SOME PROBLEMS ON MANIFOLDS WITH LOWER BOUNDS ON RICCI CURVATURE By Zhixin Wang A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of Mathematics-Doctor of Philosophy 2024 ABSTRACT In this work, we delve into geometric analysis, particularly examining the interplay between lower bounds on Ricci curvature and specific functionals. Our exploration begins with an investigation into the implications of Yamabe invariants for asymptotically PoincarΓ©-Einstein manifolds and their conformal boundaries under conditions of 𝑅𝑖𝑐 β‰₯ βˆ’(𝑛 βˆ’ 1)𝑔. We establish a relationship wherein the type II Yamabe invariant of the conformal compactification of the manifold is bounded below by the Yamabe invariant of its conformal boundary. Additionally, we focus on compact manifolds with boundary where 𝑅𝑖𝑐 β‰₯ 0 and 𝐼 𝐼 β‰₯ 1, obtaining partial results concerning Wang’s conjecture. Copyright by ZHIXIN WANG 2024 ACKNOWLEDGEMENTS This thesis represents not only my work at the keyboard but also a milestone in my life, made possible with the support of several individuals. First and foremost, I would like to express my deepest gratitude to my advisor, Xiaodong Wang, whose expertise, understanding, and patience, added considerably to my graduate experience. I appreciate his vast knowledge and skill in differential geometry. Without his guidance, this thesis would not have been possible. I would also like to thank my committee members, Kitagawa Jun, Thomas Parker, and Willie Wong, for their insightful comments and suggestions. Their feedback significantly improved the quality of this work. I am deeply grateful to my family for their love, patience, and unwavering support. I would also like to extend my heartfelt thanks to roommate Yuhan Jiang and friends from the group SMDJ8, whose camaraderie and support have been instrumental throughout this journey. Their understanding, laughter, and encouragement kept me grounded and motivated during the most challenging times. Your friendship means the world to me, and I am truly grateful to have each of you in my life. This journey would not have been possible without all of you. Thank you. iv TABLE OF CONTENTS CHAPTER 1 YAMABE INVARIANTS FOR ASMPTOTICALLY POINCARΓ‰-EINSTEIN MANIFOLDS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1 CHAPTER 2 LIOUVILLE TYPE THEOREMS ON MANIFOLDS WITH LOWER CURVATURE BOUND . . . . . . . . . . . . . . . . . . . . . . . . . . 27 CHAPTER 3 A LOG-SOBOLEV INEQUALITY . . . . . . . . . . . . . . . . . . . . 58 BIBLIOGRAPHY . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73 v CHAPTER 1 YAMABE INVARIANTS FOR ASMPTOTICALLY POINCARΓ‰-EINSTEIN MANIFOLDS Roughly speaking, a PoincarΓ©-Einstein manifold is a non-compact manifold characterized by nega- tive constant Ricci curvature and the admission of a conformal compactification. The investigation of PoincarΓ©-Einstein manifolds is underpinned by a fundamental principle: the intricate interplay between the manifold’s boundary and its interior. Given that we employ conformal transformations in defining PoincarΓ©-Einstein manifolds, a natural inquiry arises concerning the existence of con- formal invariants that exemplify this principle. Such inequality was introduced in [CLW17] with certain restrictions. Through collaborative efforts with X. Wang, we successfully eliminated these constraints, resulting in a comprehensive and unrestricted conclusion [WW21], [WW22]. This chapter will delve into the examination of these inequalities. 1.1 Asymptotically PoincarΓ©-Einstein manifold PoincarΓ©-Einstein manifolds, which serve as the foundation for the AdS/CFT correspondence framework , have been the subject of extensive research over the past three decades, yielding significant advances in both mathematics and physics (see [Biq05], for instance). The concept of the PoincarΓ©-Einstein manifold emerges from an observation rooted in hyper- bolic space (H𝑛, 𝑔𝐻). Utilizing the conformal ball model, this space can be effectively represented (1βˆ’|π‘₯|2)2 𝑑π‘₯2), wherein 𝑑π‘₯2 denotes the Euclidean metric. Through the application of the , (H𝑛, 𝑔𝐻) can be conformally compactified to the unit disk within Eu- as (B𝑛, conformal factor (1βˆ’|π‘₯|2)2 4 4 clidean space. The boundary of this compactified space is commonly termed the "boundary at infinity" or the "conformal boundary." By summarizing this distinctive property in conjunction with the Ricci curvature equation 𝑅𝑖𝑐𝑔𝐻 = βˆ’(𝑛 βˆ’ 1)𝑔𝐻, we arrive at the comprehensive definition of PoincarΓ©-Einstein manifolds: Definition 1.1.1. 𝑋 is the interior of a compact manifold ¯𝑋 with boundary 𝑀. (𝑋, 𝑔+) is called a 𝐢3,𝛼 PoincarΓ©-Einstein manifold if 𝑔+ is a noncompact complete metric, 𝑅𝑖𝑐𝑔+ = βˆ’(𝑛 βˆ’ 1)𝑔+ (1.1.1) 1 and 𝑔 = 𝜌2𝑔+ can be 𝐢3,𝛼 extended to ¯𝑋 by a boundary defining function 𝜌, i.e. 𝜌 ∈ 𝐢∞( ¯𝑋), 𝜌 > 0 in 𝑋, 𝜌 = 0 and d𝜌 β‰  0 on πœ• 𝑋. πœ• 𝑋, together with the conformal class [𝜌2𝑔(cid:12) If the Ricci curvature equation 1.1.1 is replaced by 𝑅𝑖𝑐𝑔+ = βˆ’(𝑛 βˆ’ 1)𝑔+ + π‘œ(𝜌2), we arrive at the (cid:12)πœ• 𝑋], is called conformal infinity. definition for asymptotically PoincarΓ©-Einstein manifolds. Apart from hyperbolic space (B𝑛, 𝑔H), which serves as the prototype, PoincarΓ©-Einstein mani- folds also come in different ways. Example 1.1.1. Perturbation from (B𝑛, 𝑔H) Let β„Ž be the standard round metric on Sπ‘›βˆ’1. The work by J.Lee and C.Graham showed that if we perturb the metric on Sπ‘›βˆ’1 slightly to β„Žβ€², then there exists a corresponding 𝑔′ satisfying (1.1.1) and (Sπ‘›βˆ’1, β„Žβ€²) as its conformal boundary. [GL91] Example 1.1.2. Let (𝑁 π‘›βˆ’1, 𝑔𝑁 ) be a compact manifold without boundary, and 𝑅𝑖𝑐𝑁 = βˆ’(𝑛 βˆ’ 2)𝑔𝑁 , then is a PoincarΓ©-Einstein manifold with compactification [0, 1] Γ— 𝑁. (R Γ— 𝑁, 𝑑𝑑2 + cosh2(𝑑)𝑔𝑁 ) Note that the conformal boundary is 𝑁 Γ— {Β±1}. It has negative scalar curvature and is not connected. We will revisit this example later, as it serves to illustrate how the conformal boundary significantly influences the geometry of the entire manifold. Given an asymptotically PoincarΓ©-Einstein manifold, we want to study its geometry near con- formal boundary. We start with the following result in [Lee94]. Theorem 1.1. Let (𝑋, 𝑔+) be asymptotically PoincarΓ©-Einstein manifold with (𝑀, β„Ž) as its confor- mal boundary. For any β„Žβ€² ∈ [β„Ž], there exists a boundary defining function 𝜌 so that near conformal boundary 𝑔 takes the form 𝑔 = 1 𝜌2 (π‘‘πœŒ2 βŠ• β„ŽπœŒ) (1.1.2) where β„Ž0 = β„Žβ€². In particular, |π‘‘πœŒ| 𝜌2𝑔+ = 1. 2 This is called Graham-Lee normal form. 𝜌 is a distance function for ¯𝑔, and its curvature can be computed using Riccati equation, Gauss-Codazzi equation and Codazzi-Mainardi equations. Pick local coordinates {π‘₯𝑖} for πœ• 𝑋 = 𝑀, and {π‘₯0 = 𝜌, π‘₯𝑖} form local coordinates for 𝑋 near conformal boundary. Apply (1.2.4), the traceless-Ricci curvature 𝐸 = 𝑅𝑖𝑐𝑔+ βˆ’ as 𝑅𝑔+ 𝑛 𝑔+ are given in [BMW13] 2πœŒπΈπ‘– 𝑗 = βˆ’πœŒβ„Žβ€²β€² 𝑖 𝑗 + πœŒβ„Ž π‘π‘ž β„Žβ€² 𝑖 𝑝 β„Žβ€² 𝑗 π‘ž βˆ’ 𝜌 2 β„Ž π‘π‘ž β„Žβ€² π‘π‘ž β„Žβ€² 𝑖 𝑗 + (𝑛 βˆ’ 2)β„Žβ€² 𝑖 𝑗 + β„Ž π‘π‘ž β„Žβ€² π‘π‘ž β„Žπ‘– 𝑗 + 2πœŒπ‘…π‘–π‘(β„ŽπœŒ)𝑖 𝑗 𝐸𝑖0 = 1 2 𝐸00 = βˆ’ π‘π‘ž) β„Ž π‘π‘ž (βˆ‡π‘ž β„Žβ€² 1 2 β„Ž π‘π‘ž β„Žβ€²β€² 𝑖 𝑝 βˆ’ βˆ‡π‘– β„Žβ€² 1 4 π‘π‘ž + β„Ž π‘π‘ž β„Žπ‘˜π‘™ β„Žβ€² π‘π‘˜ β„Žβ€² π‘žπ‘™ + 1 2𝜌 β„Ž π‘π‘ž β„Žβ€² π‘π‘ž (1.1.3) where β€² denotes πœ• πœ• 𝜌 . Set 𝜌 = 0 in the first equation, and we get (𝑛 βˆ’ 2)β„Žβ€² + π‘‘π‘Ÿβ„Ž (β„Žβ€²)β„Ž = 0 This implies β„Žβ€² = 0, and therefore 𝑀 is totally geodesic in ( ¯𝑋, ¯𝑔). In particular 𝑀 is umbilical in ( ¯𝑋, Λœπ‘”) for any conformal compactification Λœπ‘” since the property of umbilicus is invariant under conformal change. Take derivative πœ• πœ• 𝜌 π‘˜ times to the first equation of (1.1.3), and we get (𝑛 βˆ’ 1 βˆ’ π‘˜)πœ• π‘˜ 𝜌 β„Ž + π‘‘π‘Ÿβ„Ž (πœ• π‘˜ 𝜌 β„Ž)β„Ž = πœ• π‘˜βˆ’1 𝜌 (2𝜌𝐸)𝜌=0 + (terms containning πœ•π‘™ 𝜌 with 𝑙 < π‘˜) Now suppose 𝐸 ≑ 0, i.e. (𝑀, 𝑔+) is PoincarΓ©-Einstein. For π‘˜ < 𝑛 βˆ’ 1, the coefficients for πœ• π‘˜ 𝜌 β„Ž is non-zero. By induction, we could solve for πœ• π‘˜ 𝜌 β„Ž and thus get expansion for ¯𝑔 near conformal boundary up to order 𝑛 βˆ’ 2 if 𝑛 is even and 𝑛 βˆ’ 1 if 𝑛 is odd. For example: i)πœ• π‘˜ 𝜌 β„Ž = 0 for π‘˜ odd and π‘˜ < 𝑛 βˆ’ 1; ii) if 𝑛 is even, then π‘‘π‘Ÿβ„Žπœ•π‘›βˆ’1 𝜌 β„Ž = 0 and πœ•π‘›βˆ’1 𝜌 β„Ž is not determined. (See Proposition 2.7 in [Woo16], for example. The statement there is only for 𝑛 odd, but the argument works also for even 𝑛’s for orders below 𝑛 βˆ’ 1). In particular, we can find the second order term (cid:16) βˆ’ 2 π‘›βˆ’3 𝑅𝑖𝑐 (β„Ž) βˆ’ π‘…β„Ž 2(π‘›βˆ’2) (cid:17) β„Ž , βˆ’ 1 2 β„Ž, if 𝑛 β‰₯ 4; if 𝑛 = 3. β„Žβ€²β€² =   ο£³ (1.1.4) 1.2 Conformal Invariants In this section I will introduce basic formulas under conformal change and then introduce Yamabe conformal invariants. 3 The Yamabe problem can be thought of as a continuation of uniformization theorem. For 2-dimensional spaces, all Riemannian surfaces are locally confomally Euclidean, and we have the uniformization theorem Theorem 1.2. Simply connected Riemann surface is biholomorphic to one of the following: β€’ Β―C = S2 β€’ C β€’ {𝑧 ∈ C : |𝑧| < 1} As a result, all compact Riemannian surfaces admit a conformal metric of constant Gaussian curvature. In dim>3, Weyl tensor is conformal invariant, thus obstruction for being locally conformally flat. But we could still ask the following: can we find a metric of constant scalar curvature within each conformal metric class. This is what Yamabe problem is about. Given a Riemannian manifold (𝑀 𝑛, 𝑔) with 𝑛 β‰₯ 3 and local coordinates {π‘₯𝑖}. Under conformal change ¯𝑔 = 𝑒2𝑔, the new Levi-Civita connection can be calculated by Β―βˆ‡π‘‹π‘Œ = βˆ‡π‘‹π‘Œ + 𝑋𝑒 𝑒 π‘Œ + π‘Œ 𝑒 𝑒 𝑋 βˆ’ 𝑔(𝑋, π‘Œ ) 𝑒 βˆ‡π‘’ (1.2.1) The conformal change of Hession can be computed as ¯𝑔 (cid:0) Β―βˆ‡π‘‹ Β―βˆ‡ 𝑓 , π‘Œ (cid:1) Β―βˆ‡2 𝑓 (𝑋, π‘Œ ) = ¯𝑔 (cid:0) Β―βˆ‡π‘‹ 1 𝑒2 βˆ‡ 𝑓 , π‘Œ (cid:1) 2𝑋𝑒 𝑒3 βˆ‡ 𝑓 + 2𝑋𝑒 𝑒3 βˆ‡ 𝑓 + 1 𝑒 = βˆ‡2 𝑓 (𝑋, π‘Œ ) βˆ’ = ¯𝑔 (cid:0) βˆ’ = ¯𝑔 βˆ’ (cid:16) Β―βˆ‡π‘‹ βˆ‡ 𝑓 , π‘Œ (cid:1) 1 𝑒2 1 𝑒2 (βˆ‡π‘‹ βˆ‡ 𝑓 + (𝑋𝑒 Β· π‘Œ 𝑓 + 𝑋 𝑓 Β· π‘Œ 𝑒) + 𝑋𝑒 𝑒 βˆ‡ 𝑓 + (βˆ‡ 𝑓 )𝑒 𝑒 𝑋 βˆ’ 𝑔(βˆ‡π‘’, βˆ‡ 𝑓 ) 𝑒 𝑋 𝑓 𝑒 (cid:17) βˆ‡π‘’), π‘Œ 𝑔(𝑋, π‘Œ ) 4 which is Β―βˆ‡2 𝑓 (𝑋, π‘Œ ) = βˆ‡2 𝑓 βˆ’ 1 𝑒 (𝑑𝑒 βŠ— 𝑑𝑓 + 𝑑𝑓 βŠ— 𝑑𝑒) + 𝑔(βˆ‡π‘’, βˆ‡ 𝑓 ) 𝑒 𝑔 (1.2.2) Using the formula above, the Riemannian curvature can be computed as ¯𝑅𝑖 𝑗 π‘˜π‘™ = 𝑒2(𝑅𝑖 𝑗 π‘˜π‘™ βˆ’ π‘”π‘–π‘˜π‘‡π‘—π‘™ + 𝑔 π‘—π‘™π‘‡π‘–π‘˜ βˆ’ 𝑔𝑖𝑙𝑇𝑗 π‘˜ βˆ’ 𝑔 𝑗 π‘˜π‘‡π‘–π‘™) where 𝑇𝑖 𝑗 = βˆ‡π‘–βˆ‡ 𝑗 𝑒 𝑒 βˆ’ 2 βˆ‡π‘–π‘’βˆ‡ 𝑗 𝑒 𝑒2 + |𝑑𝑒|2 2𝑒2 𝑔𝑖 𝑗 Taking trace yields the formula for Ricci curvature and scalar curvature ¯𝑅𝑖 𝑗 = 𝑅𝑖 𝑗 βˆ’ (𝑛 βˆ’ 2)( ¯𝑅 = (cid:16) 1 𝑒2 𝑅 βˆ’ 2(𝑛 βˆ’ 1) βˆ‡π‘–βˆ‡ 𝑗 𝑒 𝑒 Δ𝑒 𝑒 βˆ’ 2 βˆ‡π‘–π‘’βˆ‡ 𝑗 𝑒 𝑒2 ) βˆ’ ( βˆ’ (𝑛 βˆ’ 4) (𝑛 βˆ’ 1) + (𝑛 βˆ’ 3) Δ𝑒 𝑒 |𝑑𝑒|2 𝑒2 (cid:17) |𝑑𝑒|2 𝑒2 )𝑔𝑖 𝑗 For scalar curvature we usually take the form ¯𝑔 = 𝑒 4 π‘›βˆ’2 𝑔, and it takes the form ¯𝑅 = π‘’βˆ’ 𝑛+2 π‘›βˆ’2 (βˆ’ 4(𝑛 βˆ’ 1) 𝑛 βˆ’ 2 Δ𝑒 + 𝑅𝑒) (1.2.3) (1.2.4) Remark 1.2.1. For 𝑛 = 2, we use the conformal change ¯𝑔 = 𝑒2πœ™π‘”. The scalar curvature transforms by ¯𝑅 = π‘’βˆ’2πœ™ (βˆ’2Δ𝑒 + 𝑅) (1.2.5) The operator 𝐿𝑔 (𝑒) := βˆ’ conformal invariance. Let ¯𝑔 = 𝑒 4 4(π‘›βˆ’1) π‘›βˆ’2 Δ𝑒 + 𝑅𝑒 is called conformal Laplacian. It has the following π‘›βˆ’2 𝑔. Suppose there is a third conformal metric 𝑔′ = 𝑣 4 π‘›βˆ’2 𝑔 = ( 𝑣 𝑒 ) 4 π‘›βˆ’2 ¯𝑔. Then by (1.2.4) 𝑅′ = π‘£βˆ’ 𝑛+2 π‘›βˆ’2 𝐿𝑔 (𝑣) = ( 𝑣 𝑒 β‡’ 𝐿𝑔 (𝑣) = 𝑒 𝑛+2 π‘›βˆ’2 𝐿 ¯𝑔 ( π‘›βˆ’2 𝐿 ¯𝑔 ( )βˆ’ 𝑛+2 𝑣 𝑒 ) 𝑣 𝑒 ) (1.2.6) Suppose 𝑋 𝑛 has a boundary Ξ£π‘›βˆ’1 and let 𝜈 be the outer normal vector. Under conformal change ¯𝑔 = 𝑒 4 π‘›βˆ’2 𝑔, the new normal vector becomes ¯𝜈 = π‘’βˆ’ 2 π‘›βˆ’2 𝜈. Using (1.2.1), the second fundamental 5 form 𝐼 𝐼 and mean curvature changes by ¯𝐼 𝐼 (𝑋, π‘Œ ) = 𝑒 2 π‘›βˆ’2 (cid:2)𝐼 𝐼 (𝑋, π‘Œ )) + ¯𝐻 = π‘’βˆ’ 2 π‘›βˆ’2 (𝐻 + Now we can define Yamabe invariants. πœ•π‘’ πœ•πœˆ 𝑔(𝑋, π‘Œ )(cid:3) (1.2.7) 2 (𝑛 βˆ’ 2)𝑒 πœ•π‘’ πœ•πœˆ ) 2(𝑛 βˆ’ 1) (𝑛 βˆ’ 2)𝑒 Definition 1.2.1. Suppose (𝑀 𝑛, 𝑔) is a Riemannian manifold without boundary, the Yamabe in- variant is defined to be where π‘Œ (𝑀, [𝑔]) = inf π‘’βˆˆπ»1 (𝑀),𝑒≠0 𝐸𝑔 (𝑒) 𝑒 2𝑛 π‘›βˆ’2 dV) π‘›βˆ’2 𝑛 (∫ 𝑀 𝐸𝑔 (𝑒) = ∫ 𝑀 4(𝑛 βˆ’ 1) 𝑛 βˆ’ 2 |βˆ‡π‘€π‘’|2 + 𝑅𝑀𝑒2dV (1.2.8) Remark 1.2.2. Pick a sequence of functions which blows up locally and it can be shown that π‘Œ (𝑀, [𝑔]) ≀ π‘Œ (S𝑛, π‘‘πœƒ2) (1.2.9) where π‘‘πœƒ2 is the round metric in S𝑛. See [Aub76]. If we write 𝑔′ = 𝑒 4 π‘›βˆ’2 𝑔 for 𝑒 > 0, the integral in (1.2.8) can be rewritten as ∫ 𝑀 4(π‘›βˆ’1) π‘›βˆ’2 |βˆ‡π‘”π‘’|2 + 𝑅𝑔𝑒2dV (∫ π‘›βˆ’2 dV) 𝑀 𝑒 2𝑛 π‘›βˆ’2 𝑛 = = ∫ 𝑀 4(π‘›βˆ’1) π‘›βˆ’2 Δ𝑔𝑒 + 𝑅𝑔𝑒)dV 𝑒(βˆ’ (∫ 𝑒 2𝑛 𝑀 𝑅𝑔′dV𝑔′ π‘›βˆ’2 dV) π‘›βˆ’2 𝑛 ∫ 𝑀 (Vol(𝑀, 𝑔′)) π‘›βˆ’2 𝑛 (1.2.10) For general 𝑒 ∈ 𝐻1(𝑀), we can take |𝑒| and approximate it with positive functions in 𝐻1. An equivalent definition for Yamabe invariant is thus derived π‘Œ (𝑀, [𝑔]) = inf π‘”β€²βˆˆ[𝑔] ∫ 𝑀 𝑅𝑔′dV𝑔′ (Vol(𝑀, 𝑔′)) π‘›βˆ’2 𝑛 where [𝑔] represents the conformal class of 𝑔. Derived from this definition, it becomes evident that these two quantities remain invariant under conformal transformations, underscoring their pivotal role in the realm of conformal geometry research. 6 The Euler-Lagrangian equation for (1.2.8) is given by 𝐿𝑔 (𝑒) = πœ†π‘’ 𝑛+2 π‘›βˆ’2 (1.2.11) In conjunction with (1.2.4), the minimizer obtained from this equation provides a metric with constant scalar curvature. Consequently, the existence of the minimizer resolves the problem introduced at the beginning of this section. However, it is worth noting that in (1.2.8), we employ the 𝐿 2𝑛 π‘›βˆ’2 (𝑋) norm in the denominator, and 2𝑛 π‘›βˆ’2 represents the critical power for Sobolev embedding. While boundedness is assured, compactness is not guaranteed. To address this challenge, we employ the β€œlowering index" technique. We define a new functional as π‘Œπ‘ž (𝑀, [𝑔]) = inf π‘’βˆˆπ»1 (𝑀),𝑒≠0 𝐸𝑔 (𝑒) (cid:17) 2 𝑝 𝑒 𝑝d𝑉 (cid:16)∫ 𝑀 (1.2.12) where 𝑝 < 2𝑛 π‘›βˆ’2. These values of 𝑝 are strictly below the critical Sobolev conjugate. Using the standard argument, the existence of minimizers 𝑒 𝑝 follows from the compactness of the inclusion 𝐻1(𝑀) βŠ‚ 𝐿 𝑝 (𝑀), and these 𝑒 𝑝’s satisfy the Euler-Lagrangian 𝐿𝑔 (𝑒) = πœ†π‘’ π‘βˆ’1 (1.2.13) If we further impose the condition |𝑒| 𝑝 = 1, then πœ† 𝑝 = π‘Œπ‘ž (𝑀, [𝑔]) in (1.2.12). Similar for (1.2.8). Trudinger[Tru68] and Aubin[Aub76] demonstrated that βˆ₯𝑒 𝑝 βˆ₯πΏπ‘Ÿ is uniformly bounded for some π‘Ÿ > 2𝑛 π‘›βˆ’2 provided the inequality in (1.2.9) is strict. Consequently, 𝑒 𝑝 converges to a smooth solution 𝑒 of (1.2.11), and 𝑒 is a minimizer for (1.2.11). Thus, the primary challenge is reduced to establishing the strict inequality in (1.2.9), except for standard spheres. This problem was ultimately resolved by R. Schoen, who utilized the positive mass theorem to construct an appropriate test function. Combining all the elements above, we arrive at the following theorem: Theorem 1.3. Let (𝑀, 𝑔) be a compact manifold without boundary. Thenπ‘Œ (𝑀, [𝑔]) ≀ π‘Œ (S𝑛, [π‘‘πœƒ2]) with inequality iff round metric on S. As a result, there exists a metric 𝑔′ ∈ [𝑔] such that 𝑅𝑔′ is constant. 7 For a comprehensive exploration of this problem, refer to [LP87] or Chapter 5 of [SY94]. For manifolds with boundary (𝑀, Ξ£, 𝑔), we can ask the following two questions. Fix a conformal class [𝑔], can we find 𝑔′ ∈ [𝑔] so that: I) 𝑅𝑔′ = constant, 𝐻𝑔′ = 0; or II) 𝑅𝑔′ = 0, 𝐻𝑔′ = constant. These two are called Type I and Type II Yamabe problem respectively. As in Yamabe problem, we can define the following two functional Definition 1.2.2. π‘Œ (𝑋, 𝑀, [𝑔]) = 𝑄(𝑋, 𝑀, [𝑔]) = inf π‘’βˆˆπ»1,𝑒≠0 inf π‘’βˆˆπ»1,𝑒≠0 (∫ 𝑋 (∫ 𝑀 𝐸 (𝑒) 𝑒2𝑛/(π‘›βˆ’2)) (π‘›βˆ’2)/𝑛 𝐸 (𝑒) Type I 𝑒2(π‘›βˆ’1)/(π‘›βˆ’2)) (π‘›βˆ’2)/(π‘›βˆ’1) Type II (1.2.14) where 𝐸 (𝑒) = ∫ 𝑋 4(𝑛 βˆ’ 1) 𝑛 βˆ’ 2 |βˆ‡π‘’|2 + 𝑅𝑒2dV + 2 ∫ 𝑀 𝐻𝑒2dS As before set 𝑔′ = 𝑒 4 π‘›βˆ’2 𝑔 for 𝑒 > 0, then 4(𝑛 βˆ’ 1) 𝑛 βˆ’ 2 |βˆ‡π‘’|2 + 𝑅𝑒2dV + 2 𝑒(βˆ’ 4(𝑛 βˆ’ 1) 𝑛 βˆ’ 2 𝑅𝑔′dV𝑔′ + 2 Δ𝑒 + 𝑅𝑒)dV + 2 ∫ Ξ£ 𝐻𝑔′dS𝑔′ ∫ 𝑋 ∫ 𝑋 ∫ 𝑀 = = ∫ 𝐻𝑒2dS 𝑀 ∫ 𝑀 𝑒2(𝐻 + 2(𝑛 βˆ’ 1) 𝑒(𝑛 βˆ’ 2) πœ•π‘’ πœ•πœˆ dS And (1.2.14) can be rewritten as π‘Œ (𝑀, Ξ£, [𝑔]) = inf π‘”β€²βˆˆ[𝑔] 𝑄(𝑀, Ξ£, [𝑔]) = inf π‘”β€²βˆˆ[𝑔] ∫ 𝑅𝑔′ + 2 ∫ 𝐻𝑔′ 𝑀 Ξ£ Vol(𝑀, 𝑔′) (π‘›βˆ’2)/𝑛 Type I ∫ 𝑀 𝑅𝑔′ + 2 ∫ Ξ£ Area(Ξ£, 𝑔′|Ξ£) (π‘›βˆ’2)/(π‘›βˆ’1) 𝐻𝑔′ Type II So these two minimum are conformal invariants. The corresponding Euler-Lagrangian equations are computed to be 𝑇 𝑦 𝑝𝑒 𝐼 : 𝑇 𝑦 𝑝𝑒 𝐼 𝐼 :   ο£³ 𝐿𝑔 (𝑒) = πœ†π‘’ 𝑛+2 π‘›βˆ’2 πœ•π‘’   ο£³ 𝐿𝑔 (𝑒) = 0 πœ•πœˆ + π‘›βˆ’2 2(π‘›βˆ’1) 𝐻𝑒 = 0 πœ•π‘’ πœ•πœˆ + π‘›βˆ’2 2(π‘›βˆ’1) 𝐻𝑒 = πœ†π‘’ 𝑛 π‘›βˆ’2 8 (1.2.15) (1.2.16) So the minimizer of Type I and Type II Yamabe invariants solves the corresponding Yamabe problems respectively by (1.2.4) and (1.2.7). Again, by picking suitable test functions we have 𝑄(𝑀, Ξ£, [𝑔]) ≀ 𝑄(B𝑛, Sπ‘›βˆ’1, [𝑑π‘₯2]) π‘Œ (𝑀, Ξ£, [𝑔]) ≀ ¯𝑄(S𝑛 +, Sπ‘›βˆ’1, [𝑑𝑠2]) (1.2.17) And strict inequality implies the existence of minimizers by β€œlowering index" method. These problems are only partially solved. For Type I Yamabe problem, the strict inequality was verified in the following cases [Esc92b] β€’ 𝑛 = 3, 4, 5; β€’ 𝑛 β‰₯ 6 and πœ• 𝑀 = Ξ£ is not umbilic. For Type II, Escobar verified the following in [Esc92a] β€’ 𝑛 > 6 and 𝑋 has a nonumbilic boundary point; β€’ 𝑛 β‰₯ 6, with 𝑋 locally flat and πœ• 𝑋 unbilic; β€’ 𝑛 = 4, 5 and πœ• 𝑋 is umbilic; β€’ 𝑛 = 3. A substantial amount of work has been dedicated to addressing these two problems; nevertheless, some cases still remain open. See, for instance, [Alm12], [Che09], [BC09], [Mar05], [Mar07], and others. Recall that PoincarΓ©-Einstein manifolds have umbilical boundaries. Apply these results and direct arguments give us (see [CLW17]) Theorem 1.4. Let 𝑋 𝑛, 𝑔+ be 𝐢3,𝛼 PoincarΓ©-Einstein manifold satisfying one of the following β€’ 3 ≀ 𝑛 ≀ 5 β€’ 𝑛 β‰₯ 6 and 𝑋 is spin 9 Then there exists a conformal compactification ¯𝑔 = 𝜌2𝑔+ which is a minimizer for π‘Œ ( ¯𝑋, 𝑀, [ ¯𝑔]). Furthermore, ¯𝑔 has constant scalar curvature and totally geodesic curvature. Theorem 1.5. Let 𝑋 𝑛, 𝑔+ be 𝐢3,𝛼 PoincarΓ©-Einstein manifold satisfying one of the following β€’ 3 ≀ 𝑛 ≀ 7 β€’ 𝑛 β‰₯ 8 ad 𝑋 is spin β€’ 𝑛 β‰₯ 8 and 𝑋 is locally conformally flat Then there exists a conformal compactification ¯𝑔 = 𝜌2𝑔+ which is a minimizer for 𝑄( ¯𝑋, 𝑀, [ ¯𝑔]). Furthermore, ¯𝑔 has vanishing scalar curvature and constant mean curvature. Remark 1.2.3. Some other approaches has been used to construct solutions to (1.2.16). Thus solutions will provide us with metric of zero scalar curvature and constant mean curvature, but they are not necessarily minimizers of Type II Yamabe invariant. See [Xu23]. 1.3 A Sharp Inequality Having established the Yamabe invariants in the previous section, we will now formulate inequalities that establish a connection between the geometry of the boundary and the interior. The work is initialized in [GH17] Theorem 1.6. Let (𝑋, 𝑔+) be a PoincarΓ©-Einstein manifold satisfying one of the following π‘Ž) 3 ≀ 𝑛 ≀ 5, or 𝑏) 𝑋 𝑖𝑠 𝑠𝑝𝑖𝑛. Let ( ¯𝑋, 𝑀, ¯𝑔) be its compactification and ˆ𝑔 = ¯𝑔(cid:12) (cid:12)𝑀. Then 𝑛 𝑛 βˆ’ 2 π‘Œ (𝑀, [ ˆ𝑔]) ≀ π‘Œ ( ¯𝑋, 𝑀, [ ¯𝑔])𝐼 ( ¯𝑋, 𝑋, ¯𝑔)2, if 𝑛 β‰₯ 4 12πœ‹ πœ’(𝑀) ≀ π‘Œ ( ¯𝑋, 𝑀, [ ¯𝑔])𝐼 ( ¯𝑋, 𝑋, ¯𝑔)2, if 𝑛 = 3 where 𝐼 ( ¯𝑋, 𝑋, ¯𝑔) = Vol(𝑀, ˆ𝑔)1/(π‘›βˆ’1)/Vol( ¯𝑋, ¯𝑔)1/𝑛. Moreover, if the equality holds, then ¯𝑔 is Einstein and ˆ𝑔 has constant scalar curvature. 10 This inequality tells in a certain sense that for PoincarΓ©-Einstein manifolds, the conformal geometry of the whole manifold can be controlled by the geometry of the conformal boundary. While the inequality represents a significant breakthrough in PoincarΓ©-Einstein manifold research, it has limitations, notably that 𝐼 ( ¯𝑋, 𝑀, 𝑔) isn’t conformally invariant. In [CLW17], X. Chen, M. Lai, and F. Wang introduced a new inequality (1.3.1) using 𝑄( ¯𝑋, 𝑀, 𝑔) instead of π‘Œ ( ¯𝑋, 𝑀, 𝑔). Theorem 1.7. Let (𝑋, 𝑔+) be a PoincarΓ©-Einstein manifold with compactification ( ¯𝑋, 𝑀, [ ¯𝑔]). Suppose (𝑋, 𝑔+) satisfies one of the conditions in Thm1.5 then 𝑄( ¯𝑋, 𝑀, [ ¯𝑔]) β‰₯ 2 βˆšοΈ‚ (𝑛 βˆ’ 1) 𝑛 βˆ’ 2 π‘Œ (𝑀, 𝑔(cid:12) (cid:12)𝑀) 𝑖 𝑓 𝑛 β‰₯ 4 𝑄( ¯𝑋, 𝑀, [ ¯𝑔]) β‰₯ 4√︁2πœ‹ πœ’(𝑀) 𝑖 𝑓 𝑛 = 3 (1.3.1) Moreover, the equality holds iff (𝑋, 𝑔+) is isometric to hyperbolic space (H𝑛, 𝑔H). Sketch of proof By Thm1.5, the minimizer for 𝑄( ¯𝑋, 𝑀, [ ¯𝑔]) can be achieved. Say ¯𝑔 = 𝜌2𝑔+, without loss of generality. Use 𝑅𝑖𝑐𝑔+ = βˆ’(𝑛 βˆ’ 1)𝑔+ and (1.2.4) and traceless Ricci of ¯𝑔 is given by ¯𝐸 = βˆ’(𝑛 βˆ’ 2) πœŒβˆ’1 (cid:104) Β―βˆ‡2𝜌 βˆ’ (cid:105) (Ξ” ¯𝑔 𝜌) ¯𝑔 1 𝑛 Integrating 𝜌| ¯𝐸 | ¯𝑔dV ¯𝑔 by parts yields ∫ ¯𝑋 𝜌| ¯𝐸 | ¯𝑔dV ¯𝑔 = ∫ 𝑀 1 𝜌 (cid:2)πœ•πœˆ | Β―βˆ‡πœŒ|2 ¯𝑔 + 1 𝜌 (1 βˆ’ | Β―βˆ‡πœŒ|2 ¯𝑔)πœ•πœˆ 𝜌(cid:3)dS ˆ𝑔 (1.3.2) where 𝜈 is outer normal and ˆ𝑔 = ¯𝑔| 𝑀. Since ¯𝑔 has zero scalar curvature, by (1.2.4) 2πœŒΞ” ¯𝑔 𝜌 = 𝑛(| Β―βˆ‡πœŒ|2 ¯𝑔 βˆ’ 1) By calculation in [Gra16], the equation above, together with (1.1.4) will give us local expansion for 𝜌 near conformal boundary: πœ•πœˆ 𝜌 = 1, πœ•2 𝜈 𝜌 = βˆ’ 1 𝑛 βˆ’ 1 ¯𝐻, πœ•3 𝜈 𝜌 = 1 𝑛 βˆ’ 2 ˆ𝑅 βˆ’ 1 𝑛 βˆ’ 1 ¯𝐻2 where ˆ𝑅 is the scalar curvature for ˆ𝑔 and ¯𝐻 is the mean curvature. Plug this into the integration above, we get 2 (𝑛 βˆ’ 2)2 ∫ 𝑋 𝜌| ¯𝐸 |2 ¯𝑔dV ¯𝑔 = ∫ 𝑀 11 (cid:16) 1 𝑛 βˆ’ 2 ¯𝐻2 βˆ’ 1 𝑛 βˆ’ 2 ˆ𝑅 (cid:17)dS ˆ𝑔 (1.3.1) follows by noting that ¯𝐻 is constant since ¯𝑔 minimizes Type II Yamabe invariant. β–‘ This inequality tells in a certain sense that for PoincarΓ©-Einstein manifolds, the conformal geometry of the whole manifold can be controlled by the geometry of the conformal boundary. However, their findings were constrained by two primary limitations. Firstly, their work rested upon the assumption that the minimizer of the second type Yamabe invariant could be realized. Further- more, their approach was confined to PoincarΓ©-Einstein manifolds, i.e. 𝑅𝑖𝑐𝑔+ = βˆ’(𝑛 βˆ’ 1)𝑔+. It’s important to note that many of the properties associated with PoincarΓ©-Einstein manifolds extend to asymptotically PoincarΓ©-Einstein manifolds with 𝑅𝑖𝑐𝑔+ β‰₯ βˆ’(𝑛 βˆ’ 1)𝑔+. The proof in [CLW17] highly depends on the vanishing of traceless Ricci curvature, so their method fails in general setting. In collaboration with X. Wang, we overcame these limitations, yielding the following result [WW21][WW22]. This inequality highlights the intricate relationship between the manifold’s boundary and its interior, aligning with our guiding principle. Theorem 1.8. (𝑋, 𝑔+) asymptotically PoincarΓ©-Einstein manifold with compactification ( ¯𝑋, 𝑀, ¯𝑔). Suppose 𝑅𝑖𝑐𝑔+ β‰₯ βˆ’(𝑛 βˆ’ 1)𝑔+ and the conformal infinity has nonnegative Yamabe invariant, then βˆšοΈ‚ (𝑛 βˆ’ 1) 𝑛 βˆ’ 2 𝑄( ¯𝑋, 𝑀, [ ¯𝑔]) β‰₯ 2 (cid:12)𝑀) 𝑖 𝑓 𝑛 β‰₯ 4 π‘Œ (𝑀, ¯𝑔(cid:12) (1.3.3) 𝑄( ¯𝑋, 𝑀, ¯𝑔) β‰₯ 4√︁2πœ‹ πœ’(𝑀) 𝑖 𝑓 𝑛 = 3 Moreover, the equality holds iff (𝑋, 𝑔+) is isometric to hyperbolic space (H𝑛, 𝑔H). Proof. The proof consists of three parts. First, we will define modified Yamabe quotients and subsequently a quantity derived from it, playing a role analogous to 𝜌|𝐸 |2 in [CLW17]. Next, we will analyze the asymptotic behavior of the function introduced in the initial step. Finally, we will prove a sequence of inequalities for each modified Yamabe quotients, the limit of which will yield the desired inequality. Finally we are going to prove rigidity, which in essence comes from [CLW17]. Let ˆ𝑔 = ¯𝑔| 𝑀. Throughout the proof, operators and tensors with a + are defined with respect to 𝑔+, those with a bar are defined with respect to ¯𝑔, and those with a hat are defined with respect to ˆ𝑔. 12 Step 1 From Cor1.1 in the next section, 𝑀 is connected. By Thm1.3 we can pick a β„Ž ∈ [ ¯𝑔| 𝑀] so that π‘…β„Ž is constant. Take Graham-Lee normal form (1.1.2). Lee [Lee94] constructed a function a positive smooth function πœ™ on 𝑋 s.t. Ξ”+πœ™ = π‘›πœ™ and near πœ• 𝑋 πœ™ = πœŒβˆ’1 + π‘…β„Ž 4 (𝑛 βˆ’ 1) (𝑛 βˆ’ 2) 𝜌 + π‘œ (cid:16) 𝜌2(cid:17) He further proved that |π‘‘πœ™|2 + βˆ’ πœ™2 ≀ 0 in the following way. Since we assume π‘Œ (𝑀, [β„Ž]) β‰₯ 0 and π‘…β„Ž is constant, π‘…β„Ž β‰₯ 0. By a direct calculation, |π‘‘πœ™|2 βˆ’ πœ™2 has a continuation extension to 𝑀 and |π‘‘πœ™|2 + βˆ’ πœ™2 ≀ 0 on 𝑀. By Bochner formula we have Ξ”+(|π‘‘πœ™|2 + βˆ’ πœ™2) = 2𝑔+(βˆ‡+Ξ”+πœ™, βˆ‡+πœ™) + 2|βˆ‡2 +πœ™|2 = 2𝑛|π‘‘πœ™|2 + + 2|βˆ‡2 +πœ™|2 + + 2𝑅𝑖𝑐+(βˆ‡+πœ™, βˆ‡+πœ™) βˆ’ = 2(𝑅𝑖𝑐+(βˆ‡+πœ™, βˆ‡+πœ™) + (𝑛 βˆ’ 1)|π‘‘πœ™|2 + + 2𝑅𝑖𝑐+(βˆ‡+πœ™, βˆ‡+πœ™) βˆ’ 2(πœ™Ξ”+πœ™ + |βˆ‡+πœ™|2 +) + βˆ’ 2|βˆ‡+πœ™|2 |Ξ”+πœ™|2 1 𝑛 2 𝑛 +πœ™|2 +) + 2(|βˆ‡2 |Ξ”+πœ™|2) + βˆ’ + β‰₯ 0 As a result |π‘‘πœ™|2 + βˆ’ πœ™2 ≀ 0 on 𝑋. Consider the metric 𝑔 := πœ™βˆ’2𝑔+ on ¯𝑋 . Its scalar curvature is (cid:101) given by 𝑅+ + 2 (𝑛 βˆ’ 1) πœ™βˆ’1Ξ”+πœ™ βˆ’ 𝑛 (𝑛 βˆ’ 1) πœ™βˆ’2 |π‘‘πœ™|2 + (cid:101)𝑅 = πœ™2 (cid:16) β‰₯ πœ™2 (𝑅+ + 𝑛 (𝑛 βˆ’ 1)) β‰₯ 0 (cid:17) Moreover, by a direct calculation the boundary is totally geodesic. We consider the following modified energy functional ˜𝐸 ( 𝑓 ) = 𝐸𝑔 ( 𝑓 ) βˆ’ ∫ 𝑋 (𝑅+ + 𝑛 (𝑛 βˆ’ 1)) πœ™2 𝑓 2𝑑𝑣𝑔. (1.3.4) Note that (𝑅 + 𝑛 (𝑛 βˆ’ 1)) πœ™2 ∈ πΆπ‘šβˆ’3,𝛼 (cid:16) 𝑋 (cid:17) under our assumptions. More explicitly, by (1.3.4) ˜𝐸 ( 𝑓 ) = ∫ 𝑋 (cid:20) 4 (𝑛 βˆ’ 1) 𝑛 βˆ’ 2 |𝑑𝑓 |2 𝑔 + (cid:101) (cid:16) (cid:101)𝑅 βˆ’ (𝑅 + 𝑛 (𝑛 βˆ’ 1)) πœ™2(cid:17) (cid:21) 𝑓 2 𝑑𝑣𝑔 β‰₯ 0. Since 𝑅+ + 𝑛 (𝑛 βˆ’ 1) β‰₯ 0, we have 𝐸𝑔 ( 𝑓 ) β‰₯ (cid:101)𝐸 ( 𝑓 ) . 13 (1.3.5) (1.3.6) For 1 < π‘ž ≀ 𝑛/(𝑛 βˆ’ 2), consider Λœπœ†π‘ž := inf (cid:101)𝐸 ( 𝑓 ) 𝑀 | 𝑓 |π‘ž+1 π‘‘πœŽπ‘” (cid:16)∫ . (cid:17) 2/(π‘ž+1) (1.3.7) Lemma 1.3.1. Since ˜𝐸 ( 𝑓 ) β‰₯ 0, limπ‘žβ†—π‘›/(π‘›βˆ’2) (cid:101)πœ†π‘ž = (cid:101)πœ†π‘›/(π‘›βˆ’2). Proof of lemma Pick a minimizing sequence 𝑒𝑖 for 𝑄(𝑋, 𝑀, [ ¯𝑔]). For each 𝑒𝑖, ˜𝐸 (𝑒𝑖) lim π‘žβ†— 𝑛 π‘›βˆ’2 (∫ 𝑀 ˜𝐸 (𝑒𝑖) π‘’π‘ž+1 𝑖 dS ˆ𝑔) = 2 π‘ž+1 (∫ 𝑀 𝑒 2(π‘›βˆ’1) π‘›βˆ’2 𝑖 dS ˆ𝑔) π‘›βˆ’2 π‘›βˆ’1 As a result lim supπ‘žβ†—π‘›/(π‘›βˆ’2) (cid:101)πœ†π‘ž ≀ Λœπœ†π‘›/(π‘›βˆ’2). Since ˜𝐸 (𝑒) β‰₯ 0, by HΓΆlder inequality ˜𝐸 (𝑒) π‘’π‘ž+1dS ˆ𝑔) β‰₯ 2 π‘ž+1 (∫ 𝑀 (∫ 𝑀 𝑒 ˜𝐸 (𝑒) 2(π‘›βˆ’1) π‘›βˆ’2 dS ˆ𝑔) π‘›βˆ’2 π‘›βˆ’1 Area(𝑀, ˆ𝑔) (π‘›βˆ’1) βˆ’ 2 π‘›βˆ’2 π‘ž+1 As a result Λœπœ†π‘ž β‰₯ Λœπœ† 𝑛 π‘›βˆ’2 Area(𝑀, ˆ𝑔) (π‘›βˆ’1) βˆ’ 2 π‘›βˆ’2 π‘ž+1 . Take a limit, and we have lim inf π‘žβ†—π‘›/(π‘›βˆ’2) (cid:101)πœ†π‘ž β‰₯ Λœπœ†π‘›/(π‘›βˆ’2) β–‘ Since ˜𝐸 ( 𝑓 ) β‰₯ 0, it is easy to see that limπ‘žβ†—π‘›/(π‘›βˆ’2) (cid:101)πœ†π‘ž = (cid:101)πœ†π‘›/(π‘›βˆ’2). Therefore, it suffices to prove the above theorem for π‘ž < 𝑛/(𝑛 βˆ’ 2). Since the trace operator 𝐻1 (cid:16) 𝑋 (cid:17) β†’ πΏπ‘ž+1 (Ξ£) is compact for π‘ž < 𝑛/(𝑛 βˆ’ 2), by standard elliptic theory, the above infimum πœ†π‘ž is achieved by a smooth, positive function 𝑓 s.t. and ∫ Ξ£ 𝑓 π‘ž+1π‘‘πœŽ = 1 βˆ’ 4(π‘›βˆ’1) π‘›βˆ’2 Ξ” 𝑓 + 𝑅 𝑓 = (𝑅 + 𝑛 (𝑛 βˆ’ 1)) πœ™2 𝑓 πœ• 𝑓 πœ•πœˆ = πœ†π‘ž 𝑓 π‘ž 4(π‘›βˆ’1) π‘›βˆ’2 on 𝑋, on 𝑀. (1.3.8) (1.3.9)   ο£³ By the conformal invariance of the conformal Laplacian, we have (cid:16) 𝑓 πœ™βˆ’(π‘›βˆ’2)/2(cid:17) 𝐿𝑔 = πœ™βˆ’(𝑛+2)/2𝐿𝑔 ( 𝑓 ) = (𝑅 + 𝑛 (𝑛 βˆ’ 1)) 𝑓 πœ™βˆ’(π‘›βˆ’2)/2. 14 In other words, 𝑒 := 𝑓 πœ™βˆ’(π‘›βˆ’2)/2 satisfies the following equation βˆ’Ξ”π‘”+ 𝑒 = 𝑛 (𝑛 βˆ’ 2) 4 𝑒. (1.3.10) Write 𝑒 = π‘£βˆ’(π‘›βˆ’2)/2. Then Equivalently Δ𝑔+ 𝑣 βˆ’ 𝑛𝑣 = 𝑛 𝑣 = 𝑛 Δ𝑔+ 2 2 Ξ¦ with Ξ¦ = π‘£βˆ’1 (cid:16) π‘£βˆ’1 (cid:16) |𝑑𝑣|2 𝑔+ |𝑑𝑣|2 𝑔+ βˆ’ 𝑣2(cid:17). + 𝑣2(cid:17) . Lemma 1.3.2. We have (cid:16) div+ π‘£βˆ’(π‘›βˆ’2)βˆ‡+Ξ¦ (cid:17) = 2π‘£βˆ’(π‘›βˆ’2)𝑄, (1.3.11) where 𝑄 = (cid:12) (cid:12) (cid:12) (cid:12) βˆ‡2 +𝑣 βˆ’ Ξ”+𝑣 𝑛 𝑔+ (cid:12) (cid:12) (cid:12) (cid:12) 2 + + 𝑅𝑖𝑐+ (βˆ‡+𝑣, βˆ‡+𝑣) + (𝑛 βˆ’ 1) |βˆ‡+𝑣|2 + β‰₯ 0. All the computation is done with respect to 𝑔+, but we drop the subscript to simplify the presentation. Proof. As 𝑣Φ = |βˆ‡+𝑣|2 βˆ’ 𝑣2, we have, by using the Bochner formula 1 2 (𝑣Δ+πœ™ + 2 βŸ¨βˆ‡+𝑣, βˆ‡πœ™βŸ©+ + πœ™Ξ”+𝑣) = (cid:12) + + βŸ¨βˆ‡+𝑣, βˆ‡+Ξ”+π‘£βŸ©+ + 𝑅𝑖𝑐+ (βˆ‡+𝑣, βˆ‡+𝑣) βˆ’ 𝑣Δ+𝑣 βˆ’ |βˆ‡+𝑣|2 + 2 +𝑣(cid:12) (cid:12)βˆ‡2 (cid:12) (Ξ”+𝑣)2 𝑛 Ξ”+𝑣 𝑛 = = = + βŸ¨βˆ‡+𝑣, βˆ‡+Ξ”+π‘£βŸ©+ + 𝑣Δ+𝑣 βˆ’ 𝑛 |βˆ‡+𝑣|2 + + 𝑄 (Ξ”+𝑣 βˆ’ 𝑛𝑣) + βŸ¨βˆ‡+𝑣, βˆ‡+ (Ξ”+𝑣 βˆ’ 𝑛𝑣)⟩+ + 𝑄 1 2 ΦΔ+𝑣 + 𝑛 2 βŸ¨βˆ‡+𝑣, βˆ‡+Φ⟩+ + 𝑄 Thus, or ΔΦ+ = (𝑛 βˆ’ 2) π‘£βˆ’1 βŸ¨βˆ‡+𝑣, βˆ‡+Φ⟩+ + 2𝑄 (cid:16) div+ π‘£βˆ’(π‘›βˆ’2)βˆ‡+Ξ¦ (cid:17) = 2π‘£βˆ’(π‘›βˆ’2)𝑄 β‰₯ 0. div (cid:16) π‘£βˆ’(π‘›βˆ’2)βˆ‡Ξ¦ (cid:17) plays the role of traceless Ricci 𝐸 in [CLW17]. 15 β–‘ Step 2 In this part we are going to figure out the asymptotical expansion for terms in div (cid:16) π‘£βˆ’(π‘›βˆ’2)βˆ‡Ξ¦ (cid:17). We now consider the metric 𝑔 = 𝑒4/(π‘›βˆ’2)𝑔+. Since 𝑒 = 𝑓 πœ™βˆ’(π‘›βˆ’2)/2, we also have 𝑔 = 𝑓 4/(π‘›βˆ’2) πœ™βˆ’2𝑔+ = 𝑓 4/(π‘›βˆ’2) 𝑔. (cid:101) As πœ• 𝑋 is totally geodesic w.r.t. 𝑔 and 𝑔 is conformal to (cid:101) 𝑔, we know that πœ• 𝑋 is umbilic w.r.t. 𝑔 and (cid:101) its mean curvature, in view of the boundary condition of (1.3.9), is given by 𝐻 = πœ†π‘ž 2 𝑓 π‘žβˆ’ 𝑛 π‘›βˆ’2 . (1.3.12) Set 𝜌 = 𝑒2/(π‘›βˆ’2) = π‘£βˆ’1. By a direct calculation, the equation (1.3.10) becomes, using 𝑔 as the background metric 2πœŒΞ”πœŒ = 𝑛 (cid:16) |βˆ‡πœŒ|2 βˆ’ 1(cid:17) . (1.3.13) Let 𝑑 be the geodesic distance to Ξ£ w.r.t. 𝑔. We need the following lemma which is essentially contained in [CLW17]. Lemma 1.3.3. Near Ξ£ = πœ• 𝑋, we can write 𝑔 = 𝑑𝑑2 + 𝑔𝑖 𝑗 (𝑑, π‘₯) 𝑑π‘₯𝑖𝑑π‘₯ 𝑗 , where {π‘₯1, Β· Β· Β· , π‘₯π‘›βˆ’1} are local coordinates on Ξ£. Then 𝜌 = 𝑑 βˆ’ 𝐻 2 (𝑛 βˆ’ 1) 𝑑2 + 1 6 (cid:18) 𝑅Σ 𝑛 βˆ’ 2 βˆ’ (cid:19) 𝐻2 𝑛 βˆ’ 1 𝑑3 + π‘œ (cid:16) 𝑑3(cid:17) . In particular, πœ• πœ•πœˆ (cid:104) πœŒβˆ’1 (cid:16) |βˆ‡πœŒ|2 βˆ’ 1(cid:17)(cid:105) |Ξ£ = 𝑅Σ 𝑛 βˆ’ 2 βˆ’ 𝐻2 𝑛 βˆ’ 1 . Proof. For completeness, we present the proof showing that the Einstein condition is not required. In local coordinates |βˆ‡πœŒ|2 = Ξ”πœŒ = (cid:19) 2 (cid:18) πœ• 𝜌 πœ•π‘‘ πœ•2𝜌 πœ•π‘‘2 + + 𝑔𝑖 𝑗 πœ• 𝜌 πœ•π‘₯𝑖 √ 𝐺 πœ• log πœ•π‘‘ πœ• 𝜌 πœ•π‘₯ 𝑗 πœ• 𝜌 πœ•π‘‘ , + 1 √ 𝐺 πœ• πœ•π‘₯𝑖 (cid:18) 𝑔𝑖 𝑗 √ 𝐺 (cid:19) . πœ• 𝜌 πœ•π‘₯ 𝑗 16 Restricting (1.3.13) on Ξ£ on which both 𝜌 and π‘Ÿ vanish with order 1 yields πœ• 𝜌 πœ•π‘‘ |Ξ£ = 1. Differentiating (1.3.13) in 𝑑 yields 2 𝑛 (cid:18) πœ• 𝜌 πœ•π‘‘ Ξ”πœŒ + 𝜌 (cid:19) Ξ”πœŒ πœ• πœ•π‘‘ = 2 πœ• 𝜌 πœ•π‘‘ Evaluating both sides on Ξ£ yields πœ•2𝜌 πœ•π‘‘2 + 2𝑔𝑖 𝑗 πœ•2𝜌 πœ•π‘₯π‘–πœ•π‘‘ πœ• 𝜌 πœ•π‘₯ 𝑗 βˆ’ π‘”π‘–π‘˜ 𝑔 𝑗𝑙 πœ•π‘”π‘˜π‘™ πœ•π‘‘ πœ• 𝜌 πœ•π‘₯𝑖 πœ• 𝜌 πœ•π‘₯ 𝑗 . (1.3.14) (cid:32) πœ•2𝜌 πœ•π‘‘2 + 2 𝑛 √ 𝐺 πœ• log πœ•π‘‘ (cid:33) |Ξ£ = 2 πœ•2𝜌 πœ•π‘‘2 |Ξ£. Thus πœ•2𝜌 πœ•π‘‘2 |Ξ£ = 1 𝑛 βˆ’ 1 √ 𝐺 πœ• log πœ•π‘‘ |Ξ£ = βˆ’ 𝐻 𝑛 βˆ’ 1 . Differentiating the formula for Ξ”πœŒ we get (cid:32) πœ•3𝜌 πœ•π‘‘3 + (cid:32) πœ•3𝜌 πœ•π‘‘3 + Ξ”πœŒ|Ξ£ = πœ• πœ•π‘‘ = √ 𝐺 πœ•2 log πœ•π‘‘2 √ 𝐺 πœ•2 log πœ•π‘‘2 + + √ 𝐺 πœ• log πœ•π‘‘ (cid:33) πœ•2𝜌 πœ•π‘‘2 |Ξ£ (cid:33) 𝐻2 𝑛 βˆ’ 1 |Ξ£ Differentiating (1.3.14) in π‘Ÿ and evaluating on Ξ£, we obtain 2 𝑛 (cid:18) πœ•2𝜌 πœ•π‘‘2 Ξ”πœŒ + 2 πœ• πœ•π‘‘ (cid:19) Ξ”πœŒ |Ξ£ = 2 (cid:19) 2 (cid:18) πœ•2𝜌 πœ•π‘‘2 |Ξ£ + 2 πœ•3𝜌 πœ•π‘‘3 |Ξ£ = 2𝐻2 (𝑛 βˆ’ 1)2 + 2 πœ•3𝜌 πœ•π‘‘3 |Ξ£. Using the previous formulas, we arrive at πœ•3𝜌 πœ•π‘‘3 |Ξ£ = 2 𝑛 βˆ’ 2 (cid:32) 𝐻2 𝑛 βˆ’ 1 + By a direct calculation, we also have √ 𝐺 πœ•2 log πœ•π‘‘2 (cid:33) |Ξ£ . √ 𝐺 πœ•2 log πœ•π‘‘2 |Ξ£ = βˆ’π‘…π‘–π‘ (𝜈, 𝜈) βˆ’ 𝐻2 𝑛 βˆ’ 1 . Therefore πœ•3𝜌 πœ•π‘‘3 |Ξ£ = βˆ’ 2 𝑛 βˆ’ 2 𝑅Σ 𝑛 βˆ’ 2 βˆ’ = 𝑅𝑖𝑐 (𝜈, 𝜈) 𝐻2 𝑛 βˆ’ 1 , where we used the Gauss equation in the last step. The second identity follows from a direct calculation. β–‘ 17 Step 3 In this part we will use lemma1.3.3 in (1.3.11) to get the main result. Integrating the identity (1.3.11) on π‘‹πœ€ = {𝑑 β‰₯ πœ€} yields ∫ 2 π‘‹πœ€ π‘£βˆ’(π‘›βˆ’2)𝑄𝑑𝑣𝑔+ = ∫ πœ• π‘‹πœ€ π‘£βˆ’(π‘›βˆ’2) πœ•Ξ¦ πœ•πœˆ π‘‘πœŽπ‘”+ . Since 𝑔+ = πœŒβˆ’2𝑔, we obtain by a direct calculation ∫ πœ• π‘‹πœ€ π‘£βˆ’(π‘›βˆ’2) πœ•Ξ¦ πœ•πœˆ+ π‘‘πœŽπ‘”+ = ∫ πœ• π‘‹πœ€ πœ• πœ•πœˆ (cid:104) πœŒβˆ’1 (cid:16) |βˆ‡πœŒ|2 βˆ’ 1(cid:17)(cid:105) π‘‘πœŽπ‘”. Therefore ∫ 2 π‘‹πœ€ Letting πœ€ β†’ 0, we obtain, in view of Lemma 1.3.3 πœ• π‘‹πœ€ π‘£βˆ’(π‘›βˆ’2)𝑄𝑑𝑣𝑔+ = ∫ πœ• πœ•πœˆ (cid:104) πœŒβˆ’1 (cid:16) |βˆ‡πœŒ|2 βˆ’ 1(cid:17)(cid:105) π‘‘πœŽπ‘”. ∫ 2 𝑋 π‘£βˆ’(π‘›βˆ’2)𝑄𝑑𝑣𝑔+ = ∫ Ξ£ (cid:18) 𝑅Σ 𝑛 βˆ’ 2 βˆ’ (cid:19) 𝐻2 𝑛 βˆ’ 1 π‘‘πœŽπ‘” (1.3.15) The rest of the argument is the same as in [WW21]. We present it for completeness. By (1.3.12) and the Holder inequality again ∫ Ξ£ 𝐻2π‘‘πœŽ = = ≀ = (cid:18) πœ†π‘ž 2 (cid:18) πœ†π‘ž 2 (cid:18) πœ†π‘ž 2 (cid:18) πœ†π‘ž 2 (cid:19) 2 ∫ Ξ£ (cid:19) 2 ∫ Ξ£ (cid:19) 2 (cid:18)∫ 𝑓 2(π‘žβˆ’ 𝑛 π‘›βˆ’2) 𝑓 2(π‘›βˆ’1)/(π‘›βˆ’2) π‘‘πœŽ 𝑓 2(π‘žβˆ’ 1 π‘›βˆ’2) π‘‘πœŽ (cid:19) 2(π‘žβˆ’ 1 π‘›βˆ’2 )/(π‘ž+1) 𝑓 π‘ž+1π‘‘πœŽ 𝑉 (Ξ£, 𝑔) ( 𝑛 π‘›βˆ’2 βˆ’π‘ž)/(π‘ž+1) Ξ£ (cid:19) 2 𝑉 (Ξ£, 𝑔)( 𝑛 π‘›βˆ’2 βˆ’π‘ž)/(π‘ž+1) . Plugging the above inequality into (1.3.15), we obtain ∫ 2 𝑋 π‘£βˆ’(π‘›βˆ’2)𝑄𝑑𝑣𝑔+ ≀ πœ†2 π‘ž 4 (𝑛 βˆ’ 1) 𝑉 (Ξ£, 𝑔) ( 𝑛 π‘›βˆ’2 βˆ’π‘ž)/(π‘ž+1) βˆ’ 1 𝑛 βˆ’ 2 ∫ Ξ£ π‘…Ξ£π‘‘πœŽ. (1.3.16) When 𝑛 = 3, this implies π‘žπ‘‰ (Ξ£, 𝑔) (3βˆ’π‘ž)/(π‘ž+1) β‰₯ 32πœ‹ πœ’ (Ξ£) . πœ†2 18 In the following, we assume 𝑛 > 3 . By (1.3.8) and the HΓΆlder inequality ∫ 1 = 𝑓 π‘ž+1π‘‘πœŽ Ξ£ (cid:18)∫ Ξ£ ≀ 𝑓 2(π‘›βˆ’1)/(π‘›βˆ’2) π‘‘πœŽ (cid:19) (π‘ž+1) (π‘›βˆ’2) 2(π‘›βˆ’1) 𝑉 (Ξ£, 𝑔) π‘›βˆ’π‘ž (π‘›βˆ’2) 2(π‘›βˆ’1) = 𝑉 (Ξ£, 𝑔) (π‘ž+1) (π‘›βˆ’2) 2(π‘›βˆ’1) 𝑉 (Ξ£, 𝑔) π‘›βˆ’π‘ž (π‘›βˆ’2) 2(π‘›βˆ’1) Thus 𝑉 (Ξ£, 𝑔)βˆ’ π‘›βˆ’π‘ž (π‘›βˆ’2) (π‘›βˆ’2) (π‘ž+1) ≀ 𝑉 (Ξ£, 𝑔) . Plugging this inequality into (1.3.16) yields π‘£βˆ’(π‘›βˆ’2)𝑄𝑑𝑣𝑔+ (cid:34) ∫ 2 𝑋 π‘›βˆ’1 𝑉 (Ξ£, 𝑔) π‘›βˆ’3 4 (𝑛 βˆ’ 1) π‘›βˆ’1 𝑉 (Ξ£, 𝑔) π‘›βˆ’3 4 (𝑛 βˆ’ 1) (cid:101)πœ†2 π‘žπ‘‰ (Ξ£, 𝑔) 2(π‘›βˆ’π‘ž (π‘›βˆ’2) ) (π‘›βˆ’3) (π‘ž+1) βˆ’ (cid:20) (cid:101)πœ†2 π‘žπ‘‰ (Ξ£, 𝑔) 2(π‘›βˆ’π‘ž (π‘›βˆ’2) ) (π‘›βˆ’3) (π‘ž+1) βˆ’ 4 (𝑛 βˆ’ 1) π‘›βˆ’1 π‘›βˆ’3 (𝑛 βˆ’ 2) 𝑉 (Ξ£, 𝑔) 4 (𝑛 βˆ’ 1) (𝑛 βˆ’ 2) π‘Œ (Ξ£, [𝛾]) (cid:35) π‘…Ξ£π‘‘πœŽ ∫ Ξ£ . (cid:21) ≀ ≀ Therefore (cid:101)πœ†2 π‘ž β‰₯ 4 (𝑛 βˆ’ 1) (𝑛 βˆ’ 2) π‘Œ (Ξ£) 𝑉 (Ξ£, 𝑔)βˆ’ 2(π‘›βˆ’π‘ž (π‘›βˆ’2) ) (π‘›βˆ’3) (π‘ž+1) . Finally let π‘ž β†— 𝑛 π‘›βˆ’2 and we arrive at the desired inequality in Theorem 1.8. Step 4 Suppose the equality in (1.3.3) holds for (𝑋, 𝑔+) as in Thm1.8. Let (𝑋, 𝑀, ˆ𝑔) be its conformal boundary and ˆ𝑔 = ¯𝑔| 𝑀. If 𝑄( ¯𝑋, 𝑀, [ ¯𝑔]) = 𝑄(B𝑛, Sπ‘›βˆ’1, 𝑑π‘₯2), then π‘Œ (𝑀, [ ˆ𝑔]) = π‘Œ (Sπ‘›βˆ’1, [π‘‘πœƒ2]) from the equality. By Thm1.3, (𝑀, [ ˆ𝑔]) is the round metric on Sπ‘›βˆ’1. Then Thm1.10 implies that (𝑋, 𝑔+) is the standard hyperbolic space. Now we suppose 𝑄( ¯𝑋, 𝑀, [ ¯𝑔]) < 𝑄(B𝑛, Sπ‘›βˆ’1, 𝑑π‘₯2). In this case the minimizer for Type II Yamabe invariant can be realized, say ¯𝑔 = 𝜌2𝑔+. Note that we defined Λœπœ†π‘ž and proved inequality for π‘›βˆ’2 in step 3. This is because we are not sure whether minimizer for 𝑄( ¯𝑋, 𝑀, [ ¯𝑔]) exists. each π‘ž < 𝑛 Now since we have got minimizer, we can run the previous method directly and get ∫ 2 𝑋 𝑣2βˆ’π‘›π‘„dV𝑔+ ≀ 𝑉 (𝑀, ˆ𝑔) π‘›βˆ’1 π‘›βˆ’3 4(𝑛 βˆ’ 1) (cid:104) 𝑄( ¯𝑋, 𝑀, [ ¯𝑔])2 βˆ’ 4(𝑛 βˆ’ 1) 𝑛 βˆ’ 2 π‘Œ (𝑀, [ ˆ𝑔]) (cid:105) 19 Given the assumption that equality holds in (1.3.3), it follows that equality also holds in (1.3.6). This implies 𝑅+ = βˆ’π‘›(𝑛 βˆ’ 1). Combining this with 𝑅𝑖𝑐+ β‰₯ βˆ’(𝑛 βˆ’ 1)𝑔+, we deduce 𝑅𝑖𝑐+ = βˆ’(𝑛 βˆ’ 1)𝑔+. We now find ourselves in a situation analogous to that in [CLW17], and their approach is applicable here as well. For the sake of completeness, we provide a detailed proof. We also get 𝑄 ≑ 0 and thus βˆ‡2 +𝑣 = Ξ”+𝑣 𝑣2 𝑔+. Compute βˆ‡2 +𝜌 βˆ‡2 +𝑣(𝑋, π‘Œ ) = 𝑔+ (cid:17) (cid:16) (βˆ‡+)𝑋 βˆ‡+ 𝑛 𝑔+. Recall 𝑔 = 𝜌2𝑔+ = 1 1 𝜌 1 𝜌2 βˆ‡+𝜌), π‘Œ (cid:17) (βˆ‡+)𝑋 ( , π‘Œ (cid:16) (cid:16) (cid:17) 𝑔+ (βˆ‡+)𝑋 βˆ‡+𝜌, π‘Œ + 2 = βˆ’π‘”+ = βˆ’ 1 𝜌2 (𝑋 𝜌) (π‘Œ 𝜌) 𝜌3 Taking trace and we get Ξ”+𝑣 = βˆ’ 1 𝜌2 Ξ”+𝜌 + 2 |βˆ‡+𝜌|+ 𝜌3 Substitute the above two equations into βˆ‡2 +𝑣 = Ξ”+𝑣 𝑛 𝑔+ and we have βˆ‡2 +𝜌 = 1 𝑛 (cid:0)Ξ”+𝜌 βˆ’ |βˆ‡+𝜌|2 + 𝜌 (cid:1)𝑔+ + 2 𝜌 π‘‘πœŒ βŠ— π‘‘πœŒ We now aim to express the preceding equation in terms of 𝑔 = 𝜌2𝑔+. Use (1.2.2), the three equation above give us We obtain Use (1.2.3) and we get βˆ‡2𝜌 = Ξ”πœŒ 𝑛 𝑔 βˆ‡π‘–βˆ‡π‘–βˆ‡ 𝑗 𝜌 = 1 𝑛 βˆ‡ 𝑗 (Ξ”πœŒ) (1.3.17) (1.3.18) 𝑅𝑖𝑐+ = 𝑅𝑖𝑐 βˆ’ (𝑛 βˆ’ 2) (cid:0)πœŒβˆ‡2( ) βˆ’ 2𝜌2(𝑑 ) βŠ— (𝑑 1 𝜌 βˆ‡2𝜌 + (cid:0) Ξ”πœŒ 𝜌 |π‘‘πœŒ|2 𝜌2 1 𝜌 2Ξ”πœŒ π‘›πœŒ βˆ’ 1 𝜌 1 𝜌 |π‘‘πœŒ|2 𝜌2 + (𝑛 βˆ’ 1) (cid:1)𝑔 = 𝑅𝑖𝑐 + (𝑛 βˆ’ 2) = 𝑅𝑖𝑐 + (𝑛 βˆ’ 1) (cid:0) )(cid:1) βˆ’ (cid:0)πœŒΞ” 1 𝜌 + (𝑛 βˆ’ 3) 𝜌2|𝑑 ( 1 𝜌 )|2(cid:1)𝑔 (cid:1)𝑔 (1.3.19) We use (1.3.17) in the last equality. Recall that the scalar curvature of 𝑔 is 0, (1.2.3) implies 𝑅+ = βˆ’π‘›(𝑛 βˆ’ 1) = βˆ’π‘›(𝑛 βˆ’ 1)|βˆ‡πœŒ|2 + 2(𝑛 βˆ’ 1) πœŒΞ”πœŒ 20 So we have 𝑅𝑖𝑐 ≑ 0. So βˆ‡π‘–βˆ‡π‘–βˆ‡ 𝑗 𝜌 = βˆ‡ 𝑗 βˆ‡π‘–βˆ‡π‘– 𝜌 + 𝑅𝑖𝑐𝑖 𝑗 βˆ‡π‘– 𝜌 = βˆ‡ 𝑗 (Ξ”πœŒ) (1.3.20) Compare (1.3.18) and (1.3.20), we have βˆ‡(Ξ”πœŒ) = 0, and therefore Ξ”πœŒ is constant. Use lemma 1.3.3 and (1.3.13), we get the constant Ξ”πœŒ = βˆ’ 𝑛 𝐻 β‰  0. If not, Ξ”πœŒ = 0 and 𝜌 = 0 on πœ• 𝑋, which implies 𝜌 ≑ 0, which is impossible. Set 𝑀 = βˆ’ π‘›βˆ’1 π‘›βˆ’1 𝐻 where 𝐻 is the mean curvature for 𝑔. Apparently 𝑛𝐻 𝜌, then 𝑀 satisfies Integrate (Δ𝑀)2 Δ𝑀 𝑀 = 1 in 𝑋 = 0 on 𝑀 (1.3.21) πœ•π‘€ πœ•πœˆ = π‘›βˆ’1 𝑛𝐻 on 𝑀   ο£³ 𝑛 βˆ’ 1 𝑛 Vol( ¯𝑋, 𝑔) = 𝑛 βˆ’ 1 𝑛 ∫ 𝑋 (Δ𝑀)2dV𝑔 (cid:2)(Δ𝑀)2 βˆ’ |βˆ‡π‘€|2(cid:3)dV𝑔 ∫ 𝑋 ∫ = = = ( 𝐻 𝑀 𝑛 βˆ’ 1 𝑛 πœ•π‘€ dS ˆ𝑔 πœ•πœˆ )2 ∫ 𝑀 1 𝐻 dS ˆ𝑔 where we used Reilly’s formula in the third line. Therefore we arrive at ∫ πœ• 𝑋 𝑛 βˆ’ 1 𝐻 dS ˆ𝑔 = 𝑛Vol( ¯𝑋, 𝑔) Recall that 𝑅𝑖𝑐 = 0. We conclude that (𝑋, 𝑀, 𝑔) is isometric to Euclidean ball by [Mul87]. β–‘ 1.4 Geometry on Conformal Boundary Affects Geometry of the Interior As discussed in the first section, a fundamental principle guiding the research on (asymptot- ically) PoincarΓ©-Einstein manifolds is to comprehend the intersection between the geometry of (𝑋, 𝑔+) and the geometry of its conformal boundary. In this section, I will introduce preliminary results utilized in the preceding section and demonstrate how our Thm1.8 exemplifies this principle. We start with a toplogy result. 21 Theorem 1.9. Let (𝑋, 𝑔+) asymptotically PoincarΓ©-Einstein manifold and 𝑅𝑖𝑐 β‰₯ βˆ’(𝑛 βˆ’ 1)𝑔+. If one connected component of its boundary has non-negative Yamabe invariant, then π»π‘›βˆ’1(𝑋; Z) = 0. In [WY99], E. Witten and S. Yau established the above theorem under the assumption that one boundary component has a positive Yamabe invariant. They introduced the brane action defined by 𝐿 (Ξ£) = Area(Ξ£) βˆ’ 𝑛𝑉 (Ξ©) (1.4.1) where Ξ£ = πœ•Ξ©, and Ξ© is a domain in 𝑋. Given the conditions outlined in the theorem, and assuming a strictly positive Yamabe invariant, they demonstrated the following: 1) 𝐿 (Ξ£) admits a minimum through local calculations; 2) there exists a minimum in each nontrivial homology class if the boundary has a component of positive scalar curvature. Therefore π»π‘›βˆ’1(𝑋; Z) = 0. Later M.Cai and G. Galloway proved the zero Yamabe invaraint case using Riccati equation and Busemann functions. Let 𝜌 be a boundary defining function and Ξ£πœ– = {𝜌(π‘₯) = πœ– }. They consider a new Busemann function given by π›½πœ– (π‘₯) = 𝑑 (Ξ£πœ– , π‘œ) βˆ’ 𝑑 (π‘₯, Ξ£πœ– ) 𝛽(π‘₯) = lim πœ–β†’0 π›½πœ– (π‘₯) (1.4.2) Using the Riccati equation, they successfully demonstrated Δ𝛽 β‰₯ 𝑛 βˆ’ 1, provided the conformal boundary has a zero Yamabe invariant. If 𝑋 has more than one end, a carefully chosen ray can be constructed. Let 𝑏 be the usual Busemann function associated with this ray, resulting in 𝛽 + 𝑏 ≀ 0 with equality at an interior point. Since 𝑅𝑖𝑐 β‰₯ βˆ’(𝑛 βˆ’ 1)𝑔+, we have Δ𝑏 β‰₯ βˆ’(𝑛 βˆ’ 1). Now, 𝛽 + 𝑏 is a subharmonic function with an interior maximum point, implying 𝛽 + 𝑏 ≑ 0. This equality leads to the splitting (𝑋 = R Γ— Ξ£, 𝑔+ = 𝑒2π‘Ÿ + β„Ž). At π‘Ÿ = βˆ’βˆž, we encounter a cusp, which contradicts the asymptotic PoincarΓ©-Einstein condition. By standard topology argument, we have Corollary 1.1. Under the same assumption as stated in the preceding theorem, it follows that πœ• 𝑋 is connected. 22 By this corollary, manifolds in Thm1.8 will have connected boundary and brings us no trouble. Apart from topology, the conformal geometry of the boundary also affects the metric inside. We have the following rigidity result Theorem 1.10. Let (𝑋, 𝑔+) be an asymptotically PoincarΓ©-Einstein manifolds with 𝑅𝑖𝑐 β‰₯ βˆ’(𝑛 βˆ’ 1)𝑔+. Suppose (𝑋, 𝑔+) has round sphere as its conformal boundary, then (𝑋, 𝑔+) is isometric to the hyperbolic space (H𝑛, 𝑔H). The theorem was initially established by Q. Jie in [Qin03] for 𝑛 ≀ 7. In the context of hyperbolic spaces, we can consider the upper plane model. Q. Jie observed that if (𝑋, 𝑔+) has a round sphere as its conformal boundary, we can construct coordinate functions and utilize them to apply conformal transformations, resulting in an uncompact manifold with Rπ‘›βˆ’1 as its boundary, akin to the upper plane model. Moreover, the scalar curvature is non-negative for the new metric. Consequently, we can glue two such manifolds along Rπ‘›βˆ’1 to obtain an asymptotically flat manifold ( Λœπ‘‹, Λœπ‘”) with non-negative scalar curvature. Notably, its Arnowitt-Deser-Misner (ADM) mass π‘š 𝐴𝐷 𝑀 = 0. The positive mass theorem [SY79a][SY79b] then implies that ( Λœπ‘‹, Λœπ‘”) is the Euclidean space. The general case was solved by S.Dutta and M.Javaheri in [DJ10], where they used a totally different method. Thm1.8 serves as a compelling illustration of this principle. In the context of our theorem, as the conformal boundary becomes rounder and rounder, i.e., π‘Œ (𝑀, [𝑔]) β†— π‘Œ (Sπ‘›βˆ’1, π‘‘πœƒ2), the second inequalities in both (1.2.17) and (1.3.3) compellingly lead to 𝑄( ¯𝑋, 𝑀, [𝑔]) β†— 𝑄(B𝑛, Sπ‘›βˆ’1, 𝑑π‘₯2), representing the compactification of (H𝑛, 𝑔𝐻). Therefore, our result can be interpreted as follows: as the conformal boundary approaches the standard sphere, the interior becomes increasingly β€œclose" to the standard hyperbolic space. In the context of Thm1.5, where the second inequality in (1.2.17) is strictly satisfied except for the case of (B𝑛, 𝑑π‘₯2), the rigidity theorem can be derived from Thm1.8. The challenge lies in the fact that we still lack a complete solution to the type II Yamabe problem. I’d like to mention another result by G.Li, Q.Jie and Y.Shi [LQS14] 23 Theorem 1.11. For any πœ– > 0, 𝑛 β‰₯ 4, there exists 𝛿 > 0 so that for any PoincarΓ©-Einstein manifold (𝑋, 𝑔+), one gets |𝐾𝑔+ + 1| ≀ πœ– for all sectional curvature 𝐾, provided π‘Œ (𝑀, [ ˆ𝑔]) β‰₯ (1 βˆ’ 𝛿)π‘Œ (Sπ‘›βˆ’1, [𝑑𝑠2]). This theorem and Thm1.8 complement each other. 1.5 Some Discussions on Compact Manifolds with Boundary It is a natural question if the inequality in Theorem 1.8 holds for a compact Riemannian manifold (𝑀 𝑛, 𝑔) with 𝑅𝑖𝑐 β‰₯ βˆ’ (𝑛 βˆ’ 1) and Ξ  β‰₯ 1. We are motivated by the observation that some results for conformally compact manifolds follow from results for compact Riemannian manifolds by a limiting process. As an illustration, consider the following theorem by Lee. Theorem 1.12. (Lee [Lee94]) Let (𝑋 𝑛, 𝑔+) be a conformally compact manifold whose conformal in- finity has nonnegative Yamabe invariant. If 𝑅𝑖𝑐 (𝑔+) β‰₯ βˆ’ (𝑛 βˆ’ 1) 𝑔+ and (𝑋 𝑛, 𝑔+) is asymptotically Poincare-Einstein, then the bottom of spectrum πœ†0 (𝑋 𝑛, 𝑔+) = (𝑛 βˆ’ 1)2 /4. When the Yambabe invariant of the conformal infinity is positive, Lee’s theorem follows from the following result for compact Riemannian manifolds. Theorem 1.13. Let (𝑀 𝑛, 𝑔) be a compact Riemannian manifold with 𝑅𝑖𝑐 β‰₯ βˆ’ (𝑛 βˆ’ 1). If along the boundary Ξ£ := πœ• 𝑀 we have the mean curvature 𝐻 β‰₯ 𝑛 βˆ’ 1, then the first Dirichlet eigenvalue πœ†0 (𝑀) β‰₯ (𝑛 βˆ’ 1)2 4 . This theorem has a simple proof. Let π‘Ÿ be the distance function to Ξ£. By standard method in Riemannian geometry, we have in the support sense. A direct calculation yields Ξ”π‘Ÿ ≀ βˆ’ (𝑛 βˆ’ 1) Δ𝑒(π‘›βˆ’1)π‘Ÿ/2 ≀ βˆ’ (𝑛 βˆ’ 1)2 4 𝑒(π‘›βˆ’1)π‘Ÿ/2. 24 This implies πœ†0 (𝑀) β‰₯ (π‘›βˆ’1)2 4 (for technical details see [Wan02]). We can deduce Lee’s theorem from Theorem 1.13 when the conformal infinity has positive Yamabe invariant in the following way. As explained in Section 2, we pick a metric β„Ž on the conformal infinity with positive scalar curvature and then we have a good defining function π‘Ÿ s.t. near the conformal infinity 𝑔+ has a nice expansion (1.1.2). Then a simple calculation shows that the mean curvature of the boundary of π‘‹πœ€ := {π‘Ÿ β‰₯ πœ€} satisfies 𝐻 = 𝑛 βˆ’ 1 + π‘…β„Ž 2 (𝑛 βˆ’ 2) πœ€2 + π‘œ (cid:16) πœ€2(cid:17) . As π‘…β„Ž > 0, we have 𝐻 > 𝑛 βˆ’ 1 if πœ€ is small enough. By Theorem, πœ†0 (π‘‹πœ€) β‰₯ (π‘›βˆ’1)2 that πœ†0 (𝑋) β‰₯ (π‘›βˆ’1)2 . It follows . As the opposite inequality was known by [Maz88], we have πœ†0 (𝑋) = (π‘›βˆ’1)2 4 . 4 4 When the conformal infinity has zero Yamabe invariant, the situation is more subtle. But by an idea in Cai-Galloway[CG99], a similar argument still works (cf. [Wan02]). We now come back to Theorem 1.8. By the asymptotic expansion (1.1.2) the second fundamental form of πœ• π‘‹πœ€ satisfies Ξ + = (1 + 𝑂 (πœ€)) 𝑔+, i.e. all the principal curvatures are close to 1. This leads us to consider a compact Riemannian manifold (𝑀 𝑛, 𝑔) with 𝑅𝑖𝑐 β‰₯ βˆ’ (𝑛 βˆ’ 1) and Ξ  β‰₯ 1 on its boundary Ξ£ and ask the question whether the inequality 𝑄 (𝑀, Ξ£, 𝑔) β‰₯ 2 βˆšοΈ„ (𝑛 βˆ’ 1) (𝑛 βˆ’ 2) π‘Œ (Ξ£) if 𝑛 β‰₯ 4; (1.5.1) 𝑄 (𝑀, Ξ£, 𝑔) β‰₯ 4√︁2πœ‹ πœ’ (Ξ£) if 𝑛 = 3 holds. The answer turns out to be no in general. To construct a counter example, we consider the (1βˆ’|π‘₯|2)2 𝑑π‘₯2. For 0 < 𝑅 < 1, the hyperbolic space using the ball model B𝑛 with the metric 𝑔H = 4 Euclidean ball (cid:40) π‘₯ ∈ B𝑛 : |π‘₯|2 = (cid:41) π‘₯2 𝑖 ≀ 𝑅 𝑛 βˆ‘οΈ 𝑖=1 25 is a geodesic ball in (B𝑛, 𝑔H) and the boundary has 2nd fundamental form Ξ  = 1+𝑅2 2𝑅 𝐼. We now consider (cid:40) 𝑀 = π‘₯ ∈ B𝑛 : |π‘₯|2 = 𝑖 + π‘˜π‘₯2 π‘₯2 𝑛 ≀ 𝑅 (cid:41) , π‘›βˆ’1 βˆ‘οΈ 𝑖=1 where π‘˜ > 0 is close to 1. Then (𝑀, 𝑔H) is a compact hyperbolic manifold with boundary and on its boundary we have Ξ  β‰₯ 1 if π‘˜ is sufficiently close to 1 by continuity. Since Ξ£ with the induced metric is rotationally symmetric, it is conformally equivalent to the standard sphere Sπ‘›βˆ’1. Thus, π‘Œ (Ξ£) = π‘Œ (cid:0)Sπ‘›βˆ’1(cid:1). But when π‘˜ β‰  1, the boundary is not umbilic with respect to the Euclidean metric (cid:16)B𝑛, Sπ‘›βˆ’1(cid:17). and hence not with respect to 𝑔H either. By [Esc92a] and [Mar07], 𝑄 (𝑀, Ξ£, 𝑔H) < 𝑄 It follows that the inequality (1.5.1) is false. Therefore, for a compact Riemannian manifold (𝑀 𝑛, 𝑔) with 𝑅𝑖𝑐 β‰₯ βˆ’ (𝑛 βˆ’ 1) and Ξ  β‰₯ 1 on its boundary Ξ£, it is more subtle to estimate its type II Yamabe invariant in terms of the boundary geometry. It is an interesting question and we do not have an explicit conjecture. Let us mention that in a similar setting, namely for a compact (𝑀 𝑛, 𝑔) with 𝑅𝑖𝑐 β‰₯ 0 and Ξ  β‰₯ 1 on its boundary Ξ£, there is a well-formulated conjecture [Wan19] on the type II Yamabe invariant in terms of the boundary area. 26 CHAPTER 2 LIOUVILLE TYPE THEOREMS ON MANIFOLDS WITH LOWER CURVATURE BOUND One problem that lies in the center of geometric analysis is to understand how geometric conditions, such as curvature and fundamental forms, exert influence over the solutions of partial differential equations. In [Wan19], X.Wang proposed a conjecture that for manifolds with boundary, if the Ricci curvature is nonnegative and second fundament form is positive, then a series of elliptic PDEs doesn’t admit non-constant solutions. Throughout this chapter, we will always assume that πœ• 𝑀 = Ξ£ is connected. 2.1 Preparation X. Wang has posed the following conjecture in [Wan19]: Conjecture 2.1 (Wang’s conjecture). Let (𝑀, πœ• 𝑀 = Ξ£, 𝑔) be a compact Riemannian manifold with boundary. Suppose 𝑅𝑖𝑐 β‰₯ 0 on 𝑀, and 𝐼 𝐼 β‰₯ 1 on Ξ£ where 𝐼 𝐼 is the second fundamental form, then the following PDE Δ𝑒 = 0 πœ•π‘’ πœ•πœˆ admits no non-constant positive solution provided πœ†(π‘ž βˆ’ 1) ≀ 1 and π‘ž ≀ 𝑛 = βˆ’πœ†π‘’ + π‘’π‘ž on Ξ£π‘›βˆ’1 on 𝑀 𝑛 (2.1.1) π‘›βˆ’2 unless (𝑀, Ξ£, 𝑔) is isometric to (B𝑛, Sπ‘›βˆ’1, 𝑑π‘₯2), π‘ž = 𝑛 π‘›βˆ’2 and 𝑒 is given by π‘’π‘Ž = (cid:104) 2 𝑛 βˆ’ 2 1 βˆ’ |π‘Ž|2 1 + |π‘Ž|2|π‘₯|2 βˆ’ 2π‘₯ Β· π‘Ž (cid:105) π‘›βˆ’2 2 for some π‘Ž ∈ B𝑛. The conjecture was proposed for the following reasons. Consider the following functional πΈπ‘ž,πœ† (𝑒) = 𝑀 |βˆ‡π‘’|2dV + πœ† ∫ ∫ (∫ Ξ£ π‘’π‘ž+1dS)2/(π‘ž+1) Ξ£ 𝑒2dS 𝑠(π‘ž, πœ†) = inf π‘’βˆˆπ»1 (𝑀),𝑒≠0 𝐸 (𝑒) 27 (2.1.2) For convenience, we drop the index (π‘ž, πœ†) when it brings no confustion. By definition, 𝑠(π‘ž, πœ†) ≀ 𝐸 (1) = πœ†|Ξ£| π‘žβˆ’1 1+π‘ž Fix arbitrary 𝑒 and take derivative in the direction of 𝑣, we have πœ• πœ•π‘‘ 𝐸 (𝑒 + 𝑑𝑣)(cid:12) (cid:12)𝑑=0 = (∫ Ξ£ ∫ (cid:104)(cid:0) 𝑀 4 π‘ž+1 1 π‘’π‘ž+1) ∫ 2βŸ¨βˆ‡π‘’, βˆ‡π‘£βŸ© + 2πœ† ∫ ∫ 𝑒𝑣(cid:1) ( π‘’π‘ž+1) 2 π‘ž+1 ∫ Ξ£ Ξ£ π‘£π‘’π‘ž(cid:105) Ξ£ ∫ 𝑀 + 𝑒 βˆ’ 𝐸 (𝑒) ( π‘’π‘ž+1) 1βˆ’π‘ž π‘ž+1 ∫ Ξ£ 𝑣 (cid:0) πœ•π‘’ πœ•πœˆ π‘’π‘ž+1) 1βˆ’π‘ž 1+π‘ž π‘’π‘ž(cid:1)(cid:105) ∫ Ξ£ 𝑣Δ𝑒 + βˆ’ 2( (|βˆ‡π‘’|2 + πœ†π‘’2)) ( = 𝑀 2 (∫ Ξ£ π‘’π‘ž+1) 2 π‘ž+1 (cid:104) ∫ 𝑀 (2.1.3) (2.1.4) Suppose 𝑒 is a critical point of 𝐸, then πœ• Δ𝑒 = 0 in 𝑀 and πœ•π‘’ πœ•πœˆ + 𝑒 βˆ’ 𝐸 (𝑒)(∫ Ξ£ πœ•π‘‘ 𝐸 (𝑒 + 𝑑𝑣)(cid:12) (cid:12)𝑑=0 = 0 for all smooth 𝑣, and therefore π‘’π‘ž+1)π‘’π‘ž = 0. Since 𝐸 is invariant under scaling, we could scale 𝑒 to get rid of the coefficients before π‘’π‘ž, yielding (2.1.1). In summary, (2.1.1) arises as the Euler-Langrangian equation for (2.1.2). If π‘ž < 𝑛 π‘›βˆ’2, the trace embedding 𝐻1(𝑀) β†’ πΏπ‘ž+1(Ξ£) is compact (see theorem 6.2 chapter 2 in [Nec11], for example), thereby enabling the attainment of the minimizer denoted as π‘’π‘ž,πœ†. Let us now consider a fixed value of π‘ž0. As the parameter πœ† decreases, the weight of ∫ 𝑀 |βˆ‡π‘’|2dV becomes increasingly prominent. To ensure that π‘’π‘ž0,πœ† attains minimizer, a concomitant decrease in ∫ 𝑀 |π‘’π‘ž0,πœ†|2dV is expected. For example, we have the following lemma: Lemma 2.1.1. Let (𝑀, Ξ£, 𝑔) be a compact Riemannian manifold with boundary. Suppose 𝑠(π‘ž, πœ†) is achieved by constants for some π‘ž ≀ 𝑛 π‘›βˆ’2. Then for any πœ‡ < πœ†, 𝑠(π‘ž, πœ†) is only achieved by constants. Proof. For any fixed 𝑒 ∈ 𝐻1(𝑀) and 𝑒 β‰  0, πΈπ‘ž,πœ‡ (𝑒) is linear function in πœ‡, and therefore concave. Since 𝑠(π‘ž, πœ‡) is the infimum of concave functions, 𝑠(π‘ž, πœ‡) is also a concave function in πœ‡. Suppose 𝑠(π‘ž, πœ†) is achieved by constant, we have 𝑠(π‘ž, πœ†) = πœ†|Ξ£| π‘žβˆ’1 π‘ž+1 . We also have 𝑠(π‘ž, 0) = 0. By concavity, π‘ π‘ž,πœ‡ β‰₯ πœ‡|Ξ£| π‘žβˆ’1 π‘ž+1 for πœ‡ < πœ†. At the same time, we have π‘ π‘ž,πœ‡ ≀ πΈπ‘ž,πœ‡ (1) = πœ‡|Ξ£| 1βˆ’π‘ž 1+π‘ž . So 𝑠(π‘ž, πœ‡) is achieved by constants. 28 Suppose 𝑠(π‘ž, πœ‡) is achieved by some non-constant 𝑒 and πœ‡ < πœ†, i.e. πΈπ‘ž,πœ‡ (𝑒) = πΈπ‘ž,πœ‡ (1) = 𝑠(π‘ž, πœ‡). Since 𝑒 is non-constant, we must have ∫ π‘žβˆ’1 π‘ž+1 . Use 𝑒 ∫ Ξ£ 𝑀 |βˆ‡π‘’|2 > 0 and therefore < |Ξ£| 𝑒2 𝑒 (π‘ž+1) )2/(π‘ž+1) (∫ Ξ£ as the test function for (π‘ž, πœ†), we obtain πΈπ‘ž,πœ† (𝑒) = = Ξ£ π‘’π‘ž+1)2/(π‘ž+1) 𝑀 |βˆ‡π‘’|2 + πœ† ∫ ∫ (∫ Ξ£ 𝑀 |βˆ‡π‘’|2 + πœ‡ ∫ ∫ (∫ Ξ£ Ξ£ π‘’π‘ž+1)2/(π‘ž+1) 𝑒2 𝑒2 + (πœ† βˆ’ πœ‡) ∫ Ξ£ 𝑒2 π‘’π‘ž+1)2/(π‘ž+1) (∫ Ξ£ <πœ‡|Ξ£| π‘žβˆ’1 π‘ž+1 + (πœ† βˆ’ πœ‡)|Ξ£| π‘žβˆ’1 π‘ž+1 = 𝑠(π‘ž, πœ†) which is contradiction since we assumer 𝑠(π‘ž, πœ†) is achieved by constant. β–‘ Note that π‘’π‘ž,0 and 𝑒1,πœ† are both constants. Therefore, for values of 1 < π‘ž < 𝑛 π‘›βˆ’2, an intriguing possibility emerges: for each value of π‘ž, there might exist a threshold πœ†π‘ž such that when πœ† < πœ†π‘ž, the minimizer π‘’πœ†,π‘ž will take constant values. This phenomena was first found as Beckner inequality [Bec93]: Theorem 2.1 (Beckner’s inequality). For unit disk with Euclidean metric and 𝑦 ∈ 𝐻1(B𝑛), we have π‘žβˆ’1 π‘ž+1 π‘›βˆ’1( 𝑐 ∫ Sπ‘›βˆ’1 π‘’π‘ž+1dS) 2 π‘ž+1 ≀ 1 πœ† ∫ B𝑛 |βˆ‡π‘’|2dV + ∫ Sπ‘›βˆ’1 𝑒2dS (2.1.5) provided that 1 ≀ π‘ž ≀ 𝑛 π‘›βˆ’2 and πœ†(π‘ž βˆ’ 1) ≀ 1, where π‘π‘›βˆ’1 = 2πœ‹(π‘›βˆ’1)/2/Ξ“ (cid:0)(𝑛 βˆ’ 1)/2(cid:1) is the volume of 𝑛 βˆ’ 1 round sphere. It follows from the inequality that the minimizers of πΈπœ†,π‘ž for unit ball are exclusively realized by constant functions. This intriguingly gives rise to the conjecture that (2.1.1) admits no non-constant solutions. When πœ†(π‘ž βˆ’ 1) > 1, however, the 𝑒 ≑ 1 is no longer minimizer for unit balls. The second variation of πΈπœ†,π‘ž at 𝑒 ≑ 1 in the direction of 𝑣 is πœ•2 πœ•π‘‘2 πΈπœ†,π‘ž (1 + 𝑑𝑣)(cid:12) (cid:12)𝑑=0 = βˆ’ 2 |Ξ£|2/(π‘ž+1) ∫ (cid:104) βˆ’ ∫ (cid:0) 𝑣Δ𝑣 + Ξ£ 𝑀 |βˆ‡π‘£|2 βˆ’ πœ†(π‘ž βˆ’ 1) πœ•π‘£ + πœ†π‘£(cid:1)𝑣 βˆ’ πœ†π‘žπ‘£2(cid:105) πœ•πœˆ ∫ 2 |Ξ£|2/(π‘ž+1) We can pick 𝑣 to be the function associated to the first Steklov eigenvalue. 𝑣2(cid:105) = 𝑀 Ξ£ (cid:104) ∫ (2.1.6) 29 Definition 2.1.1. The first Steklov eigenvalue is πœ† = inf π‘’βˆˆπ»1 (𝑀),𝑒≠0 ∫ 𝑀 |βˆ‡π‘’|2 ∫ 𝑒2 Ξ£ The corresponding Euler-Lagrangian equation, or the eigenfunction equation, is Δ𝑒 = 0 in 𝑀 πœ•π‘’ πœ•πœˆ = πœ†π‘’ on Ξ£ (2.1.7) (2.1.8) It’s well known that the first Steklov eigenvalue for unit ball is 1 with 𝑛 eigenfunctions given by coordinate functions. Pick 𝑣 to be Steklov eigenfunction in (2.1.6), and we have πœ•2 πœ•π‘‘2 πΈπœ†,π‘ž (1 + 𝑑𝑣)(cid:12) (cid:12)𝑑=0 = 2(1 βˆ’ πœ†(π‘ž βˆ’ 1)) 𝑐2/(π‘ž+1) π‘›βˆ’1 ∫ Sπ‘›βˆ’1 𝑣2 As a consequence, the minimization of πΈπœ†,π‘ž by 𝑒 ≑ 1 is unsuccessful if πœ†(π‘ž βˆ’ 1) > 1. However, the minimizer exist since the trace embedding is compact. This implies that the minimizer is a non-constant solution of (2.1.1). Therefore, the condition πœ†(π‘ž βˆ’ 1) ≀ 1 is crucial and cannot be improved. These insights serve to clarify the conjecture for B𝑛. It’s noteworthy to mention that the conjecture is fully resolved for unit balls as demonstrated in [GL23]. A natural progression from here is to delve into the intricate connection between geometric properties and the behavior of solutions of (2.1.1). This exploration is driven by the question of how geometric attributes influence the solutions of PDEs. A similar problem was resolved: Theorem 2.2 (B.VΓ©ron and L.VΓ©ron [BV91]). Let (𝑀, Ξ£, 𝑔) be a compact Riemannian manifold with boundary. βˆ’Ξ”π‘’ + πœ†π‘’ = π‘’π‘ž πœ•π‘’ πœ•πœˆ = 0 on 𝑀 𝑛 on Ξ£π‘›βˆ’1 admits no non-constant solutions provided that πœ† > 0, 1 < π‘ž < 𝑛+2 π‘›βˆ’2 and 𝑅𝑖𝑐 β‰₯ (π‘›βˆ’1)(π‘žβˆ’1)πœ† 𝑛 In the context of Wang’s conjecture, there is famous Escobar conjecture. 30 (2.1.9) 𝑔. Conjecture 2.2 (Escobar conjecture). Let (𝑀, Ξ£, 𝑔) be a compact Riemannian manifold with boundary. Suppose 𝑅𝑖𝑐 β‰₯ 0 on 𝑀, and 𝐼 𝐼 β‰₯ 1 on Ξ£. Then ∫ 𝑀 |βˆ‡π‘’|2 β‰₯ ∫ Ξ£ 𝑒2 (2.1.10) i.e. the first Steklov eigenvalue is no less than 1. Under the condition of non-negative sectional curvature Escobar’s conjecture was completely solved by C.Xia and C.Xiong in ([XX19]). The insights gleaned from these findings, in conjunction with the outcomes concerning B𝑛, culmi- nate in Wang’s conjecture. If this conjecture holds true, it would lead to fascinating geometric implications. For instance, an intriguing consequence would be an upper bound on the area of Ξ£. Conjecture 2.3. Let (𝑀, Ξ£, 𝑔) be as in conjecture 2.1. Then π΄π‘Ÿπ‘’π‘Ž(Ξ£) ≀ π΄π‘Ÿπ‘’π‘Ž(Sπ‘›βˆ’1) (2.1.11) Moreover, this inequality would only be realized by unit spheres as the boundary of unit disks. . We can view this as an extension of the Bishop volume comparison theorem. Consider the πΈπ‘ž, 1 π‘žβˆ’1 . For π‘ž < 𝑛 π‘›βˆ’2 , 𝑠(π‘ž, 1 π‘žβˆ’1) can be achieved for some smooth function. If conjecture 2.1 holds true, then the only possible minimizer are constants, which implies (π‘ž βˆ’ 1) ∫ (∫ Ξ£ 𝑀 |βˆ‡π‘’|2 + ∫ Ξ£ 𝑒(π‘ž+1))2/(π‘ž+1) 𝑒2 π‘žβˆ’1 π‘ž+1 β‰₯ |Ξ£| for any 𝑒 ∈ 𝐻1(𝑀) and 𝑒 β‰  0. Let π‘ž β†˜ in above inequality and we get 𝑀 |βˆ‡π‘’|2 + ∫ ∫ 2 π‘›βˆ’2 𝑒2(π‘›βˆ’1)/(π‘›βˆ’2)) (π‘›βˆ’2)/(π‘›βˆ’1) 𝑒2 Ξ£ (∫ Ξ£ β‰₯ |Ξ£| 1 π‘›βˆ’1 (2.1.12) Recall the definition for type II Yamabe invariant in (1.2.14). Since 𝑅𝑖𝑐 β‰₯ 0 and 𝐼 𝐼 β‰₯ 𝑔|Ξ£, we get 𝑄(𝑀, Ξ£, 𝑔) β‰₯ 4(π‘›βˆ’1) π‘›βˆ’2 𝑠( 𝑛 π‘›βˆ’2, π‘›βˆ’2 2 ). By the second inequality in (1.2.17), we finally arrive at (2.1.11). 31 The relationship between curvature and volume has a long and storied history in geometry. For manifolds without boundaries, the Bishop comparison theorem is a pivotal result. It asserts that a lower bound on the Ricci curvature results in an upper bound on the volume of geodesic balls. In the context of manifolds with boundaries, where we encounter second fundamental forms, a lower bound for this form could imply an upper bound for the distance from the boundary to the interior. When coupled with the Bishop comparison method, it naturally leads us to surmise such upper bounds on volume and area. However, tackling this problem is notably challenging. The conjecture we’ve presented offers a promising avenue to approach and potentially solve this intriguing problem. A lot of work has been invested in exploring this conjecture; however, a significant portion of its components remain unsolved. As I see it, there are two primary challenges that contribute to the difficulty of addressing this conjecture. Firstly, there exists a lack of comprehensive understanding regarding how Ricci curvature influ- ences the solutions of (2.1.1). A notable advance in this direction was made in [GHW19], where it was demonstrated that by strengthening the Ricci curvature assumption to π‘ π‘’π‘π‘‘π‘–π‘œπ‘›π‘Žπ‘™ π‘π‘’π‘Ÿπ‘£π‘Žπ‘‘π‘’π‘Ÿπ‘’ β‰₯ 0, a beneficial weight function emerges. This weight function proves advantageous during inte- gration by parts, effectively nullifying bothersome boundary terms. This paves the way for the derivation of a weighted version of Reilly’s formula, leading to partial results. Unfortunately, this method falters when the assumption is relaxed to non-negative Ricci curvature. In fact, no results exist in this setting. I have obtained results in general Riemannian manifolds using different techniques. However, it’s currently unclear how Ricci curvature, or even sectional curvature, influences the estimation in my approaches. I’m working to get over these difficulties to get a uniform estimate under curvature assumptions. The second primary challenge emerges when π‘ž β†— 𝑛 π‘›βˆ’2, particularly when the equality is (Ξ£) ↩→ 𝐻1(𝑀) becomes merely continuous reached. Notably, at π‘ž = 𝑛 π‘›βˆ’2, the embedding 𝐿 2(π‘›βˆ’1) π‘›βˆ’2 32 without the compactness property, raising uncertainties about the existence of the minimizer. It’s worth emphasizing that in this instance, (2.1.2) takes on a form reminiscent of the second-type Yamabe invariant defined in the preceding section. These intricacies contribute to the heightened complexity of addressing this conjecture. Given this challenge, it might be worthwhile to concentrate on cases where π‘ž is close to 1. When we take the derivative with respect to 𝑝 at 𝑝 = 1, we obtain a log-Sobolev inequality. The verification of the log-Sobolev inequality would provide confidence in Wang’s conjecture. I’m working in this direction and partial results are obtained. The results mentioned above will be presented in the following sections. 2.2 Pseudo Differential Operator We start with an exploration of Dirichlet-to-Neumann operator. Using the property of Dirichlet- to-Neumann operator we can derive a non-existence theorem for (2.1.1) on general manifold with boundary without adding any restriction for curvature. In this section all the integration and Sobolev spaces will be with respect to Ξ£ unless stated otherwise. Definition 2.2.1. 𝐷𝑁 : 𝐻1(Ξ£) βˆ’β†’ 𝐿2(Ξ£) 𝐷𝑁 ( 𝑓 ) = πœ•π‘’ πœ•πœˆ (2.2.1) where 𝑒 is the harmonic extension for 𝑓 , i.e. Δ𝑒 = 0 in 𝑀 𝑒|Ξ£ = 𝑓 on Ξ£ It’s well known that 𝐷𝑁 is a first order elliptic pseudo differential operator. See [Tay96] chapter 1, for example. As a result 𝐢1βˆ₯βˆ‡ 𝑓 βˆ₯ 𝐿2 ≀ βˆ₯𝐷𝑁 ( 𝑓 ) βˆ₯ 𝐿2 ≀ 𝐢2βˆ₯βˆ‡ 𝑓 βˆ₯ 𝐿2 (2.2.2) Define ˜𝐿2(Ξ£) := { 𝑓 ∈ 𝐿2(Ξ£) : ∫ Ξ£ have βˆ₯ 𝑓 βˆ₯𝑙2 ≀ 𝐢 βˆ₯βˆ‡ 𝑓 βˆ₯ 𝐿2 for 𝑓 ∈ ˜𝐻1(Ξ£). Then (2.2.2) can be rewritten as 𝑓 = 0} and ˜𝐻1(Ξ£) = 𝐻1(Ξ£) ∩ 𝐴. Using PoincarΓ© lemma, we 𝐢1βˆ₯ 𝑓 βˆ₯𝐻1 ≀ βˆ₯𝐷𝑁 ( 𝑓 ) βˆ₯ 𝐿2 ≀ 𝐢2βˆ₯ 𝑓 βˆ₯𝐻1 (2.2.3) 33 for 𝑓 ∈ ˜𝐻1(Ξ£). Suppose 𝑓 ∈ 𝐻1(Ξ£) and 𝐷𝑁 ( 𝑓 ) = 0. By (2.2.2) and employing the standard bootstrapping strategy, 𝑓 must be a smooth function. Then, by the definition of 𝐷𝑁 and the maximal principle, the harmonic extension of 𝑓 remains constant, implying 𝑓 itself is constant. If we restrict to ˜𝐻1(Ξ£), then 𝑓 must identically vanish. Consequently, 𝐷𝑁 is injective when restricted to ˜𝐻1(Ξ£). Further more, 𝐷𝑁 is self-adjoint on 𝐢∞(Ξ£). Let 𝑓 , 𝑔 ∈ 𝐢∞(Ξ£) and 𝑒, 𝑣 be their harmonic extension to 𝑀 respectively. Then ∫ Ξ£ 𝐷𝑁 ( 𝑓 )𝑔 = = = ∫ Ξ£ ∫ 𝑀 ∫ Ξ£ πœ•π‘’ πœ•πœˆ 𝑔 βŸ¨βˆ‡π‘’, βˆ‡π‘£βŸ© πœ•π‘£ πœ•πœˆ 𝑓 = ∫ Ξ£ 𝑓 𝐷𝑁 (𝑔) (2.2.4) Let me introduce theorem 5.5 in chapter 3 of [LM90]. Theorem 2.3. Let 𝐸 be a hermitian vector bundle with connection over a compact Riemannian manifold, Ξ“(𝐸) the smooth sections for 𝐸. Suppose 𝑃 : Ξ“(𝐸) β†’ Ξ“(𝐸) is elliptic and self-adjoint, then there is an 𝐿2-orthogonal direct sum decomposition: Ξ“(𝐸) = π‘˜π‘’π‘Ÿ (𝑃) βŠ• πΌπ‘š(𝑃) The statement is for hermitian bundles, but the argument works for real bundles with inner product structure as well. We already found the kernel of 𝐷𝑁 is given by constant R. Given 𝑔 ∈ 𝐢∞(Ξ£), 𝑔, which lies in ˜𝐿2(Ξ£). By the theorem above, 𝐷𝑁 is its 𝐿2-orthogonal projection to R is 𝑔 βˆ’ 1 Ξ£ surjective from 𝐢∞(Ξ£) ∩ ˜𝐿2(Ξ£) β†’ 𝐢∞(Ξ£) ∩ ˜𝐿2(Ξ£), and therefore bijective. Now assume 𝑔 ∈ ˜𝐿2(Ξ£) which doesn’t have to be smooth, then there exists a sequence {𝑔𝑖} ∈ 𝐢∞(Ξ£) so that 𝑔𝑖 β†’ 𝑔 in 𝐿2(Ξ£). We can further assume that ∫ Ξ£ 𝑔𝑖 = 0 by taking 𝑔𝑖 βˆ’ 1 |Ξ£| 𝑔𝑖 instead. Then ∫ Ξ£ ∫ Ξ£ there exists 𝑓𝑖 so that 𝐷𝑁 ( 𝑓𝑖) = 𝑔𝑖. Use the first inequality in (2.2.3), { 𝑓𝑖} is a Cauchy sequence and therefore converges to some 𝑓 ∈ 𝐻1(Ξ£). It easy to see that 𝐷𝑁 ( 𝑓 ) = 𝑔, and thus 𝐷𝑁 is actually surjective. Combine everything above, and we arrive at 34 Lemma 2.2.1. When viewed as a map from 𝐻1(Ξ£) to 𝐿2(Ξ£), the image of 𝐷𝑁 is ˜𝐿2(Ξ£), and its kernel is R. Furthermore, 𝐷𝑁 is self-adjoint when restricted to 𝐢∞(Ξ£). For the Laplace equation, the existence of Green’s function is a key tool for solving the equation. Given a compact manifold Ξ£ without boundary, there exists a unique function 𝐺 (π‘₯, 𝑦) satisfying Δ𝑦𝐺 (π‘₯, 𝑦) = 𝛿π‘₯ (𝑦). Consequently for any 𝑓 ∈ 𝐿2(Ξ£) and ∫ Ξ£ 𝑓 = 0, 𝑒(π‘₯) := ∫ Ξ£ 𝐺 (π‘₯, 𝑦) 𝑓 (𝑦)𝑑𝑦 solves Δ𝑒 = 𝑓 . In the context of Dirichlet-to-Neumann operator, a comparable kernel is anticipated. For 𝑒 ∈ 𝐢∞(Ξ£) ∩ ˜𝐿2, define 𝑇 (𝑒) = 𝐷𝑁 βˆ’1(𝑒 βˆ’ 1 ∫ |Ξ£| Ξ£ 𝑒) (2.2.5) It defines a 𝑇 is well-defined by lemma 2.2.1, and 𝑇 (𝑒) is also 𝐢∞(Ξ£), and thus in D (Ξ£). bilinear form 𝐡(𝑒, 𝑣) = ∫ Ξ£ 𝑇 (𝑒)𝑣. Obviously 𝐡 satisfies the conditions in theorem 2.2.15 (explicit formulation will be given at the end of this section), and therefore there exists a kernel 𝐾 ∈ D (Σ×Σ) such that 𝐡(𝑒, 𝑣) = ∫ Ξ£ 𝑇 (𝑒)(π‘₯)𝑣(π‘₯)𝑑π‘₯ = ∫ Σ×Σ 𝑒(π‘₯)𝑣(𝑦)𝐾 (π‘₯, 𝑦)𝑑π‘₯𝑑𝑦 for any 𝑒, 𝑣 ∈ 𝐢∞(Ξ£). Since 𝐷𝑁 is a first order elliptic pseudo-differential operator, 𝑇 is elliptic of order βˆ’1, and we have the following estimate for 𝐾 from chapter 1, section 2 in [Tay96] Lemma 2.2.2. 𝐾 is 𝐢∞ off the diagonal in Ξ£ Γ— Ξ£, and |𝐾 | ≀ 𝐢𝑑 (π‘₯, 𝑦)2βˆ’π‘› (2.2.6) Ξ£ has dimention 𝑛 βˆ’ 1, and therefore ∫ Σ×Σ |𝑒(π‘₯)𝑣(𝑦)𝐾 (π‘₯, 𝑦)|𝑑π‘₯𝑑𝑦 < ∞. Therefore we could apply Fubini theorem ∫ Ξ£ 𝑇 (𝑒)(π‘₯)𝑣(π‘₯)𝑑π‘₯ = ∫ Ξ£ 𝑣(𝑦) (cid:0) ∫ Ξ£ 𝑒(π‘₯)𝐾 (π‘₯, 𝑦)𝑑π‘₯(cid:1) 𝑑𝑦 Fix 𝑒 and view 𝑣 as the test function, we have 𝑇𝑒(π‘₯) = ∫ Ξ£ 𝐾 (π‘₯, 𝑦)𝑒(𝑦)𝑑𝑦 35 i.e. 𝑒 βˆ’ 1 ∫ |Ξ£| Ξ£ 𝑒 = ∫ Ξ£ 𝐾 (π‘₯, 𝑦)𝐷𝑁 (𝑒) (𝑦)𝑑𝑦 (2.2.7) Next we are going to use this expression to prove a non-existence theorem. For convenience, we scale 𝑒 so that ∫ Ξ£ π‘’π‘ž+1 = 1, and (2.1.1) becomes on 𝑀 𝑛 = βˆ’πœ†π‘’ + 𝑠(π‘ž, πœ†)π‘’π‘ž on Ξ£π‘›βˆ’1 (2.2.8) Δ𝑒 = 0 πœ•π‘’ πœ•πœˆ π‘’π‘ž+1 = 1 ∫ Ξ£ Use (2.2.7) for (2.2.8), (cid:0)𝑒 βˆ’ 1 ∫ Ξ£ Ξ£ 𝑒(π‘₯)(cid:1) = ∫ Ξ£ 𝐾 (π‘₯, 𝑦) (βˆ’πœ†π‘’ + π‘ π‘’π‘ž) (𝑦)𝑑𝑦 (2.2.9) By (2.2.6), 𝐾 (π‘₯, Β·) is 𝐿 𝑝 for any 𝑝 < π‘›βˆ’1 π‘›βˆ’2. Since Ξ£ is compact, we can find a 𝐢 independant of π‘₯ such that βˆ₯𝐾 (π‘₯, Β·)βˆ₯ 𝑝 ≀ 𝐢, βˆ€π‘₯ ∈ Ξ£ In the remaining of this section, 𝐢 is a constant that depends on metric and π‘ž, but not πœ†. Let π‘βˆ— = 𝑝 π‘βˆ’1 be the conjugate of 𝑝, and we have π‘βˆ— > 𝑛 βˆ’ 1. Then by HΓΆlder inequality, the left hand side of (2.2.9) can be bounded by 𝐿𝐻𝑆 ≀ 𝐢 βˆ₯ βˆ’ πœ†π‘’ + π‘ π‘’π‘ž βˆ₯ π‘βˆ— ≀ πΆπœ†(βˆ₯𝑒βˆ₯ π‘βˆ— + βˆ₯π‘’π‘ž βˆ₯ π‘βˆ—) (2.1.3) is used for the second inequality. Let 𝑀 = sup 𝑒, and 0 < 𝑑 < 1. 𝐿𝐻𝑆 ≀ πΆπœ†π‘€ 𝑑 (βˆ₯𝑒1βˆ’π‘‘ βˆ₯ π‘βˆ— + βˆ₯π‘’π‘žβˆ’π‘‘ βˆ₯ π‘βˆ—) (2.2.10) We aim to bound the right-hand side by ∫ Ξ£ π‘’π‘ž+1 = |Ξ£|. By HΓΆlder’s inequality, this is achievable when (π‘ž βˆ’ 𝑑) π‘βˆ— ≀ π‘ž + 1, which is equivalent to π‘ž ≀ 𝑑 π‘βˆ—+1 π‘βˆ—βˆ’1 . Given that π‘ž < 𝑛 π‘›βˆ’2, we can choose 0 < 𝑑 < 1 and π‘βˆ— > 𝑛 βˆ’ 1 to meet the requirement. 36 For the left hand side, we might take π‘₯ to be the maximal point for 𝑒. Again, by HΓΆlder inequality and ∫ Ξ£ π‘’π‘ž+1 = |Ξ£|, we have 1 |Ξ£| ∫ Ξ£ 𝑒 ≀ 1. Together, we have 𝑀 βˆ’ 1 ≀ πΆπœ†π‘€ 𝑑 Since 𝑑 < 1, we arrive at the following Lemma 2.2.3. Let 𝑒 be a solution of (2.2.8). Then we have βˆ₯𝑒βˆ₯∞ ≀ 𝐢 (𝑀, 𝑛, π‘ž) (2.2.11) provided that πœ† < 1, where 𝑁 is a constant that depends on 𝑛, π‘ž and 𝐢 (𝑀, 𝑛, π‘ž) depends on the 𝑛, π‘ž and Riemannian manifold 𝑀. Based on this 𝐿∞ estimate, we can prove a non-existence theorem: Theorem 2.4. For each 1 < π‘ž < 𝑛 π‘›βˆ’2, there exists πœ†π‘ž so that (2.2.8) only admits constant solutions for πœ† ≀ πœ†0. As a consequence, for these πœ†β€™s π‘ πœ†,π‘ž = πœ†π΄(Ξ£) π‘žβˆ’1 π‘ž+1 We start with a lemma Lemma 2.2.4. Let 𝑒 be a harmonic function on (𝑀, Ξ£, 𝑔), then where πœ‡ is the first Steklov eigenvalue. ∫ 𝑀 |βˆ‡π‘’|2 ≀ πœ‡ ∫ Ξ£ πœ•π‘’ πœ•π‘› | |2 (2.2.12) Proof. Note that the inequality above is invariant under translation, so it suffices to prove the case ∫ 𝑀 𝑒 = 0. ∫ 𝑀 |βˆ‡π‘’|2 = ≀ ≀ ∫ 𝑒 πœ•π‘’ πœ•π‘› Ξ£ ∫ πœ– 2 πœ– 2πœ‡ Ξ£ ∫ 𝑒2 + 1 2πœ– |βˆ‡π‘’|2 + ∫ | Ξ£ |2 πœ•π‘’ πœ•π‘› ∫ 1 2πœ– πœ•π‘’ πœ•π‘› | |2 Ξ£ 𝑀 37 Therefore (2πœ– βˆ’ πœ– 2 πœ‡ ) ∫ 𝑀 |βˆ‡π‘’|2 ≀ ∫ Ξ£ πœ•π‘’ πœ•π‘› | |2 The infimum of the quadratic on the left hand side is achieved for πœ– = πœ‡ and, and the lemma follows. Proof of the theorem ∫ πœ‡ 𝑀 |βˆ‡π‘’|2 ≀ = = β–‘ ∫ Ξ£ ∫ Ξ£ ∫ πœ•π‘’ πœ•π‘› | |2 (π‘ π‘’π‘ž βˆ’ πœ†π‘’) πœ•π‘’ πœ•π‘› (π‘ π‘žπ‘’π‘žβˆ’1 βˆ’ 1)|βˆ‡π‘’|2 (2.2.13) 𝑀 ∫ ≀ πœ† (π‘žπ‘’π‘žβˆ’1 βˆ’ 1)|βˆ‡π‘’|2 ≀ πΆπœ† 𝑀 ∫ 𝑀 |βˆ‡π‘’|2 Therefore, if πœ† is small, 2.1.2 admit no non-constant minimizer. β–‘ In Theorem 2.4, we investigated the solutions of (2.2.8), which arises as the minimizer for the functional. The same method can be applied to examine solutions for (2.1.1), which are not necessarily minimizers, but the trade-off is that π‘ž < π‘›βˆ’1 π‘›βˆ’2. Theorem 2.5. For each 1 < π‘ž < π‘›βˆ’1 π‘›βˆ’2, there exists πœ†π‘ž so that the equation only admits constant solutions for πœ† ≀ πœ†0. Proof. The method is similar, and I will only show the different parts. Recall that for equation (2.2.8) we have ∫ Ξ£ π‘’π‘ž+1 = |Ξ£| and we can control the right hand side of estimate (2.2.10). But for (2.1.1) we have to derive such an estimate. Integrate (2.1.1) by parts and we have By HΓΆlder inequality 0 = ∫ 𝑀 Δ𝑒 = ∫ Ξ£ βˆ’πœ†π‘’ + π‘’π‘ž+1 ∫ π‘’π‘ž = πœ† ∫ 𝑒 ≀ πœ†( π‘’π‘ž) 1 π‘ž |Sπ‘›βˆ’1| π‘žβˆ’1 π‘ž π‘’π‘ž ≀ πœ† π‘ž π‘žβˆ’1 |Sπ‘›βˆ’1| 38 ∫ ∫ β‡’ (2.2.14) Now run the method for the previous theorem, the estimate (2.2.10) becomes 𝐿𝐻𝑆 ≀ 𝐢 (πœ† + 1) (βˆ₯𝑒1βˆ’π‘‘ βˆ₯ π‘βˆ— + βˆ₯π‘’π‘žβˆ’π‘‘ βˆ₯ π‘βˆ—) where 0 < 𝑑 < 1 and π‘βˆ— > 𝑛 βˆ’ 1. We aim to bound the right-hand side by ∫ Ξ£ 𝑑 π‘βˆ— π‘βˆ—βˆ’1, which implies π‘ž < π‘›βˆ’1 π‘›βˆ’2. (π‘ž βˆ’ 𝑑) π‘βˆ— ≀ π‘ž. This is feasible if π‘ž ≀ π‘’π‘ž, which requires β–‘ Remark 2.2.1. In theorem 2.4, we initiate with an πΏπ‘ž+1 bound, while in theorem 2.5, we can only derive an πΏπ‘ž estimate. This is the rationale behind assuming π‘ž < π‘›βˆ’1 π‘›βˆ’2 instead of π‘ž < 𝑛 π‘›βˆ’2. To end this section, I will give explicit formulation of Schwartz kernel theorem. Let 𝑀 be two compact Riemannian manifolds. We can define the following seminorms on 𝐢∞(𝑀) by |𝑒|π‘˜ := sup π‘₯βˆˆπ‘€ βˆ‘οΈ π›Όβ‰€π‘˜ |βˆ‡π›Όπ‘’(π‘₯)| These seminorms give topology to 𝐢∞(𝑀). A linear map 𝑇 from 𝐢∞(𝑀) to R is continuous provided that there exists some 𝐢 and π‘˜ for all 𝑒 ∈ 𝐢∞(𝑀). Let D denote the space of distribution on 𝑀, i.e. all the continuous maps in 𝑇 (𝑒) ≀ 𝐢 |𝑒|π‘˜ the sense as above. Suppose there is another Riammnian manifold 𝑁, and a map 𝑇 : 𝐢∞(𝑀) β†’ D (𝑁) 𝑇𝑒 is a continuous operator on 𝐢∞(𝑁), and thus giving rise to a bilinear form 𝐡 by the following: 𝐡 : 𝐢∞(𝑀) Γ— 𝐢∞(𝑁) β†’ R 𝐡(𝑒, 𝑣) = βŸ¨π‘‡π‘’, π‘£βŸ©, 𝑒 ∈ 𝐢∞(𝑀), 𝑣 ∈ 𝐢∞(𝑁) Finally, define 𝑒 βŠ— 𝑣 ∈ 𝐢∞(𝑀 Γ— 𝑁) by 𝑒 βŠ— 𝑣(π‘₯, 𝑦) := 𝑒(π‘₯)𝑣(𝑦), π‘₯ ∈ 𝑀, 𝑦 ∈ 𝑁 Given all these preparations, we have the Schwartz kernel theorem 39 Theorem 2.6 (Schwartz kernel theorem). For any 𝐡 as in above, there exists a distribution 𝐾 ∈ D (𝑀 Γ— 𝑁) so that for 𝑒 ∈ 𝐢∞(𝑀) and 𝑣 ∈ 𝐢∞(𝑁) we have 𝐡(𝑒, 𝑣) = βŸ¨π‘’ βŠ— 𝑣, 𝐾⟩ (2.2.15) 2.3 Bootstrapping Strategy The 𝐿∞ estimate (2.2.11) can also be derived using standard bootstrapping strategy, and it’s more straightforward. In this section (π‘ž, πœ†) will dropped for 𝑠(π‘ž, πœ†) and πΈπ‘ž,πœ† (𝑒) for simplicity. 𝐢 will be a constant that does not depend on πœ† or π‘ž and might change from line to line. All the integral and norms will be on the boundary Ξ£. We start from π‘₯0 = π‘ž + 1, and choose π‘₯π‘˜ inductively by If 𝑒 ∈ 𝐿π‘₯π‘˜ for π‘₯π‘˜ β‰₯ π‘ž + 1, then πœ•π‘’ βˆ’ 1 𝑛 βˆ’ 1 (2.3.1) π‘₯π‘˜ π‘ž . By HΓΆlder inequality we have = 1 π‘ž π‘₯π‘˜+1 π‘₯π‘˜ πœ•π‘› = βˆ’πœ†π‘’ + π‘ π‘’π‘ž ∈ 𝐿 πœ•π‘’ πœ•π‘› ≀ πœ†βˆ₯𝑒βˆ₯ π‘₯π‘˜ βˆ₯ π‘₯π‘˜ π‘ž π‘ž βˆ₯ + 𝑠βˆ₯π‘’π‘ž βˆ₯ π‘₯π‘˜ π‘ž ≀ πœ†|Ξ£| π‘ž π‘₯π‘˜ βˆ’ 1 π‘₯π‘˜ βˆ₯𝑒βˆ₯π‘₯π‘˜ + 𝑠βˆ₯𝑒βˆ₯ π‘ž π‘₯π‘˜ ≀ πΆπœ†(βˆ₯𝑒βˆ₯π‘₯π‘˜ + βˆ₯𝑒βˆ₯ π‘ž π‘₯π‘˜ ) We used (2.1.3) in the third line. Since π‘ž ≀ 𝑛 π‘ž π‘₯π‘˜ π‘›βˆ’2, π‘₯π‘˜ is increasing, and π‘₯π‘˜ β‰₯ π‘ž + 1. Therefore, is bounded from both below and above, and that’s why the constant 𝐢 in the third line can βˆ’ 1 π‘₯π‘˜ be made independent of π‘ž. Since Dirichlet-to-Neumann operator is elliptic of order 1 (see chapter 1 of [Tay96], for example), we have βˆ₯𝑒βˆ₯ π‘₯π‘˜ π‘ž ,1 ≀ 𝐢 (βˆ₯ πœ•π‘’ πœ•π‘› βˆ₯ π‘₯π‘˜ π‘ž + βˆ₯𝑒βˆ₯ π‘₯π‘˜ π‘ž ) ≀ 𝐢 (πœ† + 1) (βˆ₯𝑒βˆ₯π‘₯π‘˜ + βˆ₯𝑒βˆ₯ π‘ž π‘₯π‘˜ ) By Sobolev embedding theorem on the boundary and our choice of π‘₯π‘˜ , we hav βˆ₯𝑒βˆ₯π‘₯π‘˜+1 ≀ 𝐢 βˆ₯𝑒βˆ₯π‘₯π‘˜/π‘ž,1 ≀ 𝐢 (cid:0)βˆ₯𝑒βˆ₯π‘₯π‘˜ + βˆ₯𝑒βˆ₯ (cid:1) π‘ž π‘₯π‘˜ (2.3.2) We used the assumption that πœ† < 1 in the second inequality. Note that the constant 𝐢 might change for different π‘˜. But we are only taking finite bootstripe steps, so we can pick a universal constant 𝐢. 40 Lemma 2.3.1. The sequence π‘₯π‘˜ will be negative in 𝐾 (𝑛, π‘ž) steps, and 𝐾 (𝑛, π‘ž) depends only on the dimension and an upper bound for π‘ž. This will imply an 𝐿∞ bound by Sobolev inequality. Proof of lemma: Let π‘¦π‘˜ = 1 π‘₯π‘˜ and rewrite 2.3.1 as π‘¦π‘˜+1 βˆ’ π‘¦π‘˜ = (π‘ž βˆ’ 1)π‘¦π‘˜ βˆ’ 1 𝑛 βˆ’ 1 (2.3.3) which implies π‘¦π‘˜+1 < π‘¦π‘˜ if π‘¦π‘˜ < the inequality. So by induction π‘¦π‘˜ < 1 (π‘žβˆ’1)(π‘›βˆ’1) 1 (π‘žβˆ’1)(π‘›βˆ’1) . By our assumption that π‘ž < 𝑛 π‘›βˆ’2 we see that 𝑦0 satisfy and π‘¦π‘˜+1 < π‘¦π‘˜ βˆ€π‘˜ > 0. We need to calculate how many steps it take so that π‘¦π‘˜ < 0. Again from (2.3.3) π‘¦π‘˜+1 βˆ’ π‘¦π‘˜ is decreasing, which means that π‘¦π‘˜ is decreasing faster and faster. So it takes at most 𝑦0 𝑦0 βˆ’ 𝑦1 = 1 π‘ž+1 π‘›βˆ’(π‘›βˆ’2)π‘ž (π‘ž+1)(π‘›βˆ’1) = 𝑛 βˆ’ 1 𝑛 βˆ’ (𝑛 βˆ’ 2)π‘ž steps to make π‘₯π‘˜ negative. Let 𝐾 (𝑛, π‘ž) be the least integer larger than what we want in the lemma. Use (2.3.2) and do induction, we have (2.3.4) π‘›βˆ’1 π‘›βˆ’(π‘›βˆ’2)π‘ž , and this 𝐾 (𝑛, π‘ž) is β–‘ βˆ₯𝑒βˆ₯π‘₯π‘˜+2 ≀ 𝐢 (βˆ₯𝑒βˆ₯π‘₯π‘˜+1 + βˆ₯𝑒βˆ₯ (cid:16) ≀ 𝐢 (βˆ₯𝑒βˆ₯π‘₯π‘˜ + βˆ₯𝑒βˆ₯ π‘ž π‘₯π‘˜+1) π‘ž π‘₯π‘˜ ) + (βˆ₯𝑒βˆ₯π‘₯π‘˜ + βˆ₯𝑒βˆ₯ π‘ž π‘₯π‘˜ )π‘ž(cid:17) (cid:16) ≀ 𝐢 (βˆ₯𝑒βˆ₯π‘₯π‘˜ + βˆ₯𝑒βˆ₯ π‘ž π‘₯π‘˜ ) + 2π‘ž (βˆ₯𝑒βˆ₯ π‘ž π‘₯π‘˜ + βˆ₯𝑒βˆ₯ (cid:17) π‘ž2 π‘₯π‘˜ ) (2.3.5) ≀ 𝐢 (βˆ₯𝑒βˆ₯π‘₯π‘˜ + βˆ₯𝑒βˆ₯ π‘ž2 π‘₯π‘˜ ) Β· Β· Β· ≀ 𝐢 (βˆ₯𝑒βˆ₯π‘₯0 + βˆ₯𝑒βˆ₯ π‘žπ‘˜+2 π‘₯0 ) where in the third line we used the inequality (π‘Ž + 𝑏)π‘ž ≀ 2π‘ž (π‘Žπ‘ž + π‘π‘ž). βˆ₯𝑒βˆ₯∞ ≀ 𝐢 (𝑀, 𝑛, π‘ž) βˆ₯𝑒βˆ₯π‘₯0 (2.3.6) 41 Remark 2.3.1. Note that if we π‘ž is bounded away from 𝑛 π‘›βˆ’2, then from (2.3.4) we can have a uniform bound for 𝐾 that doesn’t depend on π‘ž. However, the constant 𝐢 in (2.3.2) does depends on π‘ž and we fails to get a universal estimate. If we could find a universal constant, then (2.3.6) becomes βˆ₯𝑒βˆ₯∞ ≀ πΆπœ†π‘ βˆ₯𝑒βˆ₯π‘₯0 where both 𝑁 and 𝐢 are independent of π‘ž. And (2.2.13) is ∫ πœ‡ 𝑀 |βˆ‡π‘’|2 ≀ πœ† ∫ 𝑀 (π‘žπ‘’π‘žβˆ’1 βˆ’ 1)|βˆ‡π‘’|2 ≀ πΆπœ†(π‘ž(πΆπœ†) 𝑁 (π‘žβˆ’1) βˆ’ 1) ∫ 𝑀 |βˆ‡π‘’|2 We will be able to track how the critical πœ†π‘ž changes with π‘ž, and it can easily seen that πœ†π‘ž β†’ ∞ as π‘ž β†’ 1. Remark 2.3.2. One might inquire whether the bounds established in these two sections can be made universal when Ricci curvature and the second fundamental form are bounded below. Our interest lies in understanding how geometric conditions impact the solutions of PDEs. However, unlike the Laplacian operator, obtaining estimates for the Dirichlet-to-Neumann operator (𝐷𝑁) proves challenging. For instance, a comparison theorem for the heat kernel in terms of Ricci curvature is established in [CY81] and [LY86]. Under Ricci curvature restrictions, both lower and upper bounds for eigenvalues of the Laplacian operator can be derived. For further details, see Chapter 3 of [SY94] or [Li12]. However, these methods cannot be directly extended to the Dirichlet-to-Neumann operators. The lack of knowledge regarding how geometric conditions affect Dirichlet-to-Neumann operators poses a significant challenge in Conjecture 2.1. 2.4 Estimate of the Infimum In the previous two sections we derived 𝐿∞ estimate and then non-existence theorem for (2.2.8). Note that (2.2.8) comes from the Euler-Lagrangian equation of functional (2.1.2, so the non-existence theorem gives us 𝑠(π‘ž, πœ†) = πœ†|Ξ£| 1βˆ’π‘ž 1+π‘ž for certain (π‘ž, πœ†)’s. If we closely examine the 42 functional, we could get better estimate. Let (𝑀, Ξ£, 𝑔) be an arbitrary manifold with boundary. Throughout this section βˆ‡ and βˆ‡ denote gradient on 𝑀 and Ξ£ respectively. Integration without lower indices denotes integration on Ξ£. We begin with a lemma: Lemma 2.4.1. For 1 < π‘ž ≀ π‘ž0 < 𝑛+1 π‘›βˆ’1, and ∫ 𝑓 π‘ž+1 = |Ξ£|, there exist a constant 𝐢 that only depends on the metric and π‘ž0 so that ∫ Ξ£ 𝑓 2π‘ž (log 𝑓 )2 ≀ 𝐢 ∫ Ξ£ |βˆ‡ 𝑓 |2 (2.4.1) Proof. We first show that π‘₯2π‘ž (log π‘₯)2 ≀ 𝐢 ((π‘₯ βˆ’1)2 + |π‘₯ βˆ’1|2π‘ž1) where π‘ž1 is chosen to lie in (π‘ž0, 𝑛+1 π‘›βˆ’1) and 𝐢 only depends on π‘ž0. This could be seen by looking into the following three cases. 𝑖)π‘₯ ∈ (0, 𝑖𝑖)π‘₯ ∈ [ 1 2 ) : π‘₯2π‘ž (log π‘₯)2 is bounded above and (π‘₯ βˆ’ 1)2 is bounded below; 1 2 , 2] : π‘₯2π‘ž is bounded above and (log π‘₯)2 is bounded by (π‘₯ βˆ’ 1) 𝑖𝑖𝑖)π‘₯ ∈ (2, ∞) : π‘₯2π‘ž (log π‘₯)2 ≀ π‘₯2π‘ž0 (log π‘₯)2 and therefore uniformly bounded by |π‘₯ βˆ’ 1|2π‘ž1. Then we need to bound ∫ ( 𝑓 βˆ’ 1)2 and ∫ | 𝑓 βˆ’ 1|2π‘ž1 by ∫ Ξ£ to β€œmodify" the power for ∫ | 𝑓 βˆ’ 1|2π‘ž1. Namely, use πœƒ = 1 π‘ž1 |βˆ‡ 𝑓 |2. We might apply HΓΆlder inequality in HΓΆlder inequality: ∫ | 𝑓 βˆ’ 1|2π‘ž1 ≀ βˆ₯ 𝑓 βˆ’ 1βˆ₯ 2π‘ž1πœƒ π‘ž2 βˆ₯ 𝑓 βˆ’ 1βˆ₯ 2(1βˆ’πœƒ)π‘ž1 2 ≀ 𝐢 βˆ₯ 𝑓 βˆ’ 1βˆ₯2 π‘ž2 (2.4.2) π‘›βˆ’3 . This is where we use the assumption that π‘ž0 < 𝑛+1 where π‘ž2 = 2 π‘›βˆ’1. So π‘ž2 is strictly 2βˆ’π‘ž1 below the Sobolev conjugate. The last inequality follows from ∫ 𝑓 π‘ž+1 = |Ξ£| and HΓΆlder inequality < 2(π‘›βˆ’1) again, and it’s easy to see that 𝐢 can be chosen independent of π‘ž < π‘ž1. Now it suffices to show that βˆ₯ 𝑓 βˆ’ 1βˆ₯π‘ž2 ≀ 𝐢 βˆ₯βˆ‡ 𝑓 βˆ₯2 for 𝑓 satisfying ∫ 𝑓 π‘ž+1 = 𝐴(Ξ£). This comes from a generalized PoincarΓ¨ ∫ 𝑒 defined as the average of 𝑒 over Ξ£, then inequality. Let ⨏ 𝑒 := 1 |Ξ£| ⨏ βˆ₯ 𝑓 βˆ’ ( 𝑓 π‘Ÿ) 1 π‘Ÿ βˆ₯π‘ž2 ≀ 𝐢 βˆ₯βˆ‡ 𝑓 βˆ₯2 (2.4.3) for π‘Ÿ < π‘ž0 + 1 < 2(π‘›βˆ’1) π‘›βˆ’3 and 𝐢 only depends on π‘ž1 and π‘ž2.(In our case and π‘ž2 depends on π‘ž0, so 𝐢 only depends on π‘ž0). The proof is by contradiction and modified from the standard 43 ⨏ proof. Suppose (2.4.3) is not true, and we can find a sequence π‘Ÿπ‘– and 𝑓𝑖 so that βˆ₯βˆ‡ 𝑓𝑖 βˆ₯2 β†’ 0 and βˆ₯ 𝑓𝑖 βˆ’ ( converges in πΏπ‘ž2. Since βˆ₯βˆ‡ 𝑓𝑖 βˆ₯2 β†’ 0, the sequence 𝑓𝑖 βˆ’ ( 𝑓 π‘Ÿπ‘– 𝑖 )1/π‘Ÿπ‘– βˆ₯π‘ž2 = 1. By compactness, we might pick a subsequence so that 𝑓𝑖 βˆ’ ( 𝑓 π‘Ÿπ‘– 𝑖 )1/π‘Ÿπ‘– 𝑓 π‘Ÿπ‘– 𝑖 )1/π‘Ÿπ‘– converges in 𝐻1 to a constant function. So we might write 𝑓𝑖 = π‘Žπ‘– + β„Žπ‘– where π‘Žπ‘– are constants and β„Žπ‘– β†’ 0 in 𝐻1. As a β„Žπ‘Ÿπ‘– 𝑖 )1/π‘Ÿπ‘– by triangle inequality and it follows that 𝑖 )1/π‘Ÿπ‘– βˆ₯π‘ž2 = 1. ∫ | 𝑓 βˆ’ 1|2 𝑓 π‘Ÿπ‘– β–‘ 𝑓 π‘Ÿπ‘– 𝑖 )1/π‘Ÿπ‘– ≀ π‘Žπ‘– + ( β„Žπ‘Ÿπ‘– 𝑖 )1/π‘Ÿπ‘– . This contradicts with βˆ₯ 𝑓𝑖 βˆ’ ( β„Žπ‘Ÿπ‘– 𝑖 )1/π‘Ÿπ‘– ≀ ( 𝑓 π‘Ÿπ‘– 𝑖 )1/π‘Ÿπ‘– | ≀ |β„Žπ‘– | + ( can be estimated in a similar way. result, π‘Žπ‘– βˆ’ ( ⨏ | 𝑓𝑖 βˆ’ ( ⨏ ⨏ ⨏ ⨏ ⨏ ⨏ ⨏ In the lemma above, the base point is 1 and we measured distance from 𝑓 to 1. That’s why we have the three cases in the proof above. In order to apply this lemma we need a different normalization from (2.2.8). on 𝑀 𝑛 = βˆ’πœ†π‘’ + π‘Ÿ (π‘ž, πœ†)π‘’π‘ž on Ξ£π‘›βˆ’1 (2.4.4) Δ𝑒 = 0 πœ•π‘’ πœ•πœˆ π‘’π‘ž+1 = |Ξ£| ∫ Ξ£ If we assume 𝑒 is a minimzer for 𝑠(π‘ž, πœ†), multiply this equation by 𝑒 and integrate by parts, it’s easy to see that 𝑠(π‘ž, πœ†) = π‘Ÿ (π‘ž, πœ†)|Ξ£| π‘žβˆ’1 π‘ž+1 and thus πœ† β‰₯ π‘Ÿ (π‘ž, πœ†) (2.4.5) Theorem 2.7. For 1 < π‘ž ≀ π‘ž0 < 𝑛+1 π‘›βˆ’1, there exists a constant 𝐢 depending only on 𝑛, π‘ž0 and the metric so that 𝑠(π‘ž, πœ†) = πœ†|Ξ£| 1βˆ’π‘ž 1+π‘ž provided πœ†(π‘ž βˆ’ 1) < 𝐢 (2.4.6) This method only works for π‘ž < 𝑛+1 π‘›βˆ’1. The case π‘ž < 𝑛 π‘›βˆ’2 will be dealt in the next section from a different viewpoint. We’ll need the following lemma: 44 Lemma 2.4.2 (Pohozaev ideneity). Let (𝑀, Ξ£, 𝑔) be a compact Riemannian manifold with bound- ary, and 𝑔′ = 𝑔|Ξ£. Suppose 𝑒 is a smooth function and 𝑋 is a smooth vector field. Then ∫ 𝑀 βŸ¨βˆ‡βˆ‡π‘’ 𝑋, βˆ‡π‘’βŸ© βˆ’ 1 2 |βˆ‡π‘’|2div𝑔 𝑋 + (𝑋𝑒)Δ𝑒 = ∫ Ξ£ πœ•π‘’ πœ•π‘› ( βŸ¨π‘‹, βˆ‡π‘’βŸ© βˆ’ 1 2 |βˆ‡π‘’|2βŸ¨π‘‹, (cid:174)π‘›βŸ©) (2.4.7) Proof. By direct calculation div(𝑋𝑒)βˆ‡π‘’ = βŸ¨βˆ‡βˆ‡π‘’ 𝑋, βˆ‡π‘’βŸ© + βŸ¨π‘‹, βˆ‡βˆ‡π‘’βˆ‡π‘’βŸ© + (𝑋𝑒)Δ𝑒 div( 1 2 |βˆ‡π‘’|2𝑋) = 1 2 |βˆ‡π‘’|2div𝑋 + βŸ¨βˆ‡π‘’, βˆ‡π‘‹ βˆ‡π‘’βŸ© = 1 2 |βˆ‡π‘’|2div𝑋 + βŸ¨π‘‹, βˆ‡βˆ‡π‘’βˆ‡π‘’βŸ© By taking the difference and applying integration by parts to the left-hand side, we obtain the desired equality. β–‘ Proof. Throughout this proof 𝐢 denotes some constants that only depends on π‘ž0 and the metric 𝑔. And it might change from line to line. By Pohozaev identity (2.4.7) for harmonic function 𝑒 and arbitrary smooth vector field 𝑋 we have ∫ 𝑀 1 2 (⟨ Β―βˆ‡ Β―βˆ‡π‘’ 𝑋, Β―βˆ‡π‘’βŸ© βˆ’ | Β―βˆ‡π‘’|2div𝑔 𝑋) = πœ•π‘’ πœ•π‘› (cid:12)Ξ£ = (cid:174)𝑛 in this equality and note that Β―βˆ‡π‘‹ is bounded by compactness, | Β―βˆ‡π‘’|2βŸ¨π‘‹, (cid:174)π‘›βŸ©) βŸ¨π‘‹, βˆ‡π‘’βŸ© βˆ’ 1 2 Ξ£ ( ∫ Fix a 𝑋 satisfying 𝑋(cid:12) πœ•π‘’ πœ•π‘› πœ•π‘’ πœ•π‘› ∫ )2 + (| Β―βˆ‡π‘’|2div𝑔 𝑋 βˆ’ 2⟨ Β―βˆ‡ Β―βˆ‡π‘’ 𝑋, Β―βˆ‡π‘’βŸ©) 𝑀 ∫ )2 + 𝐢 | Β―βˆ‡π‘’|2 𝑀 (2.4.8) ∫ Ξ£ |βˆ‡π‘’|2 = ≀ ∫ Ξ£ ∫ ( ( Ξ£ ∫ ≀ 𝐢 πœ•π‘’ πœ•π‘› ( )2 Ξ£ The last inequality follows by lemma 2.2.4. Adding πœ†2βˆ’π‘Ÿ 2 πœ† ∫ 𝑀 | Β―βˆ‡π‘’2| β‰₯ 0 (by (2.4.5)) to the right hand side and using (2.4.4), it becomes the following ∫ |βˆ‡ 𝑓 |2 ≀ 𝐢 ( ∫ πœ•π‘’ πœ•π‘› )2 + πœ†2 βˆ’ π‘Ÿ 2 πœ† ∫ ∫ 𝑓 2π‘ž βˆ’ 2πœ†π‘Ÿ ∫ (π‘Ÿ 𝑓 π‘ž+1 βˆ’ πœ† 𝑀 | Β―βˆ‡π‘’2|) 𝑓 π‘ž+1 + πœ†2∫ ∫ 𝑓 2)] ( = 𝐢 [(π‘Ÿ 2∫ πœ†2 βˆ’ π‘Ÿ 2 πœ† ∫ + = 𝐢 [π‘Ÿ 2( 𝑓 2π‘ž + 𝑓 2) + π‘Ÿ ( πœ†2 βˆ’ π‘Ÿ 2 πœ† ∫ βˆ’ 2πœ†) 𝑓 π‘ž+1] 45 𝑓 2) (2.4.9) Consider the function πœ™(π‘₯) = π‘Žπ‘ž+1+π‘₯ + π‘Žπ‘ž+1βˆ’π‘₯ for π‘₯ ∈ [0, π‘ž βˆ’ 1] and π‘Ž > 0. Taylor expansion implies that for some πœƒ ∈ [0, π‘ž βˆ’ 1] πœ™(π‘ž βˆ’ 1) = 2π‘Žπ‘ž+1 + πœ™β€²β€²(πœƒ) 2 (π‘ž βˆ’ 1)2 =2π‘Žπ‘ž+1 + ≀2π‘Žπ‘ž+1 + (π‘ž βˆ’ 1)2 2 (π‘ž βˆ’ 1)2 2 (log π‘Ž)2π‘Žπ‘ž+1(π‘Žπœƒ + π‘Žβˆ’πœƒ) (log π‘Ž)2(π‘Ž2π‘ž + π‘Ž2) Using this estimate in 2.4.9, ∫ |βˆ‡ 𝑓 |2 ≀ 𝐢 [ (π‘žβˆ’1)2 2 ∫ π‘Ÿ 2( ( 𝑓 2π‘ž + 𝑓 2) log 𝑓 +π‘Ÿ ( βˆ’π‘Ÿ 2 πœ† ∫ +2π‘Ÿ βˆ’πœ†) 𝑓 π‘ž+1] ∫ 𝐢 2 (π‘ž βˆ’ 1)2π‘Ÿ 2( ≀ ≀ 𝐢 (π‘ž βˆ’ 1)2π‘Ÿ 2 ∫ 𝑓 2π‘ž (log 𝑓 )2 + ∫ 𝑓 2(log 𝑓 )2) (2.4.10) 𝑓 2π‘ž (log 𝑓 )2 Using Lemma 2.4.1 and (2.4.5), (2.4.10) becomes ∫ |βˆ‡ 𝑓 |2 ≀ 𝐢 (π‘ž βˆ’ 1)2πœ†2 ∫ |βˆ‡ 𝑓 |2 and 𝑓 , and therefore 𝑒, will be constant if (π‘ž βˆ’ 1)πœ† is small. β–‘ Remark 2.4.1. Note that (2.4.8) was obtained in (2.2.3) by directly utilizing the ellipticity of the Dirichlet-to-Neumann operator. However, it is challenging to discern how curvature conditions come into play in that method. (2.4.8) is more likely to be connected to geometry, and the problem is how to construct a nice vector field. This provides some information, but not precisely what we are seeking. This idea will come back later in a log-Sobolev inequality. One might hope to get a uniform bound for the constant in Dirichlet-to-Neumann operator, but such an estimate doesn’t exist even under the condition of positive sectional curvature and 𝐼 𝐼 β‰₯ 1 where 𝐼 𝐼 is the second fundamental form. Actually, consider the ellipse {(π‘₯, 𝑦)(cid:12) (cid:12)π‘₯2 + π‘˜ 2𝑦2 ≀ 1} Under scaling of the metric ¯𝑔 = π‘˜ βˆ’2𝑔, the second fundamental form can be arbitrarily large. At the same time both | Β―βˆ‡ 𝑓 |2 = π‘˜ 2|βˆ‡ 𝑓 |2 and ( πœ•π‘’ πœ•π‘› )2 are scaled by the same factor. So we might πœ• ¯𝑛 )2 = π‘˜ 2( πœ•π‘’ 46 forget about the restriction on the 𝐼 𝐼. The normal vector and tangent vector are (cid:174)𝑛 = (π‘₯, π‘˜ 2𝑦) and (cid:174)𝑣 = (π‘˜ 2𝑦, π‘₯) respectively. For 𝑒 = π‘₯ we calculate as follows ∫ πœ•π‘’ πœ•π‘› ( )2 = 4 ((1, 0) Β· βˆšοΈ„ (cid:174)𝑛 | (cid:174)𝑛| )2 π‘˜ 4𝑦2 + π‘₯2 π‘˜ 4𝑦2 dx ∫ 1 0 ∫ 1 = 4 π‘₯2 π‘˜βˆšοΈ(π‘˜ 2(1 βˆ’ π‘₯2) + π‘₯2) (1 βˆ’ π‘₯2) 4 π‘˜ πœ•π‘› )2 = ∫ 1 β‰₯ 4. So ∫ |βˆ‡π‘’|2 can’t be uniformly bounded by ∫ ( πœ•π‘’ 0 ∫ 1 1 βˆ’ π‘₯2 4 π‘˜ dx √ ≀ = π‘₯ 0 (2.4.11) Also note that ∫ |βˆ‡π‘’|2 + ( πœ•π‘’ curvature and 𝐼 𝐼 restriction. Consider 𝑒 = 𝑦, we see that ∫ ( πœ•π‘’ ∫ |βˆ‡π‘’|2, either. πœ•π‘› )2 under πœ•π‘› )2 can’t be uniformly bounded by Remark 2.4.2. The idea of this proof comes from the paper by Ou and Lin [LO23]. We translate |βˆ‡ 𝑓 |2 can be bounded by combination of ∫ their work as follows: ∫ 𝑀 | Β―βˆ‡π‘’|2. These two terms are β€œdifference" of 𝐿 𝑝 norms by (2.2.8), and this β€œdifference" can be bounded by ∫ |βˆ‡ 𝑓 |2. In [LO23], this β€œdifference" is measured by the ratio of 𝐿 𝑝 norm. We treated it differently by taking πœ•π‘› )2 and ∫ ( πœ•π‘’ Ξ£ Ξ£ Ξ£ the subtraction. 2.5 An ODE Approach Consider π‘ πœ†,π‘ž βˆ’ 𝐴(Ξ£) π‘žβˆ’1 π‘ž+1 πœ† ≀ 0. If this inequality is strict, πΈπ‘ž,πœ† must have a non-constant minimizer. Using this minimizer, we are going to show that the strict negativity is preserved along some curve of πœ†, π‘ž which looks like (2.5.1) in the theorem below. But we know from the work of [GW20] that for unit ball, π‘ž = 𝑛 π‘›βˆ’2, πœ† = π‘›βˆ’2 2 , πΈπ‘ž,πœ† only admits constant minimizer, which gives some restriction on π‘ž, πœ†. Lemma 2.5.1. Let (𝑀, Ξ£, 𝑔) be a Riemannian manifold with boundary so that 𝐴(Ξ£) = 1. Suppose 𝑠(π‘ž0, πœ†0) βˆ’ πœ†0 ≀ βˆ’πœ– < 0 for some (πœ†0, π‘ž0), then this inequality remains valid along the curve π‘ž + 1 π‘ž βˆ’ 1 = 𝐢 (πœ† βˆ’ πœ–) 47 (2.5.1) for πœ† ≀ πœ†0, where 𝐢 = π‘ž0+1 (π‘ž0βˆ’1)(πœ†0βˆ’πœ–) is chosen so that (πœ†0, π‘ž0) is on the curve. Proof. Along the curve (2.5.1), 𝑠(πœ†) and π‘ž(πœ†) are functions of πœ† only. It suffices to show that for πœ†1 satisfying 𝑠(πœ†1) = πœ†1 βˆ’ πœ–, 𝑠(πœ†) βˆ’ πœ† will be decreasing along the curve in the βˆ’πœ† direction near πœ†1. Note that 𝑠(πœ†1) = πœ†1 βˆ’ πœ– implies the existence of a non-constant minimizer for πΈπ‘ž1,πœ†1 satisfying (2.2.8), where π‘ž1 = π‘ž(πœ†1). Denote it by 𝑒. Fix this 𝑒, and we want to show that πΈπ‘ž(πœ†),πœ† (𝑒) decreases fast enough in βˆ’πœ† direction along the curve (2.5.1), and the theorem follows since 𝑠(π‘ž(πœ†, πœ†) ≀ πΈπœ†,π‘ž(πœ†) (𝑒) for all πœ†. Namely, we need to prove the following inequality: πœ• πœ• (βˆ’πœ†) (πΈπœ†,π‘ž(πœ†) (𝑒) βˆ’ πœ†) < 0 One calculates that πœ• πœ•π‘ž ( ∫ Ξ£ We assumed that 𝑓 π‘ž+1) 2 π‘ž+1 = ( ∫ 𝑓 π‘ž+1) 2 π‘ž+1 [βˆ’ Ξ£ 𝑓 π‘ž1+1 = 𝐴(Ξ£) = 1, so ⨏ Ξ£ 2 (π‘ž + 1)2 ∫ log Ξ£ 𝑓 π‘ž+1 + ∫ Ξ£ 2 π‘ž + 1 𝑓 π‘ž+1 log 𝑓 ∫ Ξ£ 𝑓 π‘ž+1 ] πœ• πœ•πœ† πΈπœ†,π‘ž(πœ†) (𝑒)(cid:12) (cid:12)πœ†1 = ∫ Ξ£ 𝑓 2 βˆ’ 2𝑠(πœ†1) π‘ž1 + 1 π‘žβ€²(πœ†1) ∫ Ξ£ 𝑓 π‘ž1+1 log 𝑓 So it suffices to show By HΓΆlder inequality and 𝐴(Ξ£) = 𝑓 π‘ž+1 log 𝑓 β‰₯ 0. Also note that ∫ ∫ Ξ£ Ξ£ ∫ Ξ£ 𝑓 2 βˆ’ ⨏ ∫ π‘žβ€²(πœ†1) 𝑓 π‘ž+1 log 𝑓 > 1 2𝑠(πœ†1) π‘ž + 1 𝑓 π‘ž+1 = 1, we have ∫ Ξ£ Ξ£ 𝑓 π‘₯ is a strict convex function in π‘₯, we have Ξ£ 𝑓 π‘ž+1+πœ– β‰₯ 1 for πœ– > 0 and it follows that (2.5.2) ∫ Ξ£ 𝑓 2 + (π‘ž βˆ’ 1) ∫ Ξ£ 𝑓 π‘ž+1 log 𝑓 > ∫ Ξ£ 𝑓 π‘ž+1 = 1 This inequality is strict since 𝑒 is not constant. So (2.5.2) holds provided βˆ’2π‘žβ€²(πœ†) (πœ† βˆ’ πœ–) π‘ž + 1 β‰₯ π‘ž βˆ’ 1 48 The solution for the equality is exactly (2.5.1), and we finish the proof. β–‘ Corollary 2.1. For (B𝑛, Sπ‘›βˆ’1, 𝑔) the standard metric, π‘ž ≀ 𝑛 π‘›βˆ’2, 𝑠(π‘ž, πœ†) is achieved only by constant functions for (π‘ž βˆ’ 1) (πœ† βˆ’ 𝑛 βˆ’ 2 2(𝑛 βˆ’ 1) ) < 𝑛 βˆ’ 2 𝑛 βˆ’ 1 (2.5.3) Proof. We will first scale the metric so that the area of the boundary is 1, namely consider (B𝑛, Sπ‘›βˆ’1, 𝑔 = π‘˜ 2𝑔) where π‘˜ = 𝐴(Sπ‘›βˆ’1) 1 1βˆ’π‘› .The function πΈπ‘ž,πœ† (𝑒) changes as follows 𝐸 π‘ž,πœ† (𝑒) = ∫ B𝑛 |βˆ‡π‘”π‘’|2dVol𝑔 + πœ† ∫ Sπ‘›βˆ’1 π‘’π‘ž+1dS𝑔) (∫ 2 π‘ž+1 Sπ‘›βˆ’1 𝑒2dS𝑔 = π‘˜ π‘›βˆ’2βˆ’ 2(π‘›βˆ’1) π‘ž+1 B𝑛 |βˆ‡π‘”π‘’|2dVol𝑔 + π‘˜πœ† ∫ ∫ Sπ‘›βˆ’1 π‘’π‘ž+1dS𝑔) (∫ 2 π‘ž+1 Sπ‘›βˆ’1 𝑒2dS𝑔 = π‘˜ π‘›βˆ’2βˆ’ 2(π‘›βˆ’1) π‘ž+1 𝑄 π‘˜πœ†,π‘ž (𝑒) (2.5.4) By the work of ([GW20]) for unit ball, (2.2.8) admit only constant solutions for πœ† = π‘›βˆ’2 After the scaling (2.5.4), for ¯𝑔 = π‘˜ 2𝑔 this critical point becomes (π‘ž, πœ†) = ( 𝑛 2 , π‘ž = 𝑛 π‘›βˆ’2. 2π‘˜ ). Suppose π‘›βˆ’2, π‘›βˆ’2 Β―π‘ π‘ž0,πœ†0 < πœ†0 for some (πœ†0, π‘ž0) satisfying πœ†1 < (𝑛 βˆ’ 2) (π‘ž + 1) 2π‘˜ (𝑛 βˆ’ 1) (π‘ž βˆ’ 1) (2.5.5) When πœ– is small enough, Β―π‘ π‘ž0,πœ†0 < πœ†0 βˆ’ πœ–. And by Theorem 2.5.1 this inequality remains valid along (2.5.1) for πœ† < πœ†0. In particular, we let π‘ž = 𝑛 π‘›βˆ’2, then πœ† = π‘ž + 1 𝐢 (π‘ž βˆ’ 1) + πœ– = (𝑛 βˆ’ 1) (π‘ž0 βˆ’ 1) (πœ†0 βˆ’ πœ–) π‘ž0 + 1 + πœ– (2.5.6) If (2.5.5) holds, we can make πœ– small so that πœ† < π‘›βˆ’2 2π‘˜ , which is contradiction since we must have ¯𝑠 π‘›βˆ’2 2π‘˜ , 𝑛 π‘›βˆ’2 = π‘›βˆ’2 2π‘˜ . Now transfer this result back to the unit ball and the proof is finished. β–‘ 49 Remark 2.5.1. For standard balls Wang’s conjecture been completely solved in [GL23], and the theorem above is also included. In the preceding sections, we demonstrated that for π‘ž < 𝑛 π‘›βˆ’2, there exists a corresponding πœ† such that 𝑠(π‘ž, πœ†) is only achieved by constants. Let’s fix one such pair as (π‘ž0, πœ†0) and employ a similar argument to the one in the corollary above. This will yield a result similar to Theorem 2.7, but with π‘ž0 < 𝑛 π‘›βˆ’2 instead of π‘ž0 < 𝑛+1 π‘›βˆ’1. These two approaches are distinct and offer different perspectives on Wang’s conjecture. Remark 2.5.2. The proof of the corollary does not rely on the specific structure of the manifold. The only instance where (B𝑛, Sπ‘›βˆ’1, 𝑔) is involved is at the critical point ( 𝑛 π‘›βˆ’2, π‘›βˆ’2 2 ). Therefore, our method is applicable to any manifold as long as one can compute such a critical point. The challenge lies in determining how to obtain such a point under curvature restrictions. When π‘ž = 𝑛 π‘›βˆ’2, the problem is related to type II Yamabe problem. A breakthrough in Yamabe problem might help us find a critical (π‘ž, πœ†). 2.6 Critial Power Case If π‘ž = 𝑛 π‘›βˆ’2, the trace operator is only continuous and fails to be compact, making the existence of minimizers more challenging. In this section, I will derive some existence results for the minimizer of 𝑠 (cid:0) 𝑛 π‘›βˆ’2, πœ†(cid:1). It’s difficult to determine whether these minimizers become constants for small πœ†β€™s. Lemma 2.6.1. For any compact Riemannian manifold with boundary and πœ† β‰₯ 0, 4(𝑛 βˆ’ 1) 𝑛 βˆ’ 2 𝑠( 𝑛 𝑛 βˆ’ 2 , πœ†) ≀ π‘Œ (B𝑛, Sπ‘›βˆ’1, 𝑑π‘₯2) (2.6.1) Proof. Fix a point 𝑝 ∈ Ξ£. We can find a small neighborhood π‘ˆ of 𝑝 ∈ 𝑀 so that π‘ˆ = π΅π‘›βˆ’1(𝛿)Γ—(0, 𝛿) for some small 𝛿. Fix a cut-off function πœ™ so that πœ™ = 1 in π΅π‘›βˆ’1(𝛿/2) Γ— (0, 𝛿/2) and vanishes outside π‘ˆ. Let {π‘₯𝑖}π‘›βˆ’1 𝑖=1 be the coordinates for π΅π‘›βˆ’1(𝛿) and 𝑑 coordinate for (0, 𝛿). Define π‘£πœ– = ( πœ– (πœ– + 𝑑)2 + |𝑦|2 ) π‘›βˆ’2 2 (2.6.2) 50 Recall that we can use πœ™π‘£πœ– as test functions to establish the type II Yamabe inequality (1.2.17). 𝑅𝑒2𝑑𝑉 and ∫ Ξ£ ∫ 𝑀 we can demonstrate that ∫ 𝑀 (cid:16) π‘›βˆ’2 2(π‘›βˆ’1) 𝐻 βˆ’ πœ† The differences between the functional of the type II Yamabe problem and 𝐸 𝑛 π‘›βˆ’2 ,π‘ž lie in the terms 𝑒2𝑑𝑆. Note that 𝐻 and 𝑅 remain bounded for a fixed metric. If (cid:17) 𝑒2𝑑𝑉 and ∫ Ξ£ complete. Let 𝑑𝑉𝐸 and 𝑑𝑆𝐸 be the volume form with respect to Euclidean space. 𝑒2𝑑𝑉 vanish as πœ– β†’ 0 for 𝑒 = πœ™π‘£πœ– , then the proof is ∫ 𝑀 (πœ™π‘£πœ– )2𝑑𝑉𝑔 ≀ ∫ 𝑣2 πœ– 𝑑𝑉𝑔 π‘ˆ ∫ ≀ 𝐢 𝑣2 πœ– 𝑑𝑉𝐸 π‘ˆ ∫ 𝛿 0 ∫ 𝛿 0 ∫ 𝛿 0 ∫ 𝛿 πœ– 0 = 𝐢 = 𝐢 ≀ 𝐢 = 𝐢 ∫ π΅π‘›βˆ’1 (𝛿) ∫ 𝑑+πœ– ) π΅π‘›βˆ’1 ( 𝛿 πœ– π‘›βˆ’2 (𝑑 + πœ–)π‘›βˆ’3 πœ– 2 (1 + 𝑑)π‘›βˆ’3 𝑑𝑑 𝑑𝑑 ≀ πΆπœ– 2 ( πœ– (πœ– + 𝑑)2 + |𝑦|2 )π‘›βˆ’2𝑑𝑦𝑑𝑑 πœ– π‘›βˆ’2 (𝑑 + πœ–)π‘›βˆ’3 1 (1 + |𝑧|2)π‘›βˆ’2 𝑑𝑧𝑑𝑑 We used change of variable in the third and fifth line. Similarly, we have ∫ Ξ£ (πœ™π‘£πœ– )2𝑑𝑆𝑔 ≀ ∫ 𝑣2 πœ– Ξ£ ∫ ≀ 𝐢 π΅π‘›βˆ’1 (𝛿) 𝑣2 πœ– 𝑑𝑆𝐸 ( πœ– πœ– 2 + |𝑦|2 )π‘›βˆ’2𝑑𝑧 ∫ π΅π‘›βˆ’1 ∫ = 𝐢 = 𝐢 π΅π‘›βˆ’1 ( 𝛿 πœ– ) πœ– (1 + |𝑧|2) (π‘›βˆ’2) 𝑑𝑧 ≀ πΆπœ– (πœ™π‘£πœ– ) 2(π‘›βˆ’1) π‘›βˆ’2 β‰₯ ∫ Ξ£ ∫ 2(π‘›βˆ’1) π‘›βˆ’2 πœ– 𝑣 𝑑𝑆𝑔 π΅π‘›βˆ’1 (𝛿/2) ∫ π΅π‘›βˆ’1 (𝛿/2) ∫ β‰₯ 𝐢 = 𝐢 πœ– ( πœ– 2 + |𝑦|2 ) (π‘›βˆ’1) 𝑑𝑆𝐸 1 + |𝑧|2 ) (π‘›βˆ’1) 𝑑𝑧 β‰₯ 𝐢 1 ( π΅π‘›βˆ’1 (𝛿/(2πœ–)) Let πœ– β†’ 0, and the three estimates prove the lemma. If π‘ž < 𝑛 π‘›βˆ’2, such a bound doesn’t exist. 51 (2.6.3) β–‘ Lemma 2.6.2. 𝑠(π‘ž, πœ†) β†’ ∞ as πœ† β†’ ∞ for π‘ž < 𝑛 π‘›βˆ’2. Proof. We prove by contradiction. Suppose not. Then there exists πœ†π‘– β†’ and 𝑒𝑖 such that πΈπ‘ž,πœ†π‘– (𝑒𝑖) < 𝐢. Without loss of generality, we might assume βˆ₯𝑒𝑖 βˆ₯ πΏπ‘ž+1 (Ξ£) = 1. Then we must have ∫ 𝑀 |βˆ‡π‘’π‘– |2 < 𝐢 and ∫ 𝑒2 𝑖 β†’ 0. By Alauglu theorem we can find a subsequence, still denoted by 𝑒𝑖, such that Ξ£ π‘›βˆ’2, by compactness, we can pick a further β–‘ subsequence so that βˆ₯𝑒0βˆ₯ πΏπ‘ž+1 (Ξ£) = lim βˆ₯𝑒𝑖 βˆ₯ πΏπ‘ž+1 (Ξ£) = 1, which is a contradiction. 𝑖 = 0. Since π‘ž < 𝑛 𝑒2 𝑒𝑖 ⇀ 𝑒0 and ∫ Ξ£ 0 = lim ∫ 𝑒2 Ξ£ According to the computations in Lemma 2.6.1, we have π‘£πœ– β†’ 0 in 𝐿2(Ξ£), but they do not 2(π‘›βˆ’1) converge in 𝐿 where 𝑠( 𝑛 π‘›βˆ’2 (Ξ£). This elucidates why the argument fails for π‘ž = 𝑛 π‘›βˆ’2. In this critical case, π‘›βˆ’2, πœ†) is bounded in πœ†, the dynamics are quite different. The key observation is that if π‘›βˆ’2, πœ†) stops increasing for large πœ†, it is likely minimized through a sequence of functions that blow up somewhere, with their 𝐿2(Ξ£) norms tending to zero. This phenomenon is akin to what 𝑠( 𝑛 is observed in (2.6.2). Consequently, 𝑠( 𝑛 π‘›βˆ’2, πœ†) admits no minimizer in this scenario, not even π‘›βˆ’2, πœ†) keeps increasing in πœ†, functions like (2.6.2) are ruled out as minimizing sequences. This exclusion opens up the possibility of obtaining a minimizer. These constants. On the contrary, if 𝑠( 𝑛 observations can be made concrete by the following theorem. Theorem 2.8. i):If there exists πœ‡ < πœ† such that 𝑠( 𝑛 π‘›βˆ’2, πœ†) = 𝑠( 𝑛 π‘›βˆ’2, πœ‡), then 𝑠( 𝑛 π‘›βˆ’2, πœ†) doestn’t admit any minimizer. ii): If there exists πœ† < πœ‡ such that 𝑠( 𝑛 π‘›βˆ’2, πœ†) < 𝑠( 𝑛 π‘›βˆ’2, πœ‡), then 𝑠( 𝑛 π‘›βˆ’2, πœ†) admits a minimizer. Proof. Part i): Suppose 𝑠( 𝑛 π‘›βˆ’2, πœ†) admit a minimizer 𝑒. Then 𝑠( 𝑛 𝑛 βˆ’ 2 , πœ‡) ≀ 𝐸 𝑛 π‘›βˆ’2 ,πœ‡ (𝑒) < 𝐸 𝑛 π‘›βˆ’2 ,πœ† (𝑒) = 𝑠( 𝑛 𝑛 βˆ’ 2 , πœ†) which contradicts our assumption. Part ii): Let 𝑒𝑖 be a minimizing sequence for 𝑠( 𝑛 π‘›βˆ’2, πœ†). We can scale 𝑒𝑖 so that βˆ₯𝑒𝑖 βˆ₯ 𝐿 𝑝 = 1, where 𝑝 = 2(π‘›βˆ’1) π‘›βˆ’2 . 52 Then lim π‘–β†’βˆž (cid:16) ∫ 𝑀 |βˆ‡π‘’π‘– |2 + πœ† ∫ Ξ£ (cid:17) 𝑒2 𝑖 = 𝑠( 𝑛 𝑛 βˆ’ 2 , πœ†) Use these 𝑒𝑖 as test-function for πœ‡, and we have ∫ 𝑀 |βˆ‡π‘’π‘– |2 + πœ‡ ∫ Ξ£ 𝑒2 𝑖 ≀ 𝑠( 𝑛 𝑛 βˆ’ 2 , πœ‡) By Alaoglu theorem and compactness we can get a subsequence so that 𝑒𝑖 ⇀ 𝑒 in 𝐻1(𝑀), 𝑒𝑖 ⇀ 𝑒 in 𝐿 𝑝 (Ξ£), 𝑒𝑖 β†’ 𝑒 in 𝐿2, 𝑒𝑖 β†’ 𝑒 a.e. in 𝑀 Use these 𝑒𝑖 as test-function for πœ‡, ∫ 𝑀 |βˆ‡π‘’π‘– |2 + πœ‡ ∫ Ξ£ 𝑒2 𝑖 β‰₯ 𝑠( 𝑛 𝑛 βˆ’ 2 , πœ‡) Take the difference between the (2.6.4) and (2.6.5), and we get (2.6.4) (2.6.5) (2.6.6) (πœ‡ βˆ’ πœ†) ∫ Ξ£ 𝑒2 = lim(πœ‡ βˆ’ πœ†) ∫ Ξ£ 𝑒2 𝑖 β‰₯ 𝑠( 𝑛 𝑛 βˆ’ 2 , πœ‡) βˆ’ 𝑠( 𝑛 𝑛 βˆ’ 2 , πœ†) > 0 This is where we used our assumption. This rules out possibility that 𝑒 ≑ 0, which happens in the proof of lemma 2.6.1. Next we are going to show 𝑒 minimizes 𝑠( 𝑛 π‘›βˆ’2, πœ†). Since 𝑒𝑖 ⇀ 𝑒 in 𝐿 𝑝 (Ξ£), we have βˆ₯𝑒βˆ₯ 𝑝 ≀ 1. Let 𝑣𝑖 = 𝑒𝑖 βˆ’ 𝑒. By a result in [BL83], 1 = lim βˆ₯𝑒𝑖 βˆ₯ 𝑝 𝑝 = lim βˆ₯𝑒 + 𝑣𝑖 βˆ₯ 𝑝 𝑝 = βˆ₯𝑒βˆ₯ 𝑝 𝑝 + lim βˆ₯𝑣𝑖 βˆ₯ 𝑝 𝑝 Note that 𝑒 β‰  0, so βˆ₯𝑒βˆ₯ 𝑝 ≀ 1, lim βˆ₯𝑣𝑖 βˆ₯ 𝑝 ≀ 1. Consequently 1 ≀ lim βˆ₯𝑒βˆ₯2 𝑝 + lim βˆ₯𝑣𝑖 βˆ₯2 𝑝 ≀ βˆ₯𝑒βˆ₯2 𝑝 + lim 1 𝑠( 𝑛 π‘›βˆ’2, πœ†) 1 lim 𝑠( 𝑛 π‘›βˆ’2, πœ†) ∫ (cid:0) 𝑀 |βˆ‡π‘£π‘– |2 + πœ† ∫ Ξ£ (cid:1) 𝑣2 𝑖 ∫ 𝑀 |βˆ‡π‘£π‘– |2 = βˆ₯𝑒βˆ₯2 𝑝 + 𝑣𝑖 can be estimated using (2.6.4), ∫ 𝑀 |βˆ‡π‘’|2 + πœ† ∫ Ξ£ 𝑒2 + lim ∫ 𝑀 |βˆ‡π‘£π‘– |2 = 𝑠( 𝑛 𝑛 βˆ’ 2 , πœ†) 53 where we used 𝑣𝑖 ⇀ 0 ∈ 𝐻1(𝑀) and 𝑣𝑖 β†’ 0 in 𝐿2(Ξ£). Combine the two equations above and we get 𝑀 |βˆ‡π‘’|2 + ∫ ∫ βˆ₯𝑒βˆ₯2 𝑝 Ξ£ 𝑒2 ≀ 𝑠( 𝑛 𝑛 βˆ’ 2 , πœ†) So 𝑒 is a minimizer. β–‘ Remark 2.6.1. The proof of part ii) comes from [BN83] where the H.Brezis and L.Nirenberg proved similar results for a different equation βˆ’Ξ”π‘’ = 𝑒 𝑝 + πœ†π‘’ on 𝑀 𝑒 > 0 on 𝑀 𝑒 = 0 on Ξ£ Corollary 2.2. For the unit disk, 𝑠( 𝑛 π‘›βˆ’2, π‘›βˆ’2 2 ) admits only constant minimizer for πœ† < π‘›βˆ’2 2 , and admit no minimizer for πœ† > 𝑛 π‘›βˆ’2. Proof. It’s well known that 𝑠( 𝑛 π‘›βˆ’2, π‘›βˆ’2 2 ) = π‘›βˆ’2 2 |Ξ£| 1 π‘›βˆ’1 . Then the result follows from the two theorems above and lemma 2.1.1. β–‘ For the critical power case, 𝑠( 𝑛 π‘›βˆ’2 , πœ†) has a strong relationship with type II Yamabe problem. Use standard argument and we can get a similar existence theorem Theorem 2.9. If 4(π‘›βˆ’1) π‘›βˆ’2 𝑠( 𝑛 π‘›βˆ’2, πœ†) < π‘Œ (B𝑛, Sπ‘›βˆ’1, 𝑑π‘₯2), then it admits a minimizer. Proof. The trick is again β€œlowering the index". For each π‘ž < 𝑛 π‘›βˆ’2, 2.2.8 admits a solution π‘’π‘ž (it might be constant). If π‘’π‘žβ€™s are uniforma bounded above, then the ellipticity of Dirichlet-to- Neumann operator implies a universal upper bound for π‘’π‘ž in 𝐢 π‘˜ for any π‘˜ ∈ Z+. Consequently π‘ π‘ž converges to a solution of (2.2.8) for bound. Suppose on the contrary that there exits π‘žπ‘˜ β†’ 𝑛 𝑛 π‘›βˆ’2. So it suffices to show that there doen’t exist such a 𝐿∞ π‘›βˆ’2, π‘’π‘˜ and π‘π‘˜ ∈ Ξ£ so that π‘’π‘˜ minimizes πΈπ‘žπ‘˜,πœ† and π‘šπ‘˜ := π‘’π‘˜ ( π‘π‘˜ ) = supπ‘₯βˆˆπ‘€ 𝑒(π‘₯) β†’ ∞. The idea is that we are going to show that by scaling π‘’π‘˜ will β€œconverge" locally around 𝑝 to a solution of 𝑃𝐷𝐸 in upper plane, and this contradicts the 54 assumption 𝑠( 𝑛 π‘›βˆ’2, πœ†) < π‘Œ (B𝑛, Sπ‘›βˆ’1, 𝑑π‘₯2). For convenience of readers, I will restate the equations for π‘’π‘˜ Ξ”π‘’π‘˜ = 0 πœ•π‘’π‘˜ πœ•πœˆ π‘’π‘žπ‘˜+1 π‘˜ = 1 = βˆ’πœ†π‘’π‘˜ + π‘ π‘’π‘žπ‘˜ π‘˜ ∫ Ξ£ on 𝑀 𝑛 on Ξ£π‘›βˆ’1 (2.6.7) By compactness we might assume π‘π‘˜ β†’ 𝑝 ∈ Ξ£. We might pick local coordinate upper ball π‘ˆπ‘ (2πœ–) := B𝑛 (2πœ–) ∩ {π‘₯𝑛 β‰₯ 0} centered at 𝑝, where {π‘₯𝑖}1β‰€π‘–β‰€π‘›βˆ’1 is local normal coordinate for 𝑝 ∈ Ξ£ and π‘₯𝑛 is the coordinate in normal direction. Let π›Ώπ‘˜ = π‘š1βˆ’π‘žπ‘˜ is defined in π‘ˆπ‘ ( πœ– π›Ώπ‘˜ ) for large 𝑖’s with radius πœ– π›Ώπ‘˜ β†’ ∞. By (2.6.7), 𝑣 π‘˜ locally satisfies π‘’π‘˜ (π›Ώπ‘˜ π‘₯ + π‘π‘˜ ). Then 𝑣 π‘˜ , and 𝑣 π‘˜ = 1 π‘šπ‘˜ π‘˜ 1 π‘π‘˜ πœ•π‘— (π‘Žπ‘– 𝑗 π‘˜ πœ•π‘–π‘£ π‘˜ ) = 0 on π‘ˆπ‘ ( πœ– π›Ώπ‘˜ ) βˆ’ πœ•π‘£ π‘˜ πœ•π‘₯𝑛 + π‘π‘˜ 𝑣 π‘˜ = π‘ π‘£π‘žπ‘˜ π‘˜ on π‘ˆπ‘ ( πœ– π›Ώπ‘˜ ) ∩ {π‘₯𝑛 = 0} (2.6.8) where π‘Žπ‘– 𝑗 π‘˜ (π‘₯) = 𝑔𝑖 𝑗 (π›Ώπ‘˜ π‘₯ + π‘π‘˜ ) β†’ 𝛿𝑖 𝑗 π‘π‘˜ (π‘₯) = √︁det 𝑔(π›Ώπ‘˜ π‘₯ + π‘π‘˜ ) β†’ 1 π‘π‘˜ = πœ†π‘šβˆ’π‘žπ‘˜ π‘˜ β†’ 0 In the equation above we have an additional βˆ’ in front of πœ•π‘£ π‘˜ πœ•π‘₯𝑛 because πœ• πœ•π‘₯𝑛 is in the inner normal direction instead of outer normal direction. Note that π‘’π‘˜ π‘šπ‘˜ has uniform 𝐿∞ bound by its definition. Since they satisfy a similar equation Ξ”π‘’π‘˜ = 0 on 𝑀 πœ•π‘’π‘˜ πœ•πœˆ + π‘π‘˜π‘’π‘˜ = π‘ π‘’π‘žπ‘˜ π‘˜ onΞ£ So π‘’π‘˜ are uniform bounded in any πΆπ‘˜ (𝑀) norm by ellipticity of Dirichlet-to-Neumann operator. 𝑣 π‘˜ are defined in π‘ˆπ‘ ( πœ– π›Ώπ‘˜ π‘˜ πœ•π›Όπ‘’π‘˜ (π›Ώπ‘˜ π‘₯ + π‘π‘˜ ) for π›Ώπ‘˜ β†’ 0. So 𝑣 π‘˜ are also uniformly bounded in any 𝐢 π‘˜ (π‘ˆπ‘ (𝑅)) for any fixed 𝑅. So we could pick a sub-sequence so that 𝑣𝑖 β†’ 𝑣 so ) and πœ•π›Όπ‘£ π‘˜ (π‘₯) = 𝛿|𝛼| that Δ𝑣 = 0 on H𝑛 + πœ•π‘£ πœ•πœˆ = 𝑠𝑣 𝑛 π‘›βˆ’2 on Rπ‘›βˆ’1 55 (2.6.9) where 𝜈 is the outer normal direction. (I failed to derive this convergence from (2.6.8) directly since Dirichlet-to-Neumann operator is global, while 𝑣 π‘˜ is only defined locally. Also, the Schauder estimates can’t be applied directly.) Now if 𝜈 has enough decay at infinity, we could multiply (2.6.9) by 𝑣 on both sides and integrate to get ∫ H𝑛 + |βˆ‡π‘£|2 = 𝑠 ∫ Rπ‘›βˆ’1 2(π‘›βˆ’1) π‘›βˆ’2 𝑣 (2.6.10) Also, as a limit we probably have ∫ Ξ£ + has vanishing mean curva- ture and scalar curvature. Use 𝑣 as the test function for type II Yamabe problem and we get 2(π‘›βˆ’1) π‘›βˆ’2 ≀ 1. Note that H𝑛 𝑣 π‘Œ (H𝑛 +, Rπ‘›βˆ’1, 𝑑π‘₯2) ≀ 4(π‘›βˆ’1) π‘›βˆ’2 𝑠. However, it’s well known that the upper half plane and unit disk are conformally equivalent through 𝐹 (π‘₯1, Β· Β· Β· , π‘₯𝑛) = 1 Β· Β· Β· βˆ’ π‘₯2 𝑛) (2.6.11) 𝐹 : B𝑛 β†’ H𝑛 1 𝑃 πΉβˆ—π‘”H𝑛 = (2π‘₯1, Β· Β· Β· , 2π‘₯π‘›βˆ’1, 1 βˆ’ π‘₯2 4 𝑃2 𝑔B𝑛 := πœ™ 4 π‘›βˆ’2 𝑔B𝑛 where 𝑃 = π‘₯2 1 + Β· Β· Β· + π‘₯2 π‘›βˆ’1 + (1 βˆ’ π‘₯𝑛)2. Use π‘£πœ™ as test function for B𝑛, and we will get π‘Œ (B𝑛, Sπ‘›βˆ’1, 𝑑π‘₯2) = π‘Œ (H𝑛 +, Rπ‘›βˆ’1, 𝑑π‘₯2) ≀ 4(𝑛 βˆ’ 1) 𝑛 βˆ’ 2 𝑠( 𝑛 𝑛 βˆ’ 2 , πœ†) which contradicts to assumption 𝑠( 𝑛 π‘›βˆ’2 , πœ†) < π‘Œ (B𝑛, Sπ‘›βˆ’1, 𝑑π‘₯2). Next we will fix the gaps in the argument above. Let 𝑔′ = 𝑔|Ξ£. We can compute in local coordinate ∫ π΅π‘›βˆ’1 𝑝 ( πœ– π›Ώπ‘˜ ) π‘£π‘žπ‘˜+1 π‘˜ √︁det(𝑔′)( π‘π‘˜ + π›Ώπ‘˜ π‘₯)𝑑π‘₯ = ∫ √︁det(𝑔′) (π‘₯) π‘˜ π‘šπ‘žπ‘˜+1 π›Ώπ‘›βˆ’1 π‘˜ ∫ π΅π‘›βˆ’1 π‘π‘˜ (πœ–) ≀ π‘š (π‘žβˆ’1)(π‘›βˆ’1)βˆ’π‘žβˆ’1 π‘˜ π‘’π‘žπ‘˜+1 π‘˜ 𝑑𝑦 π‘’π‘žπ‘˜+1 π‘˜ 𝑑𝑉𝑔′ Ξ£ We used change of variable 𝑦 = π‘π‘˜ + π›Ώπ‘˜ π‘₯ in the first line. Since π‘ž < 𝑛 π‘›βˆ’2, π‘š (π‘žβˆ’1)(π‘›βˆ’1)βˆ’π‘žβˆ’1 π‘˜ ≀ 1 for = π‘š ((π‘›βˆ’2)π‘žβˆ’π‘› π‘˜ 56 large π‘˜β€™s. Since Similarly, we compute ∫ π‘ˆ 𝑝 ( πœ– π›Ώπ‘˜ ) π‘Žπ‘– 𝑗 π‘˜ (π‘₯)πœ•π‘–π‘£ π‘˜ πœ•π‘— 𝑣 π‘˜ π‘π‘˜ (π‘₯)𝑑π‘₯ = ∫ π‘ˆ 𝑝 ( πœ– ) 𝛿 ) π‘˜ 𝛿2 π‘˜ π‘š2 π‘˜ ∫ (cid:0)𝑔𝑖 𝑗 πœ•π‘–π‘’π‘˜ πœ•π‘— π‘’π‘˜ (cid:1) (π›Ώπ‘˜ π‘₯ + π‘π‘˜ )𝑑π‘₯ = π‘š (π‘›βˆ’2)π‘žβˆ’π‘› π‘˜ |βˆ‡π‘’π‘˜ |2𝑑𝑉𝑔 π‘ˆ π‘π‘˜ (πœ–) Since √︁det(𝑔′)( π‘π‘˜ + π›Ώπ‘˜ π‘₯) β†’ 1 and π‘Žπ‘– 𝑗 π‘˜ β†’ 𝛿𝑖 𝑗 , and 𝑣 π‘˜ converges uniformly in any compact subset, ≀ π‘š (π‘›βˆ’2)π‘žβˆ’π‘› π‘˜ βˆ₯βˆ‡π‘’π‘˜ βˆ₯ 𝐿2 (𝑀) ≀ 𝐢 by Fatou’s lemma we arrive at ∫ Rπ‘›βˆ’1 ∫ 𝑣 2(π‘›βˆ’1) π‘›βˆ’2 ≀ 1 (2.6.12) |βˆ‡π‘£|2 ≀ ∞ H𝑛 + +, and 𝑣 𝑅 (π‘₯) = πœ‚( π‘₯ Let πœ‚(π‘₯) be a cut-off function in H𝑛 𝑅 )𝑣. Then with the two bounds above we can verify that 𝑣 𝑅 β†’ 𝑣 in 𝐻1(H𝑛 +) and 𝐿 2(π‘›βˆ’1) π‘›βˆ’2 (Rπ‘›βˆ’1) by showing ∫ H𝑛 + ∫ Rπ‘›βˆ’1 |βˆ‡(𝑣 βˆ’ 𝑣 𝑅)|2 β†’ 0 |𝑣 βˆ’ 𝑣 𝑅 | 2(π‘›βˆ’1) π‘›βˆ’2 β†’ 0 Multiply πœ‚( π‘₯ 𝑅 )𝑣 to (2.6.9) and integrate by part, we get ∫ H𝑛 + βŸ¨βˆ‡π‘£, βˆ‡π‘£ 𝑅)⟩ = 𝑠 ∫ Rπ‘›βˆ’1 𝑣 𝑛 π‘›βˆ’2 𝑣 𝑅 Let 𝑅 β†’ ∞, and we get (2.6.10). β–‘ In view of the previous results, for a fixed Riemannian manifold with boundary we can ask the following questions: i) does 4(𝑛 βˆ’ 1) 𝑛 βˆ’ 2 𝑠( 𝑛 𝑛 βˆ’ 2 , πœ†) β†’ π‘Œ (B𝑛, Sπ‘›βˆ’1, 𝑑π‘₯2) as πœ† β†’ ∞ ii) does there exists a πœ†0 so that 𝑠( , πœ†0) = 𝑠( 𝑛 𝑛 βˆ’ 2 , πœ†) for all πœ† > πœ†0 𝑛 𝑛 βˆ’ 2 𝑛 𝑛 βˆ’ 2 iii) does there exists a πœ†1 so that 𝑠( , πœ†) admits only constant minimizer for all πœ† < πœ†1 My guess is all these are correct, but I have not found a way to solve these. 57 CHAPTER 3 A LOG-SOBOLEV INEQUALITY In this chapter, I will introduce the log-Sobolev inequality, which is closely related to Wang’s conjecture. The validity of the log-Sobolev inequality provides a key insight into the confidence we can place in Wang’s conjecture. 3.1 Motivation for Log-Sobolev inequality Let (𝑀, πœ• 𝑀 = Ξ£, 𝑔) a Riemannian manifold with boundary. Wang’s conjecture 2.1 relies on the boundedness of the trace operator: 𝐻1(𝑀) ↩→ πΏπ‘ž+1(Ξ£) for π‘ž ≀ 𝑛 π‘›βˆ’2. The embedding is compact 𝑛 π‘›βˆ’2, ultimately losing compactness. Consequently, the conjecture becomes more challenging as π‘ž increases. For the critical power, the when the inequality is strict, but it weakens as π‘ž approaches existence of the minimizer is uncertain due to the loss of compactness. Due to this, one might be interested in examining the behavior for small values of π‘ž. If Wang’s conjecture is true, in its setting we have ∫ 𝑀 |βˆ‡π‘’|2 + πœ† ∫ Ξ£ 𝑒2 β‰₯ |Ξ£| π‘žβˆ’1 π‘ž+1 πœ†( π‘’π‘ž+1) 2 π‘ž+1 ∫ Ξ£ for πœ†(π‘ž βˆ’ 1) ≀ 1 and π‘ž ≀ 𝑛 π‘›βˆ’2. Let πœ† = 1 π‘žβˆ’1, and we get π‘ž βˆ’ 1 |Ξ£| ∫ 𝑀 |βˆ‡π‘’|2 + 1 ∫ |Ξ£| Ξ£ 𝑒2 β‰₯ ( 1 ∫ |Ξ£| Ξ£ π‘’π‘ž+1) 2 π‘ž+1 Notice that the equality always holds for π‘ž = 1. Now for an arbitrary 𝑒 ∈ 𝐻1(𝑀) satisfying ∫ Ξ£ 𝑒2 = Ξ£, take the limit π‘ž β†’ 1, and we arrive at ∫ 2 𝑀 |βˆ‡π‘’|2 β‰₯ ∫ Ξ£ 𝑒2 log(𝑒2) Consider the functional 𝐸 (𝑒) := 2 ∫ ∫ |βˆ‡π‘’|2 βˆ’ 𝑒2 log(𝑒2), 𝑀 Ξ£ where 𝑒 ∈ 𝐻1(𝑀), ∫ Ξ£ 𝑒2 = |Ξ£| (3.1.1) (3.1.2) 58 Then 𝐸 (𝑒) is bounded below, which will be proved in the next section. Its Euler-Lagrangian equation is Δ𝑒 = 0 πœ•π‘’ πœ•πœˆ = 𝑒 log 𝑒 + πœ†π‘’ πœ† comes from the Lagrangian multiplier. Note that the second equation is not linear, and we could scale 𝑒 to kill πœ†π‘’ to get Δ𝑒 = 0 πœ•π‘’ πœ•πœˆ = 𝑒 log 𝑒 (3.1.3) Just as Wang’s conjecture, we can pose the following conjecture Conjecture 3.1. Let (𝑀, πœ• 𝑀 = Ξ£, 𝑔) be a compact Riemannian manifold with boundary. Suppose 𝑅𝑖𝑐 β‰₯ 0 on 𝑀, and 𝐼 𝐼 β‰₯ 1 on Ξ£ where 𝐼 𝐼 is the second fundamental form, then the following PDE on 𝑀 𝑛 Δ𝑒 = 0 πœ•π‘’ πœ•πœˆ = 𝑒 log 𝑒 admits no solution other than 𝑒 ≑ 1. Consequently, ∫ 2 𝑀 |βˆ‡π‘’|2 β‰₯ ∫ Ξ£ 𝑒2 log(𝑒2) for 𝑒 ∈ 𝐻1(𝑀) and ∫ Ξ£ 𝑒2 = |Ξ£|. (3.1.4) (3.1.5) 3.2 Log-Sobolev Inequality in General Manifold In this section, I will derive a log-Sobolev inequality for general Riemannian manifolds with boundary. Although a log-Sobolev inequality can be obtained using Theorem 2.4 and a similar argument as in the previous section, I will employ a different approach that provides additional information. These methods are modified from the work of [Rot81a], [Rot81b] and [Rot86], where O.Rothaus studied log-Sobolev inequality for manifolds without boundary. 59 For arbitrary 𝜌 > 0, consider the functional 𝐸𝜌 (𝑒) := 𝜌 ∫ ∫ |βˆ‡π‘’|2 βˆ’ 𝑒2 log(𝑒2), 𝑀 Ξ£ where 𝑒 ∈ 𝐻1(𝑀), ∫ Ξ£ 𝑒2 = |Ξ£| (3.2.1) Lemma 3.2.1. π‘ πœŒ > βˆ’βˆž for any 𝜌 > 0, and the infimum can be achieved. 𝑠(𝜌) = inf π‘’βˆˆπ»1 (𝑀) 𝐸𝜌 (𝑒) Proof. Throughout the proof 𝐢 is a constant independent of and 𝑒 and might change from line to line. Without loss of generality we might assume 𝑒 > 0. Fix 0 < πœ– < 2 𝑒2𝑑𝑆 = 1 and log is a concave function, we mighe use Jensen’s inequality for the measure 𝑒2 |Ξ£| ∫ Ξ£ and function π‘’πœ– , and we get π‘›βˆ’2. Since we assume 𝑑𝑆 1 ∫ |Ξ£| Ξ£ 𝑒2 log 𝑒2𝑑𝑆 = ≀ 2 πœ– 2 πœ– ∫ log π‘’πœ– (cid:0) Ξ£ log (cid:0) ∫ Ξ£ 1 Ξ£ 𝑒2 |πœ– | 𝑑𝑆(cid:1) 𝑒2+πœ– 𝑑𝑆(cid:1) Use boundedness of trace operator, 2 πœ– ∫ log (cid:0) Ξ£ 1 Ξ£ 𝑒2+πœ– 𝑑𝑆(cid:1) ≀ log (cid:0)𝐢 βˆ₯𝑒βˆ₯2+πœ– (cid:1) 𝐻1 (𝑀) 2 πœ– 2 + πœ– πœ– ≀ 𝜌βˆ₯𝑒βˆ₯2 ≀ log(βˆ₯𝑒βˆ₯2 𝐻1 (𝑀)) + 𝐢 𝐻1 (𝑀) + 𝐢 (3.2.2) In the last line we used that 𝛼π‘₯ βˆ’ log π‘₯ is bounded below in π‘₯ for any fixed 𝛼 > 0. The proof demonstrating the achievability of the infimum is standard. We can pick a minimizing sequence 𝑒𝑖. 𝑠(𝜌) + 1 β‰₯ 𝐸𝜌 (𝑒𝑖) = 𝐸𝜌/2(𝑒𝑖) + 𝜌 2 ∫ 𝑀 |βˆ‡π‘’π‘– |2 β‰₯ 𝑠( 𝜌 2 ) + 𝜌 2 ∫ 𝑀 |βˆ‡π‘’π‘– |2 So 𝑒𝑖 are bounded in 𝐻1(𝑀). By Alauoglu theorem we can pick a subsequence that converges weakly to 𝑒 in 𝐻1(𝑀), and thus ∫ 𝑒2 𝑖 log 𝑒𝑖 π‘›βˆ’2, and we have |(π‘₯2 log π‘₯)β€²| = |(2 log π‘₯ + 1)π‘₯| < 𝐢 (1 + π‘₯1+πœ– ) for all 𝑀 |βˆ‡π‘’π‘– |2. We only need to check that ∫ converges. Fix πœ– < 1 𝑀 |βˆ‡π‘’|2 ≀ lim ∫ Ξ£ 60 π‘₯ > 0.Therefore ∫ | Ξ£ 𝑒2 𝑖 log 𝑒𝑖 βˆ’ ∫ Ξ£ 𝑒2 𝑗 log 𝑒 𝑗 | ≀ 𝐢 ≀ 𝐢 ∫ Ξ£ ∫ Ξ£ |𝑒𝑖 βˆ’ 𝑒 𝑗 | max{1 + 𝑒1+πœ– 𝑖 , 1 + 𝑒1+πœ– 𝑗 } |𝑒𝑖 βˆ’ 𝑒 𝑗 | (cid:0)1 + 𝑒1+πœ– 𝑖 + 1 + 𝑒1+πœ– 𝑗 (cid:1) ≀ βˆ₯𝑒𝑖 βˆ’ 𝑒 𝑗 βˆ₯ 𝐿2 (Ξ£) (1 + βˆ₯𝑒𝑖 βˆ₯ 𝐿2+2πœ– + βˆ₯𝑒 𝑗 βˆ₯ 𝐿2+2πœ– ) ≀ 𝐢 (1 + βˆ₯𝑒𝑖 βˆ₯𝐻1 (𝑀) + βˆ₯𝑒 𝑗 βˆ₯𝐻1 (𝑀)) βˆ₯𝑒𝑖 βˆ’ 𝑒 𝑗 βˆ₯ 𝐿2 (Ξ£) ≀ 𝐢 βˆ₯𝑒𝑖 βˆ’ 𝑒 𝑗 βˆ₯ 𝐿2 (Ξ£) β–‘ Now we can look at the Euler-Lagrangian equation for 𝑠(𝜌). By a similar computation as the previous section, we get Δ𝑒 = 0 on 𝑀 πœ•π‘’ πœ•πœˆ = πœ†π‘’ log 𝑒 in Ξ£ where πœ† = 2 𝜌 (3.2.3) For any 𝜌 > 0, the function 𝑠(𝜌) is bounded and increasing with respect to 𝜌. Let π‘’πœŒ be its minimizer. It is expected that ∫ 𝑀 |βˆ‡π‘’πœŒ | decreases to 0 as 𝜌 approaches infinity. There might exist a critical 𝜌0 such that 𝑠(𝜌0) is achieved by 𝑒 ≑ 1, enabling the establishment of a log-Sobolev inequality. See section 2 of [Rot81b]. Actually, we have the following stronger theorem. Theorem 3.1. There exists a πœ†0 so that for πœ† < πœ†0, (3.2.3) admits no solution other than 𝑒 ≑ 1. Proof. The proof is similar to theorem 2.4. First we want to bound ∫ we get ∫ Ξ£ 𝑒 log 𝑒 = 0. Then Ξ£ 𝑒. Integrate (3.2.3) by parts, ∫ 𝑒 + 𝑒 Σ∩{𝑒β‰₯𝑒} 𝑒 log 𝑒 (3.2.4) ∫ ∫ Ξ£ 𝑒 = Σ∩{𝑒<𝑒} ∫ ≀ 𝑒|Ξ£| + = 𝑒|Ξ£| + Σ∩{𝑒β‰₯𝑒} ∫ Σ∩{𝑒≀𝑒} 𝑒 log 𝑒 ≀ 𝐢1 61 Let 𝐾 (π‘₯, 𝑦) be the Schwarz kernel for Dirichlet-to-Neumann operator. From section 3.2 we know that for 𝑝 < π‘›βˆ’1 π‘›βˆ’2, βˆ₯𝐾 (π‘₯, Β·)βˆ₯ 𝐿 𝑝 (Ξ£) ≀ 𝐢 for all π‘₯ ∈ Ξ£. Let 0 < 𝑑 < 1, π‘βˆ— = 𝑝 π‘βˆ’1 > 1 and 𝑀 = supπ‘₯∈Σ 𝑒 |𝑒(π‘₯) βˆ’ 1 ∫ |Ξ£| Ξ£ 𝑒| = πœ†| ∫ Ξ£ 𝐾 (π‘₯, 𝑦)𝑒(𝑦) log 𝑒(𝑦)𝑑𝑦| ≀ πΆπœ†βˆ₯𝑒 log 𝑒βˆ₯ π‘βˆ— ≀ πΆπœ†π‘€ 𝑑 βˆ₯𝑒1βˆ’π‘‘ log 𝑒βˆ₯ π‘βˆ— Apparently, there exists a constant 𝐢2( π‘βˆ—, 𝑑) such that π‘₯1βˆ’π‘‘ 𝑝 βˆ— (log π‘₯) π‘βˆ— ≀ 𝐢2 + π‘₯ for all π‘₯ > 0 provided that (1 βˆ’ 𝑑) π‘βˆ— < 1, which is achievable. At maximal point, we have 𝑀 βˆ’ 𝐢1 ≀ |𝑒(π‘₯) βˆ’ 1 ∫ 𝑒| Ξ£ |Ξ£| ∫ ≀ πΆπœ†π‘€ 𝑑 ( 𝐢2 + 𝑒) ≀ πΆπœ†π‘€ 𝑑 Therefore 𝑒 is bounded provided πœ† is bounded above. Then use lemma 2.2.4, we have Ξ£ ∫ |βˆ‡π‘’|2 ≀ 𝐢 ∫ 𝑀 πœ•π‘’ πœ•πœˆ | |2 𝑒 log 𝑒 Ξ£ ∫ Ξ£ ∫ Ξ£ 𝑒 πœ•π‘’ πœ•πœˆ πœ•π‘’ πœ•πœˆ = πΆπœ† ≀ πΆπœ† = πΆπœ† ∫ 𝑀 |βˆ‡π‘’|2 Note constant 𝐢 is independant of πœ† as long as πœ† is bounded above. Therefore ∫ small πœ†, thus 𝑒 ≑ 1. 𝑀 |βˆ‡π‘’|2 = 0 for β–‘ 3.3 Flow Method for Manifolds without Boundary It’s well known that on manifolds without boundary we can solve 𝑒𝑑 = Δ𝑒 and 𝑒 converges to ∫ 𝑒0. If we run this flow and keep track of how ∫ |βˆ‡π‘’|2 decreases in 𝑑, hopefully the constant 1 |Ξ£| we can get something. Actually, this idea works for Gaussian measure π‘‘πœ‡ = It’s well known that ∫ R𝑛 π‘‘πœ‡ = 1, i.e. π‘‘πœ‡ is a probability on R𝑛. Define ∫ E( 𝑓 ) := 𝑓 π‘‘πœ‡ R𝑛 Λ†Ξ” 𝑓 := Ξ” 𝑓 βˆ’ ⟨π‘₯, βˆ‡ 𝑓 ⟩ 1 (2πœ‹)𝑛/2 π‘’βˆ’ | π‘₯ | 2 2 𝑑π‘₯ on R𝑛. 𝑒(𝑑, π‘₯) = 𝑃𝑑 𝑓 := E𝜁 (cid:0) 𝑓 (π‘’βˆ’π‘‘π‘₯ + √︁1 βˆ’ π‘’βˆ’2𝑑 𝜁)(cid:1) 62 Then 𝑒 defined as above solves 𝑒𝑑 = ˆΔ𝑒. Using this flow we can show that Theorem 3.2. If 𝑓 is 𝐢1(R𝑛), E( 𝑓 ) = 1 and E(|βˆ‡ 𝑓 |)2 ≀ ∞, then E( 𝑓 2 log 𝑓 2) ≀ 2E(|βˆ‡ 𝑓 |2) This method can be carried to Riemannian manifolds as follows Theorem 3.3. Let (𝑀, 𝑔) be a compact Riemannian manifold without boundary. Suppose 𝑅𝑖𝑐 β‰₯ (𝑛 βˆ’ 1)𝑔, then for all 𝑓 ∈ 𝐻1(𝑀), 𝑓 > 0 and ∫ 𝑓 = |𝑀 |, we have 𝑀 1 2(𝑛 βˆ’ 1) ∫ |βˆ‡ 𝑓 |2 𝑓 ∫ β‰₯ 𝑓 log 𝑓 If we pick 𝑓 2 in the inequality, we get ∫ 1 𝑛 βˆ’ 1 ∫ |βˆ‡ 𝑓 |2 β‰₯ 𝑓 2 log 𝑓 Proof. Let 𝑒 be solutions of 𝑒𝑑 = Δ𝑒 𝑒(0, Β·) = 𝑓 (3.3.1) Then Note that 𝑒 β†’ πœ• πœ•π‘‘ ∫ (𝑒 log 𝑒) = (log 𝑒 + 1)𝑒𝑑 = (log 𝑒 + 1)Δ𝑒 |βˆ‡π‘’|2 𝑒 𝑒 log 𝑒 = βˆ’ ∫ 𝑀 𝑀 πœ• πœ•π‘‘ ∫ 𝑓 𝑀 |𝑀 | = 1 in 𝐻1(𝑀). Integrate in time, ∫ ∞ ∫ 0 𝑀 |βˆ‡π‘’|2 𝑒 𝑑𝑉 𝑑𝑑 = ∫ 𝑀 𝑓 log 𝑓 (3.3.2) Use Bochner’s formula, we compute πœ• πœ•π‘‘ |βˆ‡|2 = 2βŸ¨βˆ‡π‘’π‘‘, βˆ‡π‘’βŸ© = 2βŸ¨βˆ‡Ξ”π‘’, βˆ‡π‘’βŸ© = Ξ”|βˆ‡π‘’|2 βˆ’ 2|βˆ‡2𝑒|2 βˆ’ 2𝑅𝑖𝑐(βˆ‡π‘’, βˆ‡π‘’) (3.3.3) ≀ Ξ”|βˆ‡π‘’|2 βˆ’ 2|βˆ‡2𝑒|2 βˆ’ 2(𝑛 βˆ’ 1)|βˆ‡π‘’|2 63 Let 𝑣 = √ 𝑒, then 4 πœ• πœ•π‘‘ |βˆ‡π‘£|2 = πœ• πœ•π‘‘ |βˆ‡π‘’2| 𝑒 1 𝑒 πœ•|βˆ‡π‘’|2 πœ•π‘‘ Δ𝑒|βˆ‡π‘’|2 𝑒2 βˆ’ = 4Ξ”|βˆ‡π‘£|2 = Ξ” |βˆ‡π‘’|2 𝑒 Ξ”|βˆ‡π‘’|2 𝑒 Ξ”|βˆ‡π‘’|2 𝑒 = = βˆ’ 2 βˆ’ 2 βŸ¨βˆ‡|βˆ‡π‘’|2, βˆ‡βŸ© 𝑒2 βŸ¨βˆ‡|βˆ‡π‘’|2, βˆ‡βŸ© 𝑒2 1 𝑒 + |βˆ‡π‘’|2Ξ” |βˆ‡π‘’|4 𝑒3 βˆ’ + 2 Δ𝑒|βˆ‡π‘’|2 𝑒2 Take difference between the two equations above and use (3.3.3), 4( πœ• πœ•π‘‘ βˆ’ Ξ”)|βˆ‡π‘£|2 ≀ βˆ’ 2(𝑛 βˆ’ 1)|βˆ‡π‘’|2 𝑒 2(𝑛 βˆ’ 1)|βˆ‡π‘’|2 𝑒 = βˆ’ = βˆ’ 2(𝑛 βˆ’ 1)|βˆ‡π‘’|2 𝑒 ≀ βˆ’ 2(𝑛 βˆ’ 1)|βˆ‡π‘’|2 𝑒 2|βˆ‡2𝑒|2 𝑒 βˆ’ + 2 βˆ’ βˆ’ 2 𝑒 2 𝑒 βˆ‘οΈ 1≀𝑖, 𝑗 ≀𝑛 βˆ‘οΈ 1≀𝑖, 𝑗 ≀𝑛 βŸ¨βˆ‡|βˆ‡π‘’|2, βˆ‡βŸ© 𝑒2 𝑒𝑖 𝑗 𝑒𝑖𝑒 𝑗 𝑒 𝑖 𝑗 βˆ’ 2 (cid:0)𝑒2 |𝑒𝑖 𝑗 βˆ’ 𝑒𝑖𝑒 𝑗 𝑒 |2 Integrate this inequality in both space and time, ∫ ∞ ∫ 0 𝑀 2(𝑛 βˆ’ 1)|βˆ‡π‘’|2 𝑒 βˆ’ β‰₯ = = βˆ’ Ξ”)|βˆ‡π‘£|2𝑑𝑉 𝑑𝑑 4( πœ• πœ•π‘‘ πœ• πœ•π‘‘ ∫ 4 ∫ ∞ ∫ 𝑀 0 ∫ ∞ ∫ 0 ∫ ∞ 0 (cid:0) 𝑀 πœ• πœ•π‘‘ ∫ = lim π‘‘β†’βˆž ∫ = βˆ’ 𝑀 𝑀 |βˆ‡ 𝑓 |2 𝑓 𝑑𝑉 |βˆ‡π‘£|2𝑑𝑉 𝑑𝑑 4|βˆ‡π‘£|2𝑑𝑉 (cid:1) 𝑑𝑑 𝑀 4|βˆ‡π‘£(𝑑, Β·)|2𝑑𝑉 βˆ’ ∫ 𝑀 4|βˆ‡π‘£(0, Β·)|2𝑑𝑉 βˆ’ 2 |βˆ‡π‘’|4 𝑒3 𝑖 𝑒2 𝑒2 𝑗 𝑒2 (cid:1) + (3.3.4) (3.3.5) Combine (3.3.3) and (3.3.5), we get desired result. β–‘ 3.4 Sectional Curvature Results Cut(Ξ£), which is a closed set in the interior of 𝑀 and is of measure zero. Consider πœ“ := 𝜌2 βˆ’ Let 𝜌(π‘₯) = 𝑑 (π‘₯, 𝜎) be the distance from the boundary. It’s smooth away from the cut locus 𝜌2 2 . If (𝑀, Ξ£, 𝑔) is assumed to have non-negative sectional curvature and 𝐼 𝐼 β‰₯ 1, then by the Hessian 64 comparison theorem (cf. [Kas82]), βˆ’βˆ‡2πœ™ β‰₯ 𝑔 Furthermore, πœ“ has nice property near the boundary πœ“Ξ£ = 0 πœ•πœ“ πœ•πœˆ = βˆ’1 These prove advantageous when we use βˆ‡πœ™ as the testing field in (2.4.7). But the problem is cut locus. To overcome this difficulty, in [XX19] the C.Xia and C.Xiong has the following construction. Theorem 3.4. Suppose (𝑀, Ξ£, 𝑔) has non-negative sectional curvature and 𝐼 𝐼 β‰₯ 1. Fix a neigh- borhood C of Cut(Ξ£) in the interior of 𝑀. Then for any πœ– > 0, there exists a smooth non-negative function πœ“πœ– on 𝑀 such that πœ“πœ– = πœ™ on 𝑀 \ C and βˆ’βˆ‡2πœ“πœ– β‰₯ (1 βˆ’ πœ–)𝑔 (3.4.1) In [GHW19], the authors use this function in Wang’s conjecture and get the following Theorem 3.5 (Q.Guo, F.Hang and X.Wang). Let (𝑀, Ξ£, 𝑔) be as in Wang’s conjecture. Then the only positive solutions to (2.1.1) is constant if (π‘ž βˆ’ 1)πœ† ≀ 1 provided 2 ≀ 𝑛 ≀ 8 and 1 < π‘ž ≀ 4𝑛 5π‘›βˆ’9. Consequently, for these (π‘ž, πœ†). π‘ž βˆ’ 1 |Ξ£| ∫ 𝑀 |βˆ‡π‘’|2 + 1 ∫ |Ξ£| Ξ£ 𝑒2 β‰₯ ( 1 ∫ |Ξ£| Ξ£ π‘’π‘ž+1) 2 π‘ž+1 (3.4.2) Their method also works for (3.1.3). Theorem 3.6. Let (𝑀, Ξ£, 𝑔) be as in Wang’s conjecture. Suppose 2 ≀ 𝑛 ≀ 8 Then the only positive solution to (3.1.3) is 𝑒 ≑ 1. Proof. Let 𝑒 be a solution to (3.1.3). Let π‘Ž, 𝑏 bw two constants that will be determined later. Set 𝑒 = π‘£βˆ’π‘Ž, then |βˆ‡π‘£|2 𝑣 Δ𝑣 = (π‘Ž + 1) πœ•π‘£ πœ•πœˆ = 𝑣 log 𝑣 65 (3.4.3) We have the following two lemmas from [GHW19]: Lemma 3.4.1. Suppose πœ™|Ξ£ = 0 and πœ•πœ™ πœ•πœˆ = βˆ’1, then for any smooth 𝑣 and 𝑏 ∈ R ∫ (1 βˆ’ 1 𝑛 )(Δ𝑣)2π‘£π‘πœ™ + 𝑏 2 πœ™π‘£π‘βˆ’2|βˆ‡π‘£|2 (cid:0)3𝑣Δ𝑣 + (𝑏 βˆ’ 1)|βˆ‡π‘£|2(cid:1) π‘£π‘βˆ‡2πœ™(βˆ‡π‘£, βˆ‡π‘£) βˆ’ |βˆ‡π‘£|2π‘£π‘Ξ”πœ™ βˆ’ 𝑏 2 |βˆ‡π‘£|2π‘£π‘βˆ’1βŸ¨βˆ‡π‘£, βˆ‡πœ™βŸ© (3.4.4) 𝑀 ∫ = 𝑀 + (cid:0)|βˆ‡2𝑣 βˆ’ Δ𝑣 𝑛 𝑔|2 + 𝑅𝑖𝑐(βˆ‡π‘£, βˆ‡π‘£)(cid:1)π‘£π‘πœ™ βˆ’ ∫ Ξ£ 𝑣𝑏 |βˆ‡Ξ£π‘£|2 Lemma 3.4.2. The proof of the first lemma is similar to that of usual Reilly’s formula, and the proof of the second one is based on Pohozaev identity (2.4.7). Under the same assumptions as in lemma 3.4.1, we have ∫ 𝑀 1 2 = π‘£π‘βˆ‡2πœ™(βˆ‡π‘£, βˆ‡π‘£) + (𝑣Δ𝑣 + ∫ Ξ£ 𝑣𝑏 (|βˆ‡Ξ£π‘£|2 βˆ’ ( πœ•π‘£ πœ•πœˆ )2) 𝑏 2 |βˆ‡π‘£|2)π‘£π‘βˆ’1βŸ¨βˆ‡π‘£, βˆ‡πœ™βŸ© βˆ’ 1 2 𝑣𝑏 |βˆ‡π‘£|2Ξ”πœ™ Apply these two lemmas for 𝑣 in (3.4.3), we get respectively 𝑄 := = (cid:0)(1 βˆ’ Δ𝑣 𝑛 )(π‘Ž + 1)2 + (cid:0)|βˆ‡2𝑣 βˆ’ 1 𝑛 𝑔|2 + 𝑅𝑖𝑐(βˆ‡π‘£, βˆ‡π‘£)(cid:1)π‘£π‘πœ™ 𝑏(3π‘Ž + 𝑏 + 2) 2 ∫ (cid:1) βˆ’π‘£π‘βˆ‡2πœ™(βˆ‡π‘£, βˆ‡π‘£) + |βˆ‡π‘£|2π‘£π‘Ξ”πœ™ + π‘£π‘βˆ’2|βˆ‡π‘£|4πœ™ 𝑀 𝑏 2 |βˆ‡π‘£|2π‘£π‘βˆ’1βŸ¨βˆ‡π‘£, βˆ‡πœ™βŸ© + ∫ Ξ£ 𝑣𝑏 |βˆ‡Ξ£π‘£|2 ∫ + 𝑀 and ∫ π‘£π‘βˆ‡2πœ™(βˆ‡π‘£, βˆ‡π‘£) + (cid:0)π‘Ž + 1 + 𝑏 2 (cid:1)βˆ‡π‘£|2π‘£π‘βˆ’1βŸ¨βˆ‡π‘£, βˆ‡πœ™βŸ© βˆ’ 1 2 𝑣𝑏 |βˆ‡π‘£|2Ξ”πœ™ 𝑀 1 2 ∫ πœ•π‘£ πœ•πœˆ Combine these two equalities to eliminate terms involving βŸ¨βˆ‡π‘£, βˆ‡πœ™βŸ©, we get 𝑣𝑏 (|βˆ‡Ξ£π‘£|2 βˆ’ ( )2) = Ξ£ (3.4.5) (3.4.6) (3.4.7) 𝑄 =(cid:0)(1 βˆ’ 1 𝑛 )(π‘Ž + 1)2 + + + ∫ 𝑀 ∫ Ξ£ βˆ’ βˆ’ π‘Ž + 1 + 𝑏 π‘Ž + 1 + 𝑏/2 𝑏/4 π‘Ž + 1 + 𝑏/2 𝑣𝑏 ( ∫ (cid:1) π‘£π‘βˆ’2|βˆ‡π‘£|4πœ™ 𝑏(3π‘Ž + 𝑏 + 2) 2 𝑀 π‘£π‘βˆ‡2πœ™βŸ¨βˆ‡π‘£, βˆ‡π‘£βŸ© + π‘Ž + 1 + 3𝑏/4 π‘Ž + 1 + 𝑏/2 |βˆ‡π‘£|2π‘£π‘Ξ”πœ™ πœ•π‘£ πœ•πœˆ )2 + π‘Ž + 1 + 3𝑏/4 π‘Ž + 1 + 𝑏/2 𝑣𝑏 |βˆ‡Ξ£π‘£|2 66 Set π‘Ž + 1 + 3𝑏 4 = 0 to eliminate terms involving Ξ”πœ™ and |βˆ‡Ξ£π‘£|2, and take πœ™ to be πœ™πœ– as in theorem 3.4, we get where π‘„πœ– ≀ (cid:0) (5𝑛 βˆ’ 9 βˆ’ (𝑛 + 9)π‘Ž) (π‘Ž + 1) ∫ 9𝑛 ∫ βˆ’ (1 βˆ’ πœ–) 𝑣𝑏 |βˆ‡π‘£|2 + ∫ (cid:1) π‘£π‘βˆ’2|βˆ‡π‘£|4πœ“πœ– 𝑀 π‘£π‘βˆ‡2πœ“(βˆ‡π‘£, βˆ‡π‘£) + 𝑀\C ∫ Ξ£ 𝑣𝑏 ( πœ•π‘£ πœ•πœˆ )2 C ∫ 𝑀 π‘„πœ– := (cid:0)|βˆ‡2𝑣 βˆ’ Δ𝑣 𝑛 𝑔|2 + 𝑅𝑖𝑐(βˆ‡π‘£, βˆ‡π‘£)(cid:1)π‘£π‘πœ“πœ– Now let πœ– β†’ 0 and then shrink C. Notice that Ξ”πœ™ ≀ βˆ’π‘” whenever its smooth. It yields 𝑄 ≀ (cid:0) (5𝑛 βˆ’ 9 βˆ’ (𝑛 + 9)π‘Ž) (π‘Ž + 1) πœ•π‘£ πœ•πœˆ 𝑣𝑏 |βˆ‡π‘£|2 + 𝑣𝑏 ( 9𝑛 βˆ’ ∫ ∫ )2 𝑀 Ξ£ ∫ (cid:1) 𝑀 π‘£π‘βˆ’2|βˆ‡π‘£|4πœ“πœ– (3.4.8) where 𝑄 := ∫ 𝑀 (cid:0)|βˆ‡2𝑣 βˆ’ Δ𝑣 𝑛 𝑔|2 + 𝑅𝑖𝑐(βˆ‡π‘£, βˆ‡π‘£)(cid:1)π‘£π‘πœ“ Compute ∫ Ξ£ 𝑣𝑏 ( πœ•π‘£ πœ•πœˆ )2 as follows = πœ† ∫ 𝑀 ∫ Ξ£ 𝑣𝑏 ( πœ•π‘£ πœ•πœˆ )2 = πœ† ∫ Ξ£ 𝑣𝑏+1 log 𝑣 πœ•π‘£ πœ•πœˆ 𝑣𝑏+1 log 𝑣Δ𝑣 + (𝑏 + 1)𝑣𝑏 log 𝑣|βˆ‡π‘£|2 + 𝑣𝑏 |βˆ‡π‘£|2 = πœ† ∫ 𝑀 (π‘Ž + 𝑏 βˆ’ 2)𝑣𝑏 log 𝑣|βˆ‡π‘£|2 + 𝑣𝑏 |βˆ‡π‘£|2 Plug this equality in (3.3.5), 𝑄 ≀ (cid:0) (5𝑛 βˆ’ 9 βˆ’ (𝑛 + 9)π‘Ž) (π‘Ž + 1) ∫ 9𝑛 ∫ (cid:1) 𝑀 + (πœ† βˆ’ 1) 𝑣𝑏 |βˆ‡π‘£|2 + πœ†(π‘Ž + 𝑏 βˆ’ 2) π‘£π‘βˆ’2|βˆ‡π‘£|4πœ“πœ– 𝑀 ∫ 𝑀 𝑣𝑏 log 𝑣|βˆ‡π‘£|2 We want π‘Ž+π‘βˆ’2 = 0 since we don’t know the sign for ∫ 𝑀 we get π‘Ž = 2, 𝑏 = βˆ’4. Additionally, we aim for (5π‘›βˆ’9βˆ’(𝑛+9)π‘Ž)(π‘Ž+1) 𝑣𝑏 log 𝑣|βˆ‡π‘£|2. Together with π‘Ž+1+ 3𝑏 4 = 0, , which imposes the condition 9𝑛 𝑛 ≀ 9. This completes our theorem. β–‘ Corollary 3.1. Under the same assumptions, ∫ 2 𝑀 |βˆ‡π‘’|2 β‰₯ ∫ Ξ£ 𝑒2 log(𝑒2) 67 for 𝑒 ∈ 𝐻1(𝑀) and ∫ Ξ£ 𝑒2 = |Ξ£| Remark 3.4.1. If we take derivative with respect to π‘ž at π‘ž = 1 for (3.4.2), just as we did in section 4.1, we get the desired log-Sobolev inequality (3.1.1). But using their method, we also proves non-existence of non-constant solutions to (3.1.3), which is stronger. In [GHW19], the authors applied maximal principle for 𝑛 = 2 and proved the following Theorem 3.7 (Q.Guo, F.Hang and X.Wang). Let (𝑀, Ξ£, 𝑔) be as in Wang’s conjecture and 𝑛 = 2. Then the only positive solutions to (2.1.1) is constant if (π‘ž βˆ’ 1)πœ† ≀ 1 provided π‘ž β‰₯ 2. This maximal principle also works for our case. But since it’s fully covered by the previous result, I won’t include it here. 3.5 Ricci Curvature Results An obstacle in both Wang’s Conjecture 2.1 and Conjecture 3.1 is the lack of a comprehensive understanding of how Ricci curvature affects the Dirichlet-to-Neumann operator. Although some partial results have been obtained under the assumption of sectional curvature β‰₯ 0, as discussed in the previous section and presented in [GHW19], no progress has been made under the condition 𝑅𝑖𝑐 β‰₯ 0. In this section, I will present a result in this direction. Theorem 3.8. (𝑀 𝑛, Ξ£, 𝑔) a Riemannian manifold with boundary. Suppose 𝑛 ≀ 8, 𝑅𝑖𝑐 β‰₯ 0 on 𝑀, II β‰₯ 𝑔Σ and 𝑅𝑖𝑐Σ β‰₯ (π‘›βˆ’2)𝑔Σ on Ξ£, then there exists a πœ†0 that only depends on the dimension so that on 𝑀 𝑛 = πœ†π‘’ log 𝑒 on Ξ£π‘›βˆ’1 Δ𝑒 = 0 πœ•π‘’ πœ•πœˆ ∫ Ξ£ 𝑒2dS ≀ 𝐴(Ξ£) (3.5.1) admit no non-constant solution provided πœ† ≀ πœ†0 68 Proof. Let 𝑒 = π‘£βˆ’π›½, then 𝑣 satisfy the following Δ𝑣 = (1 + 𝛽) |βˆ‡π‘£|2 𝑣 on 𝑀 𝑛 πœ•π‘£ πœ•πœˆ = πœ†π‘£ log 𝑣 on Ξ£π‘›βˆ’1 (3.5.2) Define 𝐸𝑖 𝑗 = 𝑣𝑖 𝑗 βˆ’ Δ𝑣 𝑛 𝛿𝑖 𝑗 and 𝐿𝑖 𝑗 = 𝑣 βˆ’ |βˆ‡π‘£|2 𝑣𝑖𝑣 𝑗 𝑛𝑣 𝛿𝑖 𝑗 . From the work of [LO23], we have (π‘£π‘ŽπΈπ‘– 𝑗 𝑣𝑖) 𝑗 β‰₯ π‘£π‘Ž [𝐸𝑖 𝑗 + π‘Ž + 2(𝛽 + 1) π‘›βˆ’1 𝑛 2 𝐿𝑖 𝑗 ]2 + π‘π‘£π‘Žβˆ’2|βˆ‡π‘£|4 := 𝑄 (3.5.3) where 𝑐 = 𝑛 βˆ’ 1 𝑛 (𝛽 + 1) 2 + 2𝛽 βˆ’ 𝑛 𝑛 βˆ’ 𝑛 βˆ’ 1 4𝑛 [π‘Ž + 2(𝛽 + 1) 𝑛 βˆ’ 1 𝑛 ]2 (3.5.4) In [LO23] Ou and Lin work on the unit ball. The calculation is essentially the same, and the only difference is that we used Bochner formula and finally get an inequality. Pick a frame {𝑒}1β‰€π›Όβ‰€π‘›βˆ’1 along the boundary, and let 𝑒𝑛 = 𝜈 be the outer normal. Integrate by parts, and we have ∫ 𝑀 div(cid:0)π‘£π‘ŽπΈ (βˆ‡π‘£, Β·)(cid:1)dV = = ∫ Ξ£ ∫ Ξ£ π‘£π‘ŽπΈ (βˆ‡π‘£, 𝜈)dS π‘£π‘ŽπΈ (βˆ‡Ξ£π‘£, 𝜈)dS + ∫ Ξ£ We calculate 𝐴 and 𝐡 as follows. π‘£π‘ŽπΈ (π‘£π‘›πœˆ, 𝜈)dS = 𝐴 + 𝐡 (3.5.5) 𝐴 = = ≀ ∫ Ξ£ ∫ Ξ£ ∫ Ξ£ π‘£π‘Žβˆ‡2𝑣(βˆ‡Ξ£π‘£, 𝜈)dS π‘£π‘Ž βŸ¨βˆ‡Ξ£π‘£, βˆ‡Ξ£π‘£π‘›βŸ© βˆ’ π‘£π‘ŽII(βˆ‡Ξ£π‘£, βˆ‡Ξ£π‘£)dS (3.5.6) πœ†π‘£π‘Ž log 𝑣|βˆ‡Ξ£π‘£|2 + (πœ† βˆ’ 1)π‘£π‘Ž |βˆ‡Ξ£π‘£|2dS As for 𝐡, we have Ξ” = ΔΣ + ( πœ• πœ•πœˆ )2 + 𝐻 πœ• πœ•πœˆ on Ξ£. Using (3.5.2), 𝐡 can be calculated as 69 ∫ Ξ£ ∫ Ξ£ ∫ Ξ£ ∫ 𝐡 = = = ≀ Ξ£ )dS π‘£π‘Žπ‘£π‘› π‘£π‘Žπ‘£π‘› (𝑣𝑛𝑛 βˆ’ Δ𝑣 𝑛 (cid:16) (𝑛 βˆ’ 1)(1 + 𝛽) 𝑛 (𝑛 βˆ’ 1)(1 + 𝛽) 𝑛 (𝑛 βˆ’ 1)(1 + 𝛽) 𝑛 |βˆ‡π‘£|2 𝑣 βˆ’ ΔΣ𝑣 βˆ’ 𝐻𝑣𝑛 (cid:17) π‘£π‘Žβˆ’1𝑣𝑛 (|βˆ‡Ξ£π‘£|2 + 𝑣2 𝑛) βˆ’ π»π‘£π‘Žπ‘£2 𝑛 βˆ’ π‘£π‘Žπ‘£π‘›Ξ”Ξ£π‘£dS π‘£π‘Žβˆ’1𝑣𝑛 (|βˆ‡Ξ£π‘£|2 + 𝑣2 𝑛) βˆ’ (𝑛 βˆ’ 1)π‘£π‘Žπ‘£2 𝑛 βˆ’ π‘£π‘Žπ‘£π‘›Ξ”Ξ£π‘£dS (3.5.7) ∫ = (𝛽 + 1)(𝑛 βˆ’ 1) 𝑛 πœ†(π‘Ž + 1 + + πœ†3 (𝛽 + 1)(𝑛 βˆ’ 1) Combine (3.5.5), (3.5.6) and (3.5.7), we have 𝑛 Ξ£ π‘£π‘Ž+2 log3 𝑣 + πœ†π‘£π‘Ž |βˆ‡Ξ£π‘£|2dS )π‘£π‘Ž log 𝑣|βˆ‡Ξ£π‘£|2 βˆ’ (𝑛 βˆ’ 1)πœ†2π‘£π‘Ž+2 log2 𝑣 ∫ 𝑄 ≀ (𝛽 + 1)(𝑛 βˆ’ 1) 𝑛 Ξ£ πœ†(π‘Ž + 2 + + πœ†3 (𝛽 + 1)(𝑛 βˆ’ 1) 𝑛 π‘£π‘Ž+2 log3 𝑣 + (2πœ† βˆ’ 1)π‘£π‘Ž |βˆ‡Ξ£π‘£|2dS )π‘£π‘Ž log 𝑣|βˆ‡Ξ£π‘£|2 βˆ’ (𝑛 βˆ’ 1)πœ†2π‘£π‘Ž+2 log2 𝑣 (3.5.8) Set π‘₯ = (1+𝛽)(π‘›βˆ’1) 𝑛 It becomes , and π‘Ž = βˆ’2 βˆ’ π‘₯ to kill the first term on the right hand side of the above inequality. 𝑄 ≀ ∫ Ξ£ βˆ’(𝑛 βˆ’ 1)πœ†2π‘£βˆ’π‘₯ log2 𝑣 + πœ†3π‘₯π‘£βˆ’π‘₯ log3 𝑣 + (2πœ† βˆ’ 1)π‘£βˆ’π‘₯βˆ’2|βˆ‡Ξ£π‘£|2dS (3.5.9) We further require 𝑃 β‰₯ 0 to make sure 𝑄 β‰₯ 0. (3.5.4) becomes 𝑃(π‘₯) = βˆ’ 𝑛2 βˆ’ 10𝑛 + 1 4𝑛(𝑛 βˆ’ 1) π‘₯2 βˆ’ 1 𝑛 π‘₯ βˆ’ 𝑛 βˆ’ 1 𝑛 (3.5.10) Next we estimate the middle term in (3.5.9) and lower the power for log3 𝑣. ∫ πœ†3π‘₯ Ξ£ π‘£βˆ’π‘₯ log3 𝑣dS = πœ†2π‘₯ = πœ†2π‘₯ ∫ Ξ£ ∫ 𝑀 π‘£βˆ’π‘₯βˆ’1 log2 𝑣𝑣𝑛dS (cid:16) (βˆ’π‘₯ βˆ’ 1) log2 𝑣 + 2 log 𝑣 (cid:17) π‘£βˆ’π‘₯βˆ’1|βˆ‡π‘€ 𝑣|2dV We pick π‘₯ > 0 so that βˆ’π‘₯ βˆ’ 1 < 0. Using (βˆ’π‘₯ βˆ’ 1) log2 𝑣 + 2 log 𝑣 ≀ 𝑐 for constant 𝑐 > 0. Note that π‘Ž only depends on dimension, and therefore 𝑐. So the above equality becomes 70 ∫ πœ†3π‘₯ Ξ£ π‘£βˆ’π‘₯ log3 𝑣dS ≀ πœ†2𝑐π‘₯ ∫ π‘£βˆ’π‘₯βˆ’2|βˆ‡π‘€ 𝑣|2 𝑀 ∫ 𝑐π‘₯πœ†3 βˆ’π‘₯ + 𝛽 𝑐(𝑛 βˆ’ 1)π‘₯πœ†3 π‘₯ βˆ’ 𝑛 βˆ’ 1 Ξ£ = = ∫ Ξ£ π‘£βˆ’π‘₯ log 𝑣dS π‘£βˆ’π‘₯ log 𝑣 (3.5.11) The second equality can be derived if we multiply (3.5.2) by π‘£βˆ’π‘₯βˆ’1 and integrate by parts, which gives πœ† ∫ Ξ£ π‘£βˆ’π‘₯ log 𝑣 = (βˆ’π‘₯ + 𝛽) ∫ 𝑀 π‘£βˆ’π‘₯βˆ’2|βˆ‡π‘€ 𝑣|2. Let 𝑀 = π‘£βˆ’π‘₯/2. Then (3.5.9) and (3.5.11) give us ∫ 𝑄 ≀ 4 π‘₯2 Finally, we want to bound ∫ Ξ£ Ξ£ (2πœ† βˆ’ 1)|βˆ‡Ξ£π‘€|2 βˆ’ (𝑛 βˆ’ 1)πœ†2𝑀2 log2 𝑀 βˆ’ 𝑐π‘₯2πœ†3 2(π‘₯ βˆ’ 𝑛 βˆ’ 1) 𝑀2 log 𝑀dS (3.5.12) 𝑀2 log 𝑀dS by ∫ Ξ£ |βˆ‡Ξ£π‘€|2dS using theorem 3.3 since the coefficient for 𝑀2 log 𝑀 is positive and 𝑅𝑖𝑐Σ β‰₯ 𝑔Σ by our assumption, where 𝑅𝑖𝑐Σ is Ricci curvature on Ξ£. ∫ 𝑀2 2(π‘›βˆ’1) 𝛽 , and π‘₯ 𝛽 < 2 provided π‘₯ β‰₯ 𝑛+1 . By Ξ£ 𝐴(Ξ£) ∫ 𝑀2 𝐴(Ξ£) ≀ 1, and therefore the tail Ξ£ Before that, let us figure out the sign for log and our assumption that ∫ Ξ£ term in theorem 3.3 could be ignored. Now (3.5.12) is HΓΆlder inequality for 𝑒2 ≀ 𝐴(Ξ£), . 𝑀2 = 𝑒 dS 𝐴(Ξ£) π‘₯ ∫ 𝑄 ≀ 4 π‘₯2 𝑐(𝑛 βˆ’ 1)π‘₯2πœ†3 2(π‘₯ βˆ’ 𝑛 βˆ’ 1) 𝑛+1 ≀ π‘₯ < 𝑛 βˆ’ 1 and 𝑃(π‘₯) β‰₯ 0. If we put π‘₯ = 𝑛 βˆ’ 1, 𝑃(𝑛 βˆ’ 1) = (9βˆ’π‘›)(π‘›βˆ’1)2 )|βˆ‡Ξ£π‘€|2 βˆ’ (𝑛 βˆ’ 1)πœ†2𝑀2 log2 𝑀 (2πœ† βˆ’ 1 βˆ’ 4𝑛 Ξ£ where 2(π‘›βˆ’1) (3.5.13) , so admissible π‘₯ could be found provided 2 ≀ 𝑛 ≀ 8. After picking such a π‘₯ that only depends on dimension, 𝑐 in (3.5.11) is also determined. If πœ† > 0 is small enough, we have from (3.5.13) that 0 ≀ 0 and therefore 𝑒 ≑ 1. Corollary 3.2. Let (𝑀, Ξ£, 𝑔) as in theorem 3.8, then ∫ 𝑀 |βˆ‡π‘’|2dV β‰₯ πœ†0 ∫ Ξ£ 𝑒2 log 𝑒dS for ∫ Ξ£ 𝑒2dS = 𝐴(Ξ£). β–‘ (3.5.14) 71 Proof. For each πœ† > 0, we can show that ∫ 𝑀 |βˆ‡π‘’|2dV βˆ’ πœ† ∫ Ξ£ 𝑒2 log 𝑒dS (3.5.15) is bounded below for ∫ Ξ£ 𝑒2 = 𝐴(Ξ£) and the infimum can be achieved. Let π‘Žπœ† be the infimum and 𝑒 be the minimizer. From the Euler-Lagrangian equation, we have Δ𝑒 = 0 on 𝑀 πœ•π‘’ πœ•πœˆ = πœ†π‘’(log 𝑒 + π‘Žπœ†) on Ξ£ 𝑒2dS = 𝐴(Ξ£) ∫ Ξ£ We can scale to get rid of π‘Žπœ†, namely take 𝑣 = π‘’π‘Žπœ†π‘’ ≀ 𝑒. The equation for 𝑣 is Δ𝑣 = 0 on 𝑀 πœ•π‘£ πœ•πœˆ 𝑣2dS ≀ = πœ†π‘£ log 𝑣 ∫ ∫ on Ξ£ 𝑒2dS = 𝐴(Ξ£) Ξ£ The last inequality holds because as the infimum, π‘Žπœ† ≀ 0. For πœ† ≀ πœ†0, 𝑣 is constant by theorem Ξ£ 3.8, and so is 𝑒. Therefore π‘Žπœ† = 0 and the proof is done. β–‘ 72 BIBLIOGRAPHY [Alm12] [Aub76] [BC09] [Bec93] SΓ©rgio Almaraz. β€œConvergence of scalar-flat metrics on manifolds with boundary un- der a Yamabe-type flow”. In: Journal of Differential Equations 259 (2012), pp. 2626– 2694. https://api.semanticscholar.org/CorpusID:119601382. Thierry Aubin. β€œEquations differentielles non lineaires et probleme de Yamabe con- cernant la courbure scalaire”. In: Journal de MathΓ©matiques Pures et AppliquΓ©es 55 (1976), pp. 269–296. https://api.semanticscholar.org/CorpusID:124706153. Simon Brendle and S. Chen. β€œAn existence theorem for the Yamabe problem on manifolds with boundary”. In: arXiv: Differential Geometry (2009). https://api. semanticscholar.org/CorpusID:14140491. William Beckner. β€œSharp Sobolev Inequalities on the Sphere and the Moser–Trudinger Inequality”. In: Annals of Mathematics 138.1 (1993), pp. 213–242. issn: 0003486X. http://www.jstor.org/stable/2946638 (visited on 12/17/2023). [Biq05] Olivier Biquard. β€œAdS/CFT Correspondence: Einstein Metrics and Their Conformal Boundaries”. In: 2005. https://api.semanticscholar.org/CorpusID:260448448. [BL83] HaΓ―m BrΓ©zis and Elliott Lieb. β€œA relation between pointwise convergence of functions and convergence of functionals”. In: Proceedings of the American Mathematical Society 88.3 (1983), pp. 486–490. [BMW13] Eric Bahuaud, Rafe Mazzeo, and Eric Woolgar. β€œRenormalized volume and the evolution of APEs”. In: Geometric Flows 1 (2013). https://api.semanticscholar.org/ CorpusID:118052946. [BN83] [BV91] HaΓ―m BrΓ©zis and Louis Nirenberg. β€œPositive solutions of nonlinear elliptic equations involving critical Sobolev exponents”. In: Communications on pure and applied mathematics 36.4 (1983), pp. 437–477. Marie-FranΓ§oise Bidaut-VΓ©ron and Laurent VΓ©ron. β€œNonlinear elliptic equations on compact Riemannian manifolds and asymptotics of Emden equations”. In: Inventiones mathematicae 106 (1991), pp. 489–539. https://api.semanticscholar.org/CorpusID: 189831176. [CG99] Mingliang Cai and Gregory J. Galloway. β€œBoundaries of zero scalar curvature in the AdS / CFT correspondence”. In: Advances in Theoretical and Mathematical Physics 3 (1999), pp. 1769–1783. https://api.semanticscholar.org/CorpusID:15792435. [Che09] Szu-yu Sophie Chen. β€œConformal Deformation to Scalar Flat Metrics with Constant In: arXiv: Differential Mean Curvature on the Boundary in Higher Dimensions”. 73 Geometry (2009). https://api.semanticscholar.org/CorpusID:115164151. [CLW17] Xuezhang Chen, Mijia Lai, and Fang Wang. β€œEscobar–Yamabe compactifications for Poincaré–Einstein manifolds and rigidity theorems”. In: Advances in Mathematics (2017). https://api.semanticscholar.org/CorpusID:119675719. [CY81] [DJ10] [Esc92a] [Esc92b] [GH17] Jeff Cheeger and Shing-Tung Yau. β€œA lower bound for the heat kernel”. In: Com- https: munications on Pure and Applied Mathematics 34 (1981), pp. 465–480. //api.semanticscholar.org/CorpusID:123113025. Satyaki Dutta and Mohammad Javaheri. β€œRigidity of conformally compact mani- folds with the round sphere as the conformal infinity”. In: Advances in Mathemat- ics 224.2 (2010), pp. 525–538. issn: 0001-8708. https://www.sciencedirect.com/ science/article/pii/S0001870809003727. JosΓ© F. Escobar. β€œConformal deformation of a Riemannian metric to a scalar flat metric with constant mean curvature on the boundary”. In: Annals of Mathematics 136 (1992), pp. 1–50. https://api.semanticscholar.org/CorpusID:124359611. JosΓ©F. Escobar. β€œThe Yamabe problem on manifolds with boundary”. In: Jour- nal of Differential Geometry 35 (1992), pp. 21–84. https://api.semanticscholar.org/ CorpusID:118296174. Matthew J. Gursky and Qing Han. β€œPoincarΓ©-Einstein metrics and Yamabe invariants”. In: arXiv: Differential Geometry (2017). https://api.semanticscholar.org/CorpusID: 119675824. [GHW19] Qianqiao Guo, Fengbo Hang, and Xiaodong Wang. β€œLiouville-type theorems on man- ifolds with nonnegative curvature and strictly convex boundary”. In: Mathematical Research Letters (2019). https://api.semanticscholar.org/CorpusID:209324410. [GL23] [GL91] [Gra16] [GW20] Pingxin Gu and Haizhong Li. β€œA proof of Guo-Wang’s conjecture on the uniqueness of positive harmonic functions in the unit ball”. In: 2023. https://api.semanticscholar. org/CorpusID:259261842. β€œEinstein metrics with prescribed conformal C.Robin Graham and John M Lee. infinity on the ball”. In: Advances in Mathematics 87.2 (1991), pp. 186–225. issn: 0001-8708. https://www.sciencedirect.com/science/article/pii/000187089190071E. C. Robin Graham. β€œVolume renormalization for singular Yamabe metrics”. In: arXiv: Differential Geometry (2016). https://api.semanticscholar.org/CorpusID:119175275. Qianqiao Guo and Xiaodong Wang. β€œUniqueness results for positive harmonic func- tions on B𝑛 satisfying a nonlinear boundary condition”. In: Calculus of Variations 74 and Partial Differential Equations 59.5 (2020), p. 146. [Kas82] [Lee94] Atsushi Kasue. β€œA Laplacian comparison theorem and function theoretic properties of a complete Riemannian manifold”. In: Japanese journal of mathematics. New series 8.2 (1982), pp. 309–341. John M. Lee. β€œThe spectrum of an asymptotically hyperbolic Einstein manifold”. In: Communications in Analysis and Geometry 3 (1994), pp. 253–271. https://api. semanticscholar.org/CorpusID:6645577. [Li12] Peter Li. Geometric analysis. Vol. 134. Cambridge University Press, 2012. [LM90] H. Blaine Jr. Lawson and M. L. Michelsohn. β€œSpin Geometry (Pms-38), Volume 38”. In: 1990. https://api.semanticscholar.org/CorpusID:222238844. [LO23] [LP87] [LQS14] [LY86] [Mar05] [Mar07] Daowen Lin and Qianzhong Ou. β€œLiouville type theorems for positive harmonic functions on the unit ball with a nonlinear boundary condition”. In: Calculus of Variations and Partial Differential Equations 62.1 (2023), p. 34. John M. Lee and Thomas H. Parker. β€œThe Yamabe problem”. In: Bulletin of the American Mathematical Society 17 (1987), pp. 37–91. https://api.semanticscholar. org/CorpusID:263850898. Gang Li, Jie Qing, and Yuguang Shi. β€œGap phenomena and curvature estimates In: Transactions of the American for Conformally Compact Einstein Manifolds”. Mathematical Society 369 (2014), pp. 4385–4413. https://api.semanticscholar.org/ CorpusID:55568779. Peter Li and Shing-Tung Yau. β€œOn the parabolic kernel of the SchrΓΆdinger operator”. In: Acta Mathematica 156 (1986), pp. 153–201. https://api.semanticscholar.org/ CorpusID:120354778. Fernando C. Marques. β€œExistence results for the Yamabe problem on manifolds with boundary”. In: Indiana University Mathematics Journal 54 (2005), pp. 1599–1620. https://api.semanticscholar.org/CorpusID:119502597. Fernando C. Marques. β€œConformal deformations to scalar-flat metrics with constant mean curvature on the boundary”. In: Communications in Analysis and Geometry 15 (2007), pp. 381–405. https://api.semanticscholar.org/CorpusID:55020586. [Maz88] β€œThe Hodge cohomology of a conformally compact metric”. Rafe Mazzeo. In: Journal of Differential Geometry 28 (1988), pp. 309–339. https://api.semanticscholar. org/CorpusID:118675774. 75 [Mul87] β€œCompact hypersurfaces with constant higher order mean Antonio Ros Mulero. curvatures.” In: Revista Matematica Iberoamericana 3 (1987), pp. 447–453. https: //api.semanticscholar.org/CorpusID:118736552. [Nec11] Jindrich Necas. Direct methods in the theory of elliptic equations. Springer Science & Business Media, 2011. [Qin03] Jie Qing. β€œOn the rigidity for conformally compact Einstein manifolds”. In: arXiv: Differential Geometry (2003). https://api.semanticscholar.org/CorpusID:6385348. [Rot81a] OS Rothaus. β€œDiffusion on compact Riemannian manifolds and logarithmic Sobolev inequalities”. In: Journal of functional analysis 42.1 (1981), pp. 102–109. [Rot81b] Oskar S Rothaus. β€œLogarithmic Sobolev inequalities and the spectrum of SchrΓΆdinger operators”. In: Journal of functional analysis 42.1 (1981), pp. 110–120. [Rot86] [SY79a] OS Rothaus. β€œHypercontractivity and the Bakry-Emery criterion for compact Lie groups”. In: Journal of functional analysis 65.3 (1986), pp. 358–367. Richard M. Schoen and Shing-Tung Yau. β€œComplete manifolds with nonnegative scalar curvature and the positive action conjecture in general relativity.” In: Proceedings of the National Academy of Sciences of the United States of America 76 3 (1979), pp. 1024–5. https://api.semanticscholar.org/CorpusID:39981834. [SY79b] Richard M. Schoen and Shing-Tung Yau. β€œOn the proof of the positive mass conjecture in general relativity”. In: Communications in Mathematical Physics 65 (1979), pp. 45– 76. https://api.semanticscholar.org/CorpusID:54217085. [SY94] [Tay96] [Tru68] Richard M. Schoen and Shing-Tung Yau. 1994. https://api.semanticscholar.org/CorpusID:117881865. In: Lectures on Differential Geometry. Michael E. Taylor. Partial Differential Equations II: Qualitative Studies of Linear Equations. 1996. https://api.semanticscholar.org/CorpusID:117308876. Neil S. Trudinger. β€œRemarks concerning the conformal deformation of riemannian In: Annali Della Scuola Normale Superiore structures on compact manifolds”. Di Pisa-classe Di Scienze 22 (1968), pp. 265–274. https://api.semanticscholar.org/ CorpusID:58889899. [Wan02] Xiaodong Wang. β€œA new proof of Lee’s theorem on the spectrum of conformally compact Einstein manifolds”. In: Communications in Analysis and Geometry 10 (2002), pp. 647–651. https://api.semanticscholar.org/CorpusID:56041650. 76 [Wan19] Xiaodong Wang. β€œOn Compact Riemannian Manifolds with Convex Boundary and Ricci Curvature Bounded from Below”. In: The Journal of Geometric Analysis (2019), pp. 1–16. https://api.semanticscholar.org/CorpusID:199511301. [Woo16] Eric Woolgar. β€œThe rigid Horowitz-Myers conjecture”. In: Journal of High Energy Physics 2017 (2016), pp. 1–27. https://api.semanticscholar.org/CorpusID:119273094. [WW21] [WW22] [WY99] [Xu23] [XX19] Xiaodong Wang and Zhixin Wang. β€œOn a Sharp Inequality Relating Yamabe Invari- ants on a Poincare-Einstein Manifold”. In: 2021. https://api.semanticscholar.org/ CorpusID:237491059. Xiaodong Wang and Zhixin Wang. β€œA Sharp Inequality Relating Yamabe Invariants on Asymptotically Poincare-Einstein Manifolds with a Ricci Curvature Lower Bound”. In: 2022. https://api.semanticscholar.org/CorpusID:246294976. Edward Witten and Shing-Tung Yau. β€œConnectedness of the boundary in the AdS / CFT correspondence”. In: Advances in Theoretical and Mathematical Physics 3 (1999), pp. 1635–1655. https://api.semanticscholar.org/CorpusID:6446516. Jie Xu. The Boundary Yamabe Problem, II: General Constant Mean Curvature Case. 2023. arXiv: 2112.05674 [math.DG]. Chao Xia and Changwei Xiong. β€œEscobar’s Conjecture on a Sharp Lower Bound for the First Nonzero Steklov Eigenvalue”. In: Peking Mathematical Journal (2019). https://api.semanticscholar.org/CorpusID:197430742. 77