals will be considered in more detail in Chapters 5 and 6. Ricci solitons were actually introduced first in the physics literature, where nontrivial ones were called quasi-Einstein metrics. So it is perhaps not surprising that some aspect of duality theory is related to Ricci solitons. We discuss gradient Ricci solitons in the form of warped products with tori and the Buscher duality transformation of these special solitons. We conclude the chapter with a summary of results and open problems on Ricci solitons. A fundamental aspect of Ricci solitons is that they occur as singularity models, i.e., limits of dilations of a singular solution. In particular, for metrics with nonnegative curvature operators and whose scalar curvature attains its maximum in space and time, the limits of Type II singularities are steady Ricci solitons. One proof of this relies on the matrix differential Harnack inequality and the strong maximum principle for tensors, whose proofs are given in Part II. In the chapter on differential Harnack inequalities in Part II we shall motivate the consideration of the Harnack quadratic by differentiating the expanding gradient Ricci soliton equation to obtain the matrix Harnack quantity, which vanishes in certain directions on expanding gradient solitons. Chapter 2. We discuss the Kahler-Ricci flow, which is simply the Ricci flow on Kahler manifolds. In the compact case, the Ricci flow preserves the Kahler structure of the metric. Because of the interaction of the complex structure with the evolving metric, a rich field has developed in the study of the Kahler-Ricci flow. We begin by giving a basic introduction to Kahler geometry. This introduction is not meant as a replacement of the standard texts on Kahler geometry, but rather an attempt to make the book more self-contained. We encourage the novice to read other texts (some of these are cited in the notes and commentary) either before or in conjunction with this chapter. We emphasize local coordinate calculations in holomorphic coordinates in the style of the book by Morrow and Kodaira [275]. To put the study of the Kahler-Ricci flow in a broader context, we give a brief summary of some results on the existence and uniqueness of KahlerEinstein metrics. Many of the results in this field are deep and we encourage the interested reader to consult the original papers or other sources. Our study of the Kahler-Ricci flow begins with the fundamental result of H.-D. Cao on the long-time existence and convergence on closed manifolds. Long-time existence holds independently of whether the first Chern class is negative, zero, or positive. Convergence to a Kahler-Einstein metric holds in the cases where the Chern class is either negative or zero. There has been substantial progress on the Kahler-Ricci flow in both the compact and the complete noncom pact cases. We briefly survey some results in this area.
CONTENTS OF PART I OF VOLUME TWO
xix
There are a number of interesting Kahler-Ricci solitons. We present the construction of some of these solitons, including the Koiso soliton. All of the known examples have some sort of symmetry. It is hoped that the study of Kahler-Ricci solitons may shed some light on the problem of formulating weak solutions to the Kahler-Ricci flow as a way of canonically flowing past a singularity. Finally we discuss nonnegativity of curvature and the Kahler-Ricci flow. A particularly natural condition is nonnegative bisectional curvature. This condition is preserved under the Kahler-Ricci flow and H.-D. Cao has proved a differential matrix Harnack estimate under this curvature condition (assuming bounded curvature). The trace form of the matrix estimate may be generalized to a differential Harnack estimate which ties more closely to the heat equation. We present a family of such inequalities and discuss some applications. Chapter 3. The study of the limiting behavior of solutions begins with Hamilton's Cheeger-Gromov-type compactness theorem for the Ricci flow, which is the subject of this chapter. One considers sequences of pointed solutions to the Ricci flow and one attempts to extract a limit of a subsequence. Such sequences arise when studying singular solutions by taking sequences of points and times approaching the singularity time and dilating the solutions by the curvatures at these basepoints. In order to extract a limit, we assume that the injectivity radii at the basepoints and the curvatures everywhere are bounded. By Shi's local derivative bounds, we get pointwise bounds on all the derivatives of the curvatures. This enables us to prove convergence in Coo on compact subsets for a subsequence. We prove the compactness theorem for solutions (time-dependent) from the compactness theorem for metrics (time-independent), which will be proved in the next chapter. We also consider a local version of the compactness theorem and discuss the application of the compactness theorem to the existence of singularity models for solutions of the Ricci flow assuming an injectivity radius estimate (such an estimate holds for finite time solutions on closed manifolds by Perelman's no local collapsing theorem). The outline of the proof of the compactness theorem is as follows. One first proves a compactness theorem for pointed Riemannian manifolds with bounded injectivity radii, curvatures and derivatives of curvature. To prove the compactness theorem for pointed solutions {M,9k (t), OkhEN to the Ricci flow from this, one observes the following. By Shi's estimate and the compactness theorem for pointed Riemannian manifolds there exists a subsequence 9k (to) which converges for a fixed time to. The bounds on the curvatures and their derivatives also imply that the metrics 9k (t) are uniformly equivalent on compact time intervals and the covariant derivatives of the metrics 9k (t) with respect to a fixed metric 9 are bounded. The compactness theorem for solutions then follows from the Arzela-Ascoli theorem. We also briefly discuss the Cheeger-Gromov-type compactness theorems for both Kahler metrics and solutions of the Kahler-Ricci flow. The only
xx
CONTENTS OF PART I OF VOLUME TWO
issue in applying the Riemannian compactness theorem is showing that the limiting metric/solution is Kahler; fortunately this is easily handled. Chapter 4. In this chapter we first give an outline of the proof of the Cheeger-Gromov compactness theorem for pointed Riemannian manifolds. The proof of the compactness theorem for pointed manifolds is rather technical and involves a few steps. The main step is to define, after passing to a subsequence, approximate isometries Wk from balls B (Ok, k) in Mk to balls B (Ok+l, k + 1) in Mk+l. The manifold Moo is defined as the direct limit of the directed system Wk. Convergence to Moo and the completeness of this limit follows from Wk being approximate isometries. The ideas in the proof of the main step are as follows. In each of the manifolds Mk in an appropriate subsequence, starting with the origins Ok = x2, one constructs a net (sequence) of points {xl:} :~~) which will be the centers of balls Bk of the appropriate radii (technically one considers balls of different radii for what follows). By passing to a subsequence and appealing to the Arzela-Ascoli theorem repeatedly, we may assume these Riemannian balls have a limit as k ---t 00 for each Q. Furthermore, we may also assume that the balls cover larger and larger balls centered at Ok and that the intersections of balls Bk and B~ are independent of k in the limit. Choosing frames at the centers of these balls yields local coordinate charts HI: (this depends on a decay estimate for the injectivity radius and our choice of the radii of the balls) and we can define overlap maps
J~/3 =
(He)
-1 0
HI:-
By passing to a subsequence, we may assume the J~/3 converge as k ---t 00 for each Q and {3. The local coordinate charts also define maps between manifolds by F[J = Hfo(Hk)-l . Now we can define approximate isometries Fke : B (Ok, k) ---t Me by taking a partition of unity and averaging the maps F[J. Technically, this is accomplished using the so-called center of mass and nonlinear averages technique. This brings us to the remaining step, which is to show that these maps are indeed approximate isometries. Chapter 5. In Volume One we saw the integral monotonicity formula for Hamilton's entropy for solutions to the Ricci flow on surfaces with positive curvature. There we also saw various curvature pinching estimates, the gradient of the scalar curvature estimate, and higher derivative of curvature estimates. Other monotonicity-type formulas, for the evolution of the lengths and areas of stable minimal geodesics and surfaces, yielded injectivity radius estimates in various special cases in low dimensions. In a generalized sense, all of these estimates may be thought of as monotonicity formulas. In this chapter we address Perelman's energy formula. One of the main ideas here is the introduction of an auxiliary function, which serves several purposes. It fixes the volume form, it satisfies a backward heat-type equation, it is used to understand the action of the diffeomorphism group, and it relates to gradient Ricci solitons. We discuss the first variation formula for the energy functional and the modified Ricci flow as the gradient flow for
CONTENTS OF PART I OF VOLUME TWO
xxi
this energy. The nonexistence of steady and expanding breathers on closed manifolds, originally proved by one of the authors, may be proved using the energy functional and an associated invariant. We also discuss the classical entropy and its relation to Perelman's energy. Chapter 6. In this chapter we discuss Perelman's remarkable entropy functional. This functional is actually an energy-entropy quantity which combines Perelman's energy with the classical entropy using a positive parameter 7 which, in the context of Ricci flow, plays the dual roles of the scale and minus time. We compute the first variation of the entropy and derive its monotonicity under the Ricci flow coupled to the adjoint heat equation. The entropy formula, via the consideration of test functions concentrated at points, leads to a volume noncollapsing result for all solutions to the Ricci flow on closed manifolds, called no local collapsing. This result also yields a strong injectivity radius estimate and rules out the formation of the cigar soliton as a finite time singularity model on closed manifolds. By minimizing the entropy functional over all functions satisfying a constraint and then minimizing over all scales 7, we obtain two geometric invariants, one depending on the metric and scale and one depending only on the metric. The consideration of these invariants is useful in proving the nonexistence of nontrivial shrinking breather solutions on closed manifolds. We also provide an improved version of the no local collapsing theorem, also due to Perelman. Our presentation is based on the diameter bound result of Topping, whose proof we also sketch. Finally we discuss variational formulas for the modified scalar curvature, which is an integrand for Perelman's entropy functional, the second variation of energy and entropy functionals, and also a matrix Harnack-type calculation for the adjoint heat equation coupled to the Ricci flow. Chapter 7. In this chapter we give a detailed introduction to Perelman's reduced distance, also called the i-function. To reduce technicalities, we first consider the analogous function corresponding to fixed Riemannian metrics, which is simply the function d (p, q)2 /47. In the Ricci flow case we first consider the C-Iength. After presenting some basic properties of the C-Iength, we compute its first variation formula and discuss the existence of C-geodesics, which are the critical points of the C-Iength. Associated to the C-Iength is the L-distance, which is obtained by taking the infimum of the C-Iength over paths with given endpoints. We compute the first space- and time-derivatives of the L-distance. This is partially analogous to the Gauss lemma in Riemannian geometry. Next we compute the second variation formula for the C-Iength, and motivated by space-time considerations, we express the formula in terms of Hamilton's matrix Harnack quadratic. This second variation formula yields an estimate for the Hessian of the L-distance (and hence for the reduced distance). Next we derive a number of differential equalities and inequalities for the reduced distance. These inequalities are the basis for the use of the reduced distance in the study of singularity formation under
xxii
CONTENTS OF PART I OF VOLUME TWO
the Ricci flow. We also consider the reduced distance in the special cases of Einstein solutions, and more generally, gradient Ricci solitons. There is a whole space-time geometry associated to the .c-Iength and reduced distance. We consider the notions of .c-Jacobi fields and .c-exponential map. We derive properties of these objects including the .c-Jacobi equation, bounds for .c-Jacobi fields, the .c-cut locus, and .c-Jacobian. We derive bounds for the reduced distance, its spatial gradient, and its time-derivative. Since the reduced distance is a Lipschitz function, we recall the basic properties of Lipschitz functions and formulate the precise sense in which differential inequalities for the reduced distance hold. Chapter 8. We discuss applications of the study of the reduced distance to the study of finite time singularities for the Ricci flow. First we consider the reduced volume associated to a static metric. This is simply the integral of the transplanted Euclidean heat kernel using the exponential map based at some point. In the case of nonnegative Ricci curvature, the static metric reduced volume is monotonically nonincreasing. This corresponds to the fact that the reduced volume integrand is a weak subsolution to the heat equation. With analogies to no local collapsing in mind, we relate the static metric reduced volume to volume ratios of balls. Next we consider Perelman's reduced volume for the Ricci flow. For all solutions of the Ricci flow on closed manifolds, the reduced volume is monotonically nondecreasing. We present various heuristic proofs and then justify these proofs using the basic properties of the reduced distance as a Lipschitz function and the .c-Jacobian developed in the previous chapter. We prove a weakened version of the no local collapsing theorem using the reduced volume monotonicity. This proof is somewhat technical since one needs some estimates for the .c-exponential map. Its advantage over the entropy proof given in Chapter 6 is that it holds for complete solutions of the Ricci flow on noncompact manifolds with bounded curvature. Perelman's no local collapsing theorem tells us that singularity models in dimension 3 are ancient K-solutions. To obtain more canonical limits, one often needs to take backward limits in time and rescale to obtain new ancient K-solutions. The reduced distance function may be used to show that certain backward limits of ancient K-solutions are shrinking gradient Ricci solitons. In dimension 3 such a soliton must either be a spherical space form, the cylinder 8 2 x JR, or its Z2-quotient. This has important consequences for singularity formation in dimension 3. Chapter 9. In this chapter we give a survey of some of the basic 3manifold topology which is related to the Ricci flow. Appendix A. We review the contents of Volume One and other aspects of basic Riemannian geometry and Ricci flow. Appendix B. In many cases, solutions to the Ricci flow limit to Ricci solitons. We present some low-dimensional results of this type for certain
CONTENTS OF PART I OF VOLUME TWO
xxiii
classes of complete solutions with bounded curvature on noncom pact manifolds diffeomorphic to Euclidean space. We also discuss the mean curvature flow and the cross curvature flow. Appendix C. This is a glossary of terms related to the study of the Ricci flow.
CHAPTER 1
Ricci Solitons The art of doing mathematics consists in finding that special case which contains all the germs of generality. - David Hilbert The last thing one knows when writing a book is what to put first. - Blaise Pascal
The notion of soliton solutions to evolution equations first appeared in connection with the modelling of shallow water waves and the Kortewegde Vries equation [131]. In this context, a soliton is, in the memorable phrase of Scott-Russell, a 'wave of translation', i.e., a solitary water wave moving by translation, without losing its shape. More generally, we now think of solitons as self-similar solutions, i.e., solutions which evolve along symmetries of the flow. In the case of the Ricci flow, these symmetries are scalings and diffeomorphisms. In this chapter we study Ricci solitons and in particular the special case of gradient Ricci solitons. In later chapters we shall further see how these special solutions of the Ricci flow motivate the general analysis of the Ricci flow through monotonicity formulas and their subsequent applications. In particular, Ricci solitons have inspired the entropy and Harnack estimates, the space-time formulation of Ricci flow, and the reduced distance and reduced volume. Furthermore, the entropy and reduced volume monotonicity formulas have the geometric application of no local collapsing, which is fundamental in the study of singularities (see the diagram below). Gradient Ricci solitons also model the high curvature regions of singular solutions. This motivates trying to classify gradient Ricci solitons, especially in low dimensions.! Gradient Ricci solitons
/
I Entropy I
/
f-----t
'\.
I Harnack I
f-----t
I Space-time I
'\.
'\. f-----t
I Red.
dist. & vol.
I
/
...
No local collapsing 1 "Red. dist. & vo!." is an abbreviation for reduced distance and volume, to be discussed in Chapter 7.
1. RICCI SOLITONS
2
Some highlights of this chapter are a systematic description of the equations obtained from differentiating the gradient Ricci soliton equation, the constructions of the Bryant steady Ricci soliton and the rotationally symmetric expanding Ricci soliton with positive curvature operator, examples of homogeneous Ricci solitons, introduction to Perelman's energy and entropy functionals via gradient Ricci solitons, and the Buscher duality transformation. 1. General solitons and their canonical forms
We begin by recalling the following. DEFINITION 1.1 (General Ricci soliton). A solution g(t) of the Ricci flow on Mn is a Ricci soliton (or self-similar solution) if there exist a positive function u(t) and a I-parameter family of diffeomorphisms c.p(t) : M ~ M such that
g(t) = dt)c.p(t)*g(O).
(1.1)
Let VJ1et denote the space of Riemannian metrics on a differentiable manifold M, and let niff denote the group of diffeomorphisms of M. Consider the quotient map 7r : VJ1et ~ VJ1etjniff x R+, where R+ acts by scalings. One verifies that a Ricci soliton is a solution 9 (t) of the Ricci flow for which 7r (g (t)) is independent of t, i.e., stationary. We start by looking at what initial conditions give rise to Ricci solitons. Differentiating (1.1) yields
(1.2)
-2 Rc (g(t))
= ir(t)c.p(t)*gO + u(t)c.p(t)* (CXgo) ,
where go = g(O), C denotes the Lie derivative, X is the time-dependent vector field such that X (c.p( t) (p)) = (c.p( t)(p)) for any p EM, and ir ~ ~~.
it
1.2. For obvious reasons, we say that g(t) is expanding, steady, or shrinking at a time to if ir (to) is > 0, = 0, or < 0, respectively. DEFINITION
Since Rc(g(t))
= c.p(t)* Rc(go), we can drop the pullbacks in (1.2) and
get
(1.3) where X (t) = u(t)X (t). Although go is independent of time, both ir(t) and X(t) may depend on time. For example, static Euclidean space (Mn, g(t)) == (l~n, gcan) is a stationary solution to the Ricci flow and, as such, is a steady Ricci soliton; however, this solution may also be considered as a Ricci soliton which expands or shrinks modulo diffeomorphisms. In particular, given any function u (t) > 0 with dO) = 1, consider the diffeomorphisms c.p(t) : ]Rn ~ ]Rn defined by c.p(t) (x) = u(t)-1/2x for x E ]Rn. Then c.p(t)*gcan = u(t)-lgcan. Since g(t) == gcan, we may rewrite this as
(1.4)
g(t) = gcan = u(t)c.p(t)*g(O).
1.
GENERAL SOLITONS AND THEIR CANONICAL FORMS
3
By choosing u (t) so that o-(t) changes sign, this soliton may both expand and shrink at different times. On the other hand, in general, a Ricci soliton solution is evolving purely by scaling (modulo diffeomorphisms). Since the time-derivative of the metric is equal to the negative of twice the Ricci tensor, which is scale-invariant, it is thus natural to ask whether one can put the general soliton equation in a canonical form where u (t) is equal to a linear function. The following gives us a condition under which we may assume that a Ricci soliton defined as in (1.1) has such a form. PROPOSITION 1.3 (Canonical form for a general soliton). Let (Mn, g( t)) be a Ricci soliton, and assume that the solution of the Ricci flow with initial metric go = g(O) is unique among soliton solutions. Then there exist diffeomorphisms 'I/J(t) : M --t M and a constant c E ~ such that (1.5)
g(t) = (1
+ ct) 'I/J(t)*go.
PROOF. Differentiating (1.3) with respect to time gives (1.6) Case 1. If a(t) == 0, then u(t) = 1 + ct for some constant c. Hence, by (1.1), we may simply take 'I/J(t) =
(1.7)
= -X (to) ja(to) at some
CYogO = go·
Substituting (1.7) into (1.3), we have -2 Rc(go)
= Ca(t)Yo+X(t)90
for all t. Consider the vector field
Xo ~ o-(O)Yo + X(O). Then -2 Rc(go) = Cxogo. Let 'I/J(t) be the I-parameter group of diffeomorphisms generated by Xo. Then it is easy to check that g(t) = 'IjJ(t)*gO satisfies the Ricci flow with the same initial conditions go and is a steady soliton. Thus, by our uniqueness assumption for soliton solutions to the Ricci flow with initial metric go, by replacing
REMARK 1.4. The proof shows that under the uniqueness assumption, a Ricci soliton not in canonical form can be made a steady soliton. If (Mn,g(t)) is a Ricci soliton where (1.5) holds, then we say that the soliton is in canonical form. By rescaling, we may assume that c = -1,0, or 1; these cases correspond to solitons of shrinking, steady, or expanding type, respectively.
4
1.
RICCI SOLITONS
A way to circumvent the undesirableness of the uniqueness assumption in Proposition 1.3 is as follows. Since the geometry of any Ricci soliton g(t) is the same as that of go, we will start with go and then construct another Ricci soliton in canonical form and with the same initial metric go. Choose any time to and let E ~ &(to) and Xo ~ X(to), so that equation (1.3) becomes at t = to -2 Rc(go) = Ego
+ LXogo.
We will now drop the subscripts on X and g. Using indices, the equation above now reads (1.8) where Xi = gijxj are the components of X~, the covariant tensor (I-form) obtained from X by lowering indices using g. A triple (g,X,E) (or pair (g, X) if we suppress the dependence on E) consisting of a metric and a vector field that satisfies (1.8) for some constant E is called a Ricci soliton structure. 2 We say that X is the vector field the soliton is flowing along. REMARK 1.5 (Solitons and normalized Ricci flow). If (g,X) is a Ricci soliton structure on a compact manifold M, then 9 evolves purely by diffeomorphisms under the normalized (constant volume) Ricci flow. DEFINITION 1.6 (Gradient Ricci soliton). A Ricci soliton structure (g, X) is a gradient soliton structure if there exists a function f (called the potential function) such that X~ = df. In this case, (1.8) becomes (1.9)
Rij
E
+ ViVjf + "2 gij = O.
The following, whose proof is elementary, shows that given a complete gradient Ricci soliton structure, we can construct a gradient Ricci soliton in canonical form (see Theorem 4.1 on p. 154 of [111] or Kleiner and Lott [231]). In particular, the result below illustrates the sense in which a Ricci soliton structure may be regarded as initial data for a Ricci solution, i.e., for a self-similar solution to Ricci flow. PROPOSITION 1.7 (Gradient soliton structures and canonical forms). Suppose (gO, V fo, E) is a complete gradient Ricci soliton structure on Mn. Then there exists a solution g (t) of the Ricci flow with 9 (0) = go, diffeomorphisms 'P (t) with 'P (0) = idM, and functions f (t) with f (0) = fo defined for all t with (1.10)
T
(t)
~
1 + Et
> 0,
such that 2Below, we will sometimes denote a Ricci soliton structure by (M n , g, X), in order to emphasize the underlying manifold, e.g. when M is a Lie group.
1.
GENERAL SOLITONS AND THEIR CANONICAL FORMS
5
(1)
a
at cp (t) (x) =
1 7
(t) (gradgofo) (cp (t) (x)),
(2) 9 (t) is the pull-back by cp (t) of go up to the scale factor g(t) =
(1.11)
7
7
(t) ,
(t) cp(t)*gO,
(3) f (t) is the pull-back by
f (t) = fo
(1.12)
0
cp(t) =
Moreover, (1.13)
Rc (g (t))
+ V'g(t)V'g(t) f
(t)
+ ;7 g (t) = 0,
where V'g(t) denotes the covariant derivative with respect to 9 (t), and af (t) = 1gradg(t) f (t) 12 . -a t get)
(1.14)
EXERCISE 1.8. Prove Proposition 1.7. HINT: Verify the equations in the order in which they are presented. Because of the proposition above, we shall at some times consider gradient Ricci soliton structures and at other times consider gradient Ricci solitons in canonical form. Now we give a couple of examples of Ricci solitons in canonical form. We are already acquainted with solutions which evolve purely by homothety, e.g., these solutions correspond to Einstein metrics. If go is an Einstein metric with Einstein constant A (Le., Rc(go) = Ago), then
g(t) = (1 - 2At)go satisfies the Ricci flow
a g = -2 Rc(g), m
since Rc(g) = Rc(go) = Ago. An Einstein manifold with A > 0 is necessarily compact (thanks to Myers's theorem) and, as the metric approaches zero, the manifold shrinks to a point in finite time. Einstein metrics are stationary points for the normalized Ricci flow on a closed manifold. Following up on our previous discussion of static Euclidean space, we have the following. EXAMPLE 1.9 (Gaussian soliton). Regard Euclidean space as a Ricci soliton in canonical form, so that (1.4) holds with CT(t) = 1 + ct for some c E R For c '" 0, the Euclidean solution is called the Gaussian soliton. By differentiating (1.4) and multiplying the result by CT(t), we obtain (1.15)
0 = -2 Rc(gcan) = D"fgcan + cgcan,
where (1.16)
f(x)=_c 1x I2 4
1. RICCI SOLITONS
6
is the exponent in the Gaussian function. Here we used the standard identity D"jgcan = 2,1''\1 f. 2. Differentiating the soliton equation analysis
local and global
2.1. Differentiating general solitons. Let (g, X) be a Ricci soliton structure (gradient or not) on Mn: (1.17)
2~j
+ 'ViXj + 'VjXi + Egij
=
O.
Condition (1.17) places a strong condition on 9 and X. For example, contracting (1.17) with 9 (tracing) gives (1.18)
R + div X
+ ~E
= 0,
where div X ~ gij'ViXj and R is the scalar curvature. If M is closed, then this implies 2r E = --,
(1.19)
n where r ~ iM R d{L/ vol(M) denotes the average scalar curvature. Furthermore, we also have the following. LEMMA 1.10. If (g, X) is a Ricci soliton structure on Mn, then (1.20) or more invariantly,
~Xi
+ ~£ lm Xm = 0,
~Xo +Rc (x o) = 0,
where Rc : T* M - T* M is defined by Rc (a)i ~ ~£ lmam . PROOF. Taking the divergence of (1.17) and applying the second contracted Bianchi identity (VI-3.13) and the Ricci identity (Vl-p. 286b), we obtain
£) mXm + ~Xi
= g3'k ( 'Vi'VjXk - Rjik£g
= -'ViR+~£lmxm+~Xi'
where we have used (1.18). The lemma follows from cancelling the -'ViR terms. 0 Lastly, computing the scalar curvature of the evolving metric (1.11) and comparing its time-derivative (at t = 0) with its evolution under the Ricci flow gives LEMMA 1.11. If (g, X) is a Ricci soliton structure on Mn, then (1.21)
~R + 21 Rc 12 = CxR - ER.
2.
DIFFERENTIATING THE SOLITON EQUATION
7
PROOF. The left-hand side (LHS) is just the usual expression for ~~ under the Ricci flow. To obtain the equality, we compute ~~ by way of the equality R(Tl.p*g) = -r- 1
(1.22)
b.R - (\1 R, X)
+ 21 Rc 12 + cR = O.
We rewrite this as (1.23) b. ( R +
~c) - \\1 ( R + ~c) ,X) + 21 Rc + ~ g 12 -
c(R + ~c) = O.
At any point Xo E M such that R (xo) = R min , we have 21Rc +~gI2
_ c(R + ~c)
::; O.
If (g, X) is expanding or steady, then c ~ O. Since R(xo) + ~e = R(xo) - r is nonpositive, this implies IRc +~gI2 (xo) = O. Tracing then implies nc
Rmin
= R(xo) = -2 = r,
and hence R == r = - ~e. Substituting back into (1.23), we conclude that Rc == -~g. If (g, X) is shrinking, then c < O. Applying the weak maximum principle to (1.22), we have R ~ O. By the strong maximum principle, either R == 0 or R > O. If R == 0, then (1.22) would imply Rc == 0, contradicting the assumption that (g, X) is shrinking. Hence R> O. D We will return to the consideration of shrinking solitons on closed manifolds in Section 7 of this chapter.
2.2. Differentiating gradient solitons. The gradient Ricci soliton structure equation (1.24)
relates the Ricci tensor to the Hessian of f. First note that tracing this gives nc (1.25) R + b.f + 2 = o.
1. RICCI SOLITONS
8
Now we proceed to systematically differentiate the function f up to fourth order, commute pairs of covariant derivatives, trace, and apply (1.24) and the second Bianchi identity to relate derivatives of the curvature and f. We begin with considering three derivatives of f since for one derivative there is nothing to do, and a pair of covariant derivatives acting on a function commute. Since ''ih [\7 i , \7 j ] f = 0, the only nontrivial commutator is \7 i \7j\7kf - \7j\7i\7d = [\7 i , \7j ] \7kf.
By (1.24), the commutator formula, and \7g = 0, we have (1.26) Taking the trace by multiplying by gjk and using the contracted second Bianchi identity, we get (1.27) Note that (1.26) is antisymmetric in i and j, and in particular, (1.27) is the only equation obtained by tracing (1.26). Next we consider equations obtained by commuting four derivatives of f. The only essentially new equation is obtained by considering \7 i [\7 j, \7 k] \7 d. The quantity \7 i \7 j [\7 k, \7 e] f is zero and [\7 i, \7 j] \7 k \7 d yields only a standard commutator formula. We have
which implies
and hence (1.28) First we trace by gij. Commuting derivatives and applying the contracted second Bianchi identity yield b.Rke -
1
"2 \7 k \7 eR + 2~kem~m -
E
RkmRme
+ "2 Rke
= \7 iRikem \7 mf = (-\7eRkm + \7 mRke) \7 mf. Since (1.28) is antisymmetric in j and k, we have that tracing (1.28) by gjk is zero, whereas tracing by gik is equivalent to tracing by gij. The remaining trace is by gil. We leave it as an exercise for the reader to show that taking this trace yields (-\7jRkm + \7kRjm) \7 m f = 0, which is nothing new since it follows directly from (1.26).
2.
DIFFERENTIATING THE SOLITON EQUATION
9
In conclusion, the new identities we obtain by differentiating fourth order are (1.29) (1.30)
-\1iRjk
+ \1jRik =
f
up to
-Rijke\1d,
\1i\1jRkf - \1i\1kRje = \1i R jkem \1 mf - Rjkem (Rim
+ ~gim)
and the traces are 1
"2 \1i R =
(1.31 ) (1.32) (1.33)
Rie\1d,
+ \1mRkf) \1mf,
M ke = (-\1f Rkm tlR + 21Rcl 2
+ eR =
\1f· \1R,
where
1 Mkf ~ tlRkf - "2\1k \1e R
+ 2RikfmRim -
RkmRmf
e
+ "2Rkf'
The last equation, which is the trace of the equation above it, is a special case of (1.21). REMARK 1.14. The quantities in (1.29) and (1.32) appear in the matrix Harnack quadratic discussed in Part II of this volume (see also subsection 4.2 of Appendix A). We will finish this subsection with an application of Lemma 1.10 to gradient solitons. Substituting X = \1 f in (1.20) yields tl(\1d)
+ Rifg fm \1 mf = o.
On the other hand, commuting covariant derivatives gives tl(\1d) = \1i(tlf)
+ Riklf\1ff.
Combining these equations with (1.25) and (1.9) yields 0= \1iR + 2gkf\1 i \1 kf\1d
+ e\1d =
\1 i (R + 1\1 fl2
+ ef),
proving the following. PROPOSITION 1.15 (Constant gradient quantity on solitons). 1f(g, \1 f, e) is a gradient Ricci soliton structure on a manifold Mn, then (1.34)
R + 1\1 fl2
+ ef == const
is constant in space. Consequently, by (1.25), (1.35)
R + 2tlf - 1\1 fl2 - ef
== const.
Formula (1.34) is used in the study of the geometric properties of gradient Ricci solitons; see Chapter 9 of [111] for an exposition. A significance of the quantity on the LHS of (1.35) will be exhibited in the use of (5.42) and (5.43) to prove energy monotonicity in Chapter 5. With additional hypotheses, we can determine the constant in the proposition above (see also Theorem 20.1 in [186]).
10
1. RICCI SOLITONS
COROLLARY 1.16. If (g, V J) is a steady gradient Ricci soliton structure on Mn with positive Ricci curvature and if the scalar curvature attains its maximum at a point 0, then
R+
IVfl 2 = R(O).
PROOF. We have Rij = -ViVjf > 0 and by Proposition 1.15, R+IV fl2 is constant. On the other hand, at 0 we have 0 = Vi IV fl 2 = -2RijgJ'k Vkf, which implies that V f (0) = O. 0 REMARK 1.17. If we only assume the Ricci curvature is nonnegative, then there is a counterexample by simply taking the flat metric on the xyplane and letting f be the x-coordinate, in which case Rij = - Vi V j f = 0, so that R + IV fl2 = const > R(O) = O. Quantities which are constant on gradient Ricci solitons are useful in the study of their geometries; see Chapter 9 of [111] for an exposition of some examples of this. 2.3. Ricci soliton structures and exterior differential systems. We end this section with a local analysis of Ricci soliton structures. From (1.17), which defines a Ricci soliton structure (g, X), one might hope to obtain more stringent tensoriallocal conditions on 9 and/or X by a combination of differentiation, contracting, and equating mixed partials. However, no further lower-order (Le., first-order in X and second-order in g) identities arise this way. This follows from applying the machinery of exterior differential systems to the soliton equation (1.17), written in local coordinates as a system of PDE that is second-order in the entries of 9 and first-order in the components of X. The theory of exterior differential systems - in particular, the Cartan-K8.hler Theorem3 - is able to predict when a system of PDE has solutions, and how large the solution space is, provided the system passes a test indicating that it is involutive. 4 In addition, if the system is involutive, then any k-jet of a solution can be extended to a (k + I)-jet of a solution, and so on to a convergent power series solution. In the case of (1.17), in order to obtain an involutive system, one has to reduce the enormous size of the solution space (which is due to the diffeomorphism invariance of the soliton condition) by adding the requirement that the local coordinates are harmonic functions with respect to the metric g. Since ~ = gij (OiOj - rfj Ok) , this is a first-order condition on the entries of 9 and takes the form (1.36)
gijrfj = 0,
where rfj are the Christoffel symbols. This is just another variation on DeTurck's trick for proving short-time existence for the Ricci flow (see 3See Theorem III.2.2 and Corollary II1.2.3 on pp. 80-88 in [36] or Theorem 7.3.3 on pp. 254-256 in [223]. 4See Chapter III, pp. 58-65 of [36] or Chapter 7, p. 256 in [223].
3. WARPED PRODUCTS AND 2-DIMENSIONAL SOLITONS
11
Sections 4.3 and 5 of Chapter 3 of [108]). Once this condition is added to (1.17), the system becomes involutive (see [221] for the proof). Furthermore, involutivity implies that Cauchy problems for the augmented system (1.17) and (1.36) are locally solvable; for example, the I-jet of 9 may be arbitrarily prescribed along a hypersurface transverse to a given vector field X. Thus, Ricci soliton structures (g, X) locally depend, modulo diffeomorphisms, on n(n + 1) functions of n - 1 variables (i.e., the components of g, and their transverse derivatives, along the hypersurface). 3. Warped products and 2-dimensional solitons
In this section, we review some of the examples of 2-dimensional Ricci solitons, such as the cigar, which have been constructed to date; the Bryant soliton is discussed in Section 4 of this chapter and examples of Kahler-Ricci solitons will be discussed in the next chapter. In Section 1 of Appendix B we will give some conditions under which the Ricci flow converges to some of the solitons discussed here. Recall from Lemma 5.96 on p. 168 of Volume One that any ancient solution of the Ricci flow on a surface of positive curvature that attains its maximum curvature in space and time is the cigar. Similarly, any ancient solution of the Ricci flow on a manifold of positive curvature operator that attains its maximum curvature in space and time is a steady gradient Ricci soliton. In dimension 3, it is conjectured that such a soliton is the Bryant soliton. This is one reason for focusing our attention on the cigar and Bryant solitons. 3.1. Solitons and Killing vector fields on surfaces. If (g, X) is a soliton structure on a surface M2, then X is a conformal vector field. This is simply because ViXj + VjXi = - (R + c:) gij' If (g, 'V f) is a gradient soliton, then J ('V f) is a Killing vector field (where J : T M ~ T M is the complex structure, defined as counterclockwise rotation of tangent vectors by 90°). To see this, we observe that since 'V f is a conformal vector field on a surface,
(CJ('\lf)g)ij
= Vi (Jj'Vkf) + 'Vj (JikVkf) = Jj'Vi'Vkf + Jik'Vj'Vkf = - (
~ + ~) (Jj gik + Jik9jk) =
0,
where the components J! are defined by J (8~i) ~ J! 8~j' That is, J ('V f) is a Killing vector field. In dimension 2 we have the following. LEMMA 1.18. A surface with a Killing vector field is locally a warped product. In particular, a gradient soliton on a surface is locally a warped product.
12
1.
RICCI SOLITONS
PROOF. Let (M2, g) be a Riemannian surface (not necessarily complete) with a Killing vector field K. Let x E M and define a smooth unit speed path'Y: (ro - E, ro + E) -+ M by 'Y (r) = I~~~~I ("( (r)) , 'Y (ro) = x. Define a I-parameter family of smooth paths !3r : (00 - E, 00 + E) -+ M by /3r (0) = K (!3r (0)), !3r (00 ) = 'Y (r), and a I-parameter family of smooth unit speed paths 'Yo : (ro - E, ro + E) -+ M by 'Yo (r) = I~~~~I ("(0 (r)), 'Yo (ro) = !3ro (0). Note that 'Yoo = 'Y. CLAIM. !3r (0) = 'Yo (r). PROOF OF CLAIM. This follows from
J (K) ] [ K, IJ (K)I
1
1
1
2
= IJ (K)I [K, J (K)]- 21J (K)1 3K IJ (K)I J (K) = 0
since [K, J (K)] = 0 and K IJ (K)1 2 = O. Hence (r,O) defines local coordinates on a neighborhood of x. Define f (r) ~ IKI (!3r) , where we are using the fact that IKI is constant on !3r. The metric is given by / J (K)
9 = \ IJ (K)I' =
J (K))
2
/
IJ (K)I dr + 2 \
J (K) ) K, IJ (K)I drdO
+ (K, K) dO
2
dr 2 + f (r)2 d0 2.
o It can be shown on a surface that the integral curves of a Killing vector field J (\7 f) are closed loops. In particular one can prove the following without using the Uniformization Theorem (see [86]).
THEOREM 1.19 (Shrinking surface soliton is spherical). If (M2,g) is a shrinking gradient soliton on a closed surface, then 9 has constant positive curvature. More generally, if (M2, g) is a shrinking gradient soliton on a closed, orient able 2-dimensional orbifold with isolated singularities, then 9 is rotationally symmetric with positive curvature. When the 2-orbifold is bad, i.e., not covered by a smooth surface, a unique Ricci soliton exists, which is a nontrivial shrinking gradient Ricci soliton (see [180] and [372]). Topologically, a closed, bad 2-orbifold has either one or two singular points [343]. REMARK 1.20. There do not exist complete shrinking Ricci solitons on noncompact surfaces.
3.2. Warped products. Many known examples of gradient Ricci solitons have been constructed in the form of warped products. In particular, we consider a metric on the product Mn+1 = I x Nn of the form (1.37)
3.
WARPED PRODUCTS AND 2-DIMENSIONAL SOLITONS
13
where r is the standard coordinate on an interval I c JR, 9 is a given metric on an n-dimensional manifold N, and w(r) > 0 is the warping function, which scales distances along the N-factors in the product. Several well-known metrics may be expressed as warped products, at least on an open subset. When N is the unit circle S1, with 9 = d(P for the standard coordinate e, then w(r) = r, w(r) = sinr, w(r) = sinhr give, respectively, the Euclidean plane, the round sphere (with Gauss curvature +1), and the hyperbolic plane (with Gauss curvature -1) with the origin omitted. When N = sn with the standard metric 9 = gcan of constant curvature +1, then w(r) = r gives the standard (flat) metric on JRn+1 with the origin omitted. In each of these cases, the metric extends smoothly across the origin. When N = sn, Lemma 2.10 on p. 29 in Volume One (Le., Lemma A.2 of this volume) gives sufficient conditions for smoothly closing off the Riemannian manifold M, by a point at one (or both) ends of the interval I. Namely, assuming we wish to close off as r ~ 0, we need limr-+o w(r) = 0 and limr-+o w'(r) = 1 (the prime denotes the derivative with respect to r). For constructing Ricci solitons, it is convenient that 9 be a sufficiently 'nice' metric on N; for the rest of this section, we will assume that 9 is an Einstein metric, with Rc(g) = pg. An easy calculation in moving frames gives the following (see Proposition 9.106 on p. 266 of Besse [27]). LEMMA 1.21 (Ricci tensor and Hessian of warped product). If 9 is an Einstein metric on a manifold Nn, with Einstein constant p, then the Ricci tensor of the warped product metric (1.37) is given b'lf (1.38)
w"
Rc(g) = -n-dr2 W
+ (p -
ww" - (n - 1)(w')2) g.
Furthermore, if f is any function of the radial coordinate r, then the Hessian of f with respect to 9 is given by (1.39)
"V"Vf
=
j"(r)dr2 +ww'f'g.
J:
For example, if f(r) = w(t)dt, then "V "V f = w'(r)g. Conversely, by an argument of Cheeger and Colding, warped products may essentially be characterized by the existence of a function whose Hessian is some function times the metric; see §1 in [71] for details. 3.3. Constructing the cigar and other 2-dimensional solitons. For the sake of starting with something familiar, before we study the Bryant soliton, we will briefly repeat the construction of the cigar metric, which is discussed at greater length in Chapter 2 of Volume One. Suppose we wish to construct a complete, steady, rotationally symmetric gradient soliton metric on JR 2. Such a metric will be a warped product (1.37), 5In the case where (N, g) is the unit n-sphere, this follows from (1.58).
1. RICCI SOLITONS
14
where N = Sl, and it is natural to assume that Rc(g) + 'V 2 f = 0 for a radial function f(r). Using (1.38), (1.39) with p = 0 and n = 1 gives (1.40) Integrating
wi" - w" = 0 = w(w' f' - w").
wi" - w' f' = 0 gives
f'
(1.41)
=
for some constant a, whereupon w' f' (1.42)
2aw
- w" integrates to
w' - aw 2 = b.
Using the closure conditions w(O) = 0, w'(O) = 1, we get b = 1. Integrating, we obtain a smooth odd function w(r) whose type depends on the sign of a: • for a = 0, w(r) = r, giving the fiat metric; 1 • for a = a 2 , w(r) = -tan(ar); a 1 2 • for a = -a , w(r) = -tanh(ar). a The third case is the cigar soliton (see [180], [373]) (1.43)
gcig =
dr
212
+ 2" tanh a
2
(ar)dO ,
where by (1.41) we see that the potential function may be taken to be f(r) = -2Iogcosh(ar). The Gauss curvature is K = 2a 2 sech2 (ar) > O. In the second case, the metric (see [125]) (1.44)
1 2 2 = dr 2 + 2" tan (ar)dfJ ,
gxpd
a o < r < 7r/ (2a) , which we call the exploding soliton, is not complete, since tan( ar) ---. 00 at a finite distance away from the origin. The potential function is f (r ) = - 2 log cos (ar) and the Gauss curvature is K = -2a 2 sec 2 (ar) < O. The above steady gradient solitons are defined on topological disks. By taking b = -1, we have the following steady solitons on the punctured disk. For a = -a 2 , taking w(r) =
!a coth(ar) yields
9 = dr 2
1
2 (ar)d0 2 , + 2"coth a
which has potential f(r) = -2logsinh(ar) and K = -2a 2 csch2 (ar) < O. For a = a 2 , taking w(r) =
!a cot(ar) yields
9 = dr 2
1
2 (ar)d0 2 , + 2"cot a
which has potential f (r) = 2 log sin( ar) and curvature K = 2a 2 cot( ar) > O. Both of the above metrics are incomplete because of their behavior near r = O.
3.
WARPED PRODUCTS AND 2-DIMENSIONAL SOLITONS
Likewise, taking b
15
= 0, we have the ODE w' = aw 2. If a = -0: 2, then
w(r) = 0:2;+C' f(r) = -2 log (0:2r+C) and K=
-(r+C:-2)2'
EXERCISE 1.22. Show that if w is a solution to (1.42), i.e., w' - aw 2 = b, then v ~ 1/w satisfies v' +bv 2 = -a. Related to the above exercise, in Section 9 of this chapter we shall see Buscher duality exhibited in the above solitons. EXERCISE 1.23. Determine all of the solutions of the rotationally symmetric steady gradient Ricci soliton equation on a surface. 3.4. A metric transformation on surfaces related to circle actions. It is interesting to search for duality transformations, besides Buscher duality, for Ricci solitons. One transformation for rotationally symmetric metrics on surfaces is the following, which is related to the study of collapsing sequences of solutions of the Ricci flow on 3-manifolds. Given a rotationally symmetric metric 9 = dr 2 + w (r)2 d(P on a surface M2, we may define another metric on M2 by
9 ~ dr 2 +
w (r 0:2w (r)
r
+1
d0 2.
A more general transformation is given by Cheeger [69J. The inverse transformation is g=dr 2 + w(r)2 d0 2 . 1-0:2w(r)2' which is defined as long as 0 < w (r) < 1/0:. EXERCISE 1.24. Consider the SI action on (M2,g) x SI (1/0:), where SI (p) ~ lR/271pZ, defined by ¢(r,O,"1)~(r,O+¢,"1+¢),
Show that
¢ESI.
9 is the quotient metric on [(M2,g) x SI (1/0:)] /SI.
SOLUTION. See Proposition 2 of [103J. We now consider some examples. The inverse transformation of the cigar, ~ . 2( o:r )dO 2, 9 cig = dr 2 + 21smh 0: is the hyperbolic metric of constant curvature K == -0: 2. The transformation of the exploding soliton, gxpd = A
dr 2 + 21sm . 2( o:r )d02 ,
0: is the spherical metric of constant curvature K == 0: 2. Since gxpd is only defined for 0 < r < 7r / (20:) , we see that 9xpd is a hemisphere. In summary, we have the following types of examples:
1. RlCCI SOLITONS
16
(1) a constant negative curvature metric on a surface transforming to a steady soliton, (2) an incomplete steady soliton on a surface transforming to an incomplete metric with constant positive curvature. 3.5. Classifying 2-dimensional solitons. PROPOSITION 1.25 (Surface solitons conformal to ]R2). The only complete steady gradient solitons conformal to the standard metric on ]R2 are the cigar and the fiat metric.
1.26. We have not assumed the curvature is bounded or has a sign. Note that since a steady gradient soliton is an ancient solution, by applying the maximum principle to the evolution equation for the scalar curvature, one sees that a complete steady Ricci soliton on a surface with curvature bounded from below is either flat or has positive curvature (see [111]). For a similar result to Proposition 1.25, see Corollary 1.28 below. REMARK
Consider the pullback of the steady gradient soliton metric under the map from the cylinder 51 x ]R to ]R2 defined by PROOF.
x = eU cosv,
y = e U sinv,
where (u, v) are coordinates on 51 x ]R and 51 = ]R/27l"Z. The pullback is a steady gradient soliton on the cylinder, and the vector field X that it is flowing along is conformal. Hence, the complexification6 of X is a holomorphic vector field and is of the form h(w){)I{)w, where w = u + iv and h(w) is a 27l"-periodic entire (analytic) function. Since X is the gradient of a function f, then X can have no closed orbits, and hence any zeros of h must be simple (by virtue of appealing to a local power series expansion). Consequently, critical points of f can only be local maxima and minima, so that indexp (\7 f) = 1 at any critical point p. Since X (51 x]R) = 0, the Poincare-Hopf Theorem, which says that the sum of the indices of \7 f is the Euler characteristic, implies that f has no critical points, hence h has no zeros. The periodicity then implies that h is a constant. Moreover, X must be a constant times 01 ou, since otherwise the corresponding vector field X on ]R2 has orbits which spiral into the origin. Hence we know that X = I up to multiple, and then by the argument of subsection 3.1 of this chapter, J(X) = 0100 is a Killing field, and so by Lemma 1.18 the metric on ]R2 is rotationally symmetric (and a warped product). Then we may appeal to the above calculation of solutions 0 to (1.40).
ro or
6The complexification of a vector field X is defined + iJ (X)). For example, the complexification of r =
~ (X
~
gr
to
xx 88
+
be y
gy
X(l,O)
is
Z
gz
(x :x + y gy + i (Y go; - xgy )) . See the next chapter for a more detailed discussion of
Kahler manifolds (real surfaces are I-complex dimensional Kahler manifolds). Since we are on a surface, the complexification of a conformal vector field is a holomorphic vector field.
4.
CONSTRUCTING THE BRYANT STEADY SOLITON
17
We have the following from Theorem 13 on p. 61 and Theorem 10 on p. 55 of Huber [210] (see also §10 of Li [251]). THEOREM 1.27 (Conformal structure of surface with finite total curvature). If (M2,g) is a complete Riemannian surface with fM K_dl1 < 00, where K_ ~ max {-K,O} , then (M,g) is conformal to a closed Riemannian surface with a finite number of points removed. Furthermore, by the Cohn- Vossen inequality
1M K+dl1 ~ 1M K- dl1 + 27rX (M) <
00,
where K+ ~ max {K,O} . By the first part of the above theorem, a complete noncom pact surface with positive curvature, which we know is diffeomorphic to the plane, must be conformal to the plane. From this and Proposition 1.25 we conclude the following.
°
COROLLARY 1.28 (Steady surface soliton with R > is cigar). If (M2, g) is a complete steady gradient Ricci soliton with positive curvature, then (M, g) is the cigar. This result gives us a classification of complete steady Ricci solitons on surfaces with curvature bounded from below since such solutions are either flat or have positive curvature.
4. Constructing the Bryant steady soliton We may generalize the cigar metric to a rotationally symmetric steady gradient Ricci soliton in higher dimensions on ]Rn+l by setting N = sn, the unit sphere with constant sectional curvature +1. 7 As the following calculations parallel unpublished work of Robert Bryant for n = 2, we will refer to the complete metrics obtained as Bryant solitons. The Bryant soliton is a singularity model for the degenerate neckpinch, a finite time singularity which is expected to form for some (nongeneric) initial data on closed manifolds. (See Section 6 in Chapter 2 of Volume One.) REMARK 1.29. A singularity model is a long existing solution (Le., the time interval of existence has infinite duration) of the Ricci flow obtained from a singular solution (Le., a solution defined on a maximal time interval [0, T)) of the Ricci flow as the limit of dilations about a sequence of points and times approaching the singularity time T. 7For background on Einstein metrics which are warped products over I-dimensional bases, see the notes and commentary at the end of this chapter.
1. RICCI SOLITONS
18
4.1. Setting up the ODE for Bryant solitons. With p = n - 1, substituting (1.38) and (1.39) into the steady gradient soliton condition Rc(g) + \1\1 f = 0 gives the following system of ODE for wand f:
" !" = n~, w
(1.45) Since R !:If -
=
WW ' f'
= ww" -
(n - 1)(1 - (w ' )2).
-!:If for a steady gradient soliton, Proposition 1.15 implies that
1\1 fl2
= C, a constant. On the other hand, tracing (1.39) gives
!:If
f'
=!" + n~. w I
Using this expression for !:If, we obtain w 2!"
Eliminating (1.46)
+ nww' !, -
w 2 (I') 2
= Cw 2.
f" and w" using (1.45) gives a first integral of our ODE system: 2nww' !, + n(n - 1)(1 - (W')2) - w 2(f')2 = Cw 2.
The analysis of solutions to (1.45) is simplified by using variables that are invariant under the symmetries of the system (Le., translating r, translating f, and simultaneously scaling rand w). We choose new dependent variables x and y and independent variable t (not to be confused with time), such that •
X'T
• I -w f' , Y'Tnw
I
W ,
dr
dt~-.
w
From (1.46) we have nx 2 - y2
(1.47)
+ n (n -1) = Cw 2.
Then by (1.45)
dx
2
dt = ww" = ww' !, + (n - 1)(1 - (w') ),
dy dt Thus the
ODE
I
= wy =
I
I
-ww f .
system (1.45) becomes
dx
dy dt = x(y - nx). ' Substituting C = 0 in the first integral (1.47) gives an invariant hyperbola y2 - nx 2 = n(n - 1) for the system (1.48); see Figure 1 on the next page. (1.48)
REMARK
-
dt
1.30. When n = 1, equation (1.48) becomes dx dt = x (x - y),
(1.49)
which implies x
~; = 2x (x -
2
=x -xy+n-1
+y =
dy dt = x(y - x),
const. Assuming, x, y
~
1 as t
~ -00,
we have
1) . The solution of (1.49) with x (t) decreasing is given by 1
x
= 1 + e2t '
1
y
= 2 - 1 + e2t .
4.
CONSTRUCTING THE BRYANT STEADY SOLITON
19
FIGURE 1. Phase portrait for system (1.48), n = 2, drawn using Maple.
Since ftlogw
= x,
we have w
r
=
= (e- 2t + 1)-1/2.
J
wdt
= arcsinh
= In ( et + J 1 + e2t ) (e t ) .
Hence et = sinh r, so that w (r) = (csch 2r the cigar soliton. EXERCISE 1.31. For n and x (t) increasing?
=
Moreover
r 1/2
+1
= tanh r, and we have
1, what is the solution with x, y
~
1 as t
~
-00,
We will establish the existence of a complete steady gradient Ricci soliton on ]Rn+l by doing phase plane analysis on the system (1.48). Note that the stationary solutions of (1.48) satisfy y = nx and x 2 - 1 = 0, so that the stationary points are (x, y) = (1, n) and (x, y) = (-1, -n). These are both
20
1. RICCI SOLITONS
saddle points since the linearization of (1.48) at (1, n) is8 du dt = - (n - 2) u - v, dv
- = -nu+v dt
'
and the negative of this at (-1, -n) Since we want the metric to close up smoothly as r ---t 0, by Lemma A.2, we will need
.9
x
---t
1 and
y
---t
n.
Therefore, we will limit our attention to the two trajectories that emerge from the saddle point (1, n) as t increases. (These, plus the point itself, comprise the unstable manifold of the saddle point; the stable manifold is the hyperbola.) Given either of these trajectories, we can reconstruct functions w, rand f of t by successively integrating dw (1.50) - = x dt, dr = w dt, df = (nx - y)dt. w Furthermore, we can choose constants of integration such that w '\. 0, r '\. 0, and df /dr ---t 0 as t ---t -00. It then follows that f(r) and w(r) smoothly extend to even and odd functions of r, respectively (cf. [217], Proposition 5.2). 4.2. Phase plane analysis of the right-hand trajectory. The linearization of (1.48) at the saddle point shows that the right-hand trajectory emerges from the saddle point into the region where dx / dt > 0 and dy / dt < o. Because trajectories along the boundary of this region only point into the region, x continues to increase and y continues to decrease. Linearization also shows that y - nx is initially negative along the trajectory, and the equation d dx -(y - nx) = x(y - nx) - n dt dt shows that y - nx is negative and monotone decreasing. Because x ~ 1 and y - nx < -C for some positive constant C when t is sufficiently large, 8The linearization of the general system dx dt = f(x,y),
dy dt=g(x,y)
at (x,y) is di dt
= ax (x,y)x+
of
_
of _ oy (x,y)y,
djj dt
= ax (x, y) x + oy (x, y) y.
og
_
og
_
9The eigenvalues of the matrix
are 2 and 1- n, with corresponding eigenvectors (-l,n) and (1,1), respectively.
4. CONSTRUCTING THE BRYANT STEADY SOLITON
21
the differential equation for Y in (1.48) shows that there exists to such that Y :::; 0 for t 2: to. Therefore,
dx 2 ->x +n-1 dt and limt/T x =
+00 for some finite T > to.
PROPOSITION
1.32. The metric associated to this trajectory is incom-
plete. PROOF.
(1.51)
We will show that there exists a constant a E (0,1) such that a x<-- T-t
for t sufficiently close to T. From this inequality, it follows by integration that w :::; C (T - t) -a for some constant C and hence that r = w dt is finite as t /' T. To establish the upper bound on x, suppose we knew that there is a positive constant 0 and a tl E (t, T) such that
J
(1.52)
y
Then dx/dt 2: (1
+ Ox :::; 0 for
+ 0)x 2 + n -
d -arctan dt
all t 2: tl.
1, and this implies that
(x) > ~-1 -Ja(n-1) a'
a ~ 1/(1 + 0).
Integrating from t to T on both sides and solving for x give the inequality (1.51). To establish (1.52), note that d dt (y
+ ox) = x(y - nx) + 0(x 2 - xy + n -
1).
When y + Ox = 0, the right-hand side equals (0 2 - n)x 2 + (n - 1)0. Let Xo be the x-intercept of the trajectory, and let 00 > 0 be small enough so that (1.53)
(05 - n)x 2
+ (n -
1)00 < 0 for all x > Xo.
Let (Xl, Yl) be any point on the trajectory such that and choose 0 :::; 00 sufficiently small so that Yl + OXI :::; that this inequality persists for all later values of t.
Xl
o.
> Xo and Yl < 0, Then (1.53) shows D
4.3. Phase plane analysis of the left-hand trajectory. We now turn to the analysis of the other trajectory emerging from saddle point (1, n). By the same reasoning as before, we see that x is decreasing, Y is increasing, and Y - nx is positive along this trajectory. Moreover,
x
---4
0+
and
Y ---4
+00
as t increases. In order to prove that the corresponding metric is complete, we need to know the limiting behavior of wand r. To do this, we introduce
22
1. RICCI SOLITONS
new phase plane coordinates
Y~ y'n(n-l).
x X ~ y'n-, y As x
---t
0+ and y
---t
y
+00, X
---t
0+
and
Y
---t
0+.
(The constants are chosen so that the invariant hyperbola becomes the unit circle in the XY-plane.) With a new independent variable s such that ds ~ y dt, we have
~~ = X 3 -
(1.54)
X
~: = Y
+ ay2,
(X2 - aX) ,
where a ~ 1/ y'n. In these coordinates, the first integral corresponds to a Lyapunov function L ~ X2 + y2 which satisfies dL/ds = 2X2(L -1). Since L < 1 for this trajectory, L is strictly decreasing but dL/ds 2 2L(L - 1). So, the trajectory approaches the origin exponentially as s ---t +00. Since (1.55)
= xdt =
d(logw)
1 dY y'nX -1 Y'
it suffices to find the limiting behavior of X in terms of Y. Note that for X and Y small, (1.54) implies dX 2 r::::: -X+aY ds '
dY
ds
r:::::
as s
---t
-
In particular, we expect -X following. LEMMA
+ ay2
---t
°
-aXY.
+00. Indeed, we have the
1.33.
and
where a =
l/vin.
PROOF. One might try to apply I'Hopital's Rule to find the limit of X /y2. However,
(1.56)
2
tsx _~ (X2X2- - aX1 ) +
.!l.. y2 - y2 ds
1 'nX2 - X
v'·
shows that this is not straightforward, since the limit of X/Y 2 is also involved in the right-hand side. Let g(s) = X and h(s) = y2 and apply the Cauchy mean value formula (see, e.g., [10]):
g'(c)(h(a) - h(b)) = h'(c)(g(a) - g(b))
4. CONSTRUCTING THE BRYANT STEADY SOLITON
23
for some c between a and b. Letting b --t +00 gives g'(c)jh'(c) = g(a)jh(a). Substituting in from (1.56) gives
~ (x~2_-a~) + v'nX ; - X Is=c =
2:
2
Is=a'
Multiplying through by X 2 - aX (evaluated at c) and taking a (which forces c --t +00) give
X . 11m y2 +a
s ...... +oo
--t
+00
= O.
However, to get zero on the right-hand side, we need to first know that Xjy 2 is bounded as s --t +00. We know that Xjy2 2 O. To obtain an upper bound, we use the differential equation :s
(~)
=a -
~ (X2 -
2aX + 1)
.
Since the trajectory approaches the origin in the XY-plane exponentially, there exists a k > 0 such that d (X) ( -X - a ) (e-k s -~ -~ < - ) 1 + ae- ks ~ .
Comparing with the solution of the
ODE
dujds = u(e- ks - 1) + ae- ks
shows that Xjy 2 is bounded above. In particular, for s 2 So
~ (s) ~ u (s) = e-(ke-kS+ s)
(1:
eke-kS+Sae-ksds + e-k
~ (so))
.
The second limit, which gives the next term in the power series expansion of X in terms of Y near the origin, follows by applying the same arguments to the equations
d (X -y4a y2 )
ds
(
2
= -1 - 3X + 4aX
)
(X _y4aY2)
- ~ y2 (X2 -2aX)
and
o EXERCISE
1.34. Verify the second limit.
24
1. RICCI SOLITONS
Now we derive the asymptotic behavior of wand r. We have d X dy(logw+logY) = Y(X-a)'
which by Lemma 1.33 approaches zero as Y ~ O. Hence, wY approaches some positive constant G as Y ~ O. Next, note that dr
dY
w
vn(n - 1)(X2 - aX)
shows that r is monotone increasing as Y decreases to zero, because X - a = a(nx - y)/y is negative along this trajectory. Furthermore, the limiting behavior of X and w implies that there is a constant c such that dr
w
dY
vn(n - l)X (a - X)
-- =
>
G
----===-----
v'n"=1(1 + c)y3
for Y sufficiently close to zero. We can similarly find a constant c' such that dr dY
2(1
+ c')G
--<----';:==~
v'n"=1y3
for Y sufficiently close to zero. Thus, we see by integration that r is O(y-2) as Y approaches zero. In particular, r ~ +00. Therefore, the metric is complete. This proves the following. THEOREM 1.35 (Existence of Bryant soliton). There exists a complete, rotationally symmetric steady gradient soliton on ~n+1 which is unique up to homothety. REMARK 1.36. Since w = 0 (y-l) and r = O(y- 2 ), we have w (r) = o (rl/2) . This tells us the Bryant soliton is like a paraboloid. Throughout this section, by x = 0 (yP) we mean there exists a positive constant G such that G-1yp :::; x :::; Gyp. This is an abuse of notation since usually x = 0 (yP) means Ixl :::; GyP for some G. 4.4. Geometric properties of Bryant solitons. It is interesting to contrast the asymptotic behavior of the curvature for these metrics with that of the cigar metric. Recall from Chapter 2 of Volume One that the Gauss curvature of the cigar falls off exponentially as a function of the distance to the origin; in fact, since w(r) is asymptotically constant as r ~ 00, the cigar is asymptotic to a cylinder. On the other hand, in a higher-dimensional warped product, the n-dimensional volume of the sphere at distance r from the origin is w(r)nvol(sn), where vol(sn) is the volume of the standard nsphere. Since w(r) = O(rl/2), we see that the sphere volume is unbounded. From the limits in Lemma 1.33, we obtain the asymptotics of the derivatives of w as functions of r: Wi
= X = v'n -
1 X/Y
= O(Y) = O(r-l/2)
25
4. CONSTRUCTING THE BRYANT STEADY SOLITON
and
w"
l'
~
-
xy +n -1 w
= 0(_y3) = 0(_1"-3/2)
(1.57)
as
x2
00.
A moving frames calculation like that mentioned above lO shows that the sectional curvatures of the Bryant solitons are (1.58)
VI
=
1- (w'? W
V2
2'
w"
= --, w
where VI is the curvature for planes tangent to the spheres, and V2 for planes tangent to the radial direction. In particular, 2VI is the eigenvalue of the curvature operator corresponding to the eigenspace El ~
{Q 1\ ,B : Q,,B E AI (sn
X
{1"})}
and 2V2 is the eigenvalue corresponding to the eigenspace
E2 ~
{Q 1\ dr : QE Al (sn
X
{r})} .
From the asymptotic behavior of wand its first two derivatives, we obtain VI = 0(1'-1) and V2 = 0(1'-2). So, the sectional curvatures decay inverse linearly in terms of the distance from the origin. Like the cigar, the curvature is positive: LEMMA 1.37 (Curvature of Bryant soliton). The curvature operator of the Bryant soliton is strictly positive (i. e., VI > 0 and V2 > 0) away from the origin, and these curvatures have a positive limit at the origin. Moreover, VI
= 0(1'-1) and
V2
= 0(1'-2).
PROOF. The positivity of VI for l' > 0 follows from the fact that w' = x is strictly less than 1. To show that V2 is positive for l' > 0, we will show that w" is negative. From (1.57), it suffices to show that x 2 - xy + n - 1 is negative, or equivalently that X 2 - y'riX + y2 is negative. Linearizing the system (1.48) about the saddle point shows that x 2 - xy + n - 1 is initially negative when the trajectory emerges. Moreover,
.!!..-(X2 _ ds
Vn + y2) = X2(X2 + y2 -
1) + (X2 _ 1)(X2 - ynX
+ y2)
shows that if X 2 - y'ri + y2 were ever equal to zero in the middle of the trajectory, it would have a negative derivative. So, w" is negative for all l' > O. At the origin, all sectional curvatures are equal. The scalar curvature R satisfies R + IV' fl2 = C, and we know C must be positive because
R = n(n -
l)VI
+ 2nv2
lOSee, for example, the solution to Exercise 1.188 in [111].
1. RICCI SOLITONS
26
is positive everywhere else. The curvatures approach zero as r ---t +00, so R has a positive maximum value at some point. Because f' = (y - nx)/w is positive for r > 0, this must occur at the origin. D We end this section by noting that the preceding construction can be generalized to the case where the fiber is a product of an Einstein manifold and a sphere (see [219]). THEOREM 1.38 (Steady solitons on doubly-warped products). Given an Einstein manifold (Nm, 9N) with positive Ricci curvature, there exists a 1parameter family of complete steady gradient soliton metrics on ~n+l x N in the form of doubly-warped productsll,' 9 = dr 2 + v(r)2gcan where
gcan
is the standard metric on
+ w(r)2gN,
sn, n ~ 1.
These solitons have positive Ricci curvature when r > O. However, the sectional curvature along the copies of N can be negative.
5. Rotationally symmetric expanding solitons In Section 4 in Chapter 2 of Volume One, we described the construction of a I-parameter family of rotationally symmetric expanding gradient Ricci solitons on ~2 (see also Gutperle, Headrick, Minwalla, and Schomerus [174]). These solitons are asymptotic to a cone (with any cone angle in the interval (0,27T') possible) and have curvature exponentially decaying as a function of the distance to the origin (see Exercise 4.15 and Corollary 9.60 of [111]). In this section we consider the problem of constructing rotationally symmetric expanding gradient Ricci solitons on ~n+l for n ~ 2. In particular, we generalize the construction from the previous section on the Bryant soliton to show that there exists a I-parameter family of rotationally symmetric complete expanding gradient solitons on ~n+l.
5.1. Setting up the ODE for expanding gradient solitons. To start, we begin with the expanding gradient soliton condition
Rc(g)
+ VV f + Ag =
0,
E
A = "2 > O.
Substituting (1.38) and (1.39) into this equation gives
o=
-n~" dr 2 + (p - ww" - (n - I)(w')2) 9 w
+ (I" dr 2 + ww' fig) + A (dr2 + w 2g) , is the warping function and 9 is the metric on sn with Einstein
where w( r) constant p. Taking p = n - 1 and collecting the components of the metric
llThis construction was inspired by a similar construction for Einstein metrics by Berard-Bergery [24].
5. ROTATIONALLY SYMMETRIC EXPANDING SOLITONS
on jRn+1, we get a system of two second-order of r:
f" + >. = nw" /w,
ww' f'
ODES
for
27
f and w as functions
+ >.w 2 = ww" + (n -
1)((w' )2 - 1).
Again, Proposition 1.15, together with equation (1.25), provides a first integral for this system:
f" + n(w' f' /w) -
(f')2 - 2>.f = C.
We can reduce the order of the system by again introducing variables that are invariant under the symmetries of translating r and translating f:
Y =;=• nw I - w f' .
. I x=;=w,
Because simultaneous scaling of rand w is no longer a symmetry of the system, we must also retain w as a variable, yielding the following firstorder system:
= xw, dx/dt = x 2 - xy + >.w 2 + n -1, dy/dt = x(y - nx) + >.w 2 ,
dw/dt (1.59)
where the parameter t is related to r by dt = w- 1 dr. (Notice that the first integral is not invariant under the translation symmetries, so it does not give an invariant manifold for this system.) Clearly, solutions of the steady soliton system (1.48) are also solutions of (1.59) for which w = O. (Of course, in the steady case, the warping function is not identically zero, but it is recovered by the quadrature d(log w) = x dt.) To get a metric that closes up smoothly at the origin, we need a trajectory that emerges from the singular point P = (0,1, n) as r increases. (We take w,x,y, in that order, as coordinates on the phase space JR.3.) Linearization at P shows that there is a 2-dimensional unstable manifold passing through P, whose tangent space at P is spanned by the vectors (1,0,0) and (0, -1, n)P The trajectory corresponding to the steady soliton lies in this surface and is tangent to (0, -1, n) at P. Of course, we are only 12The linearization of (1.59) at P is diD
_
ill =W, dx
dt
= (2 -
d-y
= -nx- +-y.
dt
n) x - ;ii,
Note that the last two equations appeared in the linearization for the steady soliton equation and the eigenvalues of the matrix
-1) ( 2-n -n 1
are 2 and 1- n, with corresponding eigenvectors (-l,n) and (1,1), respectively.
28
1. RICCI SOLITONS
interested in solutions in the half-space where w is positive. There is a 1parameter family of these trajectories in the unstable manifold, and they are tangent to the vector (I, 0, 0) as they exit from P. (This vector is associated to the positive eigenvalue closest to zero.) Among these is the hyperbolic metric, for which f' = 0, y = nx, and the warping function is w
= Vnj>.sinh (V>.jnr) ,
y
= ncosh (V>.jnr) .
5.2. Analysis of a I-parameter family of trajectories. For the rest of this section, we will consider only the I-parameter family of trajectories that lie in the quadrant of the unstable manifold of P, between the hyperbolic metric and the steady soliton, i.e., those for which wand y - nx are positive for small values of r. Among these is the flat Gaussian soliton on lR.n +1 described in Example 1.9; the corresponding solution of (1.59) is
w = r,
x = 1,
y=n
+ >.r2.
In order to show that the metrics corresponding to our family of trajectories are complete and to study their curvatures, we again introduce rescaled coordinates
w W::§=-,
. vn(n - 1)
x X::§=vn-,
y
y=;=
y
y
We introduce a new independent variable s such that ds system (1.59) becomes
.
= y dt, whereupon
d: = W(X2 _ >'W2), dX ds = X 3
(1.60)
-
X
~~ = Y(X2 -
+ ay2 + >.( vn -
X)W2,
a::§= Ijvn,
aX - >'W2).
In these coordinates, P = (0, a, VI - ( 2 ), and the hyperbolic trajectory has X identically equal to a, approaching the critical point H = (Ij-Jri"X, a, 0) as r --t 00. The trajectories we are considering lie in the unstable manifold of P, and for small r, in the region where X < a and W > o. The flat trajectory satisfies Y = v'11=1 X and X 2 - aX + >.W 2 = O. LEMMA
1.39. These trajectories remain in the region defined by
0< W:S Ij~.
0< X < a, PROOF.
Along the plane where X = a,
dXjds = a 3
-
a
+ ay2 + >'(vn -
a)W2.
Let Q stand for the quantity on the right, which must be negative for r small. The following equation shows that Q remains negative if X < a:
dQjds = aXy2(X - a)
+ >.( vn -
a)(X2 - ( 2)W2 - >.QW 2.
5.
ROTATIONALLY SYMMETRIC EXPANDING SOLITONS
29
This establishes that X < a always along these trajectories, and the equation dX/ds = ay2 + AVnW 2 when X = 0 shows that X remains positive. Let L = X 2 + y2. Then dL/ds
shows that L
~
= 2(L -
1)(X2 - AW2)
+ 2AVnW2(X -
1 is preserved. Finally, when X
a)
< a,
dW/ds ~ W(a 2 - AW 2),
showing that
W cannot exceed a/../'X = 1/vn:>...
D
LEMMA 1.40. As s ~ +00, these trajectories approach the origin in W XY coordinates, and the corresponding metrics are complete. PROOF. Because the right-hand sides of the system (1.60) are polynomial, parameter s is unbounded along these trajectories. Because the trajectories are bounded within the region given by Lemma 1.39, each trajectory must limit to either the critical point H or to the origin 0 as s ~ +00. However, linearization shows that the tangent space of the stable manifold of H intersects the closure of the region given in Lemma 1.39 only in the line through H tangent to the hyperbolic trajectory. Thus, all the trajectories under consideration limit to the origin. To show that the metrics corresponding to these trajectories are complete, note that (1.61 )
J J dr =
W ds =
J
X2
~~W2 .
Along these trajectories, W must eventually become a decreasing function of s. Then it suffices to show that there is a point (Wo, Xo, Yo) along one of our trajectories such that the limit
fWO
!~Je
dW AW2 - X2
is infinity; this follows from AW2 - X2 < AW2.
D
The flat trajectory divides our chosen quadrant of the unstable manifold near S into regions of negative sectional curvature (bordering the hyperbolic trajectory) and regions of positive sectional curvature (bordering the steady soliton trajectory). The following lemma shows that these signs persist; thus, there exist I-parameter families of rotationally symmetric expanding solitons of strictly positive and of strictly negative sectional curvature. PROPOSITION 1.41. Excluding the fiat metric, the sectional curvatures of these metrics are either strictly positive or strictly negative for all r.
30
1. RICCI SOLITONS
PROOF. VI =
In these coordinates, the sectional curvatures are
y2 - (n - 1)X2 1 - (W')2 = -----;---'------:-~;o__ w2 n(n - 1)W2 '
for 2-planes tangent to and perpendicular to, respectively, the orbits of the rotational symmetry. The evolution equations for the numerators are 1d 2 --(Y - (n - l)X 2 ) 2 ds
= (X 2 -,XW 2 -
aX)(Y 2 - (n - l)X 2 )
+ a(n -l)X( _(X2
+ y2 + n,XW2 -
vnX))
and 1d 2 r:::: --(-(X +Y 2 +n,XW 2 -ynX)) 2ds = _(X2 - ,XW2 + ~(aX - 1))(X2
+ y2 + n,XW 2 - vnX) a (Y2 - (n - l)X 2) . + "2X
Observing that the sign of the first nonconstant factor in the second line of each equation is positive, we conclude that the signs VI > 0, V2 > 0 and VI < 0, V2 < 0 are both preserved. 0 We will now obtain the asymptotic behavior of the curvature relative to distance r. PROPOSITION 1.42. Assuming that the sectional curvatures are negative, then they both decay at least quadratically in r, and for large r, the warping function w(r) is bounded between two linear functions of r. PROOF. First, we establish that, as the trajectories approach the origin, the distance r is asymptotic to a constant times l/W. We will do this by establishing a positive lower bound for W 2 / X, by examining the equations
~ ds
(W2) = W 2 (X2 _ ,XW2 + 1- (ay2
X
X
+ ,Xy'nW2))
X'
~
(Y2) = y2 (X2 _ ,XW2 + 1- 2aX _ (ay2 dsX X
+ ,Xy'nW2))
X·
For r sufficiently large, we can bound the nonconstant terms in the parentheses (those that do not involve dividing by X) by a small number c. Then, if we set Y ~ ay2/X and W ~ 'xy'nW 2/X, we have
:s (W
+ Y) ~
(W
+ Y)
(1 - c -
Y-
W) .
Hence, W + Y ~ c for some positive constant c. On the other hand, because sectional curvature VI is negative, y2 / X ~ (n-1)X, and so lims -++ oo Y = o.
5. ROTATIONALLY SYMMETRIC EXPANDING SOLITONS
31
Hence W 2 / X 2: C1 for some positive C1, for r sufficiently large. Thus, the denominator in the integral (1.61) is bounded below, 2
2
AW - X
W4
2
2: AW - (C1)2'
and it follows that r = O(I/W). The orbital sectional curvature satisfies 111
=
y2-(n-l)X2 X2 W2 > --- > --n(n - I)W2 nW2 n(ct)2
and so decays at least as fast as 1/r2. The equation 112
(1.62)
+ (n -
1)111
=-
(X2 - aX + AW2) W2
shows that in order for 112 to also decay to zero, it is necessary that lim X/W 2 s---+oo
(1.63)
=
AVn.
We already know that X/W 2 is bounded above, and we can then apply the Cauchy mean value theorem (as in Lemma 1.33) to the equation 2
.!LX ds
-
dW2 -
~
(X2 - AW 2 + a~
-1) + AVn
-----'-------;::-------:--::-::-,.".---''----
X2 -AW2
ds
to obtain the desired limit. Next, dividing (1.62) by W 2 gives -112 111 W2 = (n - 1) W2
X2
+ W4 +
A - a~ W2 .
In order to show that 112 decays at least as fast as 1/r2, we need only show that the last term is bounded. This follows from the differential equation
:. C-:l) ~(_1+2,\W2 + A(X -
-3X2)
(,,-:l)
X
y2
a) W2 - a 2 W4'
where the boundedness of the terms on the second line follows from the limit (1.63) and the negative curvature condition y2 ~ (n - I)X2. To establish the last assertion, it suffices to show that dw / dr = x = v:n=T X/Y is bounded above and below for r sufficiently large. First, because of (1.62) we can say that -d (W2) = (X2
ds
Y
+ aX -
W2 AW2)-
Because of the limit (1.63), d W2 - l o g - < CW4 ds Y-
W2 < 2X2_.
Y-
Y
32
1.
RICCI SOLITONS
for some constant C > O. Because dr = W ds and W = 0 (1/ r ), then d W2 - l o g - < Cr- 3 dr Y ,
from which it follows that W 2 /Y is bounded above for large r. Thus, there are positive constants Cl, C2 such that
for sufficiently large
T,
and the bounds on X/Y follow.
o
Much more detailed information about the solutions in the positive and negative curvature cases (as well as the steady case) has been obtained by Robert Bryant (in unpublished notes, using different coordinates). In particular, Bryant proved PROPOSITION 1.43 (Bryant). Each positive curvature solution defines a complete, rotationally symmetric expanding soliton, whose sectional curvature decays proportionally to 1/r2. 6. Homogeneous expanding solitons 6.1. Existence. Homogeneous spaces are among the nicest examples of Riemannian manifolds. 13 Einstein metrics are among the nicest examples of Riemannian metrics. It is thus natural to ask if a given homogeneous space M n = G / K admits a G-invariant Einstein metric. If M is closed, the answer is frequently 'yes'. For example, consider the following result of B6hm and Kerr [29].
THEOREM 1.44. Let Mn be a closed, simply-connected homogeneous space. If n < 12, then M admits a homogeneous Einstein metric. On the other hand, Wang and Ziller's example SU(4)/SU(2) of a 12dimensional homogeneous space that admits no homogenous Einstein metric shows that the answer is 'no' in general and that the dimension restriction above is sharp [365]. Every known example of a noncompact, nonfiat homogeneous space that admits an Einstein metric is isomorphic to a solvable Lie group S. (See [198] and [317].) Moreover, for all known examples, the Einstein metric is of standard type. DEFINITION 1.45. A left-invariant metric g on a solvable Lie group S, regarded as an inner product on the Lie algebra.5, is said to be of standard type if the orthogonal complement with respect to g of the derived algebra [.5,.5] forms an abelian subalgebra a of .5. 13See Chapter 7 of [27], for example.
6. HOMOGENEOUS EXPANDING SOLITONS
33
Many noncom pact homogeneous spaces admit no Einstein metrics whatsoever. For instance, there is the following result of Milnor [266, Theorem 2.4]. THEOREM 1.46. If the Lie algebra of a Lie group N is nilpotent but not commutative, then the Ricci curvature of any left-invariant metric on N has mixed sign.
The Ricci soliton structure equation (1.64)
Rc
=
-Ag - £Xg
illustrates the sense in which a Ricci soliton may be regarded as a generalization of an Einstein metric. Therefore, it is natural to look for Ricci solitons on homogeneous spaces, such as nonabelian nilpotent Lie groups, that do not admit any Einstein metric. Existence and uniqueness of Ricci solitons on such spaces has been investigated by Lauret [244]. To describe his results, we must recall some notation from the cohomology of a Lie algebra g, relative to its adjoint representation. (See [232] and [369].) A k-cochain is a skew-symmetric k-linear map 9 x ... x 9 --t g. Denote the vector space of all k-cochains by C k , noting the natural identifications CO = g, C 1 = End(g) = g* ® g, and C 2 = A2 (g*) ® g. The Lie algebra cohomology of 9 relative to its Chevalley complex,
Hk(g) = ker(8 k )jim(8 k -
1 ),
is derived from the coboundary operators 15k : C k
8k(A)(X 1 , .•• ,XHr)
--t
CHI defined by
,,+k (-1)1 [Xj,A(X 1 , ... ,Xj, ... ,XHl)]
=
~
A
l~j~k+1
+ l~i<j~k+1
In particular, one has 80(X)(Y) = -[Y, X] = adx(Y) for all X, Y E g, and
81 (A)(X, Y)
= [X, A(Y)]- [Y, A(X)]- A([X, Y])
for all A E End(g) and X, Y E g. A derivation of 9 is an element ofker(8r). Let g be a left-invariant metric on a simply-connected Lie group G. Regarding the Ricci curvature Rc of g as an endomorphism, Lauret considers the condition
(A E JR),
(1.65) which is easily seen to be equivalent to (1.66)
Rc
= -AI +D
(A E JR, DE ker(8r)).
Notice that equations (1.65) and (1.66) relate the geometry of (G,g) to algebraic data of g. This turns out to be a productive point of view. We now survey some of the results it generates, referring the reader to [244] for the detailed proofs.
34
1.
RICCI SOLITONS
One begins with the observation that conditions (1.64)-(1.66) are equivalent for a nilpotent group. LEMMA 1.47. Let N be a simply-connected nilpotent Lie group with Lie algebra n and a left-invariant metric g. Then there exist A E lR and X E n solving the Ricci soliton structure equation (1.64) if and only if there exist A E lR and a derivation D E ker(8d solving (1.66). DEFINITION 1.48. A standard metric solvable extension of (n, g) is a pair (5, g), where 5 = aEBn is a solvable Lie algebra such that [', ']slnxn = [', ']n and 9 is an inner product of standard type such that [5,5]s = n =a..l and
9Inxn =
g.
Recalling Definition 1.45, note that for 9 to be of standard type is equivalent to a being abelian. The following result relates Ricci soliton structures on nilpotent Lie groups with Einstein metrics of standard type on solvable Lie groups. THEOREM 1.49. Let N be a simply-connected nilpotent Lie group with Lie algebra n and a left-invariant metric g. Then (N, g) admits a Ricci soliton structure if and only if (n, g) admits a standard metric solvable extension (5 = aEBn, g) such that the simply-connected solvable Lie group (8, g) is Einstein. Although its proof is nonconstructive, Theorem 1.49 is a very useful criterion. For example, any generalized Heisenberg group [25] and many other two-step nilpotent Lie groups admit a Ricci soliton structure. On the other hand, if n is characteristically nilpotent, then N admits no such structure. (See [244] for more examples.) If a Ricci soliton structure does exist on N, it is essentially unique. THEOREM 1.50. Let N be a simply-connected nilpotent Lie group with Lie algebra n. If g and g' are both Ricci soliton metrics, then there exist a> 0 and TJ E Aut(n) such that g' = aTJ(g). REMARK 1.51. If one regards a Ricci soliton structure as a 3-tuple (N, g, X) satisfying (1.64), then uniqueness of X must be understood modulo addition of a Killing vector field. (See Example 1.57 below.) Some partial answers are known regarding the existence of Ricci soliton structures on broader classes of Lie groups. For example, if G is semisimpIe, then condition (1.66) implies condition (1.64). However, (1.66) has no solutions other than Einstein metrics: THEOREM 1.52. A left-invariant metric 9 on a simply-connected semisimple Lie group G satisfies (1.66) if and only if D = 0, hence if and only if g is Einstein. It is not known in general whether or not a noncompact, semisimple Lie group admits an Einstein metric.
6. HOMOGENEOUS EXPANDING SOLITONS
35
Lauret also gives a variational characterization of Ricci solitons [244]. Roughly speaking, this says that any simply-connected nilpotent Lie group (N, g) admitting a Ricci soliton structure is a critical point of a functional that measures, in a certain sense, how far (N, g) is from being Einstein. We now make this precise. It will be more convenient to fix an inner product and allow the Lie algebra brackets to vary. (Compare with Examples 1.63 and 1.64 below.) Specifically, fix an n-dimensional inner product space (n, (-,.)) and let A = A2 (n*) ® n denote the space of all skewsymmetric bilinear forms on n. (Compare to C2 ;2 im( 01) defined above.) Let N denote the subspace of all nilpotent elements of A that satisfy the Jacobi identity. Then N may be regarded as the space of all nilpotent Lie brackets on n. To each p, EN, associate the simply-connected Lie group NJ.t and the left-invariant metric gJ.t on NJ.t determined by (-, .). Notice that N is invariant under the natural action of the group GL(n). Thus the GL(n) orbit of p, E N corresponds to all simply-connected homogeneous nilpotent Lie groups N isomorphic to NJ.t, and the O(n) orbit of p, corresponds to all (N, g) isometric to (NJ.t, gJ.t). The fixed metric (.,.) on n induces a metric (-,.) on A defined by
(p"v) =
L (p,(Xi,Xj),Xk) (V(Xi,Xj),Xk)
i,j,k
when {Xl, ... , Xn} is any orthonormal basis of n. Let
Using the Ricci endomorphism, Lauret defines a functional F : Nl
---t
lR by
and proves THEOREM 1.53. Let p, E N l . Then (NJ.t, gJ.t) satisfies (1.64)-(1.66) (i.e. admits a Ricci soliton structure) if and only if p, is a critical point ofF.
To interpret this result, observe that E(p,) = IRc(gJ.t) - *R(gJ.t)II 2 measures the trace-free part of the Ricci endomorphism, i.e., how far gJ.t is from being Einstein. But for p, E N l , one has
Hence a local minimum p, of F in Nl should correspond to a homogeneous space (NJ.t' 9J.t) that is closest to Einstein among all near by (Nv, gv).
36
1. RICCI SOLITONS
6.2. Construction. The first explicit examples of Ricci soliton structures on Lie groups were constructed by Baird and Danielo [15] and independently by Lott [256]. We discuss [15] in this section and [256] in subsection 6.3, below. Baird and Danielo discovered soliton structures on 3-dimensional manifolds by studying semiconformal maps to Riemannian surfaces. Significantly, their constructions give the first known examples of nongradient soliton structures. We will only describe two of their conclusions, referring the reader to [15] for details of the method.
1.54. Let N denote the 3-dimensional Heisenberg group nil3 . Recall that N may be represented as the group of upper-triangular matrices EXAMPLE
under matrix multiplication. It is a general fact 14 that any simply-connected nilpotent Lie group is diffeomorphic to ]Rn. So give ]R3 its standard coordinates (Xl, X2, X3) and define the frame field
a
a
a
a
Fl - 2 F2 = 2 ( - - Xl-), F3 = 2 - . - OXl' OX2 OX3 OX3 It is easy to check that all brackets [Fi' Fj ] vanish except [Fl' F2] = -2F3, hence that (Fl' F 2, F3) is a nil 3 -geometry frame. 15 The connection I-forms may be displayed as VFIFI VFIF2 ( VF2H VF2F2 VF3Fl VF3F2 Using the dual field
define a left-invariant metric on N by 9 = 4 (wI ® wI + w 2 ® w2 + w3 ® w3) . Recalling the standard formula
(R(X, Y)Y, X) =
~ I(ad X)*Y + (ad Y)* XI 2 3
- 41[X, Y]I
2
-
I
((ad X)* X, (ad Y)*Y)
2 ([[X, Y], Y], X)
-
I
2 ([[Y, X], X], Y),
it is straightforward to compute that Rc(g) = -2(w l ® wI) - 2(w 2 ® w2) + 2(w 3 ® w3). l4The exponential map of a connected, simply-connected, nilpotent Lie group is a diffeomorphism. For instance, see [200j. l5See Volume One, Chapter 1, Sections 3-4.
6.
HOMOGENEOUS EXPANDING SOLITONS
37
Define a vector field
X
1 2
1 1 -X2F2 - (-XIX2 2 2
= --xIFI -
+ x3)F3.
It is then easy to see that
= £Xg + 3g,
-2 Rc(g)
hence that (N, g, X) is a Ricci soliton structure. REMARK 1.55. Because the I-form metrically dual to X is not closed, it follows that X =1= grad f for any soliton potential function f. REMARK 1.56. Compact locally homogeneous manifolds with nil3 geometry occur as mapping tori of Y A
:
T2
--t
T2 induced by A =
(~ ~)
E
8L(2, Z) with k =1= 0. The left-invariant metric 9 is compatible with any compact quotient, but the soliton structure is never compatible with compactification. Indeed, the scalar curvature of 9 is R = -1/2, while every compact Ricci soliton of nonpositive scalar curvature is Einstein. EXAMPLE 1.57. Let S denote the simply-connected 3-dimensional solvable Lie group sol3 = lR >
F2 =
2(e-Xl~+eXl~), 8X2
8X3
F3 =
2(e-Xl~_eXl~). 8X2
8X3
Its bracket relations are [FI' F2J = -2F3, [F2' F3J = 0, and [F3, FIJ = 2F2. 80 (FI' F 2, F3) is a sol3-geometry frame. The connection I-forms may be displayed as
Using the dual field
define a left-invariant metric on S by
9 = 4(w l ® wI)
+ 8(w 2 ® w2 ) + 8(w 3 ® w3 ).
1. RICCI SOLITONS
38
The Ricci tensor of this metric is simply
Rc(g) = -8(w l ® wI). Given any J-l E
X = J-l
[-H -
~,
define a vector field
e-x1x3F2
+ e-X1x3F3] + (1- J-l) [FI -
eX1 x2F2 - eX1 x2 F3].
The computation implies
-2 Rc(g) = £Xg + 4g, and hence (S, g, X) is a Ricci soliton structure. REMARK 1.58. As in the previous example, X potential function f.
f
grad f for any soliton
REMARK 1.59. Compact locally homogeneous manifolds with so13 geometry are mapping tori of T A : T2 -+ T2 induced by A E SL(2, Z) with eigenvalues A_ < 1 < A+. As in the previous example, the soliton structure cannot descend to any compact quotient, because the scalar curvature of 9 is R = -2.
6.3. Type III singularity models. An important reason for studying shrinking or steady solitons is that they can provide valuable information about finite time singularities of Ricci flow. For example, the (Type I) neckpinch singularity is modeled in all dimensions n 2: 3 by the shrinking gradient cylinder soliton (~ x
sn-l, 9 =
ds 2 + 2 (n - 1) gean, X
= grad(s2/4)).
(See Section 5 in Chapter 2 of Volume One and [7, 8].) The conjectured (Type II) degenerate neckpinch (Section 6 in Chapter 2 of Volume One) is expected to be modeled on the Bryant soliton, discussed above. We shall now see that homogeneous expanding solitons can model infinite time behavior of Ricci flow.
t
E
DEFINITION 1.60. A Type III solution of Ricci flow (Mn,g(t)) exists for [0, 00) (Le., is immortal) and satisfies
sup
tl Rm I < 00.
Mx[O,oo)
There are many examples of Type III solutions (e.g., manifolds with 3 ni1 or so13 geometry) that collapse with bounded curvature. As t -+ 00, these examples exhibit pointed Gromov-Hausdorff convergence to lower-dimensional manifolds. (See Sections 6 and 7 in Chapter 1 of Volume One.) For such solutions, it is not possible to form a limit solution (M~, goo(t)) in a naive way. However, Lott has shown that such solutions may have limits, properly understood, which turn out to be expanding homogeneous solitons [256]. We will describe only two of his results, omitting
6.
HOMOGENEOUS EXPANDING SOLITONS
39
many details. To discuss convergence of Ricci flow solutions, it is necessary to anticipate some material from Chapter 3. In particular, we refer the reader to Definition 3.6 for the notion of Cheeger-Gromov convergence in the COO-topology of a sequence of pointed solutions to the Ricci flow. Now suppose that (M n , g(t)) is a Type III solution of Ricci flow. Fix an origin x E M. For each s > 0, there is a rescaled pointed solution of Ricci flow (M,gs(t),x) defined for t E [0,00) by
1 gs(t) = -g(st). s Lott proves the following: THEOREM 1.61. Let (Mn,g(t)) be a Type III solution of Ricci flow. If the limit (M~,
goo(t), x oo ) = lim (M, gs(t), x) s-+oo
exists, then (Moo, goo(t), x oo ) is an expanding Ricci soliton.
In dimension n = 3, one does not have to assume existence of a limit. THEOREM 1.62. Let (M 3,g(t)) solve Ricci flow on a simply-connected homogeneous space M3 = GIK. Here, G is a unimodular Lie group and K is a compact isotropy subgroup. Then there exists a limit
which is an expanding homogeneous soliton on a (possibly different) Lie group.
Note that the diffeomorphisms with respect to which the convergence of Definition 3.6 occurs become singular as s ~ 00. We will illustrate the content of Theorem 1.62 by relating it to Examples 1.54 and 1.57 (isometric versions of which were also discovered independently by Lott). EXAMPLE 1.63. Let (N, g, X) denote the nil 3 Ricci soliton structure of Example 1.54. With respect to the frame f3 = (Fl' F2 , F3) constructed there, one may regard the fixed metric 9 as the matrix g{3 =
(04 04 0)0 . 004
Now consider the time-dependent frame a(t) = f3A(t) given by
A(t) = (
~ -2a (t) 2
o a(t)
o
a(t))
o , o
a(t) =
J
1 t12
2/ 3
.
1. RICCI SOLITONS
40
With respect to the frame a(t), one obtains the identification ga(t) =
0 0)
4/ 3 !C (
0
o
1C2/3
0
0
1C2/3
.
The limit soliton metric is
goo(t) = 3tga(t)· EXAMPLE 1.64. Let (8, g, X) denote the so13 Ricci soliton structure of Example 1.57. With respect to the frame f3 = (FI' F2, F3) constructed there, regard the fixed metric 9 as the matrix
gp=
(H ~)
Now consider the time-dependent frame a(t)
A(t)
( 0 _1 0)
= a(t)
o
02 0
0
= f3A(t) a(t)
,
given by
=
a(t)
{ft.
With respect to the frame a(t), one gets the identification
(ICo 0 0) I
ga(t)
=
4
0
I 0
0 I lC 4
.
The limit soliton metric is
goo(t) = 4tga(t)· Lott also discovered a 4-dimensional example [256]. We will describe its (time-independent) Ricci soliton structure, leaving construction of the corresponding Ricci flow solution as an interesting exercise for the reader. EXAMPLE 1.65. Let N denote the simply-connected 4-dimensional nilpotent Lie group nil 4 with bracket relations
and all other [Fi' Fj ] = O. The frame field defined in standard coordinates (Xl, X2, X3, X4) on
a F1=-a' Xl
a F2 =-a' X2
a F3 =-a' X3
,J
~ (-~4 -F4 -~4 -~4 ~F2 F3) . 2
0
- F2
0
Fl - F3
by
a a a F4 =Xl+X2+-a a X2 a X3 X4
realizes these relations. The connection I-forms are
(VpF-) =
]R4
F2
PI
0
7.
WHEN BREATHERS AND SOLITONS ARE EINSTEIN
41
Using the dual field define a left-invariant metric on S by
9
= wI
®
wI
+ w2 ® w2 + w3 ® w3 + w4 ® w4 .
Its Ricci tensor is
1 Rc(g) = __ (wI ® wI)
2 Define a vector field
X
1 3 + _(w ® 2
w3 )
-
(w 4 ® w4 ).
= -2XIFI + (-3X2 + XIX4)F2 + (-4X3 + X2X4)F3
- X4F4.
Then one has
LXg
= _2(wl
® wI) - 3(w 2 ® w2 )
-
4(w 3 ® w3 )
-
(w 4 ® w4 )
and thus
-2 Rc(g) = LXg + 3g. Therefore, (N, g, X) is a Ricci soliton structure. 7. When breathers and solitons are Einstein In this section, we review some of the significant results to date on classifying the breathers and Ricci soliton metrics that exist on a given type of manifold. As noted in the introduction, a shrinking or expanding soliton on a closed manifold will evolve purely by diffeomorphisms under the normalized Ricci flow. More generally, we define a breather for the (normalized or un normalized) Ricci flow to be a solution g(t) for which there exists a period T and a diffeomorphism ¢ such that
g(t + T) = ¢*g(t). So a breather solution is a periodic orbit in the space of metrics modulo diffeomorphisms, as compared to a Ricci soliton, which is a fixed point. The following result is the analogue of Proposition 1.13 for breather solutions to the normalized flow (see [218]). PROPOSITION 1.66 (Steady and expanding breathers are Einstein). Any breather or soliton for the normalized Ricci flow on a closed manifold Mn is either Einstein with constant scalar curvature R ::; 0 or it has positive scalar curvature. PROOF. Under the normalized flow, the scalar curvature satisfies aR 0 2 (1.67) -a = b.R + 21Rcl 2 + -R(R - r), t n o
where Rc denotes the traceless part of the Ricci tensor and r is the average scalar curvature. By compactness in space and periodicity in time, there exists a point p E M where R attains its global minimum R min . Since b.R 2: 0 and aR/at = 0 at p, then (1.67) implies that Rmin(Rmin - r) ::; 0,
42
1. o
with equality only if Rc
= o.
RICCI SOLITONS
Thus, if Rmin < 0, then R is constant at this o
time. By applying the same argument at every point, we get Rc == 0 and the metric is Einstein. Otherwise, assume R min = 0; by (1.67), we know that !1R :-::; 0 at p. Let 0 be the open set where !1R < 0 at this time. If 0 is nonempty, then R can only attain its infimum on 0 (which is zero) at a point on a~. Applying the Hopf maximum principle shows that the outward normal derivative of R must be negative there; but this is impossible because V'R = o there. Thus, 0 is empty and !1R 2:: 0 everywhere. Since R achieves its maximum somewhere, by the maximum principle R is constant (in fact, identically zero), and the metric is Einstein by applying the same argument M~~. 0 REMARK 1.67. Note that a breather with positive scalar curvature on a closed manifold is a shrinking breather. In Chapter 6 we will describe Perelman's result that shrinking breathers on closed manifolds are Ricci solitons. The proof of this uses his entropy formula. In Chapter 5 another proof that steady or expanding breathers are Einstein will be given. It WM shown in Proposition 5.10 of Volume One that the only solitons for the normalized flow on compact surfaces were constant curvature metrics. We will now prove the generalization of this to compact 3-manifolds, by first obtaining a curvature pinching estimate for the sectional curvatures. As a corollary to Theorem A.31, we have the following (see Hamilton [186] and one of the authors [218]). COROLLARY 1.68 (Hamilton-Iveyestimate). Assume the normalization infxEM3 v (x, 0) 2:: -Ion the initial metric, where v (x, t) denotes the smallest eigenvalue of the curvature operator. There exists a continuous positive nondecreasing function 'IjJ : IR -+ IR with 'IjJ (u) lu decreasing for u > 0 and 'IjJ (u) lu -+ 0 as u -+ 00, such that for any solution (M 3 ,g (t)) of the Ricci flow on a closed 3-manifold, we have v 2:: -'IjJ (R). That is, Rm 2:: -'IjJ (R) id,
where id : A2
-+
A2 is the identity.
PROOF. By Theorem A.31, wherever v < 0, we have R 2:: Ivl(log Ivl- 3). In particular, if v :-::; -e6 , then R 2:: ~ Ivllog Ivl. The function f (u) = u log u is increMing for u 2:: 1 Ie and hence has an inverse f- 1 : [-1 Ie, 00) -+ [lie, 00). If v :-::; -e6 , then R 2:: 3e 6 and Ivl :-::; 1-1 (2R) . If we let
'l/J (u)
= { f- 1 (2u)
e6
~f u 2:: 3e6 , If u < 3e6 ,
then v 2:: -'l/J (R) at all points in space and time. It is easy to see that the function 'IjJ hM all of the properties claimed in the statement of the corollary. 0
7.
WHEN BREATHERS AND SOLITONS ARE EINSTEIN
43
REMARK 1.69. A version of Corollary 1.68 for the Ricci tensor may be proved directly from the evolution of the curvature operator eigenvalues, using the maximum principle for systems. See [218] for details. COROLLARY 1.70 (Ricci flow on 3-manifolds: R bounds Rm). For any
solution (M3, 9 (t)) of the Ricci flow on a closed 3-manifold, if R is uniformly bounded, then IRml is also uniformly bounded. PROOF. If R :::; C, then 1/ ;::: -C' for some constant C'. If A ;::: J1 ;::: 1/ are the eigenvalues of Rm, then R = A + J1 + 1/ and A :::; R - 21/ :::; C + 2C'. 0
A nice application of the Hamilton-Ivey estimate is the following (see [218]). THEOREM 1.71 (Shrinking breathers on closed 3-manifolds are Einstein).
The only solitons (or breathers) for the normalized Ricci flow on a closed connected 3-manifold M are constant sectional curvature metrics. The example of Koiso's shrinking soliton (see Section 7 of Chapter 2) shows that this result cannot be extended to dimension 4. PROOF. By Proposition 1.66, either the metric is Einstein (which is equivalent to constant curvature in dimension 3) or Rmin > 0; in the latter case, the breather is a shrinking breather. By Corollary 1.68, for the unnormalized Ricci flow,
Rm > _ 1/J (R) id -7 0 as R -7 00 R R ' as t -7 T, where [0, T) is the maximal time interval of
and Rmin (t) -7 00 existence. So the sectional curvature becomes asymptotically nonnegative under the unnormalized flow. Since we assume our solution to be either a soliton or a breather, the curvature must have been nonnegative to begin with. In [179]' Hamilton has shown that either the sectional curvature becomes strictly positive immediately or M splits locally as a product of a l-dimensional flat factor and a surface with positive curvature, and this splitting is preserved by the flow. In the former case, we know 9 converges to a metric of constant positive sectional curvature under the normalized flow. To rule out the latter case, consider the evolution equation for r under the normalized flow:
!J
RdJ1= -
J(~~,RC-~R9)dJ1'
where the volume is fixed at one and the pointwise inner product is given by contraction using the metric. We can calculate the integrand on the right as
44
1. RICCI SOLITONS
Because of the flat factor, R2 = 21 Rc 12 , and therefore dr / dt = r 2/3. This implies that r increases without bound, which is impossible for a soliton or a breather. 0 A similar (but more elaborate) argument based on a pinching set can be used to prove that a nontrivial soliton for the normalized Ricci flow on a compact Kahler surface must have curvature at least as negative as Koiso's example; see [222].
8. Perelman's energy and entropy in relation to Ricci solitons The notion of gradient Ricci soliton has motivated the discovery of monotonicity formulas for the Ricci flow, which in turn have useful geometric applications. Here we consider some monotone integral quantities. In particular, in Chapter 5 we shall further study Perelman's energy functional: (1.68)
where (M n , g) is a closed Riemannian manifold and f : M ---+ lR. As we will see in (5.31) and (5.41), this functional is nondecreasing under the following set of evolution equations (see below for a motivation for considering (1.70)): (1.69) (1.70)
agij _ -2R·· at ~J' af 2 at = -ilf + IV' fl - R.
In particular, we have the following. THEOREM 1.72 (Energy monotonicity). For any solution (g (t) , f (t)) of (1.69)-(1.70) on a closed manifold M n , we have
:tF(g(t),J (t)) = 2 Hence -9t F (g(t) , f (t))
(1.71)
1M I~j + V' V'jfI2 e- f dp, ~ o. i
= 0 at some time to if and only if Rij + V'i V'jf == 0
at time to. That is, 9 (to) is a steady gradient Ricci soliton flowing along V' f (to). We can motivate the consideration of equations (1.69)-(1.70) by seeing how it relates to a steady soliton 9 flowing along a gradient vector field V' f and in canonical form. By (1.14) and (1.25),
~~ = IV' fl2 = IV' fl2 -
R - ilf,
which is (1.70). A similar consideration for shrinking gradient solitons can be used to motivate the study of Perelman's entropy functional W (g, f, T) discussed in
8.
PERELMAN'S ENERGY AND ENTROPY IN RELATION TO RICCI SOLITONS
Chapter 6 and the associated equations for g, f and r analogous to F we define in (6.1) the entropy:
W(g, f, r)
~ 1M
[r (R + IV' f12)
+f
-
45
> 0. In particular,
n] (47rr)-n/2 e-f dJ.L.
Under the system of equations
a
(1.72)
fJt gij
= -2Rij ,
(1.73)
af 2 n at=-f).f+lV'fl-R+2r,
(1.74)
dr =-1 dt '
we shall show in (6.17) the following. THEOREM 1.73 (Entropy monotonicity). If (g (t), f (t), r(t)), r (t) is a solution of (1.72)-(1.74) on a closed manifold Mn, then
> 0,
d
dt W(g(t), f(t), r(t)) =
1M 2r I~j + V'iV'jf -
;; 12 (47rr)-n/2 e- f dJ.L
~ 0.
ft
Note that the right-hand side (RHS) vanishes, i.e., W(g(t), f(t), r(t)) = 0, if and only if 9 (t) is a shrinking gradient Ricci soliton. The above monotonicity formulas beautifully display the utility of considering Ricci solitons. REMARK 1.74. Assuming that (M n , 9 (t)) is a shrinking gradient Ricci soliton in canonical form (1.13) with>' = -1, we have Rc (g (t)) where ~;
+ V'g(t)V'g(t) f
1
(t) - 2r g (t) = 0,
= -1, and hence (1.14) implies f satisfies (1.73).
Finally we consider the expander entropy of Feldman, Ilmanen, and one of the authors [143]. Define the functional W+ (g, f+, r)
~ 1M
Under the system (1. 75) (1.76) (1.77) we have the following.
[r (R + IV' f+12) - f+
+ n]
(47rr)-n/2 e- f +dJ.L'
1. RICCI SOLITONS
46
THEOREM 1.75 (Expander entropy monotonicity formula). For a solution (g (t), f+ (t), T(t)), T (t) > 0, of (1.75)-(1.77) on a closed manifold Mn,
d
dt W+(g(t), f +(t), T(t)) =
1M 2T I~j + '\h'Vjf+ + ;; 12 (47rT)-n/2 e- f +dj.L ~ O.
Here ftW+(g(t),J+(t),T(t)) = 0 if and only if g(t) is an expanding gradient Ricci soliton. Recall that from Proposition 1.7
8;j ~ Rij
(1.78)
+ \//vjf + 2ETgij ~ 0,
where T (t) ~ c:t+ 1 > 0 and ~ denotes an equality which holds for a gradient Ricci soliton (Mn,g(t),f(t),E) in canonical form. EXERCISE
1. 76. Show that
(1) (1.79)
(2) (1.80)
Moreover, if f has a critical point in space, then C (t) is independent of t.
9. Buscher duality transformation of warped product solitons In this section we describe an interesting duality transformation for solutions of the modified Ricci flow on certain warped products which in particular take gradient Ricci solitons to gradient Ricci solitons. The consideration of these warped products with tori of potentially infinite dimensions leads to Perelman's energy functional. 9.1. A metric duality transformation. Let (Mn,g) be a Riemannian manifold and let (Pq, h) be a flat manifold such as a torus or Euclidean space. Given a function A : M ---t (0,00), consider the warped product manifold (M, g) XA (P, h) which is the Riemannian manifold (M x P,g + Ah). The Buscher duality transformation (see [38]' [39]) takes the metric to the metric Ug~g+A-lh.
(This is a special case of T -d uality in string theory.) Although its definition is simple, the transformation has some surprising properties. Let {xi} ~=1
9. BUSCHER DUALITY TRANSFORMATION
47
and {ya}~=l be local coordinates on M and 'P, respectively, such that ha{3 ~
h
(a~o., ~ ) = fJa{3.
We then have
P9ij = 9ij, P9a(3
=
AfJa{3,
and the rest of the components are zero, where p
.p(8 8) 9 8xi '8xj ,
P
9ij =;=
.p(8 8) 9 8ya' 8y{3 ,
9a{3 =;=
and similarly for ~ 9. The explicit formulas below are from Haagensen [175] (see also (1.38) or §J in Chapter 9 of Besse [27] for curvature formulas for warped products). LEMMA
1.77. The Christoffel symbols of p9 are pr~. ~J
= r~.
~J'
p {3 _ {31 . ria - fJa2V'~logA,
A
.
=
prij
p .
= -fJa {3 '2 V'~ log A,
pr~{3
=
r~{3
pr~j
= 0,
and likewise, the Christoffel symbols of the dual metric ~9 are
~rfj
= rfj ,
~r{3 = -fJ{3a ~2 V' . log A , ~a
~
. r~{3
~
1·
= fJ a {3 2A V'~ log A, ~rl'{3 a = ~riaJ. = ~r~. = 0 • ~J
LEMMA
1.78. The Ricci tensor of p9 is 9iven by
~ [~logA+ ~ lV'logAI 2] fJa{3,
PRa{3
=-
PRai
= 0,
P~j
= Rij - ~ V'i V' j log A - ~ V'i log A V' j log A,
and the Ricci tensor of ~9 is
~ Ra{3 = - 2~ ( -~ log A + ~ IV' log A12) fJa{3, ~Raz. -- 0 , M
~j
M
= ~j +
q q 2V'iV'j log A - '4V'i log AV'j log A.
48
1.
RICCI SOLITONS
LEMMA 1.79. The scalar curvature of P9 and ~ 9 are given by PR =
R -
q~logA _
~ R = R + q~ log A EXERCISE 1.80. Show that if A P
Rij =
~j
q (q: 1) IVlogA1 2 ,
- q (q 4+ 1) IV log AI2 .
~ exp (-~f) (i.e., log A
+ ViVjf -
1 q
q + 1 IV f1 2 , q
~ R = R - 2~f -
q + 1 IV fl2 . q
R
-~f), then
-Vi/Vjf,
+ 2~f -
PR =
=
Hence lim R P = R
q-+oo
+ 2~f -
IV fl2 ,
are Perelman's modified Ricci tensor and scalar curvature (see also (5.18) and (5.19)). As a consequence of the above exercise, if h has unit volume, we then have
lMxP Reg) dJ.L. g = 1M (R + 2~f - q; 1 1v f12) e- f dJ.Lg = 1M (R + q ~ 1 IV fl2 ) e- f dJ.Lg.
Taking q ~
00,
this limits to F (g, f) defined in (1.68).
9.2. Buscher duality. For a warped product solution of the modified Ricci flow of the above type we have the following. THEOREM 1.81 (Buscher duality). If Pg
(t) = 9 (t)
+ exp (-~ f q
t
E
(t))
t
dyO< Q9 dyO<,
0<=1
I, satisfy the modified Ricci flow
(1.81)
:t
Pgab = -2
e
Rab + 2 PVa PVb P
where P
~ 9 (t) + exp
Gf (t)) t,
dyo 0 dyo
9. BUSCHER DUALITY TRANSFORMATION
49
satisfy (1.82) where
Uc/> ~ ~ c/> + J. REMARK
1.82. Note that ~V o:f = Uv o:f =
o.
The functions 11 c/> and Uc/> are called dilatons and the transformation from 11 c/> to Uc/> is called the dilaton shift. The Buscher duality transformation takes solutions of the modified Ricci flow with dilaton ~ c/> to solutions of the modified Ricci flow with dilaton Uc/>. LEMMA
1.83 (Buscher duality preserves modified Ricci tensor after dila-
ton shift).
URo:.8 + 2 Uv 0: ttV.8 Uc/> = URo:i + 2 Uv 0: "Vi Uc/>
=
~2
( 11 Ro:.8 + 2 I1V 0: I1V.8 11 c/> ) ,
~ Ro:i + 2 I1V 0: I1Vi 11 c/> = 0,
URij+2 UVi "Vj Uc/>= l1~j+2 I1Vi I1Vj 11c/>. REMARK
1.84. If log A = - ~ f, then
~ Rij + 2 ~Vi
I1Vj 11c/> = Rij + ViVjf -
~VdVjf + 2V iVj ~c/>, q
1 q
tt Rij + 2 UVi UVj Uc/> = Rij - ViVjf - - VdVjf + 2V i Vj Uc/>. Suppose limit as q ~
%t 00,
~ gab = -2 11 Rab so that ~ c/> = O. By Exercise 1.80, taking the the equation for the metric gij (t) on the base manifold is
a
atgij
= -2 (Rij + ViVjf).
Note that d/-l (11 g) = e- f d/-l/\ dyl /\ ... /\ dyq. The effective action of (11 g,11 c/» is (e.g., see Alvarez and Kubyshin [2], equation (26))
iMXP (Reg) + 41v 11c/>12) e- 2b4>d/-l eg) = iM (R + 2b.f - q: 1 IV fl2 + 41v ~ c/>12) e- 2b4>e- f d/-l,
S eg,11 c/» =
and similarly for the dual pair (Ug,U c/» • (Note that by taking c/>~ = 0, we have S (l1g, 0) = fMxp R (l1g) d/-lb g.) Buscher duality preserves the effective action:
50
1. RICCI SOLITONS
9.3. Examples. The Buscher dual of a rotationally symmetric soliton solution on a surface is the soliton For example,
dr 2 + tanh 2 r d0 2
dr 2 + coth 2 r d0 2
dr 2 + tan2 r d0 2
dr 2 + cot 2 r d0 2
dr 2 + r 2d0 2
dr 2 + r- 2d0 2
dr 2 + d0 2
dr 2 + d0 2
Let (Nm, h) be an Einstein metric with Rc = £g, c E R Consider the doubly-warped product metric on R2 X Nm: ~g ~ ds 2 + F (s)2 d0 2
-+ G (s)2 h,
where s is the radial coordinate, 0 is the coordinate on the circle, and F and G are positive functions. The gradient Ricci soliton equation ~ Rab + 2 ~Va ~VbI = 0 becomes (see equation (1) in [219]) the following system for the triple (F, G, J): (1.83)
G"
F"
-21" = -C - F'
(1.84)
F" = -n F'G' - l' F',
(1.85)
C =
G
G"
c G2 - (n - 1)
(G')2 G
-
F'G' G' FG - l' G .
The Buscher dual metric is 1 ~g = ds 2 + --2d02
F(s)
+ G (s)2 h.
We know that ~g is a steady gradient Ricci soliton if ~ 9 is, and so we leave it to the reader to verify that indeed the triple (*, G, 1 - log F) is a solution to (1.83)-(1.85) if (F, G, J) is. 10. Summary of results and open problems on Ricci solitons
In this section we collect some known results and open problems about the properties and classification of gradient Ricci solitons. Besides the discussion earlier in this chapter, one may consult [111] for some of the proofs.
10.
SUMMARY
OF
RESULTS AND OPEN PROBLEMS ON RICCI SOLITONS
51
10.1. Gradient Ricci solitons on surfaces. The following is a compendium of known results about complete 2-dimensional gradient Ricci solitons with bounded curvature (shrinkers, steadies, and expanders). In proving them, one may use the fact that if (M 2 ,g) admits a nontrivial Killing vector field X which vanishes at some point 0 E E, then (M,g) is rotationally symmetric. (1) A shrinker has constant positive curvature. In particular, the underlying surface is compact. (2) A steady is either flat or the cigar. (3) A compact expander has constant negative curvature (if X (M) < 0, then there are no nonzero conformal Killing vector fields). (4) An expander with positive curvature is rotationally symmetric and unique up to homothety (see [241] and [111]). In part (4), one can apply the arguments of subsection 3.1 of this chapter and the following result (see [286] and [111]). THEOREM 1.85. If (Mn,g) is a gradient Ricci soliton on a noncompact manifold with Rij ~ ERgij for some E > 0, where R ~ 0, then R decays exponentially in distance to a fixed origin. In particular, if (M2, g) is an expanding gradient Ricci soliton on a surface, then R decays exponentially. A complete proof, using a more direct method, is given in [241]. PROBLEM 1.86. Are there any other expanders on a surface diffeomorphic to ]R2 besides the positively curved rotationally symmetric expander and the hyperbolic disk? PROBLEM 1.87. Do all complete 2-dimensional gradient Ricci solitons have bounded curvature? Are all complete 2-dimensional Ricci solitons gradient? Note that in dimension 3 there are complete homogeneous expanding solitons which are not gradient; see Section 5 of this chapter.
10.2. Gradient Ricci solitons on 3-manifolds. In dimension 3 we have the following results. This subsection is abbreviated since we pose some more problems in the next subsection for dimensions at least 3. (1) (Perelman) Any nonflat shrinker with bounded nonnegative sectional curvature is isometric to either a quotient of the 3-sphere or a quotient of S2 x R In particular, any shrinker with bounded positive sectional curvature is isometric to a shrinking solution with constant positive sectional curvature. (2) There exists a rotationally symmetric steady with positive sectional curvature, namely the Bryant soliton. (3) There exists a rotationally symmetric expander with positive sectional curvature.
52
1.
RICCI SOLITONS
PROBLEM 1.88. Are there any 3-dimensional steady gradient solitons besides a flat solution, the Bryant soliton, and a quotient of the product of the cigar and 1R? This is equivalent to asking if a steady gradient Ricci soliton with n = 3 and sect (g (t)) > 0 is isometric to a Bryant soliton.
10.3. Gradient Ricci solitons in higher dimensions. Here we assume n ~ 3. Note that any expanding or steady Ricci soliton on a closed n-dimensional manifold is Einstein. PROBLEM 1.89. Is a shrinking gradient Ricci soliton with n Rm (g (t)) > 0 compact?
~
4 and
PROBLEM 1.90. Is a compact shrinking gradient Ricci soliton, n ~ 4, and Rm (g (t)) > 0 isometric to a shrinking constant positive sectional curvature solution?16 PROBLEM 1.91. Are there n-dimensional steadies with positive curvature operator besides the Bryant soliton? PROBLEM 1.92. Are there any n-dimensional expanders with positive curvature operator besides the rotationally symmetric one? PROBLEM 1.93. Does there exist an expander with n ~ 3 and positively pinched Ricci curvature, i.e., Rij ~ ERgij for some E > 0, where R> O? By Theorem 1.85, such an expander has R decaying exponentially.
11. Notes and commentary Section 1. The first occurrence of the notion of a Ricci soliton in the literature is in Friedan [145], where non-Einstein Ricci solitons are called quasi-Einstein metrics. See Besse [27] for a comprehensive treatment of Einstein manifolds. The Gaussian soliton first appeared in §2.1 of Perelman [297]. In the case of surfaces we have encountered Ricci solitons in Chapter 5 of Volume One. There, solitons motivated several aspects of Hamilton's original proof of convergence of the Ricci flow for surfaces with R (go) > 0, including the scalar curvature, Harnack, and entropy estimates, as well as the estimate which shows the metric approaches a soliton. See in Volume One, Corollary 5.17 on p. 115, Proposition 5.57 on p. 145, Proposition 5.39 on p. 134, and Corollary 5.35 on p. 130. Expanding solitons are of interest because they constitute borderline cases for Hamilton's Harnack inequality. (Expanding solitons also model the formation of certain Type III singularities.) In fact, as we mentioned in Section 2 of this chapter and will discuss in more detail in Part II, the consideration of quantities which vanish on solitons led to the discovery of the Harnack quadratic. Steady solitons, where the isometry class of the metric is independent of t E (-00,00), occur as limits of Type II singularities. 16T he answer to this problem is 'yes' and follows from the recent work of B6hm and Wilking [30] (see Part II of this volume).
11.
NOTES AND COMMENTARY
53
Section 4. Einstein metrics are special cases of Ricci solitons. In general, if a warped product metric 9 is Einstein, then the fiber ('P, g) is Einstein. If the base is I-dimensional, then provided w (r) is not constant, the metric 9 = dr 2 + w (r)2 9 satisfies Rc (g) = )..g if and only if w' (r)2
+ ~w (r)2 =
_P_, n n-I where Rc (9) = pg and n = dim 'P. Without loss of generality (e.g., up to scaling and diffeomorphism), we have the following cases:
I
p =
~~i I).. = R~) I w (r) I
9
- (n - 1)
-n
0
-n
eT
including hyperbolic cusp
n-I
-n
sinhr
including hyperbolic space
n-I
0
r
Ricci flat cone
n-I
n
smr
including sphere
0
0
1
Ricci flat product
coshr including hyperbolic with two ends
As a consequence, we have the following (see Theorem 9.110 on p. 268 of [27]). THEOREM 1.94 (Einstein warped products over I-dimensional base). If a warped product (M, g) over a I-dimensional base, with the dimension of the fiber at least 2, is a complete Einstein manifold, then either (1) 9 is a Ricci flat product, (2) M is topologically a cone on P and the fiber is Einstein with positive scalar curvature, or (3) the base is JR, the fiber is Einstein with nonpositive scalar curvature, and 9 has negative scalar curvature.
CHAPTER 2
Kahler-Ricci Flow and Kahler-Ricci Solitons Symmetry, as wide or narrow as you may define its meaning, is one idea by which man through the ages has tried to comprehend and create order, beauty, and perfection. - Hermann Weyl
The Kahler-Ricci flow is simply an abbreviation for the Ricci flow on Kahler manifolds. In this chapter we first review some basic definitions and properties for Kahler manifolds and survey some of the fundamental results on the existence of Kahler-Einstein metrics. Then we discuss elementary properties of the Kahler-Ricci flow and state some of the fundamental longtime existence and convergence results. We also give a survey of KahlerRicci solitons. Some other highlights of this chapter are an exposition of tensor calculations in holomorphic coordinates, the proof of the long-time existence and convergence of the Kahler-Ricci flow on Kahler manifolds with Cl < 0, construction of the Koiso solitons and other U(n)-invariant solitons, proofs of differential Harnack estimates under the assumption of nonnegative holomorphic bisectional curvature, and a survey of uniformization-type results for complete noncompact Kahler manifolds with positive bisectional curvature. We assume the reader either has some knowledge of Kahler geometry or will read other references on Kahler geometry along with this chapter, so we make no attempt to be completely self-contained. l The latter chapters do not depend on this chapter and the reader interested only in Riemannian Ricci flow may skip this chapter. 1. Introduction to Kahler manifolds
In this section we introduce the basic concepts of complex and Kahler manifolds including the point of view of having an almost complex structure on an even-dimensional Riemannian manifold satisfying natural properties. Let M be a real 2n-dimensional differentiable manifold. A system of holomorphic coordinates on M is a collection {Zi : Ui ---t Zi (Ui) c en} , where {Ui} is a cover of M and Zi are homeomorphisms such that the maps
Zi
0
zjl : Zj (Ui n Uj ) ---t Zi (Ui n Uj )
lSee the notes and commentary at the end of this chapter for some references on complex manifolds and Kahler geometry. 55
56
2. KAHLER-RICCI FLOW
are holomorphic (complex analytic), and hence biholomorphic, whenever
Ui n Uj i= 0. Two systems of holomorphic coordinates {Ui' Zi} and {Vj, W j} on M are equivalent if whenever Ui n Vj i= 0, Wj 0 Zi- 1 :
Zi (Ui
n Vj)
----t
Wj
(Ui n Vj)
is a biholomorphism. A complex structure on a differentiable manifold M is an equivalence class of systems of holomorphic coordinates. A complex manifold is simply a differentiable manifold with a complex structure. A complex manifold has a natural real analytic structure and is orientable (using holomorphic coordinates {zO: ~ xO: + Ryo:} an orientation is
:=1 '
{/xr, i!;r, ... ,
determined by requiring that the frame a~n , a~n } be positively oriented). In this chapter and only in this chapter we shall use Mn to denote a complex manifold M of complex dimension n (half of the real . ) ,.I.e., n:::;=. d'Ime MId' · d ImenSlon = 2' Im]R M . A real sub manifold N of a complex manifold M n is a complex submanifold if for every pEN there exist holomorphic coordinates {ZO:}:= 1 in a neighborhood U of p such that
N
nU = {q E U : zk+
1
(q) = ... = zn (q) =
o} .
Clearly a complex submanifold is itself a complex manifold. In the above definition, k is the complex dimension of N. Some elementary examples of complex manifolds are (1) complex Euclidean space C n , (2) complex projective space cpn, (3) complex submanifold of cpm = algebraic manifold = zero set of a finite number of homogeneous polynomials, (4) complex torus C n Ir, where r is a lattice. A complex structure defines at each point p E M a map J : T Mp ----t T Mp by J = d (z-l 0 R 0 z) , where z are holomorphic coordinates defined in a neighborhood of p (this definition is independent of the choice of z). Since R 0 R = - iden, it is easy to see J2 = - idTM . In general, given an even-dimensional differentiable manifold M, an automorphism J : T M ----t T M is called an almost complex structure if J2 = - idTM . (A manifold with an almost complex structure is called an almost complex manifold.) Hence a complex structure induces an almost complex structure. We say that an almost complex structure is integrable if there exists a complex structure which induces the almost complex structure. Given an almost complex manifold (M, J), the Nijenhuis tensor is defined by
N J (X, Y) ~ [JX, JYJ - J [JX, YJ - J [X, JYJ - [X, YJ for X, YET M. By the Newlander-Nirenberg Theorem, a necessary and sufficient condition that an almost complex structure be integrable is that the Nijenhuis tensor vanish.
1. INTRODUCTION TO KAHLER MANIFOLDS
57
It is interesting that 8 6 admits an almost complex structure. In fact 8 2 and 8 6 are the only even-dimensional spheres which admit almost complex structures (see §8 of Chern [95] for example). The following is a longstanding unsolved question. 2.1 (Existence of complex structure on 8 6 ). Does 8 6 admit a complex structure? More generally, one may ask which closed evendimensional manifolds admit almost complex structures but not complex structures. PROBLEM
Let (M, J) be an almost complex manifold. A vector field V is an infinitesimal automorphism of the almost complex structure if the Lie derivative of J with respect to V is zero, i.e., (2.1)
LVJ
= O.
Note that (LVJ) (W)
= LV (JW) - J (LVW) = [V, JW]- J ([V, W]),
so (2.1) is equivalent to J ([V, W]) = [V, JW] for any vector field W. There are various equivalent ways to define a Kahler manifold. We say that a Riemannian manifold (M,g) with an almost complex structure J : T M ~ T M is a Kahler manifold if the metric 9 is J-invariant (or Hermitian): 9 (JX, JY) = 9 (X, Y)
and J is parallel: \7 J = 0
or equivalently, \7 x (JY) = J (\7 x Y)
for all X, Y, where \7 is the Riemannian covariant derivative. 2 The metric 9 is called a Kahler metric. Note that almost complex structures which yield Kahler manifolds are necessarily integrable. Indeed, by the Newlander-Nirenberg Theorem, we only need to check that the Nijenhuis tensor vanishes for a Kahler manifold: NJ (X, Y)
= \7 JxJY -
\7 JyJX - J (\7 JXY - \7yJX)
- J (\7 x JY - \7 JY X) - \7 x Y
=-
[J (\7 JX Y) - \7 JxJY]
+ [J(\7yJX) -
+ \7 Y X
+ [J (\7 JY X)
- \7 JyJ X]
\7 y J(JX)]- [J(\7xJY) - \7xJ(JY)]
=0.
A complex manifold (M, J) is called a Kahler manifold if it admits a Kahler metric. 2Since 0
= (V xJ) (Y) = V x
(JY) - J (V x Y).
58
2. KAHLER-RICCI FLOW
Given an almost complex manifold (M, J), the complexified tangent bundle is TeM ~ T M ®lR C. For each p E M we may extend the almost complex structure J to a complex linear map
Jc : TeMp
~
TeMp.
Since (Jc)2 = - idTcM p ' the eigenvalues of Jc are Rand define the holomorphic tangent bundle by
Tl,o M ~ {V
E
R.
We
TeM : Je (V) = HV}
and the anti-holomorphic tangent bundle by
TO,l M ~ {V E TeM : Je (V) = -HV} . This gives us a decomposition
TeM = Tl,o M EB TO,l M. A vector in Tl,o M is called type (1,0) and a vector in TO,l M is called type (0,1) . Alternatively, we can obtain the above decomposition by decomposing a (real) tangent vector VET M as V = Vl,o + VO,l, where
Vl,o ~
'12 (V -
HJV)
E
Tl,o M,
VO,l ~
'12 (V + HJV)
E
TO,l M.
I.e.,
Tl,oM = {Vl,o: V
E
TM},
TO,lM = {VO,l : V
E
TM}.
Since TeM is the complexification of a real vector space, the complex conjugate (bar operation), X + A Y ~ X - AY, where X, YET M, is defined on TeM. Given Z E TeM, the real part of Z is defined by Re (Z)
~ ~ (Z + 2) .
Note that if VET M, then 2 Re (Vl,O)
= 2 Re (VO,l) = V.
The almost complex structure satisfies
Je (if) = Je (V),
V
E
TeM,
and we have Tl,oM = TO,lM and TO,lM = Tl,oM. The complexified cotangent bundle is TcM = (TeM)* = T* M ®lR C, which decomposes into
Tc M = Al,oM EBAO,lM, where Al,oM ~ (Tl,oM)* and AO,lM ~ (TO,lM)*.
1. INTRODUCTION TO KAHLER MANIFOLDS
59
A covariant tensor of type (p, q) is a section of ®p,q M ~ (®P A1,0 M) ® (®q AO,lM).3 The complex conjugate extends to the tensor bundles and
Q9p,qM=Q9q,PM. We say that a differential (p + q)-form "1 is of type (p, q) if it is a section of the vector bundle AP,qM ~ (APT1,OM) A (AqTO,lM) c Ap+qTcM. We denote the space of (p, q)-forms by Op,q (M). We say that a (p,p)-tensor (or form) "1 is real if ij = "1. When we are considering complex manifolds, it is most natural to carry out calculations in holomorphic coordinates. Let {zC<} be local holomorphic coordinates. We may write zC< ~ xC< + Ayc<, where xC< and yC< are realvalued functions. Define dzC< ~ dxC<
+ HdyC< ,
diC< ~ dxC< - HdyC< and
8~C< ~ ~ (8~C< - H 8~C< ~ ~ (8~C< + H
8:C<) , 8:C<) ,
so that dzC<
(8~f3 )
=
6$,
diC<
(8~f3 )
=
6$,
(8~f3 ) = diC< (8~f3 ) = o. -JJa and ib = b· The holomorphic
dzC< Note that
-£a
=
tangent bundle
Tl,o M is locally the span of the vectors {8~"'} :=1 . Given two overlapping local holomorphic coordinates {zc<} and {wf3} ,the transition matrix relating 8 d 8 . {8z"'}n h· h t· fi 8z'" an 8w/3 IS 8w/3 c<,f3=1' w lC sa IS es 8 8wf3 =
n
L
c<=l
8 zC< 8 8wf3 . 8 zC<·
The anti-holomorphic tangent bundle yo,l M is locally the span of
{ib} ;=1.
NOTATION 2.2. Henceforth we shall use the Einstein summation convention where each pair of repeated indices consisting of an upper index and
a lower index is summed from 1 to n (sometimes, as a reminder, we include the summation symbol, however we always sum over repeated indices unless otherwise indicated). 3Not to be confused with a (P, q)-tensor in Riemannian geometry which is a section of (®P T* M) ® (®q TM).
2. KAHLER-RICCI FLOW
60
EXERCISE
2.3. Given a (p, q)-form
. ~ ~ dZ i1 ./ - n. 'nl"'tp)l"')q
n -
1\
•••
1\
dz ip
1\
1\
1\
1\
dZ}l
1\
1\
•••
1\
1\
dz}q ,
show that its complex conjugate is given in local coordinates by ii == (-I)pq n. ,. . dz iI A·,· A dz jq A dZ'il A··· A dz'ip E nq,p (M) . ./ . 'nl'''tp)l''')q The exterior differentiation operator d maps Ap,qM into AP+1,qM EB Ap,q+l M. Corresponding to this decomposition of the image of d, we have
d~8
+ a,
where
8: AP,qM
~
Ap+1,QM,
a: Ap,q M
~
AP,Q+1 M.
We extend the Riemannian metric 9 complex linearly to define ge : TeMp x TeMp
~
C.
Similarly 'Ve : TeM x Coo (TeM)
~
Coo (TeM) is the complex linear extension of 'V : T M x Coo (T M) ~ Coo (T M) with the convention that 'VxY ~ 'V (X, Y) and ('Veh Y ~ 'Ve (X, Y). The complex linear extension of Rm is denoted by Rme . Let
Since ge (V, W) = ge Hermitian condition:
(V, W)
= ge
(W, V) , these coefficients satisfy the
Similarly, we define
90.(3
~ ge (8~0.' 8~(3 )
,
We claim 90.(3 = 90i.i3 = O. Indeed, if X, Y E T1,0 M, then ge (X, Y)
= ge (JX, JY) = ge (yCIx, yCIy) = -ge (X, Y),
which implies ge (X, Y) = O. Similarly, if X, Y E TO,l M, then we also have ge (X, Y) = O. Thus, in local holomorphic coordinates, the Kahler metric takes the form ge = 90.i3 ( dzo. 0 dz(3 + dz(3 0 dZo.) . EXERCISE 2.4. Show that if a a is real if and only if
=
Aao.i3dzo. A dz(3 is a (1, I)-form, then
1.
INTRODUCTION TO KAHLER MANIFOLDS
61
Given a J-invariant symmetric 2-tensor b, we define a J-invariant 2-form ,8 by ,8(X,Y)~b(JX,Y).
(We check that ,8 (Y, X) = b (JY, X) = b (-Y, JX) = -,8 (X, Y).) We call,8 the associated 2-form to the symmetric 2-tensor b. The Kahler form w, on a Riemannian manifold (M, J,9) with an almost complex structure and whose metric is Hermitian, is defined to be the 2-form associated to 9: w (X, Y)
~ 9
(JX, Y),
which is a real (1, I)-form. EXERCISE 2.5. Show that if 9 is a Kahler metric, then w is a closed 2-form. In fact, w is parallel. HINT: See [27], Proposition 2.29 on p. 70. SOLUTION: We compute (\7 xw) (Y, Z) = X (w (Y, Z)) - w (\7 x Y, Z) - w (Y, \7 x Z)
= X (9 (JY, Z)) - 9 (J (\7 x Y) , Z) - 9 (JY, \7 x Z) = (\7 x9) (JY, Z) = 0, where we used the definition of w, J (\7 x Y) = \7 x (JY) , and 9 is parallel. On a Kahler manifold (M, J,9), in local holomorphic coordinates, the Kahler form is n
W
=
R
L
9o.fjdzo. 1\ d-z{3.
o.,{3=1
b)
REMARK 2.6. In our convention, dzo. I\d-z{3 (-/z-r, = !8~8r Note that for the standard Euclidean dx 2 + dy2 metric on en, we have 9o.fj = !8o.{3 and 9c = (dzo. 0 d-zo. + d-zo. 0 dzo.) .
! r::=1
Since 9 is Kahler, we have the Kahler identities:
G
(2.2)
G
Gz"Y 9o.fj = Gzo. 9"Yfj,
which are equivalent to dw
[w]
= O. The real cohomology class
E H1,1(M;~) ~ H2(M;~)
is called the Kahler class of w. EXERCISE 2.7 (Characterization of Kahler condition). If (M, J,9) is a triple consisting of a Riemannian manifold, an almost complex structure, and a Hermitian metric, then dw = 0 if and only if \7 J = O. That is, 9 is Kahler if and only if the real (1, I)-form w is closed.
2. KAHLER-RICCI FLOW
62
An oriented Riemannian surface (M,9) has a natural complex structure and 9 is a Kahler metric with respect to this complex structure. In particular, define the almost complex structure J : T M --t T M as counterclockwise rotation by 90° with respect to the orientation and metric (clearly J2 = - idT M)' Since dimlR M = 2, we have dw = 0, which implies \1 J = O.
2. Connection, curvature, and covariant differentiation In this section we consider the connection, curvature, and covariant differentiation on a Kahler manifold (M n , J,9). Our emphasis, in the style of the book by Morrow and Kodaira [275], is to calculate geometric quantities in local holomorphic coordinates, where the formulas are particularly elegant and simpler than their Riemannian counterparts. These calculations shall prove useful in our study of the Kahler-Ricci flow. The Christoffel symbols of the Levi-Civita connection, defined by
8 ~~('Y 8 ( ) \1 IC o~'" 8z/3 -;- L...J r Ot/3 8z-Y 'Y=1
(\1 c)
0
-
Y 8) + r 'Ot/3 8z'Y '
n(r'Y --8 + r'Y-8) --
8 == ~
oz'" 8z/3 . L...J -y=1
Ot/3 8z'Y
Ot/3 8z'Y
'
etc., are zero unless all the indices are unbarred or all the indices are barred. To see this, from the Riemannian formula for the Christoffel symbols and since we are complex linearly extending the covariant derivative, we have the following using (2.2). 2.8 (Kahler Christoffel symbols). Let 9 Ot (3 be defined by 9'Y(39 Ot (3 Then, in holomorphic coordinates, we have
LEMMA 8~.
(
2.3)
-y
=
_ 1 -y6 (8 _ 8 _ 8 ) _ 'Y6 8 _ 8zOt9/3li + 8z/39Otli - 8zli9Ot/3 - 9 f)zOt9/3li
r Ot/3 - '2 9
and
Y -- r'Y/3Ot r 'Ot/3 (in fact, the last equality is equivalent to the Kahler condition).4 Similarly
r~(3 = ~ 9 -y6 (f)~Ot 9(36 + f)~/3 9Ot6 - 8~li 9Ot(3) r~(3
= 0,
= 0, and so forth. EXERCISE
2.9. Show that Y r '(i(3
-
-
r'YOt/3'
4We leave this as an exercise or see for example Theorem 5.1 in [275].
2. CONNECTION, CURVATURE, AND COVARIANT DIFFERENTIATION
63
SOLUTION. We compute 1
_ 1 18 (0
r aiJ - 29
_ {jiQ9f38
0
+ ozf3 9M -
0 _) _ 18 0 _ oz89af3 - 9 oz0l.9f38
- 0 = 9'Y8_ - = r'Y ozOl. 9f38 0I.f3. Similarly to ge and 'Ve, we may define Rme and Rce as the complex multilinear extensions of Rm and Rc. The components of the curvature (3, 1)-tensor Rme are defined by 00)0. 8 0 Rm e ( ozOl. 'ozf3 OZ'Y =:= ROI.iJ'Y Oz8
+ ROI.iJ'Y OziS'
00)0. 8 0 Rm e ( OzOI. 'ozf3 oz1 =:= ROI.iJ1 Oz8
+ ROI.iJ1 OziS'
is
0
is
0
etc. As a (4,0)-tensor the components of Rme, which are defined by ROI.iJ'Y iS
~ Rme (O~OI.' 0~f3' O~'Y' 0~8 )
,
satisfy etc. The only nonvanishing components of the curvature (3, 1)-tensor are
R~iJ'Y ' R~iJ1' R~f3'Y ' R~f3i . In particular, R~f3'Y = R~f3i = 0, etc. Hence, the only nonvanishing components of the curvature (4, O)-tensor are ROI.iJ'YiS' ROI.iJ18 , Raf3'YiS' Raf3i8.
Since a~o< r~'Y = 0, we have
(2.4) and thus we have the following lemma. 2.10 (Kahler Rm). In holomorphic coordinates, the components of Rme are 9iven by LEMMA
(2.5)
We have the identities ROI.iJ'YiS
= R'YiJOI.iS = ROI.iS'YiJ = R'YiSOI.iJ
and ROI.iJ'Y iS = Rf3a81· The vanishing of some of the components of the curvature (4, O)-tensor is related to the following.
64
KAHLER-RICCI FLOW
2.
EXERCISE 2.11 (J-invariance of curvature). Show that for a Kahler manifold Rm (X, Y) is J-invariant, i.e., Rm (X, Y) J Z = J (Rm (X, Y) Z) . SOLUTION. We compute Rm (X, Y) JZ = \lx\lyJZ - \ly\lxJZ - \l[X,y]JZ = \lx (J\lyZ) - \ly (J\lxZ) -
J (\l[X,y]Z)
= J (\lx\lyZ) - J (\ly\lxZ) - J (\l[X,y]Z) =
J (Rm (X, Y) Z) .
The components of Rcc are defined similarly:
(8~a' 8~/3 ) Rc c (8~a' 8~/3 )
Rc C
=;= Rai3' =;= R a /3,
etc. Tracing (2.4), we see that components of the Ricci tensor Rai3 = R~i3a are given by LEMMA 2.12 (Kahler Rc).
(2.6)
82
Rai3 = - 8z a8z/3 log det (9",s) .
It is easy to see that Rai3 = R/3o. and Ra/3 = RQi3 = O. EXERCISE 2.13. Show that for a Kahler manifold Rc is J-invariant, i.e., Rc (JX, JY) = Rc (X, Y). The Ricci form p is the 2-form associated to Rc, p (X, Y) =;=
1
2 Rc (JX, Y),
which is also a real (1, I)-form. EXERCISE 2.14. Show that if 9 is a Kahler metric, then the Ricci form p is a closed 2-form.
HINT: See Proposition 2.47 on p. 74 of [27]. The real de Rham cohomology class [2~P] =;= Cl (M) is the first Chern class of M. It is a beautiful fact that [2~P] only depends on the complex structure of M. The Ricci form, which is a real (1, I)-form, is in holomorphic coordinates: p = RRai3dza 1\ dz/3.
From (2.10) below and r~/3
= r~a 8
we have
8
8 za RWr = 8 z/3 Ra;'fl
2. CONNECTION, CURVATURE, AND COVARIANT DIFFERENTIATION
which is equivalent to the Ricci form being closed: dp (2.6) as p = -v'-Iaalogdet (9['8) . The complex scalar curvature is defined to be
= o.
65
We may express
R -;- 9 o13R0(3· ...!...
The complex scalar curvature is one-half of the Riemannian scalar curvature. (Exercise: Prove this.) 2.15. In this chapter, R always denotes the complex scalar curvature whereas in the rest of the book, R always denotes the Riemannian scalar curvature. In this chapter we shall usually refer to the complex scalar curvature simply as the scalar curvature. NOTATION
Holomorphic coordinates {ZO} are said to be unitary at a point p if 9 (a~Ct' (p) = 80 (3. That is, {a~a (p)} is a unitary frame for T 1,o Mp. In such a frame we have at p,
:=1
ib)
n
(2.7)
Ro13 =
L Ro/388
8=1
and 0=1
Note that for the standard Euclidean metric 1 n ge = 2 (dw O 0 diiP + diiP 0 dw O )
L
0=1
on
en the new coordinates {zo ~ ~Wo} :=1 are unitary.
{a},
1
Given Z E T 1,oM let X ~ Re(Z) = (z + z) E TM. The holomorphic sectional curvature in the direction Z is defined to be
(2.8)
K (Z)=Rm(X,JX,JX,X)
IXI 4
e.
Rm(Z+Z,J=I(Z - Z) ,J=I(Z - Z) ,Z+Z)
41Z1 4 Rme
(Z, Z, Z, Z)
IZI 4 Hence Ke (Z) = Rme (V, V, V, V) , where V ~ Zj IZI. In particular, if {ZO} is unitary at p, then Ke (-£0) = RO{io{i. We say that the holomorphic sectional curvature is positive (respectively, nonnegative) if Ke (Z) > 0 (respectively, Ke (Z)
~ 0) for all Z
{a}.
T 1,o M By (2.8), positive (respectively, nonnegative) Riemannian sectional curvature implies positive (respectively, nonnegative) holomorphic sectional curvature. Given Z,W E T 1,oM let X ~ Re(Z) and U ~ Re(W). The (holomorphic) bisectional curvature in the directions (Z, W) is defined
{a},
E
66
2. KAHLER-RICCI FLOW
to be J(,
(Z W) == Rm(X, JX,JU,U) = Rme (Z,Z, W, w)
e,
.
IZI 21WI 2
IXI21U12
Clearly we have Ke (Z, Z) = Ke (Z). We say that the bisectional curvature is positive (respectively, nonnegative) if Ke (Z, W) > 0 (respectively, Ke (Z, W) 2: 0) for all Z, WE Tl,O M Clearly positive (respectively, nonnegative) bisectional sectional curvature implies positive (respectively, nonnegative) holomorphic sectional curvature. We also have the following (see Proposition 1 on p. 32 of [270] for example).
{o} .
LEMMA 2.16. Po.sitive (respectively, nonnegative) Riemannian sectional curvature. implies positive (respectively, nonnegative) bisectional sectional curvature.
Similarly to the notation for the components of the complex Riemann curvature tensor, we define
V o.R{3"y8TJ
~ ((Vc) {}~'" Rm e) (0~{3' 0~'Y ' 0~8 ' O~TJ )
.
The second Bianchi identity says that
(2.9) since - ) + ( V {} Rme )(0 - 0 - 0 - 0) ( V {} Rme )( - 0 - 0 - 0 0 {}zO oz{3' oi'Y' oz8' oiTJ {}z'Y ozo.' oz{3' oz8' oiTJ
+ (V a;P {} Rme) (o~z'Y , -00 zo. , 00ZU
0
o~zTJ )
=0
(Le., V o.R{3"y8ii + V (3R;yo.8ii + V"yRo.{38ii = 0, where V"yRo.{38ii = 0). Taking the trace, we have
(2.10) Note that Vo.R{3;y = ~R{3;y - r~{3R8"y' The volume form is d/1g ~ ~wn, where w n ~ w 1\ n = dime M, which in local coordinates is wn = n! ( R) n det(go.,B)dz l
1\
di l
1\ ... 1\
... 1\
dz n 1\ din.
The volume is
(2.11)
Vol (M) = Volg (M) =
~ {
n·JM
The total scalar curvature may be written as (2.12)
w (n times) and
wn .
2.
CONNECTION, CURVATURE, AND COVARIANT DIFFERENTIATION
Let V i3 (2.13)
~
67
V 8 / 8z i3' The Laplacian acting on tensors is given by
1 ~ ~ "2l~/3 (VOtVi3
+ Vi3VOt)
1 ="2 (VOtVa
+ VaVOt).
NOTATION 2.17. At this point we have begun to use the extended Einstein summation convention where not only pairs of repeated indices consisting of an upper index and a lower index are summed but also pairs of repeated lower indices consisting of a barred index and an unbarred index are summed. For example, aOta ~ L:~,/3=1 gOti3aOti3' Formulas which are expressed using this convention hold in the literal sense in unitary holomorphic coordinates at a point p. Usually we use the extended Einstein summation convention in lieu of computing in unitary coordinates at a point. In this sense, the formulas in (2.7) and throughout this chapter hold in arbitrary holomorphic coordinates. EXERCISE 2.18. Show that the Laplacian defined above in (2.13) is onehalf of the Riemannian Laplacian. Acting on functions, the Laplacian is A
u =
g
Oti3n v
n 01 V
i3 = g
Oti3
since, acting on functions, VOl V i3 = 8z~~z~ is a real-valued function on M, then (2.14)
{j2
8z Ot 8z/3 '
= V i3V
01'
We also note that if ¢
~ Igrad¢1 2 = gOti3vOt¢Vi3¢ ~ IV¢1 2 ,
where the LHS is the Riemannian gradient. Likewise, (2.15)
1 2 = 1V OtVi3¢ 12 "2IHess¢1
+ IV
Ot
V/3¢1 2,
where the LHS is the Riemannian Hessian. In our calculation of evolution equations under the Kahler-Ricci flow we shall often use the following. LEMMA 2.19 (Kahler commutator formulas). On a Kahler manifold we have the following commutator formulas for covariant differentiation acting on tensors of type (1,0), (0,1), and (1,1) , respectively: (2.16) (2.17) (2.18)
V 01 V i3a-y - V i3 V Ota-y
= - R~i3-yao,
= -R~i3;ybiS = R~a}iS' VOl V i3 a-yiS - V i3 V Ota-yiS = - R Ot i3-yf/aT/iS + R Oti3T/iS a-yf/' VOl V i3 b;y - V i3V Otb;y
Analogous formulas hold for higher degree tensors and forms (see Exercise 2. 22).
REMARK 2.20. Note that ROti3T/iS = -ROt i3iST/' which exhibits the consistency of the formula above with the analogous formula in Riemannian geometry. We have also used the extended Einstein summation convention.
2. KAHLER-RICCI FLOW
68
Using the fact that the Christoffel symbols are zero unless all of the indices are unbarred or all are barred, we compute PROOF.
and
Va V /Ja'Y = OaO/Ja'Y - r~'Y0/Jao,
r~'Yao) .
V 13 Vaa'Y = 013 (Oaa'Y Hence, using (2.4), we have
Va V /Ja'Y - V13 V aa'Y = o/Jr~'Yao = - R~/J'Yao, which is (2.16). Equation (2.17) is the complex conjugate of (2.16). If a is a (1, I)-tensor, then we compute V /Ja'Y1, = 0/Ja'Y1, - r~1,a'Ye,
Vaa'Y1,
r~'YaTJ1,
= Oaa'Y1, -
and
Va V/Ja'Y1, = Oa (o/Ja'Y1, V 13 Vaa'Y1,
r~1,a'Ye)
-
r~'Y (o/JaTJ1, - r~1,aTJe)
,
= 013 (Oaa'Y1, - r~'YaTJ1,) - r~1, (Oaa'Ye - r~'YaTJe) .
Hence, after some cancellations, we obtain in holomorphic coordinates {za} at a point p where it is unitary,
Va V /Ja'Y1, - V13 Vaa'Y1, = -Oar~1,a'Ye + 0/Jr~'YaTJ1, = Ra/Je1,a'Ye - Ra/J'YfiaTJ1,'
o LEMMA 2.21 (Va and V (3 commute). We have [Va, V{3] = 0 acting on tensors of any type. That is, if a is a (p, q)-tensor, then
Va V{3a'Yl "''Y1'1,l ... 1,q - V{3 V aa'Yl "''Y1'1,l ...1,q = O. PROOF.
We have
V{3a 'Yl"''Y1'Ol'''Oq - - -- o{3a'Yl···'Y1'Ol···Oq - - -
r TJ a - {3'Yj 'Yl"''Yj-1TJ'Yj+l''''Y1'dl'''Oq '
It is easiest to compute the second covariant derivative in normal holomorphic coordinates centered at any given point p E M, where r~{3 (p) = O. In these coordinates we have at p,
Va V{3a'Yl ···'Y1'1,1···1,q = Oa (0{3a'Y1···'Y1'1,l ... 1,q -
- a o{3a -
a
- - -
'Y1"''Y1'd1'''dq
r~'Yj a'Yl "''Yj-1 TJ'Yj+1 "''Y1'1,1 ...1,q)
(a r a
TJ ) {3'Yj
a'Y1"''Yj-1 TJ'Yj+1"''Y1'01'''Oq' - -
2. CONNECTION, CURVATURE, AND COVARIANT DIFFERENTIATION
69
Hence V' a V'(3a,Yl' "'Yp81 .. ·8q
-
V' (3 V' aa'Yl .. ''Yp81 ···8q
-- (8(3r'1/a'Yj - {)a r'1/(3'Yj ) a'Yl···'Yj-l'1/'Yj+1···'Yp,h···8q· - -
Now from (2.3) we have at p,
8(3r~'Yj
-
8ar~'Yj
= g'1/5.. (8(38ag'Yj5.. - 8a8(3g'Yj5..) = 0,
o
and the lemma follows.
EXERCISE 2.22. Compute the commutator formula for the operator [V' a, V',a] ~ V' a V',a - V',aV' a acting on (p, q)-tensors and forms. SOLUTION TO EXERCISE 2.22. If a is a (p, q)-tensor, then V' a V' ,aa'Yl"''Yp81 ... 8q
-
V',a V' aa'Yl ''''Yp81 •.. 8q
p
L R a,a'Yifj a'Yl···'Yi-l'1/'YHl···'Yp8 ···8
=-
1
q
i=l
q
+L
Ra,a'1/8j a'Yl···'Yp81 ···8j _ 1 fj8j+l···8q·
j=l
The commutator [~, V',a] , acting on functions, is ~ V',af
1
= "2 (V' a V'ii + V'ii V' a) V',af =
1
1
"2 V' a V',a V'iif + "2 V',a V' ii V' af 1
= "2R'Y,aV'''ri + V',a~f.
(2.19)
LEMMA 2.23 (Kahler V' and ~ commutator). The commutator [V' a, ~l ,
acting on (0, I)-forms, is given by (2.20)
1
1
V'a~a,a - ~V'aa,a = "2V'a R8,aa8 - "2Ra8V'8 a,a + R a;Y8,aV''Y a8'
PROOF. We compute
+ V';y V' 'Y) a,a - (V''Y V';y + V';YV''Y) V' aa,a = V''Y (V' a V';y - V';y V' a) a,a + (V' a V';y - V';y V' a) V''Ya,a = V''Y (R a;Y8,aa8) - R a;Y'Y8V'8 a,a + R a;Y8,aV''Y a8 = V' aR8,aa8 - Ra8 V' 8a,a + 2R a;Y8,a V' 'Y a8'
2 (V' a~a,a - ~ V' aa,a) = V' a (V' 'Y V';y
where we used [V'a'V''Yl
= 0 and V''YRa;Y8,a = V'aR8,a from (2.9).
0
REMARK 2.24. By taking the complex conjugate of the lemma above,
we have (2.21)
2.
70
KAHLER-RICCI FLOW
Since the Kahler-Ricci flow is a heat-type equation for Kahler metrics, some evolution equations we shall derive later in this chapter use the following commutator formula. If a is a time-dependent (0, I)-form, then under the Kahler-Ricci flow,
[:t
-~, \/0:] a{J = ~\/ o:R8{Ja8 - ~RO:8\/8a{J + RO:18{J\/"(a8'
Another basic formula is the commutator of the heat operator and the Hessian. First, by (2.19) and (2.20), we have for any function f on M,
\/ 0: \/ {J~f
= ~L \/0: \/ {Jf ~ ~ \/ 0: \/{Jf
(2.22)
+ Ro:{J81 \/"( \/8f -
1 1 2 R o:8\/8\/{Jf - 2R"({J\/ 0: \/1f,
where ~L is called the (complex) Lichnerowicz Laplacian. If function also of time, then (2.22) tells us \/ 0: \/ {J (:t -
(2.23)
~)
f = (%t -
~L )
f is a
\/0:\/{Jf.
Note that since \/0: \/ {J = 8o:8f3 acting on functions, we have when acting on functions.
3t (\/0:\/(J) = 0
REMARK 2.25. Formula (2.23) has the following Riemannian analogue. For a time-dependent function f and under the Ricci flow, we have
(2.24) where
\/i\/j (%t ~L,
~)
f = (%t -
~L) \/i\/jf,
defined by ~LVij ~ ~Vij
+ 2Rkijf.Vkf. -
~kVjk
- RjkVik,
is the (Riemannian) Lichnerowicz Laplacian acting on symmetric 2-tensors (see Lemma 2.33 on p. 110 of [111]). In fact, (2.24) is a special case of (2.23). 3. Existence of Kahler-Einstein metrics In this section we discuss what is known about the existence and uniqueness of Kahler-Einstein metrics, which are canonical (constant Ricci curvature) metrics on Kahler manifolds. The Kahler-Einstein equation is elliptic whereas the Kahler-Ricci flow, discussed beginning in the next section, may be considered as its parabolic analogue. First recall the following 88-Lemma, which is a consequence of the Hodge decomposition theorem. LEMMA 2.26 (d-exact real (1, I)-form is 88 of real-valued function). Let M be a closed Kahler manifold. Ifw is an exact real (1, I)-form, then there exists a real-valued function 'IjJ such that Ff.88'IjJ = w. That is, n
82 8zo:8zf3 'IjJ
= wo:{J,
3.
EXISTENCE OF KAHLER-EINSTEIN METRICS
where W = HWo./3dzo. 1\ dz/3 and wo./3
71
= W/3ii'
PROOF. This is a standard result in the theory of Kahler manifolds; see the book by Zheng [383] for example. 0 REMARK 2.27. More generally, we have the following. Let b be a (p, q)form, where p, q > O. If b is d-closed and either d-, 8-, or 8-exact, then there exists a (p - 1, q - I)-form
(2.25)
P = 27rw.
Since Po H Rc (Yo) 0./3 dzo. 1\ dz/3 is a real (1, I)-form in the same cohomology class 27rCl (M) as 27rW = 27rHwo./3dzo.l\ dz/3, by Lemma 2.26 there exists a real-valued function f such that Rc (Yo)o./3 - 27rwo./3
= 80.8/3J.
Therefore equation (2.25), which in local coordinates is Ro./3 equivalent to
80.8/3f
= 27rwo./3' is
= Rc (yo)o./3 - Ro./3 2
2
8 ) 8 = - 8 z 0.8z/3 log det ( (Yo)'Y8 + 8 z 0.8z/3 log det (Y'Y8) . Since M is closed, this implies det
(y 8)
(1')
log
det (Yo)'Y 8
=
f + log C
for some constant C > O. Since the real (1, I)-forms wo and ware in the same cohomology class, using Lemma 2.26 again, we see that there exists a real-valued function
Y'Y8
= (Yo)'Y 8 + 8'Y 86
Thus we may rewrite (2.25) as a complex Monge-Ampere equation: det ((yo)'Y 8 + 8'Y86
----'---:---~-'- - C ef
det ((yo)'Y 8)
-
.
72
2.
KAHLER-RICCI FLOW
Yau's proof of Theorem 2.28 involves solving the fully nonlinear equation above by using the continuity method. The proof was a tour de force. One says that a Kahler metric 9 is Kahler-Einstein if p = >.w for some >. E R If M n admits a Kahler-Einstein metric g, then
(We have>. = ~, where the (complex) scalar curvature R is constant.) Therefore, a necessary condition for the existence of a Kahler-Einstein metric on M is that its first Chern class have a sign. By having a sign we mean that ct (M) = 0, < 0, or > 0 if there exists a real (1, I)-form in the first Chern class which is zero, negative definite, or positive definite, respectively. COROLLARY 2.29 (Cl = 0: existence of Kahler Ricci flat metrics). If (Mn, go) is a closed Kahler manifold with Cl (M) = 0, then there exists a Kahler metric 9 with [w] = [wo] such that Rc (g) == O. Kahler Ricci flat metrics are called Calabi-Yau metrics and Kahler manifolds with Cl (M) = 0 are called Calabi-Yau manifolds. Another consequence of Theorem 2.28 is that if Cl (M) < 0 (respectively, Cl (M) > 0), then in each Kahler class there exists a metric with negative (respectively, positive) Ricci curvature. When Cl (M) < 0, we have the following result about the existence of Kahler-Einstein metrics conjectured by Calabi and proved by Aubin [11,12] and Yau [378, 379]. Calabi proved that such a Kahler-Einstein metric is unique if it exists [43]. THEOREM 2.30 (Cl < 0 Calabi conjecture: R < 0 Kahler-Einstein metrics). If Mn is a closed complex manifold with ct (M) < 0, then there exists a Kahler-Einstein metric 9 on M, which is unique up to homothety (scaling), with negative scalar curvature. A consequence of Theorem 2.30 is the following Chern number inequality:
~
(-It 2 (n + 1) cl(M)n-2 c2 (M). n (See Yau [378] and also Corollary 9.6 on p. 226 of [383] for an exposition.) (-It cl(Mt
REMARK 2.31. Another consequence of Theorem 2.30 is that if (M2, g) is a closed Kahler surface homotopically equivalent to CP2, then M2 is biholomorphic to Cp2 (see Yau [378]; earlier related work in all dimensions was done by Hirzebruch and Kodaira [204]). If Cl (M) > 0, however, there are obstructions to the existence of KahlerEinstein metrics. An example is the Futaki invariant (see [147]). On a closed manifold Mn with Cl > 0 (Le., a Fano manifold), fix a Kahler metric 9 such that [w] is a positive real multiple of cl(M). (This is the so-called
3. EXISTENCE OF KAHLER-EINSTEIN METRICS
73
canonical case.) By scaling the metric, we may assume [w] = cl(M). By Lemma 2.26 there exists a smooth function f : M --+ IR such that p-
27l'W
=
Raar
(One can make f unique by the normalization fM e- f d/-t = 1.) Let 1}(M) denote the space of (real) holomorphic vector fields on M. The Futaki functional F[wJ : 1}(M) --+ C is defined by
F[wJ (V)
~ 1M V (I) d/-t = 1M (V, \7 f)
d/-t.
Futaki [147] showed that F[wJ is well defined, i.e., that it depends only on the homology class [w]. It is then clear that if M admits a Kahler-Einstein metric, then F[wJ vanishes. However, Tian has shown that 1}(M) = 0 (which implies F[wJ = 0) does not imply there exists a Kahler-Einstein metric. The following uniqueness result in the Cl > 0 case was proved by Bando and Mabuchi [21]. THEOREM 2.32 (Uniqueness of Kahler-Einstein metrics). Let (Mn,g) be a closed Kahler manifold with Cl (M) > O. The Kahler-Einstein metric (with positive scalar curvature), if it exists, is unique up to scaling and the pull back by a biholomorphism of M.
The following result was proved by Andreotti and Frankel [144] for n = 2, Mabuchi [260] for n = 3, and Mori [271] and Siu and Yau [336] in all dimensions; Mori proved a more general algebraic-geometric result. Work on characterizing cpn was done by Kobayashi and Ochiai [237]. THEOREM 2.33 (Frankel Conjecture). If(Mn, g) is a closed Kahler manifold with positive bisectional curvature, then M n is biholomorphic to cpn.
In fact, if the bisectional curvature is nonnegative everywhere and positive at some point, then M is biholomorphic to cpn. When n = 2 and Cl (M) > 0, we have the following (see Tian [345]). 2.34 (Cl > 0 surfaces: R > 0 Kahler-Einstein metrics). If M2 is a closed complex surface with Cl (M) > 0 and the Lie algebra of the automorphism group is reductive, then there exists a Kahler-Einstein metric g on M with positive scalar curvature. THEOREM
2.35. Note that such surfaces are biholomorphic to CP2 blown up at p points, where 3 ~ p ~ 8. On the other hand, for n ~ 2, cpn blown up at 1 or 2 points does not admit a Kahler-Einstein metric (see p. 156 of Lichnerowicz [254] and Yau [376]). See Section 7 of this chapter for the existence of Kahler-Ricci solitons on CP2 blown up at 1 or 2 points. REMARK
There are a number of additional works related to stability and the existence of Kahler-Einstein metrics with Cl > O. Notably Aubin [14], Siu [335]' Nadel [281]' Tian [346], Donaldson [129], [130]' and Phong and
74
2. KAHLER-RICCI FLOW
Sturm [304]. The existence of Kahler-Einstein metrics on complete noncompact manifolds with Cl < 0 and Cl = 0 has been well studied (see Cheng and Yau [94] and Tian and Yau [349], [350] for example). For work on the existence of singular Kahler-Einstein metrics on certain classes of closed Kahler manifolds where Cl does not have a sign, see Tsuji [360] (for some further recent work see Cascini and La Nave [60] and Song and Tian [337]). Kahler-Einstein metrics are closely related to Kahler-Ricci solitons and hence the Kahler-Ricci flow which will be discussed next in this chapter.
4. Introduction to the Kahler-Ricci flow In this section we introduce the Kahler-Ricci flow system and its equivalent formulation as a single parabolic Monge-Ampere equation. We discuss some basic estimates which may be proved using the maximum principle. 4.1. The Kahler-Ricci flow equation. Let (Mn, J) be a closed manifold with a fixed almost complex structure. Given a Riemannian metric g, we may define a 2-tensor w by w (X, Y) ~ 9 (JX, Y). Recall that when 9 is Hermitian, w is antisymmetric (Le., defines a 2-form) and w is called the Kahler form. If a solution 9 (t) to the Ricci flow ~gij = -2~j is Hermitian at some time t, then w satisfies the equation gtW = -2p at that time, where p = p (t) is the Ricci form of 9 (t) . Hence
~ (dw) = d (~w) at at
= -2dp = 0
whenever 9 (t) is Kahler (we can define even when 9 is not Hermitian). This suggests that if 9 (0) is Kahler, then under the Ricci flow 9 (t) is Kahler for all t 2: O. Consider the Kahler-Ricci flow equation (2.26)
a
at gai3
= -Rai3
for a 1-parameter family of Kahler metrics with respect to J, which is obtained from the Ricci flow by dropping the factor of 2. Now we derive the parabolic complex Monge-Ampere equation to which the Kahler-Ricci flow is equivalent. For a complete initial Kahler metric with bounded curvature, we will use this scalar equation to prove the short-time existence of a solution to the initial-value problem for the Kahler-Ricci flow. On a closed manifold we will use the scalar equation to prove that the Kahler property of an initial metric is preserved under the Ricci flow and to prove the long-time existence of solutions to the Kahler-Ricci flow. By (2.26), gt [w] = - [p (t)] = - [p (0)] , so that the Kahler class of the metric at time t evolves linearly,
[w (t)] = [w (0)] - t [p (0)] , and the real (1, 1)-forms w (t) - w (0) + tp (0) are exact for t 2: O. Let g~i3 ~ gai3 (0). Using Lemma 2.26, for each t there exists a real-valued function
4.
INTRODUCTION TO THE KAHLER-RICCI FLOW
75
'-P (t) defined on all of M such that o
-
0
-
ga:fi (t) = ga:!3 + tOa:o,a log det g')'6 + oa:o,a'-P (t) .
(2.27)
By (2.6) we have
Hence, by differentiating (2.27), we obtain
Oa: 8,a (:t '-P) = - Ra:fi - oa: 8,a log det g~6 _
det
(g~6 + to')'8olog det g~iI + o')'8o'-P (t) )
= oa:o,a log
dO.
etg')'6
Hence we conclude that the Kahler-Ricci flow equation on a closed manifold is equivalent to the following parabolic (scalar) complex MongeAmpere equation: (2.28)
o'-P
~ = log
vt
det
(g~6 + to-y8o log det g~iI + o')'8o'-P (t») d
0
etg')'6
+ Cl (t)
for some function of time Cl (t) . By standard parabolic theory, given any Coo initial function '-Po on a complete Kahler manifold with bounded bisectional curvature, there exists a unique solution '-P (t) to (2.28) with '-P (O) = '-Po, defined on some positive time interval 0 ~ t ~ E. We also have the following. LEMMA 2.36 (The Kahler property is preserved under the Ricci flow). If (M n , J, go) is a closed Kahler manifold, then there exists a solution to the Kahler-Ricci flow 9 (t), 0 ~ t ~ E, for some E > 0 with 9 (0) = go. Furthermore 9 (2t) is a solution of the (Riemannian) Ricci flow. Also any solution 9 (t) of the (Riemannian) Ricci flow with 9 (t) = go must be Kahler (preserving the compatibility with the almost complex structure). PROOF. Given go, we can find a solution '-P (t), 0 ~ t ~ E, of (2.28) with (t) == O. From the derivation of (2.28), we know that 9 (t) defined by (2.27) is a solution of (2.26). Hence 9 (2t) is a solution of the Ricci flow. The last statement follows from the uniqueness of the initial-value problem for the Ricci flow. D Cl
REMARK 2.37. From the derivation of (2.28) it is clear that if we have a bounded C 4-solution '-P (t) for some Cl (t) on any complex manifold (regardless of completeness and compactness), then we get a C 2 -solution 9 (t) defined by (2.27) to the Kahler-Ricci flow.
2. KAHLER-RICCI FLOW
76
4.2. The normalized Kahler-Ricci flow equation. Let (Mn, J, go) be a closed Kahler manifold. Now we make the basic assumption (corresponding to the canonical case), holding for the rest of this section and the next section, that the first Chern class is a real multiple of the Kahler class, i.e., that
[pol
=
c[wol
for some c E JR. Note that this is possible only if the first Chern class has a sign, i.e., is negative definite, zero, or positive definite. Comparing (2.11) and (2.12), we find that c = ~, where r ~ 1M Rodj.Lga/ Volga (M) is the average (complex) scalar curvature, so that
r 1 -2 [wol = -2 [pol = 7rn
7r
Cl
(M) .
So r depends only on the cohomology class [wol, n, and The normalized Kahler-Ricci flow is
o
otga~ = -Ra~
(2.29)
r
+ ;:;,ga~'
Cl
(M).
for t E [0, T).
The solution of (2.29) can be converted to the solution (2.26) by scaling the metric and reparametrizing time, and vice versa (see Section 9.1 in Chapter 6 of Volume One or subsection 9.1 below). Hence, from Lemma 2.36, we know that the initial-value problem for (2.29) with 9 (0) = go has a solution for a short time. By a derivation similar to that of (2.28), we get the following parabolic (scalar) complex Monge-Ampere equation, corresponding to (2.29) with ga~ (t) = g~~ + Oa8f3'P (t) ,
(2.30)
o'P ot
det
(g~J + O-y 80'P )
= log
detg O-
r +;:;,'P - 10 + Cl (t)
'"(0
for some function of time
Cl
(t). Here 10 is defined by Ra~ (go) - ~g~~ =
Oa8f310; this is possible because [-po
+ ~wo] = O.
4.3. Basic evolution equations. Let 9 (t) be a solution of either the Kahler-Ricci flow or the normalized Kahler-Ricci flow. We define the potential function 1 = 1 (t) by (2.31) This equation is solvable since [- p + ~w] = 0 and by Lemma 2.26. Note that 1 is determined up to an additive constant. Taking the trace of (2.31), we have (2.32)
R-r='~.f.
4.
INTRODUCTION TO THE KAHLER-RICCI FLOW
77
Differentiating (2.3), we find that for both the Kahler-Ricci flow and the normalized Kahler-Ricci flow, the Christoffel symbols evolve by
8 r"{ at Ot{3 -
(2 .33)
"{8'r7 R -
-9
VOl.
{30'
The volume form and scalar curvature evolve according to the following. LEMMA 2.38 (Evolution of df-L and R for normalized flow). Under the normalized Kahler-Ricci flow (2.29),
8 8t df-L = (r - R) df-L and
-8R = flR + 1R 01.{3- 12 - -nr R. 8t
(2.34)
In particular, since fM (r - R) df-L = 0, the normalized Kahler-Ricci flow preserves the volume. PROOF. We first compute, using (2.29), that
8
(2.35)
88
8t log det 9"{8 = 9 "{ 8t 9"{8 = r - R.
Hence
8 8t df-L = (r - R) df-L.
The evolution of the Ricci tensor is (2.36)
8
- (88t
8t ROt/3 = -8Ot8{3
log det 9"{8
)8Ot8{3R. =
From this and
8R _ 01./3 8
at -
9
_
8
_
_
at ROt{3 - at 9Ot{3' ROt{3
D
we easily derive (2.34).
EXERCISE 2.39 (Evolution of R for unnormalized flow). Show that under the Kahler-Ricci flow %t 9Ot/3 = -ROt/3' we have %tdf-L = -Rdf-L, %tROt/3 = 8Ot8{3R, and
(2.37)
8 R -_ at
9
01./3 at 8 ROt{3_ + 1R {3_12 -_ flR Ot
+ 1R Ot{3_1 2 .
EXERCISE 2.40 (Total and average scalar curvature evolution). Show that if M n is closed, then under the Kahler-Ricci flow ft9Ot/3 = -ROt/3'
! 1M
Rdf-L =
1M (I ROt/312 - R2) df-L,
and hence we have
~: = (1M df-L) 1M (I ROt/312 -1
R2) df-L + r2.
78
2.
KAHLER-RICCI FLOW
REMARK 2.41 (I-dimensional normalized Kahler-Ricci flow). Recall the following facts, due to Hamilton [180], about the normalized Ricci flow on Riemannian surfaces £g = (r - R) g. Throughout this remark, Rand r denote the Riemannian scalar curvature and its average, respectively. Note that 9 (t) ~ 9 (!t) is a solution of the complex I-dimensional normalized Kahler-Ricci flow. The potential function f, defined by (2.32) and normalized suitably by an additive constant, satisfies (see Lemma 5.12 on p. 113 of Volume One) of (2.38) ot = b.f + r f. By the maximum principle, this implies If I ~ Ce Tt • The gradient quantity H ~ R - r + 1\7 fl2 satisfies (see Proposition 5.16 on p. 114 of Volume One) oH (2.39) at = b.H - 21MI2 + rH, where M ~ \7\7 f - !b.f· g. By the maximum principle, we have (see Corollary 5.17 on p. 115 of Volume One)
_CeTt ~ R - r ~ H ~ CeTt . This gives the exponential decay of IR - rl when r < O. The norm squared of the tensor M evolves by (see Corollary 5.35 on p. 130 of Volume One)
:t
(2.40)
IMI2 = b.IMI2 - 21\7 MI2 - 2R IMI2 .
Generalizing the I-dimensional formula (2.38) to higher dimensions, the potential satisfies a linear-type equation. (Strictly speaking, the equation is not linear since the Laplacian is with respect to the evolving metric.) LEMMA 2.42 (The potential f satisfies a linear-type equation). Under the normalized Kahler-Ricci flow on a closed manifold Mn, the potential function f, defined by (2.31) and normalized by an additive constant, satisfies (2.41 )
PROOF. From (2.31) we compute On[)(3
(it) =
:t (on[)(3f)
= on o-(3R -
=
! (Rn~
r0
-
~gn~)
- (
r) .
-:;;, otgn~ = on o (3 b.f + -:; , f
Since M is closed, it follows that (2.42)
~~ =
b.f + ~f + c(t)
for some function c (t) and the lemma follows from the fact that we have the freedom of adding a time-dependent constant in our choice of f (x, t). 0
4.
INTRODUCTION TO THE KAHLER-RICCI FLOW
79
COROLLARY 2.43 (Estimate for f). If Mn is closed, then, for the function f given by Lemma 2.42, we have If I :S Ge;;t.
(2.43)
This is the first hint that the Kahler-Ricci flow in the case where Cl (M) < o is the easiest and that the case where Cl (M) > 0 is the hardest. We now compute for f in Lemma 2.42, for the normalized Kahler-Ricci flow, that (2.44)
%t
IV' c.d1 2= illV'afl 2 - IV'a V' ,8f1 2 - IV'a V' ,6f1 2 + ~ IV'afl 2 .
Define
h ~ ilf + IV' afl 2 = R - r Similarly to (2.39), we have
+ IV'afl 2 .
LEMMA 2.44 (Ricci soliton gradient quantity evolution). For the normalized Kahler-Ricci flow on a closed manifold Mn,
~~
(2.45)
= ilh - IV'a V' ,8f1 2 + ~h.
PROOF. We compute
{) (R at (2.46)
r) =
il (R -
r)
+ IRa,6 12 -
=
il (R -
r)
+ lV' a V',6fl + -ilf + -n n
=
il (R -
r)
+ IV'a V' ,6f1 2 + :cn (R -
rR -:;;, 2
2r
r2
r --R n
r),
where we used (2.31) and (2.32). Equation (2.45) follows from summing this equation with (2.44). 0 COROLLARY 2.45 (Estimate for R). rt
rt
-Ge n :SR-r:SGe n ,
(2.47)
lV'fI2:s Ge;;t.
(2.48)
PROOF. By (2.46), the lower bound for R - r follows from
(!
-il) (R-r)
~ ~(R-r).
To get the upper bound for R-r, we observe that by the maximum principle, we have rt R - r :S h:S Gen. This also implies (2.48) since IV' fl2 = h - (R - r) :S h + Ge;;t. 0 REMARK 2.46 (Exponential decay when Cl < 0). When Cl (M) < 0, so that r < 0, (2.47) says that R approaches its average exponentially fast. This suggests that the Kahler-Ricci flow converges to a Kahler-Einstein metric. Indeed, this is Theorem 2.50 below.
2. KAHLER-RICCI FLOW
80
Here is an interesting equation due to Hamilton. LEMMA 2.47 (Ricci soliton vanisher evolution equation). For the potential function f in Lemma 2.42, we have
(2.49)
%t IV' 0 V' /3f12 =
~ IV' 0 V' /3f12 -
IV' "I V' 0 V' /3f12 - lV'i V' 0 V' /3f12
- 2 RoihJ V' a V'i fV' /3 V' 8f. PROOF.
By (2.42) and the commutator formulas, we have
%t (V'oV'rd)
=
V' 0 V'/3
= V'0V'/3
(~{) - (%tr~/3) V''Yf (~f + ~f) + V'oRtJiV''Yf.
On the other hand, for any function V'oV'/3~f =
f,
V'oV'/3V''YV'if = V''YV'oV' i V'/3f
= V' /v i V' 0 V' /3f - V'"I (Roi!3JV' 8f) =
1
'2 (V' "I V'i + V'i V' "I) V' 0 V' /3f -
V' oR/3JV' 8f - ROi/3JV' "I V'/d
1
- '2 (ROi V'"I V' /3f + RtJi V' 0 V'"If) since Hence
a
at (V' 0 V' /3f) = ~ V' 0 V' /3f
r
+ -;;, V' 0 V' /3f -
Roi!3JV' "I V' 8f
1
- '2 (ROi V' "I V' /3f + R/3i V' 0 V'"If)
o
and one easily derives (2.49) from this.
2.48. Find geometric applications of (2.49) in the study of the Kahler-Ricci flow. PROBLEM
EXERCISE
2.49. Show that when dime M = 1,
1V'0V'/3fI2 =
~ lV'iV'jf - ~~f9ijI2
and
Show also that (2.49) generalizes (2.40). 5. Existence and convergence of the Kahler-Ricci flow In this section we present some of the proofs of the basic global existence and convergence results for the Kahler-Ricci flow due to H.-D. Cao [46].
5.
EXISTENCE AND CONVERGENCE
81
5.1. Cao's existence and convergence theorem. When the first Chern class has a definite sign, either negative, zero, or positive, Cao proved that the normalized Kahler-Ricci flow exists for all time. Let KRF and NKRF denote the Kahler-Ricci flow and the normalized Kahler-Ricci flow, respectively.
THEOREM 2.50 (NKRF: C1 <, =, > 0 global existence). Let (Mn,go) be a closed Kahler manifold with
(1) either C1 (M) < 0, C1 (M) = 0, or C1 (M) > 0, and (2) r;;- [wo] = 21l" C1 (M). Then there exists a unique solution g(t) of the normalized Kahler-Ricci flow defined for all t E [0,00) with g(O) = go. When the first Chern class is nonpositive, Cao proved that the normalized Kahler-Ricci flow converges to a Kahler-Einstein metric in the same
Kahler class as the initial metric. THEOREM 2.51 (KRF: C1 SO convergence). Let 9 (t) be a solution of the normalized Kahler-Ricci flow, as in Theorem 2.50, with ct (M) S o. Then 9 (t) converges exponentially fast in every Ck-norm to the unique KahlerEinstein metric goo in the Kahler class [wo]. Note that the initial-value problem for the normalized Kahler-Ricci flow equation
o
r
otga{J = -Ra{J + ;,ga{J' ga{J (0) = g~{J' is equivalent to the following parabolic Monge-Ampere equation for the metric potential function
o
r
+ -
f(x, 0),
where (2.52)
and f(x,O) is the potential function of Ra{J(x, 0) - ~ga{J(x, 0) defined by (2.31). To prove Theorem 2.50 and Theorem 2.51, it suffices to prove the long-time existence of
5. EXISTENCE AND CONVERGENCE
81
5.1. Cao's existence and convergence theorem. When the first Chern class has a definite sign, either negative, zero, or positive, Cao proved that the normalized Kahler-Ricci flow exists for all time. Let KRF and NKRF denote the Kahler-Ricci flow and the normalized Kahler-Ricci flow, respectively. THEOREM 2.50 (NKRF:c1 <,=,> 0 global existence). Let (Mn,go) be a closed K iihler manifold with
(1) either C1 (M) < 0, C1 (M) = 0, or C1 (M) > 0, and (2) ~ [wol = 27TC1 (M). Then there exists a unique solution g(t) of the normalized Kahler-Ricci flow defined for all t E [0,00) with g(O) = go. When the first Chern class is nonpositive, Cao proved that the normalized Kahler-Ricci flow converges to a Kahler-Einstein metric in the same Kahler class as the initial metric. THEOREM 2.51 (KRF: C1 ::; 0 convergence). Let 9 (t) be a solution of the normalized Kahler-Ricci flow, as in Theorem 2.50, with C1 (M) ::; O. Then 9 (t) converges exponentially fast in every Ck-norm to the unique KahlerEinstein metric goo in the Kahler class [wol.
Note that the initial-value problem for the normalized Kahler-Ricci flow equation
8
r
atgai3 = -Rai3 + -:;;,gai3' gai3 (0)
= g~i3'
is equivalent to the following parabolic Monge-Ampere equation for the metric potential function
8
r
+ -
f(x, 0),
where (2.52)
and f(x,O) is the potential function of Rai3(x, 0) - ~gai3(x, 0) defined by (2.31). To prove Theorem 2.50 and Theorem 2.51, it suffices to prove the long-time existence of
2. KAHLER-RICCI FLOW
82
Moser iteration, and differential Harnack estimates (see the discussion below). The corresponding theory for systems is considerably harder, sometimes tractable only under more restrictive conditions such as the nonnegativity of the bisectional curvature (see Sections 6-8 of this chapter). Since (2.53)
we have that
v(x, t)
.
7 -
8cp
at (x, t)
is also a potential function of Ro:fj(x, t) - ~9o:fj(X, t). From (2.50), (2.53), and (2.51), we compute
8v _ 8 (8CP) _ o:fj 8 _ r 8cp 8t - - 8t 8t - -9 8t 9o:fj - -:;;, 8t r n
= 6.v +-v
with the initial condition v(x,O) = f(x,O). Therefore, if we insist, as in Lemma 2.42, that the potential function f(x, t) satisfies the heat equation -& f = 6.f + !:. f, we must have n
f = _ 8cp 8t .
(2.54)
Recall from (2.43) that following.
ifi
I~ I ::;
Ce:;'t. More precisely, we have the
LEMMA 2.52 (Time-derivative estimate for cp). r t 8cp r t -Clen ::; f (x, t) = - 8t ::; C2 en ,
(2.55)
where C I
~ -
minxEMn f(x, 0) and C 2 ~ maxxEM f(x, 0).
5.2. Proof of Theorem 2.50. Theorem 2.50 is proved via a progression of estimates which culminates with a C2,O:-estimate for cp (t) on bounded time intervals. The CO-estimate is the following. LEMMA 2.53 (CO-estimate: bound for cp-uniform when r =I 0, then (2.56)
- C;n (e:;.t
-1) ::; cp(x,t)::; C:n (e:;.t -1).
Ifr = 0, then -C2t ::; cp(x, t) ::; CIt. PROOF. For the upper bound we compute
t
8cp (t r cp(x, t) = cp(x, 0) + Jo 8t (r) dr::; Jo Cle nT dr.
CI ::;
0). If
5. EXISTENCE AND CONVERGENCE
C:.n (e;-t -
If r =1= 0, the integral on the RHS is equal to obtain CIt. Similarly for the lower bound.
83
1) . When r
= 0, we 0
REMARK 2.54. The qualitative dependence of the estimate (2.56) on the sign of r should be compared to Corollaries 2.43 and 2.45.
The next estimate, a bound for the determinant of the complex Hessian of cp, is also a straightforward application of the previous result and the estimates (2.55) and (2.56) to equation (2.50). 2.55 (Estimates for the volume form-uniform when r ~ 0). If 0, then there exists a constant C ~ 1 such that
LEMMA
r
~
(2.57)
for all x E
~ < det (903 (x, t)) < C C - det (903 (x, 0)) M and t ~ o. If r > 0, then _ en (eii-t-l) det (903 (x, t)) en (eii-t-l) e r < ~er - det (903 (x, 0))
for all x E M and t PROOF.
~
O.
Applying (2.47) to (2.35), we have
~ log det (903 (x, t))
(2.58)
at
Hence, if r
=1=
Ir _ RI ~ Ce;-t.
=
det (903(X, 0))
0, then
log
det (903 (x, t)) det (903 (x, 0))
~
Cn (.!.t ) en -1 , r
-
so that
In particular, if r < 0, then e
en r
t)) < det (9~a(X, ~/J
~
en
e - -;:- .
- det (903(X, 0))
When r
=
det g", -(x,t) det g",i3(x,O)
0, equation (2.58) is not strong enough to uniformly estimate . In this case we use
a
det (903 (x, t))
at
det (903 (x, 0))
- log
=r -
R
af = -!:l.f = - -
at '
which implies log
det (903 (x, t)) ( ( )) = det 903 x,O
f (x, t) + f (x, 0) .
2. KAHLER-RICCI FLOW
84
By the uniform bound (2.43) on
I,
we conclude
1 det (9ai3(X, t))
-::;-< for some
C~
o
1.
2.56. Alternatively, when r and (2.56) to equation (2.50), we have REMARK
log
det 9ai3(X, t) = det 9ai3(X, 0)
=1=
0, applying the estimates (2.55)
locp r I !:It (x, t) - -ncp(x, t) + I(x, 0) u
~ 11/(·,0)1100 (e;;t + le;;t - 11 + 1) ,
11/(·,0)1100' In particular, ifr < 0, then e- 2I1f (-,O)lIoo < det (9ai3(X, t)) < e2I1f (.,O)lIoo
since max{C1 ,C2 } ~
- det (9ai3(X, 0)) Lemma 2.55 is the first step towards proving that V a V i3CP is bounded and that 9ai3(X, t) is equivalent to 9ai3(x, 0), which in particular implies that 9ai3(X, t) is always positive definite. Next we estimate the trace of 9ai3(X, t) with respect to 9ai3(X, 0). Let
(2.59)
Y(x, t) ~ 9ai3 (X, 0)9ai3(X, t)
be the trace-type quantity we want to estimate. As we shall see below, a bound for Y (t) will imply a C 2-estimate for cp (t) . From (2.52) we have Y
(2.60)
= n + ~g(O)CP,
n = 9ai3 (t)9ai3(0)
+ ~g(t)CP.
Hence an estimate for Y implies an estimate for ~g(O)CP. Let Aa denote the eigenvalues of 9ai3(t) with respect to 9ai3(0). Then (2.61)
Y
= E~=lAa
and the eigenvalues of ( 8za:'tz13 ) with respect to 9ai3(0) are Aa -1. If Y ~ C, then as long as 9ai3(t) is positive-definite, Aa ~ C for each 0::. On the other hand, by Lemma 2.55, we have 1 n C ~ Aa ~ C, a=l where for r ~ 0, the constant C is independent of time, whereas for r > 0, C depends on time but remains bounded as long as the solution exists (though the bound may tend to 00 as time approaches 00). Hence there exists a constant c > 0 such that Aa ~ c, where c is independent of time for r ~ 0 and may depend on time for r > O. So indeed, 9ai3(t) remains positivedefinite as long as the solution exists and we have c ~ Aa ~ C', for some
II
5.
EXISTENCE AND CONVERGENCE
85
constant C' < 00. Thus a bound on Y shall imply an estimate of the complex Hessian of
for some C' <
00.
By abuse of notation, we shall call this the C 2 -estimate.
REMARK 2.57. A quantity similar to Y also played an important role in the later work of Donaldson [128] on the Hermitian-Einstein flow. We now turn to estimating Y. First we apply the heat operator to log Y. LEMMA 2.58. We have (2.63)
where
{)) 9"Y8(t)903(t)RO(0) ( - - ~ log Y < "Y
at
-
Y
r
+ -n'
~ ~ ~g(t) and R~/ (0) ~ 9/3"1 (0) R~8"1(0).
PROOF. From (2.59) and (2.29), we compute 1 {) -903(0)R /3-(t) Y = 03(0)_ (t) = 9 {)t og y9 {)t 03 Y0< Thus, to prove the lemma, it suffices to show that £l
(2.64)
~I
+ !:.Y n
(2.65)
Given any point x E M, we will calculate in a local holomorphic coordinate system which is normal with respect to the metric 9(0) at x, so that i}a9/31 (x, 0) = 0 and 903 (x, 0) = fJo3. To simplify notation, we adopt the convention that the quantities below are at time t, unless there is a (0) after them, in which case they are at time O. Since from (2.4), at x, _
{)2
{)2
R~8J,1 (0) = _903 (0) {)z"Y{)z8 9J,1/3 (0) = - {)z"Y{)z8 9J,1Q (0) , we compute that the Laplacian of Y at x is given by
~Y =
9"Y8
{)2
{)z"Y{)z8
(903(0)90<-) /3
_ "Y8 _ {)2 _() - 9 903 {)z"Y{)z89/30< 0
+9
"Y8 03(0) {)2 _ 9 {)z"Y{)z8903
____
=
(2.66)
9"Y8903R~t(0) + 903 (0)9"Y8 {)z"Y{)z8 903
= 9"Y8R~/ (0)903 - 903 (O)Ro3
+9
03( ) 81 >.fj {) {)_ 0 9 9 {)z8 90
where we used (2.5) and (2.67)
{)2
2. KAHLER-RICCI FLOW
86
(using the Kahler identities (2.2) to get the second equality in (2.67)). Note that
We claim that the Cauchy-Schwarz inequality gives (2.68)
IV'YI2 ~ Y
(g (ot.8 9"(8g0ji. ogo.8 OgVp ) oz"( oz8
.
By applying (2.68) to (2.66), we then obtain (2.65), as desired. To prove (2.68), we first observe that we may further assume go.8(x, t)
= >"0 80/3
is diagonal, where >"0 is defined above. The xas
=Y
LHS
of (2.68) can be written at
( (O)v(3 "(8 oji. ogo(3 OgVp) 9 9 9 oz"( oz8
.
Thus the claimed inequality (2.68) and the lemma are both proved.
0
Next we prove the key C 2-estimate (2.62) via an application of the maximum principle to (2.63). 2.59 (C2-estimate for
87
5. EXISTENCE AND CONVERGENCE
(1) (Uniform estimate when Cl ~ 0) If Cl (M) ~ 0, then there exists a constant C < 00 depending only on the initial metric such that
(2.69)
Y(x, t)
~
C
for all (x, t) EM x [0, T). Hence there exists C f < of T, such that the complex Hessian of cp satisfies
00,
independent
IVov~cp(t)lg(o) ~ C f
(2.70)
on M x [O,T). (2) (Time-dependent estimate when Cl > 0) If Cl (M) > 0, then there exist constants C and C f , both depending on 9 (0) and T < 00, such that the above estimates (2.69) and (2.70) hold on M x [0, T).
2.60. Estimate (2.70) implies IAg(O)cpl ~ C f • On the other hand, by Lemma 2.53, we have Icpl ~ Co for some Co < 00. Hence, by standard elliptic theory, we have for any a E (0,1), IIcpllcl.<> ~ Cl for some C 1 < 00 depending on a. However at this stage of the proof, it is not clear whether IHess cpl is bounded, where Hess denotes the real Hessian, but we do not need this. REMARK
By the discussion above, regarding inequality (2.62), we only need to bound Y from above. Again, we calculate in local holomorphic coordinates around x where go~(x, 0) = 80 {3, 8~"Tgo~(x, 0) = 0, and go~(x, t) = >'080{3. By (2.63) and >'0 ~ Y, we have PROOF.
(2.71)
( -a - uA) 1og Y < at
Roii"/'Y (0)
-
Y
>'0 + -r < C 1 ~ L.." -1 + -r >."/ n ,,/=1 >."/ n'
where Cl is a constant depending only on a lower bound of the bisectional curvatures of g(O). To control the bad terms on the RHS above, we consider equation (2.54) . (2.72)
(ata) - cp = - f A
Acp = -
1 L ~' n
f - n+
,,/=1
"/
where the second equality follows from (2.60). Consider the modified quantity w = logY - (C1 + l)cp. Combining (2.71) and (2.72), we have (2.73)
(ata) w<-""-+C2 - L.." >. ' --A
n
,,/=1
1
"/
where C 2 depends on C 1 and Ilflloo (if r ~ 0, then C 2 is independent of time, and if r > 0, then C 2 depends on time). By the maximum principle, (2.73) implies that w can be bounded above on M x [0, T) by a constant C depending on T.
88
2.
KAHLER-RICCI FLOW
To get a bound for w independent of T when r ~ 0, we need to work harder. When r ~ 0, we shall use the term - E"Y >..\ to dominate (from below) a function of w which approaches -00 as w ~ 00. Using equation (2.50), f = - %tc.p, and (2.61), we have
f (x, t) - f(x, 0)
det 9cr[J(X, t) (0) , et 9cr[J x,
r
+ -c.p(x, t) = -log d n
and hence
y e!(x,t)- !(x,O)+~
II : 2: (II : )~ (2: :a) 2: Acr . cr
=
a
"Y
"Y~a
"Y
n-l
a
"Y
using a standard inequality (we dropped a factor of nnl_2 since it makes the inequality easier to see). Since f and c.p are uniformly bounded, this implies that there exists a constant C3, which only depends on the initial data, so that
y
(2.74) Notice that
eW
=
~C ~
e-(Cl +l)
(2.75)
eW
3 (
:0)
n-J
We then have
~ (C4~
:J-J,
where C4 > 0 only depends on the initial data. Combining (2.73) and (2.75), we have
(ata) -
(2.76)
-
~
- 1 +C2 . w < - -1e n~ C4
-
This implies that, at any time t where W max (t) ~ (n - 1) log (C2C4 ) , we have 1tWmax (t) ~ O. Hence, by the maximum principle, sup w (x, t) xEfiA
~ max {sup w (x, 0), (n -
1) log (C2 C4 )}
xEfiA
for all t E [0, T), and hence Y is also bounded from above. When r ~ 0, both the constant C2 and the function c.p are uniformly bounded independent of T; hence Y is uniformly bounded independent of T. 0 Now we can complete the proof of Theorem 2.50. PROOF OF THEOREM 2.50. Let c.p (t), t E [0, T), be a solution of the NKRF (2.50) on a closed Kahler manifold (Mn, 90), where T is the maximal time of existence. If T < 00, then by the C 2 -estimate (Le., the estimate for the complex Hessian of c.p), the metrics 9(t) are uniformly equivalent to 9(0)
5. EXISTENCE AND CONVERGENCE
89
on the time interval [0, T). Once we have a uniform C2,a-estimate of cp on [0, T) (we shall prove the C 2 ,a- estimate in the next subsection), by choosing a time to < T close enough to T, the NKRF (2.50) with initial condition cp (to) can be solved on [to, to + E] , where Edepends on the C 2 ,a- estimate of cp but not to. If we choose to = T -~, this implies that we have a solution cp (t) for t E [0, T + ~]. This contradicts T being the maximal time of existence; hence the theorem is proved. D We shall supply the details for the C 2 ,a- estimate in the next subsection.
c
5.3. The C 2 ,a- estimate of cpo We now proceed to derive the 2 ,a_ estimate for cp, which by (2.50) and (2.54) is a solution to the complex Monge-Ampere equation: 0 log det ( gaiJ
+ CPaj3 )
= h
0 . + log det gaiJ =;= h,
and (2.77)
-
h(x, t)
= - f(x, t) + f(x, 0)
r
- -cp(x, t).
n By (2.70), there exists a constant A > 0 such that the Kahler metric gaiJ g~iJ + cp aj3 satisfies
~
(2.78) By the fact that the bounded function f satisfies t:::.f = R - r and that we have the scalar curvature bound (2.47), standard Lq theory (which only requires the uniform boundedness of the coefficient matrix (gaiJ) from above and below) implies that Ilfllco(M) + IIVV flILq(M) is bounded and hence by (2.77), Ilhllco(M) + IIVVhIILq(M) is bounded (uniformly in t when Cl (M) ::; 0) for any q < ;:0 (independent of t). Note that h is globally defined whereas h is only locally defined. However, for compactly contained open subsets U of a holomorphic coordinate chart of go, Ilhllco(u)+ IIVVhIlLq(u) and Ilhllco(u) + IIVVhIILq(U) are equivalent. Let B(R) denote the Euclidean ball of radius R centered at the origin in en. Since M is compact, there exists a finite collection of open sets {Uk}~~l and normal holomorphic coordinates Zk = {zn:=l defined on Uk (independent of t) such that No
B (3RkO) C Zk (Uk)
and
U z;;l (B (RkO)) = M, k=l
where RkO > O. Hence it suffices to prove the C2,a- estimate for cp in each open set z;;l (B (RkO)) assuming that (2.79)
Ilhllco(uk)
+ IIVVhIlLq(uk)
::; C
< 00.
From now on we work in a fixed coordinate chart. More precisely, we use Zk to push forward our discussion to the Euclidean ball B (3RkO)' For
90
2. KAHLER-RICCI FLOW
simplicity we drop the indices k in our notation below. Since 9 is Kahler, we may write 9Ct /3 locally as the complex (Hermitian) Hessian u Ct /3 of a function u. We shall show that the second derivative of u has bounded Holder norm. Now the equation reads locally as log det (u Ct /3) = h.
(2.80)
It is convenient to write log det as a function F(p), where p lies in the domain of positive definite Hermitian symmetric matrices. An important property that we shall make full use of is that F(p) = log det(p) is a concave function of p, a fact which can be easily checked. Taking the derivative of (2.80), we have
8F
- 8- uCt/3-y = h-y, PCt{3 82F
8
-8 - UCt/3-yUJ.Lii,,(
8F
+ -PCt(3 8_uCt/3-y"( = h-y"(
PCt{3 PJ.LV for each 'Y = 1, ... ,n. By the concavity of F we have (2.81)
On the other hand recall that 8F
--=U
a{3-
8Pa/3
=9
a{3-
,
where (uCt/3) is the inverse of (u a/3) = (9a/3).5 Therefore we can rewrite (2.81) as ~uw ~
(2.82)
h-y"(,
where
u-y"( and ~u denotes the Laplacian with respect to the metric 9Ct /3' For any R ~ Ro, let w
(2.83)
M(s)
~
= sup wand
m(s)
=
B(sR)
inf w. B(sR)
Also define the oscillation function:
w(sR)
~
M(s) - m(s).
The following weak form of the Harnack inequality, which holds in general for linear elliptic operators of divergence form, plays a crucial role in our estimate. One can find the proof of this result in various papers and 5T his is not different from the real version of this formula used in Volume One, which is the variation formula -logdetA = A -l)ij & Q &AQ
a
for any invertible matrix A ij .
(
a
5.
EXISTENCE AND CONVERGENCE
91
books on PDE, such as Moser [276], Morrey [274], Gilbarg and Trudinger [155], Han and Lin [195] (e.g., see Theorem 4.15 on p. 83 of [195]). THEOREM 2.61 (Harnack inequality). Let function such that
U
:
B(3Ro)
---+
lR be a C 2
1
Ao (tSa.a) ~ (ua.a) ~ Ao (tSa.a) forsomeA o E [1,00). Suppose that a nonnegative function v E W 2,2(B(3Ro)) and 9 E LQ(B(3Ro)), for some q > m/2, satisfy ~uv ~
9
in the weak sense in B(3Ro), where ~u denotes the Laplacian with respect to the metric ua.a. Then for any 0 < (j ~ T < 1 and 0 < p < m~2' there exists a constant C = C(p, q, 2n, Ao, (j, T) < 00 such that for any p ~ 2Ro, 1
(2.84)
(~r v (y)P dY) P ~ C ( inf v + i-rg-lIgIlLq(B(P») . pm JB(7'p) B(Op)
REMARK 2.62. The reason for why we can apply this theorem to (ua.a) = (ga.a) =
(g~.a+CPa.a)
is that the C 2 -estimate for cP yields (2.78). Note that
~u
(M(2) - w)
~
-h"(i and -h"(i E Lq for all q
~
00 (e.g.,
< 00), so we may apply the above theorem to M(2) -w (for example, with m = 2n, p = 2R,6 and (j = T = ~, so that (jp = Tp = R) to obtain for any q> nand 0 < p < that there exists C = C(p, q, n, A) < 00 such that q
n:1
1
r (M(2) - w (Y))PdY) P (R~ JB(R) n
~ C (M(2) -
(2.85)
M(1)
+ R 2 (q;n) Ilh"(iIILq(B(2R»)
since infB(R) (-w) = M(1). On the other hand, the concavity of F implies
F(Ui](X)) ~ F(Ui](Y))
8F
+ 8p .., (Ui] (y)) (Ui](X) -
Ui](Y))'
~J
Namely we have (2.86) The following linear algebra fact enables us to estimate w(R). 6Note that
R::; Ro.
92
2.
KAHLER-RICCI FLOW
LEMMA 2.63 (Linear algebra). There exist unitary vectors r}, . .. , rN E en with the property that, as Hermitian symmetric positive definite matrices, N
(giJ(y)) = Lav(y)rv®rv .
(2.87)
v=1
Here av(y) E lR and lA ~ av(y) ~ AA for some constant A > O. Moreover we may assume that the first n vectors r}, ... , rn form a unitary basis of en . REMARK 2.64. Let {ei}~=1 denote the standard basis for en and write rv ~ l:~=1 (rv)i ei for each 1/. By (2.87) we mean that N
(gil(y)) = L av(y) (rv)i (rv)1. v=1
EXERCISE 2.65. Prove the above linear algebra lemma. Define
Wv ~ Hess(u) (rv, rv) = uiJ(rvMrvk Now we let Mv(s) and mv(s) denote the quantities M(s) and m(s) defined by (2.83) using Wv instead of w. Then (2.86) implies N
L av(y)(wv(y) - wv(x)) ~ h(y) - h(x).
(2.88)
v=1
Choosing x E B(2R) to be a point where wdx) implies
= ml(2), this in particular
al(y)(wl(y) - Wl(X)) ~ h(y) - h(x) + L av(Y)(wv(x) - wv(y)) v~2
and hence
WI(y) - ml(2)
~ C(A, A) (RIIV'hllco + L(Mv(2) - wv(Y))) , v~2
where we used al(y) ~ Thus
lA' and av(y) ~ AA and wv(x) ~ Mv(2) for 1
{ (WI (Y) - ml(2))PdY) p (R; n iB(R)
~ C(A, A)RIIV'hllco 1
(2.89)
+ C(A, A) L (R;n v~2
( iB(R)
(Mv(2) - wv(y))PdY) P
1/
~ 2.
93
5. EXISTENCE AND CONVERGENCE
On the other hand, applying (2.85) to bound the V-norm of Mv(2) - WV , we have for each /J 2: 2, 1
r
(R; n JB(R) (Mv(2) - Wv (Y))PdY) (2.90)
:::; C ( Mv(2) - Mv(1)
P
+ R 2(q;n) II Hess(h) (rv, rv )IILq(B(2R))) .
Combining (2.89) and (2.90), we have
(2.91)
r (WI (y) - m1(2))PdY) (R; JB(R)
1
P
n
:::; C
(max (Mv(2) - Mv(l)) + RllVhllco + R 2(q;n) IIVVhIILq) . v:=::2
Now let
w(sR)
~ twv(SR) ~ tv=l (sup Wv v=l B(sR)
inf wv)
B(sR)
N
= L:(Mv(s) -mv(s)).
v=l We then have 1
r (WI(Y) - ml(2))p) P (R;n JB(R) (2.92)
:::; C (W(2R) - w(R)
+ RllVhllco + R 2(q;n) IIVVhIlLq)
On the other hand, from (2.85) we also have 1
r
(R; n JB(R) (Ml(2) - WI (y))p) P (2.93)
:::; C (W(2R) - w(R)
+ R 2(q;n) IIVVhIlLq)
.
.
94
2. KAHLER-RICCI FLOW
Putting these together, we obtain 1
wl(2R) = (
s; (
V01~(R) LIR) (Ml(2) - ml(2))p)'
V01~(R) LIR/W1 (Y) - ml(2))p
r 1
1
+ ( V01~(R)
LIR) (Ml(2) - (Y))P) , WI
~ C (W(2R) -
w(R)
+ RIIVhlico + R 2(q;n) IIVVhIILq) .
Since there is nothing special about the index 1, summing the corresponding upper bounds for w/I(2R) implies w(2R)
~ C (W(2R) -
with a different constant C
w(R)
+ RllVhllco + R 2 (q;n) IIVVhIILq)
< 00. We conclude the following.
LEMMA 2.66 (Oscillation estimate). There exists 6 < 1 (i. e., 6 = 1 such that lor any R ~ Ro we have on B(3Ro), (2.94)
w(R) ~ 6· w(2R) 2(q-n)
+ RllVhllco + R
2(q-n)
q
b)
_
IIVVhIILq.
_
Now since R q IIVVhIlLq(U) and IIVhllco(u) are bounded by (2.79), the Holder continuity of VVu on B(Ro) can be derived from (2.94) by a standard argument; see Moser [276], or Corollary 4.18 on p. 91 and Lemma 4.19 on p. 92 of Han and Lin [195], for example. Finally, the Holder continuity of VVu is equivalent to the Holder continuity of VV'P' 5.4. Proof of Theorem 2.51. Finally, we give the proof of Theorem 2.51, i.e., the proof of the convergence of the normalized Kahler-Ricci flow in the case where Cl < O. Assume, without loss of generality, that r = -n, which can be achieved by scaling the initial metric go. Notice that we have that I = - ~ satisfies
a
atl = b.g(t)1 - I and I/(x, t)1 ~ Ce- t and (2.48).) That is,
IVII (x, t)
~
Ce- t . (The last inequality is by
for some Cl < 00. 2 Now b.g(t) = gCt/3 az~az{3 and gCt{J = g~{J
+ 'PCt{J'
So the C2,Ct-estimate for
'P implies a CCt-estimate for the coefficients gCt{J. Thus we may apply the
6.
SURVEY OF SOME RESULTS FOR THE KAHLER-RICCI FLOW
95
parabolic Schauder estimate (e.g., Theorem 5 on p. 64 of Friedman [146]) to obtain Ilfll c 2 ,<>(M) ~ C2e- t for some C 2 < 00. Iterating the Schauder estimate, we have IlfII C 2m '<>(M) ~ Cme- t for some constants Cm < 00 and all mEN. This implies the estimate ~ C2 e- t and hence implies the exponential convergence of I ~ II C 2m ,<>(M)
'P (', t)
---t 'Poo (.) in Coo as t ---t 00 for some smooth function 'Poo. This proves that the normalized Kahler-Ricci flow converges in Coo to a Kahler-Einstein metric with negative scalar curvature. Theorem 2.51 is proved.
6. Survey of some results for the Kahler-Ricci How 6.1. Closed Kahler manifolds with nonnegative bisectional curvature. Using the short-time existence of the Kahler-Ricci flow, the result of Mori, Siu and Yau (Theorem 2.33) was generalized by Bando [19] when n = 3 and Mok [269] for n ~ 4.7 Mok also used techniques from algebraic geometry. THEOREM 2.67 (Kahler manifolds with nonnegative bisectional curvature). If (Mn,g) is a closed Kahler manifold with nonnegative bisectional
g)
curvature, then its universal cover (Mn, is isometrically biholomorphic to the product of complex Euclidean space, compact irreducible Hermitian symmetric spaces of rank at least 2, and complex projective spaces with Kahler metrics of nonnegative bisectional curvature. REMARK 2.68. Note that the above classification is up to isometry. Any complex projective space admits a metric with constant holomorphic sectional curvature (Le., the Fubini-Study metric). The proof of the theorem above uses the following result, proved by Bando for n = 3 and Mok for n > 4. We discuss this result further in Section 8 below. THEOREM 2.69 (KRF: nonnegative bisectional curvature is preserved). If(Mn,g(O)) is a closed Kahler manifold with nonnegative bisectional curvature, then the solution g (t) to the Kahler-Ricci flow has nonnegative bisectional curvature for all t ~ O. If in addition g (0) has positive Ricci curvature at one point, then g (t) has positive holomorphic sectional curvature and positive Ricci curvature for all t > O. Using the existence of a Kahler-Einstein metric, the convergence in the case of positive bisectional curvature was settled by Chen and Tian [87],
[88]. 7The case of nonnegative curvature operator was considered by eao and one of the authors [50].
96
2.
KAHLER-RICCI FLOW
THEOREM 2.70 (KRF: compact positive bisectional curvature). Suppose (Mn, 9 (0)) is a closed Kahler manifold with nonnegative bisectional curvature everywhere and positive bisectional curvature at a point. Then the solution 9 (t) to the normalized Kahler-Ricci flow, which has positive bisectional curvature for all t > 0, converges exponentially fast to the Fubini-Study metric of constant holomorphic sectional curvature on cpn. REMARK 2.71. Without using the existence of a Kahler-Einstein metric, Cao, Chen, and Zhu [49J proved a uniform curvature estimate (see Theorem 2.92). 6.2. Uniformization of noncom pact Kahler manifolds with nonnegative bisectional curvature. In this subsection we recall Yau's fundamental conjecture on the uniformization of complete noncompact Kahler manifolds with nonnegative bisectional curvature. CONJECTURE 2.72 (Noncompact Kahler uniformization Kc (V, W) > 0). If (Mn, g (0)) is a complete noncompact Kahler manifold with positive bisectional curvature, then M is biholomorphic to C n . Using the Kahler-Ricci flow on noncompact manifolds, Chau and Tam [65J proved the following result, which affirms Yau's conjecture in the case of bounded curvature and maximum volume growth. THEOREM 2.73 (KRF: noncompact positive bisectional curvature). If n (M , g (0)) is a complete noncompact Kahler manifold with bounded positive bisectional curvature and maximum volume growth, then M is biholomorphic to cn. There have been a number of works on the Kahler-Ricci flow on noncompact manifolds with positive bisectional curvature. For example, the reader may consult Shi [331], Tam and one of the authors [290], [292J, and Chen and Zhu [79], [80J. 6.3. Limiting behavior of the Kahler-Ricci flow on closed manifolds. There are also the following results about the limiting behavior of the Kahler-Ricci flow due to Sesum [323]. THEOREM 2.74. If (Mn, 9 (t)) , t E [0,00), is a solution to the KahlerRicci flow on a closed manifold with uniformly bounded Ricci curvature, then for any sequence ti -+ 00 there exists a subsequence such that (M, g (t + ti)) converges to (M~,goo (t)), where goo (t) is a solution to the Kahler-Ricci flow. The convergence is outside a set of real codimension 4. When n = 2, Sesum improved the above result to the following. THEOREM 2.75. If (M 2 ,g(t)) , t E [0,00), is a solution to the KahlerRicci flow on a closed manifold with uniformly bounded Ricci curvature, then for any sequence ti -+ 00 there exists a subsequence such that (M, 9 (t + ti)) converges to (M ~, goo (t)) , where goo (t) is a K ahler-Ricci soliton. The convergence is outside a finite number of points.
7. EXAMPLES OF KAHLER-RICCI SOLITONS
97
For some other recent work on the Kahler-Ricci flow the reader is referred to Phong-Sturm [305], [306], Chen [84J, Chen and Li [85], Song and Tian [337], Cascini and La Nave [60]' Tian and Zhang [351]' and [234J. 7. Examples of Kahler-Ricci solitons In this section, we provide a brief and regrettably incomplete sampling of some results on Kahler-Ricci solitons. These special solutions model singularities of the Kahler-Ricci flow. A Kahler-Ricci soliton is a Kahler manifold (Mn, g, J) such that the soliton structure equation
(2.95)
1
Rc+,\g + 2LXg
=
a
holds for some constant ,\ E IR and some real vector field X which is an infinitesimal automorphism (2.1) of the complex structure J. Note that X is an infinitesimal automorphism if and only if its (1, a)-part is holomorphic: a = '\7 o.X~ = a~'" Xf3. One imposes this requirement for the following reason. As we saw in Lemma 2.36, a solution of Ricci flow that starts with a Kahler metric on a complex manifold remains Kahler with respect to the same complex structure. On the other hand, if 'Pt is any family of diffeomorphisms of M, then each pullback 'Pt (g) is Kahler with respect to the complex structure 'Pt(J). Now consider the evolving metric h(t) := (1 + ,\t)'Ptg, where 'Pt is the family of diffeomorphisms generated by 2(1~At) X. If X is an infinitesimal automorphism of the complex structure, then 'Pt (J) = J, which implies that h( t) remains Kahler with respect to the same complex structure. Furthermore, using (2.95), it is easy to see that h solves the Kahler-Ricci flow:
gth = - Rc(h). One may also define a Kahler-Ricci soliton to be a Kahler manifold (M n , g, J) together with a constant ,\ E IR and a real vector field X satisfying the complex soliton equation (2.96) Equation (2.96) is equivalent to the conjunction of equation (2.95) and the statement that X is holomorphic. Notice that if we restrict our attention to gradient solitons (so that X is the gradient of a real-valued function), then (2.96) is equivalent to (2.95) without any extra hypotheses. (See §2.2 of [142J for the detailed argument.) 7.1. Existence and uniqueness. Any Kahler metric satisfying (2.96) with X = a is Kahler-Einstein. In this sense, Kahler-Einstein metrics may be regarded as trivial Kahler-Ricci solitons. 8 So if no Kahler-Einstein metric exists, a natural replacement is a Kahler-Ricci soliton. In fact, existence of a Kahler-Einstein metric and a nontrivial gradient Kahler-Ricci soliton 80f course, there is nothing 'trivial' about Kii.hl.er-Einstein metrics!
98
2.
KAHLER-RICCI FLOW
are mutually exclusive: applying the Futaki functional F[wJ to the holomorphic vector field X = grad f, one gets
Moreover, Kahler-Ricci solitons on a compact Kahler manifold (Mn, J) are unique up to holomorphic automorphisms. (See Tian and Zhu [352, 353, 354J.) Specifically, we have the following theorem. THEOREM 2.76 (Uniqueness of Kahler-Ricci solitons). Let (Mn, J) be a compact Kahler manifold. If metrics 9 and 9' on M satisfy (2.95) with respect to holomorphic vector fields X and X', respectively, then there is an element CT in the identity component of the holomorphic automorphism group such that 9 = CT*g' and X = (CT-I)*X'. For recent results on uniqueness and other properties of noncom pact Kahler-Ricci solitons, see [63, 64J, [35J and [78J. One might ask, therefore, whether there exists either a Kahler-Einstein metric or else a Kahler-Ricci soliton on every compact Kahler manifold Mn with CI (M) > O. The answer is yes if n ~ 2. A compact complex surface with CI > 0 is p2#k p2 for some k E {O, 1, ... , 8}. (Here and below, pn = cpn is complex projective space.) A Kahler-Einstein metric exists for k = 0 and 3 ~ k ~ 8. (See Theorem 2.34.) In the remaining cases k = 1,2, there is a (non-Einstein) Kahler-Ricci soliton. (See [239], [366], [47], and Section 7.2 below.) In higher dimensions, however, the answer is no. There exist 3-dimensional compact complex manifolds that admit no Kahler-Einstein metric and no holomorphic vector fields, hence no KahlerRicci soliton structure. (See [346, §7J as well as [215, 216J and [278J.) More generally, Tian and Zhu have exhibited a holomorphic invariant that generalizes the Futaki invariant and acts as an obstruction to the existence of a Kahler-Ricci soliton metric [354J on a compact complex manifold (Mn, J). 7.2. The Koiso solitons. As was noted in Proposition 1.13 or Proposition A.32, all compact steady or expanding solitons are Einstein. This is not true for shrinking solitons. The first examples of nontrivial (Le. nonEinstein) compact shrinking solitons were discovered by Koiso [239J and independently by Cao [47J. These are Kahler metrics on certain k-twisted projective-line bundles pI ~ Fk -* pn-I first described by Calabi [44J. We will discuss their construction in considerable detail, because it serves as a prototype for later examples. We begin with Calabi's bundle construction. pn-I is covered by n charts (CPa: Ua ~ cn-I), whereUa = {[XI, ... ,XnJ E pn-I: Xa =F O} Xa-l Xa±l ~) • (lXT . an d CPa.. [Xl"",Xn J f---t (~ , ••• , - - , , ••• , vve wn't e Xa InXa: Xo; Xo: Xo. a stead of x here and in the next paragraph in order to simplify some formulas below.) In particular, one may define complex projective space by
7. EXAMPLES OF KAHLER-RICCI SOLITONS
99
lpm-l = (Il~=ICPo(Uo))/~, where, for example,
CPI (UI ) :3 (Zl, ... , Zn-l )
~
1 Z2 Zn-l) ( -, -, ... , - E CP2 (U2). Zl Zl Zl
Given kEN, we formally identify pI = e u {oo} and define the k-twisted bundle :Fk = (Il~=l (Uo X pI)) / "', where Uo x pI :3 ([Xl, .. . , Xnl;~) '" ([YI, .. . , Ynl; 1]) E U(3 X pI if and only if [Xl' ... ' xnl = [YI, ... , Ynl and 1] = (~)k~ for each Q. Equivalently, one may define where, for example, 1 Z2 Zn-l k I (-, -, ... , - ; Zl () E CP2(U2) x p . Zl Zl Zl Notice that 80 = {[Xl, ... ,xnl; o} and 8 00 = {[Xl, ... , xnl; oo} are two global sections of Fk. The key to constructing Kahler-Ricci solitons on F'J: (as well as examples on other topologies to be considered below) will be to find a Kahler potential on en \ {o} satisfying certain symmetries and boundary conditions. To see = Fk\(80 U 8 00 ) and define 'IjJ : en\{O} - t so that why this is so, let
CPI(UI ) x p
I
:3 (Zl, ... , Zn-l;
()
~
Fr:
Fr:
'IjJ: (XI, ... ,Xn ) f---+ ([xI, ... ,xnl;x~) if Xo =1= O. It is easy to see that ([Xl, ... , xnl; X~) '" ([Xl, ... , xnl; x~) whenever Xo =1= 0 and x(3 =1= 0, hence that 'IjJ is well defined. The map 'IjJ is clearly surjective. If 'IjJ(XI, ... ,xn) = 'IjJ(YI, ... ,Yn), where, say Xo =1= 0 and Y(3 =1= 0, then
(Xl, ... , Xa-l , Xa+l , ... , Xn; X~) ~ (YI , ... , Y(3-1 , Y(3+1 , ... , Yn; Y~). Xa Xa Xa Xa Y(3 Y(3 Y(3 Y(3 The equivalence relation ~ then implies that Y~ = x~, hence that Y(3 = (}x(3 for some k-th root of unity (). Because [Xl, ... , xnl = [YI, ... , Yn], it follows that Y'Y = (}x'Y for all T = 1, ... , n, hence that 'IjJ is a k-to-one map. Therefore, a Kahler potential P on e n \ {O} will induce a well-defined Kahler metric on provided that 8ap((}xI, . .. , (}x n ) = 8ap(XI, ... , xn). With these considerations in mind, our method will be to construct a suitable Kahler potential P : en \ {O} - t IR whose asymptotics as Izi - t 0 and Izi - t 00 ensure that the induced metric extends smoothly to 8 0 and 8 00 • If we are interested in shrinking solitons, what properties should P possess? Let's suppose that P determines a Kahler metric g. As above, we take the Kahler and Ricci forms to be w = Agai3 dz o 1\ dz(3 and p = A Rai3 dz a 1\ dz(3, respectively. Then (locally) we have
Fr:
82
g-a(3 - 8za8z(3 P
2. KAHLER-RICCI FLOW
100
and
82
= - 8z Ci 8z!3 log det g,
RCi[J
exactly as in (2.6). If Q denotes the soliton potential function with gradient vector field X, then equation (2.96) reduces to
82
8zCi 8z!3 (log det 9 - Q - AP) = O. Because we are interested in shrinking solitons, we may assume that A < O. Then (modifying the Kahler potential P by an element in the kernel of V'V if necessary) we may assume that Q = log det 9 - AP. This will give us a soliton provided X = grad Q is holomorphic, that is, provided that
o = ~X!3 =~ ( !3'Y~Q) . 8zCi 8zCi 9 8z'Y Substituting Q = log det 9 - AP, we obtain a single fourth-order equation for the scalar function P, namely (2.97)
8~Ci
[gf3'Y
8~'Y (log det 9 -
AP)]
= O.
To proceed, we adopt the Ansatz that the potential P is invariant under the natural U(n) action in the sense that it is a function of r = log I:~=1 IzCi l2 alone. In this case, setting cP = Pr, we have (2.98)
gCi[J
= e- rcp8Ci!3 + e- 2r (CPr - cp)ZCi Z!3,
so our P will be a Kahler potential if and only if cP and CPr are everywhere positive. Now (following [47] and [142]) we can write (2.97) as the fourthorder ODE n n n Prrrr - 2P;rr T + nrrrr - ( n - 1) P;r p2 + 1\'( rrrrrr - p2rr ) = O. rr r We shall see that only two of the four arbitrary constants in its solution are geometrically significant. Q = ~ZCi will be holomorTo simplify (2.99), notice that XCi = gCi[J~~ z rrr phic if and only if Qr = p,Prr for some p, E lR. Since 9 will be Kahler-Einstein if X = 0, we may assume that p, i= O. Substituting Q = log det 9 - AP, one then obtains
( ) 2.99
(logCPr)r + (n -l)(logCP)r - P,CPr - Acp - n
= 0,
which is a second-order equation for cP, hence a third-order equation for P. (This integration can also be accomplished by standard ODE techniques.) Because CPr > 0 everywhere, one may regard r as a function of cP and hence F(cp). One finds (remarkably) that F satisfies a linear may write CPr equation
n-l
F' + (-cP
-
p,)F - (n+ Acp)
= 0,
7.
EXAMPLES OF KAHLER-RICCI SOLITONS
101
whose solution is
r.pr = r.pl-neJ.l.'P(1/ + >..In + nIn-l), where
1/
is another arbitrary constant and In
r.pr = I/r.p
I-n
= J r.pne-J.l.'P dr.p. This leads to
J.I.'P >.. >.. + /1. ~ n! j"J+1-n e - -r.p - l+n L..J -., /1. 'I-' , /1.
/1.
j=O
J.
which is a separable first-order equation for r.p. The third and fourth arbitrary constants arise when one finds the implicit solution r( r.p) and then integrates r.p to obtain P. These are geometrically insignificant, because they disappear in (2.97); the two geometric degrees of freedom are the parameters /1. i= 0 and 1/. In summary, what we have thus far accomplished is to construct a potential function for a U(n)-invariant Kahler-Ricci soliton metric (2.98) on Cn\{O}, hence a possibly incomplete Kahler-Ricci soliton metric on What remains is to choose /1. and 1/ in order to get a complete metric on FJ:. Our choices of the two parameters will be determined by the two boundary conditions as r -4 ±oo. Since r.pr > 0, we may define a < b E [0,00] by a = limr---+_oor.p(r) and b = limr-+oo r.p(r). If a > 0, one may write P(r) = ar + p(e Qr ) in a neighborhood of Izl = 0, with p smooth at zero, p(O) = 0, p'(O) > O. Similarly, if b < 00, one may write P(r) = br + q(e-f3r ) in a neighborhood of Izl = 00, with q smooth at zero, q(O) = 0, q'(O) > O. One then takes advantage of the following observation.
Jr.
LEMMA 2.77 (Calabi). Assume that k > 0 and a, bE (0,00). (1) When a = k, the potential P(r) = ar + p(e Qr ) induces a smooth Kahler metric on a neighborhood of So in FJ:. Any pI in So has area a7r. (2) When (3 = k, the potential P(r) = br + q(e- f3r ) induces a smooth Kahler metric on a neighborhood of Soo in FJ:. Any pI in Soo has area b7r. For the proof, see [44] or [142, Lemma 4.2]. With more work, one finds that it is possible to satisfy both boundary conditions by appropriate choices of /1. and 1/. (See [47] or adapt the arguments in [142, §4.1].) Normalizing by fixing>.. = -1, these choices yield a = n - k and b = n + k. Since one needs a > 0, one obtains a unique gradient shrinking Kahler-Ricci soliton on FJ: for each k = 1, ... , n - 1. These are the Koiso solitons. 7.3. Other U(n)-invariant solitons. The construction we have described in Section 7.2 above has natural generalizations allowing the discovery of other explicit Kahler-Ricci soliton examples. Namely, one searches the 2-dimensional (/1., 1/) parameter space of U (n)- invariant Kahler potentials
102
2. KAHLER-RICCI FLOW
P(r) on en\{O} for those whose behavior at the boundary izl = 0 implies that 1_: the metric is completed by adding a smooth point at r = -00; 2_: the metric is completed by adding an orbifold point at r = -00; 3_: the metric is completed by adding a Ipm-l at r = -00; or 4_: the metric is complete as r - t -00; and whose behavior at the boundary Izl = 00 implies that 1+: the metric is completed by adding a smooth point at r = +00; 2+: the metric is completed by adding an orbifold point at r = +00; 3+: the metric is completed by adding a pn-l at r = +00; or 4+: the metric is complete as r - t +00. Of course, not all combinations of these alternatives are globally compatible. For example, it is easy to see that the growth condition cpr > 0 prohibits completing the metric by adding a Ipm-l at Izl = 0 and a smooth point at Izl = 00. Nonetheless, this has been a productive line of research. In the remainder of this section, we will survey some of its results. It is possible to add a smooth point at Izl = 0 and to construct a unique steady Kahler-Ricci soliton on en that is complete as Izl - t 00. In complex dimension n = 1, this is just the cigar soliton discovered by Hamilton and discussed in Chapter 2 of Volume One; the examples in higher dimensions are due to Cao [47]. These solitons have the following asymptotic behavior: in the sphere s2n-1 at metric distance r » 0 from Izl = 0, the Hopf fibers U(1) . z have diameter 0(1), while the pn-l direction has diameter O(Vr). Accordingly, one calls this cigar-paraboloid behavior. It is also possible to add a smooth point at Izl = 0 and to construct expanding Kahler-Ricci solitons on en that are complete as Izl - t 00. There is in fact a 1-parameter family (en,go)o>o of such examples in each dimension, due to Cao [48]. (It is a heuristic principle that expanding solitons are easier to find than their shrinking cousins. For these examples, satisfying the boundary condition at Izl = 0 reduces the parameter space by one dimension, but completion as Izl - t 00 comes for free.) Each soliton (en, go) is asymptotic as Iz I - t 00 to the Kahler cone (en \ {O}, go), where the metric go is induced by the Kahler potential P(r) = eOr jB. For each k = 2,3, ... , the authors of [142] add an orbifold point at Izl = 0 and a Ipm-l at Izl = 00 to construct a unique shrinking Kahler-Ricci soliton on an orbifold, which is called Yk' The compact orbifold Yk may be regarded as Ipm jZk branched over the origin and the pn-l at infinity. The orbifold singularity at the origin is modeled on en jZk. For each dimension n ~ 2 and k = 1, ... , n -1, the authors of [142] add a k-twisted pn-l at Izl = 0 and construct a unique shrinking Kahler-Ricci soliton metric that is complete as Izl - t 00. The resulting soliton has the topology of the complex line bundle e ~ L"!:..k - pn-l characterized by (CI' [E]) = -k, where CI is the first Chern class of the bundle and E ~ pI is a positively-oriented generator of H2(pn-l; Z). (For example, the total
8.
KAHLER-RICCI FLOW WITH NONNEGATIVE BISECTIONAL CURVATURE
103
space of L"!:..I is simply Cn blown up at the origin.) As Izl --t 00, the soliton metric is asymptotic to a Kahler cone (C n\{0},90)/Zk, where 0 = O(n,k). For each dimension n 2: 2, Cao [47] adds an n-twisted IF-I at Izl = 0 and constructs a unique complete steady Kahler-Ricci soliton on the total space of the bundle C '---t L"!:..n - pn-I. The metric exhibits cigar-paraboloid behavior at infinity. For each dimension n 2: 2 and k = n + 1, ... , the authors of [142] add a pn-I at Izl = 0 and construct a I-parameter family of complete expanding Kahler-Ricci solitons on C '---t L"!:..k - IF-I. The solutions are parameterized by 0 > 0, where (cn\{0},90)/Zk is the asymptotic Kahler cone at infinity.
8. Kahler-Ricci flow with nonnegative bisectional curvature The study of the Kahler-Ricci flow of Kahler metrics with nonnegative bisectional curvature is somewhat analogous to the study of the Riemannian Ricci flow of metrics with nonnegative curvature operator. One aim is to uniformize Kahler metrics with nonnegative bisectional curvature in both the compact and noncompact setting. In particular, one would like to flow such metrics to canonical metrics, or to infer the existence of canonical metrics from the long-time behavior of the flow. One would also like to deduce properties of the underlying complex structure of the Kahler manifold, and when possible, classify the manifold up to biholomorphism.
8.1. Nonnegative bisectional curvature is preserved. Consider the Kahler-Ricci flow 9a-'!3 = -RafJ. We shall prove that the Kahler-Ricci flow preserves the nonnegativity of the bisectional curvature. As in the real case, the key is Hamilton's weak maximum principle for tensors. This result was proved first by Bando for n S 3 and by Mok in any dimension. (See Theorem 2.69 above.) The result was also extended to the complete noncompact case by W.-X. Shi under the additional assumption of the bisectional curvature being bounded. We say that a Kahler metric has quasi-positive Ricci curvature if the Ricci curvature is nonnegative everywhere and positive at some point.
gt
THEOREM 2.78 (Nonnegative bisectional curvature preserved). The nonnegativity of the bisectional curvature is preserved under the Kahler-Ricci flow on closed Kahler manifolds. Moreover, if the initial metric also has quasi-positive Ricci curvature, then both the Ricci curvature and the holomorphic sectional curvature are positive for metrics at positive time. The basic computation in the proof of the above result is the following.
104
2.
KAHLER-RICCI FLOW
PROPOSITION 2.79 (Evolution equation for the curvature). Under the Kahler-Ricci flow, (2.100)
(:t - ~)
Ro.iJrJ
= Ro.jiIlJ R J.1.i3'""(V -
Ro.ji'""(vR J.1.i3I1J
+ Ro.i3l1ji R J.1.v'""(J
1
- "2 (Ro.ji R J.1.i3'""(J + RJ.1.i3 R o.ji'""(J + ~jiRo.i3J.1.J + RJ.1.J R o.i3'""(ji) . REMARK 2.80. The Riemannian analogue of this formula is given by Lemma 6.15 on p. 179 of [108].
In the proof of the proposition we find it convenient to use a formula relating ordinary derivatives and covariant derivatives at the center of normal holomorphic coordinates. LEMMA 2.81 (Relation between ordinary and covariant derivatives). If 'fJ is a closed (1, I)-form, then, at the center of normal holomorphic coordi-
nates, we have
(2.101)
{p
\7 i3 \7 0. 'fJ'""(J
=
{)zo.{)z{3 'fJ'""(J + 'fJ)..J R o.i3'""()'" {)2
(2.102)
\7 0. \7 i3'fJ'""(J
=
{)zo.{)z{3 'fJ'""(J + 'fJ'""().,Ro.i3)..J.
PROOF. We compute that at the center of normal holomorphic coord inates, \7i3 \7 0. 'fJ'""(J = {)i3 \1 0. 'fJ'""(J - r~o \7 0. 'fJ,""(e
= {)i3 ({)o.'fJ'""(J - r;'""('fJgJ) = {)i3{)o.'fJ'""(J - {)i3 r ;'""('fJgJ {)2
{)zo.{)z{3 'fJ'""(J
+ R~i3'""('fJgJ'
where we used (2.4) in the last line; this proves (2.101). Note that (2.102) 0 is just the conjugate of (2.101). Now we give the PROOF OF PROPOSITION 2.79. We compute the evolution equation for Ro.i3'""(J at any point x and time t using normal holomorphic coordinates {zo.} centered at x with respect to 9 (t) . In such coordinates, a;;~ (x, t) = Recall from (2.5) that R ____ {)2go.i3 o.{3'""(O {) z'""( {)ZO
+ 9
pu {)go.u {)gpi3 {)z'""( {) ZO .
o.
8. KAHLER-RICCI FLOW WITH NONNEGATIVE BISECTIONAL CURVATURE 105
This implies
8
8 (8) 8tga~
= 'V"I 'V &Ra~ -
(2.103)
8 2
2
8t Raih & = - 8z"l8z0
=
8Z"l8zoRa~
Ra>..R"I&>"~'
where we used (2.102). Since (2.103) is tensorial, it holds in any holomorphic coordinate system. We wish to compare the above formula with t:.Ra~"I& ~
1
2" ('V /1- 'V jl + 'V jl 'V /1-) Ra~"I&'
To this end we compute (apply the second Bianchi identity (2.9) and commute covariant derivatives (Exercise 2.22)) 'V"I'VJRa~ = 'V"I'V&Ra~/1-jl = 'V"I'VjlRa~/1-&
= 'V jl 'V /1-Ra~"I& - ~jl/1-iJRa~v&
~jlaiJRv~/1-&
+ R"Ijlv~RaiJ/1-&
+ R"Ijlv&Ra~/1-iJ
and
'V jl 'V /1-Ra~"I& = 'V /1- 'V jlRa~"I& + R/1-jlaiJRv~"I& - R/1-jlv~RaiJ"I&
+ R/1-jl"liJRa~v& - R/1-jlv&Ra~"IiJ = 'V /1- 'V jlRa~"I& + RaiJRv~"I& + R"IiJRa~vJ - Rv&Ra~"IiJ'
Rv~RaiJ"I&
Combining the formulas above yields
8 8t Ra~"I& = 'V jl 'V /1-Ra~"I&
- ~jlaiJRv~/1-&
+ R"Ijlv~RaiJ/1-&
+ R"IjlvJRa~/1-iJ - Ra>..R"I&>"~ = t:.Ra~"I& - ~jlaiJRv~/1-& + R"Ijlv~RaiJ/1-& + R"IjlvJRa~/1-iJ - ~iJRa~v&
1
- 2" (RaiJRv~"IJ + Rv~RaiJ"(& + ~iJRa~v& + Rv&Ra~"IiJ)
,
o
and the proposition follows.
As a consequence of the proposition we have the following evolution equations for the bisectional curvature and the Ricci tensor. COROLLARY
(! -
2.82 (Bisectional curvature evolution).
t:. ) RaOt"l'Y =
t
(IRajlv'Y12 - IR ajl"liJl2
+ RaOtvjlR/1-iJ"I'Y)
/1-,v=l n
(2.104)
- L Re (RajlR/1-Ot"l'Y + ~jlRaOt/1-'Y) . /1-=1
Here Re(A) = ~(A +..4) denotes the real part of a complex number A.
2. KAHLER-RICCI FLOW
106
PROOF.
Indeed, substituting j3
= a and 8 = l' in (2.100), we have
(%t - A ) Rcxory'Y = RCXilv'YRJ.l.c,'Yv - RCXil'YvRJ.l.c,v'Y
+ Rcxc,vilRJ.l.v'Y'Y
1
- .2 (RcxilRJ.l.c,'Y'Y + RJ.l.c,Rcxil'Y'Y + ~ilRcxc,J.I.'Y + RJ.I.'YRcxc,'Yil) . o COROLLARY 2.83 (Ricci tensor evolution). The Ricci tensor satisfies the Lichnerowicz heat equation:
a
at RcxfJ = ARcxfJ + RcxfJ'YiS R 6'Y -
(2.105)
Rcx'YR'YfJ
= ALRcxfJ·
REMARK 2.84. More generally, we say that a real (1, I)-tensor hcxfJ satisfies the Lichnerowicz heat equation if a 1 1 athcxfJ = ALhcxfJ ~ AhcxfJ + RcxfJ'YiS h6'Y - .2 Rcx'Yh'YfJ - .2 R 'YfJ hcx'Y'
See also (2.22). PROOF.
Summing (2.100) over 'Y = 8 from 1 to n, we have
(%t - A ) Rcxi3 =
t
(%t - A ) Rcxi3'Y'Y
+ R6'YRcxfJ'YiS
k=l
= RCXilV'YRJ.l.fJ'Yv -
RCXil'YvRJ.l.fJv'Y
+ Rcxi3vilRp.v + RcxfJ'YiSR6'Y
1
- .2 (RCXil R p.i3 + Rp.i3 Rcxil + ~ilRcxfJp.'Y + Rp.'YRcxfJ'Yil) = RcxfJ'YiS R 6'Y - R cxil RJ.l.i3'
o
after cancelling terms to get the last equality.
REMARK 2.85. Equation (2.105) may also be derived from (2.36), (2.10) and commuting covariant derivatives. In particular, 1 ARcxi3 = .2 (V'Y V'Y + V'Y v\) Rcxi3
1
=
1
.2 V'Y V fJRcx'Y + .2 V'Y V cxR'YfJ 1
1
= V cx V fJR - .2 R'YfJcxiS R 6'Y + .2 R6fJ R cxiS 1
(2.106) That is,
1
+ .2 RcxiS R 6i3 - .2 R'Ycxj36 R 'YiS = V cx V i3 R - RcxfJ'YiSR6'Y + RcxiSR6fJ. V cx V fJR = ALRcxfJ·
Based on the evolution equation (2.104) and Hamilton's maximum principle for tensors (see Chapter 4 of Volume One or Part II of this volume) we present the following.
8.
KAHLER-RICCI FLOW WITH NONNEGATIVE BISECTIONAL CURVATURE
107
PROOF OF THEOREM 2.78. (I) We first prove that the nonnegativity of the bisectional curvature is preserved under the flow. Analogous to Theorem 4.6 on p. 97 of Volume One, by the Kahler version of the maximum principle for tensors (see Proposition 1 in §4 of [19]), we need to show that the quadratic on the RHS of (2.104), i.e., (2.107)
QOdi'Y"Y
~
n
L (IRctjLlI"Y1 p.,I1=l
2 -IRctjL")'iiI 2
+ RctiilljLRp.ii")'''Y)
n
- L Re (RctjLRp.ii")'''Y + ~jLRctiiP."Y) , p.=1
satisfies the null eigenvector assumption. That is, we assume Rctii,,),"Y = 0 for some a and 'Y at some point x, and we shall prove that (2.108) at x. First observe that since Rctii'Y"Y = 0 at x and the bisectional curvatures are nonnegative, we have at x, n
n
L RctjLRp.ii")'''Y = p.=1 L ~jLRctiip."Y = o.
p.=1
By (2.107), in order to prove (2.108) at x, it suffices to show that n
n
L RctiilljLRp.ii,,),"Y ~ p.,I1=l L (I R ctjL")'iiI p.,I1=l
(2.109)
2
-IRctjLlI"Y1 2 ) .
We shall prove (2.109) below, but first we show how the positivity of the Ricci tensor and holomorphic sectional curvatures follow from the quasipositivity of the Ricci curvature at t = O. (2) Recall that the Ricci tensor Rct/3 satisfies the Lichnerowicz heat equation, so that by taking a = {3 in (2.105), we have
a
at R ctii =
(2.110)
b.Rctii
+ Rctii'Y8 R 8"Y -
Rct"Y~ii'
Since the nonnegativity of the bisectional curvature is preserved, by applying Hamilton's strong maximum principle for tensors to (2.110) (see Theorem A.53 and also Part II of this volume), the Ricci tensor becomes positive for all positive time. Now suppose there exists a space-time point (xo, to), with to > 0, at which some holomorphic sectional curvature R ctiictii is zero. Since the holomorphic sectional curvature is nonnegative everywhere, by (2.104), we have at (xo, to),
o~
(:t - b.)
R ctiictii
=
t
p.,I1=l
(2 1R ctiilljL 12 - IRctjLctiil 2) ,
2. KAHLER-RICCI FLOW
108
where we used I::=1 Re (Rap.R/-,aaa + Rap.Raa/-,a) other hand, by (2.109) with Q = I, we have n
=
a at (xo, to). On the
n
L
L
IR aavp.12 ~
IR ap.ailI 2 .
/-"v=l
Hence we conclude Raavp. = RaP.ail = a at (xo, to) for all j.L, v. This in turn implies Raa = a, which is a contradiction. (1) continued. We now verify (2.109). Consider the following Hermitian symmetric form defined by the bisectional curvatures: (-a a +sX'-a a +sX'-a a +sY'-a a +sY ) ~a Q(X,y,s)~Rme za
~
~
~
for X, Y E r 1,0 M and s E R At a point where the bisectional curvatures are nonnegative and Raa'Yi = a, we have Q(X, Y, s) ~ a and Q(X, Y, a) = a for all X, Y E 1,0 M and s E R Therefore the second variation at s = a is nonnegative:
r
d2
a:::; ds 2
1
= Rme
_
8=0
Q(X, Y, s)
( -a8) + X,X, az'Y' az'Y
+ 2 Re ( Rme
( X,
Rme
a~a ' Y, a~'Y )
(a8-) Y, Y aza' aza'
+ Rme ( X,
a~a' a~'Y' Y) )
.
In terms of a unitary (1, a)-frame {ei}i=l' we may write this as Ri3'YiXi Xj
+ 2 Re ( RiajiXiyj + Ria'Y3XiYj) + Raai3 yiyj ~ a,
where X ~ I:~1 Xiei and Y ~
I:j=l yjej. By Lemma 2.86 below, we have
n
n
L Ri3'YiRaaj'i ~ L (l~ajiI2 -I Ria'Y31 2) . i,j=l i,j=l The claimed inequality (2.109) follows and this completes the proof of Theorem 2.78. D LEMMA 2.86. Let Q(X, Y) be a Hermitian symmetric quadratic form
defined by Q(X, Y) = Ai3Xi Xj
+ 2 Re ( BijXiyj + Di3XiYj) + Gi] yiyj.
If Q is semi-positive definite, then n
L
i,j=l
Ai3 Gj'i
~
n
L
i,j=l
(I B ij12 -I Di312).
9.
MATRIX DIFFERENTIAL HARNACK ESTIMATE
109
For the proof of this lemma, which is elementary in nature, we refer the reader to Mok [269]. 9. Matrix differential Harnack estimate for the Kahler-Ricci flow In this and the next section we discuss various differential Harnack estimates for the Kahler-Ricci flow and their geometric applications. Differential Harnack estimates for the Riemannian Ricci flow will be discussed in Part II. For Kahler-Ricci flow, a fundamental result is H.-D. Cao's differential Harnack estimate for solutions with nonnegative bisectional curvature (see [46]). Define (2.111) a J RaP Z {X)aP ~ at RaP + RaiRyP + '\l'YRaPX'Y + '\l iRaPX'Y + RaP'YJX'Y X + -tfor any {l,O)-vector X
= X'Y 8~"" and where Xi ~ X'Y.
THEOREM 2.87 (Kahler matrix Harnack estimate). If (Mn,g{t)) is a complete solution to the Kahler-Ricci flow with bounded nonnegative bisectional curvature, then (2.112) for any {I, O)-vector X.
This result may be considered as the space-time analogue of Theorem 2.78. We shall also see a similar analogy for the Riemannian Ricci flow in Part II, where Hamilton's matrix differential Harnack estimate will appear as the space-time analogue of the result that nonnegative curvature operator is preserved under the Ricci flow. 9.1. Trace differential Harnack estimate for the Kahler-Ricci flow. Taking the trace of the estimate (2.112) leads to the so-called trace differential Harnack estimate, after applying the second Bianchi identity. COROLLARY 2.88 (The Kahler trace differential Harnack estimate). Let (Mn,g{t)) be a complete solution to the Kahler-Ricci flow with bounded nonnegative bisectional curvature. Then (2.113) PROOF. By (2.112) we have O ~ 9apZa(3- -- 9ap
R ata R a(3- + 1R a(3_1 2 + ~v 'Y RX'Y + ~v i RXi + R'YfJ-X'YXJ + t'
and (2.113) follows from (2.37).
0
2. KAHLER-RICCI FLOW
110
> 0, the {l,O)-vector minimizing the LHS of (2.113) is X'Y = I - (Rc- )'YP\7 pR, where (Rc-ItP R'YP = t5~. Hence, ifRc > 0, then (2.113) When Rc
is equivalent to
aR at
+ Rt
_ (Rc- I )'Y5 \7 R\7-R > O. 'Y
r
Since R'Y5 ~ Rg'Y5' we have - (Rc- I 5 ~ {}
(2.114)
at log (tR) - 1\7 log (tR)1
2
6-
_~g'Y5, {}
and hence
1
= at log R + t
- 1\7 log RI 2 ~ o.
Without assuming Rc > 0, we still obtain (2.114) by taking X'Y = -~\7'YR in (2.113) and using RaP ~ R9ap . COROLLARY 2.89
(Integrated form of Kahler trace Harnack estimate).
If (Mn,g(t)) is a complete solution to the Kahler-Ricci flow with bounded nonnegative bisectional curvature, then for any Xl, X2 E M and 0 < II < t2, we have R (X2' t2) > -tl e _l~ ----'----'4 R(XI,tl)-t2 ' where ~ = ~ (Xl, tl; X2, t2) ~ inf'Y fttl2 Ii' (t)I;(t) dt, and the infimum is taken over all paths 'Y : [tl, t2J ~ M with 'Y (tl) = Xl and 'Y (t2) = x2· By the fundamental theorem of calculus and (2.114), we have for any'Y : [tl, t2J ~ M with 'Y (tI) = Xl and 'Y (t2) = X2, PROOF.
log
t2R (X2' t2) lt2 d R( ) = -d [log tR b (t) , t) J dt tl
XI. tl
t
tl
= 1lt2 [(:tlogtR) ('Y(t),t) + (\7logtR,i'(t))g(t)] dt
~ lt2
[1\7l0gtRI;(t)
tl
+ (\7l0gtR,i'(t))g(t)]
dt
~ -~ lt21i' (t)I;(t) dt.
tl The corollary follows from taking the infimum over all 'Y.
o
For the normalized Kahler-Ricci flow, we have the following. COROLLARY 2.90 (Integrated trace Harnack for normalized Kahler-Ricci flow). If (Mn,g(t)) is a solution to the normalized Kahler-Ricci flow on a closed manifold with nonnegative bisectional curvature, then for any Xl, X2 E M and 0 < tl < t2,
(2.115)
where
~
is as above and r
~
0 is the average (complex) scalar curvature.
> eii-tl-l R EMARK 2 .91 . Note t h at l-e-ii-tl l-e-nrt 2 - ~. en 2-1
111
9. MATRIX DIFFERENTIAL HARNACK ESTIMATE
Before we prove the corollary, we first recall how to go from the KahlerRicci flow to the normalized Kahler-Ricci flow on closed manifolds. Let 9 (t) be a solution to ~gQJ3 = -ROti3 and let 9 (t) ~ 'IjJ (t) 9 (t) , where 'IjJ (t) is to be defined below. We compute
_.
{)
{)t9Oti3 = -'ljJROti3 since ROti3 we have
+ 'ljJgOti3
= ROti3' Hence if we define a new time parameter i by -Nt = i %t'
{) -
{)igOti3
= -
~-
R
Oti3 + 'ljJ2 gOti3'
In particular, to obtain the normalized Kahler-Ricci flow, where mains in the same Kahler class, we set ~ (t)
f
'IjJ (t)2
n
9 (t) re-
,
where f is the average scalar curvature of 9 (t), which is independent of time since 9 (t) stays in the same Kahler class. Thus we take 'IjJ (t) ~ (1 - ~t) -1 . Since di = 'ljJdt, we may take i ~ -~ log (1- ~t) . That is, t
(1 - e-~l) .
=~
2.90. Let 9 (i) be a solution of the normalized Kahler-Ricci flow %t 9Ot i3 = -ROti3 + ~9Oti3' Then PROOF OF COROLLARY
is a solution of the Kahler-Ricci flow and we have the estimate ~~;;::;~ >
~e-~A. This implies
where
Li (Xl, i I ; X2, i2) = inflt21 ddT' "f
since
tl
t
(t)1
2
g(t)
dt
= inf "f
ft = 'IjJ-9t, 9 (t) = 'IjJ-I 9 (i) , and dt = 'IjJ-Idi.
(21
itl
d~ (i) 12 dt
ii(i)
di
o
Since 9 (t) has nonnegative bisectional curvature, and in particular it has nonnegative Ricci curvature, under the normalized Kahler-Ricci flow we have itgOti3 ~ ~gOti3' which implies 9 (t) ~ e~(t-tl)g (tI) for t ~ ti' Hence (2.115) implies
112
2.
KAHLER-RICCI FLOW
by taking 'Y (t) to be a minimal geodesic, with respect to 9 (tt), joining to X2 with speed Ii' (t)l g(t1) = ce-f;(t-tt}, where
Xl
r(
t»)-l dg(h) (XI,X2). c=r;, 1-e- nret 2-1
which is the estimate we would have obtained from (2.115) by using e;;;(t-h) ~ e;;; (t2 -t1) and taking 'Y (t) to be a minimal geodesic joining Xl to X2 with speed I'Y· (t)1 g(h) -- d9 (t1)(X1,X2) t2-t1 .
9.2. Application of the trace estimate. A beautiful application of the trace differential Harnack estimate and Perelman's no local collapsing theorem is the following uniform bound for the curvatures of a solution of the normalized Kahler-Ricci flow on a closed manifold with nonnegative bisectional curvature. This proof is due to Cao, Chen, and Zhu [49] and gives a simple proof of an estimate of Chen and Tian [87], [88], who proved convergence of the Ricci flow for the normalized Kahler-Ricci flow on closed manifolds with positive bisectional curvature (Theorem 2.70). THEOREM 2.92 (NKRF: Kc (V, W) ;:::: 0 curvature estimate). If(Mn,go) is a closed Kiihler manifold with !f:: [wo] = 27rCl (M) and nonnegative bisectional curvature, then the solution g (t) to the normalized K iihler-Ricci flow, with g (0) = go, has uniformly bounded curvature for all time. PROOF. Without loss of generality we may assume the solution is nonflat and the average scalar curvature r is equal to n, independent of time. Given any time t > 1, there exists a point y E M such that R (y, t + 1) = n. By (2.116) we have for any X E M
n R(x,t)
=
R(y,t+1) 1-e- t (d;(t)(X,y)) R(x,t) ;:::: 1- e-(t+1) exp -4(1- e- l )
Since 11~e(t:1) ~ ~=:=~ = 1 + e- l for t ;:::: 1, we conclude for any (2.117)
.
X
E M,
9. MATRIX DIFFERENTIAL HARNACK ESTIMATE
113
In particular, if x E B t (y, 1) , then
R (x, t) :S n (1
+ e- 1) exp
(4 (1 ~
e- 1))
.
Hence, since g (t) has nonnegative bisectional curvature, the curvature of g (t) is bounded9 in Bg(t) (y, 1). By Perelman's no local collapsing theorem, which holds for a solution to the normalized Kahler-Ricci flow since its statement is scale-invariant and since the corresponding solution to the Kahler-Ricci flow must blow up in finite time, we conclude that there exists a constant K, > 0 depending only on the initial metric such that Volg(t) (Bg(t) (y, 1)) 2:
K,.
In particular, K, > 0 is independent of t > 1 and the choice of y E M such that R (y, t + 1) = n. We may obtain a uniform diameter bound for g (t) by Yau's argument. In particular, let x E M be a point with d (x, y) = d 2: 2. Since Rc 2: 0, by the Bishop-Gromov relative volume comparison theorem, we have (2.118) Volg(t) (Bg(t) (x, d + 1)) - Volg(t) (Bg(t) (x, d - 1)) (d + It - (d - It Volg(t) (Bg(t)(x,d-l)) :S (d-lt
C(n)
< - -d - . Since Bg(t) (y, 1) c Bg(t) (x, d + 1) \Bg(t) (x, d - 1) and Bg(t) (x, d - 1) C Bg(t) (y, 2d - 1) , by (2.118), we have Volg(t) (M) 2: Volg(t) (Bg(t) (y, 2d - 1))
2: Volg(t) (Bg(t) (x,d -1)) 2:
Volg(t) (Bg(t) (y, 1)) C (n) d.
Taking d = diamg(t) (M) , we have
. C (n) Volg(t) (M) C (n) Volg(t) (M) dlamg(t) (M) < --~--=::...:~--:- < . - Volg(t) (Bg(t) (y, 1)) K, Since under the normalized flow, Volg(t) (M) is constant, we obtain a uniform upper bound C for the diameter of g (t). Hence (2.117) implies
R (x, t) :S n (1 + e-1 ) exp ( 4 (1 ~2e- 1 )) which is our desired uniform estimate for R.
,
o
9We actually only need an upper bound on the scalar curvature for Perelman's no local collapsing theorem (Theorem 6.74).
114
2. KAHLER-RICCI FLOW
9.3. Proof of the matrix Harnack estimate. In this subsection we prove Theorem 2.87. Let
The following computation is the one corresponding to Proposition 2.79. PROPOSITION 2.93 (Evolution of the Harnack quantity Zo.(3)' Suppose that a vector field X satisfies
--
1
'\lJX'Y = '\l oXy = R'YJ + i 9'Y J, '\l oX'Y = '\lJXy = 0, and
Then Zo.13 = Z (X)o.13 defined by (2.111) satisfies the evolution equation:
(:t - Do) Zo.13 = Ro.13'YJZ,yo - ~ (Ro.1Z'Y13 + R'Y13 Zo.1) (2.119)
+ po.J'YPo131 -
Po.1JP13'Yo -
2
i Z o.13 '
The proposition follows from Lemmas 2.94 and 2.95 below. Assuming the proposition, we now prove the differential Harnack estimate, i.e., Proposition 2.93. Applying Zo.13 to a (1, O)-vector field W, we have the following general formula:
(:t - Do) (Zo.13W o. w13 ) = ( (:t - Do) Zo.(3) Wo.W13 + Zo.13 ((:t - Do) (wo.W13)) + '\lTZo.13'\lf (Wo.W13) + '\l fZo.13'\lT (Wo.W13)
,
where W13 ~ W.B. Thus, if we have a null vector W of Zo.13 at a point (xo, to) and if we extend W locally in space and time so that at (xo, to) (2.120) (2.121)
W 0:,1--'?i = Wa ,fJ = 0, (J
(~ at - Do) Wo. = 0'
9. MATRIX DIFFERENTIAL HARNACK ESTIMATE
115
where Wa = 9,6aW,6, then (2.119) implies
(:t - Ll) (ZQ~WaW,6) = (RQ~1'8Zc5'Y) Wa W,6
+ (PQ8I'P£~'Y - PQ8'YPc5~I') Wa W,6
1
2
- 2 (RQ'YZI'~ + ZQ'YRI'~) Wa W,6 - tZQ~WaW,6 = RQ~1'8Zc5'YWaW,6 + IMI'812 -IMI'c512. Here and Now we use the facts that ZQ~ ~ 0 (at least on all of M x [0, toD and W is a null vector of ZQ~. An algebraic fact, similar to Lemma 2.86 and using a second variation computation similar to that in the proof of Theorem 2.78, shows that
RQ~1'8Zc5'YWaW,6 ~ IMI'812 -IMI'c512. Thus at a point where W satisfies (2.120)-(2.121), we have
Hence, by the maximum principle, on a closed manifold we have ZQ~ ~ 0 on all of space and time. In the complete noncom pact case, one can adapt the proof in Part II of this volume of Hamilton's matrix Harnack estimate for complete solutions to the Ricci flow with nonnegative curvature operator to this Kahler setting without significant modifications. Now we give the two lemmas which are needed to complete the proof of Proposition 2.93. LEMMA
2.94.
(:t - Ll) (LlRQ~ + RQ~1'8R;yc5) 1
1
= 2LlRQpRp~ + 2RQpLlRp~ + 2RQP'I'Rp~,'Y 1 + 2Rc5'Y (\lJ\lI'RQ~
+ \l1'\l8RQ~)
1
-
Ll (RQ'YRI'~)
+ 2(\lJ\lI'RQ~+ \ll' \l8RQ~-RQPI'8Rp~-RQpRp~1'8)Rc5'Y (2.122)
+ 2RQ~1'8Rc5pRP'Y
a
+ RQ~1'8 at Rc5'Y.
2. KAHLER-RICCI FLOW
116
PROOF.
Let hOtij be a real (1, I)-tensor. Using (2.33), we compute
%t ('V s'V-yhOtij) =
(2.123)
'VS'V -y (%t hOtij) - 'V s ( (%t
r~Ot ) hpij) -
(%t
r~/3 ) 'V -yhOtp
= 'VS'V-y (%thOtij)
+ 'VS (gPu'V-yROtuhpij) + gUP'VSRiju'V-yhOtP
= 'VS'V-y (%thOtij)
+ 'VS'V-yROtphpij
+ 'V-yROtp'VShpij + 'VSRpij'V-yhOtp. Taking the complex conjugate of (2.123), we have
:t
(2.124)
('V-Y'VShOtij) = 'V-y'VS (%thOtij)
+ 'V-Y'VSRpijhOtp
+ 'VSRpij'V-yhOtp + 'V-yROtp'VShpij' Next we compute, using (2.123), (2.124) and tracing, that
a at
a [g-Ys ('Vs'V-yhOtij+'V-y'VshOtij) ] (~hOtij) = '12 at 1
=
-
a
'2 g-Yo at
('Vs'V-yhOtij
+ 'V-y'VshOtij)
1
+ '2 Ro;y ('V S'V -yhOtij + 'V -y 'V ShOtij) =
~ (:t hOtij ) + ~Ro;y ('Vs'V-yhOtij + 'V-y'VshOtij) 1
+ '2 ('V;y 'V -yROtphpij + 'V -yROtp'V ;yhpij + 'V ;yRpij'V -yhOtp) 1
+ '2 ('V-y'V;yRpijhOtp + 'V;yRpij'V-yhOtp + 'V-yROtp'V;yhpij). In particular, simplifying and taking hOtij to be the Ricci tensor, we have
(! - ~) (~ROtij) = ~ (ROtij-ySRo;y) - ~ (ROt;yR-yij) 1
+ '2 Ro;y ('V S'V -y ROtij + 'V -y 'V SROtij ) 1
+ '2 ('V;y 'V -yROtpRpij + 'V -y 'V ;yRpijROtp) + 'V -yROtp 'V ;yRpij + 'V ;yRpij 'V -yROtp, On the other hand, using (2.103), we compute
a
at (ROtij-ysR;yo)
= ('V -y 'V SROtij - ROtpRpij-yS) R;yo
+ 2ROtij-ySRopRp;y.
a
+ ROtij-yS at R;yo
9. MATRIX DIFFERENTIAL HARNACK ESTIMATE
117
Hence
(:t -
Ll ) ( LlRai3 + Rai3-y;SR-yo)
1
= 2 (\71 \7 -y R apRpi3 + \7 -y \71 Rpi3 Rap ) + 2\7 -yRap \7 ;yRpi3 1
+ 2RO;Y (\7;S\7-y R ai3 + \7-y\7;SRai3) +
(\7 -y \7 ;SRai3 - Rap Rpi3-Y;S ) R-yo
Ll (Ra;yR-Yi3)
a
+ Rai3-Y;S at R-yo
+ 2Rai3-y;SRopRP"Y. Finally we obtain (2.122) from the commutator equations: 1
2 (\7-y\7;SRai3 -
\7-y\7;SRai3) =
1
-2 (R-y;SapRpi3 -
R-y;Spi3Rap)
and \7;y \7 -yRap - \7 -y \7 ;yRap
= - (R-y-yauRup + R-y-ypuRau) = RauRup - RupRau = 0,
which imply
o LEMMA
2.95.
(:t -
Ll ) (\7 -y R ai3 X -Y)
1
2 (Rap \7-yRpi3 + \7-y RapRpi3 + \7 pRai3R-yp) X-Y + (\7 pRauR-Yi3up - Rau-yp \7 pRui3) X-Y + \7-y (Rai3puRup) X-Y
= -
(2.125)
+ \7-yRai3
PROOF.
(:t - Ll) X-Y - \7;S\7-yRai3\7oX-Y - \7o\7-yRai3\7;SX-Y.
We compute using (2.105)
(:t - Ll)
\7 -y R ai3 = \7 -y
(:t - Ll) Rai3 +
(\7 -yLl - Ll \7 -y) Rai3
- (!r~a) Rpi3 = \7-y (Rai3p;SRop - RapRpi3) + gpif\7-yRau Rpi3 (2.126)
1
+ 2 (-Rui3\7 -yRau + Rau \7-y R ui3) - R-ypau \7 pR ui3 + R-ypui3 \7 pRau -
1
2R-yu \7 uRai3'
2. KAHLER-RICCI FLOW
118
Here we also used (2.33) and the following general identity: 1 (V'-yDo. - Do. V' -y) haj3 = "2 V' p (V' -y V' p - V' pV' -y) haj3 1
+ "2 (V'-y V' p 1
V' pV' -y) V' phaj3
R-ypau h qj3
+ R-Y{iuj3 h au)
+ "2 (- R-yu V' qh a j3 -
R-ypau V' phqj3
= "2 V' p ( 1
(2.127)
+ R-ypqj3 V' phau )
1
= "2 (-V'-y R au hqj3 + V' -yRqj3hau) - R-ypau V' phqj3
1
+ R-ypqj3 V' phau - "2 R-yu V' qhaj3'
where we used the second Bianchi identity (2.9). Simplifying (2.126), we have
(%t - Do.) V'-y R aj3 = - ~ (Rqj3 V'-yRau + Rau V'-yRqj3 + R-yu V'qRaj3) + R-ypqj3 V' pRau -
R-ypau V' pRqj3
+ V' -y (Raj3pSR8p)
.
Equation (2.125) now follows from this and the general formula
(%t - Do.) (V'-y R aj3X-Y) = [ (%t - Do.) (V'-y R aj3) ] X-y + V'-y R aj3 (%t - Do.) X-y - V'SV'-y R aj3V'8 X -Y - V'8V'-y R aj3V'SX-Y.
o EXERCISE
2.96. Prove Proposition 2.93 using Lemmas 2.94 and 2.95.
10. Linear and interpolated differential Harnack estimates In this section we consider a differential Harnack estimate related to the estimate of H.-D. Cao considered in the previous section. This estimate has applications in the study of the geometry and function theory of noncom pact Kahler manifolds with nonnegative bisectional curvature. Let (M n , 9 (t)), t E [0, T), be a complete noncompact solution of the Kahler-Ricci flow with bounded nonnegative bisectional curvature. By Shi's theorem, given an initial metric which is complete with bounded nonnegative bisectional curvature, such a solution exists, at least for some short time T > 0, with
IV' Rm (x, t) I :s
c
t 1/ 2
for some C < 00. Analogous to the Riemannian case (see Remark 2.25 in this chapter or Theorem 10.46 on p. 415 of [111]), we consider a solution to the linearized Kahler-Ricci flow. That is, we let haj3 be a Hermitian symmetric (1, I)-tensor satisfying the Kahler-Lichnerowicz Laplacian heat equation:
10.
LINEAR AND INTERPOLATED DIFFERENTIAL HARNACK ESTIMATES
119
(2.128)
A bound for reasonable solutions is given by the following result (see Lemma 1.2 and Proposition 1.1 in Ni and Tam [290]). PROPOSITION 2.97 (Exponential bound for h Ot i3)' Suppose a solution hOti3 of (2.128) satisfies, for some constants A and B, the following inequalities:
IhOti3 (x, 0) I ::; eA (1+ r o(x))
(2.129) and (2.130)
Then there exists a constant C <
00
such that
IhOti3 (x, t) I ::; eC (1+r o(x)). Furthermore, if (hOti3 (x, 0)) 20, then (hOti3 (x, t)) 2 0 for all t > O. The analogue of the linear trace differential Harnack estimate for the Riemannian Ricci flow, Theorem A.57, is as follows (see Theorem 1.2 on p. 633 of Ni and Tam [290]). The Kahler linear trace differential Harnack quadratic is defined by (compare with (A.27))
Z (h, V)
~ ~gOti3 (V'13 div (h)Ot + V'
01
div (h)i3)
+ ROti3h/3a.
+ gOti3 ( div (h)Ot Vi3 + div (h)i3 VOl) + hOti3 V/3 Va. + ~, where V is a vector field of type (1,0), H ~ gOti3 hOti3' and (2.131)
div (h)Ot ~ g'Yi3V''YhOti3'
THEOREM 2.98 (Kahler linear trace differential Harnack estimate). Suppose that (Mn,g(t)), t E [O,T), is a complete solution of the Kahler-Ricci flow with bounded nonnegative bisectional curvature and (hOti3) 20 is a solution of the Kahler-Lichnerowicz Laplacian heat equation (2.128) satisfying (2.129) and (2.130). Then Z (h, V) 20 on M x [0, T) for any vector field V of type (1,0) . The proof of this theorem requires a number of calculations which we state and prove. In these calculations the theme is to derive a heat-type equation for each of the quantities under consideration. We then need to combine terms in a good way so that we obtain a supersolution to the heat equation. The way this is accomplished, as in the Riemannian case, is to look for terms which vanish on gradient Kahler-Ricci solitons.
2. KAHLER-RICCI FLOW
120 LEMMA
2.99. We have the following equations and their complex conju-
gates: (1) (2.132)
(:t - ~)
div (h)a
(:t - ~)
(ga J3 \7J3 div (h)a)
= R/LiJ \7//ha{.L + h{.L// \7 aR/LiJ -
~RaiJ div (h)//,
(2) (2.133)
= R/La \7 {.L div (h)a + \7 aR/LiJ \7//ha{.L
+ \7 aR/LiJ \7 ah{.L// + R/LJ\7 a \7//ha{.L + h{.L// \7 a \7 aR/LiJ. PROOF.
Using (2.128) and (2.131), we compute
(! - ~) \7-y haJ3 = \7 (:t - ~)haJ3 + (\7-y~ -y
= \7 -y ( RaJ3e8hU 1
-
~ \7-y) haJ3 -
(:t r~a )
hliJ3
~ (RaeheJ3 + ReJ3hae) ) 1
- "2 \7 -y R a8 hliJ3 + "2 \7 -yRliJ3 h a8 - R-Yfia8\7.,.,hliJ3
+ R-yfiliJ3\7.,.,ha8 -
1 "2R-Y8\7lihaJ3
+ \7 -yRa8hliJ3'
since
2 (\7 -y~
-
~ \7 -y) haJ3
= \7-y (\7.,., \7 fi + \7 fi \7.,.,) haJ3 - (\7.,.,\7 fi + \7 fi \7.,.,) \7 -yhaJ3 = \7.,., (\7 -y \7 fi - \7 fi \7 -y) haJ3 + (\7 -y \7 fi - \7 fi \7 -y) \7.,.,haJ3
+ R-YfiliJ3 ha8 ) R-Yfia8\7.,.,hliJ3 + R-YfiliJ3\7.,.,ha8
= \7.,., ( - R-yfia8 hliJ3
- R-Yfi.,.,8\7li h aJ3 = -\7-yRa8hliJ3 + \7 -yRliJ3ha8 - 2R-Yfia8\7.,.,h liJ3 + 2R-YfiliJ3\7.,.,ha8 - R-y8\7lihaJ3· Simplifying, we have
(:t - ~)
\7-y h aJ3
= \7-y R aJ3e8 hlie - R-Yfia8\7.,.,hliJ3 + R-YfiliJ3\7.,.,ha8 1
+ RaJ3e8\7 -yhlie - "2 (Rae\7 -y heJ3 + ReJ3\7-yhae + R-Y8\7lihaJ3) . Since div (h) a
= g-YJ3\7-yhaJ3' taking the trace and cancelling terms, we have
(:t - ~)
div(h)a = g-yJ3
(:t - ~)
\7-y h aJ3 + R/3'Y\7-y h aJ3
= \7 a Re8 h lie + Rlifi \7.,.,ha8 -
which is (2.132).
1
"2 Rae\7 {3heJ3'
10. LINEAR AND INTERPOLATED DIFFERENTIAL HARNACK ESTIMATES
121
Next we verify (2.133). Using (2.132) and (2.21), we compute
(:t - ~) = (:t - ~)
(V 13 div (h)a)
V 13
div (h)a
= V 13 ( Rf..liJ V vhajl
+ (V t3~ -
+ hjlv VaRf..liJ -
~ V13) div (h)a
~RaiJ div (h)v)
+ ~V t3RJa div (h)8 - ~Rt38VJ div (h) a + Rt3,Ja Vi div (h)8 = V t3Rf..liJ V vhajl + V t3hjlV VaRf..liJ + Rf..liJ V 13 V vha'i + hjlV V 13 V aRf..liJ -
~ RaiJ V13 div (h) v
~Rt38 VJ div (h)a + Rt3,Ja Vi div (h)8·
Tracing and cancelling terms, we have
(:t - ~) (:t - ~)
(gat3V 13 div (h)a)
= gat3
(V 13 div (h)a)
+ R,Ba V 13 div (h) a
= V aRf..liJ V vhajl + V ahjlV V aRf..liJ + R,Ba V 13 div (h) a + Rf..liJ V a V vhail
+ hilv Va VaRf..liJ, D
which is (2.133).
EXERCISE 2.100. Write down the formulas for the complex conjugate equations to (2.132) and (2.133).
Now let for c
>0
and
hat3 ~ hat3 + cgat3· Since Z is of the form Z = A+BaVa +BaVa + hat3 Va V,B and hat3 2: cgat3 > 0, at each (x, t) we have that Z attains its minimum for some V. By taking the first variation, we immediately see that (2.134) Differentiating this, we have (2.135) (2.136)
+ V ,Bhai V, + hai V,B V, = 0, V 13 div (h)a + V t3hai V, + hai V 13 V, = o.
V,B div (h)a
122
2. KAHLER-RICCI FLOW
In each of the above instances, we also have the complex conjugate equations; we leave it to the reader as an exercise to write these down. Recall that Cao's Kahler matrix differential Harnack quadratic is (2.111): Za~
=
fl.Ra~
Ra~
+ Ra~-y8Ry6 + V-yRa~V-y + V-yRa~V-y + Ra~-y8V-yV6 + -t-·
From (2.132) and (2.133) and their conjugate equations, while substituting in (2.134), (2.135), (2.136) and their conjugate equations, we obtain
(~ at -fl.) Z = Z aJJf-Ih-a aJJ + h -yu:«v (2.137)
1( - t -h-y8 (V6 V-y
a
V--y - R a-y- - ~g t a-y-)
(v-v" a
u
R-r au - ~g-r\ t au)
+ V-yV6) + 2R6-yh-Y8 + 2 (H t+ en) + 2eR)
+ h-Y8 V Q V-y Va V6 .
From (2.134) and the trace of (2.136), we have ~ 11Z = Ra~hQ/3 - 2ha~VQV/3 - 2h/3QVaV~
Substituting this into the
RHS
+
H
+ en t
+eR.
of (2.137) yields
Z aJJf-Ih-a+h --~g Oi.JJ -yu:«v a V--R -y a-y t a-y-)
(V-V"-R-r-~g-r) a u au t au
>0 - .
Hence we have for the minimizer V satisfying (2.134)
(:t - fl.) (t z) 2: 2
O.
By applying the maximum principle (see pp. 639-640 of [290] for details), we may conclude that t 2 Z 2: 0 for all t > O. We have the following matrix differential Harnack estimate due to one of the authors [287]. THEOREM 2.101 (Matrix interpolated differential Harnack estimate). Let (Mn, 9 (t)) be a complete solution of the e-speed Kahler-Ricci flow
(2.138) where e > 0, with bounded nonnegative bisectional curvature, and let u be a positive solution of the forward conjugate heat equation au (2.139) at = fl.u + eHu. Then for any (1, O)-form V we have
(2.140) Equivalently, f (2.141)
~
log u satisfies fa~
1
+ eRa~ + tga~ 2: O.
10. LINEAR AND INTERPOLATED DIFFERENTIAL HARNACK ESTIMATES
123
PROOF. First we observe that the equivalence of (2.140) and (2.141) follows from the fact that the minimizing {I, O)-form Va for the LHS of (2.140) is equal to -~ and dividing (2.140) by u. We compute
af 2 at = b.f + IY' -til + ER. Using (2.22) and commuting a pair of derivatives, we have
+ IY'_rf12 + ER) = b.LfOt.i3 + EY' a Y'i3 R + f a,rfi3;y + f a;yfi3"Y + Y';yfY'"Yfai3 + Y'"YfY' ;yfOt.i3 + Ra;yoi3Y'JfY'"Yf.
:t fai3 = Y' a Y'i3 (b.f (2.142)
Using the analogue of (2.105) for the E-speed Kahler-Ricci flow,
a
at ROt.i3 = Eb.LR ai3 = EY' a Y' i3 R , we then compute
:t
(fOt.i3
+ EROt.i3 +
t
+ EROt.i3) + fa"Yfi3;Y + fOt.;yfi3"Y + Y';yfY'''Yfai3 + Y'''YfY';yfai3 + Ra;yoi3Y'JfY'''Yf + E2 b.ROt.i3 + E2 ROt.i3"YJRo;y - E2 Ra;yR"Yi3
9ai3) = b.L (JOt.i3
1 E - t29Ot.i3 - Rai3'
t
Hence letting SOt.{3 ~ fai3
+ ERai3 + t9ai3
(! - b.L) SOt.i3
+ E2
=
fa"Yfi3;Y
and using (2.106), we have
( b.Rai3 + Rai3"YJ R O;Y +
:t
Rai3 )
- EY';yfY'''YROt.i3 - EY'''YfY';yROt.i3 + Ra;yoi3Y'JfY'''Yf + Y';yfY'"Y S ai3 + Y'"YfY' ;ySOt.i3
+ ~Sa;y (fi3"Y +~
ER"Yi3 - t9"Yi3 )
(fa;y - ERa;y - t9a;y) Si3"Y'
Now for the E-speed Kahler-Ricci flow, by Cao's Kahler matrix differential Harnack estimate (2.111) with Xa = -~Y'af, we have
1
o ::; b.Rai3 + Rai3"YJ R o;y + Et ROt.i3 1 1 - -Y';yfY'''YROt.i3 - -Y'''YfY';yRai3 E
E
1 + 7. Ra;yoi3 Y' J fY' "Y f. E
124
2.
KAHLER-RICCI FLOW
Hence
(:t -
b.. L ) 8 a(3 2 fa,d(3'Y +
+ \!'Yf\!"(8a(3 + \!"(f\!'Y8a(3
~8a'Y (f(3"( -
+~
cR"((3 - t g"((3)
(fa'Y - cRa'Y - tga'Y) 8(3"(.
The estimate (2.141) follows from an application of the maximum principle; 0 see [287] for details. Tracing (2.140), i.e., multiplying by ga(3 and summing, we have COROLLARY 2.102 (Trace interpolated differential Harnack estimate). Under the hypotheses of Theorem 2.101, nu b..u + cuR + + gaf3 (uaV(3 + u(3Va + uVaV(3) 20,
t
which, by taking Va = - ~ and then dividing the resulting expression by u, implies the equivalent inequality n b..log u + cR + -t > - O.
J
Let N ~ M u log udf..t be the (classical) entropy of u. We have under (2.139), $tdf..t = -cRdf..t (since $t detg"(8 = -cRdetg"(8)' and
~ = 1M (b..u + cRu + (logu) b..u) df..t
1M (b..logu+cR)udf..t 2 -!f 1M udf..t.
=
In other words, (2.143) 11. Notes and commentary Some books containing material on or devoted to complex manifolds and Kahler geometry, in essentially chronological order, are Weil [370], Chern [95], Goldberg [157], Kobayashi and Nomizu [236], Morrow and Kodaira [275], Griffiths and Harris [166]' Aubin [13]' Kodaira [238]' Besse [27], Siu [334]' Mok [270]' Tian [347]' Wells [371], and Zheng [383]. We refer the reader to these books for the proper study of Kahler geometry. For the Ricci flow on real2-dimensional orbifolds, see L.-F. Wu [372] and [112]. For the Ricci flow on noncompact Riemannian surfaces, see Wu [373]'
11. NOTES AND COMMENTARY
125
Daskalopoulos and del Pino [119]' Hsu [206], [207], and Daskalopoulos and Hamilton [120].
CHAPTER 3
The Compactness Theorem for Ricci Flow Although this may seem a paradox, all exact science is dominated by the idea of approximation. - Bertrand Russell
The compactness of solutions to geometric and analytic equations, when it is true, is fundamental in the study of geometric analysis. In this chapter we state and prove Hamilton's compactness theorem for solutions of the Ricci flow assuming Cheeger and Gromov's compactness theorem for Riemannian manifolds with bounded geometry (proved in Chapter 4). In Section 3 of this chapter we also give various versions of the compactness theorem for solutions of the Ricci flow. Throughout this chapter, quantities depending on the metric gk (or gk (t)) will have a subscript k; for instance, 'V'k and Rmk denote the Riemannian connection and Riemannian curvature tensor of gk. Quantities without a subscript depend on the background metric g. Often we suppress the t dependence in our notation where it is understood that the metrics depend on time while being defined on a space-time set. Given a sequence of quantities indexed by {k}, when we talk about a subsequence, most of the time we shall still use the indices {k} although we should use the indices
{jk} .
1. Introduction and statements of the compactness theorems
Given a sequence of solutions (Mr, gk (t)) to the Ricci flow, Hamilton's Cheeger-Gromov-type compactness theorem states that in the presence of injectivity radii and curvature bounds we can take a Coo limit of a subsequence. The role of the compactness theorem in Ricci flow is primarily to understand singularity formation. This is most effective when the compactness theorem is combined with monotonicity formulas and other geometric and analytic techniques, in part because these formulas and techniques enable us to gain more information about the limit and sometimes enable us to classify singularity models. This has been particularly successful in low dimensions. In latter parts of this volume we shall see some examples of 127
3. THE COMPACTNESS THEOREM FOR RICCI FLOW
128
this:
ICompactness Theorem I INo local collapsing I 1
+---
I]\10notonicity I
./
ISingularity analysis I In general, there are three scenarios in which we shall apply the compactness theorem for the Ricci flow. The compactness result may be applied to study solutions (M n , 9 (t)) to the Ricci flow defined on time intervals (0', w) , where w ::; 00 is maximal, i.e., the singularity time. To understand the limiting behavior of the solution 9 (t) as t approaches w, we shall take a sequence of times tk ~ wand consider dilations of the solution 9 (t) about the times tk and a sequence of points Ok E M by defining (3.1)
gk (t) = Kkg (tk
+ KJ:1 t ) ,
where Kk = IRm (Ok, tk)1 is the norm of Rm (g (tk)) at the point Ok. We are interested in determining when there exists a subsequence of pointed solutions to the Ricci flow (M, gk (t) , Ok) which limits to a complete solution (M~, goo (t) ,000 ) . This limit solution reflects some aspects of what the singularity looks like near (Ok, tk). Similarly, when 0' = -00 for solution (M, 9 (t)), which arises when we already have a (first) limit solution of a finite time singularity, we may consider sequences tk ~ -00 and take a second limit, now backward in time. Yet other limits that we shall consider arise from dimension reduction on a limit solution. Here tk remains fixed whereas Ok tends to spatial infinity. Many of the topics in this volume are related to the study of the geometry (and topology) of the limits of these solutions when they exist. 1.1. Definition of convergence. Now we review the definition of Coo_ convergence on compact sets in a smooth manifold Mn. By convergence on a compact set in CP we mean the following.
3.1 (CP-convergence). Let K c M be a compact set and let {gdkEN' goo, and 9 be Riemannian metrics on M. For p E {O} UN we say that gk converges in CP to goo uniformly on K if for every c > 0 there exists ko = ko (c) such that for k ~ ko, DEFINITION
sup sup IV o (gk - goo)lg
< C,
O::;o::;pxEK
where the covariant derivative V is with respect to g. Note that since we are on a compact set, the choice of metric 9 on K does not affect the convergence. For instance, we may choose 9 = goo. In regards to Coo-convergence on manifolds, with the noncompact case in mind, we have the following. We say that a sequence of open sets {UdkEN in a manifold Mn is an exhaustion of M by open sets if for any compact set K c M there exists ko EN such that Uk ::) K for all k ~ ko.
1.
INTRODUCTION; STATEMENTS OF THE COMPACTNESS THEOREMS
129
DEFINITION 3.2 (CC)()-convergence uniformly on compact sets). Suppose {UdkEN is an exhaustion of a smooth manifold Mn by open sets and gk are Riemannian metrics on Uk. We say that (Uk,gk) converges in Coo to (M, goo) uniformly on compact sets in M iffor any compact set K c M and any p > 0 there exists ko = ko (K,p) such that {gdk>k _ 0 converges in CP to goo uniformly on K. In order to look at convergence of manifolds which come from dilations about a singularity, we must ensure that the form of convergence can handle diameters going to infinity. When this happens, a basepoint, or origin, is carried along with the manifold and the Riemannian metric to distinguish what parts of the manifolds in the sequence we are keeping in focus. This allows us to compare spaces that either have diameters going to infinity or are noncompact. DEFINITION 3.3 (Pointed manifolds and solutions). A pointed Riemannian manifold is a 3-tuple (Mn, g, 0), where (M, g) is a Riemannian manifold and 0 E M is a choice of point (called the origin, or basepoint). If the metric g is complete, the 3-tuple is called a complete pointed Riemannian manifold. We say that (Mn, g (t), 0), t E (a,w) , is a pointed solution to the Ricci flow if (M,g(t)) is a solution to the Ricci flow. REMARK 3.4. In [187] Hamilton considered marked Riemannian manifolds (and marked solutions to the Ricci flow), where one is also given a frame F = {e a } at 0 orthonormal with respect to the metric g (0) with o E (a,w). Since for most applications, the choice of frame is not essential, we restrict ourselves to considering pointed Riemannian manifolds in this chapter.
:=1
Convergence of pointed Riemannian manifolds is defined in a way which takes into account the action of basepoint-preserving diffeomorphisms on the space of metrics. DEFINITION 3.5 (COO-convergence of manifolds after diffeomorphisms).
A sequence {(Mk,gk,Ok)}kEN of complete pointed Riemannian manifolds converges to a complete pointed Riemannian manifold (M~, goo, 0 00 ) if there exist
(1) an exhaustion {UkhEN of Moo by open sets with 0 00 E Uk and (2) a sequence of diffeomorphisms q,k : Uk ---t Vk ~ q,k (Uk) C Mk with q,k (0 00 ) = Ok such that (Uk,q,t: [gklvk]) converges in Coo to (Moo,goo) uniformly on compact sets in Moo. We shall also call the above convergence Cheeger-Gromov convergence in Coo. The corresponding definition for sequences of pointed solutions of the Ricci flow is given by the following.
130
3.
THE COMPACTNESS THEOREM FOR RICCI FLOW
DEFINITION 3.6 (COO -convergence of solutions after diffeomorphisms).
A sequence {(Mk' 9k (t) ,0k)}kEN' t E (a, w) , of complete pointed solutions to the Ricci flow converges to a complete pointed solution to the Ricci flow (M~, 900 (t) ,000 ) , t E (a, w) , if there exist (1) an exhaustion {UdkEN of Moo by open sets with 0 00 E Uk, and (2) a sequence of diffeomorphisms
(1) (uniformly bounded geometry) IV~Rmklk:::; Cp for all p 2:: 0 and k where C p < independent of k and
on Mk 00
is a sequence of constants
1.
INTRODUCTION; STATEMENTS OF THE COMPACTNESS THEOREMS
131
(2) (injectivity radius estimate) injgk (Ok) 2: for some constant
/'0
/'0
> O.
Then there exists a subsequence {jd kEN such that {(Mjk,9jk,Ojk)}kEN converges to a complete pointed Riemannian manifold (M~, 900,000 ) as k~ 00.
For sequences of solutions to the Ricci flow the corresponding convergence theorem takes the following form. THEOREM 3.10 (Compactness for solutions). Let {(Mk' gk (t), Ok)}kEN' (a, w) 3 0, be a sequence of complete pointed solutions to the Ricci flow such that
t
E
(1) (uniformly bounded curvatures) IRmklk ~ Co
on Mk x (a,w)
for some constant Co < 00 independent of k and (2) (injectivity radius estimate at t = 0) injgdO) (Ok) 2: for some constant
/'0
/'0
> O.
Then there exists a subsequence {jkhEN such that {(Mjk' 9jk (t) ,0jk)}kEN converges to a complete pointed solution to the Ricci flow (M~, goo (t) ,000 ), tE (a,w), as k~oo. . Note that the second theorem only supposes bounds on the curvature, not bounds on the derivatives of the curvature. This is because, for the Ricci flow, if the curvature is bounded on (a, w) , then all derivatives of the curvature are bounded at times t > a (see Chapter 7 of Volume One or Theorems A.29 and A.30 of this volume).l In particular all derivatives of the curvature are bounded at time t = 0 and we can apply Theorem 3.9 to {(M k, gk (0), Ok) hEN· In the next section we follow the proofs of Hamilton in [187]. We shall assume Theorem 3.9, which will be proven in Chapter 4. We will show that if there is a subsequence such that (M k,9k (0) ,Ok) converges to a complete limit (M~, 900 (0) ,000 ) , then there is a subsequence (Mk' gk (t) ,Ok) which converges at all times. 1The bounds on the derivatives of Rm get worse as t
-+
Q.
3. THE COMPACTNESS THEOREM FOR RICCI FLOW
132
2. Convergence at all times from convergence at one time In this section we give the proof that the compactness theorem for Ricci flow (Theorem 3.10) follows from the compactness theorem at time t = (Theorem 3.9). This is done by showing that bounds on the metric and covariant/time-derivatives of the metric at time t = extend to bounds on the metric and covariant derivatives of the metric at subsequent times in the presence of bounds on the curvature and covariant derivatives of curvature (for all time). This is shown in subsection 2.1 below. The Arzela-Ascoli theorem is then used to show that these bounds on the covariant/timederivatives of the metric imply that a subsequence converges to a solution of the Ricci flow for all times (in subsection 2.2.2 below).
°
°
2.1. Uniform derivative of metric bounds for all time. In order to extend the convergence at one time to convergence at all times, the following derivative bounds need to be shown. LEMMA 3.11 (Derivative of metric bounds at one time to all times). Let M n be a Riemannian manifold with a background metric g, let K be a compact subset of M, and let gk be a collection of solutions to the Ricci flow defined on neighborhoods of K x [,8, 'l/J], where to E [,8, 'l/J]. Suppose that
(1) the metrics gk (to) are all uniformly equivalent to 9 on K, i. e., for all V E TxM, k, and x E K, C- 1 g (V, V) :::; gk (to) (V, V) :::; Cg (V, V) , where C < 00 is a constant independent of V, k, and x; and (2) the covariant derivatives of the metrics gk (to) with respect to the metric 9 are all uniformly bounded on K, so that
for all k and p :2: 1, where C p < 00 is a sequence of constants independent of k; and (3) the covariant derivatives of the curvature tensors Rmk (t) of the metrics gk (t) are uniformly bounded with respect to the metric gk (t) on K x [,8, 'l/J] :
(3.2)
IV'~ Rmklk :::; C~ for all k and p :2: 0, where C~ is a sequence of constants independent of k.
Then the metrics gk (t) are uniformly equivalent to 9 on K x [,8, 'l/J], e.g., (3.3)
B (t, to)-l g(V, V) :::; gk (t) (V, V) :::; B (t, to) g(V, V),
where
B (t , t a) = C e 2v'n-1Cblt-tol ,
2.
CONVERGENCE AT ALL TIMES FROM CONVERGENCE AT ONE TIME
133
and the time-derivatives and covariant derivatives of the metrics gk (t) with respect to the metric 9 are uniformly bounded on K x [.8, '1/1], i.e., for each (p, q) there is a constant Cp,q independent of k such that
(3.4) 1
8q gd) -'\lP t -< Cp,q 8tq
1-
for all k. REMARK 3.12. Since we often assume bounds on Rm whereas the metric evolves by Rc, we note -In=lIRmlg:::; Rc:::; In=lIRmlg· Since we often interchange 9 and gk norms, we recall the following elementary fact. LEMMA 3.13 (Norms of tensors with respect to equivalent metrics). Suppose that the metrics 9 and h are equivalent:
C-Ig:::; h:::; Cg. Then for any (p, q)-tensor T, we have
(3.5)
ITlh :::; C(p+q)/2I T lg·
PROOF. We can diagonalize 9 and h so that gij = The assumption implies C- I :::; Ai :::; C for all i. Then
bij
and
h ij
=
Aibij.
... h kpl p hidl ... hiqjqT.kl··:kPT~l···~p ITI2h = "h L...J klll tl···tq Jl···Jq
< "
L...J
Tkl··:kPT.kl··:kP. 1l···1q
1l···1q
kl···kpjil···iq
o PROOF OF LEMMA 3.11. For the first part, since
8
at gk (t) (V, V) = -2 Rc k (t) (V, V) and IRck (t) (V, V)I :::; In=lCb9k (t) (V, V), we can estimate the time-derivatives 1
8 I at loggk (t) (V, V) =
We have proved
(3.6)
1-2RC k (t)(V,V)1 ~, gk (t) (V, V) :::; 2vn - 1Co·
134
3. THE COMPACTNESS THEOREM FOR RICCI FLOW
where C ~ 2Jn=lCb. Now we compute
C It 1
-
tol
~
1:
1 1
%t log 9k (t) (V, V) 1 dt
~ 11:1 %t log9k (t) (V, V) dtl -iiog 9k (tl) (V, V) 1 9dto) (V, V) , or equivalently, e-Clt1-toI9k (to) (V, V) ~ 9k (tl) (V, V) ~ eClh-tol9k (to) (V, V). Hence we have C-le-Clt1-toI9 (V, V) ~ 9k (it) (V, V) ~ CeClt1-tol9 (V, V) . This completes the proof of (3.3). For the second part we need to estimate the space- and time-derivatives of 9k (t). We begin with estimating the first-order covariant derivatives of 9k (t) . Note that
8 V' a (9k)bc = 8x a (9k)bc -
d
d
r ab (9k)dc - r ac (9k)bd ,
so if we take the right combination, we see that
(9k)ec (V' a (9k)bc
+ V'b (9k)ac -
V' c (9k)ab)
= 2 (rk)~b - r~b - (9k)ec r~c (9khd - r~b - (9k)ec ric (9k)ad (3.7)
= 2 (rk)~b -
+ (9k)ec r~ (9k)ad + (9k)ec r~c (9k)bd
2r~b'
This implies that
(3.8) From we have
(3.9) Hence the tensors V'9k (t) and rk (t) - r are equivalent. We recall that the derivative of the Christoffel symbols (see, for instance, (6.1) on p. 175 of Volume One) is
:t (rk)~b
= -
(9k)cd [(V'k)a (RCk)bd
+ (V'k)b (RCk)ad -
(V'k)d (Rck)abl·
2. CONVERGENCE AT ALL TIMES FROM CONVERGENCE AT ONE TIME
135
So as tensors, we find that I:t (fk - r)lk ::; 3lVd Rc k)lk ::; Thus
3Jn=1C~ It
1 -
tol
~
1:
1
I:t (fk (t) - r)lk dt
~ 11: ~ Ifk
3Jn=1C~.
1
:t (fk (t) - f) dtl k
(t1) - flk - Ifk (to) - flk .
Hence we have a bound Ifk (t) - flk ::; 3Jn=1C~ It - tol (3.10)
::;
3Jn=1C~ It -
tol
+ Ifk (to) -
flk
+ ~C3/2Cl
using (3.8) and (3.5). Since It - tol ::; '!f; - f3, we have in (3.3): B (t, to) ::; B ('!f;, f3) for all t E [f3, '!f;]. Thus by (3.9) and (3.10), (3.11)
where
C1 ,0 ~ B 3 / 2 ('!f;, f3) (6Jn=lC~ ('!f; - f3) + 3C3 / 2 C 1 ) This proves (3.4) for p = 1 and q = O. Next we prove inductively that for p IVP Rc kl ::; C; IVPgkl
~
+ C;'
.
1, IVPgkl::; Cp,o
and
(where C~, C~', and Cp,o are independent of k). If p and (3.10),
=
1, then using (3.8)
+ Vk RCklk fklk IRcklk + IVk RCklk)
IVRckl (t) ::; B (t, to)3/21(V - Vk) RCk
::; B (t, to)3/2 (If ::; B (t, to)3/2 (
(3Jn=1C~ I'!f; -
If the estimates hold for p < N with N p = N. First we have
~
f31
+ ~C3/2C1) Cb + C~) .
2, then we will prove them for
N
IV N
RCkl
=
LV
N- i
(V - Vk) V1- 1 RCk
N- i
(V - Vk) V1- 1 RCkl
+ Vt' RCk
i=l
N
: ; L IV i=1
+ IVt' RCkl·
3. THE COMPACTNESS THEOREM FOR RICCI FLOW
136
Note that, using (3.7), we can rewrite V' - V'k = r - rk as a sum of terms of the form V'9k. When i = 1, we can bound IV'N-l (V' - V'k)Rckl by a sum of terms of the form IV'N- j 9kllV'j RCkl, 0 ~ j ~ N - 1. When 2 ~ i ~ N, we can bound IV'N-i (V' - V'k) V'1- 1 RCkl by a sum of terms of the form IV'N-i-j+19kllV'jV'~-1 RCkl, 0 ~ j ~ N - i. We can also bound lV'jV'1- 1 RCkl = I((V' - V'k)
+ V'k)j V'1- 1 RCkl
by a sum of terms which are products of lV'i+i-l Rc kl ' 0
~ f ~
j, and
severallV't'9kl, 1 ~ f ~ j. By the assumption of Lemma 3.11, the induction assumption and the equivalence of 1·1 and 1·l k , we get IV'N Rc k 1 ~
C'/v IV'N 9k 1+ C'jJ.
Now we turn to bounding IV'N 9kl. Since 9 does not depend on t,
8
N
N
8t V' 9k = -2V' RCk and 8 1V' N9k 12 8t
8 8 V' N 9k, V' N = 2 \/ 8t 9k) ~ 18t V' N9k 12
= 41V'N RCkl2 + IV'N 9kl 2 ~
+ 1V' N9k 12
(1 + 8 (C'/v )2) IV'N 9kl 2 + 8 (C'jJ) 2 .
Integrating the above differential inequality of 1V' N 9k 12 , we get (compare with (7.47)) IV'N 9kl 2 (t)
~ e(1+8(C~y)(t-to) (IV'N 9kl 2 (to) +
8 (C'jJ)2
).
1 + 8 (C'/v) 2
This implies IV'N 9k (t)1 ~ CN,O, and the induction proof is complete, as well as (3.4) for the q = 0 case. Note that the above proof of bounding 1V' N Rc k 1 can be used to show that IV'PV'%Rckl, IV'PV'%Rkl, and IV'PV'%Rmkl are bounded independent of k. aq aq - 1 When q ~ 1, then atq V' P9dt) = V'P atq-l (-2Rcdt)). Using the evoV'P 9k (t) is lution equation of the curvature Rm k (t), we know that bounded by a sum of terms which are products of
l%t:
1V'PI V'%l Rm k 1(t) ,
1V'PV'% Rc k I,
and
I
1V'PV'%Rk 1.
D
2.
CONVERGENCE AT ALL TIMES FROM CONVERGENCE AT ONE TIME
137
2.2. Convergence at all times from convergence at one time. 2.2.1. The Arzela-Ascoli theorem. With uniform derivative bounds on the metrics in the sequence, the compactness theorem will follow from the Arzela-Ascoli theorem. LEMMA 3.14 (Arzela-Ascoli). Let X be a a-compact, locally compact Hausdorff space. If {fkhEN is an equicontinuous, pointwise bounded sequence of continuous functions fk : X --t lR, then there exists a subsequence which converges uniformly on compact sets to a continuous function foo : X --t R The reader is reminded that a-compact simply means that the space is a countable union of compact sets, and hence any complete Riemannian manifold satisfies the assumption. COROLLARY 3.15 (Metrics with bounded derivatives preconverge). Let (Mn, g) be a Riemannian manifold and let K c Mn be compact. Furthermore, let p be a nonnegative integer. If {gd kEN is a sequence of Riemannian metrics on K such that sup O~Q:~p+1
sup 1V'Q:gkl ::; C <
00
xEK
and if there exists 8 > 0 such that gk (V, V) 2:: 8g (V, V) for all VET M, then there exists a subsequence {gk} and a Riemannian metric goo on K such that gk converges in CP to goo as k --t 00.
PROOF (SKETCH). We need to show that {(gk)behEN form an equicontinuous family. We use the fact that in a coordinate patch
V' a (gk)be
=
oxoa (9k )be - r dab (gk)de - raed (gk)bd .
Thus if IV'gkl is bounded, then Ia~a (gk)bel is bounded for each a, b, c in each coordinate patch. Hence, by the mean value theorem, the (gk)be form an equicontinuous family in the patch and there is a subsequence which converges to (goo)be. Since K is compact, we may take a finite covering by coordinate patches and a subsequence which converges for each coordinate patch. We have thus constructed a limit metric. Note that the uniform upper and lower bounds on the metrics gk ensure that goo is positive definite. Similarly, we can use the bound on 1V'2gkl to get bounds on second derivatives of the metrics Iax~~xd (9k) be I in each coordinate patch and thus show the first derivatives are an equicontinuous family. Taking a further subsequence, we get convergence in C 1 . Higher derivatives are similar. 0
2.2.2. Proof of the compactness theorem for solutions assuming the compactness theorem for metrics. We will now use Corollary 3.15 together with Lemma 3.11 to find a subsequence which converges and complete the proof of Theorem 3.10. Recall that we have assumed Theorem 3.9 and hence there is a subsequence {(Mk, gk (0), Ok)} which converges to (M~, goo, 0 00 ) .
138
3.
THE COMPACTNESS THEOREM FOR RICCI FLOW
We shall show that there are metrics 900 (t), for t E (a,w), such that 900 (0) = 900 and {(Mk' 9k (t) ,Ok)} converges to (Moo, 900 (t) ,000 ) in Coo. Since {(Mk' 9k (0) ,Ok)} converges to (Moo, 900,000 ) , there are maps
gt
3. Extensions of Hamilton's compactness theorem In this section we give several variations of Theorem 3.10. 3.1. Local compactness theorems. From the proof of the compactness Theorem 3.9 given in the next chapter and the proof of Theorem 3.10 given above, without too much difficulty one sees that a local version of Theorem 3.10 holds. In particular, we have the following. THEOREM 3.16 (Compactness, local version). Let {(M k,9k (t), Ok)}kEN' t E (a, w) :3 0, be a sequence of complete pointed solutions to the Ricci flow.
If there exist p > 0, Co < 00, and IRmklk::; Co
/'0
in
> 0 independent of k such that Bgk(o)
(Ok,p) x (a,w)
and injgdO) (Ok) ~ /'0,
then there exists a subsequence such that {(Bgk(o) (Ok,P),9k(t),Ok)}kEN converges as k ---t 00 to a pointed solution (B~,900 (t) ,000 ), t E (a,w), in Coo on any compact subset of Boo x (a, w). Furthermore Boo is an open manifold which is complete on the closed ball Bgoo(o) (000 , r) for all r < p. EXERCISE 3.17. Prove Theorem 3.16. A simple consequence of Theorem 3.16 is (see [186]) the following corollary.
COROLLARY 3.18 (Compactness theorem yielding complete limits). Let {(M k,9k(t),Ok)}kEN' t E (a,w) :3 0, be a sequence of complete pointed solutions to the Ricci flow. Suppose for any r > 0 and c > 0 there exist constants Co (r, c) < 00 such that IRm klk ::; Co (r, c)
on
Bgk(o)
(Ok, r) x (a + c, w - c)
3.
EXTENSIONS OF HAMILTON'S COMPACTNESS THEOREM
139
for all kEN. We assume injgk(O) (Ok) ~ "0 for some "0 > O. Then there exists a subsequence {(Mk' gk (t) ,Ok)} which converges to a complete solution to the Ricci flow (M~, goo (t) ,000 ) , t E (a, w) . REMARK 3.19. Note that the limit solution (M~, goo (t), 0 00 ) may not have bounded curvature. Without the injectivity radius estimate, we may use the trick of locally pulling back the solutions by their exponential maps (since the pulled-back solutions satisfy an injectivity radius estimate). We have the following. COROLLARY 3.20 (Local compactness without injectivity radius estimate). Let {(M k,gk (t) , Ok)} kEN' t E (a, w) :3 0, be a sequence of complete solutions to the Ricci flow with IRmklk ~ Co
in Bgk(o) (Ok, p) x (a,w).
Then there exists a subsequence such that {(BTokMk
(a, c) ,(expR(O))* gd t )) ,a} kEN'
where
c~ min {p,7f/JCo} ,
converges to a pointed solution (B~,goo (t) ,000 ) , t E (a,w), on an open manifold which is complete on the closed ball Bgoo(o) (000 , r) for all r < c. REMARK 3.21. There is a similar result for geodesic tubes; see §25 of Hamilton [186].
. 3.2. Compactness for Kahler metrics and solutions. Without much difficulty, the compactness theorems apply to Kahler manifolds and solutions of the Kahler-Ricci flow (see also Cao [48] and Theorem 4.1 on pp. 16-17 of Ruan [314]). THEOREM 3.22 (Compactness for Kahler metrics). Let {(M~n, gk, Ok)} be a sequence of complete pointed Kahler manifolds of complex dimension n. Suppose 1\7~Rmklk ~ Cp on Mk for all p ~ 0 and k, where Cp < 00 is some sequence of constants independent of k, and injgk (Ok) ~ "0 for some constant "0 > O. Then there exists a subsequence {jkhEN such that {(Mjk' gjk' Ojk)}kEN converges to a complete pointed complex n-dimensional Kahler manifold (M~,goo,Ooo) as k ~ 00. See the ensuing proof for the meaning of the convergence of the complex structures Jk ~ J oo . PROOF. Since Kahler manifolds are Riemannian manifolds, we can apply the Cheeger-Gromov Compactness Theorem 3.9 to obtain a pointed limit (M~,goo,Ooo) which is a complete Riemannian manifold. So the only issue is to show that the limit is Kahler. Let Jk denote the complex structure of (Mk' gk) . We have for each kEN,
(1)
Jf = -
idTM k ,
140
3.
THE COMPACTNESS THEOREM FOR RICCI FLOW
(2) (gk 0 Jk) (X, Y) ~ gk (JkX, JkY) = gk (X, Y) for all X, YET Mk, (3) V' kJk = O. Since {(Mk' gk, Ok)} converges to (Moo, goo, 0 00 ) , there are diffeomorphisms ~k : Uk ~ Vk such that ~kgk ~ goo uniformly on compact sets. From (1) we know (~klL 0 Jk 0 (~k)*' as (1, I)-tensors, are uniformly bounded on any compact set K c Moo with respect to the metrics ~kgk' Since the ~kgk are equivalent to goo on K, we conclude that (~klL 0 Jk 0 (~kt, as (1, I)-tensors, are uniformly bounded on any compact set K c Moo with respect to the metric goo. Note that (3) is equivalent to V'
((~klL 0 Jk 0 (~k)*)
= 0 for all
p.
From the proof of Corollary 3.15, there exists a subsequence (~kl L 0 Jk 0 (~k)* converging in Coo, as a (1, I)-tensor on compact sets, to a smooth map J oo : T Moo ~ T Moo. Since
(I') 0 = ((~klL 0 Jk 0 (~k)*)2 + idTMoo ~ J~ + idTM"" , (2') 0 = (~kgk) 0 ((~klL 0 Jk 0 (~k)*) - ~kgk ~ goo 0 J oo - goo, (3') 0 = V'
= O. We conclude that 0
Applying Theorem 3.10, we obtain the following corresponding result for the Kahler-Ricci flow. THEOREM 3.23 (Compactness theorem for the Kahler-Ricci flow). Let { (M~n , gk (t) , Ok) } , t E (Q, w) 3 0, be a sequence of complete pointed solutions to the Kahler-Ricci flow of complex dimension n. Suppose
IRmklk ~ Co for some constant Co <
00
on Mk x (Q,w)
independent of k and that
(Ok) ~ /'0 for some constant /'0 > O. Then there exists a subsequence of solutions such that {(Mk,gk (t) ,Ok)} converges to a complete pointed complex ndimensional solution to the Kahler-Ricci flow (M~, goo (t) ,000 ) , t E (Q, w) , as k ~ 00. injgk(O)
PROOF. By Theorem 3.10, there exists a subsequence which converges to a Riemannian solution (Moo, goo (t) ,000 ) , t E (Q,w), to the Ricci flow. From the previous theorem, (Mk, gk (0) ,Jk) converges to (Moo, goo (0) ,Joo ) as Kahler manifolds for some complex structure J oo . Now by assumption, gk (t) remains Kahler with respect to Jk. Hence gk (t) 0 Jk = gk (t) and V'gk(t)Jk = 0 for all t E (Q,w), which implies goo(t) 0 J oo = goo(t) and V'goo(t)Joo = 0 for all t E (Q,w). That is, goo (t) remains Kahler with respect to J oo . 0
3.
EXTENSIONS OF HAMILTON'S COMPACTNESS THEOREM
141
3.3. Compactness for solutions on orbifolds. Note that Definitions 3.5 and 3.6 can be easily generalized to orbifolds. (See [343] for the definition of orbifold.) We have the following generalizations of Theorems 3.9 and 3.10 to orbifolds. The version for metrics is THEOREM 3.24 (Compactness theorem for metrics on orbifolds). Let {(Mk' gk, Ok)} be a sequence of complete pointed Riemannian orbifolds and let Ek be the singular set of Mk. Suppose that (i) IV~ Rmk Ik ~ C p on Mk for all p 2:: 0 and k, where C p < 00 are constants independent of k, and (ii) Vol gk B9k (Ok, ro) 2:: vo for all k, where ro > 0 and vo > 0 are two constants independent of k. Then either of the following hold. (1) limk--+oodgk(Ok, E k ) > O. In this case there exists a subsequence {(Mk j , gkj' Ok j )} which converges to a complete pointed Riemannian orbifold (M~,goo,Ooo) with IV~oo Rmgool goo ~ C p and Volgoo Bgoo(Ooo,ro) 2:: vo. Furthermore 0 00 is a smooth point in Moo. (2) limk--+oodgk(Ok,Ek) = o. In this case there exists a subsequence {(Mkj' gkj' Okj )} such that limj--+oo d( Okj' Ekj ) = o. If we choose O~j E Ekj with dgk(Okj'O~) = d(Okj,Ekj ), then a subsequence of {(Mkj,9kj,0~j)} converges to a complete pointed Riemannian orbifold (Moo, goo, 0 00 ) with IV~oo Rmg<XJ Ig<XJ ~ C p and Volgoo Bgoo(Ooo,ro) 2:: vo· Furthermore 0 00 is a singular point in Moo.
The version for solutions of Ricci flow is the following. THEOREM 3.25 (Compactness theorem for solutions on orbifolds). Let {(M k , gk(t), Ok)}, t E (a,w), be a sequence of complete pointed orbifold solutions of the Ricci flow. Let Ek be the singular set of Mk. Suppose that (i) 1 Rmklk ~ Co on Mk x (a,w) for all k, where Co < 00 is a constant independent of k, and (ii) Volgk(o) Bgk(o) (Ok, ro) 2:: vo for all k, where ro > 0 and vo > 0 are two constants independent of k. Then either of the following hold. (1) limk--+oodgk(O) (Ok, E k ) > o. In this case there exists a subsequence {(Mkj' gk j (t), OkJ} which converges to a complete pointed orbifold solution of the Ricci flow (M~,goo(t),Ooo) with 1 Rmg""lg"" ~ Co on Moo x (a,w) and Volg<XJ(o) Bgoo(o) (000 , ro) 2:: vo. Furthermore 0 00 is a smooth point in Moo. (2) limk--+oodgk(O) (Ok, E k ) = O. In this case there exists a subsequence {(Mkj,9kj(t),Okj)} such that limj--+oo d(Okj,Ekj) = O. Furthermore if we choose O~j E Ekj with dgk(o) (Okjl O~j) = dgk(o) (Okj' EkJ, then there is a subsequence of { (Mk j , gkj (t),
O~j )}
which converges to a complete pointed
orbifold solution of the Ricci flow (M~, goo(t), 0 00 ) with 1 Rm goo Igoo ~ Co
142
3. THE COMPACTNESS THEOREM FOR RICCI FLOW
on Moo x (o:,w) and Volgoo(o) Bgoo(o) (000 , ro) singular point in Moo.
~
Vo. Furthermore 0 00 is a
Idea of the proofs. Theorem 3.25 can be proved from Theorem 3.24 in the same way as we have proved Theorem 3.10 from Theorem 3.9. On the other hand, Theorem 3.24 can be proved with some modification of the proof of Theorem 3.9 to handle the singularity (see [257]). Note that the Bishop-Gromov volume comparison theorem holds for orbifolds. Fix r > 0; for any k and qk E Mk with dgk(o) (Ok, qk) ::; r, we have Volgk(o) Bgk(o) (qk, ro) ~ VI, where VI is a positive constant independent of k but depending on Vo, r, ro, n, Co. This implies that there exists rl independent of k such that Bgk(o) (qk, rd has the orbifold topological model B n IG (qk) , where B n is the unit ball in Euclidean space centered at the origin and G (qk) cO (n) is a discrete subgroup with rank IG (qk)1 bounded independent of k. The existence of rl implies that we can modify the choice of X'lr in Definition 4.26 and Proposition 4.22 so that the ball Bk ~ B (xl:, Nl< 12) has the orbifold topological model B n IG (xl:) . The key observation in the proof of Theorem 3.24 is that we can choose a subsequence of orbifolds so that the groups G (xl:) and their actions on Bn are independent of k. We can then use the balls Bk, B k, Bk to build the limit orbifold. 4. Applications of Hamilton's compactness theorem
In this section we discuss some applications of Theorems 3.10 and 3.16. We will see more applications of the compactness theorems later in this volume. 4.1. Singularity models. Theorem 3.16 may be applied to study singular, nonsingular, and ancient solutions of the Ricci flow. For example, let (Mn,g(t)) , t E [O,T), where T E (0,00]' be a complete solution to the Ricci flow. Given a sequence of points and times {(Xk' tk)}kEN' let Kk ~ IRm (Xk' tk)l. We say that the sequence {(Xk' tk)} satisfies an injectivity radius estimate if there exists LO > 0 independent of k such that injg(tk) (Xk) ~ LoK;;I/2. Given a complete solution of the Ricci flow, we can obtain a local limit of dilations provided we have an injectivity radius estimate and a local bound on the curvatures after dilations. 3.26 (Existence of singularity models). Let (M n , 9 (t)), t E (0:, w), be a complete solution to the Ricci flow. Given a sequence of points and times {(Xk' tk)}kEN' let Kk ~ IRm (Xk' tk)1 > 0 and COROLLARY
gk (t) ~ Kkg (tk
+ K;;lt) .
Suppose that the sequence {(Xk' tk)} satisfies an injectivity radius estimate, i.e., injg(tk) (Xk) ~ LoK;;I/2,forsomeLo > 0, and suppose thatO:k,wk,O:oo,w oo
~ 0 with O:k ~ 0: > 0, Wk ~ Woo, and [tk - ~,tk +~] 00
C
(o:,w) are such
4.
APPLICATIONS OF HAMILTON'S COMPACTNESS THEOREM
that there exist positive constants p ::; (3.12) sup
00
and C <
{I~~I (x, t) : (x, t) E B9(tk) (Xk' k) x [tk -
00
143
where
; : ' tk
+ ~]} ::; CKk.
Then there exists a subsequence of the dilated solutions (B9(tk) (Xk' pK;;I/2) ,gdt) ,Xk) which converges to a solution (B~, goo (t) ,x oo ) on an open manifold on the time interval (-a oo , woo], which is complete on the closed ball Bgoo(o) (xoo,r) for all r < p. In particular, if p = 00, then the solution (Boo, goo (t), xoo) is complete. REMARK 3.27. By definition, we let B9(tk) (Xk'
00)
= M.
In Chapter 6 we shall show that the injectivity radius estimate in the corollary above for the solution (Mn, g (t)) , t E [0, T), on a closed manifold with T < 00, is a consequence of Perelman's no local collapsing theorem. THEOREM 3.28 (Local injectivity radius estimate for finite time singular solutions). Let (Mn,g(t)), t E [O,T), T < 00, be a solution to the Ricci flow on a closed Riemannian manifold. There exists fJ > such that if (xo, to) EM x [0, T) is a point and time satisfying 1 IRm (x, to)1 ::; 2" in Bg(to) (xo, p) P for some p > 0, then inj (xo) ~ fJp. g(to)
°
For finite time singular solutions on closed manifolds we have the following. COROLLARY 3.29 (Local singularity models for finite time singular solutions). Suppose (Mn, 9 (t)) , t E [0, T), T < 00, is a solution on a closed Riemannian manifold. If there exist positive constants p ::; 00 and C < 00 such that (3.12) holds, where ak, 13k ~ 0, ak - 4 a > 0, and 13k - 4 {3, then
there exists a subsequence such that (B9(tk) (Xk' pK;t/2) ,gk (t) 'Xk) converges to a solution (B~,goo (t) ,X oo ) , t E (-a,{3) , to the Ricci flow on an open manifold which is complete on the closed ball Bgoo(o) (xoo,r) for all r < p. In particular, if p = 00, then (Boo, goo (t) ,x oo ) is complete.
4.2. 3-manifolds with positive Ricci curvature revisited. As an application of Theorem 3.10 we give a proof of the following result of Hamilton, which is a consequence of Theorem 6.3 on p. 173 of Volume One.
°
THEOREM 3.30 (Closed 3d Rc > manifolds are diffeomorphic to space forms). If (M3,gO) is a closed Riemannian 3-manifold with positive Ricci curvature, then M admits a metric with positive constant sectional curvature.
144
3. THE COMPACTNESS THEOREM FOR RICCI FLOW
PROOF. Let (M, 9 (t)), t E [0, T), be the maximal solution to the Ricci flow with 9 (0) = go. Recall that T ~ ~ (Rmin (0))-1 < 00. The proof relies on three main estimates. (1) (Positivity of Ricci is preserved.) Rc (g (t)) > 0 for all t 2: O. (2) (Strong curvature pinching.) There exist C < 00 and 8 > 0 depending only on go such that
IRc -!Rgi < CR-8
R
-
on M x [0, T) (see inequality (6.37) on p. 190 of Volume One). (3) (Injectivity radius estimate.) There exists to > 0 depending on n, T, go such that if (x, t) EM x [0, T) and r E (0,1] are such that IRm (', t)1 ~ r- 2 in Bg(t) (x, r) , then injg(t) (x) 2: toT (see Theorem 3.28 or Corollary 6.62). The main technique is to dilate and apply the compactness theorem. Choose (Xk, tk) with Xk E M and tk ---t T such that
Rk ~ R (Xk' tk) =
max
M3 x [O,tk]
R 2:
max
M3 x [O,tkl
IRml
---t
00
as k ---t 00. Consider the sequence of solutions (M, gk (t) ,Xk) , t E (-tkRk' 0], where gk (t) ~ Rkg (tk + R;lt) . Let Tk ~ R~ 1/2, which is bounded above by 1 for k large enough. We have IRm (', tk)1 ~ T;2
in B9(tk) (Xk' Tk)'
Thus by (3), injg(tk) (Xk) 2: toTk, which is equivalent to injgk(o) (Xk) 2: to· Hence Hamilton's compactness theorem implies that, for a subsequence, (M, gk (t) ,Xk) converges to (M~, goo (t) ,xoo), a complete solution defined for t E (-00,0]. We claim that (Moo, goo (t)) has constant positive sectional curvature. In particular, Moo has bounded diameter and hence is diffeomorphic to M. So the theorem follows. First note that Rgoo (xoo, 0) = limk-+oo R9k (Xk' 0) = limk-+oo 1 = 1. Hence Rgoo (x,O) > 0 for x contained in a neighborhood of Xoo' Estimate (2) says that
IRc(gk) - !R(gk)gkl (t) < CR-8R( )-8 R (gk) k gk . Since the convergence of (M, gk (t) ,Xk) is in Coo on compact subsets, we have on the subset of Moo where R (goo) > 0,
IRc(gk) - !R (gk) gkl R (gk)
IRc (goo) - !R (goo) goo I
~--~~~----~---t~------~------~
R (goo)
(we have swept under the rug the fact that in Cheeger-Gromov convergence one must pull back by appropriate diffeomorphisms from an exhaustion of Moo to M; we leave it to the reader to justify the arguments in this proof).
4.
APPLICATIONS OF HAMILTON'S COMPACTNESS THEOREM
145
On the other hand, CR;;8 R (9k)-8 ----+ o· R (goo)-8 = O. Hence we conclude that Rc(goo) = kR (goo) goo on the subset of Moo where R (goo) > O. On the other hand, the contracted second Bianchi identity implies R (goo) = const in any connected subset of the set where R (goo) > O. Hence we conclude R (goo) == 1 on all of Moo, so that Rc (goo) == kgoo on Moo. 0 4.3. Ricci flow on closed surfaces with X > O. Now we give various proofs, which are variations on a theme, of the following consequence of Theorem 5.77 on p. 156 of Volume One. THEOREM 3.31. If (M 2 ,go) is a closed Riemannian surface with positive Euler characteristic, then a smooth solution 9 (t) of the Ricci flow with 9 (0) = go exists on a maximal time interval [0, T) with T < 00. Moreover, there exists a sequence {(Xk,tk)} with tk ----+ T such that gk (t) ~ Rk9 (tk + R;;1t) , with Rk = R(Xk' tk), converges to a solution (M2, goo (t)) with constant positive curvature. First we recall the ideas of some proofs from Volume One. The first proof relies on the monotonicity of the quantity lV'iV'jf - ~LlfgijI2. PROOF #1. In this proof, which is the original proof of Hamilton, we actually recall the exponential convergence (not just sequential convergence) in Coo of the normalized flow. Consider the normalized flow %t g = (r - R) 9 on the maximal time interval [0, T). In the rest of this proof we abuse natation by using 9 (t) , t E [0, T), to stand for the normalized solution rather than the unnormalized solution in the statement of the theorem. One can prove that T = 00. If R (g (to)) > 0 for some to < 00, then we may combine the entropy and Harnack (or Bernstein-Bando-Shi derivative) estimates to show that there exist c > 0 and C < 00 such that O
on M x [to, 00). Let Mij ~ V'iV'jf - ~Llfgij, where Llf ~ R - r. From the equation (see Corollary 5.35 on p. 130 of Volume One)
!!.at IMI2 = LlIMI2 - 21V'MI2 ~
2R IMI2
LlI.l\;f12 - 2c IMI2
(using the lower bound for R), we have
IMI
(3.13)
for some C1 < Volume One)
00.
~ C1 e- ct
We also have for pEN (see Corollary 5.63 on p. 149 of
(3.14)
for some Cp > 0 and Cp < 00. By the diffeomorphism invariance of the estimates (3.13) and (3.14), this implies that the modified equation
a
{)tg =
(r - R) 9 + £.'il/g
146
3.
THE COMPACTNESS THEOREM FOR RICCI FLOW
converges exponentially fast in Coo to a gradient shrinker. The nonexistence of nontrivial gradient shrinkers (Proposition 5.21 on p. 118 of Volume One) then implies that under the original equation itg = (r - R) g, the solution converges exponentially fast in Coo to a constant curvature metric. Finally, for any go we may modify the entropy estimate to show that there exists to < 00 such that R (g (to)) > o. 0 The next proof uses Hamilton's isoperimetric estimate. PROOF #IIA. Suppose that M2 is diffeomorphic to the 2-sphere. Given an embedded loop, separating M into two connected components Ml and M2, the isoperimetric ratio of, is defined by CH (r)
~ L (r)2 (Area~M1) + Area ~M2))
and the isoperimetric constant of (M, g) is CH (M,g) ~ infCH (r) ~ 47r. 'Y
Then (see Theorem 5.88 on p. 162 of Volume One) under the Ricci flow
d dt CH (M, 9 (t)) ~ 0, so that CH (M,g(t)) ~ CH (M,go)
> O.
On the other hand, in the presence of a curvature bound, the isoperimetric constant bounds the injectivity radius by inj (M,g)
~ (4~ax CH (M,9)) 1/2,
where Kmax ~ maxM K (K is the Gauss curvature). Hence we may dilate about a sequence {(Xk' tk)} approaching the singularity of the unnormalized flow 9 (t), as in the proof of Theorem 3.30, and apply Hamilton's compactness theorem to obtain a limit solution (M~, goo (t)) . This limit solution is a complete ancient solution with bounded positive curvature. In the case of a Type IIa singularity, by choosing {(Xk' tk)} suitably, the limit is an eternal solution (attaining the supremum of R in space-time), which must be the cigar soliton
(R2, t~:t!~~ ).However, the isoperimetric estimate is pre-
served in the limit,2 which contradicts the existence of such a limit. Hence the singularity is Type 1. In this case the limit (Moo,goo (t)) is compact (see [186J or Proposition 9.16 of [111]) and has constant entropy, which implies that it is a gradient shrinker and hence is a constant curvature solution. 0 2More precisely, the limit being a cigar soliton implies that limi~oo C H (M2, 9 (ti)) 0, which leads to a contradiction.
=
4.
APPLICATIONS OF HAMILTON'S COMPACTNESS THEOREM
147
REMARK 3.32. In the above proof, we could have replaced Hamilton's isoperimetric estimate by Perelman's no local collapsing theorem in Chapter 6, which enables the application of the compactness theorem and at the same time rules out the formation of the cigar soliton singularity model. Now we give a proof that Type I limits are round 2-spheres using Perelman's entropy (see Chapter 6 for properties used in the proof below; the reader may wish to come back to this part after reading that chapter). PROOF #IIB, USING PERELMAN'S ENTROPY FOR LAST STEP. For the last step in the above proof, we may use Perelman's entropy instead of Hamilton's entropy. This has the advantage that the entropy is defined for solutions with curvature changing sign, so that we may apply it to the original solution 9 (t), t E [0, T), rather than the limit solution. We assume that 9 (t) forms a Type I singularity. Let W(g (t), f (t), T (t)) denote the entropy with T ~ T - t, which is defined for T E (0, T]. Taking f to be the constant h (t) ~ -log vOl~7)(M)' so that it satisfies the constraint
f M (47rT)-l e- h dJ.1 = 1, we see that J.1 (g (t) ,T (t)) ::; W(g,
47rT
h, T) ::; TRmllJ{ (t) - log Volg(t) (M) - 2.
In particular, by the long-time existence theorem for the Ricci flow on the 2-sphere, we have Volg(t) (M) = 87r (T - t). Hence we have an upper bound for the J.1-invariant: J.1 (g (t) , T (t)) ::; C - 2 + log 2,
where we have used the Type I assumption (T - t) RmllJ{ (t) ::; C. In particular, by the monotonicity of J.1, 1£ J.1 (g (t) , T (t)) ~ 0, the limit J.1T ~ lim J.1 (g (t) , T (t) ) t-+T
exists. Dilate the solution about (Xk' tk) with gk (t) = Rkg (tk + R"k1t) as in the proof of Theorem 3.30. By the scaling property of J.1 we have J.1 (gk (t) ,Rk (T - tk) - t)
= J.1 (g (tk + R"k1t) ,T - tk - R"k 1t ) .
Thus for each t E (-00, woo), the (maximal) time interval of existence of the limit solution (M~,goo (t)) , J.1T
== lim J.1 (gdt) ,RdT - tk) - t) = J.1 (goo (t) ,woo - t) . k-+oo
(We may assume Rk (T - tk) ~ Woo converges. Here we also used the continuity of J.1 (g, T) in 9 and the fact that the convergence, after the pull-back by diffeomorphisms, of gk (t) to goo (t) is globally pointwise in Coo since M~ ~ M is compact.) Now the theorem follows from the result that a solution having constant J.1 is a gradient shrinker. D
148
3. THE COMPACTNESS THEOREM FOR RICCI FLOW
5. Notes and commentary
Some basic references for compactness theorems for Riemannian metrics are Cheeger [70], Greene and Wu [165], Gromov [169], and Peters [300]. A survey of compactness theorems in Riemannian geometry has been given in Petersen [301]. The compactness theorem for Ricci flow in this chapter was proven by Hamilton in [187] and was used to classify singularities and nonsingular solutions in [186], [190] and [297]. Cheeger-Gromov theory was also directly used to study the Ricci flow in Carfora and Marzuoli [57]. Further compactness theorems on the Ricci flow which extend Hamilton's results can be found in [257] and [156]. It should be noted that there has been much work to ensure the injectivity radius bound for dilations of singularities, most notably by Hamilton [186] and Perelman [297]. Additional work on injectivity radius estimates has been done by Wu [372] and by the authors of [112] in the case of 2-dimensional orbifolds and [109] for sequences of solutions with almost nonnegative curvature operator.
CHAPTER 4
Proof of the Compactness Theorem We think in generalities, but we live in details. - Alfred North Whitehead There is no royal road to geometry. - Euclid
1. Outline of the proof
We now prove the compactness Theorem 3.9. This is a fundamental result in Riemannian geometry and does not require the Ricci flow. The compactness theorem is in the spirit of Cheeger [70] and Gromov [169] (see also Greene and Wu [165]' Peters [300], and the book [37]). We follow the proof for pointed sequences converging in Coo given by Hamilton [187]; as Hamilton notes there, things are easier because we can assume bounds on all covariant derivatives of the curvature. Theorem 3.9 will be proved in several steps. It is outlined as follows. STEP A: Construct a sequence of coverings of each manifold Mk which we can compare to each other. The covers should consist of balls BI: C Mk with a number of properties, most notably: • they are diffeomorphic to Euclidean balls, and for each fixed O! they have the same radii for all sufficiently large k, • they are numbered sequentially in O! starting from balls centered at the origin to balls with centers further and further away from the origin, • if we take smaller radii (BI: c BI:), they are disjoint, and if we take larger radii (BI: C BI: c BI:), they contain their neighbors, • we can bound the number of balls intersecting a given ball (the most is I (n, Co), where Co is the curvature bound), and • we can bound the number of these balls that it takes to cover a large ball in Mk, independent of k for k large (it takes fewer than A (r) balls to cover B (Ok, r) if k 2': K (r)). The specifics of this are contained in Lemma 4.18. This process is carried out in Section 3 of this chapter. STEP B: Use our nice covering to construct maps Ffd : BI: ~ M£. We do this by taking the inverse of the exponential map on the ball BI: C Mk, identifying the tangent space of Mk at the center of the ball BI: with Euclidean space and with the tangent space of M£ at the center of B't and 149
4.
150
PROOF OF THE COMPACTNESS THEOREM
then mapping by the exponential map to Me. We do this for each ball. This is done in subsection 4.1 below. STEP C: Use nonlinear averaging to glue together the maps Ffj to obtain maps Fke : B (Ok, 2k) --+ Me which take Ok to Oe (done in subsection 4.2 below). By taking subsequences, we can ensure that compositions of Fk,k+1 are approximate isometries which are getting closer to isometries as k goes to 00. STEP D: We form the limit manifold M~ as the direct limit of the directed system {Fk,k+1 : B (Ok, 2k) --+ B (Ok+1, 2k+l)} . The coordinates of B (Ok, 2k) then form coordinates for the limit Moo, i.e., for each coordinate Hk : EO: --+ B (Ok, 2k) there is a coordinate for the limit manifold defined as H~,k = h 0 Hk : EO: --+ Moo, where Ik is the inclusion of B (Ok, 2k) into Moo. Furthermore, for each coordinate EO: of B (Ok, 2k) there are Riemannian metrics 9k,e which are obtained by the pullbacks F"kege. Since Fke are approximate isometries, the sequence is equicontinuous and thus by the Arzela-Ascoli theorem we can find a convergent subsequence and get local metrics g~ k' It is not hard to see that these metrics form a Riemannian metric 900' on Moo via the coordinate charts H~ k . We can then show that the limit metric is complete and (Moo, 900' 0 00 ) s~tisfies the theorem, where 0 00 is the equivalence class of the base points in the direct limit. This is done in subsection 4.4 below. At many points in this construction we will take a subsequence; to simplify notation, at each stage the sequence will be re-indexed to continue to be k. 2. Approximate isometries, compactness of maps, and direct limits In this section we shall introduce some basic concepts which are essential to the construction of the limit manifold (M~, 900,000 ) . 2.1. Approximate isometries. In the following, the notation ITlg means the length at a point of the tensor T with respect to the Riemannian metric 9. DEFINITION 4.1 (Approximate isometry). For any 0 < c < 1 and p E NU {O}, a smooth map {[> : (M n ,9) --+ (Nn, h) is an (c,p)-pre-approximate isometry if sup I{[>*h - 91g ~ c, xEM
sup sup 1\7~ ({[>*h) l:'So::'SpxEM
I
~ c.
g
An (c,p)-pre-approximate isometry is an (c,p)-approximate isometry if it is a diffeomorphism and
I
sup ({[>-1)* g - hi xEN
h
~ c,
2.
APPROXIMATE ISOMETRIES, COMPACTNESS OF MAPS, DIRECT LIMITS 151
i.e., <1>-1 :
(N, h)
~ (M,
g) is also an (E,p)-pre-approximate isometry.
Note the condition I(<1>-1)* 9 - hlh ~ 10 is equivalent to Ig - <1>*hlcp*h ~ 10 and IVI: [(<1>-1)* g] Ih ~ 10 is equivalent to IV~*hglcp*h ~ E. Another way to express the condition sUPXEM 1<1>*h - gig ~ 10 is
j)
o<1>ao<1>b 12 _ ([ ik O<1>a] o<1>b k) (O<1>C [ .e O<1>d] 1hab oxi oxj - gij 9 habg oxi oxj - I j oxk hcdi' oxf. - I k =
I(d<1»T (d<1»
- idl2 ,
where id is the identity map on T M, the transpose comes from the two metrics hand g, and 1F12 = trace (F2) . Approximate isometries allow pointwise bounding of metric tensors as follows. PROPOSITION 4.2 (Approximate isometries and norms). Let 10 E (D, 1). (i) If<1>: (Mn,g) ~ (Nn,h) is an (E,D)-pre-approximate isometry and X is a vector field on M, then
IXI!*h ~
(1 + E) IXI~. (ii) If <1> : (Mn,g) ~ (Nn,h) is an (10, D)-approximate isometry and X is a vector field on M, then 1 2 2 2 1 + 10 IXlcp*h ~ IXl g ~ (1 + E) IXlcp*h' PROOF. (i) Using the Cauchy-Schwarz inequality on the tensor space, we find that
(<1>*h)ij XiXj
= ((<1>*h)ij - 9ij) XiXj + gijXiXj ~ 1<1>*h - gig' gijXiXj
+ gijXiXj
+ E) gijXiX j . (1 + E) IXI!*h follows ~ (1
(ii) Inequality IXI~ ~ from (i) and the fact that <1> -1 : (N, h) ~ (M, g) is also an (10, p)- pre-approximate isometry. 0 The following now immediately follows from Lemma 3.13. COROLLARY 4.3 (Norms of tensors). If <1> : (Mn, g) ~ (Nn, h) is an
(10, D)-approximate isometry, then for any (p, q)-tensor field T on M we have (4.1) (1 + E)-(P+q)/2ITlcp*h ~ ITlg ~ (1 + E)(P+Q)/2ITlcp*h. The following proposition shows how (10, D)-approximate isometries deform distances by small amounts. PROPOSITION 4.4 (Distances). If <1> : (M n , g) ~ (Nn, h) is an (10, D)-
pre-approximate isometry, then <1> (Bg (xo, r)) C Bh (<1> (xo) , (1
+ 10)1/2 r)
.
152
4. PROOF OF THE COMPACTNESS THEOREM
x
PROOF. Suppose
E
Bg (xo, r). Then
dh (
~ o:[a,b]--->M inf ~ o:[a,b]--->M inf ~ (1
lb lb a
h (
a
+ c) 9 (0,0))1/2 dt
+ c)I/2 dg (x, xo),
o
where a (a) = x and a (b) = Xo.
We shall need the following propositions about how approximate isometries affect tensors and how to compose approximate isometries.
I). Given pEN and ql, q2 E N U {O}, there exists a positive constant Cp ,Ql,q2 < 00 such that if
r-l
IV~Tlg ~ IV~*hTlg + cCp ,Ql,Q2 L
(4.2)
k=O
for all 0 < r
~
Ivi*hTI 9
p.
PROOF. We begin by proving (4.2) for p = 1. Note that fg - f
*h is a global (1,2)-tensor field since it is the difference of two connections. By (3.8) and (4.1) with C = 1 + c, we find
Ifg - f*hl g (4.3)
~
(1
+ c)3/2Ifg -
f*hl*h
~ ~ (1 + c)3/2IVg*h
~ ~ (1 + c)3IVg
since c < 1. We have
The above expression has one term of the form V <1>* hT and ql the form (f9 - f <1>* h) * T. Hence we get the estimate
+ q2 terms of
+ q2) Ifg - f*hl g ITlg IV*hTlg + 12c (ql + q2) ITl g ,
IVgTlg ~ IV*hTl g + (ql ~
using (4.3). We now induct on p. Suppose (4.2) is true for any tensor and p ~ p. Certainly we may choose Cp +1,Ql,Q2 to be greater than Cp .Q1.Q2 and so we
2.
APPROXIMATE ISOMETRIES, COMPACTNESS OF MAPS, DIRECT LIMITS 153
need only prove inequality (4.2) for r = p + 1. Then = IVPV TI IVP+1TI 9 9 9 9 9
p-1
:::;
IV~*h V gTlg + cCp,Ql,q2+1 L
IVi*h V gTI
k=O
9
by the inductive hypothesis. For the first term on the
IV~*h VgTlg
:::;
IV~*h (Vg -
V.p*h) Tig
RHS,
+ I(V.p*h)P+1 Tig'
In turn, a sum of terms of the form
I(V.p*h)k (rg - r.p*h)l g I(V.p*h)P-kTl g , where 0 :::; k < p, bounds the first term. As was shown in the proof of Lemma 3.11,
k+1 Ivi*h (rg - r.p*h) 1
:::;
Ck
9
L IV~*hgl
j=l
k+1 :::; C k
'"
L..J
9
(1 + c)(2+j)/2Ivj * .p
gl
h .p*h
j=l
2(kH)/2
-< Ck
v'22-1 c
for some constant Ck depending on k, where we have used that 4? is an (c, p + 1 )-approximate isometry and 0 :::; k :::; p. Hence we have bounded the term IV~*h VgTlg' Similarly, we can bound the terms Ivi*h VgTI, where O:::;k:::;p-1. Putting it all together, we see that in the estimate of IV~+lTlg' the coefficients of Ivi*hTlg are all bounded and all have an c in them except
IV~~hTlg . Thus we get P
IV~+1Tlg :::; IV~~hTI + cCp+1,Ql,Q2 L 9
k=O
IVi*hTI . 9
This completes the proof of the induction.
o
COROLLARY 4.6 (Norms of covariant derivatives of tensors, II). There exist C p,O,Q2 for p, q2 E N, such that if 4? : (M n , g) ~ (Nn, h) is an (c,p)approximate isometry, then for any (0, Q2)-tensor T field on N
I'J; ('T) I. oS for each 1 :::; r :::; p.
(1
+ £)<"+,,)/ 2 (l'JhTlh + cCp,a,.. ~ 1'J~t)
4.
154
PROOF OF THE COMPACTNESS THEOREM
The following proposition about the composition of approximate isometries will be used in the construction of the directed system (see subsection 2.3 of this chapter for the definition) when proving Theorem 3.9. PROPOSITION 4.7 (Composition of approximate isometries, I). There
exist Cp for pEN such that if
sup
O:::;r:::;pxEMl sup
sup
O:::;r:::;p xEM2
IVa (
IV2 ((
where we have denoted V i
~
90) Igo ::; co + c1 Cp,
(
Ig2 ::; c1 + coCp,
V gi .
PROOF. Using (4.1), we estimate
l<po
IVa (
+ cOCp,0,2
90) Igo
(IVI (
I: IV~ k=O
IVa (
90)l go
91)l g1
(
::; (1 + co)(r+2)/2 (c1 + cOCr,0,2rc1) + co. By symmetry we have
I(
IVa ((
921g2 ::; (1 +cdco +c1,
(
The proposition is proved for Cp = 2(p+2)/2 (1
+ pCp,0,2) .
o
COROLLARY 4.8 (Composition of approximate isometries, II). If
-t
are (Ei,p)-approximate isometries for i
(M a,90)
-t
(Mf+l,9H1)
= 0,1, ... k, then
(M k+l,9k+1) is a (Cp L:7=oci,p)-app roximate isometry.
PROOF. We shall induct on the number of compositions. By induction,
0 .•. 0
2.
APPROXIMATE ISOMETRIES, COMPACTNESS OF MAPS, DIRECT LIMITS ~
4.7, we have for 0
r
~
155
p
1\70((cPk 0 cPk-l 0 · · · 0 cPo)* 9k+l - 90) Igo = 1\70((cPk-l 0 . . . 0 cPo)* cP k9k+l - 90) Igo k-l
k
~ Cp LCi + ckCp ~ Cp LCi. i=O
i=O
Similarly, by induction, cPk o· .. 0 cPl is a ( Cp Ef=l Ci, P)-approximate isometry. By Proposition 4.7,
l\7k+l (( (cPk =
0'"
0
cPl
0
cPO)-l) * 90 - 9k+l)
l\7k+l ([((cPkO"'OcPl)-lrJ k
(cP
1
9
k+l
o r90-9k+l) 1
9 k+l
1
k
~ Cp LCi + coCp ~ Cp LCi. i=l Thus cPk
0 •.• 0
cPo is a
i=O
(C
p
E~=o Ci,P )-approximate isometry.
o
2.2. Compactness of maps. We shall need a version of the ArzelaAscoli theorem which applies to maps. We define Coo-convergence on compact sets for maps in Euclidean space as follows.
DEFINITION 4.9 (CP-convergence of maps). Let U and V be two open sets in jRn and let K cUbe a compact set. We say that a sequence of maps cPk : U ---t V converges to a map cP oo : U ---t V in CP on K if for every C > 0 there exists ko = ko (c, p) such that sup sup l\7r (cPk (x) - cP oo (x))1
O::;r::;pxEK
~ C
for k 2 ko·
Note that the norm given is the Euclidean norm and \7 is the gradient with respect to the Euclidean metric. DEFINITION 4.10 (COO-convergence of maps uniformly on compact sets). Let U and V are two open sets in jRn. A sequence of maps cPk : U ---t V converges to a map cP oo : U ---t V in Coo uniformly on compact sets if for any compact set K C U and any p > 0 there exists kl = kl (K,p) such that {cPkh>k _ 1 converges to cP oo in CP on K. The following is a corollary to the Arzela-Ascoli theorem, Lemma 3.14. COROLLARY 4.11 (Compactness of sequence of isometries). Let U and V be two bounded open sets in jRn. Let {9k hEN and {hd kEN be Riemannian metrics on U and V, respectively, such that the 9k and hk are all uniformly equivalent to the Euclidean metric and all of their derivatives (covariant derivatives with respect to the Euclidean metric) are bounded. If the cPk : (U,9k) ---t (V, hk) are isometries, then there is a subsequence of cPk which
156
PROOF OF THE COMPACTNESS THEOREM
4.
converges in Coo uniformly on compact sets to a Coo diffeomorphism cI>00 : U ---t V. PROOF (SKETCH). Let {xa} and {ya} be the standard Euclidean coordinates on U and V, respectively. Since cI>k are isometries, we have (4.4)
8 (cI>kt 8 (cI>k),B 8x a 8xb (hk)a,B .
(gk)ab =
Thus
_8(cI>k)a 8 (cI>k),B(h) ( )ab n8xa 8xb k a,B gk
at:/'
and the partial derivatives are all bounded. Thus there is a subsequence of {cI>d which converges to a map cI>oo. By symmetry the same argument applies to {cI>;;I}; hence cI>oo is invertible. Taking the derivatives of both sides of (4.4), we get that
8 8 2 (cI>k)a8(cI>k),B h 8xc (gk)ab = 8xc8xa 8xb (k)a,B
+
8(cI>k)a8 2 (cI>k),B h 8xa 8xc8xb ( k)a,B
+ 8 (cI>kt 8 (cI>k),B 8 (cI>k)!' ~ (hk) 8xa
8xb
8x c 8y!'
a,B .
From this equation we can express ~;;;2: as a polynomial function of 1 a (gk) ab' ay"l a (h k) a,B an d ---axaa(d' usmg . (gk) ab' (gk-1)ab ,(h) k a,B' (h-k )a,B ' axe symmetry in the usual way (see §5 of [187] for the explicit formula). Thus
I~;;;2: I can
be bounded. By differentiating the formula for ~;!'a~: and using induction, we can bound all higher derivatives of (cI>kt . This implies the corollary. 0 2.3. Review of direct limits. Let {(Ab h)}kEN be a sequence of topological spaces and open embeddings: II A 2 ---t h A 1 ---t
• • • ---t
A k ---t Ik A k+ 1
---t • •• •
Consider the compositions
hi
~
1£-1 01£-2 0 ... 0
defined for k ::; €, where hk
~
fk+l
id Ak : Ak
---t
0
fk : Ak
---t
Ai
A k , the identity map. Clearly
I£m 0 hi = hm for all k ::; € ::; m. That is, ({ AkhEN , {hih~i) is a directed system of topological spaces (see Definition 15.1 in [159]). We will use II to denote disjoint union. DEFINITION 4.12 (Direct limit). The direct limit is
limAk = (IIkAk)/ "', ~
2.
APPROXIMATE ISOMETRIES, COMPACTNESS OF MAPS, DIRECT LIMITS
157
where x y if x E Ak and y E Ae for some k, fEN and either Ae (x) = Y (if k ~ f) or !ek (y) = x (if f ~ k). The relation", is an equivalence relation. The topology on limAk is the quotient topology. "oJ
Note that direct limits can be defined for more general directed systems, but this is sufficient for our needs. Let l-e : Ae <----+ IhAk denote the inclusion map and 7r : llkAk ---t limAk denote the quotient map and define (4.5) the inclusion map into the direct limit. The topology on limA k is the finest ~
topology such that the maps Ie : Ae
---t
limA k are continuous for all fEN. ~
Since the maps A are one-to-one for all kEN, the maps Ae are one-to-one for all k ~ f. This implies the following. m
LEMMA 4.13. The maps Ie are one-to-one for all fEN, and for all f we have Ie = 1m 0 !em.
~
PROOF. The first statement is obvious. The second statement follows from l-e (xe) l-m (fem (xe)) (in Definition 4.12, we have suppressed the iden0 tification of x E At' with its image l-e (x) E llkAk)' "oJ
The following facts about direct limits are elementary. LEMMA 4.14 (An open cover for the direct limit). If Ue c Ae is an open set, then Ie (Ue) c limAk is open, i.e., Ie are open maps. Thus {Ie (AeHeEI'! ~
forms an open cover oflimAk . ~
A are open maps for all k, we have that 7r- 1 [Ie (Ue)] = !em (Ue) U (fme)-1 (Ue)
PROOF. Since the
U
U
m<e is open in llkAk. Hence Ie (Ue) C limA k is open.
o
~
This has implications about the structure of compact sets in the limit. COROLLARY 4.15 (Compact sets in the direct limit). If K C limAk ~
is compact, then for k large enough, K Kk C A k·
= h (Kk) for some compact set
PROOF. K is covered by {Ie (AeHeEI'! and since it is compact, it is also covered by a finite number of these. Since Ie (Ae) C 1m (Am) if f ~ m, we see that K is in the image of h for large enough k. Since Ik is a homeomorphism onto its image, Kk = Ik 1 (K) is compact. 0 Recall that a topological space is called second-countable if its topology has a countable base.
158
PROOF OF THE COMPACTNESS THEOREM
4.
COROLLARY 4.16 (Second-countable direct limits). If each Ak is the countable union of compact sets, then so is limA k . In particular, limAk is second-countable. 00
PROOF. If Ak =
U Kke
where Kkf is compact, then
e=1
00
00
U
00
UU
h (Ak) = h (Kke) k=1 k=le=1 which is a countable union of compact sets. limAk = --t
LEMMA 4.17 (Direct limit of Hausdorff spaces is Hausdorff). (1) If x E limA k , there exist f and Xe E Ae such that Ie (xe)
o = x.
--t
(2) If each Ak is Hausdorff, then limA k is Hausdorff. --t
PROOF. (1) Since x E limAk, there must be Xe E IIkAk such that --t
7f
(xe) = x, and thus Xe E Ae for some f. (2) Given x i- Y E limAk, there exists Xk E Ak and Ye E Ae for some --t
k, fEN such that h (Xk) = x and Ie (ye) = y. Assume without loss of generality that f ~ k. Define Xe ~ fke (Xk) E Ae. Since x i- y, we have Xe i- Yeo Since Ae is Hausdorff, there exist disjoint open neighborhoods N x and Ny of Xe and Ye, respectively. Since Ie is one-to-one and open, we conclude that Ie{Nx ) and Ie{Ny ) are disjoint open neighborhoods of x and Y, respectively, in limA k . 0 --t
The direct limit satisfies the following universal property (see Proposition 15.3 in [159]). For any space X and maps 'ljJk : Ak --+ X such that
'ljJe
0
fke = 'ljJk,
there exists a unique map W : limA k --+ X such that
3. Construction of good coverings by balls 3.1. Overview. In this section we shall prove the following lemma, which we will need in order to construct maps between different manifolds. The reader is warned that in this section we will be using superscripts which usually do not represent exponents. We shall need five radii of balls so that the smallest are disjoint and so that all others cover and are successively larger to allow for maps between the intersections. This section is quite. technical and the proofs may be skipped in the first reading.
LEMMA 4.18 (Existence of good coverings by balls). There exist a subsequence of {(M k,gk, Ok)}, convex geodesic balls Bk C Bk c Bk c Bk c
3.
CONSTRUCTION OF GOOD COVERINGS BY BALLS
159
Bk C
(Mk' gk), and functions A (r), K (r), I (n, Co), where Co is the curvature bound in Theorem 3.9 and A and K are nondecreasing in r, so that the following hold.
(1)
Bk, Bk, Bk, Bk, and Bk
are concentric, i.e., they have the same center, with center denoted as xk such that x~ = Ok. (2) Hk and HZ are disjoint for a # {3. (3) The exponential map eXPxk oLk : BOt ---t Bk is a diffeomorphism, where BOt is a ball in IRn and Lk : IRn ---t TXkMk is a linear isometry defined using an orthonormal frame at x k . Moreover, for each a the ball Bk is geodesically convex for k large enough (depending on a). (4) We have the containment
B (Ok,r) C
U 13k OI:-:;A(r)
if k ? K (r). (5) The number of {3 such that B~ n Bk =1= 0 is fewer than I (n, Co) . (6) If a, (3 < A (r), then Bk n BZ is either empty for all k ? K (r) or nonempty for all k ? K (r) . (7) If Bk n BZ =1= 0, where a, {3 :::; A (r) and k ? K (r), then Bk c B~ and
-
.... /3
Bk C Bk ·
The proof of the above lemma will occupy the rest of the section. In order to prove the lemma, we will need the following result on how the injectivity radius can decay in relation to distance. On a complete manifold with bounded curvature the injectivity radius at a point can decay at most exponentially in distance. A strong partial result in this direction was first obtained by Cheng, Li, and Yau [93]. Later, Cheeger, Gromov, and Taylor [75] obtained the following stronger estimate using different techniques. Recall that inj (x) denotes the injectivity radius at x. PROPOSITION 4.19 (Injectivity radius decay estimate). Let (M n , g) be a complete Riemannian manifold with sectional curvatures IKI :::; Co and injectivity radius inj (0) ? 1..0 > O. Then there exist constants a = a (n, Co) > o and C = C (n, Co) < 00 such that for any x E M inj (x) ? I-L [d (x, 0),1..0], where
(4.6)
. { 1..0, l}n . e-Cr . I-L [r,l..o ] =;=. a . mm
REMARK 4.20. The reason for the unnatural looking exponent n in the estimate is that the result is proved by using relative volume comparisons, and the injectivity radius inj bounds the volume comparable to injn whereas the volume V bounds the injectivity radius comparable only to V. That is, the exponent n arises from the conversion from injectivity radius to volume and then back again to injectivity radius.
4.
160
PROOF OF THE COMPACTNESS THEOREM
3.2. Choice of ball centers. In this subsection we will find centers for the balls which will make up the necessary covers. Define A [r] as
(4.7)
A[r]
~ ~J.L[r,/,o] = ;
.min{/'o,l}n.e- cr ,
where D = D (n, /'0) is a large constant, to be chosen later, depending only on n and the uniform lower bound /'0 for inj (Ok)' Note that A is a decreasing function of r. We shall choose D large enough so that A[0] $ 1. We shall work on an individual manifold (M n ,9,0) and then apply the result to (M'k,9klOk). Choose a sequence of points {xo<}:=o in M with N E N U {O, oo}, which we call a net, as follows. Let x O = 0 and rO ~ d (xO,O) = O. Let
Sl ~ {x EM: B (x, A [d (x, 0)]) n B (xO, A [rO]) = 0}. If Sl is empty, then B (xO, 2A [rO]) = M and we take N = 0 and stop choosing points. Since the balls in the definition of Sl are open, Sl is a closed set. Hence, if Sl is nonempty, then there exists a point xl E Sl such that rl ~ d (x1,0) = d (Sl,O). Since xl E sl, we have
B (xl, A [rID
n B (xO, A [rO]) = 0.
Next let
S2 ~ {x EM: B (x, A [d (x, 0)]) n B (x i3 , A [r i3 ]) = 0 for ,i3 = 0, 1} . Again, since S2 C Sl is a closed set, if S2 is nonempty (if S2 is empty, we take N = 1 and stop), we may choose x 2 E S2 such that r2 ~ d (x 2, 0) = d (S2, 0). We have
B (xi3, A [r i3 ])
n B (xl', A [rl'])
= 0
for,i3
=I 'Y E {O, 1, 2} .
By induction, assuming that the points xO, xl, x 2, ... ,xo<-l have been chosen, let
SO< ~ {x EM: B(x, A[d (x, 0)]) n B(xi3 , A [r i3 ]) = 0, ,i3 = 0, ... , a -1}. Note that SO< c So<-l. If SO< is nonempty (if SO< is empty, we take N = a-1 and stop), choose xO< E SO< such that rO< ~ d (xo< , 0) = d (SO< , 0) . Then
B (x i3 , A [r i3 ])
n B (xl', A [rl']) = 0
for,i3
=I 'Y E {O, 1,2, ... , a}.
LEMMA 4.21 (Existence of covers with bounds on the number of balls). Let (Mn, 9) be a complete Riemannian manifold with sectional curvatures IKI $ Co and injectivity radius inj (0) ~ /'0 > O. For each r > 0 there exists a nondecreasing function A (r) such that the finite collection
{B(xo<,2A[ro<]): 0 $ a $ A(r)} forms a cover for B (0, r) and rO< > r if a > A (r). Furthermore, we can choose A (r) to depend only on n, r, Co and /'0 (in particular, not depend on the manifold M).
3. CONSTRUCTION OF GOOD COVERINGS BY BALLS
161
There are two steps to the argument. The first is that the balls 0: :S A'} cover for some A'. The second is that we can find a bound A for A' which is independent of the manifold. Let A' (r) ~ max {o: : rQ :S r}. Note that this definition ensures that rQ > r if 0: > A' (r) . We will now show that {B (xQ, 2A. [rQ]) : 0 :S 0: :S A' (rn form a cover of B (0, r). Consider p E B (0, r) and let s ~ d (p, 0). If pis not covered by these balls, then d (p, xQ) 2: 2A. [rQJ for all 0: :S A' (r) . Hence B(p,A.[s]) is disjoint from the balls {B(xQ,A.[rQ]): O:S O::S A'(sn since if q E B (p, A. [s]), then for 0: :S A' (s) < A' (r), PROOF.
{B (xQ, 2A. [rQ]) : O:S
d (q, xQ) 2: d (p, xQ) - d (p, q) 2: 2A. [rQJ - A. [sJ 2: A. [rQJ since A. is decreasing. This implies that p E SA'(s)+1. This is a contradiction because r A '(s)+1 must be the minimal distance from to a point in SA'(s)+1, but since r A '(s)+1 > s, the minimum should have been s = d (O,p). In order to estimate A' (r), we shall use the curvature bound to get volume estimates. The Rauch Comparison Theorem [72J and our injectivity assumption imply that there is a number E = E (n, Co) depending only on the dimension and the upper curvature bound of Co such that
°
Vol (B (xQ, A. [rQJ)) 2: EA. [rQt for all 0: :S A' (r). By the Bishop Volume Comparison Theorem, there is a number M = M (n, Co) depending only on the dimension and the lower curvature bound of -Co such that Vol (B (0, r)) :S M exp Since A. [rJ :S A. [rQJ for
0:
[(n - 1) vi'Cor] .
:S A' (r), we have
A'(r)
L
Vol (B (xQ, A. [rQJ)) 2: (A' (r) + 1) EA. [rt· Q=o Since the balls B (xQ, A. [rQ]) are disjoint and are contained in B (0, r + A. [0]), we also have that A'(r)
L
Vol (B (xQ, A. [rQJ)) :S Vol (B (0, r + A. [OJ))
Q=o :S M exp [(n -1)
vi'Co (r + A. [0])] .
Thus we get the bound
(A' (r) + 1) EA. [rr :S M exp
[(n - 1) vi'Co (r + A. [0])]
or '() Mexp[(n-l)VCo(r+A.[O])] A r :S EA. [rr - 1. We can take A (r) to be the least integer greater than this number. Since rQ is increases as 0: increases, we see that if 0: > A (r), then rQ 2: r. 0
162
4.
PROOF OF THE COMPACTNESS THEOREM
In summary we have constructed a sequence of balls such that the following holds: PROPOSITION 4.22 (Good cover of a Riemannian manifold). Let (Mn, g) be a complete Riemannian manifold with sectional curvatures IKI ~ Co and injectivity radius inj (0) 2: /'0 > o. Then there exists a net of points {xa} ~=o ' where N E Nu { 00 } , and a nondecreasing function A (r) (which depends only on n, Co and /'0) such that (1) r a = d (x a , 0) is a nondecreasing function of a, (2) the balls {B (x a , A [raJ)} ~=o are disjoint, and
(3) B (0, r) C
Ua::;A(r)
B (x a , 2A [raj) .
3.3. Application to the sequence of manifolds. We are now ready to apply our construction of the coverings by balls to the sequence M k. For each Mk we can construct nets {xn as in Proposition 4.22. Let r~ ~ d (x~, Ok) . We next show that the A [r~] are bounded for each a, so we can find a subsequence which converges.
PROPOSITION 4.23 (Bounds on the distance of the centers to the origins). r~ ~ 2aA [0] and r~ 2: A [0] for all a i= O. PROOF. The worst case scenario is if the construction is a string of balls such that the centers are on a distance-minimizing geodesic. In this case, the distance r~ ~ A [0] + 2A + A [r~] ~ 2aA [0]. 0
E3:i [rf]
COROLLARY 4.24 (Convergence of the distance of the centers to the origins). There exists a subsequence {(Mk,9k, Ok)}kEN and positive numbers {r~} aEN such that, for each a, r~ ---* r~ as k ---* 00. Hence there is a function K (a) such that if k 2: K (a), then
~A [r~] ~ A [r~] ~ 2A [r~]. PROOF. This is a simple diagonalization argument.
o
COROLLARY 4.25. Let K' (r) ~ max {K (a) : a ~ A (r)}. Then if a ~ A (r) and k 2: K' (r), then
~A [r~] ~ A [r~] ~ 2A [r~]. PROOF. If k 2: K' (r) and a
~
A (r), then k 2: K (a).
o
We will denote A [r~] by Aa . Now define the following collection of balls: DEFINITION 4.26 (Various size balls). Define iJ~ ~ B (x~ , Aa /2) ,
Bka
--
B
iJ~ ~ B (x~ , 4A a) ,
(XOk' 205e 20cC Aa) ,
3.
CONSTRUCTION OF GOOD COVERINGS BY BALLS
163
where c and C are defined by A [r] =l= ce- cr as in (4.7), i.e., a . mm . { /'0, 1}n . c =;=. D
(4.8)
The strange radii is to ensure that Lemma 4.18 is true and it will later be clear why these numbers were chosen. We easily see the following properties: PROPOSITION 4.27 (Disjointness of smaller balls and covering of larger balls). If k 2: K' (r) and a :::; A (r), then B~ are disjoint and B (Ok, r) C Ua::O;A(r)
B~.
PROOF. It follows from Corollary 4.25 that
B~
c
B (Xk' A [r~]),
B (xk' 2A [r~])
c B~.
The proposition then follows immediately from Proposition 4.22.
0
PROPOSITION 4.28 (Bound on index of intersecting balls). For any a ~ 0 there exists an integer I (a, n) (independent of k) such that if
B~nB~ # for some
13
and k, then
13 :::;
0
I (a, n) .
PROOF. This follows easily from Lemma 4.21. If y E B~
r~ (4.9)
n B~, then
+ d (Xk'Y) + d (y, x~) :::; r~ + 5A a + 5A,B :::; r~ + lOA [0], :::; d (Ok,xk)
so by Proposition 4.23,
r~ :::; (2a + 10) A [0] since A [0] :::; 1. By Lemma 4.21 we have take I (a, n) =l= A ((2a + 10) A[0]).
13 :::;
A ((2a
+ 10) A [0]).
Now just
o
PROPOSITION 4.29 (Stability of the intersections of balls). There is a subsequence {M k,9k} such that for every pair (a, 13) there is a number K (a, 13) such that either B~ n B~ is empty for all k ~ K (a, 13) or B~ n B~ is nonempty for all k ~ K (a, 13) . PROOF. Let {(ae,J3e)}eENU{O} be an ordering of the elements of the countable set (NU {O}) x (NU {O}). We construct inductively a nesting sequence of subsets Ke c N. Let Ko be an infinite subset such that B~o n B~o intersect always or never for all k E Ko. This can be done since either the balls intersect for infinitely many k, and we take Ko to be the set of such k, or they do not intersect for infinitely many k', and we take Ko to be the set of these k'. Now for each subsequent f we can construct an infinite subset Ke of Ke-1 such that B~i nB~i intersect always or never for all k EKe. We now consider the collection of {Ke} eENu{O} as nested subsequences of N and take a diagonal subsequence. Define a function K : (NU {O}) x (NU {O}) ---+ N
164
4. PROOF OF THE COMPACTNESS THEOREM
such that K (ae, f3t.) E Kt. is a monotone function of f. If k ~ K (at., f3e), then k EKe. Hence either B~i n B~i is empty for all k ~ K (at., f3t.) or it is nonempty for all k ~ K (at., f3e). 0 DEFINITION 4.30.
Let K (r) ::§= max (K' (r), {K (a, (3) : a, f3 :S A (r)}).
By definition, we have that if a, f3 :S A (r), then K (a, (3) :S K (r). Note also that K (r) is increasing. This proves statement (6) of Lemma 4.18.
3.4. Estimates on inclusions of intersecting balls. Now to complete the proof of Lemma 4.18 (only statements (5) and (7) remain to be proved), we recall Definition 4.26 and prove the following proposition. If B~nB~ -f3 -f3 C Bk and B~ C B k .
PROPOSITION 4.31.
we have
B~
PROOF.
=1= 0,
where k
~K
(max {r~, r~}), then
Recall estimate (4.9), which gives
r~ :S r~ + lOA [0] . Now, we can estimate Af3 by Af3
~ ~A [r~] ~ ~A [r~ + lOA [0]].
Recall that the definition of A is A [r]
So A [0]
= c and
A [r~
= ce- cr .
+ lOA [0]] = e- IOcC A [r~].
Hence
Af3 > ~e-lOCC A [rO:] > ~e-lOcC AO: - 2 k - 4
or A0: :S 4e lOcC Af3.
Choose ayE B~
n B~.
Then for each x E B~ we get
d (x,x~) :S d (x,x~)
+ d (x~, y) + d (y,XO < 5AO: + 5AO: + 5A,B :S 20e 10cC Af3 + 20e 10cC A,B + 5A f3 :S 45 e lOcC Af3.
Similarly, if x E jj~, then
d (x,x~) :S d (x,x~)
+ d (xk'Y) + d (y,x~) < 45elOcC A + 5A + 5A f3 0:
0:
:S 45elOcC (4elOCC A(3 ) + 5 (4elOCC A(3 ) + 5A f3 :S 205e 20cC Af3 .
o
4. THE LIMIT MANIFOLD
(M~,goo)
165
Finally, we want to ensure that the Bk are embedded geodesic balls and to ensure that the requirements of Proposition 4.53 are satisfied, so we need that 205e 20cC )..0t < 41Oe 2occ ).. [rOt] < ! inJ' (x Otk), k - 3 which is ensured if 1230e 20cC ).. [rk] ~ J.£ [rk,
/'0]
or by (4.6), equivalently, 1230e 20cC /D ~ 1,
i.e.,
20c'C ~ Dlog(D/1230),
where c' ~ a·min {/,o, l}n (so that by (4.8) we have c ~ c' / D). Since D log D goes to infinity as D ---t 00, we can choose D large enough to satisfy this inequality and also such that c = d / D is less than 1. Finally, we can make D large enough so that the balls have radius less than 7r / (6VCO) , and hence are convex by Corollary 4.47, and so that the balls have radius less than cd VCO as in Proposition 4.32. To complete the proof of statement (5) of Lemma 4.18, we need to show that the number of {3 such that B~ n Bk i= 0 is fewer than I (n, CO) . For any {3 such that B~ n Bk i= 0, B~ c Bk. As in the proof Lemma 4.21, we can estimate the volume of Bk from above by a multiple of ()..Ott and the volume of each B~ from below by a multiple of ()..Ott; this will give the bound I (n, Co).
4. The limit manifold
(M~,
900)
We can now construct limit overlap maps of balls and limit metrics; we shall use these to take a direct limit and find the limit Riemannian manifold (M~,900). Let B k , Bk , Bk , A(r), K(r), and I (n,Co) be as in Lemma 4.18.
4.1. Local metrics on balls, transition functions, and their limits. Given r > 0, consider the ball B (Ok, r) . For this section we shall always assume that k 2: K (r). Note that by Lemma 4.18, B (Ok, r) is covered by the finite collection of balls 13k, where Q ~ A (r). For each Q ~ A (r) we shall construct maps
Ffj : Bk ---t Me using the exponential maps of Mk and Me. In addition, we wish to average these maps in such a way that the maps limit to the identity, in a sense, as k, £. ---t 00. We first look at the metrics in normal coordinates in balls and obtain convergence of the metrics locally. Choose linear isometries
Lk : IRn
---t
Tx~Mk'
We can then define all of the diffeomorphisms H- Ot . E-Ot ---t B- Ot k .
k,
4.
166
PROOF OF THE COMPACTNESS THEOREM
as restrictions of the map
expxooLk, k where g~, EO:, and EO: are the appropriately sized Euclidean balls centered at the origin (the three maps are the same, but they are defined on different domains).1 We have the following standard result (for the proof see Corollary 4.12 in [187]). PROPOSITION 4.32 (The 1'\7£ Rml :S Ce imply lamgl :S Cm in normal coordinates). Let (M n , g) be a Riemannian manifold. Let p E M and ro E (O,! inj (p)) . Assume that for all f ~ 0 there are constants Ce < 00 such that Then in the normal coordinates {xi} on B (p, ro) there are constants Ce depending on n, inj (p), Co, . .. ,Ce and a constant Cl depending only on n such that for any multi-index a with lal ~ 1
1
"2 (Oij) :S
(gij) :S 2 (Oij)
and
in B (p,min {cI/VCQ,ro}).
The {
(Bk' (Hr) -1) }
o:~A(r)
form coordinate charts covering B (Ok, r) C
Mk. Since
(-0/3)
d3 =;=. Hk * gk gk
are Riemannian metrics on E/3 in normal coordinates with uniformly bounded curvatures, by Proposition 4.32, all partial derivatives of the metrics are uniformly bounded. Hence by the Arzela-Ascoli theorem there is a subsequence converge uniformly in Coo on compact sets to a limit Rieso that the mannian metric ~ defined locally on E/3. We use our convention that the subsequence is still indexed by k. We have transition maps on Mk defined as follows. Recall that for a, {3 :S A (r), if n BZ =1= 0 for some k, then it is true for all k ~ K (r) by Lemma 4.18. If Bk n BZ =1= 0, then it makes sense to define the maps
tt
Bk
by
J~/3 ~
(RZ)-1 oHk ,
J~/3 ~
(i1Z)-1 oRr.
The maps J~/3 and J~/3 are embeddings; since for all" gy C E'Y C E'Y, J~/3 is a restriction of J~/3. IThe balls B k , EO,
13k , and Hk
are given by Definition 4.26; consequently the radii of
E"', and E'" are equal to 5>."', 45e lOcC >."', and 205e 20cC >. "', respectively.
4. THE LIMIT MANIFOLD
(M~,goo)
167
J:
To obtain the local convergence of f3 as k -+ 00, we take some further subsequences. Since the J:f3 are Riemannian isometries between llk ~ (H;:r 9k and ~ for each k and since the derivatives of the metrics are bounded, we have that for each pair a, f3 ~ A (r) such that B;: n B~ i- 0 for all k 2 K (r), there exists a subsequence such that the J:f3 converge to a limit transition map J~ : ECt -+ ef3 in Coo uniformly on compact sets (by Corollary 4.11). In fact, J:;! is a f3 is a restriction of Riemannian isometry between g~ and gf!. Since J:f3, we also have that J:f3 converges to a map J:;!. We then diagonalize the sequences so that J:f3 converges for every a and {3. Notice that since J~Ct 0 f3 = idf3 : Ef3 -+ ef3, the identity embedding, we must have that
J:
J:
J!Ct 0
J~ = idf3 .
4.2. Constructing approximate isometries Fke;r of large balls in Mk into Me. We can now construct approximate isometries Fke;r between the ball B (Ok, r) C M k and an open set in Mf. for sufficiently large k and f. (We shall use some results about the center of mass given in Section 5 of this chapter.) The following is the main result of this subsection. PROPOSITION 4.33 (Existence of an approximate isometry on a large ball). For every r > 0, c > 0, and p > 0 there exists ko = ko (r,c,p) > 0 such that for k, f > ko there is a diffeomorphism
Fke;r : B (Ok, r)
-+
Fke;r (B (Ok, r))
c
Me
which is an (c,p)-approximate isometry.
Let FfJ. : B;:
-+
Be
(from a ball in Mk to a ball in Mf.) be defined by
Ct . H eO Ct (HkCt )-1 . F kf.7 Roughly speaking, we construct the desired map Fkf.;r by averaging the local maps Ffj for a ~ A (r). Notice that in terms of the (inverse) coordinate systems (Ef3,Hf) and (e f3 where f3 satisfies B~nBk i- 0, the map
,iif),
Ffj corresponds to the map
Ct .. Ef3 F kf.,f3
-+
E-f3
between Euclidean balls defined by Ffj,f3
~ =
(4.10)
(iif) (iif)
H~
-1 0
Ffj
-1 0
He (H;:r 1
0
0
0
Hf
168
4.
PROOF OF THE COMPACTNESS THEOREM
Hence we have the following local property which is a key step to Proposition 4.33. PROPOSITION 4.34 (Local maps converge to the identity in a sense). If a and j3 are such that Bk n Be =f. 0 for k sufficiently large, then the maps F[d,{3 : E{3 ~ jj{3 converge to the identity (inclusion) map id{3 as k, £ ~ 00. N ow we proceed to average the local maps to construct a map on a large ball. To apply Proposition 4.53 on averaging maps, we need to construct a partition of unity subordinate to the covering {Bk} o<::;A(r) of B (Ok, r) as
follows. For a :S A (r) let '!j;0< be a smooth function which is 1 on EO< c EO< and 0 outside of EO< (EO< = (H k )-1 Bk). We construct a partition of unity on B (Ok, r) by letting '!j;0<
0
(Hk)-1 (x)
'f 1
cpl:(x)~ { L-'Y::;A(r)'!j;'Yo(Hn- (x)
o
E
1 X
if x
BO< k'
~
B k,
where a :S A (r). By Lemma 4.18(4) the denominator is no less than 1 and by Lemma 4.18(5) the number of terms in the denominator which are well defined is bounded by I (n, Co) , independent of k. To ensure that the basepoint is preserved, we need the partition of unity function (
H£
B2
~ IR
to be, for 0 < a :S A (r) ,
<po< (x) == k
.
I
Xk (x) . '!j;0<
'!j;oo(H2f 1 (x)
+
0
(Hk )-1 (x)
L- Xk (x).'!j;'Y o (Hn- 1 (x) O<'Y::;A(r)
if x E B k ,
o
while for a = 0, if x E
B£,
The collection of functions {
4. THE LIMIT MANIFOLD
(M~,goc)
(Ei3, He)
Notice that with respect to the coordinates where a f= 0, can be expressed as a
. =;=
c/>k,{3
i3 Hk =
a
c/>k
0
X Jk80 . '1/'
'ljJ
;30
0
Jk
c/>~,(3
the map c/>f. (x) ,
k
Jaa
0
+ LO<"Y::;A(l') X 0
so with respect to the coordinates (E(3, to a function
'a
0
0
169
80,
J k . 'IP"Y
,3"Y '
0
Jk
He) , we have that c/>k.(3 converges
defined by v 0 },(30 . • I,a 0 },(30 A 00 'f' 00
A.a. 'f' 00,(3 =;=
'ljJ
0
130
0
J oo
~ + wO<"Y::;A(1') X0
(30!3"Y . 0 J oo
J oo . 'ljJ"Y
When a = 0, we have
The definition of Fke;1' is as follows. For x E B (Ok, r), we define (4.11)
~ cm {FEe (x), Fff (x), ... , Fte(1') (x)} EMf
Fke;1' (x)
to be the center of mass using the weights c/>k (x); by the choice of balls in Lemma 4.18 we can apply Proposition 4.53 and conclude the existence of Fkf;1' when k, f are large enough. The map Fkf;1' is smooth with all its derivatives IV7P Fkf;1'1 bounded by constants Cp +1 independent of k. From the construction of the weights c/>k we have Fke;1' (Ok) = 0(. Note by Proposition 4.53 and the definition of the center of mass, Fke;1' satisfies (i) Fkf;1' (x) is the minimizer y E Me of A(1')
Ix (y)
=
L c/>k (x) d~f (y, Fke (x)),
a=O
(ii) Fkf;1' (x) is the solution y
E Me of
.4(1')
L ¢k (x) exp;1 Ffj (x) = 0,
(4.12)
a=O
where the exponential map is with respect to ge. With respect to the coordinates (E!3, equation (4.12) can be written in E(3 as
He) ,
L a::;.4(1')
c/>k
0
He (X) exp-1
(j
Fkf;roHk (X)
Ffj
0
He (X) = O.
4.
170
PROOF OF THE COMPACTNESS THEOREM
Define the local versions of Fkejr by
G~ejr ~ We may pull back to
L
o<~A(r)
ej3
cPr;:
(Hff)
-1 0
via the map 0
Fkejr
He.
0
Hff to get
He (X) exp;~.
(X)
F ke ,j3 (X)
= 0,
kl,r
(Hff)
where now the map exp is with respect to the metric 9~ = * ge. To see the limiting behavior of F kejr when k, P are large, we note that since Proposition 4.34 implies for each {3 that we have
Fke ,j3
~
idj3
when k, P ~
00,
then by Proposition 4.54, we have G~ejr ~ idj3 on any compact subset of E/3 in Coo. We have proved the following. PROPOSITION 4.35 (The maps Fkejr converge to id in a sense). For every r > 0, E > 0, and pEN U {O} there exists ko = ko (r, E,p) such that for
(3:::;A(r),
!Vp (G~ejr -
idj3)! :::;
E
for all k, P ~ ko, where V and 1·1 are the covariant derivative and norm with respect to the Euclidean metric on E/3.
As a corollary we have the following. COROLLARY 4.36 (Fkejr is a local diffeomorphism). There exists ko = ko (r) such that if k,P ~ ko, then FkejrlB!3 is a diffeomorphism for each k (3:::;A(r). PROOF. If G~e-r is sufficiently close to the identity map, then it must be injective since its derivative is nonsingular. 0 Now we turn to proving that given (c,p) and r, for k, P large enough, Fke,r is an (E,p)-pre-approximate isometry. First we have the following general result. LEMMA 4.37 (Limit of almost-identity pullbacks). Let cPk : U ~ U C jRn be diffeomorphisms, let id : U ~ U be the identity map, and let {hk} kEN and hoo be Riemannian metrics on U. Suppose hk and hoo are uniformly equivalent to the Euclidean metric for all kEN and their derivatives (covariant derivatives with respect to the Euclidean metric) are uniformly bounded. If cPk ~ id and hk ~ hoo in Coo uniformly on compact sets, then for every E > 0, pEN, and compact set K c U, there exists ko = ko (C, p, K) such that if k ~ ko, then
4.
THE LIMIT MANIFOLD
(M~,goo)
171
where 1·1 is the Euclidean norm and \7 is the Euclidean covariant derivative (i.e., partial derivative). PROOF. Let x = let
=
{Xi}
be the standard Euclidean coordinates on U and (
1;xji
have that ~ ----+ 8f and ----+ 0 uniformly on K where a is multi-index with lal ;: : 2. Since hk ----+ hoo in Coo uniformly compact sets, we have that 8 (hk(X»Qb ----+ 8 (h oo (x»Qb uniformly on K for any a Now (8X)Ci (8X)Ci . Q
Q
* - hoo = (8 0,
\7
r (
----+
0
= O.
* ) (80: (( ) 8
lo:l::;r
80: (hk)ab 8
,
where 8o:,i,j is a sum of terms of the form
80:1 (hk)ab (8xtl with lall+la21+la31 = rand on K. The lemma is proved.
80: 2 8
80:3 8
(8xt2 8x i
la21 ;: : 1. Hence \7r (
900)
----+
0 uniformly D
With this we are ready to prove the following. LEMMA 4.38 (Fkl,r is an (c,p)-pre-approximate isometry). For anyc > 0 and p > 0 there exists K, = K, (c, p) such that
1\7~k (Fkl;r9l - 9k) I gk ~ c for all q ~ p if k, R. ;:::: isometry.
K,
(c, p) . Hence Fkl,r is an (c, p) -pre-approximate
PROOF. We work in a coordinate chart
(E/1, He) . By Proposition 4.35,
for any c > 0 there exists ko = ko (r, c) such that l\7q ( G~l;r - id/1) I < c if k,R.;:::: ko. By Proposition 4.32, the metrics = * 9k are uniformly equiv-
(He)
ge
E/1.
alent in the Coo-norm to the Euclidean metric on Thus it suffices to estimate the partial derivatives of Fkl -r9l - 9k using the Euclidean metric. Since
G~h ----+ id/1 and 9~ ----+ 9~,
r
we
~ay use Lemma 4.37 to conclude that
(G~l;r ~~ ----+ 9~ in the Coo-Euclidean norm as k, R. -
timates now follow from the fact that 9~ as k ----+ 00.
----+
00.
The desired es-
9~ in the Coo-Euclidean norm D
172
4.
PROOF OF THE COMPACTNESS THEOREM
Next we turn to prove that Fkf.;r is a diffeomorphism. Since Ff.k is the inverse of Fre, by Proposition 4.34, for each (3 we have Fre,(3 ~ id(3 and FCk ,(3 ~ id(3 when k, £ ~ 00. Then by a simple argument using Proposition 4.54 we conclude that Ff.k.roFkC,r and Fkf.,roFf.k,r both approach the identity map when k, £ ~ 00. It follows that Fkf.,r is invertible. Now it follows from the inverse function theorem and Proposition 4.35 that Fke~r is an (c',p)-pre-approximate isometry. Hence, given (c,p) and r, there is a ko such that Fkf;r is an (c, p )-approximate isometry for k, £ ~ ko. Proposition 4.33 is proved. 4.3. The directed system. We are now in a position to construct a directed system whose direct limit will give us the limit manifold (M~, goo). We first show that, after passing to a subsequence, the existence of approximate isometries whose compositions are also approximate isometries, as close as we like to isometries. PROPOSITION 4.39 (Metrics are almost isometric on large balls). There
exists a subsequence
{(Mk' gk) } J
there exists jo
J
JEN
such that for any c > 0 and pEN
= jo (c,p) EN such that if j > jo, then there exist maps Wj: B (Ok j ,2j ) ~ B (OkJ+1'2j+l)
with
Wj (Ok j ) = OkJ+l such that for any £ E N the composition map Wj,f.
~ Wj+C-I 0 ' "oWj+1 oW j: (B (Ok j ,2 j ) ,gkj ) ~
(B (Ok jH,2 jH ) ,gkjH )
is an (into) (c,p)-approximate isometry. PROOF. We may assume that Cj is increasing as j increases and that
Co ~ 1. We shall inductively define the subsequence
{(M k., gk) } J
J
JEN
. It
j)-
is sufficient to construct a sequence {Wj }jEN such that Wj is a ( C;12- j , approximate isometry. In this case we can use Corollary 4.8 to see that Wr,f. is a (Cr ~r~;-l C i- 12- i , r )-approximate isometry. In fact, since Cj is increasing in j, we have
r+f.-l 00 00 1 2- i < C ~ C:- 12- i < ~ 2- i = 21- r Cr ~ C:~ t r~ t -~ , i=r i=r i=r which implies that
(2 l - r , r )-approximate isometry. We also have by Wo (B (Oka' 1)) C B (Ok!' (1 + COl) 1/2) C B (Ok!, 2)
wr,f
Proposition 4.4 that since
is a
4.
THE LIMIT MANIFOLD
(M~,goo)
173
Similarly, we have
again since C r 2: 1. Hence, given c > 0 and p > 0, we can take jo ::;:: max (1 -log2 E',p) . For j = 0, make ko large enough so that Fko£; 1 is a (Co 1 , 0) -approximate isometry for any f 2: ko (we can do this by Proposition 4.33). By induction, suppose we can do this up to k r . We then make kr+1 large enough so that Fkr+1£;2r+1 is a (C;~12-(r+1), r + I)-approximate isometry for any f 2: k r+1' Now choose w1• ::;:: Fkrkr+l;2r. 0 Let us re-index the subsequence taken in the previous proposition so that it is once again indexed by k and the index of W coincides with the index of M. We may now take the final subsequence to get metrics on B (Ok,2 k ) C Mk which will become the limit metric. Since Wj,£ are approximate isometries, we may consider the sequence {W J* £gj+£}OO ,
£=0
of Rie-
mannian metrics on B (OJ, 2j) . Since Wj,e are (c, p )-approximate isometries independent of f, the metrics W;,egj+£ are uniformly bounded together with its derivatives and so there is a subsequence in f so that they converge to a limit metric gj,oo on B (OJ, 2j) . We can use this argument diagonally to get the following proposition. PROPOSITION 4.40 (Existence of almost-isometric limiting metrics on large balls). There exist a subsequence {kj } ~1 and Riemannian metrics gkj,oo on B (Ok j , 2kj) such that for every c > 0 and p 2: 0 there exists jo = jo (E',p) such that
for all r :S p and f 2: 0 if j 2: jo.
PROOF. This essentially follows from Lemma 4.37 again.
o
We again re-index, replacing Wkj with Wj, which equals Wkj+l 0 . . . 0 Wkj+2 0 Wkj+l 0 Wk j in the old notation, so that we have a sequence of maps Wj : B (OJ, 2j) ---t B (0}+1, 2}+1) (note that if j corresponds to k j , then we have shrunk the ball of radius 2kj to the ball of radius 2j ). We note that Wj : (B (OJ,2 j ) ,gj,oo) ---t (B (0}+1,2 j +1) ,g}+l,oo) is an isometry since
Iw;gj+1,oo - gj,ool :S Iw; (gj+1,oo - W;+I'" W;+£_lgj+e) 1 + IW;W;+1 ... W;+£_lgj+£ - gj,ool and both terms on the RHS go to zero as f
---t
00.
174
4.
PROOF OF THE COMPACTNESS THEOREM
4.4. Construction of the limit. We are now ready to construct the limit manifold (M~, 900) . Topologically, we take the direct limit
M~ ~ ~B (Ok,2 k ), where the directed system comes from the maps Wk. Note that since Wk are approximate isometries, they must be open embeddings. Hence Moo is a Hausdorff space by Lemma 4.17. We recall the embeddings h : B (Ok,2 k ) --+ Moo defined in (4.5). The coordinate maps HI: : gl< --+ B (Ok, 2k) induce coordinate maps H~,k ~ h 0 HI: : E Ol --+ Moo. Note that the transition maps
f3 )-1 (Hoo,k
0
Ol Hoo,k+r = (f3)-1 h 0 Hk
(3)-1
= (Hi
(4.13)
0
0
h+r 0 H kOl+r
Wk,r
0
HI:
are Coo diffeomorphisms (when the domain and range are suitably restricted) and hence they induce a Coo structure on Moo. Furthermore, since Wk,r are isometries between 9k,00 and 9k+r,00, we easily see that the transition maps are isometries and there exists a metric 900 on Moo such that I k900
= 9k,00·
We now show that {(M k,9k, Ok)} converges to (M~, 900, 0 00 ) . Given a compact set K c Moo, it must be contained in h [B (Ok,2 k )] for some k > 0 and hence must also be contained in Ii [B (Oi, 2i)] for all f. ~ k. We now claim that for every p there exists ko = ko (K,p) such that for any € > 0 sup xEK
for all a get
~
Iv~oo
p, k
~
IVOl (900 -
(1;;1)* 9k) I
9=
ko. This follows by pulling this expression back by h to
(900 - (1;;1)* 9k) 1900 =
=
Ilk [V~oo (900 IV~k'oo (9k,00 -
(1;;1)* 9k)] 1/;900 9k)1 9k,00
and by using Proposition 4.40. Hence the maps 1;;1 satisfy the requirements of Definition 3.5 and we have shown the following. PROPOSITION 4.41 (Convergence to a limit). {(M k,9k,Ok)} converges
to (M~,900,000).
Furthermore, we can show that the limit metric is complete. PROPOSITION 4.42 (The limit is complete). The metric 900 is complete. PROOF. Any closed geodesic ball 13 c Moo is contained in the image h [B (Ok,2 k )] for some kEN. Recall that h is an open embedding, which
5.
CENTER OF MASS AND NONLINEAR AVERAGES
175
implies r;;1 (B) is closed and bounded, and hence compact since gk is complete. Therefore B is the image of a compact set, and hence compact. We are done because if closed metric balls are compact, then the metric is complete (see, for instance, [72]). 0
5. Center of mass and nonlinear averages In this section we review some standard work on the convexity of the distance function and properties of the center of mass. The treatment here follows mostly that in Buser and Karcher [40], although we address additional issues related to proving Coo-convergence. In this section we adopt the convention 7r / ~ 00 when ~ 0 and we assume that geodesics have constant speed, i.e., they are parametrized proportional to arc length.
(2JK)
K
5.1. Derivatives of the distance function and exp-l. Let (Mn,g) be a Riemannian manifold and let d (x, y) denote the distance between x and y. Fix x E M and consider the function f (y) ~ ~d2 (x, y). We can write f as an integral by letting 'Y (r) be a minimal geodesic from 'Y (0) = x to 'Y (1) = y, so that
f(y) =
~
11
gh,"y)dr
since geodesics have constant speed. The quantity 9 ("y, "y) is constant and equal to the square of the length of the geodesic. The gradient of f can be expressed as follows. LEMMA 4.43 (Gradient of the distance squared function). If y is not in the cut locus of x, then grad f (y)
= - exp;1 x
E
TyM.
PROOF. Let 'Y (r) be the unique minimal geodesic from 'Y (0) = x to 'Y (1) = y; then by the Gauss lemma, grad f (y) = "y (1). It follows easily from the uniqueness of solutions of the geodesic equation that expy (-"y (1)) = x, so that -"y (1) = exp;1 x. The lemma is proved. 0 Given Y E TyM, let a : (-E, E) --t M be the geodesic with a (0) = y and it (0) = Y (where E > 0 is sufficiently small for later purposes). Since y is not in the cut locus of x, there exists a smooth family of unique minimal geodesics 'Ys : [0,1] --t M, such that 'Ys (0) = x and 'Ys (1) = a (s) for s E (-E, E) . Then 'Yo = 'Y is the minimal geodesic joining x and y. Define (7 : (-E, E) X [0,1] --t M by (7
(s, r) ~ 'Ys (r)
= expx
[r exp,;1 a (s)]
= eXPa(s)
((1 - r) exp~ts) x) ,
176
4.
PROOF OF THE COMPACTNESS THEOREM
where we are considering the curves "Is both in terms of geodesics from x and geodesics from a (s) . Note that
8(J 8(J . 8s (s,O) = 0, 8s (s, 1) = a (s), 8(J -1 81' (s, 1) = - eXPa:(s) x. The second derivative of
f
has the following expression.
LEMMA 4.44 (Hessian of the distance squared function). The Hessian of
f is given by
where J (1') is the Jacobi field along the geodesic between x and y parametrized on l' E [0,1] such that J (0) = 0 and J (1) = Y E TyM. PROOF. Given Y E TyM, define "Is and (J as above. Let J s (1') ~ ~~ (s, 1') be the Jacobi field along "Is; then J s (0) = 0 and J s (1) = a (s). Using 81T( -1 x, we comput e 81' s, 1) = - eXPa:(s) \7yexp;1 x
= - \7{)/8s ~~ (S,1')1
(s,1')={O,I)
= - (\78/81'JO) (1).
o The following is essentially the Hessian comparison theorem. LEMMA 4.45 (Hessian comparison). If the sectional curvatU1'e of (M n , g) is bounded above by K, then there exists a constant C = C (K) > 0 such
that for any y E B (4.14)
(x, (2VK)) 1r /
(Hess!) (Y, Y)
not in the cut locus we have
= -g (\7y exp;1 x, Y)
~ C 1Y12,
Y E TyM.
PROOF. By Lemma 4.44 we need to estimate
g((\78/8r J) (l),J(l)) Note that we can write Y and c E R Then
d id ="21 dr [g(J(r),J(r))] 1'=1 = IJI dr IJI 1'=1' 1
= y-L +ci' (1) , where Y -L J (1') = J-L (1')
is perpendicular to 'Y (1)
+ C1''Y (1'),
where J-L (1') is a Jacobi field satisfying J-L (0) = 0, J-L (1) = y-L and J-L (1') ..l 'Y (1'). Now it is clear that we only need to estimate IJI IJII1'=1 assuming that Y is orthogonal to 'Y (1) .
t.
177
5. CENTER OF MASS AND NONLINEAR AVERAGES
We compute using the Jacobi equation
~ IJI = !£ dr2
dr
(9 (\7 IJI
a/ar J , J))
(\7 a/arJ, J)2 9 (\7 alar \7 a/arJ, J) j\7a/arJj2 =IJI 3 + IJI + IJI 9
9 (\7 a/ ar J,J)2
IJI
3
9(R(J,i)i,J)
IJI
-
j\7 alar Jj2
IJI
+
'
so
::2 IJI + K bl 21J1 = IJI- 1 (IJ121i1 2K
- 9 (R (J, i) i, J))
+ 1J1-3 (IJI2j\7 a/arJj2 ~
where we used J (r) ..1
-y (r)
9
(\7 alar J, J)2)
0, to conclude that
IJ1 21-y12 K -
9 (R (J,
-y) -y, J)
~
O. The corresponding ODE for cp (r) is
cp" + K
(Iii = d (x, y)
1-y12 cp = 0
is a constant since 'Y is a geodesic) and has solutions
cp (r) = cp (0) cS K1 i'12 (r) where
+ cpt (0) sn Khl 2 (r),
fi sin ( JK,r ) { sn/i; (r) = r ~ sinh (J-K,r)
if if if
K,
> 0,
K,
= 0,
K,
< 0,
and COS
cs/i; (r)
=
{
(JK,r)
1 cosh (J-K,r)
if if if
K,
> 0,
K,
= 0,
K,
< O.
The functions sn/i; (r) and cs/i; (r) are the solutions to cp" + K,cp = 0 with sn/i; (0) = 0, sn~ (0) = 1, cS/i; (0) = 1, and cs~ (0) = O. Note that J /IJI is a unit vector (and has a limit as r ----t 0) and that
\7 alarJ (0) is well defined. From
i. IJI
= 9 (\7 a/arJ,
ir
I~I)
we know that the
limit limr-+o+ IJI is also well defined. Now we compare IJI (r) with the solution cp (r) which satisfies cp (0) = IJI (0) = 0 and cpt (0) = limr-+o+ i.IJI. Note that
cp (r) = cpt (0) snKhl 2 (r)
4. PROOF OF THE COMPACTNESS THEOREM
178
and that ¢ (r) is nonnegative for r E [0,1] when K ~ 0 or when K > Ii'I ~ 1r. Assuming Ii'I < 7r if > 0, we compute
JK
JK
K
0 and
(IJI' ¢ - IJI ¢')' = IJI" ¢ - IJI ¢" = (::2 IJI + K 1i'12IJI) ¢ 2: 0 for all r E [0, 1] . Integrating this from 0 to r gives us
IJI' (r) ¢ (r) -IJ (r)1 ¢' (r) 2: IJI' (0) ¢ (0) + IJ (0)1 ¢' (0) = 0, that is, for r E (0,1]' IJI' (r) 2: IJ (r)1 :
(4.15)
(~}.
Hence
IJI.!!:.-IJII dr
provided
2:
IJI 2(1) ¢' (1) ¢ (1)
r=1
JK 11'1 < 7r when K> O. This proves 1
-g (Vy exp; x, Y) CS K1
d
= IJId IJI r
I
2:
r=1
. 12 (1)
Note that"! is positive either when K sn 2 (1) KhI
JKIi'ol
cSK bl 2
snKIi'12 ~
(1)
2
(1) WI .
0 or when K > 0 and
D
<7r/2.
Recall that a C 2 function ¢ is (strictly) convex if its Hessian is positive definite: VV ¢ > O. COROLLARY 4.46 (Local convexity of the distance squared function). Suppose the sectional curvatures of (Mn, g) are bounded above by K. Then
the function f (y) ~ !d2 (x, y) is convex for any y E in the cut locus of x. PROOF.
This follows directly from (4.14).
B(x, 7r/ (2JK))
not
D
We also have COROLLARY 4.47 (Convexity of small enough balls). Suppose the sectional curvatures of (Mn, g) are bounded above by K. Then the ball B (0, r)
is convex if r
~ min {inj 0, 7r /
(2JK) }.
PROOF. Suppose x, y E B (0, r) . Let 'Y (t) be the constant speed minimal geodesic between x and y. We simply need to show that d ("( (t), 0) < r for every t. Consider the function f (z) = !d (0, z)2 . By Corollary 4.46, we have that
5.
CENTER OF MASS AND NONLINEAR AVERAGES
which implies that the maximum of f
h' (t))
179
occurs at the endpoints. Hence
d(O,'Y(t)):S max{d(O,x),d(O,y)} < r.
o The next lemma will be used in proving the smooth dependence of the center of mass in the next subsection. PROPOSITION 4.48 (On the derivatives of exp-1). Let (Mn, g) be a Riemannian manifold such that all derivatives of the curvature are bounded: l\7 t Rml :S Gt
for f
°
= 0,1,2, ....
There is a constant c (n) > such that for any p EM and x, y E B (p, r1), where r1 :S min inj (p) , c/ JCo} , if x is not in the cut locus of y, then (i) we have
{t
(4.16)
1\7~1 \7;; exp;l xl
:S Ct1+t2+1
for f1' f2
°
= 0,1,2, ... ,
where Ct = Ct (n, inj (p) ,f, Go, . .. , Gt) > are constants independent of x and y, and \7 y and \7 x are the covariant derivatives with respect to y and x, respectively; (ii) when x, y ~ p* E B (p, r1), we have (4.17)
(\7 x exp;l x : TxM
~ TyM) ~ (id : Tp.M ~ Tp.M) ,
(\7y exp;l x: TyM ~ TyM) ~ (-id: Tp.M ~ Tp.M) , where we use parallel translation to identify TxM and TyM with
Tp.M and to define the convergences above. PROOF. (i) Let w = {w k } be normal coordinates on B (p, r1). By Proposition 4.32 (Corollary 4.12 in [187]) we have in the coordinates w, (4.18)
~ (8ij ) :S (gij) :S 2 (8ij )
and
I(:;r:tgijl
:S Clol'
where a is a multi-index. In particular the Christoffel symbols (4.19)
rfj
satisfy
ij
80 k I I (8w)Or :S Cjol+1'
N ow we consider the exponential map exp : T M ~ M with expy z = x for Z E T M using the coordinates w; here we abuse notation in that x = (xk) stands for both a point in M and its coordinates in the coordinate system w (and the same for y). Define f (r, y, z) , :S r :S 1, by d2 fk k dfi dfj (4.20) dr 2 + r (f) dr dr = 0,
°
ij
fk (O,y,z) = yk, dfk k dr (O,y,z)=z.
4. PROOF OF THE COMPACTNESS THEOREM
180
Uk
Then (1.y,z)) theorem to
(Xk) =
expy z. We will apply the implicit function
F(x,y,z) ~ f(l,y,z)-x to prove that z = is a smooth function of (x, y) and that expy 1 x has the required derivative estimates. Consider the boundary value problem for the first-order ODE expy 1 x
d:Ifk _ hk dr ,
d~k +
rr
j
(1) hihj = 0,
fk (O,y,z) = yk, hk (O,y,z)
= zk.
From f(r,y,z) E B(p,rt) , Ih(r,y,z)l g = Izlg(p) = d(y,x), and !(Iij)::::; (9ij), we have Ifk(r,y,z)1 ::::; V2rl and Ihk(r,y,z)1 ::::; V2d(y,x). From the smooth dependence property for ODE (see Theorem 4.1 on p. 100 of Hartman [196]), fk (r, y, z) is a smooth function of (r, y, z) . Using the proof of Theorem 3.1 on p. 95 of [196] and (4.19), it follows from an induction argument on the order of derivatives that for r E [0, 1] ,
8 0 + 13
(4.21)
(8y)Ct (8z)f3 fk (r, y, z) ::::; CICtI+If3I+l'
8Ct +f3
(8yt (8z)f3 hk (r,y,z) ::::; CICtI+lf3l+1' Actually the proof of Theorem 3.1 on p. 95 of [196] implies the estimate above for Inl + 1,81 = 1. Let "'fy,z be the geodesic with "'fy,z (0) = y and (fr "'fy,z) (0) = z. Let Jy,z,z (r) denote the Jacobi field along the geodesic "'fy,z with Jy,z,z (0) = 0 and (:1rJy,z,z) (0) = Z E TyM. Then the covariant partial derivative in the direction is
z
Dzf (1, y, z) (i)
= :sf (1, y, z + Si)ls=o = Jy,z,z (1).
To show Dzf (1, y, z) : TyM ---- Texpy zM is invertible, we prove that there is a constant Co > 0 such that IJy,z,z (1)1 ~ Co Iii. This follows from the Rauch comparison theorem; here we give a proof using (4.15). As in the proof of Lemma 4.45, it suffices to prove that IJy,z,z (1)1 ~ Co Iii for those which
z
are orthogonal to z in TyM. From (4.15), we have (iJY;i:r)l)' r = I-I z sncolzl2 (r ) . S'mce I'Imr-->O IJy,z,z(r)1 ¢(r)
A. ( )
'I'
IJy,z,zl (1) ~ ¢ (1)
~ 0, where
1 we h ave IJy,z,z(r)1 > 1 an d ¢(r) _
----,
= Iii sncolzl2 (1)
~ Co
Iii·
181
5. CENTER OF MASS AND NONLINEAR AVERAGES
Using JzJ ::; 2c/VCo, we can choose Co ~ sn4c2 (1). We have proved (4.22) Now we can apply the implicit function theorem to F (x, y, z) -x. From Dyf
(1, y, z)
~
f (1, y, z)
{)z
+ Dzf (1, y, z) {)y = 0,
Dzf
(1, y, z)
~:
- id = 0,
we can take higher-order derivatives of the equations above to get formulas c 8"+ lj z ' f h . 1 d" 8",+{3 fk (1 lor (8y)"'(8xi III terms 0 t e partIa envatIves 0 f (8y)"'(8z)i3 ,y, z ) and (Dzf
(1, y, z))-l . From (4.21) and (4.22) we can estimate
{)Ot+!3z ({)yt ({)x)!3
_
by induction on the order of derivatives. From (4.18) we know that the bounds of the covariant derivatives I"\7~1 "\7~2 Z I follow from the bounds of 8",+i3z I I(8y)'" (8x)i3 .
(ii) Let w be normal coordinates on B (p*, r*) for sufficiently small r* and let 9ij ~ 9 (8~i' 8~J) . From Theorem 4.10 in [187] we have (9ij) --t (b"ij) on B (p*, r*) as r* --t O. Hence the geodesic equation (4.20) in the coordinate chart w has a solution (r, y, z) which converges in C 1 to (r, y, z) ~ y+rz as r* --t O. So for x, y E B (p*, r*) , exp;l x converges in C 1 to exp~~ x ~ x - y as r * --t O. The estimate (4.17) follows from (9ij) --t (b"ij) on B (p*, r *) and
1
{)
-1
100
{)-1
{)xi expooy x = ei and {)yi expooy x = -ei, where ei = (0" .. , 1" .. ,0) is the unit vector in i-th direction. The lemma now is proved. 0 4.49. (i) Under the assumptions of Proposition 4.48, let P* E B (p, rl) and let w be normal coordinates on B (p*, r*) for sufficiently small r*. Then when x, y --t P*, we have that exp;l x converges in C 1 to the map x - y, where x = {xk} stands both for a point in M and its coordinates in the coordinate system w (and the same for y = {yk}). (ii) Suppose h is another metric on M. From the proof of Proposition 4.48 it is not difficult to see that the map (exp9)-1 : M x M --t T M is close to (exph) -1 : M x M --t T M on any compact set in C k when 9 is very close to h on any compact set in C k , for any kEN. REMARK
182
4.
PROOF OF THE COMPACTNESS THEOREM
5.2. Nonlinear averages. Let (Mn,g) be a complete Riemannian manifold with sectional curvatures bounded above by K. Let p E M and
q1,.·.,qk E B(p,r), where r < min{1inj(p)'6~}' Let J-L1, ... ,J-Lk be nonnegative numbers with J-L1 + ... + J-Lk > O. We define the center of mass with weights J-L1,· .. , J-Lk, cm {q1, ... , qk} = cm(JL1, ... ,JLk) {q1, ... , qd, as the minimizer of 1>: M 1
1> (q) ~ "2
(4.23)
R,
---+ k
L J-Li d2 (q, qi) . i=l
LEMMA 4.50 (Existence of center of mass). Let p E M and q1, ... , qk E B (p, r) for some r < 6~' Suppose inj (q) > 3r for all q E B (p, r). Then there exists a unique minimizer cm {q1, ... , qk} of 1> in M. Furthermore we have cm {ql, .. . , qk} E B (p, 2r) and
uniformly in J-L 1, ... , J-Lk· PROOF. It is clear that for any q E M \ B (p, 2r) , we have 1> (q) > 1> (p) . Hence the minimizer of 1> exists and must be contained in B (p, 2r) . Note that if q E B (p, 2r), then q E B (qi, 3r). Since B (qi, 3r) C B( qi, 7r / (2VK)), by Lemma 4.43, the functions q 1---+ ~d2 (q, qd are strictly convex in B (p, 2r) . Since the weights J-Li are nonnegative and J-Ll +- . '+J-Lk > 0,1> is strictly convex in B (p, 2r) . Hence the minimizer must be unique. To see the last statement, we apply the first part of the statement to B (q*, r *) in place of B (p, r), where r * is small. We get that when ql,···,qk E B(q*,r*), we have cm(JL1, ... ,JLkdql, ... ,qk} E B(q*,2r*) for all
0
(J-Ll, ... ,J-Lk). By Lemma 4.43 we have k
grad1>(q)
=-
LJ-Liexp;lqi. i=l
The minimizer occurs at a point q where the gradient of 1> is zero, so that k
L J-Li exp;l qi = O. i=l The following proposition tells us about the derivatives of the center of mass.
5. CENTER OF MASS AND NONLINEAR AVERAGES
183
4.51 (Dependence of cm on weights and points). Suppose (Mn, g) is a Riemannian manifold such that all of the derivatives of the curvature are bounded: PROPOSITION
IV'I!
Rml
"'.5:.
Cl
for.e
= 0,1,2, ... .
Let iiI, . .. , lik be nonnegative weights with iiI + ... + lik > 0. There exists a constant c(n) E (O,~) such that for any p E M, if inj (q) > 3r for all q E B (p, r) , where r
< ~. Then we have the following.
(i) (Bounds on the derivatives of cm) The unique center of mass cm(J.t1, ... ,J.tk)
{ql,"" qkl
is a smooth function of ql, ... , qk E B (p, r) and iiI, ... , lik' The V'~V'~ -covariant derivatives of cm(J.t1 ,... ,J.tk) {ql, ... , qk} , with respect to ql, ... ,qk and iiI, ... , lik, satisfy
IV'~V'~ cm(J.t1, ... ,J.tk) {ql," . , qk}1
(4.24)
"'.5:.
0lal+I13I+1'
B B ) where V' q = (V' q1' ... , V' qk) and V' J.t = ( BJ.t1'···' BJ.tk and qal+I13I+1 are constants depending on n, inj (p), lal+I.BI, and Co,· .. ,qal+lf3I+1' (ii) Forql, ... ,qk E B(p,r) such thatqI, ... ,qk ---+ q* E B(p,r) (i.e., the points tend to each other), we have (a) (change in a weight has negligible effect on cm)
IV' J.ti cm(J.t1, ... ,J.tk) {ql, ... ,qkll
---+
0,
(b) (effect of the change in a point on cm) ( V'qi CmeJ.t1, ... ,J.tk)
---+ (z:/Li . j=1 J.t]
{ql, ... , qk} : Tqi M
id : Tq.M
---+ Tcm(
1'1,···,l'k
){q1, ... ,qk}M)
---+ Tq.M) ,
(c) (effect of the change in a weight and point on cm)
aa
( V' qj lij ---+
(L
k1
cm(J.tl,. .. ,J.tk)
.
i=1 J.t,
{ql, . .. , qkl : Tq;M
id : Tq.M
---+ Tcmeu
u ){q1 .... ,qd M )
r1>-"' 'rk
---+ Tq.M) .
The convergences above are defined using parallel translation to identify Tqi M with Tq. M. PROOF.
(i) We apply the implicit function theorem to the family of
maps defined by k
Gq1, ... ,qk'~1. ... ,J.tk
(q) ~
L Iii exp;1 qi· i=1
184
4. PROOF OF THE COMPACTNESS THEOREM
By the previous lemma, cmC/Ll, ... ,/Lk) {ql, ... ,qk} is the unique solution of the equation
G (q) ==. G ql,···,qk./Ll,···,/Lk (q) Consider the partial derivative
= 0.
k
\1qG = LJLi\1qexp;l qi: TqM
---7
TqM.
i=l
By Lemma 4.45, \1 qG is positive definite with smallest eigenvalue being bounded from below by a constant depending only on Co and JLl,·.·, JLk. It follows from the implicit function theorem that the unique solution cmC/Ll, ... ,/Lk) {ql, ... , qd is continuous in ql, ... , qk and JL 1, ... , JLk· To see that the \1~\1~-covariant derivatives of cmC/Ll, ... ,/Lk) {ql, ... , qk} are bounded, we compute the other partial derivatives of G:
Hence (4.25)
(\1qi G , 8~j
G) + \1qG· (\1q; cm {ql, ... ,qk}, \1/Lj cm {ql, ... , qd)
= 0,
where q = cm {ql, ... , qd. Thus
(\1 qi cm {qI, ... , qk}, \1/Lj cm {ql, ... , qk})
= - (\1 qG)-l ( \1q;G, 8~j
G) .
This and Proposition 4.48(i) implies (4.24) when lal + 1,81 = l. To bound the higher derivatives of cm { ql, ... , qk} , we argue inductively on the order of the derivative lad + 1,81. We take the appropriate derivatives of (4.25) of order lal + 1,81 - 1 with respect to ql,···, qk and JLl,···, JLk so that \1 qG . \1~\1~ cmC/Ll, ... ,/Lk) {ql, ... ,qd appears in the resulting equality. Then
\1~\1~ cmC/Ll ,... ,/Lk) {ql, ... , qk} can be expressed in terms of \1~~ exp;l qi with i l ::; lal + I,BI , \1~~\1 q exp;;l qi with £2::; lal + 1,81, \1~l\1~lcmC/Lh .. '/Lkdql, ... ,qd with lall + 1,811 ::; lal + 1,81- 1, and (\1qG)-l. Now it is easy to see from Proposition 4.48(i) that 1\1~\1~ cmC/Ll, ... ,/Lk) {ql, ... , qk}1 are bounded by constants (\~I+I,BI+l depending on n, inj (p), lal + 1,81 , and Co,· .. , qal+I,BI+1· (ii) When ql, ... , qk ---7 q*, by Lemma 4.50, we have a~; G ---7 O. By Proposition 4.48(ii) we have \1 q;G
---7
JLi id and \1 qG
---7
-
(I:~=l JLi) id.
5.
CENTER OF MASS AND NONLINEAR AVERAGES
185
This proves the first two convergences. Next we estimate 8 V qi -8 cm(J.!l,- .. ,J.!k) {ql,' .. , qd . J.,Li Since we have V qi 8~i G
= V qi eXPql qi
8
--+
id, by taking V qj-derivative of
8
-8 G + VqG· -8 cm{ql, ... ,qd
J.,Lj J.,Lj in (4.25) and taking the limit, we get
=
0
(t
J.,Li) V qj 88. cm(J.!l, ... ,J.!k) {qI, .. . , qk} = O. i=1 J.,LJ This proves the third convergence. id -
D
REMARK 4.52. (i) Note that in Euclidean space ]Rn, we have the following formula for the center of mass 1 cm(J.!l, ... ,J.!k) {qI, ... ,qk} = (J.,Llql + ... + J.,Lkqk). J.,Ll + ... + J.,Lk It is clear that there are many derivatives of the form
V~V~ cm(J.!l, ... ,J.!k) {ql, ... , qd , where lal + 1,81 ~ 2, whose lengths do not approach 0 as ql, ... ,qk --+ q* E ]Rn. (ii) Suppose h is another metric on M. From the proof of Proposition 4.51 and Remark 4.49(ii), it is not difficult to see that, as a function of (J.,Ll, ... , J.,Lk, ql, ... ,qk), the center of mass map cm9(J.!I,···,J.!k ) {ql, ... , qd is close to cm h(J.!l,···,J.!k ) {ql, ... , qk} on any compact set in Coo when g is very close to h on any compact set in Coo. We can use the center of mass to average maps. We have the following. PROPOSITION 4.53 (Averaging maps). Let (Nn,h) and (Mn,g) be Riemannwn manifolds such that all derivatives of the curvature of M a're bounded: e ~ Ce for £ = 0,1,2, ....
IV Rml
Let J.,Li (x) , i = 1, ... , k, be a finite sequence of smooth nonnegative functions on N with compact support in Ui C N and with bounded derivatives. Let W be an open set with closure W C UiJ.,Li l (0,00). Let Fi : Ui --+ M be a finite sequence of smooth functions with bounded derivatives. Suppose that, such that for any j with for any Xo E W, there exist io and ro E (0,
6Jco)
Xo E J-ljl (0,00) we have Fj (xo) E B (Fio (xo) , !ro) and inj (q) > 3ro for all q E B (Fio (xo) , ro). Then there is a function F : W --+ M defined uniquely by minimizing L:~=l J-li (x) d2 (F (x), Fi (x)) . In particular, F (x) satisfies
!
k
(4.26)
LJ.,Li (x)exP'Ftx) Fi (x) = O. i=l
186
4.
PROOF OF THE COMPACTNESS THEOREM
Furthermore F (x) is smooth and its derivatives Vi F (x) are bounded by constants depending on the bounds for IVii l1i (x) I with fi ::; f, the bounds for IVii Fi (x)1 with fi ::; f, and n, f, and Co,···, CHI'
PROOF. Fix Xo E W. By assumption, there exists a small neighborhood Vxo C W of xo, io, ro, and jl,'" ,jm such that Fjk (x) E B (Fio (xo) , ro) for k = 1, ... , m and all x E Vxo and such that I1j (x) = 0 for all j i= jl, ... ,jm and x E Vxo' By Lemma 4.50, for x E V we can define Fxo (x) on Vxo to be cm(". ""11 (xl
". ) {FJ' 1 (x), ... ,FJ· (x)} 7n
""W]7n
which is the composition of cm(. .) {qjl' ... , qjm} and I1jk = I1jk (x) , J.LJI,···,J.L]m qjk = Fjk (x). From the uniqueness of the center of mass it is easy to see that for any two points Xl, X2 E W, FXl (x) = FX2 (x) on VX1 n VX2 ' Hence this defines F : W - 4 M. We now prove the second part of the lemma. Let Xo E W; then, on VXo ' F (x) is the composition of cm(". .) {qjl'" . ,qjrrJ and I1jk = I1jk (x) , ""JI "'·,J.LJm qjk = Fjk (x). It follows from the chain rule and Proposition 4.51(i) that F (x) is smooth on Vxo and that F (x) on Vxo has the required derivative 0 bounds. We also used the following convergence property. PROPOSITION 4.54 (Average by cm of maps limiting to id limits to id). Let Bl C B2 be two open subsets ofJRn and let gk, where kEN, be a family of Riemannian metrics on B 2. Assume all derivatives of the curvatures of gk are uniformly bounded and gk - 4 goo on any compact set in B2 in Coo. Suppose Ft: : Bl - 4 B2, for a = 1, ... ,A, are sequences of smooth maps such that Ft: - 4 id uniformly on compact sets in C l for each a as k - 4 00. Let 11k be partitions of unities for each k on B l . For any compact set K C Bl we can define Fk : K - 4 B2 for k sufficiently large by letting Fk (x) be the center of mass of Ft: (x) with weight 11k (x) with respect to metric gk., i.e., Fk (x) is defined by A
L 11k (x) eXPF~(X) Ft: (x) ~ 0, a=l
where the exponential map eXPFk(X) is with respect to gk. Then Fk converges to id as k - 4 00 uniformly on any compact set in Bl in C l .
PROOF. Because some of IV~V~ cm(J.Ll, ... ,J.Lk) {ql,"" qk}l, lal + 1/31 ~ 2, do not approach 0 as ql, . .. ,qk - 4 q*, we will not prove this proposition using the composition employed in the proof of Proposition 4.53. By Remark 4.52(ii), cm9(k l() A()) {F1 (x), ... , Ft (x)} can be made arbitrarily close J.Lk x ,···,J.Lk X
to
cm9(ool() A()) J.Lk x ,···,J.Lk x
{F1 (x), ... , Ft (x)}
on any compact set x EKe Bl in
Coo when we choose k large enough. On the other hand, fix Xo E Bl and let in (B2' goo) . It follows from Remark
w be normal coordinates centered at Xo
6. NOTES AND COMMENTARY
187
4.49(ii) that by choosing k large and x to be in a very small neighborhood of Xo, we can make cm9(ool() A()) {Ff (x), ... , Ft" (x)} arbitrarily close /l-k x ,···,/l-k x
in the C 1-topology to the Euclidean center of mass
1() 1 A() (llk(x)Ff(X)+"'+Il:(x)Ft"(x)) , Ilk x + ... + Ilk x where we have identified the point FI: (x) E Bl with its coordinates in the coordimite chart w. Since Ilk (x) + ... + 11: (x) == 1, we have
1() 1 A() (llk(x)Ff(X)+"'+Il:(x)Ft"(x))-x Ilk x + ... + Ilk X
= Ilk (x) (Ff (x) - x) + ... + 11: (x) (Ft" (x) - x), which clearly converges to 0 on any compact set within the coordinate chart Now the proposition is proved. 0
w in C 1 when k ---+ 00.
6. Notes and commentary For some additional references on compactness theorems not cited in the previous chapter, see Cheeger and Gromov [73], [74], Gao [152]' Yang [374], [375], Anderson [4] and Anderson and Cheeger [6].
CHAPTER 5
Energy, Monotonicity, and Breathers Truth is ever to be found in the simplicity, and not in the multiplicity and confusion of things. - Sir Isaac Newton The most beautiful thing we can experience is the mysterious. It is the source of all true art and science. - Albert Einstein
Much of the 'classical' study of the Ricci flow is based on the maximum principle. In large part, this is the point of view we have taken in Volume One. As we have seen in Section 8 in Chapter 5 of Volume One, a notable exception to this is Hamilton's entropy estimate, which holds for closed surfaces with positive curvature. 1 Even in this case, the time-derivative of the entropy is the space integral of Hamilton's trace Harnack quantity, which satisfies a partial differential inequality amenable to the maximum principle. 2 Indeed, this fact is the basis for Hamilton's original proof by contradiction of the entropy estimate which uses the global in time existence of the Ricci flow on surfaces. 3 Originally, Hamilton's entropy was a crucial component of the proofs for the convergence of the Ricci flow on surfaces and the classification of ancient solutions on surfaces. Via dimension reduction, the latter result has applications to singularity analysis in Hamilton's program on 3-manifolds. An interesting direction is that of finding monotonicity formulas for integrals of local geometric quantities. Beautiful recent examples of this are Perelman's energy and entropy estimates in all dimensions. We briefly touched upon these estimates in Section 8 of Chapter 1 (Theorems 1.72 and 1.73) to motivate the study of gradient Ricci solitons. Perelman's energy is the time-derivative of a classical entropy ((5.64) in Section 4 below). Observe how the resulting calculation in Perelman's proof of the upper bound for the maximum time interval of existence of the gradient flow (Proposition 5.34) is reminiscent of Hamilton's proof of his entropy formula. In fact this upper bound says that a modified classical entropy is increasing (see (5.67)). Monotonicity formulas usually have geometric applications. In particular, Perelman proved that any breather on a closed manifold is a Ricci soliton of the same type. This statement includes the shrinking case which remained open until his work; previously, we have seen the proofs of the ISee [108], Proposition 5.44, for the case of curvature changing sign. 2See (5.70). 3See Theorem 5.38. 189
190
5. ENERGY, MONOTONICITY, AND BREATHERS
expanding and steady cases in Proposition 1.13. To prove the nonexistence of nontrivial breathers, Perelman needed to do a separate study of each type of breather. However, in each case, the method is the same: introduce a new functional, study its properties, and apply them to the proof that there are no nontrivial breathers of each type. All such functionals have three basic characteristics: • they are nondecreasing along systems of equations including the Ricci flow, • they are invariant under diffeomorphisms and/or homotheties, • their critical points are gradient Ricci solitons (of a different type in each case). Moreover, Perelman's functionals are successive modifications of his initial functional :F and are motivated by the consideration of gradient Ricci solitons of each type. So it is important to study the cases of the proofs successively in order to see how the evolutions of the functionals are used and how to modify the functionals gradually to define the entropy functional, which is the key to proving the shrinking case and where the proof follows essentially the same steps as the other two cases but uses the new functional. In this chapter, we shall discuss in detail the energy functional, its geometric applications and its relation with classical entropy; in the next chapter we study Perelman's entropy and some of its geometric applications. The style of this chapter is that of filling in the details of §§1-2 of Perelman [291] in the hopes of aiding the reader in their perusal of [291]. Throughout this chapter Mn is a closed n-manifold. 1. Energy, its first variation, and the gradient flow
The Ricci flow is not a gradient flow of a functional on the space 9J1et of smooth metrics on a manifold Mn with respect to the standard L2-inner product. 4 On the other hand, variational methods have played major roles in geometric analysis, partial differential equations, and mathematical physics. It was unusual that the Ricci flow, a natural geometric partial differential equation, should appear to be an exception to this. Perelman's introduction of the :F functional (defined below) solved the important question of whether the Ricci flow can be seen as a gradient flow. More precisely, as we shall see in this and the following section, the Ricci flow is a gradient-like flow; it is a gradient flow when we enlarge the system. The key to solving the question above is to look for functionals whose critical points are Ricci solitons, that is, fixed points of the Ricci flow modulo diffeomorphisms and homotheties (so that the ambient space in which we consider Ricci flow is 9J1et/Diff x ~+ instead of 9J1et). This is consistent with the point of view we adopted in Chapter 1 on Ricci solitons. 4An exception is when n Kahler-Ricci flow.
=
2 (see Appendix B of [111]), and more generally, for the
1.
ENERGY, ITS FIRST VARIATION, AND THE GRADIENT FLOW
191
1.1. The energy functional F. Let COO(M) denote the set of all smooth functions on a closed manifold Mn. We define the energy functional F : met x COO(M) ~ lR by (5.1) Note, in addition to the metric, the introduction of a function f. This embeds the space of metrics in a larger space. We shall sometimes follow the physics literature and call f the dilaton. Since ~ (e- f ) = (-~f + l~fI2) e- f , we see from fM ~ (e- f ) dJ.L = 0 that (5.2) So we have two other expressions for the energy: (5.3)
1M (R + ~f)e- f dJ.L = 1M (R + 2~f - I~ fI2)e- f dJ.L.
F(g, f) =
(5.4)
The second way of expressing the energy is motivated by the pointwise formula (5.43) in subsection 2.3.2 below. LEMMA 5.1 (Elementary properties of F). (1) Dirichlet-type energy. The geometric aspect of F is reflected by F (g, 0) = fM RdJ.L being the total scalar curvature and the function theory aspect of F is reflected by expressing it as
where w = e- f /2 \ which is a Dirichlet energy with a potential term. (2) Diffeomorphism mvariance. For any diffeomorphism 'P of M, we have F('P*g,f
0
'P) =F(g,f).
(3) Scaling. For any c > 0 and b \.
F (c 2 g, f + b)
=
cn - 2 e- b F (g, f).
EXERCISE 5.2. Prove the properties for the energy in the lemma above.
1.2. The first variation of F. We use the symbol 8 to denote the variation of a tensor. We shall denote the variations of the metric and dilaton as 8g = v E Coo (T* M @sT*M) and 8f = h E Coo (M), and we
5. ENERGY, MONOTONICITY, AND BREATHERS
192
define V ~ gij Vij. Routine calculations give
k 1 kl 8vrij (g) ="2 g (\1iVjl
(5.6)
p
(5.7)
8vrpj
(5.8)
8(v,h) (e-fdl-l)
+ \1jVil -
\1IVij) ,
1
= "2\1jV, =
(~ -h)e-fdl-l.
We calculate the last one, for example,
= -e-fh dl-l + e- f ~gijV'ij dl-l =
8(v,h) (e- f dl-l)
(5.9)
(~ -
h) e- f dl-l.
LEMMA 5.3 (First variation of F). Then the first variation of F can be expressed as
(5.10)
8(v,h)F(g,f)
1M Vij(Rij + \1i\1j f)e- f dl-l + 1M (~ -h) (2D.J-I\1JI +R)e- fd l-l, =-
2
where 8(v.h)F (g, f) denotes the variation oj F at (g, f) in the direction (v, h), i.e., 8(v,h)F(g'f) PROOF.
~
:
s
I
F(g+sv,J+sh).
8=0
Recall (Vl-p. 92), i.e., P !:l·r R tJ·· -- R Ppij -_!:l t Up r Pij - U pj
q rP + r ijq r ppq - r pj iq'
so that
8~j = \1p
(8rfj) - \1i (8r:j) .
Since \1 i\1j = o/lj - rfjOk as an operator acting on functions, we have
8 (\1 i\1j f)
(8rfj) \1pJ.
= \1 i\1j (8f) -
Hence, using (5.7),
8(~j + \1 i \1 j f) =
\1 p
(8rfj) - (8rfj) \1 PJ + \1i
= ef \1 p
(e -f 8rfj) + \1 \1 i
j
(\1 j
(8 f) - 8r:j)
(h - ~) .
We then compute
(5.11)
8 [(~j
+ \1i\1jf) e- f dl-l]
+e-f\1i\1j(h-~) + \1i\1j f) e- f (~ - h)
= [\1p(e- f 8rfj ) + (~j
1dl-l.
1.
ENERGY, ITS FIRST VARIATION, AND THE GRADIENT FLOW
193
So using (5.8),
8 [(R+~J)e-fdl-l]
= gij 8 [(Rij + 'Vi'VjJ) e- f dl-l] - 8gij . (Rij + 'Vi'VjJ) e- f dl-l
=
['Vp(e-fgij8rfj)+e-f~(h- ~)+(R+~J)e-f(~ -h)]dl-l
- Vij . (Rij
+ 'Vi'VjJ) e- f dl-l.
Note that 8rfj is a tensor and we do not need an explicit formula for it in the rest of the proof. By the Divergence Theorem, we have
8(v,h)F(g,J)
1M 8 [(R+~J)e-fdl-l] = 1M (-~ (e- f ) + (R + ~J) e- f ) (~ - 1M Vij . (Rij + 'Vi'VjJ) e- f dl-l, =
h) dl-l
o
from which the lemma follows. REMARK 5.4. By (5.11), the variation of (~j vergence when h = ~:
+ 'Vi'VjJ) e- f dl-l
is a di-
8 [( ~j + 'Vi 'V jJ) e- f dl-l] = 'V p (e -f 8rfj) dl-l. Note also the factor ~ - h in front of the second term in the RHS of (5.10). The significance of when this factor vanishes will be seen in subsection 1.4 below. By (5.8) we have LEMMA 5.5. Define the measure
dm ~ e- f dl-l. If the variations of 9 and f keep the measure dm fixed. that is, 8(v,h) (dm) = 0, then (5.12)
V=2h.
As a consequence of Lemma 5.3, we have COROLLARY 5.6 (Measure-preserving first variation of F). For variations (v,h) with 8(v,h) (e-fdl-l) = 0, we have
(5.13)
8(v,h)F (g, J) = -
1M Vij (~j + 'Vi'VjJ) e- f dl-l.
Notice in formula (5.10) for 8(v,h)F (g, J) the occurrence of the terms
(5.14)
RiJ ~ (Rcm)ij ~ ~j
(5.15)
Rm ~ R
+ 'Vi'Vjf,
+ 2~f -I'V fl2 .
5. ENERGY, MONOTONICITY, AND BREATHERS
194
The first quantity vanishes on steady gradient solitons flowing along V f, whereas the second appeared in (5.4).5 We call RiJ and R m the modified Ricci curvature and modified scalar curvature, respectively; they are natural quantities from the perspective of the Ricci flow. We can rewrite
F(g,f)
= fM9ijRiJe-fdP, = fM Rme-fdp,
and when V = 2h. 1.3. The modified Ricci and scalar curvatures. In this subsection we digress by showing RiJ and Rm are natural quantities. Consider a closed Riemannian manifold (Mn, g) and a metric 9 = e- ~ f 9 conformal to g. Let ~j = Rc (g)ij' ~j = Rc (g)ij' fl = R (g), and R = R (g) . The Ricci and scalar curvatures are related by (see for example subsection 7.2 of Chapter 1 in [111] or (A.2) and (A.3) in this volume) (5.16)
Rij
= ~j +
(1-~) ViVjf + ~fl.f9ij + n -; 2VifVjf n n n
n -; 21Vfl2 gij' n
'fracing this yields
(5.17)
R = e~f (R + 2 (n - 1) fl.f _ (n - 1) (n - 2) n n2
IV f12)
.
The volume forms are related by dp,g = e- f dp, and the total scalar curvature of 9 is given by
fM Rdp,g
= fM e- n;;:2 f ( R + (n - l~;n - 2) IV f12) dp"
where we integrated by parts, i.e., we used
f e- n;;:2 f fl.fdp, = n - 2 f e- n;;:2 f IV fl2 dp,. 1M n 1M Now consider the Riemannian product (Mn, g) x (Tq, h q) ,where (Tq, h q) is a flat unit volume q-dimensional torus. The formulas for the Ricci curvature and scalar curvature of metric e-n!qf (g + h q') are given by (5.16) and (5.17), respectively, where we replace n by n + q. If we take the limit as q ~ 00 while fixing (Mn,g) , then we obtain Perelman's modified Ricci tensor: (5.18) and Perelman's modified scalar curvature: (5.19)
lim R (e - n!qf (g + h q )) = R
q-->oo
+ 2fl.f - IV fl2 ,
5Earlier we also encountered these quantities in Chapter 1.
1.
ENERGY, ITS FIRST VARIATION, AND THE GRADIENT FLOW
195
where we think of Rc (e- n!qf (g + h q)) and R (e - n!qf (g + hq)) as quantities on M since they are independent of the point in Tq. The total scalar curvatures of x Tq, e- n!qf (g + h q)) limit to Perelman's F functional:
(M
+ hq)) df-L
lim {
R (e - n!q f (g
=
{ R (e - n!qf (g
q--+oo lMxTq lim {
q--+oo 1M 1Tq
_ 2 f e ~ (g+hq)
+ hq)) e- 2f df-Lhqdf-Lg
= 1M (R + IV' f12) e- f df-L =F(g,f)· Note that
gijR7] = R+!::..f = R m -!::..f + lV'fI2. There is an analogue of the contracted second Bianchi identity for R7] and Rm. In particular we compute V'i R 7] and 1
2V'j R
m
=
= V'i~j + V'iV'jV'd =
~V'jR + V'j!::..f + Rjk V'kf
1 2 1 1 V'j!::..f - 2V'j lV'fl + 2V'j R = 2V'j R+ V'j!::..f - V'jV'kfV'kf,
which imply (5.20) To understand this formula further, we define
V'*m : Coo (T* M ®s T* M)
---t
Coo (T* M)
by
(V'*ma)j LEMMA
measure dm
~
V'iaij - ajiV'd.
5.7. The operator V'*m is the adjoint of -V' with respect to the = e- f df-L. -
PROOF. For any symmetric 2-tensor aij and I-form bi ,
1M aij (-V'i) bje- f df-L = 1M bjV'i (aije- f ) df-L
= 1M bj (V'iaij - aij V'd) e- f df-L = 1M bj (V'*ma)j e- f df-L.
o Thus (5.20) implies the following, which is the analogue of the contracted second Bianchi identity.
5.
196
ENERGY, MONOTONICITY, AND BREATHERS
LEMMA 5.8 (Modified contracted second Bianchi identity). (5.21)
1.4. The functional pTl and its gradient flow. Unlike F (g, 1) , we can obtain a functional of just the metric 9 by fixing a measure dm on a closed manifold Mn; by a measure we mean a positive n-form on M.6 Define pn : 9J1et --t ~ by (5.22) where (5.23)
j
.
=;=
log
(dJ.L dm ) .
'." REMARK 5.9. The expression (5.23) makes sense because, given a fixed measure dm on Mn, we can define the bijection
COO (A nT* M) --t Coo (M) , W I-t c.p, where c.p is defined so that W = c.pdm (here we have used the fact that AnT;M ~ ~). Thanks to this, it is possible to define the quotient of two n-forms; e.g., if WI = c.pldm and W2 = c.p2dm, where c.p2 > 0, then we set
WI ..:... c.pI W2 c.p2 Without using the notation j, we can write the energy of the metric 9 as
Using the modified Ricci and scalar curvatures, we can rewrite
r
(g)
=
1M gij RiJdm = 1M Rmdm.
REMARK 5.10. Let c.p : M --t M be a diffeomorphism. Note that in general r(c.p*g) 'Ir(g). That is, by fixing the measure dm, we get pn (g) , which breaks the diffeomorphism invariance of F (g, 1) . In subsection 3.1 of this chapter we shall solve this problem by considering a functional>' (g) which is diffeomorphisminvariant. 6For a calculational motivation for fixing the measure, see the notes and commentary at the end of this chapter.
2. MONOTONICITY OF ENERGY FOR THE RICCI FLOW
197
From (5.13) we have 8v :F"'" (g)
(5.24) where
=-
1M Vij (~j + '\l/VjJ) dm,
f is given by (5.23). The L2-inner product on 9J1et, using the metric
9 and the measure dm, is defined by
(aij,bij)m (g)
~ 1M (aij,bij)gdm.
Then by (5.24) we have
'\l:F"'" (g) = where is
f
-(~j
+ '\li'\ljJ),
is given by (5.23). Hence (twice) the positive gradient flow of Fm
(5.25) (5.26) We can also write the above system as (5.27)
~g at .. = ~J
-2
[R-. + '\l.'\l.J log (dJ.l)] dm' ~J
~
We shall call an equation of the form (5.25) by itself, for some function f, a modified Ricci flow. It is clear from taking Vij = -2 (~j + '\li'\ljJ) in (5.13) that we obtain the following. PROPOSITION 5.11 (P evolution under modified Ricci flow). Suppose 9 (t) is a solution of (5.25)-(5.26). Then (5.28)
:t:F"'"(g(t)) = 2
1M I~j + '\li'\ljfI2 e- f dJ.l.
This is Perelman's monotonicity formula for the gradient flow of P . We may rewrite (5.28) as
:t:F"'" =
:t 1M
Rmdm = 2
1M IR iJI2 dm.
Note that for a general measure dm, solutions to the initial-value problem for the gradient flow may not exist even for a short time; however, as we shall see, this will not cause us problems in applications.
2. Monotonicity of energy for the Ricci flow For monotonicity formula (5.28) to be useful, we need a corresponding version for solutions of the Ricci flow. In this section we show that solutions to equations (5.25) and (5.26), if they exist, differ from solutions of the Ricci flow by the pullback by time-dependent diffeomorphisms. Thus ~~is gives a . . .. monotonicity formula for the energy of the Ricci flow.
198
5.
ENERGY, MONOTONICITY, AND BREATHERS
2.1. A coupled system equivalent to the gradient flow of P. There is a coupled system, i.e., (5.29)-(5.30), induced from the gradient flow (5.25)-(5.26) obtained simply by computing the evolution equation for f = log (d/1,j dm) . As we shall see, this coupled system is equivalent to the gradient flow. LEMMA 5.12 (Measure-preserving evolution of f under modified RF). The function f (t) in a solution (g (t) , f (t)) of the gradient flow of P (5.25) and (5.26) satisfies the following equation: af at = -~f - R. PROOF. We calculate a ( -dJ.L ) -a f = -log at at dm
1 i' agij = _g3 i . = _g3_ (~. + 'V-'V-!).
at
2
3
t
3
o Related to the above calculation, we have the following. EXERCISE 5.13. Show that if forms, then
WI
~ log (WI)
(t) and
= %t WI
_
W2
(t) are time-dependent n-
%t W2 ,
at W2 WI w2 where the quotient of two n-forms is defined as in Remark 5.9. Hence we consider the coupled modified Ricci flow
a
(5.29)
at gij
(5.30)
af at
= -2(~j
=
-~f -
+ 'Vi'Vj!), R.
Note that the first equation is a modified Ricci flow equation whereas the second equation is a backward heat equation. LEMMA 5.14. The coupled modified Ricci flow equations (5.29)-(5.30) are equivalent to the gradient flow (5.27). PROOF. If g (t) is a solution to (5.27), then by Lemma 5.12, (g (t), f (t)) , where f = log (dJ.L/dm) , is a solution to the system (5.29)-(5.30). Conversely, if (g (t) , f (t)) is a solution to the system (5.29)-(5.30), then dm ~ e- f dJ.L satisfies
~ (dm) = (- af at at
- R - b. f ) e- f dJ.L
= o·
' that is, 9 (t) is a solution to (5.27) with dm as defined above. Hence, by (5.28), if (g (t), f (t)) is a solution to (5.29)-(5.30), then (5.31)
:t:F(g(t) , f (t))
=2
1M I~j + 'V i'V j fI2 e- f dJ.L.
o
2.
MONOTONICITY OF ENERGY FOR THE RICCI FLOW
199
2.2. Correspondence between solutions of the gradient flow and solutions of the Ricci flow. 2.2.1. Converting a solution of the gradient flow to a solution of Ricci flow. We first show that solutions of the gradient flow, if they exist, give rise to solutions of the Ricci flow with the same initial data (Lemma 5.15). In particular, suppose we have a solution (g (t) (t)) of the flow (5.25) and (5.26) on [0, T]; then we can obtain a solution g(t) of the Ricci flow on [0, T] by modifying 9 (t) by diffeomorphisms generated by the gradient of I (t) .
,I
LEMMA 5.15 (Perelman's coupling for Ricci flow). Let (g(t),/(t)) be a solution of (5.25) and (5.26) on [0, T]. We define a I-parameter family of diffeomorphisms W(t) : M ~ M by
d
-
dt w(t) = V' g(t)i(t) ,
(5.32)
(5.33)
w(o)
= idM .
Then the pullback metric g(t) = w(t)*g(t) and the dilaton f(t) = satisfy the following system:
ag = -2 Rc
( 5.34)
at at
af
(5.35)
= -I)"f
10 w(t)
'
+ IV' fl 2 -
R.
REMARK 5.16. Basically we can see this from the facts that LVI9 = 2V'V' f and Lv f f = IV' fl2 . For the sake of completeness we give the detailed calculations below. PROOF. First note that by Lemma 3.15 of Volume One the system of ODE (5.32)-(5.33) is always solvable. We compute
aa! = (a-) a! + w*
w* (LVgfg)
To obtain the equation for
= -2w* (Rc (g)) = -2Rc (g).
M, we compute
af = a(! w) = alow / (v f' w aw) at at at +\ 'at g 0
0
f)
= (-!il - Ii) = -I)"f - R
0
w+ I(vI) wi; 0
+ IV' fl2 ,
where barring a quantity indicates that it corresponds to 9 (t) .
0
So a solution to the gradient flow (5.25)-(5.26) yields a solution to the Ricci flow-backward heat equation system (5.34)-(5.35). Note that we can first solve the Ricci flow (5.34) forward in time and then solve (5.35) backward in time to get a solution of (5.34)-(5.35); this will be useful in applications.
200
5. ENERGY, MONOTONICITY, AND BREATHERS
2.2.2. Converting a solution of Ricci flow to a solution of the gr'adient flow. Now we show the converse of Lemma 5.15 by reversing the procedure ofthe last subsection. Given a solution 9 (t) of the Ricci flow (5.34) on [0, T], we can construct a solution (g(t), f(t)) of the gradient flow (5.25) and (5.26) on [0, T] by modifying the solution g(t) by diffeomorphisms. In doing so, we also need to solve a backward heat equation with initial data at time T.
tt
5.17. Let g(t) be a solution of the Ricci flow = -2 Rc on [0, T] and let h be a function on M. (i) We can solve the backward heat equation backwards in time LEMMA
of
at
= -~f
f{T) =
2
+ l\7fl -
t
R,
E
[0, T],
h·
(ii) Given a solution f (t) to the equation above, define the I-parameter family of diffeomorphisms cI>(t) : M ---t M by d dt cI>(t)
(5.36)
=
-\7g(t)f(t),
cI>(0) = idM,
which is a system of ODE and hence is solvable on [0, T]. 7 Then the pulledback metrics g(t) = cI>(t)*g(t) and the pulled-back dilaton J(t) = f 0 cI>(t) satisfy (5.29) and (5.30).
(i) Let 7 = T - t. To get the existence of solutions to equation (5.35), we simply set PROOF.
(5.37) and compute that (5.38)
au = ~u-Ru
-
07
'
which is a linear parabolic equation and has a solution on [0, T] with initial data at 7 = O. Indeed, (5.38) follows from
au au of ( 2) 07 = - at = u at = u -~f + 1\7fl - R = ~u -
Ru.
(ii) Let g(t) be a solution of the Ricci flow and let f(t) be a solution of equation (5.35). One can verify that they satisfy (5.29) and (5.30) as in the proof of Lemma 5.15. D 2.2.3. The adjoint heat equation. Let 9 (t) be a solution of Ricci flow and let D ~ %t - ~ be the heat operator acting on functions on M x [0, T] , where M x [0, T] is endowed with the volume form dp,dt. Its adjoint is
(5.39)
D*
==. -
~ at - ~ + R
7Again see Lemma 3.15 of Volume One.
2. MONOTONICITY OF ENERGY FOR THE RICCI FLOW
201
since
faT 1M bOadp,dt = faT 1M b (:t - ~ )
adp,dt
1M [a ( - :t - ~) bdp, - ab %t dP,] dt = faT 1M aO* bdp,dt
= faT
for C 2 functions a and bon M x [0, TJ with compact support in M x (0, T), where we used %t dp, = - Rdp,. By (5.38), if (g (t), f (t)) is a solution to (5.34)-(5.35), then u = e- f satisfies the adjoint heat equation (also known as the conjugate heat equation)
(5.40) It is often better to think in terms of u than in terms of f since u satisfies the adjoint heat equation. In particular, the fundamental solution to the adjoint heat equation is important. 2.3. Monotonicity of F for the Ricci flow. In this subsection we give two proofs of the monotonicity of energy for Ricci flow. In the next section we give an application of this formula to the nonexistence of nontrivial breather solutions. 2.3.1. Deriving the monotonicity of F from the mono tonicity of :pn. By the diffeomorphism invariance of all the quantities under consideration, the monotonicity formula for the gradient flow implies a monotonicity formula for the Ricci flow. This involves a function f (t) obtained by solving the backward heat equation (5.35). LEMMA 5.18 (F energy monotonicity). If (g (t), f(t)) is a solution to (5.34)-(5.35) on a closed manifold Mn, then
(5.41)
:t F (g (t) ,f(t)) = 2
1M IRij + \7i\7j fI2 e- f dp,.
PROOF. Since (g(t),f(t)) is a solution to (5.34)-(5.35), (g(t),](t)) , defined by g(t) ~ cp*(t)g(t) and J(t) = f(t) 0 cp(t), where cp(t) satisfies (5.36), is a solution to (5.29)-(5.30). Now F (g, f) = F (g, l) , so that by (5.31), we have
d d dt F (g (t) ,J (t)) = dt F (9 (t) , f-(t))
= 2 f litj + ViVj]l= e- f djl
1M
=2
9
1M I~j + \7i\7j fI2 e- f dp,.
202
5.
ENERGY, MONOTONICITY, AND BREATHERS
o 2.3.2. Deriving the monotonicity of F from a pointwise estimate. This second approach to the energy monotonicity formula is based on the pointwise formula (5.43), which is a simpler version of the evolution equation for Perelman's backward Harnack quantity (6.22). Let (g (t), f (t)) be a solution to (5.34)-(5.35). Let u = e- f and V ~ (2D..f -
(5.42)
IV fl2 + R)u = Rmu,
where R m is the modified scalar curvature defined by (5.15),8 so that F=
1M Vd/-L.
LEMMA 5.19 (Bochner-type formula for V). If (g (t), f (t)) is a solution to (5.34)-(5.35) and if u = e- f, then we have the pointwise differential equality: (5.43) This calculation, which we carry out below, is in a similar spirit to that of the calculations for the differential Harnack quantities considered in §10 of Chapter 5 in Volume One and Part II of this volume. To obtain (5.41) from the lemma, we compute
:t F (g (t) , f (t))
=
! 1M
V d/-L
1M (:t V - RV) d/-L = 1M 21 R ij + Vi V jfI 2ud/-L.
=
PROOF OF THE LEMMA. Using definition (5.42) and gij %tr~j = 0, a direct calculation shows that
~Rm = ~(2D..f - IV fl2 + R) 8t
8t
(M)
= 4~jViVjf + 2D.. 8t
- 2RijVdVjf - 2V
= 4RijViVjf - D..(2D..f -IV fl2
(M) 8t
.Vf
M + at
+ R) + D..IV fl2 + 2V D..f . V f
2 8R - 2~jVdVjf - 2V(IVfl - R)Vf + at
-
D..R.
From the above we have
(:t + D.) R
m
=
21~j + V i V j fl 2 + 2V R m . V f.
8The above V is not to be confused with our earlier V, which was the trace of the variation v of g.
3.
STEADY AND EXPANDING BREATHER SOLUTIONS REVISITED
203
On the other hand,
av at+~V-RV=
(8t+~Rm a R m ) u+ (au at +~u-Ru ) R m +2'VRm ·'Vu.
Plugging in the equation for
(gt
+~) Rm and using (5.40), we have
The last two terms cancel each other since 'V f lemma.
= - 'Vu/u,
which yields the 0
REMARK 5.20 (Backward heat-type equation for modified scalar curvature). From the proof of the lemma, we have (5.44) Note the similarity to the equation ~~ = ~R + 21Rcl 2 , except now we have a backward heat-type equation.
3. Steady and expanding breather solutions revisited A solution g(t) of the Ricci flow on a manifold Mn is called a Ricci breather if there exist times tl < t2, a constant QI > 0 and a diffeomorphism cp : M -+ M such that
When QI = 1, QI < 1, or QI > 1, we call g(t) a steady, shrinking, or expanding Ricci breather, respectively. Recall that g(t) is a Ricci soliton (or trivial Ricci breather) if for each pair of times tl < t2 there exist QI > 0 and a diffeomorphism cp : M -+ M (QI and cp will in general depend on tl and t2) such that g(t2) = Qlcp*g(tl). Note that if we consider the Ricci flow as a dynamical system on the space of Riemannian metrics modulo diffeomorphisms and homotheties, the Ricci breathers correspond to the periodic orbits whereas the Ricci solitons correspond to the fixed points. Since the Ricci flow is a heat-type equation, we expect that there are no periodic orbits except fixed points. A nice application of the energy monotonicity formula is the nonexistence of nontrivial steady or expanding breather solutions on closed manifolds (§2 of [297]). This was first proved by one of the authors in [218] (see Proposition 1.66 in this volume). In the next chapter we shall see the application of Perelman's entropy formula to prove shrinking breather solutions on closed manifolds are gradient Ricci solitons (§3 of [297]). Hence we confirm the above expectation.
5.
204
ENERGY, MONOTONICITY, AND BREATHERS
3.1. The infimum A of F. Suppose we have a steady breather solution to the Ricci flow with 9 (t2) = rp* 9 (tl) for some tl < t2 and diffeomorphism rp. One drawback of the energy monotonicity formula is that in general the solution f to (5.35) has f (t2) =1= f (td 0 rp, so that in general, F (g (t2)' f (t2)) =1= F (g (tl)' f (td)· By taking the infimum of F among f, we obtain an invariant of the Riemannian metric 9 which avoids this trouble. DEFINITION 5.21 (A-invariant). Given a metric 9 on a closed manifold
M n , we define the functional A : 9J1et --+ lR by (5.45)
Taking w = e- f /2, we have (5.46)
A(g) = inf {9(9, w) :
1M w dp, = 1, w > o} , 2
where, as in (5.5),9 (5.47)
Thus, when we fix 9 and minimize F (g, 1) among f, we are minimizing a Dirichlet-type functional and we get an eigenfunction-type equation for w. Aspects of this point of view are discussed in the next two lemmas. Note that the variation of 9 (g,') is given by
~O(0,h)9 (g, w) =
r 1M
r 1M
(4"'Vw·"'Vh + Rwh) dp, = (-4flw + Rw) hdp" 2 where h = ow. Hence the Euler-Lagrange equation for (note that we dropped the positivity condition on w) A (g)
~ inf { 9 (g, w) : 1M w 2 dp, = 1}
is (5.48)
Lw ~ -4flw
+ Rw = A (g) w.
LEMMA 5.22 (Existence and regularity of minimizer of Q). There exists a unique minimizer Wo (up to a change in sign) of (5.49)
inf { 9 (g, w) :
1M w dp, = 1} . 2
The minimizer Wo is positive and smooth. Moreover, 9In view of Lemma 5.1(1), the monotonicity of:F exhibits a dichotomy, it is analogous to both the monotonicty of the total scalar curvature under its gradient flow, g = - 2 (Rc - %g) , and the monotonicity of the Dirichlet energy under its gradient flow, the backward heat equation ftw = -~w. In this sense, the monotonicity of :F exhibits a beautiful synthesis of geometry and analysis.
it
3. STEADY AND EXPANDING BREATHER SOLUTIONS REVISITED
205
(1) the minimum value ),(g) of9(g,w) is equal to ),1(g), where),1 (g) is the lowest eigenvalue of the elliptic operator -4~ + R, and (2) Wo is the unique positive eigenfunction of (5.50)
-4~wo
+ Rwo =
),1
(g) Wo
with L2 -norm equal to 1. PROOF. To establish the existence of a minimizer Wo of (5.46), one takes a minimizing sequence {Wi}~1 of (5.46) in W 1,2 (M). There then exists a subsequence {Wi}~l which converges to Wo E W 1 ,2 (M) weakly in W 1,2 (M) and strongly in L2 (M) (by the Sobolev embedding theorem). Since
o ~ 1M 1\7 (Wi -
WO)12 dJL
= 1M I\7WiI 2 dJL + 1M l\7wol 2 dJL - 2 1M (\7Wi' \7wo) dJL, weak convergence in W 1,2, we have limi-too fM (\7wi' \7wo) dJL
by the fM l\7wol 2 dJL exists, hence
{ l\7wol 1M
2
dJL
~ li~inf ~-t00
{ l\7wil 1M
2 dJL.
On the other hand, by the strong convergence of {Wi}~1 in L2, we have .lim { Rw;dJL = { RW5 dJL ,
~->001M
1M
{ W5dJL = lim ( w;dJL = 1. 1M ~-t001M Hence Wo is a minimizer of (5.46) in W 1,2 (M) , and Wo is a weak solution to the eigenfunction equation (5.48). By standard regularity theory, Wo E Coo. We also have that any minimizer is either nonnegative or nonpositive, since otherwise ± Iwol is a distinct smooth minimizer which agrees with Wo on an open set, contradicting the unique continuation property of solutions to second-order linear elliptic equations. We now prove Wo is unique up to a sign. Without loss of generality, we may assume below that Wo is nonnegative. Call a minimizer w of 9 with fM w 2dJL = 1 a normalized minimizer. If the nonnegative normalized minimizer is not unique, then there exist two normalized minimizers Wo ~ 0 and WI ~ 0 with fM wow 1dJL = O. Then W2 = awo +bWI is also a normalized minimizer for all a, b E lR such that a2 + b2 = 1. Indeed, since Wo and WI satisfy the linear equation (5.50), so does W2 = awo+bwI, and fM w~dJL = 1. Now it not hard to see that there exist a and b such that W2 changes sign. In particular, if there are points x and y such that WI (x) = CWo (x) and WI (y) = dwo (y), where e =1= d and Wo (x) > 0 < Wo (y), then by choosing a and b with a2 + b2 = 1 such that a + bc and a + bd have opposite signs, we have that W2 (x) = (a + be) Wo (x) and W2 (y) = (a + bd) Wo (y) have opposite signs, which is a contradiction. Hence Wo is unique.
206
5.
ENERGY, MONOTONICITY, AND BREATHERS
Finally we show Wo > O. By the Hopf boundary point lemma (see Lemma 3.4 of Gilbarg and Trudinger [155]), if Wo = 0 somewhere, then there exists a point Xo E an, where 0. = {x EM: Wo (x) > O}, such that an satisfies the interior sphere condition at Xo, so that w (xo) = 0 and l\7w (xo)1 f:. 0, which is a contradiction to Wo ~ O. Finally, properties (1) and (2) follow easily. 0 The existence of a unique positive smooth minimizer Wo of g (g, w) under the constraint iM w 2dji = 1 implies the existence of a unique smooth minimizer fo of F(g, .) under the constraint iM e- f dji = 1. From (5.50) we see the following. LEMMA 5.23 (Euler-Lagrange equation for minimizer of F). The minimizer fo = -2 log Wo of F (g,.) is unique, Coo, and a solution to
,\ (g) = 2/).fo -1\7foI2
(5.51)
+ R.
That is, the modified scalar curvature is a constant, i.e., R m == ,\ (g) . Note that from setting v = 0 in (5.10), for the minimizer f of (5.45), we have
o(o,h)F (g, J)
= - 1M h (2/).f - 1\7 fl2 + R) e- f dji
for all h such that iM he- f djig = O. We can also obtain (5.51) directly from this. We summarize the properties of the functional ,\ on a closed manifold Mn.
(i) (Lower bound for ,\) '\(g) is well defined (i.e., finite) since
r
min R (x)· e- f dji = min R (x) ~ R min . xEM 1M xEM In particular, ,\ (g) ~ Rmin. (ii) (Diffeomorphism invariance) If cp : M --t M is a diffeomorphism, then '\(cp*g) = '\(g).
F(g, J)
~
(iii) (Existence of a smooth minimizer) There exists f E Coo (M) with iM e- f dji = 1 such that '\(g) = F(g, J)' i.e., (5.52)
'\(g)
= 1M (R + 1\7 fI2)e- f dji.
(iv) (Upper bound for ,\) We have
(5.53)
'\(g)
~ Vol ~M) 1M Rdji.
This can be seen by choosing
1M e-fdjig = 1 and '\(g)
f
= log Vol(M), which satisfies
~ 1M(R+ l\7fI 2)e- f dji.
3.
STEADY AND EXPANDING BREATHER SOLUTIONS REVISITED
207
(v) (Scaling) A (cg) =
C- 1A (g).
3.2. The monotonicity of A. Let (Mn,9(t» , t E [O,T], be a solution of the Ricci flow on a closed manifold. In this subsection we discuss some properties related to the continuity and monotonicity of A(g(t». Such properties are key to the proof of the nonexistence of nontrivial expanding or steady breathers. First we show that A(9(t» is a continuous function on [t1' t2]' This is a consequence of the following elementary result (see also Craioveanu, Puta, and Rassias [118] or Chapter XII of Reed and Simon [310]).10 LEMMA 5.24 (Effective estimate for continuous dependence of A on 9). If 91 and 92 are two metrics on M which satisfy 1 1 + c 91 ~ 92 ~ (1
+ c) 91
and
R (91) - c ~ R (92) ~ R (91)
+ c,
thenlI A (g2) - A (gl)
~
((1 + c)~+1 -
(1
+ c)-n/2) (1 + ct/2 (A (gl) -
+ ((1 + 8) max IRg2 -
Rgil
minRgJ
+ 28 max IRg11) (1 + ct/2 ,
where 8 ---t 0 as c ---t 0. 12 In particular, A : 9J1et with respect to the C2-topology.
---t
lR is a continuous function
PROOF. The proof is straightforward but slightly tedious. First note that (1 + c)-n/2 df..L9l ~ df..L92 ~ (1 + ct/ 2 df..Lgl' If w is a positive function on M, then in view of (5.46), we compute (writing a· b - c· d = a (b - d) + (a - c) d)
1M w2df..L9IQ(g2,W) - 1M w2df..L92Q(91 , w) = 4 1M w 2df..L9l (1M lV'wl~2 df..L92 - 1M lV'wl;l df..L9l) + 4 (1M w 2df..L9l - 1M W2df..L92) 1M lV'wl~l df..L9l + 1M w 2df..L9l (1M R92 w 2df..L92 - 1M R9l w 2df..L9l ) + (1M w 2df..L9l - 1M w 2df..L92) 1M Rgl w 2df..L9l , 10T hanks to
[231] for this last reference. lITo denote the dependence on gi, we use the subscript R (gI). 12See the proof for an explicit dependence of 8 on c.
gi
instead of (gi). So R91
=
208
5. ENERGY, MONOTONICITY, AND BREATHERS
so that
1M w 2d/L9l g (92, w) - 1M w 2d/L92 g (91, w) : :; 4((1 + E) ~+1 - 1) 1M w 2d/L9l 1M lV'wl;l d/L9l + 4 (1 - (1 + E)-n/2) 1M w2d/L91 1M lV'wl;l d/L9l + 1M w 2d/L9l 1M w 2 + 11M w 2 ( 1 -
(I
(R92 - R g1 )
~~:: I+ I(~~:: - 1) R9l I) d/L9l
~~::) d/L9l 111M R9l w 2d/L9l I·
(In the above estimates we took into account that R may change sign.) Let fJ ~ max {(I + Et/ 2 - 1,1- (1 + E)-n/2} , so that fJ ----t 0 as E ----t O. Since 1
1-
dJ.L92 dJ.L91
I
< fJ , we have
-
Hence
g (92, w)
g (91, w)
fM w 2d/L92
fM w 2d/L9l
:::; 4 ((1 + E)~+1 _ (1
+ ((1 + fJ) max IRg2
+ E)-n/2)
- Rgli
Taking w to be a minimizer for
fM IV'W~;l d/L9l fM W d/Lg2
+ 2fJmax IRg11) (1 + Et/ 2 .
g (91, .), we have
A (92) - A(91)
:::; 4 ((1
+ E)~+1 _
+ ((1 + fJ) max IRg2
(1
+ E)-n/2)
- R9lI
(1
+ Et/ 2 fM IV'W~;l d/L9l
+ 2fJ max IRgll)
fM W d/L9l (1 + Et/ 2 .
3.
STEADY AND EXPANDING BREATHER SOLUTIONS REVISITED
209
The result now follows from 4 fM
l'Vwl~l d/-Lgl = 9(91, w)
fM w 2d/-L9l ~
fM R9l w2d/-L9l fM w 2d/-L9l fM w 2d/-L9l A(91) - min Rgl .
o The monotonicity of F (9 (t) ,f (t)) under the system (5.34)-(5.35) implies the monotonicity of A (9 (t)) under the Ricci flow. LEMMA 5.25 (A monotonicity). If 9 (t), t E [0, T], is a solution to the
Ricci flow, then d 2 dtA(9(t)) ~ ;;:A 2(9(t)),
and A(9(t)) is nondecreasin9 in t E [0, T]. Here the derivative sense of the lim inf of backward difference quotients.
1t
is in the
REMARK 5.26. See the next subsection for the case where A(9(t)) is not strictly increasing. PROOF. Given to E [0, T], let fo be the minimizer of F (9 (to) , J), so that A (9 (to)) = F (9 (to), f (to))· Solve
a
2
atf=-R-/).f+I'Vfl,
(5.54)
f(to)=fo,
°
backward in time on [0, to]. Then itF (9 (t),j (t)) ~ for all t ~ to. Since the constraint fMe-fd/-L is preserved under (5.54), we have A(9(t)) ~ F(9(t),f(t)) for t ~ to. This, (5.41), and ,X (9 (to)) = F(9(to),f(to)) imply both (5.55)
A (9 (t)) ~ F (9 (t) , f (t)) ~ F (9 (to) , f (to)) = A (9 (to) )
and the following:
! (5.56)
I ~
A(9 (t) ) t=to
I
:t F (9 (t) , f (t)) t=to
= 21M IRij + 'Vi'Vj fI2 e- f d/-Lg(to)
~2
r
.!. (R + /).J)2 e- f d/-Lg(to)
JMn
~ ~ (1M (R + /).J) e- f d/-Lg(to)) 2 =
~A2(9(tO))' n
where f = fo is the minimizer. Hence, from either (5.55) or (5.56), we see that A (9 (t)) is nondecreasing under the Ricci flow. 0 EXERCISE 5.27. Prove (5.56).
5.
210
ENERGY, MONOTONICITY, AND BREATHERS
SOLUTION TO EXERCISE 5.27. We compute 13
~ >. (g (t) ) I dL
~ lim inf >. (g (to)) - >. (g (to - h))
t=to
h . f F (g (to) , fo) - F (g (to - h) , f (to - h))
h--+O+
> 1. -
lmlll
h--+O+
h
'
where fo is the minimizer for F (g (to),·) and f (t) is the solution to (5.54). On the other hand, we conclude by (5.41) that the last expression is equal to 2 fM I~j + "V/vjfoI2 e- fOdpg(to)· 3.3. There are no nontrivial steady breathers. As an application of the monotonicity of the diffeomorphism-invariant functional>' we prove the nonexistence of nontrivial steady breathers.
LEMMA 5.28 (No nontrivial steady breathers on closed manifolds). If (Mn,g(t)) is a solution to the Ricci flow on a closed manifold such that there exist tl < t2 with>' (g (tl)) = >. (g (t2)) , then 9 (t) is a steady gradient Ricci soliton, which must be Ricci flat. In particular, a steady Ricci breather on a closed manifold is Ricci flat. PROOF. Note that if 9 (t) is a steady Ricci breather with g(t2) = t.p*g(tl) for some tr < t2 and diffeomorphism t.p : M - 4 M, then >.(g (t2)) = >.(g (tl)). Hence we only need to prove the first part of the lemma. Suppose that for a solution 9 (t) to the Ricci flow there exist times tl < t2 such that >.(g (t2)) = >.(g (tl)). Let 12 be the minimizer for F at time t2 so that F (g (t2) , h) = >. (g (t2)). Take f (t) to be the solution to the backward heat equation (5.35) on the time interval [tl, t2J with the initial data f (t2) = 12. By the monotonicity formula (5.41) and the definition of >. we have 14 >. (g (tl)) :S F (g (tl), f (tl)) :S F (g (t), f (t)) :S F (g (t2) ,h)
= >. (g (t2))
for all t E [tr, t2J . Since>. (g (tl)) = >. (g (t2)) and>' (g (t)) is monotone, we have F (g (t), f (t)) = >. (g (t)) == const for t E [tr,t2J. Therefore the solution f(t) is the minimizer for F(g(t),·) and ftF (g (t), f (t)) == 0, so by (5.41) we have
1M IRij + "Vi"VjfI2 e- f dp (t) == 0 for all t E [tI, t2J . Thus (5.57)
~j
+ "Vi"Vjf =
0 for t E [tl, t2J.
In particular, g( t) is a steady gradient Ricci soliton flowing along "V f (t) l3Here d1_ denotes the lim inf of backward difference quotients. 14This is the same as (5.55). l5See (1.9), where a gradient soliton is steady if c = o.
.15
3.
STEADY AND EXPANDING BREATHER SOLUTIONS REVISITED
Note by (5.51) that
f satisfies the equation 21),.f -
On the other hand, R
211
1\7 fl2 + R = A (g) .
+ I),.f = 0, so that 1\7 fl2 + R = -A (g).
However, integrating, we have
-A (g) so that A (g)
= 1M (1\7 fl2 + R) e- f df.L = A (g),
=°and I),.f =1\7 fl2 =-R. Note that then 0= 1M (I),.f -
1\7 f12)
e f df.L = -21M
1\7 fl2 e f df.L
implies that f is constant and hence 9 is Ricci flat by (5.57). Alternatively, we could have argued that since I),.f = 1\7 fl2 2:: 0, f is subharmonic and hence constant. D REMARK 5.29. Even when M is noncompact, we have constant for gradient Ricci solitons; see Proposition 1.15.
1\7 fl2 + R
is
3.4. Nonexistence of nontrivial expanding breathers. Recall that A (g) is not scale-invariant, e.g., A (cg) = c- 1 A (g). Thus we define the normalized A-invariant:
.x (g) ~ A (g) . Vol (M)2/n . It is easy to see that .x (cg) = .x (g) for any c > 0, so the invariant .x is potentially useful for expanding and shrinking breathers. We shall prove the monotonicity of .x (g (t)) under Ricci flow when it is nonpositive. For (5.58)
this reason it is most useful for expanding breathers. Recall that by (5.56), we have
(5.59)
:t A (g (t)) 2:: 2 1M
I~j + \7i\7jfI2 e- f df.L,
where -9t A (g (t)) is defined as the lim inf of backward difference quotients. 16 Let V ~ V (t) ~ Volg(t) (M). From (5.59), we compute
!.x (g (t)) =! [A(9(t)). V (t)2/n] = V2/ndA + ~V~-lA dV dt
2:: 2v2/n 1M
n
dt
I~j + \7i\7jfI 2 e- fd f.L+ ~AV~-l 1M (__.R)df.L,
16This also applies to the time derivatives below in this argument.
212
5.
ENERGY, MONOTONICITY, AND BREATHERS
Hence
~ V- 2/ n ~ >.. (g (t)) ~ 1M I~j + \7i\7jf - ~ (R + ~f) gijl2 e- f dp, +
r ~(R+~f)2e-fdp,
JMn
- ~nJrM (R + ~f) e- f dp,· ~V JrM Rdp,. Recall from (5.53) that 1M (R
+ ~f) e- f dp, ::;
I
~dp,.
Assuming >'(t)::; 0, so that IM (R+~f)e-fdp,::; 0, we have
(5.60)
~V-2/n:t>"(g(t))- 1MI~j+\7i\7jf-~(R+Llf)9ijI2 e-fdp,
~ ~ 1M (R + ~f)2 e- f dp, - ~ (1M (R + ~f) e- f dp, ) 2 ~ 0 since IM e- f dp, = 1. Hence LEMMA 5.30. Let 9 (t) be a solution to the Ricci flow on a closed manifold Mn. If at some time t, >.. (t) ::; 0, then
(5.61)
d-
dt>.(g(t))
~ 2V2/ n 1M I~j + \7i\7j f - ~ (R + ~f) gijl2 e- f dp, ~ 0, where V = Volg(t) (M), f (t) is the minimizer for F (g (t),·), and the timederivative is defined as the lim inf of backward difference quotients. By (5.61), if -it>.. (g (t)) = 0, then 9 (t) is a gradient Ricci soliton. This is reminiscent of the fact that under the normalized Ricci flow, the minimum scalar curvature is nondecreasing as long as it is nonpositive, whereas under the un normalized Ricci flow, the minimum scalar curvature is always nondecreasing (see Lemma A.20). However these two facts appear to be quite different in nature. To apply the above monotonicity to the expanding breather case, we need to produce a time to where>.. (g (to)) < O. This is accomplished by looking at the evolution of the volume. Below we also give another proof of Lemma 5.28 using>" (g (t)).
3.
STEADY AND EXPANDING BREATHER SOLUTIONS REVISITED
213
LEMMA 5.31. Expanding or steady breathers on closed manifolds are
Einstein. PROOF. Let (Mn, 9 (t)) be an expanding or steady breather with g(t2) = acp*g(td for some tl < t2 and a 2 1. We have 5. (g (t2)) = 5. (g (tl))' Let V (t) ~ Volg(t) (M) . Since V (t2) 2 V (tt), we have for some to E (tI, t2),
o~
dd t
It=to log V (t) = - f~ ~~J.l (to) ~ -). (g (to)) . to
By Lemma 5.30, if 9 (t) is not a gradient Ricci soliton, then r1t5. (g (to)) > < 0 for some to < to. Now since 5. (g (t)) is increasing whenever it is negative, we have
oand we have 5. (g (to))
5. (g (t2)) = 5. (g (td)
~
5. (g (t~)) < 0,
which implies). (g (t)) ~ ). (g (t2)) < 0 for all t E [tl, t2]' Hence 5. (g (t)) is nondecreasing, which implies 5. (g (t)) is constant. By (5.61), we have 1 n
Rij + "\h\ljf - - (R +!:!..f) gij == 0, and since we are in the equality case of (5.60), we also have
(5.62)
R +!:!..f
=
Cl (t)
= const (depending on time).
That is, we still conclude that 9 (t) is an expanding or steady gradient Ricci soliton. Now let (Mn,g (t)) be an expanding or steady gradient Ricci soliton. Recall 2!:!..f + R - 1"\1 fl2 = C 2 (t) = const. This, combined with (5.62), implies
!:!..f - 1"\1 fl2 = const. Since
1M (!:!..f -
1"\1 f12) e- f dJ.l =
0,
1"\1 fl2 == O. Thus, by the strong maximum principle (or since now 0 = fM (!:!..f - 1"\1 f12) ef dJ.l = -2 fM 1"\1 fl2 ef dJ.l), we conclude that f == const. Hence ~j - ~R9ij == 0 and 9ij is Einstein. (When n = 2, our
we have !:!..f -
conclusion is vacuous.)
0
REMARK 5.32. As a corollary of the above result, we again see that expanding or steady solitons on closed manifolds are Einstein. In the case of shrinking solitons on closed manifolds, using the entropy functional, we shall see in the next chapter that they are necessarily gradient shrinking solitons.
214
5.
ENERGY, MONOTONICITY, AND BREATHERS
Note that on a shrinking breather we have V (t2) < V (tl) for t2 > tl' In particular, it is possible that A (g (t)) > 0 for all t E [tl, t2] (on the other hand, if A (g (to)) < 0 for some to E [tl' t2], then the proof above implies that a shrinking breather is Einstein), which causes difficulty in extending the proof above to the shrinking case; in the next chapter this problem is solved by the introduction of Perelman's entropy. (Note that for an Einstein manifold with R == r = const, under the constraint I e- f dp, = 1 we have
F (g, J) with equality if and only if .x (g) = r Vol (g)2/n > 0.)
=r+ f ==
1M IV' fl2 e- f dp, 2: r
log Vol (g)
=
const. Hence, if r
> 0, then
EXERCISE 5.33 (Behavior of .x on products). Compute.x of spheres and products of spheres. Show that .x (t) of a shrinking S2 x Sl under the Ricci flow approaches 00 as t approaches the singularity time. What happens if we start with S2 x S2, where the S2's have different radii? What is the behavior of .x for the product of Einstein spaces (or Ricci solitons)?
4. Classical entropy and Perelman's energy Define the classical entropy on a closed manifold Mn by
N
(5.63)
~ 1M fe- f dp, = -
1M ulogu dp"
where u ~ e- f . Under the gradient flow (5.29)-(5.30), we have
dN = dt
(5.64)
f afe-fdp,=_ f (R+t1f)e- f dp, 1M at 1M
= -F.
That is, the classical entropy is the anti-derivative of the negative of Perel-
man's energy. In this section we show that, by an upper bound for F, a modification of N is monotone. For comparison, we discuss Hamilton's original proof of surface entropy monotonicity, the entropy formula for Hamilton's surface entropy, the fact that the gradient of Hamilton's surface entropy is the matrix Harnack, and Bakry-Emery's logarithmic Sobolev-type inequality.
4.1. Monotonicity of the classical entropy. The following gives us an upper bound for the time interval of existence of the Ricci flow in terms of 1M dm and the initial value of P. Equivalently, it also implies the monotonicity of the classical entropy (see also [356], pp. 74-75). PROPOSITION 5.34 (Upper bound for F in terms of time to blow up).
Suppose that (g(t), f (t)) is a solution on a closed manifold Mn of the gradient flow for P, (5.25)-(5.26), for t E [0, T). Then we have (5.65)
r(g(O)) ::;
2~ 1M dm,
4. CLASSICAL ENTROPY AND PERELMAN'S ENERGY
215
that is, T <
f
n
- 2Fm(g(0) ) 1M
dm.
The proposition is a consequence of the following.
5.35 (Monotonicity formula for the classical entropy N). If (g (t), f (t)) , t E [0, T), is a solution of the gradient flow (5.29)-(5.30) on a closed manifold Mn, then LEMMA
d
2(f 1M dm
(g(t)) ~;;,
dt r
(5.66)
r(g(t))
~
)-1
r
(g(t))2,
2 (Tn_ t) !M e- f dJ.l.
By (5.64), this implies the following entropy monotonicity formula:
~ (N-
(5.67)
(i!Me-fdJ.l)log(T-t))
~O.
REMARK 5.36. Following §6.5 of [356], we may adjust the entropy quantity on the LHS of (5.67) by adding a constant and define
fj
~N-
(i !M e- f dJ.l) (log [47l' (T - t)]
+ 1).
Then we still have ~ ~ 0, whereas fj has the property that for a fundamental solution u = e- f limiting to a 8-function as t ~ T, we have fj ~ 0 as t ~ T. PROOF OF THE LEMMA.
ddt r
(g(t))
=2
From (5.28), we have
f I~j + \7i\7jfI2 dm ~ ~ f
n 1M
1M
?:
HL (R+b.f)dm),/ Ldm
=
~ (!M dm)
The solution of the ODE with limt--+T x (t)
(R + b.J)2 dm
= 00
-1
r
dx
(g(t))2 .
-=cx dt
2
is
1
x (t) Hence, taking c = ~
(JM r
= c (T - t)'
dm) -1 , we get (g(t))
~ 2 (Tn_ t)!M dm. o
216
5. ENERGY, MONOTONICITY, AND BREATHERS REMARK
5.37.
diFm
is somewhat reminiscent of Hamilton's (1) The formula above for formula for the evolution of the time-derivative dN/ dt of his entropy N (g) ~ f M2 R log RdA on a positively curved surface evolving by Ricci flow (see [180]). (2) We can rewrite (5.66) as
1M (2tJ.f -1\7 fl2 + R where
T ~
T -
;:.) e- f dp, :::; 0,
t.
4.2. Hamilton's surface entropy. Recall that the normalized surface entropy for a closed surface (M2, g) with positive curvature is defined by N(g)
~ 1M log (RA) Rdp"
where A is the area. Let (M2, 9 (t)) , t E [0, T), be a solution, on a maximal time interval of existence, of the Ricci flow on a closed surface with R > 0. In this subsection we give two proofs of the monotonicity of N (g (t)) ~ N (t) . 4.2.1. Hamilton's original proof of surface entropy monotonicity. The time-derivative of N (t) is given by dN dt =
(5.68)
f 1M QRdp"
where
Q ~ tJ.log R + R -
r
and r is the average scalar curvature. On the other hand, since ~~ = r2 and by a similar computation to (Vl-5.38),
a
at Q 2: tJ.Q + 2 (\7 log R, \7Q)
+ Q2 + 2rQ.
By the long-time existence theorem (Proposition 5.19 of Volume One), r = T~t and Area (g (t)) = 47rX (T - t), where X denotes the Euler characteristic of M. Differentiating (5.68) with respect to time, we compute that Z ~ satisfies
dI:
~~ = 1M (:t Q )
Rdp, +
1M Q :t (Rdp,)
2:
1M (Q2 + 2rQ) Rdp,
2:
fM ~dP, (1M QRdP,)
= _1_Z2+ 47rX
_2_Z,
T- t
2
+ 2rZ
4.
CLASSICAL ENTROPY AND PERELMAN'S ENERGY
217
where we integrated by parts and used Holder's inequality and the GaussBonnet formula. Thus
~ ( s -2Z) -4 > _1_ ( s -2Z)2 , ds 7l"X
(5.69)
where s ~ T~t. From this we conclude that if s02 Z (so) > 0 for some So < 00, then s-2 Z (s) ---+ 00 as s ---+ Sl for some Sl < 00. In other words, if Z (to) > 0 for some to < T, then Z (t) ---+ 00 as t ---+ t1 for some t1 < T. This contradicts our assumption that the solution exists on [0, T). Hence Z (t) ::; 0 for all t and we have proved the following. THEOREM 5.38 (Hamilton's surface entropy monotonicity). For a solution of the Ricci flow on a closed surface with R > 0, we have dN (t) < 0 dt for all t E [0, T).
Note that, from (5.69), we have t ~ (T - t)2 Z (t) is nondecreasing (since X > 0) and hence there is a constant C > 0 such that (T - t)2 Z (t) 2: -C for all t E [0, T). By (5.68), Z= dN dt
=_ { IVRI 2 dA+ { (R-r)2dA,
1M
R
1M
and we have
REMARK 5.39. An inequality of the above type is often referred to as a reverse Poincare inequality. 4.2.2. Entropy formula for Hamilton's surface entropy. Define the potential function f (up to an additive constant) by !:l.f = r - R. In [97J the monotonicity of the entropy was proved by relating its time-derivative to Ricci solitons via an integration by parts using the potential function (Proposition 5.39 in Volume One). In particular, we have
Note that Rij = ~Rgij and r = ~ 1M RdA = (T - t)-l. We have purposely written this formula to more resemble Perelman's formulas (5.41) and (6.17).
218
5.
ENERGY, MONOTONICITY, AND BREATHERS
4.2.3. The gradient of Hamilton's surface entropy is the matrix Harnack quantity. A less well-known fact is that the gradient of Hamilton's entropy in the space of all metrics with the L 2-metric is the matrix Harnack quantity: (5.70)
8v N (g)
=
1M Vij ( -~ log R· gij + ViVj log R -
~R9ij) dA,
where 8g = v (see Lemma 10.23 of [111] and use N (g) - E (g) is a constant). In the space of metrics in a fixed conformal class, the gradient is the trace Harnack quantity. Note that the same relation is true relating the entropy and the trace Harnack quantity for the Gauss curvature flow of convex hypersurfaces in Euclidean space [96]. 4.3. Bakry-Emery's logarithmic Sobolev-type inequality. The proofs of Hamilton's surface entropy formula and Perelman's energy formulas are formally similar to the proof of Bakry and Emery of their logarithmic Sobolev-type inequality [18].
PROPOSITION 5.40. Let (M n , g) be a closed Riemannian manifold with Rc 2:: K for some constant K > O. If u is a positive function on M, then
1M ulogudJ.L::; 2~ 1M u IVlogul 2 dJ.L + log (VOI~M) 1M UdJ.L) 1M udJ.L. PROOF. (See [104] for more details of the computations.) Consider the solution v to the heat equation aa:. = ~v with v (0) = u. The solution v exists for all time and
t~~
V
= Vol ~M)
1M udJ.L.
J
Define E (t) ~ M v log vdJ.L. Then
(5.71)
t~~ E (t) = 1M udJ.L . log ( Vol ~M) 1M UdJ.L) .
We have
~~ = -
1M (Vv, V log v) dJ.L = - 1M v IV log vl
2
dJ.L ::; O.
Note that limt ..... oo ~f (t) = O. Using %t log v = ~logv + IVlogvl2, we compute
ddt2 E2 = 2 1M f v ( IVVlog vl 2
+ Rc (V log v, Vlogv) ) dJ.L.
Using our assumption Rc 2:: K, we find
d2 E dE -d >-2K2 dt . t -
5. NOTES AND COMMENTARY
By limt-+oo ~f (t)
dE
219
= 0 and (5.71), we have
roo d2E
roo dE
= 2Klog ( Vol ~M)
1M UdJ.t) 1M udJ.t -
dt (0) = -
Jo
dt2 (t) dt ~ 2K Jo
dt (t) dt 2KE (0).
Hence
- 1M u IV' log ul 2 dJ.t ~ 2K log ( Vol ~M) 1M UdJ.t) 1M udJ.t - 2K 1M u log udJ.t and the proposition follows.
D
5. Notes and commentary Subsection 1.1. As we remarked earlier, the function 1 is also known as the dilatonj in the physics literature there are numerous references to Perelman's energy functional (see Green, Schwarz, and Witten [162]' Polchinksi [307]' Strominger and Vafa [341] for example), although Perelman is the first to consider it in the context of Ricci flow. The Ricci flow is the 1-loop approximation of the renormalization group flow (see Friedan [145]). Subsection 1.2. For a computational motivation for fixing the measure, see also §4 in Chapter 2 of [111], where Perelman's functional is motivated starting from the total scalar curvature functional. In particular, let 89 = v. The variation of the total scalar curvature is
8
1M RdJ.t = 1M (diV (divv) - ~v - Rc·v + R ~) dJ.t = - 1M (Rc - ~ 9) .vdJ.t.
This says that V' (JM RdJ.t) = - Rc +~ 9, where the gradient is calculated with respect to the standard L2- metric. To try to find a functional F with V'F = - Rc, we want to get rid of the ~9 term. Now this term is due to the variation of dJ.t. So we consider the distorted volume form e- f dJ.t and assume its variation is O. Hence
8
1M Re- f dJ.t = 1M (8R) e- f dJ.t = 1M (div (divv) -
and now we have the extra terms sate for this by considering
8
JM
~v -
Rc ·v) e- f dJ.t
(div (div v) - ~ V) e- f dJ.t. We compen-
1M 1V'112 e- f dJ.t = 1M (8 IV'112) e- f dJ.t = 1M (-v (V'I, V' f) + V'1 . V'V) e- f dJ.t,
220 using
5. ENERGY, MONOTONICITY, AND BREATHERS
¥- =
8F = 8
h ~ 8f. Integrating by parts yields
1M Re- f d/-L + 8 1M 1V'112 e- f d/-L = - 1M Vij (~j + V'iV'jf) e- f d/-L.
Although the Ricci tensor is not strictly elliptic in g, one can ask if the RHS of equation (5.27)
:t
gij = -2
[~j + V'iV'jlog (:~)]
is elliptic in g. The answer is still no. In particular, if 8g
= v,
then
Hence
Since we have 8
(-2
[Rij
+ V'iV'jlog (:~)])
= D.vij - V'iV'kVjk - V'jV'kVik
+
lower-order terms,
where the last pair of terms form a Lie derivative of the metric term. However the second-order operator on the RHS is still not elliptic in v.
CHAPTER 6
Entropy and No Local Collapsing Everything should be made as simple as possible, but not simpler. - Albert Einstein Disorder increases with time because we measure time in the direction in which disorder increases. - Stephen Hawking Close, but no cigar. - Unknown origin
By combining Perelman's energy and the classical entropy in a suitable way, we obtain the entropy functional W, which we shall discuss in this chapter. This is implemented with the introduction of a positive scalefactor T. The advantage which the addition of this scale-factor yields is that from the functional W we can understand aspects of the local geometry of the manifold, e.g., volume ratios of balls with radius on the order of ..ji. Perelman's entropy is also the integral of his Harnack quantity for fundamental solutions of the adjoint heat equation. 1 As such, one can integrate in space the Harnack partial differential inequality to give a proof of the monotonicity formula for W. Note that here Perelman's Harnack quantity is directly related to the entropy whereas in Hamilton's earlier work on surfaces, Hamilton's Harnack quantity is related to the time-derivative of his entropy. 2 This monotonicity formula can be used to prove that shrinking breathers must be shrinking gradient Ricci solitons. More importantly, this monotonicity formula will be fundamental in proving Hamilton's little loop conjecture or what Perelman calls the no local collapsing theorem. In this chapter, we shall discuss in detail the entropy estimates, the two functionals J1 and 1/ associated to W, and their geometric applications. We discuss the logarithmic Sobolev inequality, which is related to the entropy functional. We will also give different versions/proofs of the no local collapsing theorem. In the last part of this chapter we shall discuss some interesting calculations related to entropy. Throughout this chapter Mn denotes a closed n-dimensional manifold. 1. The entropy functional W and its monotonicity
Let (Mn,g(t)) , t E [O,T], be a solution of the Ricci flow on a closed manifold. Note that by the proof of Lemma 5.31, we have that when lSee (6.21). 2Equation (6.21) as compared to (5.70). 221
222
6. ENTROPY AND NO LOCAL COLLAPSING
>.(g(t)) ~ 0 for some t E [tl, t2J, even shrinking breathers, i.e., solutions with 9 (t2) = acp*g (tl) and a < 1, are Einstein solutions (trivial Ricci solitons). In order to handle shrinking breathers when >.(g(t)) > 0 for all t E [tl' t2], we need to generalize the monotonicity formula for the functional F to a monotonicity formula for a functional related to shrinking breathers. This is the monotonicity formula for the entropy W. In this section we introduce the entropy Wand discuss its monotonicity. We also give a unified treatment of energy and entropy in the last subsection. 1.1. The entropy W, its first variation and the gradient flow. 1.1.1. The entropy W. Let 9Jtet denote the space of smooth Riemannian metrics on a closed manifold Mn. We define Perelman's entropy functional W : 9Jtet x COO(M) x jR+ --+ jR by
W(g, 1, T)
(6.1)
(6.2)
~ 1M [T (R + 1\7112) + 1 =
n] (47rT)-n/2 e- f d/-l
1M [T (R + 1\7112) + 1 - n] ud/-l,
where 3 (6.3) This is a modification of the energy functional F (g, j) , which we considered in the last chapter, where we have now introduced the positive parameter T. By (5.1), we have
(6.4)
W(g,j, T) = (47rT)-n/2 (TF (g, j)
+ 1M (f -
n) e- f d/-l)
= (47rT)-n/2 (TF(g,j) +N(f)) - n
(6.5)
1M ud/-l,
where the second equality is obtained using definition (5.63). As we shall see, T plays the dual roles of understanding the geometry of (M, g) at the distance scale -IT and representing a constant minus time for solutions (M,g (t)) of the Ricci flow. 4 The functional W has the following elementary properties. (i) (Scale invariance) W is invariant under the scalings T f--t CT and 9 f--t cg, i.e.,
W(cg, 1, CT) = W(g, 1, T).
(6.6)
(ii) (Diffeomorphism invariance) If cP : M
--+
M is a diffeomorphism,
then
W(g, 1, T) = W(cp*g, cP* 1, T), where cp*g is the pulled-back metric and cp* 1 = 1 0 CPo 3This u is not to be confused with the u = e- f in Chapter 5. 4We may think of T as physically representing temperature (see §5 of [297]).
1.
THE ENTROPY FUNCTIONAL W AND ITS MONOTONICITY
1.1.2. The first variation of W. Let 8g 8f = h, and let 87 = (. Since
(6.7)
8 (udp,)
= v E C2(M, T* M
223
® T* M), let
= 8 ((47r7)-n/2 e- f dp,) = ( -;:. (- h + ~)
udp"
where V ~ trg v = gi j vij , the measure (47r7)-n/2e- f dp, preserving variations satisfy
V
n
(6.8)
--(-h+ - =0. 27 2
We find it convenient to write the variation of W so that this quantity is one of the factors. In particular, we have LEMMA 6.1 (Entropy first variation formula). The first variation of W at (g, f, 7) can be expressed as follows: 8W(v,h,() (g, f, 7)
(6.9)
= 1M (- 7Vij + (gij) (~j + \/i\/jf -
(~ -
+ 1M 7
h - ;;) (R
+ 2~f -
2~9ij)udp, 1\ /
fl2
+
f - ; -
1)
udp,.
PROOF. It follows from the first variation formula (5.10) of F with respect to 9 and f (keeping 7 fixed) that 8( v,h,O) (7( 47r7) -n/2 F (g, f) )
=-
1M 7Vij(~j
+ 1M 7
(~ -
+ \/i\/jf)udp,
h)
(2~f -1\/fI2 + R) udp,.
Next we calculate the first variation of the remaining term of W with respect to 9 and f (again keeping 7 fixed), 8(v,h,O) ((47r7)-n/21M (f-n)e-fdP,)
= 1M
[(1+n-f)h+
~(f-n)] udp,.
Now the term from the variation of W with respect to 8(o,o,() (1M [7(R +
= 1M
[( 1-
1\/ f12) + f
7
is
- n] (47r7)-n/2e- f dp, )
~) ((R + l\/fI2) -
;; (f - n)] udp,.
6. ENTROPY AND NO LOCAL COLLAPSING
224
Combining the above three formulas and simplifying a little, we get 8(v,h,() W
(g, f, 7)
= - 1M 7Vij(~j + V/Vjf)udJ.L + 1M 7
(~
+ 1M [h +
-
h) (2b. f - IV fl2 + R + f
(1 - ~) ((R + IV f12) -
~ n) udJ.L
;; (f - n)] udJ.L.
We rewrite the above expression as 8(v,h,() W
(g, f, 7)
= 1M (- 7Vij
+ (gij) (Rij + Vi Vj f)udJ.L
+ 1M7(~ -h-;;)
(2b.f-IVfI2+R+f~n)udJ.L
+ 1M -((R + b.f)udJ.L + 1M + 1M ( 1 -
~( (2b.f -IV fl2 + R) udJ.L
~) ((R + IV fI2)udJ.L + 1M hudJ.L,
which, by combining terms, we further simplify to 8(v,h,() W
(g, f, 7)
= 1M (- 7Vij + (gij) (Rij + ViVjf)udJ.L + 1M7(~ -h- ;;)
(2b.f-IVfI2+R+f~n)udJ.L
+ 1M (n - 1) ( (b.f -IV f12) udJ.L + 1M hudJ.L. Since ( is a constant, (6.9) follows from a rearrangement and the integration by parts identity: fM (b.f REMARK
Rij
-IV f12) e- f dJ.L = O.
0
6.2. Analogous to (5.14) and (5.15), the terms
+ ViVjf -
1 -2 gij 7
and
R + 2b.f -IVfl
2
f-n
+-7
in (6.9) are natural quantities vanishing/constant on shrinking gradient Ricci solitons. 1.1.3. The gradient flow of W. When we require that the variation (v, h, () satisfies V n ( = -1 and '2 - h - 27 ( = 0,
1. THE ENTROPY FUNCTIONAL W AND ITS MONOTONICITY
225
i.e., the variation preserves the measure (41l"7)-n/2 e- f dp, on M, we obtain from (6.9) the gradient flow (assuming ~; = -1)
a
at gij = -2 {~j + '\h\1jf) ,
(6.10) (6.11)
af at
(6.12)
~: = -1.
=
-b..f - R
n
+ 27'
LEMMA 6.3 (Entropy monotonicity for gradient flow). If (g{ t), f (t), 7{ t)) is a solution to the system (6.10)-{6.12), then
d dt W{g{t), f{t), 7{t))
=
(6.13) PROOF.
1M 27 I~j + '\h\1jf - 2~9ij12 udp, 2: O.
This follows directly from substituting Vij = -2 {~j h
=
+ '\h\1jf) , n
-b..f - R + 27'
(= -1
into (6.9) and using the facts
~ -h- 2:(=0
and
1M (b..f-I V fI 2)e-f dp,=0. D
1.2. Coupled evolution equations associated to Wand monotonicity of W. 1.2.1. The coupled evolution equations associated to W. As in Chapter 5, there is a system of evolution equations for the triple (g, f, 7) (see (1.3) of [297]) (6.14)
(6.15)
a
at gij = af
-2~j, 2
at = -b..f + IVfl - R+
n
27'
d7 =-1 dt ' whose solution differs from the solution to (6.10)-{6.12) by diffeomorphisms. This leads to the following theorem, which says that (6.16)
d dt W{g{t), f{t), 7{t)) 2: O. Another motivation for studying this system of equations, from considering gradient Ricci solitons, was discussed in Section 8 of Chapter 1.
226
6.
ENTROPY AND NO LOCAL COLLAPSING
THEOREM 6.4 (Entropy monotonicity for Ricci flow). Let (g(t),J(t),7(t)), t E [0, T], be a solution of the modified evolution equations (6.14), (6.15), and (6.16). Then the first variation of W along this solution is given by the following: d dt W(g(t), f(t), 7(t))
(6.17) (6.18)
1M 27 I~j + 'Vi'Vjf - 2~9ij12 . udJ.L = 27 1M I~j ~ ('Vi'Vju - 'Viu:jU + 2~ U9ij ) 12 . udJ.L 2 O. =
-
EXERCISE 6.5. Show that the line of reasoning of deriving Theorem 6.4 from Lemma 6.3 is rigorous.
(6.19)
is exactly the matrix Harnack quantity for a solution U of the backward heat equation. In particular, this same expression for positive solutions of the backward heat equation appeared in Hamilton's derivation of the monotonicity formulae for the harmonic map heat flow, mean curvature flow, as well as the Yang-Mills flow; see [184].5 Equation (6.17) is Perelman's entropy monotonicity formula and implies that W(g(t), f(t), 7(t)) is strictly increasing along a solution of the modified coupled flow except when g(t) is a shrinking gradient Ricci soliton (since 7> 0 in (6.17)), where it must flow along 'Vf and where W(g(t),f(t),7(t)) is constant. This monotonicity is also fundamental in understanding the local geometry of the solution 9 (t) to the Ricci flow as we shall see in the proof of the no local collapsing theorem. The function f allows us to localize and the parameter 7 tells us at what distance scale to localize (yfT) .. Since we do not give the details of how to transform between the systems (6.10)-(6.12) and (6.14)-(6.16), we also compute (6.17) directly through the following exercise. EXERCISE 6.7 (Deriving the formula for ~~.
dJi from ft)· Use the equation for ft to derive
SOLUTION TO EXERCISE 6.7. The effect of the extra term as compared to (5.30) is to add
+~
in (6.11)
-;:. 1M (R+ l'VfI2) e-fdJ.L 5For an exposition of the matrix Harnack estimate asssociated to (6.19), see [104J.
1.
THE ENTROPY FUNCTIONAL
to (5.31), so that we get
dF
dt
W
AND ITS MONOTONICITY
227
r IRj + 'Yi'Yjfl 21 n e- dJ.L - 27 F .
= 2 1M
Similarly, (5.64) becomes
dN = -F +!:. dt
r e- I dJ.L -
271M
!:.N 27
under (6.10)-(6.11). Hence, by (6.5), i.e., W =(47r7)-n/2 (7F(g, f) and iM udJ.L
+N) - n 1M udJ.L
= const, we have
d: = :t [(47r7)-n/2 (7F +N)] = (47r7)-n/2
(!:. (7F +N) _ F + 7 dF + dN) 27
dt
dt
'
so that
(47r7t/2d: = 271M IRj + 'Yi'YjfI2 e- I dJ.L - 2F + ;:. 1M e- I dJ.L
= 271M IRi j + 'Yi'Yjf -
2~9ij12 e- I dJ.L.
1.2.2. Second proof of the monotonicity ofW from a pointwise estimate. Analogously to subsection 2.3.2 of Chapter 5 we again derive (6.17) using a pointwise evolution formula. Let 6 (6.20)
In Part II of this volume we shall see that v is nonpositive when u is a fundamental solution (for this reason v is also called Perelman '8 Harnack quantity). Note that (6.21)
W (g, f, 7)
= 1M vdJ.L.
We shall show the following below. LEMMA 6.8 (Perelman's Harnack quantity satisfies adjoint heat-type equation). Under (6.14)-(6.16) (6.22)
where 0*
O*v
= -27 I~j + 'Yi'Yjf -
2~9ij12 u,
= -It - Ll + R is the adjoint heat operator defined in (5.39).
6This quantity v is not to be confused with v = 8g.
228
6.
ENTROPY AND NO LOCAL COLLAPSING
Theorem 6.4 then follows from
dW dt
= { (~_ R) vd/-L = { (-0* - 1:1) vd/-L
1M
= 2T ~ o.
1M
at
1M I~j + "h'Vjf - 2~9ij12 ud/-L
Before we prove (6.22), we need the following lemma concerning the function u = (47rT)-n/2 e- f as defined in (6.3) (compare with (5.40)). LEMMA 6.9 (u is a solution to the adjoint heat equation). The evolution equation (6.15) of f is equivalent to the following evolution equation of u : (6.23)
O*u
= o.
PROOF. We calculate
o We now show that we can apply the computation in subsection 2.3.2 of Chapter 5 to derive the evolution equation (6.22) for v. PROOF OF (6.22). Let Ibe defined by e- i ~ u. Then 1= f+~ log (47rT) and (g,1) satisfies (5.34) and (5.35). By (5.43), we have that the quantity
V~ (R + 21:11 -IVI12) u = (-21:1 log u -IVlogul 2 + R) u satisfies the equation
Note that
v= [T (R + 21:11 -Iv112) -logu (6.24)
~ log(47rT) -
= TV - (logu + ~ log(47rT) + n) u.
We compute using O*u 0* (ab)
= 0 and the general formula = bO*a + aO*b -
2 (Va, Vb) - Rab
n] u
1. THE ENTROPY FUNCTIONAL W AND ITS MONOTONICITY
229
that 0* [(lOgU + ~ log(47f7)
+ n) U] n
= uO* log u + log uO*u - 2 (V log u, Vu) - Ru log u + 27 U n
2
= -Ru-IVlogul U+ 27U, since
0* logu =
~(-
:t -~) + U
IVlogul 2 + Rlogu
= -R + IVlogul 2 + Rlogu. From (6.24), we compute O*v = 70*V + V
+ IVlogul 2 u + Ru -
n
27 U
= -271~j - ViVj logul 2 u
+ (-2~logu -IVlogul 2 + R) u 2
+ IV log ul u + Ru -
n
27 U 2
= -271~j - ViVj logul u
+ 2 (-~logu + R) u -
n 27 u.
o
Completing the square, we obtain (6.22).
1.3. A unified treatment of energy F and entropy W. We finish this section with several exercises. In total, this unifies part of the discussion of shrinking, steady, and expanding gradient Ricci solitons, which correspond to entropy, energy, and expander entropy on a closed manifold Mn (see the definition below), respectively.7 In this subsection we use the following convention: c E JR, and if c i= 0, we take 7 (t) = ct, whereas if c = 0, we take 7 (t) = 1. We only consider all t E JR such that 7 (t) > 0. Let (6.25) Define the c-entropy by
(6.26)
W€(g, f, 7)
~ 1M (7 (R + IV f12) =
c (f - n)) (47f7)-n/2 e- f dj.t
1M Veudj.t,
7Notation: For the most part we use a hat on M to emphasize that we are considering a fixed metric instead of a solution to the Ricci flow.
230
6. ENTROPY AND NO LOCAL COLLAPSING
where u is defined in (6.3). When c < 0, this is Perelman's entropy; when c = 0, this is Perelman's energy; and when c > 0, this is called the expander entropy. 8 The definition of Vc is motivated by the following exercise. EXERCISE 6.10 (Harnack Vc as an integrand for We). Let (Mn,g (t),J (t)) be a gradient Ricci soliton in canonical form. By Proposition 1. 7, the pair (g (t), I (t)) satisfies
(6.27)
Sfj
~ ~j + V/vjl + 2Cr gij
= 0,
where c E R Note that if c ::f. 0, then 9 (t) and I (t) are defined for all t such that r (t) > 0, whereas if c = 0, then 9 (t) and I (t) are defined for all t E (-00, (0) . Show that Vc (g (t), I (t), r (t)) is constant in space. 6.11 (We and the Gaussian soliton). Consider the Gaussian soliton (lRn,!}JE, Ie), where!}JE = L:~=1 (dxi)2 and EXERCISE
Ie (x, t)
~
1
_1~~2
°
Ixl2 -""4t
for t >
°
if c
< 0,
for t E lR if c = 0, for t <
°
if c
> 0.
Check that
8 2 Ie
c
8 x~'8 x J. + -dij 2r =
°
for all t such that r (t) = ct > 0, and thus S[j == 0. Show Vc = 0, so that We = 0. It is useful to keep this example in mind, which reflects the Euclidean heat kernel, when considering the function theory aspects of the material in this chapter and especially the chapter on Perelman's differential Harnack estimate in the second part of this volume. EXERCISE
6.12 (Basic properties of We). Show that on a closed Rie-
mannian manifold
(Mn, g) ,
(1)
We(g, I, r) = (47f)-n/2 where 9 ~ r- 1 g, (2) for any constant c>
1M (Rg + IVII~ - c (f - n)) e- f djLg,
°
We (cg, I, cr) = We(g, I, r), and (3) for any diffeomorphism cp : M ~
We (cp*g, 1 0 cp, r)
=
M, We(g,
I, r).
8These identifications are true up to constant factors.
1.
THE ENTROPY FUNCTIONAL
AND ITS MONOTONICITY
W
231
EXERCISE 6.13 (First variation formula for We.). Show that, on a closed manifold Mn, if
8g
(6.28)
= v,
8j
= h,
=(
8r
at (g, j, r), then (6.29)
{ ((-rVi j + (gij) Sfj ) 8(v,h,() We;(g, j, r) = 1M + (~ _ h _ (V~ + c) udJ-l,
¥f)
where Sfj is defined by (6.27). In particular, if
then
8(v,h,()We;(g,j,r) =
(6.30)
1M (-rVij + (gij)SfjudJ-l.
SOLUTION TO EXERCISE 6.13. We compute
8v~j = V'p (8rfj)
- V'i
(8r~j)
and
8(v,h) (V'iV'jf) = V'iV'j (8f) - (8rfj) V'pJ. Adding these two formulas together, we get
8(v,h) (Rij
+ V'iV'jf)
= V'p (8rfj) - (8rfj) V'pj + V'i (V'j (8f) - 8r~j) (6.31)
=efV'p(e-f8rfj)+V'iV'j(h-
~).
Tracing this formula, remembering to take the variation of %s r = (imply
lj,
and using
8(v,h,() [r (R + /1f) (47rr)-~ e- f dJ-l] = r (-8gij . (~j
+ V'iV'jf) + gij . 8 (~j + V'iV'jf)) (47rr)-~ e- f dJ-l
+r(R+/1f)(47rr)-~ ((l-~)~-h+ ~)e-fdJ-l =r [
-Vi.J
(~.J + V'iV'·f) e- f + V'P (e- f .£r1!.) J as n
+e- f /1 (h -~)
+ (R+ /1f) e- f (~- h - n22~)
1(4
7rr )-!!:d 2 J-l.
232
6. ENTROPY AND NO LOCAL COLLAPSING
Since We(g, f, 7) = iM (7 (R + /).f) - c (f - n)) (47r7)-n/2 e- f dJ.L, integrating this by parts and applying the divergence theorem, we have
d
ds We(g, f, 7) -Vij (~j
~ LT
+ \1/Vjf) e- f
+ (R + /).f) e- f
(t - h -
~)
n22~)
+t. (e- / ) (h +~ (-ch - c (f - n) (t - h - ~)) e- f - 7Vij (~j
+ \1i\1j f + 2~gij) e- f +7 (R + 2/).f -1\1fI 2) e- f (t - h -~)
(t - h -~) +( (R + /).f) e- f + (~ (/).f -1\1fI2 + ~) e- f -c (f - n -1) e- f
Now, integrating by parts tells us again that the terms on the last line are
1M ( ((R + /).f) + ~ (/).f -1\1 fl2 + ~)) e- f (47r7)-~ dJ.L = 1M ( (R + /).f + ~~) udJ.L. Substituting this into the above formula yields (6.9). In the next two exercises we give a unified proof of the monotonicity formulas for entropy, energy, and expander entropy. EXERCISE 6.14 (Monotonicity of We from the first variation formula). When we require that the variation (v, h, () satisfies ( = c and -h- ~( = o in (6.30) (e.g., preserves the measure (47r7)-n/2 e- f dJ.L), this leads to the following gradient flow:
t
(6.32) (6.33) (6.34)
{)
at gij = {)f
at =
-2 (Rij
+ \1i\1jf) ,
nc -/).f - R - 27'
d7 dt = c.
Show that if (g (t), f (t), 7 (t)) is a solution of the above system, then
1t We (g (t), f (t)
,7
(t)) =
271M ISij 12 udJ.L.
1.
THE ENTROPY FUNCTIONAL
W
AND ITS MONOTONICITY
233
EXERCISE 6.15 (Evolution of Ve and monotonicity of We). Consider the gauge transformed version of (6.32)-(6.34) on a closed manifold Mn:9
8
8t gij = -2~j,
(6.35)
81
8t = -~I
(6.36)
dT dt
(6.37)
2
- R + 1\711 -
nE
2T'
= E.
Let Ve ~ Vcu. Show that if (g (t) ,J (t), T (t)) is a solution of the system above, then (6.38) Also show that this implies
:t
1M IRij + \7i\7jl + ;Tgijl2 ud/-L = 2T1M ISf l ud/-L ~ o.
We (g (t), I (t), T (t)) = 2T
j
2
The next exercise relates the first variation of We to the linear trace Harnack quantity. EXERCISE 6.16 (Variation of We and linear trace Harnack). Show that, for any symmetric 2-tensor W on a closed manifold Mn, we have:
(6.39)
1MWijSfjud/-L = 1M Z (w, \7 f) ud/-L,
where Z is the linear trace Harnack inequality defined in (A.27) (6.40) and W ~ gi j wij . In particular, if c5(v,h,() ((47rT)-n/2 e- f d/-L) c5(v,h,()We (g,I,T)
=
= 0, then
1M Z(W, \7f) ud/-L,
where
Wij ~ -TVij
+ (gij
= -T2c5(v,() (T-1g) .
HINT. Use the identity
1M \7j\7iWije- f d/-L = 1M \7iWij\7jle- f d/-L = 1M Wij (\7d\7 j l - \7i\7j f) e- f d/-L. In the last two exercises in this subsection, we first rewrite We (g, I, T) and then we use the new formula to give a lower bound for We (g, I, T). 9These equations are equivalent to (6.35)-(6.37) after pulling back by diffeomorphisms generated by the vector fields V' f (t) .
6.
234
ENTROPY AND NO LOCAL COLLAPSING
EXERCISE 6.17. Let (6.41) so that w 2 = u. Show that (6.42)
We (g, j, T) = .
r (T (Rw2 + 41\7wI2) (log (w + ~ 10g(47rT) + n) w
1M ~
(6.43)
2)
+E
)
dp,
2
Ke (g, w, T) .
SOLUTION TO EXERCISE 6.17. We obtain (6.42) from substituting the definition of w,
j = -2 log w -
'2n log (47rT) ,
\7w
and
1 2
= --w\7 j
into (6.26). EXERCISE 6.18 (Lower bounds for We). Let (Mn,g) be a closed Riemannian manifold and let Rmin ~ infxEM R (x). Suppose that (g, j, T) satisfies the constraint
1M (47rT)-n/2 e- f dp, = 1. Show that for T > °the following
hold. (1) If E
(6.44)
> 0,
then
We(g, j, T) ~ TRmin - ~ Vol (g)
+E
(i
log (47rT)
+ n) >
-00.
(2) If E < 0, then
We(g, j, T) ~ -2G lEI
+ TRmin + E (i log (47rT) + n) > -00,
where
2T
-'-l€I Vol ...!...
G
()-2/n 9
lEI + 2TGs (M, g)'
and Gs (M,g) is the constant in the Sobolev inequality (6.66). Hence we conclude that when E < 0, for any A < 00, there exists a constant G (g, E, A) < 00 such that
We(g,j,T)
~
-G(g,E,A)
for T E [A-I, A] and j E Goo (M) with fM (47rT)-!j e- f dp, SOLUTION TO EXERCISE 6.18. (1) If expanding case, then (6.44) follows from
(6.45)
E
>
= 1.
0, which corresponds to the
r ulogu dp, ~ -!e Vol (g) >
-00.
1M
(2) If E < 0, which corresponds to the shrinking case, then the logarithmic Sobolev inequality implies that the entropy has a lower bound (see Section 4
235
2. THE FUNCTIONALS Jl. AND v
of this chapter). In particular, by taking a = I~ in (6.65) below, we have for c < 0,
Wc(g, j, r) 2:
1M (4r l\7wl
+ rRmin + c (6.46)
where
2:
2
+ cw 2 10g (w 2) ) dp,
(i log (41T"r) + n)
-21cl C (~~ ,g) + rRmin + c
(i log (41T"r) + n) >
-00,
-2/n 41cl 2r ) _ 2r C ( ~,g - ~ Vol (g) + 2rn2e2Cs (M,g)
is as in Lemma 6.36.
(Mn ,g) ,Jensen's inequality says that if
Applying Jensen's inequality with
J
(6.47)
= xlogx, we see that for any posi-
1M ulogu dp, 2: -log (Vol (M)) .
This is an alternative estimate to (6.45). 2. The functionals p, and
1/
Similarly to defining A (g) using the energy F (g, f) , we define two functionals p, and 1/ using the entropy W(g, j, r). In this section we first discuss the elementary properties of p, and 1/ obtained directly from the properties of the entropy- W. We show using the logarithmic Sobolev inequality that p,(g, r) is finite and we show that a constrained minimizer j of W(g, j, r) exists. We end this section with the monotonicity of p,. 2.1. The diffeomorphism-invariant functionals p, and 1/. Let u ~ (41T"r)-n/2 e- f as in (6.3). We define a subset X of rotet x COO(M) x R+ by (6.48)
X
=
{(g,J, r) :
1M udp, =
I} .
Note that if (g, j, r) E X, then (cg, j, cr) E X for all c > 0, and (<J?*g, <J?* j, r) E X for any diffeomorphism <J? : M ~ M. We consider the restriction of W to X. Given (g, r), we first minimize W(g, j, r) over j with (g, j, r) E X to get p,(g, r), and then we minimize p,(g, r) among r > to get 1/ (g).
°
6. ENTROPY AND NO LOCAL COLLAPSING
236
DEFINITION 6.20 (Infimum invariants J-l (g, r) and v jR+ --t jR and v : 9J1et --t jR are defined by
(g». The function-
als J-l : 9J1et x
(6.49) (6.50)
J-l(g, r)
~
inf {W(g, f, r) : f E Coo (M) satisfies (g, f, r) E X} ,
v(g) ~ inf {J-l(g, r) : r E
jR+} .
Note that we do not assume that the two functionals J-l(g, r) and v(g) take finite values. We will see later that J-l(g, r) is always finite for any given 9 and r, and in the important case where A(g) > 0 (where the proof of the nonexistence of nontrivial expanding breathers cannot be applied to the shrinking ones), we have that v(g) is finite. We have the following elementary properties of these two functionals. (i) (Continuous dependence of J-l on 9 and r) J-l(g8' r) is a continuous function of (s, r) for any C 2 family g8. 10 (ii) (Continuous dependence of v on g) V(g8) is a continuous function for any C 2 family g8. (iii) (Scale invariance) It follows from the scaling property of W that we have
J-l(g, r) v(g)
= J-l (cg, cr), = v(cg).
(iv) (Diffeomorphism invariance) Since W is invariant under a diffeomorphism <J> : M --t M, we have
J-l(g, r) = J-l (<J>* g, r) , v(g) = v(<J>*g). Compare (i) and (ii) with Lemma 5.24. EXERCISE
6.21. Prove properties (i)-(iv) above.
Using the fact that the variation of W : X 8(O,h,O) W
(g, f, r) = -
where h satisfies (6.49),
1M rh (2tl.f + f -
--t
jR
with respect to
(; + 1)
-IV fl2 +
f is
R) udJ-l,
fM hudJ-l = 0, we have the Euler-Lagrange equation of r (2tl.f -
IV fl2 + R) + f
- n
=C
for some constant C. If fT is a minimizer of (6.49) (we will see the existence of fT in the next subsection), then it follows that
J-l (g, r) =
1M [r (2tl.fT -
IV fTI2 + R) + fT
- n] (47rr)-n/2 e- JT dJ-l,
and hence C = J-l (g, r) for a minimizer. Therefore we have the following. I°It is easy to see that I-'(g, r) is semi-continuous in g. This is a consequence of the fact that if a function hex, y) is continuous, then infyEY hex, y) is upper semi-continuous in x.
2. THE FUNCTIONALS
f.L
AND
237
v
LEMMA 6.22 (Euler-Lagrange for minimizer). The Euler-Lagrange equation of (6.49) is (6.51)
For the minimizer fT of (6.49), (6.52)
r (211fT - IV fTI2
+ R) + fT
- n = J.L (g, T) .
Compare (6.52) to the equation 26.f - IV fl2 + R = A (g) for the minimizer f of F (g,.). In terms of w ~ (41rT)-n/4 e- f /2 as in (6.41), a simple computation shows that J.L is the lowest eigenvalue of the nonlinear operator: (6.53)
N (w) ~ -4r6.w + TRw -
(% 10g(41rT) + n) w -
2w logw
= J.L (g, T) w.
2.2. The finiteness of J.L and the existence of a minimizer a consequence of Exercise 6.18 we have the following.
f. As
LEMMA 6.23 (Finiteness of J.L). For any given 9 and T > 0 on a closed manifold Mn,
J.L(g, T) >
(6.54)
-00
is finite. PROOF. Since J.L (Tg, T) = J.L (g, 1), we may assume without loss of generality that r = 1. We need to show that for any metric 9 there exists a constant c = c(g) such that (6.55)
W (g, f, 1) =
1M (R + IV fl2 + f -
n) (41r)-n/2 e- f dJ.L
~c
for any smooth function f on M satisfying (41r) -n/2 1M e- f dJ.L = 1. As in (6.41) with T = 1, let w = (41r)-n/4e- f / 2. By (6.42), the lemma is equivalent to showing that W (g, f, 1)
=
1M (41Vw12 + (R -
~
1i (g,w) ~ c
210gw -
for any w > 0 such that 1M w 2dJ.L = 1. Since R - n ~ infxEM R (x) - n > exists C < 00 such that
-00,
%log (41r) -
n) w 2) dJ.L
it suffices to show that there
1M w 2 10gwdJ.L ::; 2 1M IVwl2 dJ.L + C for all w > 0 with 1M w 2dJ.L = 1. This follows from the logarithmic Sobolev inequality (6.65),11 which we state in the next subsection. D 11With a = 2.
238
6.
ENTROPY AND NO LOCAL COLLAPSING
Next we prove the existence of a smooth minimizer for (6.49); compare the proof below with the proof of Lemma 5.22. LEMMA 6.24 (Existence of a smooth minimizer for W). For any metric g on a closed manifold Mn and T > 0, there exists a smooth minimizer fT of W (g,., T) over X. PROOF. Again assume T = 1. The lemma will follow from showing that there is a smooth positive minimizer WI for 1-l (g, w) under the constraint iM w2dJ.Lg = 1. A smooth minimizer II of W (g,., 1) is then given by II = - 2 log WI - ~ log (471") (see Rothaus [312]). We give a sketch of the proof. Suppose W is such that 1-l (g, w) :S C 1. Then the above considerations imply that since iM w 2dJ.Lg = 1,
C1
~ 1-l (g,w) = 1M (41V'wI2 + (R -
~ 2 1M lV'wl2 dJ.L -
2logw -
~ log (471") -
n) w2) dJ.L
C2,
where we used (6.65) below with a = 1. Hence any minimizing sequence for 1-l (g, .) is bounded in W 1,2 (M). We get a minimizer WI in W 1,2 (M) and by (6.53), WI is a weak solution to
-4~Wl + RWI - 2WllogWl - (~lOg(471") + n) WI = J.L(g, l)Wl. By elliptic regularity theory, we have WI E Coo (see Gilbarg and Trudinger [155] for a general treatise on second-order elliptic POE). Finally, one can prove that WI > 0; see [312] for more details. 0 REMARK 6.25. We also have (6.56)
J.L(g, T)
~ inf {W(g, f, T) : f
E W 1 ,2(M),
1M udJ.L = I} ,
where Coo is replaced by W 1,2 in (6.49) and u is defined in (6.3).
2.3. Monotonicity of J.L. Let (g (t), T (t)) , t E [0, T], be a solution of (6.14) and (6.16) with T (t) > 0. For any to E (0, T], let f (to) be the minimizer of
{W(g (to), f, T (to)) : f E Coo (M) satisfies (g,J, T) E X} and solve (6.15) for f (t) backwards in time on [0, to]. By the monotonicity formula, we have d dt W (g (t) , f (t) , T (t)) ~
°
for t E [0, to]. Note that the integral constraint (6.48) is preserved by the modified coupled equations (6.14)-(6.16). This can be seen from the following calculation: (6.57)
.!!:.- { (471"T)-n/2 e- f dJ.L = { (au dt 1M 1M at
RU) dJ.L = 1M { (-O*u -
~u) dJ.L = 0.
2. THE FUNCTIONALS J.L AND
239
II
Hence we have
(6.58)
I-" (g (t), 7 (t)) ::; W (g (t), f (t), 7 (t)) ::; W (g (to) , f (to) , 7 (to)) = I-" (g (to) , 7 (to) )
for t E [0, to]. The above inequality implies
(6.59)
~lt=tol-"(9(t)'7(t)) 2:
:tlt=to W(g(t),J(t),7(t)) 2: °
in the sense of the lim inf of backward difference quotients. This inequality holds for all to E [0, T] . Actually we have
:t It=to I-" (g (t) , 7 (t) )
2:
1M 27 (to) IRi j (to) + V'iV'jf (to) - 27 ~to)g (to)i I
2
j
x (41l"7 (to))-~e-f(to)dl-"g(to) for the minimizer f (to) of {W(g (to)," 7 (to)) : (g (to)," 7 (to)) EX}. Hence, from either (6.58) or (6.59), we have the following monotonicity formula for
1-". LEMMA 6.26 (I-"-invariant monotonicity). Let (g (t) ,7 (t)) , t E [0, T] , be a solution of (6.14) and (6.16) on a closed manifold Mn with 7 (t) > 0. For all tl ::; t2 ::; T, we have
°: ;
(6.60) In particular, (6.61) for t E [O,T] and r
> 0.
The following exercise continues our discussion in subsection 1.3 of this chapter. Again u ~ (41l"7) - ~ e- f . EXERCISE 6.27 (I-"<:-invariant monotonicity). Define the I-"<:-invariant on a closed manifold Mn:
I-"<:(g, 7)
~ inf { W<:(g, f, 7) : f
E COO (M),
1M udl-" = 1} .
(1) Show that for any c> 0,
1-"<: (cg, C7) = 1-"<: (g, 7) . (2) Show that if (M, 9 (t)) is a solution to the Ricci flow on a time interval Ie JR, c E JR, and ~; = c with 7 (t) > 0, then 1-"<: (g (t), 7 (t)) is monotonically nondecreasing on I. That is, for t2 2: tl,
6. ENTROPY AND NO LOCAL COLLAPSING
240
(3) Improving on the previous part, show that for any times tl, t2 E I with tl :::; t2, 1-lE: (g (t2) , 7 (t2)) - 1-lE: (g (t 1) , 7 (t 1) )
:> [ ' 27 (t) 1M 1Rij (t) + 'Ili 'Iljf (t) + £~7(;;;j 12 u (t) dl'g(t)dt, where f (t) is the minimizer of WE: (g (t), . [tl' t2J .
,7
(t)) for each t
E
2.4. The behavior of I-l under Cheeger-Gromov convergence. If (Nn, g) is a complete, noncompact Riemannian manifold, we generalize the definition of the functionall-l to noncompact manifolds by
where W (g, f, 7) is defined as in (6.1) and the infimum is taken over smooth functions with compact support satisfying the constraint. 12
Suppose that we (N:;;, gOC)) xoo) in the Coo Cheeger-Gromov sense. Then
LEMMA 6.28 (I-l under Cheeger-Gromov convergence).
have (N/:, gk, Xk) for any 7> 0,
->
I-l (goo, 7) ~ lim sup I-l (gk, 7) .
k---+oo
By definition, there exists an exhaustion {UdkEN of Noo by open sets with Xoo E Uk and diffeomorphisms Vk ~
Nk with
->
(Noo, goo) in Coo on
1(
47r7) -n/2 e - J dl-l goo = 1.
N=
Then by the diffeomorphism invariance of W, for all kEN large enough, we have fNk (47r7)-n/2 e- Jo;1 dl-l(;I)*goo = 1 and
W(
7) makes sense since Uk :::) supp (e- J/
2)
when k is large.
12Implicitly it is understood in the discussion here that we are considering W as a function of w ~ (47rT)-n/4 e -f/2 E c,:, (N) so that there is no problem with f = 00. As usual, we use the convention that w 2 log w = 0 when w = O.
241
2. THE FUNCTIONALS J.I. AND v
Now 119klvk - (<1>;;1)* 900 II Cf (cI>k(supp(f)),(cI>;lrgoo)
= II<1>k [9klvk ]
-
9001I Cl (suPP(f),goo)
-0. Hence ILk (47rT)-n/2 e- Jocl>;l df..L( cI>;l)* goo - Lk (47rT)-n/2 e- Jocl>;l df..L9k ::; (47rT)-n/2
r e-Jocl>;l iN, k
::; 119klvk - (<1>;;1)*
x
r
iNk
(1-
I
df..L9k ) df..L -1' df..L( ~k .... -l). goo (cI>k ) goo
90011~~cI>k(SUPP(f))'(cI>;1)*goo)
(47rT)-n/2 e- Jocl>;1 df..L(cI>-l)' k goo
-0, which implies
We conclude that
W (Noo , 900' j, T) =
J~ W (Uk, <1>k
[9klvk] , j, T)
= k~oo lim W(Nk ,9k,jo<1>;;!,T) 2:: lim sup f..L (9k, T) . k~oo
To see why the last inequality is true, note that the functions log Ck satisfy Ck - 1 and the constraints
!k
~ j
0
<1>;; 1+
r (47rT)-n/2 e- Jkdf..L9k = 1
iNk
and W (Nk ,9k, j
0
<1>;;1, T) = CkW (Nk,9k, jk, T) - Ck logCk
2:: Ckf..L (9k, T) - Ck log Ck· Since e- J/2 is an arbitrary nonnegative function with compact support and since it satisfies the constraint with respect to 900' we obtain f..L(900,T) 2:: lim sup f..L (9k,T) . k~oo
D
242
6.
ENTROPY AND NO LOCAL COLLAPSING
3. Shrinking breathers are shrinking gradient Ricci solitons Let (Mn,g(t)) , t E [O,T), be a shrinking Ricci breather on a closed manifold with g(t2) = aq,*g(iI), where t2 > tl and a E (0,1) . As discussed at the beginning of Section 1 of this chapter, we only need to show that when >.(g(t)) > for all t E [tl, t2], the shrinking breather is a shrinking Ricci soliton. By Koiso's examples (subsection 7.2 of Chapter 2), such solutions need not be Einstein. In this section we give two variations on the proof that shrinking breathers on closed manifolds are shrinking gradient solitons. The first proof, which also appears in Hsu [208], involves fewer technicalities in that it uses J-L instead of 1/. The first proof also does not use the assumption>. (g(t)) > O. On the other hand, the second proof requires some knowledge of the asymptotic behavior of J-L.
°
3.1. First proof using functional J-L. The following result of Perelman rules out periodic orbits for the Ricci flow in the space of metrics modulo diffeomorphisms and scalings. THEOREM 6.29 (Shrinking breathers are gradient solitons). A shrinking breather for the Ricci flow on a closed manifold must be a gradient shrinking Ricci soliton. FIRST PROOF. Let (M,g(t)) be a shrinking Ricci breather with g(t2) aq,*g(tl), where t2 > tl and a E (0,1). Define T
(
=
t2 - atl t ..... 1 - t, )
...!...
-a
so that ~; = -1,
and T (t2) = aT (tl). By Lemma 6.24, there is a minimizer
12 for
{W(g (t2),f, T (t2)) : f E Coo (M) satisfies (g, f, T) E X}, so that W (g (t2), 12, T (t2)) = J-L (g (t2) , T (t2)). Define f (t) to solve (6.15) on [t}, t2] with f (t2) = h. By the monotonicity formula (6.17) and the definition of J-L, we have J-L
(g (tJ), T (tl)) ::; W (g (tl) , f (tl) , T (tl)) ::; W (g (t), f (t), T (t)) ::; W (g (t2) , 12, T (t2)) = J-L (g (t2) , T (t2))
for all t E [tl, t2]. Since g(tl) = aq,*g(t2) and T (t2) diffeomorphism and scale invariance of J-L, we have J-L
(g (tl) , T (tl)) =
J-L
= aT (tl),
(g (t2) , T (t2)).
This and the fact that W (g (t) ,I (t) , T (t)) is monotone implies W (g (t), f (t), T (t))
= J-L (g (t), T (t)) == const
by the
3.
SHRINKING BREATHERS ARE SHRINKING GRADIENT RICCI SOLITONS
for t E [tt, t2J. Thus W (g (t), f (t), T (t))
It
f (t) == 0,
243
is the minimizer for W (g (t), f (t), T (t)) and so by (6.17), we have
1M I~j + '\h'Vjf for all t E [tl' t2J . We conclude that goo (6.62) ~j + V/vjf - 2~
~; 12 e- JdJl == 0
= 0 for t E [tl' t2J.
Since a breather is a periodic solution of the Ricci flow (modulo diffeomorphisms and homotheties), by the uniqueness of solutions to the Ricci flow on closed manifolds, the behavior of 9 (t) on [tt, t2J determines completely the behavior of 9 (t) on its whole time interval of existence. This is why, from (6.62), which is valid on [tl' t2J , one can deduce that 9 (t) is a breather on [O,T). D
3.2. Asymptotic behavior of Jl and finiteness of //. In the second proof of Theorem 6.29 given below we need the finiteness of //, which in turn depends on A(g(t)) > 0 and the following asymptotic behavior of Jl. We have shown that for each 9 and T > 0, Jl (g, T) is finite. However we have yet to study the behavior of Jl (g, T) as T -+ 00 or T -+ O. Recall that
A(g)
=
Ad-4~ + R) = inf {1M (R + IVfI2)e- JdJl : 1M e- JdJl = I}.
Since Jl and Ware modifications of A and F, we can prove the following. LEMMA 6.30 (Jl
-+ 00
as T -+
00
when A > 0). If A(g)
lim Jl (g, T)
T-+oo
> 0, then
= +00.
REMARK 6.31. The idea of the proof is that when T -+ 00, the F term in the expression (6.4) for W dominates, so if inf F > 0, then inf W -+ 00 as T -+ 00. PROOF. By Lemma 6.24, for any T > 0, there exists a Goo function fT with fM (47rT)-n/2 e-f.r dJl = 1 such that
Jl(g, T) = W(g, fn T) =
1M [T (R + IV fT12) + fT - n] (47rT)-n/2 e-f. dJl. r
We add a constant to fT so that it satisfies the constraint for F (g, .) (instead
1
of W(g,·, T)) and we define ~ fT+~ log (47rT) so that fM e- i dJl = 1. Then by the logarithmic Sobolev inequality (e.g., Corollary 6.38 with b = 1), we have
Jl(g, T)
= 1M [T (R + IV112) + 1 - ~ log (47rT) -
~ 1M (TR + (T -
~ (7 -
1)
1)
n] e- i dJl
IV112) e- i dJl- ~ log (47rT) -
1M (R + IV112) e- i dJl + R
min -
n - Gdg)
~ log (47rT) -
n - Gl (g).
6.
244
Hence, if
7 ~
ENTROPY AND NO LOCAL COLLAPSING
1, we have
/-L(g, 7)
(6.63)
~
(7 - 1) ,\(g) -
Since '\(g) > 0, we have limT->oo /-L (g, 7)
n
2 log 7 -
C2 (g) .
o
= +00.
EXERCISE 6.32. Show that if '\(g) < 0, then limT->oo /-L (g, 7) = -00. In particular, if '\(g) < 0, then v (g) = -00. SOLUTION TO EXERCISE 6.32. Since '\(g) = 1 and
< 0,
there exists
fo
with
fM e- fOd/-L
a,* 1M Define 1,*
fo -
(R + IV' fol 2) e- fOd/-L < 0.
~ log (471'7) so that fM(471'7)-n/2 e- Jd/-L = 1. We have for all
7>0 /-L (g, 7)
50 W (g, 1,7) 50 7 1M
= a7 +
7 ---+
(R + 1V'112) + 1 -
n] (471'7)-n/2 e- Jd/-L
(R + IV' fol2) e-fod/-L + l 1M (471'7)-n/ 2d/-L (471'7 ) -n/2
since xe- x 50 ~ for x When
1M [7
=
e
> 0.
Vol (g) ,
The result follows from a < 0.
0+, we have
LEMMA 6.33 (Behavior of /-L(g,7) for
7
small). Suppose (Mn,g) 'ts a
closed Riemannian manifold. (i) There exists f > such that
°
/-L (g, 7)
<
° for all 7
E (0, f).
(ii)
The proof of Lemma 6.33 will be given elsewhere. From Lemmas 6.30 and 6.33, we have COROLLARY 6.34. If ,\ (g) > 0, then v (g) is well defined and finite. Also, there exists 7 > such that v (g) = /-L (g, 7) .
°
3.3. Monotonicity of v and the second proof. The following lemma is the monotonicity property of v(g(t)) along the Ricci flow g(t). LEMMA 6.35 (v-invariant monotonicity). Let (Mn, 9 (t)) , t E [0, T), be a solution to the Ricci flow on a closed manifold. (1) The invariant v(g(t)) is nondecreasing on [0, T), as long as v(g(t)) is well defined and finite.
3.
SHRINKING BREATHERS ARE SHRINKING GRADIENT RICCI SOLITONS
245
°
(2) Furthermore, if A (g (t)) > and ifv(g(t)) is not strictly increasing on some interval, then g(t) is a gradient shrinking Ricci soliton. (3) Ifv(g(to)) = -00 for some to, then v(g(t)) = -00 for all t E [0, to]. PROOF. (1) Given any
°
~
iI < t2 < T, we shall show that
(6.64)
>
Since by assumption, V(g(t2)) such that
-00,
for any E >
°
there exist
12
and T2
W(g(t2), 12, T2) ~ V(g(t2)) + E. Let (f(t), T(t)) , t E [0, t2], be a solution of the backward heat-type equation (6.15) with f(t2) = 12 and T(t2) = T2. By the monotonicity formula (6.17), we have l3
W(g(t2), f(t2), T (t2))
~
W(g(tt}, f(tl), T (tl)),
where equality holds if and only if 1 ~j
°
+ "h'ljf - 2Tgij =
for all t E (tI, t2)'
This implies
V(g(t2))
+ E ~ W(g(t2), f(t2), T (t2))
The result follows since
E
>
°
~
W(g(tl), f(iI), T (tl))
~
V(g(tl))'
is arbitrary.
(2) Suppose V(g(tl)) = V(g(t2)) for some tl < t2' Since A (g (t)) Corollary 6.34, there exist 12 and T2 such that
> 0, by
W(g(t2), 12, T2) = V(g(t2)). In this case, by repeating the argument in (1), we obtain W (g (t), f (t), T (t))
= v(g(t)) == const
for all t E [tl, t2]' As in the proof of Theorem 6.29, we can conclude that g(t) is a gradient shrinking Ricci soliton. (3) If v(g(to)) = -00, then for any N > -00 there exist fa and TO such that W (g (to), fa, TO) ~ N. Let (f(t), T(t)) , t E [0, to], be the solution of (6.15) with f(to) = fa and T(tO) = TO. For all t E [0, to],
v(g(t)) Since N >
-00
~
W (g (t), f (t), T (t))
~
W(g(to), f(to), T (to))
is arbitrary, we conclude v(g(t)) =
-00
~ N.
for all t E [0, to].
D
Using the v-invariant instead of the J.L-invariant, we can give a SECOND PROOF OF THEOREM 6.29. As we stated at the beginning of this section, we only need to consider a shrinking breather g( t) with g( tt} = a 0, t E [tl, t2]' From the elementary properties (iii) and (iv) of J.L and v in subsection 2.1 of this chapter, we have v(g(tt}) =
6.
246
ENTROPY AND NO LOCAL COLLAPSING
1I(g(t2)). Now the theorem follows from Corollary 6.34 and Lemma 6.35(1)(2), which tell us that 1I (g (t)) is monotone and finite and characterizes when 1I (g (t)) is constant. 0 4. Logarithmic Sobolev inequality In this section we give a proof of the logarithmic Sobolev inequality which we have used earlier. The logarithmic Sobolev inequality is related to the usual Sobolev inequality and has the advantage of being dimensionless.
4.1. Logarithmic Sobolev inequality on manifolds. LEMMA 6.36 (Log Sobolev inequality, version 1). Let (Mn, g) be a closed Riemannian manifold. For any a > 0, there exists a constant C (a, g) (given by (6.67)) such that if'P > satisfies iM 'P2dJ.l = 1, then
°
1M 'P210g 'PdJ.l :::; a 1M 1V''P12 dJ.l + C (a, g) .
(6.65)
PROOF. Recall that the Sobolev inequality (see Lemma 2 in [245]) that
if
iM 'P2dJ.l = 1, then (assume n > 2)
(6.66) where V = Voig (M). Note that for en = ~e we have en log x :::; x2/n for all x> 0, so that
1M 'P2+~ dJ.l :::; E1M{ 'P2+~ dJ.l + ~ 1M( 'P2 dJ.l ,
1M
en { 'P210g 'PdJ.l:::; {
E
for any E > 0, since 'P1+~ 'P :::; E'P2( l+~)
+ ~'P2.
By Holder's inequality,
n-2
1M 'P2'P~dJ.l:::; (1M 'Pn2~2dJ.l) n- (1M 'P2dJ.l) Hence, using
iM 'P2dJ.l =
2
TO
1, we have
n-2
en
1M 'P210g 'PdJ.l :::; E (1M 'P n2~2 dJ.l) n- + ~ :::; C (~, g) (1M 1V''P12 dJ.l + v- 2/ n ) + ~. s
Inequality (6.65) follows by choosing (6.67)
C (a,g ) -- aV -2/n
+ an2 e2C4s (M ,g)"
Now we have proved the lemma when n > 2. We leave the n an exercise. EXERCISE 6.37. Prove the above lemma when n
= 2 case as
= 2.
Making the substitution 'P = e-¢>/2 in (6.65), we have the following.
0
4.
LOGARITHMIC SOBOLEV INEQUALITY
247
COROLLARY 6.38 (Log Sobolev inequality, version 2). For any b > 0, there exists a constant C (b, g) such that if a function > satisfies fM e-dJ-t = 1, then (6.68)
4.2. Logarithmic Sobolev inequality on Euclidean space. We give a proof of Gross's logarithmic Sobolev inequality on Euclidean space [170]. Although this result will not be used elsewhere in Part I of this volume, we include it here since it is both fundamental and elegant. THEOREM 6.39. For any nonnegative function > E W 1,2 (JRn) , we have
Note that the above inequality is scale-invariant, that is, the inequality is preserved under multiplication of > by a positive constant. Also, if fRn >2dx = 1, then the inequality says that fn~.n >2 log > dx ~ fRn IV>12 dx. The following consequence of Gross's logarithmic Sobolev inequality is actually equivalent to it. (We leave the proof of the equivalence to the reader.) COROLLARY 6.40. If fn~.n (47f'r)-n/2 e- f dx (6.69)
kn
(r
In particular, taking r
IV fl2 + f
= 1, then
- n) (47f'r)-n/2 e- f dx 2: O.
= 1/2, we have
(6.70)
provided fRn (27f')-n/2 e- f dx = 1. Moreover, if we can perform an integration by parts, then we may rewrite (6.69) as (6.71)
kn
(r (2D.f
-I VfI 2 ) + f
- n) (47f'r)-n/2 e- f dx 2: O.
REMARK 6.41. Compare the LHS of (6.69) with the entropy (6.1) and compare the integrand on the LHS of (6.71) with Perelman's differential Harnack quantity (6.20). PROOF OF THE COROLLARY. We shall prove just the case where r = 1/2 since the general case follows from making the change of variables x ~ (2r)-1/2 x. Let > be defined by f = ~ - 210g>, so that e-f = e-lxI2/2. >2
248
6.
and \1 f
ENTROPY AND NO LOCAL COLLAPSING
= x - 2'2/. We compute in
(~I\1 fl2 + f
= 2in where dv
- n) (27r)-n/2 e- f dx
(~ Ixl2 cj>2 -
cj> X . \1 cj> + 1\1 cj>1 2 - cj> 210g cj> - %cj>2) dv,
= (27r)-n/2 e-lxI2/2dx. Now integrating by parts yields
r
J/Rn
_cj> x. \1cj>dv
=~
r
2 J/Rn
ncj>2dv -
~
r
2 J/Rn
Ixl 2 cj> 2dv,
so that we have the identity
Hence (6.72)
in
(~I\1fI2+f-n) (27r)-n/2e-fdx=2in (1\1cj>1 2 -cj>210g cj»dv,
with the constraint
Since log U/Rn cj> 2dv) ity, we have
= 0, by (6.72) and Gross's logarithmic Sobolev inequal-
o EXERCISE 6.42. Show that Gross's logarithmic Sobolev inequality for Euclidean space implies that Euclidean space (JR n , 91E) has nonnegative entropy: W (9IE, f, 7) ~ 0 and
JL (9IE, 7)
= O.
Now we give Beckner and Pearson's proof of Gross's logarithmic Sobolev inequality, which is a consequence of the following [23]. PROPOSITION 6.43. If J/R n 'Ij; (x)2 dx (6.73)
~ log (_2_ 4
r
7ren J/Rn
=
1\1'1j; (x)1 2dX)
1, then
~
r
J/Rn
(log 1'Ij; (x)l) 'Ij; (x)2 dx.
Note that this inequality is scale-invariant. We first show that (6.73) implies Gross's logarithmic Sobolev inequality.
4. LOGARITHMIC SOBOLEV INEQUALITY
249
6.39 FROM THE PROPOSITION. Given f such that
PROOF OF THEOREM
r (27r)-n/2 e- dx = 1,
(6.74)
f
J~n
let 'I/J ~ (27r)-n/4 e-f/2, so that 10g'I/J Then (6.73) implies
~ log (7r~n
(6.75)
~-
In ~ IV'
= -~ - %log (27r)
and J~n 'l/J 2dx
= l.
fl2 e- f (27r)-n/2 dX)
r (L2 + ?!:4 log (27r)) e- f (27r)-n/2 dx,
J~n
so that
~
r lV'fI2e-f (27r)-n/2dx ~ en2 exp{-~nknr fe-f (27r)-n/2 dX}.
2kn
We claim
r
(6.76) en exp {-~ fe- f (27r)-n/2 dX} 2 n J~n which implies the
T
Since J~n (27r)-n/2 e- f dx
?!:exp{~
+f
fl2
(n - f) e- f (27r)-n/2 dx,
= 1,
- n) e- f (27r)-n/2 dx
~ O.
inequality (6.76) is equivalent to
r
(?!:-f)e- f (27r)-n/2 dX } n J~n 2
2
r
J~n
= 1/2 case of (6.69):
In (~IV'
(6.77)
~
~?!:+
r
(?!:-f)e- f (27r)-n/2 dx . 2 J~n 2
If we let du = e- f (27r) -n/2 dx and 9 = ~ - f, then the above inequality becomes
9dU} ~ 1 + ~ gdu. n J~n n J~n 1 + a for all a E R
exp This follows from e a
~
{~
r
r
D
Now we present the PROOF OF PROPOSITION
6.43. By Jensen's inequality, if J~N
IFl r dx =
1, then
(p - r)
r
J~N
(log IFI) IFl r dx =
r
J~N
Slog
log
(IFIP- r ) IFl r dx
(IN IFIP-
r
IFl r dX) = log
for p ~ r > O. The L 2-Sobolev inequality says 2 IIFIILN* S
AN
l
~N
IV'FI 2 dx,
(IN IFIP dX)
6. ENTROPY AND NO LOCAL COLLAPSING
250
2N r(N) )2/N I By Sterl'mg '£ h N * -- N-2 were an d A N -- ( r(N/2) nN(N-2)' s ormu1a, r(N) rv .j2iiNN-!e- N , so that AN rv n;N for N large, where r is the Gamma function. Hence, if fIRN 1F12 dx = 1, then by above two inequalities
log ( AN
IN IV'FI2 dX) ~ ~* log (IN IFI dX) ~ ~ IN (log IF!) 1F12 dx, N•
where we used ~. (N* - 2) = Given f : ]Rn ~ ]R with In~.n by
t.
f 2 dx =
1, let N
= nf and define F : ]RN
~ ]R
i
F (x) ~
IIf (Xk) , k=l
where x
=
(Xl, ... ,
Xi), Xk
E ]Rn for
k
= 1, ... , f.
Now
V'F(x) = (V'f(XI) V'f(Xi)) F (x) f (xI) , ... , f (Xi) . Hence
and
i
since In~.n f (x)2 dx
=
1. Using In~nl F (x)2 dx
=
II fIRnf (Xk)2 dXk = 1, we k=l
have
5. NO FINITE TIME LOCAL COLLAPSING
251
Now
f
JlRnl =
(log IF (x)l)
lnl t IR
IF (x)12 dx
(f (xk)2Iog Ilk (xk)1
k=l
IIf (Xi)2) dX1 ... dxR.
i#k
= f f (log If (x)J) f (x)2 dx.
JlRn
Hence
for all fEN. Recall that AnR. implies log
rv
7r:nR.' which by taking the limit as f
(~ f IV f (x)1 2 dX) 7ren JlRn
-t
00,
2: i f f (x)2Iog If (x)1 dx. n
JlRn
This completes the proof of the proposition.
o
5. No finite time local collapsing: A proof of Hamilton's little loop conjecture In this section we first define the notion of ~-noncollapsed at scale r and show its equivalence to the injectivity radius estimate. We then prove Perelman's celebrated no local collapsing theorem and indicate its equivalence to Hamilton's little loop conjecture. We end this section by showing the existence of singularity models for solutions of the Ricci flow on closed manifolds developing finite time singularities corresponding to sequences of points and times with curvatures comparable to their spatial maximums. Perelman's no local collapsing theorem solves a major stumbling block in Hamilton's program for the Ricci flow on 3-manifolds. In particular, it provides a local injectivity radius estimate which enables one to obtain singularity models when dilating about finite time singular solutions of the Ricci flow on closed manifolds of any dimension. The no local collapsing theorem also rules out the formation of the cigar soliton as a singularity model. 14 The above two consequences of the no local collapsing theorem, together with Hamilton's singularity theory in dimension 3, imply that necks exist in all finite time singular solutions on closed 3-manifolds. This, together with Hamilton's analysis of nonsingular solutions, leads one to hope/expect that 14More precisely, in dimension 3 it rules out singularity models which are quotients of the product of the cigar soliton and the real line.
252
6.
ENTROPY AND NO LOCAL COLLAPSING
Ricci flow with surgery may lead to the resolution of Thurston's geometrization conjecture. Perelman's deeper analysis of 3-dimensional singularity formation greatly strengthens this expectation. I5 5.1. K-noncollapsing and injectivity radius lower bound. 5.1.1. K-noncollapsing on a Riemannian manifold. Let (Mn, complete Riemannian manifold.
g)
be a
DEFINITION 6.44 (K-noncollapsed). Given p E (0,00] and K > 0, we say that the metric 9 is K-noncollapsed below the scale p if for any metric ball B(x,r) with r < p satisfying I Rm(y)1 ::; r- 2 for all y E B(x,r), we have VolB(x,r) > rn -
(6.78)
K.
If 9 is K-noncollapsed below the scale 00, we say that 9 is K-noncollapsed at all scales. Complementarily, we give the following. DEFINITION 6.45 (K-collapsed). We say that 9 is K-collapsed at the scale r at the point x if I Rm(y) I ::; r- 2 for all y E B(x,r) and Vol B(x, r) < K. rn
(6.79)
The metric 9 is said to be K-collapsed at the scale r if there exists x E M such that 9 is K-collapsed at the scale r at the point x. Thus 9 is not K-noncollapsed below the scale p if and only if there exists r < p and x E M such that 9 is K-collapsed at the scale r at the point x. REMARK 6.46.
(1) If Mn is closed and flat, then 9 cannot be K-noncollapsed at all scales since I Rm I = 0 ::; r- 2 for all r and Vol B(x, r) ::; Vol
(M)
so that limr-->oo Vol~~x,r)
= 0 for all x E M.
g)
(2) If (Mn, is a closed Riemannian manifold, then for any p > 0 there exists K > 0 such that 9 is K-noncollapsed below the scale p. We have the following elementary scaling property for K-noncollapsed metrics. LEMMA 6.47 (Scaling property of K-noncollapsed). If a metric 9 is Knoncollapsed below the scale p, then for any a > 0 the metric a 2 g is Knoncollapsed below the scale ap. I5Some aspects of Perelman's singularity theory are discussed in Chapter 8 of Part I and also in Part II of this volume.
5.
NO FINITE TIME LOCAL COLLAPSING
253
PROOF. We leave it as an exercise to trace through the definition of Knoncollapsed and verify that the lemma follows from the scaling properties: Bg(x, r) = B Q 2 g(X, ar), I RmQ 2 g(Y) I = a- 2 1 Rmg(Y)I, and VoI Q 2 g B Q 2 g(X, ar) = anVolgBg(x,r). 0 The next lemma says the property of being K-noncollapsed below the scale p is preserved (stable) under pointed Cheeger-Gromov limits. LEMMA 6.48 (K-noncollapsed preserved under limits). Let {(M k,gk, Ok)} be a sequence of pointed complete Riemannian manifolds. Suppose that there
exist K > 0 and p > 0 so that each (Mk' 9k) is K-noncollapsed below the scale p. Furthermore assume that (Mk' gk, Ok) converges to (M~, goo, 0 in the pointed Cheeger-Gromov C 2-topology. Then the limit (Moo, goo) is K-noncollapsed below the scale p. (0 )
PROOF. This is because the distance function, the curvature, and the volume all converge under the limit. In particular, suppose x E Moo and r < p are such that I Rmgoo (Y)I :::; r- 2 for all Y E Bgoo (x, r). Then for every £ E (O,r), there exists k(£) EN such that I Rmgk(y)1 :::; (r_£)-2 for all y E B9k (x, r - E) and for all k 2: k (E) . Since each gk is K-noncollapsed below the scale p, we have Volgk(Bgk(x, r-£)) 2: K (r - Et for all k 2: k (£). Taking the limit as k ---t 00, we have Volgoo(Bgoo(x,r - E)) 2: K(r - £t. Letting £ ---t 0, we then conclude that Volgoo(Bgoo(x,r)) 2: Kr n as desired. 0 Recall that if Rc 2: 0 on a complete Riemannian manifold (Mn,
g) ,
then for p E M fixed, Vol~~p,r) is a nonincreasing function of r. When M is noncompact, the notion of K-noncollapsed at all scales is closely related to another invariant of the geometry of infinity called the asymptotic volume ratio, which we now define.
g)
DEFINITION 6.49 (Asymptotic volume ratio). Let (Mn, be a complete noncompact Riemannian manifold with nonnegative Ricci curvature. The asymptotic volume ratio is defined as the limit of volume ratios by
(6.80) where
AVR(g) Wn
~
lim VoIB(p,r) < 00, r->oo wnrn
is the volume of the unit ball in
jRn.
We say that (M,g) has
maximum volume growth if AVR(g) > O. REMARK 6.50. The asymptotic volume ratio is independent of the choice of basepoint p EM. EXERCISE 6.51 (AVR> 0 implies noncollapsed). Show that if (Mn,
g)
is a complete noncompact Riemannian manifold with Rc 2: 0 and AVR (g) > 0, then 9 is K-noncollapsed on all scales for K = Wn AVR (g) .
254
6.
ENTROPY AND NO LOCAL COLLAPSING
The next exercise shows that on small enough scales 9 is K-noncollapsed for some K. EXERCISE 6.52. Show that for any Riemannian manifold (Mn,
M, and have
K
< W n , there exists p (x) > 0 such that IRml ~ r- 2
and
g) , x E
for every r E (0, P (x)], we
in B (x, r)
Vol B(x, r)
---'-----'> K. rn SOLUTION TO EXERCISE 6.52. This follows from the facts that lim r2 sup IRml = 0 r-+O
B(x,r)
and lim VoIB(x,r) = rn
Wn.
r-+O
REMARK 6.53. In some sense Exercise 6.52 is a local version of Remark 6.46(2).
5.1.2. K-noncollapsing and injectivity radius lower bound. We now show that K-noncollapsing and a lower bound of the injectivity radius are equivalent. LEMMA 6.54. Let (Mn,g) be a complete Riemannian manifold and fix p
E (0,00]. (i) If the metric 9 is not K-collapsed below the scale p for some K > 0, then there exists a constant 8 = 8 (n, K) which is independent of that for any x E M and r < p, if IRml ~ r- 2 in B (x, r), then inj (x) ~ 8r. (ii) Suppose that for any x E M and r < p with IRml ~ r- 2 in B (x, r) we have inj (x) ~ 8r for some 8 > O. Then there exists a constant K = K (n, 8) , independent of p and g, such that 9 is not K-collapsed below the scale p. p and
9 such
PROOF. (i) Let B (x, r) be a ball satisfying IRml ~ r- 2 in B (x, r) for some r ~ p. Consider the metric r-2g on B (x,r) = B r -2 g (x, 1). Since 9 is not K-collapsed on B (x,r), we have IRmr -2 g I ~ 1 in B r -2 g (x, 1) and Voig B (x, r) n ~ K. r By a result of Cheeger, Gromov, and Taylor (see Theorem A.7), there exists 8 = 8 (n, K) such that injr-2g (x) ~ 8. Hence inj (x) ~ 8r. (ii) Again let B (x, r) be a ball satisfying IRmI ~ r- 2 in B (x, r) for some r ~ p, and consider the metric r-2g on B (x, r) = B r -2 g (x, 1). We have IRmr -2 g l ~ 1 and injr-2g (x) ~ 8. By the Bishop-Gromov volume (or Volr -2 g B r -2 g (x, 1) =
5. NO FINITE TIME LOCAL COLLAPSING
255
Rauch) comparison theorem (comparing (Br-2g (x, 1), r-2g) with the ball of radius 8 in the unit sphere (1)), there exists K = K (n, 8) such that Voig B (x, r) n = Volr -2 g B r -2 g (x, 1) ~ K. r
sn
o 5.1.3. K-noncollapsing in Ricci flow. DEFINITION
6.55. We say that a complete solution (Mn,g(t)) , t
E
[0, T), to the Ricci flow, where T E (0,00], is K-noncollapsed below the
scale p if for every t E [0, T), 9 (t) is K-noncollapsed below the scale p. If M is closed, Tl < 00, and Co ~ sUPMx[O,Td IRml < 00, then, using the metric equivalence e- 2(n-l)Co 9 (0) ::; 9 (t) ::; e2(n-l)Co g (0) for t E [0, T 1 ),16 we see that for every p E (0,00) there exists K = K (n, Co, 9 (0), TI, p) > such that the solution 9 (t) is K-noncollapsed below the scale p. Hence we are interested in K-noncollapsing near T when the solution forms a singularity at time T. We shall see that when T < 00 and M is closed, Perelman's monotonicity of entropy implies that for all p > the solution is K-noncollapsed below the scale p for some K = K(n, 9 (0), T, p) > 0. In §4.1 of [297] Perelman also gave the following.
°
°
DEFINITION 6.56 (Locally collapsing solution). Let (Mn,g(t)) , t E [0, T), be a complete solution to the Ricci flow, where T E (0,00]. The solution 9 (t) is said to be locally collapsing at T if there exists a sequence of points Xk EM, times tk -+ T, and radii rk E (0,00) with rVtk uniformly bounded (from above) such that the balls B9(tk) (Xk' rk) satisfy (1) (curvature bound comparable to the radius of the ball)
IRm [g (tk)]1 ::; r k2
in
B9(tk)
(Xk' rk),
(2) (volume collapse of the ball) . Volg(tk) hm
k-+oo
B9(tk)
rk'
(Xk' rk)
=0.
EXERCISE 6.57. It is interesting to consider solutions to the Ricci flow which are defined on a time interval of the form (0, T) with the curvature becoming unbounded as t -+ 0+. For example, consider an initial metric 90 on a surface which is Coo except for a conical singularity. We expect a smooth solution 9 (t) of the Ricci flow to exist on some time interval (0, T) with 9 (t) -+ 90 as t -+ 0+. It is interesting to ask if solutions on closed manifolds can locally collapse as t -+ 0+. In view of this, for a solution defined on (0, T) , formulate the notion of locally collapsing at time 0.
16For the proof of this metric equivalence, see Corollary 6.50 on p. 204 of Volume One. The argument there is essentially repeated in the proof of the inequalities in (3.3) of this volume.
256
6.
ENTROPY AND NO LOCAL COLLAPSING
5.2. The no local collapsing theorem and its proof. 5.2.1. No local collapsing theorem and little loop conjecture. One of the major breakthroughs in Ricci flow is the following. THEOREM 6.58 (No local collapsing-A). Let g(t), t E [0, T), be a smooth solution to the Ricci flow on a closed manifold Mn. If T < 00, then for such that g(t) is K,any p E (0,00) there exists K, = K,(n,g(O),T,p) > noncollapsed below the scale p for all t E [0, T).
°
We shall prove this theorem in the next subsubsection. Actually Theorem 4.1 of [297] states the result a bit differently. THEOREM 6.59 (No local collapsing-B). If M is closed and 9 (t) is any solution on [0, T) with T < 00, then 9 (t) is not locally collapsing at T. REMARK 6.60. We leave it as an exercise to show that Theorems 6.58 and 6.59 are equivalent.
Hamilton's little loop conjecture says the following (see §15 of [186]). Let (M n , g (t)) , t E [0, T), be a smooth solution to the Ricci flow on a closed manifold. There exists 8 = 8 (n,g (0)) > such that for any point (x, t) EM x [0, T) where
°
1 IRm (g (t))1 ::::; W2
in
Bg(to)
(x, W)
for some W > 0, we have injg(t) (x) ~
8W.
Note that the role of the positive number K, in the definition of K,noncollapsed is similar to the role of 8 in the injectivity radius lower bound which is used in the statement of Hamilton's little loop conjecture. Rephrasing the little loop conjecture (LLC) a little differently, we have the following equivalence between no local collapsing (NLC) at T and the little loop conjecture. LEMMA 6.61 (NLC and LLC are equivalent). Let (Mn, 9 (t)) , t E [0, T), be a smooth complete solution to the Ricci flow where T E (0,00]. The following two statements are equivalent.
°
(i) (Little loop conjecture) For any C > there exists 8 > if (x, t) E M x [0, T) and W E (0, v'ct] satisfy 1 IRm(t)1 ::::; W2
in
Bg(t)
°
such that
(x, W),
then (6.81)
injg(t)
(x) ~
8W.
(ii) (No local collapsing) The solution g(t) is not locally collapsing at T.
5. NO FINITE TIME LOCAL COLLAPSING
257
PROOF. (i) ==> (ii). We prove (ii) by contradiction. Suppose 9 (t) is locally collapsing at T. Then there exists a sequence of times tk / T and a sequence of metric balls B9(tk)(Xk, rk) such that r2 (1) C for some C < 00,
t: : ;
(2) IRm(g (tk))1 ::; r-;;2 in Bg(tk) (Xk' rk), (3) Volg(tk) Bg~k) (Xk' rk) '\. 0 as k ~ 00. rk Hence by (i) we have injg(tk) (Xk) ~ ork for all k, where 0> 0 is independent of k. By Lemma 6.54(ii), the volume collapsing statement (3) above cannot be true, a contradiction. (ii) ==> (i). We also prove (i) by contradiction. If (i) is not true, then there exists C > 0 and a sequence of points and times (Xk' tk) EM x [0, T) and Wk E (0, JCtk] satisfying 1 . IRm (tk)1 ::; W2 In Bg(tk) (Xk' Wk) k
and
injg(tk) (Xk) Wk
'\.
O.
k) Lemma 6.54 (1.).Impl'Ies t h at Volg(tkJ B9(tk)(Xk,W witk '\. 0 as k ~ 00. T hus 9 (t ) IS. locally collapsing at T and we have a contradiction. The lemma is proved.
o It follows from Theorem 6.59 and Lemma 6.61 that Hamilton's little loop conjecture holds for solutions of the Ricci flow on closed manifolds forming finite time singularities. COROLLARY 6.62. Let g(t), t E [0, T), be a smooth solution to the Ricci flow on a closed manifold Mn. If T < 00, then the little loop conjecture holds. That is, for any C > 0 there exists 0 > 0 such that if (x, t) E M x [0, T) and W E (0, JOt] satisfy 1 IRm (t)1 ::; W2 in Bg(t) (x, W),
then we have injg(t) (x)
~
oW.
The little loop conjecture illustrates the essence of no locally collapsing from the injectivity radius perspective. For convenience we give the following DEFINITION 6.63 (Local injectivity radius estimate). We say that a complete solution (Mn,g(t)), t E [O,T), to the Ricci flow satisfies a local injectivity radius estimate if for every p E (0,00) and C < 00, there exists c = c (p, C, 9 (t)) > 0 such that for any (p, t) E M x [0, T) and r E (0, p] which satisfy IRm (., t)1 ::; Cr- 2 in Bg(t) (p, r),
258
6.
ENTROPY AND NO LOCAL COLLAPSING
we have injg(t) (p) 2: cr. Corollary 6.62, i.e., Perelman's no local collapsing theorem, implies that if (Mn,g (t)), t E [0, T), is a solution of the Ricci flow on a closed manifold with T < 00, then g(t), t E [0, T), satisfies a local injectivity radius estimate. 5.2.2. Proof of No Local Collapsing Theorem 6.58. The idea of the proof is that if a metric 9 is K-collapsed at a point x at a distance scale r for K small and r bounded, then W (g, f, r2) is negative and large in magnitude, e.g., on the order of log K, for f concentrated in a ball of radius r centered at x. This contradicts the monotonicity formula for p, (g (t), r (t)) . PROOF OF THEOREM 6.58 ASSUMING PROPOSITION 6.64. We shall say that r is the (space) scale of p, (g, r2); the justification for this terminology occurs below. Since T /2 < T, by the remarks after Definition 6.55, there exists KO = KO (n, 9 (0) , T, p) > 0 such that g(t) is Ko-noncollapsed below the scale p for all t E [0, T /2]. On the other hand, if t E [T /2, T), then for any 0 < r ::; p, we have t + r2 E [T /2, T + p2), and by the monotonicity formula (6.61), we have p, (g (t) , r2) 2: p, (g (0) , t (6.82)
2:
+ r2)
inf TE[T/2,T+p2]
p,(g(O),r)
~
-CI(n,g(O),T,p) >-00
since T < 00. In summary, by the monotonicity formula, since the p,invariant of the initial metric is bounded from below at scales bounded from above and below, the p,-invariant of the solution after a certain amount of time (say T /2) is bounded from below at all bounded scales. The theorem will follow from the important observation that if a Riemannian metric is K-collapsed at some scale r for K small, then its p,-invariant is negative and large in magnitude at the time scale r2 (see Proposition 6.64 below). If x E M, t E [T /2, T), and r E (0, p] are such that Rmg(t) 1 in 2 2 Bg(t) (x, r) , then RCg(t) 2: -Cl (n) r- and Rg(t) ::; Cl (n) r- in Bg(t) (x, r) , where Cl (n) = n (n - 1) . So by (6.82) and (6.83),
I
-CI(n,g(O),T,p)::;p, (g(t),r
2)
::;log
Volg(t) Bg(t) (x, r)
r
n
I ::;
+Cdn,p).
We conclude that Volg(t) Bg(t) (x, r)
rn
> -
(
Kl n, 9
(0)
"
T
)
P
>0
,
where Kl (n, 9 (0), T, p) = e- C1 (n,g(0),T,p)-C2(n,p). The theorem follows with the choice K(n, g(O), T, p) ~ min {KO, Kl}. 0 Now we turn to bounding the p,-invariant from above by volume ratios in a Riemannian manifold. PROPOSITION 6.64 (p, controls volume ratios). Let p E (0,00). There
exists a constant C2 = C2 (n, p) < 00 such that if
(Mn, g)
is a closed
5. NO FINITE TIME LOCAL COLLAPSING
259
Riemannian manifold, pENt and r E (0, pJ are such that Rc ~ and R ~ Cl (n) r- 2 in B (p, r), then ~ 2) J-L ( g,r
(6.83)
~log
Vol B (p, r) r
n
-Cl
(n) r- 2
+C2 (n,p).
That is,
In particular, if for some K > the scale r, then
°
and r E (0, pJ the metric f) is K-collapsed at
J-L (f), r2) ~ log K Proof. As in (6.41) with
T
+ C2 (n, p) .
= r2, define the positive function
W
by
(6.84) From the definition (6.56) of J-L as an infimum of W, we have by rewriting W (f),f,r 2) in terms of W (compare to (6.42)), (6.85)
J-L (f), r2)
~ 1M r2 (41V'w12 + Rw 2) dJ-L + 1M (f -
n) w 2dJ-L
~ J( (f),w,r2) ,
where iM w 2dJ-L = 1 and f = -2 log w- ~ log (47rr2) . While making the convention that f (y) w 2 (y) = when w (y) = 0, we claim that (6.85) holds for nonnegative Lipschitz functions w satisfying iM w 2dJ-L = 1. To see this claim, first by (6.56) we know that (6.85) holds for positive Lipschitz functions w satisfying i M w 2dJ-L = 1. Now given any nonnegative Lipschitz function w satisfying iM w 2dJ-L = 1, define for c E (0,1),
°
we;~Ce;(w+c),
where the constant Ce; is defined by iM w;dJ-L = 1. Clearly lime;->oCe; = 1, We; is a positive Lipschitz function, and hence J-L (f), r2) ~ J( (f), We;, r2) for each c E (0,1). Using lime;->oclogc = and fe; = -2 log We; - ~ log (47rr2) in the definition of K (f),we;,r 2) , we have lime;->oK (f), We;, r2) = K (f),w,r 2). The claim is proved. Now let ¢ : [0,00) --t [0, 1J be a standard cut-off function with ¢ = 1 on [0, 1/2J , ¢ = on [1,00), and WI ~ 3. Assume pENt and r E (0, pJ are such that we have the curvature bounds Rc ~ -Cl (n) r- 2 and R ~ Cl (n) r- 2 in B (p, r) for Cl (n) = n (n - 1) . We make a judicious choice for f so that the RHS of (6.85) reflects the local geometry at p with respect to the metric f). In particular, let
°
°
(6.86)
6. ENTROPY AND NO LOCAL COLLAPSING
260
where dg(x,p) is the distance function and the constant c = c(n,g,x,r) is chosen so that fM w 2df.l = 1. Note that w is a Lipschitz function and definition (6.84) implies (6.87) (We abuse notation and write ¢ (x)
to the volume ratio LEMMA
Vol ~n(p,r)
= ¢ (dg(;,P))
.)
The constant c is related
by the following.
6.65. There exists C3 (n, p) <
00
such that for any r
E
(0, p),
C ( ) < < I Vol B (p, r) 3 n, P _ c _ og n . r r Proof of the lemma. (1) Since fM w 2df.l = 1, Iog Vol B n(p, r)
(6.88)
-
eC = (411T2rn/21M ¢ (d
(6.89)
(:,P)) 2df.l (x).
Applying ¢ ::; 1 and supp (w) C B (p, r), we have
eC
::;
(47rr2) -n/2 Vol B (p, r) ,
which implies
c::; Iog
(6.90)
Vol B (p, r) r
n
.
(2) On the other hand, since ¢ = 1 on [0,1/2] , by (6.89) we have c 2:
I VolB (p,r/2) -"2n Iog (4) 7r + og rn .
Since Rc 2: -Cl (n) r- 2 in B (p, r) and r ::; p, by the Bishop-Gromov relative volume comparison theorem, there exists C4 (n, p) < 00 such that Vol B (p, r') ::; C4 (n, p) VolB (p, r /2) . Thus c 2: -C3 (n, p)
(6.91)
+ log
VolB (p, r) r
n
.
This completes the proof of the lemma. To prove the proposition, we estimate the two terms
1M r2 (4JVwJ 2 + Rw 2) df.l + 1M fw 2df.l = K (g, w, r2) + n on the
RHS
of (6.85) separately. First we have
1M r2 Rw 2df.l ::;
Cl
(n)
261
5. NO FINITE TIME LOCAL COLLAPSING
c B(p,r),
since r2R::; cdn) in B(p,r), supp(w) (6.86), we have a.e.,
and fMW2df..L
=
1. By
r2 lV'wl2 ::; (471"r2) -n/2 e- c 1¢' 12 ::; 9 (471") -n/2 e~c r
<
9 (471") -n/2 eC3 (n,p)
using WI ::; 3, lV'dl we have
VolB(p,r)
= 1 a.e.,
,
and (6.88). Since lV'wl has support in B (p, r),
1M 4r21V'w12 df..L ::; 9 (471")-n/2 eC3 (n,p). We conclude that the energy part of the RHS of (6.85) is bounded from above:
1M r2 (41V'wI2 + Rw 2) df..L ::;
Cl
(n)
Now we consider the entropy part. Since have
+ 9 (471")-n/2 eC3 (n,p). f
= c -log (¢2) , by (6.87), we
IM fW2d f..L::; 1M (c-log (¢2)) w 2df..L
= cSince -x log x ::;
1/e and ¢
(411T2) -n/2 e- c 1M log (¢2) ¢2df..L.
has support in B (p, r) , we have l7
1M log (¢2) ¢2df..L 2:
-~ Vol B (p, r).
Hence
We conclude that ~ 2) VolB(p,r) f..L ( g,r ::;log n +Cdn,p), r
where C 2 (n, p) ~ (9 + ~) (471") -n/2 eC3 (n,p) part of the proposition. 17Our
convention is 0 . log 0
=
+ C1 (n).
0 since lim._o clog €
This proves the first
= o.
Also, log
B (p, r/2) , so in fact fM log (4)2) 4>2dp, ~ -~ Vol (B (p, r) - B (p, r/2)).
(4)2) =
0 in
262
6.
ENTROPY AND NO LOCAL COLLAPSING
The second part follows from that fact that if for some K > 0 and r > 0 the metric 9 is K-collapsed at the scale r, then there exists p E M such that _Vo_l_B~(p~,-.-!..r)
rn
< - K.
The proposition is proved. 0
(Mn, 9) ,
EQUIVALENT PROPOSITION. For every Riemannian manifold T E (0,00), C1 and C2, there exists c = c (n, T, C1, C2) > 0 such that if for some r E (0, VT] and A < 00 we have J.L (9, r2) ~ -A, then for any p E M with Rc ~ -C1r- 2 in B (p, r) and R ~ C 2r- 2 in B (p, r), we have
VolB (p, r) where K =
C
~
Kr n ,
(n, T, C1, C 2) e- A .
PROOF. The contrapositive is Proposition 6.64. In particular, the conT E (0,00), C1 and C2, there exists a trapositive is: for every
(Mn, 9) ,
constant C = C (n, T, C1, C 2) such that if for some r E (0, VT] and K > 0 there exists p E M with Rc ~ -C1r- 2 in B (p, r), R ~ C 2r- 2 in B (p, r) , and Vol B (p, r) < Kr n , then J.L (9, r2) < log K + C. 0 REMARK 6.66. From examining the proof of Proposition 6.64, in the Equivalent Proposition we can replace the condition Rc ~ -C1 r- 2 in B (p, r) by VolB (p, r) ~ C 1 VolB (p, r/2) since the only place where we used the Ricci curvature lower bound is for the relative volume comparison. That is, there exists Co = CO (n, T, Cd > 0 such that VolB (p, r) ~ Kor n , where KO = CO (n, T, C 1 ) e- A • By assuming a lower bound on 11 (9) instead of J.L the lower bound of Ricci curvature assumption.
(9, r2) , we may remove
PROPOSITION 6.67. For every Riemannian manifold
(Mn, 9),
T E
(0,00), and C 1, there exists c = c (n, T, Cd > 0 such that if for some r E (0, VT] and A < 00 we have 11 (g) ~ -A, then for any p E M with R ~ C 1r- 2 in B (p, r), we have VolB (p, r) where K =
C
~
Kr n ,
(n, T, C 1) e- A .
PROOF. If Vol B (p, r) ~ 3n Vol B (p, r /2), then by the above remark, the proposition holds. So we assume that Vol B (p, r) > 3n Vol B (p, r /2) . Since lims ...... o+ vo~~;~,s) = 1, there exists kEN such that Vol B (p, r /2k) ~ 3n Vol B (p, r /2 k +!) and Vol B (p, r /2i) > 3n Vol B (p, r /2i+1) for all 0 ~ i <
5.
NO FINITE TIME LOCAL COLLAPSING
k. We can apply Remark 6.66 to the ball B ~o
263
(p, r 12k) and get Vol B (p, r 12k)
;:::
(r 12kt . Hence Vol B (p, r) > 3n Vol B (p, r 12)
> 3n (k-l) Vol B ;::: 3n (k-l) ~o
(p, r 12k)
(r 12k)
n
~Orn. - 2n
>
Hence the proposition holds with c (n, T, G1 ) 6.66 we have ~O = CO (n, T, G1 ) e- A .
= co(n2~,cl) , since
by Remark 0
5.3. Application of ~-noncollapsing to the analysis of singularities. We present some applications to end this section. In the next section we give an improvement of the no local collapsing theorem. 5.3.1. Existence of finite time singularity models. An injectivity radius estimate (Corollary 6.62) implies that one can apply Hamilton's CheegerGromov-type compactness theorem to obtain the existence of singularity models for singular solutions with finite singularity time. (Recall that the definition of a singularity model is given in Remark 1.29.) THEOREM 6.68 (Existence of singularity models). Let g(t), t E [0, T), be a smooth solution to the Ricci flow on a closed manifold Mn with T < 00. Suppose that there exists a sequence of times ti /" T, points Pi E M, and a constant G < 00 such that (6.92) (6.93)
Ki =§: IRm g(ti) (Pi) I ---+
IRmg(t)(x)1
~ GKi
00,
for all x E M and t
< ti.
Then there exists a subsequence of the sequence of dilated solutions18 gi(t) =§: Ki . 9 (ti
+
;J
such that (Mn, gi(t),pd converges to a complete ancient solution to the Ricci flow (M~,goo(t),poo) in the sense of Goo-Cheeger-Gromov convergence. Furthermore there exists ~ > 0 such that 900 (t) is ~-noncollapsed on all scales. PROOF. By Perelman's no local collapsing theorem, we have an injectivity radius estimate at the points Pk with respect to the metrics gk (0). Hence by Hamilton's compactness theorem (Theorem 3.10), there exists a subsequence such that (Mn,gi(t),Pi) converges to a complete ancient solution (M~, goo(t),poo) to the Ricci flow. Since 9 (t) is ~-noncollapsed on the scale JT for all t E [0, T), we have gi(t) is ~-noncollapsed on the scale 18 As usual we denote a subsequence of i still by i rather than ij to simplify our notation.
264
6.
ENTROPY AND NO LOCAL COLLAPSING
v'KiT for all t E [-Kiti,KdT-ti))' Since limi ...... oov'KiT = 00, goo(t) is K-noncollapsed on all scales from Lemma 6.48 (regarding a limit property of sequences of K-noncollapsed solutions). 0 5.3.2. Ruling out the cigar as a finite time singularity model. Finally, we observe that Perelman's no local collapsing theorem implies that the cigar (product with any flat solution like lRn - 2 or a torus Tn-2) cannot be a limit of dilations about a finite time singularity as in Theorem 6.68. This is because the cigar (product with any flat solution) is not K-noncollapsed on all scales for any K > O. An easy way to see this is that the cylinder S1 x lR is a limit of the cigar. Clearly, S1 x IR (product with any flat solution) is not K-noncollapsed on all scales for any K > O. By the property of K-noncollapsed being preserved under limits, this implies the same for the cigar (product with any flat solution). 6. Improved version of no local collapsing and diameter control
In this section we give a proof of Perelman's improvement of his no local collapsing theorem to the case where one assumes, in the ball to be shown to be noncollapsed, only the scalar curvature has an upper bound. We also present the work of Topping [357] on diameter control. We end this section with a variation on the proof of Perelman's no local collapsing theorem. 6.1. Improved version of no local collapsing. We first revisit and revise Proposition 6.64. Let (Mn, be a closed Riemannian manifold and
g)
r > O. Again we shall consider the inequality (6.85) for J.l (g, r2) and the test function w defined by (6.86). It is easy to see that the proof of Lemma 6.65 yields the following, where the estimate now involves Vol B (p, r /2) , which we had previously estimated in terms of Vol B (p, r) under a local Ricci curvature lower bound assumption. LEMMA 6.69 (c and the volume ratio). The constant c in (6.86) satisfies the following bounds: 1 < (4 2)-n/2 -c < 1 VoIB(p,r) 7rr e - VoIB(p,r/2),
(694) . Equivalently,
(6.95)
-~log(47r)+IOg VOIB;~,r/2) ~c~-~IOg(47r)+log VOI~~p,r).
Using the above lemma, we obtain the following (the proof is similar to the proof of Proposition 6.64). PROPOSITION 6.70 (Bounding J.l by the scalar curvature and volume ratio). The J.l-invariant has the following upper bound in terms of local geometric quantities.
For any closed Riemannian manifold (Mn,
g) , point
6.
PE
M,
(696)
.
IMPROVED VERSION OF NLC AND DIAMETER CONTROL
265
and r > 0, we have
J-l g,r
where R+
r
r2 R+dJ-l) VolB(p,r) ( 36 JB(p,r) VolB(p,r) - og rn + + VolB(p,r) VolB(p,r/2)'
(A 2) <1 ~
max {R, o} is the positive part oj the scalar curvature.
REMARK 6.71. Proposition 6.64 follows from the above statement and the Bishop-Gromov relative volume comparison theorem. PROOF. We estimate each of the three terms on the RHS of (6.85) separately, where w is chosen by (6.86). Since WI ::; 3, I\i'dl = 1 a.e., and supp (
4r21M
l\i'wl2 dJ-l ::; 41M (47rr2) -n/2 e- c I
VolB(p,r/2)
where we used (6.94) to obtain the last inequality. We also have
r2 1M Rw 2dJ-l = r21M R (47rr2) -n/2 e-C
<
r
r2
- Vol B (p, r /2)
(~) 2 dJ-l
R+dJ-l,
J B(p,r)
where we used the fact that
1M log (w 2) w 2dJ-l since iM w 2dJ-l upper bound:
~ -log Vol B (p, r)
= 1 and supp (w 2) C B (p, r) . Hence the third term has the
1M Jw 2dJ-l = 1M
(-~ log (47rr2)
-log (w2)) w 2dJ-l::; log VOl~~p,r).
Summing the above three inequalities yields the proposition.
o
Motivated by the expression on the RHS of (6.96), we define the maximal
(Mn, g) by
function on a closed Riemannian manifold 2
MR(p, r) ~ sup y, lB8 ( ) O<s::;r
Since lim
s--+O
p, S
0
S2
Vol B (p, 8)
1
B(p,s)
1
B(p,s)
R+ dJ-l.
R+ dJ-l = 0,
266
6. ENTROPY AND NO LOCAL COLLAPSING
the quantity MR{P, r) is a well-defined finite number for 0 < r < if rl ::; r2, then MR{P, rl) ::; MR{p, r2). By (6.96), we have (6.97)
A (
J.t g, s
2)<1 VolB{p,s) - og sn
+
{36
+
00.
Clearly,
M ( )) VolB{p,s) R p, s VolB (p, s/2)"
Interestingly, the factor v~fk~~:~j~) on the RHS above does not prevent one from estimating volume ratios by J.t and MR only under a scalar curvature upper bound. We shall prove the following.
(JVtn, g)
6.72 (Bounding volume ratios by Vr and M R). If is a closed Riemannian manifold and 0 < s ::; r, then PROPOSITION
where
In particular, if R ::; cdn) r- 2 in B (p, r), then MR (p, r) ::; cdn) , so that
Vol B (p, s) > -3n (36+C1 (n)) vr(Y) sn - e e. 6.73. Given p E is nonincreasing. REMARK
JVt, the function r
~ e-3n36evr(Y)e-3nMR(p,r)
If v~fkB:,~s2) ::; 3n (in this case we say that at p the volume doubling property holds at scale s), then (6.97) implies PROOF.
and the estimate follows. VolB(p s) 3n h . l' If Vol B(p,~/2) ~ ,t en smce Imk->oo
VolB p,s/2 k Vol B p,s/2k+ 1 )
1'lJ h h VolB{p,s/2k) 3n d VOlB{P,S/2i~ k E 1~ sue t at VolB{p,S/2k+ 1 ) : : ; an VolB(p,s/2 +) 1
Applying (6.97) to B
>
2n
=
3n
&
h . ,t ere eXists
lor a
110
. k ::; '/, < .
(p, s/2 k ) , we get
k)2) VolB(p,s/2k) J.t ( g, ( s/2 ::; log {s/2kt
)) + ( 36 + MR (p,k s/2
n
·3.
6.
IMPROVED VERSION OF NLC AND DIAMETER CONTROL
267
Hence
VoIB(p,s) > (~)n VoIB(p,s/2) Sn 2 (s/2t
> (~)nk VolB (p,s/2 k) -
(s/2k)n
2
2: (~) nk e-3n36eJL(g,s2/22k)e-3nMR(p,s/2k) 2:
(~) nk e-3n36evr(g)e-3nMR(p,r).
o
The theorem again follows in this case.
Now we apply Proposition 6.72 to solutions of the Ricci flow and obtain the following improvement of the no local collapsing theorem. THEOREM 6.74 (No local collapsing theorem improved). Let (Mn, 9 (t)) , t E [0, T), be a solution to the Ricci flow on a closed manifold with T < 00 and let p E (0,00). There exists a constant K, = K, (n, 9 (0) , T, p) > Osuch that if p E M, t E [0, T), and r E (0, p] are such that R ::;
r- 2 in
Bg(t)
(p, r),
then Volg(t) Bg(t) (p, s) >
_....::....0..."----.::....0....:_ _
for all
°< s ::;
sn
-
K,
r.
PROOF. By (6.61) and the definition of Vn we have Vr
(9 (t)) 2: v v'p2+T (9 (0))
for r E (0, p] and t E [0, T). Then the theorem follows from Proposition 6.72. 0 6.2. Diameter control. In this subsection we show how ideas related to the previous subsection can be used to obtain a diameter bound for solutions of the Ricci flow in terms of the L(n-l)/2- norm of the scalar curvature. This result is due to Topping [357] and our presentation essentially follows his ideas. Recall that Proposition 6.72 implies that if (,Mn, is a closed Riemannian manifold, then for any p E ,Mn and
(6.98)
VoIB(p,s) > sn -
°<
g)
s
< 00, we have
e-3n36ev(g)e-3nMR(P,r)
,
using v (g) ~ infrE[o,oo) J.L (g, 1") ::; Vr (9) for all r. Recall that by Corollary 6.34, if the lowest eigenvalue is positive, i.e., A(9) ~ Al (-4~g + Rg) > 0, then v (g) > -00.
268
6.
ENTROPY AND NO LOCAL COLLAPSING
THEOREM 6.75 (Topping). Let n 2': 3 and let mannian manifold with 1/ (9) > -00. Then diam
(Mn, 9)
be a closed Rie-
(M, 9) ~ max {~~ ,6e3n37 e-V(g)} 1M R:21 dp"
where Wn is the volume of the unit n-ball in Rn.
PROOF. Let 8 (9) ~ min {~, ~ (9) e- 3n } , where ~ (9) ~ e-3n36ev(g). Note that lims->o Vol~Jp,s) = Wn and lims->oo Vol~;p,s) = 0 since M is closed. Hence for any point p E M, there exists s(p) > 0 such that Vol~~'t~(p))
=
8 (9) and Vol~;p,s) 2': 8 (9) for s E (0, s(p)]. Applying inequality (6.98), we have
MR(p, s(p)) 2': l.
(6.99)
This implies there exists s' (p) ~ s(p) such that
(s'(p))2 VoIB(p, s'(p))
r
>1
R d
JB(p,s'(p))
+ p, -
.
Applying the Holder inequality, we have
VoIB(p,s'(p)) < (s' (p))2 -
~
r
R d
J B(p,s'(p))
+
P, 2
(r
R:21 dP,)
n-l
[VoIB(p,s'(p))]
~=i ,
J B(p,s'(p))
1
so that
VoIB(p,s'(p)) n;-l --:---:-:-":-:------7-':"':" ~ R+ dp,. (s'(p))n-l B(p,s'(p)) We have proved that for every p EM, there exists s' (p) > 0 such that 8 (A) '( ) < VoIB(p, s'(p)). '( ) 9 s P _ (s'(p))n s p =
VoIB(p, s'(p)) 1 < (s'(p)t-
1
n-l
R+2 dp"
B(p,s'(p))
where the first inequality follows from the definition of s(p) and s'(p) ~ s(p). To finish the proof of the theorem, let "f be a minimal geodesic whose length is the diameter of 9). One can show that there exists a countable
(M,
(possibly finite) number of points Pi E "f such that B (pi, s' (pd) are disjoint and cover at least 1/3 of"f (Vitali covering-type theorem). Then
~diam(M'9) ~ L2s'(Pi) ~ i
I
8 2(A) L 9
~ 8 ~9) 1M R:21 dp,.
i
r
JB(Pi,S'(Pi))
R:21 dp,
6.
IMPROVED VERSION OF NLC AND DIAMETER CONTROL
269
The theorem follows from plugging in the definition of 8 (g) .
D
Now we can apply Theorem 6.75 to the Ricci flow and obtain the following. COROLLARY 6.76. Let n ~ 3 and let (Mn, g(t)), t E [0, T), be a solution of the Ricci flow on a closed manifold with T < 00. Assume that A(g(O)) > O. Then there exists C = C(n,g(O)) > 0 such that
(6.100)
diam (M,g(t))
~ C 1M R+ (t) n 2
1
d/-lg(t)·
PROOF. Note that by the monotonicity of the A-invariant we have A(g(t)) ~ A(g(O)) > 0,
and hence the theorem is applicable. Now the corollary follows from ~ l/(g(O)).
1/
(g (t))
D
6.3. A variation on the proof of no local collapsing. In this subsection we give a modified proof that the no local collapsing theorem follows from entropy monotonicity using a local eigenvalue estimate. We also give a heat equation proof of a less sharp form of the global version of this eigenvalue estimate. 6.3.1. Modified proof of no local collapsing theorem. Recall Cheng's sharp upper bound for the first eigenvalue Al of the Laplacian -~ on balls with a lower bound on the Ricci curvature [92].
THEOREM 6.77 (Cheng, local eigenvalue comparison). Let (Mn,g) be a complete Riemannian manifold with Rc(g) ~ -(n - l)g. Then for any
PEM, (6.101) where BlHIn (1) is the open ball of radius 1 in hyperbolic space lHIn of sectional curvature -1. Here Al denotes the first eigenvalue of the Laplacian with the Dirichlet boundary condition.
EXERCISE 6.78. Suppose (Mn,g) is a complete Riemannian manifold with Rc(g) ~ (n - l)Kg, where K ~ O. Given r > 0, determine an upper bound for Al (B(p, r)) in terms of the corresponding model space. We now give the modified proof of no local collapsing using the above eigenvalue estimate and Jensen's inequality. Given the monotonicity of /-l, the first proof we presented relies on inequality (6.83) giving an upper bound for /-l in terms of the volume ratio; it is this inequality for which we give a second proof. Recall from (6.85) that we have for a Riemannian manifold
ENTROPY AND NO LOCAL COLLAPSING
6.
270
J-t(9, r2)
~ 1M r2(41V'wI2 + Rw 2) dJ-t
- 1M (log (w 2) + ~ log(47rr2) + n) w2 dJ-t for all w with fM w 2dJ-t = 1. Using (6.47), we have for any w with supp (w) B(p, r) and fM w2dJ-t = 1 that
c
- 1M log(w 2)w2dJ-t ~ logVolB(p,r). By assumptions Rc (g) ~ - (n - 1) r- 2 and R ~ n (n - 1) r- 2 in B (p, r) ,19 we have for any w with supp (w) c B(p, r) and fM w 2dJ-t = 1,
1M Rw2 dJ-t ~ n (n -
1) .
Let 9 = r- 2g; then Bg(p, 1) = B(p,r) ~ Bg(p,r). By Theorem 6.77 and the Rayleigh principle for eigenvalues, inf
J:
f . lV'wl2 dJ-t-
supp(w)CBg(P,I)
2d M w J-tg
=.AI (Bg(p, 1)) ~ .AI (Bnnn (1)).
9
Hence we have
Therefore log Vol ~ip, r)
~ J-t(g, r2) -
4.AI (Bnnn (1)) - n (n - 2)
+ ~ log(47r).
This provides the needed estimate to replace (6.83).
6.3.2. A heat equation proof of a global eigenvalue estimate. We now recall a global version of Cheng's Theorem 6.77. THEOREM 6.79 (Cheng's eigenvalue estimate, global).
complete noncompact Riemannian manifold with Rc (g) .A 1
(_~) < (n -
4
~
If (Mn,g)
is a
-(n - l)g, then
1)2
It turns out that a weaker version of this estimate, i.e., .AI ~ n(n4-1) , can be proved using the energy/entropy computation of Perelman for the fixed metric case; we give this proof below. First we state a formula which is implicit in [283]. 19We choose these constants for our curvature bounds since they are implied by
_r- 2 :$ sect :$ r- 2 •
6.
IMPROVED VERSION OF NLC AND DIAMETER CONTROL
271
LEMMA 6.80. Let U be a positive solution to the heat equation
(:t-A)U=O on a fixed Riemannian manifold then
(Mn,g). If f = -logu, that is,
U
= e- f ,
(6.102)
ftJ = Af -1V'fI 2 , we calculate ~ 1M IV' fl2u dp, = ~ 1M (AI) udp,
PROOF. (1) Using
= 1M (2Af -IV' f12) Audp, = 21M (AfAu+ V'iV'jfV'dV'ju)dp,. Now integrating by parts yields
1M Af Audp,
= - 1M V' f . V' Audp, = - 1M (V' f·
A V'u - !4jV'dV'ju) dp,
= 1M (V'V' f . V'V' U
-
u!4j V'dV' j I) dp,
= 1M (-u lV' i V'jfI2 - V'iV'jfV'dV'ju - u!4jV'dV'jf) dp,. The lemma immediately follows from combining the above two formulas. (2) Alternatively, one can integrate the following formula to get (6.102) (see (2.1) and Lemma 2.1 in [283]):
(:t - A) (u (2Af -IV' fI2))
= -2u lV'i V'jfI2 - 2u!4jV'dV'jf,
which follows directly from the calculation:
:t (2Af - IV' f12) = 2A (Af - IV' f12) - 2V' f . V' (Af -IV' f12) = A (2Af -1V'fI2) - 2V'f· V' (2Af -1V'fI2)
- A IV' fl2
+ 2V' f
. V' Af
= A (2Af -1V'fI2) - 2V'f· V' (2Af -1V'fI2)
- 21V'i V'jfI2 - 2!4jV'dV'jf.
o Now we can prove the following weaker version of Cheng's Theorem 6.79.
6.
272
ENTROPY AND NO LOCAL COLLAPSING
g)
PROPOSITION 6.81 (Weaker version using the heat equation). If (Mn, is a complete noncompact Riemannian manifold with Rc (g) ~ -(n - l)g, then Al :S n(n4-1) .
SKETCH OF PROOF. Assume that 0 (see [67J or [117J for example). Now we let u : M x [0,00) ---+ lR be the solution of the heat equation
JM
(~ at -~) U=o ' U(o) =
Al
= u. Since e- 1(0)/2 =
~
=
and equivalently (6.103)
Applying the above lemma, we have at t
= 0,
1M (l\7i\7jfI2 + ~j\7d\7jJ)udp, = ! 1M 1\7 fl2u dp, = 1M (2~f -1\7 f12) ~udp, = 1M 4AI~udp, = o.
- 2
Since Rc ~ -(n -l)g and
o ~ 1M (~(~J)2 =
l\7i\7jfI2
(n
~ *(~J)2, we then have
-1) l\7 fI2 ) udp,
1M (~(4A~+2AII\7fI2+~I\7fI4) -(n-1)I\7fI 2)UdP,.
Noting that at t = 0,
4AI we obtain (6.104)
o~
=
4
1M 1\7
1
12 2 1 4 -AI-4(n-1)AI+4n M l\7fl udp,. n
7.
SOME FURTHER CALCULATIONS RELATED TO :F AND
W
273
Applying the Holder inequality,
1M 1\7 fl4u dJ-l ~ (1M 1\7 fl2u dJ-l) = 16A~, 2
which we substitute into (6.104) to conclude
A
1$
n{n-l) 4
.
o 7. Some further calculations related to F and W
In this section we discuss some interesting computations related to energy and entropy including variational formulas for the modified scalar curvature, the second variation of energy and entropy, and a matrix Harnack calculation for the adjoint heat equation. 7.1. Variational structure of the modified scalar curvature. 7.1.1. Variation of the modified scalar curvature. We now give yet another proof of the variation formula for the F functional (5.10) when 8f ~ h = using the pointwise formula for the variation of the modified scalar curvature. The formulas for Vo ~ R + 2tl.f - 1\7 fl2 and F below should generalize to Vc and We (see (6.25) and (6.26) for the definitions of Vc and We).
¥
LEMMA 6.82 (Measure-preserving variations of R + 2tl.f - 1\7 fl2 and the linear trace Harnack). If 8g = v and 8f = where V ~ gi j vij , on a
¥,
manifold Mn, then (6.105)
8(v,Jf) (R+2tl.f-l\7fI 2) = \7i\7 jvij
+ VijRij -
2\7ivik \7 kf + vij\7d\7jf - 2Vij (~j
+ \7i\7j f) .
For the proof see the more general Lemma 6.85 below. Since under our assumptions 8 (e- f dJ-l) = 0, we have, using (6.105) and the identity (6.39) with € = 0, that :
1M (8(v,Jf) (R+2tl.f-l\7fI 2))e-f dJ-l = - 1M Vij (Rij + \7i\7jf) e- f dJ-l. =
This is the special case of (5.10) where h
=
¥.
REMARK 6.83. If (Mn,g(t)) , t E (-00,00), is a gradient Ricci soliton flowing along \7 f so that Rij + \7 i \7 j f = 0 and if 8g(t)~v(t)~0,
tE(-OO,oo),
274
6.
ENTROPY AND NO LOCAL COLLAPSING
is a bounded nonnegative solution to the Lichnerowicz Laplacian heat equation ltv = dLV, where dL is defined by (VI-3.6), then under measurepreserving variations (v(t),h(t)), where 8(v(t),h(t)) (e-f(t)dllg(t)) = 0, we have 8(v(t),h(t))
(R + 2df - 1\7 f12) (t)
~ O.
This follows from the linear trace Harnack estimate for ancient solutions. EXERCISE 6.84. Show that extending (6.105) to nonmeasure-preserving variations, we have the following. If 89 = v and 8f = h, then
+ 2df -1\7 f12) = div (divv) + (v,Rc) - 2 (divv)· \7f + vij\7d\7j f (R
8(v,h)
(6.106)
+ 2 (d -
\7 f . \7) (h -
~)
- 2Vij (I4j
+ \7 i \7j f) .
Note that (6.105) generalizes to the following, where \7 f is replaced by a vector field X. LEMMA 6.85. If 89 = v and 8X =
then
V2V,
Vt) (R + 29ij \7i X j -IXI 2) = \7i\7jvij + VijRij -
8(v,
2\7i v ik X k + VijXiXj
+ \7i Xj + \7jXi ).
- Vij (2I4j
PROOF. This follows from (VI-p. 69d):
8v R = -dv + \7i\7jvij - VijI4j,
VV) (-IXI2) = V· ·X·X· -
8( v'""2
ZJ
Z
J
X· \7v ,
and 8(v,
Vt) (29ij \7i X j) = 2\7i (:8 Xi) -
2 (:89ij) \7iXj - 29ij
= dv - 2Vij\7iXj - 2\7ivikXk
(:8 r f
j)
Xk
+ \7kvXk.
o EXERCISE 6.86 (Generalizing Exercise 6.84). Show that if 89 8X = Y, then Lemma 6.85 generalizes to 8(v,Y)
(R + 29ij\7iXj
= v and
-IXI 2) = div (divv) + V· Rc -2 div (v) . X + v (X, X) - Vij (2I4j + \7iXj + \7jXi) + 2 div
A different calculation yields
(Y - ~
\7V) - 2X .
(Y - ~
\7V ) .
7. SOME FURTHER CALCULATIONS RELATED TO F AND W LEMMA
6.87. Under the equations %tgij
275
= -2J4j and
a
at Xi = t:.dXi - \1 iR + 2 (\1 X X)i , we have (%t
+ t:. -
2X . \1) (R + 2gij\1iXj -IXI 2 )
= 21J4j + ~ (\1iXj + \1jXi) 12 -
~ l\1 Xj i
\1j Xil 2 .
This extends the calculation (6.22) of Perelman.
7.1.2. An extension of the monotonicity formula. A generalization of Perelman's energy monotonicity formula is given by the following. LEMMA
6.88. If %tgij = -2J4j and
~{ for some a
E
= t:.f + a (R
+ 2t:.f -
1\1 f12) ,
JR, then
(%t - (2a + 1) t:. + 2a\1 f . \1) (R + 2t:.f -1\1 f12) = 21J4j + \1i\1jfI2 . PROOF.
We have by direct calculation,
%t 1\1 fl2 = 2Rij\1d\1jf + 2\1 f . \1 [t:.f + a (R + 2t:.f - 1\1 fI2)] = (2a
+ 1) t:.1\1 fl2 -
2 (2a + 1) l\1 i\1jfI2 - 4aJ4j\1d\1jf
+2\1f· \1 [a (R-I\1fI 2)] and
%t (t:.f) = 2J4j\1i\1jf + t:. (t:.f + a (R + 2t:.f -1\1 fI2))
= (2a + 1) t:. (t:.f) + 2J4j\1i\1j f + t:. (a (R - 1\1 fI2)) . Hence
(%t - (2a + 1) t:. ) (R + 2t:.f - 1\1 f12) = -2at:.R + 21Rijl2
+ 4J4j\1i\1jf + 2t:. (a (R -1\1 fI2))
+ 2 (2a + 1) l\1i\1j fI2 + 4aJ4j\1d\1jf -
2\1 f· \1 [a (R -1\1 fI2)]
= 21J4j12 + 4J4j\1i\1jf + 21\1i\1j fI2 - 2a\1 f . \1 (R + 2t:.f -1\1 f12) and the lemma follows. From this we deduce the following.
0
6.
276
ENTROPY AND NO LOCAL COLLAPSING
COROLLARY 6.89. If %tgij
= -2~j
~~ = ~f + a for some a E
~,
(R
and
+ 2~f - IV f12)
,
then the energy on a closed manifold M n satisfies
r (R + 2~f -IV f12) e- f d/l
dF (g (t), f (t)) = ~ dt dt 1M
1M I~j + Vi V jfl 2 e- f d/l (1 + a) 1M (R + 2~f -IV f12)
=2 -
2
e- f d/l.
So if a ::; -1, then :t
1M (R+2~f -I V fI
2)
e-fd/l
~ O.
EXERCISE 6.90. Put a positive constant E in the above formulas; more precisely, consider the E-entropy We and determine equations for f and T such that a monotonicity formula for We holds. Perhaps one should consider the set of equations
a at gij = -2~j af at
= ~f -:. (f -!!:.) + a 2
T
(R
+ 2~f -IV fl2
-:. (f - n)) T
dT
-=E
dt for a ::; -l.
7.2. Second variation of energy and entropy. As is typical for energy-type functionals, we consider the second variation of Perelman's energy and entropy functionals. We are particularly interested in what sense critical points of the entropy functional are stable, i.e., have nonnegative second variation. Suppose ag af = v and =h
as
as
on a closed manifold Mn. Equation (6.31) implies the following:
:s [(~j + Vi Vj f) e-
f dj.t ]
(6.107)
= Vp (e- f ~r~.) as ~J
+ [ViVj
(h -
d/l
~) - (~j + ViVjf) (h - ~)] e- f d/l.
7. SOME FURTHER CALCULATIONS RELATED TO:F AND W
277
Differentiating (5.10) again and using (6.106) and (6.107), we have
d2 d82:F(g, f)
1M (:8 Vij - 2VikVjk) (~j + V'/\i'jf)e- I dp, - 1M Vij V'p (e-I :8 rfj) dp, - 1M Vij [V'iV'j (h - ~) - (~j + V'iV'jf) (h - ~)] e- I dp, + 1M [:8 (~ - h)] (2~f -1V'fI 2 + R) e- I dp, + 1M (~ - h) (div (divv) + (v, Rc) - 2 (divv) . V' f + vijV'dV'jf) e- I dp, + 1M (~ - h) (2 (~ - V' f . V') ( h - ~) - 2Vij (~j + V'i V' j f)) e - I dp, =-
+
IM(~ -h)2(2~f-lV'fI2+R)e-IdP,.
Integrating by parts, we have
1M Vij [V'iV'j (h - ~) - (~j + V'iV'jf) (h - ~)] e- I dp, = 1M (h- ~) [V'iV'j (Vije- / ) -Vij(~j+V'iV'jf)e-/] dp, = r (h _ V) ( div (divv) + (v, Rc) ) e-I dp,
(6.108)
1M
- 2
-2 (divv) . V' f
2
1M Vij (~j + V'iV'jf) (h -
+ vijV'dV'jf
~) e- I dp,.
Hence ~:F(g, f) is equal to
- 1M (:8 Vij - 2VikVjk) (~j + V'iV'jf)e- 1dp, + 1M (:8 rfj) (V'p v ij)e- I dp,+2 1M IV' (h- ~)12 e-Idp, + 41M Vij (Rij + V'iV'jf) (h - ~) e- I dp, +
L[:8 (~ -h)
+2
+ (~
-h)'] (2~f-IVfI2+R)e-fd~
r (diV (divv) + (v, Rc)
1M - 2 (div v) . V' f + Vij V' dV' j f
)
(V2 _
h) e-I dp,
.
278
6. ENTROPY AND NO LOCAL COLLAPSING
If 9 is a steady gradient soliton flowing along
V' f,
then
+ V'iV'jf = 0, 2flf -1V'fI 2 + R = 0, ~j
(6.109)
Writing the second variation this way, we see the appearance of the linear trace Harnack quadratic. Note that the first line of (6.109) is independent of h whereas for (pointwise) measure-preserving variations (v, h) of (g, f), the last two lines of (6.109) vanish. This shows that given Vij, the variation (v, h) minimizing ~F(g, I) while preserving the (integral) constraint fM e- f dJ.-t = 1 has the last two lines of (6.109) nonpositive; not surprisingly, the last two lines are less than or equal to a negative norm squared, as we shall show below. Using (6.108), the second line in (6.109) may be rewritten as
1M [V'iV'j (h - ~)] Vije- f dJ.-t = 2 1M (div v - v (V'I)) . [V' ( h - ~)] e- f dJ.-t.
- 2
Hence for a steady gradient soliton 9 flowing along
:2
F (g,1)
V' f,
1M (V'iVjP - ~V'PVij) (V'pvij)e-fdJ.-t + 2 1M (div v - v (V'I)) . [V' ( h - ~)] e- f dJ.-t =
+21M IV' (h- ~)12 e-fdJ.-t. Completing the square, we have
:2
F (g, f)
=
1M (V'iVjP - ~V'PVij) (V'pVij) e- f dJ.-t
-~
{ Idivv-v(V'1)1 2 e- f dJ.-t
21M
+2
1M I~ (divv - v (V'I)) + V' (h - ~) 12 e-f dJ.-t.
7. SOME FURTHER CALCULATIONS RELATED TO :F AND W
279
If a is a 1-form, then the minimizer of the energy
E(h)
~ 1M Idh+aI 2 dJ..l
is given by
!:lh = - div (a) . Thus, for a steady gradient soliton 9 flowing along V f,
::2 F (g, f) ~ 1M (viVjp - ~VPVij) (VpVij) e- dJ..l f
-! f
(6.110)
21M
Idivv-v(Vf)1 2 e- f dJ..l
+ ~ 1M Idivv - v (V f) + Vwl 2 e- f dJ..l, where
!:lw = - div (divv - v (V f)) and with equality in (6.110) if and only if
(6.111)
!:l(h-
~) =-~div(divv-v(Vf)).
Since
1M (divv - v (V f), Vw) dJ..l = - 1M w div (div v - v (V f)) dJ..l = 1M w!:lwdJ..l = - 1M IVwl2 dJ..l,
we conclude
1)
d f( ViVjp - 2VpVij (Vpvij)dJ..l- 21M 1f IVwl 2 dJ..l. ds2F(g,f) ~ 1M 2
Since
1M ViVjpVpVijdJ..l
= - 1M VjpViVpVijdJ..l = 1M Idiv v 12 dJ..l + 1M Vjp (- RpqVqj + RipjqViq) dJ..l,
we obtain the following. 6.91. Let (M n , g) be a steady gradient Ricci soliton flowing along V f on a closed manifold. If = v and ~ = h, then PROPOSITION
(6.112)
.!...-F(g, f) ds 2
1s
~ f (-~ IVvI2 + IdivvI 2 - ~ IVwl2 1M
- RpqVqjVjp + RipjqViqVjp
with equality if and only if (6.111) holds.
) dJ..l,
280
6.
ENTROPY AND NO LOCAL COLLAPSING
REMARK 6.92. Note that
(Vi - Vd) (Vj - vjf) Vij
= div (divv) + (Rc, v) - 2 div (v) . V f + vijVdVjf - (Ri,j + ViVjf) Vij' Also,
and
For any closed Riemannian manifold
J
by (5.10), (5.51), and M (~- h) e-fdp,
(Mn, g) with variation gsg = v, = 0, we have
!.x (g) = - 1M Vij (Ri,j + ViVjf) e- f dp" where f is the minimizer of :F (g,.). Thus the critical points of .x (g) are the steady gradient Ricci solitons flowing along the gradient of the minimizer. Since, given v, equality holds in (6.112) when h satisfies (6.111), we obtain the following second variation formula proved in [53].
(Mn,g)
THEOREM 6.93 (Cao, Hamilton, and Ilmanen). Let Ricci flat manifold and let v be a symmetric 2-tensor. If 9 (s)
d s=o .x (g (s)) 2
d
2
s
1
=
1 (1
= 9+ sh, then
1
. 2 - -21Vwl 2 + Ri,pjqViqVjp ) dp, --2IVvl 2 + Idlvvl
.
M
=
be a closed
1M (L v, v) dp"
where w is defined by (up to an additive constant)
!!l.w
~
div (div v)
and
Lv
~ ~!!l.v - ~.c(diVV)~g + Rm (v).
Similarly, one can compute the variation of v (9)
~
inf p, (g, 7) = inf W (g, f, 7) ,
TE[O,oo)
T,f
where JM e- f dp, = (47r7t/ 2 , on an Einstein manifold with positive scalar curvature (see [53]). THEOREM 6.94 (Cao, Hamilton, and Ilmanen). Let
(Mn,g)
be a closed
Einstein manifold with positive scalar curvature, i.e., Rc = 2~g where 7 > 0,
7. SOME FURTHER CALCULATIONS RELATED TO :;: AND W
and let v be a symmetric 2-tensor. If 9 (s) = d22 -d s
1
II
1 (1
r(~). Vol M M
(9 (s)) =
8=0
+ Vol ( M) _
~(
then
1
. 2 - -2 1V' w l2) dp, --21V'vl 2 + Idlvvl
L(
R;jktVijV,'+
1
Vol
2n
9 + sh,
281
r VdP,)
(Nt) lM.
2
4~ W 2 ) d~
'
where w is defined uniquely by ~w
w
+-
2r
~
div (divv) ,
7.3. A matrix Harnack calculation for the adjoint heat equation. We also have the following Harnack-type calculation (see [287]). We may think of equation (6.113) below as a Bochner-type formula which says that the evolution of the modified Ricci tensor is given by a backward Lichnerowicz Laplacian heat operator with Hamilton's matrix Harnack quadratic as the main term on the RHS. PROPOSITION
6.95 (Matrix Harnack formula for adjoint heat equation).
Under the system
a
at 9ij = - 2Ri,j , of
2
nc
at
= -~f - R + IV' fl - 2r'
dr dt
= C,
we have
(:t + ~L
- 2V'f· V') (V'iV'jf
+ Ri,j + 2Cr9ij)
=2(~LR"-~V"V"R+R-kR'k+ CD .. ) 2 2r
(6.113)
t)
t)
t)
.I.
Lot)
+ 2 ((PiPj + Pjpd V'd + RkijPV'dV'kf) - (V'iV'k!+Rik+ 2Cr9ik) (-V'jV'kf+Rjk+ 2Cr 9jk )
- ( - V'iV'k! + Rik + 2Cr9ik) (V'jV' k! + Rjk + 2Cr9jk) . PROOF.
We have
(:t + ~L) V'iV'j
V'P
282
6.
ENTROPY AND NO LOCAL COLLAPSING
for any function
(:t
+!:1L )
~j = 2!:1L~j.
Hence
(:t +
+ ~j + 2£7 9ij )
!:1L) (V /'1 j I
(a I)
£2 = 2!:1L~j + ViVj ( -R + IV II 2) - 2 at r ij Vi! - 2729ij = 2 ( !:1LRij -
C ;~j
~ViVjR + RkijIVkIVi!)
+ 2VkViVjIVkI + 2ViVkIVj Vki
+ 2 (V'iRjl + VjRil -
c2
c
7
7
V'eRij) V'i! - 2 29ij - -Rij
= 2Vk (V'iV'ji + Rij + 2£79ij) V'ki
+ 2 (!:1L~.J - ~V'·V'·R 2 ~ J + ~kR·k J + ~~.) 27 J + 2 ((Pilj + Pjei) V'i! + RkijeV'kIV'i!) + 2V'iV'kiV'j V'ki -
2£
2~kRjk - -~j 7
£2 -229ij 7
o
and the result follows from rewriting the last line. REMARK 6.96. In the evolution equation for the RHS by any function of t.
I,
we may replace -
~~
on
Since the matrix Harnack quadratic is the space-time Riemann curvature, we ask the following. PROBLEM 6.97. Is there a space-time interpretation of equation (6.113)? For convenience, we define Ttj
~ Rij -
V'iV'ji + 2:9ij
and
1 c H (X)ij ~ !:1L~j - 2V'iV'jR + RikRjk + 27 ~j + (Pilj
+ Pjli) Xl + RkijlXkXe.
Then we may rewrite (6.113) as
(:t
+!:1L - 2V'I .
V)
where Sfk is defined in (6.27).
Sfj = 2H (V' f)ij - SfkTJk -
SjkTi~'
7. SOME FURTHER CALCULATIONS RELATED TO :F AND W EXERCISE
283
6.98. Under the system of equations
a
ot 9ij
a
at x dr dt
= -2Rij , = -~X
+ Rc (X) -
\7 R
+ 2 (\7 X·X) ,
= c,
generalize the above calculation to find (:t
+~L -
X· \7 - LX)
(2~j + \7iXj + \7jXi + ~9ij).
There is also a corresponding formula where \7 f is replaced a closed 1-form X. When X is not closed, there are a couple of extra terms in the calculation which have a dX factor. A Kahler version of this calculation of (6.113) was obtained by one of the authors. There is a computation analogous to (6.113) which one can perform for solutions of a forward heat-type equation. In particular (see [287]). LEMMA
6.99. If on a manifold Mn,
a
at 9ij = -2~j,
~~ = ~f -
R-J\7fJ2,
then
satisfies (:t -
~L + 2\7 f
. \7) Zij
= Yij
- Zik ( Rjk
- ( \7i\7kf +
+ \7j\7kf + ;t9jk)
~k + ;t9ik) Zjk,
where
1 Yij ~ \7i\7jR + 2\7kR ij\7kf + 2RkijR.\7k!\7ef + RikRjk + t~j·
One of the authors has made the following conjecture. CONJECTURE 6.100. If (Mn, 9 (t)) and f (t) are solutions to the equations in Lemma 6.99, where 9 (t) has bounded nonne9ative curvature operator, then
Yij 2::
o.
284
6. ENTROPY AND NO LOCAL COLLAPSING
8. Notes and commentary Section 5. For T < 00 the condition on rVtk just says that the rk are uniformly bounded. In Definition 6.56, condition (1) may be replaced by (I') For some C < 00, IRm [g (tk)ll ::; C r;,2 in B (Pk, rk). We call the original definition the 'first definition' and the definition with (I') replacing (1) the 'second definition'. If 9 (t) is locally collapsing according to the first definition, then taking C = 1, it is locally collapsing according to the second definition with the same rk. On the other hand, if gij (t) is locally collapsing according to the second definition, then the following hold. (i) If C ::; 1, then 9 (t) is locally collapsing according to the first definition with the same rk. (ii) If C ~ 1, then set 1'k ~ rk/v'C ::; rk. Then IRm [g (tk)ll ::; 1';,2 in B(Pk,1'k), and we still have limk->oo1';,nVoIB(pk,1'k) = 0, so that gij (t) is locally collapsing according to the first definition with
Tk = rk/v'C.
CHAPTER 7
The Reduced Distance How thoroughly it is ingrained in mathematical science that every real advance goes hand in hand with the invention of sharper tools and simpler methods which, at the same time, assist in understanding earlier theories and in casting aside some more complicated developments. - David Hilbert Technical skill is mastery of complexity while creativity is mastery of simplicity. - Chris Zeeman
In [297] Perelman introduced a new length (energy-like) functional for paths in the space-times of solutions of the Ricci flow, called the .L:-length. The naturalness of this functional can be justified both by the space-time approach and the various differential inequalities that the quantities associated to the .L:-length satisfy, which we shall show in this chapter. A fundamental inequality is the monotonicity of the reduced volume. As we shall see in the next chapter, this monotonicity leads to a second proof of (weakened) no local collapsing for finite time singularities. We emphasize that, unlike the entropy proof, the proof of weakened no local collapsing in Chapter 8 using the reduced volume also holds for complete noncompact solutions with bounded sectional curvature. Recall that no local collapsing provides a local injectivity radius estimate and at the same time rules out the formation of the cigar soliton singularity model. Besides bringing comparison geometry and integral monotonicity into Ricci flow, some original aspects of Perelman's work on the reduced distance function are as follows. (1) A space-time distance-like function which is not always nonnegative. (2) Bochner formulas, i.e., partial differential inequalities, which are geometry (e.g., curvature) independent, i.e., these formulas and inequalities hold in any dimension and are independent of the initial metric. (3) Pointwise monotonicity adapted to the space-time geometry. This describes in some sense how, for any solution to the Ricci flow, the geometry improves as time increases. (4) Using the space-time geometry to understand point-picking and compactness, in particular, understanding the structure of ancient 285
286
7. THE REDUCED DISTANCE
Ii:-solutions, finite-time singularity models, and high curvature regions of the solution. (5) Relating Ricci flow and aspects of function theory and the heat equation, for example, the analysis of the fundamental solution of the adjoint heat equation coupled to the Ricci flow. In this chapter we discuss the basic properties of the .c-Iength and the associated distance functions for complete (not necessarily compact) solutions to the backward Ricci flow with bounded sectional curvature. l 1. The .c-Iength and distance for a static metric
One of the primary antecedents of Perelman's .c-Iength and reduced distance f. is the work of Li-Yau on differential Harnack inequalities for the heat equation on a Riemannian manifold with a static metric. 2 With this in mind we start by summarizing the properties of the energy functional for paths in a Riemannian manifold and various monotonicity and comparison results. The purpose for this is to compare properties associated to the linear heat equation with respect to a static metric to properties of the nonlinear case of metrics evolving by Ricci flow and to show a strong analogy between these two cases. The fact that the case of the heat equation is less technical facilitates the presentation of some of the underlying ideas. Let (Mn,g) be a complete Riemannian manifold. Given a C 1-path , : [71, 72] ~ M, 71 ~ 0, joining two points, i.e., ,(71) = p and ,(72) = q, we define its energy by En,,. h)
~
f
JTI~~I: dr.
Convention: By saying that a path, (7) is C k , where k we really mean, ( is a C k function of (J" ~ 2JT.
u;)
= 1 or 2,
This energy functional, which is well-suited for studying the heat equation, is equivalent to the usual energy for paths (see for example §12 of Milnor's book [265]). Indeed, making the change of variables (J" = 2JT yields
(7.1)
£Tl,T2
(r) =
{2.JF21 dd, 12 d(J".
12Ft
(J"
9
Keeping in mind that we shall be discussing parabolic equations, we require that the parameter of the path be given by the time variable 7. In particular, if we want to study the (reversed) concentration process of the fundamental ISee the last section for some of the notational conventions we use in this chapter. 2The later work of Hamilton on the matrix differential Harnack for the Ricci flow is also important in this development.
1.
THE .c-LENGTH AND DISTANCE FOR A STATIC METRIC
solution to the heat equation, which is a delta function at M, we take 1'1 = 0 so as to define
l'
287
= 0 at a point
p E
(7.2)
£ (T) ~
r
Jo
Id'12 d1'
.,fT d1'
{j
for CI-paths , : [0, rj ---+ M from p to arbitrary points q E M. From (7.1) we see that the critical points, are of the form
,(1') = (3 (2.,fT) , where (3 is a constant speed geodesic. Hence, in Euclidean space, ,(1') = 2JTV for some V E ~n, and its graph (, (1') , 1') is a parabola. This is one justification for the JT factor in (7.2): parabolas are more suited to the heat equation. Note that . d, d(3 (7.3) hm .,fT-d (1') = -d (0) E TpM. T--O l' a A
EXERCISE 7.1. Show that if, is a critical point of (7.2), then 1
"\lxX+-X=O 21'
where X =~. SOLUTION. From (7.1) we find that the critical points satisfy
"\l vrX (.,fTX) = O. Given a basepoint p (at time 0) define a space-time distance function on M x (0,00) by L (q, r) ~ inf £ (T) , '"Y
where the infimum is taken over all CI-paths , : [0, rj ---+ M with, (0) and, (r) = q. An elementary computation using (7.1) yields
(7.4)
L ( r) q,
=
=p
d (p, q)2 2...jT ,
where d is the distance with respect to 9 and the infimum of £ (,) is obtained by a minimal geodesic, from p to q with
(7.5)
1
1 d(p,q)
d, (1')1
d1'
{j
JT2...jT·
With the Euclidean heat kernel in mind we define the reduced distance:
(7.6)
f( r)=L(q,r)=d(p,q)2 q, . 2...jT 4r
If we wish to remove the time-dependence in (7.6), then we may define the enlarged distance:
(7.7)
L (q, r) ~ 2VfL (q, r) = 4rf (q, r) = d (p, q)2 .
7. THE REDUCED DISTANCE
288
EXERCISE 7.2 (L-distance and £. between two space-time points). Show using (7.1) that infl' £rl,r2 h) , where the infimum is taken over all CI-paths 'Y with 'Y(Td = p and 'Y(T2) = q, is attained by 'Y(T) = !3(2y'T) , where
!3 : [2..jTl,2yfT2J d(p,q) 2JT2-2..jTl'
H
--+
is a minimal geodesic with constant speed 1~~ Ig
Nt
ence C () L (p,rt) ( q, T2 ) :::;=..In f (...rl,r2 'Y l'
(7 . 8)
==
=
d (p, q )2 2;;;:;; 2 r,;; yT2 -
yTI
We may then define the reduced distance by
£.
(7.9)
r (q, (p,l).
T2) ~
L(p,rt) (q, T2) = d (p, q)2 2yfT2 + 2..jTl 4 (T2 - Td
We compute under the assumption RCg 2: 0 (7.10)
(:r
+ ~ ) (L -
2nr) =
~ (d 2 ) -
2n = 2
(d~d + IVdl 2 -
n) ::; 0
in the weak sense,3 where we used IVdl = 1 a.e. and the Laplacian Comparison Theorem (A.9): d~d::; n - 1. That is, L - 2nr is a subsolution of
the backward heat equation in the weak sense. REMARK 7.3. It is useful to keep in mind the examples of flat tori to see why one cannot prove a stronger statement (see subsection 9.5 of this chapter).
The role of the reduced distance in the study of the heat equation is exhibited by the Li-Yau differential Harnack inequality (A.12), which implies that for any positive solution u of the heat equation on a complete Riemannian manifold with RCg 2: 0,
U(X2,T2) (T2)-n/2 { } ( ) 2: exp -£'(xl,rt} (X2, T2) , U Xl, Tl
Tl
where Tl < T2. A similar, but slightly more complicated, statement holds when RC g 2: -Kg, where K 2: O. See [253] for details.
2. The £-length and the L-distance Let
(Nn, Ii (t)) , t
E
(a,w), be a solution to the Ricci flow. From this
we can easily obtain a solution
(Nn, h (T)) 8
8Th
to the backward Ricci flow
= 2Rc
cp on space-time with compact support, .£00 1M (- ~~ + ~cp) (L - 2n1') dp.d1' ~ O.
3That is, for any nonnegative C 2 function
See subsection 9.5 of this chapter for a justification of (7.10).
2. THE .c-LENGTH AND THE L-DISTANCE
289
by reversing time. In particular, if w < +00, let 7 ~ w -t, so that (N, h (7)) is a solution to the backward Ricci flow on the time interval (O,w - a).4
2.1. Space-time motivation for the L:-Iength. We begin by motivating the definition of the L:-Iength for the Ricci flow as a renormalization of the length with respect to Perelman's potentially infinite Riemannian metric on space-time. Given N E N, define a metric on.N ~ Nn x SN X (0, T) by
(7.11)
h
~ hijdxidx j + 7hOl(3dy
OI
dy(3 +
(~ + R) d7 2 ,
where hOl(3 is the metric on SN of constant sectional curvature 1/ (2N) and R denotes the scalar curvature of the evolving metric h on N. Here we have used the convention that {xi} ~=l will denote coordinates on the N factor, {yOl}~=1 coordinates on the SN factor, and xo ~ 7. Latin indices i,j, k, ... will be on N, Greek indices a, f3, ,,(, ... will be on S N, and 0 represents the (minus) time component. Choosing N large enough so that ~ + R > 0 implies that the metric h is Riemannian, i.e., positive-definite. In local coordinates,
(7.12)
hij = hij ,
(7.13)
hOl(3 = 7h Ol(3,
(7.14)
hO~
(7.15)
hiO = hiOi = hOio = O.
-
-
=
N
27
-
+ R, -
Let 1'(s) ~ (x(s), y(s), 7(S)) be a shortest geodesic, with respect to the metric h, between points P ~ (xo, Yo, 0) and q ~ (Xl, YI, 7q ) E .N. Since the fibers SN pinch to a point as 7 ---t 0, it is clear that the geodesic 1'( s) is orthogonal to the fibers SN. (To see this directly, take a sequence of geodesics from Pk ~ (xo, YI, 1/k) to q and pass to the limit as k ---t 00.) Therefore it suffices to consider the manifold jJ' ~ N x (0, T) endowed with the Riemannian metric:
(7.16)
Ii
~ hijdxidxj + (~ + R) d7 2 .
(This metric is dual to the metric considered in [100J.) For convenience, denote x(s) ~ "((s). Now we use s = 7 as the parameter of the curve. Let.:y (7) ~ ~; (7). The length of a path 1 (7) ~ ("((7), 7) , with respect to the metric Ii, is given by the following: 4We shall consider the case where Q = -00 (in which case we define w - Q ~ +00). On the other hand, if w = +00 and Q = -00, we may simply take T = -to However, for the backward Ricci flow we are not as interested in the case where w = +00 and Q > -00.
290
7. THE REDUCED DISTANCE
Lengthji (i)
= fo
Tq
J~
+ R + 11' (r)1 2 dr
r q V[ii. ( . 2) 2;V/1 + 2r N R+ 1,(r)1 dr
= Jo
= fo =
Tq
~ (1 + ~ (R + 11' (r)1
2)
+ a (N- 2 )) dr
fo Tq ~dr + fo Tq J2~ (R + Ii' (r)1
= J2Nrq + v'~N
2)
dr +
fo Tq {fa (N-
fo Tq yT (R + 11' (r)1 2 ) dr + ..j'iTqa (N-
3/ 2 )
dr
3/ 2 ) •
The calculation indicates that as N --+ 00, a shortest geodesic should approach a minimizer of the C-Iength functional defined by
Cb)
~ fo Tq yT (R b (r), r) + 11' (r)I~(T)) dr.
Note that the definition of Cb) only depends on the data of (N, h). EXERCISE 7.4 (Levi-Civita connection of the potentially infinite metric) . Consider the metric h on Nn x (0, T) defined in (7.16) by (7.12), (7.14), and hio = 0 (without the SN factor). The components of the Levi-Civita connection N'fj of h are defined by
N'fj
8 8xa
~ = ~ Nrc ~c aXb
~
ab aX '
c=O
where x O = r. Show that
and
(N = - 2r + R ) ~j, (N )-1 21 ViR , = 2r + R -1
N-O
r ij
N-O riO N-O
roo=
(N-+R )-1 -1 (aR R) 1 -+--. 2r ar r 2r 2
2.
THE C-LENGTH AND THE L-DISTANCE
291
In particular, Nf'!b are independent of N, whereas .
hm
N-O
r··~J
N-+oo .
hm .
'
N-O
riO = 0,
N-+oo
hm
= 0
r oo = - 1-
N-O
N-+oo
27
2.2. The £-length. A natural geometry on space-time (in the sense of lengths, distances and geodesics) is given by the following. DEFINITION 7.5 (£-length). Let (Nn, h (7)),7 E (A, 0), be a solution to the backward Ricci flow h = 2 Rc, and let 'Y : [71, 72J ~ N be a piecewise C 1-path,5 where h,72J c (A,O) and 71 2: O. The £-length of'Y is6
-t
(7.17)
£("()
~ £h ("() ~ 1T2 Vr (R("((7) ,7) + Idd'Y (7)1 T1 h(T) 2
)
d7.
7
Later we shall take
71
= 0 and call 72 = f.
REMARK 7.6. Taking 71 = 0, the subsequent degeneracy introduced by the Vi factor in (7.17) reflects the infinite speed of propagation of the Ricci flow (as a nonlinear heat-type equation for metrics). We also note the formal similarity between R+ 1~ 12 and the quantity R+ 1'\7112 which we considered for gradient Ricci solitons and which also appeared in the definitions of energy and entropy; this seems like more than just a coincidence. The £-length is defined only for paths defined on a subinterval of the time interval where the solution to the backward Ricci flow exists. Note that £ may be negative since the scalar curvature may be negative somewhere. This is in contrast to the energy defined in Section 1 above for a static metric. Often we shall use the following conventions: (7.18) We may rewrite £ as (7.19)
£("() =
r..fi2 (:2 R(f3((]'),(]'2/4) + 1~f3 ((]')1 2 ) d(]'.
J2v'T1
This is especially useful in the case
(]'
71
h(u 2 /4)
= O.
Because of the 1~ 12 term on the RHS of (7.17), £ ("() looks more like an energy than a length. Another way to obtain £, which is related to the 5That is, 'Y
("42) is a C 1function of u.
6 R (f (T), T) is just a notation meaning vature of (N", h (T)) .
Rh(T)
(T (T)), where
Rh(T)
is the scalar -cur-
292
7.
THE REDUCED DISTANCE
above approach of renormalizing the Riemannian length functional, is as follows. We define the space-time graph
l' : h, T2J -t N
h, T2J
x
of the path, by i'(T) ~ (,(T),T), so that ~~ (T) = (~; (T),1). Note that the parameter T, of which, is a function, also serves as time; so it is natural to consider its graph. Define the space-time metric h ~ h + RdT2. In general, this metric is indefinite since R may be negative somewhere. We easily compute
£ (,) Using
0"
= 2y'T,
=
1
72
71
Id-
12
y'T d; (T) h dT.
we may rewrite the £-length as
where O"i ~ 2fo" i = 1,2. That is, £ (,) is the energy of the space-time path l' with respect to the space-time metric h and the new time parameter 0". If 0:: [Tl,T2J-t Nand f3: h,T3J-t N are paths with 0: (T2) = f3(T2), then we define the concatenated path 0: '---' f3 : h, T3J -t N by
We have the following additivity property. LEMMA 7.7 (Additivity of the £-length). (7.20)
£ (0: '---' (3) = £ (0:)
+ £ (f3) .
However, the £-length of a path, is not invariant under reparametrizations of ,. We leave it to the reader to make the easy verification of this fact. The following bound on £ is elementary. LEMMA 7.8 (Lower bound for the £-length). (7.21)
£ (,)
~
-2 (3/2 T2 - Tl3/2) 3
. mf
R.
NX[7I,T21
1
72
This follows directly from £ (,) ~
y'TRinf (T) dT, where Rinf (T)
~
71
infNx{7} R. The Riemannian counterpart of estimate (7.21) is the obvious fact that the length of a path is nonnegative.
292
7.
THE REDUCED DISTANCE
above approach of renormalizing the Riemannian length functional, is as follows. We define the space-time graph
i: h,T2]----t N
x h,T2]
1) .
of the path 'Y by i (T) ~ b (T) , T), so that ~~ (T) = (~~ (T) , Note that the parameter T, of which 'Y is a function, also serves as time; so it is natural to consider its graph. Define the space-time metric h ~ h + RdT2. In general, this metric is indefinite since R may be negative somewhere. We easily compute
[b) Using a
=
=
1 Vi Id-d~
12
T2 Tl
(T) h dT.
2yfi, we may rewrite the [-length as
2,;r;.,
i = 1,2. That is, I:- ('Y) is the energy of the space-time path where ai ~ i with respect to the space-time metric h and the new time parameter a.
T2] ----t Nand {3 : [T2' T3] ----t N are paths with 0: (T2) then we define the concatenated path 0: '-../ {3 : [TI' T3] ----t N by If 0: :
[TI'
=
{3 (T2),
We have the following additivity property. LEMMA 7.7 (Additivity of the [-length). (7.20)
[ (0: '-../ {3)
=
I:- (0:)
+[
({3) .
However, the £-length of a path 'Y is not invariant under reparametrizations of "f. We leave it to the reader to make the easy verification of this fact. The following bound on I:- is elementary. LEMMA 7.8 (Lower bound for the £-length). (7.21)
(3/2 - 3/2) mf .
[ ('Y) ~ -2 T2 3
TI
Nxh,T2l
1
R.
T2
This follows directly from [b) ~
yfiRinf
(T) dT, where
Rinf
(T)
~
T1
infNx{T} R. The Riemannian counterpart of estimate (7.21) is the obvious fact that the length of a path is nonnegative.
2. THE L:-LENGTH AND THE L-DISTANCE
293
2.3. The L-distance function. Just as for the usual length functional (perhaps it is better to compare with the energy functional), one gives the following definition. DEFINITION 7.9 (L-distance). Let (Nn, h (T)), T E (A, 0), be a solution to the backward Ricci flow. Fix a basepoint pEN. For any x E Nand T > 0, define the L-distance by
L (x, T) ~ L~p,o) (x, T) ~ i~f £ h) , where the infimum is taken over all CI-paths 'Y : [0, T] ~ N joining p to x (the graph i joins (p,O) to (x, T)). We call an £-length minimizing path a minimal £-geodesic. We also define
L (x, T) ~ L~,o) (x, T) ~ 2.;TL (x, T) .
(7.22)
Note that the L-distance defined above may be negative. To help the reader have a feeling for the L-distance function, we present some exercises. EXERCISE 7.10 (Scaling properties of £ and L). Let (Nn, h (T)) be a solution to the backward Ricci flow, 'Y : [Tl' T2] ~ N a CI-path, and c > a constant. Show that for the solution it (f) ~ ch (c-lf") and the path 1: [CTl' CT2] ~ N defined by 1 (f) ~ 'Y (c- 1 f) , we have
°
£it (i) = JC£h h)·
Consequently,
L~,o) (q, f) = JCL~p,o) (q, c-lf) . EXERCISE 7.11 (£ and L on Riemannian products). Suppose that we are given a Riemannian product solution (Nrl x N;:2, hI (T) + h2 (T)) to the backward Ricci flow and a C 1-path 'Y = (a, (3) : [Tl' T2] ~ Nl X N 2. Show that
Hence
L~~1~:22,O) (ql, q2, T)
=
L~~l'O) (ql, T) + L~;2'O) (q2, T) .
It is useful to keep in mind Euclidean space as a basic example; more generally we have EXERCISE 7.12 (L-distance for Ricci flat solutions). Let (Nn, h (T) = h o) be a static Ricci flat manifold and let pEN be the basepoint. Show that given any q E Nand f > 0, the £-length of a C 1-path 'Y : [0, f] ~ N from p to q is
£h) =
10
2"fi T
I~; (a 2 /4)
1da, 2
which is the same as (7.2). Hence a minimal £-geodesic 'Y is of the form (7.23)
'Y (T) =
f3 (2.;T) ,
294
7.
THE REDUCED DISTANCE
where (3 : [0, 2v'r] ~ Nn is a minimal constant speed geodesic with respect to ho joining p to q. Thus L ( f) = d (p, q)2 q, 2VT
(7.24)
For reference below, we have L (q, f) ~ d (p, q)2 and £ (q, f) ~ 2~L (q, f)
=
d~~)2 ; the definition of £ will be given again in (7.87). SOLUTION TO EXERCISE 7.12. This exercise is a special case of the discussion in Section 1 above. We also note that d, (7) 12 7 1d
(7.25)
=
g(T)
7
1dd(3 (0') 12 0'
=
2 , IVlg(o)
g(u2/4)
where V ~ limT-+o yT~ (7) = limu-+o ~ (0'). We leave it to the reader to check that for .c-geodesics defined on a subinterval [71,72] c [0, T] , we still have
71~' (7) 12
(7.26)
== const .
g(T)
7
2.4. Elementary properties of L. In this subsection (M n , 9 (7)) ,7 E [0, T], shall denote a complete solution to the backward Ricci flow, and p EM shall be a basepoint. We will assume the curvature bound (7.27)
max
(x,T)EMx[O,Tj
{IRm (x, 7)1, IRc (x, 7)1} ~ Co
< 00.
The curvature bound assumption is written in this way for the convenience of stating later estimates. We prove some elementary CO-estimates for the L-distance and lengths of .c-geodesics, relating them to the Riemannian distance; we shall use these estimates often later. First recall from (3.3) in Lemma 3.11 that for 71 < 72 and x E M, e- 2CO (T2- Tl) 9 (72, x) ~ 9 (71, x) ~ e2CO (T2- TI) 9 (72, x) .
LEMMA 7.13 (.c and Riemannian distance). Let, : [0, 1'] ~ M, l' E
(0, TJ, be a C 1 -path starting at p and ending at q. (i) (Bounding Riemannian distance by .c) For any
d~(o) (p, ,
(7))
7 E
[0,1'] we have
~ 2JTe 2COT ( .c (,) + 2n3Co1'3/2) .
In particular, when M is noncompact, for any l' E (0, T], we have
lim L (q, f)
q-+oo
= q-+oo lim 2VfL (q, f) = +00.
(ii) (Bounding speed at some time by.c) There exists 7* E (0, f) such that
d, 12 7* I-d (7*) 7
g(T.)
Id(3
= -d (0'*)
0'
12 g(T*)
~ 2
nCo _ r;.c (T) + -3- 7, y 7 1
2. THE .c-LENGTH AND THE L-DISTANCE
295
where f3 (cr) ~ ')' (T), cr = 2.jT, and cr* ~ 2..rr;. (iii) (Bounding L by Riemannian distance) For any q E M and L (q, T-)
:::;
e
(p, q) 2-/T
2Co1' d;(1')
f
> 0,
2nCo -3/2
+ -3- T
.
7.14. In each of the estimates above, on the RHS one may think of the first term as the main term and the second term as an error term. Recall by (7.24) that if (Mn, 9 (T) = 90) is Ricci fiat and')' : [0, f] --t M is a minimal .c-geodesic from p to q, then REMARK
d~o (p,,), (T)) = 2y'TL(')'(T) ,T) = 2y'T.c (')'I[O,rl) = ~d~o (p,q) , and for all
T*
E
(0, f) , T*
Id')' (T*)1 dT
2
=
g(r.)
~o (p,q). 4T
Hence for Ricci fiat solutions, .e (')' (T), T) defined in (7.87) is constant (= Ad;o (p, q)) along .c-geodesics. PROOF. (i) Let a- = 2v'T and f3 (a-) ~ ')' (7'). The idea is to first bound the energy of f31 [0,2y'T] . By splitting the formula for .c into two time intervals, we see that
1
2y'T 1 df3 - (0-) 12 dao dOg(q2/4) 2";¥
=.c(')')-
(7.28)
:::; .c (')')
r J2y'T
Idd~(o-)I cr
2
r VfR(')'(7'),f)d7' Jo l'
da--
g(q2/4)
+ 2n3Co f3/2,
since R ~ -nCo. Hence, since 9 (0) :::; e2Cor 9 (7') for 7' E [0, T] , we have
d;(o) (p, ')' (T)) :::; e2Cor ( :::; e2Cor . 2y'T
2 r y'T dd~ (0-) dO-) Jo cr g(q2/4) 1
+ 2~Co f3/2)
(ii) From the proof of (i) we have (take
T
.
= f in (7.28))
2..;¥ 1df3 --=- (a-) 12 dO- :::; -1. c (')') 2..,fi 0 dcr g(q2/4) 2..,fi
-1
2
2y'T 1d~ (a-) 12 r daJo dcr g(q2/4)
:::; 2.J!e2cor ( .c (')')
1
1
nG0 f. +_ 3
296
7. THE REDUCED DISTANCE
By the mean value theorem for integrals, there exists
T*
E (0, f) such that
d;3 12 1 nCo (0'*) ::; 2 ,=C h) + - 3 f. I-d 0' g(T.) V T
(iii) Let 'f} : [0, 2JT] ~ M be a minimal geodesic from p to q with respect to the metric 9 (f) . Then
o 3. The first variation of C-Iength and existence of C-geodesics Now that we have defined the C-length, we may mimic basic Riemannian comparison geometry in the space-time setting for the Ricci flow. We compute the first variation of the 'c-length and find the equation for the critical points of C (the C-geodesic equation). We also compare this equation with the geodesic equation for the space-time graph (with respect to a natural space-time connection) and prove two existence theorems for C-geodesics. 3.1. First variation of the 'c-Iength. Let (Nn, h (T)), T E (A, 0), be a solution to the backward Ricci flow. Consider a variation of the C 2-path , : h, T2J ~ N; that is, let G: h,T2J x (-e,e) ~N be a C2-map such that GI[Tl,T2jX{O}
= ,.
Convention: We say that a variation G (.,.) of a C 2-path , is C 2 if G s) is C 2 in (0', s) .
C:t '
Define
,s
~ GI[Tl,T2jX{S} :
[T1, T2J ~ N for -e < S < e. Let
. aG (T, s) = a,s aT (T)
x (T, s) =;= aT
and Y (T, s)
. as aG (T, s) = a,s as (T)
=;=
be the tangent vector field and variation vector field along tively. The first variation formula for C is given by
,s (T) , respec-
3.
FIRST VARIATION OF .c-LENGTH AND EXISTENCE OF .c-GEODESICS
297
LEMMA 7.15 (L:-First Variation Formula). Given a C 2-Jamily oj curves "fs : h, r2] ~ N, the first variation oj its L:-length is given by
1 d
1
'-
2(8yL:)("fs)~2dsL:("(s)= yrY·X
+
(7.29)
IT2 Tl
lT2 VTY' (~\7R - ~X - \7xX - 2Rc (X)) dr, 2 2r Tl
where the covariant derivative \7 is with respect to h (r) . REMARK 7.16. We use the notation (8y L:) ("(s) since tsL: ("(s), at a given value of s, depends only on "fs and Y along "fs. PROOF. We compute in a similar fashion to the usual first variation formula for length (see [72], p. 4ff for example)
d L: ("fs) = -d d -d
(7.30)
s
S
=
l
lT2 VT (R ("(S (r), r) + 1~ a"f (r) 12) dr ur
T2
h(T)
Tl
VT ((\7 R, Y) + 2 (\7y X, X)) dr ;
Tl
here ( . , . ) = h (r) ( . , . ) denotes the inner product with respect to h (r) . Using [X, Y] = [~~, ~~] = 0 and = 2Rc, we have
:Th
d
(\7yX, X) = (\7x Y ,X) = dr [h(Y,X)]- (Y, \7xX) - 2Rc(Y,X). Hence
1
d 2dsL:("(s) =
lT2 VT (12 (\7R,Y) + dr(Y,X)-(Y,\7x d X )-2Rc(Y,X) ) dr Tl
and integration by parts yields
l T2 VT-dd (Y, X) dr Tl
r
I1T2
= --2
Tl
1 '- (Y, X) dr yr
+ VT (Y, X) I~~ .
The lemma follows from the above two equalities.
D
REMARK 7.17. In comparison, the Riemannian first variation of arc length formula on is
(Mn, g)
(7.31)
where "fu : [0, b]
U~
d {b b du L ("(u) = - Jo (U, \7TT) ds + (U, T)l o ,
~ M is
a I-parameter family of paths, T
:u "fu, and ds is the arc length element.
~ ~ / I~ I'
298
7. THE REDUCED DISTANCE
3.2. The £-geodesic equation. The £-first variation formula leads us to the following. DEFINITION 7.18 (£-geodesic). If 'Y is a critical point of the £-length functional among all C 2 -paths with fixed endpoints, then 'Y is called an £-geodesic.
By the £-first variation formula, 7.19 (£-geodesic equation). Let (Nn, h (T)) , T E (A, 0), be a solution to the backward Ricci flow. A C 2-path 'Y : [Tl' T2J ~ N is an £-geodesic if and only if it satisfies the £-geodesic equation: 1 1 (7.32) V' x X - "2 V'R + 2 Rc (X) + 2T X = 0, COROLLARY
where X (T) ~ ~ (T). For the four terms in (7.32), (1) is the usual term in the geodesic equation, (2) comes from the variation of R in £, (3) comes from :rh, and (4) comes from -IT in £ via integration by parts. In local coordinates, the £-geodesic equation is d2'Yi i d'Y j d'Yk 1 i' i' d'Yk 1 d'Yi (7.33) 0 = dT2 +fjk ("( (T), T) dT dT -"2 h 3V'jR+2h 3 Rjk dT + 2T dT' where 'Yi = xi 0 'Y. We find it convenient to use the notation for the covariant derivative along the curve 'Y. Multiplying (7.32) by T yields
£.
D
T
vT dT (vTX) - 2V'R + 2vT Rc (vTX) = O.
(7.34)
Since the covariant derivative along the curve can be written as D dT V = V'x V , we may write D
vT dT (vTX) = vTV'X (vTX) = V'.;rx (vTX) along 'Y (T) , where the last two terms require extending vector field in a neighborhood of 'Y (T) . Note that
V'.;rx (vTX) = TV'XX
V (T) = -ITX
to a
+ vT (d~ vT) X 1
= TV'XX + "2X, which is different from -ITV' .;rxX because the curve. (7.35)
-IT must be differentiated along Using the convention (7.18) and Z (0') ~ d~~) = -ITX, we get 0'2 V'zZ - SV'R+O'Rc(Z) = O.
3.
FIRST VARIATION OF C-LENGTH AND EXISTENCE OF C-GEODESICS
299
EXAMPLE 7.20 (C-geodesics on Einstein solutions). Let (NO', ho (r)) , r E (0,00), be a 'big bang' Einstein solution to the backward Ricci flow
with RCho (r) = h~~). Then Rho (r) = ~ and the C-geodesic equation (7.32) is (7.36)
\lx X
3
+ 2rX = 0,
so that \l ~ (r3/2 ~)
= 0. Note that since \l is independent of scaling and ho (r) = rho (1), we have \lho(r) = \lho(l) is independent of r. Clearly the constant paths, where X = 0, are C-geodesics. More generally, reparametrize
, and define the path {3 by (3 (p) = , (f (p)) , where (7.37)
= l' (1-1 (r)).
r 3/ 2
Then /3 (p) implies
c:t =
~(f(p))f'(p)
\l/3(p)/3(p) =
=
~~(f(p))f(p)3/2, so that (7.36)
\lr3/2~(r) (r3/2~; (r))
= OJ
i.e., {3 is a constant speed geodesic with respect to ho (1) . Since solutions of (7.37) are given by f (p) = ~( 4 ) , the C-geodesics are of the form Po-P
,(r) = {3
(~-~), JTo VT
defined for r E (0,00), where (3 : (-00,00) geodesic with respect to ho (1) . Note that d, 1 1-(r) dr ho(r)
=VT I'(3
No is a constant speed
(2JTo VT2) 11 ---
-r 3/ 2 ho(l)
const r
That is,
r21~'(r)12 =:const. r ho(r) (Compare with (7.26) for the Ricci flat case.) In particular,
I~r
(r)1 ho(1) =
C:3/~t. In any case, the speed of ~ (r) tends to infinity as r - 0, whereas the speed of for all
~ (r) tends to zero as r -
00. Note
J7iooo I~
ro E (0,00), whereas J;o I~ (r)1 ho(l) dr = 00.
(r)1 ho(l) dr <
00
EXERCISE 7.21. Determine the C-geodesics for Einstein solutions with negative scalar curvature. What multiple of 1~ (r) 12 is constant? h(r)
300
7.
THE REDUCED DISTANCE
SOLUTION TO EXERCISE 7.21. Consider a maximal Einstein solution to the backward Ricci flow h (7),7 E [0, T), where Rc = -2(T1_r)h is negative. In this case h (7) = TT r h (0) (which is easy to see from independent of 7). The £-geodesic equation (7.32) for, (7) is V'x X
(7.38) Let (3 (p)
~
, (f (p)) , where
+ (~27
I
RCh(r)
being
_1_) X= O.
T-7
is defined by
(:f - T~ f) df)
I' (1-1 (7)) = exp ( / r
=y'T(T-7),
tii
J I (p)(T - I (p)).
Then /3 (p) ~ fp = (f (p)) (p) . Equation (7.38) and the definition of I imply
i.e.,
I' (p)
7 ~
I
=
V' (3(p)/3 (p) = V'
Let
f'(p)~(r) (I' (p) ~~ (7))
= !" (p) d,
(7) _
d7
=
I' (p).
~ (I' (p) ~
I' (p) I' (p) (~ __1_) d, (7) (If(P) ( T-7 ~) df)) .d, (7) T - 7
27
exp
d7
1__
~
b
=0. That is, (3 (p) is a constant speed geodesic with respect to h (0) . We make the rationalizing substitution x ~ I (p), so that
p
=
I ..;'T(P)f'
J
(p) dp (T - I (p))
=2
I
dx T - x2
=
_1 10 ( v'T g
x) .
v'T + v'T - x
That is,
I (p) = x2 = T
(_e~-=T:-P_-_1)
2
evTp + 1
or
1-1 (7) =
_1 log
v'T
(v'T + vfT) . v'T-vfT
Using, (7) = (3 (/- 1 (7)) and the fact that (3 is constant speed with respect to h (0) , we compute
Id'12 =T-71/3(1-1(7))12 d7 her)
T
d (1-1)
h(O)
const
2
d7
-
since h (7) = TT r h (0) and d (1-1)
1
d7
I' (p)
1 vfT (T - 7)"
= --=-----
T7(T-7)'
3.
FIRST VARIATION OF C-LENGTH AND EXISTENCE OF C-GEODESICS
301
EXERCISE 7.22. Estimate the speed of an .c-geodesic for a solution (Mn, 9 (r)) , r E [0, T), to the backward Ricci flow with
C IRm(x,r)1 ~ T-r
on M x [O,T).
Hint: What does the Bernstein-Bando-Shi estimate say about IV'RI? 3.3. Space-time approach to the .c-geodesic equation. We now compare the .c-geodesic equation for "{ with the geodesic equation for the graph i (r) = ("( (r), r) with respect to the following space-time connection (see also Lemma 4.3 in [100]): (7.39) (7.40) (7.41) (7.42)
where i,j,k 2 1 (above and below), and the rest of the components are zero. It is instructive to compare the Christoffel symbols above with the symbols Nt of the Levi-Civita connection Nfj for the metric It introduced in Exercise 7.4. For k 2 1, note that t~b = Nt~b is independent of N, whereas r-ab0= ·hmN-+oo N-O r ab for all a, b 2 o. Let r = r(O") ~ 0"2/4, Le., 0" ~ 2..,fT. We look for a geodesic, with respect to the space-time connection defined above, of the form
t
/3(0") ~ ("((r(O")), 0"2/4), where "( : [rl' r2J
/3i
M is a path. For convenience, let f3(0") ~ "((r(O")), ~ xi 0f3 ~ f3i for i = 1, ... , n, and /30 ~ xO 0/3 (so that /30 (0") = 0"2/4). --t
By direct computation, we have
0" d"{k dO" ="2 dr ' d/3° 0" = dO" 2'
df3k
and
We justify the change of variables from r to 0" via the geodesic equation with respect to t by showing that the time component of /3 satisfies the geodesic
THE REDUCED DISTANCE
7.
302
equation: 2 -0
~ + '"" d17 2 L...J
O~i,j~n
-.
(t 9~J.
0
-.
2
2
d(3~ d(3J = ~ (~) + t80 (jj (17)) (~)2 dl7 dl7 d17 2 4 2
jj)
1
1
= 2 - 2 ( ~2 )
(17)2
2 = O.
(This last equation justifies defining the time component of jj (17) as 17 2/4, and in particular, the change of variables 17 = 2ft.) For the space components, the geodesic equation with respect to t says that for k = 1, ... , n,
d2jjk 0= d172
+
L
r ij dl7
O~i,j~n
d2(3k
= d172 +
L
_k djji djj j r ij dl7 dl7 k d(3i d(3j dl7
l~i,j~n
+2
L
l~i~n
- k d(3i djjO dl7
riO dl7
- k djjO djjO dl7 dl7·
+ roo
This is equivalent to
o = (~)2 d2,l (7(17)) + '"" r~. (~d,i (7(17))) (~d,j (7(17))) 2 d7 2 L...J ~J 2 d7 2 d7 l~i,j~n
+~
(dd~ (7(17))) + 2 L
Rf
l~i~n
which, after dividing by
d2,k 0= d7 2 (7(17))
+
7
(~~~ (7(17))) (~) - ~ (~r Vk R,
= 172 /4, implies d,i
L
d,j
1 (d,k
rfj d7 (7(17)) d7 (7(17)) + 27
d7 (7(17))
)
l~i,j~n
+ 2 '"" L...J
k d,i 1 k Ri d7 (7(17)) - 2V R.
l~i~n
That is, in invariant notation and with X ~ ~, we have
VxX -
1
1
2VR + 2Rc(X) + 27X = 0,
which is the same as (7.32). Thus C-geodesics correspond to geodesics defined with respect to the space-time connection. In particular, ,(7) is an C-geodesic if and only if (3 (17) ~ , (17 2/4) is a geodesic with respect to the space-time connection V. Since t~b = limN ..... oo Nt~b' we also conclude that the Riemannian geodesic equation for the metric h on Nn x (0, T) (defined in Exercise 7.4) limits to the 17 = 2ft reparametrization of the C-geodesic equation as N ---t 00. EXERCISE 7.23 (Motivation for change of time variable). Show that if ---t N x [0, T] is a geodesic, with respect to the connection V, with
jj : [0, a-]
3.
FIRST VARIATION OF L-LENGTH AND EXISTENCE OF L-GEODESICS
/30 (0) = a and constant A.
d!: (0") i= a for 0" > 0, then /30 (0")
303
= A0"2 for some positive
SOLUTION TO EXERCISE 7.23. If /30 (0") = 7(0") , then the time component of the geodesic equation with respect to '\7 is
o= =
d2/30 d0"2
+
'" L....J
(- 0
O~i,j~n
d/3i d/3j dO" dO"
-)
r ij 0 f3
d27 1 (d7)2 d0"2 - 27 dO"
r?j
since = 0 when i ~ 1 or j ~ 1, and 7 (0") > 0 and ~~ (0") > 0 for 0" > 0, we have
d d7 _ log dO" dO"
r80
B
dT
d
~~
27
dO"
= - 2~' Hence, assuming
= da = da = - log ..jT
'
so that
d7 = C..jT dO" for some constant C > O. Since 7 (0) = 0, we conclude 0"2 7 (0") = C 2 4. 3.4. Existence of C-geodesics. Our next order of business is to establish the existence of solutions to the initial-value problem for the C-geodesic equation. In this subsection (M n , g (7)) , 7 E [0, T] , is a complete solution to the backward Ricci flow with curvature bound max {IRml, IRcl} ~ Co < 00 on M x [0, T] . We shall use the following LEMMA 7.24 (Estimate for speed of C-geodesics). Let (M n ,g(7)), 7 E [0, T], be a solution to the backward Ricci flow with bounded sectional curvature. There exists a constant C (n) < 00 depending only on n such that given 0 ~ 71 ~ 72 < T, if'Y : h, 72] ---+ M is an C-geodesic with lim ..jTdd'Y (7) T->Tl 7
=V
E
T-Y(TI)M,
then for any 7 E [71,72] ,
2
71 d'Y (7)1 d7 g(T)
~ e6CoT 1V12 + .
C(n)T -1 mm {T - 72, Co }
(e 6COT - 1) ,
where Co is as in (7.27) and 1V12 ~ 1V1;(Tl) . PROOF. Let O"i ~ 2y1ri. Define f3 : [0"1,0"2] so that
---+
2
(7.43)
lim 1df3 1 a-+al dO" g(a 2 /4)
= 1V12 .
M by f3 (0") = 'Y (0"2/4) ,
304
7. THE REDUCED DISTANCE
Since 72 < T, by (7.27) and the Bernstein-Bando-Shi derivative estimate (Theorem Vl-p. 224), there exists a constant C (n) < 00 such that
IV R (x, 7)1 ~
(7.44)
~ C2
C(n)Co Jmin {T
-
72,
COl}
for all (x, 7) EM x h, 72J. From the .c-geodesic equation (7.35), we compute
.!!..-I d(31 2 deI deI
=
g(u2/4)
(7.45)
09 (d(3 d(3)
OeI
deI' deI
= -eIRc
+ 2 / V rJ1t d(3 \
d(3 d(3) ( deI' deI
d(3) d~ deI' deI
eI 2
d(3)
/
+ 4 \ VR, deI
.
Applying the bounds (7.27) and (7.44) on the curvature and its first derivative, we have
-d Id(31 <2VTCo Id(31 +TC2 Id(31 deI deI g(u2/4) deI g(u2/4) deI g(u 2 /4) 2
2
2 < 3VTCo Id(31 deI g(u2/4)
(7.46)
+ C~ T3/2. 4Co
In view of the above ordinary differential inequality (7.46), given positive constants Cl and C2, consider the ODE (7.47) Then for positive solutions A (eI) , log (cIA
+ C2) (eI)
= log (cIA
+ C2) (eIl) + Cl (eI -
eII) ,
which implies
A (eI) =
eC1(u-ud A
(eIl)
+ ~:
(eC1(U-UI) -
1) .
Hence, comparing the solution to the ODI (7.46) with (7.43) to the solution to the ODE (7.47) with A (eII) we have
= IVI2 and taking Cl = 3VTCo and C2 = ~T3/2,
~ e 6Co v'T( Vr-v'Tl) IV 2 + C~~ . (e 6CO v'T( Vr-v'Tl) - 1) Id(312 deI g(u2/4) 12Co l
(7.48) since 0 ~ 71 ~ (7.44) of C2.
< e6CoT IVI2 + -
7 ~ 72
C~T (e 6CoT - 1)
12C3
< T. The lemma follows from this and the definition 0
LEMMA 7.25 (.c-geodesic IVP-existence). Let (Mn,9 (7)), 7 E [0, TJ, be a complete solution to the backward Ricci flow with bounded sectional
3.
FIRST VARIATION OF L:-LENGTH AND EXISTENCE OF L:-GEODESICS
305
curvature. Given a space-time point (p,71) E M x [0, T) and a tangent vector V E TpM, there exists a unique C-geodesic 'Y : [71, T) ~ M with lim VTdd'Y (7) 7~71
7
= V.
NOTATION 7.26. We shall usually denote 'Y as 'Yv. PROOF. The main idea of the proof is to change variables from 7 to a and to apply standard ODE theory. In local coordinates, the C-geodesic equation (7.35) for /3 (a) = 'Y (a 2 /4) is
(7.49) The above system of ODE is of the form 2 dda /3 2
= F (d/3 da,/3,a ) ,
where F is a smooth function. From elementary ODE theory, given any point p E M and tangent vector V E TpM, there exists a path
/3 : [2JTI, 2JTl + cJ
~
M
solving (7.49) with ~~ (2Ft) = V for some c > O. We prove by contradiction that /3 can be extended to an C-geodesic on [2Ft,2.jT). Suppose the maximal time interval of existence of /3 is
I Ig(0-2/4) : :; C. This
[2Ft, 2JT2) for some 72 < T. By Lemma 7.24, we have ~~
implies /3 (a) converges as a ~ 2JT2 to some q E M using the assumption that the solution g (7) is complete. Hence we can extend /3 beyond 2JT2. This is a contradiction and the lemma is proved. 0 We end this section by proving a lemma about the existence of Cgeodesics between any two space-time points (this may also be proved without using the calculus of variations). LEMMA 7.27 (Existence of minimal C-geodesics). Let (Mn, g (7)), 7 E
[0, TJ, be a complete solution to the backward Ricci flow with bounded sectional curvature. Given p, q E M and 0 :::; 71 < 72 < T, there exists a smooth path 'Y (7) : [71, 72J ~ M from p to q such that'Y has the minimal C-length among all such paths. Furthermore, all C-length minimizing paths are smooth C-geodesics. PROOF. Here we sketch a proof using the direct method in the calculus of variations. Let 'Yi : [71, 72J ~ M, i EN, be a minimizing sequence for the C-length functional. That is, C hi) ~ inf..y C (i') as i ~ 00, where the infimum is taken over all C1- paths i' : h, 72J ~ M from p to q. We have
1
72
71
VTldd'Yi (7)1 7
2 g(7)
d7:::; Chi) -
~ (7~/2 - 7~/2) 3
inf MX[71,T2)
R(X,7):::; C.
306
7.
Letting
Uj
~
THE REDUCED DISTANCE
2y'Tj and (3i (u) ~ 'Yi (u 2 /4) , the above formula says
1 0" 0"1
2
1
d{3i du
12 g(0"2/4)
du < C. -
From standard theory in the calculus of variations, we can conclude that there exists a subsequence such that 'Yi converges to a path 'Yoo : h,72] ---+ M with C (roo) = inf.-y C (1'). All C-length minimizing paths satisfy the Cgeodesic equation in the weak sense. By standard theory again, we have that such paths are smooth. 0
4. The gradient and time-derivative of the L-distance function In this section (Mn, 9 (7)), 7 E [0, T], shall again denote a complete solution to the backward Ricci flow satisfying the pointwise curvature bound max {IRml ,IRcl} ~ Co < 00 on M x [0, T], and p E M shall denote a basepoint. 4.1. L is locally Lipschitz. Before we study \l Land
fi,
we prove that, as a consequence of Lemma 7.13, L is locally Lipschitz. As is the case with this and the following one-sided Lipschitz result, we shall prove effective estimates. When we prove L is locally Lipschitz in the space variables, the following one-sided Lipschitz property will be used. LEMMA 7.28 (One-sided locally Lipschitz in time). Given a < 70 < T, let c ~ min {ffi, TIc?' lo} > O. For any 71 < 72 in (70 - c, 70 + c), qo EM, and q E Bg(o) (qO, c) , we have
where
PROOF. Let'Y : [0,72] ---+ M be a minimal C-geodesic from p to q. We define the piecewise linearly reparametrized path T/ : [0,71] ---+ M by
where
for 7 E L (q, 7J)
+ 72 -
271 ~ 7
Although T/ is only piecewise smooth, we still have C (T/). (We will use this fact a few times later.) Hence, since
[271 - 72,71].
~
27
4. GRADIENT AND TIME-DERIVATIVE OF THE L-DISTANCE FUNCTION
IRI
307
:SnC o,
L(q, 71) :S .c b) +
r
1
{T2 J2TI-T2
..jT
J2TI-T2
..jT
(R b (7) , 7) + 1~' (7) 12geT) ) d7 7
(R b (1> (7)) , 7) + 1~' 7
(1) (7)) . 4> (7) 12 ) d7 geT)
2nCo ( 723/2 - (271 - 72) 3/2) :S L (q, 72) + -32nCo (3/2 + -371 (7.50)
+2
(2 71 - 72 )3/2)
l
T2 -/1>-1 (7) Id-.2 (7) 12 d7. 2TI-T2 d7 g(q..-l(T))
Recall that f3 (a) is defined by (7.18). By Lemma 7.13(ii), (iii) there exists 7* E (0,72) such that
(7.51) Since 72 :S T - c:, by Shi's derivative estimate we have
IV R (x, 7)1 :S
C(n)Co vmin {T - 72, COl}
~ C2
for any (X,7) EM x [0,72]. From equation (7.46), we have
~ 1 df312
<
da da g(a2/4) -
3VTCo 1 df312
da g(a2/4)
+ C~ T3/2. 4Co
Integrating the above inequality over [a*, a] , we get for all a E [0,2y1T2], (7.52)
df3 1 :S e3v'TCola-a*ll d f3 (a*)1 + C~~e3v'TCola-a*l. Ida g(a2/4) da g(T.) 12Co 2
Noting that 1>-1 (7) = T2T2 using (7.52) and (7.51)
2
+ 71 :S 7 for 7 E
[271 - 72,72] , we can estimate
7.
308
THE REDUCED DISTANCE
where
C
..!...
3 -;-
2 4Coe+6CoT ( 2COT2 d~(T2) (p, q) e e 4T2
+
2nCo 3 T2
C~T) + 12C6
(using 10" - 0"*1 ::; 2VT). Combining this with (7.50), we obtain 7 L (q, Tl) - L (q, T2)
2nCo (3/2 (2 TI-702 )3/2) <3- To2 -
-
2nCo (3/2 + -3Tl - (2 TI- T2)3/2)
+ 2C3 (Ti/ 2 -
(2Tl - T2)1/2)
::; C1 (T2 - Td , where
o When combined with the uniqueness of the boundary-value problem for the L:-geodesic equation on small time intervals, Lemma 7.13 can also be used to prove the following; we shall give the proof elsewhere. LEMMA 7.29 (Short L:-geodesics are minimizing). Given V E TpM, there exists T* > 0 such that 'yvl[o,T.] is a minimal L:-geodesic. The next lemma and Rademacher's Theorem (Lemma 7.110) imply that L is differentiable almost everywhere on M x (0, T). LEMMA 7.30 (L is locally Lipschitz). The function L : M x (0, T) - t R is Lipschitz with respect to the metric 9 (T) + dT2 defined on space-time. PROOF. For any 0 < TO < T and qo E M, let E ~ min {f%, T~;o, lo} > O. Then for any Tl < T2 in (TO - E, TO + E) and ql, q2 E Bg(o) (qO, E),
IL (q1, Tl)
- L (q2, T2)1 ::;
IL (ql, Tl) -
L (q2, Tdl
+ IL (q2, Tl) -
L (q2, T2)1·
To prove that L is Lipschitz near (qo, TO), it suffices to prove (1) and (2) below. (1) L (', Td is locally Lipschitz in the space variables uniformly in Tl E (TO - E, TO + E). Let dT denote the distance function with respect to the metric 9 (T) and let 'Y : [0, TIl - t M be a minimal L:-geodesic from p to ql. Let a : h, Tl + do (q1, q2) 1- t M be a minimal geodesic of constant speed 1, with respect to 9 (0) , joining ql to q2. Then 'Y "'-...-/ a :
[0, Tl
+ do (ql, q2)l - t M
4. GRADIENT AND TIME-DERIVATIVE OF THE L-DISTANCE FUNCTION
309
is a piecewise smooth path from p to q2. We estimate, using I~~ (7)1:(T) :::; e2Co T 1dOl (7) 12 = e2Co T that g(O)
dT
L (q2, 71
,
+ do (qI, q2) )
:::; .c (')') +
1
+dO (Ql,q2) Vi (R (Q (7),7)
T1
+ 1ddQ 7
T1
2 (nCo
:::;L(ql,71)+ :::; L (ql, 71)
(7)1
+3 e2CoT )
(
2
)
d7
geT)
3/2 3/2) (71+ dO(ql,q2)) -71
+ C 1do (qI, q2) .
By Lemma 7.28 we have
L (q2, 7t)
+ do (ql, q2)) + C 1do (ql, q2) :::; L (ql, 7t) + C 1 do (ql, q2) :::; L (ql, 71) + C 1d (ql, q2),
:::;
L (q2, 71
Tl
where we have used do (ql, q2) :::; eCoT dTl (ql, q2). By the symmetry between ql and q2 we get
IL (q2, 71)
-
L (ql. 71)1
:::;
Cl dTl (ql. q2) .
(2) L (q, .) is locally Lipschitz in the time variable uniformly in q E Bg(o) (qO, c). For any 71 < 72 in (70 - c, 70 + c), let I : [0, 71l ~ M be a minimal C-geodesic from p to q and let f3 : h, 72l ~ M be the constant path f3 (7) = q. Then I f3 : [0, 72l ~ M is a piecewise smooth path from p to q. Hence 0"""./
L (q, 72)
:::;
.c (f) +.c (f3) = C (')') + 1T2 ViR (q, 7) d7 T1
o < - L (q,71 ) + 2nC 3 ( 723/2
:::; L (q, 71)
+ C1 (72 -
_
3/2) 71
71),
where C 1 depends only on Co and T. Combining this with Lemma 7.28, we obtain where (7.53)
o 7.31. L is differentiable almost everywhere on M x (0, T) and L E Wl~'~ (M x (0, T)). COROLLARY
PROOF.
See Lemmas 7.110 and 7.111.
o
7. THE REDUCED DISTANCE
310
4.2. Gradient of L. We compute the gradient of L via the first variation formula for C. Since L (', r) is not smooth in general, the gradient is defined in the barrier sense as described below. Let 'Y : [0, r] --+ M be a minimal L:-geodesic from p to q so that L (q, r) = C ('Y). For any point x in a small neighborhood U of q and any r E (r - c, r + c) with small c > 0, let 'Yx,r : [0, r] --+ M be a smooth family of paths with 'Yx,r (0) = p, 'Yx,r (r) = x and 'Yq,'f = 'Y. (Recall that our definition of a smooth variation says that
'Yx,r (~2) is a smooth function of (0', x, r) .) Define L : U x (r - c, r
+ c) --+ R
by
L(x,r) =C("tx,r)' Then L (x, r) is a smooth function of (x, r) when r > 0, L (x, r) ~ L (x, r) for all (x, r) E U x (r - c, r + c), and L (q, r) = L (q, r). That is, the function L (.,.) is an upper barrier for L (".) at the point (q, r) . Given a vector Y (r) at q, let q (s) be a smooth path in U with q (0) = q and ~ (0) = Y (r). Consider the smooth I-parameter family of paths 'Ys ~ 'Yq(S),T : [0, r] --+ M. Let Y (r) ~ %s Is=o 'Ys (r) denote the variation vector field along 'Y (r). By (7.29), (7.32), and Y (0) = 0, we have
~ dl s=OL(q(s),r) ~ VL(q,r)·Y(r)= ds = (8 y C)("t) =2v'=r Y(r)·X(r). Hence
v L (q, r) =
2../fX (r) . It follows from Lemma 7.30 that L (.,7) is differentiable a.e. on M. Suppose L(·,r) is differentiable at q. Since L(.,r) is an upper barrier for L(',r) at the point q, it is easy to see that VL(q,r) = VL(q,r) = 2../fX (r). Suppose there is another minimal C-geodesic 'Y' : [0, r] joining p to q. Then we can construct another barrier function L' as above; the same proof will imply V L (q, r) = 2../fX' (r), where X' (r) = ~ (r). Now both 'Y and 'Y' satisfy the same C-geodesic equation and 'Y (r) = 'Y' (f) = q and ~ (f) = ~ (r) = V L (q, r) . By the standard ODE uniqueness theorem, we conclude that 'Y (r) = 'Y' (r) for r E [0, r]. Hence if L (', r) is differentiable at q, then the minimal C-geodesic joining (P,O) to (q, r) is unique. Convention: If the function L (', f) is not differentiable at q, then by writing V L (q, r) = 2..jTX (r),8 we mean that there is a smooth function L satisfying L (x, r) ~ L (x, r) for x E U, L (q, r) L (q, r), and V L (q, r) = 2..jTX (r). We have proved the following. 8Note that X (f) depends on the choice of minimal .c-geodesic, which may not be unique.
4.
GRADIENT AND TIME-DERIVATIVE OF THE
L-DISTANCE
FUNCTION
311
LEMMA 7.32 (Gradient of L formula). The spatial gradient of the Ldistance function is given by
V'L (q, f)
(7.54)
= 2VfX (f),
where X (f) = ~ (f), for any minimal C-geodesic I : [0, f] --t M joining p to q. Furthermore if L (', f) is differentiable at q, the minimal C-geodesic joining (p,O) to (q, f) is unique.
REMARK 7.33. The analogy of (7.54) in Riemannian geometry is as follows. Let dp(x) ~ d(x,p) and suppose dp is smooth at q E
(Mn,g).
Define I : [0, b] --t M to be the unique unit speed minimal geodesic from p to q. By the first variation formula (7.31), for any U E TqM, (V'dp(q),U)
=
d~lu=o L(,u) = h(b),U),
provided the IU : [0, b] --t M satisfy 10 = I, IU (0) = p and U. That is, V'dp (q) = '1' (b).
Iu Iu=o IU (b) =
Taking the norm of (7.54), (7.55) IV'LI2 (q, f) = 4f IX (f)1 2 = -47
R(q, f) + 4f (R (q, f) + IX (f)1 2) .
The reason we rewrite this in a seemingly more complicated way is that both Rand R + IX (f)1 2 are natural quantities. 9 4.3. Time-derivative of L. Next we compute the time-derivative of L. This time we need to choose Ix,r used in subsection 4.2 above a little more carefully. Given (q, f), let I : [0, f] --t M be a minimal C-geodesic from p to q so that L (q, f) = C (,). We first extend I to a smooth curve I : [0, f + oS] --t M for some oS > 0, and then we choose a smooth family of curves Ix,r to satisfy Ix,r (0) = p, I'Y(r),r = II[O,r] and Ix,r (T) = x. Define
(7.56) for (x, T) E U x (f - oS, f + oS) . We compute, using the chain rule and (7.54),
O£( _)_O£(,(T),T) aT q, T aT
=
d~ I
r=T
[£ (, (T) , T)]
- V' £ . X
r=T
=
:7 IT~' [[ v/f (R ("(f), f) + 1~>f)l}fl- 2v/f IX
(7')1'
= Vf (R (, (f), f) + IX (f)1 2) - 2Vf IX (f)1 2 . It follows from Lemma 7.30 that L (q,.) is differentiable a.e. on (0, T). As discussed in subsection 4.2 above, if L (q,') is differentiable at f, then 9As it is the integrand of the L:-Iength,
..fi (R + IXn
is a natural quantity.
7. THE REDUCED DISTANCE
312
~~ (q, f) ~~ (q, f). If L (q,') is not differentiable at 1', then by writing ~~ (q, f) = JT h (f), f) + IX (1')1 2) - 2JT IX (1')1 2, we mean (this is our convention below) that there is a smooth function L (x, I) satisfying L (x, I) ~ L (x, I) for x E U and I E (1' - €, l' + €), L (q, f) = L (q, f) and (q, f) = JT h (f), f) + IX (1')12) -2JT IX (1')1 2 . Now we have proved
(R
fi
(R
7.34 (Time-derivative of L formula). The time-derivative of the L-distance function is given by LEMMA
~~ (q, f) = -# (R (q, f) + IX (1')12) + 2#R (q, f),
(7.57)
where X (f) = ~ (f) , for any minimal C-geodesic 1 : [0,1'] to q.
---+
M joining p
In the case where (Mn,g (I) == go) is Ricci fiat, we have 1 (I) = f3 (2JT) , where f3 : [0,2vT] ---+ M is a constant speed Riemannian geodesic with respect to go. Thus hand L (q, I)
~=
Jr/3 (2JT) and 1/3 (2JT) 1 == ~~. On the other
= d~~2. Hence
aa,L (q,,)_ = - d41'3/2 (q, p) 2 '= = -VI
1
d , _1 2 d, (I) ,
agreeing with (7.57).
5. The second variation formula for C and the Hessian of L Recall that the second variation of arc length formula of a geodesic 1 is (7.58)
2
d 2 1u=o L (,U) du
= fob (1\7"yU1 2- (\7"yU, i)2 - (Rm (U, i) i, U)) ds + (\7u U, i) Ig, where IU : [0, b] ---+ M is parametrized by arc length s and satisfies 10 = 1 and U ~ Iu=o IU' This formula is fundamental to Riemannian geometry for a variety of reasons. For example, on a complete Riemannian manifold, any two points can be joined by a minimal geodesic, in which case 2 PdU 1u=o L (,U) ~ for endpoint-preserving variations. The second variation of arc length can also be used to bound from above the Hessian of the distance function. In this section we consider the analogous second variation formula for C-Iength. The first variation of C-Iength determines the C-geodesic equation and is related to the space-time connection. The second variation formula for Clength at an C-geodesic 1 : [0,1'] ---+ M is related to the space-time curvature and hence Hamilton's matrix Harnack quadratic. In the case of a minimalCgeodesic, it also gives an upper bound for the Hessian of the barrier function
Ju
°
5. THE SECOND VARIATION FORMULA FOR C AND THE HESSIAN OF L
313
t defined in subsection 4.2 of this chapter. At a point q where L (', r) is C2, the second variation formula gives an upper bound for the Hessian of the L-distance function and tracing this estimate yields an estimate for the Laplacian of L. The Hessian upper bound will be very important in discussing the weak solution formulation later. In this section (M n , 9 (7)), 7 E [0, T], will denote a complete solution to the backward Ricci flow satisfying the pointwise curvature bound max {IRml ,IRcl} :S Co < 00 on M x [0, T], and p E M shall denote a basepoint. 5.1. The second variation formula for C. Let r E (0, T) and let ---t M be an C-geodesic from p to q. Let IS : [0, r] ---t M, s E ( -E, E), be a smooth family of paths with 10 (7) = I (7). Recall that our convention about the smoothness of a variation IS of I is that !3s (0') ~ IS ( ~2) is required to be a smooth function of (0', s) . It is easy to see that the nondifferentiability of IS at 7 = causes no trouble in the following calculation. This would also be clear if we use (7.19) to do the calculation. Define ~ (7) ~ X (7, s) and ~ (7) ~ Y (7, s) , so that [X, Y] = 0.10 We I : [0, r]
°
also write Y (7) ~ Y (7, 0). Note that Y (U42, (0', s) . Recall the first variation formula (7.30)
s)
is a smooth function of
which holds for all s E (-E, E). Differentiating this again, we get
(8~C) b) ~ d2~s~IS) Is=o =
Io -IT (Y (Y (R)) + 2 (V'yV'yX,X) + 21V'yXI f
2)
d7.
Now since [X, Y] = 0, (V'yV'yX, X)
= (V'yV'xY, X) =
(R(Y,X) Y, X)
+ (V'xV'yY, X).
Hence (7.59)
(8~C) b) = {f -IT (
Jo
Y (Y (R)) + 2 (R (Y, X) Y, Xl ) d7. +2 (V' x V'y Y, X) + 21V'y XI
10Alternately, given any vector field Y along "I, there exists a family of paths "Is such that ~Is=o = Y. In this case, we extend Y by defining ~ = Y for s E (-c:,c:), so that [X,YJ = O. Technically, X and Y are sections of the bundle G*TM on [O,fJ x (-c:,c:), where G (7, s) ~ "Is (7).
314
7. THE REDUCED DISTANCE
On the other hand, we compute d dT (VyY,X) = (VxVyY,X) + (VyY, VxX)
+ ~~ (VyY,X) + (
(:T V)
y
Y,X).
Now ~ = 2Rc and
((:T V) y y,x)
= 2 (Vy Rc) (Y,X) - (Vx Rc) (Y, Y).
Hence d dT (Vy Y, X)
(7.60)
= (V x Vy Y, X) + (Vy Y, V x X) + 2 Rc (Vy Y, X) + 2 (Vy Rc) (Y, X) - (V x Rc) (Y, Y).
Suppose (7.61)
Y (0)
=0
(this and the fact that ..jTX (T) has a limit as T --t 0 are used to get the third equality below). Then applying (7.60) to (7.59) and integrating by parts, we compute
(&}.c) b) = foT .jT (Y (Y (R)) + 2 (R (Y, X) Y, X) + 21Vy X12) dT +2 =
1 T
o
d~ (VyY,X) - (VyY, VxX) - 2Rc (VyY, X) ) -2 (Vy Rc) (Y, X) + (V x Rc) (Y, Y)
'- (
yT
~
foT .jT (Y (Y (R)) + 2 (R (Y, X) Y, X) + 21VyX12) dT
+2
1 T
o
(VyY, VxX) - 2Rc(VyY,X) ) dT -2 (Vy Rc) (Y, X) + (V x Rc) (Y, Y)
'- (
-
yT
+ 2.jT (Vy Y, X) I~ - foT '=
= 2YT (VyY
X)
+
,
+2
1 T
o
'- (
yT
Jr
(Vy Y, X) dT
1T .jT (Y(Y(R))-VYY.VR 0
)
+2 (R (Y, X) Y,X) + 21VyXI 2
dT
(VyY, [VxX + 2Rc (X) - !VR + 2~X]) ) dT -2 (Vy Rc) (Y, X) + (V x Rc) (Y, Y) -
= 2Vf (VyY, X) +
foT .jT (V~,yR + 2 (R (Y,X) Y, X) + 21VyX12) dT
+ foT .jT (-4 (V y Rc )(Y, X) + 2 (V x Rc )(Y, Y)) dT,
5.
THE SECOND VARIATION FORMULA FOR
I:.
AND THE HESSIAN OF
L
315
where we used the £-geodesic equation (7.32) to get the last equality; in the above, \7~yR ~ Y (Y (R)) - (\7yY) (R) ,
= Hess (R) (Y, Y).
That is, LEMMA 7.35 (£-Second variation - version 1). Let f E (0, T) and let "I : [0, r] ---t M be an £-geodesic from p to q and let Y ~ gs "Is for some smooth variation "Is of "I with Y (0) = O. The second variation of £-length is given by (8~£) ("t) =
+
(7.62)
2v'r (\7yY,X) (f)
i
T
o
r= ( \7~yR + 2 (R (Y, X) Y, X) + 21\7y XI 2 ' -4 (\7y Rc) (Y, X) + 2 (\7 x Rc) (Y, Y)
y 7
REMARK 7.36. Note that by (7.54) we have (8£) ("t)
)
d7.
= 2.jTX (f) . Hence
whose value only depends on Y (7) defined along "I (7). This is analogous to considering (Hess f) (Y, Y) = YY (f) - \7 y Y . \7 f. We now rewrite the £-second variation formula in a better form, which relates to Hamilton's matrix Harnack quadratic, Le., the space-time curvature. ll Since
!
[Rc (Y (7) , Y (7))]
= (:7 RC)
(Y, Y) + (\7 x Rc )(Y, Y) + 2 Rc (\7 x Y, Y) ,
integrating by parts, we have
- foT -IT ( =
:7
RC) (Y, Y) d7
foT -IT (;7 Rc (Y, Y) + (\7 x Rc) (Y, Y) + 2 Rc (\7 x Y, Y) )
d7
- -ITRc (Y, Y)I~.
llSee Section 5 of Chapter 8 for the reason why the space-time curvature is Hamilton's matrix quadratic.
7.
316
Hence (7.62) and Y (0)
THE REDUCED DISTANCE
= 0 imply
1
2 (6}.c) h) -
Vf(V'yY,X)
-
+ JTRc(Y,Y)I~
1T JT ((:r Rc+ 2~ RC) (Y, Y) + ~V'},YR) dr + 1T JT ((R (Y, X) Y, X) -IRc (Y)12) dr + 1T JT (-2 (V'y Rc) (Y, X) + 2 (V' Rc) (Y, Y)) dr + 1T JTIV'x Y + Rc (Y)1 2 dr, =
X
where we used V' x Y = V'y X. Let H (X, Y) denote the matrix Harnack expression
H (X, Y) (7.63)
~ -2 (:r RC) (Y, Y) -
V'},yR + 21Rc (Y)1 2 -
~ Rc (Y, Y)
- 2 (R (Y, X) Y, X) - 4 (V' x Rc) (Y, Y) + 4 (V'y Rc) (Y, X) .
By substituting the definition of H (X, Y) in the above formula, we obtain (6}.c)
h) -
=-1T
(7.64)
2Vf (V'yY,X) (f) + 2VfRc (Y, Y) (f)
JTH(X,Y)dr
+
1T
2JTIV'xY+Rc(Y)12dr.
An even nicer form is LEMMA 7.37 (.c-Second variation - version 2). Let l' E (0, T) and let ---t M be an .c-geodesic. If Y (r) ~ IS (r) , for a smooth variation IS of I, satisfies Y (0) = 0, then
ts
I : [0,1']
(7.65)
(6}.c)
h) -
-1T
JTH (X, Y) dr +
2Vf (V'yY,X) (f) + 2VfRc (Y, Y) (f)
1T
2JT IV' x Y + Rc (Y) -
= IY;1 2
2~ YI2 dr.
PROOF. Since lV'x Y +Rc(Y) -
2~Y12
1 r 2 1 d 2 1 2 = lV'x Y +Rc(Y)1 - 2rdr IYI + 4r21YI ,
1 r
= lV'x Y + Rc(Y)12 - - ((V'x Y , Y) + (Rc (Y), Y)) + -42 1YI2
5. THE SECOND VARIATION FORMULA FOR C AND THE HESSIAN OF L
317
we have
where we integrated by parts. Note that the assumption of the lemma IY (T)12 = 0. The lemma implies that (~2) is smooth in (J and limT->o now follows from (7.64). 0
Y
Jr
We now consider a special case of this formula. As above, let 'Y : [0, f] --+ M be an ,C-geodesic. Fix a vector Y r E T-y(r)M and define a vector field Y (T) along 'Y by solving the following ODE along -"(:
V'x Y = -Rc(Y)
(7.66)
I
+ 2TY'
TE
[0,1'],
Y (f) = Yr. Note that any vector field along 'Y can be considered as a variation vector field. In particular, we may extend Y (T) to Y (T, s) for some smooth variation of 'Y. REMARK
7.38. Equation (7.66) is equivalent to
JrY
which essentially says is parallel with respect to the space-time connection. Note that if 'Ys : [0,1'] --+ M is a I-parameter family of paths such that X = 'Ys and Y = :s 'Ys, then [X, Y] = and (7.66) may be rewritten as
°
:T
( V'
X+ Rc - 2~g) (Y) = 0,
which is reminiscent of the gradient shrinker equation. From (7.66) we compute (7.67)
ddT 1Y12
Solving this
ODE,
=
dd [g (Y, Y)] T
= 2 (V' x Y, Y) + 2 Rc (Y, Y) = ~ 1Y12 .
we have
(7.68) Thus Y (0)
= O. Hence by (7.65) we have the following.
T
7. THE REDUCED DISTANCE
318
LEMMA 7.39 (.c-Second variation under (7.66)). If f E (0, T), YT E T'Y(T)M and Y is a solution to (7.66) with Y (f) = YT, then the second variation of .c-length is given by (8~.c)
(7.69)
=-
h) - 2Vf (\7yY, X) (7) + 2VfRc (Y, Y) (f)
loT .,fTH (X, Y) d7
112
+ IY
5.2. Hessian comparison for L. Corresponding to the .c-second variation formula is an upper bound for the Hessian of the L-distance function, which we derive in this subsection. Given any (q, f), let I : [0,1'] --+ M be a minimal .c-geodesic from p to q so that L (q, f) = .c (,), Fix a vector Y E TqM - { and define the vector field Y (7) along I to be the solution to (7.66) with Y (f) = Y. Let IS : [0,1'] --+ M be a smooth family of curves for s E (-c, c) with
IT}
I
d,s (7) = Y (7) and (\7yY) (f) = 0. ds s=o . Then there exists a small neighborhood U of q, 8 E (0, c], and a smooth family of curves IX,T : [0,7] --+ M for (X,7) E U x (1' - 8, l' + 8) satisfying IX,T (0)
= p,
''Ys(T),T
= IS,
IX,T (7)
and
=x
for s E (-8,8). We define L (x, 7) ~ .c hX,T)' Then L hs (f), f) = .c hs). Since L (', .) is an upper barrier function for the L-distance function L (.,.) at (q, f), we have (Hess(q,T) L) (Y, Y) ::; (Hess(q,T)
when L (', f) is C 2 at q. Since (\7y Y) (f) =
°
L) (Y, Y)
and
dd 2 I .c hs) , s s=o S s=o combining this with Lemma 7.39, we get the Hessian Comparison Theorem for L. ( Hess(q,T)
L) (Y, Y)
2
= dd 2
I L hs (f) ,f) =
2
COROLLARY 7.40 (Inequality for Hessian of L). Given l' E (0, T), q E M, and Y E TqM, let I : [0,1'] --+ M be a minimal .c-geodesic from p to q. The Hessian of the L-distance function L (', f) at q has the upper bound (7.70)
r
(Hess(q,T)L) (Y,Y)::; - Jo .,fTH(X,Y)(7)d7+
IY(1')1 2
.JT
- 2VfRc (Y, Y) (f), where Y (7) is a solution to (7.66) with Y (f) = Y and H is the matrix Harnack expression defined in (7.63). Equality in (7.70) holds when L (', f) is C 2 at q and Y (7) is the variation vector field of a family of minimal .c-geodesics.
5.
THE SECOND VARIATION FORMULA FOR
C
AND THE HESSIAN OF
L
319
If L (., r) is not C 2 at q, the above inequality is understood in the barrier sense; this is our convention below. More precisely, there is a smooth
function
L (., r)
defined near q such that ( Hess(q,T)
equality (7.70) and
L (., r)
i) (Y, Y)
satisfies in-
is an upper barrier function for L (., r) with
L (q, r) = L (q, r) . LEMMA 7.41 (Upper bound for Hessian of L). Fix To E (0, T). Given (0, To], q E M, and Y E TqM, the Hessian of the L-distance function L (., r) at q has the upper bound
rE
where C2 is a constant depending only on n, To, T, Co (Co 2: sUPMx[o,T]IRml is as in (7.27)).
:T
PROOF. From Shi's derivative estimate and the equation for Rc, there is a constant Cl depending on n, T - To, and Co such that Rcl ' IVV RI , IV Rcl , and IVV Rml are all bounded by CIon M x [0, To]. From (7.52) and Lemma 7.13(ii) and (iii), we get
Ifr
e 6CoT I fTX (7 )1 2 + 12C5 C~T e6CoT I.jTX (r)12 < V'* * g(T.)
2
nCo) C T_e6CoT 1 < e 6CoT ( --c, (r) + r + _2
-
2Vf
12C5
3
nCor + __ e2COT < e 6CoT ( d2 -
<
-
3
(1 +
_ g(T)
4r
d;(T) (p, q)) C
r
nco) (p q) + r 3
,
C T_e 6CoT + _2 12C5 2
2,
where C2 is a constant depending only on n, To, T, and Co. Hence, from (7.63), we have
IH (X, Y)I (r) S
+
CllY (r)12
+ CO IY (r)12
coH
d;(f)~' q)) C,IY (T) I'
+ c, S C2
1+
~
r
(1 + ~(f)T(P,q)) c, ·IY
(1 +.!. + r
d;(T)
~,q))
rr
(T)I'
IY (r)12.
320
7. THE REDUCED DISTANCE
Plugging the above estimate into (7.70) and using we get
(Hess(q,r)
L) (Y, Y):::; ( C2 + C2
~(r) (p,q)
Vi
IY (7)12 = ~ IY (f)1 2 ,
1)
+ Vi IY (f)1
2
.
o 5.3. Laplacian comparison theorem for L. Tracing the Hessian comparison theorem for L in (7.70), we obtain the Laplacian comparison theorem for L. We adopt the notation in Corollary 7.40. Let {Eil~=l be an orthonormal basis at q = 'Y (f) . For each i, we extend Ei to a vector field Ei (7) ,7 E [0, f] , along'Y to solve the ODE (7.66) with Ei (f) = E i . Below, Ei will stand for either Ei E TqM or Ei (7) , which will be clear from the context. Taking Y = Ei in (7.70) and summing over i, we have n
ilL (q, f) = L Hess(q,r) L (Ei' Ei ) i=l
n
- 2VfL Rc (Ei' Ei) (f) . i=l We compute, without assuming (Ei (f), E j (f)) =
dij,
that
d d d7 (Ei' E j ) (7) = d7 [g (7) (Ei (7), Ej (7))]
= 2Rc(Ei,Ej) + (\1XEi,Ej) + (Ei' \1xEj) = 2Rc(Ei,Ej) + (-RC(Ed + 2~Ei,Ej)
+ (Ei' -Rc(Ej) + 217Ej) 1
= - (Ei' E j ) (7). 7
Hence (7.72)
for all 7 E [0, fl. When i = j, we recover (7.68), which says lEi (7)12 = 7/f.
5. THE SECOND VARIATION FORMULA FOR
.c AND THE HESSIAN OF
L
321
Using the Einstein summation convention (expressions with i repeated are summed from 1 to n), we now simplify L~1 H (X, E i ) (7):
i=1
= -2
(:7
RC) (Ei' Ei ) - \7 Ei \7 EiR + 21Rc (Ei)12 -
- 2 (R (Ei' X) Ei, X)
+ 4 (\7 Ei Rc) (Ei' X)
~ Rc (Ei' Ei)
- 4 (\7 x Rc) (Ei' Ei)
8
7
7
= -2[Rc (Ei' Ed] +4Rc(\7xEi,Ei) - -=~R+2-= IRcl 87 7 7 17
7
77
7
- --=R + 2-= Rc (X, X) = -2!
7
.
+ 4-=7 dlV (Rc) (X)
2
7
- 4-=\7 xR 7
(~R) + 4Rc ( -Rc (Ei) + 2~Ei,Ei) - ~~R+ 2~ IRcl 2
1 7 7
- -=R + 2-= Rc (X, X) - 2-=\7xR 7 7 7
=
~7 (_2 8R _ ~R _ 21Rcl 2 87
R 7
+ 2Rc (X,X)
- 2\7 x R) .
Recall from (Vl-p.274) that Hamilton's trace Harnack expression is 8R R (7.73) H(X)~H(X)(7)~--8 -2\7R·X+2Rc(X,x)-7
(in (Vl-p.274) let t 21Rcl 2 , we get
=
7
and replace X by -X). Using ~~
-7
=
-~R-
n
LH(X,Ei) (7) = ~H(X).
(7.74)
i=1
7
Hence ~L(q,f')::; -
Io -_-H(X)d7+ f 73/2
o
7
1
n
7
y7
= --=K +
£
-
n £
-2..foR(q,f')
y7
'= 2yf'R(q,f') ,
where (7.75) We call K the trace Harnack integral. Dropping the bars on the have the Laplacian comparison theorem for L. LEMMA
'Y :
[0,7]
(7.76)
---+
7.42 (Inequality for ~L). Given 7 E (0, T) and q M be a minimal C-geodesic joining p to q. Then 1 n ~L (q, 7) ::; --K + . r::; - 2...fiR. 7
y7
E
7'S,
we
M, let
322
7. THE REDUCED DISTANCE
As before, if L (', r) is not C2 at q, by our convention the above inequality is understood in the barrier sense. 6. Equations and inequalities satisfied by Land
.e
In this section (Mn,g(r)), r E [O,Tj, will be a complete solution to the backward Ricci flow satisfying the curvature bound max {IRml ,IRcl} ~ Co < 00 on M x [0, T], and p E M will be a basepoint. Given a point q E M and f E (0, T), let, : [0, f] -+ M be a minimalC-geodesic from p to q and let X (r) ~ ~. 6.1. The C-Iength integrand and the trace Harnack quadratic. Now using the C-geodesic equation (7.32), we compute the evolution of the C-Iength integrand as
:r = =
(R (r (r), r) + IX (r)I;(7)) oR or +VR·X+2Rc(X,X)+2(Vx X ,X) oR or + VR· X + 2Rc(X,X) + \/ VR - 4Rc(X) oR +2VR· X -
=-
Or
1) -:;.X,X
1 2 2Rc(X,X) - -IXI .
r
Hence
d~ (R + IXI2)
(7.77)
=
-H (X) - ~ (R + IXI2) ,
where H (X) is defined in (7.73). In another form, we have
Recall that
K = K (r, f) =
loT r 3/ 2H (X) dr
defined in (7.75). Multiplying (7.77) by r 3 / 2 and integrating (by parts for the second equality), we get
-K(r,f)
= loT [r3/2 = f3/2
! (R +
+ r 1/2 (R + IXI2) ] dr
(R (, (f), f) + IX (7)12) - ~ loT r 1/2 (R + IXI2) dr
= f3/2 (R (r (f) ,f)
Hence
IXI2)
+ IX (f)1 2)
-
~C (r) .
6. EQUATIONS AND INEQUALITIES SATISFIED BY LAND l LEMMA
323
7.43 (L-distance and trace Harnack quadratic). Let f E (0, T). - t M is a minimal C-geodesic from p to q. Then
Suppose, : [0, f]
f3/2 (R (q, f)
(7.78)
+ IX (f)1 2)
= -K (" f)
+ ~L (q, f) .
Using (7.78), we can rewrite (7.57) and (7.55) as follows. For convenience we include (7.76) below as (7.81). LEMMA 7.44. Let f E (0, T), let, : [0, f] - t M be a minimal C-geodesic from p to q, and let K = K (" f) be given by (7.75). Then, at (q, f) ,
(7.79)
8L 1 1 -=-K--L+2 'TR 8T f 2f YT,
(7.80)
2 _ 4 2 IVLI = -4TR- - K +-L
VT VT' ilL < - -1 K + -n - 2v!f. f R. f VT
(7.81)
These may seem like strange ways to rewrite equations (7.57) and (7.55). A motivation is given by (7.76), which involves the integral K of Hamilton's trace Harnack quadratic, which naturally arises from the second variation formula for C-Iength. Note that besides K, these formulas do not explicitly contain the quantity X. When the minimal C-geodesic is not unique, the quantities f;:, IVLI 2 , ilL and K are defined using the choice of,.
6.2. Inequalities for L. Combining (7.79) and (7.80) with (7.81), we get LEMMA
7.45. At (q, f) the L-distance function L (x, T) satisfies
8L 1 n < -IlL- - L + 8T 2f VT' 8L 1 2 1 . '= n 8T - ilL + 2VT IV LI - 2fL - 2yfR + VT ~ 0.
-
(7.82) (7.83)
Equation (7.82) already exhibits the advantage of the L-distance over the usual distance function in Riemannian geometry in the setting of the Ricci flow. It is a subsolution to a backward heat equation. Recall that £ (x, T) = 2JTL (x, T). Then at (q, f)
8£ :::; 2v!f. (-ilL - I_L 8T 2T
+~) + ~L = VT
VT
-Il£
+2n,
so that 7.46 (£ - 2nT supersolution of heat equation). At (q, f) the function £ (x, T) satisfies LEMMA
(7.84)
8£ 8T
-
+ ilL:::; 2n,
7.
324
THE REDUCED DISTANCE
that is, (7.85) Now we prove a lemma which describes the limiting behavior of L (q, r) as r -+ 0+. We will use it to estimate the minimal value of 2~L(., r) (and likewise L) over M. LEMMA 7.47 (L tends to its Euclidean value as r
-+
0). We have
Jim L(q,r) = (dg(o) (p,q))2. r--+O+
(7.86)
Hence L satisfies lim L (q, r) 1'--+0+ [dg(o) (p, q)]2
/2-/7
=
1.
PROOF. We only need to prove the first equality above; the second equality follows by definition. In Lemma 7.13(i), let 7 = 72 = r and let 'Y: [0, r] -+ M be a minimal C-geodesic between p and q. We have
d;(o)(p, q) :S e2Co1' (L(q, r) for any
r E (0, T).
+ 4n3Cor2)
Taking the limit lim1'--+o+ of the above inequality, we get
lim1'--+o+ L(q,r) ;::: (dg(o) (p,q))2. To see the other direction of the inequality, choosing 7.13(iii), we have
72
= r in
Lemma
- r) < 4nCo 2 2'" - 2 L(q - (p , q) , 3- r + e ,-,or dg(T) for any
r E (0, T).
Taking
r -+ 0+, we get
Jim L (q, r) :S d;(o) (p, q) . T--+O+
o
The lemma is proved. Next we estimate the minimum value of L (q, r) - 2nr over M.
LEMMA 7.48 (Monotonicity and estimate for minM L). Let (M n , 9 (7)) , 7 E [0, T], be a complete solution to the backward Ricci flow with bounded sectional curvature. (i) The function min (L (q, r) - 2nr) qEM
is a nonincreasing function of r. (ii) For all f E (0, T) , min L (q, f) :S 2nf.
qEM
6. EQUATIONS AND INEQUALITIES SATISFIED BY LAND f
PROOF.
325
(i) From (7.85) we have
:T (L - 2nT)
+ ~ (L -
2nT) :::; O.
If L (.,.) were C 2 , then the lemma would follow from the maximum principle. Since we only know that L (', .) is locally Lipschitz, we proceed with the maximum principle argument with some care. Define the function h : (0, T) -+ ~ by
h(T) ~ min (L (x, T) - 2nT) . xEM
CLAIM.
For any
> 0 there exists qT EM such that
T
h(T) = L (qn T) - 2nT, and h( T) is a continuous function. If M is closed, then the claim follows from L(x, T) being a continuous function when T > O. When M is not closed, the claim follows from Lemma 7.13(i), which says that limx-+oo L (x, T) = +00. This and the local Lipschitz property of L (x, T) imply the claim. Now we estimate the right lim sup derivative of hat f E (0, T):
d+h(_) ...!...l' T
dT
"7'
lmsup
h(f + s) - h(f) s
s-+O+
. L(qHs, f + s) - L(qr, f) = 1lmsup s-+O+
L(qr, f . :::; 1lmsup
2 n
-
S
+ s) - L(qr, f) s
s-+O+
-
2
n,
where the last inequality follows from the definition of qHs' If L is C 2 at (qr, f), then by using (7.84), we have . L(qr,f+s) -L(qr,f) _ oL( -) 11m sup - ~ qr, T s-+O+
S
uT
:::; 2n - ~L(qr, f) :::; 2n. The reason why ~L(qr, f) 2: 0 is that qr is a minimum and smooth point of L(·,f). We have proved that d;Th(f):::; 0 when L is C 2 at (qr,f). When L is not C 2 at (qr, f).!. let L be a smooth barrier function of L at (qr, f) as in (7.56). Setting L(x, T) ~ 2..jTLix, T), we se~ that qr is a minimum point of the locally defined function L(., f) since L is a barrier function of L from above at (qr, f) and qr is a minimum point of L(., f).
7. THE REDUCED DISTANCE
326
Hence t1t(q'f, f) ~ O. We have
d+h(_)
-d r r
~
l' L(q'f, l' + s) - L(q'f, f) Imsup 8 ..... 0+ s
. L(q'f' l' + s) - L(q'f' f) _ < _ 1Imsup s
8 ..... 0+
aL (q'f,r_) .
~
ur
Then using (7.84), which holds for the barrier function
at
t,
we have
~
or (q'f, f) ~ 2n - t1L(q'f' f) ~ 2n.
We have proved that d;Th (f) ~ 0 when L is not C 2 at (q'f, f). Hence we have proved that d;Th (f) ~ 0 for all l' E (0, T). By the monotonicity principle for Lipschitz functions stated in §3 (Lemma 3.1) of [179], h(r) is nonincreasing. (ii) This follows from (i) and Jim h(f)
T ..... O+
= Tlim min L (q, f) ..... OqEM ~ !im L (p, f) = (dg(o) (p,p))2 T ..... O
=0. D
6.3. The reduced distance function £. To get even better equations than those in Lemma 7.45, we introduce the reduced distance function £. DEFINITION
(7.87)
7.49. The reduced distance £ is defined by
£(x,r) ~
1
1 -
2JT L (x,r) = 4rL(x,r).
Let l' E (0, T) and let "Y : [0,1'] ---+ M be a minimal £-geodesic from p to q and let K = K b, f) be defined as in (7.75). By (7.79), (7.80), and (7.81), we have at (q, f) , (7.88)
a£ =_1_K_~ R of 21'3/2 f + ,
(7.89)
1 £ IV'£I 2 = -R- K +1'3/2 1"
(7.90)
1 n t1£ < - - - K + - -R. -
21'3/2
21'
From these equations (which involve the trace Harnack integral K), (7.86) and Lemma 7.48, we easily deduce the following which do not involve K.
6.
EQUATIONS AND INEQUALITIES SATISFIED
BY L
AND
I.
327
LEMMA 7.50 (Reduced distance - partial differential inequalities). At (q, '1') the reduced distance e(x, 7) satisfies
ae _ ~e + l\lel 2 - R + ~ >0 a7 2T - , 2 e-n 2~e -I\lel + R + -_- :s; 0, 7
(7.91) (7.92)
ae
(7.94)
lim 1'-->0+
(7.95)
e
n
<0 a7 '1' 2'1' - , a£ 2 e 2a7+I\I£1 -R+:f=O, -+~e+---
(7.93)
e(q,'1') =1 [dg(o) (p, q)] 2/ 4'1' '
min £ (q,'1')
qEM
= 41_ minL(q,'1'):S;~. 2
7 qEM
REMARK 7.51. (i) Note that (7.94) is the only possible equality obtainable from (7.88) and (7.89) which does not involve K. (ii) In the inequalities above, the direction of the inequality depends on the sign of the coefficient in front of the term ~e. The reason for this is that, analogous to considering ~ (d 2 ) on a Riemannian manifold with nonnegative Ricci curvature, ~£ has an upper bound (7.90). The above equations for e demonstrate, in the context of Ricci flow, the superiority of the reduced distance over the Riemannian distance function. It is a space-time notion of distance which is a subsolution of Laplace-type and heat-type equations. Note that (7.91) is a forward heat superequation whereas (7.93) is a backward heat subequation. One would think that the backward heat subequation is more natural since it is associated to the backward Ricci flow, but we shall find the forward superequation (7.91) very useful. REMARK 7.52. (i) If we replace the inequality by an equality in (7.91), we obtain equation (6.14) for 1 used in the study of the entropy W:
!1 - 61+ 1\1112 - R+ ~
=0.
(ii) Compare (7.92) with the equation (6.52): 26
for the minimizer
1 - 1\1112 + R + 1 7
n =
~ J.L (9, 7) 7
1 of W (9, " 7) .
If we divide by 2y'f in the LHS of (7.65), we get (\I\le + Rc) (Y, Y) at smooth points of e. Motivated by the special case of shrinking solitons, we
328
7.
THE REDUCED DISTANCE
may write formula (7.65) at smooth points of f as (7.96)
(~7\7f + Rc - 2~g) (Y, Y)(1')
r vr Y)
r Vft: I\7X Y
~Y12 dT.
1,G H (X, dT + + Rc (Y) _ 2y T Jo Jo f 2T EXERCISE 7.53. What does the trace of this equality say?
=-
We have the following property of the reduced distance f. LEMMA 7.54 (Reduced distance as T -t 0). Given V E TpM, let [0, T] -t M be the C-geodesic with lim7'-+o+ JT~ (T) = V. Then lim f 7'-+0+
(7.97)
'v :
hv (f) ,f) = 1V1~(0) .
PROOF. Since lim7'-+o JT~ (T) = V and 'V I[O,Tj is a minimal C-geodesic when l' is small (see Lemma 7.29), we have lim fhv (T) ,T) 7'-+0+
r v!f (R(IV (7) ,7) + Idd'VT (7)1 7'-+0+ 2y T Jo = lim 21;;; r v!f. ~T 1V1~(0) d7 = 1V1~(0) . yTJo = lim
1;;;
2 )
d7
g(')'(1'),1')
7'-+0+
o We conclude this subsection with a few exercises concerning the reduced distance f. EXERCISE 7.55 (C-triangle inequality). Show that for any path I
h,T2]-t M, L (, (T2) , T2) - L (, (T1) , T1) = 2VTi.f h (T2) , T2) - 2JTif h (T1) , TI)
= Ch)
-1:2 vrl\7f- ~~12
dT.
REMARK 7.56. Given points (iiI, 1'1) and (i12, 1'2), we may define
Cd ((q1, 1'd, (q2, 1'2))
= LUil,Tl) (q2, 1'2) ~
inf {C h) : I (1'i) =
qi,
i = 1, 2} .
Then the above formula implies
+ Cd ((q1, T1), (q2, T2))
2: Cd ((p, 0), (q2, T2)) . Equality holds if and only if for some minimal C-geodesic I : h, T2] -t M Cd ((p, 0), (q1, T1))
we have ~ (p, 0).
= \7 f along I, where f is defined with respect to the basepoint
6.
EQUATIONS AND INEQUALITIES SATISFIED
BY LAND
e
329
The next exercise characterizes when an integral curve of f is an £geodesic. EXERCISE 7.57 (Integral curves of V'f). Show that for any solution
(Mn,g (7)),7 E [0, T), to the backward Ricci flow, a smooth integral curve 'Y of V' f is an £-geodesic if and only if gradient vector field.
:7 (V' = °along f)
'Y, where V' f is the
SOLUTION TO EXERCISE 7.57. In view of the £-geodesic equation (7.32), we compute 1 1 V'veV'f- "2V'R+2Rc(V'f) + 27 V'f =
~V' (lV'fI2 -
R
+~) +2Rc (V'f).
Applying the identity (7.94) for f, we obtain V'veV'f -
1
1
"2 V'R + 2 Rc (V'f) + 27 V'f
= -V' (;;)
+ 2 Rc (V'f)
a
= - a7 (V'f) , where V'f is the gradient of f (considered as a vector field), Le., (V'f)i = gijV' jf. We also note the following consequence of (7.91) which we shall revisit later when considering the reduced volume monotonicity. EXERCISE 7.58 (Pointwise monotonicity of reduced volume integrand). Show that for a solution g = 2 Rc of the backward Ricci flow,
:7
(:7
=
+ £ve)
(-~ 27
((47r7)-n/2 e-edJ.l)
af + R -1V'fI 2 + ~f) (47r7)-n/2 e-edJ.l
a7
::; 0, where £ denotes the Lie derivative. In what sense is the inequality true? Note that f is only a Lipschitz function. See Section 9 of this chapter for some hints.
330
7.
THE REDUCED DISTANCE
6.4. The growth of the reduced distance function £. The growth of £, in particular the lower bound of £, will be used to justify some technical issues later (for example, the proof of Lemma 7.130). The next lemma follows from Lemma 7.13(i) and (iii). LEMMA 7.59 (Bounds for the reduced distance). Let (Mn,g(r)) , r E [0, T], be a complete solution to the backward Ricci flow with bounded sectional curvature and let p be a basepoint. Then for any (q, T) EM x (0, T) ,
1 2 nCOT _ e 2CoT 2 4Te2COTdg(0) (p, q) - -3- :::; £ (q, r):::; 4T dg(r) (p, q)
nCoT
+ -3-·
Ig; I
Next we bound 1V'£12 and by £, and hence we can bound them by d;(r) (p, q). These estimates will be improved in Lemmas 7.64 and 7.65 when we assume the curvature operator is nonnegative. LEMMA 7.60 (Bounds for first derivatives of f). Suppose (Mn,g(r)), r E [0, T), is a complete solution to the backward Ricci flow with bounded sectional curvature. Then for any T E (0, T) there exist positive constants A 2 1 and C 1 depending only on T, n, T and Co which satisfy the following properties. For any q E M and r E (0, T], we have £ (q, r) + Ar 2 0,
(i)
(ii)
I;: I (q, r) :::; ~l (£ (q, r) + Ar) . PROOF. (i) Let 1(7') be a minimal L:-geodesic from (p,O) to (q, r) and let X(7') ~ ~. By Lemma 7.13(ii) we have
+ £ (q,f) E (0, f). Here and below C\ and A are constants depending only IJ:r;;X(r*)1 2
:::;
nC\ f
for some r* on f, n, T and Co and its value may change from line to line. Thus using Shi's derivative estimate and integrating equation (7.45) (0- = 2Vf),
d~ IJfX(7')I:(iT /4) = -o-Rc (JfX(7') , JfX(7')) + :2 (V'R, JfX(7')) 2
on [r, r*], as in the proof of Lemma 7.28, we get for any r E [0, f] Iy'TX(r)1 2
It follows from V'£(q,r) (7.98)
In particular £ (q, r)
:::;
(nC\f+f(q,f)
= X(r)
IV'£I 2 (q, r)
+ Ar 2
°
+C\f) i;l.
that
C\ :::; -:;:-
(£(q, r) + Ar
for all (q, r).
A
)
.
6. EQUATIONS AND INEQUALITIES SATISFIED BY LAND £
331
(ii) From (7.94) we have at (q,7), 2. af I -< ~ + ill + !IV'fI Ia7 2 27 2
Using (7.98), we get at (q,7) that
af I ~ Co + ill + (\ (f + A7) Ia7 27 7
If + A712 + A7 + -G\ ( + A7~) ~ (61+ !) (f + (A + C~ + !:) 7) ,
~ Co
+
7
7
where we have used f (q, 7)
+ A7
Thus (ii) follows from taking C1
f
7
°
~
C1 + '2
for all (q,7) in the last inequality.
= 61 +!
and A
= A+ GC?+!;4. 1+2'
The lemma 0
is proved. From Lemma 7.41 we get LEMMA 7.61 (Hessian of reduced distance). Suppose (Mn, g(7)),
7 E
[0, T), is a complete solution to the backward Ricci flow with bounded sectional curvature. Fix To E (0, T). Given l' E (0, To] and q E M, the Hessian of the reduced distance function f (', f) at q has the upper bound
(7.99)
C2
f ~ . '= '2y7
Hess(q 'f)
+
1 + C2 d ;('f) (p, q) 27
,
where C 2 is a constant depending only on n, To, T, SUPMx[o,T]IRml. 7.62. Let
6.5. Estimates for f when Rm ~ 0. As we shall show below, when (Mn, g (7)),7 E [0, T), has bounded nonnegative curvature operator, there are better estimates for Hess(q,'f) f, IV'fI 2 , and 1g~ I. Let l' E (0, T) and q E M. Let 'Y : [0, 1'] ~ M be a minimal .c-geodesic from p to q and let X (7) ~ ~. Let Y be a solution to (7.66). Hamilton's matrix Harnack estimate implies H (X, Y) (7)
~ - (~ + T ~ 7 )
Rc (Y, Y) (7)
(! + _1_) R IY (7)12 T-7 ~ -~ (1 + _7_) IY (1')1 2R T-7
~
-n
7
7
7. THE REDUCED DISTANCE
332
since
IY (7)12 = ~ IY (1')12.
Hence by (7.70),
(Hess(q,T) L) (Y (1'), Y (1'))
~ ~ IY (1')1 2 foT y'T (1 + T ~ 7) Rd7 + IY;1 2 :s =
*(1+
T~
r) IY(r)I' 1." ,j7 (R+ I~;I') d7+ IV;1'
[(~ + ~n~) £ (q,r) + ~] IY(1')1 2 .
We have proved (compare with Lemma 7.41) 7.63. Let (Mn,g(7)), 7 E [O,T), be a solution to the backward Ricci flow with bounded nonnegative curvature operator. We have for any l' E (0, T), LEMMA
1 Hess(q,T) £ ~ n ( -;;.
+T
1 ) _ l' £ (q, 1')
1 + 2T'
Recall that Hamilton's trace Harnack estimate says that H (X) (7) Hence K
h, 1') K
~ - (~+ T ~ 7) R (1 (7),7).
defined in (7.75) satisfies
h, 1') ~
-
foT 73/ 2 (~ + T ~ 7) R h (7),7) d7
~ -1."
7 1/' T
~
7
(Rb(7) ,7) +
I~;I:(T}T
T
~ - - T_Lh(1'),1')· -7
Therefore, from (7.89), we have at (q,1') 2
1
£
T
IV£I ~ -R+ 1'3/2T_1' L h(1'),1')+-;;.
2T) £.
1 ( l+ _1' ~-R+-;;. T In particular,
LEMMA 7.64. Let (Mn,g (7)),7 E [0, T), be a solution to the backward Ricci flow with bounded nonnegative curvature operator. If l' E (0, (1 - c) T) for some c E (0,1), then for any q E M
6. EQUATIONS AND INEQUALITIES SATISFIED BY LAND
where C
= 1 + ~.
If T
= 00,
e
333
then for any (q, f) EM x (0, (0)
IV£1 2 (q, f)
+ R (q, f)
::;
~£ (q, f) . 7
Hence, for an ancient solution with bounded Rm tance bounds Rm. Now we can estimate 1 1.
~
0, the reduced dis-
g;
LEMMA 7.65. Let (Mn, 9 (7)) , 7 E [0, TJ, be a solution to the backward Ricci flow with bounded nonnegative curvature operator. If f E (0, (1 - c) T) for some c E (0,1), then at any (q, f) EM x (0, (0),
8£1 < C+ 1 ~ 187 2 f' where C = 1 +~. 1fT = 00, then at any (q,f) EM x (0,00),
I~ 87 IOg£1 -< ~ f'
( 7.100)
and for any
°<
71 < 72 and q E M, (71)2 ::; f(q,72)::; (72)2 72 £ (q, 71) 71
(7.101) PROOF.
From (7.94) we have at (q, f) that
8£ + ~ IV£1 2 = ~R- ~£. 87 2 2 2f Hence by Lemma 7.64 and £ (x, 7) > 0, we have
I;~ I ::;
~ (IV£1 2 + R) + 2~£
If T = 00, we can choose c arbitrarily close to 1 and get 1:7" log £1 ::; ~. Since log £( x, 7) is a locally Lipschitz function of 7 > 0, we have
(q, 72) I ::; 17"2 ~d7 = log (72) 2 , IIOg £f(q,7I) 7 71 T1
and (7.101) follows.
0
6.6. Reduced distance under Cheeger-Gromov convergence. Finally we discuss the convergence of the reduced distance under CheegerGromov convergence. Let {(M k,9k (7) ,Pk)}kEN and (M~, goo (7) ,Poo) , 7 E [0, T], be complete pointed solutions to the backward Ricci flow satisfying the curvature bound
max {IRmgk I, IRc gk I} ::; Co < 00
on Mk x [0, TJ
for all kEN. Suppose that (Mk,gd7)'Pk) ---+ (Moo, 900 (7),Poo) on the time interval [0, T] in the Coo Cheeger-Gromov sense; that is, there exist
334
THE REDUCED DISTANCE
7.
an exhaustion {UdkEN of Moo by open sets with Poo E Uk and diffeomorphisms ~k : Uk ~ Vk ~ ~k (Uk) C Mk with ~k (Poo) = Pk such that
(Uk, ~k (9d [O,T].
7)lvk))
~ (Moo, 900 (7))
in Coo on compact sets in Moo x
LEMMA 7.66 (£ under Cheeger-Gromov convergence). Under the setup above, for any (q, f) E Moo x (0, T), we have £9(00 0)
Poo,
(q, f) = k--+oo lim £9(k 0) (~k (q), f). Pk,
The convergence is uniform on compact subsets of Moo x (0, T). Furthermore, the convergence is uniform in Coo on compact subsets of the open set of points at which £(;00,0) is Coo.
'1 :
PROOF. Let [0, 1'] ~ Moo be a minimal .e-geodesic, with respect to goo, joining Poo to q. By the convergence of the sequence of solutions {gk (7)} in C 2 on compact sets, we have £9(00 0)
poo,
(q, f)
('1) =
=
1;:;.e9oo 2y 7
2::
lim sup £9(pk
k--+oo
1;:; lim .e 2y 7 k--+oo 9k
0) (~dq),
(~k 0
'1)
f).
k,
Next we prove the opposite inequality. Since for any compact set K
c
Moo, (7.102) we know for every q E Moo that d 9k (r)
(Pk, ~k (q)) =
dcJ?j;(9klvk)(r)
(Poo, q) ~
d 900 (r)
(Poo, q)
as k ~ 00. For sufficiently large k, let 'Yk : [0, 1'] ~ Mk be the minimizing .e-geodesic with respect to 9k from Pk to ~k (q) . By Lemma 7.13(iii) and (i), we have .e("(k) = L(~k'O) (~k (q) ,f) :S C2, and for any
7
E [0,1'] ,
d9k (0) (Pk, 'Yk (7)) :S C2 for k large enough, where C2 is a constant independent of k and 7 (depending on n, 1', T, Co, d 900 (r) (Poo, q)). Hence ~k1 ("(k ([0,1'])) stays inside some compact set /(1 C Moo independent of k. For k sufficiently large, we have L9(k 0) (~k (q), f) = .e ..... ( I ) (~k1 ("(k)) . By (7.102), we get ""'k 9k Vk
Pk,
liminf £9(k
k--+oo
Pk,
0)
(~dq) ,f)
1 ;:; liminf .e..... ( I ) (~k1 0 'Yk) 2y 7 k--+oo ""'k 9k Vk 1 = ;:;liminf.e (~k10'Yk) 2::£9(00 0) (q,1'). 2y 7 k--+oo 900 poo,
=
7. THE i-FUNCTION ON EINSTEIN SOLUTIONS AND RICCI SOLITONS
The second equality above needs uniform bound .jT
I~ I:k (7) :S
335
C2 , which
follows from Lemma 7.13(ii) and an argument similar to that in the proof of Lemma 7.28. The last inequality above follows from the fact that ~kl 0 'Yk is a curve from Poo to q. We have shown the desired opposite inequality and the lemma is proved (we leave it as an exercise to prove the uniform convergence). 0 7. The f-function on Einstein solutions and Ricci solitons 7.1. f function on an Einstein solution with positive scalar curvature. We consider an Einstein solution (Mn, 9 (7)) of the backward Ricci flow with positive scalar curvature defined on a time interval containing O. We first consider the case that the solution is smooth at 7 = 0 and then generalize to the case that the solution becomes singular as 7 '\. O. From the scalar curvature evolution equation ~~ = - ~ R2, we have
R (7) _ _----::-1-----,,- R(o)-1 + ~
(7.103)
1+
2:1R (0) R(O)
and
g(7)=
(1+~R(0))9(0),
Given a curve 'Y : [0, fJ
-+
7E
(-2Rn(0)'OO).
M from p to q, we have
[y:r (R(')(T),T) + I:;[J dT = r VT ( 27 + (1 + Jo (0) + n n
L:(,)=
Let a
27 R(O)) 1dd'Y 12
:1
R
7
) d7.
g(O)
= 2.jT, so that .jT n 12..ft a2 =da 1oT----'-::----=-d7 R (0) + 2 2n . R (0) + a -1
2:
0
-1
2
= n..;f (1 _ tan- 1 ( J27R (0) In)) J2fR (0)
(7)2 d7 under the x (7) d7 = d is given by x (7) = >(7) J; >(:) 1d7' Similarly the
Recall that the minimum of the functional
J;
constraint minimum of (7.104)
J; ¢ (7)
In
.
E b)
~
1 T
o
Id
¢ (7) d'Y 12 7
g(O)
d7
X
7. THE REDUCED DISTANCE
336
is
d~(O) (p, q)
J; (T)-l dT and the minimizer is given by a minimal geodesic '"'I with speed
d'"'ll 1 dg(o) (p, q) IdT g(O) = ( T) J; (T) dT' -1
Caveat: Here we have assumed that the improper integral convergent. Hence inf
r ..,fT (1 + n
J;
(T)-l dT is
I
2T R (0)) dd'"'ll2 dT T g(O)
Jo
d~(O) (p, q)
- J; T-
1/ 2
(1 + ~ R (0)r 1 dT' = 2.jT to get
Again we make the rationalizing substitution u
loo
T
T
(2
-1/2 1 + TR 0 n ()
)-1 d
2 n 10 2 R(O)
.,ff
T---
-
0
1
dU 2nR(0)-1 +u 2
= vIn-: 2nR (0) -1/2 tan -1 (
J7L
2v'T
v2nR (0)-
) 1/2'
We conclude LEMMA 7.67. Let (Mn, g (T)) be an Einstein solution of the backward Ricci flow on a time interval containing 0 with R (0) > O. Let p E M be the basepoint. For all f E (0,00) , we have
_
tan-l (J2TR(0) In))
_(
L(q,T)=nJi- 1-
R (0)1/2 d~(O) (p, q)
+
ffn tan In particular, (7.105)
f.
f
(q,)
J2TR(0) In
=?!: 2
(1-
1 ( J2TR (0) In) .
tan-1 (J2TR(0) In))
J2TR(0) In
J2TR(0) In
+ tan- 1 ( J2TR (0) In)
d~(O) (p,q) 4f
.
7. THE (I-FUNCTION ON EINSTEIN SOLUTIONS AND RICCI SOLITONS
337
Since by (7.103),
= R(O) -1
2fR(O) n
R(f)
,
we may rewrite (7.105) as
f(q f) = ~ '2
(7.106)
(1 _
+ d~(O) (p, q)
tan- 1 (8 (f))) 8
(T)
4T
8 (f) tan- 1 (8 (f))'
where 8
(f)
R(O) _ 1 R (f) .
~
= 00.
Now we consider the extreme case: R (0)
7.68. Let (M n , 9 (T)) ,T > 0, be an Einstein solution of the backward Ricci flow. Suppose that limr-->o R (T) = 00 so that R (T) = ;~, Rc (T) = r 9 (T) and 9 (T) = T9 (1). Although the metric 9 (0) is not defined, we may still consider (p, 0) , p E M, as the basepoint for defining £', Land f as before. We have for 'Y : [0,1'] ---t M from p to q, EXERCISE
1
-
£, ("()
= nVf +
2
r T3/21 ~'Y 1 Jo
dT.
T g(l)
Show by considering the paths
i(T) = {
~(T/r7)
T
:S 'fl,
T
> 'fl,
where (3 : [0,1] ---t M is a constant speed geodesic with respect to 9 (1) joining p to q, and letting 'fl ---t 0, that (7.107)
inf 'Y
1
1' T 3 / 2
Id-.:J.. 12
where the infimum is taken over 'Y : [0,1'] (7.108)
dT
= 0,
dT g(l)
0
L (q, f)
---t
M joining p to q. That is,
= nVf.
Hence for an Einstein solution 9 (T) of the backward Ricci flow with lim R (T)
r-->O
= 00,
we have f(q,1')
n
= "2'
Using the rule for the f function on product spaces (see Exercise 7.11), if we have a product solution (M x N,9 (T) + h (T)), where (Mn,9 (T)) is an Einstein solution of the backward Ricci flow with limr-->o Rg (T) = 00 and where (N"t, h (T)) is Ricci flat, then g+h ( f(Pl,P2,0) q1,
_) _
q2, T
n
-"2 +
dh
(p2, q2)2 41'
338
THE REDUCED DISTANCE
7.
For example, we may take (M 2 ,g (7)) to be an evolving 2-sphere and (Nl, h (7)) to be the line to get the cylinder 8 2 x R EXERCISE 7.69. Assuming the solution (Mn,g (7)) is defined only for
7> 0, rewrite equation (7.106) using the metric 9 (1) instead of 9 (0) . Show that this equivalent form is consistent with (7.108). We follow up on the above exercise by translating time in our Einstein solution so that R (0) --t 00 in (7.103), and correspondingly, R (7) --t 2~' Since we then have d;(O) (p, q) --t 0, we choose to rewrite the formula for £ in terms of d;(T) (p, q) using
d;(O) (p,q) R(r) 1 d;(T) (p, q) = R (0) = 1 + ~ R (0)' In particular, from (7.105), we have £
r (q,)
(1-
=~ 2
tan-l (V27R(0)
V27R(0)
In
In))
R(r)d;(T) (p,q)
1
+----:--J~:::::::::::::;:::::;::::~-::--;:::::::=-r===:::::;::::=;=
In) 2ffnV7R (0)
tan- 1 (V27R (0)
n 2
--t -
as R (0)
--t
00.
EXERCISE 7.70. Let (Mn,g(t)), t E [O,T), be a maximal shrinking Einstein solution of Ricci flow so that R (t) = 2(T~t)' Given any (Xi, ti) E M x (0, T), we define a solution gi (7) ~ 9 (ti - 7) of the backward Ricci flow and i(x."" t.) (x, t) = £9(;x,.,. 0) (x, ti - t) . Check that
+ 4 (T -
t)
From the above remarks, if~i,td (x, t)
J
--t
t;-t T-ti
tan- 1
~ as i
--t
J
00
t;-t . T-ti
if ti
--t
T.
7.2. The £ function on a steady gradient Ricci soliton. Consider a steady gradient Ricci soliton g(7) =
a;;
(x) = - (grad 9o fo) (
= idM·
7. THE i-FUNCTION ON EINSTEIN SOLUTIONS AND RICCI SOLITONS
339
Given a path I' (r) , its C-Iength is
C (')') = =
r-
Jo
Vi (R (')' (r) , r)
l v'T (
R (I'r
+ 1~I' 12 r
) dr
g(r)
(-y(T)) ,0) + 1(l'r)o :~[O}T.
Let (3 (r) ~ CPr (')' (r)). Note that since g(r) = cp;gO, geometrically, a point (x, r) is the same as the point (CPr (x), 0). That is, (')' (r), r) is the same as ((3 (r), 0) . We have
.
(3 (r)
. d(3 dr
'7'
=
dl' (CPr) * dr
+
aCPr or ((3 (r)) ,
so that Hence
C (I') = =
lof Vi (R (r), 0) + 1/3 (r) + (gradgofo) (r))I:(o)) dr lof Vi(R((3(r),O)+I(gradgofo)((3(r))I~(o))dr ((3
r-
+ Jo
Vi
((3
(1/3(r)1 g(O) +2//3(r),(grad go fo) ((3 (r))) ) dr. \ g(O) 2
Now from (1.34) we have
R ((3 (r) ,0) + I(grad gofo)((3 (r) )I~(o) = {)
(7.110)
independent of (3 (r) , and
(/3(r),(grad go fo) ((3 (r)))g(O) =
d~ Uo((3(r))).
Hence, letting (j = 2y'T and a ((j) ~ (3 ((j2 /4) and integrating by parts, we have an alternate formula for the C-Iength on a steady gradient Ricci soliton. LEMMA
7.71. On a steady gradient Ricci soliton Rc (go)
we have
r. (-y) =
~Cf3/2+2v'¥10 (a (2v'¥)) +J.'v'¥
(I::
(u{O) -
+ \l\l fo =
0,
10 (a (U))) du,
where {) is given in (7.110), a((j) = CPu2/4 (I' (u;)) and CPr is defined by
(7.109).
Later we shall see that the shrinking case gives us a more explicit formula. Now we compute the geodesic equation by taking a variation Y = 8a of a which vanishes at the endpoints. We have
7. THE REDUCED DISTANCE
340
COROLLARY 7.72. The variation of the C-length on a steady gradient Ricci soliton Rc (go) + '\1'\1 fo = 0 is given by
where '\1 is the Levi-Civita connection for g (0) . Hence the C-geodesic equation is
da
'\1 do -d du
0"
1
+ -2'\1fo = o.
This also implies that the following directional derivative vanishes:
da dO"
2 (l da dO" lg(O) +
fi) 0
-
0.
That is, along an C-geodesic, the square of its speed with respect to the potential function is constant. EXERCISE
0"
plus
7.73. Show that the above equation is equivalent to (7.35): 0"2
o= D z Z -
8'\1 R + 0" Rc (Z) ,
!
where Z ~ JT~. Note that on a gradient soliton '\1 R
= Rc ('\1 f) .
In the rest of this subsection we consider C-geodesics on the cigar on ~2. The scalar curvature of the cigar solution to the backward Ricci flow
!!.9.. aT =2Rc '
is given by 4
R(X,y,7)
= 1 +e 4(2 x +y2)" T
Let r2 = x 2 + y2. The C-Iength of a radial path 'Y (7), 0 :S 7 :S f, with r (7) ~ r h (7)), is C h)
('" (4
1
= Jo ..jT 1 + e4Tr2 + e-4T + r2
(dr)2) d7 d7.
Define s (7) ~ sinh -1 (e 2T r (7)) (which is the distance to the origin with respect to 9 (7)) so that
£b) = [
(48eCh 2 S+
(~;
-
2 tanh
s)') v'Tdr.
7. THE i-FUNCTION ON EINSTEIN SOLUTIONS AND RICCI SOLITONS
341
Setting 0" ~ 2.Ji, we have
( 0"2 sech 28 + (d8 £ (-y) = J{2Vf 0 dO" - 0" tanh 8
)2) dO"
( d 8dO" ) 2 -20"tanh8 d8) = Jo{2Vf (0"2+ dO" dO". Those £-geodesics (i.e., the critical points of £) that emanate from the origin are given by radial paths 8 = 8 (0") which satisfy d2 8 d0"2 - tanh 8 = 0,
(0) = O.
8
Indeed, if 88 = v, then 8£ (-y) =
d8 dv Jo{2Vf ( 2 dO" dO" -
= fo
2Vf
( -2v
( d8 20" vtanhf 8 dO"
dV)) + tanh 8 dO"
dO"
::~ + 2v tanh 8) dO".
Multiplying the £-geodesic equation above by ~! and integrating, we get ( :;) 2 _
2 log cosh 8 =
8f
(0) 2 ,
or equivalently,
:; =
va + 2
2 log cosh 8,
where a = 8 f (0). Note that ~! ~ a and for 8 large, ~! ~ j2S. We leave it as an exercise to check that 8 ~ c;2 = 27. In comparison, the radial £-geodesics on a cylinder of any dimension sn-l X lR satisfy 8 = 2a.Ji (see (7.23)). REMARK 8
= 0,
7.74. The solutions of the linearized £-geodesic equation at
Le.,
d2 8 d0"2 8
8
= 0,
(0) = 0,
are 8 (0") = a sinh 0". 7.75. Determine the qualitative properties of the reduced distance £ on the Bryant soliton. PROBLEM
7. THE REDUCED DISTANCE
342
(Nn,
7.3. f function on a gradient shrinker. Let h (t)) , -00 < t < 1, be a shrinking gradient Ricci soliton in canonical form as given in Proposition 1.7 with c = -1. Define l' ~ 1 - t, and let h(r)~h(1-r)
(7.111)
and
f(r)~j(1-r),
0<1'<00.
Then h (1') is a solution of the backward Ricci flow on the maximal time interval (0,00). Since h(t) = (1 - t) 0
h (1') = r
f (1') = f (1) 0
=
~Rh(l) (
:t
:1'
-~ (grad h(l)f (1))
0
7.76. Note that if h (1') is a shrinking gradient soliton flowing = -1\lfI2, the along \If, then Rc+\I\lf - 2~h = O. Since f satisfies gradient vector field \If satisfies REMARK
U
:1' (\If) = -2Rc(\lf) Given a path, : [0, f] .c (r)
=
+ \I (~~) = -2Rc(\lf) ---7
N from
\l1\lfI2 =
-~\lf.
p to q, its .c-length is
loT JT (Rh(T) (r (1')) + 11' (r)I~(T») dr
= loT JT
(~Rh(l) (
l'
I(
Let f3 (1') ~
. f3 (1') which implies
8
= (
THE i-FUNCTION ON EINSTEIN SOLUTIONS AND RICCI SOLITONS
7.
343
Hence
2 .fi(~Rh(l) ({3 (7)) + 71/3 (7) + ~ (gradh(l)i (1)) ({3 (7))1 ) d7 Jo h(l)
£ (f) = (T
7
=
7
foT 7- 1/ 2 (Rh(l) ({3 (7)) + I(gradh(l)f (1)) ({3 (7))1~(1)) d7
r
r
2
d
Jo 7 / 1/3 (7)1 h(l) d7 + Jo 2.fi d7 (f ({3 (7),1)) d7 = foT 7- 1/ 2(Rh(l) ({3 (7)) + I(gradh(l)i) ({3 (7), 1)1~(1) -
+
+
3
/ J(T7 o 3
2
f ({3 (7), 1))d7
2 1/3(7)12
d7+2v'Tf({3(f),1). h(l) On a shrinking gradient Ricci soliton we have (7.113)
+ I(gradh(l)i) ({3 (7), 1)1~(1) -
Rh(l) ({3 (7))
f ({3 (7),1)
= 6.
Hence
£ (f) =
2FT (i ({3 (f), 1) + 6) + foT 73/ 2 1/3 (7)1:(1) d7
and 1;:;£ (f) 2y7
= f ({3 (f), 1) + 6 + 1;:; {T 73/ 2 1/3 (7)1 2 d7. 2y7
Jo
h(1)
We have f (~;l ({3 (f)), f)
f ({3 (f), 1) Note that from (7.107),
=
(7.114)
lo0 T 73/ 2 {3 (7)1 2h(l) d7 = 0,
inf (3
= f (f (f), f).
1.
where the infimum is taken over all {3 : [0, f] -+ N joining p to {3 (f) = ~T (q) implies '"Y (f) = q, we conclude
~T
(q) . Since
LEMMA 7.77 (Reduced distance on shrinker). For a shrinking gradient Ricci soliton as in Proposition 1.7 with c = -1,
f (q, f) =
(7.115)
where f is defined in (7.111) and f (q, f)
f (q, f) + 6,
6
is from (7.113). That is,
= f ({3 (f), 1) = f (~T (q), 1) + 6,
where {3 (7) ~ ~T (f (7)) and
CPT
is defined by (7.112).
REMARK 7.78. Note that for fiat Euclidean space, thought of as a shrinking (Gaussian) soliton, the potential f.
f
=
I~~ given in
(1.16) is the same as
EXERCISE 7.79. Show that for a shrinking gradient Ricci soliton, the paths '"Y (7) = ~l-T (x) =
344
7. THE REDUCED DISTANCE
Following up on details related to the previous exercise, we have EXERCISE
7.80. In regards to (7.114), show that although for CPT (q)
a smooth minimizer of inf,6
J; 73/ 2 1/J (7) 12
h(1)
i= P
d7, where the infimum is taken
over paths joining p to CPT (q) , does not exist (which is related to h (0) not being well-defined), any minimizing sequence (3i of paths joining p to CPT (q) limits to the constant path 7 1--+ CPT (q) . That is, any minimizing sequence 'Yi of paths joining p to q for 2~£ ("() limits to the path 7 1--+ cp.;:-1 (CPT (q)). Note that in general, qo ~ limT->o cp.;:-1 (CPT (q)) i= p. We may think of this as saying that the minimal geodesic starting at (p,O) immediately jumps to (QQ,O) and then becomes a constant path in the geometric sense. Caveat: The solution is undefined at the 'big bang' time 7 = O. We now present another proof of Lemma 7.77, following an original idea of one of the authors [289]. (The proof given above is also inspired by his line of reasoning.) Given a path 'Y : [0,7'] -+ N, we have
d~ (v'Tf("((7) ,7)) =v'T(~ + ~~ +Vf'i) =
(7.116)
v'T
(L + of07 + ~ IVfl 2+ ~ lil 2- ~ b - Vf12) . 27
2
2
2
Now for a gradient shrinker in canonical form, where (1.14) holds, i.e.,
~~ = -IVfI 2,
(7.117)
assuming f has a critical point,12 we may normalize propriate constant) so that
f
(by adding the ap-
R + IV fl2 - ~ f == O.
(7.118)
7
Substituting this with (7.117) into (7.116), we have
d~ (v'Tf
("( (7),7))
=
~v'T (R + lil2 - Ii - Vf12) .
Hence
f(,,((7'),7')= 1££(,,()_ 1£ 2y7
Jor v'Tli(7)-Vf("((7),7)1~(T)d7.
2y7 Given a point q E N, we may take 'Y : (0, f'] -+ N to be the path with i (7) = V f (7) for all 7 E (0,7'] and 'Y (7') = q. We have 1 f ("( (7'),7') = 2.jf£ ("() 2:: f ("( (7') ,7') . Since 7 = 0 is the big bang time, f ("( (7') ,7') is independent of the basepoint chosen. 12If N is compact, this assumption is always satisfied (though it is also satisfied for the shrinking Gaussian soliton).
8.
.c-JACOBI FIELDS AND THE .c-EXPONENTIAL MAP
On the other hand, taking 'Y to be a minimal C-geodesic with 'Y (-1') we have
f ("( (f) ,f) = f
("( (f), f) -
1,s ('" JT Ii' (T) - V f
2yT
Jo
345
= q,13
("( (T), T)I~(T) dT
::; £ ("( (f) ,f) . Since 'Y (f) = q is arbitrary, we conclude that defined by (7.118). EXERCISE 7.81. Let
f =
f on N x (0,00) for
f
(Nn, h (t)) , -00 < t < 1, be a shrinking gradient
Ricci soliton in canonical form and consider (N, h (T)) , T E (0,00), where h (T) ~ h (1 - T). Show that for any p, q E Nand l' > 0, if we take any sequence Ti ---t and minimal C-geodesics 'Yi : h,1'] ---t N with 'Yi h) = p and 'Yi (f) = q, then a subsequence 'Yi converges to a minimal C-geodesic 'Y : (0,1'] ---t N with 'Y (f) = q. In particular, independent of the choice of pEN, for any path (3 : (0,1'] ---t N with 'Y (f) = q, we have
°
C ((3) ~ C ('Y) .
8. C-Jacohi fields and the C-exponential map Continuing our mimicry of Riemannian comparison geometry, in this section we derive the C-Jacobi equation, which is a linear second-order ODE along an C-geodesic, and we derive an estimate for the norm of an C-Jacobi field. We also discuss the C-exponential map, its Jacobian, called the CJacobian, and briefly mention the C-index lemma. These results will be of crucial importance to our discussion of the reduced volume and its applications in the next chapter. Throughout this section (Mn, 9 (T)), T E [0, T], will denote a complete solution to the backward Ricci flow satisfying the curvature bound max {IRml ,IRcl} ::; Co < 00 on M x [0, T], and p E M is a basepoint. Before discussing Ricci flow, we first recall some basic Riemannian geometry that is relevant to the material in this section. A good reference for comparison Riemannian geometry, besides Cheeger and Ebin [72], is Milnor's book [265]; in our setting, the Riemannian path energy is analogous to
(Nt, g)
2,
the C-Iength. Let 'Y : [a, b] ---t be a unit speed geodesic, let i' ~ and let ds denote the arc length element. The index form is defined by (7.119)
I (V, W)
~
lb
((V-yV, V-yW) - (R (V,i') i', W))ds,
where V and W are vector fields along 'Y perpendicular to i' and vanishing at the endpoints. By the second variation of arc length formula, under these assumptions on V and W, we have t5~,w L ("() 13See the exercise below.
= I (V, W) ,
346
7. THE REDUCED DISTANCE
where L (-y) denotes the length of"f. Integrating by parts on (7.119), we may express this as I (V, W) =
-l
b
(\71'\71' V
+ R (V, 1h', W) ds.
Recall that a vector field J along 'Y is a Jacobi field if \71'\71'J + R(J,1)1
= o.
(If J vanishes at the endpoints of 'Y, then I (J, W) = 0 for all W.) Equivalently, a Jacobi field is the variation vector field of a 1-parameter family of geodesics. Given a unit speed geodesic 'Y : [a, b] ~ M, the set of Jacobi fields along'Y is isomorphic to T'Y(so)M x T'Y(so)M, for any So E [a, b]; each Jacobi field J is determined by the initial data J (so) = Jo and (\71'J) (so) = Jl for Jo, Jl E T'Y(so)M.
The Index Lemma says that if'Y : [a, b] ~ M is a unit speed geodesic without conjugate points, then among all vector fields along 'Y perpendicular to 1 with prescribed values at the endpoints, the unique such Jacobi field minimizes the index form; i.e., given A E T'Y(a)M and B E T'Y(b)M perpendicular to 1, the Jacobi field J with J (a) = A and J (b) = B satisfies (7.120)
I (J, J) ~ I (W, W)
for all W perpendicular to 1 and such that W (a) = A and W (b) = B. Equality in (7.120) holds if and only if W = 1. In particular, for any W =1= 0 perpendicular to 1 with W (a) = 0 and W (b) = 0, we have I (W, W) > O. If 'Y has no conjugate points in the interior, but possibly one at b, then (7.120) still holds although J may not be unique, Furthermore, we have I (W, W) ~ 0 for any W perpendicular to 1 with W (a) = 0 and W (b) = 0, where equality holds if and only if W is a Jacobi field.
8.1. £-Jacobi fields. Let 'Y : [0, f'] ~ M be an £-geodesic, where f' E (0, T), and let X (7) ~ ~ be its tangent vector field. DEFINITION 7.82 (£-Jacobi field). An £-Jacobi field along an £-geodesic 'Y is the variation vector field of a smooth 1-parameter family of £-geodesics 'Ys, s E (-E, E), for some E > 0, all defined on the same time interval as 'Yo = 'Y.
Let X (7, s) ~ ~, Y (7, s) ~ ~,and let Y (7) ~ Y (7, 0) be an £-Jacobi field along "f. Using the £-geodesic equation (7.32), we compute
\7x (\7xY) = \7x (\7yX) = R(X, Y)X
= R (X, Y) X + \7y
+ \7y (\7xX)
(~\7 R -
2 Rc (X) -
2~ X) .
8.
l:-JACOBI FIELDS AND THE .c-EXPONENTIAL MAP
347
Thus we have a linear second-order ODE for the C-Jacobi field Y (7), called the C-J acobi equation: (7.121)
\Ix (\lx Y )
1
= R(X, Y)X + 2\1y (\lR)
- 2 (\lyRc) (X)
1 - 2Rc (\lxY) - -\lxy' 27
Since 7 = 0 is a singular point because of the ~ factor in the last term, we rewrite the equation as D,fox (\I ,foxY)
=7
(\Ix (\lx Y ) +
2~ \lXY)
= R (JTX, Y) JTx
7
+ "2\1y (\lR)
- 2JT (\ly Rc) (JTX) - 2JTRc (\I ,foxY) .
Let Z(O") ~ y'TX(7), where 0" = 2y'T and (3 (0") = 'Y (0"2/4) . Then Z(O") ~~ and we can rewrite the C-Jacobi equation for Y (7) as \I z (\I z Y) = -20" Rc (\I z Y)
(7.122)
=
+ R (Z, Y) Z 0"2
- 20" (\ly Rc) (Z)
+2
\ly (\I R),
where we view Y (0"2/4) as a function of 0". Suppose Z(O) = limT-to y'TX = V E T'""((o)M. We have the following by solving the initial-value problem for (7.122). LEMMA 7.83. Given initial data Yo, Yl E T'""((o)M, there exists a unique solution Y (7) of (7.121) with Y (0) = Yo and (\I zY) (0) = Y1. Since (7.121) is linear, the space of C-Jacobi fields along an C-geodesic 'Y is a finite-dimensional vector space, isomorphic to T'""((o)M x T'""((o)M. REMARK 7.84. If the solution (M n ,9 (7) C-Jacobi equation (7.121) says
= 90) is Ricci fiat, then the
1 \Ix (\lx Y ) = R(X, Y)X - 27 \lxY.
That is, we obtain the Riemannian Jacobi equation for 90, D,fox (\I ,foxY)
= R (JTX, Y)
JTXj
i.e., \I z (\I z Y) = R (Z, Y) Z.
On the other hand, if 9 (7) is Einstein and satisfies Rc = 2~9, then \Ix (\lx Y )
= R (X, Y) X
3
- 27 \lxy'
7. THE REDUCED DISTANCE
348
We now rewrite the £-Jacobi equation in a more natural way in view of the space-time geometry associated to the Ricci flow. Consider the quantity Rc g(r) (Y) . The time-dependent symmetric 2-tensor Rc g(r) is defined on all of M whereas Y is a vector field along the path '"Y ( r) in M. In local coordinates, RCg(r) (y)i = gijRjkyk, so actually we are considering Rc as a (1, I)-tensor. Caveat: When we take the time-derivative of Rc, we consider it as a (2,O)-tensor and then raise an index to get a (1, I)-tensor! By V x [Rc (Y)] (ro) we simply mean the covariant derivative along '"Y (r) of the vector field RCg(ro) (Y h (r))) at r = TO. In this respect the vector field RCg(ro) (Y h (r))) along '"Y (r) should be distinguished from RCg(r) (Y h (r))) , where in the latter case the Ricci tensor depends on time. Combining the equations 14
D;. [Rcg(r) (Y)] = (:rRC) (Y)+ (VxRc)(Y) +Rc(VxY)-2Rc 2 (y) and
Dd~ (V~r)x) =
Vx(VyX) +
= Vx
(:r V)y X
(VyX) + (VyRc) (X)
+ (V x Rc) (Y) - (V Rc) where (V Rc) (~, we have (7.123)
1") (Z) ~ (V Rc) (Z, Y, X), and commuting derivatives,
D.!L (Rc (Y) d.,.
=
(~, 1") ,
+ Vy X)
(:r RC) (Y) + 2 (V x Rc) (Y) + Rc (V x Y) - 2 Rc
+ Vy (VxX) + R(X, Y)X + (Vy Rc) (X) - (VRc)
2
(Y)
(~'1")
,
where we used Vx (VyX)
= Vy (VxX) + R(X, Y)X.
Substituting the £-geodesic equation 1
1
o = V x X - "2 V R + 2 Rc (X) + 2r X in (7.123) and adding this to 1 ) =-Y--VxY 1 1 Dd ( --Y d.,. 2r 2r2 2r ' 14In accordance with the caveat above, :.,. Rc denotes the derivative of Rc as a (2,0)tensor and (:.,. Rc) (y)i ~ gij (t.,. RC)jk yk.
8 . .L:-JACOBI FIELDS AND THE .L:-EXPONENTIAL MAP
349
we have D..!!. (RC(Y) + V'yX dr
~Y) 27
= (:7RC) (Y)+2(V'xRc)(Y)-2Rc 2 (y) +
1
"2 V'y (V'R) - 2 (V'y Rc) (X) - Rc (V'y X)
+ R (X, Y) X +
(V'y Rc) (X) - (V'Rc)
(~, 1)
1 1 + 272 Y - :;:-V'yX.
This may be rewritten as D..!!. (RC (Y) dr
=
+ V' y X
-
~ Y) 27
(aa7 RC) (Y) + ~2 V' y (V'R) - Rc
2
(Y) +
~ Rc (Y) 27
- 2 (V'y Rc) (X) + 2 (V' x Rc) (Y) + R (X, Y) X + (V'Rc)
(1' 1) - (V'Rc) (~, 1)
- Rc (RC(Y) + V'yX -
~y) 27
-~ 7
(RC(Y) + V'yX -
~Y). 27
Define the matrix Harnack expression
J (Y)
::§:: -
(aa7 RC) (Y) - ~2 V' y (V'R) + Rc
2
(Y) -
~ Rc (Y) 27
+ 2 (V'y Rc) (X) - 2 (V' x Rc) (Y) - R (X, Y) X
(1'1) + (V'Rc) (~'1)' so that (note (- (V'Rc) (1'1) + (V'Rc) (~'1)) (Y) = 0) - (V'Rc)
(J(Y),Y) =
~H(X,y).
Thus we have the following. LEMMA
7.85. The £-Jacobi equation is equivalent to
(7.124)
+ Rc +~) (RC (Y) + V' x Y ( D..!!. ~ 7
~ Y) ~
=
-J (Y) ,
where we have replaced V' y X by V' x Y. EXERCISE
7.86. Rewrite the above equation using Uhlenbeck's trick.
350
7. THE REDUCED DISTANCE
8.2. Bounds for L:-Jacohi fields. Let c > 0 and let IS : [0, f] -+ E (-€, €), be a smooth I-parameter family of L:-geodesics. In this subsection we adopt the notation of subsection 8.1 above. Assume Y (0, s) = for s E (-€, €) (for simplicity we may assume IS (0) = I (0) for all s). We shall estimate from above the norms of L:-Jacobi fields Y (7) = Y (7, 0). By the first variation formula for the L:-Iength and the L:-geodesic equation, we have for s E (-€, c),
M, s
o
fJyL: (Ts) = 2Vf (Xs, Ys ) (f). We differentiate this again to get
= 2Vf (\7 x Y, Y) (f) + 2Vf (X, \7y Y) (f) , where we used \7y X = \7 x Y. (fJ~L:) (T)
Now the derivative of the norm squared of the L:-Jacobi field is
d~ 17=1' IYI
2
= :717=1' IY (7)1;(7) = 2 (\7 x Y, Y) (f) = 2Rc (Y, Y)(f)
(7.125)
1
+ v'T (fJ~L:) (T) -
+ 2Rc (Y, Y) (f)
2 (X, \7yY) (f),
which is expressed in terms of the second variation of L:. Let field along I which satisfies the ODE
(\7xY)
(7.126)
(7) = -Rc
(Y (7)) + 2~Y (7),
Y be a
vector
7 E [o,f] ,
Y(f)=Y(f).
(7.127)
(The first equation is the same as (7.66).) As in (7.68),
Iy (7)1
(7.128)
2
=
~ IY (f)1 2 •
In particular, Y (0) = 0 = Y (0) . Now we further assume that the IS are minimal L:-geodesics for each s E (-€, c). Let is : [0, f] -+ M be a I-parameter variation of I with
81
-
-8 is = Y, s s=o
is (f) =
IS (f)
and is (0) = IS (0) ;
this is possible because Y (0) = Y (0) and Y (f) = Y (f). Then L: (is) ~ L: (Ts) for all s, and equality holds at s = O. Hence
(fJ~L:) (T) S (fJ~L:) (T), where equality holds if Y is an L:-Jacobi field. Combining this with (7.125), we get
d~ 17=1' IYI 2 S 2 Rc (Y, Y)(f) + J:r (fJ~L:) (T) -
2 (X, \7yY) (f).
8. 'c-JACOBI FIELDS AND THE ,C-EXPONENTIAL MAP
By (7.69), since (7.126) holds and Y (0)
351
= 5, we have
(8~C) (r) - 2Vf (X, VyY) (f) =Note that
loT VTH (X,Y) dT+ IY~W -2VfRc(Y,Y) (f).
is (f)
= IS
(f) implies
VyY (f) = VyY (f). Hence
7.87 (Differential inequality for length of C-Jacobi field). Let IS : [0, T2J --t M, where T2 E (0, T), S E (-E, E), and E > 0, be a smooth family of minimal C-geodesics with Ys (0) = 5 for S E (-E, E). Then for any LEMMA
l' E
(0, T2J
the C-Jacobi field Y (T)
~I
(7.129)
dT
7=1'
=
¥: Is=o satisfies the estimate
r
IYI 2~ -~ VTH (X,Y) dT+ IY~)12, Vi Jo T
Y satisfies
(7.126) and (7.127) and quantity defined in (7.63).
where
H(X, Y)
is Hamilton's Harnack
Note that the only place where we used an inequality (versus an equality)
(8~C) (r) ~ (8~C) (r) . Hence equality holds in (7.129) if and only if the vector field Y satisfying (7.126) and (7.127) is an C-Jacobi in our derivation is
field. Then
d~ 17=1' 1Y12 = d~ 17=1' lyI 2= IY ~)12 in (7.129), and
(7.130)
J; vrH (X, Y) dT = O. From (7.125) we get
:T 17=1' 1Y12 = 2Rc (Y, Y) (f) + Jr (HessL) (Y, Y) (f) = IY ~)12
Applying Hamilton's matrix Harnack inequality to (7.129), we get LEMMA 7.88 (Estimate for time-derivative oflength of C-Jacobi field). If the solution (Mn, 9 (T)) , T E [0, TJ , to the backward Ricci flow has bounded nonnegative curvature operator and the C-Jacobi field Y (T) along a minimal C-geodesic I : [0, T2J --t M satisfies Y (0) = 5, then for C E (0,1) and l' E (0, min {T2, (1 - c)T}],
d~ 17=1' log IYI 2~ ~ (C£ (r (f), f) + 1), where C = ~. 1fT =
00,
then
352
7. THE REDUCED DISTANCE
7.89. Note that for Euclidean space, .c-Jacobi fields satisfy (and £ is constant along .c-geodesics). REMARK
IY (T)12 = const ·T. In particular, d~ log IY (T)1 2 =
f:
Since 9 (T) has nonnegative curvature operator, Hamilton's matrix inequality holds and we have for any T E [O,fJ, PROOF.
H(X,y) (T) +(~+ T~T) Rc (y,y) (T) ~ O. Since Rc T E
~
Iy (T)1 2
0 and
=
¥IY (1')1 2, from
l'
~
(1 - c) T, we get for
[0,1'],
H(X, Y) (T) ~ - (~+ T ~ T) R(-y(T) ,T) Iy (T)1 T
2
2
T (T _ T) R (, (T) , T) IY (f) I ~
1 2 --=R (, (T), T) IY (1')1 . CT
Then (7.129) implies
~
dd I _1Y12 T T=T
(
2 r JTR (-y (T), T) dT + 1) IY (1')1 CyTJO T
~
~ (~£(-y(f),f)+1) IY~)12, since I is a minimal C-geodesic. Hence d I -d T
1(2-£ (-y (f) ,f) + 1) .
_ log IYI 2 ~ -=-
T=T
T
C
Finally, we leave it as an exercise to check that when T in essence take C = 1 in the inequality above.
=
00,
one can 0
8.3. The .c-exponential map. The .c-exponential map Cexp: TM x [O,T)
--+
M
is defined by Cexp (V, f) ~ Cexpv (f) ~ IV (f), where IV (T) is the .c-geodesic with limT->o JT~~ (T) = V E TpM (and IV (0) = p). Given 1', define the C-exponential map at time l' Cf'exp: TM
--+
M
by Cf'exp (V) ~ IV (f).
353
8. L:-JACOBI FIELDS AND THE .c-EXPONENTIAL MAP
EXAMPLE 7.90 (.L:-exponential map on a Ricci flat solution). To get a feel for the .L:-exponential map, we first consider a Ricci fiat solution (Mn,9 (7) = 90). Here, by (7.23), for V E TpM,
.L: exp (V, f) = exp ( 2v1rV)
,
where exp is the usual exponential map of (M, 90) with basepoint p. Note that for a Ricci flat solution, the .L:-exponential map has the scaling property:
= .L:exp (cv, ;)
.L:exp (V, f)
for any c > O. However this is not true for general solutions of the Ricci flow. The .L:-exponential map at f = 0 is related to expg(O) which is the usual (Riemannian) exponential map with respect to the metric 9 (0) . LEMMA 7.91 (.L:-exponential map as f ---+ 0). Let (Mn, 9 (7)) , 7 E [0, T] , be a complete solution to the backward Ricci fiow with bounded sectional curvature. Given V E TpM, as f ---+ 0, the .L:-exponential map tends to the Riemannian exponential map of 9 (0) in the following sense:
(7.131)
!im .L:exp ( 1"-+0
1G V,
2y 7
f) =
expg(O)
(V) .
From the proof we can see that the convergence in (7.131) can be made into
Coo -convergence. Motivated by the Ricci flat case, we define the path f3 : [0, 1]
PROOF.
---+
M by f3(p) so that f3 (1)
= .L: exp
~ .L:exp ( 2y1G7 V ,p2f)
= "1_1 V (p2f) , 2,j¥
C.i¥ V, f) . (Note that f3 depends on f but we do not
emphasize this in our notation.) We have (7.132)
df3 () d ( ( 2_)) 2 _ d'Y ~ v ( 2-) -d p = -d 'Y_1_V P 7 = p7 d P7 P P 2,j¥ 7
.
Hence the .L:-geodesic equation (7.32) becomes O=V'
1 ~ 2pr dp
(3 ) 1 d (3 ) --V'R+2Rc 1 ( -1_d( ----=2p7 dp 2 2p7 dp
1 1 df3 +--_-_-. 2 2p 7 2p7 dp
Multiplying this by 4p2f2 yields for p E [0,1] , (7.133)
df3 2 2 (df3) V' ~! dp - 2p f V'R + 4pf Rc dp
= O.
The covariant derivative and Ricci tensor are with respect to 9 (p2f) . Since
1
d'Y_1_V lim yT
1"-+0
~,j¥ 7
(7) =
G V, 2y7
354
7. THE REDUCED DISTANCE
we have df3 d'Y~v lim -d (p) = lim 2pf ~..;¥ (p2f) = V p--->O P p--->O r independent of f. Hence, by taking the limit of (7.133) as f "V~~O) fp = 0 and
--t
0, we have
dp
P 1-+ !im £ exp (
1;:; V, p2f) 2yr
r--->O
= !im f3 (p) r--->O
is a constant speed geodesic with respect to g (0) and with initial vector V. Thus when we evaluate it at p = 1, we get that the limit is exp9(0) (V). 0 Next we compute the differential of £fexp. LEMMA
7.92. For f E (0, T), £f exp is differentiable at V and the tan-
gent map
is given by
=
(£fexp). (W)
J(f),
where J(r) is the £-Jacobifield along £exp(V,r) with J(O)
=0
~
and
do-
J(2)
1
4
0'=0
= W.
PROOF. We have a family of £-geodesics £ exp(V for s E (-c, c). By the definition of £-Jacobi field,
J(r)
~
dd 1 s
8=0
1
r
E
[0, fJ,
£exp(V +sW,r)
is an £-Jacobi field. Since £ exp(V + sW, 0)
dd
+ s W, r),
£exp(V + sW,r)
= p, we have
J(O)
= O. From
= V + sW,
0- r=O
where
0-
= 2..j7, we get by taking dd
fa 18=0' J(r)
1
= W.
0- r=O
Note that the tangent map of £1' exp at V is given by D [£1'exp(V)] (W)
=
:SI8=0 £exp(V + sW,f).
The lemma follows. As a simple corollary of the lemma we have
o
8. C-JACOBI FIELDS AND THE C-EXPONENTIAL MAP
355
COROLLARY 7.93. Fix f E (0, T) and consider the £-exponential map £f' exp : TpM - t M. Then V is a critical point of the map £f' exp if and only if there is a nontrivial£-Jacobi field J(7) along £exp(V, 7), 7 E [0, f], such that J(O) = 0 and J(f) = O. PROOF. If V is a critical point of £f' exp, then there exists W such that D [£exp(V, f)] (W) = O. By the lemma, £exp(V + SW,7) is the required £-Jacobi field. On the other hand, if we have a nontrivial £-J acobi field J1 (7) with J 1(0) = 0 and J1(f) = 0, then we define W ~ d~IT=oJ1(7). Since J 1(7) is nontrivial, we have W =1= o. By the uniqueness of solutions of the initial-value problem for £-Jacobi fields, we know that
is Is=o
D [£f'exp(V)] (W)
=
J(f)
=
J1(f)
= o.
o
We see that V is a critical point of £f' expo
The Hopf-Rinow theorem in Riemannian geometry can be generalized for £-geodesics and the £-exponential map. The proof of this result will appear elsewhere. LEMMA 7.94 (£'-Hopf-Rinow). Suppose (Nn, h (7)),7 E [0, TJ, is a solution to the backward Ricci flow satisfying the curvature bound IRm (x, 7)1 ~ Co < 00 for (x, 7) E N x [0, T]. The following are equivalent:
(1) for every 7 E [0, T), the metric h (7) is complete; (2) £exp is defined on all ofTpNx [O,T) forsomepEN; (3) £ exp is defined on all of TpN x [0, T) for all pEN. Moreover,
(4) any of the above statements implies that given any two points p, q and 0 ~ 71 < 72 < T, there is a minimal £-geodesic T : [71,72] - t N joining p and q. 8.4. £-cut locus. We have the following simple lemma which is analogous to the corresponding theorem in Riemannian geometry. LEMMA 7.95 (When £-geodesics stop minimizing). Given V E TpM, there exists 7V E (0, T] such that
£ ( TVI[O,Tl) = L (Tv (7),7)
for all 7 E [0, Tv)
and £, ( TVI[o,Tl)
where TV : [0, T)
-t
> L (Tv (7),7)
for any 7 E (Tv, T),
M is the £-geodesic with limT--+O JT~ (7)
= V.
PROOF. The existence of Tv follows from the additivity property for concatenated paths (7.20). The fact that 7V > 0 follows from Lemma 7.29.
o
7.
356
THE REDUCED DISTANCE
That is, if TV < T, then
is the first time the £-geodesic rV : [0, T) -+ TV = T if and only if rV is minimal. The lemma establishes that either rV is minimal or there exists a first positive time TV past which rV does not minimize. Let r : [0, T) -+ M be an £-geodesic with r (0) = p. We say that a point (, (f) , f) , l' E (0, T), is an £-conjugate point to (p,O) along r ifthere exists a nontrivial £-Jacobi field along r which vanishes at the endpoints (p,O) and (,(1'), f). A point (q, f) is an £-conjugate point to (p, 0) if (q, f) is £-conjugate to (p, 0) along some minimal £-geodesic r (T) , T E [0,1'], from p to q. If r (T) = rV (T) for some V E TpM, then this is equivalent to V being a critical point of the £-exponential map Dr exp (see Corollary 7.93). TV
M stops minimizing. On the other hand,
DEFINITION 7.96. (i) The £-cut locus of (p,O) in the tangent space of space-time is defined by £C(p,O) ~ {(V, TV) : V E TpM} ,
where TV is defined above. Since TV > 0, we have £C(p,O) C TpM x (0, T]. (ii) The £-cut locus of (p, 0) at time l' E (0, T] in the tangent space is defined by (iii) Define n(p,O) (f) ~ {V E TpM : TV > f}. In words, O(p,O) (f) is the open set of tangent vectors at p for which the corresponding £-geodesic minimizes past time f. Note that n(p,O) (f) is not necessarily star-shaped in the sense that if V E n(p,O) (f) , then a V E n(p,O) (7) for any a E (0,1). However n(p,O) (f) is an open subset in TpM and n (T2) en (Td if Tl < T2· Next we define the £-cut locus of the map Dr exp for l' E (0, T) . DEFINITION 7.97. (i) The £-cut locus of (p,O) at time l' is defined by £ Cut(p,O) (f) ~ {£-fexp (V): V E TpM and
TV
= f}.
(ii) We define £ CutZp,O) (f) to be the set of points q E M such that there are at least two different minimal £-geodesics on [0,1'] from p to q. (iii) We define .cCut(p,O)(f) to be the set of points q such that (q,f) is £-conjugate to (p, 0) . EXERCISE 7.98. Show that for l' E (0, T) we have q i £ Cut(p,O) (f) if and only if for every minimal £-geodesic r : [0,1'] -+ M joining p to q, we may extend r as a minimal £-geodesic past time f. There is a characterization of £-cut locus points analogous to the characterization of cut locus points in Riemannian geometry. The first lemma below can be proved using the locally Lipschitz property of the £-distance L, while the second lemma below can be proved via a calculation similar to
8.
C-JACOBI FIELDS AND THE C-EXPONENTIAL MAP
357
the proof of the Riemannian index lemma. We will give the details of the proof elsewhere. LEMMA 7.99 (.L:-cut locus). (i) .L:Cut(p,O)(f) = .L:Cutzp,O)(f) U .L:Cut(p,O)(f) and is closed. (ii) .L:CutZp,O)(f) has measure O. (iii) .L:Cut(p,O)(f) is closed and has measure o. The proof of the above lemma depends on an index lemma for .L:-length. Define the .L:-index form .L: I(Y, W) by
(7.134) .L: I(Y, W)
~
l:
b
l
v'T
1
!V'yV'wR+ (R(Y,X)W,X) + (V' x Y, V' x W) - (V'y Rc)(W, X)
dT.
- (V'w Rc) (Y, X) + (V' x Rc) (Y, W) Note that the second variation of .L:-length (7.62) is related to the .L:-index form by
(6~.L:)
(7.135)
h) =
2v'T(V'yy,X)I~~ +2.L:I(V,V).
On the other hand, the .L:-index form is related to the .L:-Jacobi equation by .L: I(Y, W)
= JT (V' x Y, W) I~~
r
b
[I
V'x (V'x Y ) - Rm(X, Y)X - !V'y (V'R)
- JTa JT \ +2(V'yRc)(X)+2Rc(V'xY)+2~V'xY This can be proved by integrating by parts on the term in (7.134).
)] ,W
dT.
..jT (V' x Y, V' x W)
LEMMA 7.100 (.L:-index lemma). Let'"Y be an .L:-geodesic from (p, Ta) to (q, Tb) such that there are no points .L:-conjugate to (p, Ta) along '"Y. For any piecewise smooth vector field W along'"Y with W(Ta) = 0, let Y be the unique .L:-Jacobi field such that Y(Ta) = W(Ta) = 0 and Y(Tb) = W(Tb). Then .L: I(Y, Y) :::; .L: I(W, W)
and the equality holds if only if Y = W. Here we have used the obvious generalization of the definition of .L:-conjugate point with (p,O) replaced by (p, Ta) . Lemma 7.99 implies the following. COROLLARY 7.101 (Differentiability of L away from .L:-cut locus). Given f E (0, T) and q E M, suppose that there is only one minimal .L:-geodesic '"Y joining (p,O) and (q, f) and suppose that (q, f) is not an .L:-conjugate point of (p, 0), i.e., (q, f) is not an .L:-cut locus point. Then the L-distance L(·,·) and reduced distance £ are C 2 -difJerentiable at (q, f).
7. THE REDUCED DISTANCE
358
PROOF. Suppose limT->o JT~(O) = Vj the hypothesis implies by Lemma 7.92 that there is some E > and some small neighborhood Uv of V E TpM such that the map
°
(C exp, id) : Uv X (1' ~ E, l' + E) ~ M x (1' - E, l' + E) , (C exp, id) (W, r) = (C T exp (W) ,r) is a local diffeomorphism. For each WE Uv and r* E (1'-E, 1'+E), we claim the curve Cexp(W, r), r E [0, r*l, is a minimal C-geodesic. Hence using the local diffeomorphism property, there are an El > 0, a small neighborhood Uq of q, and a family of minimal C-geodesics Iq,T* smoothly depending on the endpoint q E Uq and 'r* E (1' - EI, l' + EI). Now L (q, r*) = C (,q,T.) is a smooth function of (q, r*) , L is differentiable near (q, f). We now prove the claim by contradiction. If the claim is false, then there is a sequence of points (Wi,ri) ~ (V, f) such that Cexp(Wi,r), r E [O,ril, is not a minimal C-geodesic. Let Cexp(Wi' r), r E [0, ril, be a minimal C-geodesic from p to Cexp(Wi, ri). As in the proof of Lemma 7.28, using Lemma 7.13(ii) and the C-geodesic equation, it is easy to show that IWil
g(p,O)
is bounded. Hence there is a subsequence Wi ~ Woo. If Woo =J V, then we get two minimal C-geodesics IV and IWoo joining (p,O) and (q, f), which contradicts the assumption of the lemma. If Woo = V, then (C exp, id) cannot be a local diffeomorphism since (C exp(Wi , rd, ri) The claim is proved and the lemma is proved.
= (C exp(Wi' ri), ri) . 0
As a simple consequence, we have (7.136)
is a diffeomorphism. Note that M\CCut(p,O)(1') is open and dense in M. Now we end this subsection by rewriting the formula for the C-index form in terms of Hamilton's matrix Harnack quadratic. Using
d~ [Rc (Y, W)] = (! RC) (Y, W) + (\7 x Rc) (Y, W) + Rc (\7 x Y, W) + Rc (Y, \7 x W) , we may write (7.134) as
l
CI(Y, W) = -
+
l
d
Tb
Ta
JT dr [Rc (Y, W)] dr
(,t Rc) (Y, W) + Tb
Ta
y'T
+ (R (Y, X) W, X) -
~\7y\7wR - (Rc (Y), Rc (W))
- (\7 y Rc) (W, X)
(\7w Rc) (Y, X) + 2 (\7 x Rc) (Y, W)
+ (Rc (Y) + \7 x Y, Rc (W) + \7 x W)
dr.
8 . .c-JACOBI FIELDS AND THE .c-EXPONENTIAL MAP
359
Rearranging terms and integrating by parts, we express this as follows. Define the symmetric 2-tensor Q by Q (Y, W)
~ (:r RC) (Y, W) + ~ Vy V w R 1 + 2r Rc (Y, W) -
(Rc (Y) ,Rc (W))
.
+ (R (Y, X) W, X) - (Vy Rc) (W, X) (Vw Rc) (Y, X) + 2 (V x Rc) (Y, W)
and
S (Y)
~
Rc (Y)
+ Vx Y -
1 -Y. 2r
Then we have LEMMA 7.102 (C-index form). The C-index form CI(Y, W) can be written as CI(Y, W)
= - ViRc (Y, W)I~: +
+
l
Tb
~
Vi
l
TaTb
ViQ (Y, W) dr
((S(Y)'S(W))+(S(Y)'~))
+ \~, / Y s (W) + (y'\f) ~
Note that, assuming [X, Y]
dr.
= 0 and [X, W] = 0, we have
d~ ((Y'rW )) = ~ (RC + Sym (VX) - 2~g) (Y, W), where Sym (V X)ij ~ ! (ViXj + VjXi) = !CXg. 8.5. C-Jacobian. First we recall the Jacobian in Riemannian geometry. Let (Mn, be a Riemannian manifold, let p EM, and given V E TpM
g)
with IVI = 1, let ')'v : [0, 8V) ~ M be the maximal unit speed minimal geodesic with 'Y (0) = V. Take {Ed?:11 to be an orthonormal frame at p perpendicular to V and define Jacobi fields {Ji (8)} along"YV so that Ji (0) = 0 and (VvJd (0) = Ei. The Jacobian J is defined by J hv (8))
~
Jdet ((Jd 8) , Jj (8)) ),
where ((Ji (8), Jj (8))) is an (n - 1) x (n - 1) matrix. Note that (7.137)
lim J hv (8)) = 1. 8n - 1
s-+O
Let d17sn-l denote the volume form on the unit (n - I)-sphere in TpM, which naturally extends to TpM -
pin
M. We define the
{O}, and let Cut (p) be the cut locus of
(n - I)-form d17 on M\ (Cut (p) U {p}) by d17 ~ (exp;1)* d17sn-l,
360
7. THE REDUCED DISTANCE
where the exponential map is restricted to inside the cut locus in the tangent space. Then the volume form of 9 on M\ (Cut (p) U {p}) is given by dl1g
= J hv (8)) dr Ada.
The volume forms of the geodesic spheres S (p, r) at smooth points, are given by das(p,r)
~ {x
EM: d (x, p)
= r} ,
= J hv (8)) da.
The Jacobian is related to the mean curvatures of the geodesic spheres and the Ricci curvatures of the metric 9 on M by the following formulas:
8
8r logJ
=H
and
~H = -Rc (~,~) -lhl 2 8r 8r 8r < _ Rc (~ ~) _ H2 -
8r'8r
n-l'
where h is the second fundamental form of S (p, r). The Bishop-Gromov volume comparison theorem may be proved this way (see [111] for example). Now we turn to the case of Ricci flow. Let ')'v (7), 7 E [0, T), be an L:-geodesic emanating from p with limT-+o VT'Y (7) = V. Let Jr (7), i 1, ... , n, be L:-Jacobi fields along ')'v with
Jr (0) = 0 and
('VvJr) (0) = EP,
where {En ~=l is an orthonormal basis for TpM with respect to 9 (0). Note that Jr (7) is a smooth function of V and 7 > 0 since g( 7) is smooth. Via the orthonormal basis {EP} ~=l we can identify TpM with IRn. Since D (L:exp(V, 7)) (EP) = Jr (7) (see Lemma 7.92), the Jacobian of the L:exponential map L:Jv (7) E IR (called the L:-Jacobian for short) is the square root of the determinant (computed using the inner products on the tangent spaces from the Riemannian metric 9 (7)) of the basis of L:-J acobi fields: (Jr (7), ... , (7)). That is,
J;:
L:Jv (7) ~ It is clear that L: Jv (7) is a smooth function of (V, 7) when 7 equivalent way of describing L: J v (7) is to define
L: J v (7) dx (V) ~ [( L:T exp(V))* dl1g( T,L:
T
> o. Another
exp(V)) ] '
where dx is the standard Euclidean volume form on (TpM,g(O,p)).
8.
£-JACOBI FIELDS AND THE £-EXPONENTIAL MAP
361
To get a feeling of the £-Jacobian, we calculate an example. EXAMPLE 7.103 (£-Jacobian of Ricci flat solution). Recall the fact that if (M n , 9 (7) = 90) is a Ricci flat solution, then an £-geodesic is of the form "t (7) = (3 (2ft) where (3 ((1) is a constant speed geodesic. Then an £-Jacobi field is of the form JV (7) = K (2ft) where K ((1) is a Riemannian Jacobi field along (3 ((1) with respect to 90. . JV () 2;;; ¥ (2y'T) Hence, by ch oosmg n 7 = V 7 1d /3 (" '-)1 ' d"
£Jv (7)
2yT
9(0)
= 2ftJv (2v'T)
,
where Jv is the Jacobian of the Riemannian exponential map of 90. Since by (7.137), lim Jv ((1) - 1 (1n-1 - ,
u->o+
we have lim £Jv (7) T->O+
= 2n.
7n/2
Note that for Euclidean space R n we have £Jv (7)
= 2n7 n/ 2 .
As suggested by Example 7.103 and (7.131), we now prove the following lemma. LEMMA 7.104 (£-Jacobian as 7 ~ 0). Let (Mn,9 (7)) be a solution of the backward Ricci flow with bounded sectional curvature. We have the following asymptotics for the £-Jacobian at 7 = 0: lim £Jv (7)
(7.138)
T->O+
= 2n.
7 n/ 2
PROOF. Let Ei (7) denote the parallel translation of with respect to 9 (0). Since Jr (0)
= 0 and
E?
(\7 d: Jr) (0) =
along "tv (7)
(\7vJr) (0) =
E?, then by the definition of derivative, . IJr (7) - 2ft E i (7)1 (0) . IJr (7) - (1Ed7) 1 (0) 11m 9 = 11m 9 T->O+ 2ft T->O+ (1
= O.
Hence lim £JV}7) = lim 7- n/ 2
T->O+
7n 2
T->O+
det ((2v'TEi (7), 2v'TEj (7))g(0))
= 2n. D
For the proof of the no local collapsing result via the L-distance in Chapter 8 we need the following properties of the £-Jacobians.
362
7. THE REDUCED DISTANCE
7.105 (Time-derivative of £-Jacobian). Let (M n , g (7)), 7 E [0, T], be a solution of the backward Ricci flow with bounded sectional curvature. Along a minimizing £-geodesic /'v (7) , 7 E [0, Tv), with /'v (0) = p, where Tv is defined in Lemma 7.95, for 0 < f < Tv the £-Jacobian £ J v (7) satisfies PROPOSITION
d ) (f) (-log£Jv d7
(7.139)
~
n --= 27
1 K,
-3
27"2
where K = Kbv, f) is defined by (7.75). Equality in (7.139) holds at the point /,v(f) only if
Rc bV(f), f)
(7.140)
+ (Hess f) bv(f), f)
= g
bV2~)' f).
7.106. The proof of (7.139) is closely modeled on that of the classical Bishop-Gromov volume comparison theorem. Here we follow the derivation using £-Jacobi fields. There are other ways to prove volume comparison such as in Li [246]. REMARK
REMARK
7.107. If we let it
= 2...;T and u = 2ft, then (7.139) says
Compare this with (7.76). PROOF OF PROPOSITION 7.105. Choose an orthonormal basis {Ei (f)} of TW{f)M. Since there is no point on /,v(7) , 0 ~ 7 ~ f, which is £conjugate to (p,O) along /,v, we can extend Ei (f) to an £-Jacobi field Ei (7) along /'v for 7 E [0, f] with Ei (0) = O. Actually for the same reason, we know that both {Jt (7)} E T-yv{r)M and {Ei(7)} E Tw(r)M are linearly independent when 7 E (0, fl. We can write n
Jt (f) =
LA{Ej (f) j=l
for some matrix
(Ai) E GL (n, ~). Then
for all 7 E [0, fJ, since we cannot have a nontrivial £-Jacobi field vanishing at the endpoints 7 = 0, f.
9. WEAK SOLUTION FORMULATION
363
Now we compute that the evolution of the £-Jacobian along I'v is given by
The last inequality is due to (7.129). Here the along I'V satisfying \l X Ei
Ei (f)
where
= Ei (f) and H
Ei (r)
are the vector fields
- ) 1= - Rc ( Ei + 2r Ei,
(x, Ei) (r) is the matrix Harnack quadratic
given by (7.63). By (7.72), we have have
(Ei' Ej) (r)
n
~H (X,Ei)
(r)
=
¥Oij,
and by (7.74), we
= ~H(X) (r).
t=l
So
(d~ log £ J) (f) ~ - 2'1'~/2 loT r 3/2H (X) dr + ; 1
n
= - 2'1' 3 / 2K + 2'1'.
(7.141)
If equality in (7.139) holds, then we have equality in (7.129) for each Y (r) = Ei (r), i = 1,··· ,n. By (7.130), we have ) 2Rc ( Ed'1'),Ed'1')
for each i. Since (7.140).
1 ( ) lEi '(f) + y'f(HessL) Ei (f) ,Ei(T) = 1'
Ei (f) =
r
Ei (f) can be chosen arbitrarily, this implies 0
9. Weak solution formulation The purpose of this section is to prove the integration by parts inequality (7.148) for the reduced distance f and to give the inequalities we proved for f a weak interpretation. We first recall some of the well-known results in real analysis which we shall need. An excellent reference for properties of
7.
364
THE REDUCED DISTANCE
Lipschitz functions and other aspects of real analysis on lRn is the book by Evans and Gariepy [139]. Many of the results in their book easily extend to Riemannian manifolds; when this is the case, we state the extensions without proof. In this section we shall assume that is a complete Riemannian manifold.
(Mn, g)
9.1. Locally Lipschitz functions. Recall the definition of differentiability on Riemannian manifolds. DEFINITION 7.108 (Differentiable function). A function differentiable at p E M if there exists a linear map
f :M~
lR is
Lp: TpM ~ lR such that lim If (expp (X)) - f (p) - Lp (X)I = x-o IXI
o.
When this is the case, by definition we write
If (expp (X)) - f (p) - Lp (X)I
=
0
(IXi)
X ~
as
o.
This implies for every X E TpM that the directional derivative
Dxf ~ lim f (expp (sX)) - f (p) = Lp (X) 8-0
S
exists. REMARK 7.109. Note that differentiability can be defined more generally for differentiable manifolds, but in the definition above we chose to endow the manifold with a Riemannian metric. Now we list three results in Evans and Gariepy's book. The first is Theorem 2 in §3.1.2 on p. 81 of [139]. LEMMA 7.110 (Rademacher's Theorem). Let
(M,g)
be a Riemannian
manifold. If f : M ~ lR is a locally Lipschitz function, then f is differentiable almost everywhere with respect to the Riemannian (Lebesgue) measure. Secondly, Theorem 5 in §4.2.3 on p. 131 of [139]. LEMMA 7.111 (Locally Lipschitz is equivalent to being in Wl~':l Let U be an open set in a Riemannian manifold
(M, g).
locally Lipschitz if and only if f E Wl~~oo (U) . Thirdly, Theorem 2 in §2.4.2 on p. 76 of [139].
Then
f :U ~
lR is
9.
and
WEAK SOLUTION FORMULATION
365
LEMMA 7.112 (Hausdorff dimension of a Lipschitz graph). Let
(Mr, 91)
Ml ~ M2
is locally
(MH\ 92)
be two Riemannian manifolds. If f :
Lipschitz in the sense that for any p E M1, there is an open neighborhood Up of p and a constant Cp such that dg2 (J (qd ,f (q2)) ~ Cpdg1 (ql, q2) for any ql, q2 E Up, then 'Hdim
{(x, f (x)) : x
E
Ml} = n,
where 'Hdim denotes the Hausdorff dimension. In particular, the (n + m)dimensional Riemannian measure vanishes: measMlxM2 {(x,
f (x)) : x E Ml} = O.
Later we shall recall some more basic results, especially about convex functions, as we need them. Now we give a proof that integration by parts holds for locally Lipschitz functions. We say a vector field v on M is locally Lipschitz if for any p E M and local coordinates {xi} in a neighborhood of p, we have for each i that the function vi (x) is locally Lipschitz, where v (x) vi (x) a~i. It is well known that integration by parts holds for Lipschitz functions.
'*'
LEMMA 7.113 (Integration by parts for Lipschitz functions). Let f be a locally Lipschitz function on M and let v be a locally Lipschitz vector field on M. Suppose that at least one of f and v has compact support. Then
1M f div v dJ.Lg = - 1M v . V' f dJ.Lg. Here div v and v . V' f are defined with respect to
9.
PROOF. We prove the lemma in the case where v has compact support; the other case can be proved similarly.15 Rademacher's Theorem says that both derivatives div v and V' f exist almost everywhere. Since v has support in some compact set K, we have lvi, Idiv vi E L oo (K), and
f, IV' fl
E
L~c (M) , so the integrals in the lemma make sense. We can
choose a smooth partition of unity {
1M f div (
7. THE REDUCED DISTANCE
366
Let U1 be an open neighborhood of the support of v satisfying U1 CC U. 16 We can apply Theorem 1 in §6.6.1 on p. 251 of [139] to I and v. This tells us that for any c > 0 there exists a C 1 function Ie and a C1 vector field Ve = v~ a~t defined on U such that Ve has compact support in U1 and
(x) or '\lIe (x) i= '\l I (x)} ~ c, U1 : Ve (x) i= v (x) or '\ljVe (x) i= '\ljV (x) for some j}
meas {x E U1 : meas {x E
188I~ (x)1 ~ C (n) Lip (j, U)
for all i,
188V~J (x) I ~ C (n) Lip (vi, U)
for all i, j,
sup xEV
sup xEV
Ie (x) i= I
~
c,
X
x
where meas is the n-dimensional Riemannian (Lebesgue) measure on U and Lip (j, U) is the Lipschitz constant of I on U. By the divergence theorem, which clearly holds for C 1 functions, we have
f Ie div vedJ-Lg =
lUI
-
f Ve· '\l IedJ-Lg.
lUI
Taking c '\. 0, we get JUI I div vdJ-Lg = - JUI V . '\l IdJ-Lg and hence
1M I div v dJ-Lg = - 1M v . '\l I dJ-Lg. o In summary, the divergence theorem for Lipschitz functions follows from a standard approximation result. EXERCISE 7.114. Prove Lemma 7.113 in the case where has compact support.
I,
instead of v,
9.2. Convex functions. Convex functions have nice differentiability properties. Recall the notion of derivatives on JRn. 7.115 (First and second derivatives). Let U C JRn be an open set. Given a continuous function I : U --t JR, we say DEFINITION
(i) I has first derivative DI (x) E JRn at x E U if
II(y)-I(x)-DI(x)·(y-x)l=o(ly-xl)
as
y--tx.
This definition agrees with Definition 7.108. E M nxn at x E U, where M nxn is the set of n x n matrices, if there exists a vector D I (x) E JRn
(ii) I has second derivative D2 I (x)
16UI CC U means UI is compact and UI C U. In this case we say that UI is compactly contained in U.
9.
WEAK SOLUTION FORMULATION
367
(the first derivative) such that If (y) - f (x) - D f (x) . (y - x) -
=o(ly-xI2)
as
~ (y -
x)T . D2 f (x) . (y -
x)1
y-tx.
We also recall the following related notion (see p. 167 of [139] for example). DEFINITION 7.116 (Locally bounded variation). Let U be an open set in ~n. A function fELloe (U) has locally bounded variation if for every open set U1 C C U, sup
{luI
f div ¢dx :
C;
where the lower index c in case we write f E BVloc (U) .
¢ E C~ (U1 ; ~n), I¢I ~ 1} < 00, indicates having compact support. In this
Clearly if fECI (U) , then f E BVloc (U) since
r
JUI
fdiv¢dx
=
-1
Vf· ¢dx
UI
~ meas(U1)'
sup
IVfl < 00
SUPP
for all U1 CC U. On the other hand, as a partial converse, note that as remarked on p. 166 at the beginning of Chapter 5 in [139]' " ... a BV function is 'measure theoretically C 1 . '" The following lemma is well known; see Theorem 1(i) in §6.3 on p. 236 of [139] for the proof of (i), Aleksandrov's Theorem in §6.4 on p. 242 of [139] for the proof of (ii), and Theorem 3 in §6.3 on pp. 240-241 of [139] for the proof of (iii). Let B (r) C ~n denote the ball ofradius r centered at
O. -t
LEMMA 7.117 (Regularity properties of convex functions). Let f : B (r) be a convex function. Then (i) f is locally Lipschitz, (ii) (Aleksandrov's Theorem) f has second derivative D2 f (x) for a.e. x E B (r), (iii) g~ has locally bounded variation for each i, and
~
D2 fELloe (B (r); Mnxn) . REMARK 7.118. Note that it follows easily from the convexity of f that
D2 f (x) ~ 0 when it exists. We consider the approximation of continuous functions, in particular convex functions, by smooth functions via the convolution with mollifiers. A standard mollifier is 'fJ : ~n - t ~ defined by 'fJ (x)
~
{ae 1/(lx I2 -1)
o
if if
Ixl < 1, Ixl ~ 1,
368
7. THE REDUCED DISTANCE
where a > 0 is chosen so that fB(l) 'TJdx = 1. This function is Coo with support contained in B (1) c ~n and all derivatives vanishing for Ixl ~ 1 including Ixl = 1. Let f : B (r) ~ ~ be a continuous function. For E E (0, r) define the mollified function
by
r
fe(Y)=;=~
(7.142)
JB(e)
En
f(y+z)'TJ(~)dz. E
It is a standard fact that fe is Coo.
We now prove a lemma about convex functions. LEMMA 7.119 (Mollifiers and derivatives). Let f : B (r) ~ ~ be a continuous function. Let a~i and ay~~yj denote the standard partial derivatives on ~n.
(i) If f has first derivative Df (x) at x
(aaf~ (X))n
lim e->O+
Y
E B (r), then
= Df (x).
i=l
(ii) If f has second derivative D2 f (x) at x
E B (r), then
(iii) If f is a convex function on B (r), then fe is a convex function on B(r-E). (iv) If f E Lfoc' where p E [1, (0), then fe converges to f in Lfoc' PROOF.
(i) Define
1 by
f (x + y) =;= f (x) Then
1(y)
+ Df
has first derivative D 1(0) =
mollified function
(x) . Y + 1(y) .
O.
It suffices to prove that the
Ie (y) satisfies lime->o+ Ut (0) = 0 for each i. Note that 1 fe(Y)=n E
1 -
f(z)'TJ
(z- - y) dz
B(y,e)
E
and
ale (0) = ayt
_~ En
r
JB(e)
1(z) (a'TJ.) (~) dz. t aZ
E
E
9. WEAK SOLUTION FORMULATION
369
Since Di(O) = 0, given any Cl > 0 there exists 0> 0 such that li(y)1 ::; El Iyl whenever Iyl ::; o. Hence, if c ::; 0, then
{)i~ (0) {)y~
: ; ~ 1r cn
::; El
Ii (z) 1.1
B(e)
r
{)T].I (:') dz
{)z~
c
E
1:'1.1 {)T]·I (:') d (:') {)z~
c
1B(e)
E
E
= C1 Cl,
where Cl
~
fB(I)
IYI'I~I (y) dy
is independent of E. This implies
lim
e~O+
{){)i~ (0) = O. y~
(ii) Write I (x + y)
~ I (x) + DI (x) . y + ~yT . D2 I (x) . y + i (y).
i
°
i
i
Then (y) has first derivative D (0) = and second derivative D2 (0) = O. It suffices to prove that for all (i,j) the mollified function ie (y) satisfies
lime~o+ a~:£~j (0) = O.
Note that
nIl In(z) T] (z- - y) dz
Ie (y) = n E
and {)2 ie {)yi{)yj
(0) =
1
B(y,e)
r
c n 1B(e)
E
({)2T])
n
I
(z)
{)zi{)zj
(Z) dz €
E2'
Since Di (0) = 0, given any El > 0 there is a 0> 0 such that Ii (y)1 ::; El1yl2 when Iyl ::; O. Hence when E ::; 0,
where C2
~
~ ayia~j (0)
= 0 as c ---+ 0+.
fB(I)
lyI2Ia;2!JyJ I (y) dy is independent of
E.
This implies that
370
7.
THE REDUCED DISTANCE
(iii) For 0 :s A :s 1 and x, y
E
B (r - E) , we compute
fe(Ax+(I-A)y)=2.. { f(A(X+Z)+(I-A)(y+z))1J(~)dz ~hw E
:s2.. { (Af(x+z)+(I-A)f(y+z))1J(~)dz En JB(e) E = Afe (X)
+ (1 -
A) fe (y) .
(iv) See Theorem 6 on p. 630 of Evans [131].
9.3. Functions with Hessian upper bound. Recall that is a complete Riemannian manifold. First we give a definition.
o
(Mn, g)
DEFINITION 7.120 (Hessian upper bound in support sense). Let C E R and W C M be an open set. A continuous function f : W --+ R has the Hessian upper bound C in the support sense if for any pEW and any E > O·there is a neighborhood U of p and a C2 local upper barrier function t.p : U --+ R such that t.p (P) = f (p), f (x) :S t.p (x) for all x E U, and VV t.p (p) :S C + E. We denote this by Hess supp (f) :S C. REMARK 7.121. Clearly we can generalize the above definition to the case where C : W --+ R is a function. However we shall only need the case where C is a constant. It is easy to see that when f is C 2 , Hess supp (f) :S C implies VV f (p) :S C for each pEW. The following elementary lemma enables us to study the differentiability properties of functions with a Hessian upper bound via the theory of convex functions in Euclidean space. The idea is that we can add a suitable smooth function to a function with a Hessian upper bound to make it concave. Let Bp (r) C TpM denote the open ball of radius r centered at 0, and let B(p, r) C M denote the open geodesic ball of radius r centered at p E M.
LEMMA 7.122 (Functions with a Hessian upper bound). Let W C M be an open set, let f : W --+ R be a continuous function, and let pEW. (i) IfHess supp (f) :S C, then for any c > 0 there exists r > 0 and a C 2 function 'l/J on B (p, r) such that Hess supp (f + 'l/J) :S -c on B (p, r). (ii) If Hess supp (f) :S 0 on W, then - f is a convex function on (W, g) in the Riemannian sense, i. e., the restriction of - f to each geodesic segment in W is convex. (iii) Let r :S min { 1, injiP )} and x: B (p, r) --+ Bp (r) C TpM be normal coordinates. If there exist constants C 1 , C2 < 00 such that for each q E B (p, r) there exists a local upper barrier function t.pq for f at q satisfying
IVt.pq I (q) :S C1
and VVt.pq (q) :S C2g (q) ,
9. WEAK SOLUTION FORMULATION
371
then there exists a C2 function 'l/J (x) defined on Bp (2r) C TpM such that (J 0 exp; 1 +'l/J) (x) is a concave function on Bp (r) with respect to the Euclidean metric 9 (p). In particular, f is locally Lipschitz on B (p, r) .
° r >°
(i) We may assume C + c > since otherwise we are done. Since \1\1 [d(x,p)2] (p) = 2g(p) , where 9 is the metric, by the continuity PROOF.
of \1\1 [d(x,p)2] near p, there exists
(depending only on 9 and how
big of a ball centered at p fits inside W) such that \1\1 [d (X,p)2] 2: 9 in
= - (C + c) d (x, p)2 . Then Hess supp ('l/J) = \1\1 'l/J ::; - (C + c) g.
B (p, r) C W. Let 'l/J (x)
Since Hess supp (f) ::; c, we conclude Hess supp (f + 'l/J) ::; -c on B (p, r). Note that we may also assume I\1'l/J I ::; C'on B (p, r) for some constant C ' < 00. (ii) Let'Y : [0, a] ~ W be any geodesic parametrized by arc length. To verify that - f is a convex function on (W, g) in the Riemannian sense, we need to show that f 0 'Y : [0, a] ~ lR is a concave function. Since f 0 'Y also satisfies Hesssupp (f 0 'Y) ::; on [0, a] , this reduces the original problem to a I-dimensional problem. For every point So E (0, a) and c > there exists a C2 function cp (s ) defined on a subinterval (so - 8, So + 8) such that
°
°
f 0 'Y (so) = cp (so)
and
f 0 'Y (s) ::; cp (s )
for all s E (so - 8, So + 8) , and cp" (so) ::; c. Let
cp (s) ~ cp (so) + cp' (so) (s - so) + c (s - sO)2 . cp (so) = cp (so), cp' (so) = cp' (so), and cp" (s) == 2c > c 2: cp" (so).
Note that Hence there exists 81 E (0,8) such that
f 0 'Y (so) = cp (so)
f 0 'Y (s) < cp (s) {so}. We claim that f 0 'Y (s) ::; cp (s) for all and
for s E (so - 81, So + 8d s E [0, a]. By taking c ~ 0, the claim then implies
f 0 'Y (s) ::; f 0 'Y (so) + cp' (so) (s - so) for all so, s E [0, a]. We conclude f 0 'Y (s) is concave on [0, a]. Finally, suppose the claim is false; then there exists SI E [0, a] - {so} such that f 0 'Y (SI) = cp (SI)' Suppose S2 E (so, SI) is a minimum point of f 0 'Y (s) - cp (s) on [so, SI]. Then
f 0 'Y (s) 2: f 0'Y (S2) + [cpl (so) + 2c (S2 - so)] (s - S2) + c (s - S2)2 on [so, SI]. On the other hand, by our hypothesis on f 0 'Y, there exists a C 2 (7.143)
function CP2 (s) defined for s near S2 such that
f 0'Y (s) ::; f 0'Y (S2)
+ cp~ (S2)(S -
S2)
+ ~ (s -
S2)2 ,
7. THE REDUCED DISTANCE
372
which contradicts (7.143). This completes the proof of the claim and part (ii) . (iii) By hypothesis, for every q E B (p, r) there exists a G 2 function
(8~~~:j where G' <
ffj
~:%) (q) = (Hess g (
00
is independent of q E B (p, r). Since 1V'
Iffj I(q) ~ Gil, where Gil < 00 is independent of q, there exists a constant G3 independent of q such that the matrix (8~~~;;j -
at q. Choose 'ljJ (x) function for
- G3c5ij )
is negative definite
2 -
= -G3dg(p) (x,p) . Then
-
f + 'ljJ at
2 -
q, and since 8~i;tJ
= -2G3c5ij, we have
By repeating the proof of (ii), where we replace the geodesic, in (ii) by any
(f
straight line in Bp (r) , we see that 0 exp; 1 +(fi 0 exp; 1 ) (x) is a convex function of x E Bp (r). Finally, since - (J 0 exp; 1 +'ljJ) is a convex function on Bp (r) , we have - (J 0 exp;l +'ljJ) is locally Lipschitz on Bp (r), and hence f is locally Lipschitz on B (p, r) . D REMARK 7.123. Let f : B (p, r) ~ IR be a continuous function, where pENt and r < inj (p). Suppose there exist G1 , G2 < 00 such that for each q E B (p, r) there is a local upper barrier function
V'V' f (q) ~ D2 f (0) and the Laplacian I:l.f (q) ~ I:l.gf (q) ~ trlx=o (D2 f) (x). It is clear that (V'V' f) (X, Y) , I:l.f E Lfoc (B (p, r)) for any continuous vector fields X and Y on B (p, r). 17Here we have abused notation. When we write f (q), we treat f (x), we actually mean fox-I.
Nt, and when we write
f
ru;
a function on
9. WEAK SOLUTION FORMULATION
373
(ii) Let q E B (p, r) and let coordinates x be the same as in (i), where 'V 'V f (q) = D2f (0) exists. For any c > 0 we define the Coo function on TqM, _ 1 T C T ipq,e(X)=f(O)+Df(O).x+"2 x ''V'Vf(q),x+"2x ·x. It is clear that f (0) 0, and
= rpq,e (0) , f (x) ::; rpq,e (x) in a small neighborhood of
'V'Vrpq,e (q) = 'V 'V f (q) + cg (q) . Thus if 'V'Vf (q) ::; kqg (q), then Hes1:i supp (f) (q) ::; kq. (iii) Again assuming 'V 'V f (q) = D2 f (0) exists, for any c > 0, define 1 T c T
We have f (0) = 0 there exists a C 2 local upper barrier function ip : U -> lR such that 'V 'V ip (q) ::; (kq + 6) 9 (q) . Since
'V 'V f (q) - cg (q) Taking 6 -> 0 and c
->
= 'V'V
0, we conclude
'V 'V f (q) ::; kqg (q) , and tracing, we have b.f (q) ::; nkq. 9.4. The equivalence of notions of supersolution for nonsmooth functions. Note that (Ntn, is a complete Riemannian manifold. Recall the following definition.
g)
DEFINITION 7.124 (Weak-type notions of supersolution). Let f : Nt -> lR be a continuous function. (i) Suppose f satisfies Hesssupp (f) ::; C in the support sense. Then f is said to satisfy b.f ::; k in the support sense for some continuous function k if for every pENt there are a neighborhood U of p and a constant C I which have the following property. For any c > 0 and any q E U there are a constant r > 0 and C 2 function ip : B (q, r) -> lR such that ip (q) = f (q), f (x) ::; ip (x) for all x E B (q, r), I'Vip I (q) ::; C I , and
f}.ip (q) ::; k + c.
(ii) A continuous function f : Nt -> lR is said to satisfy b.f ::; k in the weak sense for some function k E Lfoc (Nt) if for any nonnegative C 2 function ip with compact support, we have
(7.144)
1M f b.ipdllg ::; 1M ipkdllg·
374
7.
THE REDUCED DISTANCE
(iii) A continuous function f : M ---t IR is said to satisfy b..f ~ k in the viscosity sense for some continuous function k if for every p E M and any C 2 function r.p : U ---t IR on some neighborhood of p satisfying r.p (p) = f (p), f (x) 2: r.p (x) for all x E U, we have b..r.p (p) ~ k. (iv) A continuous function f : M ---t IR is called a supersolution of b..f ~ k for some continuous function k if for every p E M, any r < inj (p) , and every C 2 function r.p on B (p, r) with b..r.p = k and r.plaB(p,r) = flaB(p,r) we have r.p ~ f on B (p, r) .
i
Now we can prove the following. LEMMA 7.125 (Equivalence of notions of supersolution). Let k : M ---t IR be a continuous function. (i) Let f : M ---t IR be a continuous function with Hess supp (f) ~ Cp < 00 on B (p, ~ inj (p)) for each p EM. Suppose that for each q E B (p, inj (p)) there is a local upper barrier function r.pq for f near
i
q satisfying 1'Vr.pql (q) ~
Cp < 00. We have for any r.p E C~ (M)
1M f b..r.pdp,g ~ 1M b..f . r.pdp,g.
(7.145)
In part~cular if b..f ~ k in the support sense, then f satisfies b..f ~ k in the weak sense. (ii) f satisfying b..f ~ k in the weak sense is equivalent to f satisfying b..f ~ k in the viscosity sense; they are both equivalent to f being a supersolution of b..f ~ k. PROOF. (i) Let 'IjJ be a C2 function on M. Then having b..f ~ k in the support sense is equivalent to b.. (f + 'IjJ) ~ k + b..'IjJ in the support sense, and b..f ~ k in the weak sense is equivalent to b.. (f + 'IjJ) ~ k + b..'IjJ in the weak sense. We use a partition of unity {cPoJ to rewrite r.p = r.pcPa, so that we only need to verify inequality (7.145) and (7.144) when r.p has small compact support, say in B (p, ~r) for some p E M, where r < min { 1,
L
injJp)}
as determined by Lemma 7.122(iii). From Lemma 7.122(iii) and by adding to f another concave C 2 function if necessary, we may assume that f (q) satisfies Hess supp (f) ~ -1 on B (p, r) and that f (x) is a concave function on Bp (r) C TpM in normal coordinates {xi} on B (p, r) . From Lemma 7.117(ii) and Lemma 7.119(ii), there are smooth functions fe (x) such that fe (x)
D2f(x) for x a.e. on Bp
---t
(~r) , (a~:!J~j (x)) ---t 7.119(iii), (a~:!J~j (x)) ~ 0 for
f (x) uniformly on Bp
(~r). By Lemma
all x E Bp (~r) . Hence for any 61 > 0 and any C 2-test function r.p supported in Bp (~r), there exist a sequence ck ---t 0+ and a set W 01 C Bp (~r) with meas (W01 )
~
61 such that (aJ;f (x))
---t
D f (x) and
(::i~:j
(x))
---t
9 WEAK SOLUTION FORMULATION
D2 f (X) uniformly on Bp (~r) \ W81. This in turn implies that b.gfek uniformly on Bp ar) \ W81. We compute using b.fek ::; 0,
375 ---t
Since meas (W81) ::;
b.gf
---t
0+,
°1
fA
1M
fb.
f
1
1Bp(2r)
b.f·
= fA b.f·
If b.f ::; k in the support sense, then by Remark 7.123(iii), f satisfies b.f ::; k a.e. on M. Combining this with (7.145), we get
1M f b. o. Actually, the equivalence of the notions of weak and viscosity supersolutions 0 was first proved by Ishii and Lions [214]. 9.5. Comparison theory for Riemannian distance d and reduced distance f. As a simple application of the results in the previous subsection we consider the distance function d on a Riemannian manifold and the reduced distance f of a solution to the backward Ricci flow.
376
THE REDUCED DISTANCE
7.
9.5.1. Weak differentiability of the distance function d on a Riemannian manifold. Let be a Riemannian manifold with sect (9) 2: -K. Let
(Mn,g)
p E M and define dp (x) ~ d (p, x) . Since lV'dp (x)1 = 1 for a.e. x EM and since the Hessian and Laplacian comparison theorems (A.9) hold pointwise a.e., dp satisfies
~dp ::; (n - 1) JK coth ( JK dp) in the support sense from
Definition 7.124(i). Applying Lemma 7.125 in the previous subsection to dp (x) , we obtain the following (see for example, [316]' [246] or Theorem 1.128 in [111]). LEMMA 7.126 (Laplacian comparison). Let
(M, g)
be a complete Rze-
mannian manifold with sect (g) 2: - K for some K 2: O. (i) If K > 0, then
~dp ::; (n - 1) JK coth ( JKdp) in the weak sense.
(ii) If K = 0, then ~dp ::; (iii) (iv)
nil
m the weak sense.
p
~dp E Leoc (M) . iM dp~r.pdJ.Lg ::; iM ~dp . r.pdJ.Lg for any r.p
E
C; (M) .
REMARK 7.127. It is not clear to us whether ~dp E L Ioc for r > 1 when dimension n 2: 3. When n = 2, near the point p, we have ~dp f/:. L Ioc for r 2: 2. Below we give an example to show that in general,
for r.p E
C; (M) .Hence d
p
does not belong to the Sobolev space W 2 ,1
(M) .
be the flat torus and choose p = O. We consider = 2n outside of the cut locus, For r.p = 1 we have which is Cut (p) = 8 ([-1, 1t)/ Let Tn ~ [-1, 1t /
r'V
d~ to facilitate our discussion. We have ~d~ r'V
•
r ~d~. r.pdJ.L = 2 + n i= 0 = lTnr d~. ~r.pdJ.L. lTn n l
The underlying reason for this phenomenon is the nonsmoothness of d~ on the cut locus of p. Let 7;;n ~ [-1 + c:, 1 for c: > 0 (which has piecewise linear boundary). Let II denote the outward normal of 87;;n c Tn. Then
c:t
9
WEAK SOLUTION FORMULATION
Note that lime->o+ J8T.n d~ ("\I
r
("\Id; . 1/)
} 8T.n
r
377
= 0 (since n pairs of opposing bound-
dp ("\Id p . 1/)
} 8T.n
r
} 8T.n
since dp "\I dp = (1 - c) 1/. Hence
t
r
r
("\Id; . 1/) O+ } 8T.T! . JTn-l e 1=1
(xl, ... , xi-I, 1. xi +!, ... , xn) da.
We conclude that
2n
r
This reflects the fact that, in the sense of weak derivatives, I::.d~ ::; 2n and the distribution [I::.d~] has its singular part supported on the cut locus.
9.5.2. Weak differentiability of the reduced d~stance functzon £. In this subsection (Mn, 9 (T)) , T E [0, T] , is a solution to the backward Ricci flow with the curvature bound max {IRcl, IRml} ::; Co < 00 on M x [0, T]. Let p E M and let £ (x, T) be the f-distance function defined using basepoint (p, 0) . From the gradient estimate of £ in space-time and from the Hessian and Laplacian pointwise estimates of f in space (Lemma 7.61 and (7.90)), £ (', T) satisfies (7.90) in the support sense from Definition 7.124(i). Lemma 7.125 in the previous subsection applies to f (x, T) as a function of the space variable. We have the following. PROPOSITION 7.128 (Regularity properties of the reduced distance). For every T E (0, T), (i) f (q, T) is locally Lipschitz in the varwble q, (ii) 1::.£ (q, T) E Lfoc (M), (iii) for any smooth nonnegatzve function
1M £ (., T) I::.
w~th
respect to 9 (T) .
Note that (i) gives an abstract proof that f(., T) is locally Lipschitz, which is different from the effective proof given earlier. From Lemma 7.113, we have
1M f (-, T) I::.
7. THE REDUCED DISTANCE
378
(i) The inequality 8£ 2 n 87 -~£+IV'£I -R+ 27 ~O
holds in the weak sense on M x (0, T) , i.e., (7.146)
1: 1M 2
[V'£' V'
(~! + 1V'£12 -
R + ;:.)
~0
for any nonnegative C2 function
£-n
2~£ -1V'£12 + R+ - - ~ 0 7
holds in the weak sense on M x {7} , 2. e., (7.147)
1M [-2V'£' V'
Let 0 be any nonnegative locally Lipschitz function which satisfies the decay conditions:
o(q, 7), 1V'0 (q, 7)1 ~ ~ec -cd~(o)(p,q) for some constant c > O. We now show that (7.146) and (7.147) hold for
(p, r) ~ Cecr
for some constant C > O. By Lemmas 7.59 and 7.60 we know that both integrals in (7.146) and (7.147) are finite. Let {4>o,} be a partition of unity on M with compact support. Because of the finiteness of the integrals it suffices to show that for each a the two inequalities hold for 0:0, which is Lipschitz and has compact support in space. Since 4>0:0 can be approximated by smooth functions along with its first derivative outside a set of arbitrary small measure, by a proof similar to that of Lemma 7.113 we conclude that for any a, (7.146) and (7.147) hold for
0:0. Hence we have proved that (7.146) and (7.147) hold for any nonnegative locally Lipschitz function which satisfies the decay conditions 0 (q, 7), 1V' 0 (q, 7) 1 ~ c- 1e -cd~(o) (p,q) for some constant c > O. Note by Lemmas 7.59 and 7.60 that we can choose
o
10.
NOTES AND COMMENTARY
379
LEMMA 7.130 (Integration by parts inequality for i). We have (7.148)
1M (boe -l ve I
2)
e-fdp, 2 O.
PROOF. Since integration by parts holds for Lipschitz functions, we can write Proposition 7.128(iii) as (7.149) where O. Taking
1M IViI e- fdp,:::; 1M e-fboidp,. 2
o ~
As a simple consequence of the above lemma and g~ 20 a.e., we obtain the following.
- boi + 1Vi 12 -
R
+
COROLLARY 7.131. For a solution to the backward Riccifiow (M n , 9 (7)), 7 E
[0, T] , with bounded sectional curvature, we have
(7.150)
(7.151) 10. Notes and commentary 1. This chapter is a discussion of §7.1 and §7.2 of Perelman's [297]. 2. The reader may also consult Rugang Ye's notes on the i-function [382], which we have partially used as a source. See also the appendix of [134] for a brief discussion of reduced distance. 3. Notational conventions. In this chapter we have endeavored to maintain a consistent convention for the notation we have used. In particular we have used the following notation when discussing solutions to the forward and backward Ricci flow: 9): a static Riemannian manifold,
(Mn,
(Nn, h (t) ): an arbitrary (not necessarily complete or with bounded curvature) solution to the Ricci flow, (Nn,h(7)) , 7 E (A,n): an arbitrary solution to the backward Ricci flow,
380
7. THE REDUCED DISTANCE
(Mn,g(T)): a solution to the backward Ricci flow, usually defined for T E [0, T], complete, and satisfying the pointwise curvature bound max {IRmI , IRcl} ::; Co < 00 on M x [0, T].
CHAPTER 8
Applications of the Reduced Distance Now...
the basic principle of modern mathematics is to achieve a complete fusion [of]
'geometric' and 'analytic' ideas - Jean Dieudonne
In this chapter we give some geometric applications of the reduced distance for Ricci flow. We give two proofs of the monotonicity of the reduced volume. This result is the Ricci flow analogue of the Bishop-Gromov volume comparison theorem in Riemannian geometry. A beautiful and striking aspect of this monotonicity formula is that unlike Hamilton's matrix Harnack inequality and most other monotonicity formulas, no curvature assumption is needed. Using the reduced volume monotonicity, we prove a weakened no local collapsing theorem and we also prove that certain backward limits of ancient ",-solutions are gradient shrinkers. All of these results are due to Perelman. Throughout this chapter we shall use to denote a complete
(Mn, [})
Riemannian manifold and (Mn, 9 (T)) to denote a solution of the backward Ricci flow. 1. Reduced volume of a static metric
We begin by defining the reduced volume of a static metric since this is technically easier than the Ricci flow case yet it still exhibits many of the ideas. 1.1. The reduced volume for a static metric and its monotonicity when Rc ~ O. Consider the following functional for a complete Riemannian manifold Given a point p EM, define the static reduced volume by
(Mn, [}) .
(8.1)
if ([}, T)
'* 1M (47rT)-n/2 e- d(x,p)2/ 4T dJ.l (x).
This geometric invariant depends on [}, T and p. Clearly if is positive. In general, we can think of this as the integral of the Euclidean heat kernel transplanted to a manifold via the exponential map. If the Ricci curvatures of are bounded below by a constant, then by the Bishop-Gromov volume comparison theorem, which gives an upper bound for the volumes of balls, the integral defining if ([}, T) converges for all T > 0 even when M is noncompact.
(M, [})
381
8.
382
APPLICATIONS OF THE REDUCED DISTANCE
EXERCISE 8.1. Show that if (Mn'9) is a complete noncompact Riemannian manifold with RC g ~ - K for some K E ~, then the integral defining if (9, T) converges for all T > 0. In Euclidean space if is the integral of the heat kernel, which is the constant 1. Note also that for any (Mn'9) and p EM, lim if (9, T) = 1
(8.2)
T~O
essentially since manifolds are locally Euclidean. EXERCISE 8.2. Prove (8.2). Let
u (x, T) ~ (41l'T)-n/2 e- d (x,p)2/ 4T , which is a Lipschitz function, and let d (x) ~ d (x, p). We can think of
if (9, T) =
1M u (x, T) dJ.L (x) as a weighted volume centered at p with the
radial weight function u. As T ---+ 0, the weight u concentrates at p and as u diffuses throughout M.
T ---+ 00,
REMARK 8.3. If M is closed, then the upper bound
if (9, T) ~
1M (41l'T)-n/2 dJ.L (x) = (41l'T)-n/2 Vol (9)
implies that limT~oo if (9, T)
= 0.
9)
Now assume that (Mn, is complete with nonnegative Ricci curvature. Since RC g ~ 0, the Bishop-Gromov volume comparison theorem says that the volume ratio r- n Vol B (p, r) is a nonincreasing function of r. It is thus natural to expect that if (9, T) is a nonincreasing function of T since as T increases, the weighting favors larger radii. Indeed we have LEMMA 8.4 (Static reduced volume monotonicity). If (Mn'9) is complete with RC g ~ 0, then
d~ if (9,r) = 1M (:T - ~) udJ.L ~ 0.
(8.3)
In particular, by (8.2),
if (9, T)
~ 1
for all T
> 0.
REMARK 8.5. Clearly this lemma implies limT~oo V (9, T) E [0,1] exists. PROOF. We compute that u is a subsolution, in the weak sense, to the heat equation:
( ~ _ ~) u = u (_~ + ~ + (d~d + IV'dI2) _ ~ IV'dI2) ~
(8.4)
~
~
0,
~
~
~
1. REDUCED VOLUME OF A STATIC METRIC
383
where we used I\i'dl = 1 a.e. and the Laplacian comparison theorem, i.e., dfld::; n - 1. Note that the Laplacian comparison theorem is equivalent to the Bishop-Gromov volume comparison theorem. 0 It is useful to keep the following simple examples in mind when pondering Lipschitz continuous sub- and super-solutions of the heat equation.
8.6 (Heat equation on 8 1 ). Let Ml the standard metric 9 = d0 2 • Consider the function EXAMPLE
f (0, t)
02 t + 2'
=
= 81 =
0 E (-71",71"] and t
~/
(271"Z) with
E R
For each fixed 0, f (0, .) is a smooth (linear) function of time and, for each fixed t, we have f (., t) is Lipschitz on 8 1 and Coo except at 0 = 71". Moreover f is a solution to the heat equation almost everywhere. In particular,
8f 8 2f 8t (0 , t) = 80 2 (0, t) = 1 for all 0 E 8 1
-
{71"} and t
E R On the other hand,
f f(0,t)dO=271">O. ls!
ddt
This is consistent with the fact that f is not a subsolution of the heat equation in the weak sense (as we shall now see, f is a supersolution). Note that
8f 80 (0, t) = 0
for all 0 E (-71",71") and t E ~ (and ~ is undefined for 0 = 71"). In particular, for each t E ~, ~ (., t) has a jump discontinuity at 0 = 71". In the sense of distributions, we have ~ (0, t) = 0 and
82 f
(8.5)
80 2
(.,
t)
= 1 - 271" ·6n ,
where 6n is the Dirac 6-function centered at 0 distributions,
8f
82 f
= 71".
Hence, in the sense of
82 f
at = 802 + 271" . 6n ~ 802.
EXERCISE
8.7. Prove (8.5).
SOLUTION TO EXERCISE
f ls!
8.7. For any C 2 function cp: 8 1
n 02 cp" (0) dO = 02cp' (O)l 2 2 -n
-
f Ocp' (0) dO ls!
f cp(O)dO- Ocp(O)I:n ls! = f cp(0)dO-271"cp(7I"). ls! =
---+ ~
we have
384
8.
APPLICATIONS OF THE REDUCED DISTANCE
That is,
fJ2
80 2
((P) 2
= 1 - 27r . b7r •
The square torus is abo a nice concrete example for which we can compute the static reduced volume explicitly. EXAMPLE 8.8 (Static reduced volume for square torus). Consider the torus M n ~ lRn / (2Zr with the standard fiat metric [} = dxy + ... + dx~. A fundamental domain for the covering lR n -> M is D ~ (-1, l]n. Let p = 0 be the origin so that d (x, p) = Ixl for x E (-1, 1]n = M. The static reduced volume is!
(8.6) using the change of variables a decre1lliing function of T.
x=
2ft. From (8.6) it is clear that V ([}, T) is
1.2. Static reduced volume and volume ratios. We may think of Vas the static manifold analogue of Perelman's reduced volume for the Ricci fiow (see (8.16) defined later in this chapter), which is defined similarly with d2/4T replaced by the reduced distance function .e. The monotonicity formula (8.3) is analogous to Perelman's monotonicity of the reduced volume (8.28). In the Ricci fiat case, V is the same as Perelman's reduced volume (see Exercise 7.12 and (8.16)). Intuitively the static reduced volume V says something about volume ratios r- n Vol B (p, r) at scales r rv ft. Motivated by these elementary yet a posteriori considerations, we now relate V to the volume ratio r- n Vol B (p, r) under the assumption that the Ricci curvatures of [} are nonnegative. be a complete Riemannian manifold with RC g > O. We Let
(Mn, [})
divide the integral (8.7)
V into two parts:
V ([}, T) = f
udJ1 +
JB(p,r)
f.
udJ1.
JM-B(p,r)
For the first term on the RHS, just using the obvious fact that e- r2 /47 ~ 1, we have
f
udJ1
~
(47rT)-n/2 Vol B (p, r) .
JB(p,r)
lSince the torus is (Ricci) flat. the static metric reduced volume is the same Ricci flow reduced volume of Perelman.
88
the
REDUCED VOLUME OF A STATIC METRIC
1
385
Let A (8) denote the volume of the geodesic (n - l)-sphere of radius 8 centered at p. Since RCg 2: 0, we may apply the Bishop-Gromov volume comparison theorem (see (A.8)), which says that for 8 2: r,
8n-l VolB(p,r) n-l A() A() 8::; r n-l::; n n 8 , r r
(8.8)
to estimate the second term on the RHS of (8.7):
{.
JM-B(p.r)
udp, =
/.00 (47fT)-n/2 e- s2 / 4r A (8) d8 . ,.
< n Vol B (p, r) rn =
100 r
(4 7fT )-n/2 e -s2/4r 8 n- 1d8
n -n/2 Vol B (p, r)
-27f
rn
100 r2
e
-'1/
rJ
n-2 2
d rJ,
4r
where we made the change of variables rJ ~ ~~. Hence we have ) (4 )_n/2 VolB (p,r) V-(-g, T::; 7f Tn /2
-_
(8.9)
+ -2n 7f _n/2
4r
Vol B (p, r) (( r2 )n/2
-47fT
rn
J~n
n -n/21°O -'1/ n-2 d ) + -7f e rJ 2 rJ 2 r2
.
4r
This tells us that a lower bound for ratios of balls. Note that
1= {
1
VolB (p,r) 00 -'1/ n-2 d rn 2. e rJ 2 rJ
7f-n/2e-lxI2 dx
V yields
a lower bound for the volume
= nW n7f- n/2
2
roo e-'1/rJ n~2 drJ,
Jo
where Wn is the volume of the unit Euclidean n-ball. Hence LEMMA 8.9 (The static reduced volume is bounded by volume ratios). If
(Mn, g)
is complete w1.th RCg 2: 0, then for all r
V(g,T)::; VolB(p,r) rn
(8.10)
>0
and
T
> 0,
((~)n/2 +~). 47fT
wn
Thus, for r ::; p, the static reduced volume controls the volume ratio in the sense that
VolB(p,r) -IV- (- 2) n 2: Cn g, P , r
(8.11) where have (8.12)
Cn
~
(20r)
n
+ WIn'
Note also that if for some p E Vol B (p, r) n r
::; K,
M and r > 0 we
386
8.
APPLICATIONS
OF
THE REDUCED DISTANCE
then
-( 2) 0< V [}, K,a r
I_an
:::;
K,
2
(47rt Hence if a < 2/n, then assumption (8.12), as
/2 K,
+-. K,
Wn ---t 0, implies
V ([}, K,a r 2)
---t
O. 1.3. The noncompact case and the asymptotic volume ratio. Consider the case where is complete and noncompact with RC g ~ O. It is natural to believe that the limit as 7 ---t 00 of the static reduced volume V ([}, 7) is related to the limit as r ---t 00 of the volume ratios; as we now show, this is indeed the case. Inequality (8.10) implies
(Mn, [})
V ([},7)
lim
:::; inf VoIB(p,r) r>O
T--+oo
where
wnrn
= AVR([}) ,
AVR(9)~ lim VoIB(p,r)
= lim A(s) wnrn s-->oo nwns n- l as in (6.80). Next we show the opposite inequality. Since, by (8.8), A (r) ~ nWn AVR ([}) r n- 1, r--+oo
we have for all
V ([}, 7) =
7
> 0,
1
00
(47r7)-n/2 e- s2 / 4T A (s) ds
~ nWn AVR ([})
1
00
(47r7)-n/2 e- s2 / 4T sn-1ds
= AVR ([}).
Therefore the limit, as 7 tends to infinity, of the static reduced volume is the asymptotic volume ratio.
(Mn,[})
LEMMA 8.10 (Asymptotic limit of V is AVR). If noncompact Riemannian manifold with RC g ~ 0, then lim
T--+oo
is a complete
V (9, 7) = AVR ([}).
REMARK 8.11. When n = 2, (8.9) says for any r > 0 (8.13)
V- (~g,7 ) :::; Area B 2(p, r) (r2 -4 7rr 7
+ e _ 4-.r2 )
•
Note that the function F (x) ~ x + e- x , x ~ 0, is an increasing function and its minimum value of 1 is attained at x = O. In particular, as in (8.10) the upper bound in (8.13) improves as 7 increases (for fixed r > 0).
2. Reduced volume for Ricci flow In this section (Mn, 9 (7)) , 7 E [0, Tj , will be a complete solution to the backward Ricci flow satisfying the curvature bound IRm (x, 7)1 :::; Co < 00 for (X,7) EM x [O,Tj.
2. REDUCED VOLUME FOR RICCI FLOW
387
2.1. Volumes of geodesic spheres in M. We motivate the definition of the reduced volume by computing the volume of geodesic spheres in the potentially infinite-dimensional manifold introduced in subsection 2.1 of Chapter 7. In particular, let P = (xo, Yo, 0) , T E (0, T) , and
(M, 9)
Bg (p, v'2NT) C M ~ M
X
SN
X
(0, T)
denote the ball centered at P with radius v'2NT with respect to the metric:
9 ~ 9ijdxi dx j + T903d yO
v'2NT = dg (w,p) = dg ((x, y, Tw), (xo, Yo, 0)) = dg ((x, y, Tw), (xo, y, 0)). Hence, letting 'Y (T) ~ bM (T), y, T), T E [0, Tw], with 'Y (0) = (xo, y, 0) and 'YM (Tw) = w, we have
v'2NT = infLength g ('Y) l'
=inf ( 'YM
( ) 8.14
k J;w vIr (R+i'YM(T)1 2)dT) +v'2NTw + 0 (N- 3 / 2)
1 -_ V ~ 2NTw + v'2NL(x, Tw)
+0
( N -3/2) ,
where
L(x, Tw)
~ ~~ foTW vir (R + i'YM (T)12) dT
and the infimum is taken over 'YM : [0, Twl -+ M with 'YM (0) = Xo and 'YM (Tw) = x. Therefore for any w = (x, y, Tw) E oBg(p, v'2NT),
Fw = Vf -
1
2NL(x,Tw)
+0
(N- 2).
This implies that the geodesic sphere oBg (p, v'2NT) , with respect to 9, is O(N-l)-close to the hypersurface M x SN EXERCISE
X
{T}.
.
8.12. Justify the equality (8.14).
Note that since the fibers SN pinch to a point as T -+ 0, if w = (X,y,Tw) E oBg(p,v'2NT), then any point in {x} x SN X {Tw} also lies on the sphere oBg (p, v'2NT) . We have that the volume of oBg (p, v'2NT)
388
8 APPLICATIONS OF THE REDUCED DISTANCE
is roughly (since the sphere has small curvature for N large) the volume of the hypersurface M x SN x {f} in M and its volume can be computed as Volg BBg
(p, J2Nf)
~ f
JaB g(p,v'2Nr)
dJ-L9M(Tw ) (x) /\ T;;:/2dJ-LSN (y)
~Vol(SN,9SN) 1M (..Jf- 2~L(X'Tw)+O(N-2))N dJ-LgM(r) ~ wN ( J2Nf) N 1M (1 -2;.;fL(x, f) + O(N- 2)) N dJ-LgM(r) , where WN is the volume of the unit sphere SN (recall that 9sN has constant sectional curvature 1/ (2N) , i.e., radius J2N). Observing that lim N--.oo
(1 _ 1 2N y
£L(x, f) T
+ O(N- 2 )) N
. ( 1 - -1 -L(x 1 _) = hm T) N
N--.oo
2V¥
'
N
r) -e(-) = e--l-L(x 2../¥ ' = e X,T
,
one can prove (8.15)
Volg (BBg
(p, v'2Nf) )
( J2Nf)N+n
= (2N)-n/2 WN
(1M f-n/2e-e(x,r)dJ-LgM(r) +O(N-
1 )).
In particular, we obtain the geometric invariant
f f- n/ 2e- e(x,r)dJ-L gM(T)_ JM for f E (0, T) . EXERCISE 8.13. Make the above arguments rigorous (especially the approximations) and in particular prove (8.15).
2.2. Definition of Perelman's reduced volume. Thus we are led to the following. DEFINITION 8.14 (Reduced volume for Ricci flow). Let (Mn, 9 (T)) , T E [0, T], be a complete solution to the backward Ricci flow with bounded curvature. The reduced volume functional is defined by
V (T) ~ 1M (47fT)-n/2 exp [-R (q, T)] dJ-Lg(T) (q)
(8.16) for
T
E (0, T)
.
2. REDUCED VOLUME FOR RICCI FLOW
389
See Lemma 8.16(ii) below for why V (r) is well defined even when M is noncompact. In the case of a Ricci flat solution 9 (r) == g, we have
L
(41rr)-n/2 exp { _ d (:~q)2} dl'g.
if (r) =
(Compare with (8.1).) We now give heuristic (e.g., unjustified) proofs of the reduced volume monotonicity. Provided we can differentiate (8.16) under the integral sign, we obtain dV (r) dr
(8.17)
~ ({ (47rr)-n/2 e-£(-,T)df..l (T)) dr 1M 9
=
~
1M! ((47rr)-n/2 e-£(·,T)df..lg(T))
=
1M
{
(-~ 2r
a£ ar
+ R)
(47rr)-n/2 e-£df..l,
where ~ denotes an unjustified inequality.2 By applying inequalities (7.91) and (7.148), which results in (7.150), we obtain
~~ (r) ::; 1M (Iv£1
2 -
.6.£) (47rr)-n/2 e-£df..l::; O.
Recall that for any vector field X, we have £xdf..l = div (X) df..l. In particular, £\7 hdf..l = .6.h df..l for any C 2 function h. Define the first-order differential operator acting on time-dependent tensors and forms:
D
a
dr ~ ar +£x.
EXERCISE 8.15. Show that for any C 1-vector field X and C 1 function >, one of them with compact support, under the backward Ricci flow (Mn,g(r)), we have :r
1M >df..l = 1M ~ (>df..l) =
1M (:r > + X . v> + >R + >div (X)) df..l.
SOLUTION TO EXERCISE 8.15. The result follows from tTdf..l and the integration by parts identity
= Rdf..l
1M (X· v> + >div (X)) df..l = O. Now we consider again the time-derivative of the reduced volume of (Mn, 9 (r)) under the backward Ricci flow and we heuristically discuss some 2For the justification, see the proof of Theorem 8.20 below.
390
8. APPLICATIONS OF THE REDUCED DISTANCE
formulas to be considered rigorously later. Let X = 'V l. Again, provided we can differentiate under the integral sign, we compute
dV -Q dr =
j
j
(-~ 2r
M
~
dD r
M
((4trr )-n/2 e-i(.,T)dJ.Lg(T) )
- 8r + R- l'Vll + ~l) (47rr)-n/2 e-idJ.L 2
{)£
0,
where we used (7.91) to obtain the last inequality. Note that we actually have the pointwise inequality
~
((47rr)-n/2 e- i dJ.L)
= ( - 2: - ;:
R
+ -1'VlI 2 + ~l)
(47rr)-n/2 e-idJ.L
~ 0.
Pulling this back to the tangent space TpM, we have
!
[(47rr)-n/2 e -i(W(T),T).L:Jv (r)]
~
°
as in (8.22) below. 2.3. Monotonicity of reduced volume: A proof using the .L:Jacobian. In this subsection we give a rigorous proof of the reduced volume monotonicity. Recall that the open set
O(r) = O(p,O)(r)
~
{V
E
TpM : 'TV> r}
C
TpM
is given by Definition 7.96(iii) and satisfies O(p,O) (r2) c O(p,O) (r1) if r1 Recall from (7.136) that the .L:-exponential map restricted to O(r), .L:Texp: O(p,O) (r)
--+
< r2.
M\.L:Cut(p,o)(r)
is a diffeomorphism. If O(p,O) (r) = TpM for some r > 0, then .L: Cut(p,O) (r) = o and M n is diffeomorphic to Euclidean space. Since .L:Cut(p,O)(r) has measure zero in (M,g(r)), we have by the definition of the .L:-Jacobian and (7.136),
V (r) =
j
(47rr)-n/2 exp [-l (q, r)] dJ.Lg(T) (q)
M\C CutCI',O) (T)
(8.18)
=
r
(47rr)-n/2 e -i(')'v(T),T).L:Jv(r)dx(V),
JnCp,Q)M where 'Yv is the .L:-geodesic emanating from p with limT--+o+ ..Jiiv (r) = V and .L: J v (r) is the .L:-J acobian associated to the .L:-geodesic 'Yv. Here dx is the volume form on TpM with respect to the Euclidean metric g(O,p), .L:T exp(V) = 'Yv (r) , and .L: J v (r) dx (V)
= (.L:T exp)* dJ.Lg(c7' exp(V),T)'
2. REDUCED VOLUME FOR RICCI FLOW
391
We will use the convention CJv (7) ~ 0
for 7 ~ TV.
We can then write the reduced volume as
if (7) =
(8.19)
r
(41IT)-n/2 e-f(-Yv(T),T) CJV (7) dx (V).
iTpM
We compute the evolution of f along a minimal C-geodesic 'YV(7) for 7 < TV, where V E TpM. For q = 'YV(7), 7 E [0, TV), the function f(·,·) is smooth in some small neighborhood of (q,7); hence the following derivatives of e at such (q,7) exist. Recall from (7.78) that
o ::;
73/ 2(R + IXI2) (7) = -K + ~L (q,7), where K = K (7) is the trace Harnack integral defined by (7.75). Thus (8.20) Recall equation (7.88):
af
1
K -a7 = -S/2 27
1 - -f+R, 7
and from (7.54) recall that
\If(q,7) = 'Yv (7) = X (7). Hence the derivative of the reduced distance along a minimal C-geodesic is given by af d d7 [fhv(7),7)] = a7 +\If·X 1 f 2 =-K--+R+IXI 3 2 27 /
7
= _~7-3/2K
(8.21)
2
by (8.20). The following lemma can be viewed as an infinitesimal Bishop-Gromov volume comparison result for the Ricci flow geometry. The striking part is that no curvature assumption is needed. 8.16 (Pointwise monotonicity along C-geodesics). Suppose that (Mn, 9 (7)), 7 E [0, T], is a complete solution to the backward Ricci flow with bounded curvature. (i) For any V E TpM and 0 < 7 < TV, LEMMA
(8.22)
d~
[(47r7)-n/2 e-f('YV(T),T) CJ v (7)] ::; 0,
where equality holds if equality in (7.139) holds.
2. REDUCED VOLUME FOR RICCI FLOW
391
We will use the convention .c J v (T)
~
0 for T ;::: TV.
We can then write the reduced volume as
V(T)
(8.19)
= (
(47rT)-n/2 e -£(W(T),T).cJV(T)dx(V).
JTpM
We compute the evolution of f along a minimal .c-geodesic 'IV (T) for T < TV, where V E TpM. For q = 1'V(T), T E [O,TV), the function f(·,·) is smooth in some small neighborhood of (q, T); hence the following derivatives of f at such (q, T) exist. Recall from (7.78) that
o~
T3 / 2 (R + IXI2) (T) where K
= -K + ~L (q, T),
= K (T) is the trace Harnack integral defined by
(7.75). Thus
(8.20) Recall equation (7.88):
and from (7.54) recall that
'Vf (q, T)
= 'Yv (T) = X (T).
Hence the derivative of the reduced distance along a minimal .c-geodesic is given by d af dT [fhv(T),T)] = aT +'Vf·X 1 f 2 =-K--+R+IXI 2T 3/2
T
= _!T- 3 / 2 K 2
(8.21)
by (8.20). The following lemma can be viewed as an infinitesimal Bishop-Gromov volume comparison result for the Ricci flow geometry. The striking part is that no curvature assumption is needed. 8.16 (Pointwise monotonicity along .c-geodesics). Suppose that (Mn,9 (T)), T E [0, T], is a complete solution to the backward Ricci flow with bounded curvature. (i) For any V E TpM and 0 < T < TV, LEMMA
(8.22)
d~
[(47rT)-n/2 e-£(')'V(T),T).cJV (T)]
~ 0,
where equality holds if equality in (7.139) holds.
8. APPLICATIONS OF THE REDUCED DISTANCE
392
(ii) For any V
E TpM and 0
< 7 < T,
(47l'7)-n/2 e-l('YV(1').1') I:- Jv (7) ::; 7l'-n/2e-IVI~(o,p).
(8.23)
Hence, even for a complete solution on a noncompact manifold, the reduced volume is well defined. PROOF.
(i) Recall from (7.139) that
(d~ logl:-Jv )
(7) ::;
2: - ~7-~K.
From this and (8.21), we compute
d~
[(47l'7)-n/2 e-l(-rv(1'),1') I:-Jv (7)]
= (47l'7)-n/2 e-l(-rv(1'),1') I:- JV (7)
(-!!...27 -
df d7
+ ~ log I:- J V ) d7
::; O.
(ii) It follows from (8.22) that for any 0 <
7
< Tv, we have
(47l'7)-n/2 e-lhv(1').1') I:- Jv (7)
::; lim (47l'71)-n/2e-l(-rvh),1'J)I:-JV(7I) 1'1 .....0+
= lim 7'} .....
(8.24)
=
[(47l'71)-n/2 I:-Jv
0+
(71)] e-limT1--+o+l(iv(1'1),1'1)
7l'-n/2 e -1V1 2 ,
where in the last equality we have used (7.138) and (7.97). If 7 the statement is obvious.
~ Tv,
then 0
An immediate consequence of the above lemma is the following fundamental result: the monotonicity of the reduced volume. 8.17 (Reduced volume monotonicity). Suppose (Mn, g(7)), 7 E [0, T], is a complete solution to the backward Ricci flow with the curvature bound IRm (x, 7)1 ::; Co < 00 for (x, 7) EM x [0, T]. Then COROLLARY
(i) lim1'-->o+ V (7) = 1. (ii) The reduced volume is nonmcreasmg:
V (7I) ~ V (72)
(8.25)
for any 0 < 71 < 72 < T, and V(7)::; 1 for any 7 E (O,T). (iii) EquaZzty m (8.25) holds if and only zf (M,g (7)) is isometric to Euclidean space (l~n, 9lE), regarded as the Gaussian soliton. PROOF.
(i) From equation (7.131) it follows that lim
1'-->0
n(pO)
'
(7) = TpMn.
2. REDUCED VOLUME FOR RlCCI FLOW
393
Since 0Cp,O) (Tl) = 0 (Tl) ::J 0 (T2) for Tl < T2, we have limT-+o+ XnCT) = 1, where Xn denotes the characteristic function of the set O. We compute lim T-+O+
V (T) =
lim ( (47rT)-n/2 e-e(rVCT),T) C Jv (T) dx (V) JnCT)
T-+O+
= (
lim [xn(T) (47rT)-n/2 e-e C'YVCT),T) C Jv (T)] dx (V)
JTpM T-+O+
= (
1 . 7r-n/2e-1V12 dx (V)
= 1,
JTpM
where we used (8.24).
(ii) From (8.22), we have for any 0 < Tl < T2 and V
E
0(T2),
T;n/2e-eC'YVCTl),TI)CJv (Td 2: T;n/2e-eC'YVCT2),T2)CJv (T2) , so
V (T2) = (
(47rT2)-n/2e-lC'YV(T2),T2) C Jv (T2) dx (V)
JnCT2) :::; ( (47rTl)-n/2e-lC'YVCT1),Tl) C Jv (Tl) dx (V) Jnb)
:::; (
(47rTl)-n/2e-l('YV(TI),Tl) CJv (Tl) dx (V)
In(Tl)
= V (Tl) , where we used 0 (Tl) ::J 0 (T2)' Note that from (8.25) and (i).
V (T) :::;
1 for any T > 0 follows
(iii) We prove this statement in two steps by first showing that g(T) is a shrinking gradient Ricci soliton and then showing that (M, 9 (T)) is Euclidean space. If V (Td = V (T2) for a pair of times 0 < Tl < T2, then for T E (Tl' T2) and V E O(T), we have that equality in (8.22) holds:
d~ [(47rT)-n/2 e-£C'YVCT),T) CJv (T)]
= 0,
which, by the proof of Lemma 8.16, implies that we have equality in (7.139). Hence, by (7.140), we get (8.26)
Rc (-Tv (T), T)
+ (Hesse) (TV(T), T) = 9 (-Tv (T), T) 2T
for all V E O(T) and T E [T1,T2J, and where e is Coo at (TV(T),T) for all such (V, T). Since V (Tl) = V (T2) , we have 0 (Tl) = 0 (T2)' Suppose there exists VI E TpM such that Tvl :::; Tl. Since 0 (Tl) -# 0, there exists V2 E TpM such that Tv2 > Tl. Since the function V I-t Tv is a continuous function, there exists V3 E TpM such that TV3 E (Tl' T2) . Thus V3 E 0 (Tl) - 0 (T2) , which is a contradiction. Therefore, for every V E TpM, we have Tv> Tl, so that 0 (Tl) = TpM (and hence 0 (T) = TpM
394
8.
APPLICATIONS OF THE REDUCED DISTANCE
for all 7 E [TI,72])' This implies f. is Coo on M x [TI,72] and (8.26) holds on M x [71,72], Thus 9 (7) is a shrinking gradient Ricci soliton. Given 70 E [TI,72] , by Proposition 1.7, the shrinking gradient Ricci soli9 (70) , \l f. (70) , - ~) may be put into a canonical timeton structure dependent form (1.11) defined for all t < 70,
(M,
g(t) = f (t) r.p(t)*g (70) , 70 where g(t) is a solution of the Ricci flow, by (1.10), i.e., f (t) = 70 - t (c = - ';0)' and r.p(t) is a I-parameter family of diffeomorphisms with r.p (0) = idM . By the uniqueness of complete solutions of the Ricci flow with bounded curvature (see Chen and Zhu [82]) and since g(O) = 9 (70) , we have 9 (7) =
9 (70 - 7),
so that (8.27) Since IRm [g (7)]1 :::; Co < (8.27) we have
00
for 7 E [0, T] (we just use this for 7 small), by
sup IRm [g (70)]1 = sup IRm [r.p(70 - 7)*g (70)]1 M M 7
7
:::; - sup IRm [g (7)]1:::; Co70 M 70 for all 7 E (0,70]' Hence IRm [g (70)]1 == 0. Since 0(70) = TpM, M is diffeomorphic to jRn. Part (iii) follows since a flat shrinking gradient Ricci soliton on jRn must be the Gaussian soliton. 0 REMARK 8.18. (i) The Riemannian analogue of Corollary 8.17(i) is lim Vol B (p, r) = 1. wnrn
r-+O
(ii) Note that for the shrinking gradient Ricci soliton 9 (7) in subsection 7.3 of Chapter 7, the metric 9 (0) is not well-defined. The monotonicity of the reduced volume can be easily generalized to the following. For any fixed measurable subset A c TpM, we can define D(A, 7) to be the set of vectors V E A such that TV > 7, i.e.,
D(A, 7)
~
{V
E
A: TV > 7} = An 0(7).
It is clear that D(A,7) satisfies D(A,72)
c D(A,71) if 72 > 71.
COROLLARY 8.19 (.c-relative volume comparison). Suppose (Mn, g(7)), 7 E [0, T], is a complete smooth solution to the backward Ricci flow with the curvature bound IRm (x, 7)1 :::; Co < 00 for (X,7) E M x [0, T]. Define for any 7 E (0, T) and any measurable subset A c TpM, VA
(7)
~
{ } £:,. exp(D(A,r))
(41l'7)-n/2 exp [-f. (q, 7)] dJ.Lg(r) (q).
394
8.
APPLICATIONS OF THE REDUCED DISTANCE
for all l' E h,T2])' This implies f is Coo on M x h, T2J and (8.26) holds on M x [1'1, T2J . Thus 9 (1') is a shrinking gradient Ricci soliton. Given TO E h, T2J , by Proposition 1.7, the shrinking gradient Ricci soliton structure 9 (TO) , V f (TO) , - r~) may be put into a canonical timedependent form (1.11) defined for all t < TO,
(M,
g(t) = f (t) t.p(t)*g (TO), TO where g(t) is a solution of the Ricci flow, by (1.10), i.e., f (t) = TO - t (c = - ';0)' and t.p(t) is a I-parameter family of diffeomorphisms with t.p (0) = idM . By the uniqueness of complete solutions of the Ricci flow with bounded curvature (see Chen and Zhu [82]) and since g(O) = 9 (TO), we have
9 (TO - 1'),
9 (1') =
so that
t.p(TO - T)*g (TO) = TO 9 (1')
(8.27)
for
l'
E (0, TOJ.
l'
Since IRm [g (T)JI :::; Co < (8.27) we have
00
for
sup IRm [g (To)JI M
l'
E [0, TJ (we just use this for
l'
small), by
= sup IRm [t.p(TO - T)*g (To)JI M
l'
l'
:::; - sup IRm [g (1')]1 :::; Co-
TO
TO O. Since 0(1'0) = TpM, M is
M
for all l' E (0, TOJ. Hence IRm[g(To)JI = diffeomorphic to ]Rn. Part (iii) follows since a flat shrinking gradient Ricci soliton on ]Rn must be the Gaussian soliton. 0 REMARK 8.18. (i) The Riemannian analogue of Corollary 8.17(i) is lim VolB (p, r)
= 1.
wnrn
r-.O
(ii) Note that for the shrinking gradient Ricci soliton 9 (1') in subsection 7.3 of Chapter 7, the metric 9 (0) is not well-defined. The monotonicity of the reduced volume can be easily generalized to the following. For any fixed measurable subset A c TpM, we can define D(A, 1') to be the set of vectors V E A such that TV > 1', i.e.,
D(A, 1')
~
{V
E
A: TV > T} = An 0(1').
It is clear that D(A,T) satisfies D(A,T2)
c D(A, 1'1) if 1'2 > 1'1.
COROLLARY 8.19 (C-relative volume comparison). Suppose (Mn,g(1')), l' E [0, TJ, is a complete smooth solution to the backward Ricci flow with the curvature bound IRm (x, 1')1 :::; Co < 00 for (x,1') E M x [0, TJ. Define for any l' E (0, T) and any measurable subset A c TpM, VA
(1')
~
(471'T)-n/2 exp [-f (q, T)J dJ.tg(r) (q).
{
J.c
r
exp(D(A,r))
2.
Then for any
71
REDUCED VOLUME FOR RICCI FLOW
395
< 72,
PROOF. By the definition of the C-Jacobian we know that for any L1 function f on M
r
JL,. exp(D(A,r))
f(y) dJ.tg(r)(Y)
=
r
JD(A,r)
f(C r exp(V))C JV(7) dx(V).
(We have used this change of variables formula for A = TpM in previous sections.) We have
VA
(72) =
S
r
JD(A,"T2)
r r
JD(A,"T2)
S
(47r72)-n/2e-l(-YV("T2),"T2)CJv(72)dx(V) (47r7I)-n/2 e -l(w(rt},TJ.) CJv (71)
dx (V)
(47r71)-n/2e-l(w(71),rl) C Jv (7I)
dx (V)
JD(A,71)
= VA (71).
o The above can be thought of as a relative volume comparison theorem for the Ricci flow. This is along the lines of the generalization by Shunhui Zhu in [384] (see also Theorem 1.135 in [111] for example). 2.4. Monotonicity of reduced volume revisited. Now we give another proof of the monotonicity of the reduced volume without using the C-Jacobian. Recall that under the evolution equations
a = 2I4j,
a7 gij with
7
a
2 n a7f - 6f + IVfl - R+ 27
= 0,
> 0, we have
d~ 1M 7- n/ 2e- f dJ.tg = 0. In comparison, by (7.146), the reduced distance .e is a subsolution to the above equation for f. We use this fact to give another proof of the monotonicity of reduced volume. THEOREM 8.20 (Monotonicity of the reduced volume: second proof). Let (Mn, g (7)) , 7 E [0, T], be a complete solution to the backward Ricci flow satisfying the curvature bound IRm (x, 7)1 Co < 00 for (x, 7) EM x [0, T]. Then for any 7 E (0, T), the reduced volume V (7) is differentiable and nonincreasing:
s
(8.28)
396
8. APPLICATIONS OF THE REDUCED DISTANCE
PROOF. We justify the differentiation under the integral sign in equality (8.17). Consider the difference quotient for the reduced volume integrand:
Note that
if (T) = (47r)-n/2
(8.29)
dd
T
lim (
h-+OJM
so that the time-derivative of if (T) exists if the limit on the RHS exists. At any point (q, T) where f is differentiable (e.g., for each T, a.e. on M), we have
= T- n / 2 exp [-f (q, T)] (_~
lim
h-+O
2T
_
~f +
uT
R) .
Recall by Lebesgue's dominated convergence theorem that if we can show there exists a function W (q, T) such that for T > 0 there exists CT > 0 where
I
(8.30)
w(q, T)
lim (
h-+OJM
on M
lim
J M h-+O
h)dj-Lg(T)
(q).
Thus, provided we have (8.30), d d
T
if (T) = (47r)-n/2 = (47r)-n/2
(
lim
J M h-+O
1M T-
n / 2 exp
[-f (q, T)] ( -;:. -
;~ + R) dj-Lg(T) (q)
::; 0, where the last inequality follows from (7.151). This is the reduced volume monotonicity formula. To see (8.30), we first observe that
q, T,
h)
==. (T + h)-n/2 -i(q,T+h) dj-Lg(T+h) (q) _ ed () j-Lg(T)
q
T
-n/2 -i(q,T)
e
is a locally Lipschitz function of h near h = 0, for T > 0 fixed. (Note that
2. REDUCED VOLUME FOR RICCI FLOW
397
analysis (see Corollary 15 on p. 110 of [313]), we have
iP(q,7,li)
=
K(iP 1(q,7,li) -
=
~ {k ~
[(7 + h)-n/2 e-£(q,T+h) d/-Lg(T+h) (q)] dh
h io ah
d/-Lg(T) (q)
Kfak (-2(7:h)
=
(8.31)
iP 1(q,7,0))
-
~! (q,7+h) +R(q,7+h))
x (7 + h)-~ e-f(q,T+h) d/-Lg(T+h) (q) dh.
d/-Lg(T) (q) By Lemma 7.59, we have exp [-f (q, f)] ~ exp
(8.32)
(
-e-
2C 0
T dg2(O) (p, q) 4f
nCof)
+ -3-
,
which decays exponentially quadratically in terms of the distance function. Also we have gTd/-L9(T) IT=1' (q) = R (q, f) d/-Lg(1') (q), so that we have the following bounds for the volume form: e
-nColhl
< d/-L9(T+h) (q) < nColhl -
d/-L9(T) (q) - e
.
By Lemma 7.59, we also have
< 2CoT d;(T) (p, q) (. q,7 - e 47 f) (
-)
nCof
+ -3-'
so it follows from Lemma 7.60(ii) that
af (q,7) I _ ~ --=C 1 (f (q, f) + Af) I-a 7 T=T 7 C-1 ( e2CoT d;(1') (p, q) < - f 4f
(8.33) where A <
00.
Hence, for f
_
I _~ 2-
af ( -) a q,7
7
(8.34)
7
n C1 < -+- 2f f x
--!!
7
= 7 + h,
2
(
+ (nco -3+ A)-) 7 '
we have
+ R (q,7-)I--~ -£(q,1') d/-Lg(1') (q) 7 e d ( ) /-Lg(T) q
d e 2CoT ;(T) (p, q)
4f
+ (nco -3+ A) 7-)
(p, q) exp (-2COTd;(O) -e 4f
+nCOf) 3
+n C,0
enCol1'-TI .
On the other hand, by the curvature lower bound and the Bishop-Gromov volume comparison theorem,
(8.35)
2. REDUCED VOLUME FOR RICCI FLOW
397
analysis (see Corollary 15 on p. 110 of [313]), we have 1 Cf.J(q, 7, h) = h (Cf.Jl(q,7,h) - Cf.J 1 (q,7,0))
[(7 + h)-n/2 e-l(Q,r+h)dJ.tg(r+h) (q)] dh
= ~ [h ~ h Jo 8h
~foh (-2(7: h) -
=
(8.31)
dJ.tg(r) (q)
;: (q,7+h)+R(q,7+h))
x (7 + h)-!j e-l(q,r+h) dJ.tg(r+h) (q) dh.
dJ.tg(r) (q) By Lemma 7.59, we have exp [-£ (q, f)] ~ exp
(8.32)
(
-e-
2C 0
T dg2(O) (p, q) 4f
ncof)
+ -3-
,
which decays exponentially quadratically in terms of the distance function. Also we have trdJ.tg(r) Ir=r (q) = R (q, f) dJ.tg(r) (q), so that we have the following bounds for the volume form:
< dJ.tg(r+h) (q) < nColhl (e. dJ.tg(r) q) -
-nColhl
e
-
By Lemma 7.59, we also have
£ (_) 2CoT d;(r) (p, q) q,7 ~ e 47
nCof
+ -3-'
so it follows from Lemma 7.60(ii) that 1
8£ (q,7) I ~
u7
-=-
_ ~ Cl (£(q, f) r=r 7
< Cl (e2CoTd;(r) (p, q)
(8.33) where A <
-
00.
f
4f
Hence, for f =
7
+
(nco 3
+
A)
f) ,
+ h, we have
_u78£ (q,7-) + R (q,7-) I--!j e-l(q,r) ddJ.tg(r) (q) ( ) I -~ 27J.tg(r) q 7
!l
(8.34)
+ Af)
< n
-
x
2f
--!!
7
2
2CoTd;(r) (p, q) 4f
+ Cf
1 (
exp
(-
e
-e
A) -) rr + (nCo -3- + 7 + nLlO
2COT d;(O) (p, q) 4f
+ -3ncof)
nColr-rl
e
.
On the other hand, by the curvature lower bound and the Bishop-Gromov volume comparison theorem,
(8.35)
398
8.
APPLICATIONS OF THE REDUCED DISTANCE
for some constant C4 < 00. From (8.34), (8.35), and (8.31), it is easy to see that I~(q, 7, h)1 is bounded by an integrable function on M, independent of h small enough. 0 EXERCISE 8.21. Let (Mn,g) be a Riemannian manifold, p E M, and assume that expp : TpM ~ M is a diffeomorphism (M is then diffeomorphic to ]Rn). Define CPs: M ~ M by CPs : expp (V) ~ expp (eSV) for V E TpM. Show that {CPs} sElR is a I-parameter group of diffeomorphisms and
where r (x) ::§= d (x,p). SOLUTION TO EXERCISE 8.21. Another way to define CPs is CPs (x) = expp (e S exp;l x) . We have CPSI (CP S2 (x)) = expp (e SI exp;l 0 expp (e S2 exp;l x)) = expp (e SI +S2 exp;l x) = CPSI+S2 (x).
Note r (CPs (x)) = eS lexp;l xl. We compute
(:8
CPS) (x) = (dexpp)es exp;l x = (rVr) (CPs (x))
(e S exp;l x)
= V (r;)
(CPs (x)).
Now we consider a Ricci flow analogue of the above discussion. Given p E M and 70 > 0, assume that S1(p,0) (70) = TpM. For 7 such that S1(p,O) (7) = TpM, define by that is, We compute (:7 ¢T) (x)
=
(! £T
expp)
((L:To expp) -1 (x))
= X (¢T (x)) = (Vi) (¢T (x)) , where X ::§= d~ TV (7) for V
= (£TO exppr 1 (x).
3.
NO LOCAL COLLAPSING VIA REDUCED VOLUME MONOTONICITY
399
3. A weakened no local collapsing theorem via the monotonicity of the reduced volume In this section, (Mn, 9 (t)) , t E [0, T), shall denote a complete solution to the Ricci flow with T < 00 and SUPxEM, tE[o.tI]IRmg (x, t)1 < 00 for any tl < T (i.e., the curvatures are bounded, but possibly not uniformly as t ~ T, as in the case of a singular solution). We fix a time To E (t, T) and a basepoint Po EM. Let g(T)~g(To-T).
Then (Mn, 9 (T)) , T E [0, To], is a solution to the backward Ricci flow with initial metric 9 (0) = 9 (To) and bounded sectional curvature. Let C b) denote the C-Iength of a curve ,,(, let L : M x (0, Tol ~ R denote the L-distance, let £ : M x (0, Tol ~ R denote the reduced distance, and let if : (0, Tol ~ (0,00) denote the reduced volume, all with respect to 9 (T) and the basepoint (po, 0) .
3.1. A bound of the reduced distance. The following lower bound of the reduced volume will be used in the proof of the Weakened No Local Collapsing Theorem 8.26. LEMMA 8.22 (Lower bound for if at initial time). (i) (£ upper bound) Fix an arbitrary ro > 0. There exists a constant Cl > 0, depending only on rO, n, T, and sUPMx [0,T/2] RCg(t), and there exists qo E M such that
£ (q, To) :::; Cl
for every q E Bg(o) (qO, ro) .
(ii) (if lower bound) Suppose there exist rl > 0 and VI > Volg(o) Bg(o) (w, rl) ;::::
°such that
VI
for all w EM. Then there exists a constant C2 > 0, depending only on rl, VI, n, T, and supM x [O,T /2] Rc g( t), such that
if (To) ;:::: C 2 · PROOF. (i) By Lemma 7.50 there exists qo EM such that 3
£ (qO, To - T/2) = min f (q, To - T/2) :::; ~2. qEM
For any q E Bg(o) (qO, ro) , let f3 : [To - T /2, Tol ~ M be a constant speed minimal geodesic from qo to q with respect to 9 (0) . Defining Co
3This corresponds to time t
~
sup RCg(t), Mx[0,T/2]
= T /2.
3.
NO LOCAL COLLAPSING VIA REDUCED VOLUME MONOTONICITY
399
3. A weakened no local collapsing theorem via the monotonicity of the reduced volume In this section, (Mn,g (t)) , t E [0, T), shall denote a complete solution to the Ricci flow with T < 00 and SUPXEM, tE[O,td IRmg (x, t)1 < 00 for any tl < T (Le., the curvatures are bounded, but possibly not uniformly as t --+ T, as in the case of a singular solution). We fix a time To E (t, T) and a basepoint Po EM. Let g(1')~g(To-1').
Then (Mn, 9 (1')), l' E [0, To], is a solution to the backward Ricci flow with initial metric 9 (0) = 9 (To) and bounded sectional curvature. Let C b) denote the C-Iength of a curve ,,(, let L : M x (0, To] --+ 1R denote the L-distance, let f. : M x (0, To] --+ 1R denote the reduced distance, and let V : (0, To] --+ (0,00) denote the reduced volume, all with respect to 9 (1') and the basepoint (po, 0) .
3.1. A bound of the reduced distance. The following lower bound of the reduced volume will be used in the proof of the Weakened No Local Collapsing Theorem 8.26. LEMMA 8.22 (Lower bound for
V at initial time).
(i) (f. upper bound) Fix an arbitrary ro > O. There exists a constant Cl > 0, depending only on ro, n, T, and SUPM x [0,T/2] RCg(t), and there exists qo E M such that
f. (q, To) ~ C1
for every q E Bg(o) (qO, ro) .
(ii) (V lower bound) Suppose there exist rl > 0 and VI > Volg(o) Bg(o) (w,rl) 2:
°such that
VI
for all w EM. Then there exists a constant C2 > 0, depending only on rI, VI, n, T, and SUPM x [0,T/2] RCg(t), such that
V (To) 2: C2· PROOF. (i) By Lemma 7.50 there exists qo EM such that 3
f.(qO, To - T /2) = min f. (q, To - T /2) ~ ~. qEM 2 For any q E Bg(o) (qO, ro), let f3 : [To - T /2, To] --+ M be a constant speed minimal geodesic from qo to q with respect to 9 (0). Defining
Co
~
sup Mx[0,T/2]
3This corresponds to time t
= T /2.
RCg(t),
400
8. APPLICATIONS OF THE REDUCED DISTANCE
we have 1·1;(7') ::; eCoT I'I;(TO) = eCOT 1'1~(0) acting on vector fields for [To - T /2, To] . We can estimate £, (13) as follows: £,
E
(;3)
::; {TO .jT (Rg(7') (13 (7)) lTo-T/2
d
+ 1 12 ) 7
g(7')
d7
::; -2 (T~/2 - (To - T /2)3/2) sup Rg(t) 3 Mx[0,T/2] =
7
~ (T~/2 _ (To _ T /2)3/2)
(
+ eCoT iTO
sup Rg(t) Mx[0,T/2]
To-T/2
.jT Id1312 d 7
g(O)
d7
+ 4eCoT (dg(o) ~;' qo) )2)
(G
COT ::;3~T3/2 no+ 4e T2 r 5) . Let a : [0, To - T /2] - t M be a minimal £'-geodesic with a (0) a (To - T /2) = qo. Then
= Po
and
= 2JTo - T/2· f (qO, To - T/2) ::; nJT/2.
£, (a)
Consider the concatenated path: 'Y (7)
== (a '-' 13)(7) = { a (7) if t
E [0, To - T/2]' 13(7) iftE[To-T/2,To].
.
This path is well defined and piecewise smooth. We have 1 f (q, To) ::; 2JTo£' (T)
::;
1
= 2JTO [£, (a) + £, (;3)]
~ [nJT/2 + ~ T 3/ 2 ( nCo + 4e~~ r5) ]
~ Cl (ro, n, T, Mx[0,T/2] sup Rc get») . (ii) Choosing
ro = rl in (i), we have
V (To) = 1M (41l'To)-n/2 e-f.(q,TO)d/Lg(o) (q) 2: {
1Bg(o) (qo,rI)
2: {
1B9
(0)
(41l'To)-n/2 e-f.(q,TO)d/Lg(o) (q)
(41l'To) -n/2 e -Cl d/Lg(O) (q) (qO,rl)
2: VI (41l'T)-n/2 e- Cl
~ C2 (Vl,r 1,n,T, Mx[0,T/2] sup RCg(t»). o
3.
NO LOCAL COLLAPSING VIA REDUCED VOLUME MONOTONICITY
401
3.2. The weakened no local collapsing theorem. In Lemma 8.9 we saw how the reduced volume of a static metric bounds the volume ratios of balls from below. Similarly, the reduced volume monotonicity for solutions of the Ricci flow enables one to prove a weakened form of the no local collapsing theorem, which we first encountered in Chapter 6 using entropy monotonicity. DEFINITION 8.23 (Strongly K-collapsed). Let K > 0 be a constant. We say that a solution (Mn,g(t)), t E [O,T), to the Ricci flow is strongly K-collapsed at (qO, to) EM x (0, T) at scale r > 0 if (1) (curvature bound in a parabolic cylinder) IRmg (x, t)1 ::; for all x E Bg(to) (qO, r) and t E [max {to - r2, O} ,to] and (2) (volume oj ball is K-collapsed)
fx
Volg(to)
Bg(to) (qO,
r) <
_....:::...0...'::':""---"-''--''-'-_ _
rn
K.
Given an r > 0, if for any to E [r2, T) and any qo E M the solution 9 (t) is not strongly K-collapsed at (qO, to) at scale r, then we say that (M, 9 (t)) is weakly K-noncollapsed at scale r. Recall that the reduced volume V (7) has the upper bound 1. When the solution is strongly K-collapsed, we shall obtain a better upper bound Jor V which tends to 0 as K tends to O. THEOREM 8.24 (Main estimate for weakened no local collapsing). Let (Mn,g (t)) , t E [0, T), be a complete solution to the Ricci flow with T < 00 and suppose SUPMx[o,tl]IRml < 00 Jor any tt < T. Then there exists Cl = Cl (n) E (O,~] depending only on n such that iJ Jor some KIln::; Cl (n) , the solution 9 (t) is strongly K-collapsed at (p*, t*) at scale r, where t* > and r < v'f;, then the reduced volume V* oj g* (7) ~ 9 (t* - 7) with basepoint p* has the upper bound
t
where and -!...
¢(c,n)-;-
exp Un (n - 1)) nl2 12 c
(47rt
+wn -dn-2)
REMARK 8.25. Note that lime-+o ¢ (c, n)
n-2 2
e
_ n-2 2
(
1)
exp -2 ~
.
yc
= O.
We will prove this theorem in the next subsection. This theorem gives a proof of the following weakened no local collapsing theorem. THEOREM 8.26 (Weakened no local collapsing). Let (Mn,g (t)) , t E [0, T), be a complete solution to the Ricci flow with T < 00. Suppose (1) sUPMx[o,h]IRml
< 00 Jor
any
tl
and
402
8. APPLICATIONS
OF THE REDUCED DISTANCE
(2) there exist rl > 0 and VI > 0 such that Volg(o) Bg(o) (x, rd ;::: VI for all x EM. Then there exists K, > 0 dependin9 only on rI, VI, n, T, and supM x [0,T/2] Rc g(t) such that 9 (t) is weakly K,-noncollapsed at any point (P*, t*) EM x (T /2, T) at any scale r < y'T/2. Note that if M is a closed manifold, then assumptions (1) and (2) of the theorem are automatically true. PROOF OF THEOREM 8.26. Let CI (n) be as in Theorem 8.24. Suppose is strongly K,-collapsed at a point (p*, t*) E M x (T /2, T) at a scale r < y'T/2 with K,I/n ::; CI (n) . Let c ~ K,I/n. Consider the backward solution of the Ricci flow
9 (t)
(1') ~ 9 (t* - 1') with basepoint p*. Since cr2 ::; r2 ::; t*, by the monotonicity of the reduced volume V*, we have V* (t*) ::; l%. (cr2) . 9*
By Lemma 8.22(ii), choosing (po, To) Applying Theorem 8.24 yields
0< C2 (rI' VI, n, T, ::;
= (P*, t*) , we
have
l%. (t*)
;::: C 2 > O.
sup Rc g(t») Mx[0,T/2]
exp(kn(n-1)) n/2 _n-2 ( 1) /2 c + Wn-I ()n-2 n - 2 2 e 2 exp - 2 r;:.
(471't
This implies a positive lower bound for c
.
yc
= K,I/n
and the theorem is proved.
o
REMARK 8.27. Under the assumptions of Theorem 8.26, (Mn,g(t)), t E [0, T /2], has bounded geometry; so it is easy to see that there exists K,I > 0 depending only rI, VI, n, T, and SUPMx [0,T/2] Rc g(t) such that 9 (t) is weakly K,I-noncollapsed at (p*,t*) EM x (0,T/2] at any scale r < y'T/2.
3.3. Bounding reduced volume from above when the solution is strongly K,-collapsed. This subsection will be devoted to the proof of Theorem 8.24, which follows directly from Propositions 8.28 and 8.30 below. Note that the assumptions on 9 (t) translate to the following assumptions on 9* (1'): (8.36) (8.37)
3.
NO LOCAL COLLAPSING VIA REDUCED VOLUME MONOTONICITY
Shi's local derivative estimate implies that there is a constant depending only on n such that
IVgoR (x, r)1 ~ rc~
(8.38)
for x E Bgo(o)
(P*'~)
C2
=
403 C2
(n)
and r E [0, 1r2] .
From (8.19) we can write the reduced volume of g* (r) as
V* (r) =
VI (r) + V2 (r),
where
Vdr) ~ (
(47rr)-n/2 e-i(W(T),T).cJv (r)dx(V) ,
V2 (r) ~ (
(47rr)-n/2 e-i(W(T),T).cJv(r)dx(V).
JlVlgo(o)~Cl/4 JlVlgo(o»Cl/4
Here both the reduced distance .e bv (r) , r) and the are defined with respect to g* (r) . CI
.c-Jacobian .c Jv (r)
PROPOSITION 8.28. Under the assumptions of Theorem 8.24, there exists CI (n) E (0,1]. depending only on n, such that if c = ".l/n ~ CI (n),
=
then
- (2)
VI cr
<
exp (in (n
-
(27rt
-
1)) n/2
/2
c
The idea of the proof is to show that for some choice of Cl, IV (r) is contained in Bg(o) (P*, r /2) and.e (!'V (cr2) ,cr2) has a lower bound independent of c when IVlg.(o) ~ c- I / 4 . The proposition then follows easily. LEMMA 8.29. Suppose (P*, t*), where t* > (8.36) holds, i.e., such that
IRm go (x,r)1
E
2
~ 2
Then there exists CI = c = ".l/n ~ CI (n), then 'Yv (r)
1
t, and r < ..;t;. are such that
for all x E Bg.(o)(p*,r) and r E [O,r ].
r
CI
(n) E (0,1] depending only on n such that if
Bg.(o) (p*,r/2) for any V
E
B;.(p.,o) (0,c- 1/ 4 ) and r
E
[0,cr2 ]
.
PROOF. We prove the lemma by contradiction. Suppose
V
E
Bgo(Po,O)
(0, c-
I / 4)
and r' E (0,cr2] is the first time such that 'Yv (r') E aBg.(o) (p*,r/2). Let V (r) ~ y'Tiv(r) = y'TX (r), so that limT-+o V (r) = V. Since 'Yv ([0, r'l) E Bg.(o) (p*, r/2) and r' E [0, 1r2] , (8.36) and (8.38) imply
n-1
Rc go > - -r2-
and
404
8. APPLICATIONS OF THE REDUCED DISTANCE
along
IV,
and hence we can use (7.48) to get 4
Therefore by Holder's inequality,
(t
ltv (T)lg.(T) dT) 2
~ loTI T- 1/2dT loTI v'Tliv (T)I;.(T) dT
~ 2v'cr loTI T- 1/ 2W (T)I;.(T) dT ~ (2v'cr)2 (e6(n-l)Cc- 1 / 2 +
We get the last inequality by choosing
C~C
(e6(n-l)c -
12(n-1)2
CI ~ ~
1))
such that
4(e6(n-l)ft l/2 + 12 (n~t2- 1)2 (e6(n-l)f _ 1)) <- ~16 for all t E [0, cll. Hence
If we also require Cl ~ 6~' then T' ~ cr2 ~ ~~. Since IRmg • (x, T)I ~ ~ for all x E Bg.(o) (p*, r) and T E [0, r2], we have 9* (x, T) ~ ~9* (x, 0) for T E [0, T'l and x E Bg.(o) (P., r). Hence
This contradicts
IV
(T')
E
8Bg .(0) (P., r /2) . The lemma is proved.
D
We now give a proof of Proposition 8.28. 4Note that (3 (u) ~ 'Yv (u 2/4) satisfies ~ = and C2 =;t.
V. In (7.48) we take Co = ~, T = c:r2,
3. NO LOCAL COLLAPSING VIA REDUCED VOLUME MONOTONICITY PROOF OF PROPOSITION
8.28. Let
let ,,;,1/n ~ Cl. If V E Bg*(p*,O) geodesic, then by Lemma 8.29,
(O',c-
Cl
1/ 4 )
405
be chosen as in Lemma 8.29 and and "YV I[O,er2] is a minimal C-
In the integral defining VI (cr2) , it follows from (8.18) that we only need to consider those vectors V minimal C-geodesic. Hence
E Bg*(p.,O) (O',C 1/ 4 )
for which ''YV I[O,er2] is a
where we have used 9* (cr2) ~ 29* (0) , and also VOIBg~~)(p*,r) ~ cn in the last inequality. 0 Finally we give an estimate for PROPOSITION
from above.
8.30. Under the assumptions of Theorem 8.24, we have
ii2 (cr2) ~ wn-dn PROOF.
V2 (cr2)
2) n;2 e- n;2 exp { -
2~}'
By (8.23), we have
(47l'cr2rn/2 e-e('YV(er2),er2) CJ v (7) ~ exp
(-1V1;.(0))'
406
8. APPLICATIONS OF THE REDUCED DISTANCE
Therefore
Vdcr2) ~ (
exp
JlVlg.(o)?Cl/4
=
1
00
Wn-l
e
_r2 n-l
e- 1 / 4
2
1 2
Noting that r n - e-2 r V2
~
(-1V1!.(0)) dx (V)
n-2
r
dr.
n-2
(n - 2)-2- e--2- , we get
(cr2 ) ~ Wn-l (n -
n-2 n-2 ( 2- exp 2)-2e--
1)
2JE .
o 4. Backward limit of ancient K-solution is a shrinker DEFINITION 8.31 (ancient K-solution). Let K be a positive constant. A complete ancient solution (M n , g(t)) , t E (-00,0]' of the Ricci flow is called an ancient K-solution (or K-solution for short) if it satisfies the following three conditions. (i) g(t) is nonflat and has nonnegative curvature operator for each t E (-00,0]. (ii) There is a constant C < 00 such that Rg (x, t) ~ C for all (x, t) E M x (-00,0]. (iii) g(t) is K-noncollapsed on all scales for all t E (-00,0]; i.e., for any p > and for any (p, t) E M x (-00,0]' if IRmg (x, t)1 ~ p-2 for all x E Bg(t) (p, p) , then
°
Volg(t) Bg(t)
(p, p)
-..=...:..."'--~--
pn
> K. -
If the curvature bound condition (ii) in the definition is replaced by the requirement that 9 (t) satisfies the trace Harnack inequality
8R
at + 2V R . X + 2 Rc (X, X)
~
° for all X,
we say that 9 (t) is a K-solution with Harnack. In Part II of this volume we will prove that in dimension n = 3 the notions of K-solution with Harnack and ancient K-solution are equivalent. In this section we prove that in all dimensions certain backward limits of ancient K-solutions are nonftat shrinking gradient Ricci solitons (Theorem 8.32 below). The proof will take several steps, which will be carried out in the following subsections. 4.1. Statement of the theorem. Let (Mn,g(t)), t E (-00,0], be a K-solution. For to E (-00,0] we define a solution to the backward Ricci flow (M, g (T)), T E [0,00), by
g (T)
~
9 (to - T).
4.
BACKWARD LIMIT OF ANCIENT It-SOLUTION IS A SHRINKER
407
*
Given Po E M, we have the reduced length .e (q, r) .e9 (q, r) , £-Jacobian £ J v (r) £ J~ (r) , and the reduced volume V (r) defined with respect to 9 (r) using the basepoint Po. For any r > 0, define dilated backward solutions:
*
(8.39)
gr((})
*
r- 1 . g(r(}),
for (} E [0,00).
Let qr E M be a point such that
.e(qn r) ~
n
"2
(by Lemma 7.50 such a point always exists). The following is Proposition 11.2 in [297]. THEOREM 8.32. -+ 00 and A > 1, there exists a subsequence, still denoted by ri, such that (Mn,gri((}),qr.), (} E (A-I, A) , converges in the Cheeger-Gromov sense to a complete nonflat shrinkmg gradient Ricci soliton (M~, goo ((}), qoo). (2) By choosing a sequence of Ak -+ 00 and using a diagonalization argument, we have for any ri -+ 00 that there exists a subsequence such that (M,gri((}),qrJ, (} E (0,00), converges in the CheegerGromov sense to a complete nonflat shrinking gradient Ricci soliton, which we also denote by (Moo, goo ((}), qoo). Since the trace Harnack estimate holds for the sequence, it also holds for goo ((}); hence the limit (Moo, goo ((})) is a ",,-solution with Harnack.
(1) For any sequence ri
In dimension n = 3, because of the equivalence between ",,-solutions with Harnack and ancient ",,-solutions, goo ((}) has bounded curvature. REMARK 8.33. When n sectional curvature.
~
4, it is not clear to us if the limit has bounded
Before we begin the proof of Theorem 8.32, we end this subsection with some elementary properties about the change of the reduced length .e and the £-Jacobian £Jv (r) under scaling (8.39). Let ,ft; (-) be the £geodesic, with respect to the solution gn satisfying ,ft; (0) = Po E M and limo-+o JOi'ft; ((}) = W. The reduced length .e9 .,. (q, (}) and the £-Jacobian £ JV" ((}) shall be defined with respect to gr ((}) using the basepoint Po. LEMMA 8.34 (Elementary scaling properties). For any r > 0 and (} E [0,00), we have
= I~ (r(}) ,
(8.40)
10=V ((})
(8.41)
.e9 .,. (q, (}) =.e9 (q, r(}) ,
(8.42)
r- n / 2 £ J~v ((}) = £ J~ (r(}) .
408
8. APPLICATIONS OF THE REDUCED DISTANCE
PROOF. Given that we know how curvature changes under scaling, it is easy to check that T~ (r(}) satisfies the .c-geodesic equation for the solution gr ((}) . We compute using the change of variable 7 = r(} that
· ytJ r;;(}dT~ (r(}) 11m 9->0+ d(}
1·1m
=
1'->0+
,=dT~ (7) y . '-r = y '-V r. dr
yr
By the uniqueness of the initial-value problem for the .c-geodesic equation, we get (8.40). Given any curve a(f), f E [O,rO], from (po,O) to (q,rO) , the curve
ar
(0) ~ , (rO) , 0E [0, ()] , joins (po, 0) to (q, ()). We compute _1_.c9 (a) = _1_ r 9 Vf (R9 (T) + 1d~ 12 2VTO Jo
2VTO
=
dr
~r9.;;e (r-l R9 (9) + rr(} Jo
2y
) df
9(1') 1
1da: 12
T
-
r JO (R9T(9)_+ 1dar 12 ) 2...;0 Jo dO 9T(9)
=
1 2...;o.c
__ 1
d(} 9T ( 9)
9
9T
) rdO
dO
(a r ) .
(m] ,
Since £9 (q,rO) = infa [2J;:e.c 9 (a)] and £9T (q,(}) = inf{3 [2~.c9T we conclude that £9 (q, r(}) = £9T (q, ()) and the minimizing .c-geodesics are related by (8.40). From (8.40) we get .c;gexP: (TpoM,g (O,po)) ~ (M,g (rO)) where V ~ T~ (r(}) ,
.cZ
T
exp: (TpoM, r-lg (O,po)) ~ (M, r-lg (r())) where..;TV ~ T~ (r(}).
Hence the Jacobian of the above two maps are related by (8.40).
0
4.2. The blowdown limit of gri(O). Recall by Lemma 7.64 that the estimate
IV£ (q, r)12 + R (q, r)
(8.43)
~ 3£ (q, r) r
holds for the solution 9 (r). We have the following consequence in regards to the space-time points (qr, r) . LEMMA 8.35 (Estimates for £ and R in large neighborhoods of (qr, r)). For any c > and A> 1, there exists 6 > such that for any r > 0,
°
°
£ (q,f) ~ 0- 1
for all (q, f)
E
and
fR (q, f) ~ 6- 1
B 9 (r) (qr, \Ie-lr ) x [A- 1 r, Ar] .
4. BACKWARD LIMIT OF ANCIENT PROOF. T
~-SOLUTION
IS A SHRINKER
409
First we shall prove that there exists 6 > 0 such that for any
> 0, £ (q, f) ::;
(8.44) for all (q, f) E Bg(r) (qr.
~6-I
v'CIT) X[A-IT, AT] . From
(8.43), since our an-
cient solution has Rm ;::: 0 (in particular, R;::: 0)
3 I\J yI£ (q,f) Ig(r)_::; V/3 44f_ .
v'E:- IT)
, let l' (8) be a minimal normal geodesic from qr to q with respect to the metric 9 (T). Since £ is locally Lipschitz, we have
For any q E B g( r) (qn
IyI£ (q, T) -
I Jo
rd9(T)(q,q(r))
ylf (qn T) ::;
I\J yI£ b (8), T) Ig(r) d8
< /3.~= f3. -V~ V~ Since £(qr, T) ::;
(8.45)
~,
by the above estimate, we have
e(q, T) S
(
fJ + Ii)'
for q E
B,(T)
(q" velT) .
From Lemma 7.65, we have for q E M ·f-E[A-I ] A - 2
f(q,f) A2 ·f- [ A 1 f(q,T)::; ITET, T. (Note that the inequalities in the above two lines have the same form.) Hence for q E Bg(r) (qn
v'E:-IT)
f(q,;') S
and f E [A-IT, AT] ,
A2
(fJ + Ii)
2
Now (8.44) is proved by choosing
.-1
= 3A2
(fJ + Ii)
2
Now f R (q,f) ::; 6- 1 follows directly from (8.43) and (8.44); the lemma is proved. 0 For any sequence
Ti
--t
00,
consider the sequence of solutions
410
8. APPLICATIONS OF THE REDUCED DISTANCE
> 0, Lemma 8.35, after parabolic rescaling 9 (T) by Ti, yields the curvature bound 5- 1 = 5(n,E,A)-1 for gri(O) on B gTi (l) (qri ,VE- 1) x [A-1,A]. Applying Lemma 8.35 with E = 1 and A = 2, we obtain
For any E
IRmgTi (q, 1)1 ::; 5- 1 = 5 (n, 1,2)-1
for q E B gTi (l) (qri' 1).
Since g(O) is K-noncollapsed on all scales, we have gri(O) is K-noncollapsed on B gTi (l) (qri' 1) and the injectivity radius estimate inj gTi(l) (qr,) ~ 51 (n, K). Now we can apply Hamilton's Cheeger-Gromov-type Compactness Theorem 3.10 to the sequence of solutions gri (0) to the backward Ricci flow to get
(Mn,gri(O),qrJ ~ (M~,goo(O),qoo)
for 0 E [A-1,A].
The limit is a complete solution to the backward Ricci flow. Since each gri (0) satisfies the trace Harnack estimate, the limit goo (0) satisfies the trace Harnack estimate. Note that goo(O) is K-noncollapsed on all scales, has nonnegative curvature operator, and satisfies inj goo(l) (qoo) ~ 51 (n, K) . To finish the proof of Theorem 8.32, we need to show that for each 0, goo (0) is a nonflat shrinking gradient Ricci soliton.
4.3. The limit of reduced length fdq,O). Let fdq, 0) ~ fgTi (q, 0), [A -1, A] , denote the reduced distance of the solution gr; (0) with respect to the basepoint Po. LEMMA 8.36 (Limit of the reduced distance). (i) The limit
oE
exists in the Cheeger-Gromov sense on Moo x [A-I, A].5 (ii) The limitfoo(q,e) is a locally Lipschitz junction on Moo x [A-1,A] and V goo((J)foo(q, 0) and a~'f (q, e) exist a.e. on Moo x [A-I, A]. (i) Suppose 0, by Lemma 8.35, (8.43), (7.100), and the scaling property (8.40), we obtain for all i, 0 E [A-I, A], and q E BgTi (1) (qri , vel) eM, the estimates PROOF.
0::; fi(q,O)::; 5- 1,
(8.46)
IVgTi((J)fi(q, e)I~Ti(O) 1
::; 35- 1 ,
8fi 8e (q,O ) I ::; 25 -1 .
Hence for i large enough, fi (
-t
foo(q, e) on the closed set Bgoo(l) (qoo, tVC1)
X
5See the proof below for what we mean by convergence in the Cheeger-Gromov sense.
4.
BACKWARD LIMIT OF ANCIENT t;;-SOLUTION IS A SHRINKER
411
[A -1, A] for some subsequence. Since £ is arbitrary, (i) then follows from a diagonalization argument. The convergence of the pulled-back reduced distance functions .ei(
o
2g; = R- 1V'.e1
In the next lemma we show that the equality preserved under the limit.
2 -
*" is
LEMMA 8.37 (Properties of the limit of the reduced distance functions).
(i) We have (8.47) (ii) For any smooth compactly supported
we have (8.48)
{A {
JA-l JMoo
J'f + g~.eoo· (8-Rgoo+29 V'
V' goo
PROOF. (i) Equation (7.94) tells us that scaling, we have
a.ei
2 ao
2g; + 1V'.e1
.e + 1V' g'Ti(9).ei 2 Rg'Ti(9) + -0 i
1
-
R+ *" = O. By
2 -
= O.
It suffices to prove that oo 8.ei ao (
1) IY(O)l 1) IY (O)l g'Ti(9) . n8- + 2
( ) (Y(O),Y(O)) ~ ( n.edq,O) Hess(q,9).ei 0 ~ A (
1
2
+ 20
g 'Ti(9)
2
Since 91', (
K
c
8. APPLICATIONS OF THE REDUCED DISTANCE
412
there exists a neighborhood Bgoo«(J)
(q, 1injgoo«(J) (q)) and a smooth function
F on K such that for each 0 E (A -1, A) the function
c;dx, 0)
~ F (x, 0) -
fi ( cJ>i
0
eXPgoc(l,q) (x) ,0)
l6
is convex on B goo (l,q) (0, injgoo(l) (q)) C (TqM oo , 900 (1, q)). Since fi (·,0) is differentiable almost everywhere, V' goo (l,q)C;i (x, 0) a.e. on B goo (l,q) (0,
-t
V' goo(l,q) (F (x, 0) - foo (eXP9oo(1,q) (x), 0))
1~ injgoo(l) (q))
V'goc(l,q)fi(cJ>i oexPgoo (l,q) (x)
,0)
by Theorem D6.2.7 in [202]. Hence
converges to V'goo(1,q)foo (exP9oo (l,q) (x)
,0)
Bgoo (l,q) (0, l6 injgoo(l) (q)) and V' gTi«(J)fi (cJ>i eXPgoc (l,q) (x) ,0) converges to V' goo «(J)foo (exP 9oo (l,q) (x) ,0) a.e. on B goo (l,q) (0, l6 injgoo(l) (q)) .
a.e. on
0
Because q is an arbitrarily point on Moo, we have proved that
(8.49)
lV'gT/il;Ti«(J) (cJ>i (-) ,0)
-t
lV'gocfool!oo«(J) (-,0)
a.e. on Moo for each 0 E (A-I, A) . From 2~+1V' gTi«(J)fiI2_RgTi«(J)+j = 0 and (8.49) we know ~ (cJ>i (-),0) converges a.e. on Moo for each 0 E (A-I, A) . Since fi (cJ>i (-) ,0) converges to foo (·,0) uniformly, we conclude
Bfi ( () ) BO cJ>i· ,0
-t
Bfoo ( ) BO ·,0
a.e. on Moo for each 0 E (A-I, A) .
°
(ii) For any smooth compactly supported function cp(q, 0) ~ on Moo x (A -1, A), for i large enough we can extend cp (cJ>; 1 (ql) , 0) by 0 to a smooth function, still denoted by cp (cJ>;1 (ql) ,0) , which has compact support on M x (A-I, A). Using the Lipschitz test function e- fi (qllJ)CP (cJ>;1 (ql) ,0) in (7.146), we get
and
i~l 1M (~~ +
V'gT/i· V'gTiCP - RgTi
+ ~)
x e-fi(~i(q),(J)cp (q, 0) dJ.L~:gTi«(J) (q) Taking the limit i
- t 00,
we obtain (8.48). The lemma is proved.
dO ~
O. 0
4. BACKWARD LIMIT OF ANCIENT I\:-SOLUTION IS A SHRINKER
413
4.4. The limit of the reduced volume. Note that the limit goo(O) is defined for 0 E [A-I, A] . Instead of considering the reduced volume of this limit, we define the function
which will play the role of the reduced volume; formally this is the reduced volume using the limit function foo. The fact that Voo(O) is finite follows from the following lemma. LEMMA
8.38.
(i) We have
(ii) Voo(O) is a constant contained in (0,1) . (iii) For any 1/J(O) which has compact support in (A-I, A), fA f (47rO)-n/2 e-£oo (q,8) 1/J' (O)dJ1,goo (8) (q)dO JA-l JMoo
(8.50)
= fA Voo (O)1/J'(O)dO = 0, JA-l
where 1/J' (0) = d1/J / dO. (i) Let Ti be a subsequence such that both gTi and fi(q, 0) converge. By (8.23), we have PROOF.
(47rTio)-n/2e-l'(')'V(Ti8),Ti8) .c Jv (TiO) ::; (47r) -n/2 e -lvl~(o).
Then by Lebesgue's dominated convergence theorem and (8.40), we have .lim V(TiO) t-+oo
= f lim ((47rTio)-n/2 e- l (')'V(Ti 8),T;8) .cJv (TiO) dJ1,g(o) (V))
iJRn t-+oo
= f JMoo
(47rO)-n/2.lim (Ti -n/2 e - l i ("fy"fiv(8),8) dJ1,g(Ti 8)) t-+oo
= f (47rO)-n/2 e- loo (q,8) .lim dJ1,gT(8)
JJRn
= f
t-+oo
(47rO)-n/2
e-£oo (q,8)dJ1,goo (8)
•
(q)
= Voo(O).
JMoo
Since
V(T)
is a monotone decreasing function, we have
(8.51) In particular, Voo(O) is independent of O.
414
8. APPLICATIONS OF THE REDUCED DISTANCE
(ii) Note that V(oo) < 1 follows from Corollary 8. 17(iii). To see V(oo) 0, we compute using (8.46) and 0 = 1 that V(Ti)
>
= f (47r)-n/2 e- li (q,1)dJ.Lgd1)(q) 1M
'
~ f
IfB gT,.
(47r)-n/2 e- li (q,1)dJ.L (1)
T
(1)(q)
g,
(qTi ,e:- 1/2 )
~ (47rA)-n/2 e-o- 1 VoIgT, (1) B gTi (l) (qTi,c- 1/ 2)
= (471' A) -n/2 e _0-1 . Ti-n/2
U
I g(Td B g(Ti) ( qTi' Ti1/2 C-1/2)
vO
•
By Lemma 8.35, we have R (q, Ti) ~ 0~1 on the ball Bgh) (qTi' Til/2c-1/2). It follows from g(T) being ~-noncollapsed on all scales (choosing the scale r = min { Ti1/ 2C 1/2, Ti1/ 201/ 2} ) that I B g(Ti) ( qTi' Ti1/2 C-1/2) Ti-n/2'\T vO g(Td
Hence
~ ~.
(mIll . { c -1/2 , ud/2})n .
V(Ti) ~ (47rA)-n/2e-O-l~. (min {c- 1/ 2,01/2})n
and V(oo)
> O.
(iii) For any 1jJ(O) which has compact support in (A-I, A), we compute
(47rO)-n/2 e- loo (q,(J) 1jJ' (O)dJ.Lg oo ((J) (q)dO j A f A-l1Moo
= jA Voo(O)1jJ'(O)dO = Voo(O) fA 1jJ'(O)dO = O. A-I
lA-I
o 4.5. The limit is a shrinking gradient soliton. Let 1jJ(O) ~ 0 be a smooth function only of 0 with compact support in (A-I, A). Applying Stokes's theorem for Lipschitz functions, we get
L~1 iMoo (47rO)-~ (a~; - Rgoo + ;) e- loo (q,(J)1jJ (0) dJ.Lgoo((J) (q) dO = jA f A-11M
=0,
(47rO)-~ e- loo (q,(J)1jJ' (0) dJ.Lgoo((J) (q) dO
4. BACKWARD LIMIT OF ANCIENT It-SOLUTION IS A SHRINKER
415
where 'l/J(O) 2: 0 is an arbitrarily smooth function with compact support in (A-I, A). For any smooth compactly supported
~) -loo(q,B)
_ fA f ({)f oo _ 0- JA-I JMoo {)O Rgoo
+ 20
fA f ({)foo 2: J A-I J Moo {)O - Rgoo
+ 20
e
n)
e
-loo(q,B)
(
Note that
fA f ({)foo JA-l JMoo {)O - Rgoo
n)
e
n)
e
+ 20
Hence
fA f ({)foo JA-I JMoo {)O - Rgoo
+ 20
-loo(qB) '
-loo(q B) ( ) _ '
for any smooth compactly supported
(8.52)
{)foo 2 {)O - I::1go,/oo + IV' goofool - Rgoo
n
+ 20 = O.
Hence foo (q, 0) is a smooth function by standard regularity theory for parabolic PDE (see G. Lieberman [255]' Chapters 5 and 6) and foo (q, 0) satisfies (8.52) in the classical sense. Let U oo ~ (47rO)-~ e- loo , 0* ~ :B - 1::1 + R goo ' and
Voo ~ (O(2I::1foo
-1V'f ooI2 + Rgoo) + foo - n) U oo ·
Equation (8.52) implies O*u oo = as used to obtain (6.22) to get (8.53)
o. Hence we can apply the same calculation
O*voo = -20 IRC (900)ij
+ V'iV'jfoo -
It follows from (8.47) and (8.52) that Voo hence
2 210 (900)ij I U oo ·
= o.
Thus O*voo
=0
and
(8.54)
i.e., 900(0) is a shrinking gradient Ricci soliton on Moo for 0 E (A-I, A) .
416
8. APPLICATIONS OF THE REDUCED DISTANCE
4.6. The limit is nonflat. We argue by contradiction. If goo (0) is flat for some 00, then Rg-x.(Bo) = 0 and we get from (8.54) that 1 '\h\ljR.oo (q, ( 0) = 200 (goo)ij (00).
(8.55)
Taking the trace of the above equation, we get /).goo(Bo)£oo (q, ( 0 ) ging this into
=
w' Plug-
we obtain (8.56)
It follows from Lemma 7.59 that (one estimates £i and takes the limit
as i
~
00) 1 2 nCoOo £00 (q, ( 0) ~ 40oe2CoBo dg"",(O) (pO, q) - - 3 -
so that £00 (q, ( 0 ) must have a minimum point PI E Moo. On the other hand, (8.55) implies that £00 (q, (0) is a strictly convex function on (Moo, goo(Oo)). It is well known that a strictly convex function has at most one critical point. 6 Hence PI is the only critical point of £ex,(-, ( 0 ) on Moo, so that M~ is diffeomorphic to ]Rn. Since, by assumption, (Moo,goo(Oo)) is a flat manifold, it is then isometric to (]Rn,!lIE) under
4~o
£00 (x,Oo) =
(x 1 )2 + ... +
Equation (8.56) implies (cr)2
4~o (xn)2 + crx 1 + ... + cnx n + Cn+l·
+ ... + (c n )2 = 00lcn+!
and hence
£00 (x, 1) = 4010 ( x 1 + 200Cl )2 + ... + 4010 (x n + 20ocn) 2 . From the definition of Voo(O)
= fM"" (47l'0)-~ e-£oo (q,B) dJ..L9oo (B) (q), we
compute Voo(Oo) = 1 by using the formulas for £00 (x, ( 0) and goo(Oo). This contradicts Voo(Oo) = V(oo) < 1 in (8.51). Hence goo(O) is not flat.
8.39. In dimension 3 this shrinking gradient soliton limzt has bounded sectional curvature. PROPOSITION
6Suppose foo (q, (0) has two different critical points P and Pl. Let 'Y (s) , 0 ::; s ::; so, be a minimal geodesic, with respect to goo (00), from P to Pl. We have by (8.55) that d
Is=so
0= d/oo('Y(s),Oo) s=o =
to
= 10 'iJ 2 foo which is a contradiction.
(SO d 2
10
ds 2foo C'Y(s),00)ds
to 201 b' (S)I~oc(lIo) ds > 0,
h' (s) ,'Y' (s)) ds = 10
5
PERELMAN'S FORMALISM IN POTENTIALLY INFINITE DIMENSIONS
417
PROOF. The sequence (M 3,9Ti(O),qT;)' 0 ~ 1, is a sequence of ancient ~-solutions with bounded curvature for each i. The convergence 9T i (0) 900(0) for 0 E [A-I, A] implies that R gTi (qT;' 1) - Rgoo (qoo, 1). Since the limit 900(0) is nonflat, by the strong maximum principle, Rgoo(l) (qoo) > O. By the compactness theorem of the set of 3-dimensional ancient ~-solutions (see Part II of this volume), we know that a subsequence
(M, R;T~(1,qTi)9Ti (0), qTi) converges to a limit (Moo, goo(O), iioo). The limit has bounded sectional curvature. Using the definition of Cheeger-Gromov convergence, it is easy to
(M,
check that R;~(1,qoo)9Ti (0), qT;) converges to the limit On the other hand it follows from Theorem 8.32 that
(M, R;~(1,qoc)9T; (0), qTi) -
(Moo, goo(O), iioo).
(Moo, R;~(1,qoo)9oo(O), qoo) .
Hence, by the uniqueness of the Cheeger-Gromov pointed limit, 900(0) = Rgoo(l,qoo)goo(O) for 0 E [1, A], which has bounded curvature. We have proved that the shrinking gradient soliton 900 (0) is complete and has nonflat bounded nonnegative curvature. 0 5. Perelman's Riemannian formalism in potentially infinite dimensions
Here we discuss Perelman's potentially infinite Riemannian metric in more detail. We discuss the calculation of its curvature tensor, the Ricci flatness (modulo renormalization) of its metric, and a geometric interpretation of Perelman's entropy formula. Although this section does not belong to the linear flow of this chapter, it provides a unifying viewpoint for various components of this volume. In particular, we have the following. (1) In Section 2.1 of Chapter 7 we considered the Riemannian metric 9 on M = M n x SN X (0, T) and showed that the renormalization of its Riemannian length is the C-Iength. 7 (2) In subsection 3.3 of Chapter 7 we showed that the C-geodesic equation is the same (up to the time reparametrization (T = 2JT) as the geodesic equation of the space-time connection defined by (7.39)(7.42), which is the limit, as N - 00, of the Levi-Civita connections of the Riemannian metrics 9 defined by (7.11) (see Exercise 7.4). Thus the C-geodesic equation is the limit, as N - 00, of the geodesic equations of the Riemannian metrics g. (3) In subsection 2.1 of this chapter we showed how to obtain the reduced volume functional from a renormalization of the volumes of geodesic balls with respect to the metric 9 on M. 7Here we have switched the notation from
it to 9 and N to M.
8. APPLICATIONS OF THE REDUCED DISTANCE
418
Recall from (7.11) that given a solution (Mn,9(T)) , T E (O,T), of the backward Ricci flow, Perelman [297] introduced the manifold M = M X SN X (0, T) with the following metric: gij
= 9ij, ga{3 = T9a{3, goo =
~ + R,
gia
= giQ = gao = 0,
i.e.,
9 = 9ij dx i dxi + T9a{3dy ady{3 + ( R + ~) dT2, where the metric 9a{3 on SN has constant sectional curvature 2kr. 5.1. Riemann curvature tensor of (M, g). We shall apply the following two steps to compute the Riemann curvature tensor of the manifold (M,g). First we treat (M,9) as a hypersurface in the manifold M = M x (0, T) with the metric
9 = 9ijdx i dxJ +
(R + ~) dT2.
We compute the curvature of the manifold (M, g) using the Gauss equations and Koszul's formula. Secondly, we consider the manifold (M, g) as a warped product with base (M, g) and fiber SN. We then use O'Neill's formulas to compute the curvature of g. This method of computation essentially follows Guofang Wei [368]. Let Oi ~ a~t denote the coordinate vector fields on the M factor, let Or ~ and let
t,
1I
=
1
(R+ ~)1/2
a l'
be the unit normal vector field of M x {T} C M. Direct computation, using [on oil = 0, gives
(8.57) By the formula for the evolution of the metric of a hypersurface evolving in the direction of its normal with speed lOr I = (R + ~) 1/2 , we have8 2~j
= Or9ij = 2101'1' II ij ,
where II denotes the second fundamental form of M x {T} C M. 9 Therefore
II(oi' 0)) = IIi) =
1 N
(R+ 21')
1/2~j,
and the Levi-Civita connections of 9 and 9 are related by 8See the proof of (B.13) with the mean curvature H replaced by laTI. 9In other parts of this volume we have sometimes used h instead of II to denote the second fundamental form.
5. PERELMAN'S FORMALISM IN POTENTIALLY INFINITE DIMENSIONS
V' aJ1j = V' aJ1j - II ijll = V' a/)j -
419
1 N -1/2 Rtjll. (R + 2r)
By the Gauss equations, we have
(R( Oi, OJ )Ok, Ot) = (R(Oi, OJ)Ok, Ot) - (II(oi, Ot), II(Oj, Ok)) + (II(Oj, Ot), II(oi, Ok)) 1 = (R(Oi,Oj)Ok,Ot) - ( N) (~tRjk - RjtRtk). R+ 2r By Koszul's formula for the Levi-Civita connection of g, i.e., 2 (V x Y, Z) = X (Y, Z)
+Y
+ ([X, Yl, Z) -
(X, Z) - Z (X, Y) ([X, Z], Y) - ([Y, Zl, X) ,
where the inner products are with respect to g, we have 1 (8.58) V' a,1I = N 1/2 Rc (Oi) , (R + 2r) 1 (8.59) V'v ll = - ( N) V'R. 2 R+ 2r To derive the two formulas above, we used (Vaill, II) = 0, 2 (Vaill,ok) = II (Oi,Ok) = (R+
r\ V'v ll, Ok ) =
-
(V vII, II) =
0, and
~) -1/2 Or9ik,
V'k R ([II, Ok], II) = - 2 (R + ~)
(these follow from Koszul's formula and (8.57)). Applying another covariant derivative to (8.58) and (8.59), we have
r
-
\V'a·V'vll,Oj ,
)=
r )= -\V'VV'aill,Oj
1 2(R+
fr)
2V'i RV' j R-
1 ( N)V'iV'jR, 2 R+ 2r
N)
1 ( 1 N (RteRtj-orRtj), 2 OrR -2:2 Rij+ 2(R+fr) T R+ 2r
-(V[ai,v]II,Oj)=-4(R 1 !:1.)2V'i RV' j R, + 2r where, to obtain the second formula, we used
(VaT (Rc (Oi)) , OJ) = orRtj -
(Rc (Oi) , VaT OJ) = orRtj - RuRtj,
(note that (VaTOj,ok) = ~Or (OJ, Ok) = Rjk), and the third formula follows from (8.57) and (8.59). Hence the curvatures in the normal direction are given by
420
8. APPLICATIONS OF THE REDUCED DISTANCE
(R(8i, //)//, 8j ) =
(R(8i, 8J)8k, //)
[(8TR - 2N2) Rij 2(R+ ~) T 1
2
+ -21V'iRV'jR]
+
(R~-tf:) (-8TRij-~V'iV'jR+RiiR£j),
=
(Vai Vaj 8k - VaJ Va;8k, //) 1
--------:-3/.,-2 (Rjk V'iR - Rik V'jR) 2 (R+ ~) 1 -
N
(R + 2T)
1/2
(V'iRjk - V'JRik) .
In terms of 8i and 8T , this says
(R(8i , 8T)8n 8j )
= 2 (R
~ -tf:) [(8TR -
+ ~ V'i R · V'J R ]
~V'iV'jR - 8T~j) ,
+ ( Ri£R£J (R(8i, 8j )8k•8T) =
:2 )~j
( 1 N) (Rjk V'iR - Rik V'jR) - (V'iRjk - V'jRik) .
2 R+ 2T Note that the curvatures of 9 are the components of Hamilton's matrix Harnack expression in the following sense:
(R(8i , 8j )8k, 8£) == ~jk£ mod O(N- 1 ), /-
)
1
1
\R(8i,8T)8n 8j == -8T~j - "2V'iV'jR+Rie R£j - 2TRij mod O(N
-1
)
1
== 6RiJ - "2V'iV'jR + 2~kjeRkf 1
1
- RieR£j - 2T ~j mod O(N- ), (R(8i ,8j )8k,8T) == V'iRjk - V'j~k mod O(N- 1 ). Taking the trace gives the entries of the trace Harnack expression: -
_
-1
Rc(8i , 8j ) = ~j mod O(N ), _ 1 -1 Rc(8i.8T) = 2V'j R mod O(N ),
_ 18R Rc(8n 8T) ="2 8t
1
= "2L\.R+ IRcl
2
mod O(N- 1 ).
Now consider (M, g) as a warped product:
M =B
xj2
F,
where the base manifold is (B, gB) = (M, g), the fiber is (F, gp) = (SN, go./3), and f = Fr. We use O'Neill's formula to compute the curvatures of (M, g).
5. PERELMAN'S FORMALISM IN POTENTIALLY INFINITE DIMENSIONS
421
LEMMA 8.40 (O'Neill). Let X, Y, Z be vector fields on B and let U, V, W be vector fields on F. Then
(1) R(U, V)W = K(U, V)W -IGI~ (9F(U, W)V - 9F(V, W)U), (2) R(X, V)Y
= -:}(V xG, Y)V = -:} Hessg (I) (X, Y)V,
(3) R(X, Y)V = R(V, W)X = 0, (4) R(X, V)W = R(X, W)V = f(V, W)VxG,
(5) R(X, Y)Z is the same on either B or M. Here, K denotes the curvature tensor of (F, 9F) and 1
G = gradf =
r;; (
2yT
R+
N) 8
T •
2T
Let {OQJ~=1 be a basis of tangent vectors on SN and 9a{3 ~ 9 (Oa,O{3)' By (3) we have
R(8i ,8j )Oa = 0,
R(8i ,8T )t9a = 0,
R(Oa,0{3)8i = 0,
R(Oa,0{3)8T = 0,
and hence
(R(8i, 8j )Oa, 8k) =0, (R(8i,8})t9a,8T ) =0, (R(8i,8})t9a,0{3) =0, (R(8i, 8T)Oa, 8j ) = 0, (R(8i, 8T)t9a, 8T) = 0, (R(8i, 8T)t9a, 0(3) = 0, (R(Octl 0(3)8i, O-y) = 0, (R(Oa, 0(3)8n O-y) = 0. By (1) we have -
R(Oa, 0(3)t9-y = K(Oa, 0(3)t9-y
-IGlg2 [9(0(i) 0-y)0{3 -
9(0{3, O-y)t9a],
and hence
(R(Oa, 0(3)t9-p 08) g = T (K(Oa, 0(3)t9-y, 08) 9 1
- 4 (R+ ~) [9(Oa, 0-y)9 (0{3 , 08) - 9(0{3,0-y)9(Oa,08)]
TR
= 2N (R +
== since IGI~ = 4T(R~if)'
°
~) (9{3-y9a8 - 9a-y9(38)
mod O(N- 1 ),
9 (0{3, 08)
=
T9 (0{3, 08) , and 1
(K(Oa,0{3)t9-y,08)g = 2N (9{3-y9a8 - 9a-y9(38)'
8. APPLICATIONS OF THE REDUCED DISTANCE
422
By (2) we have
R(Oi, (}o)Oj
1
= - JT Hessg ( JT)(Oi, OJ)Oo,
1 R(Oi, (}O)OT = - JT Hessg( JT)(Oi, OT )00' 1 R(OT, (}O)OT = - JT Hessg( JT)(on OT )00'
and hence
\R(Oi, (}o)Oj, (}{3) =
( 1 2 R+
!1../~"j90{3 2T
== 0 mod O(N- 1 ), \R(Oi,(}o)On(}{3)
= - ( 1 !1..)90{3'ViR
\ R(On (}o)on (}{3)
= - ( 1 N) (OTR + R) 90{3
4 R+ 2T == 0 mod O(N-l), 4 R+ 2T
T
== 0 mod O(N- 1 ). By (5) we have
= R(Oi, OJ)Ok, R( Oi, OT)OT = R( Oi, OT )on R(Oi, OJ)Ok
and hence
\R(Oi,Oj)Ok,oe) = (R(Oi,Oj)Ok,Oe) 1
= (R(Oi,Oj)Ok,Oe) - R+!1.. (RuRjk -
Rje~k)
2T
== (R(Oi,Oj)Ok,Oe) mod O(N- 1 ), \ R(Oi, OJ)Ok, OT)
= (R(Oi, OJ)Ok, OT) 1
= 2 (R + ~) (Rjk'Vi R - ~k 'VjR) - ('ViRjk - 'Vj~k) == -('ViRjk -
'Vj~k) mod O(N-l),
5. PERELMAN'S FORMALISM IN POTENTIALLY INFINITE DIMENSIONS
423
(R(Oi' OT )Or, OJ) = (R(Oi' OT)Or, OJ) = 2 (R
~ ~)
+ (Rtj
-
(( oTR -
~ViVjR -
2~2 )
Rij
+ ~ ViR· V j R)
OTRtj)
1
== - 2r Rtj + 6 g Rtj + 2~kjf.Rkf. 2 - 2ViVjR 1 - Rij mod 0 (N -1) .
The components of the curvature tensor of this metric coincide (modulo N- 1 ) with the components of Hamilton's matrix Harnack expression. 5.2. Ricci flatness of (M, g). Taking the trace of the above formulas for the curvature tensor yields ~ ) RC(Oi, OJ) = 9 kf./\ R(Oi' Ok)Of., OJ
+
=
+ /\ R(Oi' //)//, OJ )
~ga~ (R(Oi' ()a)8~, OJ)
1 2 [(OTR - 2N2) RC(Oi' OJ) 2(R+~) r
+
(R~ ~)
(-OTR ij -
+ ~ViRVjR]
~ViVjR+2Rtf.Rf.j)
== 0 mod O(N- 1 ), RC(Oi' OT) = gkf. (R(Oi' Ok)Of., OT) 1 . = ( N)RtVjR 2 R+ 2T == 0 mod O(N- 1 ), RC(Or,OT) = If. (R(OT' Ok)Of., OT) +
+ (R(Oi' //)//, OT) + ~ga~ (R(Oi' (}a)8~, OT)
+ (R(or,//)//,OT)
~ga~ (R(or, (}a)8~, OT)
1 4 (R + ~)
IVRI 2
== 0 mod O(N- 1 ), RC(Oi' (}a) = If. (R(Oi' Ok)Of., ()a) + (R(Oi' //)//, ()a) + ~g/'~ (R(Oi' ()/' )8~, (}a) =0,
8. APPLICATIONS OF THE REDUCED DISTANCE
424
Rc(8r , (Ja) = II (R(8r , 8k)8l, (Ja) +
+ (R(8r , 11)11, (Ja)
~g-Y/3 (R(8,., (J-y)(}/3, (Ja)
=0, Rc((Jo, (J/3)
= II (R((Ja, 8k)al, (J/3) + (R((Ja, 11)11, (J/3) + ~g-yt5 (R((Ja, (J-y)(Jt5, (J/3)
_ 1 (8 R _ 2R2) ga/3
-2(R+~)2
N
r
== 0 mod O(N-l). Hence all of the components of the Ricci tensor are equal to zero (modulo N- 1 ).
Finally, taking the trace again yields the scalar curvature of g:
R = gOO Roo + gtj Rij
=
2 (R
1
+ ~)
2
+ go/3 Ro/3
(-R8r R+ IVRI 2 _ R2) _ T
( 1 N) (6R+ R).
2 R
+ 2r
T
5.3. Geometric interpretation of Perelman's entropy integrand. Let (Mn, 9 (T)) be a solution to the backward Ricci flow and let 1 satisfy (6.15). In §6 of [297] Perelman also gave a geometric interpretation of the integrand (6.20), i.e., T
(2fll -IV11 2 +R) + 1 -
n,
of his entropy W (g, I, T) . In this subsection we discuss this interpretation. Define the diffeomorphism
'P
= 'PJ,N : M
--t
M
by
'PJ,N:
(X,y,T) ~ (x,y,
(1- ~) T).
Clearly, limN --+00 'P J,N = idM' i.e., for N large, 'P J,N is close to the identity. Consider the pulled-back metric
gm ~ ('PJ,N)* g, which we think of as a perturbation of the metric g. By definition,
gm (X, y)
=
9 (( 'P J,N )* X, ('P J,N ) * y) .
We first compute the components of the metric
gm.
5. PERELMAN'S FORMALISM IN POTENTIALLY INFINITE DIMENSIONS LEMMA
where
425
8.41. 2 -0f - -f
+ 0 (N- 1 ) aT T = N + R _ 2 of _ £ + 0 (N-l) , 2T aT T
(1)
-m 900
(2)
9iG =
(3)
9il = 9ij
(4)
9~/j =
(5)
9~
(6)
9~0 = 0,
= 900 -
-
:!i + +0
0 (N-l) ,
(N-l) = 9ij
+0
(N-l) ,
(1- ~) 9a/j,
= 0,
tr 9ij = 2Rtj.
°
(1 - ¥)
Let f = f (x, T) ~ T. In the formulas below, denotes the time index; i, j, k, ... denote indices on M; a, (3, " ... denote indices on SN; and a, b, c, ... denote arbitrary indices. The pulled-back metric is given by PROOF.
9;;b (x, y, T) =
~~: (x, y, T) ~~: (x, y, T) 9cd (cp (x, y, T)),
where z = x, y, or T. Using the formulas for 9ij, 9a/j, 900, and 9ia = 0, we obtain the following.
=
9iQ
9ao (1)
900 (x, y, T)
=
(~O (x, y, T) ) 2900 (cp (x, y, T) )
(0-)2 900 (x, y, f)
= o~
(1- 2fN _ 2TOf)2 (N (1- 2f)-1 +R) NOT 2T N = (1 + 2(-~ - ~ ~~) + 0(N-
=
2 ))
X (
R+
~ (1 + ~) + 0 (N-l) )
N ( 1+2f) +-2 N (2f =R+- - -2TOf) - - +O(N -1 ) 2T N 2T N NOT of = R + -N - -f - 2+ 0 (N -1 ) 2T T aT - (x, y, T ) - -f - 2of = 900 - + 0 (-1) N . T
0T
=
426
8. APPLICATIONS OF THE REDUCED DISTANCE
(2)
a-c a- o 9iO(x,y,r) = a'P· (x,y,r) a'P (x,y,r)gco(
a- o
= a'P· (x,y,r) a'P (x,y,r)900(x,y,f) xt r
=
_~ af.t r N ax
==-
f ) (N (1- 2f)-1 +R) (1- 2fN _ 2ra N ar 2r N
~ :~ r (1 + 0 (N- 1 )) (~ + 0 (1) )
::i
(x,r) +0 (N- 1 ).
(3)
a-c
a- d
a- k
a-I.
9iJ· (x,y,r) = a'P· xJ (x,y,r)gcd(
= a'P· xt (x, y, r) a'P· xJ (x, y, r)gkl.(x, y, f)
a-0
a- o
+ a'P· xt (x, y, r) a'P· xl (x, y, r) 900 (x, y, f) _ _ = 9ij (x, y, r) = 9ij (x,r)
+
(2r)2 af af N axi axj
(N2r (1 -
2f)-1 N
+R
+ 0 (N- 1).
(4)
a-c
a- d
9~,B (x, y, r) = a;a (x, y, r) a;,B (x, y, r)gcd (
- ( -) (1
=9a,B x,y,r =
f N - (x,y,r ) . - 2 (x,r)) 9a,B
(5)
9fc! (x, y, r) = aa
a- o
+ a'P· xt
(x, y, r)gOa(x, y, f)
)
5. PERELMAN'S FORMALISM IN POTENTIALLY INFINITE DIMENSIONS
427
(6)
a- c
a- d
9:;0 = a;a (x,y,r) ;
a- o
= ;
(x,y,r)9cd(
(x, y, r) 9aO(X, y, f)
=0.
o Now define the I-parameter family of diffeomorphisms 'ljJT:M~M
by
a
ar'ljJT = (V f) ('ljJT (x), r). Consider the diffeomorphism
defined by ~ (x, y, r) ~ ('ljJT (x) , y, r)
and the diffeomorphism defined by Let 9
m . ,T.*-m =;='J!
9 ,
which is a Riemannian metric on M. We compute the components of the metric gm. First note that
2'ljJ; (Hess gf) = 'ljJ; (Cgradg Ig) = C./'l'r.• (grad f ) ('ljJ;g) g
= Cgrad",;g(foliT) ('ljJ;g) = 2Hess,p~g
(J o~).
LEMMA 8.42. Let (Mn,g (r)) be a solution of the backward Ricci flow and let f satisfy (6.15)
af = fl.f -IVfI 2 + R -~. ar 2r The (spatial) components gij satisfy the following:
+ 2VVf))ij mod O(N- 1 ) == 2 Rc(gmIMx{y}x{T})ij + 2ViVjf 0 ~ mod
:r gij == ('ljJ;(2Rc(g)
O(N- 1 ),
428
8. APPLICATIONS OF THE REDUCED DISTANCE
where vm is the covariant derivative associated to the metric giJ on M. The other components 01 gm are given by
goo = 900 - IV112 = N _ 2(~1 -IV/1 2+ R -~) 27
=
g:{3 =
giQ =
L -IV/1 2+ 0 (N-l) 7
~ (~ - [7(2~/-IV/12+R)+I-nl) +O(N- 1 ),
(1 - ~)
9~ =
-V(Vh).(8i)!
g:;O = g~ = PROOF.
27
9a{3,
+ 0 (N-l) ,
o.
We have
Hence
o m 07gij
=
0 ('l/JTg *) ij 07
= ('I/J;
+ O( N -1)
(~~))ij + (C(1Pr).(V'f)('I/J;g))ij+O(N-
1
)
=='I/J;(2Rc(g))ij +2ViVjloW mod O(N- 1 )
== 2 Rc(gmIMx{y}x{T})ij + 2ViVj loW mod O(N- 1 ). We compute
goo =
9m (q, * (OT) , q, * (OT) ) = 9m ((V 1,0,1), (V 1,0,1)) = IV /12 - 21V /12 + 900 = 900 -IV/12.
5. PERELMAN'S FORMALISM IN POTENTIALLY INFINITE DIMENSIONS
429
Using (6.15), we find that
goo = goo -IVfI 2 = ~ (~ - [T(2~f -IVfI 2 + R) + f
-
nl) + 0 (N-1) ,
g~ = gm (q,* (8aJ, q,* (8{3)) = gm (8a, 8(3) = g;:{3 (7/JT (x) , y, T) =
gill =
(1 - q,) 2f;
ga{3(7/JT(X), y, T),
gm(q,* (8i ) , q,* (8T)) = gm((7/JT)* (8i ) ,8T)
= -VC1PT)*C8i)f
+ O(N- 1 ),
where 8a ~ 8~o. and 8{3 ~ -/yp. We leave the proofs of the formulas for the rest of components to the reader as exercises. 0
MI
Let 'He C M denote the hypersurface {(x, y, T) E T = c} ('He is simply a time-slice) with the metric induced by gm, where c is a constant. By the definition of the metric gm, we have the following: df-L1t c
= df-Lgrn'J /\ df-Lgmo.{3
N
Using (1-~) 2' = e- f the hypersurface is df-L1t c
+ O(N- 1 ),
we see that the volume form df-L1t c of
= Tlf (e-foq, (7/J;df-LCM,g)) /\ df-LsN + O(N- 1)) .
To find the scalar curvature of the metric on 'He induced by the metric gm, first we calculate the Ricci tensor Rc of the metric on 'He induced by the metric gm. Second, we pull back the tensor Rc by the diffeomorphism induced by q,. LEMMA 8.43. The Christoffel symbols f~b corresponding to the metric on 'He induced by the metric gm are given by
mod 0 (N- 1) ,
_ k r ijk = rij(M,g) A
Aa r ij
ft?a = ~{J
Ak
= 0,
-
ria
t5~Vd N
_ Vk f rAka {3 = Tga {3N
r a{3 = r a{3(S N ), A"(
"(
= 0,
mod 0
(N-2) ' ( -2) ,
mod 0 N
430
8. APPLICATIONS OF THE REDUCED DISTANCE
where rt(M, 9) and r~,8(SN) denote the Christoffel symbols of (M, 9) and SN, respectively. PROOF.
Recall that the induced metric gm
~
gml 1tc on ftc is given by
'9ij (x, y, T) = gij (x, y, T) + 0 (N-l) = 9ij (x, T) + 0 (N- 1 ) , -m( X,y,T ) = ( 1- 2f(x,T))_ 901.,8 N 901.,8 ( X,y,T ) , g~(X,y,T)=O.
We compute the Christoffel symbols r~b as follows (where p denotes an index in the M and SN directions, i, j, k, £ denote indices in the M direction only, and ct, f3, 'Y, 8 denote indices in the SN direction only):
rk - 21 (-m)kp(8' -m + 8.J9ip -m 9 19jp ij -
8P9ij -m)
1 (-m)kl(!l -m !l -m !l -m) =2 9 vi9jl + Vj9il - Vl9ij == r~j(M,9) mod 0 (N-l) ,
rOi.ij -- 2 1 (-m)OI.p(8-m 8. -m 8 -m) 9 19Jp + J9ip - P9ij 1 (-m)OI.o(8 -m + 8j9io -m - 809ij -m) =2 9 i9jo =0,
rOi.i,8 = 2 1 (-m)OI.p(8 -m -m - ) 9 i9,8p + 8,89ip - 8p9i,8 =
-m 8 -m 9 21-01.1'(8 i9,81' + /39i-y -
81'9i,8 -m)
8$\1d 2 = --r;r- mod 0 (N- ),
l-0I.1'!l -m _
= 29
vi9,81'
rkiOi. -- 21 (-m)k 9
-m + 80I.9ip -m - 8P9iOl. -m) 1901.p
p (8-
1 (-m)kl(8 -m 8 -m 8 -m) =2 9 i901.l + 0I.9il - 'l9iOl. =0, p (8 -m 8 -m rk01.,8 -- 21 (-m)k 9 0I.9,8p + /3901.p -
8p901.,8 -m )
1 (-m)kl(8 -m 8 -m 8 -m ) =2 9 0I.9,8l + /3901.l - 1901.,8 _ \1kf -2 = T901.,8N mod 0 (N ),
5. PERELMAN'S FORMALISM IN POTENTIALLY INFINITE DIMENSIONS
r'Ya(3 --
431
-m + 8(39ap -m - a.p9a(3 -m) "21 (-m)'Yp(8 9 a9(3p
-m + 8(39ao -m - 8o9a(3 -m) = "21 (-m)'Yo(8 9 a9(3o = r:(3(sn).
D
Therefore the components of the Ricci tensor given by
/lpq
of
(Hc, gmlr£J
are
and
(Note that Rc (SN)a(3 = ~!./9a(3.) Thus, the scalar curvature hypersurface
(Hc, 9mlr£J ' where 9m = ~*gm,
is (denote
R of the
gm =l= gmlr£J
8. APPLICATIONS OF THE REDUCED DISTANCE
432
so that .
A
R
_
=
.
n
kf. 8x~ 8x] ~ 9 8'ljJk 8'ljJf. ~ (Rpq T T p,q=l
+ (1 _
p
~1-1 g: (Rc (sN).~ _ Tg"~~V112 + T9"~ll./)
( 2f)-1 9r
== 2b.f - 1\/ fl2 + R + 1 - N == 2b.f -1\/fI2 + R+ ==
q
8'ljJT 8'ljJT + \/p\/qf) 8xi 8x j
2~
a{3
mod O( N- 1)
N-1
~ga{3 mod O(N- l )
(1 + ~) (N -1) mod O(N- 1 )
~ (N; 1 + r(2b.f -1\/fI2 + R) + f)
mod O(N- l ).
This is the same as the formula on p. 13 of [297] except for the -1 in Nil. If we instead choose the metric 9 on SN so that Ra{3(SN) = !ga{3, then we would obtain the exact formula. The integrand for the entropy W (g,J, r) is related to the scalar curvature of the hypersurface (ftc, gml?-£J by the following formula:
r(2b.f-l\/fI2+R)+f-n=rR-n-N;1 mod O(N-l). 6. Notes and commentary
Section 2. Corollary 8.17 is from §7.1 of [297]. Section 3. Theorem 8.26 is from §7.3 of [297]. A localized version of the theorem is given by Perelman in §8.2 of [297]. Section 4. Theorem 8.26 is from §11.2 of [297].
CHAPTER 9
Basic Topology of 3-Manifolds The purpose of this chapter is to introduce certain well-known facts in 3-manifold topology which are related to the Ricci flow. It is by no means meant to be a complete survey of the subject, and we have omitted many important results in the field. Due to space limitations, we will not provide proofs and instead will refer the reader to the literature. Unless mentioned otherwise, we shall assume in this chapter that all 3manifolds are connected, orientable, and with possibly nonempty boundary. The n-dimensional sphere, n-dimensional ball and n-dimensional Euclidean space are denoted by sn, B n , and ]Rn, respectively. 1. Essential 2-spheres and irreducible 3-manifolds
1.1. Topological, PL and smooth categories. The fundamental work of Moise [268] in 1952 shows that any topological homeomorphism of an open set in ]R3 into ]R3 can be CO approximated by piecewise-linear (PL) homeomorphisms. As a consequence, he proved that any topological 3manifold can be triangulated and that there is a unique PL structure on any topological 3-manifold. Different proofs of the Moise theorem can be found in Bing [28] and Shalen [327]. It is shown in Hirsch [203] and Munkres [279] that the PL and smooth categories in dimension 3 are equivalent. For the rest of this chapter, we assume that all manifolds and maps between them are smooth. 1.2. Sphere decompositions and irreducibility. The study of 2spheres in 3-manifolds probably began with the Schoenflies problem. It asks if any smoothly embedded 2-sphere in Euclidean 3-space must bound a 3-ball. The affirmative solution of the problem in dimension 3 by Alexander [1] is one of the milestones in the field. We say a 2-sphere in a 3-manifold is essential if it does not bound a 3-ball. Essential 2-spheres are closely related to the connected sum decomposition. Indeed, if a 3-manifold M is a connected sum M = Ml#M2, where neither Ml nor M2 is S3 or B 3 , then the decomposing 2-sphere is essential. On the other hand, if S2 is an essential 2-sphere in M which decomposes the manifold into two pieces Nl and N 2 , i.e., S2 is separating, then this gives a connected sum decomposition M = Ml #M2. Namely, we simply take Mi to be M capped off by a 3-ball. If the essential2-sphere S2 does not separate M, i.e., M\S2 is connected, then 433
434
9. BASIC TOPOLOGY OF 3-MANIFOLDS
one can again obtain a connected sum decomposition M = (S2 X SI)#M'. This gives more information than a separating 2-sphere. Here is a way to see the S2 x SI factor in M. Take an embedded arc A whose endpoints lie in S2 in such a way that A starts from one side of S2, ends at the other side, and has no other intersection with S2. We may assume that, after an isotopy, the endpoints of A are the same. Thus there is an embedded SI in M which intersects the 2-sphere transversely in one point. Let N be a small regular neighborhood of S2 U SI. Then it is easy to see that N is homeomorphic to (S2 X S 1 ) \ { open 3-ball }. In particular, this shows that the boundary of N is a 2-sphere. This gives the connected sum decomposition. (To see the topology of N, the reader may try to consider the corresponding problem in dimension two: replace S2 by SI inside a surface. In this case, we have two simple loops a and b inside an oriented surface so that a intersects b in one point transversely. Then it is an elementary exercise in topology to show that the regular neighborhood N(aUb) is (SI x SI)\{ open 2-disk }. It turns out that this fact holds in all dimensions. ) Thus essential 2-spheres correspond to connected sum decompositions. A 3-manifold is called irreducible if each embedded 2-sphere bounds a 3ball. If a 3-manifold M is not irreducible, it contains an essential 2-sphere. Using the operations above, one concludes that either M = S2 X SI or M = Ml#M2, where neither Ml nor M2 is S3 or B3. Now one asks if each of the factor 3-manifolds is irreducible or not. Continuing in this way, one bumps into the question of whether this decomposition process for a compact 3-manifold stops after finitely many steps. This was resolved affirmatively by Kneser in 1929 [233] for compact triangulated 3-manifolds. THEOREM 9.1 (Kneser). Let M3 be a compact triangulated 3-manifold. Then M can be decomposed into a connected sum
M~#M~#··· #M~#(S2 x SI)# ... #(S2 x SI),
where each Mi is irreducible.
A counterexample to the Poincare conjecture is usually called a 'fake' 3-sphere. An interesting consequence of Kneser's Theorem is the following statement: if there exists a counterexample to the Poincare conjecture in dimension 3, then there exists an irreducible fake 3-sphere. This holds because the fundamental group of a connected sum is the free product of the fundamental groups of the factors. One calls two 2-spheres in a 3-manifold parallel if they are disjoint and bound a region homeomorphic to S2 x [0,1]. Kneser's theorem states that for any compact triangulated 3-manifold, there is an integer k such that any collection of more than k disjoint essential 2-spheres in the manifold must contain a pair of parallel 2-spheres. Since Kneser's finiteness theorem is so deeply related to Hamilton's program, we will indicate the basic ideas of its proof here. (See [201] for a complete proof. Note too that the proof
2. INCOMPRESSIBLE SURFACES AND GEOMETRIZATION CONJECTURE
435
was generalized by W. Haken to normal surface theory [176].) Let us fix a triangulation T of a compact 3-manifold M. Let t be the number of tetrahedra in the triangulation. Kneser proved that if n > 6t + 1HI (M; Z2) I, where H l (M;Z2) is the first homology group of M with Z2 coefficients, then any n disjoint essential 2-spheres contain a parallel pair. Here is the basic argument. Suppose {SI, ... ,Sn} is a collection of n essential disjoint 2-spheres in M. By isotopy and topological surgeries, one may find a new collection of n essential disjoint 2-spheres, still denoted by {SI, ... , Sn}, that are in 'nice position' with respect to the triangulation. Here, 'nice position' means that the intersection of each 2-sphere with each tetrahedron consists of a disjoint union of geometric triangles and quadrilaterals. These are called normal surfaces. There are only seven normally isotopic triangles and quadrilaterals in a tetrahedron. This shows that inside a tetrahedron (73, all but at most six components of (73 \ (SI U ... U Sn) are parallel regions. It follows from this fact by a simple computation that there are two parallel 2-spheres if n > 6t + 1HI (M; Z2) I· 1.3. Irreducible 3-manifolds. Irreducible 3-manifolds joined in connected sums may be regarded as building blocks for 3-manifolds. Most familiar 3-manifolds are irreducible. For instance, the complement of a knot in S3 is irreducible by Alexander's theorem. Also, if a covering space N of a 3-manifold M is irreducible, then M is irreducible. This is due to the fact that 2-spheres are simply connected. Thus any 2-sphere in M can be lifted to a 2-sphere in N. Now using the irreducibility of N, one produces a 3-ball in N bounding the lifted 2-sphere. Using the Brouwer fixed point theorem, one then shows that the 3-ball is mapped injectively to M by the covering map. This proves M is irreducible. In particular, if the universal cover of a 3-manifold is R3, then the manifold is irreducible. This shows, for instance, that all flat (e.g., SI x SI X SI) and all hyperbolic 3-manifolds are irreducible. A deep result of Meeks and Yau [263] shows that the converse is also true. Namely, if a manifold is irreducible, then all covering spaces of it are irreducible. In [264] Milnor proved that the connected sum decomposition of a compact 3-manifold is unique up to self-homeomorphism of the 3-manifold. But the decomposition in Kneser's theorem is in general not unique up to isotopy of the 3-manifold. This is due to the action of the diffeomorphism group of the 3-manifold on the decomposition. For instance, the manifold (S2 x SI )#(S2 X SI) has many nonisotopic essential separating 2-spheres, due to the large diffeomorphism group of the manifold. 2. Incompressible surfaces and the geometrization conjecture 2.1. Incompressible surfaces and Haken manifolds. The success of the study of 2-spheres in 3-manifolds prompted people to look for more general surfaces. Evidently, surfaces that can be contained inside a coordinate chart are not going to be interesting. Haken introduced the following
436
9. BASIC TOPOLOGY OF 3-MANIFOLDS
important concept. A compact, connected, properly embedded surface F in a 3-manifold M is said to be incompressible if one of the following conditions holds:
(1) F i= 52 or B2 and the inclusion map induces an injective homomorphism in the fundamental groupj or (2) F = 52 is an essential 2-spherej or (3) F = B2 and 8F is not null homotopic in 8M. An orientable Haken 3-manifold is a compact and irreducible 3-manifold M that admits a two-sided incompressible surface. This definition is the same as saying that M is a compact, orient able , and irreducible manifold which contains an incompressible surface other than lRJID2 • The reason is that if a compact, irreducible, orientable 3-manifold M contains JRJP>2 as an incompressible surface, then M = JRJP>3. Also, if a compact surface is incompressible in M and is one-sided, then the boundary of a regular neighborhood of the surface is a two-sided incompressible surface. Furthermore, since we assume the 3-manifold is orientable, two-sided surfaces are the same as orient able surfaces. Haken manifolds constitute a huge portion of all of 3-manifolds. For instance, if a compact orientable 3-manifold M is irreducible and has nonempty boundary, then M is Haken. Also, if a closed, irreducible 3-manifold has positive first Betti number, it is Haken. The homeomorphism classification of Haken manifolds is considered to be solved. It is due to the deep work of F. Waldhausen [363] in 1968. Among the many results he proved, the following stands out as one of the most striking. THEOREM 9.2 (Waldhausen). Two homotopically equivalent closed Haken manifolds are homeomorphic. 2.2. Torus decompositions and the geometrization conjecture. A Seifert 3-manifold (also called a Seifert space) is a compact 3-manifold admitting a foliation whose leaves are 51. Although this was not the original definition by Seifert in 1931, subsequent work of Epstein [136] shows that this simpler definition is equivalent to Seifert's original formulation. If a compact 3-manifold admits an 51 action without global fixed points (i.e., no point is fixed by all elements in 51), then the manifold is a Seifert space. Seifert 3-manifolds have been classified. In particular, there exist Seifert manifolds which are irreducible, non-Haken 3-manifolds and have infinite fundamental group. In 1976, Thurston constructed closed hyperbolic 3manifolds that are not Haken. Suppose M is a closed, irreducible, and orient able 3-manifold. A natural step after the connected sum decomposition is to look for incompressible tori. This is called the torus decomposition. A compact, irreducible 3-manifold is called geometrically atoroidal if every incompressible torus is isotopic to a boundary component. (If the manifold is closed, this simply means that there are no incompressible tori.) The torus decomposition theorem of Jaco
2.
INCOMPRESSIBLE SURFACES AND GEOMETRIZATION CONJECTURE
437
and Shalen [224] and Johannson [225] says the following. (For simplicity, we state the result for closed manifolds only.) THEOREM 9.3 (Jaco and Shalen, Johannson). Suppose M is a closed, orientable, and irreducible 3-manifold. Then there exzsts a possibly empty disjoint union of incompressible tori in M that decomposes M into pieces which are either Seifert 3-manifolds or geometrically atoroidal 3-manifolds. Furthermore, the minimal such collection of tori is unique up to isotopies. Thurston's work on the geometrization of 3-manifolds addresses the geometries underlying the Seifert pieces and the geometrically atoroidal pieces. Thurston proved that if N is a non-Seifert, geometrically atoroidal manifold appearing in the decomposition above (so that it has nonempty boundary), then the interior of the manifold N admits a complete hyperbolic metric of finite volume. Also, it is proved that the interior of any compact Seifert 3-manifold admits a complete, locally homogeneous Riemannian metric of finite volume. Thus, the remaining issue is the geometry of a closed, irreducible, geometrically atoroidal 3-manifold. The geometrization conjecture of Thurston for a closed, irreducible, orientable 3-manifold M states that there is an embedding of a (possibly empty) disjoint union of incompressible tori in M such that every component of the complement is either a Seifert space or else admits a complete Riemannian metric of constant curvature and finite volume. Using Thurston's theorem and the torus decomposition theorem of Jaco and Shalen and of Johannson, one can reduce the geometrization conjecture to the following form. Suppose M is a closed, irreducible, orientable 3manifold without any incompressible tori. Conjecture I: If the fundamental group of M is infinite and does not contain any subgroup isomorphic to Z EEl Z, then M admits a hyperbolic metric. Conjecture II: If the fundamental group of M contains a subgroup isomorphic to Z EEl Z, then M is a Seifert space. Conjecture III: If the fundamental group of M is finite, then M admits a spherical (constant positive curvature) metric. Topologists have made great progress toward resolving these conjectures. First of all, if the manifold is Haken, Conjecture I was shown to be valid by Thurston. (See also McMullen [262] and Otal [294].) Conjecture II for Haken manifolds was shown to be valid by the work of Gordon and Heil [158], Johannson [225]' Jaco and Shalen [224], Scott [318] and Waldhausen [364]. Conjecture II for non-Haken manifolds was solved affirmatively by Casson and Jungreis [61] and Gabai [149] in 1992. Furthermore, Gabai, Myerhoff, and Thurston [150] proved that if a closed, irreducible 3-manifold M is homotopic to a hyperbolic 3-manifold N, then M is homeomorphic to N. This gives evidence that Conjecture I holds. Note that Conjecture III implies the Poincare conjecture. Indeed, if M is a simply-connected 3-manifold, then by Kneser's theorem, we may
438
9 BASIC TOPOLOGY OF 3-MANIFOLDS
assume that M is irreducible. Since M has a trivial fundamental group, it is geometrically atoroidal. (By definition, if a manifold contains an incompressible torus, its fundamental group must contain Z EB Z, the fundamental group of the torus.) Thus, if Conjecture III holds, M admits a metric of constant positive sectional curvature. Thus M must be S3. 2.3. Examples of 3-manifolds and their geometries. There are eight locally homogeneous Riemannian geometries in dimension 3. Besides the well-known constant curvature metrics S3, JR3, and '}-£3, the rest of the five geometries are given by the standard metrics on S2 xJR, '}-£2 xJR, SL(2, JR), nil, and sol. (The Ricci flow of homogeneous metrics on S3 is discussed in Section 5 of Chapter 1 in Volume One.) H~e is a description of the three non product geometries. The geometry SL(2, JR) may be regarded as the universal cover of the Lie group SL(2, JR) with a metric invariant under left multiplication. The nilpotent 3dimensional Lie group nil is the Heisenberg group of strictly upper-triangular 3 x 3 matrices. (The Ricci flow on nil is discussed in Section 7 of Chapter 1 in Volume One.) The geometry sol is that of the 3-dimensional solvable Lie group defined as the semi-direct product of JR2 with JR, where the action of
t E JR on JR2 is given by the matrix
(~ e~t).
(The Ricci flow on sol is
discussed in Section 7 of Chapter 1 in Volume One.) A closed, orient able manifold with S2 x JR geometry must be either S2 x Sl or lRJP>3#JRF. It is interesting to note that JRF#JRIP3 is the only connected-sum 3-manifold admitting a locally homogeneous metric. The product of a closed surface of genus at least two with the circle admits an '}-£2 x Sl metric, i.e., the product of their standard metrics given by the uniformization theorem. The unit tangen!..bundle over a surface of negative Euler characteristic is a 3-manifold with SL(2, JR) geometry. More generally, any circle bundle over a surface of negative Euler characteristic admits the SL(2, JR) geometry if the Chern class of the bundle is nonzero. A circle bundle over the torus with nonzero Chern class is a 3-manifold with nil geometry. A torus bundle over the circle whose monodromy is a linear map with distinct real eigenvalues has sol geometry. It can be shown that any closed 3-manifold with one of these five geometries is finitely covered by one of the examples just mentioned. For more information, see the excellent reference [319]. A closed 3-manifold with sol geometry is not a Seifert 3-manifold. It admits a nontrivial torus decomposition coming from the torus fiber. On the other hand, any manifold admitting one of the six geometries S3, JR3, S2 X JR, '}-£2 X JR, SL(2, JR), or nil is a Seifert manifold. We will say a compact 3-manifold M is a topological graph manifold if it admits a torus decomposition such that each complementary piece is a Seifert manifold. (This is a standard definition in low-dimensional topology.)
3. DECOMPOSITION THEOREMS AND THE RICCI FLOW
439
In particular, this implies that the boundary of M is a possibly empty collection of tori. Graph manifolds appear in the work of Cheeger and Gromov as those 3-manifolds admitting an oF-structure. Note that the definition of graph manifold that appears in Cheeger and Gromov's work [74] is slightly different from the one above. To be more precise, a graph manifold in the sense of Cheeger-Gromov is a closed 3-manifold admitting a decomposition by (not necessarily incompressible) tori such that each complementary piece is a Seifert space. It can be shown that if M is a graph manifold in the sense of Cheeger-Gromov, then it is either a topological graph manifold or else a connected sum of topological graph manifolds with 8 2 x 8 1 factors and lens spaces. At any rate, there are no fake 3-spheres or fake 3-balls embedded inside a Cheeger-Gromov graph manifold. 3. Decomposition theorems and the Ricci flow Recall that a solution (M3, g( t)) , t E [0, T), to the Ricci flow is said to develop a singularity at time T E (0,00) if the norm of the Riemann curvature tensor becomes infinite at some point or points of the manifold as t / T. (See Corollary 7.2 of Volume One.) A typical situation in which a finite time singularity develops is the neckpinch. It is important to note that the formation of neckpinch singularities may be triggered more by the (local) nonlinearity of the Ricci flow PDE than by the (global) topology of the underlying manifold. In any case, here is a heuristic description. (For precise statements, see Section 5 of Chapter 2 in Volume One, as well as the recent papers of Angenent and one of the authors [7, 8].) Suppose a 3-manifold M contains a separating 2-sphere. Then under the Ricci flow, a region homeomorphic to 8 2 x lR may develop in M such that, as t / T, the sectional curvatures become infinite precisely along the hypersurface identified with 8 2 x {O}. In this evolution, the geometry of the region identified topologically with 8 2 x lR asymptotically approaches the cylinder 8 2 x lR with its standard product metric. Hamilton developed a program of applying Ricci flow techniques to general 3-manifolds and analyzed the singularities which may arise (see especially [186], [189], and [190]).1 Some of Hamilton's ideas are as follows; we first consider [186]. Via point-picking arguments and assuming an injectivity radius estimate, one dilates about singularities and takes limits using the Cheeger-Gromov-type compactness theorem for solutions of the Ricci flow to obtain so-called singularity models, which are nonflat ancient solutions of the Ricci flow. In dimension 3 these ancient solutions have nonnegative lSome other papers in which Hamilton developed his program to approach Thurston's geometrization conjecture by Ricci flow are as follows: characterizing spherical space forms [178], weak and strong maximum principles for systems [179]' ancient 2-dimensional solutions and surface entropy monotonicity [180] (as used in [186]), matrix Harnack estimate [181] and its applications to eternal solutions [182], and the compactness theorem [187]. (These are only partial descriptions that reflect aspects of the papers' relevance to Hamilton's 3-manifold program.)
440
9. BASIC TOPOLOGY OF 3-MANIFOLDS
sectional curvature. By the strong maximum principle, the universal covers of the ancient solutions either split as the product of a surface solution with ~ or have positive sectional curvature in which case they are diffeomorphic to either 8 3 or ~3.2 In the case of splitting, Hamilton proved that the ancient surface solution is either a round shrinking 8 2 (and the universal cover of the singularity model is hence geometrically a shrinking round product cylinder; by definition we say that in this case a neck singularity forms) or it has a backward limit which is the cigar soliton. Note that Perelman's no local collapsing theorem rules out the last case of a cigar. In the case when the universal cover of the singularity model has positive sectional curvature and is diffeomorphic to ~3, the covering is trivial. This ancient solution is either Type I or has backward limit which is a steady Ricci soliton on a topological ~3. In the latter case, the asymptotic scalar curvature ratio is infinite, and by dimension reduction, there exists a sequence of points tending to spatial infinity for which the corresponding dilations of the solution limit to an ancient product solution, which again must be a shrinking round cylinder.3 In this case the singularity model is expected to be the positively curved and rotationally symmetric Bryant soliton and the forming singularity is expected to be a degenerate neckpinch. In summary, we should have that at the largest curvature scale the dilations yield the Bryant soliton, whereas at lower scales dilations yield round product cylinders. This agrees with the fact that the dimension reduction of the Bryant soliton, and more generally a 3-dimensional gradient steady soliton with positive curvature which is /'i,noncollapsed at all scales, is a round product cylinder. On the other hand, the former case of a Type I ancient solution with positive sectional curvature, if it exists, also dimension reduces to a round cylinder. Thus a consequence of Hamilton's 3-dimensional singularity formation theory and Perelman's no local collapsing theorem is that if a finite time singularity forms on a closed 3-manifold, then either M is diffeomorphic to a spherical space form or a neckpinch forms. In [186] Hamilton studies 3-dimensional singularity formation by considering regions in the solution where the scalar curvature is comparable to its spatial maximum. He studies the regions where the scalar curvature is not comparable to its spatial maximum by the technique of dimension reduction. For example, when a 3-dimensional steady Ricci soliton singularity model forms, Hamilton proved an injectivity radius estimate to obtain a second limit which is either a shrinking round product cylinder or the product of a cigar with R (Again no local collapsing rules out the latter case.) The 2When the universal cover of the singularity model has positive sectional curvature and is diffeomorphic to 8 3 , M admits a metric with positive sectional curvature (e.g., 9 (t) for t large enough) and hence is topologically diffeomorphic to a spherical space form. In this case the singularity model must be geometrically a shrinking spherical space form with its underlying manifold diffeomorphic to M. 3With the help of no local collapsing.
3. DECOMPOSITION THEOREMS AND THE RICCI FLOW
441
regions with curvature comparable to their spatial maximum are geometrically close to a 3-dimensional steady Ricci soliton, whereas some regions with curvature not comparable to their spatial maximum are geometrically close to a shrinking round product cylinder. In [189] Hamilton developed surgery theory and formulated a version of Ricci flow with surgery. Although this theory was developed for solutions on closed 4-manifolds with positive isotropic curvature, its higher aim was clearly a surgery theory for 3-manifolds. Indeed, the class of 4-manifolds with positive isotropic curvature is flexible enough to allow for connected sums. Many of the techniques developed in [189] applied to the setting of the Ricci flow on closed 3-manifolds. Limiting arguments and the study of ancient solutions were developed by Hamilton with the aim of enabling surgery. A contradiction argument using limiting techniques was proposed to show that for suitable surgery parameters, the set of surgery times is discrete, and in particular, do not accumulate in finite time. Unfortunately, as was known to some mathematicians working in the field of Ricci flow and as pointed out in [298], there was an error in this part of Hamilton's argument. In the recent work of Perelman [297], [298], building on Hamilton's theory, Ricci flow behavior (especially singularity formation) on 3-manifolds is carefully examined and classified. The overall picture is subtle and technical, with some of the foundations being discussed in this volume. In the following two simplified examples (the first of which continues our discussion above), we try to convey some of its topological flavor, omitting most of the details. The formation of neckpinch singularities (as described above) in a certain sense reflects the topological connected-sum decomposition of the underlying manifold. Indeed, suppose a neckpinch with two ends occurs on a region identified with 8 2 x lR. Hamilton proposes a surgery process as follows [189].4 One does surgery near the large ends of the long, thin tubes in that part of the manifold identified with 8 2 x JR, capping these off with round 3-balls. Note that Hamilton's theory predicts the existence of such tubes where the curvature is very large at the center and slowly decreases as one moves away from the center along the relatively very long length of the tube. In fact his theory predicts that as one approaches the singularity time, the tube becomes arbitrarily close to an exact cylinder and its size slowly increases as one moves away from the center to an arbitrarily much larger but still very small size. 5 Note that Perelman's surgery process in [298] is a modification of the surgery process proposed earlier by Hamilton. 6 4More precisely, he considers the 4-dimensional version of this. 5More precisely, the tube is conformally close to a round product cylinder, where the conformal factor changes very slowly as one moves away from the center. 6Huisken and Sinestrari have considered an analogue of Hamilton's surgery theory for the mean curvature flow.
442
9. BASIC TOPOLOGY OF 3-MANIFOLDS
After the surgery, one continues the Ricci flow on the resulting (possibly disconnected) manifold, taking the glued (smoothed) metric as initial data. Heuristically, this surgery procedure corresponds to Kneser's sphere decomposition theorem. It is known that a 2-sphere removed from the neck may bound a 3-ba11. Thus for an arbitrary initial 3-manifold M, there is no guarantee that this surgery process will stop in finitely many steps. The finiteness theorem of Kneser is in some sense related to a finiteness conjecture of Hamilton for Ricci flow. The latter states that if one runs the unnormalized Ricci flow on a closed 3-manifold and performs geometrictopological surgeries whenever the Ricci flow develops a finite time singularity, then after finitely many such surgeries, the Ricci flow will have a nonsingular solution for all time. (Since one discards S3 and S2 x SI factors, this solution may well be empty.) Using Kneser's finiteness theorem, one sees that if Hamilton's conjecture does not hold, then all but finitely many geometric-topological surgeries at finite time singularities of the Ricci flow must split off 3-spheres. (In this regard, see the following recent papers: Perelman [299] and Colding and Minicozzi [116].) In this context, Hamilton's finiteness conjecture may be regarded as a geometric refinement of Kneser's theorem. Another type of Ricci flow behavior on 3-manifolds reflects the torus decomposition. This was first noticed in the work of Hamilton [190]. For simplicity, we recall Hamilton's formulation. He assumes that a solution to the normalized Ricci flow on a closed 3-manifold M exists for all time t E [0, 00) with uniformly bounded curvature. In this scenario, as time approaches infinity, it may happen that a manifold M can be decomposed into two parts M = Mthin U Mthick. In the components of Mthin, the metrics are collapsing with bounded curvature. (Recall that a manifold is said to be collapsible if it admits a sequence of Riemannian metrics of uniformly bounded curvature and volumes tending to zero.) In the components of Mthick, the metrics converge to complete hyperbolic metrics of finite volume. Furthermore, using minimal surface techniques, Hamilton proves that the fundamental group of Mthin n Mthick injects into the fundamental group of M. By the work of Cheeger and Gromov on collapsing manifolds, one concludes that Mthin is a Cheeger-Gromov graph manifold and hence a connected sum of Seifert spaces and sol-geometry manifolds. As a consequence, Hamilton was able to establish the geometrization conjecture under the (restrictive) hypotheses of long-time existence and uniformly bounded curvature. Perelman's recent work [298]' in conjunction with Shioya and Yamaguchi [332], claims to establish a similar picture without the assumption of bounded curvature. 4. Notes and commentary
We would like to thank Ian Agol for carefully reading this chapter and for pointing out mistakes in a draft version. We also note that Milnor's
4. NOTES AND COMMENTARY
443
[267] and Morgan's [272] are two excellent survey papers introducing Ricci flow on 3-manifolds and giving an overall picture of the current state of the research in the field.
APPENDIX A
Basic Ricci Flow Theory Heat, like gravity, penetrates every substance of the universe, its rays occupy all parts of space. The object of our work is to set forth the mathematical laws which this element obeys. The theory of heat will hereafter form one of the most important branches of general physics. - Joseph Fourier
In this appendix we recall some basic Ricci flow notation, formulas, and results, mostly from Volume One. Unless otherwise indicated, all page numbers, theorem references, chapter and section numbers, etc., refer to Volume One. Some of the results below are slight modifications of those stated therein. If an unnumbered formula appears on p. of Volume One, we refer to it as (Vl-p. v.); if the equation is numbered 0 .•, then we refer to it as (Vl-O .•). The reader who has read or is familiar with Volume One may essentially skip this chapter, referring to it only when necessary.
v.
1. Riemannian geometry 1.1. Notation. Let (M, g) be a Riemannian manifold. Throughout this appendix we shall often sum over repeated indices and not bother to raise (or lower) indices. For example, aijbij ~ gikgjiaijbki.
• If a is a l-form, then a~ denotes the dual vector field. Conversely, if X is a vector field, then X b denotes the dual l-form. • T M, T* M, A2T* M, and S2T* M denote the tangent, cotangent, 2-form, and symmetric (2, D)-tensor bundles, respectively. • r, V', and .6. denote the Christoffel symbols, covariant derivative, and Laplacian, respectively. • R, Rc, and Rm denote the scalar, Ricci, and Riemann curvature tensors, respectively. • r often denotes the average scalar curvature (assuming M is compact). • The upper index on the Riemann (3, l)-tensor is lowered into the 4-th position: ~jki = R7]k gmi . • ..\ ~ f.l ~ 1/ denote the eigenvalues of the Riemann curvature operator of a 3-manifold, in decreasing order. • d = dist, diam, and inj denote the Riemannian distance, diameter, and injectivity radius, respectively. • L, A = Area, and V denote length, area, and volume, respectively. 445
A. BASIC RICCI FLOW THEORY
446
• trg denotes the trace with respect to 9 (e.g., of a symmetric (2,0)-
tensor).
• sn usually denotes the unit n-sphere.
• For tensors A and B, A * B denotes a linear combination of contractions of the tensor product of A and B. • ~ denotes an equality which holds on gradient Ricci solitons. • ~ denotes an equality which holds on expanding gradient Ricci solitons.
1.2. Basic Riemannian geometry formulas in local coordinates. In Ricci flow, where the metric is time-dependent, it is convenient to compute in a local coordinate system. Let (Mn,g) be an n-dimensional Riemannian manifold. Almost everywhere we shall assume the metric 9 is complete. Let {xi} be a local coordinate system and let 8i ~ a~.. The components of the metric are gij ~ 9 (8i , 8 j ) . The Christoffel symbols for the Levi-Civita connection, defined by "ai8j ~ rfj8k, are k
r ij
(Vl-p. 24)
="21 g kl (8i g j l + 8 j g i l
where (gij) is the inverse matrix of curvature (3, I)-tensor, defined by
(gij) .
- 8lgij ) ,
The components of the Riemann
8 8) 8 . l 8 R ( 8x i '8x j 8x k =;= R ijk 8x l
(Vl-p. 286)
'
are (Vl-p. 68) The Ricci tensor is given by (Vl-p. 92)
D
. -
.L"'IJ -
P RPpij -_!l Up r ij
-
!l·rPpj
u~
+ r ijq r ppq -
q rP r pj iq'
The scalar curvature is R = gij ~j. If M is oriented and the local coordinates { xi} ~1 have positive orientation, then the volume form is (Vl-p. 70) The Bianchi identities. (1) First Bianchi identity: (Vl-3.l7)
0 = ~jkl + ~klj
+ ~ljk'
(2) Second Bianchi identity: (Vl-3.l8) where we take "Rm, cyclically permute the first three indices of "q~jkl (components), and sum to get zero.
447
1. RIEMANNIAN GEOMETRY
(3) Contracted second Bianchi identity: . 1 (VI-3.13) ,p~J = 2'ViR, which is obtained from (VI-3.18) by taking two traces (e.g., multiplying by gqlgjk and summing over q, e,j, k). 1.3. Cart an structure equations. For metrics with symmetry, such as rotationally symmetric metrics, it is convenient to calculate with respect to a local orthonormal frame, also called a moving frame. Let {ed ?=1 be a local orthonormal frame field in an open set U c Mn. Denote the dual orthonormal basis ofT* M by {wi} ~=1 so that 9 = E?=1 wi 0 wi. The connection I-forms w{ E 0 1 (U) are defined by n
'VXei ~ LwI (X) ej, j=1
(Vl-p. 106a)
for all i = 1, ... , n and X E Coo (T Mlu). They are antisymmetric: wI The first and second Cartan structure equations are
-w;.
(Vl-p. 106b)
dw i
(Vl-p. 106c)
Rm ~~
=
= wj !\ w;,
= O~ ~
j = dw1~ - w~~ !\ wk·
The following formula is useful for computing the connection I-forms:
(A.l)
k
wi (ej)
= dw i (ej, ek) + dwJ. (ei' ek) - dw k (ej, ei).
1.4. Curvature under conformal change of the metric. Let 9 9 be two Riemannian metrics on a manifold Mn conformally related by 9 = e2u g, where u : M - t R If {ed?=1 is an orthonormal frame field for g, then {ei}~1 , where ei = e-uei, is an orthonormal frame field for g. The Ricci tensors of 9 and 9 are related by and
(A) .2
if (- -)
c el, ei = e
-2u ( Rc (ee, ei) + (2 - n) 'Vel'VeiU - 8li flu ) + l'Vul 2 (2 _ n) 8il~ - (2 - n) eR.{u) edu) .
Tracing this, we see that the scalar curvatures of 9 and (A.3)
R = e- 2u (R - 2 (n -
9 are related by
1) flu - (n - 2) (n - 1) l'VuI 2 )
.
For derivations of the formulas above, which are standard, see subsection 7.2 of Chapter 1 in [111] for example. 1.5. Variations and evolution equations of geometric quantities. The Ricci flow is an evolution equation where the variation of the metric (i.e., time-derivative of the metric) is minus twice the Ricci tensor. More generally, we may consider arbitrary variations of the metric. Given a variation of the metric, we recall the corresponding variations of the Levi-Civita connection and curvatures. (In Volume One, see Section 1 of Chapter 3 for the derivations, or see Lemma 6.5 on p. 174 for a summary.)
A. BASIC RICCI FLOW THEORY
448
A.I (Metric variation formulas). Suppose that 9 (s) is a smooth I-parameter family of metrics on a manifold Mn such that :sg = v. (1) The Lem-Civita connection r of 9 evolves by LEMMA
(VI-3.3)
= ~ £p ( 'V'/ilkVjp + 'V'j'V'pVik - 'V'i'V'pVjk - 'V'j'V'kVip )
(VI-p. 69a)
2g
q -RqiJkVqp - R ijpVkq
.
(3) The Ricci tensor Rc of 9 evolves by
8 8s Rjk
(VI-3.5) (VI-p. 69b)
1 = '2 gpq ('V' q'V'jVkp + 'V' q'V' kVJp - 'V' q'V' pVjk - 'V' j 'V' kVqp) =
-~ [~LVjk + 'V'j'V'k (trgv) + 'V'J (8v)k + 'V'k (8v)j]
,
where ~L denotes the Lichnerowicz Laplacian of a (2, a)-tensor, which is defined by (~LV)jk ~ ~Vjk
(VI-3.6)
+ 2gqpR~jkVrp -
gqPRjpvqk - gqPRkpVjq.
(4) The scalar curvature R of 9 evolves by (VI-p. 69c) (VI-p. 69d)
where V ~ gi j vij is the trace of v. (5) The volume element df.L evolves by
8 V 8s df.L = '2 df.L.
(VI-p. 70)
(6) Let "Is be a smooth family of curves with fixed endpoints in M n and let Ls denote the length with respect to 9 (s). Then :s Ls ("(s)
(VI-3.8)
=
~
1
v (T, T) dO'
1.
-1
('V'TT, U) dO',
18
-'-- a1s -'-- as' a th T -;h were O' zs. arc leng, au' an d U -;1.6. Commuting covariant derivatives. In deriving how geometric quantities evolve when the metric evolves by Ricci flow, commutators of covariant derivatives often enter the calculations. (For the following, see p. 286 in Section 6 of Appendix A in Volume One.) If X is a vector field, then
(VI-p. 286a)
1. RIEMANNIAN GEOMETRY
449
If 0 is a I-form, then
(VI-p. 286b) More generally, if A is any (p, q)-tensor field, one has the commutator (VI-p. 286) [V',~, V'.J J
Al1···lq ....:... kl···kp --;-
V'.V' ·Al1···lq ~
kl···kp
J
V' ·V'·Al1···lq
-
~
J
kl···kp
q
=
p
~ R~r:
L..J
Al1· .. lr-l mlr+l· .. l q _ ~ R"!': Al1 .. ·lq ~Jm kl· .. k p L..J ~Jks kl···ks-l m ks+l' ·kp ·
r=1
8=1
1. 7. Lie derivative. Because of the diffeomorphism invariance of the Ricci flow, the effect of infinitesimal diffeomorphisms on tensors, e.g., the Lie derivative, enters the Ricci flow. (See p. 282 in Section 2 of Appendix A in Volume One.) The Lie derivative of the metric satisfies
(£Xg) (Y, Z) = 9 (V'yX, Z)
(VI-p. 282a)
+ 9 (Y, V' zX)
for all vector fields X, Y, Z. In local coordinates
(£Xg)ij = (£Xg)
(VI-p. 282b) In particular, if X
(a~i' a~j )
= V'iXj + V'jXi.
= V'I is a gradient vector field, then (£Vjg)ij = 2V'iV'jf.
1.8. Bochner formulas. (See p. 284 in Section 4 of Appendix A in Volume One.) The rough Laplacian denotes the operators ~:
where T%M ~ (VI-p. 284a)
Coo (T%Mn)
®P T* M 0 ®q T M,
~
Coo (T%Mn) ,
defined by
n
(~A)(Y1"'" Yp; 01, ... , Oq) =
L
(V'2 A) (ei' ei, Y1, ... , Yp; 01, ... , Oq)
i=1
for all (p, q)-tensors A, all vector fields Y1, ... , Yp, and all covector fields 01 , . .. , Oq, where {ed~=1 is a (local) orthonormal frame field. The Hodgede Rham Laplacian -~d : [!P (M) ~ [!P (M) is defined by -~d ~
(VI-p. 284b)
d8 + 8d.
In particular, if 0 is a I-form, then ~dO
(VI-p. 284c) For any function
1:M
=
Rc (0) .
~ ~
~ V' 1 =
(A.4)
~O -
V' ~I + Rc (V' f)
and (A.5)
~ 1\1112
= 21V'V' 112 + 2 Rc (V'I, V' f) + 2V'1 . V' (~f) ,
A. BASIC RICCI FLOW THEORY
450
where the dot denotes the metric inner product, i.e., X . Y gijXiyj.
=
If -9tgij (
= (X, Y) =
-2~j on M x (a,w) and f: M x (a,w) -+~, then
~-
:t) IV fl2
= 21VV fl2 + 2V f . V ( ( ~ -
!)
f) .
1.9. The cylinder-to-ball rule. The following is an obvious modification of Lemma 2.10 on p. 29 of Volume One.
°
LEMMA A.2 (Cylinder-to-ball rule). Let < L ::; 00 and let 9 be a warped-product metric on the topological cylinder (0, L) x sn of the form
9 = dr 2 + w (r)2 gcan, where w : (0, L) -+ ~+ and gcan is the canonical round metnc of radius 1 on sn. Then 9 extends to a smooth metric on B
(6, L)
(as r -+ 0+) if and
only if
(V1-2.16)
lim w(r)
r ......O+
= 0,
lim w' (r) = 1,
(V1-2.17)
r ...... O+
and
d2k w
(V1-2.18)
lim d 2k (r) = r ...... O+ r
°
for all kEN.
1.10. Volume comparison. We recall the Bishop-Gromov volume comparison (BGVC) theorem. THEOREM A.3 (Bishop-Gromov volume comparison). Let (Mn,g) be a complete Riemannian manifold with Rc ~ (n - 1) K, where K E ~. Then for any p E M, VoIB(p,r) VolKB(PK,r) zs a nomncreasing function of r, where PK is a point in the n-dimensional simply-connected space form of constant curvature K and YOlK denotes the volume in the space form. In particular
(A.6) for all r > 0. Given p and r is isometric to B(PK, r).
If Rc
~
> 0,
equality holds in (A.6) if and only if B(p, r)
0, we then have the following.
COROLLARY A.4 (BGVC for Rc ~ 0). If (Mn,g) is a complete Riemannian manifold with Rc ~ 0, then for any p E M, the volume ratio Vol~~p,r) is a nonincreasing function of r. We have Vol~~p,r) ::; Wn for all r > 0, where Wn is the volume of the Euclidean unit n-ball. Equality holds if and only if (Mn, g) is isometric to Euclidean space.
1.
RIEMANNIAN GEOMETRY
451
As a consequence, we have the following characterization of Euclidean space. COROLLARY A.5 (Volume characterization ofll~n). If (Mn, g) is a complete noncompact Riemannian manifold with Rc 2:: 0 and if for some p EM, lim VolB (p, r) = r-+oo
Wn ,
rn
then (M, g) is isometric to Euclidean space. The following result about the volume growth of complete manifolds with nonnegative Ricci curvature is due to Yau (compare with the proof of Theorem 2.92). COROLLARY A.6 (Rc 2:: 0 has at least linear volume growth). There exists a constant c (n) > 0 depending only on n such that zf (Mn, g) is a complete Riemannian manifold with nonnegative Ricci curvature and p E Mn, then Vol B (p, r) 2:: c (n) VolB (p, 1) . r for any r E [1,2 diam (M)).1 The asymptotic volume ratio of a complete Riemannian manifold (Mn,g) with Rc 2:: 0 is defined by
AVR(g)~ lim VoIB(p,r),
(A.7)
r-+oo
wnrn
where Wn is the volume of the unit ball in ]Rn. By the Bishop-Gromov volume comparison theorem, AVR (g) ~ 1. Again assuming Rc (g) 2:: 0, we have for 8 2:: r, (1) 8n - 1
A (8) ~ A (r) r n where A (8)
~
1'
Vol8B (p, 8),
(2) (A.8)
VoIB(p,r) > Wnrn
A(8) nwn 8 n -
> AVR(g). 1 -
We have the following relation between volume ratios and the injectivity radius in the presence of a curvature bound (see for example Theorem 5.42 of [111]). THEOREM A.7 (Cheeger, Gromov, and Taylor). Given c > 0, ro > 0, and n EN, there exists /'0 > 0 such that if (Mn , g) is a complete Riemannian manifold with Isect I ~ 1 and if p E M is such that VolB (p, ro) n
rO
2:: c,
lWe allow the noncompact case where diam (M) =
00.
A. BASIC RICCI FLOW THEORY
452
then inj (p) 2:
LO·
1.11. Laplacian and Hessian comparison theorems. Given K E lR and r > 0, let (n - 1) ..JK cot ( ..JKr )
HK (r)
~
{
(n _ 1)
°
~~th ( VJKlr)
if K
> 0,
if K
= 0,
if K < 0,
2:!K.
where if K > we assume r < The function HK (r) is equal to the mean curvature of the (n - 1)-sphere of radius r in the complete simplyconnected Riemannian manifold of constant sectional curvature K. THEOREM A.8 (Laplacian comparison). Let (Mn,g) be a complete Riemannian manifold with Rc 2: (n - 1) K, where K E lR. For any p E M n and x E Mn at which dp (x) is smooth, we have
Ildp (x) ::; HK (dp (x)) .
(A.9)
On the whole manifold, the Laplacian comparison theorem (A.9) holds in the sense of distributions. That is, for any nonnegative Coo function <.p on Mn with compact support, we have
r
JMn
dp (x) Il<.p (x) dJL(x) ::;
r
JMn
cK(dp(x))<.p(x)dJL(x).
The following is a special case of the Hessian comparison theorem. THEOREM A.9 (Hessian comparison theorem). If (Mn, g) is a complete Riemannian manifold with sect 2: K, then for any point p E M the distance function satisfies (A.lO)
at all points where dp is smooth (i.e., away from p and the cut locus). On all of M the above inequality holds in the sense of support functions. That is, for every point x E M and unit tangent vector V E TxM, there exists a C 2 function v : (-E, E) ~ lR with v (0) = dp (x) , dp (expx (tV)) ::; v (t), and 2
ddt 2 It=O v (t)::; n
for t E (-E, E),
~ IHK (d
p
(x)).
Note that n~l HK (r) is equal to the principal curvature of the totally umbillic (n - 1)-sphere of radius r in the complete simply-connected Riemannian manifold of constant sectional curvature K. In fact Ildp is the mean curvature of the distance sphere in M whereas VV dp is the second fundamental form of the distance sphere in M.
1.
RIEMANNIAN GEOMETRY
453
1.12. Li-Yau differential Harnack estimate. Let u : Mn x [0, 00) ---t lR be a positive solution to the heat equation ~~ = ~u on a complete Riemannian manifold (Mn, g) . Define I by
u ~ (47rt)-n/2 e- f ,
(A.11)
so that
Eft =
~I -1V'/12 -~.
THEOREM A.lO (Li-Yau differential Harnack estimate). II (M n , g) has nonnegative Ricci curvature, then ~I
(A.12)
n
81
2
- -2t = -8t + IV'/I -< O.
Integrating (A.12) yields the following sharp version of the classical Harnack estimate: (A.13) for all Xl, X2 E M and t2 > tl' If u = H is a fundamental solution centered at a point X E M, then taking tl ---t 0 implies the Cheeger-Yau estimate:
I (y, t) ~ d (~'ty)2
(A.14)
In terms of u, the positive solution to the heat equation, on a complete Riemannian manifolds with nonnegative Ricci curvature, we have
8
8t logu -1V'logul and
2
n
= ~logu ~ - 2t
{d
U (X2' t2) > (t2) ---:----'----'- -n/2 exp - (Xl, X2)2} . U(XI,tl) - tl 4(t2-tl) For a fundamental solution u H (y, t)
= H,
~ (41rt)-n/2 exp { _ d (~ty)2} .
1.13. Calabi's trick. In this subsection we give an example of Calabi's trick which is useful in the study of heat-type equations and analytic aspects of the Ricci flow. In particular, a slight modification of the discussion below applies to the proof of the local first derivative of curvature estimate for the Ricci flow (see Theorem A.30). First, let us recall some facts about the distance function. Let (Mn, g) be a Riemannian manifold. Given p EM, the distance function r (x) ~ d (x, p) is Lipschitz on M with Lipschitz constant 1. Let Cut (p) denote the cut locus of p and let Cp ~ {V E TpM : d (p,expp (V)) =
IVI},
A. BASIC RICCI FLOW THEORY
454
so that Cut (p) zero. We have
= expp (aCp ). The cut locus is a closed set with measure expplintCp : Cp \aCp
is a diffeomorphism. Let field on TpM -
aIar =
--t
M\ Cut (p)
I!I L:~1 xi a~i denote the unit radial vector
{5} . If x rt. Cut (p) U {p}, then r is smooth at x, 'V'r (x) =
(exppL alar, and l'V'r (x)1 = 1. Suppose that we have a function F : M x [0, T) differential inequality
(~ at -
(A.15)
--t
lR which satisfies the
2 < - C - F2
6.) F
for some constant C. For an example of such a function, see the proof of the local first derivative of curvature estimate in Part II of this volume. If M is a closed manifold, then we can apply the maximum principle to F to obtain the estimate
F (x, t) :::; C coth (Ct) , where the RHS is the solution to the ODE 1ft = C2 - P with limt'\.o f (t) = +00. On the other hand, if M is noncompact, then one way of obtaining an estimate for F is to localize the equation by introducing a cut-off function. In particular, suppose (M n , g) satisfies Rc ~ - (n - 1) K for some K > O. Let 'f/ : [0, 00) --t lR be a smooth nonincreasing function satisfying 'f/( s) = 1 for 0 :::; s :::; ~ and 'f/(s) = 0 for s ~ 1. We may assume 0 ~ 'f/' ~ -6Jri
(A.16)
- CoJri :::; 'f/" :::; Co,
and
where Co is a universal constant. 2 Given p E M and A 'f/ ( d(~t»)· Recall that in M\ ({p} U Cut (p))
'V'd (', p) 1 = 1,
1
> 0, define ¢(x)
6.d (" p) :::; (n - 1) v'K coth ( v'Kd ( " p)) ,
where the Laplacian estimate follows from (A. g) and the assumption Rc - (n - 1) K. Hence at points x rt. Cut (p) , ¢ is smooth and (A.17)
~
1'V'¢1 2 :::;C¢,
6.¢=
~
~'f/,.6.d+ 12'f/"I'V'dI2~-CV¢,
where C depends on A (and where we used (A.16), 'f/' :::; 0, and the fact that the support of 'f/' is contained in B (p, A) \B (p,
(d(:»)
4) ).
We calculate that for x ¢
(:t -
6.) (¢F) = ¢2
rt. Cut (p) ,
(:t -
6.) F - ¢ (6.¢) F - 2¢'V'¢· 'V' F
:::; -2'V'¢. 'V' (¢F)
+ ¢2C2 -
2Let ( be a cut-off function with 0 ~ (' ~ -3 and constant, and define 'f/ = (2.
¢2 F2 - ¢F6.¢ + 2FI'V'¢12
ICI
S; C, where C is a universal
455
1. RIEMANNIAN GEOMETRY
(the above calculation holds for any C 2 function
[
~) + 2'V
(
2
1 ( C1 - (
for some constant C 1 <
[
00.
+ C
2)
Hence
~) + 2'V
°
;t (Cft2 - (t
Since limt-+o (t
Cft 2 - (t
+ 2t
°
at a maximum point of t
t
~
C2
on M x [0, T), where C2 depends only on C in (A.1S) and A. In particular,
F(x,t) ~
C2
T
on B (p, A/2) x [0, T). EXERCISE
A.11. What happens to the constant C2 as A
--t
oo?
In the above we have assumed that at the choice (xo, to) of maximum point of t
[0, d(xo,p)] --t M be a unit speed minimal geodesic joining Xo to p. Consider q ~ 'Y (d(xo,p) - (5) for any small <5 = d(q,p) > 0. We have d(xo, q) + <5 = d(xo,p) and also q is not in the cut locus of Xo since <5 > 0. Consider the function 'Y:
q) G (x, t) =;=. t'f'/ (d(X, A Since d(x, q)
+ <52: d (x,p)
+ (5) F (x, t) .
and 'f'/ is nonincreasing, we have G (x, t) ~ (t
at any point (x, t) where F (x, t) 2: 0. Since G (xo, to) = to
A. BASIC RICCI FLOW THEORY
456
(xo, to). However, since de, q) is C2 at Xo, we may apply the maximum principle to the equation for G. One checks (Exercise: Prove this) that G (x, t) ~ C~
on M x [0, T), where C~
--t
C 2 as 8 --t O. Hence
max_ (t¢F)
Mx[O,t)
~
C2 •
2. Basic Ricci flow The (unnormalized) Ricci flow equation on a manifold Mn is (A.18) whereas its cousin, the (volume-preserving) normalized Ricci flow equation on a closed manifold, is (A.19) where r (t) ~
(fM RdJ-l/ fM dJ-l) (t) is the average scalar curvature.
REMARK A.12. The variation of a Riemannian metric in the direction of the Ricci curvature was considered by Bourguignon (see Proposition VIII.4 of [32]). A fundamental work using a nonlinear heat-type equation (the harmonic map heat flow) is by Eells and Sampson [135]. Substituting h = -2 Rc into Lemma A.l yields the following result, which gives the evolution equations for the Levi-Civita connection and curvatures under the Ricci flow (A.18). (See Corollary 6.6 (1) on p. 175, Lemma 6.15 on p. 179, and Lemmas 6.9 and 6.7 on p. 176 of Volume One.) COROLLARY A.13 (Ricci flow evolutions). Suppose 9 (t) is a solution of the Ricci flow: g = -2 Rc.
tt
(1) The Levi-Civita connection
ata r ijk =
(VI-6.1)
-g
kl
r
C<;;JiRjl
of 9 evolves by
+ \1jRi l
- \1lRij).
(2) Under the Ricci flow, the (4,0)-Riemann curvature tensor evolves by (VI-6.17)
ata Rijkl = f::.Rijkl + 2 (Bijkl -
where (VI-6.16)
(Rf Rpjkl
Bijlk
+ Bikjl -
Bifjk)
+ R~RiPkl + R~Rijpf. + R~RiJkP) ,
2 BASIC RICCI FLOW
457
(3) The Ricci tensor Rc of 9 evolves by (VI-6.7) In dimension 3, this equation becomes (VI-6.1O)
:t Rjk
=
f:l.Rjk
+ 3RRjk -
6g pq RjpRqk
+ (21Rc12 -
R2) gjk·
(4) The scalar curvature R of 9 evolves by oR at = f:l.R + 21Rcl 2 .
(VI-6.6)
(5) The volume form dj.l evolves by
o
(VI-6.5)
otdj.l
=
-Rdj.l.
(6) If 'Yt is a smooth I-parameter family of geodesic loops, then
()I
d dt L t 'Yt t=T
=-
1 ''(T
Rc
(o'Yos'T O'YT) os
ds,
where s is the arc length parameter. Part (6) is similar to Lemma 5.71 on p. 152 of Volume One. EXERCISE A.I4. Derive the corresponding formulas for the normalized Ricci flow. For example, under (A.I9) we have oR 2 at = f:l.R + 21Rcl -
2 -:;;,rR.
2.1. Short- and long-time existence. Any smooth metric on a closed manifold will flow uniquely, at least for a little while (Theorem 3.13 on p. 78 of Volume One). THEOREM A.15 (Short-time existence for M closed). If (M n , go) i8 a closed Riemannian manifold, then there exi8t8 a unique 8olution 9 (t) to the Ricci flow defined on 80me p08ztive tzme interval [0, E) 8uch that 9 (0) = go. As long as the curvature stays bounded, the solution exists (Corollary 7.2 on p. 224 of Volume One). THEOREM A.I6 (Long-time existence for M closed). If(Mn,g(t)), t E (0, T), where T < 00, i8 a 80lutwn to the Ricci flow on a clo8ed manifold with sUPMx(O,T) IRml < 00, then the 8olution 9 (t) can be uniquely extended past time T. In the theorem above the condition sUPMx(O,T) IRml < 00 may be replaced by sUPMx(O,T) IRcl < 00; this was proved by Sesum [321]. W.-X. Shi generalized Hamilton's short-time existence theorem to complete solutions with bounded curvature on noncompact manifolds.
A. BASIC RICCI FLOW THEORY
458
THEOREM A.17 (Short-time existence on noncompact manifolds). Let
M be a noncompact manifold and let 90 be a complete metnc with bounded sectwnal curvature. There exists a complete solution 9 (t) , t E [0, T), of the Ricci flow wzth 9 (0) = 90 and curvature bounded on compact time intervals. This solutwn zs unzque in the class of complete solutions with curvature bounded on compact time intervals.
2.2. Maximum principles for scalars, tensors and systems. A form of the scalar maximum principle useful for the Ricci flow is the following (Theorem 4.4 on p. 96 of Volume One). THEOREM A.lS (Scalar maximum principle: ODE to PDE). Let u : M n x [0, T) ---+ lR. be a C 2 function on a closed manifold satisfying
au at ~ ~g(t)U + (X, V'u) + F (u) and u (x, 0) ~ C for all x E M, where 9 (t) is a I-parameter family of metncs and F is locally Lipschitz. Let 'P (t) be the solution to the initialvalue problem d'P dj=F('P), 'P(O) =
c.
Then u (x, t) ~ 'P(t) for all x E M and t E [0, T) such that 'P (t) exists. Since under the Ricci flow, by (VI-6.6) ~~ 2': ~R + ~R2, we have the following (see the proof of Lemma 6.53 on pp. 209-210 of Volume One). COROLLARY A.19 (Scalar curvature lower bound). Let (Mn, 9 (t)), where be a solution of the Ricci flow for which the maximum principle holds. If infxEMn R (x, to) ~ P > 0 for some to E [0, T), then
o ~ t < T,
Rinf
(t) ~ inf R (x, t) 2': xEM
1 1 - -
p
2 ( -n t
- to
).
In particular, g (t) becomes singular m finite time. Moreover, we have the following. LEMMA A.20 (Minimum scalar curvature monotonicity).
(1) Under the unnormalized Ricci flow the minimum scalar curvature is a nondecreasing function of time. (2) Under the normalized Ricci flow the minzmum scalar curvature is nondecreasmg as long as it is nonpositive.
2.
PROOF. Let p (t)
~
BASIC RICCI FLOW
459
R min (t). Under the unnormalized Ricci flow,
> ~p2 > O. dt - n Under the normalized Ricci flow, dp 2 dt ~ ;;,P (p - r) ~ 0 dp
o
as long as p S 0 (note that p - r S 0 always).
The weak maximum principle as applied to symmetric 2-tensors says the following (see Theorem 4.6 on p. 97 of Volume One).3 THEOREM A.21 (Maximum principle for 2-tensors). Let g (t) be a smooth I-parameter family of Riemannian metrics on a closed manifold Mn. Let a(t) E Coo (T* M ®s T* M) be a symmetric (2, O)-tensor satzsfying the semilinear heat equation
a
at a
~ ~g(t)a + /3,
where /3 (a, g, t) is a symmetric (2, O)-tensor which is locally Lipschitz in all its arguments and satisfies the null eigenvector assumption that /3 (V, V) (x, t)
=
(/3ij vivj) (x, t) ~ 0
at any point and time (x, t) where (aijWiW j ) (x, t) ~ 0 for all Wand (aijVj) (x, t)
= O.
If a (0) ~ 0 (that is, if a (0) is positive semidefinite), then a (t) t ~ 0 such that the solution exists.
~
0 for all
Applying this to the evolution equation (Vl-6.10) for the Ricci tensor in dimension 3 yields the following (Corollary 6.11 on p. 177 of Volume One). COROLLARY A.22 (3d positive Ricci curvature persists). Let g (t) be a solution of the Ricci flow on a closed 3-manifold with g (0) = go· If go has positive (nonnegative) Ricci curvature, then g (t) has positive (nonnegative) Ricci curvature for as long as the solution exists.
2.3. Uhlenbeck's trick. Uhlenbeck's trick allows us to put the evolution equation (Vl-6.17) satisfied by Rm into a particularly nice form. (See pp. 180-183 in Section 2 of Chapter 6 in Volume One.) Let (Mn,g(t)) , t E [0, T), be a solution to the Ricci flow with g (0) = go. Let V be a vector bundle over M isomorphic to T M, and let "0 : V ---+ T M be a bundle isomorphism. Then if we define a metric h on V by (V1-p. 181a)
h~
"0 (go) ,
we automatically obtain a bundle isometry (V1-p. 181b)
"0 :
(V, h)
---+
(T M, go) .
3The statement we give here is slightly stronger, since in the null eigenvector assumption we also assume that (Qi] Viv j ) (x, t) ~ 0 at (x, t) .
A. BASIC RICCI FLOW THEORY
460
LEMMA A.23. If we evolve the zsometry L (t) by
a
at L = RcoL, L (0) = LO ,
(VI-6.19a) (VI-6.19b) then the bundle maps
L(t) : (V, h)
~
(T M, 9 (t))
remain isometries. 4
We define the Laplacian acting on tensor bundles of T M and V by D..D ~
trg (V'D
0
V'D),
where (V'Dh (~) = L- 1 (V'x (L(O)) acting on sections of V and V'D is naturally extended to act on tensor bundles. For x E M and X, Y, Z, WE Vx , the tensor (Vl-p. 182) is defined by (VI-6.20)
(L* Rm) (X, Y, Z, W)
~
Rm (L (X) , L(Y), L(Z), L(W)).
Let Rabed denote the components of L* Rm with respect to an orthonormal basis of sections of V. We have (VI-6.21)
ata Rabcd = D..DRabed + 2 (Babcd -
Babde
+ Baebd -
Badbe) ,
where B abed
...!...
-;- -
heghfhRaebf R egdh·
We may rewrite the above equation in a more elegant way. (See pp. 183-187 in Section 3 of Chapter 6 in Volume One.) Let 9 be a Lie algebra endowed with an inner product (-, .). Choose a basis {cpO:} of 9 and let C~/3 denote the structure constants defined by [cpO:, cp/3] ~ L.'Y C~/3 cp'Y. We define the Lie algebra square L# E 9 08 9 of L by (VI-6.24)
(L#)O:/3 ~ CJ8C~( L'YcL8(.
For each x E Mn, the vector space A2T; M can be given the structure of a Lie algebra 9 isomorphic to so (n). Given U, V E A 2 T;M, we define their Lie bracket by (VI-6.25) THEOREM A.24 (Rm evolution after Uhlenbeck's trick). If 9 (t) is a solution of the Riccz flow, then the curvature L* Rm defined in (VI-6.20) evolves by (VI-6.27)
ata (L* Rm) =
D..D
(L* Rm)
+ (L* Rm) 2 + (L* Rm) #.
4The statement here is the one we intended in Claim 6.21 of Volume One.
2 BASIC RICCI FLOW
461
(See pp. 187-189 in Section 4 of Chapter 6 in Volume One.) The PDE (VI-6.27) governing the behavior of Rm corresponds to the ODE (VI-6.28) In dimension 3, if Mo is diagonal, then M (t) remains diagonal, and its eigenvalues satisfy
dA
2
dt = A
(VI-6.32)
dJ.l dt
+ I.LV,
2
= J.l + AV,
dv 2 dt = v
+ AJ.l.
From now on we shall assume A 2:: J.l 2:: v, a condition which is preserved under the ODE. Theorem 4.8 on p. 101 of Volume One applied to the Riemann curvature operator Rm yields the following. THEOREM A.25 (Maximum principle for Rm: ODE to PDE). Let 9 (t) be a solution to the Ricci flow on a closed manifold Mn and let K (t) be a closed subset of E ~ A2V 08 A 2V for all t E [0, T) satisfying the following properties: (1) the space-time track UtE[O,T) (K (t) x {t}) zs a closed subset of E x
[0, T); (2) K (t) is invariant under parallel translation by '\7 (t) for all t E [0, T); (3) Kx (t) ~ K (t)n7r- 1 (x) is a closed convex subset of Ex for all x E M and t E [0, T); and (4) every solution M of the ODE (VI-6.28) with Rm (to) E Kx (to) defined in each fiber Ex remains in Kx (t) for all t 2:: to and to E [0, T). If (L* Rm) (0) E K (0), then (L* Rm) (t) E K (t) for all t E [0, T).
2.4. 3-manifolds with positive Ricci curvature. The following famous theorem of Hamilton started the Ricci flow (RF). (See Theorem 6.3 on p. 173 of Volume One.) THEOREM A.26 (RF on closed 3-manifolds with Rc > 0). Let (M 3 ,go) be a closed Riemannian 3-manifold of positive Ricci curvature. Then a unique solution 9 (t) of the normalized Ricci flow wzth g (0) = go exists for all positive tzme; and as t ---* 00, the metrics g(t) converge exponentially fast in every Ck-norm. kEN, to a metric goo of constant positive sectional curvature. The above result generalizes to orbifolds (see [191]).
2 BASIC RICCI FLOW
461
(See pp. 187-189 in Section 4 of Chapter 6 in Volume One.) The PDE (Vl-6.27) governing the behavior of Rm corresponds to the ODE (Vl-6.28) In dimension 3, if Mo is diagonal, then M (t) remains diagonal, and its eigenvalues satisfy d)"
(Vl-6.32)
2
+ J-L//,
dt
=)..
dJ-L dt
= J-L + )..//,
d// dt
= // + )"J-L.
2
2
From now on we shall assume ).. ~ J-L ~ //, a condition which is preserved under the ODE. Theorem 4.8 on p. 101 of Volume One applied to the Riemann curvature operator Rm yields the following. THEOREM A.25 (Maximum principle for Rm: ODE to PDE). Let g (t) be a solution to the Ricci flow on a closed manifold M n and let K (t) be a closed subset of E ~ A2 V 08 A 2 V for all t E [0, T) satisfying the following properties:
(1) the space-time track UtE[O,T) (K (t) x {t}) is a closed subset of Ex [0, T); (2) K (t) is invarwnt under parallel translation by ~ (t) for all t E [0, T); (3) Kx (t) ~ K (t)n1T- 1 (x) is a closed convex subset of Ex for all x E M and t E [0, T); and (4) every solution M of the ODE (Vl-6.28) with Rm (to) E Kx (to) defined in each fiber Ex remains m Kx (t) for all t ~ to and to E [0, T).
If (t* Rm) (0) E K (0), then (t* Rm) (t) E K (t) for all t E [0, T). 2.4. 3-manifolds with positive Ricci curvature. The following famous theorem of Hamilton started the Ricci flow (RF). (See Theorem 6.3 on p. 173 of Volume One.) THEOREM A.26 (RF on closed 3-manifolds with Rc > 0). Let (M 3 , go) be a closed Riemannian 3-manifold of positive Ricci curvature. Then a unique solution g (t) of the normalized Ricci flow with g (0) = go exists for all positive time; and as t ---? 00, the metrics g(t) converge exponentwlly fast in every k -norm, kEN, to a metr'tc goo of constant positive sectional curvature.
c
The above result generalizes to orbifolds (see [191]).
A BASIC RICCI
462
FLOW THEORY
THEOREM A.27 (RF on closed 3-orbifolds with Rc > 0). If (V 3 ,go) zs a closed Riemannian 3-orbifold of positive Ricci curvature, then a unique solution g (t) of the normalized Ricci flow wzth g (0) = go exists for all t > 0, and as t ---+ 00, the g(t) converge to a metric goo of constant positive sectional curvature. In particular, V 3 is dzJJeomorphic to the quotient of 53 by a finite group of isometries. One of the main ideas in the proof of Theorem A.26 is to apply Theorem A.25 to obtain pointwise curvature estimates which lead to the curvature tending to constant as the solution evolves. From now on we shall assume that A (t) 2:: J.L (t) 2:: v (t) are solutions of the ODE system (VI-6.32). The evolutions of various quantities and their applications to the Ricci flow on closed 3-manifolds via the maximum principle for systems are given as follows.
(1) d dt (v
(A.20)
+ J.L) = v 2 + J.L2 + (v + J.L) A 2:: 0,
with the inequality holding whenever J.L + v 2:: preserved in dimension 3 under the Ricci flow.
o.
So Rc > 0 is
(2) (A.21) If go has positive Ricci curvature, then so does g (t) and there exists a constant C 1 < 00 such that (A.22)
A (Rm)
~
C 1 [v (Rm)
+ J.L (Rm)].
(3) If v + J.L > 0, then
! Cv +~ ~:)l-' ) log
= 8 (v + A _ J.L) _ (1 _ 8) (v + J.L) J.L + (J.L - v) A + J.L2 V+J.L+A J.L2 ~ 8 (v + A - J.L) - (1 - 8) -v---'+-J.L-+-A
+ A-
J.L ~ A ~ 2C1 J.L and ~ 2:: vt;..f1: 2:: enough so that 1~6 ~ 12b2, we have Since v
6b1
1
d ( -log dt (v
A-V
+ J.L + A)1-6
)
~
o.
,
choosing 8
> 0 small
463
2 BASIC RICCI FLOW
So if go has positive Ricci curvature, then there exist constants C fJ > 0 such that A (Rm) - 1/ (Rm) C (A.23) Rl-li ::;.
< 00 and
We shall call (A.23) the 'pinching improves' estimate. Next we consider estimates for the derivatives of Rm.
2.5. Global derivative estimates. For solutions to the Ricci flow on a closed 3-manifold with positive Ricci curvature, we have the following estimate for the gradient of the scalar curvature. (See Theorem 6.35 on p. 194 of Volume One.) THEOREM A.28 (3-dimensional gradient of scalar curvature estimate). Let (M 3 ,g (t)) be a solution of the Ricci flow on a closed 3-manifold with g (0) = go. If Rc (go) > 0, then there exist {J,"8 > 0 depending only on go such that for any {3 E [0, {J], there exists C depending only on {3 and go such that
IV'RI2
R3
< {3R-8/2 + C R- 3 . -
After a short time, the higher derivatives of the curvature are bounded in terms of the space-time bound for the curvature. (See Theorem 7.1 on pp. 223-224 of Volume One.) THEOREM A.29 (Bernstein-Bando-Shi estimate). Let (Mn, g (t)) be a solution of the Ricci flow for whzch the maximum principle applies to all the quantztzes that we consider. (This is true in particular if M is compact.) Then for each 0: > 0 and every mEN , there exists a constant C (m, n, 0:) dependzng only on m, and n, and max {o:, I} such that if IRm (x, t)lg(t) ::; K
0:
for all x E M and t E [0, K],
then for all x E M and t E (0, ~],
(Vl-p.
224 )
()I l"mR v m x, t g(t)::;
C(m,n,o:)K tm / 2
With all of the above estimates and some more work, one obtains Theorem A.26. Finally we mention that an important local version of Theorem A.29 is the following. THEOREM A.30 (Shi-Iocal first derivative estimate ). For any 0: > 0 there exists a constant C (n, K, r, 0:) depending only on K, r, 0: and n such that if Mn zs a manifold, p E M, and g (t), t E [0,7], 0 < 7 ::; 0:/ K, zs a solutwn to the Ricci flow on an open neighborhood U of p containing Bg(o) (p, r) as a compact subset and if IRm (x, t)1 ::; K for all x E U and t E [0,7],
A BASIC RICCI FLOW THEORY
W-I
then
Inv
(A.24)
R
( )1 C (n, K, r, a) = C(n, vXr, a)K m y, t:s; jt jt
for all (y,t) E Bg(D) (p,r/2) x (0,7]. Gwen m addition {3 > 0 and"( > 0, if also "(/vX:S; r:S; {3/vX, then there exists C(n,a,{3,"() such that under the above assumptwns K IV'Rml < C (n, a, {3, "() ;;
-
m
Bg(D)
vt
(p,r/2) x (0,7].
For a proof and applications, see W.-X. Shi [329], [330], Hamilton [186], [111], or Part II of this volume.
2.6. The Hamilton-Ivey estimate. The following result reveals the precise sense in which all sectional curvatures of a complete 3-manifold evolving by the Ricci flow are dominated by the positive sectional curvatures. (See [186] or Theorem 9.4 on p. 258 of Volume One.) THEOREM A.31 (3d Hamilton-Ivey curvature estimate). Let (M3,g(t)) be any solution of the Rzccz flow on a closed 3-mamfold for 0 :s; t < T. Let v (x, t) denote the smallest eigenvalue of the curvature operator. If infxEM v (x, 0) ~ -1. then at any pomt (x, t) EM x [0, T) where v (x, t) < 0, the scalar curvature is estimated by (A.25)
R
~ Ivl (log Ivl
+ log (1 + t) -
3) .
2.7. Ricci solitons. If 9 is a Ricci soliton, then
for some p E (VI-5.16)
Rc-!!"'g = LXqg n lR and I-form X. Under this equation we have
~ (R -
p)
+ (V' (R -
p), X)
+ 21Rc _~gI2 + ~ (R -
p)
= O.
Using this formula, in Proposition 5.20 on p. 117 of Volume One, the following classification result for Ricci solitons was proved. (See Chapter 1 of this volume for the relevant definitions.) PROPOSITION A.32 (Expanders or steadies on closed manifolds are Einstein). Any expanding or steady Rzcci soliton on a closed n-dzmenswnal manifold is Emstem. A shrinking Riccz solzton on a closed n-dzmenswnal manifold has positive scalar curvature. In dimension 2, all solitons have constant curvature. tion 5.21 on p. 118 of Volume One.)
(See Proposi-
PROPOSITION A.33 (Ricci solitons on closed surfaces are trivial). If (M2, 9 (t)) is a self-similar solution of the normalized Rzcci flow on a Riemanman surface, then g(t) == g(O) is a metric of constant curvature.
3. BASIC SINGULARITY THEORY FOR RICCI FLOW
465
3. Basic singularity theory for Ricci flow The knowledge of which geometry aims is the knowledge of the eternal - Plato Geometry is knowledge of the eternally existent. - Pythagoras And perhaps, posterity will thank me for having shown it that the ancients did not know everything - Pierre Fermat
In this section we review some basic singularity theory as developed by Hamilton and discussed in Volume One. 3.1. Long-existing solutions and singularity types. For the following, see pp. 234-236 in Section 1 of Chapter 8 in Volume One. DEFINITION A.34. • An ancient solution is a solution that exists on a past time interval (-00, w). • An immortal solution is a solution that exists on a future time interval (a, 00 ) . • An eternal solution is a solution that exists for all time (-00, (0). DEFINITION A.35 (Singularity types). Let (Mn, 9 (t)) be a solution of the Ricci flow that exists up to a maximal time T ~ 00.
• One says (M, 9 (t)) forms a Type I singularity if T < sup
(T - t) IRm (', t)1 <
00
and
00.
Mx[O,T)
• One says (M, 9 (t)) forms a Type IIa singularity if T < sup
(T - t) IRm (', t)1
00
and
00
and
= 00.
Mx[O,T)
• One says (M, 9 (t)) forms a Type lIb singularity if T = sup
t IRm (', t)1 =
00.
Mx[O,oo)
• One says (M, 9 (t)) forms a Type III singularity if T sup
t IRm (" t)1 <
= 00 and
00.
Mx[O,oo)
To this we add the following. A.36 (More singularity types). If 9 (t) is defined on (0, TJ, then
DEFINITION
where T <
00,
• one says (M,g(t)) forms a Type lIe singularity as t sup Mx(O,T]
t IRm (-, t)1 =
00;
-t
0 if
466
A. BASIC RICCI FLOW THEORY • one says (M, 9 (t)) forms a Type IV singularity as t sup
---t
0 if
t IRm (', t)1 < 00.
Mx(O,T]
We have the following examples of singularities. A neckpinch forms a Type I singularity (Section 5 in Chapter 2 of Volume One). A degenerate neckpinch, if it exists, forms a Type IIa singularity (Section 6 in Chapter 2 of Volume One). Many homogeneous solutions form Type III singularities (Chapter 1 of Volume One). CONJECTURE A.37 (Degenerate neckpinch existence). There exzst solutions to the Ricci flow on closed manifolds which form degenerate neckpinches. The analogue of the above conjecture has been proved for the mean curvature flow [9]. CONJECTURE A.38 (Nonexistence of Type lIb on closed 3-manifolds). If (M3, 9 (t)) , t E [0, 00), is a solution to the Rzcci flow on a closed 3-manifold, then 9 (t) forms a Type III singularity. Similar to the division of types for finite time singular solutions, we may divide ancient solutions into types. DEFINITION A.39 (Ancient solution types). Let (M n , 9 (t)) be a solution of the Ricci flow defined on (-00,0). • We say (M, 9 (t)) is a Type I ancient solution if sup
ItliRm (" t)1 < 00.
Mx(-oo,-l]
• We say (M, 9 (t)) is a Type II ancient solution if sup
ItliRm (., t)1
= 00.
Mx(-oo,-l]
3.2. Ancient solutions have nonnegative curvature. Every ancient solution (of any dimension) has nonnegative scalar curvature. (See Lemma 9.15 on p. 271 of Volume One.) LEMMA A.40 (Ancient solutions have R ~ 0). Let (Mn,g (t)) be a complete anczent solution of the Ricci flow. Assume that the functwn Rmin (t) ~ infxEMn R (x, t) is finite for all t ~ 0 and that there is a continuous function K (t) such that Isect [g (t)]1 ~ K (t). Then 9 (t) has nonnegative scalar curvature for as long as it exists. A particular consequence of the Hamilton-Ivey estimate is that ancient 3-dimensional solutions of the Ricci flow have nonnegative sectional curvature. (See Corollary 9.8 on p. 261 of Volume One.) COROLLARY A.41 (Ancient 3-dimensional solutions have Rm ~ 0). Let (M 3,g (t)) be a complete ancient solution of the Ricci flow. Assume that there exists a continuous functwn K (t) such that Isect [g (t)]1 ~ K (t). Then 9 (t) has nonnegative sectwnal curvature for as long as it exists.
3
BASIC SINGULARITY THEORY FOR RICCI FLOW
467
3.3. Trace Harnack inequality. Given a surface (M2,g) with positive curvature, the trace Harnack quantity is defined by (VI-5.35)
Q=
~logR
+R
- r
a
2
= at logR -IVlogRI .
(Also see Lemma 5.35 on p. 144 of Volume One.) We have the following differential Harnack estimate of Li-Yau-Hamiltontype. (See Corollary 5.56 on p. 145 of Volume One.) COROLLARY A.42 (2d trace Harnack evolution and estimate). On any solution of the normalized Ricci flow (A.19) on a complete surface with bounded positive scalar curvature, Q satzsfies the evolutionary inequality (VI-5.38)
For the unnormalized flow (A.18), the analogous quantity (Vl-p. 169a)
-
Q=
~ log R
a log R + R = at
IV log RI
2
satisfies (VI-5.57)
By the maximum principle, -
Q (x, t)
(Vl-p.169b)
1
+t
~ 0
for all x E M and t > O. In all dimensions, we have the following. (See Proposition 9.20 on p. 274 of Volume One.) PROPOSITION A.43 (Trace Harnack estimate). If (Mn, 9 (t)) is a solution of the Ricci flow on a complete manifold with bounded positive curvature operator, then for any vector field X on M and all times t > 0 such that the solution exists, one has (Vl-p.274)
aR
R
at + t + 2 (VR,X) + 2Rc(X,X)
;::: O.
The proof of Proposition A.43 will be given in Part II. When n = 2, by choosing the minimizing vector field X = - R- 1 V R, it can be seen that (Vl-p.274) is equivalent to (Vl-p.169b). One also has Corollary 9.21 on p. 274 of Volume One, namely COROLLARY A.44 (Trace Harnack consequence, tR monotonicity). If (Mn,g(t)) is a solution of the Ricci flow on a complete manifold with bounded curvature operator, then the function tR is pointwzse nondecreasmg for all t ;::: 0 for which the solution exists. If (M,g (t)) zs also ancient, then R ztself is pointwise nondecreasing.
A BASIC RICCI FLOW THEORY
468
3.4. Surface entropy formulas. The surface entropy N is defined for a metric of strictly positive curvature on a closed surface M2 by (Vl-p. 133) Let
f
N(g)
~
r
t:.f
=R
Rlog Rdp,. 1M2 be the potential function, defined up to an additive constant by
(VI-5.8)
- r.
(See Lemma 5.38 on p. 133 and Proposition 5.39 on p. 134 of Volume One.) PROPOSITION A.45 (Surface entropy formula). If (M2, 9 (t)) 2S a solutwn of the normalized Ricci flow on a compact surface wzth R (., 0) > 0, then (VI-5.25) (Vl-p. 134)
dN dt
=_
r
1M2
=_
r
r
IV'RI2 dA+ (R-r)2dA R 1M2
V'R + R V' fl2 dA 1M2 R
- 2
1
fM21V'V' f - ~t:.f . gl2 dA
::; o. 3.5. Ancient 2-dimensional solutions. 3.5.1. Examples. (See pp. 24-28 in Section 2 of Chapter 2 in Volume One.) Hamilton's cigar soliton is the complete Riemannian surface (lR?, 9E), where . dx @ dx + dy @ dy (VI-2.4) gE =;= 1 + x2 + y2 This manifold is also known in the physics literature as Witten's black hole. In polar coordinates (VI-2.5)
1 + r2
If we define
(Vl-p. 25a)
s
~ arcsinh r = log (r + J 1 + r2)
,
then we may rewrite gE as (VI-2.7)
gE
= ds 2 + tanh2 sd(P.
The scalar curvature of gE is
4
(Vl-p. 25b)
RE
= 1 + r2 =
4 cosh2 s
16
(e s
+ e- s )2·
(See pp. 31-34 in subsection 3.3 of Chapter 2 in Volume One.) Let h be the flat metric on the manifold M2 = lRxst, where st is the circle of radius 1. Give M2 coordinates x E lR and () E st = lR/27rZ. The Rosenau
3
BASIC SINGULARITY THEORY FOR RICCI FLOW
469
solution or sausage model (see [311] or [141]) of the Ricci flow is the metric g = u . h defined for t < 0 by (Vl-2.22)
u(x,t) = .x-I sinh (-.xt) , cosh x + cosh .xt
where .x > O. LEMMA A.46 (Rosenau solution and its backward limit). The metric defined by (Vl-2.22) for t < 0 extends to an anczent solution with positive curvature of the Rzcci flow on S2. The Rosenau solution is a Type II ancient solution which gives rzse to an eternal solution zf we take a lzmzt lookmg infimtely far back in time. In particular, if one takes a lzmit of the Rosenau solution at either pole x = ±oo as t ~ -00, one gets a copy of the cigar soliton. Note that the sausage model is an ancient Type II solution which encounters a Type I singularity. 3.5.2. Classification results. The following provides a characterization of the cigar soliton. (See Lemma 5.96 on p. 168 of Volume One.) LEMMA A.47 (Eternal solutions are steady solitons, 2d case). The only ancient solution of the Ricci flow on a surface of strictly positive curvature that attains its maximum curvature in space and time is the cigar (lR2,g~ (t)). The following classifies 2-dimensional complete ancient Type I solutions. (See Proposition 9.23 on p. 275 of Volume One.) PROPOSITION A.48 (Nonflat Type I ancient surface solution is round S2). A complete ancient Type I solution (N2, h (t)) of the Riccz flow on a surface zs a quotzent of ezther a shrmking round S2 or a flat ]R2. We have the following result for 2-dimensional Type II solutions. (See Proposition 9.24 on p. 277 of Volume One.) PROPOSITION A.49 (Type II ancient solution backward limit is a steady, 2d case). Let (M2, g (t)) be a complete Type II anczent solutwn of the Rzccz flow defined on an interval (-00, w). where w > O. Assume there exists a function K (t) such that IR I :S K (t). Then either g (t) is flat or else there exists a backwards limzt that is the cigar solzton. Combining the above results, we obtain the following. lary 9.25 on p. 277 of Volume One.)
(See Corol-
COROLLARY A.50 (Ancient surface solutions). Let (M 2 ,g (t)) be a complete ancient solution defined on (-00, w), where w > o. Assume that its curvature is bounded by some functwn of time alone. Then either the solution is flat or it is a round shrinkmg sphere or there exists a backwards limit that is the cigar.
470
A BASIC RICCI FLOW THEORY
3.6. Necklike points in Type I solutions. (See Section 4 in Chapter 9 of Volume One.) We say that (x, t) is a Type I c-essential point if c (Vl-p. 262a) IRm(x,t)1 ~ T-t > O. We say that (x, t) is a <5-necklike point if there exists a unit 2-form 0 at (x, t) such that (Vl-p. 262b)
IRm -R (0 ® 0)1 ::; <5IRml.
The following result can be used to show that necks must form in Type I solutions where the underlying manifold is not diffeomorphic to a spherical space form. (See Theorem 9.9 on p. 262 of Volume One.) A.51 (Necklike points in 3d Type I singular solutions). Let (M3, 9 (t)) be a closed solution of the Riccz flow on a maximal time interval o ::; t < T < 00. If the normalized flow does not converge to a metric of constant positive sectional curvature, then there exists a constant c > 0 such that for all 7 E [0, T) and <5 > 0, there are x E M and t E [7, T) such that (x, t) is a Type I c-essential point and a <5-necklike point. THEOREM
The analogous result for ancient solutions is the following. (See Theorem 9.19 on p. 272 of Volume One.) THEOREM A.52 (Necklike points in 3d Type I ancient solutions). Let (M3,g(t)) be a complete ancient solution of the Ricci flow with positwe sectional curvature. Suppose that
sup
Itll' R (x, t) < 00
Mx(-oo,Oj
for some 'Y > O. Then either (M, 9 (t)) is isometric to a spherical space form or else there exists a constant c > 0 such that for all 7 E (-00,0] and <5 > 0, there are x E M and t E (-00,7) such that (x, t) is an ancient Type I c-essential point and a <5 -necklike point. 4. More Ricci flow theory and ancient solutions In this section we summarize some additional basic aspects of Ricci flow. We warn the reader that 'basic' here (and in the previous sections) means neither 'should be obvious' nor 'should be known by every graduate student' nor 'should be easy to prove'. 4.1. Strong maximum principle. For solutions with nonnegative curvature operator, the strong maximum principle implies a certain type of rigidity in the case where the curvature operator is not strictly positive (see [179] or Theorem 6.60 in [111]).
4.
MORE RICCI FLOW THEORY AND ANCIENT SOLUTIONS
471
THEOREM A.53 (Strong maximum principle for Rm). Let (M n , g(t)) , t E [0, T), be a solutwn of the Ricci flow with nonnegative curvature operator. There exists J > such that for each t E (0, J), the set
°
Image (Rm [g (t)])
c
A2T* M
%s a smooth subbundle which is invariant under parallel translation and constant in time. Moreover, Image (Rm [g (x, t)]) is a L%e subalgebra of A2T;M ~ so (n) for all x E M and t E (0, J).
As an application of the strong maximum principle we have the following classification result due to W.-X. Shi. THEOREM A.54 (Complete noncompact 3-manifolds with Rc ~ 0). If (M3, g(t)) , t E [0, T), is a complete solution to the Ricci flow on a 3manifold with nonnegative sectional (Ricci) curvature, then for t E (0, T) the universal covering solution (;\it 3 ,g(t)) is either
(1) ]R3 with the standard flat metric, (2) the product (.N2, h (t)) x ]R, where h (t) is a solution to the Ricci flow with positive curvature and.N2 is diffeomorphic to either S2 or]R2 or (3) g(t) and g(t) have positive sectional (Ricci) curvature and hence ;\it3 is diffeomorphic to S3 or]R3 (in the former case M3 is diffeomorphic to a spherzcal space form). 4.2. Hamilton's matrix Harnack estimate. Motivated by the consideration of expanding gradient Ricci solitons, Hamilton proved the following (see [181]). THEOREM A.55 (Matrix Harnack estimate for RF). If (M n , g( t )), t E [0, T), %s a complete solution to the Ricci flow with bounded nonnegative curvature operator, then for any I-form WE COO (AIM) and 2-form U E Coo (A2 M) , we have (A.26) where and P pij
=i= V'p~j
-
V' i!lpj.
REMARK A.56. See the discussion in Section 2 of Chapter 1 for a motivation for defining Mij and Ppij. Chposing an orthonormal basis of cotangent vectors {w a } ~=I at any point (x, t), letting W = wa and U = wa A X for any fixed I-form X, and summing over a, yield the trace Harnack estimate for the Ricci flow (Proposition A.43).
A. BASIC RICCI FLOW THEORY
472
The following is a generalization of the trace Harnack estimate (see [105] and [290]). THEOREM A.57 (Linear trace Harnack estimate). Let (Mn,g(t)) and h (t) , t E [0, T), be a solution to the linearized Ricci flow system:
8
8t gij
=
-2~),
8
-hi)' at = (b..Lh) t).. such that (M, 9 (t)) is complete with bounded and nonnegative curvature operator. h (0) 2: 0, and [h (t)[g(t) ~ C for some constant C < 00. Then h (t) 2: 0 for t E [0, T) and for any vector X we have (A.27) where H = gij hi) . Indeed, (A.27) generalizes Hamilton's trace Harnack estimate since we may take hi) = Rij (under the Ricci flow we have Bt~j = (~L RC)i)·
4.3. Geometry of gradient Ricci solitons. The asymptotic scalar curvature ratio of a complete noncompact Riemannian manifold (Mn, g) is defined by ASCR (g) = lim sup R (x) d (x, 0)2, d(x.O)--->oo
where 0 E M is a choice of origin. This definition is independent of the choice of O. Theorem 9.44 on p. 354 of [111]: THEOREM A.58 (Asymptotic scalar curvature ratio is infinite on steady solitons, n 2: 3). If (Mn, g, f) , n 2: 3, is a complete steady gradient Riccz soliton with sect (g) 2: 0, Rc (g) > 0, and if R (g) attains its maximum at some pomt, then ASCR (g) = 00. Theorem 8.46 on p. 318 of [111]: THEOREM A.59 (Dimension reduction). Let (Mn,g(t)), t E (-oo,w), 0, be a complete noncompact ancient solution of the Riccz flow with bounded nonnegative curvature operator. Suppose there exzst sequences Xi E M , 1r'. ---> 00 ,and A- ---> 00 such that dO(p,Xi) > A-1. and 1. Ti -
w
>
(A.28)
R(y, 0) ~ r;2
for all y E BO(Xi, Ari).
Assume further that there exists an injectivity radius lower bound at (Xt' 0); namely, injg(O) (Xi) 2: &ri for some & > o. Then a subsequence of solutions (Mn, r;2g(r;t), Xi) converges to a complete limit solution (M~, goo (t), xoo) which is the product of an (n - l)-dzmenswnal solutwn (wzth bounded nonnegative curvature operator) with a line.
4.
MORE RICCI FLOW THEORY AND ANCIENT SOLUTIONS
4;3
Theorem A.58 says the following. (See Chapter 6 for a definition of K-noncollapsed. ) COROLLARY A.50 (Dimension reduction of steady solitons). If (Mn,g,f) , n ~ 3, zs a complete steady gradient Riccz sohton whzch is K-noncolla]JsPc! on all scales for some K > and 1.f sect (g) ~ 0, Rc (g) > 0, and zf R VI) attams its maxzmum at some point, then a dilatzon about a sequence of points tendmg to spatial infinity at time t = converges to a complete solutzon (M~, goo (t), x oo ) which zs the product of an (n - l)-dzmenszonal solutwn5 with JR.
°
°
Proposition 9.45 on p. 355 of [111]:
°
PROPOSITION A.51 (Rc > expanders have AVR > 0). If (Mil, 9 (t)) , t > 0, is a complete noncompact expanding gradient Ricci soliton with Rc > 0, then AVR (g (t)) > 0. Theorem 9.56 on p. 362 of [111]: THEOREM A.52 (Steady or expander with pinched Ricci has R exponential decay). If (Afn , g) is a gradient Riccz soliton on a noncompart mamfold with pmched Ricci curvature in the sense that Rij ~ =:RgtJ for some E > 0. where R ~ 0, then the scalar curvature R has exponential decay. Theorem 9.79 on pp. 375-376 of [111]: THEOREM A.63 (Classification of 3-dimensional gradient shrinking solitons with Rm ~ 0). In dimension 3, any nonflat complete shrmkmg gradzent Ricci soliton with bounded nonnegative sectional curvature is ezther a quotzent of the 3-sphere or a quotient of 8 2 x R
4.4. Ancient solutions. Theorem 10.48 on p. 417 of [111]:
°
THEOREM A.54 (Ancient solution with Rm ~ and attaining sup R is steady gradient soliton). If (Mn , 9 (t)) , t E (-00, w) , is a complete solutzon to the Ricci flow with nonnegative curvature operator. posztive Rzccz curvature, and such that SUPMx (-oo,w) R ·lS attained at some pomt in space and tzme, then (M n, 9 (t)) zs a steady gradient Rzcci soliton. Analogous to the above result is the following:
°
THEOREM A.65 (Immortal solution with Rm ~ and attaining sup tR is gradient expander). If (Mn,g (t)), t E (0,00), is a complete solution to the Ricci flow with nonnegative curvature operator, positive Rzcci curvature, and such that sUPMx(O,oo) tR is attained at some point in space and time, then (Mn,g(t)) zs an expanding gradient Ricci soliton. Proposition 9.29 on p. 344 of [111]: 5With bounded nonnegative sectional curvature.
A. BASIC RICCI FLOW THEORY
474
PROPOSITION A.66 (n 2: 2 backward limit of Type II ancient solution with Rm 2: and sect> 0). Let (Mn,g(t)) , t E (-oo,w) , w E (0,00], be a complete Type II ancient solution of the Rzcci flow with bounded nonnegative curvature operator and positive sectional curvature. Assume either
°
(1) M is noncompact, (2) n is even and M is orientable, or (3) 9 (t) is K-noncollapsed on all scales. Then there exists a sequence of pmnts and times (Xi, ti) with ti -> -00 such that (M,gi (t) ,Xi), where gi(t) ~ ~g (ti + Hilt), limits in the Coo pointed Cheeger-Gromov sense to a complete nonflat steady gradient Ricci soliton (M~, goo (t) ,x oo ) with bounded nonnegative curvature operator.
Theorem 9.30 on p. 344 of [111]: THEOREM A.67 (Ancient has AVR = 0). Let (Mn,g(t)), t E (-00,0], be a complete noncompact nonflat anczent solution of the Rzccz flow. Suppose g(t) has nonnegative curvature operator and sup (x,t)EM x (-00,01
IRm g(t) (x) Ig(t) < 00.
Then the asymptotic volume ratio AVR(g(t))
=
°
for all t.
Theorem 9.32 on p. 345 of [111]: THEOREM A.68 (Type I ancient has ASCR = 00). If(Mn,g(t)), -00 < t < w, is a complete noncompact Type I ancient solution of the Ricci flow with bounded positive curvature operator, then the asymptotic scalar curvature ratio ASCR(g(t)) = 00 for all t. 5. Classical singularity theory In this section we continue the discussion of Hamilton's singularity theory and recall some further results concerning the classifications of singularities, especially in dimension 3. An exposition of some of these results, which were originally proved by Hamilton in [186] and [190]' is given in [111]. One of the differences between Hamilton's and Perelman's singularity theories is that in Hamilton's theory, singularities are divided into types, e.g., for finite time singularities, Type I and Type IIa. In Perelman's theory, a more natural space-time approach is taken where singularity analysis is approached via the reduced distance function. Throughout most of this section we shall consider the case of dimension 3. We first consider the case of Type I singularities, which was essentially treated in Volume One (see Theorems A.51 and A.52 above). Applying the compactness theorem and the classification of Type I ancient surface solutions to Theorem A.51 yields the following.
5.
CLASSICAL SINGULARITY THEORY
475
THEOREM A.69 (3d Type I - existence of necks). If (M3,g(t)) is a Type I singular solution of the Rzcci flow on a closed 3-manifold on a maximal time interval ~ t < T < 00, then there exists a sequence of points and times (Xi, ti) with ti - t T such that the corresponding sequence of dilated solutions (M3, gi (t), Xi) converges to the geometric quotient of a round shrinking product cylinder S2 x ~.
°
For the rest of this section we consider Type IIa singular solutions. In this case we invoke Perelman's no local collapsing theorem (see Chapter 6). This has the following two effects on Hamilton's theory. It enables one to apply the compactness theorem to the dilation of Type IIa singular solutions. It rules out the formation of the cigar soliton as a product factor in a singularity model. Classical point picking plus the no local collapsing theorem yield the following result (see Proposition 8.17 of [111]). PROPOSITION A.70 (Type IIa singularity models are eternal). Choose any sequence Ti / T. For a Type IIa singular solution on a closed manifold satisfying JRmJ ~ CR+C
(A.29)
and for any sequence {(Xi, td}, where ti
(A.30)
(Ti - ti) R (Xi, ti)
=
-t
T, satisfying6
max (Ti - t) R (X, t) ,
Mx[O,TiJ
the sequence (M, gi (t), Xi) , where
(A.31)
gi (t) ~ ~g (ti
+ Rilt)
with ~ ~ R (Xi, td,
preconverges to a complete eternal solution
(M~, goo (t), xoo), t
E
(-00,00),
with bounded curvature. The singularity model (Moo, goo (t)) zs nonflat, noncollapsed on all scales for some I'\, > 0, and satisfies R(goo (t)) = 1 = R(goo)
sup
1'\,-
(Xoo, 0) .
Moo x ( -00,00) If n = 3, then the singularzty model has nonnegative sectional curvature. In particular, we have that if (M3, g (t)) is a Type IIa singular solution of the Ricci flow on a closed 3-manifold, then by Theorem A.64, the
smgularity model
(M~, goo (t)) obtained in Proposition A.70 zs a steady
gradient soliton. Since scales for some I'\,
(Moo, goo (t))
is nonflat and I'\,-noncollapsed on all
> 0, the sectional curvatures of goo (t) are positive. 7 Now
by Theorem A.58, the asymptotic scalar curvature ratio of
(Moo, goo (t))
is
6This is a special case of the point picking method described in subsection 4.2 of Chapter 8 in Volume One. 7In the splitting case, we obtain a cigar, contradicting the no local collapsing theorem.
4;6
A BASIC RICCI FLOW THEORY
equal to infinity. Thus we can apply dimension reduction, Theorem A.59, to get a second limit which splits as the product of a surface solution and a line. This second limit is a shrinking round product cylinder 52 x R8 Hence, as a consequence of Hamilton's singularity theory and Perelman's no local collapsing theorem, we have the following result, which complements Theorem A.59. THEOREM A.71 (3d Type IIa - existence of necks). If (M 3 ,g(t)) is a Type IIa singular solutwn of the Ricc~ flow on a closed 3-mamfold, then there exists a sequence of points and times (Xi, ti) such that the corresponding sl'quence (/\/1. Yi (t) ..1"/) COrlveryes to a round shrinking product cylinder 52 x
JR. A precursor to the above result is Theorem 9.9 in Volume One, which basically says that even for a Type IIa singular solution on M3 x [0, T), there exists a sequence of points (Xi, ti) with ti ~ T whose curvatures satisfy (T - ti) [Rm (Xi, ti)[ ~ C for some C > 0 independent of i and at the points (Xi, ti) the curvature operators approach that of 52 xlR after rescaling. However Theorem 9.9 in Volume One does not directly imply the existence of a cylinder limit because, for the sequence (M, gl (t) , Xi) , it not a priori clear that the curvatures are bounded in space at finite distances from X t independent of i, even at time O. The reason for this is that globally, the curvature of gi (0) , whose norm is 1 at X t , may be unbounded since the solution is Type IIa whereas the point (Xi, tt) may, for example, have curvature (T - ti) [Rm (Xi. td[ :S C for some C < 00 independent of i. Since finite time singularities are either Type I or Type IIa, we obtain the existence of necks for all finite time singular solutions on closed 3-manifolds.
80t herwise we again get a cigar limit.
APPENDIX B
Other Aspects of Ricci Flow and Related Flows 1. Convergence to Ricci solitons Given that convergence to a soliton plays a role in proving the convergence of the Ricci flow on compact surfaces (see Chapter 5 ill Volume One), it is reasonable to ask if noncompact steady solitons playa role as limiting geometries for the Ricci flow on complete surfaces (or higher-dimensional manifolds) . Since steady soliton solutions are, by definition, evolving by diffeomorphism, even if we start near a soliton metric, we cannot expect pointwise convergence of the Ricci flow unless we take the diffeomorphisms into account. We use the following notion of convergence [373]: DEFINITION B.l. Let g(t) satisfy the Ricci flow on a noncompact manifold Mn for 0 :S t < 00 and let 9 be a metric on M. We say that g(t) has modified subsequence convergence to 9 if there exist a sequence of times ti ---? 00 and a sequence of diffeomorphisms ¢t of M such that ¢ig(fi) converges uniformly to 9 on any compact set.
Since the Ricci flow is conformal on surfaces and preserves completeness, it makes sense to begin with metrics in the conformal class of the Euclidean metric, i.e., metrics of the form (B.1)
We can describe the overall 'shape' of such metrics in terms of the aperture and circumference at infinity. First, the aperture is defined by
A( 9 ) -;- l'1m L(aB (0, r)) , 1'~00 21Tr ...!...
°
where the B (0, r) are geodesic balls about some chosen point E IR? (The choice does not affect the value of the limit.) For example, the flat metric has aperture one, the cigar metric has aperture zero, and surfaces in 1R3 that are asymptotic to cones have aperture 0:/(21T), where 0' is the cone angle. The circumference at infinity is defined by
Coo(g)
~
sup inf {L(aD) I K is compact, D is open and KeD}. K
D
For flat and conical metrics, Coo is infinite, while Coo is finite for the cigar. Let R_ = max{ -R, a}. 477
478
B
OTHER ASPECTS OF RICCI FLOW AND RELATED FLOWS
THEOREM B.2 (Wu [373]). Let go be a complete metric on ]R2, of the form (B.I), with Rand lV'ul (measured using go) bounded, and I R_ dJ-l finite. Then a solution to the Ricci flow exists for all time, the aperture and circumference at infinity are preserved, and the metric has bounded subsequence convergence as t ---t 00. If R(go) > 0 and A(go) > 0, then the limit is flat. If R(go) > 0 and Coo(go) < 00, then the limit is the cigar. Note that Coo being finite implies that A = O. However, there are plenty of complete metrics with A = 0 and Coo = 00, for which the limit of the Ricci flow is not classified. An example of a surface of positive curvature with A = 0 and Coo = 00 is the paraboloid (M2,g) ~ {(x,y,z) E]R3: z = x 2 +y2}. OUTLINE OF PROOF. Short-time existence follows from the BernsteinBando-Shi estimates. As with the Ricci flow on compact surfaces (see Section 3 of Chapter 5 in Volume One), long-time existence is proved by using a potential f such that D.f = R (where D. denotes the Laplacian with respect to g) and examining the evolution of the quantity h ~ R + lV'fI2.
In fact, for metrics of the form (B.I) we can use
8u
-
at
f = -u. Because
=D.u=-R
and
%t lV'ul 2 = D.1V'uI 2 -
21V' 2uI 2,
we have 8h = D.h _ 21MI2
8t
'
where M is the symmetric tensor with components Mij
= V'iV'jU + ~Rgij.
Long-time bounds for lV'ul and R (and higher derivatives) follow. The bounds on IRI imply that the metric remains complete; in particular, the length of a given curve at time t > 0 is bounded above and below by multiples of its length at time zero. By a theorem of Huber [210], the hypothesis I R_ dJ-l < 00 implies that I RdJ-l ::; 47l'X (M) on a complete surface M. In particular, I IRI dJ-l is finite, and this is preserved by the Ricci flow. Finite total curvature, together with bounds on IV'RI at any positive time, imply that R decays to zero at infinity. One then shows that Coo (g) and A (g) are preserved under the flow. In the special case when R(go) > 0, limt-+oo eu(x,y,t) exists pointwise and is either identically zero or positive everywhere. In the latter case, 8u/8t = - R implies that 00 Rdt is bounded, and hence the limiting metric is flat. In the general case, we may define diffeomorphisms (/>t(x, y) = e- u(0,0,t)/2(x, y), so that g(t) = c/>;g(t) is constant at the origin. Then uniform bounds on
10
1.
CONVERGENCE TO RICCI SOLITONS
479
the derivative of the conformal factor for 9 give subsequence convergence, on any compact set, to a metric g. If R(go) > 0 and Coo < 00, then using the Bernstein-Bando-Shi estimates, one can show that after a short time T, JIR2 IMI 2dJ.L is bounded uniformly in time (where dJ.L indicates measure with respect to the evolving metric). The evolution equation for IMI 2dJ.L then implies that
1 (l2 00
J
21V'M12
+ 3RIMI2dJ.L) dt < 00.
Thus, either M vanishes or R vanishes for the limiting metric g. If M = 0, then 9 is a gradient soliton; by Proposition 1.25, it is either flat or the cigar metric. However, Coo < 00 precludes a flat limit. If R(go) > 0 and A > 0, then one may use the Harnack inequality to show that tRmax is bounded. It follows that the limit 9 is flat in this 0 case. One may also study the Ricci flow for a solution of the form (B. 1) in terms of the nonlinear diffusion equation satisfied by the conformal factor v ~ eU , 8v (B.2) 8t = .6. log v, where .6. is the standard Laplacian on ]R2. (Note that v the cigar, where we write x = (x, y).)
= l/(k + Ix1 2) for
REMARK B.3. As pointed out by Angenent in an appendix to [373], the equation (B.2) is a limiting case of the porous medium equation
8v = .6.vm 8t as the positive exponent m tends to zero. For, substituting t = gives (B.3)
r:
~~ = .6. (vm and taking m
~
T /m
in (B.3)
1)
0 gives 8v/8T = .6. (log v).
In connection with Ricci flow on ]R2 we also have the following result, which says that metrics that start near the cigar converge to a cigar (in the same conformal class), but under weaker assumptions than in the above theorem. THEOREM B.4 (Hsu [206]). Suppose that vo E Lfoc(]R2) for p > 1, 0 ~ vo ~ 2/(,6lxI 2) for,6 > 0, and vo - 2/(,6(lxI 2 + ko)) E Ll(]R2) for ko > O. Then there exists a unique positive solution of (B.2), defined for 0 < t < 00, such that limt-+o v = Vo in LIon any compact set and such that
t~~ e2{3t v (e 2{3t m
Ll(]R2), for some kl > o.
x, t)
= ~(lxI2 + k1)-1
ISO
B.
OTHER ASPECTS OF RICCI FLOW AND RELATED FLOWS
In higher dimensions, there are not many results. However, consider complete warped product metrics on ]R3, of the form (B.4) where gean is the standard metric on S2. (Recall that the sectional curvatures VI. V2 of such metrics are given in terms of'l17 by (1.58).) For such metrics, we call prove convergence to a soliton if the curvature is positive and bounded and the manifuld "opens up" like a paraboloid. THEOREM B.5 (Ivey [220]). Let 9 be a complete metric on form (B.4). Suppose
]R3
of the
(B.5) and s'uppose
(B.6) for positzve constants C and Z, and (B.7)
liminf ('--'00
(aDr w
2)
> O.
Then the solution of the Rz('cz flow 'Wzth g(O) addztion. (B.~)
limsup (-->=
(aDr w 2 )
=9
exists for all time. If, zn
< 00,
then the flow converyes to a mtationally symmetric steady gradzent soliton, zn the sense of the C=-Cheeger-Gromov topology (see Theorem 3.10).
Intuitively, the condition (B. 7) means that the area of a sphere centered at the origin grows at least as fast as for a paraboloid, while (B.8) means that the sphere area grows no faster than a paraboloid. In other words, the metric becomes fiat as r - t 00, but not too fiat. (By contrast, the result of Shi [330], which gives convergence of the Ricci fiow to a fiat metric, assumes that sectional curvatures fall off like r-(2+c).) The condition V2 :::; ZVl means that as the sectional curvature Vl along the planes tangent to the spheres becomes fiat as r - t +00, the sectional curvature V2 of the perpendicular planes also becomes fiat. Of course, for the Bryant soliton of subsection 3.2 of Chapter 1, V2 falls off much faster than VI. However, it is not difficult to construct other metrics that satisfy the conditions in the theorem; see [220] for details. For metrics on ]R3 of the form (B.4), the warping function satisfies a quasilinear heat equation
aw
at = w
, , 1 - (w l )2 -
w
=
-(Vl
+ V2)W.
1. CONVERGENCE TO RICCI SOLITONS
481
However, the time-derivative a/at under the Ricci flow and the radialderivative a/or do not commute; in fact,
a a oJ [ at' or = 2V2 or . We now outline the proof of Theorem B.5; again, details may be found in [220]. First, long-time existence is proved by obtaining an estimate, depending only C and Z, for Iww"'l. Then, convergence to a soliton is proved by examining the evolution of the quantity
Q ~ R+ ((VI +V2)W/w,)2. Comparing with (1.58) shows that Q coincides with R+ 1\7112 when 9 is the Bryant soliton. Thus, Q is constant for the soliton. In general, it evolves by the equation
(B.9)
(8t8) - ~ Q
- 2
where
(8Q )2
(ww')2
= - (1
+ (w')2 +WV2)2 8r
( WV2 w'
WW'
)
oQ
+ 1 + (w')2 + WV2 (VI + V2) or'
~ = (~)2 8r
_
2W ' W
~ 8r
is the Laplacian with respect to 9 for functions depending on rand t only. Given (B.5) and (B.6), the paraboloid condition (B.8) is equivalent to Q having a positive lower bound. In fact, (B.6) implies that (B.10)
(ww')2 ~ (1
+ Z)2/Q.
As part of the proof of long-time existence, one shows that conditions (B.5) and (B.6) persist in time. These, together with (B.10), imply that
_~) ~ _ A (oQ) 2 _ (~ at Q or
B 1 8Q I
Or
for some positive constants A, B. Applying the maximum principle to the corresponding inequality for rP = l/Q shows that the lower bound on Q persists in time. Finally, existence of a limiting metric is proven using the compactness theorem, and showing that it is a nontrivial soliton comes by appealing to Theorem A.64 and the positive lower bound for Q. REMARK B.6. The generalization of Theorem B.5 to rotationally symmetric Ricci flow in higher dimensions should be straightforward. It may even be possible to generalize at least the long-time existence to the nonrotationally-symmetric case by finding a generalization for the quantity Q (open problem). It would also be interesting to find conditions under which the flow converges to the product of the cigar metric with the real line.
482
B. OTHER ASPECTS OF RICCI FLOW AND RELATED FLOWS
2. The mean curvature flow In this section we give a brief introduction to the mean curvature flow and some monotonicity formulas. It is interesting to compare these monotonicity formulas for the mean curvature flow to those for the Ricci flow. We also refer the reader to the books by Ecker [132] and X.-P. Zhu [386]. 2.1. Mean curvature flow of hypersurfaces in Riemannian manifolds. Let ('pn+l, 9P) be an orientable Riemannian manifold and let M n be an orientable differentiable manifold. The first fundamental form of an embedded hypersurface X : M ---+ P is defined by 9 (V, W) ~ 9p (V, W)
for V, W E TxX (M). More generally, for an immersed hypersurface, we define 9 (V, W) ~ (X*V,X*W) for V, W E TpM. Let 1/ denote the choice of a smooth unit normal vector field to M. The second fundamental form is defined by h (V, W) ~ (Dvl/, W)
=-
(DvW, 1/)
for V, WE TxX (M), where D denotes the Riemannian covariant derivative of (P, 9P) and ( , ) ~ 9p ( , ). To get the second equality in the line above, we extend W to a tangent vector field in a neighborhood of x and use (1/, W) == O. In particular, 0 = V (1/, W) = (Dvl/, W) + (DvW, 1/). The mean curvature is the trace of the second fundamental form: n
H~ Lh(ei,ei), i=l
where {ed is an orthonormal frame on X (M) . A time-dependent immersion X t = X (', t) : M ---+ P, t E [0, T), is a solution of the mean curvature flow (MCF) of a hypersurface in a Riemannian manifold if
(B.ll)
ax (p, t) = H. . (x (p, t)) ~ -H (p, t) . 1/ (p, t), at
p E M, t
E
[0, T),
where ii is called the mean curvature vector. When the X t are embeddings, we define M t ~ X t (M). From now on we shall assume that the X t are embeddings, although for the most part, the following discussion holds for immersed hypersurfaces. Let {xi} ~=1 denote local coordinates on M so that {¥xt} are local coordinates on M t . We have
9ij
~ 9 (~~, ~~) = (~~, ~~),
hij
~ h (aaX , aax) xt xJ
=
/aax,DMI/) = xt ax)
\
- /
\
DM aax, ,1/) . ax' x J
2.
THE MEAN CURVATURE FLOW
483
Note that H = gij h ij and
(B.12)
D Mil = hjkg
ke
axJ
ax ax e'
We have the following basic formulas for solutions of the mean curvature flow. LEMMA B. 7 (Huisken). The evolution of the first fundamental form (induced metric), normal, and second fundamental form are given by the following:
a
(B.13)
at gij = -2Hhij ,
(B.14)
DltY at
(B.15)
ata hij =
= \lH, ke \l/VjH - Hhjkg hei
= t:"hij -
+ H (Rmp )lIijll 2Hhikhkj + Ihl 2 hij
+ hij (Rc p) hik (Rc P ) jk - hkj (Rc P )ik + 2 (Rm P ) kije hke + hki (Rm P ) II kj II + hkj (Rm P ) IIkill - Di (Rcp)jll - D j (RcP)ill + DII (RCP)iJ'
(B.16)
111/ -
a
(B.17)
at H = t:"H + Ihl
(B.18)
!!.-dH
2
H
+H
(Rc p)1II/ ,
at ,.. = -H 2dH,..,
where \l is the covariant derivative with respect to the induced metric on the hypersurface and dp, is the volume form of the evolving hypersurface.
REMARK B.8. Technically, we should consider these tensors (or sections of bundles) as existing on the domain manifold M. However, we shall often view them as tensors on the evolving hypersurface M t = X t (M). The time-derivative of the unit normal is expressed slightly differently since it is actually the covariant derivative of II in the direction ~~ along the path t I--t X t (p) for p E M fixed. On the other hand, if we view 9 and h as on the fixed manifold M, we have the ordinary time-derivatives, whereas if we view 9 and h as on the evolving M t , then the time-derivatives are actually Da. at
PROOF. While carrying out the computations below, keep in mind that the inner product of a tangential vector with a normal vector is zero. The
-18!
B OTHER ASPECTS OF RICCI FLOW AND RELATED FLOWS
evolution of the metric is given by
%t gil = ( D%;. (aa~) , ~~ ) + ( ~~ , D:~ (aa~)) = -H (DQLV' aT' =
-2Hh i.!,
aa~) x - H(aa~' x Dax ax) v) J
g;) (g;, v) = O. This is (B.13).
where we used (v, = Since (~~, v) = 0 and
X 0= aat (v, aax~X ) = (Daxv, (-Hv)) at aax~ ) + \/v,DcJx a,' the normal v evolves by
··aHaX Dax v = gtJ_. at axt -. ax) = V' H,
(B.19)
where V' H is the gradient 011 the hypersurface of H. This is (B.14). The evolution of the second fundamental form is (we use [~"'i, g~] = 0)
~ hiJ = -aat (Dax v) ax' aa~, Xl
ot
(Daxat (Dax aX) axJ,Daxv) J ,v) - (Dax axi ax axi ax at = _ (Dax (DQJ£ (ax)) ,v) _ (Rmp (ax, ax) ax,v) ax' ax) at at axt ax.! ax ,Daxv) - (DQL ax' axl at = -
(Dax ax' (aa~v xl + HDax ax) v) ,v) + H(Rmp (v, aa~) xt aax, xJ v) - (DQL V'H) ax' aa~, xJ kC ax k aH 2H = aa·a . + Hhjkg ( Daxa ",v ) +H(Rmp)ViJV-fiJ·a k xt xJ ax' x< x = V'iV'jH - Hhjklfhfi + H(Rmp)vijv' =
where we used (B.ll), (B.19), and (B.12), and where
ax ' ax) fiJk = 9kf ( D QL -a ax"" ax' x J are the Christoffel symbols and V' is the covariant derivative with respect to the induced metric on the hypersurface. This is (B.15).
2 THE MEAN CURVATURE FLOW
-185
'fracing the above formula, we see that the mean curvature evolves by
This is (B.17). Equation (B.18) follows from (B.13) and
Finally we rewrite the evolution equation for hij to see (B.16). Recall that the Gauss equations say that for any tangent vectors X, Y, Z, W on M, (RmM) (X,Y,Z,W)
=
(Rmp) (X,Y,Z,W)
+ h (X, W) h (y, Z) -
h (X, Z) h (Y, W) .
'fracing implies (in index notation)
(RCM)il = (RcP)if - (Rmp)vifv
+ Hhif -
h;e·
The Codazzi equations say for any X, Y, Z E T M, (\lxh) (Y, Z) - (\lyh) (X, Z)
=-
(Rmp (X. Y) Z, v) .
In index notation, this is \l ihkj
= \l khij
- (Rm P )ikjv .
'fracing over the Y and Z components in the hypersurface directions, we have \l H - div (h)
=-
Rcp (v).
To obtain (B.16) from (B.15) we note that, by using the Codazzi equations and Gauss equations, we obtain
(B.20) where Rcp (v)
~
(Rcp))vdx) is a I-form on M. Note that
(rp
)r = -~v (:~, ~;) = j
-hi)
and similarly k
(rp )iv
1 kl / ax ax) axi ' axl
="29 v \
k£
=9
hit·
486
B. OTHER ASPECTS OF RICCI FLOW AND RELATED FLOWS
Thus, considering Rc P as a 2-tensor on 'P and changing its covariant derivative to the one with respect to gp in the formula (B.20), we have
'h'VjH
= \h'Vkhkj - Di (Rcp)jv - (fp)rj (Rcp)w - (fp)7v (RCP)jk = \1 /iJ khkj - Di (Rc p) jv
+ hij (Rcp)w -
hf (RcP)jk'
Commuting derivatives and applying the Codazzi and Gauss equations, we compute
\1i\1jH = \1k\1ihkj - (RCM)uh£j - (RmM)ikj£hk£ - Di (RcP)jv
+ hij (Rcp)w - hf (RcP)jk = \1 k\1 khij - \1 k (Rm P )ikjv - (Hhu - h7£) h£j - (hi£hkj - hijhk£) h k£
+ (Rm P )vuv h£j - (Rm P )ikj£ hk£ - Di (Rcp)jv + hij (Rcp)w - hf (RcP)jk = b..hij - Hh7j + Ihl 2 hij - Dk (Rm P )ikjv + hki (Rmp)vkjv + hkk (RmP)ivjv - h1 (Rmp)ikj£ - (Rc p)u h£j + (Rm P )vuv h£j - (Rm P )ikj£ hk£ - Di (Rcp)jv + hij (Rcp)w - hf (RCP)jk' - (Rc p)u h£j
where h~ ~ hikhkj and in the third line (Rmp )ikjv dx i 0 dx k 0 dx j is considered as a 3-tensor. This is known as Simons' identity. Hence, under the mean curvature flow
ata hij = \1i\1jH -
k£ Hhjkg h£i + H (Rmp )vijv
= b..hij - 2Hh;j
+ Ihl 2 hij
+ Dv (RcP)ij - Di (Rcp)jv (Rc P )i£ h£j + (Rm P ) vi£v h£j - (Rm P )ikj£ hkl
- Dj (RCP)iv -
+ hij (Rcp)w - hf (RcP))k + hki (Rmp)vkjv - h1 (Rmp)ikj£' where we used the second Bianchi identity:
o
2
THE MEAN CURVATURE FLOW
487
EXERCISE B.9. Compute gt (lhl2 - ~H2) . See §5 of Huisken [212] for a study of under what conditions on (P, 9P) the pinching estimate
Ihl 2- .!. H2 ~ C . H 2- t5 n
holds for some 8 > 0 and C <
00.
EXERCISE B.lD. Let Xt : Mn -+ pn+l be a hypersurface evolving by mean curvature flow where the metric 9p (t) on P evolves by Ricci flow. Compute gt hij . HINT. The terms - Di (Rc p) jll - Dj (Rc P )ill + DII (Rc P )ij are cancelled by new terms introduced by the Ricci flow. Note that the above terms represent the evolution of the Christoffel symbols under the Ricci flow. 2.2. Huisken's monotonicity formula for mean curvature flow. When the ambient manifold pn+l is euclidean space IR.n+1, Lemma B. 7 says the following.
LEMMA B.ll. The evolution of the first fundamental form (i.e., induced metric), normal, and second fundamental form are given by the following:
a
(B.2l)
at 9ij
(B.22)
Daxl/ at
(B.23)
athij
(B.24)
%tH
(B.25)
~dJ-t = at
a
= - 2Hhij,
= VH, 2
= tlhij - 2Hhikh kj + Ihl hij , = tlH + Ihl2 H, -H 2 dJ-t.
For the mean curvature flow there is a monotonicity formula due to Gerhard Huisken (see [213], Theorem 3.1). THEOREM B.12 (Huisken's monotonicity formula). Let X t : Mn -+ n IR. +1 , t E [0, T), be a smooth solution to the mean curvature flow (B.ll). Then any closed hypersurface M t ~ X t (M) evolving by MCF satisfies (B.26)
.!i
r (47f7)-~ e-~dJ-t
= -
dtlMt
where
7 ~
(H _
(x,I/))2 (47f7)-~ e-~dJ-t, 27
T - t.
PROOF. Let u (B.27)
r
lMt
~
r
~
atlMt
n
(47f7)-"2 e-
udJ-t
=
r
lMt
~ 47"
•
Using (B.ll) and (B.25), we compute
(~+ H (X, 1/) _ IX~2 _ H2) udJ-t. 27
27
47
488
B. OTHER ASPECTS OF RICCI FLOW AND RELATED FLOWS
Using the facts that (B.28) where XT ~ X - (X, v) v (~ is the Laplacian with respect to the induced metric), we have
r
}Mt
H (X, v) udJL
=-
r
}Mt
= 1Mt
(X,
~X) udJL
(IV XI 2 11, + ~ (V IXI 2 , vu)) dJL
r (n1l - ~ Ix
=
}Mt
27
I
T 2 u)
dJL.
Multiplying this by 2~ and adding the difference (which is zero) into (B.27), we have
~
r udJL }Mt r (H (X, v) _ (X, ~)2 _ H2) udJL =
dt}Mt
7
=_
47
r (H _(x,V))2 udJL.
}Mt
27
o EXERCISE
B.13. Verify the formulas in (B.28). B.13. We have {)2 X k aX ViVjX = axiaxj - r ij ax k
SOLUTION TO EXERCISE
so that
Hence Tracing this yields ~X Next we compute
=-
IVXI and
2
ViVjX = -hijv. H v. ··ax aX ax' axJ
..
= l J - · - · =lJ gi · =n J
2 THE MEAN CURVATURE FLOW
-189
Formula (B.26) is useful for studying Type I singularities of the mean curvature flow, where
(T - t)
sup
Ihl 2 < 00.
Mnx[o.T)
For Type I singularities there is a natural way to rescale the 1vICF equation:
X-(;:\ p, t) Then
~
1 X (p, t), J2 (T - t)
1 where t (t) ~ log y'T=t T - t
Mi ~ Xi (M n ),
dX -Hi/+X dt and (B.26) becomes the normalized lllonotonicity formula: -_ =
From this one can prove the following (see Proposition 3.4 and Theorem 3.5 in [213]). THEOREM B.14 (MCF convergence to self-similar). Suppo.~e X t : Mn ~ lRn+l. t E [0, T), 2S a Type I smgular solutwn to the mean curvature flow. For every sequence of times ti ~ 00, there exists a subsequence such that Mi, converges smoothly to a smooth immersed limit hypersurface Moo wh2ch is selJ-s2mzlar:
Huiskell's monotonicity formula was generalized by Hamilton [184] as follows. Let X t : M n ~ pN, t E [0, T), be a submanifold of a Riemannian manifold (pN, g) evolving by the mean curvature flow l gt X = ii and suppose u : pN x [0, T) ~ lR is a positive solution of the backward heat equation:
au at = -b..pu.
Then
!!.- [(T - t)(N-n)/2 dt
r UdP,] r Iii - (\7 log u)~ 12 r tr ~ (\7\7 log + 2 (T1- t ) g) }Xt(M)
= _ (T - t)(N-n)/2
udp,
}Xt{M)
- (T -
t)(N-n)/2
u
udp"
}Xt(M)
where dl-l is the volume form of the submanifold X t (M) and tr ~ denotes the trace restricted to the normal bundle of X t (M) C pN. By Hamilton's 1In all ('odimensions the mean curvature vector fj is defined by tracing the second fundamental form, which takes values in the normal bundle.
490
B
OTHER ASPECTS OF RICCI FLOW AND RELATED FLOWS
matrix Harnack inequality for the heat equation (see Part II of this volume), if (pN, g) has parallel Ricci tensor and nonnegative sectional curvature, then \7\7logu + 2(-1-t)g ~ 0 and hence
!i dt
[(T - t)(N-n)/2 (
UdJ.t] ::;
o.
}Xt(M)
Equality holds if and only if
ii -
(\7logu)1-
=0
and
EXERCISE B.15. Prove Hamilton's generalization of Huisken's monotonicity formula. 2.3. Monotonicity for the harmonic map heat flow. Similar inequalities hold for the harmonic map heat flow and the Yang-Mills heat flow. The original monotonicity formula for the harmonic map heat flow was discovered by Struwe [342] and extended by Chen and Struwe [90] and Hamilton [161]. The monotonicity formula for the Yang-Mills heat flow appears in [184]. THEOREM B.16 (Struwe and Hamilton). Let (Mn,g) and (Nm,h) be Riemannzan manzfolds where M zs closed. If F : (M, g) ~ (N, h) zs a solutwn to the harmonic map heat flow
8F
at =
!:l.g,h F ,
U : M ~ ~ is a posztive solutwn to the backward heat equation and r ~ T - t, then :t (r
a;: = -!:l.gU,
1M IdFI2 UdJ.t) = -2r 1M I!:l.F + \ d:, dF )1 udJ.t - 2r 1M (\7i\7jU - \7 iUu\7 j u + 2~ U9ij ) diF· djFdJ.t, 2
where diF ~ dF (8/8x i ). See the above references for applications of the monotonicity formula to the size of the singular set of a solution.
3. The cross curvature flow In this section we present the details to results for the cross curvature flow stated in subsection 4.2 of Chapter 3 of Volume One.
490
B.
OTHER ASPECTS OF RICCI FLOW AND RELATED FLOWS
matrix Harnack inequality for the heat equation (see Part II of this volume), if (pN , g) has parallel Ricci tensor and nonnegative sectional curvature, then \7\7 log u + 2(i-t)g ~ 0 and hence
~ [(T - t)(N-n)/2 dt
r
UdP,]
s o.
}Xt(M)
Equality holds if and only if ....
H - (\7logu)
.1
=0
and (\7\7 log U
+ 2 (Tl_ t)g).l = O.
EXERCISE B.15. Prove Hamilton's generalization of Huisken's monotonicity formula. 2.3. Monotonicity for the harmonic map heat flow. Similar inequalities hold for the harmonic map heat flow and the Yang-Mills heat flow. The original monotonicity formula for the harmonic map heat flow was discovered by Struwe [342] and extended by Chen and Struwe [90] and Hamilton [161]. The monotonicity formula for the Yang-Mills heat flow appears in [184]. THEOREM B.16 (Struwe and Hamilton). Let (Mn, g) and (Nm, h) be Riemannian manifolds where M is closed. If F : (M,g) - t (N, h) zs a solution to the harmonic map heat flow
8F
at = D.g,h F , M and T
U:
~ is a positive solution to the backward heat equatzon ~~ = -D.gu, T - t, then
-t
~
See the above references for applications of the monotonicity formula to the size of the singular set of a solution. 3. The cross curvature flow In this section we present the details to results for the cross curvature flow stated in subsection 4.2 of Chapter 3 of Volume One.
3. THE CROSS CURVATURE FLOW
491
3.1. The cross curvature tensor. Let (M 3 ,g) be a Riemannian 3manifold with either negative sectional curvature everywhere or positive sectional curvature everywhere. Recall the cross curvature tensor c is defined on p. 87 of Volume One by
(B.29)
r,· ~tJ
==. det E
(E- l ) .. tJ
= ~J.LipqJ.LjrS E E 2 pr qs
-_ 81 J.Lpqk J.LrsfRUpq R kjr s ,
(B.30)
where Eij ~ ~j -1Rgij is the Einstein tensor, det E ~ ~:~!,:, and J.Lijk are the components of dJ.L with indices raised. The equalities in the definition of Cij are a consequence of the following identities. LEMMA
B.17 (Elementary curvature identities).
(a) (B.31) (b)
d et
E (E
-l) ij -- 8J.L 1 pqk rsfv. R. J.L .LL1,fpq kJrs·
(a) We compute in an oriented orthonormal basis so that J.L123 = 1 and Eij is diagonal. In such a frame, we have En = -R2332, etc. Given f, define a and b so that abf is a cyclic permutation of 123. Then PROOF.
J.L123 =
J.Lpqk J.L rsf Rkjrs
= 2J.Lpqk RkJab = 2J.Lpqa 8j Rabab + 2J.Lpqb 8j Rbaab =
2 (J.Lpqa 8j -
J.Lpqb 8j ) Eu
=2~~~-~~~-~~~+~~~)~ =
2
(81 (8~ -
= 2Eqf. 8~
-
8~8J)
2EPf.
- 8~ (8J - 81 8;)) Eu
8J.
(b) By part (a), we have 1
8J.L
pqk rsf.R R _ 1 v. Emf (l
1 ( Eqf. RUjq - EPf~fpj .) = 4 1
= -2"Epr ~rpj. In a frame as in part (a), an
aij
~
-1 Epr ~rpj is diagonal and
= - -1E 22 R1221 2
= E22E33
-
1 2
-E
33
R1331
= detE (E- l )n'
492
B
OTHER ASPECTS OF RICCI FLOW AND RELATED FLOWS
and similarly for
a22
and
o
a33.
A nice property of the cross curvature tensor is the following result due to Hamilton.
LEMMA B.18 (Bianchi-type identity for Cij). If (M 3,g) has negatwe sectzonal curvature (or positive sectional curvature), then c zs a metrzc and (B.32)
This zmplies id:
(M 3 ,c)
---t
(M 3 ,g)
is a harmonic map.
REMARK B.19. We may think of this result a::; dual to the fact that if Rc is positive (negative) definite, then the identity map id : (Mn,g) ---t (Mn,±Rc) is a harmonic map (Corollary 3.20 on p. 86 of Volume One). PROOF. Using \liEiJ = 0 (which follows from the contracted second Bianchi identity), we compute (c-1)ij\liCjk
=
(detE)-1 Eij\li (detE(E- 1)Jk)
=
_ (det E) 1 \l k det E
1
= "2 \l k log det c
="21 (c- 1) ~j \lkCij, where det c ~ det Ci] / det gil . Now given two Riemannian metrics c and 9 on a manifold M, the Laplacian of the identity map id : (M, c) ---t (M, g) is given by
(~id)k = (c- 1 )ij (r~j
-
~fj)
= - (c_ 1)ke (c- 1)ij (\liCje
(B.33)
+ \ljCif -
\leCij) = 0,
rfj
where and ~fj denote the Christoffel symbols of 9 and c, respectively. 0 Here we used (B.32). REMARK B.20. The above proof is the solution to Exercise 3.23 of Volume One. 3.2. The cross curvature flow and short-time existence. We say that a I-parameter family of 3-manifolds (M 3 , 9 (t)) with negative sectional curvature is a solution of the cross curvature flow (XCF) if
8
()tg
= 2c.
Likewise, if g(t) has positive sectional curvature, we say that a solution if
8
8t g
= -2c.
We have the following result due to Buckland [34].
(M 3 ,g(t))
is
3. THE CROSS CURVATURE FLOW
493
THEOREM B.21 (XCF short-time existence). If (M3,gO) zs a closed 3manifold with ezther negative sectional curvature everywhere or positive sectzonal curvature everywhere, then a solution 9 (t), t E [0, E), to the cross curvature flow with 9 (0) = go exzsts for a short time. PROOF. We only consider the case of negative sectional curvature and leave the case of positive sectional curvature as an exercise for the reader. By Remark 3.4 on p. 69 of Volume One, if %.sgzl = 'Vi) is a variation of the metric gZJ' then
8
1
at Rijkf = '2 ("
i"
f'Vjk
+ "J "k'Vif -
"l"
k'V)f - " j " f'V 1h')
1
+ '2 gpq (Rijkp'Vqf + RiJPf.'Vqk) ,
(B.34) so that
where the dots denote terms with 1 or fewer derivatives of 'V. Applying (B.31) to (B.30), we obtain
8 Cij -_ 8 1 ( 8s 8 R if.pq ) /-Lpqk /-L TSe R kJTS 8s
= _~ ( 8
_
~
(
8
8 2 'Vfp 8x i 8x q
8 2 'Vfp 8x j 8x q
+
+
8 R kJTS ) + 81 ( 8s
/-L
pqk TSe
/-L
8 2 'Vzq _ 8 2 'Vfq _ 8 2vip ) Emf. 8x i 8xp 8x R8x q 8 x i'.8x p
8 2 'Vjq _ 8 2 'Vfq _ 8 2 'Vjp ) Emf 8x f 8.rP 8xj8xp 8x f 8xq
TJ L
LtRpq
(8 P8q
+ ... _
8P 8q)
J m
(8 P 8q l
m
_
m
J
8P 8q ) m
l
+ .... Thus the linearization of the map X which takes 9 to 2c is a second-order partial differential operator. Its symbol is obtained from 2%sCiJ by replacing 8~i by a cotangent vector (i in the second-order terms:
here DX (g) denotes the linearization of X at g. Since the sectional curvature is negative, Emf is positive, which ill turn implies that the eigenvalues of the symbol are nonnegative. In analyzing the symbol (JDX (g) ((), without loss of generality we may assume (1 = 0 and (2 = (3 = O. Then
494
B. OTHER ASPECTS OF RICCI FLOW AND RELATED FLOWS
Hence, if we take el ® e3, e2
® e3
+ e3 ® e2
as
el ® e2
eI,
a basis for
+ e2 ® el, el ® e3 + e3 ® el, e2 ® e2, e3 ® S2T* M,
then
-2ElmVlm
Vll
+ EmiW.m + El1vl1 + El1vl2 + Ellvl3
- E lf V£2 -E lf Vf3
I7DX (g) (()
El1v22 Ellv33 El1v23 E 22 v22
+ E33 v33 + 2E23 v23
- El2v22 - El3v23 - El2v23 - El3 v33 Ellv22 Ellv33 El1V23
The symbol
as
a matrix is given by
I7DX (g) (()
0 0 0 0 0 0
=
0 0 0 0 0 0
E22
0 0 0 0 0 0
E33
_E12
0
2E 23 _E13
0
_E13
_E12
Ell
0
0 0
El1
0 0
0
El1
Since this matrix is upper triangular, its eigenvalues are 0 and El1, which are nonnegative. To eliminate the degeneracy, we apply DeT'urck's trick. As in (3.29) on define the p. 80 of Volume One, given a fixed torsion-free connection vector field W = W by
t,
(g, t)
Wk
==. gpq
(rk _ pq
tkpq. )
Consider the second-order operator
Y (g) We have
-=;=
CWg.
3. THE CROSS CURVATURE FLOW
495
Hence laDY (g)
(() vjij = gPq ((i(pVqj
+ c5jI VIi -
= c5i1 VIj
assuming (1
+ (j(pVqi -
(i(jVpq )
c5iI c5j 1 gpq v pq ,
= 0 and (2 = (3 = O. In other words,
aDY (g)
(()
Vll
Vll - V22 - V33
VI2
VI2
VI3
VI3
V22
0 0 0
V33 V23
or in matrix form,
aDY (g)
(()
=
1
0
0 0 0 0 0
1
0 0
-1
-1
0 0 0 0 0
0 0 0 0 0
1
0 0 0 0
0 0 0
0 0 0 0 0 0
Hence
aD (X
+ Y) (g) (() =
1
0
0 0 0 0 0
1
0 0
0 0 0 0
E22 -1 _EI2
-1
0
2E 23 _E I 3
_EI3
_EI2
E33
1
0
0 0 0
Ell
0
0 0
Ell
0 0
0
Ell
which has all positive eigenvalues. Hence the equation (B.35)
a at g = 2c+ CWg
is strictly parabolic. Hence for any metric go with negative sectional curvature, a solution 9 (t) to (B.35) with 9 (0) = 90 exists for a short time. Similarly to the case of the Ricci flow, we solve the following ODE at each point in M (see (3.35) on p. 81 of Volume One):
(B.36)
a
8t'Pt =
-w* ,
'Po = id, where W* (t) is the vector field dual to W (t) with respect to 9 (t) . Pulling back 9 (t) by the diffeomorphisms 'Pt, we obtain a solution (B.37) to the cross curvature flow with 9 (0)
= go.
o
-196
B
OTHER ASPECTS OF RICCI FLOW AND RELATED FLOWS
REMARK B.22. The above proof follows [34], where it is pointed out that the proof of short-time existence in [106] is incomplete. 3.3. Monotonicity formulas for the XeF. Given short-time existence of th(' flow, one would like to prove long-time existence and convergence of the flow on closed 3-manifolds. Although this problem is still open, we have the following result due to Hamilton, from [106], which is Proposition 3.24 on p. 88 of Volume One. PROPOSITION B.23 (Monotonicity formulae). If (M 3 , 9 (t)) zs a solutzon to the cross curvature flow with negative sectional curvature on a closed 3mamfold. then
(l) (volume of Einstein tensor increases)
ata Vol (E)
(B.38)
~ 0,
(2) (mtegral difference from hyper'bolic decreases)
!i f (~( g ) _ dt .fMJ 3 g~)
(B.39)
i)
E) 1/3) d/-l <- O.
(det det 9
REMARK B.24. Applying the arithmetic-geometric mean inequality to the eigenvalues of Eij with respect to gij, we see that the integrand in part (2) is nonnegative, and it is identically zero if and only if Eij == Eg ij , that is, if and only if gij has constant sectional curvature.
!
In the remainder of this section, we prove the above two monotonicity formulas. We begin by computing the evolution of the Einstein tensor. LEMMA B.25.
gt E
i)
=
V' k V' f (Eke Eij - Eik EJi') - det E
g~)
- C Eij,
where C ~ gij Cij .
PROOF. Taking q = j in (B.3l), we may rewrite the Einstein tensor as E mn
(B.40)
From (B.40) and (B.34) with
l/lijm IIkCnR -_ -41"" ijkf· I""
Vi)
= 2Cij.
we compute
1 ijm kln (n n n ata E mn = - 4/-l /-l v i VeCjk + v j V' kCif. -
1 ijm /-l kfn 9 pq - 4/-l -
-
lIijmllklnV'.V' C I"" '/, k J£
I""
(R
ijkpCqf
V'i V' kCjl
-
V'j V' fCik )
+ R iJP£Cqk) 2CEmn
~lIijrn/lklngpqR" C 2 I"" I"" ~JP£ qk
-
2cgmn ,
497
3. THE CROSS CURVATURE FLOW
J1 ijrn J1kfn cJ f.
= ~J1ijrn J1.7pq J1kfn J./rs EprEqs =~
(_giPgm q
+ lqgm p )
= E 1i.-E mll _ E
1Il
( _l"gnb
+ lsgnr.)
EprEqs
E m l..·.
Hence J1 ijm J1ken\1t \1k('jf
= \1 1 \1 k
( Etk E mn - EtIl Emk) .
With thb, the lemma follows from the identity 1 .. k' _l/tJT1II/·,n gpq R·· C +CEmn=detEgrnn 2"" ,.., . t.7pf qk .
To see this, we choose a basis where gi.7 = iS tj ' Eij and cij are diagonal, J1123 = 1, and RiJkf -I- 0 only if (i, j) = (k, £) as unordered pairs. We compute 1
2J1
iJ'l keJ pqR J1 9 ijpfCqk
=
J1
kf.1 pq
9
R 23pPCqk
= R2323 C22 - R2:J32 C33
= _Ell (C22
+ ('3:~),
so that !JliJIJ1kf.lgpqRijpfCqk + CEll = Ellcll = detE, similarly for the other diagonal components. We leave it as an exercise to check that the off-diagonal components are zero. 0 As we shall show below, part (1) of Proposition B.23 follows from the following more general computation. Let
LEMMA
B.26. For any
0' E
lR
(B.41)
~~ .1M (det Er~ dJ1 = 0' .1M (~ ITijk + (1 -
IAtJkl2E-l =
20')
Tjik
1:-1 -
0' ITt
1~;-1 ) (det E t dJ1
.1M (det E t C dJ1.
(E- I ). (E- 1 ). (E- 1 ) Aijk Apqr. tp Jq kr
498
B. OTHER ASPECTS OF RICCI FLOW AND RELATED FLOWS PROOF.
:t 1M
We compute using the evolution equation for Eij that (det E)a dJ.l
= 1M (det E) a
( a ( (E- 1) ij
:t
Eij - gij
(!
gij ) )
+ C)
dJ.l
1M (det Et (E- 1)ij 'h\le (Eke Eij - Eik Eje) dJ.l + 1M (det E)a (a ((E- 1)ij (- det Egij - CEij) + 2C) + C) dJ.l = -a 1M \lk [(detE)O (E- 1)ij] (Eke\leEij - Eje\leEik) dJ.l + (1 - 2a) 1M (det Et C dJ.l. =a
Since \lk [(detE)a (E- 1)ij] = (detE)O
(a (E- 1)pq (\lkEpq) (E- 1)ij -
(E- 1)iP (E- 1)jq \lkEpq) ,
we have
1M \lk [(det E)a (E- 1)ij] (Eke\leEij - Eje\le Eik ) dJ.l = a 1M (detE)a (E-1)pq \lkEpq (Eke (E- 1)ij \leEij - \liEik) dJ.l - 1M (detE)a (E- 1)iP (E- 1)jq \lk Epq (Eke\leEij - EifO\leEik) dJ.l, and the lemma follows from \liEik
= O.
To see the monotonicity in the a
D
= 1/2 case, we decompose the 3-tensor
Tijk into its irreducible components. In particular, the orthogonal group 0(3) with respect to the metric E-1 acts on the bundle of 3-tensors which
are symmetric in the last two components. The irreducible decomposition is given by
where the coefficients
-lo and ~ are chosen so that U is trace-free:
3. THE CROSS CURVATURE FLOW
499
Since this decomposition is orthogonal with respect to E- 1 , we have
ITiJk - Tjikl2 = IUijk _ Ujikl2 E-l E-l
(B.42)
+ 1_~EikTj + ~EjkTil2 2
2
= IU ijk - Ujikl:_l + ITil~_l
(B.43)
Taking
0:
~ dt
E-l
.
= 1/2 in the lemma, we conclude that
1
(detE)1/2 dJ.£
M
= ~ llUijk - Ujikl2 4
E-l
M
(detE)1/2 dJ.£;::: 0
and part (1) of Proposition B.23 follows. We now give the proof of part (2) of Proposition B.23. Define
J
~ 1M (gi j:ij -
(det E)1/3) dJ.£,
where gijEij = -~R. We compute
d dt
1. M
(lJ Eij) dJ.£ =
1 [( a M
a.. at gij ).. EtJ + gij at EtJ
. )C] dJ.£ + (lJ.Eij
1M (2Cij Eij - gij (det E gij + C EiJ) + (gi j Eij) c) dJ.£ = 3 1M det E dJ.£.
=
Combining this equation with (B.41) for
dJ dt
= _~
1 (~ITijk
3 M
2
0:
= 1/3 and (B.43), we conclude
- Tjikl2 - ~ ITil2 _ ) (det E)1/3 dJ.£ E-l 3 E 1
+ 1M det E dJ.£ - ~ 1M (det E) 1/3 C dJ.£ =
-~6
1 M
(IU ijk - Ujikl2 E-l
+ ~3 Ir l2E_1 )
(det E)1/3 dJ.£
- 1M (~ - (detc)1/3) (detE)1/3 dJ.£ :::; 0,
where we used detE
= (detc)1/3 (detE)1/3.
3.4. Cross curvature solitons. A solution to the XCF is a cross curvature soliton if there exists a vector field V and >. E ~ such that at some time
(B.44)
More generally, a solution to the XCF is a cross curvature breather if there exist times t1 < t2, a diffeomorphism 'P : M -+ M, and a > 0 such
500
B OTHER ASPECTS OF RICCI FLOW AND RELATED FLOWS
that (B.45) We have the following nonexistence result. B.27 (XCF breathers are trivial). If (M3, 9 (t)) zs a cross curvature breather with negative sectwnal curvature, then 9 (t) has constant sectional curvature. LEMMA
PROOF.
We have
d r1M ·gZJciJdJ-l > 0,
dt Vol (g (t)) =
so that the breather equation (B.45), i.e., 9 (t2) = w.p*g (tJ) for tl < t2, implies a > 1. On the other hand, J (g (t2)) = a 1/ 2J (g (tJ)) 2': 0, which contradicts the monotonicity formula (B.39) unless J (g (t2)) = J (g (td) = 0, in which case 9 (t) has constant sectional curvature. 0 4. Notes and commentary
Lemma B.7 is Lemma 3.3, Theorem 3.4 and Corollary 3.5(i) in [212]. See l\la and Chen [259] for a study of the cross curvature flow for certain classes of metrics on sphere and torus bundles over the circle. For work on the stability of the cross curvature flow at a hyperbolic metric, see Young and one of the authors [235].
APPENDIX C
Glossary adjoint heat equation. When associated to the Ricci flow, the equation is
au
at + Dou -
Ru
= O.
For a fixed metric, the adjoint heat equation is just the backward heat equation ~~ + Dou = O. ancient solution. A solution of the Ricci flow which exists on a time interval of the form (-00, w) , where w E (-00,00]. Limits of dilations about finite time singular solutions on closed manifolds are ancient solutions which are K-noncollapsed at all scales. For this reason a substantial part of the subject of Ricci flow is devoted to the study of ancient solutions. asymptotic scalar curvature ratio (ASCR). For a complete noncompact Riemannian manifold, ASCR(g) = limsup R(x)d(x,O)2,
°
d(x,O)--->oo
where E Mn is any choice of origin. This definition is independent of the choice of EM. For a complete ancient solution of the Ricci flow 9 (t) on a noncompact manifold with bounded nonnegative curvature operator, ASCR (g (t)) is independent of time. The ASCR is used to study the geometry at infinity of solutions of the Ricci flow and in particular to perform dimension reduction when ASCR = 00. asymptotic volume ratio (AVR). On a complete noncompact Riemannian with nonnegative Ricci curvature, AVR is the limit of the monotone quantity Vol ~n(p,r) as r ~ 00. This definition is independent of the choice of p EM. Ancient solutions with bounded nonnegative curvature operator have AVR = O. Expanding gradient Ricci solitons with positive Ricci curvatures have AVR > O. backward Ricci flow. The equation is
°
a aT g = 2Rc. Usually obtained by taking a solution 9 (t) to the Ricci flow and defining T (t) ~ to - t for some to. Bernstein-Bando-8hi estimates (also BB8 estimates). Shorttime estimates for the derivatives of the curvatures of solutions of the Ricci flow assuming global pointwise bounds on the curvatures. Roughly, given 501
502
C. GLOSSARY
a solution of the Ricci flow, if IRm (t)lg(t) :::; K on a time interval [0, T) of length on the order of K- 1 , then IV m Rm (t)lg(t) :::;
~;:/~
on that interval. The BBS estimates are used to obtain higher derivative of curvature estimates from pointwise bounds on the curvatures. In particular, they are used in the proof of the Coo compactness theorem for sequences of solutions assuming only uniform pointwise bounds on the curvatures and injectivity radius estimates. Bianchi identity. The first and second Bianchi identities are
+ Rjkil + Rkijl = 0, ViRjklm + VjRkilm + VkRijlm = 0, Rijkl
respectively. The Bianchi identities reflect the diffeomorphism invariance of the curvature. In the Ricci flow they are used to derive various evolution equations including the heat-like equation for the Riemann curvature tensor. Bishop---Gromov volume comparison theorem. An upper bound for the volume of balls given a lower bound for the Ricci curvature. This bound is sharp in the sense that equality holds for complete, simply-connected manifolds with constant sectional curvature. Bochner formula. A class of formulas where one computes the Laplacian of some quantity (such as a gradient quantity). Such formulas are often used to prove the nonexistence of nontrivial solutions to certain equations. For example, harmonic i-forms on closed manifolds with negative Ricci curvatures are trivial. In the Ricci flow, Bochner-type formulas (where the Laplacian is replaced by the heat operator) take the form of evolution equations which yield estimates after the application of the maximum principle. bounded geometry. A sequence or family of Riemannian manifolds has bounded geometry if the curvatures and their derivatives are uniformly bounded (depending on the number of derivatives). breather solution. A solution of the Ricci flow which, in the space of metrics modulo diffeomorphisms, is a periodic orbit. Bryant soliton. The complete, rotationally symmetric steady gradient Ricci soliton on Euclidean 3-space. The Bryant soliton has sectional curvatures decaying inverse linearly in the distance to the origin and hence has ASCR = 00 and AVR = 0. It is also expected to be the limit of the conjectured degenerate neckpinch. Buscher duality. A duality transformation of gradient Ricci solitons on warped products with circle or torus fibers. Calabi's trick. A typical way to localize maximum principle arguments is to multiply the quantity being estimated by a cut-off function depending on the distance to a point. Calabi's trick is a way to deal with the issue of the cut-off function being only Lipschitz continuous (since the distance function is only Lipschitz continuous).
503
C. GLOSSARY
Cartan structure equations. Given an orthonormal frame {ei} and dual coframe {w j }, they are the identities satisfied by the connection 1forms
{w;} and the curvature 2-forms {Rm 1}:
The Cartan structure equations are useful for computing curvatures, especially for metrics with some sort of symmetry. They may also be used for general calculations in geometric analysis. Cheeger-Gromov convergence. (See compactness theorem.) Christoffel symbols. The components of the Levi-Civita connection with respect to a local coordinate system: .
k
k
'iJ 8 / 8x i8/8xJ = rij 8/8x . The variation formula for the Christoffel symbols is the first step in computing the variation formula for the curvatures. The evolution equation for the Christoffel symbols is also used to derive evolution equations for quantities involving covariant derivatives. cigar soliton. The rotationally symmetric steady gradient Ricci soliton on the plane defined by
The scalar curvature of the cigar is Rr.
=
4 2
1+x +y
2
2
= 4 sech s,
where s is the distance to the origin. Note that gr. is asymptotic to a cylinder and Rr. decays exponentially fast. Perelman's no local collapsing theorem implies the cigar soliton and its product cannot occur as a limit of a finite time singularity on a closed manifold. classical entropy. An integral of the form M flog fdJ-l. collapsible manifold. A manifold admitting a sequence of metrics with uniformly bounded curvature and maximum injectivity radius tending to zero. compactness theorem (Cheeger-Gromov-type). If a pointed sequence of complete metrics or solutions of Ricci flow has uniformly bounded curvature and injectivity radius· at the origins uniformly bounded from below, then there exists a subsequence which converges to a complete metric or solution. The convergence is after the pull-back by diffeomorphisms, which we call Cheeger-Gromov convergence. conjugate heat equation. (See adjoint heat equation.)
J
c.
504
GLOSSARY
cosmological constant. A constant c introduced into the Ricci flow equation:
o at gij = -2 (Rij + cgi)) .
i
The case c = is useful in converting expanding Ricci solitons to steady Ricci solitons. cross curvature flow. A fully nonlinear flow of metrics on 3-manifolds with either negative sectional curvature everywhere or positive sectional curvature everywhere. curvature gap estimate. For long existing solutions, a time-dependent lower bound for the spatial supremums of the curvatures. See Lemmas 8.7. 8.9, and 8.11 in [111]. curvature operator. The self-adjoint fiberwise-lillear map Rm: A2MTI ~ A2M
n
defined by Rm (a)i) ~ RijkfCtfk. curve shortening flow (CSF). The evolution equation for a plane curve given by
-ox = at
-/'\,1)
'
where K, hi the curvature and 1/ is the unit outward normal. It is useful to compare the CSF with the Ricci flow (especially on surfaces). degenerate neckpinch. A conjectured Type IIa singularity on the nsphere where a neck pinches at the same time its cap shrinks leading to a cusp-like singularity. Such a singularity has been proven by Angenent and Velazquez to occur for the mean curvature flow. DeTurck's trick. (See also Ricci-DeTurck flow.) A method to prove short-time existence of the Ricci flow using the Ricci-DeThrck flow. differential Harnack estimate. Any of a class of gradient-like estimates for solutions of parabolic and heat-type equations. dilaton. (See Perelman's energy.) dimension reduction. For certain classes of complete, noncompact solutions of the Ricci flow, a method of picking points tending to spatial infinity and blowing down the corresponding pointed sequence of solutions to obtain a limit solution which splits off a line. Einstein-Hilbert functional. The functional of Riemannian metrics: E (g) = RdJ.L, where R is the scalar curvature. Einstein metric. A metric with constant Ricci curvature. Einstein summation convention. The convention in tensor calculus where repeated indices are summed. Strictly speaking the summed indices should be one lower and one upper, but in practice we do not always bother to lower and raise indices. energy. (See Perelman's energy.) entropy. (See classical entropy, Hamilton's entropy for surfaces, and Perelman's entropy.)
IMn
c.
GLOSSARY
505
eternal solution. A solution of the Ricci flow existing on the time interval (-00,00). Note that eternal solutions are ancient solutions which are immortal. expanding gradient Ricci soliton (a.k.a. expander). A gradient Ricci soliton which is evolving by the pull-back by diffeomorphisms and scalings greater than 1. Gaussian soliton. Euclidean space as either a shrinking or expanding soliton. geometrization conjecture. (See Thurston's geometrization conjecture.) gradient flow. The evolution of a geometric object in the direction of steepest ascent of a functional. gradient Ricci soliton. A Ricci soliton which is flowing along diffeomorphisms generated by a gradient vector field. Gromoll-Meyer theorem. Any complete noncompact Riemannian manifold with positive sectional curvature is diffeomorphic to Euclidean space. Hamilton's entropy for surfaces. The functional E (g) = fM2log R· RdJ.t defined for metrics on surfaces with positive curvature. Hamilton-Iveyestimate. A pointwise estimate for the curvatures of solutions of the Ricci flow on closed 3-manifolds (with normalized initial data) which implies that, at a point where there is a sufficiently large (in magnitude) negative sectional curvature, the largest sectional curvature at that point is both positive and much larger in magnitude. In dimension 3 the Hamilton-Ivey estimate implies that the singularity models of finite time singular solutions have nonnegative sectional curvature. harmonic map. A map between Riemannian manifolds f: (Mn,g)---+ (Nm, h) satisfying tlg,hf = 0, where tlg,h is the map Laplacian. (See map Laplacian. ) harmonic map heat flow. The equation is %f = tlg,hf. Harnack estimate (See differential Harnack estimate.) heat equation. For functions on a Riemannian manifold: ~~ = tlu. This equation is the basic analytic model for geometric evolution equations including the Ricci flow. heat operator. The operator ~ - tl appearing in the heat equation. Hodge Laplacian. Acting on differential forms: tld = - (d& + &d) . homogeneous space. A Riemannian manifold (M n , g) such that for every x, y E M there is an isometry t- : M ---+ M with d x) = y. Huisken's monotonicity formula. An integral monotonicity formula for hypersurfaces in Euclidean space evolving by the mean curvature flow using the fundamental solution to the adjoint heat equation. immortal solution. A solution of the Ricci flow which exists on a time interval of the form (a, 00), where a E [-00,00). isoperimetric estimate. A monotonicity formula for the isoperimetric ratios of solutions of the Ricci flow. Examples are Hamilton's estimates
c.
506
GLOSSARY
for solutions on closed surfaces and Type I singular solutions on closed 3manifolds. Jacobi field. A variation vector field of a I-parameter family of geodesics. Kahler-Ricci flow. The Ricci flow of Kahler metrics. Note that on a closed manifold, an initial metric which is Kahler remains Kahler under the Ricci flow. ~-noncollapsed at all scales. A metric (or solution) which is ~ noncollapsed below scale p for all p < 00. ~-noncollapsed below the scale p. A Riemannian manifold satisfying Vol~~x,r) ~ ~ for any metric ball B(x, r) with IRm I ~ r- 2 in B (x, r) and r < p. ~-solution (or ancient ~-solution). A complete ancient solution which is ~-noncollapsed on all scales, has bounded nonnegative curvature operator, and is not flat. In dimension 3, a large part of singularity analysis in Ricci flow is to classify ancient ~-solutions. L-distance. A space-time distance-like function for solutions of the backward Ricci flow obtained by taking the infimum of the £-length. The L-distance between two points may not always be nonnegative. i-distance function. (See reduced distance.) £-exponential map. The Ricci flow analogue of the Riemannian exponential map. £-geodesic. A time-parametrized path in a solution of the backward Ricci flow which is a critical point of the £-length functional. £-Jacobi field. A variation vector field of a I-parameter family of £-geodesics. £-length. A length-like functional for time-parametrized paths in solutions of the backward Ricci flow. The £-length of a path may not be positive. Laplacian (or rough Laplacian). On Euclidean space the operator ~ = E~l 8(~~)2. On a Riemannian manifold, the second-order linear differential operator ~ = gij·~h'Vj acting on tensors. Levi-Civita connection. The unique linear torsion-free connection on the tangent bundle compatible with the metric. (Also called the Riemannian connection. ) Lichnerowicz Laplacian. The second-order differential operator ~L acting on symmetric 2-tensors defined by (VI-3.6), i.e., ~LVij ~ ~Vjk
+ 2gqp R~jk vrp -
gqp R Jp Vqk - gqp R kp V Jq .
Lichnerowicz Laplacian heat equation. The heat-like equation = (~LV)ij for symmetric 2-tensors. linear trace Harnack estimate. A differential Harnack estimate for nonnegative solutions of the Lichnerowicz Laplacian heat equation coupled to a solution of the Ricci flow with nonnegative curvature operator, which generalizes the trace Harnack estimate. (See trace Harnack estimate.) &tVij
C. GLOSSARY
507
little loop conjecture. Hamilton's conjecture which is essentially equivalent to Perelman's no local collapsing theorem (e.g., the conjecture is now a theorem). Li-Yau-Hamilton (LYH) inequality. (See differential Harnack estimate.) locally homogeneous space. (See also homogeneous space.) A Riemannian manifold (Mn, g) such that for every x, y E M there exist open neighborhoods U of x and V of y and an isometry 1-: U - V with t-{x) = y. A complete simply-connected locally homogeneous space is a homogeneous space. locally Lipschitz function. A function which locally has a finite Lipschitz constant. logarithmic Sobolev inequality. A Sobolev-type inequality which essentially bounds the classical entropy of a function by the L2- norm of thf:l first derivative of the function. long existing solutions. Solutions which exist on a time interval of infinite duration. long-time existence. The existence of a solution of the Ricci flow on a closed manifold as long as the Riemann curvature tensor remains bounded. By a result of Sesum, in the above statement the Riemann curvature tensor may be replaced by the Ricci tensor. map Laplacian. Given a map f : (Mn, g) - (Nm, h) between Riemannian manifolds, Ag,hf is the trace with respect to 9 of the second covariant derivative of f. (See (3.39) in Volume One.) matrix Harnack estimate. In Ricci flow a certain tensor inequality of Hamilton for solutions with bounded nonnegative curvature operator. A consequence is the trace Harnack estimate. One application of the matrix Harnack estimate is in the proof that an ancient solution with nonnegative curvature operator and which attains the space-time maximum of the scalar curvature is a steady gradient Ricci soliton. maximum principle. (Also called the weak maximum principle.) The first and second derivative tests applied to heat-type equations to obtain bounds for their solutions. The basic idea is to use the inequalities Au ~ 0 and \7u = 0 at a spatial minimum of a function u and the inequality !JJt ~ 0 at a minimum in space-time up to that time. This principle applies to scalars, tensors, and systems. maximum volume growth. A noncom pact manifold has maximum volume growth if for some point p there exists c > 0 such that Vol B (p, r) ~ ern for all r > O. This is the same as AVR > 0 (when Rc ~ 0). mean curvature flow. The evolution of a submanifold in a Riemannian manifold in the direction of its mean curvature vector. modified Ricci flow. An equation of the form = -2 (Rc +VV f), where f is a function on space-time. Often coupled to an equation for f such as Perelman's equation Wt = -R - Af.
Itg
508
C. GLOSSARY
monotonicity formula. Any formula which implies the monotonicity of a pointwise or integral quantity under a geometric evolution equation. Examples are entropy and Harnack estimates. J.t-invariant. An invariant of a metric and a positive number (scale) obtained from Perelman's entropy functional W (g, I, T) by taking the infimum over all I satisfying the constraint fMn (47rT)-n/2 e- f dJ.t = 1. There is a monotonicity formula for this invariant under the Ricci flow. neckpinch. A finite time (Type I) singular solution of the Ricci flow where a region of the manifold asymptotically approaches a shrinking round cylinder sn-l X lR. Sufficient conditions for initial metrics on sn for a neckpinch have been obtained by Angenent and one of the authors. no local collapsing. (Also abbreviated NLC.) A fundamental theorem of Perelman which applies to all finite time solutions of the Ricci flow on closed manifolds. It says that given such a solution of the Ricci flow and a finite scale p > 0, there exists a constant K. > 0 such that for any ball of radius r < p with curvature bounded by r- 2 in the ball, the volume ratio of the ball is at least K.. We say that the solution is K.-noncollapsed below the scale p. v-invariant. A metric invariant obtained from the J.t-invariant by taking the infimum over all T > O. This invariant may be -00. null-eigenvector assumption. A condition, in the statement of the maximum principle for 2-tensors, on the form of a heat-type equation which ensures that the nonnegativity of the 2-tensor is preserved under this heattype equation. parabolic equation. In the context of Ricci flow, a heat-type equation (which is second-order). In general, parabolicity of a nonlinear partial differential equation is defined using the symbol of its linearization. Perelman's energy. The functional
This invariant appeared previously in mathematical physics (e.g., string theory) and I is known as the dilaton. Perelman's entropy. The following functional of the triple of a metric, a function, and a positive constant:
W(g,I,T) = fMn (T (R+
IVII2) + (f -
n)) (47rT)-n/2 e- f dJ.t.
Perelman's Harnack (LYH) estimate. A differential Harnack (e.g., gradient) estimate for fundamental solutions of the adjoint heat equation coupled to the Ricci flow. Perelman solution. The non-explicit 3-dimensional analogue of the Rosenau solution. The Perelman solution is rotationally symmetric and has positive sectional curvature. Its backward limit as t ~ -00 is the Bryant soliton.
C. GLOSSARY
509
Poincare conjecture. The conjecture that any simply-connected closed 3-manifold is diffeomorphic to the 3-sphere. (In dimension 3, the topological and differentiable categories are the same.) Hamilton's program and Perelman's work aim to complete a proof of the Poincare conjecture using Ricci flow. positive (nonnegative) curvature operator. The eigenvalues of the curvature operator are positive (nonnegative). potentially infinite dimensions. A device which Perelman combined with the space-time approach to the Ricci flow to embed solutions of the Ricci flow into a potentially Ricci flat manifold with potentially infinite dimension. preconvergent sequence. A sequence for which a subsequence converges. quasi-Einstein metric. The mathematical physics jargon for a nonEinstein gradient Ricci soliton. Rademacher's Theorem. The result that a locally Lipschitz function is differentiable almost everywhere. reaction-diffusion equation. A heat-type equation consisting of the heat equation plus a nonlinear term which is zeroth order in the solution. reduced distance. The distance-like function for solutions of the back(q, 7). (See L-distance.) Partly ward Ricci flow defined by f (q, 7) = motivated by consideration of the heat kernel and the Li-Yau distance function for positive solutions of the heat equation. reduced volume. For a solution to the backward Ricci flow, the timedependent invariant
2FrL
V (7) = {
iMn
(47rT)-n/2 e-£(Q,T)djLg(T) (q).
Ricci flow. The equation for metrics is %t 9ij = - 2~j. This equation was discovered and developed by Richard Hamilton and is the subject of this book. Ricci soliton. (See also gradient Ricci soliton.) A self-similar solution of the Ricci flow. That is, the solution evolves by scaling plus the Lie derivative of the metric with respect to some vector field. Ricci tensor. The trace of the Riemann curvature operator: n
Rc (X, Y)
= trace (X ~ Rm (X, Y) Z) =
L (Rm (ei' X) Y, ei) . i=l
Ricci-DeTurck flow. A modification of the Ricci flow which is a strictly parabolic system. This equation is essentially equivalent to the Ricci flow via the pull-back by diffeomorphisms and is used to prove short-time existence for solutions on closed manifolds. Riemann curvature operator. (See curvature operator.) Riemann curvature tensor. The curvature (3, i)-tensor obtained by anti-commuting in a tensorial way the covariant derivatives acting on vector
C. GLOSSARY
510
fields. Formally it is defined by Rm (X, Y) Z
'*' VxV'yZ - V'yVxZ - '\7[X,YjZ.
Rosenau solution. An explicit rotationally symmetric ancient solution on the 2-sphere with positive curvature. Its backward limit as t ~ -00 (without rescaling) is the cigar soliton. scalar curvature. The tra.ce of the Ricci tensor: R = L:~=l Rc (ei' ei) = gij~j.
sectional curvature. The number K (P) = (Rm (el' e2) e2, el) associated to a 2-plane in a tangent spaee P c TxM, where {el' e2} is an orthonormal basis of P. self-similar solution. (For the Ricci flow, see Ricci soliton.) Shi's local derivative estimate. A local estimate for the covariant derivatives of the Riemann curvature tensor. short-time existence. The existence, when it holds, of a solution to the initial-value problem for the rued flow on some nontrivial time interval. For example, for a smooth initial metric on a closed manifold. shrinking gradient Ricci soliton (a.k.a. shrinker). A gradient Ricci soliton which is evolving by the pull-back by diffeomorphisms and scalings legs than 1. singular solution. A solution on a maximal time interval. If the maximal titne interval [0, T) is flnite, then sup
IRml <
00.
Mnx[O,T)
singUlarity. For exatnple, if T is the singular (Le., maximal) time, we say that the solution forms a singularity at time T. singularity model. The limit of dilations of a singular solution. For finite time singular solutions, singularity models are ancient solutions. For infinite time singUlar solutions, singularity models are immortal solutions. singUlarity time. For a solution, the time T E (0,00], where [0, T) is the maximal time interval of existence. Sobolev inequality. A class of inequalities where the Lq-norms of functions are bounded by the V-norms of their derivatives, where p and q are related by the dimension and number of derivatives. space form. A complete Riemannian manifold with constant sectional curvature. space-time. The manifold Mn x T of a solution (Mn,g (t)) defined on the time interval T. space-time connection. Any of a class of natural connections on the tangent bundle of space-time spherical space form. A complete Riemannian manifold with positive constant sectional curvature. Such a manifold is, after scaling, isometric to a quotient of the unit sphere by a group of linear isometries.
C. GLOSSARY
511
steady gradient Ricci soliton. A gradient Ricci soliton which is evolving by the pull-back by diffeomorphisms only. In particular, the metrics at different times are isometric. strong maximum principle. For the scalar heat equation it says that a nonnegative solution which is zero at some point (xo, to) vanishes for all points (x, t) with t :::; to. For the curvature operator under the Ricci flow Hamilton's strong maximum principle says that given a solution with nonnegative curvature operator, for positive short time t E (0,8) the image of the curvature operator image (Rm (t)) c A2T* Mn is invariant under parallel translation and constant in time. At each point (x, t) E Mn x (0, 8) the image of the curvature operator is a Lie sub algebra of A2T* M~ ~ 50 (n). tensor. A section of some tensor product of the tangent and cotangent bundles. Thurston's geometrization conjecture. The conjecture of William Thurston that any closed 3-manifold admits a decomposition into geometric pieces. This subsumes the Poincare conjecture. total scalar curvature. (See Einstein-Hilbert functional.) trace Harnack estimate. Given a solution of the Ricci flow with nonnegative curvature operator, the estimate
8R 8t
R
+ t + 2 (V'R, V) + 2 Rc (V, V)
2:
°
holding for any vector V. Taking V = 0, we have ~ (tR (x, t)) 2: 0. Type I, IIa, lIb, III singularity. The classification of singular solutions according to the growth or decay rates of the curvatures. For T < 00, Type I is when sup (T - t) IRmI < 00 Mx[O,T)
and Type IIa is when (T - t) IRml
sup
= 00.
Mx[O,T)
For T
= 00, Type III is when sup
t IRml <
00
Mx[O,T)
and Type lIb is when sup
t IRml
= 00.
Mx[O,T)
Examples of Type I singularities are shrinking spherical space forms and neckpinches. Uhlenbeck's trick. In the Ricci flow it is the method of pulling back tensors to a bundle isomorphic to the tangent bundle with a fixed metric. This method simplifies the formulas for various evolution equations involving the Riemann curvature tensor, its derivatives and contractions. variation formula. Given a variation of a geometric quantity such as a metric or a submanifold, the corresponding variation for an associated
512
C. GLOSSARY
geometric quantity. For example, given a variation of a metric, we have variation formulas for the Christoffel symbols and the Riemann, Ricci, and scalar curvature tensors. volume form. On an oriented n-dimensional Riemannian manifold, the n-form df-l = Jdet (gij) dx 1 /\ ... /\ dxn in a positively oriented local coordinate system. The integral of the volume form is the volume of the manifold. volume ratio. For a ball B (p, r) , the quantity VolB (p, r) rn W-functional. (See Perelman's entropy.) warped product metric. Given Riemannian manifolds (Mn, g) and (Nm, h), the metric on the product M x N of the form 9 (x) + f (x) h (y) , where f : M ~ lR. The natural space-time metrics are similar to warped products. Witten's black hole. (See cigar soliton.) Yamabe flow. The geometric evolution equation of metrics: %tgij = -Rgij. When n = 2, this is the same as the Ricci flow. A technique applied to the Yamabe flow described in this book is the Aleksandrov reflection method. It is expected that the Yamabe flow evolves metrics on closed manifolds to constant scalar curvature metrics. Yamabe problem. The problem of showing that in any conformal class there exists a metric with constant scalar curvature. This problem has been solved through the works of Yamabe, Trudinger, Aubin, and Schoen. Schoen's work uses the positive mass theorem of Schoen and Yau.
Bibliography [1] Alexander, J.W. On the subdivision of 3-space by a polyhedron. Proc. Nat. Acad. Seci, USA, 10, 6-8, 1924. [2] Alvarez, E.; Kubyshin, Y. Is the string coupling constant invariant under T-duality? hep-th/9610032. [3] Anderson, Greg; Chow, Bennett, A pinching estimate for solutions of the linearized Ricci flow system on 3-manifolds. Calculus of Variations 23 (2005), no. 1, 1-12. [4] Anderson, Michael T. Degeneration of metncs with bounded curvature and applicatwns to cntical metrics of Rtemannian functionals. In Differential geometry: Riemannian geometry (Los Angeles, CA, 1990), 53-79, Proc. Sympos. Pure Math., 54, Part 3, Amer. Math. Soc., Providence, RI, 1993. [5] Anderson, Michael T. Geometrization of 3-manifolds via the Ricci flow. Notices Amer. Math. Soc. 51 (2004), no. 2, 184-193. [6] Anderson, Michael T.; Cheeger, Jeff. C"'-compactness for manifolds with Riccz curvature and mJectivity radius bounded below. J. Differential Geom. 35 (1992), no. 2, 265-281. [7] Angenent, Sigurd B.; Knopf, Dan. An example of neckpinching for Ricci flow on sn+l. Math. Res. Lett. 11 (2004), no. 4, 493-518. [8] Angenent, Sigurd B.; Knopf, Dan. Precise asymptottCS for the Ricci flow neckpmch. arXiv:math.DG/0511247. [9] Angenent, Sigurd B.; Velazquez, J. J. L. Degenerate neckpinches in mean curvature flow. J. Reine Angew. Math. 482 (1997), 15-66. [10] Apostol, Tom M. Calculus. Vol. I: One-vanable calculus, wtth an introduction to lmear algebra. Second edition, Blaisdell Publishing Co., Ginn and Co., Waltham, Mass.-Toronto, Ont.-London, 1967. [11] Aubin, Thierry. Equations du type Monge-Ampere sur les varietes kiihlenennes compactes. C. R. Acad. Sci. Paris Ser. A-B 283 (1976), no. 3, Aiii, A119-A121. [12] Aubin, Thierry. Equations du type Monge-Ampere sur les varietes kiihlenennes compactes. Bull. Sci. Math. (2) 102 (1978), no. 1, 63-95. [13] Aubin, Thierry. Nonlinear analysis on manifolds. Monge-Ampere equations. Grundlehren der Mathematischen Wissenschaften, 252. Springer-Verlag, New York, 1982. [14] Aubin, Thierry. Reduction du cas positif de l'equation de Monge-Ampere sur les vanetes K iihlenennes compactes d la demonstration d 'une inegalite. J. Funct. Anal. 57 (1984), 143-153. [15] Baird, Paul; Danielo, Laurent. Three-dimensional Ricci solitons which project to surfaces. Preprint. [16] Bakry, D. Diffusion on compact Riemannian manifolds and logarithmic Sobolev inequalitzes, J. Funet. Anal. 42 (1981) 102-109. [17] Bakry, D.; Concordet, D.; Ledoux, M. Optimal heat kernel bounds under loganthmic Sobolev inequalittes. ESAIM Probab. Statist. 1 (1995/97), 391-407 (electronic).
513
514
BIBLIOGRAPHY
[18] Bakry, D.j Emery, Michel. Diffusions hypercontractives. (French) [Hypercontractive diffusions} Seminaire de probabilites, XIX, 1983/84, 177-206, Lecture Notes in Math., 1123, Springer, Berlin, 1985. [19] Bando, Shigetoshi. On the classification of three-dzmenszonal compact Kaehler manifolds of nonnegative bisectional curvature. J. Differential Geom. 19 (1984), no. 2, 283-297. [20] Bando, Shigetoshi. Real analyticity of solutions of Hamilton's equation, Math. Zeit. 195 (1987), 93-97. [21] Bando, Shigetoshij Mabuchi, Toshiki. Umqueness of Einstein Kahler metrics modulo connected group actions. Algebraic geometry, Sendai, 1985, 11-40, Adv. Stud. Pure Math., 10, North-Holland, Amsterdam, 1987. [22] Barenblatt, Grigory I. On self-similar motions of compressible fluid in a porous medium. Prikladnaya Matematika i Mekhanika (Applied Mathematics and Mechanics (PMM)), 16 (1952), No.6, 679-698 (in Russian). [23] Beckner, Williamj Pearson, Michael. On sharp Sobolev embedding and the logarithmic Sobolev inequality. Bull. London Math. Soc. 30 (1998), no. 1, 80-84. [24] Berard-Bergery, L. Sur de nouvelles varieties riemannienes d'Einstein. Publ. Inst. E. Cartan # 4, U. de Nancy (1982). [25] Berndt, Jiirgenj Tricerri, Francoj Vanhecke, Lieven. Generalized Hezsenberg groups and Damek-Ricci harmonic spaces. Lecture Notes in Mathematics, 1598. SpringerVerlag, Berlin, 1995. [26] Bernstein, Sergi N. A limitation on the moduli of a sequence of derivatives of solutions of equations of parabolic type. (Russian) Dokl. Akad. Nauk SSSR 18 (1938) 385-388. [27] Besse, Arthur, Einstein manifolds. Ergebnisse der Mathematik und ihrer Grenzgebiete, 10. Springer-Verlag, Berlin, 1987. [28] Bing, R. H. An alternative proof that 3-manifolds can be triangulated. Ann. of Math. (2) 69 (1959), 37-65. [29] Bohm, Christophj Kerr, Megan M. Low-dzmenszonal homogeneous Einstein manifolds. Trans. Amer. Math. Soc. 358 (2006), no. 4, 1455-1468. [30] Bohm, Christophj Wilking, Burkhard. Manifolds with positive curvature operators are space forms. arXiv:math.DG/0606187. [31] Bohm, Christophj Wilking, Burkhard. Nonnegatwely curved manifolds wzth finite fundamental groups admit metrics with positive Ricci curvature. Preprint. [32] Bourguignon, Jean-Pierre. Une stratification de l'espace des structures nemanniennes. (French) Compositio Math. 30 (1975), 1-41. [33] Bourguignon, Jean-Pierre. Ricci curvature and Emstem metrics. Global differential geometry and global analysis (Berlin, 1979), pp. 42-63, Lecture Notes in Math., 838, Springer, Berlin-New York, 1981. [34] Buckland, John A. Short-tzme existence of solutions to the cross curvature flow on 3-manifolds. Proc. Amer. Math. Soc. 134 (2006), no. 6, 1803-1807. [35] Bryant, Robert. Gradient Kahler Rwcz solitons. arXiv:math.DG/0407453. [36] Bryant, R. L.j Chern, S. S.j Gardner, R. B.j Goldschmidt, H. L.j Griffiths, P. A. Exterior differential systems. Mathematical Sciences Research Institute Publications, 18. Springer-Verlag, New York, 1991. [37] Burago, D.j Burago, Y.j Ivanov, S. A course in metric geometry, Grad Studies Math. 33, Amer. Math. Soc., Providence, RI, 2001. Correctzons of typos and small errors to the book "A Course in Metric Geometry": http://www.pdmi.ras.rujstaff/burago. html#English [38] Buscher, T. H. A symmetry of the string background field equations. Phys. Lett. B194 (1987), 59-62. [39] Buscher, T. H. Path integral denvation of quantum duality in nonlmear sigma models. Phys. Lett. B201 (1988), 466.
BIBLIOGRAPHY
515
[40] Buser, Peter; Karcher, Hermann. Gromov's almost flat manifolds. Asterisque, 81. Societe Mathematique de France, Paris, 1981. [41] Cabre, Xavier. Nondivergent elliptic equations on manifolds with nonnegative curvature. Comm. Pure Appl. Math. 50 (1997) no. 7, 623-665. [42] Calabi, Eugenio. An extens~on of E. HopI's maximum principle with an application to R~emanman geometry. Duke Math. J. 24 (1957),45-56. [43] Calabi, Eugenio. On Kahler mamfolds w~th vanishing canonical class. Algebraic geometry and topology. A symposium in honor of S. Lefschetz, pp. 78-89. Princeton University Press, Princeton, N. J., 1957. [44] Calabi, Eugenio. Metriques kahleriennes et fibres holomorphes. Ann. Sci. Ecole Norm. Sup. (4) 12 (1979), no. 2, 269-294. [45] Cao, Huai-Dong. Deformation of Kahler metNcs to Kahler-Einstein metNcs on compact Kahler manifolds. Invent. Math. 81 (1985), no. 2, 359-372. [46] Cao, Huai-Dong. On Harnack's mequaMies for the Kahler-Ricc~ flow. Invent. Math. 109 (1992), 247-263. [47] Cao, Huai-Dong. Existence of gmdient Ricci-Kahler solitons. In Elliptic and parabolic methods in geometry (Minneapolis, MN, 1994), 1-16; A K Peters, Wellesley, MA,1996. [48] Cao, Huai-Dong. Limits of solutions to the Kahler-Ricc~ flow. J. Differential Geom. 45 (1997), no. 2, 257-272. [49] Cao, Huai-Dong; Chen, Bing-Long; Zhu, Xi-Ping. Ricci flow on compact Kahler manifolds of positive bisectional curvature. C. R. Math. Acad. Sci. Paris 337 (2003), no. 12, 781-784. [50] Cao, Huai-Dong; Chow, Bennett. Compact Kahler manifolds with nonnegative curvature opemtor. Invent. Math. 83 (1986), no. 3, 553-556. [51] Cao, Huai-Dong; Chow, Bennett. Recent developments on the Ricci flow. Bull. Amer. Math. Soc. (N.S.) 36 (1999), no. 1, 59-74. [52] Cao, Huai-Dong; Chow, Bennett; Chu, Sun-Chin; Yau, Shing-Tung, editors. Collected papers on Ricci flow. Internat. Press, Somerville, MA, 2003. [53] Cao, Huai-Dong; Hamilton, Richard S.; Ilmanen, Tom. Gaussian densities and stability for some Ricci solitons. arXiv:math.DG/0404165. [54] Cao, Huai-Dong; Ni, Lei. Matrix Li-Yau-Hamilton estimates for the heat equation on Kahler manifolds. Math. Ann. 331 (2005), no. 4, 795-807. [55] Cao, Huai-Dong; Sesum, Natasa. The compactness result for Kahler Ricci solitons. arXiv:math.DG/0504526. [56] Cao, Huai-Dong; Zhu, Xi-Ping. A complete proof of the Pomcare and geometrization conjectures - application of the Hamilton-Perelman theory of the Ricci flow. Asian J. Math. 10 (2006), 165-498. [57] Carfora; M.; Marzuoli, A. Model geometNes in the space of Riemanman structures and Hamilton's flow. Classical Quantum Gravity 5 (1988), no. 5, 659-693. [58] Carlen, E. A.; Kusuoka, S.; Strook, D. W. Upper bounds for symmetric Markov tmnsition functions. Ann. Inst. H. Poincare Probab. Statist. 23 (1987) 245-287. [59] Carron, G. Inegalites isoperimetriques de Faber-Kmhn et consequences. Actes de la Table Ronde de Geometrie Differentielle (Luminy, 1992), Semin. Congr., 1, Soc. Math. France, Paris 1996, 205-232. [60] Cascini, Paolo; La Nave, Gabriele. Kahler-Ricc~ Flow and the minimal model progmm for proJective varietzes. arXiv:math.AG/0603064. [61] Casson, Andrew; Jungreis, Douglas. Convergence groups and Seifert fibered 3mamfolds. Invent. Math. 118 (1994), no. 3, 441-456. [62] Chang, Shu-Cheng; Lu, Pengo Evolution of Yamabe constants under R~cc~ flow. Preprint. [63] Chau, Albert; Tam, Luen-Fai. Gmdient Kahler-Rwci solitons and a umformization conjecture. arXiv:math.DG/0310198.
516
BIBLIOGRAPHY
[64] Chau, Albert; Tam, Luen-Fai. A note on the uniformization of gradient Kahler-Ricci solitons. Math. Res. Lett. 12 (2005), no. 1, 19-21. [65] Chau, Albert; Tam, Luen-Fai. On the complex structure of Kahler manifolds w~th nonnegative curvature. arXiv:math.DG/0504422. [66] Chau, Albert; Tam, Luen-Fai. Nonnegatively curved Kahler manifolds with average quadrahc curvature decay. arXiv:math.DG/0510252. [67] Chavel, Isaac. Eigenvalues in R~emannian geometry. Including a chapter by Burton Randol. With an appendix by Jozef Dodziuk. Pure and Applied Mathematics, 115. Academic Press, Inc., Orlando, FL, 1984. [68] Chavel, Isaac. Riemannian geometry - a modern introduction. Cambridge Tracts in Mathematics, 108. Cambridge University Press, Cambridge, 1993. [69] Cheeger, Jeff. Some examples of manifolds of nonnegative curvature. J. Differential Geometry 8 (1973), 623-628. [70] Cheeger, Jeff. Fimteness theorems for Rzemannian manifolds. Amer. J. Math. 92 (1970), 61-74. [71] Cheeger, Jeff; Colding, Tobias H. Lower bounds on Ricci curvature and the almost rigidity of warped products. Ann. of Math. (2) 144 (1996), no. 1, 189-237. [72] Cheeger, Jeff; Ebin, David G. Comparison theorems in Riemannian geometry. North-Holland Mathematical Library, Vol. 9. North-Holland Publishing Co., Amsterdam-Oxford; American Elsevier Publishing Co., Inc., New York, 1975. [73] Cheeger; Jeff; Gromov, Mikhail. Collapsing Riemannian manifolds whzle keepmg their curvature bounded, J, J. Differential Geom. 23 (1986) 309-346. [74] Cheeger; Jeff; Gromov, Mikhail. Collapsing Riemannwn manifolds while keeping their curvature bounded, II, J. Differential Geom. 32 (1990) 269-298. [75] Cheeger; Jeff; Gromov, Mikhail; Taylor, Michael. Finite propagation speed, kernel eshmates for functions of the Laplace operator, and the geometry of complete Riemanman mamfolds, J. Differential Geom. 17 (1982) 15-53. [76] Cheeger; Jeff; Yau, Shing-Thng. A lower bound for the heat kernel. Comm. Pure Appl. Math. 34 (1981), no. 4, 465-480. [77] Chen, Bing-Long; Zhu, Xi-Ping. Complete Riemannian manifolds with pointwise pinched curvature. Invent. Math. 140 (2000), no. 2, 423-452. [78] Chen, Bing-Long; Zhu, Xi-Ping. A property of Kahler-Ricci solitons on complete complex surfaces. Geometry and nonlinear partial differential equations (Hangzhou, 2001), 5-12, AMS/IP Stud. Adv. Math., 29, Amer. Math. Soc., Providence, RI, 2002. [79] Chen, Bing-Long; Zhu, Xi-Ping. On complete noncompact Kahler manifolds with positive bisectional curvature. Math. Ann. 327 (2003) 1-23. [80] Chen, Bing-Long; Zhu, Xi-Ping. Volume growth and curvature decay of positively curved Kahler manifolds. arXiv:math.DG/0211374. [81] Chen, Bing-Long; Zhu, Xi-Ping. Riccz flow with surgery on four-manifolds wzth positive isotropic curvature. arXiv:math.DG/0504478. [82] Chen, Bing-Long; Zhu, Xi-Ping. Uniqueness of the Ricci flow on complete noncompact manifolds. arXiv:math.DG/0505447. [83] Chen, Xiuxiong. The space of Kahler metrics. J. Differential Geom. 56 (2000), no. 2, 189-234. [84] Chen, Xiuxiong. On the lower bound of energy functional El (J) - a stability theorem on the Kahler Ricci flow. arXiv:math.DG/0502196. [85] Chen, Xiuxiong; Li, H. The Kahler-Ricci flow on Kahler manifolds with 2-traceless bisectional curvature operator. arXiv:math.DG/0503645. [86] Chen, Xiuxiong; Lu, Peng; Tian, Gang. A note on uniformization of Riemann surfaces by Ricci flow. Proc. Amer. Math. Soc. 134 (2006),3391-3393. [87] Chen, Xiuxiong; Tian, Gang. Rzcci flow on Kahler-Einstem surfaces. Invent. Math. 147 (2002), no. 3, 487-544.
BIBLIOGRAPHY
517
[88] Chen, Xiuxiongj Tian, Gang. Rtcci flow on Kiihler-Emstem mamfold. Duke Math. J. 131 (2006), 17-73. [89] Chen, Ya-Zhe. Second-order pambolic partial differential equations, Beijing University Press, 2003, Beijing. [90] Chen, Yun Meij Struwe, Michael. Existence and partial regularity results for the heat flow for harmonic maps. Math. Z. 201 (1989), no. 1, 83-103. [91] Cheng, Hsiao-Bing. A new Li-Yau-Hamilton estimate for the Ricci flow. Comm. Anal. Geom. 14 (2006), 551-564. [92] Cheng, Shiu-Yuen. Eigenvalue comparison theorems and tts geometnc applications. Math. Zeit. 143 (1975), no. 3, 289-297. [93] Cheng, Shiu-Yuenj Li, Peterj Yau, Shing-Thng. On the upper estimate of the heat kernel of a complete Riemannian manifold, Amer. J. Math. 103 (1981), no. 5, 10211063. [94] Cheng, Shiu-Yuenj Yau, Shing-Thng. On the existence of a complete Kahler metric on noncompact complex manifolds and the regularity of Fefferman's equation. Comm. Pure Appl. Math. 33 (1980), 507-544. [95] Chern, Shiing Shen. Complex manifolds without potential theory. With an appendix on the geometry of chamcteristic classes. Second edition. Universitext. SpringerVerlag, New York-Heidelberg, 1979. [96] Chow, Bennett. On Harnack's inequality and entropy for the Gaussian curvature flow. Comm. Pure Appl. Math. 44 (1991), no. 4, 469-483. [97] Chow, Bennett. On the entropy estimate for the Ricct flow on compact 2-orbifolds. J. Differential Geom. 33 (1991) 597-600. [98] Chow, Bennett. Interpolating between Li-Yau 's and Hamtlton's Harnack inequalities on a surface. J. Partial Diff. Equations (China) 11 (1998) 137-140. [99] Chow, Bennett. On an alternate proof of Hamilton's matrix Harnack inequaltty of Li-Yau type for the Rtcci flow. Indiana Univ. Math. J. 52 (2003), 863-874. [100] Chow, Bennettj Chu, Sun-Chin. A geometnc mterpretation of Hamilton's Harnack mequality for the Ricct flow, Math. Res. Lett. 2 (1995), 701-718. [101] Chow, Bennettj Chu, Sun-Chin, A geometnc approach to the linear tmce Harnack inequality for the Ricct flow, Math. Res. Lett. 3 (1996), 549-568. [102] Chow, Bennettj Chu, Sun-Chinj Lu, Pengj Ni, Lei. Notes on Perelman's papers on Ricci flow. Unpublished. [103] Chow, Bennettj Glickenstein, Davidj Lu, Pengo Metric tmnsformatwns under collapsmg of Riemanman mamfolds. Math Res Lett. 10 (2003), No. 5-6, 737-746. [104] Chow, Bennettj Guenther, Christine. Monotonicity formulas for pambolic equatwns m geometry. In preparation. [105] Chow, Bennettj Hamilton, Richard S. Constmined and linear Harnack inequalities for pambolic equattons. Invent. Math. 129 (1997), 213-238. [106] Chow, Bennettj Hamilton, Richard S. The cross curvature flow of 3-manifolds wtth negatwe sectional curvature. Thrkish Journal of Mathematics 28 (2004), 1-10. [107] Chow, Bennettj Knopf, Dan. New Lt-Yau-Hamilton inequalities for the Ricci flow via the space-time approach. J. of Diff. Geom. 60 (2002), 1-54. [108] Chow, Bennettj Knopf, Dan. The Ricci flow: An introduction. Mathematical Surveys and Monographs, AMS, Providence, RI, 2004. [109] Chow, Bennettj Knopf, Danj Lu, Pengo Hamilton's injectivity mdius esttmate for sequences with almost nonnegative curvature opemtors. Comm. Anal. Geom. 10 (2002), no. 5, 1151-1180. [110] Chow, Bennettj Lu, Pengo The time-dependent maxtmum principle for systems of pambolic equations subject to an avotdance set. Pacific J. Math. 214 (2004), no. 2, 201-222.
518
BIBLIOGRAPHY
[111] Chow, Bennett; Lu, Peng; Ni, Lei. Hamilton's Ricci flow. Lectures in Contemporary Mathematics, 3, Science Press and Graduate Studies in Mathematics, 77, American Mathematical Society (co-publication), 2006. [112] Chow, Bennett; Wu, Lang-Fang. The Rzcci flow on compact 2-orbifolds with curvature negative somewhere. Comm. Pure Appl. Math. 44 (1991), no. 3, 275-286. [113] Chu, Sun-Chin. Basic Properties of Gradient Ricci Solztons. In Geometric Evolution Equations, Contemporary Mathematics 367, S.-C. Chang, B. Chow, S.-C. Chu. C.S. Lin, eds., American Mathematical Society, 2005. [114] Chu, Sun-Chin. Geometry of .'i-dimensional gradient Ricci solitons wzth positive curvature. Comm. Anal. Geom. 13 (2005), no. 1, 129-150. [115] Cohn-Vossen, S. Kiirzete Wege und Totalkriimmung auf Fliichen. Compo Math. 2 (1935), 69-133. [116] Colding, Tobias; Minicozzi, William P. II, Estimates for the extinction time for the Ricci flow on certain .'i-manifolds and a question of Perelman. J. Amer. Math. Soc. 18 (2005), 561-569. [117] Courant, R.; Hilbert, David. Methods of mathematical physics. Vols. I and II. Interscience Publishers, Inc., New York, N.Y., 1953 and 1962 (reprinted in 1989). [118] Craioveanu, Mircea; Puta, Mircea; Rassias, Themistocles M. Old and new aspects in spectral geometry. Mathematics and its Applications, 534. Kluwer Academic Publishers, Dordrecht, 2001. [119] Daskalopoulos, Panagiota; del Pino, Manuel A. On a singular difJuszon equation. Comm. Anal. Geom. 3 (1995), 523-542. [120] Daskalopoulos, Panagiota; Hamilton, Richard S. Geometnc estimates for the loganthmic fast diffusion equation. Comm. Anal. Geom. 12 (2004), 143--164. [121] Daskalopoulos, Panagiota; Sesum, Natasa. Eternal Solutions to the Ricci Flow on ]R2. arXiv:math.AP /0603525. [122] Davies, E. B. Heat kernel and spectral theory, Cambridge Univ. Press, Cambridge, 1989. [123] DeThrck, Dennis, Deforming metrics in the dzrection of their Ricci tensors, J. Differential Geom. 18 (1983) 157-162. [124] DeTurck, Dennis M. Deforming metncs in the direction of thezr Ricci tensors, improved version, Collected Papers on Ricci Flow, H.-D. Cao, B. Chow, S.-C. Chu, and S.-T. Yau, eds. Internat. Press, Somerville, MA, 2003. [125] Dijkgraaf, Robbert; Verlinde, Herman; Verlinde, Erik. Stnng propogation in a black hole geometry. Nucl. Phys. B 371 (1992), 269-314. [126] Ding, Yu. Notes on Perelman's second paper. http://math.ucLeduryding/ perelman. pdf. [127] do Carmo, Manfredo Perdigao. Riemannian geometry. Translated from the second Portuguese edition by Francis Flaherty. Mathematics: Theory & Applications. Birkhiiuser Boston, Inc., Boston, MA, 1992. [128] Donaldson, Simon K. Anti self-dual Yang-Mills connections over complex algebraic surfaces and stable vector bundles. Proc. London Math. Soc. (3) 50 (1985), no. 1, 1-26. [129] Donaldson, Simon K. Scalar curvature and stability of toric vanetzes. J. Differential Geom. 62 (2002), no. 2, 289-349. [130] Donaldson, Simon K. Lower bounds on the Calabi functional. J. Differential Geom. 70 (2005), no. 3, 453-472. [131] Drazin, P.; Johnson, N. Solitons: an introduction. Cambridge, 1989. [132] Ecker, Klaus. Regularity theory for mean curvature flow. Progress in Nonlinear Differential Equations and their Applications, 57. Birkhiiuser Boston, Inc., Boston, MA,2004. [133] Ecker, Klaus. A local monotonicity formula for mean curvature flow. Ann. of Math. (2) 154 (2001), no. 2, 503-525.
BIBLIOGRAPHY
519
[134J Ecker, Klaus; Knopf, Dan; Ni, Lei; Topping, Peter. Local mono tonicity and mean value formulas for evolvmg Riemannian manifolds. Preprint. [135J Eells, James, Jr.; Sampson, J. H. Harmonic mappings of Riemannian manifolds. Amer. J. Math. 86 (1964), 109-160. [136J Epstein, D. B. A. Periodic flows on three-manifolds. Ann. of Math. (2) 95 1972 66-82. [137J Evans, Lawrence. Partial differential equations. AMS. Providence, 1998. [138J Evans, Lawrence C. Entropy and partial differentzal equations. Lecture Notes at UC Berkeley. 212 pp. http://math.berkeley.edur evans/entropy.and.PDE.pdf [139J Evans, Lawrence; Gariepy, R. F. Measure theory and fine properties of functions. CRC Press, Boca Raton, 1992. [140J Fabes E. B.; Garofalo, N. Mean value properties of solutions to parabolic equations with variable coefficients, Jour. Math. Anal. Appl. 121 (1987), 305-316. [141J Fateev, V. A. ; Onofri, E.; Zamolodchikov, A. B. The sausage model (integrable deformations of 0(3) sigma model). Nucl. Phys. B 406 (1993), 521-565. [142J Feldman, Mikhail; Ilmanen, Tom; Knopf, Dan. Rotationally symmetric shrinking and expanding gradient Kahler-Ricci solitons. J. Differential Geom. 65 (2003), no. 2, 169-209. [143J Feldman, Mikhail; Ilmanen, Tom; Ni, Lei. Entropy and reduced distance for Ricci expanders. arXiv:math.DG/0405036. [144J Frankel, Theodore. Manifolds with positive curvature. Pacific J. Math. 11 (1961), 165-174. [145J Friedan, Daniel Harry. Nonlinear models in 2 + c dimensions. Ann. Physics 163 (1985), no. 2, 318--419. [146J Friedman, Avner. Partial differential equations of parabolic type. Robert E. Krieger Publishing Company, 1983, Malabar, Florida. [147J Futaki, Akito. An obstruction to the existence of Einstein-Kahler metrics. Invent. Math. 73 (1983), no. 3, 437--443. [148J Futaki, Akito. Kahler-Einstein metrics and integral invariants. Lecture Notes in Math., 1314, Springer-Verlag, Berlin, 1988. [149J Gabai, David. Convergence groups are F'uchsian groups. Ann. of Math. (2) 136 (1992), no. 3, 447-510. [150J Gabai, David; Meyerhoff, G. Robert; Thurston, Nathaniel. Homotopy hyperbolic 3-manifolds are hyperbolic. Ann. of Math. (2) 157 (2003), no. 2, 335--43l. [151J Gage, M.; Hamilton, Richard S. The heat equation shrinking convex plane curves. J. Differential Geom. 23 (1986), no. 1, 69-96. [152J Gao, L. Zhiyong. Convergence of Riemannian manifolds; Ricci and L n/2 -curvature pinching. J. Differential Geom. 32 (1990), no. 2, 349-38l. [153J Garofalo, N.; Lanconelli, E. Asymptotic behavior of fundamental solutions and potential theory of parabolic operators with variable coefficients. Math. Ann. 283 (1989) no. 2, 211-239. [154J Gastel, Andreas; Kronz, Manfred. A family of expanding Ricci solitons. Variational problems in Riemannian geometry, 81-93, Progr. Nonlinear Differential Equations Appl., 59, Birkhhauser, Basel, 2004. [155J Gilbarg, David; Trudinger, Neil S. Elliptic partial differential equations of second order. Reprint of the 1998 edition. Classics in Mathematics. Springer-Verlag, Berlin, 200l. [156J Glickenstein, David. Precompactness of solutions to the Ricci flow in the absence of injectivity radius estimates. Geom. Topol. 7 (2003), 487-510 (electronic). [157J Goldberg, Samuel I. Curvature and homology. Revised reprint of the 1970 edition. Dover Publications, Inc., Mineola, NY, 1998. [158J Gordon, C. MeA.; Heil, Wolfgang. Cyclic normal subgroups of fundamental groups of 3-mamfolds. Topology 14 (1975), no. 4, 305-309.
520
BIBLIOGRAPHY
[159] Gray, Brayton. Homotopy theory. An introduction to algebraic topology. Pure and Applied Mathematics, Vol. 64. Academic Press [Harcourt Brace Jovanovich, Publishers], New York-London, 1975. [160] Grayson, Matthew A. Shortening embedded curves. Ann. of Math. (2) 129 (1989), no. 1, 71-111. [161] Grayson, Matthew; Hamilton, Richard S. The formation of singularities in the harmomc map heat flow. Comm. Anal. Geom. 4 (1996), no. 4, 525-546. [162] Green, M.B.; Schwarz, J.H.; Witten, E. Superstring Theory, Volumes 1 & 2. Cambridge University Press, Cambridge, 1987. [163] Greene, R. E.; Wu, H. Coo convex functions and manifolds of positive curvature. Acta Math. 137 (1976), no. 3-4, 209-245. [164] Greene, R. E.; Wu, H. Functwn theory on manifolds which possess a pole. Lecture Notes in Mathematics, 699. Springer, Berlin, 1979. [165] Greene, R. E.; Wu, H. Lipschitz convergence of Riemannian manifolds. Pacific J. Math. 131 (1988), no. 1, 119-141. [166] Griffiths, Phillip; Harris, Joseph. Principles of algebraic geometry. Reprint of the 1978 original. Wiley Classics Library. John Wiley & Sons, Inc., New York, 1994. [167] Grigor'yan, Alexander. Upper bounds of derivatzves of the heat kernel on an arbitrary complete manifold. J. Flmct. Anal. 127 (1995), no. 2, 363-389. [168] Grigor'yan, Alexander. Gaussian upper bounds for heat kernel on arbitrary manifolds, J. Differential Geom. 45 (1997), 33-52. [169] Gromov, Misha. Metnc structures for Riemannian and non-Rzemanman spaces. Based on the 1981 French original. With appendices by M. Katz, P. Pansu and S. Semmes. Translated from the French by Sean Michael Bates. Progress in Mathematics, 152. Birkhiiuser Boston, Inc., Boston, MA, 1999. [170] Gross, Leonard. Logarithmic Sobolev inequalities. Amer. J. Math. 97 (1975), no. 4, 1061-1083. [171] Guenther, Christine M. The fundamental solution on manifolds with time-dependent metncs. J. Geom. Anal. 12 (2002), no. 3, 425-436. [172] Guenther, Christine; Isenberg, Jim; Knopf, Dan, Stabilzty of the Ricci flow at Ricciflat metrics, Comm. Anal. Geom. 10 (2002), no. 4, 741-777. [173] Gursky, Matthew. The Weyl functional, de Rham cohomology, and Kiihler-Einstein metrics, Ann. of Math. 148 (1998), 315-337. [174] Gutperle, Michael; Headrick, Matthew; Minwalla, Shiraz; Schomerus, Volker. Spacetzme energy decreases under world-sheet RG flow. arXiv:hep-thj0211063. [175] Haagensen, Peter E., Duality Transformations Away Prom Conformal Points. Phys. Lett. B 382 (1996), 356-362. [176] Haken, Wolfgang. Theorie der Normalfliichen. Acta Math. 105 (1961), 245-375. [177] Hamilton, Richard S. Harmonic maps of manifolds with boundary. Lecture Notes in Mathematics, Vol. 471. Springer-Verlag, Berlin-New York, 1975. [178] Hamilton, Richard S. Three-manifolds with positive Ricci curvature. J. Differential Geom. 17 (1982), no. 2, 255-306. [179] Hamilton, Richard S. Four-manifolds with positive curvature operator. J. Differential Geom. 24 (1986), no. 2,153-179. [180] Hamilton, Richard S. The Ricci flow on surfaces. Mathematics and general relativity (Santa Cruz, CA, 1986), 237-262, Contemp. Math., 71, Amer. Math. Soc., Providence, RI, 1988. [181] Hamilton, Richard S. The Harnack estimate for the Ricci flow. J. Differential Geom. 37 (1993), no. 1, 225-243. [182] Hamilton, Richard S. Eternal solutions to the Ricci flow, J. Differential Geom. 38 (1993), 1-11. [183] Hamilton, Richard. A matrix Harnack estimate for the heat equation, Comm. Anal. Geom. 1 (1993), 113-126.
BIBLIOGRAPHY
521
[184] Hamilton, Richard. Monotonicity formulas for parabolic flows on mamfolds, Comm. Anal. Geom. 1 (1993), 127-137. [185] Hamilton, Richard S. Convex hypersurfaces with pinched second fundamental form. Comm. Anal. Geom. 2 (1994), no. 1, 167-172. [186] Hamilton, Richard S. The formatwn of singularitzes in the Rzcci flow. Surveys in differential geometry, Vol. II (Cambridge, MA, 1993), 7-136, Internat. Press, Cambridge, MA, 1995. [187] Hamilton, Richard S. A compactness property for solutions of the Ricci flow. Amer. J. Math. 117 (1995), no. 3, 545-572. [188] Hamilton, Richard S. Harnack estimate for the mean curvature flow, J. Diff. Geom. 41 (1995) 215-226. [189] Hamilton, Richard S. Four-manifolds with posztwe zsotropic curvature. Comm. Anal. Geom. 5 (1997), no. 1, 1-92. [190] Hamilton, Richard S. Non-singular solutwns of the Ricci flow on three-manifolds. Comm. Anal. Geom. 7 (1999), no. 4, 695-729. [191] Hamilton, Richard S. Three-orbifolds with positwe Ricci curvature. In Collected Papers on Ricci Flow, H.-D. Cao, B. Chow, S.-C. Chu, and S.-T. Yau, eds. Internat. Press, Somerville, MA, 2003. [192] Hamilton, Richard S. Differenhal Harnack estimates for parabolic equatzons, preprint. [193] Hamilton, Richard S.; Sesum, Natasa. Properties of the solutions of the conJugate heat equation. arXiv:math.DG/0601415. [194] Hamilton, Richard S.; Yau, Shing-Tung. The Harnack estimate for the Ricci flow on a surface-revisited. Asian J. Math. 1 (1997), no. 3, 418-421. [195] Han, Qing; Lin, Fanghua. Elliptic Partial Differenhal Equations. AMS/New York University Press, 1997. [196] Hartman, Philip. Ordmary dlfferentzal equations. Corrected reprint of the second (1982) edition. Birkhii.user, Boston, MA. [197] Hass, Joel; Morgan, Frank. Geodesics and soap bubbles in surfaces. Math. Z. 223 (1996), no. 2, 185-196. [198] Heber, Jens. Noncompact homogeneous Einstem spaces. Invent. Math. 133 (1998), no. 2, 279-352. [199] Hebey, Emmanuel. Introductwn a l'analyse non lmeaire sur les varietes. Diderot Editeur, Arts et Sciences, 1997. [200] Helgason, Sigurdur. Differential geometry, Lle groups, and symmetnc spaces. Pure and Applied Mathematics, 80. Academic Press, Inc. [Harcourt Brace Jovanovich], New York-London, 1978. [201] Hempel, John. 3-Manifolds. Ann. of Math. Studies, No. 86. Princeton University Press, Princeton, N. J.; University of Tokyo Press, Tokyo, 1976. [202] Hiriart-Urruty, Jean-Baptiste; Lemarechal, Claude. Fundamentals of convex analysis. Springer, 2001. [203] Hirsch, Morris W. Obstructwn theories for smoothmg mamfolds and maps. Bull. Amer. Math. Soc. 69 (1963), 352-356. [204] Hirzebruch, F.; Kodaira, K. On the complex proJectwe spaces. J. Math. Pures Appl. 36 (1957), 201-216. [205] Hormander, Lars. Notwns of convexlty. Birkhauser, 1994. [206] Hsu, Shu-Yu. Large time behaviour of solutwns of the Ricci flow equation on ]R2. Pacific J. Math. 197 (2001), no. 1, 25-41. [207] Hsu, Shu-Yu. Global existence and uniqueness of solutions of the Ricci flow equation. Differential Integral Equations 14 (2001), no. 3, 305-320. [208] Hsu, Shu-Yu. A slmple proof on the nonexzstence of shrinkmg breathers for the Ricci flow. arXiv: math.DG/0509087.
522
BIBLIOGRAPHY
[209] Huang, Hong. A note on Morse's index theorem for Perelman's C-length. arXiv:math.DG /0602090. [210] Huber, Alfred. On subharmonic functions and dzfferential geometry in the large. Comment. Math. Helv. 32 (1957), 13-72. [211] Huisken, Gerhard. Flow by mean curvature of convex surfaces mto spheres, J. Differential Geom. 20 (1984), no. 1, 237-266. [212] Huisken, Gerhard. Rzccz deformatwn of the metNc on a Rzemannian manifold. J. Differential Geom. 21 (1985), no. 1,47-62. [213] Huisken, Gerhard. Asymptotic behavior for singularities of the mean curvature flow. J. Differential Geom. 31 (1990), no. 1, 285-299. [214] Ishii, Hitoshi; Lions, Pierre-Louis. Viscosity solutions of fully nonlinear second-order elliptic partial differential equations. J. Differential Equations 83 (1990), no. 1, 2678. [215] Iskovskih, V. A. Fano threefolds. 1. Izv. Akad. Nauk SSSR Ser. Mat. 41 (1977), no. 3, 516-562, 717. [216] Iskovskih, V. A. Fano threefolds. II. Izv. Akad. Nauk SSSR Ser. Mat. 42 (1978), no. 3, 506-549. [217] Ivey, Thomas. On solztons for the Ricci Flow. PhD thesis, Duke University, 1992. [218] Ivey, Thomas. Riccz solztons on compact three-mamfolds. Diff. Geom. Appl. 3 (1993), 301-307. [219] Ivey, Thomas. New examples of complete Ricci solztons. Proc. Amer. Math. Soc. 122 (1994), 241-245. [220] Ivey, Thomas. The Ricci flow on radwlly symmetric ]R3. Comm. Part. Diff. Eq. 19 (1994), 1481-1500. [221] Ivey, Thomas. Local exzstence of Rzcci solitons. Manuscripta Math. 91 (1996), 151162. [222] Ivey, Thomas. Rzccz solitons on compact Kahler surfaces. Proc. Amer. Math. Soc. 125 (1997), no. 4, 1203-1208. [223] Ivey, Thomas A.; Landsberg, J. M. Cartan for begmners: differentwl geometry via moving frames and exterior differential systems. Graduate Studies in Mathematics, 61. American Mathematical Society, Providence, RI, 2003. [224] Jaco, William; Shalen, Peter B. Seifert fibered spaces m 3-manifolds. Geometric topology (Proc. Georgia Topology Conf., Athens, Ga., 1977), pp. 91-99, Academic Press, New York-London, 1979. [225] Johannson, Klaus. Homotopy equivalences of 3-manifolds with boundaries. Lecture Notes in Mathematics, 761. Springer, Berlin, 1979. [226] Juutinen, Petri; Lindqvist, Peter; Manfredi, Juan J. On the equwalence of viscoszty solutwns and weak solutwns for a quasi-linear equation. SIAM J. Math. Anal. 33 (2001), no. 3, 699-717. [227] H. Karcher, Rzemanman center of mass and mollzfier smoothing, Comm. Pure Appl. Math. 30 (1977), 509-54l. [228] Karp, Leon; Li, Peter. The heat equatwn on complete Riemannian manifolds. Unpublished manuscript. [229] Kasue, A. A compactzfication of a mamfold with asymptotically nonnegative curvature. Ann. Scient. Ec. Norm. Sup. 21 (1988), 593-622. [230] Kato, Tosio. Perturbation theory for 1m ear operators. Reprint of the 1980 edition. Classics in Mathematics. Springer-Verlag, Berlin, 1995. [231] Kleiner, Bruce; Lott, John. Notes on Perelman's papers. May 16, 2006. http://www.math.lsa.umich.edurlott/riccifiow /perelman.html [232] Knapp, Anthony W. Lie groups, Lie algebras, and cohomology. Mathematical Notes, 34. Princeton University Press, Princeton, NJ, 1988. [233] Kneser, H. GeschlofJene Flachen in dreidimensionalen Mannigfaltikeiten, Jahresbericht der Deut. Math. Verein. 38 (1929), 248-260.
BIBLIOGRAPHY
523
[234J Knopf, Dan. Positimty of Ricc~ curvature under the Kiihler-Ricc~ flow. Commun. Contemp. Math. 8 (2006), no. 1, 123-133. [235J Knopf, Dan; Young, Andrea. Asymptotic stability of the cross curvature flow at a hyperbolic metric. Preprint. [236J Kobayashi, Shoshichi; Nomizu, Katsumi. Foundations of differential geometry. Vols. I & II. Reprint of the 1963 and 1969 originals. Wiley Classics Library. A WileyInterscience Publication. John Wiley & Sons, Inc., New York, 1996. [237J Kobayashi, Shoshichi; Ochiai, Takushiro. Charactenzatwns of complex projective spaces and hyperquadncs. J. Math. Kyoto Univ. 13 (1973), 31-47. [238J Kodaira, Kunihiko. Complex manifolds and deformatwn of complex structures. Translated from the 1981 Japanese original by Kazuo Akao. Reprint of the 1986 English edition. Classics in Mathematics. Springer-Verlag, Berlin, 2005. [239J Koiso, Norihito. On rotationally symmetnc Hamilton's equatwn for KiihlerEinstein metncs. Recent topics in differential and analytic geometry, 327-337, Adv. Stud. Pure Math., 18-1, Academic Press, Boston, MA, 1990. [240J Kotschwar, Brett. Hamilton's grad~ent est~mate for the heat kernel on complete mantfolds. Proc. A.M.S. To appear. [241J Kotschwar, Brett. A note on the umqueness of complete expanding Rtcci solitons m 2-d. Preprint. [242J Krylov, N. V.; Safonov, M. V. A property of the solutions of parabolic equations with measurable coefficients (Russian), Izv. Akad. Nauk SSSR Ser. Mat. 44(1980), no. 1, 161-175. [243J Ladyzenskaja, O. A.; Solonnikov, V. A.; Uralceva, N. N. Lmear and quasilmear equations of parabolic type. (Russian) Translated from the Russian by S. Smith. Translations of Mathematical Monographs, Vol. 23, American Mathematical Society, Providence, R.I., 1967. [244J Lauret, Jorge. Ricci soliton homogeneous mlmanifolds. Math. Ann. 319 (2001), no. 4,715-733. [245J Li, Peter. On the Sobolev constant and the p-spectrum of a compact Riemannian mamfold. Ann. Sc. Ec. Norm. Sup. 4e serie, t. 13 (1980),451-469. [246J Li, Peter. Lecture notes on geometric analysis, RIMGARC Lecture Notes Series 6, Seoul National University, 1993. [247J Li, Peter. Large time behavior of the heat equation on complete manifolds with nonnegative Ricct curvature. Ann. of Math. (2) 124 (1986), no. 1, 1-2l. [248J Li, Peter. Lecture notes on geometric analysis. http://math.uci.edu/ pli/lecture.pdf. [249J Li, Peter. Lecture notes on heat kernel. Lecture Notes at UCI taken by Jiaping Wang. [250J Li, Peter. Lecture notes on harmonic functions. To appear. [251J Li, Peter. Curvature and function theory on Rtemanman mamfolds. Surveys in Differential Geometry: Papers dedicated to Atiyah, Bott, Hirzebruch, and Singer. Vol VII, International Press (2000), 375-432. [252J Li, Peter; Tam, Luen-Fai; Wang, Jiaping. Sharp bounds for the Green's function and the heat kernel. Math. Res. Lett. 4 (1997), no. 4, 589-602. [253J Li, Peter; Yau, Shing-Tung. On the parabolic kernel of the Schrodmger operator. Acta Math. 156 (1986), no. 3-4, 153-201. [254J Lichnerowicz, Andre. Geometrie des groupes de transformatwns. (French) Travaux et Recherches Mathematiques, III, Dunod, Paris, 1958. [255J Lieberman, Gary M. Second order parabolic dtfferential equatwns. World Scientific Publishing Co., River Edge, NJ, 1996. [256J Lott, John. On the long-time behavior of type-III Ricci flow solutions. arXiv:math.DG/0509639. [257J Lu, Pengo A compactness property for solutions of the Ricci flow on orbifolds. Amer. J. Math. 123 (2001), no. 6, 1103-1134. [258J Lu, Pengo Unpublished.
524
BIBLIOGRAPHY
[259) Ma, Li; Chen, Dezhong. Examples for cross curvature flow on 3-manifolds. Calc. Var. Partial Differential Equations 26 (2006), no. 2, 227-243. [260) Mabuchi, Toshiki. 1(:3 -actwns and algebrazc threefolds wzth ample tangent bundle. Nagoya Math. J. 69 (1978), 33-64. [261) Mabuchi, Toshiki. K -energy maps integratmg Futakz invanants. Tohoku Math. J. 38 (1986), 575-593. [262) McMullen, C. Renormalizatwn and 3-manifolds which fiber over the czrcle. Annals of Mathematics Studies 12, Princeton U. Press, 1996. [263) Meeks, William H., III; Yau, Shing Tung. Topology of three-dzmensional manifolds and the embedding problems m mmzmal surface theory. Ann. of Math. (2) 112 (1980), no. 3, 441-484. [264) Milnor, John. A umque factonzation theorem for 3-mamfolds. Amer. J. Math. 84 (1962) 1-7. [265) Milnor, John W. Morse theory. Based on lecture notes by M. Spivak and R. Wells. Annals of Mathematics Studies, No. 51. Princeton University Press, Princeton, N.J., 1963. [266) Milnor, John. Curvatures of left invariant metncs on Lie groups. Advances in Math. 21 (1976), no. 3, 293-329. [267) Milnor, John. Towards the Pomcare conjecture and the classification of 3-mamfolds. Notices AMS 50 (2003), 1226-1233. [268) Moise, E. E., Affine structures m 3-mamfolds. Ann. of Math. (2) 56 (1952), 96-114. [269) Mok, Ngaiming. The umformzzation theorem for compact Kahler manifolds of nonnegative holomorphtc bisectional curvature. J. Differential Geom. 21 (1988), no. 2, 179-214. [270) Mok, Ngaiming. Metric ngzdzty theorems on Hermztian locally symmetnc manifolds. Series in Pure Mathematics, 6. World Scientific Publishing Co., Inc., Teaneck, NJ, 1989. [271) Mori, Shigefumi. Projective manifolds with ample tangent bundles. Ann. of Math. (2) 110 (1979), no. 3, 593-606. [272) Morgan, John. Recent progress on the Pomcare conjecture and the classification of 3-manifolds. Bull. Amer. Math. Soc. 42 (2005), 57-78. [273) Morgan, John; Tian, Gang. Rzcci flow and the Pomcare conjecture. arXiv:mat.h.DG/0607607. [274) Morrey, C.B. Multzple mtegrals in the calculus of vanations. Springer-Verlag, Berlin, 1966. [275) Morrow, James; Kodaira, Kunihiko. Complex mamfolds. Holt, Rinehart and Winston, Inc., New York-Montreal, Que.-London, 1971. [276) Moser, Jiirgen. On Harnack's theorem for ellzptzc difJerential equations. Comm. Pure Appl. Math. 14 (1961), 577-59l. [277) Moser, Jiirgen. A Harnack mequalzty for parabolic difJerentzal equations, Comm. Pure Appl. Math. 11(1964) 101-134. [278) Mukai, Shigeru. Fano 3-folds. Complex projective geometry (Trieste, 1989/Bergen, 1989),255-263, London Math. Soc. Lecture Note Ser., 119, Cambridge Univ. Press, Cambridge, 1992. [279) Munkres, James. Obstructwns to the smoothmg of pzecewzse-dzfJerentiable homeomorphzsms. Ann. of Math. (2) 12 (1960), 521-554. [280) Myers, Sumner; Steenrod, Norman. The group of zsometries of a Rzemannian manifold. Annals Math. 40 (1939), 400-416. [281) Nadel, Alan M. Multiplzer ideal sheaves and Kahler-Einstem metrics of positive scalar curvature. Ann. of Math. (2) 132 (1990), no. 3, 549-596. [282) Nash, John. Continmty of solutions of parabolic and ellzptzc equations, Amer. J. Math. 80 (1958), 931-954.
BIBLIOGRAPHY
525
[283] Ni, Lei. The entropy formula for linear heat equation. Journal of Geometric Analysis, 14 (2004), 85-98. Addenda, 14 (2004), 369-374. [284] Ni, Lei. A mono tonicity formula on complete Kahler manifolds with nonnegatwe btsectional curvature. J. Amer. Math. Soc. 17 (2004), no. 4, 909-946. [285] Ni, Lei. Monotomctty and Kahler-Ricci flow. In Geometric Evolution Equations, Contemporary Mathematics, Vol. 367, S.-C. Chang, B. Chow, S.-C. Chu. C.-S. Lin, eds. American Mathematical Society, 2005. [286] Ni, Lei. Ancient solution to Kahler-Riccz flow. Math. Res. Lett. 12 (2005), 633-654. [287] Ni, Lei. A matrix Li- Yau-Hamilton estimate for Kahler-Ricct flow. J. Diff. Geom. 75 (2007), 303-358. [288] Ni, Lei. A note on Perelman's Lt- Yau-Hamilton mequality. Comm. Anal. Geom. 14 (2006). [289] Ni, Lei. Unpublished. [290] Ni, Lei; Tam, Luen-Fai. Plurtsubharmonic functwns and the Kahler-Rtcci flow. Amer. J. Math. 125 (2003), 623-645. [291] Ni, Lei; Tam, Luen-Fai. Plurtsubharmomc functwns and the structure of complete Kahler mamfolds with nonnegative curvature. J. Differential Geom. 64 (2003), no. 3,457-524. [292] Ni, Lei; Tam, Luen-Fai. Kahler-Rtcctflow and the Pomcare-Lelong equatwn. Comm. Anal. Geom. 12 (2004), 111-141. [293] Nitta, Muneto. Conformal sigma models with anomalous dimensions and Ricci solitons. Modern Phys. Lett. A 20 (2005), no. 8, 577-584. [294] Otal, J.-P. Le theoreme d'hyperbolisation pour les varieUs fibrees de dimension 3. Aterisque 235, 1996. [295] Pali, Nefton. Characterization of Einstein-Fano manifolds via the Kahler-Rtcctflow. arXiv:math.DG /0607581. [296] PattIe, R. E. Diffusion from an mstantaneous point source with a concentrationdependent coefficient. Quart. J. Mech. Appl. Math. 12 (1959), 407-409. [297] Perelman, Grisha. The entropy formula for the Ricci flow and its geometrtc applicattons. arXiv:math.DG/0211159. [298] Perelman, Grisha. Ricci flow wtth surgery on three-manifolds. arXiv:math.DG/ 0303109. [299] Perelman, Grisha. Fmite extmctton ttme for the soluttons to the Ricci flow on certain three-mamfolds. arXiv:math.DG /0307245. [300] Peters, Stefan. Convergence of Rtemanman mamfolds. Compositio Math. 62 (1987), no. 1,3-16. [301] Petersen, Peter. Convergence theorems in Riemanman geometry. Comparison geometry (Berkeley, CA, 1993-94), 167-202, Math. Sci. Res. Inst. Publ., 30, Cambridge Univ. Press, Cambridge, 1997. [302] Petersen, Peter. Riemannian geometry. Graduate Texts in Mathematics, 171. Springer-Verlag, New York, 1998. [303] Petrunin, A.; Thschmann, W. Asymptotic flatness and cone structure at infinity. Math. Ann. 321 (2001),775-788. [304] Phong, D. H.; Sturm, Jacob. Stability, eneryy functwnals, and Kahler-Einstein metrics. Comm. Anal. Geom. 11 (2003), no. 3, 565-597. [305] Phong, D.H.; Sturm, Jacob. On the Kahler-Ricci flow on complex surfaces. arXiv:math.DG /0407232. [306] Phong, D.H.; Sturm, Jacob. On stabthty and the convergence of the Kahler-Ricci flow. arXiv:math.DG /0412185. [307] Polchinksi, Joe, String theory. Vols. I and II. Cambridge Monographs on Mathematical Physics. Cambridge University Press, Cambridge, 1998. [308] Poor, Walter A. Differential geometric structures. McGraw-Hill Book Co., New York, 1981.
526
BIBLIOGRAPHY
[309] Protter, Murray H.; Weinberger, Hans F. Maximum principles in differential equations. Corrected reprint of the 1967 original. Springer-Verlag, New York, 1984. [310] Reed, Michael; Simon, Barry. Methods of modern mathematical physics, Analysis of operators, Volume IV. Academic Press, New York, 1978. [311] Rosenau, Philip. On fast and super-fast diffusion. Phys. Rev. Lett. 74 (1995),10561059. [312] Rothaus, O. S., Logarithmic Sobolev inequalities and the spectrum of Schrodmger operators. J. Funct. Anal. 42 (1981), 110-120. [313] Royden, Halsey L. Real analysis. Third edztion. Macmillan Publishing Company, New York, 1988. [314] Ruan, Wei-Dong. On the convergence and collapsmg of Kiihler metncs. J. Differential Geom. 52 (1999), no. I, 1-40. [315] Saloff-Coste, Laurent. Umformly elliptic operators on Riemannian manifolds. J. Differential Geom. 36 (1992), no. 2, 417-450. [316] Schoen, R.; Yau, S.-T. Lectures on differential geometry. Lecture notes prepared by Wei Yue Ding, Kung Ching Chang [Gong Qing Zhang], Jia Qing Zhong and Yi Chao Xu. Translated from the Chinese by Ding and S. Y. Cheng. Preface translated from the Chinese by Kaising Tso. Conference Proceedings and Lecture Notes in Geometry and Topology, I. International Press, Cambridge, MA, 1994. [317] Schueth, Dorothee. On the 'standard' condition for noncompact homogeneous Einstein spaces. Geom. Dedicata 105 (2004), 77-83. [318] Scott, Peter. A new proof of the annulus and torus theorems. Amer. J. Math. 102 (1980), no. 2, 241-277. [319] Scott, Peter. The geometries of 3-manifolds. Bull. London Math. Soc. 15 (1983), no. 5,401-487. [320] Sesum, Natasa. Convergence of Kiihler-Einstein orbifolds. J. Geom. Anal. 14 (2004), no. I, 171-184. [321] Sesum, Natasa. Curvature tensor under the Ricci flow. Amer. J. Math. 127 (2005), no. 6, 1315-1324. [322] Sesum, Natasa. Limiting behaviour of the Ricci flow. arXiv:math.DGj0402194. [323] Sesum, Natasa. Convergence of a Kiihler-Ricc~ flow. Math. Res. Lett. 12 (2005), no. 5-6, 623-632. [324] Sesum, Natasa. Convergence of the Ricci flow toward a soliton. Comm. Anal. Geom. 14 (2006), no. 2, 283-343. [325] Sesum, Natasa; Tian, Gang. Bounding scalar curvature and diameter along the Kiihler-Riccz flow (after Perelman). Preprint. [326] Sesum, Natasa; Tian, Gang; Wang, Xiaodong, Notes on Perelman's paper on the entropy formula for the Ricci flow and its geometric applwations. June 23, 2003. [327] Shalen, Peter B. A "piecewise-linear" method for tnangulating 3-manifolds. Adv. in Math. 52 (1984), no. I, 34-80. [328] Sharafutdinov, V. A. The Pogorelov-Klingenberg theorem for manifolds that are homeomorph~c to !R.n. (Russian) Sibirsk. Mat. Z. 18 (1977), no. 4, 915-925, 958. [329] Shi, Wan-Xiong. Deforming the metric on complete Rzemannian manifolds. J. Differential Geom. 30 (1989), no. I, 223-301. [330] Shi, Wan-Xiong. R~cci deformatwn of the metric on complete noncompact R~emann ian manifolds. J. Differential Geom. 30 (1989), no. 2, 303-394. [331] Shi, Wan-Xiong. Ricci flow and the uniformization on complete noncompact Kiihler manifolds. J. Differential Geom. 45 (1997), no. I, 94-220. [332] Shioya, Takashi; Yamaguchi, Takao. Collapsmg three-mamfolds under a lower curvature bound. J. Differential Geom. 56 (2000), no. I, 1-66. [333] Simon, Leon. Schauder estimates by scaling. Calc. Var. Partial Differential Equations 5 (1997), no. 5, 391-407.
BIBLIOGRAPHY
527
[334] Siu, Yum Tong. Lectures on Hermitzan-Emstein metrics for stable bundles and Kahler-Emstem metrics. DMV Seminar, 8. Birkhauser Verlag, Basel, 1987. [335] Siu, Yum Tong. The existence of Kahler-Einstein metrics on manifolds with positive anhcanonical line bundle and a suitable fimte symmetry group. Annals Math. 127 (1988), 585-627. [336] Siu, Yum Tong; Yau, Shing-Tung. Compact Kahler mamfolds of positive bisectwnal curvature. Invent. Math. 59 (1980), no. 2, 189-204. [337] Song, Jian; Tian, Gang. The Kahler-Riccz flow on surfaces of positive Kodaira dzmension. arXiv:math.DG/0602150. [338] Song, Jian; Weinkove, Ben. Energy functwnals and canomcal Kahler metncs. Preprint. [339] Souplet, Philippe; Zhang, Qi. Sharp gradzent estimate and Yau's Lwuville theorem for the heat equation on noncompact manifolds. arXiv:math.DG/0502079. [340] Stein, Elias. Smgular integrals and difJerenhability propertzes of functions. Princeton Univ. Press, 1970. [341] Strominger, Andrew; Vafa, Cumrun. Microscopic ongin of the Bekenstein-Hawking entropy. arXiv:hep-th/9601029. [342] Struwe, Michael. On the evolution of harmonic maps in higher dimensions. J. Differential Geom. 28 (1988), no. 3, 485-502. [343] Thurston, William P. Three-dimensional geometry and topology. Vol. 1. Edited by Silvio Levy. Princeton Mathematical Series, 35. Princeton University Press, Princeton, NJ, 1997. [344] Tian, Gang. On Kahler-Einstein metrics on certam Kahler manifolds wzth C 1 > o. Invent. Math. 89 (1987), no. 2, 225-246. [345] Tian, Gang. Calabi's conjecture for complex surfaces with positive first Chern class. Invent. Math. 101 (1990), no. 1, 101-172. [346] Tian, Gang. Kahler-Emstem metncs with positive scalar curvature. Invent. Math. 130 (1997), no. 1, 1-37. [347] Tian, Gang. Canomcal metrics m Kahler geometry. Notes taken by Meike Akveld. Lectures in Mathematics ETR Zurich. Birkhauser Verlag, Basel, 2000. [348] Tian, Gang; Yau, Shing-Tung. Kahler-Einstein metrics on complex surfaces wdh C 1 > o. Comm. Math. Phys. 112 (1987), no. 1, 175-203. [349] Tian, Gang; Yau, Shing-Tung. Complete Kahler mamfolds with zero Ricci curvature. I. J. Amer. Math. Soc. 3 (1990), no. 3, 579-609. [350] Tian, Gang; Yau, Shing-Tung. Complete Kahler manifolds with zero Rzccz curvature. II. Invent. Math. 106 (1991), no. 1, 27-60. [351] Tian, Gang; Zhang, Zhou. Preprint. [352] Tian, Gang; Zhu, Xiaohua. Umqueness of Kahler-Ricci sohtons on compact Kahler manifolds. C. R. Acad. Sci. Paris Ser. I Math. 329 (1999), no. 11, 991-995. [353] Tian, Gang; Zhu, Xiaohua. Umqueness of Kahler-Riccz solitons. Acta Math. 184 (2000), no. 2, 271-305. [354] Tian, Gang; Zhu, Xiaohua. A new holomorphic mvanant and umqueness of KahlerRzccz solztons. Comment. Math. Relv. 77 (2002), no. 2, 297-325. [355] Tian, Gang; Zhu, Xiaohua. Preprint. [356] Topping, Peter. Lectures on the Riccz flow. London Mathematical Society Lecture Note Series (No. 325). Cambridge University Press, 2006. [357] Topping, Peter. Diameter control under Ricci flow. Comm. Anal. Geom. 13 (2005) 1039-1055. [358] Topping, Peter. Rzcci flow compactness via pseudolocalzty, and flows with mcomplete imhal metncs. Preprint. [359] Tso, Kaising. On an Aleksandrov-Bakelman type maximum principle for secondorder parabolic equatzons. Comm. Partial Diff. Equations 10 (1985), no. 5, 543-553.
528
BIBLIOGRAPHY
[360] Tsuji, Hajime. Existence and degeneratwn of Kiihler-Emstein metncs on mmimal algebrazc varietzes of general type. Math. Ann. 281 (1988), no. 1, 123-133. [361] Varadhan, S. R. S. On the behavwr of the fundamental solutzon of the heat equation with variable coefficients. Comm. Pure Appl. Math. 20 (1967),431-455. [362] Varopoulos, N. Hardy-Littlewood theory for semigroups. J. Funet. Anal. 63 (1985), 240-260. [363] Waldhausen, FriedheIm. On irreduczble 3-manifolds which are sufficiently large. Ann. of Math. (2) 87 (1968), 56-88. [364] Waldhausen, Friedholm. Cruppen mzt Zentrum und 3-dimenswnale Mannigfaltzgkezten. Topology 6 (1967), 505-517. [365] Wang, McKenzie Y.; Ziller, Wolfgang. Exzstence and nonexistence of homogeneous Emstem metrics. Invent. Math. 84 (1986), no. 1, 177-194. [366] Wang, Xu-Jia; Zhu, Xiaohua. Kiihler-Rzccz solitons on toric manzfolds with positzve first Chern class. Adv. Math. 188 (2004), no. 1,87-103. [367] Watson, N. A. A theory of temperatures in several vanables. Jour Proc. London Math. Soc. 26 (1973), 385-417. [368] Wei, Guofang. Curvature formulas m Sectwn 6.1 of Perelman's paper (math. DC/0211159). http://www .math. ucsb.edu/ -wei/Perelman.html [369] Weibel, Charles A. An mtroductwn to homologzcal algebra. Cambridge Studies in Advanced Mathematics, 38. Cambridge University Press, Cambridge, 1994. [370] Weil, Andre. Introductwn Ii l'etude des variet€s kiihleriennes. (French) Publications de l'Institut de Mathematique de l'Universite de Nancago, VI. Actualites Sci. Ind., no. 1267. Hermann, Paris, 1958. [371] Wells, R. 0., Jr. Dzfferentzal analysis on complex manifolds. Second edition. Graduate Texts in Mathematics, 65. Springer-Verlag, New York-Berlin, 1980. [372] Wu, Lang-Fang. The Riccz flow on 2-orbzfolds wzth posztive curvature. J. Differential Geom. 33 (1991), no. 2, 575-596. [373] Wu, Lang-Fang. The Rzccz flow on complete ]R2. Comm. Anal. Geom. 1 (1993), no. 3-4, 439-472. [374] Yang, Deane. Convergence of Riemannian manifolds wzth integral bounds on curvature, I. Ann. Sci. Ecole Norm. Sup. (4) 25 (1992),77-105. [375] Yang, Deane. Convergence of Riemannian manifolds with mtegral bounds on curvature, II. Ann. Sci. Ecole Norm. Sup. (4) 25 (1992), 179-199. [376] Yau, Shing-Tung. On the curvature of compact Hermitian manifolds. Invent. Math. 25 (1974), 213-239. [377] Yau, Shing-Tung. Harmonic functwns on complete Riemannian manifolds. Comm. Pure Appl. Math. 28 (1975), 201-228. [378] Yau, Shing-Tung. Calabi's conjecture and some new results zn algebrazc geometry. Proc. Nat. Acad. Sci. U.S.A. 74 (1977), no. 5, 1798-1799. [379] Yau, Shing-Tung. On the Ricci curvature of a compact Kiihler manifold and the complex Monge-Ampere equatwn. I. Comm. Pure Appl. Math. 31 (1978), no. 3, 339-411. [380] Yau, Shing-Tung. On the Harnack inequalitzes of partzal dzfferentzal equations, Comm. Anal. Geom. 2 (1994), no. 3, 431-450. [381] Yau, Shing-Tung. Harnack mequality for non-self-adJomt evolution equations, Math. Res. Lett. 2 (1995), no. 4, 387-399. [382] Ye, Rugang. On the i-function and the reduced volume of Perelman. http://www.math.ucsb.edu;-yer/reduced.pdf [383] Zheng, Fangyang. Complex differential geometry. AMS/IP Studies in Advanced Mathematics, 18. American Mathematical Society, Providence, RI; International Press, Boston, MA, 2000. [384] Zhu, Shunhui. The comparison geometry of Ricci curvature. In Comparison Geometry, Grove and Petersen, eds. MSRI Publ. 30 (1997), 221-262.
BIBLIOGRAPHY
529
[385] Zhu, Xiaohua. Kiihler-Rzcci soliton typed equations on compact complex manifolds with Cl (M) > O. J. Geom. Anal. 10 (2000), no. 4, 759-774. [386] Zhu, Xi-Ping. Lectures on mean curvature flows. AMS/IP Studies in Advanced Mathematics, 32. American Mathematical Society, Providence, RI; International Press, Somerville, MA, 2002.
Index
adjoint heat equation, 201, 228 Alexander's Theorem, 433 almost complex manifold, 56 almost complex structure, 56 integrable, 56 ancient solution, 465 anti-holomorphic tangent bundle, 58 aperture, 477 approximate isometry, 150 arc length evolution, 457 variation formula, 448 Arzela-Ascoli Theorem, 137 associated 2-form, 61 asymptotic scalar curvature ratio, 472 is infinite on steady solitons, 472 asymptotic volume ratio, 253, 386, 451 of expanding soliton, 473 of Type I ancient solutions, 474 average nonlinear, 182 average scalar curvature, 6
breather solution, 41 Bryant soliton, 17, 440 Buscher duality transformation, 46 Calabi conjecture, 71 Calabi's trick, 453 Calabi-Yau metric, 72 canonical case, 73, 76 Cartan-Kahler Theorem, 10 center of mass, 182 Cheeger-Gromov convergence, 129 Cheeger-Gromov-Taylor, 159 Cheeger-Yau estimate, 453 Cheng's eigenvalue comparison, 269 Chevalley complex, 33 cigar soliton, 14, 102, 440, 468, 469 cigar-paraboloid behavior, 102 circumference at infinity, 477 Classic Dimension, xiv classical entropy, 124, 214 monotonicity, 215 Cohn-Vossen inequality, 17 collapsible manifold, 442 compactness of isometries, 155 compactness theorem for Kahler metrics, 139 for Kahler-Ricci flow, 140 for metrics, 130 for solutions, 131, 138 local version, 138 complex manifold, 56 complex projective space, 98 complex structure, 56 complex submanifold, 56 complexified cotangent bundle, 58 complexified tangent bundle, 58 concatenated path, 292 conjugate complex, 58
backward heat equation, 198 backward Ricci flow, 288 Bakry-Emery log Sobolev inequality, 218 Bianchi identity second,66 bisectional curvature, 65 nonnegative, 95, 96, 103, 118 Bishop volume comparison theorem, 450 Bishop-Gromov volume comparison, 450 boundary condition Dirichlet, 269 bounded geometry, 130 bounded variation locally, 367 531
Index
adjoint heat equation, 201, 228 Alexander's Theorem, 433 almost complex manifold, 56 almost complex structure, 56 integrable, 56 ancient solution, 465 anti-holomorphic tangent bundle, 58 aperture, 477 approximate isometry, 150 arc length evolution, 457 variation formula, 448 Arzela-Ascoli Theorem, 137 associated 2-form, 61 asymptotic scalar curvature ratio, 472 is infinite on steady solitons, 472 asymptotic volume ratio, 253, 386, 451 of expanding soliton, 473 of Type I ancient solutions, 474 average nonlinear, 182 average scalar curvature, 6
breather solution, 41 Bryant soliton, 17, 440 Buscher duality transformation, 46 Calabi conjecture, 71 Calabi's trick, 453 Calabi-Yau metric, 72 canonical case, 73, 76 Cart an-Kahler Theorem, 10 center of mass, 182 Cheeger-Gromov convergence, 129 Cheeger-Gromov-Taylor, 159 Cheeger-Yau estimate, 453 Cheng's eigenvalue comparison, 269 Chevalley complex, 33 cigar soliton, 14, 102, 440, 468, 469 cigar-paraboloid behavior, 102 circumference at infinity, 477 Classic Dimension, xiv classical entropy, 124, 214 monotonicity, 215 Cohn-Vossen inequality, 17 collapsible manifold, 442 compactness of isometries, 155 compactness theorem for Kahler metrics, 139 for Kahler-Ricci flow, 140 for metrics, 130 for solutions, 131, 138 local version, 138 complex manifold, 56 complex projective space, 98 complex structure, 56 complex submanifold, 56 complexified cotangent bundle, 58 complexified tangent bundle, 58 concatenated path, 292 conjugate complex, 58
backward heat equation, 198 backward Ricci flow, 288 Bakry-Emery log Sobolev inequality, 218 Bianchi identity second, 66 bisectional curvature, 65 nonnegative, 95, 96, 103, 118 Bishop volume comparison theorem, 450 Bishop-Gromov volume comparison, 450 boundary condition Dirichlet, 269 bounded geometry, 130 bounded variation locally, 367 531
532
conjugate heat equation, 201 connected sum decomposition, 433 convergence of maps C= on compact sets, 155 CP, 155 convergence on Il~?, 478 convex, 178 convex function, 367 coupled modified Ricci flow, 198 cross curvature breather, 499 cross curvature flow, 492 monotonicity formula, 496 short-time existence, 493 cross curvature soliton, 499 cross curvature tensor, 491
88- Lemma,
70 derivation of a Lie algebra, 33 derivative first, 366 second, 366 derivative of metric bounds, 132 diameter control, 268 differentiable function, 364 dilaton, 49, 191 dilaton shift, 49 dimension reduction, 440, 472 direct limit, 156 directed system, 156 Dirichlet boundary condition, 269 doubly-warped product, 26, 50 effective action, 49 Einstein metric, 5, 53 homogeneous, 32 standard type, 32 Einstein summation convention, 59 extended, 67 energy of a path, 286 energy functional, 44, 191 entropy, 124 entropy functional, 222 Euler-Lagrange equation for minimizer, 237 existence of minimizer, 238 entropy monotonicity for gradient flow, 226 for Ricci flow, 226 e-entropy, 229 essential 2-sphere, 433 eternal solution, 465
INDEX
Euclidean space characterization of, 451 exhaustion, 128 expanding breathers nonexistence, 213 expanding gradient soliton, 46 expanding soliton, 2 exploding soliton, 14 exponential map derivatives of, 175 exterior differential systems, 10 .1'-functional, 191 Fano manifold, 72 first Chern class, 64 first derivative, 366 first fundamental form, 482 first variation of .1',192 of W, 223 of We, 231 Frankel Conjecture, 73 fundamental solution of the heat equation, 453 funny way, 282 Futaki functional, 73, 98 Futaki invariant, 72 Gaussian soliton, 5 geometrically atoroidal, 436 geometrization conjecture, 437 197 gradient flow of gradient soliton structure, 4 graph space-time, 292 graph manifold Cheeger-Gromov, 439 topological, 438 Gromov-Hausdorff convergence, 38 Gross's logarithmic Sobolev inequality Beckner-Pearson's proof, 249
r,
Haken 3-manifold, 436 Hamilton's surface entropy formula, 217 gradient is matrix Harnack, 218 monotonicity, 217 harmonic map heat flow, 490 monotonicity formula, 490 Harnack inequality classical, 91 heat equation, 453 fundamental solution, 453 Hermitian metric, 57
INDEX
Hessian upper bound, 370 holomorphic coordinates, 55 holomorphic sectional curvature, 65 holomorphic tangent bundle, 58 homogeneous space, 32 Huisken's monotonicity formula, 487 Hamilton's generalization, 489 hyperbolic space, 269 immortal solution, 465 incompressible surface, 436 index form Riemannian, 345 Index Lemma Riemannian, 346 infinitesimal automorphism, 57, 97 injectivity radius estimate, 142, 159 local, 257 integral curves of \1£, 329 integration by parts for Lipschitz functions, 365 involutive system, 10 irreducible 3-manifold, 434 Jacobi field Riemannian, 346 Jacobian Riemannian, 359 Jensen's inequality, 235 Kahler class, 61 Kahler form, 61 Kahler identities, 61 Kahler manifold, 57 Kahler metric, 57 Kahler-Einstein metric, 72 Kahler-Ricci flow normalized, 76 K-collapsed at the scale r, 252 /'i- noncollapsed below the scale p, 252 Killing vector field, 11 Kneser Finiteness Theorem, 434 Koiso soliton, 44, 98 KRF,81 .c-cut locus, 356 L-distance, 293 and trace Harnack, 323 for Ricci flat solutions, 294 gradient of, 311 Hessian of, 318
533
is locally Lipschitz, 308 Laplacian Comparison Theorem, 321 time-derivative of, 312 triangle inequality, 328 .c-exponential map, 352 as T -+ 0,353 for Ricci flat solution, 353 .c-geodesic. 298 estimate for speed of , 303 existence for IVP, 305 on Einstein solution, 299 .c-geodesic equation, 298 .c-geodesics short ones are minimizing, 308 .c-Hopf-Rinow Theorem, 355 .c-index form. 359 .c-index lemma, 357 .c-Jacobi equation, 347 .c-Jacobi field, 346 time-derivative of, 351 .c-Jacobian, 360 of Ricci flat solution, 361 time-derivative of, 362 .c-Iength, 291 additivity, 292 first variation formula, 297 lower bound, 292 scaling property, 293 second variation of, 316 A-invariant, 204 lower bound, 206 monotonicity, 209 second variation of, 280 upper bound, 206 Laplacian for a Kahler metric, 67 Hodge-de Rham, 449 rough,449 Laplacian comparison Riemannian, 376 Laplacian Comparison Theorem, 452 for L-distance, 321 L-distance as T -+ 0,324 monotonicity of minimum, 324 supersolution to heat equation, 323 Levi-Civita connection, 446 evolution, 456 Kahler evolution, 77 of a Kahler metric, 62 variation formula, 448 Li-Yau inequality, 453 Lichnerowicz Laplacian, 448
534
for a Kahler metric, 70, 106, 118 Lie algebra cohomology, 33 Lie algebra square, 460 Lie derivative, 449 linear trace Harnack and We, 233 linearized Kahler-Ricci flow, 118 Lipschitz graph, 365 Little Loop Conjecture, 256, 257 locally bounded variation, 367 locally collapsing, 255 locally homogeneous geometries, 438 locally Lipschitz, 364 vector field, 365 logarithmic Sobolev inequality, 246 of L. Gross, 247 mapping torus, 37, 38 matrix Harnack formula for adjoint heat equation, 281 matrix Harnack quadratic, 9 for Kahler-Ricci flow, 109 maximal function, 265 maximum principle strong tensor, 107, 471 weak scalar, 458 weak tensor, 459 maximum volume growth, 96, 253, 507 mean curvature, 482 mean curvature flow, 482 convergence to self-similar, 489 evolution equations under, 483 measure, 196 minimal 'c-geodesic, 293 existence, 305 minimizer of :F Euler-Lagrange equation for, 206 existence of, 205 modified Ricci curvature, 194 modified Ricci flow, 197 modified Ricci tensor, 48 modified scalar curvature, 48, 194 evolution of, 203 first variation, 273 modified subsequence convergence, 477 mollifier, 367 Monge-Ampere equation complex, 75, 76 Monge-Ampere equation complex, 71 monotonicity of We, 233 monotonicity formula
INDEX
for static metric reduced volume, 382 for the gradient flow, 197 JL-invariant, 236 as T -+ 0,244 as T -+ 00, 243 finiteness, 237 monotonicity, 239 under Cheeger-Gromov convergence, 240 necklike point, 470 neckpinch, 439 Newlander-Nirenberg Theorem, 56 Nijenhuis tensor, 56 nilpotent Lie group, 34 NKRF,81 no local collapsing, 256, 440 improved version, 267 proof of, 258 normal holomorphic coordinates, 68 normal surfaces, 435 II-invariant, 236 monotonicity, 244 null eigenvector assumption, 459 open problem, 282 orbifold, 12, 141 bad, 12 parabola, 287 parallel 2-spheres, 434 Perelman's energy functional, 44 Perelman's entropy functional, 45, 222 Perelman's equations coupled to Ricci flow, 225 Perelman's Harnack quantity, 227 evolution, 227 pinching improves, 463 PL category, 433 Poincare conjecture, 437 pointed Riemannian manifold, 129 pointed solution to the Ricci flow, 129 pointwise convergence in CP , 128 uniformly on compact sets, 129 pointwise monotonicity along 'c-geodesics, 391 porous medium equation, 479 potential function, 76 potentially infinite dimensions and conformal geometry, 194 quasi-Einstein metric, 52
INDEX
Rademacher's Theorem, 364 real (p, p )-form, 59 real part of complex vector, 58 reduced distance, 326 for Einstein solution, 336 of a static manifold, 384 on shrinker, 343 partial differential inequalities, 327 regularity properties, 377 under Cheeger-Gromov convergence, 334 reduced volume for Ricci flow, 388 static, 381 reduced volume monotonicity heuristic proof, 389 proof of, 392, 395 Ricci breather, 203 Ricci curvature quasi-positive, 103 Ricci flow geometry, 391 Ricci flow on surfaces revisited, 145 Ricci form, 64 Ricci soliton, 2, 464 canonical form, 3 Ricci soliton structure, 4 Ricci tensor, 446 evolution, 457 of a Kahler metric, 64 variation formula, 448 Riemann curvature tensor variation formula, 448 Riemann tensor, 446 evolution, 456, 460 evolution for Kahler metric, 104 of a Kahler metric, 63 Riemannian distance function convexity of, 178 derivatives of, 175 Rosenau solution, 469 scalar curvature, 446 evolution, 457 Kahler evolution, 77 of a Kahler metric, 65 variation formula, 448 Schoenflies problem, 433 second derivative, 366 second fundamental form, 482 Seifert 3-manifold, 436 self-similar solution, 2
535
semisimple Lie group, 34 separating 2-sphere, 433 short-time existence compact manifolds, 457 noncompact manifolds, 458 shrinking breathers are gradient solitons, 242 shrinking gradient soliton, 45 shrinking soliton, 2 Simons' identity, 486 singularity model, 439 existence of, 143, 263 smooth category, 433 soliton Kahler-Ricci, 96, 97 space-time metric, 387 static reduced volume, 381 steady breathers nonexistence, 210 steady gradient soliton, 44 steady soliton, 2 Sterling's formula, 250 strongly K-collapsed, 401 supersolution, 374 support sense, 373 surface entropy, 468 surgery Ricci flow with surgery, 441 T-duality, 46 tensor of type (p, q), 59 3-manifolds with positive Ricci curvature revisited, 143 topological category, 433 torus decomposition, 436 trace Harnack quadratic, 467 for Kahler-Ricci flow, 109 Type I essential point, 470 Type I singularity, 489 Type III solution, 38 Dhlenbeck's trick, 459 Dniformization Theorem, 12,438 unitary frame, 65 vicosity sense, 374 volume doubling property, 266 volume form, 446 evolution, 457 Kahler evolution, 77 of a Kahler metric, 66 variation formula, 448 volume ratio, 382
536
W-functional, 222 warped product, 12, 46, 480 weak sense, 373 weakened no local collapsing, 401 Witten's black hole, 468 Yau's unifomization conjecture, 96
INDEX