Limit Theorems for low dimensional generalized transformations
1University of Maryland, College Park, MD, USA
2Chern Institute of Mathematics and LPMC, Nankai University, Tianjin, China
3Yeshiva University, New York, NY, USA
4Jagiellonian University, Krakow, Poland
)
Abstract.
We consider generalized transformations such that the base map satisfies a multiple mixing local limit theorem and anticoncentration large deviation bounds and in the fiber we have actions with or which are exponentially mixing of all orders. If the skewing cocycle has zero drift, we show that the ergodic sums satisfy the same limit theorems as the random walks in random scenery studied by Kesten and Spitzer (1979) and Bolthausen (1989). The proofs rely on the quenched CLT for the fiber action and the control of the quenched variance. This paper complements our previous work where the classical central limit theorem is obtained for a large class of generalized transformations.
Key words and phrases:
transformations; Kesten–Spitzer process; quenched and annealed limit theorems; local time1991 Mathematics Subject Classification:
Primary: 60F05. Secondary: 37D30, 60J55, 60K351. Introduction
This work is a continuation of our study on generalized transformations, following previous work [11, 12, 13]. Our main innovation in this paper is to provide several limit theorems in the low dimensional setting, complementing to the higher dimensional case in [12].
1.1. Results
Let be a smooth map of a manifold preserving a measure and be an action on a manifold preserving a measure
Definition 1.1.
enjoys multiple exponential mixing of all orders if there is such that for each there are constants such that for all zero mean functions , for all
(1.1) |
where is the gap
Remark 1.2.
In case , there are plenty of examples of multiply exponentially mixing systems see e.g. the discussion in [11]. In the case our main example is the following: , where is a cocompact lattice, is the Cartan action on , and is the Haar measure. More generally one can consider subactions of Cartan actions on where is a semisimple Lie group with compact factors and is a cocompact lattice. In particular, we can take , and let be an action by a two dimensional subgroup of the group of diagonal matrices.
Let be a smooth map. We study the map given by
(1.2) |
Note that preserves the measure and that
Let be a sufficiently smooth function with and
We want to study the distribution of when the initial condition is distributed according to
We shall assume that is ergodic and satisfies the CLT for Hölder functions.
Definition 1.3.
We say that satisfies the mixing local limit theorem (MLLT) if for any sequences , with and such that for any cube and any continuous functions
where is a non-degenerate Gaussian density and the convergence is uniform once is fixed and range over compact subsets of and .
Definition 1.4.
satisfies multiple mixing local limit theorem (MMLLT) if for each for any sequence , with , for any family of sequences with for any cubes and any continuous functions for any sequences such that (with ),
where Moreover, the convergence is uniform once is fixed, range over compact subsets of and range over a compact subset of for every .
Definition 1.5.
satisfies the anticoncentration large deviation bound of order if there exist a constant and a decreasing function such that and for any numbers , for any unit cubes centered at
(1.3) |
The class of maps satisfying the MMLLT and the large deviation anticoncentration bounds includes in particular the maps which admit Young towers with exponential tails, see [21, 19].
Theorem 1.6.
Suppose that , enjoys multiple exponential mixing of all orders, satisfies the CLT for smooth observables and satisfies the MMLLT and the anticoncentration large deviation bound. Then there is such that converges as to the normal distribution with zero mean and variance
Theorem 1.7.
Suppose that , enjoys multiple exponential mixing of all orders, satisfies the CLT for smooth observables and satisfies the MMLLT and the anticoncentration large deviation bound. Then there is a constant such that
(1.4) |
where and are independent, has standard normal distribution and
(1.5) |
where is the local time of the standard Brownian Motion at time and spatial location
Remark 1.8.
If then the exponential mixing condition can be weakened to sufficiently fast polynomial mixing, see Theorem 5.1 in Section 5. It seems more difficult to weaken the mixing assumption in two dimension. We do not pursue this topic since we do not know examples of smooth actions which are mixing at a fast polynomial (rather than exponential) speed.
Remark 1.9.
The asymptotic variance in Theorems 1.6 and 1.7 has similar form (see (3.10) and (4.5)). Namely let Then
In dimension if then there is an function such that
This was shown in the discrete case by Rudolph, [20] (Proposition 2 and Lemma 7), but the proof applies with no changes in the continuous case.
It is shown in [12, Theorem 6.4] that for volume preserving exponentially mixing flows, the quadratic form is not identically zero. Therefore the set of functions of vanishing asymptotic variance is a proper linear subspace. The proof of Theorem 6.4 in [12] works also for higher dimensional (not necessarily volume preserving actions) provided that
(i) there are constants such that for all and , and
(ii) there exists a slowly recurrent point, that is, a point such that for all positive constants and there exists so that for all and for all with , , where is the ball of radius centered at . (We note that for exponentially mixing volume preserving actions, almost all points are slowly recurrent, see [14].)
We believe that in many cases the vanishing of the asymptotic variance entails that the ergodic sums of satisfy the classical CLT but this will be a subject of a future work.
Remark 1.10.
Results analogous to the above theorems can be proved in case is an action of () and is a piecewise smooth map satisfying the appropriate assumptions such as the CLT, MMLLT, and anticoncentration large deviations bounds. Since the results as well as the proofs are virtually the same, we omit the case of discrete actions. One can also take to be a subshift of finite type, see the discussion below.
1.2. Discussion
The first results about transformations pertain to so called random walks in random scenery. In this model we are given a sequence of i.i.d. random variables. Let be a simple random walk on independent of s. We are interested in This model could be put in the present framework as follows. Let be a set of sequences where where are basis vectors in is the Bernoulli measure with for all is the space of sequences , is the product of distribution functions of , and are shifts and For random walks in random scenery, Theorem 1.7 is due to [15], and Theorem 1.6 are due to [4]. The results of [15] and [4] are extended to more general actions in the fiber (still assuming the random walk in the base) in [7, 8].
In the context of dynamical systems, Theorem 1.7 was proven in [17] under the assumption that we have a good rate of convergence in the Central Limit Theorem for . In the present paper we follow a method of [4] which seems more flexible and allows a larger class of base systems.
We also note that if or has a drift, one has a classical CLT (that is, converges to a Gaussian distributions). These results are proven in [11, 12]. The paper [11] also studies mixing properties of transformation defined by (1.2). In particular [11] shows that in many cases the mixing of the whole product is exponential. We observe that in the case where enjoys multiple exponential mixing, one can obtain the classical CLT by applying the results of [3]. The CLT of Björklund and Gorodnik [3] also plays a key role in our proof and we review it in Section 2.
In the case where the base map is uniformly hyperbolic the skew product map is partially hyperbolic. It is possible that generic partially hyperbolic maps enjoy strong statistical properties including the multiple exponential mixing and the Central Limit Thereom. However, such a result seems currently beyond reach (some special cases are considered in [1, 6]). In present setting the exotic limit theorems such as our Theorems 1.7 and 1.6 require the zero drift assumption, which is a codimension condition.
2. Björklund - Gorodnik CLT
In order to prove our results we use the strategy of [4] replacing the Feller Lindenberg CLT for iid random variables by a CLT for exponentially mixing systems due to [3]. More precisely we need the following fact.
Proposition 2.1.
Let , be a sequence of measures on and let be a family of real valued functions on so that is uniformly bounded and . Set Suppose that
(a) where
(b) For each for each
where is the ball of radius around .
(c) where
Then, as , converges weakly to the normal distribution with zero mean and variance
3. Dimension two
3.1. Reduction to quenched CLT
Here we prove Theorem 1.6.
Consider a function satisfying
(3.1) |
for all
Given define the measures on by
(3.2) |
and functions by
(3.3) |
Proposition 3.1.
The proposition will be proven later. Now we shall show how to obtain Theorem 1.6 from the proposition.
The remaining part of this section is devoted to the proof of Proposition 3.1. It suffices to verify the conditions of Proposition 2.1.
Property (a) is clear, since
3.2. Property (b)
Let
(3.5) |
Lemma 3.2.
If for each , , then there are sets such that for all the measures satisfy property (b) and
Proof.
Given let By the assumption of the lemma, On the other hand, for ,
where the first inequality holds since and the second one holds because . Since is arbitrary, the result follows.
Let
(3.6) |
Lemma 3.3.
For each there is a constant such that for each and ,
(3.7) |
Proof.
Since we have
where the last step uses the anticoncentration large deviation bound.
3.3. Property (c)
Theorem 4.7 in [11] implies that
(3.8) |
where
(3.9) |
and is the limit Gaussian density of (that is, converges as to the normal distribution with density ). We note that the integral (3.9) converges by the exponential mixing of and (3.1). Furthermore, , which can be seen from the following formula (whose proof is standard and so is omitted)
By Chebyshev’s inequality, to establish property (c), it suffices to show that
(3.11) |
Note that
where
Fix a large and let be the number of indices such that for all The number of terms where is . Using the trivial bound , we see that the contribution of terms with is which is negligible.
Next, we consider the contribution of the terms with . Without loss of generality, we will assume that
(3.12) |
We distniguish cases based on the relative position of with respect to and
Case 1. We claim that for each there is such that if
(3.13) |
then
(3.14) |
The proof of (3.14) is based on [11]. First, by [11, Lemma 3.3], we can assume that and without loss of generality we assume = 0. Then we have
where . Note that for as above (3.11) simplifies to
(3.15) |
The remaining part of the proof of (3.14) closely follows the lines of [11, Theorems 4.6, 4.7] and so we only give a sketch. Decompose into a countable disjoint family of small squares . Let be the center of square . Then
where
Fixing , letting and using the MMLLT and (3.13), we find
Case 2. We claim that
(3.16) |
To prove (3.16) take as in (3.16) and write
Then we have
Decompose into unit boxes with center . Then, we find
where Applying (1.3), we find
(3.17) | ||||
Let us consider the special case when and (without loss of generality we say that in this case ). Then we have
(3.18) |
Note that the last expression is symmetric in and . Thus without loss of generality we can assume (indeed, the multiplier can be incorporated into the constant ). Denoting , , , we obtain
Note that the multiplier accounts for all choices of . Thus
where in the third inequality we used that and the last inequality follows from direct integration. Thus we have verified that the terms corresponding to and are in .
To complete the proof of (3.16), we need to estimate the contribution of the sum corresponding to all and . To this end, we will distinguish two cases based on whether the following assumption holds or not:
(3.19) |
For any given fixed , let denote the sum over that do not satisfy (3.19) and let denote the sum for satisfying (3.19).
First, we assume that (3.19) fails. Then, we have
Now we can perform the summation with respect to first. Denote , , . By the exponential decay of we get
Summing the last displayed formula for , we obtain a term exactly as in the estimate of (3.18).
Now we assume that (3.19) holds. Then observing that and using the exponential decay of , we find that
with a fixed constant . Thus we conclude
Summing over all choices of , we obtain a
term in .
This completes the proof of
(3.16).
Case 3. We claim that
(3.20) |
The proof of (3.20) is similar to that of (3.16). Indeed, the right hand side of (3.17) is now replaced by
We consider the special case first. Then (3.18) is replaced by
(3.21) |
which is bounded by the right hand side of (3.18) (perhaps with a different constant ) and hence is in . The contribution of general , can be bounded as before. (3.20) follows.
The upshot of the above three cases is that the leading term only comes from indices satisfying (3.13). Summing for all indices satisfying (3.13), we obtain where we have used (3.14) and the fact that
Next we claim that
(3.22) |
Indeed similarly to the case of (3.13) considered above we see that the main contribution to our sums comes from the terms where the pairs and are not intertwined. Note that if we choose a random permutation of then the probability of non intertwining is 1/3 (since the second element should come from the same pair as the first one). Therefore there are 24/3=8 permutations 111These eight permutations corresponds to all possible choices of signs () in the inequalities contributing to our sum which explains the additional factor of 8 in (3.22).
To complete the proof of property (c), it remains to verify that the contribution of terms with is negligible. First note that is impossible by the definition of . Finally, if , the we can repeat a simplified version of the proof for using the MLLT instead of the MMLLT. For example, let us consider the case . Since , we must have or . Without loss of generality assume . Then, using the fact that due to the exponential mixing of , on the set , we obtain from the MLLT that
Hence the total contribution terms with is in . This completes the proof of property (c). We have finished the proof of Proposition 3.1 and hence the proof of Theorem 1.6.
4. Dimension one
4.1. Reduction to the limit theorem for the variance.
Here we prove Theorem 1.7.
We use the decomposition (3.4). Since satisfies the CLT for smooth observables, converges weakly, and so this sum is negligible after rescaling by . It remains to study the ergodic sum of .
Following the ideas of the previous section, we let
(4.1) |
where
Thus in the notation of Proposition 2.1, we have
In contrast with Theorem 1.6, does not satisfy a weak law of large numbers. Instead we have
Proposition 4.1.
There is a constant so that converges in law as to the random variable given by (1.5).
Before proving Proposition 4.1, let us use it to derive Theorem 1.7. We start with the following analogue of Proposition 3.1:
Lemma 4.2.
Proof of Lemma 4.2 assuming Proposition 4.1.
Clearly, for a large set of ’s and so (a) holds. Recalling that , property (c) holds trivially for all . It remains to show property (b).
Note that is non-negative and non-atomic at zero, i.e. for any there is so that and so by Proposition 4.1,
Thus we can assume that is such that . Next, we define
As in the proof of Lemma 3.3, we find that for every , every , and every we have . Then choosing , the Markov inequality gives . Thus we can assume that for all with , holds. By the anticoncentration large deviation bounds, we can also assume that since the set of points where this condition fails has negligible measure. In summary, the measure of the set of ’s satisfying that
tends to as For any such , we have
if . Property (b) and the lemma follow.
Proof of Theorem 1.7 assuming Proposition 4.1.
Recall that By Lemma 4.2 and Proposition 2.1, there is a sequence of subsets with such that for every sequence , the distribution of w.r.t. converges to the standard normal (note that this time properties (a) and (c) are immediate).
In fact, we have the stronger statement that
(4.2) |
where is given by (1.5), is standard normal and and are independent. This follows from Proposition 4.1, Lemma 4.2 and the asymptotic independence of and which comes from the fact that depends only on while is asymptotically independent of . More precisely, let and be two continuous compactly supported test functions. Then
where the first and the fourth equalities hold since the third inequality holds by Proposition 2.1, and the last equality holds by Proposition 4.1. This proves (4.2).
4.2. Proof of Proposition 4.1
We are going to use the following well known fact (see e.g. [2] Chapter 1.7, Problem 4). If is a sequence of random variables so that for every
(4.3) |
exists and
(4.4) |
then converges weakly to a random variable . Furthermore, for every and is uniquely characterized by its moments.
Now we explain our strategy of proof of Proposition 4.1 (a similar strategy was used in [18]). We prove that there is a constant depending on the transformation and the observable so that satisfies (4.3) (with meaning integral w.r.t. ) and (4.4) with constants that do not depend on and . Consequently, there is a random variable so that converges weakly to for any transformation satisfying the assumptions of Theorem 1.7. Recall from §1.2 that one such example is the one dimensional random walk in random scenery, in which case by the result of [15], converges weakly to . Thus and it has to be the limit for all transformation satisfying the assumptions of Theorem 1.7. It remains to check (4.3) and (4.4).
Recall that satisfies the MMLLT with a Gaussian density . Let be the standard deviation of this Gaussian random variable, that is , where is the standard Gaussian density.
Lemma 4.3.
For , (4.3) holds with
(4.6) |
We note that the above integral can be performed explicitly. Namely . This shows that was chosen correctly. Indeed, in case of the simple random walk in random scenery, formula (1.2) in [15] shows that while by the main result of [15], the conclusion of Proposition 4.1 holds for simple random walk in random scenery.
Proof of Lemma 4.3.
Fix some . We need to show that for large enough,
In the following proof, we will choose a small , a large a small and finally . There will be finitely many restrictions on these parameters, we can choose the most strict one.
Choose a partition of into sets of diameter and fix some elements . We have
(4.7) |
where
Split the sum in (4.7) as , where corresponds to the terms satisfying
-
(A1)
-
(A2)
-
(A3)
.
and stands for the terms where at least one of the conditions (A1)–(A3) is violated. Write .
We start by estimating . Let us write if there are constants , and so that for large enough, . We claim that
(4.8) |
Indeed, by the continuity of we can choose so small that the difference between the LHS and the RHS of (4.8) does not exceed
so (4.8) follows from the anticoncentration large deviation bound.
Next, writing , and using the MMLLT, we find
(4.9) |
Noting that by (A1) and (A3), , we see that can be replaced by in (4.9) without invalidating . Then the only term depending on is under the integral with respect to . We can approximate the sum over by a Riemann integral:
(4.10) |
where
Combining (4.9) and (4.10), we conclude
where the factor comes from taking into account both and . Further decreasing and increasing if necessary and recalling (3.9), we can replace the first sum by .
Now replacing two Riemann sums with the corresponding Riemann integrals, we obtain
In order to complete the proof of (4.6), it suffices to show that for large and for small
(4.11) |
We will estimate the contribution of the terms where exactly one of the assumptions (A1)–(A3) fail (the cases when more than one fail are similar and easier).
Suppose that (A1) fails and, moreover, (the other cases are similar). Using the anticoncentration large deviation bounds instead of the MMLLT, we obtain that the corresponding sum is bounded by
if is small enough.
Let us assume that (A2) fails. Then again using the anticoncentration large deviation bounds instead of the MMLLT, we obtain that the corresponding sum is bounded by
for large.
Finally assume that (A3) fails. Then we use the exponential mixing of to derive
Proceeding as in the case of we obtain that the sum corresponding to indices when (A3) is violated is bounded by . This completes the proof of the lemma.
Lemma 4.4.
For any , (4.3) holds with
(4.12) |
where is the set of all two-to-one mappings from to (that is, for all ).
Proof of Lemma 4.4.
We follow the strategy of the proof of Lemma 4.3.
Fix some . We need to show that for large enough,
Now we have
where
Let us order the numbers as and denote . As before, we write the sum in (4.2) as , where corresponds to the terms satisfying
-
(A1’)
for every ,
-
(A2’)
for all
-
(A3’)
for all
and write . As in (4.8),
(4.14) |
where now means that the difference between the LHS and the RHS an be made smaller than provided that and are small enough and and are large enough. To compute , we use the MMLLT for times . Note however that the range of summation for different ’s are different and so it is important to keep track of the index so that . To this end, define to permutation (uniquely defined by the tuple if (A1’) holds) so that . Writing and and using the MMLLT we obtain
(4.16) | |||||
By (A1’) and (A3’), can be replaced by (the biggest odd integer not bigger than ) for in the subscripts of in (4.16). Consequently, only appears in (4.16). As before, we compute
Thus we arrive at
where refers to the sum for , , satisfying (A1’), (A3’). Next, observe that defined by is a two-to-one mapping. Let be the set of all such mappings. The summand in the last displayed formula only depends on through and . Furthermore, for any given and , there are exactly corresponding tuples . Thus
Now we are ready to replace the last two sums (Riemann sums) by the corresponding Riemann integral with and . To simplify the notation a little, we denote , thus is a 2-to-1 mapping from to (the set of all such mapping is denoted by ). We obtain
Substituting to in the above integrals, we obtain
As in Lemma 4.3, is negligible. This completes the proof of Lemma 4.4.
To finish the proof of Proposition 4.1, it remains to verify (4.4). To this end, first note that since decays quickly, there is a constant so that for every and every ,
Noting that , we have
The above integral can be computed explicitly and is equal to Noting that , we find that whence (4.4) follows. This completes the proof of Proposition 4.1.
5. Polynomially mixing flows in dimension one.
In this section, we extend the result of Theorem 1.7 to some flows with polynomial mixing rates. We use the setting of [10]. Recall that a flow is called partially hyperbolic if there is a invariant splitting and positive constants such that for each
(i)
(ii) For any unit vectors we have
For partially hyperbolic flows the leaves of are tangent to the leaves of an absolutely continuous foliation
Fix constants . Let denote the collection of sets which belong to one leaf of and satisfy
for all , where is the neighborhood of the boundary of We say that the sets from have bounded geometry.
Fix Let denote the set of linear functionals of the form
where is a probability density on , We shall often use the existence of almost Markov decomposition established in [10, Section 2]: if , and are large enough and and are small enough, then given and we can decompose
(5.1) |
with and for some
We say that a measure is u-Gibbs if its conditional measures on unstable leaves have smooth densities. The existence of almost Markov decomposition implies that u-Gibbs measures belong to the convex hull of for appropriate choice of (see [5, §11.2]).
From now on we shall fix constants involved in the definition of which satisfies (5.1) and whose convex hull contains u-Gibbs measures. For the sake of simplicity, we will write instead of
We say that the unstable leaves become equidistributed at rate if there is such that for all we have
(5.2) |
where is the reference invariant measure for We note that if (5.2) holds with , then belongs to the convex hull of , whence it is a u-Gibbs measure (in fact, it is the unique SRB measure for , see [10, Corollary 2]).
Theorem 5.1.
Theorem 1.7 remains valid if the assumption that enjoys exponential mixing of all orders is replaced by the assumption that is partially hyperbolic and unstable leaves become equidistributed with rate for some and .
Corollary 5.2.
Indeed, according to [9, Theorem 3], for Anosov flows, unstable leaves are equidistributed on at rate where as
The proof of Theorem 5.1 requires a modification of Proposition 2.1. We shall use the same notation as in that proposition.
Proposition 5.3.
Suppose that is partially hyperbolic with unstable leaves equidistributed at rate for some Suppose that for some such that
(5.3) |
and we have
(a) and ;
(b) for every ,
(c)
(d)
Then as
Lemma 5.4.
Proof.
The fact that conditions (a) and (b) hold for arbitrary are verified as before. Condition (d) is immediate from the definition of In order to verify condition (c) we note that by Proposition 4.1, for any , as Therefore it is enough to check that if and are sufficiently small, then
is close to 1, where
and is the local time defined by (3.6). Note that
Therefore, using the anticoncentration large deviation bound, we conclude that
Now the result follows by the Markov inequality provided that
Proof of Proposition 5.3..
We divide the interval into big blocks of size separated by small blocks of size where the parameters will be chosen later. Let denote the -th big block and be the union of the small blocks. Let Note that We claim that as provided that
(5.4) |
Indeed
(5.5) |
According to [10, Theorem 2], Therefore property (b) shows that the integral of the RHS of (5.5) with respect to is . To integrate with respect to we divide into unit intervals. Noting that the mass of each interval is and the number of intervals is , we conclude that which tends to zero under (5.4).
Thus the main contribution to the variance comes from the big blocks. Let
Fix any . We will show that there is a sequence , depending only so that for each we have
where is the number of big blocks. Let with .
Lemma 5.5.
If
(5.6) |
then
where
Note that by property (b) and (5.6), we have Hence and so the estimate of Lemma 5.5 can be rewritten as
Repeating this process, we obtain
(5.7) |
Next we claim that
(5.8) |
provided that
(5.9) |
Indeed, The second term here is at most
Summing over using assumption (b) we are left with
where in the last inequality we also used assumption (a).
This proves that (5.9) implies (5.8).
(5.8) shows in particular that
Also due to assumption (c) of Proposition 5.3, while
provided that
(5.10) |
Plugging these estimates into (5.7) we conclude that for all we have
(5.11) |
if and satisfy (5.4), (5.6), (5.9), and (5.10). Thus we need and to satisfy
Since can be chosen arbitrary close to and can be chosen arbitrarily close to , the above inequalities are compatible if (5.3) holds. It then follows that (5.11) holds on the convex hull of which includes This completes the proof of Proposition 5.3 modulo Lemma 5.5.
Proof of Lemma 5.5.
Fix arbitrary Then
where
Next
Note that due to the equidistribution of unstable leaves.
To estimate we split
where includes the terms where and includes the other terms.
To estimate , we note that
where with
Hence using the equidistribution of unstable leaves, we obtain
It follows that
Finally
Summing over and using the fact that , we obtain the result.
Acknowledgement
C.D. was partially supported by Nankai Zhide Foundation, and AMS-Simons travel grant. D.D. was partially supported by NSF DMS 1956049. A.K. was partially supported by NSF DMS 1956310. P.N. was partially supported by NSF DMS 2154725.
References
- [1] Avila A., Gouezel S., Tsujii, M. Smoothness of solenoidal attractors, Discrete Contin. Dyn. Syst. 15 (2006) 21–35.
- [2] Billingsley, P., Convergence of Probability Measures Wiley (1968).
- [3] Bjorkund M., Gorodnik A. Central Limit Theorems for group actions which are exponentially mixing of all orders, J. Anal. Math. 141 (2020) 457–482.
- [4] Bolthausen E. A central limit theorem for two-dimensional random walks in random sceneries, Ann. Probab. 17 (1989) 108–115.
- [5] Bonatti C., Diaz L. J., Viana M. Dynamics beyond uniform hyperbolicity. A global geometric and probabilistic perspective, Encyclopaedia Math Sciences 102 (2005) Springer, Berlin, xviii+384 pp.
- [6] Castorrini R., Liverani C. Quantitative statistical properties of two-dimensional partially hyperbolic systems, Adv. Math. 409 (2022), part A, paper 108625, 122 pp.
- [7] Cohen G., Conze J.–P. CLT for random walks of commuting endomorphisms on compact abelian groups, J. Theoret. Prob. 30 (2017) 143–195.
- [8] Cohen G., Conze J.–P. On the quenched functional CLT in random sceneries, Studia Math. 269 (2023) 261–303.
- [9] Dolgopyat D. On decay of correlations in Anosov flows, Ann. of Math. 147 (1998) 357–390.
- [10] Dolgopyat, D. Limit theorems for partially hyperbolic systems, Trans. AMS 356 (2004) 1637–1689.
- [11] Dolgopyat D., Dong C., Kanigowski A., Nándori P. Mixing properties of generalized transformations, Israel J. Math. 247 (2022) 21–73.
- [12] Dolgopyat D., Dong C., Kanigowski A., Nándori P. Flexibility of statistical properties for smooth systems satisfying the central limit theorem Invent. Math. 230 (2022) 31–120.
- [13] Dolgopyat D., Dong C., Kanigowski A., Nándori P. Notes on Bjorklund-Gorodnik Central Limit Thereom, preprint.
- [14] Dolgopyat D., Fayad B., Liu S. Multiple Borel-Cantelli lemma in dynamics and multilog law for recurrence, J. Mod. Dyn. 18 (2022) 209–289.
- [15] Kesten H., Spitzer F. A limit theorem related to a new class of self-similar processes, Z. Wahrsch. Verw. Gebiete 50 (1979) 5–25.
- [16] Kleinbock, D. Y., Margulis, G. A.: Bounded orbits of nonquasiunipotent flows on homogeneous spaces. AMS Transl. 171 (1996) 141–172.
- [17] Le Borgne S. Exemples de systemes dynamiques quasi-hyperboliques a decorrelations lentes, preprint; research announcement: CRAS 343 (2006) 125–128.
- [18] Nándori, P., Szász, D. Lorentz Process with shrinking holes in a wall Chaos 22 (2012) paper 026115.
- [19] Pène F. Planar Lorentz process in a random scenery, Ann. Inst. Henri Poincare Probab. Stat. 45 (2009) 818–839.
- [20] Rudolph, D,Asymptotically Brownian skew products give non-loosely Bernoulli K-automorphisms, Invent. Math. 91 (1988) 105–128.
- [21] Szász D., Varjú T. Local limit theorem for the Lorentz process and its recurrence in the plane, Erg. Th. Dynam. Sys. 24 (2004) 257–278.