Skip to main content
Log in

Gaussian fields, equilibrium potentials and multiplicative chaos for Dirichlet forms

  • Published:
Potential Analysis Aims and scope Submit manuscript

Abstract

For a Dirichlet form \((\mathcal {E},\mathcal {F})\) on L2(E;m), let \(\mathbb {G}(\mathcal {E})=\{X_{u};u\in \mathcal {F}_{e}\}\) be the Gaussian field indexed by the extended Dirichlet space \(\mathcal {F}_{e}\). We first solve the equilibrium problem for a regular recurrent Dirichlet form \(\mathcal {E}\) of finding for a closed set B a probability measure μB concentrated on B whose recurrent potential \(R\mu ^{B}\in \mathcal {F}_{e}\) is constant q.e. on B (called a Robin constant). We next assume that E is the complex plane \(\mathbb {C}\) and \(\mathcal {E}\) is a regular recurrent strongly local Dirichlet form. For the closed disk \(\bar B(\textbf {x},r)=\{\textbf {z}\in \mathbb {C}:|\textbf {z}-\textbf {x}|\le r\}\), let μx, r and f(x, r) be its equilibrium measure and Robin constant. Denote the Gaussian random variable \(X_{R\mu ^{\textbf {x}.r}}\in \mathbb {G}(\mathcal {E})\) by Yx, r and let, for a given constant γ > 0, \(\mu _{r}(A,\omega )={\int \limits }_{A} \exp (\gamma Y^{\textbf {x},r}-(1/2)\gamma ^{2} f(\textbf {x},r))d\textbf {x}.\) Under a certain condition on the growth rate of f(x, r), we prove the convergence in probability of μr(A, ω) to a random measure \(\overline {\mu }(A,\omega )\) as r 0. The possible range of γ to admit a non-trivial limit will then be examined in the cases that \((\mathcal {E}.\mathcal {F})\) equals \((\frac 12{\textbf {D}}_{\mathbb {C}},H^{1}(\mathbb {C}))\) and \((\textbf {a},H^{1}(\mathbb {C}))\), where a corresponds to the uniformly elliptic partial differential operator of divergence form.

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.

Similar content being viewed by others

References

  1. Barlow, M.T., Grigor’yan, A., Kumagai, T.: On the equivalence of parabolic Harnack inequalities and heat kernel estimates. J. Math. Soc. Japan 64, 1091–1146 (2012)

    Article  MathSciNet  Google Scholar 

  2. Berestycki, N.: An elementary approach to Gaussian multiplicative chaos. Electro. Commun. Probab 22(27), 1–12 (2017)

    MathSciNet  MATH  Google Scholar 

  3. Beurling, A., Deny, J.: Dirichlet spaces. Proc. Nat. Acad. Sci. U.S.A. 45, 208–215 (1959)

    Article  MathSciNet  Google Scholar 

  4. Cameron, R.H., Martin, W.T.: Transformations of Wiener integrals under translations. Annals Math. 45, 386–396 (1944)

    Article  MathSciNet  Google Scholar 

  5. Chen, Z.Q., Fukushima, M.: Symmetric markov processes, time change and boundary theory, Princeton university press (2011)

  6. Deny, J.: Méthodes hilbertiennes Et Théorie Du Potentiel. In: Potential Theory, Centro Internationale Mathematico Estivo Edizioni Cremonese Roma (1970)

  7. Doob, J.L: Stochastic Processes Wiley (1953)

  8. Duplantier, B., Sheffield, S.: Liouville quantum gravity and KPZ. Invent. Math. 185, 333–393 (2011)

    Article  MathSciNet  Google Scholar 

  9. Dynkin, E.B: Theory of Markov Processes Pergamon Press (1960)

  10. Dynkin, E.B.: Processes, Markov Vol II Springer (1965)

  11. Friedman, A.: Partial Differential Equations of Parabolic Type Prentice-Hall Inc (1964)

  12. Fukushima, M.: Dirichlet forms and markov processes North-Holland/Kodansha (1980)

  13. Fukushima, M.: Logarithmic and linear potentials of signed measures and Markov property of associated Gaussian fields. Potential Anal 49, 359–379 (2018)

    Article  MathSciNet  Google Scholar 

  14. Fukushima, M., Oshima, Y.: Recurrent Dirichlet forms and Markov property of associated Gaussian fields. Potential Anal. 49, 609–633 (2018)

    Article  MathSciNet  Google Scholar 

  15. Fukushima, M., Oshima, Y., Takeda, M.: Dirichlet Forms and Symmetric Markov Processes, de Gruyter, 1994 2nd revised Edition (2010)

  16. Ikeda, N., Watanabe, S.: Stochastic Differential Equations and Diffusion Processes north Holland/Kodansha (1980)

  17. Itô, K.: Probability Theory in Japanese Iwanami-Shoten (1953)

  18. Jones, P.W.: Quasiconformal mappings and extendability of functions in Sobolev spaces. Acta Math. 147, 71–78 (1981)

    Article  MathSciNet  Google Scholar 

  19. Kahane, J. -P.: Sur le chaos multiplicatif. Ann.Sci Math. Québec 9(2), 105–150 (1985)

    MathSciNet  MATH  Google Scholar 

  20. Littman, W., Stampacchia, G., Weinberger, H.F.: Regular points for elliptic equations with discontinuous coefficients. Ann. Scuola Norm.Sup.Pisa III 17, 43–77 (1963)

    MathSciNet  MATH  Google Scholar 

  21. Mandrekar, V.S., Gawarecki, L.: Stochastic analysis for gaussian random processes and fields CRC press (2015)

  22. Marcus, M.B., Rosen, J.: Markov Processes Gaussian processes and local times Cambridge Univ. Press (2006)

  23. McKean, H.P.: Brownian motion with a several-dimensional time. Theory Probab. Appl. 8, 335–354 (1963)

    Article  MathSciNet  Google Scholar 

  24. Oshima, Y.: On the equilibrium measure of recurrent Markov processes, Osaka. J. Math. 15, 283–310 (1978)

    MathSciNet  MATH  Google Scholar 

  25. Pitt, L.D.: A Markov property for Gaussian processes with a multidimensional parameter (1971)

  26. Port, S.C., Stone, C.J.: Brownian motion and classical potential theory academic press (1978)

  27. Rhodes, R., Vargas, V.: Gaussian multiplicative chaos and appliacations; A review. Probab. Surv. 11, 315–392 (2014)

    Article  MathSciNet  Google Scholar 

  28. Röckner, M.: Generalized Markov fields, Dirichlet forms. Acta. Appl. Math 3, 285–311 (1985)

    Article  MathSciNet  Google Scholar 

  29. Sheffield, S.: Conformal weldings of random surfaces: SLE and the quantum gravity zipper. Ann. probab. 44, 3474–3545 (2016)

    MathSciNet  MATH  Google Scholar 

  30. Silverstein, M.L.: Symmetric Markov Processes, Lecture Notes in Math, vol. 426. Springer, Berlin Heidelberg (1974)

    Book  Google Scholar 

  31. De La Vallée Poussin, Ch.-J: Le Potentiel Logarithmique Gauthier-Villars (1949)

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Yoichi Oshima.

Additional information

Publisher’s Note

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

This work was supported by Research Institute for Mathematical Science, A joint Usage/Research Center located in Kyoto University

Appendices

Appendix: Proof of Proposition 4.7

1.1 Proof of Proposition 4.7 (i)

Using the function \(\check r_{1}(\textbf {x},\textbf {y}),\textbf {x},\textbf {y}\in F,\) in Eq. 4.13, define

$$\check {r_{1}^{1}}(\textbf{x},\textbf{y})=\check r_{1}(\textbf{x},\textbf{y}),\quad \check {r_{1}^{n}}(\textbf{x},\textbf{y})=\int \check r_{1}(\textbf{x},\textbf{z})\check r_{1}^{n-1}(\textbf{z},\textbf{y})m_{F}(d\textbf{z}),\ n\ge 2.$$

\(\check {r_{1}^{n}}(\textbf {x},\textbf {y})\) is the density function of the kernel \(\check R^{n}(\textbf {x},d\textbf {y})\) on \((F,{\mathcal B}(E))\) with respect to mF. \(\check {R_{1}^{n}}(\textbf {x},d\textbf {y})\) is mF-symmetric and \(\check {R_{1}^{n}}1_{F}(\textbf {x})=1, \textbf {x}\in F,\) so that \(\widetilde m_{F}\check {R_{1}^{n}}=\widetilde m_{F}.\) Consequently,

$$\check {R_{1}^{n}}(\textbf{x},A)-\widetilde m_{F}(A)={\int}_{F} [\check {R_{1}^{n}}(\textbf{x},A)-\check {R_{1}^{n}}(\mathbf{x}^{\prime},A)]\widetilde m_{F}(d\mathbf{x}^{\prime}),\quad A\in {\mathcal B}(F).$$

Denote by ||μ|| the total variation of a signed measure μ on F, We then get from the above identity and an estimate [14, (3.4)]

$$ \sup_{\textbf{x}\in F}||\check {R_{1}^{n}}(\textbf{x},\cdot)-\widetilde m_{F}(\cdot)||\le 2\gamma^{n},\quad \text{for some constant}\ \gamma\in (0,1). $$
(7.1)

Therefore, if we let

$$ \check r^{(\pm)}(\textbf{x},A)={\int}_{A} {\sum}_{n=1}^{\infty} (\check {r_{1}^{n}}(\textbf{x},\textbf{y})-1/m(F))^{\pm} m_{F}(d\textbf{y}),\ \textbf{x}\in F,\ A\in {\mathcal B}(F), $$
(7.2)

then, \(\check r^{(+)}(\textbf {x},A),\ \check r^{(-)}(\textbf {x},A)\) are positive kernels on \((F,{\mathcal B}(F))\) satisfying \(\sup _{\textbf {x}\in F}\check r^{(\pm )}(\textbf {x},F)<\infty \) and, for any \(\varphi \in L^{\infty }(F;m_{F})\),

$$ \check R\varphi(\textbf{x})={\int}_{F}\check r^{(+)}(\textbf{x},d\textbf{y})\varphi(\textbf{y}) - {\int}_{F}\check r^{(-)}(\textbf{x},d\textbf{y})\varphi(\textbf{y})\ \text{for}\ m\text{-a.e.}\ \textbf{x}\in F. $$
(7.3)

on account of Eq. 4.14. This identity can be readily verified to hold also for φL2(F; mF).

Define \(r^{(\pm )}(\textbf {x},\textbf {y})={\int \limits }_{F}{\int \limits }_{F\times F} r^{g}(\textbf {x},\textbf {z}) {\check r}^{(\pm )}(\textbf {z}, d\textbf {w}) r^{g}(\textbf {w},\textbf {y}) m_{F}(d\textbf {z})\) for \(\textbf {x},\textbf {y}\in \mathbb {C}\), r(±)(x, y) are symmetric and \(\mathbb {M}^{g}\)-excessive for each variable x and y. \({\int \limits }_{\mathbb {C}} r^{(+)}(\textbf {x},\textbf {y})h(\textbf {y})m(d\textbf {y})\) is finite for each \(\textbf {x}\in \mathbb {C}\) for any non-negative bounded Borel function h on \(\mathbb {C}\) vanishing outside a bounded set, because Rgh is bounded on \(\mathbb {C}\) by Lemma 3.1 and so \(\psi (\textbf {z})={\int \limits }_{F} \check r^{(+)}(\textbf {z},d\textbf {w})R^{g}h(\textbf {w})\) is bounded on F by a constant C > 0, and furthermore \({\int \limits }_{\mathbb {C}} r^{(+)}(\textbf {x},\textbf {y})h(\textbf {y})m(d\textbf {y})= R^{g}(1_{F}\cdot \psi )(\textbf {x})\le C R^{g}g(\textbf {x})=C\) in view of [14, (3.28)]. Therefore r(±)(x, y) is finite for m-a.e.y and hence q.e.\(\textbf {y}\in \mathbb {C}\).

We see from Eq. 4.14 that, for φL2(F; mF), \(\check R\varphi =\check R_{1}\varphi -{\langle }\widetilde m_{F},\varphi \>+ \check R_{1}\check R\varphi \). Consider any \(\mu \in {\mathcal S}_{0}^{g,(0)}\) with \(\mu (\mathbb {C})<\infty .\) Since \(R^{g}\mu \in L^{2}(\mathbb {C},m_{F})\) and \({\langle }\widetilde m_{F},R^{g}\mu \>=\frac {1}{m(F)}{\langle }R^{g}g, \mu \>= \mu (\mathbb {C})/m(F),\) we have

$$ H_{F}\check R(1_{F}R^{g}\mu)(\textbf{x})=R^{g}(1_{F}R^{g}\mu)(\textbf{x})-\mu(\mathbb{C})/m(F) + R^{g}(1_{F}\check R(1_{F}R^{g}\mu))(\textbf{x}),\quad \textbf{x}\in \mathbb{C}, $$

which combined with Eqs. 4.12 and 7.3 implies that Rμ admits an expression Eq. 4.26 by a kernel \(\frak r(\textbf {x},\textbf {y})\) defined by

$$ \begin{array}{@{}rcl@{}} \frak r(\textbf{x},\textbf{y})&=&{\int}_{F} r^{g}(\textbf{x},\textbf{z})r^{g}(\textbf{z},\textbf{y})m_{F}(d\textbf{z})+ r^{(+)}(\textbf{x},\textbf{y})-r^{(-)}(\textbf{x},\textbf{y})\\&&+r^{g}(\textbf{x},\textbf{y})-2/m(F),\quad \textbf{x}, \textbf{y}\in \mathbb{C}. \end{array} $$
(7.4)

\(\frak r(\textbf {x},\textbf {y})\) is symmetric and, for each \(\textbf {x}\in \mathbb {C},\) it is a difference of Mg-excessive functions finite for q.e. \(\textbf {y}\in \mathbb {C}.\) This property for the first term of the righthand side can be verified in a similar way to the proof for other terms given previously.

Proof of Proposition 4.7 (ii)

We take x, yB(S − 1) with |xy| > η. Since there exists a constant M2 such that \(|\widehat R\mu ^{\textbf {y},r_{2}}|\leq M_{2}\) on B(x, η/8) for any r2 < η/8 by Lemma 4.6, the stated uniform boundedness of \({\langle }\mu ^{\textbf {x},r_{1}}, R\mu ^{\textbf {y},r_{2}}\> ={\langle }\mu ^{\textbf {x},r_{1}}, \widehat R\mu ^{\textbf {y},r_{2}}\>\) holds true. To prove Eq. 4.27, we first show that

$$ \lim_{r_{1},r_{2}\downarrow 0} {\langle}\mu^{\textbf{x},r_{1}}, R^{g}\mu^{\textbf{y},r_{2}}\>= r^{g}(\textbf{x},\textbf{y}), $$
(7.5)

for m × m-a.e. (x, y) ∈ B(S − 1) × B(S − 1) ∩{(x, y) : |xy| > η}.

Since \(e^{-t} p_{t}(\textbf {z},\textbf {w})\leq {p^{g}_{t}}(\textbf {z},\textbf {w})\leq p_{t}(\textbf {z},\textbf {w})\), we can use Eq. 4.1 to find for any ε > 0 a positive t0 satisfying

$$ {{\int}_{0}^{t}} {p^{g}_{s}}(\textbf{z},\textbf{w})ds \leq {{\int}_{0}^{t}} {K_{2}\over s} e^{-9 k_{2} \eta^{2}/16s}ds <{\varepsilon} $$
(7.6)

for any tt0 and z, wE such that |zw| > 3η/4. In particular, \({{\int \limits }_{0}^{t}} {\langle }\mu ^{\textbf {x},r_{1}}, {P^{g}_{s}}\mu ^{\textbf {y},r_{2}}\>ds<{\varepsilon }\).

Let M1 be a constant satisfying \(R^{g} \mu ^{\textbf {y},r_{2}}\leq M_{1}\) on \(\mathbb {C}\setminus B(\textbf {y},\eta /2)\) for all r2η/8. Such constant M1 exists by Lemma 4.6. By Eq. 4.1 and the tail estimate Eq. 4.32, we may assume that, by taking smaller t0 > 0 if necessary,

$$ {\int}_{\mathbb{C}\setminus B(\textbf{x},\eta/2)}{p^{g}_{t}}(\textbf{z},\textbf{w})m(d\textbf{w})\leq {\int}_{\{|\textbf{w}|>3\eta/8\}} {K_{2}\over t} e^{-k_{2}|\textbf{w}|^{2}/t} d\textbf{w} \leq K_{2}\sqrt{\pi/k_{2}t} e^{-9\eta^{2}k_{2}/64t}< {{\varepsilon}\over M_{1}} $$

for all tt0 and zB(x, η/8). In particular,

$$ \begin{array}{@{}rcl@{}} &&{\langle}\mu^{\textbf{x},r_{1}} {P^{g}_{t}}, 1_{\mathbb{C}\setminus B(S-1/2)} R^{g} \mu^{\textbf{y},r_{2}}\>\\ &&=\int\int {p^{g}_{t}}(\textbf{z},\textbf{w}) 1_{\mathbb{C}\setminus B(S-1/2)}(\textbf{w}) R^{g}\mu^{\textbf{y},r_{2}}(\textbf{w})m(d\textbf{w})\mu^{\textbf{x},r_{1}}(d\textbf{z})\\ &&\leq M_{1} {\int}_{B(\textbf{x},\eta/8)} \mu^{\textbf{x},r_{1}}(d\textbf{z}){\int}_{\mathbb{C}\setminus B(\textbf{x},\eta/2)}{p^{g}_{t}}(\textbf{z},\textbf{w})m(d\textbf{w})<{\varepsilon}. \end{array} $$
(7.7)

Put \(D(\textbf {y})=B(S-1/2)\setminus \overline {B(\textbf {y},\eta /2)}\). Since \({p^{g}_{t}}(\textbf {z},\textbf {w})\leq (K_{2}/t)e^{-9k_{2}\eta ^{2}/64t}\) for any zB(x, η/8) and wB(y, η/2) and Rg1B(S− 1)M4 on \(\mathbb {C}\) for some constant M4 by Lemma 3.1 (i),

$$ {\langle}\mu^{\textbf{x},r_{1}} {P^{g}_{t}}, 1_{B(\textbf{y},\eta/2)} \cdot R^{g}\mu^{\textbf{y},r_{2}}\> \!\leq\! {K_{2}\over t} e^{-9k_{2}\eta^{2}/64 t} {\langle}\mu^{\textbf{y},r_{2}},R^{g} 1_{B(\textbf{y},\eta/2)}\> \!\leq\! {K_{2} M_{4}\over t} e^{-9k_{2}\eta^{2}/64t}\!<\!{\varepsilon} $$
(7.8)

for any t < t0 by taking smaller t0 if necessary. Further, since the distance between F and B(x, η/8) exceeds 1/2, we get by putting \(A_{t}={{\int \limits }_{0}^{t}} 1_{F}(X_{s})ds\),

$$ \begin{array}{@{}rcl@{}} P_{t_{0}}(\textbf{z}, D(\textbf{y}))-P^{g}_{t_{0}}(\textbf{z},D(\textbf{y}))&=&{\mathbb E}_{\textbf{z}}\left[(1-e^{-A_{t_{0}}})1_{D(\textbf{y})}(X_{t_{0}})\right]\leq {\mathbb E}_{\textbf{z}}\left[{\int}_{0}^{t_{0}} 1_{F}(X_{s}) ds\right]\\ &\leq& {\int}_{0}^{t_{0}} {K_{2} m(F)\over s} e^{-k_{2}/4s}ds \quad\text{for any}\ \textbf{z}\in B(\textbf{x},\eta/8). \end{array} $$

Hence we may also assume that

$$ {\langle}\mu^{\textbf{x},r_{1}}, (P_{t}-{P^{g}_{t}})(1_{D} R^{g}\mu^{\textbf{y},r_{2}})\><{{\varepsilon}} \quad\text{for all}\ t\leq t_{0}, $$
(7.9)

because \(R^{g} \mu ^{\textbf {x},r_{2}}\le M_{1}\) on D(y).

Therefore, in the decomposition

$$ \begin{array}{@{}rcl@{}} {\langle}\mu^{\textbf{x},r_{1}}, R^{g} \mu^{\textbf{y},r_{2}}\>&=&{{\int}_{0}^{t}}{\langle}\mu^{\textbf{x},r_{1}}, {P^{g}_{s}}\mu^{\textbf{y},r_{2}}\>ds+{\langle}\mu^{\textbf{x},r_{1}} ,{P^{g}_{t}} R^{g}\mu^{\textbf{y},r_{2}}\>\\ &=&{{\int}_{0}^{t}}{\langle}\mu^{\textbf{x},r_{1}}, {P^{g}_{s}}\mu^{\textbf{y},r_{2}}\>ds+{\langle}\mu^{\textbf{x},r_{1}} {P^{g}_{t}}, 1_{\mathbb{C}\setminus B(S-1/2)} \cdot R^{g}\mu^{\textbf{y},r_{2}}\>\\ &&+{\langle}\mu^{\textbf{x},r_{1}} {P^{g}_{t}}, 1_{B(\textbf{y},\eta/2)} \cdot R^{g}\mu^{\textbf{y},r_{2}}\>+{\langle}\mu^{\textbf{x},r_{1}}, ({P^{g}_{t}}-P_{t})(1_{D(\textbf{y})} R^{g}\mu^{\textbf{y},r_{2}})\>\\ &&+{\langle}\mu^{\textbf{x},r_{1}},P_{t}(1_{D(\textbf{y})} R^{g}\mu^{\textbf{y},r_{2}})\>, \end{array} $$

the sum of the first four terms of the righthand side is smaller than 4ε for any r1, r2 ∈ (0, η/8) and tt0.

Since pt(z, w) is uniformly continuous relative to (z, w) on \(\overline {B(\textbf {x},\eta /8)}\times \overline {D}(\textbf {y})\), by putting \(\delta (t,r_{1})=\sup \{|p_{t}(\textbf {z},\textbf {w})-p_{t}(\textbf {x},\textbf {w})|:\textbf {z}\in \overline {B(\textbf {x},r_{1})},\textbf {w}\in \overline {D}(\textbf {y})\}\), we can see that the difference of the last term of the righthand side and \({\int \limits }_{D(\textbf {y})} p_{t}(\textbf {x},\textbf {w}) R^{g}\mu ^{\textbf {y},r_{2}}(\textbf {w})m(d\textbf {w})\) is smaller than M1δ(t, r1) which converges to zero as r1 0 for each t < t0. Furthermore, for \(f^{\textbf {x}}_{t}(\textbf {w})=1_{D(\textbf {y})}(\textbf {w})p_{t}(\textbf {x},\textbf {w})\), \(R^{g} f_{t}^{\textbf {x}}\) is \(\mathcal {E}\)-harmonic on B(y, η/8) by Lemma 3.1 and continuous there as in the proof of Lemma 4.4. Consequently, \( \lim _{r_{2}\to 0}{\int \limits }_{D(\textbf {y})} p_{t}(\textbf {x},\textbf {w}) R^{g}\mu ^{\textbf {y},r_{2}}(\textbf {w})m(d\textbf {w})=\lim _{r_{2}\to 0}{\langle }\mu ^{\textbf {y},r_{2}}, R^{g} f_{t}^{\textbf {x}}\>=R^{g} f_{t}^{\textbf {x}}(\textbf {y}). \) Accordingly

$$ \limsup_{r_{1},r_{2}\downarrow 0} |{\langle}\mu^{\textbf{x},r_{1}}, R^{g}\mu^{\textbf{y},r_{2}}\>-R^{g} f_{t}^{\textbf{x}}(\textbf{y})|<4{\varepsilon} $$
(7.10)

for any tt0 and any x, yB(S − 1) with |xy| > η.

Thus, to verify Eq. 7.5, it suffices to show that \(\lim _{t\to 0} R^{g} f_{t}^{\textbf {x}}(\textbf {y})=\lim _{t\to 0} P_{t}(1_{D(\textbf {y})}\cdot r^{g}(\cdot ,\textbf {y}))(\textbf {x})=r^{g}(\textbf {x},\textbf {y})\) for m × m-a.e.(x, y) ∈ B(S − 1) × B(S − 1) ∩{|xy| > η}. For any yB(S − 1), let \(E_{1}(\textbf {y})=\{\textbf {x}:r^{g}(\textbf {x},\textbf {y})<\infty \}.\) As \(\mathbb {C}\setminus E_{1}(\textbf {y})\) is polar and

$$ {P^{g}_{t}}(1_{D(\textbf{y})}r^{g}(\cdot,\textbf{y}))(\textbf{x})\leq\ P_{t}(1_{D(\textbf{y})}r^{g}(\cdot,\textbf{y}))(\textbf{x})\leq e^{t} {P^{g}_{t}}(1_{D(\textbf{y})}r^{g}(\cdot,\textbf{y}))(\textbf{x}), $$

it is enough to show that \(\lim _{t\to 0} {P^{g}_{t}}(1_{D(\textbf {y})}\cdot r^{g}(\cdot ,\textbf {y}))(\textbf {x})=r^{g}(\textbf {x},\textbf {y})\) for any xD(y) ∩ E1(y). Since rg(⋅, y) is \(\mathbb {M}^{g}\)-excessive and 1D(y)(Xt)rg(Xt, y) is right continuous at t = 0 a.s.ℙx for xD(y) ∩ E1(y), we have

$$ \begin{array}{@{}rcl@{}} r^{g}(\textbf{x},\textbf{y})\wedge n&=&\lim_{t\to 0} {\mathbb E}_{\textbf{x}}^{g}\left[1_{D(\textbf{y})}(X_{t}) r^{g}(X_{t},\textbf{y})\wedge n\right] \leq \mathop{\underline{\lim}}_{t\to 0} {\mathbb E}_{\textbf{x}}^{g}\left[1_{D(\textbf{y})}(X_{t}) r^{g}(X_{t},\textbf{y})\right]\\ &\leq&\mathop{\overline{\lim}}_{t\to 0} {\mathbb E}_{\textbf{x}}^{g}\left[1_{D(\textbf{y})}(X_{t}) r^{g}(X_{t},\textbf{y})\right] \leq\lim_{t\to 0} {\mathbb E}_{\textbf{x}}^{g}\left[r^{g}(X_{t},\textbf{y})\right]=r^{g}(\textbf{x},\textbf{y}),\ \ n\ge 1. \end{array} $$

By letting \(n\to \infty \), we arrive at Eq. 7.5.

We shall next show that, for the kernels r+(x, y) and r(x, y) appearing in the proof of Proposition 4.7 (i),

$$ \lim_{r_{1},r_{2}\downarrow 0} {\langle}\mu^{\textbf{x},r_{1}}, R^{(\pm)}\mu^{\textbf{y},r_{2}}\>= r^{(\pm)}(\textbf{x},\textbf{y}), $$
(7.11)

for m × m-a.e. (x, y) ∈ B(S − 1) × B(S − 1). Here we let \(R^{(\pm )}\mu (\textbf {x})={\int \limits }_{\mathbb {C}} r^{(\pm )}(\textbf {x},\textbf {z})\mu (d\textbf {z}),\ \textbf {x}\in \mathbb {C}.\) Consider the function on \(\mathbb {C}\) defined by \(\textbf {x}i^{\textbf {y},r_{2}}_{+}(\textbf {z})=1_{F}(\textbf {z}){\check R}^{(+)}(1_{F}R^{g}\mu ^{\textbf {y},r_{2}})(\textbf {z}),\ \textbf {z}\in \mathbb {C}.\) Since \(R^{g}\mu ^{\textbf {y},r_{2}}(\textbf {z})\) is bounded in zF and r2 by Lemma 4.6 (i) and \(\check {R}^{(+)}\) is a bounded linear operator on \(L^{\infty }(F;m_{F})\), there exists a constant M > 0 such that for any \(\textbf {z}\in \mathbb {C}, r_{2}\in (0,\eta /8)\),

$$ \textbf{x}i^{\textbf{y},r_{2}}_{+}(\textbf{z})\le M,\ R^{g}\textbf{x}i_{+}^{\textbf{y},r_{2}}(\textbf{z})={\int}_{F} r^{g}(\textbf{z},\textbf{w})\textbf{x}i_{+}^{\textbf{y},r_{2}}(\textbf{w})m(d\textbf{w})=H_{F}(R^{g}\textbf{x}i_{+}^{\textbf{y},r_{2}})(\textbf{z})\le M. $$
(7.12)

In view of definition, we have the identity \(R^{(+)} \mu ^{\textbf {y},r_{2}}=R^{g} \textbf {x}i_{+}^{\textbf {y},r_{2}}\). Accordingly, as in the previous proof of Eq. 7.5, we can decompose \({\langle }\mu ^{\textbf {x},r_{1}}, R^{(+)}\mu ^{\textbf {y},r_{2}}\>\) as

$$ \begin{array}{@{}rcl@{}} {\langle}\mu^{\textbf{x},r_{1}},R^{(+)}\mu^{\textbf{y},r_{2}}\> &=&{{\int}_{0}^{t}}{\langle}\mu^{\textbf{x},r_{1}}, {P^{g}_{s}}\textbf{x}i_{+}^{\textbf{y},r_{2}}\>ds+{\langle}\mu^{\textbf{x},r_{1}} {P^{g}_{t}}, 1_{\mathbb{C}\setminus B(S-1/2)} \cdot R^{g}\textbf{x}i_{+}^{\textbf{y},r_{2}}\>\\ &&+{\langle}\mu^{\textbf{x},r_{1}}, ({P^{g}_{t}}-P_{t})(1_{B(S-1/2)} R^{g}\textbf{x}i_{+}^{\textbf{y},r_{2}})\> \\&&+{\langle}\mu^{\textbf{x},r_{1}},P_{t}(1_{B(S-1/2)} R^{g}\textbf{x}i_{+}^{\textbf{y},r_{2}})\>. \end{array} $$

For any ε > 0, we can take t1 such that the first term of the righthand side is less than ε for any t ∈ (0, t1) as Eq. 7.6 because of dist(F, B(x, r1)) > 1/2 and the bound Eq. 7.12. Because also of the bound Eq. 7.12, we can take t1 such that the second term is less than ε for any t ∈ (0, t1) as Eq. 7.7. Further, as Eq. 7.9, we may suppose that the third term is less than ε for all tt1.

Since \(\sup \{|p_{t}(\textbf {z},\textbf {w})-p_{t}(\textbf {x},\textbf {w})|:\textbf {z}\in B(\textbf {x},r_{1}), \textbf {w}\in B(S-1/2)\}\to 0\) as r1 → 0, \(\lim _{r_{1}\to 0} {\langle }\mu ^{\textbf {x},r_{1}},P_{t}(1_{B(S-1/2)}R^{g} \textbf {x}i^{\textbf {y},r_{2}}_{+})\> = P_{t}(1_{B(S-1/2)}R^{g} \textbf {x}i^{\textbf {y},r_{2}}_{+})(\textbf {x})\) uniformly in r2η/8. Put \(h^{\textbf {x}}_{t}(\textbf {w})=1_{F}(\textbf {w}){\check R}^{(+)}R^{g} (1_{B(S-1/2)} p_{t}(\cdot , \textbf {x}))(\textbf {w})\). Since \(h^{\textbf {x}}_{t}\) vanishes outside of F, we can see as before that \(R^{g}h^{\textbf {x}}_{t}(\textbf {w})\) is continuous on B(S − 1) and consequently

$$ \lim_{r_{2}\to 0} P_{t}(1_{B(S-1/2)}R^{g} \textbf{x}i^{\textbf{y},r_{2}}_{+})(\textbf{x})=\lim_{r_{2}\to 0}{\langle}\mu^{\textbf{y},r_{2}}, R^{g} h^{\textbf{x}}_{t}\>=R^{g} h^{\textbf{x}}_{t}(\textbf{y}). $$

Therefore, as Eq. 7.10, \( \limsup _{r_{1},r_{2}\downarrow 0} |{\langle }\mu ^{\textbf {x},r_{1}},R^{(+)}\mu ^{\textbf {y},r_{2}}\>- R^{g}h_{t}^{\textbf {x}}(\textbf {y})|<3{\varepsilon } \) for any tt1.

As \(R^{g} h^{\textbf {x}}_{t}(\textbf {y})=R^{(+)}(1_{B(S-1/2)} p_{t}(\cdot ,\textbf {x}))(\textbf {y})=P_{t} (1_{B(S-1/2)}\cdot r^{(+)}(\cdot ,\textbf {y}))(\textbf {x}),\) and r(+)(⋅, y) is \(\mathbb {M}^{g}\)-excessive and finite q.e., we obtain similarly to the above proof of Eq. 7.5, that \(\lim _{t\to 0} R^{g} h^{\textbf {x}}_{t}(\textbf {y})=r^{(+)}(\textbf {x},\textbf {y})\) for q.e.xB(S − 1) for each yB(S − 1), and consequently, the validity of Eq. 7.11 for R(+) and r(+). In the same way Eq. 7.11 for R(−) and r(−) is valid.

It remains to prove

$$ \lim_{r_{1},r_{2}\downarrow 0} {\langle}\mu^{\textbf{x},r_{1}}, Q\mu^{\textbf{y},r_{2}}\>= q(\textbf{x},\textbf{y}), $$
(7.13)

for m × m-a.e. (x, y) ∈ B(S − 1) × B(S − 1). Here q(x, y) is the first term of the righthand side of Eq. 7.4 and \(Q\mu (\textbf {x})={\int \limits }_{\mathbb {C}} q(\textbf {x},\textbf {z})\mu (z), \textbf {z}\in \mathbb {C}.\) But this can be shown in exactly the same way as the proof of Eq. 7.11 using \(1_{F}(\textbf {z})R^{g}\mu ^{\textbf {y},r_{2}}(\textbf {z})\) in place of \(\textbf {x}i_{+}^{\textbf {y},r_{2}}(\textbf {z}).\)

Appendix: Proof of Proposition 5.4

Assume that (aij(x)) is a family of C1 functions on \(\mathbb {C}\) with Hölder continuous derivative satisfying Eq. 1.3. Let \(b_{i}(\textbf {x})={\sum }_{i,j=1}^{2} \partial a_{ij}(\textbf {x})/\partial x_{j}\) and L be the infinitesimal generator corresponding to the form a:

$$Lu(\textbf{x})={\sum}_{i,j=1}^{2} {\partial\over \partial x_{i}}\left( a_{ij}(\textbf{x}){\partial u\over \partial x_{j}}\right) ={\sum}_{i,j=1}^{2} a_{ij}(\textbf{x}){\partial^{2} u\over \partial x_{i}\partial x_{j}}+{\sum}_{i=1}^{2} b_{i}(\textbf{x}){\partial u\over \partial x_{i}}. $$

Let us fix an open disk G containing \(\overline {B(S+1)}.\) A function Γ(x, y) is said to be a fundamental solution of L on G if it satisfies − LΓ(x, y) = δ(xy) weakly, that is, for all \(u\in {C^{1}_{c}}(G)\),

$$ {\int}_{G}{\sum}_{i,j=1}^{2} a_{ij}(\textbf{x}) {\partial u(\textbf{x})\over \partial x_{i}}{\partial {\Gamma}(\textbf{x},\textbf{y})\over \partial x_{j}} d\textbf{x}=u(\textbf{y}), \quad \forall \textbf{y}\in G. $$
(7.14)

For any fixed yG, let \( L_{0}u(\textbf {x})={\sum }_{i,j=1}^{2} a_{ij}(\textbf {y}){\partial ^{2} u\over \partial x_{i}\partial x_{j}}. \) Then Γ0(x, y) defined by Eq. 5.19 is a fundamental solution of L0 on G. We shall briefly describe a construction of a fundamental solution of L from the parametrix Γ0(x, y) as is stated in [11, §5.6] under the condition that the coefficients of L are Hölder continuous.

Since \(a_{ij}\in {C^{1}_{b}}(G)\), the function k0(x, y) = (LL00(x, y) satisfies, for some constant K1 > 0, |k0(x, y)|≤ K1/|xy|, ∀x, yG. Define \(k^{(n)}_{0}(\textbf {x},\textbf {y})\) by \(k^{(1)}_{0}(\textbf {x},\textbf {y})=k_{0}(\textbf {x},\textbf {y})\) and \(k^{(n)}_{0}(\textbf {x},\textbf {y})={\int \limits }_{G}k_{0}(\textbf {x},\textbf {z}) k^{(n-1)}_{0}(\textbf {z},\textbf {y})d\textbf {z}\). Then \(|k^{(2)}_{0}(\textbf {x},\textbf {y})|\leq K_{2} \log (1/|\textbf {x}-\textbf {y}|)+K_{3}\) and \(|k^{(3)}_{0}(\textbf {x},\textbf {y})|\leq K_{4}\) for some constants K2, K3 and K4. Put \(K^{(n)}_{0}f(\textbf {x})={\int \limits }_{G} k^{(n)}_{0}(\textbf {x},\textbf {y})f(\textbf {y})d\textbf {y}\).

A fundamental solution Γ(x, y) of L on G can be constructed by

$$ {\Gamma}(\textbf{x},\textbf{y})={\Gamma}_{0}(\textbf{x},\textbf{y})+{\int}_{G} {\Gamma}_{0}(\textbf{x},\textbf{z}) {\Phi}(\textbf{z},\textbf{y})d\textbf{z}+\sum \alpha_{i}(\textbf{x}){\upbeta}_{i}(\textbf{y}) $$
(7.15)

for suitable continuous functions Φ(x, y), αi(x) and βi(y). In order to make Γ to satisfy − LΓ(x, y) = δ(xy), Φ(x, y) needs to be a solution of the following Fredholm integral equation.

$$ {\Phi}(\textbf{x},\textbf{y})=k_{0}(\textbf{x},\textbf{y})+{\int}_{G} k_{0}(\textbf{x},\textbf{z}){\Phi}(\textbf{z},\textbf{y})d\textbf{z}+\sum L\alpha_{i}(\textbf{x}){\upbeta}_{i}(\textbf{y}). $$
(7.16)

Note that \(k^{(n)}_{0}(\textbf {x},\textbf {y})\) is continuous on G for any n ≥ 3. Let us take a continuous function

$$ g(\textbf{x},\textbf{y})= k_{0}^{(4)}(\textbf{x},\textbf{y})+k_{0}^{(5)}(\textbf{x},\textbf{y})+k_{0}^{(6)}(\textbf{x},\textbf{y}) + \sum (K_{0}^{(3)}+K_{0}^{(4)}+K_{0}^{(5)})(L\alpha_{i})(\textbf{x}){\upbeta}_{i}(\textbf{y}). $$

Here αi = βi = 0 for all i if λ = 1 is not an eigenvalue of the dual operator \((K^{*}_{0})^{(3)}\) on Cb(G) of \(K^{(3)}_{0}\) defined by \((K_{0}^{*})^{(3)}f(\textbf {x})={\int \limits } (k_{0}^{*})^{(3)}(\textbf {x},\textbf {y})f(\textbf {y})d\textbf {y}\) with \(k_{0}^{*}(\textbf {x},\textbf {y})=k_{0}(\textbf {y},\textbf {x})\), while, if λ = 1 is an eigenvalue, then αii are chosen to satisfy (g(⋅, y), ψj) = 0 for all eigenfunctions {ψj} corresponding to the eigenvalue λ = 1 of \((K^{*}_{0})^{(3)}\). Then the Fredholm equation \( w(\textbf {x},\textbf {y})=K^{(3)}_{0} w(\textbf {x},\textbf {y})+g(\textbf {x},\textbf {y}) \) has a unique continuous solution w(x, y) for any yG. Using this solution, the unique solution of Eq. 7.16 is given by \( {\Phi }(\textbf {x},\textbf {y})=k_{0}(\textbf {x},\textbf {y})+k^{(2)}_{0}(\textbf {x},\textbf {y})+k^{(3)}_{0}(\textbf {x},\textbf {y})+w(\textbf {x},\textbf {y}). \) We notice that, according to the construction of Γ from Γ0 by Eq. 7.15,

$$ {\Gamma}(\textbf{y},\textbf{z})-{\Gamma}_{0}(\textbf{y},\textbf{z})\text{is bounded in}\ (\textbf{y},\textbf{z})\in G\times G. $$
(7.17)

We now proceed to a proof of Eq. 4.19 with κ = Λ/λ. For xB(S − 1) and 0 < 5rt ≤ 1/3, let μx, r be the equilibrium measure for \(\overline {B(\textbf {x},r)}\) relative to the admissible set \(F=\overline {B(S+1)}\setminus B(S)\) for the Dirichlet form a on \(H^{1}(\mathbb {C})\). We first show that the logarithmic potential

$$U\mu^{\textbf{x},r}(\textbf{y})={1\over \pi}\int \log{1\over |\textbf{y}-\textbf{z}|}\mu^{\textbf{x},r}(d\textbf{z}),\ \textbf{y} \in \mathbb{C},$$

of μx, r has the properties

$$ {\langle}\mu^{\textbf{x},r},U\mu^{\textbf{x},r}\> <\infty\quad \text{and}\quad U\mu^{\textbf{x},r}\in L^{2}_{\text{loc}}(\mathbb{C}). $$
(7.18)

Since μx, r is a measure of 0-order finite energy for the perturbed form ag of a by g = 1F, so it is for the perturbed Dirichlet integral (1/2)D(u, u) + (u, u)g.

Denote by \(\acute {\mathbb {M}}\) the planar Brownian motion. \(\acute {R}^{g}(\textbf {x},\textbf {y})\) and \(\acute {R}^{\mathbb {C}\setminus F}(\textbf {x},\textbf {y})\) denote the 0-order resolvent density of the subproces of \(\acute {\mathbb {M}}\) by \(\exp [-{{\int \limits }_{0}^{t}} I_{F}(X_{s})ds]\) and that of the part of \(\acute {\mathbb {M}}\) on the set \(\mathbb {C}\setminus F\), respectively. Then \(\acute {R}^{\mathbb {C}\setminus F}(\textbf {x},\textbf {y})\le \acute {R}^{g}(\textbf {x},\textbf {y})\) so that

$${\langle}\mu^{\textbf{x},r}, \acute{R}^{\mathbb{C}\setminus F}\mu^{\textbf{x},r}\>\!\le\! {\langle}\mu^{\textbf{x},r}, \acute{R}^{g}\mu^{\textbf{x},r}\>\!<\!\infty,\ \text{and}\ \acute R^{\mathbb{C}\setminus F}\mu^{\textbf{x},r}\!\in\! H_{0,e}^{1}(\mathbb{C}\setminus F)\!\subset\! \text{BL}(\mathbb{C})\!\subset\! L_{\text{loc}}^{2}(\mathbb{C}).$$

According to the fundamental identity of the logarithmic potential (cf. [13, (2.13)]),

$$U\mu^{\textbf{x},r}(\textbf{y})=\acute R^{\mathbb{C}\setminus F}\mu^{\textbf{x},r}(\textbf{y})+\acute H_{F} U\mu^{\textbf{x},r}(\textbf{y})-W_{F}(\textbf{y}),\quad \textbf{y}\in \mathbb{C},$$

which readily implies Eq. 7.18.

Define \({\Gamma }\mu ^{\textbf {x},r}(\textbf {y})={\int \limits } {\Gamma }(\textbf {y},\textbf {z})\mu ^{\textbf {x},r}(d\textbf {z}), y\in \mathbb {C}\). Γ0μx, r is defined similarly. Since Γ0(x, y) is bounded by \(K_{5} \log (1/|\textbf {x}-\textbf {y}|)+K_{6}\) for some constants K5 and K6, we have \({\Gamma }_{0}\mu ^{\textbf {x},r}\in L^{2}_{\text {loc}}(\mathbb {C})\) by Eq. 7.18. By Eq. 7.17, this also holds for Γ in place of Γ0.

Put A− 1(y) = (aij(y)). Since the weak derivative ∇Γ0μx, r is given by

$$ \nabla {\Gamma}_{0} \mu^{\textbf{x},r}(\textbf{w})={\int}_{G}{1\over \pi ({\det}(A^{-1}(\textbf{y})))^{1/2}} {A^{-1}(\textbf{y})(\textbf{w}-\textbf{y})\over {}^{t}(\textbf{w}-\textbf{y})A^{-1}(\textbf{y})(\textbf{w}-\textbf{y})}\mu^{\textbf{x},r}(d\textbf{y}), $$

we get

$$ \begin{array}{@{}rcl@{}} {\int}_{G} |\nabla {\Gamma}_{0}\mu^{\textbf{x},r}(\textbf{w})|^{2}d\textbf{w}&\leq& K_{7} \int\int\int{1\over |\textbf{w}-\textbf{y}||\textbf{w}-\textbf{z}|}d\textbf{w} \mu^{\textbf{x},r}(d\textbf{y})\mu^{\textbf{x},r}(d\textbf{z})\\ &\leq& K_{8} \int\int \log{1\over |\textbf{y}-\textbf{z}|} \mu^{\textbf{x},r}(d\textbf{y})\mu^{\textbf{x},r}(d\textbf{z})+K_{9}. \end{array} $$

which is finite by Eq. 7.18. Consequently Γ0μx, r ∈BL(G). By Eq. 7.15, Γμx, r also belongs to the space BL(G). Since the disk G is an extendable domain for BL-functions ([18]), there exists \({\Psi }\in \text {BL}(\mathbb {C})\) such that Ψ|G = Γμx, r.

In what follows, we let T = S − 1/4. By virtue of Lemma 3.8, it holds that

$$\widehat R\mu^{\textbf{x},r}-H_{\mathbb{C}\setminus B(T)}{\widehat R}\mu^{\textbf{x},r}=R^{B(T)}\mu^{\textbf{x},r}\quad \text{q.e.}$$

Further, if we let \(\mathcal {F}_{e,B(T)}=\{u\in \text {BL}(\mathbb {C}): \widetilde u =0\text {q.e. on}\ \mathbb {C}\setminus B(T)\},\) then

$$ R^{B(T)}\mu^{\textbf{x},r}\in \mathcal{F}_{e,B(T)},\ \text{and}\ \textbf{a}(R^{B(T)}\mu^{\textbf{x},r},v)={\langle}\mu^{\textbf{x},r}, \widetilde v\>,\ \forall v\in \mathcal{F}_{e,B(T)}.$$

Define \({\Psi }_{B(T)}(\textbf {y})={\Psi }(\textbf {y})-H_{\mathbb {C}\setminus B(T)}{\Psi }(\textbf {y}),\textbf {y}\in \mathbb {C}\). As \({\Psi }\in \text {BL}(\mathbb {C})\), \({\Psi }_{B(T)}\in \mathcal {F}_{e,B(T)}\) and \(H_{\mathbb {C}\setminus B(T)}{\Psi }\) is a-harmonic on B(T), namely, \(\textbf {a}(H_{\mathbb {C}\setminus B(T)}{\Psi },v)=0,\forall v\in \mathcal {F}_{e,B(T)}\). Since Ψ equals Γμx, r on G and Γ is a fundamental solution of L on G, we have \( \textbf {a}({\Psi }_{B(T)},v)={\langle }\mu ^{\textbf {x},r}, v\>, \forall v\in \mathcal {F}_{e,B(T)}\cap C_{c}(B(T)).\) Therefore \(\textbf {a}(R^{B(T)}\mu ^{\textbf {x},r}-{\Psi }_{B(T)}, R^{B(T)}\mu ^{\textbf {x},r}-{\Psi }_{B(T)})=0,\) which in turn implies

$$ \widehat R\mu^{\textbf{x},r}(\textbf{y})-{\Gamma}\mu^{\textbf{x},r}(\textbf{y})=H_{\partial B(T)}\widehat R\mu^{\textbf{x},r}(\textbf{y})-H_{\partial B(T)}{\Gamma}\mu^{\textbf{x},r}(\textbf{y}),\text{for a.e.} \textbf{y}\in B(T)\setminus \overline B(\textbf{x},r), $$
(7.19)

where \(\widehat R\mu ^{\textbf {x},r}\) is a version of Rμx, r introduced in Lemma 4.4.

By Lemma 4.6,

$$ \begin{array}{@{}rcl@{}} \sup_{\textbf{y}\in B(T)}\sup_{\textbf{x}\in B(S-1),0<r<1/8} |H_{\partial B(T)} \widehat R\mu^{\textbf{x},r}(\textbf{y})|&\le& \sup_{\textbf{z}\in \partial B(T)}\sup_{\textbf{x}\in B(S-1),0<r<1/8} |\widehat R\mu^{\textbf{x},r}(\textbf{z})| \\&=:&\ell_{1}<\infty. \end{array} $$

By Eq. 7.15, Γ(y, z) is jointly continuous on G × G off the diagonal set, and consequently

$$\sup_{\textbf{y}\in B(T)} \sup_{\textbf{x}\in B(S-1), 0<r<1/8}|H_{\partial B(T)}{\Gamma}\mu^{\textbf{x},r}(\textbf{y})|\le \sup_{\textbf{y}\in \partial B(T), z\in B(S-1/2)} |{\Gamma}(\textbf{y},\textbf{z})|=:\ell_{2}<\infty. $$

Therefore, it follows from Eq. 7.19 and Lemma 4.4 that

$$\sup_{\textbf{x}\in B(S-1), 0<r<1/8} \sup_{\textbf{y}\in B(T)\setminus B(\textbf{x},r)}|\widehat R\mu^{\textbf{x},r}(\textbf{y})-{\Gamma}\mu^{\textbf{x},r}(\textbf{y})| \le \ell_{1}+\ell_{2} <\infty.$$

By taking Eq. 7.17 into account, it holds further that \({\Gamma }\mu ^{\textbf {x},r}(\textbf {y})-{\Gamma }_{0}\mu ^{\textbf {x},r}(\textbf {y})\) is bounded uniformly in xB(S − 1) and 0 < r < 1/2.

Hence, there exists a constant K9 such that, for xB(S − 1) and 0 < 4rt < 1/2,

$$ \begin{array}{@{}rcl@{}} &&\max\left\{\widehat R\mu^{\textbf{x},r}(\textbf{y}):\textbf{y}\in \partial B(\textbf{x},t)\right\}\leq \max\left\{{\Gamma}_{0}\mu^{\textbf{x},r}(\textbf{y}): \textbf{y}\in \partial B(\textbf{x},t)\right\}+K_{9}\\ &&\leq {\Lambda\over \pi}\max\left\{{\int}_{B(\textbf{x},r)}\log{\Lambda\over |\textbf{y}-\textbf{z}|^{2}}\mu^{\textbf{x},r}(d\textbf{z}): \textbf{y}\in \partial B(\textbf{x},t)\right\}+K_{9}. \end{array} $$

Since (3/4)t ≤|yz|≤ (5/4)t for any yB(x, t) and zB(x, r), the last expression in the above display is dominated by

$$ \begin{array}{@{}rcl@{}} &&{\Lambda\over \pi}\min\left\{{\int}_{B(\textbf{x},r)}\log{25{\Lambda}\over 9|\textbf{y}-\textbf{z}|^{2}}\mu^{\textbf{x},r}(d\textbf{z}): \textbf{y}\in \partial B(\textbf{x},t)\right\}+K_{9}. \\ &&\leq {\Lambda\over \lambda}\min\left\{{\lambda\over \pi}{\int}_{B(\textbf{x},r)}\log {\lambda\over |\textbf{y}-\textbf{z}|^{2}}\mu^{\textbf{x},r}(d\textbf{z}): \textbf{y}\in \partial B(\textbf{x},t)\right\}\\&&\quad+ {\Lambda\over \pi}(\log (25{\Lambda}/9)-\log\lambda)+K_{9}\\ &&\leq {\Lambda\over \lambda} \min\{{\Gamma}_{0}\mu^{\textbf{x},r}(\textbf{y}): \textbf{y}\in \partial B(\textbf{x},t)\}+K_{10}\\ &&\leq {\Lambda\over \lambda} \min\{\widehat R\mu^{\textbf{x},r}(\textbf{y}): \textbf{y}\in \partial B(\textbf{x},t)\}+K_{9}+K_{10} \end{array} $$

for \(K_{10}=K_{9}+({\Lambda }/\pi )(\log (25{\Lambda }/9)-\log \lambda )\). Therefore Eq. 4.19 holds for κ = Λ/λ and C2 = K9 + K10.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Fukushima, M., Oshima, Y. Gaussian fields, equilibrium potentials and multiplicative chaos for Dirichlet forms. Potential Anal 55, 285–337 (2021). https://doi.org/10.1007/s11118-020-09858-0

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11118-020-09858-0

Keywords

Mathematics Subject Classification (2010)

Navigation