Skip to main content
Log in

Mass concentration and aging in the parabolic Anderson model with doubly-exponential tails

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

Abstract

We study the non-negative solution \(u=u(x,t)\) to the Cauchy problem for the parabolic equation \(\partial _t u={\varDelta }u+\xi u\) on \(\mathbb Z^d\times [0,\infty )\) with initial data \(u(x,0)=\mathsf{1}_0(x)\). Here \({\varDelta }\) is the discrete Laplacian on \(\mathbb Z^d\) and \(\xi =(\xi (z))_{z\in \mathbb Z^d}\) is an i.i.d. random field with doubly-exponential upper tails. We prove that, for large t and with large probability, most of the total mass \(U(t):=\sum _x u(x,t)\) of the solution resides in a bounded neighborhood of a site \(Z_t\) that achieves an optimal compromise between the local Dirichlet eigenvalue of the Anderson Hamiltonian \({\varDelta }+\xi \) and the distance to the origin. The processes \(t\mapsto Z_t\) and \(t \mapsto \tfrac{1}{t} \log U(t)\) are shown to converge in distribution under suitable scaling of space and time. Aging results for \(Z_t\), as well as for the solution to the parabolic problem, are also established. The proof uses the characterization of eigenvalue order statistics for \({\varDelta }+\xi \) in large sets recently proved by the first two authors.

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. Astrauskas, A.: Extremal theory for spectrum of random discrete Schrödinger operator. II. Distributions with heavy tails. J. Stat. Phys. 146(1), 98–117 (2012)

    Article  MathSciNet  MATH  Google Scholar 

  2. Astrauskas, A.: Extremal theory for spectrum of random discrete Schrödinger operator. III. Localization properties. J. Stat. Phys. 150(5), 889–907 (2013)

    Article  MathSciNet  MATH  Google Scholar 

  3. Astrauskas, A.: From extreme values of i.i.d. random fields to extreme eigenvalues of finite-volume Anderson Hamiltonian. Probab. Surv. 13, 156–244 (2016)

    Article  MathSciNet  MATH  Google Scholar 

  4. Billingsley, P.: Convergence of Probability Measures, 2nd edn. Wiley, New York (1999)

    Book  MATH  Google Scholar 

  5. Biskup, M., König, W.: Long-time tails in the parabolic Anderson model with bounded potential. Ann. Probab. 29(2), 636–682 (2001)

    Article  MathSciNet  MATH  Google Scholar 

  6. Biskup, M., König, W.: Screening effect due to heavy lower tails in one-dimensional parabolic Anderson model. J. Stat. Phys. 102(5/6), 1253–1270 (2001)

    Article  MathSciNet  MATH  Google Scholar 

  7. Biskup, M., König, W.: Eigenvalue order statistics for random Schrödinger operators with doubly-exponential tails. Commun. Math. Phys. 341(1), 179–218 (2016)

    Article  MATH  Google Scholar 

  8. Carmona, R., Molchanov, S.A.: Parabolic Anderson problem and intermittency. Mem. Am. Math. Soc. 108, 518 (1994)

    MathSciNet  MATH  Google Scholar 

  9. Fiodorov, A., Muirhead, S.: Complete localisation and exponential shape of the parabolic Anderson model with Weibull potential field. Electron. J. Probab. 19(58), 1–27 (2014)

    MathSciNet  MATH  Google Scholar 

  10. Grimmett, G.: Percolation, 2nd edn. Springer, Berlin (1999)

    Book  MATH  Google Scholar 

  11. Gärtner, J., den Hollander, F.: Correlation structure of intermittency in the parabolic Anderson model. Probab. Theory Relat. Fields 114, 1–54 (1999)

    Article  MathSciNet  MATH  Google Scholar 

  12. Gärtner, J., König, W., Molchanov, S.: Geometric characterization of intermittency in the parabolic Anderson model. Ann. Probab. 35(2), 439–499 (2007)

    Article  MathSciNet  MATH  Google Scholar 

  13. Gärtner, J., Molchanov, S.: Parabolic problems for the Anderson model I. Intermittency and related topics. Commun. Math. Phys. 132, 613–655 (1990)

    Article  MathSciNet  MATH  Google Scholar 

  14. Gärtner, J., Molchanov, S.: Parabolic problems for the Anderson model II. Second-order asymptotics and structure of high peaks. Probab. Theory Relat. Fields 111, 17–55 (1998)

    Article  MathSciNet  MATH  Google Scholar 

  15. König, W.: The Parabolic Anderson Model, Pathways in Mathematics. Birkhäuser, Basel (2016)

    Google Scholar 

  16. König, W., Lacoin, H., Mörters, P., Sidorova, N.: A two cities theorem for the parabolic Anderson model. Ann. Probab. 37(1), 347–392 (2009)

    Article  MathSciNet  MATH  Google Scholar 

  17. Lacoin, H., Mörters, P.: A scaling limit theorem for the parabolic Anderson model with exponential potential. In: Deuschel, J.-D., et al. (eds.) Probability in Complex Physical Systems in Honour of Erwin Bolthausen and Jürgen Gärtner, Springer Proceedings in Mathematics, vol. 11, pp. 153–179. Springer, Berlin (2012)

    Google Scholar 

  18. Martin, J.B.: Linear growth for greedy lattice animals. Stoch. Proc. Appl. 98, 43–66 (2002)

    Article  MathSciNet  MATH  Google Scholar 

  19. Molchanov, S.: Lectures on random media. In: Bakry, D., Gill, R.D., Molchanov, S. (eds.) Lectures on Probability Theory, Lecture Notes in Mathematics, vol. 1581, pp. 242–411. Springer, Berlin (1994)

    Chapter  Google Scholar 

  20. Mörters, P.: The parabolic Anderson model with heavy-tailed potential. In: Blath, J., Imkeller, P., Rœlly, S. (eds.) Surveys in Stochastic Processes, Proceedings of the 33rd SPA Conference in Berlin, 2009. EMS Series of Congress Reports (2011)

  21. Mörters, P., Ortgiese, M., Sidorova, N.: Ageing in the parabolic Anderson model. Ann. Inst. Henri Poincaré (B) Prob. Stat. 47(4), 969–1000 (2011)

    Article  MathSciNet  MATH  Google Scholar 

  22. Muirhead, S., Pymar, R.: Localisation in the Bouchaud–Anderson model. Stoch. Proc. Appl. 126(11), 3402–3462 (2016)

    Article  MathSciNet  MATH  Google Scholar 

  23. Molchanov, S., Ruzmaikin, A.: Lyapunov exponents and distributions of magnetic fields in dynamo models. In: Freidlin, M. (ed.) The Dynkin Festschrift: Markov Processes and Their Applications, pp. 287–306. Birkhäuser, Basel (1994)

    Chapter  Google Scholar 

  24. Resnick, S.I.: Extreme Values, Regular Variation, and Point Processes. Springer, New York (1987)

    Book  MATH  Google Scholar 

  25. Sznitman, A.-S.: Brownian Motion, Obstacles and Random Media. Springer, Berlin (1998)

    Book  MATH  Google Scholar 

  26. Sidorova, N., Twarowski, A.: Localisation and ageing in the parabolic Anderson model with Weibull potential. Ann. Probab. 42(4), 1666–1698 (2014)

    Article  MathSciNet  MATH  Google Scholar 

  27. van der Hofstad, R., König, W., Mörters, P.: The universality classes in the parabolic Anderson model. Commun. Math. Phys. 267(2), 307–353 (2006)

    Article  MathSciNet  MATH  Google Scholar 

  28. van der Hofstad, R., Mörters, P., Sidorova, N.: Weak and almost sure limits for the parabolic Anderson model with heavy-tailed potential. Ann. Appl. Probab. 18, 2450–2494 (2008)

    Article  MathSciNet  MATH  Google Scholar 

Download references

Acknowledgements

We thank an anonymous referee for several helpful comments and suggestions. The research of WK and RdS was supported by the German DFG Project KO 2205/13-1 “Random mass flow through random potential.” MB was partially supported by the Simons Foundation and the Mathematische Forschungsinstitut Oberwolfach as a Simons visiting professor in Summer 2015, as well as by NSF Grant DMS-1407558 and GAČR Project P201/16-15238S.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Marek Biskup.

Additional information

© 2017 Marek Biskup, Wolfgang König and Renato S. dos Santos. Reproduction, by any means, of the entire article for non-commercial purposes is permitted without charge.

Appendices

Appendix: A tail estimate

In this section, we prove (7.18) for \(\widehat{Y}_t\) given by (7.21) using an approach from [7]. We will strongly rely on Assumption 2.1. The first step concerns the asymptotic for the upper tail of \(\xi \).

Lemma 12.1

For any \(\varepsilon >0\), there exists \(t_0>0\) such that, for all \(t\ge t_0\),

$$\begin{aligned} t^d \,\text{ Prob } \left( \xi (0) > \widehat{a}_t + s d_t \right) \le \mathrm{e}^{-s (1- \varepsilon )} \qquad \forall s \ge 0. \end{aligned}$$
(12.1)

Proof

Recall the definition of F in (2.1). Note that \(t^d = \exp (\mathrm{e}^{F(\widehat{a}_t)})\) to write

$$\begin{aligned}&-\ln \left\{ t^d \text{ Prob } \left( \xi (0) > \widehat{a}_t + s d_t \right) \right\} \nonumber \\&\quad = \mathrm{e}^{F(\widehat{a}_t)} \left( \mathrm{e}^{F(\widehat{a}_t + s d_t)- F(\widehat{a}_t)} - 1 \right) \nonumber \\&\quad \ge \mathrm{e}^{F(\widehat{a}_t)} \left\{ F(\widehat{a}_t + s d_t)- F(\widehat{a}_t) \right\} , \end{aligned}$$
(12.2)

where in the last inequality we used \(\mathrm{e}^x-1 \ge x\). Using (2.2) and the Mean Value Theorem, we obtain \(F(\widehat{a}_t + sd_t)-F(\widehat{a}_t) \ge s d_t (1-\varepsilon ) / \rho \) for all \(s \ge 0\) if t is large enough. Since \(d_t = \rho \mathrm{e}^{-F(\widehat{a}_t)}\), (12.1) follows from (12.2). \(\square \)

Lemma 12.1 will allow us to reduce the sum in (7.18) to \(|x| \le 6 d \theta t / d_t\).

Corollary 2

For any \(\eta \in \mathbb R\), \(\theta \in (0,\infty )\),

(12.3)

Proof

Recall that \(\max _{x \in B_{\widehat{N}_t}} \xi (x) \ge \lambda ^{{\scriptscriptstyle {({1}})}}_{B_{\widehat{N}_t}}\) by (5.5). Using \(a_t = \widehat{a}_t - \chi + o(1)\) and \(\chi \le 2d\), we obtain, for each \(L \in \mathbb N\),

(12.4)

by Lemma 12.1 and (2.6). Since the last integral converges to 0 as \(L \rightarrow \infty \), the claim (12.3) follows. \(\square \)

To control the sum in (7.18) with \(|x| \le t 6 d \theta / d_t\), we will use the following lemma.

Lemma 12.2

There exist \(c_0, \varepsilon >0\) such that, for all large enough t and all \(s \ge 0\),

$$\begin{aligned} \frac{t^d}{(2\widehat{N}_t)^d} \text{ Prob } \left( \widehat{Y}_t(0) > s \right) \le 4 \, \mathrm{e}^{-c_0 s} + t^{-\varepsilon }. \end{aligned}$$
(12.5)

Before we prove Lemma 12.2, let us finish the proof of (7.18).

Proof of (7.18)

By Corollary 2, we may restrict the sum over \(|x| \le t 6 d \theta / d_t\). Fix \(\eta \in \mathbb R\). Taking \(n\ge \theta |\eta |\), if \(|x| \ge nt\) then \(|x|/(\theta t) + \eta \ge 0\). Thus we may bound, by Lemma 12.2,

(12.6)

for a constant \(c_2>0\) and all large enough t. To conclude (7.17), note that the right-hand side of (12.6) converges as \(t \rightarrow \infty \) to

$$\begin{aligned} 4 \int _{|z| \ge n} \mathrm{e}^{-c_0 \left( \frac{|z|}{\theta } + \eta \right) } \mathrm{d}\,z, \end{aligned}$$
(12.7)

which converges itself to 0 as \(n \rightarrow \infty \). \(\square \)

The remainder of this section is dedicated to the proof of Lemma 12.2. Note that, by Assumption 2.1, \(\xi (0)\) has a density f with respect to Lebesgue measure given by

$$\begin{aligned} f(r) = \left\{ \begin{array}{ll} F'(r) \exp \left\{ F(r)-\mathrm{e}^{F(r)} \right\} , &{} r > {\text {essinf}}\,\xi (0),\\ 0 &{} \text {otherwise.} \end{array}\right. \end{aligned}$$
(12.8)

The following bound holds for f.

Lemma 12.3

Fix a finite \({\varLambda }\subset \mathbb Z^d\) and two functions \(\alpha , \varphi :{\varLambda }\rightarrow \mathbb R\). Then, as \(t \rightarrow \infty \),

$$\begin{aligned}&\prod _{x \in {\varLambda }} \frac{f(\widehat{a}_t + \varphi (x) + \alpha (x) d_t)}{f(\widehat{a}_t + \varphi (x))} \nonumber \\&\quad \le \exp \left\{ -(1+ o(1))\sum _{x \in {\varLambda }} \alpha (x) \mathrm{e}^{\frac{\varphi (x)}{\rho }} + o(1) \mathcal L_{\varLambda }(\varphi )\right\} \end{aligned}$$
(12.9)

where \(\mathcal L_{\varLambda }(\varphi )\) is as in (5.9). If \(\alpha (x) \ge 0\) and \(|\varphi (x)| \le M\), then o(1) only depends on M. If \(|\alpha (x)| \vee |\varphi (x)| \le M\), then equality holds in (12.9) with o(1) only depending on M.

Proof

This follows by the arguments in the proof of [7, Lemma 7.5]. \(\square \)

Fix now \(c_0 := \tfrac{1}{4} \mathrm{e}^{-2(d+1)/\rho }\); this will the constant appearing in (12.2). The following corollary is a convenient rephrasing of (12.9):

Corollary 3

There exists \(t_0 >0\) such that, for all \(t \ge t_0\), \(s \ge 0\), \({\varLambda }\subset \mathbb Z^d\) and all \(\alpha , \varphi :{\varLambda }\rightarrow \mathbb R\) with \(\alpha (x) \ge 0\), \(-2(d+1) \le \varphi (x) \le 1\),

$$\begin{aligned} \prod _{x \in {\varLambda }} \frac{f(\widehat{a}_t + \varphi (x) + s \alpha (x)d_t)}{f(\widehat{a}_t + \varphi (x))} \le \exp \left\{ - 2 c_0 s \sum _{x \in {\varLambda }} \alpha (x) + \mathcal L_{{\varLambda }}(\varphi ) \right\} . \end{aligned}$$
(12.10)

We can now prove Lemma 12.2:

Proof of Lemma 12.2

For \(t >0\) such that \(a_t > {\text {essinf}}\,\xi (0)+1\), define the continuous map

$$\begin{aligned} \mathcal F_{t,s}(r) := \left\{ \begin{array}{ll} r &{}\quad \text { if } r \le a_t-1,\\ r - s d_t &{}\quad \text { if } r \ge a_t + s d_t,\\ \text {linear, } &{}\quad \text { otherwise.} \end{array}\right. \end{aligned}$$
(12.11)

Then \(\mathcal F_{t,s}\) is bijective with inverse

$$\begin{aligned} \mathcal F^{-1}_{t,s}(r) := \left\{ \begin{array}{ll} r &{}\quad \text { if } r \le a_t-1,\\ r + s d_t &{}\quad \text { if } r \ge a_t,\\ \text {linear, } &{}\quad \text { otherwise.} \end{array}\right. \end{aligned}$$
(12.12)

Let \(\xi _{t,s}(x) := \mathcal F_{t,s}(\xi (x))\). Then \(\xi _{t,s}(x)\) has a density with respect to \(\xi (x)\) given by

$$\begin{aligned} \frac{\mathrm{d}\,\xi _{t,s}(x)}{\mathrm{d}\,\xi (x)}(r) = \left\{ \begin{array}{ll} 1 &{} \text { if } r \le a_t -1, \\ \left( 1 + s d_t \right) ^{\mathsf{1}\{r < a_t\}} \frac{f(\mathcal F^{-1}_{t,s}(r))}{f(r)} &{} \text { otherwise.} \end{array}\right. \end{aligned}$$
(12.13)

Recalling that \(\lambda ^{{\scriptscriptstyle {({1}})}}_{B_R}(\xi )\) denotes the principal Dirichlet eigenvalue of \({\varDelta }+ \xi \) in \(B_R\), define

$$\begin{aligned} G_{t,s} := \left\{ \xi :\, \lambda ^{{\scriptscriptstyle {({1}})}}_{B_{R_t}}(\xi ) > a_t + s d_t, \, \mathcal L_{B_{R_t}}(\xi -\widehat{a}_t) \le \ln 2, \, \max _{x \in B_{R_t}} \xi (x) \le \widehat{a}_t + 1 \right\} . \end{aligned}$$
(12.14)

Since \(\xi (x)-s d_t \le \xi _{t,s}(x) \le \xi (x)\), \(\xi \in G_{t,s}\) implies \(\xi _{t,s} \in G_{t,0}\). Write

(12.15)

where E denotes expectation with respect to environment law \(\text{ Prob }\). Bound the middle term in (12.15) by

$$\begin{aligned} \left( 1 + s d_t \right) ^{\left| B_{R_t}\right| } \le \mathrm{e}^{s d_t (2R_t+1)^d} \le \mathrm{e}^{s c_0} \end{aligned}$$
(12.16)

for large t by (5.11). For the product term, define \(\varphi (x) := \xi (x) - \widehat{a}_t \le 1\) on \(\mathcal G_{t,0}\), and \(\alpha (x) \in [0,1]\) by the equation \(\xi (x) + s d_t \alpha (x) = \mathcal F^{-1}_{t,s}(\xi (x))\). Note that, if \(\alpha (x) \ne 0\), then \(\varphi (x) > a_t - 1 -\widehat{a}_t \ge -2(d+1)\) for large t; thus, by Corollary 3,

$$\begin{aligned} \prod _{x \in B_{R_t} :\, \xi (x)> a_t -1} \frac{f(\mathcal F^{-1}_{t,s}(\xi (x))) }{f(\xi (x))} \le 2 \exp \left\{ -2 c_0 s \sum _{x \in B_{R_t} :\, \xi (x) > a_t -1} \alpha (x) \right\} \end{aligned}$$
(12.17)

since \(\mathcal L_{B_{R_t}}( \varphi ) \le \ln 2\) on \(G_{t,0}\). Moreover, by (5.5), on \(\mathcal G_{t,0}\) we have \(\xi (x) > a_t\) for some \(x \in B_{R_t}\) and thus also \(\alpha (x) = 1 \). Noting now that, by (12.1) and Lemma 6.4 of [7],

$$\begin{aligned} \text{ Prob } \left( \lambda ^{{\scriptscriptstyle {({1}})}}_{B_{R_t}}(\xi ) > a_t +s d_t \right) \le \text{ Prob } \left( \xi \in G_{t,s} \right) + o(t^{-(d+\varepsilon _0)}) \end{aligned}$$
(12.18)

for some \(\varepsilon _0>0\), we obtain by (12.1412.18)

$$\begin{aligned} \text {Prob} \left( \lambda ^{{\scriptscriptstyle {({1}})}}_{B_{R_t}}(\xi ) \ge a_t + s d_t \right) \le 2 \mathrm{e}^{-c_0 s} \text {Prob} \left( \lambda ^{{\scriptscriptstyle {({1}})}}_{B_{R_t}}(\xi ) \ge a_t \right) + o(t^{-(d+\varepsilon _0)}). \end{aligned}$$
(12.19)

To pass the estimate to \(\lambda ^{{\scriptscriptstyle {({1}})}}_{B_{\widehat{N}_t}}(\xi )\), note first that, by Lemma 7.6 of [7],

$$\begin{aligned} \limsup _{t \rightarrow \infty } \frac{t^d}{(2 R_t)^d}\text {Prob} \left( \lambda ^{{\scriptscriptstyle {({1}})}}_{B_{R_t}}(\xi ) \ge a_t \right) \le 1, \end{aligned}$$
(12.20)

and thus for large t the right-hand side of (12.19) is at most \(3 \, \mathrm{e}^{-c_0 s} (2 R_t/t)^d + o(t^{-(d+\varepsilon _0)})\). Moreover, by Lemma 7.7 of [7] applied to \(t_L : = a_L-\widehat{a}_L +s d_L\) and \(R'_L := (\ln _2 L)^2\),

$$\begin{aligned} \frac{t^d}{(2 \widehat{N}_t)^d} \text {Prob} \left( \lambda ^{{\scriptscriptstyle {({1}})}}_{B_{\widehat{N}_t}}(\xi ) \ge a_t + sd_t\right) \le \widehat{N}^{-d}_t + 4 \, \mathrm{e}^{-c_0 s} + o(t^{-\varepsilon _0}) \end{aligned}$$
(12.21)

for t large enough, noting that \(o(L^{-d})\) and o(1) in equation (7.27) of [7] are uniform on the sequence \(t_L\). Note that the factor 2 multiplying \(R_t\) and \(\widehat{N}_t\) here and not in [7] appears since our boxes have side-length \(2R+1\) while theirs R. Recalling that \(\widehat{N}_t \gg t^{\beta }\) for some \(\beta > 0\) and taking \(\varepsilon := \varepsilon _0 \wedge (\beta d)\), the lemma is proved. \(\square \)

Appendix: Compactification

Let \(\mathfrak {E} := (\mathbb R\times \mathbb R^d) \cup [0,\infty )\) be equipped with a metric \(\mathbf {d}\) defined by setting, for \(\theta , \theta ' \in [0,\infty )\) and \((\lambda , z), (\lambda ', z') \in \mathbb R\times \mathbb R^d\),

$$\begin{aligned} \begin{aligned} \mathbf {d}(\theta , \theta ')&:= \left| \theta - \theta '\right| , \qquad \; \mathbf {d}(\theta , (\lambda , z)) := \mathrm{e}^{-\lambda } + \left| \frac{|z|}{1 \vee \lambda }-\theta \right| ,\\ \mathbf {d}((\lambda , z), (\lambda ', z'))&:= \mathrm{e}^{-\lambda \wedge \lambda '} \left( 1 - \mathrm{e}^{-|\lambda - \lambda '| - |z - z'|} \right) + \left| \frac{|z|}{1 \vee \lambda }-\frac{|z'|}{1 \vee \lambda '}\right| .\\ \end{aligned} \end{aligned}$$
(13.1)

One may verify that \(\mathbf {d}\) is indeed a metric under which \(\mathfrak {E}\) is separable, complete and locally compact. Moreover:

Lemma 13.1

For any \((\theta , \eta ) \in (0,\infty ) \times \mathbb R\), the set \(\mathcal H^\theta _\eta \subset \mathfrak {E}\) defined in (7.15) is relatively compact.

Proof

Note that the closure of \(\mathcal H^\theta _\eta \) in \(\mathfrak {E}\) is given by

$$\begin{aligned} \overline{\mathcal H^\theta _\eta } = \left\{ (\lambda , z) \in \mathbb R\times \mathbb R^d :\, \lambda - \frac{|z|}{\theta } \ge \eta \right\} \cup [0,\theta ]. \end{aligned}$$
(13.2)

Fix a sequence \(({\varXi }_n)_{n \in \mathbb N}\) in \(\overline{\mathcal H^\theta _\eta }\) and consider the following three cases:

  1. 1.

    \({\varXi }_n \in [0, \theta ]\) for infinitely many n;

  2. 2.

    There is an infinite subsequence \({\varXi }_{n_j} = (\lambda _j, z_j) \in \mathbb R\times \mathbb R^d\) and \((\lambda _j)_{j\in \mathbb N}\) is bounded, implying that \(\{{\varXi }_{n_j} :\, j \in \mathbb N\}\) is contained in a compact subset of \(\mathbb R\times \mathbb R^d\);

  3. 3.

    There is an infinite subsequence \({\varXi }_{n_j} = (\lambda _j, z_j) \in \mathbb R\times \mathbb R^d\) and \(\lim _{j \rightarrow \infty }\lambda _j =\infty \). Note that \(\limsup _{j \rightarrow \infty } |z_j|/\lambda _j \le \theta \).

As is directly checked, in each case there exists a subsequence converging in \(\mathfrak {E}\) to a point of \(\overline{\mathcal H^\theta _\eta }\), thus proving the claim. \(\square \)

We finish the section with the following important property of \(\mathfrak {E}\).

Lemma 13.2

For any compact set \(K \subset \mathfrak {E}\), there exist \(\theta \in (0,\infty )\) and \(\eta \in \mathbb R\) such that \(K \cap (\mathbb R\times \mathbb R^d) \subset \mathcal H^{\theta }_\eta \).

Proof

Cover each \(x \in K\) with an open set \(\mathcal H^{\theta _x}_{\eta _x} \cup [0, \theta _x)\) for some \(\theta _x >0, \eta _x \in \mathbb R\). Use compactness to extract a finite subcover corresponding to \(x_1, \ldots , x_N\) and set \(\theta := \max _{i=1}^N \theta _{x_i}\), \(\eta := \min _{i=1}^N \eta _{x_i}\) to obtain the result. \(\square \)

Appendix: Properties of the cost functional

In this section we prove Lemmas 7.5, 7.6, 7.8 and 7.9.

Proof of Lemma 7.5(i)

Fix \(\theta _0 < \theta _1\) and set \((\lambda _i, z_i)={\varXi }^{{\scriptscriptstyle {({1}})}}_\vartheta (\mathcal P)(\theta _i)\), \(i=0, 1\). Then

$$\begin{aligned} \theta _0(\lambda _1 - \lambda _0) \le |\vartheta (\lambda _1, z_1)| - |\vartheta (\lambda _0, z_0)| \le \theta _1 (\lambda _1 - \lambda _0) \end{aligned}$$
(14.1)

by the definition of \({\varPsi }^{{\scriptscriptstyle {({1}})}}_\vartheta (\mathcal P)\), so that all three functions are non-decreasing. Now, if \((\lambda _0, z_0)\ne (\lambda _1, z_1)\), then one of the inequalities above is strict, since otherwise \(\lambda _1 = \lambda _0\), \(|\vartheta (\lambda _1, z_1)|= |\vartheta (\lambda _0, z_0)|\) and we would have \((\lambda _i, z_i) \in \mathfrak {S}^{{\scriptscriptstyle {({1}})}}_{\vartheta }(\mathcal P)(\theta _j)\) for all \(i, j \in \{0,1\}\), implying that \((\lambda _1, z_1) = (\lambda _0,z_0)\) by the definition of \({\varXi }^{{\scriptscriptstyle {({1}})}}_\vartheta (\mathcal P)\). This concludes the proof. \(\square \)

Proof of Lemma 7.5(ii)

We will first consider the case \(|{\text {supp}}\,(\mathcal P)|<\infty \). We may assume \(|{\text {supp}}\,(\mathcal P)| \ge 2\) since otherwise there is nothing to prove.

Consider first the case \(i=1\). \({\varPsi }^{{\scriptscriptstyle {({1}})}}_\vartheta (\mathcal P)\) is continuous as the pointwise maximum of finitely many continuous functions. Lemma 7.5(i) implies that \({\varXi }^{{\scriptscriptstyle {({1}})}}_\vartheta (\mathcal P)\) jumps finitely many times, and thus has left limits; let us to show that it is càdlàg. Fix \(\theta _0 > 0\) and let \((\lambda _0, z_0) := {\varXi }^{{\scriptscriptstyle {({1}})}}_\vartheta (\mathcal P)(\theta _0)\). Note first that, if \((\lambda , z) \in \mathfrak {S}^{{\scriptscriptstyle {({1}})}}_\vartheta (\mathcal P)(\theta _0)\), then \(\psi ^\vartheta _\theta (\lambda ,z) \le \psi ^\vartheta _\theta (\lambda _0,z_0)\) for all \(\theta \ge \theta _0\) because \(\lambda \le \lambda _0\) by definition. On the other hand, if \((\lambda , z) \notin \mathfrak {S}^{{\scriptscriptstyle {({1}})}}_\vartheta (\mathcal P)(\theta _0)\), then there exists \(\delta _{\lambda , z}>0\) such that \(\psi ^\vartheta _\theta (\lambda , z) < \psi ^\vartheta _\theta (\lambda _0, z_0)\) for all \( \theta \in [\theta _0, \theta _0+\delta _{\lambda , z}]\). Setting \(\delta >0\) to be the smallest among these, we can see that

$$\begin{aligned} (\lambda _0, z_0) \in \mathfrak {S}^{{\scriptscriptstyle {({1}})}}_\vartheta (\mathcal P)(\theta ) \subset \mathfrak {S}^{{\scriptscriptstyle {({1}})}}_\vartheta (\mathcal P)(\theta _0) \;\; \forall \, \theta \in \left[ \theta _0, \theta _0+\delta \right] \end{aligned}$$
(14.2)

implying \({\varXi }^{{\scriptscriptstyle {({1}})}}_\vartheta (\mathcal P)(\theta ) = {\varXi }^{{\scriptscriptstyle {({1}})}}_\vartheta (\mathcal P)(\theta _0)\) for all \(\theta \in [\theta _0, \theta _0+\delta ]\), i.e., \({\varXi }^{{\scriptscriptstyle {({1}})}}_\vartheta (\mathcal P)\) is right-continuous.

Assume now by induction that the statement of Lemma 7.5(ii) has been proved in the case \(|{\text {supp}}\,(\mathcal P)|<\infty \) for all \(i \le k-1\), \(k \ge 2\). Note that, by the definition of \({\varPhi }^{{\scriptscriptstyle {({k}})}}_\vartheta \),

$$\begin{aligned} {\varPhi }^{{\scriptscriptstyle {({k}})}}_\vartheta (\mathcal P)(\theta )= \sum _{{\varXi }\in {\text {supp}}\,(\mathcal P)} \mathsf{1}_{\left\{ {\varXi }^{{\scriptscriptstyle {({1}})}}_\vartheta (\mathcal P)(\theta )= {\varXi }\right\} } {\varPhi }^{{\scriptscriptstyle {({k-1}})}}_\vartheta (\mathcal P_{{\varXi }})(\theta ) \end{aligned}$$
(14.3)

where \(\mathcal P_{{\varXi }}(\cdot ) := \mathcal P( \cdot {\setminus } \{{\varXi }\})\). Since \({\varXi }^{{\scriptscriptstyle {({1}})}}_\vartheta (\mathcal P)\) is càdlàg, it follows from the induction hypothesis that \({\varPhi }^{{\scriptscriptstyle {({k}})}}_\vartheta (\mathcal P)\) is also càdlàg. To prove in addition that \({\varPsi }^{{\scriptscriptstyle {({k}})}}_\vartheta (\mathcal P)\) is continuous, we only need to show that, if \({\varXi }_0:={\varXi }^{{\scriptscriptstyle {({1}})}}_\vartheta (\mathcal P)(\theta -)\ne {\varXi }^{{\scriptscriptstyle {({1}})}}_\vartheta (\mathcal P)(\theta )=:{\varXi }\), then \({\varPsi }^{{\scriptscriptstyle {({k-1}})}}_\vartheta (\mathcal P_{{\varXi }_0})(\theta ) = {\varPsi }^{{\scriptscriptstyle {({k-1}})}}_\vartheta (\mathcal P_{{\varXi }})(\theta )\); but this follows from the definition of \({\varPsi }^{{\scriptscriptstyle {({k-1}})}}_\vartheta \) since, by the continuity of \({\varPsi }^{{\scriptscriptstyle {({1}})}}_\vartheta (\mathcal P)\), \(\psi ^\vartheta _\theta ({\varXi }_0) = \psi ^\vartheta _\theta ({\varXi })\). This finishes the proof in the case \(|{\text {supp}}\,(\mathcal P)| < \infty \).

The case \(|{\text {supp}}\,(\mathcal P)|=\infty \) can be reduced to the previous one as follows. First note that we may substitute \((0,\infty )\) by [ab] with \(0< a< b <\infty \) arbitrary. Fix \(i \in \mathbb N\). Since \(\mathcal H^a_\eta \uparrow \mathbb R\times \mathbb R^d\) as \(\eta \rightarrow -\infty \), \(\mathcal H^b_{\eta }\) is relatively compact and \(\mathcal P^\vartheta \in \mathscr {M}_{\mathrm{P}}\), there exists an \(\eta \in \mathbb R\) such that \(i \le |{\text {supp}}\,(\mathcal P^\vartheta ) \cap \mathcal H^a_{\eta }| \le \mathcal P^\vartheta (\mathcal H^b_{\eta }) < \infty \). Noting that, on [ab], \({\varPhi }^{{\scriptscriptstyle {({i}})}}_\vartheta (\mathcal P) = {\varPhi }^{{\scriptscriptstyle {({i}})}}_\vartheta (\mathcal P')\) where \(\mathcal P'(\cdot ):=\mathcal P(\cdot \cap \{(\lambda , z) :(\lambda , \vartheta (\lambda , z)) \in \mathcal H^b_\eta \})\), we fall into the previous case.

For the last statements, note that the proof above shows that \({\varXi }^{{\scriptscriptstyle {({i}})}}_\vartheta (\mathcal P)\) jumps finitely many times in each compact interval \([\theta _1, \theta _2] \subset (0,\infty )\). Moreover, if we have \(\vartheta ({\varXi }^{{\scriptscriptstyle {({1}})}}_\vartheta (\mathcal P)(\theta _1)) \ne 0\) and \({\varXi }^{{\scriptscriptstyle {({1}})}}_\vartheta (\mathcal P)\) is constant in \([\theta _1, \theta _2]\), then \({\varPsi }^{{\scriptscriptstyle {({1}})}}_\vartheta (\mathcal P)\) is strictly increasing in \([\theta _1, \theta _2]\). \(\square \)

Proof of Lemma 7.6

We first consider the case \(1 \le |{\text {supp}}\,(\mathcal P)|<\infty \). By Proposition 3.13 of [24], for t large enough there exist bijections \(T_t : {\text {supp}}\,(\mathcal P) \rightarrow {\text {supp}}\,(\mathcal P_t)\) such that

$$\begin{aligned} \lim _{t\rightarrow \infty }\sup _{{\varXi }\in {\text {supp}}\,(\mathcal P)} {\text {dist}}\,(T_t({\varXi }),{\varXi }) = 0. \end{aligned}$$
(14.4)

Letting \(\mathcal T_t(\lambda , z):=(\lambda , \vartheta _t(\lambda , z))\), by (7.45) and \({\text {supp}}\,(\mathcal P) \cap \mathbb R\times \{0\}=\emptyset \) we also have

$$\begin{aligned} \lim _{t \rightarrow \infty } \sup _{{\varXi }\in {\text {supp}}\,(\mathcal P)} {\text {dist}}\,( \mathcal T_t \circ T_t ({\varXi }), {\varXi }) = 0, \end{aligned}$$
(14.5)

and \(\mathcal T_t \circ T_t\) is a bijection onto \({\text {supp}}\,(\mathcal P_t^{\vartheta _t})\). In particular, \(\mathcal P_t^{\vartheta _t} \rightarrow \mathcal P\).

To characterize the jump times of our processes, the following definition will be useful: For \(\vartheta : \mathbb R\times \mathbb R^d\), \({\varXi }_i = (\lambda _i, z_i) \in \mathbb R\times \mathbb R^d\), \(i=0,1\), and \(\theta > 0\), let

$$\begin{aligned} \mathcal F^{\vartheta }_\theta ({\varXi }_1, {\varXi }_0) := \left\{ \begin{array}{ll} \frac{\left| \vartheta ({\varXi }_1)\right| - \left| \vartheta ({\varXi }_0)\right| }{\lambda _1 - \lambda _0} &{} \text { if } \lambda _1 > \lambda _0 \text { and } \psi ^\vartheta _\theta ({\varXi }_1) < \psi ^\vartheta _\theta ({\varXi }_0), \\ \infty &{} \text { otherwise.}\end{array} \right. \end{aligned}$$
(14.6)

When \(\vartheta (\lambda , z)=z\), we omit it from the notation.

We now proceed with the proof. Let \(a_0 := a\) and, recursively for \(\ell \in \mathbb N\),

$$\begin{aligned} a_{\ell } := \inf \left\{ \theta > a_{\ell -1} :\, \exists \, 1 \le i \le \left| {\text {supp}}\,(\mathcal P)\right| , {\varXi }^{{\scriptscriptstyle {({i}})}}_\vartheta (\mathcal P)(\theta ) \ne {\varXi }^{{\scriptscriptstyle {({i}})}}_\vartheta (\mathcal P)(a_{\ell -1})\right\} . \end{aligned}$$
(14.7)

Note that \({\varXi }^{{\scriptscriptstyle {({i}})}}(\mathcal P)\) jumps finitely many times. Indeed, for \(i=1\) this follows by Lemma 7.5(i), and for \(i \ge 2\), by induction using (14.3). Thus \(\ell _* = \ell _*(a, \mathcal P):= \inf \{\ell \ge 0 :\, a_{\ell +1}=\infty \} < \infty \).

We proceed by induction on \(\ell _*\), starting with \(\ell _* = 0\). Since \(\mathcal P\in \widetilde{\mathscr {M}}^a_\text {P}\), the values \(i \mapsto \psi _a({\varXi }^{{\scriptscriptstyle {({i}})}}(\mathcal P)(a))\) are all distinct, which together with (14.414.5) implies that \({\varXi }^{{\scriptscriptstyle {({i}})}}_{\vartheta _t}(\mathcal P_t)(a) = T_t({\varXi }^{{\scriptscriptstyle {({i}})}}(\mathcal P)(a))\) for all i when t is large enough. In particular, (14.4) implies the result in the case \(\ell _* =0\). Assume by induction that, for some \(L \in \mathbb N\), the statement has been proved for all \(a' \in (0, \infty )\) and \(\mathcal P' \in \widetilde{\mathscr {M}}^{a'}_{\text {P}}\) satisfying \(|{\text {supp}}\,(\mathcal P')| < \infty \) and \(\ell _*(a', \mathcal P') \le L-1\), and suppose that \(\ell _*=\ell _*(a, \mathcal P)=L\) (in which case necessarily \(|{\text {supp}}\,(\mathcal P)|\ge 2\)).

Note now that, because \(\mathcal P\in \widetilde{\mathscr {M}}^{a}_{\text {P}}\), there exists a unique \(i_1\) such that both \({\varXi }^{{\scriptscriptstyle {({i_1}})}}(\mathcal P)\) and \({\varXi }^{{\scriptscriptstyle {({i_1+1}})}}(\mathcal P)\) jump at \(a_1\) while \({\varXi }^{{\scriptscriptstyle {({i}})}}(\mathcal P)\) is continuous at \(a_1\) for all \(i \notin \{i_1, i_1+1\}\). Moreover, \({\varXi }^{{\scriptscriptstyle {({i_1}})}}(\mathcal P)(a_1)\) is the point \({\varXi }\in {\text {supp}}\,(\mathcal P)\) minimizing \(\mathcal F_a({\varXi }, {\varXi }^{{\scriptscriptstyle {({i_1}})}}(\mathcal P)(a))\) [cf. (14.6)], while \(a_1 = \mathcal F_a({\varXi }^{{\scriptscriptstyle {({i_1}})}}(\mathcal P)(a_1), {\varXi }^{{\scriptscriptstyle {({i_1}})}}(\mathcal P)(a))\) and \({\varXi }^{{\scriptscriptstyle {({i_1+1}})}}(\mathcal P)(a_1) = {\varXi }^{{\scriptscriptstyle {({i_1}})}}(\mathcal P)(a)\).

Let \(a^t_\ell \), \(\ell ^t_*\) be the analogues of \(a_\ell \), \(\ell _*\) for \({\varXi }^{{\scriptscriptstyle {({i}})}}_{\vartheta _t}(\mathcal P_t)\), and fix \(a' \in (a_1, a_2) \cap \mathbb Q\). By (14.414.5) and the previous discussion, when t is large enough, \({\varXi }^{{\scriptscriptstyle {({i}})}}_{\vartheta _t}(\mathcal P_t)\) does not jump in \([a, a']\) for all \(i \notin \{i_1, i_1+1\}\), \({\varXi }^{{\scriptscriptstyle {({i_1}})}}_{\vartheta _t}(\mathcal P_t)(a^t_1) = T_t({\varXi }^{{\scriptscriptstyle {({i_1}})}}(\mathcal P)(a_1))\), and \({\varXi }^{{\scriptscriptstyle {({i_1+1}})}}_{\vartheta _t}(\mathcal P_t)(a^t_1) = {\varXi }^{{\scriptscriptstyle {({i_1}})}}_{\vartheta _t}(\mathcal P_t)(a) = T_t({\varXi }^{{\scriptscriptstyle {({i_1}})}}(\mathcal P)(a))\). Moreover, \(a^t_2> a' > a^t_1\) and

$$\begin{aligned} a^t_1= & {} \mathcal F_a^{\vartheta _t}( {\varXi }^{{\scriptscriptstyle {({i_1}})}}_{\vartheta _t}(\mathcal P_t)(a^t_1), {\varXi }^{{\scriptscriptstyle {({i_1}})}}_{\vartheta _t})(a) ) \nonumber \\= & {} \mathcal F_a(\mathcal T_t \circ T_t({\varXi }^{{\scriptscriptstyle {({i_1}})}}(\mathcal P)(a_1)), \mathcal T_t \circ T_t({\varXi }^{{\scriptscriptstyle {({i_1}})}}(\mathcal P)(a))), \end{aligned}$$
(14.8)

allowing us to conclude, by (14.5),

(14.9)

Define now a time change \(\sigma _t:[a, a']\rightarrow [a, a']\) by setting

$$\begin{aligned} \sigma _t(a)=a, \;\;\; \sigma _t(a_1)=a^t_1, \;\;\; \sigma _t(a')=a' \;\;\; \text { and linear otherwise.} \end{aligned}$$
(14.10)

Then, by the previous discussion together with (14.4), (14.5) and (14.9),

$$\begin{aligned} \lim _{t \rightarrow \infty } \, \sup _{1 \le i \le |{\text {supp}}\,(\mathcal P)|} \, \sup _{\theta \in [a, a']} \, \left| \sigma _t(\theta ) - \theta \right| \vee \left| {\varPhi }^{{\scriptscriptstyle {({i}})}}_{\vartheta _t}(\mathcal P_t)(\sigma _t(\theta )) - {\varPhi }^{{\scriptscriptstyle {({i}})}}(\mathcal P)(\theta )\right| = 0. \end{aligned}$$
(14.11)

Since \(\ell _*(a', \mathcal P) = L-1\) and \(\mathcal P\in \widetilde{\mathscr {M}}^{a'}_{\text {P}}\), by the induction hypothesis we can extend \(\sigma _t\) to \([a, \infty )\) in such a way that (14.11) holds with \([a, a']\) substituted by \([a, \infty )\), finishing the proof in the case \(|{\text {supp}}\,(\mathcal P)|<\infty \).

Consider now the case \(|{\text {supp}}\,(\mathcal P)| = \infty \). We may assume without loss of generality that \(c_*\) in (7.46) is not larger than 1. Let us first show (7.47). Fix \(k \in \mathbb N\) and a point \(b \in (a, \infty ) \cap \mathbb Q\). Note that, since \(\mathcal P\in \widetilde{\mathscr {M}}^{a}_{\text {P}}\), b is a continuity point of \({\varPhi }^{{\scriptscriptstyle {({i}})}}(\mathcal P)\) for all \(1 \le i \le k\). Let \(\eta \in \mathbb R\) be negative enough such that, for all t large enough,

$$\begin{aligned}&k \le \left| {\text {supp}}\,(\mathcal P) \cap \mathcal H^{a}_\eta \right| \nonumber \\&\quad = \left| {\text {supp}}\,(\mathcal P_t) \cap \mathcal H^{a}_\eta \right| \le \mathcal P_t(\mathcal H^{2b/c_*}_\eta ) = \mathcal P(\mathcal H^{2b/c_*}_\eta ) < \infty , \end{aligned}$$
(14.12)

which is possible because \(\mathcal P\in \mathscr {M}_{\mathrm{P}}\) and \(\mathcal P_t \rightarrow \mathcal P\). Moreover, since \({\text {supp}}\,(\mathcal P) \cap \mathbb R\times \{0\} = \emptyset \), by (7.457.46) we may also assume that

$$\begin{aligned} k \le \left| {\text {supp}}\,(\mathcal P_t^{\vartheta _t}) \cap \mathcal H^a_\eta \right| \end{aligned}$$
(14.13)

and

$$\begin{aligned} {\text {supp}}\,(\mathcal P_t^{\vartheta _t}) \cap \mathcal H^b_\eta \subset \mathcal T_t\left( {\text {supp}}\,(\mathcal P_t) \cap \mathcal H^{2b/c_*}_\eta \right) , \end{aligned}$$
(14.14)

where \(\mathcal T_t\) is defined right before (14.5). Now (14.1214.14) imply that, on [ab], \({\varPhi }^{{\scriptscriptstyle {({i}})}}(\mathcal P) = {\varPhi }^{{\scriptscriptstyle {({i}})}}(\mathcal P')\) and \({\varPhi }^{{\scriptscriptstyle {({i}})}}_{\vartheta _t}(\mathcal P_t) = {\varPhi }^{{\scriptscriptstyle {({i}})}}_{\vartheta _t}(\mathcal P'_t)\) for all \(1 \le i \le k\), where \(\mathcal P'(\cdot ) := \mathcal P(\cdot \cap \mathcal H^{2b/c_*}_\eta )\) and analogously for \(\mathcal P'_t\). Since \(\mathcal P'_t \rightarrow \mathcal P'\), (7.47) follows by the previous case and Theorem 16.2 of [4]. The convergence \(\mathcal P_t^{\vartheta _t} \rightarrow \mathcal P\) follows from (14.14), (7.45) and \(\mathcal P_t \rightarrow \mathcal P\) (note that b, \(\eta \) above can be taken arbitrarily large, respec. negative). \(\square \)

Proof of Lemma 7.8

By Lemma 7.7, it suffices to show \((4) \Rightarrow (1)\). Arguing as at the end of the proof of Lemma 7.6, we reduce to the case \(|{\text {supp}}\,(\mathcal P)| < \infty \). Denote by \(\pi :\mathbb R\times \mathbb R^d \rightarrow \mathbb R^d\) the projection on the second coordinate, i.e., the map \(\pi (\lambda , z) := z\). For \(z \in \pi ({\text {supp}}\,(\mathcal P))\), set \(\lambda _z := \max \{\lambda :(\lambda , z) \in {\text {supp}}\,(\mathcal P)\}\) and define \(\widehat{\mathcal P} := \sum _{z \in \pi ({\text {supp}}\,(\mathcal P))} \delta _{(\lambda _z, z)}\). Note that \(\pi \) is injective over \({\text {supp}}\,(\widehat{\mathcal P})\), and that \({\varXi }^{{\scriptscriptstyle {({1}})}}(\mathcal P) = {\varXi }^{{\scriptscriptstyle {({1}})}}(\widehat{\mathcal P})\). By (14.414.5), when t is large enough, \(\pi \) is injective over the support of \(\widehat{\mathcal P}_t:= \widehat{\mathcal P} \circ T_t^{-1}\), and moreover \({\varXi }^{{\scriptscriptstyle {({1}})}}_{\vartheta _t}(\mathcal P_t)(\theta ) = {\varXi }^{{\scriptscriptstyle {({1}})}}_{\vartheta _t} (\widehat{\mathcal P}_t)(\theta )\) for all \(\theta \in [a,b]\). This concludes the proof. \(\square \)

Proof of Lemma 7.9

For \((\lambda , z) \in \mathbb R\times (\mathbb R^d {\setminus } \{0\})\), let

$$\begin{aligned} \mathcal A(\lambda , z) := \left\{ (\lambda ', z') \in \mathbb R\times \mathbb R^d :\, \begin{array}{l} \psi _a(\lambda ', z')> \psi _a(\lambda , z) \text { or}\\ \psi _a(\lambda ', z') = \psi _a(\lambda , z) \text { and } \lambda '>\lambda \end{array} \right\} . \end{aligned}$$
(14.15)

By the definition of \(\mathcal P^\vartheta \), \(\mathcal F_a^{\vartheta }(\mathcal P, \lambda , z) = \mathcal P^\vartheta \left\{ \mathcal A(\lambda , \vartheta (\lambda , z)) \right\} \). Since \(\vartheta _t(\lambda _t, z_t) \rightarrow z_*\) by (7.45) and \(\mathcal P_t^{\vartheta _t} \rightarrow \mathcal P\) by Lemma 7.6, we may assume that \(\vartheta _t(\lambda , z) = z\) for all \((\lambda ,z) \in \mathbb R\times \mathbb R^d\).

Now, since \(\mathcal P\in \widetilde{\mathscr {M}}^a_{\text {P}}\), \(\mathcal F_a(\mathcal P,\lambda _*, z_*) = \mathcal P\left\{ \mathcal H^a_{\psi _a(\lambda _*, z_*)} \right\} \) and there exists a \(\delta >0\) such that

$$\begin{aligned} \mathcal P\left\{ \overline{\mathcal H^{a}_{\psi _a(\lambda _*, z_*) - \delta }} \right\} = 1 + \mathcal P\left\{ \overline{\mathcal H^a_{\psi _a(\lambda _*, z_*) + \delta }} \right\} . \end{aligned}$$
(14.16)

On the other hand, since \(\mathcal P_t \rightarrow \mathcal P\) and \((\lambda _t, z_t) \rightarrow (\lambda _*, z_*)\), when t is large we also have

$$\begin{aligned} \mathcal P_t \left\{ \overline{\mathcal H^a_{\psi _a(\lambda _*, z_*)\pm \delta }} \right\} = \mathcal P\left\{ \overline{\mathcal H^a_{\psi _a(\lambda _*, z_*)\pm \delta }} \right\} \end{aligned}$$
(14.17)

and

$$\begin{aligned} (\lambda _t, z_t) \in \mathcal H^{a}_{\psi _b(\lambda _*, z_*)-\delta } {\setminus } \mathcal H^a_{\psi _a(\lambda _*, z_*)+\delta }. \end{aligned}$$
(14.18)

In particular, for all t large enough,

$$\begin{aligned} \begin{aligned} \mathcal P_t \left\{ \mathcal A(\lambda _t, z_t) \right\}&= \mathcal P_t \left\{ \overline{\mathcal H^a_{\psi _a(\lambda _*, z_*)+\delta }} \right\} \\&= \mathcal P\left\{ \overline{\mathcal H^a_{\psi _a(\lambda _*, z_*)+\delta }} \right\} = \mathcal P\left\{ \mathcal H^a_{\psi _a(\lambda _*, z_*)} \right\} , \end{aligned} \end{aligned}$$
(14.19)

concluding the proof. \(\square \)

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Biskup, M., König, W. & dos Santos, R.S. Mass concentration and aging in the parabolic Anderson model with doubly-exponential tails. Probab. Theory Relat. Fields 171, 251–331 (2018). https://doi.org/10.1007/s00440-017-0777-x

Download citation

  • Received:

  • Revised:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00440-017-0777-x

Mathematics Subject Classification

Navigation