Skip to main content
Log in

Density of states for the Anderson model on nested fractals

  • Published:
Analysis and Mathematical Physics Aims and scope Submit manuscript

Abstract

We prove the existence and establish the Lifschitz singularity of the integrated density of states for certain random Hamiltonians \(H^\omega =H_0+V^\omega \) on fractal spaces of infinite diameter. The kinetic term \(H_0\) is given by \(\phi (-{\mathcal {L}}),\) where \({\mathcal {L}}\) is the Laplacian on the fractal and \(\phi \) is a completely monotone function satisfying some mild regularity conditions. The random potential \(V^\omega \) is of alloy-type.

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

Availability of data and materials

Not applicable. No datasets were generated or analyzed during the current study.

References

  1. Akkermans, E., Chen, J.P., Dunne, G., Rogers, L.G., Teplyaev, A.: Fractal AC circuits and propagating waves on fractals. Analysis, probability and mathematical physics on fractals, 557–567, Fractals Dyn. Math. Sci. Arts Theory Appl., 5, World Sci. Publ., Hackensack, NJ (2020)

  2. Alonso-Ruiz, P., Baudoin, F., Chen, L., Rogers, L.G., Shanmugalingam, N., Teplyaev, A.: Besov class via heat semigroup on Dirichlet spaces I: Sobolev type inequalities. J. Funct. Anal. 278(11), 108459 (2020)

    MathSciNet  Google Scholar 

  3. Alonso-Ruiz, P., Baudoin, F., Chen, L., Rogers, L.G., Shanmugalingam, N., Teplyaev, A.: Besov class via heat semigroup on Dirichlet spaces II: BV functions and Gaussian heat kernel estimates. Calc. Var. Partial Differ. Equ. 59(3), 103 (2020)

    MathSciNet  Google Scholar 

  4. Alonso-Ruiz, P., Baudoin, F., Chen, L., Rogers, L.G., Shanmugalingam, N., Teplyaev, A.: Besov class via heat semigroup on Dirichlet spaces III: BV functions and sub-Gaussian heat kernel estimates. Calc. Var. Partial Differ. Equ. 60(5), 170 (2021)

    MathSciNet  Google Scholar 

  5. Alonso-Ruiz, P., Baudoin, F., Chen, L., Rogers, L.G., Shanmugalingam, N., Teplyaev, A.: BV functions and fractional Laplacians on Dirichlet spaces. arXiv:1910.13330

  6. Balsam, H., Kaleta, K., Olszewski, M., Pietruska-Pałuba, K.: IDS for subordinate Brownian motions in Poisson random environment on nested fractals: existence and Lifschitz tail. preprint (2022)

  7. Benderskii, M., Pastur, L.: On the spectrum of the one-dimensional Schrödinger equation with random potential. Mat. Sb. 82, 245–256 (1970)

    Google Scholar 

  8. Bourgain, J., Kenig, C.E.: On localization in the continuous Anderson–Bernoulli model in higher dimension Invent. Invent. Math. 161, 389–426 (2005)

    MathSciNet  Google Scholar 

  9. Carmona, R., Lacroix, J.: Spectral Theory of Random Schrödinger Operators. Probability and Its Applications. Birkhäuser Boston Inc., Boston (1990)

    Google Scholar 

  10. Chaumont, L., Uribe Bravo, G.: Markovian bridges: weak continuity and pathwise constructions. Ann. Probab. 39(2), 609–647 (2011)

    MathSciNet  Google Scholar 

  11. Chen, J.P., Hinz, M., Teplyaev, A.: From non-symmetric particle systems to non-linear PDEs on fractals. Stochastic partial differential equations and related fields, 503–513, Springer Proc. Math. Stat., 229, Springer, Cham (2018)

  12. Chen, Z.-Q., Song, R.: Two sided eigenvalue estimates for subordinate processes in domains. J. Funct. Anal. 226, 90–113 (2005)

    MathSciNet  Google Scholar 

  13. Chung, K.L., Zhao, Z.: From Brownian Motion to Schrödinger’s Equation. Springer, New York (1995)

    Google Scholar 

  14. Combes, J.M., Hislop, P.D.: Localization for some continuous, random Hamiltonians in d-dimensions. J. Funct. Anal. 124, 149–180 (1994)

    MathSciNet  Google Scholar 

  15. Dekkers, A., Rozanova-Pierrat, A., Teplyaev, A.: Mixed boundary valued problems for linear and nonlinear wave equations in domains with fractal boundaries. Calc. Var. Partial Differ. Equ. 61(2), 75 (2022)

    MathSciNet  Google Scholar 

  16. Demuth, M., van Casteren, J.A.: Stochastic Spectral Theory for Self-adjoint Feller Operators. A Functional Analysis Approach. Birkhäuser, Basel (2000)

    Google Scholar 

  17. Friedberg, R., Luttinger, J.: Density of electronic energy levels in disordered systems. Phys. Rev. B 12, 4460–4474 (1975)

    MathSciNet  Google Scholar 

  18. Fukushima, M.: Dirichlet forms, diffusion processes, and spectral dimensions for nested fractals. In: Ideas and Methods in Stochastic Analysis. Stochastics and Applications, pp. 151–161. Cambridge University Press, Cambridge (1992)

  19. Fitzsimmons, P.J., Hambly, B.M., Kumagai, T.: Transition density estimates for Brownian motion on affine nested fractals. Commun. Math. Phys. 165(3), 595–620 (1994)

    MathSciNet  Google Scholar 

  20. Fukushima, M.: On the spectral distribution of a disordered system and a range of a random walk. Osaka J. Math. 11, 73–85 (1974)

    MathSciNet  Google Scholar 

  21. Fukushima, M., Nagai, H., Nakao, S.: On an asymptotic property of spectra of a random difference operator. Proc. Jpn. Acad. 51, 100–102 (1975)

    MathSciNet  Google Scholar 

  22. Gebert, M., Rojas-Molina, C.: Lifshitz tails for the fractional Anderson model. J. Stat. Phys. 179, 341–353 (2020)

    MathSciNet  Google Scholar 

  23. Germinet, F., Hislop, P.D., Klein, A.: Localization for Schrödinger operators with Poisson random potential. J. Eur. Math. Soc. (JEMS) 9(3), 577–607 (2007)

    MathSciNet  Google Scholar 

  24. Hinz, M., Rozanova-Pierrat, A., Teplyaev, A.: Non-Lipschitz uniform domain shape optimization in linear acoustics. SIAM J. Control Optim. 59(2), 1007–1032 (2021)

    MathSciNet  Google Scholar 

  25. Kumagai, T., Nakamura, C.: Lamplighter random walks on fractals. J. Theor. Probab. 31(1), 68–92 (2018)

    MathSciNet  Google Scholar 

  26. Hinz, M., Teplyaev, A.: Dirac and magnetic Schrödinger operators on fractals. J. Funct. Anal. 265(11), 2830–2854 (2013)

    MathSciNet  Google Scholar 

  27. Kaleta, K., Olszewski, M., Pietruska-Pałuba, K.: Reflected Brownian motion on simple nested fractals. Fractals 27(6), 19501041–29 (2019)

    MathSciNet  Google Scholar 

  28. Kaleta, K., Pietruska-Pałuba, K.: Integrated density of states for Poisson–Schrödinger perturbations of subordinate Brownian motions on the Sierpiński gasket. Stoch. Process. Appl. 125(4), 1244–1281 (2015)

    Google Scholar 

  29. Kaleta, K., Pietruska-Pałuba, K.: Lifschitz singularity for subordinate Brownian motions in presence of the Poissonian potential on the Sierpiński triangle. Stoch. Process. Appl. 128(11), 3897–3939 (2018)

    Google Scholar 

  30. Kaleta, K., Pietruska-Pałuba, K.: Lifschitz tail for alloy-type models driven by the fractional Laplacian. J. Funct. Anal. 279(5), 108575 (2020)

    MathSciNet  Google Scholar 

  31. Kaleta, K., Pietruska-Pałuba, K.: Lifschitz tail for continuous Anderson models driven by Lévy operators. Commun. Contemp. Math. 23(6), 2050065 (2021)

    Google Scholar 

  32. Kigami, J.: Time changes of the Brownian motion: poincaré inequality, heat kernel estimate and protodistance. Mem. Am. Math. Soc. 259(1250), v+118 (2019)

    Google Scholar 

  33. Kirsch, W., Martinelli, F.: On the density of states of Schrödinger operators with a random potential. J. Phys. A 15, 2139–2156 (1982)

    MathSciNet  Google Scholar 

  34. Kirsch, W., Martinelli, F.: Large deviations and Lifshitz singularity of the integrated density of states of random Hamiltonians. Commun. Math. Phys. 89, 27–40 (1983)

    MathSciNet  Google Scholar 

  35. Kirsch, W., Simon, B.: Lifshitz tails for periodic plus random potentials. J. Stat. Phys. 42(5/6), 799–808 (1986)

    MathSciNet  Google Scholar 

  36. Kirsch, W., Veselić, I.: Lifshitz tails for a class of Schrödinger operators with random breather-type potential. Lett. Math. Phys. 94, 27–39 (2010)

    MathSciNet  Google Scholar 

  37. Klopp, F.: Weak disorder localization and Lifshitz tails. Commun. Math. Phys. 232, 125–155 (2002)

    MathSciNet  Google Scholar 

  38. Klopp, F.: Weak disorder localization and Lifshitz tails: continuous Hamiltonians. Ann. Henri Poincaré 3, 711–737 (2002)

    MathSciNet  Google Scholar 

  39. Kumagai, T.: Estimates of transition densities for Brownian motion on nested fractals. Probab. Theory Relat. Fields 96(2), 205–224 (1993)

    MathSciNet  Google Scholar 

  40. Kumagai, T.: Anomalous random walks and diffusions: from fractals to random media. In: Proceedings of the International Congress of Mathematicians, Seoul 2014. Vol. IV, 75–94, Kyung Moon Sa, Seoul (2014)

  41. Kusuoka, S.: Dirichlet forms on fractals and products of random matrices. Publ. RIMS Kyoto Univ. 25, 659–680 (1989)

    MathSciNet  Google Scholar 

  42. Lindstrom, T.: Brownian motion on nested fractals. Mem. Am. Math. Soc. 83(420), iv+128 (1990)

    MathSciNet  Google Scholar 

  43. Luttinger, J.M.: New variational method with applications to disordered systems. Phys. Rev. Lett. 37, 609–612 (1976)

    MathSciNet  Google Scholar 

  44. Mezincescu, G.: Bounds on the integrated density of electronic states for disordered Hamiltonians. Phys. Rev. B 32, 6272–6277 (1985)

    Google Scholar 

  45. Nagai, H.: On an exponential character of the spectral distribution function of a random difference operator. Osaka J. Math. 14, 111–116 (1977)

    MathSciNet  Google Scholar 

  46. Nakao, S.: On the spectral distribution of the Schrödinger operator with random potential. Jpn. J. Math. 3, 111–139 (1977)

    Google Scholar 

  47. Nieradko, M., Olszewski, M.: Good labeling property of simple nested fractals (2021) arXiv:2110.15921 to appear in Journal of Fractal Geometry

  48. Okura, H.: On the spectral distributions of certain integro-differential operators with random potential. Osaka J. Math. 16(3), 633–666 (1979)

    MathSciNet  Google Scholar 

  49. Olszewski, M.: Estimates for the transition densities of the reflected Brownian motion on simple nested fractals. Prob. Math. Statist. 39(2), 423–440 (2019)

    MathSciNet  Google Scholar 

  50. Pastur, L.A.: The behavior of certain Wiener integrals as \(t\rightarrow \infty \) and the density of states of Schrödinger equations with random potential. Teoret. Mat. Fiz. 32, 88–95 (1977). (Russian)

    MathSciNet  Google Scholar 

  51. Pietruska-Pałuba, K.: The Lifschitz singularity for the density of states on the Sierpinski gasket. Probab. Theory Relat. Fields 89(1), 1–33 (1991)

    MathSciNet  Google Scholar 

  52. Pietruska-Pałuba, K.: The wiener sausage asymptotics on simple nested fractals. Stoch. Anal. Appl. 23(1), 111–135 (2005)

    MathSciNet  Google Scholar 

  53. Pietruska-Pałuba, K., Stos, A.: Poincaré inequality and Hajlasz–Sobolev spaces on nested fractals. Studia Math. 218(1), 1–28 (2013)

    MathSciNet  Google Scholar 

  54. Romerio, M., Wreszinski, W.: On the Lifshitz singularity and the tailing in the density of states for random lattice systems. J. Stat. Phys. 21, 169–179 (1979)

    Google Scholar 

  55. Reed, M., Simon, B.: Methods on Modern Mathematical Physics. Vol. 4: Analysis of Operators. Academic Press, Cambridge (1978)

    Google Scholar 

  56. Schilling, R., Song, R., Vondraček, Z.: Bernstein Functions. Walter de Gruyter, Berlin (2010)

    Google Scholar 

  57. Shima, T.: Lifschitz tails for random Schrödinger operators on nested fractals. Osaka J. Math. 29, 749–770 (1992)

    MathSciNet  Google Scholar 

  58. Simon, B.: Lifshitz tails for the Anderson model. J. Stat. Phys. 38, 65–76 (1985)

    Google Scholar 

  59. Stollmann, P.: Caught by Disorder: Bound States in Random Media. Birkhäuser, Boston (2001)

    Google Scholar 

  60. Sznitman, A.S.: Lifschitz tail and Wiener sausage on hyperbolic space. Commun. Pure Appl. Math. 42, 1033–1065 (1989)

    MathSciNet  Google Scholar 

  61. Sznitman, A.S.: Lifschitz tail on hyperbolic space: Neumann conditions. Commun. Pure Appl. Math. 43, 1–30 (1990)

    MathSciNet  Google Scholar 

Download references

Acknowledgements

We thank the referee for a careful reading of the paper and useful comments.

Funding

Research was supported by the National Science Centre, Poland, Grants No. 2015/17/B/ST1/01233 and 2019/35/B/ST1/02421.

Author information

Authors and Affiliations

Authors

Contributions

All the authors equally contributed to this article.

Corresponding author

Correspondence to Kamil Kaleta.

Ethics declarations

Conflict of interest

The authors have no competing interests to declare that are relevant to the content of this article.

Ethical approval

Not applicable.

Additional information

Publisher's Note

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

Research was supported in parts by the National Science Centre, Poland, Grants No. 2015/17/B/ST1/01233 and 2019/35/B/ST1/02421.

Appendices

Appendix A: Technical lemmas on subordinate processes

We need the following technical results.

Lemma A.1

Let \({\mathcal {K}}^{\langle \infty \rangle }\) be an USNF with the GLP and let the assumption (B) hold. For every \(t>0\) and \(a>1\) we have

$$\begin{aligned} \sum _{M=1}^{\infty } \sup _{x \in {\mathcal {K}}^{\langle \infty \rangle }} {\textbf{P}}^x \left[ \sup _{s\le t} |X_s - x| > a^M \right] <\infty . \end{aligned}$$

Proof

The proof follows the lines of that of [28, Lemma 2.3]. The only difference is that now we use the Euclidean distance instead of the geodesic one, and apply the sub-Gaussian estimates from [19, Lemma 5.6, Remark 3.7]. At the end, we also use the summation property (2.20) as before. \(\square \)

Let us recall that \(\Delta _{M,i}\) for \(1 \le i \le N\) denote the M-complexes in \({\mathcal {K}}^{\langle M+1 \rangle }\), see Definition 2.1 (9).

Lemma A.2

Let \({\mathcal {K}}^{\langle \infty \rangle }\) be an USNF with the GLP and let the assumption (B) hold. We have the following.

  1. (a)

    For

    $$\begin{aligned} C(M,t):= \sup _{x,y \in {\mathcal {K}}^{\langle M \rangle }} \sum _{\begin{array}{c} y^{\prime } \in \pi _{M}^{-1}(y) \\ y' \notin {\mathcal {K}}^{\left\langle M+1 \right\rangle } \end{array}} p(t,x,y'), \quad t>0, \quad M \in {\mathbb {Z}} \end{aligned}$$
    (A.1)

    we have

    $$\begin{aligned} \sum _{M=1}^{\infty } C(M,t) < \infty , \quad t>0; \end{aligned}$$

    in particular, for every \(t>0\), \(C(M,t) \rightarrow 0\) as \(M \rightarrow \infty \).

  2. (b)

    For every \(t>0\),

    $$\begin{aligned} \sum _{M=1}^{\infty } \frac{1}{\mu \left( {\mathcal {K}}^{\langle M \rangle }\right) } \int _{{\mathcal {K}}^{\langle M \rangle }} \left| p(t,x,x) - p_M(t,x,x) \right| \mu (\textrm{d}x) < \infty ; \end{aligned}$$
    (A.2)

    in particular

    $$\begin{aligned} \frac{1}{\mu \left( {\mathcal {K}}^{\langle M \rangle }\right) } \int _{{\mathcal {K}}^{\langle M \rangle }} \left| p(t,x,x) - p_M(t,x,x) \right| \mu (\textrm{d}x) \rightarrow 0, \quad as \ M \rightarrow \infty .\nonumber \\ \end{aligned}$$
    (A.3)

Proof

By Tonelli’s theorem and [49, formula (3.3) and Lemma 3.5], we get

$$\begin{aligned} C(M,t)&= \sup _{x,y \in {\mathcal {K}}^{\langle M \rangle }} \int _0^{\infty } \sum _{\begin{array}{c} y^{\prime } \in \pi _{M}^{-1}(y) \\ y' \notin {\mathcal {K}}^{\left\langle M+1 \right\rangle } \end{array}} g(t,x,y') \eta _t(\textrm{d}u) \\&\le c_1 \int _0^{\infty } L^{-d M} \left( \frac{L^M}{t^{1/d_w}} \vee 1\right) ^{d - \frac{d_w}{d_J-1}} \exp \left( -c_2 \left( \frac{L^M}{t^{1/d_w}} \vee 1\right) ^{\frac{d_w}{d_J-1}}\right) \eta _t(\textrm{d}u) \\&\le c_1 \int _0^{L^{d_{w}M}} L^{-d M} \left( \frac{L^M}{u^{1/d_w}}\right) ^{d - \frac{d_w}{d_J-1}} \exp {\left( -c_2 \left( \frac{L^M}{u^{1/d_w}}\right) ^{\frac{d_w}{d_J-1}}\right) } \eta _t(\textrm{d}u) \\&\ \ \ \ \ \ \ \ \ \ \ \ \ \ + c_3 L^{-d M} \eta _{t} \left( L^{d_{w}M}, \infty \right) . \end{aligned}$$

Further steps of the proof follow exactly the reasoning in the proof of [28, Lemma 2.5 (b)].

The proof of (b) is similar to that of [28, Lemma 2.7], i.e. we start with

$$\begin{aligned}&\frac{1}{\mu \left( {\mathcal {K}}^{\langle M \rangle }\right) } \int _{{\mathcal {K}}^{\langle M \rangle }} \left| p(t,x,x) - p_M(t,x,x) \right| \mu (\textrm{d}x) \\&\qquad \le \frac{N}{\mu \left( {\mathcal {K}}^{\langle M \rangle }\right) } \int _{{\mathcal {K}}^{\langle M \rangle }} \sup _{\begin{array}{c} x^{\prime } \in \pi _{M}^{-1}(x) \\ x' \in {\mathcal {K}}^{\langle M+1 \rangle } \backslash {\mathcal {K}}^{\langle M \rangle } \end{array}} p(t,x,x') \mu (\textrm{d}x)\\&\qquad \quad + \frac{1}{\mu \left( {\mathcal {K}}^{\langle M \rangle }\right) } \int _{{\mathcal {K}}^{\langle M \rangle }} \sum _{\begin{array}{c} x^{\prime } \in \pi _{M}^{-1}(x) \\ x' \notin {\mathcal {K}}^{\langle M+1 \rangle } \end{array}} p(t,x,x') \mu (\textrm{d}x)\\&\qquad =: {\mathcal {A}}_M(t) + {\mathcal {B}}_M(t) \end{aligned}$$

and first observe that \({\mathcal {B}}_M(t)\) is a term of convergent series by part (a). To prove that the same is true for \({\mathcal {A}}_M(t)\), we only need to modify the estimate in the proof of the quoted lemma. The difference is that the domain of the integration in \({\mathcal {A}}_M(t)\) has to be divided into two different sets \(E_M^1\) and \(E_M^2\). In the present general case, we cannot use geodesic balls. We will apply the graph distance here.

Let us denote \(V_M^{\langle M \rangle } = \{v_1,\ldots , v_k\}\) and

$$\begin{aligned} E_M^1 = \bigcap _{i=1}^{k} \left\{ y \in {\mathcal {K}}^{\langle M \rangle }: d_{\lfloor M/2\rfloor } (y, v_i) > 1\right\} , \quad E_M^2 = {\mathcal {K}}^{\langle M \rangle } \backslash E_M^1. \end{aligned}$$
(A.4)

Recall that \(\sup _{(x,y) \in {\mathcal {K}}^{\langle \infty \rangle } \times {\mathcal {K}}^{\langle \infty \rangle }} p(t,x,y) < \infty \) by Lemma 2.1 (a). Since \(E_M^2\) consists of k \(\lfloor M/2 \rfloor \)-complexes (each one attached to one of the vertices from \(V_M^{\langle M \rangle }\)), we have that

$$\begin{aligned} \frac{\mu \left( E_M^2\right) }{\mu \left( {\mathcal {K}}^{\langle M \rangle }\right) } = \frac{k \cdot N^{\lfloor M/2 \rfloor }}{N^M}. \end{aligned}$$
(A.5)

In consequence, the part of \({\mathcal {A}}_M(t)\) including the integral over \(E_M^2\) is a term of a convergent series.

When \(x \in E_M^1\) and \(x^{\prime } \in \pi _{M}^{-1}(x) \backslash {\mathcal {K}}^{\langle M \rangle }\), then \(d_{\lfloor M/2 \rfloor } (x, x') >2\), so from [27, Lemma A.2] we have \(|x-x'|> c_4 L^{M/2}\). Using the subordination formula for the density p(txy) and the subgaussian estimates from [39] for the density g(uxy), we then get

$$\begin{aligned} \frac{1}{\mu \left( {\mathcal {K}}^{\langle M \rangle }\right) } \int _{E_M^1} \sup _{\begin{array}{c} x^{\prime } \in \pi _{M}^{-1}(x) \\ x' \in {\mathcal {K}}^{\langle M+1 \rangle } \backslash {\mathcal {K}}^{\langle M \rangle } \end{array}} p(t,x,x') \mu (\textrm{d}x) \le c_5 \int _0^{\infty } u^{-d_s / 2} \textrm{e}^{-c_6 \left( \frac{L^{M/2}}{u^{1/d_w}}\right) ^{\frac{d_w}{d_J-1}}} \eta _t(\textrm{d}u). \end{aligned}$$

Collecting these estimates, we obtain

$$\begin{aligned} {\mathcal {A}}_M(t) \le c_7 \left( N^{-M/2}+ \int _0^{\infty } u^{-d_s / 2} \textrm{e}^{-c_6 \left( \frac{L^{M/2}}{u^{1/d_w}}\right) ^{\frac{d_w}{d_J-1}}} \eta _t(\textrm{d}u)\right) , \end{aligned}$$

with the constants \(c_6, c_7\) independent of M, and from this point the proof can be continued in the same manner as that of [28, Lemma 2.7]. \(\square \)

Appendix B: Proofs of the statements from Sect. 3

Proof of Proposition 3.1

Fix \(t>0\). For \(\mu \)-almost all \(x \in {\mathcal {K}}^{\langle M+1\rangle },\) by Lemmas 2.2(a),  3.1, and the inclusion \(\pi _{M+1}^{-1}(x) \subset \pi _M^{-1}\left( \pi _M(x)\right) \), we may write

$$\begin{aligned}&p_{M+1}(t,x,x) {\textbf{E}}_{M+1,t}^{x,x} [{\mathbb {E}}_{{\mathbb {Q}}} \textrm{e}^{-\int _0^t V^{\omega }_{M+1}(X^{M+1}_s)\textrm{d}s} ]\\&\quad = \sum _{x' \in \pi _{M+1}^{-1} (x)} p(t,x,x') {\textbf{E}}_{t}^{x,x'} \left[ {\mathbb {E}}_{{\mathbb {Q}}} \textrm{e}^{-\int _0^t V^{\omega }_{M+1}(\pi _{M+1}(X_s))\textrm{d}s} \right] \\&\quad \le \sum _{x' \in \pi _{M}^{-1} \left( \pi _M(x)\right) } p(t,x,x') {\textbf{E}}_{t}^{x,x'} \left[ {\mathbb {E}}_{{\mathbb {Q}}} \textrm{e}^{-\int _0^t V^{\omega }_{M+1}(\pi _{M+1}(X_s))\textrm{d}s} \right] \\&\quad \le \sum _{x' \in \pi _{M}^{-1} \left( \pi _M(x)\right) } p(t,x,x') {\textbf{E}}_{t}^{x,x'} \left[ {\mathbb {E}}_{{\mathbb {Q}}} \textrm{e}^{-\int _0^t V^{\omega }_{M}(\pi _{M}(X_s))\textrm{d}s} \right] . \end{aligned}$$

Using this estimate, the definition of \(\Lambda _{M^*}^{N}(t,\omega )\) and Lemma 2.2(b), we may now follow the argument in the second part of the proof of [28, Theorem 3.1], getting that

$$\begin{aligned} {\mathbb {E}}_{{\mathbb {Q}}}[\Lambda _{M+1}^{N, V_{M+1}^\omega }(t)] \le {\mathbb {E}}_{{\mathbb {Q}}}[ \Lambda _{M}^{N, V_M^\omega }(t)], \quad M \in \mathbb {Z_+}. \end{aligned}$$

Since \({\mathbb {E}}_{{\mathbb {Q}}}\Lambda _{M}^{N, V_M^\omega }(t) \ge 0\), it converges to a finite limit \(\Lambda (t)\) as \(M \rightarrow \infty \). This completes the proof. \(\square \)

Proofs of Lemmas 3.2 and 3.3 given below are of technical nature. They follow the steps and ideas from the proofs of Proposition 3.1 and Lemmas 3.1\(-\)3.2 in [28]. The main difference here is that we now work with different type of random potentials and the state space is now a general USNF (let us emphasize that the Sierpiński triangle which was studied in the quoted paper is one of the simplest planar nested fractals with high regularity). This causes some extra geometric issues, which are solved by using the graph distance (the geodesic metric may not be defined at all) and the comparison principle from [27, Lemma A.2]. Therefore, we will follow only the main steps of the proofs which are affected by these changes, focusing on the most critical differences.

Proof of Lemma 3.2

We first show (a). By following the argument at the beginning of [28, Proposition 3.1] we obtain that for each fixed \(t>0\) there exists a constant \(c=c(t)\) such that

$$\begin{aligned} {\mathbb {E}}_{{\mathbb {Q}}} \left( \Lambda _{M}^{D, V^\omega }(t) - \Lambda _{M}^{N, V^\omega } (t) \right) ^2 \le c \left( R_{1,M}(t) + R_{2,M}(t)\right) , \quad M \in {\mathbb {Z}}_{+}, \end{aligned}$$

where

$$\begin{aligned} R_{1,M}(t)&= \frac{1}{\mu \left( {\mathcal {K}}^{\langle M \rangle }\right) } \int _{{\mathcal {K}}^{\langle M \rangle }} p(t,x,x) {\textbf{E}}_{t}^{x,x} \left[ t\ge \tau _{{\mathcal {K}}^{\langle M\rangle }} \right] \mu (\textrm{d}x)\\ R_{2,M}(t)&= \frac{1}{\mu \left( {\mathcal {K}}^{\langle M \rangle }\right) } \int _{{\mathcal {K}}^{\langle M \rangle }} \sum _{x' \in \pi _M^{-1}(x), x' \ne x} p(t,x,x') \mu (\textrm{d}x) \\&= \frac{1}{\mu \left( {\mathcal {K}}^{\langle M \rangle }\right) } \int _{{\mathcal {K}}^{\langle M \rangle }} (p_M(t,x,x) - p(t,x,x)) \mu (\textrm{d}x). \end{aligned}$$

Note that the above bound does not depend on \(\omega \). By Lemma A.2 (b), \(R_{2,M}(t)\) is the term of a convergent series, so we only need to estimate \(R_{1,M}(t)\).

Denote the vertices from \({\mathcal {V}}_M^{\langle M \rangle }\) as \(v_i\), \(1\le i\le k\), and let \(\Delta _{\left\lfloor M/2 \right\rfloor ,v_i} \subset {\mathcal {K}}^{\langle M \rangle }\), be the \(\left\lfloor M/2 \right\rfloor \)-complex attached to \(v_i\). If the process starts from \(x \in {\mathcal {D}}_M:={\mathcal {K}}^{\langle M \rangle } \backslash \bigcup _{i=1}^{k} \Delta _{\left\lfloor M/2 \right\rfloor ,v_i}\), then, using [27, Lemma A.2], we have

$$\begin{aligned} \left\{ t\ge \tau _{{\mathcal {K}}^{\langle M \rangle }}\right\} \subseteq \left\{ \sup _{0<s\le t} d_{\left\lfloor M /2 \right\rfloor }(x,X_s)>2\right\} \subseteq \left\{ \sup _{0<s\le t} |x-X_s| > c_{1} L^{M/2}\right\} , \end{aligned}$$

with a constant \(c_1\), independent of M. We also have

$$\begin{aligned}{} & {} \left\{ \sup _{0<s\le t} |x-X_s|> c_{1} L^{M/2} \right\} \subseteq \left\{ \sup _{0<s\le t/2} |x-X_s|> c_{1} L^{M/2} \right\} \\{} & {} \qquad \cup \left\{ \sup _{t/2<s\le t} |x-X_s| > c_{1} L^{M/2} \right\} . \end{aligned}$$

Recall that \(\mu \left( {\mathcal {K}}^{\langle M \rangle } \backslash {\mathcal {D}}_M \right) = k N^{\left\lfloor M/2 \right\rfloor }\). Then, by using the upper bound in Lemma 2.1(a), the formula (2.25) and the symmetry of the bridge measure, we get

$$\begin{aligned} \int _{{\mathcal {K}}^{\langle M \rangle }} p(t,x,x)&{\textbf{P}}^{t}_{x,x} \left[ t\ge \tau _{{\mathcal {K}}^{\langle M\rangle }} \right] \mu (\textrm{d}x)\\&\le c_{2} \mu \left( {\mathcal {K}}^{\langle M \rangle } \backslash {\mathcal {D}}_M \right) + \int _{{\mathcal {D}}_M} p(t,x,x) {\textbf{P}}^{x,x}_{t}\left[ t\ge \tau _{{\mathcal {K}}^{\langle M \rangle }}\right] \mu (\textrm{d}x)\\&\le c_3 \left( N^{M/2} + N^M \sup _{x \in {\mathcal {K}}^{\langle M \rangle }} {\textbf{P}}^{x}\left[ \sup _{0<s\le t/2} |x-X_s|>c_{1} L^{M/2}\right] \right) . \end{aligned}$$

Therefore

$$\begin{aligned} R_{1,M}(t) \le c_3 \left( N^{-M/2} + \sup _{x \in {\mathcal {K}}^{\langle M \rangle }} {\textbf{P}}^{x}\left[ \sup _{0<s\le t/2} |x-X_s|>c_{1} L^{M/2}\right] \right) , \end{aligned}$$

which is, by Lemma A.1, a term of a convergent series. The proof of (a) is completed.

The proof of the first convegence in (b) follows the lines of that of (a) and it is omitted. We only show the second convergence.

First observe that

$$\begin{aligned}{} & {} \left| {\mathbb {E}}_{{\mathbb {Q}}} \left( \Lambda _{M}^{D, V^\omega } (t) - \Lambda _{M}^{D, V_M^\omega } (t)\right) \right| \\{} & {} \quad \le \frac{1}{\mu \left( {\mathcal {K}}^{\langle M\rangle }\right) } \int _{{\mathcal {K}}^{\langle M \rangle }} p(t,x,x){\textbf{E}}^{x,x}_{t} \left[ \left| F(t,M)\right| ; t<\tau _{{\mathcal {K}}^{\langle M \rangle }}\right] \mu (\textrm{d}x), \end{aligned}$$

where

$$\begin{aligned} F(t,M)&= {\mathbb {E}}_{{\mathbb {Q}}} \bigg ( \exp \bigg ( -\int _0^t \sum _{y \in {\mathcal {V}}_{0}^{\langle \infty \rangle }} \xi _y(\omega ) W\left( X_s, y \right) \ \textrm{d}s \bigg ) \\ {}&- \exp \bigg ( -\int _0^t \sum _{y \in {\mathcal {V}}_{0}^{\langle M\rangle }}\xi _y(\omega ) \sum _{y' \in \pi _{M}^{-1}(y)} W\left( X_s, y' \right) \ \textrm{d}s \bigg ) \bigg ) . \end{aligned}$$

By using exactly the same notation and argument as in the proof of part (a), we may write

$$\begin{aligned} \frac{1}{\mu \left( {\mathcal {K}}^{\langle M\rangle }\right) }&\int _{{\mathcal {K}}^{\langle M \rangle }} p(t,x,x) {\textbf{E}}^{x,x}_{t} \left[ \left| F(t,M)\right| ; t<\tau _{{\mathcal {K}}^{\langle M \rangle }}\right] \mu (\textrm{d}x) \\&\le \frac{c_4 \mu \left( {\mathcal {K}}^{\langle M\rangle } \backslash {\mathcal {D}}_M \right) }{\mu \left( {\mathcal {K}}^{\langle M\rangle }\right) } + \frac{1}{\mu \left( {\mathcal {K}}^{\langle M\rangle }\right) } \int _{{\mathcal {D}}_M} p(t,x,x){\textbf{E}}^{x,x}_{t} \left[ \left| F(t,M)\right| ; t<\tau _{{\mathcal {K}}^{\langle M \rangle }}\right] \mu (\textrm{d}x) \\&= c_4 \frac{k N^{\left\lfloor M/2\right\rfloor }}{N^M} + \frac{1}{\mu \left( {\mathcal {K}}^{\langle M\rangle }\right) } \int _{{\mathcal {D}}_M} p(t,x,x){\textbf{E}}^{x,x}_{t} \left[ \left| F(t,M)\right| ; t<\tau _{{\mathcal {K}}^{\langle M \rangle }}\right] \mu (\textrm{d}x). \end{aligned}$$

We see that the first term converges to 0 as \(M \rightarrow \infty \). It is sufficient to show that the second term goes to zero as well; denote it by I(tM). We have

$$\begin{aligned} I(t,M)&\le \frac{1}{\mu \left( {\mathcal {K}}^{\langle M\rangle }\right) } \int _{{\mathcal {D}}_M} p(t,x,x){\textbf{E}}^{x,x}_{t} \left[ \left| F(t,M)\right| ; t<\tau _{{\mathcal {D}}_M }\right] \mu (\textrm{d}x) \\&\quad + \frac{1}{\mu \left( {\mathcal {K}}^{\langle M\rangle }\right) } \int _{{\mathcal {D}}_M} p(t,x,x){\textbf{P}}^{x,x}_{t} \left[ t\ge \tau _{{\mathcal {D}}_M } \right] \mu (\textrm{d}x) \\&=: I_1(t,M) + I_2(t,M). \end{aligned}$$

To estimate \(I_1(t,M)\) we will use the inequality \(|\textrm{e}^{-x}-\textrm{e}^{-y}| \le |x-y|, x,y>0\), the fact that W and \(\xi _y\) are nonnegative.

$$\begin{aligned}&\left| F(t,M)\right| \\&\le {\mathbb {E}}_{{\mathbb {Q}}} \left| \exp \left( -\int _0^t \sum _{y \in {\mathcal {V}}_{0}^{\langle \infty \rangle }} \xi _y(\omega ) W\left( X_s, y \right) \ \textrm{d}s \right) \right. \\&\quad \left. - \exp \left( -\int _0^t \sum _{y \in {\mathcal {V}}_{0}^{\langle M\rangle }}\xi _y(\omega ) \sum _{y' \in \pi _{M}^{-1}(y)} W\left( X_s, y' \right) \ \textrm{d}s \right) \right| \\&\le {\mathbb {E}}_{{\mathbb {Q}}} \left| \int _0^t \sum _{y \in {\mathcal {V}}_{0}^{\langle \infty \rangle }} \xi _y(\omega ) W\left( X_s, y \right) \ \textrm{d}s - \int _0^t \sum _{y \in {\mathcal {V}}_{0}^{\langle M\rangle }}\xi _y(\omega ) \sum _{y' \in \pi _{M}^{-1}(y)} W\left( X_s, y' \right) \ \textrm{d}s \right| . \end{aligned}$$

Now, observe that for every \(x \in {\mathcal {K}}^{\langle \infty \rangle }\) we have

$$\begin{aligned} \sum _{y \in {\mathcal {V}}_{0}^{\langle \infty \rangle }} \xi _y(\omega ) W\left( x, y \right) = \sum _{y \in {\mathcal {V}}_{0}^{\langle M \rangle }} \xi _y(\omega ) W\left( x, y \right) + \sum _{y \in {\mathcal {V}}_{0}^{\langle \infty \rangle } \setminus {\mathcal {V}}_{0}^{\langle M \rangle }} \xi _y(\omega ) W\left( x, y \right) \end{aligned}$$

and

$$\begin{aligned} \sum _{y \in {\mathcal {V}}_{0}^{\langle M\rangle }} \sum _{y' \in \pi _{M}^{-1}(y)} \xi _{y'}(\omega ) W\left( x, y' \right){} & {} = \sum _{y \in {\mathcal {V}}_{0}^{\langle M\rangle }} \xi _{y}(\omega ) W\left( x, y \right) \\{} & {} \quad + \sum _{y \in {\mathcal {V}}_{0}^{\langle M\rangle }} \sum _{y' \in \pi _{M}^{-1}(y) \backslash \{y\}} \xi _{y'}(\omega ) W\left( x, y' \right) . \end{aligned}$$

By this observation, the Fubini theorem and the fact that all the lattice random variables together with the profile function W are nonnegative, we get that the above expectation can be estimated above by

$$\begin{aligned}&{\mathbb {E}}_{{\mathbb {Q}}}\int _0^t \left( \sum _{y \in {\mathcal {V}}_{0}^{\langle \infty \rangle } \setminus {\mathcal {V}}_{0}^{\langle M \rangle }} \xi _y(\omega ) W\left( X_s, y \right) \right. \left. + \sum _{y \in {\mathcal {V}}_{0}^{\langle M\rangle }} \sum _{y' \in \pi _{M}^{-1}(y) \backslash \{y\}} \xi _{y'}(\omega ) W\left( X_s, y' \right) \right) \textrm{d}s \\&\qquad \le 2 \mathbb {E_Q}\xi \int _{0}^{t} \sum _{y \in {\mathcal {V}}_{0}^{\langle \infty \rangle } \backslash {\mathcal {V}}_{0}^{\langle M \rangle }} W\left( X_s, y \right) \textrm{d}s. \end{aligned}$$

As \(X_s \in {\mathcal {D}}_M\) for \(0\le s \le t\) and \(y \in {\mathcal {V}}_{0}^{\langle \infty \rangle } \backslash {\mathcal {V}}_{0}^{\langle M \rangle }\), we have \(d_{\left\lfloor M/2 \right\rfloor }(X_s, y)>2\) for \(0\le s\le t\). This gives

$$\begin{aligned} I_1(t,M)&\le \frac{2\kappa }{\mu \left( {\mathcal {K}}^{\langle M\rangle }\right) } \int _{{\mathcal {D}}_M} p(t,x,x) {\textbf{E}}^{x,x}_{t} \left[ \int _0^t \sum _{y \in {\mathcal {V}}_{0}^{\langle \infty \rangle } \backslash {\mathcal {V}}_{0}^{\langle M \rangle }} W(X_s(\omega ), y) \textrm{d}s ; t<\tau _{{\mathcal {D}}_M }\right] \mu (\textrm{d}x)\\&\le c_5 \sup _{z \in {\mathcal {K}}^{\langle \infty \rangle }} \sum _{y \in {\mathcal {V}}_0^{\langle \infty \rangle } \backslash B_{\lfloor M/2 \rfloor }(z,1)} W(z,y), \end{aligned}$$

with a constant \(c_5>0\) independent of M. By (W3) this is a term of a convergent series.

To show that \(I_2(t,M)\) is a term of a convergent series, we just follow the steps from the proof of (a) for an appropriate set \({\mathcal {D}}^{'}_M \subset {\mathcal {D}}_M\). The proof of (b) is finished. \(\square \)

Proof of Lemma 3.3

The proof of this lemma follows the main steps from the proof of [28, Lemma 3.2], but it substantially differs in some details which are critical for the argument. We will focus on explanation of these differences.

First note that due to Lemma 3.2(a), similarly as in the proof of [28, Lemma 3.2], we only need to establish (3.2). We consider a family of measures \((\nu _M)_{M \in \mathbb {Z_+}}\) given by

$$\begin{aligned} \nu _M:= \left( \frac{1}{\mu \left( {\mathcal {K}}^{\langle M\rangle }\right) } \int _{{\mathcal {K}}^{\langle M\rangle }} p(t,x,x) {\textbf{P}}_{t}^{x,x} \mu (dx) \right) ^{\otimes 2} \otimes {\mathbb {Q}}^{\otimes 3}, \quad M \in {\mathbb {Z}}_{+}, \end{aligned}$$
(B.1)

defined with the product space \({\widetilde{\Omega }} = D([0,t],{\mathcal {K}}^{\langle \infty \rangle })^2 \times \Omega ^3\), and nondecreasing sequence of positive integers \(\left( a_M\right) _{M \in {\mathbb {Z}}_{+}}\) such that \(a_M \le M\), \(M \in {\mathbb {Z}}_{+}\). Its values will be chosen later in the proof.

For \(m \in {\mathbb {Z}}_{+}\) we set

$$\begin{aligned} V^{\omega ,m}(x):= \sum _{y \in {\mathcal {V}}_{0}^{\langle \infty \rangle } \cap B_m(x,1)} \xi _{y}(\omega ) W(x,y) \end{aligned}$$
(B.2)

and

$$\begin{aligned} {\widetilde{V}}^{\omega ,m}(x):= \sum _{y \in {\mathcal {V}}_{0}^{\langle \infty \rangle } \setminus B_m(x,1)} \xi _{y}(\omega ) W(x,y), \end{aligned}$$
(B.3)

where the ball \(B_m(x,1)\) is taken in the m-graph metric, i.e. for \(x \in {\mathcal {K}}^{\langle \infty \rangle } \backslash {\mathcal {V}}^{\langle \infty \rangle }_m\) it is equal to \(\Delta _m(x)\) – the only m-complex containing x; for \(x \in {\mathcal {V}}^{\langle \infty \rangle }_m\) it is a sum of m-complexes attached to x (there are \(\text {rank}(x) \in \{1,2,3\}\) of them). We also denote for \(M \in {\mathbb {Z}}_{+}\)

$$\begin{aligned} F_M(w,\omega ):= \textrm{e}^{-\int _0^t V^{\omega ,a_M}(X_s(w)) \textrm{d}s} \quad \text {and} \quad {\widetilde{F}}_M(w,\omega ):= \textrm{e}^{-\int _0^t {\widetilde{V}}^{\omega ,a_M}(X_s(w)) \textrm{d}s}. \end{aligned}$$

With this notation we have

$$\begin{aligned} {\mathbb {E}}_{{\mathbb {Q}}} \left[ \Lambda _M^D - {\mathbb {E}}_{{\mathbb {Q}}} \Lambda _M^D \right] ^2&= \int _{{\widetilde{\Omega }}} \prod _{i=1}^{2} \left( F_M(w_i, \omega _0) {\widetilde{F}}_M(w_i, \omega _0) - F_M(w_i, \omega _i) {\widetilde{F}}_M(w_i, \omega _i) \right) \\&\ \ \ \ \ \ \ \ \ \ \ \ \ \ \times {\textbf{1}}_{\{t<\tau _{{\mathcal {K}}^{\langle M \rangle }}(w_i)\}} \cdot \textrm{d}\nu _M (w_1,w_2,\omega _0,\omega _1,\omega _2) \\&=: \int _{{\widetilde{\Omega }}} {\mathcal {X}}(w_1,w_2,\omega _0,\omega _1,\omega _2) \textrm{d}\nu _M (w_1,w_2,\omega _0,\omega _1,\omega _2). \end{aligned}$$

We now consider the partition of \({\widetilde{\Omega }}\) into three disjoint sets:

$$\begin{aligned} D_0^M&:= \left\{ (w_1,w_2) \in D([0,t],{\mathcal {K}}^{\langle \infty \rangle })^2 : \text {for every } s_1,s_2 \in [0,t] \ d_{a_M} \left( X_{s_1}(w_1),X_{s_2}(w_2)\right)>2 \right\} \times \Omega ^3,\\ D_1^M&:= \bigg \{(w_1,w_2) \in D([0,t],{\mathcal {K}}^{\langle \infty \rangle })^2 : d_{a_M+3}\left( X_{0}(w_1),X_{0}(w_2)\right) >2 \text { and there exist } s_1,s_2 \in (0,t] \\&\ \ \ \ \ \text { such that } d_{a_M}\left( X_{s_1}(w_1),X_{s_2}(w_2)\right) \le 2 \bigg \} \times \Omega ^3,\\ D_2^M&:= {\widetilde{\Omega }} \backslash \left( D_0^M \cup D_1^M\right) . \end{aligned}$$

We will integrate \({\mathcal {X}}\) over each of these sets separately. Let us point out that all these sets are now defined with the m-graph metric \(d_m(x,y)\). We also use this opportunity to correct the definition of the sets \(D_0^M\) and \(D_1^M\) on p. 1272 in [28]: they also should be defined with \(s_1,s_2 \in [0,t]\) as above instead of single \(s \in [0,t]\). Also, \(a_M\) should be used in the definition of \(D_0^M\) instead of \(c_M\).

By following the argument in the proof of the quoted lemma, we get

$$\begin{aligned}{} & {} \prod _{i=1}^{2} \left( F_M(w_i,\omega _0) {\widetilde{F}}_M(w_i, \omega _0) - F_M(w_i,\omega _i) {\widetilde{F}}_M(w_i, \omega _i) \right) \\{} & {} \quad \le \prod _{i=1}^{2} \left( F_M(w_i,\omega _0) - F_M(w_i,\omega _i) \right) \\{} & {} \qquad + 2 - \left( {\widetilde{F}}_M(w_1, \omega _0) {\widetilde{F}}_M(w_2, \omega _2) + {\widetilde{F}}_M(w_2, \omega _0) {\widetilde{F}}_M(w_1, \omega _1) \right) . \end{aligned}$$

For a given \(M \in {\mathbb {Z}}_{+}\) and a path \(X_s(w)\), the functional \(F_M(w,\cdot )\) depends only on those fractal lattice points from \(V_{0}^{\langle \infty \rangle }\) which are in the set \(X_{[0,t]}^{a_M}(w):= \bigcup _{s\in [0,t]} B_{a_M}(X_s(w),1)\). From the definition of \(D_0^M\) we see that on this set we have \(X_{[0,t]}^{a_M}(w_1) \cap X_{[0,t]}^{a_M}(w_2) = \emptyset \) and therefore, the random variables \(F_M(w_1,\omega _0) - F_M(w_1,\omega _1)\) and \(F_M(w_2,\omega _0) - F_M(w_2,\omega _2)\) are \({\mathbb {Q}}^{\otimes 3}\)-independent. In consequence,

$$\begin{aligned} \int _{D_0^M} \prod _{i=1}^{2} \left( F_M(w_i,\omega _0) - F_M(w_i,\omega _i) \right) {\textbf{1}}_{\{t<\tau _{{\mathcal {K}}^{\langle M \rangle }}(w_i)\}} \textrm{d}\nu _M (w_1,w_2,\omega _0,\omega _1,\omega _2) =0 \end{aligned}$$

and, by following the argument in the proof of the cited lemma, including Jensen’s inequality and the assumption that all lattice random variables are nonnegative and integrable, we obtain

$$\begin{aligned} \int _{D_0^M} {\mathcal {X}}\ \textrm{d}\nu _M \le c t \, \mathbb {E_Q}\xi \sup _{x \in {\mathcal {K}}^{\langle \infty \rangle }} \sum _{y \notin B_{a_M}(x,1)} W(x,y), \end{aligned}$$
(B.4)

for a constant \(c>0\), independent of M.

On the set \(D_1^M\) we have

$$\begin{aligned} \sup _{s \in (0,t]} d_{a_M} \left( X_0(w_i), X_s(w_i) \right) >2, \quad \text {for }i=1 \text { or } i=2. \end{aligned}$$
(B.5)

This can be seen as follows (see Fig. 1 for illustration). From the definition of \(D_1^M\) we have that \(d_{a_M+3} \left( X_0(w_1), X_0(w_2) \right) >2\), what means that \(X_0(w_1)\) and \(X_0(w_2)\) are in separate \((a_M+3)\)-complexes (light grey complexes in the figure). If, on the contrary to (B.5), for both \(i \in \{1,2\}\) and all \(s \in (0,t]\) were \(d_{a_M} \left( X_0(w_i), X_s(w_i) \right) \le 2\), then for both \(i=1,2\) the entire path \(X_{s}(w_i)\) would be bounded inside \(B_{a_M}(X_0(w_i),2)\) (dark grey complexes in the figure). That would mean that these two paths of the process up to time t are not closer to each other than in two different \(a_M\)-complexes attached to the vertices of a common \((a_M+3)\)-complex (white in the figure). In fact, they would be in separate \((a_M+1)\)-complexes inside two different \((a_M+2)\)-complexes. This would mean that \(d_{a_M} \left( X_{s_1}(w_1), X_{s_2}(w_2) \right) \ge 6\) for all \(s_1, s_2 \in (0,t]\), because the path realizing the graph distance must pass through another \(a_M\)-complex inside \((a_M+1)\)-complex containing \(X_{s_1}(w_1)\), then through another \((a_M+1)\)-complex inside \((a_M+2)\)-complex containing \(X_{s_1}(w_1)\), so at least two more \(a_M\)-complexes. By symmetry, it must go next through at least three \(a_M\)-complexes inside \((a_M+2)\)-complex containing \(X_{s_2}(w_2)\).

Fig. 1
figure 1

\(X_0(w_i)\) are in separable \((a_M+3)\)-complexes (light grey). If (B.5) does not hold, then the entire paths \(X_s(w_i)\), \(s \in [0,t]\), must be restricted to the respective dark grey \(a_M\)-complexes, contradicting the definition of \(D_1^M\) (colour figure online)

By [27, Lemma A.2] this implies that

$$\begin{aligned} \sup _{s \in (0,t]} |X_0(w_i), X_s(w_i)| \ge c_1 L^{a_M }, \quad \text {for }i=1 \text { or } i=2, \end{aligned}$$

with a constant \(c_1\) independent of M, and since \(0\le F_M(w_i,\omega _k) \le 1\), \(i=1,2\), \(k=0,1,2\), the integral over \(D_1^M\) can be estimated by

$$\begin{aligned} \frac{1}{\mu \left( {\mathcal {K}}^{\langle M\rangle }\right) } \int _{{\mathcal {K}}^{\langle M\rangle }} p(t,x,x) {\textbf{P}}_{t}^{x,x} \left[ \sup _{s \in (0,t]} |X_0(w_i), X_s(w_i)| \ge c_1 L^{a_M } \right] \mu (\textrm{d}x).\qquad \end{aligned}$$
(B.6)

Using the same argument as in the same step of the proof of the cited lemma, we then get that the above expression is less than or equal to

$$\begin{aligned} c_2 \sup _{x \in {\mathcal {K}}^{\langle \infty \rangle }} {\textbf{P}}^x \left[ \sup _{s \in (0,t/2]} |X_0(w_i), X_s(w_i)| \ge c_1 L^{a_M } \right] , \end{aligned}$$
(B.7)

with some \(c_2\), independent of M.

Since the integrand \({\mathcal {X}}\) is not bigger than 1, it is enough to estimate the measure of \(D_2^M\). We have

$$\begin{aligned} \nu _M(D_2^M) \le \frac{ c_3 \mu \{(x,y) \in {\mathcal {K}}^{\langle M \rangle } \times {\mathcal {K}}^{\langle M \rangle }: d_{a_M+3}\left( x,y\right) \le 2 \}}{\left( \mu ({\mathcal {K}}^{\langle M \rangle })\right) ^2} \le \frac{ 4 c_3 N^{a_M+3}}{N^{M}},\qquad \nonumber \\ \end{aligned}$$
(B.8)

with some \(c_3\) independent of M.

We may choose \(a_M=\lfloor M/4 \rfloor \). Then (B.4) is a term of a convergent series by (W3) and (B.7) is a term of convergent series by Lemma A.1. \(\square \)

Rights and permissions

Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Balsam, H., Kaleta, K., Olszewski, M. et al. Density of states for the Anderson model on nested fractals. Anal.Math.Phys. 14, 23 (2024). https://doi.org/10.1007/s13324-024-00880-8

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s13324-024-00880-8

Keywords

Mathematics Subject Classification

Navigation