You are on page 1of 178
Lennart Carleson COMPLEX BYNAMICS he 6) Springer-verlag Lennart Carleson Theodore W. Gamelin Complex Dynamics With 28 Figures Springer-Verlag New York Berlin Heidelberg London Paris Tokyo Hong Kong Barcelona Budapest Preface Complex dynamics is today very much a focus of interest. Though several fine expository articles were available, by P. Blanchard and by M.Yu. Lyubich in particular, until recently there was no single source where students could find the material with proofs. For anyone in our position, gathering and organizing the material required a great deal of work going through preprints and papers and in some cases even finding a proof. We hope that the results of our efforts will be of help to others who plan to learn about complex dynamics and perhaps even lecture. Meanwhile books in the field are beginning to appear. The Stony Brook course notes of J. Milnor were particularly welcome and useful. Still we hope that our special emphasis on the analytic side will satisfy a need. This book is a revised and expanded version of notes based on lectures of the first author at UCLA over several Winter Quarters, particularly 1986 and 1990. We owe Chris Bishop a great deal of gratitude for supervising the production of course notes, adding new material, and making computer pictures. We have used his computer pictures, and we will also refer to the attractive color graphics in the popular treatment of H.-O. Peitgen and P. Richter. We have benefited from discussions with a number of colleagues, and from suggestions of students in both our courses. We would vi Preface particularly like to acknowledge contributions from Peter Jones and M. Shishikura. Any reader familiar with the area will recognize the exposition of quasiconformal mappings from Ahlfors’ book. It is often difficult to trace particular results to the rightful owners, particularly in such a rapidly developing area where so much seems to flow by word of mouth. We apologize for any inadequacy and for omissions. L. Carleson and T.W. Gamelin Los Angeles, March, 1992 Contents Preface I. Il. Conformal and Quasiconformal Mappings 1. Some Estimates on Conformal Mappings. ....... 2.The Riemann Mapping... . . 3.Montel’s Theorem ....... 4. The Hyperbolic Metric... . . 5. Quasiconformal Mappings . . . 6. Singular Integral Operators . . 7. The Beltrami Equation... .. Fixed Points and Conjugations 1. Classification of Fixed Points . 2. Attracting Fixed Points ... . 3. Repelling Fixed Points... .. 4. Superattracting Fixed Points . 5. Rationally Neutral Fixed Points 6. Irrationally Neutral Fixed Points ............ 7. Homeomorphisms of the Circle 27 27 31 32 33 35 41 47 viii III. IV. VI. VII. VIII. Contents Basic Rational Iteration 53 1.The Julia Set 2... es 53 2. Counting Cycles .. 1... . ee ee ees 58 3. Density of Repelling Periodic Points .........-.- 63 4.Polynomials . 2... 6 ee 65 Classification of Periodic Components 69 dSullivamis theorem 0 69 2. The Classification Theorem .......-----205 74 3. The Wolff-Denjoy Theorem ........---.-++-: 79 Critical Points and Expanding Maps 81 1. Siegel Disks... 0 0 ee 81 2.Hyperbolicity 6. 6. ee 89 3.Subhyperbolicity . 2... ee 91 4. Locally Connected Julia Sets 2... 2... 0020 es 93 Applications of Quasiconformal Mappings 99 1. Polynomial-like Mappings ......--------5- 99 2.Quasicircles ©... ee 101 3.Herman Rings... . 2... - eee ee es 103 4. Counting Herman Rings ........----++--- 105 5. A Quasiconformal Surgical Procedure ........- 106 Local Geometry of the Fatou Set 109 1. Invariant Spirals... 2... ee ee 109 2.Repelling Arms... 22. - ee ee ee 113 3. John Domains... 2... 2-6 ee es 117 Quadratic Polynomials 123 1. The Mandelbrot Set... 2.20.0. eee ee eee 123 2. The Hyperbolic Components of M .......---- 133 3) Greens PUNCtION Of Je 6 136 4.Green’s Function of M.... 2... .0-2 0+ eee 139 5. External Rays with Rational Angles .........- 142 6. Misiurewicz Points 7. Parabolic Points Epilogue References Index Symbol Index Contents ix 161 163 171 175 I Conformal and Quasiconformal Mappings We discuss some topics that are not included in the standard in- troductory graduate course in complex analysis. For the most part, we assume only the background provided by elementary graduate courses in real and complex analysis. There are a few exceptions. We use covering spaces and at one point the uniformization theorem, which can be found in Chapters 9 and 10 of [A3]. A readable discus- sion of the Poisson kernel and Fatou’s theorem is given in Chapter 1 of [Ho]. In discussing the Beurling transform we appeal to the Calderén-Zygmund theorem, which can be found in [St] or [A2]. 1. Some Estimates on Conformal Mappings A mapping f is called conformal if it is analytic and one-to-one. Such mappings are also called univalent. Let S be the collection of univalent functions in the open unit disk A = {|z| < 1} such that f(0) =0 and f’(0) = 1. The compactness properties of this class are of crucial importance for our study. THEOREM 1.1 (area theorem). Let g(z) = 1/z+bo + biz+... be univalent in A (with a pole at z = 0). Then Xyn|bn|? < 1. 2 I. Conformal and Quasiconformal Mappings Proof. For 0 nar). r n=l Since area D, > 0, taking r — 1 gives the result. 0 THEOREM 1.2. If f(z) = 2+ D822 an2” €S, then |a2| < 2. Proof. Define g(z) = 1//f(z?) = 1/z—@22/2+---. If g(z1) = g(22), then f(z?) = f(#), a= 23, and z; = +22. But g is odd, so 11 = 22. Hence g is univalent. The area theorem gives |a2| < 2. 0 THEOREM 1.3 (Koebe one-quarter theorem). If f € S, then the image of f covers the open disk centered at 0 of radius one-quarter, that is, f(A) > A(0, 1/4). Proof. Fix a point c, and suppose f # c in A. Then ai. =2z+ (a+ 5)22+- belongs to S. Applying Theorem 1.2 twice, we obtain. 1 << |a|+ le| 1 a+ 2|<242=4 a ¢c The Koebe function f(z) = z/(1— 2)? = nz” maps the disk A to the slit plane C\(—00, —1/4]. This shows that one-quarter is optimal. The Koebe one-quarter theorem and the Schwarz lemma combine to give 1 ne dist(0,0f(A))<1, feS, where “dist” denotes distance. To prove the upper estimate, note that the image of f € S cannot cover the closed unit disk, or else f~! would map A to a proper subdomain, contradicting f'(0) = 1. If we 1.1. Some Estimates on Conformal Mappings 3 translate and scale, we obtain dist(f (zo), Of (A(Go, 6))) > 6|f’(z0)|/4 from the lower estimate whenever f is univalent on the open disk A(z, 6). This leads to the following. THEOREM 1.4. If f is univalent on a domain D, and 2 € D, then 1 : : 5 qf’ 20) ldist(20, 0D) S dist(f (zo), A(f(D))) < 4If"(z0)|dist(z0, 0D). Proof. The lower estimate follows immediately from the statement preceding the theorem, and the upper estimate is obtained by ap- plying the lower estimate to f~! at the point f(z). 0 We aim now at proving some stronger distortion results. THEOREM 1.5. If f € S, then zf"(z) le? 4lz| f(@) 1-[eP> 1-feP Proof. Fix ¢ € A and consider f((2+.0/0.+ €)) - FO) F Se eee “) = PFO Then F € S, so by Theorem 1.2, |a2(¢)| < 2. A computation shows aa(6) = 5{(1- Ete - 2c}. 2 =ztag(C)jz*t+---. The assertion follows. 0 THEOREM 1.6 (distortion theorem). If f € S, then 1- 1+ |z| aor SOS Gas mm A ll a ae alt -|2)2, eA. Proof. We apply the Koebe one-quarter theorem (or the earlier dis- tance estimate) to the function F' appearing in the proof of Theorem 1.5, to obtain dist(f(z), f(@A)) 2 ila -\2?), 264. The lower estimate for |f’(z)| in the distortion theorem then gives the result. 0 THEOREM 1.8. If f(z) = 2+ D222 an2" €S, then |an| < en’. Proof. Using the distortion theorem to estimate |f(z)|, we obtain lan | = af F@) 4, < Qi Jigar 241 | ~ (1-17)? 1-n O0 D extends continuously to the closed disk A. Proof. First suppose extends continuously. Let z € OD, and let {Zn} be a sequence in OD converging to z. Taking Cn € w+ (zp) and passing to a subsequence, we assume ¢, — ¢. By continuity, #(¢) = z. The arc yn = W([Cn,¢]) is a connected subset of OD containing 2n and z. Since 7 is uniformly continuous, diam(7,) — 0. Hence OD is locally connected at z. For the other direction, suppose OD is locally connected and as- sume 00 = 7(0) € D. Fix ¢o € OA. For p > 0 consider yp = {C € A: IC — Go| = p}. Define L(p) = I \w"(Olla|, the length of the image curve (7p) in D. By the Cauchy-Schwarz inequality L(p)?

0. Now I, separates D into two connected components, one containing oo = 7(0). Let Dn be the bounded component of D\I'n. We claim that D, is contained in a bounded component of C\(I',ULn). Indeed, otherwise we could draw a simple arc from a fixed point zo € Dn to oo in C\([nULn), followed by another arc from oo to 29 in D crossing I’, exactly once, to obtain a simple closed Jordan curve in C\L, which separates a, and Bn, contradicting the connectedness of Ly. Since diam(I, U Ln) — 0, also diam(D,,) — 0, and hence 7 is continuous at Co. Let D be the domain at the left of Figure 1, whose boundary includes a union of spikes emanating from 0 at rational angles, so that (say) the spike at angle 2xp/q has length 1/q. Since OD is locally connected, the Riemann mapping ~ of A onto D extends continuously to OA. However the point 0 € OD is the inverse image under w of a Cantor set on the unit circle. Thus even a continuous Riemann mapping can have “bad ” behavior. Recall that a Stolz angle at Co € OA is a sector in A with vertex at and aperture strictly less than 77, bisected by the radius (see Figure 2). The property of Stolz angles we require is that if u is a harmonic function, u > 0, such that lim inf u(¢) > 1 as ¢ tends to a boundary arc on one side of ¢) with endpoint at Co, then liminf u(¢) > « > 0 as ¢ tends to Cp through any fixed Stolz angle. This is easiest to see in the upper half-plane, where a Stolz angle corresponds to a sector with vertex 29 € R of the form {n < arg(z — x0) < —n}. Here one compares u with the harmonic function arg(z — xo) to obtain the estimate with ¢ = n/n. 8 I. Conformal and Quasiconformal Mappings For the local study of a conformal mapping at a boundary point zo, we will appeal several times to the following immediate consequence of two well-known theorems of Lindeléf. THEOREM 2.2. Let D C © be simply connected, suppose OD has more than one point, and let (¢) map A conformally to D. Let y be a Jordan are in D except for one endpoint z) € OD. Then the curve w-! oy terminates in a point Go € OA, and ¥(¢) > 20 a8 ¢ > Co inside any Stolz angle at Co. Proof. Let Tbe the curve #~!(7\{zo}) in A, which clusters on OA. We first want to prove that I’ has a unique cluster point 9 on OA. For this we could use Fatou’s theorem, but let us give a more geometric proof. We wish to use some conformal invariant and let us consider a Dirichlet integral. We may assume D is bounded. Define 1 1/3 (2) = (tog* aoales) ; A simple computation shows that independently of < > 0, p(s) = ff \wsPdedy (og ye -c, (C=F(0), for 0, <0 < 69. Using the Schwarz inequality we find 82 1 bh Ch 0 which contradicts D(F) < M. Thus accumulates at only one point Co of the circle, and we can assume that Co = 1. OF ag! 1\?/3 — > = _ a (re’’)| rdr > 6( tog :) 2nC, 1.3. Montel’s Theorem 9 FIGURE 2. Stolz angles. For p < 1 let I’, be the subarc of T' lying in the annulus p < |¢| < 1 with one endpoint ¢, on the circle |¢] = p and the other at the terminal point 1 of I. Then I’, divides the slit annulus {p < |¢| < 1}\(—1, —p) into two components. Let U be (say) the top component. Let w(¢) be harmonic in U, equal to 1 on I’, and equal to 0 on the rest of OU. Let = e(p) be the supremum of |*)(¢) — zo| over ¢ ET, , so that ¢ > 0 as p > 1. Assume as we may that D C {\z| < 1/2} so that |y(¢) — zo| < 1. By the maximum principle and the fact that log |b(¢) — zo] < 0 on U, log |(¢) — 20] < w(¢) loge in U. We obtain a minorant w*(¢) of w(¢) by replacing I’, by the arc A from 1 to —1 along the bottom half of {|¢| = 1} and defining w*(¢) harmonic in the ring R = {p < |¢| < 1} slit along (—1,—p). Then w*(¢) = 0 on OR except on A where w* = 1. The estimate above now holds in U, with w* replacing w. According to our preliminary remarks, lim inf w*(¢) > c > 0 as ¢ — 1 inside any fixed Stolz angle S. Hence lim sup |4h(¢) — zo] < €°, S361 so that w(¢) — 20 as asserted. 0 3. Montel’s Theorem It was P. Montel (1911) who formulated the notion of a normal family of meromorphic functions and realized that the modular function 10 J. Conformal and Quasiconformal Mappings could be used to derive the criterion for normality which bears his name. This provided a key ingredient for the work of Fatou and Julia on complex iteration theory later in the decade. Let R and S be Riemann surfaces. A holomorphic map P : R > Ss is a covering map if every w € S lies in a coordinate disk U such that each connected component of P~!(U) is mapped conformally by P onto U. Let S© be the universal covering surface of S, obtained from S by regarding all loops around boundary points as nontrivial, and let P : S% — S denote the universal covering map (see [Sp],[A3)]). We will use the fact that any map f from a simply connected surface D to S can be lifted to a map f:D—S®, satisfying Po f ae The value f(z0) € P~'(f(20)) can be specified arbitrarily, and this determines the lift f uniquely. The other lifts of f are the maps yo f ; where ¢ is a covering transformation, that is, a conformal self-map of $© satisfying Poy = P. One consequence of the lifting property is that if f is itself a covering map of a simply connected surface D onto S, then the lift f : D > S® is one-to-one and onto, so that D is conformally equivalent to S°°. THEOREM 3.1. If D = C\{0,1} is the thrice-punctured sphere, then D®& is conformally equivalent to the open unit disk A. Proof. Since A is conformally equivalent to the upper half-plane H = {y > 0}, it suffices to find a covering map of H onto D. We construct such a map, the so-called modular function, as follows. Let E = {z: 0 1/2}. In view of the Riemann mapping theorem, there is 7 mapping E to H fixing 0, 1,00. Let E* denote the reflection of E through the circle {|z — 1/2| = 1/2}. By the Schwarz reflection principle, we can extend 7 to a conformal mapping of EU E* to C\(—00, 0] U [1, 00). By continuing to reflect we can extend 7 to all of {0 < x < 1,y > 0} taking its values in C\{0, 1}. By reflecting across the vertical lines {x =n} for n an integer, we can extend 7 to all of H. From the construction one sees that the extended 7 is a covering map of H over D. 0 THEOREM 3.2 (Montel’s theorem). Let F be a family of meromorphic functions on a domain D. If there are three fixed values that are omitted by every f € F, then F is a normal family. Proof. We may assume that D is a disk, and, by composing with a 1.4, The Hyperbolic Metric 11 Mobius transformation, we may assume that the functions in F do not assume the values 0, 1,00. Let S = C\{0,1}. By Theorem 3.1, there is a covering map y : A > S. Let f : D > A be a litt of f EF, so that fow =f. Then {f: f € F} is a normal family, by Theorem 1.9, and this implies that F is normal. O The uniformization theorem states that every simply connected Riemann surface is conformally equivalent to exactly one of the unit disk A, the complex plane C, or the Riemann sphere C. Moreover, every Riemann surface has one of these three as its universal cover. The three cases are referred to as hyperbolic, parabolic, and elliptic, respectively. For a proof of the uniformization theorem see [A3]. We will require the following criterion for a surface to be hyperbolic. THEOREM 3.3. Let R be a Riemann surface that has a nonconstant meromorphic function omitting at least three values. Then R®, its universal covering surface, is conformally equivalent to the unit disk, with any given point corresponding to the origin. Proof. Let P : R® — R be the universal covering map. If f on R omits three values, then fo P is a nonconstant meromorphic function on R® that omits three values, say {0, 1,00}. Using a covering map of A onto C\{0,1}, we can lift f o P to a nonconstant function from R® to the the open unit disk A. Thus R® is not conformally equivalent to C or C, and consequently R is hyperbolic. 0 4. The Hyperbolic Metric A conformal mapping of A onto itself has the form w = e(z — a)/(1 — Gz) for some 0 < 6 < 2m and |a| < 1. A computation yields dw|_ 1—|w/? az es lzien from which we see that 2\|dz| 2|dw| dp = —— = ——— °“7-—eP 1—TwP is invariant under the mapping. The metric dp is called the hyperbolic (or Poincaré) metric on A. From the hyperbolic metric, we get a 12 I. Conformal and Quasiconformal Mappings distance p(z1,22) in the obvious way by integrating along curves from z; to zg and taking the infimum. One checks that 1+(|z| ica 7 p(0, 2) = log Using a covering map P : A — S, we define a hyperbolic met- tic dps on any hyperbolic Riemann surface S by declaring that P induces an isometry at every point. In other words, we set 2\dw dps(z) = ana z=P(w)eS. If y is a local determination for P7', this formula becomes dps(z) = i- wep zes. Since dp is invariant under conformal self-maps of A, this is indepen- dent of the choice of the branch y of P~', and it is also independent of the choice of the covering map P. As an example, using the con- formal map ¢(z) = (z — 4)/(z +4) of the upper half-plane H onto the unit disk A, we compute the hyperbolic metric of the upper half-plane to be doate) =, z=atiy,y>0. A holomorphic map f : R — S allows us to pull back differentials on S to differentials on R by the obvious formula (chain rule). Thus we can pull back the hyperbolic metric dps on S to a metric on R, which we denote by f*(dps). THEOREM 4.1. Suppose f maps a hyperbolic Riemann surface R holomorphically into a hyperbolic surface S. Then f*(dps) < dor, ps(f(z1),f(z2)) < pr(zi,z2), mee R, with strict inequality unless f lifts to a Mébius transformation map- ping A onto A. 1.4. The Hyperbolic Metric 13 Proof. The second estimate is just the integrated version of the first. In the case R = S = A, the first boils down to G. Pick’s (1915) invariant form of the Schwarz lemma, _ 2 ey < zea. This is obtained from the Schwarz lemma by pre- and postcomposing with the appropriate Mobius transformations. For the general case, we check the estimate at a fixed point zp € R. Let wo = f(z0), and consider covering maps P: A Rand Q: A S satisfying P(0) = 20, Q(0) = wo. We lift f to an analytic function F : A > A satisfying QoF = foP, F(0) =0. It suffices to check that F*(dp(¢)) < dp(z) at z= 0. Now ae FAS) = Reyes FEM so the estimate becomes |F’(0)| < 1. This holds, with equality only when F maps A conformally onto A. 0 THEOREM 4.2. Suppose RC S. Then dppr > dps with strict inequal- ity unless R= S. Proof. Apply Theorem 4.1 to the inclusion map. 0 THEOREM 4.3. Let D be a domain in C, and for z € D, let 6(z) denote the distance from z to OD. Then if D is simply connected, 5a < dpp(z) < oat. (4.1) For general domains as z — 0D, 1+ 0(1) |dz| 5(z) log(1/6(2) 2! < dpp(z) < 25)" (4.2) Proof. For the right-hand sides, observe that D’ = A(z, 6(zo)) C D, so by Theorem 4.2, dpp(zo) < dpp:(z) = 2|dz|/6(zo). For the the left-hand side of (4.1), let y» map A to D with ¥(0) = z. By the Koebe one-quarter theorem, 7 takes all values in A(zo, |y‘(0)|/4), and hence 6(z0) > |~'(0)|/4, which gives the desired inequality. 14 I. Conformal and Quasiconformal Mappings To prove the left-hand side of (4.2), let 21 be a point in OD closest to zo and choose two other points z2 and 23 € OD. Assume z; = 0, z = 1, and 23 = oo. Let D’ be the complement of {0, 1,00}. Then dpp > dpp’, so we only have to estimate dpp: near z; = 0. Let y:H- D’ be the modular function, constructed in the proof of Theorem 2.1. Then p(w + 2) = p(w), so that w can be expanded as a Laurent series in e’™”. Since ~ maps a neighborhood of too in the strip {0 < Rew < 2} ontoa punctured neighborhood of 00, w has a simple pole, and z(w) = 1/(w) has a simple zero, 2(w) = bye + dboe2i™™ 4 ---, Imw > 0, where b, # 0. Solving for w, we obtain w(z) = e log z + analytic, a 1 + analytic. Tt dz wiz Now z(w) is a covering map of H over D’, so it can be used to express the hyperbolic metric of D’ in terms of that of H, (dw| dw |dz| |dz| [i+ O(1) Imw dz|Imw(z) __|2|log(1/lzI) log(1/|z|)}’ from which the estimate follows. 5 dpp(2) = We end this section with an estimate for Green’s function which will be useful for the study of the Mandelbrot set. The estimate, which appears in [CaJ], is a simple variant of estimates for harmonic measure and Green’s function in the thesis of A. Beurling (pp. 29-30 of [Beu]). THEOREM 4.4. Let D be simply connected, and let 6(z) be the dis- tance from z € D to OD. Let G(z, 20) be Green’s function for D with pole at 2 € D. IfzeD satisfies G(z, z0) <1, then there is an arc ¥ from z to z inD such that G(z, 20) 0 Selz0)=-+ ff So ag an "ds In view of the identity dgdn 2 a< |¢j/ 1 as p — 2. These estimates allow us to extend S to all of L”, for 1 < p < oo. The usual convolution argument with an approximate identity shows that if a distribution f has distribution derivatives f, and f; in L? for some 1 < p < ov, then (the extended) S satisfies S(fz) = fz. A proof of these facts is included in Ahlfors’ exposi- tion [A2] of quasiconformal mapping. For a detailed exposition of singular integral operators, see [St]. IA 7. The Beltrami Equation Fix 0 < k < 1, and let L*(k, R) denote the measurable functions on C bounded by k& and supported in A(0, R). We let QC1(k, R) denote the continuously differentiable homeomorphisms f of C such that fz = uf, for some wp € L©(k,R), normalized so that f(z) = z+ O(1/z) as z — oo. Note that f is analytic for |z| > R. Lemma 7.1. If f € QC1(k, R), then the inverse function f~! belongs to QC!(k, 4R). Proof. This follows directly from the Koebe one-quarter theorem and the formula for the Beltrami coefficient of f~! given in Section 5. 0 Let f € QC1(k, R), and consider the Cauchy-Green formula for a disk, fi=— #) d¢ - all... BO adn, l2| R the O(1/z) term does not contribute to the first integral. We can replace f(¢) by ¢ in the first integral, and this leads to oe * Moone = £09) dean =Tf;(z). (7.1) Thus if we set g = f, — 1 and use fz = pfz, we obtain = S(fz) = S(ufz) = S(ug) + S(H)- In terms of the operator U,(g) = S(ug), 9 EL, this equation becomes (1 - U,)9 = S(w)- The operator U,, has L?-norm ||Uy\|p S kKp < 1forp > 2 sufficiently close to 2. We fix p = p(k) > 2, once and for all, so that kK, <1. Then ||U,,|| < 1 and J —U, is invertible, with inverse oo (I-U,)" = X un, r-U,)M < Sl)" = 1—-kKp Thus we can solve for g, to obtain a= (1 -U,)8(u) € L?. (7.2) This leads to L?-estimates for g, -1 (aR?) lIgllp < WZ — Un)" IIpll Sllpllelle S aR Ky (7.3) Since the Jacobian of f is given by Js = |fel? — [fel? = (1 le? )1feI?, ~ 1.7. The Beltrami Equation 21 Jy belongs to L?/? on bounded sets. This has the following conse- quence. THEOREM 7.2. Let p = p(k) be as above. If f € QC(k, R) and Ec A(0, R), then area f(E) < C(areaE)!~*/?, areaE < C(area f(E))!~?/?, where C' depends only on k and R. Proof. By Hélder’s inequality applied with exponent p/2, aren f(E) = ff Stray < (ff, Péody) “Crea Since f, = 9 +1, the L?-estimate (7.3) for g gives the first estimate. Lemma 7.1 permits us to apply the first estimate to f~', and this yields the second with a larger constant. 0 We can also prove a uniform estimate on the modulus of continuity of f. THEOREM 7.3. Let p = p(k) be as above. If f € QC1(k, R), then [fla) — f(@)| < Cla - 2|*??”, lz, [za] < R, where C' depends only on k and R. Proof. We use the formula (7.1) and find |z1 = 2a|% 1 1/q Ls) ~ flea) < lax ~ 2al+ ZIltll ff oan an) where q = p/(p—1) is the conjugate index of p. We split the integral into three pieces Dy = {|¢ — zi| < $|z1 — zal}, De = {| — 29] < sla — zg|}, and D3 = C\(D, U Dg). Since |¢ — z2| > |z1 — z2|/2 on Dj, the integral over D, is controlled by I d&dn E I rdrd@ _ ela — 29/29. IC — le ie Di OSr< la — ze] 22 I. Conformal and Quasiconformal Mappings The integral over D2 is treated similarly. The integral over D3 is controlled by rdrd@ req lan —2al0, andp= p(k) > 2 as above. For p € L™(k,R), there is a function f on C, normalized so that f(z) =2+O(1/z) at co, with distribution derivatives satisfying the Beltrami equation fz = pfz, and such that fz and f, — 1 belong to L?. Any such f is unique. The solution f is a homeomorphism of C, which is analytic on any open set on which wp = 0. If wé C? and Hz € C}, then f eC’. Proof. The proof of existence is now easy. Define g € L? by (7.2), and define f(z) =2+Tg +H). Since T is a convolution operator with kernel 1/z locally in L}, f is continuous. Moreover, f is normalized at oo, and fe = both, fe = 14+S(ng +H) =1t9 in the sense of distributions, so f satisfies the Beltrami equation. The defining formula for T shows that f is analytic on any open set on which p = 0. To prove the uniqueness, suppose F is another solution of the Beltrami equation with Fz, F,—-1 € LP. Then G = F,—1 satisfies G = S(F;). (We use here the fact that if h has distribution derivatives in LP, then h, = S(hz).) As before we obtain G=(I-U,)7!S(u), so G coincides with g, F, with f,, and Fz with fz. Thus F = f+c, and the normalization at oo gives c = 0. Now suppose pz is smooth. Following [A2], to obtain the smooth- ness of f we consider the differentiated Beltrami equation (fez = W(fz)2 + Meds - If f were C} we could write y = log f, and the equation would become ys = Hz + Hz- Using the Cauchy-Green formula as before, 1.7. The Beltrami Equation 23 applied to ¢ instead of f and using y(co) = 0, we obtain y = T(y3;). Consequently yp, = S(yz) = S(uyz) + S(uz), and (I - Uy) pz = S(uz). This is the same type of equation as we dealt with above. We now wish to run this argument backwards. Define b= (I-U,) 'S(uz) € L?, which satisfies v= S(up) + S(uz). Define gp =T (uv + pz). Since py + wz € L” is supported on A(0, R), the proof of Theorem 7.3 shows that y is Hélder continuous with exponent 1 — 2/p. The distribution derivatives of y are oz = woth, gz = S(up+puz) =, which belong to L?. Thus if h = e, we are justified in calculating the distribution derivatives of h to be h, = e®y, and hz = e?yz = e? (ups + Wz) = how + hu, = (hu)z. Hence hdz + (hu)dz is exact, and fl) = [ hdz + (hu)dz is well-defined. Since y (hence h) is continuous and yp is C!, f is C!. Moreover, we have f, = h and fz = hu, so fz = pf, as desired. For |z| > R, wh = 0 so that f is analytic there. Since y(oo) = 0, h(z) = 1+ O(1/z) at oo, and f(z) — z is analytic at oo. Adding a constant to f, we can arrange that f(z) = z+ O(1/z) at oo. Also Jp = (1 = |MP?)IA)? = (1 — |n?)1e?| #0. Thus f is locally one-to-one, and f is one-to-one near oo, so f is glob- ally one-to-one, and f is a homeomorphism of the Riemann sphere to itself. Thus f € QC!(k, R). To complete the proof it suffices to show that f is a homeomor- phism, even when y is not smooth. For this we approximate yu point- wise a.e. by smooth Beltrami coefficients un € L™(k, R), with cor- responding fn € QC1(k, R). On account of the uniform Hélder es- timates of Theorem 7.3, the sequence f,, is equicontinuous, and we 24 J. Conformal and Quasiconformal Mappings may assume f, converges uniformly to F on C. By Lemma 7.1 and Theorem 7.3 the inverses f,! are also equicontinuous, so f, 1 con- verges uniformly to an inverse for F, and F is a homeomorphism. Let gn = (I — Upn)~1(S(un)) as before. Choose s > p so that kK, < 1. Everything we have done works in L* as well as in L?, and in particular (J — Uyn)? is defined and bounded on L*, so that gn is bounded in L*. Since py — p ae. (and since pn € L~(k, R)), En converges in L” for all r < 00, and also fingn converges in L’ for all r <8. Thus gn = S(tngn) + S(t) converges in L?, say to g. Since fn @ F uniformly, and (fn)z = gn +1, we see that F, = g+1 in the sense of distributions. Also (fn)z = Hn(fn)z converges to 1g in L?, so F; = pg in the sense of distributions. Thus F coincides with the unique solution f of the Beltrami equation with L? derivatives, and we are done. 0 We let QC(k, R) denote the family of homeomorphisms f of C that are normalized solutions of the Beltrami equation with Beltrami coefficient js € L°(k, R) as in Theorem 7.4. The proof of Theorem 7.4 shows the following. THEOREM 7.5. Suppose fn € QC(k, R) has Beltrami coefficient Un, and f € QC(k, R) has Beltrami coefficient . If Un > p a.e., then fn — f uniformly, and (fnr)z — 1 and (fn)z converge respectively to fe—1 and to fz in L? forp= p(k) as above. Furthermore the inverse functions f,,! converge uniformly to Puc Theorem 7.5 shows that f depends continuously on any continu- ous parameters. We also have analytic dependence of f on analytic parameters. THEOREM 7.6. Suppose that y:(z,t) € L®(k, R) depends analytically on one or several parameters t, for each fixed z € C. Then the cor- responding f(z,t) also depends analytically on t. Proof. First note that if u(z,t) depends analytically on t for each fixed z, then on account of the uniform bounds on 1 and the uniform estimates provided by the Schwarz inequality, the functions move analytically in the Banach space L® with the parameter t. Since the supports of y:(z,t) are uniformly bounded, the functions p.(z,t) also move analytically in L? with t. Thus g = (J — U,)71S(u) moves ~ 1.7. The Beltrami Equation 25 analytically in L? with t, and T(g+ 4) moves analytically with t, in the norm of uniform convergence on C. Hence f(z, t) = z+T(ug+p) depends analytically on t. 0 We complete this chapter with some remarks. The uniform estimates derived in Theorems 7.2 and 7.3 clearly apply to f € QC(k, R) as do the computation rules for quasicon- formal mappings and also Lemma 7.1. Also, we can assign to any yu € L®(k, R) a measurable ellipse field, and the arguments involv- ing ellipse fields carry over to f € QC(k,R), as can be seen by approximating by smooth Beltrami coefficients. We have normalized our quasiconformal maps so they look like z+ O(1/z) near infinity. Given y, the corresponding f is unique up to compositions with Mébius transformations, so instead of this normalization we could normalize by fixing any three points of the sphere. We do not have to restrict ourselves to functions which are holo- morphic in {|z| > R}. By composing with a Mobius transformation, it will always be enough to simply assume f is holomorphic in some disk. Although this restriction can also be removed (see [A2]), in our applications this condition will always be satisfied. The way we will use quasiconformal mappings to study dynamical systems is as follows. Suppose g is a smooth function with certain dynamical behavior, and suppose we can construct an ellipse field E that is invariant under g, corresponding to a measurable Bel- trami coefficient py with |u| < k < 1. Let y solve the corresponding Beltrami equation, and set f = yogog™!. Then f has the same dynamical behavior as g, and moreover f is analytic. Indeed, y~! maps infinitesimal circles to E, g maps E to E, and y then maps E back to infinitesimal circles, so that f maps infinitesimal circles to infinitesimal circles, and f is analytic. II Fixed Points and Conjugations The study of complex dynamical systems begins with the descrip- tion of the local behavior near fixed points. We are concerned with the existence of canonical coordinate systems at fixed points. The coordinatizing functions play an important role, both locally and globally. 1. Classification of Fixed Points Suppose zo is a fixed point of an analytic function f, that is, f(zo) = zo. The number \ = f’(zo) is called the multiplier of f at zo. We classify the fixed point according to 4 as follows: Attracting : |A| < 1. (If A = 0 we refer to a superattracting fixed point.) Repelling : |A| > 1. Rationally neutral : |A| = 1 and A” = 1 for some integer n. Irrationally neutral : |A| = 1 but A” is never 1. It will be convenient to denote the iterates of a function f by Ge fond fae feof 28 Il. Fixed Points and Conjugations Suppose zo is an attracting fixed point for f. If |A| < p <1, then |f(2)—20| < plz—20| on some neighborhood of zo. Then | f"(z)—z0| < p|z — zo|, and consequently the iterates f” converge uniformly to zo on the neighborhood. We define the basin of attraction of 20, denoted by A(zo), to consist of all z such that f"(z) is defined for all n> 1and f(z) — zo. Thus for ¢ > 0 small, A(zo) coincides with the union of the backward iterates f~"(A(zo,€)), and consequently A(zo) is open. The connected component of A(zo) containing 2 is called the immediate basin of attraction of zo and denoted by A*(2)- We say that a function f :U — U is (conformally) conjugate to a function g : V — V if there is a conformal map ¢ : U — V such that g = yo fog}, that is, such that (f(z) = 9((2))- (1-1) The maps f and g can be regarded as the same map viewed in different coordinate systems. The definition implies the iterates an and g” are also conjugate, g” = po fog}, as are au andgpe when defined, g7! = yof to vy}. Note that maps fixed points of f to fixed points of g, and the multipliers at the corresponding fixed points are equal. A basin of attraction for f is mapped by a conjugating map ¢ onto a basin of attraction for g. Suppose we are given f(z) = 20 + Mz — 20) + A(z — 20)P + with p > 2. Near 2p the function f “looks like” g(C) = A¢ in the case \ # 0, and like g(¢) = aC? in the case A = 0, a # 0, where ¢ = 2-2. Does there always exist ap conjugating f to g? It turns out that the answer depends on f and in particular on the multiplier . We will show that a conjugation exists in the cases of attracting and repelling fixed points. In the case of an irrationally neutral fixed point, a conjugation exists unless \ is “very close” to roots of unity. In the case of a rationally neutral fixed point, a conjugation does not exist in general, but a conjugation to another canonical form exists in a large domain with the point as a cusp on its boundary. If \ # 0 and 4 is not a root of unity, then the conjugation » is unique up to a scale factor. To prove this, it suffices to show that any conjugation of f(z) = Az to itself is a constant multiple of z. Suppose y(z) = az + agz? +--+ is such a conjugation, so that v(Az) = Ap(z). Substituting power series and equating coefficients, IL.1. Classification of Fixed Points 29 we obtain a,\” = Aan, so that a, = 0 for n > 2, and y(z) = az. In the superattracting case 4 = 0, the same method shows that any conjugation to ¢? is unique up to multiplication by a (p — 1)th root of unity. The functional equation of a conjugation of z? to itself is (2?) = y(z)?, and comparing power series we find that a where a} = @1. When ) is an nth root of unity, any conformal map of the form p(z) = zh(z") = biz + bnyi2z"t! +--+ conjugates Az to itself. In particular, the identity map z with multiplier \ = 1 is conjugated to itself by any conformal y. We must consider other normal forms. The idea of conformal conjugation was introduced by E. Schréder (1871) to study iteration of rational functions. Equation (1.1) and its variants are referred to as Schréder’s equation. There was some continuing interest in the iteration of rational functions in connec- tion with algorithms for approximating roots, specifically Newton’s method. Schréder was interested in finding effective methods for computing iterates. Schréder (1870), and also A. Cayley (1879), found the basins of attraction for Newton’s method applied to quad- ratic polynomials (see the example to follow), and both mention the cubic case as an interesting problem. (According to Douady (1986), it was a question from a student in 1978 about the convergence of New- ton’s method that initially aroused the interest of J. Hubbard, and “by contamination” of Douady himself, in rational iteration theory.) For color pictures of domains of attraction associated with Newton’s method applied to certain cubic polynomials, see Maps 66, 77 and 78 on pages 91, 116 and 117 of [PeR]. Schréder emphasized how addition and multiplication formulae for trigonometric and elliptic functions give rise to conjugations. For in- stance, the double angle formula for tangent shows that 2 = —itan¢ conjugates 22/(1+ z”) to 2¢. The double angle formula for cosine is behind the following example. EXAMPLE. The polynomial P(z) = 2?—2 has a superattracting fixed point at oo. Consider the conformal map h(¢) = ¢ + 1/¢ of {\¢| > 1} onto C\[—2,2]. The identity P(A(¢)) = A(¢)? — 2 = h(C?) gives h-10 Poh = C?, and P(z) is conjugate to ¢?. Thus the dynamics of P(z) on C\[—2, 2] are the same as those of a rLrt—~—S~™=S the iterates of any ¢, |¢| > 1, under ¢? tend to oo, so do the iterates under P of any z € C\[—2,2]. Evidently [—2, 2] is invariant under P, so the basin of attraction of oo for P is A(oo) = C\{-2, 2]. 30 Il. Fixed Points and Conjugations EXAMPLE. The preceding example extends as follows. Let Tn(z) = cos(narccos z) be the nth Tchebycheff polynomial, and set F,(z) 2T,(z/2), which is monic. The coordinate change z = h(¢) = ¢ 1/¢ =e +e7™ = 2cosw yields + F,,(h(C)) = 2.cos(n arccos(z/2)) = 2cos(nw) = h(¢”). Thus h implements a conjugation of F, and ¢”. EXAMPLE. Let S$ denote the unit square in the ¢-plane with corners 0,1,1+7%,% and let h be the conformal map taking S to the upper half-plane, with 0,1,1 +7 being mapped to co, —1,0 respectively. Symmetry considerations show h maps ito +1 and (1+ %)/2 toi. We continue h by reflection to a meromorphic function on the whole ¢-plane mapping unit-squares alternately to upper and lower half- planes. Then h is doubly periodic, with periods 2 and 2i and with a double pole at each period point. The function h(2¢) is also doubly periodic, and by comparing poles and expansion at 0 one checks that ATTA) = ACI + 4)/2)) _ (h(G)? + 1)? A HCCC) — (HUG) + 1) -4h(C)((G)? = 1) Thus the rational function f(z) = (2? + 1)?/42(z? — 1) satisfies the Schroder equation h(2¢) = f(h(¢)), and h implements a conjugation of f(z) and multiplication by 2, at least where h is univalent. In any event this makes the dynamic behavior of f transparent. In particular, one sees that a point z is iterated by f to the repelling fixed point of f at oo if and only if z = h(C) for some dyadic point ¢ = 2-im +27*ni. Note that h is not univalent at 0, and the multiplier of f at co is 4. The function h is the Weierstrass P-function. Several other elliptic functions were treated by Schréder, and this specific example was given by S. Lattés (1918). h(2¢) EXAMPLE. We apply Newton’s method (actually due in this form in 1690 to J. Raphson) to a quadratic polynomial P(z) with simple zeros. We are iterating f(z) = 2 — P(z)/P’(z), which has superat- tracting fixed points at the zeros of P(z) and a repelling fixed point at co. The Mobius transformation ¢ = (z) sending the two zeros of P(z) to 0 and co respectively, and oo to 1, conjugates f(z) to C. Since the midpoint of the line segment joining the zeros of P(z) is mapped by f to co, the midpoint is sent by ¢ to the preimage —1 of II.2. Attracting Fixed Points 31 +1. Thus the perpendicular bisector of the line segment joining the zeros of P(z) is mapped by ¢ to the unit circle. We conclude that the basins of attraction for Newton’s method are the respective open half-planes on either side of the bisector. 2. Attracting Fixed Points The easiest to treat are the attracting fixed points that are not super- attracting. The following linearization theorem is due to G. Koenigs (1884). THEOREM 2.1. Suppose f has an attracting fired point at zo, with multiplier satisfying 0 < |A| < 1. Then there is a conformal map ¢ = 9(z) of a neighborhood of z onto a neighborhood of 0 which conjugates f(z) to the linear function g(¢) = XC. The conjugating function is unique, up to multiplication by a nonzero scale factor. Proof. Suppose zo = 0. Define yp(z) = AT" f"(z) = z +--+. Then yp satisfies gnof =A "Ft! = Nong. eee —“—™—SSs—s—sSSCSCsCSCSCsC“(“(‘(‘;‘“™SE"UC‘Cia__i_CtisczsCCN conjugation. To show convergence note that for 6 > 0 small If(z)-Az| 0 so small that p = (|A|+C6)?/|A| < 1, and we obtain _ ™(z)/? rC|z/? Inti (Zz) — Yn(z)| = EK yaw ©) s ot < ° a | for |z| < 6. Hence y,(z) converges uniformly for |z| < 6, and the conjugation exists. The uniqueness assertion was already noted in the preceding section. 0 32 Il. Fixed Points and Conjugations The conjugation y constructed above is normalized so that ¢' (20) = 1. The uniform convergence shows that if f depends analytically on a parameter, so does the (normalized) conjugation . The Schréder equation (1.1) satisfied by the conjugating function y becomes elf (2) = Ave). (2.1) This equation allows us to extend analytically to the entire basin of attraction A(z), by the formula y(z) = y(f"(z))/A", where 7 is cho- sen large enough so that f”(z) belongs to a coordinate neighborhood of zp. The extended ¢ is well-defined and satisfies the same functional equation (2.1). Note that y(z) = 0 if and only if f"(z) = 20 for some n>1. The branch of y~! mapping 0 to zo can be continued until we meet a critical point of f or leave the domain of f. If f is a polynomial, or rational, then the range of y covers the entire complex plane. The Riemann surface of y~! is a branched covering surface over C, on which the action of f can be described as a multiplication by A and a shift from one sheet to another. EXAMPLE. Suppose f(z) is a finite Blaschke product of order d, with a simple zero at z = 0: d 2) =e, J] 2-2. fey = ete ae The basin of attraction A(0) of the fixed point 0 is the open unit disk A. The conjugating function y is an infinite-to-one mapping of A onto the complex plane C, whose zero set accumulates on all of OA. The functional equation (2.1) for y shows that the critical points of y are the critical points of f and all their inverse iterates. 3. Repelling Fixed Points The existence of a conjugating map for a repelling fixed point follows immediately from the attracting case. For suppose f(z) = 20 +Xr(z— 2) +++: where |A| > 1. Then f~!(z) = 20 + (z — 20)/A +--> has an attracting fixed point at 29. Any map conjugating foi(z) to ¢/A also conjugates f(z) to AC. 11.4. Superattracting Fixed Points 33 EXAMPLE. If P(z) is a polynomial with a repelling fixed point at the origin, the corresponding map h = y~! can be defined on the entire plane by the functional equation h(AC) = P(h(¢)). It is an entire function of finite order, and in fact it satisfies an estimate of the form logd log |h < 7 = ogi SGo+ilcl", 7 =i ES d= deg P. As a special case, the monomial 2” has a repelling fixed point at z = 1 with multiplier m. The coordinate change € = y(z) = logz conjugates z” to m¢. The corresponding entire function is h(¢) = eS. 4. Superattracting Fixed Points In the superattracting case the existence of a conjugation was first proved by L.E. Boettcher (1904). THEOREM 4.1. Suppose f has a superattracting fired point at zo, f(z) = 20 + ap(z — 20)? +---, dp #0, p> 2. Then there is a conformal map ¢ = y(z) of a neighborhood of zy onto a neighborhood of 0 which conjugates f(z) to ¢?. The conjugating function is unique, up to multiplication by a (p—1)th root of unity. Proof. Suppose zo = 0. For |z| small there is C > 1 such that | f(z)| < C|z|?. By induction, writing f"t! = f”o f and using p > 2, we find that If(2)i<(ClzP", lel $6, so f"(z) — 0 super-exponentially. If we change variables by setting w = cz where c?-! = 1/ap then we have conjugated f to the form f(w) = w?+---. Therefore we may assume ap = 1. We wish to find a conjugating map y(z) = z+--- such that y(f(z)) = y(z)?, which is equivalent to the condition that pofop l=CP. Let en(z) = f(z)" = (2 PS aL)", which is well defined in a neighborhood of the origin. The y,’s satisfy n+l yn-10f=(f" bof)” =f, 34 II. Fixed Points and Conjugations so if Yn — ¢ then ¢ satisfies yo f = ¢? and so is a solution. To show that {yn} converges, we write frt) = fo f™ and note that enti _ (ro ft\P” _ n))\p-" ae ( P= a+ our? 14+ O(p-")O(|z/?"C*") = 1+ O(P") I if |z| < 1/C. Thus the product co Il Pn+1 Pn converges uniformly for |z| < ¢ < 1/C, and this implies {yn} con- verges. Hence ¢ exists. The uniqueness statement was noted in Sec- tion 1.0 Again if f depends analytically on a parameter, then so does the conjugation y constructed above. In this case the functional equation satisfied by the Boettcher coordinate function, allows us to extend y analytically only until we meet a critical point of f. However, the functional equation log |y(f(z))| = plog |y(z)| allows us to extend log |p(z)| to the entire basin of attraction A(z0) of zo. The extended function is a negative harmonic function, except for logarithmic poles at all inverse iterates of zo. Consider the case in which f is a p-sheeted cover of the immediate basin of attraction A*(zo) onto itself. The only pole of log |y(z)| is now zo, and log|y(z)| = log|z — 20| + O(1) at 20. Furthermore, the functional equation shows that log |y(z)| > 0 as z + 0A*(zo). Except for sign, these are precisely the properties that characterize Green’s function G(z, zo) for A*(zo) with pole at zo, and we obtain log |y(z)| = —G(z,20), 2 € A*(20)- EXAMPLE. A polynomial P(z) = az4+--- with d > 2 and a # 0 has a superattracting fixed point at oo. Replacing 0 by oo in Boettcher’s II.5. Rationally Neutral Fixed Points 35 theorem, we see that P(z) is conjugate to ¢? near oo. The conjugating map has the form y(z) = cz + O(1), with a simple pole at oo. Again the functional equation log|p(P(z))| = dlog|p(z)|, zl > B, allows us to extend log |y| to the entire basin of attraction A(co) of oo. The basin of attraction has no bounded components, by the maximum principle, so that A(oo) = A*(oo) is connected. The ex- tended log |p| is harmonic on A(oo), and log |yp(z)| > 0 as z tends to the boundary. This time log |p(z)| = log |2| + log |e|+o(1), —_|2| + 00. Now Green’s function for an exterior domain D with pole at oo is characterized as the positive harmonic function on D which tends to zero at OD (or at least at all regular boundary points of 0D), and which has the form G(z, 00) = log|z| + o + o(1) at oo. The constant o appearing here is called Robin’s constant, and the logarithmic capacity of 0D can be defined as e~°. (See [Ts].) In the case at hand we see that log |y(z)| is Green’s function for A(co) with pole at oo, and log|c| is Robin’s constant. The logarithmic capacity of 0A(oo) is 1/|e| = Ja|~/(4-), 5. Rationally Neutral Fixed Points Suppose \” = 1 and f(z) = Az + az?t1 +---, a #0. We consider three cases: 1A=1,p=1, 2X 1 pe, 3. A7=1, AF 1. CaSE 1. We assume f(z) = z+a2?+--+ with a 0. By conjugating f by y(z) = az, we may assume a = 1. Let us next move 0 to oo 36 Il. Fixed Points and Conjugations by conjugating by the inversion z + —1/z. The conjugated map has the expansion b gz) =2+lt ote (5.1) near oo. We shall consider two methods for proving there is a map y conjugating g to the translation z > z+ 1. The first is due to Fatou and uses the asymptotic behavior of g”. The second involves quasiconformal mappings. First observe that if Co > 0 is sufficiently large, then the half-plane {Rez > Co} is invariant under g, and Reg"(z) >Rez +5, Rez >Cp,n2>1. This is proved easily by induction, as is the upper bound in ; a Rez>Co,n>1. These estimates are valid whenever g(z) = z+ 1+ o(1), and they will be used later. In the first method we define yp by Yn(z) = 9"(z) —n— blogn, Rez > Co. From A k+1 k = ee gz) = 9"(z) +14 Fat (=) we obtain (z) — yx (2) = bflog k — lo (k+N]+ gy +0 i ) =0(;) Pk+1 Priz) = g 3 9F(2) Re) kK) where the estimates are all independent of z. It follows that n-1 len(z) — 2] < ber(z) - 2] + YO lees (z) — ¢e(z)] = O(log n) k=1 for Rez > Cy. Now to prove that y,(z) converges, we estimate gnei(z)— Gn(z) = blogn —blog(n + 1) +g"*(z) — g"(z) - 1 bb 1 ~ ne" ghz) o(;3) 1 1 1 _ of blogn+Yn(z) nl +0(3) b / : 1 logn i 2° |blog n 4 en(2)l) + o( ) =o( 1 y IL.5. Rationally Neutral Fixed Points 37 Ay AY y=C loghl > > X > 2] {0 > > -Y ; ~ a i y =—C logh! FIGURE 1. Parabolic attracting regions at oo and at 0. Hence & |y~n+1(z) — Yn(z)| < 00, and yn converges, say to y. Since the yn’s are univalent, so is y. From the definition, ¢n(g(z)) = Yn4i(z) + 1+ blog(1 + 1/n). Hence yog = y +1, and ¢ conju- gates g to the translation z > z+ 1. Using the functional equation y(g(z)) = y(z) +1, we can continue y analytically to any domain Q on which g is defined and satis- fies g(Q) C © and Reg"(z) > +00, z € ©. For instance, from the asymptotic form of g at oo we see that for any 6 > 0 we can choose Cs so large that 2 = {|y| > —éa + Cs} U {a > Co}, the union of three half-spaces, is invariant under g, and g on 2 is conjugate to translation. With some more effort we can construct an invariant domain 2 with smooth boundary so that the part of 2 in the half-plane x < —2 consists of the domain above the curve y = Clog|z| and the symmetric domain below the curve y = —C log |z|, where C > 0 is appropriately large, as at the left of Figure 1. Indeed, from g(z) = utiv = 2+1+iy+O(1/|2|), we obtain for z on the upper boundary curve that il 1 log |u| = log |x| — jel + o(aa): so that A A v > 9 Ty = Cloelel~ F CHA 1 Cloglul + G- —+O(cR) > C log |u| IV 38 II. Fixed Points and Conjugations for C large enough. If we connect the two logarithm curves as in Figure 1, we have an invariant , on which g is conjugate to trans- lation. A similar estimate in the right half-plane shows that for C sufficiently large and any t > 0 the domain % defined by x > Co and |y| < Cloga +t is invariant under g. Note that the ’s fill out the half-plane x > Co. Thus for any C; > C' the domain (* defined by x > Cp and |y| < Ci logz is invariant under g, and moreover it captures eventually the forward orbit of every point in the half-plane z>Cp. The inversion z — —1/z carries 2 to the cardioid-shaped region pictured at the right of Figure 1, which we refer to loosely as the at- tracting petal for the fixed point. Inside the attracting petal f is con- jugate to z > z/(1—z). At the cusp the boundary curves are tangent to the positive real axis. In fact, the images of the logarithm curves defined above have the form y = +C? log(1/x) + O(y”) as x — 0+. The iterates of any point in the cardioid tend to 0 in a direction tan- gent to the negative real axis. Furthermore they all eventually enter and approach 0 through the image of *, which forms a narrow cor- ridor bounded by curves of the form y = +C,2" log(1/|2|) + Oly?) as z > 0-. We now describe the second method for dealing with this case. As before, assume g is given by (5.1), and let V = {Rez > A} bea half-plane such that g(V) C V. Let L be the vertical line bounding V. Then L’ = g(L) is a smooth curve which is approximately the translation of L by 1 to the right. Let S denote the vertical strip in the z-plane bounded by L and L’. In the ¢ = € + in plane, let ¥ be the honest strip {0 < € < 1}. We construct a diffeomorphism h: 0S by setting h(in) = A + in and h(1 + in) = g(A + in) and then filling in these boundary values in a smooth way, uniformly as Im¢ — +o. For technical reasons, let h be conformal on some disk in &. We now extend h(¢) to a diffeomorphism of the right half-plane onto V by iterating the formula h(¢ + 1) = g(h(¢)). The analyticity of g guarantees that the ellipse field (Beltrami coefficient) j1 of h is invariant under translation by 1, 1(¢ + 1) = u(¢). Extend yp to the whole ¢-plane to be periodic. Then jz is smooth except at oo, and |u| 1, and then le(g"(z))| ~ lal”, which grows faster that n®. We conclude that a = 1, and ¢ is the desired conjugation. CASE 2. Suppose now that 2’ = f(z) = z+az2?t!+--- with a #0 and p > 1. As before we may assume a = 1. Define z = cl/P and a crip for 0 < arg¢,arg¢’ < 2m and z,z’ restricted to an appropriate sector of aperture 27/p. Then C=C +p? + 071”). We renormalize again to remove the p and then change variables by ¢ =-1/z and ¢’ = —1/z’. This gives zg =g(z)=2414+O(l2|71/”). The first method in Case 1 can be used here again, but it becomes much more complicated since not only logn enters but also other terms involving n*/? for 1 < k < p. On the other hand, the second method involving quasiconformal maps goes through in this case without change. Moreover, the discussion of the invariant domain 2 remains intact, except that now the invariant Q is bounded by the curves y = +C|z|!~!/? in some half-space « < —M. 40 II. Fixed Points and Conjugations RS FIGURE 2. Pattern of attracting petals for z + z* and —z + 24, Another proof covering Case 2 proceeds as follows. First we kill off one by one the higher order Taylor coefficients of f(z) = z+ aptly. Ap p22? +? +++ by conjugations of the form z — z+ az, Only the coefficient of z2?+! refuses to die, and we obtain for any arbitrarily large prescribed q a conjugation of f(z) to F(z) = z+ 2Pt! 4 Az?t) + O(|z2|%). The coefficient A is actually a conformal conjugation invariant, though not the only one (see [Vo]). If now we fix g > 2p+1, then the method used in Case 1 applied to F(z) converges to the desired conjugation. (We mention as another approach the proof in [Mi2], in which a cou- ple of preliminary smoothings near oo lead to rapid convergence.) The picture is as follows. Instead of having a single attracting petal, there are now p attracting petals, each within a sector of aper- ture 2n/p, as in Figure 2. The angles of the boundary rays of the sectors are given in terms of the coefficient a of 2+! by 6, = me tk ck 0. But {\"} is dense in the unit circle so h(Cie) = A(G2e") for every 6. This implies h(¢,z) = h(¢2z) for |z| < 1, and since h/(0) = 1, a =¢.5 THEOREM 6.2. A solution h to the Schréder equation (6.1) exists if and only if the sequence of iterates {f"} is uniformly bounded in some neighborhood of the origin. 42 Il. Fixed Points and Conjugations Proof. If h exists then f"(z) = h(A"A7!(z)) is obviously bounded. On the other hand, if |f"| < M for all n, then we can define ym . gn(z) = > Dd? F (2): 0 Then {yn} is a uniformly bounded sequence of analytic functions, so contains a convergent subsequence. Since (no f = An + O(1/n), any limit of the Yn’s satisfies yo f = Ay. From d = f'(0) we have gi,(0) = 1, and y’(0) = 1. Thus h = y7} is a solution to (6.1). 0 One immediate consequence of Theorem 6.2, which is not so easy to see directly, is that if f is topologically conjugate to Az near 0, then it is conformally conjugate. Indeed, if there is a topological mapping h, h(0) = 0, such that h-ofoh=Xé, then for 6 > 0 small the images of the disks A(0, 6) under h are invariant under f, so the iterates of f are uniformly bounded on a neighborhood of 0. THEOREM 6.3. There exists a \ = 2" so that the Schréder equation (6.1) has no solution for any polynomial f. Proof. Let f(z) = 244...4z, and suppose there is a conjugation h defined on A(0, 6). Consider the d” fixed points of f”, that is, the roots of ft(z) — 2a 2h 4-4 (A"- Dz =0. One root is 0. Label the others 21,...,2ar—1 and note that since f(z) = h(A"h71(z)) has only one zero in A(0, 4), 25 ¢ A(0, 6) for peo de chug of < J] lel = 1-2" We now construct a A for which this is impossible, thus contradicting the existence of h. Suppose qi < q2 < --: is an increasing sequence of integers, and set 6 = 72, 27% and A= exp(2770). Then jl- "| ww DMV +1 Taking logarithms, we find that dhe S$ C(5)0™. IL.6. Irrationally Neutral Fixed Points 43 If we define inductively qj,’s to grow very rapidly, say with log qx41 > k2%, this inequality is violated for any d and 6 > 0.0 An example as above was first given in 1917 by G.A. Pfeifer. The work was continued by H. Cremer, who proved in 1938 that if |A| = 1 and lim inf |\” — 1|!/” = 0, then there is an analytic function f(z) = Az +--+ such that the Schréder equation (6.1) has no solution. In a seminal paper in 1942, C.L. Siegel gave the first example of a unimodular \ for which (6.1) is solvable. A real number 6 is Diophantine if it is badly approximable by rational numbers, in the sense that there exist c > 0 and p < co so that Dic 6-— | >— 6.2 | qi gt (6.2) for all integers p and q, q 4 0. This occurs if and only if A = e278 satisfies A" -1p>en0#, n>1. For fixed yp > 2, the condition (6.2) holds for a.e. real number 6. Indeed, if E is the set of @ € [0,1] such that |@ —p/q| < q-+ infinitely often, then the measure of E£ is estimated in the obvious way by |E| < $0 2-q*-q=O(n?*) 30. q=n In particular, almost all real numbers are Diophantine. By a theorem of J. Liouville [HaW)], any algebraic surd 6 of degree m satisfies (6.2) for p >m. THEOREM 6.4 (Siegel). If 0 is Diophantine, and if f has fixed point at 0 with multiplier e?"°, then there exists a solution to the Schréder equation (6.1), that is, f can be conjugated near 0 to multiplication by @2m9, Proof. We want to solve h(Az) = f(h(z)). If we define f and h by f(z) = Az + f(z) and A(z) = z+ h(z), then the equation can be written h(Az) — Ah(z) = f(R(z))- (6.3) Siegel’s original method was to expand both sides in power series 44 II. Fixed Points and Conjugations using 00 0 A(z) = Slane”, f(z) = Yo bne” n=2 n=2 to inductively obtain equations of the form a, (A" — d) = An(a2,°°+5 Qn—1,b2,:+*;bn), n > 2. Using |A” — | > cn! he was able to estimate the a,,’s and prove the power series converges. This is not easy. We take a different route, following the proof in [SiM]. We shall use KAM theory (A.N. Kolmogorov, V.I. Arnold, J. Moser) in a simple case. We consider coordinate changes ~ around z = 0, normalized so that ¥(z) = z+ w(z), where p(z) = O(2?). Instead of finding a solution to (6.1) right away, we build a ~ so that yo fop=g(z) =rz2+G(2), (6.4) where g is smaller than f is some sense. (Note that g = 0 solves the problem.) We then repeat the procedure with f replaced by g and ~ defined on a slightly smaller disk. Careful estimates show we obtain the desired h in the limit on a disk of positive radius. So suppose f(z) is as above. The idea is to replace (6.3) by its linear version, by replacing h in the right-hand side by z, bO2z) — A¥(2) = F(2)- (6.5) This can easily be solved for av, 7 an : 02) = Vy” j=2 and our hope is that the corresponding g defined by (6.4) is smaller than f. We estimate g and g’ using the assumptions ne r(1 — 2n) for z € OA(0,r(1 — )) and (z) = 0 only at 0 in this disk. Now consider how g = y~!o0 fow maps A(0,r(1— 4n)). First y takes this disk into A(0,r(1 — 3n)), then f takes this to A(0,r(1 — 2n)) if 6 < n, and finally ~~! takes A(0,r(1 — 2n)) into A(0,r(1—17)). Thus g maps A(0,r(1—47)) into A(0,r(1 — 7)). We now estimate g. From (6.4) we have G(2) + bz + G) = A(z) + fle + ¥), and using (6.5) we obtain G2) = bz) — bz4 §) + fle+ dB) - fe). We wish to estimate g in A(0,r(1—4n)). Let C be the maximum of |g| in this disk. Then C < sup|i!|-Ca + sup|f(z +0) - f(z)| 0 small enough and require 7 < ¢1, then the first condition is satisfied, and the third condition follows from the second. Suppose now 1,69 have been chosen to satisfy these conditions. Define sequences Trtl = tn(1— 5mm), Mn+ = Mn/2; code Ht2Q- 442), ont The required condition codn < nt? is now easy to verify by induc- tion. If it holds for n, then 827 (ut) Q— (H+?) < Peta Ht2) g(t?) = tt? 2 Cobn+i = © m+1? and it holds for n+1. We have also inductively constructed sequences {un} and {gn} with go = f and gn = Ua! 0Gn-1° Yn, OF equivalently gn = Vg! opal o---opp lo fodroyo-:-odn. If R=ro[](1—5mm) > 0, then |9),| < dntn/(1- Mn) > 0 on A(0, R), SO gn — Az on the disk. Thus {1 0--- on} converges to a mapping h which conjugates f to Az, as desired. 0 For quadratic polynomials P(z) = e292 4 2? precise conditions are known for the existence of a conjugation. Such exists if and only if y log gn+1 oo nal &n where {pn/dn} is the sequence of rational approximations to 0 com- ing from its continued fraction expansion. The sufficiency of this condition was proved by A.D. Brjuno (1965), and the necessity was established by J.-C. Yoccoz (1988). In Section V.1 we give an ele- mentary proof, due to Yoccoz, of the existence of the conjugation for almost all 0. II.7. Homeomorphisms of the Circle 47 7. Homeomorphisms of the Circle Any orientation preserving homeomorphism f of the unit circle can be expressed in the form f(t) = e?"*(, where F is an increasing homeomorphism of the real line R that satisfies F(t+1) = F(t)+1. The lift F of f is unique up to adding an integer. We wish to assign a rotation number to f, which will measure the average advance of F over an interval of length 1, that is, the average speed of F’. LEMMA 7.1. Let F be as above, and suppose F(0) > 0. Fit p> 1. If m= m(p) is the first integer such that F™(0) > p, then <=+—. nm m a < lim inf F"(0) < limsup ~ < 1) tec on n—co =o m m a. Proof. If0 0 is arbitrary, and j is the integral part of s, then s+ F(0)—1 < j+F(0) = F(j) < F(s) < F(j+1) = F(0)+j+1 < s+ F(0)+1. We check by induction that s + k(F(0) - 1) Fem-1)(9) < F*(s)0,k>1, < kp< FY"), k>1. Write n = km-+q where 0 < gq < m. Then kp + q(F(0) — 1) < F4%kp) < F4(F*"(0)) = F"(0) = FAK(FH™N(0)) < FU*(kp) < kp + (q+k)(1+ F(0)). This gives k F(0O)-1 Fro k; k+ Pp, ¢FO)—1) . FO) . kp 144 F(0)). n n ~ nN n n Since k/n — 1/m and q/n — 0 as n — 00, the lemma follows. 0 There is a similar estimate if F(0) < 0. It follows from the estimate that a(F) = lim F"(0) no oN exists. We call a(F’) the rotation number of F’. We define the rotation number a(f) of f to be the residue class of a(F’) modulo 1. This is independent of the lift F’. One checks that the translation t — t + 6 has rotation number 6, so the rotation fo(¢) = e?"°¢ has rotation number 6 (mod 1). From the estimates in the lemma it is clear that the rotation number of 48 Il. Fixed Points and Conjugations F depends continuously on any parameters on which F' depends continuously, as does the rotation number of f. We have F"(t)/n — a(F) for any real number t. This follows from the periodicity modulo 1 of F and the estimate F"(0) < Fr(t) < F"(0) +1 for0 t+ao+e for 0 < t < 1, then F"(0) > n(ao + €); and a(F) > ao + €. Similarly, if F(t) 1. Then there is € > 0 such that if f is any homeomorphism of the circle II.7. Homeomorphisms of the Circle 49 with rotation number a(f) = a, which extends to be analytic and univalent on the annulus {1/0 < |w| < o} and satisfies |f(w) — e?taW| < © there, then f is conformally conjugate to the rotation e"4y on the annulus {1/V/o < |w| < Vo}. Proof. It will be convenient to work with the lifts of functions to a strip rather than functions on an annulus. Let S, be the horizontal strip {x + iy : |y| < p}. For g(z) defined on S, let |\g||, denote the supremum of |g(z)| over Sp. Replacing f by its lift, we assume f(z) is analytic on S, and satisfies f(z + 1) = f(z) +1. We may assume that f(x) is increasing on R. Write f(z) = z+a-+ f(z). Then f is periodic and has a Laurent expansion +00 f= > bpe2™inz, en 2TP < |e2Mi2| < e2np, n=—00 We wish to find a univalent function h on S,/2 that satisfies h~! 0 f oh(z) =z+a, that is, h(z +a) —A(z) =a+ f(h(z)). As before, we consider the linearized version V(z +a) - V(z) = a4 f(z). (7.1) In Theorem 6.4 the multiplier was just the coefficient of z in the series expansion, the term corresponding to a + bo here. To simplify matters we assume for the moment that bp = 0, that is, the constant term of f is just the rotation number. The linearized equation is then solved by ¥(z) = z + f(z), where + b - A w(z) = > oa i io N= eta, n#0 We assume that 1 = a— 1 : prog —4n < 37. We wish to obtain an estimate of the form ||g||p—4) < Co6*n-#-1. The proof is then completed as before by iterating, with 6? providing the margin of victory. Equating the expressions a(2) + bolg(z)) = 2 +.0+ Gz) + dole +a + G(2)), doz) +a + f(do(2)) = 2 + dolz) +a + f(z + Yole)), vo(9(z)) F(Yo(2)) we find using (7.3) that 4(z) = [f(z + dolz)) — F@)) — fo(z + + G(2)) — ole +] + bo Since @ is also the rotation number of g, g(x) is equal to x + a for some 2, so that g(x) = 0 for some x € R. Thus bp is estimated by the terms in square brackets, and it suffices to estimate these. From the Schwarz inequality and (7.2) we obtain oe r 25 _ |f(z+ Yolz)) - FS Gy lMolle—an < C28 # Set C = ||9l|p-an- Then C < 3n and bo (z + a + G(z)) — Pol(z +.)| < 2069S < Cy6n# 0. Hence C < 2[C262n-#7! + Cydn-#—1C]. If 6 < nt1/(4C1), we can solve for C as before and get the desired estimate. II.7. Homeomorphisms of the Circle 51 For the iteration procedure the essential condition on 6 has the form 5 < cyn4*?. We choose 1m = p/2"**, pn = p—4(m+++-+Mn-1); 3/2 and 6, = 67). and a, and f with \Ifllo < 6, , we obtain a sequence of conjugations {vn} on S,, and a sequence gn = Wrto...ou; lo fowyo...oUn. By induction the corresponding g,,’s satisfy ||Gnllo, < 6n. Then gn converges uniformly to z + @ on Sp/2, and 71 0---0 Up converges on S,/2 to a univalent mapping h which conjugates f to translation by a0 Then for 6; sufficiently small, depending only on p Ill Basic Rational Iteration We consider a rational mapping R(z) and use it to partition the sphere into two disjoint invariant sets, on one of which R(z) is well- behaved (the Fatou set), on the other of which R(z) has chaotic behavior (the Julia set). The first milestone of the theory is a theorem of Fatou and Julia that the repelling periodic points are dense in the Julia set. From this follows the homogeneous nature of the Julia set. In the words of Julia, “la structure de E’ est la méme dans toutes ses parties.” 1. The Julia Set Let R(z) be rational, and assume always that R has degree d > 2. Thus R = P/Q where P,Q are polynomials with no common factors and d = max(deg P,degQ) > 2. Then R(z) is a d-fold branched covering of the Riemann sphere, and in fact every d-to-1 conformal, branched covering of C comes from some such R. We define the iterates of R as before, R? = Ro R, R" = R"!o R and note that R” has degree d”. Given zo, the sequence of points {z,} defined by zn = R(zn-1) is called the (forward) orbit of 29. The chain rule gives the useful 54 Ill. Basic Rational Iteration identity (R")'(20) = Ri (zo) Ri (21) ++ R'(2n-1)- The point zo is called periodic if zn = zo for some n. This occurs if and only if zo is a fixed point for R”. The minimal n is its period. The orbit {21, 22, ++) 2n = zo} is called a cycle. We classify the cy- cle as attracting, repelling, rationally neutral or irrationally neutral according to the type of fixed point for R". Thus for instance the cycle is attracting if and only if J] |R'(z;)| < 1, and this does not depend on the choice of point in the cycle. The point zo is called preperiodic if z, is periodic for some k > 0, and strictly preperiodic if it is preperiodic but not periodic. Recall that a critical point of R is a point on the sphere where R is not locally one-to-one. These consist of solutions of R'(z) = 0 and of poles of R of order two or higher. The images of the critical points are the critical values of R. The order of a critical point z is the integer m such that R is (m + 1)-to-1 in a punctured neighborhood of z. If z is not a pole this is its multiplicity as a zero of R’. There are 2d — 2 critical points, counting multiplicity. One way to see this is to compose R with a Mobius transformation, to reduce to the case where oo is neither a critical point nor a critical value, and R(oo) = 0. Then R(z) => = where a, 3 # 0, and so R(z) Q(2)P'(z) — P(2)Q"(z) _ — at + Q(z)? Q(z)? The Fatou set F of R is defined to be the set of points zo € Cc such that {R”} is a normal family in some neighborhood of zo. The Julia set J is the complement of the Fatou set. Thus 20 € F if and only if the family {R”} is equicontinuous on some neighborhood of zo, with respect to the spherical metric of C. Evidently the Fatou set is an open subset of C, and the Julia set is compact. If R has an attracting fixed point zp, then the basin of attraction A(z) is contained in the Fatou set. On the other hand, since the iterates of R converge to zo on A(zo) but not on its complement, the iterates cannot be normal on any open set meeting OA(zo), and OA(zo) is included in the Julia set. We will see later (Theorem 2.1) that 7 = 0A(z0). IIL.1. The Julia Set 55 EXAMPLE. Consider R(z) = z?. Then R"(z) = 2?” converges to 0 in {|z| < 1} and to co in {|z| > 1}. The Julia set is the unit circle {l2| = 1}. EXAMPLE. Let R(z) = z” — 2. We have seen in Section II.1 that the basin of attraction of 00 coincides with C\[—2, 2], so the Julia set 7 is the closed interval [—2, 2]. The preceding two examples may be misleading. Julia sets are al- most always intricate fractal sets, which are conformally self-similar. See the figures in Chapter VIII of the Julia sets of R(z) = z* +c for various other values of c. EXAMPLE. Take A, |A| = 1, such that R(z) = Az + z? is conjugate to an irrational rotation Az near 0 (such exist by Theorem II.6.4). Then {R"} is normal in neighborhoods of 0 and oo, so there are components Up and Ux of F containing these points. Since oo is an attracting fixed point and R"(0) 4 oo, they must be distinct. Hence R"(Up) never hits Ux, and {R"} is uniformly bounded on Up. Also Up is simply connected, since by the maximum principle {R"} is bounded within any closed curve in Up. In view of Theorem II.6.2, we see that Up is characterized as the largest domain containing 0, invariant under R, on which the conjugation map exists. A simply connected component of the Fatou set in which R is conjugate to an irrational rotation is called a Siegel disk. An example is illustrated in Section V.1. THEOREM 1.1. The Julia set J contains all repelling fired points and all neutral fixed points that do not correspond to Siegel disks. The Fatou set F contains all attracting fixed points and all neutral fixed points corresponding to Siegel disks. Proof. This is obvious, in view of Theorem II.6.2. 0 In the third example above we saw that a disk around 0 was in the Fatou set for some |\| = 1, but if |A| > 1 then it is a repelling fixed point so is in the Julia set. Thus F and J may move quite discontinuously as R varies. We shall see other examples of this later. 56 III. Basic Rational Iteration THEOREM 1.2. The Julia set J is nonempty. Proof. Recall our standing assumption that R has degree d > 2. Suppose J = 0. Then {R"} is a normal family on all of C, and so there is a subsequence {nj} such that R™(z) — f(z) for some analytic function f from C to C. Since f is analytic on all of Citisa rational function. If f is constant then the image of R™ is eventually contained in a small neighborhood of the constant value, which is impossible since R” covers C. If f is not constant, eventually R™ has the same number of zeros as f (apply the argument principle), which is also impossible since R” has degree d”. 0 We say that a set E is completely invariant if both it and its complement are invariant. Since R is onto, this occurs if and only if R-(E) =E£. THEOREM 1.3. The Julia set J is completely invariant. Proof. Clearly R~!(F) € F. Suppose zp € F, and suppose {Ruths R" o R} converges uniformly (in the spherical metric) on a neigh- borhood of zo. Since R maps open neighborhoods of z) onto open neighborhoods of R(zo), {R"} converges uniformly on a neighbor- hood of R(zo). It follows that R(zo) € F, and R(F) C F. Hence F is completely invariant, as is its complement J.0 THEOREM 1.4. For any N > 1, the Julia set of R coincides with that of RN. Proof. The Fatou sets for R and RN are the same, since {R"} is normal on an open set U if and only if {R} is normal on U. 0 Let z € J and let U be any neighborhood of z. By Montel’s theorem (Theorem 1.3.2), the sequence {R”} on any such U omits a set E, containing at most two points. THEOREM 1.5. The set Ez, is independent of z (so shall be denoted as E). If E is a singleton, we can conjugate tt to co, and then R(z) is a polynomial. If E consists of two points, we can conjugate these to 0 and co, and then R(z) is Cz4 or Cz~4. In all cases, E is contained in the Fatou set F. III.1. The Julia Set 57 Proof. By the definition R-1(E,) C E,. If Ez is one point a, then R(a) = a. Take a = oo. Since R7!(00) = oo, there are no other poles, and R is a polynomial. Clearly E, is independent of z. If Ez consists of two points, we may assume these to be 0,00 and either R(0) = 0, R(co) = oo or R(0) = 00, R(co) = 0. In the first case, R is a polynomial with 0 as its only zero, so R(z) = C'z?. Similarly, R(z) = Cz~¢ in the second case. 0 The set E is called the exceptional set of R. It follows immediately from the definition of FE that if z ¢ E then J is adherent to the inverse orbit Un>1R~"(z) of z. Since E is disjoint from 7, we obtain in particular the next two theorems as corollaries. THEOREM 1.6. The backward iterates of any z € J are dense in J. This theorem can be used as a basis for computing J (the inverse iteration method). It is effective when the degree of R is small, so that the number of inverse iterates does not grow uncontrollably fast. THEOREM 1.7. Any nonempty completely invariant subset of J is dense in J. If D is a union of components of F that is completely invariant, then J = OD. Proof. For the second statement, note that OD is a completely in- variant subset of 7. 0 THEOREM 1.8. The Julia set J contains no isolated points, that is, J is a perfect set. Proof. Take zp € J and U an open neighborhood of zo. First assume zo is not periodic and choose z; with R(z1) = zo. Then R"(zo) # 21 for all n. Since z1 € J, backward iterates of z; are dense in J, so there is a ¢ € U with R™(¢) = 21. Thus ¢ € JNU and ¢ F 2%. Next suppose R"(zo) = 2 for some minimal n. If zo were the only solution of R"(z) = zo then zo would be a superattracting fixed point for R", contradicting zo) € J. Hence there is 2} # z with R"(z1) = zo. Furthermore R(zo) # z1 for all j since otherwise it would hold for some 0 < j 1 and |(R")'(z0)| < 1, this ar- gument shows each component of A*(zo) contains a critical point of R". Since (R")'(z) = J] R’(R?(z)), A*(zo) must also contain a critical point of R. O REMARK. We observe for later reference that the critical point p produced above, for an attracting cycle that is not superattracting, has distinct iterates R*(p), k > 1. This is because R is one-to-one on Up. Since there are 2d — 2 critical points, counting multiplicity, Theo- rem 2.2 shows that the number of attracting cycles is at most 2d— 2. A proof of Theorem 2.2 can be based on Koenigs’ coordinate func- tion y and the (locally defined) inverse y=! near 0. Recall that y maps A*(z9) onto C, y(zo) = 0, and y’(zo) # 0. Since A* omits three points and hence cannot be mapped conformally onto C, the radius 60 III. Basic Rational Iteration of convergence p of the power series for y7(¢) at 0 is finite. There must be a critical point for R on the boundary of gy l(\¢| 0, that is, such that R"(z) tends to zo through P. As in the case of attracting cycles, the basin of attraction is an open subset of the Fatou set F, whose boundary coincides with the Julia set. The immediate basin of attraction is the connected component of the basin of attraction containing P. The definitions are extended in the obvious way to arbitrary ratio- nally neutral cycles, as follows. Suppose {20, ..., 2n-1} is a parabolic cycle of length n, so that the multiplier of R” at the fixed point zo is a primitive mth root of unity. Let P be an attracting petal for R”” at zo. We define the basin of attraction of R containing the petal P to consist of all z such that R*(z) € P for some k > 1. In this case, the sets R’(P), 0 < j < nm, are disjoint attracting petals at the z;’s, and the basin of attraction of R containing P is the union of the basins of attraction of R"” containing the R/(P)’s. The imme- diate basin of attraction containing the petal P is the union of the nm connected components of F containing the R?(P)’s. THEOREM 2.3. If zo is a rationally neutral periodic point, then each immediate basin of attraction associated with the cycle of zo contains a critical point. Proof. Replacing R by RN, we may assume R(zo) = 20 and R'(z0) = 1. Let A* be the immediate basin of attraction containing a petal P at zo. Let y be the Fatou coordinate function defined on P, so that g is univalent and conjugates R to a translation: y(R(z)) = y(z) +1. Since R(A*) = A*, the functional equation allows us to extend ¢ to A*, where it satisfies the same identity. Since y(P) covers a right half- plane, y maps A” onto the entire complex plane. Now y has a critical point wo in A*, else we could define a branch of y~!, which would be a meromorphic function on the entire complex plane omitting the Julia set, hence constant. For n large, R"(wo) is in P, where y has no critical points, so by replacing wo by the largest iterate that is a critical point, we can assume that R(wo) is not a critical point of III.2. Counting Cycles 61 y. From the functional equation we then have y’(R(wo))R’(wo) = y'(wo) = 0, so that wo is a critical point of R. 0 THEOREM 2.4. The number of attracting cycles plus the number of rationally neutral cycles (counting their natural multiplicity) is at most 2d — 2. Proof. This follows immediately from Theorems 2.2 and 2.3, since the basins of attraction are disjoint, and since there are at most 2d — 2 critical points. O Finally we treat irrationally neutral cycles by means of an analytic perturbation. LEMMA 2.5. Let Ri(z) be a family of rational functions depending analytically on a parameter t. Let zp be a periodic point of Ri, of period m, and suppose the multiplier of the cycle of zo is #1. Then for t near to there is a unique periodic point z(t) of Ry of period m near zo. The periodic point z(t) depends analytically on t, and 2z(to) = zo. The multiplier X(t) of the cycle is analytic, and z(t) and A(t) extend analytically along any path in the t-plane along which R; is defined provided X(t) remains bounded away from 1, so that z(t) is periodic of period m. Proof. Set Q(z,t) = R7*(z) — z, and let (R{")'(z) be the z-derivative of Ri" (z). A point z is periodic for R; with period dividing m if and only if Q(z,t) = 0. If the period is exactly m, then the multiplier of the cycle is \(t) = (Rf")'(z). Now the z-derivative of Q(z,c) is (R@)'(z) — 1. Thus the implicit function theorem guarantees that if the multiplier of the cycle of zp is not 1, the equation Q(z,t) = 0 has a unique solution z(t) near zo for t near to, and z(t) depends analytically on t, as does the multiplier (t) of the cycle of z(t). The period of each z(t) divides m, and the set of t-values for which it is equal to m is evidently open, so the period must remain constant along any path. 0 THEOREM 2.6. The number of attracting cycles plus half the number of neutral cycles with multiplier 4 1 is at most 2d — 2. Proof. We consider an analytic perturbation R; of R = Ro. Under 62 III. Basic Rational Iteration an analytic perturbation we can expect half the neutral cycles to become attracting. To foreclose the possibility that the multiplier of the perturbed cycle is constant, we choose our perturbation to be an analytic deformation of R to 2, which has no neutral cycles. If R= P/Q, then such a deformation is given for instance by Ri(z) = (1 — t)P(z) + tz4)/[(1 — t)Q(z) + t], defined for all complex t for which R; has degree d (a cofinite subset of C). Let zo be a neutral periodic point of R, with period m and with multiplier Ao # 1. Let z(t) be as in the preceding lemma, with mul- tiplier A(t). Since we can continue z(t) and A(t) analytically along some path to 1, and |A(1)| 4 1, A(t) is not constant. Let E, be the set of @ such that |A(pe!)| < 1. Then the length of E, tends to 7 as p70. Define these sets for one point in each neutral cycle. Suppose there are N such cycles. Let Ej,) be the corresponding set and let xj,9() be the corresponding characteristic function on the circle. Then Noa op 1 = - >-N- X x ff x3,0(8)d0 2 SN e(p), where e(p) — 0 as p — 0. Hence for p small, there is a 6 with ¥ x;,0(0) = N/2. If p is small enough, then the old attracting cycles of R still correspond to attracting cycles for Ry and at least half of the neutral cycles do also. By Theorem 2.4, R; has at most 2d —2 attracting cycles and this completes the proof. 0 Theorems 2.4 and 2.6 combined yield immediately the finiteness theorem for nonrepelling cycles, with the estimate 6d — 6. The proof method we have followed is due to Fatou. THEOREM 2.7. The total number of attracting and neutral cycles is at most 6d — 6. The sharp value for Theorem 2.7 is 2d — 2, due to M. Shishikura (1987), who obtained the result for his Master’s thesis. He showed, using quasiconformal surgery, that there is an analytic perturbation for which all neutral cycles become attracting (see [Be2]). We prove this later for polynomials (Theorem VI.1.2), in which case the sharp estimate of d — 1 bounded nonrepelling cycles is due to A. Douady (1982). IIL.3. Density of Repelling Periodic Points 63 3. Density of Repelling Periodic Points We can now prove the theorem, obtained independently by Julia and Fatou, which was mentioned at the beginning of the chapter. THEOREM 3.1. The Julia set J is the closure of the repelling periodic points. Proof. Suppose there is an open disk U that meets J and that con- tains no fixed points of any R™. We may assume U contains no poles of R nor critical values of R. If f1, fg are two different branches of R7! on U, then since there are no solutions of R™(z) = z in U, _R-fh z-h R"- fp 2z-fi omits the values 0, 1,00 in U. By Montel’s theorem, {gn} is normal and hence so is {R"}, a contradiction. Thus periodic points are dense in J. Since there are only a finite number of attracting and neutral cycles, and since J is perfect, the repelling cycles are dense. 0 In THEOREM 3.2. Let U be open, and suppose UN J #0. Then there is an integer N such that RN(UN J) = J. Proof. Let 29 € UN J be a repelling periodic point with period a rrrrrrrlr—s=—“‘=O™COCOCOCOCNSNC:iCiéCé:C‘C ‘i sststis)‘ac.éCOrtCtCéiiCtsCN R"G-Y(V) c R'(V), and by Theorem 1.5 J CO\E CUR, R™(V). Since J is compact, the Heine-Borel theorem implies J C R/"(V) for some j. Thus the theorem holds with N = nj. 0 Theorem 3.2 shows that J is “almost everywhere” conformally self-similar. More explicitly, if z € J has no critical point in its forward orbit, then within any neighborhood of any other point of J there is an open subset of J that is the conformal image of a neighborhood of zo. An open subset A of C is a conformal annulus if A can be mapped conformally to a genuine annulus {r < |¢| < s}. The modulus of A is defined to be log(s/r) times a normalization factor 1/27, and this is a 64 III. Basic Rational Iteration conformal invariant. A compact subset E of C is uniformly perfect. if E has at least two points and if the moduli of the conformal annuli in C\E which separate E are bounded. As an application of the Fatou—Julia theorem, we prove the following theorem of R. Maiié and L.F. da Rocha (1992). THEOREM 3.3. The Julia set J of R is uniformly perfect. Proof. Suppose there is a sequence {An} of conformal annuli in ce with moduli tending to 00, such that both components of C\An meet J. Let En be the component of C\A, with the smaller diameter (measured in the spherical metric). The diameter of E,, tends to 0, as can be seen by noting that there are closed geodesics (conformal circles) in A, whose hyperbolic lengths tend to 0 as n — oo, and by using the usual comparison of metrics. Let %n map the open unit disk conformally onto Ay U En, so that tn(0) € Ep. Then Ky = wn} (En) is a continuum in A containing 0. The modulus of A\K, also tends to oo, since it is the same as that of An. Hence the diameter of Kn tends to 0, as above. Fix 6 > 0 small, and assume that the diameter of Ep, is less than 6. In view of Theorem 3.2, there is a first integer kp such that the diameter of R*"(E,) exceeds 6. If gn = Rk o Wn, then the diameter of gn(Kn) is at least 6. Now the diameter of R*"(E,) is at most C6, where C is the Lipschitz constant of R (with respect to the spherical metric). Fix four points of 7 which are of distance greater than Cé from each other. Since gn(A\Kn) © F and gn(Kn) has diameter at most C6, each gn omits at least three of the four (fixed) points. Consequently {gn} is a normal family, and the diameter of gn(Kn) tends to 0. This contradiction establishes the theorem. 0 A simple scaling and normal families argument shows that confor- mal annuli of large modulus contain genuine annuli of large modulus. Thus the compact set E is uniformly perfect if and only if there is c > 0 such that for any finite z € E andr > 0 (and r < To when co ¢ E), the euclidean annulus {cr < |z— 20| < r} meets EF. Uniformly perfect sets were introduced by A.F. Beardon and Ch. Pommerenke (1979), who showed that if oo ¢ E, the compact set E is uniformly perfect if and only if the hyperbolic metric of the com- plement of E is comparable to the reciprocal of the distance to the boundary, that is, the factor log(1/6(z)) appearing in Theorem 1.4.3 TII.4. Polynomials 65 can be omitted. There are many other characterizations of uniformly perfect sets. One characterization is in terms of linear density of log- arithmic capacity (see [Po2]). This guarantees that Green’s function of any fixed component of the Fatou set extends to the boundary to be Hélder continuous, and this in turn implies that Julia sets have positive Hausdorff dimension. For more details, see Section VIII.3, where Hélder continuity is proved in the context of quadratic poly- nomials. 4. Polynomials In this section we consider the case where R = P is a polynomial of degree d > 2. We have a superattracting fixed point at oo. The iterates of P are bounded on the bounded components of ¥, by the maximum principle, so the basin of attraction A(oo) is connected. By Theorem III.1.7, the Julia set coincides with 0A(co). The filled-in Julia set of P, denoted by K, is defined to be the union of the Julia set 7 and the bounded components of F. Thus z € K if and only if the iterates P"(z) are bounded. This property can be taken as a basis for drawing computer pictures of K and of J. For K, color z purple if |P"(z)| < C for 1 r} of the level curve is invariant under P, and P maps it d-to-one onto {G > rd}. For r large, y(z) is defined on {G > r} and maps it conformally onto {|¢| > e"}. The equation y(z) = (y(P(z)))!/4 allows us to extend p(z) to {G > r/d} provided no critical point of P belongs to this domain. Now there are two cases to consider. If there is no critical point 66 III. Basic Rational Iteration FIGURE 1. Level lines of Green’s function with 7 totally disconnected. of P in A(oo), we can continue indefinitely to all of A(oo), and y maps A(oo) conformally onto the complement {|¢| > 1} of the closed unit disk in the ¢-plane. In particular, A(oo) is simply connected, and the Julia set J = 0A(oo) is connected. Otherwise we extend y until we reach a level line {G = r} of Green’s function that contains a critical point of P. The situation is then as follows. The domain {G > r} is simply connected and mapped by ¢ conformally onto {|¢| > e"}. The domain forms several cusps at the critical point, and ¢(z) approaches different values as z approaches the critical point through different cusps. The level line {G =r} consists of at least two simple closed curves that meet at the critical point. Within each of these curves there are points of J, or else G would be harmonic and positive, hence constant within the curve. Hence J is disconnected. In fact, in this case TI has uncountably many connected components. This can be seen by noting that the critical points of G are the critical points of P and all their inverse iterates, and by following the splitting of level curves at each such critical point. We have proved in particular the following. THEOREM 4.1. The Julia set J is connected if and only if there is no finite critical point of P in A(oo), that is, if and only if the forward orbit of each finite critical point is bounded. At the other extreme we have the following. III.4. Polynomials 67 THEOREM 4.2. If P"(q) — oo for each critical point q, then the Julia set J is totally disconnected. Proof. Let D be a large open disk containing J such that P(C\D) C\D. Choose N so large that P% maps the critical points of P to C\D. For n > N there are no critical values of P" in D, so that all inverse branches P~” are defined and map D conformally into D. Let zo € J. Then P"(zo) € J, and we define f, to be the inverse branch of P” which maps P"(zo) to zo. The f,’s are uniformly bounded on a neighborhood of D, hence they form a normal family there. Since fn(z) accumulates on J for z € DN A(oo), any limit function f maps D1 A(oc) into J. Since J contains no open sets (Theorem 1.9), f must be constant. Hence f(D) has diameter tending to zero. Since fn(OD) is disjoint from 7, {zo} must be a connected component of J, and J is totally disconnected. 0 If P is quadratic, then Theorems 4.1 and 4.2 cover all cases, and the corresponding Julia set is either connected or totally discon- nected. We study quadratic polynomials in detail in Chapter VII, in connection with the Mandelbrot set. IV Classification of Periodic Components We focus on the behavior of a rational function R(z) on the Fatou set F. Our aim is twofold: to show that every component of F is iterated eventually to a periodic component, and to classify the action of R(z) on periodic components. 1. Sullivan’s Theorem As before, we assume that the degree of R is d > 2. The image of any component of the Fatou set F under R is a component of F, and the inverse image of a component of F is the disjoint union of at most d components of F. The dynamics of R can be understood in part by determining how the various components of F are moved about by R. Consider a fixed component U of F. There are several possibilities for the orbit of U under R. 1. If R(U) =U, we call U a fixed component of F. 2. If R"(U) =U for some n > 1, we call U a periodic component of F. The minimal n is the period of the component. If n = 1, we have a fixed component. 70 IV. Classification of Periodic Components 3. If R™(U) is periodic for some m > 1, we call U a preperiodic component of F. 4. Otherwise, all {R"(U)} are distinct, and we call U a wandering domain. We have already seen several examples of fixed and periodic com- ponents of F, and we will soon see strictly preperiodic components. IN. Baker (1976) has shown that some entire functions have wan- dering domains. However, according to D. Sullivan (1985) this is not possible for rational functions, and our main goal in this section is to prove this. First we make some preliminary observations. On any component U of F, R is a branched cover of U over RU) with at most d sheets. A component of F is completely invariant if and only if R is a d-sheeted branched covering of the component over itself. THEOREM 1.1. If U is a completely invariant component of F, then QU = J, and every other component of F is simply connected. There are at most two completely invariant components of F. Proof. If U is completely invariant, then OU = J, by Theorem III.1.7. Moreover the sequence {R"} omits the open set U on C\U, so {R"} is normal there, and C\U Cc F. Since U is connected, each compo- nent of C\U is simply connected. If U is furthermore simply con- nected, then since R is a d-to-1 mapping, U must contain d—1 critical points. Since there are only 2d—2 critical points altogether, there can be at most two simply connected completely invariant components. Oo THEOREM 1.2. The number of components of the Fatou set can be 0, 1, 2, or co, and all cases occur. Proof. Suppose F has only finitely many components and let Up be one. Consider a chain of inverse images R(U_1) = Uo, R(U-2) = U_,:+:. Eventually we reach U_, with R(U_n) = U_, for some 0 areaU, < oo, which implies that the only possible limit functions of {R"} on Up are constants. In fact, (R”)’ — 0 on compact subsets for otherwise each U, would contain a disk with diameter bounded uniformly away from 0. Replacing Un by Unim, we may assume no U,, contains a critical point of R. We claim that R maps each U,, one-to-one onto U,+1. For this it suffices to prove that each {U;,} is simply connected. Let 7 be a closed curve in Up. Since (R")’ — 0 on Up, the diameters of the images R"(-y) tend to 0. Now each pole of R is contained in a component of F that is mapped to V, and the images of 7 are not mapped to V, so that if the diameter of R"(7) is sufficiently small, say for n > N, there are no poles of R in the bounded components of C\R"(7). Then the iterates R+* are all analytic inside C\R™ (7), and by the maximum modulus principle they are uniformly bounded there, so that the bounded components of C\.R (7) are included in F, hence in Uy. Then R'(y) is homotopic to a constant path in Uy. Since there are no critical points, RN is a covering map of Up over Un, and the homotopy can be lifted to a homotopy of 7 to a constant in Uo. Hence Up is simply connected, as is each U;, by the same argument. We will now construct a family of quasiconformal mappings { ft} with t € C’, so that ia o Ro f; is an N-dimensional analytic family of distinct rational maps. Since they have the same degree as R, this gives a contradiction if we take N > 2d — 1. a .LrCS—S—ti‘“—O—ss—s—s—eSsrsSsr—s—sNOCSCirzsSCSC(i“CisSNzsaCNCisiCN‘a‘CCN)SCSsCizds¥CD For t = (t1,...,t~w) in CN, |t| < 6, we define a Beltrami coefficient 72 IV. Classification of Periodic Components (ellipse field) 2 on Do by N a . = > te", z=re® € Dy. For n > 1, set Dn = R"(Do). Since the D,’s are disjoint and R maps each D, conformally onto Dn+1, we can extend the ee field to UD,, to be invariant under R. Each component of R7 1(UD,) is mapped finite-to-one onto one of the D,’s, so we can further extend the ellipse field to R~!(UD,), and proceeding backwards to all in- verse images of UD, to be invariant under R. Since R is conformal, the extension does not increase ||f||.0. For other z set = 0, and then the ellipse field is everywhere invariant. Note that » = 0 on Uo\Do. Let. f:(z) = f(z,t) be the solution of the Beltrami equation, nor- malized so f(z,t) = z + 0(1) at oo. Now (suppressing the t’s) let g =f. — 1 and recall from Section 1.7 that g = (I= Uy)7*S(u) = S(u) + Up(S(u)) + Up(S(H)) + Since ||j4\loo = O(\t|), also ||S(u)|l2 = O((#l), |]Uulle = (lel), and consequently g = S(u) + Olle) | in L?. Thus with ¢ = € + in = pe? we calculate for z € D\Dpo that 29 at; _ 9S(u) Oty —id “5 fee : Spkdntes(@) = satel) t=0 t=0 where the y;’s are analytic in D, and cj = —2e9+?(7+1)/(j +2) £0. It follows that if A = (Aj) € CX, A #0, then for some z € D\Do To ageo) #0, (1.1) because otherwise AjCj Lon =- Do Ag;, € <2] <2, which contradicts the uniqueness of Laurent expansions. From the invariance of the ellipse field, we see that R; = fro Ro (fe)7} is analytic. (See the remark at the end of Chapter I.) It is furthermore d-to-1. Thus it is a rational mapping of the same degree IV.1. Sullivan’s Theorem 73 as R. By Theorem I.7.6, f; depends analytically on the parameter t, and moreover fo is the identity z. Hence R; depends analytically on the parameter t, and Ro = R. Thus the (normalized) coefficients of R; are holomorphic functions of t and agree with those of R at t = 0. Write, assuming R(oo) = 1, aq(t)z4 + ag-124 | +--+ + ao(t) Ril) = a + bai(h2t!--- + bolt)” where aq(0) = 1 and a;,(t), 6;(t) are holomorphic in {|t| < 6}. Let V be the connected component containing 0 of the analytic subvariety of {|t| < 6} determined by the equations a;(t) = a;(0), bj(t) = 6;(0). There are only 2d—1 equations here, and so V has positive dimension. (For background on analytic varieties, see [GuR].) Let r € V. Then f,oRo(f,)~! = R. If zo is a fixed point of R”, then so is f-(zo), and vice versa, since f;(zo) = fr(R"(z0)) = R"(fr(z0))- If zo is a repelling fixed point for R” then so is f;(zo). Hence the set A, of such fixed points satisfies f;(An) = An. However, since fo is the identity and A, is discrete, f; must fix each point of An. But UA, is dense in 7, by Theorem II.2.8, and so f;(z) = z on J. Now f(z) is analytic on Up\Dpo and f,(z) = z on OUp C J. It follows that f;(z) = z in Up\Do. (For an elementary argument, map Up conformally to the unit disk and observe that f,(z) — z becomes analytic on an annulus {r < |¢| < 1} and tends to 0 as |¢| — 1, so is identically 0 on the annulus.) Now changing notation, we have f(z,7) — z = 0 for all z € Uo\Do, rt €V. Since g = f, — 1, we obtain g(z,7) =0, z€Up\Do, TEV. (1.2) If 7) is a regular point of V, there is an analytic map ¢ — 1(C) of a disk A(0,6) into V with 7(0) = 79 and 7/(0) = A, |A| = 1. From (1.2) and the chain rule we obtain 0 Sr Ajse (2,70) =0, 2 € Up\Do. Ot; Here the unit vector \ depends on 79, and passing to a limit of such \’s as T) — 0 we obtain a contradiction to (1.1). We conclude there are no wandering domains. 0 74 IV. Classification of Periodic Components 2. The Classification Theorem Having ruled out the possibility of wandering domains, we now know that every component of the Fatou set is periodic or preperiodic. Our next goal is to classify periodic components. We have already seen some examples: domains containing an attracting fixed point, and Siegel disks. There are various other possibilities. A periodic component U of period n of the Fatou set F is called parabolic if there is on its boundary a neutral fixed point ¢ for R” with multiplier 1, such that all points in U converge to ¢ under iteration by R". The domains U, R(U),..., R"-1(U) form a parabolic cycle. Their union is the immediate basin of attraction associated with an attracting petal at ¢. A periodic component U of period n of the Fatou set is called a Herman ring (or an Arnold ring) if it is doubly connected and R” is conjugate to either a rotation on an annulus or to a rotation followed by an inversion. We shall see in Chapter VI that Herman rings can occur and that moreover there can be only finitely many of them. The definition of Siegel disk is similarly extended to include periodic components U that are simply connected, on which some R” is conjugate to a rotation. Siegel disks and Herman rings are referred to as rotation domains. THEOREM 2.1. Suppose U is a periodic component of the Fatou set F. Then exactly one of the following holds: 1. U contains an attracting periodic point. 2. U is parabolic. 3. U is a Siegel disk. 4. U is a Herman ring. Proof. We may assume U is fixed by R. Since J has more than two points, U is hyperbolic. The proof will be organized as a series of lemmas in the context of analytic maps of hyperbolic domains. Let p = pu denote the hyperbolic metric on U. The first alternative of Theorem 2.1 is covered by the following lemma. LEMMA 2.2. Suppose U is hyperbolic, f : U — U is analytic, and f is not an isometry with respect to the hyperbolic metric. Then either IV.2. The Classification Theorem 75 f"(z) — OU for all z € U, or else there is an attracting fired point for f inU to which all orbits converge. Proof. Since f is not an isometry, p(f(z),f(w)) < p(z,w) for all z,w € U. In particular for any compact set K C U, there is a constant k = k(K) <1 such that (f(z), F(w)) S kplz,w), 2,we K. Suppose there is 29 € U whose iterates z, = f"(zo) visit some com- pact subset L of U infinitely often. Take K to be a compact neigh- borhood of LU f(L). Then p(zm+2,2m+1) < kp(2m+41, 2m) Whenever 2m € L, and this occurs infinitely often, so p(zn41,2n) — 0. Thus by continuity any cluster point ¢ € L of the sequence {z,} is fixed by f, and in fact is an attracting fixed point since p(f(z),¢) < kp(z,¢) in some neighborhood of ¢. Since the iterates of f form a normal family, they converge on U to ¢. 0 The third and fourth alternatives of Theorem 2.1 are covered by the following. LEMMA 2.3. Suppose U is hyperbolic, f : U — U is analytic, and f is an isometry with respect to the hyperbolic metric. Then exactly one of the following holds: 1. f"(z) — OU for all z EU. 2. f™(z) =z for all z €U and some fixed m > 1. 3. U is conformally a disk, and f is conjugate to an irrational rotation. 4. U is conformally an annulus, and f is conjugate to an irra- tional rotation or to a reflection followed by an irrational ro- tation. 5. U is conformally a punctured disk, and f is conjugate to an irrational rotation. Proof. Since f is an isometry with respect to the hyperbolic metric, f is a conformal self-map of U. Suppose first that U is simply connected. Let y map U conformally onto the open unit disk A. Then S = yo f oy™! is a conformal self- map of A, a Mobius transformation. If S has fixed points on the unit 76 IV. Classification of Periodic Components circle, then |S"| — 1 on A, and (1) holds. If S has a fixed point in the disk, we may assume it is at the origin. Then S$ is a rotation, and either (2) or (3) holds. Now assume U is not simply connected. Let yp : A + U be the universal covering map, and let G be the associated group of covering transformations, the group of conformal self-maps g of A satisfying pog =p. The lift of f to the unit disk via is a Mobius transfor- mation F, which satisfies poF = fo. Let be the group obtained by adjoining F to G. Assume first I’ discrete (orbits accumulate only on 0A). Since no iterate f* of f is the identity on U, no iterate F* of F belongs to G. Since T is discrete, this implies gF*(0) — OA uniformly in g € G. Hence f*(z9) + OU, and (1) holds. The final possibility is that I is not discrete. Let T be the closure of I in the (Lie) group of conformal self-maps of A, and let To be the connected component of F containing the identity. If g € G then also FgF~' € G, since po(FogoF") = fopogoF t= fopoF |! =fof tow=y. It follows that I’, and hence Ig, also conjugates G to itself. Since G is discrete and Tp is connected, hgh~! = g for all hE To and g €G, and every g € G commutes with every h € To. Lema. If A and B are two conformal self-maps of the open unit disk A which commute, and A is not the identity, then B belongs to the one-parameter subgroup generated by A. Proof. There are three cases: Case 1. Suppose A has a fixed point in A. We may assume the fixed point is 0, so that A(z) = ez. Then eB(0) = (AB)(0) = (BA)(0) = B(0). Since A is not the identity, B(0) = 0, and B has the form ez. Case 2. Suppose A has two fixed points on {|z| = 1} which are different. We can map the problem to the right half-plane with the fixed points going to 0 and oo, and A(z) = Az for some A>0, AF 1. ‘As above, B either fixes each of 0,00 or interchanges them. In the second case, B(z) = /z for some > 0 and does not commute with A. Therefore B fixes these points and B(z) = pz. IV.2. The Classification Theorem 7 CASE 3. Suppose A has one fixed point on {|z| = 1}. Again map the problem to the right half-plane with oo fixed. Then one sees easily that A(z) = z+ Ai for some real A # 0 and B(z) = z +p for some weR.O Mobius transformations corresponding to these three cases are called elliptic, hyperbolic, and parabolic, respectively. It is easy to show that every such transformation preserving the disk (except the identity) is one of these three types. Proof of Lemma 2.3 (continued). Choose h € Tg which is not the identity. Since G commutes with h it belongs to the one-parameter group generated by h. Since G is discrete and infinite we conclude that G has the form {g”}%,.. This means that the fundamental group of U is isomorphic to the integers, and U is doubly connected. Since U is hyperbolic, U cannot be a punctured plane, and U is either an annulus or a punctured disk. One of the alternatives (2), (4) or (5) must hold. 0 There is one more technical obstacle to surmount before complet- ing the proof of the classification theorem. It is handled by the fol- lowing variant of the “snail lemma”. The reason for the nomenclature will become apparent from the proof, which stems from Section 54 of [Fa2]. LEMMA 2.4. Suppose U is hyperbolic, f : U — U is analytic on U and across OU, and f"(zo) + OU for some 2 € U. Then there is a fixed point ¢ € OU for f such that f"(z) > ¢ for all z € U. Either ¢ is an attracting fixed point, or ¢ is a parabolic fixed point with multiplier f'(¢) = 1. Proof. By Theorem 1.4.3, the spherical distances between the iterates 2 and 241 of zo tend to 0. Thus the limit set of {z,,} is a connected subset of OU, and furthermore, by continuity of f, any limit point is a fixed point for f. Since the fixed points of f are isolated, we conclude that z, — ¢ for some fixed point ¢ € OU of R. The orbit of every other z € U also converges to ¢, since it remains a bounded hyperbolic distance from the orbit of zo. The fixed point ¢ is not repelling, since z, — ¢. Suppose f’(¢) = e?" where @ is rational. Then U is contained in 78 IV. Classification of Periodic Components the basin of attraction associated with one of the petals at ¢. Since the local rotation f at ¢ induces a cyclic permutation of the petals at ¢, and since f leaves U invariant, in fact f induces the identity permutation, and f’(¢) = 1. Suppose f’(¢) = 2 where @ is irrational. Assume ¢ = 0. Let zg € U, and let V be a relatively compact subdomain of U such that V is simply connected and zp and z; = f(z) belong to V. Since f is univalent near 0 and f” — 0 uniformly on V, we can assume each f” is univalent on V. Then f(z) £"(20)' is also univalent on V, Yn(zo) = 1, and 0 ¢ Yn(V). Let be the Riemann map from A to V, 7(0) = zo. Then Lr is univalent on A, hn(0) = 0, hi,(0) = ¢i,(20)¥/(0), and hn omits —1. Thus the function hn/h!,(0) belongs to S and omits —1/hi,(0). The Koebe one-quarter theorem implies |h},(0)| < 4. Since Sisa normal family, the sequence {hn/h',(0)} is normal on A, as is i Consequently {yn} is normal on V. We claim that all limit functions of {gn} are nonconstant. For if |~n — 1| < 6 then f"(V) would be included in a narrow angle with vertex 0, and if this angle were smaller than 6/3 then since f(z) = ez near 0, f"*1(V) would be disjoint from f"(V), contrary to hypothesis. Thus ¥7,(zo) is bounded away from 0. From Koebe’s theorem again we deduce that there is p > 0 such that Yn(V) con- tains a disk centered at 1 of radius p. Thus f"(V) contains a disk centered at zn of radius p|zn|- Choose N so that the disks of radius p/2 centered at the points e2rim9 Q < m < N, cover an annulus containing the unit circle. Since zm41 = 272m + 0(|2m|), the disks centered at zm of radius plZm|,n< mes n+N, cover an open annulus containing z, and 2n+41 for n large. Hence Uf”(V) contains a punctured neighborhood of 0, and 0 is an isolated point of OU. But then Theorem 11.6.2 implies that f is conjugate to a rotation about 0, contradicting f"(z) — 0. We conclude that f’(¢) cannot be irrational. 0 Qn(z) = zeéV, Proof of Theorem 2.1 (continued). The proof is now complete, in view of the following observations. Since R” has degree > 1, no power of R can coincide with the identity, and case (2) of Lemma 2.3 is ruled out. Since 7 has no isolated points, case (5) of Lemma IV.3. The Wolff-Denjoy Theorem 79 2.3 is also impossible for R. Finally, since there are no attracting periodic points in 7, Lemma 2.4 produces a parabolic fixed point in OU whenever f"(z) > OU. 0 EXAMPLE. Suppose R is a finite Blaschke product of degree d > 2. We have seen in Section III.1 that the Julia set 7 of R is either the entire unit circle or a totally disconnected subset of the circle. There are now four possibilities. If R has an attracting fixed point on 0A, or if R has a parabolic fixed point on 0A with only one attracting petal, then 7 is totally disconnected. If R has an attracting fixed point ¢ € A, then 1/€ € C\A is also an attracting fixed point, and J separates the basins of attraction, so J = OA. The only remaining possibility is that R has a parabolic fixed point on 0A with two attracting petals, in which case again J = OA. One can check that each of these four cases already occurs for Blaschke products of degree two. In all cases, the periodic points of R that are not fixed are repelling. 3. The Wolff-Denjoy Theorem Before leaving this area let us prove the following beautiful related theorem of J. Wolff (1926) and A. Denjoy (1926). This brief proof was discovered by A.F. Beardon (1990). THEOREM 3.1. Let f : A — A be analytic, and assume f is not an elliptic Mébius transformation nor the identity. Then there isa ¢ A such that f"(z) > a@ for allz eA. Proof. The theorem is easy if f is a Mobius transformation, and so we assume that f is not an isometry with respect to the hyperbolic metric, and further that f(0) 4 0. If the orbit of 0 visits any compact set infinitely often, Lemma 2.2 provides a fixed point in A. Thus we may assume that the orbit of 0 accumulates on 0A. The problem is to show that the orbit can accumulate at only one point a of 0A. This is again easy if f extends continuously to 0A. The main point of the theorem is that no continuity is assumed. Define f-(z) = (1—¢) f(z), which maps to a compact subset of A. Let z- be the fixed point of f,, and let D, be the hyperbolic disk centered at z- with radius p(0, z-). Since f. is contracting, f-(D-) C 80 IV. Classification of Periodic Components D-. Now D, is a euclidean disk with 0 on its boundary. Any limit D of the D.’s is a euclidean disk with 0 on its boundary, and f(D) C D. Thus the point of tangency of D and 0A is the only possible limit point on 0A of the orbit of 0. 0 Another approach to the Wolff-Denjoy theorem provides an illu- minating application of the Herglotz formula. We replace A by the upper half-plane H. We are now considering f =utiv as a mapping of H to itself, with v = _Im f > 0, and we must show that either every orbit tends to 00, or else every orbit is bounded. The Herglotz repre- sentation of the positive harmonic function v in the upper half-plane is v(a,y) = cy + vf do(s) y>0, (3.1) ~oo (a — 8)? + y?’ where c > 0 and a is a positive measure satisfying [- do(s) a coo 1+? , We may assume o # 0. Suppose c > 1. From (3.1) we have v(z,y) > y, so that the half- planes {y > yo} are invariant under f. These correspond to the invariant tangent disk D to OA in the preceding proof. Let zn = 2n + iyn be an orbit. Then either yn increases to +00, or else Yn increases to some finite limit value yoo, in which case (3.1) shows that |¢n| — oo. In any event |zn| — 00. Now suppose 0 < c < 1. In this case orbits are bounded. To see this, we use the Herglotz formula for f, f()zee+ [ 2) -co §$—Z% +6, zéEH, where 6 is real. The problem is to obtain an estimate of the form |f(z) — ez| = 0(z) for large |z|. This can be done by expressing f as 1 A cet | - «| da(s) + f da(s) +0! \sj>als—z 8§ AS—2 do(s) ore Joos s(s — 2) +O) as |z| — oo and by estimating carefully the integral that appears here. f(z) ll V Critical Points and Expanding Maps Critical points and their forward orbits play a key role in complex dynamical systems. The forward orbit of the critical points is dense in the boundary of any Siegel disk and Herman ring. If the critical points and their iterates stay away from the Julia set, the mapping is expanding on the Julia set, and the Julia set becomes more tractable. 1. Siegel Disks Let CP denote the (finite) set of critical points of R. The postcritical set of R is defined to be the forward orbit Un>0R"(CP) of the critical points. We denote by CL the closure of the postcritical set of R. This set is important because on its complement all branches of R7", n > 1, are locally defined and analytic. As we have already seen for polynomials (Section III.4), many basic properties of the dynamics and structure of 7 are determined by the critical points and the postcritical set. By Theorems III.2.2 and III.2.3, each attracting and parabolic cycle of components of the Fatou set F contains a critical point. Siegel disks and Herman rings contain no critical point, however in a certain sense they can be associated to critical points. We begin with the following theorem. 82 V. Critical Points and Expanding Maps The proof is essentially that of Section 31 of [Fa2] and predates the establishment of the existence of rotation domains. THEOREM 1.1. IfU is a Siegel disk or Herman ring, then the bound- ary of U is contained in the closure of the postcritical set of R. Proof. Let U be a rotation domain that is invariant under R, and suppose that C'L does not contain 0U. Let D be an open disk disjoint from CL which meets OU. We assume also that D is disjoint from some open invariant subset V # @ of U. Define fn to be any branch of R-” on D. Since the fn’s omit V, they form a normal family on D. Now R is one-to-one on U, so there are other components of R71(U). Since inverse iterates of any fixed point of J are dense in J, there is for suitable m > 1 a component W of R-™(U) distinct from U that meets D. If z € DNW, then f;(z) and f,(z) belong to different components of ¥ for j # k, or else they would belong to a periodic component, which could not be iterated eventually to W then U. Hence f,(z) tends to J for z € DOW, and since J has no interior, any normal limit of the f,’s is constant on DOW. On the other hand, since the f;’s are rotations of U, any normal limit is nonconstant on DMU. This contradiction establishes the theorem. ao THEOREM 1.2. If the postcritical set of R is finite and there are no superattracting cycles, then J =C. Proof. Siegel disks and Herman rings require that CL be infinite, since CL is dense on the boundary. So does a parabolic component, since it contains a critical point and the iterates of this point are distinct. According to the remark after Theorem II.2.4, an attract- ing component that is not superattracting has a critical point with infinite forward orbit. Thus if CL is finite, the only possibility for a periodic component of ¥ is a superattracting cycle. By hypothesis, there are no such components, and so F is empty and J = C.a EXAMPLE. Take R(z) = 1—2/z?. The critical points of R are 0 and oo, which are iterated as follows: 0-0 -1--1l--l--:::. Thus R can have no superattracting cycles, and J = Cc. V.1. Siegel Disks 83 THEOREM 1.3. If there are two completely invariant components of F, then these are the only components of F. Proof. As remarked earlier (in the proof of Theorem IV.1.1), each completely invariant component of F contains d — 1 critical points, and there are no other critical points. Since rotation domains have no critical points, each completely invariant component is attracting or parabolic. Moreover, the iterates of each critical point tend to an attracting or parabolic cycle. In view of Theorem 1.1, there can be no Siegel disks or Herman rings, nor can there be other parabolic or at- tracting components, since they would require further critical points. Thus the completely invariant components are the only components of F.0 We show in Section VI.4 that if there are two completely invariant components of F, one of which has an attracting fixed point, then the Julia set 7 is a simple closed Jordan curve. It follows from Theorem VI.2.1 that if the two completely invariant components both have attracting fixed points, then 7 is a quasicircle. EXAMPLE. For 0 < 6 < 1, the function R(z) = z/(z? — bz +1) hasa parabolic fixed point at z = 0 with multiplier +1 and an attracting fixed point at b with multiplier 1 — b?. The attracting petal at 0 and the immediate basin of attraction at b contain the critical points —1 and +1, respectively. Since R has degree two and is two-to-one on each immediate basin of attraction, each immediate basin is com- pletely invariant. By Theorem 1.3, there are no other components of F. As mentioned above, J is a simple closed Jordan curve, passing ° through 0 and oo. For b = 0, we have only the parabolic fixed point at 0, which has two completely invariant attracting petals. The re- spective petals include the right and left half-planes. Since the Julia set separates the petals, it coincides with the (extended) imaginary axis. The classification theorem shows that the iterates of any point of F either remain in a compact subset of F or converge to a parabolic cycle in J. It follows that the boundary OU of any Siegel disk U lies in the closure of the forward orbits of those critical points that lie on 7. We claim that in fact there is a single critical point in J whose forward orbit is dense in QU. Indeed, since there are only 84 V. Critical Points and Expanding Maps OE Critical point FIGURE 1. Iterates of a critical point delineate a Siegel disk. finitely many critical points, there is a critical point gq € J whose forward orbit is dense in a relatively open subset W # @ of OU. Replacing W by UR"(W), we can assume that W is invariant under R. Let » : A > U be a conjugation of multiplication by A on A and R on U. Let S be the set of ¢ € 0A at which ~ has a radial limit belonging to W. Since W is open and nonempty, S has positive length on 0A, and further S is invariant under the rotation ¢ > AC. If Cp is any point of Lebesgue density of S, then so is each rotate "Co, and evidently every point of OA is a point of Lebesgue density of S, so that S has full measure. It follows that W is dense in OU, and the claim is established. EXAMPLE. Consider P(z) = A(z — 2”/2) where = exp(27i(V5 — 1)/2). The golden mean (V5 — 1)/2 is a favorite choice of rotation angle for obtaining computer-generated pictures of Siegel disks. From its continued fraction expansion ee ae 1 ee ie I+ it is the worst approximable number in the sense of Diophantine ap- proximation (see p. 164 of [HaW]). By Siegel’s Theorem II.6.4, the fixed point 0 of P is the center of a Siegel disk. Figure 1 shows 20,000 iterates of the critical point 1 under P, which delineate the Siegel disk. (A similar picture of a Herman ring is given later in Section V.1. Siegel Disks 85 FIGURE 2. Siegel disk and enlargement at critical point. VI.3.) On the left of Figure 2 is the filled-in Julia set for P, and on the right an enlargement around the critical point 1. An interesting question concerns whether the critical point is an accessible bound- ary point of A(oo) or not, that is, the end of a curve in A(oo). This seems to be the case in Figure 2. Let R(z) = Azt+agz2?+---, where 0 < |A| < 1, and suppose z = h(¢) conjugates the multiplication ¢ + A¢ to R(z). If y is Koenigs’ coor- dinate function (appropriately normalized), then h is the branch of vy! mapping 0 to 0. Schréder’s equation for the conjugation assumes the form AG/A)= RMAC), Mel 1. Let y be Koenigs’ coordinate function, normalized by (0) = 1. Then y depends analytically on A. Since conjugating functions are unique up to dilations, and since h-1(1, A) = 1, we have h-\(z,A) = pr(z)/pa(1). Thus h-}, and also h, depend analytically on 4. The coefficient estimate of Theorem 1.1.8 gives |ax(A)| < ek?|ai(A)| < 4ek?, so each a,(A) extends analytically to 0, and the series expansion extends h analytically to the bidisk |¢| < 1,|A| < 1. Furthermore, the coefficient estimates, or the distortion theorem, show that for each fixed ¢, h(¢, A) is bounded in A. (It is easy to show that |h| < 4.) The analyticity of h(¢,) in A can also be seen in a more ele- mentary way, following Yoccoz, by considering the sequence of poly- nomials Qn(A) = A" Px(1). Thus Qi = 1/2, Qo = 1/2 — +/8, V.1. Siegel Disks 87 Q3 = 1/2—A/8— 2/8 + 3/16 — 4/128, ..., and in general Qn41 = Qn — °7!Q?./2. One shows directly that |Qn(A)| < 4 for |A| < 1, and Qn(A) > a1(A). Let E be the set of e € OA such that ai(A) > 0 as A e®?. Then E has zero measure. (This is an elementary fact, which can be proved by applying Fatou’s lemma to the integrals of log la; (re®)| as r — 1.) Suppose e” ¢ E. Choose A, € A such that \, > e! and ay(Ax) + a # 0. Let ¥(¢) be any limit of the univalent functions h(C, Ax). Then (0) = a, and v is univalent. The functional equation for the conjugation h is AAG, A) = Pa(h(GA)), Ie 3 log +Co l-r for all r > r(A). In order to place the critical point 1 on the boundary of the Siegel disk, we want to show that 1 is a limit point of h(r, e%) as r — 1. For this, it suffices to show lim sup u(r, e®) = oo. Thus the estimate above is what we need, except that it is not uniform in A. 88 V. Critical Points and Expanding Maps For ) fixed, 0 < |A| < 1, the function 1 — h(z, A) can be expressed as the quotient of two functions in S, namely, h(z, A)/ai(A) and h(z, A)/[ax(A)(1 — A(z, A))]- The upper and lower estimates in the distortion theorem (Theorem I.1.6) then yield (1 — lel)? lz| (= lel)? |L—h(z,A)] 2 (ae : W+lay? 2 I Hence u(r, A) < 2 log log 4. eee This estimate persists for the radial boundary values at rA=e, Fix 0 clog for some k > N}. 1-r, It suffices to show that Ay has full measure for all N. Indeed, then NAn has full measure, and for almost all @ we obtain a sequence of r’s tending to 1 such that h(r, 9) 1. For arbitrary 4) consider \ = pe!%, Fix M large, and let I be the interval of length M(1 — |A|) centered at e%, The Poisson kernel P,(6) = (1 — |Al?)/|e® — Al? satisfies En! ¢ [rod saat +s for any subset E of 0A, where | -| denotes angular measure. Choose k so large that k > N and rz > r(A). From the Poisson formula we obtain 5 los imn +Cy < ulre,dA) = [ rOueneD < clog 7 = + I. Py(B)u(rase) Se < clog + + 20g || PA(O)57 + los < log fer zeueay TT FF] + log 4. Solving for |Ay N1|/|Z| and letting rx — 1, we obtain A I 1 ee sam (3-¢- 8) ae for some suitable cp. Hence Ay has positive density everywhere. By Lebesgue’s differentiation theorem, Ay has full measure. 0 V.2. Hyperbolicity 89 2. Hyperbolicity A rational function R is hyperbolic (on J) if there is a metric o(z)|dz| smoothly equivalent to the spherical metric in a neighborhood of 7, with respect to which R is expanding. If co ¢ 7, this means that o(R(z))|R(z)| 2 A(z), zz ET, (2.1) for some fixed A > 1. Then the identity o(R"(z))(RY(z) _ Ty o(R*1(2)) o(2) pao (R*(z)) shows that o(R"(z))|(R")'(z)| > A"o(z) for all n > 1. R'(RK(2)) LEMMA 2.1. Suppose co ¢ J. Then the following are equivalent. 1. R is hyperbolic on J. 2. There exist a > 0 and A > 1 such that |(R")'| > aA" on J for alln > 1. err r—S=s—=tsess—S—s—s—~=s—=é—e™—ONCsSSsKsisisCS=isésitrésaCizs a/c* for some 0 < c < 1. If z and w lie ina disk in V, and if R-* is any analytic branch on the disk, then |R-K(z) — R-(w)| < vee =u Hence ea < Co|R-*(z) — R*(w)| < Cic*|z — wl. ) 1 R(R*F(w)) V.3. Subhyperbolicity 91 Using the chain rule, we obtain (R-"Y(w) RR n-1 (R-")'(z) = I R(R- mEKay ym A (1 +c lz — w|O(1)) _ i i |z- iO ; where the bounds are independent of z, w,n, and the branch of R~”. This shows that branches of R~” are uniformly close to being affine on small disks. Now fix ¢ > 0 small, and note that disks A(z,¢), with z € J, have at least a fixed proportion (say 6 > 0) of their area in F. Let zm € J. Choose n large, and set z, = R"(zo). Now R~"(A(zn;€)) is approximately a disk with at least 6/2 of its area in F, and its diameter tends to 0 as n — oo. Hence zp cannot be a point of full area density of 7. By the Lebesgue differentiation theorem, J has zero area. 0 It is conjectured that for rational maps either J has zero area or J = C. There is a strong analogy with Kleinian groups, which was recognized early by Fatou (1906) and which is still not com- pletely understood (see [Su2], [ErL]). The corresponding conjecture for Kleinian groups is that the limit set of a Kleinian group has zero area. 3. Subhyperbolicity It is useful to consider a slightly weaker form of the notion of an expanding map. Assume oo ¢ J. Consider distances measured by o(z)|dz| where o > c > 0, as above, and denote the corresponding distance function by d,(z1, 22). We allow o to blow up at a finite number of exceptional points ay, ...,a@q but assume Cc Sn oF for some 3 < 1, so that d,(z,a:) < oo. Under these conditions, we say o is an admissible metric. Note that elz — w| < dy(z,w) < Clz- |"? 92 V. Critical Points and Expanding Maps for z and w near J. We call R subhyperbolic if R is expanding on 7 for some admissible metric, that is, there is A > 1 such that (2.1) holds in a neighborhood of J. If R is expanding on a neighborhood of J with respect to any metric (equivalent to the euclidean metric), then there are no neutral periodic points in 7. Indeed any periodic point zp then has an open neighborhood that is mapped to a compact subset of itself by the appropriate power R~™, and this implies \(R™)'(zo)| > 1. In the subhyperbolic case, there are further restrictions. THEOREM 3.1. Suppose J # C. Then R is subhyperbolic if and only if each critical point in J has finite forward orbit, while each critical point in F is attracted to an attracting cycle. Proof. Suppose R is subhyperbolic. The condition (2.1) shows that each critical value in 7 is an exceptional point for the metric a, as is any iterate of an exceptional point. Hence CL 7 is finite. By Theorem V.1.1, there are no rotation domains in F. By the remark preceding the theorem, there are no parabolic periodic points. We conclude that F consists only of basins of attraction for attracting cycles. For the converse, assume the condition on CL holds. Then all the components of ¥ are associated with attracting cycles, and there is an open neighborhood V of J such that R-'(V) C V and VNCL C J. We are assuming oo ¢ J. Let {a}, ...,@g} be the critical values in J and their successive iterates. We build an N-sheeted branched covering surface W over V branched over the a;’s so that any branch of R7!, defined locally from W to W, can be continued analytically along any path in W. For this there should be N/n(a;) branch points over aj, each of order n(a;). The integers n(aj) are chosen so that, if a € R-4(aj) and v(a) is the order of R at a, then n(aj) is an integral multiple of n(a)v(a). A moment's reflection shows this is always possible, the key point being that the cycles in CLO J contain no critical point or we would have a superattracting cycle. (If for instance there is only one critical point in 7, of order n, take W to be n-sheeted, with a single branch point of order n over each _ The universal covering surface of W is the open unit disk A, by Theorem I.3.3. Any locally defined inverse branch of R™! on W lifts to an analytic map of A onto a proper subset. Hence branches of R7! V.4. Locally Connected Julia Sets 93 are locally strict contractions with respect to the hyperbolic metric of W. Parametrize the kth sheet of W by z and express the hyperbolic metric there as pz(z)|dz|. Define o(z) = ¥ pr(z). Then o is smooth except at the branch points, and o(z) ~ |z — a;|71+1/n(@) near aj. The branches of R~! are strictly contracting with respect to each px(z)|dz|, hence with respect to o(z)|dz|, so R is strictly expanding. Qo The inverse R~! of R remains strictly contracting with respect to any small perturbation of the hyperbolic metric on W. If we perturb p to be a constant times |dw| at branch points, where w is the local coordinate satisfying z — aj = w’(s), and then sum over sheets, we obtain an admissible metric o for which R is strictly expanding, such that o is smooth away from the a;’s, and o(z) is a constant multiple of |z — aj|7141/"() near aj. EXAMPLE. Consider P(z) = 2? + 7. Our metric should have singu- larities on the postcritical orbit i i — 1 — —i. Define 1 0) = Taries er so that o(P(z)) = |z[2 2? +.1)71/? [22+ 2i|-1/2. The inequality to be proved, o(P(z))|P"(z)| > Ao(z), can be simplified to |z + (é— \i< 4/A”, which should hold for z near J. If z € J, then |z| < V3, and then |z + (i —1)| < V3 + V2 < 3.2. So the inequality holds on TI with A= /4/8.2>1.1. In Section VII.3 we extend the local affine estimate used in the proof of Theorem 2.3 to subhyperbolic maps. From this it follows that 7 has zero area also when R is subhyperbolic. 4. Locally Connected Julia Sets In this section we return to the case where R = P is a polynomial of degree d > 2. Recall (Section III.4) that A(oo) is connected and J = OA(co). We are interested in determining when J is locally connected. 94 V. Critical Points and Expanding Maps THEOREM 4.1. If the polynomial P is subhyperbolic on J, and if J is connected, then J is locally connected. Proof. Since J is connected, A(oo) is simply connected. By Theo- rems 3.1 and III.4.1, A(oo) contains no (finite) critical points. The conjugation of P to ¢4 at co extends to an analytic map y : A(oo) > {|¢| > 1} satisfying y(P(z)) = ¢(z)4. For fixed R > 1 we parametrize closed curves Yn by am(e®) =p (RY), 0S OS 27, n> 1. The yn’s tend to J as n — oo, and P is a d-to-1 cover of Yn41 over ‘m. Let o be an admissible metric for which P is strictly expanding near J. If 0 < c < 1 is the contraction constant for P~! near J, then for n large, dy (ne), Yn+1(€)) cde(P(mn(e?)), Perms (€”))) = cdo(>n-1(e*), yn(e*)). This shows that y,(e) converges uniformly as n — oo. The limit function maps the circle continuously onto J. Hence J is locally con- nected, and in fact the conformal map vy! extends to a continuous map of {|¢| > 1 } onto A(oo) UJ. IA THEOREM 4.2. Suppose P is a polynomial with finite postcritical set, such that no finite critical point is periodic. Then the Julia set J is a dendrite, that is, a compact, pathwise connected, locally connected, nowhere dense set that does not separate the plane. Proof. No components for F except A(oo) are possible, and Theorem 4.1 applies. 0 EXxampLe. If P(z) = 2? — 2, then 0 > -2 - 2 + 2—--,, so the Julia set is a dendrite. In fact, we already know J = [—2, 2]. A more interesting case is P(z) = z* + i, where 0 is iterated to a cycle of length two:0 > i ~ i-l1 > -it > i-1 — ---. See Figure 3. A crucial ingredient of the proof of local connectedness is the exis- tence of an expanding metric. When there are parabolic cycles, such a metric cannot exist in a full neighborhood of the Julia set. How- ever, under certain conditions such metrics exist in the part of A(co) bordering 7. We follow Exposé No. X of [DH2]. V.4. Locally Connected Julia Sets 95 FIGURE 3. Julia set for 2? +i. For simplicity, suppose there is a parabolic fixed point 29 € J and that the forward iterates of each critical point either tend to an attracting cycle or to z9 through an attracting petal. In view of the description in Section II.5 of the behavior of P near a parabolic fixed point, we see that we can construct an open set U containing J\{zo} such that U is invariant under branches of P~!, UM A(oo) = A(0, R)NA(co) for some large R, and UN A(z, €) is a finite union of narrow cusps terminating at zo and tangent to repelling directions. For ¢ > 0 sufficiently small, P is expanding with respect to the euclidean metric |dz| in each of these cusps, as can be seen from a simple calculation. Curves in A(oo) terminating at zo do not have finite length with respect to dpy. To remedy this, we aim to modify dpy to be a multiple of |dz| near zo. This is done as follows. Choose M so large and « so small that P(z) is expanding in a neighborhood of any point of P~!(A(zo,¢)) MU from the metric 96 V. Critical Points and Expanding Maps dpy to the metric M|dz|. Set V = P-1(U), an open set containing J\{z0} on which P is locally expanding with respect to dpy. De- fine a metric do(z) on V by setting do(z) = dpy(z) on V\A(20,€) and do(z) = min{dpy(z), M|dz|} on VO A(zo,€). The condition on M guarantees that P is locally expanding with respect to 7 on V\A(zo,€)- Since P is expanding with respect to both M|dz| and dpy in VNA(zo,€), it is expanding with respect to their minimum, hence it is locally expanding on V with respect to do. Integrating along paths in V and taking an infimum, we obtain a uniformly bounded metric d,(z,w) on V for which P is strictly expanding (though not uniformly expanding). Such a metric can also be constructed when there are several parabolic cycles, and even when there are critical points on J so long as they are strictly preperiodic, as in Section 3. With such an expanding metric in hand, we obtain the following. THEOREM 4.3. Let P be a polynomial with connected Julia set J, such that each critical point of P belonging to J is preperiodic. Then T is locally connected. Proof. Assume for convenience there is only one parabolic point zo, with p invariant petals. Let do(z) be as above. Let the curves Yn be as in the proof of Theorem 4.1, so that P is a o-expanding map of each yp onto Yn—1. Let Qn be the geodesic arc in A(oo) from in (e”) to Yni(e®), and for 1 < j 0 be small. Suppose first that more than 3n/4 of the aj’s meet A(z0,€)+ We arrange these in k blocks of consecutive iterates, with mj arcs in the jth block. Then k 3n/4. Estimating with the aid of the Fatou coordinate for a repelling direction at zo, we see that for any fixed 0 < 6 < 1/p, the o-length of a; is magnified by a factor at least O(m}**) over the jth block. Thus the g-length of ap is at most O((m1-- -mg)~“+9)) times the o-length of ay. Now m1 --- mx is minimized, subject to the constraints ™; >1 and m, +--+: + mx fixed, when all the m,’s except one are 1, with minimum m,+---+mg—(k—1) > n/2. Thus the o-length of ap is at most O(n~"'+9)) times the length of a1. Suppose, on the other hand, that at least n/4 of the a;’s lie outside A(zo,€). On these a;’s, P is uniformly expanding, so the o-length of a, is at most c-/4 times the length of a1, where C > 1 is an expansion constant. In any event we obtain do (Yn+1(€%), In (€*)) = O(n-"+4)), the estimate being uniform in 0. It follows that the curves yn converge uniformly in V.4. Locally Connected Julia Sets 97 the o-metric. Since dg(zn, Wn) — 0 implies [zn — wnl — 0, the 7n’s converge also uniformly in the euclidean metric, and the limit 7 is locally connected. 0 EXAMPLE. Suppose the Fatou set of the polynomial P consists of two completely invariant components A(oo) and W. Then all d—1 finite critical points of P belong to W, and the preceding theorem applies. The Riemann map from A to A(oo) extends to map 0A continuously onto J. Since J is adherent to W, the Riemann map must be one-to-one on 0A, and 7 is a simple closed Jordan curve. This applies for instance to the polynomial z? + 1/4, which has a parabolic fixed point at 1/2 with one attracting petal. The Julia set of 2? + 1/4 is the cauliflower set in Section VIIL.1, Figure 5. Lest the impression be left that 7 is always locally connected, we prove the following theorem of Douady and Sullivan (cf. [Sul]). THEOREM 4.4. Suppose the quadratic polynomial P has an irra- tionally neutral fixed point that does not correspond to a Siegel disk. Then the Julia set J is a connected but not locally connected set. Proof. We use the result, proved in the next chapter (Theorem VI.1.2), that a polynomial of degree d has at most d—1 attracting or neutral cycles (not including 00). In the case at hand, there are no other attracting or neutral cycles than the irrationally neutral fixed point, so that F = A(oo). We use also the fact that A(oo) is simply con- nected. By Theorem III.4.1, this is equivalent to the boundedness of CL. This in turn follows from Theorem VIII.1.3, and it can also be established by a direct argument. We assume z = 0 is the irrationally neutral fixed point. Let y be the conjugation of P to ¢? at oo. Since A(oo) has no finite critical points, y extends to a conformal map of A(co) onto {|¢| > 1}, whose inverse yp : {|¢| > 1} + A(oo) satisfies P((¢)) = (C7). Suppose ~ extends continuously to {|¢| > 1}. Then y71(0) is a closed subset of the unit circle {|¢| = 1} of zero length. Let 7 > 0 be small, and let D,, denote the set of z € A(oo) such that z = ¥(C) for some ¢ € A satisfying dist(¢,~1(0)) < n. Evidently Dy is open, and the continuity of ~ shows that D, shrinks to {0} as 7 decreases to 0. Thus we can assume that the branch f of P~! satisfying f(0) = 0 is defined and analytic on a disk containing D,. 98 V. Critical Points and Expanding Maps We claim that D, is invariant under f. Indeed, let zo € Dy, and write z9 = W(Co)- Choose 0, |€o| = 1, such that w(€o) = 0 and |Go — &o| < . Let 21 = f (20). Then z; = ~(¢1), where C=0,0 = VO. Set £1 = V€o, where we use the same branch of the square root. Since ¢; and €; are approximately in the same direction, |¢1+&1| > 1. Hence [C1 — £1] < |G — €1lli + al = G7 — &fl = [G0 — Sol <2 Also P(W(E1)) = ¥(E2) = P(Eo) = 0. Hence w(E1) is either 0 or —A, and since €; is near ¢ and (G1) = 21 is near 0, (1) = 0. Thus z1 € Dn, and D,, is invariant under f. Now D, includes A(oo) {|z| < 6} for some 6 > 0, by the uniform continuity of p. Thus if z € A(oo), |2| < 6, then all iterates f(z) are in D,. Since J = 0A(oo) has no interior points and F = A(oo), Dy is dense in {|z| < 6}. Thus by continuity, f™(z) € Dy for all z,|z| < 6, and all m > 1. By Theorem II.6.2, f can then be conjugated to a rotation, hence so can P, contrary to hypothesis. This contradic- tion shows that does not extend continuously to {|¢| = 1}. By Carathéodory’s theorem, J = 0A(oo) is not locally connected. 0 VI Applications of Quasiconformal Mappings One of the basic ideas behind the use of quasiconformal mappings is to consider two dynamical systems acting in different parts of the plane and to construct a new system that combines the dynamics of both. This procedure is called quasiconformal surgery. 1. Polynomial-like Mappings Let U; and U2 be bounded, open, simply connected domains with smooth boundaries, such that U; C U2. Let f(z) be holomorphic on U, and map U, onto U2 with d-fold covering, so that f maps QU, onto U2, that is, f is proper. Following Douady and Hubbard (1985), we call the triple (f;U1,U2) polynomial-like. This name is motivated by the following. THEOREM 1.1. If (f;U1, U2) is polynomial-like of degree d, then there are a polynomial P of degree d and a quasiconformal mapping y with y(z) = 2+0(1) near oo, such that f = yo Poy! on Us. Proof. We construct a Beltrami coefficient (ellipse field) for y as follows. Take p > 1. Let (z) map C\U2 conformally to {|¢| > pt} with &(co) = oo. We wish to extend ® to C\U; so that ® maps 100 VI. Applications of Quasiconformal Mappings U2\U, one-to-one smoothly onto {p<\dl< p*}, so that for z € 0U1, O(z)4 = &(f(z)). To do this, note that f describes 0U2 d times as z runs through QU}, so ® can be defined on OU; and then extended arbitrarily in a smooth way to U2\U;. We define _ f(z), zeEU,, 9(2) = { 1 o@(z)4, ze C\U;. Define an ellipse field by making it circles in C\U2 and defining it to be g-invariant on oo U 9-"(C\U2). n=1 This can be done uniquely as the sets g(U2)\U2, Uo\Ui, 971 (U2\Ui), ..., are disjoint, and each is mapped d-to-1 onto the preceding one by g. Furthermore, the circles are distorted only on the first iteration (inside U2\U1), since g = f is analytic in U;. The eccentricities of the ellipses do not change after the first iteration and so are uniformly bounded. Hence |p| < k < 1. We define » = 0 on any remaining part of C. Let y solve the Beltrami equation corresponding to pz. Then P = yogop™! maps infinitesimal circles to circles and is a d-to-1 mapping, hence a polynomial of degree d. Finally, g = gp loPoy=f inh, as desired. 0 We can use Theorem 1.1 to prove the statement used in the proof of Theorem V.4.4. THEOREM 1.2 (Douady). A polynomial P of degree d has at most d—1 nonrepelling cycles in the finite plane. Proof. Let M be the set of points occurring in nonrepelling cycles. By Theorem III.2.7, M is finite. Choose a polynomial Q such that Q(z) = 0 for all z € M, and }/ Re (Q'(z;)/P"(z;)) < 0 for all neutral cycles {21,..., 2m}. To see that this is possible, note that P’(z;) #0 and that the conditions on Q are linear, so the conditions can be satisfied if the degree of Q is large enough. Suppose € > 0 is small, and define f = P + €Q. Then since Q = 0 on M, the nonrepelling V1.2. Quasicircles 101 cycles of P are also cycles of f, and they must be attracting, since for any neutral cycle for P, Q'(z3) Slog |P'(z;)| + So log P'(z;) O+e> Re (3) + O(c?) <0 l+e Vlog | f(z) I MI if ¢ is small enough. Hence all nonrepelling cycles for P are attracting for f. Assume P is monic, take p large, and set Uz = A(0, p*), U1 = f-(Ug). If € is sufficiently small, then f ~ P ~ 24 near {|z| = p}, so f is a d-to-1 proper mapping from U; to U2, and (f;U1, U2) is polynomial-like. We can assume M is included in Uj. By Theorem 1.1, f is conjugate to a polynomial S of degree d, and every at- tracting cycle for f must be attracting for S. By Theorem TII.2.2, a polynomial has at most d— 1 attracting cycles (not counting oo). Thus P had at most d— 1 nonrepelling cycles. 0 Since oo is an attracting fixed point of order d — 1, Douady’s theorem above is the same as Shishikura’s theorem mentioned in Section III.2 in the case of a polynomial. EXAMPLE. Suppose P is a quadratic polynomial that has a cycle corresponding to a Siegel disk, U) + U2 > --- > Um > U7: The components of F in the cycle are mapped one-to-one onto them- selves by P. Each Uj in the cycle is the image of U;—1 and one other bounded component V; of F. Each V; has two distinct preimage components, each of these has two further distinct preimage com- ponents, and so on, so that the inverse images of each V; form a tree. By Theorem 1.2, there are no attracting or neutral cycles other than the Siegel cycle (and oo). From the classification theorem, we conclude that each bounded component of F is eventually iterated by P to the Siegel cycle. 2. Quasicircles A Jordan curve I is called a quasicircle if it is the image of a circle under a quasiconformal homeomorphism of the sphere. Quasicircles can be geometrically characterized by the “three-point property”: 102 VI. Applications of Quasiconformal Mappings FIGURE 1. P(z) = 27 +%/2, J is a quasicircle. there is an M > 0 such that if z1, 22, z3 € [and 22 is on the “shorter” arc between 2 and 23, then |z1 — 22| + |z2 — 23] S Lr [A2]. THEOREM 2.1 (Sullivan). Suppose that the Fatou set of a rational function R has exactly two components and that R is hyperbolic on the Julia set J. Then J is a quasicircle. Proof. The proof is very similar to that of Theorem 1.1 (see also the proof of Theorem VII.5.1). Suppose the two components are D, and D2. By considering R? if necessary, we may assume each component is completely invariant. They are simply connected, by Theorem IV.1.1. In D; construct a curve 72 with no critical points in the annular domain between 72 and J. Let y1 = R71(749). Since R is hyperbolic we can assume that 2 lies between 7, and 7. The domain between 7 and J corresponds to Uj, the domain between 72 and J to Us. As before, we construct g; in D; and the analogous Beltrami coefficient on D,. We do the same in D2. We set pp = 0 elsewhere and solve the Beltrami equation for y. Again Ro = yogo |! is analytic, and deg Ry = deg R = d. Evidently Ro has critical points of order d— 1 at two points, since g; and gp do, and hence is equivalent to 24, Thus J is the image of the circle {|z| = 1} under composed with a Mobius transformation. 0 V1.3. Herman Rings 103 EXAMPLE. Suppose P is a quadratic polynomial with an attracting fixed point. By Theorem III.2.2 the immediate basin of attraction A* of the fixed point contains the critical point, so P maps A* two- to-one onto itself, and A* is completely invariant, as is A(oo). By Theorem V.1.3 these are the only components of F. Furthermore, P is hyperbolic on J, by Theorem V.2.1. Thus Theorem 2.1 applies, and 7 is a quasicircle. An example is shown in Figure 1. This picture was produced by iterating P and coloring black those points z for which |P"(z)| < 4 for n = 1,2,..., 100. 3. Herman Rings The first examples of Herman rings were produced by M. Herman (1984), who based their existence on Arnold’s theorem (Theorem IL.7.2). For a > 1 real, consider R(z) = e?72?(z — a)/(1 — a2). It is an orientation-preserving homeomorphism of the unit circle, and it is uniformly close to the rotation z — 27° > for large a. Now the rotation number of R moves continuously with a and @, it is increasing in 0, and it converges uniformly to 6 (mod 1) asa — +00. For a fixed Diophantine number a, we can therefore find 6 = 6(a) so that R has rotation number a. Moreover 6(a@) — a as a — oo. Hence for large a we have R(z) = e?™%2(z) where ® is arbitrarily close to 1. The hypotheses of Theorem II.7.2 are met, and R is conjugate to a rotation in some annulus containing the unit circle. The circle is contained in a fixed component U of the Fatou set, and U cannot be attracting or parabolic (since the circle is invariant) nor a Siegel disk (since 0 and oo are attracting). Thus U must be a Herman ring. Actually, from the global version of Arnold’s theorem we obtain a Herman ring for any a > 1 and any 6 such that R has a Diophantine rotation number. If for instance a = 4, the finite critical points of R are the roots of z(z? — (19/8)z + 1), which are 7) = 0, r2 0.547, and r3 1.828. The first is superattracting, so by Theorem V.1.1, OU must be contained in the closure of the iterates of the second and third. Figure 2 shows 20,000 iterates of each critical point for a randomly chosen value 6 = 0.5431245. We give a second proof of the existence of Herman rings, due to 104 VI. Applications of Quasiconformal Mappings foo. ; oY N J \ f o NU FIGURE 2. Iterates of critical points delineate a Herman ring. M. Shishikura (1987), which transforms a Siegel disk into a Herman ring with the same rotation number. THEOREM 3.1. Suppose a rational function R has a Siegel disk, on which R is conjugate to multiplication by = 28 Then there ex- ists another rational function with a Herman ring, on which it is conjugate to multiplication by the same multiplier X. Proof. Let U denote the Siegel disk of R, say with center 0, and let R(z) = R(@). Then R has a Siegel disk U with multiplier e277 at 0. Thus U and U are conformally equivalent to the unit disk, and R and R act there by multiplication by e279 and e278, respectively. For 0 < 7,7 < 1 let 7 C U correspond to a circle {\¢| = r} via the conjugating map, and let 7 be an analogous curve in U correspond- ing to 7. Define an orientation-reversing homeomorphism w from an annular neighborhood of 7 to an annular neighborhood of ¥ so that ~ in z-coordinates corresponds to the map ¢ > r7/¢ in ¢-coordinates. Then 7(7) = 7, and ¥(R(z)) = R(ib(z)) for z € 7. Extend 7 to be a diffeomorphism from the domain inside y onto the domain outside 7. We may assume ~(0) = oo and that w is analytic in a neighborhood of 0. Define a smooth function g on C by setting g = R outside y and g = w-!o Row inside y. These definitions coincide near ¥, so that g is analytic outside a compact subset of U. Let E be the ellipse field which is circles outside 7 and V1.4. Counting Herman Rings 105 which is mapped to circles by 7 inside 7. Since R and R are analytic, this field is invariant under g. Let y solve the Beltrami equation and observe that Ro = pogoy | is rational (of degree 2d). Let A be the annulus that is the part of U outside y, and let W = (A). Since g maps A onto itself, Ro(W) = W, and W is part of a component of F\ (Ro). This component is either a Siegel disk or a Herman ring. The first is impossible. Indeed RG cannot be normal in either complementary component of W, since R” and R® are not normal outside U and U, respectively. Since the action of Ro on v(7) is conjugate to the action of R on 7, and the rotation number of R on 7 is 6, the rotation number of Ro on v7) is also 0, and Ro must be conjugate to multiplication by eee Herman ring. 0 1 4. Counting Herman Rings By means of quasiconformal surgery, Shishikura showed there are at most d — 2 Herman rings, and this bound is sharp. Here we are concerned only with showing that the number of Herman rings is finite. THEOREM 4.1. The number of cycles of Herman rings of a rational function of degree d is at most 4d + 2. Proof. The proof is by counting parameters. The number of inde- pendent quasiconformal deformations of the system to nonconjugate systems cannot exceed the number of parameters describing the con- jugacy classes. The line of argument is due to Sullivan (cf. [Do3]), and it is similar to that used in Theorem IV.1.3. Let U be a Herman ring, and let E be a compact invariant subset of U which is conformally equivalent to the annulus {1 < ¢ < R}. Consider the function v on E which in ¢-coordinates is given by C2/|¢|2. For 0 < t <1 the ellipse field corresponding to the Beltrami coefficient tv is invariant under rotation and has major axis parallel to the direction of rotation. One can argue geometrically that any solution of the Beltrami equation increases the modulus of the annu- lus. This can also be seen by noting that a solution in ¢-coordinates is given explicitly on the annulus by \opPe/A-H¢, Consider N such annuli Uj, ..., Uy from different cycles of Herman 106 VI. Applications of Quasiconformal Mappings rings. Define E; and v; as above, and for t € RN satisfying 0 < tj < 1,1< 7 0. Define h to be the conformal map of the inside of 72 onto the disk {|¢| < p™} sending wo to 0, and extend h smoothly to map the annular region between +; and 72 homeomorphically to the annulus {p™ < |¢| < p}, so that h(z)™ = h(R(z)) for z € 11. Define g(z) to be h-!(h(z)™) inside 7, and set g(z) = R(z) elsewhere in G. Thus g combines the dynamics of R outside 7 and of h-}oh™ ~ ¢™ inside 72, as in the proof of Theorem VI.2.1. VL.5. A Quasiconformal Surgical Procedure 107 Define an ellipse field to be circles on the part of U outside 71, and extend it to G to be g-invariant and to be circles outside UR~"(U). Let 7 be the solution of the corresponding Beltrami equation. Then Ro = Wogow! is a rational function, which has the same dynamics as g. Thus Rp is m-to-1 on Y(U) and has a superattracting fixed point as its only critical point in y(U), so Ro is conjugate to c™. 0 Note that the Julia set of Ro is the ~-image of the Julia set of R, and the dynamics of Ro outside y(U) are the same as those of R outside U. We are now in a position to extend Theorem V.4.3 to cover attracting fixed points. THEOREM 5.2. Suppose that each critical point of R on J is strictly preperiodic. If U is a simply connected component of F that has an attracting periodic point, then OU is locally connected. Proof. We may assume U has a fixed point and R is an m-to-1 branched cover of U over itself. Replacing R by the rational function Ro of the preceding theorem, we may also assume that R on U is conjugate to ¢™ on A. Then we define the curves 7 in U exactly as in the proof of Theorem V.4.1, so that R is an m-fold cover of Yn+1 over Yn. As in Theorem V.4.3, we have a bounded metric on the part of U near J for which R is expanding. Now the proof of Theorem V.4.3 goes through unaltered. 0 THEOREM 5.3. If there are two completely invariant components of the Fatou set, at least one of which has an attracting fized point, then the Julia set J is a simple closed Jordan curve. Proof. In this case, J is adherent to each component of F, so J is connected. Since F contains all the critical points of R, Theorem 5.2 applies, and J is locally connected. The Riemann map from A to either component of F extends continuously to & and is easily seen to be one-to-one on 0A. 0 EXAMPLE. For 0 < 6 < 1, the Julia set of z/(z? — bz + 1) is a simple closed Jordan curve (see the example in Section V.1). When d = 2, the hypothesis of the existence of an attracting fixed point in Theorem 5.3 is superfluous. In fact, if R is a rational func- 108 VI. Applications of Quasiconformal Mappings tion of degree d = 2, then R has at most one parabolic fixed point (parabolic fixed points are of multiplicity > 2). So if F has two com- pletely invariant parabolic components, then they share the same fixed point, with two attracting petals, the fixed point has multi- plicity three, and there can be no other fixed points. Now we claim that if R has degree d = 2 and a single fixed point in C (necessarily of multiplicity three), then R is conjugate to z/(z? +1), and conse- quently (Section V.1) the Julia set J of Ris acircle in the extended plane. To see this, we place the fixed point at z = 0 and its other preimage at 00, so we have conjugated R to the form z/(az?—bz+c). Since the fixed point is parabolic, c = 1. If we place a critical point at z = 1, we obtain a = 1. Since z/(z2 — bz +1) has a fixed point at z = b, we have also b = 0. EXAMPLE. The disguised Blaschke product (327 + 1)/ (2? +) is con- jugate to z/ (z? +1), as it has only the one parabolic fixed point at z=1. VII Local Geometry of the Fatou Set We focus on components of the Fatou set that are simply connected, and we study the geometry and dynamics near the boundary. 1. Invariant Spirals We know by Theorem III.3.1 that repelling periodic points are dense in the Julia set 7. Let zo be such a repelling periodic point, with mul- tiplier \ = (R™)'(zo). Locally R™(z) behaves like z — 29 > A(z — 20) near zg, and this seems to indicate that the Fatou set of R con- tains a logarithmic spiral with endpoint at zo. We can then conclude that 7 has a very complicated structure as soon as arg(R™)'(zo) is generically irrational. We may assume that zo is a repelling fixed point, and we assume also that zo is on the boundary of an invariant component U of the Fatou set F. We wish to construct curves in U terminating at zo. Our strategy is as follows. Let g be the branch of R~! fixing zo, and denote A. = {|z — 20| < €}. For e small, g is a contraction mapping of Ac, with multiplier 1/A at its attracting fixed point zp. Let wo € UNA be near 29. At issue is whether some iterate w, = g"(wo) lies in the same component of UM A, as wo. If so, we join wo to wn by an arc yo in UM Ag, and the union of the images of yo under g’”, j > 0, 110 VIL. Local Geometry of the Fatou Set forms a curve y in UNA. terminating at zo. The curve ¥ is invariant under g”, and R"(7) > 7. Suppose there are finitely many connected components Uj, ..., Um of UNA, that fill out a neighborhood of zo in U, that is, that include all points of UNA, for p > 0 sufficiently small. We may assume that each Uj; contains 29 on its boundary. Evidently each g(Uj) is either a subset of some U, or disjoint from U. If z € Us is near zo, then w = R(z) belongs to some U;, and g(U;) C Ug. In view of the finiteness assumption, g thus determines a permutation of the U;’s, and consequently some iterate g” leaves invariant each U;. Now the argument in the preceding paragraph shows that every wo € UNA, can be joined to zo by a curve y in UN A; satisfying R"(y) D7. With more effort, we establish the following. THEOREM 1.1. Let U be a simply connected and completely invariant component of the Fatou set F, and let zo € OU be a repelling fixed point of R. Then there exists n > 1 such that any wo € U sufficiently close to 29 can be joined to z by a curve y inU satisfying R"(y) D 7- Proof. We continue with the notation introduced above. The com- plete invariance of U guarantees that g maps UM A, into itself. Fix wo € UN Ae, and define w; = g? (wo): Fix a point b € U, b ¢ Ac, and consider Green’s function G(z, b) = G(z). In terms of the conformal map w(¢) = z of the open unit disk A onto U, (0) = b, we have G(z) = log |¥(¢)|. Now B= wp loRopisa d-sheeted covering of A over A, hence a Blaschke product of degree d. If p < 1 is such that B(¢) has no zeros for p < |¢| < 1, then, for c > 0 sufficiently small, log |¢| < clog|B(¢)| on the boundary of the annulus and hence on the entire annulus. In terms of Green’s function for U, this becomes G(z) > cG(R(z)), z€U, z near OU. Iterating this, we obtain G(w;) 2 cG(wo). We also choose cp > 0 so small that G(z,wo) > coG(z,6) on a small circle centered at b. By the maximum principle, the estimate persists for all z € U outside the circle, and we obtain G(w;, wo) > coc! G(wo)- (1.1) Now suppose the points wj all lie in different components of UNA¢. Note that |w; — 20| < C|A|~? for some C > 0. Let 7; be the geodesic VII.1. Invariant Spirals 111 curve from wo to w;, for which we have the estimate of Theorem 1.4.4. Fix N large. For C|A|~“ N, let 6;(r) be the distance to OU of the point where 7; meets the circle {|z—z| = r} for the last time. The +;’s must pass through different arcs of UM {|z— z0| = r}, or else we could join some w; to a wy by an arc in UM A,. Thus N+p YS 6;(r) < 2nr, g=N4+1 and by the Cauchy—Schwarz inequality, Nip N+p N+p 2 S 6;(r 2ar <(¥ 4m)( > am) <9" > ae g=N+1 j=. Wi © g=N41 Hence Nip 2 E I |az| >— 2 Ff & — >cNp*. ', 55(l2\) Cl|A-N The integral over one of the curves 7; must exceed c, Np. If we take p =N, choose such a 7;, and use the estimate of Theorem 1.4.4, we obtain for this j the estimate j=N41 G(w;, wo) < 3exp(—e1N*/2). (1.2) On the other hand, from (1.1) and j < 2N we have G(w;,wo) 2 coc?’ G(wo), and this contradicts (1.2) for N large. Hence for some N < j < j+n< 2N, wy; can be joined to wj+n by an arc in UM Ag. The union of the images of the arc under iterates of g” forms a curve in UNA, starting at w; and terminating at 20. The image 7¥ of this curve under R? starts at wo, terminates at 20, and satisfies R"(7y) D 7. It remains to show that n can be chosen independent of wo, and this part of the argument is topological. Replacing R by an iterate, we can assume there is already a path 6 in UM A, terminating at zg and satisfying g(3) C 8. We repeat the above discussion, and we want to show that w; and w),1 lie in the same component of UNA, for some j, since then we obtain 7 satisfying R(y) > 7. We assume that the points w;,, are in different components of UN A, for 0 < k 7- Proof. The proof of Theorem 1.1 carries through with minor modifi- cations. In this case, U is included in a cusp at Zo, and the discussion of the Fatou coordinate, applied to g, shows for ¢ small that g maps UNA. into itself. Moreover, |wz — 20| ~ 1/k1/™, where m is the number of petals at zo. This time 7; satisfies je x oS ogN log N +c3N +4, ng j(l2|) ~ 2m Jon-m where c) > 0. As before, the estimate of Theorem 1.4.4 leads to a contradiction for large N. Again we obtain a curve 7 in UNA; joining w; to wp for some k > j. Consider the Fatou coordinate function y for g, which conjugates g to translation by 1. The curve ¢(7) goes from y(w;) to p(w) +k-j, while y(g(7)) is the translate of p(y) by 1. If k—j 2 2, the curve must intersect its translate, and we find a curve from w; to wj+1 in Un A.. As before the various iterates generate a curve from wo to zo, this time satisfying R(y) Dy. 9 Theorems 1.1 and 1.2 were proved in [Ca2]. The proof meth- ods are related to those of the Ahlfors-Carleman—Denjoy theorem VII.2. Repelling Arms 113 (see [Beu], [Fu]). A-E. Eremenko and G.M. Levin (1989) have ob- tained invariant curves in U terminating at repelling fixed points, without the hypothesis of simple connectivity. They treat polynomi- als, with U = A(oo), and they apply the subharmonic version of the Ahlfors-Carleman—Denjoy theorem to the composition of the inverse of Koenigs’ coordinate function (an entire function of finite order; see Section II.3) and Green’s function. 2. Repelling Arms Let U be an invariant, simply connected component of the Fatou set F, and let z € OU be a fixed point of R(z). Let V be an open subset of U, and suppose there is a conformal map ¢ = y(z) of V into a horizontal strip S = {a < Im¢ < b} whose image includes a left half-strip {¢ € S: Re¢ < —C}, such that z — z as Re¢ — —oo with Im¢ fixed; z + J as Im¢ — a or Im¢ — 5; and y conjugates R to the translation ¢ > ¢ + 1, p(R(z)) =9(z) +1, 2€V, RizjeV. In this situation, we say that V is a repelling arm of F terminating at zg. The quotient space V/R is conformally equivalent to an annulus with modulus b — a, so that the height of the strip is uniquely de- termined. The coordinate y is unique, up to a translation. Applying Lemma IV.2.4 to R7!, we see that the fixed point zo is either a re- pelling fixed point or a parabolic fixed point with multiplier 1. If zo is parabolic, then V approaches zp through a gap between consecutive attracting petals. A periodic repelling arm terminating at a periodic point is defined in the obvious way. The topological argument at the end of the proof of Theorem 1.1 shows that periodic repelling arms terminating at the same point have the same period. Moreover, if the fixed point is parabolic, and if a cycle of periodic repelling arms terminates at the fixed point through the same cusp, then in fact the cycle consists of only one repelling arm. EXAMPLE. Let B(z) be a finite Blaschke product of degree d > 2 whose Julia set coincides with the entire unit circle 0A, and suppose zo € OA is a repelling fixed point of B. Let t be Koenigs’ coordinate function at 20, so that 7(zo) = 0, and r(B(z)) = A7(z) in a neigh- 114 VII. Local Geometry of the Fatou Set borhood of zo. Since arg B(z) is increasing around 0A, the multiplier ) is positive, A > 1. Since B leaves OA invariant, we can assume T maps A to the lower half-plane. Then (2) = a ican (principal branch) coordinatizes two repelling arms terminating at 20, one in A and the other in C\A, each mapped to a half-strip of height 7/ log A. THEOREM 2.1. Suppose U is an invariant component of F that is simply connected, let yp: A > U be the Riemann mapping, and set B=w!oRoy, a finite Blaschke product. Suppose Cy € OA is a fixed point for B(¢). Then w(¢) has a nontangential limit zo at Co, and z is a fixed point for R(z) that is either repelling or parabolic with multiplier 1. If Co is a repelling fixed point for B(¢), then for p > 0 sufficiently small, Y({AN |C — Go| < p}) is a repelling arm for R(z). Otherwise, Co is a parabolic fixed point for B(¢) with Julia set AA, and z is a parabolic fixed point for R(z) which has U as an attracting petal. Proof. Since B is proper, it is a finite Blaschke product. Assume Co = 1 is fixed by B. Let rn be any sequence of radii increasing to 1 such that u(rn) converges, say to 20 # co. Now the conformal self-map (¢ — Tn)/(1 — Tn¢) of A maps Tn to 0 and B(r,) to Brn) tm _ (Bln) = Dra WI = 1 1—rnB(rn) Tn{(B(rn) — 1)/(tn — 1)] + V whose modulus tends to Bi(1j)-1 | Bi(1) + | Thus the hyperbolic distance in A from rn to B(rn) is uniformly bounded, and the hyperbolic distance in U from W(rn) to p(B(rn)) = R(w(rn)) is also bounded. The comparison of the hyperbolic and euclidean metrics in Theorem 1.4.3 shows that |R((Tn)) — wW(rn)| > 0. In the limit we obtain R(zo) = 20. Since R has only a finite number of fixed points, 7(r) must accumulate at 20 as T increases to 1, and w(¢) has nontangential limit z at 1. If the fixed point Cp € OA is repelling, we compose w with the logarithm of Koenigs’ coordinate function, as in the example above, VII.2. Repelling Arms 115 to obtain a repelling arm at zo. This implies 2 is either repelling or parabolic with multiplier 1. Suppose finally that Co is not repelling. Then B"(¢) — Co on A, so that R"(z) > OU for z € U. By Lemma IV.2.4, R"(z) converges on U to a parabolic fixed point z; of R(z) with multiplier 1. By joining an appropriate ¢ € A to B(C) and iterating, we obtain a curve in A terminating at ¢o along which #(¢) has limit z). Lindelof’s theorem implies z1 = zo. Thus in this case zo is a parabolic fixed point and U is an attracting petal at zo. Consider two arcs 7, and y_ in U starting at zo and terminating at points 24,2. € J which lie near zp in the respective repelling cusps on each side of the attracting petal at zo. These curves divide U into three pieces, one containing the attracting direction, the other two Ux lying between 7 and the part of 7 in the repelling cusps. In view of our analysis of parabolic fixed points, we can choose 7+ and y— so that U, and U_ are invariant under g. Let ¢4 be the terminal points on OA of the curves #7! (74) as z > z4. The domains w1(Ux) are contiguous to the two arcs between (9 and ¢4, and these arcs must be on different sides of Co. The curve y~1(g"(y+)) in U4 terminates at ¢o and another point ¢,. Since g"(U;) decreases, Cn Moves monotonically towards Co, hence converges to a point ¢*. Since g"(z;) moves towards 2 in the repelling cusp, one sees that any nontangential limit of ~ on the arc between ¢* and ¢o must coincide with zo, hence ¢* = Co. Evidently B(¢n41) = Gn, so the direction along the arc from Co to ¢4 is repelling for B. Similarly, the direction from Co to ¢_ is repelling, so Co is a parabolic fixed point for B with two petals. 0 EXAMPLE. Consider R(z) = z? + 1/4. The Julia set J of R is the cauliflower set depicted in Figure 5 of Section VIII.1, which was shown in Section V.4 to be a simple closed Jordan curve. On the bounded component of the Fatou set F, R is conjugate to a Blaschke product on A of degree 2, with one parabolic fixed point on 0A corresponding to the parabolic fixed point of R at 1/2. According to the discussion in Section VI.5, R is in fact conjugate to B(¢) = (3¢? + 1)/(3 + ¢?). The unbounded component A(co) of F has a repelling arm at the parabolic fixed point 1/2, which issues from 1/2 along the real axis in the positive direction. The action of R on A(oo) is conjugate to the Blaschke product ¢?, which has a superattracting 116 VII. Local Geometry of the Fatou Set fixed point at 0 (corresponding to oo), and a repelling fixed point at ¢ = 1 (corresponding to the parabolic fixed point at z= 1 /2). The argument in Theorem 2.1 is reversible, as follows easily from Lindeldf’s theorem. A periodic repelling arm in an invariant compo- nent U arises from a unique repelling periodic point on 0A, and the period of the arm coincides with the period of the periodic point. In the case where OU is locally connected, we can say more. THEOREM 2.2. Suppose U is an invariant component of the Fatou set F that is simply connected, and suppose OU is locally connected. Let»: AU be the Riemann mapping, with associated Blaschke product B = plo Row. Suppose zo € OU is a repelling periodic point for R, or that zo is a parabolic periodic point and U is not an attracting petal for zo. Then wW}(z0) = {C1 Cw} ts finite. Each ¢; is a repelling periodic point for B and determines a periodic repelling arm in U that terminates at zo. These N periodic repelling arms fill out a neighborhood of zo in U. Proof. We assume 2p is a repelling fixed point (the parabolic case is almost the same, and there is also an analogous result if U is an attracting petal for zo). We use the notation of Section 1, and we take 6 > 0 sosmall that |b(¢)—¥(€)| < €/2 for ¢,€ € A satisfying |¢—€| < 6. fz =v(6) €UN Aga, then the image under of AN A(C, 6) is contained in UN Ag. Since only finitely many of these sets ANA(¢, 6) can fit disjointly into A, there are only finitely many components Uj, ..., Un of UNA, that contain zo in their boundary, and these fill out a neighborhood of zo in U. By the discussion preceding Theorem 1.1, there is in each Uj a curve 7; invariant under some iterate of g and terminating at 29. Then w71(4) terminates at a point ¢; € 0A, and if R"(7;) D 7, then ¢; is a fixed point of B”. By our hypothesis, ¢; cannot be a parabolic fixed point for B, so ¢; is a repelling periodic point. It determines a repelling arm terminating at 20, via Theorem 2.1, through which 7; approaches 29. The curve 7; can be chosen so that its initial arc passes through any two prescribed points of Uj. Consequently w7! has the same limit along the iterates of these two points, and for each w € Uj, w1(gk"(w)) tends to ¢; as k — oo. Let p > 0 be small, and let Vj = w(A(G,p) A). Then Vj is a repelling arm for some R”, which is contained in U; and which terminates at zo. We claim that V,..,Vn are the only repelling VIL.3. John Domains 117 arms in U terminating at zo. Indeed, if V is such a repelling arm, say for R”, then the iterates of w € V under g” eventually enter some U;, and since w-!(gk"(w)) tends to ¢;, the arm V is the same arm as V;. Suppose & € w~1(zo) is distinct from the ¢;’s. We may assume that p is less than the distance from to the V;’s, and that there is r > 0 small such that |7(¢) — zo| > r on the circular arc {|¢ — Go| = p} NA. Let W be the component of UN A, containing the image under ¢ of the end of the radial segment terminating at 0. By what has been established, with e replaced by r, W contains a repelling arm terminating at zo. However, this contradicts the fact that W is disjoint from the V;’s. We conclude that wh (20) = {G,..., Cw}, and it follows easily that the V,’s fill out a neighborhood of zp in U. 0 For a color illustration of such a configuration (in the Mandelbrot set), with 29 arms, see Map 38 on page 82 of [PeR]. 3. John Domains A domain D is called a John domain if there exists c > 0 such that for any zo € D, there is an arc y joining zo to some fixed reference point wo € D satisfying dist(z,0D) > clz — 20], z€7. (3.1) If 00 € OD, we use the spherical metric to measure the distance. If two smooth boundary arcs meet at a corner with a positive an- gle, the condition (3.1) is satisfied near the vertex. If however the two boundary arcs meet tangentially at an outward-pointing cusp, the condition (3.1) cannot be satisfied. Similarly, if D is a simply connected component of the Fatou set that has a repelling arm ter- minating at a parabolic point, then D is not a John domain. A simply connected John domain has a locally connected bound- ary. This can be seen easily from the proof of Carathéodory’s theorem given in Section 1.2. The crosscuts I’, constructed there have lengths tending to 0, and the condition (3.1) then shows that the diameters of the pieces of D outside the crosscuts tend to 0. This implies that the Riemann map from A to D extends continuously to 0A. The image of a John domain under a quasiconformal homeomor- phism of C is evidently a John domain. Thus the two complemen- 118 VII. Local Geometry of the Fatou Set tary components of a quasicircle are John domains. It can be shown, conversely, that if the two complementary components of a simple closed Jordan curve are John domains, then the curve is a quasicircle. Thus John domains can be regarded as “one-sided quasidisks.” For this and further background material on John domains, see [NaV]. Towards improving upon Theorem VI.2.1, we prove the following. THEOREM 3.1. Suppose R is subhyperbolic on the Julia set J. Then any simply connected component of the Fatou set F is a John do- main. We carry out the proof first in the hyperbolic case, where the ideas are clearer, and then we indicate how to adapt the proof to the subhyperbolic case. Proof (hyperbolic case). Let R be hyperbolic, and suppose U isa simply connected component of F. Each component of F is iterated eventually to an attracting cycle of components. Replacing R by an iterate, we may assume that R(U) contains an attracting fixed point. Since R is proper on U, R(U) is also simply connected. Since R is conformal on QU, it suffices to prove that R(U) is a John domain. Thus we may assume that U itself has an attracting fixed point. Replacing R by the rational function Ro of Theorem VI.5.1, which is also hyperbolic, we may further assume that R on U is conjugate to C™ on A. Let »: A — U conjugate ¢” to R(z). By Theorem VI.4.2 (or by the proof of Theorem V.4.3), y extends continuously to OA. We will show that the geodesics {a(re®) :0 0 so that all the inverse iterates R~* are almost affine on any disk of radius p centered in a neighborhood of 7, that is, so that the estimate used in the proof of Theorem V.2.2 holds. Set V = W({1—6 < |¢| < 1}), where 6 > 0 is so small that the affine mapping estimate holds on disks centered in V of radius p. Since ~ is uniformly continuous, we can assume that any two points of V on VII.3. John Domains 119 the same conformal ray are within distance p of each other. Set co = inf {dist(z, J) : 2 € V, R(z) ¢ V}- Then co < p. Suppose 2 lies on the ray between wo and 2. It suffices to check the estimate (3.1) for zo and z; in V. Let n > 1 satisfy R"(21) € V and R°+1(z,) ¢ V. Then the disk centered at R"(z1) of radius co is disjoint from J. By the affine mapping property, the image of this disk under the appropriate branch of R~” covers a disk centered at z of radius say half of co/|(R”)'(z1)|. Hence : i co dist(z1,7) >= 2 [Ral (3.2) On the other hand, since R"(zo) and R"(z1) lie on the same ray, they are within distance p of each other, and the affine mapping es- timate yields |29 — z1| < 2p/|(R")'(z1)|- Combined with the estimate (3.2) above, this gives (3.1), and the theorem is proved when R is hyperbolic. 0 To carry over the proof to the subhyperbolic case, we require a version of the affine mapping estimate. For this, it is natural to work with an admissible metric, and we use the specific metric 7 con- structed in the proof of Theorem V.3.1, as modified in the ae after the proof. Thus o(z) is a constant multiple of |z — aj|~ 5) near each of its singularities a; , where 0 < 3; <1, and elsewhere a(z) is smooth. Let D, denote differentiation with respect to the a-metric, defined by (De R)(2) = Thus R dilates infinitesimal distances at z in the o-metric by a factor of |(D,R)(z)|. The chain rule is valid, n-1 (Do R")(2) = [] (Do R)(RN(2)), k=0 as is the inverse function rule, (D,R~!)(R(z)) = 1/(DoR)(z). From the precise form of o at its singularities, one checks that |D,R| is smooth in a neighborhood of J. The affine mapping estimate is now as follows. 120 VII. Local Geometry of the Fatou Set LEMMA 3.2. Suppose R is subhyperbolic on J. Let o be the admissible metric introduced above. Then there is a neighborhood V of J such that (DRM) = g(a wevn (DeR=*y(ay ~ bt )O(1), wevi,nei. Proof. Here we can specify R-"(z) to be an arbitrary branch, and then the determination of R~"(w) is that obtained by continuing R-"(z) along the shortest path between z and w in the o-metric (irrespective of singularities of R~”). Since R is strictly expanding with respect to a, the inverse branches satisfy |(D-R7!)(z)| <¢<1 in a neighborhood of J. The neighborhood can be chosen to be invariant under the inverse branches. By integrating, we obtain dg(R7*(z), R-*(w)) < chde(z,w) whenever the shortest path between z and w lies in the neighborhood of J and in particular whenever z is near w. Thus _ (DoR)(R*)(2)) | [Pome - (DeR)(R*)(w)) (Do R)(R-*)(w)) (DoR)(R*)(w)) < O(R*\(2)- (RY) < Codg(R*(z), R-*(w)) < Coc*de(z, v). 1 Using the chain rule as in the proof of Theorem V.2.3, we obtain (Dek Mw) = | cdg (z,w (Dene) ~ LLL + Pdole, tn, which yields the desired estimate. 0 Proof of Theorem 3.1 (subhyperbolic case). We aim to show that U satisfies the John condition (3.1), with distances measured in the g-metric rather than the euclidean metric. From this, it is straight- forward to conclude that U is a John domain, and the details are left as an exercise. (Note that a problem arises in comparing the metrics only at singularities, so one can focus on o(z) = |z|~8 inside the unit disk.) Since |D,(R)| is smooth, bounded above and away from zero near J, the o-geometry of U near the boundary is the same as that of VII.3. John Domains 121 R(U), so we may proceed as before and assume that R(z) on U is conjugate to ¢™ on A. Again the conjugation extends continuously to OU, hence is uniformly continuous with respect to the o-metric. Choose p > 0 as before, with R-* almost affine on o-disks of radius p, and choose V as before. Define co in terms of the o-metric, so that if z € V but R(z) ¢ V, then the o-distance from z to J is at least co. For z; on the ray between wo and zo, we obtain this time an estimate of the form co 1 21D Vayl °°) diste(z1, 7) > Since R"(zp) and R"(z:) lie on the same ray, they are contained in a o-disk of radius p, as is the ray connecting them. We may assume that this disk also contains the shortest path between the points. If there are no singular points a; in this disk, the analytic branch of R-™ sending R"(z1) to z also sends R"(z9) to zo. This yields an estimate of the form _ dg(20521) < TE RMAD| (3.4) If there is a singular point a; in this disk, we connect R"(z1) to R"(zo) by a path in the disk of o-length at most 3p, such that the continuation of the branch of R-" mapping R"(z1) to 2; along the path sends R"(zo) to zo. By integrating along this path, we obtain again an estimate of the form (3.4). Now (3.3) and (3.4) show that U is a John domain in the o-metric, as required. 0 Suppose P(z) is a subhyperbolic polynomial with a connected Ju- lia set 7. We claim that the boundary of each bounded component U of the Fatou set F is a quasicircle. Indeed, OU is locally connected, so the Riemann map from A to U extends continuously to 0A. Since QU is included in the boundary of a single component A(oo) of its complement, the Riemann map is one-to-one on 0A, and OU is a simple closed Jordan curve. Now A(oo) is also a simply connected John domain. From the condition (3.1) for A(oo), one deduces easily that (3.1) holds for the larger domain C\U. If we invoke now the characterization of quasicircles cited earlier (but not proved), we de- duce that OU is a quasicircle. In the case of hyperbolic polynomials, this result is subsumed by a theorem of M.V. Yakobson (1984). 122 VII. Local Geometry of the Fatou Set EXAMPLE. Each component of the body of Douady’s rabbit (Figure 4 in Section VIII.1) is bounded by a quasicircle. Theorem 3.1 and the application come from [CaJ]. Under certain hypotheses, Theorem 3.1 remains valid for parabolic components of F (see [CaJY]). Herman has proved (cf. [Do3]) that if a Siegel disk associated with a quadratic polynomial has the critical point on its boundary, and if the multiplier has the form ) = e279 where @ satisfies the Diophantine condition (6.2) of Section II.6 for some exponent j4, then the boundary of the Siegel disk is a quasicircle if and only if the Diophantine condition is satisfied with p = 2. Vill Quadratic Polynomials We are interested in the dependence of a dynamical system on pa- rameters, and we focus on the maps z — z? +c. This leads to the Mandelbrot set M in parameter space, which has a universal char- acter in the sense that similar-looking sets show up when one stud- ies very general parameter dependence. One striking feature of M is that shapes of certain of the Julia sets J in dynamic space (z- space) are reflected in the shape of M near the corresponding points in parameter space (c-space). 1. The Mandelbrot Set A quadratic polynomial P(z) can be conjugated by z' = az to a monic polynomial z* + az + 3. This can be further conjugated by a translation z’ = z+ b to move any given point to 0. If we move one of the fixed points to 0 we have conjugated P to the form Az + 2, where is the multiplier of the fixed point. This does not determine the conjugacy class uniquely, as we can place the second fixed point at 0. If we move the critical point to 0 we have conjugated P to the form P,(z) = 27 +6, 124 VIII. Quadratic Polynomials and now it is easy to check that different c’s correspond to different conjugacy classes. Thus we can regard the c-plane as representing conjugacy classes of quadratic polynomials (that is, it is a “moduli space”). We are interested in how the dynamic behavior of P. de- pends on the parameter c. The following dichotomy is an immediate consequence of Theorems III.4.1 and III.4.2. THEOREM 1.1. If P®(0) — 00, then the Julia set J. is totally discon- nected. Otherwise P®(0) is bounded, and the Julia set is connected. The set of parameter values c such that P®(0) is bounded is of special interest. It is called the Mandelbrot set and denoted by M. Thus c € M if and only if 0 does not belong to the basin of attraction of the superattracting fixed point at oo. THEOREM 1.2. The Mandelbrot set M is a closed simply connected subset of the disk {|c| < 2}, which meets the real azis in the inter- val [—2,1/4]. Moreover, M consists of precisely those c such that |P"(0)| <2 for alln > 1. Proof. If |c| > 2, one shows by induction that [PR(O)| Z lel(le|— 1", MEL so |P"(0)| + co and c ¢ M. Thus |ce| < 2 force M. Suppose |P”(0)| = 2+ 6 > 2 for some m > 1. If |c| = |Pe(0)| > 2, then c ¢ M. If |c| < 2, then |Pz"*1(0)| > (2+ 6)? -2>2+46. Proceeding by induction, we obtain |Pmr+k(Q)| > 2+ 4*§ — oo, and again c ¢ M. This proves the final statement of the theorem, from which it follows that M is closed. By the maximum principle, C\M has no bounded components, so C\M is connected, and M is simply connected. If c is real, then P(x) — x has no real roots if c > 1/4, one root at 1/2 if c= 1/4, and two real roots if ¢ < 1/4. Ife > 1/4, P2(0) is increasing and must go to infinity, since any finite limit point would satisfy P(x) = 2. Ifc < 1/4, leta = (1+ V1 — 4c) /2 be the larger root of P,(x) — x. If additionally c 2 —2, one checks that a > |c| = |P.(0)|. Then |P2(0)| < @ implies |Pn+1(0)| = [P2(0)? +e] < atc= a, and the sequence is bounded. Thus MnR = [-2,1/4]. 0 VIII.1. The Mandelbrot Set 125 FIGURE 1. The Mandelbrot set M. The theorem suggests a simple algorithm for computing M. If |P"(0)| < 4 for 1 < n < 1000, color the pixel c black, otherwise color it white. See Figure 1. We review the various possibilities for the Julia set % of P.. By Douady’s Theorem VI.1.2, there is at most one periodic cycle of bounded components of F,. By Sullivan’s Theorem IV.1.3, every bounded component of F; is eventually iterated into the cycle. Since Herman rings do not arise for polynomials, the classification theorem of Section IV.2 gives the following four possibilities for c € M. 1. There is an attracting cycle for P.. Either there is an attracting fixed point, in which case there is only one bounded component of F,. Or the cycle has length two or more, in which case there are infinitely many bounded components of F,. By Theorem V.4.1, J is locally connected. 2. There is a parabolic cycle for P.. Either there is a parabolic fixed point with multiplier 1, in which case there is only one bounded component of F,. Or the cycle of parabolic compo- nents of F, has length two or more, in which case there are infinitely many bounded components of F.. The former case 126 VIII. Quadratic Polynomials occurs only for one value of c, namely, c = 1 /4 (see Figure 5 to follow). By Theorem V.4.3, Jc is again locally connected. 3. There is a cycle of Siegel disks. This case has been discussed in Section VI.1. 4. There are no bounded components of F,. This occurs for in- stance for c = —2, where J = [—2,2], or for c = i, as in the figure in Section V.4. It may happen (Theorem V.4.4) that Jc is not locally connected. THEOREM 1.3. For each 2, |A| <1, there is @ unique c= c(A) such that P, has a fixed point with multiplier d. The values c for which Po has an attracting fixed point form a cardioid CCM, anddC Cc OM. If céC, then Jc is a quasicircle. Proof. The fixed points of P, are at 2 = (1+ V1 —4c)/2, and the multiplier at z. is A(c) = 2zc. Since ¢ = A/2 — 2/4, the condition |X(c)| < 1 corresponds in the c-plane to the cardioid C= {d/2-7/4: [Al < 1. It is a subset of M, called the main cardioid of M. Since the func- tion \/2 — \?/4 is one-to-one on the closed unit disk, we have the uniqueness assertion of the theorem. According to the discussion at the end of Section VI.2, the Julia set J is a quasicircle for c € C. Let W be the component of the interior of M containing C’. By Theorem 1.2, the polynomials fn(c) = P(0) are uniformly bounded on W. They converge to the attracting fixed point z, on C, hence on account of analyticity to the fixed point z- on all of W. However, if c ¢ C, then 2; is repelling, and we cannot have P?(0) > Ze unless P?(0) = 2c for n large. Since this can occur only on a countable set, we conclude that W = C. 0 The figure in Section VI.2 corresponds to c = 1/2, which is inside the main cardioid. Another example is given in Figure 6 to follow. The multiplier function A(c) = 1— V1 — 4c maps the closure of the cardioid homeomorphically onto the closed unit disk. The boundary OC consists of precisely the parameter values c for which P, has a neutral fixed point. These are favorite choices for generating com- puter pictures. The Siegel disks appearing in Section V.1 correspond VIIL.1. The Mandelbrot Set 127 FIGURE 2. Attracting flowers. to a value c on the boundary of the main cardioid with multiplier A =e? where 6 = (\/5 — 1)/2 is the golden mean, as discussed in Section V.1. For a multicolored picture of the Siegel disk correspond- ing to this c-value, see Map 25 on page 77 of [PeR]. The filled-in Julia sets corresponding to parabolic fixed points with multipliers X = exp(27i/3) and A = exp(27i/5) are depicted in Figure 2. Now we consider the more general case of attracting cycles. THEOREM 1.4. Suppose there is an attracting cycle of length m for P,. Then a belongs to the interior of M. If W is the component of the interior of M containing a, then P, has an attracting cycle {21(c),--) 2m(c)} of length m for allc € W, where each z;(c) depends analytically on c. Proof. Let z:(a) be an attracting periodic point of period m for a. Applying the implicit function theorem to Q(z,c) = P!"(z) — z, we obtain an attracting periodic point 2;(c) for P. of period m, which depends analytically on c in a neighborhood of a. In particular, a belongs to the interior of M, say to the component W. The sequence f;(c) = P3™(0) is bounded hence normal on W, and it converges at a to some point in the cycle of z1(a), say to z;(a) itself. Since z1(c) is attracting, f;(c) converges to 2;(c) for c near a. By normality f;(c) converges on W to some analytic function g(c), which satisfies 128 VIII. Quadratic Polynomials Q(g(c),¢c) = 0 near a hence on W. Now g(c) can be a repelling periodic point for fixed c ¢ W only if Pi™(0) actually coincides with g(c) for j large. This occurs on an at most countable set. Since the multiplier A(c) of the cycle is analytic, we conclude that |A| < 1 on W, that is, the cycle of g(c) is attracting for all c € W. Further, the period of g(c) is exactly m for all c € W (as in Lemma III.2.5). The functions P?(g(c)), 0 < i < m, give analytic selections of the points of the cycle. 0 The set of c for which P, has an attracting cycle was studied by R. Brooks and J. Matelski in 1978. The components of the inte- rior of M associated with attracting cycles are called the hyperbolic components of the interior of M since, by Theorem V.2.2, P. is hy- perbolic precisely when c ¢ M or P has an attracting cycle. It is not known whether the hyperbolic components fill out the interior of M, but we will soon see that they are dense in M. Brooks and Matelski arrived at the work of Fatou and Julia from a problem on discrete subgroups of PSL(2,C), and their rough computer picture was the first picture of the Mandelbrot set. Working about the same time and completely independently, B. Mandelbrot obtained progres- sively sharper pictures of the Mandelbrot set, which called attention to the intricate nature of the set. One of the hyperbolic components is the main cardioid C, corre- sponding to cycles of length one. For cycles of length two, we have Q(z,0c) =(2+e2te-z=(P- zt QP tzt+eF)). Discarding fixed points, we obtain the equation 27 +z2+c+1=0 for periodic points of period two. The multiplier is given by (P2)\\(z) = 422 +4ze = 4c+4 modulo (2+z+e+1). Thus there is an attracting two-cycle if and only if \4c+ 4| <1, and we obtain a single hyperbolic component which is a disk centered at —1 of radius 1/4, tangent to the main cardioid. Note again that the multiplier \(c) maps the hyperbolic component conformally onto the open unit disk. We show in Section 2 that each hyperbolic component of the interior of M contains a unique point c such that P, has a superattracting cycle, and that the corresponding multiplier map c — X(c) maps the component conformally onto the unit disk. Of special interest are the parameter values c for which P, has a superattracting cycle. The critical point 0 must belong to the cycle, VIII.1. The Mandelbrot Set 129 FIGURE 3. Airplane set, c + —1.754878. so these are precisely the c’s for which 0 is periodic, that is, the solutions of P”(0) = 0. The first few polynomials are P2(0) = ¢, P?(0) = ct’, P30) = c+423 +4, PAO) = c+e%42c3 + 5ct 4 6c° + 68 + 4c7 +. 8. For n = 1, the equation reduces to c = 0, and we obtain the super- attracting fixed point at 0 for Po(z) = z?. For n = 2, the equation becomes c? + c = 0. This gives an additional solution c = —1, for which P_;(z) = z* — 1 has a superattracting cycle 0 + —1 — 0 of period two. See Figure 8. The superattracting cycles of period three correspond to the so- lutions of P3(0) = 0, excluding \ = 0. These are the roots of c? + 2c? +c+1 = 0, which are approximately —1.755 and —0.123+0.7491. The real root belongs to the largest component of the interior of M located on the “main antenna” towards the left of M. The filled-in Julia set for the real root is the “airplane set” pictured in Figure 3. The two complex roots belong to the largest components of the interior of M above and below the main cardioid. The filled-in Julia set corresponding to a complex root is pictured in Figure 4. This set is known as “Douady’s rabbit.” THEOREM 1.5. The values of c € M corresponding to superattracting cycles cluster on the entire boundary OM. In particular, the interior of M is dense in M. Proof. Let U be a disk that meets 0M, such that 0 ¢ U. Suppose U contains no c for which 0 is periodic. Consider a branch of /—c defined on U. We have P”(0) # /—ce, or else P?+1(0) = 0 and 0 is periodic. Thus f,(c) = P2(0)/\/—c omits the values 0,1,00 on 130 VIII. Quadratic Polynomials FIGURE 4. Douady’s rabbit, c ¥ —0.122561 + 0.744862. U, hence is a normal sequence on U. But since U meets 0M, it contains both points c with fn(c) bounded and with fn(c) > 00, so the sequence cannot be normal. 0 Let us see how the Julia set changes shape as c moves along the real axis. If we move c to the right of 1/4, it leaves the Mandelbrot set and the Julia set becomes totally disconnected. At the right of Figure 5 we have a picture of Je for c = 0.251. The black points are those for which |P”(0)| < 4 for n < 500. At the left of Figure 5 is the cauliflower set, corresponding to c = 1/4. In this case J is a simple closed Jordan curve (Section V.4), though it cannot be a quasicircle due to the cusps. We show in Figure 6 the Julia set Je for c = —3/5. We are back in the main cardioid, so J is a quasicircle, symmetric with respect to R. Moving back to the left, at the left edge of the main cardioid we arrive at the point c = —3/4 with multiplier —1, so that P, has a parabolic fixed point —1/2. There are two petals at —1/2, which cycle back and forth. The filled-in Julia set is pictured in Figure 7. When we continue to the left of —3/4, the fixed point bifurcates to an attracting cycle of length two, corresponding to the two petals. VIII.1. The Mandelbrot Set 131 ee oe oe FIGURE 5. c = 0.25 (cauliflower set), and c = 0.251 (totally disconnected). FIGURE 6. c = —0.6, quasicircle. 132 VIII. Quadratic Polynomials FIGURE 7. c = —0.75, parabolic cycle of period two. FIGURE 8. c = —1, superattracting cycle of period two. This process is called “budding,” and the point —3/4 is the “root” of the bud. There is budding at each point of OC with rational mul- tiplier. The parabolic fixed point with multiplier a primitive mth root of unity splits into an attracting cycle of length m, sprouting a “bud” which is a hyperbolic component of the interior of M tangent to OC. For c = —1, we have the superattracting cycle 0 > -1-0 pictured in Figure 8. The basic shape of the Julia set is preserved as we cross from c = —3/5 over c = —3/4 toc = —1. At c = —5/4 we have a parabolic cycle of petals of order 4, and there is further budding. Continuing to decrease c gives a sequence (CO Cleo of parameter values corresponding to parabolic cycles of order 2”. In the complementary intervals, P, has attracting cycles of order 2”. This behavior is known as the period doubling VIII.2. The Hyperbolic Components of M 133 of Feigenbaum, and cn — Coo © —1.401. In the interval [—2, coo], periods of many different orders occur. A point c € M is called a Misiurewicz point if 0 is strictly preperi- odic, that is, P"(0) = P*(0) for some n > k > 0, but P”(0) 4 0 for all n > 1. By Theorem V.2.3, these are exactly the points where P, is subhyperbolic without being hyperbolic on Z%. By Theorem V.4.2, the Julia set JZ is a dendrite, and, by Theorem VII.3.1, C\s isa John domain. Since P, is expanding with respect to some metric, there are no neutral cycles, and the cycle of P*(0) is repelling. The Misiurewicz points are located in a complicated pattern in the antenna area of M. Both c = —2 and c = i are Misiurewicz points. We have seen that the Julia set of z? — 2 is [-2,2]. The Julia set of z* +1 is pictured in Section V.4. Since the Misiurewicz points are the zeros of a sequence of polynomials, they are countable. The proof of Theorem 1.5 can be modified to show that they accumulate on all of OM. We will see this in another way in Sections 5 and 6, where we show that these points belong to 0M and are terminal points of arcs in the complement of M. 2. The Hyperbolic Components of M We return to the polynomial Q(z,c) = P?"(z) — z in z and c. The zero set V of Q is a one-dimensional complex analytic variety in C?. It consists of all (z,c) such that z is a periodic point of P, whose period divides m. Note that Q is a monic polynomial in both z and in c, so that if (z,c) € V and one of |z| or |c| tends to infinity, then so does the other. The projection of V onto the c-plane is a branched covering map with 2” sheets. Fix one of the hyperbolic components W of the interior of M that corresponds to attracting cycles of length m, and let z(c) be the analytically varying periodic point of P, from Theorem 1.4. Let Ww be the subset of V of points (z1(c),c) for c € W, and let Vo be the irreducible branch of V containing W. We can view Vo as a Riemann surface spread over the c-plane, on which (P2")/(z) is analytic. Let V be the subset of Vo where |(P”)'(z)| < 1. Then OV consists of piecewise analytic curves on which |(P2")/(z)| = 1. As c € W tends to a € OW, we have (zi(c),c) > OV, or else we could continue 21(c) analytically to a neighborhood of a to give attracting periodic points, contradicting a € 0M. Thus W is a connected component of 134 VIII. Quadratic Polynomials V. The projections of the analytic arcs in OW are piecewise analytic. (At worst they might develop cusps where (P2”)'(z) = 1.) Thus OW is piecewise analytic. Since W U OW has connected complement, OW has no multiple points, and OW must be a simple closed Jordan curve. Thus QW is also a simple closed Jordan curve, and the branch z21(c) extends continuously from W to W U OW (including to the cusps). Let p(c) denote the multiplier of the attracting cycle for c € W. Since p(c) = (P™)'(aa(c)), plo) extends continuously to OW, where it has unit modulus. Consequently, the extended multiplier function p(c) is a continuous finite-to-one map of W onto the closed unit disk A, which maps W analytically onto A. Our main goal in this section is to prove the theorem of Douady-Hubbard-Sullivan (see [Do1]) that this covering is simple, so that the multiplier function p(c) parametrizes W. THEOREM 2.1. If W is a hyperbolic component of the interior of M, then the multiplier p(c) of the attracting cycle of Pc, c © W, maps W conformally onto the open unit disk A. It extends continuously to @W and maps W homeomorphically onto the closed disk A. Proof. Fix a € W. Let Ui, -.. Um-—1, Um = Up be the cycle of compo- nents of F containing the attracting cycle. We assume Up contains the critical point 0, so P, is two-to-one from Up onto U;, and P, maps each other U; conformally onto Uj41. Let g map Up conformally to A, with the fixed point of P™ going to 0. Then yo Prog lisa Blaschke product of degree two, mapping 0 to 0. After normalization of y, the Blaschke product has the form moon} = _.¢+% (po Pi" ov (0) = BW) = Se lel <1, for some \ € A. Since is the multiplier of the fixed point at 0 for By(¢), and this is invariant, we have \ = p(a). The idea now is to reverse this procedure. For |A| < 1—e, we combine the dynamics of P,(z) outside the U;’s with By inside, to obtain c() so that P.(,) has multiplier p(c(A)) = A, and so that e(A) depends continuously (!) on A. Since c(A) must have a branch point at a critical value of p, the continuity of c(A) shows that p(c) has no critical points, and hence p is one-to-one. The construction of c() is done by quasiconformal surgery as in the proof of Theorem VII-4.1, VIII.2. The Hyperbolic Components of M 135 except that here the Beltrami equation is solved with a continuous parameter \. Attention must also be paid to the complication that there is an invariant cycle rather than a fixed component. So fix ¢ > 0 small, let 1 By'(AA,) so that Boy, = yo P”, choosing the branch so that y) is continuous in A. Extend y, smoothly to a diffeomorphism from Ag to By! (A;), to vary smoothly also with \, so that y, = y on y 1(A,). Define A; C U; for 1 < 7 < mso that An = Ay and P, maps each A; conformally onto Aj41, 1 < 7 < m. Define gy(z) = Pa(z) for z ¢ UA;, and define g,(z) on UA; so that it maps each A; into Aj+1, and so that g¥"(z) = y1(By(p,(z))) on Ao. Thus gy(z) = Pa(z) on Ay U-+++UAm—1, and gy(z) = Pr™+}(g%(z)) on Ap. Consider the ellipse field that is circles outside of UA; and that is invariant under gy. Under iteration by gy, points hit the annular region where gy is not conformal at most once, so the ellipse field is distorted at most once on any orbit, and consequently the Beltrami coefficients py are bounded strictly less than 1 in modulus. Moreover jz, moves continuously with \. Let ~, solve the Beltrami equation, with the usual normalization #(z) = z+ 0(1) at oo. Then fy = 4, 09 op! is analytic, and fy(z) = 2? + O(1) at 00, so fy(z) = 2? + e(A) = Pxy)(z). Since gy has an attracting cycle of length m, so does fy), and the multipliers are the same. Thus p(c(A)) = A. Finally the solutions {, depend continuously on the parameter A, so that c(A) depends continuously on A. By the initial remarks we have proved the theorem. 0 If P, has a neutral cycle of length m, then a € OW for some hyperbolic component W associated with attracting cycles of length m. Indeed, let zo be a point in the cycle, and let Vo be an irreducible branch of V containing (zo,a). Then (P2")'(z) cannot be constant on Vo, or else since Vo projects onto C there would be a neutral cycle for all P., an absurdity. Thus (zo, a) is in the boundary of some component W of V (defined earlier). The projection W of W into the c-plane is a hyperbolic component of the interior of M, and a € OW. If W is associated with attracting cycles of length m, then every 136 VII. Quadratic Polynomials c € OW with p(c) #1 has a neutral cycle of length m, which is the analytic continuation of the attracting cycles of W. If p(c) = 1, then P, has a parabolic cycle whose length divides m. If the length of the cycle of P, is a proper divisor of m, then c lies on the boundary of another hyperbolic component, from which W is obtained by “bud- ding.” The point on OW where p(c) = 1 is called the root of W, and the point of W where p(c) = 0 is called the center of W. EXAMPLE. There are three superattracting cycles of period three, hence three hyperbolic components corresponding to attracting three- cycles. One of these is the largest bulb on the top of the main car- dioid. At the root of the component, the cycle of period three coa- lesces to a parabolic fixed point with multiplier 27/3 depicted in Figure 2 of Section 1. The second hyperbolic component is located in the symmetric position below the main cardioid. The third hyper- bolic component is the largest component centered on the real axis between —2 and the Feigenbaum point. For a color close-up of this component of the Mandelbrot set, see Map 32 on p. 89 of [PeR]. The root of the component is a cusp, and it corresponds to a parabolic cycle of length three and multiplier 1. 3. Green’s Function of % For the analytic study of P., the conjugation map ¢-(z) at co is the natural tool, pe(Pe(z)) = e(z)*. It is uniquely determined and has the form yo(z) = z + 0(1) at co. Recall from the superattracting case of Chapter II that ge(2) = lim, P22)? ”, n—00 which can be written et= =I (eae) Il (14 mor) n=1 n=0 The functional equation for ge allows us to extend log |y¢(z)| har- monically to A,(co), where it coincides with Green’s function G-(z) for A;(oo) with pole at oo. From the discussion in Section [1.4 we have the following. VIII.3. Green’s Function of Je 137 THEOREM 3.1. The logarithmic capacity of the Julia set J. is equal to 1 for all c. In this case the functional equation for Green’s function is Ge(Pe(z)) =2Ge(z), 2 € Ac(0o). By the discussion in Section III.4, there are two cases that occur. If c € M, there are no critical points of P. (or G-) in A-(oo), and the equation y-(z) = V¢-(P-(z)) allows us to extend ,(z) to a conformal map of all of A-(co) onto {|¢| > 1}. On the other hand, if c ¢ M, then 0 € A,(co), and y,(z) extends analytically to the exterior {G_- > G_(0)} of the level curve of Green’s function passing through the critical point 0, mapping it conformally onto {|¢| > eFe(0)}. The level curve forms a figure-eight with cusp at 0, as in Section III.4, and y,(z) approaches different values as z approaches 0 from different sides. Recall that Green’s lines are the orthogonal lines to the level curves of Green’s function. If 0 € M, these are just the curves in A,(co) mapped by ¢-,(z) to rays in conjugation space. Even if c ¢ M, they are well-defined, and they cluster at the boundary of A,(0oo), except for at most countably many that meet a critical point of G,. We define R(6,K,), called an external ray of K,, to be the Green’s line that corresponds to the ray Re? in conjugation space, that is, the Green’s line with initial segment y7!(Re?™!*), eG < R < oo. Thus Green’s lines are parametrized by a parameter @ from 0 to 1 (mod 1), where as usual we measure angles in turns rather than radians. The action of P. on Green’s lines corresponds to doubling the angle in conjugation space, always reduced mod 1, zER(OK,) => P,(z) € R(20,Ke). The dynamics of P, on dynamic space are closely related to the dynamics of the doubling transformation in @-space. If the ray R(0, XK.) terminates at a point 29 € K., we say that 6 is an external angle of K, at zo. Ifc € M and P. has an attracting or parabolic cycle, then all external rays terminate, and all points of % have at least one external angle. If c ¢ M, then K, = J is totally disconnected, so R(0, J.) terminates at a point of J. as soon as it is defined. 138 VIII. Quadratic Polynomials We extend the definition of Green’s function G for a domain D by declaring it to be zero in the complement of D. With this convention in mind, we show that G, is Holder continuous. THEOREM 3.2. For every A > 0, there exists a = a(A) > 0 so that Green’s function G- is uniformly a-Hélder continuous for |c| < A. Proof. Assume A > 10. Let 2 € C\K- and let 6(z) = dist(z, Ke). Let zo € K- be the point closest to z, and let S be the straight line segment from zo to z. Take N = N(z) to satisfy |P"(w)| < A for all w € S$ and all n < N, while |PN(z1)| > A for some 21 € S. Using P!(z) = 2z and the chain rule, we see that \(P)/(w)| < (24) for all n < N and w € S. Also, |P"(zo)| < 2VA for all n, or else the iterates would escape to oo. The mean value theorem then implies [PN (2)| < 2VA + 2% ANO(A1). But |P%(z;)| > A, hence 2N AN6(z) > 1, and 2N AN6(z) > 1. For a = log 2/(log 2 + log A) we then have 6(z)* > 27%, so that G(z) = G(PN(2))2-™ < Mbl2)", where M depends only on A. Consider two points 21, 22 and suppose 6(z1) > 6(22). We want to prove |G(z1) — G(22)| $ Cla — 22!" If |z1-22| > 36(21) this follows from the estimate above. If |z1-22| S 36(21); just use the fact that G(z) is a positive harmonic function in the disk A(z, 6(z1)), to conclude |G(a1) - Glza)|_ S CoM6(z1)°lz1 — 221/6(z1) S Cla — al". 9 The preceding proof extends easily to polynomials of higher de- gree, and the corresponding theorem is due to N. Sibony (seminar talk, 1981). The idea of the proof has been used by J.E. Fornaess and N. Sibony (1992) to obtain Hélder continuity of Green’s func- tion associated with certain polynomial mappings (complex Hénon mappings) on C2. It can be shown that Green’s function of a domain with uniformly perfect boundary is Holder continuous, so Theorem IIL3.3 leads to Hélder continuity also in the case of iteration of ra- tional functions. VIIL4. Green’s Function of M 139 The Julia sets J for |c| < A have Hausdorff dimension at least as large as the exponent a(A) of the preceding theorem. This follows from a theorem of Carleson (1963), that a compact subset of R’™ is a set. of removable singularities for a-Hélder continuous harmonic functions if and only if it has zero (m— 2+ a)-dimensional Hausdorff measure. If c € M, we can prove directly that G,(z) is Hélder continuous with exponent a = 1/2. In this case the map f(w) = 1/y71(1/w) is a normalized univalent function on A, belonging to S. The estimate of Theorem I.1.7 for f leads immediately to an estimate of the form G.(z) < C6(z)!/? for z in a neighborhood of K,, and the Hélder estimate follows as in the preceding proof. Note that the exponent 1/2 is sharp, as can be seen by taking c = —2 € M, for which 7, is the interval [—2, 2]. THEOREM 3.3. If c, > c, then the corresponding Green’s functions G_,, (2) converge uniformly on C to G.(z). Thus G.(z) is jointly con- tinuous in c and z. Proof. The uniform Holder estimates show that the sequence of func- tions G,,,(z) is equicontinuous on compact sets. Let H be a uniform limit on compacta of a subsequence. Then 4 is continuous, and H is harmonic on the set {H > 0}. By the maximum principle, there are no bounded components of the set {H > 0}, just one unbounded component. Since the conjugating functions y, depend analytically onc, G.,,(z) = log |¥¢,, (z)| converges uniformly to G.(z) = log |p-(z)| for |z| large. It follows that H = G, on A,(oo), hence everywhere. Since the limit H is unique, we have uniform convergence of the original sequence. 0 4. Green’s Function of M Suppose c € C\M. As long as G,(z) > G.(0) > 0, the function (c(z) is well-defined and analytic. This holds in particular for z = c, since G,(c) = 2G.(0) > G,(0) > 0. Hence (0) = gle) = Il (1+ ma) 140 VIII. Quadratic Polynomials is analytic, @ has a simple pole at co, and log |®(c)| = Ge(c) = 2G,(0). By Theorem 3.3, G.(c) = 0asc > M, and consequently Lr M. By the argument principle, ® assumes every value in C\A exactly once on C\M, and ® maps C\M conformally onto the exterior of the closed unit disk C\A. In particular, C\M is simply connected, and we obtain the theorem of Douady and Hub- bard (1982), proved independently by N. Sibony (cf. [DH1)), that M is connected. Since ®(c) = c+ O(1) as c > &, Green’s function of C\M, which is log |®(c)|, has the form log |e|+0(1) at oo. Hence Robin’s constant is 0, and M has capacity 1. We state these results formally. THEOREM 4.1. The Mandelbrot set M is connected and has logarith- mic capacity equal to 1. In analogy to the external rays of Ke, we define the external rays of M to be Green’s lines for C\M, that is, the inverse images of radial lines under ®. These are denoted by R(8,M), 6 between 0 and 1 (mod 1). If the ray terminates at c € M, we say that @ is an external angle of M at c. Observe that the external rays of M are related to the external rays of K, in the sense that cE R(O,M) = cE R(0, Ke). It is not known whether every external ray of M terminates, nor is it known whether the inverse 7) of ® extends continuously to map QA onto AM. This latter question is equivalent by Carathéodory’s theorem to M being locally connected. A great deal of work has gone into this question. It may be of interest to note that this would follow if there were some uniform estimate of escape times for c in terms of its distance to M. THEOREM 4.2. Let 6(c) = dist(c,M). If there exists a decreasing function F(x) > 1 on (0, 1] such that [ F(ax)dx < co 0 and [PN >5 for N > F(d(e)), then M is locally connected. VUI.4. Green’s Function of M 141 Proof. Let w : {|¢| > 1} + C\M be the inverse map to ®(c). If we can show that |y'(¢)| < H(|¢|) for some decreasing function H(r) on (1, 00) such that 2 | H(r)dr < 0, 1 then 7 is continuous up to the boundary, and we are done. 2G) N be the first integer larger than F(5(c o- Since G(P.(z)) = G(z), |PN(c)| > 5 implies that G.(c) > Co2-%. For ¢ = &(c) we a G.(c) = log |¢| ~ |¢| — 1, so Q-F(4(e)) < g-N+1 < Cr(|C| _ 1): The Koebe one-quarter theorem gives |'(¢)|(|¢|—1) < 46(c), so that 2-PUWOII-D/9) < EA (\¢| — 1). Setting t = C;(|¢| — 1) and solving s = 2~F**) for s, we find that 1 1 1 I) — R, and then compute the derivative of zw(c) by the chain rule, 2) = 0; 2j41 = Qazi, + 1. The approximator for 6(c) is |zy|"!|zn| log |zy|, and we can obtain good estimates for the er- ror factor. The resolution of the pictures in [Pe] obtained by the distance-estimator algorithm is striking. 5. External Rays with Rational Angles The external rays with @ rational play a special role. We aim to show that these rays hit M at well-defined points. If @ = p/q with p and q relatively prime, then the cases with q odd or even are very different. If q is odd, the ray terminates at a point c € M with the property that 0 belongs to a parabolic component of F(Pz). If q is even, it ends at a point c for which 0 is strictly preperiodic, that is, at a Misiurewicz point. The reason stems from a simple result on binary expansions. Let 6 € [0, 1] have binary expansion 6 = )"{* 0;2~’, where each 0; is 0 or 1. The function P, acts on external rays by doubling @ (and reducing modulo 1). This corresponds to shifting the sequence of bi- nary coefficients backwards (and chopping). Evidently the sequence {6;} is periodic if and only if 29 = 0 (mod 1) for some m > 1. The sequence is preperiodic if and only if 2"6 = 2*@ (mod 1) for some n>k2>0. VIII.5. External Rays with Rational Angles 143 THEOREM 5.1. Let 0 = 9° 0;2~) be as above, withO <6 < 1. Then 6 is rational if and only if the sequence {0;} is preperiodic. If 0 = p/q, where p and q are relatively prime integers, then q is odd if and only if {0;} is periodic, and q is even if and only if {0;} is strictly preperiodic. Proof. Suppose 6 = p/q = p/(2*r) is rational, with r odd, p and q relatively prime. Choose (by Euler’s generalization of Fermat’s little theorem) m so that 2” = 1 (mod r). Then alm+k)g — ok9 + ok ing = 26 + jp = 246, so 0 is preperiodic. If moreover q is odd, then k = 0, and @ is periodic. Conversely, if 6 is periodic, there are integers m,n such that 20 = 0+n. Then 0 = n/(2™—1) is rational, and the denominator must be odd. If 6 is preperiodic, then 2*6 is periodic, and @ is again rational, with even denominator if k > 1.0 THEOREM 5.2. If @ is rational, then R(0,M) terminates at a point aé OM. If 6 has even denominator, then a is a Misiurewicz point, whereas if 9 has odd denominator, then P, has a parabolic cycle. Proof. The proof, which is long, proceeds in outline as follows. If dist(c, J.) — 0 as c tends to a along R(6,M), it is rather easy to see that a is a Misiurewicz point and 6 has even denominator. If dist(c, J-) does not converge to 0 as c tends to a along R(6, M), it is again easy to see that a parabolic component of the Fatou set is created for P,. The main problem concerns the relation between the parabolic periodic points and the ray R(0,Kq), which eventually yields the result that 9 has odd denominator. So let @ be rational with period ¢, say 2°+*@ = 2*0, where £ > 1 and k > 0 are minimal. If c € R(8,.M), then both P*(c) and P£+*(c) belong to R(2*6, Jc). Moreover, these points have uniformly bounded hyperbolic distance from each other in C\.%. To see this, it suffices by Theorem 1.4.2 to bound the hyperbolic distance between them in the simply connected subdomain {G, > G_(0)} of A-(oo). This subdomain is mapped conformally to the unit disk by e%«)/y.(z), and the points are mapped to points with radii r+! and ré+*+! on the same radial segment, where r = e~@¢), The hyperbolic distance between these points in the unit disk is bounded, independent of r (for fixed @). 144 VIII. Quadratic Polynomials Fix a cluster point a € 0M of R(8, M). Suppose first there are points cj in R(6,M) converging to a such that the distance from cj to Te, tends to 0. The lower estimate of Theorem 1.4.3 featuring 6log(1/5) shows that hyperbolic disks of uniformly bounded hyper- bolic radius have euclidean diameters tending to 0 as the centers approach the boundary. Thus |PE**(c;) - re (cj)| > 0, and in the limit we obtain P{t*(a) = Pk(a). Thus a is preperiodic. Now the cycles containing 0 are superattracting and correspond to points in the interior of M. Since a € 0M and since 0 is its only predecessor, a must be strictly preperiodic, that is, a is a Misiurewicz point. Thus k > 1, and @ has even denominator. Next we analyze the case in which there are points cj in R(8, M) converging to a such that the distances from c; to Jc, are bounded away from 0. This situation occurs for instance for the rays R(0, Jc) as c > 1/4 decreases to 1/4, as illustrated in Figure 5, Section 1. The disconnected set J. closes up, and the limit ray goes into the “interior.” The point c on the ray converges to 1/4, strictly inside the filled-in Julia set Ky/4, which has the parabolic fixed point z = 1/2 on its boundary. Assuming dist(cj, Je, ) 2 6 > 0, Green’s function Ge, (2) is positive and harmonic in A(c;,6) and G.,(cj) — 0. Hence G.(z) = 0 on A(a, 6), and a belongs to a bounded component U of the Fatou set of P,. Since a € OM, U is not an attracting component. Since a € U, there are no Siegel disks. Hence U is in the basin of attraction of a parabolic cycle in Ja. Suppose z; € R(4,Ka) clusters at go € Ja. Arguing as before, this time in dynamic space, we see that qo must satisfy P£+*(qo) = P(qo). Since the solutions of this equation form a finite set, the ray must actually terminate at qo. Then R(2*6,Ka) terminates at the point q = PK(qo), and q is a periodic point with period dividing ¢. Our first goal is to show that is a parabolic periodic point. Suppose not. Since there is at most one attracting or neutral cy- cle, the cycle of q is repelling. As cj — 4, R(O, Je,) > R(0,Ka) in the sense that every compact part of R(6,Kq) is approached by R(O, Jc,). The appropriate branch of P;* has an attracting fixed point at q, so there is a disk A(q,¢) in the basin of attraction of g mapped into a disk of smaller radius by P7’. By the implicit func- tion theorem (see Lemma III.2.5), P-‘ has an attracting fixed point q(c) near q = q(a) for all c near a, whose basin of attraction includes VIII.5. External Rays with Rational Angles 145 A(q,€) and which maps A(q, €) into a compact subset of itself. For c near a the ray R(2*6, J.) remains in the disk A(q,¢) for a long pa- rameter interval, and the ray is invariant under Pf, so the ray must terminate at q(c). Thus for c near a the rays R(2*6, J.) are uniformly bounded away from the parabolic cycle. On the other hand, c; be- longs to R(O, Jc;) and c; > a, so Pretk(cs) belongs to R(2*6, Tc; ) and converges to P™+*(a). Consequently P+#(a) is bounded away from the parabolic cycle. This contradicts the fact that a is in the basin of attraction of the cycle. We conclude that q is a parabolic periodic point. The rays R(2/+*9, K,) terminate at P3(q), so the condition on 6 gives P£(q) = q. Since Pf leaves R(2*6,Kq) invariant, it does not rotate the petals at g, and (P£)/(q) = 1. Now the solution set of the pair of equations P£(z) = z, (P®)/(z) = 1, is a complex analytic variety in (z,c)-space. For each fixed z, the variety is bounded (in fact, c € M), so that c is locally constant on the variety. For fixed c, there are only finitely many solutions z. It follows that the variety is finite, and there are only finitely many possibilities for a. Since there are also only finitely many solutions of P{*(a) = P(a), there are only finitely many possibilities for the cluster points a of the ray R(0,M). It follows that the ray terminates, and the first statement of the theorem is proved. It remains to show in the parabolic case that 6 has odd denominator. For this, we begin by showing that the terminal point qo of the ray R(0,Kq) is actually a parabolic periodic point, and then we will argue that 6 is periodic. We parametrize R(0, Jo,) by z;(t) = CCE’) t > 0, so that Ge, (z;(t)) = t. Let q be the first point in the parabolic cycle that is a cluster point of the rays R(6, J.,) as they come in from oo. By this we mean that for « > 0 small and j large, there is t; > 0 such that |z(t;) — qi| = €, while the distance from 2;(t) to other points of the parabolic cycle is bounded away from 0 for all ¢ > tj, uni- formly in j. There is then 8 > 0 such that IPS (z(t) — 2,(t)| > 6 for t > t;. Consider g;(¢) = Pe, (1/4), which is univalent for IC] < exp(—G-,(0)) = pj. Note that t; > G.,(cj) = 2G.,(0) for j large, so lg; “(2 (t))| = e* < pj for t > t,, and the points 95 (z(t), t > tj, are uniformly bounded in the hyperbolic met- ric of the disk A(0,;) on which g; is univalent. Now if t > t; and 95 (25 (t)) =re?™, then 95 ' (PS (2i(t))) = 1% 29 is ata uniformly bounded hyperbolic distance from 95 (z(t) in the disk A(0, p;). 146 VIII. Quadratic Polynomials Since |g; (re?"") — 9; (r2“e2"#®)| > B, we see from the distortion theo- rem scaled to A(0, p;) that for some small 6 > 0 the image of A(0, p;) under g; contains a disk centered at Yc, (re?) of radius 6. Here 6 depends on ¢ and f but not on j. Thus for j large, R(0, Jc;) stays at a uniform distance 6 from Tey, and hence at a uniform distance from Jq, until entering an e-neighborhood of qi. It follows that qi must coincide with the terminal point qo of R(0,Ka). Thus qo isa parabolic periodic point. (With a little additional effort, we could establish that qo = q. The limit of the rays Ricj, Te,) is a ray which traverses R(, Ka) to the parabolic point q, then makes a sharp turn depending on the number of petals, and continues upstream along the flow lines in the basin of attraction of g to the critical value a and beyond.) For j > 1, the rays R(2’0,K) terminate at P3£(qo). Take n so that né > k. The preperiodicity condition shows that these rays coincide for j > n, and in particular they terminate at the same point P£(qo). Since P£ maps this family of rays to itself, and since P£ is conformal near go, there can be only one ray in the family. Hence R(0,Kq) is invariant under P£. We conclude that 26 = 6, and @ is periodic. 0 Note in the parabolic case that the period ¢ of @ coincides with the number of petals of the parabolic cycle of Pa. Indeed, the proof shows that @ divides the total number N of petals (= number cusps). By Theorem III.2.3, each petal contains a critical point of PN, so that the orbit of a hits each petal. Thus P, permutes the cusps cyclically, and f=N. Theorem VII.2.2 shows that there are a finite number of repelling arms for P£ terminating at any parabolic periodic point q, and these arms fill out the unbounded component of the Fatou set near q. Using Lindel6f’s theorem, we deduce that through each repelling arm travels exactly one external ray of Kq that terminates at q. Now P® leaves each cusp between consecutive petals invariant, it permutes the rays terminating at qg, and it preserves the circular order of the rays. It follows that each ray is invariant under P!, and furthermore the same number of external rays terminate through each cusp at each point of the cycle. It turns out that at most two rays terminate through each cusp. More precise information will be given in Theorem 7.1. VIIL.5. External Rays with Rational Angles 147 3/7 4/7 5/7 / \o FIGURE 9. External rays terminating at period-three parabolic points of M. EXAMPLE. Consider the external angles whose binary representa- tions have period three. There are exactly six of them, 1/7, ...,6/7, and each corresponding ray terminates at a parabolic point a € OM for which there is a cycle of attracting petals of length three (see Fig- ure 9). Either P, has a parabolic fixed point with multiplier e+?"/3, or Py has a parabolic cycle of period three and multiplier 1. The equations give only three possibilities for a, and these are the roots of the three hyperbolic components corresponding to attracting three- cycles. The rays R(1/7,M) and R(2/7, M) terminate at the point of tangency of the main cardioid and the big bulb at the top of the main cardioid. The two rays approach their terminal point from opposite directions and cut the Mandelbrot set into two pieces. Their reflec- tions in the lower half-plane are the rays R(5/7, M) and R(6/7, M). The rays R(3/7, M) and R(4/7, M) approach the root r of the third component from above and below the real axis. (See the example in Section 2.) In the dynamic plane, the two rays R(3/7,K,) and R(4/7,K,) approach a parabolic periodic point q on the real axis of period three from the respective half-planes. There is only one attracting petal at q. This provides the simplest example for which there is more than one ray terminating through the same cusp at a parabolic periodic point. 148 VIII. Quadratic Polynomials 6. Misiurewicz Points We turn now to a discussion of the converse to Theorem 5.2, that all Misiurewicz points and all points in M corresponding to parabolic cycles are the endpoints of rays R(0,.M) with @ rational. This in- volves the construction of curves in the complement of M that ter- minate at the point, and in turn this involves the deformation of paths in the Fatou sets F (P,) as we vary the parameter c. In the case of Misiurewicz points, we are deforming rays that terminate at repelling periodic points, which is considerably easier than deform- ing rays terminating at parabolic periodic points. (Viewed this way, repelling points are attractive.) In this section, we give a reason- ably complete treatment of the repelling (Misiurewicz) case. In the final section, we content ourselves with describing the results and indicating a deformation method for the parabolic case. We begin with a more detailed description of the Julia set corre- sponding to a Misiurewicz point. Let a be a Misiurewicz point, with critical orbit 04a > Py(a) 9° > Pia) > pi+k(a) = PE (a), so that a enters the cycle at the kth iteration and the cycle has period m. Since P, is subhyperbolic, the cycle is repelling. Since the only preimage of a is 0, the critical value a does not belong to the cycle, and k > 1. By Theorem V.4.2 the Julia set Ja = Ka isadendrite, and the inverse of the conjugating map Ya extends continuously to map the unit circle onto Ja. Thus every external ray R(0, Ka) terminates at a point of Ka, and every point of Kq has at least one external angle. By Theorem VII.2.2, there are a finite number N of repelling arms at P#(a), and these fill out the Fatou set near Pk(a). From Lindeléf’s theorem (Theorem 1.2.2) we see that through each repelling arm there is exactly one external ray terminating at P*(a). The appro- priate inverse branch of P-* maps these onto N arms of the Fatou set that fill out the Fatou set near a, each of which has exactly one external ray terminating at a. Let the external angles of Ka at a be 61,...,4v. The successive images of the ith arm under P, then have external angles 26; ,40;,..., all reduced mod 1. The external angles are circularly ordered, and since Pq preserves the order, so does the doubling operation on the external angles. Thus if one of the re- pelling arms at P(a) has period @, they all do, and consequently VIII.6. Misiurewicz Points 149 2k+f9, = 26; for 1 < i < N, that is, each of the 6,’s is strictly preperiodic, k is the first integer such that 2*6; is periodic, and the periods are all the same, equal to the period @ of the repelling arms. EXAMPLE. Consider the ray R(1/6,.M), which terminates at some aéM with Ima > 0. The binary representation 1/6 ~ (0,0, 1,0, 1, 0,1,...) is strictly preperiodic. From this representation we deduce that P3(a) = P,(a). Now P3(c) — P(c) = eMe-+ 2)(c + 1)%(c2 +1). The roots 0 and —1 corresponding to superattracting cycles can be discarded. The ray cannot terminate at —2, which is at the left tip of M. The only possibility is that R(1/6,.M) terminates at i. The arms at P;(i) = i—1 have period two. Since i— 1 — —i, there can be only one arm at each of the points i — 1 and —i, so these points lie at tips of the dendrite, as does a = i. See the figure in Section V.4. EXAMPLE. Let a be a Misiurewicz point such that P?(a) is a fixed point for P,, that is, such that P3(a) = P?(a). Since P3(c)— P2(c) = 8 +4c7 +68 +608 +.4ct = cA(c+2)(c82 + 2c? +2c+2), c we find upon discarding the superattracting cycle exactly four possi- bilities. The root a = —2 corresponds to the critical orbit 0 — —2 3 2—2-—.---, so P,(a) is already a fixed point. The cubic factor has one real root ro —1.54369 and two complex roots. One can verify that the rays with external angles 5 7 aa oO LOO tee) ea aa 0, Onl On 07s) Fy ~ 110,101... 35 + (1,0,0,1,0,1,0,...) correspond to a Misiurewicz point a on the real axis with two re- pelling arms, one above and one below the real axis, which are inter- changed by Pa, and that this a-value coincides with ro. The complex roots correspond to rays with external angles 1/4 ~ (0,0,1,1,1,...) and 3/4 ~ (1,1,0,0,0,...). Now we wish to transfer information about the Julia set at a to the parameter plane, to obtain information about M at a. We continue with the notation above. Consider again the polynomial Q(z,c) = P™(z) — z in z and c. Denote by zo = zo(a) the first point P*(a) of the repelling cycle, 150 VIII. Quadratic Polynomials and denote by (a) the multiplier of the cycle, so that \A(a)| > 1. Then zo is a fixed point of P;", so Q(z0,a) = 0. As before, the implicit function theorem guarantees that the equation Q(z,c) = 0 has a unique solution zo(c) near 2o(a) for c near a, which depends analytically on c. The point zo(c) is a repelling periodic point of period m for P., with multiplier A(c) depending analytically on c. Let h, map a disk {|¢| < 4} to a neighborhood of zo(c), so that he is a conjugation of P”” and multiplication by X(c): hel M(c)6) = Pr(he(G))s Ae) (2) = he (Pe*(2)), Ish < 6 We normalize h, so that h/,(0) = 1, and then h, depends analytically on the parameter c. Fix an external angle 6; of Ka at a, and let ¢ = 2*6;, so that R(@,Kq) terminates at zo(a). We consider the rays R(¢,K-) for ¢ near a. These vary analytically with c outside any neighborhood of 20(a). Near zo(c) the rays are invariant under P-™, and the relation Po™(2) = he(s7yhe'(2)) shows that the rays terminate at 2o(c), uniformly for c near a. For fixed t > 0, let 2:(c) be the point on the ray R(d, Ke) that satisfies G(z1(c)) = t, that is, pelze(c)) = ee’, t>0. Then z;(c) depends continuously on ¢ and c, 2(c) depends analyti- cally on c for each fixed t, and z(c) tends to zo(c) uniformly for ¢ near a as t decreases to 0. Now zo(c) does not coincide identically with P(c) near a, or we would have P™+*(c) = P§(c) identically. Thus for some v = 1 and A, #0, we have 2o(c) — P¥(c) = Av(e- a)” + Olle - a\’*1). c For t > 0 small, the equation 2(c) — P*(c) = 0 has v solutions ci(t),--,¢v(t) counting multiplicities, depending continuously on t and tending to a as t — 0. In particular, PE 4 (ci(t)) € Ro, Ke,(4))- By applying the appropriate branch of Poly, we obtain c(t) € R(8;,Ke,(t)): Hence ci(t) € R(6;,M), and it follows that R(8;,M) terminates at a. We have proved the following. VIII.6. Misiurewicz Points 151 THEOREM 6.1. If a is a Misiurewicz point, then there are a finite number of external rays R(0;,Ka) of Ka that terminate at a. Each external angle 0; of Kq at a is rational and strictly preperiodic. More- over, each ray R(0;,.M) of the Mandelbrot set terminates ata € M. In particular, every Misiurewicz point belongs to OM and is the ter- minal point of R(@,M) for some strictly preperiodic 0. We wish to analyze the situation at the point a more carefully. We aim to show that the Julia set J is asymptotically similar at a to the Mandelbrot set M, and in particular if C\.% has N arms at a, then C\M also has N arms at a. We begin by showing that the multiplicity v above is 1. LEMMA 6.2. With the notation above, zo(c) — P#(c) has a simple zero at a, as does P™+*(c) — Pk(c). Proof. From Pex(t) (PE (ca(t))) = Be,ce) (24(ex(€))) = ete?™**, we have, since ¢;(t) € R(6;,Ke,(x)), that e,(t)(ci(t)) = e/ e243, Since ® is one-to-one, the c;(t)’s coincide, call their common value c(t). Thus z(c) — PE(c) = O(le-e(t)|"), — e > e(2). Now ¢-(z) = ete? + O(|z — :(c)]), so (c)”* = y(PE(c)) = ete? + O(le—e(t)|”), > e(t). But ®(c)”* has nonzero derivative at c = c(t). We conclude that v =1, and zo(c) — P#(c) has a simple zero at a. We calculate ge) PHO] = la) - 2 [PMO -aale)]_ so the second statement is equivalent to the first. O For c near a, define Z(c) = haohe'(PS(c)) = zo(a) + Olle — al). 152 VIII. Quadratic Polynomials Since hi,(0) = 1, we have h>1(P¥(c)) = P¥(c) — 29(e) + O(|P(c) — z0(c)|*). From the preceding lemma, hz1(P¥(c)) has a simple zero at a, so Z(c) is univalent near a. We wish to compare the orbits P/+*(c) and PJ™(Z(c)), j 2 1. We compare the orbits in conjugation space, where the dynamics are simply multiplication by A(c) and (a), respectively. We have defined Z(c) so that the starting points Cy = hoi (PE(c)) = hz} (Z(c)) are the same, and at time jm we have X(c)/o and d(a)¢o, respectively. We iterate until we are about to leave a disk of fixed size, say {|¢| < Ci}, at the Jth iteration, where J = J(c). For Co we have the estimate IGol = |ha*(Z(€))| 2 Cole — al, since hz! is analytic and hg *(z0(a)) = 0. Thus Cole—al|A(a)|? < C1, and J < C2 log le~ al" Since (c) is an analytic function of c, we have Ac)/A(a) = 1+ O(|c — a|), and thus J kolacan’ ft - (12) Cy|1 - (1+ Olle = al)” | [X(c)"Go — Xa)" Col lA 1 C3\c — a| log jenal’ IA Going back to the z-plane, we obtain an estimate of the form \Pm**(e) — PIM(Z(e))| ” is translation by 1 in the é-coordinate. Let S_ be the subset of S' represented by parameter values a +e 4 in at most No steps by Pe for all c near a. We claim that if c is near a and Z(c) € Se, then c ¢ M. Indeed the estimate (6.1) shows that when eventually the iterates P3™(Z(c)) hit VIIL7. Parabolic Points 153 T., the corresponding iterates Pj™+*(c) hit Ve, from whence in No more steps they move to |z| > 4 and then beyond towards oo. Thus the image of S under Z~! is asymptotic to an “arm” of C\M at a. Now suppose that ¢ = 2*6; is the external angle of the ray termi- nating at zo(a) through S. Since the ray R(¢,Ka) is periodic in €- space, it lies in S2_ for € > 0 small. For c(t) near a the ray R(¢, Ke) passes through T,. Hence for c(t) sufficiently close to a, we obtain from (6.1) that Pj™(Z(c(t))) € Ve, and consequently Z(c(t)) € S. Thus the ray R(9;, M) terminates at a through the set Z~!($). The argument shows in fact that the hyperbolic distance between points of R(9;,Kaq) and R(d;,M), with respect to C\M, tends to 0 as the rays tend to their common terminal point a. We have proved the following version of a theorem of Tan Lei (Exposé No. XXIII of [DH2)}). THEOREM 6.3. Let a € M be a Misiurewicz point. Let 0,...,0N be the external angles of Ja ata, and let L; be the arm of Fa containing R(9;,Ja), 80 that S; = P¥(L;) is a repelling arm at P¥(a) as above, and Ly,..., Lyn fill out the Fatou set near a. Parametrize L; by ¢ in a strip aj < Im¢ < b;, Re¢ — —oo, and for e > 0 small let Lj. be the subset of the arm corresponding to a; +e < Im¢ < b; — e. Let Z be as above, and define the conformal map g = Z~! 0 P#, so that g(a) =a and g'(a) # 0. Then g(L;.) eventually lies in C\M, and the ray R(0;,M) terminates at a through g(Lj.). One can note that computer-generated pictures of certain pieces of the Mandelbrot set in parameter space look strikingly similar to pieces of Julia sets in dynamic space, with swirling arms configured in the same pattern. See particularly Figures 4.22 and 4.23 on pages 204-206 of [Pe], and also the figures in [Ta]. An important step in the proof of Shishikura [Sh2] that the boundary of the Mandelbrot set has Hausdorff dimension two hinges on the local similarity of the Mandelbrot set and Julia sets. 7. Parabolic Points As noted before, the analysis of parabolic points of M is more dif- ficult than that of Misiurewicz points. We summarize the situation 154 VIII. Quadratic Polynomials in Theorems 7.1 and 7.2, without supplying detailed proofs. For a complete discussion, see [DH2]. THEOREM 7.1. Suppose P, has a parabolic cycle of length m with multiplier . If X #1, then between any pair of consecutive petals at any point q of the cycle, there is exactly one external ray of Ka that terminates at q, and these are permuted cyclically by Pa. IfX=1 and a # 1/4, then there are two external rays terminating at any point of the cycle, and the set of pairs is permuted cyclically by Pa. In any case, the corresponding external angles are all rational and periodic, with period £ = mN, where N is the number of petals at q. The external rays of Ka in dynamic space are again related to those of the Mandelbrot set in parameter space. By Theorem 5.2, every external ray R(0,M) of the Mandelbrot set with periodic 0 terminates at a parabolic point a « M. Conversely, every parabolic point a is the terminal point of such a ray. THEOREM 7.2. Suppose that a € M is such that P, has a parabolic periodic point. Except in the trivial case a = 1/4 (corresponding to a parabolic fixed point, with one ray R(0,M) terminating at a), there are exactly two rays R(A1,M), R(G2, M) terminating at a. The external angles 01,02 coincide with the angles of the external rays of Ka tangent to and on either side of the petal containing a. We conclude by focusing on one part of the proofs of Theorems 7.1 and 7.2, the existence of an external ray of M that terminates at the parabolic point a. This will serve to shed light on the phenomena behind the results and what kind of arguments go into the proof. For £ the period of a petal, as above, we expand PM(z) = Yraj(o(z- 9), where ao(a) = 0 and ay(a) = 1. If the flower at q has N petals, then a2(a) = ++: = an(a) = 0 and ay+1(a) # 0. We will consider only the special case N = 1; that is, we assume there is only one petal at q. By moving the point near q where (P£)! = 1 to 0 and scaling, we can then assume that our family P£(z) has the expansion flac) = vo? +242 + BOA + O(z) = yc +2+ Fe: VIII.7. Parabolic Points 155 Here +(c)? is analytic and not identically zero, y(a)? = 0, and 7(c) is the principal branch of the square root. The point z = 0 is a parabolic fixed point for f(z,a), which belongs to the Julia set and also to the boundary of the unbounded component of the Fatou set of f(z, a). Let z§ belong to this component and be very close to z = 0. We denote by I* a compact subset of this component containing zj, which will be specified later. The target set I* will be situated near 0 in the repelling arm of the Fatou set containing z}, and it will be large enough so that the following statement holds. Main deformation construction. Let zo(c) be any function analytic near a so that z0(a) is in the parabolic petal of 0. Then there is a curve I in parameter space, terminating at a, so that for c € T, c# a, there exists n = n(c) such that f"(zo(c),c) €I*. In our application, zo(c) is the point corresponding under the con- jugation to the critical point 0. Thus if c € T is near a, the critical point is iterated first to a point in a fixed compact subset of the unbounded component of the Fatou set of f(z,a) and then, by con- tinuity, to oo. Hence I is in the complement of the Mandelbrot set. By Lindeléf’s theorem, this implies there exists a ray R(6,M) which terminates at a. Since the iterates of z9(a) eventually approach 0 through a narrow cusp tangent to the negative real axis, we focus on initial values zo(c) that lie in some fixed domain 2 containing —c, so that x ~ —e and lyl < on 9. We study parameter values c for which y(c) = a + i2 where a > 0 and || < a. In fact, we shall see we can take |3| < Ca? or |B| < Ca? log(1/a), depending on whether Im B(a) is 0 or not. We regard a as the main parameter, and we seek parameter values c = c(q) so that the parabolic fixed point at 0 splits into two fixed points, above and below the real axis, opening a window through which the iterates pass from (2 in the left half-plane to I'* in the right. As the window narrows, the number of iterations required to pass through it becomes large. Now the iteration Znt1 = Zn +7(c)? + F(zns0) (7.1) can be approximated by the flow dz(t) ie ¥(c)? + F(z,0), 2(0) =z =x2o+iyoEN. = (7.2) 156 VIII. Quadratic Polynomials The equation (7.2) is easily solved for t, - ee t=t(z,c) = [ yc)? + F(G,¢)" We are interested in the point iy where the solution curve meets the imaginary axis. The corresponding parameter value t is given by t= I + [. + f Wer aa e The first integral is analytic in c. In the second, we expand the integrand 1/[7(c)? + F(¢,¢)] as 1 _ Bee Boyere , __ O6") Hort P+ (VC? +67)? © (V(c)? + 07)?” and we integrate, thereby obtaining t= A(z d+ ats + BOE +f aS aay ar 210) ve) Jo (ce)? — 8? + O(s%)’ where A is analytic and uniformly bounded for Re7(c)? > 0. Note also that for 7 = a+i6 as above, |dB/dy| < Cly|°-? for some 6 > 0, and |dA/dy| < C/|11- For fixed a > 0, the condition on 6 that the curve pass through the origin is that mB, ReB(c) 1, ImB(c) 1 mort 2 MyeR* 2 ioe ® Using |7(c)|? = a? and Arg (c)? = 28/a, we see that the expression on the left moves continuously from positive to negative as B in- creases between the limits +Ca?log(1/a). Hence it is zero for some B & (2/n)Im B(a)a” log(1/a). For || < Ca? log(1/a), the curve crosses the imaginary axis at iy, where Im A(z, ¢) ye 58 —a?Im A(zo, c) + aGRe B(c) — a” log(1/a) Im B(c). Using the estimates mentioned above, we obtain dy 7 7 2 tO: VIII.7. Parabolic Points 157 In particular, the y-value of the crossing point is an increasing func- tion of B. Now, let f\ (z,c) determine a similar flow in the opposite direction, starting at Z and approaching the imaginary axis from the right. We assume f has the same leading term but with opposite sign: f(z,c) = —y(c)? +2 — 27 + B(c)z* + O(24) = z - (y(c)? + F(z, 0). We conjugate z — —z and use the same argument. In this case, the y-value of the crossing point on the imaginary axis is a decreasing function of 3. Thus we conclude the following. THEOREM 7.3. Given z(c), Rezo < 0 and Z(c), ReZo(c) > 0, as above, there exists a Jordan curve T of parameter values c terminat- ing at c = a and satisfying a > 0 and |B| < Ca? log(1/a), where y(c) = a+ iP, so that the two solution curves of dz(t) aed yc)? + F(z,c), Rez <0, z(0) = z0(c), BO = -y(o?- FG), Reé20, #0) = 400), meet on the imaginary azis. We can also apply the analysis to f~!(z,c), which has the expan- sion 2-40)? ~ (z= ¥(0)?)? — (Ble) — 2)(z ~ 4(0)?)? + O((z - ¥(0))*). After a conjugation 2’ = z — 7(c)®, this assumes the form f(z,c) treated above. The lines of flow of f~!(z,c) arrive from the right, cross over the vertical line Re (z — 7(c)?) = 0, and hit the imaginary axis at a crossing point that is a decreasing function of 3. Thus we obtain the main deformation result, in the case of flows, with I'* the single point z. When we consider iterations rather than flows, the situation is quite analogous, but the technical changes are somewhat tedious. We face two problems. (a) For the discrete iteration, we cannot expect to hit the imaginary axis, but to come within C|y|?. This is the rea- son why I* must be considered. 158 VIII. Quadratic Polynomials (b) We wish to use the flow to follow the iteration. However, the (Euler) approximation (7.2) to (7.1) is not accurate enough. Let us deal first with the second problem. Instead of using (7.2), we consider the flow dw(t) | dt which can be regarded as a (backward) Runge-Kutta approximation for (7.2). 4(c)? + w? + (Ble) - 1)w®, (7.3) Lemma. If 29 = xo + iyo, where —€ < to < —Ja and |yo| < |xol, then z0+ f (z0,.¢) dz | = 1+ O(\z0!?). x 0 y(c)2 + 22 + (B(c) — 1)28 Proof. Expand the integrand, integrate the leading terms in closed form (as before), and make the obvious estimates. 0 Now let We be a rectangular window centered at —Va, of width (say) 3a, and of height sufficient so that the flow lines of (7.3) starting in 2 pass through W,. Take Cy € Wa on such a flow line, and consider backward iterates ¢; = f(¢j+1)- We compare the polygonal line P, starting at Co and passing through (1,...,¢n to the backward flow w(—t) starting at the same point (9. From (3, = Gj — G41 +O(G 41) and Re (Gj — Cj41) ~ |¢j — Gj+i|, we obtain d¢ > = -1+0 |?) = —-n+O : [ y(c)? + Ce + (B(c) = 1)¢3 nt o> Gil ) n+ (l¢nl) Since the integrand behaves like 1/¢?, this shows that w(—n) = Gat O(|Cn|3). We conclude that if zo € 9, there is a point wo satisfying |wo — 20| = O(e3), such that the forward flow (7.3) starting at wo follows closely the iterates of zp and meets them in W,. Now let Dg be a similar window, centered at 0, of width (say) 3a? and height Ca” log(1/a). The curvature of the flow lines of (7.2) is determined by Pz naa (22 + 3B(c)2” ae dz dt’ VIII.7. Parabolic Points 159 The curve Im (2z+3B(c)z?+--+) = 0 is analytic at 0 with horizontal tangent. Above this curve, where the imaginary part is positive, the flow lines are concave up, and below it they are concave down. Thus there is a cusp domain at 0 of the form U = {Rez < 0, |Imz| < co|z|?}, co independent of +, such that whenever an iterate Zm lies above U, the succeeding iterate zm41 lies below the flow line through 2m; and similarly when 2, lies below U. The iterates z,, are chan- neled closer to the real axis than the flow lines, and consequently they also cross the imaginary axis through the window Dy. Without supplying details, we describe now our strategy for con- structing I’. For convenience, assume zo(c) = zo € (2. The iterates f"(zo,c) pass through W, and cross over the imaginary axis in Da, moving from left to right. By the same token, the inverse iterates of 26 eventually traverse D, from right to left. By varying the param- eters, we find ag and (% so that for yp = ag + if and corresponding co, y(co) = 70, @ forward iterate f"(zo,co) meets a backward iter- ate f~™(z2),co) within Da. Then f"*™(zp,co) = 2. Let I* be a compact tubular neighborhood of a curve in the repelling arm of 6 joining z} to f—1(zg,a). In terms of the coordinate for the repelling arm (Section VII.2), we can take I* to be a thin horizontal rectan- gle with these two points situated near the ends of the coordinate rectangle. Now f~™(I*, co) is a tubular domain containing f”(zo, co) near one end, and our task is to vary the parameters a and (3 so as to steer through the (continuously varying) tube to the other end, that is, to obtain a curve a; + i = y% = 7(c), 0 < t < 1, so that f"(zo, ce) € f-™(I*, ¢) and so that f"(zo,c1) = f7"- (z8c1). Then f"*™+1(z9,c1) = 23, and we may repeat this process, noting that c tends to a as t — oo, to obtain the desired curve I. This concludes the outline of the main deformation construction. Epilogue Complex dynamics leads in many different directions, and there are a number of interesting and important facets that we have not even touched upon. We would have liked to have discussed the Hausdorff dimension of Julia sets. Although we have not mentioned ergodic theory, it plays today an important role. The idea of holomorphic motions, for which a prototype is a family of Julia sets depending analytically on a parameter, leads to fruitful developments. A the- ory of iteration of entire functions has been extensively developed, and a theory of iteration of analytic functions in several variables is beginning to take shape. For further orientation we mention the expository article of Eremenko and Lyubich (1990), and for several variables the recent paper of Fornaess and Sibony (1992). References [AJ] [A2] [A3] [Ar] [Ba] [Be] [Be2] L.V. Ahlfors, Complex Analysis, McGraw-Hill, 1953 L.V. Ahlfors, Lectures on Quasiconformal Mappings, Van Nostrand, 1966 L.V. Ahlfors, Conformal Invariants: Topics in Geometric Function Theory, McGraw-Hill, 1973 V.I. Arnold, “Small denominators. I: On the mappings of the circumference onto itself,” Izv. Akad. Nauk SSSR Ser. Mat., 25, 1961, pp. 21-86; Amer. Math. Soc. Translations (2), 46, 1965, pp. 213-284 IN. Baker, “An entire function which has wandering do- mains,” J. Austral. Math. Soc. (Series A), 22, 1976, pp. 173-176 A.F. Beardon, “Iteration of contractions and analytic maps,” J. London Math. Soc., 41, 1991, pp. 141-150 A.F. Beardon, Iteration of Rational Functions, Springer- Verlag, 1991 164 References [BeP] [Beu] [Bl] [Bo] [Bra] [Bro] [BrM] [Brjul] [Brju2] [Cal] [Ca2] AF. Beardon and Ch. Pommerenke, “The Poincaré metric of plane domains,” J. London Math. Soc., 41, 1979, pp. 475-483 A. Beurling, The Collected Works of Arne Beurling, Vol- ume 1, Complex Analysis, L. Carleson, P. Malliavan, J. Neuberger and J. Wermer (editors), Birkhauser, 1989 P. Blanchard, “Complex analytic dynamics on the Rie- mann sphere,” Bull. Amer. Math. Soc., 11, 1984, pp. 85- 141 LE. Boettcher, “The principal laws of convergence of it- erates and their applications to analysis (Russian), Izv. Kazan. Fiz.-Mat. Obshch., 14, 1904, pp. 155-234 B. Branner, “The Mandelbrot set,” In: Chaos and Fractals: The Mathematics Behind the Computer Graphics, R.L. Devany and L. Keen (editors), Proceedings of Symposia in Applied Mathematics, Vol. 39, Amer. Math. Soc., 1989, pp. 75-105 H. Brolin, “Invariant sets under iteration of rational func- tions,” Arkiv fér Mat., 6, 1965, pp. 103-144 R. Brooks and J.P. Matelski, “The dynamics of 2-generator subgroups of PSL(2,C), In: Riemann Surfaces and Re- lated Topics: Proceedings of the 1978 Stony Brook Con- ference, I. Kra and B. Maskit (editors), Annals of Math. Studies No. 97, Princeton Univ. Press, 1981, pp. 65-71 A.D. Brjuno, “Convergence of transformations of differen- tial equations to normal forms,” Dokl. Akad. Nauk USSR, 165, 1965, pp. 987-989 A.D. Brjuno, “Analytical form of differential equations,” Trans. Moscow Math. Soc., 25, 1971, pp. 131-288 L. Carleson, “Removable singularities of continuous har- monic functions on R™,” Math. Scand., 12, 1963, pp. 15- 18 L. Carleson, Complex Dynamics, UCLA Course Notes, 1990 [CaJ] CaJY] Cal] Ca2] Cr] Del] De2] Dol] Do2] Do3] DH] DH2| References 165 L. Carleson and P. Jones, “On coefficient problems for uni- valent functions and conformal dimension,” Duke Math. J., 66, 1992, pp. 169-206 L. Carleson, P. Jones and J.-C. Yoccoz, “When is Julia John?,” in preparation, 1992 A. Cayley, “The Newton-Fourier imaginary problem,” Amer. J. Math., 2, 1879, p. 97 A. Cayley, “Application of the Newton-Fourier method to an imaginary root of an equation,” Quarterly J. Pure Appl. Math., 16, 1879, pp. 179-185 H. Cremer, “Uber der Haufigkeit der Nichtzentren,” Math. Ann., 115, 1938, pp. 573-580 A. Denjoy, “Sur l’itération des fonctions analytiques,” C.R. Acad. Sci. Paris, 182, 1926, pp. 255-257 A. Denjoy, “Sur les courbes définies par les équations différ- entielles 4 la surface du tore,” J. Math., 11, 1932, pp. 333- 375 A. Douady, “Systemes dynamiques holomorphes,” Sémin- aire Bourbaki, Volume 1982-83, exposé no. 599, Astérisque, 105-106, 1983, pp. 39-63 A. Douady, “Julia sets and the Mandelbrot set, In: The Beauty of Fractals, H.-O. Peitgen and P. Richter, Springer- Verlag, 1986, pp. 161-173 A. Douady, “Disques de Siegel et anneaux de Herman,” Séminaire Bourbaki, Volume 1986-87, exposé no. 677, As- térisque, 152-153, 1987, pp. 151-172 A. Douady and J.H. Hubbard, “Itération des polynémes quadratiques complexes,” C.R. Acad. Sci. Paris, 294, 1982, pp. 123-126 A. Douady and J.H. Hubbard, “Btude dynamique des poly- nomes complexes, (Premiére Partie) ,” Publ. Math. d’Orsay 84-02, 1984; “(Deuxiéme Partie),” 85-02, 1985 166 References DH3] Ec] ELe] ELyu] Fal] Fa2] Fi] FoS] Fu] Gh] GuR] [HaW] A. Douady and J.H. Hubbard, “On the dynamics of polyno- mial-like mappings,” Ann. Sci. Ecole Norm. Sup., 18, 1985, pp. 287-343 J. Ecalle, “Théorie itérativ: introduction a la théorie des invariants holomorphes,” J. Math. Pure Appl., 54, 1975, pp. 183-258 AE. Eremenko and G.M. Levin, “Periodic points of poly- nomials,” Ukrainian Math. J., 41, 1989, pp. 1258-1262 A.E. Eremenko and M.Yu. Lyubich, “The dynamics of an- alytic transformations,” St. Petersburg Math. J., 1, 1990, pp. 563-633 P. Fatou, “Sur les solutions uniformes de certaines équa- tions fonctionnelles,” C.R. Acad. Sci. Paris, 143, 1906, pp. 546-548 P. Fatou, “Sur les équations fonctionnelles,” Bull. Soc. Math. France, 47, 1919, pp. 161-271; 48, 1920, pp. 33- 94, pp. 208-314 Y. Fisher, “Exploring the Mandelbrot set, In: The Science of Fractal Images, H.-O. Peitgen and D. Saupe (editors), Springer-Verlag, 1988, pp. 287-296 JE. Fornaess and N. Sibony, “Complex Hénon mappings in C? and Fatou-Bieberbach domains,” Duke Math. J., 65, 1992, pp. 345-380 W.H.J. Fuchs, Topics in the Theory of Functions of One Complex Variable, van Nostrand, 1967 E. Ghys, “Transformation holomorphe au voisinage d’une courbe de Jordan,” C.R. Acad. Sci. Paris, 289, 1984, pp. 385-388 R. Gunning and H. Rossi, Analytic Functions of Several Complex Variables, Prentice-Hall, 1965 G.H. Hardy and E.M. Wright, An Introduction to the The- ory of Numbers, Clarendon Press, 1938 References 167 Hei] M. Herman, “Sur la conjugaison différentiable des difféo- morphismes du cercle & des rotations,” Publ. Math. IHES, 49, 1979, pp. 5-233 He?2] M. Herman, “Exemples de fractions rationnelles ayant une orbite dense sur la sphére de Riemann,” Bull. Soc. Math. France, 112, 1984, pp. 93-142 He3] M. Herman, “Are there critical points on the boundaries of singular domains?,” Comm. Math. Phys., 99, 1985, pp. 593-612 Ho] K. Hoffman, Banach Spaces of Analytic Functions, Prentice- Hall, 1962 HuW] W. Hurewicz and H. Wallman, Dimension Theory, Prince- ton Univ. Press, 1941 Jul G. Julia, “Mémoire sur l’itération des fonctions ration- nelles,” J. Math. Pures Appl. (7th series), 4, 1918, pp. 47-245 Ke] L. Keen, “Julia sets,” In: Chaos and Fractals: The Math- ematics Behind the Computer Graphics, R.L. Devany and L. Keen (editors), Proceedings of Symposia in Applied Mathematics, Vol. 39, Amer. Math. Soc., 1989, pp. 57-74 Ko] G. Koenigs, “Recherches sur les intégrales de certaines équations fonctionnelles,” Ann. Sci. Ecole Norm. Sup. (3rd series), 1, 1884, pp. Supplément 3-41 Lat] S. Lattés, “Sur l’itération des substitutions rationnelles et les fonctions de Poincaré,” C.R. Acad. Sci. Paris, 166, 1918, pp. 26-28 Lav] P. Lavaurs, “Une description combinatoire de l’involution définie par M sur les rationnels & dénominateur impair,” C.R. Acad. Sci. Paris, 303, 1986, pp. 143-146 Le] L. Leau, “Etude sur les équations fonctionnelles & une ou plusiérs variables,” Ann. Fac. Sci. Toulouse, 11, 1897, pp. E.1-E.110 168 References Lyu] Mal] Ma?2] Mii] MaR] MSS] MaN] [Mit] [Miz] [NaV] [Pe] [PeR] M.Yu. Lyubich, “The dynamics of rational transforms: the topological picture,” Russian Math. Surveys, 41, 1986, pp. 43-117 B. Mandelbrot, “Fractal aspects of the iteration of z > dz(1—2z) and z for complex 4 and z,” Annals of New York Acad. Sci., 357, 1980, pp. 249-259 B. Mandelbrot, “Fractals and the rebirth of iteration the- ory, In: The Beauty of Fractals, H.-O. Peitgen and P. Rich- ter, Springer-Verlag, 1986, pp. 151-160 R. Mané, “On the instability of Herman rings,” Inven- tiones Math., 81, 1985, pp. 459-471 R. Maiié and L.F. da Rocha, “Julia sets are uniformly perfect,” Proc. Amer. Math. Soc., 116, 1992, pp. 251-257 R. Maiié, P. Sad and D. Sullivan, “On the dynamics of rational maps,” Ann. Sci. Ecole Norm. Sup., 16, 1983, pp. 193-217 N.S. Manton and M. Nauenberg, “Universal scaling behav- ior for iterated maps in the complex plane,” Comm. Math. Phys., 89, 1983, pp. 555-570 J. Milnor, “Self-similarity and hairiness in the Mandelbrot set, In: Computers in Geometry and Topology, M.C. Tan- gora (editor), Lecture Notes in Pure and Applied Mathe- matics, Vol. 114, Marcel Dekker, 1989, pp. 211-257 J. Milnor, Dynamics in One Complex Variable: Introduc- tory Lectures, Institute for Math. Sci., SUNY Stony Brook, 1990 R. Nakki and J. Vaisala, “John disks,” Expositiones Math., 9, 1991, pp. 3-43 H.-O. Peitgen, “Fantastic deterministic fractals, In: The Science of Fractal Images, H.-O. Peitgen and D. Saupe (editors), Springer-Verlag, 1988, pp. 168-218 H.-O. Peitgen and P. Richter, The Beauty of Fractals, Springer-Verlag, 1986 [Pf] [Pol] [Po2] [Po3] Scl] Sc2] Shi] Sh2] si] SiM] [Sp] [St] [Sul] References 169 G.A. Pfeifer, “On the conformal mapping of curvilinear angles. The functional equation ¢[f(zx)] = ai¢(z),” Trans. A.M.S., 18, 1917, pp. 185-198 Ch. Pommerenke, Univalent Functions, Vandenhoeck and Ruprecht, 1975 Ch. Pommerenke, “Uniformly perfect sets and the Poincaré metric,” Arch. Math., 32, 1979, pp. 192-199 Ch. Pommerenke, “On conformal mapping and iteration of rational functions,” Compler Variables Th. Appl., 5, 1986, pp. 117-126 E. Schréder, “Ueber unendlichviele Algorithmen zur Auflé- sung der Gleichungen,” Math. Ann., 2, 1870, pp. 317-365 E. Schréder, “Ueber iterirte Functionen,” Math. Ann., 3, 1871, pp. 296-322 M. Shishikura, “On the quasiconformal surgery of rational functions,” Ann. Sci. Ecole Norm. Sup., 20, 1987, pp. 1-29 M. Shishikura, “The Hausdorff dimension of the boundary of the Mandelbrot set and Julia sets,” Annals of Math., 1993 C.L. Siegel, “Iteration of analytic functions,” Annals of Math., 43, 1942, pp. 607-612 C.L. Siegel and J.K. Moser, Lectures on Celestial Mechan- ics, Springer-Verlag, 1971 G. Springer, Introduction to Riemann Surfaces, Addison- Wesley, 1957 E.M. Stein, Singular Integrals and Differentiability Prop- erties of Functions, Princeton Univ. Press, 1970 D. Sullivan, “Conformal dynamical systems,” In: Geomet- ric Dynamics, J. Palis (editor), Lecture Notes in Math., Vol. 1007, Springer-Verlag, 1983, pp. 725-752 170 [Su2] [Ta] [Ts] [Vo] [Yo2] References D. Sullivan, “Quasiconformal homeomorphisms and dy- namics, I, Solution of the Fatou-Julia problem on wander- ing domains,” Annals of Math., 1985, 122, pp. 401-418 Tan Lei, “Similarity between the Mandelbrot set and J ulia sets,” Comm. Math. Phys., 1990, 134, pp. 587-617 M. Tsuji, Potential Theory in Modern Function Theory, Maruzen, 1959 S.M. Voronin, “Analytic classification of germs of confor- mal mappings (C,0) — (C,0) with identity linear part,” Functional Analysis Appl., 15, 1981, pp. 1-13 J. Wolff, “Sur l’itération des fonctions bornées,” C.R. Acad. Sci. Paris, 182, 1926, pp. 200-201 M.V. Yakobson, “The boundaries of certain normality do- mains for rational maps,” Russian Math. Surveys, 39:6, 1984, pp. 229-230 J.-C. Yoccoz, “Conjugation différentiable des difféomor- phismes du cercle dont le nombre de rotation vérifie une condition diophantienne,” Ann. Sci. Ecole Norm. Sup., 17, 1984, pp. 333-359 J.-C. Yoccoz, “Linéarisation des germes de difféomorphis- mes holomorphes de (C,0),” C.R. Acad. Sci. Paris, 306, 1988, pp. 55-58 Index admissible metric, 91 Ahlfors, L.V., vi, 19 airplane set, 129 area theorem, 1 Arnold, V.L, 44, 48 Arnold ring, 74 attracting cycle, 54 attracting direction, 40 attracting fixed point, 27, 31 attracting flower, 127 attracting periodic point, 54 attracting petal, 38 Baker, I.N., 70 basin of attraction, 28 of attracting cycle, 58 of parabolic cycle, 60 Beardon, A.F., 64, 79 Beltrami coefficient, 16 Beltrami equation, 16 Beurling, A., 14 Beurling transform, 18 Bieberbach, L., 5 binary expansion, 142 Bishop, C., v Blanchard, P., v Boettcher, L.E. (Béttcher, Botkher), 33 boundary scanning method, 65 Brooks, R., 128 Brjuno, A.D., 46 budding, 132 Carathéodory, C., 6 Carathéodory’s theorem, 6 Carleson, L., 86, 139 Cauchy-—Green formula, 17 cauliflower set, 97, 115, 130 Cayley, A., 29 completely invariant set, 56 conjugation, 28 covering map, 10 Cremer, H., 43 critical point, 54 172 Index closure of forward orbit, CL, 81 cycle, 54 attractive, 148 repulsive, 173 da Rocha, L.F., 64 deBranges, L., 5 dendrite, 94 Denjoy, A., 48, 79 dilatation, 16 Diophantine number, 43 distance-estimator algorithm, 141 distortion theorem, 3 Douady, A., 29, 62, 97, 99, 134, 140 Douady’s rabbit, 122, 129 ellipse field, 16 entire functions of finite order, 33 Eremenko, A.E., 113, 161 exceptional set, 57 expanding map, 89 external angle, 137, 140 external ray, 137, 140 Fatou, P., 53, 62, 63, 91 Fatou set, F, 54 Feigenbaum point, 133, 141 filled-in Julia set, K, 65 fixed point, 27 Fornaess, J.E., 138, 161 golden mean, 84, 127 Green’s function, 14, 34 of Julia set, 136 of Mandelbrot set, 139 Green’s lines, 137 Hausdorff dimension, 65, 139 Herman, M., 48, 86, 103, 122 Herman ring, 74, 103 Hubbard, J., 29, 99, 134, 140 hyperbolic components of inte- rior of M, 128, 133 center, 136 root, 132 hyperbolic metric, 11 hyperbolic mapping, 89 immediate basin of attraction, 28, 58 inverse iteration method, 57 Jacobian, Js, 15 John domain, 117 Jones, P., vi, 86 Julia, G., 53, 63 Julia set, 54 area, 90 capacity, 137 connectedness, 66 dimension, 139 equals C, 58, 82 filled-in, 65 hyperbolic, 89 local connectedness, 93 uniformly perfect, 64 KAM theory, 44 Kleinian group, 91 Koebe one-quarter theorem, 2 Koenigs, G., 31 Kolmogorov, A.N., 44 Kuusalo, T., 48 Lattés, S., 30 Levin, G.M., 113 Lindeléf, E., 6 Liouville, J., 43 Littlewood, J.E., 5 locally connected set, 6 Lyubich, M.Yu., v, 161 main cardioid, 126 Mandelbrot, B., 128 Mandelbrot set, 124 connectedness, 140 Maié, R., 64 Matelski, J.P., 128 Milnor, J., v, 141 Misiurewicz point, 133 modular function, 10 Montel, P., 5, 9 Montel’s theorem, 10 Moser, J., 44 multiplier of fixed point, 27 multiplier of periodic point, 54 neutral fixed point, 27 irrational, 27, 41 rational, 27, 35 Newton’s method, 30 normal family, 5 orbit, 53 parabolic component, 74 parabolic fixed point, 37 Peitgen, H.-O., v period doubling of Feigenbaum, 132 Index 173 periodic component of Fatou set, 69, 74 Periodic point, 54 petal, 38 Pick, G., 13 Pfeifer, G.A., 43 Poincaré, H., 48 Poincaré (hyperbolic) metric, ll polynomial-like mapping, 99 Pommerenke, Ch., 64 postcritical set, 81 preperiodic component of F, 70 preperiodic point, 54 quasicircle, 101 quasiconformal mapping, 16 k-quasiconformal mapping, 16 quasiconformal surgery, 99 Raphson, J., 30 rationally neutral fixed point, 27, 35 repelling arm, 113 repelling fixed point, 27, 32 repelling periodic point, 54 dense in Julia set, 63 repulsive cycle, 172 Richter, P., v Riemann mapping theorem, 5 Robin’s constant, 35 rotation domain, 74 rotation number, 47 Schréder, E., 29 Schréder equation, 29, 41 174 Index Shishikura, M., vi, 62, 104, 105, 153 Sibony, N., 138, 140, 161 Siegel, C.L., 43 Siegel disk, 55 snail lemma, 77 Stolz angle, 7 subhyperbolic mapping, 92 Sullivan, D., 15, 70, 97, 105, 134 superattracting fixed point, 27, 33 superattracting cycle, 54, 128 Tan Lei, 153 Thurston, W., 141 uniformization theorem, 11 uniformly perfect set, 64 univalent mapping, 1 universal covering surface, 10 wandering domain, 70 Wolff, J., 79 Wolff-Denjoy theorem, 79 Yakobson, M.V., 121 Yoccoz, J.-C., 46, 48, 86 Symbol Index f™, m-fold iterate f o---0 f (m times), 27 OE, boundary of E E, closure of E A, open unit disk {|z| < 1} A(z,7r), open disk {|z — z9| < r} C, complex plane CG, extended complex plane C U co H, upper half-plane {Im z > 0} R, real line S, universal covering surface of S, 10 dpp(z), hyperbolic (Poincaré) metric of D, 12 C', continuously differentiable QC(k, R), normalized k-quasiconformal maps analytic off A(0, R), 24 QC1(k, R), C1N QC(k, R), 19 S, normalized univalent functions, 1 F, Fatou set, 54 J, Julia set, 54 K, filled-in Julia set, 65 M, Mandelbrot set, 124 R(6,K), external ray of K, 137 R(0, M), external ray of M, 140 A(z), basin of attraction of zo, 28 A*(zo), immediate basin of attraction of zo, 28 CL, closure of the postcritical set, 81 D,, derivative with respect to o-metric, 119 P., Pe(z) = 2? +c, 123 Fe, Fatou set of z? +¢ Je, Julia set of 27 +¢ K, filled-in Julia set of z? +¢ Universitext (continued) Nikulin/Shafarevich: Geometries and Groups @ksendal: Stochastic Differential Equations Rees: Notes on Geometry Reisel: Elementary Theory of Metric Spaces Rey: Introduction to Robust and Quasi-Robust Statistical Methods Rickart: Natural Function Algebras Rotman: Galois Theory Rybakowski: The Homotopy Index and Partial Differential Equations Samelson: Notes on Lie Algebras Schiff: Normal Families Logical Number Theory I: An Introduction Self-Reference and Modal Logic Stillwell: Geometry of Surfaces Stroock: An Introduction to the Theory of Large Deviations Sunder: An Invitation to von Neumann Algebras Tondeur: Foliations on Riemannian Manifolds Verhulst: Nonlinear Differential Equations and Dynamical Systems Zaanen: Continuity, Integration and Fourier Theory

You might also like