Fractal and Multifractal Time Series - Docs
Fractal and Multifractal Time Series - Docs
Contents
1 Definition of the Subject and Its Importance 4
2 Introduction 4
1
6.1 The Structure Function Approach and Singularity Spectra . . 26
6.2 Wavelet Transform Modulus Maxima (WTMM) Method . . . 28
6.3 Multifractal Detrended Fluctuation Analysis (MF-DFA) . . . 29
6.4 Comparison of WTMM and MF-DFA . . . . . . . . . . . . . . 32
7 Statistics of Extreme Events in Fractal Time Series 33
7.1 Return Intervals Between Extreme Events ........... 33
7.2 Distribution of Extreme Events . . . . . . . . . . . . . . . . . 34
8 Simple Models for Fractal and Multifractal Time Series 36
8.1 Fourier Filtering . . . . . . . . . . . . . . . . . . . . . . . . . . 36
8.2 The Schmitz-Schreiber Method ................. 37
8.3 The Extended Binomial Multifractal Model .......... 38
8.4 The Bi-fractal Model . . . . . . . . . . . . . . . . . . . . . . . 39
9 Future Directions 40
10 Bibliography 41
Glossary
• Time series: One dimensional array of numbers (xi), i = 1,...,N,
representing values of an observable x usually measured equidistant
(or nearly equidistant) in time.
2
by self-similarity, i. e., a magnification of a small part is statstically
equivalent to the whole.
3
does not change with time. Non-stationarities like monotonous,
periodic, or step-like trends are often caused by external effects. In a
more general sense, changes in the dynamics of the system also
represent non-stationarities.
4
2 Introduction
The characterisation and understanding of complex systems is a difficult
task, since they cannot be split into simpler subsystems without tampering
the dynamical properties. One approach in studying such systems is the
recording of long time series of several selected variables (observables),
which reflect the state of the system in a dimensionally reduced
representation. Some systems are characterised by periodic or nearly
periodic behaviour, which might be caused by oscillatory components or
closed-loop regulation chains. However, in truly complex systems such
periodic components are usually not limited to one or two characteristic
frequencies or frequency bands. They rather extend over a wide spectrum,
and fluctuations on many time scales as well as broad distributions of the
values are found. Often no specific lower frequency limit – or, equivalently,
upper characteristic time scale – can be observed. In these cases, the
dynamics can be characterised by scaling laws which are valid over a wide
(possibly even unlimited) range of time scales or frequencies; at least over
orders of magnitude. Such dynamics are usually denoted as fractal or
multifractal, depending on the question if they are characterised by one
scaling exponent or by a multitude of scaling exponents.
The first scientist who applied fractal analysis to natural time series is
Benoit B. Mandelbrot [1, 2, 3], who included early approaches by H.E. Hurst
regarding hydrological systems [4, 5]. For extensive introductions
describing fractal scaling in complex systems, we refer to [6, 7, 8, 9, 10, 11,
12, 13]. In the last decade, fractal and multifractal scaling behaviour has
been reported in many natural time series generated by complex systems,
including
5
• technical time series (internet traffic, highway traffic, and neutronic
power from a reactor),
• physics data (also going beyond time series), e. g., surface roughness,
chaotic spectra of atoms, and photon correlation spectroscopy
recordings.
6
stationarities are caused by external or internal effects that lead to either
continuous or sudden changes in the average values, standard deviations or
regulation mechanism. They are a major problem for the characterisation of
the dynamics, in particular for finding the scaling properties of given data.
Following the description of important properties of fractal and
multifractal time series and the definition of our primary quantities of
interest in Section 3, we focus on methods for the analysis of self-affine or
(mono-)fractal data in Sections 4 and 5. While Section 4 describes
traditional approaches for the analysis of stationary time series, Section 5
discusses more recently established methods applicable for non-stationary
data. Section 6 describes techniques for multifractal time series analysis.
The consequences of fractal scaling behaviour on the statistics of extreme
events and their return intervals are presented in Section 7, and a few
standard models for fractal and multifractal time series are described in
Section 8 before an outlook on future directions in Section 9.
3 Fractal and Multifractal Time Series
3.1 Fractality, Self-Affinity, and Scaling
The topic of this article is the fractality (and/or multifractality) of time
series. Since fractals and multifractals in general are discussed in many
other articles of the encyclopedia, the concept is not thoroughly explained
here. In particular, we refer to the articles ... and ... for the formalism
describing fractal and multifractal structures, respectively.
In a strict sense, most time series are one dimensional, since the values
of the considered observable are measured in homogeneous time intervals.
Hence, unless there are missing values, the fractal dimension of the support
is D(0) = 1. However, there are rare cases where most of the values of a time
series are very small or even zero, causing a dimension D(0) < 1 of the
support. In these cases, one has to be very careful in selecting appropriate
analysis techniques, since many of the methods presented in this article are
not accurate for such data; the Wavelet Transform Modulus Maxima
technique (see Section 6.2) is the most advanced applicable method.
Even if the fractal dimension of support is one, the information
dimension D(1) and the correlation dimension D(2) can be studied. As we
will see in Section 6.1, D(2) is in fact explicitly related to all exponents
studied in monofractal time series analysis. However, usually a slightly
different approach is employed based on the notion of self-affinity instead
7
of (multi-) fractality. Here, one takes into account that the time axis and the
axis of the measured values x(t) are not equivalent. Hence, a rescaling of
time t by a factor a may require rescaling of the series values x(t) by a
different factor aH in order to obtain a statistically similar (i. e., self-similar)
picture. In this case the scaling relation
x(t) → aHx(at) (1)
holds for an arbitrary factor a, describing the data as self-affine (see, e. g.,
[6]). The Hurst exponent H (after the water engineer H.E. Hurst [4])
characterises the type of self affinity. Figure 1(a) shows several examples of
self-affine time series with different H. The trace of a random walk
(Brownian motion, third line in Fig. 1(a)), for example, is characterised by H
= 0.5, implying that the position axis must be rescaled by a factor of 2 if the
time axis is rescaled by a factor of 4. Note that self-affine series are often
denoted as fractal even though they are not fractal in the strict sense. In this
article the term ”fractal” will be used in the more general sense including all
data, where a Hurst exponent H can be reasonably defined.
The scaling behaviour of self-affine data can also be characterised by
looking at their mean-square displacement. Since the mean-square
displacement of a random walker is known to increase linear in time, hx2(t)i
∼ t, deviations from this law will indicate the presence of self-affine scaling.
As we will see in Section 4.4, one can thus retrieve the Hurst (or self-
affinity) exponent H by studying the scaling behaviour of the mean-square
dispalcement, or the mean-square fluctuations hx2(t)i ∼ t2H.
8
Considering the increments ∆xi = xi − xi−1 of a self-affine series, (xi), i =
1,...,N with N values measured equidistant in time, one finds that the ∆xi can
be either persistent, independent, or anti-persistent. Examples for all cases
are shown in Fig. 1(b). In our example of the random walk with H = 0.5
(third line in the figure), the increments (steps) are fully independent of
each other. Persistent and anti-persistent increments, where a positive
increment is likely to be followed by another positive or negative increment,
respectively, are also leading to persistent integrated series .
For stationary data with constant mean and standard deviation the
autocovariance function of the increments,
. (2)
with an exponent 0 < γ < 1. Figure 2(b) shows C(s) for one configuration
with γ = 0.4. This type of behaviour can be modelled by the Fourier filtering
technique (see Section 8.1). Long-term correlated, i. e. persistent, behaviour
9
of the ∆xi leads to self-affine scaling behaviour of the xi, characterised by H =
1 − γ/2, as will be shown below.
10
regarded as non-stationarities here, too. Hence, if crossovers in the scaling
behaviour of data are observed, more detailed studies are needed to find
out the cause of the crossovers. One can try to obtain homogenous data by
splitting the original series and employing methods that are at least
insensitive to monotonous (polynomially shaped) trends.
To characterise a complex system based on time series, trends and
fluctuations are usually studied separately (see, e. g., [20] for a discussion).
Strong trends in data can lead to a false detection of long-range statistical
persistence if only one (non-detrending) method is used or if the results are
not carefully interpreted. Using several advanced techniques of scaling time
series analysis (as described in Chapter 5) crossovers due to trends can be
distinguished from crossovers due to different regulation mechanisms on
fast and slow time scales. The techniques can thus assists in gaining insight
into the scaling behaviour of the natural variability as well as into the kind
of trends of the considered time series.
It has to be stressed that crossovers in scaling behaviour must not be
confused with multifractality. Even though several scaling exponents are
needed, they are not applicable for the same regime (i. e., the same range of
time scales). Real multifractality, on the other hand, is characterised by
different scaling behaviour of different moments over the full range of time
scales (see next section).
11
due to different long-term correlations of the small and large fluctuations. In
this case the probability density function of the values can be a regular
distribution with finite moments, e. g., a Gaussian distribution. The
corresponding shuffled series will exhibit non-multifractal scaling, since all
long-range correlations are destroyed by the shuffling procedure. Randomly
shuffling the order of the values in the time series is the easiest way of
generating surrogate data; however, there are more advanced alternatives
(see Chapter 8). If both kinds of multifractality are present, the shuffled
series will show weaker multifractality than the original series.
A multifractal analysis of time series will also reveal higher order
correlations. Multifractal scaling can be observed if, e. g., three or four-point
correlations scale differently from the standard two-point correlations
studied by classical autocorrelation analysis (Eq. (2)). In addition,
multifractal scaling is observed if the scaling behaviour of small and large
fluctuations is different. For example, extreme events might be more or less
correlated than typical events.
12
defined by the (auto-) covariance function C(s) = hx˜i x˜i+si or the (auto-)
correlation function C(s)/hx˜2i i, see also Eq. (2).
As already mentioned in Section 3.2, the ˜xi are short-term correlated if
C(s) declines exponentially, C(s) ∼ exp(−s/t×), and long-term correlated if
C(s) declines as a power-law C(s) ∝ s−γ with a correlation exponent 0 < γ < 1
(see Eqs. (3) and (5), respectively). As illustrated by the two examples
shown in Fig. 2, a direct calculation of C(s) is usually not appropriate due to
noise superimposed on the data ˜xi and due to underlying non-stationarities
of unknown origin. Non-stationarities make the definition of C(s)
problematic, because the average hxi is not well-defined. Furthermore, C(s)
strongly fluctuates around zero on large scales s (see Fig. 2(b)), making it
impossible to find the correct correlation exponent γ. Thus, one has to
determine the value of γ indirectly.
Spectral analysis, however, does not yield more reliable results than
autocorrelation analysis unless a logarithmic binning procedure is applied
to the double logarithmic plot of S(f) [21], see also Fig. 3. I. e., the average of
logS(f) is calculated in successive, logarithmically wide bands from anf0 to
an+1f0, where f0 is the minimum frequency, a > 1 is a factor (e. g., a = 1.1), and
the index n is counting the bins. Spectral analysis also requires stationarity
of the data.
13
4.3 Hurst’s Rescaled-Range Analysis
The first method for the analysis of long-term persistence in time series
based on random walk theory has been proposed by the water construction
engineer Harold Edwin Hurst (1880-1978), who developed it while working
in Egypt. His so-called rescaled range analysis (R/S analysis) [1, 2, 4, 5, 6]
begins with splitting of the time series (˜xi) into non-overlapping segments ν
of size (time scale) s (first step), yielding Ns = int(N/s) segments altogether.
In the second step, the profile (integrated data) is calculated in each
segment ν = 0,...,Ns − 1,
. (7)
By the subtraction of the local averages, piecewise constant trends in the
data are eliminated. In the third step, the differences between minimum and
maximum value (ranges) Rν(s) and the standard deviations Sν(s) in each
segment are calculated,
. (8)
Finally, the rescaled range is averaged over all segments to obtain the
fluctuation function F(s),
for , (9)
where H is the Hurst exponent already introduced in Eq. (1). One can show
[1, 24] that H is related to β and γ by 2H ≈ 1+β = 2−γ (see also Eqs. (6) and
(14)). Note that 0 < γ < 1, so that the right part of the equation does not hold
unless 0.5 < H < 1. The relationship does not hold in general for multifractal
data. Note also that H actually characterises the self-affinity of the profile
function (7), while β and γ refer to the original data.
The values of H, that can be obtained by Hurst’s rescaled range analysis,
are limited to 0 < H < 2, and significant inaccuracies are to be expected close
to the bounds. Since H can be increased or decreased by 1 if the data is
integrated (˜ ) or differentiated (˜xi → x˜i −x˜i−1), respectively, one
can always find a way to calculate H by rescaled range analysis provided the
data is stationary. While values H < 1/2 indicate long-term anti-correlated
behaviour of the data ˜xi, H > 1/2 indicates long-term positively correlated
14
behaviour. For power-law correlations decaying faster than 1/s, we have H
= 1/2 for large s values, like for uncorrelated data.
Compared with spectral analysis, Hurst’s rescaled range analysis yields
smoother curves with less effort (no binning procedure is necessary) and
works also for data with piecewise constant trends.
X
Y (j) = x˜i, j = 0,1,2,...,N, (10)
i=1
and study how the fluctuations of the profile, in a given time window of size
s, increase with s. The procedure is illustrated in Fig. 4 for two values of s.
We can consider the profile Y (j) as the position of a random walker on a
linear chain after j steps. The random walker starts at the origin and
performs, in the ith step, a jump of length ˜xi to the right, if ˜xi is positive, and
to the left, if ˜xi is negative.
To find how the square-fluctuations of the profile scale with s, we first
divide each record of N elements into Ns = int(N/s) non-overlapping
segments of size s starting from the beginning (see Fig. 4) and another Ns
non-overlapping segments of size s starting from the end of the considered
series. This way neither data at the end nor at the beginning of the record is
neglected. Then we determine the fluctuations in each segment ν.
In the standard FA, we obtain the fluctuations just from the values of the
profile at both endpoints of each segment ν = 0,...,Ns − 1,
. (12)
15
. (13)
By definition, F2(s) can be viewed as the root-mean-square displacement of
the random walker on the chain, after s steps (the reason for the index 2 will
become clear later). For uncorrelated xi values, we obtain Fick’s diffusion
law F2(s) ∼ s1/2. For the relevant case of long-term correlations, where C(s)
follows the power-law behaviour of Eq. (5), F2(s) increases by a power law,
where the fluctuation exponent α is identical with the Hurst exponent H for
mono-fractal data and related to γ and β by
2α = 1 + β = 2 − γ. (15)
16
. (16)
Here, ψ(t) is a so-called mother wavelet, from which all daughter wavelets
ψτ,s(t) = ψ((t − τ)/s evolve by shifting and stretching of the time axis. The
wavelet coefficients Lψ(τ,s) thus depend on both, time position τ and scale s.
Hence, the local frequency decomposition of the signal is described with a
time resolution appropriate for the considered frequency f = 1/s (i. e.,
inverse time scale).
All wavelets ψ(t) must have zero mean. They are often chosen to be
orthogonal to polynomial trends, so that the analysis method becomes
insensitive to possible trends in the data. Simple examples are derivatives of
a
Gaussian, 2), like the Mexican hat wavelet
−ψGauss(2) and the Haar wavelet,
2, and 0 otherwise. It is straightforward to construct Haar wavelet that are
orthogonal to linear, quadratic and cubic trends, e. g.,
¯ ¯ ¯
FWT1(ν,s) ≡ LψHaar(1) (νs,s) = Y ν(s) − 2Y ν+1(s) + Y ν+2(s), and (18)
L ¯ ¯ ¯ ¯
FWT2(ν,s) ≡ ψHaar(2) (νs,s) = Y ν(s) − 3Y ν+1(s) + 3Y ν+2(s) − Y ν+3(s) (19) for
constant, linear and quadratic detrending, respectively. The generalization
17
for higher orders of detrending is obvious. The resulting mean-square
fluctuations ) are averaged over all ν to obtain the mean
fluctuation F2(s), see Eq. (13). Figure 5 shows typical results for WT analysis
of long-term correlated, short-term correlated and uncorrelated data.
Regarding trend-elimination, wavelet transform WT0 corresponds to
standard FA (see Section 4.4), and only constant trends in the profile are
eliminated. WT1 is similar to Hurst’s rescaled range analysis (see Section
4.3): linear trends in the profile and constant trends in the data are
eliminated, and the range of the fluctuation exponent α ≈ H is up to 2. In
general, WTn determines the fluctuations from the nth derivative, this way
eliminating trends described by (n−1)st-order polynomials in the data. The
results become statistically unreliable for scales s larger than one tenth of
the length of the data, just as for FA.
18
trends in a detrending procedure. This is done by estimating a polynomial
trend ) within each segment ν by least-square fitting and subtracting
this trend from the original profile (‘detrending’),
. (20)
The degree of the polynomial can be varied in order to eliminate constant
(m = 0), linear (m = 1), quadratic (m = 2) or higher order trends of the
profile function [15]. Conventionally the DFA is named after the order of the
fitting polynomial (DFA0, DFA1, DFA2, ...). In DFAm, trends of order m in the
profile Y (j) and of order m − 1 in the original record ˜xi are eliminated. The
˜
variance of the detrended profile Y s(j) in each segment ν yields the mean-
square fluctuations,
. (21)
As for FA and discrete wavelet analysis, the ) are averaged over all
segments ν to obtain the mean fluctuations F2(s), see Eq. (14). Calculating
F2(s) for many s, the fluctuation scaling exponent α can be determined just
as with FA. Figure 6 shows typical results for DFA of the same long-term
correlated, short-term correlated and uncorrelated data studied already in
Fig. 5.
We note that in studies that include averaging over many records (or
one record cut into many separate pieces by the elimination of some
unreliable intermediate data points) the averaging procedure (13) must be
performed for all data. Taking the square root is always the final step after
all averaging is finished. It is not appropriate to calculate F2(s) for parts of
the data and then average the F2(s) values, since such a procedure will bias
the results towards smaller scaling exponents on large time scales.
If F2(s) increases for increasing s by F2(s) ∼ sα with 0.5 < α < 1, one finds
that the scaling exponent α ≈ H is related to the correlation exponent γ by α
= 1−γ/2 (see Eq. (15)). A value of α = 0.5 thus indicates that there are no (or
only short-range) correlations. If α > 0.5 for all scales s, the data are long-
term correlated. The higher α, the stronger the correlations in the signal are.
α > 1 indicates a non-stationary local average of the data; in this case, FA
fails and yields only α = 1. The case α < 0.5 corresponds to long-term anti-
19
correlations, meaning that large values are most likely to be followed by
small values and vice versa. α values below 0 er not possible. Since the
maximum value for α in DFAm is m + 1, higher detrending orders should be
used for very non-stationary data with large α. Like in FA and Hurst’s
analysis, α will decrease or increase by one upon additional differentiation
or integration of the data, respectively.
Small deviations from the scaling law (14), i. e. deviations from a straight
line in a double logarithmic plot, occur for small scales s, in particular for
DFAm with large detrending order m. These deviations are intrinsic to the
usual DFA method, since the scaling behaviour is only approached
asymptotically. The deviations limit the capability of DFA to determine the
correct correlation behaviour in very short records and in the regime of
small s. DFA6, e. g., is only defined for s ≥ 8, and significant deviations from
the scaling law F2(s) ∼ sα occur even up to s ≈ 30. They will lead to an
overestimation of the fluctuation exponent α, if the regime of small s is used
in a fitting procedure. An approach for correction of this systematic artefact
in DFA is described in [34].
The number of independent segments of length s is larger in DFA than in
WT, and the fluctuations in FA are larger than in DFA. Hence, the analysis
has to be based on s values lower than smax = N/4 for DFA compared with smax
= N/10 for FA and WT. The accuracy of scaling exponents α determined by
DFA was recently studied as a function of the length N of the data [42]
(fitting range s ∈ [10,N/2] was used). The results show that statistical
standard errors of α (one standard deviation) are approximately 0.1 for N =
500, 0.05 for N = 3 000, and reach 0.03 for N = 10 000. Findings of long-term
correlations with α = 0.6 in data with only 500 points are thus not
significant; and α should be at least 0.55 even for data of 10 000 points.
A generalization of DFA for two-dimensional data (or even higher
dimensions d) was recently suggested [33]. The generalization works well
when tested with synthetic surfaces including fractional Brownian surfaces
and multifractal surfaces. The two-dimensional MFDFA is also adopted to
analyse two images from nature and experiment, and nice scaling laws are
unravelled. In the 2d procedure, a double cumulative sum (profile) is
calculated by summing over both directional indices analogous with Eq.
(10), Y (k,l) =
. This surface is partitioned into squares of size s × s with in-
20
dices ν and µ, in which polynomials like
are fitted. The fluctuation function
F2(s) is again obtained by calculating the variance of the profile from the fits.
21
The procedure is also required if the characteristic time scale of
shortterm correlations shall be studied with DFA. If consistent (corrected)
s× values are obtained based on DFAm with different m, the existence of a
real characteristic correlation time scale is positively confirmed. Note that
lower detrending orders are advantageous in this case, since the observed
crossover time scale might become quite large and nearly reach one
forth of the total series length (N/4), where the results become statistically
inaccurate.
We would like to note that studies showing scaling long-term
correlations should not be based on DFA or variants of this method alone in
most applications. In particular, if it is not clear whether a given time series
is indeed long-term correlated or just short-term correlated with a fairly
large crossover time scale, results of DFA should be compared with other
methods. For example, one can employ wavelet methods (see, e. g., Section
5.2). Another option is to remove short-term correlations by considering
averaged series for comparison. For a time series with daily observations
and possible short-term correlations up to two years, for example, one
might consider the series of two-year averages and apply DFA together with
FA, binned power spectra analysis, and/or wavelet analysis. Only if these
methods still indicate long-term correlations, one can be sure that the data
are indeed long-term correlated.
As discussed in Section 3.3, records from real measurements are often
affected by non-stationarities, and in particular by trends. They have to be
well distinguished from the intrinsic fluctuations of the system. To
investigate the effect of trends on the DFAm fluctuation functions, one can
generate artificial series (˜xi) with smooth monotonous trends by adding
polynomials of different power p to the original record (xi),
For the DFAm, such trends in the data can lead to an artificial crossover in
the scaling behaviour of F2(s), i. e., the slope α is strongly increased for large
time scales s. The position of this artificial crossover depends on the
strength A and the power p of the trend. Evidently, no artificial crossover is
observed, if the detrending order m is larger than p and p is integer. The
order p of the trends in the data can be determined easily by applying the
different DFAm. If p is larger than m or p is not an integer, an artificial
22
crossover is observed, the slope αtrend in the large s regime strongly depends
on m, and the position of the artificial crossover also depends strongly on m.
The artificial crossover can thus be clearly distinguished from real
crossovers in the correlation behaviour, which result in identical slopes α
and rather similar crossover positions for all detrending orders m. For more
extensive studies of trends with non-integer powers we refer to [34, 35].
The effects of periodic trends are also studied in [34].
If the functional form of the trend in given data is not known a priori, the
fluctuation function F2(s) should be calculated for several orders m of the
fitting polynomial. If m is too low, F2(s) will show a pronounced crossover to
a regime with larger slope for large scales s [34, 35]. The maximum slope of
logF2(s) versus logs is m + 1. The crossover will move to larger scales s or
disappear when m is increased, unless it is a real crossover not due to
trends. Hence, one can find m such that detrending is sufficient. However, m
should not be larger than necessary, because shifts of the observed
crossover time scales and deviations on short scales s increase with
increasing m.
23
Most published results report short-term anti-correlations and no
longterm correlations in the sign series, i. e., αsign < 1/2 for the non-
integrated signs si (or αsign < 3/2 for the integrated signs) on low time scales
and αsign → 1/2 asymptotically for large s. The magnitude series, on the other
hand, are usually either uncorrelated αmagn = 1/2 (or 3/2) or positively
longterm correlated αmagn > 1/2 (or 3/2). It has been suggested that findings
of αmagn > 1/2 are related with nonlinear properties of the data and in
particular multifractality [31, 46, 47], if α < 1.5 in standard DFA. Specifically,
the results suggest that the correlation exponent of the magnitude series is
a monotonically increasing function of the multifractal spectrum (i. e., the
singularity spectrum) width of the original series (see Section 6.1). On the
other hand, the sign series mainly relates to linear properties of the original
series. At small time scales s < 16 the standard α is approximately the
average of αsign and αmagn, if integrated sign and magnitude series are
analysed. For α > 1.5 in the original series, the integrated magnitude and
sign series have approximately the same two-point scaling exponents [46].
An analytical treatment is presented in [47].
24
• the Modified Detrended Fluctuation Analysis (MDFA) [54], which is
essentially a mixture of old FA and DFA,
. (23)
As in case of DFA, MDFA can easily be generalised to remove higher order
trends in the data. Since the fitting polynomials in adjacent segments are
not related, Y˜s(j) shows abrupt jumps on their boundaries as well. This
leads to fluctuations of F2(s) for large segment sizes s and limits the
maximum usable scale to s < N/4 as for DFA. The detection of crossovers in
the data, however, is more exact with MDFA (compared with DFA), since no
correction of the estimated crossover time scales seems to be needed [42].
25
The Fourier-detrended fluctuation analysis [57] aims to eliminate slow
oscillatory trends which are found especially in weather and climate series
due to seasonal influences. The character of these trends can be rather
periodic and regular or irregular, and their influence on the detection of
long-range correlations by means of DFA was systematically studied
previously [34]. Among other things it has been shown that slowly varying
periodic trends disturb the scaling behaviour of the results much stronger
than quickly oscillating trends and thus have to be removed prior to the
analysis. In case of periodic and regular oscillations, e. g., in temperature
fluctuations one simply removes the low frequency seasonal trend by
subtracting the daily mean temperatures from the data. Another way, which
the Fourier-detrended fluctuation analysis suggests, is to filter out the
relevant frequencies in the signals’ Fourier spectrum before applying DFA
to the filtered signal. Nevertheless, this method faces several difficulties
especially its limitation to periodic and regular trends and the need for a
priori knowledge of the interfering frequency band.
To study correlations in data with quasi-periodic or irregular oscillating
trends, empirical mode decomposition (EMD) was suggested [58]. The EMD
algorithm breaks down the signal into its intrinsic mode functions (IMFs)
which can be used to distinguish between fluctuations and background. The
background, estimated by a quasi-periodic fit containing the dominating
frequencies of a sufficiently large number of IMFs, is subtracted from the
data, yielding a slightly better scaling behaviour in the DFA curves.
However, we believe that the method might be too complicated for wide-
spread applications.
Another method which was shown to minimise the effect of periodic and
quasi-periodic trends is based on singular value decomposition (SVD) [59,
60]. In this approach, one first embeds the original signal in a matrix whose
dimension has to be much larger than the number of frequency components
of the periodic or quasi-periodic trends obtained in the power spectrum.
Applying SVD yields a diagonal matrix which can be manipulated by setting
the dominant eigen-values (associated with the trends) to zero. The filtered
matrix finally leads to the filtered data, and it has been shown that
subsequent application of DFA determines the expected scaling behaviour if
the embedding dimension is sufficiently large. None the less, the
performance of this rather complex method seems to decrease for larger
26
values of the scaling exponent. Furthermore SVD-DFA assumes that trends
are deterministic and narrow banded.
The detrending procedure in DFA (Eq. (20)) can be regarded as a
scaledependent high-pass filter since (low-frequency) fluctuations
exceeding a specific scale s are eliminated. Therefore, it has been suggested
˜
to obtain the detrended profile Y s(j) for each scale s directly by applying
digital highpass filters [61]. In particular, Butterworth, Chebyshev-I,
Chebyshev-II, and an elliptical filter were suggested. While the elliptical
filter showed the best performance in detecting long-range correlations in
artificial data, the Chebyshev-II filter was found to be problematic.
Additionally, in order to avoid a time shift between filtered and original
profile, the average of the directly filtered signal and the time reversed
filtered signal is considered. The effects of these complicated filters on the
scaling behaviour are, however, not fully understood.
Finally, a continuous DFA method has been suggested in the context of
studying heartbeat data during sleep [55, 56]. The method compares
unnormalised fluctuation functions F2(s) for increasing length of the data. I.
e., one starts with a very short recording and subsequently adds more
points of data. The method is particularly suitable for the detection of
change points in the data, e. g., physiological transitions between different
activity or sleep stages. Since the main objective of the method is not the
study of scaling behaviour, we do not discuss it in detail here.
5.7 Centered Moving Average (CMA) Analysis
Particular attractive modifications of DFA are the Detrended Moving
Average (DMA) methods, where running averages replace the polynomial
fits. The first suggested version, the Backward Moving Average (BMA)
method [50, 51, 52], however, suffers from severe problems, because an
artificial time shift of s between the original signal and the moving average
is introduced. This time shift leads to an additional contribution to the
detrended profile Y˜s(j), which causes a larger fluctuation function F2(s) in
particular for small scales in the case of long-term correlated data. Hence,
the scaling exponent α is systematically underestimated [62]. In addition,
the BMA method preforms even worse for data with trends [63], and its
slope is limited by α < 1 just as for the non-detrending method FA.
27
It was soon recognised that the intrinsic error of BMA can be overcome
by eliminating the artificial time shift. This leads to the Centred Moving
˜
Average (CMA) method [53], where Y s(j) is calculated as
, (24)
replacing Eq. (20) while Eq. (21) and the rest of the DFA procedure
described in Section 5.3 stay the same. Unlike DFA, the CMA method cannot
easily be generalised to remove linear and higher order trends in the data.
It was recently proposed [42] that the scaling behaviour of the CMA
method is more stable than for DFA1 and MDFA1, suggesting that CMA
could be used for reliable computation of α even for scales s < 10 (without
correction of any systematic deviations needed in DFA for this regime) and
up to smax = N/2. The standard errors in determining the scaling exponent α
by fitting straight lines to the double logarithmic plots of F2(s) have been
studied in [42]; they are comparable with DFA1 (see end of Section 5.3).
Regarding the determination of crossovers, CMA is comparable to DFA1.
Ultimately, the CMA seems to be a good alternative to DFA1 when analysing
the scaling properties in short data sets without trends. Nevertheless for
data with possible unknown trends we recommend the application of
standard DFA with several different detrending polynomial orders in order
to distinguish real crossovers from artificial crossovers due to trends. In
addition, an independent approach (e. g., wavelet analysis) should be used
to confirm findings of long-term correlations (see also Section 5.4).
6 Methods for Multifractal Time Series Analysis
This chapter describes the multifractal characterisation of time series, for
an introduction, see Section 3.4. The simplest type of multifractal analysis is
based upon the standard partition function multifractal formalism, which
has been developed for the multifractal characterisation of normalised,
stationary measures [6, 12, 64, 65]. Unfortunately, this standard formalism
does not give correct results for non-stationary time series that are affected
by trends or that cannot be normalised. Thus, in the early 1990s an
improved multifractal formalism has been developed, the wavelet
transform modulus maxima (WTMM) method [66, 67, 68, 69, 70], which is
based on wavelet analysis and involves tracing the maxima lines in the
continuous wavelet transform over all scales. An important alternative is
28
the multifractal DFA (MF-DFA) algorithm [32], which does not require the
modulus maxima procedure, and hence involves little more effort in
programming than the conventional DFA. For studies comparing methods
for detrending multifractal analysis (multifractal DFA (MF-DFA) and
wavelet transform modulus maxima (WTMM) method), see [32, 71, 72].
for , (25)
where τ(q) is the Renyi scaling exponent and q is a real parameter that can
take positive as well as negative moments. Note that τ(q) is sometimes
defined with opposite sign (see, e. g., [6]). A record is called monofractal (or
self-affine), when the Renyi scaling exponent τ(q) depends linearly on q;
otherwise it is called multifractal. The generalised multifractal dimensions
D(q) (see also Section 3.4) are related to τ(q) by D(q) = τ(q)/q − 1, such that
the fractal dimension of the support is D(0) = −τ(0) and the correlation
dimension is D(2) = τ(2).
In time series, a discrete version has to be used, and the considered data
(xi), i = 1,...,N may usually include negative values. Hence, setting Ns =
int(N/s) and 1 we can define [6,
12],
Ns−1
X
Zq(s) = |X(ν,s)|q ∼ sτ(q) for s > 1. (26)
ν=0
Inserting the profile Y (j) and FFA(ν,s) from Eqs. (10) and (11), respectively,
we obtain
Ns−1 Ns−1
X o
n 2 q/2 X q/2
29
ν=0 ν=0
Comparing Eqs. (27) with (13), we see that this multifractal approach can
be considered as a generalised version of the Fluctuation Analysis (FA)
method, where the exponent 2 is replaced by q. In particular we find
(disregarding the summation over the second partition of the time series)
We thus see that all methods for (mono-)fractal time analysis (discussed in
Chapters 4 and 5) in fact study the correlation dimension D(2) = 2α − 1 = β =
1 − γ (see Eq. (15)).
It is straightforward to define a generalised (multifractal) Hurst
exponent h(q) for the scaling behaviour of the qth moments of the
fluctuations [64, 65],
with h(2) = α ≈ H. In the following, we will use only h(2) for the standard
fluctuation exponent (denoted by α in the previous chapters), and reserve
the letter α for the Ho¨lder exponent.
Another way to characterise a multifractal series is the singularity
spectrum f(α), that is related to τ(q) via a Legendre transform [6, 12],
30
wavelet transform with continuous basis functions as defined in Section 5.1,
Eq. (16). Note that in this case the series ˜xi are analysed directly instead of
the profile Y (j) defined in Eq. (10). Using wavelets orthogonal to mth order
polynomials, the corresponding trends are elliminated.
Instead of averaging over all wavelet coefficients Lψ(τ,s), one averages,
within the modulo-maxima method, only the local maxima of |Lψ(τ,s)|. First,
one determines for a given scale s, the positions τj of the local maxima of |
W(τ,s)| as function of τ, so that |Lψ(τj − 1,s)| < |Lψ(τj,s)| ≥ |Lψ(τj + 1,s)| for j =
1,...,jmax. This maxima procedure is demonstrated in Fig. 7. Then one sums up
the qth power of the maxima,
jmax
X
Z(q,s) = |Lψ(τj,s)|q. (32)
j=1
The reason for the maxima procedure is that the absolute wavelet
coefficients |Lψ(τ,s)| can become arbitrarily small. The analysing wavelet
ψ(x) must always have positive values for some x and negative values for
other x, since it has to be orthogonal to possible constant trends. Hence
there are always positive and negative terms in the sum (16), and these
terms might cancel. If that happens, |Lψ(τ,s)| can become close to zero. Since
such small terms would spoil the calculation of negative moments in Eq.
(32), they have to be eliminated by the maxima procedure.
In fluctuation analysis, on the other hand, the calculation of the
variances F2(ν,s), e. g. in Eq. (11), involves only positive terms under the
summation. The variances cannot become arbitrarily small, and hence no
maximum procedure is required for series with compact support. In
addition, the variances will always increase if the segment length s is
increased, because the fit will always be worse for a longer segment. In the
WTMM method, in contrast, the absolute wavelet coefficients |Lψ(τ,s)| need
not increase with increasing scale s, even if only the local maxima are
considered. The values |Lψ(τ,s)| might become smaller for increasing s since
just more (positive and negative) terms are included in the summation (16),
and these might cancel even better. Thus, an additional supremum
procedure has been introduced in the WTMM method in order to keep the
dependence of Z(q,s) on s monotonous. If, for a given scale s, a maximum at a
certain position τj happens to be smaller than a maximum at τj0 ≈ τj for a
lower scale s0 < s, then Lψ(τj,s) is replaced by Lψ(τj0,s0) in Eq. (32).
31
Often, scaling behaviour is observed for Z(q,s), and scaling exponents
τˆ(q) can be defined that describe how Z(q,s) scales with s,
• Step 1: Calculate the profile Y (j, Eq. (10), by integrating the time
series.
• Step 3: Calculate the local trend for each of the 2Ns segments by a
least-square fit of the profile. Then determine the variance by Eqs.
(20) and (21) for each segment ν = 0,...,2Ns − 1 Again, linear, quadratic,
32
cubic, or higher order polynomials can be used in the fitting
procedure, and the corresponding methods are thus called MF-DFA1,
MF-DFA2, MF-DFA3, ...) [32]. In (MF-)DFAm [mth order (MF-)DFA]
trends of order mp in the profile (or, equivalently, of order m − 1 in
the original series) are eliminated. Thus a comparison of the results
for different orders of DFA allows one to estimate the type of the
polynomial trend in the time series [34, 35].
• Step 4: Average over all segments to obtain the qth order fluctuation
function
, (34)
This is the generalization of Eq. (13) suggested by the relations
derived in Section 6.1. For q = 2, the standard DFA procedure is
retrieved. One is interested in how the generalised q dependent
fluctuation functions Fq(s) depend on the time scale s for different
values of q. Hence, we must repeat steps 2 to 4 for several time scales
s. It is apparent that Fq(s) will increase with increasing s. Of course,
Fq(s) depends on the order m. By construction, Fq(s) is only defined for
s ≥ m + 2.
For very large scales, s > N/4, Fq(s) becomes statistically unreliable
because the number of segments Ns for the averaging procedure in step 4
becomes very small. Thus, scales s > N/4 should be excluded from the fitting
procedure determining h(q). Besides that, systematic deviations from the
scaling behaviour in Eq. (35), which can be corrected, occur for small scales
s ≈ 10.
33
The value of h(0), which corresponds to the limit h(q) for q → 0, cannot
be determined directly using the averaging procedure in Eq. (34) because of
the diverging exponent. Instead, a logarithmic averaging procedure has to
be employed,
. (36)
Note that h(0) cannot be defined for time series with fractal support, where
h(q) diverges for q → 0.
For monofractal time series with compact support, h(q) is independent
of q, since the scaling behaviour of the variances ) is identical for
all segments ν, and the averaging procedure in Eq. (34) will give just this
identical scaling behaviour for all values of q. Only if small and large
fluctuations scale differently, there will be a significant dependence of h(q)
on q: If we consider positive values of q, the segments ν with large variance
F2(ν,s) (i. e. large deviations from the corresponding fit) will dominate the
average Fq(s). Thus, for positive values of q, h(q) describes the scaling
behaviour of the segments with large fluctuations. On the contrary, for
negative values of q, the segments ν with small variance ) will
dominate the average Fq(s). Hence, for negative values of q, h(q) describes
the scaling behaviour of the segments with small fluctuations. Figure 8
shows typical results obtained for Fq(s) in the MF-DFA procedure.
Usually the large fluctuations are characterised by a smaller scaling
exponent h(q) for multifractal series than the small fluctuations. This can be
understood from the following arguments: For the maximum scale s = N the
fluctuation function Fq(s) is independent of q, since the sum in Eq. (34) runs
over only two identical segments. For smaller scales the averaging
procedure runs over several segments, and the average value Fq(s) will be
dominated by the F2(ν,s) from the segments with small (large) fluctuations if
q < 0 (q > 0). Thus, for ) with q < 0 will be smaller than
Fq(s) with q > 0, while both become equal for s = N. Hence, if we assume an
homogeneous scaling behaviour of Fq(s) following Eq. (35), the slope h(q) in
a log-log plot of Fq(s) with q < 0 versus s must be larger than the
corresponding slope for Fq(s) with q > 0. Thus, h(q) for q < 0 will usually be
larger than h(q) for q > 0.
However, the MF-DFA method can only determine positive generalised
Hurst exponents h(q), and it already becomes inaccurate for strongly
34
anticorrelated signals when h(q) is close to zero. In such cases, a modified
(MF)DFA technique has to be used. The most simple way to analyse such
data is to integrate the time series before the MF-DFA procedure. Following
the MF-DFA procedure as described above, we obtain a generalised
˜
fluctuation functions described by a scaling law with h (q) = h(q) + 1. The
scaling behaviour can thus be accurately determined even for h(q) which
are smaller than zero for some values of q.
The accuracy of h(q) determined by MF-DFA certainly depends on the
length N of the data. For q = ±10 and data with N = 10 000 and 100 000,
systematic and statistical error bars (standard deviations) up to ∆h(q) ≈
±0.1 and ≈ ±0.05 should be expected, respectively [32]. A difference of
h(−10) − h(+10) = 0.2, corresponding to an even larger width ∆α of the
singularity spectrum f(α) defined in Eq. (30) is thus not significant unless
the data was longer than N = 10 000 points. Hence, one has to be very
careful when concluding multifractal properties from differences in h(q).
As already mentioned in the introduction, two types of multifractality in
time series can be distinguished. Both of them require a multitude of scaling
exponents for small and large fluctuations: (i) Multifractality of a time series
can be due to a broad probability density function for the values of the time
series, and (ii) multifractality can also be due to different longrange
correlations for small and large fluctuations. The most easy way to
distinguish between these two types is by analysing also the corresponding
randomly shuffled series [32]. In the shuffling procedure the values are put
into random order, and thus all correlations are destroyed. Hence the
shuffled series from multifractals of type (ii) will exhibit simple random
behaviour, hshuf(q) = 0.5, i. e. non-multifractal scaling. For multifractals of
type (i), on the contrary, the original h(q) dependence is not changed, h(q) =
hshuf(q), since the multifractality is due to the probability density, which is
not affected by the shuffling procedure. If both kinds of multifractality are
present in a given series, the shuffled series will show weaker
multifractality than the original one.
35
methods are rather equivalent. Besides that, the main advantage of the
MFDFA method compared with the WTMM method lies in the simplicity of
the MF-DFA method. However, contrary to WTMM, MF-DFA is restricted to
studies of data with full one-dimensional support, while WTMM is not. Both,
WTMM and MF-DFA have been generalised for higher dimensional data, see
[33] for higher dimensional MF-DFA and, e. g., [70] for higher dimensional
WTMM. Studies of other generalisations of detrending methods like the
discrete WT approach (see Section 5.2) and the CMA method (see Section
5.7) are currently under investigation [75].
36
, (37)
This behaviour is shown in Fig. 10. The exponent γ is the correlation
exponent from C(s), and the parameters aγ and bγ are independent of q. They
R
can be determined from the normalization conditions for Pq(r), i. e., Pq(r)dr
R
= 1 and rPq(r)dr = Rq. The form of the distribution (37) indicates that
return intervals both well below and well above their average value Rq
(which is independent of γ) are considerably more frequent for long-term
correlated than for uncorrelated data. It has to be noted that there are
deviations from the stretched exponential law (37) for very small r
(discretization effects and an additional power-law regime) and for very
large r (finite size effects), see Fig. 10. The extent of the deviations from Eq.
(37) depends on the distribution of the values xi of the time series. For a
discussion of these effects, see [79].
(38)
37
For uncorrelated records, τq(x|r0)/Rq = 1 (except for discreteness effects
that lead to τq(x|r0)/Rq > 1 for x > 0, see [84]). Due to the scaling of Pq(r|r0),
also τq(x|r0)/Rq scales with r0/Rq and x/Rq. Small and large return intervals
are more likely to be followed by small and large ones, respectively, and
hence τq(0|r0)/Rq = Rq(r0)/Rq is well below (above) one for r0/Rq well below
(above) one. With increasing x, the expected residual time to the next event
increases. Note that only for an infinite long-term correlated record, the
value of τq(x|r0) will increase indefinitely with x and r0. For real (finite)
records, there exists a maximum return interval which limits the values of x,
r0 and τq(x|r0).
38
independently and identically distributed (i.i.d.) data (xi) with Gaussian or
exponential distribution density function P(x) the integrated distribution
GR(m) converges to a double exponential (Fisher-Tippet-Gumbel)
distribution (often labelled as Type I) [85, 86, 87, 88, 89], i. e.,
(40)
for R → ∞, where α is the scale parameter and u the location parameter. By
the method of moments those parameters are given by and u =
mR − neα with the Euler constant ne = 0.577216 [88, 90, 91, 92]. Here mR and
σR denote the (R-dependent) mean maximum and the standard deviation,
respectively. Note that different asymptotics will be reached for broader
distributions of data (xi) that belong to other domains of attraction [88]. For
example, for data following a power-law distribution (or Pareto
distribution), P(x) = (x/x0)−k, GR(m) converges to a Fr´echet distribution,
often labelled as Type II. For data following a distribution with finite upper
endpoint, for example the uniform distribution P(x) = 1 for 0 ≤ x ≤ 1 , GR(m)
converges to a Weibull distribution, often labelled as Type III. We do not
consider the latter two types of asymptotics here.
Numerical studies of fractal model data have recently shown that the
distribution P(x) of the original data has a much stronger effect upon the
convergence towards the Gumbel distribution than the long-term
correlations in the data. Long-term correlations just slightly delay the
convergence of GR(m) towards the Gumbel distribution (40). This can be
observed very clearly in a plot of the integrated and scaled distribution
GR(m) on logarithmic scale [80].
Furthermore, it was found numerically that (i) the maxima series (mj)
exhibit long-term correlations similar to those of the original data (xi), and
most notably (ii) the maxima distribution as well as the mean maxima
significantly depend on the history, in particular on the previous maximum
[80]. The last item implies that conditional mean maxima and conditional
maxima distributions should be considered for improved extreme event
predictions.
39
8 Simple Models for Fractal and Multifractal Time
Series
8.1 Fourier Filtering
Fractal scaling with long-term correlations can be introduced most easily
into time series by the Fourier-filtering technique, see, e. g., [93, 94, 95]. The
Fourier filtering technique is not limited to the generation of long-term
correlated data characterised by a power-law auto-correlation function C(s)
∼ x−γ with 0 < γ < 1. All values of the scaling exponents α = h(2) ≈ H or β = 2α
− 1 can be obtained, even those that cannot be found directly by the fractal
analysis techniques described in Chapters 4 and 5 (e. g. α < 1). Note,
however, that Fourier filtering will always yield Gaussian distributed data
values and that no nonlinear or multifractal properties can be achieved (see
also Sections 3.4, 5.5, and Chapter 6). In Section 5.4, we have briefly
described a modification of Fourier filtering for obtaining reliable short-
term correlated data.
For the generation of data characterised by fractal scaling with β = 2α−1
[93, 94] we start with uncorrelated Gaussian distributed random numbers
xi from an i.i.d. generator. Transforming a series of such numbers into
frequency space with discrete Fourier transform or FFT (for suitable series
lengths N) yields a flat power spectrum, since random numbers correspond
to white noise. Multiplying the (complex) Fourier coefficients by f−β/2, where
f ∝ 1/s is the frequency, will rescale the power spectrum S(f) to follow Eq.
(6), as expected for time series with fractal scaling. After transforming back
to the time domain (using inverse Fourier transform or inverse FFT) we will
thus obtain the desired long-term correlated data ˜xi. The final step is the
normalization of this data.
The Fourier filtering method can be improved using modified Bessel
functions instead of the simple factors f−β/2 in modifying the Fourier
coefficients [95]. This way problems with the divergence of the
autocorrelation function C(s) at s = 0 can be avoided.
An alternative method to the Fourier filtering technique, the random
midpoint displacement method, is based on the construction of self-affine
surfaces by an iterative procedure, see, e. g. [6]. Starting with one interval
with constant values, the intervals are iterative split in the middle and the
midpoint is displaced by a random offset. The amplitude of this offset is
40
scaled according to the length of the interval. Since the method generates a
self-affine surface xi characterised by a Hurst exponent H, the differentiated
series ∆xi can be used as long-term correlated or anti-correlated random
numbers. Note, however, that the correlations do not persist for the whole
length of the data generated this way. Another option is the use of wavelet
synthesis, the reverse of wavelet analysis described in Section 5.1. In that
method, the scaling law is introduced by setting the magnitudes of the
wavelet coefficients according to the corresponding time scale s.
. (41)
41
distribution and is clearly correlated. However, due to the exchange
algorithm the perfect long-term correlations of the new data sequence were
slightly altered again. So the procedure is repeated: the new sequence is
Fourier transformed followed by spectrum adjustment, and the exchange
algorithm is applied to the Fourier back-transformed data set. These steps
are repeated several times, until the desired quality (or the best possible
quality) of the spectrum of the new data series is achieved.
. (42)
42
It is easy to see that h(1) = 1 for all values of a and b. Thus, in this form the
model is limited to cases where h(1), which is the exponent Hurst defined
originally in the R/S method, is equal to one.
In order to generalise this multifractal cascade process such that any
value of h(1) is possible, one can subtract the offset ∆h = ln(a+b)/ln(2) from
h(q) [99]. The constant offset ∆h corresponds to additional long-term
correlations incorporated in the multiplicative cascade model. For
generating records without this offset, we rescale the power spectrum.
First, we fast-Fourier transform (FFT) the simple multiplicative cascade
data into the frequency domain. Then, we multiply all Fourier coefficients
by f−∆h, where f is the frequency. This way, the slope β of the power spectra
S(f) ∼ f−β is decreased from β = 2h(2)−1 = [2ln(a+b)−ln(a2+b2)]/ln2 into β0 =
2[h(2)−∆h]−1 = −ln(a2+b2)/ln2. Finally, backward FFT is employed to
transform the signal back into the time domain.
, (45)
or vice versa,
(46)
43
Both versions of this bi-fractal model require three parameters. The
multifractal spectrum is degenerated to two single points, thus its width can
be defined as ∆α = α1 − α2.
9 Future Directions
The most straightforward future direction is to analyse more types of time
series from other complex systems than those listed in Chapter 2 to check
for the presence of fractal scaling and in particular long-term correlations.
Such applications may include (i) data that are not recorded as function of
time but as function of another parameter and (ii) higher dimensional data.
In particular, the inter-relationship between fractal time series and spatially
fractal structures can be studied. Studies of fields with fractal scaling in time
and space have already been performed in Geophysics. In some cases
studying new types of data will require dealing with more difficult types of
non-stationarities and transient behaviour, making further development of
the methods necessary. In many studies, detrending methods have not been
applied yet. However, discovering fractal scaling in more and more systems
cannot be an aim on its own.
Up to now, the reasons for observed fractal or multifractal scaling are
not clear in most applications. It is thus highly desirable to study causes for
fractal and multifractal correlations in time series, which is a difficult task,
of course. One approach might be based on modelling and comparing the
fractal aspects of real and modelled time series by applying the methods
described in this article. The fractal or multifractal characterisation can thus
be helpful in improving the models. For many applications, practically
usable models which display fractal or transient fractal scaling still have to
be developed. One example for a model explaining fractal scaling might be a
precipitation, storage, and runoff model, in which the fractal scaling of
runoff time series could be explained by fractional integration of rainfall in
soil, groundwater reservoirs, or river networks characterised by a fractal
structure. Also studies regarding the inter-relation between fractal scaling
and complex networks, representing the structure of a complex system, are
desirable. This way one could gain an interpretation of the causes for fractal
behaviour.
Another direction of future research is regarding the linear and
especially non-linear inter-relations between several time series. There is
great need for improved methods characterising cross-correlations and
44
similar statistical inter-relations between several non-stationary time
series. Most methods available so far are reserved to stationary data, which
is, however, hardly found in natural recordings. See [101] for a very recent
approach analysing fractal cross-correlations in non-stationary data. An
even more ambitious aim is the (time-dependent) characterisation of a
larger network of signals. In such a network, the signals themselves would
represent the nodes, while the (possibly directed) inter-relations between
each pair represent the links (or bonds) between the nodes. The properties
of both, nodes and links can vary with time or change abruptly, when the
represented complex system goes through a phase transition.
Finally, more work will have to be invested in studying the practical
consequences of fractal scaling in time series. Studies should particularly
focus on predictions of future values and behaviour of time series and
whole complex systems. This is very relevant, not only in hydrology and
climate research, where a clear distinguishing of trends and natural
fluctuations is crucial, but also for predicting dangerous medical events on-
line in patients based on the continuous recording of time series.
10 Bibliography
References
[1] B.B. Mandelbrot, J.W. van Ness, Fractional Brownian motions, fractional
noises and applications, SIAM Review 10, 422 (1968).
45
[4] H.E. Hurst, Long-term storage capacity of reservoirs, Transactions of the
American Society of Civil Engineering 116, 770 (1951).
[5] H.E. Hurst, R.P. Black, Y.M. Simaika, Long-term storage: an experimental
study (Constable and Co., Ltd, London, 1965).
[14] C.-K. Peng, J. Mietus, J.M. Hausdorff, S. Havlin, H.E. Stanley, A.L.
Goldberger, Long-range anti-correlations and non-Gaussian behaviour of
the heartbeat, Phys. Rev. Lett. 70, 1343 (1993).
46
[18] G.E.P. Box, G.M. Jenkins, G.C. Reinsel, Time-series Analysis (Prentice
Hall, New Jersey 1994).
[24] G.A. Hunt, Random Fourier transforms, Trans. Amer. Math. Soc. 71, 38
(1951).
[30] C.-K. Peng, S.V. Buldyrev, S. Havlin, M. Simons, H.E. Stanley, A.L.
Goldberger, Mosaic organization of DNA nucleotides, Phys. Rev. E 49,
1685 (1994).
47
[31] Y. Ashkenazy, P. Ch. Ivanov, S. Havlin, C.-K. Peng, A. L. Goldberger, H.
E. Stanley, Magnitude and sign correlations in heartbeat fluctuations, Phys.
Rev. Lett. 86, 1900 (2001).
[33] G.-F. Gu, W.-X. Zhou, Detrended fluctuation analysis for fractals and
multifractals in higher dimensions, Phys. Rev. E 74, 061104 (2006)
[35] K. Hu, P.Ch. Ivanov, Z. Chen, P. Carpena, H.E. Stanley, Effect of trends
on detrended fluctuation analysis, Phys. Rev. E 64, 011114 (2001).
48
[43] S. Bahar, J.W. Kantelhardt, A. Neiman, H.H.A. Rego, D.F. Russell, L.
Wilkens, A. Bunde, F. Moss, Long range temporal anti-correlations in
paddlefish electro-receptors, Europhys. Lett. 56, 454 (2001).
49
[54] K. Kiyono, Z.R. Struzik, N. Aoyagi, F. Togo, Y. Yamamoto, Phase
transition in a healthy human heart rate, Phys. Rev. Lett. 95, 058101
(2005).
[63] L. Xu, P.Ch. Ivanov, K.Hu, Z. Chen, A. Carbone, and H.E. Stanley,
Quantifying signals with power-law correlations: a comparative study of
detrended fluctuation analysis and detrended moving average techniques,
Phys. Rev. E 71, 051101 (2005).
50
[65] E. Bacry, J. Delour, J.F. Muzy, Multifractal random walk , Phys. Rev. E
64, 026103 (2001).
51
(2008).
[77] A. Bunde, J.F. Eichner, J.W. Kantelhardt, S. Havlin, The effect of longterm
correlations on the return periods of rare events, Physica A 330, 1
(2003).
[83] G.F. Newell, M. Rosenblatt, Ann. Math. Statist. 33, 1306 (1962).
[84] D. Sornette, L. Knopoff, The paradox of the expected time until the next
earthquake, Bull. Seism. Soc. Am. 87, 789 (1997).
[85] R.A. Fisher, L.H.C. Tippett, Limiting forms of the frequency distribution
of the largest or smallest member of a sample, Proc. Camb. Phil. Soc. 24,
180 (1928).
52
[87] J. Galambos, The asymptotic theory of extreme order statistics, (John
Wiley and Sons, New York, 1978).
[95] H.A. Makse, S. Havlin, M. Schwartz, H.E. Stanley, Method for generating
long-range correlations for large systems, Phys. Rev. E 53, 5445 (1996).
53
[100] J.W. Kantelhardt, E. Koscielny-Bunde, D. Rybski, P. Braun, A. Bunde, S.
Havlin, Long-term persistence and multifractality of precipitation and
river runoff records, J. Geophys. Res. (Atmosph.) 111, D01106 (2006).
Figure 1: (a) Examples of self-affine series xi characterized by different Hurst exponents H = 0.9, 0.7, 0.5, 0.3 (from top to bottom). The data has
been generated by Fourier filtering (see Section 8.1) using the same seed for the random number generator. (b) Differentiated series ∆xi of the data
from (a); the ∆xi are characterized by positive long-term correlations (persistence) with γ = 0.2 and 0.6 (top curve and second curve), uncorrelated
behaviour (third curve), and anti-correlations (bottom curve), respectively.
Figure 2: Comparison of the autocorrelation functions C(s) (decreasing functions) and fluctuation functions F2(s) (increasing functions) for
shortterm correlated data (top) and long-term correlated data (γ = 0.4, bottom). The asymptotic slope H ≈ α = 0.5 of F2(s) clearly indicates the
absence of long-term correlations, while H ≈ α = 1 – γ/2 > 0.5 indicates the presence long-term correlations. The difference is much harder to
observe in C(s), where there are more statistical fluctuations and even negative values (e.g., above s = 150 in (a) and between s = 300 and 400 in
(b), not shown). The data have been generated by an AR process (Eq. (4)) and Fourier filtering (Section 8.1) for (a) and (b), respectively. The
dashed lines indicate the theoretical curves.
54
frequency
Figure 3: Spectral analysis of a fractal time series characterized by long-term correlations with γ = 0.4 (β = 0.6). The expected scaling behaviour
(dashed line indicating the slope –β) is observed only after binning of the spectrum (circles). The data has been generated by Fourier filtering (see
Section 8.1).
55
Figure 4: Illustration of the Fluctuation Analysis (FA) and the Detrended Fluctuation Analysis (DFA). For two segment durations (time scales) s
=
100 (a) and 200 (b), the profiles Y(j) (blue lines; defined in Eq. (10)), the values Y(νs) used for the FA in Eq. (11) (green circles), and least square
quadratic fits to the profiles used in DFA (red lines) are shown.
56
Figure 5: Application of discrete wavelet transform (WT) analysis on uncorrelated data (black circles), long-term correlated data (γ = 0.8, α = 0.6,
red squares), and short-term correlated data (summation of three AR processes, green diamonds). Averages of F2(s) averaged over 20 series with N
= 216 points and divided by s1/2 are shown, so that a horizontal line corresponds to uncorrelated behaviour. The blue open triangles show the result
for one selected extreme configuration, where it is hard to decide about the existence of long-term correlations (figure prepared by Mikhail
Bogachev).
57
Figure 6: Application of Detrended Fluctuation Analysis (DFA) on the data already studied in Fig. 5 (figure by Mikhail Bogachev).
58
Figure 7: Example of the Wavelet Transform Modulus Maxima (WTMM) method, showing the original data (top), its continuous wavelet
transform (grey scale coded amplitude of wavelet coefficients, middle), and the extracted maxima lines (bottom) (figure taken from [67]).
59
Figure 8: Multifractal Detrended Fluctuation Analysis (MF-DFA) of data from the binomial multifractal model (see Section 8.3) with a = 0.75.
Fq(s) is plotted versus s for the q values given in the legend; the slopes of the curves correspond to the values of h(q). The dashed lines have the
slopes of the theoretical slopes h(±∞) from Eq. (42). 100 configurations have been averaged.
60
Figure 9: Illustration for the definition of return intervals rq between extreme events above two quantiles (thresholds) q1 and q2 (figure by Jan
Eichner).
61
Figure 10: Normalized rescaled distribution density functions RqPq(r) of r values with Rq = 100 as a function of r/Rq for long-term correlated data
with γ = 0.4 (open symbols) and γ = 0.2 (filled symbols; we multiplied the data for the filled symbols by a factor 100 to avoid overlapping curves).
In (a) the original data was Gaussian distributed, in (b) exponentially distributed, in (c) power-law distributed with power −5.5, and in (d)
lognormally distributed. All four figures follow quite well stretched exponential curves (solid lines) over several decades. For small r/Rq values a
power-law regime seems to dominate, while on large scales deviations from the stretched exponential behaviour are due to finite-size effects (figure
by Jan Eichner).
62
0 365 730 1095 1460 1825 2190 2555 2920
i [days]
Figure 11: Illustration for the definition of maxima mR within periods of R = 365 values (figure by Jan Eichner).
63