Genericity on submanifolds and application to Universal hitting time statistics
Abstract.
We investigate Birkhoff genericity on submanifolds of homogeneous space , where and are fixed integers. The submanifolds we consider are parameterized by unstable horospherical subgroup of a diagonal flow in . As long as the intersection of the submanifold with any affine rational subspace has Lebesgue measure zero, we show that the trajectory of along Lebesgue almost every point on the submanifold gets equidistributed on . This generalizes the previous work of Frączek, Shi and Ulcigrai in [8].
Following the scheme developed by Dettmann, Marklof and Strömbergsson in [3], we then deduce an application to universal hitting time statistics for integrable flows.
1. Introduction
Let be a probability measure preserving system, where is a Borel measurable space with probability measure , and is an -action (or -action) preserving . Assume that is ergodic, then Birkhoff’s ergodic theorem (cf. [4]) asserts that, for any ,
(1.1) |
for -almost every . In particular, (1.1) holds for any , where denotes the collection of all compactly supported and continuous functions on . Therefore, (1.1) implies that for -almost every ,
(1.2) |
in the weak* topology on the set of all probability measures on . Here is the Dirac measure on .
For , we say that is Birkhoff generic with respect to if satisfies (1.2). Given a Radon measure on (possibly singular to ), if -almost every is Birkhoff generic with respect to , we say that is Birkhoff generic with respect to . It is then natural to ask the following
Question 1.1.
Under what conditions the measure is Birkhoff generic with respect to ?
This question had been previously studied in the case of , is the Lebesgue measure on and in [10]. It was shown that for any , any invariant ergodic probability measure is Birkhoff generic with respect to . This result was strengthened later in [9]. An analogous question was also studied in the context of moduli space of translation surfaces in [29].
We consider Question 1.1 in the setting of homogeneous dynamics. Let , where is a Lie group and is a lattice in . Here and hereafter let be the -invariant probability measure on . Let be a one-parameter flow in . Assume that , where is a one-parameter unipotent flow, that is, the adjoint action of on the Lie algebra of is unipotent. In this case, Ratner’s uniform distribution theorem [20] says that for any , is Birkhoff generic with respect to , where is the -invariant probability measure supported on the orbit closure . By Ratner’s orbit closure theorem [19], these orbit closures are all homogeneous. Thus this provides a satisfactory answer to Question 1.1.
On the other hand, when , here and hereafter is a one-parameter diagonal flow on , that is, the adjoint action of on Lie algebra of is semisimple, a description of Birkhoff genericity of under the flow is much harder. Indeed, the question of describing the orbit closures of diagonal action remains open (cf. [15][27, Conjecture 1]).
Nevertheless, there is a natural class of probability measures on that are interesting to study with respect to Question 1.1. This class of measures is given as follows. We define the unstable horospherical subgroup with respect to by
where is the identity element of . Let be a submanifold and be a normalized bounded supported volume measure of (here and hereafter by normalized measure, we mean that is renormalized to be a probability measure). It has been proved in [21][22][23][28] that if satisfies certain algebraic conditions, then the translation of the measure under converges weakly to as . That is, for any ,
As in Question 1.1, it is curious to ask the following
Question 1.2.
Assume that a bounded supported normalized volume measure of a submanifold is such that the translate of under is equidistributed with respect to , is it true that is also Birkhoff generic with respect to ?
Roughly speaking, Question 1.2 is answered when the manifold is considerably "large" compared to the unstable horosphere subgroup of . In [25], Shi considered the situation where is a semisimple Lie group, is the expanding subgroup (cf. [24]) of and is a normalized bounded supported Haar measure on . By [21], satisfies the assumption of Question 1.2. Shi showed that is also Birkhoff generic with respect to (), and thus gave an affirmative answer to Question 1.2. In the special case where and , authors in [13] also obtained the effective convergence rate of (1.2).
In [8], the authors considered the setting where , and . One of the main results in [8] asserts that if is a curve in that intersects any affine rational line in a Lebesgue null set, then a normalized bounded supported volume measure on is Birkhoff generic with respect to . By the equidistribution result of translation of such under in [3], the result in [8] also gives an affirmative answer to Question 1.2 in the case where is a curve.
Let . In a recent preprint [12, Theorem 1.4], it is shown that when the natural measure on the planar line gets equidistributed under , then for almost every point in , the orbit is dense in . This supports an affirmative answer to Question 1.2.
The aim of this paper is to generalize the genericity results in [8] to and gives an affirmative answer to Question 1.2 when the manifold satisfies certain diophantine condition. We also deduce an application of our results to the statistics of universal hitting time for integrable flows.
1.1. Notations
From now on, any vector in the Euclidean space will be taken to be a column vector, and we will use boldface letters to denote vectors and matrices. Also, a.e. will be the shorthand for Lebesgue almost everywhere. will denote Lebesgue measure of measurable subsets of Euclidean space or absolute value of real numbers. will denote the standard Euclidean norm and the sup norm of a vector or matrix. Throughout this article, for two matrices and , will denote matrix multiplication.
For , will denote the space of by real matrices. is the direct product of m copies of .
Fix integers , . Let , , and . It is well known that is a lattice in and is a lattice in .
Let . Denote the -invariant probability measure on . Note that the action of on is given by
where and with . The multiplication law in is given by
is naturally embedded into by
Fix an . For and , denote
(1.3) | |||
(1.4) | |||
(1.5) |
For a column vector or a matrix , let (or be the first rows (or last rows) of . For example, if , then
With the above notations, the unstable horospherical subgroup of in is
Lastly, for a map , we write .
We also write
(1.6) |
1.2. Main results
For any and , define the probability measure
(1.7) |
As before, we say that is Birkhoff generic with respect to if converges to in the weak*-topology as .
One of our main results is the following:
Theorem 1.3.
Let be a bounded open subset. Let be a -map satisfying for any . Assume that for any ,
(1.8) |
then for Lebesgue a.e. , is Birkhoff generic with respect to .
Using the observation that if a trajectory equidistributes with respect to , then all other parallel trajectories will also equidistribute with respect to (see Lemma 2.3), we can remove the assumption that and strengthen Theorem 1.3 to the following
Corollary 1.4.
Let be a bounded open subset of . Let be a map. If for any ,
then for Lebesgue a.e. , is Birkhoff generic with respect to .
We also obtain the following variants of Theorem 1.3.
Theorem 1.5.
Let be a bounded open subset of . Let be a map. Let . If for any ,
(1.9) |
Then for Lebesgue a.e. , is Birkhoff generic with respect to .
Remark 1.6.
Remark 1.7.
Corollary 1.8.
Let be a bounded open subset of . Let be a map. Let . If for any ,
then for Lebesgue a.e. , is Birkhoff generic with respect to .
For , let be the column vector such that -th row of is and others are .
Corollary 1.9.
Let be a bounded open subset of . Let be a smooth map such that the map has a nonsingular differential at Lebesgue almost every . Let be a map. Assume that for any ,
then for Lebesgue a.e. , is Birkhoff generic with respect to .
1.3. Ingredients of the proof
Proof of Theorem 1.3 follows the similar strategy as in [8]. However, some new ingredients are required. We need the description of orbit closures of in . This can be done using Ratner’s orbit closure theorem following the approach in [3].
We need to construct a suitable mixed height function in our situation, which measures the distance of point to the cusp and singular submanifolds.
1.4. Overview
In Section 3, we will investigate the orbit closure of in using Ratner’s orbit closure theorem.
In Section 4, we prove that for a.e. , the limit measure is invariant under the unipotent group . This enables us to apply Ratner’s measure classification theorem.
In Section 5 and Section 6, we will construct mixed height function for , and give a proof of its uniform contraction property.
In Section 9, we deduce an application to universal hitting time statistics.
2. Reductions and proof of Theorem 1.3
In this section, assuming several Theorems/Propositions/Lemmas that will be proved later, we give a proof of Theorem 1.3.
By Proposition 4.1, for a.e. , after possibly passing to a subsequence the weak* limit of is -invariant. From the definition of (see (1.7)), it follows that is also -invariant. Hence for a.e. , is -invariant. Note that is an epimorphic subgroup of . By [17], as is a probability measure invariant under , is -invariant. By Ratner’s measure classification theorem, any invariant and ergodic probability measure is supported on an orbit closure of on .
A consequence of Ratner’s orbit closure theorem (Theorem 3.1) shows that any orbit closure of is either
(1) the whole , or
(2) concentrated in a proper closed submanifold for some , where
Therefore, it remains to show that for a.e. , is a probability measure on and for any .
Let
(2.1) |
where , and is a by matrix whose -th entry is . Here the choice of is flexible, we just choose a value for that is convenient for us.
Since there are only finitely many such that , is a closed set with Lebesgue measure zero. Thus to prove Theorem 1.3, it suffices to prove it for a closed cube contained in . Now let’s fix a closed cube .
Let be a measurable subset of . For any , we define the average operator by
where is the characteristic function of .
The key proposition, which ensures that is a probability measure putting zero mass on for a.e. , is the following:
Proposition 2.1.
Let . Let be a map satisfying for any . Suppose that
(2.3) |
Then for any , there exists a compact subset and such that for any ,
(2.4) |
It will be proved in Lemma 6.1 that condition (2.3) in Proposition 2.1 follows from condition (1.8) in Theorem 1.3.
Proposition 2.1 will be proved in Section 6. Combining Borel-Cantelli lemma, a direct consequence of Proposition 2.1 is the following:
Proposition 2.2.
Under the assumptions of Theorem 1.3, for a.e. , by possibly passing to a subsequence, converges to a probability measure on in weak*-topology as , and for any .
Proof.
Fix an and . By Proposition 2.1, we can choose a compact subset of such that (2.4) holds for any . Let and apply Borel-Cantelli lemma to the collection of the sets
We can find a measurable subset of with full measure such that for any , for all sufficiently large . Therefore, for any , and . Let , then has full Lebesgue measure in , and for any , and .
To complete the proof, we let . Then has full Lebesgue measure in and the proposition holds for all . ∎
Proof of Theorem 1.3.
(1) a probability measure on , and for any ;
(2) -invariant.
Since is an epimorphic subgroup of , and is a -invariant probability measure on , is -invariant by [17, Theorem 1].
By Ratner’s measure classification theorem [19], any ergodic component of such is supported on an orbit closure of on . Theorem 3.1 describes all the possible orbit closures of on : either it is or it is concentrated on for some .
Since for any , we conclude that for a.e. , . ∎
We note the following
Lemma 2.3.
Assume that for some ,
and for some , as , then
3. Orbit closure
In this section, we will classify all orbit closures of in following [3]. Recall that and .
Consider a base point . Since is a simple Lie group, an application of Ratner’s orbit closure theorem gives the following theorem describing the orbit closure of in :
Theorem 3.1.
The orbit closure is if and only if for any , .
By Ratner’s orbit closure theorem ([20]), for any , there exists a closed subgroup of containing such that
and admits an -invariant probability measure.
It can be checked that if there exists such that , then , where
(3.1) |
This is a closed submanifold of of codimension . In this case, the orbit does not equidistribute in .
The converse of Theorem 3.1 will follow from Lemmas 3.2-3.4. We will follow the proof strategy of [3, Theorem 3]. Let , be as above.
Lemma 3.2.
There is a linear subspace such that , where is a subset of such that for any element of , each row vector of is a vector in .
Proof.
Let . Because , for any , we have It follows that is -invariant and . For any , we have , so . Since is -invariant, . Therefore .
Let , then for any , and any
Let , we obtain . Recall that consists of all trace zero matrices. Let be the matrix with in the -th entry and zero for all other entries. Then for any , . Since , for any we have as well. Therefore, is invariant under left multiplication of all real matrices. Since left multiplication is row operation, there is a linear subspace such that . ∎
Let be the natural projection map and
Lemma 3.3.
Let be the linear subspace of obtained by the Lemma 3.2. Then is a lattice in and .
Proof.
By Lemma 3.2, and is closed and admits an -invariant probability measure, therefore is a lattice in .
By [18, Corollary 8.28], is a lattice in , that is, is a lattice in . Thus has a basis belonging to , and it follows that is a lattice in .
Recall that . Now consider . Again by [18, Corollary 8.28], is a lattice in . Therefore is a finite index subgroup of . Pick a such that is invertible, then . ∎
Lemma 3.4.
Let be the linear subspace of obtained by Lemma 3.2. If for any , . Then and hence, .
Proof.
Suppose , then . Since is a lattice in , there exists a nonzero . Since , we can choose to be a suitable integral multiple of such that , this contradicts to the assumption of the lemma. ∎
4. Unipotent invariance
The collection of all probability measures on the one point compactification of is a compact space in weak*-topology. Therefore, for any , after possibly passing to a subsequence, we have
for some probability measure on . Throughout this section, the function is assumed to be and satisfy .
Proposition 4.1.
For a.e. , is -invariant.
Proof.
Since is a bounded open subset of , it is enough to prove the proposition for a.e. in an open cube of .
We may choose an open interval such that . For , let be the matrix with in -th entry and zero otherwise.
If are two real numbers linearly independent over , then the closure of the subgroup generated by is .
Therefore, given , without loss of generality, it suffices to prove that for a.e. , the limit measure is invariant under .
Theorem 4.2.
[13, Theorem 3.1] Let be a probability space. Let be a bounded measurable function. Suppose that there exist and such that for any ,
(4.1) |
then given any , for -a.e. ,
Lemma 4.3.
There exist such that for any ,
Proof.
In the following proof, for positive valued functions , we write if there exists a positive constant depending only on and (these are fixed throughout the proof) such that . Also, for any positive number , we let denote a group element in a -neighborhood of in .
Without loss of generality, we assume that . For , consider the interval
such that . For any , and any ,
Note that
where the last equality follows by mean value theorem (for simplicity of notations, by uniform boundedness of on , we write for with arbitrary ). Since
we have
Likewise,
Since , is Lipschitz, and hence
Therefore,
(4.2) |
Now we estimate . Note that
As is Lipschitz,
Since and overlap except for a length of , we have
Now we consider the partition such that with for , and . By (4), we have
The above estimate holds for any . Now the lemma follows from the above estimate and Fubini’s theorem. ∎
5. Margulis’ height function
In this section, we will recall the definition of Margulis’ height function on and its uniform contraction property .
Margulis’ height function was first introduced in [6] and later developed in several papers (see for example [1][25]). It measures the depth of elements of into the cusps. It has been used to study equidistribution problem for certain unbounded functions (cf.[6][7][16]) and random walks on homogeneous spaces (cf.[1][5]).
We start with the vector space , where acts on naturally.
Let be a lattice in . We say that a subspace of is -rational if is a lattice in . For any -rational subspace , denote or the volume of . Note that is the norm of in , where is a -basis of . If , we set .
For any lattice , we define for ,
Proposition 5.1.
There exists a continuous map and such that for a bounded open box in , for all large enough and for any unimodular lattice of ,
and there exists such that
Moreover, a measurable subset of is precompact if there exists such that
Proof.
The function above is the Margulis’ height function that we need in our setting.
Remark 5.2.
The function satisfies Lipschitz property as follows: For any bounded neighborhood of of , there exists such that for any , any ,
Indeed, is the maximum of operator norms of elements in acting on .
6. Mixed height function
In this section, we will construct a mixed height function, which is crucial for us to prove Proposition 2.1. The main result of this section is the following:
Proposition 6.1.
Let be a map from to satisfying for any . For any , any closed cube (for , see (2)), there are sufficiently large (depending on , ) and measurable function such that the following hold:
(1) For any , is compact;
(2) For any , if and only if ;
(3) Given any , a box with , where and for all . There exists such that for any , one has
(4) There exists depending on , for any , any and any with , one has
(5) There exists such that the following holds: for any and any box with satisfying either and for all , or and , one has
Remark 6.2.
The function in Proposition 6.1 is the desired mixed height function.
From now on until the end of this section, we will fix a closed cube . Recall that the finite number is defined as in (2.1). By the choice of , and the fact that there are only finitely many satisfying , we obtain such that
(6.1) |
Remark 6.3.
By (6.1), we can choose a closed neighborhood of such that is a closed cube contained in and satisfies
Next we construct a suitable function measuring the distance to the closed submanifold . For , consider the quotient space
where if and only if . One can directly verify that is an equivalence relation.
Lemma 6.4.
For any , there exists at most one such that
(6.2) |
Proof.
Suppose there are two vectors and in satisfying (6.2) such that . Then
But since , . This is a contradiction. ∎
Definition 6.5.
Let . For any , we say that exists if satisfies (6.2) in the place of . By convention, we set if it does not exist. Define the function
(6.3) |
Remark 6.6.
Lemma 6.7.
is a well-defined function on . Moreover, is lower semi-continuous.
Proof.
Take any and any .
If exists, then . Note that
Thus exists and .
If does not exist, same argument as above shows that does not exist neither.
If exists, it is locally constant. Therefore, is lower semi-continuous. ∎
Let be a number satisfying Proposition 5.1. Let and be a sufficiently large number (to be specified later). Define
(6.4) |
We will prove that satisfies properties (1)-(5) of Proposition 6.1.
Proof of Proposition 6.1 (1).
Since
is precompact. As is continuous and is lower semicontinuous, is closed and thus compact. ∎
Proof of Proposition 6.1 (2).
If for some , , then as . Let . By definition of , we have for some . Note that . Hence .
Conversely, if , then by definition of , we have . Choose any such that , then . Therefore, ∎
Notations. Let’s fix some simplified notations for the rest of the proof. We will fix an till the end of this section. In the following, is a sufficiently large number.
-
•
For any , any , denote .
-
•
If exists, denote .
-
•
For any , , , let
We note that if exists, then .
-
•
For any differentiable function , by mean value theorem in several variables, for any , there is a such that
Since the functions that we consider have bounded first derivative on a bounded set, we will omit this for simplicity.
-
•
For , let denote the column vector in with in -th row and elsewhere. Let denote the usual inner product of column vectors.
Lemma 6.8.
Proof.
Proof of Proposition 6.1 (3).
If , by Remark 6.6, we have . Since is continuous and bounded on compact sets, there exists such that for any ,
Therefore,
Assume . If does not exist, then by definition, . Now we suppose that exists.
Case 1: , where .
Choose , by Lemma 6.8, the Lipschity continuity of , and the choices of and , we have
Hence, by the assumption of Case 1,
Case 2: . Then
By construction of , we have . Combining Case 1 and Case 2, we have
Therefore,
For , by Remark 5.2, we have for large enough ,
By the above, let , then
∎
Proof of Proposition 6.1 (4).
If does not exist, then .
To prove property (5) of Proposition 6.1, we record the following lemmas:
Lemma 6.9.
[25, Lemma 4.8] Let and . Let , , where is an interval with for each . Let be a measurable function. Then
(6.5) |
Lemma 6.10.
Let be a measurable function. Suppose that there exists such that for any ,
Then for any , there exists such that
Proof.
This is a direct generalization of [8, Lemma 6.10]. ∎
The following lemma is a special case of [11, Lemma 3.3].
Lemma 6.11.
Let be a bounded open subset of , and let be such that for some constants , one has
where denote the sup norm of a function on . Then for any box (or ball) , any , one has
with .
Lemma 6.12.
Let be an integer. Let be a map. Given , for any , let . Then
where superscript denotes the transpose of the vector.
Proof.
Since , . Using chain rule, it can be verified that
As , we have . This proves the lemma. ∎
Lemma 6.13.
Let be an integer. Given real numbers , and a partition of such that . There is (depending only on the partition) such that the following holds: For any vector (here superscript denote the transpose of the corresponding vector) satisfying
-
•
For any , ;
-
•
For any , ;
-
•
For any , ;
-
•
For any , .
If we denote , then for any ,
Proof.
If , then the lemma is trivial. Now we assume that . Let be such that , then . Without loss of generality, we may assume that and . Choose a satisfying
-
•
, for ;
-
•
if , or ;
-
•
if and .
Note that for any , . Therefore, for any , we have the lower bound
where in the second inequality we apply Cauchy-Schwartz inequality. Also for any , we have the upper bound
On the other hand, for any , we have . Therefore, for any ,
∎
Roughly speaking, Lemma 6.13 says that one can find a suitable rotation depending only on the partition of such that for any vector , as long as there is a coordinate of with large enough absolute value, the absolute value of all coordinates of the new vector are bounded below by a suitable constant.
Proof of (5) of Proposition 6.1.
Now we assume and let be a sufficiently large number (to be specified later). By Lemma 6.9,
By Proposition 5.1, for sufficiently large, there exists such that for any ,
(6.6) |
Note that since
by definition of , it remains to estimate the following integral for any :
(6.7) |
Define for any ,
Since for , is dominated by , by Lemmas 6.14, 6.15 given as follows, we have
As we choose , recall that , we have
This finishes the proof of property (5) of Proposition 6.1, modulo Lemmas 6.14, 6.15.
∎
Lemma 6.14.
Let be the box as in Proposition 6.1 (5). There is sufficiently large such that for any ,
Proof.
We will prove that for sufficiently large, for any ,
For , denote .
Case 1: , where .
By definition of (cf.(2.1)), the choice of and Lipschity continuity of , we have for any ,
By Lemma 6.8, the choices of the sidelength of the box and ,
Choose large enough such that , we obtain
Case 2: .
Then by the choice of , we have
Therefore,
Combining cases 1 and 2, the lemma is proven. ∎
Lemma 6.15.
There exists sufficiently large such that for any ,
Proof.
We fix for the rest of the proof. For , denote . Since , for simplicity we denote .
Case 1: , where .
Case 2: .
Recall that . For any , we have , where is the unit ball in . We may choose large enough such that for any , any , we have , where is given as in Remark 6.3. Define a function on by
Note that .
We will apply Lemma 6.13 to find such that after the change of basis , for
(6.8) |
we have
(6.9) |
Applying Lemma 6.11 to , since preserves Lebesgue measure, we obtain that for any ,
where is a constant depending only on and . Choose large enough, by Lemma 6.10, with ,
This prove the lemma. Therefore, it remains to achieve (6.9). Consider the function , where . We have
(6.10) |
Let , define the partition of by
Note that by definition, . Using (6.10), the choice of , and the estimate
the following holds for any :
-
•
For any , where , ,
-
•
For any ,
-
•
For any ,
-
•
For any ,
Applying Lemma 6.13 with , and given as above, we obtain (Since the partition depends only on , depends only on ), such that after the change of basis , the vector satisfies the following: For any ,
(6.11) |
By the assumption of Case 2, and the choice of , it is elementary to verify that
(6.12) |
Moreover, using the expression (6.10), we obtain
(6.13) |
Also, note that
(6.14) |
Now we choose as in (6), by (6.11)(6)(6.13)(6.14), (6.9) is achieved. This finishes the proof of the lemma. ∎
7. Proof of Proposition 2.1
Following a general strategy developed in [8, Section 6.6], we derive Proposition 2.1 from Proposition 6.1.
Let be a locally compact, second countable Hausdorff topological space. Let be a compact box in . Let be a continuous map. Let be a continuous map and we write as for .
Let . For every , let be a partition of elements in into countably many subboxes with positive Lebesgue measure. By construction, is a filtration. For any , let denote the atom in containing .
Let be a measurable map. Assume that satisfies the following conditions:
(1) satisfies contraction hypothesis: There exist and such that for any and any atom in ,
(7.1) |
(2) satisfies Lipschitz property: There exists a constant such that for any , any , and any ,
(7.2) |
(3) is bounded on , that is, there exists such that
(7.3) |
For any and a measurable subset of , define
where is the indicator function of .
Lemma 7.1.
[8, Lemma 6.20] For any , there exist and such that for , and any ,
proof of Proposition 2.1.
Recall that is a closed cube in . We may assume that is large enough such that is less than the length of each side of .
8. Proof of variants of Theorem 1.3
Given , we may write
(8.1) |
where , , , . For , we can write
(8.2) |
Since , it is clear that the set of such that is a proper algebraic subvariety of and hence, it has Lebesgue measure zero.
Therefore we can assume that , and
We may write
where
Lemma 8.1.
For a.e. , there is an open neighborhood of contained in and an open subset of , such that the map defined by
(8.3) |
is a diffeomorphism.
Proof.
For any such that , there is a neighborhood of , for any , the map is well defined and differentiable. Therefore is an open subset. Let
for any such that . Now we verify that is the inverse of , that is, we need to verify that for ,
(8.4) |
Note that as , we have Therefore, left hand side of (8.4) is , which is equal to the right hand side of (8.4). ∎
Proof of Theorem 1.5.
Choose satisfying Lemma 8.1. By Lemma 2.3, it suffices to prove that for a.e. , the point
(8.5) |
is Birkhoff generic with respect to . For , define
Applying Corollary 1.4 to for , we obtain that if for any ,
(8.6) |
then for a.e. , is Birkhoff generic with respect to .
Proof of Corollary 1.8.
Proof of Corollary 1.9.
Fix an at which the map has a nonsingular differential. It is enough to prove the corollary for a.e. in a neighborhood of . Choose a neighborhood of such that for any , as in (8.1) we can write
where and . This can be done by smoothness of .
Since , , that is,
In particular, we have
We may write
where
By Lemma 2.3, for any , is Birkhoff generic with respect to if and only if
is Birkhoff generic with respect to .
By assumption, the map
has nonsingular differential at . Thus the map
(8.10) |
also has nonsingular differential at .
Shrink the neighborhood of if necessary, we can assume that there exists an open subset of such that is a diffeomorphism. Denote the inverse of .
9. Application to universal hitting time statistics for integrable flows
9.1. An adapted form of Corollary 1.9
Following notations of [3], for , let
Theorem 9.1.
Let be a bounded open subset of and be a map. Let be a smooth map such that the map has a nonsingular differential at Lebesgue almost every . Assume that for any ,
Then for Lebesgue a.e. , is Birkhoff generic with respect to .
Proof.
For any , denote . Choose such that for any
Note that since ,
Applying Corollary 1.9, we obtain that for a.e. , is Birkhoff generic with respect to , and the theorem follows. ∎
9.2. Universal hitting time
Let be a measurable space with probability measure , and be a measure-preserving dynamical system. Given some target set , it is natural to study how often a -trajectory along random initial data intersects this target set. On the other hand, another question is to consider a sequence of randomized target sets whose "size" shrink to zero, and study the distribution of intersection times of a random -trajectory with these shrinking targets. The interested reader is referred to [3] and the references therein for a survey of the history of aforementioned questions.
In the setting of universal hitting time statistics for integrable flows (cf. [3]), the above question is studied when the measurable space is a dimensional torus and is a linear flow on the torus. Now let and be fixed integers. In this article, the sequence of target sets we consider is a sequence of union of many bounded codimensional one balls in , whose radius shrink to zero. More precisely, let be a bounded open subset of . Consider the following smooth functions:
where is the unit one sphere in . For the functions above, we assign to any the following:
-
•
initial position of the flow;
-
•
direction of the flow;
-
•
direction of the -th target ball;
-
•
center of the -th target ball.
With these functions, we define the flow
(9.1) |
From now on, we fix a map from to such that for all ,
(9.2) |
and is smooth on for a singular point . For , fix a bounded open subset . For any , denote the -level target set with to be
where
Note that parameterizes the size of the target. For any , let be the set of hitting times defined by
This is a discrete subset of , and we label its elements by
By Santalo’s formula (cf. [2]), if is such that the components of are not rationally related, then for any , the normalized -th return time to target is
where is the mean return time (cf. [3, Section 2]), and
Definition 9.2.
The smooth map is regular if the push forward of Lebesgue measure on under is absolutely continuous with respect to the Haar measure on .
The following is a corollary of Theorem 9.1:
Corollary 9.3.
Let be a bounded open subset of . Let be a regular smooth map. Let be a map. If for any ,
(9.3) |
then for Lebesgue a.e. , is Birkhoff generic with respect to .
Proof.
By assumption, is a regular smooth map from to . By Sard’s theorem, the set of critical points of has Lebesgue measure . For any which is not a critical point of , there exists an open neighborhood of such that is a diffeomorphism of to some open subset of .
Therefore, the map has nonsingular differentials for all . Then we apply Theorem 9.1 to and the corollary follows. ∎
Definition 9.4.
The -tuple of smooth functions is -generic if for any , we have
To state our theorem precisely, we need some preparations. Our notations follows from [3, Section 6]. Given , denote . For and , define . Let be the matrix of the linear transformation
(9.4) |
where for .
For any , we define
(9.5) |
Let . For and , write . Our main theorem of this section is the following.
Theorem 9.5.
Let be a bounded open subset of . For , let be given as in the beginning of this section. Let be a bounded open subset of . For each , assume that
(1) ,
(2) for all ,
(3) for a.e. , the boundary has Lebesgue measure .
(4) is a smooth positive function of .
Also assume that is regular and is -generic. Then for any , any for , the following holds: For a.e. ,
Note that in [3, Theorem 2], the authors proved that for each , there is some random variable in such that the -th normalized hitting time converges to in distribution as . Unlike [3], in Theorem 9.5 we are interested in the question that given a fixed initial , when the target is shrinking, how often the -th normalized hitting time is bounded by some given constant.
For any , any real numbers , following [3, Eq. (8.9)], we define
Given any real numbers for , following [3, Eq.(8.10)], we define
where
For any , denote
Lemma 9.6.
[3, Lemma 17] For every , and , .
Proof.
By [3, Lemma 14,16], it suffices to prove that for every and , has Lebesgue measure zero. Now since for any , we have
By assumption, for a.e. , , the lemma follows. ∎
We are now ready to prove Theorem 9.5.
Proof of Theorem 9.5.
The proof of Theorem 9.5 is almost the same as the proof of [3, Theorem 2], except that here we use the equidistribution result for the average along trajectory, while in [3], the equidistribution of translation of the average over a bounded open subset in horospherical subgroup is used.
Acknowledgment. I would like to thank Yitwah Cheung, Nimish Shah, Ronggang Shi and Barak Weiss for helpful discussions and comments. I thank Jim Cogdell for pointing out some inaccuracies in the draft. I also thank anonymous referee for pointing out many inaccuracies of the paper, as well as providing a lot of helpful suggestions to improve this article.
References
- [1] Yves Benoist and Jean-Francois Quint. Random walks on finite volume homogeneous spaces. Invent. Math., 187(1):37–59, 2012. MR 2874934
- [2] N. Chernov. Entropy, Lyapunov exponents, and mean free path for billiards. J. Statist. Phys., 88(1-2):1–29, 1997. MR MR1468377
- [3] Carl P. Dettmann, Jens Marklof, and Andreas Strömbergsson. Universal hitting time statistics for integrable flows. J. Stat. Phys., 166(3-4):714–749, 2017. MR MR3607587
- [4] Manfred Einsiedler and Thomas Ward. Ergodic theory with a view towards number theory, volume 259 of Graduate Texts in Mathematics. Springer-Verlag London, Ltd., London, 2011. MR 2723325
- [5] Alex Eskin and Gregory Margulis. Recurrence properties of random walks on finite volume homogeneous manifolds. In Random walks and geometry, pages 431–444. Walter de Gruyter, Berlin, 2004. MR 2087794
- [6] Alex Eskin, Gregory Margulis, and Shahar Mozes. Upper bounds and asymptotics in a quantitative version of the Oppenheim conjecture. Ann. of Math. (2), 147(1):93–141, 1998. MR 1609447
- [7] Alex Eskin, Gregory Margulis, and Shahar Mozes. Quadratic forms of signature and eigenvalue spacings on rectangular 2-tori. Ann. of Math. (2), 161(2):679–725, 2005. MR 2153398
- [8] Krzysztof Frączek, Ronggang Shi, and Corinna Ulcigrai. Genericity on curves and applications: pseudo-integrable billiards, Eaton lenses and gap distributions. J. Mod. Dyn., 12:55–122, 2018. MR 3808209
- [9] Michael Hochman and Pablo Shmerkin. Equidistribution from fractal measures. Invent. Math., 202(1):427–479, 2015. MR 3402802
- [10] Bernard Host. Nombres normaux, entropie, translations. Israel J. Math., 91(1-3):419–428, 1995. MR 1348326
- [11] D. Y. Kleinbock and G. A. Margulis. Flows on homogeneous spaces and Diophantine approximation on manifolds. Ann. of Math. (2), 148(1):339–360, 1998. MR 1652916
- [12] Dmitry Kleinbock, Nicolas de Saxcé, Nimish A Shah, and Pengyu Yang. Equidistribution in the space of 3-lattices and dirichlet-improvable vectors on planar lines. arXiv preprint arXiv:2106.08860, 2021.
- [13] Dmitry Kleinbock, Ronggang Shi, and Barak Weiss. Pointwise equidistribution with an error rate and with respect to unbounded functions. Math. Ann., 367(1-2):857–879, 2017. MR 3606456
- [14] A. Malcev. On the representation of an algebra as a direct sum of the radical and a semi-simple subalgebra. C. R. (Doklady) Acad. Sci. URSS (N.S.), 36:42–45, 1942. MR 0007397
- [15] Gregory Margulis. Problems and conjectures in rigidity theory. In Mathematics: frontiers and perspectives, pages 161–174. Amer. Math. Soc., Providence, RI, 2000. MR 1754775
- [16] Gregory Margulis and Amir Mohammadi. Quantitative version of the Oppenheim conjecture for inhomogeneous quadratic forms. Duke Math. J., 158(1):121–160, 2011. MR 2794370
- [17] Shahar Mozes. Epimorphic subgroups and invariant measures. Ergodic Theory Dynam. Systems, 15(6):1207–1210, 1995. MR 1366316
- [18] M. S. Raghunathan. Discrete subgroups of Lie groups. Springer-Verlag, New York-Heidelberg, 1972. Ergebnisse der Mathematik und ihrer Grenzgebiete, Band 68. MR 0507234
- [19] Marina Ratner. On Raghunathan’s measure conjecture. Ann. of Math. (2), 134(3):545–607, 1991. MR 1135878
- [20] Marina Ratner. Raghunathan’s topological conjecture and distributions of unipotent flows. Duke Math. J., 63(1):235–280, 1991. MR 1106945
- [21] Nimish A. Shah. Limit distributions of expanding translates of certain orbits on homogeneous spaces. Proc. Indian Acad. Sci. Math. Sci., 106(2):105–125, 1996. MR 1403756
- [22] Nimish A. Shah. Equidistribution of expanding translates of curves and Dirichlet’s theorem on Diophantine approximation. Invent. Math., 177(3):509–532, 2009. MR 2534098
- [23] Nimish A. Shah. Limiting distributions of curves under geodesic flow on hyperbolic manifolds. Duke Math. J., 148(2):251–279, 2009. MR 2524496
- [24] Ronggang Shi. Expanding cone and applications to homogeneous dynamics. International Mathematics Research Notices, 2015. MR 4105521
- [25] Ronggang Shi. Pointwise equidistribution for one parameter diagonalizable group action on homogeneous space. Trans. Amer. Math. Soc., 373(6):4189–4221, 2020. MR 4251297
- [26] Ronggang Shi and Barak Weiss. Invariant measures for solvable groups and Diophantine approximation. Israel J. Math., 219(1):479–505, 2017. MR 3642031
- [27] George Tomanov. Actions of maximal tori on homogeneous spaces. In Rigidity in dynamics and geometry (Cambridge, 2000), pages 407–424. Springer, Berlin, 2002. MR 1919414
- [28] Pengyu Yang. Equidistribution of expanding translates of curves and Diophantine approximation on matrices. Invent. Math., 220(3):909–948, 2020. MR 4094972
- [29] Jon Chaika and Alex Eskin. Every flat surface is Birkhoff and Oseledets generic in almost every direction. J. Mod. Dyn., 9:1–23, 2015.