1 Introduction
By a general result of Green and Tao [Reference Green and TaoGT12], all orbits of Diophantine flows on any nilmanifold become equidistributed at polynomial speed. Their approach is an extension of Weyl’s method, based on induction on the number of steps, but the rate of decay in their theorem is not explicit and presumably far from optimal. Flaminio and Forni also established estimates for the quadratic polynomial speed of equidistribution of nilflows on higher-step nilmanifolds [Reference Flaminio and ForniFF14] called quasi-abelian (filiform). This is the simplest class of nilmanifolds of arbitrarily higher-step structures, and it has an application in proving the bound of an exponential sum called the Weyl sum. It is notable that the bound obtained from Flaminio and Forni is established only almost everywhere but comparable with the results by Wooley [Reference WooleyT15] from number theory (see also [Reference Bourgain, Demeter and GuthBDG15]).
In this paper we extend the result for quasi-abelian to a non-renormalizable class of nilflows on higher-step nilmanifolds under Diophantine conditions on the frequencies of their toral projections (see Definition 5.15).
For any $\alpha = (\alpha _1,\ldots ,\alpha _n) \in \mathbb {R}^n$ , for any $N \in \mathbb {N}$ and every $\delta>0$ , let
For every $\nu>1 $ , let $D_n(\nu ) \subset (\mathbb {R}\backslash \mathbb {Q})^n$ be the subset defined as follows: the vector $\alpha \in D_n(\nu )$ if and only if there exists a constant $C(\alpha )>0$ such that, for all $N \in \mathbb {N}$ for all $\delta>0$ ,
The Diophantine condition $D_n(\nu )$ contains the set of simultaneously Diophantine vectors so that $D_n(\nu )$ has a full measure for sufficiently large $\nu \geq 1$ .
For a set of generator $\mathfrak {G}_\alpha $ of $\mathfrak {n}$ , there exist an element $X_\alpha \in \mathfrak {G}_\alpha $ and a codimension 1 ideal $\mathfrak {I}$ such that $X_\alpha \notin \mathfrak {I} \subset \mathfrak {n}$ . We assume that the Lie algebra satisfies the transversality condition if
where $C_{\mathfrak {I}}(X_{\alpha }) = \{Y \in \mathfrak {I} \mid [Y,X_{\alpha }] = 0\}$ is the centralizer.
Under this hypothesis, the rate of convergence for ergodic average of nilflows $(\phi _{X_{\alpha }}^t)$ under Diophantine conditions $\alpha \in D_n(\nu )$ is polynomial for almost all points, as a function of step size and total number of elements of Lie algebras.
Theorem 1.1. Let $(\phi _{X_{\alpha }}^t)$ be a nilflow on a k-step nilmanifold M on $n+1$ generators such that the projected toral flow $(\bar \phi _{X_{\alpha }}^t)$ is a linear flow with frequency vector $\alpha := (1,\alpha _1, \ldots , \alpha _{n} ) \in \mathbb {R} \times \mathbb {R}^{n}$ . Assume that the Lie algebra satisfies the transversality condition and $\alpha \in D_n(\nu )$ for some $1\leq \nu \leq {k}/{2}$ . Then there exists a Sobolev norm $\|\cdot \|$ on the space $C^\infty (M)$ of smooth function on M and for every $\epsilon>0$ there exists a positive measurable function $K_{\epsilon } \in L^p(M) $ for all $p \in [1,2)$ , such that the following bound holds. For every smooth zero-average function $f \in C^\infty (M)$ , for every $T \geq 1$ , for almost all $x \in M$ ,
where $S_{\mathfrak {n}}(k)$ is a higher order polynomial determined by the structure of $\mathfrak {n}$ . Specifically, if $n_i$ is the number of elements in $\mathfrak {n}$ with step size i,
In the general higher-step nilmanifold, no renormalization for nilflows is known. Instead, based on the theory of unitary representations for the nilpotent Lie group (Kirillov theory), it is possible to choose a proper scaling operator on the space of invariant distributions. Compared to earlier work on the quasi-abelian case [Reference Flaminio and ForniFF14], the main novelty of our results lies in generalization of the scaling method to the general Lie algebra satisfying a transversality condition.
The transversality condition enables the measure estimate (§5) for the return orbit. This condition is sufficient, and in principle there are no obstructions to a generalization to arbitrary nilflows with Diophantine frequencies and all points $x \in M$ , except that this would require new approaches to estimation other than a Borel–Cantelli type argument. On the other hand, the necessity of the condition explains that the total number of elements in the basis cannot grow too fast as the step size gets larger: it grows almost linearly in the number of steps and generators.
We can view this phenomenon in the following way. If the growth of the number of elements in lower steps (generated by basis) is too large, then it lacks the dimensions to count the measure of the return orbit on the transverse manifold. For instance, we observe this phenomenon in free nilpotent Lie algebras. Even a small number of generators creates a large number of elements in the lower level under small steps of commutations, which behave in a completely different way than strictly triangular and quasi-abelian. We present such an example in the appendix to motivate our condition.
The above theorem is supplemented by its corollary on the strictly triangular nilmanifold $M_{k}^{(k)}$ . Let $N_{k}^{(k)}$ denote a step-k nilpotent Lie group on k generators. Up to isomorphism, $N_{k}^{(k)}$ is the group of upper triangular unipotent matrices
with one-dimensional center. The next result states that the rate of equidistribution for nilflows on the triangular nilmanifold $M_{k}^{(k)}$ decays at a polynomial speed with exponent, and cubically as a function of number of steps.
Corollary 1.2. Let $(\phi _{X_\alpha }^t)$ be a nilflow on k-step strictly triangular nilmanifold M on k generators such that the projected toral flow $(\bar \phi _{X_\alpha }^t)$ is a linear flow with frequency vector $\alpha := (1,\alpha _1, \ldots , \alpha _{k-1} ) \in \mathbb {R} \times \mathbb {R}^{k-1}$ . Under the condition that $\alpha \in D_n(\nu )$ for some $1\leq \nu \leq {k}/{2}$ , there exists a Sobolev norm $\|\cdot \|$ on the space $C^\infty (M_k^{(k)})$ of smooth function on $M_k^{(k)}$ and there exists a positive measurable function $K_\epsilon \in L^p(M_k^{(k)}) $ for all $p \in [1,2)$ and for every $\epsilon>0$ , such that the following bound holds. For every smooth zero-average function $f \in C^\infty (M_k^{(k)})$ , for almost all $x \in M$ and for every $T \geq 1$ ,
We also establish the uniform bound for the step-3 strictly triangular nilmanifold case. The result holds for all points by counting arguments for linearly divergent close return orbits under Roth-type Diophantine condition. The step-3 case (as well as the filiform case [Reference ForniF16]) is a good example for deriving a simplified proof.
Theorem 1.3. Let $(\phi _X^t)$ be a nilflow on the 3-step nilmanifold M on three generators such that the projected toral flow $(\bar \phi _X^t)$ is a linear flow with frequency vector $v:= (1,\alpha ,\beta )$ of Diophantine condition with exponent $\nu = 1+\epsilon $ for all $\epsilon>0$ . For every $s> 26$ , there exists a constant $C_s$ such that for every zero-average function $f \in W^s(M)$ , for all $(x,T) \in M \times \mathbb {R}$ , we have
In the last section, we present exponential mixing of hyperbolic nilautomorphisms as a main application. Exponential mixing of ergodic automorphisms and its applications to the central limit theorem on compact nilmanifolds was proven by Gorodnik and Spatzier [Reference Gorodnik and SpatzierGS14]. Their approach was based on the result of Green and Tao [Reference Green and TaoGT12], and mixing follows from the equidistribution of the exponential map called box map satisfying certain Diophantine conditions. Our result also shows specific exponent of exponential mixing depending on the structure of nilmanifolds, which follows from equidistribution results and a renormalization argument on partial hyperbolic automorphisms (hyperbolic on the projected torus). However, they are limited to a special class of nilautomorphisms due to lack of hyperbolicity on the group of automorphisms on general nilpotent Lie algebras (cf. triangular step 3 with three generators).
Our work leaves open questions.
Problem 1. Can we find uniform bounds on the average width under the transversality condition? That is, can we prove Theorem 1.1 with uniform bound for all points?
Problem 2. Give an effective bound on ergodic averages on any nilmanifold.
It is still conjectured that the result may hold for all points on any nilmanifolds. However, our argument for averaged width can not be improved to control the slow growth of displacement for the return orbit on general higher-step cases.
This paper is organized as follows. In §2 we define structures of nilmanifolds and nilflows. In §3 we find Sobolev estimates on solutions of the cohomological equation and on invariant distributions as an application of Kirillov theory of unitary representations of nilpotent groups. In §4 we introduce the notion of average width and prove a Sobolev trace theorem. In §5 we prove an effective equidistribution theorem for good points by a Borel–Cantelli argument. In §6 we prove bounds on the average width of an orbit segment of nilflows by gluing all the irreducible representations. In §7 we introduce a uniform width estimate under a Roth-type Diophantine condition based on counting return time directly without a good point argument for the width estimate in §5.4. Finally, in §8, as an application, we prove the mixing of nilautomorphisms.
2 Nilflows on higher-step nilmanifolds
In this section we review nilpotent Lie algebras, groups and basic structures. We recall Kirillov theory and representation theory.
2.1 Background on nilpotent Lie groups and Lie algebras
Let N be a connected, simply connected k-step nilpotent Lie group with Lie algebra $\mathfrak {n}$ with $n+1$ generators. Let $\Gamma $ be a cocompact lattice in N. The quotient $N/\Gamma := M$ is then a compact nilmanifold on which the left action of $N$ is given by translations. Denote by $\mu $ the N-invariant measure on M.
For $j = 1,\ldots , k$ , let $\mathfrak {n}_j$ denote the descending central series of $\mathfrak {n}$ :
where $Z(\mathfrak {n})$ is the center of $\mathfrak {n}$ . In this setting, there exists a strong Malcev basis through the filtration $(\mathfrak {n}_{i})_{i=1}^k$ strongly based at the lattice $\Gamma $ (see [Reference Corwin and GreenleafCG90]). That is, given a basis
with $\xi := \xi ^{(1)} \in \mathfrak {n}_1 \backslash \mathfrak {n}_2$ and $\eta _i^{(l)} \in \mathfrak {n}_l \backslash \mathfrak {n}_{l+1} $ for $i = 1,\ldots , n_l$ , we have the following assertions.
-
(1) If we drop the first l elements of the basis, we obtain a basis of a subalgebra $\mathfrak {n}$ of codimension l.
-
(2) For each j, the elements in order $\eta _1^{(j)},\ldots \eta _{n_j}^{(j)},\ldots \eta _1^{(k)},\ldots \eta _{n_k}^{(k)}$ form a basis of an ideal $\mathfrak {n}_j$ of $\mathfrak {n}$ .
-
(3) The lattice $\Gamma $ is generated by $\{x, y_{1}^{(1)}, \ldots , y_{n_k}^{(k)} \}$ with
$$ \begin{align*}x := \exp(\xi),\quad y_{n_i}^{(j)}:= \exp(\eta_{n_i}^{(j)}).\end{align*} $$
For any nilpotent Lie algebra $\mathfrak {n}$ , there exists a codimension- $1$ subalgebra $\mathfrak {I}$ where
Then $\mathfrak {I}$ is an ideal and $[\mathfrak {n},\mathfrak {n}] \subseteq \mathfrak {I}$ (see [Reference HumphreysH73, Ch. 3, p. 12] and [Reference Corwin and GreenleafCG90, Lemma 1.1.8]). For convenience, we write dimension $a = \dim (\mathfrak {I}) = n_1+\cdots + n_k$ and set $n = n_1$ .
Definition 2.1. An adapted basis of the Lie algebra $\mathfrak {n}$ is an ordered basis $(X,Y) := (X,Y_1,\ldots , Y_a) $ of $\mathfrak {n}$ such that $X \notin \mathfrak {I}$ and $Y:= (Y_1, \ldots , Y_a)$ is a basis of $\mathfrak {I}$ .
A strongly adapted basis $(X,Y) := (X,Y_1,\ldots , Y_a) $ is an adapted basis such that the following assertions hold.
-
(1) The system $(X,Y_1, \ldots , Y_n)$ is a system of generators of $\mathfrak {n}$ , hence its projection is a basis of the abelianization $\mathfrak {n}/[\mathfrak {n},\mathfrak {n}]$ of the Lie algebra $\mathfrak {n}$ .
-
(2) The system $(Y_{n+1},\ldots , Y_a)$ is a basis of the ideal $[\mathfrak {n},\mathfrak {n}]$ .
2.2 Nilmanifolds and nilflows
Every nilmanifold M is a fiber bundle over a torus. In fact, the group $N^{ab} = N/[N,N]$ is abelian, connected and simply connected, hence isomorphic to $\mathbb {R}^{n+1}$ , and $\Gamma ^{ab} = \Gamma /[\Gamma , \Gamma ]$ is a lattice in $N^{ab}$ . Thus, we have a natural projection $pr_1: M \rightarrow \mathbb {T}^{n+1}$ .
We introduce two fibrations of the nilmanifold M. Let $M_2 \simeq N_2/\Gamma _2$ with $N_2 = \exp (\mathfrak {n}_2)$ and its lattice $\Gamma _2$ . Then there exists an exact sequence
Another fibration arises from the canonical homomorphism $N \rightarrow N/N' \approx \langle \exp \xi \rangle $ . For $\theta \in \mathbb {T}^1$ , the fiber $M_\theta ^a = {pr_2}^{-1}(\theta )$ is a local section of the nilflow on M:
On the nilmanifold M, the nilflow $\phi ^t_X$ generated by $X \in \mathfrak {n}$ is the flow obtained by the restriction of this action to the one-parameter subgroup $(\exp tX)_{t \in \mathbb {R}}$ of N, with
The projection $\bar {X}$ of X is the generator of a linear flow $\psi _{\bar {X}} := \{\psi ^t_X\}_{t \in \mathbb {R}}$ on $\mathbb {T}^{n+1} \approx \mathbb {R}^{n+1} \backslash \bar {\Gamma }$ defined by
The canonical projection $ {pr_1}:M \rightarrow \mathbb {T}^{n+1}$ intertwines the flows $\phi ^t_X$ and $\psi ^t_{\bar {X}}$ .
We recall the following result.
Theorem 2.2. [Reference Auslander, Green and HahnAGH63]
The following assertions are equivalent.
-
(1) The nilflow $(\phi ^t_X)$ on M is ergodic.
-
(2) The nilflow $(\phi ^t_X)$ on M is uniquely ergodic.
-
(3) The nilflow $(\phi ^t_X)$ on M is minimal.
-
(4) The projected flow $(\psi ^t_{\bar {X}})$ on $\overline M = N^{ab} / \Gamma ^{ab} \simeq \mathbb {T}^{n+1}$ is an irrational linear flow.
Consider the set of indices
Let $\alpha = (\alpha _i^{(j)}) \in \mathbb {R}^{J}$ and $X:= X_{\alpha }$ be a vector field on M defined
and equivalently we write
For $\theta \in \mathbb {T}^1$ , let $M_\theta ^a = {pr_2}^{-1}(\theta )$ denote a fiber over $\theta \in \mathbb {T}^1$ of the fibration $pr_2$ . The transverse section $M_\theta ^a$ of the nilflow $\{\phi ^t_{X_{\alpha }}\}_{t\in \mathbb {R}}$ corresponds to
Lemma 2.3. The flow $(\phi ^t_{X_{\alpha }})_{t\in \mathbb {R}}$ on M is isomorphic to the suspension of its first return map $\Phi _{\alpha , \theta }: M_\theta ^a \rightarrow M_\theta ^a$ . For every $(i,j) \in J$ and $r \geq 1$ , there exists a polynomial $p_{i,r}^{(j)}(\alpha , \mathbf {s})$ for $\mathbf {s} = (s_i)_{i=1}^a \in \mathbb {R}^a$ such that the rth return map $\Phi _{\alpha , \theta }^r$ is written as follows.
On the coordinates of $\ \mathbf {s} = (s^{(j)}_i)$ for $\Gamma \exp (\theta \xi )\exp (\sum _{(i,j)\in J} s^{(j)}_i\eta ^{(j)}_i) \in M_\theta ^a$ ,
and for $r \in \mathbb {N}$ ,
Proof. By (7), we have
By the Baker–Campbell–Hausdorff formula, there exist polynomials $p_i^{(j)}(\alpha , s)$ with
Since $x \in \Gamma $ , we conclude
The formula implies that $t =1$ is a return time of the restriction of the flow to $M_\theta ^a \subset M$ . The formula for $r \in \mathbb {N}$ follows by induction.
2.3 Kirillov theory and classification
Kirillov theory yields the complete classification of irreducible unitary representations of N. All the irreducible unitary representations of nilpotent Lie groups are parametrized by the coadjoint orbits $\mathcal {O} \subset \mathfrak {n}^*$ . A polarizing (or maximal subordinate) subalgebra for $\Lambda \in \mathfrak {n}^*$ is a maximal isotropic subspace $\mathfrak {m} \subset \mathfrak {n}$ which is a subalgebra of $\mathfrak {n}$ . It is well known that for any $\Lambda \in \mathfrak {n}^*$ there exists a polarizing subalgebra $\mathfrak {m}$ for a nilpotent Lie algebra $\mathfrak {n}$ (see [Reference Corwin and GreenleafCG90, Theorem 1.3.3]). Let $\mathfrak {m}$ be a polarizing subalgebra for given linear form $\Lambda \in \mathfrak {n}^*$ . Then the character $\chi _{\Lambda ,\mathfrak {m}}: \exp {\mathfrak {m}}\rightarrow S^1$ is defined by
Given the pair $(\Lambda ,\mathfrak {m})$ , we associate the unitary representation
where the induced representation $\pi _{\Lambda }$ is defined by
These unitary representations are irreducible up to equivalence, and all unitary irreducible representations are obtained in this way. It is known that $\Lambda $ and $\Lambda '$ belong to the same coadjoint orbit if and only if $\pi _{\Lambda ,\mathfrak {m}} $ and $\pi _{\Lambda ',\mathfrak {m}'} $ are unitarily equivalent and $\pi _{\Lambda ,\mathfrak {m}}$ is irreducible whenever $\mathfrak {m}$ is maximal subordinate for $\Lambda $ . We write $\pi _\Lambda \simeq \pi _{\Lambda '}$ if $\Lambda $ and $\Lambda '$ are in the same coadjoint orbit.
Since the action of N on M preserves the measure $\mu $ , we obtain a unitary representation $\pi $ of N. The regular representation of $L^2(M)$ of N decomposes as a countable direct sum (or direct integral) of the irreducible, unitary representation $H_\pi $ , which occurs with at most finite multiplicity
The derived representation $\pi _{*}$ of a unitary representation $\pi $ of N on a Hilbert space $H_\pi $ is a representation of the Lie algebra $\mathfrak {n}$ on $H_\pi $ defined as follows. For every $X \in \mathfrak {n}$ ,
We recall that a vector $v \in H_\pi $ is of $C^\infty $ -vectors in $H_\pi $ for representation $\pi $ if the function $g \in N \mapsto \pi (g)v \in H_\pi $ is of class $C^\infty $ as a function on N with values in a Hilbert space.
Definition 2.4. The space of Schwartz functions on $\mathbb {R}$ with values of $C^\infty $ vectors for the representation $\pi '$ on $H'$ is denoted $\mathcal {S}(\mathbb {R}, C^\infty (H'))$ . It is endowed with the Fréchet topology induced by the family of seminorms
and defined as follows: for all $f \in \mathcal {S}(\mathbb {R}, C^\infty (H'))$ ,
Lemma 2.5. [Reference Flaminio and ForniFF07, Lemma 3.4]
As a topological vector space
where $\mathcal {S}(\mathbb {R}, C^\infty (H'))$ is Schwartz space.
Suppose that $\mathfrak {n} = \mathbb {R} X\oplus \mathfrak {I} $ with its codimension- $1$ ideal $\mathfrak {I}$ , and $N = \mathbb {R} \ltimes N'$ with a normal subgroup $N'$ of N. Let $\pi '$ be a unitary irreducible representation of $N'$ on a Hilbert space $H'$ . Each irreducible representation $H_\pi $ is unitarily equivalent to $L^2(\mathbb {R},H',\,dx)$ , and the derived representation of $\pi _*$ of the induced representation $\pi = \text {Ind}_{N'}^N(\pi ')$ has the following description.
For $f \in L^2(\mathbb {R},H',\,dx)$ , the group $\mathbb {R}$ acts by translations and its representation is polynomial in the variable x. For any $Y \in \mathfrak {n}'$ ,
We define its degree $d_Y \in \mathbb {N}$ with respect to the representation $\pi _{*}(Y)$ to be the degree of the polynomial. Let $(d_1,\ldots , d_a)$ be the degrees of the elements $(Y_1, \ldots , Y_a)$ , respectively. The degree of the representation $\pi $ is defined as the maximum of the degrees of the elements of any basis.
3 The cohomological equation
In this section we prove an a priori Sobolev estimate on Green’s operator for the cohomological equation $Xu = f$ of nilflows with generator X. Then we estimate the bounds of Green’s operator by invariant distributions with rescaled Sobolev norm.
3.1 Distributions and Sobolev space
Let $L^2(M)$ be the space of complex-valued, square integrable functions on M. Given an ordered basis $\mathcal {F}$ of $\mathfrak {n}$ , the transverse Laplace–Beltrami operator is a second-order differential operator defined by
For any $\sigma \geq 0$ , let $|\cdot |_{\sigma ,\mathcal {F}}$ be the transverse Sobolev norm defined as follows: for all functions $f \in C^\infty (M)$ , let
Equivalently,
The completion of $C^\infty (M)$ with respect to the norm $|\cdot |_{\sigma ,\mathcal {F}}$ is denoted $W^\sigma (M,\mathcal {F})$ and the distributional dual space (as a space of functional with values in $H'$ ) to $W^\sigma (M)$ is denoted
We denote by $C^\infty (H_\pi )$ the space of $C^\infty $ vectors of the irreducible unitary representation $\pi $ . Following the notation in (11), let $W^\sigma (H_\pi ) \subset H_\pi $ be the Sobolev space of vectors, endowed with the Hilbert space norm in the maximal domain of the essential self-adjoint operator $(I+\pi _{*}( \Delta _{\mathcal {F}}))^{{\sigma }/{2}}$ . That is to say, for every $f \in C^\infty (H_\pi )$ and $\sigma>0$ ,
where $\pi _{*}(\Delta _{\mathcal {F}})$ is determined by derived representations.
3.2 A priori estimates
The distributional obstruction to the existence of solutions of the cohomological equation
in an irreducible unitary representation $H_\pi $ is the normalized X-invariant distribution.
Definition 3.1. For any $X \in \mathfrak {n}$ , the space of X-invariant distributions for the representation $\pi $ is the space $\mathcal {I}_X(H_\pi )$ of all distributional solutions $D \in D'(H_\pi )$ of the equation $\pi _{*}(X)D = XD = 0$ . Let
be the subspace of invariant distributions of order at most $\sigma \in \mathbb {R}^+$ .
By [Reference Flaminio and ForniFF07], each invariant distribution D has Sobolev order equal to $1/2$ , that is, $D \in W^{-\sigma }(H_\pi )$ for any $\sigma> 1/2$ .
For all $\sigma> 1$ , let $\mathcal {K}^\sigma (H_\pi ) = \{f \in W^\sigma (H_\pi ) \mid D(f) = 0 \in C^\infty (H_\pi ), \text { for any } D \in W^{-\sigma }(H_\pi ) \}$ be the kernel of the X-invariant distribution on the Sobolev space $W^\sigma (H_\pi )$ . Green’s operator $G_X:C^\infty (H_\pi ) \rightarrow C^\infty (H_\pi )$ with
is well defined on the kernel of distribution $\mathcal {K}^\infty (H_\pi )$ on $C^\infty (H_\pi )$ . In fact, for any $f \in \mathcal {K}^\infty (H_\pi )$ , we have $\int _{\mathbb {R}}f(t)\,dt = 0 \in C^\infty (H')$ and
Now we define the generalized (complex-valued) invariant distribution on smooth vector $C^\infty (H_\pi )$ .
Lemma 3.2. The invariant distribution is generalized in the following sense. For every invariant distribution D, there exists a linear functional $\ell : C^\infty (H_\pi ) \rightarrow \mathbb {C}$ such that for every function $f \in C^\infty (H_\pi ) \subset L^2(\mathbb {R}, H')$ ,
Furthermore, $\ell \in W^{-s}(H_\pi )$ for $s> 1/2$ and
Proof. We construct a linear functional $\ell $ in (14) as follows. Let $\chi \in C_0^\infty (\mathbb {R})$ be a smooth function with compact support with unit integral over $\mathbb {R}$ . Given an invariant distribution $D \in \mathcal {I}_X(H_\pi )$ , let us define $\ell (v) = D (f_v)$ for $f_v = \chi v$ and $v\in C^\infty (H')$ .
Firstly, we prove that $\ell $ is well defined. Let $\chi _1 \neq \chi _2 \in C_0^\infty (\mathbb {R})$ be functions with compact support such that $\int _{\mathbb {R}} \chi _1(t)\,dt = \int _{\mathbb {R}} \chi _2(t)\,dt = 1$ . Note that there exists $\psi \in C_0^\infty (\mathbb {R})$ such that $\chi _1 - \chi _2 = \psi '$ with $\psi (t) = \int _{-\infty }^t (\chi _1(x) - \chi _2(x))\,dx.$
Then we have
and $\chi _1(t) v - \chi _2(t) v$ is an X-coboundary for every $v \in C^\infty (H')$ . Hence, $D(\chi _1(t) v - \chi _2(t) v) = 0$ , which implies that $D(\chi _1(t)v) = D(\chi _2(t)v)$ . Therefore, $\ell (v)$ does not depend on the choice of $\chi $ and the functional $\ell $ is well defined.
Next, we verify that $\ell $ is a distribution on $C^\infty (H')$ . It suffices to prove that $\ell $ is bounded and continuous. For $v \in C^\infty (H')$ and $s> 1/2$ ,
In the representation, $\pi _*(Y_i)$ acts as multiplication of the polynomial $p_i(t)$ on $L^2(\mathbb {R},H')$ . By definition of the Sobolev norm in representation, there exists a non-zero constant
such that
Hence, for $\ell \in W^{-s}(H')$ ,
Therefore, $\ell $ is continuous on $C^\infty (H')$ .
To prove equality (15), for any $f \in C^\infty (H_\pi )$ , we observe $f(t) - \chi (t)( \int _{\mathbb {R}} f(x)\,dx )$ has zero average, hence it is a coboundary with smooth transfer function. Since distribution D is invariant under translation, we obtain
and
Furthermore,
Since D is an invariant distribution for translation and $\int _{\mathbb {R}} \chi = 1$ , we conclude
Let $\mathcal {O}$ be any coadjoint orbit of maximal rank. For all $(X,Y)\in \mathfrak {n} \times \mathfrak {n}_{k-1}$ and $\Lambda \in \mathcal {O}$ , the skew-symmetric bilinear form
does not depend on the choice of linear form $\Lambda \in \mathcal {O}$ .
Let
Here we recall estimates for Green’s operator.
Lemma 3.3. [Reference Flaminio and ForniFF07, Lemma 2.5]
Let $X \in \mathfrak {n}$ and $Y \in \mathfrak {n}_{k-1}$ be any operator such that $B_\Lambda (X,Y) \neq 0$ . The derived representation $\pi _*$ of the Lie algebra $\mathfrak {n}$ satisfies
Theorem 3.4. [Reference Flaminio and ForniFF07, Theorem 3.6]
Let ${\delta _{\mathcal {O}}}:=\delta _{\mathcal {O}}(X)>0$ , and let $\pi $ be an irreducible representation of $\mathfrak {n}$ on a Hilbert space $H_\pi $ . If $f \in W^s(H_\pi )$ , $s>1$ , and $D(f) = 0$ for all $D \in \mathcal {I}_X(H_\pi )$ , then $G_Xf \in W^r(H_\pi )$ , for all $r < (s-1)/k$ , and there exists a constant $C := C(X,k,r,s)$ such that
3.3 Rescaling method
Definition 3.5. The deformation space of a k-step nilmanifold M is the space $T(M)$ of all adapted bases of the Lie algebra $\mathfrak {n}$ of the group N.
The renormalization dynamics is defined as the action of the diagonal subgroup of the Lie group on the deformation space. Let $\rho : = (\rho _1,\ldots , \rho _a) \in (\mathbb {R}^{+})^a$ be any vector with rescaling condition $\sum _{i=1}^a {\rho _i} = 1$ . Then there exists a one-parameter subgroup $\{A_t^{\rho } \}$ of the Lie group of $SL(a+1,\mathbb {R})$ defined as follows:
The renormalization group $\{A_t^{\rho } \}$ preserves the set of all adapted bases. However, it is not a group of automorphisms of the Lie algebra since generated elements in the center $Z(\mathfrak{n})$ are not preserved in higher steps. Therefore, on higher-step nilmanifolds, the dynamics induced by the renormalization group on the deformation space is trivial (it has no recurrent orbits).
Definition 3.6. Given any adapted basis $\mathcal {F} = (X,Y)$ , the rescaled basis $\mathcal {F}(t) = (X(t),Y(t)) = \{e^tX, e^{-\rho _1t}Y_1, \ldots , e^{-\rho _at}Y_a\}$ of $\mathcal {F}$ is a basis of Lie algebra $\mathfrak {n}$ satisfying (18).
Let $(d_1,\ldots , d_i)$ be the degrees of the elements $(Y_1, \ldots , Y_i)$ , respectively. For any $\rho = (\rho _1, \ldots , \rho _a ) \in \mathbb {R}^a$ , let
Definition 3.7. The scaling factor $\rho = (\rho _1, \ldots ,\rho _a)$ is called homogeneous if the vector $\rho $ is proportional to the vector $d = (d_1,\ldots , d_a) $ of degree of $(Y_1, \ldots ,Y_a)$ . That is, under homogeneous scaling, $\unicode{x3bb} _{\mathcal {F}}(\rho ) = {\rho _i}/{d_i} $ for all $1 \leq i \leq a$ .
For all $i = 1, \ldots , a$ , denote
This is the coefficient appearing in (13). Set
Let $\mathfrak {U}(\mathfrak {n})$ be the enveloping algebras of $\mathfrak {n}$ . The generator $\delta $ is the derivation on $\mathfrak {U}(\mathfrak {n}')$ obtained by extending the derivation $\text {ad}(X)$ of $\mathfrak {n}'$ to $\mathfrak {U}(\mathfrak {n}')$ . From the nilpotency of $\mathfrak {n}$ it follows that for any $L \in \mathfrak {U}(\mathfrak {n}')$ there exists a first integer $[L]$ such that $\delta ^{[L]+1}L = 0$ .
Recall that $\mathfrak {I}$ is a codimension- $1$ ideal of $\mathfrak {n}$ used in §2.3.
Lemma 3.8. For each element $L \in \mathfrak {I}$ with degree $[L] = i$ , there exists $Q_j \in \mathfrak {U}(\mathfrak {n})$ such that $\pi _{*}(L) = \sum _{j=0}^i({1}/{j!})\pi ^{\prime }_{*}(Q_j)x^j$ and $[Q_j] = [L]+1-j$ .
Proof. Firstly, we fix elements X and Y as stated in Lemma 3.3. For convenience, we normalize the constant of $\pi _*(Y)$ by 1. That is, there exists $X, Y \in \mathfrak {n}$ such that
We now replace the expansion of $\pi _{*}(L) := \sum _{j=0}^i \frac {1}{j!} \Lambda _L^{(j)}(\mathcal {F})x^j$ by choosing elements $Q_i$ in the enveloping algebra $\mathfrak {U}(\mathfrak {n})$ .
For the coefficient of top degree, denote $Q_i = ({1}/{i!})\text {ad}_X^i(L) \in \mathfrak {n}$ . For degree $i-1$ , we set $Q_{i-1} = \text {ad}_X^{i-1}(L) - Q_iY \in \mathfrak {U}(\mathfrak {n})$ such that
Repeating this process up to degree 0, there exists $ Q_{l} \in \mathfrak {U}(\mathfrak {n})$ such that for $0 < l <i$ ,
and
Definition 3.9. If A is self-adjoint on a Hilbert space H and $\langle Au,u\rangle \geq 0$ for every $u \in H$ , then A is called positive, denoted by $A \geq 0$ .
Remark 3.10. For two self-adjoint operators A and B, $A \geq B$ if and only if $A - B \geq 0$ . Suppose that A and B are bounded operators and commute. Then $A \geq 0$ and $B \geq 0$ implies that $AB \geq 0$ . By the spectral theorem, there exists a unique, self-adjoint square root $A^{1/2} = \int _{\sigma (A)} x^{1/2} dE_A(x)$ . Since A and B commute, $A^{1/2}$ commutes with B. Then
Also, if $0 \leq A \leq B$ , then $A^2 \leq B^2$ ( $B^2-A^2 = B(B-A) + (B-A) A $ ).
Recall that for any positive operators A and B,
Lemma 3.11. For any $r\geq 1$ and $a \geq 1$ , there exists constant $C(a,r)>0$ such that
Proof. It suffices to prove that there exists $C = C(a,r)>0$ such that
since this implies $\Delta (t)^{4r} \leq C^2(\sum _{i=1}^aY_i(t)^{4r})^2$ by the remark.
We prove (24) by induction. If $r =1$ , then
By (22), for each i and j,
Then there exists $C_0 = (a+1)$ such that
Assume that the statement holds for large r. Then, since $\Delta (t)^{2}$ and $\sum _{i=1}^aY_i(t)^4$ are positive, by (25), there exists $C_1(a,r)$ such that
Note that the following inequality holds: for any $r \geq 1$ ,
This inequality is proved by showing that
Since $Y_i(t)^4, Y_j(t)^4$ are all positive, the last inequality holds.
Then, by (26),
Setting $C_2(a,r) = C_1(a,r)(a+1)^2$ ,
Therefore, induction holds and we finish the proof.
For the cohomological equation $X(t)u = f$ , denote its Green’s operator $G_{X(t)}$ . The following theorem states an estimate for the rescaled version of Theorem 3.4.
Theorem 3.12. For $r>1$ , let $s> r(k+1)+1/4$ . For any $f \in \mathcal {K}^s(M) $ , there exists $C_{r,k,s}>0$ such that the following holds: for any $t> 0$ ,
Proof. First, we estimate the bound on Green’s operator with the Sobolev norm.
By Lemma 3.3, there exists a rescaled operator $Y(t) \in \mathfrak {n}_{k-1}$ with
where $\delta _{\mathcal {O}}(t) =\delta _{\mathcal {O}} e^{-\rho _Y t}>0$ .
By the Cauchy–Schwarz inequality, for any $l \in \mathbb {N}$ ,
If $\alpha>1$ , then by Hölder’s inequality,
For all $\alpha>1$ , we set
By Hölder’s inequality and change of variables for $ x' = {\delta _{\mathcal {O}}(t)}x$ and $s' = {\delta _{\mathcal {O}}(t)}s$ ,
Let $E_x:C^\infty (H_\pi ) \rightarrow C^\infty (H')$ be the linear operator defined by $E_xf = f(x)$ (cf. Lemma 2.5). Then the action of $\pi _{*}(Y(t))$ on $C^\infty (H_\pi )$ can be rewritten as
Let $L(t)$ be a rescaled element of $L \in \mathcal {F}$ . By definition of representation (13) and by Lemma 3.8 for rescaled basis $\mathcal {F}(t)$ , for any $L(t)$ there exists $Q_j(t) \in \mathfrak {U}(\mathfrak {n})$ such that
For Green’s operator $G_{X(t)}$ , since $E_xG_{X(t)} = \int _{-\infty }^{x}E_s \,ds$ for all $x \in \mathbb {R}$ ,
Combining (29) and (32), there exists a constant $C^{\prime }_{\alpha ,j}>0$ such that
Note that by binomial formula, there exists $R_{j}(t) \in \mathfrak {U}(\mathfrak {n})$ such that
In particular, $R_j(t)$ is product of $Q_i(t)$ s and $[R_{j}(t)] = 2r([L]+1)-j$ . Specifically, to compute the transverse Laplacian, here we assume that $L(t) = (Y_i(t))^{4r}$ for each element $Y_i \in \mathcal {F}$ . Then, by (33), for any $\alpha>1$ ,
Therefore, combining with Lemma 3.11,
Then, there exists $C' = C'(a,\alpha ,r,X)> 0$ such that
By interpolation, for all $s> r(k+1)+1/4$ , there exists a constant $C_{r,s}: = C_{r,s}(k,X)>0$ such that
By the choice of Y, we obtain $\rho _Y = \unicode{x3bb} _{\mathcal {F}}$ , which finishes the proof.
3.4 Scaling of invariant distributions
We introduce the Lyapunov norm and compare bounds between the Sobolev dual norm and the Sobolev Lyapunov norm of invariant distribution in every irreducible, unitary representation.
For all $t \in \mathbb {R}$ and $\unicode{x3bb} := \unicode{x3bb} _{\mathcal {F}}(\rho )$ defined in (19), let the operator $U_t: L^2(\mathbb {R},H') \rightarrow L^2(\mathbb {R},H')$ be the unitary operator defined as follows:
We will compare the norm estimate of invariant distributions with respect to scaled basis
by the unscaled norm $|D|_{-r,\mathcal {F}}$ .
Theorem 3.13. For $r \geq 1$ and $s> r(k+1)$ , there exists a constant $C_{r,s}>0$ such that for all $t>0$ , the following bound holds:
Proof. Assume the same hypothesis for $L \in \mathcal {F}$ and $L(t)$ in the proof of Theorem 3.12. By Lemma 3.8, there exists an $(i-j+1)$ th-order $Q_j \in \mathfrak {U}(\mathfrak {n})$ with
Then there exists $C>0$ such that
Since $[L] \leq k$ , by unitarity
Hence, for any $s> r(k+1)$ ,
Theorem 3.14. For $r \geq 1$ and $s> r(k+1)$ , there exists a constant $C_{r,s}>0$ such that for any $\unicode{x3bb}>0$ and $t>0$ , the invariant distribution defined in (14) satisfies
Proof. Recall the functional $\ell $ defined in the Lemma 3.2. For $f \in C^\infty (H_\pi )$ ,
Then, by unitarity (37),
and
Definition 3.15. (Lyapunov norm)
For any basis $\mathcal {F}$ and all $\sigma> 1/2$ , define the Lyapunov norm
The following lemma is immediate from the definition of the norm.
Lemma 3.16. [Reference Flaminio and ForniFF14, Lemma 4.15]
For all $t \geq 0$ and $\sigma> 1/2$ , we have
We conclude this section by introducing a useful inequality that follows from the Theorem 3.14:
4 A Sobolev trace theorem
In this section the notion of the average width (40), which is an average measure of close returns along an orbit, is introduced and we prove a Sobolev trace theorem for nilpotent orbits. According to this theorem, the uniform norm of an ergodic integral is bounded in terms of the average width of the orbit segment times the transverse Sobolev norms of the function, with respect to a given basis of the Lie algebra.
4.1 Sobolev a priori bounds
Assume $\mathcal {F}(t) = (X(t),Y(t))$ is a rescaled basis. For any $x \in M$ , let $\phi _{x,t} : \mathbb {R} \times \mathbb {R}^a \rightarrow M$ be the local embedding defined by
Lemma 4.1. For any $x \in M$ , $t \geq 0$ , and $f \in C^{\infty }(M)$ , we have
where q is polynomial in s of degree at most $k-1$ and $|q_l(\mathrm{\mathbf{s}},t)| \leq |q_l(\mathrm{\mathbf{s}},0)|$ for all $t \geq 0$ .
Proof. For $h \in \mathbb{R}$ and $1 \leq i \leq a$ , let $\textbf {s}+h_i$ denote a sequence with $(\textbf {s}+h_i)_i = s_i +h$ and $(\textbf {s}+h_i)_j = s_j$ if $i \neq j$ . By definition,
and we plan to rewrite $f\circ \phi _{x,t}(\tau ,\textbf {s}+h_i)$ in a suitable manner for differentiation.
Fix $i$ . Then for $j >i$ ,
By the Baker–Campbell–Hausdorff formula, we set
Choose $j = i+1$ and observe that all the terms of h are combined into an exponential. Iteratively, we will repeat this process from $j = i+1$ to a until all the terms of h are pushed back. That is, we conclude
For convenience, we write coefficient function $q_l(\textbf{s},t)$ in polynomial degree at most k for $\textbf{s}$ such that
We conclude the proof by choosing $S_i = Y_{i}(t)+\sum _{l>i}^aq_l(\textbf{s},t)Y_{l}(t)$ . Also, commutation between rescaled elements yields $[Y_i(t),s_{j}Y_{j}(t)] = s_{j}e^{-\rho_{ij} t}Y_{k}(t)$ for some $\rho_{ij} >0$ . This implies that the term $q_l(\textbf{s},t)$ for each $l >i$ includes exponential terms with negative exponent so that it decreases for $t \geq 0$ .
Let $\triangle _{\mathbb {R}^a}$ be the Laplacian operator on $\mathbb {R}^a$ given by
Given an open set $O \subset \mathbb {R}^a$ containing the origin, let $\mathcal {R}_O$ be the family of all a-dimensional symmetric rectangles $R \subset [-(1/2), 1/2]^a \cap O$ that are centered at the origin. The inner width of the set $O \subset \mathbb {R}^a$ is the positive number
where Leb is Lebesgue measure on R. The width function of a set $\Omega \subset \mathbb {R} \times \mathbb {R}^a$ containing the line $\mathbb {R} \times \{0\}$ is the function $w_\Omega : \mathbb {R} \rightarrow [0,1]$ defined as follows:
Definition 4.2. Consider the family $\mathcal {O}_{x,t,T}$ of open sets $\Omega \subset \mathbb {R} \times \mathbb {R}^a$ satisfying the two conditions
and $\phi _{x,t}$ is injective on the open set $\Omega \subset \mathbb {R}^a$ . The average width of the orbit segment of the rescaled nilflow $\{\phi _{x,t}(\tau ,0) \mid 0 \leq t \leq T\}$ ,
is a positive number.
The following lemma is derived from the standard Sobolev embedding theorem under the rescaling argument.
Lemma 4.3. [Reference Flaminio and ForniFF14, Lemma 3.7]
Let $I \subset \mathbb {R}$ be an interval, and let $\Omega \subset \mathbb {R} \times \mathbb {R}^a$ be a Borel set containing the segment $I \times \{0\} \subset \mathbb {R} \times \mathbb {R}^a$ . For every $\sigma>a/2$ , there is a constant $C_s>0$ such that for all functions $F \in C^{\infty }(\Omega )$ and all $\tau \in I$ , we have
The following theorem indicates the bound of the ergodic average of the scaled nilflow $\phi _{X(t)}^\tau $ with width function on general nilmanifolds (see also [Reference Flaminio, Forni and TanisFFT16, Theorem 5.2] for twisted horocycle flows).
Theorem 4.4. For all $\sigma> a/2 $ , there is a constant $C_{\sigma }>0$ such that
Proof. Recall that for any self-adjoint operators A and B,
Since $|s_i| \leq \tfrac 12$ and $t \geq 1$ , by Lemma 4.1, each polynomial $q_{j}$ is bounded in $\textbf {s}$ and t. Then, essentially by skew-adjointness of $Y_i(t)$ , there exists a large constant $C>1$ with
Since operators on both sides are essentially self-adjoint,
Thus, there is a constant $C_\sigma>0$ such that
By Lemma 4.3, we can see that for $\sigma>a/2$ , setting $F(\tau ,0) = f\circ \phi _{X(t)}^\tau (x)$ ,
5 Average width estimate
This section is devoted to the proof of estimates on the averaged width of orbits of nilflows. Compared with the quasi-abelian case (see [Reference Flaminio and ForniFF14, Lemma 2.4]), there are no explicit expressions for the return map on transverse higher-step nilmanifolds. Instead, we calculate the differential of displacement and estimate the measure of close return orbits with respect to the rescaled vector fields.
Strategy.
-
(1) In §5.1 we introduce basic settings. Since the flow commutes with the centralizer, we take the quotient map to obtain a local diffeomorphism. It is remarkable to see that we set a tubular neighborhood to consider close return orbits on the quotient space. This calculates the measure of the set of close return orbits so-called almost periodic set. (See (54) and Lemma 5.4.)
-
(2) The range of the differential of the displacement map coincides with the range of the adjoint map $ad_{X_\alpha }$ . This is one important reason why a transverse condition is needed. Without this condition, there could be a direction that the return orbit that does not reach on the transverse manifold, which contradicts the idea that the set of almost periodic points should have a small measure up to rescaling vector $\rho $ . (See Lemma 5.8.)
-
(3) In §5.2 we prove a bound on the average width by ergodic averages of cut-off functions. By Definition 5.9, we classify the type of close return orbits by growth of local coordinates. The width of the function does not vanish on such a set and it is injective under the restricted domain. (See Lemmas 5.10 and 5.11.)
-
(4) Finally, in §5.3 and 5.4, we follow the known estimates from [Reference Flaminio and ForniFF14], which are necessary for proving bounds of ergodic averages in §6. In particular, the definition of a good point (Definition 5.21) means the set of points whose the width along the direction transverse to the flow cannot be too small and we prove the complement of the set of good points has a small measure. (See Lemma 5.22.)
5.1 Almost periodic points
Let $X_{\alpha }$ be the vector field on M defined in (7). Recall formula (8):
Let us introduce special type of condition for the Lie algebra $\mathfrak {n}$ required for width estimates.
Definition 5.1. The nilpotent Lie algebra $\mathfrak {n}$ satisfies the transversality condition if there exists a basis $(X_\alpha , Y)$ of $\mathfrak {n}$ such that
where $\mathfrak {G}_\alpha = (X_\alpha , Y^{(1)}_i)_{1 \leq i \leq n}$ is a set of generator, $\text {Ran}(\text {ad}_{X_{\alpha }})=\{Y \in \mathfrak {I} \mid Y= \text {ad}_{X_{\alpha }}(W), W \in \mathfrak {I} \}$ and $C_{\mathfrak {I}}(X_{\alpha }) = \{Y \in \mathfrak {I} \mid [Y,X_{\alpha }] = 0\}$ is the centralizer.
It is clear that the set of generators is included neither in the range of $\text {ad}_{X_{\alpha }}$ nor in the centralizer $C_{\mathfrak {I}}(X_{\alpha })$ . We will restrict $\mathfrak {n}$ to satisfying the condition (42) in the rest of sections.
Remark 5.2. The transversality condition implies that the displacement (or distance between x and $\Phi ^r_{\alpha ,\theta }(x)$ ), induced by return map $\Phi _{\alpha ,\theta }$ , should intersect the set of centralizers transversally. That is, the measure of the set of close return orbits in the transverse manifold $M_\theta ^a$ should not be invariant under the action of flow. This condition is crucial in estimating the almost periodic orbit (54) under rescaling of basis in Lemma 5.8.
Recall that $M_\theta ^a$ denotes the fiber at $\theta \in \mathbb {T}^1$ of the fibration $pr_2: M \rightarrow \mathbb {T}^1$ . $\Phi _{\alpha ,\theta }$ denotes the first return map of the nilflow $\{\phi _{X_{\alpha }}^t\}$ to the transverse section $M_\theta ^a$ , and $\Phi ^r_{\alpha ,\theta }$ denotes the rth iterate of the map $\Phi _{\alpha ,\theta }$ . Let G denote the nilpotent Lie group with its lattice $\Gamma $ defining $M_\theta ^a = \Gamma \backslash G$ . G acts on $M_\theta ^a$ by right action and the action of G extends to $M_\theta ^a \times M_\theta ^a$ .
Define a map $\psi _{\alpha ,\theta }^{(r)} : M_\theta ^a \rightarrow M_\theta ^a \times M_\theta ^a$ given by $\psi _{\alpha ,\theta }^{(r)}(x) = (x,\Phi ^r_{\alpha ,\theta }(x))$ . By its definition, the map $\Phi ^r_{\alpha ,\theta }$ commutes with the action of the centralizer $C_{G} = \exp (C_{\mathfrak {I}}(X_{\alpha })) \subset G$ and its action on the product $M_\theta ^a \times M_\theta ^a$ commutes with $\psi _{\alpha ,\theta }^{(r)}$ . That is, for $c \in C_{G}$ and $x = \Gamma g$ ,
Then the quotient map is well defined on
Setting. (i) In $M_\theta ^a \times M_\theta ^a$ , we set the diagonal $\Delta = \{(x,x) \mid x \in M_\theta ^a\}$ which is isomorphic to $M_\theta ^a $ by identifying $(x,x)$ with $x \in M_\theta ^a$ . Given $(x,x) \in \Delta $ , the tangent space of the diagonal is $ T_{(x,x)}\Delta := \{(v,v) \mid v \in T_{x} M_\theta ^a\}$ , and its normal space is defined by $(T_{(x,x)}\Delta )^\perp = \{(v,-v) \mid v \in T_x M_\theta ^a \} = T_{(x,x)}\Delta ^\perp $ . On the tangent space at $(x,x) \in M_\theta ^a\times M_\theta ^a$ , it splits as
For any $w_1, w_2 \in T_{x}M_\theta ^a$ ,
(ii) Given $x = \Gamma h_1, y = \Gamma h_2 \in M_\theta ^a$ , define a set $\Delta _{(x,y)} = \{(xg,yg) \mid g \in G \} \subset M_\theta ^a \times M_\theta ^a$ for $(xg,yg) = (\Gamma h_1 g, \Gamma h_2 g)$ and $\Delta _{(x,y)}^\perp = \{(xg,yg^{-1}) \mid g \in G \}$ that contains $(x,y)$ . For $\psi _{\alpha ,\theta }^{(r)}(x) = (x,\Phi ^r_{\alpha ,\theta }(x))$ , its tangent space in $M_\theta ^a \times M_\theta ^a$ is decomposed as
Then the tangent space of the diagonal is $T_{(x,\Phi ^r_{\alpha ,\theta }(x))}\Delta _{(x,\Phi ^r_{\alpha ,\theta }(x))} = \{(v,d_x\Phi ^r_{\alpha ,\theta }(v) ) \mid v \in T_x M_\theta ^a\}$ and its normal space is identified as
By identification in (45), for $w_1= v$ and $w_2 = -d_x\Phi ^r_{\alpha ,\theta }(v)$ , we write
(iii) Now define the orthogonal projection $ \pi : M_\theta ^a \times M_\theta ^a \rightarrow M_\theta ^a \times M_\theta ^a$ along the direction of the diagonal. That is, for $(x,y) \in M_\theta ^a \times M_\theta ^a$ , there exists $(x',y')$ such that $\pi (x,y) = (x', y') \in \Delta _{(x,y)} \cap \Delta _{(x,x)}^\perp .$ Then
Define a map $F^{(r)}: M_\theta ^a \rightarrow M_\theta ^a \times M_\theta ^a$ given by $F^{(r)} = \pi \circ \psi _{\alpha ,\theta }^{(r)}$ . In the local coordinate, by identification (46) and (47),
By (43) and the definition of $F^{(r)}$ , we have $F^{(r)}(xc) = F^{(r)}(x)c$ for $c \in C_G$ . Then for all $r \in \mathbb {Z}$ , $F^{(r)}$ induces a quotient map $F_C^{(r)}: M_\theta ^a / C_G \rightarrow M_\theta ^a \times M_\theta ^a / C_G$ . From (48), the range of the differential $DF_C^{(r)}$ is determined by $I-D\Phi ^r_{\alpha ,\theta }$ (refer to Figure 1).
In the next lemma, we verify the range of the differential map $DF_C^{(r)}$ .
Lemma 5.3. For all $r \in \mathbb {Z} \backslash \{0\}$ , the range of $I-D\Phi ^r_{\alpha ,\theta }$ on $\mathfrak {I}/C_{\mathfrak {I}}(X_{\alpha }) $ coincides with $ \text {Ran}(\text {ad}_{X_{\alpha }})$ and the Jacobian of $F_C^{(r)}$ is non-zero constant.
Proof. Recall that $\Phi ^r_{\alpha ,\theta }$ is the rth return map on $M_\theta ^a$ . We find the differential in the direction of each $Y^j_i$ for fixed i and j. For $x \in M$ , set a curve $\gamma ^x_{i,j}(t) = x\exp (tY_i^{(j)})\exp (rX_{\alpha }). $ Note that
and
By definition, $ ({\partial \Phi ^r_{\alpha , \theta }}/{\partial s_{i}^{(j)}})(x) = {d}/{dt}(\gamma ^x_{i,j}(t))\mid _{t=0}$ and we have $I-D\Phi ^r_{\alpha ,\theta } = I-\sum _{(i,j) \in J}({\partial \Phi ^r_{\alpha , \theta }}/{\partial s_{i}^{(j)}})$ . Then
Therefore, the range of $I-D\Phi ^r_{\alpha ,\theta }$ is contained in $\text {Ran}(\text {ad}_{X_{\alpha }})$ .
Conversely, $({1-e^{-\text {ad}_{X_{\alpha }}}})/{\text {ad}_{X_{\alpha }}} = \sum _{k=0}^\infty ({(-1)^k})/({(k+1)!})(\text {ad}_{X_{\alpha }})^k$ is invertible and
Therefore, we conclude that the range of $I-D\Phi ^r_{\alpha ,\theta }$ is $ \text {Ran}(\text {ad}_{X_{\alpha }})$ .
If $\sum _{(i,j) \in J} s_{i}^{(j)}Y_{i}^{(j)} \in C_{\mathfrak {I}}(X_{\alpha })$ , then $(I-D\Phi ^r_{\alpha ,\theta })(\sum _{(i,j) \in J} s_{i}^{(j)}Y_{i}^{(j)}) = 0$ and the kernel of $I-D\Phi ^r_{\alpha ,\theta }$ is $C_{\mathfrak {I}}(X_{\alpha })$ . That is, $I-D\Phi ^r_{\alpha ,\theta }$ is bijective on $\mathfrak {I}/C_{\mathfrak {I}}(X_{\alpha }) $ . Thus, by (49), the Jacobian of $I-D\Phi ^r_{\alpha ,\theta }$ is non-zero constant, and this concludes the proof.
Setting (continued). (iv) Set the submanifold $\mathcal {S} \subset M_\theta ^a \times M_\theta ^a$ that consists of the diagonal $\Delta $ and the coordinates of generators in normal (transverse) directions. Denote its quotient $\mathcal {S}_C = \mathcal {S}/C_G \subset M_\theta ^a \times M_\theta ^a / C_G$ . Then, following Lemma 5.3, we obtain the transversality of $F_C^{(r)}$ to $\mathcal {S}_C$ . For every $p \in (F_C^{(r)})^{-1}(\mathcal {S}_C)$ , the transversality holds on the tangent space:
(v) Denote the Lebesgue measure $\mathcal {L}^{a+1} (= \mathrm {vol}_{M})$ on the nilmanifold M and the conditional measure $\mathcal {L}_\theta ^{a} (= \mathrm {vol}_{M_\theta ^a})$ on the transverse manifold $M_\theta ^a$ . On the quotient space $M_\theta ^a/C_G$ , we write the measure $\mathcal {L}_\theta ^{c} (= \mathrm {vol}_{M_\theta ^a/C_G})$ . Similarly, we set the conditional measure $\mu _\theta ^{a}(= \mathrm {vol}_{M_\theta ^a \times M_\theta ^a})$ on the product manifold and $\mu _\theta ^{c}(= \mathrm {vol}_{M_\theta ^a \times M_\theta ^a/C_G})$ on its quotient space.
Denote the image of $F^{(r)}$ by $M_{\theta ,r}^a:= F^{(r)}(M_\theta ^a) \subset M_\theta ^a \times M_\theta ^a$ and $M_{\theta ,r,C}^a:= F_C^{(r)}(M_\theta ^a/C_G)$ . We write its conditional Lebesgue measure $\mu _{\theta ,r}^{a}:= \mu _\theta ^{a}|_{M_{\theta ,r}^a}$ and $\mu _{\theta ,r}^{c}:= \mu _\theta ^{c}|_{M_{\theta ,r,c}^a}$ , respectively.
For any open set $U_{\mathcal {S}_C} \subset M_\theta ^a \times M_\theta ^a / C_G$ , we write the pushforward measure $(F_C^{(r)})_* \mathcal {L}_\theta ^{c}$ as
By the compactness of $M_\theta ^a$ (or $M_\theta ^a/C_G$ ), the above expression is finite. By Lemma 5.3, the Jacobian of $F_C^{(r)}$ is constant and $(F_C^{(r)})_* \mathcal {L}_\theta ^{c} = \mu _{\theta ,r}^{c}$ is Lebesgue.
By invariance of the action of the centralizer, for any neighborhood $U_{\mathcal {S}} \in M_\theta ^a \times M_\theta ^a$ with $U_{\mathcal {S}_C} = U_{\mathcal {S}}/C_G$ ,
and by definition of conditional measure,
Let d be a distance function in $M_\theta ^a \times M_\theta ^a$ , and we abuse notation d for the induced distance on $M_\theta ^a \times M_\theta ^a/C_G$ . Let $U_{\delta } = \{ z \in M_\theta ^a \times M_\theta ^a \mid d(z,\mathcal {S}) < \delta \}$ be a $\delta $ -tubular neighborhood of $\mathcal {S}$ and $U_{\delta ,C} = \{ z \in M_\theta ^a \times M_\theta ^a/C_G \mid d(z,\mathcal {S}_C) < \delta \}$ be its quotient.
Define the almost-periodic set (set of rth close return) on the diagonal
Since $F^{(r)}$ commutes with $C_G$ , $AP^{r}(\mathcal {U}_\delta )/C_G = \{ x \in M_\theta ^a/C_G \mid d(F_C^{(r)}(x),\mathcal {S}_C) < \delta \}$ and $ \mathcal {L}_\theta ^{a}(AP^{r}(\mathcal {U}_\delta )) = \mathcal {L}_\theta ^{c}(AP^{r}(\mathcal {U}_\delta )/C_G) $ .
The following volume estimate of almost-periodic set holds.
Lemma 5.4. Let $U_{\delta ,C}$ be any tubular neighborhood of $\mathcal {S}_C$ in $M_\theta ^a \times M_\theta ^a/C_G$ . For all $r \in \mathbb {Z} \backslash \{0\}$ , the conditional measure $ \mathrm {vol}_{M_\theta ^a}$ of $AP^{r}(\mathcal {U}_\delta )$ is given as follows:
Proof. By setting (v), it suffices to prove $\mathcal {L}_\theta ^{c}(AP^{r}(\mathcal {U}_\delta )/C_G) = \mu _{\theta ,r}^{c}(U_{\delta } \cap M_{\theta ,r,C}^a)$ . Note that $(F_C^{(r)})^{-1}(AP^{r}(\mathcal {U}_\delta )/C_G) = \{ x \in M_\theta ^a / C_G \mid d(z,\mathcal {S}_C) < \delta \}$ if $z = F_C^{(r)}(x)$ for some $x \in M_\theta ^a/C_G$ , otherwise it is an empty set.
Then $(F_C^{(r)})^{-1}(AP^{r}(\mathcal {U}_\delta )/C_G) = (U_{\delta ,C} \cap M_{\theta ,r,C}^a) $ . Thus, by definition of the pushforward measure, the equality holds.
Recall that $F_C^{(r)}: M_\theta ^a/C_G \rightarrow M_{\theta ,r,C}^a$ has non-zero constant Jacobian if $r \neq 0$ by Lemma 5.3 and it is a local diffeomorphism. Thus, by the transversality of $F_C^{(r)}$ , in a small tubular neighborhood U, $F_C^{(r)}$ is covering.
Lemma 5.5. For any $z \in U\cap M_{\theta ,r,C}^a$ , there exist a finite number of pre-images of $F_C^{(r)}$ .
Proof. If we suppose that $(F_C^{(r)})^{-1}(z)$ contains infinitely many different points, then since the manifold $M_\theta ^a$ is compact (and $M_\theta ^a/C_G$ is compact), there exists a sequence of pairwise different points $x_i \in (F_C^{(r)})^{-1}(z)$ , which converges to $x_0$ . We have $(F_C^{(r)})(x_0) = z$ and, by the inverse function theorem, the point $x_0$ has a neighborhood $U'$ in which $F_C^{(r)}$ is a homeomorphism. In particular, $U'\backslash \{x_0\} \cap (F_C^{(r)})^{-1}(z) = \emptyset $ , which is a contradiction.
Set $N_r(z) = \#\{x \in M_\theta ^a/C_G \mid F_C^{(r)}(x) = z\}$ as the number of pre-images of $F_C^{(r)}$ . The number $N_r(z)$ is independent of the choice of $z \in U\cap M_{\theta ,r,C}^a$ since Jacobian is constant and the degree of the map is invariant (see [Reference Dubrovin, Fomenko and NovikovDAS, §3]).
We now introduce the volume estimate of the $\delta $ -neighborhood $U_{\delta ,C}$ .
Proposition 5.6. The following volume estimate holds: for any $r \neq 0$ , there exists $C:= C(M_\theta ^a)>0$ such that
Proof. Let $U \subset M_\theta ^a \times M_\theta ^a/C_G$ be a tubular neighborhood of $\mathcal {S}_C$ that contains $U_{\delta ,C}$ with the following condition:
If $U\cap M_{\theta ,r,C}^a = \emptyset $ , then there is nothing to prove since $U_{\delta ,C}\cap M_{\theta ,r,C}^a = \emptyset $ . Assume $z \in U\cap M_{\theta ,r,C}^a$ and let $\{J_k\}_{k \geq 1}$ be connected components of $(F_C^{(r)})^{-1}(U \cap M_{\theta ,r,C}^a)$ . We first claim that $F_C^{(r)}|_{J_k}$ is injective.
Given $z \in U\cap M_{\theta ,r,C}^a$ , assume that there exist $x_1 \neq x_2 \in J_k$ for some k such that $z = F_C^{(r)}|_{J_k}(x_1) = F_C^{(r)}|_{J_k}(x_2)$ . Let $\gamma : [0,1] \rightarrow J_k$ be a path that connects $\gamma (0) = x_1$ and $\gamma (1) = x_2$ . Set the lift of the path $\tilde \gamma = F_C^{(r)}|_{J_k} \circ \gamma : [0,1] \rightarrow U$ . Then $\tilde \gamma (0) = \tilde \gamma (1) = z$ and $\tilde \gamma $ is a loop in U. Since U is simply connected, $\tilde \gamma $ is contractible and there exists a homotopy of the path $g_s: [0,1] \rightarrow U$ such that $g_0 = \tilde \gamma $ is homotopic to a constant loop $g_1 = c$ by fixing two end points $F_C^{(r)}|_{J_k}(x_1) = F_C^{(r)}|_{J_k}(x_2) = z$ for $s \in [0,1]$ .
Note that $F_C^{(r)}|^{-1}_{J_k} \circ g_s$ is a lift of homotopy $g_s$ , and a lift of $g_0$ is $\gamma = F_C^{(r)}|^{-1}_{J_k}(\tilde \gamma )$ with fixed end points $x_1$ and $x_2$ . By homotopy continuity, $g_s$ also keeps the same end points $x_1$ and $x_2$ fixed for all $s \in [0,1]$ . Since $g_1$ is constant loop and its lift should be a single point, $\gamma $ is homotopic to a constant. Since the end points of $\gamma $ are fixed, they have to be constant, but this leads a contradiction. Therefore, we have $x_1= x_2$ .
By injectivity of $F_C^{(r)}|_{J_k}$ , we obtain
Furthermore, we obtain the following equality:
Since the volume of $M_\theta ^a/C_G$ is a finite,
Assume that $(F_C^{(r)})^{-1}(U_{\delta ,C}) = \bigcup _{k=1}^{N_r} (F_C^{(r)}|_{J_k})^{-1}(U_{\delta ,C})$ . Then by (56) and definition of conditional measure,
By the last equality, together with condition (55),
Therefore, combining (57) and (59), there exists $C>0$ such that
Definition 5.7. For any basis $Y = \{Y_1,\ldots Y_a \}$ of codimension- $1$ ideal $\mathfrak {I}$ of $\mathfrak {n}$ , let $I:=I(Y)$ be the supremum of all constants $I' \in (0,\tfrac 12)$ such that, for any $x \in M$ , the map
is a local embedding (injective) on the domain
For any $x, x' \in M$ , set local distance $d_*$ (measured locally in the Lie algebra) on the transverse section $M_\theta ^a$ along the $Y_i$ direction by $d_{Y_i}({x,x'}) = |s_i|$ if there is $\textbf {s}: = (s_1,\ldots , s_a) \in [-I/2,I/2]^a $ such that
otherwise $d_{Y_i}({x,x'}) =I$ .
Recall the projection map $pr_1 : M \rightarrow \mathbb {T}^{n+1}$ onto the base torus. On the transverse manifold, for all $\theta \in \mathbb {T}^1$ , let $pr_\theta : M_\theta ^a \rightarrow \mathbb {T}^n$ be the restriction to $M_\theta ^a$ . Then, by applying formula (10) to projection to the base torus,
For any $L \geq 1$ , $r \in \mathbb {Z}$ , $x \in M_\theta ^a $ and given scaling factor $\rho = (\rho _1,\ldots , \rho _a) \in [0,1)^a$ , we define
We note that the distance $d_{Y_i}(\Phi ^r_{\alpha , \theta }(x),x)$ on the generator level does not depend on the choice of x. For this reason, we split the cases $\epsilon _{r,L}$ and $\delta _{r,L}(x)$ to control the distance separately on higher steps.
The condition $\epsilon _{r,L} < \epsilon < I$ and $\delta ' < \delta _{r,L}(x) < \delta < I$ are equivalent to saying
for some vectors $\textbf {s}: = (s_1,\ldots , s_a) \in [-I/2,I/2]^a $ such that
For every $r \in \mathbb {Z} \backslash \{0\}$ and $j \geq 0$ , let ${AP^r_{j,L}} \subset M$ be a set defined as follows:
In the next lemma, the Lebesgue measure of the set of almost-periodic points ${AP_{j,L}^{r}}$ on M is estimated by the volume of the $\delta $ -neighborhood $\mathcal {U}_\delta $ .
Lemma 5.8. For all $r \in \mathbb {Z} \backslash \{0\}$ , $j \in \mathbb {N}$ , $L \geq 1$ , the $(a+1)$ dimensional Lebesgue measure of the set $AP^r_{j,L}$ can be estimated as follows: there exists $C>0$ such that
Proof. Without loss of generality, we assume that $AP^r_{j,L} \neq \emptyset $ .
By Tonelli’s theorem,
Recall the definition $AP^{r}(\mathcal {U}_\delta )$ in (54). Choose $\delta = {I^{a-n}}/{2^{j(a-n)}} L^{-\sum _{i=n+1}^a\rho _i}$ and set $ \mathcal {U}_{\delta }^{L,j}:= \mathcal {U}_{\delta }$ . Then we claim that $ {AP_{j,L}^{r}} \cap M_\theta ^a \subset AP^{r}(\mathcal {U}_{\delta }^{L,j})$ .
By identification of $x \in M_\theta ^a$ with $ (x,x)$ in the diagonal $\Delta \subset M_\theta ^a \times M_\theta ^a$ , the local distance $d_{Y_i}(\Phi ^r_{\alpha , \theta }(x),x)$ is identified by the distance function d in the product $M_\theta ^a \times M_\theta ^a$ . Thus, $x \in AP^r_{j,L} \cap M_\theta ^a$ implies that $d(F^{(r)}(x),\mathcal {S}) < \delta $ . That is, $AP^r_{j,L} \cap M_\theta ^a \subset AP^r(\mathcal {U}_\delta ^{L,j})$ . By Lemma 5.4, the volume estimate follows
Finally, by Proposition 5.6,
Thus the proof follows from formula (65).
5.2 Expected width bounds
We prove a bound on the average width of an orbit on a nilmanifold with respect to a scaled basis. This subsection follows along the lines of [Reference Flaminio and ForniFF14, §5.2]. To complete of the proof, we repeat the similar arguments in nilmanifolds under transverse conditions.
For $L\geq 1$ and $r \in \mathbb {Z} \backslash \{0\}$ , let us consider the function
Define the cut-off function $J_{r,L} \in \mathbb {N}$ by the formula
The function $h_{r,L}$ is
For every $L\geq 1$ , let $\mathcal {F}_{\alpha }^{(L)}$ be the rescaled strongly adapted basis
For $(x,T) \in M \times \mathbb {R}$ , let $w_{\mathcal {F}_{\alpha }^{(L)}}(x,T)$ denote the average width of the (scaled) orbit segment
We prove a bound for the average width of the orbit arc in terms of the following function:
Definition 5.9. For $t \in [0,T]$ , we define a set of points $\Omega (t) \subset \{t\}\times \mathbb {R}^a$ as follows.
Case 1. If $\phi ^t_{X_{\alpha }^{(L)}}(x) \notin \bigcup _{|r|=1}^{[TL]}\bigcup _{j>0} AP^r_{j,L}$ , let $\Omega (t)$ be the set of all points $(t,s_1,\ldots ,s_a)$ such that
If $\phi ^t_{X_{\alpha }^{(L)}}(x) \in \bigcup _{|r|=1}^{[TL]}\bigcup _{j>0} AP^r_{j,L}$ , then we consider two subcases.
Case 2-1. If $\phi ^t_{X_{\alpha }^{(L)}}(x) \in \bigcup _{|r|=1}^{[TL]}\bigcup _{j>J_{r,L}} AP^r_{j,L}$ , let $\Omega (t)$ be the set of all points $(t,\textbf {s})$ such that
Case 2-2. If $\phi ^t_{X_{\alpha }^{(L)}}(x) \in \bigcup _{|r|=1}^{[TL]}\bigcup _{j\leq J_{r,L}} AP^r_{j,L}\backslash \bigcup _{|r|=1}^{[TL]}\bigcup _{j>J_{r,L}} AP^r_{j,L}$ , let l be the largest integer such that
and let $\Omega (t)$ be the set of all points $(t, \textbf {s})$ such that
We set
Lemma 5.10. The restriction to $\Omega $ of the map
is injective.
Proof. Assume there exist $t'\geq t$ and $(t,s_1,\dotsc,s_n) \neq (t',s'_1, \dotsc, s'_n)$ such that
By considering the projection on the base torus, we have the following identity:
which implies $t \equiv t'$ modulo $\mathbb {Z}$ . As $\phi _{X_{\alpha }}^{tL} = \phi ^t_{X_{\alpha }^{(L)}}$ , the number $r_0 = t' - t$ is a non-negative integer satisfying $r_0 \leq TL$ ; hence $r_0 \leq [TL]$ .
If $r_0=0$ , then $t' = t$ and $s^{\prime }_i = s_i$ for all $1\leq i \leq a$ . Then injectivity is obtained by definition of I. Assume that $r_0\neq 0$ . Let $p,q \in M_\theta ^a$ . Then we have
From identity (72) we have
where $P_i$ is polynomial expression following from the Baker–Cambell–Hausdorff formula.
Note that $P_i(s_i,s_i') = 0$ if $i = 1,\ldots , n$ and $|P_i(s_i,s_i')| \leq \sum _{l = 1}^\infty 1/2|s_ls_l'|^l$ for $i>n$ . Since $|s_i|, |s_i'| \leq {I}/{4} \ll 1$ ,
where $I_\epsilon = \sum _{l=1} ({I}/{4})^l = {I}/({4-I}) < I/3.$ Thus for all $i \in \{1,\ldots ,a\}$ ,
and
For the same reason, from formula (74) we also obtain that
By defining $j_0 \in \mathbb {N}$ as the unique non-negative integer such that
and by the Definition 5.7, we have $p \in AP^{r_0}_{j_0,L}$ and $q \in AP^{-r_0}_{j_0,L}$ .
If $j_0> J_{r_0,L} = J_{-r_0,L},$ then $p,q \in \bigcup _{|r|=1}^{[TL]}\bigcup _{j \geq J_{r,L}}$ . It follows that the sets $\Omega (t)$ and $\Omega (t')$ are both defined in case 2-1. Hence by (75),
which is a contradiction.
If the map in formula (71) fails to be injective at points $(t,s)$ and $(t',s')$ with $t \geq t'$ , then there are integers $r_0 \in [1,TL], j_0 \in [1,J(|r_0|)]$ and $\theta \in \mathbb {T}^1$ such that the points p and q satisfy
In this case, the sets $\Omega (t)$ and $\Omega (t')$ are both defined according to case (2-2). Let $l_1$ and $l_2$ as the largest integers such that
In case (2-2), we have
which also leads to a contradiction because $l_1,l_2> j_0$ deduce the contradiction
Hence, the injectivity is proved.
We simply re-prove the bound on the averaged width (see [Reference Flaminio and ForniFF14, Lemma 5.5]) in the general settings (under transversality conditions) by combining with Lemma 5.10.
Lemma 5.11. For all $x \in M$ and for all $T, L \geq 1$ we have
Proof. The width function $w_{\Omega }$ of the set $\Omega $ is given by
and implies that
By the definition of the function $H_{L}^T$ in formula (70), we have
From the definition (40) of the average width of the orbit segment $\{x\exp {(tX_{\alpha }^{(L)})} \mid 0 \leq t \leq T\}$ , we have the estimate
Lemma 5.12. For all $r \in \mathbb {Z} \backslash \{0\}$ and for all $L \geq 1$ , the following estimate holds:
Proof. It follows from Lemma 5.8 that for $r \neq 0$ and for all $j \geq 0$ , the Lebesgue measure of the set $AP^r_{j,L}$ satisfies the following bound:
From formula (68), it follows that
By the estimate in formula (79), we immediately have that
By the definition of the cut-off in formula (67) we have the bound
and by an estimate on a geometric sum
5.3 Diophantine estimates
We review the simultaneous Diophantine condition.
Definition 5.13. A vector $\alpha \in \mathbb {R}^n\backslash \mathbb {Q}^n$ is simultaneously Diophantine of exponent $\nu \geq 1$ , say $\alpha \in DC_{n,\nu }$ if there exists a constant $c(\alpha )>0$ such that, for all $r\in \mathbb {N} \backslash \{0\} $ ,
Definition 5.14. For any basis $\bar {Y}:= \{ \bar {Y}_1, \ldots , \bar {Y}_n\} \subset \mathbb {R}^n$ , let $\bar {I}:= \bar {I}(\bar {Y})$ be the supremum of all constants $\bar {I'}>0$ such that the map
is a local embedding on the domain
For any $\theta \in \mathbb {R}^n$ , let $[\theta ] \in \mathbb {T}^n$ be its projection onto the torus $\mathbb {T}^n := \mathbb {R}^n/\mathbb {Z}^n$ and let
if there is $\textbf {s}:=(s_1,\ldots , s_n) \in [-\bar {I}/2, \bar {I}/2]^n$ such that
otherwise we set $|\theta |_1 = \cdots = |\theta |_n = \bar {I}$ .
Definition 5.15. (Counting for close return time)
Let $\sigma = (\sigma _1,\ldots ,\sigma _n) \in (0,1)^n$ be such that $\sigma _1+\cdots +\sigma _n = 1$ . For any $\alpha = (\alpha _1,\ldots ,\alpha _n) \in \mathbb {R}^n$ , for any $N \in \mathbb {N}$ and every $\delta>0$ , let
For every $\nu>1 $ , let $D_n(\bar {Y},\sigma , \nu ) \subset (\mathbb {R}\backslash \mathbb {Q})^n$ be the subset defined as follows: the vector $\alpha \in D_n(\bar {Y},\sigma , \nu )$ if and only if there exists a constant $C(\bar {Y},\sigma , \alpha )>0$ such that for all $N \in \mathbb {N}$ for all $\delta>0$ ,
The Diophantine condition for counting return time implies a standard simultaneous Diophantine condition.
Lemma 5.16. [Reference Flaminio and ForniFF14, Lemma 5.9]
Let $\alpha \in D_n$ . For all $r \in \mathbb {Z}\backslash \{0\}$ , we have
For any vector $\sigma = (\sigma _1, \ldots , \sigma _n) \in (0,1)^n$ such that $\sigma _1+\cdots +\sigma _n = 1$ , let
Lemma 5.17. [Reference Flaminio and ForniFF14, Lemma 5.12]
For all bases $\bar {Y} \subset \mathbb {R}^n$ , for all $\sigma = (\sigma _1, \ldots , \sigma _n) \in (0,1)^n$ such that $\sigma _1+\cdots +\sigma _n = 1$ and for all $\nu \geq 1$ , the inclusion
holds under the assumption
The set $D_n(\bar {Y},\sigma ,\nu )$ has full measure if
In dimension $1$ , the vector space has unique basis up to scaling. The following result is immediate.
Lemma 5.18. [Reference Flaminio and ForniFF14, Lemma 5.13]
For all $\nu \geq 1$ the following identity holds:
Let $\mathcal {F}_\alpha := (X_\alpha , Y)$ be a basis and let $\bar {Y} = \{\bar {Y}_1,\ldots , \bar {Y}_{n} \} \in \mathbb {R}$ denote the projection of the basis of codimension- $1$ ideal $\mathfrak {I}$ onto the abelianized Lie algebra $\bar {\mathfrak {n}} := \mathfrak {n}/[\mathfrak {n},\mathfrak {n}] \approx \mathbb {R}^{n}$ . For $\rho = (\rho _1,\ldots ,\rho _a) \in [0,1)^a$ , we write a vector of scaling exponents
Let $\alpha _1 = (\alpha ^{(1)}_1, \ldots , \alpha ^{(1)}_{n} ) \in D_n(\bar {Y}, \bar {\rho }/|\bar {\rho }|,\nu )$ . For brevity, let $C(\bar {Y}, \bar {\rho }/|\bar {\rho }|,\alpha _1)$ denote the constant appearing in (80) for $\alpha _1 \in D_n(\bar {Y}, \bar {\rho }/|\bar {\rho }|,\nu )$ . Let
We prove the upper bound on the cut-off function in formula (67). Let $I = I(Y)$ and $\bar {I} = \bar {I}( \bar {Y})$ be the positive constant introduced in Definitions 5.7 and 5.14. We observe that $I \leq \bar {I} $ since the basis $\bar {Y}$ is the projection of the basis $Y \subset \mathfrak {n}'$ and the canonical projection commutes with the exponential map. Then the following logarithmic upper bound holds.
Lemma 5.19. [Reference Flaminio and ForniFF14, Lemma 5.14]
For every $\rho \in [0,1)^a$ , for every $\nu \leq 1/|\bar {\rho }|$ and for every $\alpha \in D_n(\bar {Y}, \bar {\rho }/|\bar {\rho }|,\alpha )$ , there exists a constant $K>0$ such that for all $T \geq 1$ and for all $r \in \mathbb {Z} \backslash \{0\}$ , the following bound holds:
Proof. By Lemma 5.16 and by the definition of $\epsilon _{r,L}$ in formula (62), it follows that, for all $T>0, L\geq 1$ and for all $r \in \mathbb {Z} \backslash \{0\}$ , we have
It follows by the bound above and by the definition of the cut-off function (67) that
Assume that there exists $\nu \in 1/|\bar {\rho }|$ such that $\alpha _1 \in D_n(\bar {Y},\bar {\rho }/|\bar {\rho }|,\nu )$ . For brevity, we introduce the following notation:
Theorem 5.20. [Reference Flaminio and ForniFF14, Theorem 5.15]
For every $\rho \in [0,1)^a$ , for every $\nu \leq 1 / |\bar {\rho }|$ such that $\alpha _1 = \alpha _i^{(1)} \in D_n(\bar {Y},\bar {\rho },\nu ) $ there exists a constant $K'>0$ such that for all $T>0$ and for all $L \geq 1$ , the following bound holds:
Proof. By the definition of $H_{L}^T$ in formula (70), the statement follows from Lemmas 5.12 and 5.19. In fact, for all $r \in \mathbb {Z} \backslash \{0\}$ and $j \geq 0$ , by definition (64) the set $AP_{j,L}^{r}$ is non-empty only if $\epsilon _{r,L}<({I}/{2})$ . Since $\nu \leq 1/|\bar {\rho }|$ , it follows from the definition of the Diophantine class $D_n$ that
The main idea of the proof above follows from Lemma 5.12 which is based on the estimate to the upper bound of the measure of the almost periodic set $AP^r_{j,L}$ . In Lemma 5.8, this bound is independent of choice of transverse section $M_\theta ^a$ .
5.4 Width estimates along orbit segments
In this subsection we introduce the definition of good points, which is crucial in controlling the average width estimate.
Definition 5.21. For any increasing sequence $(T_i)$ of positive real numbers, let $h_i \in [1,2]$ denote the ratio $\log {T_i}/[\log T_i]$ for every $T_i \geq 1$ . Set $N_i = [\log T_i]$ and $T_{j,i} = e^{jh_i}$ for integer $j \in [0,N_i]$ .
Let $\zeta>0$ and $w>0$ . A point $x \in M$ is $(w,T_i,\zeta )$ -good for the basis $\mathcal {F}_\alpha $ if, setting $y_i = \phi _{X_{\alpha }}^{T_i}(x)$ , then for all $i \in \mathbb {N}$ and for all $0 \leq j \leq N_i$ ,
Under the transversality condition, the following proof is still valid and extendable from the quasi-abelian case. For the completion of the subsection, we repeat the proof of the following lemma.
Lemma 5.22. [Reference Flaminio and ForniFF14, Lemma 5.18]
Let $\zeta>0$ be fixed and let $ (T_i)$ be an increasing sequence of positive real numbers satisfying the condition
Let $\rho \in [0,1)$ with $\sum \rho _i =1$ . Then the Lebesgue measure of the complement of the set $\mathcal {G}(w, (T_i), \zeta )$ of $(w,(T_i),\zeta )-$ good points is bounded above. That is, $\text { there exists } K>0$ such that
Proof. For all $i \in \mathbb {N}$ and for all $j = 0, \ldots , N_i$ , let
By definition we have
By Lemma 5.11 for all $z \in \mathfrak {S}_{j,i}$ we have
It follows that
By the maximal ergodic theorem, the Lebesgue measure $\mathrm {meas}(\mathfrak {S}_{j,i})$ of the set $\mathfrak {S}(j,i) $ satisfies the inequality
Let $\mathcal {H} = \mathcal {H}(Y,\rho ,\nu )$ denote the constant defined in formula (83). By Theorem 5.20, since by hypothesis $\nu \leq 1/|\bar {\rho }|$ and $\alpha \in D_n(\bar {\rho }/|\bar {\rho }|,\nu )$ , there exists a constant $K':= K'(a,n,\nu )>0 $ such that the following bound holds:
Hence, by the definition of the $T_{j,i}$ , we have
Thus, for some constant $K'' $ , we have
By (86), for some constant $K'''>0$ ,
By sub-additivity of the Lebesgue measure, we derive the bound
By formula (85), the above estimate concludes the proof.
Remark 5.23. Lemma 5.22 proves that the Lebesgue measure of the complement of a good set is bounded along the sequence $(T_i^{-\zeta })_{i \in \mathbb {N}}$ with a certain rate of decay. This explains that the Lebesgue measure of the set of $x \in M$ with a small average width along unit time (or slow divergence along close return orbits) is bounded by the ergodic integral of cut-off functions for length 1. In particular, this lemma is necessary to handle bounds of ergodic averages of flows under rescaled time in §6.2, which is a counterpart argument to the use of the Sobolev constant that is only valid for the proof of (renormalizable) Heisenberg nilflows (cf. [Reference Flaminio and ForniFF06]).
6 Bounds on ergodic average
In this section, we prove the bound for deviation of ergodic averages of nilflows, for each function on irreducible representations. This bound will be derived from the calculations for the coboundary in §3 and the average width estimates for good points in §5.
We shall introduce assumptions on coadjoint orbits $\mathcal {O} \subset \mathfrak {n}^*$ .
Definition 6.1. A linear form $\Lambda \in \mathcal {O}$ is integral if the coefficients $\Lambda (\eta _i^{(m)})$ , $(i,m) \in J $ , are integer multiples of $2\pi $ . Denote by $\widehat M$ the set of coadjoint orbits $\mathcal {O}$ of integral linear forms $\Lambda $ .
There exist coadjoint orbits $\mathcal {O} \subset \mathfrak {n}^*$ that correspond to unitary representations which do not factor through the quotient $N/\exp {\mathfrak {n}_k}$ , $\mathfrak {n}_k \subset Z(\mathfrak {n})$ . Such coadjoint orbits and unitary representations are called maximal (see [Reference Flaminio and ForniFF07, §2.2]).
Let $\widehat M_0$ be subset of all coadjoint orbits of forms $\Lambda $ such that $\Lambda (\eta _i^{(m)}) \neq 0$ for $m = k$ . This space has maximal rank and $\Lambda (\eta _i^{(m)}) \neq 0, \text { for all } (i,m) \in J $ .
Definition 6.2. Given a coadjoint orbit $\mathcal {O} \in \widehat M_0$ and a linear functional $\Lambda \in \mathcal {O}$ , let us denote by $\mathcal {F}_{\alpha ,\Lambda }$ the completed basis obtained by completion of $(X_{\alpha}, \eta_*^{(1)},\dotsc,\eta_*^{(k)})$ . For all $t \in \mathbb {R}$ , we write the scaled basis $\mathcal {F}_{\alpha ,\Lambda }(t)$ as
For any $\mathcal {O} \in \widehat M_0$ , let $H_{\mathcal {O}}$ denote the primary subspace of $L^2(M)$ which is a direct sum of subrepresentations equivalent to $\text {Ind}_{N'}^N(\Lambda )$ . For the adapted basis $\mathcal {F}$ , set
6.1 Coboundary estimates for rescaled basis
Recall the definition of degree of $Y_{i}^{(m)}$ and
For any linear functional $\Lambda $ , the degree of the representation $\pi _{\Lambda }$ only depends on its coadjoint orbits. We denote the scaling vector $\rho \in ({\mathbb {R}^+})^J$ such that
Assume that the number of bases of $\mathfrak {n}$ with degree $k-m$ is $n_m$ . Define
We have $\delta (\rho ) \leq \unicode{x3bb} (\rho )$ . This inequality is strict unless one has homogeneous scaling
Lemma 6.3. There exists a constant $C>0$ such that, for all $r \in \mathbb {R}^+$ and for any function $f \in W^r(H_{\mathcal {O}}) $ , we have
Proof. For all $(m,i) \in J$ , we have
We note that $c^{(j)}_l = 0$ for $j = k$ and for some l, which is determined by the commutation relation. Setting $C = \max _{(i,j)\in J^+ }\{|c_i^{(j)}|\} $ ,
For $x \in M$ , let $\gamma _x$ be the Birkhoff average operator
Consider the decomposition of the restriction of linear functional $\gamma _x$ to $W^r_0(H_{\mathcal {O}},\mathcal {F}_{\alpha ,\Lambda }(t))$ as an orthogonal sum $\gamma _x = D(t)+R(t) \in W^{-r}_0(H_{\mathcal {O}},\mathcal {F}_{\alpha ,\Lambda }(t))$ of an $X_{\alpha }$ -invariant distribution $D(t)$ and an orthogonal complement $R(t)$ .
Theorem 6.4. Let $r> (k + 1)(a/2 + 1) + 1/4 $ . For $g \in W^r(H_{\mathcal {O}},\mathcal {F}_{\alpha ,\Lambda }(t))$ and for all $t \geq 0$ , there exists a constant $C_r^{(1)}>0$ such that
Proof. Fix $t \geq 0$ and set $ D = D(t), R = R(t)$ for convenience. Let $g \in W^r(H_{\mathcal {O}},{\mathcal F}_{\alpha,\Lambda}(t))$ . We write $g = g_D + g_R$ , where $g_R$ is the kernel of $X_{\alpha }$ -invariant distributions and $g_D$ is orthogonal to $g_R$ in $W^r(H_{\mathcal {O}},{\mathcal F}_{\alpha,\Lambda}(t))$ . Then $g_R$ is a coboundary and $R(g_D)=0$ . Let $f = G^{X_\alpha (t)}_{X_\alpha ,\Lambda } (g_R)$ . From $D(g_R)=0$ ,
By the Gottschalk–Hedlund argument,
By Theorem 4.4 and Lemma 6.3, for any $\tau> a/2+1$ , there exists a positive constant $C_r$ such that for any $z \in M$ ,
By Theorem 3.12, if $r> (k+1)\tau +1/4$ , then
By orthogonality, we have $|g_R|_{r,\mathcal {F}_{\alpha ,\Lambda }(t)} \leq |g|_{r,\mathcal {F}_{\alpha ,\Lambda }(t)}$ .
Corollary 6.5. For every $r> (k+1)(a/2+1)+1/4 $ , there is a constant $C_r^{(2)}>0$ such that the following holds for every $\mathcal {O} \in \widehat I_0$ and every $x \in M$ :
Proof. For all $x \in M$ , we have
It follows from Theorem 6.4 applied to the orthogonal decomposition of $\gamma _x = D(0)+R(0)$ .
6.2 Bounds on ergodic averages in an irreducible subrepresentation
In this subsection we derive the bounds on ergodic averages of nilflows for functions in a single irreducible subrepresentation.
For brevity, let us set
Proposition 6.6. Assume $(T_i)_{i \in \mathbb{N}}$ is an increasing sequence of positive real numbers $ \geq 1$ . Let $0 < w < I(Y)^a$ and $\zeta>0$ . Given $r> (k+1)(a/2+1)+1/4 $ , there exists a constant $C_r(\rho )$ such that for every $\mathcal {G}(w, (T_i), \zeta )$ -good point $x \in M$ and all $f \in W^r(H_{\mathcal {O}},\mathcal {F})$ , we have
Proof. By group action for scaling (18), a sequence of frames $\mathcal {F}(t_j) = A_{\rho }^{t_j}\mathcal {F} $ is chosen with other scaling factors $\rho _i t_j$ on elements of Lie algebras $Y_i$ . Then, as j increases from 0 to N, the scaling parameter $t_j$ becomes larger, while the scaled length of the arc becomes shorter, approaching $1$ . Let $\phi _{X_j}^s$ denote the flow of the scaled vector field $e^{t_j}X = X(t_j)$ .
For each $j = 0,\ldots , N$ , let $\gamma = D_j + R_j$ be the orthogonal decomposition of $\gamma $ in the Hilbert space $W^{-r}(H_\pi ,\mathcal {F}(t_j))$ into an $X_{\alpha }$ -invariant distribution $D_j$ and an orthogonal complement $R_j$ . For convenience, we denote by $|\cdot |_{r,j}$ and $\|\cdot \|_{r,j}$ respectively the transversal Sobolev norm $|\cdot |_{r,\mathcal {F}(t_j)}$ and Lyapunov Sobolev norm $\|\cdot \|_{r,\mathcal {F}(t_j)}$ relative to the rescaled basis $\mathcal {F}(t_j)$ .
Let us set $N_i = [\log T_i]$ and $t_{j,i} := T_{j,i} = \log T_i^{j/N_i}$ for integer $j \in [0,N_i]$ . We observe that $N_i < \log T_i < N_i + 1$ . For simplicity, we will omit the index $i \in \mathbb {N}$ and set $T = T_i$ , $N = N_i$ for a while within the proof and lemmas of this subsection.
Our goal is to estimate $|\gamma |_{-r,\mathcal {F}_\alpha } = |\gamma |_{-r,0}$ (the norm of the distribution of unscaled bases). By the triangle inequality and Corollary 6.5,
We now estimate $|D_0|_{-r,0}$ . By definition of the Lyapunov norm and its bound (39), for $-s < -r<0$ ,
Since $D_j + R_j = D_{j-1}+R_{j-1}$ , observe that $D_{j-1} = D_j + R^{\prime }_j$ , where $R^{\prime }_j$ denotes the orthogonal projection of $R_j$ on the space of invariant distributions $W^{-r}(H_{\mathcal {O}},\mathcal {F}(t_{j-1}))$ . By definition of the Lyapunov norm,
By Lemma 6.9, equivalence of norms gives
By Lemma 3.16, for any $X_{\alpha }$ -invariant distribution D and for all $t_j \geq t_{j-1}$ ,
Since $\mathcal {F}(t_j) = A_{\rho }^{t_j-t_{j-1}}\mathcal {F}(t_{j-1}) $ and $t_{j} - t_{j-1} = \log T/N$ , we obtain
From (97) we conclude by induction
From (95) and the above, we conclude that there exists a constant $C_r(\rho )$ such that
Here we provide the proofs of our supplementary lemmas.
Lemma 6.7. For any $r>a/2$ , there exists a constant $C_r>0$ such that for all good points $x \in \mathcal {G}(w, (T_i), \zeta )$ , we have
Proof. By definition of norm,
It suffices to find the bound of orbit segment with respect to rescaled bases.
For all $i \in \mathbb {N}$ , set $t_j = t_{j,i}$ . By Definition 5.21, for $x \in \mathcal {G}(w, (T_i), \zeta )$ and $y_i = \phi _{X_{\alpha }}^{T_i}(x)$
Note that the orbit segment $(\phi ^t_{{X_{\alpha }}}(x))_{0\leq t \leq T} $ coincides with the orbit segment $(\phi ^\tau _{{X_{\alpha }}(t_N)}(x))_{0\leq \tau \leq 1} $ of length 1 since ${X_{\alpha }}(t_N) = {X_{\alpha }}(\log T) = T{X_{\alpha }}$ . Then by Theorem 4.4,
Therefore, by the inequality (100),
Lemma 6.8. For every $r> 2(k+1)(a/2+1)+1/2 $ , there is a constant $C_r(\rho )>0$ such that for every good point $x \in \mathcal {G}(w, (T_i), \zeta )$ , we have
Proof. The orbit segment $(\phi ^t_{{X_{\alpha }}}(x))_{0\leq t \leq T} $ has length $T^{l/N}$ with respect to the generator ${X_{\alpha }}(t_{N-l}) = {X_{\alpha }}((1-l/N)\log T) = T^{1-l/N}{X_{\alpha }}$ . Thus, by Theorem 6.4 with $e^{(1-\delta (\rho ))t_{N-l}} = T^{(1-l/N)(1-\delta (\rho ))}$ , we obtain
Let $C = 2C^{(1)}_rC_r(\mathcal {O}) w^{-1/2}$ . Since $N_i = [\log T_i]$ and $N_i \leq \log T_i \leq N_i+1$ , we have ${T_i}^{1/{(N_i+1)}} \leq e \leq {T_i}^{1/{N_i}}.$ By setting $T = T_i$ , we obtain
By (88), we have $1+\unicode{x3bb} -\delta (\rho )-\rho _Y/2 \geq 1-\rho _Y/2>1/2$ , thus the geometric series converges.
Lemma 6.9. There exists a constant $C:= C(r)>0$ such that, for all $j = 0, \ldots , N,$
Proof. From (86), $t_j-t_{j-1} \leq 2$ and observe that $\mathcal {F}(t_j) = A^{t_j - t_{j-1}}\mathcal {F}(t_{j-1})$ . Passing from the frame $\mathcal {F}(t_{j-1})$ to $\mathcal {F}(t_j)$ , it can be verified that distortion of the corresponding transversal Sobolev norm is uniformly bounded.
Let
be the collection of maximal integral coadjoint orbits.
Remark 6.10. Let $\sigma = (\sigma _1, \ldots , \sigma _n) \in (0,1)^n$ be such that $\sigma _1+ \cdots + \sigma _n = 1$ . For simplicity, we choose $\sigma _i = 1/n$ from now on (see Definition 5.13 or Lemma 5.17).
Theorem 6.11. For any $\Lambda \in \widetilde M_0$ , let $\nu \in [1, 1 + (k/2-1)({1}/{n}) ]$ . Then, for any $r> (k+1)(a/2+1)+1/4$ , there exists a constant $C(\sigma , \nu )$ satisfying the following condition. For every $\epsilon> 0$ there exists a constant $K_\epsilon (\sigma , \nu )>0$ such that, for every $\alpha _1 = (\alpha _1^{(1)},\ldots ,\alpha ^{(1)}_n) \in D_n(\sigma , \nu )$ and for every $w \in (0, I(Y)^a]$ , there exists a measurable set $\mathcal {G}_{\Lambda }(\sigma ,\epsilon ,w)$ satisfying the estimate
Then for every $x \in \mathcal {G}_{\Lambda }(\sigma ,\epsilon ,w)$ , for every $f \in W^r(H_{\mathcal {O}},\mathcal {F})$ and $T \geq 1$ we have
Proof. If the coadjoint orbit $\mathcal {O}$ is integral and maximal with full rank, then the optimal exponent of $\rho = (\rho _i^{(m)})$ will be attained by the following homogeneous scaling:
Let us set $\zeta = 2\delta (\rho )/3 -{\unicode{x3bb} }/{3}$ . Given $\epsilon>0$ and for all $i \in \mathbb {N}$ , set $T_i = i^{(1+\epsilon )\zeta ^{-1}} $ . Then there exists a constant $K_\epsilon (\rho )>0$ such that
Let $\mathcal {G} = \mathcal {G}_{\Lambda }(\sigma , \epsilon , w) = \mathcal {G}(w,(T_i),\zeta )$ be the set of $(w,(T_i),\zeta )$ -good points for the basis $\mathcal {F}_{\alpha }$ . Then the estimate in formula (103) follows from Lemma 5.22 and the definition of good points. By Proposition 6.6, for all $x \in \mathcal {G}$ and for every $f \in W^r(H_{\mathcal {O}},\mathcal {F})$ , estimate (94) holds true. Given $T \in [T_i, T_{i+1}]$ ,
Let $C = C_{r}(\rho )C_r(\mathcal {O}) /w^{1/2}$ . The first term is estimated by formula (94):
For the second term, let us set $\gamma = (1+\epsilon )\zeta ^{-1}$ and observe that $\gamma ^{-1} = \zeta (1+\epsilon )^{-1} \geq (1-\epsilon )\zeta .$ We have
By the estimates on the terms $(I)$ and $(\textit{II}\kern1pt)$ , the proof is complete.
Remark 6.12. If $\mathcal {O}$ is integral but not maximal, then the restriction of $\Lambda $ factors through an irreducible representation of the ( $k-1$ )-step nilpotent group $N/\exp {n_k'}$ . Then $\mathfrak {n}/ \mathfrak {n}_k$ is a polarizing subalgebra for subrepresentation and reduces to the maximal integral case. Since the growth rate is determined by the scaling factors and the exponent $\unicode{x3bb} $ is determined by the step size and number of elements, the highest exponent is obtained by the integral maximal full rank case.
6.3 General bounds on ergodic averages
In this subsection the bounds on ergodic averages for a function on a single irreducible subrepresentation obtained in §6.2 are extended to all sufficiently smooth functions.
Definition 6.13. For every $\mathcal {O} \in \widehat M_0$ , we define $|\mathcal {O}| = \max _{\eta _i \in \mathfrak {n}_k}|\Lambda (\eta _i^{(k)})|. $
Note that $|\mathcal {O}|$ does not depend on the choice of $\Lambda $ and $|\mathcal {O}| \neq 0$ by maximality. We specifically choose an element $\eta _{*}^{(k)}$ of degree k such that
Lemma 6.14. For every $\mathcal {O} \in \widehat M_0$ and for every $\Lambda \in \mathcal {O}$ , we have
Proof. The return time of the flow $X_\alpha $ to any orbit of the codimension- $1$ subgroup $N' \subset N$ is 1. Hence, by Definition 5.7, we have $I(Y_\Lambda ) = 1/2$ for the basis. By (82), we have $C(\alpha _1)\geq 1$ . Then, from the definition of the constant $H(Y,\rho ,\alpha )$ , we obtain
Corollary 6.15. For every $\mathcal {O} \in \widehat M_0, \Lambda \in \mathcal {O}, w>0$ and $\epsilon>0$ , let
Then, for every $w>0$ and $\epsilon>0$ , the set
has measure greater than $1-Cw\epsilon ^{-1},$ with $C = 2^{-a+1}K_\epsilon (\sigma ,\nu )C(\alpha _1)(1+\log C(\alpha _1))$ . Furthermore, if $\epsilon '<\epsilon $ we have $\mathcal {G}(\sigma ,\epsilon ,w) \subset \mathcal {G}(\sigma ,\epsilon ',w)$ .
Proof. Recall that $|\Lambda (\mathcal {F})|$ is an integral multiple of $2\pi $ . By Lemma 6.14, inequality (103) and definition of $w_{\Lambda }$ , we have
where $C' = 2^{a+1}K_\epsilon (\sigma ,\nu )C(\alpha _1)(1+\log C(\alpha _1))$ . Since $|\Lambda (\mathcal {F})| = 2\pi l $ is bounded by $(2l)^{a-1}$ ,
The last statement on the monotonicity of the set follows from the analogous statement in Theorem 6.11.
In every coadjoint orbit, we will make a particular choice of a linear form to sum up estimates of the bound for each irreducible subrepresentation in terms of higher regularity of norms.
Definition 6.16. For every $\mathcal {O} \in \widehat M_0$ , we define $\Lambda _{\mathcal {O}}$ as the unique integral linear form $\Lambda \in \mathcal {O}$ such that
The existence and uniqueness of $\Lambda _{\mathcal {O}}$ follow from
and the form $\Lambda \circ \text {Ad}(\exp (tX_{\alpha })$ is integral for all integer values of $t \in \mathbb {R}$ .
Lemma 6.17. There exists a constant $C(\Lambda )>0$ such that the following holds on the primary subspace $C^{\infty }(H_{\mathcal {O}})$ . Given a basis $\mathcal{F}_{\alpha} = \mathcal{F}_{\alpha,\eta} = (X_\alpha, \eta)$ ,
Proof. Let $x_0 = -\Lambda _{\mathcal {O}}(\eta _{*}^{(k-1)})/|\mathcal {O}|$ . Then there exists a unique $\Lambda ' \in \mathcal {O}$ such that $\Lambda '(\eta ^{(k-1)}_{*}) = 0$ given by $\Lambda ' = \Lambda \circ \text {Ad}(e^{x_0X_{\alpha }})$ . The element $W \in \mathfrak {I}$ is represented in the representation as multiplication operators by the polynomials
By the definition of the linear form, the identity $[X_{\alpha }, \eta _{*}^{(k-1)}] = \eta _{*}^{(k)}$ implies
From (105), we have
Then we obtain
For any $\Lambda \in \mathfrak {n}^*$ the transversal Laplacian for a basis $\mathcal {F}$ in the representation $\pi _\Lambda $ is the operator of multiplication by the polynomial and derivative operators
Hence, for any $(m,j) \in J$ ,
In view of (106), the constant operators $\Lambda '(\eta _{j}^{(m)})$ are given by polynomial and derivative expressions of degree k in the operators $P(\Lambda ',\eta _{j}^{(m)})$ . Then we obtain the estimate
Since the representations $\pi _{\Lambda '}$ and $\pi _{\Lambda _{\mathcal {O}}}$ are unitarily intertwined by the translation operator by $x_0$ , and since constant operators commute with translations, we also have
Since $x_0$ is bounded by a constant depending only the step size k, the norms of the linear maps $\text {Ad}(\exp (\pm x_0X_{\alpha }))$ are bounded by a constant depending only on k. Therefore, $|\Lambda _{\mathcal {O}}(\mathcal{F}_{\alpha,\eta})| \leq C_2(k)|\Lambda '(\mathcal{F}_{\alpha,\eta})|$ and we finish the proof.
Corollary 6.18. There exists a constant $C'(\Lambda )$ such that for all $\mathcal {O} \in \widehat M_0$ and for any sufficiently smooth function $f \in H_{\mathcal {O}}$ ,
where $l = 2k(r-1) + (1/2)ak(k+1)$ .
Proof. From definition (93) we have $C_r(O) \leq (1+|\Lambda _{\mathcal {O}}(\mathcal{F}_{\alpha,\Lambda_O})|)^{l_1}$ with $l_1 = 4(r-1)$ . By formula (104) and inequality $|\Lambda_{O}(\mathcal{F}_{\alpha,\Lambda_O})| \leq |\Lambda_{O}(\mathcal{F}_\alpha)|$ , we have
with $l_2 = l_1 + a(k+1)$ . By Lemma 6.17 we have
Proposition 6.19. Let $r> (k+1)(a+1)+1/4$ . Let $\sigma = (1/n, \ldots , 1/n) \in (0,1)^n$ be a positive vector. Let us assume that $\nu \in [1,1+ (k/2-1){1}/{n}]$ and let $\alpha \in D_n(\sigma ,\nu )$ . For every $\epsilon>0 $ and $w>0$ , there exists a measurable set $\mathcal {G}(\sigma ,\epsilon ,w)$ satisfying
such that for every $x \in \mathcal {G}(\sigma ,\epsilon ,w)$ , $f \in W^r(M)$ and any $T \geq 1$ we have
Proof. Let $\tau := r - ak/2> (a/2+1)(k+1)+1/4 $ . Let $f \in W^\tau (M,\mathcal {F})$ and let $f = \sum _{\mathcal {O} \in \widehat M_0}f_{\mathcal {O}}$ be its orthogonal decomposition onto the primary subspace $H_{\mathcal {O}}$ . Recall that for each $\mathcal {O} \in \widehat M_0$ , the constant $w_{\Lambda_O}$ is provided in (104) and by equivalence from definition (102) the set
has measure greater than $1-Cw\epsilon ^{-1}$ by Corollary 6.15.
If $x \in \mathcal {G}(\sigma ,\epsilon ,w)$ , then by Theorem 6.11 and Corollary 6.18, for every $\mathcal {O} \in \widehat M_0$ and all $T \geq 1$ ,
For any $\tau>0$ and any $\epsilon '>0$ , by Lemma 6.17 and orthogonal splitting of $H_{\mathcal {O}}$ we have
Thus the proof ends by linearity.
Proof of Theorem 1.1
Under the same hypothesis of Proposition 6.19, for $i \in \mathbb {N}$ let $w_i = 1/2^iC$ and $\mathcal {G}_i = \mathcal {G}(\sigma ,\epsilon ,w_i)$ . Set $K_{\epsilon }(x) = 1/{w_i}^{1/2}$ if $x \in \mathcal {G}_i \backslash \mathcal {G}_{i-1}$ . By Proposition 6.19, the sets $\mathcal {G}_i$ are increasing and satisfy $\mathrm {meas}( \mathcal {G}_i^c) \leq 1/2^i\epsilon $ . Hence, the set $\mathcal {G}(\sigma ,\epsilon ) = \bigcup _{i \in \mathbb {N}}\mathcal {G}_i$ has full measure and the function $K_{\epsilon }$ is in $L^p(M)$ for every $p \in [1,2)$ .
Proof of Corollary 1.2
Recall that the k step strictly triangular nilpotent Lie algebra $\mathfrak {n}$ has dimension $\tfrac 12k(k+1)$ with one-dimensional center. If the coadjoint orbit $\mathcal {O}$ is integral and maximal, then the optimal exponent will be attained by formula (87). Recall that the scaling vector $\rho = (\ldots , \rho _i^{(m)},\ldots )$ is obtained by
and, in particular, by choosing homogeneous scaling,
Then we verify
By repeating the same argument for proof of Theorem 1.1, this concludes the proof of Corollary 1.2.
7 Uniform bound of the average width in the step- $3$ case
In this section we prove Theorem 1.3, uniform bound of the effective equidistribution of nilflows on a strictly triangular step-3 nilmanifold. On this structure it is possible to derive a uniform bound of ergodic averages under the Roth-type Diophantine condition due to the linear divergence of nearby orbits. This argument is based on counting principles of close return times which substitutes the necessity of good points (see also [Reference ForniF16] for the $3$ -step filiform case).
7.1 Average width function
Let N be a step-3 nilpotent Lie group on three generators as introduced in (3). We denote its Lie algebra $\mathfrak {n} $ , with its basis $\{X_1,X_2,X_3,Y_1,Y_2,Z\}$ satisfying the commutation relations
As introduced in §2, $ \{\phi ^t_V\}_{t\in \mathbb {R}}$ is a measure preserving flow generated by $V:=X_1+\alpha X_2+\beta X_3$ and $(1,\alpha ,\beta )$ satisfies the standard simultaneous Diophantine condition (Definition 5.13).
By definition of the average width (see Definition 4.2), for any $t \geq 0$ and for any $(x,T) \in M \times [1,+\infty )$ we will construct an open set $\Omega _{t}(x,T) \subset \mathbb {R}^6$ which contains the segment $\{(s,0,\ldots ,0)\mid 0 \leq s \leq T\}$ such that the map
is injective on $\Omega _{t}(x,T)$ . Injectivity fails if and only if there exist vectors
such that
Let us denote $r = s'-s$ and $\tilde {x_i} = {x_i}' - {x_i}, \tilde {y_i} = {y_i}' - {y_i}$ and $\tilde z = z' - z$ . Let $c_\Gamma>0$ denote the distance from the identity of the smallest non-zero element of the lattice $\Gamma $ . Let us assume that
so that $\tilde x_i, \tilde y_i \in [-({c_\Gamma }/{2}), ({c_\Gamma }/{2})]$ .
For all $t \geq 0$ and $s \in [0,T]$ , let us adopt the notation
From the identity in formula (110), we derive the identity
Projecting the above identity on the base torus, we obtain
which implies that $re^t$ is return time for the projected toral linear flow at most distant from $e^{-t/3}{c_\Gamma }/{2}$ .
Let $R_t(x,T)$ denote the set of $r \in [-T,T]$ such that equation (112) on the projected torus has a solution $\tilde x_2, \tilde x_3 \in [-({c_\Gamma }/{2}), ({c_\Gamma }/{2})]$ . Then for every $r \in R_t(x,T)$ , the solution $\tilde x_i: = \tilde x_i(r)$ of the identity in formula (110) is unique. Given $r \in R_t(x,T)$ , let $\mathcal {S}{(r)}$ be the set of $s \in [0,T]$ such that there exists a solution of identity (110) satisfying (111).
Recall that $w_{\Omega _t(r)}(s)$ is the (inner) width function along the orbit $\phi _x(s, \cdot )$ for $s \in [0,T]$ .
Lemma 7.1. The following average width estimate holds: for every $T>1$ , there exists a constant $C_\alpha>0$ such that
Proof. We approximate the average width by estimating counting close return orbits. By definition $\mathcal {S}{(r)}$ is a union of intervals $I^*$ of length at most
To count the number of such intervals, we will choose certain points where the distance is minimized. As long as $|\tilde x_2(r)| \geq e^{-5t/6}$ , for each component $I^*$ of $\mathcal {S}{(r)}$ , there exists $s^* \in I^*$ solution of the equation $\tilde {y_1}(t,s) = \tilde {y_1}(r)+e^{(5/6)t}s\tilde {x_2}(r)$ (see Figure 2). The same argument holds for $|\tilde x_3(r)|$ . Let $\mathcal {S}^*(r)$ be the set of all such solutions. Its cardinality can be estimated by counting points.
Claim. There exists a constant $C_\alpha>0$ such that
Proof. Note that $s^*$ is a point which minimizes the distance between an orbit and its close return, say,
If either distance $|\tilde {y_1}(t,s)|$ or $|\tilde {y_2}(t,s)|$ dominates the other, then it reduces to simply finding a solution to single equation. Otherwise, we assume $|\tilde {y_1}(t,s)| = |\tilde {y_2}(t,s)|$ and solve the equation for s. We distinguish the following two cases, but in either case we can restrict either $\tilde y_1(r) = 0$ or $\tilde y_2(r) = 0$ for convenience.
By solving the equation $|\tilde {y_1}(t,s)| = |\tilde {y_2}(t,s)|$ , there exists $s^* \in I^*$ such that
From the bound
we obtain
Thus we prove the claim.
For every $r \in R_t(x,T)$ and every $s \in [0,T]$ , we define the function
and set
Now we define the set of narrow width by
Under the above construction, the map $\phi _x$ is injective on $\Omega _t(x,T)$ . The open set $\Omega _t(r) \cap \Omega _t(-r) $ are narrowed near both endpoints of the return time r so that their images in M have no self-intersections under return times r and $-r$ .
By the definition of inner width and by construction of the set $\Omega _t(r)$ we have that
It follows that for every subinterval $I^* \subset S(r)$ we have (using the definition of $\delta _{r}$ )
By the upper bound on the length of interval $I^*$ and on the cardinality of the set $\mathcal {S}^*(r)$ we finally derive the conclusion.
Recall from Definition 5.13 that we choose simultaneously Diophantine number $\alpha \in \mathbb {R}^2\backslash \mathbb {Q}^2$ of exponent $\nu \geq 1$ .
Lemma 7.2. Given the Diophantine condition of exponent $\nu \geq 1$ , there exists a constant $C_{\alpha }:= C(\alpha )>0$ such that all solutions of formula (112) satisfy the lower bound
Proof. By projected identity (112) on the base 3-torus, assume
Then we set $re^t =q \in \mathbb {Z}$ and there exists $(p_1,p_2)\in \mathbb {Z}^2$ such that $ p_1-q\alpha = e^{-t/3} \tilde x_2,\ p_2-q\beta = e^{-t/3} \tilde x_3$ . By the Diophantine condition, there exists a constant $C(\alpha )$ such that
which proves the statement.
For every $n \in \mathbb {N}$ , let $R^{(n)}_t(x,T) \subset R_t(x,T) $ be characterized by
Lemma 7.3. For all $\epsilon>0$ , if the frequency of the projected linear flow satisfies the Diophantine condition of exponent $ \nu = \sqrt {2}+\epsilon $ , then there exists $C_\epsilon>0$ such that
Proof. Under a Diophantine condition of exponent $\nu \geq 1$ , from inequality (80) and the definition of $R^{(n)}_t(x,T)$ , we have
It suffices to show $(Te^t)^{1-({1}/{\nu })} $ is less than or equal to the desired bound. From Lemma 7.2,
In the limit as $\nu \rightarrow \sqrt {2}$ ,
Approximating $\|(x_2,x_3)\| \sim 1/ 2^n$ ,
By combining counting return time and width estimates, we obtain a uniform bound.
Proposition 7.4. There exists a constant $C_\epsilon (V)>0$ such that
Denote the average width of the orbit segment with length 1 by
Corollary 7.5. Let $\phi ^V_t$ be a nilflow on step-3 strictly triangular M generated by V such that the projected flow on $\mathbb {T}^3$ satisfies the Diophantine condition of Roth type with $\nu \in [1, \sqrt {2}+\epsilon ]$ . For every $\epsilon>0$ there exists a constant $C_{\epsilon }(V)>0$ such that
Proof of Theorem 1.3
By Corollary 7.5, we do not rely on the good points technique and Lyapunov norm. An improved bound of remainder term R in Theorem 6.4 can be obtained:
We revisit the backward iteration scheme introduced in the proof of Theorem 6.6. We have
and
Changing the length to 1 and by uniform width bound from Corollary 7.5,
Then, by inductive argument resembling (98),
Therefore
Finally, we sum up all the functions on irreducible representation $H_{\mathcal {O}}$ , which only increase the regularity accordingly.
8 Application: mixing of nilautomorphisms
In this section, as a further application of main equidistribution results, we verify an explicit bound for the rate of exponential mixing of hyperbolic automorphisms relying on a renormalization argument.
Let $\mathfrak {F}_{2,3}= \{X_1, X_2, Y_1, Z_1, Z_2\}$ be a step-3 free nilpotent Lie algebra with two generators with commutation relations
The group of automorphisms on Lie algebras induces an automorphism on the nilmanifold
and we consider a hyperbolic automorphism T with an eigenvalue $\unicode{x3bb}>1$ with corresponding eigenvector $V = X_1 + \alpha X_2$ satisfying Diophantine condition $(1,\alpha )$ on base torus $\mathbb {T}^2$ . By direct computation, the following renormalization holds:
Theorem 8.1. Let $(\phi _{V}^t)$ be a nilflow on a 3-step nilmanifold $M = \mathfrak {F}_{2,3} / \Gamma $ such that the projected toral flow $(\bar \phi _V^t)$ is a linear flow with frequency vector $v:= (1,\alpha )$ in the Roth-type Diophantine condition (with exponent $\nu = 1+\epsilon $ for all $\epsilon>0$ ). For every $s> 12$ , there exists a constant $C_s$ such that for every zero-average function $f \in W^s(M)$ , for all $(x,T) \in M \times \mathbb {R}$ , we have
The detailed computation follows similarly to §7. The only difference with respect to the step-3 filiform case [Reference ForniF16] is that it has an extra element in the center which is redundant in the actual calculation on width, only raising required regularity of zero-average function.
The proposition below was first proved by Gorodnik and Spatzier in [Reference Gorodnik and SpatzierGS14].
Proposition 8.2. The hyperbolic nilautomorphism T is exponential mixing.
Proof. Let $f, g \in C^1(M)$ be smooth. Define $\langle f,g \rangle = \int _M fg d\mu. $ Since the Haar measure is invariant under $\phi _V^t$ ,
By integration by parts,
Therefore,
By renormalizing the flow,
and
Therefore, by the result of equidistribution (123),
Acknowledgement
The author deeply appreciates Giovanni Forni for his illuminating suggestions and guidance. He is grateful to Livio Flaminio and Rodrigo Treviño for making several comments which improved an earlier draft of this paper. He further thanks Krzysztof Frączek for careful advice on the revision of this work. Part of the paper was written when the author visited the Institut de Mathematiques de Jussieu-Paris Rive Gauche in Paris, France. He also acknowledges Anton Zorich for hospitality during the visit. He is grateful to Xuesen Na, Davi Obata, and Davide Ravotti for helpful discussions and thanks Jacky Jia Chong, J.T. Rustad and Lucia D. Simonelli for their encouragement. Lastly, the author is grateful to the referee for helpful comments and suggestions for improvement in the presentation of this work. This research was partially supported by the NSF grant DMS 1600687 and by the Center of Excellence ‘Dynamics, mathematical analysis and artificial intelligence’ at Nicolaus Copernicus University in Toruń.
A Appendix
In this appendix we introduce a specific example of a nilpotent Lie algebra which goes beyond our approach introduced in §5.
A.1 Free group type of step 5 with three generators
In this example, we will show the failure of transversality condition. This only means that we cannot apply our theorem but we do not know whether the conclusion holds or not.
Let $\mathfrak {F}_{n}$ be a free nilpotent Lie algebra with n generators and $(\mathfrak {F}_{n})_{k+1}$ be the ( $k+1$ )th subalgebra in the central series, following the notation in (4). Denote by $\mathfrak {F}_{n,k}:= \mathfrak {F}_{n}/(\mathfrak {F}_{n})_{k+1}$ the quotient of the free algebra with n generators $\mathfrak {F}_{n}$ , which is finite- dimensional.
Definition A.1. Let $\mathfrak {n}$ be a nilpotent Lie algebra satisfying the generalized transversality condition if there exists basis $(X_\alpha , Y_\Lambda )$ of $\mathfrak {n}$ for each irreducible representation $\pi _\Lambda ^{X_\alpha }$ such that
where $C_{\mathfrak {I}}(\pi _\Lambda ^{X_{\alpha }}) = \{Y \in \mathfrak {I} \mid \Lambda ([Y,X_{\alpha }]) = 0\}$ .
The generalized transversality condition implies the existence of a completed basis for each irreducible representation $\pi _\Lambda ^{X_\alpha }$ of non-zero degree. That is, given the adapted basis $\mathcal {F} = (X,Y_1,\ldots , Y_a)$ , there exists reduced system $\bar {\mathcal {F}} = (X, Y_1',\ldots , Y^{\prime }_{a'})$ satisfying transversality condition (42) and $\pi _\Lambda ^X(Y^{\prime }_{m}) = 0$ for all $a' \leq m \leq a$ .
Now we will investigate an example that fails the transversality condition as well as that in the sense of representation.
Let $\mathcal {F} = (X, Y^{(j)}_{i})$ be basis of $\mathfrak {F}_{5,3}$ with generators $\{X_1, X_2, X_3\}$ with the following relations:
with
The rest of the elements in $\mathcal{F}_{5,3}$ are generated by some commutation relations but these are not listed here. In general, we write elements $Y^{(i)}_j \in \mathfrak {n}_{i}\backslash \mathfrak {n}_{i+1}$ and $Y^{(5)}_i \in Z(\mathfrak {n})$ for all i. By the Jacobi identity
For fixed $\alpha _i$ and $\beta _i$ , let
and set to be the $\mathfrak {I}$ ideal of $\mathfrak {F}_{5,3}$ codimension 1, not containing V.
Proposition A.2. $\mathfrak {F}_{5,3}$ does not satisfy the generalized transversality condition for some irreducible representation.
Proof. To find the centralizer in the Lie algebra, for $a_i, b_i \in \mathbb {R}$ , set
Then it contains
By linear independence, all the coefficients vanish and we are left with
Therefore, there is no non-trivial element in $C_{\mathfrak {I}}(V) \cap {\mathfrak {n}_2 \backslash \mathfrak {n}_3}$ . Since the range of $\text {ad}_V$ has rank 2, this model does not satisfy the transversality condition.
We now verify that the generalized transversality condition is not satisfied on some irreducible representation. By Schur’s lemma, an irreducible representation $\pi _{\Lambda }^{V}$ acts as a constant on the center $Z(\mathfrak {n})$ .
Assume $\pi _*(W_i) = s_i I \neq 0$ for some $W_i \in Z(\mathfrak {n})$ . Then it is possible to choose an element $L_i \in {\mathfrak {n}_2 \backslash \mathfrak {n}_3}$ such that
where $(a_i,b_i,c_i)$ are non-proportional for each i, and
However, on the given irreducible representation, any linear combination of $L_1, L_2$ and $L_3$ does not give a trivial relation. If $s_1 L_1 + s_2 L_2 + s_3 L_3 \in C_{\mathfrak {I}}(\pi _\Lambda ^{V})$ , then
The system of equations has trivial solution $(t=0)$ by linear independence of each coefficient. Then there does not exist any element of $\mathfrak {n}_2 \backslash \mathfrak {n}_3$ that has degree 0. However, the range of $\text {ad}_V$ has rank 2 and the generalized transversality condition cannot be satisfied in this example.