INTEGRO-DIFFERENTIAL OPERATORS: CONNECTIONS TO DEGENERATE ELLIPTIC EQUATIONS AND SOME FREE BOUNDARY PROBLEMS By Reshma Menon A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of Mathematics — Doctor of Philosophy 2020 ABSTRACT INTEGRO-DIFFERENTIAL OPERATORS: CONNECTIONS TO DEGENERATE ELLIPTIC EQUATIONS AND SOME FREE BOUNDARY PROBLEMS By Reshma Menon In this dissertation, we study aspects of integro-differential operators, and how they relate to different types of equations. In each case, we use information and results about the operators in a lower dimension to analyse an equation in a higher dimension, and vice- versa. We begin in chapter 1 with an introduction to the operators and equations we will be considering. In Chapters 2 and 3, we discuss certain integro-differential operators of functions in a relatively smooth space like C1,α(Rn). However, to understand more about the structure of these operators, particularly about the measure associated with them, we study certain equations in a higher dimension such as degenerate elliptic equations in the upper half space. We analyse the solution of such an equation and its gradient, followed by estimates on its Green’s function and Poisson kernel. These estimates then help reveal some properties of the measure associated with the integro-differential operator in the lower dimension. The structure of the degenerate elliptic equations is similar to that of uniformly elliptic equations, but with an additional complexity of a term which involves distance to the boundary. This degeneracy complicates the analysis; as such, the classical techniques of finding pointwise es- timates as mentioned above do not work so well anymore. So we provide some revised results for the same. Thus understanding an equation in a higher dimension gives us information about an integro-differential operator in a lower dimension. In Chapters 4 and 5, we prove some results about the solutions of free boundary problems in Rn+1 × [0, T ], where the free boundary for a fixed time t can be seen as the graph of a function over a sphere. This time, we connect the solution of the free boundary problem to the solution of a parabolic equation on the sphere – that is, in a lower dimension. This parabolic equation involves an integro-differential operator, which has a min-max representation that is consistent with all the results about viscosity solutions of parabolic equations in Rn. We modify these results for parabolic equations on the sphere, which then gives us existence and uniqueness results about the free boundary problem in a higher dimension. For Amma, the most resilient person I know. iv ACKNOWLEDGMENTS First and foremost, my sincerest gratitude to my dissertation advisor, Russell Schwab – for guiding me through this journey, answering my stupid questions with more patience than I deserve, knowing when to overlook or reprimand my procrastination and balancing these reactions perfectly, encouraging my non-research career choice in teaching for my own happiness, and above all, always standing up for me as a true mentor and friend. Thank you to my dissertation committee – Jun Kitagawa for being instrumental to my understanding of my comprehensive exam syllabus with his teaching and class notes and for being the most chill professor to get a drink with; Keith Promislow for never failing to brighten my day with a big smile and a genuine compliment whenever we crossed paths; and Jeffrey Schenker for always giving my grievances and suggestions a fair hearing as a co-member on committees we have served on together, and the graduate director of the mathematics department. Graduate school to me was initially a stepping stone to what I really wanted – a teaching career where I could have more freedom to plan and teach my own classes, design courses and policies. I could not have imagined having a multitude of these opportunities in the past 6 years itself, leading me to secure a job that I still cannot believe I landed. The people I owe this to are my ultimate teaching mentors, Tsvetanka Sendova and Andrew Krause. Tsveta, thank you for leaving your office door open for me to waltz into anytime to unload my good and bad thoughts on you, for celebrating all my successes and offering a quick pick-me-up from my setbacks, and for giving me the best no-nonsense advice which always worked, regarding teaching, graduate school life, or just human interaction. Andy, than k you for the million conversations about classroom teaching and assessment, for all the last-minute meetings about how-tos for a class, and being editor-in-chief for all my pieces of v writing about my teaching. I owe you both so much, for encouraging and believing in me and getting me to where I am today. To those who have unknowingly been my teaching mentors over the years, I have learnt so much from all of you – Sharon Griffin, Ryan Maccombs, Jeanne Wald, Jane Zimmerman, and Sue Allen. Sue, thank you for all the wine and cheese dinners and the laughs we shared, and for making it so effortless to be your lead TA. I will always cherish our unique friendship. There are many teachers who have inspired my journey as a mathematician and a teacher. Some were mentors, some my own teachers, and some were colleagues. I will be forever grateful towards Casim Abbas for giving me this opportunity to start graduate school at Michigan State University, despite my mixed-levels-of-impressive portfolio. My first year of graduate school at MSU saw my continued love for mathematics thanks to Teena Gerhardt and Ignacio Uriarte-Tuero, who are both responsible in their own ways with their enthusiasm and kindness for making me fall in love with the fields of Algebraic Topology and Analysis respectively. Before I came to even dream of a PhD in mathematics as an option for me, I am thankful to both Sanjay Shinde Sir and Felix Almeida Sir from St. Xavier’s College for inspiring me to fall in love with mathematics and believing that I could get a degree in the subject even though I was registered in the Arts stream. Thanks to Ms. Chethna Amin from Udayachal High School who taught me everything I know about public speaking and life as a young adult, for being the first cheerleader to all my dreams, and for the beautiful bond we continue to share across generations and oceans. A million thanks to all my colleagues at Don Bosco Institute of Technology for giving me some of the best years of my professional life and fortifying my belief that teaching is my life’s calling – especially Sheetal Dighe, Minirani Nair, Anice Matthew, Rohini Chandramouli, Revathy Sunderraman, and Sameer Hadkar. In the last year of my PhD program, I decided to take a peek into the world of Mathe- vi matics Education. I was fortunate to work with people who welcomed me into their research group like old friends, and were open and respectful towards my untrained opinions. Thank you Jack, Mari, Shiv, Bob, Jihye, Sarah, Sofia, Valentin, and Yao – I have enjoyed being a part of your work in the past year, and it has made me so proud to be able to contribute in whatever way I can. Special thanks to Jack Smith for nudging me about my dissertation every chance he got and hence keeping me on my toes. For opening the door to these op- portunities for me, I am ever grateful to Shiv Smith Karunakaran, who became my de facto mentor in this field. Thank you, Shiv, for taking a chance by including me in your research group, for always being generous with your time and advice, especially with navigating the job market. Thanks to Monica Smith Karunakaran for all the helpful job market advice, especially navigating the two-body problem as a female mathematician, and for teaching me everything I’ve learnt about how to teach a brilliantly coordinated course. I want to make a special mention of the staff in the mathematics department, especially those who serve as a liaison between graduate students and faculty. Leslie, for patiently responding to all the phone calls and emails from a very nervous international student; Ami, for our long walks and gossip sessions; Estrella, for her regular kind check-ins and always adding the extra touch of care; Carolyn, for being patient with my endless questions about reimbursements; and Debie, for her warm smile and strong presence in C212. This has been a long and arduous journey, which would have been impossible without the presence of many, many friends in my life who inspire and cheer for me alike. I have been blessed with some beautiful life-long friendships, especially the ones I had back home in India and have traveled halfway across the world with me. Elisha, we are truly lucky to have come to Michigan together, and if it wasn’t for you, I would have never survived the homesickness of those initial years. Maya, Suraj, and Ankita, with whom there is always vii more laughter than I can handle. Ankita, for finding the silver lining and reminding me about things to be thankful for, and Maya, for being the calm voice of reason, I am so happy that my two oldest friends are in a time zone closer to me. I am even more blessed to have friends scattered across the globe, long distance friend- ships which have stood the test of time and distance. Sreya, for being the 4 AM friend irrespective of time zones; Pari, for long weekend conversations about book recommenda- tions and reminiscing about the 90s; Garima and Cheryl, for keeping all the Sundays free for long video calls across 3 continents; Shruti, for honestly calling me out when I am slacking yet telling me what I should be proud of; and Nikita, for salvaging my waning belief in romantic comedies. Thanks to Ajay and Kasturi Aunty for all the help when I was planning my first move to the US, and to Abhishek, for being a fierce friend and pushing me to apply for a PhD in the first place. Uprooting your very well-loved life with an abundance of friends, family, and colleagues you can count on and moving to the other side of the planet is an incredibly difficult choice, made harder once you realise going back home will cost you 3 paychecks and is not an option you have presently available. Thankfully, my army of friends in East Lansing made this easier over time. First, my PhD soulmate, Hitesh – for pushing me through qualifying exams and panic attacks, fighting with me like only family can, inspiring me to make better rajma and broadening my definition of Broadway, thank you for walking every step with me as a true partner in this program. My first roommate and co-parent to my cats, Sarvie, thank you and your family for giving me a home away from home. Thank you to Rani, for all the long conversations and being my most insightful friend no matter what I need advice on; Charlotte, for your supportive outrage when unfair things happened and inspiring my fitness goals; Sarah Klanderman, for being my quals-study partner and introducing me viii to so many of my American firsts; Christos, for always keeping it real; Eylem, for always fighting for what’s right; Allison, for the best surprises and presents; and Sarah Vedolich, for being such a wonderful caretaker to my cats. Thank you to my friends Salil, Charuta, Amit, and Manasi for all the wonderful evenings filled with nostalgia for home. Thank you to Ben and Lauren for bringing music and so much positivity into my life; Guillermo for further inspiring my love for real analysis; Sebastian and Michael for welcoming me into their office space and regaling me with wonderful tales about their little ones; and Rajinder for introducing me to the Pomodoro method for productivity. Thanks to Danika for always reminding me of my awesomeness; Chloe for being the most wonderfully kind neighbour and friend; and Will, Teja, and Grace for our weekly dinners and giving me the community that makes East Lansing home. I am also grateful for the company and advice from good friends like Livy, Stevie, Rami, Sami, Mollee, Darshan, Farhan, Emily, Leo, and Rodrigo. Tyler, thank you for being my go-to person with my analysis questions, for being the kindest and most righteous person I know – inspiring me to be better at my job and in life, for reminding and showing me every single day that we’re on the same team. I am grateful to have you as my life-partner and best friend, and for your ever-loving and supportive parents and grandparents who are now my family. To my sister, Rasya, thank you for believing I am a rockstar and keeping me humble by not saying it out loud. To my parents, Suresh and Jayasree Menon, thank you for raising us with the belief that education always comes first, for pushing my reluctant self to complete my Master’s degree, for unflinchingly letting me go halfway across the world to follow my dreams, and for never failing to encourage our unconventional career choices as women. Finally, I am also grateful for my two cats, Don Quixote and Leopold, for all their warmth and cuddles during the long hours of isolated dissertation writing. ix TABLE OF CONTENTS LIST OF FIGURES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xii KEY TO SYMBOLS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xiii Chapter 1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.1 Degenerate elliptic equations and integro-differential operators . . . . . . . . 1.2 Parabolic equations and some free boundary problems . . . . . . . . . . . . . 2.1 Uniformly elliptic equations Chapter 2 Background for degenerate elliptic equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.1.1 Green’s function . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.1.2 Dirichlet-to-Neumann maps 2.2 Degenerate Elliptic Equations . . . . . . . . . . . . . . . . . . . . . . . . . . 2.2.1 Green’s function . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.2.2 Dirichlet-to-Neumann maps Chapter 3 Connection to Dirichlet-to-Neumann (D-to-N) maps . . . . . . 3.1 Introduction to our equation and coefficients . . . . . . . . . . . . . . . . . . 3.2 Estimates for the Green’s function . . . . . . . . . . . . . . . . . . . . . . . . 3.2.1 Interior estimates . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.2.2 Boundary estimates . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.2.3 Relationship with harmonic measure . . . . . . . . . . . . . . . . . . 3.3 D-to-N maps of the weighted equation (main result and proof) . . . . . . . . Chapter 4 Background on free boundary problems . . . . . . . . . . . . . . 4.1 Introduction and background . . . . . . . . . . . . . . . . . . . . . . . . . . 4.2 Definitions and examples . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.2.1 General Definitions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.2.2 4.3 Results to be modified . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Some examples of free boundary problems 1 1 5 9 10 12 18 20 22 25 27 27 30 30 31 44 45 51 51 56 56 61 66 5.1 Chapter 5 Parabolic equations and free boundary problems . . . . . . . . Introduction to the problem and results . . . . . . . . . . . . . . . . . . . . . 5.1.1 Main results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.1.2 Details of two phase free boundary problem on the sphere . . . . . . 5.2 . . . . . . . . . . . . Integro-differential representations of certain operators 5.3 Comparison theorem and existence for parabolic viscosity solutions on the sphere . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 88 5.3.1 Using the assumptions of section 5.2 . . . . . . . . . . . . . . . . . . 88 5.3.2 Comparison results . . . . . . . . . . . . . . . . . . . . . . . . . . . . 101 5.3.3 Existence (Perron’s method) . . . . . . . . . . . . . . . . . . . . . . . 108 73 73 74 77 79 x 5.4 Revisiting the main results . . . . . . . . . . . . . . . . . . . . . . . . . . . . 112 . . . . . . . . . . . . . . . . . 114 . . . . . . . . . . . . . . . . . . . . 117 5.4.1 Different notions of viscosity solutions 5.4.2 Vertical shifts in the intersurface 5.4.3 Correspondence between viscosity solutions of the free boundary evo- lution and viscosity solutions of the parabolic equation . . . . . . . . 118 5.4.4 Propogation of the modulus of continuity . . . . . . . . . . . . . . . . 120 REFERENCES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 123 xi LIST OF FIGURES Figure 3.1: Boundary conditions for lemma 3.2.2 . . . . . . . . . . . . . . . . . . . . 33 Figure 3.2: Different locations for Bd(X0) . . . . . . . . . . . . . . . . . . . . . . . . 34 Figure 4.1: One phase Hele-Shaw on the half-space . . . . . . . . . . . . . . . . . . . 62 Figure 4.2: Two phase Hele-Shaw on an infinite strip . . . . . . . . . . . . . . . . . . 63 Figure 5.1: Two phase free boundary problem on the sphere . . . . . . . . . . . . . . 78 xii KEY TO SYMBOLS • Rn+1 + , the half space in dimension n + 1, i.e. Rn+1 0}. We sometimes use, N = n + 1 for brevity. + = {X = (x, y) : x ∈ Rn, y ∈ R, y > + of side-length R. R = QR ∩ R+ N . + = Rn, the boundary of the half space. , the partial derivative with respect to xi. For a function u, we will also • ∂RN • ∂i = ∂ ∂xi sometimes denote this as ui or uxi. Further, ∇u = (ux1, ux2,··· , uxn). • QR, a cube centred at 0 in RN • Q+ • Sn, the n-dimensional sphere. • ∂ν, the normal derivative. The notation is also used later for the co-normal derivative. • A, a uniformly elliptic matrix, and ˜A, the matrix with weights. In this work, ˜A = yaA, where a ∈ (0, 1) or a ∈ (−1, 1), as specified. • p.v., Principal Value integral, i.e. if x0 is a point of singularity for an integral(cid:82) Ω f dx, then p.v. • σ, the Lebesgue measure. • ||f||p, the Lp-norm, i.e. Ω |f (x)|pdx ||f||p := • Lp(Ω) = {f|f : Ω → R, ||f||p < ∞}. • Lp(Ω, w), the weighted Lp-space. • Lip(Ω) = {f : Ω → R, f Lipschitz}. • H1,p(Ω), the Sobolev space of Lp functions whose weak first derivatives are also in Lp. • D(u, v), the Dirichlet form, i.e. (cid:82) Ω aij(x)u(x)v(x)dx. • cap(E), the capacity of a set E. • [f ]Cγ , the γ-th H¨older seminorm of f , i.e. [f ]Cγ = sup x,y∈Ω x(cid:54)=y |f (x) − f (y)| |x − y|γ . (cid:90) (cid:90) f dx = lim ε→0 Ω Ω\Bε f dx (cid:19)1/p (cid:18)(cid:90) xiii • C1,γ(Ω) = {f : Ω → R : ||f||L∞ + ||∇f||L∞ + [∇f ]Cγ < ∞}. (equivalently) C1,γ(Ω) = f ∈ L∞(Ω) : sup z∈Ω r−1−γ sup r>0 inf P (x)=c+p·x c∈R, p∈Rn ||f − P||L∞(Br(z)) < ∞  . • For γ ∈ (0, 1), δ > 0, m > 0, the convex set K(γ, δ, m) is defined as K(γ, δ, m) := {f ∈ C1,γ(Ω) : f (x) ≥ δ ∀x ∈ Ω and ||f||C1,γ (Ω) ≤ m}. • K(γ, δ) = K(γ, δ, m). (cid:91) (cid:91) m>0 (cid:91) γ∈(0,1) m>0 • K(δ) = K(γ, δ, m). • For f : Rn → R, Df = {(x, y) ∈ Rn+1, 0 < y < f (x)} • Γf = {(x, y) ∈ Rn+1, y = f (x)} • Similarly, when δ ≤ f ≤ L, f = {(x, y) ∈ Rn+1, 0 < y < f (x)} and D− D+ f = {(x, y) ∈ Rn+1, f (x) < y < L} • ν is the inward facing normal to the boundary of the set D+ • f , and ∂+ ν u = lim t→0+ ∂− ν u = − lim t→0+ u(X0 + tν(X0)) − u(X0) , u(X0 − tν(X0)) − u(X0) t . t xiv Chapter 1 Introduction 1.1. Degenerate elliptic equations and integro-differential operators The first area of my work is in degenerate elliptic equations on a half-space and their Dirichlet to Neumann (D-to-N) maps expressed as integro-differential operators. The D-to-N map appears in wide range of contexts in analysis, probability, Calder´on or inverse problems, and mathematical physics. For example, the D-to-N map arises naturally in the study of operators that describe boundary processes of diffusions in a bounded domain. The map is also central to studying free boundary problems, acting as the natural quantity that drives the free boundary. It is a fundamental object in problems relating voltage to current, where the D-to-N map takes the voltage on the boundary and gives the resulting current density on the boundary. When D-to-N maps are expressed as integro-differential operators, I study the properties of the measure associated with this representation; in particular its relationship with the Lebesgue measure and the density in question. To analyse these properties I proved some estimates of the Green’s function associated with these equations, as well as the connection between the Green’s function and the harmonic measure for the equation. There is a connection between certain partial differential equations on some “nice” domain 1 (like the half space) of the form Lu = −∂j(aij∂iu) = 0 in Ω u = f on ∂Ω (1.1.1) and operators of the kind (cid:90) ∂Ω I(f, x) = f (h) − f (x)µ(x, dh) (1.1.2) for aij and µ that satisfy certain conditions. For example, one possible connection between the two is that the operator I in (1.1.2) will often arise as the D-to N map for (1.1.1). These operators have generated considerable interest lately and it is important to understand when the two situations overlap. We can ask the following two questions – for what µ does the representation in (1.1.2) hold? And what can we say about the order of I, i.e. is I(u(r·), x) = rαI(u, rx)? for some α? Classically, it is known that when L in (1.1.1) is the Laplacian, then the corresponding D-to-N map I will have the structure in (1.1.2), and it is the 1/2-Laplacian of the boundary data f . Subsequently, it is equally natural to consider the α/2−Laplacian for some α instead of the 1/2−Laplacian. As for µ, it is typically reflected in a weight such as µ(x, dh) ≈ |h|−(n+α)dh in this case, as opposed to µ(x, dh) ≈ |h|−(n+1)dh in case of the 1/2-Laplacian. However, to make the connection between the D-to-N map I and the equation in (1.1.2), one must study a weighted equation, which consists of L = div(ya∇u). This has been studied by many authors, for example, Caffarelli and Silvestre [4] and in the context of probability theory, by Song and Vondraˇcek in [36], and also in the book, ‘Bernstein functions. Theory and applications’ [31] by Schilling, Song, and Vondraˇcek. 2 The main goal of the following chapters (2 and 3) is to generalize these results for opera- tors that are not translation invariant, and thus we present work on the weighted extensions Lu = div(yaA∇u), where a ∈ (−1, 1), and the matrix A(X) = (aij(X)) satisfies the follow- ing conditions: • Uniform ellipticity i.e. ∃λ, Λ > 0 for all X, ξ ∈ Rn+1 + , λ|ξ|2 ≤ aij(X)ξiξj ≤ Λ|ξ|2. • A modified version of the Dini-continuity given in [22, Section 3]; we have for X = (x, y) and Z = (z, s) |yaA(X) − taA(Z)| ≤ ω(|X − Z|), where ω is a type of Dini-modulus of continuity, particular to our degenerate equation. ω satisfies the properties in [22, Section 3], and additionally, (cid:90) y−a ω(|X − Z|) |X − Z|N dX ≤ C. Ω Thus, the equation we will study in this work is: Lu = −∂j(yaaij∂iu) = 0 in Rn+1 on ∂Rn+1 u = f + + = Rn (1.1.3) The uniformly elliptic equation (a = 0) has been studied by Guillen, Kitagawa, and Schwab [23]. In this case solutions behave linearly at the boundary, yet the analysis is quite delicate. However, in the weighted setting, we lose uniform ellipticity and the equation degenerates as we approach the boundary. As such, the techniques of [23] no longer apply, the normal derivative can blow up, solutions behave like y1−a instead of being linear, and the analysis becomes more complicated as the proofs to many foundational results need to 3 be adapted to these new challenges. So far we have been able to overcome these technical difficulties to the following result. Theorem 1.1.1. In equation (1.1.2) above, µ satisfies the following conditions (a) µ is absolutely continuous with respect to the surface measure, σ, i.e. for all X ∈ ∂Ω, µ(X,·) has a density, µ(X, dh) = K(X, h)σ(dh). (b) There exist universal constants C1, C2 > 0 so that ∀X, h ∈ ∂Rn+1 + , X (cid:54)= h K(X, h) ≤ C2|X − h|−n+a−1. This extends the uniformly elliptic case treated in [23, Theorem 1.1] to the degenerate elliptic case, for which I have established part (a) and the bound in part (b) for a ≥ 0. The proof of the result in [23] relies on two key steps, namely the relationship between the harmonic measure and the Green’s function found in [3], as well as boundary estimates for the Green’s function found in [22]. When a < 0, we see that writing analogous proofs to [22, Lemmas 3.1,3.2] seem to fail as the estimates blow up the closer you get to the boundary. To prove the first condition on µ, we prove two important lemmas. The first of these gives a relationship between the harmonic measure and the Green’s function – not unlike the well-known result for the uniformly elliptic equation (a = 0) which comes from [3, Lemma 2.2] – but with a change reflecting the term ya. Namely, for x ∈ Rn, y ∈ Rn+1 + \ Bsr for some s > 1, we have constants c1, c2 such that C1r(n+1)+a−2G(y, x + rν(x)) ≤ ωy(∂Ω ∩ Br(x)) ≤ C2r(n+1)+a−2G(y, x + rν(x)) (1.1.4) where ν(x) represents the inward normal vector at x ∈ ∂Rn+1 + . The above result for certain 4 classes of degenerate elliptic equations has been established in [20]. The second result is to obtain lower and upper estimates for the Green’s function associ- ated with the operator L, especially as we get close to the boundary. This will in turn give us the bounds in condition (b) of theorem 1. Thus, there are constants C1, C2 that depend only on λ, Λ, n, ω such that C1 δp(X)δp(Y ) |X − Y |(n+1)−a ≤ G(X, Y ) ≤ C2 δp(X)δp(Y ) |X − Y |(n+1)−a (1.1.5) for a good choice of p, where δ(X) = dist(X, ∂Rn+1 + ). For this, I have obtained the upper bounds when a ≥ 0 with p = 1 − a by first proving that for the solutions of the degenerate elliptic equation in a cube with specific boundary conditions, we have δa(X)|∇u(X)| ≤ C, where C = C(λ, Λ, n, ω). The proof here follows in a similar fashion to the one given by Gr¨uter and Widman in [22, Lemma 3.2] for uniformly elliptic equations. However, the proof only works when a ≥ 0, as when a < 0 the estimate we get on δa(X)|∇u(X)| seems to blow up as you get close to the boundary. 1.2. Parabolic equations and some free boundary problems We analyse a function u : Rn+1 × [0, T ] → R, which is harmonic on the sets {u > 0},{u < 0} ⊂ Rn+1 × [0, T ]. These sets have boundary ∂{u > 0} which, for a fixed time t, is a hypersurface in Rn+1. There is an additional constraint on u which arises naturally from either physical principles or energy minimization, and it concerns the balance of the normal derivatives ∂+ n u+ and ∂− n u− in the inward direction along the free boundary, which is given by the boundary velocity. More formally, for our two-phase problem, we seek a function u 5 that solves under the velocity condition ∆u = 0 in {u > 0} ∪ {u < 0} (1.2.1) ∂{u > 0} moves with normal velocity V = |∇u+| − |∇u−|. (1.2.2) This model includes a special case of the stationary two phase problem under the condition that u does not depend upon time and the balance condition becomes |∇u+| − |∇u−| = 1. (1.2.3) It also includes the one phase version that usually carries the name Hele-Shaw, and that corresponds to ignoring ∂{u < 0} and setting the velocity condition to be V = |∇u+|. I work with these canonical examples, and the goal is to find a solution u that satisfies (1.2.1) above with either of the conditions in (1.2.2) or (1.2.3). Viscosity solutions give a way of handling this when other techniques fail, and there is substantial literature on the variants of these equations in [1, 5, 27, 28]. Recently, the work of Chang-Lara, Guillen, and Schwab [9] gives a new technique which reduces the variants in the family of problems with conditions (1.2.2) and (1.2.3) to an equivalent problem of the integro-differential type by considering the hypersurface ∂{u(·, t) > 0} as the graph of a function f : Rn+1 × [0, T ]. Then, for a sufficiently smooth f , the equation (1.2.1) with the condition in (1.2.2) is equivalent to ∂tf = H(f, x) on Rn × [0, T ] f (·, 0) = f0 on Rn (1.2.4) 6 where H : C1,γ → Cγ, and H(f ) = (|∇U + f |)(cid:112)1 + |∇f|2. The main technique f | − |∇U− used to show the equivalence of solving (1.2.1) under condition (1.2.2) with solving (1.2.4) was given in [9]; the key idea is to use that the operator H has a special structure of the min-max form H(f, x) = min i {aij + cijf (x) + bij ·∇f (x) + p.v. max j (cid:90) Rn f (x + h)− f (x)µij(x, dh)} (1.2.5) for an appropriate family of aij, bij, cij and µij. Equations that admit a similar min-max form are frequently amenable to a large collection of tools from the viscosity solutions context. The assumption that ∂{u > 0} = graphf is not ideal; we see that it does not appear as a requirement in [1, 5, 11, 12, 27, 28]. However, it gives a natural reduction to explore new techniques, especially with an assumption that {u > 0} is a star-shaped domain with respect to X = 0. Thus it makes sense to expand the ideas in [9] to the case of the functions u, f defined on the sphere Sn instead of Rn+1. My work in this area is about expanding this new technique; in particular, when the functions u, f are defined over Sn × [0, T ]. So if time t is fixed, then f (·, t) : Sn → [δ,∞). As such, any previous assumption about translation invariance of the domain or operators now becomes a matter of rotational invariance. The advantage of using the approach of (1.2.4) is that it now allows us to generalise these techniques to many more variations in the free boundary problems described above. One important variant of the Hele-Shaw type problem is to have ∆u = ρ(u) with monotone ρ in {u > 0} and the condition V = k(x, t)|∇u+| in (1.2.3); this was used as a semilinear model for tumor growth [30]. Another equation to study is when ∆u = g(x) in (1.2.1) and the 7 Batchelor-Prandtl equation is a special case of this with the condition (1.2.2) [9, 16, 17]. The second area of my work is in free boundary problems, which have an additional layer of complexity as their domain and boundary are not fixed; but rather, the domain of the function is an unknown in the equation, it is in fact the set where the solution is positive. Furthermore in the type of problem we study, it may evolve in time. When expressed as a function over the d-dimensional sphere, this boundary is a hypersurface that may not be regular. One physical example of this is the motion of a pressurized fluid through a medium with friction, which is modeled by the Hele-Shaw type equations. In fact, we study a generalized version of Hele-Shaw that allows for two phases of a fluid. 8 Chapter 2 Background for degenerate elliptic equations Overall , we are interested in learning more about the connection between partial differential equations on some “nice” domain of the form and operators of the kind Lu = −∂j(aij∂iu) = 0 in Ω u = f on ∂Ω (cid:90) ∂Ω I(f, x) = f (h) − f (x)µ(x, dh) for aij and µ that satisfy certain conditions. One of the possible connections we know of is that the operator I will often arise as the D-to N map for the equation above. Once we know we can express the D-to-N map as an integro-differential operator, we are further interested in studying the properties of the measure µ associated with this representation. We also want to investigate the order of I, i.e. is I(u(r·), x) = rαI(u, rx) for some α? These questions have already been answered for the case of an equation where the coeffi- cients aij are uniformly elliptic, and hence the operator L is translation invariant. Here the 9 order is 1, and the operator is the half-Laplacian, −∆1/2. In this work, we want to show we can create non translation invariant operators of order α, where we have α = 2s = 1 − a for a ∈ (−1, 1). This is to be done by the process of a weighted D-to-N map for degen- erate elliptic equations in the half space with variable coefficients, which is the main goal for chapters 2 and 3. In particular, we will be able to express I in the degenerate case as an integro-differential operator like in (1.1.2), and establish some properties of this measure µ. In the second part of this background chapter, we will provide some background about equations with weighted coefficients with results mainly from [4], [19], and [20]. But first, in the first part of this chapter, we will provide some background about uni- formly elliptic equations, listing the known results about the integro-differential operators of their D-to-N maps. The proofs of these results rely heavily of some results of the Green’s function for uniformly elliptic equations, so we start by first providing a list of results from [22] and [3], and then go on to describe what is known about the structure of µ for the uniformly elliptic operators from [23]. 2.1. Uniformly elliptic equations In this section we will provide some background for the uniformly elliptic equation and the corresponding Green’s function. Most of the following results are standard and appear in many places, but as [3, 22] was a main reference in this dissertation, we will list the relevant definitions and provide references to results only from [3, 22]. We first define the notions of uniformly elliptic equations and operators. Definition 2.1.1. A : Rn → Rn is called uniformly elliptic if ∃ positiveand integro- differ- 10 ential operators constants λ, Λ such that λ|ξ|2 ≤ (cid:104)A(x)ξ, ξ(cid:105) ≤ Λ|ξ|2 ∀x, ξ ∈ Rn (2.1.1) A partial differential equation of the type Luφ = − div(A∇uφ) = 0 in Ω ⊂ Rn uφ = φ on ∂Ω (2.1.2) where A is uniformly elliptic; is a uniformly elliptic partial differential equation. One of the classic and simplest examples of a uniformly elliptic equation is when A = Id, making L = −∆, the Laplacian operator. The set-up for Dirichlet-to-Neumann maps (henceforth D-to-N maps) is as follows. Definition 2.1.2. Given the equation (2.1.2) as above with φ ∈ C1,α(∂Ω), we define a map andintegro − dif f erentialoperatorsI : C1,α(∂Ω) → Cα(∂Ω) as φ (cid:55)→ ∂νuφ, (2.1.3) where ν(x) is the inward normal vector to ∂Ω at x. The simplest possible case of the map I is when A = Id and the domain is Ω = Rn+1, the upper half space. In this case, the inward normal to the boundary is (0, 0,··· , 1), we will have ∂νu = uy. Here u is the harmonic extension of φ, and it is well known that I = (−∆)1/2. This D-to-N map I is the one that has a certain inteand integro- differential operatorsgro- differential representation by the results in [14]. To prove our main goal which is to describe properties of the special measure associated with I, we will need another important measure 11 which is used in the context of a partial differential equation, i.e. the harmonic measure. Another key factor in the proofs we will present are estimates on the Green’s function of the equation. We define tand integro- differential operatorshese concepts below. Definition 2.1.3. Given the operator L as in (2.1.2) and φ ∈ C(∂Ω), there exists a unique uφ that solves (2.1.2). Hence, for a fixed x ∈ Ω, the mapping x (cid:55)→ uφ(x) is well defined. By the comparison principle, this is a non-negative linear functional on C(∂Ω) (i.e. the mapping φ (cid:55)→ uφ(x)). The unique Borel measure that represents this functional is called the L-harmonic measure, and we denote it by ωx. In other words, ωx is uniquely characterized by ∀φ ∈ C(∂Ω), uφ(x) = 2.1.1. Green’s function (cid:90) φ(z)ωx(dz) ∂Ω and integro- differential operators The primary tool of all of our analyses is to understand the boundary behaviour of the Green’s function. So in what follows, I will give the reader the known results of the Green’s function as well as my new results. Definition 2.1.4. Given an operator L and a function f (in some appropriate function space), suppose u is the unique solution of the equation Lu = f u = 0 in Ω on ∂Ω Then the Green’s function is the unique function such that u can be represented as u(x) = G(x, z)f (z)dz (cid:90) Ω 12 The equation (2.1.2) above can be rewritten by using u(cid:48) = u − φ. Then Lu(cid:48) = L(u − φ) = Lu − Lφ = Lφ u(cid:48) = u − φ = 0 in Ω on ∂Ω (cid:90) (cid:90) G(x, z)Lφ(z)dz Thus the Green’s function here is the unique function such that u(cid:48) can be represented as u(cid:48)(x) = G(x, z)Lφ(z)dz =⇒ u(x) = φ(x) + Ω Ω One of first results we are interested in and will later modify and use in this work is the relationship between the Green’s function and the harmonic measure. Proposition 2.1.5. [3, Lemma 2.2] Let {ωx}x∈Ω be the L−harmonic measure where L is as given in the divergence equation in (2.1.2). Then there are universal constants r0, C1, C2, and s > 1 such that for any r ∈ (0, r0), z ∈ ∂Ω, and x ∈ Ω \ Bsr(z) ⊂ Rn, the following holds: C1rn−2G(x, z + rν(z)) ≤ ωx(∂Ω ∩ Br(z)) ≤ C2rn−2G(x, z + rν(z)) and integro- differential operators Further, we will be interested in studying the behaviour of the Green’s function in the interior of the domain, as well as if an how it can be approximated when you get closer to the boundary of the domain. Of course, the latter will depend on how regular the boundary of the domain is. Therefore in the results that follow, we will discuss the estimates on the Green’s function starting with some crude domains that fulfil some minimum requirements (e.g. the cone condition in definition 2.1.6). At first, we will see that we find some very basic H¨older regularity up to the boundary. However, as we improve the conditions on the 13 boundary (e.g. the exterior ball condition (see definition 2.1.7)) we get an improved estimate near the boundary on both the Green’s function and it’s gradient. Naturally, these results are not special to a Green’s function, but to any solution with zero boundary data. Definition 2.1.6. [22, Assumption 1.6] A set Ω is said to satisfy the exterior cone condition if ∃h > 0, 0 < θ < π/2 such that the following is true: for each z ∈ ∂Ω, ∃ a cone C = C(z, h, θ) such that Ω◦ ∩ C = ∅, Ω ∩ C = {z}. Here, C(z, h, θ) denotes a cone with cusp z, height h, and opening angle θ. Definition 2.1.7. A set Ω is said to satisfy the exterior sphere condition if ∃r > 0 such that the following is true: for each z ∈ ∂Ω, there exists a sphere of radius r B = B(z, r) such that Ω◦ ∩ B = ∅, Ω ∩ B = {z}. For a domain that satisfies the exterior cone condition, we can get the following estimates on the solution and the Green’s functions as we approach the boundary. Lemma 2.1.8. [22, Lemma 1.7] Let r > 0 and Dr = Br(0) \ C(0, r/2, θ). Consider the weak solution ur of the following equation with the given boundary conditions Lur = 0 ur = 0 ur = 1 in Dr in ∂C(0, r/2, θ) in ∂Br(0) Then, ∃ K(n, λ, Λ, θ) > 0, α(n, λ, Λ, θ) ∈ (0, 1) such that ∀x ∈ Dr ur(x) ≤ K |x|α rα 14 The above lemma zooms into a portion of the domain close to the boundary, by con- structing cone at a point x0 on the boundary and taking a little ball around the point which of course overlaps with a little part of the domain close to the boundary. For convenience, this point is taken to be x0 = 0 and this works for any operator L with bounded, measurable coefficients. We look at Br(0) because the ellipticity class is invariant by rescaling, i.e. if we have ˜aij(x) = aij(rx) or ˆaij(x) = aij(x/r), then the ellipticity constants of ˜aij, ˆaij are also λ, Λ. The lemma redefines the boundary values in this little domain Dr which comprises a small neighbourhood around this boundary point. As mentioned earlier, this is a very basic H¨older estimate. This lemma then leads us to a a result about the boundary regularity of the Green’s function, as below. Theorem 2.1.9. [22, Theorem 1.8] There are constants K(n, λ, Λ, θ, diam Ω, ∂Ω) > 0, α(n, λ, Λ, θ) ∈ (0, 1) such that ∀x, z ∈ Ω G(x, z) ≤ Kδα(z)|x − z|2−n−α where δ(z) = dist(z, ∂Ω). If we improve the boundary of our domain to one that satisfies the exterior sphere con- dition, we will get better estimates for the solutions and the Green’s function. But first, we also require some special assumptions on the coefficients aij [22, Section 3]. Definition 2.1.10. The coefficient matrix A is called Dini-continuous if it satisfies, |aij(x) − aij(z)| ≤ ω(|x − z|) (2.1.4) 15 where ω : R+ → R+ is non-decreasing and it satisfies the doubling condition, ω(2t) ≤ Kω(t) for some K > 0 and all t > 0 and 0 t (cid:90) 1 ω(t) dt < ∞. (2.1.5) (2.1.6) With coefficients as given above, we have the following two lemmas from [22] regarding the size of the gradient of the solution. These estimates are then used to study the behaviour of the Green’s function close to the boundary. Lemma 2.1.11. [22, Lemma 3.1] Suppose u is a bounded solution of Lu = 0 in Ω. with L as in (2.1.2) and Dini-continuous coefficients aij as described above. Then ∃ K = (n, λ, Λ, ω, Ω) > 0 such that for any x ∈ Ω |∇u(x)| ≤ Kδ−1(x) sup |u|, Ω where δ−1(x) := dist(x, ∂Ω). In the above result we assume no special conditions about the domain Ω. As such, we note that when you get closer to ∂Ω, it is quite possible for the gradient to blow up. The following lemma describes a bound of the gradient of the solution on a much nicer and integro- differential operatorsdomain, i.e. an annulus. Lemma 2.1.12. [22, Lemma 3.2] Let u be the solution of the Dirichlet problem (with Dini 16 continuous coefficients aij) Lu = 0 u = 0 u = 1 in Dr = B2r \ Br in ∂Br in ∂B2r Then ∃ K = (n, λ, Λ, ω) > 0 such that for any x ∈ Dr, r ∈ (0, 1], |∇u(x)| ≤ K r . The above lemmas are used powerfully in proving the following theorem about the es- timates of the Green’s function. If we have a domain that satisfies the exterior sphere condition, we can construct an annulus of a suitable radius at every point on the boundary of the domain, with the inner sphere touching the boundary on the outside. We then em- ploy the same techniques used in the proof of 2.1.9 to prove the boundary regularity of the Green’s function. Theorem 2.1.13. [22, Theorem 3.3] Let Ω be a domain that satisfies the exterior sphere condition and let L satisfy (2.1.1)–(2.1.6). Let G be the corresponding Green’s function. Then the following inequalities are true for any x, z ∈ Ω (a) G(x, z) ≤ K|x − z|2−n, K = K(n, λ, Λ) (b) G(x, z)δ(x) ≤ K|x − z|1−n, K = K(n, λ, Λ, ω, Ω) (c) G(x, z) ≤ Kδ(x)δ(z)|x − z|−n, |∇xG(x, z)| ≤ K|x − z|1−n, (d) 17 |∇xG(x, z)| ≤ Kδ(z)|x − z|−n, |∇x∇zG(x, z)| ≤ K|x − z|−n, (e) (f ) 2.1.2. Dirichlet-to-Neumann maps We seek to study the relationship between the equation (1.1.1) and the operator in (1.1.2) in a degenerate elliptic setting. However, we already have a lot of information about this I, which arises as the Dirichlet-to-Neumann map in the case of the uniformly elliptic equation (see definition 2.1.2). In this section, we provide some background and results about I as a D-to-N map, its integro-differential representation, and properties of the associated measure. To begin with, this map I is not only well defined from C1,α(∂Ω) to Cα(∂Ω), but it also satisfies the global comparison property as defined below. Definition 2.1.14. The global comparison property (GCP) for I : C1,α(X) → C0(X) re- quires that for all f, g ∈ C1,α(X) such that f (x) ≤ g(x) for all x ∈ X and f (x0) = g(x0) for some x0 ∈ X, then the operator I satisfies I(f, x0) ≤ I(g, x0). That is to say that I preserves the ordering on the functions f, g on X at any points where the graphs of f, g touch. In the case of the divergence equations, L, and hence also I are linear operators. It was proved in the 1960s, by Courr`ege [14] and Bony-Courr`ege-Priouret [2], through linearity and the global comparison property, that I must be an integro-differential operator of the form (cid:90) I(φ, x) = b(x) · ∇φ(x) − p.v. φ(x + h) − φ(x)µ(x, dh). ∂Ω Following is the main theorem from [23]. We will prove the analogous result for degenerate elliptic equations in the next chapter. 18 Theorem 2.1.15. [23, Theorem 1.1] Suppose L is the divergence operator as in (2.1.2) with Ω = Rn+1 a family of measures parametrised by x, µ(x, dh) such that ∀φ ∈ C1,α(Rn), + and I is as defined via (2.1.2) and (2.1.3), then there exists a vector field, b, and (cid:90) Rn I(φ, x) = (cid:104)b(x),∇φ(x)(cid:105) − Further, µ satisfies φ(x + h) − φ(x) − 1B1(x)(h)(cid:104)∇φ(x), h − x(cid:105) µ(x, dh). (i) For all x ∈ Rn, µ(x,·) has density µ(x, dh) = K(x, h)σ(dh), (ii) There exists universal c1 > 0, c2 ≥ c1 such that for all x, h ∈ Rn, x (cid:54)= h, c1|x − h|−n−1 ≤ K(x, h) ≤ c2|x − h|−n−1 In this dissertation, we explore the answers to these questions in a setting where the coefficients are not uniformly elliptic; in fact they come attached with variable coefficients and weights making them degenerate as we get closer to the boundary of the domain. As a result, we are dealing with operators that are no longer translation invariant. This means many of the classically known results, particularly about the Green’s function may no longer be valid or have modified versions to suit the new conditions. In the next section, we discuss some degenerate elliptic equations with weights similar to the ones in our work, and well known results about the Green’s function of these equations. 19 2.2. Degenerate Elliptic Equations Consider the following type of degenerate elliptic equation in divergence form Lu = −∂i(˜aijuj) = 0 in Rn+1 + on Rn u = f (2.2.1) The coefficients ˜aij will be real-valued, measurable, symmetric, and satisfy λ|Z|2w(X) ≤ ˜aij(X)ZiZj ≤ Λ|Z|2w(X) (2.2.2) for all X = (x, y), Z = (z, s) ∈ Rn+1 + . The weight w(X) will be a non-negative, measurable function satisfying Muckenhoupt’s condition, or the A2 condition, which is 1 dX ≤ C. (2.2.3) (cid:90) (cid:90) (cid:18) 1 (cid:19)(cid:18) 1 (cid:19) Here the supremum is taken over all Euclidean balls B and(cid:82) a set E, we denote w(E) =(cid:82) w(X) B w(X)dX sup B |B| B |B| B 1dX = |B|. Further, for E w(X)dX. Another way of writing the equation above would be to denote ˜A(X) = w(X)A(X), where A is uniformly elliptic. Now if w(X) satisfies the A2 condition (2.2.3), then following are two well known facts about the measure w(X)dX: (a) w(X)dX and σ(dX) are mutually absolutely continuous. (b) w(B(X, 2r)) ≤ cw(B(X, r)) (doubling condition. The above description of the degenerate elliptic equation is taken from [19]. In this section, we will list many of the results in [19], many of which we will adapt later in chapter 20 3 to our particular version of the degenerate elliptic equation given in (2.2.1). First, we start by listing some commonly known function spaces with their definitions which will show up in multiple results in this section and the next chapter. Definition 2.2.1. (Some function spaces): (a) Lp(Ω) is the set of all functions f such that their Lp-norm, ((cid:82) (b) Lp(Ω, w) is the Lebesgue class with the norm ((cid:82) Ω |f (x)|pdx)1/p, is finite. Ω |f (x)|pw(x)dx)1/p, or the weighted Lp-space. (c) Lip(Ω) is the set of all functions f on Ω that satisfy the Lipschitz condition, i.e. |f (x) − f (y)| ≤ M|x − y| for some M . (d) If we consider the inclusion from Lip(Ω) → [Lp(Ω, w)]n+1 given by the mapping f (cid:55)→ (f,∇f ) = (f, fx1, fx2,··· , fxn), then H1,p(Ω) denotes the closure of the image of Lip(Ω) in [Lp(Ω, w)]n+1. Essentially, H1,p(Ω) is the space of Lp functions on Ω whose weak first derivatives are also in Lp(Ω). (e) H1,p 0 (Ω) denotes the closure of compactly supported functions in the image of Lip(Ω) in [Lp(Ω, w)]n+1. i.e. f ∈ H1,p 0 (Ω) if and only if there exist functions fm ∈ Lip(Ω) such that fm are all compactly supported in Ω and fm → f in H1,p(Ω). We further describe a notion of convergence in these Sobolev spaces, and the capacity of a set, which later appears in the interior estimates of the Green’s function. Definition 2.2.2. Let K ⊂ Ω. We say that u ≥ c on K in the H1,2(Ω) sense if ∃ ϕj ∈ Lip(Ω) such that ϕj ≥ c ∀x ∈ K and ϕj −→ u in H1,2(Ω). 21 Definition 2.2.3. The Dirichlet form D : H1,2(Ω) × H1,2(Ω) → R is defined by D(u, v) := ˜aiju(x)v(x)dx. Definition 2.2.4. Let K ⊂ Ω be compact. The capacity of K in Ω is given by Ω cap(K) := inf{D(u, u) : u ∈ H1,2 0 (Ω), u ≥ 1 on K in the H1,2(Ω) sense.} For an open set Θ in Ω, cap(Θ) := sup{cap(K) : K compact, K ⊂ Θ}. 2.2.1. Green’s function Definition 2.2.5. The Green’s function for the degenerate equation is defined in a similar manner as the elliptic case. Given the equation Lu = f in Ω with L as in (2.2.1), the Green’s function is the unique function such that u can be represented as (cid:90) (cid:90) Ω (cid:90) u(x) = G(x, z)f (z)dz As before, we can rewrite this equation u(cid:48) = u − φ and get u(x) = φ(x) + G(x, z)Lφ(z)dz Ω Remark 2.2.6. Another way of looking at the Green’s function is that is we fix z ∈ Ω, then the Green’s function denoted by G(x, z) is the weak solution of Lxu(x) = δz(x) as a function 22 of x, where δz is the unit mass at z. Thus, we claim that LxG(x, z) = δz(x).This is because Ω Ω (cid:90) (cid:90) (cid:90) (cid:90) (cid:90) δz(x)u(x)dx. (by the definition of δz) G(x, z)f (x)dx (by the definition of G(x, z)) G(x, z)Lxu(x)dx (since Lxu(x) = f (x)) LxG(x, z)u(x)dx (since L is self-adjoint) LxG(x, z)u(x)dx first, u(z) = Next, u(z) = = = δz(x)u(x)dx = (cid:90) Ω =⇒ Ω Ω Ω Thus we have, as claimed, LxG(x, z) = δz(x). Now we list results from [19, Section 3] about the size of the Green’s function for the degenerate elliptic equation described in (2.2.1), which will be useful for finding the estimates of the Green’s function to the case of the specific degenerate elliptic equation we will look at in the next chapter. Lemma 2.2.7. [19, Lemma 3.1] If B(x, 2r) ∈ Ω and z ∈ ∂B(x, r), then G(x, z) ∼= 1/ cap(B(x, r)). Lemma 2.2.8. [19, Lemma 3.2] If x ∈ Ω and 3 2 r ≤ dist(x, ∂Ω) ≤ 8r, then cap(B(x, r)) (cid:39) w(B(x, r))/r2. Using these lemmas, we get an estimate for the Green’s function in the interior of the domain. 23 Theorem 2.2.9. [19, Theorem 3.3] Let BR ⊂ Ω and x, z ∈ BR/4. Denote r = |x− z|. Then (cid:90) R G(x, z) (cid:39) s2 w(B(x, s)) r ds s . (2.2.4) The above result gives interior estimates for the Green’s function, and we will still need results about the behaviour of the Green’s function as we approach the boundary, which will take up the entire first section in chapter 3. Additionally, we also need to describe some relationship between the Green’s function and the harmonic measure (same as defined in definition 2.1.3) for the equation (2.2.1). There is such a result in [20] which concerns nontangentially accessible (NTA) domains. One of the conditions in the definition of an NTA domain Ω is that ∃A > 1, r0 > 0 such that ∀r, 0 < r < r0, ∀z ∈ ∂Ω, ∃z(cid:48) ∈ Ω such that |z − z(cid:48)| < Ar and B(z(cid:48), r/A) ⊂ Ω. In a sense, this is an interior ball condition analogous to the exterior sphere condition described in definition 2.1.7. Lemma 2.2.10. [20, Lemma 3] Let z ∈ ∂Ω, z(cid:48) ∈ Ω be as above in the definition of an NTA domain, and let {ωx}x∈Ω be the L-harmonic measure where L is as given in (2.2.1). x ∈ Ω \ B(z, 4Ar) then If G(z(cid:48), x) (cid:39) ωx(∂Ω ∩ B(z, r)) r2 w(B(x, r)) This is of course analogous to proposition 2.1.5, by considering z(cid:48) = z + rν(z). 24 2.2.2. Dirichlet-to-Neumann maps Now that we are equipped with sufficient background about degenerate elliptic equations, we will look at the D-to-N maps of these equations. A canonical result of this is the frac- tional Laplacian which we mentioned in Chapter 1. Let us consider a fairly simple type of degenerate elliptic equation in the half-space in Rn+1. For X ∈ Rn+1 where x ∈ Rn, and y > 0. Also note that ∂Rn+1 = Rn. + , we denote X = (x, y), Lu = − div(ya∇u) = 0 u = f in Rn+1 + in Rn (2.2.5) This is an equation that looks like (2.2.1), and the weights as given in (2.2.2) are ya. where a ∈ (−1, 1). Thus, ˜aij(X) = aij(X)w(X) = yaId. Thus w(X) = w(x, y) = ya and A(x) = Id, which is uniformly elliptic. Among the many authors to study this equation in the context of the fractional Laplacian are Caffarelli and Silvestre [4]. Definition 2.2.11. The fractional Laplacian of a function f : Rn → R is expressed by (cid:90) (−∆)sf (x) = Cn,s f (x) − f (z) |x − z|n+2s dz Rn where the parameter s ∈ (0, 1) and Cn,s is some normalization constant. One of the main results of [4] is the connection between the fractional Laplacian and the D-to-N map for the equation (2.2.5). The D-to-N map is given by the co-normal derivative of the solution of the equation. Definition 2.2.12. The co-normal derivative for an equation is defined with the help of Green’s identity. In particular, it is the term that appears in the integral on the boundary. 25 For the above equation (2.1.2), using, integration by parts, we have − div(ya∇u) · ϕdX = ya∇u · ∇ϕdX − ya∇u · ϕ · νdS (cid:90) Rn+1 + (cid:90) Rn+1 + (cid:90) Rn Since the inward facing normal to the upper half space is (0, 0, . . . , y), the last term above is (cid:82)Rn ϕ · (yauy)dS. The co-normal derivative is the term multiplied to the test function, i.e. −yauy. Theorem 2.2.13. [4, Section 3] For u, the solution of (2.2.4), we have up to a constant factor, that the co-normal derivative is the fractional Laplacian of order s = (1 − a)/2, i.e. yauy = (−∆)sf (x) lim y→0+ Note that s ∈ (0, 1) =⇒ a ∈ (−1, 1). yauy = (−∆) 1−a 2 f (x) = Cn,s lim y→0+ (cid:90) Rn f (x) − f (z) |x − z|n+1−a dz This also fits with our earlier mention of the D-to-N map in the case of the Laplacian, i.e. when a = 0, we have I = (−∆)1/2. Now for this equation, say we would like to develop a Poisson formula P to explicitly solve (2.2.5) i.e. we want P that satisfies for X = (x, y) ∈ Rn+1 + , Z ∈ ∂Rn+1 + = Rn (cid:90) Rn u(X) = P (X, Z)f (Z)dZ We will also denote this Poisson Kernel P as P (X, Z) = PX (Z). Another notable comment about the Poisson kernel P is that it is the co-normal derivative of the Green’s function. 26 Chapter 3 Connection to Dirichlet-to-Neumann (D-to-N) maps 3.1. Introduction to our equation and coefficients In this work, we will be proving many of the above results for a special case of the degenerate elliptic equation set in the upper half-space. For some f ∈ C1,α(Rn+1 + ), let u solve the following equation: Lu = − div(yaaijui) = 0 u = f in Rn+1 + in Rn (3.1.1) where [aij] = A is measurable, symmetric, and uniformly elliptic. At times, we will use the notation ˜aij to denote yaaij, or ˜A to denote yaA. Thus, ˜aij are like the weighted coefficients in (2.2.2) with weights w(X) = ya. Notice that when a = 0, we get the uniformly elliptic case that was discussed in chapter 2. However, if a (cid:54)= 0, then this equation degenerates as you get closer to the boundary, i.e. Rn, where y = 0. Now, we will consider a special version of Dini-continuity for our coefficients ˜aij, namely for any bounded subset Ω ⊂ Rn+1 + we have (cid:90) Ω y−a|˜aij(X) − ˜aij(Z)| |X − Z|N dX ≤ C (3.1.2) 27 where X, Z ∈ Rn+1 + , X = (x, y), Z = (z, s), a ∈ (0, 1). The reason for imposing this extra condition on the coefficients will become evident in the proofs of the results about the Green’s function to this equation, particularly lemma 3.2.2. Before we look at the D-to-N map for this equation, we will need a definition for the co-normal derivative in this setting. Definition 3.1.1. The co-normal derivative for an equation is defined with the help of Green’s identity, just as in the definition 2.2.12. In particular, it is the term that appears in the integral on the boundary. For the above equation (3.1.1), using, integration by parts, we have (cid:90) Rn+1 + − div(yaaijui) · ϕdX = (cid:90) Rn+1 + yaaijui · ϕjdX − (cid:90) Rn yaaijui · ϕ · νdS Since the inward facing normal to the upper half space is (0, 0, . . . , en+1), the last term above j an+1,juj)dS. The co-normal derivative is the term multiplied to the test is (cid:82)Rn ϕ · (ya(cid:80) function, i.e. −ya(cid:80) j an+1,juj. Remark 3.1.2. There are, in essence, two choices for the “normal” derivative. This is because for the divergence equation, there are two possible normal vectors we can consider. One of them is the natural inward normal vector we see geometrically, which in the case of the half-space is just (0,··· , 0, 1). We can see that this is the normal vector taken into consideration for the definition 2.2.12 of the co-normal derivative for the equation 2.2.5. The other choice of normal vector comes from the equation, i.e. (an+1,1,··· , an+1,n+1) as in the definition 3.1.1 above. This is still an inward normal vector, but it is weighted by the coefficients yaaij. Since aij are uniformly elliptic, the distance between this normal vector and the conventional geometric normal vector are comparable up to a constant. 28 The Dirichlet-to-Neumann map for this equation (3.1.1) is given by this weighted normal derivative, i.e. I(f, x) = −ya∂νu = −yauy. This map I is not only well defined from C1,α(∂Ω) to Cα(∂Ω), but it also satisfies the global comparison property as defined in definition 2.2.1. Again, here in (3.1.1), L is a linear operator and hence so is I. By a result of Bony-Courr`ege-Priouret [2] in the 1960s using linearity and the global comparison property, it is known that I can be expressed as an integro-differential operator of the form (cid:90) I(φ, x) = b(x) · ∇φ(x) − p.v. φ(x + h) − φ(x)µ(x, dh). ∂Ω The main result in this chapter is a modification of the theorem 2.1.15 from [23] for this particular elliptic equation which degenerates as you get close to the boundary. Theorem 3.1.3. Suppose L is the divergence operator as in (3.1.1) with Ω = Rn+1 and a ∈ (0, 1). If I is as defined via (2.1.2) and (2.1.3), then we know there exists a vector field, b, and a family of measures parametrised by x, µ(x, dh) such that ∀φ ∈ C1,α(Rn), + φ(x + h) − φ(x) − 1B1(x)(h)(cid:104)∇φ(x), h − x(cid:105) µ(x, dh). (cid:90) Rn I(φ, x) = (cid:104)b(x),∇φ(x)(cid:105) − Further, µ satisfies (i) µ is absolutely continuous with respect to the surface measure, σ, i.e. for all x ∈ Rn, µ(x,·) has density µ(x, dh) = K(x, h)(dh), (ii) There exists universal C > 0 such that for all x, h ∈ Rn, x (cid:54)= h, K(x, h) ≤ C|x − h|−n+a−1. 29 The proof of this theorem relies on many properties and results about the Green’s function associated with (3.1.1), which we will outline in the next section. 3.2. Estimates for the Green’s function We start by finding estimates for the Green’s function of (3.1.1) as defined in definition 2.1.4, first in the interior of our domain, and then closer to the boundary. 3.2.1. Interior estimates Since we have the the weighted coefficients, i.e. yaaij, we know that the A2 weights for (3.1.1) are w(X) = ya. The following result for the interior estimate for the Green’s function are a direct result of applying the results in [19, Section 3]. Since our domain is a half-space, we will consider cubes that are contained in the domain instead of the spheres/balls that appear in lemmas 2.2.7, 2.2.8 for the proofs. Following is the modified version of theorem 2.2.9, which is [19, Theorem 3.3]. Theorem 3.2.1. Let X, Z ∈ Q Then 1/4 R = {(m, n+3/4) : M = (m, n) ∈ QR} and let r = |X−Z|. (cid:90) R G(X, Z) (cid:39) w(B(X, t)) r t2 dt t The proof of this theorem relies on the two aforementioned lemmas 2.2.7, 2.2.8. Using this theorem and w(X) = w(x, y) = ya we can approximate the size of the Green’s function for our degenerate equation if we let P = (p, q) ∈ B(X, t), then w(B(X, t)) = w(P )dP = qadP (cid:39) Ctn+1+a (cid:90) (cid:90) B(X,t) B(X,t) 30 (cid:90) R t2 w(B(X, t)) r dt t = Then (cid:90) R t2 |X−Z| Ctn+1+a (cid:90) R |X−Z| Ct−n−adt (cid:39) C|X − Z|−n−a+1 dt t = But 1 − n − a = 2 − (n + 1) − a = 2 − N − a and thus ∃ C = C(R, n) G(X, Z) (cid:39) C|X − Z|2−N−a One can also write an alternate lengthier proof using the techniques in [29]. 3.2.2. Boundary estimates In this section, we will modify some of the results in [22] to suit the degenerate elliptic equation in (3.1.1). Recall from section 2.1.1 that first, in order to find estimates on the Green’s function, it is helpful to get some estimates on the gradient of the solution. We saw what with a domain that satisfies the exterior cone condition defined in definition 2.1.6, we get some crude H¨older-type estimates on the solution of the equation, and subsequently the Green’s function, as seem in lemma 2.1.8 and 2.1.9. After modifying the coefficients, when we have no assumptions on the boundary, we see in lemma 2.1.11 that the gradient of the solution can possibly blow up as you get closer to the boundary. Finally, we considered a domain with the exterior sphere condition; by first proving some estimates for the uniformly elliptic equation with Dini-continuous coefficients on an annulus (lemma 2.1.12), we are then able to get estimates for the Green’s function and it’s gradient up to the boundary in theorem 2.1.13. Since the equation we are dealing with is set in the half-space Rn+1 + , the boundary of our domain i.e. Rn, is incredibly nice. It satisfies both the interior and exterior cone and sphere conditions. However, the roadblock in our work is due to the degeneracy of the equation as 31 you get closer to the boundary. To be able to uniformly study points close to the boundary, we make use of cubes in Rn+1. Particularly, we look at the intersection of cubes centred at the origin with the half-space. Of course, the operator in (3.1.1) is not translation invariant in general, but for any fixed distance from the boundary, we do have translation invariance. This is why it is acceptable to look at cubes centred around 0 alone, as we can “slide” the cubes “horizontally” along the boundary Rn without changing the equation. Our goal is to be able to prove the estimates for the Green’s function as given in theorem 3.2.7. For this, we will need some specific estimates on the solution of (3.1.1) and its gradient. In the following results, we make some discoveries regarding how the gradient of the solution and the solution itself behaves close to the boundary. In lemma 3.2.2 and corollary 3.2.6, we see that when we restrict the values of a to (0, 1), we can get the estimates we want which lead us to the estimates for the Green’s function in the following theorem 3.2.7. Lemma 3.2.2. Suppose u solves the following equation in Q+ 4R = Q4R ∩ Rn+1 + given in (3.1.1)) Lu = 0 in Q+ 4R u = 0 on Q+ u = 1 on ∂Q+ 0 ≤u ≤ 1 on (Q+ + 2R ∩ ∂Rn+1 4R \ ∂Rn+1 4R \ Q+ + 2R) ∩ ∂Rn+1 + Then ∃K = K(n, λ, Λ, ω) such that for any X = (x, y) ∈ Q+ R, |∇u(X)| ≤ K y−a R1−a (with L as (3.2.1) (3.2.2) Here, Lu = − div(yaA∇u), with A uniformly elliptic, and a ∈ (0, 1). (We will sometimes denote ˜A = yaA.) 32 u = 1 Q+ 4R Lu = 0 t (x, y) Q+ R Q+ 2R 0 ≤ u ≤ 1 u = 0 0 ≤ u ≤ 1 Figure 3.1: Boundary conditions for lemma 3.2.2 Remark 3.2.3. We shed light on some things we may take for granted in this proof, which we will elaborate upon later. (a) We will first consider the case where R = 1, and see later that we can re-scale to get the result (see Remark 3.2.4). (b) We do not yet know if Schauder theory applies to (3.2.1) with the given coefficients, but we can fix this with a limiting argument (see Remark 3.2.5). For now, we shall assume that u ∈ C1 4 ) and proceed with the proof. loc(Q+ Proof. ya|∇u(X)| =: M < ∞ Let sup X∈Q+ 1 Let X0 = (x0, t) ∈ Q+ 1 be such that ya|∇u(X0)| ≥ 1 2 M , i.e. X0 is a point where this value is particularly large. Consider a ball of radius d around X0, where d ≤ 1/2 will be determined later. We c (Bd(X0)) such that η ≡ 1 on Bd/2(X0), |∇η| ≤ k1d−1 and choose a cut-off function η ∈ C∞ 33 # q X0 "! Q+ 1 u = 0 t = 1 # q X0 "! t = 1/2 t = 0 Figure 3.2: Different locations for Bd(X0) |∇2η| ≤ k2d−2. (k1, k2 are universal constants.) Note that we also have uη ≡ 0 on ∂Q+ 4 . Indeed, if for X0 = (x0, t) we have t > 1/2 then Bd(X0) is safely contained inside Q+ 4 and since η = 0 outside Bd(X0), uη = 0 on ∂Q+ 4 . On the other hand, if t < 1/2 then part of Bd(X0) may be outside Q+ 4 , but the only part of the boundary it intersects is where Q+ different places that Bd(X0) could be with X0 ∈ Q+ 2 ∩ ∂Rn+1 1 and given d < 1/2. + , where u = 0. The figure above shows the Now suppose F is the Green’s function in Q+ to the operator with the constant coefficients ˜aij L0 = − div(taaij Then for Z = (z, s) ∈ Q+ 0 ∇), and L0F = 0 in Q+ (cid:90) 4 we use η(·)F (·, Z) as a test function to get 4 for the equation in (3.2.1) corresponding 0 (X) = ˜aij(X0) = taaij(X0). i.e we have 4 , F ≡ 0 on ∂Q+ 4 . Hence, ηF is a valid test function. 0 = Q+ 4 ˜aij(X)uxi(X)[η(X)F (X, Z)]xj dX = ˜aijui(ηF )j (cid:90) Q+ 4 34 We will use the second notation for brevity. ˜aijui(ηF )j ˜aij 0 ui(ηF )j + (cid:90) Q+ 4 ˜aij 0 ui(ηjF + ηFj) + (˜aij − ˜aij 0 )ui(ηF )j (˜aij − ˜aij 0 )ui(ηF )j (cid:90) Q+ 4 (cid:90) (cid:90) (cid:90) Q+ 4 Q+ 4 Q+ 4 0 = = = (cid:90) Q+ 4 We will use integration by parts and note that uη ≡ 0 on ∂Q+ 4 , as is F . 0 = − 0 uηjFi− ˜aij 0 uηjiF − ˜aij 0 uηiFj− ˜aij ˜aij 0 uηFji + (˜aij−˜aij 0 )ui(ηF )j (cid:90) Q+ 4 (cid:90) Q+ 4 (cid:90) Q+ 4 (cid:90) Q+ 4 Since −aij 0 Fji(·, Z) = δZ , we have (cid:90) (cid:90) Q+ 4 ˜aij 0 uηiFj (3.2.3) u(Z)η(Z) = + ˜aij 0 uηjiF + ˜aij 0 uηjFi + Q+ 4 0 − ˜aij)ui(ηF )j (˜aij (cid:90) (cid:90) Q+ 4 Q+ 4 Now we differentiate (take gradient) (3.2.3) with respect to Z and set Z = X0. Note that η(X0) = 1 and ∇η(X0) = 0 since η ≡ 1 on Bd/2(X0). (cid:90) ∇u(X0) = taaij 0 uηj∇Z Fi(·, X0) + taaij 0 uηji∇Z F (·, X0) Q+ 4 taaij 0 uηi∇Z Fj(·, X0) (taaij 0 − yaaij)ui[ηj∇Z F (·, X0) + η∇Z Fj(·, X0)] (3.2.4) (cid:90) (cid:90) (cid:90) Q+ 4 Q+ 4 Q+ 4 + + = I1 + I2 + I3 + I4 35 Since aij 0 is just a constant coefficient and ta is a fixed constant multiplied to it, the estimates we have on the Green’s function F from [37] are as follows: |∇Z F (X, X0)| ≤ t−aC1|X − X0|1−N , and ∇Z Fxi(X, X0) ≤ t−aC2|X − X0|−N For I1 and I3, since ηj = ηi = 0 on Bd/2(X0) as well as outside Bd(X0), we can integrate 4 . Also using the fact that u ≤ 1, in this case we have the over the annulus Bd \ Bd/2 ∩ Q+ following estimates (cid:90) |I1| ≤ taΛk1d−1C2t−a|X − X0|−N dX (cid:90) d r−N rN−1dr = Λk1C2d−1 ln(2) = Cd−1 Bd\Bd/2 = Λk1C2d−1 d/2 Similarly, we also have |I3| ≤ Cd−1 For I2, we will just integrate over Bd by using the fact that η ≡ 0 outside Bd and hence ηij ≡ 0 too. (cid:90) Bd |I2| ≤ taΛk2d−2C1t−a|X − X0|1−N dX = Λk2C1d−2 r1−N rN−1dr = Cd−1 (cid:90) d 0 As for I4, we know that ui(X) ≤ M y−a by the assumption and recall we have the condition (cid:90) + ∩Bd Rn+1 y−a|˜aij(X) − ˜aij(Z)| |X − Z|N dX ≤ C 36 Therefore, (cid:90) (cid:90) ( ˜A(X0) − ˜A(X))ui[ηj∇Z F (·, X0) + η∇Z Fj(·, X0)] k1d−1C1t−a|X0 − X|1−N + t−aC2|X0 − X|)−N(cid:105) dX |I4| ≤ ≤ Bd Bd ( ˜A(X0) − ˜A(X))M y−a(cid:104) (cid:90) y−a ˜A(X0) − ˜A(X) |X0 − X|N Bd ≤ C(cid:48)M t−a dX (since d ≥ |X0 − X|) Putting all the above estimates together, we have |∇u(X0)| ≤ Cd−1 + C(cid:48)M t−a (cid:90) + C(cid:48)M ta d Bd =⇒ 1 2 M ≤ ta|∇u(X0)| ≤ C (cid:40) d : C(cid:48)(cid:90) y−a Bd ˜A(X0) − ˜A(X) |X0 − X|N dX ≤ 1/4 Let d0 = sup Then, (cid:90) Bd y−a dX ˜A(X0) − ˜A(X) |X0 − X|N y−a ˜A(X0) − ˜A(X) |X0 − X|N (cid:41) , choose d = min{d0, 1/2}. dX 1 2 M ≤ C M ≤ C ta d ta d 1 4 M + ≤ C sup{2, 1/d0} = K (since X0 ∈ Q+ 1 , ta < C) Note that in order to be controlled by this universal constant, it is important that a ≥ 0, hence our assumption that a ∈ (0, 1). Finally, this gives us ya|∇u(X)| ≤ K M = sup X∈Q1 =⇒ |∇u(X)| ≤ Ky−a. (for any X = (x, y) ∈ Q+ 1 ) 37 make a change of variables as follows: values as in (3.1.1). Now if instead we look at Q+ Remark 3.2.4. Scaling u solves the equation Lu(X) = − div(yaA(X)∇u(X)) = 0 in Q+ 4 , then X ∈ Q+ (cid:19) (cid:18) X 2 ∩ ∂Rn+1 4 \ ∂Rn+1 4 \ Q+ ˆu = 1 on ∂Q+ 0 ≤ˆu ≤ 1 on (Q+ Let u(RZ) = ˆu(Z) =⇒ u(X) = ˆu Then, we will have ˆu = 0 on Q+ 2 ) ∩ ∂Rn+1 + 4R with all the boundary 4R =⇒ X R = Z ∈ Q+ 4 . We R + + Since we have the same boundary conditions, we must ask - what equation does ˆu solve in R∇Z ˆu, and since X = (x, y) = RZ = Q+ 4 if u solves (3.1.1) in Q+ (Rz, Rs), we have ya = Rasa. If we also denote aij(X) = aij(RZ) = ˆaij(Z), and we know 4R? Note that ∇X u = 1 that ˆA will satisfy the same conditions as A, then ˆu solves the following equation in Q+ 4 R1−a (saˆaij(Z)∇Z ˆu(Z)) = 0 − 1 Repeat all the calculations in the proof to see that the factor 1 R1−a remains throughout, thus giving us the result in lemma 3.2.2. Remark 3.2.5. Limit argument Now we don’t really know whether u ∈ C1 coefficients ˜Am(X) = ˜Am(x, y) = max{ 1 4 ). However, we can consider ∀m ∈ N the loc(Q+ m , ya} · A(x, y), where A is uniformly elliptic. Note 38 that ˜Am → ˜A in the L∞-norm. Also, for every m, ˜Am(X) ≥ 1 m A(X) =⇒ ˜aij m(X)ξiξj ≥ 1 m λ|ξ|2 ∀X, ξ ∈ Rn+1 + and as such the equation Lmum = −(˜aij m(um)i)j = 0 with the same boundary values given in (3.1.1) is non-degenerate. If fact, each ˜Am is uniformly elliptic, and hence we know that each um ∈ C1(Q+ 4 ) by Schauder. Also, um ∈ H1(Q+ 4 ) as it solves the elliptic PDE. On the other hand, we also know that u which solves (3.1.1) is in the weighted Sobolev space H1(Q+ H1(Q+ 4 , w), where w(X) = ya is the weight. We could also say that um ∈ H1(Q+ 4 , w) ⊂ H1(Q+ 4 ). 4 , w) as If we proceed as in the proof above, then we see that we have |∇um| ≤ Ky−a, which is a uniform bound independent of m, and we also know that for all m, 0 ≤ um ≤ 1. Thus the sequence um is uniformly bounded in H1(Q+ a u(cid:63) ∈ H1(Q+ 4 , w) such that a subsequence umk converges weakly to u(cid:63) in H1(Q+ 4 , w), a Hilbert space, which means there is 4 , w) (cid:98) L2(Q+ 4 , w). On 4 , w). Thus, not only does umk → u(cid:63) 4 ) such that umk → v strongly in L2(Q+ 4 ). 4 , w). Hence the uniform the other hand, we also know that H1(Q+ 4 ), there is also a v ∈ L2(Q+ weakly in L2(Q+ Therefore, we must have u(cid:63) = v, and umk → u(cid:63) strongly in H1(Q+ bound on |∇um| also applies to its limit, |∇u(cid:63)|. Finally, what guarantees that this u(cid:63) which is the limit of um is the same function u that solves the equation (3.1.1)? Using ϕ as a test function 39 (cid:90) Q+ 4 (cid:90) (cid:90) = lim m→∞ Q+ 4 lim m→∞ m(um)iϕj − ˜aiju(cid:63) [˜aij i ϕj] m(um)iϕj − ˜aij(um)iϕj + ˜aij(um)iϕj − ˜aiju(cid:63) ˜aij i ϕj (cid:90) Q+ 4 ˜aij((um)i − u(cid:63) i )ϕj = lim m→∞ ≤ lim m→∞|| ˜Am − ˜A||L∞ m − ˜aij)(um)iϕj + lim (˜aij m→∞ |∇um||∇ϕ| (cid:90) Q+ 4 Q+ 4 (cid:90) yaΛ|∇um − ∇u(cid:63)||∇ϕ| Q+ 4 + lim m→∞ ≤ lim m→∞|| ˜Am − ˜A||L∞||um|| m→∞||um − u(cid:63)|| H1(Q+ + lim ||∇ϕ||L∞|Q+ 4 | 4 ,w) H1(Q+ CN Λ||∇ϕ||L∞ = 0 Thus, we have 0 = lim m→∞ (cid:90) Q+ 4 4 ,w) (cid:90) Q+ 4 ˜aij m(um)iϕj = aiju(cid:63) i ϕj Thus, u(cid:63) solves the equation (3.1.1) weakly. But u solves (3.1.1), and the solution is unique, which means u = u(cid:63). As for the argument with freezing the coefficients, we have ˜A0,m = ˜Am(X0) = max (cid:26) 1 (cid:27) , ta m · A(x0, t) which means we either have ˜aij 0,m(X) = taA(x0, t) or ˜aij 0,m(Z) = 1 m A(x0, t). In the proof above, we only considered the former case, however, it is easy to see that if we consider the latter, we would have the Laplacian multiplied by a constant which gives us the 40 same estimates for the Green’s function, up to a universal constant. Corollary 3.2.6. Let u solve the same equation above (3.1.1). Then ∃K = K(n, λ, Λ, ω) such that ∀X = (x, y) ∈ Q+ R, u(X) ≤ K y1−a R1−a Proof. For X ∈ Q+ R, let X(cid:63) ∈ ∂Rn+1 + be such that |X − X(cid:63)| = δ(X) = y. If [X, X(cid:63)] is the line joining X and X(cid:63), |u(X) − u(X(cid:63))| |X − X(cid:63)| ≤ sup Z∈[X,X(cid:63)] |∇u(Z)| ≤ K y−a R1−a (using the lemma above) Since we know that X(cid:63) ∈ Q+ 2R ∩ ∂Rn+1 + , we have u(X(cid:63)) = 0, which gives u(X) ≤ K y−a R1−a|X − X(cid:63)| = K y1−a R1−a . Theorem 3.2.7. Let G be the Green’s function for (3.1.1) with a ∈ (0, 1). Then ∃K1, K2 that depend only on n, µ, λ, ω such that ∀ X, Z ∈ RN + (i) G(X, Z) ≤ K1δ(X)1−a|X − Z|1−N (ii) G(X, Z) ≤ K2δ(X)1−aδ(Z)1−a|X − Z|−N +a Proof. (i) Let X be fixed and look at G as a function of Z alone, i.e. G(·) = G(X,·). Let R = |X − Z|/5. Consider the following 2 cases: (a) δ(Z) ≥ R. This means X, Z are safely inside the domain. So we can apply the 41 interior estimates to get G(X, Z) = K|X − Z|2−N−a R1−a R1−a (cid:16) X−Z (cid:17)1−a ≤ Kδ(Z)1−a|X − Z|2−N−a ≤ K1δ(Z)1−a|X − Z|1−N 5 (b) δ(Z) ≤ R. In this case, choose Z(cid:63) ∈ ∂RN + such that |Z − Z(cid:63)| = δ(Z). With the point Z(cid:63) as the center, consider Q+ 4R(Z(cid:63)). Let uR be as in lemma 3.1. + , we have G(P ) ≤ K|X − P|2−N−a since Now, for any point P ∈ ∂Q+ 4R \ ∂RN we have interior estimates. Also, since uR = 1 here, we have G(P ) ≤ K|X − P|2−N−auR(P ). We also know that |X − Z| ≤ |X − P| + |P − Z| ≤ |X − P| + 4R =⇒ |X − Z| ≤ 5|X − P| =⇒ |X − P|2−N−a ≤ |X − Z|2−N−a Thus, we have G(·) ≤ K|X − Z|2−N−auR(·) for any point on ∂Q+ 4R ∩ ∂RN the other hand, on the bottom boundary, i.e., on Q+ and uR ≥ 0, and thus we have G(·) ≤ K|X − Z|2−N−auR(·). 4R \ ∂RN + . On + , we have G ≡ 0, Thus, not only are G(·) and K|X − Z|2−N−auR(·) both solutions of Lu = 0 in 4R(Z(cid:63)), we also have G(·) ≤ K|X − Z|2−N−auR(·) on ∂Q+ Q+ 4R(Z(cid:63)). By compari- 42 son principle, we get G(P ) ≤ K|X − Z|2−N−auR(P ) ∀P ∈ Q+ 4R(Z(cid:63)) In particular, this statement is true for Z ∈ Q+ 4R(Z(cid:63)). From lemma 3.1, we get G(X, Z) ≤ K|X − Z|2−N−auR(Z) ≤ K|X − Z|2−N−a|uR(Z) − uR(Z(cid:63)| ≤ K|X − Z|2−N−a|∇(Z)|δ(Z) ≤ K|X − Z|2−N−a δ(Z)1−a R1−a = K1δ(Z)1−a|X − Z|1−N (ii) The proof of this is obtained from (i) the same way in which we prove (i) from the interior estimates. Remark 3.2.8. Recall that the Poisson kernel P is one that helps explicitly solve the equa- tion, which is this case is (3.1.1) i.e. we want P that satisfies for X = (x, y) ∈ Rn+1 (z, 0) ∈ Rn + , Z = u(X) = P (X, Z)f (Z)dZ (cid:90) Rn We will use the notation PX (Z) = P (X, Z) at times. This Poisson Kernel can be rec- ognized as the co-normal derivative of the Green’s function. Since from the above theorem, we have an estimate for the Green’s function as G(X, Z) ≤ Cy1−as1−a|X − Z|−N +a. Using 43 this estimate we will directly compute the co-normal derivative of the Green’s function from the definition 3.1.1. We will use the notation an+1,j = (cid:126)an+1, X = (x, y), Z = (z, s) PX (Z) = ya(cid:88) j an+1,jGxj (X, Z) G(X + y(cid:126)an+1, Z) − G(X, Z) = ya lim y→0 ya−1[Cy1−as1−a|X − Z|−N +a] y ≤ lim y→0 ≤ Cs1−a|X − Z|−N +a (3.2.5) 3.2.3. Relationship with harmonic measure As seen in the case of the uniformly elliptic equation, one of the other results about the Green’s function which is crucial to the proof of our main result is the relationship between the Green’s function and harmonic measure. We recall from the previous chapter (stated as lemma 2.2.10) that we know what this relationship is directly from [20, lemma 3]. However, since we do not need the assumption of an NTA domain with our domain being very smooth and regular, we will adapt this result to the half-space to establish the connection between the Green’s function and harmonic measure that we will call upon later. Recall that in section 3.2, we worked out that w(B(X, r)) = Crn+1+a = rN +a. Thus applying [20, lemma 3] to (3.1.1), we get Lemma 3.2.9. [20, lemma 3] Let Z = (z, s) ∈ Rn = ∂Rn+1 + , r > 0, and {ωx}x∈Ω be the L- harmonic measure where L is as given in (3.1.1). Also let ν(Z) be the inward normal vector at Z (i.e. ν(Z) = (0,··· , 0, 1)), and let Z(cid:48) r = Z + rν(Z). Then, if X ∈ Rn+1 + \ B(Z, 4r), we 44 have G(X, Z(cid:48) r) (cid:39) ωX (∂Ω ∩ B(Z, r)) r2 w(B(x, r)) = r2−N−aωX (∂Ω ∩ B(Z, r)) In other words, we can also say that ∃C1, C2 > 0 such that C1r2−N−aωX (∂Ω ∩ B(Z, r)) ≤ G(X, Z + rν(Z)) ≤ C2r2−N−aωX (∂Ω ∩ B(Z, r)) which is exactly the analogous result to [3, Lemma 2.2] that we need (stated in this work as 2.1.5). 3.3. D-to-N maps of the weighted equation (main result and proof ) The main goal of this chapter and the previous one had been to modify the theorem 2.1.15 to include the degenerate elliptic equations with weights. The work of [14] already confirms that the D-to-N map I associated with the equation (3.1.1) can be represented as the integro- differential operator I(φ, x) = (cid:104)b(x),∇φ(x)(cid:105) − (cid:90) Rn φ(x + h) − φ(x) − 1B1(x)(h)(cid:104)∇φ(x), h − x(cid:105) µ(x, dh). (3.3.1) We aim to prove the properties of the measure µ in the following main result: Theorem 3.3.1. Suppose L is the divergence operator as in (3.1.1) with Ω = Rn+1 and I is as defined via (2.1.2) and (2.1.3), then it is known that there exists a vector field, b, and a family of measures parametrised by x, µ(x, dh) such that ∀φ ∈ C1,α(Rn), we have the + representation in (3.3.1). The measure µ in this expression satisfies 45 (a) µ is absolutely continuous with respect to the Lebesgue measure, i.e. for all X = (x, 0) ∈ Rn, µ(x,·) has a density, µ(x, dh) = K(x, h)(dh). (We will use the notation (x, 0) = x since we are talking about points on the boundary, Rn.) (b) There exist universal constants C1, C2 > 0 so that ∀x, h ∈ Rn, x (cid:54)= h K(x, h) ≤ C2|x − h|−n+a−1. Proof. First, we will want to know if µ is comparable to the Lebesgue measure on Rn. To do this, fix X = (x, 0) = x ∈ Rn, show that µ(x,·) is absolutely continuous with respect to dh on Rn \ {x}. We will do this by showing absolute continuity on the set Rn \ Br(x) for any arbitrary r > 0, and then we can exhaust Rn \ {x} by a union of such sets. Thus, fix r > 0 and any set E ⊂ Rn \ Br(x) with |E| = 0. Our goal is to show that µ(E) = 0. We fix δ > 0, and find a countable cover {B(xj, rj)}∞ (cid:80)∞ in Rn. Now let φ ∈ C2(Rn) be such that 0 ≤ φ ≤ 1∪∞ j < δ, and we will write Bj = B(xj, rj) for brevity. Note that all of these balls are j=1 of E by open geodesic balls such that j=1 rn . j=1Bj If δ is sufficiently small, then we will have φ ≡ 0 on Br/2(x). This is because since E ⊂ Rn \ Br(x), the worst case scenario is that the balls cover all of Rn \ Br(x), and maybe some part of Br(x). But if δ is small, then the radii of the balls are really small, and so they won’t cross over too much, and we can have Bj in such a way that none of them intersect = 0 on Br/2(x), we have φ ≡ 0 there. Thus in (3.3.1), Br/2(x). Since 0 ≤ φ ≤ 1∪∞ φ(X) = 0, ∇φ(x) = 0. Also, φ(x) = 0 when x ∈ Br/2(x) and so we have j=1Bj 46 Rn\Br/2(x) φ(z)µ(z, dz) (3.3.2) (cid:90) (cid:90) I(φ, x) = = φ(z)µ(z, dz) Rn (cid:90) Let {ωX} X ∈ Rn+1 + , X∈Rn+1 + be the L-harmonic measure for L given by (3.1.1), and recall for any Uφ(X) = φ(z)ωX (dz) ∂Rn We are now going to employ lemma 3.2.9, and we choose any t > 0 as for the X = x ∈ Rn, we have X(cid:48) = X + tν(X) ∈ Rn+1 + always. Then we have ωX+tν(X)(∂Ω ∩ Brj (xj)) ≤ CrN +a−2 j G(X + tν(X), Xj + rjν(Xj)) (3.3.3) Where G is the Green’s function and C only depends on the ellipticity of the equation. This works because xj are far away enough from x, and so Xj + rjν(Xj) will be far away enough from X + rν(X). Thus we have the estimate 47 Uφ(X + tν(X)) = ≤ Rn (cid:90) ∞(cid:88) φ(z)ωX+tν(X)(dz) ωX+tν(X)(Rn ∩ Brj (zj)) ∞(cid:88) ∞(cid:88) rN−2+a j=1 j=1 j rN−2+a j ≤ C ≤ C j=1 ≤ Crt1−aδ (φ ≤ 1∪Bj ) (from (3.3.3)) G(X + tν(X), Xj + rjν(Xj)) t1−a · r1−a j |X + tν(X) − (Xj + rjν(Xj))|N−2+a (from theorem 3.2.7 (ii)) (the distance between Xj + rjν(Xj) and X + rν(X) is some constant that only depends on r) From the definition, we have I is the weighted normal derivative, i.e. I(φ, X) = ∂νUφ I(φ, X) = ta lim t→0 Uφ(X + tν(X)) − Uφ(X) t Crt1−aδ − 0 t1−a = lim t→0 = Crδ But Bj covers E, so we can take a sequence of functions φk ∈ C2(Rn) such that 1E ≤ φk ≤ 1∪Bj and that they decrease pointwise to 1E, so we have (cid:90) (cid:90) (cid:90) Rn Rn I(φk, X) = = ≥ Rn\Br/2(X) φk(Y )µ(X, dY ) φk(Y )µ(X, dY ) (φk ≡ 0 on Br/2(X)) 1Eµ(X, dY ) = µ(X, E) =⇒ µ(x, E) ≤ I(φk, X) ≤ Crδ. (from earlier) 48 This shows µ is absolutely continuous with respect to the Lebesgue measure, this estab- lishing part (a) of the theorem. Now, in order to prove part (b), we use the Poisson kernel. From 3.2.8, we expect (cid:90) Uφ(X) = P (X, Z)φ(Z)dZ Rn and we also have an analogous estimate to the Poisson kernel similar to [4, Section 2.4] from (3.2.5), which is PX (Z) ≤ C s1−a |X − Z|n+1−a In the following calculation, we denote X + tν(X) = X(cid:48) ∈ Rn+1 + for X = (x, 0) ∈ Rn, and thus for X(cid:48) = (x, t) and ξ = (ξ, 0) ∈ Rn we have PX(cid:48)(ξ) ≤ Ct1−a|X(cid:48) − ξ|−N +a = Ct1−a(cid:16)|x − ξ|2 + t2(cid:17)−n−1+a 2 Thus I(φ, x) = lim t→0 = lim t→0 ta−1(cid:0)Uφ(X + tν(X)) − Uφ(X)(cid:1) (cid:18)(cid:90) (cid:19) (cid:90) PX(cid:48)(Z)φ(Z)dZ ta−1 t1−a|X(cid:48) − Z|−n−1+adZ ta−1 Rn Rn |X − Z|−n−1+adZ (cid:90) ≤ lim t→0 ≤ C Rn (Since Uφ(X) = φ(X) = 0 and using the definition of the Poisson kernel) (Since X(cid:48) = X + ν(X), |X − Z| and |X(cid:48) − Z| are comparable) Also, since X = (x, 0), Z = (z, 0) ∈ Rn, we can say |X − Z| = |x − z|. Now from (3.3.2), (cid:90) Rn I(φ, x) = (cid:90) Rn φ(z)µ(x, dz) ≤ C 49 |x − z|−n−1+aφ(z)dz The above statement is true for every φ on Rn such that φ ≡ 0 on Br/2(x). Thus, for E ⊂ Rn \ Br/2(x), we have (cid:90) (cid:90) (cid:90) E µ(E) ≤ 1Eµ(x, z) = Rn Rn 1E|x − z|−n−1+adz = |x − z|−n−1+adz Since this is true for every set E ⊂ Rn \ Br/2(x), and we can exhaust Rn by the union over r of such sets Rn \ Br/2(x), we have proved part (b) of the theorem. Remark 3.3.2. The above “proof” is comprised of two different proofs of the same result. We note here that using the estimates on the Poisson Kernel in the second half of the proof would be enough to shove both parts (a) and (b) of the theorem. However, we also include an alternate proof of why µ (cid:28) σ in the first half of the above proof. Remark 3.3.3. In the proof of the above theorem, we have used in the definition of the weighted normal derivative the inward normal vector which is the natural geometric candi- date as explained in 3.1.2. However, we reiterate here that this theorem and it’s proof also works identically using the co-normal derivative defined with the help of the equation in def- inition 3.1.1. Since the operators we use fulfil the global comparison property, the canonical derivative to investigate is always the one that comes from the natural geometric normal. 50 Chapter 4 Background on free boundary problems 4.1. Introduction and background The second project in this dissertation studies some two-phase free boundary problems that are similar to Hele-Shaw equations. To do this, we will be generalising some recent work that used methods from integro-differential parabolic equations. We shall now briefly describe what these equations are. The simplest type of such an equation is as follows. Consider a function U : Rn+1 × [0, T ] → R, which is harmonic (i.e ∆U = 0) on the sets {U > 0},{U < 0} ⊂ Rn+1 × [0, T ]. These sets have boundary ∂{U > 0} which, for a fixed time t, is a hypersurface in Rn+1, and may not be regular. Thus these equations have an additional layer of complexity as their domain and boundary are not fixed; but rather, the domain of the function is an unknown in the equation and changes with time. There is an additional constraint on u which arises naturally from either physical principles or energy minimization, and it concerns the balance of the normal derivatives ∂+ n U + and ∂− n U− in the inward direction along the free boundary, which is given by the boundary velocity. More formally, for our two-phase problem, we seek a function u that solves 51 ∆U = 0 in {U > 0} ∪ {U < 0} (4.1.1) under the velocity condition ∂{U > 0} moves with normal velocity V = |∇U +| − |∇U−|. (4.1.2) This model includes a special case of the stationary two phase problem under the condition that U does not depend upon time and the balance condition becomes |∇U +| − |∇U−| = 1. (4.1.3) It also includes the one phase version that usually carries the name Hele-Shaw, and that corresponds to ignoring ∂{U < 0} and setting the velocity condition to be V = |∇U +|. I work with these canonical examples, and the goal is to find a solution u that satisfies (4.1.1) above with the condition in (4.1.2). The work presented in this dissertation applies to more general equations in which the stationary equation for U us given by F1(D2U ) = 0 F2(D2U ) = 0 V = G(∂+ n U, ∂− n U ) in {U (·, t) > 0} in {U (·, t) < 0} in ∂{U > 0} (4.1.4) where F1, F2 are uniformly elliptic rotationally invariant nonlinear operators (see the section 4.2 for precise definitions). The example stated before in equations (4.1.1)-(4.1.3) is a special case of the same, where both F1(D2U ) = F2(D2U ) = ∆U . Numerous works have studied these types of equations and their solutions in the case 52 that U : Rn+1 → R and some of them include [9, 11, 18, 27]. Viscosity solutions give a way of handling these equations when other techniques fail, and there is substantial literature on the variants of these equations in [1, 5, 27, 28]. Recently, the work of Chang-Lara, Guillen, and Schwab [9] gives a new technique which reduces the variants in the family of problems with the condition (4.1.2) to an equivalent problem of the integro-differential type by considering the hypersurface ∂{U (·, t) > 0} as the graph of a function f : Rd × [0, T ] → R. Then, for a sufficiently smooth f , the equation (4.1.1) with the condition in (4.1.2) is equivalent to ∂tf = H(f, x) on Rn × [0, T ] f (·, 0) = f0 on Rn (4.1.5) where H : C1,γ → Cγ. In our example of (4.1.2), we will have (cid:113) f | − |∇U− f |) H = (|∇U + 1 + |∇f|2. This particular H is a form of the Hamilton-Jacobi equation, however, before the methods of viscosity solutions are applicable to our case, we will need more information about the structure of |∇U + f | as operators on f . f |,|∇U− In chapters 4 and 5 of this dissertation, we will be looking at the case where the free boundary set ∂{U (·, t) > 0} is given again by the graph of a function, but that the function is now a function on Sn × [0, T ] instead of f : Rn × [0, T ], which was the assumption for the graph in the previous works. This graph assumption is a technical restriction for the methods herein. In this regard, the point of the work in this part is to extend the work of [9] to the case of above by expanding this new technique, with the free boundary assumed 53 to be the graph of a function over the sphere Sn instead of Rn. Thus we have that U, f are functions defined over Sn × [0, T ], and when time t is fixed, then f (·, t) : Sn → [δ,∞). As such, any previous assumption about translation invariance of the domain or operators now becomes a matter of rotational invariance. (4.1.5) is a parabolic equation, but now we shall describe a little more explicitly the examples of parabolic equations that play a role for the free boundary analysis. In the earlier chapters regarding degenerate elliptic equations, we considered the canonical operator to be the 1/2-Laplacian, which can also be considered the canonical integro-differential operator in this context. Recall that −(−∆)1/2f (x) = cn (cid:16) (cid:90) Rn f (x + h) − f (x) − 1B1 (h)|h|−n−1(cid:17) dh. The canonical parabolic equation in our context is the 1/2-heat equation given by ∂tf + (−∆)1/2f = g Just like in the earlier chapters, the Dirichlet to Neumann operator in Rn+1 is given by the linear operator −(−∆)1/2. Define an operator that is suitable for (4.1.2) using the normal derivative i.e. f (cid:55)→ I(f ) = ∂νUf . Here Uf is the unique function that solves a one-phase problem, namely ∆Uf = 0 Uf = 0 Uf = 0 in {0 < y < f (x)} on {y = f (x)} on {y = 0} (4.1.6) The condition Uf = 1 on {y = 0} can simply be interpreted as there being some background 54 pressure in the system. (As earlier, we denote X ∈ Rn+1 as X = (x, y) where x ∈ Rn, y ∈ R. The main technique used to show the equivalence of solving (4.1.1) under condition (4.1.2) with solving (4.1.5) was given in [9]. It turns out that the set up above is in some sense a non-linear version of the D-to-N mapping, and the key idea is to use that the operator H has a special structure of the min-max form H(f, x) = min i {aij + cijf (x) + bij ·∇f (x) + p.v. max j (cid:90) Rn f (x + h)− f (x)µij(x, dh)} (4.1.7) for an appropriate family of aij, bij, cij and µij. Equations that admit a similar min-max form are frequently amenable to a large collection of tools from the viscosity solutions context. Hence, in the next chapter, we will utilize the solutions of ∂tf − min i {aij + cijf (x) + bij · ∇f (x) + p.v. max j f (x + h) − f (x)µij(x, dh)} = g (cid:90) Rn to deduce existence, uniqueness, and some low regularity results for the solutions of the free boundary problems described above. The assumption that ∂{U > 0} = graph f is not ideal; we see that it does not appear as a requirement in [1, 5, 11, 12, 27, 28]. However, it gives a natural reduction to explore new techniques, especially with an assumption that {U > 0} is a star-shaped domain with respect to X = 0. Thus it makes sense to expand the ideas in [9] to the case of the functions u, f defined on the sphere Sn instead of Rn+1. The advantage of using the approach of (4.1.5) is that it now allows us to generalise these techniques to many more variations in the free boundary problems described above. One important variant of the Hele-Shaw type problem is to have ∆U = ρ(U ) with monotone ρ 55 in {U > 0} and the condition V = k(x, t)|∇U +| in (4.1.3); this was used as a semilinear model for tumor growth [30]. Another equation to study is when ∆U = g(x) in (4.1.1) and the Batchelor-Prandtl equation is a special case of this with the condition (4.1.2) [9, 16, 17]. 4.2. Definitions and examples Since the main goal of this project is to extend results owing to the connections between certain parabolic equations and free boundary problems using integro-differential operators, we use this section to provide definitions of the different equations and operators, as well as detailed descriptions of the existing results in the field. There is a tremendous amount of work on integro-differential parabolic equations [6, 7, 8, 10, 32, 33, 34, 35]. For this project, the main works that contain most of the techniques which we generalize to the context of rotationally invariant equations on Sn is that of Silvestre [34] and [9]. 4.2.1. General Definitions In this section, we provide a general list of definitions of various concepts, operators, and their properties which are present in this chapter and the next one. First, we start with the definitions of some of the properties of the integro-differential operators which appear in the parabolic equations we study. Definition 4.2.1. As a function, Rx : Sn → Sn is a rotation such that Rx(0) = x. The rotation operator Rx, which acts on functions on Sn (say U : Sn → R), is defined for a fixed x as RxU (·) := U ◦ Rx(·) 56 Definition 4.2.2. We say that an operator J : C1,γ(Sn) → C0(Sn) is rotationally invariant if ∀f ∈ C1,γ(Sn) we have J(Rxf, z) = J(f, Rxz). Definition 4.2.3. (GCP). We say that J : C1,γ(Sn) → C0(Sn) has the global compari- son property (GCP) if ∀ f, g ∈ C1,γ(Sn) such that f (x) ≤ g(x) ∀x ∈ Sn, and for some x0, f (x0) = g(x0), the operator J satisfies J(f, x0) ≤ J(g, x0). In other words, J preserves the ordering of functions on Sn wherever their graphs touch. We also say that J has GCP at x0 if the above property only holds for one fixed x0, instead of all possible x0. Recall that we are generally interested in equations that involve uniformly elliptic opera- tors, and hence we shall define what a uniformly elliptic operator means in this context, for which we first need to define Pucci operators. Definition 4.2.4. (Extremal Operators) For a function u that is second-differentiable at X, the second order (λ, Λ) Pucci operators are defined as M+, M− via M−(D2U, X) = min λId≤B≤ΛId tr(BD2U (X)) = Λ M+(D2U, X) = max λId≤B≤ΛId tr(BD2U (X)) = λ ei + λ ei + Λ ei ei (cid:88) (cid:88) ei≤0 ei≤0 (cid:88) (cid:88) ei>0 ei>0 Where D2U is the Hessian matrix of all the second partial derivatives of u, {ei}i=1,2,···d+1 are the eigenvalues of D2U (X). Definition 4.2.5. (Uniformly Elliptic) We define uniform ellipticity for F that is either linear of non-linear in the following ways: • When F is linear, i.e. when F (D2U,∇U ) = tr(AD2U ) + B · ∇U , then we need ||B||L∞ ≤ Λ and λId ≤ A ≤ ΛId to say F is uniformly elliptic, and 57 • when F is non-linear, if for all U, V ∈ C2, M−(D2U − D2V ) − Λ|∇U − ∇V | ≤ F (D2U,∇V ) − F (D2V,∇V ) ≤ M+(D2U − D2V ) + Λ|∇U − ∇V | The integro-differential form with which we can represent our operators consists of a linear operator on C1,γ(Sn), but we can later restrict this operator to some special subsets of C1,γ(Sn) to suit the parabolic equations we work with. We shall now list some definitions for these special subsets. Definition 4.2.6. (C1,γ(Ω)) The γ−th H¨older semi-norm of f : Ω → R (where Ω ⊂ Sn is given by |f (x) − f (y)| |x − y|γ [f ]γ := sup x,y∈Ω x(cid:54)=y The C1,γ-norm of f is given by ||f||C1,γ (Ω) := ||f||L∞ + ||∇f||L∞ + [∇f ]Cγ The set C1,γ(Ω) consists of all functions f : Ω → R whose C1,γ-norm is finite. Equivalently, we can also say that C1,γ(Ω) = f ∈ L∞(Ω) : sup z∈Ω ||f − P||L∞(Br(z)) < ∞  . r−1−γ sup r>0 inf P (x)=c+p·x c∈R, p∈Rn 58 Definition 4.2.7. For γ ∈ (0, 1), δ > 0, m > 0, the convex set K(γ, δ, m) is defined as K(γ, δ, m) := {f ∈ C1,γ(Ω) : f (x) ≥ δ ∀x ∈ Ω and ||f||C1,γ (Ω) ≤ m} Further, K(γ, δ) = K(δ) = (cid:91) (cid:91) m>0 (cid:91) γ∈(0,1) m>0 K(γ, δ, m) and K(γ, δ, m). Definition 4.2.8. (Upper Gradient). Suppose K ∈ C1,γ(Sn) is an open convex subset and φ : K → R is Lipschitz. Then the upper gradient of φ at f ∈ K in the direction of g ∈ C1,γ(Sn) is defined as φ0(f ; g) := lim sup t(cid:38)0 t φ(f + tg) − φ(f ) This can be viewed as a function φ0 : K × C1,γ(Sn) → R Definition 4.2.9. (Subdifferential). Let φ be as in definition 4.2.3. The Clarke differential (or the generalized gradient) of φ at f ∈ K is a subset of(cid:0)C1,γ(Sn)(cid:1)(cid:63) given by ∂φ(f ) := C1,γ(Sn) (cid:96) ∈(cid:16) (cid:110) (cid:91) f∈K (cid:17)(cid:63) |∀ψ ∈ C1,γ(Sn), φ0 ≥ (cid:104)(cid:96), ψ(cid:105)(cid:111)  and [∂φ]K := hull ∂φ(f ) (the convex hull) Definition 4.2.10. (C1,γ-semi-concave) Given γ ∈ (0, 1], m > 0, a Lipschitz function f : Sn → R is said to be C1,γ-semi-concave with constant m if there is a real-valued function, r : Sn → R, such that f (x) = inf y∈Sn {r(y) + m|x − y|1+γ} f is said to be C1,γ-semi-convex with constant m if (−f ) is C1,γ-semi-concave with constant 59 m. Remark 4.2.11. The original definition above is for functions from Rn → R, but we can see that we can do an easy replacement of Rn by Sn and the definition still holds. Indeed, since Sn is a manifold, we have charts from all open subsets U ∈ Sn to B1 ∈ Rn. Let us call this map ΦU : U (⊂ Sn) → B1 ⊂ Rn for an arbitrary open set. We will use x, y ∈ Rn and p, q ∈ Sn and let ΦU (p) = x, ΦU (q) = y. Then we can say that f : Sn → R is C1,γ-semi-concave if f ◦ Φ−1 U : Rn → R is C1,γ-semi-concave by the usual definition. Thus, ∃r(cid:48) : Rn → R, a real-valued function such that f (p) = f ◦ Φ−1 U (x) = inf y∈Rn = inf q∈Sn = inf q∈Sn {r(cid:48)(y) + m|x − y|1+γ} (cid:110) r(cid:48)(ΦU (q)) + m|ΦU (p) − ΦU (q)|1+γ(cid:111) (cid:110) (r(cid:48) ◦ ΦU )(q) + m · C|p − q|1+γ(cid:111) Thus, we now have a real-valued function r = r(cid:48) ◦ ΦU : Sn → R that satisfies the condition in definition 4.2.5, with the constant m · C > 0, where C depends on the Lipschitz constant of ΦU . Definition 4.2.12. (Pointwise or punctually C1,γ). Let γ ∈ (0, 1] and m > 0 be fixed. We say that f is pointwise m − C1,γ at x0 (denoted f ∈ m − C1,γ(x0)) if ∇f (x0) exists, |f (x0)| ≤ m, |∇f (x0)| ≤ m, and ∃r > 0 such that ∀x ∈ Br(x0),|f (x) − f (x0) − ∇f (x0) · (x − x0)| ≤ m|x − x0|1+γ 60 4.2.2. Some examples of free boundary problems To support the motivation behind studying the two-phase free boundary problem on the sphere, we shall give some examples of free boundary problems, their D-to-N maps, and the parabolic equations associated with them. (a) One phase Hele-Shaw on the half-space As before, we will use the notation X = (x, y) for any X in the half space Rn+1 + , with x ∈ Rn, y > 0. We consider functions f : Rn → R which are continuous, non-negative, bounded and bounded away from zero. To any such function, we can associate a domain Df = {(x, y) ∈ Rn+1 + , 0 < y < f (x)} We can also define the hypersurface Γf , which is the graph of f as Γf = {(x, y) ∈ Rn+1 + , y = f (x)} Now let Uf : Df → R be the unique bounded solution to the Dirichlet problem ∆Uf = 0 Uf = 1 Uf = 0 in Df on {y = 0}, i.e. Rn on Γf (4.2.1) Now, if ν is the inward facing unit normal to Γf (i.e. it points towards Df ), then we 61 V = ∂+ ν U ∆U = 0 Df U = 1 Γf y = 0 Figure 4.1: One phase Hele-Shaw on the half-space define the normal derivative ∂+ ν U as follows for X0 ∈ Γf , ∂+ ν U = lim t→0+ U (X0 + tν(X0)) − U (X0) t , and I(f, x) := ∂+ ν Uf (x, f (x)). (4.2.2) Now, we can put a time evolution into the setup to make a free boundary problem, i.e. consider U : Rn+1 × [0, T ] → R, which solves ∆U = 0 U = 1 V = ∂+ ν U in {u > 0} on {y = 0}, i.e. Rn on ∂{u > 0} (4.2.3) (4.2.3) is the one-phase Hele-shaw problem on the upper half space. For a sufficiently smooth f , U = Uf is a solution of (4.2.2) if and only if f is a solution to the parabolic equation ∂tf := I(f, x) (cid:113) 1 + |∇f|2, on Rn × [0, T ] (4.2.4) 62 U− = −1 ∆U− = 0 ∆U + = 0 U + = 1 D− f D+ f y = L Γf y = 0 V = |∂+ ν U| − |∂− ν U| Figure 4.2: Two phase Hele-Shaw on an infinite strip where I is as given in (4.2.2), and it is is translation invariant, non-linear, non-local, and it enjoys GCP. (b) Two phase Hele-Shaw on an infinite strip + . Now for every f : Rn → R such that f is We first fix an upper boundary L > 0 in Rn+1 continuous, non-negative, and bounded away from 0 and L, i.e. 0 < δ ≤ f ≤ L−δ < L. We then define two domains for {U > 0} and {U < 0} f = {(x, y) ∈ Rn+1 D+ D− f = {(x, y) ∈ Rn+1 and Γf = {(x, y) ∈ Rn+1 + , 0 < y < f (x)}, + , f (x) < y < L}, + , y = f (x)} Now we suppose U + : D+ f → R and U− : D− f → R are respectively the unique bounded 63 solution to the Dirichlet problem ∆U± = 0 U± = ±1 U± = 0 in D± on {y = 0} and {y = L} respectively f (4.2.5) on Γf Then a Lipschitz function G : (0,∞)2 → R which satisfies the monotonicity condition λ0 ≤ ∂ ∂a G(a, b), G(a, b) ≤ Λ0 ∂ ∂b will give the normal velocity of the flow depending on ∂± 0}. ∂+ ν U is as given above in (4.2.2), and ∂− ν U is given by ν U along the boundary ∂{U > ∂− ν U = − lim t→0+ U (X0 − tν(X0)) − U (X0) t (4.2.6) Now, we can talk about the two-phase free boundary problem along an infinite strip as follows: ∆U = 0 U = 1 U = −1 in {U (cid:54)= 0} on {y = 0}, i.e. Rn on {y = L} on ∂{U > 0} (4.2.7) ν U, ∂− ν U ) ν U (x, f (x)), ∂− If we define H(f, x) as H(f, x) = G(∂+ V = G(∂+ ν U (x, f (x)), then similarly as in the one-phase problem, a sufficiently smooth f solving the parabolic equation (cid:113) ∂tf := H(f, x) 1 + |∇f|2, on Rn × [0, T ] (4.2.8) 64 is equivalent to U = Uf solving the free boundary problem (4.2.7). (c) Prandtl-Batchelor as a non-linear integro-differential equation on the sphere. Finally, we come to an example where the free boundary is not given by the graph of a function over Rn but by the graph of a function over a sphere. The Prandtl-Batchelor is a two dimensional model in fluid mechanics that models a vortex patch occupying a convex region, the patch being surroundedby a steady flow. The interface of the vortex patch is what plays the role of the free boundary. We denote the stream function of the flow by u, and it solves the following equation ∆U = 0 ν U|2 − |∂− |∂+ ∆U = 1 ν U|2 = 1 in {U > 0} in {U < 0} on ∂{U > 0} (4.2.9) Now consider any function f : Sn → R, which is non-negative, bounded, and bounded away from zero, and satisfies 0 < f (x) ≤ f (x) for a given positive and continuous function f . Thus we can define the sets D− f := {x ∈ Rn + 1 : x = re, e ∈ Sn, 0 ≤ r < f (x)} f := D− D+ \ D− f f 65 Suppose U± f : D± f → R give the unique solution to the Dirichlet problem ∆U− = 1 U− = 0 ∆U + = 0 U + = 0 U + = 1 in D− on ∂D− f f (4.2.10) in D+ f on ∂D− on ∂D− f f and we define an operator P : C2(Sn) → C0(Sn) as P(f, x) := |∂+ ν U (x, f (x))|2 − |∂− ν U (x, f (x))|2. P admits GCP, and if there is a sufficiently smooth f : Sn → R solving P(f, x) = 1 in x ∈ Sn, then the radial graph of f is the boundary of a vortex patch in a Batchelor-Prandtl flow, and U± f gives the positive and negative phases of the stream function. Since the Laplacian operator and the boundary conditions on the domain are both rotationally invariant, the operator P will also remain invariant under the action of rotations on C2(Sn). 4.3. Results to be modified Here we will state some of the main results from [9], which are about the equation (4.1.4). There is a notion of weak solutions for both the free boundary problems and the parabolic equations and these are called viscosity solutions. We will define these notions in detail later, in sections 5.4.1 and 5.3.1 respectively. Theorem 4.3.1. [9, Theorem 1.1] If F1, F2 are uniformly elliptic and rotationally invariant in the Hessian variable, G is Lipschitz and monotone, and f : Rd × [0, T ] → [δ,∞), Uf : 66 Rd+1 × [0, T ] → R are such that ∀t ∈ [0, T ], Γ(t) = ∂{Uf (·, t) > 0} = graph f (·, t), then, (a) Uf is a viscosity solution of the free boundary evolution, (4.1.4) if and only if f is a viscosity solution of the fractional parabolic equation (4.1.5). (b) If additionally, graph f (·, t) = ∂{Uf (·, t)} enjoys a modulus of continuity, i.e. |f (x, 0) − f (y, 0)| ≤ ω(|x − y|), then ∀t ∈ [0, T ] this modulus is preserved for f (·, t) and hence for ∂{Uf (·, t). (c) If graph f (·, t) = ∂{Uf (·, t)} enjoys a modulus of continuity, i.e. |f (x, 0) − f (y, 0)| ≤ ω(|x − y|), then there exists a unique viscosity solution to both (4.1.4) and (4.1.5). We refer the reader to [9] for the proof of the above theorem. The main tool used in the analysis of the above results is what could be called a non-linear version of the D-to-N operator I, defined via the inward normal derivative of Uf which is given by I(f, x) := ∂νUf (x, f (x)) (4.3.1) for the one-phase problem. For the two phase problem, we define the D-to-N operators as I+(f, x) := ∂+ ν Uf (x, f (x)), I−(f, x) := ∂− ν Uf (x, f (x)) One of the results that leads to the proof of theorem 4.3.1 is the property of I which we list in the following result. 67 Theorem 4.3.2. [9, Theorem 1.4] If F is uniformly elliptic and rotationally invariant in the Hessian variable, I is as in (4.3.1), and γ ∈ (0, 1) is fixed, then ∃γ(cid:48) with 0 < γ(cid:48) < γ so that I : K(γ, δ, m) → Cγ(cid:48) (Rd) and I is locally Lipschitz. Also, I enjoys the following representation for aij, bij, cuj, µij(dh) which are all independent of x: (cid:110) aij + cijf (x) + bij · ∇f (x) f (x + h) − f (x) − 1B1 (h)∇f (x) · h (cid:17) (cid:27) µij(dh) I(f, x) = min (cid:90) i max j (cid:16) Rd + The bounds of aij, bij, cuj, µij(dh) all depend on the bounded set {f : f ≥ δ,||f|| C1,γ (Rd) ≤ m}. Furthermore, ∃C > 0 such that |aij|,|bij| ≤ C, −C ≤ cij ≤ 0, and (cid:90) min Rd (cid:110)|h|1+γ, 1 (cid:111) (cid:90) Rd µij(dh) ≤ C and µij(dh) ≤ C The proof of the above result can also be found in [9]. Inorder to prove these properties about I and subsequently H as given in (4.1.5), one can study a general class of operators that satisfies certain conditions, which we outline in the assumption below. Assumption 4.3.3. (i) 0 < γ < 1 is fixed. 68 (cid:32)(cid:83) (cid:83) δ>0 m>δ (cid:33) K(γ, δ, m) (ii) J : → C0(Rd). (iii) For each δ and m, J is a Lipschitz mapping on the sets K(γ, δ, m), whose Lipschitz constant, C(δ, m) increases as δ decreases or m increases. (iv) J satisfies GCP. (v) J is translation invariant. (vi) If f ∈ K(γ, δ, m) and c > 0 is a constant, then ∀x ∈ Sn, J(f + c, x) ≤ J(f, x). (vii) J enjoys the operator splitting property: ∃C = C(γ, δ, m), and a modulus of continuity ω with ω(R) → 0 as R → ∞ such that for all f, g ∈ K(γ, δ, m), for R > 1, ||J(f,·) − J(g,·)||L∞(BR) ≤ C (cid:16)||f − g||C1,γ (B2R) + ω(R)||f − g|| L∞(Rd) (cid:17) + ω(R) This operator J will enjoy a special min-max structure, and we will state the relevant result below. However, first, the proof for deriving this structure relies on the following the results from [25]. Theorem 4.3.4. [25, Theorem 1.10] If I : C2 b (Rd) → C0 b (Rd) (bounded functions in C2, C0, respectively) satisfies the conditions (iii), (iv), and (v) of Assumption 4.3.3, then there exists a family {fab, Lab}a,b∈K(I), that depends only on I, where for all a, b, fa,b are constants, and 69 Lab are linear translation invariant operators mapping I : C2 b (Rd) → C0 b (Rd) of the form L(u, x) =Cu(x) + B · ∇u(x) + tr(AD2u(x)) (cid:90) u(x + h) − u(x) − 1B1(0)(h)∇u(x) · hµ(dh) + Rd i.e. with constant coefficients. Further, ∀u ∈ C2 b (Rd) and x ∈ Rd we have I(u, x) = min a {fab + Lab(u, x)}. max b Finally, for a universal C, and for all fab and Lab, we have (cid:90) |fab| + |Aab| + |Bab| + |Cab| + min{1,|h|2}µab(dh) ≤ C||I|| Lip,C2 b . →C0 b (4.3.2) Rd The above theorem relies on the following fact from [25], which elaborates on the char- acterization of linear functionals (whose codomain is R) which have the GCP. Lemma 4.3.5. [25, Lemma 3.9] Assume that β ∈ [0, 3). let (cid:96) : Cβ linear functional which has the GCP with respect to 0. For β ≥ 2 and any u ∈ Cβ b (Rd) → R be a bounded b (Rd) ∩ C2(0), we have the representation (cid:90) (cid:104)l, u(cid:105) =C(cid:96)u(0) + (B(cid:96),∇u(0)) + tr(A(cid:96)D2u(0)) u(h) − u(0) − χB1(0) (cid:104)∇u(0), h(cid:105) µ(cid:96)(dh) + Rd This representation is unique, which means that if there were ˜C, ˜B, ˜A, and ˜µ a measure in 70 Rd \ {0} all such that (cid:104)l, u(cid:105) = ˜Cu(0) + ( ˜B,∇u(0)) + tr( ˜AD2u(0)) (cid:90) u(h) − u(0) − χB1(0) (cid:104)∇u(0), h(cid:105) ˜µ(dh) + Rd for all u, then ˜C = C(cid:96), ˜B = B(cid:96), ˜A = A(cid:96), and ˜µ = µ(cid:96). Furthermore, if β < 2 and u ∈ Cβ(Rd) ∩ C1(0), then A(cid:96) = 0, and if β < 1 then B(cid:96) = 0 and the integrand on the right can be replaced by just u(h) − u(0). The proof of the theorem 4.3.4 above employs a lot of information from [25, Lemma 3.6] and its proof, particularly the part which says that for β ∈ [1, 2) (as it is in the case of C1,γ, which is the case relevant to us) and u ∈ Cβ(Rd) ∩ C1(0), (cid:104)l, u(cid:105) = C(cid:96)u(0) + (B(cid:96),φ,∇u(0)) + where Pφ,η,u is given by (cid:90) Rd u(h) − Pφ,η,uµ(cid:96)(dh), (4.3.3) Pφ,η,u,x(·) = u(x) + φ(· − x)(∇u(x),· − x) if β ∈ [1, 2). (4.3.4) Finally, we state the result from [9] about the structure of the operator J that satisfies the assumption 4.3.3. Theorem 4.3.6. (Translation invariant min-max) If J follows all the conditions in the assumption 4.3.3, then J admits a min-max representation as follows: ∀f ∈ K(γ, δ, m), J(f, x) = min g∈K(γ,δ,m) max L∈L(K(γ,δ,m)) {J(g, x) + L(f − g, x)} (4.3.5) 71 in which each L is also translation invariant. This means L ∈ Linv which is the class that contains all linear operators of the form L(f, x) =cf (x) + b · ∇f (x) (cid:90) + Rd f (x + h) − f (x) − 1B1 (h)∇f (x) · hµ(dh), (4.3.6) where each of c, b, µ are independent of x. Furthermore, given a C1 > 0, ∃C2 > 0 such that for all J with a Lipschitz norm bounded by C1, all such c, b, µ resulting from an L ∈ Linv, we have |c|,|b| < C2, and (cid:90) Rd (cid:90) Rd\BR max{|h|1+γ, 1}µ(dh) ≤ C2, and µ(dh) ≤ C2ω(R). (4.3.7) The class of operators, Linv in (4.3.6) and the min-max in (4.3.5) both depend on γ, δ, m via K(γ, δ, m). 72 Chapter 5 Parabolic equations and free boundary problems 5.1. Introduction to the problem and results The goal of this chapter is to reproduce the results from [9] as mentioned above in section 4.3; particularly theorems 4.3.1 and 4.3.2, but this time for the specific equation given in (4.1.4) with the following details: (a) F1 = F2 = ∆, (b) G = |∇U + f | − |∇U− f |, (c) and for functions f defined over the sphere Sn instead of Rd. Thus, consider the following free boundary problem for U : Rn+1 → R ∆U = 0 V = |∇U +| − |∇U−| in {U (·, t) > 0} and {U (·, t) < 0} in ∂{U > 0} (5.1.1) Once again, we consider the hypersurface ∂{U (·, t) > 0} as the graph of a function f , but this time f : Sn × [0, T ]. Then, for a sufficiently smooth f , the equation (5.1.1) is equivalent 73 to ∂tf = H(f, x) = (|∇U + (cid:113) f |) f | − |∇U + f (·, 0) = f0 where H : C1,γ(Sn) → Cγ(Sn). 5.1.1. Main results 1 + |∇f|2 on Sn × [0, T ] (5.1.2) on Sn Theorem 5.1.1. For the equations given above, if we have f : Sn × [0, T ] → [δ,∞), Uf : Sn × [0, T ] → R are such that ∀t ∈ [0, T ], Γ(t) = ∂{Uf (·, t) > 0} = graph f (·, t), then, (a) Uf is a viscosity solution of the free boundary evolution, (5.1.1) if and only if f is a viscosity solution of the fractional parabolic equation (5.1.2). (b) If additionally, graph f (·, t) = ∂{Uf (·, t)} enjoys a modulus of continuity, i.e. |f (x, 0) − f (y, 0)| ≤ ω(|x − y|), then ∀t ∈ [0, T ] this modulus is preserved for f (·, t) and hence for ∂{Uf (·, t)}. (c) If graph f (·, t) = ∂{Uf (·, t)} enjoys a modulus of continuity, i.e. |f (x, 0) − f (y, 0)| ≤ ω(|x − y|), then there exists a unique viscosity solution to both (5.1.1) and (5.1.2). Once again, the precise definitions of the notions of viscosity solutions that are mentioned in these results will be given later, in sections 5.4.1 and 5.3.1. Now we can define the D-to-N 74 operator I, via the inward normal derivative of Uf which is given by I+(f, x) := ∂+ ν Uf (x, f (x)), I−(f, x) := ∂− ν Uf (x, f (x)) for the two-phase problem. Here ∂+ ν Uf are as defined in (4.2.2), (4.2.6) respectively. ν Uf , ∂− Since the gradient of Uf at a point is orthogonal to the level curve of Uf at the same point, we have that the normal to the curve at point is given by the gradient. i.e. we have ν = ± ∇Uf |∇Uf|, and since ∂νUf = ∇Uf · ν, we have ∂νUf = |∇Uf|. Similarly, ∂+ ν Uf = |∇U + f |, and ∂− ν Uf = |∇U− f |. This of course, establishes the connection between I and H, and hence one of the results that leads to the proof of theorem 5.1.1 is the property of I which we list in the following result. Theorem 5.1.2. If I is as given above, and γ ∈ (0, 1) is fixed, then ∃γ(cid:48) with 0 < γ(cid:48) < γ so that I : K(γ, δ, m) → Cγ(cid:48) (Sn) and I is locally Lipschitz. Also, I enjoys the following representation for aij, bij, cuj, µij(dh) which are all independent of x: 75 (cid:110) (cid:90) i j (cid:16) + Sn I(f, x) = min max aij + cijf (x) + bij · ∇f (x) f (x + h) − f (x) − 1B1 (h) (cid:68)∇f (x), exp−1 x (h) (cid:69)(cid:17) (cid:27) µij(dh) The bounds of aij, bij, cuj, µij(dh) all depend on the bounded set {f : f ≥ δ,||f||C1,γ (Sn) ≤ m}. Furthermore, ∃C > 0 such that |aij|,|bij| ≤ C, −C ≤ cij ≤ 0, and (cid:90) min Sn (cid:110)|h|1+γ, 1 (cid:111) (cid:90) Sn µij(dh) ≤ C and µij(dh) ≤ C We shall give the proofs of these results in section 5.4. However, we will give a brief outline of the steps involved in the proof. (1) We will first study a general class of operators acting on the functions on the sphere (with some reasonable level of smoothness). The connection of these operators to the problem at hand is that the H described in (5.1.2) will belong to this class of operators. Moreover, this H depends on the D-to-N map, I. We will analyse these operators in terms of their enjoyment of the min-max structure, which is the result given by [9, Theorem 1.4], modified here as 5.1.1. (2) Next, we will study the existence and uniqueness properties of viscosity solutions to parabolic equations on the sphere. These equations will involve the operators mentioned in the previous step, as given by (5.1.2). (3) We will go on to show how that the viscosity solutions of the two different equations (free boundary problem, i.e. (5.1.1) and the parabolic equation, i.e. (5.1.2)) produce 76 an equivalent outcome, accounting for the correspondence between the free boundary in (5.1.1) and the graph of the function f from (5.1.2). This is the equivalence of solutions listed in [9, Theorem 1.1], modified here as 5.1.2. (4) Finally, we will prove that a modulus of continuity for the initial data will be propagated for all time. This will lead us to part (c) of the theorem 5.1.1 The work in this chapter will focus primarily on steps (1) and (2), and the detailed expla- nations of steps (3) and (4) can be directly modified from [9], although we given a brief explanation in section 5.4. 5.1.2. Details of two phase free boundary problem on the sphere Here we give a few more technical details of the equations in question along with some pictures. Consider any function f : Sn → R, which is non-negative, bounded, and bounded away from zero. Thus, we have r0, δ, L > 0 so that 0 < r0 < δ ≤ f ≤ L − δ < L. Now consider the following sets: f := {X ∈ Rn+1 : X = re, e ∈ Sn, 0 ≤ r < f (x)} D+ D− f := {X ∈ Rn+1 : X = re, e ∈ Sn, f (x) ≤ r < L} Γ− f := {X ∈ Rn+1 : X = re, e ∈ Sn, r = f (x)} Now suppose we have a function Uf : Sn × [0, T ] → R that solves the following free 77 ∆Uf = 0 D+ f ∆Uf = 0 D− f f (e) e Sn Uf = 1 Uf = −1 BL V = |∂+ ν Uf| − |∂− ν Uf| Figure 5.1: Two phase free boundary problem on the sphere boundary problem ∆Uf = 0 Uf = 0 Uf = 1 Uf = −1 V = |∂+ ν Uf| − |∂− ν Uf| in D± f on Γf on Bδ(0) on BL(0) on Γf Now consider the parabolic equation: ∂tf =(cid:0)|∂+ ν Uf| − |∂− ν Uf|(cid:1)(cid:113) 1 + |∇f|2 on Sn × [0, T ] (5.1.3) (5.1.4) The main result adapted from [9] describes the connection between the parabolic equa- tions and the (two-phase) free boundary problem as given in the equations above. 78 Note that the one-phase problem is given by ∆Uf = 0 Uf = 1 in Df in Br0 Uf = 0 on graph f (5.1.5) 5.2. Integro-differential representations of certain operators Before we prove the results for the specific two-phase free boundary problem in this work, we will first consider a broad abstract class of operators that fulfil certain conditions. Such an operator will appear in the parabolic equation as given in (4.1.5). In the next two sections, we will prove a min-max representation for the operators, existence and uniqueness of solutions, as well as the comparison principle for the equations involving these abstract operators. In the final section 5.4, we will elaborate on why the operators in two phase equation we have chosen definitely falls into the category of operators we will describe in this section, and as such it will be evident why all the subsequent results will be applicable to (5.1.1), (5.1.2) as well. Assumption 5.2.1. We look at the broad set of operators that act on functions on Sn and satisfy the following assumptions: (i) 0 < γ < 1 is fixed. (cid:32)(cid:83) (cid:83) (cid:33) (ii) J : K(γ, δ, m) → C0(Sn). δ>0 m>δ (iii) For each δ and m, J is a Lipschitz mapping on the sets K(γ, δ, m), whose Lipschitz constant, C(δ, m) increases as δ decreases or m increases. 79 (iv) J satisfies GCP. (v) J is rotational invariant. (vi) If f ∈ K(γ, δ, m) and c > 0 is a constant, then ∀x ∈ Sn, J(f + c, x) ≤ J(f, x). In this section we will first record some of the relevant results about the structure and properties of J, but in the case when J acts on functions in Rn. All of the results below can be found in [25]. Proposition 5.2.2. If (cid:96) : K(γ, δ, m) → R is a bounded linear function that enjoys GCP at x0 = 0, then ∃b, c ∈ R and a measure µ so that ∀f ∈ K(γ, δ, m) (cid:96)(f ) = cf (0) + b · ∇f (0) + 0 (h) µ(dh), (cid:90) Sn (f (h) − f (0) −(cid:68)∇f (0), exp−1 (cid:69) where b, c, and µ satisfy the same bounds as in the above theorem, but none of them depend on x. This result for functions in Rn and its proof can both be found in [24, Lemma 3.6]. However, for functions defined on the sphere, the Taylor polynomial given in (4.3.4) will now change to Pφ,η,u,x(·) = u(x) + φ(Rx(·))(∇u(x), exp−1 x (·)) The techniques for the proof in [24, Lemma 3.6] will follow in a straightforward manner. We also note that since φ ∈ C1,γ(Sn), the term with the gradient (∇u(x), exp−1 x (·)), will decay 80 at the rate of |x − ·|1+γ, which is integrable because we will have (cid:90) Sn max{|h|1+γ, 1}µ(dh) ≤ C. Theorem 5.2.3. (Rotationally invariant min-max) If J follows all the assumptions above, then J admits a min-max representation as follows: ∀f ∈ K(γ, δ, m), J(f, x) = min g∈K(γ,δ,m) max L∈L(K(γ,δ,m)) {J(g, x) + L(f − g, x)} (5.2.1) in which each L is also rotationally invariant. This means L ∈ Linv which is the class that contains all linear operators of the form L(f, x) =cf (x) + b · ∇f (x) f (Rx(h) − f (x) −(cid:68)∇f (x), exp−1 x (h) (cid:69) µ(dh), (5.2.2) (cid:90) + Sn where each of c, b, µ are independent of x. Furthermore, given a C1 > 0, ∃C2 > 0 such that for all J with a Lipschitz norm bounded by C1, all such c, b, µ resulting from an L ∈ Linv, we have |c|,|b| < C2, and max{|h|1+γ, 1}µ(dh) ≤ C2. (5.2.3) (cid:90) Sn The class of operators, Linv in (5.2.2) and the min-max in (5.2.1) both depend on γ, δ, m via K(γ, δ, m). The consequences of the above theorem are as follows: 81 • Following as in [9], this identifies a natural class of extremal operators for J as M + inv = max L∈Linv L(f, x), and M− inv = min L∈Linv L(f, x) (5.2.4) Just by the definition, we can use the min-max expression for J and write J(f, x) − J(g, x) = min {J(g(cid:48), x) + L(f − g(cid:48), x)} − J(g, x) g(cid:48) max L ≤ max L L(f − g, x) (Taking g(cid:48) = g in the minimum) and J(f, x) − J(g, x) = J(f, x) − min {J(f(cid:48), x) + L(g − f(cid:48), x)} L f(cid:48) max L(g − f, x) ≥ − max L (Taking f(cid:48) = f in the minimum) L(f − g, x) = min L These extremal operators are defined specifically to produce, ∀f, g ∈ K(γ, δ, m), the following inequalities: M− inv(f − g, x) ≤ J(f, x) − J(g, x) ≤ M + inv(f − g, x) (5.2.5) We note that because of the rotational invariance we have that ∀x ∈ Rn, M± inv(f, x) = M±(f ◦ Rx, 0) • Secondly, due to the rotational invariance of J, we see that all of the desired prop- erties of Linv can be obtained from studying the (possibly non-linear) functional j : K(γ, δ, m) → R with j(f ) := J(f, 0). 82 This functional, j, also satisfies GCP based at x0 = 0. So once J is fixed, the class L can be further restricted to include only those L that are in the Clarke differential of j (from definition 4.2.4), i.e. those L such that L(f, x) = (cid:96)(f ◦ Rx) for some choice of (cid:96) ∈ [∂j]K(γ,δ,m) Thus, we can now define a new extremal operator, depending explicitly on J and K(γ, δ, m) as M + J,K(γ,δ,m)(f, x) = and M− J,K(γ,δ,m)(f, x) = max (cid:96)∈[∂j]K(γ,δ,m) min (cid:96)∈[∂j]K(γ,δ,m) ((cid:96)(f ◦ Rx)) ((cid:96)(f ◦ Rx)) (5.2.6) Now for the operator given above, we note that the inequalities in (5.2.4) still hold, they are again, rotation invariant, and they also serve as extremal operators of J, which we talk about in the result below. Proposition 5.2.4. If J is fixed, and M± M± J,K(γ,δ,m) also obey the inequalities in (5.2.5), and M± mappings from C1,γ(Sn) → C0(Sn) with a Lipchitz norm bounded by that of J. J,K(γ,δ,m) are defined as in (5.2.6) above, then J and J,K(γ,δ,m) are Lipschitz functions, as Proof. Firstly, from [13, Proposition 2.1.2], we know that [∂j]K(γ,δ,m) (cid:54)= ∅. This is a result of the Hahn-Banach Theorem, which asserts that every positively homogeneous and subad- ditive functional on a vector space majorises a linear functional on the vector space. In this case, the vector space being C1,γ(Sn) × C1,γ(Sn) and j0 is the positively homogeneous and subadditive functional, so there exists some linear functional (cid:96) so that for all ψ ∈ C1,γ(Sn), 83 j0(f ) ≥ (cid:104)(cid:96), ψ(cid:105), which means [∂j]K(γ,δ,m) (cid:54)= ∅ as claimed. Secondly, j and ∂j enjoy a mean value property due to a theorem by Lebourg [13, Proposition 2.3.7], which says that for any points x, y ∈ X and f Lipschitz on an open set containing the line segment [x,y], ∃ u ∈ (x, y) such that f (y) − f (x) ∈ (cid:104)∂f (u), y − x(cid:105). Thus, for f, g ∈ K(γ, δ, m) and j Lipschitz, we can say that there is an element, (cid:96) ∈ [∂j]K(γ,δ,m) such that j(f ) − j(g) = (cid:96)(f − g). Now we take a max over [∂j]K(γ,δ,m), and again from [13, Proposition 2.1.2], [∂j]K(γ,δ,m) is weak-(cid:63) compact, thus giving us ∀f, g ∈ K(γ, δ, m), j(f ) − j(g) ≤ max (cid:96)∈[∂j]K(γ,δ,m) (cid:96)(f − g) A similar argument gives us the lower inequality in (5.2.5). To see why M± are Lipschitz, consider J,K(γ,δ,m)(f1, x) − M + M + J,K(γ,δ,m)(f2, x) = ≤ max (cid:96)∈[∂j]K(γ,δ,m) max (cid:96)∈[∂j]K(γ,δ,m) max ((cid:96)(f1 ◦ Rx)) − (cid:96)∈[∂j]K(γ,δ,m) {((cid:96)(f1 ◦ Rx)) − ((cid:96)(f2 ◦ Rx))} ((cid:96)(f2 ◦ Rx)) = max c,b,µ (cid:90) −(cid:68) + x ∇f1(x) − R−1 {c(f1(x) − f2(x)) + b · [R−1 (cid:68) f1(Rx(h)) − f2(Rx(h)) − f1(x) + f2(x) Sn x ∇(f1(x)), exp−1 R−1 x ∇(f1(x)), exp−1 R−1 x ∇f2(x)] (cid:69) 0 (h) + 0 (h) (cid:69) µ(dh)} Owing to the fact that f1, f2 ∈ K(γ, δ, m) and due to the bound on µ given in (5.2.3), 84 all the terms above are bounded up to a constant by the C1,γ-norm of f1 − f2. Proof of theorem 5.2.3: Proof. From the previous proposition, we can see that ∀f, g ∈ K(γ, δ, m), j(f ) ≤ max (cid:96)∈[∂j]K(γ,δ,m) (j(g) + (cid:96)(f − g)) Thus j(f ) ≤ min g∈K(γ,δ,m) max (cid:96)∈[∂j]K(γ,δ,m) (j(g) + (cid:96)(f − g)) (by taking minimum over g) On the other hand, by letting g = f in the minimum, min g∈K(γ,δ,m) max (cid:96)∈[∂j]K(γ,δ,m) (j(g) + (cid:96)(f − g)) ≤ j(f ) Since J is rotation invariant, we see that J(f, x) = j(Rxf ). Now, any (cid:96) ∈ [∂j]K(γ,δ,m) enjoys the comparison principle, and thus we have that (cid:96) has a representation as in (5.2.2), and thus the proof follows. Corollary 5.2.5. For each γ, δ, and m fixed, we can extend the operator J (with respect to the set K(γ, δ, m)) to a function on all of C1,γ(Sn). We define ˜JK(γ,δ,m) : C1,γ(Sn) → C0(Sn) via ˜JK(γ,δ,m)(f, x) = min g∈K(γ,δ,m) max (cid:96)∈[∂j]J,K(γ,δ,m) {j(g) + (cid:96)((f ◦ Rx) − g)} (5.2.7) ˜J is Lipschitz, it enjoys GCP, and ˜J = J on K(γ, δ, m). (It also satisfies all of the other assumptions.) 85 Proof. We can see that ˜J is Lipschitz because M + is Lipschitz and from the following: ˜J(f1) − ˜J(f2) = min h∈K(γ,δ,m) max (cid:96)∈[∂j]J,K(γ,δ,m) {j(h) + (cid:96)((f1 ◦ Rx) − h)} − min g∈K(γ,δ,m) max k∈[∂j]J,K(γ,δ,m) {j(g) + k((f2 ◦ Rx) − g)} = min h∈K(γ,δ,m) max (cid:96)∈[∂j]J,K(γ,δ,m) {j(h) + (cid:96)((f1 ◦ Rx) − h)} − ≤ − ≤ max k∈[∂j]J,K(γ,δ,m) {j(g(cid:63)) + k((f2 ◦ Rx) − g(cid:63))} (Assume g = g(cid:63) is the optimizer for the second minimum) {j(g(cid:63)) + (cid:96)((f1 ◦ Rx) − g(cid:63))} max (cid:96)∈[∂j]J,K(γ,δ,m) max k∈[∂j]J,K(γ,δ,m) {j(g(cid:63)) + k((f2 ◦ Rx) − g(cid:63))} (setting h = g(cid:63) in the first min) max (cid:96)∈[∂j]J,K(γ,δ,m) {j(g(cid:63)) + (cid:96)((f1 ◦ Rx) − g(cid:63)) − j(g(cid:63)) − (cid:96)((f2 ◦ Rx) − g(cid:63))} ((cid:96)(f1 ◦ Rx) − (cid:96)(f2 ◦ Rx)) ≤ max (cid:96)∈[∂j]J,K(γ,δ,m) = M +(f1 − f2) Further, ˜J satisfies GCP because j, (cid:96) satisfy GCP at x0 = 0. Corollary 5.2.6. If x is fixed, and γ, δ, m are given, and f ∈ m − C1,γ(x), ˜JK(γ,δ,m)(f, x) is classically defined via (5.2.6). Proof. Since f is m − C1,γ(x), then by definition, there are two C1,γ functions f + and f− such that they touch f above and below respectively at x. In other words, f± ∈ C1,γ(Rn), and ∀y, f−(y) ≤ f (y) ≤ f +(y) with f−(x) = f (x) = f +(x). By Proposition 5.2.2, we know 86 that if a bounded linear functional (cid:96) has GCP at a point then (cid:96)(Rxf ) = L(f, x). We can work this out explicitly from (cid:69) µ(dh) 0 (h) (cid:96)(f ◦ Rx) =c(f ◦ Rx)(0) + b · ∇(f ◦ Rx)(0) b, R−1 x ∇f (x) (f ◦ Rx)(h) − f ◦ Rx(0) −(cid:68)∇(f ◦ Rx)(0), exp−1 (cid:68) (cid:69) f (Rxh) − f (x) −(cid:68) f (Rxh) − f (x) −(cid:68)∇f (x), exp−1 x ∇f (x), exp−1 R−1 (cid:69) 0 (h) x (h) µ(dh) µ(dh) (cid:69) =cf (x) + (cid:104)Rx · b,∇f (x)(cid:105) + Sn =cf (x) + (cid:90) (cid:90) (cid:90) + Sn + Sn =L(f, x), with the corresponding bL = Rx · b and the same c. Thus the formula in (5.2.7) holds classically, i.e. as in theorem 5.2.3 with the estimate on µ that appears in 5.2.3. Proposition 5.2.7. If J is as in Assumption 5.2.1 and M + J,K(γ,δ,m) is as defined in (5.2.5) then ∀x ∈ Rn, M + J,K(γ,δ,m)(1, x) ≤ 0 As a result, we see that ∀L ∈ Linv, we have c ≤ 0 in (5.2.2). Proof. Note that in (5.2.2), L(1, 0) = c. Because of the rotation invariance of M + J,K(γ,δ,m), it is enough to show that M + J,K(γ,δ,m)(1, 0) ≤ 0. It can also be observed here that the consequence of the result follows from having c = L(1, 0) = max (cid:96)∈[∂j] (cid:96)(R01) = M + J,K(γ,δ,m)(1, 0) 87 Suppose we write m0 to denote the Lipschitz functional m0(f ) = M + J,K(γ,δ,m)(f, 0) = max (cid:96)∈[∂j] (cid:96)(f ), where j = j(f ) = J(f, 0). Assumption 5.2.1 says that J(f + c, x) ≤ J(f, x) for c > 0 and f ∈ K(γ, δ, m). Thus j0(f ; 1) = lim sup x→0+ By definition, ∂j(f ) := {(cid:96) ∈(cid:16) C1,γ(Rn)) and [∂j]K := hull j(f + s · 1) − j(f ) ≤ 0. s (cid:17)(cid:63) |∀ψ ∈ C1,γ(Rn), j0(f ; ψ) ≥ (cid:104)(cid:96), ψ(cid:105)} (cid:91)  ∂j f∈K Thus, if (cid:96) ∈ [∂j]K(γ,δ,m), then (cid:96)(1) ≤ 0. So from the definition of M + can say that M + J,K(γ,δ,m)(1, 0) = m0(1) ≤ 0. J,K(γ,δ,m) in (5.2.5), we 5.3. Comparison theorem and existence for parabolic viscosity solutions on the sphere 5.3.1. Using the assumptions of section 5.2 Now, we look at operators of the type J that satisfy the above assumptions, and for such J, we prove uniqueness for the following equation: ∂tf = J(f ) in Sn × (0, T ] f (·, 0) = f0 on Sn (5.3.1) Definition 5.3.1. We say that f is a viscosity subsolution of (5.3.1) if f is upper semi- 88 continuous, inf f > 0, and if f has the following property: For all (x, t) ∈ Sn × (0, T ] for which there exists a function φ ∈ C1,γ+β(Sn) in space and C1 in time (for some β > 0 with γ + β ≤ 1), and δ ≤ φ ≤ L− δ such that for some t > r > 0, f − φ attains a global maximum over Sn × (t − r, t], then φ must satisfy ∂tφ(x, t) ≤ J(φ, x) On the other hand, we say g is a viscosity supersolution if g is lower semi-continuous, inf g > 0, and if we replace the above properties with g − φ attains a minimum and ∂tφ(x, t) ≥ J(φ, x) Of course, f is a viscosity solution if is satisfies both of the above. Lemma 5.3.2. If f is a viscosity subsolution of (5.3.1), and suppose (x, t) ∈ Sn × (0, T ) is a point such that f − φ attains a maximum at the point (x, t) for some φ that is punctually C1,γ+β(x) in space and C1 in time, with β > 0, γ + β ≤ 1. Then there exists a choice of δ0 small enough and m0 large enough such that ˜JK(γ,δ0,m0)(φ, x) is classically defined and ∂tφ(x, t) ≤ ˜JK(γ,δ0,m0)(φ, x) Remark 5.3.3. The analogous result holds for g that are supersolutions and for those φ we also require them to additionally satisfy inf φ > 0. Note that this is not a problem with the subsolution because f − φ ≤ 0 =⇒ δ ≤ f ≤ φ =⇒ inf φ > 0. Remark 5.3.4. ˜J is defined in (5.2.6) and we need this because φ is not necessarily in the 89 set, K(γ, δ0, m0). Also, this result will hold for all δ < δ0 and m > m0. Proof. The well defined nature of ˜JK(γ,δ0,m0)(φ, x) is because of Corollary 5.2.6, from which we can also see that we only need φ ∈ C1,γ(x) and not γ + β. The extra regularity will come in later. Now, since φ is punctually C1,γ+β at x, this means there are two C1,γ+β functions f +, f−, such that ∀y ∈ Rn, f−(y) ≤ φ(y) ≤ f +(y) and f−(x) = φ(x) = f +(x) Now for each r > 0, we can define a function φr such that f +(y) if y ∈ Br(x) otherwise. φr(y) = φ(y) Thus, we also have the ordering f− ≤ φ ≤ φr ≤ f + and ∇f−(x) = ∇φ(x) = ∇φr(x) = ∇f +(x) (they all touch at the point x). Finally, we can also assume wlog that ∂tφ(x, t) = ∂tf +(x, t) f + ∈ C1,γ+β(Rn) is a function such that f − f + attains a maximum at (x, t). Now, f + ≥ φ and inf φ > 0, and also, f + ∈ C1,γ+β. These facts imply that there exists some δ0 and m0 such that f + ∈ K(γ, δ0, m0) = K (we call is so for simplicity). Thus by definition of 90 viscosity subsolution, ∂tφ(x, t) = ∂tf +(x, t) ≤ J(f +, x) = ˜JK(f +, x) Now we use Corollaries 5.2.5 and 5.2.6, which tell us that ˜J has many of the properties of J and satisfies many of the same inequalities. ˜JK(f +, x) ≤ ˜JK(φr, x) + M + ≤ ˜JK(φ, x) + M + J,K(f + − φr, x) J,K(φr − φ, x) + M + (since ˜J also satisfies (5.2.4)) J,K(f + − φr, x) All of the above operators are well defined on these functions, as they are all punctually C1,γ(x). If we can show that the last two terms → 0 as r → 0, then we complete the proof. This is where we will need the slightly higher regularity. Now following the definitions: J,K(φr − φ, x) = max M + (cid:96)∈[∂j]K (cid:96)((φr − φ) ◦ Rx) ≤ max L∈Linv L(φr − φ, x) 91 Also, from Theorem 5.2.3, we obtain the following bounds: L((φr − φ, x) =c(((((((( (φr(x) − φ(x)) − b · (((((((((( (∇φr(x) − ∇φ(x)) (cid:90) (cid:90) φr(Rx(h)) − φr(x) − 1Br(x) φ(Rx(h)) − φ(x) − 1Br(x) (cid:68)∇φr(x), exp−1 (cid:68)∇φ(x), exp−1 x (h) Sn + − (cid:69) Sn (cid:69) x (h) µ(dh) µ(dh) (at x, we have φr = φ,∇φr = ∇φ) (cid:90) ≤ Br(x) (Br(x) since outside of this set, ∇φr − ∇φ = 0) Cφ|h|1+γ+βµ(dh) (Cφ is the constant depending on φ, φr ∈ C1,γ+β) We can make a similar statement about f + − φr. Now, for each L, there is a µ and hence M + J,K is a max over a family of the measures µ, arising from the set [J]K, and each of these µ satisfy the uniform bound in (5.2.3). Suppose we say meas(K) := {µ|∃(cid:96) ∈ [∂j]K s.t. µ corresponds to (cid:96) as in Proposition 5.2.2} Then we can say that we have the following bounds J,K(φr − φ, x) + M + M + J,K(f + − φ, x) ≤ (cid:90) Cφ|h|1+γ+βµ(dh) (cid:90) |h|1+γµ(dh) Br ≤ Cφrβ Br (cid:90) Br ≤ Cφrβ sup µ∈meas(K) ≤ C2 · Cφ · rβ 92 min{|h|1+γ, 1}µ(dh) Hence, as r → 0, we see that M + J,K(φr − φ, x) + M + J,K(f + − φ, x) → 0 and thus we have proved the lemma. We now propose that the inf and sup convolutions that are appropriate for the parabolic equation on Sn × [0, T ] are the following: φε(x, t) := sup y=Rzx,d(z,0)≤c0,t(cid:48)∈[0,T ] φ(y, s) − 1 2ε φε(x, t) := y=Rzx,d(z,0)≤c0,t(cid:48)∈[0,T ] inf φ(y, s) + 1 2ε (cid:16)||x − y||2 + |t − t(cid:48)|2(cid:17) (cid:16)||x − y||2 + |t − t(cid:48)|2(cid:17) (5.3.2) (5.3.3) Another way to write these (and a notations we will often use) is as follows: φε(x, t) := sup d(z,0)≤c0,s∈[−t,T−t] φ(Rzx, t + s) − 1 2ε φε(x, t) := inf d(z,0)≤c0,s∈[−t,T−t] φ(Rzx, t + s) + 1 2ε (cid:16)||Rz||2 + |s|2(cid:17) (cid:16)||Rz||2 + |s|2(cid:17) Lemma 5.3.5. We note the following properties of the sup/inf-convolutions: (i) The sup-convolution given in (5.3.2) is semi-convex and the inf-convolution in (5.3.3) is semi-concave. (ii) the sup and inf convolution are Lipschitz functions. (iii) Both φε and φε converge to φ pointwise as ε → 0. (iv) If u and v are a subsolution and supersolution to the equation ∂tu(x, t) = J(u, x) in Sn respectively, then uε and vε are also a subsolution and supersolution respectively to the same equation. 93 Proof. (i) It is easy to see from the definition directly that the inf-convolution, φε is semi- concave. We can write the sup-convolution as φε(x) = − inf d(z,0)≤c0,s∈[0,T ] −φ(Rzx, s) + 1 2ε (cid:16)||x − Rzx||2 + |t − s|2(cid:17)(cid:27) Thus, −φε is semi-concave by definition, which makes φε semi-convex. We can also see this from knowing that φε + |X|2/2ε is convex (where X = (x, t)). Similarly, we can see that φε is semi-concave. (ii) Consider for X1 = (x1, t1), X2 = (x2, t2) ∈ Sn × [0, T ], and let Y ε 1 = (y1, s1) = (Rzεx1, s1) be the point at which the supremum is realized for X1, or φε(x1, t1) = φ(Rzεx1, s1) − 1 2ε (cid:16)||x1 − Rzεx1||2 + |t1 − s1|2(cid:17) . On the other hand, we also have (cid:16)||x2 − Rzx2||2 + |t2 − s|2(cid:17)(cid:27) φε(x2, t2) = sup d(z,0)≤c0,s∈[0,T ] ≥ φ(Rzεx1, s1) − 1 2ε φ(Rzx2, t2) − 1 2ε (cid:16)||x2 − Rzεx1||2 + |t2 − s1|2(cid:17) where zε 1 is such that Rzx2 = Rzε 1 x1 for some z ∈ Bc0. Thus, we can write φε(x1, t1) − φε(x2, t2) ≤((((((( (cid:26) (cid:26) (cid:16)||x1 − Rzε x1||2 + |t1 − s1|2(cid:17) (cid:16)||x2 − Rzεx1||2 + |t2 − s1|2(cid:17) 1 φ(Rzεx1, s1) − 1 2ε − ((((((( φ(Rzεx1, s1) + ≤ C1|x1 − x2| + C2|t1 − t2| ≤ C||(x1, t1) − (x2, t2)|| 1 2ε 94 (C = C(Sn, T )) thereby showing that φε is Lipschitz. We can similarly show that φε is Lipschitz. (iii) This proof is based on the one given in [15, Lemma A.5]. Assume ψ is an appropriate function that touches φε from above at (x, t) = X, and that Dψ(x, t) = q ∈ Rn+1, D2ψ(x, t) = M ∈ Sd+1, the set of symmetric (d + 1) × (d + 1) matrices. We will first show that this means ∃ some ψ(cid:48) that touches φ at X + qε and φε(X) + ε 2 |q|2 = φ(X + qε) Indeed, suppose Y ε (x,t) is the point such that the supremum is achieved in the definition of φε(X). We will use the notation Y ε (x,t) = Y ε = (yε, tε), i.e. φε(x, t) = φ(yε, tε) − 1 2ε (||yε − x||2 + |tε − t|2) From similar calculations as in [15], we get Y ε = X + qε. Thus, we have φε(X) + ε 2 |q|2 = φ(Y ε) − 1 2ε ε 2 |q|2 |X − Y ε|2 + |q|2 = φ(X + qε) − ε 2 |q|2 + ε 2 = φ(X + qε) 95 We also know that |q|ε → 0 when ε → 0 because φ(X) ≤ φε(X) = φ(Y ε) − 1 2ε |X − Y ε|2 =⇒ 1 2ε |X − Y ε|2 ≤ φ(Y ε) − φ(X) ≤ 2||φ||L∞ =⇒ |X − Y ε| ≤(cid:112)4ε||φ||L∞ → 0 as ε → 0 Finally, we see that (cid:110) lim sup ε→0 φε(X) + |q|2(cid:111) ε 2 = lim sup ε→0 ≤ φ(X) ≤ lim sup ε→0 φ(X + qε) φε(X) (φ is USC) (since we have φε ≥ φ) This forces lim supε→0 ε 2|q|2 = 0, and eventually, we can see that lim sup ε→0 φε(X) = lim sup ε→0 φ(X + qε) − ε 2 |q|2 = φ(X) (iv) Suppose u, v are respectively the subsolution, supersolution of the given equation, i.e., we have in the viscosity sense, that ∂tu(x, t) ≤ J(u, x), and ∂tv(x, t) ≥ J(v, x). We shall use u◦Rz to denote the function u◦Rz(x, t) = u(Rzx, t), where Rz : Sn → Sn is the rotation on the sphere that takes 0 to z. Note that we have J(u◦Rz, x) = J(u, Rzx) 96 since J is a rotation invariant operator. So as u is a subsolution, we should have ∂t(u ◦ Rz)(x, t) = ∂tu(Rzx, t) ≤ J(u, Rzx) = J(u ◦ Rz, x). (5.3.4) Now, suppose φ is the required function such that uε − φ has its maximum at a point (x0, t0) = X0, i.e., we have uε ≤ φ and uε(x0, t0) = φ(x0, t0). Suppose that in the definition of uε(x, t), zε, sε are the parameters so that the supremum is realized for (x, t), and in particular, zε 0, sε 0 are the parameters for the supremum in the definition of uε(x0, t0). Then we have uε(x0, t0) = u(Rzε 0 x0, t0 + sε 0) − 1 2ε Further, ∀(x, t) ∈ Sn × [0, T ], since uε is the sup (cid:16)||Rzε 0 0|2(cid:17) ||2 + |sε φ(x, t) ≥ uε(x, t) ≥ u(Rzε 0 x, t + sε and φ(x0, t0) = uε(x0, t0) = u(Rzε 0 x0, t0 + sε (cid:16)||Rzε (cid:16)||Rzε 0 0 0) − 1 2ε 0) − 1 2ε 0|2(cid:17) 0|2(cid:17) . ||2 + |sε ||2 + |sε Let φ(cid:48) = φ + cε, where cε = 1/2ε(||Rzε 0 ||2 + |sε 0|2). From earlier, we have ∂tu ◦ Rzε 0 (x, t) ≤ J(u ◦ Rzε 0 , x). 97 Note that φ(cid:48) touches u ◦ Rzε 0 (· + (0, sε 0)) from above at x0, i.e. we have φ(cid:48)(x, t) ≥ u(Rzε and φ(cid:48)(x0, t0) = u ◦ Rzε 0 (x0, t0 + sε 0) 0 x, t + sε 0) = u ◦ Rzε 0 (x, t + sε 0) Since u is the viscosity subsolution to a rotation invariant equation, we have from (5.3.4) ∂tφ(cid:48)(x0, t0) ≤ J(φ(cid:48), x0). Also, since cε (or sε 0) is constant, the function uε solves in the viscosity sense ∂t(uε)(x, t) = ∂t(uε + cε)(x, t) ≤ J(uε + cε, x) ≤ J(uε, x) from the properties of J. We have already seen from the proof of part (iii) that cε → 0 as ε → 0. See previous proof. We also need to know how the family of operators described in Assumption 5.2.1 will operate on rescaled versions of smooth bump functions. Lemma 5.3.6. For x0 ∈ Sn, if we define smooth functions Φ and ΦR as Φ(x) = |x − x0|2 1 + |x − x0|2 , and ΦR(x) = Φ (cid:16) x (cid:17) R Then, for a fixed J, given and δ > 0, m > δ, ρ > 0,∃R > 1, with R = R(J, ρ, δ, m), so that J,K(γ,δ,m)(ΦR, x) ≤ ρ M + sup x∈Sn Proof. First, we will observe the following facts about ΦR 98 0 ≤ ΦR(x) ≤ 1 and ||∇ΦR||L∞ ≤ C R (cid:12)(cid:12)(cid:12)ΦR(Rhx) − ΦR(x) −(cid:68)∇ΦR(x), exp−1 (1 + |x − x0|2)2 =⇒ |∇Φ| ≤ 1 =⇒ |∇ΦR| ≤ C 2(x − x0) x (h) R (cid:69)(cid:12)(cid:12)(cid:12) ≤ C|h|2 R2 Note that ∇Φ = The first fact is owing to the fact that since |∇Φ| ≤ 1, we have |∇ΦR| ≤ C/R when we rescale. The second fact follows from the Taylor expansion of ΦR. Now, from the estimates in Theorem 5.2.3, we can choose R large enough so that, uni- formly across [∂J]K(γ,δ,m) (cid:90) Sn (cid:12)(cid:12)(cid:12)ΦR(Rhx) − ΦR(x) − 1B1(x)(h) (cid:68)∇ΦR(x), exp−1 (cid:69)(cid:12)(cid:12)(cid:12) µ(dh) x (h) |h|2 R2 µ(dh) ≤ ρ 2 (cid:90) ≤ C Sn Next, given this t we can choose R large enough so that uniformly across b ∈ [∂j]K(γ,δ,m) (actually, it is (cid:96) ∈ [∂j]K(γ,δ,m), but there is a one to one correspondence between b and (cid:96) from Proposition 5.2.2), we have b·∇ΦR ≤ ρ/2. Finally, thanks to Proposition 5.2.7, we see that for any of the constants, c, that appear in (5.2.2), we have cΦR(x) ≤ 0. Now if we add everything up, we get (cid:96)(f ◦ Rx), and if we take the max then by the definition of M + J,K(γ,δ,m), we have the required inequality after taking the supremum. 99 Lemma 5.3.7. Given any δ, ρ, C > 0, m > δ, ∃R = R(J, ρ, δ, m) so that ∀h > 0, ∀(x, t) ∈ Sn × [0, T ], Ψ(x, t) = C + hΦR(x) + hρt is a classical, strict supersolution of ∂tΨ > M + J,K(γ,δ,m)(Ψ) Proof. It is a direct calculation to show that J,K(γ,δ,m)(C + hΦR, x) ≤ hM + M + JK(γ,δ,m)(ΦR) Indeed, L(C + hΦR) =c(C + hΦR) + hb · ∇ΦR (cid:90) hΦR(Rxy) − hΦR(x) − 1B1(x)(y)h + Sn (cid:68)∇ΦR, exp−1 x (y)µ(dy) (cid:69) =c · C + hL(ΦR, x) ≤ hL(ΦR, x) (since c ≤ 0 by Proposition 5.2.7 and C > 0 is given) Now we invoke Lemma 5.3.5 with ρ/2 and get M + J,K(γ,δ,m)(Ψ) = M + J,K(γ,δ,m)(C + hΦR) ≤ hM + J,K(γ,δ,m)(ΦR) ≤ hρ/2 < ∂t(Ψ) 100 5.3.2. Comparison results Lemma 5.3.8. If δ > 0, m > δ are fixed, and w : Sn × [0, T ] → R is a bounded, upper semi-continuous function such that, in the viscosity sense, ∂tw ≤ M + w+(·, t) ≤ sup Sn J,K(γ,δ,m)(w) w+(·, 0) then, sup Sn×[0,T ] Proof. First, we will start the proof by assuming supSn w+(·, 0) = 0. Then, we need to show that supSn×[0,T ] w+ ≤ 0. Suppose we assume, for the sake of contradiction, that supSn×[0,T ] w+ > 0. Since for h, C, ρ > 0, Ψ as described in Lemma 5.3.6 is strictly above 0, we can choose h and C in a way so that w+ − Ψ attains a global maximum for some t > 0. If we use Ψ as a test function and apply the definition of a viscosity subsolution for w+, we get a contradiction i.e. ∂tΨ ≤ M + J,K(γ,δ,m)(Ψ), which contradicts what we proved in Lemma 5.3.7. On the other hand, suppose we do not necessarily assume that supSn w+(·, 0) = 0. Now we can replace w by the function ˜w = w − sup Sn w+(·, 0) Thus, let c = supSn w+(·, 0) ≥ 0, we see that w = ˜w + c, and so in the viscosity sense, ∂tw = ∂t ˜w. Also, since c ≥ 0, by Proposition 5.2.7 (which tells us that c ≤ 0 in the 101 expression for L), just like in the proof of Lemma 5.3.7, we get M + J,K(γ,δ,m)(w) = M + J,K(γ,δ,m)( ˜w + c) ≤ M + J,K(γ,δ,m)( ˜w) Hence, in the viscosity sense, we have ∂t ˜w = ∂tw ≤ M + J,K(γ,δ,m)(w) ≤ M + J,K(γ,δ,m)( ˜w) and now, we can apply the first case again since we have supSn ˜w+(·, 0) = 0, which gives us with the assumption above ˜w+(·, t) ≤ 0 sup Sn×[0,T ] =⇒ sup Sn×[0,T ] w+(·, t) ≤ sup Sn w+(·, 0) Proposition 5.3.9. Suppose f, g : Sn × [0, T ] → R are bounded, and for some δ > 0, f, g ≥ δ, and they are respectively a subsolution and supersolution in the viscosity sense, of the equation ∂t(u) = J(u), then f ε(x, 0) ≤ gε(x, 0) =⇒ f ε(x, t) ≤ gε(x, t) ∀(x, t) ∈ Sn × [0, T ], (5.3.5) where f ε, gε are as their definitions given in (5.3.2), (5.3.3). Proof. We first note from lemma 5.3.5 that f ε, gε are given in (5.3.2) and (5.3.3), then the 102 following inequalities are satisfied in the viscosity sense: ∂tf ε ≤ J(f ε), and ∂tgε ≥ J(gε) Let ε ∈ (0, 1) be fixed, and we assume ∀x ∈ Sn, f ε(x, 0) − gε(x, 0) ≤ 0. We want to show that f ε(x, t) − gε(x, t) ≤ 0 ∀(x, t) ∈ Sn × [0, T ] we will assume for the sake of contradiction that there exists a time t(cid:63) ∈ (0, T ] such that sup (x,t)∈Sn×[0,T ] (f ε(x, t) − gε(x, t)) = sup x∈Sn (f ε(x, t(cid:63)) − gε(x, t(cid:63))) = m0 > 0 Note that this cannot happen at t = 0 because of the assumption in this case (when t = 0, we have f ε − gε ≤ 0). Since ε is fixed, and m0 is given, we can invoke Lemma 5.3.6 with m = cε−1 and C = cm0, i.e. have some Ψ = cm0 + hΦR + hρt (where R = R(J, δ, ρ, cε−1)) as a strict supersolution. classically for all (x, t) ∈ Sn × [0, T ] i.e. we know Ψ exists such that the following inequality holds ∂tΨ > M + J,K(γ,δ/2,cε−1) (Ψ) Now we can translate Ψ around so it touches f ε − gε from above. Since we are assuming that f ε − gε attains a positive supremum, and Ψ is always positive, we know they can touch if we translate,i.e. we can choose C so that ∃(xε 0) ∈ Sn × (0, T ] such that 0, tε f ε − gε ≤ Ψ, and (f ε − gε)(xε 0, tε 0) = Ψ(xε 0, tε 0) 103 By definition, for any ε > 0, f ε and −gε are semi-convex, which means that they both have tangent paraboloids of opening 1/ε touching from below. Thus this is also true for f ε − gε. Now since a smooth function Ψ touches f ε − gε from above at the point (xε we have that both f ε and −gε must be C1,1 at the point (xε ∂tf ε, ∂tgε,∇f ε,∇gε classically at (xε 0). Thus, by Lemma (5.3.2), we see that the equations hold classically for f ε and −gε. For f ε, we use gε +Ψ as the test function, and for gε, the test function is f ε − Ψ. Furthermore, by Proposition 5.2.4 and Lemma 5.3.2, we get classically 0). Thus, we can evaluate 0, tε 0), 0, tε 0, tε ∂tf ε − ∂tgε ≤ ˜JK(γ,δ,cε−1)(f ε, xε 0) − ˜JK(γ,δ,cε−1)(gε, xε 0) ≤ M + J,K(γ,δ,cε−1) (f ε − gε, xε 0) ≤ M + J,K(γ,δ,cε−1) (Ψ, xε 0) < ∂tΨ (Since M + J,K(γ,δ,cε−1) enjoys GCP.) Since we know that at (xε 0), we have ∂tf ε − ∂tgε = ∂tΨ, this gives us a contradiction. So 0, tε we can conclude that for each ε fixed, sup (x,t)∈Sn×[0,T ] (f ε(x, t) − gε(x, t)) ≤ 0 i.e. ∀(x, t) ∈ Sn × [0, T ], f ε(x, t) ≤ gε(x, t) Corollary 5.3.10. As a consequence of the above proposition, we have that if f, g are a subsolution and supersolution respectively of ∂tu = J(u), and f ε, gε are the sup-convolution 104 and inf-convolution respectively, then sup (x,t)∈Sn×[0,T ] [f ε(x, t) − gε(x, t)] ≤ sup x∈Sn [f ε(x, 0) − gε(x, 0)] (5.3.6) Proof. We assume that supx∈Sn (f ε(x, 0) − gε(x, 0)) = cε > 0. Now let ˜wε(x, t) = f ε(x, t) − gε(x, t) − cε Since cε is the supremum of f ε − gε at 0, we obviously have ˜wε(x, 0) ≤ 0, and now we are back to the case in the previous proposition. Using the same argument as before leads us to the conclusion that sup (x,t)∈Sn×[0,T ] [f ε(x, t) − gε(x, t) − cε] ≤ 0 i.e. sup (x,t)∈Sn×[0,T ] [f ε(x, t) − gε(x, t)] ≤ sup x∈Sn [f ε(x, 0) − gε(x, 0)] Remark 5.3.11. We want to be able to make similar statements like (5.3.5), (5.3.6) about f, g. However, though we know that f and g are LSC and USC respectively, we do not know about the regularity of these functions yet. Thus if in (5.3.6) we take the limit as ε → 0, the limit commutes with the supremum on the left (since it is the supremum over all (x, t) ∈ Sn × [0, T ]) but not on the right because we are on the boundary. Thus we get sup (x,t)∈Sn×[0,T ] [f (x, t) − g(x, t)] ≤ sup x∈Sn [f ε(x, 0) − gε(x, 0)] , 105 when what we really want is sup (x,t)∈Sn×[0,T ] [f (x, t) − g(x, t)] ≤ sup x∈Sn [f (x, 0) − g(x, 0)] , In order to take limits on the right of (5.3.6) we will need the extra assumption in the following proposition. Proposition 5.3.12. Let f and g be a subsolution and supersolution of (5.3.1) respectively, and assume f (−, 0) ≤ g(−, 0) in a somewhat uniform way, i.e. ∀ε > 0,∃ δ > 0 s.t. f (x, t) ≤ g(y, s) + ε ∀|x − y| < δ,|t|,|s| < δ. (5.3.7) Then lim ε→0 sup x∈Sn f ε(x, 0) − gε(x, 0) ≤ 0. Proof. Let ρ > 0. Then we will show that lim sup ε→0 sup x∈Sn f ε(x, 0) − gε(x, 0) ≤ ρ By the definitiosn of f ε, gε, we know that ∃ , xε, tεxε, tε such that f ε(x, 0) = f (xε, tε) + gε(x, 0) = g(xε, tε) − (cid:18) 1 (cid:18) 1 2ε 2ε (cid:19) (cid:19) |xε − x|2 + |xε − x|2 + 1 2ε 1 2ε |tε|2 |tε|2 Since f ε → f, gε → g as ε → 0, this means that entire term in parentheses above → 0. Also, xε, xε → x and tε, tε → 0. Now for ε = ρ/3, ∃δ > 0 so that (5.3.7) holds. 106 Now we can choose ε small, and |xε − x|, |xε − x|, |tε|, |tε| are all < δ. Thus we get f ε(x, 0) ≤ f (xε, tε) + ρ/3 ≤ g(xε, tε) + ρ/3 + ρ/3 ≤ gε(x, 0) + ρ/3 + ρ/3 + ρ/3. (as f ε → f ) (because of (5.3.7)) (as gε → g) Putting together all the above inequalities, we can take the supremum over x ∈ Sn and then the limit supremum as ε → 0 to get lim sup ε→0 sup x∈Sn [f ε(x, 0) − gε(x, 0)] ≤ ρ. Since ρ > 0 was arbitrary, we have f ε(x, 0) − gε(x, 0) ≤ 0. lim ε→0 sup x∈Sn Theorem 5.3.13. Let f, g be a subsolution and supersolution respectively of (5.3.1), and assume f (x, 0) ≤ g(x, 0) on Sn in a somewhat uniform way, i.e. suppose (5.3.7) holds. Then f (x, t) ≤ g(x, t) ∀(x, t) ∈ Sn × [0, T ]. The proof of the theorem follows easily if you consider remark 5.3.11, thus by putting together the results from the previous two propositions. Remark 5.3.14. If we need to start the previous theorem with just the assumption that f (x, 0) ≤ g(x, 0) ∀ x ∈ Sn, then we need to additionally assume that f, g are bounded and uniformly continuous so that (5.3.7) holds. 107 5.3.3. Existence (Perron’s method) The proof for existence will follow just like in [34] (which comes from ideas in [26], [15]), with a minor modification just like in [9]. In the Perron method, the solution is given by u(x) = supv∈Sϕ v(x), where Sϕ consists of all subsolutions. This u will be a subsolution in the interior, it may not match the values at the boundary, i.e. u(y) = ϕ(y),∀y ∈ ∂Ω. Definition 5.3.15. If ∃ y ∈ ∂Ω such that there is supersolution wy and wy(y) = 0 whereas wy(x) > 0 ∀x (cid:54)= y, then such y are called regular points, and the function wy is called a barrier function. At such points we also get u(x) → ϕ(y) as x → y. We will sketch the proof for existence for (5.3.1). But first, we will list some results and definitions from [26]. Note that we will always have f, g ∈ C(Sn × (0, T ]). Proposition 5.3.16. [26, Proposition 2.3] If f is a subsolution and g is a supersolution of (5.3.1), f ≤ g on Sn×(0, T ], then ∃ a viscosity solution h such that f ≤ h ≤ g on Sn×(0, T ]. Next, we will give the definitions of the relaxed supremum/infimum of a function. Definition 5.3.17. [26, Section 2] Let Ω ⊂ Rn be open. Then for any function f : Ω → R we can define f (cid:63)(X) = lim sup r↓0 {f (Z) : Z ∈ B(X, r)} for X ∈ Ω f(cid:63)(X) = lim inf r↓0 {f (Z) : Z ∈ B(X, r)} for X ∈ Ω These functions are upper semicontinuous and lower semicontinuous respectively, and we can see that f(cid:63) ≤ f ≤ f (cid:63). 108 Theorem 5.3.18. [26, Theorem 3.1] Suppose f, g are respectively viscosity subsolution and supersolution to (5.3.1), and we have f (cid:63)(X) ≤ g(cid:63)(X) for X ∈ Sn. Then we will also have f (cid:63) ≤ g(cid:63) on Sn × (0, T ]. Remark 5.3.19. Under the assumptions of this theorem, there is a modulus on continuity, ω0, such that f (cid:63)(x) − f(cid:63)(z) ≤ ω0(|x − z|) for x, z ∈ Sn. We can also now find a modulus of continuity ω so that this is true for X = (x, t), Z = (z, s) ∈ Sn × (0, T ]. Theorem 5.3.20. [26, Theorem 3.2] If f is a viscosity subsolution, and g is a viscosity supersolution to (5.3.1), such that f ≤ g in Sn × (0, T ] and f = g on Sn, then ∃ viscosity solution h ∈ C(Sn × (0, T ]) such that f ≤ h ≤ g in Sn × (0, T ]. We can now sketch the proof of existence using the following ideas from [34]. Proof. Let f0 be a uniformly continuous function in Sn. We will prove that there exists a continuous function f : Sn × [0, T ] → R that solves (5.3.1) and f (−, 0) = f0. Perron’s method gives us a continuous viscosity solution in the interior by taking a relaxed infimum over the family of supersolutions of the equation. Now we need to show that this infimum is continuous at t = 0 and the initial condition f (−, 0) = f0 is satisfied. Assume f0 : Sn → R is uniformly continuous and f0 ≥ δ. Let F be the set of all supersolutions h such that there is some modulus of continuity ω so that for every x, z ∈ Sn and 0 < s ≤ T h(z, s) > f0(x) − ω(|z − x| + s) Suppose b is a smooth bump function such that b(0) = 1, b ≤ 1, and supp b = B1. For an arbitrary x0 ∈ Sn, let bδ(x) = b((x − x0)/δ) Then depending on the modulus of continuity 109 for f0, ∀ε > 0, ∃δ > 0 so that U0(x) := bδ(x)f0(x0) + (1 − bδ(x)) sup Sn f0 − ε ≤ f0(x) L0(x) : bδ(x)f0(x0) + (1 − bδ(x)) infSn f0 + ε ≥ f0(x) Thus, near x0, f0 is trapped between U0 and L0 with a gap of width 2ε. Since U0 and L0 are smooth functions (because b is smooth), |∇U0| and |∇L0| are both bounded by some constant C. Thus we can construct a supersolution and subsolution respectively by U (x, t) := U0(x) + Ct L(x, t) := L0(x) − Ct Going back to the equation (5.3.1), J in our case will only include terms like I(f ), G(I(f )), which are Lipschitz. We also have J being Lipschitz regardless from Assumption 5.2.1. Moreover J(U0), J(L0) have min-max representations consisting of integro-differential terms which will be controlled by C (from the estimates for µ in Theorem 5.2.3). Thus, we have Ut − J(U ) = C − J(U0) ≥ 0 Lt − J(L) = −C − J(L0) ≤ 0. (we can choose C thus) So U and h ∈ F are supersolutions and L is a subsolution to (5.3.1), and they are uniformly continuous functions. L(x, t) = L0(x) − Ct ≤ f0(x) − Ct ≤ h(z, s) + ω(|x − z| + s) − Ct 110 We can choose δ small enough so that |x − z| < δ, |s|,|t| < δ =⇒ ω(|x − z| + s) − Ct < ε. The comparison principle in proposition 5.3.12 gives L ≤ h in Sn×[0, T ] for all lower barriers L, which means F is bounded below. Moreover, F (cid:54)= ∅ because every upper barrier U ∈ F. Indeed we have already said why U is a supersolution, but also U (z, s) = U0(z) + Cs ≥ f0(x) + f0(z) − f0(x) + Cs = f0(x) − ω(|x − z| + s) F is a non-empty set that is bounded below, so we can find an infimum. In fact, here we use the relaxed infimum h(cid:63), i.e. h(cid:63)(x, t) = lim inf r→0 inf |x−z| 0, we can define these functions U, L, and conse- quently h(cid:63). We have already seen that ∀h ∈ F, ∀ lower barriers L L(x, t) ≤ h(z, s) + ω(|x − z| + s) − Ct = h(z, s) + ω(|x − z| + |s − t|) =⇒ L(x, t) ≤ lim inf r→0 = lim inf r→0 inf |x−z| 0, we have trapped a solution h(cid:63) in between a subsolution and a supersolution. Since we can make this ε as small as possible, this means that h(cid:63) is uniformly continuous and we have h(cid:63)(x0, 0) = f0(x0). Since x0 was arbitrary, we have h(cid:63)(−, 0) = f0. This proves part (c) of the main result, i.e. theorem 5.1.1. 5.4. Revisiting the main results So far, we have proved numerous properties for the broad class of operators J which satisfy the conditions in assumption 5.2.1, and then proved uniqueness for the viscosity solution of the equation (5.1.2), wherein such an operator J makes an appearance. As mentioned in the introduction in chapter 4, the reason for studying these operators and equations is that there is a strong connection between the solutions to the parabolic equations of the type in (4.1.5) and the free boundary problem of the type in (4.1.4), and this connection is explained in [9, Theorems 1.1, 1.4] which were mentioned earlier in section 4.3. The modified versions of these results for the sphere were given in section 5.1.1. 112 In particular, we are interested to see if these results apply to the case of the free boundary problem (5.1.1) and parabolic equation (5.1.2) introduced in section 5.1.1. Recall that the parabolic equation in question is ∂tf =(cid:0)|∂+ ν U| − |∂− ν U|(cid:1)(cid:113) 1 + |∇f|2 on Sn × [0, T ] suggesting that the corresponding operator J in this situation for which we proved all the results in sections 5.2 and 5.3 is given by J(f, x) =(cid:0)|∂+ ν U (x, f (x))| − |∂− ν U (x, f (x))|(cid:1)(cid:113) 1 + |∇f (x)|2 (5.4.1) But of course, inorder to apply all of the results in sections 5.2 and 5.3 to this operator J and subsequently to the equations (5.1.1) and (5.1.2), we need to first make sure that the chosen operator J from (5.4.1) satisfies all the properties listed in assumption 5.2.1. We check these properties below. Indeed, if we fix 0 < γ < 1, and if f ∈ (cid:32)(cid:83) , (cid:83) (cid:33) K(γ, δ, m) , then f ∈ C1,γ(Sn) and δ>0 m>δ δ0 < f < M for some δ0, M > 0. U and f are at least C1,γ ( [21, Section 8.11]), therefore their first derivatives are H¨older continuous, and thus J(f ) ∈ C0(Sn). Furthermore, in [21, Theorem 8.33] we see that ||U||C1,γ depends in a linear way on the C1,γ-norm of the boundary of the domain, which in our case is given by f . This implies that there is a Lipschitz dependence from f → ∂νUf . This and due to the fact that the mapping f (cid:55)→(cid:112)1 + |∇f (x)|2 is bounded and Lipschitz on each of the sets K(γ, δ, m, L) (where L is the upper bound for f : Sn → R as described in (5.1.4)), we can see that J is a Lipschitz operator. Further, since all terms in J are rotationally invariant, so is J, and we can also see that the rest of the 113 assumptions work out easily as in [9, Theorem 7.10]. In the following sections, we list some results that follow analogously from the results in [9, Sections 9,10,11], which in turn are useful for proving the main result, theorem 5.1.1. We do not provide any full proofs for the results as they can all be easily adapted from the proofs in [9], but we will mention some ideas behind the proofs and leave some quick remarks where relevant. 5.4.1. Different notions of viscosity solutions We will elaborate a little more on the topic of the equivalence between solving the parabolic equations of the type in (5.1.2) and the free boundary problem of the type in (5.1.1), but in order to do so, we will need a new notion of viscosity solution. In this notion, we look at test interfaces in lieu of test functions. We shall henceforth denote the set Σ0 = Sn × [δ, L]. A test interface S in [a, b] where 0 < a ≤ b < T will be described as follows. Definition 5.4.1. A test interface is a hypersurface S ⊂ Σ0×[a, b], such that each time slice S(t) is a positive distance away from ∂Σ0 and it separates Σ0 into two connected components, say S(t)+ and S(t)−, i.e. we will have Σ0 \ S(t) = S+ ∪ S−. Recall the definitions of D± f from (5.1.1), and we will also have in the two phase case f ⊂ S(t)+ and S(t)− ⊂ D− D+ f . Now, given a test interface S in [a, b] we will define the function US : Σ0 × [a, b] → R as the function which for each fixed time t ∈ [a.b] is the unique solution to the Dirichlet problem 114 ∆US = 0 US = 0 in S± on S US = 1 US = −1 on Bδ(0) on BL(0) (5.4.2) Below we will give the definitions of the classical subsolution (and supersolution), as well as the original and new notions of viscosity solutions. Definition 5.4.2. A function U : Σ0 × [a, b] → R is said to be a classical subsolution (respectively supersolution) of (5.1.1) if • The set ∂{U > 0} is a differentiable submanifold of Σ0 × [a, b] with codimension 1, and each time-slice ∂{U > 0} ∩ Σ0 × {t} (a ≤ t ≤ b) is a codimension 1 differentiable submanifold of Σ0. Also, U is twice differentiable in space and differentiable in time in Σ0 × [a, b] \ {U (cid:54)= 0}. • For each fixed t, the function U (·, t) solves, in the viscosity sense ∆U ≤ 0 in {U > 0} and {U < 0} (respectively ≥) and if V denotes the normal velocity of ∂{U > 0} (in the outer normal direction) then V ≤ |∂+ ν U| − |∂− ν U| (respectively ≥) U is a classical solution of (5.1.1) if it is both a supersolution and a subsolution. Now we move on to the definition of a viscosity solution. 115 Definition 5.4.3. A function U is said to be touched from above at (X0, t0) by S if S is a test interface in [t0 − τ, t0] for some τ > 0 such that {U > 0} ∩ {t0 − τ ≤ t ≤ t0} ⊂ S+ and (X0, t0) ∈ ∂{U > 0} ∩ S Definition 5.4.4. A viscosity subsolution (respectively supersolution) of the two-phase equa- tion (5.1.1) is an upper semicontinuous function (respectively lower semicontinous function) U : Σ0 × [a, b] → R, and it is required to have the following properties. U ≤ 1 (respectively U ≥ 1) on Bδ(0), U ≤ −1 (respectively U ≥ −1) on BL(0), And it satisfies the following relations in the viscosity sense −∆U ≥ 0 in {U > 0}◦ and {U < 0}◦ (respectively ≤), and for any test interface S touching U from above at (X0, t0) ∈ ∂{U > 0} we have VS(X0, t0) ≤ |∂+ ν Us| − |∂− ν Us| (respectively ≥), (5.4.3) where Us is as given in equation (5.4.2). We will use the above definitions 5.4.2, 5.4.3, and 5.4.4 for classical and viscosity solutions of (5.1.1), and in [9], it was shown that this is equivalent to the definitions given in [27] 116 5.4.2. Vertical shifts in the intersurface It will be necessatry to understand how ∂U± ν US varies with the vertical shifts of S, and thus we shall introduce some notation for such shifts. Given a test interface S, we define the intersurface Sh, which will result from shifting S in the upward direction by h as follows Sh := {X = (x, xn+1)|(x, xn+1 − h) ∈ S}, provided h > 0; if h < 0, then Sh results from shifting S down by |h|. The resulting variation in ∂+ ν US is recorded below. Lemma 5.4.5. Let S be a test interface. Then there is a constant C = C(S) such that for all sufficiently small h and any (X, t) ∈ Sh, we have |∂+ ν USh (X − hen+1, t) − ∂+ ν US(X, t)| ≤ C|h|, where both US and USh are functions given by (5.4.2). The proof of the above lemma involves looking at a function ˜U (X, t) := USh (X − hen+1, t), defined in S+ ∩ {(X, t)|X = (x, xn+1), xn+1 ≥ h}, and then applying the maximum principle comparing ˜U and US in S+ \ {0 ≤ xn+1 ≤ h}. One can use the above estimate for the one-phase problem to then obtain the bound for the two-phase problem relevant to our equation, i.e. (cid:12)(cid:12)(cid:12)|∇U + Sh | − |∇U− Sh | − |∇U + S | + |∇U− S |(cid:12)(cid:12)(cid:12) ≤ CS|h|. 117 This can be done using a similar approach to the one in [9, Lemma 6.1]. 5.4.3. Correspondence between viscosity solutions of the free boundary evolution and viscosity solutions of the parabolic equation In this subsection, we state the results that ultimately help establish the equivalence between the viscosity solutions of the free boundary evolution in (5.1.1) and viscosity solutions of the parabolic equation given by (5.1.2). These results will show that a viscosity solution to the parabolic equation will yield a viscosity solution to the free boundary problem (in the sense of definition 5.4.4). Remark 5.4.6. We also note that we can look at these equations in the context of the level-set equations. In this case, we look at the set ∂{U > 0} as the zero level set of the function U , or we can also consider it the domain and the boundary as a function of time, i.r. ∂{U (·, t) > 0} = ∂Ω(t). Assume that we parametrize t by some defining function Φ(X, t), then we have ∂Ω(t) = {Φ(·, t) = 0}. Now, if the normal velocity is given by V and the outer normal vector is ν, then the level-set equation becomes ∇Φ · (νV ) + ∂tΦ. Further, as Φ is the defining function of a level set, we have ν = ± ∇Φ |∇Φ| this reduces the flow to ∂tΦ = ±V |∇Φ|. To be depending on whether Φ > 0 or Φ < 0. consistent with the parametrization, we choose Φ > 0, and also Φ(X, t) = f (x, t) − xn+1. 118 With this and using the normal velocity condition given in (5.1.1), we finally obtain (5.1.2), i.e. ∂tf =(cid:0)|∂+ ν U| − |∂− ν U|(cid:1)(cid:113) 1 + |∇f|2 on Sn × [0, T ] Lemma 5.4.7. [9, Lemma 9.9] If f is a globally Lipschitz viscosity subsolution (respectively supersolution) of (5.1.2) in Sn × [a, b], then Uf is a viscosity subsolution (respectively su- persolution) of the free boundary evolution(5.1.1) in Σ0 × [a, b] (in the sense of definition 5.4.4). For the proof of this lemma (see [9, Lemma 9.9] for full proof), we start with the Γf being the graph of f and also the set ∂{Uf > 0}, and a test interface S touching Uf from above at a point, and we try to find an intermediate test interface between S and Γf which is also the graph of some function that lies between Uf and US. This is possible because of the Lipschitz nature of f and constructing a function using sup-convolutions. The intermediate function, which touches the graph of f at the same point as the test interface touches the level set of U , will satisfy the equation (5.1.2) by the definition of viscosity solutions, and we get many more inequalities using the comparison principle, which leads us to Uf being the viscosity subsolution (respectively supersolution) of the free boundary evolution (5.1.1). Lemma 5.4.8. [9, Lemma 9.10] If f is a viscosity subsolution (respectively supersolution) of (5.1.2) in Sn × [a, b], then Uf is a viscosity subsolution (respectively supersolution) of the free boundary evolution(5.1.1) in Σ0 × [a, b]. Note that this lemma does not have the extra assumption of f being globally Lipschitz as the previous lemma 5.4.7. To prove this lemma, we look at the sup-convolution of f , which is in fact a Lipschitz function, and apply lemma 5.4.7. This will tell us that the sup- convolution of Uf is a viscosity subsolution of (5.1.1), and we can now make some limiting 119 arguments to show that Uf is a viscosity subsolution to (5.1.1). Now the next result will tell us how the test functions for the lower dimensional, non-local parabolic equations will yield to test functions of the free boundary problem. Proposition 5.4.9. [9, Proposition 9.11]Let φ be an admissible test function touching f from above at some t0 ∈ [a, b] and x0 ∈ Sn. Then Uφ touches Uf from above at (X0, t0) where X0 = (x0, f (x0, t0)) The proof of the above proposition is a straightforward consequence of the comparison principle and can be found in [9, Lemma 9.11]. Lemma 5.4.10. [9, Lemma 9.12] Let U be a viscosity subsolution (respectively supersolution) of (5.1.1) in Σ0 × [a, b] whose free boundary is given as the graph of some upper semicon- tinuous (lower semicontinuous) function f , then f is a viscosity subsolution (respectively supersolution) of (5.1.2) in Sn × [a, b]. Using the previous proposition and the level set equations for (5.1.1) (see remark 5.4.6), we can prove this lemma. Remark 5.4.11. All the given results in this section 5.4.3 will give the proof for part (a) of the main result, i.e. theorem 5.1.1. 5.4.4. Propogation of the modulus of continuity In this final section, we will give an explanation for the remaining parts of the main result, theorem 5.1.1, which pertain to the modulus of continuity. In particular, we can show that the modulus of continuity of the initial data will be preserved by the fractional parabolic equation. This result follows without difficulty once the comparison theorem for viscosity solutions has been established, which we have done in section 5.3. 120 Lemma 5.4.12. Let J be as in assumption 5.2.1, T > 0 and f : Sn × [0, T ] → R, a continuous viscosity solution of ∂tf = J(f ) f (x, 0) = f0(x) in Sn × [0, T ] in Sn If f0 is continuous with the modulus of continuity ω(·), then the same will be true of f (·, t) for all t ∈ [0, T ]. In particular, we have the estimate |f (x, t) − f (z, t)| ≤ ω(|x − z|). Proof. Let z ∈ Sn be fixed, and consider the function w(x, t) := (Rzf )(x, t) − f (x, t) = f (Rzx, t) − f (x, t). Then by assumption, w(x, 0) ≤ ω(||Rz||), ∀x ∈ Sn Now, f being a viscosity solution is a subsolution as well as a supersolution, and since J is rotationally invariant, we can also say that Rzf is a viscosity solution. Thus by an argument similar to the one in the proof of proposition 5.3.9, we get ∂tw ≤ M + J (w). 121 Now w satisfies all the assumptions of lemma 5.3.8, thus we have w+ ≤ sup Sn w+(·, 0) sup Sn×[0,T ] i.e. w(x, t) ≤ ω(||Rz||), ∀(x, t) ∈ Sn × [0, T ] =⇒ f (Rzx, t) − f (x, t) ≤ ω(||Rz||). Since z ∈ Sn was arbitrary, we have proved the lemma. Note that the above lemma is the proof of part (b) of the main result, i.e. theorem 5.1.1. 122 REFERENCES 123 REFERENCES [1] I. Athanasopoulos, L. Caffarelli, and S. Salsa. Regularity of the free boundary in parabolic phase-transition problems. Acta Math., 176(2):245–282, 1996. [2] Jean-Michel Bony, Philippe Courr`ege, and Pierre Priouret. Sur la forme int´egro- diff´erentielle du g´en´erateur infinit´esimal d’un semi-groupe de Feller sur une vari´et´e diff´erentiable. C. R. Acad. Sci. Paris S´er. A-B, 263:A207–A210, 1966. [3] L. Caffarelli, E. Fabes, S. Mortola, and S. Salsa. Boundary behavior of nonnegative solutions of elliptic operators in divergence form. Indiana Univ. Math. J., 30(4):621– 640, 1981. [4] Luis Caffarelli and Luis Silvestre. An extension problem related to the fractional Lapla- cian. Comm. Partial Differential Equations, 32(7-9):1245–1260, 2007. [5] Luis A. Caffarelli. A Harnack inequality approach to the regularity of free boundaries. III. Existence theory, compactness, and dependence on X. Ann. Scuola Norm. Sup. Pisa Cl. Sci. (4), 15(4):583–602 (1989), 1988. [6] H´ector Chang Lara and Gonzalo D´avila. Regularity for solutions of non local parabolic equations. Calc. Var. Partial Differential Equations, 49(1-2):139–172, 2014. [7] H´ector Chang-Lara and Gonzalo D´avila. Regularity for solutions of nonlocal parabolic equations II. J. Differential Equations, 256(1):130–156, 2014. [8] H´ector A. Chang-Lara and Gonzalo D´avila. H¨older estimates for non-local parabolic equations with critical drift. J. Differential Equations, 260(5):4237–4284, 2016. [9] H´ector A. Chang-Lara, Nestor Guillen, and Russell W. Schwab. Some free boundary problems recast as nonlocal parabolic equations. Nonlinear Anal., 189:11538, 60, 2019. [10] H´ector A Chang-Lara and Dennis Kriventsov. Further time regularity for nonlocal, fully nonlinear parabolic equations. Communications on Pure and Applied Mathematics, 70(5):950–977, 2017. [11] Sunhi Choi, David Jerison, and Inwon Kim. Regularity for the one-phase Hele-Shaw problem from a Lipschitz initial surface. Amer. J. Math., 129(2):527–582, 2007. [12] Sunhi Choi, David Jerison, and Inwon Kim. Local regularization of the one-phase Hele- Shaw flow. Indiana Univ. Math. J., 58(6):2765–2804, 2009. [13] F. H. Clarke. Optimization and nonsmooth analysis, volume 5 of Classics in Applied Mathematics. Society for Industrial and Applied Mathematics (SIAM), Philadelphia, PA, second edition, 1990. 124 [14] Philippe Courrege. Sur la forme int´egro-diff´erentielle des op´erateurs de c∞ k dans c satis- faisant au principe du maximum. S´eminaire Brelot-Choquet-Deny. Th´eorie du Potentiel, 10(1):1–38, 1965. [15] Michael G. Crandall, Hitoshi Ishii, and Pierre-Louis Lions. User’s guide to viscosity solutions of second order partial differential equations. Bull. Amer. Math. Soc. (N.S.), 27(1):1–67, 1992. [16] Daniela De Silva, Fausto Ferrari, and Sandro Salsa. Two-phase problems with dis- tributed sources: regularity of the free boundary. Anal. PDE, 7(2):267–310, 2014. [17] Daniela De Silva, Fausto Ferrari, and Sandro Salsa. Perron’s solutions for two-phase free boundary problems with distributed sources. Nonlinear Anal., 121:382–402, 2015. [18] Joachim Escher and Gieri Simonett. Classical solutions of multidimensional Hele-Shaw models. SIAM J. Math. Anal., 28(5):1028–1047, 1997. [19] E. Fabes, D. Jerison, and C. Kenig. The Wiener test for degenerate elliptic equations. Ann. Inst. Fourier (Grenoble), 32(3):vi, 151–182, 1982. [20] E. B. Fabes, C. E. Kenig, and D. Jerison. Boundary behavior of solutions to degenerate elliptic equations. In Conference on harmonic analysis in honor of Antoni Zygmund, Vol. I, II (Chicago, Ill., 1981), Wadsworth Math. Ser., pages 577–589. Wadsworth, Belmont, CA, 1983. [21] David Gilbarg and Neil S. Trudinger. Elliptic partial differential equations of second order. Classics in Mathematics. Springer-Verlag, Berlin, 2001. Reprint of the 1998 edition. [22] Michael Gr¨uter and Kjell-Ove Widman. The Green function for uniformly elliptic equa- tions. Manuscripta Math., 37(3):303–342, 1982. [23] Nestor Guillen, Jun Kitagawa, and Russell W Schwab. Estimates for dirichlet-to- neumann maps as integro-differential operators. Potential Analysis, pages 1–39, 2019. [24] Nestor Guillen and Russell W. Schwab. Min-max formulas for nonlocal elliptic operators. Calc. Var. Partial Differential Equations, 58(6):Art. 209, 79, 2019. [25] Nestor Guillen and Russell W. Schwab. Min–Max formulas for nonlocal elliptic operators on Euclidean Space. Nonlinear Anal., 193:111468, 2020. [26] Hitoshi Ishii. On uniqueness and existence of viscosity solutions of fully nonlinear second-order elliptic PDEs. Comm. Pure Appl. Math., 42(1):15–45, 1989. [27] Inwon C. Kim. Uniqueness and existence results on the Hele-Shaw and the Stefan problems. Arch. Ration. Mech. Anal., 168(4):299–328, 2003. [28] Inwon C. Kim and Norbert Poˇz´ar. Viscosity solutions for the two-phase Stefan problem. Comm. Partial Differential Equations, 36(1):42–66, 2011. 125 [29] W. Littman, G. Stampacchia, and H. F. Weinberger. Regular points for elliptic equations with discontinuous coefficients. Ann. Scuola Norm. Sup. Pisa (3), 17:43–77, 1963. [30] Antoine Mellet, Benoˆıt Perthame, and Fernando Quir´os. A Hele-Shaw problem for tumor growth. J. Funct. Anal., 273(10):3061–3093, 2017. [31] Ren´e L. Schilling, Renming Song, and Zoran Vondraˇcek. Bernstein functions, volume 37 of De Gruyter Studies in Mathematics. Walter de Gruyter & Co., Berlin, 2010. Theory and applications. [32] Russell W. Schwab and Luis Silvestre. Regularity for parabolic integro-differential equa- tions with very irregular kernels. Anal. PDE, 9(3):727–772, 2016. [33] Joaquim Serra. Cσ regularity for concave nonlocal fully nonlinear elliptic equations with rough kernels. arXiv preprint arXiv:1405.0930, 2014. [34] Luis Silvestre. On the differentiability of the solution to the Hamilton-Jacobi equation with critical fractional diffusion. Adv. Math., 226(2):2020–2039, 2011. [35] Luis Silvestre. Regularity estimates for parabolic integro-differential equations and ap- plications. In Proceedings of the International Congress of Mathematicians—Seoul 2014. Vol. III, pages 873–894. Kyung Moon Sa, Seoul, 2014. [36] Renming Song and Zoran Vondraˇcek. Potential theory of subordinate killed Brownian motion in a domain. Probab. Theory Related Fields, 125(4):578–592, 2003. [37] Kjell-Ove Widman. Inequalities for the Green function and boundary continuity of the gradient of solutions of elliptic differential equations. Math. Scand., 21:17–37 (1968), 1967. 126