This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

A note on a deterministic property to obtain the long run behavior of the range of a stochastic process

Maher Boudabra111Mathematics department, King Fahd University for Petroleum and Minerals, KSA    Binghao Wu222School of Mathematics, Monash University, Australia
Abstract

A Brownian motion with drift is simply a process VtηV^{\eta}_{t} of the form Vtη=Bt+ηtV^{\eta}_{t}=B_{t}+\eta t where BtB_{t} is a standard Brownian motion and η>0\eta>0 333The case η<0\eta<0 is deducible by remarking Vη(t)=Vη(t)V^{-\eta}(t)=-V^{\eta}(t). In [13], the authors considered the drifted Brownian motion and studied the statistics of some related sequences defined by certain stopping times. In particular, they provided the law of the range Rt(Vη)R_{t}(V^{\eta}) of VηV^{\eta} as well as its first range process θVη(a)\theta_{V^{\eta}}(a). In particular, they investigated the asymptotic comportment of Rt(Vη)R_{t}(V^{\eta}) and θVη(a)\theta_{V^{\eta}}(a). They proved that if VtηV_{t}^{\eta} is a Brownian motion with a positive drift η\eta then its range Rt(Vη)=sup0stVtηinf0stVtηR_{t}(V^{\eta})=\sup_{0\leq s\leq t}V_{t}^{\eta}-\inf_{0\leq s\leq t}V_{t}^{\eta} is asymptotically equivalent to ηt\eta t. In other words

Rt(Vη)tta.eη.\frac{R_{t}(V^{\eta})}{t}\overset{a.e}{\underset{t\rightarrow\infty}{\longrightarrow}}\eta. (1)

In this paper, we show that (1) follows from a striking deterministic property. More precisely, we show that the long run behavior of the range of a deterministic continuous function is obtainable straightaway from that of the function itself. Our result can be deemed as the continuous version of a similar one appeared in [1].

1 Introduction

The range of a real valued process X=(Xt)tTX=(X_{t})_{t\in T} is defined to be the "measure" (in a certain sense) of the set of values generated by this process up to time tt. We shall denote the range of the process XX by Rt(X)R_{t}(X). When XX is discrete, i.e X=(Xn)nX=(X_{n})_{n\in\mathbb{N}}, Rn(X)R_{n}(X) simplifies to the number of explored sites up to time nn by the process XX, i.e Rn(X)=|{X0,,Xn}|R_{n}(X)=|\{X_{0},...,X_{n}\}|. The range of discrete processes (random walks, Markov chains etc…) has been investigated in several seminal works [4, 5, 2, 15, 16]. In particular, we cite Kesten-Spitzer-Whitman theorem [12] which says that the average of the range of a nearest neighbor simple random walk (referred to as its speed in [1]) converges to the probability of never returning to the origin (the starting point more generally). That is, let (ξn)n(\xi_{n})_{n} be a sequence of i.i.d Redmacher random variables of parameter pp and set Xn=ξ1++ξnX_{n}=\xi_{1}+...+\xi_{n}. Then

Rn(X)nna.e(Xnever returns to the origin).\frac{R_{n}(X)}{n}\overset{a.e}{\underset{n\rightarrow\infty}{\longrightarrow}}\mathbb{P}(X\,\,\text{never returns to the origin}). (2)

Note that (Xnever returns to the origin)\mathbb{P}(X\,\,\text{never returns to the origin}) is simply 2p1=|𝔼(X1)|2p-1=|\mathbb{E}(X_{1})|. In [1], the authors showed that (2) is a deterministic property, and no requirement to involve randomness to get it. In fact they showed the following result.

Theorem 1.

[1] if x=(xn)nx=(x_{n})_{n} is any deterministic sequence mimicking a nearest neighbor random walk, i.e |xn+1xn|1|x_{n+1}-x_{n}|\leq 1 for all nn, then if

xnnn+\frac{x_{n}}{n}\underset{{\scriptscriptstyle n\rightarrow+\infty}}{\longrightarrow}\ell

then

Rn(x)nn+||.\frac{R_{n}(x)}{n}\underset{{\scriptscriptstyle n\rightarrow+\infty}}{\longrightarrow}|\ell|.

Theorem (1) shows that the asymptotic behavior of the range of any nearest neighbor random walk follows immediately from the asymptotic behavior of the random walk itself. This deterministic result covers also the case of random walks in random environments [11]. An immediate consequence of theorem (1) is that if XX is a random walk for which the Birkoff ergodic theorem [10] applies then

Rn(X)nn+|𝔼(X1)|.\frac{R_{n}(X)}{n}\underset{{\scriptscriptstyle n\rightarrow+\infty}}{\longrightarrow}|\mathbb{E}(X_{1})|. (3)

In continuous time, the range of a process X=(Xt)t0X=(X_{t})_{t\geq 0} is defined similarly. It is the Lebesgue measure of the interval [sup0stXsinf0stXs][\sup_{0\leq s\leq t}X_{s}-\inf_{0\leq s\leq t}X_{s}]. In other words

Rt(X)=sup0stXsinf0stXs.R_{t}(X)=\sup_{0\leq s\leq t}X_{s}-\inf_{0\leq s\leq t}X_{s}. (4)

One should notice that the path of Rt(X)R_{t}(X) is non-decreasing. Thus, we can define its right continuous inverse

θX(a):=inf{t0Rt(X)>a}.\theta_{X}(a):=\inf\{t\geq 0\mid R_{t}(X)>a\}.

Note that Rt(X)R_{t}(X) and θX(a)\theta_{X}(a) share the same monotonicity. In addition, the following two relations encodes the duality in between them:

  • Rt(X)<aθX(a)<t.R_{t}(X)<a\Longleftrightarrow\theta_{X}(a)<t.

  • θX(Rt(X))t.\theta_{X}(R_{t}(X))\geq t.

A typical process to study its range in continuous time is obviously the Brownian motion as well as its derivatives [13, 6, 14]. The density of the range of a standard Brownian motion B=BtB=B_{t} is computed in [4] for a fixed time tt while the Laplace transform of θB(a)\theta_{B}(a) is given in [6, 14]. In [13], the authors considered the case of a positive drifted Brownian motion Vtη:=Bt+ηtV_{t}^{\eta}:=B_{t}+\eta t where η>0\eta>0. In section 4,4, they provided the distribution functions of Rt(Vη)R_{t}(V^{\eta}) and θVη(a)\theta_{V^{\eta}}(a). In addition, they proved two limit theorems for θVη(a)\theta_{V^{\eta}}(a) that can be seen as a law of large numbers and a central limit theorem. What matters for us in this work is their result about the long run behavior of the range. They showed that the range of VtηV_{t}^{\eta} is asymptotically equivalent to ηt\eta t (a.e), or equivalently

Rt(Vη)tta.eη.\frac{R_{t}(V^{\eta})}{t}\overset{a.e}{\underset{t\rightarrow\infty}{\longrightarrow}}\eta.

However, their proof is quite technical and it is based on the study of the right continuous inverse process θVη(a):=inf{t0Rt(Vη)>a}\theta_{V^{\eta}}(a):=\inf\{t\geq 0\mid R_{t}(V^{\eta})>a\}, combined with the use of the subadditive ergodic theorem. Their approach will be discussed in the sequel. What we have discovered is that the comportment of the range of VtηV_{t}^{\eta} at infinity does not require that whole machinery to obtain. In fact, it is a deterministic property that follows directly from the behavior of the underlying process. More generally, we show that the long run behavior of the range of a function f:[0,)f:[0,\infty)\rightarrow\mathbb{R} is deducible from the long run behavior of the function itself. To this end, recall that the range of ff is defined exactly as in (4), i.e

Rt(f)=sup0stfsinf0stfs.R_{t}(f)=\sup_{0\leq s\leq t}f_{s}-\inf_{0\leq s\leq t}f_{s}.

When ff is continuous then its range path is also continuous. Moreover, when ff s monotonic then its range increases over the time. Finally note that

Rt(f)=Rt(f).R_{t}(f)=R_{t}(-f). (5)

2 Results and applications

Now, we are ready to state the main result of the paper which is inspired from the work in [1] but has more consequences.

Theorem 2.

Let f,ψ:[0,)f,\psi:[0,\infty)\rightarrow\mathbb{R} be two functions such that ψ(t)\psi(t) is positively increasing to infinity. If

f(t)ψ(t)t0\frac{f(t)}{\psi(t)}\underset{t\rightarrow\infty}{\longrightarrow}\ell\not=0 (6)

then

Rt(f)ψ(t)t||.\frac{R_{t}(f)}{\psi(t)}\underset{t\rightarrow\infty}{\longrightarrow}|\ell|.
Proof.

Without loss of generality we may assume \ell is positive thanks to (5). Let ε(0,)\varepsilon\in(0,\ell). The assumption (6) yields

ψ(t)(ε)f(t)ψ(t)(+ε)\psi(t)(\ell-\varepsilon)\leq f(t)\leq\psi(t)(\ell+\varepsilon) (7)

for tt greater than some positive TT. As ψ(t)\psi(t) is positive then ψ(t)(ε)Rt(f)\psi(t)(\ell-\varepsilon)\leq R_{t}(f). On the other hand, we have

Rt(f)RT(f)+ψ(t)(+ε)R_{t}(f)\leq R_{T}(f)+\psi(t)(\ell+\varepsilon) (8)

since f(t)>0f(t)>0 for t>Tt>T. By combining (7) and (8) we get

2εRt(f)ψ(t)RT(f)ψ(t)+ε.-2\varepsilon\leq\frac{R_{t}(f)}{\psi(t)}-\ell\leq\frac{R_{T}(f)}{\psi(t)}+\varepsilon.

Since RT(f)ψ(t)\frac{R_{T}(f)}{\psi(t)} will vanish at infinity, we can find TTT^{\sharp}\geq T such that

2εRt(f)ψ(t)2ε-2\varepsilon\leq\frac{R_{t}(f)}{\psi(t)}-\ell\leq 2\varepsilon

for all tt greater than TT^{\sharp} which completes the proof by taking into account the case <0\ell<0. ∎

Note that theorem (2) covers also the case when \ell is infinite. This is done by a minor change in the previous proof. Using the same assumptions and notations of theorem (2), the case when \ell is zero is similar and it is the subject of the next theorem.

Theorem 3.

If

f(t)ψ(t)t0\frac{f(t)}{\psi(t)}\underset{t\rightarrow\infty}{\longrightarrow}0 (9)

then

Rt(f)ψ(t)t0.\frac{R_{t}(f)}{\psi(t)}\underset{t\rightarrow\infty}{\longrightarrow}0.
Proof.

The assumption (9) yields

ψ(t)εf(t)ψ(t)ε-\psi(t)\varepsilon\leq f(t)\leq\psi(t)\varepsilon (10)

for tt greater than some positive TT. In particular we get

Rt(f)RT(f)+2ψ(t)εR_{t}(f)\leq R_{T}(f)+2\psi(t)\varepsilon (11)

The inequality (11) yields

Rt(f)ψ(t)RT(f)ψ(t)+2ε.\frac{R_{t}(f)}{\psi(t)}\leq\frac{R_{T}(f)}{\psi(t)}+2\varepsilon.

Since RT(f)ψ(t)\frac{R_{T}(f)}{\psi(t)} will vanish at infinity, we can find TTT^{\dagger}\geq T such that

0Rt(f)ψ(t)3ε0\leq\frac{R_{t}(f)}{\psi(t)}\leq 3\varepsilon

for all tt greater than TT^{\dagger} which completes the proof. ∎

The long run behavior of the the range of the drifted Brownian motion VtηV^{\eta}_{t} follows easily by looking at the behavior of VtηV^{\eta}_{t} itself followed by applying theorems (2) and (3).

Corollary 4.

If VtηV_{t}^{\eta} is a Brownian motion with a drift η\eta (possibly η=0\eta=0) then

Rt(Vη)tta.e|η|.\frac{R_{t}(V^{\eta})}{t}\overset{a.e}{\underset{t\rightarrow\infty}{\longrightarrow}}|\eta|.
Proof.

Recall that Vtη=Bt+ηtV_{t}^{\eta}=B_{t}+\eta t where BtB_{t} is a standard Brownian motion. Then by the law of large numbers for Brownian motion (See [9] for example), BtB_{t} is almost surely negligible against tt, that is Bttta.e0\frac{B_{t}}{t}\overset{a.e}{\underset{t\rightarrow\infty}{\longrightarrow}}0. Thus Vtηtta.e|η|\frac{V_{t}^{\eta}}{t}\overset{a.e}{\underset{t\rightarrow\infty}{\longrightarrow}}|\eta| . Therefore by applying theorems (2) and (3) we get

Rt(Vη)tta.e|η|\frac{R_{t}(V^{\eta})}{t}\overset{a.e}{\underset{t\rightarrow\infty}{\longrightarrow}}|\eta|\,\,\,\,\,\, (12)

In particular we obtain the behavior of the range of BtB_{t}

Rt(B)tta.e0\frac{R_{t}(B)}{t}\overset{a.e}{\underset{t\rightarrow\infty}{\longrightarrow}}0

which is not covered in [13]. The reason we think is because the authors derived the limit (12) by proving fist the following limit of θVη(a)\theta_{V^{\eta}}(a)

θVη(a)aaa.e1|η|\frac{\theta_{V^{\eta}}(a)}{a}\overset{a.e}{\underset{a\rightarrow\infty}{\longrightarrow}}\frac{1}{|\eta|}

which itself requires computations in terms of 1η\frac{1}{\eta} (valid only when η0\eta\neq 0). Then they deduced the asymptotic behavior of the range Rt(Vη)R_{t}(V^{\eta}) leaning on the fact that θVη(a)\theta_{V^{\eta}}(a) is the right continuous process of Rt(Vη)R_{t}(V^{\eta}). Now by theorems (2) and (3), the asymptotic comportment of the range of VηV^{\eta} are obtained independently of θVη(x)\theta_{V^{\eta}}(x) and follows exclusively from the behavior of the underlying process VtηV^{\eta}_{t}.

Now we show that the behavior of θVη(a)\theta_{V^{\eta}}(a) is obtainable from that of Rt(Vη)R_{t}(V^{\eta}) and vice versa. The main tool to do so is the following result relating the asymptotic behaviors of a non decreasing function and its generalized inverse. For a survey about generalized inverses, we refer the reader to [3].

Proposition 5.

Let ξ:+\xi:\mathbb{R}_{+}\rightarrow\mathbb{R} be a non decreasing function and let ξ\xi^{\dagger} be its generalized inverse, i.e

ξ(y)=inf{xξ(x)>y}.\xi^{\dagger}(y)=\inf\{x\mid\xi(x)>y\}. (13)

If

ξ(x)xx\frac{\xi(x)}{x}\underset{x\rightarrow\infty}{\longrightarrow}\ell

then

ξ(x)xx1.\frac{\xi^{\dagger}(x)}{x}\underset{x\rightarrow\infty}{\longrightarrow}\frac{1}{\ell}.
Proof.

The statement is straightforward when ξ\xi is continuously increasing since in such a case ξ(y)=ξ1(y)\xi^{\dagger}(y)=\xi^{-1}(y). Otherwise fix a positive ε\varepsilon (less than \ell if >0\ell>0) and let τ\tau be a large number so that

x(ε)ξ(x)x(+ε)x(\ell-\varepsilon)\leq\xi(x)\leq x(\ell+\varepsilon)

for xτx\geq\tau. It is not hard to see that for x>τx>\tau, the curve of ζ\zeta is squeezed between x+ε\frac{x}{\ell+\varepsilon} and xε\frac{x}{\ell-\varepsilon} 444Jumps (resp. flat parts) of ξ\xi becomes flat parts (resp. jumps) for ζ\zeta. In other words

x(+ε)ξ(x)x(ε)\displaystyle\frac{x}{(\ell+\varepsilon)}\leq\xi^{\dagger}(x)\leq\frac{x}{(\ell-\varepsilon)} if >0\ell>0 (14)
xεξ(x)xε\displaystyle-\frac{x}{\varepsilon}\leq\xi^{\dagger}(x)\leq\frac{x}{\varepsilon} if =0\ell=0

for xx larger than τ\tau. The result follows by considering the reciprocals in (14). ∎

Remark 6.

Another common definition of the generalized inverse is ξ(y):=inf{xξ(x)y}\xi^{*}(y):=\inf\{x\mid\xi(x)\geq y\}. proposition (5) is still valid if we use ξ\xi^{*} since ξ=ξ\xi^{*}=\xi^{\dagger} a.e 555It is because the set of discontinuity of a non decreasing function is countable. The only difference is that the formula (13) generates a right continuous function while ξ\xi^{*} generates a left continuous one. A very known usage of the left continuous inverse is when ξ\xi represents the cumulative distribution function of some distribution μ.\mu. In such a case, ξ\xi^{*} is used to simulate a random variable of the same law μ\mu. More precisely

ξ(Uni(0,1))μ.\xi^{*}(\text{Uni}(0,1))\sim\mu.

That is, the uniform distribution over (0,1)(0,1) is mapped to a random variable of distribution μ\mu. Note also that proposition (5) works in both directions since (ξ)=ξ\left(\xi^{\dagger}\right)^{\dagger}=\xi a.e. In particular, \ell could be infinite.

A straightforward consequence of proposition (5) is the equivalence between the comportment of θVη(x)\theta_{V^{\eta}}(x) and Rt(Vη)R_{t}(V^{\eta}). This appeared only as implication in [13].

Proposition 7.

We have

θVη(x)xxa.e1η.\frac{\theta_{V^{\eta}}(x)}{x}\overset{a.e}{\underset{x\rightarrow\infty}{\longrightarrow}}\frac{1}{\eta}.

In particular

θB(x)xxa.e+.\frac{\theta_{B}(x)}{x}\overset{a.e}{\underset{x\rightarrow\infty}{\longrightarrow}}+\infty.

The following diagram illustrates the difference between our approach and that in [13].

Behavior of VtηV_{t}^{\eta}Behavior of Rt(Vη)R_{t}(V^{\eta})Behavior of θVη(x)\theta_{V^{\eta}}(x)
Figure 1: The continuous bent arrows illustrate our approach while the dashed ones illustrate that in [13].

Another application of proposition (5) is the recovery of the limit theorem for renewal processes (We refer the reader to [8] for more details). That is, let NtN_{t} be a renewal counting process associated to some renewal sequence TnT_{n} defined by

Nt=sup{nTnt}.N_{t}=\sup\{n\mid T_{n}\leq t\}. (15)

Note that the RHS of (15) is simply n=1𝟏{Tnt}\sum_{n=1}^{\infty}\mathbf{1}_{\{T_{n}\leq t\}}. The process NtN_{t} is the generalized inverse of the renewal sequence TnT_{n} (the right continuous one). More precisely, TnT_{n} is obtainable from NtN_{t} by

Tn=inf{tNtn}.T_{n}=\inf\{t\mid N_{t}\geq n\}.

That is, we recover the following equivalence

Tnnna.eγNttta.e1γ\frac{T_{n}}{n}\overset{a.e}{\underset{n\rightarrow\infty}{\longrightarrow}}\gamma\Longleftrightarrow\frac{N_{t}}{t}\overset{a.e}{\underset{t\rightarrow\infty}{\longrightarrow}}\frac{1}{\gamma}

valid even in the case when the sequence of inter-arrivals TnTn1T_{n}-T_{n-1} is ergodic. Both theorems (2) and (3) can be seen as the continuous version of the main result stated in [1]. However, the discrete version can follow easily from the continuous case. That is, let (xn)\left(x_{n}\right) be a sequence of integers mimicking a simple walk on \mathbb{Z} and let rnr_{n} be the range of (xn)\left(x_{n}\right), i.e the number of explored sites up to time nn by (xn)\left(x_{n}\right). Let xtx_{t}^{\star} be the function obtained from (xn)\left(x_{n}\right) by connecting the dots. It is clear that

|Rt(x)rt|2.|R_{t}(x^{\star})-r_{\lfloor t\rfloor}|\leq 2.

Hence, in case of existence of limits, we obtain

limnrnn=limtrtt=limtRt(x)t.\lim_{n\rightarrow\infty}\frac{r_{n}}{n}=\lim_{\lfloor t\rfloor\rightarrow\infty}\frac{r_{\lfloor t\rfloor}}{\lfloor t\rfloor}=\lim_{t\rightarrow\infty}\frac{R_{t}(x^{\star})}{t}.

A typical process derived from Brownian motion is the so called Bessel process. Bessel processes are important as they interfere in financial modeling of stock market prices etc. A nn-dimensional Bessel process, denoted by BES(n)BES(n), is defined by

BES(n)=𝑩tBES(n)=\|\boldsymbol{B}_{t}\|

where \|\cdot\| is the Euclidean norm and 𝐁t\mathbf{B}_{t} is a nn-dimensional Brownian motion 666People often add the starting point as a subscript.. Another alternative definition of Bessel processes uses SDE theory. More precisely, a Bessel process is the solution of the following SDE

dXt=dBt+n12dtXt.dX_{t}=dB_{t}+\frac{n-1}{2}\frac{dt}{X_{t}}.

We refer the reader to [7] for more details about the subject. The determination of the statistic of the range of a Bessel process is not an easy task, not to mention its long run behavior. Now, theorems (2) and (3) allows us to obtain the asymptotic behavior of such a range by knowing only the behavior of the process itself. However, we provide a result that covers more than the case of Bessel processes.

Proposition 8.

Let 𝐁t=(𝐁t(1),𝐁t(2),,𝐁t(n))\boldsymbol{B}_{t}=(\boldsymbol{B}_{t}^{(1)},\boldsymbol{B}_{t}^{(2)},...,\boldsymbol{B}_{t}^{(n)}) be an nn-dimensional Brownian motion and let 0<p+0<p\leq+\infty. Then

Rt(𝑩p)tta.e0(a.e)\frac{R_{t}(\|\boldsymbol{B}\|_{p})}{t}\overset{a.e}{\underset{t\rightarrow\infty}{\longrightarrow}}0\,\,\,\,\,\,(a.e)

where 𝐁p\|\boldsymbol{B}\|_{p} is the pthp^{th} norm 777 It is not a true norm when p<1p<1. We use the word by abuse of of terminology. of 𝐁t\boldsymbol{B}_{t}, i.e

𝑩p:={(|Bt(1)|p+|Bt(2)|p++|Bt(n)|p)1pp<+max1in|𝑩t(i)|p=+.\|\boldsymbol{B}\|_{p}:=\begin{cases}\left(|B_{t}^{(1)}|^{p}+|B_{t}^{(2)}|^{p}+...+|B_{t}^{(n)}|^{p}\right)^{\frac{1}{p}}&p<+\infty\\ \max_{1\leq i\leq n}|\boldsymbol{B}_{t}^{(i)}|&p=+\infty.\end{cases}
Proof.

Let p<+p<+\infty. The law of large number implies that

|Bt(j)|ptpta.e0\frac{|B_{t}^{(j)}|^{p}}{t^{p}}\overset{a.e}{\underset{t\rightarrow\infty}{\longrightarrow}}0

for j=1,,nj=1,...,n. It follows that

𝑩pt=(|Bt(1)|ptp+|Bt(2)|ptp++|Bt(n)|ptp)1pta.e0.\begin{aligned} \frac{\|\boldsymbol{B}\|_{p}}{t}&=\left(\frac{|B_{t}^{(1)}|^{p}}{t^{p}}+\frac{|B_{t}^{(2)}|^{p}}{t^{p}}+...+\frac{|B_{t}^{(n)}|^{p}}{t^{p}}\right)^{\frac{1}{p}}\\ &\overset{a.e}{\underset{t\rightarrow\infty}{\longrightarrow}}0\end{aligned}.

For p=+p=+\infty we have

𝑩pt(|Bt(1)|t+|Bt(2)|t++|Bt(n)|t)ta.e0.\begin{aligned} \frac{\|\boldsymbol{B}\|_{p}}{t}&\leq\left(\frac{|B_{t}^{(1)}|}{t}+\frac{|B_{t}^{(2)}|}{t}+...+\frac{|B_{t}^{(n)}|}{t}\right)\\ &\overset{a.e}{\underset{t\rightarrow\infty}{\longrightarrow}}0\end{aligned}.

Therefore, in both cases

Rt(𝑩p)tta.e0.\frac{R_{t}(\|\boldsymbol{B}\|_{p})}{t}\overset{a.e}{\underset{t\rightarrow\infty}{\longrightarrow}}0.

Corollary 9.

The range of a Bessel process is negligible against the time.

The following result illustrates the relation between a deterministic function ff and its underlying supremum.

Proposition 10.

Let f,ψ:[0,)f,\psi:[0,\infty)\rightarrow\mathbb{R} be two functions such that ψ(t)\psi(t) is positively increasing to infinity and let >0\ell>0. If

f(t)ψ(t)t\frac{f(t)}{\psi(t)}\underset{t\rightarrow\infty}{\longrightarrow}\ell (16)

then

sup0stfsf(t)t1.\frac{\sup_{0\leq s\leq t}f_{s}}{f(t)}\underset{t\rightarrow\infty}{\longrightarrow}1.
Proof.

From the assumption, we know for every ε>0\varepsilon>0 there exists a large enough TT such that, when t>Tt>T

ψ(t)(ε)f(t)ψ(t)(+ε).\psi(t)(\ell-\varepsilon)\leq f(t)\leq\psi(t)(\ell+\varepsilon). (17)

Also,

1=f(t)f(t)sup0stfsf(t)1=\frac{f(t)}{f(t)}\leq\frac{\sup_{0\leq s\leq t}f_{s}}{f(t)} (18)

for t>Tt>T. On the other hand, we have

sup0stfsf(t)sup0sTfs+supTstfsf(t)sup0sTfs+ψ(t)(+ε)f(t)\frac{\sup_{0\leq s\leq t}f_{s}}{f(t)}\leq\frac{\sup_{0\leq s\leq T}f_{s}+\sup_{T\leq s\leq t}f_{s}}{f(t)}\leq\frac{\sup_{0\leq s\leq T}f_{s}+\psi(t)(\ell+\varepsilon)}{f(t)} (19)

Letting tt go to infinity, sup0sTfsf(t)\frac{\sup_{0\leq s\leq T}f_{s}}{f(t)} will converge to 0 as f(t)f(t) diverges to ++\infty. ψ(t)(+ε)f(t)\frac{\psi(t)(\ell+\varepsilon)}{f(t)} will converge to 1+ε1+\frac{\varepsilon}{\ell}. Therefore,

1lim inftsup0stfsf(t)lim suptsup0stfsf(t)1+ε1\leq\liminf_{t\to\infty}\frac{\sup_{0\leq s\leq t}f_{s}}{f(t)}\leq\limsup_{t\to\infty}\frac{\sup_{0\leq s\leq t}f_{s}}{f(t)}\leq 1+\frac{\varepsilon}{\ell} (20)

The inequality (19) holds for any ε>0\varepsilon>0. Thus,

sup0stfsf(t)t1.\frac{\sup_{0\leq s\leq t}f_{s}}{f(t)}\underset{t\rightarrow\infty}{\longrightarrow}1. (21)

Corollary 11.

Under the same assumptions of Proposition (10), we have

sup0stfsψ(t)t\frac{\sup_{0\leq s\leq t}f_{s}}{\psi(t)}\underset{t\rightarrow\infty}{\longrightarrow}\ell

and

limt+Rt(f)ψ(t)=limt+Rt(supf)ψ(t)=.\lim_{t\rightarrow+\infty}\frac{R_{t}(f)}{\psi(t)}=\lim_{t\rightarrow+\infty}\frac{R_{t}(\sup f)}{\psi(t)}=\ell.

In other words, the four functions f,sup0stfs,Rt(f),Rt(supf)f,\sup_{0\leq s\leq t}f_{s},R_{t}(f),R_{t}(\sup f) have the same long run behavior under the assumptions of proposition (10). Note that proposition (10) is still valid when =0\ell=0 provided that ff becomes eventually positive after certain time. The following result is a consequence of proposition (10). It concerns the underlying supremum process of a drifted Brownian motion.

Proposition 12.

Let V~tη=sup0stVsη\widetilde{V}_{t}^{\eta}=\sup_{0\leq s\leq t}V_{s}^{\eta}. Then

V~tηtta.eη\frac{\widetilde{V}_{t}^{\eta}}{t}\overset{a.e}{\underset{t\rightarrow\infty}{\longrightarrow}}\eta

and

Rt(V~η)tta.eη.\frac{R_{t}(\widetilde{V}^{\eta})}{t}\overset{a.e}{\underset{t\rightarrow\infty}{\longrightarrow}}\eta.


A corollary of proposition (12) is that the supremum of a Brownian motion, as well as its range are negligible against the time evolution. However, it would be interesting to derive estimates for the range like the law of iterated logarithm for Brownian motion, and on top of that investigate whether such potential properties are deterministic or not.

Acknowledgments

The authors would like to thank Greg Markowsky for helpful communications.

References

  • [1] M. Boudabra and G. Markowsky. Remarks on the speeds of a class of random walks on the integers. Discrete Mathematics, 344(8):112436, 2021.
  • [2] K. S. Chong, R. Cowan, and L. Holst. The ruin problem and cover times of asymmetric random walks and brownian motions. Advances in Applied Probability, 32(1):177–192, 2000.
  • [3] A. de La Fortelle. A study on generalized inverses and increasing functions part i: generalized inverses. 2015.
  • [4] W. Feller. The asymptotic distribution of the range of sums of independent random variables. The annals of mathematical statistics, pages 427–432, 1951.
  • [5] P. W. Glynn. On the range of a regenerative sequence. Stochastic Processes and Their Applications, 20(1):105–113, 1985.
  • [6] J. P. Imhof. On the range of brownian motion and its inverse process. The Annals of Probability, 13(3):1011–1017, 1985.
  • [7] G. F. Lawler. Notes on the bessel process. Lecture notes. Available on the webpage of the author, pages 523–531, 2018.
  • [8] K. V. Mitov and E. Omey. Renewal processes. In Renewal Processes, pages 1–51. Springer, 2014.
  • [9] P. Mörters and Y. Peres. Brownian Motion, volume 30. Cambridge University Press, 2010.
  • [10] A. N Shiryaev. Probability-1, volume 95. Springer, 2016.
  • [11] F. Solomon. Random walks in a random environment. The Annals of Probability, 3(1):1–31, 1975.
  • [12] F. Spitzer. Principles of random walk, volume 34. Springer Science & Business Media, 2001.
  • [13] E. Tanré and P. Vallois. Range of brownian motion with drift. Journal of Theoretical Probability, 19(1):45–69, 2006.
  • [14] P. Vallois. Decomposing the brownian path via the range process. Stochastic Processes and Their Applications, 55(2):211–226, 1995.
  • [15] P. Vallois. The range of a simple random walk on \mathbb{Z}. Advances in Applied Probability, 28(4):1014–1033, 1996.
  • [16] P. Vallois and Charles S Tapiero. Range reliability in random walks. Mathematical Methods of Operations Research, 45:325–345, 1997.