Keywords

These keywords were added by machine and not by the authors. This process is experimental and the keywords may be updated as the learning algorithm improves.

1 Notations and Main Result

Let \(V = {\mathbb{R}}^{d}\) be the d-dimensional Euclidean space endowed with the scalar product \(< x,y >=\sum _{ i=1}^{d}x_{i}y_{i}\) and the corresponding norm \(\vert x\vert = {(\sum _{i=1}^{d}\vert x_{i}{\vert }^{2})}^{1/2}\). We denote by G = GL(V ) (resp. H = Aff(V )) the linear (resp. affine) group of V and we fix a probability measure μ (resp. λ) on G (resp. H) such that μ is the projection of λ. We consider the affine stochastic recursion (S) on V defined by

$$\displaystyle{ X_{n+1} = A_{n+1}X_{n} + B_{n+1},\quad X_{0} = x, }$$
(S)

where \((A_{n},B_{n})\) are i.i.d. random variables with law λ, hence A n (resp. B n ) are i.i.d. random matrices (resp. vectors). We assume that (S) has a stationary solution R which satisfies in distribution

$$\displaystyle{R = A{R}^{1} + B}$$

where R 1 has the same law as R and is independent of (A, B). We are interested in the “asymptotic shape” of the law ρ of R. Our focus will be on the case d > 1. For d = 1, corresponding results are described in Sect. 4.

We denote by η ∗ θ the convolution of a probability measure η on H with a positive Radon measure θ on V i.e. \(\eta {\ast}\theta =\int \delta _{hx}\,d\eta (h)\,d\theta (x)\). Also η n denotes the nth convolution iterate of η. With these notations, the law ρ n of X n is given by \(\rho _{n} {=\lambda }^{n} {\ast}\delta _{x}\), and a λ-stationary (probability) measure ρ satisfies λ ∗ ρ = ρ.

We denote by Ω the product space \({H}^{\otimes \mathbb{N}}\), by \(\mathbb{P}\) the product measure \({\lambda }^{\otimes \mathbb{N}}\) on Ω, and by \(\mathbb{E}\) the corresponding expectation operator. Provided that

$$\displaystyle{\mathbb{E}(\vert \log \vert A\vert \vert ) + \mathbb{E}(\vert \log \vert B\vert \vert ) < \infty,}$$

it is well known (see [12] for example) that a λ-stationary measure ρ exists and is unique if the top Lyapunov exponent

$$\displaystyle{L_{\mu } =\lim _{n\rightarrow \infty }\frac{1} {n}\,\mathbb{E}(\log \vert A_{n}\ldots A_{1}\vert )}$$

is negative. For informations on products of random matrices we refer to [2, 5, 9].

Since the properties of μ play a dominant role for the “shape” of ρ, we give now a few corresponding notations. Let S (resp. T) be the closed subsemigroup of G (resp. H) generated by the support suppμ (resp. suppλ) of μ (resp. λ) and write

$$\displaystyle{S_{n} = A_{n}\ldots A_{1},\quad k(s) =\lim _{n\rightarrow \infty }{(\mathbb{E}(\vert S_{n}{\vert }^{s}))}^{1/n}\;(s \geq 0).}$$

Then logk(s) is a convex function on \(I_{\mu } =\{ s \geq 0;k(s) < \infty \}\) and we write \(s_{\infty } =\sup \{ s \geq 0;k(s) < \infty \}\).

We denote by \({\mathbb{S}}^{d-1}\) (resp. \({\mathbb{P}}^{d-1}\)) the unit sphere (resp. projective space) of V and observe that in polar coordinates:

$$\displaystyle{V \setminus \{0\} = {\mathbb{S}}^{d-1} \times \mathbb{R}_{ +}^{{\ast}}.}$$

If \(\dot{V }\) denotes the factor space of \(V \setminus \{0\}\) by the group \(\{\pm Id\}\), we have also

$$\displaystyle{\dot{V } = {\mathbb{P}}^{d-1} \times \mathbb{R}_{ +}^{{\ast}}.}$$

For \(x \in V \setminus \{0\},\;g \in G\), we write \(\tilde{x}\) (resp. \(\bar{x}\)) for the projection of x ∈ V on \({\mathbb{S}}^{d-1}\) (resp. \({\mathbb{P}}^{d-1}\)), \(g \cdot \tilde{ x}\) (resp. \(g \cdot \bar{ x}\)) for the projection of gx on \({\mathbb{S}}^{d-1}\) (resp. \({\mathbb{P}}^{d-1}\)).

For some moment conditions on μ, the quantity \(\gamma (g) =\sup (\vert g\vert,\vert {g}^{-1}\vert )\) will be used. The dual map g  ∗  of g ∈ GL(V ) is defined by < g  ∗  x, y > = < x, gy > and the push-forward of μ by \(g \rightarrow {g}^{{\ast}}\) will be denoted μ  ∗ .

From now on, we assume d > 1. An element g ∈ G is said to be proximal if g has a unique simple dominant eigenvalue \(\lambda _{g} \in \mathbb{R}\) with \(\vert \lambda _{g}\vert =\lim _{n\rightarrow \infty }\vert {g}^{n}{\vert }^{1/n}\). In this case we have the decomposition \(V = \mathbb{R}w_{g} \oplus V _{g}^{<}\) where w g is a dominant eigenvector and V g  <  a g-invariant hyperplane. We say that a semigroup of G satisfies condition i − p if this semigroup contains a proximal element and does not leave any finite union of subspaces invariant.

One can observe that, if d > 1, the set of probability measures μ on G such that S satisfies condition i − p is open and dense in the weak topology. Also, condition i − p is satisfied for S if and only if it is satisfied for the closed subgroup Zc(S), the Zariski closure of S, which is a Lie group with a finite number of components. Thus condition i − p is in particular satisfied, if Zc(S) = G.

It is known that, if the probability measure μ satisfies \(\mathbb{E}(\vert \log \vert A\vert \vert ) < \infty \) and \(\mathrm{supp}\mu\) generates a closed semigroup S satisfying condition i − p, then the top Lyapunov exponent of μ is simple (see [2]). In this case logk(s) is strictly convex and analytic on [0, s [ (see [9]). Also the set \({S}^{\mbox{ prox}}\) of proximal elements in S is open and the set of corresponding positive dominant eigenvalues generates a dense subgroup of \(\mathbb{R}_{+}^{{\ast}}\). Furthermore, the action of μ on \({\mathbb{P}}^{d-1}\) has a unique μ-stationary measure ν and \(\mathrm{supp}\nu\) is the unique S-minimal subset of \({\mathbb{P}}^{d-1}\); the set \(\Lambda (S) =\mathrm{ supp}\nu\) is the closure of \(\{\bar{w}_{g};g \in {S}^{\mbox{ prox}}\}\) and has positive Hausdorff dimension.

Under condition i − p and for the action of S on \({\mathbb{S}}^{d-1}\), there are two cases I, II, say, regarding the existence of a convex S-invariant cone in V. In case I (non-existence), the inverse image \(\tilde{\Lambda }(S)\) of \(\Lambda (S)\) in \({\mathbb{S}}^{d-1}\) is the unique S-minimal invariant set in \({\mathbb{S}}^{d-1}\). In case II (existence), \(\tilde{\Lambda }(S)\) splits into two symmetric S-minimal subsets \(\tilde{\Lambda }_{+}(S)\) and \(\tilde{\Lambda }_{-}(S)\).

Returning to the affine situation, we need to consider the compactification \(V \cup \mathbb{S}_{\infty }^{d-1}\) of V by the sphere at infinity \(\mathbb{S}_{\infty }^{d-1}\) and we identify \(\tilde{\Lambda }(S)\) (resp. \(\tilde{\Lambda }_{+}(S),\tilde{\Lambda }_{-}(S)\)) with the corresponding subset \(\tilde{{\Lambda }}^{\infty }(S)\) (resp. \(\tilde{\Lambda }_{+}^{\infty }(S),\tilde{\Lambda }_{-}^{\infty }(S)\)) of \(\mathbb{S}_{\infty }^{d-1}\). We observe that \(\mathbb{S}_{\infty }^{d-1}\) is H-invariant and the corresponding H-action reduces to the G-action.

If h = (g, b) is such that | g |  < 1, then h has a fixed point x ∈ V, and this point is attractive, i.e. for any \(y \in V,\;\lim _{n\rightarrow \infty }{h}^{n}y = x\). The set Δ a (T) of such attractive fixed points of elements h ∈ T plays an important role in the description of \(\mathrm{supp}\,\rho\), for ρ λ-stationary.

On the other hand, if for some s > 0 we have k(s) > 1 and condition i − p is satisfied, then one can show the existence of g ∈ S with \(\lim _{n\rightarrow \infty }\vert {g}^{n}{\vert }^{1/n} > 1\). This implies that \(\mathrm{supp}\rho\) is unbounded, if suppλ has no fixed point in V.

We have the following basic (see [10], Proposition 5.1)

Proposition 1.

Assume \(\mathbb{E}(\log \gamma (A)) +\log \vert B\vert ) < \infty \) and

$$\displaystyle{L_{\mu } =\lim _{n\rightarrow \infty }\frac{1} {n}\,\mathbb{E}(\log \vert S_{n}\vert ) < 0.}$$

Then \(R_{n} =\sum _{ 1}^{n}A_{1}\ldots A_{k-1}B_{k}\) converges \(\mathbb{P}\) -a.e. to

$$\displaystyle{R =\sum _{ 1}^{\infty }A_{ 1}\ldots A_{k-1}B_{k},}$$

and for any \(x \in V,\;X_{n}\) converges in law to R. If \(\beta \in I_{\mu }\) satisfies \(k(\beta ) < 1, \mathbb{E}(\vert B{\vert }^{\beta }) < \infty \) , then \(\mathbb{E}(\vert R{\vert }^{\beta }) < \infty \) .

The law ρ of R is the unique λ-stationary measure on V. The closure \(\overline{\Delta _{a}(T)} = \Lambda _{a}(T)\) in V is the unique T-minimal subset of V and Λ a (T) = supp  ρ. If the semigroup S satisfies condition i − p and supp λ has no fixed point in V, then ρ(W) = 0 for any affine subspace W. Furthermore, if T contains an element (g,b) with \(\lim _{n\rightarrow \infty }\vert {g}^{n}{\vert }^{1/n} > 1\) , then Λ a (T) is unbounded.

The first part of the proposition is well known (see for example [12]).

For s ≥ 0, we denote by l s (resp. h s) the s-homogeneous measure (resp. function) on \(\mathbb{R}_{+}^{{\ast}}\) given by \({l}^{s}(dt) = {t}^{-(s+1)}dt,\;{l}^{0} = l\) (resp. \({h}^{s}(t) = {t}^{s}\)). We observe that the cone of Radon measures on \(\dot{V }\) which are of the form η ⊗ l s with η a positive measure on \({\mathbb{P}}^{d-1}\) is G-invariant. Also \(g(\eta \otimes {l}^{s}) = (\rho _{s}(g)\eta ) \otimes {l}^{s}\) with

$$\displaystyle{\rho _{s}(g)\eta =\int \vert gx{\vert }^{s}\delta _{ g\cdot x}d\eta (x).}$$

One can show that if the subsemigroup S associated to μ satisfies condition i − p and s ∈ I μ , there exists a unique probability measure ν s on \({\mathbb{P}}^{d-1}\) such that equation \(\mu {\ast}{(\nu }^{s} \otimes {l}^{s}) = k{(s)\nu }^{s} \otimes {l}^{s}\) is satisfied. Furthermore ν s gives mass zero to any projective hyperplane and \(\mathrm{{supp}\,\nu }^{s} = \Lambda (S)\).

We denote by \({\tilde{\nu }}^{s}\) the unique symmetric positive measure on \({\mathbb{S}}^{d-1}\) with projection ν s on \({\mathbb{P}}^{d-1}\) and (in case II) by \(\tilde{\nu }_{+}^{s},\;\tilde{\nu }_{-}^{s}\) its normalized restrictions to \(\tilde{\Lambda }_{+}(S),\;\tilde{\Lambda }_{-}(S)\) hence \({\tilde{\nu }}^{s} = \frac{1} {2}(\tilde{\nu }_{+}^{s} +\tilde{\nu }_{ -}^{s})\). Then we have

$$\displaystyle{\mu {\ast}{(\tilde{\nu }}^{s} \otimes {l}^{s}) = k{(s)\tilde{\nu }}^{s} \otimes {l}^{s}}$$

and

$$\displaystyle{\mu {\ast}(\tilde{\nu }_{+}^{s} \otimes {l}^{s}) = k(s)\tilde{\nu }_{ +}^{s} \otimes {l}^{s},\quad \mu {\ast} (\tilde{\nu }_{ -}\otimes {l}^{s}) = k(s)\tilde{\nu }_{ -}^{s} \otimes {l}^{s}.}$$

If there exists \(\alpha \in I_{\mu }\) such that k(α) = 1, the measures \({\tilde{\nu }}^{\alpha } \otimes {l}^{\alpha },\;\tilde{\nu }_{+}^{\alpha } \otimes {l}^{\alpha },\;\tilde{\nu }_{-}^{\alpha }\otimes {l}^{\alpha }\) enter in an essential way in the description of the “shape” of ρ. We need first to discuss the action of S on \(\mathbb{S}_{\infty }^{d-1}\), if supp ρ is unbounded. In this case \(\overline{\mathrm{supp}\,\rho } \cap \mathbb{S}_{\infty }^{d-1}\) is a non trivial closed S-invariant set, hence three cases can occur, in view of the above discussion of minimality.

Case I::

S has no proper convex invariant cone and \(\overline{\Lambda _{a}(T)} \supset \tilde{ {\Lambda }}^{\infty }(S)\).

Case II’::

S has a proper convex invariant cone and \(\overline{\Lambda _{a}(T)} \supset \tilde{ {\Lambda }}^{\infty }(S)\).

Case II”::

S has a proper convex invariant cone and \(\overline{\Lambda _{a}(T)}\) contains only one of the sets \(\tilde{\Lambda }_{+}^{\infty }(S),\;\tilde{\Lambda }_{-}^{\infty }(S)\), say \(\tilde{\Lambda }_{+}^{\infty }(S)\) hence \(\tilde{\Lambda }_{-}^{\infty }(S) \cap \overline{\Lambda _{a}(T)} = \emptyset \).

The push-forward of a measure η on V by the dilation x → tx (t > 0) will be denoted t. η. For d > 1, our main result in [10] is the following

Theorem 1.

With the above notations, assume that S satisfies condition i − p,  that T has no fixed point in V, that L μ < 0, and that there exists α > 0 with k(α) = 1 and \(\mathbb{E}(\vert A{\vert {}^{\alpha }\gamma }^{\delta }(A)) < \infty,\; \mathbb{E}(\vert B{\vert }^{\alpha +\delta }) < \infty \) for some δ > 0. Then supp  ρ is unbounded and we have the following vague convergence on \(V \setminus \{0\}\) :

$$\displaystyle{\lim _{t\rightarrow 0+}{t}^{-\alpha }(t.\rho ) = C{(\sigma }^{\alpha } \otimes {l}^{\alpha }) = \Lambda,}$$

where \(C > 0{,\;\sigma }^{\alpha }\) is a probability on \({\mathbb{S}}^{d-1}\) and the Radon measure Λ satisfies μ ∗ Λ = Λ. Moreover,

$${\displaystyle{\sigma }^{\alpha }\ =\ \left \{\begin{array}{@{}l@{\quad }l@{}} {\tilde{\nu }}^{\alpha } \quad &\text{in Case I}, \\ C_{+}\tilde{\nu }_{+}^{\alpha } + C_{-}\tilde{\nu }_{-}^{\alpha }\text{ for some }C_{+},C_{-} > 0\quad &\text{in Case II}^{\prime}, \\ \tilde{\nu }_{+}^{\alpha } \quad &\text{in Case II}\textquotedblright. \end{array} \right.}$$

The measures \({\tilde{\nu }}^{\alpha } \otimes {l}^{\alpha }\) (case I), \(\tilde{\nu }_{+}^{\alpha } \otimes {l}^{\alpha }\) and \(\tilde{\nu }_{-}^{\alpha }\otimes {l}^{\alpha }\) (cases II’, II”) are minimal μ-harmonic measures.

The above convergence is valid on any Borel function f with \({\sigma }^{\alpha } \otimes {l}^{\alpha }\) -negligible set of discontinuities such that \(\vert w{\vert }^{-\alpha }\vert \log \vert w{\vert }^{1+\varepsilon }\vert f(w)\vert \) is bounded for some ε > 0, hence

$$\displaystyle{\lim _{t\rightarrow 0+}{t}^{-\alpha }\mathbb{E}(f(tR)) = \Lambda (f).}$$

The theorem shows that ρ belongs to the domain of attraction of a stable law, a fact conjectured by F. Spitzer. It plays a basic role in the study of slow diffusion for random walk in a random medium on \(\mathbb{Z}\) (see [7]), and also in extreme value theory for GARCH processes. The proof of the theorem shows that the above convergence is valid on the sets \(H_{w}^{+} =\{ x \in V;< x,w > 1\}\) for \(w \in V \setminus \{0\}\) under the weaker hypothesis \(\mathbb{E}(\vert A{\vert }^{\alpha }\log \gamma (A) + \vert B{\vert }^{\alpha +\delta }) < \infty \). Then, using [1], it follows that the theorem is valid if \(\alpha \notin \mathbb{N}\). Actually, [1] implies also the validity of the theorem under the above weaker hypothesis, in the following situations:

Case I and:

\(\alpha \notin 2\mathbb{N}\),

Case II” and:

α > 0,

Case II’ and:

\(C_{+} = C_{-},\;\alpha \notin 2\mathbb{N}\).

As observed in [14], the condition \(C_{+} = C_{-}\) occurs if ρ is symmetric, in particular if the law of B is symmetric (for example if B is Gaussian). In the context of extreme value theory the convergence stated in the theorem says that ρ has “multivariate regular variation”. This property is basic for the development of the theory for “ARCH processes” (see [14]).

The proof given in [10] (Theorem 6) is long. For a short survey of earlier work, see [8]. Here we will give a sketch of a few main points of the proof.

2 Some Tools for the Proof of the Theorem

2.1 The Renewal Theorem for Products of Random Matrices (d > 1)

We use the notations already introduced above: μ is a probability measure on G = GL(V ), S the closed subsemigroup of G generated by supp μ,  L μ the top Lyapunov exponent of μ,  ν the μ-stationary measure on \({\mathbb{P}}^{d-1}\) etc. Under condition i − p, the following is the d-dimensional analog of the classical renewal theorem (see [4]) and follows from the general renewal theorem of Kesten [13] for Markov random walks on \(\mathbb{R}\).

Theorem 2.

Assume that the semigroup S associated with μ satisfies condition i − p,   that log γ(g) is μ-integrable, and that \(L_{\mu }\,=\,\lim _{n\rightarrow \infty }\frac{1} {n}\int \log \vert g\vert {d\mu }^{n}(g)\,>\,0\) . Then, for any w ∈ V, \(\sum _{0}^{\infty }{\mu }^{k} {\ast}\delta _{w}\) is a Radon measure on \(\dot{V }\) and we have

$$\displaystyle{\lim _{w\rightarrow 0}\,\sum _{0}^{\infty }{\mu }^{k} {\ast}\delta _{ w}\ =\ \frac{1} {L_{\mu }}{\nu }^{0} \otimes l.}$$

in the sense of vague convergence. This convergence is also valid on any bounded continuous function f on \(\dot{V }\) with \(\sum _{-\infty }^{\infty }\sup \{\vert f(w)\vert;{2}^{l} \leq \vert w\vert \leq {2}^{l+1}\} < \infty \) .

As proved in [10], if S satisfies i − p, s ∈ I μ and \(\int \vert g{\vert }^{s}\log \gamma (g)d\mu (g) < \infty \), then the top Lyapunov exponent \(L_{\mu }(s) =\lim _{n\rightarrow \infty }\frac{1} {n}\int \vert g{\vert }^{s}\log \vert g\vert {d\mu }^{n}(g)\) exists, is simple and satisfies \(L_{\mu }(s) = \frac{k^{\prime}(s_{-})} {k(s)} < \infty \). Also there exists a unique positive function e s on \({\mathbb{P}}^{d-1}\) such that ν s(e s) = 1 and

$$\displaystyle{\mu {\ast}\delta _{w}({e}^{s} \otimes {h}^{s}) = k(s)({e}^{s} \otimes {h}^{s})(w).}$$

Then, using [13] again, we have the following result which includes information on the fluctuations of S n w:

Theorem 3.

Assume that \(L_{\mu } < 0,\;\alpha \in I_{\mu }\) exists with \(\alpha > 0,\;k(\alpha ) = 1,\int \vert g{\vert }^{\alpha }\log \gamma (g)d\mu (g) < \infty \) , and S satisfies condition i − p. Then we have the following vague convergence on \(\dot{V }\) , for any \(w \in \dot{ V }\)

$$\displaystyle{\lim _{t\rightarrow 0+}{t}^{-\alpha }\sum _{ 0}^{\infty }{\mu }^{k} {\ast}\delta _{ tw}\ =\ \frac{({e}^{\alpha } \otimes {h}^{\alpha })(w)} {L_{\mu }(\alpha )} {\nu }^{\alpha } \otimes {l}^{\alpha }.}$$

This convergence is actually valid on any continuous function f on \(\dot{V }\) such that \(f_{\alpha }(w) = \vert w{\vert }^{-\alpha }f(w)\) is bounded and \(\sum _{-\infty }^{\infty }\sup \{f_{\alpha }(w);{2}^{l} \leq \vert w\vert \leq {2}^{l+1}\} < \infty \) .

In particular for some A > 0 and any w ∈ V

$$\displaystyle{\lim _{t\rightarrow \infty }{t}^{\alpha }\,\mathbb{P}\{\sup _{n\geq 1}\vert S_{n}w\vert > t\}\ =\ A({e}^{\alpha } \otimes {h}^{\alpha })(w).}$$

The last formula is the so-called Cramér estimate of ruin in collective risk systems if d = 1 [4].

For the convergence proof in Theorem 1, we will need an analogue of Theorem 3 with \(\dot{V }\) replaced by \(V \setminus \{0\}\). For \(u \in {\mathbb{S}}^{d-1}\), the function e α(u) can be lifted to \({\mathbb{S}}^{d-1}\) and we have

$$\displaystyle{\int \vert gu{\vert }^{\alpha }\,\frac{{e}^{\alpha }(g.u)} {{e}^{\alpha }(u)} \ {d\mu }^{n}(g) = 1}$$

for any \(n \in \mathbb{N}\). Hence the family of probability measures \(\vert gu{\vert }^{\alpha }\frac{{e}^{\alpha }(g.u)} {{e}^{\alpha }(u)} {d\mu }^{\otimes n}(g)\) with \(g = g_{1}\ldots g_{n}\) defines a projective system on the spaces G  ⊗ n and one can consider the projective limit \(\mathbb{Q}_{u}^{\alpha }\) on \({G}^{\otimes \mathbb{N}}\). Referring again to [13], we get the following

Theorem 4.

Assume μ and α are as in Theorem  3 . Then, for any \(u \in {\mathbb{S}}^{d-1}\) , we have the vague convergence

$$\displaystyle{\lim _{t\rightarrow 0+}{t}^{-\alpha }\sum _{ 0}^{\infty }{\mu }^{k} {\ast}\delta _{ tu}\ =\ \frac{1} {L_{\mu }(\alpha )}{e}^{\alpha }(u)\ \tilde{\nu }_{u}^{\alpha } \otimes {l}^{\alpha },}$$

where \(\tilde{\nu }_{u}^{\alpha }\) is a probability measure on \({\mathbb{S}}^{d-1}\) and \(\tilde{\nu }_{u}^{\alpha } \otimes {l}^{\alpha }\) is a μ-harmonic Radon measure on \(V \setminus \{0\}\) . The convergence is valid on any continuous function f such that \(f_{\alpha }(w) = \vert w{\vert }^{-\alpha }f(w)\) is bounded and satisfies

$$\displaystyle{\sum _{-\infty }^{\infty }\sup \{\vert f_{\alpha }(w)\vert;{2}^{l} \leq \vert w\vert \leq {2}^{l+1}\} < \infty.}$$

There are two cases:

Case I: :

\(\tilde{\nu }_{u}^{\alpha } =\tilde{\nu }\) has support \(\tilde{\Lambda }(S)\) .

Case II: :

\(\tilde{\nu }_{u}^{\alpha } = p_{+}^{\alpha }(u)\tilde{\nu }_{+}^{\alpha } + p_{-}^{\alpha }(u)\tilde{\nu }_{-}^{\alpha }\) , where \(p_{+}^{\alpha }(u)\) (resp. \(p_{-}^{\alpha }(u)\) ) is the entrance probability under \(\mathbb{Q}_{u}^{\alpha }\) of S n ⋅ u into the convex envelope of \(\tilde{\Lambda }_{+}(S)\) (resp. \(\tilde{\Lambda }_{-}(S)\) ).

These results improve earlier ones by Kesten [12] and Le Page [16].

2.2 A Spectral Gap Property for Convolution Operators (d > 1)

As above we consider the operator P on \(\dot{V }\) defined by Pf(w) = (μ ∗ δ w )(f) and its action on s-homogeneous functions. The Euclidean norm on V extends to a norm on the wedge product \({\bigwedge }^{2}V\): For x, y, x′, y′ ∈ V, we put

$$\displaystyle{< x\wedge y,x^{\prime}\wedge y^{\prime} >\:=\ \text{det}\left (\begin{array}{cc} < x,x^{\prime} >& < x,y^{\prime} >\\ < y, x^{\prime} > & < y, y^{\prime} > \end{array} \right).}$$

This allows to consider the distance δ on \({\mathbb{P}}^{d-1}\) defined by \(\delta (x,y) = \vert x \wedge y\vert \), where x, y correspond to unit vectors \(\tilde{x},\tilde{y}\) in \({\mathbb{S}}^{d-1}\). We will denote by \(H_{\varepsilon }({\mathbb{P}}^{d-1})\) the space of \(\varepsilon\)-Hölder functions on \({\mathbb{P}}^{d-1}\) with respect to the distance δ. We write

$$\displaystyle{[\varphi ]_{\varepsilon } =\sup _{x\not =y}\frac{\vert \varphi (x) -\varphi (y)\vert } {\delta {(x,y)}^{\varepsilon }},\quad \vert \varphi \vert =\sup _{x}\vert \varphi (x)\vert,\quad \vert \varphi \vert _{\varepsilon } = [\varphi ]_{\varepsilon } + \vert \varphi \vert,}$$

and we observe that \(\varphi \rightarrow \vert \varphi \vert _{\varepsilon }\) defines a norm on \(H_{\varepsilon }({\mathbb{P}}^{d-1})\).

If \(z \in \mathbb{C},\;z = s + \mathit{it}\), and the z-homogeneous function f on \(\dot{V }\) is of the form \(f =\varphi \otimes {h}^{z}\), with \(\varphi \in H_{\varepsilon }({\mathbb{P}}^{d-1})\), the action of P on f defines an operator P z on \(\varphi\) by

$$\displaystyle{Pf = {P}^{z}\varphi \otimes {h}^{z},\quad \mbox{ i.e.}\quad {P}^{z}\varphi (x) =\int \varphi (g \cdot x)\,\vert gx{\vert }^{z}\ d\mu (g).}$$

Then we have the following (see [10], Theorem A)

Theorem 5.

Let d > 1 and assume that the closed subsemigroup S generated by supp  μ satisfies condition i − p. For s ∈ I μ , assume \(\int \vert g{\vert {}^{s}\gamma }^{\delta }(g)d\mu (g) < \infty \) for some δ > 0. Then, for any \(\varepsilon > 0\) sufficiently small, the operator P s on \(H_{\varepsilon }({\mathbb{P}}^{d-1})\) has a spectral gap, with dominant eigenvalue k(s):

$$\displaystyle{{P}^{s} = k(s){(\nu }^{s} \otimes {e}^{s} + U_{ s}),}$$

where ν s ⊗ e s is the projection on \(\mathbb{C}{e}^{s}\) defined by \({\nu }^{s},{e}^{s}\) and U s is an operator with spectral radius less than 1 which commutes with \({\nu }^{s} \otimes {e}^{s}\) . Furthermore, if \(\mathfrak{I}z = t\not =0,\;z = s + \mathit{it}\) , then the spectral radius of P z is less than k(s).

If s = 0,  P s reduces to convolution by μ on \({\mathbb{P}}^{d-1}\) and convergence to the unique μ-stationary measure ν 0 = ν was a basic property studied in [5]. In this case the spectral gap property is a consequence of the simplicity of the top Lyapunov exponent of μ (see [2, 9]). The spectral gap properties of P s are basic ingredients for the study of precise large deviations for the product of random matrices \(S_{n} = A_{n}\ldots A_{1}\) (see [16, 18]). Here the theorem will be used for the study of s-homogeneous P-eigenmeasures on \(\dot{V }\) and \(V \setminus \{0\}\). In the context of V ∖ {0} we need to replace \({\mathbb{P}}^{d-1}\) by \({\mathbb{S}}^{d-1}\) and to use an analogous theorem (see [10]).

2.3 A Choquet-Deny Property for Markov Walk

Here (S, δ) is a compact metric space and P is a Markov kernel on \(S \times \mathbb{R} = Y\) which commutes with the \(\mathbb{R}\)-translations and acts continuously on the space \(C_{b}(S \times \mathbb{R})\) of continuous bounded functions on \(S \times \mathbb{R}\). Such a set of datas will be called a Markov walk on \(\mathbb{R}\). We define for \(t \in \mathbb{R}\) the Fourier operator P it on C(S) by

$$\displaystyle{{P}^{\mathit{it}}\varphi (x) = P(\varphi \otimes e_{\mathit{ it}})(x,0),}$$

where e it is the Fourier exponential on \(\mathbb{R},\;e_{\mathit{it}}(r) = {e}^{\mathit{itr}}\). For \(t = 0,\;{P}^{\mathit{it}} = {P}^{0}\) is equal to \(\bar{P}\), the factor operator on S defined by P. We assume that for \(\varepsilon > 0\;{P}^{\mathit{it}}\) preserves the space of \(\varepsilon\)-Hölder functions \(H_{\varepsilon }(S)\) on (S, δ) and is a bounded operator on \(H_{\varepsilon }(S)\).

We denote \([\varphi ]_{\varepsilon } =\sup _{x\not =y}\frac{\vert \varphi (x)-\varphi (y)\vert } {\delta {(x,y)}^{\varepsilon }},\;\vert \varphi \vert =\sup _{x}\vert \varphi (x)\vert \) for ϕ ∈ C(S). Moreover, we assume that P it and P satisfy the following condition D:

  1. 1.

    For any \(t \in \mathbb{R}\), one can find \(n_{0} \in \mathbb{N},\;\rho (t) \in [0,1[\) and C(t) ≥ 0 for which

    $$\displaystyle{[{({P}^{\mathit{it}})}^{n_{0} }\varphi ]_{\varepsilon } \leq \rho (t)[\varphi ]_{\varepsilon } + C(t)\vert \varphi \vert.}$$
  2. 2.

    For any \(t \in \mathbb{R}\), the equation \({P}^{it}\varphi = {e}^{i\theta }\varphi,\;\varphi \in H_{\varepsilon }(S),\;\varphi \not =0\), has only the trivial solution \({e}^{i\theta } = 1,\;t = 0,\;\varphi =\) constant.

  3. 3.

    For some \(\delta > 1:\; M_{\delta } =\sup _{x\in S}\int \vert a{\vert }^{\delta }P((x,0),d(y,a)) < \infty \).

Conditions 1 and 2 above imply that \(\bar{P}\) has a unique stationary measure π and the spectrum of \(\bar{P}\) in \(H_{\varepsilon }(S)\) is of the form \(\{1\}\, \cup \, \Delta \), where Δ is a compact subset of the open unit disk (see [11]). They imply also that for any t ≠ 0, the spectral radius of P it is less than one.

If \(Y =\dot{ V }\), P is the convolution operator by μ on \(\dot{V } = {\mathbb{P}}^{d-1} \times \mathbb{R}_{+}\;(d > 1)\), hence \(S = {\mathbb{P}}^{d-1}\) and \(\mathbb{R}_{+}^{{\ast}} =\exp \mathbb{R}\). Theorem 5 implies that condition D is satisfied if I μ  ≠ 0 and condition i − p is valid.

Furthermore, for s ∈ I μ one can also consider the Markov operator Q s on \(\dot{V }\) defined by

$$\displaystyle{Q_{s}f = \frac{1} {k(s){e}^{s} \otimes {h}^{s}}P(f{e}^{s} \otimes {h}^{s}).}$$

If for some \(\delta > 0,\;\int \vert g{\vert {}^{s}\gamma }^{\delta }(g)d\mu (g) < \infty \), Theorem 5 implies that conditions D are also satisfied by Q s .

We will say that a Radon measure θ on \(Y = S \times \mathbb{R}\) is translation-bounded if for any compact \(K \subset Y\) there exists C(K) > 0 such that \(\theta (K + t) \leq C(K)\) for any \(t \in \mathbb{R}\), where K + t is the set obtained from K by translation with t. Then we have the following Choquet-Deny type property

Theorem 6.

With the above notations if the Markov operator P on \(Y = S \times \mathbb{R}\) satisfies the condition D. Then any translation-bounded P-harmonic measure on Y is proportional to π ⊗ l with l = dt.

This theorem can be used for \(Y =\dot{ V }\) and P = Q α if 0 < α < s .

2.4 A Weak Renewal Theorem

As in the Sect. 2.3, we consider a Markov walk P on \(\mathbb{R}\) with compact factor space S, a probability ν on S such that ν ⊗ l is P-invariant. A path starting from S for this Markov chain will be denoted \((X_{n},V _{n})\) with \(X_{n} \in S,\;V _{n} \in \mathbb{R}\) and the canonical probability measure on the paths starting from x ∈ S will be denoted by \({}^{a}\mathbb{P}_{x}\). We write also \({}^{a}\mathbb{P}_{\nu } {=\int \, }^{a}\mathbb{P}_{x}d\nu (x)\).

For a non negative Borel function on \(S \times \mathbb{R}\), we write \(U\psi =\sum _{ 0}^{\infty }{P}^{k}\psi\). We observe that if \((x,t) \in S \times \mathbb{R}\), ψ = 1 K , then (x, t) is the expected number of visits to K starting from \((x,t) \in S \times \mathbb{R}\). In other words \(U\psi (x,t) = \mathbb{E}_{x}\left (\sum _{0}^{\infty }\psi (X_{k},t + V _{k})\right )\). Then we have the following weak analogue of the renewal theorem.

Proposition 2.

Suppose that ψ is a bounded, non-negative and compactly supported Borel function on \(S \times \mathbb{R}\) . Further suppose that the potential \(U\psi =\sum _{ 0}^{\infty }{P}^{k}\psi\) is locally bounded and that, for any \(\varepsilon > 0\),

$$\displaystyle{{\lim _{n\rightarrow \infty }}^{a}\mathbb{P}_{ v}\left \{\left \vert \frac{V _{n}} {n} -\gamma \right \vert >\varepsilon \right \} = 0\quad \mbox{ with}\quad \gamma < 0}$$

holds true. Then

$$\displaystyle{\lim _{t\rightarrow \infty }\frac{1} {t}\int _{0}^{t}ds\int _{ s}U\psi (x,s)d\nu (x) = \frac{1} {\vert \gamma \vert }\int \int _{-\infty }^{\infty }\psi (x,s)d\nu (x)ds.}$$

If ψ is a non-negative Borel function on S such that \(\lim _{t\rightarrow \infty }U\psi (x,t) = 0\;\nu\) -a.e., then \(\psi = 0\;\nu \otimes l\) -a.e.

3 Elements of Proof of Theorem 1

3.1 Convergence for Radon Transforms

For a finite measure η on V we write \(\hat{\eta }(w) =\eta (H_{w}^{+})\) where \(u = tw,\;t > 0,\;u \in {\mathbb{S}}^{d-1},\;H_{w}^{+} =\{ x \in V;< x,w >> 1\}\). We observe that \(\hat{\eta }\) can be considered as an integrated form of the Radon transform of η. Observe that \(\widehat{\mu {\ast}\eta }(w) = {(\mu }^{{\ast}}{\ast}\delta _{w})(\hat{\eta })\), hence convolution equations on G ×V can be transformed to functional equations for Radon transforms.

We will not be able to apply directly the renewal Theorem 4 to the convolution equation \(\lambda {\ast}\rho =\rho\) corresponding to \(R = A{R}^{1} + B\) but rather to functional equations for \(\hat{\rho }\) and μ  ∗ . We denote by ρ 1 the law of R − B and we begin with the

Proposition 3.

With the hypothesis of Theorem  4 , we denote by \({}^{{\ast}}\tilde{\nu }_{u}^{\alpha }\) the positive kernel on \({\mathbb{S}}^{d-1}\) given by Theorem  4 and associated with μ . Then one has the equations on \(V \setminus \{0\}\)

$$\displaystyle{\rho =\sum _{ 0}^{\infty }{\mu }^{k} {\ast} (\rho -\rho _{ 1}),\qquad \hat{\rho }(w) =\sum _{ 0}^{\infty }({({\mu }^{{\ast}})}^{k} {\ast}\delta _{ w})(\hat{\rho }-\hat{\rho }_{1}).}$$

For \(u \in {\mathbb{S}}^{d-1}\) , if \(\alpha \in ]0,s_{\infty }[,\;k(\alpha ) = 1\) , the function \(t \rightarrow {t}^{\alpha -1}(\hat{\rho }-\hat{\rho }_{1})(u,t)\) is Riemann-integrable on ]0,∞[ and one has, with \(r_{\alpha }(u) =\int _{ 0}^{\infty }{t}^{\alpha -1}(\hat{\rho }-\hat{\rho }_{1})(u,t)dt\)

$$\displaystyle {\lim _{t\rightarrow \infty }{t}^{\alpha }\hat{\rho }(u,t) = \frac{{}^{{\ast}}{e}^{\alpha }(u)} {L_{\mu }(\alpha )} \ {}^{{\ast}}\tilde{\nu }_{ u}^{\alpha }(r_{\alpha }) = C({\sigma }^{\alpha } \otimes {l}^{\alpha })(H_{ u}^{+}),}$$

where C ≥ 0 and the probability σ α on \(\tilde{\Lambda }(S)\) satisfies \(\mu {\ast}{(\sigma }^{\alpha } \otimes {l}^{\alpha }) {=\sigma }^{\alpha } \otimes {l}^{\alpha }\) . There exists b > 0 such that \(\mathbb{P}\{\vert R\vert > t\} \leq b{t}^{-\alpha }\) . Furthermore supp  ρ is unbounded and: In case I: \({\sigma }^{\alpha } {=\tilde{\nu } }^{\alpha }\) , in case II: \({C\sigma }^{\alpha } = C_{+}\tilde{\nu }_{+}^{\alpha } + C_{-}\tilde{\nu }_{-}^{\alpha },\;C_{+},C_{-}\geq 0\) .

3.1.1 Sketch of Proof

Since \(\vert {g}^{{\ast}}\vert = \vert g\vert \), the function k(s) is equal to the corresponding function for \({\mu }^{{\ast}}\), condition i − p is satisfied for \({\mu }^{{\ast}}\) and \(L_{{\mu }^{{\ast}}}(\alpha ) = L_{\mu }(\alpha )\). We observe that the stationarity equation \(R - B = A{R}^{1}\) can be written in distribution as \(\rho -\rho _{1} =\rho -\mu {\ast}\rho\). Also \(\rho (\{0\}) = 0\), hence we get

$$\displaystyle{\rho =\sum _{ 0}^{\infty }{\mu }^{k} {\ast} (\rho -\rho _{ 1}),\qquad \hat{\rho }(w) =\sum _{ 0}^{\infty }({({\mu }^{{\ast}})}^{k} {\ast}\delta _{ w})(\hat{\rho }-\hat{\rho }_{1})}$$

on \(V \setminus \{0\}\).

In order to use Theorem 4, we need to regularize \(\hat{\rho }-\hat{\rho }_{1}\) by multiplicative convolution on \(\mathbb{R}_{+}^{{\ast}}\) with 1[0, 1], hence to consider

$$\displaystyle{{r}^{\alpha }(u,t) = \frac{1} {t}\int _{0}^{t}{x}^{\alpha -1}(\hat{\rho }-\hat{\rho }_{ 1})(u,x)dx.}$$

Clearly \(\vert {r}^{\alpha }(u,t)\vert {\leq \alpha }^{-1}{t}^{\alpha -1}\). By using the conditions \(\mathbb{E}(\vert A{\vert }^{\alpha +\delta }) < \infty \) and \(\mathbb{E}(\vert B{\vert }^{\alpha +\delta }) < \infty \), one can show the existence of \(\delta ^{\prime} > 0,\;c(\delta ^{\prime}) > 0\) such that for t ≥ 1, 

$$\displaystyle{\vert {r}^{\alpha }(u,t)\vert \leq c(\delta ^{\prime}){t}^{-\delta ^{\prime}}.}$$

Then Theorem 4 can be applied to \(f_{\alpha }(w) = {r}^{\alpha }(u,t)\), whence, by a Tauberian argument as in [6], we get the convergence of \({t}^{\alpha }\hat{\rho }(u,t)\) towards \({ \frac{1} {L_{\mu }(\alpha )}{}^{{\ast}}{e}^{\alpha }(u)}^{{\ast}}\tilde{\nu }_{ u}^{\alpha }(r_{\alpha })\). From the existence of \(\alpha \in ]0,s_{\infty }[\) with k(α) = 1, one can deduce the existence of g ∈ S with | g |  > 1, hence supp ρ is unbounded.

The above formulae and the description of \({}^{{\ast}}{e}^{\alpha }{,\;\sigma }^{\alpha }\) in terms of \({\tilde{\nu }}^{\alpha },\tilde{\nu }_{+}^{\alpha },\tilde{\nu }_{-}^{\alpha }\) give the harmonicity equation \(\mu {\ast}{(\sigma }^{\alpha } \otimes {l}^{\alpha }) {=\sigma }^{\alpha } \otimes {l}^{\alpha }\). The boundedness of \({t}^{\alpha }P\{\vert R\vert > t\}\) follows from the convergence of \({t}^{\alpha }\hat{\rho }(u,t)\).

3.2 Homogeneity at Infinity of ρ

The boundedness of \({t}^{\alpha }P\{\vert R\vert > t\}\) stated in Proposition 3 implies that the family of Radon measures \(\{{t}^{-\alpha }(t.\rho );t \in \mathbb{R}_{+}\}\) is relatively compact in the vague topology.

Proposition 4.

Given the situation of Theorem  1 , assume that η is a vague limit of a sequence \(t_{n}^{-\alpha }(t_{n}\cdot \rho )\) as \(t_{n} \rightarrow \infty \) . Then η is translation-bounded and satisfies μ ∗η = η. If η and \(\sigma \otimes {l}^{\alpha }\) satisfy

$$\displaystyle{\eta (H_{u}^{+}) = (\sigma \otimes {l}^{\alpha })(H_{ u}^{+}),}$$

for any \(u \in {\mathbb{S}}^{d-1}\) and some positive measure σ on \({\mathbb{S}}^{d-1}\) , then \(\eta =\sigma \otimes {l}^{\alpha }\) .

This proposition is based on the moment conditions satisfied by R, A, B, and on Theorem 6. Using furthermore Propositions 4 and 3, we get the

Theorem 7.

With the hypothesis of Theorem  1 , we have the following vague convergence

$$\displaystyle{\lim _{t\rightarrow 0+}{t}^{-\alpha }(t.\rho ) = \Lambda = C{(\sigma }^{\alpha } \otimes {l}^{\alpha }),}$$

where C ≥ 0.

The above convergence is also valid on any Borel function f such that the set of discontinuities of f is \({(\sigma }^{\alpha } \otimes {l}^{\alpha })\) -negligible and such that for some \(\varepsilon > 0\) , the function \(\vert w{\vert }^{-\alpha }\vert \log \vert w\vert {\vert }^{1+\varepsilon }\vert f(w)\vert \) is bounded.

3.3 Positivity of \(C_{+},C_{-}\)

We need to consider processes (dual to X n ) and taking values in \((V \setminus \{0\}) \times \mathbb{R}\) or \({\mathbb{S}}^{d-1} \times \mathbb{R}\) and we write

$$\displaystyle{S^{\prime}_{n} = A_{n}^{{\ast}}\ldots A_{ 1}^{{\ast}}.}$$

Let M be a S  ∗ -minimal subset of \({\mathbb{S}}^{d-1}\) i.e. \(M =\tilde{ \Lambda }({S}^{{\ast}})\) in case I and \(M =\tilde{ \Lambda }_{+}({S}^{{\ast}})\) (or \((\tilde{\Lambda }_{-}({S}^{{\ast}}))\) in case II. We denote by \(\Lambda _{a}^{{\ast}}(T)\) the set of \(u \in {\mathbb{S}}^{d-1}\) such that the projection of ρ on the line \(\mathbb{R}u(u \in {\mathbb{S}}^{d-1})\) is unbounded in direction u. The following is the essential step in the discussion of positivity.

Proposition 5.

With the hypothesis of Theorem  1 , if \(\Lambda _{a}^{{\ast}}(T) \supset M\) , then for any u ∈ M

$$\displaystyle{C_{M}(u) =\lim _{t\rightarrow \infty }{t}^{\alpha }\mathbb{P}\{ < R,u >> t\} > 0.}$$

In order to explain the main points of the proof, we need to introduce some notations. We observe that R n satisfies the recursion

$$\displaystyle{< R_{n+1},w >=< R_{n},w > + < B_{n+1},S^{\prime}_{n}w >,}$$

hence \((S^{\prime}_{n}w,r+ < R_{n},w >)\) is a Markov walk on \(V \setminus \{0\} \times \mathbb{R}\) based on \({\mathbb{S}}^{d-1} \times \mathbb{R}\). If we write

$$\displaystyle{t^{\prime} = {r}^{-1},\;w = \vert w\vert u,\;p = r\vert w{\vert }^{-1}}$$

with \(u \in {\mathbb{S}}^{d-1}\) this Markov walk can be expressed on \(({\mathbb{S}}^{d-1} \times \mathbb{R}) \times {\mathbb{R}}^{{\ast}}\) as

$$\displaystyle{u_{n+1} = g_{n+1}^{{\ast}}.u_{ n},\;p_{n+1} = \frac{p_{n}+ < b_{n+1},u_{n} >} {\vert g_{n+1}^{{\ast}}u_{n}\vert },\;t^{\prime}_{n+1} = t^{\prime}_{n}{(\vert g_{n+1}^{{\ast}}u_{ n}\vert p_{n+1}p_{n}^{-1})}^{-1}.}$$

We denote by \({}^{{\ast}}\hat{P}\) the corresponding Markov kernel. Since \((S^{\prime}_{n}w,r+ < R_{n},w >)\) has equivariant projection S′ n w on \(V \setminus \{0\}\), we have \({}^{{\ast}}\hat{P}{(}^{{\ast}}{e}^{\alpha } \otimes {h}^{\alpha }) {= }^{{\ast}}{e}^{\alpha } \otimes {h}^{\alpha }\), hence we can consider the new relativized kernel \({}^{{\ast}}\hat{P}_{\alpha }\) and the corresponding Markov walk \((u_{n},p_{n},t^{\prime}_{n})\) over the chain \((u_{n},p_{n}) \in X = M \times \mathbb{R}\).

We denote

$${\displaystyle{}^{{\ast}}{q}^{\alpha }(u,g) = \vert {g}^{{\ast}}u{\vert}^{\alpha }\frac{{}^{{\ast}}{e}^{\alpha }({g}^{{\ast}}.u)} {{}^{{\ast}}{e}^{\alpha }(u)} }$$

and for h = (g, b) ∈ H, 

$$\displaystyle{{h}^{u}p = \frac{1} {\vert {g}^{{\ast}}u\vert }(p+ < b,u >);}$$

then the Markov kernel \({}^{{\ast}}\hat{{Q}}^{\alpha }\) of the chain (u n , p n ) is given by

$${\displaystyle{}^{{\ast}}\hat{{Q}}^{\alpha }\varphi (u,p) =\int \varphi {({g}^{{\ast}}\cdot u,{h}^{u}p)}^{{\ast}}{q}^{\alpha }(u,g)d\lambda (h).}$$

We have L μ (α) > 0 and M is minimal, hence it is easy to show that \({}^{{\ast}}\hat{{Q}}^{\alpha }\) has a unique stationary measure κ on X, and with respect to the Markov measure \({}^{{\ast}}\hat{\mathbb{Q}}_{u}^{\alpha }\) on X ×Ω we have \(\mathbb{E}_{u}^{\alpha }{(\log }^{+}\vert p\vert ) < \infty \) and \(\limsup _{n\rightarrow \infty }\vert S^{\prime}_{n}u\vert \vert p_{n}\vert = \infty \). We observe that, since L μ (α) > 0, the Markov walk \((u_{n},p_{n},t_{n}^{\prime})\) on \(X \times {\mathbb{R}}^{{\ast}}\) has negative drift, in additive notation.

The condition \(\Lambda _{a}^{{\ast}}(T) \supset M\) implies

$$\displaystyle{\kappa (M\times ]0,\infty [) > 0,\;\limsup _{n\rightarrow \infty }\vert S^{\prime}_{n}u\vert p_{n} = \infty,}$$

for p > 0.

We now consider the following \(\mathbb{N} \cup \{\infty \}\)-valued stopping time τ on X ×Ω defined by

$$\displaystyle{\tau = \mathit{Inf }\{n > 1;{p}^{-1} < R_{ n},u >> 0\},}$$

and we observe that, by definition of p n :

$$\displaystyle{\tau = \mathit{Inf }\{n > 1;{p}^{-1}p_{ n}\vert S^{\prime}_{n}u\vert > 1\},}$$

hence \({p}^{-1}p_{\tau } > 0\). Hence τ (resp. \({p}^{-1}p_{\tau }\vert S^{\prime}_{\tau }u\vert \vert )\) can be interpreted as the first ladder epoch (resp. height) of the Markov walk \({p}^{-1}p_{n}\vert S^{\prime}_{n}u\vert \) (see [4]).

Using Poincaré’s recurrence theorem and \(\lim _{n\rightarrow \infty }\vert S^{\prime}_{n}u\vert = \infty \) \({}^{{\ast}}\hat{\mathbb{Q}}_{u}^{\alpha }\)-a.e. we infer that \(\tau < {\infty \;}^{{\ast}}\hat{\mathbb{Q}}_{\kappa }^{\alpha }\)-a.e.

Let \({}^{{\ast}}\hat{{P}}^{\tau }{,\;}^{{\ast}}\hat{{Q}}^{\tau }\) be the stopped kernels of \({}^{{\ast}}\hat{P}{,\;}^{{\ast}}\hat{Q}\), respectively, defined by τ and let \({}^{{\ast}}\hat{P}_{\alpha }^{\tau }{,\;}^{{\ast}}\hat{{Q}}^{\alpha,\tau }\) be the corresponding relativised Markovian kernels. Then we have the

Lemma 1.

With \(tw = u \in {\mathbb{S}}^{d-1},\;t > 0\) , we write on \(X \times {\mathbb{R}}^{{\ast}}\)

$$\displaystyle\begin{array}{rcl} & &\psi (w,p) = \mathbb{P}\{{p}^{-1} < R,u >> t\},\;\psi _{\tau }(v,p) = \mathbb{P}\{t < {p}^{-1} < R,u > < t + {p}^{-1} < R_{\tau },u >\} {}\\ & & {\psi }^{\alpha } = {{(}^{{\ast}}{e}^{\alpha } \otimes {h}^{\alpha })}^{-1}\psi,\;\psi _{\tau }^{\alpha } = {{(}^{{\ast}}{e}^{\alpha } \otimes {h}^{\alpha })}^{-1}\psi _{ \tau }. {}\\ \end{array}$$

Then \(\psi =\sum _{ 0}^{\infty }{{(}^{{\ast}}\hat{{P}}^{\tau })}^{k}\psi _{\tau }{,\;\psi }^{\alpha } =\sum _{ 0}^{\infty }{{(}^{{\ast}}\hat{P}_{\alpha }^{\tau })}^{k}\psi _{\tau }^{\alpha }\) .

The proof is analogous to the first part of Proposition 3, in order to get the Poisson equation \(\psi _{\tau } =\psi {-}^{{\ast}}\hat{{P}}^{\tau }\psi\). Since \({p}^{-1}p_{\tau } > 0\), the operator \({}^{{\ast}}\hat{{Q}}^{\alpha,\tau }\) preserves \(X_{+} = M\times ]0,\infty [\). If \(\Lambda _{a}^{{\ast}}(T) \supset M\), then \(\kappa (X_{+}) > 0\). Since \(\mathbb{E}_{\kappa }^{\alpha }{(\log }^{+}\vert p\vert ) < \infty \), one can show that the Markov kernel \({}^{{\ast}}\hat{Q}_{x}^{\alpha,\tau }\) has an ergodic stationary measure \(\kappa _{+}^{\tau }\) which is absolutely continuous with respect to \(1_{X_{+}}\kappa\). Also we have, using the interpretation of τ as a return time in the dynamical system associated with \({}^{{\ast}}\hat{\mathbb{Q}}_{x}^{\alpha }\) and the bilateral shift,

$$\displaystyle{\mathbb{E}_{0}^{\alpha }(\tau ) =\int \mathbb{E}_{ u}^{\alpha }(\tau )d\kappa _{ +}^{\tau }(u,p) < \infty,\;\gamma _{\tau }^{\alpha } = \mathbb{E}_{\kappa _{ +}^{\tau }}^{\alpha }(\log ({p}^{-1}p_{\tau }\vert S^{\prime}_{\tau }u\vert )) \in ]0,\infty [}$$

with \(\gamma _{\tau }^{\alpha } = L_{\mu }(\alpha )\mathbb{E}_{0}^{\alpha }(\tau )\).

Now we can consider the Markov walk defined by \({}^{{\ast}}\hat{P}_{\alpha }^{\tau }\) on \(X_{+} \times \mathbb{R}_{+}^{{\ast}}\). In view of the above observations we can apply Proposition 2 to \({}^{{\ast}}\hat{P}_{\alpha }^{\tau }\) and \(\kappa _{+}^{\tau } \otimes l\). We recall that, in additive notation, this Markov walk has negative drift \(-\gamma _{\tau }^{\alpha } < 0\). If for some u ∈ M we have C M (u) = 0, then for p > 0 and \(u = tw\;(t > 0)\;\lim {_{t\rightarrow \infty }\psi }^{\alpha }(w,p) = 0\).

Using Proposition 3 we get \(\lim {_{t\rightarrow \infty }\psi }^{\alpha }(w,p) = 0\) for any u = tw ∈ M. In particular, this is valid \(\kappa _{+}^{\tau }\)-a.e., hence Proposition 2 implies \(\psi _{\tau }^{\alpha } = 0\;\kappa _{+}^{\tau } \otimes l\)-a.e., i.e.

$$\displaystyle{\mathbb{P}\{t < {p}^{-1} < R,u > < t + {p}^{-1} < R_{\tau },u >\}= 0.}$$

Since \({p}^{-1} < R_{\tau },u >> 0\), we get \({p}^{-1} < R,u >\leq 0\;\kappa _{+}^{\tau } \otimes \mathbb{P}\)-a.e., i.e. \(< R,u >\leq 0\;\mathbb{P}\)-a.e. This contradicts \(\Lambda _{a}^{{\ast}}(T) \supset M\). One can show that \(\Lambda _{a}^{{\ast}}(T) = {\mathbb{S}}^{d-1}\) in cases I, II’ and \(\Lambda _{a}^{{\ast}}(T) \supset \tilde{ \Lambda }_{+}({S}^{{\ast}})\) in case II”, hence C  +  > 0.

4 The One-Dimensional Case

If d = 1, the notations and definitions introduced in Sect. 1 make sense. Then \(G = {\mathbb{R}}^{{\ast}}\) and H = H 1 is the affine group “ax + b” of the line. Condition i − p is always satisfied for any probability μ on \({\mathbb{R}}^{{\ast}}\), and the analogue of Proposition 1 is valid verbatim. For the analogue of Theorem 1 one needs to consider the possibility that S resp. μ are arithmetic, i.e. S is contained in a subset of \({\mathbb{R}}^{{\ast}}\) of the form { ± a n} for some a > 0. The function k(s) has the explicit form

$$\displaystyle{k(s) =\int \vert a{\vert }^{s}d\mu (a).}$$

Also \(L_{\mu } =\int \log \vert a\vert d\mu (a) = k^{\prime}(0)\). Then, Theorem 1 has the following analogue, with weaker moment conditions.

Theorem 8.

Assume that the probability measure λ on H 1 and μ on \({\mathbb{R}}^{{\ast}}\) satisfy the following conditions

  1. (a)

    \(\mathbb{E}(\log \vert A\vert ) < 0,\;k(\alpha ) = 1\) , for some α > 0.

  2. (b)

    S is non arithmetic and T has no fixed point.

  3. (c)

    \(\mathbb{E}(\vert B{\vert }^{\alpha }) < \infty \) and \(\mathbb{E}\vert A{\vert }^{\alpha }\vert \log \vert A\vert \vert ) < \infty \) .

Then one has the following convergences:

$$\displaystyle\begin{array}{rcl} & & \lim _{t\rightarrow \infty }{t}^{\alpha }\mathbb{P}\{R > t\} = C_{+} {}\\ & & \lim _{t\rightarrow \infty }\vert t{\vert }^{\alpha }\mathbb{P}\{R < -t\} = C_{-}. {}\\ \end{array}$$

Either \(\mathrm{supp}\rho = \mathbb{R}\) and then \(C_{+},C_{-} > 0\) or supp ρ is a half-line [c,∞[ (resp. ] −∞,c]) and then \(C_{+} > 0,\;C_{-} = 0\) (resp. \(C_{-} > 0,C_{+} = 0\) ).

With respect to [6], the main new situation occurs for the discussion of positivity of C  + , if A n  > 0 and the r.v. B n may have arbitrary sign. The proof [17] uses only the classical renewal theorem and a spectral gap property for the Markov chain p n on \(\mathbb{R}\). If \(\mathrm{supp}\lambda\) does not preserve a half-line ] − , c], one considers τ as the entrance time of p n into ]0, [. The spectral gap property gives the finiteness of \(\mathbb{E}_{p}^{\alpha }(\tau )\) for any \(p \in \mathbb{R}\); using Wald’s identity for the random walk log | S n  | , one gets the finiteness and positivity of log | S τ  | and then one concludes as for d > 1. Under stronger assumptions, the positivity of C  +  has been obtained also in the more general context of [3], using a complex analytic method for Mellin transform due to E. Landau, and familiar in analytic number theory. The positivity of \(C_{+} + C_{-}\) was obtained in [6], using P. Levy’s symmetrisation method. For an analytic proof of these facts, using also Wiener-Ikehara theorem, see ([10], Appendix). In contrast to Theorem 1 and due to the Diophantine character of the hypothesis, the convergences stated in Theorem 8 are not robust under perturbation of λ in the weak topology. From that point of view, the respective roles of stable laws and of the Gaussian law are different for d = 1 and for d > 1.