Skip to main content
Log in

Exact asymptotics of component-wise extrema of two-dimensional Brownian motion

  • Published:
Extremes Aims and scope Submit manuscript

Abstract

We derive the exact asymptotics of

$ {\mathbb {P} \left \{ \underset {t\ge 0}{\sup } \left (X_{1}(t) - \mu _{1} t\right )> u, \ \underset {s\ge 0}{\sup } \left (X_{2}(s) - \mu _{2} s\right )> u \right \} },\ \ u\to \infty , $

where (X1(t), X2(s))t, s≥ 0 is a correlated two-dimensional Brownian motion with correlation ρ ∈ [− 1,1] and μ1, μ2 > 0. It appears that the play between ρ and μ1, μ2 leads to several types of asymptotics. Although the exponent in the asymptotics as a function of ρ is continuous, one can observe different types of prefactor functions depending on the range of ρ, which constitute a phase-type transition phenomena.

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.

Institutional subscriptions

Similar content being viewed by others

References

  • Adler, R.J., Taylor, J.E.: Random Fields and Geometry. Springer Monographs in Mathematics. Springer, New York (2007)

    MATH  Google Scholar 

  • Azaïs, J., Wschebor, M.: Level sets and extrema of random processes and fields. John Wiley & Sons, New York (2009)

    Book  Google Scholar 

  • Dȩbicki, K., Hashorva, E., Ji, L., Rolski, T.: Extremal behavior of hitting a cone by correlated Brownian motion with drift. Stoch. Process. Appl. 128(12), 4171–4206 (2018)

    Article  MathSciNet  Google Scholar 

  • Dėbicki, K., Hashorva, E., Krystecki, K.: Finite-time ruin probability for correlated Brownian motions. arXiv:2004.14015 (2020)

  • Dȩbicki, K., Hashorva, E., Liu, P.: Ruin probabilities and passage times of γ,-reflected Gaussian process with stationary increments. ESAIM: Probability and Statistics 21, 495–535 (2017)

    Article  MathSciNet  Google Scholar 

  • Dȩbicki, K., Hashorva, E., Michna, Z.: Simultaneous ruin probability for two-dimensional Brownian risk model. Accepted for publication in J. Appl. Probab. (2019)

  • Dȩbicki, K., Hashorva, E., Wang, L.: Extremes of vector-valued Gaussian processes. Stochastic Processes and their Applications. In press (2020)

  • Dȩbicki, K., Ji, L., Rolski, T.: Logarithmic asymptotics for probability of component-wise ruin in two-dimensional Brownian model. Risks, 7(83) (2019)

  • Dȩbicki, K., Kosiński, K.M., Mandjes, M., Rolski, T.: Extremes of multidimensional Gaussian processes. Stochastic Process. Appl. 120 (12), 2289–2301 (2010)

    Article  MathSciNet  Google Scholar 

  • Hashorva, E.: Asymptotics bounds for multivariate Gaussian tails. J. Theoret. Probab. 18(1), 79–97 (2005)

    Article  MathSciNet  Google Scholar 

  • Hashorva, E., Ji, L.: Extremes and first passage times of correlated fractional Brownian motions. Stoch. Model. 30(3), 272–299 (2014)

    Article  MathSciNet  Google Scholar 

  • He, H., Keirstead, W.P., Rebholz, J.: Double lookbacks. Math. Financ. 8(3), 201–228 (1998)

    Article  MathSciNet  Google Scholar 

  • Honnappa, H., Jaiswal, P., Pasupathy, R.: Large deviations of gaussian extremes on convex sets. Manuscript. https://web.ics.purdue.edu/pasupath/PAPERS/ldextremes.pdf (2020)

  • Ji, L.: On the cumulative Parisian ruin of multi-dimensional Brownian motion models Accepted for publication in Scandinavian Actuarial Journal (2020)

  • Kou, S., Zhong, H.: First-passage times of two-dimensional Brownian motion. Adv. Appl Prob. 48, 1045–1060 (2016)

    Article  MathSciNet  Google Scholar 

  • Lieshout, P., Math, M.: Mandjes. Tandem Brownian queues. Methods Oper. Res. 66, 275–298 (2007)

    Article  MathSciNet  Google Scholar 

  • Mandjes, M.: Large Deviations for Gaussian Queues: Modelling Communication Networks. Wiley, Chichester (2007)

    Book  Google Scholar 

  • Marshall, A.W., Olkin, I.: A multivariate exponential distribution. J.Amer Statist. Assoc. 62, 30–44 (1967)

    Article  MathSciNet  Google Scholar 

  • Metzler, A.: On the first passage problem for correlated Brownian motion. Stat. Probab. Lett. 80, 277–284 (2010)

    Article  MathSciNet  Google Scholar 

  • Piterbarg, V.I.: Asymptotic Methods in the Theory of Gaussian Processes and Fields, Volume 148 of Translations of Mathematical Monographs. RI, American Mathematical Society, Providence (1996)

    Google Scholar 

  • Piterbarg, V.I.: High extrema of Gaussian chaos processes. Extremes 19(2), 253–272 (2016)

    Article  MathSciNet  Google Scholar 

  • Resnick, S.: Extreme Values, Regular Variation and Point Processes. Springer-Verlag, Berlin (1987)

    Book  Google Scholar 

  • Rogers, L.C.G., Shepp, L.: The correlation of the maxima of correlated Brownian motions. J. Appl. Prob. 43, 880–883 (2006)

    Article  MathSciNet  Google Scholar 

  • Shao, J., Wang, X.: Estimates of the exit probability for two correlated Brownian motions. Adv. Appl. Prob.(45):37–50 (2013)

  • Teunen, M., Goovaerts, M.: Double boundary crossing result for the Browian motion. Scandinavian Actuarial Journal, (2):139–150 (1994)

  • van der Hofstad, R., Honnappa, H.: Large deviations of bivariate Gaussian extrema. Queueing Syst. 93, 333–349 (2019)

    Article  MathSciNet  Google Scholar 

Download references

Acknowledgements

We are thankful to the referees for their careful reading and constructive suggestions which significantly improved the manuscript. TR & KD were partially supported by NCN Grant No 2018/31/B/ST1/00370 (2019-2022).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Lanpeng Ji.

Additional information

Publisher’s note

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

Appendix A: Proofs of Lemmas 4.1–5.1

Appendix A: Proofs of Lemmas 4.1–5.1

In this section we give proofs of Lemmas 4.1–5.1 that are the building blocks of the proofs of Theorems 2.1 and 2.3.

We begin with the analysis of the local behaviour of function \(g(t,s), (t,s)\in (0,\infty )\) at its minimizer in scenarios (i)–(iv) of Proposition 3.1, respectivelly.

Lemma A.1

Assume that μ1 < μ2. We have

  • If \(-1< \rho <\hat \rho _{1}\), then as (t, s) → (0, 0),

    $$ \begin{array}{@{}rcl@{}} g(t_A+t,s_A+s)=g_A(t_A,s_A)&+&\frac{a_1}{2}t^2(1+o(1)) - a_2 ts(1+o(1))\\ &+&\frac{a_3}{2} s^2(1+o(1)), \end{array} $$

    where, with h(ρ) := μ2 − 2(μ1 + μ2)ρ + 3μ1ρ2 > 0,

    $$ \begin{array}{@{}rcl@{}} a_1:&=&\frac{2\mu_1^3(\mu_2-2\mu_1\rho)}{h(\rho)}>0, \ \ a_2:=\frac{-2\rho\mu_1^2(\mu_2-2\mu_1\rho)^2}{h(\rho)}, \\ a_3:&=&\frac{2(\mu_2-2\mu_1\rho)^4(1-2\rho)}{h(\rho)}>0. \end{array} $$
  • If \( \hat \rho _{1} < \rho < \hat \rho _{2} \), then

    • (ii.1), as (t, s) → (0, 0), with s < t (i.e., (t + t, s + s) ∈ A),

      $$ \begin{array}{@{}rcl@{}} g(t^*+t,s^*+s)=g_L(t^*) + b_1(t-s ) (1+o(1)) + \frac{ c_1 }{2}s^2(1+o(1)), \end{array} $$

      where

      $$ \begin{array}{@{}rcl@{}} b_1&:=&\frac{(\rho-1-2\rho^2)+2\rho(\mu_2-\mu_1\rho)s^* +(1+\rho)\mu_1^2{s^*}^2}{(1-\rho)(1+\rho)^2{s^*}^2}>0,\\ c_1&:=& \frac{2 }{ {s^*}^3}\left( 1+\frac{\rho^2(\rho(1-\rho)-(\mu_2-\mu_1\rho) s^*)^2}{(1-\rho^2)^3}\right)>0. \end{array} $$
    • (ii.2), as (t, s) → (0, 0), with s > t (i.e., (t + t, s + s) ∈ B),

      $$ \begin{array}{@{}rcl@{}} g(t^*+t,s^*+s)=g_L(t^*)+b_2(s-t ) (1 + o(1)) + \frac{ c_2 }{2}t^2(1+o(1)), \end{array} $$

      where

      $$ \begin{array}{@{}rcl@{}} b_2&:=&\frac{(\rho-1-2\rho^2)+2\rho(\mu_1-\mu_2\rho)t^* +(1+\rho)\mu_2^2{t^*}^2}{(1-\rho)(1+\rho)^2{t^*}^2}>0, \\ c_2&:=& \frac{2 }{ {t^*}^3}\left( 1+\frac{\rho^2(\rho(1-\rho)-(\mu_1-\mu_2\rho) t^*)^2}{(1-\rho^2)^3}\right)>0. \end{array} $$
    • (ii.3), as (t, s) → (0, 0), with s = t (i.e., (t + t, s + s) ∈ L),

      $$ \begin{array}{@{}rcl@{}} g(t^*+t, s^*+t)=g_L(t^*)+\frac{b_0}{2}t^2(1+o(1)), \end{array} $$

      where \(b_{0}:=\frac {4}{(1+\rho ){t^{*}}^{3}}. \)

  • If \(\rho = \hat \rho _{1}\) (in this case tA = sA = t = s), then

    • (iii.1), as (t, s) → (0, 0), with s < t,

      $$ \begin{array}{@{}rcl@{}} g(t_A+t,s_A+s)&=&g_A(t_A,s_A)+\frac{a_1}{2}t^2(1+o(1))\\ &&- a_2 ts(1+o(1)) +\frac{a_3}{2} s^2(1+o(1)), \end{array} $$
    • (iii.2), as (t, s) → (0, 0), with s > t,

      $$ \begin{array}{@{}rcl@{}} g(t^* + t,s^*+s) = g_L(t^*) + b_2(s - t ) (1 + o(1)) + \frac{ c_2 }{2}t^2(1 + o(1)). \end{array} $$
    • (iii.3), as (t, s) → (0, 0), with s = t,

      $$ \begin{array}{@{}rcl@{}} g(t^*+t, s^*+t)=g_L(t^*)+\frac{b_0}{2}t^2(1+o(1)). \end{array} $$

The proof of Lemma A.1 is tedious but only involves basic calculations using Taylor expansion, and thus it is omitted.

Next we present below a generalized version of the Bonferroni’s inequality. The proof can be found in, e.g., Hashorva and Ji (2014).

Lemma A.2

Let \(({\Omega }, \mathcal {F}, \mathbb {P})\) be a probability space and A1,⋯ , An and B1,⋯ , Bm be n + m events in \(\mathcal {F}\) with n, m ≥ 2. Then

$$ \begin{array}{@{}rcl@{}} &&\sum\limits_{k=1}^n\sum\limits_{l=1}^m{\mathbb{P} \left\{ A_k\cap B_l \right \} }\ge {\mathbb{P} \left\{ \underset{l=1,\ldots, m}{\underset{k=1,\ldots,n} \bigcup }(A_k\cap B_l) \right \} }\ge\sum\limits_{k=1}^n\sum\limits_{l=1}^m{\mathbb{P} \left\{ A_k\cap B_l \right \} }\\ &&\ \ \ \qquad -\sum\limits_{k=1}^n\underset{1\le l_1<l_2\le m}{\sum}{\mathbb{P} \left\{ A_k\cap B_{l_1}\cap B_{l_2} \right \} } -\sum\limits_{l=1}^m\underset{1\le k_1<k_2\le n}{\sum}{\mathbb{P} \left\{ A_{k_1}\cap A_{k_2}\cap B_l \right \} }. \end{array} $$

1.1 A.1 Proof of Lemma 4.1

Let T0 > 0 be a fixed large constant (will be determined later). It is easily seen that

$$ \begin{array}{@{}rcl@{}} r_0(u)&\le& {\mathbb{P} \left\{ \exists_{(t,s) \in [0,T_0]^2\setminus U_1} \ X_1(t)>\sqrt u (1+\mu_1 t), X_2(s)>\sqrt u (1+\mu_2 s) \right \} }\\ &&+{\mathbb{P} \left\{ \exists_{t\ge T_0} \ X_1(t)>\sqrt u (1+\mu_1 t) \right \} }+{\mathbb{P} \left\{ \exists_{s\ge T_0} \ X_2(s)>\sqrt u (1+\mu_2 s) \right \} }. \end{array} $$

Next we consider upper bounds for each term on the right-hand side. According to Lemma 5 of Dȩbicki et al. (2019), for any fixed t, s, there exists a unique index set

$$ I(t,s) \subseteq \{1,2\} $$

such that

$$ \begin{array}{@{}rcl@{}} g(t,s)=(1+{\mu}_1 t, 1+{\mu}_2 s)_{I(t,s)} \ ({\Sigma}_{ts})_{{I(t,s)},{I(t,s)}}^{-1} \ (1+{\mu}_1 t, 1+{\mu}_2 s)_{I(t,s)}^{\top}, \end{array} $$
(25)

and

$$ \begin{array}{@{}rcl@{}} ({\Sigma}_{ts})_{{I(t,s)},{I(t,s)}}^{-1} \ (1+{\mu}_1 t, 1+{\mu}_2 s)_{I(t,s)}^{\top} > {\boldsymbol{0_{I(t,s)}}}. \end{array} $$
(26)

In the above, we use notation that if I ⊂{1, 2}, then for a vector \({\boldsymbol {a}} \in \mathbb {R}^{2}\) we denote by aI = (ai, iI) a sub-block vector of a. Similarly, if further J ⊂{1, 2}, for a matrix \(M=(m_{ij})_{i,j\in \{1,2\}}\in \mathbb {R}^{2\times 2}\) we denote by MI, J = (mij)iI, jJ the sub-block matrix of M determined by I and J. Furthermore, we write \(M_{II}^{-1}=(M_{II})^{-1}\) for the inverse matrix of MII whenever it exists.

Thus,

$$ \begin{array}{@{}rcl@{}} && {\mathbb{P} \left\{ \exists_{(t,s) \in [0,T_0]^2\setminus U_1} \ X_1(t)>\sqrt{u} (1+\mu_1 t), X_2(s)>\sqrt{u} (1+\mu_2 s) \right \} }\\ && \le {\mathbb{P} \left\{ \exists_{(t,s) \in [0,T_0]^2\setminus U_1} \ (1 + \mu_1 t, 1+\mu_2 s)_{I(t,s)}\ ({\Sigma}_{ts})_{{I(t,s)},{I(t,s)}}^{-1} \ (X_1(t),X_2(s))_{I(t,s)}^{\top} >\sqrt u g(t,s) \right \} } \\ &&= {\mathbb{P} \left\{ { \exists_{(t,s) \in [0,T_0]^2\setminus U_1} \ \frac{ Z(t,s)}{g(t,s)}>\sqrt{u} } \right \} }, \end{array} $$
(27)

where

$$ \begin{array}{@{}rcl@{}} Z(t,s):= (1+\mu_1 t, 1+\mu_2 s)_{I(t,s)}\ ({\Sigma}_{ts})_{{I(t,s)},{I(t,s)}}^{-1} \ (X_1(t),X_2(s))_{I(t,s)}^{\top} . \end{array} $$
(28)

Note that

$$ \begin{array}{@{}rcl@{}} \text{Var}\left( \frac{Z(t,s)}{g(t,s)}\right)=\frac{1}{g(t,s)}. \end{array} $$
(29)

In order to apply the Borell-TIS inequality, we first show that

$$ \underset{(t,s)\to(t^{(b)},s^{(b)})}{\limsup}\frac{{\left\lvert {Z(t,s)} \right\rvert}}{g(t,s)}<{\infty},\ \ \ \ \text{almost\ surely} $$

holds for any (t(b), s(b)) on the boundary {(t, s) : t ≥ 0, s = 0}∪{(t, s) : t = 0, s ≥ 0}. In fact, if the above does not hold for some boundary point (t(b), s(b)), then for any M > 0 there exist a sequence \(\{(t_{k},s_{k})\}_{k=1}^{\infty }\) and some measurable set E such that \((t_{k},s_{k}) \to (t^{(b)},s^{(b)}), {\mathbb {P} \left \{ {E} \right \} }>0\) and

$$ \frac{\left\lvert Z(t_{k},s_{k}) \right\rvert}{g(t_{k},s_{k})}\ge M \ \ \ \text{on}\ E $$

for all large enough k. Then we have

$$ \begin{array}{@{}rcl@{}} \text{Var}\left( \frac{Z(t_k,s_k)}{g(t_k,s_k)} \right)\ge M^2 {\mathbb{P} \left\{ {E} \right \} >0}. \end{array} $$
(30)

On the other hand, by Lemma 6 of Dȩbicki et al. (2019) we have g(t, s) = g3(t, s) for all (t, s) ∈{(t, s) : t ≥ 0, s = 0}∪{(t, s) : t = 0, s ≥ 0}, and thus by Eq. 29 and Eq. 13 we have \(\lim _{k\to {\infty }} \text {Var}\left (\frac {Z(t_{k},s_{k})}{g(t_{k},s_{k})} \right ) =0\). This is a contradiction with Eq. 30. Therefore, \(\frac { Z(t,s)}{g(t,s)}, (t,s) \in [0,T_{0}]^{2}\setminus U_{1} \) is almost surely bounded. Consequently, by the Borell-TIS inequality (see, e.g., Adler and Taylor (2007)) we have, for any fixed small constant 𝜃0 > 0

$$ \begin{array}{@{}rcl@{}} {\mathbb{P} \left\{ { \exists_{(t,s) \in [0,T_0]^2\setminus U_1 } \ \frac{Z(t,s)}{g(t,s)}>\sqrt u } \right \} }\le e^{-\frac{(\sqrt u -C_0)^2}{2} \widehat g} \end{array} $$

holds for all u such that

$$ \sqrt u>C_{0}:=\mathbb{E}\left\{ \underset{(t,s)\in [0,T_{0}]^{2}\setminus U_{1} }{\sup} \frac{ Z(t,s)}{g(t,s)}\right\}. $$

Moreover, since Xi is the standard Brownian motion,

$$ \begin{array}{@{}rcl@{}} \lim_{t\to{\infty}}\frac{X_i(t)}{1+\mu_i t} = 0\ \ \ \ \mathrm{almost\ surely}, \end{array} $$

showing that the random process \(\frac {X_{i}(t)}{1+\mu _{i} t}, t\ge T_{0}\) has almost surely bounded sample paths on \([T_{0},{\infty })\). Again by the Borell-TIS inequality

$$ \begin{array}{@{}rcl@{}} {\mathbb{P} \left\{ { \exists_{t\ge T_0} \ X_i(t)>\sqrt u (1+\mu_i t) } \right \} }\le e^{-\frac{(\sqrt u -C_i)^2}{2}\frac{(1+\mu_i T_0)^2}{T_0}} \end{array} $$

holds for all \(\sqrt u>C_{i}:=\mathbb {E}\left \{ \sup _{t\in [T_{0},{\infty })} \frac {X_{1}(t)}{1+\mu _{i} t}\right \}\). Since for all large enough T0 it holds that \( \frac {(1+\mu _{i} T_{0})^{2}}{T_{0}} >\widehat g, \) the claim for r0(u) is established.

Below we consider r1(u). Since (tA, sA) ∈ A, we have from Proposition 3.1 that for any chosen small 𝜃0

$$ \begin{array}{@{}rcl@{}} g(t,s)=g_A(t,s), \ \ (t,s)\in U_1 \subset A, \end{array} $$

and further (cf. Eq. 28)

$$ \begin{array}{@{}rcl@{}} Z(t,s)&=& (1+\mu_1 t, 1+\mu_2 s)\ {\Sigma}_{ts}^{-1} \ (X_1(t),X_2(s))^{\top}\\&& =: h_1(t,s) X_1(t) + h_2(t,s) X_2(s), \ \ (t,s)\in U_1, \end{array} $$

with

$$ \begin{array}{@{}rcl@{}} h_1(t,s) =\frac{(1+\mu_1 t)s-\rho s (1+\mu_2 s))}{ts-\rho^2 s ^2},\ \ \ h_2(t,s) =\frac{(1+\mu_2 s)t-\rho s (1+\mu_1 t))}{ts-\rho^2 s ^2}. \end{array} $$

Thus, similarly to Eq. 27 we conclude that

$$ \begin{array}{@{}rcl@{}} r_1(u)\le {\mathbb{P} \left\{ { \exists_{(t,s) \in U_1 \setminus \triangle^{(1)}_u \times \triangle_u^{(2)} } \ \frac{ Z(t,s)}{g_A(t,s)}>\sqrt u } \right \} } . \end{array} $$
(31)

Since h1(t, s), h2(t, s), gA(t, s), (t, s) ∈ U1 are all smooth functions and

$$ \begin{array}{@{}rcl@{}} \mathbb{E}\left\{ (X_i(t_1)-X_i(t_2))^2\right\}={\left\lvert t_1-t_2 \right\rvert}, \ \ i=1,2 \end{array} $$

one can check that, for all (t1, s1), (t2, s2) ∈ U1,

$$ \begin{array}{@{}rcl@{}} \mathbb{E}\left\{ \left( \frac{Z(t_1,s_1)}{g_A(t_1,s_1)}-\frac{Z(t_2,s_2)}{g_A(t_2,s_2)}\right)^2\right\}\le \text{Const}\cdot ({\left\lvert {t_1-t_2} \right\rvert}+{\left\lvert {s_1-s_2} \right\rvert}). \end{array} $$

Therefore, an application of the Piterbarg’s inequality in Dȩbicki et al. (2017)[Lemma 5.1] (see also Piterbarg (1996) [Theorem 8.1] or Piterbarg (2016) [Theorem 3]) yields that

$$ \begin{array}{@{}rcl@{}} r_1(u)\le {\mathbb{P} \left\{ { \exists_{(t,s) \in U_1 \setminus \triangle^{(1)}_u \times \triangle_u^{(2)} } \ \frac{ Z(t,s)}{g_A(t,s)}>\sqrt u } \right \} }\le C_3 u^{3/2} e^{-\frac{ u }{2} \widetilde g_u}, \end{array} $$
(32)

where C3 > 0 is some constant which does not depend on u and

$$ \begin{array}{@{}rcl@{}} \widetilde g_u : = \underset{(t,s)\in U_1 \setminus \triangle^{(1)}_u \times \triangle_u^{(2)} }{\inf} g_A(t,s). \end{array} $$

Moreover, we have from (i) of Lemma A.1 that for all (tA + t, sA + s) ∈ U1

$$ \begin{array}{@{}rcl@{}} g_A(t_A+t,s_A + s)&\!\ge\!& g_A(t_A,s_A)+\frac{a_1}{2}(1 - \varepsilon)\left( (1 - \rho^2) t^2+\left( \rho t+\frac{\mu_2-2\mu_1\rho}{\mu_1}s\right)^2 \right. \\ &&\left.\ \ \ \ \ +\left( \frac{\mu_2-2\mu_1\rho}{\mu_1}\right)^2\left( \frac{\mu_2-2\mu_1\rho}{\mu_1} (1-2\rho)-1\right) s^2 \right) \end{array} $$
(33)

holds with some small ε > 0, where for all \(-1< \rho <\hat \rho _{1}\) (see also the proof of (b).(i) in Lemma 9 of Dȩbicki et al. (2019) for ρ > 0)

$$ \begin{array}{@{}rcl@{}} \frac{\mu_2-2\mu_1\rho}{\mu_1} (1-2\rho)-1>0. \end{array} $$

Thus

$$ \begin{array}{@{}rcl@{}} \widetilde g_u \ge g_A(t_A,s_A) +\frac{a_1}{2}(1 - \varepsilon)\min\left( (1-\rho^2), \left( \frac{\mu_2-2\mu_1\rho}{\mu_1}\right)^2\left( \frac{\mu_2-2\mu_1\rho}{\mu_1} (1 - 2\rho) - 1\right) \right)\frac{(\ln(u))^2}{u}. \end{array} $$

Inserting the above to Eq. 31 completes the proof.

1.2 A.2 Proof of Lemma 4.2

We first analyze the summand pj, l; u. We set

$$ \begin{array}{@{}rcl@{}} \boldsymbol{b}_{j,l;u} =(a_{j;u}, b_{l;u})^{\top}, \ \ \ \ a_{j;u}=1+\mu_1 (t_A+\frac{jT}{u}), \ \ b_{l;u}=1+\mu_2 (s_A+\frac{lS}{u}). \end{array} $$
(34)

It follows that

$$ \begin{array}{@{}rcl@{}} p_{j,l;u}& = &{\mathbb{P} \left\{ { \underset{s\in[0,S]}{\exists_{t \in [0, T]}} \begin{array}{lll} X_{1}(t_{A}+\frac{jT}{u}+\frac{t}{u})> a_{j;u} \sqrt{u} +\frac{\mu_{1}}{\sqrt u} t \\ X_{2}(s_{A}+\frac{lS}{u}+\frac{s}{u})> b_{l;u} \sqrt{u} +\frac{\mu_{2}}{\sqrt u} s \end{array} } \right \} } \\ & = &{\mathbb{P} \left\{ { \underset{s\in[0,S]}{\exists_{t \in [0, T]}} \begin{array}{lll} X_{1}(t_{A}+\frac{jT}{u})+ X_{1}(t_{A}+\frac{jT}{u}+\frac{t}{u})- X_{1}(t_{A} + \frac{jT}{u})> a_{j;u} \sqrt{u} +\frac{\mu_{1}}{\sqrt u} t \\ X_{2}(s_{A}+\frac{lS}{u})+ X_{2}(s_{A}+\frac{lS}{u}+\frac{s}{u})- X_{2}(s_{A}+\frac{lS}{u})> b_{l;u} \sqrt{u} +\frac{\mu_{2}}{\sqrt u} s \end{array} } \right \} } . \end{array} $$
(35)

Since \((t_{A}+\frac {jT}{u}, s_{A}+\frac {lS}{u})\in A\) for all large u, the covariance matrix of \(\boldsymbol {Z}_{j,l;u}:=(X_{1}(t_{A}+\frac {jT}{u}), X_{2}(s_{A}+\frac {lS}{u}))^{\top }\) is given by

$$ \ {\Sigma}_{j,l;u}= \left( \begin{array}{ll} t_{A}+\frac{jT}{u} & \rho\ (s_{A}+\frac{lS}{u})\\ \rho\ (s_{A}+\frac{lS}{u}) &\quad s_{A}+\frac{lS}{u} \end{array} \right). $$

Thus, the density function of Zj, l; u is given by

$$ \begin{array}{@{}rcl@{}} \phi_{{\Sigma}_{j,l;u}}(\boldsymbol{w})=\frac{1}{\sqrt{(2\pi)^{2}\left\lvert {\Sigma}_{j,l;u} \right\rvert}}\exp\left( -\frac{1}{2}\boldsymbol{w}^{\top}({\Sigma}_{j,l;u})^{-1} \boldsymbol{w} \right),\ \ \ \boldsymbol{w}=(w_1,w_2)^{\top}. \end{array} $$

By conditioning on the value of Zj, l; u we rewrite Eq. 35 as

$$ \begin{array}{@{}rcl@{}} p_{j,l;u} = {\int}_{\mathbb{R}^{2}}\phi_{{\Sigma}_{j,l;u} }(\boldsymbol{w}) \mathbb{P} \left\{\left. \underset{s\in[0,S]}{\exists_{t \in [0, T]}} \begin{array}{lll} X_1(t_A+\frac{jT}{u} + \frac{t}{u})- X_1(t_A + \frac{jT}{u})> a_{j;u} \sqrt{u} +\frac{\mu_1}{\sqrt u} t - w_1\\ X_2(s_A+\frac{lS}{u} + \frac{s}{u})- X_2(s_A + \frac{lS}{u})> b_{l;u} \sqrt{u} + \frac{\mu_2}{\sqrt u} s -w_2 \end{array} \right| \boldsymbol{Z}_{j,l;u} = {\boldsymbol{w}} \right \} d{\boldsymbol{ w}}, \end{array} $$

Using change of variables \(\boldsymbol {w} ={\sqrt {u}} \boldsymbol {b}_{j,l;u}-\boldsymbol {x}/ {\sqrt {u}}\) we further obtain

$$ \begin{array}{@{}rcl@{}} p_{j,l;u}= u^{-1}{\int}_{\mathbb{R}^{2}}\phi_{ {\Sigma}_{j,l;u} }(\sqrt{u} \boldsymbol{b}_{j,l;u}-\boldsymbol{x}/ {\sqrt{u}} ) P_{j,l;u}(\boldsymbol{x}) d{\boldsymbol{x}}, \end{array} $$

where

$$ \begin{array}{@{}rcl@{}} P_{j,l;u}({\boldsymbol{x}}) :=\mathbb{P} \left\{\left. \underset{s\in[0,S]}{\exists_{t \in [0, T]}} \begin{array}{lll} X_1(t_A+\frac{jT}{u}+\frac{t}{u})- X_1(t_A+\frac{jT}{u})> \frac{\mu_1}{\sqrt u} t+\frac{x_1}{\sqrt u}\\ X_2(s_A+\frac{lS}{u}+\frac{s}{u})- X_2(s_A+\frac{lS}{u})> \frac{\mu_2}{\sqrt u} s +\frac{x_2}{\sqrt u} \end{array} \right | \boldsymbol{Z}_{j,l;u} = {\sqrt{u}} \boldsymbol{b}_{j,l;u} - \frac{\boldsymbol{x}}{ {\sqrt{u}}} \right \} . \end{array} $$

Now, we analyse Pj, l; u(x). Due to the fact that (tA, sA) ∈ A, we have for all t ∈ [0, T], s ∈ [0, S], and large enough u

$$ \begin{array}{@{}rcl@{}} t_A+\frac{jT}{u}+\frac{t}{u}\ge t_A+\frac{jT}{u}>s_A+\frac{lS}{u}+\frac{s}{u}\ge s_A+\frac{lS}{u}. \end{array} $$

Thus, by the properties of Brownian motion

$$ \begin{array}{@{}rcl@{}} P_{j,l;u}(\boldsymbol{x}) &= &\mathbb{P} \left\{ \exists_{t\in[0,T]}\ X_1(t)-\mu_1 t> x_1 \right \} \\ && \times \mathbb{P} \left\{ \left. \exists_{s\in[0,S]} \ X_2(s_A+\frac{lS}{u} + \frac{s}{u})- X_2(s_A + \frac{lS}{u})> \frac{\mu_2}{\sqrt u} s + \frac{x_2}{\sqrt u} \right | \boldsymbol{Z}_{j,l;u}= {\sqrt{u}} \boldsymbol{b}_{j,l;u} - \frac{\boldsymbol{x}}{ {\sqrt{u}}} \right \} , \end{array} $$

Next we have

$$ \phi_{{\Sigma}_{j,l;u} }({\sqrt{u}} \boldsymbol{b}_{j,l;u}\!-\boldsymbol{x}/ {\sqrt{u}} ) = \frac{1}{\sqrt{(2\pi)^{2}\left\lvert {\Sigma}_{j,l;u} \right\rvert}}\exp\left( - \frac{1}{2} ({\sqrt{u}} \boldsymbol{b}_{j,l;u}-\boldsymbol{x}/ {\sqrt{u}} )^{\top}({\Sigma}_{j,l;u})^{-1} ({\sqrt{u}} \boldsymbol{b}_{j,l;u} - \boldsymbol{x}/ {\sqrt{u}} ) \right), $$

where the exponent can be rewritten as

$$ \begin{array}{@{}rcl@{}} && ({\sqrt{u}} \boldsymbol{b}_{j,l;u}-\boldsymbol{x}/ {\sqrt{u}} )^{\top}({\Sigma}_{j,l;u})^{-1} ({\sqrt{u}} \boldsymbol{b}_{j,l;u}-\boldsymbol{x}/ {\sqrt{u}} ) \\ &&=u (\boldsymbol{b}_{j,l;u})^{\top}{\Sigma}_{j,l;u}^{-1} \boldsymbol{b}_{j,l;u} - 2 \boldsymbol{x}^{\top} {\Sigma}_{j,l;u}^{-1} \boldsymbol{b}_{j,l;u} + \frac{1}{u } \boldsymbol{x}^{\top} {\Sigma}_{j,l;u}^{-1} \boldsymbol{x} \\ &&= u g_A(t_A+\frac{jT}{u}, s_A+\frac{lS}{u}) - 2 \boldsymbol{x}^{\top} {\Sigma}_{j,l;u}^{-1} \boldsymbol{b}_{j,l;u} + \frac{1}{u} \boldsymbol{x}^{\top} {\Sigma}_{j,l;u}^{-1} \boldsymbol{x} . \end{array} $$

Define

$$ \begin{array}{@{}rcl@{}} f_{j,l;u}(\boldsymbol{x}) := \exp\left( \boldsymbol{x}^{\top}{\Sigma}_{j,l;u}^{-1} \boldsymbol{b}_{j,l;u} - \frac{1}{2u } \boldsymbol{x}^{\top}{\Sigma}_{j,l;u}^{-1} \boldsymbol{x} \right),\ \ \ \ \boldsymbol{x} \in \mathbb{R}^2. \end{array} $$

Thus, it follows that

$$ \begin{array}{@{}rcl@{}} p_1(u) = \frac{u^{-1}}{ 2\pi} \sum\limits_{j=-N_u^{(1)}}^{N_u^{(1)}} \sum\limits_{l=-N_u^{(2)}}^{N_u^{(2)}} \frac{1}{\sqrt{\left\lvert {\Sigma}_{j,l;u} \right\rvert}} \exp\left( -\frac{1}{2} u g_A(t_A + \frac{jT}{u}, s_A+\frac{lS}{u}) \right) {\int}_{\mathbb{R}^{2}} f_{j,l;u}(\boldsymbol{x})P_{j,l;u}(\boldsymbol{x}) d\boldsymbol{x}. \end{array} $$

Further, we obtain from (i) of Lemma A.1 that, for all large enough u,

$$ \begin{array}{@{}rcl@{}} g_A(t_A+\frac{jT}{u}, s_A+\frac{lS}{u}) \sim g_A(t_A,s_A)+\frac{1}{2}\left( a_1 \left( \frac{jT}{u}\right)^2-2a_2 \left( \frac{jT}{u}\right)\left( \frac{lS}{u}\right) +a_3 \left( \frac{lS}{u}\right)^2\right) \end{array} $$

holds uniformly for \(-N_{u}^{(1)}\le j\le {N}_{u}^{(1)}, {-N}_{u}^{(2)}\le l\le {N}_{u}^{(2)}.\)

Consequently, by Lemma A.3 below we obtain

$$ \begin{array}{@{}rcl@{}} \underset{u \to \infty}{\lim} \frac{p_1(u)}{ \exp\left( - g_A(t_A, s_A)u /{2} \right) } = \frac{1}{2\pi\sqrt{\left\lvert {\Sigma}_{0} \right\rvert}} \frac{\mathcal{H} (\mu_1;T)\mathcal{H}(\mu_2-2\mu_1\rho;S)}{TS} {\int}_{\mathbb{R}^2} e^{\frac{- (a_1x_1^2-2a_2 x_1x_2+a_3 x_2^2)}{4}} d \boldsymbol{x}, \end{array} $$

which gives the result for p1(u). The claim for p2(u) follows with the same arguments. \(\Box \)

Lemma A.3

For any T, S > 0

$$ \begin{array}{@{}rcl@{}} \lim\limits_{u\to{\infty}}{\int}_{\mathbb{R}^{2}} f_{j,l;u}(\boldsymbol{x})P_{j,l;u}(\boldsymbol{x}) d \boldsymbol{x}=\mathcal{H} (\mu_1;T)\mathcal{H}(\mu_2-2\mu_1\rho;S) \end{array} $$

holds uniformly for \(-{N}_{u}^{(1)}\le j\le N_{u}^{(1)}, -N_{u}^{(2)}\le l\le N_{u}^{(2)}.\)

We omit the tedious proof of Lemma A.3 since its idea is standard, i.e., it is based on finding a uniform integrable bound for the integrand and then using the dominated convergence theorem.

1.3 A.3 Proof of Lemma 4.3

Let us begin with Π1(u). It follows that

$$ \begin{array}{@{}rcl@{}} {\Pi}_1(u)& = &\sum\limits_{j=-N_u^{(1)}}^{N_u^{(1)}}\underset{ -N_u^{(2)}\le l_1< l_2\le N_u^{(2)}}{\sum}p_{j,l_1,l_2;u}\\ &=&\sum\limits_{j=-N_u^{(1)}}^{N_u^{(1)}} \sum\limits_{l=-N_u^{(2)}}^{N_u^{(2)}} p_{j,l,l+1;u}+ \sum\limits_{j=-N_u^{(1)}}^{N_u^{(1)}} \sum\limits_{l_1=-N_u^{(2)}}^{N_u^{(2)}} \sum\limits_{l_2=l_1+2}^{N_u^{(2)}} p_{j,l_1,l_2;u} =: {\Pi}_{11}(u)+{\Pi}_{12}(u). \end{array} $$

In order to deal with Π11(u) we note that

$$ \begin{array}{@{}rcl@{}} p_{j,l,l+1;u}=p_{j,l;u}+p_{j,l+1;u}-\widetilde p_{j,l;u}, \end{array} $$

where

$$ \widetilde p_{j,l;u}=\mathbb{P} \left\{ \exists_{(t,s)\in\triangle^{(1)}_{j;u}\times (\triangle_{l;u}^{(2)}\cup \triangle_{l+1;u}^{(2)})} \ X_{1}(t)> \sqrt{u}(1 + \mu_{1} t ), X_{2}(s)> \sqrt{u}(1 + \mu_{2} s ) \right \} . $$

Then we have

$$ \begin{array}{@{}rcl@{}} {\Pi}_{11}(u)=\sum\limits_{j=-N^{(1)}_u}^{N_u^{(1)}} \sum\limits_{l=-N^{(2)}_u}^{N_u^{(2)}} (p_{j,l;u}+p_{j,l+1;u}-\widetilde p_{j,l;u}). \end{array} $$

Using the same arguments as in the proof of Lemma 4.2 we obtain

$$ \begin{array}{@{}rcl@{}} \underset{u\to{\infty}}{\lim}\frac{{\Pi}_{11}(u)}{ e^{- g_A(t_A,s_A) u /2 } }= \frac{1}{\mu_1 (\mu_2-2\mu_1\rho)} \left( \frac{2\mathcal{H}(\mu_1;T)\mathcal{H}(\mu_2-2\mu_1\rho;S)}{T S} -\frac{\mathcal{H}(\mu_1;T)\mathcal{H}(\mu_2-2\mu_1\rho;2S)}{T S} \right), \end{array} $$

which gives that

$$ \begin{array}{@{}rcl@{}} \underset{S\to{\infty}}{\limsup}\underset{T\to{\infty}}{\limsup} \underset{u\to{\infty}}{\lim}\frac{{\Pi}_{11}(u)}{ e^{- g_A(t_A,s_A) u /2 } }= 0. \end{array} $$

Next we consider Π12(u) which is more involved. We have (recall Eq. 34 for aj; u, bl; u)

$$ \begin{array}{@{}rcl@{}} p_{j,l_{1},l_{2};u}&=&\mathbb{P} \left\{ \underset{ s_{2}\in[0,S]} {\underset{ s_{1} \in [0, S]} {\underset{ t \in [0, T]} \exists}} \begin{array}{lll} X_{1}(t_{A}+\frac{jT}{u}+\frac{t}{u})> a_{j;u} \sqrt{u} +\frac{\mu_{1}}{\sqrt{u}} t \\ X_{2}(s_{A}+\frac{l_{1}S}{u}+\frac{s_{1}}{u})> b_{l_{1};u} \sqrt{u} +\frac{\mu_{2}}{\sqrt {u}} s_{1}\\ X_{2}(s_{A}+\frac{l_{2}S}{u}+\frac{s_{2}}{u})> b_{l_{2};u} \sqrt{u} +\frac{\mu_{2}}{\sqrt u} s_{2} \end{array} \right \} \\ &\quad\le & \mathbb{P} \left\{ \underset{ s_{2}\in[0,S]} {\underset{ s_{1} \in [0, S]} {\underset{ t \in [0, T]} \exists}} \begin{array}{ll} X_{1}(t_{A}+\frac{jT}{u}+\frac{t}{u})> a_{j;u} \sqrt{u} +\frac{\mu_{1}}{\sqrt u} t \\ \frac{1}{2}\left( X_{2}(s_{A}+\frac{l_{1}S}{u}+\frac{s_{1}}{u})+ X_{2}(s_{A}+\frac{l_{2}S}{u}+\frac{s_{2}}{u})\right)> b_{l_{1},l_{2};u} \sqrt{u} +\frac{\mu_{2}}{2\sqrt u} (s_{1}+s_{2}) \end{array} \right \} \\&=&: P_{j,l_{1},l_{2};u}, \end{array} $$
(36)

with

$$ \begin{array}{@{}rcl@{}} b_{l_1,l_2;u} =1+\mu_2\left( s_A+\frac{l_1S}{u}+\frac{(l_2-l_1)S}{2u}\right). \end{array} $$

For notational simplicity, we shall denote

$$ \begin{array}{@{}rcl@{}} \widetilde{t_A}=t_A+\frac{jT}{u}, \ \ \widetilde{s_A}=s_A+\frac{l_1S}{u},\ \ \overline{s_A}=\widetilde{s_A}+\frac{(l_2-l_1)S}{2 u}, \ \ \widehat{s_A}=\widetilde{s_A}+\frac{(l_2-l_1)S}{4 u}. \end{array} $$

Again by conditioning on the event

$$ \begin{array}{@{}rcl@{}} E_{j,l_1,l_2;u}(x_1,x_2):=\left\{X_1(\widetilde{t_A}) = a_{j;u} \sqrt{u}-\frac{x_1}{\sqrt {u}}, \ \ \ \frac{1}{2}\left( X_2(\widetilde{s_A})+ X_2(s_A+\frac{l_2S}{u} )\right)=b_{l_1,l_2;u} \sqrt{u} -\frac{x_2}{ \sqrt u} \right\}, \end{array} $$

we have

$$ \begin{array}{@{}rcl@{}} P_{j,l_1,l_2;u}= u^{-1}{\int}_{\mathbb{R}^{2}}\phi_{{\Sigma}_{j,l_1,l_2;u} }({\sqrt{u}} \boldsymbol{b}_{j,l_1,l_2;u}-\boldsymbol{x}/ {\sqrt{u}} ) F(j,l_1,l_2;u,\boldsymbol{x}) d \boldsymbol{x}, \end{array} $$

where

$$ \begin{array}{@{}rcl@{}} {\Sigma}_{j,l_1,l_2;u} = \left( \begin{array}{ll} \widetilde{t_A} & \rho\ \overline{s_A}\\ \rho\ \overline{s_A} & \widehat{s_A} \end{array} \right), \ \ \ \boldsymbol{b}_{j,l_1,l_2;u}=(a_{j;u}, b_{l_1,l_2;u})^{\top} \end{array} $$

and

$$ \begin{array}{@{}rcl@{}} F(j,l_1,l_2;u,\boldsymbol{x}) := \mathbb{P} \left\{ \exists_{t \in [0, T]}\ X_1(t) - \mu_1 t> x_1 \right \} \ \mathbb{P} \left\{ \left. \underset{ s_2\in[0,S]} {\underset{ s_1 \in [0, S]} { \exists}} Y_{j,l_1,l_2;u}(s_1,s_2) \!>\!x_2\ \right | E_{j,l_1,l_2;u}(x_1,x_2) \right \} , \end{array} $$

where

$$ \begin{array}{@{}rcl@{}} Y_{j,l_1,l_2;u}(s_1,s_2) = \frac{\sqrt u}{2} \left( X_2(\widetilde{s_A} + \frac{s_1}{u})- X_2(\widetilde{s_A})+ X_2(s_A + \frac{l_2S}{u}+\frac{s_2}{u})- X_2(s_A + \frac{l_2S}{u}) \right)-\frac{\mu_2}{2}(s_1+s_2). \end{array} $$

Similarly as in the proof of Lemma 4.2, we obtain

$$ \begin{array}{@{}rcl@{}} \phi_{{{\Sigma}}_{j,l_1,l_2;u} }({\sqrt{u}} \boldsymbol{b}_{j,l_1,l_2;u} - \boldsymbol{x}/ {\sqrt{u}} ) = \frac{1}{\sqrt{(2\pi)^{2}\left\lvert {\Sigma}_{j,l_1,l_2;u} \right\rvert}}\exp\left( - \frac{1}{2} u \ (\boldsymbol{b}_{j,l_1,l_2;u})^{\top}{\Sigma}_{j,l_1,l_2;u}^{-1} \boldsymbol{b}_{j,l_1,l_2;u} \right)\ f_{j,l_1,l_2;u}(\boldsymbol{x}), \end{array} $$

where

$$ \begin{array}{@{}rcl@{}} f_{j,l_1,l_2;u}(\boldsymbol{x}):= \exp\left( \boldsymbol{x}^{\top} {\Sigma}_{j,l_1,l_2;u}^{-1} \boldsymbol{b}_{j,l_1,l_2;u} - \frac{1}{2u } \boldsymbol{x}^{\top} {\Sigma}_{j,l_1,l_2;u}^{-1} \boldsymbol{x} \right). \end{array} $$

Next, some elementary calculations give that

$$ \begin{array}{@{}rcl@{}} (\boldsymbol{b}_{j,l_1,l_2;u})^{\top}{\Sigma}_{j,l_1,l_2;u}^{-1} \boldsymbol{b}_{j,l_1,l_2;u} = g_A(t_A + \frac{jT}{u}, s_A + \frac{l_1S}{u}+\frac{(l_2-l_1)S}{2 u}) + \frac{ \widetilde{t_A} g_A(\widetilde{t_A}, \overline{s_A} ) -a_{j;u} ^2 }{4 (\widetilde{t_A} \widehat{s_A}-\rho^2{\overline{s_A}}^2)}\ \frac{(l_2-l_1)S}{u}. \end{array} $$

Further, note that

$$ \begin{array}{@{}rcl@{}} g_A(t_A + \frac{jT}{u}, s_A + \frac{l_1S}{u} + \frac{(l_2-l_1)S}{2 u}) = g_A(t_A + \frac{jT}{u}, s_A + \frac{l_1S}{u})+\frac{\partial g_A(t,s)}{\partial s} \mid_{(\widetilde{t_A}, \widetilde{s_A}+\theta_{l_1,l_2;u}\frac{(l_2-l_1)S}{2 u})} \frac{(l_2-l_1)S}{2 u} \end{array} $$

holds for some \(\theta _{l_{1},l_{2};u}\in (0,1)\) and

$$ \begin{array}{@{}rcl@{}} \frac{\partial g_A(t,s)}{\partial t} \mid_{(\widetilde{t_A}, \widetilde{s_A}+\theta_{l_1,l_2;u}\frac{(l_2-l_1)S}{2 u})} \ \ \to\ 0, \ \ \ \ u\to{\infty} \end{array} $$

holds uniformly for j, l1, l2 (hereafter when we write j, l1, l2 we mean \({-N}_{u}^{(1)}\le j \le {N}_{u}^{(1)}, {-N}_{u}^{(2)}\le l_{1},l_{2}\le {N}_{u}^{(2)}\))). Consequently

$$ \begin{array}{@{}rcl@{}} \exp\left( -\frac{1}{2} u \ (\boldsymbol{b}_{j,l_1,l_2;u})^{\top}{\Sigma}_{j,l_1,l_2;u}^{-1} \boldsymbol{b}_{j,l_1,l_2;u} \right) \sim \exp\left( -\frac{1}{2} u \ g_A(t_A+\frac{jT}{u}, s_A+\frac{l_1S}{u}) \right) e^{- Q_0 (l_2-l_1)S} \end{array} $$
(37)

holds uniformly for j, l1, l2 as \(u\to {\infty }\), where (by (b).(i) of Lemma 9 of Dȩbicki et al. (2019) or Lemma A.1.(i) with a1 > 0)

$$ \begin{array}{@{}rcl@{}} Q_0=\frac{t_A g_A(t_A,s_A) -(1+\mu_1 t_A)^2}{8(t_A s_A-\rho^2s_A^2)} >0. \end{array} $$

Next, we consider the uniform, in j, l1, l2, limit of the following:

$$ \begin{array}{@{}rcl@{}} \mathbb{P} \left\{ \underset{ s_2\in[0,S]} {\underset{ s_1 \in [0, S]} { \exists}} Y_{j,l_1,l_2;u}(s_1,s_2) >x_2\ \Bigg{|} E_{j,l_1,l_2;u}(x_1,x_2) \right \} , \ \ \ u\to{\infty} \end{array} $$

For the conditional mean we can derive that

$$ \begin{array}{@{}rcl@{}} \mathbb{E}\left\{ Y_{j,l_1,l_2;u}(s_1,s_2) { \mid} E_{j,l_1,l_2;u}(x_1,x_2) \right\}&=& -\frac{\mu_2}{2}(s_1+s_2) \\ &&+ \left( \frac{\rho (s_1+s_2)}{2\sqrt u}, \frac{s_1}{4\sqrt u}\right)\ {\Sigma}_{j,l_1,l_2;u}^{-1} (\boldsymbol{b}_{j,l_1,l_2;u} - \boldsymbol{x}/ {\sqrt{u}} ), \end{array} $$

which further gives that

$$ \begin{array}{@{}rcl@{}} &&\mathbb{E}{ Y_{j,l_1,l_2;u}(s_1,s_2) { \mid} E_{j,l_1,l_2;u}(x_1,x_2) } = - \frac{\mu_2}{2}(s_1 + s_2) + \frac{2\rho a_{j;u} \widehat{s_A} - \rho a_{j;u} \overline{s_A} -2\rho^2 b_{j,l_1,l_2;u} \overline{s_A} + b_{j,l_1,l_2;u} \widetilde{t_A}}{4(\widetilde{t_A} \widehat{s_A}-\rho^2 {\overline{s_A}}^2)} s_1\\ && \ \ \ \ \ \ \ \ \ \ \ +\frac{ \rho a_{j;u} \widehat{s_A}-\rho^2 b_{j,l_1,l_2;u} \overline{s_A} }{2(\widetilde{t_A} \widehat{s_A}-\rho^2 {\overline{s_A}}^2)} s_2 + \frac{ \rho \overline{s_A} s_1 -2\rho \widehat{s_A} (s_1+s_2) }{4(\widetilde{t_A} \widehat{s_A}-\rho^2 {\overline{s_A}}^2)}\ \frac{x_1}{u} + \frac{ 2\rho ^2 \overline{s_A} (s_1+s_2) - \widetilde{t_A}s_1 }{4(\widetilde{t_A} \widehat{s_A}-\rho^2 {\overline{s_A}}^2)}\ \frac{x_2}{u}\\ && \ \ \ \ \ \ \ \ \ \ \quad to -\frac{1}{2}\left( \mu_2-2\mu_1\rho\right) s_2, \ \ \ \ u\to{\infty}. \end{array} $$

For the conditional variance of the increments we have

$$ \begin{array}{@{}rcl@{}} &&\text{Var}\left\{ Y_{j,l_1,l_2;u}(s_1,s_2) - Y_{j,l_1,l_2;u}(s^{\prime}_1,s^{\prime}_2) { \mid} E_{j,l_1,l_2;u}(x_1,x_2) \right\} = \frac{\left\lvert s_1-{s}_1^{\prime} \right\rvert+\left\lvert s_2-{s}_2^{\prime} \right\rvert}{4}\\ &&\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ +\left( \frac{\rho(s_1-s^{\prime}_1 + s_2-s^{\prime}_2)}{2\sqrt u}, \frac{s_1-s^{\prime}_1}{4 \sqrt u}\right) {\Sigma}_{j,l_1,l_2;u}^{-1} \left( \frac{\rho(s_1 - s^{\prime}_1 + s_2 - s^{\prime}_2)}{2\sqrt u}, \frac{s_1 - s^{\prime}_1}{4 \sqrt u}\right)^{\top}\\ &&\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \to \frac{\left\lvert s_1-s^{\prime}_1 \right\rvert+\left\lvert s_2-s^{\prime}_2 \right\rvert}{4}, \ \ \ u\to{\infty}. \end{array} $$

Therefore, similarly as in Lemma A.3 we can show that as \(u\to {\infty }\)

$$ \begin{array}{@{}rcl@{}} &&\mathbb{P} \left\{\left. \underset{ s_2\in[0,S]} {\underset{ s_1 \in [0, S]} { \exists}} Y_{j,l_1,l_2;u}(s_1,s_2) >x_2\ \right | E_{j,l_1,l_2;u}(x_1,x_2) \right \} \ \\&&\to \mathbb{P} \left\{ \underset{ s_2\in[0,S]} {\underset{ s_1 \in [0, S]} { \exists}} \frac{1}{2}(B_1(s_1)+B_2(s_2)) -\frac{1}{2}\left( \mu_2-2\mu_1\rho\right) s_2 >x_2 \right \} . \end{array} $$

Consequently, the dominated convergence theorem gives

$$ \begin{array}{@{}rcl@{}} &&{\int}_{\mathbb{R}^{2}} f_{j,l_1,l_2;u}(\boldsymbol{x}) F(j,l_1,l_2;u,\boldsymbol{x}) d \boldsymbol{x} \\ && \to {\int}_{\mathbb{R}}e^{2\mu_1 x_1} \mathbb{P} \left\{ \exists_{t\in[0,T]}\ X_1(t)-\mu_1 t> x_1 \right \} dx_1 \\ &&\ \ \ \ \ \times {\int}_\mathbb{R} e^{2(\mu_2-2\mu_1\rho) x_2} \mathbb{P} \left\{ \underset{ s_2\in[0,S]} {\underset{ s_1 \in [0, S]} { \exists}} \frac{1}{2}(B_1(s_1)+B_2(s_2)) -\frac{1}{2}\left( \mu_2-2\mu_1\rho\right) s_2 >x_2 \right \} dx_2\\ &&=: \mathcal{H}(\mu_1; T)\ \mathcal{H}(\mu_1,\mu_2;S) \end{array} $$
(38)

holds uniformly for j, l1, l2, as \(u\to {\infty }\).

Next we derive a useful upper bound for \({\mathscr{H}}(\mu _{1},\mu _{2};S), S>0\):

$$ \begin{array}{@{}rcl@{}} \mathcal{H}(\mu_1,\mu_2;S) \le (\lfloor S\rfloor)^2 e^{Q_0 S} \mathcal{H}(\mu_1,\mu_2;1)<{\infty}. \end{array} $$
(39)

In order to prove (39), by taking j = l1 = 0, l2 = 1 we arrive at

$$ \begin{array}{@{}rcl@{}} P_{0,0,1;u}&=&\mathbb{P} \left\{ \underset{ s_2\in[0,S]} {\underset{ s_1 \in [0, S]} {\underset{ t \in [0, T]} \exists}} \begin{array}{ll} X_1(t_A+\frac{t}{u})> a_{0;u} \sqrt{u} +\frac{\mu_1}{\sqrt u} t \\ \frac{1}{2}(X_2(s_A+\frac{s_1}{u})+ X_2(s_A+\frac{S}{u}+\frac{s_2}{u}))> b_{0,1;u} \sqrt{u} +\frac{\mu_2}{2\sqrt u} (s_1+s_2) \end{array} \right \} \\ &\sim & \frac{u^{-1}}{\sqrt{(2\pi)^{2}\left\lvert {\Sigma}_{0,0,0;u} \right\rvert}} \exp\left( -\frac{1}{2} u \ g_A(t_A, s_A) \right) e^{- Q_0 S} \mathcal{H}(\mu_1; T)\mathcal{H}(\mu_1,\mu_2;S). \end{array} $$
(40)

Define, for any integers 0 ≤ m, n ≤⌊S⌋,

$$ \begin{array}{@{}rcl@{}} q_{m,n;u} := \mathbb{P} \left\{ \underset{ s_2\in[0,1]} {\underset{ s_1 \in [0, 1]} {\underset{ t \in [0, T]} \exists}} \begin{array}{cc} X_1(t_A+\frac{t}{u})> a_{0;u} \sqrt{u} +\frac{\mu_1}{\sqrt u} t \\ \frac{1}{2}(X_2(s_A+\frac{m}{u}+\frac{s_1}{u})+ X_2(s_A+\frac{S+n}{u}+\frac{s_2}{u}))> \widetilde b_{m,n;u} \sqrt{u} +\frac{\mu_2}{2\sqrt u} (s_1+s_2) \end{array} \right \} \end{array} $$

with

$$ \begin{array}{@{}rcl@{}} \widetilde b_{m,n;u}=1+\mu_2\left( s_0+\frac{m}{u}+\frac{ S+n-m}{2u}\right). \end{array} $$

Using the same arguments as in the derivation of Eq. 40 one can show that

$$ \begin{array}{@{}rcl@{}} q_{m,n;u} \sim \frac{u^{-1}}{\sqrt{(2\pi)^{2}\left\lvert {\Sigma}_{0,0,0;u} \right\rvert}} \exp\left( -\frac{1}{2} u \ g_A(t_A, s_A) \right) e^{- Q_0 (S+n-m)} \mathcal{H}(\mu_1; T)\mathcal{H}(\mu_1,\mu_2;1). \end{array} $$
(41)

Comparing Eqs. 40 and 41 we derive

$$ \begin{array}{@{}rcl@{}} \mathcal{H}(\mu_1,\mu_2;S) &\le& \sum\limits_{m=0}^{\lfloor S\rfloor-1} \sum\limits_{n=0}^{\lfloor S\rfloor-1} e^{- Q_0 (n-m)} \mathcal{H}(\mu_1,\mu_2;1)\\ &\le & (\lfloor S\rfloor)^2 e^{Q_0 S} \mathcal{H}(\mu_1,\mu_2;1). \end{array} $$

The finiteness of \({\mathscr{H}}(\mu _{1},\mu _{2};1)\) can be proved by using the Borell-TIS inequality. This justifies bound (39).

Now, we are ready to analyse the triple sum Π12(u). We have

$$ \begin{array}{@{}rcl@{}} {\Pi}_{12}(u)& = &\sum\limits_{j=-N_u^{(1)}}^{N_u^{(1)}} \sum\limits_{l_1=-N_u^{(2)}}^{N_u^{(2)}} \sum\limits_{l_2=l_1+2}^{N_u^{(2)}}\frac{u^{-1}}{\sqrt{(2\pi)^{2}\left\lvert {\Sigma}_{j,l_1,l_2;u} \right\rvert}} \\ && \ \ \ \ \ \ \ \times\exp\left( -\frac{1}{2} u \ (\boldsymbol{b}_{j,l_1,l_2;u})^{\top}{\Sigma}_{j,l_1,l_2;u}^{-1} \boldsymbol{b}_{j,l_1,l_2;u} \right) {\int}_{\mathbb{R}^{2}} f_{j,l_1;u}(\boldsymbol{x}) F(j,l_1,l_2;u,\boldsymbol{x}) d\boldsymbol{x}. \end{array} $$

Therefore, we can derive from Eqs. 3738 and 39 that

$$ \begin{array}{@{}rcl@{}} \underset{u\to{\infty}}{\lim}\frac{ {\Pi}_{12}(u)}{\exp(-ug_A(t_A,s_A)/2)} \le \text{Const} \ \sum\limits_{k=1}^{{\infty}} e^{- kQ_0 S} \frac{{\mathcal{H}(\mu_1;T) \mathcal{H}(\mu_1,\mu_2;1)(\lfloor S\rfloor)^2}}{TS} . \end{array} $$

Consequently, the above implies that

$$ \begin{array}{@{}rcl@{}} \underset{S\to{\infty}}{\limsup} \underset{T\to{\infty}}{\limsup}\underset{u\to{\infty}}{\lim}\frac{{\Pi}_{12}(u)}{\exp(-ug_A(t_A,s_A)/2)}=0. \end{array} $$

Thus, the claim for Π1(u) is established. Using similar arguments, one can further show that the claim for Π2(u) holds. \(\Box \)

1.4 A.4 Proof of Lemma 4.4

The claim for r1(u) follows from the same arguments as that for r0(u) of Lemma 4.1. Next, as in the proof of Lemma 4.1, using the Piterbarg’s inequality we can show that

$$ \begin{array}{@{}rcl@{}} r_2(u)\le C_2 u^{3/2} e^{-\frac{ u }{2} \widetilde g_u}, \end{array} $$

where C2 > 0 is some constant which does not depend on u, and thus the claim for r2(u) follows since

$$ \begin{array}{@{}rcl@{}} \widetilde g_u&=& \underset{(t,s)\in D_1\cup D_2 }{\inf} g(t,s)=\underset{s \in[s_0-\theta_0, s_0-\ln(u)/\sqrt u]\cup[ s_0+\ln(u)/\sqrt u, s_0+\theta_0]}{\inf} g_L(s)\\ &\ge& g_L(s^*) +\frac{b_0}{2}(1-\varepsilon) \frac{(\ln(u))^2}{u}, \end{array} $$

where the last inequality follows by (ii.3) of Lemma A.1. Finally, the claim for r3(u) can be proved similarly, by using Piterbarg’s inequality and (ii.1)-(ii.2) of Lemma A.1. \(\Box \)

1.5 A.5: Proof of Lemma 4.5

We first analyse the summand pj; u. Let

$$ \begin{array}{@{}rcl@{}} \boldsymbol b_{j;u}=(a_{j;u}, b_{j,u})^{\top}, \ \ \ a_{j;u}=1+\mu_1 (t^*+\frac{jT}{u}), \ \ \ b_{j;u}=1+\mu_2 (s^*+\frac{jT}{u}). \end{array} $$

Then

$$ \begin{array}{@{}rcl@{}} p_{j;u} = \mathbb{P} \left\{ \underset{ (t,s) \in \triangle_{T,S} }\exists \begin{array}{lll} X_{1}(t^{*}+\frac{jT}{u}+\frac{t}{u})> a_{j;u} \sqrt{u} +\frac{\mu_{1}}{\sqrt {u}} t \\ X_{2}(s^{*}+\frac{jT}{u}+\frac{s}{u})> b_{j;u} \sqrt{u} +\frac{\mu_{2}}{\sqrt {u}} s \end{array} \right \} . \end{array} $$

Define \(\boldsymbol {Z}_{j;u}:=(X_{1}(t^{*}+\frac {jT}{u}), X_{2}(s^{*}+\frac {jT}{u}))^{\top }\), whose density function is given by

$$ \begin{array}{@{}rcl@{}} \phi_{{\Sigma}_{j;u}}(\boldsymbol {w})=\frac{1}{\sqrt{(2\pi)^{2}\left\lvert {\Sigma}_{j;u} \right\rvert}}\exp\left( -\frac{1}{2}\boldsymbol {w}^{\top}({\Sigma}_{j;u})^{-1} \boldsymbol{w} \right),\ \ \ \boldsymbol{w}=(w_1,w_2)^{\top}, \end{array} $$

with the covariance matrix given by

$$ \ {\Sigma}_{j;u}= \left( \begin{array}{ll} t^{*}+\frac{jT}{u} & \rho\ (t^{*}+\frac{jT}{u})\\ \rho\ (t^{*}+\frac{jT}{u}) &\quad s^{*}+\frac{jT}{u} \end{array} \right). $$

By conditioning on the value of Zj; u and using change of variables \(\boldsymbol {w} ={\sqrt {u}} \boldsymbol {b}_{j;u}-\boldsymbol {x}/ {\sqrt {u}}\), we further obtain

$$ \begin{array}{@{}rcl@{}} p_{j;u}= u^{-1}{\int}_{\mathbb{R}^{2}}\phi_{ {\Sigma}_{j;u} }({\sqrt{u}} \boldsymbol{b}_{j;u}-\boldsymbol{x}/ {\sqrt{u}} ) \mathbb{P} \left\{ \underset{ (t,s) \in \triangle_{T,S} }\exists \begin{array}{lll} X_1(t)- \mu_1 t>x_1\\ X_2(s)- \mu_2 s>x_2 \end{array} \right \} d \boldsymbol{x}. \end{array} $$

Consequently, similar arguments as in the proof of Lemma 4.2 yield

$$ \begin{array}{@{}rcl@{}} p_1(u) &\sim& p_2(u)\ \sim\ \sum\limits_{j=-N_u^{(1)}}^{N_u^{(1)}} p_{j;u} \sim \frac{\mathcal{H}(T,S)u^{-1}}{\sqrt{(2\pi t^*)^2(1-\rho^2)}} \sum\limits_{j=-N_u^{(1)}}^{N_u^{(1)}} e^{-\frac{u}{2} g_L(t^*+\frac{jT}{u})}\\ &\sim& \frac{\mathcal{H}(T,S)u^{-1/2}}{T\sqrt{(2\pi t^*)^2(1-\rho^2)}} e^{-\frac{u}{2} g_L(t^*)}{\int}_{\mathbb{R}}e^{-\frac{b_0}{4}x^2} dx. \end{array} $$

This completes the proof. \(\Box \)

1.6 A.6: Proof of Lemma 4.6

First note that

$$ \begin{array}{@{}rcl@{}} &&\mathbb{P} \left\{ \exists_{(t,s) \in (t^*,s^*)+u^{-1}\triangle_{T_1+T_2,S}} \ X_1(t)> \sqrt{u}(1+ \mu_1 t ), X_2(s)> \sqrt{u}(1+ \mu_2 s ) \right \} \\ &&\ \ \le \mathbb{P} \left\{ \exists_{(t,s) \in (t^*,s^*)+u^{-1}\triangle_{T_1,S}} \ X_1(t)> \sqrt{u}(1+ \mu_1 t ), X_2(s)> \sqrt{u}(1+ \mu_2 s ) \right \} \\ &&\ \ \ \ \!+ \mathbb{P} \left\{ \exists_{(t,s) \in (t^*+\frac{T_1}{u},s^*+\frac{T_1}{u})+u^{-1}\triangle_{T_2,S}} \ X_1(t)\!>\! \sqrt{u}(1 + \mu_1 t ), X_2(s)\!>\! \sqrt{u}(1 + \mu_2 s ) \right \} . \end{array} $$

Using the same arguments as the proof of Lemma 4 5, we conclude the sub-additivity of \({\mathscr{H}}(T,S), T>0.\) Thus

$$ \begin{array}{@{}rcl@{}} \underset{T\to{\infty}}{\lim}\frac{1}{T}\mathcal{H}(T,S) = \underset{T> 0}{\inf}\frac{1}{T}\mathcal{H}(T,S) <{\infty} \end{array} $$

follows directly from Fekete’s lemma. Moreover, since by definition

$$ \begin{array}{@{}rcl@{}} \mathcal{H}(T,S)\ge {\int}_{\mathbb{R}^2}e^{\boldsymbol{x}^{\top} {\Sigma}_{*}^{-1} \boldsymbol{b}_{*} }\mathbb{P} \left\{ {\underset{ t \in [0,T]}\exists } \begin{array}{lll} X_1(t)- \mu_1 t > x_1 \\ X_2(t)- \mu_2 t >x_2 \end{array} \right \} dx_1 dx_2, \end{array} $$

the positive lower bound follows from Lemma 4.7 in Dȩbicki et al. (2018). This completes the proof. \(\Box \)

1.7 A.7 Proof of Lemma 4.7

We begin with the analysis of Π1(u). We first look at pj, l; u. Denote

$$ \begin{array}{@{}rcl@{}} &&\boldsymbol {b}_{u}=\boldsymbol {b}_{j,l,m,n;u}:=(a_{j,m;u}, b_{j,l,m,n;u})^{\top}, \\ && a_{j,m;u}=1+\mu_1 (t^*+\frac{jT+m}{u}), \ \ b_{j,l,m,n;u}=1+\mu_2 (t^*+\frac{jT+m}{u}+\frac{lS+n}{u}). \end{array} $$

It is derived that

$$ \begin{array}{@{}rcl@{}} p_{j,l;u} &\le &\sum\limits_{m=0}^{\lfloor T \rfloor-1}\sum\limits_{n=0}^{\lfloor S \rfloor-1} \mathbb{P} \left\{ \underset{s-t\in[\frac{lS}{u}+\frac{n}{u}, \frac{lS}{u}+\frac{n+1}{u} ]}{\underset{ t \in [t^*+\frac{jT}{u}+\frac{m}{u}, t^*+\frac{jT}{u}+\frac{m+1}{u}] }\exists } \ X_1(t)> \sqrt{u}(1+ \mu_1 t ), X_2(s)> \sqrt{u}(1+ \mu_2 s ) \right \} \\ &=& \sum\limits_{m=0}^{\lfloor T \rfloor-1}\sum\limits_{n=0}^{\lfloor S \rfloor-1} \mathbb{P} \left\{ \underset{s\in[0,1]}{\underset{ t \in [0, 1] }\exists } \begin{array}{lll} X_1(t^*+\frac{jT+m}{u}+\frac{t}{u}) > a_{j,m;u} \sqrt{u} +\frac{\mu_1}{\sqrt u} t \\ X_2(t^*+\frac{jT+m}{u}+\frac{lS+n}{u}+\frac{t+s}{u}) > b_{j,l,m,n;u} \sqrt{u} +\frac{\mu_2}{\sqrt u} (t+s) \end{array} \right \} \\ &=:& \sum\limits_{m=0}^{\lfloor T \rfloor-1}\sum\limits_{n=0}^{\lfloor S \rfloor-1} p_{j,l,m,n;u} . \end{array} $$

Next, we look at pj, l, m, n; u. We define

$$ \begin{array}{@{}rcl@{}} \boldsymbol{Z}_{u} &:=& \left( X_1(t^*+\frac{jT+m}{u}), \ X_2(s^*+\frac{jT+m}{u}+\frac{lS+n}{u})\right)^{\top},\\ Y_{1;u}(t) &:= &\left( X_1(t^*+\frac{jT+m}{u}+\frac{t}{u})- X_1(t^*+\frac{jT+m}{u})\right) {\sqrt u} -\mu_1 t, \\ Y_{2;u}(t,s) &:=&\left( X_2(t^*+\frac{jT+m}{u}+\frac{lS+n}{u} + \frac{t+s}{u})- X_2(t^* + \frac{jT+m}{u}+\frac{lS+n}{u})\right) {\sqrt u} -\mu_2 (t+s). \end{array} $$

Consider the conditional process

$$ \begin{array}{@{}rcl@{}} \boldsymbol {W}_{u}(t,s):=(Y_{1;u}(t), Y_{2;u}(t,s))^{\top} \mid \boldsymbol{Z}_{u}= {\sqrt{u}} \boldsymbol{b}_{u}-\frac{\boldsymbol{x}}{ {\sqrt{u}}}. \end{array} $$

We have that \( (Y_{1;u}(t), Y_{2;u}(t,s), \boldsymbol {Z}_{u}^{\top })\) is a normally distributed random vector, with mean

$$ \begin{array}{@{}rcl@{}} \boldsymbol {\widehat\mu}(t,s) := (-\mu_1 t, -\mu_2(t+s), 0, 0)^{\top} \end{array} $$

and covariance matrix given by (suppose S > 1)

$$ \begin{array}{@{}rcl@{}} \widehat{\Sigma}_{u}(t,s):= \left( \begin{array}{cccr} t&0& 0 & \rho\frac{t}{\sqrt u}\\ 0 & t+s & 0 & 0\\ 0 & 0& t^*+\frac{jT+m}{u} & \rho\ (t^*+\frac{jT+m}{u})\\ \rho\frac{t}{\sqrt u} & 0& \rho\ (t^*+\frac{jT+m}{u}) & t^*+\frac{jT+m}{u} +\frac{lS+n}{u} \end{array} \right). \end{array} $$

Thus, for the mean

$$ \begin{array}{@{}rcl@{}} \mathbb{E}\left\{ \boldsymbol {W}_{u}(t,s)\right\}& = & (-\mu_1 t, - \mu_2(t+s)) + \left( \begin{array}{cc} 0 & \rho\frac{t}{\sqrt u}\\ 0&0 \end{array} \right) \left( \begin{array}{cc} t^*+\frac{jT+m}{u} & \rho\ (t^*+\frac{jT+m}{u})\\ \rho\ (t^* + \frac{jT+m}{u}) & t^*+\frac{jT+m}{u} + \frac{lS+n}{u} \end{array} \right)^{-1} \left( {\sqrt{u}} \boldsymbol{b}_{u} - \frac{\boldsymbol{x}}{ {\sqrt{u}}}\right)\\ & = & (-\mu_1 t, -\mu_2(t+s))+ \left( \frac{\rho (b_{j,l,m,n;u} - \rho a_{j,m;u} ) t - \rho t \frac{x_2-\rho x_1}{u}}{(t^*+\frac{jT+m}{u}+\frac{lS+n}{u}) -\rho^2(t^*+\frac{jT+m}{u})} , 0\right) \\ &\quad\to & \left( - \nu(\rho) \ t, -\mu_2 (t+s) \right)^{\top}, \ \ \nu(\rho):=\frac{\rho^2-\rho+(\mu_1-\mu_2\rho)t^*}{t^*(1-\rho^2)}, \end{array} $$

as \(u\to {\infty }\), where the convergence is uniform for \(-N_{u}^{(1)}\le j\le N_{u}^{(1)}, 1\le l\le N_{u}^{(2)}\). Similarly, we can derive that, for any t1, t2 ∈ [0, T], s1, s2 ∈ [−S, S],

$$ \begin{array}{@{}rcl@{}} \text{Cov}(\boldsymbol {W}_{u}(t_1,s_1) -\boldsymbol {W}_{u}(t_2,s_2)) &\to &\text{Cov}((B_1(t_1)-B_1(t_2), B_2(t_1+s_1)-B_2(t_2+s_2))^{\top}) \end{array} $$

as \(u\to {\infty }\), uniformly for \(-N_{u}^{(1)}\le j\le N_{u}^{(1)}, 1\le l\le N_{u}^{(2)}\). Consequently, we have, as \(u\to {\infty },\)

$$ \begin{array}{@{}rcl@{}} &&\mathbb{P} \left\{ \left. \underset{s\in[0,1]}{\underset{ t \in [0, 1] }\exists } \begin{array}{lll} Y_{1;u}(t)>x_1 \\ Y_{2;u}(t,s) > x_2 \end{array} \right | \boldsymbol{Z}_{u}= {\sqrt{u}} \boldsymbol{b}_{u}-\frac{\boldsymbol{x}}{ {\sqrt{u}}} \right\} \\ && \to \ \mathbb{P} \left\{ \underset{s\in[0,1]}{\underset{ t \in [0, 1] }\exists } \begin{array}{lll} B_1(t)- \nu(\rho) t > x_1 \\ B_2(t+s) - \mu_2 (t+s) >x_2 \end{array} \right \} . \end{array} $$

Similar arguments as those in the proof of Lemma 4.2 gives that

$$ \begin{array}{@{}rcl@{}} p_{j,l,m,n;u} \sim \frac{\widehat{\mathcal{H}}(1,1)u^{-1}}{\sqrt{(2\pi)^{2}\lvert {\Sigma}_{*} \rvert}} e^{-\frac{u}{2} g_B(t^{*}+\frac{jT+m}{u}, s^{*}+\frac{jT+m}{u}+\frac{lS+n}{u})}, \end{array} $$

where (recall notation in Eq. 7)

$$ \begin{array}{@{}rcl@{}} \widehat{\mathcal{H}}(1,1)={\int}_{\mathbb{R}^2} e^{\boldsymbol{x}^{\top} {\Sigma}_{*}^{-1} \boldsymbol{b}_{*} } \mathbb{P} \left\{ \underset{s\in[0,1]}{\underset{ t \in [0, 1] }\exists } \begin{array}{lll} B_1(t)- \nu(\rho) t > x_1 \\ B_2(t+s) - \mu (t+s) >x_2 \end{array} \right \} dx_1 dx_2\in(0,{\infty}). \end{array} $$

It follows further from (ii.2) of Lemma A.1 that there exists some ε > 0 such that, for all t < s small,

$$ \begin{array}{@{}rcl@{}} g_B(t^*+ t,t^*+ s) \ge g_L(t^*)+b_2(1-\varepsilon)(s-t )+\frac{c_2 (1-\varepsilon) }{2} t^2, \end{array} $$

thus, for u sufficiently large

$$ \begin{array}{@{}rcl@{}} g_B(t^* + \frac{jT+m}{u}, t^* + \frac{jT+m}{u} + \frac{lS+n}{u}) \ge g_L(t^*) + b_2(1 - \varepsilon)\frac{lS}{u} + \frac{c_2(1-\varepsilon)}{2} \left( \frac{\widehat j T}{u}\right)^2 \end{array} $$

holds for all \(-N_{u}^{(1)}\le j\le N_{u}^{(1)}, 1\le l\le N_{u}^{(2)}, 0\le m\le \lfloor T\rfloor -1, 0\le n\le \lfloor S\rfloor -1\), where \(\widehat j=j I_{\{j\ge 0\}} +(j+1) I_{\{j<0\}}\). This implies that, for u large

$$ \begin{array}{@{}rcl@{}} e^{-\frac{u}{2} g_A(t^*+\frac{jT+m}{u}, s^*+\frac{jT+m}{u}+\frac{lS+n}{u})}\le e^{-\frac{u}{2} g_L(t^*)} \frac{\sqrt u}{T} \left( e^{- \frac{c_2(1-\varepsilon)}{4} \left( \frac{\widehat jT}{\sqrt u}\right)^2} \frac{T}{\sqrt u}\right) e^{-\frac{b_2(1-\varepsilon)}{2} lS}. \end{array} $$

Based on the above discussions we obtain

$$ \begin{array}{@{}rcl@{}} \underset{u\to{\infty}}{\lim}\frac{{\Pi}_1(u)}{u^{-1/2} \exp(- g_L(t^*){u}/{2}) } \le \frac{\widehat{\mathcal{H}}(1,1)}{\sqrt{(2\pi)^2\left\lvert {\Sigma}_{*} \right\rvert}} \ \frac{\lfloor T \rfloor \lfloor S \rfloor}{T} \underset{ l\ge 1 }{\sum} e^{-\frac{b_2(1-\varepsilon)}{2} lS} {\int}_{\mathbb{R}} e^{-\frac{c_2(1-\varepsilon)}{4} x^2}dx. \end{array} $$

Similar bounds can be found for \(\overline {\Pi }_{1}(u)\), and thus the first claim follows.

Next we consider Π21(u). For any j2 > j1 + 1 we have

$$ \begin{array}{@{}rcl@{}} q_{j_1,j_2;u}&=&\mathbb{P} \left\{ \begin{array}{cc} \exists_{(t,s) \in (t^*+\frac{j_1 T}{u},s^*+\frac{j_1 T}{u})+u^{-1}\triangle_{T,S}} \ X_1(t)> \sqrt{u}(1+ \mu_1 t ), X_2(s)\!>\! \sqrt{u}(1+ \mu_2 s ) \\ \exists_{(t,s) \in (t^*+\frac{j_2 T}{u},s^*+\frac{j_2 T}{u})+u^{-1}\triangle_{T,S}} \ X_1(t)> \sqrt{u}(1+ \mu_1 t ), X_2(s)\!>\! \sqrt{u}(1+ \mu_2 s ) \end{array} \right \} \\ &\le & u^{-1}{\int}_{\mathbb{R}^{2}}\phi_{{\Sigma}_{j_1,j_2;u} }({\sqrt{u}} \boldsymbol{b}_{j_1,j_2;u}-\boldsymbol{x}/ {\sqrt{u}} ) \overline F(j_1,j_2;u,\boldsymbol{x}) d \boldsymbol{x}=:Q_{j_1,j_2;u}, \end{array} $$

where, with \(a_{j;u}=1+\mu _{1}(t^{*}+\frac {j_{1}T}{u}), b_{j;u}=1+\mu _{2}(s^{*}+\frac {j_{1}T}{u})\),

$$ \begin{array}{@{}rcl@{}} {\Sigma}_{j_1,j_2;u} = \left( t^*+\frac{j_1T}{u}+\frac{(j_2-j_1)S}{4u} \right) \left( \begin{array}{cc} 1 & \rho \\ \rho & 1 \end{array} \right), \ \ \boldsymbol{b}_{j_1,j_2;u} =\left( \frac{a_{j_1;u}+a_{j_2;u}}{2}, \frac{b_{j_1;u}+b_{j_2;u}}{2}\right), \end{array} $$
$$ \begin{array}{@{}rcl@{}} \overline F(j_1,j_2;u,\boldsymbol{x}):= \mathbb{P} \left\{\left. \underset{(t^{\prime},s^{\prime})\in\triangle_{T,S}}{\underset{ (t,s)\in\triangle_{T,S} }\exists } \begin{array}{ccc} Y_{1;u}(t,t^{\prime}) > x_1 \\ Y_{2;u}(s,s^{\prime}) >x_2 \end{array} \right| \begin{array}{lll} Y_{3;u} = \frac{a_{j_1;u}+a_{j_2;u}}{2}\sqrt u -\frac{ x_1}{\sqrt u} \\ Y_{4;u} =\frac{a_{j_1;u}+a_{j_2;u}}{2}\sqrt u -\frac{ x_2}{\sqrt u} \end{array} \right \} , \end{array} $$

with

$$ \begin{array}{@{}rcl@{}} Y_{1;u}(t,t^{\prime})&=& \frac{\sqrt u}{2} \left( X_1(t^* + \frac{j_1T}{u} + \frac{t}{u}) - X_1(t^* + \frac{j_1T}{u}) + X_1(t^* + \frac{j_2T}{u}+\frac{t^{\prime}}{u}) - X_1(t^*+\frac{j_2T}{u})\right) - \frac{\mu_1}{2}(t + t^{\prime}), \\ Y_{2;u}(s,s^{\prime})& = & \frac{\sqrt u}{2} \left( X_2(s^*+\frac{j_1T}{u} + \frac{s}{u}) - X_2(s^* + \frac{j_1T}{u}) \!+ \! X_2(s^* + \frac{j_2T}{u} + \frac{s^{\prime}}{u})- X_2(s^* + \frac{j_2T}{u})\right) - \frac{\mu_2}{2}(s + s^{\prime}), \\ Y_{3;u} &=&\frac{1}{2} \left( X_1(t^*+\frac{j_1T}{u})+ X_1(t^*+\frac{j_2T}{u})\right), \ \ \ \ Y_{4;u} = \frac{1}{2} \left( X_2(s^*+\frac{j_1T}{u})+X_2(s^*+\frac{j_2T}{u})\right). \end{array} $$

Next we have that \((Y_{1;u}(t,t^{\prime }), Y_{2;u}(s,s^{\prime }), Y_{3;u},Y_{4;u})\) is a normally distributed random vector, with mean

$$ \begin{array}{@{}rcl@{}} \boldsymbol{\widehat\mu}(t,t^{\prime},s,s^{\prime}) = \left( -\frac{\mu_1}{2}(t+t^{\prime}), -\frac{\mu_2}{2}(s+s^{\prime}), 0, 0\right)^{\top} \end{array} $$

and covariance matrix given by (suppose T > S)

$$ \begin{array}{@{}rcl@{}} \widehat{\Sigma}_{u}(t,s)= \left( \begin{array}{cccr} \frac{t+t^{\prime}}{4}&\frac{\rho(t\wedge s+t^{\prime} \wedge s^{\prime})}{4}& \frac{t}{4 \sqrt u} & \frac{\rho t}{4 \sqrt u}\\ \frac{\rho(t\wedge s+t^{\prime} \wedge s^{\prime})}{4} & \frac{s+s^{\prime}}{4}& \frac{\rho s}{4 \sqrt u} & \frac{s}{4 \sqrt u}\\ \frac{t}{4 \sqrt u} & \frac{\rho s}{4 \sqrt u}& t^*+\frac{j_1T}{u}+\frac{(j_2-j_1)T}{4u} & \rho\ \left( t^*+\frac{j_1T}{u}+\frac{(j_2-j_1)T}{4u}\right)\\ \frac{\rho t}{4 \sqrt u} & \frac{s}{4 \sqrt u}& \rho\ \left( t^*+\frac{j_1T}{u}+\frac{(j_2-j_1)T}{4u}\right) & s^*+\frac{j_1T}{u}+\frac{(j_2-j_1)T}{4u} \end{array} \right). \end{array} $$

Similarly as before, one can get

$$ \begin{array}{@{}rcl@{}} Q_{j_1,j_2;u} \sim \frac{\widetilde{\mathcal{H}}(T,S)u^{-1}}{\sqrt{(2\pi)^2\left\lvert {\Sigma}_{*} \right\rvert}} e^{-\frac{u}{2} \left( 1+\frac{\frac{(j_2-j_1)T}{4u}}{t^*+\frac{j_1T}{u}+\frac{(j_2-j_1)T}{4u}}\right) g_L(t^*+\frac{j_1T}{u}+\frac{(j_2-j_1)T}{2u})}, \end{array} $$

as \(u\to {\infty },\) where

$$ \begin{array}{@{}rcl@{}} \widetilde{\mathcal{H}}(T,S): = {\int}_{\mathbb{R}^2}e^{\boldsymbol{x}^{\top} {\Sigma}_{*}^{-1} \boldsymbol{b}_{*} }\mathbb{P} \left\{ \underset{(t^{\prime},s^{\prime})\in\triangle_{T,S}}{\underset{ (t,s)\in\triangle_{T,S} }\exists } \begin{array}{lll} \frac{1}{2}(X_1(t)+\widetilde X_1(t^{\prime}))-\frac{\mu_1t^*-1}{4t^*} t- \frac{\mu_1}{2} t^{\prime} > x_1 \\ \frac{1}{2}(X_2(s)+\widetilde X_2(s^{\prime})) -\frac{\mu_2s^*-1}{4s^*} s- \frac{\mu_2}{2}s^{\prime} >x_2 \end{array} \right \} dx_1 dx_2\in(0,{\infty}), \end{array} $$

with \((\widetilde X_{1}, \widetilde X_{2})\) an independent copy of (X1, X2). Particularly, letting j1 = 0, j2 = 2 we can show, similarly as in Eq. 39, that

$$ \begin{array}{@{}rcl@{}} \widetilde{\mathcal{H}}(T,S) \le \widetilde{\mathcal{H}}(1,S) (\lfloor T \rfloor )^2 e^{\frac{g_L(t^*)}{8t^*} T}. \end{array} $$

Therefore, as \(u\to {\infty },\)

$$ \begin{array}{@{}rcl@{}} {\Pi}_{21} (u) &\lesssim & \frac{\widetilde{\mathcal{H}}(T,S)u^{-1/2}}{T \sqrt{(2\pi)^2\left\lvert {\Sigma}_{*} \right\rvert}} e^{-\frac{u}{2}g_L(t^*)} \left( \sum\limits_{j_1=-N_u^{(1)}}^{N_u^{(1)}} e^{-\frac{b_0}{4}\left( \frac{j_1 T}{\sqrt u}\right)^2} \frac{T}{\sqrt u}\right) \left( \sum\limits_{j_2=j_1+2}^{N_u^{(1)}} e^{-\frac{g_L(t^*)}{8t^*} (j_2-j_1)T}\right)\\ &\lesssim& \frac{\widetilde{\mathcal{H}}(1,S) \lfloor T\rfloor^2 u^{-1/2}}{T \sqrt{(2\pi)^2\left\lvert {\Sigma}_{*} \right\rvert}} e^{-\frac{u}{2}g_L(t^*)} \ {\int}_{\mathbb{R}}e^{-\frac{b_0}{4} x^2} dx \ \sum\limits_{k=1}^{{\infty}} e^{-\frac{g_L(t^*)}{8t^*} kT}. \end{array} $$

Finally, we consider Π22(u). Note that

$$ \begin{array}{@{}rcl@{}} {\Pi}_{22}(u) = \sum\limits_{j=-{N}_u^{(1)}}^{{N}_u^{(1)}} p_{j;u}+p_{j+1;u}-\widetilde{p}_{j;u}, \end{array} $$

where

$$ \begin{array}{@{}rcl@{}} \widetilde{p}_{j;u}=\mathbb{P} \left\{ \exists_{(t,s) \in (t^*+\frac{j T}{u},t^*+\frac{j T}{u})+u^{-1}\triangle_{2T,S}} \ X_1(t)> \sqrt{u}(1+ \mu_1 t ), X_2(s)> \sqrt{u}(1+ \mu_2 s ) \right \} . \end{array} $$

Consequently, the claim for Π22(u) follows directly by using Lemma 4.5. \(\Box \)

1.8 A.8 Proof of Lemma 5.1

Similarly as in Eq. 27 we obtain

$$ \begin{array}{@{}rcl@{}} P_{\theta_0,0}(u)&\le& \mathbb{P} \left\{ \exists_{(t_1,s_1)\in U_{11} } \ Z(t_1,s_1) >\sqrt u g(t_1,s_1),\ \ \exists_{(t_2,s_2)\in U_{12}} \ Z(t_2,s_2)>\sqrt u g(t_2,s_2) \right \} \\ &\le& \mathbb{P} \left\{ \exists_{(t_1,s_1)\in U_{11} } \ \overline{Z}(t_1,s_1)>\sqrt u\sqrt{g(t_0,s_0)},\ \ \exists_{(t_2,s_2)\in U_{12} } \ \overline{Z}(t_2,s_2)>\sqrt u \sqrt{g(s_0,t_0)} \right \} \\ &\le& \mathbb{P} \left\{ \exists_{(t_1,s_1)\in U_{11} , (t_2,s_2)\in U_{12}} \ \overline{Z}(t_1,s_1)+ \overline{Z}(t_2,s_2)> 2\sqrt u \sqrt{g(t_0,s_0)} \right \} , \end{array} $$

where, we used the fact that \(g(t_{0},s_{0})=g(s_{0},t_{0}) \le \inf _{(t,s)\in (U_{11}\cup U_{12})}g(t,s)\), and

$$ \begin{array}{@{}rcl@{}} \overline{Z}(t_i,s_i):= \frac{Z(t_i,s_i)}{\sqrt{\text{Var} (Z(t_i,s_i))}}= \frac{Z(t_i,s_i)}{\sqrt{g(t_i,s_i) }}, \ \ \ i=1,2. \end{array} $$

Further note that

$$ \begin{array}{@{}rcl@{}} \mathbb{E}\left\{ Z(t_0,s_0) Z (s_0,t_0)\right\}&=&\mathbb{E}\left\{ (2\mu X_1(t_0) +2(1-2\rho)\mu X_2(s_0)) (2(1-2\rho)\mu X_1(s_0) +2\mu X_2(t_0))\right\}\\ &=& 8(1+2\rho)(1-\rho)\mu. \end{array} $$

We obtain

$$ \begin{array}{@{}rcl@{}} \mathbb{E}\left\{ (\overline{Z}(t_0,s_0)+ \overline{Z}(s_0,t_0))^2\right\}&=&2+ 2\mathbb{E}\left\{ \overline{Z}(t_0,s_0) \overline{Z}(s_0,t_0)\right\}\\ &=&2+ 2\frac{\mathbb{E}\left\{ Z(t_0,s_0) Z (s_0,t_0)\right\}}{g(t_0,s_0)}\\ &=&2+ 2 (1+2\rho)<4. \end{array} $$

Thus, for sufficiently small 𝜃0 > 0,

$$ \begin{array}{@{}rcl@{}} \sigma ^2:=\underset{(t_2,s_2)\in U_{12}}{\underset{(t_1,s_1)\in U_{11}}{\sup}}\mathbb{E}\left\{ (\overline{Z}(t_1,s_1)+ \overline{Z}(t_2,s_2))^2\right\}<4, \end{array} $$

where we use continuity of the functions involved. Again, using the Borell-TIS inequality we obtain

$$ \begin{array}{@{}rcl@{}} \mathbb{P} \left\{ \exists_{(t_1,s_1)\in U_{11}, (t_2,s_2)\in U_{12} } \ \overline{Z}(t_1,s_1)+ \overline{Z}(t_2,s_2)> 2\sqrt u \sqrt{g(t_0,s_0)} \right \} \le e^{-\frac{(2\sqrt u \sqrt{g(t_0,s_0)}-C_0)^2}{2\sigma ^2}} \end{array} $$

holds for all large u such that

$$ \begin{array}{@{}rcl@{}} 2\sqrt u \sqrt{g(t_0,s_0)}>C_0:=\mathbb{E}\left\{ \underset{(t_2,s_2)\in U_{12}}{\underset{(t_1,s_1)\in U_{11}}{\sup}} (\overline{Z}(t_1,s_1)+ \overline{Z}(t_2,s_2)) \right\}. \end{array} $$

Thus, the claim follows.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Dȩbicki, K., Ji, L. & Rolski, T. Exact asymptotics of component-wise extrema of two-dimensional Brownian motion. Extremes 23, 569–602 (2020). https://doi.org/10.1007/s10687-020-00387-y

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10687-020-00387-y

Keywords

AMS Classification

Navigation