Inside singularity sets of random Gibbs measures
Julien Barral and Stéphane Seuret
INRIA Rocquencourt, Equipe “Complex”, Domaine de Voluceau Rocquencourt, 78153 Le Chesnay cedex, France
-
Abstract.
We evaluate the scale at which the multifractal structure of some random Gibbs measures becomes discernible. The value of this scale is obtained through what we call the growth speed in Hölder singularity sets of a Borel measure. This growth speed yields new information on the multifractal behavior of the rescaled copies involved in the structure of statistically self-similar Gibbs measures. Our results are useful to understand the multifractal nature of various heterogeneous jump processes.
1 Introduction
Contrary to what happens with monofractal measures (for instance uniform measures on regular Cantor sets), multifractal measures exhibit simultaneously several different behaviors at small scales. It is natural to question from which scale the multifractal structure of these measures becomes discernible and remains stable. This paper introduces a notion which provides a way to examinate the value of this critical scale. This notion, that we call growth speed in singularity sets, is naturally related with multifractal measures.
In the following, we define and study the growth speed in singularity sets for a class of statistically self-similar measures which includes random Gibbs measures. This work requires refinements of the known theoretical results on the multifractal nature of these measures. Finally, we obtain rigorous estimates of the error made when approximating the asymptotic local behavior of the measure by observing it at a fine but fixed grid.
Before making precise all these notions, let us explain what one of our main motivations was. The new multifractal properties we point out in this paper are naturally involved in the small-scale structure analysis of some jump processes recently considered in [5, 8, 9] . Typical examples of such heterogeneous jump processes are Lévy processes in multifractal time.
Performing a multifractal time change in irregular processes is a natural idea when trying to build multi-parameter processes [25, 27, 33] . Indeed, such processes yield multifractal objects with an interesting structure, that may be more realistic than classical homogeneous jump processes (for instance like Lévy processes) for the purpose of modeling multifractal discontinuous phenomena (Internet traffic [23] , variations of financial prices [25] ). Another relevant property of these processes is that they provide new illustrations of multifractal formalisms [15, 11, 28, 5] .
Our results provide tools to study these processes. Indeed, the multifractal analysis of heterogeneous jump processes in [5, 8, 9] requires to deepen our knowledge regarding statistically self-similar singular measures generated by multiplicative processes. The fact that these measures are locally equivalent to a rescaled copy of themselves is exploited in a new direction using the notion of growth speed in the Hölder singularity sets of these copies. The growth speed yields new insights on the structure of the process, which are more precise than those obtained by only considering individually these copies as the same probabilistic object. In particular, it provides a new quantitative way of distinguishing two well-known families of statistically self-similar singular measures, namely the random Gibbs measures [19] and the independent random cascades, like Mandelbrot canonical cascades [24] .
This paper focuses on random Gibbs measures, the case of the Mandelbrot canonical cascades is very different and treated in [7] . The multifractal structure of random Gibbs measures has been extensively studied ([15, 31, 20, 13, 29, 4, 14] ). This topic is concerned with the size estimation of the Hölder singularity sets of such a measure
. These sets are defined as the level sets of the pointwise Hölder exponent
. The sizes of Hölder singularity sets are measured through their Hausdorff (or packing) dimension. It can be shown that these dimensions are obtained thanks to the Legendre transform of a kind of free energy function
related to
. More precisely, let
be an integer
and
the alphabet
. Suppose that we are working on the symbolic space
endowed with the product topology and the one-sided shift transformation
. If
, the
step cylinder about
in
is denoted by
. The measures we are interested in are associated with some (random) Hölder potentiel and the dynamical system
.
The function
considered in the multifractal formalism for measures in [15, 11] is obtained as follows: For every
, let
|
(1)
|
The Legendre transform of
at
is then
.
Then the Hölder singularity set of level
is defined as
(
stands for
). The Gibbs measures we consider obey the multifractal formalism in the sense that
when
.
This property is classically implied by the existence of a probability measure
of the same nature as
and such that
is concentrated on
. This measure
is called an analyzing measure of
at
.
The existence of the measure
has another important consequence regarding the possibility of measuring how the mass of
is distributed at a given large enough scale. Indeed, a direct consequence of the multifractal formalism ([32] ) and the existence of
is that for any
and
such that
, one has
|
(2)
|
The result we establish in this paper brings precisions on these sizes estimates. We consider a refined version of the sets
by considering, for any sequence
going down to 0, the sets
|
(3)
|
It is possible to choose
so that with probability one, for all the exponents
such that
, one has
.
Since the sets sequence
is non-decreasing and
, the growth speed
in
can be defined as the smallest value of
for which the
-measure of
reaches a certain positive fraction
of the mass of
, that is the number
Now for
and
let
|
(4)
|
Heuristically, one has
i.e.
controls by above the smallest rank
from which considering the evaluation of
at any scale
smaller than
yields a correct representation of the asymptotic behavior of
.
Our results concern estimates of the growth speed of singularities sets of copies of
involved in the self-similarity property of
. To illustrate our purpose, let us describe the model of statistically self-similar measures we shall work with in the sequel. We shall consider a natural random counterpart to quasi-Bernoulli measures introduced in [26, 11] and mainly illustrated by deterministic Gibbs measures on
. We are inspired in particular by self-similar Riesz products and their random version constructed with random phases (see [13] and examples of Section 3 ).
1.1 Quasi-Bernoulli independent random measure
In the sequel
means equality in distribution.
A random probability measure
on
is said to be a quasi-Bernoulli independent random measure if there exists a constant
and two sequences of random measures
and
such that for every
,
(P1)
,
,
(P2) for every
,
,
(P3)
.
is also denoted
,
(P4)
and
are independent.
The measures
are the copies of
mentioned in the paragraphs above.
1.2 Controlling the growth speed in Hölder singularity sets of the
's
Let
be quasi-Bernoulli independent measure. For each copy
of
, the corresponding family of analyzing measures
will be defined as
is defined for
. The result we focus on is the asymptotic behavior of
|
(5)
|
For sake of simplicity, we give in this introduction a shorter version of our main result (Theorem 2 ).
Theorem A. Suppose that
is
. With probability one, for all
such that
there exists
such that if
is large enough,
.
Let us introduce the quantity
Theorem A also implies a control of
(recall ( 4 )). A stronger version (Theorem 3 ) of the following result is going to be proved.
Theorem B. Suppose that
is
. The same conclusion as in Theorem A holds if
is replaced by
.
As claimed above, Theorems A and B indeed yields new information on the multifractal structure of random Gibbs measures.
Section 2 contains new definitions and two propositions that are used in Section 3 and 5 to state and prove stronger versions of Theorems A and B. Section 4 contains the proof of results concerning the speed of convergence of
to
.
We end this introduction by giving an application of Theorem A.
1.3 An application: The Hausdorff dimension of new limsup sets
Let
be a quasi-Bernoulli independent random measure as defined previously and consider
, its projection on
. Examples of jump processes of [5, 9] are
where
is a Lévy process. Basically, if
denotes the countable set of jump points of such a process and
is a sequence decreasing to 0 such that
, the multifractal nature of these processes is closely related to the computation of the Hausdorff dimension of the sets defined for every
,
by
for some sequence
converging to 0. The set
contains the points that are infinitely often close to a jump point
at rate
relatively to
, upon the condition that
. This last condition implies that
has roughly a Hölder exponent
at scale
around
. One of the main results of [5, 10] (see also [6] ) is the computation of the Hausdorff dimension of
. Under a suitable assumption on
, it is proved in [5, 10] that, with probability one, for all
such that
and all
,
|
(6)
|
where
stands for the Hausdorff dimension. This achievement is a non-trivial generalization of what is referred to as “ubiquity” properties of the resonant system
. Ubiquity plays a role for instance in the description of exceptional sets arising in the problem of small denominators and the physical phenomenon of resonance [1, 12] . In the classical result,
is equal to the monofractal Lebesgue measure, so
, the condition
is trivial, and
(see [12] for instance).
The fact that, by Theorem A, the growth speed
behaves like
as
is a crucial issue in constructing a Cantor set of Hausdorff dimension
in
.
2 Definitions, Growth speed in singularity sets
In the sequel,
denotes the probability space on which the random variables of this paper are defined.
2.1 Measure of singularity sets: a neighboring boxes condition
Let
and
be two probability measures with supports equal to
.
With any
can be associated the integer
such that the
-adic subinterval of
naturally encoded by
is
(alternatively
). Then, if
,
stands for
. This defines an integer valued distance on
. This distance yields a notion of neighbors for cylinders of the same generation. This notion coincides with the natural one on
-adic subintervals of the same generation in
.
Let
be a positive sequence,
, and
.
We consider a slight refinement of the sets introduced in ( 3 ): For
,
|
(7)
|
This set contains the points
for which, at each scale
large enough, the
-measures of the
neighbors of
(for the distance
) belong to
. Controlling the mass of these neighbors is necessary in the proof of ( 6 ) when
is a quasi-Bernoulli independent random measure.
For
and
, let us define the quantity
|
(8)
|
Proposition 1.
Let
be a positive sequence.
If
, then
is of full
-measure.
Remark 1.
The same kind of conditions was used in [
3]
to obtain a comparison between the box [
11]
and centered [
28]
multifractal formalisms.
-
Proof.
For
and
, let us define
|
(9)
|
For
, if there exists (a necessarily unique)
such that
, this word
is denoted
. For
, let
with
One clearly has
|
(10)
|
Fix
and
. Let
be the random variable which equals
if
exists, and 0 otherwise.
The Markov inequality applied to
with respect to
yields
. Since
is constant over each cylinder
of generation
, we get
| |
Summing over
yields
. The conclusion follows from ( 10 ) and from the Borel-Cantelli Lemma. □
2.2 Growth speed in families of singularity sets
Let
be a set of indexes, and
a measurable subset of
of probability 1. Some notations and technical assumptions are needed to state the result.
For every
, we consider two sequences of families of measures
and
such that for every
, the elements of the families
and
are probability measures on
. For
,
is written
.
We consider an integer
, and a positive sequence
, as well as a family of positive numbers
. Then, remembering ( 9 ) let us consider for every
and
the sets
|
(11)
|
The sets
form a non-decreasing sequence. One then defines the growth speed of
as the quantity
|
(12)
|
This number, maybe infinite, is a measurement of the number
of generations needed for
to recover a certain given fraction (here chosen equal to 1/2) of the probability measure
. We assume that
is concentrated on
, so that
.
We assume that for every positive sequence
, there exist a random vector
, a sequence
of copies of
, a sequence
such that for
-almost every
,
|
(13)
|
where
is defined in ( 8 ). This provides us with a uniform control over
of the families of measures
.
Proposition 2 (Uniform growth speed in singularity sets).
Let
be a the sequence of positive numbers.
Let
be a sequence of integers such that
. Assume that
|
(14)
|
With probability one, for every
large enough, for every
, one has
.
-
Proof.
Fix
. As shown in Proposition 1 , for every
and every
, one can write
| |
Thus, using ( 13 ), one gets
|
(15)
|
Now ( 14 ) yields
| |
Thus, with probability one,
for every
large enough.
This, combined with ( 11 ), ( 15 ) and ( 12 ), implies that, with probability one, for all
large enough, for every
,
. □
3 Main results
3.1 Examples of quasi-Bernoulli independent measures
It is not difficult to show that, in the setting of [19] , the two following examples can be seen as random Gibbs measures associated with a random Hölder potential in the dynamical system
.
Example 1. Multinomial random measures. Let
be a positive random vector such that
almost surely, and let
be a sequence of independent copies of the vector
. Let
denote the unique measure on
such that
for
.
With probability one, the sequence of measures
defined on
by
|
(16)
|
converges weakly, as
, to a probability measure
which clearly satisfies (P1) to (P4). Here
is constructed like
, but with the vectors
instead of
.
Example 2. Random Riesz products. Let
be a 1-periodic Hölder continuous function on
and let
be a sequence of independent random variables uniformly distributed in
. Let
be the mapping
. Then consider on
the sequence of measures
whose density with respect to
is given by
|
(17)
|
Because of Theorems 3.1 and 3.2 in [19] , with probability one, the sequence
converges weakly to a probability measure
. Moreover, it is shown in [13, 4] that, because of the Hölder regularity and the 1-periodicity of
, properties (P1) to (P3) hold. Property (P4) follows from the fact that the
's are chosen independent. Here
is constructed like
, but with the phases
instead of
.
3.2 Identification of the function
and auxiliary measures
Let
be quasi-Bernoulli independent random measure. We specify the scaling function
and the family of analysing measures discussed in the Introduction.
The function
. For every
, let us define the function
where
denotes the measure associated with
like
is associated with
in formulas ( 16 ) and ( 17 ). When
we simply write
.
The same arguments as those used in [13] and [4] (mainly based on Kingman's sub-multiplicative ergodic theorem) show that, with probability one, for all
and for all
,
converges, as
, to a real number
(thus independent of
).
coincides with the number defined in ( 1 ). Moreover,
is also the limit when
of the sequence
. In particular the mapping
is deterministic.
Due the concavity of
, with probability one,
converges uniformly to
on compact sets.
Auxiliary measures. The multifractal spectrum of
is obtained thanks to the following auxiliary measures
. Let
be a subset of
with
such that the conclusions of Proposition 3 hold for all
.
For every
, for all
and for all
, let
be the probability measure with a density with respect to the measure
on
(for every
) given by
.
If
is still fixed, for every
one can consider a subsequence
such that the sequence
converges weakly to a measure
(which depends on
). This can also be done for the measures
. For every fixed
, for all
and
, a measure
is built as
.
3.3 Main results
In the sequel,
stands for the integer part of the real number
. If the function
is differentiable,
stands for the open interval
.
Theorem 1.
Let
be a quasi-Bernoulli independent random measure, and assume that
is twice continuously differentiable.
Let
a sequence of positive numbers going to 0. Assume that
the series
converges.
With probability one,
, the singularity sets
and
(defined in ( 7 )) are both of full
-measure.
Remark 2.
(1) As soon as
for some
, one has
for all
. The conclusions of Theorem 1 thus hold in this case.
In view of the law of the iterated logarithm (see [
30,
21]
), one could expect
to decrease faster toward 0. This is not the case because we impose the control of neighboring cylinders (in the sense of
) and the uniform control over the parameter
.
(2) In Examples 1 and 2,
is analytic (see [
4]
and references therein).
The next statement uses the definitions introduced in Section 2.2 . The measures
and
play respectively the role of
and
for
.
Theorem 2 (Growth speed in singularity sets).
Under the assumptions of Theorem 1 , let us choose
,
and a sequence
so that
. Let us also fix
.
For every compact subinterval
of
, with probability one, for
large enough and for all
, if
, one has
Remark 3.
Instead of a fixed number of neighbors, it is not difficult to treat the case of an increasing sequence of neighbors
, simultaneously with the speed of convergence
. This number
can then go to
under the condition that
.
Another improvement consists in replacing the fixed fraction
in ( 5 ) by a fraction
going to 1 as
goes to
. The choice
with
is convenient, as the reader can check.
Let us recall that for all integers
and
|
(18)
|
Theorem 3 (Speed of renewal of large deviation spectrum).
Under the assumptions of Theorem 1 , let us choose
for some
. Let
be a compact subinterval of
, and let
.
For every
, with probability one, for
large enough, for all
and for all
, one has
The following Propositions are useful in the sequel.
Proposition 3.
Let
be a compact subset of
, and let us fix
.
1. There exists a constant
such that
2. There exists a constant
such that with probability one
and for
large enough, for every
,
Proposition 4.
Assume that
is differentiable and that
. Let us denote
the word consisting of
consecutive zeros and
the word consisting of
consecutive
.
There are three constants
such that with probability one,
and for
large enough, for every
,
Propositions 3 and 4 are proved in Section 4 , and the theorems in Section 5 .
4 Proofs of Proposition 3 and 4
4.1 Proof of Proposition 3
1. The arguments are standard. For
and
, let us define
. As a consequence of (P1),
| |
| |
Using then (P3) and (P4), and the definition of
, one gets
It follows that the two sequences
and
are sub-additive. Consequently, the sequence
converges, as
, to its infimum denoted by
. Similarly, the sequence
converges to
. This yields that
|
(19)
|
which gives the desired conclusion since we have seen that
.
2. We invoke a property which does hold because of (P2): there exists
such that with probability one,
|
(20)
|
Fix
, a non-trivial compact subinterval of
.
For
,
and
, let us define the random variables
and
. It follows from (P1) and (P4) that
|
(21)
|
Let
, and fix
. For every integer
, we write
where
. Using again (P1) to (P4), one deduces from ( 21 ) that for every
, there exist
independent copies
of
such that
|
(22)
|
We invoke the following concentration inequality (see Lemma 1.5 of [22] )
Lemma 1.
Let
and let
be a sequence of random variable i.i.d. with a centered and bounded random variable
.
For all
,
Let us define the random variables
.
By (P2), one can find a constant
such that
. As a consequence, Lemma 1 can be applied to the bounded family
. Then choosing
yields (remember that
)
For every
, let
be a finite sequence of points of
such that
, and denote by
the set of these points. We can assume that the cardinality of
is less than or equal to
. Then
| |
| |
This implies that for every
and for
large enough,
|
(23)
|
On the other hand, remembering the proof of item 1. and ( 19 ), one has
|
(24)
|
For every
,
can be upper bounded by
| |
| |
With probability one, for
large enough, using respectively ( 22 ) and ( 24 ), this first and the third term are both bounded by a
(which does not depend on
). Using ( 23 ) and remarking that
, one gets
| |
where
is uniform over
. This yields
uniformly for
when
is large enough. The conclusion follows from ( 20 ) and from the construction of the sets
.
Let us show the second inequality of item 2. For every
,
and
, let us consider a sequence
,
, associated with
like
,
, is associated with
. Let
be defined as above, and let us consider the events
One verifies that
. We then deduce from the Borel-Cantelli Lemma that with probability one, for
large enough, if
then
holds. One concludes by using the same estimates as above.
4.2 Proof of Proposition 4
If
, the same kind of arguments as in the proof of Proposition 3 show that, with probability one,
exists, and this number is deterministic. Hence, using ( 25 ), with probability one, for every
, the limit
exists and is equal to
. Since
is a finite measure,
.
Moreover, there exists
such that for
-almost every
, for
large enough, for all
,
| |
| |
So, for
large enough,
and
, one has
| |
and
where
Let us show that
for
and
. Suppose
for some
. Remember that
for all
. Using the concavity of
, the equality
implies that
and then that
, in contradiction with our assumption
.
Finally, since
, the mapping
is continuous, and the conclusion follows from properties (P1) and (P2), the compactness of
, and the form of
.
5 Proofs of Theorems 1 , 2 and 3
Mimicking the approach in [4] and using Proposition 3 and 4 shows that if
is a compact subset of
, there exists a constant
such that, with probability one, for
large enough,
,
,
|
(25)
|
and if
is fixed, for the same constant
, with probability one, for
large enough, for
,
,
,
|
(26)
|
Before starting the proofs, let us make a last useful remark.
Remark 4.
If
and
are words of length
, and if
and
stand for their prefixes of length
, then
implies
. It implies that, given two integers
and two words
and
in
such that
, there are two prefixes
and
of respectively
and
of common length
such that
; moreover, for these words
and
, there are at most
pairs
of words in
such that
and
are respectively the prefixes of
and
.
5.1 Proof of Theorem 1
Fix
a compact subinterval of
and
a bounded positive sequence to be precised later. For
and
, let us introduce the two quantities (recall ( 8 ))
|
(27)
|
Due to Proposition 1 , we seek for a uniform control of
and
on
.
We only consider
, since the study of
is similar.
An upper bound for
: Consider
such that
, as well as two prefixes
and
of respectively
and
of common length
such that
. Let
. If
is large enough, ( 25 ) holds for both
and
. Then, using the construction of
, item 2. of Proposition 3 , (P1), (P2) and ( 25 ), one gets for
large enough
|
(28)
|
where
depends on
,
,
and
. Thus, by Remark 4 , for
large enough,
and
,
| |
| |
for some other constant
depending on
,
,
,
and
.
Let us remark that for every integer
, there are less than
integers
such that
. One thus deduces from the definition ( 27 ) (and ( 8 )) of
and from the above estimate that
|
(29)
|
where
|
(30)
|
|
(31)
|
Let us first upper bound
. ( 25 ) yields for some constant
that
where
. Using item 2. of Proposition 3 , for some constant
Since
is twice continuously differentiable, one has
independently of
(if
is small enough), and
|
(32)
|
In order to estimate
, we use the words
and
defined in Proposition 4 . For every
, a representation of the set of pairs
in
such that
is the following:
|
(33)
|
Let
. The sum
equals
| |
Let us introduce
Using ( 25 ) and property (P1) of
, one obtains another constant
such that
Then, item 2. of Proposition 3 yields (with another
)
| |
| |
| |
| |
By Proposition 4 and ( 25 ), the sum
is uniformly bounded over
and
when
is small enough.
Hence, replacing
by
,
|
(34)
|
Finally, combining ( 29 ), ( 32 ) and ( 34 ) yields
| |
| |
for some
independently of
. By our assumption on
, the choice
with
small enough so that
makes the series
converge for every
. The conclusion concerning the sets
then follows from Proposition 1 .
5.2 Proof of Theorem 2
Fix
. We use twice ( 26 ), with
and
, in order to get a control like ( 28 ). For
large enough, if
and if
and
is a prefix of
such that
, then
and ( 28 ) holds for
and
. Then, from the computations performed in the proof of Theorem 1 and from Proposition 4 , one deduces that for every compact subinterval
of
, there exist
and
such that with probability one, for
large enough, if
, for all
| |
| |
In order to apply Proposition 2 , let us define
,
and
,
,
for every
and for
,
,
for every
,
.
With these parameters the conditions of Proposition 2 are fulfilled. As a consequence, with probability one, for
large enough, for all
,
.
Let us then consider the families
and
instead of the family
and
respectively, and keep the same definitions for the other variables involved in Proposition 2 .
Then the same control as above holds for the growth speed
. Notice that here the vector
is deterministic.
5.3 Proof of Theorem 3
Fix
and
a compact subinterval of
and
. A standard Markov inequality (as in Proposition 1 ) shows that for
,
and
, one has
, where
stands for the sign of
. Then, by Proposition 4 , with probability one, one has for
large enough, for
and for
where
. One remarks that
for
large enough. On the other hand, let
Using Theorem 2 , with probability one, for
large enough, for all
,
. But, looking back at the definition of
, one remarks that
for
, that is
.
References
-
V.I. Arnol'd, Small denominators and problems of stability of motion in classical and celestial mechanics, Ups. Mat. Nauk. 18 (1963), 91–192, English transl. in Russian Math. Surveys 18 (1963), 85-191.
-
J. Barral, Continuity of the multifractal spectrum of a random statistically self-similar measure, J. Theoretic probab. 13 :1027–1060 (2000).
-
J. Barral, F. Ben Nasr, J. Peyrière, Comparing multifractal formalisms: The neighboring boxes conditions, Asian J. Math. 7:149–165 (2003).
-
J. Barral, M.O. Coppens, B.B. and Mandelbrot, Multiperiodic Multifractal Martingale Measures, J. Math. Pures Appl. 82:1555-1589 (2003).
-
J. Barral, S. Seuret, Combining multifractal additive and multiplicative chaos, accepted for publication in Commun. Math. Phys. (2004).
-
J. Barral, S. Seuret, Sums of Dirac masses and conditioned ubiquity, C. R. Acad. Sci. Paris, Ser. I 339: 787–792 (2004).
-
J. Barral, S. Seuret, Growth speed in singularity sets of Mandelbrot cascades, submitted.
-
J. Barral, S. Seuret, The multifractal nature of heterogeneous sums of Dirac masses, submitted.
-
J. Barral, S. Seuret, Lévy processes in multifractal time and fixed points of the smoothing transformation, submitted.
-
J. Barral, S. Seuret, Heterogenuous ubiquitous systems and Hausdorff dimension in
, submitted.
-
G. Brown, G. Michon, and J. Peyrière, On the multifractal analysis of measures, J. Stat. Phys. 66(3-4):775–790 (1992).
-
M.M. Dodson, Exceptional sets in dynamical systems and Diophantine approximation. Rigidity in Dynamics and Geometry (Cambridge, 2000), 77–98, Springer (2002).
-
A.H. Fan, Multifractal analysis of infinite products, J. Stat. Phys. 86(5/6):1313-1336 (1997).
-
D.-J. Feng, E. Olivier, Multifractal analysis of weak Gibbs measures and phase transition—application to some Bernoulli convolutions. Ergodic Theory Dynam. Systems 23:1751–1784 (2003).
-
T.C. Halsey, M.H. Jensen, L.P. Kadanoff, I. Procaccia, B.I. Shraiman, Fractal measures and their singularities: The characterization of strange sets, Phys. Rev. A, 33(2):1141–1151 (1986).
-
Y. Heurteaux, Estimations de la dimension inférieure et de la dimension supérieure des mesures, Ann. Inst. H. Poincaré Probab. Statist. 34:309–338 (1998).
-
S. Jaffard, The multifractal nature of Lévy processes, Probab. Theory Rel. Fields 114: 207–227 (1999).
-
J.-P. Kahane, J. Peyrière, Sur certaines martingales de Benoît Mandelbrot, Adv. Math. 22:131-145 (1976).
-
K. Khanin, Y. Kifer, Thermodynamic formalism for random transformations and statistical mechanics, in Amer. Math. Soc. Transl. (2) 171:107–140 (1996).
-
Y. Kifer, Fractals via random iterated function systems and random geometric constructions. Fractal geometry and stochastics (Finsterbergen, 1994), 145–164, Progr. Probab., 37, Birkhuser, Basel, 1995.
-
Y. Kifer, Limit theorems for random transformations and processes in random environments, Trans. Amer. Math. Soc., 350: 1481–1518 (1998).
-
M. Ledoux, M., Talagrand, Probability in Banach Spaces, Springer-Verlag, New-York.
-
J. Lévy Véhel, R.H. Riedi, TCP traffic is multifractal: a numerical study, INRIA research report, RR-3129 (1997).
-
B.B. Mandelbrot, Intermittent turbulence in self-similar cascades: divergence of hight moments and dimension of the carrier, J. Fluid. Mech. 62:331–358 (1974).
-
B.B. Mandelbrot, Fractals and Scaling in Finance: Discontinuity, Concentration, Risk, Springer (1997).
-
G. Michon, Mesures de Gibbs sur les Cantor Réguliers, Ann. Inst. Henri Poincaré: Phys. Théor. 58:267–285 (1993).
-
J.-F. Muzy, E. Bacry, Multifractal stationary random measures and multifractal random walks with log-infinitely divisible scaling laws, Phys. Rev. E 66: (2002).
-
L. Olsen, A multifractal formalism, Adv. Math. 116:92–195 (1995).
-
Pesin, Y., Weiss, H., The multifractal analysis of Gibbs measures: motivation, mathematical foundation, and examples, Chaos 7(1), 89–106 (1997)
-
W. Philipp, W. Stout, Almost sure invariance principles for partial sums of weakly dependent random variables, Mem. Amer. Math. Soc. 161 (1975).
-
D.A. Rand, The singularity spectrum
for cookie-cutters, Ergod. Th. Dynam. Sys. 9:527–541 (1989).
-
R. Riedi, An improved multifractal formalism and self-similar measures, J. Math. Anal. Appl. 189:462–490 (1995).
-
R. Riedi, Multifractal processes, Long Range Dependence: Theory and Applications, eds. Doukhan, Oppenheim and Taqqu, (Birkhäuser 2002), pp 625-715.
-
G. Zohar, Large deviation formalism for multifractals, Stoch. Proc. Appl. 79:229–242 (1999).
INRIA Rocquencourt, Equipe “Complex”, Domaine de Voluceau Rocquencourt, 78153 Le Chesnay cedex, France