Home Examples of non-Dini domains with large singular sets
Article Open Access

Examples of non-Dini domains with large singular sets

  • Carlos Kenig EMAIL logo and Zihui Zhao
Published/Copyright: April 26, 2023

Abstract

Let u be a nontrivial harmonic function in a domain D R d , which vanishes on an open set of the boundary. In a recent article, we showed that if D is a C 1 -Dini domain, then, within the open set, the singular set of u , defined as { X D ¯ : u ( X ) = 0 = u ( X ) } , has finite ( d 2 ) -dimensional Hausdorff measure. In this article, we show that the assumption of C 1 -Dini domains is sharp, by constructing a large class of non-Dini (but almost Dini) domains whose singular sets have infinite d 2 -measures.

MSC 2010: 35J25; 42B37; 31B35

1 Introduction

We consider the following question, which is inspired by a classical question asked by Bers (see the Introduction in [9]):

(Q) Suppose u is a nontrivial harmonic function in a domain D , and that u = 0 on a relatively open set of the boundary B 2 R ( 0 ) D . How big can the singular set S { X B R ( 0 ) D : u ( X ) = 0 } be ?

When D is a C 1 , 1 domain, Lin proved that S has zero ( d 1 ) -dimensional Hausdorff measure, and that S is a ( d 2 ) -dimensional set, see [11, Theorem 2.3]. Adolfsson et al. [2] (see also Kenig and Wang [5] for an alternative proof) extended the result to convex domains. This was then followed by the works of Adolfsson and Escauriaza [1] and Kukavica and Nyström [10], who proved (using different methods) the result for C 1 -Dini domains (see Definition 2.1). Recently, Tolsa [14] proved that for all C 1 domains (or Lipschitz domains with sufficiently small Lipschitz constant), the set S has zero ( d 1 ) -dimensional Hausdorff measure.

In a recent work, we proved the following theorem:

Theorem 1.1

[9] Let D be a C 1 -Dini domain in R d (see Definition 2.1) with 0 D , and let R > 0 . Suppose u is a nontrivial harmonic function in D B 50 R ( 0 ) , and that u = 0 on D B 50 R ( 0 ) . Then, the singular set

S ( u ) { X D B 50 R ( 0 ) ¯ : u ( X ) = 0 = u ( X ) }

satisfies that S ( u ) B R ( 0 ) is ( d 2 ) -rectifiable, and

d 2 ( S ( u ) B R ( 0 ) ) C < + ,

where the constant C depends on d, R, and (the upper bound of) the frequency function of u centered at 0 with radius 50 R .

In short, the theorem says that when D is a C 1 -Dini domain, the singular set at the interior and boundary, S ( u ) B R ( 0 ) , is ( d 2 ) -rectifiable, and its ( d 2 ) -dimensional Hausdorff measure is finite. We remark that a similar result for convex domains can be found in [12].

It is natural to ask whether such a fine estimate (i.e., d 2 estimate) of the singular set can be extended to more general domains, for example, Lipschitz domains with small constants. In that setting, recall Tolsa showed that S ( u ) D has surface measure zero in B R ( 0 ) (see [14]). The answer is no in general because if the domain is less regular than C 1 -Dini, the gradient of harmonic functions that vanish on an open subset of the boundary may not exist everywhere in that open set, and thus it does not make sense to talk about its d 2 -measure. The goal of this article is to provide counterexamples demonstrating that C 1 -Dini domains are indeed the optimal class of domains for which Theorem 1.1 holds. More precisely, if D is less regular than C 1 -Dini and a harmonic function u vanishes on an open subset of D , we cannot make sense of u at the boundary in general. However, in the special case when u is a nonnegative harmonic function in D that vanishes in D B 2 R ( 0 ) with 0 D , by the comparison principle (see Lemma 2.3) u is comparable to the Green’s function G in D B R ( 0 ) . Hence, for σ -almost every x D B R ( 0 ) (where σ d 1 D denotes the boundary surface measure of D ), we have

u ( x ) n G ( x ) d ω d σ ( x ) ,

where n G denotes the normal derivative of G at the boundary and d ω / d σ denotes the Poisson kernel of the harmonic measure ω (whose pole is the same as that of the Green’s function G ). Since the upper and lower densities of the Radon measure ω are defined everywhere (and take values in [ 0 , + ] ), we can use the following set[1]

p D : liminf r 0 ω ( Δ r ( p ) ) r = limsup r 0 ω ( Δ r ( p ) ) r = 0

in place of the boundary singular set of u , namely { p D : u ( p ) = 0 } . Roughly speaking, we showed that when the domain D barely fails to be C 1 -Dini, the d 2 -measure of the above set could be infinite.

Theorem 1.2

Given a monotone nondecreasing function θ : R + R + , which satisfies

(1.1) lim r 0 + θ ( r ) = 0 and 0 θ ( r ) r d t = + ,

there exist a C 1 function φ : R R and a C 1 domain

D { ( x , t ) R × R : x R , t > φ ( x ) }

such that the following holds:

  • there exists a bounded set S R containing infinitely (countably) many points such that, for each x 0 S , the modulus of continuity of φ at x 0 , denoted by α ( r ) , satisfies

    θ ( r ) α ( r ) θ ( 4 r ) ;

  • φ C 2 ( R S ¯ ) ;

  • let ω denote the harmonic measure in D, we have that

    p D : liminf r 0 ω ( Δ r ( p ) ) r = limsup r 0 ω ( Δ r ( p ) ) r = 0 graph S φ ,

    where graph S φ { x , φ ( x ) : x S } .

In particular, the set

p D : liminf r 0 ω ( Δ r ( p ) ) r = limsup r 0 ω ( Δ r ( p ) ) r = 0

is infinite.

Remark 1.3

  • We can easily extend the above example of planar domains to domains in R d , by considering the domains D ˜ D × R d 2 R d . In this case, the singular set in the boundary of D ˜ is equal to graph S φ × R d 2 and has infinite d 2 -measure.

  • In the proof, we will also show that whenever x 0 and x 0 , the function φ ( x ) is linear, and thus the boundary D = graph R φ is flat when we are sufficiently far from the center ( 0 , φ ( 0 ) ) . Therefore, it is not hard to close off D so that it becomes a bounded domain while maintaining that D has C 2 regularity except for points in graph S φ . Thus, the result also holds for bounded domains.

We remark that when D is a C 1 -Dini domain and the harmonic function u in consideration is nonnegative, the Hopf maximum principle implies that

u = ν u c > 0 at the boundary ,

where ν u denotes the normal derivative of u with the normal vector ν pointing inward, see [3]. (The Hopf maximum principle in [3] was proven for solutions to parabolic equations, but by taking a slice at a fixed positive time, the elliptic analog follows.) In particular, this implies that for C 1 -Dini domains, the singular set

p D : liminf r 0 ω ( Δ r ( p ) ) r = limsup r 0 ω ( Δ r ( p ) ) r = 0 = .

In a related work [4, Section 9], the author constructed the following example (credited to Tolsa): there exist Lipschitz domains D R 2 with small constants such that the singular set at the boundary

p D : lim r 0 ω ( Δ r ( p ) ) r exists and is equal to 0

has Hausdorff dimension as close to 1 as we want. In particular, it indicates that for Lipschitz domains, one cannot expect a better answer to (Q) than saying that the singular set has zero surface measure. (For comparison, our examples show that in order to obtain the sharp ( d 2 ) -dimensional estimate for (Q) as in [9], the assumption of C 1 -Dini domains cannot be weakened.) These Lipschitz domains are built by taking the union of cones with vertices at a fat Cantor set, whose Hausdorff dimension can be chosen sufficiently close to 1. The purpose of their example is similar to ours, but the constructions are completely different. Besides, our example of domains is better than C 1 -regular, instead of just Lipschitz, but the singular set is 0-dimensional (albeit infinite) rather than ( 1 ε ) -dimensional.

The proof of Theorem 1.2 is inspired by the work of the first-named author [6]. It was demonstrated there that one can construct a Lipschitz domain in R 2 with prescribed tangent vectors on its boundary such that its harmonic measure is given by the exponential of the Hilbert transform of that prescribed function, see, e.g., [6, Lemma 1.11]. This article is organized as follows. We recall some definitions and preliminary results in Section 2. Then, to fix ideas, we first construct Lipschitz domains with the desired properties in Section 3. These domains are explicit and not difficult to visualize. In Section 4, we construct the desired C 1 domains for every modulus of continuity θ satisfying (1.1).

2 Preliminaries

Definition 2.1

(Dini domains) Let θ : R R be a nondecreasing function that satisfies lim r 0 + θ ( r ) = 0 and

0 θ ( r ) r < .

A connected domain D in R d is a Dini domain with parameter θ if for each point X 0 D , there is a coordinate system X = ( x , x d ) , x R d 1 , and x d R such that X 0 = ( 0 , 0 ) with respect to this coordinate system, and there is a ball B centered at X 0 and a Lipschitz function φ : R d 1 R verifying the following:

  1. φ L ( R d 1 ) C 0 for some C 0 > 0 ;

  2. φ ( x ) φ ( y ) θ ( x y ) for all x , y R d 1 ;

  3. D B = { ( x , x d ) B : x d > φ ( x ) } .

The following integral will be used repeatedly in the computation of Hilbert transforms, so we state it as a lemma here.

Lemma 2.2

Let a < b be two real numbers. Suppose x [ a , b ] , we have that

(2.1) a b 1 x y d y = log x a log x b .

Proof

When x < a < y , let z y x > 0 . By a change of variables, we have

a b 1 x y d y = a x b x 1 z d z = log ( a x ) log ( b x ) .

When x > b > y , let z = x y > 0 . By a change of variables, we have

a b 1 x y d y = x a x b 1 z d z = x b x a 1 z d z = log ( x a ) log ( x b ) .

We recall the following lemmas about positive solutions to elliptic partial differential equations (for a reference, see [7]).

Lemma 2.3

(Comparison principle) Let D be a Lipschitz domain and p D , r > 0 . Let u , v 0 be two nontrivial harmonic functions in D B 2 r ( p ) such that u = v = 0 on Δ 2 r ( p ) D B 2 r ( p ) . Then, for any X D B r ( p ) ,

C 1 u v ( A r ( p ) ) u v ( X ) C u v ( A r ( p ) ) ,

where C 1 is a universal constant, and A r ( p ) denotes the corkscrew point in D B r ( p ) , that is to say, there exists a constant M > 0 only depending on the Lipschitz constant of the domain such that B r / M ( A r ( p ) ) D B r ( p ) .

Lemma 2.4

Let D be a Lipschitz domain in R d . For any X D , let G ( X , ) denote the Green’s function for the Laplacian in D with pole at X, and let ω X denote the corresponding harmonic measure. For any p D and r > 0 such that X B 2 r ( p ) , we have that

ω X ( Δ r ( p ) ) r d 1 G ( X , A r ( p ) ) r ,

where means that the two quantities are equivalent modulo two universal constants.

3 Lipschitz domains

Let H : R R be the Heaviside step function, namely

H ( x ) = 0 , x 0 1 , x > 0 .

Let { x k } be a sequence of distinct points in R { 0 } such that x k 0 . (Then, in particular, { x k } is bounded, say x k 1 .) Let c be a positive real number and { a k } be a sequence in R + such that

(3.1) c c a k < π 2 .

We define a function f : R R as follows:

f ( x ) = c a k H ( x x k ) .

Clearly,

f L = c a k = c < π 2 .

Let K be the Hilbert transform operator, modulo a constant, defined as follows:

(3.2) K h ( x ) lim ε 0 1 π h ( y ) χ x y ε x y + χ y > 1 y d y ,

whenever the limit on the right-hand side exists. Here, χ E denotes the characteristic function of the set E . Recall that the Hilbert transform maps L ( R ) functions into functions in the bounded mean oscillations (BMO) space. Simple computations show that

K ( H ( x k ) ) ( x ) = K H ( x x k ) = 1 π log x x k ,

and hence, formally, we have

K f ( x ) = c π a k log x x k .

In fact, by the assumption (3.1), we have that

c k a k H ( x x k ) f ( x ) in L ( R ) , as + .

Hence,

(3.3) K f ( x ) = lim + K c k a k H ( x x k ) = lim + c π k a k log x x k ,

where the limit is taken in the BMO space.

The function on the right-hand side of (3.3) is well-defined and continuous in R { 0 } . In fact, assume that x 0 and x x k for every k . Since log is a continuous function in R { 0 } , it is uniformly continuous on compact subset of R { 0 } . Let E be a compact subset of R { 0 } containing x such that E { x k : k N } = . We have that, for any y E ,

log y x k log y , as x k 0 ,

and the convergence is uniform. In particular, there exists k 0 N depending on E such that for any k k 0 and y E , we have that

log y x k log y + 1 .

Hence, for any m k 0 , we have

(3.4) c π k = m a k log y x k c k = m a k ( log y + 1 ) < + .

Therefore,

c π a k log x x k = lim + c π k a k log x x k

is well-defined and continuous in E . Therefore, we have shown that K f ( x ) is well-defined and continuous on R ( { x k } { 0 } ) . In addition, we have

(3.5) K f ( x ) π 2 log ( x + 1 ) , whenever x 2 .

Moreover, near each x k , we have

(3.6) K f ( x ) = c π a k log x x k + lim k 0 + c π k k 0 k k a k log x x k c π a k log x x k + e k ( x ) .

Since x k x k for every k k , and x k 0 x k , we have that

(3.7) δ k inf { x k x k : k N and k k } > 0 .

Then, as long as 0 < x x k < δ k / 2 ,

c π k k a k log x x k = c π k k x x k 1 a k log x x k + c π k k x x k > 1 a k log x x k = c π k k x x k 1 a k log 1 x x k + c π k k x x k > 1 a k log x x k c π k k x x k 1 a k log 2 δ k + c π k k x x k > 1 a k log ( x + 1 ) c π log 2 δ k + log ( x + 1 ) .

In particular, the second term in (3.6), e k ( ) , is bounded near x k , and thus

(3.8) K f ( x ) as x x k .

Let V ( x , t ) and W ( x , t ) be the Poisson integrals of f ( x ) and K f ( x ) , respectively, in the upper half plane R + 2 . (The Poisson integral of K f ( x ) is well-defined because K f ( x ) has logarithmic growth at infinity according to (3.5).) Since f ( x ) is continuous and bounded in R ( { x k } { 0 } ) , by the classical theory[2] for every x R ( { x k } { 0 } ) , we have that V ( z ) f ( x ) as z x . Since K f ( x ) is continuous in  R ( { x k } { 0 } ) , we also have that W ( z ) K f ( x ) as z x for every x R ( { x k } { 0 } ) . Moreover, let

F ( y ) c π k a k log y x k ,

and recall that F K f in BMO ( R ) and pointwise in R ( { x k } { 0 } ) . We claim that for any ( x , t ) R + 2 ,

(3.9) W ( x , t ) = P t K f ( x ) = lim + P t F ( x ) ,

where P t denotes the Poisson kernel in R + 2 with P t ( ξ ) = 1 π t ξ 2 + t 2 . Then, it immediately follows that

(3.10) W ( x , t ) = P t c π a k log x x k = c π a k log ( x , t ) ( x k , 0 ) .

In the second equality, we use the fact that the Poisson integral of log x in R + 2 is just log ( x , t ) . To prove the claim, let

δ 0 inf { ( x , t ) ( x k , 0 ) : k N } > 0 .

We let

(3.11) P t h ( x ) = P t 1 h ( x ) + P t 2 h ( x ) 1 π y x 2 δ 0 t ( x y ) 2 + t 2 h ( y ) d y + 1 π y x < 2 δ 0 t ( x y ) 2 + t 2 h ( y ) d y ,

for any allowable function h on R . For the first term, we have

P t 1 K f ( x ) P t 1 F ( x ) = 1 π y x 2 δ 0 t ( x y ) 2 + t 2 c π k > a k log y x k d y 1 π y x 2 δ 0 t ( x y ) 2 + t 2 c π k > a k log y x k d y 1 π c π k > a k t ( x y ) 2 + t 2 max log 1 δ 0 , log ( y + 1 ) d y ,

which converges to 0 as + . However, since x k 1 and log is square-integrable near the origin, for any E that is a neighborhood of the origin, we have by the Minkowski integral inequality that

K f F L 2 ( E ) = c π k > a k log y x k L 2 ( E ) c π k > a k log y x k L 2 ( E ) c π k > a k 0 ,

as + . Hence,

P t 2 K f ( x ) P t 2 F ( x ) K f F L 2 ( B x + 2 δ 0 ( 0 ) ) 1 π y x 2 δ 0 t ( x y ) 2 + t 2 2 d y 1 / 2 1 π 2 δ 0 1 / 2 K f F L 2 ( B x + 2 δ 0 ( 0 ) ) ,

which also converges to 0 as + . This finishes the proof of the claim (3.9). In particular, (3.10) implies that

W ( x , t ) as ( x , t ) converges to x k ,

with logarithmic decay.

Let z = x + i y , and we define the function

g ( z ) W ( x , y ) + i V ( x , y ) .

One can verify that W and V satisfy the Cauchy-Riemann equations and thus g is analytic in R + 2 . Let G ( z ) exp g ( z ) . Then, G is analytic in R + 2 and has a nontangential limit toward the boundary for almost every x R + 2 , since G is nontangentially bounded at almost every boundary point.

For every z R + 2 , G ( z ) = exp ( W ( x , y ) ) 0 ,

(3.12) for every x R ( { x k } { 0 } ) , G ( x ) = exp ( K f ( x ) ) = x x k c π a k ,

(3.13) as z converges to x k , G ( z ) z ( x k , 0 ) c π a k + ,

and

(3.14) arg G ( z ) = V ( x , y ) f L c < π 2 ,

where we used the maximum principle for the Poisson integral.

Let Φ denote the antiderivative of G in R + 2 . More precisely, for any z R + 2 , let γ z denote any rectifiable curve from i to z and let

Φ ( z ) = γ z G .

This function is well-defined (i.e., independent of the choice of curve) since R + 2 is simply connected. Besides, for any z = ( x , t ) R + 2 , by choosing γ z to be the line segment connecting i to z , we can easily show that

Φ ( z ) = z i 0 1 G ( γ z ( s ) ) d s z i min { t , 1 } c π < + ,

namely Φ ( z ) C . Since Φ ( z ) = G ( z ) 0 , Φ is locally a conformal mapping. We claim that Φ is injective. Assume there are two distinct points z 1 , z 2 R + 2 such that Φ ( z 1 ) = Φ ( z 2 ) . Let γ 0 denote the line segment in R + 2 connecting z 1 to z 2 . More precisely, we consider the parametrization γ 0 ( t ) = z 1 + t ( z 2 z 1 ) with t [ 0 , 1 ] . We have that

γ 0 G = Φ ( z 2 ) Φ ( z 1 ) = 0

and

γ 0 G = ( z 2 z 1 ) 0 1 G ( γ 0 ( t ) ) d t .

Hence, it follows that

0 1 G ( γ 0 ( t ) ) d t = 0 .

In particular, the real part of the above integral also vanishes, i.e.,

(3.15) 0 1 G ( γ 0 ( t ) ) cos ( arg G ( γ 0 ( t ) ) ) d t = 0 .

However, by (3.14), we have that

cos ( arg G ( z ) ) cos c > 0 for all z R + 2 .

Combined with G ( z ) 0 , this is a contradiction with (3.15). Therefore, Φ is injective.

For any z R + 2 and z x k , z 0 , by the properties of the Poisson integrals V and W , it is easy to see that

Φ ( z ) = γ z G is still well-defined ,

and moreover, it is independent of the choice of the curve γ z R + 2 ¯ . Next, we show that Φ ( z ) is well-defined as z x k and z 0 and is independent of the choice of the curve. For fixed k , let z and z be arbitrary points in R + 2 ¯ { 0 } with z , z x k for any k N such that δ max { z ( x k , 0 ) and z ( x k , 0 ) } is sufficiently small. Let γ denote a rectifiable curve in R + 2 ¯ connecting z and z such that γ does not intersect the origin or x k for any k N . Then, by (3.13),

Φ ( z ) Φ ( z ) = γ G γ γ ( t ) ( x k , 0 ) c π a k .

Since c π a k < 1 2 by the assumption (3.1), by carefully choosing the curve γ (e.g., by taking γ to be the union of an arc on B δ ( x k ) and a line segment on a ray from x k ), we can guarantee that

γ γ ( t ) ( x k , 0 ) c π a k 0 as δ 0 .

Therefore, Φ ( z ) is continuous and finite as z x k . To show Φ ( z ) is continuous at the origin, let z = ( x 0 , t 0 ) and z = ( x 0 , t 0 ) be arbitrary points in R + 2 ¯ { 0 } that are sufficiently close to the origin. Let δ max { z , z } , and clearly x 0 , x 0 , t 0 , t 0 < δ . Let t max { t 0 + δ , t 0 + δ } . Let γ 1 denote the vertical line segment between z and ( x 0 , t ) , γ 2 denote the horizontal line segment between ( x 0 , t ) and ( x 0 , t ) , and γ 3 denote the vertical line segment between ( x 0 , t ) and z , each parametrized by unit length. For each i { 1 , 2 , 3 } , we have

γ i G γ i G = γ i exp ( W ( z ) ) = γ i exp c π a k log z ( x k , 0 ) = γ i z ( x k , 0 ) c π a k .

Since we always have that

z ( x k , 0 ) Im z ,

it follows that

γ 1 z ( x k , 0 ) c π a k t 0 t t c π a k t 1 c π a k δ 1 c π a k , γ 2 z ( x k , 0 ) c π a k γ 2 t c π a k = x 0 x 0 t c π a k δ 1 c π a k ,

and the same estimate holds for γ 3 . Therefore,

Φ ( z ) Φ ( z ) = γ 1 γ 2 γ 3 G i = 1 3 γ i G δ 1 c π a k .

Hence, Φ ( z ) is continuous and finite as z 0 . To sum up, we have shown that Φ has a continuous extension to R + 2 ¯ . Using the same argument (3.15) as in the interior case, we can show that Φ is also injective on R + 2 ¯ .

We claim that Φ ( ) = . Thus, in particular, the set D Φ ( R + 2 ) is unbounded, and D = Φ ( R + 2 ) (the boundary in R + 2 , not in the Riemann sphere). Let z j be an arbitrary sequence in R + 2 such that z j . Let γ j denote the straight line segment connecting i to z j , namely γ j ( t ) = i + t ( z j i ) for t [ 0 , 1 ] . Then,

Φ ( z j ) = γ j G = ( z j i ) 0 1 G ( γ j ( t ) ) d t .

Hence,

Re Φ ( z j ) z j i = Re 0 1 G ( γ j ( t ) ) d t = 0 1 G ( γ j ( t ) ) cos ( arg G ( γ j ( t ) ) ) d t cos c 0 1 G ( γ j ( t ) ) d t .

Using (3.10) again, we have

0 1 G ( γ j ( t ) ) d t = 0 1 γ j ( t ) ( x k , 0 ) c π a k d t 0 1 ( γ j ( t ) + 1 ) c π d t ( z j + 1 ) c π .

Therefore,

(3.16) Φ ( z j ) = z j i Φ ( z j ) z j i z j i Re Φ ( z j ) z j i z j 1 c π ,

for j sufficiently large. In particular, Φ ( z j ) for any sequence z j in R + 2 such that z j .

Since Φ : R + 2 D is a conformal homeomorphism, and Φ is injective on R + 2 with D = Φ ( R + 2 ) , it follows that D is also simply connected and bounded by a simple curve. By the same argument as in [6, Theorem 1.1], at every x R + 2 where Φ ( x ) exists and is different from 0, it is a tangent vector to D at the point p Φ ( x ) ; and a set E R + 2 has measure zero if and only if Φ ( E ) D has surface measure zero. We remark that because G ( x ) is continuous in  R ( { x k } { 0 } ) , the fundamental theorem of calculus states that Φ ( x ) = G ( x ) there. Moreover, Let [ x k a , x k + b ] R be an arbitrary interval containing x k , with a and b sufficiently small satisfying 0 < a , b < δ k / 2 (recall the definition of δ k in (3.7)). Recall that k k x x k c π a k is continuous at x k . It follows that by choosing a and b sufficiently small, we can guarantee that

k k x x k c π a k > 1 2 k k x k x k c π a k > 0 for every x [ x k a , x k + b ] .

Hence,

x k a x k + b Φ ( x ) d x = x k a x k + b G ( x ) d x = x k a x k + b x x k c π a k d x x k a x k + b x x k c π a k d x 1 2 k k x k x k c π a k 1 2 k k x k x k c π a k max { a , b } c π a k .

In particular,

(3.17) x k a x k + b Φ ( x ) d x + as a , b 0 + .

Let ω denote the harmonic measure in D with pole at infinity and normalized at Φ ( 0 ) (for its precise definition and properties, see [8, Corollary 3.2 and Lemma 3.8]).[3] By the conformal invariance of the Brownian motion, we can find the explicit formula for ω: we have that

d ω ( z ) = 1 Φ ( Φ 1 ( z ) ) d σ ( z ) ,

where σ denotes the surface measure at the boundary ∂D, namely σ = 1∣∂D . In fact, let ω R + 2 denote the harmonic measure of R + 2 with pole at infinity and normalized at the origin. Clearly, d ω R + 2 = d x . For any z D , let Δ denote a surface ball of D centered at z . Then, for every z D such that Φ 1 ( z ) { x k } { 0 } , as Δ z , we have

(3.18) ω ( Δ ) 1 ( Δ ) = ω R + 2 ( Φ 1 ( Δ ) ) Φ 1 ( Δ ) Φ ( x ) d x = Φ 1 ( Δ ) d x Φ 1 ( Δ ) Φ ( x ) d x 1 Φ ( Φ 1 ( z ) ) = 1 G ( Φ 1 ( z ) ) > 0 .

Here, we use the conformal invariance of the harmonic measure and the area formula in the first equality. On the other hand, when z = Φ ( x k ) for some x k , we claim that

lim Δ z ω ( Δ ) 1 ( Δ ) = 0 .

In fact, since Φ is a homeomorphism on R + 2 , Φ 1 ( Δ ) is just an interval containing Φ 1 ( z ) = x k . By (3.17), it follows that

lim Δ z ω ( Δ ) 1 ( Δ ) = lim a , b 0 + x k a x k + b Φ ( x ) d x 1 = 0 .

In short,

z D : lim Δ z ω ( Δ ) 1 ( Δ ) = 0 { Φ ( 0 ) } = Φ ( { z k : k N } ) .

Finally, we remark that given the input f ( x ) c H ( x ) , where c is a constant with 0 < c < π / 2 and H ( x ) is the Heaviside step function, our construction produces the following simple Lipschitz domain. Intuitively, it is clear that the density of the harmonic measure in D is zero only at the vertex.

4 C 1 domains

The following lemma is just a special case of the more general Lemma 4.2 that we need later. But we introduce and prove this lemma first in order to fix ideas.

Lemma 4.1

There exists a continuous function f C ( R ) such that

  1. f is a monotone nondecreasing function with 0 f 1 and f C 1 ( R { 0 } ) ;

  2. the modulus of continuity of f at the origin, denoted by θ ( r ) , satisfies

    0 θ ( r ) r d r = + ;

  3. K f ( x ) C ( R { 0 } ) , and K f ( x ) K f ( 0 ) = as x 0 , where K denotes the Hilbert transform operator as is defined in (3.2).

Proof

Let θ : ( 0 , 1 ) R + be defined as θ ( r ) = log 2 1 r 1 . Clearly, θ is monotone increasing, lim r 0 + θ ( r ) = 0 , and

0 θ ( r ) r d r = + .

One can check that there exists x 0 = 2 1 log 2 ( 0 , 1 / 2 ) such that in ( 0 , x 0 ] , θ is concave and

(4.1) θ ( x ) = 1 x log 2 log 2 1 x 2 > 0 is monotone decreasing .

Let g ( ) be a smooth, nondecreasing function defined on [ x 0 , 1 / 2 ] such that

g ( x 0 ) = θ ( x 0 ) = log 2 , g 1 2 = θ 1 2 = 1 , g ( x 0 + ) = θ ( x 0 ) = 2 1 log 2 log 2 , g 1 2 = 0 ,

and g ( x ) g = 2 1 log 2 for all x [ x 0 , 1 / 2 ] . Finally, we define f : R R as follows:

f ( x ) = 0 , x 0 , θ ( x ) = log 2 1 x 1 , 0 < x x 0 , g ( x ) , x 0 < x < 1 2 , 1 , x 1 2 .

It is not hard to see that f satisfies (1) and (2). Next, we analyze K f ( x ) . Recall that the Hilbert transform maps bounded continuous functions into functions in the vanishing mean oscillations (VMO) space, so K f ( x ) VMO ( R ) .

When x < 0 , we have that

(4.2) π K f ( x ) = 0 1 / 2 f ( y ) x y d y + log 1 2 x ,

and clearly, K f ( x ) is continuous on ( , 0 ) . A rough estimate (simply using the monotonicity of f ( ) ) provides

(4.3) < ( 1 f ( x 0 ) ) log ( x 0 x ) + f ( x 0 ) log ( x ) π K f ( x ) log 1 2 x < + .

Moreover, we claim that

0 1 / 2 f ( y ) x y d y as x 0 .

Combined with (4.2), the claim implies that

K f ( x ) as x 0 .

In fact, since

(4.4) 0 1 / 2 f ( y ) x y d y = 0 x 0 θ ( y ) y x d y x 0 1 / 2 g ( y ) y x d y

and the second term is uniformly bounded in x , it suffices to show that

0 x 0 θ ( y ) y x d y + as x 0 .

This follows easily from Fatou’s lemma:

+ = 0 x 0 θ ( y ) y d y liminf x 0 0 x 0 θ ( y ) y x d y .

When 0 < x < x 0 , we have that

(4.5) π K f ( x ) = lim ε 0 0 x ε f ( y ) x y d y + x + ε x 0 f ( y ) x y d y + x 0 1 2 f ( y ) x y d y + log 1 2 x .

Note that

(4.6) lim ε 0 0 x ε f ( y ) x y d y + x + ε x 0 f ( y ) x y d y = f ( x ) [ log x log ( x 0 x ) ] + 0 x 0 f ( y ) f ( x ) x y d y ,

when the integral on the right-hand side is well-defined. In order to analyze the last term 0 x 0 f ( y ) f ( x ) x y d y , we break the integral into two regions y [ 0 , x ] and y [ x , x 0 ] . On the one hand, by the mean value theorem and the monotonicity of f ( ) on [ 0 , x 0 ] , we have

(4.7) 0 < x x 0 f ( y ) f ( x ) y x d y sup [ x , x 0 ] f ( x 0 x ) = f ( x ) ( x 0 x ) .

On the other hand,

(4.8) 0 < 0 x f ( y ) f ( x ) y x d y = 0 x / 2 f ( y ) f ( x ) y x d y + x / 2 x f ( y ) f ( x ) y x d y .

Since 0 f f ( x 0 ) on [ 0 , x ] , we can control the first term as follows:

(4.9) 0 < 0 x / 2 f ( y ) f ( x ) y x d y 0 x / 2 f ( x ) x y d y f ( x 0 ) 0 x / 2 1 x y d y = f ( x 0 ) log 2 ;

again, by the mean value theorem and the monotonicity of f ( ) on [ 0 , x 0 ] , we can control the second term as follows:

(4.10) 0 < x / 2 x f ( y ) f ( x ) y x d y x / 2 x sup [ x / 2 , x ] f d y f x 2 x 2 .

Combining (4.7), (4.8), (4.9), and (4.10), we conclude that

(4.11) 0 < 0 x 0 f ( y ) f ( x ) y x d y f ( x ) ( x 0 x ) + f x 2 x 2 + f ( x 0 ) log 2 .

Finally, combining (4.5), (4.6), and (4.11), we obtain

(4.12) π K f ( x ) ( 1 f ( x ) ) log ( x 0 x ) + f ( x ) log x f ( x ) ( x 0 x ) f x 2 x 2 f ( x 0 ) log 2 >

and

(4.13) π K f ( x ) ( 1 f ( x 0 ) ) log 1 2 x + ( f ( x 0 ) f ( x ) ) log ( x 0 x ) + f ( x ) log x < + .

We claim that

K f ( x ) as x 0 + .

By (4.5), (4.6) and the dominated convergence theorem (which implies that lim x 0 + x 0 1 / 2 f ( y ) x y d y = x 0 1 / 2 f ( y ) y d y and is finite), to prove the claim, it suffices to show

f ( x ) log x + 0 x 0 f ( y ) f ( x ) x y d y = f ( x ) log 1 x 0 x 0 f ( y ) f ( x ) y x d y

as x 0 + . This holds because

f ( x ) log 1 x > 0

and

liminf x 0 + 0 x 0 f ( y ) f ( x ) y x d y 0 x 0 θ ( y ) y d y = +

by Fatou’s Lemma (since f ( y ) f ( x ) y x [ 0 , + ] for every y [ 0 , x 0 ] ) and the fact that lim x 0 f ( x ) = f ( 0 ) = 0 . Finally, we also remark that since we have proven that the right-hand side of (4.6), as a principal value, is finite, we can formally write

(4.14) 0 x 0 f ( y ) x y d y f ( x ) [ log x log ( x 0 x ) ] + 0 x 0 f ( y ) f ( x ) x y d y ,

which is well-defined for every 0 < x < x 0 and decays to as x 0 + . (Recall that by (4.2) and (4.4), the decay rate of K f ( x ) as x 0 is also given by 0 x 0 f ( y ) x y d y .)

When x = x 0 , we have

(4.15) π K f ( x 0 ) = lim ε 0 0 x 0 ε f ( y ) x 0 y d y + x 0 + ε 1 / 2 f ( y ) x 0 y d y + log 1 2 x 0 = lim ε 0 0 x 0 ε f ( x 0 ) x 0 y d y + x 0 + ε 1 / 2 f ( x 0 ) x 0 y d y + 0 x 0 f ( y ) f ( x 0 ) x 0 y d y + x 0 1 / 2 f ( y ) f ( x 0 ) x 0 y d y + log 1 2 x 0 = f ( x 0 ) log x 0 + ( 1 f ( x 0 ) ) log 1 2 x 0 + 0 x 0 f ( y ) f ( x 0 ) x 0 y d y + x 0 1 / 2 f ( y ) f ( x 0 ) x 0 y d y .

For the last two terms, note that

0 < x 0 1 / 2 f ( y ) f ( x 0 ) y x 0 d y g 1 2 x 0 ,

0 < 0 x 0 / 2 f ( x 0 ) f ( y ) x 0 y d y 0 x 0 / 2 f ( x 0 ) x 0 y d y = f ( x 0 ) log 2 ,

and by the concavity of f on ( 0 , x 0 ) ,

(4.16) 0 x 0 / 2 x 0 f ( x 0 ) f ( y ) x 0 y d y f x 0 2 x 0 2 .

Therefore,

π K f ( x 0 ) f ( x 0 ) log x 0 2 + ( 1 f ( x 0 ) ) log 1 2 x 0 g 1 2 x 0 f x 0 2 x 0 2 > ,

π K f ( x 0 ) f ( x 0 ) log x 0 + ( 1 f ( x 0 ) ) log 1 2 x 0 < + .

Moreover, by (4.5), (4.6), and (4.15), we also have

(4.17) K f ( x ) K f ( x 0 ) as x x 0 .

When x 0 < x < 1 / 2 , we have

(4.18) π K f ( x ) = 0 x 0 f ( y ) x y d y + lim ε 0 x 0 x ε f ( y ) x y d y + x + ε 1 / 2 f ( y ) x y d y + log 1 2 x = 0 x 0 f ( y ) x y d y + f ( x ) log ( x x 0 ) log 1 2 x + x 0 1 / 2 f ( y ) f ( x ) x y d y + log 1 2 x .

Note that

g 1 2 x 0 x 0 1 / 2 f ( y ) f ( x ) x y d y 0

and

0 0 x 0 f ( y ) x y d y f ( x 0 ) 0 x 0 1 x y d y = f ( x 0 ) [ log x log ( x x 0 ) ] .

Therefore, we have that

(4.19) π K f ( x ) f ( x ) log ( x x 0 ) + ( 1 f ( x ) ) log 1 2 x g 1 2 x 0 >

and

(4.20) π K f ( x ) ( f ( x ) f ( x 0 ) ) log ( x x 0 ) + f ( x 0 ) log x + ( 1 f ( x ) ) log 1 2 x < + .

Moreover, by (4.18) and (4.15), we have that

(4.21) K f ( x ) K f ( x 0 ) as x x 0 + .

When x = 1 / 2 , we have

(4.22) π K f 1 2 = lim ε 0 0 1 / 2 ε f ( y ) 1 / 2 y d y + 1 / 2 1 1 / 2 y χ { y > 1 / 2 + ε } + 1 y d y = 0 x 0 f ( y ) 1 / 2 y d y + x 0 1 / 2 f ( y ) f ( 1 / 2 ) 1 / 2 y d y + log 1 2 x 0 .

Therefore,

(4.23) < log x 0 π K f 1 2 ( 1 f ( x 0 ) ) log 2 + f ( x 0 ) log x 0 < + .

Moreover, by (4.18), (4.22), and the assumption lim x 1 / 2 f ( x ) = f ( 1 / 2 ) = 1 , we have that

K f ( x ) K f 1 2 as x 1 2 .

When x > 1 / 2 ,

(4.24) π K f ( x ) = 0 1 2 f ( y ) x y d y + log x 1 2 .

Hence, we have

(4.25) < log x 1 2 π K f ( x ) log x log x 1 2 + log x 1 2 = log x < + .

Moreover, since by Lemma 2.2,

log x 1 2 log 1 2 x 0 = f 1 2 x x 0 1 1 / 2 y d y ,

and by combining (4.22) and (4.24), we show that

K f ( x ) K f 1 2 as x 1 2 + .

Next, for any nondecreasing function θ satisfying (1.1), we construct a continuous function whose modulus of continuity is given by θ .

Lemma 4.2

Let θ : R + R + be a monotone nondecreasing function such that

lim r 0 + θ ( r ) = 0 and 0 θ ( r ) r d t = + .

Let x 0 > 0 be sufficiently small (depending on θ ). There exists f C ( R ) , defined as in (4.30), which satisfies all the properties in Lemma 4.1, and moreover, the modulus of continuity of f at the origin, denoted by θ ˜ ( r ) , satisfies

(4.26) θ ( r ) θ ˜ ( r ) θ ( 4 r ) .

Proof

Let

(4.27) θ ˜ ( r ) = 1 log 2 2 r 2 r 1 t t 2 t θ ( s ) s d s d t .

Simple computations show that

(4.28) θ ( r ) θ ˜ ( r ) θ ( 4 r ) , lim r 0 + θ ˜ ( r ) = 0 ,

and

d d r θ ˜ ( r ) = 1 log 2 2 1 r 2 r 4 r θ ( s ) s d s r 2 r θ ( s ) s d s = 1 log 2 2 1 2 θ ( 2 r t ) θ ( r t ) r t d t 0 .

Let x be the largest real number such that θ ˜ ( r ) < 1 for all r [ 0 , x ) . (If θ ˜ ( r ) < 1 for all r R + , we simply let x = 1 / 2 .) Then, for any r x / 4 , we have

(4.29) d d r θ ˜ ( r ) 1 log 2 2 1 2 1 r t d t = 1 r log 2 .

Let x 0 ( 0 , x / 4 ) be sufficiently small such that θ ˜ ( x 0 ) < 1 / 2 (other than this constraint, we are free to choose x 0 as small as needed). Let g be a smooth, nondecreasing function defined on [ x 0 , x ] such that

g ( x 0 ) = θ ˜ ( x 0 ) , g ( x ) = 1 , g ( x 0 + ) = d d r θ ˜ ( x 0 ) , g ( x ) = 0 ,

and g ( r ) g . We define the function f : R R as follows:

(4.30) f ( x ) = 0 , x 0 θ ˜ ( x ) , 0 < x x 0 g ( x ) , x 0 < x < x 1 , x x

Clearly, f ( x ) satisfies (1) and (2) of Lemma 4.1, (4.26), and K f ( x ) V MO ( R ) since f is a bounded continuous function on R .

In the proof of property (3) in Lemma 4.1, we use the fact that on [ 0 , x 0 ] , the function f is monotone nondecreasing, differentiable except at the origin and concave. In the general case, here f ( x ) = θ ˜ ( x ) may not be concave in [ 0 , x 0 ] . However, after going over the estimation of K f ( x ) when 0 < x x 0 , if f is not assumed to be concave similar estimates hold once we make the following changes: in (4.7) replace f ( x ) by sup [ x , x 0 ] f , in (4.10) replace f ( x / 2 ) by sup [ x / 2 , x ] f , replace these terms accordingly in the lower bound (4.12), and replace f ( x 0 / 2 ) in (4.16) by sup [ x 0 / 2 , x 0 ] f . The rest of the proof is exactly the same as in Lemma 4.1.□

From now on, we always denote the function in Lemma 4.2 (see (4.30)) as H ˜ ( ) , which will play the same role as the Heaviside function in Section 3. (In the construction of the function in Lemma 4.2, we choose x 0 > 0 sufficiently small, depending on θ , so that (4.59) holds.)

As in Section 3, we construct a new function f as follows. Let c be a positive real number, and { a k } be a sequence in R + such that

(4.31) c c a k < π 2 .

Using the sequence { x k 2 k } [4], we define a function f : R R as follows:

(4.32) f ( x ) = c a k H ˜ ( x x k ) .

Clearly, f C ( R ) , f C 1 ( R ( { x k } { 0 } ) ) , and

f L c a k = c < π 2 .

Moreover, we can prove the following lemma:

Lemma 4.3

K f ( x ) is well-defined and continuous in R ( { x k } { 0 } ) . Near each x k , we have that

(4.33) K f ( x ) c a k K H ˜ ( x x k ) = c k k a k K H ˜ ( x x k ) ,

where the right-hand side is continuous and bounded (the bound only depends on δ k in (4.36)). In particular,

K f ( x ) as x x k .

Proof

Since

c k a k H ˜ ( x x k ) f ( x ) in L ( R )

and f C b ( R ) , we have that

(4.34) K f ( x ) = lim + K c k a k H ˜ ( x x k ) = lim + c k a k K H ˜ ( x x k ) ,

where the limit is taken in the BMO space and K f V MO ( R ) . On R { x k } , for each k , the function K H ˜ ( x k ) is pointwise defined. We claim that the limit on the right-hand side of (4.34) is well-defined and gives a continuous function on R ( { x k } { 0 } ) .

Let x R { 0 } be an arbitrary point. (If x = x k for any k , we just remove the k -th term and consider the summation k k , so we also have that x x k for every k . See (4.35).) Recall that K H ˜ is a continuous function in R { 0 } , it is uniformly continuous on compact subsets of R { 0 } . Let E be a compact subset of R { 0 } containing x , such that E { x k : k N } = . We have that for any y E ,

K H ˜ ( y x k ) K H ˜ ( y ) , as x k 0 ,

and the convergence is uniform. In particular, there exists k 0 N depending on E such that for any k k 0 and y E , we have that

K H ˜ ( y x k ) K H ˜ ( y ) + 1 .

Hence, for any m k 0 , we have

c k = m a k K H ˜ ( y x k ) c k = m a k ( K H ˜ ( y ) + 1 ) < + .

Therefore, as an absolutely convergent series of continuous functions,

c a k K H ˜ ( y x k ) = lim + c k a k K H ˜ ( y x k )

is well-defined and continuous at x .

Moreover, near each x k , we have

(4.35) K f ( x ) = c a k K H ˜ ( x x k ) + lim + c k k k a k K H ˜ ( x x k ) .

Since x k x k for every k k , we have that

(4.36) δ k inf { x k x k : k N and k k } > 0 .

Then, as long as x x k < δ k / 2 , since

x x k x k x k x x k > δ k / 2 ,

by the estimates of K H ˜ away from the origin in Lemmas 4.1 and 4.2 as well as the assumption (4.31), we can show that

K f ( x ) c a k K H ˜ ( x x k ) = lim + c k k k a k K H ˜ ( x x k )

is well-defined and continuous at x k , as an absolutely convergent series of continuous functions. In particular,

as x x k , the limit of K f ( x ) c a k K H ˜ ( x x k ) exists and is finite .

Therefore, K f ( x ) as x x k for every k N .□

Proof of Theorem 1.2

We can construct a Lipschitz domain D = Φ ( R + 2 ) using f and K f as in Section 3. (Again because K H ˜ has logarithmic growth at infinity, by (4.3), (4.25), and the analogous estimates in Lemma 4.2, the Poisson integral of K f is well-defined.) Since f C b ( R ) , its Poisson integral V ( z ) converges to f ( x ) for every x R + 2 ; the Poisson integral W ( z ) of K f ( x ) converges to K f ( x ) for every x R ( { x k } { 0 } ) , as in the paragraph after (3.8).

Moreover, we claim that for every ( x , t ) R + 2 ,

(4.37) W ( x , t ) = P t K f ( x ) = lim + P t c k a k K H ˜ ( x x k ) = lim + c k a k P t K H ˜ ( x x k ) .

(In the last equality, we simply use the linearity of the Poisson integral operator.) The proof is by studying the Poisson integral in the regions close to x and away from x , similar to the proof of the analogous claim (3.9) in Section 3. So we only sketch the key steps here. Let

δ 0 inf { ( x , t ) ( x k , 0 ) : k N } > 0 .

For the Poisson integral on the region that is 2 δ 0 -away from x (i.e., the P t 1 term in (3.11)), we use the lower and upper bounds of K H ˜ proven in Lemma 4.1, the continuity of K H ˜ at x 0 and 1/2, combined with (4.29), to obtain that

K H ˜ ( y ) log ( y + 1 ) + log 1 δ 0 + log 1 1 2 x 0 + log 1 x 0 + 1 δ 0 + K f ( x 0 ) + K f 1 2

for any y R with y > δ 0 . On the other hand, by the estimates (4.3), (4.12), and (4.13), we have that K H ˜ is square-integrable near the origin. This can be used to estimate the Poisson integral on the region that is 2 δ 0 -close to x (i.e., the P t 2 term in (3.11)). This finishes the proof of (4.37).

In particular, as z = x + i t converges to x k , by Lemma 4.3 and the property of the Poisson integral for bounded and continuous functions, we have that

W ( x , t ) c a k P t ( K H ˜ ( x x k ) ) is continuous and bounded .

Combined with the estimates of K H ˜ near the origin (see the estimates (4.2), (4.5), and (4.6), as well as the definition in (4.14)), we have that

W ( x , t ) + c a k π P t 0 x 0 θ ˜ ( y ) y d y ( x x k ) is continuous and bounded near x k .

Recall that we have shown that the function x 0 x 0 θ ˜ ( y ) y x d y is well-defined as a principal value and continuous in R { 0 } (in particular, recall that we have proven its continuity at x 0 in the proof of Lemma 4.1). So in order to estimate the Poisson integral

P t 0 x 0 θ ˜ ( y ) y d y ( x )

near the origin, let us focus on 0 x 0 θ ˜ ( y ) y x d y near the origin. When x < 0 (and x < ε 0 for some sufficiently small ε 0 ), we have that

(4.38) 0 < 0 x 0 θ ˜ ( y ) y x d y 0 x 0 1 y x d y = log x + log x 0 x < log 1 x .

When x > 0 (and x < ε 0 ), the estimate (4.11) is not enough for our purpose; instead, we claim that

(4.39) 0 < 0 x 0 θ ˜ ( y ) θ ˜ ( x ) y x d y C + 1 2 log 1 x .

In fact, it easily follows from (4.9), (4.10), and (4.29) that

(4.40) 0 < 0 x θ ˜ ( y ) θ ˜ ( x ) y x d y C 1 < + .

However, if x is sufficiently small, we have 2 x < x 0 , and hence

(4.41) 0 < x x 0 θ ˜ ( y ) θ ˜ ( x ) y x d y = x 2 x θ ˜ ( y ) θ ˜ ( x ) y x d y + 2 x x 0 θ ˜ ( y ) θ ˜ ( x ) y x d y x 2 x sup [ x , 2 x ] θ ˜ d y + 2 x x 0 θ ˜ ( y ) y x d y C 2 + θ ˜ ( x 0 ) [ log x + log x 0 x ] C 2 + 1 2 log 1 x .

The claim (4.39) then follows by combining (4.40) and (4.41). Since

0 θ ˜ ( x ) ( log x 0 x log x ) 1 2 log 1 x ,

it follows from (4.14) that

(4.42) 0 < 0 x 0 θ ˜ ( y ) y x d y = 0 x 0 θ ˜ ( y ) θ ˜ ( x ) y x d y + θ ˜ ( x ) ( log x 0 x log x ) C + log 1 x .

Combining (4.38) and (4.42), we have that

0 P t 1 < ε 0 0 x 0 θ ˜ ( y ) y d y ( x ) C + P t 1 < ε 0 log 1 ( x ) .

Note that

P t 1 < ε 0 log 1 ( x ) = P t log 1 ( x ) P t 1 ε 0 log 1 ( x ) = log 1 ( x , t ) P t 1 ε 0 log 1 ( x ) ;

and P t 1 ε 0 log 1 ( x ) is bounded when ( x , t ) is sufficiently close to the origin, since 1 ε 0 log 1 is continuous at the origin and thus

P t 1 ε 0 log 1 ( x ) 0 as ( x , t ) 0 .

Therefore, when ( x , t ) is close to the origin,

P t 0 x 0 θ ˜ ( y ) y ( x ) d y = P t 1 < ε 0 0 x 0 θ ˜ ( y ) y d y ( x ) + P t 1 ε 0 0 x 0 θ ˜ ( y ) y d y ( x ) P t 1 < ε 0 0 x 0 θ ˜ ( y ) y d y ( x ) + C 3 log 1 ( x , t ) + C .

Hence, we can use the same argument as in Section 3 to show that, for any z , z sufficiently close to x k ,

Φ ( z ) Φ ( z ) γ z , z G ( ω ) = γ z , z exp ( W ( y , s ) ) 0 as z , z x k .

That is to say, Φ ( z ) is continuous at x k .

Next, we show that Φ ( z ) is also continuous at the origin. To that end, we need to estimate W ( x , t ) near the origin. Since x k 0 , there exists k 0 N such that x k < ε 0 / 2 for every k k 0 . Then,

W ( x , t ) = c a k P t K H ˜ ( x x k ) = k < k 0 c a k P t K H ˜ ( x x k ) k k 0 c a k P t K H ˜ ( x x k ) .

As before,

k k 0 c a k P t K H ˜ ( x x k ) 1 + k k 0 c a k π P t 0 x 0 θ ˜ ( y ) y d y ( x x k ) 1 + k k 0 c a k π P t 1 < ε 0 / 2 0 x 0 θ ˜ ( y ) y d y ( x x k ) C + k k 0 c a k π log 1 ( x , t ) ( x k , 0 ) .

The constants in the inequality only depends on c = c a k and the constants C 1 and C 2 above; in particular, they are independent of ( x , t ) for ( x , t ) sufficiently close to the origin. Again we can use the same argument as in Section 3 to show that, for any z , z sufficiently close to the origin,

Φ ( z ) Φ ( z ) γ z , z G ( ω ) = γ z , z exp ( W ( y , s ) ) 0 , as z , z 0 .

That is to say, Φ ( z ) is continuous at the origin. To sum up, Φ : R + 2 D extends continuously to R + 2 ¯ D ¯ . Moreover, by the same argument as in Section 3, it is a homeomorphism.

Note that in proving Φ ( ) = , we no longer have an explicit formula for K f ( x ) to estimate the growth of the Poisson integral W ( z ) of K f ( x ) at infinity. However, by (4.3) and (4.25), we still have that K H ˜ ( x ) grows like log x whenever x 0 and x 1 / 2 . More precisely, we have that

(4.43) K H ˜ ( x ) = K H ( x ) + h ( x ) = 1 π log x + h ( x ) ,

where h ( x ) is a bounded and continuous function away from the origin. Near the origin, h ( x ) can be written as h 0 ( x ) + h + ( x ) , with h 0 being a continuous and bounded function, and

h + ( x ) = 1 π 0 x 0 θ ˜ ( y ) x y d y 1 π log x = 1 π log 1 x 1 π 0 x 0 θ ˜ ( y ) y x d y .

Recall that we have shown before that

0 < 0 x 0 θ ˜ ( y ) y x d y log 1 x + C , whenever x < ε 0 .

Thus,

C π h + ( x ) < 1 π log 1 x .

Combining the above, we have that

P t h ( x ) C + P t ( 1 < ε 0 h + ( ) ) ( x ) ,

where

P t ( 1 < ε 0 h + ( ) ) ( x ) = 1 π y < ε 0 t ( x y ) 2 + t 2 h + ( y ) d y y < ε 0 t ( x y ) 2 + t 2 2 d y 1 / 2 log x L 2 ( [ ε 0 , ε 0 ] ) . 2 ε 0 t log x L 2 ( [ ε 0 , ε 0 ] ) .

If x 2 ε 0 , we have

y < ε 0 t ( x y ) 2 + t 2 2 d y 1 / 2 2 ε 0 min t x 2 , 1 t ;

if x < 2 ε 0 , we have

y < ε 0 t ( x y ) 2 + t 2 2 d y 1 / 2 2 ε 0 t .

In particular, whenever ( x , t ) > 3 ε 0 , there is a uniform lower bound for P t h ( x ) (depending only on ε 0 ). Combining (4.37) and (4.43), we obtain the following lower bound for ( x , t ) R + 2 with ( x , t ) > 3 ε 0 :

G ( x , t ) = exp ( W ( x , t ) ) = exp c a k P t K H ˜ ( x x k ) = exp c π a k log ( x , t ) ( x k , 0 ) exp c a k P t h ( x x k ) = ( x , t ) ( x k , 0 ) c π a k exp c a k P t h ( x x k ) ( ( x , t ) + 1 ) c π ,

where the constant depends on the uniform lower bound of P t h ( x ) and c = c a k . Therefore, by the same argument as in (3.16), we have that Φ ( ) = . In particular, D = Φ ( R + 2 ) , where D denotes the topological boundary of D in R 2 , not the boundary in the Riemann sphere.

As in Section 3, we know that G ( x ) is continuous on R ( { x k } { 0 } ) , and thus

Φ ( x ) exists and is equal to G ( x ) .

Next, we will show that exp ( i f ( x ) ) is a unit tangent vector field to D at p = Φ ( x ) for every x R (including when x = x k and x = 0 ). Thus, the property f C b ( R ) C 1 ( R ( { x k } { 0 } ) ) implies that D is C 1 -regular everywhere, and it is also C 2 regular everywhere except at the countably infinite set { Φ ( x k ) : k N } { Φ ( 0 ) } . (At each x k , the modulus of continuity for f ( x ) is comparable to θ ( ) , which fails the Dini condition.) Recall that we relabel the function defined in (4.30) of Lemma 4.2 as H ˜ . Hence, by the definition of f ( x ) via H ˜ in (4.32), it is clear that

f ( x ) = 0 when x < 0 ,

and

f ( x ) = c a k H ˜ ( x x k ) = c a k = c when x 2 .

Therefore, D is flat on { Φ ( x ) : x < 0 } and { Φ ( x ) : x 2 } .

We claim that, for each k , Φ ( x k ) = + and arg Φ ( x k ) = f ( x k ) , in the following sense:

(4.44) Φ ( x k + ε ) Φ ( x k ) ε + and arg Φ ( x k + ε ) Φ ( x k ) ε f ( x k ) as ε 0 .

In the above notation, we take the principal branch of the argument function (in fact, the argument of G ( z ) always lies in ( π / 2 , π / 2 ) , by the bound on f L ). Recall that Φ extends continuously to the boundary, and on R ( { x k } { 0 } ) , we have

G ( x ) = exp ( K f ( x ) ) exp ( i f ( x ) ) .

Therefore,

Φ ( x k + ε ) Φ ( x k ) ε = 1 ε x k x k + ε G ( x ) = 1 ε x k x k + ε exp ( K f ( x ) ) exp ( i f ( x ) ) .

By (4.33), when ε is sufficiently small, we have

(4.45) Φ ( x k + ε ) Φ ( x k ) ε = 1 ε x k x k + ε exp ( K f ( x ) ) exp ( i f ( x ) ) = 1 ε x k x k + ε exp ( c a k K H ˜ ( x x k ) ) exp c k k a k K H ˜ ( x x k ) exp ( i f ( x ) ) = exp ( i f ( x k ) ) exp c k k a k K H ˜ ( x k x k ) 1 ε x k x k + ε exp ( c a k K H ˜ ( x x k ) ) exp ( ρ ( x x k ) ) exp ( i α ( x x k ) ) ,

where we denote the real-valued functions

ρ ( τ ) c k k a k K H ˜ ( τ + x k x k ) + c k k a k K H ˜ ( x k x k ) , α ( τ ) f ( τ + x k ) f ( x k ) .

By the continuity of the functions k k a k K H ˜ ( x x k ) and f ( x ) at x k , we have that

(4.46) ρ ( τ ) , α ( τ ) 0 as τ 0 .

We will show that exp ( c a k K H ˜ ( x ) ) is integrable near the origin[5], and moreover, as ε 0 ( ε can be negative),

(4.47) 1 ε 0 ε exp ( c a k K H ˜ ( x ) ) d x + .

Once that is proven, by (4.46) and by considering the real and complex parts separately, it follows easily that

(4.48) 1 ε x k x k + ε exp ( c a k K H ˜ ( x x k ) ) exp ( ρ ( x x k ) ) exp ( i α ( x x k ) ) 1 ε x k x k + ε exp ( c a k K H ˜ ( x x k ) ) = x k x k + ε exp ( c a k K H ˜ ( x x k ) ) exp ( ρ ( x x k ) ) exp ( i α ( x x k ) ) x k x k + ε exp ( c a k K H ˜ ( x x k ) ) 1 as ε 0 .

Therefore, combining (4.45), (4.48), and (4.47), we conclude the proof of (4.44).

We first consider the case when ε < 0 . Assume, without loss of generality, that ε < x 0 / 2 . By (4.2) and the definition of H ˜ ( x ) in (4.30), for x < 0 , we have[6]

(4.49) π K H ˜ ( x ) = 0 x 0 θ ˜ ( y ) y x d y + x 0 1 / 2 g ( y ) y x d y log 1 2 x 0 x 0 θ ˜ ( y ) y x d y + 1 .

Let k 0 , N 0 , and be the natural numbers such that

(4.50) 2 k 0 x 0 < 2 k 0 + 1 , 2 N 0 + 1 < ε 2 N 0 , and 2 + 1 < x 2 .

The assumption that x ε < x 0 guarantees that N 0 > k 0 . We have[7]

(4.51) 0 x 0 θ ˜ ( y ) y x d y 0 2 k 0 + 1 θ ˜ ( y ) y x d y = i = k 0 2 i 2 i + 1 θ ˜ ( y ) y x d y i = k 0 θ ˜ ( 2 i + 1 ) 2 i 2 i + 2 = i = k 0 θ ˜ ( 2 i + 1 ) 1 1 + 2 i = i = k 0 θ ˜ ( 2 i + 1 ) 1 1 + 2 i + i = + 1 θ ˜ ( 2 i + 1 ) 1 1 + 2 i i = k 0 θ ˜ ( 2 i + 1 ) + j = 1 θ ˜ ( 2 j + 1 ) 1 1 + 2 j i = k 0 + 1 θ ˜ ( 2 i + 1 ) .

Hence, (4.49) implies that

(4.52) exp ( c a k K H ˜ ( x ) ) exp c a k π i = k 0 + 1 θ ˜ ( 2 i + 1 ) .

Therefore,

(4.53) ε 0 exp ( c a k K H ˜ ( x ) ) d x 2 N 0 + 1 0 exp ( c a k K H ˜ ( x ) ) d x = N 0 2 + 1 2 exp c a k π i = k 0 + 1 θ ˜ ( 2 i + 1 ) d x = = N 0 exp c a k π i = k 0 + 1 θ ˜ ( 2 i + 1 ) 2 a k , x 0 = N 0 exp c a k π i = k 0 + 1 θ ˜ ( 2 i + 1 ) π c a k log 2 .

Since c a k < π / 2 , we can choose x 0 so that for β ( 0 , 1 ) fixed,

(4.54) θ ( 8 x 0 ) 2 ( 1 β ) log 2 < ( 1 β ) π c a k log 2 for every k .

Thus, for every i k 0 , we have that

θ ˜ ( 2 i + 1 ) θ ˜ ( 2 k 0 + 1 ) θ ˜ ( 2 x 0 ) θ ( 8 x 0 ) ( 1 β ) π c a k log 2 ,

and thus,

ε 0 exp ( c a k K H ˜ ( x ) ) d x = N 0 exp ( β log 2 ) 2 β N 0 ε β < + .

In particular, exp ( c a k K H ˜ ( x ) ) is integrable on [ ε , 0 ] . On the other hand, as in (4.51), we can also provide a lower bound:

(4.55) 0 x 0 θ ˜ ( y ) y x d y 1 4 i = k 0 + 1 θ ˜ ( 2 i ) .

Hence,

ε 0 exp ( c a k K H ˜ ( x ) ) d x 2 N 0 + 1 0 exp ( c a k K H ˜ ( x ) ) d x = N 0 + 1 exp c a k 4 π i = k 0 + 1 θ ˜ ( 2 i ) 2 = exp c a k 4 π i = k 0 + 1 N 0 θ ˜ ( 2 i ) = N 0 + 1 exp c a k 4 π i = N 0 + 1 θ ˜ ( 2 i ) 2 2 N 0 exp c a k 4 π i = k 0 + 1 N 0 θ ˜ ( 2 i ) .

Since ε 2 N 0 , it follows that

1 ε ε 0 exp ( c a k K H ˜ ( x ) ) d x exp c a k 4 π i = k 0 + 1 N 0 θ ˜ ( 2 i ) .

Recall that

i = k 0 + 1 N 0 θ ˜ ( 2 i ) i = k 0 + 1 N 0 θ ( 2 i ) i = k 0 + 1 N 0 2 i 1 2 i θ ( r ) r d r = 2 N 0 1 2 k 0 1 θ ( r ) r d r ε / 2 x 0 / 4 θ ( r ) r d r + ,

as ε 0 , it follows that

1 ε ε 0 exp ( c a k K H ˜ ( x ) ) d x + , as ε 0 .

This finishes the proof of (4.47) for the case ε < 0 .

We next consider the case ε > 0 . Assume, without loss of generality, that 0 < ε < x 0 / 4 . As in the previous case, let k 0 , N 0 , and be the natural numbers such that

(4.56) 2 k 0 x 0 < 2 k 0 + 1 , 2 N 0 ε < 2 N 0 + 1 , 2 x < 2 + 1 .

By (4.5) and (4.6), we have that[8]

(4.57) π K H ˜ ( x ) 1 + 0 x 0 θ ˜ ( y ) θ ( x ) ˜ y x d y θ ( x ) ˜ ( log x log ( x 0 x ) ) 1 + 0 x 0 θ ˜ ( y ) θ ( x ) ˜ y x d y + θ ˜ ( x ) log 1 x .

By the estimate (4.29) and the monotonicity of θ ˜ ( ) , we have

0 x θ ˜ ( y ) θ ( x ) ˜ y x d y = 0 x / 4 θ ˜ ( x ) θ ( y ) ˜ x y d y + x / 4 x θ ˜ ( x ) θ ( y ) ˜ x y d y 0 2 1 θ ˜ ( x ) θ ( y ) ˜ x y d y + sup [ x / 4 , x ] θ ˜ x 1 + i = + 2 2 i 2 i + 1 θ ˜ ( 2 + 1 ) 2 2 i + 1 d y = 1 + i = + 2 θ ˜ ( 2 + 1 ) 2 i 2 1 + θ ˜ ( 2 + 1 )

and

x x 0 θ ˜ ( y ) θ ( x ) ˜ y x d y = x 4 x θ ˜ ( y ) θ ( x ) ˜ y x d y + 4 x x 0 θ ˜ ( y ) θ ( x ) ˜ y x d y sup [ x , 4 x ] θ ˜ 3 x + 4 x x 0 θ ˜ ( y ) y x d y 1 x 3 x + 2 + 2 2 k 0 + 1 θ ˜ ( y ) y x d y = 3 + i = k 0 2 2 i 2 i + 1 θ ˜ ( y ) y x d y 1 + i = k 0 2 θ ˜ ( 2 i + 1 ) 2 i 2 i 2 + 1 1 + i = k 0 2 θ ˜ ( 2 i + 1 ) .

Hence,

(4.58) 0 x 0 θ ˜ ( y ) θ ( x ) ˜ y x d y 1 + i = k 0 θ ˜ ( 2 i + 1 ) .

Note that (4.58) is similar to the estimate in (4.51) (modulo adding a constant). Hence, by a similar argument to (4.53), we can show that

0 ε exp ( c a k K H ˜ ( x ) ) d x = N 0 2 1 c a k π exp c a k π i = k 0 θ ˜ ( 2 i + 1 ) = N 0 exp c a k π i = k 0 θ ˜ ( 2 i + 1 ) 1 c a k π log 2 .

As in (4.54), for any β ( 0 , 1 ) fixed, by choosing x 0 smaller if necessary, we can guarantee that

(4.59) θ ( 8 x 0 ) ( 1 β ) log 2 < ( 1 β ) 1 c a k π c a k π log 2 , for every k .

Therefore, it follows that

(4.60) 0 ε exp ( c a k K H ˜ ( x ) ) d x = N 0 exp β 1 c a k π log 2 ε β < + ,

where 0 < β < β .

Next, we want to show that

1 ε 0 ε exp ( c a k K H ˜ ( x ) ) d x 1 ε 0 ε exp c a k π 2 x x 0 θ ( y ) ˜ θ ( x ) ˜ y x d y d x +

as ε 0 + . By (4.56) and the monotonicity of θ ˜ ( ) , we have

(4.61) 2 x x 0 θ ( y ) ˜ θ ( x ) ˜ y x d y 2 + 2 2 k 0 θ ( y ) ˜ θ ( x ) ˜ y x d y = i = k 0 1 2 2 i 2 i + 1 θ ( y ) ˜ θ ( x ) ˜ y x d y i = k 0 1 2 θ ˜ ( 2 i ) θ ˜ ( 2 + 1 ) 2 i + 1 2 2 i 1 2 i = k 0 1 2 [ θ ˜ ( 2 i ) θ ˜ ( 2 + 1 ) ] .

Since θ ˜ ( ) is an increasing function, we remark that i = k 0 1 2 [ θ ˜ ( 2 i ) θ ˜ ( 2 + 1 ) ] increases as increases. Hence,

(4.62) 1 ε 0 ε exp c a k π 2 x x 0 θ ( y ) ˜ θ ( x ) ˜ y x d y d x 1 ε 0 2 N 0 exp c a k π 2 x x 0 θ ( y ) ˜ θ ( x ) ˜ y x d y d x 1 ε = N 0 1 2 2 + 1 exp c a k 2 π i = k 0 1 2 [ θ ˜ ( 2 i ) θ ˜ ( 2 + 1 ) ] d x exp c a k 2 π i = k 0 1 N 0 3 [ θ ˜ ( 2 i ) θ ˜ ( 2 N 0 + 2 ) ] 1 ε = N 0 1 2 exp c a k 2 π i = k 0 1 N 0 3 [ θ ˜ ( 2 i ) θ ˜ ( 2 N 0 + 2 ) ] .

Moreover,

(4.63) i = k 0 1 N 0 3 [ θ ˜ ( 2 i ) θ ˜ ( 2 N 0 + 2 ) ] i = k 0 1 N 0 3 2 i + 1 2 i θ ( x ) ˜ θ ˜ ( 2 N 0 + 2 ) x d x = 2 N 0 + 2 2 k 0 + 1 θ ( x ) ˜ θ ˜ ( 4 ε ) x d x 4 ε x 0 θ ( x ) ˜ θ ˜ ( 4 ε ) x d x .

For each ε > 0 , we denote the positive-valued function h ε ( x ) as follows:

h ε ( x ) θ ( x ) ˜ θ ˜ ( 4 ε ) x χ { x 4 ε } .

Then, for each x > 0 , we have

h ε ( x ) θ ( x ) ˜ x , as ε 0 + .

Hence, Fatou’s lemma implies that

(4.64) liminf ε 0 + 0 x 0 h ε ( x ) d x 0 x 0 θ ˜ ( x ) x d x 0 x 0 θ ( x ) x d x = + .

Combining (4.62), (4.63), and (4.64), we conclude that

1 ε 0 ε exp c a k π 2 x x 0 θ ( y ) ˜ θ ( x ) ˜ y x d y d x + as ε 0 + .

To complete the proof that D is C 1 -regular, we also claim that

(4.65) arg Φ ( ε ) Φ ( 0 ) ε f ( 0 ) = 0 as ε 0 .

By the definition of Φ , we have

(4.66) Φ ( ε ) Φ ( 0 ) ε = 1 ε 0 ε G ( x ) = 1 ε 0 ε exp ( K f ( x ) ) exp ( i f ( x ) ) = exp ( i f ( 0 ) ) 1 ε 0 ε exp ( K f ( x ) ) exp ( i α ( x ) ) ,

where α ( x ) f ( x ) f ( 0 ) 0 as x 0 . As in (4.48), if exp ( K f ( x ) ) is integrable on [ 0 , ε ] , then

0 ε exp ( K f ( x ) ) exp ( i α ( x ) ) 0 ε exp ( K f ( x ) ) 1 as ε 0 .

Combined with (4.66), this implies (4.65). Hence, it suffices to show that

(4.67) exp ( K f ( x ) ) = exp c k a k K H ˜ ( x x k ) is integrable on [ 0 , ε ] .

For x 1 , combining the estimates of K H ˜ ( x ) in (4.2), (4.38) (when x < 0 ), (4.5), (4.6), and (4.42) (when x > 0 ), we have that

π K H ˜ ( x ) C + log 1 x .

Hence,

exp c a k K H ˜ ( x x k ) exp C c a k π exp c a k π log 1 x x k x x k c a k π ,

where the constant only depends on the upper bound of c . Therefore, to prove (4.67), it suffices to show that x x k c a k π is integrable near the origin. The latter is indeed the case when we choose x k = 2 k , and we postpone its proof to the appendix. Therefore, the claim (4.65) is proven.

Finally, let ω denote the harmonic measure of D with pole at infinity. As in Section 3, we have that

d ω ( z ) = 1 Φ ( Φ 1 ( z ) ) d z ,

and moreover,

d ω d σ ( Φ ( x k ) ) = lim Δ Φ ( x k ) ω ( Δ ) 1 ( Δ ) = 0 for every k .

Therefore,

z D : lim Δ z ω ( Δ ) 1 ( Δ ) exists and is equal to 0 { Φ ( 0 ) } = Φ ( { x k : k N } ) .

We remark that since D is C 1 -regular, it is, in particular, Ahlfors regular, i.e., there are uniform constants 0 < C 1 C 2 such that

C 1 r 1 ( Δ r ( z ) ) C 2 r for every z D and r > 0 .

Hence,

lim Δ z ω ( Δ ) 1 ( Δ ) exists and is equal to 0 lim r 0 ω ( Δ r ( z ) ) r exists and is equal to 0 .

This finishes the proof of Theorem 1.2 for the harmonic measure with pole at infinity.

Let ω (resp. ω ) denote the harmonic measure in D with pole at X D (resp. with pole at ), and let G ( X , ) (resp. G ( , ) ) denote the Green’s function of the Laplacian in D with pole at X (resp. with pole at ). By applying the comparison principle in Lemma 2.3 to G ( X , ) and G ( , ) , we have that

G ( X , ) G ( , ) ,

as long as we are dist ( X , D ) / 2 -close to the boundary. By Lemma 2.4, it follows that

ω ( Δ r ( p ) ) ω ( Δ r ( p ) ) ,

as long as X B 2 r ( p ) . Therefore,

p D : lim r 0 ω ( Δ r ( p ) ) r exists and is equal to 0 = p D : lim r 0 ω ( Δ r ( p ) ) r exists and is equal to 0 Φ ( { x k : k N } ) .

This finishes the proof of Theorem 1.2.□


Dedicated to David Jerison on his 70’th birthday.


Acknowledgments

Remarks from C.K.: I have known David for 45 years, as a close collaborator and a dear friend. When our collaboration began, at the start of our careers, it brought us great excitement, and it greatly helped to launch my professional path. Our close friendship, through both happy times and very difficult ones, has been a joy. Thank you David! From Z.Z.: I have been greatly influenced and inspired by Prof. Jerison’s work, especially how he makes connections between different fields of math. For that I am always grateful. Happy birthday, Prof. Jerison!

  1. Funding information: Carlos Kenig was supported in part by National Science Foundation (NSF) grants DMS-1800082 and DMS-2153794, and Zihui Zhao was supported by NSF grant DMS-1902756.

  2. Conflict of interest: The authors state that there is no conflict of interest.

Appendix

Lemma A.1

Assume that b k < 1 / 2 and x k = 2 k . The function

g ( x ) = x x k b k

is integrable near the origin. Moreover,

ε ε g ( x ) d x ε 1 b k .

Proof

Let ε > 0 be sufficiently small. We first prove that

ε 0 g ( x ) d x 2 ε 1 b k < + .

For each x ( ε , 0 ) and k N , we have

x x k = x k x x .

Hence

ε 0 g ( x ) d x ε 0 x b k d x = ε 1 b k 1 b k < + .

To estimate 0 ε g ( x ) d x , we assume k 0 N is such that 2 k 0 ε < 2 k 0 + 1 . Then

0 ε g ( x ) d x 0 2 k 0 + 1 g ( x ) d x = i = k 0 2 i 2 i + 1 g ( x ) d x .

Let x [ 2 i , 2 i + 1 ] be arbitrary. For every k i + 1 , we have

x x k 2 i 2 ( i + 1 ) = 2 i 1 ;

for every k i 2 , we have

x x k 2 ( i 2 ) 2 i + 1 = 2 i + 1 .

Hence

(A1) k i , i 1 x x k b k ( 2 i 1 ) k i + 1 b k ( 2 i + 1 ) k i 2 b k .

It remains to estimate

2 i 2 i + 1 x x i b i x x i 1 b i 1 d x .

To that end, let c i denote the midpoint of the interval [ 2 i , 2 i + 1 ] . Then

(A2) 2 i c i x x i b i x x i 1 b i 1 d x ( 2 i 1 ) b i 1 0 2 i 1 t b i d x = ( 2 i 1 ) b i 1 ( 2 i 1 ) 1 b i 1 b i .

Similarly,

(A3) c i 2 i + 1 x x i b i x x i 1 b i 1 d x ( 2 i 1 ) b i 0 2 i 1 t b i 1 d x = ( 2 i 1 ) b i ( 2 i 1 ) 1 b i 1 1 b i 1 .

Combining (A1), (A2), and (A3), we obtain

2 i 2 i + 1 g ( x ) d x 4 ( 2 i 1 ) 1 b k .

Therefore

0 ε g ( x ) d x i = k 0 2 i 2 i + 1 g ( x ) d x 4 i = k 0 ( 2 i 1 ) 1 b k ε 1 b k .

References

[1] V. Adolfsson and L. Escauriaza, C1,α domains and unique continuation at the boundary, Commun. Pur. Appl. Math. 50 (1997), no. 10, 935–969. 10.1002/(SICI)1097-0312(199710)50:10<935::AID-CPA1>3.0.CO;2-HSearch in Google Scholar

[2] V. Adolfsson, L. Escauriaza, and C. E. Kenig, Convex domains and unique continuation at the boundary, Rev. Mat. Iberoamericana 11 (1995), 519–525. 10.4171/RMI/182Search in Google Scholar

[3] L. Caffarelli and C. Kenig, Gradient estimates for variable coefficient parabolic equations and singular perturbation problems, Am. J. Math. 120 (1998), no. 2, 391–439. 10.1353/ajm.1998.0009Search in Google Scholar

[4] J. M. Gallegos, Size of the zero set of solutions of elliptic PDEs near the boundary of Lipschitz domains with small Lipschitz constant, Calculus Variations Partial Differential Equations 62 (2023), no. 4, 113. 10.1007/s00526-022-02426-x.Search in Google Scholar

[5] C. Kenig and W. Wang, A note on boundary unique continuation for harmonic functions in non-smooth domains, Potential Anal. 8 (1998), 143–147. 10.1023/A:1008621009597Search in Google Scholar

[6] C. Kenig, Weighted Hp spaces on Lipschitz domains, Amer. J. Math. 102 (1980), no. 1, 129–163. 10.2307/2374173Search in Google Scholar

[7] C. Kenig, Harmonic analysis techniques for second order elliptic boundary value problems, CBMS Regional Conference Series in Mathematics 83. Published for the Conference Board of the Mathematical Sciences, Washington, DC; by the American Mathematical Society, Providence, RI, 1994. 10.1090/cbms/083Search in Google Scholar

[8] C. Kenig and T. Toro, Free boundary regularity for harmonic measures and Poisson kernels, Ann. Math. 150 (1999), 369–454. 10.2307/121086Search in Google Scholar

[9] C. Kenig and Z. Zhao, Boundary unique continuation on C1-Dini domains and the size of the singular set, Arch. Rational Mech. Anal. 245 (2022), 1–88. 10.1007/s00205-022-01771-7Search in Google Scholar

[10] I. Kukavica and K. Nyström, Unique continuation on the boundary for Dini domains, P. Am. Math. Soc. 126 (1998), no. 2, 441–446. 10.1090/S0002-9939-98-04065-9Search in Google Scholar

[11] F. Lin, Nodal sets of solutions of elliptic and parabolic equations, Commun. Pur. Appl. Math. 45 (1991), 287–308. 10.1002/cpa.3160440303Search in Google Scholar

[12] S. McCurdy, Unique continuation on convex domains, Rev. Mat. Iberoam. (2022). 10.4171/RMI/1389.Search in Google Scholar

[13] E. M. Stein and T. S. Murphy, Harmonic analysis: Real-variable Methods, Orthogonality, and Oscillatory Integrals, Princeton University Press, Princeton New Jersey, 1993. 10.1515/9781400883929Search in Google Scholar

[14] X. Tolsa, Unique continuation at the boundary for harmonic functions in C1 domains and Lipschitz domains with small constant, Commun. Pure Appl. Math. 76 (2023), no. 2, 305–336. 10.1002/cpa.22025Search in Google Scholar

Received: 2022-12-02
Accepted: 2023-03-13
Published Online: 2023-04-26

© 2023 the author(s), published by De Gruyter

This work is licensed under the Creative Commons Attribution 4.0 International License.

Articles in the same Issue

  1. Research Articles
  2. Asymptotic properties of critical points for subcritical Trudinger-Moser functional
  3. The existence of positive solution for an elliptic problem with critical growth and logarithmic perturbation
  4. On some dense sets in the space of dynamical systems
  5. Sharp profiles for diffusive logistic equation with spatial heterogeneity
  6. Generic properties of the Rabinowitz unbounded continuum
  7. Global bifurcation of coexistence states for a prey-predator model with prey-taxis/predator-taxis
  8. Multiple solutions of p-fractional Schrödinger-Choquard-Kirchhoff equations with Hardy-Littlewood-Sobolev critical exponents
  9. Improved fractional Trudinger-Moser inequalities on bounded intervals and the existence of their extremals
  10. The existence of infinitely many boundary blow-up solutions to the p-k-Hessian equation
  11. A priori bounds, existence, and uniqueness of smooth solutions to an anisotropic Lp Minkowski problem for log-concave measure
  12. Existence of nonminimal solutions to an inhomogeneous elliptic equation with supercritical nonlinearity
  13. Non-degeneracy of multi-peak solutions for the Schrödinger-Poisson problem
  14. Gagliardo-Nirenberg-type inequalities using fractional Sobolev spaces and Besov spaces
  15. Ground states of Schrödinger systems with the Chern-Simons gauge fields
  16. Quasilinear problems with nonlinear boundary conditions in higher-dimensional thin domains with corrugated boundaries
  17. A system of equations involving the fractional p-Laplacian and doubly critical nonlinearities
  18. A modified Picone-type identity and the uniqueness of positive symmetric solutions for a prescribed mean curvature problem
  19. On a version of hybrid existence result for a system of nonlinear equations
  20. Special Issue: Geometric PDEs and applications
  21. Preface for the special issue on “Geometric Partial Differential Equations and Applications”
  22. Convex hypersurfaces with prescribed Musielak-Orlicz-Gauss image measure
  23. Total mean curvatures of Riemannian hypersurfaces
  24. On degenerate case of prescribed curvature measure problems
  25. A curvature flow to the Lp Minkowski-type problem of q-capacity
  26. Aleksandrov reflection for extrinsic geometric flows of Euclidean hypersurfaces
  27. A note on second derivative estimates for Monge-Ampère-type equations
  28. The Lp chord Minkowski problem
  29. Widths of balls and free boundary minimal submanifolds
  30. Smooth approximation of twisted Kähler-Einstein metrics
  31. The exterior Dirichlet problem for the homogeneous complex k-Hessian equation
  32. A Carleman inequality on product manifolds and applications to rigidity problems
  33. Asymptotic behavior of solutions to the Monge-Ampère equations with slow convergence rate at infinity
  34. Pinched hypersurfaces are compact
  35. The spinorial energy for asymptotically Euclidean Ricci flow
  36. Geometry of CMC surfaces of finite index
  37. Capillary Schwarz symmetrization in the half-space
  38. Regularity of optimal mapping between hypercubes
  39. Special Issue: In honor of David Jerison
  40. Preface for the special issue in honor of David Jerison
  41. Homogenization of oblique boundary value problems
  42. A proof of a trace formula by Richard Melrose
  43. Compactness estimates for minimizers of the Alt-Phillips functional of negative exponents
  44. Regularity properties of monotone measure-preserving maps
  45. Examples of non-Dini domains with large singular sets
  46. Sharp inequalities for coherent states and their optimizers
  47. Gradient estimates and the fundamental solution for higher-order elliptic systems with lower-order terms
  48. Propagation of symmetries for Ricci shrinkers
  49. Linear extension operators for Sobolev spaces on radially symmetric binary trees
  50. The Neumann problem on the domain in 𝕊3 bounded by the Clifford torus
  51. On an effective equation of the reduced Hartree-Fock theory
  52. Polynomial sequences in discrete nilpotent groups of step 2
  53. Integral inequalities with an extended Poisson kernel and the existence of the extremals
  54. On singular solutions of Lane-Emden equation on the Heisenberg group
Downloaded on 13.9.2025 from https://www.degruyterbrill.com/document/doi/10.1515/ans-2022-0058/html
Scroll to top button