We consider transient nearest-neighbor random walks X:={Xn}n≥0 on the half-line, whose transition probabilities are state-dependent with certain asymptotic perturbations. This is a specific case of Lamperti's random walks. Let Mt:=max{Xi,0≤i≤t} be the maximum process of X, and Tn:=inf{t≥0,Xt=n} be its inverse process. Hong&Yang (2019) provided the law of large numbers for Tn. In this paper, we study the large deviations for Mt and Tn with speeds less than n. This indicates that the perturbations slow down the random walk.
Citation: Hui Yang. Large deviations for transient random walks with asymptotically zero drifts[J]. AIMS Mathematics, 2025, 10(4): 8777-8793. doi: 10.3934/math.2025402
[1] | You Lv . Asymptotic behavior of survival probability for a branching random walk with a barrier. AIMS Mathematics, 2023, 8(2): 5049-5059. doi: 10.3934/math.2023253 |
[2] | Qingwu Gao, Wenlei Pan . Precise large deviations of aggregate claims in a nonstandard risk model with arbitrary dependence between claim sizes and waiting times. AIMS Mathematics, 2023, 8(1): 2191-2200. doi: 10.3934/math.2023113 |
[3] | Mohamed Abdelkader, Rafik Aguech . Moran random walk with reset and short memory. AIMS Mathematics, 2024, 9(8): 19888-19910. doi: 10.3934/math.2024971 |
[4] | Frithjof Lutscher, Thomas Hillen . Correlated random walks in heterogeneous landscapes: Derivation, homogenization, and invasion fronts. AIMS Mathematics, 2021, 6(8): 8920-8948. doi: 10.3934/math.2021518 |
[5] | Murugan Suvinthra, Krishnan Balachandran, Rajendran Mabel Lizzy . Large Deviations for Stochastic Fractional Integrodifferential Equations. AIMS Mathematics, 2017, 2(2): 348-364. doi: 10.3934/Math.2017.2.348 |
[6] | He Song, Longmin Wang, Kainan Xiang, Qingpei Zang . The continuity of biased random walk's spectral radius on free product graphs. AIMS Mathematics, 2024, 9(7): 19529-19545. doi: 10.3934/math.2024952 |
[7] | Najmeddine Attia . Some remarks on recursive sequence of fibonacci type. AIMS Mathematics, 2024, 9(9): 25834-25848. doi: 10.3934/math.20241262 |
[8] | Ruonan Liu, Tomás Caraballo . Random dynamics for a stochastic nonlocal reaction-diffusion equation with an energy functional. AIMS Mathematics, 2024, 9(4): 8020-8042. doi: 10.3934/math.2024390 |
[9] | Gülnur Şaffak Atalay . A new approach to special curved surface families according to modified orthogonal frame. AIMS Mathematics, 2024, 9(8): 20662-20676. doi: 10.3934/math.20241004 |
[10] | Luigi Accardi, Amenallah Andolsi, Farrukh Mukhamedov, Mohamed Rhaima, Abdessatar Souissi . Clustering quantum Markov chains on trees associated with open quantum random walks. AIMS Mathematics, 2023, 8(10): 23003-23015. doi: 10.3934/math.20231170 |
We consider transient nearest-neighbor random walks X:={Xn}n≥0 on the half-line, whose transition probabilities are state-dependent with certain asymptotic perturbations. This is a specific case of Lamperti's random walks. Let Mt:=max{Xi,0≤i≤t} be the maximum process of X, and Tn:=inf{t≥0,Xt=n} be its inverse process. Hong&Yang (2019) provided the law of large numbers for Tn. In this paper, we study the large deviations for Mt and Tn with speeds less than n. This indicates that the perturbations slow down the random walk.
Consider a nearest-neighbor random walk defined as follows. X={Xn}n≥0 is a Markov chain on N={0,1,2,⋯} with X0=0 and for n≥1, the transition probabilities are
pi:=P(Xn+1=i+1 | Xn=i)=1−P(Xn+1=i−1 | Xn=i)={1,if i=0,12+Biα,if i≥1, | (1.1) |
where α,B>0. This random walk X describes the motion of particles that starts at zero, moves on the nonnegative integers, and goes away from 0 with a larger probability than in the direction of 0. Obviously, Biα goes to 0 as i→∞. This means that the state 0 has a repelling power that decreases as the particle moves away from 0 (see [4]). This is a special case of the so-called Lamperti's random walk [7,9].
The transience and recurrence for X are well-known results in the literature (e.g., Chung [3]). For i≥1, denote
ρi=1−pipi. |
Theorem A ([3]) The random walk X defined in (1.1) is transient if and only if
∞∑n=1n∏i=1ρi<∞. |
However, this criterion does not explicitly reveal the transient/recurrent classification for sequences of the form Biα. Lamperti ([7,9]) proved a more general theorem regarding the recurrence and transience of real nonnegative processes. Here we explicitly characterize the types of Biα sequences and discuss their implications for the transience and recurrence of the random walk X. These results also follow straightforwardly from Theorem A.
Transience criterion. For sufficiently large i:
● When 0<α<1, the random walk X is transient if B>0 and recurrent if B<0.
● When α=1, the random walk X is transient if B>14 and recurrent if B≤14.
Numerous studies have investigated the limiting behavior of X depending on the sequence Biα. Lamperti [8] established the weak convergence of X to a Bessel process. The law of the iterated logarithm for X was provided in [1,12]. In [13], Voit proved the law of large numbers for X, which we restate here in our specific framework.
Theorem B ([13]) If Ei=Biα, where 0<α<1 and B>0, then
limn→∞Xnn1/(1+α)=[2B(1+α)]1/(1+α)almostsurely. |
There is a minor mistake in [13] about the limit value. Hong&Yang [6] gave the correct form and provided the limit of Tn, which is defined as follows.
For n≥0, denote
Tn=inf{t≥0,Xt=n}. | (1.2) |
Theorem C ([6]) If 0<α<1 and B>0, then
limn→∞Tnn1+α=12B(1+α)almostsurely. |
On this basis, we investigate the large deviations for the sequences of hitting times {Tn,n≥1}. Additionally, for t∈N, let
Mt=max{Xi,0≤i≤t} |
be the maximum of the random walk X up to time t. Note that Tn defined in (1.2) is the inverse process of Mt. Observing the relationship between Mt and Tn, we study the limit theorems and large deviations for Mt.
The paper is organized as follows. Section 2 presents the main results. Section 3 provides auxiliary results required for the proofs. The proofs of the main theorems are contained in Section 4.
This section presents the main results of the paper. It is divided into three parts. The first part establishes the law of large numbers for Mt, and the second part addresses the large deviation principles for Tn. Finally, we derive the large deviation principles for Mt.
Theorem 2.1. (Law of Large Numbers for Mt) When 0<α<1 and B>0, we have
limn→∞Mnn1/(1+α)=[2B(1+α)]1/(1+α)almostsurely. |
Po-Ning Chen [2] established a generalization of the Gärtner–Ellis Theorem for arbitrary random sequences. All subsequent definitions in this section are adapted from reference [2].
Definition 2.1. Define
Λn(λ)=1n1−αlogEexp(λn1−αTnn1+α) |
and
¯Λ(λ):=lim supn→∞Λn(λ),Λ_(λ):=lim infn→∞Λn(λ). |
The sup-large deviation rate function of {Tn}∞n=0 is defined as
¯Λ∗(x)=sup{λ∈R,¯Λ(λ)>−∞}{λx−¯Λ(λ)}, | (2.1) |
and the inf-large deviation rate function is defined as
Λ_∗(x)=sup{λ∈R,Λ_(λ)>−∞}{λx−Λ_(λ)}. | (2.2) |
Actually, we have ¯Λ(λ)≥Λ_(λ)>−∞ for every λ∈R (see Property 2). So the ranges of the supremum operations in (2.1) and (2.2) are always {λ∈R}. Hence, ¯Λ∗(x) and Λ_∗(x) are always defined.
Definition 2.2. Define the sup-Gärter–Ellis set as
¯G△=⋃{λ∈R,¯Λ(λ)>−∞}¯G(λ) |
where
¯G(λ)△={x∈R:lim supt↓0¯Λ(λ+t)−¯Λ(λ)t≤x≤lim inft↓0¯Λ(λ)−¯Λ(λ−t)t}. |
Define the inf-Gärter–Ellis set as
G_△=⋃{λ∈R,Λ_(λ)>−∞}G_(λ) |
where
G_(λ)△={x∈R:lim supt↓0Λ_(λ+t)−Λ_(λ)t≤x≤lim inft↓0Λ_(λ)−Λ_(λ−t)t}. |
Let us briefly remark on the sup-Gärter–Ellis set defined above. The definitions of ¯G and G_ are special cases of Definitions 3.4 and 3.5 in [2], which only require h(x)=x. By Property 2, we can see that ¯Λ′(λ) and Λ_′(λ) exist for |λ|≤2B2. So it can be derived that {x=¯Λ′(λ):|λ|≤2B2}⊂¯G and {x=Λ_′(λ):|λ|≤2B2}⊂G_.
Theorem 2.2. (Large Deviation Principles for Tn)
(1) For any closed set F⊂R, we have
lim supn→∞1n1−αlogP{Tnn1+α∈F}≤−infx∈F¯Λ∗(x) | (2.3) |
and
lim infn→∞1n1−αlogP{Tnn1+α∈F}≤−infx∈FΛ_∗(x). | (2.4) |
(2) For any (a,b)⊂¯G, we have
lim supn→∞1n1−αlogP{Tnn1+α∈(a,b)}≥−infx∈(a,b)¯Λ∗(x). |
(3) For any (a,b)⊂G_, we have
lim infn→∞1n1−αlogP{Tnn1+α∈(a,b)}≥−infx∈(a,b)Λ_∗(x). |
Theorem 2.3. (Large Deviation Principles for Mt) Define
¯I(x)=x1−α¯Λ∗(1x1+α),I_(x)=x1−αΛ_∗(1x1+α). |
Then,
(1) for any closed set F⊂R,
lim supn→∞1n1−α1+αlogP(Mnn11+α∈F)≤−infx∈F¯I(x), |
and
lim infn→∞1n1−α1+αlogP(Mnn11+α∈F)≤−infx∈FI_(x), |
(2) for any open set G,
lim infn→∞1n1−α1+αlogP(Mnn11+α∈G)≥−infx∈G∩{v,1v1+α∈G_o∩¯Go}I_(x), |
where G_o and ¯Go represent the interior of G_ and ¯G respectively.
Property 3.1. ¯Λ∗(x) and Λ_∗(x) are the sup- and inf-large deviation rate functions of {Tn}∞n=0 respectively. Denote ¯x=12B(1+α), then
(1) ¯Λ∗(x) and Λ_∗(x) are always defined;
(2) ¯Λ∗(x) and Λ_∗(x) are both convex rate function;
(3) ¯Λ∗(x) is continuous over {x∈R:¯Λ∗(x)<∞}. Likewise, Λ_∗(x) is continuous over {x∈R:Λ_∗(x)<∞};
(4) ¯Λ∗(¯x)=Λ_∗(¯x)=0.
Before proving Property 3.1, we provide some preparatory works.
Lemma A ([11], Lemma 17) Let p=(pi)i∈Z and q=(qi)i∈Z such that
pi≤qi,∀i∈Z. |
Then we can construct the random walks {Xn,n∈N} and {Yn,n∈N}, respectively associated with p and q, and starting from any common point d∈Z, such that
∀n∈N,Xn≤Ynalmostsurely. |
In [14], Ke Zhou provided explicit expressions for the generating functions of hitting times of the skip-free Markov chain on Z+. The chain's upward jumps are restricted to unit size; moreover, it starts at state 0 and is absorbed by state d. The result relevant to our case is stated as follows.
Assume that the random walk defined in 1.1 is absorbed by state d and that pi≡p∈(0,1) for 1≤i≤d−1. We denote this random walk as X∗. Let P be the transition probability matrix of X∗, and let τd−1,d denote the hitting time of state d when starting from state d−1.
Lemma B ([14]) Denote fd−1(s) as the generating function of τd−1,d; then we have
fd−1(s)=psdet[Ad−1(s)]det[Ad(s)] |
where matrix Ai(s) is the first i rows and first i lines of Id+1−P for i=d−1,d. Id+1 is (d+1)×(d+1) unit matrix.
Actually, Lemma B can be derived from the proof of Theorem 1.1 in [14]. Specifically, we have
fd−1(s)=φd(s)φd−1(s), |
where φd(s) is a symbol defined in [14], representing the generating function of the hitting time from state 0 to state d. We omit further details here.
Lemma 3.1. Let q=1−p; we have
det[Ad−1(s)]det[Ad(s)]=(ηd−3−βd−3)g3(s)−(ηd−3β−ηβd−3)g2(s)(ηd−2−βd−2)g3(s)−(ηd−2β−ηβd−2)g2(s) | (3.1) |
where η,β are the roots of equation x2−x+pqs2=0, and g2(s)=1−s2q,g3(s)=1−s2q−pqs2.
Proof. We will prove this result using mathematical induction. We can easily calculate that det[A2(s)]=1−s2q=g2(s) and det[A3(s)]=1−s2q−pqs2=g3(s). Clearly, (3.1) holds for d=3. Assume (3.1) is also hodes for d=k. To calculate det[Ak+1(s)], we expand along the last row and obtain
det[Ak+1(s)]=det[Ak(s)]−pqs2det[Ak−1(s)]. |
By the quadratic formula, we have
β=1−√1−4pqs22,η=1+√1−4pqs22, |
satisfying ηβ=pqs2 and η+β=1. Therefore, for d=k+1, we have
det[Ak(s)]det[Ak+1(s)]=det[Ak(s)]det[Ak(s)]−ηβdet[Ak−1(s)]=11−ηβdet[Ak−1(s)]det[Ak(s)]=11−ηβ(ηk−3−βk−3)g3(s)−(ηk−3β−ηβk−3)g2(s)(ηk−2−βk−2)g3(s)−(ηk−2β−ηβk−2)g2(s)=(ηk−2−βk−2)g3(s)−(ηk−2β−ηβk−2)g2(s)(ηk−2−βk−2−ηk−2β+ηβk−2)g3(s)−(ηk−2β−ηβk−2−ηk−2β2+η2βk−2)g2(s)=(ηk−2−βk−2)g3(s)−(ηk−2β−ηβk−2)g2(s)(ηk−1−βk−1)g3(s)−(ηk−1β−ηβk−1)g2(s), |
that is to say, (3.1) is true for d=k+1. In conclusion, (3.1) is true for all d. This completes the inductive step.
The assumptions 0<α<1,B>0 will be used throughout the paper. Recall the definition of Tn in (1.2); let
τi=Ti−Ti−1,fori≥1. | (3.2) |
Lemma 3.2. For λ≤2B2, we have
Eeλn2ατi→1,asn→∞ |
uniform about 1≤i≤n.
Proof. By Lemma A, let p(1)=(p1,p2,⋯,pi−1) and p(2)=(pi,⋯,pi) be (i−1)-dimensional vectors. Obviously, for every 1≤j≤i−1, pj>pi. We construct the random walks {X(1)n}n∈N and {X(2)n}n∈N, associated with p(1) and p(2) respectively, starting from i−1, reflected at 0, and absorbed at i, such that
∀n∈N,X(1)n≥X(2)nalmost surely. |
For j=1,2, denote τ(j)i=inf{n∈N,X(j)n=i} as the passage time of X(j) starting from i−1 and ending at i. Then
τ(1)i≤τ(2)ialmost surely. |
implying Esτ(1)i≤Esτ(2)i. By Lemma B and Lemma 3.1, the generating function of τ(2)i is:
fi−1(s)=Esτ(2)i=spidet[Ai−1(s)]det[Ai(s)]=spi(ηi−3−βi−3)g3(s)−(ηi−3β−ηβi−3)g2(s)(ηi−2−βi−2)g3(s)−(ηi−2β−ηβi−2)g2(s)=spi1+ηg2(s)−g3(s)g3(s)−βg2(s)γi−3η+βηg2(s)−g3(s)g3(s)−βg2(s)γi−3, | (3.3) |
where η,β are the roots of equation x2−x+piqis2=0, and
g3(s)=1−s2qi−piqis2,g2(s)=1−s2qi,γ=βη. |
By the definitions of Xn in (1.1) and τi in (3.2), the random variables τ(1)i and τi share the same distribution. Consequently, their generating functions Esτi=Esτ(1)i≤Esτ(2)i. Letting s=eλn2α and considering sufficiently large n large enough, we employ the representation τ(2)i from (3.3) to derive that
Eeλn2ατ(2)i≤Eeλn2ατ(2)n=spn1+ηg2(s)−g3(s)g3(s)−βg2(s)γn−3η+βηg2(s)−g3(s)g3(s)−βg2(s)γn−3,for∀1≤i≤n. |
Hence, to finish the proof of the lemma, we just need to verify that Eeλn2ατ(2)n→1, as n→∞. To ensure the existence of η,β, we require
s2=e2λn2α≤14pnqn, |
where pn=1−qn=12+Bnα. In other words,
λ≤12log1[(1−4B2n2α)n2α4B2]4B2↓2B2. |
Now,
γ=βη=1−√1−4pnqne2λn2α1+√1−4pnqne2λn2α∼1−2√4B2−2λnα,asn→∞, |
so limn→∞γn−3=0.
Additionally, note that asn→∞,
⋅s=eλn2α→1,⋅pn=12+Bnα→12,⋅qn=1−pn→12. |
From this, we can conclude that
η=1+√1−4pnqne2λn2α2→12,β=1−√1−4pnqne2λn2α2→12,asn→∞. |
Similarly, we have
ηg2(s)−g3(s)g3(s)−βg2(s)=η(1−s2qn)−(1−s2qn−pnqns2)(1−s2qn−pnqns2)−β(1−s2qn)→constant,asn→∞. |
Hence, for λ≤2B2,
Eeλn2ατ(2)n→1,asn→∞. |
Lemma 3.3. When |λ|≤2B2, the series
∞∑j=1λjj!(2j−3)!!(4B2)2j−12 |
are convergent.
Proof. Indeed,
(2(j+1)−3)!!(j+1)!/(2j−3)!!j!→2,asj→∞. |
Therefore, when |λ4B2|<12, i.e., |λ|<2B2, the series
∞∑j=1λjj!(2j−3)!!(4B2)2j−12 |
converges.
When λ=2B2,
j[1−((2B2)j+1(j+1)!(2(j+1)−3)!!(4B2)2(j+1)−12)/((2B2)jj!(2j−3)!!(4B2)2j−12)]=j(32j+2)→32≥r>1,asj→∞. |
By the label discriminant, we can conclude that the series ∞∑j=1(2B2)jj!(2j−3)!!(4B2)2j−12 is convergent.
Similarly, when λ=−2B2, the series ∞∑j=1(−2B2)jj!(2j−3)!!(4B2)2j−12 is also convergent.
Lemma 3.4. Suppose 0<α<1; then we have
−∞<∞∑j=1λjj!L(j)=Λ_(λ)≤¯Λ(λ)=∞∑j=1λjj!U(j)<∞ |
for |λ|≤2B2, where
L(j):=lim infn→∞1n1+(2j−1)αn∑i=1E(τi)j,U(j):=lim supn→∞1n1+(2j−1)αn∑i=1E(τi)j. |
Proof. Recalling that the random variables τi,i≥1, defined in (3.2), are independent, we observe that for any fixed λ≤2B2, the following holds:
Λn(λ)=1n1−αlogEexp(λn1−αTnn1+α)=1n1−αlogEexp(λn2αn∑i=1τi)=1n1−αn∑i=1log(1+Eeλn2ατi−1)∼1n1−αn∑i=1(Eeλn2ατi−1)=1n1−αn∑i=1(∞∑j=1E(λn2ατi)jj!)=1n1−α∞∑j=11j!(λn2α)jn∑i=1E(τi)j=∞∑j=1λjj!1n1+(2j−1)αn∑i=1E(τi)j. |
Hence,
∞∑j=1λjj!L(j)=lim infn→∞Λn(λ)=Λ_(λ)≤¯Λ(λ)=lim supn→∞Λn(λ)=∞∑j=1λjj!U(j), |
where,
L(j):=lim infn→∞1n1+(2j−1)αn∑i=1E(τi)j,U(j):=lim supn→∞1n1+(2j−1)αn∑i=1E(τi)j. |
Furthermore, by (3.3), we obtain
E(τ(2)nn2α)j=f(j)n−1(0)∼(2j−3)!!(4B2)2j−121nα,asn→∞. |
Consequently, for all 1≤i≤n, it follows that
1n1+(2j−1)αn∑i=1E(τi)j≤nE(τ(2)n)jn1+(2j−1)α=E(τ(2)n)jn(2j−1)α→(2j−3)!!(4B2)2j−12,asn→∞. |
Hence, L(j)≤U(j)≤(2j−3)!!(4B2)2j−12.
By Lemma 3.3, we have −∞<Λ_(λ)≤¯Λ(λ)<∞ for |λ|≤2B2.
Proposition 3.1. For j=1,2, L(j)=U(j).
Proof. For j=1, we have
limn→∞n∑i=1Eτin1+α=limn→∞ETnn1+α=12B(1+α), |
so, L(1)=U(1)=12B(1+α).
For j=2, we have
limn→∞n∑i=1E(τi)2n1+3α=limn→∞VarTn+n∑i=1(Eτi)2n1+3α=14B3(1+3α). | (3.4) |
By Proposition 2.1 [6],
limn→∞Eτnnα=12B. |
Hence,
limn→∞n∑i=1(Eτi)2n1+2α=14B2(1+2α). | (3.5) |
According to Theorem1.2 [6],
limn→∞Var(Tn)n1+3α=14B3(1+3α). | (3.6) |
The combination of (3.4, 3.5, and 3.6) yields:
limn→∞n∑i=1E(τi)2n1+3α=14B3(1+3α). |
Hence, L(2)=U(2)=14B3(1+3α).
Property 3.2. ¯Λ′(λ) and Λ_′(λ) exist for |λ|≤2B2.
Proof. By Lemma 3.4, ¯Λ(λ) and Λ_(λ) can be expressed as power series when |λ|≤2B2. According to the properties of power series, it follows that ¯Λ′(λ) and Λ_′(λ) exist for |λ|≤2B2.
Proposition 3.2. For every λ∈R, ¯Λ(λ)≥Λ_(λ)>−∞.
Proof. Since for every n≥0, Tn is non-negative, we have Λ_(λ)>−∞ for every λ∈R+. Let DΛ_△={λ:Λ_(λ)>−∞}, and let DoΛ_ be the interior of DΛ_. By Lemma 3.4, we know that 0∈DoΛ_, so there exists λ0<0 such that Λ_(λ0)>−∞. Hence, for every λ<λ0, by Jensen's inequality we have the following:
Eexp(λn1−αTnn1+α)=E[exp(λ0n1−αTnn1+α)]λλ0≥[Eexp(λ0n1−αTnn1+α)]λλ0. |
So,
Λ_(λ)=lim infn→∞1n1−αlogEexp(λn1−αTnn1+α)≥λλ0lim infn→∞1n1−αlogEexp(λ0n1−αTnn1+α)=λλ0Λ_(λ0)>−∞. |
Hence, we have ¯Λ(λ)≥Λ_(λ)>−∞ for every λ∈R.
The proof of Property 3.1
By Proposition 3.2, it follows that ¯Λ∗(x) and Λ_∗(x) are always defined. Properties (2) and (3) follow from the results in [2]. Specifically, ¯x=12B(1+α) is the limiting value of Tnn1+α as stated in Theorem C. The idea of the proof of (4) comes from Lemma2.2.5 in [5]. For all λ∈R, by Jensen's inequality,
¯Λ(λ)=lim supn→∞1n1−αlogEexp(λn1−αTnn1+α)≥lim supn→∞1n1−αElog[exp(λn1−αTnn1+α)]=lim supn→∞1n1−αE(λn1−αTnn1+α)=λlim supn→∞ETnn1+α=λ¯x, |
Since ¯Λ(0)=0, we obtain ¯Λ∗(¯x)=sup{λ∈R}{λ¯x−¯Λ(λ)}=0. Similarly, we also have Λ_∗(¯x)=0.
Proposition 3.3. For all x≥¯x
¯Λ∗(x)=sup{λ≥0}{λx−¯Λ(λ)},Λ_∗(x)=sup{λ≥0}{λx−Λ_(λ)} |
is a non-decreasing function. Similarly, for all x≤¯x,
¯Λ∗(x)=sup{λ≤0}{λx−¯Λ(λ)},Λ_∗(x)=sup{λ≤0}{λx−Λ_(λ)} |
is a non-increasing function.
Proof. For every x≥¯x and every λ<0,
λx−¯Λ(λ)≤λ¯x−¯Λ(λ)≤¯Λ∗(¯x)=0, |
so ¯Λ∗(x)=sup{λ≥0}{λx−¯Λ(λ)}. This also implies the monotonicity of ¯Λ∗(x) on (¯x,∞), since for every λ≥0, the function λx−¯Λ(λ) is nondecreasing as a function of x.
For every x≤¯x and every λ>0,
λx−¯Λ(λ)≤λ¯x−¯Λ(λ)≤¯Λ∗(¯x)=0, |
so ¯Λ∗(x)=sup{λ≤0}{λx−¯Λ(λ)}. This also implies the monotonicity of ¯Λ∗(x) on (−∞,¯x), since for every λ≤0, the function λx−¯Λ(λ) is nonincreasing as a function of x.
Similarly, we can also obtain analogous properties for Λ_∗(x).
Proof of Theorem 2.1
Let kn be the unique (random) integers such that
Tkn≤n≤Tkn+1, | (4.1) |
Since Tkn represents the time when the random walk first reaches position kn, and Tkn≤n, the maximum position Mn must satisfy Mn≥kn. Similarly, because n≤Tkn+1 and the definition of Tkn+1, it follows that Mn≤kn+1. Therefore, we conclude kn≤Mn≤kn+1. Hence,
knn11+α≤Mnn11+α≤kn+1n11+α. |
Note that Theorem C states that limn→∞Tkn(kn)1+α=limn→∞Tkn+1(kn)1+α=12B(1+α). As a consequence, dividing both sides of inequality (4.1) by (kn)1+α yields limn→∞n(kn)1+α=12B(1+α). Thus,
[2B(1+α)]1/(1+α)≥lim supn→∞Mnn11+α≥lim infn→∞Mnn11+α≥[2B(1+α)]1/(1+α). |
The proof is completed.
Proof of Theorem 2.2
The idea originates from the proof of Cramér's Theorem in [5] and Theorem 3 in [10]. Let F be a non-empty closed set. Note that (2.3) holds trivially if
infx∈F¯Λ∗(x)=0. |
Assume instead that
infx∈F¯Λ∗(x)>0. |
Since ¯Λ∗(¯x)=0 (see Property 3.1), ¯x must lie in the open set Fc. Let (x−,x+) denote the union of all open intervals (a,b)⊂Fc containing ¯x. Observe that x−<x+ and at least one of x− or x+ must be finite (since F is non-empty).
(1) If x− is finite and x+=+∞, then x−∈F⊂(−∞,x−). Consequently,
¯Λ∗(x−)≥infx∈F¯Λ∗(x). |
For every λ≤0,
P(Tnn1+α∈F)≤P(Tnn1+α∈(−∞,x−])=P(Tnn1+α−x−≤0)=E[ITnn1+α−x−≤0]≤E[exp{n1−αλ(Tnn1+α−x−)}]=exp{−n1−αλx−}Eexp{n1−αλTnn1+α}. |
Observe that the random variable exp{n1−αλ(Tnn1+α−x−)}≥1 on the set {Tnn1+α−x−≤0},forλ≤0, we used Chebyshev's inequality in the last inequality above. Next, by taking the logarithm on both sides of the above inequality and considering the upper limit with proper scaling, we obtain
lim supn→∞1n1−αlogP{Tnn1+α∈F}≤lim supn→∞1n1−αlogP(Tnn1+α∈(−∞,x−])≤lim supn→∞1n1−αlog{exp{−n1−αλx−}Eexp{n1−αλTnn1+α}}=−λx−+lim supn→∞1n1−αlogEexp{n1−αλTnn1+α}=−(λx−−¯Λ(λ)). | (4.2) |
The above inequality holds for all λ≤0; consequently,
lim supn→∞1n1−αlogP{Tnn1+α∈F}≤lim supn→∞1n1−αlogP(Tnn1+α∈(−∞,x−])≤infλ≤0{−(λx−−¯Λ(λ))}=−supλ≤0{λx−−¯Λ(λ)}=−¯Λ∗(x−)≤−infx∈F¯Λ∗(x), | (4.3) |
where the last equality follows from Proposition 3.3. The final inequality holds trivially since x−∈F.
By a similar argument, if x+ is finite and x−=−∞, then
lim supn→∞1n1−αlogP{Tnn1+α∈F}≤lim supn→∞1n1−αlogP(Tnn1+α∈[x+,∞))≤−¯Λ∗(x+)≤−infx∈F¯Λ∗(x) | (4.4) |
(2) If x−,x+ are all finite, then x−,x+∈F, and x−<¯x<x+.
1n1−αlogP{Tnn1+α∈F}≤1n1−αlogP{Tnn1+α∈(−∞,x−]∪[x+,∞)}≤1n1−αlogmax{2P(Tnn1+α≤x−),2P(Tnn1+α≥x+)}≤log2n1−α+max{1n1−αlogP(Tnn1+α≤x−),1n1−αlogP(Tnn1+α≥x+)}, |
Combining (4.3) with (4.4), we obtain
lim supn→∞1n1−αlogP{Tnn1+α∈F}≤max{lim supn→∞1n1−αlogP(Tnn1+α≤x−),lim supn→∞1n1−αlogP(Tnn1+α≥x+)}≤−infx∈F¯Λ∗(x). | (4.5) |
In summary, we have completed the proof of (2.3). The proof of (2.4) follows by taking the limit inferior in (4.2), (4.3), (4.4) and (4.5).
The large deviations lower bound of Tn (statements (2) and (3) in Theorem 2.2) follow directly from Theorem3.5-3.6 in [2] with h(x)=x; therefore, we omit the details here.
Proof of Theorem 2.3
Let v0=(2B(1+α))11+α, which is the limit value of Mn/n11+α in Theorem 2.1. For every v>v0, i.e. 1v1+α<1v1+α0=12B(1+α)=¯x, note that
P(Mnn11+α≥v)=P(Mn≥n11+αv)≤P(T⌊n11+αv⌋≤n)=P(T⌊n11+αv⌋(⌊n11+αv⌋)1+α≤n(⌊n11+αv⌋)1+α). |
The event Mn≥n11+αv implies that the random walk has reached n11+αv before time n, and thus
{Mn≥n11+αv}⊂{T⌊n11+αv⌋≤n}, |
where ⌊x⌋ denotes the greatest integer less than or equal to x, and ⌈x⌉ denotes the smallest integer greater than or equal to x. Taking logarithms on both sides of the inequality and taking the upper limit with proper scaling yields
lim supn→∞1(⌊n11+αv⌋)1−αlogP(Mnn11+α≥v)≤lim supn→∞1(⌊n11+αv⌋)1−αlogP(T⌊n11+αv⌋(⌊n11+αv⌋)1+α≤n(⌊n11+αv⌋)1+α)≤−¯Λ∗(1v1+α), |
where the last inequality follows from (4.3), noting that limn→∞n(⌊n11+αv⌋)1+α=1v1+α and the continuity of −¯Λ∗(x) (see Property 3.1). Multiplying both sides by v1−α gives
lim supn→∞1n1−α1+αlogP(Mnn11+α≥v)≤−v1−α¯Λ∗(1v1+α)=−¯I(v). |
Next, we derive an upper bound for P(Mn/n11+α≤v), where v<v0 (i.e., 1v1+α>1v1+α0=¯x). Observe that
P(Mn1n1+α≤v)≤P(Mn≤n11+αv)≤P(T⌈n11+αv⌉≥n), |
Thus, (4.4) implies
lim supn→∞1⌈n11+αv⌉1−αlogP(Mnn11+α≤v)≤lim supn→∞1⌈n11+αv⌉1−αlog[P(T⌈n11+αv⌉⌈n11+αv⌉1+α≥n⌈n11+αv⌉1+α)]≤−¯Λ∗(1v1+α) |
Consequently,
lim supn→∞1n1−α1+αlogP(Mnn11+α≤v)≤−v1−α¯Λ∗(1v1+α)=−¯I(v). | (4.6) |
The remainder of the proof follows similarly to the large deviations upper bound for Tn (Theorem 2.2). We have completed the proof of part (1) in Theorem 2.3.
Next, we consider the large deviation lower bound for Mn. For v<v0 (i.e., 1v1+α>¯x) satisfying 1v1+α∈¯Go, there exists a neighborhood (1v1+α−δ,1v1+α+δ)⊂¯Go. For any 0<ϵ<δ/2, we have
P(Mnn11+α<v)=P(Mn<n11+αv)≥P(T⌊n11+αv⌋>n)≥P(T⌊n11+αv⌋⌊n11+αv⌋1+α>n⌊n11+αv⌋1+α)≥P(T⌊n11+αv⌋⌊n11+αv⌋1+α>1v1+α+ϵ) |
for sufficiently large n. By Theorem 2.2, we obtain
lim supn→∞1(⌊n11+αv⌋)1−αlogP(Mnn11+α<v)≥lim supn→∞1(⌊n11+αv⌋)1−αlogP(T⌊n11+αv⌋⌊n11+αv⌋1+α>1v1+α+ϵ)≥−infx∈(1v1+α+ϵ,1v1+α+2ϵ)¯Λ∗(x)≥−¯Λ∗(1v1+α+32ϵ) |
Since 1v1+α∈¯Go, ¯Λ∗(1v1+α)<∞, and ¯Λ∗(x) is continuous at 1v1+α (see (3) in Property 3.1), letting ϵ→0 yields
lim supn→∞1n1−α1+αlogP(Mnn11+α<v)≥−v1−α¯Λ∗(1v1+α)=−¯I(v). |
Combining this with (4.6), we conclude
lim supn→∞1n1−α1+αlogP(Mnn11+α<v)=lim supn→∞1n1−α1+αlogP(Mnn11+α≤v)=−¯I(v). | (4.7) |
Similarly, for v<v0, satisfying 1v1+α∈G_o, we have
lim infn→∞1n1−α1+αlogP(Mnn11+α<v)=lim infn→∞1n1−α1+αlogP(Mnn11+α≤v)=−I_(v) | (4.8) |
Hence, for any v with 1v1+α∈¯Go∩G_o, there exists a neighborhood (1v1+α−δ,1v1+α+δ)⊂¯Go∩G_o. Assume v<v0 (the case v>v0 is analogous). Choose δ1<δ and δ2<δ, combining (4.7) with (4.8), we obtain
−I_(v+δ2)=lim infn→∞1n1−α1+αlogP(Mnn11+α<v+δ2)>lim supn→∞1n1−α1+αlogP(Mnn11+α≤v−δ1)=−¯I(v−δ1). | (4.9) |
Hence, for any ϵ′>0 and sufficiently large n,
P(Mnn11+α<v+δ2)≥exp{−n1−α1+α(I_(v+δ2)+ϵ′)} |
P(Mnn11+α≤v−δ1)≤exp{−n1−α1+α(¯I(v−δ1)−ϵ′)} |
Thus,
P(v+δ2>Mnn11+α>v−δ1)=P(Mnn11+α<v+δ2)−P(Mnn11+α≤v−δ1)≥exp{−n1−α1+α(I_(v+δ2)+ϵ′)}−exp{−n1−α1+α(¯I(v−δ1)−ϵ′)}=exp{−n1−α1+α(I_(v+δ2)+ϵ′)}{1−exp{−n1−α1+α(¯I(v−δ1)−I_(v+δ2)−2ϵ′)}} |
Since ¯I(v−δ1)>I_(v+δ2) (by (4.9)); choosing ϵ′,δ1, and δ2 such that ¯I(v−δ1)−I_(v+δ2)−2ϵ′>0 yields
lim infn→∞1n1−α1+αlogP(v+δ2>Mnn11+α>v−δ1)≥−I_(v+δ2)−ϵ′. |
By the continuity I_(v), for any ϵ>0, we can choose δ2 sufficiently small such that I_(v+δ2)<I_(v)+ϵ. Consequently,
lim infn→∞1n1−α1+αlogP(v+δ2>Mnn11+α>v−δ1)≥−I_(v)−ϵ−ϵ′. |
Since ϵ and ϵ′ are arbitrary, we conclude
lim infn→∞1n1−α1+αlogP(v+δ2>Mnn11+α>v−δ1)≥−I_(v). |
In recent years, random walks with asymptotic perturbations in transition probabilities have received widespread attention from scholars. These perturbations bring many new phenomena to random walks. For transient near-critical random walks, Voit [13] established the law of large numbers for the random walks, showing that the escape velocity of such processes is significantly slower than that of simple random walks. Building on this foundation, we derive large deviation principles for such random walks and demonstrate that their velocity order is substantially sublinear in n. This result further indicates that asymptotic perturbations reduce the wandering speed of random walks.
The author declares she has not used Artificial Intelligence (AI) tools in the creation of this article.
The author thanks the anonymous referees for their meticulous reading of the manuscript. This work was partly supported by the National Natural Science Foundation of China (Grant Nos. 12001558, 71971228, 72371261).
The author declares no conflicts of interest in this paper.
[1] |
H. Brezis, W. Rosenkrantz, B. Singer, An extension of khintchine's estimate for large deviations to a class of markov chains converging to a singular diffusion, B. Am. Math. Soc., 77 (1971) 980–982. http://dx.doi.org/10.1002/cpa.3160240506 doi: 10.1002/cpa.3160240506
![]() |
[2] |
P. N. Chen, Generalization of gartner-ellis theorem, IEEE T. Inform. Theory, 46 (2000), 2752–2760. http://dx.doi.org/10.1109/18.887893 doi: 10.1109/18.887893
![]() |
[3] | K. L. Chung, Markov chains with stationary transition probabilities, 2 Eds., New York: Springer, 1967. http://dx.doi.org/10.1002/bimj.19660080419 |
[4] |
E. Csaki, A. Foldes, P. Revesz, Transient nearest neighbor random walk on the line, J. Theor. Prob., 22 (2009), 100–122. http://dx.doi.org/10.1007/s10959-007-0137-3 doi: 10.1007/s10959-007-0137-3
![]() |
[5] | A. Dembo, O. Zeitouni, Large deviations techniques and applications, 2 Eds., Springer, 2009. http://dx.doi.org/10.1007/978-3-642-03311-7 |
[6] |
W. M. Hong, H. Yang, Cutoff phenomenon for nearest Lamperti's random walk, Methodology and Computing in Applied Probability, 21 (2019), 1215–1228. http://dx.doi.org/10.1007/s11009-018-9666-8 doi: 10.1007/s11009-018-9666-8
![]() |
[7] |
J. Lamperti, Criteria for the recurrence or transience of stochastic process. Ⅰ, J. Math. Anal. Appl., 1 (1960), 314–330. http://dx.doi.org/10.1016/0022-247X(60)90005-6 doi: 10.1016/0022-247X(60)90005-6
![]() |
[8] |
J. Lamperti, A new class of probability limit theorems, B. Am. Math. Soc., 67 (1961), 267–269. http://dx.doi.org/10.1090/S0002-9904-1961-10575-2 doi: 10.1090/S0002-9904-1961-10575-2
![]() |
[9] |
J. Lamperti, Criteria for stochastic processes Ⅱ: Passage-time moments, J. Math. Anal. Appl., 7 (1963), 127–145. http://dx.doi.org/10.1016/0022-247X(63)90083-0 doi: 10.1016/0022-247X(63)90083-0
![]() |
[10] |
P. W. Glynn, W. Whitt, Large deviations behavior of counting processes and their inverses, Queueing Syst., 17 (1994), 107–-128. http://dx.doi.org/10.1007/BF01158691 doi: 10.1007/BF01158691
![]() |
[11] |
P. Seignourel, Discrete schemes for processes in random media, Probab. Theory Rel., 118 (2000), 293–322. http://dx.doi.org/10.1007/PL00008743 doi: 10.1007/PL00008743
![]() |
[12] |
M. Voit, A law of the iterated logarithm for a class of polynomial hypergroups, Monatsh. Math., 109 (1990), 311–326. http://dx.doi.org/10.1007/BF01320696 doi: 10.1007/BF01320696
![]() |
[13] |
M. Voit, Strong laws of large numbers for random walks associated with a class of one-dimensional convolution structures, Monatsh. Math., 113 (1992), 59–74. http://dx.doi.org/10.1007/BF01299306 doi: 10.1007/BF01299306
![]() |
[14] |
K. Zhou, Hitting time distribution for skip-free markov chains: a simple proof, Stat. Probabil. Lett., 83 (2013), 1782–1786. http://dx.doi.org/10.1016/j.spl.2013.04.008 doi: 10.1016/j.spl.2013.04.008
![]() |