1 Introduction

In this paper we prove a weak type (1, 1) theorem for the maximal operator associated to a general Ornstein–Uhlenbeck semigroup. We extend the proof given by the third author in 1983 in a symmetric context. Our setting is the following.

In \(\mathbb {R}^n\) we will consider the semigroup generated by the elliptic operator

$$\begin{aligned} \mathcal L= \frac{1}{2} \sum _{i,j=1}^n q_{ij} \frac{\partial ^2}{\partial x_i \partial x_j} + \sum _{i,j=1}^n b_{ij} x_i\frac{\partial }{ \partial x_j}, \end{aligned}$$

or, equivalently,

$$\begin{aligned} \mathcal L= \frac{1}{2}\, \mathrm {tr} \big ( Q\nabla ^2 \big )+\langle Bx, \nabla \rangle ,\end{aligned}$$

where \(\nabla \) is the gradient and \(\nabla ^2\) the Hessian. Here \(Q = (q_{ij})\) is a real, symmetric and positive definite \(n\times n\) matrix, indicating the covariance of \(\mathcal L\). The real \(n\times n\) matrix \(B = (b_{ij})\) is negative in the sense that all its eigenvalues have negative real parts, and it gives the drift of \(\mathcal L\).

The semigroup is formally \( \mathcal H_t=e^{t\mathcal L}\), \({t> 0}\), but to write it more explicitly we first introduce the positive definite, symmetric matrices

(1.1)

and the normalized Gaussian measures \(\gamma _t \) in \(\mathbb {R}^n\), with \(t\in (0,+\infty ]\), having density

$$\begin{aligned} y\mapsto (2\pi )^{-\frac{n}{2}} (\text {det} \, Q_t)^{-\frac{1}{2} } \exp \left( {-\frac{1}{2} \langle Q_t^{-1}y,y\rangle }\right) \end{aligned}$$

with respect to Lebesgue measure. Then for functions f in the space of bounded continuous functions in \(\mathbb {R}^n\) one has

$$\begin{aligned} \mathcal H_t f(x)= \int f(e^{tB}x-y)\,d\gamma _t (y)\,, \quad x\in \mathbb {R}^n\,, \end{aligned}$$
(1.2)

a formula due to Kolmogorov. The measure \(\gamma _\infty \) is invariant under the action of \(\mathcal H_t\); it will be our basic measure, replacing Lebesgue measure.

We remark that \(\big ( \mathcal H_t \big )_{t> 0}\) is the transition semigroup of the stochastic process

$$\begin{aligned}\chi (x,t)= e^{tB}+ \int _0^t e^{(t-s)B} \, dW(s),\end{aligned}$$

where W is a Brownian motion in \(\mathbb {R}^n\) with covariance Q.

We are interested in the maximal operator defined as

$$\begin{aligned} \mathcal H_* f(x)= \sup _{t> 0} \big | \mathcal H_t f(x) \big |. \end{aligned}$$

Under the above assumptions on Q and B, our main result is the following.

Theorem 1.1

The Ornstein–Uhlenbeck maximal operator \(\mathcal H_*\) is of weak type (1, 1) with respect to the invariant measure \(\gamma _\infty \), with an operator quasinorm that depends only on the dimension and the matrices Q and B.

In other words, the inequality

$$\begin{aligned} \gamma _\infty \{x\in \mathbb {R}^n : \mathcal H_*f(x)> \alpha \} \le \frac{C}{\alpha }\,\Vert f\Vert _{L^1( \gamma _\infty )},\qquad \alpha >0, \end{aligned}$$
(1.3)

holds for all functions \(f\in L^1 (\gamma _\infty )\), with \(C=C(n,Q,B)\).

For large values of the time parameter, we also obtain a refinement of this result. Indeed, we prove in Proposition 6.1 that

$$\begin{aligned} \gamma _\infty \left\{ x\in \mathbb {R}^n : \sup _{t> 1}|\mathcal H_t f(x)| > \alpha \right\} \le \frac{C}{ \alpha \ {\sqrt{\log \alpha }}}\, \end{aligned}$$
(1.4)

for large \(\alpha >0\) and all normalized functions \(f\in L^1 (\gamma _\infty )\). Here \(C=C(n,Q,B)\), and this estimate is shown to be sharp. It cannot be extended to \(\mathcal H_*\), since the maximal operator corresponding to small values of t only satisfies the ordinary weak type inequality. This sharpening is not surprising, in the light of some recent results for the standard case \(Q=I\) and \(B=-I\) by Lehec [8]. He proved the following conjecture, proposed by Ball, Barthe, Bednorz, Oleszkiewicz and Wolff [2]:

For each fixed \(t>0\), there exists a function \(\psi _t=\psi _t(\alpha )\), with \(\displaystyle \lim _{\alpha \rightarrow +\infty } \psi _t (\alpha )=0\), satisfying

$$\begin{aligned} \gamma _\infty \{x\in \mathbb {R}^n : |\mathcal H_t f(x)| > \alpha \} \le \frac{\psi _t (\alpha )}{\alpha }\, \end{aligned}$$

for all large \(\alpha >0\) and all \(f\in L^1(\gamma _\infty )\) such that \( \Vert f\Vert _{L^1( \gamma _\infty )}=1\). Lehec proved this conjecture with \(\psi _t (\alpha )={C(t)}/{\sqrt{\log \alpha }}\) independent of the dimension, and this \(\psi _t\) is sharp. Our estimates depend strongly on the dimension n, but on the other hand we estimate the supremum over large t.

The history of \(\mathcal H_*\) is quite long and started with the first attempts to prove \(L^p\) estimates. When \(\big ( \mathcal H_t \big )_{t> 0}\,\) is symmetric, i.e., when each operator \(\mathcal H_t \) is self-adjoint on \(L^2 (\gamma _\infty )\), then \(\mathcal H_*\) is bounded on \(L^p (\gamma _\infty )\) for \(1<p\le \infty \), as a consequence of the general Littlewood–Paley–Stein theory for symmetric semigroups of contractions on \(L^p\) spaces [16,   Ch. III].

It is easy to see that the maximal operator is unbounded on \(L^1 (\gamma _\infty )\). This led, about fifty years ago, to the study of the weak type (1, 1) of \(\mathcal H_*\) with respect to \(\gamma _\infty \). The first positive result is due to B. Muckenhoupt [13], who proved the estimate (1.3) in the one-dimensional case with \(Q=I\) and \(B=-I\). The analogous question in the higher-dimensional case was an open problem until 1983, when the third author [15] proved the weak type (1, 1) in any finite dimension. Other proofs are due to Menárguez, Pérez and Soria [11] (see also [10, 14]) and to Garcìa-Cuerva, Mauceri, Meda, Sjögren and Torrea [7]. Moreover, a different proof of the weak type (1, 1) of \(\mathcal H_*\), based on a covering lemma halfway between covering results by Besicovitch and Wiener, was given by Aimar, Forzani and Scotto [1]. A nice overview of the literature may be found in [17,   Ch.4].

In [4] the present authors recently considered a normal Ornstein–Uhlenbeck semigroup in \(\mathbb {R}^n\), that is, we assumed that \(\mathcal H_t\) is for each \(t> 0\) a normal operator on \(L^2 (\gamma _\infty )\). Under this extra assumption, we proved that the associated maximal operator is of weak type (1, 1) with respect to the invariant measure \(\gamma _\infty \). This extends earlier work in the non-symmetric framework by Mauceri and Noselli [9], who proved that if \(Q=I\) and \(B=\lambda (R-I)\) for some positive \(\lambda \) and a real skew-symmetric matrix R generating a periodic group, then the maximal operator \( \mathcal H_* \) is of weak type (1, 1).

In Theorem 1.1 we go beyond the hypothesis of normality. The proof has a geometric core and relies on the ad hoc technique developed by the third author in [15]. It is worth noticing that, while the proof in [4] required an analysis of the special case when \(Q=I\) and \(B=(-\lambda _1, \ldots , -\lambda _n)\), with \(\lambda _j>0\) for \(j=1, \ldots , n\), and then the application of factorization results, we apply here directly, avoiding many intermediate steps, the "forbidden zones" technique introduced in [15].

Since the maximal operator \(\mathcal H_* \) is trivially bounded from \(L^{\infty }\) to \(L^{\infty }\), we obtain by interpolation the following corollary.

Corollary 1.2

The Ornstein–Uhlenbeck maximal operator \(\mathcal H_*\) is bounded on \(L^p (\gamma _\infty )\) for all \(p>1\).

This result improves Theorem 4.2 in [9], where the \(L^p\) boundedness of \(\mathcal H_*\) is proved for all \(p>1\) in the normal framework, under the additional assumption that the infinitesimal generator of \(\big ( \mathcal H_t \big )_{t> 0}\,\) is a sectorial operator of angle less than \(\pi /2\).

In this paper we focus our attention on the Ornstein–Uhlenbeck semigroup in \(\mathbb {R}^n\). In view of possible applications to stochastic analysis and to SPDE’s, it would be very interesting to investigate the case of the infinite-dimensional Ornstein-Uhlenbeck maximal operator as well (see [3, 6, 18] for an introduction to the infinite-dimensional setting). The Riesz transforms associated to a general Ornstein–Uhlenbeck semigroup in \(\mathbb {R}^n\) have been studied in the authors’ paper [5].

The scheme of the paper is as follows. In Sect. 2 we introduce the Mehler kernel \(K_t(x,u)\), that is, the integral kernel of \(\mathcal H_t\). Some estimates for the norm and the determinant of \(Q_t\) and related matrices are provided in Sect. 3. As a consequence, we obtain bounds for the Mehler kernel. In Sect. 4 we consider the relevant geometric features of the problem, and introduce in Sect. 4.1 a system of polar-like coordinates. We also express Lebesgue measure in terms of these coordinates. Sections 5, 6, 7 and 8 are devoted to the proof of Theorem 1.1. First, Sect. 5 introduces some preliminary simplifications of the proof; in particular, we restrict the variable x to an ellipsoidal annulus. In Sect. 6 we consider the supremum in the definition of the maximal operator taken only over \(t> 1\) and prove the sharp estimate (1.4). Section 7 is devoted to the case of small t under an additional local condition. Finally, in Sect. 8 we treat the remaining case and conclude the proof of Theorem 1.1, by proving the estimate (1.3) for small t under a global assumption.

In the following, we use the “variable constant convention”, according to which the symbols \(c>0\) and \(C<\infty \) will denote constants which are not necessarily equal at different occurrences. They all depend only on the dimension and on Q and B. For any two nonnegative quantities a and b we write \(a\lesssim b\) instead of \(a \le C b\) and \(a \gtrsim b\) instead of \(a \ge c b\). The symbol \(a\simeq b\) means that both \(a\lesssim b\) and \(a \gtrsim b\) hold.

By \(\mathbb {N}\) we mean the set of all nonnegative integers. If A is an \(n\times n\) matrix, we write \(\Vert A\Vert \) for its operator norm on \(\mathbb {R}^n\) with the Euclidean norm \(|\cdot |\).

2 The Mehler kernel

For \(t>0\), the difference

$$\begin{aligned} Q_\infty -Q_t =\int _t^{\infty }e^{sB}Qe^{sB^*}ds \end{aligned}$$
(2.1)

is a symmetric and strictly positive definite matrix. So is the matrix

$$\begin{aligned} Q_t ^{-1} - Q_\infty ^{-1} = Q_t^{-1}( Q_\infty -Q_t )Q_\infty ^{-1}, \end{aligned}$$
(2.2)

and we can define

$$\begin{aligned} D_t = (Q_t^{-1}-Q_\infty ^{-1} )^{-1} Q_t^{-1} e^{tB}\,,\qquad t>0. \end{aligned}$$
(2.3)

Then formula (1.2), the definition of the Gaussian measure and some elementary computations yield

$$\begin{aligned} \mathcal H_t f(x)&= (2\pi )^{-\frac{n}{2}} (\text {det} \, Q_t)^{-\frac{1}{2} } \int f(e^{tB}x-y) \exp \left[ {-\frac{1}{2} \langle Q_t^{-1}y,y\rangle }\right] dy\nonumber \\&= \Big ( \frac{\text {det} \, Q_\infty }{\text {det} \, Q_t} \Big )^{{1}/{2} } \exp \left[ {\frac{1}{2} \langle Q_t^{-1} e^{tB}x, D_t\, x-e^{tB}x\rangle }\right] \nonumber \\&\quad \times \int f(u) \, \exp \left[ {\frac{1}{2} \langle (Q_\infty ^{-1}- Q_t^{-1}) (u-D_t\, x)\,,\, u-D_t\, x\rangle }\right] d\gamma _\infty (u)\,, \end{aligned}$$
(2.4)

where we repeatedly used the fact that \(Q_\infty ^{-1}- Q_t^{-1}\) is symmetric. We now express the matrix \(D_t\) in various ways.

Lemma 2.1

For all \(x\in \mathbb {R}^n\) and \(t>0\) we have

  1. (i)

    \( D_t = Q_\infty e^{-tB^*} Q_\infty ^{-1} \);

  2. (ii)

    \(D_t = e^{tB} + Q_t e^{-tB^*}Q_\infty ^{-1}\).

Proof

  1. (i)

    The formulae (2.1) and (1.1) imply

    $$\begin{aligned} Q_\infty - Q_t= e^{tB} Q_\infty e^{tB^*} \end{aligned}$$
    (2.5)

    (see also [12,   formula (2.1)]). From (2.3) and (2.2) it follows that

    $$\begin{aligned} {D_t = Q_\infty ( Q_\infty -Q_t)^{-1}\, e^{tB},}\end{aligned}$$

    and combining this with (2.5) we arrive at (i).

  2. (ii)

    Multiplying (2.5) by \( e^{-tB^*} Q_\infty ^{-1}\) from the right, we obtain

    $$\begin{aligned} Q_\infty e^{-tB^*}Q_\infty ^{-1}- Q_t e^{-tB^*}Q_\infty ^{-1} = e^{tB}, \end{aligned}$$

    and (ii) now follows from (i).

\(\square \)

By means of (i) in this lemma, we can define \(D_t\) for all \(t\in \mathbb {R}\), and they will form a one-parameter group of matrices.

Now (ii) in Lemma 2.1 yields

$$\begin{aligned} \langle Q_t^{-1} e^{tB}x, D_t\, x-e^{tB}x\rangle = \langle Q_t^{-1} e^{tB}x, Q_t e^{-tB^*}Q_\infty ^{-1}x \rangle =\langle Q_\infty ^{-1}x ,x \rangle . \end{aligned}$$

Thus (2.4) may be rewritten as

$$\begin{aligned} \mathcal H_t f(x)&= \int K_t (x,u)\, f(u)\, d\gamma _\infty (u) \,, \end{aligned}$$

where \(K_t\) denotes the Mehler kernel, given by

(2.6)

for \(x,u\in \mathbb {R}^n\). Here we introduced the quadratic form

3 Some auxiliary results

In this section we collect some preliminary bounds, which will be essential for the sequel.

Lemma 3.1

For \(s>0\) and for all \(x\in \mathbb {R}^n\) the matrices \( D_{s }\) and \( D_{-s }= D_{s }^{-1}\) satisfy

$$\begin{aligned} e^{cs}|x| \lesssim |D_s\, x| \lesssim e^{Cs} |x|, \end{aligned}$$

and

$$\begin{aligned} e^{-Cs}|x| \lesssim |D_{-s}\, x| \lesssim e^{-cs} |x|. \end{aligned}$$

This also holds with \( D_{s }\) replaced by \(e^{-sB}\) and \(e^{-sB^*}\).

Proof

We make a Jordan decomposition of \(B^*\), thus writing it as the sum of a complex diagonal matrix and a triangular, nilpotent matrix, which commute with each other. This leads to expressions for \(e^{-sB^*}\) and \(e^{sB^*}\), and since \(B^*\) like B has only eigenvalues with negative real parts, we see that

$$\begin{aligned} \Vert e^{-sB^*} \Vert \lesssim e^{Cs} \qquad \text { and } \qquad \Vert e^{sB^*} \Vert \lesssim e^{-cs}. \end{aligned}$$
(3.1)

From (i) in Lemma 2.1, we now get the claimed upper estimates for \(D_{\pm s}\). To prove the lower estimate for \(D_s\), we write

$$\begin{aligned} |x| = |D_{-s}\, D_s\, x| \lesssim e^{-cs} |D_s\, x|. \end{aligned}$$

The other parts of the lemma are completely analogous. \(\square \)

In the following lemma, we collect estimates of some basic quantities related to the matrices \(Q_t\).

Lemma 3.2

For all \(t>0\) we have

  1. (i)

    \(\det { \, Q_t} \simeq (\min (1,t))^{n}\);

  2. (ii)

    \(\Vert Q_t^{-1}\Vert \simeq (\min (1,t))^{-1}\);

  3. (iii)

    \(\Vert Q_\infty -Q_t \Vert \lesssim e^{-ct}\);

  4. (iv)

    \(\Vert Q_t^{-1}-Q_\infty ^{-1} \Vert \lesssim {t}^{-1}\,{e^{-ct}}\);

  5. (v)

    \(\Vert \left( Q_t^{-1}-Q_\infty ^{-1} \right) ^{-1/2}\Vert \lesssim {t^{1/2}}\, e^{Ct} \).

Proof

(i) and (ii) Using (3.1), we see that for each \(t>0\) and for all \(v\in \mathbb {R}^n\)

$$\begin{aligned} \langle Q_{t} v ,v\rangle&= \left\langle \int _0^t e^{sB} Q e^{sB^*}v \, ds,v\right\rangle =\int _0^t \langle Q^{1/2} e^{sB^*}v , Q^{1/2} e^{sB^*} v\rangle \, ds\\ {}&= \int _0^t \big | Q^{1/2} e^{sB^*}v \big |^2 \, ds\simeq \int _0^t \big | e^{sB^*}v \big |^2 \, ds\\ {}&\lesssim \int _0^t e^{-cs} \, ds \,|v|^2 \simeq \min (1,t)\, |v|^2. \end{aligned}$$

Since \(\Vert \left( e^{s B^*}\right) ^{-1}\Vert = \Vert e^{-s B^*}\Vert \lesssim e^{Cs}\), there is also a lower estimate

$$\begin{aligned} \int _0^t&\big | e^{sB^*}v \big |^2 \, ds \gtrsim \int _0^t e^{-Cs} \, ds \, |v|^2 \simeq \min (1,t) |v|^2. \end{aligned}$$

Thus any eigenvalue of \(Q_t\) has order of magnitude \(\min (1,t)\), and (i) and (ii) follow.

(iii) From the definition of \(Q_t\) and (3.1), we get

$$\begin{aligned} \Vert Q_\infty -Q_t \Vert =\left\| \int _t^\infty e^{sB}Qe^{sB^*}ds\right\| \lesssim e^{-ct}. \end{aligned}$$

(iv) Using now (ii) and (iii), we have

$$\begin{aligned} \Vert Q_t^{-1}-Q_\infty ^{-1}\Vert&=\Vert Q_t^{-1}(Q_\infty -Q_t) Q_\infty ^{-1} \Vert \lesssim \Vert Q_t^{-1}\Vert \, \Vert Q_\infty -Q_t\Vert \\&\lesssim (\min (1,t))^{-1}\, e^{-ct} \lesssim t^{-1}\, e^{-ct}. \end{aligned}$$

(v) Since \(\Vert A^{1/2}\Vert = \Vert A\Vert ^{1/2}\) for any symmetric positive definite matrix A, we consider \((Q_t^{-1}-Q_\infty ^{-1})^{-1}\), which can be rewritten as

$$\begin{aligned} (Q_t^{-1}-Q_\infty ^{-1})^{-1}= (Q_\infty ^{-1}(Q_\infty -Q_t) Q_t^{-1})^{-1} =Q_t(Q_\infty -Q_t)^{-1}Q_\infty . \end{aligned}$$
(3.2)

It follows from (2.5) that \((Q_\infty - Q_t)^{-1}= e^{-tB^*} Q_\infty ^{-1}e^{-tB},\)

so that

$$\begin{aligned}\Vert (Q_\infty - Q_t)^{-1}\Vert \lesssim e^{Ct} \end{aligned}$$

as a consequence of (3.2). Inserting this and the simple estimate \(\Vert Q_t \Vert \lesssim t\) in (3.2), we obtain \(\Vert (Q_t^{-1}-Q_\infty ^{-1})^{-1}\Vert \lesssim t e^{Ct}\), and (v) follows. \(\square \)

Proposition 3.3

For \(t\ge 1\) and \(w\in \mathbb {R}^n\), we have

$$\begin{aligned} \langle (Q_t^{-1}- Q_\infty ^{-1} ) D_t \, w,\, D_t \, w \rangle&\simeq |w|^2. \end{aligned}$$

Proof

By (2.3) and Lemma 2.1 (i) we have

$$\begin{aligned} \langle (Q_t^{-1}- Q_\infty ^{-1} ) D_t \, w,\, D_t \, w \rangle&= \langle Q_t^{-1} e^{tB} w\,,\, Q_\infty e^{-tB^*} Q_\infty ^{-1} \,w \rangle \\&= \langle Q_\infty Q_t^{-1} e^{tB} w\,,\, e^{-tB^*} Q_\infty ^{-1} \,w \rangle . \end{aligned}$$

Since \(Q_\infty Q_t^{-1}=I+(Q_\infty -Q_t)Q_t^{-1}\), this leads to

$$\begin{aligned}&\langle (Q_t^{-1}- Q_\infty ^{-1} ) D_t \, w,\, D_t \, w \rangle \\&\quad = \langle e^{tB} w\,,\, e^{-tB^*} Q_\infty ^{-1} \,w \rangle + \langle ( Q_\infty - Q_t) Q_t^{-1} e^{tB} w\,,\, e^{-tB^*} Q_\infty ^{-1} \,w \rangle \\&\quad = \langle Q_\infty ^{-1} w, w\rangle + \langle e^{-tB} ( Q_\infty - Q_t) Q_t^{-1} e^{tB} w\,,\, Q_\infty ^{-1} \,w \rangle . \end{aligned}$$

Here \( \langle Q_\infty ^{-1} w, w\rangle \simeq |w|^2\). Using (2.1) and then the definition of \( Q_\infty \), we observe that the last term can be written as

$$\begin{aligned}&\left\langle \int _t^\infty e^{(s-t)B}Qe^{(s-t)B^*} \,ds\,e^{tB^*} \, Q_t^{-1} e^{tB} w \,,\, Q_\infty ^{-1} \,w \right\rangle \nonumber \\&\quad = \big \langle Q_\infty \,e^{tB^*} \, Q_t^{-1} e^{tB} w\, \,,\, Q_\infty ^{-1} \,w \big \rangle \nonumber \\&\quad = \langle \,e^{tB^*} \, Q_t^{-1} e^{tB} w\, \,,\, \,w \rangle \nonumber \\&\quad = \, \, \big |Q_t^{-1/2} e^{tB} w \big |^2. \end{aligned}$$
(3.3)

Since \(\big |Q_t^{-1/2} e^{tB} w \big |^2 \lesssim |w|^2 \) for \(t\ge 1\) by Lemmata 3.1 and 3.2 (ii), the proposition follows.\(\square \)

We finally give estimates of the kernel \(K_t\), for small and large values of t. When \(t\le 1\), one has \(\Vert (Q_t^{-1}-Q_\infty ^{-1})^{1/2}\Vert \simeq t^{-1/2} \) and \(\Vert (Q_t^{-1}-Q_\infty ^{-1})^{-1/2}\Vert \simeq t^{1/2}\), by (iv) and (v) in Lemma 3.2. Combined with (2.6), this implies

$$\begin{aligned} \frac{ e^{R( x)}}{t^{n/2}}\exp \left( -C\,\frac{|u-D_t \,x |^2}{t}\right) \lesssim K_t(x,u) \lesssim \frac{ e^{R( x)}}{t^{n/2}} \exp \left( -c\,\frac{|u-D_t\, x |^2}{t}\right) , \quad \; 0 < t\le 1.\nonumber \\ \end{aligned}$$
(3.4)

Lemma 3.4

For \(t\ge 1\) and \(x,u\in \mathbb {R}^n\), we have

$$\begin{aligned} e^{R(x)} \exp \Big [ -C \big | D_{-t}\,u- x \big |^2 \Big ] \lesssim K_t (x,u)&\lesssim e^{R(x)} \exp \Big [ - c\big | D_{-t}\,u- x\big |^2 \Big ]. \end{aligned}$$
(3.5)

Proof

This follows from (2.6), if we write \(u-D_t\, x= D_t (D_{-t}\,u- x)\) and apply Proposition 3.3 with \(w=D_{-t}\,u- x\). \(\square \)

4 Geometric aspects of the problem

4.1 A system of adapted polar coordinates

We first need a technical lemma.

Lemma 4.1

For all x in \(\mathbb {R}^n\) and \(s\in \mathbb {R}\), we have

$$\begin{aligned}&\langle B^* Q_\infty ^{-1} x, x\rangle = - \frac{1}{2}\, |Q^{1/2}\, Q_\infty ^{-1} x|^2; \end{aligned}$$
(4.1)
$$\begin{aligned}&\frac{\partial }{\partial s} D_s\, x = - Q_\infty B^*\,Q_\infty ^{-1} D_s\, x = -Q_\infty e^{-sB^*}B^*Q_\infty ^{-1} x; \end{aligned}$$
(4.2)
$$\begin{aligned}&\frac{\partial }{\partial s} R\big ( D_s\, x \big ) =\frac{1}{2}\, \big | Q^{1/2} Q_\infty ^{-1} D_s\, x\big |^2 \simeq \big | D_s\, x\big |^2 . \end{aligned}$$
(4.3)

Proof

To prove (4.1), we use the definition of \(Q_\infty \) to write for any \(z\in \mathbb {R}^n\)

$$\begin{aligned} \langle B^* z, Q_\infty z\rangle&= \int _0^\infty \langle B^* z, e^{sB}\,Q\, e^{sB^*}z\rangle \,ds\\&= \int _0^\infty \langle e^{sB^*}\, B^* z, \, Q\, e^{sB^*} z \rangle \,ds\\&=\frac{1}{2}\,\int _0^\infty \frac{d}{ds}\langle e^{sB^*}\, z,\, Q\, e^{sB^*} z \rangle \, ds\\&= -\frac{1}{2}\, |Q^{1/2}\, z|^2. \end{aligned}$$

Setting \(z=Q_\infty ^{-1} x \), we get (4.1).

Further, (4.2) easily follows if we observe that

$$\begin{aligned} \frac{\partial }{\partial s} D_s\, x = \frac{\partial }{\partial s} \left( Q_\infty e^{-sB^*} Q_\infty ^{-1} x \right) = - Q_\infty B^*\, Q_\infty ^{-1} Q_\infty e^{-sB^*} Q_\infty ^{-1} x =- Q_\infty B^*\,Q_\infty ^{-1} D_s\, x. \end{aligned}$$

Finally, we get by means of (4.2) and (4.1)

$$\begin{aligned} \frac{\partial }{\partial s} R\left( D_s\, x \right)&= \frac{1}{2} \frac{\partial }{\partial s} \langle Q_\infty ^{-1/2} D_s\, x, Q_\infty ^{-1/2} D_s\, x \rangle \\&=- \langle Q_\infty ^{-1/2} Q_\infty B^* Q_\infty ^{-1} D_s\, x, Q_\infty ^{-1/2} D_s\, x \rangle \\&= \frac{1}{2} \big | Q^{1/2} Q_\infty ^{-1}D_s\, x\big |^2\,, \end{aligned}$$

and (4.3) is verified. \(\square \)

We observe here that an integration of (4.2) leads to

$$\begin{aligned} |x- D_t \,x| \lesssim t\, |x|, \qquad 0\le t\le 1. \end{aligned}$$
(4.4)

Fix now \(\beta >0\) and consider the ellipsoid

$$\begin{aligned} E_\beta =\{x\in \mathbb {R}^n:\, R(x)= \beta \} \,.\end{aligned}$$

As a consequence of (4.3), the map \(s\mapsto R(D_s z)\) is strictly increasing for each \(0 \ne z\in \mathbb {R}^n\). Hence any \(x\in \mathbb {R}^n,\, x\ne 0\), can be written uniquely as

$$\begin{aligned} x=D_s\, {\tilde{x}} \,, \end{aligned}$$
(4.5)

for some \({\tilde{x}}\in E_\beta \) and \(s\in \mathbb {R}\). We consider s and \({\tilde{x}}\) as the polar coordinates of x. Our estimates in what follows will be uniform in \(\beta \).

Next, we shall write Lebesgue measure in terms of these polar coordinates. A normal vector to the surface \(E_\beta \) at the point \({\tilde{x}} \in E_\beta \) is \(\mathbf{{N}} ({\tilde{x}})=Q_\infty ^{-1}{\tilde{x}} \), and the tangent hyperplane at \({\tilde{x}}\) is \(\mathbf{{N}}({\tilde{x}})^\perp \). For \(s > 0\) the tangent hyperplane of the surface \(D_sE_\beta = \{D_s\, {\tilde{x}}: {\tilde{x}}\in E_\beta \}\) at the point \(D_s\,{\tilde{x}}\) is \(D_s(\mathbf{{N}}({\tilde{x}})^\perp )\), and a normal to \(D_sE_\beta \) at the same point is \(w = (D_s^{-1})^* (\mathbf{{N}}({\tilde{x}}))=D_{-s}^* Q_\infty ^{-1}{\tilde{x}} = Q_\infty ^{-1}e^{sB} {\tilde{x}}\).

The scalar product of w and the tangent of the curve \(s \mapsto D_s\,{\tilde{x}}\) at the point \(D_s\,{\tilde{x}}\) is, because of (4.2) and (4.1),

$$\begin{aligned}&\left\langle \frac{\partial }{ \partial s} D_s\, {\tilde{x}}, w \right\rangle \nonumber \\&\quad = -\langle Q_\infty e^{-sB^*}B^* Q_\infty ^{-1} {\tilde{x}},\, Q_\infty ^{-1}e^{sB} {\tilde{x}}\rangle = - \langle B^* Q_\infty ^{-1} {\tilde{x}}, {\tilde{x}}\rangle = \frac{1}{2}\, |Q^{1/2}\, Q_\infty ^{-1} {\tilde{x}}|^2 > 0. \end{aligned}$$
(4.6)

Thus the curve \(s \mapsto D_s\,{\tilde{x}}\) is transversal to each surface \(D_sE_\beta \). Let \(dS_s\) denote the area measure of \(D_sE_\beta \). Then Lebesgue measure is given in terms of our polar coordinates by

$$\begin{aligned} dx= H&(s, {\tilde{x}}) \, dS_s(D_s\, {\tilde{x}})\,ds, \end{aligned}$$
(4.7)

where

$$\begin{aligned} H(s,{\tilde{x}})= \left\langle \frac{\partial }{ \partial s} D_s\, {\tilde{x}}, \frac{w}{|w|}\right\rangle = \frac{ |Q^{1/2}\, Q_\infty ^{-1} {\tilde{x}} |^2 }{2\ | Q_\infty ^{-1} e^{sB} {\tilde{x}} |}. \end{aligned}$$

To see how \(dS_s\) varies with s, we take a continuous function \(\varphi =\varphi ({\tilde{x}}) \) on \(E_\beta \) and extend it to \({\mathbb {R}}^n \setminus \{0 \}\) by writing \(\varphi (D_s\, {\tilde{x}}) =\varphi ({\tilde{x}}) \). For any \(t>0\) and small \(\varepsilon >0\), we define the shell

$$\begin{aligned} \Omega _{t,\varepsilon } = \{D_s\, {\tilde{x}}: t<s<t+ \varepsilon , \;{\tilde{x}} \in E_\beta \}. \end{aligned}$$

Then \( \Omega _{t,\varepsilon }\) is the image under \(D_t\) of \( \Omega _{0,\varepsilon }\), and the Jacobian of this map is \(\det D_t = e^{-t{{\,\mathrm{tr}\,}}B}\). Thus

$$\begin{aligned} \int _{\Omega _{t,\varepsilon }} \varphi (x)\,dx = e^{-t{{\,\mathrm{tr}\,}}B} \int _{\Omega _{0,\varepsilon }} \varphi (D_t\, x)\,dx, \end{aligned}$$

which we can rewrite as

$$\begin{aligned}&\int _{t<s<t+ \varepsilon } \int _{{\tilde{x}} \in E_\beta } \varphi ({\tilde{x}})\, H(s,{\tilde{x}})\, dS_s(D_s\, {\tilde{x}})\,ds \\&\quad =\, e^{-t{{\,\mathrm{tr}\,}}B}\, \int _{0<s< \varepsilon } \int _{{\tilde{x}} \in E_\beta } \varphi ({\tilde{x}})\, H(s,{\tilde{x}})\, dS_s(D_s\, {\tilde{x}})\,ds. \end{aligned}$$

Now we divide by \(\varepsilon \) and let \(\varepsilon \rightarrow 0\), getting

$$\begin{aligned} \int _{ E_\beta } \varphi ({\tilde{x}})\, H(t,{\tilde{x}})\, dS_t(D_t\, {\tilde{x}}) = e^{-t{{\,\mathrm{tr}\,}}B}\, \int _{ E_\beta } \varphi ({\tilde{x}})\, H(0,{\tilde{x}})\, dS_0( {\tilde{x}}). \end{aligned}$$

Since this holds for any \(\varphi \), it follows that

$$\begin{aligned} dS_t(D_t\, {\tilde{x}}) = e^{-t{{\,\mathrm{tr}\,}}B}\, \frac{H(0,{\tilde{x}})}{H(t,{\tilde{x}})}\, dS_0( {\tilde{x}}). \end{aligned}$$

Together with (4.7), this implies the following result.

Proposition 4.2

The Lebesgue measure in \({\mathbb {R}}^n\) is given in terms of polar coordinates \((t, {\tilde{x}})\) by

$$\begin{aligned} dx = e^{-t{{\,\mathrm{tr}\,}}B}\, \frac{ |Q^{1/2}\, Q_\infty ^{-1} {\tilde{x}} |^2}{2\,| Q_\infty ^{-1} {\tilde{x}} |}\, dS_0( {\tilde{x}})\,dt\,. \end{aligned}$$

We also need estimates of the distance between two points in terms of the polar coordinates. The following result is a generalization of Lemma 4.2 in [4], and its proof is analogous.

Lemma 4.3

Fix \(\beta > 0\). Let \(x^{(0)},\; x^{(1)}\in \mathbb {R}^n \setminus \{ 0\} \) and assume \( R(x^{(0)}) > \beta /2\). Write

$$\begin{aligned} x^{(0)} = D_{ s^{(0)}} ({\tilde{x}}^{(0)} )\qquad \text { and } \qquad x^{(1)} = D_{s^{(1)}}({\tilde{x}}^{(1)}) \end{aligned}$$

with \(s^{(0)}\), \(s^{(1)}\in \mathbb {R}\) and \({\tilde{x}}^{(0)},\; {\tilde{x}}^{(1)} \in E_\beta \).

  1. (i)

    Then

    $$\begin{aligned} \big |x^{(0)} - x^{(1)}\big | \gtrsim c\, \big |{\tilde{x}}^{(0)} - {\tilde{x}}^{(1)}\big | .\end{aligned}$$
    (4.8)
  2. (ii)

    If also \(s^{(1)} \ge 0\), then

    $$\begin{aligned} \big |x^{(0)} - x^{(1)}\big | \gtrsim c\,\sqrt{\beta }\,|s^{(0)} -s^{(1)}|. \end{aligned}$$
    (4.9)

Proof

Let \(\Gamma : [0,1] \rightarrow \mathbb {R}^n \setminus \{0\}\) be a differentiable curve with \(\Gamma (0) =x^{(0)}\) and \(\Gamma (1) =x^{(1)}\). It suffices to bound the length of any such curve from below by the right-hand sides of (4.8) and (4.9).

For each \(\tau \in [0,1]\), we write

$$\begin{aligned} \Gamma (\tau ) = D_{s{(\tau )}}\,{\tilde{x}}{(\tau )} , \end{aligned}$$

with \({\tilde{x}}(\tau ) \in E_\beta \) and \({\tilde{x}}{(i)}= \tilde{x}^{(i)}\), \(s{(i)}= s^{(i)}\) for \(i=0,1\). Thus

$$\begin{aligned} \Gamma '(\tau )&= -s'(\tau ) \,\frac{\partial }{\partial s}\, {D_{s}}_{\big |s=s{(\tau )}} \,{\tilde{x}}{(\tau )} + D_{s{(\tau )}}\, {\tilde{x}}' (\tau ). \end{aligned}$$

The group property of \(D_s\) implies that

$$\begin{aligned} \frac{\partial }{\partial s} {D_{s}}_{\big |s=s{(\tau )}}= {D_{s(\tau )}} \frac{\partial }{\partial s} {D_{s}}_{\big |s=0}, \end{aligned}$$

and so

$$\begin{aligned} \Gamma '(\tau )&= {D_{s(\tau )}} \, v, \end{aligned}$$

with

$$\begin{aligned} v&= -s'(\tau ) \,\frac{\partial }{\partial s}\, {D_{s}}_{\big |s=0} \,{\tilde{x}}{(\tau )} + {\tilde{x}}' (\tau ). \end{aligned}$$

The vector \({\tilde{x}}' (\tau )\) is tangent to \(E_\beta \) and thus orthogonal to \(\mathbf{{N} }({\tilde{x}})\). Then (4.6) (with \(s=0\)) implies that the angle between \(\frac{\partial }{\partial s} {D_{s}}_{\big |s=0} \,{\tilde{x}} (\tau )\) and \({\tilde{x}}'(\tau )\) is larger than some positive constant. It follows that

$$\begin{aligned} |v|^2 \gtrsim |s'(\tau )|^2\, \Big | \frac{\partial }{\partial s}\, {D_{s}}_{\big |s=0} \,{\tilde{x}}{(\tau )} \Big |^2 + \big | {\tilde{x}}' (\tau )\big |^2 \gtrsim |s'(\tau )|^2\, \beta + \big | {\tilde{x}}' (\tau )\big |^2, \end{aligned}$$
(4.10)

where we also used the fact that, by (4.2),

$$\begin{aligned} \Big | \frac{\partial }{\partial s}\, {D_{s}}_{\big |s=0} \,\tilde{x}{(\tau )} \Big |\simeq |{\tilde{x}}(\tau )| \simeq \sqrt{\beta }. \end{aligned}$$

Since

$$\begin{aligned} |v|&= \big |{D_{-s(\tau )}}\, \Gamma '(\tau )\big | \le \big \Vert {D_{-s(\tau )}} \big \Vert \, \big | \Gamma '(\tau )\big | \lesssim e^{-C\min (s(\tau ),0)} \big | \Gamma '(\tau )\big | \end{aligned}$$

because of Lemma 3.1, we obtain from (4.10)

$$\begin{aligned} \big | \Gamma '(\tau )\big |&\gtrsim e^{C \min (s(\tau ), 0)} \ \big (\sqrt{\beta }\, |s'(\tau )| + \big | {\tilde{x}}' (\tau )\big |\big ). \end{aligned}$$
(4.11)

Next, we derive a lower bound for s(0); assume first that \(s{(0)} < 0\). The assumption \( R(x^{(0)}) > \beta /2\) implies, together with Lemma 3.1,

$$\begin{aligned} \beta /2 \le R(D_{s(0)}\, {\tilde{x}}^{(0)})&\lesssim \big | D_{ s{(0)}}\,\tilde{x}^{(0)} \big |^2\lesssim e^{c\, s(0)} \big | {\tilde{x}}^{(0)} \big |^2 \simeq e^{c\, s(0)}\beta .\end{aligned}$$

It follows that

$$\begin{aligned} s{(0)} > -{\tilde{s}}, \end{aligned}$$

for some \({\tilde{s}}\) with \(0<{\tilde{s}}<C\), and this obviously holds also without the assumption \(s(0)<0\).

Assume now that \( s(\tau ) > -{\tilde{s}}-1\) for all \(\tau \in [0,1]\). Then (4.11) implies

$$\begin{aligned} \big | \Gamma '(\tau )\big | \gtrsim \sqrt{\beta }\, |s'(\tau )| \end{aligned}$$

and

$$\begin{aligned} \big |\Gamma '(\tau )\big | \gtrsim |{\tilde{x}}'(\tau )|. \end{aligned}$$

Integrating these estimates with respect to \(\tau \) in [0, 1], we immediately see that one can control the length of \(\Gamma \) from below by the right-hand sides of (4.8) and (4.9).

If instead \( s(\tau ) \le -{\tilde{s}}-1\) for some \(\tau \in [0,1]\), we can proceed as in the proof of Lemma 4.2 in [4]. More precisely, since the image s([0, 1]) contains the interval \([-{\tilde{s}}-1, \max (s(0), s(1))]\), we can find a closed subinterval I of [0, 1] whose image s(I) is exactly the interval \([-{\tilde{s}}-1, \max (s(0), s(1))]\). Thus we may use (4.11) to control the length of \(\Gamma \) by

$$\begin{aligned} \int _0^1 \big | \Gamma '(\tau )\big |\, d\tau \ge \int _I \big |\Gamma '(\tau )\big |\, d\tau \gtrsim \sqrt{\beta }\, \int _I |s'(\tau )|\,d\tau \ge \sqrt{\beta }\, \big (\max \, (s(0), s(1))\; + \;{\tilde{s}}+1\big ). \end{aligned}$$

Here

$$\begin{aligned} \sqrt{\beta }\, \big (\max \, (s(0), s(1))\; + \;{\tilde{s}}+1\big )&\gtrsim \sqrt{\beta }\gtrsim \text {diam}\, E_\beta \ge \big |{\tilde{x}}^{(0)} - {\tilde{x}}^{(1)}\big |, \end{aligned}$$

and (4.8) follows. Under the additional hypothesis \(s(1)\ge 0\) of (ii), we have

$$\begin{aligned}{\tilde{s}}\ge \max \, (-s(0), -s(1))=-\min \,(s(0), s(1)).\end{aligned}$$

Then

$$\begin{aligned} \sqrt{\beta }\, \big (\max \, (s(0), s(1))\; + \;{\tilde{s}}+1\big )&\gtrsim \sqrt{\beta }\, \big (\max \, (s(0), s(1))-\min \,(s(0), s(1))\big )\\&= \sqrt{\beta }\, |s(0)-s(1)|, \end{aligned}$$

and (4.9) follows. \(\square \)

4.2 The Gaussian measure of a tube

We fix a large \(\beta > 0\). Define for \(x^{(1)}\in E_\beta \) and \(a>0\) the set

$$\begin{aligned} \Omega = \left\{ x \in E_\beta : \left| x - x^{(1)}\right| < a \right\} . \end{aligned}$$

This is a spherical cap of the ellipsoid \(E_\beta \), centered at \(x^{(1)}\). Observe that \(|x| \simeq \sqrt{\beta }\) for \(x \in \Omega \), and that the area of \(\Omega \) is \(|\Omega |\simeq \min \,( a^{n-1},\beta ^{(n-1)/2}) \). Then consider the tube

$$\begin{aligned} Z = \{D_s\, {\tilde{x}} :s\ge 0, \;{\tilde{x}} \in \Omega \}. \end{aligned}$$
(4.12)

Lemma 4.4

There exists a constant C such that \(\beta > C\) implies that the Gaussian measure of the tube Z fulfills

$$\begin{aligned} \gamma _\infty (Z)\lesssim \frac{a^{n-1}}{\sqrt{ \beta }}\, e^{-\beta }. \end{aligned}$$

Proof

Proposition 4.2 yields, since \(H(0,{\tilde{x}}) \simeq |{\tilde{x}}|\simeq \sqrt{\beta }\),

$$\begin{aligned} \gamma _\infty (Z) \simeq \int _0^\infty e^{-s {{\,\mathrm{tr}\,}}B} \ e^{-R(D_s\, {\tilde{x}})}\, \int _{\Omega } {H(0,{\tilde{x}})} \, dS ({\tilde{x}})\,ds \lesssim \sqrt{\beta }\,a^{n-1} \int _0^\infty e^{-s {{\,\mathrm{tr}\,}}B} e^{-R(D_s\, {\tilde{x}})}\,ds. \end{aligned}$$

By (4.3) we have

$$\begin{aligned} R ({D_s\, {\tilde{x}}})- R({\tilde{x}}) \simeq \int _0^s \big | D_{s'} \,{\tilde{x}}\big |^2 ds' \gtrsim s | {\tilde{x}}|^2\simeq s\beta , \end{aligned}$$

which implies

$$\begin{aligned} \gamma _\infty (Z) \lesssim \sqrt{\beta } \ a^{n-1}\, e^{-\beta } \int _0^\infty e^{-s {{\,\mathrm{tr}\,}}B} \, e^{-cs\beta }\, \ ds .\end{aligned}$$

Assuming \(\beta \) large enough, one has \(c\beta > -2{{\,\mathrm{tr}\,}}B\), and then the last integral is finite and no larger than \(C/\beta \). The lemma follows. \(\square \)

5 Simplifications

In this section, we introduce some preliminary simplifications and reductions for the proof of (1.3), i.e., of Theorem 1.1.

  1. (1)

    We may assume that f is nonnegative and normalized in the sense that

    $$\begin{aligned} \Vert f\Vert _{L^1( \gamma _\infty )}=1, \end{aligned}$$

    since this involves no loss of generality.

  2. (2)

    We may assume that \(\alpha \) is large, \(\alpha > C\), since otherwise (1.3) and (1.4) are trivial.

  3. (3)

    In many cases, we may restrict x in (1.3) and (1.4) to the ellipsoidal annulus

    $$\begin{aligned} {\mathcal E_\alpha }=\left\{ x \in \mathbb {R}^n:\, \frac{1}{2} \log \alpha \le R(x) \le 2 \log \alpha \, \right\} . \end{aligned}$$

    To begin with, we can always forget the unbounded component of the complement of \(\mathcal E_\alpha \), since

    $$\begin{aligned}&\gamma _\infty \{ x\in \mathbb {R}^n : \,R( x)> 2 \log \alpha \}\nonumber \\ {}&\quad \lesssim \int _{ R(x)> 2 \log \alpha } \exp (-R(x) )\, dx\, \lesssim (\log \alpha )^{(n-2)/2}\,\exp ( {- 2 \log \alpha }) \lesssim \frac{1}{\alpha }. \end{aligned}$$
    (5.1)
  4. (4)

    When \(t>1\), we may forget also the inner region where \(R(x)<\frac{1}{2} \log \alpha \). Indeed, from (3.5) we get, if \((x,u)\in \mathbb {R}^n\times \mathbb {R}^n\) with \(R(x) < \frac{1}{2} \log \alpha \),

    $$\begin{aligned} K_t (x,u) \lesssim \, e^{R(x)}< \sqrt{\alpha } < \alpha , \end{aligned}$$

    since \(\alpha \) is large. In other words, for any \((x,u)\in \mathbb {R}^n\times \mathbb {R}^n\)

    $$\begin{aligned} R(x) < \frac{1}{2} \log \alpha \; \qquad \Rightarrow \;\qquad K_t (x,u) \lesssim \alpha , \end{aligned}$$
    (5.2)

    for all \(t> 1\).

    Replacing \(\alpha \) by \(C\alpha \) for some C, we see from (3) and (4) that we can assume \(x \in {\mathcal E_\alpha }\) in the proof of (1.3) and (1.4), when the supremum in the maximal operator is taken only over \(t> 1\).

Before introducing the last simplification, we need to define a global region

$$\begin{aligned} G&=\left\{ (x,u)\in \mathbb {R}^n\times \mathbb {R}^n\,: \,|x-u|> \frac{1}{1+|x|} \right\} \end{aligned}$$

and a local region

$$\begin{aligned} L&=\left\{ (x,u)\in \mathbb {R}^n\times \mathbb {R}^n\,: \,|x-u|\le \frac{1}{1+|x|} \right\} . \end{aligned}$$

Notice that the definition of G and L does not depend on Q and B.

  1. (5)

    When \(t\le 1\) and \((x,u) \in G\), we shall see that (5.2) is still valid, and it is again enough to consider \(x\in {\mathcal E_\alpha }\).

To prove this, we need a lemma which will also be useful later.

Lemma 5.1

If \((x,u)\in G\) and \(0<t\le 1\), then

$$\begin{aligned} \frac{1}{(1+|x|)^2} \lesssim t^2 |x|^2 + | u-D_{t}\, x|^2. \end{aligned}$$

Proof

From the definition of G and (4.4) we get

$$\begin{aligned} \frac{1}{1+|x|}&\le |x- u| \le |x-D_t\, x| +|D_t\, x -u| \lesssim t | x |+|u-D_{t}\,\, x|. \end{aligned}$$

The lemma follows. \(\square \)

To verify now (5.2) in the global region with \(t\le 1\), we recall from (3.4) that

$$\begin{aligned} K_t (x,u) \lesssim \frac{e^{R(x)}}{ t^{n/2}}\, \exp \left( -c\,\frac{| u-D_t \,x|^2}{t} \right) . \end{aligned}$$

It follows from Lemma 5.1 that

$$\begin{aligned} t^2 \gtrsim \frac{1}{(1+|x|)^4}&\qquad \text {or} \qquad \frac{|u-D_{t}\, x|^2}{t} \gtrsim \frac{1}{(1+|x|)^2t}. \end{aligned}$$
(5.3)

The first inequality here implies that

$$\begin{aligned} K_t (x,u) \lesssim e^{R(x)}\, (1+| x|)^n \lesssim e^{2R(x)}, \end{aligned}$$

and (5.2) follows. If the second inequality of (5.3) holds, we have

$$\begin{aligned} K_t (x,u) \lesssim \frac{e^{R(x)}}{t^{n/2}} \exp \left( -\frac{c}{(1+|x|)^2 t} \right) \lesssim e^{R(x)}\, (1+|x|)^n, \end{aligned}$$

and we get the same estimate. Thus (5.2) is verified.

Finally, let

$$\begin{aligned} \mathcal H_*^{G} f(x)=\sup _{0 < t\le 1} \left| \int K_t (x,u)\,\chi _{{G}}(x,u)\,f(u) \, d\gamma _\infty (u) \, \right| \,, \end{aligned}$$

and

$$\begin{aligned} \mathcal H_*^{L} f(x)=\sup _{0 < t\le 1} \left| \int K_t (x,u)\,\chi _{{L}}(x,u)\,f(u) \, d\gamma _\infty (u) \, \right| \,. \end{aligned}$$

6 The case of large t

In this section, we consider the supremum in the definition of the maximal operator taken only over \(t> 1\), and we prove (1.4).

Proposition 6.1

For all functions \(f\in L^1 (\gamma _\infty )\) such that \(\Vert f\Vert _{L^1( \gamma _\infty )}=1\),

$$\begin{aligned} \gamma _\infty \left\{ x : \sup _{t> 1}|\mathcal H_t f(x)|> \alpha \right\} \lesssim \frac{1}{\alpha \sqrt{\log \alpha }}, \quad \quad \alpha >2. \end{aligned}$$
(6.1)

In particular, the maximal operator

$$\begin{aligned} \sup _{t> 1}|\mathcal H_t f(x)| \end{aligned}$$

is of weak type (1, 1) with respect to the invariant measure \(\gamma _\infty \).

Proof

We can assume that \(f\ge 0\). Looking at the arguments in Sect. 5, items (3) and (4), we see that it suffices to consider points \(x\in {\mathcal E_\alpha }\). For both x and u we use the coordinates introduced in (4.5) with \(\beta =\log \alpha \), that is,

$$\begin{aligned} x=D_{s}\,{\tilde{x}}, \qquad u=D_{ s'}\, {\tilde{u}}, \end{aligned}$$

where \({\tilde{x}}, {\tilde{u}} \in E_{\log \alpha }\) and \(s,s' \in \mathbb {R}\).

From (3.5) we have

$$\begin{aligned} K_t (x,u)&\lesssim \exp (R( x))\exp \big ( - c\, \big | D_{-t}\,u-x\big |^2 \big ) \end{aligned}$$

for \(t> 1\) and \(x,u\in \mathbb {R}^n\). Since \(x \in {\mathcal E_\alpha }\) and \(D_{-t}\,u= D_{s'-t}\, \tilde{u}, \) we can apply Lemma 4.3 (i), getting

$$\begin{aligned} \big | D_{-t}\,u-x \big | \gtrsim \big | {\tilde{x}}-{\tilde{u}}\big |, \end{aligned}$$

so that

$$\begin{aligned} \int K_t (x,u) f(u)\,d \gamma _\infty (u)&\lesssim \exp \big ( R(D_s\, {\tilde{x}}) \big ) \int \exp \big ( - c\, \big | {\tilde{x}}-{\tilde{u}}\big |^2 \big )\,f(u)\,d \gamma _\infty (u). \end{aligned}$$

In view of (4.3), the right-hand side here is strictly increasing in s, and therefore the inequality

$$\begin{aligned} \exp \big ( R(D_s\, {\tilde{x}}) \big ) \int \exp \big ( - c\, \big | {\tilde{x}}-{\tilde{u}}\big |^2 \big )\,f(u)\,d \gamma _\infty (u) > \alpha \end{aligned}$$
(6.2)

holds if and only if \(s> s_\alpha ({\tilde{x}})\) for some function \({\tilde{x}}\mapsto s_\alpha ({\tilde{x}})\), with equality for \(s=s_\alpha ({\tilde{x}})\). Since \(\alpha >2\) and \(\Vert f\Vert _{L^1 (\gamma _\infty )}=1\), it follows that \(s_\alpha ({\tilde{x}})>0\).

For some C, the set of points \(x\in {\mathcal E_\alpha }\) where the supremum in (6.1) is larger than \(C\alpha \) is contained in the set \(\mathcal A(\alpha )\) of points \(D_s\, {\tilde{x}}\in {\mathcal E_\alpha }\) fulfilling (6.2). We use Proposition 4.2 to estimate the \(\gamma _\infty \) measure of \(\mathcal A(\alpha )\). Observe that \(H(0,{\tilde{x}}) \simeq |{\tilde{x}}|\simeq \sqrt{\log \alpha }\) and that \(D_s\, {\tilde{x}}\in {\mathcal E_\alpha }\) implies \(s\lesssim 1\), so that also \(e^{-s{{\,\mathrm{tr}\,}}B} \lesssim 1\). We get

$$\begin{aligned} \gamma _\infty (\mathcal A(\alpha ) )&= \int _{\mathcal A(\alpha )\cap {\mathcal E_\alpha }} e^{-R(x)} dx \\&\lesssim { \sqrt{\log \alpha }} \int _{E_{\log \alpha }} \int _{s_\alpha ({\tilde{x}})}^{C} e^{-R(D_s\, {\tilde{x}})} \,ds\, dS({\tilde{x}}) \\&\lesssim { \sqrt{\log \alpha }} \int _{E_{\log \alpha }} \int _{s_\alpha ({\tilde{x}})}^{+\infty } \exp \left( - {R( D_{s_\alpha ({\tilde{x}})} \,{\tilde{x}}) -c\log \alpha \, (s-s_\alpha ({\tilde{x}})}) \right) \,ds\, dS({\tilde{x}}) , \end{aligned}$$

where the last inequality follows from (4.3), since \( |D_s\, {\tilde{x}}|^2\gtrsim |{\tilde{x}}|^2\simeq \log \alpha . \) Integrating in s, we obtain

$$\begin{aligned} \gamma _\infty (\mathcal A (\alpha ) )&\lesssim \frac{1}{\sqrt{\log \alpha }}\int _{E_{\log \alpha }} \exp \big (-R( D_{{s_\alpha ({\tilde{x}})}} \, {\tilde{x}}) \big ) \,dS({\tilde{x}}) . \end{aligned}$$

Now combine this estimate with the case of equality in (6.2) and change the order of integration, to get

$$\begin{aligned} \gamma _\infty (\mathcal A (\alpha ) )&\lesssim \frac{1}{\alpha \sqrt{\log \alpha }} \int \int _{E_{\log \alpha }} \exp \big ( - c\, \big | {\tilde{x}}-{\tilde{u}}\big |^2 \big ) \,dS(\tilde{x})\,f(u)\,d \gamma _\infty (u)\\&\lesssim \frac{1}{\alpha \sqrt{\log \alpha }} \int f(u)\,d \gamma _\infty (u)\,, \end{aligned}$$

which proves Proposition 6.1. \(\square \)

Finally, we show that the factor \(1/\sqrt{\log \alpha }\) in (6.1) is sharp.

Proposition 6.2

For any \(t> 1\) and any large \(\alpha \), there exists a function f normalized in \(L^1 (\gamma _\infty )\) and such that

$$\begin{aligned} \gamma _\infty \left\{ x :|\mathcal H_t f(x)| > \alpha \right\} \simeq \frac{1}{\alpha \sqrt{\log \alpha }}\,. \end{aligned}$$

Proof

Take a point z with \(R(z)=\log \,\alpha \), and let f be (an approximation of) a Dirac measure at the point \(u=D_t z\). Then, as a consequence of (3.5), \(K_t (x,u) \simeq \exp (R( x))\) when x is in the ball \(B(D_{-t}\,u,1)=B(z,1)\). We then have \(\mathcal H_t f(x)=K_t (x,u)\gtrsim \alpha \) in the set \(\mathcal B=\{x\in B(z,1)\!: R(x)>R(z) \}\), whose measure is

$$\begin{aligned} \gamma _\infty \, (\mathcal B) \simeq e^{-R(z)}\, \frac{1}{\sqrt{R(z)}} = \frac{1}{\alpha \,\sqrt{\log \,\alpha }}. \end{aligned}$$

\(\square \)

7 The local case for small t

Proposition 7.1

If \((x,u) \in L \) and \(0<t\le 1\), then

$$\begin{aligned} \big |K_{t} (x,u)\big |\lesssim \,\frac{ \exp \big ( R(x) \big ) }{ t^{n/2} }\, \exp { \left( -c\, \frac{| u-x|^2}{ t} \, \right) } \,. \end{aligned}$$

Proof

In view of (3.4), it is enough to show that

$$\begin{aligned} \frac{|u-D_t\, x|^2}{t} \ge \frac{|u-x|^2}{t} -C. \end{aligned}$$
(7.1)

We write

$$\begin{aligned}&|u-D_t\, x|^2 = |u-x +x-D_t \,x|^2 = |u-x|^2 +2 \langle u-x, x-D_t\, x\rangle + |x-D_t\, x|^2\\&\quad \ge |u-x|^2 -2 |u-x|\,|x-D_t\, x|. \end{aligned}$$

By (4.4),

$$\begin{aligned} |u-x|\, |x-D_t\, x| \lesssim |u-x|\,t\, |x|\le t \end{aligned}$$

since \((x,u) \in L\), and (7.1) follows. \(\square \)

Proposition 7.2

The maximal operator \(\mathcal H_*^{L} \) is of weak type (1, 1) with respect to the invariant measure \(\gamma _\infty \).

Proof

The proof is standard, since Proposition 7.1 implies

$$\begin{aligned} \mathcal H_*^{L} f(x)&\lesssim \sup _{0<t\le 1} \frac{\exp \big ( R(x)\big )}{ t^{n/2}} \int \exp \Big (-c\, \frac{| x-u|^2 }{ t} \, \Big ) \,\chi _{L}(x,u)\,f(u) \, d\gamma _\infty (u). \end{aligned}$$

The supremum here defines an operator of weak type (1, 1) with respect to Lebesgue measure in \(\mathbb {R}^n\). From this the proposition follows, cf. [7,   Section 3]. \(\square \)

8 The global case for small t

In this section, we conclude the proof of Theorem 1.1.

Proposition 8.1

The maximal operator \(\mathcal H_*^{G}\) is of weak type (1, 1) with respect to the invariant measure \(\gamma _\infty \).

Proof

We take f and \(\alpha \) as in items (1) and (2) of Sect. 5. Then item (5) tells us that we need only consider \(\mathcal H_*^{G} f(x)\) for \(x\in \mathcal E_\alpha \).

For \(m\in \mathbb {N}\) and \(0<t\le 1\), we introduce regions \(\mathcal S^{m}_t\). If \(m>0\), we let

$$\begin{aligned} \mathcal S^{m}_t&=\left\{ (x,u)\in G: 2^{m-1} \sqrt{t}< |u-D_t\, x |\le 2^{m} \sqrt{t} \,\right\} . \end{aligned}$$

If \(m=0\), we replace the condition \(2^{m-1} \sqrt{t}< |u-D_t\, x |\le 2^{m} \sqrt{t}\) by \(| u-D_t\, x | \le \sqrt{t}\). Note that for any fixed \(t\in (0,1]\) these sets form a partition of G.

In the set \({\mathcal S^{m}_t}\) we have, because of (3.4),

$$\begin{aligned} K_t (x,u)&\lesssim \frac{ \exp (R(x)) }{ t^{n/2} } \exp \left( { -c{2^{2 m} } }\right) . \end{aligned}$$

Then setting

$$\begin{aligned} {\mathcal K}_t^{{m}} (x,u)&= \frac{ \exp (R(x)) }{ t^{n/2} } \,\chi _{\mathcal S^{m}_t}(x,u), \end{aligned}$$

one has, for all \((x,u)\in G\) and \(0<t<1\),

$$\begin{aligned} K_t (x,u)\lesssim \sum _{m=0}^\infty \exp \left( { -c{2^{2 m} } }\right) {\mathcal K}_t^{{m}} (x,u) \,. \end{aligned}$$

Hence, it suffices to prove that for \(m = 0,1,\dots \)

$$\begin{aligned} \gamma _\infty \left\{ x\in {\mathcal E_\alpha }: \sup _{0<t\le 1} \int {\mathcal K}_t^{{m}} \! (x,u)\, f(u)\, d \gamma _\infty (u) \! >\alpha \right\} \lesssim \frac{2^{Cm}}{\alpha } , \end{aligned}$$
(8.1)

for large \(\alpha \) and some C, since this will allow summing in m in the space \(L^{1,\infty }(\gamma _\infty )\).

Fix \(m\in {\mathbb {N}}\) and assume that \((x,u) \in S_t^m\) for some \(t \in (0,1]\), so that \(|u-D_t\, x| \le 2^{m}\sqrt{t}\). Then Lemma 5.1 leads to

$$\begin{aligned} 1&\lesssim (1+|x|)^4 t^2+(1+|x|)^2 \, 2^{2m}\,t \le ((1+|x |)^2 \, 2^{2m}\,t)^2+ (1+|x|)^2 \,2^{2m}\,t. \end{aligned}$$

Consequently, a point \(x\in \mathcal E_\alpha \) satisfies

$$\begin{aligned} (1+|x|)^2 \, 2^{2m}\,t \gtrsim 1 \end{aligned}$$
(8.2)

as soon as there exists a point u with \(\mathcal K_t^{m}(x,u)\ne 0\), and then \(t\ge \varepsilon >0\) for some \(\varepsilon =\varepsilon (\alpha ,m) >0\). Hence the supremum in (8.1) will be the same if taken only over \(\varepsilon \le t\le 1\), and it follows that this supremum is a continuous function of \(x\in {\mathcal E_\alpha }\).

To prove (8.1), the idea, which goes back to [15], is to construct a finite sequence of pairwise disjoint balls \(\big (\mathcal B^{(\ell )}\big )_{\ell =1}^{\ell _0}\) in \(\mathbb {R}^n\) and a finite sequence of sets \(\big (\mathcal Z^{(\ell )}\big )_{\ell =1}^{\ell _0}\) in \(\mathbb {R}^n\), called forbidden zones. These zones will together cover the level set in (8.1). We will then verify that

$$\begin{aligned} \left\{ x\in {\mathcal E_\alpha }: \sup _{\varepsilon \le t\le 1} \int {\mathcal K}_t^{{m}} (x,u)\,f(u)\,d \gamma _\infty (u)\, \ge \alpha \right\} \subset \bigcup _{\ell =1}^{\ell _0}\mathcal Z^{(\ell )}, \end{aligned}$$
(8.3)

that for each \(\ell \)

$$\begin{aligned}&\gamma _\infty (\mathcal Z^{(\ell )}) \lesssim \frac{2^{Cm}}{\alpha } \int _{\mathcal B^{(\ell )}}f(u) \,d \gamma _\infty (u), \end{aligned}$$
(8.4)

and that the \( \mathcal B^{(\ell )}\) are pairwise disjoint. This would imply

$$\begin{aligned} \gamma _\infty \Big (\bigcup _{\ell =1}^{\ell _0} \mathcal Z^{(\ell )} \Big ) \lesssim \frac{2^{Cm}}{\alpha }\, \sum _{\ell =1}^{\ell _0} \int _{\mathcal B^{(\ell )}}f(u)\,d \gamma _\infty (u) \lesssim \frac{2^{Cm}}{\alpha } ,\end{aligned}$$

and thus also (8.1) and Proposition 8.1.

The sets \(\mathcal B^{(\ell )}\) and \(\mathcal Z^{(\ell )}\) will be introduced by means of a sequence of points \(x^{(\ell )}, \ell =1,\ldots , \ell _0\), which we define by recursion. To start, we choose as \(x^{(1)}\) a point where the quadratic form R(x) takes its minimal value in the compact set

$$\begin{aligned} \mathcal A_1 (\alpha )= \left\{ x\in {\mathcal E_\alpha } : \sup _{\varepsilon \le t\le 1} \int {\mathcal K}_t^{{m}} (x,u)\,f(u)\,d \gamma _\infty \ge \alpha \right\} . \end{aligned}$$

However, should this set be empty, (8.1) is immediate.

We now describe the recursion to construct \(x^{(\ell )}\) for \(\ell \ge 2\). Like \(x^{(1)}\), these points will satisfy

$$\begin{aligned} \sup _{\varepsilon \le t\le 1} \int {\mathcal K}_t^{{m}} (x^{(\ell )},u)\,f(u)\,d \gamma _\infty \ge \alpha . \end{aligned}$$

Once an \(x^{(\ell )},\;\ell \ge 1 \), is defined, we can thus by continuity choose \(t_\ell \in [\varepsilon , 1] \) such that

$$\begin{aligned} \int {\mathcal K}_{t_\ell }^{{m}} (x^{(\ell )},u)\, f(u)\,d \gamma _\infty \ge \alpha . \end{aligned}$$
(8.5)

Using this \(t_\ell \), we associate with \(x^{(\ell )}\) the tube

$$\begin{aligned} \mathcal Z^{(\ell )} = \left\{ D_{ s}\,\eta \in \mathbb {R}^n :\,s\ge 0,\; R{( \eta )}= R(x^{(\ell )}), \; | \eta - x^{(\ell )} |< A\, 2^{3m}\, \sqrt{t_{\ell }}\right\} , \end{aligned}$$

Here the constant \(A>0\) is to be determined, depending only on n, Q and B.

All the \(x^{(\ell )}\) will be minimizing points of R(x). To avoid having them too close to one another, we will not allow \(x^{(\ell )}\) to be in any \(\mathcal Z^{(\ell ')}\) with \(\ell ' < \ell \). More precisely, assuming \(x^{(1)}, \dots , x^{(\ell )}\) already defined, we will choose \(x^{(\ell +1)}\) as a minimizing point of R(x) in the set

$$\begin{aligned} \mathcal A_{\ell +1} (\alpha )= \left\{ x\in {\mathcal E_\alpha } \setminus \bigcup _{\ell '=1}^{\ell } \mathcal Z^{(\ell ')}:\, \sup _{\varepsilon \le t\le 1} \int {\mathcal K}_t^{{m}} (x,u)\,f(u)\,d\gamma _\infty (u) \ge \alpha \right\} , \end{aligned}$$
(8.6)

provided this set is nonempty. But if \(\mathcal A_{\ell +1} (\alpha )\) is empty, the process stops with \(\ell _0=\ell \) and (8.3) follows. We will see that this actually occurs for some finite \(\ell \).

Now assume that \(\mathcal A_{\ell +1} (\alpha ) \ne \emptyset \). In order to assure that a minimizing point exists, we must verify that \(\mathcal A_{\ell +1} (\alpha )\) is closed and thus compact, although the \(\mathcal Z^{(\ell ')}\) are not open. To do so, observe that for \(1\le \ell ' \le \ell \), the minimizing property of \(x^{(\ell ')}\) means that there is no point x in \(\mathcal A_{\ell '} (\alpha ) \) with \(R(x) < R(x^{(\ell ')})\). Thus we have the inclusions

$$\begin{aligned} \mathcal A_{\ell +1} (\alpha ) \subset \mathcal A_{\ell '} (\alpha ) \subset \left\{ x: R(x)\ge R(x^{(\ell ')})\right\} , \qquad 1\le \ell ' \le \ell . \end{aligned}$$

It follows that

$$\begin{aligned}&\mathcal A_{\ell +1} (\alpha ) = \mathcal A_{\ell +1} (\alpha )\,\cap \, \bigcap _{1\le \ell ' \le \ell } \{x: R(x)\ge R(x^{(\ell ')})\} \\&\quad =\bigcap _{\ell '=1}^{\ell } \left\{ x\in {\mathcal E_\alpha } \setminus \mathcal Z^{(\ell ')}: R(x)\ge R(x^{(\ell ')}), \; \sup _{\varepsilon \le t\le 1} \int {\mathcal K}_t^{{m}} (x,u)\,f(u)\,d\gamma _\infty (u) \ge \alpha \right\} . \end{aligned}$$

For each \(\ell ' = 1,\dots , \ell \) we have

$$\begin{aligned}&\{x\in {\mathcal E_\alpha } \setminus \mathcal Z^{(\ell ')}: R(x)\ge R(x^{(\ell ')})\} \\&\quad = \left\{ D_s\,\eta \in \mathcal E_\alpha :\, s\ge 0, \,R(\eta ) = R(x^{(\ell ')}), \;|\eta - x^{(\ell ')}| \ge A2^{3m} \sqrt{t_{\ell '}} \right\} , \end{aligned}$$

and this set is closed. It follows that \( \mathcal A_{\ell +1} (\alpha )\) is compact, and a minimizing point \(x^{(\ell +1)}\) can be chosen. Thus the recursion is well defined.

We observe that (8.2) applies to \(t_\ell \) and \(x^{(\ell )}\), and \(|x^{(\ell )}|\) is large, so

$$\begin{aligned} |x^{(\ell )}|^2\, 2^{2m}\, t_\ell \gtrsim 1. \end{aligned}$$
(8.7)

Further, we define balls

$$\begin{aligned} \mathcal B^{(\ell )}=&\{ u \in \mathbb {R}^n :\, | u -D_{t_\ell } \,x^{(\ell )} | \le 2^{m} \sqrt{t_\ell }\, \}\,. \end{aligned}$$

Because of the definitions of \(\mathcal K_t^m\) and \({\mathcal S^{m}_t}\), the inequality (8.5) implies

$$\begin{aligned} \alpha&\le \frac{\exp \left( { R(x^{(\ell )}) }\right) }{ t_\ell ^{n/2}} \int _{\mathcal B^{(\ell )} }f(u)\,d \gamma _\infty (u). \end{aligned}$$
(8.8)

It remains to verify the claimed properties of \(\mathcal B^{(\ell )}\) and \(\mathcal Z^{(\ell )}\). The arguments below follow the lines of the proof of Lemma 6.2 in [4], with only slight modifications.

Lemma 8.2

The balls \(\mathcal B^{(\ell )}\) are pairwise disjoint.

Proof

Two balls \(\mathcal B^{(\ell )}\) and \(\mathcal B^{(\ell ')}\) with \(\ell <\ell '\) will be disjoint if

$$\begin{aligned} \big | D_{t_{\ell }} \, x^{(\ell )}- D_{t_{\ell '}} \, x^{(\ell ')}\big | > 2^m (\sqrt{t_\ell }+ \sqrt{ t_{\ell '}}). \end{aligned}$$
(8.9)

By means of our polar coordinates with \(\beta =R(x^{(\ell )})\), we write

$$\begin{aligned} x^{(\ell ')} = D_{s}\, {\tilde{x}}^{(\ell ')} \end{aligned}$$

for some \({\tilde{x}}^{(\ell ')}\) with \(R({\tilde{x}}^{(\ell ')})= R(x^{(\ell )})\) and some \(s \in \mathbb {R}\). Note that \(s\ge 0\), because \(R(x^{(\ell ')})\ge R( x^{(\ell )})\). Since \(x^{(\ell ')}\) does not belong to the forbidden zone \( \mathcal Z^{(\ell )}\), we must have

$$\begin{aligned} | {\tilde{x}}^{(\ell ')}- x^{(\ell )} |\ge A 2^{3m} \sqrt{t_\ell }. \end{aligned}$$
(8.10)

We first assume that \({t_{\ell '}} \ge M\, 2^{4m} \, t_\ell \), for some \(M=M(n, Q,B)\ge 2\) to be chosen. Lemma 4.3 (ii) implies

$$\begin{aligned} \big | D_{t_{\ell }} \, x^{(\ell )}- D_{t_{\ell '}} \, x^{(\ell ')}\big | = \big | D_{t_{\ell }} \, x^{(\ell )}- D_{t_{\ell '}+s} \, {\tilde{x}}^{(\ell ')}\big | \gtrsim |x^{(\ell )}|\,( t_{\ell '}+s-t_\ell ) \gtrsim |x^{(\ell )}|\, { t_{\ell '}}, \end{aligned}$$

the last step by our assumption. Using again the assumption and then (8.7), we get

$$\begin{aligned} |x^{(\ell )}|\,{ t_{\ell '}}\gtrsim |x^{(\ell )}|\, \sqrt{M} \, 2^{2m} \sqrt{ t_{\ell }}\, \sqrt{ t_{\ell '}} \gtrsim \sqrt{M} \, 2^{m}\sqrt{ t_{\ell '}} \simeq \sqrt{M}\, 2^{m}\, ( \sqrt{ t_{\ell '}} +\sqrt{ t_{\ell }}). \end{aligned}$$

Fixing M suitably large, we obtain (8.9) from the last two formulae.

It remains to consider the case when \({t_{\ell '}} < M\, 2^{4m}\, t_\ell \). Then

$$\begin{aligned} \sqrt{t_{\ell }} > \frac{2^{-2m-1}}{\sqrt{M}} ( \sqrt{t_{\ell '}}+\sqrt{t_\ell }). \end{aligned}$$

Applying this to (8.10), we obtain (8.9) by choosing A so that \(A/\sqrt{M}\) is large enough. \(\square \)

We next verify that the sequence \((x^{(\ell )})\) is finite. For \(\ell <\ell '\), we have (8.10), and Lemma 4.3 (i) implies

$$\begin{aligned} \big | x^{(\ell ')}- x^{(\ell )} \big |&\gtrsim A\, 2^{3m} \sqrt{t_\ell }. \end{aligned}$$

Since \(t_\ell \ge \varepsilon \), we see that the distance \(\left| x^{(\ell ')}- x^{(\ell )} \right| \) is bounded below by a positive constant. But all the \( x^{(\ell )}\) are contained in the bounded set \( {\mathcal E_\alpha }\), so they are finite in number. Thus the set considered in (8.6) must be empty for some \(\ell \), and the recursion stops. This implies (8.3).

We finally prove (8.4) . Observe that the forbidden zone \(\mathcal Z^{(\ell )}\) is a tube as defined in (4.12), with \(a=A\, 2^{3m} \sqrt{t_\ell }\) and \(\beta =R(x^{(\ell )})\). This value of \(\beta \) is large since \(x^{(\ell )} \in {\mathcal E_\alpha }\), and thus we can apply Lemma 4.4 to obtain

$$\begin{aligned} \gamma _\infty ( \mathcal Z^{(\ell )})\lesssim \frac{\big ( A 2^{3m} \sqrt{t_\ell }\big )^{n-1}}{ \sqrt{ R(x^{(\ell )})}}\, \exp \left( { -{ R(x^{(\ell )}) }}\right) . \end{aligned}$$

We bound the exponential here by means of (8.8) and observe that \(R(x^{(\ell )}) \sim |x^{(\ell )}|^2\), getting

$$\begin{aligned}\gamma _\infty (\mathcal Z^{(\ell )})\lesssim \frac{1}{\alpha |x^{(\ell )}| {\sqrt{t_\ell }}} \, (A 2^{3m})^{n-1} \, \int _{\mathcal B^{(\ell )} }\!f(u)\,d \gamma _\infty (u). \end{aligned}$$

As a consequence of (8.7), we obtain

$$\begin{aligned} \gamma _\infty (\mathcal Z^{(\ell )}) \lesssim \frac{2^{m}}{\alpha } \, \big (A 2^{3m}\big )^{n-1} \, \int _{\mathcal B^{(\ell )}}f(u)\,d \gamma _\infty (u)\, \lesssim \frac{2^{Cm}}{\alpha }\, \, \int _{\mathcal B^{(\ell )}}f(u)\,d \gamma _\infty (u), \end{aligned}$$

proving (8.4). This concludes the proof of Proposition 8.1. \(\square \)