Skip to main content
Log in

Tightness for the cover time of the two dimensional sphere

  • Published:
Probability Theory and Related Fields Aims and scope Submit manuscript

A Correction to this article was published on 06 March 2020

This article has been updated

Abstract

Let \({\mathcal {C}}^*_{\epsilon ,\mathbf{S}^2}\) denote the cover time of the two dimensional sphere by a Wiener sausage of radius \(\epsilon \). We prove that

$$\begin{aligned} \sqrt{{\mathcal {C}}^{*}_{\epsilon ,\mathbf{S}^2} } -\sqrt{\frac{2A_{\mathbf{S}^2}}{\pi }}\left( \log \epsilon ^{-1}-\frac{1}{4}\log \log \epsilon ^{-1}\right) \end{aligned}$$

is tight, where \(A_{\mathbf{S}^2}=4\pi \) denotes the Riemannian area of \(\mathbf{S}^2\).

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

Change history

  • 06 March 2020

    Lemma  9.1 in the paper is incorrect as stated.

Notes

  1. We expect that the case of the torus \(\mathbf {T}^2\) could be handled by our methods, although we do not address this in the paper.

  2. An earlier version of this article claimed such a result, based on a faulty reduction from general manifolds to planar Brownion motion.

  3. Terms depending on z are absorbed in the factor \(\sqrt{L}\).

  4. The formulation in [25, Theorem 5.2] allows for the independence of \(Z_i\) and \(\bar{Z}_i\) of each other, with the change that the parameter of the Bernoulli \(B_i\) equals the variation distance. By splitting \(B_i\) in that formulation, we arrive at the current statement.

  5. As pointed out by the referee, there is a typo in the latter; in the bottom of page 538 and top of page 539, all sums of the form \(\sum _{i_1,i_2,\ldots : \sum i_j=k}\) are missing the multiplicative factor \(k!/\prod _j i_j!\). With these extra factors, the derivation in [4] gives the result claimed there.

References

  1. Abe, Y.: Second order term of cover time for planar simple random walk. arXiv:1709.08151 (2017)

  2. Arguin, L.-P.: Extrema of log-correlated random variables: principles and examples. In: Contucci, P., Giardina, C. (eds.) Advances in Disordered Systems, Random Processes and Some Applications, pp. 166–204. Cambridge University Press, Cambridge (2016)

    Google Scholar 

  3. Arguin, L.-P., Belius, D., Harper, A.J.: Maxima of a randomized Riemann zeta function, and branching random walks. Ann. Appl. Probab. 27, 178–215 (2017)

    Article  MathSciNet  Google Scholar 

  4. Belius, D., Kistler, N.: The subleading order of two dimensional cover times. Probab. Theory Relat. Fields 167, 461–552 (2017)

    Article  MathSciNet  Google Scholar 

  5. Belius, D., Rosen, J., Zeitouni, O.: Barrier estimates for a critical Galton–Watson process and the cover time of the binary tree. Annales Inst. H. Poincare 55, 127–154 (2019)

    Article  MathSciNet  Google Scholar 

  6. Birkhoff, G.: Tres observaciones sobre el algebra lineal. Univ. Nac. TucumRev Ser. A 5, 147–151 (1946)

    Google Scholar 

  7. Bolthausen, E., Deuschel, J.-D., Giacomin, G.: Entropic repulsion and the maximum of the two-dimensional harmonic crystal. Ann. Probab. 29, 1670–1692 (2001)

    Article  MathSciNet  Google Scholar 

  8. Bramson, M.: Minimal displacement of branching random walk. Z. Wahrsch. Verw. Gebiete 45, 89–108 (1978)

    Article  MathSciNet  Google Scholar 

  9. Bramson, M., Ding, J., Zeitouni, O.: Convergence in law of the maximum of the two-dimensional discrete Gaussian free field. Commun. Pure Appl. Math. 69, 62–123 (2016)

    Article  MathSciNet  Google Scholar 

  10. Biskup, M., Louidor, O.: Extreme local extrema of two-dimensional discrete Gaussian free field. Commun. Math. Phys. 345, 271–304 (2016)

    Article  MathSciNet  Google Scholar 

  11. Bramson, M., Zeitouni, O.: Tightness of the recentered maximum of the two-dimensional discrete Gaussian free field. Commun. Pure Appl. Math. 65, 1–20 (2011)

    Article  MathSciNet  Google Scholar 

  12. Brillinger, D.: A particle migrating randomly on a sphere. JTP 10, 429–443 (1997)

    Article  MathSciNet  Google Scholar 

  13. Chung, K.L.: A Course in Probability Theory, 3rd edn. Academic Press, New York (2001)

    Google Scholar 

  14. Cortines, A., Louidor, O., Saglietti, S.: A Scaling Limit for the Cover Time of the Binary Tree. arXiv:1812.10101 (2018)

  15. Dembo, A., Peres, Y., Rosen, J., Zeitouni, O.: Cover times for Brownian motion and random walks in two dimensions. Ann. Math. 160, 433–467 (2004)

    Article  MathSciNet  Google Scholar 

  16. Dembo, A., Peres, Y., Rosen, J., Zeitouni, O.: Thick points for planar Brownian motion and the Erdös–Taylor conjecture for random walk. Acta Math. 186, 239–270 (2001)

    Article  MathSciNet  Google Scholar 

  17. Dembo, A., Rosen, J., Zeitouni, O.: Convergence in law of the cover time and cover time excursion counts on the binary tree. arXiv:1906.07276 (2019)

  18. Ding, J.: Asymptotics of cover times via Gaussian free fields: bounded degree graphs and general trees. Ann. Probab. 42, 464–496 (2014)

    Article  MathSciNet  Google Scholar 

  19. Ding, J., Lee, J., Peres, Y.: Cover times, blanket times, and majorizing measures. Ann. Math. 175, 1409–1471 (2012)

    Article  MathSciNet  Google Scholar 

  20. Ding, J., Roy, R., Zeitouni, O.: Convergence of the centered maximum of log-correlated Gaussian fields. arXiv:1503.04588 (2015)

  21. Ding, J., Zeitouni, O.: A sharp estimate for cover times on binary trees. Stoch. Process. Appl. 122, 2117–2133 (2012)

    Article  MathSciNet  Google Scholar 

  22. Fitzsimmons, P., Pitman, J.: Kac’s moment formula and the Feynman–Kac formula for aditive functionals of a Markov process. Stoch. Proc. Appl. 79, 117–134 (1999)

    Article  Google Scholar 

  23. Fournier, N., Guillin, A.: On the rate of convergence in Wasserstein distance of the empirical measure. Probab. Theory Relat. Fields 162, 707–738 (2015)

    Article  MathSciNet  Google Scholar 

  24. Kistler, N.: Derrida’s random energy models. From spin glasses to the extremes of correlated random fields. In: Gayrard, V., Kistler, N. (eds.) Correlated Random Systems: Five Different Methods. Lecture Notes in Mathematics, vol. 2143, pp. 71–120. Springer, Berlin (2015)

  25. Lindvall, T.: Lectures on the Coupling Method. Wiley, New York (1992)

    MATH  Google Scholar 

  26. Mirsky, L.: Proofs of two theorems on doubly stochastic matrices. Proc. Am. Math. Soc. 9, 371–374 (1958)

    Article  MathSciNet  Google Scholar 

  27. Spivak, M.: A Comprehensive Introduction to Differential Geometry, vol. 2, 2nd edn. Publish or Perish Inc., Berkeley (1975)

    MATH  Google Scholar 

  28. Symeonidis, E.: The Poisson integral for a ball in spaces of constant curvature. Comment. Math. Univ. Carol. 44, 437–460 (2003)

    MathSciNet  MATH  Google Scholar 

  29. Zhai, A.: Exponential concentration of cover times. Electron. J. Probab. (2018). https://doi.org/10.1214/18-EJP149

  30. Zeitouni, O.: Branching random walks and Gaussian fields. In: Probability and Statistical Physics in St. Petersburg, Proceedings of Symposia in Pure Mathematics, vol. 91, pp. 437–471. American Mathematical Society, Providence (2016)

Download references

Acknowledgements

We thank two anonymous referees for a detailed reading of the paper. We particularly thank one of the referees for raising doubts concerning our reduction of the general compact manifold case.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Ofer Zeitouni.

Additional information

Publisher's Note

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

David Belius was supported by SNF grant PP00P2_176918. Part of his work was carried out while he was at the University of Zurich and at the Courant Institute.

Jay Rosen was partially supported by the Simons Foundation. This project has received funding from the European Research Council (ERC) under the European Union’s Horizon 2020 research and innovation programme (Grant Agreement No. 692452).

Appendices

Appendix I: Barrier estimates

Recall that \(\alpha (l)=\rho _{L}(L-l) - l^{\gamma }_{L}\), \(\gamma <1/2\), see (1.14) and (1.18), where

$$\begin{aligned} \rho _{L}= 2-\frac{\log L}{2L}, \end{aligned}$$
(8.1)

and recall that

$$\begin{aligned} t_{z}=\frac{\left( \rho _{L}L+z\right) ^{ 2}}{2}. \end{aligned}$$
(8.2)

Recall the notation \(I_{y }=[y,y+1)\). (Note that in [5], \(I_{y }\) is denoted \(H_{y }\).)

Lemma 8.1

For any \(k<L\) and all \( j \ge 0\) and \(0\le z\le 100k\) with \(t_{z}\in \mathbb {N}\),

$$\begin{aligned} K_{1}\,:= & {} \mathbb {P}\left( \alpha ( l) \le \sqrt{2T_{l}^{0,t_{z}}},\,l=1,\ldots ,k-1;\, \sqrt{2T_{k}^{0, t_{ z }}}\in I_{\alpha (k)+j}\right) \nonumber \\\le & {} ce^{-2k-2z-2k^{\gamma }_{L}+2j} \times \left( 1+z+ k^{ \gamma }_{L}\right) \left( 1+j \right) e^{-\frac{\left( z + k_{L}^{\gamma } - j \right) ^{2}}{4k}}. \end{aligned}$$
(8.3)

Proof of Lemma 8.1

Recall that we denote by \(T_{l}, l=0,1,2,\ldots \) the critical Galton–Watson process with geometric distribution, and let \(P^{\mathrm{GW}}_n\) denote its law when \(T_0=n\). With \(v=\rho _{L}(L-k)- k^{ \gamma }_{L}\) we rewrite \(K_{1} \) as

$$\begin{aligned} K_{1}= & {} P^{\mathrm{GW}}_{t_{z}} \left( \rho _{L}(L-l)- l^{ \gamma }_{L}\le \sqrt{2T_{l} } \,\, \text{ for } l=1,\ldots , k-1;\right. \nonumber \\&\left. \sqrt{2T_{k} }\in I_{v+ j} \right) . \end{aligned}$$
(8.4)

We show below that for any \(0<\gamma <1\), and all \(1\le l\le k-1\)

$$\begin{aligned} l_{L}^{\gamma }\le k_{L}^{\gamma }+l_{k}^{\gamma }, \end{aligned}$$
(8.5)

from which it follows that

$$\begin{aligned} \rho _{L}\left( L-l\right) - l_{L}^{\gamma }\ge \rho _{L}\left( L-k\right) - k_{L}^{\gamma }+\rho _{L}\left( k-l\right) - l_{k}^{\gamma }. \end{aligned}$$

Hence

$$\begin{aligned} K_{1}\le P^{\mathrm{GW}}_{t_{z}}\left( v+\rho _{L}\left( k-l\right) - l_{k}^{\gamma } \le \sqrt{2T_{l}},l=1,\ldots ,k-1;\sqrt{2T_{k}}\in I_{v+ j} \right) , \end{aligned}$$

and by (8.2), \(\sqrt{2t_{z}} =\rho _{L}L+ z=\rho _{L}k+v+ z + k^{\gamma }_{L} \).

Thus using [5, Theorem 1.1], with \(a=\rho _{L}k+v, x=\sqrt{2t_{z}}\) and \(b=v, y=v+ j\),

$$\begin{aligned}&K_{1}\le K_{2}=c\frac{\left( 1+ z+ k_{L}^{\gamma }\right) \left( 1+ j\right) }{k} e^{-\frac{\left( \rho _{L}k+ z + k_{L}^{\gamma } - j \right) ^{2}}{2k}}, \end{aligned}$$

We write out

$$\begin{aligned}&\left( \rho _{L}k+ z + k_{L}^{\gamma } - j \right) ^{2}/2k \nonumber \\&\quad = \left( 2k+ (z + k_{L}^{\gamma } - j) -k\log (L)/2L\right) ^{2}/2k \nonumber \\&\quad =2k+ 2(z + k_{L}^{\gamma } - j) -k\log (L)/L + \left( (z + k_{L}^{\gamma } - j) -k\log (L)/2L\right) ^{2}/2k. \nonumber \\ \end{aligned}$$
(8.6)

Using the concavity of the logarithm, we have \(k\log (L)/L \le \log k\), and using \((r-s)^{2}\ge r^{2}/2-s^{2}\) we see that

$$\begin{aligned} \left( (z + k_{L}^{\gamma } - j) -k\log (L)/2L\right) ^{2}/2k\ge (z + k_{L}^{\gamma } - j)^{2} /4k-o_{L}(1) \end{aligned}$$
(8.7)

It follows that

$$\begin{aligned} K_{2}\le c \left( 1+ z+ k_{L}^{\gamma }\right) \left( 1+j\right) e^{-2k-2z-2k_{L}^{\gamma }+2j} e^{-\frac{\left( z + k_{L}^{\gamma } - j \right) ^{2}}{4k}}. \end{aligned}$$

We now prove (8.5). This certainly holds if \(l\le k/2\), since then \(l_{L}^{\gamma }=l_{k}^{\gamma }=l^{\gamma }\). If \(L/2\le l\le k\), then (8.5) says that

$$\begin{aligned} (L-l)^{\gamma }\le (L-k)^{\gamma }+ (k-l)^{\gamma }, \end{aligned}$$
(8.8)

which follows from concavity.

Equation (8.5) holds if \(k\le L/2\), since then \(l_{L}^{\gamma }=l^{\gamma }\le k^{\gamma }= k_{L}^{\gamma }\). Finally, if \(k/2\le l\le L/2<k\), then (8.5) says that

$$\begin{aligned} l^{\gamma }\le (L-k)^{\gamma }+ (k-l)^{\gamma }. \end{aligned}$$
(8.9)

Since for \( l\le L/2\) we have \(l^{\gamma }\le (L-l)^{\gamma }\), (8.9) follows from (8.8). \(\square \)

Recall that

$$\begin{aligned} \gamma \left( l\right) =\gamma \left( l,L\right) = \rho _{L}(L-l)+ l^{ 1/4}_{L}, \end{aligned}$$
(8.10)

see (4.14).

Lemma 8.2

For all L sufficiently large, and all \(0\le z\le 10 L\) with \(t_{z}\in \mathbb {N}\),

$$\begin{aligned}&\mathbb {P}\left( \gamma (l)\le \sqrt{2T_{l}^{0,t_{z}}}\,\, \text{ for } l=1,\ldots ,L-1;\,T_{L}^{0,t_{z}}=0 \right] \nonumber \\&\quad \asymp (1+z)e^{ -2L-2z}e^{-z^{2}/4L}. \end{aligned}$$
(8.11)

Similar estimates hold if we delete the barrier condition on some fixed finite interval.

Proof of Lemma 8.2

The left hand side of (8.11) can be written in terms of the critical Galton–Watson process \(T_{l}, l=0,1,2,\ldots \) as

$$\begin{aligned}&P^{\mathrm{GW}}_{t_{z}}\left( \rho _{L}(L-l)+ l^{ 1/4}_{L}\le \sqrt{2T_{l} }, \,\, \text{ for } l=1,\ldots , L-1;\,T_{L} =0 \right) . \end{aligned}$$

Equation (8.11) follows immediately from [5, Theorem 1.1], with \(a=\rho _{L}L,\)\(x=\rho _{L}L+z\), \( b=y=0\), since \(\{ T_{L} =0\}=\{ T_{L} \in [0,1] \}\).

For the last statement, we simply note that following the proof of [5, Lemma 2.3] we can show that the analogue of [5, Theorem 1.1] holds where we skip some fixed finite interval. \(\square \)

Lemma 8.3

If \((L-k) \log L/L=o_{L}(1)\) then for all L sufficiently large, and all \(0\le z \le \log L\) with \(t_{z}\in \mathbb {N}\),

$$\begin{aligned}&\mathbb {P}\left( \rho _{L}(L-l)\le \sqrt{2T_{l}^{0,t_{z}}}\,\, \text{ for } l=1,\ldots , k;\,T_{L}^{0,t_{z}}=0 \right) \nonumber \\&\quad \le c (1+z)(L-k)^{ 1/2}e^{ -2L-2z-z^{2}/4L}. \end{aligned}$$
(8.12)

Proof of Lemma 8.3

We rewrite this in terms of the critical Galton–Watson process with geometric offspring distribution, \(T_{l}, l=0,1,2,\ldots \), as

$$\begin{aligned} P^{\mathrm{GW}}_{t_{z}}\left( \rho _{L}(L-l)\le \sqrt{2T_{l} } \,\, \text{ for } l=1,\ldots , k;\,T_{L}=0 \right) . \end{aligned}$$
(8.13)

We condition on \(T_{k}\) as follows: let \(v=\rho _{L}\left( L-k\right) \). Then

$$\begin{aligned} J_{1}:= & {} P^{\mathrm{GW}}_{t_{z}}\left( \rho _{L}\left( L-l\right) \le \sqrt{2T_{l}}, l=1,\ldots ,k,T_{L}=0\right) \\\le & {} \sum _{j=0}^{\infty }P^{\mathrm{GW}}_{t_{z}}\left( \rho _{L}\left( L-l\right) \le \sqrt{2T_{l}},l=1,\ldots ,k,\sqrt{2T_{k}}\in I_{v+j}\right) \\&\times \sup _{u \in I_{j}}P^{\mathrm{GW}}_{( v+u)^{2}/2}\left( T_{L-k}=0\right) \\= & {} \sum _{j=0}^{\infty }P^{\mathrm{GW}}_{t_{z}}\left( \rho _{L}\left( L-l\right) \le \sqrt{2T_{l}},l=1, \ldots ,k,\sqrt{2T_{k}}\in I_{v+j}\right) \\&\times \sup _{u\in I_{j}}\left( 1-\frac{1}{L-k}\right) ^{(v+u)^{2}/2}. \end{aligned}$$

Since, by (3.21),

$$\begin{aligned} P^{\mathrm{GW}}_{t_{z}}\left( \sqrt{2T_{k}}\ge 100L\right) \le ce^{-\left( \rho _{L}L+z-100L\right) ^{2}/2L}\le e^{-100L}, \end{aligned}$$

we obtain

$$\begin{aligned} J_1\le & {} \sum _{j=0}^{100L}P^{\mathrm{GW}}_{t_{z}}\left( \rho _{L}\left( L-l\right) \le \sqrt{2T_{l}},l=1,\ldots ,k,\sqrt{2T_{k}}\in I_{v+j}\right) \\&\times e^{-\frac{( v+j)^{2}}{2(L-k)}}+e^{-100L}=: J_{1,1}+e^{-100L}. \end{aligned}$$

By [5, Theorem 1.1], with \(a=\rho _{L}L\), \(x=\rho _{L}L+z\) recall (8.2), and \(b=v, y=v+j\),

$$\begin{aligned} J_{1,1}\le & {} \sum _{j=0}^{100L}c\frac{(1+z)(1+j) }{k} \sqrt{\frac{\rho _{L}L}{k( v+j)}}\,\, e^{-\frac{\left( \rho _{L}L+z-v-j\right) ^{2}}{2k}}e^{-\frac{( v+j)^{2}}{2(L-k)}}\\\le & {} c\sum _{j=0}^{100L}\frac{(1+z)( 1+j) }{k}\sqrt{\frac{L}{k(v+j)}}\,\, e^{-\frac{\left( \rho _{L}k+z-j\right) ^{2}}{2k}}e^{-\frac{\left( \rho _{L}\left( L-k\right) +j\right) ^{2}}{2(L-k)}}\\\le & {} c(1+z)e^{-\frac{ \rho _{L}^{2}L }{2}-2z}\sum _{j=0}^{100L} \frac{(1+j) }{k}\sqrt{\frac{L}{k( v+j)}}\,\,e^{-\frac{(j-z)^{2}}{2k}-\frac{j^{2}}{2(L-k)}} \\\le & {} c(1+z)e^{-2L-2 z-z^{2}/2k}\sum _{j=0}^{100L}(1+j) \frac{1}{\sqrt{ L-k}}\,\,e^{-\frac{j^{2}}{2(L-k)}}e^{jz/k}. \end{aligned}$$

But

$$\begin{aligned} e^{ -z^{2}/4k}e^{-\frac{j^{2}}{4(L-k)}}e^{jz/k}\le 1 \end{aligned}$$
(8.14)

which follows by considering separately \(j\ge 4(L-k)z/k\) and \(j< 4(L-k)z/k\), since in this case \(j<z/4\) because our condition on k implies that \((L-k)/k \) is tiny. It then follows that

$$\begin{aligned} J_{1,1}\le & {} c(1+z)e^{-2L-2 z-z^{2}/4L} \sum _{j=0}^{100L} ( 1+j)\frac{1}{\sqrt{ L-k}}\,\,e^{-\frac{j^{2}}{4(L-k)}}\\\le & {} c(1+z) e^{-2L-2 z-z^{2}/4L}\sqrt{L-k}. \end{aligned}$$

\(\square \)

Lemma 8.4

If \(k\log L/L=o_{L}(1)\) and \(m=\rho _{L}(L-k)+j\), then for all L sufficiently large, with \(m^{ 2}/2\in \mathbb {N}\),

$$\begin{aligned}&\mathbb {P}\left( \rho _{L}(L-l)\le \sqrt{2T_{l}^{0,k,m^{ 2}/2}}\,\, \text{ for } l=k+1,\ldots , L-1;\,T_{L}^{0,k,m^{ 2}/2}=0 \right) \nonumber \\&\quad \le c \left( 1+j\right) e^{-2\left( L-k\right) -2j -\frac{j^{2}}{4(L-k)}}, \end{aligned}$$
(8.15)

and if \(j\le \eta L\) and we skip the barrier from \(k+1\) to \(k+s\), with \(s\le \eta ' \log L\), for some \(\eta , \eta '<\infty \), then

$$\begin{aligned}&\mathbb {P}\left( \rho _{L}(L-l)\le \sqrt{2T_{l}^{0,k,m^{ 2}/2}}\,\, \text{ for } l=k+s,\ldots , L-1;\,T_{L}^{0,k,m^{ 2}/2}=0 \right) \nonumber \\&\quad \le c\,( 1+j+\sqrt{s}) e^{-2\left( L-k\right) -2j -\frac{j^{2}}{4(L-k)}}. \end{aligned}$$
(8.16)

Proof

We first turn to the probability in (8.15). By the Markov property, the probability in question can be rewritten in terms of the critical Galton–Watson process \(T_{l}, l=0,1,2,\ldots \) as

$$\begin{aligned} M_{1}:=P^{\mathrm{GW}}_{m^{2}/2}\left( \rho _{L}\left( (L-k)-l\right) \le \sqrt{2T_{l}}, l=1,\ldots ,L-k-1;\, T_{L-k}=0\right) . \end{aligned}$$

By [5, Theorem 1.1], with \(a=\rho _{L}\left( L-k\right) , x=m\) and \(b=y=0\),

$$\begin{aligned} M_{1}\le c\frac{\left( 1+j\right) }{L-k} \,\, e^{-\frac{\left( \rho _{L}\left( L-k\right) + j\right) ^{2}}{2\left( L-k\right) }}. \end{aligned}$$
(8.17)

We have

$$\begin{aligned}&e^{-\frac{\left( \rho _{L}\left( L-k\right) + j\right) ^{2}}{2\left( L-k\right) }} =e^{-\frac{\left( \rho _{L}\left( L-k\right) \right) ^{2}}{2(L-k)}- j\rho _{L}-\frac{j^{2}}{2(L-k)}}\\&\le ce^{-\frac{\left( 2\left( L-k\right) -\frac{\log L}{2}\frac{L-k}{L} \right) ^{2}}{2\left( L-k\right) }-2j-\frac{j^{2}}{ 3(L-k)}}\le Ce^{\log L \frac{L-k}{L}}e^{-2\left( L-k\right) -2j -\frac{j^{2}}{3(L-k)}}. \end{aligned}$$

Using the concavity of the logarithm, we get that

$$\begin{aligned} M_{1}\le c \left( 1+j\right) e^{-2\left( L-k\right) -2j -\frac{j^{2}}{3(L-k)}}, \end{aligned}$$

which gives (8.15).

For (8.16) with \(k'=k+ s\), \(v= \rho _{L}(L-k')\) we bound

$$\begin{aligned}&\mathbb {P}\left( \rho _{L}(L-l)\le \sqrt{2T_{l}^{0,k,m^{ 2}/2}}\,\, \text{ for } l=k+s,\ldots , L-1;\,T_{L}^{0,k,m^{ 2}/2}=0 \right) \nonumber \\&\quad \le \sum _{j'=0}^{\infty }P^{\mathrm{GW}}_{m^{2}/2}\left( \sqrt{2T_{s}}\in I_{v+j'}\right) M'_{1} \end{aligned}$$
(8.18)

where

$$\begin{aligned}&M'_{1}=\sup _{u\in I_{v+j'}}P^{\mathrm{GW}}_{u^{2}/2}\left( \rho _{L}\left( (L-k')-l\right) \le \sqrt{2T_{l}},l=1, \ldots ,L-k'-1;\, T_{L-k'}=0\right) . \end{aligned}$$

By [5, Proposition 1.4, Remark 2.2]

$$\begin{aligned}&P^{\mathrm{GW}}_{m^{2}/2}\left( \sqrt{2T_{s}}\in I_{v+j'}\right) \le c\sqrt{\frac{m}{(v+j')s}}\,\,e^{-\left( m-(v+j')\right) ^{2}/2s}\nonumber \\&\quad \le c\sqrt{\frac{1}{s}}\,\,e^{-\left( \rho _{L}s+(j-j')\right) ^{2}/2s}\le c\sqrt{\frac{1}{s}}\,\,e^{-2s-2(j-j')- (j-j')^{2}/3s} \end{aligned}$$
(8.19)

and as in the first part of this proof

$$\begin{aligned} M'_{1}\le c \left( 1+j'\right) e^{-2\left( L-k'\right) -2j' -\frac{j'^{2}}{3(L-k')}}. \end{aligned}$$
(8.20)

Thus

$$\begin{aligned}&\sum _{j'=0}^{\infty }P^{\mathrm{GW}}_{m^{2}/2}\left( \sqrt{2T_{s}}\in I_{v+j'}\right) M'_{1}\nonumber \\&\quad \le c e^{-2\left( L-k\right) -2j }\sqrt{\frac{1}{s}}\,\,\sum _{j'=0}^{\infty } \left( 1+j'\right) e^{ -\frac{j'^{2}}{3(L-k')}} e^{ -(j'-j)^{2}/3s}. \end{aligned}$$
(8.21)

Considering separately the cases where \(j'\le 1.1 j\) and \(j'> 1.1 j\) we see that for L sufficiently large,

$$\begin{aligned} e^{ -\frac{j'^{2}}{3(L-k')}} e^{ -(j'-j)^{2}/3s}\le ce^{ -\frac{j^{2}}{4(L-k)}} e^{ -(j'-j)^{2}/4s}, \end{aligned}$$
(8.22)

and (8.16) follows. \(\square \)

Lemma 8.5

If \(v=\rho _{L}(L-k)+u\) then for L large with \(v^{2}/2\in \mathbb {N}\), and \(1\le k,{\widetilde{k}}, u,j\le L^{3/4}\),

$$\begin{aligned}&\mathbb {P}\left( \rho _{L}(L-l)\le \sqrt{2T^{0,k, v^{2}/2}_{l}},l=k+1,\ldots ,k+{\widetilde{k}};\sqrt{2T^{0,k, v^{2}/2}_{k+{\widetilde{k}}}}\in I_{\rho _{L}(L-k-{\widetilde{k}})+j} \right) \nonumber \\&\quad \le C\frac{\left( 1+u\right) \left( 1+j\right) }{{\widetilde{k}}^{3/2}}e^{-2{\widetilde{k}} -2\left( u-j\right) -\frac{(u-j)^{2}}{4{\widetilde{k}}}}. \end{aligned}$$
(8.23)

In addition, if we skip the barrier from \(k+1\) to \(k+3\),

$$\begin{aligned}&\mathbb {P}\left( \rho _{L}(L-l)\le \sqrt{2T^{0,k, v^{2}/2}_{l}},l=k+3,\ldots ,k+{\widetilde{k}};\sqrt{2T^{0,k, v^{2}/2}_{k+{\widetilde{k}}}}\in I_{\rho _{L}(L-k-{\widetilde{k}})+j} \right) \nonumber \\&\quad \le C\frac{\left( 1+u \right) \left( 1+j\right) }{{\widetilde{k}}^{3/2}}e^{-2{\widetilde{k}}-2\left( u-j\right) -\frac{(u-j)^{2}}{4{\widetilde{k}}}}. \end{aligned}$$
(8.24)

Proof

By the Markov property, the probability in question can be rewritten in terms of the critical Galton–Watson process \(T_{l}, l=0,1,2,\ldots \) as

$$\begin{aligned}&V_{1}:=P^{\mathrm{GW}}_{ v^{2}/2}\left( \rho _{L}\left( L-k-l\right) \le \sqrt{2T_{l}},l= 1,\ldots , {\widetilde{k}};\right. \nonumber \\&\quad \left. \sqrt{2T_{ {\widetilde{k}}}}\in I_{\rho _{L}\left( L-k-{\widetilde{k}}\right) + j } \right) . \end{aligned}$$
(8.25)

This is a linear barrier of length \({\widetilde{k}}\). At the start of the barrier it is at distance u from the starting point, and at the end it is at distance j from the end point. Therefore by [5, Theorem 1.1] we have that the probability is at most

$$\begin{aligned}&c\frac{\left( 1+ u\right) \left( 1+ j\right) }{{\widetilde{k}}^{3/2}}\sqrt{ \frac{\rho _{L}\left( L-k\right) + u}{\rho _{L}\left( L-k-{\widetilde{k}}\right) + j}} \\&\quad e^{-\frac{\left( \rho _{L}\left( L-k\right) + u-\left( \rho _{L}\left( L-k-{\widetilde{k}}\right) + j\right) \right) ^{2}}{2{\widetilde{k}}}}\\&\quad \le c\frac{\left( 1+u\right) \left( 1+j\right) }{{\widetilde{k}}^{3/2}}e^{-\frac{\left( \rho _{L}{\widetilde{k}}+ u- j \right) ^{2}}{2{\widetilde{k}}}}, \end{aligned}$$

where we have bounded the square root using the fact that \(k,{\widetilde{k}}, u,j<L^{3/4}\) so that the ratio inside the square root is \(\asymp 2L/2L=1\). Using \(k,{\widetilde{k}}, u,j<L^{3/4}\) again we see that

$$\begin{aligned} e^{-\frac{\left( \rho _{L}{\widetilde{k}}+ u- j \right) ^{2}}{2{\widetilde{k}}}}\le & {} ce^{-\frac{\left( \rho _{L}{\widetilde{k}}\right) ^{2}}{2{\widetilde{k}}}}e^{ -2\left( u- j\right) }e^{-\frac{\left( u- j \right) ^{2}}{2{\widetilde{k}}}}\\= & {} ce^{-\frac{\left( 2{\widetilde{k}}-\frac{\log L}{2}\frac{{\widetilde{k}}}{L} \right) ^{2}}{2{\widetilde{k}}}}e^{ -2\left( u- j \right) }e^{-\frac{\left( u- j \right) ^{2}}{2{\widetilde{k}}}}\\\le & {} Ce^{- 2{\widetilde{k}} }e^{ -2\left( u- j \right) }e^{-\frac{\left( u- j \right) ^{2}}{2{\widetilde{k}}}}. \end{aligned}$$

This gives (8.23).

Equation (8.24) follows as in the proof of the previous Lemma. \(\square \)

Appendix II: Conditional excursion probabilities

The following is a modification of [16] adapted to our situation. Fix \(k'>k>2\) and let \({{\mathcal {G}}}_k^y\) to be the \(\sigma \)-algebra generated by the excursions from \(\partial B_{d}(y,h_{k-1})\) to \(\partial B_{d}(y,h_{k})\) (and if we start outside \(\partial B_{d}(y,h_{k-1})\) we include the initial excursion to \(\partial B_{d}(y,h_{k-1})\)). Note that \({{\mathcal {G}}}_k^y\) includes the information on the end points of the excursions from \(\partial B_{d}(y,h_{k})\) to \(\partial B_{d}(y,h_{k-1})\).

Recall that \(T_{y,l-1\rightarrow l}^{y,k\overset{n}{\rightarrow }k-1} \) is the number of excursions from \(\partial B_{d}(y,h_{l-1})\rightarrow \partial B_{d}(y,h_{l})\) during the first n excursions from \(\partial B_{d}(y,h_{k })\rightarrow \partial B_{d}(y,h_{k-1})\), see (2.11).

Lemma 9.1

For any \(L-2k\ge k '>k\ge 2\) and \(n>1\), let \(\mathcal {A}_{k'}\) denote an event, measurable on the excursions of the Brownian motion from \(\partial B_{d}(y,h_{k'})\) to \(\partial B_{d}(y,h_{k'-1})\) and on the collection \(\{T_{y,l-1\rightarrow l}^{y,k\overset{n}{\rightarrow }k-1} \}_{l=k',\ldots ,L}\) . Then,

$$\begin{aligned} \mathbb {P}(\mathcal {A}_{k'} \, | {{\mathcal {G}}}_k^y) \le \left( 1+10(k'-k)\frac{h_{k'-1}}{h_{k}}\right) ^n \mathbb {P}(\mathcal {A}_{k'}) \,. \end{aligned}$$
(9.1)

In particular, for all \(m_l ;l=k',\ldots L\), and all \(y \in \mathbf{S}^2\),

$$\begin{aligned}&\mathbb {P}(T_{y,l-1\rightarrow l}^{y,k\overset{n}{\rightarrow }k-1} = m_l ;l=k',\ldots L \, | {{\mathcal {G}}}_k^y) \nonumber \\&\quad \le \left( 1+10(k'-k)\frac{h_{k'-1}}{h_{k}}\right) ^n \mathbb {P}(T_{y,l-1\rightarrow l}^{y,k\overset{n}{\rightarrow }k-1} = m_l ;l=k',\ldots L \,) \,. \end{aligned}$$
(9.2)

The key to the proof of Lemma 9.1 is the following Lemma.

Lemma 9.2

For \(k'>k \ge 2\) and a Brownian path \(X_\cdot \) starting at \(z\in \partial B_{d}(y,h_{k'-1})\), let \(Z_l=T_{y,l-1\rightarrow l}^{y,k'-1\overset{1}{\rightarrow }k} \), \(l=k',\ldots L\), denote the number of excursions of the path from \(\partial B_{d}(y,h_{l-1})\rightarrow \partial B_{d}(y,h_{l})\), prior to \(\bar{\tau }=\inf \{t>0\, :\,X_t\in \partial B_{d}(y,h_{k})\}\), and let F denote an event measurable with respect to the path of the Brownian motion inside \(B_d(y,h_{k'})\), prior to \(\bar{\tau }\). Then, for some \(c<\infty \) and all \(\{ m_l : l=k',\ldots L \}\), uniformly in \(v\in \partial B_{d}(y,h_{k})\) and y,

$$\begin{aligned} \left| \frac{\mathbb {E}^z (F ;Z_l = m_l ,\, l=k',\ldots L \,\big |\,w_{\bar{\tau }}=v)}{\mathbb {E}^z (F; Z_l = m_l ,\, l=k',\ldots L)}-1\right| \le 10 (k'-k)\frac{h_{k'-1}}{h_k}. \end{aligned}$$
(9.3)

In words, conditioning by the endpoint of the excursion at level k has only minor influence on the probability of events involving pieces of excursions at levels larger than \(k'\).

Proof of Lemma 9.2

Without loss of generality we can take \(y=0\). Fixing \(k \ge 2\) and \(z \in \partial B_{d}(0, h_{k'-1})\) it suffices to consider \(\{m_l ,\, l=k',\ldots L \}\) for which \(\mathbb {P}^z\left( Z_{l} =m_l,\, l=k',\ldots L\right) >0\). Fix such \(\{m_l ,\, l=k',\ldots L \}\) and a positive continuous function g on \(\partial B_{d}(0,h_{k})\). Let \(\bar{\tau } = \inf \{ t:\, w_t \in \partial B_{d}(0,h_{k})\}\), \(\tau _0=0\) and for \(i=0,1,\ldots \) define

$$\begin{aligned} \tau _{2i+1}= & {} \inf \{ t \ge \tau _{2i} :\; X_t \in \partial B_{d}(0,h_{k'}) \cup \partial B_{d}(0,h_{k}) \} \\ \tau _{2i+2}= & {} \inf \{ t \ge \tau _{2i+1} :\; X_t \in \partial B_{d}(0,h_{k'-1})\}\,. \end{aligned}$$

Set \(j=m_{k'}\) and let \(Z^j_l\), \(l = k',\ldots L\) be the corresponding number of excursions by the Brownian path prior to time \(\tau _{2j}\). Then, by the strong Markov property at \(\tau _{2j}\),

$$\begin{aligned}&{\mathbb {E}}^z [ g(X_{\bar{\tau }}) ; F,Z_l = m_l,\, l = k',\ldots L ] \\&\quad = \mathbb {E}^z \left[ {\mathbb {E}}^{X_{\tau _{2j}}} (g(X_{\bar{\tau }}); Z_{k'}=0) ;F, Z^j_l =m_l,\, l = k',\ldots L, {\bar{\tau }} \ge \tau _{2j} \right] \,. \end{aligned}$$

and, substituting \(g=1\),

$$\begin{aligned}&\mathbb {P}^z \left( F,Z_l=m_l,\, l = k',\ldots L \right) \\&\quad = \mathbb {E}^z \left[ \mathbb {P}^{X_{\tau _{2j}}} ( Z_{k'}=0) ;F, Z^j_l =m_l,\, l = k',\ldots L, {\bar{\tau }} \ge \tau _{2j} \right] \;. \end{aligned}$$

Consequently,

$$\begin{aligned}&\mathbb {P}^z \left( F, Z_l =m_l,\, l = k',\ldots L \right) \inf _{x\in \partial B_{d}(0,h_{k'-1})} \frac{{\mathbb {E}}^{x} \left( g(X_{\bar{\tau }});\, Z_{k'}=0 \right) }{\mathbb {P}^x \left( Z_{k'}=0\right) }\\&\quad \le \mathbb {E}^z [ g(X_{\bar{\tau }}); F, Z_l =m_l,\, l = k',\ldots L ]\\&\quad \le \mathbb {P}^z \left( F, Z_l =m_l,\, l = k',\ldots L \right) \sup _{x\in \partial B_{d}(0,h_{k'-1})} \frac{{\mathbb {E}}^{x} \left( g(X_{\bar{\tau }});\, Z_{k'}=0 \right) }{\mathbb {P}^x \left( Z_{k'}=0\right) } \,, \end{aligned}$$

and, using again the strong Markov property at time \(\tau _{2}\),

$$\begin{aligned} {\mathbb {E}}^{x} \left( g(X_{\bar{\tau }});\, Z_{k'}=0 \right)= & {} {\mathbb {E}}^{x} \left( g(X_{\bar{\tau }}) \right) - {\mathbb {E}}^{x} \left( {\mathbb {E}}^{X_{\tau _2}} (g(X_{\bar{\tau }}));\, Z_{k'} \ge 1 \right) \\\le & {} {\mathbb {E}}^{x} \left( g(X_{\bar{\tau }}) \right) - \mathbb {P}^{x} (Z_{k'} \ge 1) \inf _{y\in \partial B_{d}(0,h_{k'-1})} {\mathbb {E}}^{y} (g(X_{\bar{\tau }})), \end{aligned}$$

with the reversed inequality if the \(\inf \) is replaced by \(\sup \). Writing \(p=\mathbb {P}^{x} (Z_{k'} \ge 1) =1-1/(k'-k)\) whenever \(x\in \partial B_{d}(0,h_{k'-1})\), c.f. (2.7), it thus follows that

$$\begin{aligned}&\mathbb {E}^z \left[ g(X_{\bar{\tau }}); F,Z_l =m_l,\, l = k',\ldots L \right] \nonumber \\&\quad \le \mathbb {P}^z\left( F,Z_l =m_l,\, l = k',\ldots L \right) \nonumber \\&\quad \cdot {\mathbb {E}}^z\left( g(w_{\bar{\tau }})\right) (1-p)^{-1} \left\{ {\sup _{x\in \partial B_{d}(0,h_{k'-1})} E^x( g(X_{\bar{\tau }})) \over \inf _{y\in \partial B_{d}(0,h_{k'-1})} E^y( g(X_{\bar{\tau }}))} -p\right\} , \end{aligned}$$
(9.4)

with the reversed inequality if the \(\inf \) and \(\sup \) are interchanged.

As in (2.8), let \(p_{B_{d}( 0, h_{k})}( z,x)\) denote the Poisson kernel for \(B_{d}( 0, h_{k})\subseteq \mathbf{S}^{ 2}\), see (2.8).

Recall that

$$\begin{aligned} {\mathbb {E}}^{z'} g (X_{\bar{\tau } }) = \int _{\partial B_{d}(0,h_{k})} p_{B_{d}(0,h_{k})}( z',u) g(u) du. \end{aligned}$$

Therefore, we get the Harnack inequality

$$\begin{aligned} \frac{\sup _{x\in \partial B_{d}(0,h_{k'-1})} {\mathbb {E}}^x( g(X_{\bar{\tau }}))}{\inf _{y\in \partial B_{d}(0,h_{k'-1})} {\mathbb {E}}^y( g(X_{\bar{\tau }}))}\le & {} \frac{\max _{x\in \partial B_{d}(0,h_{k'-1}),u\in \partial B_{d}(0,h_{k })}p_{B_{d}(0,h_{k})}( x,u)}{\min _{y\in \partial B_{d}(0,h_{k'-1}),u\in \partial B_{d}(0,h_{k })} p_{B_{d}(0,h_{k})}( y,u)}\nonumber \\= & {} \frac{\max _{y\in \partial B_{d}(0,h_{k'-1}),u\in \partial B_{d}(0,h_{k })}\sin ^{ 2} ( d( u,y)/2)}{\min _{x\in \partial B_{d}(0,h_{k'-1}),u\in \partial B_{d}(0,h_{k })} \sin ^{ 2} ( d( u,x)/2)} \nonumber \\\le & {} \frac{\sin ^{ 2}((h_k+h_{k'-1})/2) }{\sin ^{ 2}((h_k-h_{k'-1})/2) }. \end{aligned}$$
(9.5)

Writing \(\alpha =(h_k-h_{k'-1})/2\), \(\beta =h_{k'-1}\) we have

$$\begin{aligned} \frac{\sin ^{ 2}((h_k+h_{k'-1})/2) }{\sin ^{ 2}((h_k-h_{k'-1})/2) }.=\frac{\sin (\alpha +\beta ) }{\sin (\alpha ) }=\cos (\beta )+\frac{\sin ( \beta ) }{\tan (\alpha ) }. \end{aligned}$$
(9.6)

Using the bounds, valid for \(|\alpha |,|\beta |<0.1\), \(|\sin (\alpha )|\ge |2\alpha /3|\), \(|\cos \beta -1|<|\beta |/2\), we obtain that the right side of (9.6) is bounded by \(1+4h_{k'-1}/h_k\), and therefore the right side of (9.5) is bounded by \(1+10h_{k'-1}/h_k\). Substituting this bound into (9.4) and using the value of p yields the claim. \(\square \)

Proof of Lemma 9.1

This follows from Lemma 9.2 in the same manner as [16, Lemma 7.3] was derived from [16, Lemma 7.4]: using the strong Markov property, each excursion from \(\partial B_{d}(y,h_{k})\) to \(\partial B_{d}(y,h_{k-1})\) will contribute a multiplicative factor \((1+10 (k'-k)h_{k'-1}/{h_{k}})\) to the probability. We omit further details. \(\square \)

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Belius, D., Rosen, J. & Zeitouni, O. Tightness for the cover time of the two dimensional sphere. Probab. Theory Relat. Fields 176, 1357–1437 (2020). https://doi.org/10.1007/s00440-019-00940-2

Download citation

  • Received:

  • Revised:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00440-019-00940-2

Keywords

Mathematics Subject Classification

Navigation