“H'EEHS J 332‘s w- H:- ' £7; 1;; :' ff;- , '(‘- ":3 ‘ '3‘ '1 - . "“‘3- ,1) . Ih (10,): A" 1 .Z ]'-'n if" .J.‘~»I'A "Fm ’g ‘ I.» s; v‘ . - _ _. ., . P. n ! mafiréfl“ ‘-‘ :41" ‘--'%-..'.3. 5.5" $65 a an. L Uaai'vwg‘iay __I‘ MSU LIBRARIES RETURNING MATERIALS: P1ace in book drop to remove this checkout from your record. FINES will be charged if book is returned after the date stamped be1ow. UNFOLDING OF A CLASS OF SINGULAR FREE BOUNDARIES FOR THE FOUR DIMENSIONAL AXI-SYMMETRIC OBSTACLE PROBLEM BY Ahmad Feyzi Dizaji A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Mathematics 1983 / GLVQOS ABSTRACT UNFOLDINGS OF A CLASS OF SINGULAR FREE BOUNDARIES FOR THE FOUR DIMENSIONAL AXI—SYMMETRIC OBSTACLE PROBLEM BY Ahmad Feyzi Dizaji In this thesis we shall consider an elliptic free boundary value problem, We shall deal with the four dimensional aXi—symmetric obstacle problem and shall consider a class of singularities of free boundary. Specifically, we study generic perturbations of the singularity x = H(y)-—% y log y, where H(t) is a real analytic function about t = O with H(O) = O and H'(O) # 0 near the origin. This locus described by this formula represents the boundary of the region of contact between a membrane and a smooth rigid obstacle. The point of view taken is that of generic bifurcation, where one scalar parameter (2K, the height of the membrane above the obstacle at the origin) and one functional parameter (¢) is present. Our prime interest is a description of the unfolding of such singularities, their normal forms, and generic conditions for unfoldings. ACKNOWLEDGMENTS I want to express my deep gratitude to Professor Shui—Nee Chow and Professor John Mallet-Paret, my research advisors, for all their patience, expert guidance, and encouragement during the course of my research. I am especially indebted to Professor Mallet—Paret for his stimulating discussions and assistance during the preparation of this thesis. I am also grateful to my wife, Nasrin Habibi, for her patience and cooperation displayed throughout my graduate studies. I am grateful to the members of my thesis committee as well as guidance committee. I extend my appreciation to all my instructors at Michigan State University, Tehran University, high school, elementary school, and to my parents from whom I learned a great deal of wisdom and faith. I also want to thank all the peOple who want to use the science for justice and peace only. Finally, I would like to thank Mrs, Cindy Lou Smith for her patient typing of this manuscript. TABLE OF CONTENTS Chapter 1. INTRODUCTION . . . . . . . . . . . . . . . The minimal surface problem . . . . . The obstacle problem . . . . . . Some examples . . . . . . . . . . . . . . . Local solution to the obstacle problem . FORMULATION OF A PROBLEM . . . . . . . . . Local solution to the axi- -symmetric obstacle problem in R4 . . . . . . . . A class of examples . . . . . . . . . . . . Perturbation of Example 2.5: Heuristics . Description of a problem . . . . . . o . . A NECESSARY AND SUFFICIENT CONDITION FOR A LOCAL SOLUTION . . . . . . o . . . . . . An equivalent definition of a local solution of axi-symmetric problem in R4 . Proposition 3.9° Hypothesis 3.7 holds if and only if Hypothesis 3.8 holds . . . . Theorem 3.11. (u,O) in Definition 3.10 is a local solution to the obstacle problem in the sense of Definition 3.1 if and only if . . . . . . . . . . . . . Proposition 3.12° The local solution (u,G) is symmetric with respect to y-aXiS if and only if a o o o o o a a o 0 AN ANALYTIC CONTINUATION TO THE FREE BOUNDARY . . . o . . o a o . . o . ._o . Theorem 4.6. (The main Theorem) . . . . 0 Remark 4.7. (free boundary equation) , 28 28 35 47 56 59 59 68 77 83 86 9O 91 Chapter 50 A CLASS OF UNFOLDING OF THE SINGULARITY Problem 5.1. . . . . . . . . . . . . . Necessary conditions in Problem 5.1 . Remark 5.20. (a list of necessary conditions for Problem 5.1) . . . Sufficient conditions in Problem 5.1 continue Conclusions . . . . . . . . . . . . . i v Theorem 5.30. (solving the Problem 5.1). 128 132 137 148 Figure Figure Figure Figure Figure Figure Figure Figure Figure Figure Figure Figure Figure Figure Figure ~Figure Figure Figure Figure LIST 0 o o o n o o o o e o o n o o a o o n a o o o o a o o o o a a a a OF FIGURES Figure Figure Figure Figure Figure .Figure Figure Figure Figure Figure Figure Figure Figure Figure Figure Figure Figure Figure Figure Figure Figure Figure Figure Figure Figure 5.9 5.10 5.11 5.12 5.13 5.14 5.15 5.16 5.17 5.18 5.19—a 5.19-b 5.20-a 5.20—b vi 61 64 89 93 95 96 98 100 104 105 107 109 109 111 113 115 122 123 126 130 134 141 142 145 145 GLOSSARY OF NOTATIONS Rn: Euclidean n-dimensional space, the product of n copies of the real line R. 0: An open, generally bounded and connected subset of Rn. an: The boundary of Q. Bu bu Bu) I [can] 8x1 5x2 axn partial derivatives. vu: The vector if u :Rn 4 R has n 2u n Au: Z) ——E if u :R 4 R =1 6X. Y J u ,u : The partial derivatives 22 . i3 if u :R2 a R. x y BX by Cé(Q): The set of continuous functions with compact support in Q. Cm’l(5): The functions m times continuously differentiable in 6 whose mth derivatives satisfy a Halder condition with exponent X. ‘ Cm(5): The functions m times continuously differentiable in 5 2,5 1 1 H (Q), H (0), HO(Q): The usual Sobolev spaces of completion of Cm(5), 02(6), CS(Q), respectively, each one with the corresponding norm. 1. INTRODUCTION The minimal surface problem 1.1. Consider the classical minimal surface problem of finding a surface in three—dimensional space, spanning a fixed boundary loop, in such a way as to minimize the area of the surface (Plateau's problem) [l8,16,6]. ‘ X 7 ./® Figure 1.1 In particular in the non-parametric minimal surface problem [18,6,8], we suppose the surface and its boundary are graphs of functions u and g defined in subsets of R2 as follows: Given 0 C R2 a bounded domain with smooth boundary BC, and g :53 4 R, find 1 u :5 4 R such that u = g on 30 and u minimizes the area integral H A(u) = H‘ (1+ \vu‘2)2 dxdy (1.1) 0 among such functions. Suppose this problem has a solution u, sufficiently smooth. Let us derive the Euler equation satisfied by this solution. Fix v E Cé(0)7 then for any t E R A(u4—tv) 2 A(u) , (u-+tv)| = g . 6. Hence _1 0=-d—A(u+tv)| =r'f (1+)vu12) 2 (u v +u v )dxdy dt 6:0 Juq xx yy (Fréchet derivative) Integrating by parts and assuming u twice differentiable in 0 we get uX u o = - P ————-————— + -————X———— vcb.)t2w+ (1 —>.tl— (l ->.)t2)v) ==AM(Hw+(1—tflv)+(l—k)fliw+(l-tflv) g AB(tl)4-(1-X)B(t2) for o.g i g_1. So B :R 4 R is a convex and therefore continuous function which is constant on [0,1]. Thus B'(t) = B”(t) = o for t 6 [0,1]. B(t) = HQ f(tvw+ (l-t)vv) dxdy , B'(t) = 55A f'(tvw4—(l-—t)vv) ~v(w-—v) O Q The proof in fact covers the higher dimensional case where 3 : Rn. It was shown in [l] with the above 1,1 conditions that u E C (3). 8 The most interesting questions here focus on the contact set and its boundary: I [(X,y) e 0 |u(X.y) = w(x,y)l . T BI = the free boundary . The boundary al of the region of contact is called a free boundary as it is not given apriori, but rather depends on the solution, that is, it is a part of the solution to be found. Observe that the derivation of (1.2) is still valid (at least formally) in the complement Q-—I of the contact set, namely where u > t. A fundamental result on the geometry of T, in the two dimensional cases was proved by Kinderlehrer [11]: if Q c R2 is strictly convex with smooth boundary, if g E O, and if N is strictly concave and analytic, then T is an analytic Jordan curve enclosing the contact set I. More generally if the smoothness of the obstacle C2,s — is relaxed to assume only N E (Q) (keeping all other assumptions), then T is a Jordan curve admitting Cl,t a parametrization for any t < 5 (see [8]). A main purpose of this thesis is to study different ~possible geometrical shapes T of free boundaries, for a related class of problems. In particular, more complicated shapes (cusps) will arise. We shall also study how the set T varies as the data of the problem (such as m and g) are allowed to change. 1.2.1. The Dirichlet Integral If g, N and its gradiant VN are sufficiently small, we might expect the solution u and its gradient also to be sufficiently small. If this is so, then using the approximation ./1+—x2 = 14“; x2+-0(x4), near x = 0, leads us to consider the Dirichlet integral J(u) = % j lVUl2 Q as an approximation for A(u)-A(O). We may therefore consider the obstacle problem obtained from the one above by minimizing J(u) instead of A(u). The Euler equation of J(u) is Laplace's equation vu = O, in contrast to the (nonlinear) minimal surface equation (1.2) associated with A(u). In general we may consider such a problem in a space of any dimension, as follows: Given 0 C Rn a bounded domain with smooth boundary 50, g :80 4 R sufficiently smooth H N :u 4 R sufficiently smooth with w < g on 80. find (*) u :5 a R, u = g on 50, u 2 N on 6, such that u minimizes J(u) = % j lvu|2 among those functions satisfying (*). 10 We stress that the relation between the minimal surface problem (which minimizes A(u)) and the above Dirichlet problem which minimizes J(u) (or energY) is purely formal: they are two different problems. We may consider the Dirichlet problem above as a model problem as it has many features in common with the minimal surface problem, yet is simpler in many respects. In particular the Euler equation vu = 0 associated with the functional J(u) is linear. In this thesis we study the Dirichlet integral. 1.3. Existepce, uniqueness, regularity More precisely, J(u) = % I [Vulzch<, X E Q C Rn, Q is minimized over the closed convex subset K of the Sobolev space H1(Q): r 1 K = (u E H (0) :u = g on 60, u 2 W a. e.} Since K is a convex closed set and J is a convex continuous functional, it is easy to show [8] u minimizes J over K e DJ(u)(v-—u) = JP. Vu°v(V~u) 20, VV 6 K , where DJ(u) denotes the Fréchet derivative. Therefore, an equivalent formulation of the problem seeks u E K so that 11 J0 vu .v(v-u)dx'2 0, VV 6 K This new formulation is called a variational inequality. If g and an are sufficiently smooth, then without loss of generality we can assume g e O in the problem as follows: simply replace u with v = u--ul and N with a = w-ul, where ul is the solution to the boundary value problem vu = 0 in Q, u l = g. 1 1 80 (It is known [12,6] that the solution u exists if g 1 and 60 are smooth enough). Since w E K implies w-u = 0 on 80. we have DJ(u)(w-—u) = f vu °v(w-u)dx = f v(u-—ul) ov(w.-u)dx Q L =]‘ v(u-ul) -v((W-ul)- (u-ul))dX o f vv -v(wl-v)dx , Q where w = (w-u 1 ) E K = [v E Hé(0) :v 2 a a.e.]. 1 1 Thus u solves the minimizing problem with data (g,¢) if and only if v solves it with (0,a). Remark: Unless otherwise stated, from now on we shall assume g E O on 50. # We may summarize the above results in the following precisely stated problems. 12 (1) Minimization problem: Given 1 E H (0), with N g 0 on 50, minimize J(u) pg; [vulzdx Q 1 . among u E HO(0) w1th u 2 N a.e. (2) Variational problem: Given w E Hl(0), with N.§ O on 30. find u E K 1 . K = [v E HO(0) :VIZ w a.e. 1n 0] so that f vu «u(v-u)dx 2 0, VV 6 K . Q Theorem 1.3.1. These two problems have unique solutions, and they are the same solutions. # For the proof and more detail in this context see [8]. A basic regularity result is the following (we state without proof): Theorem 1.3.2 (Brezis and Kinderlehrer [1]). If u E C2(5), and N < 0 on 50, then the unique solution ' u to the above problems satisfies: u E Cl(5) and Du is uniformly Lipschitz in a. # As before the contact set and free boundary 13 are compact subsets of 3 under the assumptions of this theorem. On Q-—I the derivation of the Euler equation is valid and so Au = 0 there. It is also the case that I E {x E Q :A¢ g 0] because Au 3 0 in Q (by applying the variational inequality, see [8,20]). Finally, because u is C1, it follows immediately that v(u-—¢) = O on T. See [8] for details. Conversely we have the following result (see [2]). Proposition 1.3.3. Suppose I C Q is a compact set with boundary T which is piecewise smooth in the following sense: there is a finite subset F E T such that if x E T-—F, then near x, T is an embedded Cn"l manifold. Suppose u E Cl(5) and satisfies: Au = O and u 2 W on 0-—I u = 0 on 50 u = N and AN 3 O on I Then u is the solution of the obstacle problem; Outline of Proof: We show 5 vu -v(v-—u)dx 2 0 Q 'for every v E K. It is enough to consider v E Cl(5) 0 K, as this set is dense in K. Let Br = U B(x,r), where B(x,r) is the r—ball about X. xEF - Let also E = 1-—I and set C = Q-—B , E = E-—B , r r r r and I = I—IB. r r 14 By Sard's Theorem [ 4], for almost every r, the (n-1)-manifold aBr and T-F intersect transversally; we choose r so that this is the case. We may apply Green's Theorem (since aBr and T-—F are transverse) to obtain j Vu -v(v-u)dx Q r [ vu-Wv—qu+j vw-WV—UNX Er Ir =§ % (v-u) + (BI g—r'l‘l (v-u) — j (v¢)(v—u)dx BEr r Ir As r 4 O, the integrals over Or and Ir tend to those over 0 and I. The integrals over arr and aIr can be divided into two parts: those over F-—Br and those over the spheres BBr. The integrals over F-—Br cancel as %% = %% there, and the integrals over aEr and aIr are taken with opposite orientation. The contribution of the integrals over aBr tends to zero as r 4 0, since the integrands are bounded and measure of aBr tends to zero. Hence we get the result. # Some Examples Example 1. n = 1: A string fixed at endpoints (a,0), (b,0), where Q = (alb). 15 Y u(x) \ w(x) or a b >X Figure 1.3 It is clear that on the intervals, where u(x) > u(x), we have u”(x) = 0. So u(x) is linear. It is also clear that u”(x) = w”(x) g 0 on intervals where u(x) = u(x). In general u E C2((a,b)) as u” typically has jump discontinuities: at the endpoints of the intervals where u(x) = u(x), u”(x) jumps from $"(x) to zero. See Figure 1.3. Example 2. n = 2: Suppose Q = [(x,y) E R2 |x2+-y2< l], g E O on an, and u(x,y) = -%(x24-y2)4-c, where O < c < %. So t < 0 on 80. The radial symmetry of the obstacle and the invariance of the Dirichlet integral -under rigid rotations implies the unique solution u(x) also is radially symmetric. 16 obstacle 2 2 » ~ .7 the free boundary Figure 1.4 As a candidate for the contact set, consider the disk I = {(X.y) e R2 |X2+y2 grg< 1] of radius r for some r . If this is the contact set, 0’ 0 then in Q-—I, u(x) satisfies the Dirichlet problem 2 2 2 Au = O for r0 3 x +—y g 1 u = 0 for x2+y2 = 1 l 2 2 _ u — c-§ rO for x 4-y — r and moreover Vu = vw on T, hence ill_ 2_ 2 Br — -rO for x + v — rO , (l) 17 “U . . . . . where g; 15 the outward radial der1vat1ve. The unique solution of this Dirichlet problem is (with r2 = x2+y2 ) u(x,y) = A log r, rO g r g 1, where _ .1. 2 A — (c--2 rO)/log r0 (2) In addition (1) holds also if and only if ii = -r0, or O A: a?) (3) There exists rO and A satisfying (2) and (3) if and only if r2 r2 _ O O c — -7f log (7;) It is an easy exercise to show that for any c E (O ,%), there is a unique rO E (0,1) satisfying this relation. We claim this unique rO and A = —r8 give the solution w(x,y) . o g r 3 r0 U(X.y) = A log r , r0 3 r g 1 of the problem. This follows from Preposition 1.3.3 , once we show u > ¢ on Q-—I, that is, f(r) = -r3 log r4—% r2-c > O, for r E (rO,l). But this is easy as f(ro) = O and f'(r) > 0 on (ro,1). 2 . Again note u is not c2, as .2_% undergoes a dr jump at r = r from 18 ——— = -l to Ji- (A log r) I = 1 . 2 dr dr r-rO The anolog of this problem in dimensions n 2 3 may be considered in a similar fashion. The contact set is again a disk of radius r0, In Q-—l, u has the form u = Ar where rO depends on n. —n+2__A° Example 3. Lewy and Stampacchia [12] show that if 0 C R2 is convex and N is strictly concave and analytic with N < 0 on 80, then T is an analytic Jordan curve, Q-—I is homeomorphic to an annulus, and I is simply connected. This result can be generalized for the case of obstacle u E C2(a). See [8]. Some additional properties of the solution 1. If N < O in 50 and At < 0 in Q, then Q-—I is connected. Proof: This can be shown by strong maximum and minimum principle: (Let Aw 2 0 (g 0) in A, a bounded domain, and suppose there exists a point y E A for which w(y) = sup w(inf w). Then w is constant in A.) A Suppose C-—I is not connected, then it has some component 6 whose boundary is part of the boundary of I. Then 11— t is zero on 56 and u-—¢ > 0, A(u-—$) > O in '3, which contradicts the strong maximum principle. 19 2. Under quite general conditions [3,9,10,12] F = bl consists of smoothly parametrized arcs, possibly with cusps (in fact analytically parametrized arcs if N is analytic). In particular if G is a neighborhood of 2 and (3 0 T is a Jordan are a p01nt (XO’YO) E Q C R through (xo,yo), and N is analytic, then the arc 0 0 F admits an analytic parametrization (possibly with cusps). See [8] for more details. 3. If )1 < $2, then 0 < u2--ul < SUP(N2'-Wl)- Proof: Let Kl = {w E Hé(0) :w 2 $1} and 1 . = r . . = K2 Kw E HO(0) .w 2 $2], and put u3 m1n(ul,u2), v = max(ul,u2). Then by the variational inequality formulation I vul -v(u3-ul)dx 2 O, f vu2 :v(v-u2)dx 2 O, Q Q fi ' ._ = _ where u3 - K1 and v E K2. Since v u2 (u 3 1)! thus fqvu2 ~v(v-—u2)dx 2 O a vauz «v(u3-ul)dx g 0, Suntracting the first inequality above from the last one yields f v(u2-ul) -v(u -ul)dx S O or 3 ") J{u 1] onto the plane cut along the real axis from —1 to 1. By choosing s > 0 small, we may arrange that, in either example 1 or example 2, B maps the circles [t] = l and [t] = 2 onto Tj and a Jordan curve Cj which contains Tj in its interior, respectively. Let Ij be the (closed) region inside Tj and let Qj be the (open) region inside Cj° Schaeffer shows that B maps the annulus A = [t :1 < [t] < 2] onto 0.-—I. in a one-to—one 3 J manner. He proves that with the obstacle 111(XIY): _% (X2+Y2) I The obstacle problem has its unique solution in Gj with contact set Ij and the free boundary Fj (for sufficiently small s > O). This is accomplished by considering an equivalent (non—singular) problem in the disk \tl < 2 via the mapping B. Example 5. In [20] Schaeffer gives an example with a C” obstacle for which I has infinitely many connected components in the neighborhood of some point. 23 Local solution to the obstacle problem 1.4. The obstacle problem discussed in last sections has a local formulation as a differential equation which we define it more precisely now. Definition 1.4.1. A local solution in an open set 0‘9 0 C 191 is a 01 function u :0 4 I? satisfying u(x) 2 u(x) on O Au(x) = 0 if u(x) > w(x) A¢(x) go if u(x) = w(x) . # If u is a local solution in G, then the contact set in 0 is the set I = [X E O |u(x) = u(x)], and the free boundary in O is F = BI n 0. Observe that wu-w)=0 on L Now assume in the neighborhood of some point, P is an (n-1)-dimensional manifold (arc if n = 2, surface if n = 3) which is analytic with I lying on one side of P. Then on a one sided neighborhood V of f, contained in O-—I, we have Au = O, u > w, and on T we have u = N, v(u-—¢) = 0. In fact the initial boundary value problem 24 has a unique solution u near T in O -1 by the Cauchy—Kowalewski Theorem. In fact an explicit formula for u can be given when n = 2 (see [20]) and we shall present the method later on. The higher dimensional cases are much more difficult. However, for n even, if the solution has enough symmetry, results can be obtained. We shall study the case n = 4 for a class of axisymmetric problems. Example 6. Kinderlehrer and Nirenberg [10,8] gave an example of a local solution in which the free boundary has a cusp and away from it the free boundary is analytic. They showed for m 2 1 odd the free boundary can never have a cusp of the form 2 2m+l (y-yy Km-XO) , I<#O near some (xo,yo) in two dimensional case (0 C R2), whereas for m even such a cusp can occur. (In fact, the case m = l was first noted by Schaeffer: Example 4). Their proof involves first straightening out the cusp to a line segment by means of a conformal mapping; then an analysis of several terms of the Taylor series of u near (xo,yo), based on the equation Au = O governing u in C-—I, gives a contradiction to u 2 w. 25 Further specialization We shall study local solutions in the neighborhood of a point, say the origin. We assume the nondegeneracy condition A¢(O) < 0 (if A¢(O) > 0, then T = ¢ near 0). Then, as we see from the Definition 1.4.1, upon adding the same harmonic function h to a local solution u and obstacle W and multiplying by a positive constant c we get another local solution u1 = c(u—+h) for the new obstacle W1 = c(¢+—h). In particular a quadratic polynomial can always be chosen for h(x) to give 1 ‘(J(X) = -§ IX]2+O(IX\3) . x E Rn . We choose u(x) = -% )X)2 again as a model problem near the origin and consider only this obstacle from now on. Deleting the higher-order terms in N should not affect the resulting theory significantly, although this has yet to be established. Changing Boundary conditions 222. Schaeffer [20,21] has studied how the set I changes as the data t and g vary. He proved a theorem that if g and g are C” with t < g on 50, T a Ca curve, and Au < 0 on I, then if ml, gl are sufficiently near 3, g in the C: topology, the free 7‘ boundary of .1 of the corresponding problem will be a 26 Cf curve near I in the C0° topology (in normal coordinates about T). 'f : 44:74 solution u 1, - , obstacle Figure 1.8 Schaeffer has also pointed out the need for a generic theory of such variations of I and BI. Mallet—Paret and Chow declared that such a theory presumably could take the form of a bifurcation theory or unfolding theory for the singularities of al. Such a theory was described in the case n = 2 [14,5,15]. A significant point here is that the unfoldings one encounters are not generic in the sense of singularity theory as developed by Thom Mather, Arnold and others; only very special types of singularities can occur. This is seen for example from the result of Kinderlehrer and Nirenberg [10,8] described above in Example 6. Mallet—Paret made a detailed study of a class of singularities of a free boundary, and their bifurcation, in the two dimensional case, as the data of the system varies parametrically. Typically cusps on the free boundary become smooth, and islands may appear as in Figure 1.9. 27 singular free boundary after perturbation Figure 1.9 In this thesis we shall study the unfoldings (bifurcations) of a class of singularities of the four dimensional axi-symmetric obstacle problem. This case is the next simplest after the two-dimensional case, as harmonic functions admit a particularly nice representation. 2 Let u(x) = V(X1' xg4---:4-xn), where v is defined in a region in R2. Putting y = ,/x§+ o°°+-x: , for n = 2 and 4 we have locally Au = 0 if and only if v(xl,y) = Imf(xl+-iy), f(z) analytic (n = 2) v(xl,y) = % Imf(xl+-iy). f(z) analytic (n = 4) There is no such formula in three dimensional space. In general there are anologs of the above (n = 2,4) cases for even (but not odd) dimensions. 2. FORMULATION OF A PROBLEM Local solution to the axi—symmetric obstacle problem in R4 2.1. In this chapter we consider the aXi—symmetric obstacle problem in R4 (n = 4) with the obstacle 142 e(xl,x2,x3,x4) = --3 jg: xj , and we study the local solutions w defined near the origin. By aXi—symmetric we mean W(xl.X2.x3,x4) = V(X,y) . where x = x y = ,/X2+-x24-x Note then 1’ 2 3 4 ' ‘ 6(X1.X2,X3,X4) = -% (X24-y2) = w(x,y). We shall formulate a problem of unfolding a singular free boundary passing through 0 in this chapter. But we need to state and prove a prOposition and a lemma first. Proposition 2.2. w is a local solution of the axi-symmetric obstacle problem with obstacle e in a region for which y > 0 if and only if u = yv is a local solution to the obstacle problem (for »n = 2) with obstacle &(x,y) = ym(x,y) there. 28 29 3599;: We check the conditions of definition of local solution stated in Chapter 1. Clearly w_2 e in the region if and only if yv‘2 t there. Also it is easy to check that AR4 w = 0 when w > e if and only if A 2(yv) = 0 when yv > ym (of course in the region R y > 0) because 2 2 8 v a v 2 5v 1 A W = —-—'+ ——- + -'-— = - A (YV) R4 ax2 ayz y By y R2 Now, it is only left to check the C1 condition. Assume V(X,y) is C1 in a region for which y > 0, then its composition with y = x34—xi4—x: , i.e., . 1 w(xl,x2,x3,x4) Is also C there. . l . Conversely assume w(xl,x2,x3,x4) is C 1n the region, then w(xl,x2,O,O) = V(X,y), where x = x1 and y = x2, is also Cl there. Here is the proof complete because v is C1 in the region if and only if u = yv is C1 there. # The last proposition shows that in a simply connected domain in the upper half—plane in the noncontact set V(z) = % Imf(z), where z = X4—iy, for an analytic function f. Remark 1: If w is an axi—symmetric local solution with the above obstacle, then we can extend the function v to be even in y. We shall use the extended even 30 function v and the odd one u = yv defined in a whole neighborhood of the origin in (x,y)—plane later on. Remark 2: Following Remark 1, sometimes, we shall need to use a modification of Proposition 2.2 concerning the x-axis (y = O) entering the region (recall y > O in the Proposition). We claim here that w is C1 in a region for which y 2 0 if and only if V(X,y) is C1 1 there. To show this let w(x1,x2,x3,x ) be C in the 4 region y 2 0. Then w(xl,x2,0,0) = V(X,y), where x = x1 and y = x2, is C1 there. Conversely if V(X,y) is C1 in the region, then w(xl,x2,x3,x4) composite function of 2 2 2 w(xl,x2,x3,x4) — V(xl,y) , y — ./x24—x3+-x4 . is continuous there by simply being a . x. Moreover since .VV = 3! —1 , j = 2,3,4, clearly these oXj Y Y partial derivatives of w are continuous when y # O. . . . 8v . Since v is an even function of y, thus 3? is an odd function of y. If we now let y approach to zero, then lim g2? = 0 because %z(x,0) = lim %X(X,y) (x,y)4(0.0) ‘ j Y ..o vY x. and 7% remains bounded (between zero and one). Hence w(xl,x2,x3,x4) is C1 in the region for which y 2 O. # 31 2.3. Local solution formula Let us consider an analytic arc Fl :x = a(y) through (xo,yo), where (x0,yo) is fixed near the origin with yO > O and x0 = a(yo). We shall construct a local solution in a neighborhood of (xo,yo), in the region y > 0, for the above axi—symmetric problem, with free boundary T1 and contact set I lying on one side of F1. Let us define the conformal mapping Ll for t complex near yO as 2 = L1(t) = a(t)4—it, where 20 = xO-i-iyO E Tl. See Figure 2.1. t—plane Figure 2.1 Lemma 2.4. For a sufficiently small neighborhood V of 2 L-1 exists as a conformal mapping and 0' l Lil(x+—iy) = y for X+-iy E El. Let VO be a sufficiently small one—sided neighborhood of F1 about 32 0' and let u(x,y) = -% y(x24-y2). Then u = yv is a local solution to the obstacle problem, in the region Z y > O in a neighborhood of z with obstacle t and the Q! noncontact set in V0, the free boundary T1, and the contact set in V--VO if and only if u(x,y) = (u(x,y), for (X.y) E (V-VO) Z i 2 —1 2 u(x,y) = u(Z) = 1(zO)+Re f §[ +4(L1 (q)) ]dq , z 0 for z E VO U 11 . Proof: Since Li(t) = a’(t)+—i # O for t (as complex) near yO in the t—plane, then L1 is a conformal mapping near yO taking the real axis (near yo) to the 1' Therefore L11 exists and it is conformal in a small enough neighborhood V of 20 and LIl(x+-iy) = y 1' L1 also maps the upper or lower half (the other half) of a neighborhood of y0 conformally onto a one— arc T on T sided neighborhood V (the other one—sided neighborhood) 0 of F about 2 The proof of the first part is complete ‘1 here. 00 Necessary part: With obstacle N we have . -- _ - 1 2._§ _ i _ - _E. _' wX-iuy — -Xy+1(2 x 1-2 y ) — X(x 1y) 2 y(x 1y) = g(x~ iy) (x+ 3 iy) = %(z- 21y)(Z+ 21y) = é—(z2+4y2) 33 For 2 near 20 function f(z) = %[z2+-4(Lil(z))2] is holomorphic. Note uX--iuy is holomorphic in V0 because u is harmonic there, and it is also continuous on V6 because u is C1 in V. For 2 E T . _ 2 2 2 _ Vu — VN- Therefore uX-1uy _ 2[z *‘4Y ] _ 2] = f(z). for z E Fl. Hence 1 we have —§[zz+4(L11 N in V0. To show this observe that on F1 we have u-¢ = O, . 2 1 en where n is the unit (outer) normal to F1. Thus u > N near T1 in V0 in the region y > O. # Remark. Instead of defining u = N on V-V = I, 0 if we extend the integral formula for u to be defined in a full neighborhood of 20, then u > N holds on both sides of T1 with u = N on T1. We also observe in the proof that v(u-—¢) = 0 in the closure of the noncontact set if and only if (L—l(z))2 = (Im z)2. # Now we are going to start the main purpose of this chapter by constructing a local solution near 0 with a singular free boundary in Example 2.5. In Section 2.7 later on we shall discuss heuristically unfolding of the above free boundary by some perturbation of Example 2.5. A class of examples 2.5. We want to construct a local solution to the above axi-symmetric problem whose free boundary possesses a cusp at the origin. Singular free boundaries for axi-symmetric obstacle problem are only in the form y2 = KX4n+l or z = H(t)-—% t2n+llog|t|jzit2n+l , t real near zero Figure 2.2—a or isolated points [17], where n 2 O is an integer, H(t) is a real analytic function near t = O with H(O) = 0, H'(0) > 0; and let T = T1 U T2 be the union of the four arcs, near the origin in the z—plane given . 2 , . parametrically by z = H(t)"? t log It] :lt, 0 g lt‘ < r0. See Figure 2.2. We assume symmetry of v = - u as before about the X—axis, namely that v is even in y. In this example the contact set I will be the shaded region as shown in the Figure 2.2 "—1 H -"'"‘ “‘-“9' "'1 Figure 2.2—b 36 Following the local solution formula given in the Lemma 2.4, let us first define L(t) = H(t)—gt log t+it, t g [io:o< o} (i.e. Jig arg t < 11 I 2 2 , where log t has the branch which is real for t > O; and L(t) for t > 0 represents the right hand part of T1. We proceed in five steps as follows: (i) 'We are going to show that L maps the real axis near 0 onto T and an upper one—sided 1 neighborhood of the real axis about 0 conformally onto an upper one-sided neighborhood of F1 about 0. To see this let t move along the real axis from O to r > 0. By the branch we chose L maps [O,r] into the right hand part of P1 in a one—to—one manner. Let t move from t = r along \t] = r in the upper half plane to t = -r, then back again to t = 0 along the negative axis (see Figure 2.3). 37 Continue the function L analytically along this closed path Yr and its interior Gr° Clearly L is continuous on Er and analytic on Gr-[O]. Along the negative part of the real axis in the t—plane L has the image: L(t) = H(t)-gt log ltl—it, -rgth. Argt=7r. This image is exactly the left hand part of T1 (above the x—axis) near the origin. We now want to show that the image of Yr 0 {t :It] = r], for r small enough, stays away from the origin in the upper half-plane. Let z = L(t), t = rele, 0 g e g N, for r small enough. Then Im z = Im[t(H’(0)-—% log r4—i(l.—%§)]+-O(r2) r[(l--27T—a-)cos 9+ (H'(O) —% log r)sin e] +0(r2) - f(67r) > O . This shows that for r small enough, the image of Yr 0 (t :[t] = r] stays away from the origin in the upper half-plane. Observe also that é%(Re z) = r[(—H’(O)+g log er)sin 6 ”IT 2 + ( 9-—l)cos 9]4-O(r ) < O . =1|N Thus Re 2 decreases as Q varies from O -to F. Finally 38 2 £L§(Im z) = r[-(l-—%§)cos 9 d9 + (-H’(O)4—%~log re2)sin a]+-O(r2) < O. This shows Im 2 has only one critical point, a maximum point, at a point 9 near %, and Im z is increasing 0 on [0,90] and decreasing on [60,v]. Thus Im 2 takes its minimum values at e = 0,w. Hence L1 is one—to-one on Yr' and L(yr) is a Jordan curve. Consider the topological degree (or winding number) d(L,Gr,z) for any 2 E L(yr). This equals one in the region bounded by L(Yr) (the Jordan curve), and is zero in the unbounded component of ¢-—L(yr). Because L is analytic, it follows that L maps Gr conformally onto the region bounded by the Jordan curve L(yr). 1 Hence L— conformally maps L(Gr) onto Gr‘ Moreover L"1 is analytic at each point of the closure L(Gr) except 0, where it is continuous (ii) Near any 2 E T O -(O] we have, by Lemma 2.4, l a local solution 2 u = w(zO)-% Im j [q2+4(L_l(q))2 Zo ]dq, 2 above -{0] T1 u = u(z), z beneath T1-—[O] . 39 The overlapping solutions along Tl-[O] agree and they are independent of 20 on each side of Tl-{O]. are two such initial points This is because if z 21' 2 on the right hand part, for example, then the difference of Z u(z) = ((21) -% 1m] [q2+421dq Z 1 Z u(z) = 1(zz>-%Im]‘z [q2+4(L'121dq 2 Z is (:(zl) -1(22)-% 1m 5 2 [q2+4 Figure 2.4 40 we may here check that u = N (by taking integration path along Tl-{O]) and vu = vN on Tl-{O] by observing the fact that [L_l(z)]2 = (Im 2)2 there. 2 Also 8 u; ) an where n is the unit outer normal to T-{0] at any = A(u-N) = -AN = Ny > O on Fl-{O], z E 1‘1-{0}. This shows u-N > O on the noncontact sets corresponding to the left hand and right hand parts of Tl-[O]. Note u given above is defined in 375;? and it is the imaginary part of an analytic function in fTG;T-—[O}. Thus Au = 0 in L(Gr). Define u = N on T1 and between T1 and real axis. To show now that u is a local solution in the upper half of z-plane, near the origin, we need to show u is C1 there and u > N in L(Gr). (iii) At this step we want to show u > N in L(Gr) Z 11(2) ~N(z) = -% Im j [q 2+4>21dq+% Y(X2+y2) 1 _E-Im yO[L2(s)+-4SZ]L'(s)dS +% (Im L(t))[(Im L(t))2+ (Re L(t))2] 1 1 3 2 .t = --3 Im[§(L(t)) +4t L(t)-8 J sL(s)ds] o +%(1m L(t))[(Im L(t))2+ (Re L(t))z] 41 2 3 2 t g(Im L(t)) —2 Im(t L(t))+4 Im J‘ sL(s)ds o = --l£% r3sin3 e log3r 37 + 3% r3sin2 9 log2r((l-%§)cos 9+ H'(0)sine) w + 0(r3 log r sin2 e) = F(r.e) , where t = re16 and capital "0" is used here uniformly near (r,9) = (0,0). For 9 away from zero or w, certainly u(z)-N(z) > O for r sufficiently small. For 9 near 0 or v, we divide by sin29: F(r,e) = E r3(log r)2((1_fi)cos 6+H’(0)sine) . 2 2 v Sln 9 F - -l—6-3- r3(log r)3sin9+0(r3 logr) 3W This shows that for 9 near 0, v we have EigLfll > 0 sin e if r is small enough. But we claim that there is a neighborhood of (r,9) = (0,0) in which _——7?_ > 0. To show this we do as follows F(r.a) = (l-Ei)cos 5+-H’(0)sin9 3 2-- r r (logr) sin 9 . I 1 log r — (log r)sin e+—O( ) 42 2 . . G(r,e) = (l-7§)cos 64-H'(0)51n 9-—(log r)Sin e 2 c > 0 0 g e g F for , for some fixed (constant) rO . O < r 3 r0 . l c We can assume rO is so small that 0(IBE_?) 2 -§-. Then -?r——J:L£L%L-—§— > g-. This shows that u-N > O in the r (log r) sin a whole region L(Gr). In the next step we want to discuss Cl condition of the local solution. (iv) In order to show that u is a local solution in the upper half of z—plane near the origin, it is now left only to show that u is C1 there. By what we explained in the first two steps it is clear that u is C1 in 2 # 0 there. It is also clear that u is continuous at the origin. So the only thing left is to show that the partial derivatives of u are continuous at z = 0. To do this it is enough to show that lim :13 = a—‘1J(o,o) , lim 33 = :11 (0,0). (x,y)4(O.O) °X 5" (x,y)4(O.O) Y Y (XIY)EL(Gr) (le)€L(Gr) By looking at the formula for u defined on L(G ), we have 43 Z 2 —1 2 $1: "23—25 m f0 [q +4(L (q)) ]dq = _% Im[22+4(L-1(2))2] . a_u= -11 I... (2 [q2+4(L‘1(q))21dq BY 2 5y 0 = ..% Re[zz+-4(L—l(z))2] . Thus lim a—u=O=M(OIO) I (x,y)4(O.O) 5X ax 11m g—‘l o = SJ (0,0) (X,Y)"(OIO) y y In the next step we complete the process of constructing a local solution to the axi—symmetric problem promised at the beginning of Section 2.5. (v) In order to show that v = % u is a local solution to the axi—symmetric problem it is only left to show that v is C1 at the origin. Since u is C1 at the origin by step (iv), thus v = % u is C0 there. Thus we only need to show the partial derivatives of v are continuous at the origin. As we argued in last step we only need to show: n- 1 a _ _2 2 y Y t y Re t2-Im f 52L'(s)ds O Y Hence 2 I 2 Im J s L (s)ds limfl=—2 lim Ret 240 By t4O Im L(t) [Im L(t)]2 . r cos 29 '2 1““ 2 . , . 29 r40 "F logr Sin 94—H (0)51n 9+—(l-—7?)cos 94-O(r) because we've shown in step (i) that the minimum of the function (2 O) 44 1im 5—V = 549 (0,0) , (x,y)4(0,0) ex ex lim EX =‘§£(0,0) (x,y)4(0.0) BY BY Z y Retzz+4(L‘l(z))21—Im ] [q2+4(L‘1)21dq O -—;-r(sin 39)log r-iL r sin 3e-—3% 3W 9W (-% logr sin 94-H'(0)sin e + 5 cos 394—% H'(O)sin 3e+—0(r2) 3 + (l-%)cos 9+0(r))2 £13) = -—% occurs at 9 = 0 or T with f(O) . . 2” logr Sin 9+—H’(O)51n 9+—(l-—1S)cos E i 45 and therefore for r g r0, r0 small enough, f(e)+0(r) >%. 31 Thus lim (x,y)4(0.0) BY = o = 2&3, (0,0). Also 2 _5_V_ i 2 -1 2 _ 2 Im(t ) 8X — -2y Im[z +4(L (Z)) ] — X-——ImL(t) . Similarly we can show 6V —2 Im(t2) 8 lim 5;: 11mm: O=-O—:% (0:0) . (XIY)"(OIO) r40 This completes step (v), hence v is a local solution to the axi-symmetric problem near the origin of the z—plane. The shaded region is the contact set; T1, T2 are the free boundaries, and the rest of the neighborhood of the origin is the noncontact set in the z—plane. See Figure 2.5. Figure 2.5 46 Remark 2.6. Simple calculation shows that F1 is symmetric with respect to the y-axis if and only if H is odd. In particular then L(io) is pure imaginary for 0 > 0. Putting a(t) = L(t)-it = H(t)-—% t log t and assuming H odd, we get L(t) = a(t)4—it, which for t > 0 parametrizes the right hand part of F1. The analytic continuation of this equation along Yr yields at -r < O the equation L(-r) = a(-r)-ir = -H(r)+—r log r4-ir = -a(r)+—ir , which shows that following the analytic continuation of a along Yr from r to —r gives a(ar) = -a(r) -21(-r) a (*) Moreover the equation of the right hand part of the free boundary Fl x = a(y) = H(y) -% y log y extended for complex variable implies ,, _ ~ 2- . .. a (t) — H (Hm?t — L (t) ( ) is single—valued, in fact a”(t) is meromorphic in a neighborhood of the origin. See (**). 47 We shall show later on that the analogue of (*), (**) will also happen for the perturbed problem where the cusp is unfolded to a smooth free boundary. Perturbation of Example 2.5: Heuristics 221. Let us now physically pull the obstacle down a little bit, keeping symmetry of the problem with respect to both axes. We then expect a new local solution surface v over m near (0,0) and a new free boundary. The new contact set should be near the old one. We might anticipate, for example, that the new free boundary would consist of two analytic curves T1 and T2, and left half plane, as shown in Figure 2.6 (recall in the right that the axi—symmetric solution V(X,y) is extended to be even in y). We want the right hand part of the free Figure 2.6 48 boundary T1 of this local solution to be represented by x = a(y), an analytic even function near zero, passing through 20 = x0 = a(O) with the region to the right of F1 lying in the contact set. Fl has the equation 2 = L(t) = a(t)+-it, for t real. t-plane Figure 2.7 Let us now first construct the local solution near 20 = a(O). We extend L again on the complex numbers. Since L’(O) # 0, L_l(z) is defined as a conformal map from a neighborhood of 20 to a neighborhood of t = 0, taking the contact set (to the right of T ) to Im t < O, l and the non—contact set (to the left of Tl)_ to Im t > 0. By the formula given in Section 2.3 for u = yv, on the left hand side of Tl we have (note N(zo) = 0) 49 Z V(X,y) = v = —i Im [q2+4(L‘l(q)>2]dq. 2y 2 O for y = Im z # 0. Observe that the above formula for V(X,y) defines a real analytic function, even when y = 0. Moreover this function is even in y. This is because L-l(q) is imaginary for q real, the integral above is analytic function which is real for 2 real. Therefore 22 —1 2 . Im I [q +-4(L (q)) ]dq is real analytic and is zero 2 0 when y = Im z = 0, and in fact is odd in y by the reflection principle. When y = 0 this formula yields V(Xl,O) = 113 V(Xl.y) = -%[.— my [q2+4>21dq1 y 20 Z=Xl = w(X110)-2(L_l(xl)) . (1) On the right hand side of 1‘1 we have V(z) = :p(z). From the past theory, 2.4, 2.5, we have that V(X,y) 30 defined is C1 across the free boundary (even at y = O), 50 with A(yv) = O to the left of T1. To check that v is a local solution near all there remains is to show v > w near 2 , to the left of F and that A W = O O 1 R4 2 in the non—contact set. We show §_A§L%gfl_= 4 at any 6n point of T1 where n the normal to T1 directed away from the contact set. We have 32(v— ) 32 u-N 152(u—N) . d—=——2( )=———2 (Slnce Vu=VNIU=1W an an Y Y an = l A(u..N) = -l-AN = 4 , for y # O . Y Y ' This is also true for y = O by continuity. To see A 4 W = O observe that w is analytic in noncontact set R since V(X,y) is analytic and even in y, and hence v is an analytic function to x and y2. Also observe that A W = 0 if y # 0. Hence by continuity A W = 0 if R4 R4 y = 0 also. Here is the construction of the local solution near 20 complete. Recall that in 2.7 we assumed the local solution to the axi-symmetric problem exists near the origin of the z-plane. This local solution agrees with the one constructed near 2 0° 2.7.1. Further Assumptions. We are assuming the perturbed local solution has a free boundary as shown in Figure 2.5. In particular assume v-¢ >.0 on the interval (—zO,zO) on the real axis (i.e., this interval is in the non-contact set). Assume further that the only 51 local maximum of v(x,0)-m(x,0) in this interval is x = O, that JL (v(x,0)-—m(x,0)) # O for x # O, at , 8X 2 x E (—zo,zo), and $13 (v(0,0)-—m(0,0)) < 0. See 6X Figure 2.8, where —K = -% (V(0.0)-$(0:0)): y = —% O is described by The graph f = a (A/Y)+i(/Y, where fl 6 [o,i./1'<] . or x < O is, by symmetry x = -L(JY). At t = iJK, the Riemann surface has a branch point with two sheets, (t-i./I_<)2 As L(f’?) each other analytic on at iJR. Consid so a(t), i.e. L is an analytic function of 2 there. This is because §_iY§21 (0,0) # 0. 8X and -L(jY) are analytic continuations of near i/K, it follows that (L(t))2 is [O,iJR]. In fact (L(t))2 has a simple zero er t E [—iJK,O]. a(t) is real analytic, L(t) ha closed segment and a(t) L(t) ve analytic continuation along this a(I) . a(t)4—it = a(E)4-i€4—2 it Thus (L(t) -2 it)2 simple zero at L(E) + 2 it is analytic on [-iJR,0] with a 2 —i(/E. Consider (Ll/(t)) =((L(t)—21t)”) (a”(t))2 which is analytic on (—iJK,iJR) as L(t) # O 2 53 H there. Near iJR, L(t) = (t-in)2M(t), where M(1JR) # O, M is analytic. Hence (L”(t))2 is meromorphic near id? with a pole of order exactly three. The same thing is true at —iJK with L(t)-—2 it. Hence // 2 (a (t)) = ——3‘t)3 (t +K) where h is real analytic, non-zero on [-iJK,iJK], and even in t. For clarity we display the analytic continuations of a(jY) and L(jY). See Figure 2.9. The symmetric Figure 2.9 path passing through (O,—K) is the graph A of 54 x = 1mm. 7? 6 [0.1.7121 (*) shown in Figure 2.7. Near Y = O we have L(JY) = a(JY)+—ij§ , where a(JY) is analytic function of Y because a(t) is an even function of t. Therefore the graph of x = a(jY) = L(JY)-1JY cross the horizontal axis at x = a(O), as shown by the solid line B in Figure 2.8. Indeed, the dashed graph A (*) may be displaced to the right to obtain the solid graph B X: sL(fi)—ifi> sL(fi) . Near (a(O),0) the graph A has the form x = L(j?) = a(JY)+ LfY, so may be continued as a smooth curve (as a(jY) is analytic in Y) to the graph of x = a(fi) _ 1,)? denoted C. Similarly, near (—a(0),O) the graph B has the form x = _wfi) — ifi = _a(fi) - 21.5 so may be continued to D: x = —a(JY)+-21JY. Continuing in this fashion yields all graphs (4) X: iamY)+NwY= tam)+Nit, w 10 H 1“ - O, 2.: ll 55 as shown in Figure 2.8. We may regard x as a function of t, where t lies on a Riemann surface with infinitely many sheets (#). d2X 2 // However, from (#) we see (——§) = (a (t) dt single valued as shown above. This last result and the )2 is equation of path D are analogue of the Remark 2.6. Second step. Let us calculate the above h(t) for the unperturbed problem of Example 2.5, where K = O. 2 ) = We have t2(a”(t) (tH”(t)-—%)2. Hence I 4 h(t) = t6(a’(t>>2 = (%)2t +O(t6) (**) is real analytic and even with a fourth order zero at t = O. In the perturbed problem we have (t2+K)3(a”(t))2 = h(t,K), where the left hand side is some special perturbation of the left hand side of (**). Since the current problem is a perturbation of 2.5, we expect that h: R2 -4 It must, near (0,0), be a real analytic function even in t near the function h(t,O) = (%)2t4+-O(t6) in some sense. Motivated by the Weierstrass Preparation Theorem [7] we might expect a family h(t,K) = (%)2(t4+2A(K)t2+B(K))¢l(t2,K) of perturbations, where 01, A, B are real functions with $1 analytic in t2 satisfying A(O) = O = 8(0), ¢l(0,0) = 1. Thus 56 4 2 (aI/(t))2 = (3)2 t +2At +B ¢2< 2 t IK) I w (t2+K)3 where K, A, B E I! are near zero, K > O, B > O and ¢(O,K) is near 1. Let fl.) -2@ NW2.) , p...) (t +K) Description of a problem 2.8. In the heuristic derivation of the above formula we regarded A(KL B(K), and ¢(t2,K) as depending on the perturbation parameter K. For K = 0, an explicit local solution with A(O) = B(O) = O, 2 _ 2 _ ¢(t ,0) ~ ¢O(t ) was constructed, where ¢O(0) — 1 (Example 2.5). Now we adopt a somewhat different view, that of multi—parameter bifurcation theory. We regard A, B, K and the function ¢(-) themselves as independent parameters. They vary in the parameter space as (K,A,B,¢) E R3XZ, in a neighborhood of (O,O,0,¢O) where Z is a Banach space of functions ¢ defined as follows. Definition 2.8.1. Fix p > O and let Z be the Banach space of functions @(w) analytic on the open disk 2 . . .. 2 [ml < p , continuous on the closed disk )4) g p , such 57 that @(T) is real for T real, with the norm HNH = supidpml, m 3 pz. .1 Remark: Fix $0 E Z 2(for some p) w1th ¢O(O) = l. 2 ¢O(y ) _%.__c;__ K = A = B = O and ¢O given in (***)). This gives a Consider fo(y) = (the f corresponding to family of curves Fl : x = a(y), a”(y) = fo(y). In ¢O(y2)—1 , _ 2 1. ________ _ 2. 1 . fact a (y) _ -W (y+ y ) —- —W (Y+-No(y)). where NO is odd. Integrating twice we get 2 2 . a(y) = D4—Cy-—F (y log y-y)-; Nl(y), where N1 is odd with cubic leading term. Comparing with Example 2.5 we get D = 0, H(t) = C-F%)y-—% Nl(y). Hence given ¢O we have a typical local solution as constructed in Example 2.5, which is not unique by presence of the constant C in the formula H above. Problem 2.8.2. Fix 0, an open disk about the origin in z = x4—iy plane, small enough (say 5 E domain of the local solution in Example 2.5). Fix 20 = xo+~1yo, x0 > 0, yO > 0, on the unperturbed curve (F1 in Example 2.5). Consider A, B, K near 0, ¢ near $0 in Z, and f(y) with the branch of f for which f(yO) < O, and (a(yo),x ) near 2 O 0' a’(yo) near ii (H(Y)-% Y 109 Y)) ° Y " y=yo 58 (1) Find necessary and sufficient condition that there exists a symmetric local solution in G with part of the free boundary X = a(y) a”(y) = f(y) near yO with a'(yo), a(yo) as stated above. (2) Are the constants of integration unique? In other words, is the local solution unique for the given f(y) above? # We will devote chapter 5 to handle this problem. But first on the third and fourth chapters we develop some theory needed to go over chapter five. For this purpose we put an end to the chapter 2 and we start chapter 3 now. 3. A NECESSARY AND SUFFICIENT CONDITION FOR A LOCAL SOLUTION NEAR THE ORIGIN The main purpose of this chapter is stated in problem 3.2 below. But first, using Proposition 2.2 and the Remarks 1, 2 after, we can restate the definition of local solution of the axi—symmetric obstacle problem as follows: An equivalent definition of a local solution of axi-symmetric problem in R4 Definition 3.1. By a local solution (u,O) of the 4-dimensional axi—symmetric obstacle problem, symmetric with reSpect to x—axis, on a fixed open disk 0:: R2 centered at the origin, we mean u :0 4 R such that (i) u(X.-y) = —u(X.y) .. 1 . 1 . (11) § u(x,y) = V(X,y) is C in 0 (iii) Ell-u(x,y) = V(X,y) 2 u(x,y) in o, where m(x,y) = -% (Xzi-yz) . \ . 2, 1 (1V; if I = [(x,y) E J I? u(x,y) = w(X,Y)]I then Au = 0 in G —I. 4 59 60 Remark. In part (iv), Au = 0 in (0-—I) is equivalent to A 4 W = 0 there. To see this let R A 4 W = O in (O-—I), then u(x,y) = yw(x,y,0,0) R implies azu _ 62w azu _ aw 82w ——y—2', —ZE+Y'—‘§. ax ax By BY _ _ ii = 2E - If y — 0, then Au — 2 8y 0 because BY 18 an odd real analytic function of y. If y # 0, then 2 2 v a v 2 Av Au - y(0 + -§ + - :—) = O . 3X2 y y 01/ Conversely if Au = 0 in (O-I), then as in (the end of) 2.7 we can show A 4 W = O. R Let us now consider (as before) an analytic arc :x = al(y), near (xO,yO) E 0, where yO > O, X =a 0 l(yo) and al(y) is real analytic near yo. Problem 3.2. What are the necessary and sufficient conditions for (u,0) to be a local solution to the obstacle problem (in the sense of Definition 3.1) with El as a part of the free boundary and the contact set I to one side of T1? # We shall obtain the solution to this problem step by step throughout this Chapter. Let us start by considering the complex form of 61 I -z = Ll(t) = al(t)+ it, t E R, near yO . Extend L1 to be defined for complex t near yO as in 2.3, 2.4. As in Lemma 2.4 we can construct a local solution uO in a neighborhood of (xo,yo), with free boundary T1, and noncontact set lying on one side of P Let v Cio be a one-sided neighborhood of z on 1' O O the side of F1 in which the noncontact set lies. For Figure 3.1 the remainder of this chapter, we deal with this local solution (u ) arising from the given are Fl as O'Vo above. We shall try to extend this to a local solution on the disk 9. Before doing this, we need to state and prove the following lemma. 62 Lemma 3.3. Assume V(E 0) containing V is an open 0 connected set. Then Au = O in V with u = uO in V0 if and only if (Lil)2 defined in V0 has a (single-valued) analytic continuation in V and I (L11(q))2dq is real for c every closed path c C V. Moreover we have then i 2 z u(z) = N(zo)+-Re I2 2[q +—4(Lil(q))2]dq, z E V O which is independent of the path of integration in V. Proof: First assume u is harmonic in V, and equals 110 in V0. Then uX--1uy is holomorphic in V. In VO we have that —— _ . _ i 2 -l 2 _ vu — uX-—1uy — 2[ +4(Ll (2)) ] — f(z) is analytic, so f and (L_l)2 have analytic continuation 1 in V. Hence the integral representation of u remains valid: Z u(Z) = (h(zo)+Ref f(q)dq Z 0 Z _ l p 2 —l 2 — (:(zo)—2 ImuzZ [q +4(Ll (q)) ]dq, 2 e V o as both sides of this equation are harmonic functions agreeing on V0. If 2 can be reached from 20 by different paths 11’ v2 in V, then Re " f(q)dq = Re 5 f(q)dq because u is single—valued. "‘ 1 "'2 63 Thus Re f f(q)dq = O, which implies Y1'Y2 . -1 2 _ . Re § 21(Ll (q)) dq — O, or equivalently Yl'Yz -l 2 . . . . -1 2 . § [Ll (q)) dq is real. This implies I (Ll (q)) dq ls Yi'Yz C real for any closed path C C V. Conversely let (Lil)2 have an analytic continuation from V0 to V, and put d Z Z u(z) = N(20)-+Re V f(q)dq . O . where the integral is along some path from 20 to z in V. We assume § (L11(q))2dq is real, and therefore C f f(q)dq is imaginary for any closed path C C V. This C 2 guarantees u(z) = N(zO)+-Re E f(q)dq to be single— 2 O valued in V and independent of the path of integration there. Hence u is harmonic in V and u = 110 in V0. This completes the proof. 4 Let us now begin extending u to a local solution 0 of the obstacle problem in the sense of Definition 3.1. The first step is defining a set containing VO on which uO is extended as a harmonic function ul with the property ul > N. Definition 3.4. Consider (continuous) paths v: [0,1] 4 0' with V(O) E VO such that 64 (l) uO can be harmonically extended along the path, i.e., there exist 0 = t0 < tl < ... < tN = l, and harmonic functions uO k defined on some open disks Dk E O centered at y(tk) with uO,k{Z) = uo,k+l(z) on Dk fl Dk+l # ¢ and u0 = u0,0 on DO. See Figure 3.2 Figure 3.2 (Note this extension may not be single—valued on U Dk); and (2) uO,k(z) > w(z) on Dk’ where w(z) = -% y(x24-y2). Let Vi = {y(l) ly is a path as above with properties stated in (l) and (2)}. # Remark 1. V1 is clearly an open connected set containing V0. Following the proof of Lemma 3.3, we see that condition (1) above is equivalent to analytically continuing (Lil)2 from y(O) to V(l) (along v) with the disks Dk' Let ul(z) denote the corresponding harmonic continuation of uolz). 65 —l 2 Remark 2. ul as well as (Ll ) extended above may not be single—valued° By Lemma 3.3 ul is single- valued harmonic in Vi if and only if (Lil)2 has a single—valued analytic continuation (from V0) to Vi with § [LIl(q)]2dq = real for any closed path C (: C IV. . 1 Remark 3 o O V] is a dis .IOint union Of tWO SUbsetS P and Q defined as follows: P = [z 6 avi [there is no y :[O,l] 4 5 with y<1)= z. y(O) 6 v0, and wow) co. having properties (1) and (2) as in Definition 3.4} , Q = (avi) —P observe Q E (avi) D 50. We call P the Natural Boundary of Vi and it is closed. # Let us now extend (uO’VO) anti-symmetrically to its reflection in the lower half plane. Definition 3.5. Let V5 = {z :E 6 Vi}, _ . _ I , x w OO—[Z‘ZEVO}I Wl—Vln{(XIY)‘YZO/‘I v w2=v2’l: {(x,y)zy W) is valid. Remark 2. Note even if W1 is not connected, ul(z) can still be defined as before (by means of integration along the paths which extend into the lower half plane). However we easily see that W1 = Vi C {(x,y) :y > 0], W2 = V5 C [(x,y) :y < O}, and W W are connected if u extends to a local 1’ 2 0 solution u in 0, since u(x,O) = w(X,O) = 0. Remark 3. awl is a disjoint union of three subsets: = P ' v1 fl 8W1, V1 0 5W1 C J, and Q N 5W1. Set v2 = [z :z 6 v1}, and v = v1 U v2. If we have Wl = Vi C {(x,y) :y > 0}, then the second subset of 5W1 above is empty. Lemma 3.6. For 2 = L (t) E V we have 1 O —l - _ -l —l — 2 _ -l 2 L2 (2) — —L1 (2) and [L2 (2)] — [L1 (2)] . Proof: T1, T2 have equations 2 = Ll(t) = a(t)+ it for t real near yO > O and z = L2(t) = a(—t)+—it for t real near -yO < 0, respectively. Therefore, for t e L'lw) l O 67 Ll(t) = a(t) a? = a(t) —i't‘ = L2(_E) . -l — — -1 Putting z = L1(t) above we get L2 (z) = -t = -L1 (2) for z 6 V0. Hence [L;l(E§]2 = [Lil(z)]2 for z 6 VO . We are now going to state two hypotheses and then prove they are equivalent. Hypothesis 3.7. The functions uo and uOO in V0 and VOO respectively have a common single—valued harmonic extension u to w, which is C1 in W; and u = w and vu = vw on (v-J). Observe that v-—J C aW is the closure of the set of Natural Boundary points which do not lie on the x—axis. Note by this Hypothesis we have D u2(X,y) = u(x,y) = -u(X.-y) = —ul(x,-y) for (x,y) 6 W2 well-defined as a harmonic extension of uOO in W2 for which u2 < $2. # Hypothesis 3.8. 1) (Lil)2 defined in V0 has a single-valued analytic continuation in W1 and § (Lil(q))2dq is C real for any closed path C C Wl (which imply similar properties for (Lgl)2 in W2)° 68 2) (Lil)2 is continuous on W1, [L11(z)]2 = y2 for z = X+—iy 6 (v--J),[Lil(z)]2 = [LIl(z)]2 = [L;l(z)]2 «Z -1 2 is real for z E J, and lim Im J [L1 (q)] dq exists z4z* zO ZEWl for an 2 5 BW and equals l'-(y3--y3) for any Y * P l 3 * Q 2* 6 (vl-J), where the integration path is in Wl U {20}. # Proposition 3.9. Hypothesis 3.7 holds if and only if Hypothesis 3.8 holds, and in such a case we have vi = wl c ((x,y) =y > o]. v; = W2 C (my) =Y< 0}: with W1 and W2 open and connected. Note: Under either one of the above Hypotheses if J-—(vl-J) # ¢, which is a countable union of disjoint Open intervals on the real axis, then along any path from 20 to 20 in W U (J-(v-—J)) passing through a point z in J-(vl-J) we shall have 2 3 E0 —1 2 —3 Yo = Im f (L (q)) dq. 2o —l 2 . . —l 2 where (L ) shows, in this case, both (L ) and (L31)2 because these two are the analytic continuation of each other in W U (J-(v-—J)) by the Reflection principle. Instead of the latter integral condition we can impose an equivalent condition: 69 Proof: Assume Hypothesis 3.7 hold. Let us first show the second part (the set equation). If J = ¢, then clearly we have wl=vic ((x,y):y>0}. W2=V5C {(x,y)zy<0}- 1 If J # ¢, then u being C in W and u(z) 2 w(z) in Wl imply “.2 O on J (continuity of u is enough here). Similarly u(z) g w(z) in W2 imply u S O on J. Thus u E O on J. Hence Vi = Wl C {(x,y) :y > O], Vé = W2 C ((x,y) :y < 0}. These show W1 and W2 are open and connected. Let us now show the first part. The Hypothesis 3.7 2 is assumed, thus, by Lemma 3.3, (Lil) has a single valued analytic continuation in W1 and § (Lil(q))2dq C is real for any closed path C C W1. Since u is C1 in W, thus u(z) = @(zo)+Re J." f(q)dq = vu = f(z) = i-ifgz4-4(L-l(z))2], for z E W , 2 l l _ z u(z) = ((zOHRe j_ f(q)dq 2 O — = vu = f(z) = -i[324-4(L_l(z))2], for z E W , 2 2 2 imply (LZ1)2, (L31)2 are continuous on W1, W2 respectively. 70 If J # ¢, then putting the above values of vu equal on J we get: (Lflznz = (L;1(z))2 for z 6 J = wl n 562 . —1 Using Lemma 3.6 we can conclude [L;1(E)]2 = [L1 (2)]2 for z E J because [L;l(E)]2-[Lil(z)]2is an analytic function which is identically zero on V0, and therefore it is zero on W1. Comparing the last two equations we get [L;l(z)]2 = [L11(z)]2 for z 6 J. Hence 2 [Lil(z)]2 = [L;l(z)] is real for z E J. For any 2 E W we have _ = ' _ 1 = "TT*__— 1 vu vm ux+iuy vu uX iuy VU ' — - 2 = —%[ZZ+4(le(Z)) ]—V1L’ = 21[y2- 0}, W2 = Vé C [(X.y) zy < 0} (l) in 3.8 again implies, by Lemma 3.3, that u is a single-valued harmonic extension of uO defined in V The first integral condition in (2) of 3.8 implies O. u(z*) = lim u(z) = w(z*) = O , V 2* E J . z z* ZEW The rest of the proof goes as in the case J = ¢. Let us now conSider the subcase ¢.§ J"(Vl“J).§ J. Then (1) in 3.8 again implies, by Lemma 3.3, that u is the single—valued harmonic extension of uO at least in the component of W containing V (similarly we 1 O can argue about the lower half of the z—plane). Note J-(vl-J) are those points of J which are not limit points of the part of v above (outside of) the real 1 axis. Thus for any small enough disk, D(z*;r), neighborhood of any 2* E J-(vl-J) we have D(2*,r) n{(x,y):y(_1)3 < o} c w.. In fact J— (V 3 l"j) is open and it is a countable union of disjoint Open intervals on the real axis. Let (a,b) be the maximal interval containing 2* and contained in J-—(vl-—J). Any such point 2* in this open interval can be joined 75 to 20 6 BW by a path in W with only endpoints z*, 1 z 1° Clearly a or b e (v --J)° By the first 0 l integral condition in (2) of (3.8) we get lim u(z) = w(a) (or w(b)) = O (for detail look z-0a(or b) zEWl u ' u I . at only if part). Suppose now (al'bl) C Wl‘g Vl With u(al) = w(al) = O or u(bl) = w(b ) = 0. Then 1 2 _ u(2*) = ((20 )-— my: OIq +4(L11 O}, and by symmetry 1 1 W2 = Vé C {(x,y) :y < 0}. Since the limit in (2) of 3.8 . l 3 3 ————— eXists, and equals to ‘§(y*-yo) for z* E (vl-J), thus u(z*) = lim u(z), 2* E 5W, exists and can be zaz 'k defined, and equals to $(2*) for 2* E (vlfl-J). Since (L_l)2 is continuous on W , thus vu is continuous l 76 on W1 (for detail look at "only if" part) and therefore u is C1 there. Moreover u s O on J. Also [Lil(z)]2 = y2 for z E (187:7?) implies vu = v on (Vl’J)° Let us finally consider the subcase .I- 63;:73) = where (31::3) O J = ¢. Then any point z 6 J can be joined to 20 (and EC) byapath in W1 (and in w2 with only endpoints 20’ z perhaps in awl (and E0' 2 E 5W2 perhaps). Let us show now J C awl. Since (L_]‘)2 is continuous on W, thus for z 6 J we have Z ul(z) = (<20) ——§- Im “r [q2+4(L11)21dq 2 ((2) = o 2o u (z) = w("z‘ )—l Im (2 [ 2+4(L‘1(q))2]d w(z) — o 2 o 2 9— q 2 ‘13- ‘ Zo that is -i 3 Im (2 (L_l( ))2d and , 3yo.>. . i q q z 0 z i y3 2 Im F (L_l(q))2dq. Since (L—l)2, (L-l)2 are 3 O 02' 2 l 2 O analytic continuation of one another by the Reflection principle (use Lemma 3.6 and (2) in 3.8), thus 2 z ————————— 2 Im Ur; (L;l(Q))2dq = Im j’_ (L;l 0}, W2 V2 C [(x,y) .y < 0}. Condition (1) of 3.8 again implies of course, by Lemma 3.3, that u is a single-valued harmonic extension of uO and in W U J. The rest of the proof in W1, W2, can be stated as before. # Definition 3.10. Assume either one of Hypothesis 3.7 or 3.8. Now let us extend the function u from the set W to the set 0' as u(z), if z 6 W u(z) = . w(z), if z E 0-—W Theorem 3.11. (u,@) in Definition 3.10 is a local solution to the obstacle problem in the sense of Definition 3.1 if and only if z 1) 1im if; y3+-Im f (L-l(q))2dq] exists and y 3 O 1 z-vxl zO ZEWl equals to (L11(xl))2 for any x1 6 J, ~—---—- 2 _ — n — 2 2) 11m -J3[-y(Lll(Z))2+“% ygi-Im J (L l(q)) ] 24x1 y 20 zEWl exists and equals to i[(L11(xl))2]’, V X1 6 u. 2 l (x)> 20, Note: Since (— u-w)( = -2(L-l x Y -1 , ] at any X E J, thus any possible real zero of (L ) in J-(V-J) is not 78 simple. These zeros are isolated in J- (v -J) and they form a set S so that S U (v-J) establishes the free boundary aI for the obstacle problem. Proof: Assume u satisfies the requirement of Definition 3.1. Then v = % u is C1 in O, in particular, it is C1 on the real axis. If J = O, then nothing needs to be proved. Assume therefore J # O. For any z 6 W1 we have then Z u(z) =1)21dq 0 Hence 1 _ l 2 _1_ 2 _2; 3 z —1 2 y u(z) - —2 x +6 y —y[3 yO+Im “(2 (L1 (q)) dq], O for z 6 W1 . (*) Furthermore for any 2 6 W1 we have 1 -1 -_ A __l. - - v(§u(z)) — yuX+i( u — 2 u) — (uX+iu )— 2 11 Y Y _ i _i 2_i —l 2 __i_ 2 i — -x+2y-2yx — y[Ll (2)] +2yx -6y 2 2i —1 2 + 713 yO+Im ( (L1 (q)) dq] Y 20 = —x+—— y+-2%[—y(L'l'l(z))2+-§ yg Y z . _ 2 + ImJ (L.l(q)) dq] (**) 79 J may consist of two classes of points, one which are limit points of v-—J (the part of v outside of J) and the others are the interior points of J-(v-—J). We treat them in different ways. First let Xl be a limit point of v-—J, i.e., in the first class. Since % u(z) is continuous on W 1! ' 11 3 .z —1 2 thus by (*) we have lim '—[§ yoi-Im J (Ll (q)) dq] 24x y z __1 o ZEWl exists, where z = xi-iy. To find this limit let z* 4 x1, where 2* E (v-J), then .13 z-12_i3133_13 11m {-3- y0+1mf (Ll (q)) dq] — 3 yo+3(y*-yo) — 3 y. zaz* zO zEWl and 1 1 3 “2* —1 2 . 1 2 lim —— 3 yoi-Im 3 (L1 (q)) dq] = 1im 3-y* = 0 2*7X1 y* 20 Yieqo 2*6wl (there are such 2*) . Or we can argue as follows: 1im % u(z) = z—vx1 - ‘ 2 x . 1 _ . 1 _ 1 lim ——-u(z*) — lim —— w(z*) — "7f° Thus by (*) z*+x1y* 2*4xly* 2 lim l[l y34-Im F (L_l(q))2dq] = O, which proves y 3 0 . l 24xl 20 C" Z‘Wl 80 condition (1). (Note for such an xl we have [Lil(xl)]2 = 0). Since u(z) is C1 on the real axis also, thus by (**) , 1 —1 2 1 3 Z -l 2 lim —2[_Y(Ll (z)) +w§ y04-Im f (Ll (Q)) dq] 27:1 Y 2O ZEWl l 2 is continuous on W and exists. Recall that (L1 1 ) 2 equals to y2 on (v-—J), and lim Im I (Lil(q))2dq = z 242* zEW1 O Y3-Y3 * 3 O for 2* 6 vl-J. To compute the above limit we can make these substitutions to get it equal to . 2 . . . 11m --3 y* = O (which proves condition (2)) Y*40 Or we can argue as saying 1im v(£ u(z)) = lim VO = -x1 24xl 24x ZEWl zEWl and consequently by (**): —— Z . 1 -1 2 1 3 —1 2 _ ii: 2[-y(Ll (2)) +3 yO+Im (‘2 (L1 (q)) dq] - o , zEWl which proves (2) . Now let xl be an interior point of J-—(v-—J), i.e., in the second class. Then any point z c D(x1;r), 81 for sufficiently small radius r, can be joined to 20 or 20 in W, where only endpoints may not be in -l 2 -l 2 L2 ) W. Recall that in this case (Ll ) , ( are analytic continuation of one another across the x-axis '1>2. and we may call both as (L Again since % u(z) is continuous on W1, thus by (*) z _ - 2 lim %[% ygi-Im f (L l(q)) dQ] 22:1 z0 zEWl exists for any such x1. To compute this limit we can do as follows Z lim hi y3+1m§ (L‘1(q))2dq1 y 3 0 24X 2 O ZEW x z = lim i[% yg-i- Im i (L'l(q))2dq+ Imj‘ (L‘l(q))2dq] zax y z x _l O 26W = lim — Im j:( 1(q)) exists, z+x l zEWl where z = Xi'iY: % yg+-Im j (L—l(q))qu = O. (The latter equation was shown at the end of first part, ”only if”, of the proof of 3.9.) Moreover 82 z -1— 2— Imf:(Lc1))2dq=Im‘((L(q))dq=Im§:(Ll(q))2dq= X z -1 2 z —1 2 Im f (L (q)) dq, which shows Im I (L (q)) dq is x x an odd function, real analytic vanishing for y = O. 1 z -1 2 Thus § Im I (L (q)) dq is a real analytic function even in y. Hence Z 1% yg+ 1erz (L'l2dq1 0 lim l Z-’X y zEWl Z = — Im 5 (L'lmnqut X II 721 (D b l N Since % u(z) is C , i.e., v(% u) is C in W, thus by (**) we have 11 %[-y(L_l(z))2+% Im J“ (L'l(q))2dq] 24:1 y 20 26Wl —"'_‘ Z = lim —l§[—y(LZl(Z))2+ Im j‘ (L_l(q))2dq] 11 X1 _1 W1 exists. To compute the value of this limit, since we have real analytic function (which vanishes at y = 0) divided 83 by y2 involved, thus it is enough to consider z a Xl vertically, i.e. z = xli-iy. Then by L'Hopital's Rule — 2 lim -J§[-y(L-l(z))2+-Im ( (L_l(q))2dQ] z-vxl y xl zer -(L‘1)2 —y ga—uL‘Hz) >21 +Re(L_l(Z) )2 = lim y 2y Z"X 26Wl i Im(L-l(z))2 —y Big/(film? = lim 2 Y 24X]- ZEW:L = gnaw-1n. ))2]'+-i- Re121' l 2 . - 2 = 1[(L 1(x1))]’ where the last equality is because (L—l)2 is real on J. This proves condition (2). We discover also that (% u(z)-—m(z)| = -2(L—l(x))2, for any x E J, including x . ———— —l 2 2 the pOints of J 0 (v-J) because (L (z)) = (Im z) for these points. The rest of the note is clear now. Conversely assume conditions (1), (2) hold. If J = O, then we are done. So let J # O. Using conditions (1), (2), equations (*), (**) imply that u is C1 on the real axis. Proposition 3.12. The local solution (u,e) is symmetric with respect to y—axis if and only if 84 [L‘l12 = [L‘H—En Proof: Assume (u,0) is symmetric with respect .. _ Lu - .53.- to y—aXis. u(-X,y) — u(x,y). Then ax(x,y) — _BX( X,y), i.e., uX is odd with respect to X while u is even. 2E _ fiE._ _ - _ _ Moreover 5y(x,y) — ay( X,y). Hence ux( X,y)4-iuy( X,y) — —ux(x,y)4-iuy(x,y), that is, uX(-z)-—iuy(—z) = -(uX(z)-iuy(z)); where u(z) = z . - 2 . ((zo)4-Re f -%[q24-4(L l(q)) ]dq. Applying Cauchy— z 0 Riemann equations we get %[(—E)24-4(L-l(—E))2] = -—§-[22+4(L'l(z))2]. Thus —2i(1.'l(—E))2 = -2i(L—l(z))2, that is, [L-l(-E)]2 = [L-l(z)]2° In particular if z is imaginary, then —E = z and therefore [L'l(z)]2 is real. Conversely assume now [L-l(—E)]2 [L_l(z)]2. First notice that u(JE) = u(z) if and only if -% Im ”(2 [q2+4(L—l(q))2]dq = _% Im d(-z[q2+4(L-l(q))2]dq. Z0 Z0 N I This is true if and only if Im f [q2+-4(L-l(q))2]dq = 0, 2 -Z if and only if Im f (L-l(q))2dq = 0. Let us show now 2 "z -1 2 the last equation: Im F (L (q)) dq = ”2 C -1 2 ‘E -1 2 Im[f (L (q)) dq-E-r (L (q)) dq], where the path from 2 c z to —z is symmetric with respect to y—axis crossing 85 y—axis at c. Hence _]_( _E 2 Im I (L q)) dq 2 C Z _ __ — Im[f (L‘l(q>>2dq+-j (L'1(-q>)2<—dq)1 Z C C C— Im[f (L'1)2dqi-§ (L‘l O, yO > 0, where f(yo) < O and a”(y) = f(y) near yO as stated in Problem 2.8.2. Following Section 2.7 we want to treat a special case of Problem 2.8.2 here: Problem 5.1. Study Problem 2.8.2 in the special case when K > O. 4 .1 Lemma 5.2. Let IO be a compact segment (e.g. interval), and L be analytic in an open neighborhood of IO. Let L’(z) # O, V 2 6 IO, and also let L be one-to—one on IO. Then there exists a neighborhood of IO on which L is a conformal map. Proof: Suppose not. Then there are sequences 1’ r ' = ‘an}’ (hm), an 4 IO, bn 4 IO, w1th L(an) _L(bn). Then there are subsequences an 4 a, bn 4 b, where a, b 6 IO. 92 93 ‘-~-—-_—..— Figure 5.1 If a # b, then we get a contradiction to one—to-one assumption because L(a) = L(b). If a = b, then L(bnk)-L(ank) L’(a) = lim ———————-————-—- = 0 because L(b ) = L(a ), k4m bn -a nk nk k “k V k 5 EL Here we get a contradiction to the assumption f'(t) # O on IO. This completes the proof. # Necessary conditions in Problem 5.1 5.3. We want to find necessary conditions that there exist a local solution symmetric with respect to both axes in 6, with part of the free boundary Tl :x = a(y), a”(y) = f(y), near yO Here (a(yo),yo) 18 near 20 = (XO’YO) fixed on the unperturbed free boundary curve XO > O, yO > O, and al(yo) is near é§(H(y)-% y log y)| , f(yo) < O, y=YO / 4; ,2 f(y) = -%-4Jéfi§§3:% @(yZ), ¢ near $0 in Z; M (y +K) / A,B,K near 0, K > O 94 We begin by building up the set v on one side 0 of T1 (over) as in Lemma 2.4° Consider the complex form of Fl: 2 = L(t) = a(t)+ it, t E R near yo . Extend L to be defined for complex t near yO again as in the past. On an interval about yO (on the real axis) L is one—to—one, L' # 0, thus by Lemma 5.2. L-1 exists and is a conformal map in a neighborhood of L(yo). As long as f remains real on the real axis to the left of yo, or equivalently y4+ 2Ay2+-B remains non- negative there, we can extend F L analytically and the 1! local solution also nearby T as in Section 2.7 by 1 Lemma 5.2. Lemma 5.4° The case where t44-2At24-B has a simple root t2 = 32 E (O,yg) cannot happen. Proof: Suppose it can. The proof proceeds in three steps. First step: Without loss of generality we can assume t44-2At24—B > O for t E (B,yo). The relation “y 4_ 2 ” yl (t 4—K) a’(y2)-a'(yl) = ’ ( ) I B S y2 < y1 g'yo shows that a (y2 > a (yl) for the above yl, y2. This shows a’(y) > O for B < y < yO and a(B) < X We now want to show a(B) 2 O and O. 95 conclude that a(B)4—iB 6(3. Suppose a(fi) < 0. Then a(B)+-iB is in the second quadrant. We can extend T as the graph of function a as long as it remains 1 inside 0. We can also extend L nearby (B,yO]. The local solution built up near a(yo)+iyO can then be extended near the new Fl as shown in Figure 5.2 with Figure 5.2 the shaded region as noncontact set. By symmetry with respect to the y-axis we have the same construction starting in second quadrant with a free boundary F L2 intersecting Tl at 2 on the y—axis. Here we get a contradiction because above T1 near 2 we have u > Q, and on T in the noncontact set to the left of 2 we 2 have u = 9. This shows a(B) x O. .1— 96 Since we assume the existence of a local solution of the perturbed problem in 0, thus the shaded region is in the set W defined in Chapter 3. We are first going to show that any small circle about a(B)4-iB does not lie entirely in W. Hence such a circle must intersect the free boundary somewhere other than on F 1° t-plane Figure 5.3 Since B is a simple zero of t44-2At24-B, it (t-B)2 ; . 3 follows a”(t) = — 4IN —' I) aj(t--B)j near t = B. O 97 Integrating twice we get a(t) = a(B)+—d:l(t-B)- & . 2(t--B)2 27 d{(t-B)j, where o’l, of are real, and Tr j=o j - j L(t) = L(a)+ (a;1+i)(t-a) -% O, G:1+'i # O . 1 Then (z-—L(B))2 = g(C04-C3§3+O(§4)), where CO # 0, C3 # O, and inverting this function gives H (.A.’ 5 = (z-L(B))2[b0+b3(Z-L(B))2+0((Z-L(B))2)], where bO # 0, b3 # 0. Hence (t-B) U.) = <2-L)2 = 132+ 23b§ +b§2 rolm . 7 + 4BbOb3(z—L(B)) +o< B, y1 = B4—(y-—B)e2Wl, which are in two different Riemann sheets, then L(yl)-L(y) = $(y-—B)2 Z) d3(y-—B)J;>O j=0 for (y-B) small enough, which shows that L is locally one-to—one from a cut neighborhood of B (the interval (B,y) deleted) into a cut neighborhood of L(B) (a cusp deleted) as shown in Figure 5.4. In TVs 7” t—plane Figure 5.4 99 fact if y moves to B, then the image of y moves to L(B) on T and the image of y1 moves to L(B) l on some path P The paths T1 and T' are tangent I 1' l —1 2 _ 2 at L(B) by (*). Note [L (2*)] — (Im z*) for a point on the free boundary. Thus L-l(z*) = y* near B here. If there were a segment with left endpoint B (on the real axis) mapped onto a new part of the free I] ll boundary passing through L(B), then L = a would have to be real there which is not true. Thus Fl, Ti establish the boundary of W1 near L(B)° Ti is the new part of the free boundary, and it is easy to show directly u-¢ 2 O in the cusp deleted neighborhood of L(B) with equality occuring only on Tl, Ti (one can also refer here to the Kinderlehrer and Nirenberg cusp result [10]). This shows a(B) = 0 cannot happen by symmetry with respect to the y—axis if we argue as in Figure 5.2. Then we can extend the local solution near Ti also as we did for T1. Second Step. In this step we want to end up with a contradiction which proves the lemma. Note this Lemma involves two cases according to whether there is another root, say a, O < a < B, of t44-2At24-B or not. Because of these singularities we may consider ,— cuts along the x-axis as in Figure 3.5. In the upper lOO Figure 5.5 figure there is another simple root d, where in the lower one there is not any. Consider the analytic continuation in the upper half plane of the branch of f we were using for definition of T1 on the right hand side of B. Note t a(t) = a(B)+—j a’(¢)d¢ B t t = a(B)+-[Ta'(T)] -f Tf(T)dT B B t = a(B) +a'(B)(t-B) + f (t- g)f(g)dg B Since the zeros of L’ are isolated and L’(t) = a’(B)+ i+‘F f(;)d§ # O for —B S t‘g B, thus we can join B to d (or B to -G, or B to -B 101 in the second case in Figure 5.5) by a path y staying away from the zeros of L' with Im y(s) > O for O < s < l, y(0) = B. y(1) = a (or y(1) = -a. or y(1) = —B) , so that L is invertible locally in a neighborhood of any point y(s), O < s < l, with analytic inverse L-l. So we have an open connected region containing y((O,l)) in the upper half plane in which L is locally invertible. L is continuous on the closure of this region. The part of y near B is mapped by L in W1. Thus either there is a point on y, y(g), O < 9 < l, which is mapped by L to a point on the free boundary or all points y(s), O < s < l are mapped in W (assume the image stays inside 0 in this case, we will show A this in the third step). If there is such a point y(s), then [L_l(L(Y(Q)))]2 = [Im L(Y(g))]2; L-1(L(y(g))) = mg) , which is not true no matter which one of pairs [B,d}, {B,-o}, {B,—B) is joined by v. Therefore there is no such a y(g), and all y(s), O < s < l, are mapped in W. At the end of this step we will show that L(G) 102 (or L(—d), or L(-B) cannot be mapped inside W and they must be mapped into the free boundary by L. Assume this now. Hence -1 1 a a = L (L(a)) = iIm L(a) = i[c1+—if(C1-Q)f(§)d§ '13 (note a(B), OMB) > O) * -a = L_l(L(—G)) = :tIm L(—d) 1 +0 ' = 1pm.: fa (_a_g)f(g)dg] * -B = film-13)) = iIm L(—B) = 1 [43»:— f (—B-g>f(g)dgl B (there is no a in this case) ** Positive sign case in any one of these leads to a I contradiction immediately (in *, * leads to o = B and in ** leads to —B = B) because f has fixed sign there. In *, *’ negative sign case implies 1 .B . 1 B 2a = 33 (d—g)f(g)dg, +2a = Ti (+a+g)f(g)dg, d a l . . respectively, where I f(g) is negative. These two results cannot be true simultaneously. Hence the case of 4 2 having two roots 0 < d < B for t +—2At 4—B is ruled out. Now assume this polynomial does not have any zero in (O,B). illnllllll 103 Then ** implies = — If B(B4—g f(g)d§ which is not again possible because i f(g) g 0. This contradiction proves the Lemma if we can resolve the two questions which are left open. We conclude this step by answering the second question: Claim: L(a) gin L(-O() £177. L(—B) e 17v. Proof of the claim: We only prove the first one and one can argue similarly for the others. Suppose L(a) fi W, then (L‘l)2 has to be analytic at L(a) (note (L_l)2 here is defined in the noncontact set) by results of Chapter 3. To check this observe that l 2 2 2 2 — L”(t) = ii: ‘5 )‘t/gc‘ ) ¢dg < ans). for d the pair {6,-a], where t < B in any case above. So L(y) remains to the left of X = a(B) and beneath y = B in any case. Let us now show that L(y) does not intersect the x—axis. Suppose it does, say at L(Y(Q)). Then by Theorem 3.11 and the fact that (% u-¢)| = -2(L‘l(x))2210 x _ /\ on J, we must have L l[L(y(s))] imaginary. This cannot happen obviously if y is joining the pair [B.G}. To show it cannot occur in the other two cases we do as _ A follows. If L l[L(y(s))] is imaginary we can assume the path y intersects the imaginary axis only at 0. See Figure 5.7. Then L(O) = a(B)-—Ba’(B)-—j g f(g) 3 Figure 5.7 106 which is supposed to be real. But this is not real if Y is joining any one of the pairs [B,-o}, {B,-B} shown in Figure 5.7. This contradiction results that L(y) does not intersect x-axis in any case. Now in order to show that L(y) remains inside 0, it is enough to show that it does not intersect y-axis° Suppose it does. Then by symmetry with respect Bu 2 B to y~axis we have —;(O,y) = —2 Im(L_l(iy)) = 0 (on y-axis) by Proposition 3.12. This shows then L_l(iy) is either real or imaginary. For the pair [B,o} we can take the path y in such a way that it does not intersect the X-axis except at endpoints. This rules out the case {B,a]. For the other pairs {B,—a}, {B,—B] we can assume y intersects X—axis and y—axis only at O as in Figure 5.7. Then L-l(iy) = O, and 3 therefore L(O) = a(B)-—Ba'(5)4-f Q f(Q) = iy, for O y > O. The latter result is not true because % f(g) < O and a(B)-—Ba'(B) is real. Thus L(y) cannot intersect y—axis either. Hereby we showed L(y) remains inside 0 and in the first quadrant. Here are the third step and the proof of the Lemma both complete. é u Lemma 5.4 proved that t44-2At2+-B cannot have a simple root 0 < B < yo. Thus it has either a double root o, O < o < YC’ or no root in (O,yo). Next Lemma shows that the latter case cannot happen either. 107 Lemma 5.5. The case t44—2At24-B > O, for O < t2 < yg, cannot happen. Y Proof: Since a’(y)-a'(yo) = j f(§)d§ > O for yo 0 g y < yo, thus a'(y) is increasing to the left of y0 but it is always finite. Moreover a defined by y a(y) = a(yo)+a’(yo)(y-y0)+f (y-§)f(§)d§ , yo for O S y < yO is finite, decreasing as y moves to the left of yo. Therefore F intersects x—axis transversally (or 1 y-axis non-horizontally). This provides a contradiction as follows. The local solution built up near a(yo)4-iyo, with T1 as a piece of free boundary, is continued to the left with noncontact set over Fl. By symmetry WWW ‘Q/ WL/fl/ Jb/Wqé%%@%z O I ..//, I [7/ , // ”7/7 /' WW a( )_ i ”WW yo Yo Figure 5.8 108 with respect to x-axis (or y-axis) we must have u < w on the part of T1 below X-axis (or u > w on the right hand side of y—axis) as shown in Figure 5.8. On the other hand we must have u = W there. Here is the proof complete. # Corollary 5.6. The only possible case for f is 2 2 a -t 2 2 2 f(t) = — (t ), where O < d < y . Moreover W (t24-K)3/2 ¢ 0 yo a'(O) = o, a(O) > o, a’(y ) = j f(g)dg. 0 0 Proof: The first part is an immediate conclusion of Lemmas 5.4, 5.5. If a’(O) # 0 (note a’(O) is finite), then we get a contradiction as in Figure 5.8 in Y Lemma 5.5. This result yields a'(yo) = I O f(§)dg. We 0 can now build up the local solution near a(O) with the noncontact set to the left of F as in Section 2.7° 1 If a(O) < 0, since a'(y) = j f(g)dg shows that the o tangent to the curve Fl never becomes horizontal, thus at the intersection point of F1 with y-axis we get a contradiction, by symmetry with respect to y-axis, similar to the one we obtained at a(O) in Figure 5.8. See Figure 5.9. Hence a(O) 2 0. If a(O) = 0, then similar argument leads to a contradiction. Nearby the origin to the left of we have u > h: -on the other i1 hand on the part of the free boundary symmetric to T1 109 // Y \ 1‘ \€%;/ 1 We) 1 / . -a(yo) + iyo WWW/W J a(Yo) + iyo {é Figure 5.9 with respect to y—axis we must have u = w. This contradiction shows a(O) # 0. Hence we conclude a(O) > O and the proof is complete. # 2%” -a(yO )+iyO Figure 5.10 llO Remark 5.7. We can build up the local solution near a(O) with the noncontact set on the left hand side of T1 in the first and forth quadrants as in Section 2,7. Then T1 :2 = L(t) = a(t)+-it, t near zero, is symmetric with respect to x-axis, and z u(z) = -% Im f [g24-4(L_l(g))2]dg, for 2 near a(0) a(O) to the left of T1, where L_1 is a conformal map in a uniform neighborhood of T1 near a(O), and % u-—w = —2(L-1(X))2 2 O on the real axis to the left of a(O), where is contained in J. # Remark 5.8. Since u is a local solution to the problem, thus by Theorem 3.11 the map (L-l)2 must have analytic continuation to the left of a(O) on the real axis and it must be negative—valued until it becomes zero again (by symmetry with respect to y—axis) and we get a new point of the free boundary on the X-axis. Since (L-l)2 # 0 there, thus (L—l)2 has analytic continuation if and only if L"1 has, as long as (L-l)2 is nonzero. 4 M For the time being we want to check the necessary conditions for L (obtained already) to fulfill these requirements, so that by these conditions we can conversely construct a local solution near a(O) continuable to the left to get the local solution u eventuallyo lll Lemma 509° L"1 in 5.7 does have analytic continuation to the left of a(O) on the real axis at least on (L(iJE),a(O)) and (L_l)2 is negative-valued. L(iji) is the first critical point of (L_l)2 to the left of a(O). Moreover, the value of (L-l)2 is (i/E)2 = -K. Further % L_l(z) is indeed strictly increasing as 2 < a(O) decreases. t Proof: Note a’(t) = j f(T)dT, O t a(t) = a(o)+f (t-T)f(T)dT, L(t) = a(t)+it, O L'(O) = i # O and L is conformal in a neighborhood of 0 taking the real axis to Tl and the imaginary axis to the real axis to the left of a(O) by z = L(t) = a(t)+-it as shown in Figure 5.11 (one can check the formulas above to see there). t-plane Figure 5.11 112 Furthermore L’(t) = i+ l f(g)dg shows that l i L’(iO) > O for O < C < J? and it becomes unbounded as o 4 (JR)-. Hence L is real and strictly decreasing on [O,i/R], which can be seen from t L(t) = a(O)+—f (t-g)f(g)dg+—it also. Applying Lemma 0 5.2 for L on a sequence of intervals [O,ib/R-en)], where an 4 O, we get L a conformal mapping on a neighborhood of each [O,i(Jfi-en)] not containing iJE. Hence L"1 is a conformal mapping in a neighborhood of (L(ijfi),a(0)] not containing L(L/R). Moreover L-l takes this part of the real axis to the upper part of the imaginary axis and imlL—l is strictly increasing to the left a(O) on the closed interval [L(iJE‘,a(O)] (because of having similar property for L)° A _ Let x < a(O) be the first critical point of (L l)2 to the left of a(O). Since éi (L_1(X))2lA = X -l d -l d -l 2 L (x) d; L (x))A = O a a; L (X)\A = 0, thus x x _ A L’(L 1(x)) is infinity and by definition of L’ we _ /\ ._ ,_ A conclude that L 1(X) = iJK. Thus L(iJK) = x. # Definition 5.10. For the singularities t iJE, we assume the imaginary axis in t-plane is cut from i]? to m (upward) and from —ij? to m (downward). Define m(t), for t near zero, in the lower Riemann sheet as analytic continuation of L(t) along the paths of the 113 t form a, where L(t) = a(t)+it = a(0)+i (t—g)f(g)dg+it. - 0 Note a'(LfR) is not defined (infinity) but a(ifi) is defined as a finite number. Also m(t) = b(t)4—it, where b is the analytic continuation of a along the paths of type 0 (see Figure 5.12). # l t—plane Figure 5.12 Remark 5.11. Recall that during move from a(O) to the left, we are looking for necessary conditions for having local solution u so that they are going to be sufficient at least for constructing a local solution near a(O) continuable to the left. At this stage in order to have a local solution u it is necessary to 2 have (L_l) negative-valued on the real axis near _ _1 L(L/K) and L ‘ must be analytic in a full neighborhood 114 of this point. Let us now check these requirements for our map L_l. Lemma 5.12. L‘1 is analytic in a full neighborhood of L(ifio with (L'1)'(L(ifi<)) = o. L"1 is imaginary- valued on the real axis part of the neighborhood and 3— L-1 l the left of L(ifi), L‘ attains a local maximum j? at L(ijE). To 1 . . . . . is invertible w1th inverse m as defined in 5.10. Proof: In a sufficiently small neighborhood of ij? we have 3 2 2 ——— m 2 _ L”(t) = "777% ¢(t2) = (t—iwll—O 2 (t +-K) A0 # 0, t near iji , i an . and L(t)—L(ifii) =A_'l(t-i/E)+(t_t/E)2 '2 Ajf(t_iJE)3, 3:0 1 A6 # 0 (by integrating twice). Putting (t-iJE)2 = g, z = L(t) we get z-L(iN/Iz) = A11g2+ E A]! §2j+l = §g(§). 3:0 where g(O) # 0 and g is analytic near zero. By invoking the Inverse Function Theorem we get i = n(z-—L(iV?)), where 9(T) is analytic near zero with 5(0) = 0, SIM” 7’0. Hence t—zufik= 92(z-L(iJE)), and t = L_l(z) = iAth92(z-L(i/EO) is analytic in a full neighborhood of L(iJE). Since 9(0) = 0, \ a’(O) i 0, thus the latter equation of L_l(z) shows 115 (L_l)'(L(iJ§)) = 0. Since L-1 is imaginary—valued to the right of L(ijfi), thus in the expansion of L-1 in power series of (z —L(iJ§)) all coefficients are imaginary and L_1 is imaginary—valued to the left of L(iji) also. By looking at the formula for L"1 we see that if z-L(ij§) changes its argument by N, then t -i/§ will change its argument by almost 2w if |t-iji| is small enough. In particular by starting from 21 > L(ifiE) and ending with 22 < L(ifi) (see Figure 5.13), t-i/E will change its argument by exactly 2? by the above argument, that is, t = L_l(z) starts from t1 = L-l(zl) on the imaginary axis and ends up with t2 = L_l(zz) on the imaginary axis again. Note tl is in the upper Riemann sheet and t2 in the lower one. We observe that -% L."l attains a local maximum J? at L(iji). The last part of the Lemma is clear. Figure 5.13 116 Lemma 5.13. % L_1 is strictly decreasing as 2 < L(iJE) decreases, with inverse m, and it vanishes at m(O) = b(O) with (—i L'l)’(b(o)) a! o. b defined in 5.10 has the equation b(t) = ui—Xt-—a(t), where X = f f(g)dg is pure imaginary, u is real, and C l I X4—l S 0. Proof: If (% L_l)' vanishes at some point E e (b(O),L(ijE)), then m’(t) = Li—i-—a’(t) becomes infinity at some point i0, 0 < O < JR. This result is not true because m’ becomes infinity only at i L/K. l —l . . . . Hence i L is strictly decreaSing as z < L(iJ?) decreases with inverse m, strictly decreasing as t falls below id? on the imaginary axis. By symmetry with respect to y—axis L_l must vanish at some point and it cannot have minimum before that. This implies . . . . .f— . l —l m is strictly increaSing on [O,lVK]. Again i L cannot vanish at point m(0) with derivative zero there by m’(0) being finite. To show the formula for b observe that m” = b” is also an analytic continuation II of L” = a along the paths of type 6 (referring to Definition 5.10). Since along 0 the argument of 3 (t-—i¢§)2 is changed by 3w, thus b"(t) = —a”(t) for t near 0. By integrating twice we get b(t) = u-tlt-—a(t) and m(t) = ui—(ki—i)t-ea(t), where 117 x = b'(O) =J‘ f(g)dg. m(O) = L(“21(0) and a(0) are ’a real, therefore u is real. Since m is real-valued on [O,iji], and a(t), u are real there, thus A must be imaginary there. Moreover m being invertible on this interval implies l i x+-1 g 0 We also have the following relations m(ifi) - L(ifi) = b(ifi) = aw?) = kiffi = 2am?) -u = km? i K _ 2a(o>+2j (iv'K-§)f(§)dC.-u O and b(O) = a(O) _f gf(g)d<;. 0, Corollary 5.13. % u-w attains its absolute maximum (2k) on [b(O),a(O)] at L(iji), whereas it vanishes at a(O), b(O) and strictly monotone on each side of L(iji). Remark 5.14. By symmetry with respect to y—axis we must have —a(0) g b(O) < a(O), and this condition will be fulfilled if u (the constant of integration) ischosen real and appropriately. We have thus far imposed the following conditions also: k = j f(g)dg ll8 iV _ imaginary, u = —>.iJIZ+2a(o)+2 J" (iV’K—g)f(g)dg, O and %(X4-i) S 0. In the next Lemma we show that equality case is not necessary here. Lemma 5.15. The point m(O) = b(O) = u-a(0) belongs to the free boundary and this implies the necessary condition k+—i # O, i.e., %(l+—i) < 0. . l -l 2 . Proof: Since y u-m = —2(L ) on J, vanishes at m(O), thus m(O) E (v-J) U s, the free boundary. To show l+—i # 0, we check v(% u-w) = 0 at m(O) = b(O). Suppose x+i = 0. Then m(t) —u+a(0) = 2 4' ll . -llt -k2t -..., where kl — a (O) > O, i.e., x2 3 2 z—u+a(0) = —)\ s(l+——- s+— s + --:) if we put 1 Al kl 2 . dz _ t = 5. Since a; I — —xl # 0, thus by the Inverse mapping theorem we have -llt2 = (z-u4-a(0))+ k 2 2 —2- (z—u+a(o)) +0((z—u+a(0)) )= e(z-u+a(0)). )‘1 where e is an analytic function with 9(0) = 0, a'(O) # 0. We discover that (L_l)2 is analytic at 1.1—am) but [(L“l(x))2]’\ #0. Thus u-a(0) v(% u-$)(m(0)) # 0, which violates a necessary condition of Theorem 3.11. Hence k—ti = 0 is impossible. # 119 Lemma 5.16. Either b(O) is an isolated point of the free boundary or 1 = -2i. Proof: We discovered so far that m(t)-(u-a(0)) = z-b(O) = (1+-i)t-xlt2-12t4---- for t near zero, where 1+ i # 0 by Lemma 5.15. This means that the inverse of the function m exists and is analytic in a full neighborhood of b(O). The inverse, which is the analytic continuation of L'l, can still be called L_l: 1 1 _Z_:_b_(_0_l. [1+Tl— t+__2._ t3+ 090 t: >.+i +i 1+i A 1 l 2 3 son 2 co. +(>.—+i‘t+>.+it+ )+ 3' i.e., 1 1 t z—z _ -l _ z-b(0) 1 z—bgo) l 0 ... t _ L (2) _ xi-i [14'x+i ( li—i +x+i 1+1 + ) 2 + *1 (3:09.... (k+i)§ 1+1 where the rest of the terms are of order greater than three. Hence 2 2 3 _ —l _ z-b(0) 11(z—b(0)) 211(z—b(0)) t—L(z)—>\+i+ 3 + 5 (k+-i) (Xi-i) Let's investigate % u(z)-—m(z) near b(O) = m(O). For every 2 near b(0) in W we have 1 222 r2 -1.2, V u(z)-—w(z) = 3 y -— Im d [L (s)] ds , _ y a(o) 120 where the integration is along a path in W U {a(0)] -l 2 from a(O) to 2. Since (L ) is real valued on Z [b(O),a(O)], thus l u(z)-—o12dg=§ 1‘g b(°))2+ l 4 mm b(O) b(O) (1+ 1) (x.+i) =1(_z_-.1-;12>>_3+il (z—b(o>)4+u 3 (>~+i)2 2 (7\.+i) and Z In“? (L‘l(;))2dg= l 2[(x_h(0))2.—3- Y3] “20 (1+1) 1 +-+2[4(x-b(0))3y 2(x+i) _ 4(x—b(0))y3]+ m, 121 where the higher order terms are ignored. Note all terms on the right hand side of the latter equation have y factor because we've taken the imaginary part of the right hand side of the former equation. Thus for z E C we have Z --2- my (L‘l2d; y Z 0 —2 2 1 2 =————'HX-MOH -*Y) (>.+i)2 3 )‘1 3 2 -—4 [4(x-b<0)) —4(X-b(O))y 1+m. (1+i) 1 —2 2 2 2 1 and — u(z) -o(z) = —— (x-b(0)) +— y (1+————) y (X+i)2 3 (>.+i)2 411 3 2 - ‘——*_Z [(X-b(0)) —(x-—b(0))y ]+—"' , where the higher (X+i) order terms are ignored. The right hand sides are real analytic. Let‘s now define w(x,y) as a function of x with y as a fixed parameter as follows: (u(x,y) = —‘—27 (x-b<0))2+§y2(1+——l——2> (1+i) (1+i) At this point we are going to show l+~——£——§_2 0. Suppose (Mi) not. Then 2 2 —2 2 :h—§y(1+l—2)= ———2- (x-b<0)> . (1+i) (k+i) - for fixed small \yi # 0 122 This shows a parabola with negative minimum % y2(l+-—JL—3) (1+i) which intersects the horizontal x-axis transversally (look at Figure 5.14). Since m(x,y) has the leading Figure 5.14 terms of % u(x,y)-—w(x,y) near b(O) and 52 1 —§ (— U(x,y) —cp(x,y))l > O , ax y (b(O),O) bifurcation theory in this case [4] says that % u(x,y)-¢(x,y) behaves similarly as w(x,y) does, in the following sense, for small fixed perturbation of y from zero. % u(x,y)-—¢(x,y) as a function of X with parameter y, for any small fixed ly) # 0, intersects the xy-plane transversally and it has a negative minimum. But we know that if % u(x,y)-—¢(X,y) vanishes for any fixed small perturbation of y from zero, then it must have even multiple root with reSpect to X, i.e., % u(x,y)-—;(x,y) must touch the horizontal plane 123 tangentially, not transversally. Hence 1+ 1 2 2 O, (l+i) . l . .-l 2 . . i.e., l -—:1____2 , that is, (i 14-1) ‘2 l which is (i 1+1) equivalent to: i-lk-tl 2 l or i_lk-+1 g -1. We already knew that i-ll < —l (by Lemma 5.15), thus the case i-ll 2 0 can't happen. Therefore 1‘11 3 -2. If -; l < -2, i.e. 1+-——L——- > 0, then we'll have i . 2 (1+1) an isolated point b(O) of the free boundary because % u(z)-m(z) > O for z E V-[b(0)], where V is a small enough neighborhood of b(O). We'll show that this case cannot happen either. Finally if l+-——L—-§ = 0, _ (1+i) then i 11 = -2 will be the only possible case which works. Lemma 5.17. The case % 1 < -2, which says b(0) is an isolated point of the free boundary, can't happen. Figure 5.15 Proof: Suppose it can. Then (L—l)2 has to have analytic continuation along the real axis to the left of 124 b(O) and it has to be real—valued on the real axis. Again %u(z) —cp(Z) = -2[L_l(z)]2 > O ' for z = x < b(O) near b(O) . Since l+—i # 0, thus L_l exists and it is analytic in a neighborhood of b(O) and is continued analytically 2 to the left of b(O) because (L-l) is nonzero there. Since L-1 is pure imaginary—valued on the real axis to the left of b(O), thus L-l has the image on the lower half of the pure imaginary axis because if it had the image on the upper half of the pure imaginary axis, then m would be doublevalued on the upper part of the pure imaginary axis near zero which is not true because m is one—to—one near 0. Thus % L_l(x) < 0 for x < b(O). But then m(t) = ui—(li—i)t-—a(t) implies m'(t) = (l+2i)—a’(t)—i = i(i'1x+2)—i(% a’(t)+l) , which shows %-m'(t) < 0 for % t < 0 and It) sufficiently small, while t 4 —id§ on the pure imaginary axis implies % m’(t) a 3. Thus m'(t) must vanish some— » where between 0 and —iJE on the pure—imaginary axis. But this implies, then, m is not invertible there which is not true because If:L exists and is analytic on the u '/_ n - _12 1 ~' .' ' segment [-i.K,0] and (L ) takes its minimum not before -iJK. Hence we get contradiction and we conclude 125 that i-ll < -2 cannot happen. Here is the proof complete. # Our investigation and search on the real axis is still going on in O as long as we haven't reached to the part of the free boundary on the left symmetric to the earlier one with respect to y-axis. At this stage we realize that i_ll = —2 is the only possible case left. Then 1 = —2i and the midpoint of L(t) = a(t)+-it, m(t) = u-—a(t)-it, for t = i0, J? > O > 0 is %(L(t)+-m(t)) = % independent of G. In particular as we've already also seen b(O) = u-—a(0), i.e. % = %(a(0)+-b(0)). By the next theorem and symmetry with respect to y—axis we'll discover the midpoint % = 0. Theorem 5.18. There is a new part T2 of the free boundary passing through b(O) which is the mirror reflection of the old one (passing through a(0)) with respect to the vertical line x = g = L(ijE). Thereby u = O and b(O) = —a(O). Proof: Since 1 = —2i, thus m(t) = n-—a(t)-—it. This formula, for t real, determines a curve passing through m(O) = b(O) = u-—a(0) (as a vertex which we already know belongs to the free boundary) concave to the left and symmetric with reSpect to X—axis because a is even. 126 m which was the analytic continuation of L along the paths of type 0 determines the upper (lower) half of the new curve for negative (positive) t, and clearly for Itl small enough but % t > 0 we have m(t) > u-—a(0) = b(0) = m(O) . t—plane x/‘>‘/( ;/\l:'/‘< . M'fi-XX x mm o __ a v I I 1“ \l‘ \l‘ \/ \l \ Ix 1‘ /\ z ...... Y y i #UY) Xxxxk.xxxx I Figure 5.16 Let's try to show that the new curve is the left part of the free boundary. We have m’(t) = —a’(t)-—i # 0, m is one-to-one, for t real Thus by Lemma 5.2 there is a strip neighborhood of any segment of the real axis, centered at O, in which m is a conformal mapping taking the upper (lower) part of the neighborhood to the right (left) hand side of the new curve. The inverse of m is L_l. Along the new curve (the image of the real-axis under m) we have -l(z) = —Im z = -y. For 2 on the new curve and taking the integration path on this curve we get [L‘1(z)12-y2 = o a v6 I1(2) -cp(Z)) = o, i u(z) —o(z) = g y2—3 Im drz [L‘l(g)]2dg Y Y b(O) 2 2 2 z 2 =—y——Im (Img)dz=O, 3 Y jb(0) where the first equality is shown by computations in the proof of Theorem 3.9. Moreover 2 1 (-u—co) _ 2 _ 2 = -—252(u—'-h-)=-l-—-L92—M=AA(u—¢)=4>0 an an y an y on the new curve. The latter three results about the new curve passing through b(O) = m(O) imply that this new curve is part of the free boundary. This new free boundary is symmetric to the earlier one with respect to the vertical line X = %. Since we have symmetry with respect to both axis, thus % = O, i.e., v = O, B(O) = —a(0), u and m(t) = —a(t)-it. Moreover 128 L(ifi) = mm) = a(ifi) + mm?) = (ham/E) -i(i./—IE) => (1 = 2mm?) + 2i(ifi) = L(iffi) = a(ifi) +i(L/—IE) = % = 0 which was expected by the symmetry with respect to y-axis. Hence the result. # Remark 5.19. If we started by considering a piece of arc on the left (new) free boundary, which specifically starts with the branch of a” negative at O; m(t) = —a(t)-—it, and assumed the noncontact set on the right, then we could make a similar study and by analytic continuation of m—1 we could reach to the right free boundary: L(t) = a(t)+-it. In both cases cross signs (x) are mapped into cross signs and dot signs are mapped into dot signs in Figure 5.16. The signed area in the z—plane are contained in W. We've already seen that Q44-2Ag24-B cannot have a simple root i0, 0 g 0 3 ji, and it can't have a double root at O, iJE. Remark 5.20. (a list of necessary conditions for Problem 5.1). So far we have derived some necessary conditions for problem 5.1 as follows: 2 2 d — 2 2 ————————- ¢(t ), where A = —d , B = d (t2+K)3/2 HM 129 (iii) f f(g)d; = —2i, 0 (iv) a(O) = 3 f gf(g)dg > o G Y (v) a(O) = a(yO)-yo a’(yo>+-(OO cffdg) 0 Note using (iii) we can show that (iv) is equivalent to ifi a0)=.fi—f (Mi-QH(OdL 0 We realized that there is only one constant of integration arbitrary and the other one is imposed by the symmetry (e.g. by a’(0) = 0). Recall that in the unperturbed problem case also, given 00, the local solution was not unique by depending on one parameter c as stated in the Remark following Definition 2.8.1. # Lemma 5.21. Except perhaps near the semicircle the mapping L is conformal in a semi—circular domain T bounded by v consisting of segments [O,yo], [O,iJE] in the upper Riemann sheet, segments [-yO,0], [O,ijfi] in Ithe lower one, and the semi—circle centered at O (in both sheets) and radius yO started from yO in the upper Riemann sheet and ended with -yO in the lower one. The image will be bounded from below by [—a(0),a(0)], Fl, 12. 130 Figure 5.17 Proof: The mapping L is analytic in T, continuous up to the boundary. L takes the lower part of the boundary of T to [—a(0),a(0)] and (T1 U T2) 0 [(x,y) :y > 0}. Recall that by assumptions of problem 5.1, a(yo), a’(yo) are near x0, é§(H(y)-—% y log y)| (in unperturbed Y 0 problem) respectively. For any t on the semicircle we have for unperturbed case: t _ - _ I r _, — Lu(t) — it+-au(yo)+-(t yo)au(yO)-+J (t :)fo(g)dt Y0 = it+x + (t—y )[i(H(t) -—2— t 1og't)] o 0 dt‘ 1‘ - _ t—Y <1 2> O t (S -%7 (t—;) 0, dc 1 V. :2 131 for perturbed case: t L(t) = it+a(yO)+ (t—yo)a’(yo)+f (t—g)f(g)dg Yo iti—a(yo)+-(t-yo)a'(yo) t +% (y (t—g) 7—953 @(gzmc. 0 where the path of integration can be chosen on y starting from 0 on the positive direction. Since K, o2 are small, in comparison with y0 = |t| = lg\, thus the integrands are in both cases uniformly close; and therefore Lu(t), L(t) are also uniformly close, for all t on the semicircle. But the image of the semicircle is a simple path in the unperturbed case as we showed in Example 2.5 (Figure 2.3) . Hence the image of the semicircle under L is uniformly close to the above simple path. Now consider the topological degree (or winding number) d(L,T,Z) for any Z f L(Y). This is zero in the unbounded component of ¢-—L(y) and it is one in the major bounded component of ¢-—L(y) which is bounded by F1’ T2, [-a(0),a(0)] from below (note there may be small bounded components bounded only by the image of the semicircle under the map L). Hence L is a conformal mapping in the preimage T0 of the above major bounded component of ¢-—L(v). (One can show directly, as in Example 2.5, that L(v) is a simple curve here also). # 132 We can argue similarly about the other semi-circular domain which is mapped below T [—a(0),a(O)], F by l’ 2 the analytic continuation of L. Hence we have the following: Theorem 5.22. L"1 is a conformal mapping in a neighborhood of the origin, more Specifically in L(TO) U [—a(0),a(0)] U L(TO) U T U T2, where by bar 1 here we mean the complex conjugate. We can assume the disk 0 is inside this neighborhood. Moreover the free boundary consists of T1 U T2 only. Proof: The first part is clear from the above argument. The last part is simply because L_l(z) = i Im 2 can happen only on T l l U T2. 7 Sufficient conditions in problem 5.1 continue Remark 5.23. By using the necessary conditions derived so far and stated in the Remark 5.20, we could think of process after Corollary 5.6 as part of the establishing the sufficient conditions for having a local solution v = % u in the following sense. We could establish a local solution near a(O), T1 as explained in Remark 5.7, and we could continue it along a strip neighborhood of [—a(O),a(0)], where L_1 is analytic and imaginary valued and the local solution is bigger than w except at : a(O). We could then continue the 133 construction of the local solution to agree with the local solution near —a(O), T2 without any need to Lemmas 5.15. 5.16, 5.17, and Theorem 5.18 because we have now the formula for T2 which is represented by m, the analytic continuation of L. We can then apply Lemma 5.21 and Theorem 5.22 to 1 show that L_ is a conformal mapping in the subset of O bounded between T1 and T2. Thus the formula for the local solution can be extended to the whole disk 0. In order to show i u = v is a local solution in the sense stated in Problem 5.1, it is enough to show % ui-m > 0 in the domain bounded between F1. T2, or u-—¢ > 0 away from (—a(0),a(0)) there. # Theorem 5.24. u-—w > 0 in the domain 2: Y L(TO) U [-a(0),a(o)] U L(TO), where by bar here we mean the complex conjugation. Proof: It is enough to show u-¢ > 0 away from (—a(0),a(0)) in L(TO), because we can do similarly to show u-—¢ < 0 away from this segment in L(TO) by symmetry of L, L—l, and also because % u-m > 0 in a uniform strip around (-a(0),a(0)) bounded by T1, T2. Let us first investigate the sign of u-v along ' i e A A the image of the arc segment t = yOe , 0 g 7 g :0, 134 _-‘\~\\\\\\\\:;plane T T 2 ./fi; 1 4’ ’1’ “ \\ \ \ \‘ d ~ \ ,/ \| _‘ "_T(,— x O \ 1 / / ’ //l t—plane / I f i J? ’l 0 yo Figure 5.18 under the mapping L. Put 9 = §L . Let 1(s) = L(t(s)) = s O L(yoe yo) and F(s) = u(l(s)) —¢(l(s)). Then F(O) = O, F'(O)=V(U-tl;)').'(s)l =0, and F”(O)= S=O = —Ale'(O)H2 which is near A(u-'¢:)H7~'(S‘IHZ| s=0 4yOHké(O)H2 where the index u indicates unperturbed case. Thus u-w > 0 along 1(5) uniformly with respect to perturbation parameters. Similarly we can argue and 135 show u-¢ > 0 along the image of the arc segment ie . t = yOe , v-eo g e g v under the mapping L. To show u-¢ > 0 elsewhere on L(t), ie.., for t = yoele, 90 g e g W-eo, let zp = a(yo)+-iyo. Then for any 2 on this arc we have for unperturbed solution: Z 2 _ uu(z) = $(20)-% Im (2 [g +—4(Lul>2dc Zo Z + 2 Im in (L_l(§))2d<; Zp _ ‘ ‘ l rzp 2 —1(zO)—1(zp)—2 Im ‘20 g dC Z _ 2 Im r [>2->21dc Zp 2 .2p —1 2 — Imv‘. (Lu (QH dg Z 136 Since zp is near 20' thus w(zO)-¢(zp), 1 z 2 z -1 2 _—2— Im 5 p g dg, and —2 Im ( p(Lu (g)) dg are small. . 2O 20 Putting Lglm = t we get (L;l)2- (L'l(g>)2 = [L;1(L(t))]z—t2. Since L(t) is uniformly (with respect to t) close to Lu(t), thus L;l(L(t))-t is near zero uniformly with respect to t. Hence uu(z)-u(z) is near zero for all such 2. This implies (nu-1b)- hl-w) is near zero for all 2 on the boundary. Since uu-¢ > 0 on the noncontact set of the unperturbed problem, thus away from F1 U T2 on the arc we have u-—¢ > 0. Hence u-—¢ 2 0 on the image of the semicircle under the map L vanishing only at 2p and -Eé. Now we want to show u-w > 0 everywhere in the domain bounded by T1, F2, the segment [-a(0),a(0)], and the image of the semicircle under L. Suppose not. Then u-—m takes its minimum at some point in the interior of the domain. At such a point V(u-—¢) = 0. Since 2 u(z)—(1(2) = 3 y3-2 Imj (151(0)de a(O) 3 ‘_u fl__ -1( 2: 5_u_fl_ 2_ —l 2- thus 5X'-ax — 21m[L 2)] 0, By By — 2y 2Re[L (2)] —(L This shows [L_l(z)]2 = y2, which means 2 on F1 U T2. This contradiction completes the proof. 137 Conclusions 5.25. Summary: We want to put the answer we got for the Problem 5.1 in a summary. Given 0 near 00 in Z, and parameters K, A = -d2, B = A2 fixed, as stated in the Remark 5.20, there is only one constant of integration to be chosen for the equation a”(y) = f(y) because the other one is determined by symmetry. More specifically we observed that a'(O) = O and a(O) > 0 was to be chosen. We need also to see this point by considering the initial data at yO (as we did already) in the following way. We shall see that choosing a'(yo) near §§(H(y) _% y log y)‘ , then (ii) and (iii) in the y=YO Remark 5.20 will determine d2, ¢(O), with ___J§fllL:fL____ > O, in terms of other parameters, where (_l)r¢(r+l) (O) ¢(r+l) is the first derivative of 0 nonzero at 0 and (iv) will give a(O). Then a(yO) will be completely determined by (v), that is, we have only one constant of integration, a'(yo), to be chosen. We stress that we choose a'(yO)-é%(H(y)-% y log y)‘ as close to y=Y 0 zero as we wish. We will show that a(yo) will then be close to x0 and the answer to the Problem 5.1 will be unique. Lemma 5.26. From (ii) in the Remark 5.20 we can . 2 . . find a as a function of K, 0 (or as a function of i J 138 2 °° 2j K, a , d , d ,..., where ¢(t ) = Z d.t ). Moreover 0 l 2 j=0 j o2 Yo + + we have asymptotically 7? = (log fi:)0 (l), where 0 is ./K the usual 0 with positive leading term. y 2 2 .. 0 g @(g ) Proof: By (ii) we have a’(y )+ dg = —— 0 i0 (§2+K)3/2 y 2 d2 I O __$U£L_L__ dg. Put g - éjK. Then 0 (C2+K)3/2 Yo Yo My H(fi _ifiwsidheijfi Agra)— d. >0, 0 o (g +1)3/2 K o (§+1)3/2 (*) where the integral on the right hand side stays away from zero because ¢ is near $0 in Z. This shows d2 as a function of K, 0 (or parameters K,do,dl,d2,...). To get the asymptotic representation, observe that for the integral on the left hand side we have Yo [m g2mg) dg o (g2+1)3/2 Yo 1 2 2 K 2 2 a (K) J.“ K) 1M—7d§+ M7d§ IO (§2+1)3 2 1 (§2+1)3 2 {q _ .1 gmegz) [12 1: J 2 3/2 dg + (pm) g o (g +1) 1 yo JR -2 2 .d +i (KEN—Oak: 1 E (§2+1)3/2 139 V0 1%dwioneg— (-3 +1) MK where the first integral on the right is pOSitive and K, and the integrand bounded (away from zero) for small in the second one is: §31¢1+—¢<0)[§3 -<§2+1)3/21 §<§2+1)3/2 But (§2+-1)2-§3§-——5—1%%%———— and therefore (g 2+1) + g 3% 2 2 d <1+§2)3 Yo _J‘ffi .2IW52 4101...; l (2)g+13/2 *< O ©(o)<3§4+-3§2+ 1)d g y 2 3/2] ' +J‘ 1 §<§2+1)3/2[§3 (é +1) where the second integral on the right hand Side is pOSitive For the first integral on the <1 2:) and bounded for small K. 140 p right hand side, by going back to the t variable we have y w j 0 [ ( dg, which is bounded for small k J? (g2 +K)3 (easy to check). Substituting these results back in (*) yields Y_o ./K 2 §(K§ ) zio (1+g2) 3 2 d3 (**) Hence asymptotically (as K approaches zero) we have: Yo ¢(o) 1og -:+ 0(1) = ./K leNI Y %? = (log —9)0+(l), where 0+ is the usual 0 with j}? the leading term of 0+(l) positive (as we see above). Remark. If we put ¢(t2) = ¢(0)e(t2), then (**) will have the form Y_o 2 ,,K y 3? 79—32%dgflog—91‘al-0‘1): 0 (1+: JR co , G. e(t2)= 213.133, B.=51. Lemma 5.27. From (iii) in the Remark 5.20 we can find do as a function of K, Bj’ for all integers . \ ,. Li— 1 H _.J; j 4 1, With B (_l)r+l 2 0, where a — GO and Br+l r+l is the first nonzero among Bl, B2, B3,... . Moreover 141 do = 1 if and only if Bj = O V j 2 l or K = O = d. . 1 _ yo + 3 Asymptotically we have l-—a— — BlK[(log ——)O (l)+75+ o R 0(K log ZS2)] if B # O and (1.--l—-)(_1)r = J? l 0‘o y Y Br+lKr+l[(l°g —9)0+(l)+-O+(l)*'O(K log -9)] if jg J? 51 = B2 = on. = fir = O and Br‘l‘l 7'1 OI r 2 l“ 2 (a2 2) 2 Proof: Since F j ‘——‘_:J;—_ ¢(C )dg = clucz (g2+K)3/2 2 2 2 2 2 2 2 a — ( ) . 2 (a -C >9 _ f dg T — (—dg), where W val (€2+_K)3/2 w fol _(€2*_K)3/2 g is changed to -Q in the last integral, thus 2 2 2 (a —g ) 2 . . f(g)dg = — ¢ (C) is the first nonzero coefficient ___2L;i___ among B1, 52, B3,..., then we must have = r+l Br+l(-l) r+l 2 O with equality case only when K = O. l--0(O C1r+1(-l) More precisely u = 1 if and only if all Bj = O or K = o, i.e., if and only if ¢ 5 1 or K = o = a = A = B. Moreover asymptotically we have (by (1) above) 3:; _ (O)—l Y Y K[(log -g)0+(l)-+%+'O(K 109 —%)]. JK JK if B1 2 o, (2) and 3 5 3 ‘_ -§ ._ (—~+r-—l) y 2: l)(_l)r+l 2 r1’2i)' Kr+l[(log -9)O+(1) r+l ‘ ' JK 3 r + 42 2 +—O(K log —Q)] (rfl)(2ml) ,R if Br+l is the first nonzero among 82, B3, B4,°.. . If all Bj' j 2 l, are zero, then u = 1, a0 = 1 = ¢(O) (no matter what K is — of course small). If Bl # 0, then by (2) we have asymptotically the graph 5.20—a. If Bl = Q = 62 = ... = Br but Br+l ; O for some r 2 1, then by (2) we have asymptotically the graph 5,20—b. Figure 5.20-a Figure 5,20—b Corollary 5.28o Given K and e as in the last Remark, Lemma 5.27 shows that u (i.e., G0 = ¢(O)) can be found uniquely in terms of K, Bj = 9(3)(O), j 2 1. By Lemma 5.26 then d2 can be found in terms of these '7 parameters° 7 yo Lemma 5929, a(O) = VK(log -:)0+ (1), Moreover UK a(y ) is really close to X 0 ‘ O' 146 Proof: By the formula given for a(O) in the Remark 5.20 we have N a(O) JR-(ijj K(i./I'<- mag) d; J? _ 2 2 2 2mg (1K-..) —92—1/2((31+§) J«o 214(2th Recall that the leading term of ¢(g2) is do = i = ¢(O Thus by substituting the value of d2 from Lemma 5.26 Y we get JK(log jg)0+(l) as the leading term for a(O): VK a(O) = fi+fid§ V o (K+g )3/2 + 3 IYO {—61—- - 11¢dg V o (K+g )3/2 + % §:O(¢3/21 2 Yo 2 2 +FJ‘O (¢(C )-l)dg+jfi_‘ YO . The last term hereis % yo, therefore we only need to compare the other terms with -% w(yo). The first and the second terms on the right hand side above are negligible because of coefficients and the bounded integrals Y for small K. For the last integral % f O(¢(g2)—¢(O))d§+ O 2 yo HO series expansion of ¢(g2). The first term is close (¢(O)-—l)dg, where $(O) = d by our last power 0 enough to -% ¢l(yo) (both with leading qubic term, at least, with respect to yo), and the second term g(dO--l)yO is negligible by the results of the Lemma 5.2.7. n Here is the proof complete. f Remark. As the above proof shows, the difference _ . r 3 of a(yo), xO — au(yo) is O(K )4-O(yo) for any 0 < r < %. # What we have proved so far can be summarized into the following (main result) theorem: .. o ' =:+'r ‘ I TheOiem 5.30. Fix zO <0 iyo, A0 > O yO / on the unperturbed free boundary (in Example 2.5). L1 149 Consider A, B, K near 0, K > O, and ¢ near $0 in z, and f(y) with the branch of f for which f(yo) < O, u = n I and a(yo)+iyo near 20 xO+-iyo, a (yo) near d dy(H(y)-—% y log y)| . Then a necessary and sufficient y=YO condition that there exists a symmetric local solution in 0, an open fixed disk (independent of parameters K,A,B) about the origin contained in the domain of the local solution in Example 2.5, with part of the free boundary II x = a(y), a (y) = f(y) Tl: a’(y)| = a’(yo). a(y)! = a(yo) yo yo near yO is (i) f(t) = 3 q2"t2 ¢(t2) where A = —d2 w (t2+K)372 B = a4, a > o . yo (ii) a'(yo) = f0 f(g)dg (iii) j f(g)dg = -21 fl , nyo (iv) a(yo) = a(O)4—yoa (yO)-J gf(g)dg near 0 X0, where 1 iv? _ a(O) = 3? gfmdg = fi-f (iv‘K—§)f(§)d<; '40 0 Moreover these four conditions are consistent and given K and 9 = $%67 ¢ fixed, do = ¢(O), d2 = -A, B = A2 150 can be found uniquely as a function of the given parameters (K,a'(0),e”(0),...,S(r)(0),...). In particular the set of admissible parameters for the above necessary and sufficient conditions is not vacuous. Furthermore given the above fixed parameters the local solution is unique with only one constant of integration: a’(yo) to be chosen near aé(yo). # LIST OF REFERENCES LIST OF REFERENCES [1] H. Brezis — D. Kinderlehrer, The smoothness of solutions to non-linear variational inequalities, Indiana Math. J., 23 (1974), pp. 831-844. [2] H. Brezis, Operateurs Maximaux Monotones, North Holland, Amsterdam, 1973. [3] L.A. Caffarelli and N.M. Riviere, Smoothness and analyticity of free boundaries in variational inequality, Ann. Scuda Norm. Sup. Pisa, Ser. IV 3, 289—310 (1976). [4] S.N. Chow — J. Hale, Methods of Bifurcation theory, Springer-Verlag 1982. [5] S.—N. Chow and J. Mallet—Paret, The parameterized obstacle problem, Nonlinear Analysis, Theory, Methods and Applications, vol. 4, No. 1, pp. 73-91. [6] R. Courant — D. Hilbert, Methods of Mathematical physics, volumes I, II. [7] M. Golubitsky, Stable Mappings and their singularities, Springer—Verlag, 1973. [8] D. Kinderlehrer - G. Stampacchia, An Introduction to variational Inequalities and their Applications, Academic Press, 1980. [9] D. Kinderlehrer, The free boundary determined by the solution to a differential equation, Indiana Univ. Math. J. 25, 195—208 (1976). [10] D. Kinderlehrer and L. Nirenberg, Regularity in free boundary problems, Ann. Scuola Norm. Sup. Pisa, Ser. IV 4, 373—391 (1977). [ll] Kinderlehrer, D., The coincidence set of solutions of certain variational inequalities, Archs. ration. Mech. Analysis 40, 231-250 (1971). 151 [121 [l3] [14] [15] [l6] [171 [18] [19] [20] [21] 152 H. Lewy and G. Stampacchia, On the regularity of the solution of variational inequality, Comm. Pure Appl. Math. 22, 153-188 (1969). H. Lewy and G. Stampacchia, On existence and smoothness of solutions of some noncoercive variational inequalities, Arch. Rational Mech. Anal. 41 (1971), 241—253. J. Mallet-Paret, Generic unfoldings and normal forms of some singularities arising in the obstacle problem, Duke Math. J. 46, 645—683 (1979). J. Mallet-Paret, Generic bifurcation in the obstacle problem, Quarterly Appl. Math. Jan. 1980, pp. 355—387. C. Morrey, Multiple integrals in the calculus of variations, N.Y., Springer—Verlag, 1966. Private communication with Professor J. Mallet—Paret. T. Rado, On the problem of plateau, Springer-Verlag, 1971. W. Rudin, Real and complex Analysis, McGraw—Hill, 1974. D.G. Schaeffer, Some examples of singularities in a free boundary, Ann. Scuola Norm. Sup. Pisa, Ser. IV, 4, 133—144 (1977). D.G. Schaeffer, A stability theorem for the obstacle problem, Adv. Math. 17, 34-47 (1975). SITY LIBRARIES VER lllll E|| 2 2 8 4 1 7 0 3 0 3 9 2 1 3 NIH) 1 IIHIWHIH