Skip to main content
Log in

Harmonic measure and quantitative connectivity: geometric characterization of the \(L^p\)-solvability of the Dirichlet problem

  • Published:
Inventiones mathematicae Aims and scope

Abstract

It is well-known that quantitative, scale invariant absolute continuity (more precisely, the weak-\(A_\infty \) property) of harmonic measure with respect to surface measure, on the boundary of an open set \( \Omega \subset \mathbb {R}^{n+1}\) with Ahlfors–David regular boundary, is equivalent to the solvability of the Dirichlet problem in \(\Omega \), with data in \(L^p(\partial \Omega )\) for some \(p<\infty \). In this paper, we give a geometric characterization of the weak-\(A_\infty \) property, of harmonic measure, and hence of solvability of the \(L^p\) Dirichlet problem for some finite p. This characterization is obtained under background hypotheses (an interior corkscrew condition, along with Ahlfors–David regularity of the boundary) that are natural, and in a certain sense optimal: we provide counter-examples in the absence of either of them (or even one of the two, upper or lower, Ahlfors–David bounds); moreover, the examples show that the upper and lower Ahlfors–David bounds are each quantitatively sharp.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1

Similar content being viewed by others

Notes

  1. This is a quantitative, scale-invariant version of rectifiability, see Definition 2.2 and the ensuing comments.

  2. See, e.g., [26, Proposition 13] for the details in this context, but the proof originates in [39].

  3. The CDC is a scale invariant potential theoretic “thickness” condition, i.e., a quantitative version of Weiner regularity; see, e.g., [1].

  4. An earlier version of this work [33] gave a direct proof of the fact that (1) implies (3) in Theorem 1.1, without passing through condition (2).

  5. To guarantee the existence of the sets \(V_i\) and the fact that they are contained in \(\Omega \) we use the assumption that \(\Omega =\mathbb {R}^{n+1}{\setminus }\partial \Omega \).

References

  1. Aikawa, H., Hirata, K.: Doubling conditions for harmonic measure in John domains. Ann. Inst. Fourier (Grenoble) 58(2), 429–445 (2008)

    Article  MathSciNet  MATH  Google Scholar 

  2. Akman, M., Bortz, S., Hofmann, S., Martell, J.M.: Rectifiability, interior approximation and Harmonic Measure. Ark. Mat. 57(1), 1–22 (2019)

    Article  MathSciNet  MATH  Google Scholar 

  3. Alt, H.W., Caffarelli, L.A., Friedman, A.: Variational problems with two phases and their free boundaries. Trans. Am. Math. Soc. 282(2), 431–461 (1984)

    Article  MathSciNet  MATH  Google Scholar 

  4. Auscher, P., Hofmann, S., Lewis, J.L., Tchamitchian, P.: Extrapolation of Carleson measures and the analyticity of Kato’s square-root operators. Acta Math. 187(2), 161–190 (2001)

    Article  MathSciNet  MATH  Google Scholar 

  5. Auscher, P., Hofmann, S., Muscalu, C., Tao, T., Thiele, C.: Carleson measures, trees, extrapolation, and \(T(b)\) theorems. Publ. Mat. 46(2), 257–325 (2002)

    MathSciNet  MATH  Google Scholar 

  6. Azzam, J.: Semi-uniform domains and a characterization of the \(A_\infty \) property for harmonic measure. Preprint, arXiv:1711.03088

  7. Azzam, J., Garnett, J., Mourgoglou, M., Tolsa, X.: Uniform rectifiability, elliptic measure, square functions, and \(\varepsilon \)-approximability via an ACF monotonicity formula. Preprint arXiv:1612.02650 (2016)

  8. Azzam, J., Mourgoglou, M., Tolsa, X.: Singular sets for harmonic measure on locally flat domains with locally finite surface measure. Int. Math. Res. Not. 2017(12), 3751–3773 (2017)

    MathSciNet  MATH  Google Scholar 

  9. Azzam, J., Mourgoglou, M., Tolsa, X.: Harmonic measure and quantitative connectivity: geometric characterization of the \(L^p\)-solvability of the Dirichlet problem. Part II. Preprint arXiv:1803.07975

  10. Badger, M.: Null sets of harmonic measure on NTA domains: Lipschitz approximation revisited. Math. Z. 270(1–2), 241–262 (2012)

    Article  MathSciNet  MATH  Google Scholar 

  11. Bennewitz, B., Lewis, J.L.: On weak reverse Hölder inequalities for nondoubling harmonic measures. Complex Var. Theory Appl. 49(7-9), 571–582 (2004)

    MathSciNet  MATH  Google Scholar 

  12. Bishop, C., Jones, P.: Harmonic measure and arclength. Ann. Math. (2) 132, 511–547 (1990)

    Article  MathSciNet  MATH  Google Scholar 

  13. Bourgain, J.: On the Hausdorff dimension of harmonic measure in higher dimensions. Invent. Math. 87, 477–483 (1987)

    Article  MathSciNet  MATH  Google Scholar 

  14. Caffarelli, L., Salsa, S.: A geometric approach to free boundary problems. Graduate Texts in Math. 64. Am. Math. Soc. (2005)

  15. Carleson, L.: Interpolation by bounded analytic functions and the corona problem. Ann. Math. (2) 76, 547–559 (1962)

    Article  MathSciNet  MATH  Google Scholar 

  16. Carleson, L., Garnett, J.: Interpolating sequences and separation properties. J. Anal. Math. 28, 273–299 (1975)

    Article  MATH  Google Scholar 

  17. Christ, M.: A \(T(b)\) theorem with remarks on analytic capacity and the Cauchy integral. Colloq. Math. LX/LXI, 601–628 (1990)

    MathSciNet  MATH  Google Scholar 

  18. Dahlberg, B.: On estimates for harmonic measure. Arch. Ration. Mech. Anal. 65, 272–288 (1977)

    Article  MathSciNet  Google Scholar 

  19. David, G.: Morceaux de graphes lipschitziens et intégrales singulières sur une surface (French) [Pieces of Lipschitz graphs and singular integrals on a surface]. Rev. Mat. Iberoam. 4(1), 73–114 (1988)

    Article  Google Scholar 

  20. David, G.: Wavelets and Singular Integrals on Curves and Surfaces. Lecture Notes in Mathematics, vol. 1465. Springer, Berlin (1991)

    Book  Google Scholar 

  21. David, G., Jerison, D.: Lipschitz approximation to hypersurfaces, harmonic measure, and singular integrals. Indiana Univ. Math. J. 39(3), 831–845 (1990)

    Article  MathSciNet  MATH  Google Scholar 

  22. David, G., Semmes, S.: Singular integrals and rectifiable sets in \(\mathbb{R}^n\). Au-delĂ  des graphes lipschitziens. Asterisque 193, 152 (1991)

    MATH  Google Scholar 

  23. David, G., Semmes, S.: Analysis of and on Uniformly Rectifiable Sets. Mathematical Monographs and Surveys, vol. 38, AMS (1993)

  24. Evans, L.C., Gariepy, R.F.: Measure Theory and Fine Properties of Functions. Studies in Advanced Mathematics. CRC Press, Boca Raton (1992)

    MATH  Google Scholar 

  25. Heinonen, J., Kilpeläinen, T., Martio, O.: Nonlinear Potential Theory of Degenerate Elliptic Equations (Rev. ed.). Dover, New York (2006)

    MATH  Google Scholar 

  26. Hofmann, S.: Quantitative absolute continuity of harmonic measure and the Dirichlet problem: a survey of recent progress. To appear in special volume in honor of the 65th birthday of Carlos Kenig

  27. Hofmann, S., Le, P.: BMO solvability and absolute continuity of harmonic measure. J. Geom. Anal. 28(4), 3278–3299 (2018)

    Article  MathSciNet  MATH  Google Scholar 

  28. Hofmann, S., Le, P., Martell, J.M., Nyström, K.: The weak-\(A_\infty \) property of harmonic and \(p\)-harmonic measures implies uniform rectifiability. Anal. PDE. 10(3), 513–558 (2017)

    MathSciNet  MATH  Google Scholar 

  29. Hofmann, S., Lewis, J.L.: The Dirichlet problem for parabolic operators with singular drift terms. Mem. Am. Math. Soc. 151, 719 (2001)

    MathSciNet  MATH  Google Scholar 

  30. Hofmann, S., Martell, J.M.: \(A_\infty \) estimates via extrapolation of Carleson measures and applications to divergence form elliptic operators. Trans. Am. Math. Soc. 364(1), 65–101 (2012)

    MathSciNet  MATH  Google Scholar 

  31. Hofmann, S., Martell, J.M.: Uniform rectifiability and harmonic measure I: Uniform rectifiability implies Poisson kernels in \(L^p\). Ann. Sci. École Norm. Sup. 47(3), 577–654 (2014)

    MATH  Google Scholar 

  32. Hofmann, S., Martell, J.M.: Uniform Rectifiability and harmonic measure IV: Ahlfors regularity plus Poisson kernels in \(L^p\) implies uniform rectifiability. Preprint, arXiv:1505.06499

  33. Hofmann, S., Martell, J.M.: A sufficient geometric criterion for quantitative absolute continuity of harmonic measure. Preprint arXiv:1712.03696v1

  34. Hofmann, S., Martell, J.M.: Harmonic measure and quantitative connectivity: geometric characterization of the \(L^p\) solvability of the Dirichlet problem. Part I. Preprint arXiv:1712.03696v3 (2018)

  35. Hofmann, S., Martell, J.M., Mayboroda, S.: Uniform rectifiability, Carleson measure estimates, and approximation of harmonic functions. Duke Math. J. 165(12), 2331–2389 (2016)

    Article  MathSciNet  MATH  Google Scholar 

  36. Hofmann, S., Mitrea, D., Mitrea, M., Morris, A.J.: \(L^p\)-square function estimates on spaces of homogeneous type and on uniformly rectifiable sets. Mem. Am. Math. Soc. 245, 1159 (2017)

    MATH  Google Scholar 

  37. Hofmann, S., Mitrea, M., Taylor, M.: Singular integrals and elliptic boundary problems on regular Semmes-Kenig-Toro domains. Int. Math. Res. Not. 2010, 2567–2865 (2010)

    MathSciNet  MATH  Google Scholar 

  38. Hytönen, T., Kairema, A.: Systems of dyadic cubes in a doubling metric space. Colloq. Math. 126(1), 1–33 (2012)

    Article  MathSciNet  MATH  Google Scholar 

  39. Jerison, D., Kenig, C.: Boundary behavior of harmonic functions in nontangentially accessible domains. Adv. Math. 46(1), 80–147 (1982)

    Article  MathSciNet  MATH  Google Scholar 

  40. Kenig, C.E., Toro, T.: Harmonic measure on locally flat domains. Duke Math. J. 87(3), 509–551 (1997)

    Article  MathSciNet  MATH  Google Scholar 

  41. Lavrentiev, M.: Boundary problems in the theory of univalent functions (Russian). Math Sb. 43, 815–846 (1936); AMS Transl. Series 2 32 , 1–35 (1963)

  42. Lewis, J., Murray, M.: The method of layer potentials for the heat equation in time-varying domains. Mem. Am. Math. Soc. 114, 545 (1995)

    MathSciNet  Google Scholar 

  43. Mattila, P., Melnikov, M., Verdera, J.: The Cauchy integral, analytic capacity, and uniform rectifiability. Ann. Math. (2) 144(1), 127–136 (1996)

    Article  MathSciNet  MATH  Google Scholar 

  44. Mourgoglou, M., Tolsa, X.: Harmonic measure and Riesz transform in uniform and general domains. J. Reine Angew. Math. 758, 183–221 (2020)

    Article  MathSciNet  MATH  Google Scholar 

  45. Nazarov, F., Tolsa, X., Volberg, A.: On the uniform rectifiability of ad-regular measures with bounded Riesz transform operator: the case of codimension 1. Acta Math. 213(2), 237–321 (2014)

    Article  MathSciNet  MATH  Google Scholar 

  46. Petrosyan, A., Shahgholian, H., Uraltseva, N.: Regularity of free boundaries in obstacle-type problems. Regularity of free boundaries in obstacle-type problems. Graduate Studies in Mathematics, vol. 136. American Mathematical Society, Providence, RI (2012)

  47. Riesz, F. and M.: Über die randwerte einer analtischen funktion. Compte Rendues du Quatrième Congrès des Mathématiciens Scandinaves, Stockholm 1916, Almqvists and Wilksels, Uppsala (1920)

  48. Semmes, S.: A criterion for the boundedness of singular integrals on on hypersurfaces. Trans. Am. Math. Soc. 311, 501–513 (1989)

    Article  MathSciNet  MATH  Google Scholar 

  49. Stein, E.M.: Singular Integrals and Differentiability Properties of Functions. Princteon University Press, Princeton (1970)

    MATH  Google Scholar 

  50. Tolsa, X.: Analytic capacity, the Cauchy transform, and non-homogeneous Calderón-Zygmund theory. Progress in Mathematics, vol. 307. Birkhäuser Verlag, Basel (2014)

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Xavier Tolsa.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

S.H. was supported by NSF Grant DMS-1664047. J.M.M. acknowledges financial support from the Spanish Ministry of Economy and Competitiveness, through the “Severo Ochoa” Programme for Centres of Excellence in R&D (SEV-2015- 0554). He also acknowledges that the research leading to these results has received funding from the European Research Council under the European Union’s Seventh Framework Programme (FP7/2007-2013)/ ERC Agreement No. 615112 HAPDEGMT. In addition, S.H. and J.M.M. were supported by NSF Grant DMS-1440140 while in residence at the MSRI in Berkeley, California, during Spring semester 2017. M.M. was supported by IKERBASQUE and partially supported by the Grant MTM-2017-82160-C2-2-P of the Ministerio de Economía y Competitividad (Spain), and by IT-641-13 (Basque Government). X.T. was supported by the ERC Grant 320501 of the European Research Council (FP7/2007-2013) and partially supported by MTM-2016-77635-P, MDM-2014-044 (MICINN, Spain), 2017-SGR-395 (Catalonia), and by Marie Curie ITN MAnET (FP7-607647).

Appendix A: Some counter-examples

Appendix A: Some counter-examples

We shall discuss some counter-examples which show that our background hypotheses in Theorem 1.1 (namely, n-ADR and interior corkscrew condition) are natural, and in some sense in the nature of best possible. In the first two examples, \(\Omega \) is a domain satisfying an interior corkscrew condition, such that \(\partial \Omega \) satisfies exactly one (but not both) of the upper or the lower n-ADR bounds, and for which harmonic measure \(\omega \) fails to be weak-\(A_\infty \) with respect to surface measure \(\sigma \) on \(\partial \Omega \). In this setting, in which full n-ADR fails, there is no established notion of uniform rectifiability, but in each case, the domain will enjoy some substitute property which would imply uniform rectifiability of the boundary in the presence of full n-ADR. Moreover, these examples may be constructed in such a way that the failure of the condition (either upper or lower n-ADR) can be expressed quantitatively, with a bound that may be taken arbitrarily close to a true n-ADR bound; see (A.3) and (A.6) below.

In the last example, we construct an open set \(\Omega \) with n-ADR boundary, and for which \(\omega \in \) weak-\(A_\infty \) with respect to surface measure, but for which the interior corkscrew condition fails, and \(\partial \Omega \) is not n-UR.

Example 1

Failure of the upper n-ADR bound. In [8], the authors construct an example of a Reifenberg flat domain \(\Omega \subset \mathbb {R}^{n+1}\) for which surface measure \(\sigma = H^n\lfloor _{\,\partial \Omega }\) is locally finite on \(\partial \Omega \), but for which the upper n-ADR bound

$$\begin{aligned} \sigma (\Delta (x,r)) \le C r^n \end{aligned}$$
(A.1)

fails, and for which harmonic measure \(\omega \) is not absolutely continuous with respect to \(\sigma \). Note that the hypothesis of Reifenberg flatness implies in particular that \(\Omega \) and \(\Omega _{ext}:= \mathbb {R}^{n+1}{\setminus } \overline{\Omega }\) are both NTA domains, hence both enjoy the corkscrew condition, so by the relative isoperimetric inequality, the lower n-ADR bound

$$\begin{aligned} \sigma (\Delta (x,r)) \ge c r^n \end{aligned}$$
(A.2)

holds. Thus, it is the failure of (A.1) which causes the failure of absolute continuity: in the presence of (A.1), the results of [21] apply, and one has that \(\omega \in A_\infty (\sigma )\), and that \(\partial \Omega \) satisfies a “big pieces of Lipschitz graphs” condition (see [21] for a precise statement), and hence is n-UR. We note that by a result of Badger [10], a version of the Lipschitz approximation result of [21] still holds for NTA domains with locally finite surface measure, even in the absence of the upper n-ADR condition.

In addition, given any \(\varepsilon >0\), the construction in [8] can be made in such a way that (A.1) fails “within \(\varepsilon \)”, i.e., so that

$$\begin{aligned} \sigma (\Delta (x,r) \le C r^{n-\varepsilon }, \quad \forall \, x\in \partial \Omega ,\, r<1. \end{aligned}$$
(A.3)

Let us sketch an argument to explain why this is so; we refer the interested reader to [8] for more details.

The domain \(\Omega \) in [8] is obtained by enlarging a Wolff snowflake, that we will denote here by D. Both \(\Omega \) and D are \(\delta \)-Reifenberg flat, with \(\delta \) as small as wished in the construction (recall that Wolff snowflakes can be taken \(\delta \)-Reifenberg flat, with \(\delta \) as small as wished).

It is shown in [8, Theorem 3.1] that for all \(x\in \partial \Omega \) and \(r<1\),

$$\begin{aligned} H^n(B(x,r)\cap \partial \Omega ) \lesssim \max ( r^n, r^\alpha \mu (B(x,Cr)) ) \le \max ( r^n, \mu (B(x,Cr)) ) \end{aligned}$$
(A.4)

where \(\mu \) is some measure supported on \(\partial D\) satisfying \(\mu (B(x,r)) \gtrsim r^{n-\alpha }\) for all x in some compact set \(E \subset \partial \Omega \cap \partial D\), and some \(\alpha >0.\) In the construction in [8], the authors take \(\mu =\omega _D\), the harmonic measure for D. Further, from results of Kenig and Toro it follows that harmonic measure in a \(\delta \)-Reifenberg flat domain D satisfies

$$\begin{aligned} \omega _D(B(x,r)) \lesssim r^{n-\varepsilon } \omega _D(B(x,1)), \quad \forall \, x\in \partial D, \, r<1, \end{aligned}$$

with \(\varepsilon \rightarrow 0\) as \(\delta \rightarrow 0\) (see [40, Theorem 4.1]). As a consequence, the measure \(\mu \) satisfies

$$\begin{aligned} \mu (B(x,r)) \lesssim r^{n-\varepsilon }, \quad \forall \, x\in {\mathbb {R}}^{n+1},\, r<1, \end{aligned}$$

with \(\varepsilon \) as small as wished depending on \(\delta \). From (A.4), it follows that

$$\begin{aligned} H^n(B(x,r)\cap \partial \Omega ) \lesssim \max ( r^n, r^{n-\varepsilon } ) \le r^{n-\varepsilon }, \quad \forall \, x\in \partial \Omega , r<1. \end{aligned}$$

Example 2

Failure of the lower n-ADR bound. In [2, Example 5.5], the authors give an example of a domain satisfying the interior corkscrew condition, whose boundary is rectifiable (indeed, it is contained in a countable union of hyperplanes), and satisfies the upper n-ADR condition (A.1), but not the lower n-ADR condition (A.2), but for which surface measure \(\sigma \) fails to be absolutely continuous with respect to harmonic measure, and in fact, for which the non-degeneracy condition

$$\begin{aligned} A\subset \Delta _x:= B(x,10\delta _\Omega (x)) \cap \partial \Omega ,\quad \sigma (A)\ge (1-\eta ) \sigma (\Delta _x) \, \, \implies \quad \omega ^x(A) \ge c, \end{aligned}$$
(A.5)

fails to hold uniformly for \(x\in \Omega \), for any fixed positive \(\eta \) and c, and therefore \(\omega \) cannot be weak-\(A_\infty \) with respect to \(\sigma \). We note that in the presence of the full n-ADR condition, if \(\partial \Omega \) were contained in a countable union of hyperplanes (as it is in the example), then in particular it would satisfy the “BAUP” condition of [23], and thus would be n-UR [23, Theorem I.2.18, p. 36].

Moreover, given any \(\varepsilon >0\), the parameters in the example of [2] can be chosen in such a way that the lower ADR bound fails “within \(\varepsilon \)”, i.e., so that

$$\begin{aligned} H^n(\Delta (x,r))\gtrsim \min (r^{n+\varepsilon },r^n), \quad \forall x\in \partial \Omega . \end{aligned}$$
(A.6)

To see this, we proceed as follows. We follow closely the construction in [2, Example 5.5], with some modification of the parameters. Fix \(\varepsilon >0\), and set

$$\begin{aligned} c_k:= 2^{-k(n+\varepsilon )}. \end{aligned}$$

For \(k\ge 1\), and \(n\ge 2\), set

$$\begin{aligned} \Sigma _k:= \{(x,t)\in \mathbb {R}^{n+1}_+:\, t=2^{-k},\, x\in \overline{\Delta (0,2^{-\varepsilon k}c_k)}+ c_k {\mathbb {Z}}^n\}, \end{aligned}$$

where for \(x\in \mathbb {R}^n\), \(\Delta (x,r):= \{y\in \mathbb {R}^n:\, |x-y|<r\}\) is the usual n-disk of radius r centered at x. Define

$$\begin{aligned} \Omega := \mathbb {R}^{n+1}_+{\setminus } \left( \cup _{k=1}^\infty \Sigma _k \right) ,\qquad \Omega _k:= \mathbb {R}^{n+1}_+{\setminus } \Sigma _k, \end{aligned}$$

each of which is clearly open and connected. Notice that \(\Omega \) satisfies the interior Corkscrew condition (since the sets \(\Sigma _k\) are located at heights which are sufficently separated). Moreover, it is easy to see that \(\partial \Omega \) satisfies the upper ADR condition and that \(\mathbb {R}^n\times \{0\}\subset \partial \Omega \).

On the other hand, the lower ADR bound fails. To see this, let \(X=(x,0)\in \partial \Omega \), and choose \(\mathbf {m}_{k,x}\in {\mathbb {Z}}^n\) and \(X_k=(c_k\,\mathbf {m}_{k,x}, 2^{-k})\in \Sigma _k\subset \partial \Omega \) such that \(X_k\rightarrow X\). Set \(B_k=B(X_k, 2^{-k-2})\), and observe that \(H^n(B_k\cap \partial \Omega )/(2^{-kn})\approx 2^{-kn\varepsilon }\rightarrow 0\) as \(k\rightarrow \infty \), or equivalently

$$\begin{aligned} H^n(B_k\cap \partial \Omega )\approx r_k^{n+\varepsilon '}, \end{aligned}$$

where \(B_k\) has radius \(r_k \approx 2^{-k}\), and \(\varepsilon '=n\varepsilon \). We shall show that this behavior is in fact typical, and that (A.6) holds, with \(\varepsilon '\) in place of \(\varepsilon \).

Let \(\omega ^{(\cdot )}:= \omega ^{(\cdot )}_\Omega \) and \(\omega _k^{(\cdot )}:= \omega ^{(\cdot )}_{\Omega _k}\) denote harmonic measure for the domains \(\Omega \) and \(\Omega _k\) respectively.

Claim

\(\omega ^{(\cdot )}(F) = 0\), with \(F:=\mathbb {R}^n\times \{0\}\). Thus, in particular (A.5) fails.

It remains to verify (A.6), and the claim. As regards the former, note that for \(X=(x,0)\in F\), we have the trivial standard lower n-ADR bound \(H^n(\Delta (X,r))\gtrsim r^n\), whereas for \(X=(x,2^{-k}) \in \Sigma _k\), we have

$$\begin{aligned} H^n\lfloor _{\partial \Omega }\big (B(X,r)\big )\ge H^n\lfloor _{\partial \Omega _k}\big (B(X,r)\big )\gtrsim \left\{ \begin{array}{ll} r^n, &{} r<2^{-\varepsilon k} c_k, \\ 2^{-kn\varepsilon }c_k^{n}, &{} 2^{-\varepsilon k} c_k \le r\le c_k \\ 2^{-kn\varepsilon }r^n, &{} c_k< r\le 2^{-k+1}\\ r^n,&{} r> 2^{-k+1}. \end{array} \right. \end{aligned}$$
(A.7)

The first and fourth of these estimates are of course the standard lower n-ADR bound. For \(r\le c_k\), the second estimate is bounded below by \(2^{-kn\varepsilon }r^n\), and in turn, with \(r\lesssim 2^{-k}\), the second and third estimates are therefore bounded below by

$$\begin{aligned} 2^{-kn\varepsilon }r^n \gtrsim r^{n+n\varepsilon }= r^{n+\varepsilon '}, \end{aligned}$$

which yields (A.6) with \(\varepsilon '=n\varepsilon \) in place of \(\varepsilon \).

Let us now prove the claim. We first recall some definitions. Given an open set \(O\subset \mathbb {R}^{n+1}\), and a compact set \(K\subset O\), we define the capacity of K relative to O as

$$\begin{aligned} \text{ cap }(K,O)=\inf \left\{ \iint _O|\nabla \phi |^2\, dY:\ \phi \in C_0^\infty (O), \ \phi \ge 1 \text{ in } K\right\} . \end{aligned}$$

Also, the inhomogeneous capacity of K is defined as

$$\begin{aligned} \text{ Cap } (K)=\inf \left\{ \iint _{\mathbb {R}^{n+1}} \big (|\phi |^2+|\nabla \phi |^2\big )\, dY:\ \phi \in C_0^\infty (\mathbb {R}), \ \phi \ge 1 \text{ in } K\right\} . \end{aligned}$$

Combining [25, Theorem 2.38], [1, Theorem 2.2.7] and [1, Theorem 4.5.2] we have that if K is a compact subset of \(\overline{B}\), where B is a ball with radius smaller than 1, then

$$\begin{aligned} \text{ cap }(K,2B) \gtrsim \text{ Cap }(K) \gtrsim \sup _{\mu } \mu (K) \end{aligned}$$
(A.8)

where the implicit constants depend only on n, the sup runs over all Radon positive measures \(\mu \) supported on K, for which

$$\begin{aligned} W(\mu )(X):=\int _0^1 \frac{\mu (B(X,t))}{t^{n-1}}\,\frac{dt}{t} \le 1, \qquad \forall \, X\in {{\,\mathrm{supp}\,}}\mu . \end{aligned}$$

Fix \(k\ge 2\), and set

$$\begin{aligned} \beta =\beta _k:= 2^{k(n-1)}c_k= 2^{k(n-1)} 2^{-k(n+\varepsilon )} = 2^{-k(1+\varepsilon )}, \end{aligned}$$

by definition of \(c_k\). Our next goal is to show that

$$\begin{aligned} \mathrm{cap}\big (\overline{B(X_0,s)}\cap \Sigma _k, B(X_0,2s)\big ) \gtrsim s^{n-1},\quad X_0:=(x_0,2^{-k})\in \Sigma _k,\ \beta \le s<1. \end{aligned}$$
(A.9)

For a fixed \(X_0\) and s, write \(K=\overline{B(X_0,s)}\cap \Sigma _k\), set \(\mu = 2^{kn\varepsilon } s^{-1} \, H^n\lfloor _K,\) and note that for \(X\in K\), similarly to (A.7), we have

$$\begin{aligned} \mu \big (B(X,r)\big )\approx 2^{kn\varepsilon } s^{-1} \left\{ \begin{array}{l@{\quad }l} r^n, &{} r<2^{-\varepsilon k} c_k, \\ 2^{-kn\varepsilon }c_k^{n},&{}2^{-\varepsilon k} c_k \le r\le c_k \\ 2^{-kn\varepsilon }r^n,&{} c_k< r\le s\\ 2^{-kn\varepsilon }s^n, &{} r> s. \end{array} \right. \end{aligned}$$
(A.10)

To compute \(W(\mu )(X)\) for \(X\in K\) write

$$\begin{aligned} W(\mu )(X)&= \int _0^1 \frac{\mu (B(X,t))}{t^{n-1}}\,\frac{dt}{t}= \int _0^{2^{-\varepsilon k}c_k} +\, \int _{2^{-\varepsilon k}c_k}^{c_k} + \, \int _{c_k}^s +\, \int _s^1\\&=: \,I + II+III+IV. \end{aligned}$$

Then, since \(s\ge \beta = 2^{k(n-1)} c_k = 2^{-k(1+\varepsilon )}\),

$$\begin{aligned} I + II \lesssim 2^{kn\varepsilon } s^{-1} \left( 2^{-\varepsilon k}c_k \, + 2^{-kn\varepsilon }c_k^{n}\int _{2^{-\varepsilon k}c_k}^\infty \frac{dt}{t^n} \right) \lesssim \, 2^{\varepsilon k(n-1)} c_k s^{-1} \lesssim 1. \end{aligned}$$

Furthermore, the last two estimates in (A.10) easily imply that \(III+IV\lesssim 1\) and hence \(W(\mu )(X)\lesssim 1\) for every \(X\in K\). This, (A.8), and (A.10) imply as desired (A.9):

$$\begin{aligned} \mathrm{cap}\big (\overline{B(X_0,s)}\cap \Sigma _k, B(X_0,2s)\big ) \gtrsim \mu (K) \gtrsim s^{n-1}. \end{aligned}$$

Set

$$\begin{aligned} P_k:= \left\{ \left( x,2^{-k}-\beta \right) \in \mathbb {R}^{n+1}_+:\, x\in \mathbb {R}^n\right\} , \end{aligned}$$

and observe that for \(X\in P_k\),

$$\begin{aligned} \beta \le \delta _k(X):= {\text {dist}}(X,\partial \Omega _k) = {\text {dist}}(X,\Sigma _k)\le 2\beta . \end{aligned}$$

Recall that \(F=\mathbb {R}^n\times \{0\}\), and define

$$\begin{aligned} u(X):= \omega _k^X(F),\qquad X\in \Omega _k. \end{aligned}$$

Observe that \(u\in W^{1,2}(\Omega _k)\cap C(\overline{\Omega _k})\) since \(\partial \Omega _k\) is ADR (constants depend on k but we just use this qualitatively) and \(\chi _F\) is a Lipschitz function on \(\partial \Omega _k\). Fix \(Z_0\in P_k\) and let \(Z_0'\in \Sigma _k\) be such that \(|Z_0-Z_0'|={\text {dist}}(Z_0, \partial \Omega _k)\le 2\beta \). Let \(\Omega _{Z_0}= \Omega _k\cap B(Z_0', \frac{3}{4} 2^{-k})\), which is an open connected bounded set. We can now apply the usual capacitary estimates (see, e.g., [25, Theorem 6.18]) to find a constant \(\alpha =\alpha (n)>0\) such that

$$\begin{aligned} u(Z_0) \lesssim \exp \left( -\alpha \int _{3\,\beta }^{2^{-k-2}} \frac{ds}{s} \right) \approx \big (2^k \beta \big )^\alpha =2^{-\alpha \varepsilon k}. \end{aligned}$$

where we have used (A.9), the definition of \(\beta \), and the fact that \(u\equiv 0\) on \(\partial \Omega _k \cap B(Z_0',2^{-k-1})\). Note that the last estimate holds for any \(Z_0\in P_k\) and therefore, by the maximum principle,

$$\begin{aligned} u(x,t) \lesssim 2^{-\alpha \varepsilon k}, \qquad (x,t) \in \Omega _k, \quad t > 2^{-k}-\beta . \end{aligned}$$

In particular, if we set \(X_0:= (0,\dots ,0,1) \in \mathbb {R}^{n+1}_+\), then by another application of the maximum principle,

$$\begin{aligned} \omega ^{X_0}(F)\, \le \, \omega _k^{X_0}(F)= u(X_0) \, \lesssim 2^{-\alpha \varepsilon k}\, \rightarrow 0, \end{aligned}$$

as \(k \rightarrow \infty \), and the claim is established.

Example 3

Failure of the interior corkscrew condition. The example is based on the construction of Garnett’s 4-corners Cantor set \(\mathcal {C}\subset \mathbb {R}^2\) (see, e.g., [23, Chapter 1]). Let \(I_0\) be a unit square positioned with lower left corner at the origin in the plane, and in general for each \(k = 0, 1,2,\dots \), we let \(I_k\) be the unit square positioned with lower left corner at the point (2k, 0) on the x-axis. Set \(\Omega _0:= I_0\). Let \(\Omega _1\) be the first stage of the 4-corners construction, i.e., a union of four squares of side length 1/4, positioned in the corners of the unit square \(I_1\), and similarly, for each k, let \(\Omega _k\) be the k-th stage of the 4-corners construction, positioned inside \(I_k\). Note that \({\text {dist}}(\Omega _k, \Omega _{k+1}) = 1\) for every k. Set \(\Omega := \cup _k \Omega _k\). It is easy to check that \(\partial \Omega \) is n-ADR, and that the non-degeneracy condition (A.5) holds in \(\Omega \) for some uniform positive \(\eta \) and c, and thus by the criterion of [11], \(\omega \in \) weak-\(A_\infty (\sigma )\). On the other hand, the interior corkscrew condition clearly fails to hold in \(\Omega \) (it holds only for decreasingly small scales as k increases), and certainly \(\partial \Omega \) cannot be n-UR: indeed, if it were, then \(\partial \Omega _k\) would be n-UR, with uniform constants, for each k, and this would imply that \(\mathcal {C}\) itself was n-UR, whereas in fact, as is well known, it is totally non-rectifiable. One can produce a similar set in 3 dimensions by simply taking the cylinder \(\Omega '=\Omega \times [0,1]\). Details are left to the interested reader.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Azzam, J., Hofmann, S., Martell, J.M. et al. Harmonic measure and quantitative connectivity: geometric characterization of the \(L^p\)-solvability of the Dirichlet problem. Invent. math. 222, 881–993 (2020). https://doi.org/10.1007/s00222-020-00984-5

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00222-020-00984-5

Mathematics Subject Classification

Navigation