5gang/ftp.transfer/final.book.nofigs.Feb.20…  · Web view5.2.4 Direct empirical estimation of...

93
4/13/22 Chapter 5 1 Words: 26594 (without references), no figs: 22618 5. Cascades, dimensions and codimensions: 5.1 Multifractals and the codimension function 5.1.1 Probabilities and codimensions We have given evidence that the atmosphere is scaling over wide ranges of scale (ch. 1), we have argued that the dynamics are also scaling (ch. 2) and lead to multiplicative cascades (ch. 3), and finally we have given empirical support for this (ch. 4). Throughout our approach was to provide the minimal theoretical framework necessary for understanding the most straightforward data analyses (e.g. the trace moments method, φ λ q ). We mentioned that specification of all of the statistical moments is generally a complete statistical characterization of the process and hence this was equivalent to their specification in terms of probabilities. However, we did not go further to specify the exact relation; the moment characterization was convenient and adequarte for our purposes. We now turn to the complete formalism needed to understand the probability structure of the cascades. In this chapter, we thus continue to study the properties of cascade this time emphasizing their probability distributions and their exponents, the codimensions. 5.1.2 Revisiting the b Model We have already described the monofractal stochastic b model in ch. 3 that is said to be “monofractal” or “unifractal”, because it can be defined with the help of a unique codimension; let us examine its probability structure as the cascade develops. Recall from ch. 3 that it has one parameter c>0 and that two states specify the statistics of the multipliers me: DRAFT 1 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 2

Transcript of 5gang/ftp.transfer/final.book.nofigs.Feb.20…  · Web view5.2.4 Direct empirical estimation of...

Page 1: 5gang/ftp.transfer/final.book.nofigs.Feb.20…  · Web view5.2.4 Direct empirical estimation of c(g): the probability distribution multiple scaling (PDMS) technique. In chapter 3,

5/18/23 Chapter 5 1

Words: 26594 (without references), no figs: 22618

5. Cascades, dimensions and codimensions:

5.1 Multifractals and the codimension function

5.1.1 Probabilities and codimensions

We have given evidence that the atmosphere is scaling over wide ranges of scale (ch. 1), we have argued that the dynamics are also scaling (ch. 2) and lead to multiplicative cascades (ch. 3), and finally we have given empirical support for this (ch. 4). Throughout our approach was to provide the minimal theoretical framework necessary for understanding the most straightforward

data analyses (e.g. the trace moments method, φλq

). We mentioned that specification of all of the statistical moments is generally a complete statistical characterization of the process and hence this was equivalent to their specification in terms of probabilities. However, we did not go further to specify the exact relation; the moment characterization was convenient and adequarte for our purposes. We now turn to the complete formalism needed to understand the probability structure of the cascades. In this chapter, we thus continue to study the properties of cascade this time emphasizing their probability distributions and their exponents, the codimensions.

5.1.2 Revisiting the b Model

We have already described the monofractal stochastic b model in ch. 3 that is said to be “monofractal” or “unifractal”, because it can be defined with the help of a unique codimension; let us examine its probability structure as the cascade develops. Recall from ch. 3 that it has one parameter c>0 and that two states specify the statistics of the multipliers me:

Pr(me =λ0c)=λ0

−c (aλive)

Pr(me =0)=1−λ0−c (dead)

where λ0 is the single step (integer) scale ratio. Recall that the magnitude of the boost me =λ0c

>1 is chosen so that at each cascade step the ensemble averaged e is conserved:

me =1 ⇔ nε = 0ε

Indeed at each step in the cascade the fraction of the alive eddies decreases by the factor b =λ0

−c (hence the name “b model”) and conversely their energy flux density is increased by the

factor 1/b to assure (average) conservation. Rather than follow ch. 3 and consider how the moments change with scale, let us now consider how the probabilities evolve as we increase the

DRAFT

1

1

2

3

4

5678

9101112131415

16

17181920

21

2223

24

25

262728

2

Page 2: 5gang/ftp.transfer/final.book.nofigs.Feb.20…  · Web view5.2.4 Direct empirical estimation of c(g): the probability distribution multiple scaling (PDMS) technique. In chapter 3,

5/18/23 Chapter 5 2

number of cascade steps. After n steps, the effect of the single-step dichotomy of “dead” or “alive” is amplified by the total (n step) scale ratio λ =λ0

n:

Pr en = λ0n( )

c=λc

( )= λ0n( )

−c=λ−c (aλive)

Pr(en =0)=1− λ0n( )

c =1- λ−c (dead)

hence either the density diverges en with an (algebraic) order of singularity c, but with an (algebraically) decreasing probability, or is “calmed” down to zero.

Following the discussion (and definitions) given in section 3.2, c is the codimension of the alive eddies, hence their corresponding dimension D is:

D =d −c

(d is the dimension of the embedding space, equal to 2 in Figures 3.4, 3.9). This is the dimension of the “support” of turbulence, corresponding to the fact that after n steps the average number of alive eddies in the b - model is

Nn =λd Pr eλ =λc( )=λd−c

5.1.2 Revisiting the a Model

In ch. 3, we introduced the a model which more realistically allows eddies to be either “more active” or “less active” according to the following binomial process:

Pr(me =λ0γ+ )=λ0

−c (>1 ⇒ INCREASE)

Pr(me =λ0γ− )=1−λ0

−c (<1 ⇒ DECREASE)

where γ+, γ- correspond to boosts and decreases respectively, the b model being the special case where γ− =−∞ and γ+ =c (due to conservation <me> = 1, there are only two free parameters eq. 3.6):

λ0γ+ −c + λ 0

γ − 1 − λ 0−c( ) = 1

Taking −γ > −∞, the pure orders of singularity −γ and +γ lead to the appearance of mixed orders of singularity, of different orders −γ (γ ≤ γ ≤ +γ ) . These are built up step by step through a complex succession of −γ and +γ , values as illustrated in Figure 3.7b.

What is the behaviour as the number of cascade steps, n → ∞? Consider two steps of the process, the various probabilities and random factors are:

Pr(me =λ02γ+ )=λ0

−2c (two boosts)

Pr(me =λ0γ+ +γ− )=2λ0

−c(1−λ0−c) (one boost and one decrease)

Pr(me =λ02γ− )=(1−λ0

−c)2 (two decreases)

This process has the same probability and amplification factors as a new three-state a model with a new scale ratio of λ0 2 defined as:

DRAFT

3

2930

31

32333435

36

373839

40

41

4243

44

454647

48

4950515253

54

5556

4

Page 3: 5gang/ftp.transfer/final.book.nofigs.Feb.20…  · Web view5.2.4 Direct empirical estimation of c(g): the probability distribution multiple scaling (PDMS) technique. In chapter 3,

5/18/23 Chapter 5 3

Pr(me = λ02( )

γ+ )= λ02( )

−c (one λarγe)

Pr(me = λ02( )

γ+ +γ−( )/2)=2 λ0

2( )−c/2

−2 λ02( )

−c (interm ediate)

Pr(me = λ02( )

γ− )=1−2 λ02( )

−c/2+ λ0

2( )−c (λarγe decrease)

Iterating this procedure, after n = n+ + n- steps we find:

γn+ ,n−=

n+γ + + n−γ −

n+ + n−

, n+ = 1,...,n

Pr με = λ 0n( )

γn+ ,n−

( ) =nn+

⎝⎜⎞

⎠⎟λ 0

n( )−cn+ /n

1− λ 0n( )

−c /n

( )n−

where

nn+

⎛⎝⎜

⎞⎠⎟ is the number of combinations of n objects taken n+ at a time. This implies that

we may write:

Pr eλ0n ≥ λ0

n( )γi

( )= pi, φφ∑ λ0

n( )−ci, φ

The pij’s are the “submultiplicities” (the prefactors in the above), cij are the corresponding

exponents (“subcodimensions”) and λ0n

is the total ratio of scales from the outer scale to the smallest scale. Notice that the requirement that me =1 implies that some of the iγλ are >1 (boosts) and some are <1 (decreases), that is, some γi>0 and some γi<0. Note also that the a model has bounded singularities:

γ− ≤γ≤γ+

so that the important maximum attainable singularity γmax is equal to +γ . The final step in “renormalizing” the cascade is to replace the above n - step, 2-state cascade with ratio λ0 by a single step cascade with ratio λ = λ0

n and with n+1 states. Note that we are not saying that there is absolutely no difference between the n-state a model with ratio λ and the corresponding (n+1)-state model with λ = λ0

n: their properties will however be identical for integral powers of

λ. Finally, doing this and making the replacement λ0n → λ , and the limit λ→ ∞ , one of the terms

in the sum (eq. 11) will dominate (that with the smallest cij). Hence defining:

ci =m in ciφ⎡⎣ ⎤⎦=c γi( )

yields for λ→ ∞ :

Pr eλ ≥λγi( )=piλ−ci

where ci is the codimension and pi is the corresponding multiplicity. If we now drop the subscripts “i” (this allows for the possibility of a continuum of states, e.g., the original process being defined by a uniform or other continuous distribution) then we obtain:

DRAFT

5

57

58

59

6061

62

63

64656667

68

6970717273

7475

76

77

78

798081

6

Page 4: 5gang/ftp.transfer/final.book.nofigs.Feb.20…  · Web view5.2.4 Direct empirical estimation of c(g): the probability distribution multiple scaling (PDMS) technique. In chapter 3,

5/18/23 Chapter 5 4

Pr eλ ≥λγi( )=p γ( )λ−c γ( ) ; dcdγ

> 0

This is a basic multifractal probability relation for cascades. We now simplify this using the “~” sign which absorbs the multiplicative (pi) as well as taking into account the logarithmic number of terms in the sum (which can lead to logarithmic prefactors corresponding to “sub-codimensions”). With this understanding about the equality sign, we may thus write:

Pr eλ ≥λγ( )~λ−c γ( )

Each value of λe corresponds to a singularity of order γ and codimension c(γ); see the schematic, fig. 5.1. Note that since the smallest scale ≈ λ-1, strictly speaking the expression “singularity” applies to γ > 0 ( λe → ∞ for λ → ∞ ), when γ < 0 it is rather a “regularity”. Note that eqs. 11, 14- 16 consider probability distributions of events above a given (scaling) threshold; therefore we consider “exceedence probability distributions”, Pr(eλ≥λγ) rather than the the

standard “cumulative probability distribution function” CDF = Pr eλ < λγ( ) . However both are obviously related by Pr(eλ≥λγ) =1-CDF. Here and throughout his book will always use the term “probability distribution” in the sense of exceedence probability distribution and therefore for events above a given (scaling) threshold.

5.2 The Codimension Multifractal Formalism

5.2.1 Codimension of Singularities c(γ) and its relation to K(q)

In this section we continue our discussion of multifractal fields in term of singularities, but also relate this to its dual representation in terms of statistical moments which we already discussed in ch. 3. Contrary to the popular dimension f(a) formalism which was developped for (low dimensional) deterministic chaos (see section 5.4), we develop a codimension formalism necessary for stochastic processes, it is therefore more general than the dimension formalism.

The measure of the fraction (at resolution λ with corresponding scale l =L / λ ) of the probability space with singularities higher than γ is given by the probability distribution eq. 16. The previous section underlines this new feature; the exponent c(γ) is a function not a unique value. Rather than dealing with just a scaling geometric set of points we are dealing with a scaling function (in the limit λ → ∞ , the density of a measure); from this function we can define an infinite number of sets, e.g., one for each order of singularity γ (fig. 5.1).

Figure 5.1 here

We now derive the basic connection between c(γ) and the moment scaling exponent K(q). To relate the two; write the expression for the moments in terms of the probability density of the singularities:

p γ( ) =

dPrdγ

: c' γ( ) λoγλ( )λ−c γ( ) : λ−c γ( )

DRAFT

7

82

83848586

87

8889909192

93949596

97

98

99100101102103104105106107108109110111112113114115

116

8

Page 5: 5gang/ftp.transfer/final.book.nofigs.Feb.20…  · Web view5.2.4 Direct empirical estimation of c(g): the probability distribution multiple scaling (PDMS) technique. In chapter 3,

5/18/23 Chapter 5 5

(where we have absorbed the c’(γ)logλ factor into the “ : ” symbol since it is slowly varying, subexponential). This yields:

eλq = dPr ελ( )ελ

q∫ ≈ dγλ −c γ( )λ qγ∫where we have used λe = γλ (this is just a change of variables λe for γ , λ is a fixed parameter).

Hence:

eλq = λ K q( ) = eK q( ) log λ ≈

−∞

∫ dγeξf γ( ); ξ = logλ; f γ( ) = qγ − c γ( ); λ >>1

We see that our problem is to obtain an asymptotic expansion of an integral with integrand of the form exp(x f(γ)) where x = logλ is a large parameter and f(γ) = qγ - c(γ). These expansions can be conveniently preformed using the mathematical technique of “steepest descents” e.g. (Bleistein and Handelsman, 1986) which shows that the dominant term in the

expansion for the integral is exp xm ax

γf γ( )( )⎡

⎣⎢⎤⎦⎥ (i.e. the integral is dominated by the singularity γ

which yields the maximum value of the exponent) so that as long as x = logλ >>1 :

K q( )=m axγ

qγ−c γ( )( )

This relation between K(q) and c(γ) is called a “Legendre transform” (Parisi and Frisch, 1985); see fig. 5.2. We can also invert the relation to obtain c(γ) from K(q); just as the inverse Laplace transform used to obtain K(q) from c(γ) is another Laplace transform so the inverse Legendre transform is just another Legendre transform. To show this, consider the twice iterated Legendre transform F(q) of K(q):

F(q) =maxγ{qγ−(maxq'{q'γ−K(q')})} =maxγ;q'{γ(q−q')+ K(q')}

Taking ∂F/∂γ = 0 q = q' so that we see that F(q) = K(q). This shows that a Legendre transform is equal to its inverse, hence we conclude:

c γ( )=max

qqγ−K q( )( )

The γ which for a given q maximizes qγ - c(γ) is γq and is the solution of c’(γq) = q (fig. 5.3). Similarly, the value of q which for given γ maximizes qγ- K(q) is qγ so that:

qγ = ′c γ( )

γq = ′K q( )

This is a one-to-one correspondence between moments and orders of singularities (see figures 5.2 and 5.3). Note that if γ is bounded by γmax (for example in microcanonical cascades, γ≤d; ch. 3 or for the a model; γ ≤ γ+) there is a qmax = c’(γmax) such that for q>qmax, K(q) = qγmax -c(γmax), i.e. K(q) becomes linear in q (see Figure 5.4).

Figure 5.2, 5.3, 5.4 here

DRAFT

9

117118

119

120

121

122

123124125126

127128

129

130131132133134

135

136137

138

139140

141

142143144145146147

10

Page 6: 5gang/ftp.transfer/final.book.nofigs.Feb.20…  · Web view5.2.4 Direct empirical estimation of c(g): the probability distribution multiple scaling (PDMS) technique. In chapter 3,

5/18/23 Chapter 5 6

5.2.2 Properties of codimension functions

We have seen that for each singularity order γ, that c(γ) is the statistical scaling exponent characterizing how its probability changes with scale. The first obvious property is that due to its very definition (eq. 16) c(γ) is an increasing function of γ: c’(γ)>0. Another fundamental property which follows directly from the Legendre relation with K(q), is that c(γ) must be convex: c’’(γ)>0.

Many properties of the codimension function can be illustrated graphically. For example, consider the mean, q =1. First, applying K’(q) = γ (eq. 23) we find K’(1) = γ1 where 1γ is the singularity giving the dominant contribution to the mean (the q = 1 moment). In ch. 3 we have already defined C1 = K’(1), so that this implies C1 = γ1; the Legendre relation thus justifies the name “codimension of the mean” for C1. Also at q = 1 we have K(1) = 0 (due to the scale by scale conservation of the flux) so that from eq. 22, C1 = c(C1) as indicated in Figure 5.5 (this is a

fixed point relation). C1 is thus simultaneously the codimension of the mean of the process and the order of singularity giving the dominant contribution to the mean. Finally, applying c’(γ) = q (eq. 23) we obtain c’(C1) = 1 so that the curve c(γ) is also tangent to the line x = y (the bisectrix). If the process is observed on a space of dimension d, it must satisfy d≥ C1, otherwise, following the above, the mean will be so sparse that the process will (almost surely) be zero everywhere; it will be “degenerate”. We will see that when C1>d that the ensemble mean of the spatial averages (the dressed mean) cannot converge.

Taking ensemble averages we see that <Df> = <e>DxH ≈ λ-H where we have taken <e>= constant and Dx ≈ λ-1, H thus determines the deviation from scale by scale conservation of <Df>, it is the basic fluctuation exponent. At the level of random variables, writing eλ = λγ we have Dfλ = λγ-H so that Dfλ has the statistics of a bare cascade process with a translation of singularities by -H (see figure 5.6). This generalizes the Kolmogorov relation (f = v, H = 1/3) although we shall see that such simple change of cascade normalization (γ->γ-H) is a poor model of the observable f fields; which are best modelled by fractional integration of order H, see section 5.5. Finally, since c(γ) is convex with fixed point C1, it is possible (see Figure 5.7) to define the degree of multifractality (a) by the (local) rate of change of slope at C1, (the singularity corresponding to the mean) its radius of curvature Rc(C1) is:

Rc C1( )=1+ ′c C1( )( )

3/2

′′c C1( )

Using the general relation c’(C1)=1 we obtain Rc C1( )=23/2 / ′′c C1( ) hence we can locally (near the mean γ = C1) define a curvature parameter a from either of the equivalent relations:

a =23/2 Rc C1( )

C1

=1

C1 ′′c C1( )

These local (q = 1, γ = C1) definitions of a are equivalent to the definition via moments a = K’’(1)/K’(1) (ch. 3). For the universal multifractals (ch. 3), this becomes global (i.e., is sufficient to describe the entire c(γ) function), and we find an upper bound (maximum degree of multifractality) a =2 (a parabola, eq. 3.40). The a =0 case is the monofractal extreme b - model whose singularities all have the same fractal dimension.

DRAFT

11

148

149150151152153154155156157158159

160161162163164165166167168169170171172173174175176

177

178179

180

181182183184185

12

Page 7: 5gang/ftp.transfer/final.book.nofigs.Feb.20…  · Web view5.2.4 Direct empirical estimation of c(g): the probability distribution multiple scaling (PDMS) technique. In chapter 3,

5/18/23 Chapter 5 7

Figures 5.5, 5.6, 5.7 here

5.2.3 The sampling dimension, sampling singularity and second-order

multifractal phase transitions

The statistical properties we have discussed up until now are valid for (infinite) statistical ensembles. However, real world data are always finite so that sufficiently rare events will always be missed. In order to understand the effect of finite sample sizes, consider a collection of Ns samples each d dimensional and spanning a range of scales λ = L/l (= largest / smallest) so that for example there are λd pixels from each sample, (fig. 5.8) and introduce the “sampling dimension” Ds:

Ds =λoγNs

λoγλ

as well as the singularity γs corresponding to the largest (and rarest) value eλ,s the "sampling singularity":

γs =log ελ ,s

log λ

The relation between Ns and eλ,s is thus equivalent to the relation between their base λ logarithms i.e. between Ds and γs (eqs. 26, 27).

In order to relate γs and Ds, consider a collection of satellite images (d = 2). Our question is thus: what is the rarest event with the most extreme γs that we may expect to see on a single picture? On a large enough collection of pictures? The answer to, these questions is straightforward: there are a total of λd+Ds pixels in the sample; hence the rarest event has a probability ≈ λ-(d+Ds) (see fig. 5.8). However the probability of finding γs is simply λ-c(γs) so that we obtain the following implicit equation for γs:

c γs( )=d + D s =Ds

Ds = d+Ds is the corresponding (overall) effective dimension of our sample. More extreme singularities would have codimensions greater than this effective dimension c > Ds and are almost surely not present in our sample.

Figure 5.9, 5.10 here

As an example, we can use the aircraft data shown in fig. 3.1 to estimate the largest singularity that we should expect over a transect 213 points long. We saw that the largest normalized flux value was ≈26.5 which corresponds to an order of singularity of γs = loge/logλ = log(26.5)/log(213) = 0.364. Using the estimated multifractal parameters a = 1.8, C1= 0.06 (these are mean C1, a values for 24 flight legs 1120 km long), we find that in a d =1 section that the solution of c(γs) = 1.364 is γs = 0.396 which is very close to the observed maximum.

DRAFT

13

186187

188

189

190191192193194195

196

197198

199

200201202203204205206207

208

209210211212213

214215216217218219220

14

Page 8: 5gang/ftp.transfer/final.book.nofigs.Feb.20…  · Web view5.2.4 Direct empirical estimation of c(g): the probability distribution multiple scaling (PDMS) technique. In chapter 3,

5/18/23 Chapter 5 8

Let us now calculate the moment exponent Ks(q) for a process with Ns realizations. To do this, we calculate the Legendre transform of c(γ) but with the restriction γ≤γs; this is the same type of restriction as discussed earlier (fig. 5.4, take γmax = γs):

K s (q) =γs(q −qs)+ K(qs) , q ≥qs

K s (q) =K(q) , q ≤qs

hence at q = qs there is a jump/discontinuity in the second derivative of K:

D ′′K s = − ′′K (qs )

due to the existence of formal analogies between multifractal processes and classical thermodynamics, this is termed a “second order multifractal phase transition” (Szépfalusy et al., 1987), (Schertzer and Lovejoy, 1992).

5.2.4 Direct empirical estimation of c(γ): the probability distribution

multiple scaling (PDMS) technique

In chapter 3, we saw how to empirically verify the cascade structure and characterize the statistics using the moments, and how to determine their scaling exponent, K(q). In this chapter, we saw how – via a Legendre transform of K(q) - this information can be used to estimate c(γ). However, it is of interest to be able to estimate c(γ) directly; to do this, we start from the fundamental defining eq. 16, take logs of both sides and rewrite it as follows:

Log(Prλ(eλ > λγ)=−c(γ)Loγ(λ)+ o(1 / Loγ(λ))O(γ)

o(1/Log(λ ))O(γ ) corresponds to the logarithm of the slowly varying factors that are hidden in the “≈” sign in eq. 16 and the subscript “λ” on the probability has been added to underline the resolution dependence of the cumulative histograms. For each order of singularity γ, this equation expresses the linearity of log probability with the log of the resolution. The singularity itself must be estimated from the fluxes by:

γ=log ελ( )

log λ

We now see that things are a little less straightforward than when estimating K(q). First, the term “

o(1/Log(λ ))O(γ ) may not be so negligible, in particular for moderate λ’s, so that using

the simple approximation c γ( )≈−λoγPrλ / λoγλ may not be sufficiently accurate. Second, in eq. 32, we assumed that eλ is normalized such that <eλ> =1; if it is not, it is can be normalized by dividing by the ensemble mean: eλ->eλ/<eλ>. However from small samples, there may be factors of the order 2 in uncertainty over this so that even the estimate of γ may involve some uncertainty. In comparison, if one wants to estimate K(q), one needn’t worry about either of

these issues since (even for the un-normalized eλ) the linear relation log eλq =K q( )λoγλ is

exact (at least in the framework of the pure multiplicative cascades): K(q) is simply the slope of

the log eλq

versus logλ graph (and if the normalization is accurate, the outer scale itself can be

DRAFT

15

221222223

224

225

226

227

228229230

231

232

233234235236237

238

239240241242243

244

245246

247248249250251

252253

254

16

Daniel Schertzer, 29/11/10,
I do not remember what they said, but it was extremely, weak, better to chek it !
Page 9: 5gang/ftp.transfer/final.book.nofigs.Feb.20…  · Web view5.2.4 Direct empirical estimation of c(g): the probability distribution multiple scaling (PDMS) technique. In chapter 3,

5/18/23 Chapter 5 9

estimated from the points where the lines cross, see the examples in ch. 4). The relative simplicity of the moment method explains why in practice it is the most commonly used. c(γ) can then be estimated from K(q) by Legendre transform (either numerically or using a universal multifractal parametrization).

In order to exploit eqs. 31, 32 to directly estimate c(γ), we may thus calculate a series of histograms Prλ of the lower and lower resolution fluxes (obtained from eλ by the usual degrading / “coarse-graining”/ averaging) and use the transformation of variables eq. 32 to write the histograms in terms of γ rather than eλ. There are then two variants on this “probability distribution multiple scale” (PDMS) method (Lavallée et al., 1991). The first and simplest is

simply to ignore the typically small “slow” term in eq. 31 and simply use c γ( )≈−λoγPrλ / λoγλ ; the second performs (for each of a series of standard γ values) regressions of logPrλ versus log λ and then estimates c(γ) as the (negative) slope.

In fig. 5.10 a, b we show some examples of the first method when applied to the aircraft data discussed in ch. 3; 24 legs, each of 4000 points were used with a resolution of 280 m (i.e. a total of 24 x 4000 = 9.6 x 104 data points for each field), the absolute differences at 280 m resolution were analyzed. In ch. 3 we already considered the temperature flux and showed graphically that the transformation from eλ’s into γ’s did indeed result in an apparently stable dynamic range of the γ’s (in comparison, the range of the degraded eλ’s kept diminishing as λ decreased; see fig. 3.2a,b). Figure 5.10a, b simply shows the logλ-normalized log probability of the histograms at each resolution. The resolution was degraded systematically by factors of 2; only the first 7 iterations (i.e. a degradation of resolution factor of 128 corresponding to ≈ 36 km) are shown (the corresponding moments are shown in fig. 4.6a,b). From the relatively tight bunching of the curves, we see that the method is reasonably successful: we have effectively removed almost all of the scale dependency from the histograms. Although the curves tend to diverge somewhat at large, (and rare) values of γ (where the sample size starts to be inadequate), the basic c(γ) shape is discernible. To be a bit more quantitative, we can use the estimate of the sampling dimension Ds = logNs/logλ = log24/log4000 ≈ 0.38 to determine that the estimates are reliable only up to c = d+Ds =1.38. Also shown in the figures are straight reference lines since we shall see in the next section that one generally expects high order moments to diverge implying asymptotically linear c(γ)’s. Along with both the transverse and longitudinal wind components, and the thermodynamic fields we have also included the essentially aircraft measurement/trajectory specific fields z, p: the fluxes derived from aircraft altitude and pressure (recall we used absolute differences). These both show extreme tails (c.f. the reference slopes ≈ 3), and this in spite of the autopilot attempting to enforce an isobaric trajectory! In fact, as discussed in ch. 6 it seems that due to the aircraft response to wind turbulence, the trajectories only begin to be effectively isobaric for scales around 40 km and greater (i.e. just the resolution of the curves is shown in the figures).

Finally, we show a similar analysis of hourly raingauge data from Nime France (1972- 1975, ≈35,000 points). Again we see a good collapse to a unique c(γ), and this time there is evidence for an asymptotic linear behaviour with slope qD ≈3 (see section 5.4).

Figure 5.11 a, b, c here

DRAFT

17

255256257258

259260261262263

264265266

267268269270271272273274275276277278279280281282283284285286287288289290291292293294295296

18

Page 10: 5gang/ftp.transfer/final.book.nofigs.Feb.20…  · Web view5.2.4 Direct empirical estimation of c(g): the probability distribution multiple scaling (PDMS) technique. In chapter 3,

5/18/23 Chapter 5 10

5.2.5 Codimensions of Universal multifractals, cascades

When discussing the moment characterization of the cascades, we have already noted that the two parameters C1, a are of fundamental significance. C1 characterizes the order and codimension of the mean singularities of the corresponding conservative flux, it is the local trend of the normalized K(q) near the mean; K(q) = C1 (q-1) is the best monofractal “b model approximation near the mean” (q ≈ 1). Finally, a = K’’(1)/K’(1) characterizes the curvature near the mean. The curvature parameter a can also be defined directly from the probability exponent c(γ) by using the local radius of curvature Rc(C1) of c(γ) at the point γ = C1, i.e. the corresponding singularity (eq. 25). Finally, for the observed field f, there is a third exponent H which characterizes the deviation from conservation of the mean fluctuation Df ≈ <e>DxH ≈ DxH

since <e> =constant, it is a “fluctuation” exponent.

In ch. 3 we discussed how the universal attractors of additive processes can be used to deduce those of the multiplicative processes by studying their “generators”, Gλ = logeλ . Since multiplying fields eλ is equivalent to adding generators Gλ (for a fixed scale ratio λ ); the generators which are stable and attractive under addition are the Lévy generators discussed in ch.3 and to which we return in section 5.5. The moment exponent K(q), (see eq. 3.40) is given by:

K q( )=C1

a −1qa −q( ); a ≠1

K q( )=C1qLoγ q( ); a =1

The top formula is valid for 0≤a≤2; however as discussed in section 5.5, K diverges for all q <0 except in the special (“log-normal”) case a = 2. To obtain the corresponding c(γ), one can simply take the Legendre transformation (eq. 22) to obtain:

c γ( )=C1

γC1 ′a

+1a

⎛⎝⎜

⎞⎠⎟

′a

; a ≠1; 1 / ′a +1 / a =1

c γ( )=C1eγC1

−1⎛⎝⎜

⎞⎠⎟; a =1

where a’ is the auxiliary variable defined above (Schertzer and Lovejoy, 1987). Figs. 3.15, 5.11 show respectively the universal K(q)/C1 and c(γ) /C1 curves. Note that since a’ changes sign at a =1, for a<1, there is a maximum order of singularity γmax = C1/(1-a) so that the cascade singularities are “bounded”, whereas for a>1, there is on the contrary a minimum order γmin = -C1/(a-1) below which the prefactors in eq. 16 dominate (c(γ) = 0 for γ<γmin) but the singularities are unbounded.

If we take H = 0 (for simplicity) and use eq. 35 to express the exceedence probability

distribution of e, using γ = logeλ/logλ, we find for eλ >> λC1 / α −1( )

that

Pr(ε λ '≥ ε λ ) =

DRAFT

19

297

298299300301302303304305306307

308309310311312313

314

315

316317318

319

320321322323324325326327

328

20

, 19/01/11,
I have suppressed the H shift since it is really for non normalized cascades, not fractionally integrated ones…
Daniel Schertzer, 29/11/10,
there is old sign bug on the r.h.s. !!!
, 19/04/11,
Add: to obtain the alpha=1 case, just take limit as alpha->1 of eq. 32.
Page 11: 5gang/ftp.transfer/final.book.nofigs.Feb.20…  · Web view5.2.4 Direct empirical estimation of c(g): the probability distribution multiple scaling (PDMS) technique. In chapter 3,

5/18/23 Chapter 5 11

p eλ( )≈exp − λoγeλ / K( )′a

( ) with constant

K = α '(C1Logλ )1/α; this is a “stretched exponential”

distribution (also known as a “Weibull” distribution) for the logarithm of e We stress this because eclectic semi-empirical approaches to scaling in turbulence, climate and elsewhere

have claimed the existence of the multiscaling of the moments ( eλq ≈ λ K q( )

) while simultaneously that the distribution for eλ (not for logeλ) is a stretched exponential. Since over limited ranges of eλ, it may be difficult to empirically distinguish the two behaviours (i.e. stretched exponential for eλ or for logeλ) theoretical consistency (and clarity!) is particularly important.

Figs. 5.12 show one-dimensional realizations using the continuous cascades described in section 5.5!Unexpected End of Formula. In fig 5.12 a, we fix C1 =0.1 (a typical value for atmospheric fields, table 4.8) and we demonstrate the effect of increasing a with the same random seed so that the change in the structures with a is apparent. For low a, the series are dominated by “holes” of very low values while for high a (near the maximum, a = 2), it is rather the large singularities that dominate. Fig. 5.13 shows two-dimensional realizations with the same “Lévy hole” phenomenology. In fig. 5.12b, c we show the huge realization to realization variability by showing 10 realizations with different seeds for a = 0.2, 1.9 respectively. Note that due to severe numerical underflow problems, it is difficult to make simulations with a less than 0.2.

Figure 5.11 here

Before leaving the topic of universal multifractals, we should mention another, weak form of universality that was proposed by (Dubrulle, 1994) and (She and Levèsque, 1994); “Log-Poisson” cascades. Recall that starting with 2-state binomial processes, one can take limits which lead to either Gaussian or to Poisson processes. The corresponding binomial cascade is the a model, and in (Schertzer et al., 1991) it was shown how to obtain as a limiting case, the a =2 (“Log-Gaussian” universal cascade. By taking a different limit of the a model, one obtains “Log-Poisson” cascades which have the following form:

γ+ =c 1 − λ 0−γ−( )

K q( )=qγ+ −c+ 1−γ+

c⎛⎝⎜

⎞⎠⎟q

c

c γ( )=c 1−γ+ −γcγ−

1−λoγγ+ −γcγ−

⎛⎝⎜

⎞⎠⎟

⎛⎝⎜

⎞⎠⎟; γ ≤γ+

c γ( )=∞; γ > γ+

where λ0>1 is the cascade ratio for a single step, c>0, γ+>0 and the relation between γ+ and γ-

(top line) is from the conservation requirement of the a model (eq. 7; see (Schertzer et al., 1995)). Clearly, γ+ is the highest order singularity and c is the corresponding codimension so that the log-Poisson cascade has instrinsically a maximum singularity that it can produce. The log-Poisson cascade shares with the Levy generator universal multifractals the possibility of

DRAFT

21

329330331

332333334335336337338339340341342343344345346347348349350351352353354355356

357

358

359360361362363364365

22

Daniel Schertzer, 29/11/10,
it will be againg misleading !
Page 12: 5gang/ftp.transfer/final.book.nofigs.Feb.20…  · Web view5.2.4 Direct empirical estimation of c(g): the probability distribution multiple scaling (PDMS) technique. In chapter 3,

5/18/23 Chapter 5 12

“densififying” the cascade - i.e. it can be made continuous in scale (“infinitely divisible”; section 5.6, take the limit λ0->1), so that it could be said to have “weak” universality properties. However, the Levy generator cascades, could be termed “strongly” universal since the generator is stable and attractive as well.

The limitations of the log-Poisson cascade can be seen as soon as one considers applications. For example, in hydrodynamic turbulence, (She and Levèsque, 1994) assumed non fractal, d =1 filament-like structures for the highest order singularity along with homogeneous eddy turn over times, this leads to the parameter choice c = 2, γ+

= 2/3 (implying λ1

γ- = 3/2). In Magneto-Hydrodynamic turbulence one can argue that extreme events occur on current sheets and select c =1, γ+ = 1/2 (Grauer et al., 1994). In both cases, the use of a model with maximum order of singularity seems difficult to justify, as do the particular parameter values. In the atmosphere, (Deidda, 2000) and (Onof and Arnbjerg-Nielsen, 2009) have used the log-Poisson model for modelling rainfall, where they found - perhaps unsurprisingly - that the parameters c, γ+ vary from realization to realization, exactly as predicted by models with unbounded singularities (such as the strongly Levy generator universal cascades with a>1 where on any realization, the maximum singularity is itself simply a random variable rather than a fixed parameter).

Figures 5.12 a, b, c , 5.13 here

5.3 Divergence of Statistical Moments and extremes

5.3.1 Dressed and bare moments

We now consider the effect of spatially integrating a cascade and then taking the limit λ → ∞ . This leads to the fundamental difference between the “bare” and “dressed” cascade properties; the former have all moments finite (since by definition, for bare quantities, λ is finite) whereas the latter will generally have divergence for all moments greater than a critical value qD which depends on the dimension of space over which the process is integrated, see fig.

5.14 for a schematic.

Figure 5.14 here

In order to define the dressed flux, start by defining the L resolution flux PL A( ) over the

set A:

PL A( ) = εΛd D xA∫

We can now define the “partially dressed” flux density eλ,Λ(d ) as:

eλ,Λ(d ) =ΠΛ Bλ( )vol Bλ( )

where vol(Bλ) = λ-D is the D-dimensional volume of a ball (interval, square, cube etc.) of size L/λ and the “(fully) dressed flux density” as:

DRAFT

23

366367368369370371372373374375376377378379380381382383384

385

386

387388389390391

392393394

395396

397

398

399

400401

24

Daniel Schertzer, 29/11/10,
it seems rather a cut and paste from the Lecture Notes (?)
Page 13: 5gang/ftp.transfer/final.book.nofigs.Feb.20…  · Web view5.2.4 Direct empirical estimation of c(g): the probability distribution multiple scaling (PDMS) technique. In chapter 3,

5/18/23 Chapter 5 13

eλ,(d ) = limΛ→∞

ελ ,Λ(d )

The terms “bare” and “dressed” are borrowed from renormalization jargon and are justified because the “bare” quantities neglect the small scale interactions (<L/λ), whereas the “dressed” quantities take them into account.

Now we can use the factorization property of the cascade (see fig. 5.15): the independence of the large and small scale multiplicative factors:

eL =eλTλ εΛ/λ( )

where the operator Tλ increases scales “zooms” by a factor λ. This equation should be understood in the following way: to obtain a fine scale cascade (resolution L) we may take a lower resolution (λ) cascade and multiply each of the λ resolution boxes (balls) by independent cascade processes each developed over a range of scales L/λ and reduced in size by factors of λ. This leads to:

eL,λ d( ) = ελεΛ/λ h( )

where eλ is the usual bare density (accounting for variability at scales larger than the observation scale) and the density eL / λ (h) (accounting for variability at scales smaller than the observation scale) can be said to be “hidden” since it corresponds to the scales which we average over (see appendix 5A for more details). We shall see shortly that the small scale activity is not necessarily smoothed out!

Figure 5.15 here

5.3.2 The Divergence of the dressed moments and first order

multifractal phase transtions*

From the above, we see that the bare and dressed densities differ only by the “hidden” factor:

e∞ h( ) = limΛ→∞

εΛ/λ h( ) = Π∞ B1( )

i.e. e∞ h( ) is a fully developed, fully integrated cascade, in appendix 5A we derive the following

explicit relation for it. Hence from eqs. 39, 41:

eλ d( ) = ελΠ∞ B1( )

and taking qth moments:

eλ d( )q = ελ

q Π∞ B1( )q

Since for any q, finite λ, eλq = λ K q( )

is always finite, the finiteness of eλ d( )

q

depends on P∞ B1( )

q

. Using “trace moments” as explained in appendix 5A, we obtain the result:

DRAFT

25

402

403404405406407

408

409410411412413

414

415416417418419420

421

422

423424

425

426427

428

429

430

431

432

26

, 19/01/11,
asterix means advanced topic… or do we move this to an appendix… or simply refer to the result??
Page 14: 5gang/ftp.transfer/final.book.nofigs.Feb.20…  · Web view5.2.4 Direct empirical estimation of c(g): the probability distribution multiple scaling (PDMS) technique. In chapter 3,

5/18/23 Chapter 5 14

P∞ B1( )q ≈

O 1( ); q < qD

∞; q ≥ qD

where qD is the solution to the implicit equation:

C qD( )=D

and C(q) is the dual codimension function (eq. 3.20). This leads to:

eλ d( )q =

~ λ K q( ); q < qD

→ ∞; q ≥ qD

⎧⎨⎩⎪

where for q <qD we have absorbed the proportionality constant B1( )∞

q∏ into the “~” sign. The overall result is that the dressed and bare densities have identical scaling exponents for q <qD and only differ for q >qD. The case q = qD requires some care (see Lavallée et al. 1992); we obtain divergence. We can thus introduce the dressed exponent scaling function:

Kd q( )=K(q); q < qD

∞; q ≥qD

⎧⎨⎩The dressed and bare K(q) are therefore the same except for the extreme fluctuations

which will be much more pronounced for the dressed quantities. To calculate the corresponding dressed codimension cd(γ), we can use the Legendre transform of Kd(q) to obtain:

cd (γ)=c(γ), γ≤γD

cd (γ)=qD (γ−γD )+ c(γD ), γ > γD

where γD = K’(qD) is the critical singularity corresponding to the critical qD. This transition from convex “bare” behaviour to linear “dressed” behaviour represents a discontinuity in the second derivative of c(γ); hence a “second order multifractal phase transition” for c (for K, see below).

We now note that the microcanonical constraint γ<D precludes the possibility of dressed microcanonical cascades having a divergence of moments when averaged over sets of dimension D. To see this, recall that K(q) = C(q)(q-1) where C(q) is the monotonically increasing dual codimension function (eq. 3.28); i.e. C’(q)>0 and C(qD) = D, (eq. 46). By differentiating K(q) we thus obtain γD = K’(qD) = C’(qD) (qD-1) +D, hence (since qD >1), we have γD>D. This shows that microcanonical models - which must have γmax≤ D - cannot display divergence of moments – at least not on the spaces over which the microcanonical constraint is imposed.

A linear cd(γ) implies a power law tail on the probability distributions; this is just another way of obtaining the fundamental equivalence between divergence of moments q≥qD and "hyperbolic" or "fat tailed" behaviour of the probability distribution (for large enough thresholds s):

⟨eλ(d )q ⟩= ∞, q ≥ qD ⇔ Pr(ελ (d ) > s) ~ s−qD , s >> 1

DRAFT

27

433

434

435

436

437

438439440441

442

443444445

446

447448449450451452453454455456457458459460

461

28

Page 15: 5gang/ftp.transfer/final.book.nofigs.Feb.20…  · Web view5.2.4 Direct empirical estimation of c(g): the probability distribution multiple scaling (PDMS) technique. In chapter 3,

5/18/23 Chapter 5 15

In order to observe the algebraic probability tail however, the sample size must be sufficiently large. Let us consider the effect of varying Ds = logNs/logλ. Following the argument for eq. 28 section 5.2.3, the maximum observable dressed singularity γd,s is given by the solution

of cd (γd,s)=d + D s and by taking the Legendre transform of cd(γ) with the restriction γd < γd,s (see fig. 5.16 and 5.17) we obtain the finite sample dressed Kd,s(q):

Kd ,s (q) =γd,s(q −qD )+ K(qD ) ; q > qD

Kd ,s (q) =K(q) ; q < qD

In the limit NS → ∞ , γd ,s → ∞ , and for q>qD, Kd ,s (q)→ Kd (q)=∞ as expected. This transition

corresponds to a jump in the first derivative of the K(q):

D ′K (qD ) ≡ ′Kd ,s (qD ) − ′K (qD ) = γ d ,s − γ D =d + Ds − c(γ D )

qD

hence, this is a “first order multifractal phase transition”.

Figure 5.16, 5.17 here

5.3.3 The Physical Significance of the Divergence of Moments: the

multifractal butterfly effect, self-organized criticality

In the previous subsection we noted the equivalence between the divergence of moments eλ

q → ∞ (q>qD) and "hyperbolic" or "fat tailed" behaviour of the probability distribution Pr(eλ(d ) > s)~s−qD (eq. 50) i.e. for large enough thresholds s, an algebraic fall-off of the probability distribution. In real cascades, viscosity cuts off the cascade process at the viscous scale so that real observables are only partially dressed (section 5.3.1). This implies that the graphs of logPr versus logs will only be linear over a finite interval; for sufficiently large s, they will be truncated. However, it must not be concluded that the divergence of moments is academic - on the contrary, it has a profound physical significance. To see this, denote the inner scale l, the observing scale L, and the ratio L = L/l>>1. In this case, our previous analysis for dressed moments will be valid, and if we estimate dressed moments with q>qD (i.e. C(q)>D), then the result will be ~ L(q−1)(C(q)−D ) (appendix 5A), which can be very large and whose value will depend crucially on the small scale details, i.e., the exact value of l, etc. On the contrary, when q<qD (i.e. C(q)<D), the dressed moments will be insensitive to L and the small scale. Overall we can say that for q<qD the moments are macroscopically determined whereas for q>qD

the moments will be microscopically determined. Because of this dependence on the small scale details, the divergence of moments is a kind of “multifractal butterfly effect” similar to the “butterfly effect” in deterministic chaos.

It is worth noting that the power law probability distributions (of dressed quantities) are a basic feature of Self Organized Criticality (Bak et al., 1987). This divergent “hard” statistical behaviour indeed corresponds to the fact that rare and catastrophic events will make dominant contributions. For example, the "mean" shape of a sand pile (the prototypical example of self-

DRAFT

29

462463464

465466

467

468

469470

471

472473474

475

476

477

478

479480481482483484485486487488489490491492493494495496497

30

Page 16: 5gang/ftp.transfer/final.book.nofigs.Feb.20…  · Web view5.2.4 Direct empirical estimation of c(g): the probability distribution multiple scaling (PDMS) technique. In chapter 3,

5/18/23 Chapter 5 16

organized critical system, see fig. 5.18 for a schematic) is maintained due to avalanches of all sizes. In cascades, the flux energy is maintained by instabilities of all intensities. We have also observed that the transition from soft to hard processes (from “thin” to “fat” tailed probabilities) corresponds to the fact that microscopic activity cannot be removed to yield a purely macroscopic description of the system. Similarly, we will not be able to understand the "mean" field without understanding its extremes.

The sandpile exemplifies “classical SOC”; it produces avalanches of all sizes, and the structure of each avalanche is fractal. However, it is only produced in the “zero flux” limit, in other words, one must add grains of sand so slowly that any avalanches thus induced have the time to stop: if grains are added at a constant rate, then the scaling will be broken. If SOC is defined as a dynamical system with both fractal structures and power law probabilities, then we could consider cascade processes to be nonclassical SOC models. While potentially having many of the same statistics, cascades have the advantage that they are generated through a more realistic quasi constant flux boundary condition.

Another way to generate power law probability tails is to use Correlated Additive and Multiplicative (CAM) processes (Sardeshmukh and Sura, 2009). These processes are modelled by systems of coupled (linear) stochastic differential equations and have been used as models of atmospheric variability at various scales, the “Stochastic Linear Forcing” (SLF) approach (see appendix 10C for a detailed comparison of the SLF and scaling approaches). The mechanism is apparently quite close to additive (“dressing”) on multiplicative cascade processes.

Finally, we could mention yet another related mechanism yielding power law probability tails: compound multifractal Poisson processes (Lovejoy and Schertzer, 2006b). This was developed as a model of individual rain drop size distributions in space; it leads to a direct link between the H parameter of the controlling large scale turbulent fluxes and the individual drop probability exponent qD.

Figure 5.18 here

5.3.4 Empirical tests of divergence of moments

When a ≥1 we have “unconditionally hard multifractal processes”; they are unconditionally hard since C(q) increases without limit, hence for any observing space of dimension D, there will be a finite qD satisfying C(qD) = D; this is a consequence of the fact that there is no upper bound on the orders of singularity. Conversely, when a <1, we find that the maximum order of singularity = max(C(q)) = C1/(1-a) so that if the “dressing” (averaging/integrating) space has D> C1/(1-a) there will be no solution C(qD) = D to the equation and all the moments will converge. Since for small enough D the moments will still diverge, the a<1 multifractals are called “conditionally hard”. In ch. 4, we saw that the value of a is commonly in the range 1.5 - 2, indeed, empirically, it is almost always found to be >1, so that there are theoretical and empirical reasons to commonly expect power law probability tails.

In order to empirically test for the divergence of moments, we could look either directly at the probability distributions at a specific resolution, or look for a linear asymptote on a c(γ) plot; we have already seen (fig. 5.10a, b) some evidence for qD ≈3, 5 for various aircraft fluxes; fig. 5.10c for rainrates (qD ≈ 3); the c(γ) from these figures essentially uses information from a range of scales (the “bunching” of the curves). From the figure, we can already sense a recurring practical problem: that without knowledge of the “effective dimension of dressing” the theory

DRAFT

31

498499500501502503504505506507508509510511512513514515516517518519520521522523524

525

526527528529530531532533534535536537538539540541

32

Page 17: 5gang/ftp.transfer/final.book.nofigs.Feb.20…  · Web view5.2.4 Direct empirical estimation of c(g): the probability distribution multiple scaling (PDMS) technique. In chapter 3,

5/18/23 Chapter 5 17

gives no guidance as to at what point in c(γ) to expect the linear asymptote to begin nor the value of the asymptotic log – log slope -qD. It may well be at such low probability levels (such high γ) that enormous sample sizes are necessary to observe it. Furthermore, most sensors have difficulty measuring the amplitudes of very rare but extreme events; they are frequently truncated, so that the critical extreme power law region may be corrupted. This is especially a problem for rain gauges which tend to saturate at high rain rates. Finally, the sampling properties of such algebraic “fat tailed” probabilities are quite nontrivial and can even be sensitive to the way the probability densities are estimated from the empirical frequencies of occurrence.

Alternatively, at a single resolution, we can look for asymptotically linear graphs of log Pr versus log s. Although this is the approach used in most of the literature, including the examples given below, recent work by (Clauset et al., 2009) shows that the maximum likelihood estimator is superior, although the problem of determining the region where the power law tail begins is still nontrivial.

Figures 5.20 a, b, c show the predicted power law behaviour for velocity differences (qD

≈ 5 - 7.5 vertical, horizontal, time, see table 5.1 for an overall intercomparison). Fig 5.21 displays the same behaviour for the potential temperature differences (qD ≈ 3.3). Similarly, figure 5.21 shows evidence for qD = 5 in nondimensional raindrop distributions, and fig. 5.22 from reanalysis anomalies (geopotential and vorticity; see (Sardeshmukh and Sura, 2009) for this and for corresponding simulations using CAM noise). These graphs have been shown either for their particular high quality or for their scientific interest, in table 5.1 we give a more complete summary of the values from the literature. We see that although the values of qD

(notably for the wind) apparently depend on the direction (horizontal, vertical, time; presumably a consequence of anisotropy), that the usual variables of state typically have values in the range 5 – 7.5, with rainrate, potential temperature, geopotential height and vorticity a bit smaller (≈3). Since we also have C1 and a estimates for these fields we could determine the effective dimension of dressing from C(qD) = D using the empirical C1, a, qD to determine C(qD). When this is done, we find – due to the typically small values of C1 ( ≈ 0.05 – 0.1) that C(qD) <1 implying that the effective dressing dimension D<1. It is not clear whether this low “effective dimension of dressing” is an indication that it is rather the (fractional) dressing exponent H which is intervening or whether there is some other mechanism at work. Because in practice the origin of the power law is not obvious, for a given field, the usual practice has been to simply regard qD as an empirical characterization. Finally, we should also mention (Tuck et al., 2004), (Tuck, 2008), (Tuck, 2010) who propose that molecular velocity may also have power law tails.

Figures 5.19a, b, c, 5.20a, b, 5.21 here

DRAFT

33

542543544545546547548549550551552553554555556557558559560561562563564565566567568569570571572573574575576

34

, 19/01/11,
From Hydrology papers?
Page 18: 5gang/ftp.transfer/final.book.nofigs.Feb.20…  · Web view5.2.4 Direct empirical estimation of c(g): the probability distribution multiple scaling (PDMS) technique. In chapter 3,

5/18/23 Chapter 5 18

Field Data source type qD referenceHorizontal wind Sonic 10Hz, time 7.5 (Schmitt et al., 1994a)

Hot wire probe Inertial range 7.7 (Radulescu et al., 2002)Hot wire probe Dissipation range 5.4 (Radulescu et al., 2002)anemometer 15 minutes 7 (Tchiguirinskaia et al.,

2006)anemometer daily 7 (Tchiguirinskaia et al.,

2006)Aircraft, stratosphere

Horizontal, 40m 5.7 (Lovejoy and Schertzer, 2007)

Aircraft, troposphere

Horizontal, 280m- 36 km

≈5 Fig. 5.10

Aircraft, troposphere

Horizontal, 40m- 20 km

≈7±1 (Chigirinskaya et al., 1994)

Aircraft, troposphere

Horizontal, 100m ≈5 (Schertzer and Lovejoy, 1985)

radiosonde Vertical, 50m 5 (Schertzer and Lovejoy, 1985), (Lazarev et al., 1994)

Potential temperature

radiosonde Vertical, 50m 3.3 (Schertzer and Lovejoy, 1985)

Humidity Aircraft, troposphere

Horizontal, 280m- 36 km

≈5 Fig. 5.10

Temperature Aircraft, troposphere

Horizontal, 280m- 36 km

≈5 Fig. 5.10

Paleotemperatures Ice cores 200 years, time 4 (Lovejoy and Schertzer, 1986)

Geopotential anomalies

Reanalyses 500 mb, daily 2.7 (Sardeshmukh and Sura, 2009)

Vorticity anomalies

Reanalyses 300 mb, daily 1.7 (Sardeshmukh and Sura, 2009)

Seveso pollution Ground Concentrations

In situ measurements

2.2 Salvadori et al., 1993

Chernobyl fallout Ground Concentrations

In situ measurements

1.7 (Chigirinskaya et al., 1998; Salvadori et al., 1993)

Table 5.1a: A summary of various estimates of the critical order of divergence of moments (qD) for various meteorological fields.

DRAFT

35

577

578579

36

Page 19: 5gang/ftp.transfer/final.book.nofigs.Feb.20…  · Web view5.2.4 Direct empirical estimation of c(g): the probability distribution multiple scaling (PDMS) technique. In chapter 3,

5/18/23 Chapter 5 19

Field Data source type qD referenceRadar reflectivity of Rain

Radar reflectivity factor

1 km3 resolution 1.1 (Schertzer and Lovejoy, 1987)

Rain Rate gauges Daily, time, France

≈3 (Ladoy et al., 1993)

gauges Daily, US 1.7 - 3 (Georgakakos et al., 1994)

High resolution gauges

8 minutes ≈2 (Olsson, 1995)

High resolution gauges

15 s 2.8 – 8.5 (Harris et al., 1996)

gauges Daily, time 3.6±0.07 (Tessier et al., 1996)gauges 1 – 8 days 3.5 (De Lima, 1998)gauges Hourly, time 4.0 (Kiely and Ivanova,

1999)Gauges Daily, 4 series

from 18th century3.78±0.46 (Hubert et al., 2001)

gauges Hourly, time ≈3 Fig. 5.10c, (Schertzer et al., 2010)

gauges Hourly, time ≈3 Fig. 5.20b, (Lovejoy et al., 2010a)

High resolution gauges

15s, averaged to 30 minutes

2.23 (Verrier, 2011)

Rain drop volumes

stereophotography 10 m3 sampling volume

5 (Lovejoy and Schertzer, 2008)

Liquid Water at turbulent scales

stereophotography Total water in 40 cm cubes

3 (Lovejoy and Schertzer, 2006b)

Stream flow River gauges (France)

daily 3.2±0.07 (Tessier et al., 1996)

River gauges (US) daily 3.2±0.07 (Pandey et al., 1998; Tessier et al., 1996)

River gauges (France)

daily 2.5-10 (Schertzer et al., 2006)

Table 5.1b: A summary of various estimates of the critical order of divergence of moments (qD) for various hydrological fields.

5.3.5 Divergence of moments in laboratory turbulence

Obviously, the cascade theories developed here have implications for laboratory turbulence – at least if the Reynolds number (Re) is high enough to allow the cascade to develop over a wide enough range. Aside from its intrinsic interest, the ability to use well-controlled laboratory-scale measurements allows us to resolve the dissipation scale and test our ideas further. To this effect, we analyzed active grid wind tunnel turbulent velocity data at a Taylor micorscale Reynolds number of Rλ = 582, approaching those of the atmosphere (the usual Reynolds number used earlier is roughly the square of Rλ i.e. Re ~ 105- 106; some of the largest Re laboratory

DRAFT

37

580

581582

583

584585586587588589590

38

Page 20: 5gang/ftp.transfer/final.book.nofigs.Feb.20…  · Web view5.2.4 Direct empirical estimation of c(g): the probability distribution multiple scaling (PDMS) technique. In chapter 3,

5/18/23 Chapter 5 20

turbulence to date). The experiments are described in (Mydlarski and Warhaft, 1998), along with the main characteristics of the turbulent flow. The Kolmogorov characteristic dissipation scale, as estimated from the mean energy dissipation rate is Ldiss = 0.26 mm. Time series of the velocity fluctuations in the longitudinal direction were obtained via hot-wire anemometry and analyzed in the spatial domain by using Taylor’s hypothesis. The inertial scaling range extended more than an order of magnitude in scales and the increased sampling rate allowed a resolution of the same order of the dissipation scale Ldiss.

To study the statistics of both the inertial range energy flux and the rate of energy dissipation, we studied the velocity differences of the longitudinal signal. The probability distributions (integral of the probability density functions) of |Dvr| = |v(x) - v(x + r)| normalized by the RMS value are shown in Fig. 5.22 for two typical examples taken at the highest resolution of the data, corresponding to dissipation range (DR) separations r/ Ldiss ≈ 2, and at a spacing within the inertial (scaling) range (IR) of scales r/ Ldiss ≈ 314 . The tails of both probability distributions display a distinctive power-law distribution, with qD ≈ 5.4 and 7.7, respectively. This is confirmed from the compensated probability distributions by the tail exponent qD,V, as shown in the inset of Fig. 5.22. A deviation from the power law dependence is observed for the extreme events measured at inertial range separations, but this is due to the limitation in the observation of events over larger separations for the same length of dataset, making the observation of these extreme events less probable.

According to the cascade model, it is the dressed turbulent fluxes which have the power law tails (finite qD’s); therefore, following the discussion in section 4.1.2, the qD of the observables ought to be different depending on whether we consider dissipation or scaling range values. Recall (section 4.1.2) that for the velocity field, we had Dv ∝ εη

with h = 1/3 in the scaling range, but h = 1/2 in the dissipation range. Dv ∝ εη

implies qD,h = qD,e/h so that we expect qD,IR/ qD,DR = 3/2.

According to the probability distribution of velocity differences at IR and DR separations in fig. 5.22 we have qD,IR ≈ 7.7 and qD,DR ≈ 5.4 so that qD,IR / qD,DR = 1.43 which is close to the theoretical ratio 1.5 and is consistent with atmospheric observations (table 5.1) and the value qD,e

= qD,IR/3≈ 2.6. A summary of this study may be found in (Radulescu et al., 2002).

Figure 5.22 here

5.3.6 The classification of multifractals according to the maximum

singularity they can generate

We have seen that the microcanonical constraint places an upper bound d on the orders of singularity (section 3.2.5). We saw that consequently for microcanonical cascades, there is no divergence of dressed moments of fluxes averaged over sets with dimension d. However, more restrictive (calm) types of multifractals exist. Here we briefly discuss the geometric multifractals introduced by (Parisi and Frisch, 1985) which involve neither a probability space nor a cascade process. (Parisi and Frisch, 1985) coined the term “multifractal” and gave the following quite abstract definition (the following is translated into our notation).

DRAFT

39

591592593594595596597598599600601602603604605606607608609610611612613614615616617618619620621

622

623

624625626627628629630

40

Daniel Schertzer, 29/11/10,
again a pure cut and paste, we may keep it as is..
Page 21: 5gang/ftp.transfer/final.book.nofigs.Feb.20…  · Web view5.2.4 Direct empirical estimation of c(g): the probability distribution multiple scaling (PDMS) technique. In chapter 3,

5/18/23 Chapter 5 21

Consider a component multifractal field v(r) on a space S dimension d; (Parisi and Frisch,

1985) considered a turbulent velocity field. Define the set of points Sγhoλ⊂S as the set of all

points rhol for which:

limDr→ 0

v(rγhoλ + Dr)−v(rγhoλ)

Dr−γhoλ

is non-zero and finite. For such points, Dv ≈ Δx −γ hol

where γhol is called the “Hölder” exponent. The set Sγhol has a fractal dimension D(γhol). If many different singularities γ are present, each distributed over a set dimension D(γhol), then v(r) is a multifractal. Notice that by restricting the multifractal to geometric sets, D(γhol) is necessarily nonnegative and c(γhol) = d- D(γhol) is bounded above by the embedding dimension d. Since c(γ) is an increasing function (see section 5.2.2) this implies that the orders of singularity are bounded. Just as microcanonical multifractals have maximum orders of singularity γmax= d, the geometric multifractals have γmax= c-1(d) which is less than d due to the convexity of c(γ) and the nondegeneracy requirement C1<d (see Figure 5.23). It is not obvious how to construct such geometric multifractals, they are simply defined as an abstract (highly non-trivial!) superposition of fractal sets. The basic difficulty is that cascade processes are generally not functions defined at mathematical points, rather they are more like “Dirac d functions” - where the word “function” is a misnomer because they are not true functions but rather limits of a series of functions. This means that for cascades, at a given point r, as we increase the resolution of the “incipient singularity” γλ(r):

γλ r( ) =logελ r( )

logλ

there will generally not be a well defined small scale limit: limλ→ ∞

γλ r( )≠γhoλ r( ).

Figure 5.23a, b, 5.24 here

5.4 Multifractals and deterministic chaos,

5.4.1 Box, information, and correlation dimensions

In section 3.2.3 we introduced both the box (Dbox) and correlation (Dcor) dimensions of a set of points: the first is the exponent of the average number of disjoint boxes size L/λ needed to cover the set, while the second is the exponent of the number of point pairs separated by a distance ≤ L/λ. Since both dimensions are in common use (Dcor particularly for characterizing strange “chaotic”/”strange” attractors) such as the Mandelbrot set, fig. 2.2, let us now consider the relation between the two. First suppose that the set of interest (denoted A) can be embedded in a d-dimensional “cube” of size L; and cover the cube with a grid of λd disjoint boxes each of size l = L/λ. Denote the number of points in the ith l-sized grid box by ni,λ so that the total number of points is:

DRAFT

41

631

632633

634

635636637638639640641642643644645646647648

649

650651652

653

654

655656657658659660661662663

42

Page 22: 5gang/ftp.transfer/final.book.nofigs.Feb.20…  · Web view5.2.4 Direct empirical estimation of c(g): the probability distribution multiple scaling (PDMS) technique. In chapter 3,

5/18/23 Chapter 5 22

N = ni,λi=1

λd

If the points are from a strange attractor (such as the Lorenz attractor), then the space is the system’s phase space and (with an ergodic hypothesis) Pi,λ= ni,λ/N is an empirical frequency that approximates the probability of finding the system in the ith box at phase space resolution λ = L/l, this would be its asymptotic limit for an infinite resolution. In order to characterize the scale by scale statistics of the attractor, similarly to estimating the “trace moments” (appendix 5A) we can use a “partition function” approach to introduce the following family of measures indexed by q (Hentschel and Procaccia, 1983), (Grassberger, 1983), (Halsey et al., 1986):

mq λ( ) = Pi,λq

i =1

λ d

and with the corresponding scaling exponents;

mq λ( ) ∝ l τ q( ) ∝ λ −τ q( )

To see the meaning of mq and its exponent t(q), first consider q = 0 and adopt the convention that for any x, x0 = 1 if x>0, and x0 = 0 if x =0. In this case, m0 is simply the number of boxes needed to cover the set and t(0) = -Dbox. Next, consider q = 2; in each box, the number of points which are within a distance l of each other is equal to the number of pairs in the box: ni,λ ni,λ −1( ) / 2 ≈ni,λ

2

(for large n and ignoring constant factors). However we have Pi,λ2 ∝ ni,λ

2

so that we see that m2 is proportional to the number of point pairs within a distance l, and hence t(2) = Dcor (the correlation dimension). The above suggests the definition:

D q( )=t q( )q−1

=1

q−1λimλ→ 0

λoγmq

λoγλ⎡⎣⎢

⎤⎦⎥; λ=L/ λ

This was first proposed by (Hentschel and Procaccia, 1983) and (Grassberger, 1983) for strange attractors, and acknowledged that such expressions were in fact pointed out in (Rényi, 1970), they are “Renyi dimensions”. In the next subsection, we relate D(q) to the dual codimension function (C(q); section 3.2.8) of the multifractal probability pλ.

What about the value q = 1? In this case, since the sum of the probabilities is unity, we

have m1 = Pi,λ = 1∑ so that eq. 58 reduces to 0/0 and we must use l’Hopital’s rule to evaluate the limit q->1. We find:

D 1( )=λimλ→ 0

pi,λi=1

λd

∑ λoγpi,λ

λoγλ

⎢⎢⎢⎢

⎥⎥⎥⎥; λ=L/ λ

D(1) is thus the exponent of the information Iλ:

DRAFT

43

664

665666667668669670671

672

673

674

675676677678

679680681

682

683684685686687

688689

690

691

44

, 21/01/11,
Yes, for rain for example
, 21/01/11,
Daniel Schertzer, 29/11/10,
is it what they really do?
Page 23: 5gang/ftp.transfer/final.book.nofigs.Feb.20…  · Web view5.2.4 Direct empirical estimation of c(g): the probability distribution multiple scaling (PDMS) technique. In chapter 3,

5/18/23 Chapter 5 23

Iλ = pi,λi=1

λd

∑ λoγ pi,λ

so that Iλ ≈lDI where the information dimension DI = D(1). In the next subsection we show that t(q) = D(q-1) -K(q) so that the convexity of K(q) (appendix 3B) implies the concavity of t(q) so that D(q) is a monotonically decreasing function of q; we therefore have the hierarchy: Dbox ≤DI

≤ Dcor.

5.4.2 The singular measure (f(a), t(q)) multifractal formalism

In the previous subsection, we saw that in order to statistically characterize these complex phase space densities, the density can itself be thought of as being a realization of a multifractal probability measure; we saw that this probability density pλ at resolution λ can be defined by covering the space with λ-d sized grids (boxes) and using pλ = nλ /N for each box.

The resulting measure is a geometric (calm) multifractal since although it represents the probability of finding the system at a point in the phase space, it is not itself random at all!

(Halsey et al., 1986)wrote an influential paper proposing a notation for dealing with these “geometric attractor” multifractals. Rather than considering the density of the

multifractal measure pλ =Pλ / voλ Bλ( ) (the non-random analogue of the turbulent λe ), the measure itself Pλ (i.e. the integral of the density pλ over a ball (box) size L/λ) was considered.

Figure 5.25 here

Pλ defined in this way is really a dressed quantity, but for these multifractals the bare/dressed distinction is irrelevant. They then determined the order of singularity ad of Pλ:

Pλ = pd d x =pλλ−d ≈λ−ad

Bλ∫

(the subscript “d” was not used in the original; we have added it to underscore the dependence on the dimension of the system and to distinguish it from the totally different

Levy a). In codimension notation, we may write pλ =λγ

so that Pl = λγ-d; we thus obtain:

ad = d − γ

Each box can thus be indexed according to ad. The number of boxes at each resolution corresponding to ad can then be used to define the (box counting) dimension fd(ad):

Number Pλ =λ−ad⎡⎣ ⎤⎦=λ fd ad( )

since number = λd x probability, and probability ≈ λ-c(γ), we obtain:

fd ad( )=d −c(γ), ad =d −γ

Finally, we can define the scaling exponents td (q) for the moments:

DRAFT

45

692

693694695696

697

698699700701

702703704705

706707708709710711712

713

714715

716

717

718719

720

721

722

723

46

Daniel Schertzer, 29/11/10,
should be part of an pasted introduction that is useless here
Page 24: 5gang/ftp.transfer/final.book.nofigs.Feb.20…  · Web view5.2.4 Direct empirical estimation of c(g): the probability distribution multiple scaling (PDMS) technique. In chapter 3,

5/18/23 Chapter 5 24

Pλ,iq

i=1

λd

∑ =λ−td q( )

where the sum is over all the i balls size −1λ needed to cover the d-dimensional phase space region A. Comparison with the moments of the density pλ shows that they differ only in the ensemble averaging; since p is non random, we therefore have:

Pλ,iq

i=1

λd

∑ = pλλ−d( )

q

i=1

λd

∑ ≈λ−td q( ) =λ−d q−1( ) pλq =λK q( )−d q−1( )

this implies:

td (q) = (q −1)d − K(q) = (q −1)D(q) D(q)= d − C(q)

As long as we deal with strange attractors and study the full d -dimensional phase

space, the ad, fd ad( ), td q( )

and D(q) notation is adequate. However, if we are interested in random multifractals (involving probability spaces) then d → ∞ , or if we are interested in looking at subspaces with dimension smaller than d, the d dependence is respectively a fundamental limitation or an unnecessary complication. The turbulent γ, c(γ), K(q), C(q) notation (the “codimension” notation for short) has the advantage of being intrinsic to the process i.e. of being d - independent.

5.5 Continuous in scale multifractal modeling

5.5.1 Review

Up until now, we have discussed discrete in scale models constructed by iteratively dividing large structures (‘eddies”) into disjoint daughter “subeddies” each reduced in scale by an integer ratio λ0. The smaller eddies have intensities which are equal to those of their parents multiplied by weights which are independently and identically distributed. They yield visually weird, highly artificial simulations (see e.g. fig. 5.26). One way of seeing the problem is that for these processes, the basic cascade equation <eλq> = λK(q) only holds exactly for the special scale ratios λ = λ0

n where both λ0 and n are integers i.e. only a “countable” infinity of scale ratios; in comparison, continuous in scale processes satisfy this equation for all λ>1. Other key limitations of these “toy model” cascades include a) isotropy (self-similarity; although limited self-affine extensions are possible (Schertzer and Lovejoy, 1985)) and b) left-right (mirror) symmetry (which precludes causal processes). In addition, most multifractal models are of scale by scale conserved cascade quantities whereas the observables typically have exponents with an extra linear term (e.g. Dv = e1/3DxH with H =1/3). A somewhat different but still discrete in scale multifractal simulation method is based on iterated function systems (Levy-Véhel et al., 1995), (Basu et al., 2004); see fig 5.27 for an example in one dimension. Perhaps the most sophisticated discrete in scale approach is the Markov-Switching Multifractal (MSM) simulation method; it is available only for pure (1-D) time series (Calvet and Fisher, 2001), (Calvet and Fisher, 2008).

DRAFT

47

724

725726727

728

729

730

731

732733734735736737

738

739

740741742743744745746747748749750751752753754755756

48

, 29/11/10,
Show IFS multifractals Show multifractals on a sphere
Page 25: 5gang/ftp.transfer/final.book.nofigs.Feb.20…  · Web view5.2.4 Direct empirical estimation of c(g): the probability distribution multiple scaling (PDMS) technique. In chapter 3,

5/18/23 Chapter 5 25

This model involves causal random transitions but is still based on iterating a finite fundamental scale ratio so that it is not yet truly continuous in scale.

Figures 5.26, 5.27, 5.28, 5.29 here

To explain continuous in scale cascade processes in a rather elementary manner, let us first review some properties of Lévy variables. First introduce the “unit” (and extremal) Lévy random variable γa whose probabilities are implicitly defined by the following characteristic function:

eqγa =eqa

a−1( ); q ≥0

eqγa =∞; q < 0; a < 2

Note that a) for a = 2 we have the familiar Gaussian case and the q ≥0 formula in eq. 68 is valid for all q, b) for a = 1 we have eqγa =eqλoγq (q >0, otherwise =∞ ). An extremal Lévy random variable A with amplitude a>0 and Lévy index a therefore satisfies:

A =aγa

eqA =eaa qa

a−1 ; q ≥0

eqA =∞; q < 0; a < 2

(with corresponding exception for a = 1). Due to the additivity of second characteristic functions for any independent, identically distributed random variables (section 3.3.3), this implies that for the sum of two statistically independent Lévy variables A, B we have:

C =A+ B; ca =aa +ba

where C is also an extremal Lévy with the same a but with amplitude c. Eq. 70 expresses the “stability under addition” property of the Lévy variables and a, b, c are the corresponding amplitudes. Equation 70 shows that for Lévy variables log eqF (the second characteristic function) of the random variables F =fγa (amplitude f) are “a additive”, a property that generalizes to Lévy noises; we use this below.

To understand why only extremals must to be used to generate cascades (Schertzer and Lovejoy, 1987), (Schertzer and Lovejoy, 1991), and not the more general Lévy variables la with a<2, let us recall the general properties of Lévy variables. These have algebraic tails for both positive and negative values:

laq

→ ∞; q ≥a

p λa( )≈A+λa−a−1; λa >>1

p λa( )≈A− −λa( )−a−1 ; λa << −1

DRAFT

49

757758759760761762763764765

766

767768769

770

771

772773774

775

776777778779780781782783784

785

50

Page 26: 5gang/ftp.transfer/final.book.nofigs.Feb.20…  · Web view5.2.4 Direct empirical estimation of c(g): the probability distribution multiple scaling (PDMS) technique. In chapter 3,

5/18/23 Chapter 5 26

where p is the probability density and A+ and A- are constants that depend on an “asymmetry parameter” (the a = 2 case is the qualitatively different Gaussian case, where there is neither power law nor asymmetry). For symmetric Lévy’s, A+ = A-; whereas for maximally asymmetric Lévy’s without algebraic behaviour for l>>0, A+ =0. The latter are required to avoid the divergence of the Laplace transform for q>0, simply because an algebraic fall-off cannot tame an exponential divergence. It turns out that for a<1, the values of an extremal Lévy variables all have the same sign as that of the algebraic tail. This is not the case for a>1, although most of the values have the same sign as the tail. The qualitative difference between the a>1 and a<1 processes is more apparent if we consider the corresponding asymptotic forms of the probability densities, which can be seen in fact as a straightforward consequence of the universal multifractal codimension function (eq. 35) with C1=1 :

p γa( )≈exp −γa

′a⎛⎝⎜

⎞⎠⎟

′a⎡

⎣⎢⎢

⎦⎥⎥; γa

′a⎛⎝⎜

⎞⎠⎟

′a

>> 0;1′a+1a

=1

p γa( )≈ −γa( )−a−1

; γa << 0; 0 <a < 2

where the key point is that the auxiliary variable a’ changes sign at a =1. Note that exact closed form expressions in elementary functions for the probabilities of extremals only exist in the “inverse 26gaussian” a = ½ case; analytic symmetric Lévy’s exist for the a = 1 (Cauchy) and a = 2 (Gaussian) cases.

5.5.2 Continuous in scale cascade processes

An (extremal) white Lévy noise γa(r), can be understood as the limit of independent identically distributed extremal Lévy variables on a grid for a mesh size shrinking to zero (with appropriate normalization); this is possible because Lévy distributions are “infinitely divisible”; see (Feller, 1971)). With the help a suitably normalized convolution kernel gλ(r), it then possible

to colour this white noise to obtain (Schertzer and Lovejoy, 1987) a generator Gλ =logελ such

that eλ

q = eqΓλ = eK q( ) log λ

. We now show how to obtain a generator with the appropriate properties (including a mean codimension C1) by convolving a Levy noise γa(r) with a kernel gλ(r):

Gλ =C11/α gλ ∗γ α = C1

1/α gλ r − r′( )γα∫ r( )dr

“*” denotes “convolution”. We now review, step by step the different properties that the kernel gλ(r) and its domain of integration must satisfy in order to obtain the announced result, in particular that the multifractal eλ:

e λ =eΓλ

DRAFT

51

786787788789790791792793794795796

797

798799800801

802

803804805806

807

808809810

811

812813814

815

52

Daniel Schertzer, 29/11/10,
somethig nothing not easy at all. WHY to spend time to re-invent the wheel ?
Daniel Schertzer, 17/12/10,
possibly in appendix
Daniel Schertzer, 17/12/10,
Still not SATISFACTORY: - should be re-rodered, - the technical details should appear only at the last stage, - e.g. the question of normalization.. - still important subtlety not explained, e.g. why intergration over a pixel is not a problem to compute the caracteristic function
Daniel Schertzer, 29/11/10,
the second line of eq.88 is NOT obvious and presumably WRONG and USELESS, in any case, >>1, instead of >>0
, 21/01/11,
, 21/01/11,
There was an explanation, now just a citation… is this better?
Page 27: 5gang/ftp.transfer/final.book.nofigs.Feb.20…  · Web view5.2.4 Direct empirical estimation of c(g): the probability distribution multiple scaling (PDMS) technique. In chapter 3,

5/18/23 Chapter 5 27

will be indeed be multiscaling: eλq = λ K q( )

. This is the first property to be respected:

KΓ (q,λ ) = Log(< eqΓλ >= Log(λ )K(q)i.e. gλ(r) must be chosen so as to yield a logarithmic divergence of the second characteristic function of the generator Gλ (Schertzer and Lovejoy, 1987). Before detailing the corresponding necessary and sufficient conditions, we need to stress that for a<2, gλ must be non-negative otherwise G would a (nonextremal) mixture of extremal Lévy variables with opposite signs, see fig. 5.30 for a graphical illustration. Note that in this chapter we will only consider isotropic d dimensional multifractals, the only exception being the asymmetric causal processes needed for space-time simulations in ch. 7 we consider anisotropic extensions. Since γa(r) is statistically homogeneous, from eq. 73, the statistics of G are independent of r so that one can take r = 0 and apply the additivity of the second characteristic function (eq. 70):

KG(q)=C1

qa

a −1γλ a

a; γλ a

a≡ γλ r( )∫

ad dr

where the a-1 in the denominator comes from the definition of the unit Levy variables, eq. 68.

Figure 5.30 here

In order to obtain the desired Log(λ) divergence for KG (eq. 75), it suffices to choose gλ to be an isotropic power law with the appropriate power law from the larger scale L to the resolution L/λ:

gλ r( )=Nd ,a

−1/a1L/λ≤r≤L r−1/a

where

1B is the indicator function of the subset B i.e. 1B(r) =1 if r ∈B , = 0 otherwise. This implies:

eλq = eqΓλ = e

C1

α −1qα Nd

−1Ωd log λ; Ωd = d d ′r

′r =1∫

where Wd is the integral over all the angles in the d dimensional space (e.g. W1 = 2, W2 = 2p, W3 = 4p etc.). Note that for theoretical work it is more convenient to take L = 1 so that the smallest scale is 1/λ whereas for numerical simulations on discrete grids it is more convenient to take L = λ so that the smallest scale is 1.

From eq. 78, we see that if we choose:

Nd =Wd

then we obtain the desired nonlinear part of the multiscaling behaviour:

eλ ,uq = λ Ku q( ); Ku q( ) =

C1

α −1qα

DRAFT

53

816

817

818819820821822823824825826

827

828829830831832833834

835

836837

838

839840841842843

844

845

846

54

Daniel Schertzer, 17/12/10,
There is subtlety that is not yet explained with 1(a-1) especially for a <1
Page 28: 5gang/ftp.transfer/final.book.nofigs.Feb.20…  · Web view5.2.4 Direct empirical estimation of c(g): the probability distribution multiple scaling (PDMS) technique. In chapter 3,

5/18/23 Chapter 5 28

where Ku is the unnormalized exponent scaling function corresponding to the fact that e given by eq. 73, 74 is unnormalized (hence we temporarily add the subscript “u”). A normalized eλ,n can now be easily obtained using:

eλ ,n =ε λ ,u

ε λ ,u

so that:

eλ ,nq = λ K q( ); K q( ) = Ku q( ) − qKu 1( ) =

C1

α −1qα − q( )

as required (we temporarily add the subscript “n” to distinguish it from the unnormalized process). The above leads to cascades with the correct statistics at the finest resolution λ. Unfortunately, it turns out that while the above is correct, that due to “finite size effects” at both large but mostly at small scales, that the internal structure of the realizations is not perfectly scaling. If needed, corrections are not difficult to make, see appendix 5B. Numerical simulations based on spatially discretizing these continuous in scale cascade processes are shown in fig. 5.12, 5.13, 5.28.

5.5.3 Fractional Brownian and fractional Lévy noises

Up until now, we have concentrated our attention on the underlying turbulent fluxes which are conserved from one scale to another. We have seen that the typical observables such as the wind have fluctuations (Dv) whose statistics are related to the fluxes by a lag Dx raised to a power, the prototypical example being the Kolmogorov law: Dv = φDxH with φ = e1/3, H =1/3. If we take the qth moments of this equation, we obtain:

Sq Dx( )∝ Dxx q( ); Sq Dx( )= Dvq ; x q( )=qH −K q( )

(we have used <φλq> ≈ λK(q) and Dx ∝ λ −1

). Sq(Dx) is the qth order structure function and x(q) is its scaling exponent; the usual (2nd) order structure function (related to the spectrum) was defined in eq. 2.72 (see also section 3.2.10). Since we now know how to construct fields with the (convex) scaling exponent K(q), our problem is to produce a field with the extra linear exponent qH. Note that if the fluctuations Dv are defined simply as differences (i.e. Dv = v(x+Dx)-v(x)) then the above is only valid for 0≤H≤1; when H is outside this range other definitions of fluctuations must be used, see section 5.6.

Before the development of cascade processes, the intermittency of the flux φ was considered unimportant, and turbulence was modelled (for the purposes of statistical closure theories for example) with quasi-Gaussian statistics, this implies that K(q) = 0 so that these processes therefore had perfectly linear exponents x(q) = qH. Before discussing the fractionally integrated flux (FIF) model which gives x(q) with convex K(q), let us therefore briefly discuss the simpler processes that give rise to linear x(q).

If x(q) is linear, we anticipate that the corresponding v(r) can be modelled using an additive stochastic process. In order to assure that the process can be made continuous, we

DRAFT

55

847848849

850

851

852

853854855856857858859

860

861862863864865

866

867868869870871872873874875876877878879880881

56

Page 29: 5gang/ftp.transfer/final.book.nofigs.Feb.20…  · Web view5.2.4 Direct empirical estimation of c(g): the probability distribution multiple scaling (PDMS) technique. In chapter 3,

5/18/23 Chapter 5 29

consider, fractional Brownian motion (fBm) and its generalization, fractional Lévy motion (fLm) which are produced by linear combinations of independent Gaussian and Lévy noises with pure singularities (the noises φa(r) are of the same type as the γa(r) introduced in the previous subsection). More precisely, these are written as convolutions of noises with power laws which are extensions of integration/differentiation to fractional orders; “fractional integrals”:

v r( )=γa ∗r− d− ′H( ) =γa ′r( )

r− ′r d− ′H d d ′r∫ ; ′H =H + d /a

where γa is again a Lévy noise made of uncorrelated Lévy random variables (here they need not be extremal) and H’ is the order of fractional integration (as usual, the Gaussian case is recovered with a =2). At each point, the resulting v field is a Lévy variable and has fluctuation statistics (of Dv(Dr) = v(r+Dr) - v(r)) obeying eq. 83. When a<2, x(q) diverges for q>a These models are monofractal because the fractal codimension of any level set v(r) = T has a codimension c(T) = H (i.e. independent of T). Note that for the integral to converge, we require 0<H<1. In order to obtain processes with H outside this range, we may use eq. 84 for the fractional part and then perform ordinary (integer valued) differentiation or integration as required.

The power law convolution (eq. 8) is easier to understand if we consider it in Fourier space (by taking Fourier transforms of both sides of eq. 84). Since the Fourier transform (“F.T.”) of a singularity is another singularity (section 2.4.5; the Tauberian theorem):

r − D−H( )→FT

k −H

We can use the basic property that a convolution is Fourier transformed into a multiplication, to obtain simply:

v%k( )∝ γa± k( ) k −H

where:

v r( )→FT

v%k( ); γa r( )→FT

γa± k( )

We thus see that the convolution with power law r − D −H( )

is the equivalent to a power

law filter k −H

. However, such filters are themselves generalizations of differentiation (H<0) or

integration (H>0). To see this, recall the Fourier transform of the Laplacian: ∇2γα( )→

F .T .− k −2 γα

±

so (ignoring constant factors) that k −H

corresponds to real space ∇2( )− H /2

, i.e. for H>0 it corresponds to a negative order differentiation (i.e. integration) of order H. Let us mention that this definition of fractional derivative with the help of the Laplacian is very convenient but somewhat restrictive. In fact, it is related to an underlying microscopic (isotropic) Brownian motion. As soon as the latter are replaced by Lévy flights (i.e. additive Levy processes), strongly asymmetric fractional operators are obtained, (Schertzer et al., 2001).

DRAFT

57

882883884885886

887

888889890891892893894895896897898899

900

901902

903

904

905

906

907

908

909910911912913914915

58

, 21/01/11,
Not clear?!
, 13/07/11,
Change “D”’s to “d”
Page 30: 5gang/ftp.transfer/final.book.nofigs.Feb.20…  · Web view5.2.4 Direct empirical estimation of c(g): the probability distribution multiple scaling (PDMS) technique. In chapter 3,

5/18/23 Chapter 5 30

Figure 5.31a, b here

5.5.4 The Fractionally Integrated Flux model for nonconservative

multifractals

In analogy with the fractional Gaussian/Lévy noises and motions, non conservative multifractal fields are obtained by a fractional integration of a pure cascade process eλ (a “flux”, hence the name “Fractionally Integrated Flux”, FIF model, (Schertzer et al., 1997)):

vλ r( )=eλ ∗r d−H =eλ ′r( )d d ′rr− ′r d−H∫

we already saw that the flux itself can be modelled in the same (power convolution/fractional integration) framework: (eq. 73). Fig. 5.31 a, b shows respectively the effect of both fractional integration (H = +0.333) and fractional differentiation (H = -0.333) on the cascade shown in fig. 5.12a with C1 = 0.1, varying a.

The basic structures of the FIF and fBm and fLm processes are compared and contrasted in table 5.2. We see that the difference is simply the type of noise (Gaussian, Lévy, multifractal) which is (fractionally) integrated. Fig. 5.32 a shows a comparison of fBm, fLm and a universal multifractal process with the same H value; we can see that the fBm gives a relatively uninteresting texture. fBm is fairly limited in its possibilities since due to the central limit theorem (the Gaussian special case), a process with the same statistical properties can be produced by using singularities of quite different shapes; the details of the shape get “washed out”. In comparison as one can see in fig. 5.32a, the fLm has extremes which are on the contrary too strong: several strong mountain peaks stand out. Although far from Gaussian, most atmospheric fields as well as the topography seem to empirically have finite variance (i.e. qD>2; table 5.1) so a Fractional Lévy motion (fLm) is generally not a relevant model for those fields. On the contrary the multifractal simulation has much more interesting structures, however we are still missing the interesting ridges, valleys and other anisotropic features of real geomorphologies which we deal with in ch. 6, 7.

Figure 5.32a here

DRAFT

59

916917

918

919

920921922

923

924925926927928929930931932933934935936937938939940941942943

60

, 21/01/11,
Add an appendix showing that the strucutrree functions really do have the form qH-K(q)
, 21/01/11,
Also 1987 reference?
, 21/01/11,
MOTIONS NOT DEFINED
Page 31: 5gang/ftp.transfer/final.book.nofigs.Feb.20…  · Web view5.2.4 Direct empirical estimation of c(g): the probability distribution multiple scaling (PDMS) technique. In chapter 3,

5/18/23 Chapter 5 31

Model Field Increments Codimensions of

level sets

Monofractal

fBmv r( )=γ2 ∗r− d− ′H( )

′H = H + H2; H2 = d / 2

Dv=d

γ 2 Δr H

Dv q ∝ Δr ξ q( )

x q( ) = qH

C=H

D=d-C

Monofractal

fLmv x( )=γa ∗r− d− ′H( )

′H = H + Hα ; Hα = d / α

Dv=d

γ α Δr H

Dv q ∝ Δr ξ q( )

x q( ) =qH ; q < α∞; q > α

C=H

D=d-C

Multifractal

FIFGλ ∝C1

1/α γ α ∗ r − d − H ′α( ) ; ϕ λ = eΓλ

vλ r( )=f λ ∗r− d−H( )

H ′a =d / ′a ;1a+

1′a=1

Dv = ϕ λ Δr H

Dv q ∝ Δr ξ q( )

x q( ) = qH − K q( )

c γ( )=m axq

qγ −K q( )( )

D γ( )=d −c γ( )

Table 5.2: An intercomparison of various scaling models for v showing the essential similarities and differences in their mathematical structure, statistical properties. Here d=2 for horizontal planes and the dimension D is the fractal dimension of lines of constant v in the horizontal. The monofractal fractional Brownian motion (fBm) model involves a fractional integration of order H’ with a flux γ2(r) which is simply a (“d correlated”) gaussian white noise with variance s2. Note that the symbol a=b

d indicates equality in probability distributions, i.e. a=b

d⇔ Pr a> s( )=Pr b > s( ) for all s, “Pr” indicates “probability”. It results in fluctuations with

gaussian statistics, linear structure function exponent x(q) and altitude independent surface codimension c (or dimension D). The fLm is the generalization obtained by replacing Gaussian variables by stable Lévy variables with index a (fBm is obtained in the case a=2). These have diverging moments q for q≥a. Finally, the multifractal Fractionally Integrated Flux (FIF) model has the same structure, except that the white noises are replaced by multifractal cascades φλ

where λ is the resolution. The multifractal noise φl is the result of a continuous in scale multiplicative cascade. We again find the exponent H, although now there are an infinite number of codimensions c(γ) (or dimensions D(γ)) that depend on the threshold given by λγ (do not confuse the singularity γ with the subgenerator γa). To generalize fBm, fLm and FIF to anisotropic processes, we must replace the distances in the fractional integration denominators by anisotropic scale functions as discussed in ch. 6.

DRAFT

61

944

945946947948949950951952953954955956957958959960961962963

62

Page 32: 5gang/ftp.transfer/final.book.nofigs.Feb.20…  · Web view5.2.4 Direct empirical estimation of c(g): the probability distribution multiple scaling (PDMS) technique. In chapter 3,

5/18/23 Chapter 5 32

5.5.5 Morphologies of isotropic (self-similar) universal multifractals

Considering the “universal multifractals” vλ, defined by eq. 73, 74, 88 we see that they are isotropic (the singularities have no preferred directions, they depend only on the vector norm), they are therefore “self-similar”; “zoomed” structures will (on average) resemble the unzoomed ones. In addition, they depend on three parameters: the a, C1 which define the statistics of the generator Gλ (eqs. 73 - 76) and the H in eq. 88. While the parameter H is the order of fractional integration and quantifies the degree of scale invariant smoothing, the qualitative effects of the codimension of the mean (C1) and the Lévy index (a) are less easy to see. For H>0, the k-H filter is a kind of scale invariant smoothing, whereas for H<0 it is scale invariant roughening. This was already confirmed in the 1-D simulations fig. 5.31 a, b. To further explore the effect of varying H on the various morphologies, we performed multifractal simulations in 2D (rather than in 1D). Fig. 5.32 b systematically shows the morphologies of the structures obtained by varying these three parameters, each has the same initial random “seed” so that the basic generating white-noises are the same. For large scale geo-applications, we require simulations on the sphere (fig. 5.32 c; these are discussed in appendix 5.D).

To systematically survey the simulation parameter space (a, C1, H), we use false colours (using a “cloud” palette, fig. 5.33) and simulations with a =0.4- 2.0 in increments of 0.4, as well as C1, H from 0.05 to 0.80 in increments of 0.15, amply covering the parameter range commonly encountered in the geosciences (this may be compared with fig. 6.9 showing the same simulations but with scaling stratification). In fig. 5.34, rather than fixing a and varying C1, H we show the effect of fixing C1 and H and varying a. For fixed a, C1, we see that increasing H systematically smoothes the structures whereas for fixed a, H, varying C1 changes the “spottiness”, sparseness of structures. For reference, note that the empirically most common values of a are in the range 1.5-1.9 (the latter being appropriate for topography and cloud radiances, the former, for rain and atmospheric turbulence). The parameter C1 is often fairly low (e.g. in the range 0.05-0.15 for the wind, cloud radiances, topography; see table 4.8), although it can be large (0.25-0.7) for rain and turbulent fluxes. While the basic Kolmogorov value of H is 1/3, many fields (such as cloud radiances) are near this, while for rain this parameter is nearly zero, for topography it is in the range 0.45-0.7. From fig. 5.34 we can see that high values of C1

lead to fields totally dominated by one or two strong structures, while low a values lead to fields dominated by “Lévy holes”: large regions with extremely low values.

Figure 5.32b, c, 5.33 a,b,c,d,e, 5.34 a,b,c,d

5.6 Wavelets and fluctuations: Structure functions and other data

analysis techniques

5.6.1 Defining fluctuations using wavelets

We have seen that data analyses constantly rely on defining fluctuations at a given scale and location; the simplest definition of fluctuation at position x, scale Dx, being Dv(x,Dx) = v(x+Dx) - v(x). Note that since we typically assume that the statistics of the fluctuations are

DRAFT

63

964

965966967968969970971972973974975976977978979980981982983984985986987988989990991992993994995996

997

998

999

100010011002

64

, 11/01/11,
We introduce structrure functions earlier, I’m not sure what you want here.
Daniel Schertzer, 17/12/10,
A large part deals with somehtinf elsse: introduc a subsection on structure function?, Or change the title into “structure functions and wavelets?
Daniel Schertzer, 17/12/10,
Misleading denomination
Page 33: 5gang/ftp.transfer/final.book.nofigs.Feb.20…  · Web view5.2.4 Direct empirical estimation of c(g): the probability distribution multiple scaling (PDMS) technique. In chapter 3,

5/18/23 Chapter 5 33

independent of position, we previously suppressed the x argument. We have already mentioned (ch. 4) that other definitions of fluctuation are possible and are occasionally necessary, let us examine this a bit more closely.

Consider the statistically translationally invariant process v(x) in 1-D: the statistics are thus independent of x and this implies that the Fourier components are “d correlated”:

v%k( )v% ′k( ) =dk+ ′k v%k( )

2; v%k( )= e−ikxv x( )dx∫

If it is also scaling then the spectrum E(k) is a power law: E k( )≈ v%k( )

2≈k−b

(where here and below, we ignore constant terms such as factors of 2p etc.). In terms of its Fourier components, the fluctuation is thus:

Dv x, Δx( ) = v x + Δx( ) − v x( ) = eikx v%k( ) eikΔx − 1( )dk∫

so that the F.T. of Dv(x,Dx) is v%k( ) eikDx −1( ) . We first consider the statistics of quasi-Gaussian

processes for which C1 = 0, x(q) = Hq. Exploiting the statistical translational invariance, we drop the x dependence and relate the second order structure function to the spectrum:

Dv Δx( )

2 = 4 eikx v%k( )2 sin2 kΔx2

⎛⎝⎜

⎞⎠⎟dk∫ ≈ eikxk−β sin2 kΔx

2⎛⎝⎜

⎞⎠⎟dk∫

As long as the integral on the right converges, then the usual Tauberian argument (section 2.4.5) shows that:

Dv Δx( )2 ∝ Δxξ 2( ) ≈ eikxk−β sin2 kΔx

2⎛⎝⎜

⎞⎠⎟dk∫ ∝ Δxβ −1

so that b = x(2)+1 = 2H +1 (C1 = 0 here); see eqs. 4.18, 4.19. However, for large k, the integrand ≈ k-b which has a large wavenumber divergence whenever b<1. However, since for small k, sin2 kDx / 2( )∝ k2

, there will be a low wavenumber divergence only when b>3. In the divergent cases, although real world (finite) data will not diverge, the structure functions will no longer characterize the fluctuations, but will depend spuriously on the either the highest or lowest wavenumbers present in the data. In the quasi gaussian case, or when C1 is small, we have x(2) ≈ 2H and we conclude that the using first order differences to define the fluctuations leads to second order structure functions being meaningful in the sense that they adequately characterize the fluctuations whenever 1<b<3, i.e. 0<H<1.

Since 0<H<1 is the usual range of geophysical H values, and the difference fluctuations are very simple, they are commonly used. However, we can see that there are limitations; in order to extend the range of H values, it suffices to define fluctuations using finite differences of different orders. To see how this works, consider using second (centred) differences:

DRAFT

65

10031004100510061007

1008

100910101011

1012

101310141015

1016

10171018

1019

10201021

10221023102410251026102710281029103010311032

66

, 11/01/11,
I’m not sure what you want here
Daniel Schertzer, 17/12/10,
Problem of noation: - the ordre, when ≠1, should appear
Page 34: 5gang/ftp.transfer/final.book.nofigs.Feb.20…  · Web view5.2.4 Direct empirical estimation of c(g): the probability distribution multiple scaling (PDMS) technique. In chapter 3,

5/18/23 Chapter 5 34

Dv x, Δx( ) = v x( ) −12

v x + Δx( ) − v x − Δx( )( ) = eikx v%k( ) 1−12

eikΔx /2 + e−ikΔx /2( )⎡⎣⎢

⎤⎦⎥dk∫

= 2 eikx v%k( )sinkΔx

4⎛⎝⎜

⎞⎠⎟

2

dk∫

repeating the above arguments we can see that the relation b = x(2)+1 holds now for 1<b<5, or (with the same approximation) 0<H<2. Similarly, by replacing the original series by its running sum (a finite difference of order -1) – as done in the MFDFA technique described below – we can extend the range of H values down to -1. Since the “weather-climate plateau” (at scales longer than about two weeks, up to decades and centuries) is precisely characterized by -1<H<0, a corresponding “tendency structure function” technique is indeed useful but its introduction is postponed until ch. 10.

More generally, going beyond Gaussian processes we can consider intermiitent FIF

processes, which have v%k( )≈e%k( ) k −H

, we see that the F.T. of Dv(x,Dx) is eikDx −1( ) k−H e%k( ) .

This implies that for low wavenumbers, v%k( )≈ k 1−H e%k( ) (k<<1/Dx) whereas at high

wavenumbers, v%k( )≈ k −H e%k( ) (k>>1/Dx) hence since the second characteristic function of e(x)

has logarithmic divergences with scale for both large and small scales ( log eλq =K q( )λoγλ ),

we see that for 0<H<1, that the fluctuations Dv(Dx) are dominated by wavenumbers k ≈ 1/Dx, so that for this range of H, fluctuations defined as differences capture the variability of Dx sized structures, not structures either much smaller or much larger than Dx. More generally, since the

F.T. of the nth derivative dnv/dxn is ik( )n v%k( ) and the finite derivative is the same for small k but “cut-off” at large k, we find that nth order fluctuations are dominated by structures with k ≈1/Dx as long as 0 < H <n. This means that Dv(Dx) does indeed reflect the Dx scale fluctuations.

While finite differences are usually adequate in scaling applications, in the past twenty years there has been a development of systematic ways of defining fluctuations; wavelet analysis (for a meteorological introduction, see (Torrence and Compo, 1998), (Foufoula-Georgiou and Kumar, 1994)for a mathematical introduction, see (Holschneider, 1995)).

In wavelet analysis, one defines fluctuations with the help of a basic “mother wavelet” Y(x) and performs the convolution:

Dv x,Δx( ) = v ′x( )∫ Ψ′x − xΔx

⎛⎝⎜

⎞⎠⎟d ′x

where we have kept the notation Dv to indicate “fluctuation”. The basic “admissibility” condition on Y(x) (so that it is a valid wavelet) is that it has zero mean. If we take the

fluctuation Dv as the symmetric difference Dv = v x + Δx / 2( ) − v x − Δx / 2( ) then:

Y x( ) = δ x − 1 / 2( ) − δ x +1 / 2( )

where “d” is the Dirac delta function, then we recover the usual first difference fluctuation (eq. 90, see fig. 5.35b; due to the statistical translational invariance, this is equivalent to the difference, eq. 90). This “poor man’s” wavelet is usually adequate for our purposes. As the generality of the definition (eq. 94) suggests, all kinds of special wavelets can be introduced; for

DRAFT

67

1033

10341035103610371038103910401041

1042

1043

1044

1045104610471048

104910501051105210531054105510561057

1058

10591060

1061

1062

1063106410651066

68

Page 35: 5gang/ftp.transfer/final.book.nofigs.Feb.20…  · Web view5.2.4 Direct empirical estimation of c(g): the probability distribution multiple scaling (PDMS) technique. In chapter 3,

5/18/23 Chapter 5 35

example special orthogonal wavelets can be used which are convenient if one wishes to “reconstruct” the original function from the fluctuations, or to define power spectra locally in x rather than globally, (averaged over all x). Alternatively, we could consider the second derivative of the Gaussian which is the popular “Mexican hat” shown in fig. 5.35, along with the second finite difference wavelet:

Y x( ) =12

δ x +12

⎛⎝⎜

⎞⎠⎟ + δ x −

12

⎛⎝⎜

⎞⎠⎟

⎛⎝⎜

⎞⎠⎟

− δ x( )

The Mexican hat is thus essentially just a smoothed out version of the 2nd finite difference. Just as one can use higher and higher order finite differences to extend the range of H values, so one can simply use higher and higher order derivatives of the Mexican hat. The second finite difference wavelet is easy to implement, it is simply Dv = (v(x+Dx)+v(x-Dx/2))/2 - v(x) and has the advantage that the structure function based on this is valid for 0<H<2.

Fig. 5.35a here

Because of the usual difference structure function is restricted to H in the range 0<H<1, it is useful to introduce variants, in particular, to extend the range down to H = -1. The idea is simple: perform a running sum (i.e. an integral of order 1, increasing H by 1), and then perform a second finite difference of the running sum, so that with respect to the original function, the range of H is now -1<H<1. In terms of v, the fluctuation Dv is:

DvHaar x,Δx( ) =1

2Δxv ′x( )d ′x

x

x+ Δx /2

∫ − v ′x( )d ′xx−Δx /2

x

∫⎡⎣⎢

⎤⎦⎥

where we have added the extra 1/Dx factor so that the scaling is the same as for the poor man’s fluctuation, i.e. xHaar(q) = x(q) for processes with 0<H<1. To within the division by Dx, the result is the equivalent of using the Haar wavelet as indicated in fig. 5.35b:

Y x( ) =

1 / 2; 0 ≤ x < 1 / 2

−1 / 2; −1 / 2 ≤ x < 0

0; otherwise

It can be checked that the fourier transform of this wavelet is 2ik−1sin2 k / 4( ) so that it is better localized in Fourier space with both low and high wavenumber fall-offs (≈ k and ≈ k-1

respectively). It turns out that for analyzing multifractals, that even over their common range of validity (0<H<1), the Haar structure function does somewhat better than the poor man’s wavelet structure function (i.e. the usual structure function). This is demonstrated on numerical examples in appendix 5E. While the latter has the advantage of being relatively simple to interpret (e.g. the “typical difference” for the mean of the absolute difference), the corresponding interpretation of the Haar wavelet is less intuitive. To understand it better, including its interpretation for H<0 which is important for low frequency weather signals which typically have -1<H<0 (ch. 10), we introduce a slightly different wavelet based fluctuation that we call the “tendency fluctuation”:

DRAFT

69

10671068106910701071

1072

1073107410751076107710781079108010811082108310841085

1086108710881089

1090

1091109210931094109510961097109810991100

70

, 11/01/11,
I don’t believe that wavelets are by definition orthogonal: is that your point?!
Daniel Schertzer, 17/12/10,
there is a confustion between the methodology and the wavelets themselves !!!
Page 36: 5gang/ftp.transfer/final.book.nofigs.Feb.20…  · Web view5.2.4 Direct empirical estimation of c(g): the probability distribution multiple scaling (PDMS) technique. In chapter 3,

5/18/23 Chapter 5 36

Dv Δx( ) =1

Δx′v ′x( )d ′x ;

x

x+Δx

∫ ′v x( ) = v x( ) − v x( )

we can see that the fluctuation Dv Δx( ) defined in this way is effectively a spatial average, hence

the overbar notation; v’(x) is the transect v(x) with the mean removed: ′v x( ) = 0 . In terms of wavelets, this can be seen to be equivalent to using the wavelet:

Y x( ) = I −1/2,1/2[ ] x( ) −I − L /2,L /2[ ] x( )

L; L >> 1

where I is the indicator function:

I a,b[ ] x( )= 1 a≤x ≤b0 otherwise

see fig. 5.35 b for a schematic (note that the difference between the tendency fluctuation and wavelet is the division by Dx in eq. 99 which is done so that the scaling of the structure functions defined from tendency fluctuations is the same as for the other fluctuations).

The first term in eq. 100 represents the integral in eq. 99 whereas the second removes the mean L>>1 is the overall length of the data set). The removal of the mean in this way is necessary in order that the wavelet satisfy the admissibility condition that its mean is zero. The only essential difference between wavelet coefficients defined by eq. 100 and the “tendency fluctuation” defined in eq. 99 (where the mean is removed beforehand), is the extra normalization by Dx in eq. 99 which changes the exponent by one. The term “tendency structure function” is used since for H<0, it gives directly an estimate of the typical tendency of the fluctuations to increase or to decrease.

From the above we see that if needed, the Haar wavelet can be generalized by using the (normalized) third difference of the running sum: Dv Δx( ) = s x + Δx( ) / 3 − s x + Δx / 3( ) + s x − Δx / 3( ) − s x − Δx( ) / 3( ) / Δx which is very close to the Mexican hat wavelet (see fig. 5.34a) but which is numerically much easier to implement. As with the Mexican hat, it is valid for -1<H<2.

Wavelet mathematics are seductive - and there certainly exist areas such as speech recognition where both wavenumber and spatial localization of statistics are necessary. However, the use of wavelets in geophysics is often justified by the existence of strong localized structures which are cited as evidence that the underlying process is statistically nonstationary/ statistically inhomogeneous (i.e. in time or in space respectively). However, we have seen that cascades produce exactly such structures - the singularities - but that their statistics are nevertheless strictly translationally invariant. In this case the structures are simply the result of strong singularities but nevertheless, the local statistics are uninteresting, and they are averaged out in order to improve statistical estimates.

Before continuing to discuss other related methods for defining fluctuations, we should mention that there have been strong claims that wavelets are indispensable for analyzing multifractals, e.g. (Arneodo et al., 1999). Inasmuch as the traditional definition of fluctuations as first differences is already a wavelet, this is certainly correct. However, for most applications,

DRAFT

71

1101

1102

11031104

1105

1106

1107

1108110911101111111211131114111511161117111811191120

1121112211231124112511261127112811291130113111321133113411351136

72

Page 37: 5gang/ftp.transfer/final.book.nofigs.Feb.20…  · Web view5.2.4 Direct empirical estimation of c(g): the probability distribution multiple scaling (PDMS) technique. In chapter 3,

5/18/23 Chapter 5 37

one doesn’t need the full arsenal of wavelet techniques. At a more fundamental level however, the claim is at least debatable since mathematically, wavelet analysis is a species of functional analysis, i.e. its objects are mathematical functions defined at mathematical points. On the contrary, we have seen that the generic multifractal process – cascades - are rather densities of singular measures so that strictly speaking, they are outside the scope of wavelet analysis. Although in practice the multifractals are always cut-off at an inner scale, when used as a theoretical, mathematical tool for multifractal analysis, wavelets may therefore not always be appropriate. A particular example where wavelets may be misleading is the popular “modulus maximum” technique where one attempts to localize the singularities (Bacry et al., 1989), (Mallat and Hwang, 1992). If the method is applied to a cascade process then as one increases the resolution, the singularity localization never converges so that the significance of the method is not as obvious as is claimed and its utility is questionable.

Figure 5.35b here

5.6.2 Generalized Structure Functions and Detrended Multifractal

Fluctuation Analysis

We have seen (eq. 83) that the statistical moments of the fluctuations lead to the exponent x(q) = qH - K(q); in the special case where the fluctuations are defined by the first differences, this is the usual qth order structure function (when q ≠ 2, this is sometimes called the “generalized structure function”). While these have the advantage that they can be used to directly estimate H (= x(1)); they are often not optimal for estimating C1, a. This is because often H is much larger than C1 so that for low order moments, the qH term in x(q) is much larger in magnitude than the K(q) term (and at “large” q, e.g. q >≈3-4), the moments become spurious due to the multifractal phase transitions discussed above. In both cases, trying to estimate K(q) using x(q) and the equation K(q) = qH -x(q) can lead to large errors since the magnitude of K(q) may be of the same order as the uncertainty in x(q). The moment method applied to the turbulent flux (“trace moments”) as in ch. 4 gives better accuracy for C1, a by removing the qH term by taking the absolute fluctuations at the smallest scale. Nonetheless, the universality parameters can be estimated from x(q) in a variety of ways; for example using x(1) = H, x’(1) = H-C1 and x’’(1) = aC1.

An interesting example of the use of structure functions is shown in fig. 5.36a, b that complement the flux analyses of chapter 4 (fig. 4.6a, b). In fig. 5.36a we show the corresponding results for the longitudinal and transverse wind and pressure (i.e. the wind components parallel and orthogonal to the aircraft direction respectively). Recall that these are the fields most strongly affected by the aircraft motion; especially the pressure (the aircraft attempted to fly on an isobar). For the reasons discussed in ch. 2 - the slopes of the isoabars - the behaviour for the wind has essentially two scaling regimes with a transition at about 40 km - unsurprisingly - the pressure has poor scaling. Contrast this with fig. 5.36b where we show the corresponding plot for those fields less affected by the trajectory fluctuations, the temperature, humidity and log potential temperature; we see that the scaling is indeed very good. These analyses were used to estimate the corresponding H and C1 values given in table 4.4 and fig. 5.36c shows the corresponding x(q) functions. Notice that the curvature is small since H ≈ 0.5 while C1 ≈ 0.05. We return to discuss these results in more detail in section 6.4.1.

DRAFT

73

11371138113911401141114211431144114511461147114811491150

1151

1152

115311541155115611571158115911601161116211631164116511661167116811691170117111721173117411751176117711781179

74

, 11/01/11,
As we discussed, the DTM does not seem so wonderful and the MFDFA is essentially a variant on waveltes
Daniel Schertzer, 17/12/10,
not clear it is well placed here, because rather distractinf from the main questions, e;g. DTM has not yet been addressed
Page 38: 5gang/ftp.transfer/final.book.nofigs.Feb.20…  · Web view5.2.4 Direct empirical estimation of c(g): the probability distribution multiple scaling (PDMS) technique. In chapter 3,

5/18/23 Chapter 5 38

Figure 5.36 a,b,c here

We now discuss a variant method that defines the fluctuations in a different way, but still quantifies the statistics in the manner of the generalized structure function by assuming that the fluctuations are stationary; the “MultiFractal Detrended Fluctuation Analysis”; MFDFA, (Kantelhardt et al., 2001), (Kantelhardt et al., 2002); this is a straightforward generalization of the original “Detrended Fluctuation Analysis” (Peng et al., 1994). The method only works for 1-D sections, so consider the transect v(x) of series on a regular grid with resolution = 1 unit, N units long. The MFDFA starts by replacing the original series by the running sum:

V x( )= v ′x( )′x =1

x

As discussed above, a sum is a finite difference of order -1 so that analyzing V rather than v allows the treatment of transects with H down to -1 (we will discuss the upper bound shortly). We now divide the range into NDx = int(N/Dx) disjoint intervals, each indexed by i = 1, 2, …, NDx

(“int” means “integer part”). For each interval starting at x = iDx, one defines the “nth order” fluctuation Fn(x = iDx,Dx) = DV in the running sum of v as follows:

Fn x =iDx,Dx( )=DV =1Dx

V i−1( )Dx + φ( )−pn,i φ( )( )2

φ=1

Dx

∑⎡⎣⎢

⎤⎦⎥1/2

; Dv=Fn

Dx=DVDx

where pn,n(x) is the nth order polynomial regression of V(x) over the ith interval of length Dx. It is the polynomial that “detrends” the running sum V; the fluctuation DVn is the root mean square deviation from the regression. The F is the standard DFA notation for the “fluctuation function”, as indicated, it is in fact a fluctuation in the integrated series DV. In terms F we have Fn= DV = DvDx so in terms of the usually cited DFA fluctuation exponent a (not the Levy a!) we have Fn ≈ Dxa and a =1+H so that the (second order) spectral exponent b = 2a-1-K(2) so that the oft-cited relation b = 2a-1 ignores the intermittency correction K(2). The fluctuation Dv is very close to the nth order finite difference of v, it is also close to a wavelet based fluctuation where the wavelet is of nth order, although this - like many MFDFA results - is not completely rigorous (see however (Kantelhardt et al., 2001; Taqqu et al., 1995)). As a practical matter, in the above sum, we start with intervals of length Dx = n+1 and normalize by the number of degrees of freedom of the regression, i.e. by Dx-n not by Dx; formula 97 is the approximation for large Dx.

In the usual presentation of the MFDFA, one considers only a single realization of the process and averages powers of the fluctuations over all the disjoint intervals i. However, more generally, we can average over all the intervals and realizations to obtain the statistics:

DVn Δx( )q 1/q

= Δv Δx( )q 1/q

Δx = Δxh q( ); h(q) = 1+ ξ(q) / q

where we have used the exponent h(q) as defined in (Davis et al., 1996) and (Kantelhardt et al.,

2002) and the relation Dvq ≈ Δxξ q( )

. From our analysis above, we see that the nth order MFDFA exponent h(q) is related to the usual exponents by:

DRAFT

75

1180118111821183118411851186118711881189

1190

119111921193119411951196

1197

119811991200120112021203120412051206120712081209121012111212

1213

1214

12151216

76

Daniel Schertzer, 17/12/10,
there are several notation problems , espiccially v(x) =DV(1) is unconsistent
, 11/01/11,
OK?
Daniel Schertzer, 17/12/10,
check agin the coherenc od the formaul (see correction ton the paper copy)
Page 39: 5gang/ftp.transfer/final.book.nofigs.Feb.20…  · Web view5.2.4 Direct empirical estimation of c(g): the probability distribution multiple scaling (PDMS) technique. In chapter 3,

5/18/23 Chapter 5 39

h q( )=1+ x q( ) / q = 1+ H( )−K q( ) / q; −1< H < n

(the 1 appears because of the initial integration of order 1 in the MFDFA recipe, the method works up to H = n because the fluctuations are defined as the residues with respect to nth order polynomials of the sum V corresponding to n+1th order differences in V and hence n-1th order differences in v). The basic justification for defining the MFDFA exponent h(q) in this way is that in the absence of intermittency (i.e. if K(q) = constant = K(0) = 0), one obtains h(q) = H = constant. Unfortunately - contrary to what is often claimed - the nonconstancy of h(q) - neither implies that that the process is multifractal nor conversely does its constancy imply a monofractal process. Indeed, it suffices to consider the (possibly fractionally integrated, order H) monofractal b model where h(q) = 1+H- C1+C1/q which is not only non constant but even diverges as q approaches zero. The same divergence occurs for any universal multifractal with a<1!

During the last ten years, the MFDFA technique has been applied frequently to atmospheric data, of special note is the work on climate by A. Bunde and co-workers (e.g. (Koscielny-Bunde et al., 1998), (Kantelhardt et al., 2001), (Bunde et al., 2002), (Bunde et al., 2005), (Lennartz and Bunde, 2009) and see ch. 10). However, the exact status of the MFDFA method is not clear since a completely rigorous mathematical analysis has not been done (especially in the multifractal case), and the literature suffers from unnecessary and misleading claims to the effect that the MFDFA is necessary because by its construction it removes nonstationarities due to trends (linear, quadratic etc. up to polynomials order n-1) in the data. However, the removal of these trends only accounts for these rather trivial types of nonstationarity: the method continues to make the standard (and strong) stationarity assumptions about the statistics of the deviations which are left over after linear or polynomial detrending. In particular it does nothing to remove the most common genuine type of statistical nonlinearity in atmospheric science: the diurnal and annual cycles which still strongly break the scaling of the MFDFA statistics. Second, the corresponding wavelet or finite difference definition of fluctuations can also easily take into account of these polynomial trends (see appendix 5E for the Haar wavelet, structure function and generalizations, and ch. 10 for many applications), and thirdly, the method removes trends at all scales and locations so that this emphasis on detrending is misleading (recall that that the FIF model is strictly statistically stationary i.e. it is statistically translationally invariant over the entire range over which it is defined yet it has random trends at all scales and all orders). In other words, the MFDFA is a variant with respect to some of the wavelet methods, in particular with respect to the Haar wavelet and generalizations discussed in appendix 5E) and has the disadvantage that while it yields accurate exponent estimates, the interpretation of the fluctuations are not straightforward. In comparison, the usual difference and the (new) tendency fluctuations and structure functions have simple physical interpretations in terms of magnitudes of changes (when 1>H>0) and magnitudes of average tendencies (when -1<H<0).

5.6.3 The Double trace moment technique

Above, we reviewed the results of multifractal analysis techniques which in principle could be applied to arbitrary multifractals. They enjoyed the apparent advantage of making no assumptions about the type of multifractal being analyzed. In practice however, the techniques are overly ambitious: for a finite (and usually small) number of samples of a process, they

DRAFT

77

1217

1218121912201221122212231224122512261227122812291230123112321233123412351236123712381239124012411242124312441245124612471248124912501251125212531254

1255

1256125712581259

78

, 11/01/11,
Perhaps Yulia might comment on this section?
Page 40: 5gang/ftp.transfer/final.book.nofigs.Feb.20…  · Web view5.2.4 Direct empirical estimation of c(g): the probability distribution multiple scaling (PDMS) technique. In chapter 3,

5/18/23 Chapter 5 40

attempt to deduce an entire exponent function (an infinite number of parameters), with the result that there is considerable uncertainty in the resulting estimates of c(γ) or K(q). With the realization that physically observable multifractals are likely to belong to universality classes, it is natural to develop specific methods to directly estimate the universality parameters (H, C1, a). These parameters can then be used to determine c(γ), K(q) from the expressions for universal multifractals.

The double trace moment (DTM) technique directly exploits universality by generalizing the usual moment method (the trace moments) based on a unique exponent q, which is the order of the moment; it introduces a second exponent h by taking the high resolution field to the power h. The basic idea is to use the fact that whereas K(q) is not a pure power law in q (due to the extra linear term), the corresponding formula for the normalized h power is a pure power in the exponent h: K(h,q) = K(qh)-hK(q) = haK(1,q) (eqs. 4.12, 4.13). The method thus works by taking the fluxes at the finest resolution and then raising them to a series of powers h; these can conveniently be taken as uniformly spaced in their logs. Then h powers of the fields are then degraded in the usual way and at each of a series of scales, a fixed qth power is taken (q = 0.5, 2 are typical values which are used). We can then estimate K(h,q) and repeat for a series of h values, finally we can estimate a from a plot of logh versus logK. The intercept, (=log K(1,q) with K(1,q) = C1(qa-q)/(a-1)) can then be used to determine C1. The whole process can then be repeated with another value of q and the results compared and averaged if needed. Fig. 5.37 a, b, c shows an example of hot wire anemometer wind data at 2 kHz in the atmosphere yielding the parameters a ≈1.50±0.05, C1 =0.25±0.05 (for the energy flux e; see fig. 2.4 for a sample of the wind component data). At the far right of the figure (corresponding to large qh), the curve flattens; this is because of the multifractal phase transition, for large q, K(q) becomes linear so that K(h,q) = K(qh)-hK(q) becomes constant.

Figure 5.37 a,b,c hereFigure 5.37a,b,c here

5.7 Summary of emergent turbulent laws in Chapter 5

We introduced the codimension formalism; c(γ):

f λq = λ K q( ) ↔ Pr ϕ λ > λ γ( ) ≈ λ −c γ( )

that characterizes how the probability distributions change with scale ratio λ. It is related to the moment scaling function K(q) introduced in ch. 3 by a Legendre transformation:

K q( )=m axγ

qγ−c γ( )( )

c γ( )=maxq

qγ−K q( )( )

This implies one- to – one relations between orders of singularities and moments:

qγ = ′c γ( )

γq = ′K q( )

DRAFT

79

12601261126212631264126512661267126812691270127112721273127412751276127712781279128012811282128312841285

1286

1287

1288

12891290

1291

1292

1293

80

Page 41: 5gang/ftp.transfer/final.book.nofigs.Feb.20…  · Web view5.2.4 Direct empirical estimation of c(g): the probability distribution multiple scaling (PDMS) technique. In chapter 3,

5/18/23 Chapter 5 41

For the Levy generator “universal multifractal” cascade, (0≤a<1, 1<a≤2, for a =1, take the limit a->1) this leads to:

K q( )=C1

a −1qa −q( )

c γ( )=C1

γC1 ′a

+1a

⎛⎝⎜

⎞⎠⎟

′a

; 1 / ′a +1 / a =1

The dressed cascade properties (those of a cascade developed over a wide range of scales and then spatially averaged) generally have power law probability distributions, a kind of Self-Organized Critical (SOC) behaviour:

⟨eλ(d )q ⟩= ∞, q ≥ qD ⇔ Pr(ελ (d ) > s) ~ s−qD , s >> 1

Universal multifractal cascade processes which are continuous in scale (“infinitely” divisible”) can be constructed, using the “Fractionally Integrated Flux” (FIF) model:

G =C11/α I1/α γ α( ) = C1

1/α γα ∗ r − D−1/α( ) ; ϕ = eΓ

f =IH f( )=f ∗r− D −H( )

where γa is a unit i.i.d. Levy noise with index a and IH indicates a fractional integration of order H.

The FIF yields f with statistics:

Sq Dr( )= Dfq = Dr x q( ) ; x q( )=qH −K q( ); K q( )=C1

a −1qa −q( )

where Sq is the qth order structure function.

Appendix 5.A: Divergence of High Order Statistical Moments

In section 5.3.2 we gave an outline of the basic argument showing that while the bare moments converge for all moments q, any finite λ, that the dressed moments generally diverge for q≥qD. However, in the interest of simplicity and brevity, we skipped a few nontrivial steps.

First, recall the definition of the λ resolution flux:

Pλ A( ) = ελd D xA∫

Since we are interested in the statistics of the dressed and partially dressed density: eλ,Λ d( ) = ΠΛ Bλ( ) / vol Bλ( ) we will consider the mean of the qth power of the flux on the set A (dimension D) of the cascade constructed down to the scale L/λ :

DRAFT

81

12941295

1296

129712981299

1300

13011302

1303

1304

130513061307

1308

1309

1310

1311131213131314

1315

1316

13171318

82

, 14/07/11,
The order of epsilon and dDx has changed in the integrals, this looks confusing…
Daniel Schertzer, 17/12/10,
I believe that this ia carbon copy of the section of the lecture notes? Therefore I did not check it
Page 42: 5gang/ftp.transfer/final.book.nofigs.Feb.20…  · Web view5.2.4 Direct empirical estimation of c(g): the probability distribution multiple scaling (PDMS) technique. In chapter 3,

5/18/23 Chapter 5 42

λq∏ (A) =

A∫ Dd x ελ ⎡ ⎣

⎤ ⎦q

when q is an integer ≥1:

A∫ Dd x eλ ⎡ ⎣

⎤ ⎦q

=A∫ ⋅⋅⋅

A∫ Dd x1⋅⋅⋅ Dd xq eλ(x1)⋅⋅⋅eλ(xq )

The complexity of this multiple integral suggests the introduction of “trace moments” which are obtained by integrating over the subset of the integral obtained by taking x1 =x2 =x3 = ;

TrA (eλ )q =

A∫d qDx eλq

~Aλ

∑ eλq λ−qD =

∑λK(q)λ−qD

where Aλ is the set A at resolution λ (i.e., obtained by a disjoint covering of A with balls Bλ, eλ is the usual (bare) flux density at resolution λ). Since eλ ≥0 and we integrate over a subspace (using the fact that 1/ q

(∑ qx i ) is a decreasing function of q) we have that the trace moments are bounds on the usual moments corresponding to a Jensen’s inequality:

qλ∏ (A) ≥TrA

qeλ (q>1)

≤TrAqeλ (q< 1)

The use of trace moments rather than the usual moments has a number of advantages. First, it is defined for all q (whereas the usual moments can only be expanded as multiple integrals for positive integer q). Second, trace moments are Hausdorff measures since we can use the scaling of ⟨ qελ ⟩ to obtain a Hausdorff measure over a higher dimensional space (for convenience we have left out the inf, etc.). We anticipate that in the limit λ → ∞ they will either diverge to ∞ or converge to 0; in fact, they will have two transitions!

To see this, use box counting in the sum ∑Aλ ; there will be Dλ terms, each of value ⟨ qελ ⟩ −qDλ :

TrAλqeλ = Dλ ⋅ K(q)λ ⋅ −qDλ = K(q)−(q−1)Dλ = (q−1)(C(q)−D)λ

where we have used the dual codimension function C(q) (section 3.2.7). We recall that C(q) is monotonic in q (see fig. 3.11). Consider first the case where C(q)>D for q<1. Due to the monotonicity of C(q) this is equivalent to C1>D. In this case,

limλ→ ∞

TrAλqeλ → 0 ⇒ q

∞∏ (A) =0

DRAFT

83

1319

1320

1321

132213231324

1325

1326

1327132813291330

1331

1332

13331334133513361337133813391340

1341

134213431344

1345

84

Page 43: 5gang/ftp.transfer/final.book.nofigs.Feb.20…  · Web view5.2.4 Direct empirical estimation of c(g): the probability distribution multiple scaling (PDMS) technique. In chapter 3,

5/18/23 Chapter 5 43

for all q<1, hence the process is degenerate on the space. This implies that when C1>D, then the mean of the bare process is too sparse to be observed in the space D; in fact, the above shows

that if C1>D it is impossible to normalize the process so that the dressed mean P∞(A) is finite.Now consider the nondegenerate case C1<D. In this case, the trace moments diverge for

q<1, but this does not affect the convergence of the dressed moments (the trace moments are upper bounds here). On the other hand, for q>1, we find:

limλ→ ∞ λATr qeλ → ∞ ⇒ ∞∏ (A) → ∞

for all C(q)>D. Using the implicit definition of qD: C(qD) = D, we thus obtain:

P∞(A) → ∞; q > qD

i.e., in this case, divergence of the trace moments implies divergence of the corresponding dressed moments. This is the key result needed to complete the demonstration in section 5.3.2.

Appendix 5.B: Continuous in-scale cascades: the autocorrelation and

finite size effects

5.B.1 The internal cascade structure

Various numerical details and examples of simulations of continuous in-scale isotropic (self-similar) multifractals, were given in (Schertzer and Lovejoy, 1987) and (Wilson et al., 1991) and earlier in this chapter. (Marsan et al., 1996) showed how to extend the framework to causal (space-time) processes, and (Pecknold et al., 1996), (Pecknold et al., 1997) discusses extensions to anisotropic multifractal processes needed for example to take into account atmospheric stratification. As we detailed in section 5.5, the basic method for simulating G is to fractionally integrate a Lévy noise, i.e. to convolute it with a singularity. While this method works for large enough scale ranges (i.e. it yields simulations with statistics satisfying eλ

q = λ K q( )

with K(q) of the universal form (eq. 33) for large enough λ), there are significant deviations at small scales (“finite size effects”), especially for a>1 which is the most empirically relevant range (there are also “large scale effects” which are typically less important). In order to get an idea of the importance of these deviations, consider the theoretical form of the normalized autocorrelation function Rλ Dr( ) for isotropic cascades:

Rλ Dr( )=eλ r( )eλ r−Dr( )

eλ2

= Dr−K 2( ) ; Dr ≥1

(as derived on discrete in-scale cascades in eq. 3.28), x, D r are D dimensional position vectors and lags respectively. The above uses the convention that the process is developed over the range of scales from λ down to 1 unit so that Rλ(1) = 1 (eq. 3.28); in this appendix, since we are

DRAFT

85

13461347

1348134913501351

1352

1353

1354

13551356

1357

1358

1359

13601361136213631364136513661367

13681369137013711372

1373

137413751376

86

Page 44: 5gang/ftp.transfer/final.book.nofigs.Feb.20…  · Web view5.2.4 Direct empirical estimation of c(g): the probability distribution multiple scaling (PDMS) technique. In chapter 3,

5/18/23 Chapter 5 44

concerned with the numerical simulation of cascades over a fixed range of scales L, we take 1 unit = 1 pixel so that in 1-D, the L resolution cascade is L pixels long so that 1≤ λ ≤L. K(2) appears because the autocorrelation is a q = 2 order statistic). Since the spectrum is the Fourier transform of the autocorrelation function we find that for wavenumber k = k , the power spectrum E(k) of eλ has the scaling form given by eq. 3.29. The discussion here is a summary of key sections of (Lovejoy and Schertzer, 2010a), (Lovejoy and Schertzer, 2010b).

Fig. 5B.1 shows the power spectrum of the process for a = 2 compensated by dividing by the theoretical k-b so that the theoretically expected spectral scaling leads to horizontal straight lines on the log-log plots. From the figure, we can see that the pure (power law) fractional integration takes nearly a factor of 103 in-scale to completely converge (the largest wavenumber for a 214 point long series is k = 213 ≈ 104); see (Lovejoy and Schertzer, 2010b) for similar results for smaller a values. Although the rate of convergence improves as a decreases from 2 deviations are noticeable for all a. Also shown in the figure are the corresponding spectra of the processes obtained using the improved simulation techniques described in (Lovejoy and Schertzer, 2010a), (Lovejoy and Schertzer, 2010b). It can be seen that although the spectral results are still not perfect, that they are significantly better. Although the biases in the individual realizations are not very visible to the eye, the statistics as revealed by various scale by scale analyses are still quite biased at small scales.

Figure 5B.1 here

A simple way to examine the scaling properties of realizations of cascades developed over a finite scale ratio L is to consider two point statistics such as autocorrelation functions, or their Fourier transforms, spectra. To calculate these, we first consider the qth power of the product of the random variables:

eλ ′r( )ε λ ′r − Δr( )( )q

= eq Γ ′r( )+Γ ′r −Δr( )( ) = eqC1

1/α g ′r − ′′r( )+g ′r −Δr − ′′r( )( )γ ′′r( )d D ′′r1≤ r ≤λ

(c.f. eqs. 73, 74); we see that G ′r( ) + Γ ′r − Δr( ) is the generator of the autocorrelation. For the statistics we can define the second characteristic function (SCF) of the generator by taking ensemble averages of the above:

log eλ ′r( )eλ ′r −Dr( )( )q

=λoγ eq G ′r( )+G ′r −Dr( )( ) =C1

a −1qaS Dr( )

the entire expression is the full SCF of the log of the autocorrelation, the key function S(D r ) defined by eq. 125 is its spatial part. Note that below we do not need the more complex full two point SCF log eλ

q1 r1( )eλq2 r2( ) . Using the statistical translational invariance of the process (by

construction the noise “subgenerator” ga(r) is statistically independent of r), we can take x’ = 0 to obtain:

S Dr( )= γλ − ′′r( )+ γλ − ′′r −Dr( )( )ad ′′r

1≤ ′′r ≤λ∫ = γλ r( )+ γλ r−Dr( )( )

adr

1≤r≤λ∫

where in the far right we have taken r = -r’’ and used the fact that the domain of integration (but not necessarily g) is invariant under inversion. We have considered isotropic multifractals in a d

DRAFT

87

1377137813791380138113821383138413851386138713881389139013911392139313941395139613971398139914001401

1402

140314041405

1406

14071408140914101411

1412

14131414

88

, 21/01/11,
Note in conformity with the new notation earlier, we remover the ND factors here
Daniel Schertzer, 17/12/10,
The type of rexpression could be used for othere part of the text
Page 45: 5gang/ftp.transfer/final.book.nofigs.Feb.20…  · Web view5.2.4 Direct empirical estimation of c(g): the probability distribution multiple scaling (PDMS) technique. In chapter 3,

5/18/23 Chapter 5 45

– dimensional space; for anisotropic multifractals we must replace the vector norm in the above by the corresponding scale function (see (Pecknold et al., 1996; Pecknold et al., 1993) for anisotropic simulations; the extension of the present technical discussion to anisotropic cascades will be made elsewhere). We note that from its definition (and assumed statistical translational invariance, i.e. independence from r ’ in eq. 126), the autocorrelation function (and hence S) is symmetric under inversion: S Dr( )= S −Dr( ) . Isotropic processes are symmetric under rotation in which case S is simply a function of the vector norm Dr . In this case comparing eqs. 123, 125 we see that the theoretical S Dr( ) is:

Stheory Dr( )=Nda −1C1

λoγ e r( )e r−Dr( ) =−Nd 2a −2( )λoγ Dr + Nd 2a λoγλ

Where we have used log eλ2 =

C1

a −1( )2a λoγλ ; the corresponding logλ term is absent in the

normalized autocorrelation function. We now consider how closely Stheory Dr( ) (eq. 127) is

approximated by S Dr( ) (eq. 126).

5.B.2 The general D = 1 case

For simplicity, we start with the problem in one dimension. Due to the inversion symmetry of S(D r ) noted above, in 1-D it is sufficient to consider Dx >0 (x is a coordinate in 1-D). The general d = 1 case is:

S Dx( )= γ x( )+ γ x−Dx( )( )a dx

1< x <λ∫

= γ x( )+ γ x−Dx( )( )a+ γ −x( )+ γ −x−Dx( )( )

a⎡⎣ ⎤⎦dx1

λ

The two basic cases of interest are the symmetric acausal case with g(x) = g(-x), and the causal case with g(x) = 0 for x < 0; see ch. 9 and (Marsan et al., 1996). The corresponding condition for a space-time process in d+1 dimensions to be causal is g(r,t) = 0, t<0 where r is a d-dimensional (spatial) vector (this is equivalent to multiplying an acausal g(r,t) by a Heaviside function Q(t) such that Q(t) =0 for t<0, Q(t) =1 for t ≥0, see (Lovejoy and Schertzer, 2010b) for numerical implementations).

For the symmetric acausal case we have:

S Dx( )= γ x( )+ γ x−Dx( )( )a+ γ x( )+ γ x + Dx( )( )

a⎡⎣ ⎤⎦dx1

λ

We have already seen that for this case with g a truncated power law given in eq. 77 the normalization factor Nd =2.

For the causal case, since g(x) = 0 for x < 0, we have:

S Dx( )= γ x( )+ γ x−Dx( )( )a dx

1

λ

DRAFT

89

14151416141714181419142014211422

1423

1424

1425

1426

1427

142814291430

1431

1432143314341435143614371438

1439

144014411442

1443

90

Page 46: 5gang/ftp.transfer/final.book.nofigs.Feb.20…  · Web view5.2.4 Direct empirical estimation of c(g): the probability distribution multiple scaling (PDMS) technique. In chapter 3,

5/18/23 Chapter 5 46

Here it is easy to see that the corresponding normalization factor is simply Nd =1. Using these definitions and expansions, (Lovejoy and Schertzer, 2010a) show:

S Dx( )=C a,d( )−Nd 2a −2( )λoγDx−2a λoγλ + 2a 2

dDx−d /a⎡

⎣⎢⎤⎦⎥

−O Dx−2( )+O Dx−2m −nd /a( )+ODxλ

⎛⎝⎜

⎞⎠⎟2

where C = 2Nd(B-+B+) is constant with respect to Dx, λ and is unimportant; n and m are positive integers and d = 1, N1 = 2. The leading order expression for S is thus = -Nd(2a-2) (logDx)+ Nd2alogλ; i.e. the same as Stheory (eq. 127) with the leading Dx dependent correction

−2Ndα 2

dΔx−d /α .

Using the above expansion for S(Dx) we can obtain the following expression for the autocorrelation function:

e x( )ε x − Δx( ) ∝ Δx−K 2( ) exp −2C1

dα 2

α −1Δx−d /α⎛

⎝⎜⎞⎠⎟

where we have only kept the (Dx-d/a) correction to the leading power law term. To gauge the importance of this correction, note that for Dx as large as 100 it can still be a 10% effect.

Having shown the origin of the problem, (Lovejoy and Schertzer, 2010b) show how to remove the leading order correction; this leads to the improved statistics shown in fig. 5B.1. Also in the next subsection, we discuss some technical details about the simulations of causal cascades and cascades in higher dimensional spaces. See also ch. 9 for some space-time examples and discussion of causal versus acausal processes. In appendix 5C we give a Mathematica code to generate the corrected causal and acausal simulations in 1, and 2D.

Appendix 5B.3 Some practical (numerical) considerations

Before making numerical simulations, there are a few practical points we should mention. Numerically, it very advantageous to use transform-based convolution routines and these are periodic. In all our simulations, we therefore used periodic convolutions. Since the noise γ(x) is statistically invariant under translations along the x axis, this will not break the translational invariance, however it does mean that the left half of the simulation is statistically dependent on the right half in a way which is artificially strong (due to the constraint that the resulting e(x) is periodic). If absolutely needed, zero padding could be used to avoid the periodicity, or - almost equivalently - only one half of the simulation could be used. One consequence is that the

periodic eλ ,λ h( ) is somewhat modified with respect to its nonperiodic value. Note that although

the Fourier filter corresponding to the convolution with x −d /a

is of the type k −d / ′a

(where a’ is the auxiliary variable, 1/a’+1/a =1), there is trigonometric prefactor, whose sign is quite important for preserving the extremality of the of the subgenerator in the physical space. Therefore, this prefractor function cannot be forgotten if one wants to proceed directly from Fourier space. It therefore simpler to proceed from the physical space or to directly compute

DRAFT

91

14441445

1446

144714481449

1450

14511452

1453

14541455145614571458145914601461

1462

14631464146514661467146814691470

1471

14721473147414751476

92

Page 47: 5gang/ftp.transfer/final.book.nofigs.Feb.20…  · Web view5.2.4 Direct empirical estimation of c(g): the probability distribution multiple scaling (PDMS) technique. In chapter 3,

5/18/23 Chapter 5 47

convolutions in physical space. While this issue is indeed important for the convolution kernel g of the cascade generator G = g*γ it is not necessarily important for the kernel of the observable I = x − d−H( ) ∗eG unless the latter is also a positive definite quantity such as a passive scalar concentration. On the contrary, it may be desirable that the observable be symmetric about zero; in this case one may take an antisymmetric convolution kernel for the fractional integration as x −1−H( ) sγn x( ) (in d =1; generalizations to higher d are obvious).

Appendix 5B.4 Extensions to causal and higher dimensional

cascades

In this subsection we proceed to evaluate and intercompare the statistical accuracy of the causal and acausal simulations, as well as the effect of passing from 1-D to 2-D. We use the Dx-d/a correction method for the small Dx corrections because it is simple to implement, works reasonably well and is mathematically well grounded. Also, to evaluate the simulations, we only consider the (compensated) spectra i.e. the q = 2 statistics rather than statistics of all orders. This is sufficient since we find: a) that the statistics are more accurate for the causal case than for the acausal case, b) as expected when going from d = 1 to d = 2 the convergence is improved, c) we only consider the case a = 2 since the corresponding corrections are the largest.

Consider first the causal 1-D simulations. Fig. 5B.2 a shows that even without any corrections (blue, 2nd from top) that the causal simulations have significantly smaller deviations from pure power law behaviour than the uncorrected acausal simulations (cyan); indeed they are even comparable to the corrected acausal spectra. With the Dx-d/a corrections (green, top) we see that the causal spectrum is nearly perfect. In comparison, the corrected acausal curve (pink, reproduced from fig. 5B.1) has significantly larger deviations.

Figure 5B2a, b here

Turning our attention to simulations in 2-D, we recall the examples of isotropic acausal realizations of the process in section 5.5. As discussed there, we expect the corrections to be smaller for d = 2 when compared to d = 1, fig. 5B.2 b confirms this, indeed, we see that the uncorrected 2-D (isotropic, i.e. angle integrated) spectrum is slightly better than the corrected 1-D spectrum, but in all cases the correction method makes the spectra significantly closer to the theoretical pure power form.

Since space and time are no longer symmetric, we made (x,t) causal simulations and evaluated compensated 1-D spectra in the spatial and temporal directions separately; see fig. 5B.3 and fig. 9.3 for a corresonding realization). We see that the temporal spectra are very close to those of the 1-D causal processes (fig. 5B2 a) – for both the corrected and uncorrected cases (green, blue respectively). Similarly, the spectra of the corrected and uncorrected cases in the spatial direction (pink and cyan respectively) have much larger deviations; very close to the acausal 1-D results (fig. 5B.2 a) and compare with the 2-D acausal spectra (fig. 5B.2 b). We conclude that the correction method works well independently of the dimension of the space and that for the causal extensions that the temporal statistics have significantly smaller deviations. This is presumably because of the sharp discontinuity introduced by the Heaviside function

DRAFT

93

14771478

147914801481

1482

1483

1484

148514861487148814891490149114921493149414951496149714981499150015011502150315041505150615071508150915101511151215131514151515161517

94

, 21/01/11,
YES!
Page 48: 5gang/ftp.transfer/final.book.nofigs.Feb.20…  · Web view5.2.4 Direct empirical estimation of c(g): the probability distribution multiple scaling (PDMS) technique. In chapter 3,

5/18/23 Chapter 5 48

which roughens the simulations along the time axis, thereby somewhat compensating for the otherwise overly smooth behaviour.

Figure 5B.3 here

Appendix 5C: A Mathematica code for causal and acausal multifractal

simulations

The Mathematica code reproduced in fig. 5C.1 produces conservative (H = 0) multifractals for both acausal and causal (ch.9) fields in one and two dimensions. It implements the Dx-d/a finite size effect correction scheme described (Lovejoy and Schertzer, 2010b) in both 1D and 2D for causal and acausal processes. Lévy generates random extremal Lévy variables attention is paid to avoiding underflows when exponentiating. The basic functions epsa, epsc generate the corrected multifractal acausal, causal simulations; λ is the length of the resulting vector, a, C1 are the basic parameters. Note that as mentioned in the text, the built-in function Convolve performs a periodic convolution (no zero padding was used).

Figure 5.C.1 here

Appendix 5D: Multifractal simulations on a sphere

In this appendix, we outline how to perform isotropic fractional integrals on a sphere using spherical harmonic expansions (extensions to anisotropic multifractals on a sphere are not as straightforward as in flat space, and we do not consider them here). We require the following addition formula for spherical harmonics:

Y ∗λm m,φ( )

m∑ Yλm m',φ'( )=

2λ+14p

Pλ Dm( )=2λ+14p

Yλ0 Dm,φ( )

Where Ylm is the spherical harmonic with principle order l (conjugate to the azimuthal angle q in spherical polar coordinates) and m is conjugate to the spherical polar angle f, Pl is the l order Legendre polynomial. Note that the right hand expression is the m = 0 harmonic, there is no f dependence, it is indicated purely for notational completeness. The Dm = cosq is not m-m', but is the cosine of the angle (q) between the directions defined by m, m' and f, f':

Dm =mm '+ 1 − μ 2 1 − μ '2 cos φ − φ'( )

We are interested in the following convolution:

I HU m,φ( )= q− 2−H( )U m',φ'( )dm'dφ'−1

1

∫0

2p

DRAFT

95

15181519152015211522

1523

1524

1525152615271528152915301531153215331534

1535

1536153715381539

1540

15411542154315441545

1546

1547

1548

96

Page 49: 5gang/ftp.transfer/final.book.nofigs.Feb.20…  · Web view5.2.4 Direct empirical estimation of c(g): the probability distribution multiple scaling (PDMS) technique. In chapter 3,

5/18/23 Chapter 5 49

which is a fractional integration of order H in a space dimension 2 (we have left out normalization factors). In the above, U is the function which we wish to fractionally integrate (i.e. our noise).

We now make the following expansion:

q− 2− H( ) = σ lYl ,0 Δμ,φ( )l

sl is the m =0 component of the full spherical harmonic expansion:

s lδm0 = θ− 2− H( )Ylm Δμ,φ( )−1

1

∫0

∫ dΔμdφ

Combining the expansion of q-(2-H) with the addition formula, we obtain:

q− 2− H( ) = σ l4π

2l + 1Yl ,m

∗ μ,φ( )l ,m∑ Yl ,m μ ',φ'( )

Combining this with the expansion for the noise U:

U m,φ( )= uλ,mYλ,m m,φ( )λ,m∑

We can now obtain an expression for the fractional integral:

I HU m,φ( )= uλmsλ'4p

2λ'+1Yλm m',φ'( )Yλ'm ' m,φ( )Y∗

λ'm ' m',φ'( )dm'dφ'−1

1

∫0

2p

∫λ,m ,λ',m '∑

or:

I HU m,φ( )= uλmsλ'4p

2λ'+1dλλ'dmm 'Yλ'm ' m,φ( )

λ,m ,λ',m '∑

hence:

I HU m,φ( )= sλ4p

2λ+1uλmYλm m,φ( )

λ,m∑

We thus see that the required filter is:

sl4π

2l +1

This is the filter required for the various fractional integrations needed for the FIF model on the sphere. See fig. 5.32 c for an example of a topography simulation.

DRAFT

97

1549155015511552

1553

1554

1555

1556

1557

1558

1559

1560

1561

1562

1563

1564

1565

1566

1567

15681569

98

Page 50: 5gang/ftp.transfer/final.book.nofigs.Feb.20…  · Web view5.2.4 Direct empirical estimation of c(g): the probability distribution multiple scaling (PDMS) technique. In chapter 3,

5/18/23 Chapter 5 50

Appendix 5E: Tendency, Poor Man’s, and Haar Structure

functions and the MFDFA technique

5E.1 An intercomparison using multifractal simulations

In section 5.6 we saw that we could characterize the statistics of fluctuations by their moments using structure functions. If the fluctuations at separation Dx are defined by differences (the “poor man’s” wavelet), then it is only when 0<H<1 that they are dominated by structures of size Dx (by wavenumbers ≈ Dx-1). When H is outside this range the fluctuations reflect the statistics of much larger (H>1) or much smaller (H<0) structures present in the sample, they are independent of Dx, they “saturate”. We have seen that in the weather regime, most geophysical H parameters are indeed in this range justifying the common use of the classical difference structure function. Even more commonly, we find H>0 implying that implying that fluctuations Df tend to increase with scale Df≈ DxH. The main exception is in the low frequency weather regime (at time scales >≈ 10 days), where we generally have H<0 so that fluctuations tend to decrease with scale. We noted that the range of H over which fluctuations are usefully defined could be changed by integration and/or differentiation corresponding to changing the shape of the defining wavelet, changing its real and Fourier space localizations. In the usual wavelet framework, this is done by modifying the wavelet directly e.g. by choosing the Mexican hat or higher order derivatives of the Gaussian, or by choosing them to satisfying some special criterion such as orthogonality. Following this, the actual convolutions are calculated using with Fast Fourier (or equivalent) numerical convolution techniques.

A problem with this usual implementation of wavelets is that not only do these convolutions make the determination of the fluctuations numerically cumbersome, but that at the end, the physical interpretation of the fluctuations is no longer clear. In contrast, when 0<H<1, the difference structure function gives direct information on the typical difference (q =1) and typical variations around this difference (q = 2) and even typical skewness (q =3) or typical Kurtosis (q =4) or - if the probability tail is algebraic – of the divergence of high order moments. Similarly, when -1<H<0 the tendency structure function directly quantifies the fluctuations’s deviation from zero and the exponent, the rate at which the deviation decreases by averaging to larger and larger scales. These poor man’s and tendency fluctuations are also very easy to directly estimate from series with uniformly spaced data and - with straightforward modifications - to irregularly spaced data (see e.g. section 6.4.2). While the corresponding wavelets are not orthogonal, when wavelets are used for statistical characterizations, this is generally unimportant.

The real drawback of the difference (poor man’s) and tendency structure functions is the limited range of H over which they are useful. In this appendix, we show how to extend this range without loosing the double advantages of simplicity of implementation and simplicity of interpretation.

Before considering the Haar wavelet, (section 5.6), let us recall the definitions of the difference and tendency structure functions. The difference/ poor man’s fluctuation is thus:

Dv Δx( )( )diff= δΔxv; δΔxv = v x + Δx( ) − v x( )

DRAFT

99

1570

1571

1572

15731574157515761577157815791580158115821583158415851586158715881589159015911592159315941595159615971598159916001601160216031604160516061607

1608

100

Page 51: 5gang/ftp.transfer/final.book.nofigs.Feb.20…  · Web view5.2.4 Direct empirical estimation of c(g): the probability distribution multiple scaling (PDMS) technique. In chapter 3,

5/18/23 Chapter 5 51

where d is the the difference operator. Similarly, in ch. 10 we use the “tendency fluctuation” it can be defined as:

Dv Δx( )( )tend

=T Δxv =1

Δx′v ′x( )

x< ′x < x+Δx∑

or, with the help of the summation operator s, equivalently by:

Dv Δx( )( )tend=

1Δx

δΔxs v '( ); sv ' = ′v ′x( )′x <x∑

Dv Δx( )( )tend has a straightforward interpretation in term of the mean tendency of the data but valid for -1<H<0 (the climate range, ch.10). It is also easy to implement: simply remove the

overall mean to yield ′v x( ) = v x( ) − v x( ) and then take the mean over intervals Dx (equivalent to the mean of the differences of the running sum).

We can now define the Haar fluctuation which is a special case of the Daubechies family of wavelets (see e.g. (Holschneider, 1995) and for a recent application, (Ashok et al., 2010)). This can be done by instead taking the second differences of the mean:

Dv Δx( )( )Haar= H Δxv =

2Δx

δΔx /22 s =

2Δx

s x( ) + s x + Δx( )( ) − 2s x + Δx / 2( )( )

=2

Δxv ′x( )

x+Δx /2< ′x < x+Δx∑ − v ′x( )

x< ′x <x+Δx /2∑⎡

⎣⎢⎤⎦⎥

Where H Dx is the Haar operator. Although this is still a valid wavelet (see fig. 5.35b), it is almost trivial to calculate and below, we shall see that the technique can be used for -1<H<1. When applied without the initial running sum to v(x) directly, it can be used for series with 0<H<2, see section 6.4.2). The numerical factor (2) is needed so that the Haar fluctuation is close to the usual differences when 0<H<1, see below.

From the definitions, it is easy to obtain the relation:

H Dx = 2T Δx /2δΔx /2 = 2δΔx /2T Δx /2

which will be useful for interpreting the Haar fluctuations. Since the difference operator removes any constants, the tendency operator could be replaced by an averaging operator so that eq. 148 means that the Haar fluctuation is simply the difference of the series that has been degraded in resolution.

If needed, the Haar fluctuations can easily be generalized to higher order n by using nth

order differences on the sum s:

H Dx

n( )v =n + 1( )Δx

δΔx / n+1( )n+1 s

We can note that the nth order Haar fluctuation is valid over the range -1<H<n; it is insensitive to polynomials of order n-1, although as we see below, for analyzing scaling series, this

DRAFT

101

16091610

161116121613

16141615

16161617

16181619162016211622

1623

162416251626162716281629

1630163116321633163416351636

163716381639

102

, 31/08/11,
Corrected defintion 31/11/8
, 31/08/11,
Corrected definition, 31/8/11
Page 52: 5gang/ftp.transfer/final.book.nofigs.Feb.20…  · Web view5.2.4 Direct empirical estimation of c(g): the probability distribution multiple scaling (PDMS) technique. In chapter 3,

5/18/23 Chapter 5 52

insensitivity has no particular advantage. Note that H Dx1( ) = H Δx and H Dx

0( ) ′v =T Δx ′v hence when

applied to a series with zero mean, H Dx0( ) =T Δx .

Consider for a moment the case n = 2 which defines the “quadratic Haar fluctuation”:

Dv Δx( )( )Haarquad

=3

Δxs x + Δx( ) − 3s x + Δx / 3( ) + 3s x − Δx / 3( ) − s x − Δx( )( )

this fluctuation is sensitive to structures of size Dx-1 – and hence useful - over the range -1<H<2 and it “filters” out polynomials of order 1 (lines); this is thus essentially equivalent to the quadratic MFDFA (Multi Fractal Detrended Fluctuation Analysis, section 5.6) technique that we numerically investigate below and numerically it is much faster still there are no time consuming regression calculations. Below, we only give some limited discussion of this n =2 “quadratic Haar” wavelet since for the most common series studied here with H<1, it was generally found to be very close to the usual Haar wavelet although with the slight disadvantage of converging a little more slowly for small Dx, and having a less straightforward interpretation. On the plus side, it was a bit more accurate at the very largest factor of two Dx, see the example in section 5E.3. We could also note that if H< -1, fluctuations could be usefully defined using higher order summing operators, but we do not need them here.

Before proceeding, we could make a general practical remark about these real space statistics: most lags Dx do not divide the length of the series (L) exactly so that there is a “remainder” part. This is not important for the small Dx<<L, for each realization there are many disjoint intervals of length Dx however when L/3≈<Dx<L the statistics can be sensitive to this since from each realizaiton there will only one or two segments and hence poor statistics. A simple expedient is to simply repeat the analysis on the reversed series and average the two results. This can indeed improve the statistics at large Dx when only one or a small number of realizations are available; it has been done in the analyses below.

Figure 5E1.a, b here

To have a clearer idea of the limitations of the various fluctuations for determining the statistics of scaling functions, we now numerically compare the performance of these various fluctuations and their corresponding structure functions when applied to the characterization of multifractals. In order to easily compare their performances with those of standard spectral analysis, we will only consider the second order (q = 2) structure functions and will use a=1.8, C1 = 0.1 simulations with H in the range -7/10<H<7/10 (the most common range for geodata). These parameters yield an intermittency correction K(2) = 0.18 for the q = 2 moment; 50 simulations were averaged to estimate the ensemble spectrum. The simulations were made using the technique described in appendix 5B for the flux (i.e. H = 0); for the first fractional integration (for the generator), the Dx-d/a correction method was used. For the second fractional integration (for H ≠ 0 to obtain the field from the flux) we used a Fourier space power law filter k-H. For this fractional integration, exponential cutoffs were used at high frequencies to avoid small scale numerical instabilities (this is standard for differentiation, i.e. when filtering by k-H with H <0, it is equivalent to using the “Paul” wavelet; see (Torrence and Compo, 1998)). The series were 216 in length and were then degraded in resolution by factors of 4 to 214 to avoid residual finite size effects in the simulation at the smallest scales. In order to avoid spurious correlations

DRAFT

103

1640

16411642

1643

164416451646164716481649165016511652165316541655165616571658165916601661166216631664166516661667166816691670167116721673167416751676167716781679168016811682

104

, 31/08/11,
Corrected definition: factor of 9 with respect to previous!
Page 53: 5gang/ftp.transfer/final.book.nofigs.Feb.20…  · Web view5.2.4 Direct empirical estimation of c(g): the probability distribution multiple scaling (PDMS) technique. In chapter 3,

5/18/23 Chapter 5 53

introduced by the periodicity of the simulations, the series were split in half so that the largest scale was 213.

In fig. 5E.1a we show sample of the simulations visually showing the effect of increasing H form -7/10 to 7/10. In fig. 1Eb, we see that as expected, the compensated spectra are nearly flat, although for the lowest and highest factors of about 100.5 ≈3 there were more significant deviations; the slopes of the central portions were thus determined by regression and the corresponding exponents are shown in fig. 5.E.4. In the figure, the spectra were averaged over logarithmically spaced bins, 10 per order of magnitude (with the exception of the lowest factor of 10 where all the values were used). We also performed the regressions using log-log fits using all the Fourier components (rather than just the averaged values) in the same central range, these gave virtually identical exponent estimates (the mean absolute differences in the estimated spectral exponent b were ≈ ±0.007), and were thus not shown. Similarly, if instead of using log-log linear regression, we use nonlinear regressions over the same range yielded identical exponents to within ±0.009, and again these estimates were too similar to be worth showing. From fig. 5.E.4, one can see that there seems to be a residual small bias of about -0.04 the origin of which is not clear.

Figures 5E.2, 3, 4 here

We next consider the Haar structure function which we compare to the second order, quadratic MFDFA structure function analogue (based on the MFDFA fluctuation DvMFDFA = F/Dx where F is the usual MFDFA scaling function, see eq. 103), the former is valid over the range -1<H<1, the latter over the range -1<H<2. We can see (fig. 5E.2) that the Haar structure function does an excellent job with overall bias most around +0.01 to 0.02, (fig. 5.E.4), and that the MFDFA method is nearly as good (overall bias ≈ -0.02 to -0.04). It is interesting to compare this in the same figure with the results of the tendency structure function (H<0) and the usual difference (poor man’s) structure function (H>0), fig. 5E.3. The tendency structure function turns out to be quite accurate, except near the limiting value H = 0 with the large scales showing the largest deviations. In comparison, for the difference structure function, the scaling is poorest at the smaller scales requiring a range factor of ≈ 100 convergence for H = 1/10 and a factor ≈ 10 for H = 3/10. The overall regression estimates (fig. 5.E.4) show that the biases for both increase near their limiting values H = 0 to accuracies ≈ ±0.1 (note that they remain quite accurate if one only makes regressions around the scaling part). This can be understood since the process can be thought of as a statistical superposition of singularities of various orders and only those singularities with large enough orders (difference structure functions) or small enough orders (tendency structure functions) will not be affected by the theoretical limit H = 0. The main advantage of the Haar structure function with respect to the usual (tendency or difference) structure functions is precisely that it is valid over the whole range -1< H <1 so that it is not affected by the H =0 limit.

5E.2 The theoretical relation between poor man’s, tendency and Haar

fluctuations, hybrid structure functions

We have seen that although the difference and tendency structure functions have the advantage of having simple interpretations in terms of respectively the average changes in the

DRAFT

105

168316841685168616871688168916901691169216931694169516961697169816991700170117021703170417051706170717081709171017111712171317141715171617171718171917201721

1722

1723

17241725

106

Page 54: 5gang/ftp.transfer/final.book.nofigs.Feb.20…  · Web view5.2.4 Direct empirical estimation of c(g): the probability distribution multiple scaling (PDMS) technique. In chapter 3,

5/18/23 Chapter 5 54

value of the process and its mean value over an interval, this simple interpretation is only valid over a limited range of H values. In comparison, the Haar and MFDFA fluctuations give structure functions with valid scaling exponents over wider ranges of H. But what about their interpretations? We now briefly show how to relate the Haar, tendency and poor man’s structure functions thus giving it a simple interpretation as well.

In order to see the connection between the fluctuations, we use the “saturation” relations:

dDxv=d

Ctendv; H < 0

T Δxv=d

Cdiff v; H > 0

where =d

indicates equality in the random variables in the sense of probability distributions and Ctend, Cdiff are proportionality constants. These relations were easily verified numerically and arise for the reasons stated above; they simply mean that for series with H<0, the differences are typically of the same order as the function itself and that for H>0, the tendencies are of the same order: the fluctuations “saturate”. By applying eqs. 145 to eq. 144 we now obtain:

H Dxv = 2T Δx /2δΔx /2v=d

2T Δx /2v=d

′CtendT Δxv; H < 0

H Δxv = 2δΔx /2T Δx /2v=d

2δΔx /2v=d

′Cdiff δΔxv; H > 0

where C’tend, C’diff are “calibration” constants (only a little different from the unprimed quantities – they take into account the factor of two and the change from Dx/2 to Dx). Taking the qth

moments of both sides of eq. 152, we obtain the results for the various qth order structure functions:

Dv( )Haarq

Δv( )diffq = ′Cdiff

q ; H > 0

Δv( )Haarq

Δv( )tendq = ′Ctend

q ; H < 0

This shows that at least for scaling processes that the Haar structure functions will be the same as the difference (H>0) and tendency structure functions (H<0), as long as these are “calibrated” by determining C’diff, and C’tend. In other words, by comparing the Haar structure functions with the usual structure functions we can develop useful correction factors which will enable us to deduce the usual fluctuations given the Haar fluctuations. Although the MFDFA fluctuations are not wavelets, at least for scaling processes, the same basic argument applies, since the MFDFA and usual fluctuations scale with the same exponents, they can only differ in their prefactors, the calibration constants.

To see how this works on a scaling process, we confined ourselves to considering the root mean square (RMS) S(Dt) = <Dv2>1/2; see fig. 5E.5a,b which show the ratio of the usual S(Dt) to those of the RMS Haar and MFDFA fluctuations. We see that at small Dx’s, the ratio stabilizes after a range of about a factor 10 – 20 in scale and that it is quite constant up to the extreme

DRAFT

107

172617271728172917301731

1732

17331734173517361737

17381739174017411742

1743174417451746174717481749175017511752175317541755

108

, 31/08/11,
Increased by a factor of 4 for new defintion 31.8.11
Page 55: 5gang/ftp.transfer/final.book.nofigs.Feb.20…  · Web view5.2.4 Direct empirical estimation of c(g): the probability distribution multiple scaling (PDMS) technique. In chapter 3,

5/18/23 Chapter 5 55

factor of 3 or so (depending a little on the H value). The deviations are largely due to the slower convergence of the usual RMS fluctuations to their asymptotic scaling form. From the figure, we can see that the ratios C’Haar = S/SHaar (fig. 5E.5a) and corresponding C’MFDFA = S/SMFDFA (fig. 5E.5b) are well defined in the central region, and are near unity, more precisely in the range ½< C’Haar <2 for the most commonly encountered range of H: -4/10<H<4/10. In comparison (bottom of the figure) over the same range of H, we have 0.23> C’MFDFA >0.025 so that the MFDFA fluctuation is quite far from the usual ones. For applications, one may use the semi-empirical formulae C’Haar ≈ 1.1 e-1.65H and C’MFDFA ≈ 0.075e-2.75H which are quite accurate over the entire range -7/10<H<7/10. Although these factors in principle allow one to deduce the usual RMS structure function statistics from the MFDFA and Haar structure functions this is only true in the scaling regime. Since the corrections for Haar fluctuations are close to unity for many purposes they can be used directly.

Figure 5E.5 here

5E.3: Hybrid fluctuations and structure functions: global monthly

surface temperatures: an example with both H<0, and H>0 regimes

For pure scaling functions, the difference (1>H>0) or tendency (-1<H<0) structure functions are adequate, the real advantage of the Haar structure function is apparent for functions with two or more scaling regimes one with H>0, one with H<0. Can we still “calibrate” the Haar structure function so that the amplitude of typical fluctuations can still be easily interpreted? For these, consider eqs.152 which motivate the definition:

H hybrid ,Dxv=max dDxv,T Dxv( )

of a “hybrid” fluctuation as the maximum of the difference and tendency fluctuations; the “hybrid structure function” is thus the maximum of the corresponding difference and tendency structure functions and therefore has a straightforward interpretation. The hybrid fluctuation is useful if a unique calibration constant Chybrid can be found such that:

H hybrid ,Dxv≈dChybridH Dxv

To get an idea of how the different methods can work on real data, we consider the example of the global monthly averaged surface temperature of the earth (discussed in detail in section 10.3). This is obviously highly significant for the climate but the instrumental temperature estimates are only known over a small number of years (here, the period 1881-2008; 129 years, three different series). Fig. 10.12 shows the comparison of the difference, tendency, hybrid and Haar RMS structure functions; the latter increased by a factor Chybrid = 100.35 ≈ 2.2. It can be seen that the hybrid structure function does extremely well; the deviation of the calibrated Haar structure function from the hybrid one is ±14% over the entire range of near a factor 103 in time scale. This shows that to a good approximation, the Haar structure function can preserve the simple interpretation of the difference and tendency structure functions: in regions where the logarithmic slope is between -1 and 0, it approximates the tendency structure function whereas in regions where the logarithmic slope is between 0 and 1, the calibrated Haar structure function approximates the difference structure function.

DRAFT

109

1756175717581759176017611762176317641765176617671768

1769

1770

17711772177317741775

17761777177817791780

178117821783178417851786178717881789179017911792179317941795

110

Page 56: 5gang/ftp.transfer/final.book.nofigs.Feb.20…  · Web view5.2.4 Direct empirical estimation of c(g): the probability distribution multiple scaling (PDMS) technique. In chapter 3,

5/18/23 Chapter 5 56

Before embracing the Haar structure function let us consider it’s behaviour in the presence of nonscaling perturbations; it is common to consider the sensitivity of statistical scaling analyses to the presence of nonscaling external trends superposed on the data which therefore break the overall scaling. Even when there is no reason to suspect such trends, the desire to filter them out is commonly invoked to justify the use of quadratic MFDFA or high order wavelet techniques which eliminate linear or higher order polynomial trends. However, for this purpose, these techniques are not obviously appropriate since on the one hand they only filter out polynomial trends (and not for example the more geophysically relevant periodic trends), while on the other hand, even for this, they are “overkill” since the trends they filter are filtered at all – not just the largest scales. The drawback is that with these higher order fluctuations, we loose the simplicity of interpretation of the Haar wavelet while obtaining few advantages. Fig. 5E.6a shows the usual (linear) Haar RMS structure function compared to the quadratic Haar and quadratic MFDFA structure functions. It can be seen that while the latter two are close to each other (after applying different calibration constants, see the figure caption), and that the low and high frequency exponents are roughy the same, the transition point has shifted by a factor of about 3 so that overall they are quite different from the Haar structure function, it is therefore not possible to simultaneously calibrate the high and low frequencies.

Figure 5E.6a here

If there are indeed external trends that perturb the scaling, these will only exist at the largest scales; it is sufficient to remove a straight line (or if needed parabola, cubic etc.) from the entire data set i.e. at the largest scales only. Fig. 5E.6b shows the result when the Haar structure function is applied to the (perfectly scaling) data that has been detrended in two slightly different ways: by removing a straight line through the first and last points of the series and by removing a regression line. Since these changes essentially effect the low frequencies only, they mostly affect the extreme factor of two in scale. If we discount this extreme factor 2, then in the figure we see that there are again two scaling regimes, the low frequency one is slightly displaced with respect to the Haar structure function, but not nearly as much as for the quadratic Haar. In other words, removing the linear trends at all scales as in the quadratic Haar or the quadratic MFDFA is too strong to allow a simple interpretation of the result and it is unnecessary if one wishes to eliminate external trends.

Figure 5E.6b here

In conclusion, if all that is required are the scaling exponents, the Haar structure function and quadratic MFDFA seem to be a excellent techniques. However, the Haar structure function has the advantage of numerical efficiency, simplicity of implementation, simplicity of interpretation.

5.8 References

A. Arneodo, N. Decoster and S.G. Roux, Intermittency, Log-Normal Statistics, And Multifractal Cascade Process In High-Resolution Satellite Images Of Cloud Structure, Physical Review Letters 83(1999), pp. 1255-1258.

DRAFT

111

17961797179817991800180118021803180418051806180718081809181018111812181318141815181618171818181918201821182218231824182518261827182818291830183118321833

1834

183518361837

112

Page 57: 5gang/ftp.transfer/final.book.nofigs.Feb.20…  · Web view5.2.4 Direct empirical estimation of c(g): the probability distribution multiple scaling (PDMS) technique. In chapter 3,

5/18/23 Chapter 5 57

V. Ashok, T. Balakumaran, C. Gowrishankar, I.L.A. Vennila and A. Nirmalkumar, The Fast Haar Wavelet Transform for Signal & Image Processing, , Inter. J. of Comp Sci, Info. Security 7( 2010), pp. 126-130.

A. Bacry, A. Arneodo, U. Frisch, Y. Gagne and E. Hopfinger, Wavelet analysis of fully developed turbulence data and measurement of scaling exponents. In: M. Lessieur and O. Metais, Editors, Turbulence and coherent structures, Kluwer (1989), pp. 703-718.

P. Bak, C. Tang and K. Weiessenfeld, Self-Organized Criticality: An explanation of 1/f noise, Physical Review Letter 59(1987), pp. 381-384.

S. Basu, E. Foufoula-Georgiou and F. Porté-Agel, Synthtic turbulence, fractal interpolation and large-eddy simulations, PhysicalReview E, 70(2004), p. 026310.

N. Bleistein and R.A. Handelsman, Asymptotic Expansions of Integrals, Dover, Mineola (1986).A. Bunde, J.F. Eichner, J.W. Kantelhardt and S. Havlin, Long-term memory: a natural

mechanism for the clustering of extreme events and anomalous residual times in climate records, Phys. Rev. Lett. 94(2005), p. 048701.

A. Bunde, J. Kropp and H.J. Schellnhuber, Editors, The science of disasters: Climate disruptions, heart attacks and market crashes Springer (2002).

L.E. Calvet and A.J. Fisher, Forecasting multifractal volatility, J. of Econometrics 105(2001), pp. 27-58.

L.E. Calvet and A.J. Fisher, Multifractal Volatility" theory, forecasting and pricing, Academic Press, New York (2008) 258 pp.

Y. Chigirinskaya, D. Schertzer, S. Lovejoy, A. Lazarev and A. Ordanovich, Unified multifractal atmospheric dynamics tested in the tropics Part 1: horizontal scaling and self organized criticality, Nonlinear Processes in Geophysics 1(1994), pp. 105-114.

Y. Chigirinskaya, D. Schertzer, G. Salvadori, S. Ratti and S. Lovejoy, Chernobyl 137Cs cumulative soil deposition in Europe: is it multifractal? In: F.M. Guindani and G. Salvadori, Editors, Chaos, Fractals and models 96, Italian University Press (1998), pp. 65-72.

A. Clauset, C.R. Shalizi and M.E.J. Newman, Power law distributions in empirical data, SIAM Review 51(2009), pp. 661-703.

A. Davis, A. Marshak, W. Wiscombe and R. Cahalan, Multifractal characterization of intermittency in nonstationary geophysical signals and fields. In: J. Harding, B. Douglas and E. Andreas, Editors, Current topics in nonstationary analysis, World Scientific, Singapore (1996), pp. 97-158.

I. De Lima, Multifractals and the temporal structure of rainfall. Wageningen (1998), p. 225.R. Deidda, Rainfall Downscaling In A Space-Time Multifractal Framework, Water Resources

Research 36(2000), pp. 1779-1794.B. Dubrulle, Phys. Rev. Lett. 73(1994), p. 959.W. Feller, An Introduction to probability theory and its applications, vol. 2, Wiley, New York

(1971) 669 pp.E. Foufoula-Georgiou and P. Kumar, Editors, Wavelets in Geophysics, Academic Press, San

Diego, Ca., USA (1994).J. Gagnon, S., S. Lovejoy and D. Schertzer, Multifractal earth topography, Nonlin. Proc.

Geophys. 13(2006), pp. 541-570.K.P. Georgakakos, A.A. Carsteanu, P.L. Starvedent and J.A. Cramer, Observation and analysis

of midwestern rain rates, J. Appl. Meteorol. 33(1994), pp. 1433-1444.

DRAFT

113

183818391840184118421843184418451846184718481849185018511852185318541855185618571858185918601861186218631864186518661867186818691870187118721873187418751876187718781879188018811882

114

Page 58: 5gang/ftp.transfer/final.book.nofigs.Feb.20…  · Web view5.2.4 Direct empirical estimation of c(g): the probability distribution multiple scaling (PDMS) technique. In chapter 3,

5/18/23 Chapter 5 58

P. Grassberger, Generalized dimensions of strange attractors, Physical Review Letter A 97(1983), p. 227.

R. Grauer, J. Krug and C. Marliani, Phys. Letters A 195(1994), p. 335.T.C. Halsey, M.H. Jensen, L.P. Kadanoff, I. Procaccia and B. Shraiman, Fractal measures and

their singularities: the characterization of strange sets, Physical Review A 33(1986), pp. 1141-1151.

D. Harris, M. Menabde, A. Seed and G. Austin, Multifractal characterizastion of rain fields with a strong orographics influence, J. Geophys. Res. 101(1996), pp. 26405-26414.

H.G.E. Hentschel and I. Procaccia, The infinite number of generalized dimensions of fractals and strange attractors, Physica D 8(1983), pp. 435-444.

M. Holschneider, Wavelets : an analysis tool, Oxford : Clarendon Press, New York (1995).P. Hubert, H. Bendjoudi, D. Schertzer and S. Lovejoy, Multifractal Taming of Extreme

Hydrometeorological Events. The Extremes of the Extremes, Hubert P., Bendjoudi H., Schertzer D., Lovejoy S., 2001: Multifractal taming of ex-treme hydrometeorological events, Proceedings of a symposium on extraordinary floods held at Reyjavik, Iceland, July 2000 (2001).

J.W. Kantelhardt, E. Koscielny-Bunde, H.H.A. Rego, S. Havlin and S. Bunde, Detecting long range corelations with detrended flucutation analysis, Physica A 295(2001), pp. 441-454.

J.W. Kantelhardt et al., Multifractal detrended fluctuation analysis of nonstationary time series, Physica A 316(2002), pp. 87-114.

G. Kiely and K. Ivanova, Multifractal analysis of hourly precipitation, Phys Chem Earth Pt B 24(1999), pp. 781-786.

E. Koscielny-Bunde et al., Indication of a universal persistence law governing atmospheric variability, Phys. Rev. Lett. 81(1998), p. 729.

P. Ladoy, F. Schmitt, D. Schertzer and S. Lovejoy, Variabilité temporelle des observations pluviométriques à Nimes, Comptes Rendues Acad. des Sciences 317, II(1993), pp. 775-782.

D. Lavallée, S. Lovejoy and D. Schertzer, On the determination of the codimension function. In: D. Schertzer and S. Lovejoy, Editors, Non-linear variability in geophysics: Scaling and Fractals, Kluwer (1991), pp. 99-110.

A. Lazarev, D. Schertzer, S. Lovejoy and Y. Chigirinskaya, Unified multifractal atmospheric dynamics tested in the tropics: part II, vertical scaling and Generalized Scale Invariance, Nonlinear Processes in Geophysics 1(1994), pp. 115-123.

S. Lennartz and A. Bunde, Trend evaluation in records with long term memory: Application to global warming, Geophys. Res. Lett. 36(2009), p. L16706.

J. Levy-Véhel, K. Daoudi and E. Lutton, Fractals 2(1995), p. 1.S. Lovejoy, J. Pinel and D. Schertzer, The Global space-time Cascade structure of precipitation:

satellites, gauges and reanalyses, Advances in Water Resources (submitted 11/11)(2010a).

S. Lovejoy, J. Pinel and D. Schertzer, The Global space-time Cascade structure of precipitation: satellites, gridded gauges and reanalyses, Advances in Water Resources (in press)(2012).

S. Lovejoy and D. Schertzer, Scale invariance in climatological temperatures and the spectral plateau, Annales Geophysicae 4B(1986), pp. 401-410.

S. Lovejoy and D. Schertzer, Multifractals, cloud radiances and rain, J. Hydrol. Doi:10.1016/j.hydrol.2005.02.042(2006a).

DRAFT

115

188318841885188618871888188918901891189218931894189518961897189818991900190119021903190419051906190719081909191019111912191319141915191619171918191919201921192219231924192519261927

116

Page 59: 5gang/ftp.transfer/final.book.nofigs.Feb.20…  · Web view5.2.4 Direct empirical estimation of c(g): the probability distribution multiple scaling (PDMS) technique. In chapter 3,

5/18/23 Chapter 5 59

S. Lovejoy and D. Schertzer, Stereophotography of rain drops and compound poisson -cascade processes. Cloud conference, American Meteor. Soc., Madison, Wi. (2006b), pp. p. 14.14.11-14.14.19.

S. Lovejoy and D. Schertzer, Scale, scaling and multifractals in geophysics: twenty years on. In: J.E. A.A. Tsonis, Editor, Nonlinear dynamics in geophysics, Elsevier (2007).

S. Lovejoy and D. Schertzer, Turbulence, rain drops and the l **1/2 number density law, New J. of Physics 10(2008), pp. 075017(075032pp), doi:075010.071088/071367-072630/075010/075017/075017.

S. Lovejoy and D. Schertzer, On the simulation of continuous in scale universal multifractals, part I: spatially continuous processes, Computers and Geoscience(2010a), p. 10.1016/j.cageo.2010.1004.1010.

S. Lovejoy and D. Schertzer, On the simulation of continuous in scale universal multifractals, part II: space-time processes and finite size corrections,, Computers and Geoscience(2010b), p. 10.1016/j.cageo.2010.1007.1001.

S. Lovejoy and D. Schertzer, Low frequency weather and the emergence of the Climate. In: A.S. Sharma, A. Bunde, D. Baker and V.P. Dimri, Editors, Complexity and Extreme Events in Geosciences, AGU monographs (2012).

S. Lovejoy, A.F. Tuck and D. Schertzer, The Horizontal cascade structure of atmospheric fields determined from aircraft data, J. Geophys. Res. 115(2010b), p. D13105.

S. Mallat and W. Hwang, Singularity detection and processing with wavelets, IEEE Trans. Info. Theory 38(1992), pp. 617-643.

D. Marsan, D. Schertzer and S. Lovejoy, Causal space-time multifractal processes: predictability and forecasting of rain fields, J. Geophy. Res. 31D(1996), pp. 26,333-326,346.

L. Mydlarski and Z. Warhaft, J. Fluid. Mech. 358(1998), pp. 135-175.J. Olsson, Limits and Characteristics of the multifractal behavior of a high-resolution rainfall

time series, non-linear processes in geophysics 2(1995), pp. 23-29.C. Onof and K. Arnbjerg-Nielsen, Quantification of anticipated future changes in high

resolutiondesign rainfall for urban areas, Atmos. Res. 92(2009), pp. 350–363.G. Pandey, S. Lovejoy and D. Schertzer, Multifractal analysis including extremes of daily river

flow series for basis five to two million square kilometres, one day to 75 years,, J. Hydrol., 208(1998), pp. 62-81.

G. Parisi and U. Frisch, A multifractal model of intermittency. In: M. Ghil, R. Benzi and G. Parisi, Editors, Turbulence and predictability in geophysical fluid dynamics and climate dynamics, North Holland, Amsterdam (1985), pp. 84-88.

S. Pecknold, S. Lovejoy and D. Schertzer, The morphology and texture of anisotropic multifractals using generalized scale invariance. In: W. Woyczynski and S. Molchansov, Editors, Stochastic Models in Geosystems, Springer-Verlag, New York (1996), pp. 269 - 312.

S. Pecknold, S. Lovejoy, D. Schertzer and C. Hooge, Multifractals and the resolution dependence of remotely sensed data: Generalized Scale Invariance and Geographical Information Systems. In: M.G. D. Quattrochi, Editor, Scaling in Remote Sensing and Geographical Information Systems, Lewis, Boca Raton, Florida (1997), pp. 361-394.

S. Pecknold, S. Lovejoy, D. Schertzer, C. Hooge and J.F. Malouin, The simulation of universal multifractals. In: J.M. Perdang and A. Lejeune, Editors, Cellular Automata: prospects in astronomy and astrophysics, World Scientific, Singapore (1993), pp. 228-267.

DRAFT

117

192819291930193119321933193419351936193719381939194019411942194319441945194619471948194919501951195219531954195519561957195819591960196119621963196419651966196719681969197019711972

118

Page 60: 5gang/ftp.transfer/final.book.nofigs.Feb.20…  · Web view5.2.4 Direct empirical estimation of c(g): the probability distribution multiple scaling (PDMS) technique. In chapter 3,

5/18/23 Chapter 5 60

C.-K. Peng et al., Mosaic organisation of DNA nucleotides, Phys. Rev. E 49(1994), pp. 1685-1689.

D. Quattrochi and M. Goodchild, Editors, Scaling in Remote Sensing and Geographical Information Systems, Lewis, Boca Raton, Florida (1997).

M.I. Radulescu, L.B. Mydlarski, S. Lovejoy and D. Schertzer, Evidence for algebraic tails of probability distributions in laboratory-scale turbulence Advances in turbulence IX: proceedings of the 9th Euro. Turb. Conf. (ETC9), July 2-5, 2002., Southampton, UK (2002), p. 891.

A. Rényi, Probability Theory. , American Elsevier Publishing Company, New York (1970) 666 pp.

G. Salvadori, S. Ratti, G. Belli, S. Lovejoy and D. Schertzer, Multifractal and Fourier analysis of Seveso pollution, J. of Toxicological and Environ. Chem. 43(1993), pp. 63-76.

P.D. Sardeshmukh and P. Sura, Reconciling non-gaussian climate statistics with linear dynamics, J. of Climate 22(2009), pp. 1193-1207.

D. Schertzer et al., Extrêmes et multifractals en hydrologie : résultats, validations et perspectives, Houille Blanche 5(2006), pp. 112-119.

D. Schertzer, M. LarchevÍque, J. Duan, V.V. Yanovsky and S. Lovejoy, Fractional Fokker-Planck equation for nonlinear stochastic differential equation driven by non-Gaussian Levy stable noises, J. of Math. Physics 42(2001), pp. 200-212.

D. Schertzer and S. Lovejoy, The dimension and intermittency of atmospheric dynamics. In: B. Launder, Editor, Turbulent Shear Flow 4, Springer-Verlag (1985), pp. 7-33.

D. Schertzer and S. Lovejoy, Physical modeling and Analysis of Rain and Clouds by Anisotropic Scaling of Multiplicative Processes, Journal of Geophysical Research 92(1987), pp. 9693-9714.

D. Schertzer and S. Lovejoy, Nonlinear variability in geophysics: multifractal analysis and simulation. In: L. Pietronero, Editor, Fractals: Physical Origin and Consequences, Plenum, New-York (1989), p. 49.

D. Schertzer and S. Lovejoy, Nonlinear geodynamical variability: multiple singularities, universality and observables. In: D. Schertzer and S. Lovejoy, Editors, Non-linear variability in geophysics: Scaling and Fractals, Kluwer (1991), pp. 41-82.

D. Schertzer and S. Lovejoy, Hard and Soft Multifractal processes, Physica A 185(1992), pp. 187-194.

D. Schertzer and S. Lovejoy, Lecture Notes: Nonlinear Variability in Geophysics 3: Scaling and Mulitfractal Processes in Geophysics. Institut d'Etudes Scientifique de Cargèse, Cargèse, France (1993), p. 291 pp.

D. Schertzer and S. Lovejoy, Lecture Notes: Resolution Dependence and Multifractals in Remote Sensing and Geographical Information Systems. Course Notes, McGill University, , Montreal, June 10 (1996), p. 390pp.

D. Schertzer, S. Lovejoy and D. Lavallée, Generic Multifractal phase transitions and self-organized criticality. In: J.M. Perdang and A. Lejeune, Editors, Cellular Automata: prospects in astronomy and astrophysics, World Scientific (1993), pp. 216-227.

D. Schertzer, S. Lovejoy, D. Lavallée and F. Schmitt, Universal hard multifractal turbulence: theory and observation. In: R.Z. Sagdeev, U. Frisch, A.S. Moiseev and A. Erokhin, Editors, Nonlinar dynamics of Structures, World Scientific, Singapore (1991), pp. 213-235.

DRAFT

119

197319741975197619771978197919801981198219831984198519861987198819891990199119921993199419951996199719981999200020012002200320042005200620072008200920102011201220132014201520162017

120

Page 61: 5gang/ftp.transfer/final.book.nofigs.Feb.20…  · Web view5.2.4 Direct empirical estimation of c(g): the probability distribution multiple scaling (PDMS) technique. In chapter 3,

5/18/23 Chapter 5 61

D. Schertzer, S. Lovejoy and F. Schmitt, Structures in turbulence and multifractal universality. In: M. Meneguzzi, A. Pouquet and P.L. Sulem, Editors, Small-scale structures in 3D and MHD turbulence, Springer-Verlag, New York (1995), pp. 137-144.

D. Schertzer, S. Lovejoy, F. Schmitt, Y. Chigirinskaya and D. Marsan, Multifractal cascade dynamics and turbulent intermittency, Fractals 5(1997), pp. 427-471.

D. Schertzer, I. Tchiguirinskaia, S. Lovejoy and P. Hubert, No monsters, no miracles: in nonlinear sciences: hydrology is not an outlier!, J. of Hydrol. Sci. in press(2010).

F. Schmitt, D. Schertzer, S. Lovejoy and Y. Brunet, Empirical study of multifractal phase transitions in atmospheric turbulence, Nonlinear Processes in Geophysics 1(1994a), pp. 95-104.

F. Schmitt, D. Schertzer, S. Lovejoy and Y. Brunet, Estimation of universal multifractal indices for atmospheric turbulent velocity fields. , Fractals 3(1994b), pp. 568-575.

Z. She, S. and E. Levèsque, Universal scaling laws in fully developed turbulence, Phys. Rev. Lett. 72(1994), p. 336.

P. Szépfalusy, T. Tél, A. Csordas and Z. Kovas, Phase transitions associated with dynamical properties of chaotic systems, Physical Review A 36(1987), p. 3525.

M. Taqqu, V. Teverovsky and W. Willinger, Fractals 3(1995), p. 785.I. Tchiguirinskaia, D. Schertzer, S. Lovejoy and V. J.M., Wind extremes and scales: multifractal

insights and empirical evidence. In: P.S. Eds. J. Peinke, Editor, EUROMECH Colloquium on Wind Energy, Springer-Verlag (2006).

Y. Tessier, Multifractal objective analysis of rain and clouds. Physics, McGill (1993), p. 167.Y. Tessier, S. Lovejoy and D. Schertzer, Universal Multifractals: theory and observations for

rain and clouds, Journal of Applied Meteorology 32(1993), pp. 223-250.Y. Tessier, S.Lovejoy, P. Hubert, D. Schertzer and S. Pecknold, Multifractal analysis and

modeling of Rainfall and river flows and scaling, causal transfer functions, J. Geophy. Res. 31D(1996), pp. 26,427-426,440.

T. Torrence and G.P. Compo, A practical guide to wavelet analysis, Bull. of the Amer. Meteor. Soc. 79(1998), pp. 61-78.

A.F. Tuck, Atmospheric Turbulence: A Molecular Dynamics Perspective, Oxford University Press (2008).

A.F. Tuck, Quart. J. Roy. Meteor. Soc. 136(2010), pp. 1125-1144.A.F. Tuck, S.J. Hovde and T.P. Bui, Scale Invariance in jet streams: ER-2 data around the lower-

stratospheric polar night vortex, Q. J. R. Meteorol. Soc. 130(2004), pp. 2423-2444.S. Verrier, Modélisation de la variabilité Spatiale et temporelle des précipitations a la sub-

mésoéchelle pare une approche multifractale. LATMOS, Univ. St. Quentin-en-Yvelines (2011), p. 184.

J. Wilson, D. Schertzer and S. Lovejoy, Physically based modelling by multiplicative cascade processes. In: D. Schertzer and S. Lovejoy, Editors, Non-linear variability in geophysics: Scaling and Fractals, Kluwer (1991), pp. 185-208.

DRAFT

121

20182019202020212022202320242025202620272028202920302031203220332034203520362037203820392040204120422043204420452046204720482049205020512052205320542055205620572058

122