Random Composition of L-S-V Maps Sampled Over Large Parameter Ranges
Abstract
Liverani-Saussol-Vaienti (L-S-V) maps form a family of piecewise differentiable dynamical systems on depending on one parameter . These maps are everywhere expanding apart from a neutral fixed point. It is well known that depending on the amount of expansion close to the neutral point, they have either an absolutely continuous invariant probability measure and polynomial decay of correlations (), or a unique physical measure that is singular and concentrated at the neutral point (). In this paper, we study the composition of L-S-V maps whose parameters are randomly sampled from a range in , and where these two contrasting behaviours are mixed. We show that if the parameters are sampled with positive probability, then the stationary measure of the random system is absolutely continuous; the annealed decay rate of correlations is close (or in some cases equal) to the fastest rate of decay among those of the sampled systems; and suitably rescaled Birkhoff averages converge to limit laws. In contrast to previous studies where , we allow in our sampling distribution. We also show that one can obtain similar decay of correlation rates for , when sampling is done with respect to a family of smooth, heavy-tailed distributions.
1 Introduction
Let us consider the one-parameter family of L-S-V maps [17] , , defined as
(1) |
These maps were introduced as a simplified version of the previously studied Pomeau-Manneville family [20], retaining the essential property of having a neutral fixed point at but all having the same two intervals of monotonicity and uniformly expanding, affine, right-hand branch***It is reasonable to expect that the results one can obtain for L-S-V maps hold for the original P-M maps. However we take advantage of the affine second branch in L-S-V at various points in the proofs so extending our arguments to the non-affine case is likely to be technically complicated. Since then, L-S-V maps have become a standard example of dynamical systems with intermittency, alternating stretches of time where orbits exhibit chaotic behaviour, and long stretches of time where they are almost constant and close to zero. The uniform expansion of the maps away from the fixed point is responsible for the chaotic behaviour, while the fact that implies that it takes a long time before orbits can escape the vicinity of . It is well known that for , has a unique physical probability measure which is absolutely continuous and exhibits polynomial decay of correlations where the rate of decay depends on [23]. For , does not have any invariant absolutely continuous probability measure, but it has a unique invariant absolutely continuous infinite measure, and the Dirac delta concentrated at zero is the unique probability physical measure. This is due to the fact that for , the repulsion in a neighbourhood of zero is so small that, asymptotically, almost every orbit spends most of its time arbitrarily close to zero.
After fixing a compact interval with , we study the composition of maps sampled randomly with respect to a given probability measure on , and characterise the average statistical properties of their orbits (annealed results). The case where has been previously considered. In [4] and [3] the authors show that if the distribution is discrete and samples only finitely many values in , the annealed correlations decay at a rate equal to that of the system with the smallest which is sampled with positive probability. Other works deal with quenched results that, in contrast with annealed results, establish decay of correlations and convergence to limit laws for almost every sequence of maps sampled with respect to the given measure on [6] (see also [5] for quenched results on random composition of unimodal maps). In [2], [19], and [18] an arbitrary composition of maps (sequential random systems) from , where and satisfy some additional technical conditions, is shown to give decay of correlations at a speed bounded above by that of the system , satisfy central limit theorems, and large deviations principles.
In contrast with the above cited works, we focus on the case where and thus maps with both and are composed. In this case, there is competition between two contrasting behaviours: maps with tend to spread mass over the whole space, while maps with , although still expanding on most of the space, tend to accumulate mass at the neutral fixed point. We show that it is enough for maps with to be sampled with positive probability, to ensure that the average random system has an absolutely continuous stationary probability measure, polynomial decay of annealed correlations, and convergence to limit laws.
The case of discrete (i.e. ) was previously treated in [3] for using the Young tower approach [23] to find estimates on the decay of correlations. In this analysis, the base of the skew product is conjugated to a piecewise affine and uniformly expanding system. This allows one, after inducing on a suitable subset of the phase space, to reduce the analysis to the study of a Gibbs-Markov system [1] with countably many invertible branches. This construction is not possible in the case of a non-discrete where one can find uncountably many inverse branches all defined on measurable sets of zero measure that cover a set of positive measure, thus obstructing the construction of a countable Gibbs-Markov structure.
Our main approach in this paper is based on the renewal theory of operators which was introduced by Sarig [21] and further developed by Gouëzel [10]), and which can deal with any measure on . For the machinery to work, it is crucial to bound the distortion of the composition of different maps with parameters chosen arbitrarily from . For a single (deterministic) map this was done by Young [23] using a direct calculation, valid for all . For random maps, distortion estimates have been obtained in the case using the Koebe principle and non-positive Schwartzian derivative as in [11] and [4]. However, this technique for random maps fails when since has points where the Schwartzian derivative is positive when . In the following we give a direct estimate of the distortion, more in the spirit of [23], that encompasses the general case , including .
In the case where the interval of parameters is unbounded, i.e. , the approach above does not work as a uniform bound on the distortion is missing. However, in the special situation where is an absolutely continuous measure on with very regular density, we present a second approach that exploits the continuous distribution in parameter space by looking at the system as a mixture between a diffusion process, and a deterministic uniformly expanding system. The estimates on the (annealed) decay of correlations are obtained using the theory of Markov chains with subgeometric rates of convergence [22]. We are going to present this approach in the case where the distribution of has fat polynomial tails at infinity, meaning that large parameters are sampled with high probability.
2 Setting and Results
Let us consider the one-parameter family of L-S-V maps as in (1) above. Given with and , we consider a probability measure on the compact interval , and the product measure on . We assume without loss of generality that belongs to the topological support of (that is for any ).
These data define the random dynamical system taking skew-product form
(2) |
where the reference measure on is , with the Lebesgue measure on . Given and , we denote by . We are going to describe the annealed statistical properties of this random system, i.e. the statistical properties of averaged with respect to the reference measure . In the following, for a measurable we denote by
and for a measurable we denote with an abuse of notation
Decay of correlations.
We show that the random system has an absolutely continuous stationary probability measure and exhibits decay of correlations at polynomial speed.
Theorem 2.1.
Let , and let be a probability measure on such that lies in the topological support of . Then there is an absolutely continuous stationary measure , i.e. is invariant under . For , and all , we have
It is natural here to compute correlation integrals with respect to the reference measure since one does not, in general, know explicitly. Note that so, heuristically, we interpret this result as weak convergence of the sequence to , measured against Lipschitz test functions. An elementary calculation shows (and ) above can be replaced by the stationary measure in both integrals, obtaining the same decay rates with respect to the stationary measure, more in line with results in classical probability.
We prove this theorem using the renewal theory for operators (see Section 5.1). This approach is more standard in the study of dynamical systems with nonuniform hyperbolic properties, and relies on the bound on distortion given in Section 3. However, we have to deal with the fact that the induced map we are going to study, with , is not Gibbs-Markov, as is usually assumed to be the case. In fact, there is an uncountable partition of sets with zero measure that are mapped bijectively onto by and that cover a set of positive measure. Therefore, we have to prove the spectral properties of the operators involved directly.
In Section 6 we consider the case where is absolutely continuous with a power law distribution, and prove a decay of correlations statement for bounded measurable functions. We look at the diffusion process induced by the skew product map on the vertical fibre and use the theory of Markov chains with subgeometric rates of convergence to their stationary state. This approach does not use the bound on the distortion, whose role in the arguments is played by the diffusion and gives decay of correlations for arbitrary functions.
For and , let denote the measure on defined by
for any . Alternatively, is characterized by for any .
Theorem 2.2 ( decay of correlations for power law parameter distributions).
Let and let and consider the random composition of LSV maps where the parameters are chosen i.i.d. from . The corresponding Markov chain has a stationary probability distribution, . Let . Then for all ,
Convergence to limit laws.
Let be a Lipschitz function. Define by
where denotes projection to the second coordinate.
We denote the Birkhoff sums with respect to by
Theorem 2.3.
Let , , and . Let be a probability measure supported on with in the topological support of . Let be a Lipschitz function on with . Then there is a stable law a sequence , and a sequence such that
where the convergence is in distribution. Furthermore
-
i)
if , then is a Gaussian;
-
ii)
if , then is a stable law with index .
These limit laws assuming a compact parameter range are proved using the Nagaev-Guivarc’h approach as outlined in [9] and [12]. The spectral properties of the transfer operator of will play a crucial role to this end. The case is not addressed in this work, and is likely to be delicate since even for the deterministic map the analysis in [9] (Section 1.3) derives a Gaussian limit under normalization whereas for the limit is a stable law similar to the above. One can therefore expect that when , precise properties of the distribution of near will be needed to derive limit theorems in the random case, in contrast to the simpler results stated above.
For our argument exploits the approximation†††For a more precise statement, see the proof of Theorem 2.3 in Section 5. in distribution. A similar argument holds when . When the estimate on is more delicate, even in the deterministic case, so we do not consider that case in the current work.
Finally, we do not consider limit theorems in the case of unbounded parameter range and heavy tails, such as . Although there may be path to obtain this via the Nagaev-Guivarc’h approach, the details will require a complete reworking of the bounded range case. We suggest this for possible investigation in the future.
3 Bound on the distortion
Definition 3.1.
Suppose is a piecewise differentiable map with on any such that is differentiable. The distortion of on is defined as
We restrict our attention to and for the moment write (abusing notation),
Proposition 3.2.
There is such that for any , , and
As an immediate corollary to the previous proposition we obtain:
Corollary 3.3.
There is such that for any and interval with , mapping bijectively to
Proof.
There is and numbers where is the -th return of all points in to the set . From Proposition 3.2, we get that for every
Since one gets
∎
To prove the proposition we use the following lemma.
Lemma 3.4.
For any closed interval define
Then,
(3) |
Proof.
Fix . For
where in the estimates above we used the fact that is positive and monotonically increasing. Calling and all we need to show is that
(4) |
The left-hand side equals
and the right-hand side equals
LHSRHS | |||
The fact that LHSRHS follows from the fact that
is non-decreasing as an elementary derivative calculation shows, and the fact that LHSRHS with . This concludes the proof of (4). ∎
4 Estimates on the Return Times to the Inducing Set
Let and the probability measure obtained by restricting and then normalizing to . We use the following condition on the tails of the return time to in the renewal theory approach
(C1) |
In the diffusion driven case we are going to use the following, closely related condition: there is a non-decreasing function such that
(C2) |
In fact, for , if condition (C1) holds for all then condition (C2) holds for for all . To see this, we use summation by parts, showing
Notice that, having fixed the family , (C1) and (C2) are conditions on the measure . Sharp bounds for the expressions in (C1) and (C2) have been obtained for various in [6]. Below we prove the following proposition
Proposition 4.1.
Conditions (C1) and (C2) concern the return times of orbits to the inducing set under the map . They can be verified on a case-by-case basis in the following way (see also [6]). Call the first invertible branch of , and mod 1 on the second branch common to all the maps in the family. Define . Then
This implies that
(5) |
where the expectation is with respect to . The equality above together with computations as in [4] allow us to prove Proposition 4.1.
Proof of Proposition 4.1.
From the assumptions, . Let , so that from the definition of topological support, . It is known that if is a sequence such that , then (see for example [4]) and is a consequence of the monotonicity found in the family of functions . It follows from the Hoeffding concentration inequality for i.i.d. and bounded random variables that
so that calling
(6) |
. Now, using (5):
establishing Condition (C1) for any . As pointed out above, this implies that Condition (C2) holds with for any .
If , then one can repeat the above reasoning with to obtain
∎
5 Renewal Theory Approach
In this section we prove Theorem 2.1 on correlation decay, and Theorem 2.3 on convergence to stable laws. In Section 5.1 we treat the correlation decay using the renewal theory for transfer operators, as introduced by Sarig [21] and further developed by Gouëzel in [10], while in Section 5.2 we use the Nagaev-Guivarc’h approach to prove convergence to limit laws.
5.1 Decay of Correlations via the Renewal Theory for Transfer Operators
We apply the following general theorem that can be found in [10]. Let us denote the open unit disk by .
Theorem 5.1 (Theorem 1.1 [10]).
Let be bounded operators on a Banach space such that converges for every . Assume that:
-
•
for every , where and are bounded operators on such that ;
-
•
1 is a simple isolated eigenvalue of ;
-
•
for every , is invertible.
Let be the eigenprojection of at 1. If for some and , then for all
where is given by , and is a bounded operator satisfying
To apply the above result to our context, consider , and its transfer operator. Define operators for every , in the following way
for , where denotes the transfer operator for the skew product .
Corollary 5.2.
Assume there is a Banach space , , such that the operators and satisfy the hypotheses of Theorem 5.1. Then for any supported on , and any ,
where is the stationary measure for on .
Let us say a few words about how this result follows from the previous theorem, given the definitions above. We will see (in the proof of Proposition 5.3 below) that , and , with and the relation . We can then write
We now apply the expansion of to obtain
where the higher order terms arise from the second and third terms in the expansion of and both decay with a rate upper bounded by for all three ranges of in Theorem 5.1).
Corollary 5.2 now gives Theorem 2.1 for the restricted case of and supported on since in both integrals. We can extend to the case where is supported on as follows: Set for , for and . Observe that so and the correlation integral above becomes
leading to the result stated in Theorem 2.1 provided is supported on . The final extension to fully supported can be established using the method detailed in Gouëzel [10], Theorem 6.9.
The proposition below shows that the hypotheses of Theorem 5.1 and Corollary 5.2 are satisfied in our setup by the Banach space of functions on that are a) constant w.r.t. and b) Lipschitz w.r.t. the spatial variable . To simplify notation, we indicate these functions in terms of only and write
where
is the Lipschitz semi-norm, and is endowed with the norm . Notice that if the function is also constant w.r.t. . In particular, we can compute the skew product transfer operator as
where denotes the transfer operator associated to on with respect to . We will see this leads to a useful simplification when computing higher powers and the induced transfer operator below.
Proposition 5.3.
Let be as above. The maps and are bounded operators on . Suppose that condition (C1) is satisfied for some . The series converges on , and:
-
(i)
is invertible for every ;
-
(ii)
for , where and ;
-
(iii)
has a spectral gap, i.e. there is with , , and there is satisfying , , such that ; The non-degeneracy condition holds.
-
(iv)
.
Proof of Proposition 5.3.
Fix any , a , and any sequence and write for . Define to be the collection of maximal subintervals of where is continuous, returning to for the th time at time under . If , the points of have the same return times up to the th return and is mapped injectively and onto under . Pick , then
where the second sum is over and again we indicate by the transfer operator of . We are going to prove that satisfies a Lasota-Yorke inequality. First of all notice that
with indicating the restriction of the function to . Given two points
(7) | |||
To bound we show first that there is uniform in the choice of such that
In fact, recall that is mapped bijectively onto . By the mean value theorem, there is such that
Combining this with the bound on distortion from Corollary 3.3, there is a constant such that
This and the fact that
imply that
To bound first of all notice that
where we used that the diameter of is less then given that its points have common return times with respect to sequences from the cylinder . Then notice that
where we made use of Proposition 3.2 in the last line. We conclude that
Putting the bounds for and together, we conclude that there is (uniform) such that for every , , , and
(8) |
Applying this to , which consists of a single interval, say, we see
We note
so that
Since for each , the sets form a countable partition of , it follows that as required.
For every , let be , the countable partition of according to the first returns to under the maps . One obtains
(9) |
where in the first line, the sum is over those intervals whose th return occurs at time . An elementary calculation shows
for every which, when combined with , yields . Combining this with the estimate above gives the following Lasota-Yorke inequality.
(10) |
The operator is therefore quasi-compact (by Hennion’s theorem [13]).
We now show that has unique fixed point, giving rise to an invariant mixing measure for where is absolutely continuous with respect to Lebesgue. In particular, consider the cone of positive functions . Since is independent of
Since has bounded distortion and expansion both uniform in and , there is sufficiently large such that if
with . It is easy to conclude that if , then . It is standard to conclude that fixes a direction in , giving a mixing invariant absolutely continuous probability measure (see [15] pp 244-250). It is also easy to see that . This fact together with quasi-compactness of and mixing implies that the operator has a spectral gap.
Point (i) We first deal with . Pick . Notice that
where the summation is over , the collection of intervals (depending on ) returning to for the th time at the th step as above. The terms corresponding to and in (7) are scaled by , where is the time of the th return. In particular, arguing as in (9), we see , where does not depend on or . Hence for , is convergent, and is invertible.
For , one can show that satisfies a Lasota-Yorke inequality as we did for . This implies that the essential spectrum of is contained in the open disk , and is invertible if and only if is not an eigenvalue of . Now let lie on the unit circle but . Suppose for a contradiction that is a Lipschitz function that is an eigenfunction of with eigenvalue 1. Then taking absolute values and using the triangle inequality, we see , this inequality being strict unless for each , almost every term contributing to has the same argument. Since , this implies , so that is the leading eigenvector, say, of . We write where for all . Since , it is bounded below, so that must be Lipschitz. The condition for equality in the triangle inequality implies that for -a.e. . In particular for -a.e. , for -a.e. in each , we have
Since both sides of the equality are Lipschitz functions, this holds for all . Since the are arbitrarily short for large , we see that becomes arbitrarily small. But the above equality and the fact that maps onto imply that . Hence is a constant and we may assume . Now implies that .
Point (ii) One can verify the renewal equations
Now since converges, we deduce that is invertible. The coefficient multiplying in this last sum is
so . This also establishes the boundedness of for each .
Point (iii) Notice that , so that there is the required spectral gap and simple eigenvector with eigenvalue 1 as shown above. We have , so that this is a bounded operator by point (iv). Note also that preserves integrals, so that also preserves integrals and the operators preserve the class of non-negative functions. Since , It follows that is non-zero as required. ∎
5.2 Central Limit Theorem and Convergence to Stable Laws
The fact that has a spectral gap allows us to use the Nagaev-Guivarc’h method (as outlined in [1] [12]) to show that suitably rescaled Birkhoff sums converge to stable laws. In this approach, one first proves that suitably rescaled Birkhoff sums of observables in the induced system () converge to stable laws, and then use this result to show that the same limit theorem holds for the unfolded system (). We are not going to give full details of the proofs, but we are going to present some computations needed in the particular case that we are treating, and direct the reader to the relevant literature for the remaining standard part of the argument.
Let us call
the Birkhoff sums with respect to . We assume, as in the statement of Theorem 2.3 that is a Lipschitz function on satisfying . Recall that given , , and is the cumulative distribution function of with respect to : that is .
Proposition 5.4.
Assume is such that with is slowly varying as and for large negative . Then there is a stable law a sequence , and a sequence satisfying such that
where the convergence is in distribution. If , then is a Gaussian, while if , is a stable law of index .
Proof.
Calling , it follows that
The relation above provides the foundation to the Nagaev-Guivarc’h approach that recovers information on the characteristic function of from spectral properties of the family .
Step 1
The operator has a spectral gap in for all sufficiently small . We show how to prove this fact directly with computations analogous to those carried out in the proof of point (iii) in Proposition 5.3. However, in that proposition , i.e. , and therefore is piecewise constant on every fibre . When this is not the case, remains piecewise Lipschitz on fibres, where the fibres are partitioned according to the return time to . However, the Lipschitz constants on these pieces are typically not bounded, so the previous argument needs some extra care.
One needs to estimate , given by
where the summation is over , the collection of intervals returning to for the th time at time . Now fix in ; , and . Let be respectively the points in and that lie in . Then the contribution to coming from the interval is
We estimate the absolute value of this quantity by
The combined contribution from the second and third terms as runs over and runs over and integrated over is estimated exactly as in Proposition 5.3. It remains to estimate the combined contribution from the first term. Since the ’s are non-contracting and the th return time is , the first term may be estimated by . As before, the distortion estimates give . Hence the combined contribution to coming from the first terms in the above display (as , and vary) is bounded above by
where denotes the th return time to . The measure agrees with (where is the absolutely continuous invariant measure for , i.e. the density of is the unique fixed point of on ) up to a multiplicative factor that is uniformly bounded above and below. Hence the above displayed quantity may be estimated by
By Kac’s Lemma, the integral is finite, so the above reduces to . Taking sufficiently small, we see that the additional contribution to the estimate of from that appearing in the earlier proposition is . In particular, for sufficiently small , one obtains a Lasota-Yorke inequality analogous to the one in (10).
Step 2
The family of operators (acting on ) is continuous in . To see this, we estimate . As above, is expressed as an integral over of a countable sum (one term for each ). Fix in , and as above. Since we are considering first returns, there is exactly one interval, , in . Letting and be the preimages of and under in , the corresponding contribution to is estimated by
(where we dropped the -dependence of for clarity of the notation), which we bound as a difference of products using the triangle inequality as usual, giving rise to the sum of four terms. We use the estimates for any between and ; ; (obtained from the distortion bound). Combining these estimates, we find that the contribution to coming from the th interval from each of the four terms described above is bounded above by . Summing over and integrating, we find
As in Step 1, the integral is finite using Kac’s Lemma, so that .
Step 3
The spectral gap of together with continuity of the family implies that has a simple eigenvalue such that as .
If , then the distribution random variable with distribution is square summable, and results from [16] show that the limit law is a Gaussian (see also [12]).
If , an expansion of in can be obtained following the proof of Theorem 5.1 from [1]. The expansion will depend on the distribution . One of the main requirements here is that is in the domain of attraction of a stable law or, equivalently, that there are , a slowly varying function as , and with such that as and for . These conditions are ensured by the assumptions on with . One can then conclude the statement of the proposition as outlined in Theorem 6.1 from [1]. The gist of the argument is that approaches . This information and the expansion of found at the previous step can be used to determine the limit for the law of when . ∎
The following general theorem allows us to \sayunfold the limit theorem from the induced system to the original system .
Theorem 5.5 (Theorem 4.8 [12]).
Let be an ergodic probability preserving map w.r.t. the measure , let and be two sequences of integers which are regularly varying with positive indices, let , and let be a subset of positive measure. Denote by the probability measure induced on . Let be the first return time to , the induced transformation, a measurable function, , . Assume that in distribution, that is such that is tight and tends in probability to zero. Then in distribution.
Before proving Theorem 2.3, we show that the tails of and satisfy the hypotheses of Proposition 5.4. Recall that is a probability measure supported on with ; in the topological support of ; and is a Lipschitz function on . For the rest of this argument we will assume . Then there is a bounded number of times one can have which implies the condition for all . The case will be addressed at the very end of this section. The following proposition shows that the the distribution of is a regularly varying function independently of the choice of , i.e. it can always be written as the product of a monomial and a slowly varying function. Furthermore, the index of the distribution depends only on the minimum of the topological support of .
Proposition 5.6.
Let and be as above. Then satisfies , where is a slowly varying function.
The proof of this proposition requires a couple of lemmas. Let
In the function above, is the average size (integrating ) of a one-step displacement . The random i.i.d. nature of the composed maps implies that the number of steps needed to escape from the point to the inducing set, concentrates around a given value, and is expected to be a proxy for the value around which concentrates, where is the unique preimage of in . As a convenient abuse of notation in the following computations, we will write, when , with the convention that this means . The following lemmas make the above heuristic precise.
Lemma 5.7.
For all , there exists an such that if , then
for all | |||
In particular, if and , then
To prove this lemma we first show a concentration result for the number of steps needed to escape from the interval .
Lemma 5.8.
Let the probability measure on be as above. Let denote the interval and fix . For , let , that is the number of steps spent in .
For all sufficiently large and all ,
where .
Notice that when a random orbit enters from the first point of that is encountered lies in . Hence the lemma is estimating the number of steps spent in when entering from the left.
Proof.
Let . If , the size of each step is where . We approximate this above and below by and . The ratio of these step sizes is bounded above by , which approaches 1 as .
This allows us to compare the displacement of under the composition (that is ) to , the sum of the displacements , provided that remains in for . It is convenient (to maximize the strength of the conclusion from Hoeffding’s inequality) to rescale the displacements by a multiplicative factor of .
Let be a sequence of i.i.d. random variables, where and is distributed according to . In particular, for each . Similarly, set , so that for , the step size satisfies . We use and to refer to the distribution on the and in order to distinguish from used elsewhere in the paper. Notice also that (we drop the subscript in the following computations and note that the random variables here are and ).
If we define (dependent) random variables by , then provided remains below we have . Hence for , implies . For , we have
The equality follows from the expression for and a simple manipulation of the inequality in parentheses. The second inequality holds for all sufficiently large using the facts , and , which follows from the comparison of step sizes mentioned above; the final inequality is an application from Hoeffding’s inequality. Since , we see that , while takes values in the range with positive probability, say, so that . Thus we see the probability of spending less than steps in is (much) smaller than for large .
An analogous argument based on studying the probability that shows that the probability of spending more than steps in is smaller than for large . ∎
Proof of Lemma 5.7.
Let be fixed. Given , let , and . Notice that for large . Let be the event that starting from , the number of steps spent in is at most for each . Applying Lemma 5.8 with taken to be , provided (and hence ) is sufficiently large, has probability at least . After leaving , the position exceeds and the number of steps before hitting is bounded above by , so that the contribution to from this part of the orbit is bounded above by . Similarly, for , the contribution from is bounded above by . We claim that for all sufficiently large (with and related to as above),
(11) |
One may check that , where as . In particular, the right side of (11) exceeds for all large (and ). By the choice of , . Similarly, for all sufficiently large , for all . Also , so that . To show (11) it therefore suffices to show
(12) |
To see this, set and observe
This Riemann sum is
Notice that provided is sufficiently large, . We also see that for any , for all sufficiently small , , so that for large (and corresponding and ),
Since
the claimed inequalities (12) and hence (11) follow. We have therefore shown that for , . Now if , since , the same bound applies to . Hence we have established the second statement of the lemma. The first statement is proved in a very similar way, by using Lemma 5.8 to give lower bounds on the time spent in where . ∎
Proof of Proposition 5.6.
Since for , for all , Lemma 5.7 shows that . It therefore suffices to show that is a regularly varying function with order . However, Theorem 1.5.12 of [7] shows that if is regularly varying of order as , then is regularly varying of order . Applying this to , it suffices to show that for each . A calculation shows
Now set . If , noticing that and are conjugate Hölder exponents, we have
so that is log-convex. Since for all , for large , it follows that as and as . It follows that is regularly varying with index as , in fact
Now it follows from Theorem 1.5.11 (i) of [7] that since is regularly varying with index , for
and therefore is regularly varying with index as required. ∎
Proof of Theorem 2.3.
We can restrict to the case , where is the stationary measure. By Proposition 5.6, the cumulative distribution function, of satisfies , where is slowly varying. We showed in the proof of Lemma 5.7 that where . The same proposition applies if is replaced by so that the cumulative distribution function, of satisfies , where is again slowly varying. As before, behaves asymptotically as the inverse function of . Since and both are regularly varying of order , we see that and .
Let be the solution of as in the statement of Proposition 5.4. By that proposition, we obtain the existence of and such that and converge to stable laws. By Kac’s Lemma, we may assume .
Since converges in distribution, then it is also tight. Since is integrable with , where is the unfolding of to and equals, up to a normalizing factor, the invariant measure . By Birkhoff’s ergodic theorem tends to zero almost everywhere. This implies that tends to zero almost surely (and thus in probability). To see this, consider the set where the Birkhoff averages have limit zero. For , call the sequence of indices realizing the maxima. The sequence is non-decreasing. If is eventually constant, then . If not, then
By Theorem 5.5 the conclusion of the theorem follows when . ∎
Our final step it to consider the case . One simply observes that the statement of the Theorem 5.5 is invariant under , and , another stable law, and apply the previous argument.
6 Diffusion Driven Decay of Correlations
In this section we consider a family of distributions on the parameter space, with unbounded support. More precisely, for , is supported on , and given by
Notice that since the measure has unbounded support, the considerations made in sections 4 and 5 on the uniform bound of the distortion, do not apply. Furthermore, the fat polynomial tails of the distribution of , imply that arbitrarily large parameters are sampled somewhat frequently.
6.1 Markov chains with subgeometric rates of convergence
The arguments in sections 4 and 5 do not apply to this setup. Instead, we exploit the absolute continuity of to translate the deterministic process (2) into a Markov process on the state space and apply a theorem on Markov chains (see [22]) to prove the existence of a stationary measure and to estimate the asymptotic rate of correlation decay. To this end, consider the stationary Markov process on the probability space taking values in , where . Calling the map , then the transition probability , is given by , i.e. the push-forward under of the probability on the parameter space, or more concretely . The Markov chain defines an evolution of measures, which we denote by : , that is if has distribution , then has distribution . In case is an absolutely continuous measure, with density , is again absolutely continuous with density
(13) |
For , we define , and let us call , the probability measure conditioned on .
Definition 6.1.
A Markov chain is irreducible if there is a measure such that for every with and every
Lemma 6.2.
The Markov chain is irreducible and aperiodic.
Proof.
The only problem with these two properties is that . However, this can be fixed by removing the set from the state space, where is mod , and redefining the algebras and transition probabilities accordingly. In this modified state space, every orbit eventually enters and is spread by the diffusion. This implies that any set of positive Lebesgue measure is eventually visited by a random orbit originating from any given . Aperiodicity is a consequence of the two branches of all the maps being onto. ∎
Definition 6.3.
Given a probability distribution , and a nontrivial measure , a set is -petite if
Definition 6.4.
A function is a subgeometric rate function if there is a non-decreasing with and as such that
The main result we are going to use in this section is the following theorem on subgeometric rates of convergence of ergodic Markov chains.
Theorem 6.5 ([22]).
Suppose that is a discrete time Markov process on a general state space endowed with a countably generated -field which is -irreducible and aperiodic. Suppose that there is a petite set and a subgeometric rate function (see above for the definitions) such that
(14) |
where , and is the expectation with respect to the probability law of the Markov process conditioned on . Then the Markov process admits a stationary measure , and for almost every point
where denotes the total variation norm difference between the two measures. Furthermore, if (14) holds and is a probability measure on that satisfies
(15) |
then
For the remainder of this section, let and be fixed. Let be chosen so that .
We compute the one step probability transition density starting from a point , which will be denoted (with support ). For , define to be the value of such that . That is satisfies , or , giving an explicit form: , where we took reciprocals of both numerator and denominator to ensure positivity of the logarithms. Note that this is a decreasing function of . We can then obtain by
Recall that is supported on . By the choice of and since , this contains the sub-interval . We will denote by the step probability transition density.
In this case where has a power law distribution, we can write the transition operator (13) of the Markov process in the form:
where is taken to be 0 if or .
Remark 6.6.
For any fixed , the set
is a petite set. In the case where that we are considering, it is not hard to see that is uniformly bounded below for and . Therefore, in Definition 6.3 one can pick for and zero otherwise, with a multiple of the Lebesgue measure on .
6.2 Outline of the proof of Theorem 2.2
The rest of the section is mostly dedicated to estimating the return times to and to finding rates for which (14) holds.
Notice that is the preimage of under the second branch. Two facts play an important role in the arguments that follow: by the particular choice of , any random orbit starting from will pass through before landing to the right of it; points in are mapped to , i.e. to the left of the petite set . Any point in must visit before hitting .

Step 1. First we show that random orbits originating from tend to end up to the left of quite fast. More precisely, given any initial condition to the right of , the probability that a random orbit hasn’t entered by time is exponentially small in . This is the content of Proposition 6.13 and Lemma 6.14 below which need several other lemmas to be proven.
Step 2. Then we study the hitting times to for orbits originating on . For close to 0, by Lemma 5.7 the time to hit is of the rough order . Since for , for all , the distribution of times to enter after hitting is determined by the distribution of positions at which is hit. Lemma 6.15 assembles the prior facts to show that , conditioned on starting at is absolutely continuous with density bounded above uniformly in . This ensures that the distribution of the time between hitting and entering has a polynomial tail which, compared to the exponential tails of the times estimated in Step 1, dominate the statistics of the returns to . In Proposition 6.15 we put together all the estimates and obtain subgeometric rates for which (14) holds.
6.3 Step 1
In order to show that the hitting times to have exponential tails, we divide into the intervals and . Notice that is such that for each , the iterates of under the ’s remain in the right branch until they hit . We are going to show that
Lemma 6.7.
Let satisfy . There is such that for every
and when .
Proof.
Notice that for , , which is strictly positive, and approaches 1 as . ∎
We show that there exists such that for all , conditioned on and (so that ) then the distribution of has density bounded between and , where is the time of the first return to . To prove this, we first need an estimate on :
Lemma 6.8.
There exists such that for all , and all satisfying (that is, for any , ), one has is bounded between and .
Proof.
We have
Clearly the ratio is uniformly bounded above and below for and so it suffices to establish is uniformly bounded above and below for , and as in the statement of the claim. It is clear that the numerator exceeds the denominator, so it suffices to give an upper bound.
There exist positive numbers and such that for all . Applying this with taken to be and , we see that , so that it suffices to give a uniform upper bound for . Since , we see .
Finally , giving the required upper bound for . ∎
Lemma 6.9.
There exists such that for all the density of conditioned on , satisfies for all .
Proof.
We establish this by showing that there exists such that for all and all , .
We first observe that
(16) |
However, since is supported on , there are only finitely many non-trivial terms in the sum. Also if both belong to , then so
Hence the number of non-trivial terms in the summation for is at least the number of terms in the summation for and at most one more.
Now suppose that is the largest number such that . Since , The previous lemma establishes
If , then since
and , the previous lemma implies
Summing these inequalities yields the desired claim. ∎
The next lemma shows that if the starting density on is controlled as in equation (17) below, then the condition is invariant under the transition operator.
Lemma 6.10.
Let . There exists such that for all and for every density on satisfying
(17) |
then
for any .
Furthermore, for any , there exists such that if satisfies (17), then for all .
Proof.
Let . We start by noticing that for any
Since for any and any ; and using the hypothesis on
Set and notice that the integrand is symmetric about . Integrating by parts, the integral becomes
where and where, in the first inequality, we used the fact that derivative in the second line is negative for all for all sufficiently close to . Since
the first conclusion follows.
For ,
For any , is uniformly bounded above for and and is integrable. Similarly, on , is bounded above and the functions for are integrable over with integral uniformly bounded in , so that the second conclusion holds. ∎
The following lemma shows that the gaps between consecutive entries to are dominated by a random variable with an exponential tail. If , we define to be the time of the th re-entry to . That is and . Recall that is the subset mapped to the left of the petite set. We study the distribution of some random variables conditioned on the event . Notice that since , this is the event that the process leaves before first hitting .
Lemma 6.11.
There exists an integer-valued random variable such that for each absolutely continuous distribution on with density bounded as in (17), there exists a random variable such that the following properties are satisfied:
-
1.
If is continuously distributed on with density , then conditional on , .
-
2.
For all and all satisfying (17),
-
3.
There exists , such that ;
-
4.
Let be absolutely continuously distributed on with density satisfying (17). For any , conditioned on ; and given that , the distribution of , the position at which the system reenters , is absolutely continuous with density bounded above and below by constants that do not depend on or .
Proof.
Let be a probability density on satisfying (17). Let be distributed with density . Notice that by Lemma 6.7, the event , that is that the system leaves before entering , has probability bounded below by a constant . We define by
where denotes an independent geometric random variable with parameter , so that conclusion 1 is evident. To establish conclusions 2 and 3, it suffices to show that there exists such that for all satisfying (17), one has for all . It then follows that there exists such that for all . Then can be defined by with probability for all where is chosen so that ; and with probability .
Notice that in order that , at least one of the following must occur: the system must remain in for steps; the system must exit to a point above (so that it takes or more steps to re-enter ); or the geometric random variable must take a value of or above. Then is at most the sum of these three probabilities. The first of these has probability at most by Lemma 6.7. The third event has probability .
For the second event, note that for , it is only possible that if , that is, if . Define the operator , mapping to itself by . By Lemma 6.10, the function satisfies
(18) |
which is bounded above in a neighbourhood of by some number (which does not depend on the initial distribution). Hence for all small . In particular, the probability of hitting is bounded above by a constant multiple of , where the constant does not depend on .
To establish conclusion 4, suppose we are given that and . We additionally condition on the time taken for the system to exit and the location of the system prior to exiting . We establish bounds on the density of based on this additional information. Then the bounds without this additional conditioning are simply a convex combination of these bounds.
Thus suppose that the system exits for the first time at time , and we are given that . Since , the number of steps to reenter after leaving is one of 1, 2, …, . That is belongs to one of the intervals
for in the range 1 to where is such that . Since , we have . Since typically , one knows that may only occupy a (possibly small, depending on ) portion of , namely . Hence if , one sees that is restricted to a possibly small sub-interval of (and therefore its conditional density may not be bounded away from zero). This is the reason that we introduced the geometric random variable: to ensure that the return time variable, does not completely determine and thereby overly constrain the location of .
Conditioned on , the event has probability for and has probability . We therefore have is
Now an application of Bayes’ theorem, together with Lemma 6.8, shows that conditioned on , and , the distribution of on each interval above is uniform up to a multiplicative factor of the fixed constant , except for in which the density drops off to 0, but has the property that on , the density is within a factor of of that on . Since in the next steps, the interval is mapped linearly onto , the distribution of is uniform up to a multiplicative factor of on . ∎
The following lemma states that after each return to the set , at least a fixed positive proportion of the mass ends up in the subset of . Recall that is defined to be the time of the th reentry to .
Lemma 6.12.
There exists such that for every
Proof.
We now show that the entry time into has exponential tails.
Proposition 6.13.
Given any probability measure with density satisfying (17), there is such that for any ,
Proof.
First of all define random variables , and the random variable equal to the number of reentries to up to the first entry to (recalling that ). With these definitions, the time to enter is given by , where is the number of steps spent in after the th reentry to prior to entering (note that may be 0 if the first point of that the system enters on the th visit is in ).
Notice that if the constant in Lemma 6.10 is chosen sufficiently large, then conclusion 4 of Lemma 6.11 shows that conditional on , the density of on satisfies (17).
Lemma 6.11 and the Markov property, together with Strassen’s theorem [14, Theorem 2.4], imply that the sequence of random variables may be coupled with a sequence of independent identically distributed random variables with exponential tails in such a way that for . Lemma 6.12 implies that also has exponential tails.
The following lemma proves that Proposition 6.13 can be applied to the density of .
Lemma 6.14.
There exists and such that for all in the petite set , one has for all
There exists a such that for all in the petite set , the distribution of , the position at the first entrance to , is absolutely continuous with density satisfying
(20) |
for all (that is, the density satisfies the condition (17) extended to the full interval ).
Proof.
Let be chosen so that . For any , the time to enter is bounded above by a geometric random variable (with parameter not depending on since for all ). For any , the time to enter is bounded above by another geometric random variable by an identical argument. On the interval , the time to enter is uniformly bounded above. Summing these contributions gives the required geometric upper bounds on .
We look at the distribution of . If , then we study
(21) |
the contribution to the density on coming from points that stay in until they enter (necessarily into since ). (Note that this contribution is trivial if .)
If , we showed above that the number of steps before leaving the interval is bounded above by a geometric random variable; and is bounded above for . Combining these, we see that the density of the contribution in (21) is uniformly bounded above.
For the remaining part of the distribution, we condition on the last point of that is visited. We show that for all , conditional on leaving in a single step, the density of conditioned on and satisfies a bound of the form (20). First note, that for and , is uniformly bounded above, so that when it is mapped under iteration of the second branch back to , it gives a contribution that is uniformly bounded above (similar to (16)). It therefore suffices to show that for , satisfies a bound of the form (20) with a that does not depend on . Since the probability of entering in one step is bounded below for these ’s, it suffices to show the existence of a such that for all and all , . For in a small interval and in a small interval , one may check that is increasing in , so that as required. If either or lies outside this range, there is a uniform upper bound on , establishing the required inequality. ∎
6.4 Step 2
Proposition 6.15.
Suppose that is the minimum of the support of the measure and let . For every belonging to the petite set ,
Proof.
Notice that we can bound the return time to the petite set as where is the first time to hit , is the random time after then that it takes to hit (this may be 0), and is the random time needed to go back to starting from when is hit. It follows from Minkowski’s inequality that if , and are finite, then also is finite.
By Lemma 6.14, . To show that , Lemma 6.14 shows that the distribution of satisfies (20) and the conclusion follows from Proposition 6.13.
To estimate , we need to control the distribution of . The system may enter in a number of ways: from without previously entering ; by making a number of visits to and then entering direct from ; or by making a number of visits to and then entering from . For direct entry from , conditioning on the last point visited in , the density on is uniformly bounded above. Similarly, conditioning on the last point visited in before entering from the right, Lemma 6.9 gives a uniform upper bound on the contribution to the density of . For the points entering from , Lemmas 6.9 and 6.14 ensure that on entry to , the density satisfies (17) (with a fixed ). The sum of the densities prior to exiting is estimated in (18) and this gives a bound on the density of the last position in before exiting, which is times the bound in (17). Then the second part of Lemma 6.10 gives a uniform upper bound on this contribution to the density on of . Taken together, we have shown that the distribution of (and therefore the density of its image on ) is absolutely continuous with density bounded above by a constant that is independent of . Since for , , and Proposition 4.1 established condition (C2), we see as required. ∎
6.5 Step 3
References
- AD [01] J. Aaronson and M. Denker, Local limit theorems for partial sums of stationary random sequences generated by Gibbs-Markov maps, Stoch. Dyn. (2001), 193–237.
- AHN+ [15] R. Aimino, H. Hu, M. Nicol, S. Vaienti, et al., Polynomial loss of memory for maps of the interval with a neutral fixed point, Discrete and Continuous Dynamical Systems (2015), no. 35(3), 793–806.
- BB [16] W. Bahsoun and C. Bose, Mixing rates and limit theorems for random intermittent maps, Nonlinearity 29 (2016), 1417.
- BBD [14] W. Bahsoun, C. Bose, and Y. Duan, Decay of correlation for random intermittent maps, Nonlinearity 27 (2014), 1543.
- BBMD [02] V. Baladi, M. Benedicks, and V. Maume-Deschamps, Almost sure rates of mixing for iid unimodal maps, Annales scientifiques de l’école normale supérieure, vol. 35, Elsevier, 2002, pp. 77–126.
- BBR [19] W. Bahsoun, C. Bose, and M. Ruziboev, Quenched decay of correlations for slowly mixing systems, Transactions of the American Mathematical Society 372 (2019), no. 9, 6547–6587.
- BGT [87] N. Bingham, C. Goldie, and J. Teugels, Regular variation, Encyclopedia of mathematics and its applications, Cambridge, 1987.
- Dur [19] R. Durrett, Probability theory and examples, 5th edition ed., Cambridge, 2019.
- [9] S. Gouëzel, Central limit theorem and stable laws for intermittent maps, Probability Theory and Related Fields 128 (2004), no. 1, 82–122.
- [10] S. Gouëzel, Sharp polynomial estimates for the decay of correlations, Israel J. Math. 139 (2004), 29–65.
- Gou [07] , Statistical properties of a skew product with a curve of neutral points, Erg. Thry. Dyn. Sys. 27 (2007), 123–151.
- Gou [15] S. Gouëzel, Limit theorems in dynamical systems using the spectral method, Hyperbolic dynamics, fluctuations and large deviations, vol. 89, Amer. Math. Soc., Providence, RI, 2015, pp. 161–193.
- Hen [93] H. Hennion, Sur un théorème spectral et son application aux noyaux lipchitziens, Proc. Amer. Math. Soc. 118 (1993), 627–634.
- Lig [12] T. M. Liggett, Interacting particle systems, vol. 276, Springer Science & Business Media, 2012.
- Liv [95] C. Liverani, Decay of correlations, Annals of Mathematics-Second Series 142 (1995), no. 2, 239–302.
- Liv [96] , Central limit theorem for deterministic systems, International Conference on Dynamical Systems (Montevideo, 1995), vol. 362, 1996, pp. 56–75.
- LSV [99] C. Liverani, B. Saussol, and S. Vaienti, A probabilistic approach to intermittency, Ergodic Theory Dynam. Systems 19 (1999), 671–685.
- NP [19] M. Nicol and F. Perez Pereira, Large deviations and central limit theorems for sequential and random systems of intermittent maps, arXiv preprint arXiv:1909.07435 (2019).
- NTV [18] M. Nicol, A. Török, and S. Vaienti, Central limit theorems for sequential and random intermittent dynamical systems, Ergodic Theory Dynam. Systems 38 (2018), 1127–1153.
- PM [80] Y. Pomeau and P. Manneville, Intermittent transition to turbulence in dissipative dynamical systems, Comm. Math. Phys. 74 (1980), no. 2, 189–197.
- Sar [02] O. Sarig, Subexponential decay of correlations, Invent. Math. 150 (2002), 629–653.
- TT [94] P. Tuominen and R. L. Tweedie, Subgeometric rates of convergence of -ergodic Markov chains, Adv. Appl. Prob. 26 (1994), 775–798.
- You [99] L.-S. Young, Recurrence times and rates of mixing, Isr. J. Math. 110 (1999), 153–188.