4
Fractional Brownian Motion Versus the Continuous-Time Random Walk: A Simple Test for Subdiffusive Dynamics Marcin Magdziarz, * Aleksander Weron, and Krzysztof Burnecki Hugo Steinhaus Center, Institute of Mathematics and Computer Science, Wroclaw University of Technology, Wyspianskiego 27, 50-370 Wroclaw, Poland Joseph Klafter x School of Chemistry, Raymond and Beverly Sackler Faculty of Exact Sciences, Tel Aviv University, Tel Aviv 69978, Israel; and Freiburg Institute for Advanced Studies (FRIAS), University of Freiburg, 79104 Freiburg, Germany (Received 6 May 2009; revised manuscript received 12 October 2009; published 30 October 2009) Fractional Brownian motion with Hurst index less then 1=2 and continuous-time random walk with heavy tailed waiting times (and the corresponding fractional Fokker-Planck equation) are two different processes that lead to a subdiffusive behavior widespread in complex systems. We propose a simple test, based on the analysis of the so-called p variations, which allows distinguishing between the two models on the basis of one realization of the unknown process. We apply the test to the data of Golding and Cox [Phys. Rev. Lett. 96, 098102 (2006)], describing the motion of individual fluorescently labeled mRNA molecules inside live E. coli cells. It is found that the data does not follow heavy tailed continuous-time random walk. The test shows that it is likely that fractional Brownian motion is the underlying process. DOI: 10.1103/PhysRevLett.103.180602 PACS numbers: 05.40.Fb, 02.50.Ey, 02.70.c, 05.10.a Distinguishing between normal and anomalous diffusion is usually based on the analysis of the mean-squared dis- placement (MSD) of the diffusing particles. In the case of classical diffusion, the second moment is linear in time, whereas anomalous diffusion processes exhibit distinct deviations from this fundamental property. In particular, subdiffusive systems are characterized by the sublinear pattern hx 2 ðtÞi t , 0 << 1,[1]. The origin of subdif- fusive dynamics in a given system is often unknown. It is not always clear which model applies to a particular system [2,3], an information which is essential when diffusion- controlled processes are considered. Therefore, determin- ing the appropriate model of subdiffusive dynamics is an important and timely problem; see [26] for discussion on the origins of anomaly in the case of single protein fluctu- ations and intracellular diffusion. Two distinct processes have been proposed to account for subdiffusion. The first one is the fractional Brownian motion (FBM), [7]. FBM is a generalization of the classical Brownian motion. The MSD of FBM satisfies hx 2 ðtÞi t 2H , where 0 <H< 1 is the Hurst exponent. Thus, for H< 1=2 we obtain the subdiffusive dynamics, [8,9]. The second model of subdiffusion is the continuous-time random walk (CTRW) and the corresponding fractional Fokker-Planck equation (FFPE) [1]. In this model, a par- ticle performs random jumps whose length is given by the probability density function (PDF) with finite second mo- ment. The waiting times between consecutive jumps are assumed to follow a power law t 1 with 0 << 1. These heavy tailed waiting times correspond to nonsta- tionary increments and give rise to sublinear MSD of the particle. As a consequence, the CTRW model exhibits ergodicity breaking and aging. The MSD can be obtained either by performing an average over an ensemble of particles, or by taking the temporal average over a single trajectory [1012]. Recent advances in single molecule spectroscopy enabled single particle tracking experiments following individual particle trajectories [3,4]. These re- quire temporal, moving averages. Although temporal aver- ages of heavy tailed CTRW and FBM have been shown to differ [10,1214], the issue of determining the underlying process is still open. Motivated by growing interest in single molecule spec- troscopy, in particular, by single particle tracking, we propose a method to distinguish between mechanisms leading to subdiffusion. Introducing such a method is timely and goes beyond the very basic claims of ‘‘normal’’ vs ‘‘anomalous’’ diffusion by seeking an origin for the anomalous. We apply our theoretical approach to experi- mental data (random motion of an individual molecule inside the cell by tracking fluorescently labeled mRNA molecules in E. coli in the experiment described in details in [3]) and resolve a recent controversy on the origin of the Golding-Cox subdiffusion [12,13]. We clearly demonstrate that, unlike some claims, the observed subdiffusion cannot stem from a broad distribution of waiting times. It is likely that fractional Brownian motion is the underlying process. Subdiffusive dynamics.—We begin with recalling the two models of subdiffusion, namely, FBM and CTRW. For 0 <H< 1, FBM of index H is the mean-zero Gaussian process B H ðtÞ whose covariance function is given by EðB H ðsÞB H ðtÞÞ ¼ð' 2 =2Þðs 2H þ t 2H jt sj 2H Þ, t, s> 0. Here, ' 2 ¼ EðB 2 H ð1ÞÞ. For H ¼ 1=2, B H ðtÞ reduces to the standard Brownian motion BðtÞ. FBM is self-similar PRL 103, 180602 (2009) PHYSICAL REVIEW LETTERS week ending 30 OCTOBER 2009 0031-9007= 09=103(18)=180602(4) 180602-1 Ó 2009 The American Physical Society

Fractional Brownian Motion Versus the Continuous-Time Random Walk: A Simple Test for Subdiffusive Dynamics

Embed Size (px)

Citation preview

Page 1: Fractional Brownian Motion Versus the Continuous-Time Random Walk: A Simple Test for Subdiffusive Dynamics

Fractional Brownian Motion Versus the Continuous-Time RandomWalk:A Simple Test for Subdiffusive Dynamics

Marcin Magdziarz,* Aleksander Weron,† and Krzysztof Burnecki‡

Hugo Steinhaus Center, Institute of Mathematics and Computer Science, Wroclaw University of Technology,Wyspianskiego 27, 50-370 Wroclaw, Poland

Joseph Klafterx

School of Chemistry, Raymond and Beverly Sackler Faculty of Exact Sciences, Tel Aviv University, Tel Aviv 69978, Israel;and Freiburg Institute for Advanced Studies (FRIAS), University of Freiburg, 79104 Freiburg, Germany(Received 6 May 2009; revised manuscript received 12 October 2009; published 30 October 2009)

Fractional Brownian motion with Hurst index less then 1=2 and continuous-time random walk with

heavy tailed waiting times (and the corresponding fractional Fokker-Planck equation) are two different

processes that lead to a subdiffusive behavior widespread in complex systems. We propose a simple test,

based on the analysis of the so-called p variations, which allows distinguishing between the two models

on the basis of one realization of the unknown process. We apply the test to the data of Golding and Cox

[Phys. Rev. Lett. 96, 098102 (2006)], describing the motion of individual fluorescently labeled mRNA

molecules inside live E. coli cells. It is found that the data does not follow heavy tailed continuous-time

random walk. The test shows that it is likely that fractional Brownian motion is the underlying process.

DOI: 10.1103/PhysRevLett.103.180602 PACS numbers: 05.40.Fb, 02.50.Ey, 02.70.�c, 05.10.�a

Distinguishing between normal and anomalous diffusionis usually based on the analysis of the mean-squared dis-placement (MSD) of the diffusing particles. In the case ofclassical diffusion, the second moment is linear in time,whereas anomalous diffusion processes exhibit distinctdeviations from this fundamental property. In particular,subdiffusive systems are characterized by the sublinearpattern hx2ðtÞi � t�, 0<�< 1, [1]. The origin of subdif-fusive dynamics in a given system is often unknown. It isnot always clear which model applies to a particular system[2,3], an information which is essential when diffusion-controlled processes are considered. Therefore, determin-ing the appropriate model of subdiffusive dynamics is animportant and timely problem; see [2–6] for discussion onthe origins of anomaly in the case of single protein fluctu-ations and intracellular diffusion.

Two distinct processes have been proposed to accountfor subdiffusion. The first one is the fractional Brownianmotion (FBM), [7]. FBM is a generalization of the classicalBrownian motion. The MSD of FBM satisfies hx2ðtÞi �t2H, where 0<H < 1 is the Hurst exponent. Thus, forH <1=2 we obtain the subdiffusive dynamics, [8,9].

The second model of subdiffusion is the continuous-timerandom walk (CTRW) and the corresponding fractionalFokker-Planck equation (FFPE) [1]. In this model, a par-ticle performs random jumps whose length is given by theprobability density function (PDF) with finite second mo-ment. The waiting times between consecutive jumps areassumed to follow a power law t���1 with 0<�< 1.These heavy tailed waiting times correspond to nonsta-tionary increments and give rise to sublinear MSD of theparticle. As a consequence, the CTRW model exhibits

ergodicity breaking and aging. The MSD can be obtainedeither by performing an average over an ensemble ofparticles, or by taking the temporal average over a singletrajectory [10–12]. Recent advances in single moleculespectroscopy enabled single particle tracking experimentsfollowing individual particle trajectories [3,4]. These re-quire temporal, moving averages. Although temporal aver-ages of heavy tailed CTRW and FBM have been shown todiffer [10,12–14], the issue of determining the underlyingprocess is still open.Motivated by growing interest in single molecule spec-

troscopy, in particular, by single particle tracking, wepropose a method to distinguish between mechanismsleading to subdiffusion. Introducing such a method istimely and goes beyond the very basic claims of ‘‘normal’’vs ‘‘anomalous’’ diffusion by seeking an origin for theanomalous. We apply our theoretical approach to experi-mental data (random motion of an individual moleculeinside the cell by tracking fluorescently labeled mRNAmolecules in E. coli in the experiment described in detailsin [3]) and resolve a recent controversy on the origin of theGolding-Cox subdiffusion [12,13]. We clearly demonstratethat, unlike some claims, the observed subdiffusion cannotstem from a broad distribution of waiting times. It is likelythat fractional Brownian motion is the underlying process.Subdiffusive dynamics.—We begin with recalling the

two models of subdiffusion, namely, FBM and CTRW.For 0<H < 1, FBM of index H is the mean-zero

Gaussian process BHðtÞwhose covariance function is givenby EðBHðsÞBHðtÞÞ ¼ ð�2=2Þðs2H þ t2H � jt� sj2HÞ, t,s > 0. Here, �2 ¼ EðB2

Hð1ÞÞ. For H ¼ 1=2, BHðtÞ reducesto the standard Brownian motion BðtÞ. FBM is self-similar

PRL 103, 180602 (2009) P HY S I CA L R EV I EW LE T T E R Sweek ending

30 OCTOBER 2009

0031-9007=09=103(18)=180602(4) 180602-1 � 2009 The American Physical Society

Page 2: Fractional Brownian Motion Versus the Continuous-Time Random Walk: A Simple Test for Subdiffusive Dynamics

with respect to H [9], i.e., for every c > 0 we have

BHðctÞ ¼d cHBHðtÞ. Here, ¼d stands for ‘‘equal in distri-bution.’’ Moreover, FBM has stationary increments. Thestationary sequence of FBM increments bHðjÞ ¼ BHðjþ1Þ � BHðjÞ is very strongly correlated. One can show thatthe autocovariance function of bHð�Þ satisfies rðjÞ ¼EðbHðjÞbHð0ÞÞ� �2Hð2H � 1Þj2H�2 as j ! 1.

For the second moment of the FBMwe have EðB2HðtÞÞ ¼

�2t2H, which for H < 1=2 gives the subdiffusive dynam-ics. We assume that �2 ¼ 1. Note that one can alwaysnormalize the process in such a manner by dividing it by�> 0. The parameter � can be estimated using the prop-erty BHðtþ sÞ � BHðtÞ � Nð0; �2s2HÞ.

The second fundamental model of subdiffusive dynam-ics is the CTRWand the corresponding FFPE. A force-freeFFPE has the form [1]:

@wðx; tÞ@t

¼ 0D1��t

�1

2

@2

@x2

�wðx; tÞ (1)

with the initial condition wðx; 0Þ ¼ �ðxÞ. Here, the opera-tor 0D

1��t , 0<�< 1, is the fractional derivative of the

Riemann-Liouville type. The derivation of the above equa-tion is based on the CTRW scheme with heavy tailedwaiting times. It is easy to verify [1] that the MSD corre-sponding to wðx; tÞ is equal to t�

�ð�þ1Þ .In Eq. (1), wðx; tÞ denotes the PDF of a subdiffusive

stochastic process Z�ðtÞ. The process Z�ðtÞ can be equiva-lently written in the form of subordination [15–17]

Z�ðtÞ ¼ BðS�ðtÞÞ; (2)

where BðtÞ is the standard Brownian motion and S�ðtÞ isthe inverse �-stable subordinator independent of BðtÞ. Theinverse �-stable subordinator is defined as

S�ðtÞ ¼ inff� > 0: U�ð�Þ> tg; (3)

0<�< 1, where U�ð�Þ is the �-stable subordinator [18]with Laplace transform Eðe�uU�ð�ÞÞ ¼ e��u� . The processS�ðtÞ is �-self-similar, and therefore Z�ðtÞ is �=2-self-similar. For every jump of U�ð�Þ there is a correspondingflat period of its inverse. These flat periods of S�ðtÞ arecharacteristic for the subdiffusive dynamics and corre-spond to the heavy tailed waiting times in the underlyingCTRW scenario. The Langevin-type process (2) corre-sponding to FFPE (1) gives insight into the structure oftrajectories. Therefore, it allows to detect differences be-tween single trajectories of FBM BHðtÞ and CTRW-basedmodel Z�ðtÞ.

p Variation.—Let us now discuss the idea of p variation,p > 0, which will be our main tool in a procedure ofidentifying the type of subdiffusion. The concept of pvariation generalizes the well-known notion of total varia-tion, which has found applications in various branches ofmathematics, physics and engineering, like optimal con-trol, numerical analysis of differential equations, and cal-culus of variations [19]. Let XðtÞ be a stochastic process

observed on time interval ½0; T�. Then, for t 2 ½0; T�, the pvariation corresponding to XðtÞ is defined as

VðpÞðtÞ ¼ limn!1V

ðpÞn ðtÞ; (4)

where VðpÞn ðtÞ is the partial sum of increments of the

process XðtÞ given by

VðpÞn ðtÞ ¼ X2n�1

j¼0

��������X

�ðjþ 1ÞT2n

^ t

�� X

�jT

2n^ t

���������p

(5)

with a ^ b ¼ minfa; bg. Let us underline that VðpÞn ðtÞ is

very easy to calculate, since it is just the finite sum ofpth powers of the increments of XðtÞ. For large enough n,

VðpÞn ðtÞ approximates nicely p variation VðpÞðtÞ. When p ¼

1, Vð1ÞðtÞ reduces to the total variation.As an example let us recall the variational properties of

the standard Brownian motion. It is a well known fact thatthe total variation of Brownian motion is infinite, which isnot very surprising given the ‘‘wild’’ behavior of thetrajectories of BðtÞ. However, the quadratic variation of

BðtÞ is finite and equals Vð2ÞðtÞ ¼ t [18].It is well known that the p variation of the FBM BHðtÞ

satisfies [20]

VðpÞðtÞ ¼8<:þ1 if p < 1

H ;

tEðjBHð1Þj1=HÞ if p ¼ 1H ;

0 if p > 1H ;

(6)

The expected value in the above expression is given by

EðjBHð1Þj1=HÞ ¼ 2 12Hffiffiffi�

p �ð 12H þ 1

2Þ. Let us note that for the

considered here subdiffusive case H < 1=2, the quadratic

variation Vð2ÞðtÞ of BHðtÞ is infinite.The p variation of the Langevin process Z�ðtÞ ¼

BðS�ðtÞÞ satisfies [21]

VðpÞðtÞ ¼8<:þ1 if p < 2;S�ðtÞ if p ¼ 2;0 if p > 2:

(7)

The above formula confirms that the quadratic variation ofZ�ðtÞ is finite and equal to the inverse subordinator S�ðtÞ,[22]. We underline that in this case Vð2ÞðtÞ is a stochasticprocess and not the deterministic function as in (6).

Moreover, Vð2ÞðtÞ ¼ S�ðtÞ is an �-self-similar process.Test.—Suppose we are given one realization (time se-

ries) of some subiffusive process XðtÞ observed on the timeinterval ½0; T�. If not known, estimate the index of self-similarity of the process XðtÞ, [10,12,23,24]. Recall that theestimated self-similarity index will give us the approxi-mate value of H or �=2 depending on the type of sub-diffusion. Our goal is to verify if the subdiffusive dynamicsoriginates from the FBM process BHðtÞ or the CTRW-based model Z�ðtÞ. Using Eqs. (6) and (7) we proposethe following procedure:

PRL 103, 180602 (2009) P HY S I CA L R EV I EW LE T T E R Sweek ending

30 OCTOBER 2009

180602-2

Page 3: Fractional Brownian Motion Versus the Continuous-Time Random Walk: A Simple Test for Subdiffusive Dynamics

p variation test.—Calculate the partial sums Vð1=HÞn ðtÞ

and Vð2Þn ðtÞ, which approximate 1=H variation and 2 varia-

tion of XðtÞ, respectively. Here, H is the previously esti-mated self-similarity index. (i) If the process XðtÞ is the

FBM, then Vð1=HÞn ðtÞ � tEðjBHð1Þj1=HÞ and Vð2Þ

n ðtÞ shouldincrease with increasing n. (ii) If the process XðtÞ origi-nates from the CTRW model Z�ðtÞ, then Vð1=HÞ

n ðtÞ shouldtend to zero with increasing n, whereas Vð2Þ

n ðtÞ should

stabilize [recall that for Z�ðtÞ we have Vð2ÞðtÞ ¼ S�ðtÞ].The implementation of the above test is based on the

computation of the finite sums VðpÞn ðtÞ (5), which is rather

straightforward for analytical models as well as for em-pirical data.

For H very close to 1=2 it is necessary to take large

enough n while calculating the partial sum VðpÞn ðtÞ.

Otherwise, one can not practically distinguish the proper-ties of p variation corresponding to BH and Z�.

In practice an analyzed empirical trajectory is given as atime series Xðt1Þ; Xðt2Þ; . . . ; Xðt2N Þ. The sequence t1 <t2 < . . .< t2N represents the time points, in which positionof the test particle is observed. In such setting, N is the

largest value for which the sum VðpÞN ðtÞ can be calculated.

Then, for fixed t ¼ ti, we have VðpÞN ðtÞ ¼ P

i�1k¼1 jXðtkþ1Þ �

XðtkÞjp. Similarly, to determine VðpÞN�1ðtÞ, one has to sum up

the pth powers of the increments jXðt3Þ � Xðt1Þj; jXðt5Þ �Xðt3Þj; . . . ; jXðt2N�1Þ � Xðt2N�3Þj. Then, for fixed t ¼ ti,

with i ¼ 2jþ 1, we have VðpÞN�1ðtÞ ¼

Pjk¼1 jXðt2kþ1Þ �

Xðt2k�1Þjp. Consequently, to determine VðpÞN�2ðtÞ, one

sums up the pth powers of the increments jXðt5Þ �Xðt1Þj; jXðt9Þ � Xðt5Þj; . . . , etc. Finally, plotting VðpÞ

n ðtÞ

for different values of n and observing how it behaveswhile n increases/decreases, one can draw conclusions onthe origins of subdiffusion.First, we tested the algorithm on simulated data. We

simulated one trajectory of FBM BHðtÞ and one trajectory

of Z�ðtÞ. We demonstrate the results in Fig. 1. The qua-

dratic variation of BHðtÞ diverges, whereas the quadratic

variation of Z�ðtÞ is equal to S�ðtÞ. Moreover, the1=H variation of BHðtÞ is a linear function, while the1=H variation of Z�ðtÞ vanishes. These differences in thebehavior of variations of both subdiffusive processes BHðtÞand Z�ðtÞ allow to distinguish between mechanisms lead-ing to subdiffusion.Next, we applied the test to the Golding-Cox experi-

mental data [3]. We analyzed six two-dimensional samplepaths (all those having more than 29 ¼ 512 points, whichseems reasonable for the p variation test) from their set of27 trajectories. We examined X and Y coordinates as wellas the two-dimensional trajectories separately. The testclearly demonstrated (see the supplementary material inRef. [25] for the details, and Fig. 2 for the analysis of onesample trajectory) that the subdiffusion cannot stem fromthe CTRW model. Moreover, the test also shows that thereis no reason to reject the hypothesis that the data followsFBM. This resolves a recent controversy over the under-lying reason for the Golding-Cox subdiffusion [12,13].However, to reach a more conclusive statement on theFBM origins of the experimental data, longer trajectoriesand extended statistical analysis are necessary.The conclusion also concurs with the result of [10]

contrasting temporal average of heavy tailed CTRW withthat of FBM. We strongly believe that our approach pro-

c d

FIG. 1 (color online). In panels (a)–(b) the analysis of a simulated trajectory of FBM BHðtÞ, with Hurst index H ¼ 0:3, is presented.

Panel (a) shows the value of Vð1=HÞn ðtÞ, n ¼ 12, corresponding to the sample path of BHðtÞ (solid blue line). The dotted red line is the

theoretical 1=H variation of FBM given in Eq. (6). We observe excellent agreement between the two lines. The approximation gets

even better for larger n. Panel (b) depicts the value of Vð2Þn ðtÞ corresponding to the simulated trajectory of BHðtÞ, calculated for different

n ¼ 10; 12; . . . ; 18. We observe the rapidly increasing values of Vð2Þn ðtÞ while n increases. This demonstrates the fact that the quadratic

variation of BHðtÞ is infinite forH < 1=2. Panels (c)–(d) depict the analysis of a simulated trajectory of the process Z�ðtÞ with � ¼ 0:6.

In panel (c) we see the value of Vð1=HÞn ðtÞ corresponding to the sample path of Z�ðtÞ calculated for different n ¼ 10; 12; . . . ; 18. We

observe that Vð1=HÞn ðtÞ tends to zero while n increases. This confirms the fact that the 1=H variation of Z�ðtÞ is equal to zero. Panel (d)

shows the value of Vð2Þn ðtÞ, n ¼ 12, calculated for the simulated trajectory of Z�ðtÞ (solid blue line). The dotted red line is the trajectory

of the inverse subordinator S�ðtÞ. We observe excellent agreement between the two lines, which confirms that the quadratic variation ofZ�ðtÞ is equal to S�ðtÞ. For larger n the approximation is even better. The observed differences in the behavior of quadratic and1=H variations corresponding to BHðtÞ and Z�ðtÞ allow to distinguish between mechanisms leading to subdiffusion.

PRL 103, 180602 (2009) P HY S I CA L R EV I EW LE T T E R Sweek ending

30 OCTOBER 2009

180602-3

Page 4: Fractional Brownian Motion Versus the Continuous-Time Random Walk: A Simple Test for Subdiffusive Dynamics

vides a way, missing up to now, to look deeper intoprocesses leading to single particle diffusion.

Finally, we note that the same methodology based on pvariations can be applied to analyze another model ofsubdiffusion—random walks on fractal structures. Ourpreliminary results show that the quadratic variation of arandom walk on a Sierpinski gasket embedded in twodimensions is infinite (similar to the FBM and differentfrom the CTRW). The p variation is finite for p ¼ dw,where dw ¼ log5= log2 ¼ 2:32193 . . . is the walk dimen-sion, which corresponds to the self-similarity index of theSierpinski gasket.

The authors would like to thank Ido Golding andEdward C. Cox for providing the data. The work ofM.M. and K. B. was partially supported by thePOIG.01.03.01-02-002/08 contract.

*[email protected][email protected][email protected]@post.tau.ac.il

[1] R. Metzler and J. Klafter, Phys. Rep. 339, 1 (2000).[2] G. Guigas, C. Kalla, and M. Weiss, Biophys. J. 93, 316

(2007).[3] I. Golding and E. C. Cox, Phys. Rev. Lett. 96, 098102

(2006).[4] A. Caspi, R. Granek, and M. Elbaum, Phys. Rev. Lett. 85,

5655 (2000).[5] W. Min et al., Phys. Rev. Lett. 94, 198302 (2005).[6] M. J. Saxton, Biophys. J. 92, 1178 (2007).[7] B. B. Mandelbrot and J.W. Van Ness, SIAM Rev. 10, 422

(1968).[8] E. Lutz, Phys. Rev. E 64, 051106 (2001).[9] A. Weron, K. Burnecki, Sz. Mercik, and K. Weron, Phys.

Rev. E 71, 016113 (2005).[10] A. Lubelski, I.M. Sokolov, and J. Klafter, Phys. Rev. Lett.

100, 250602 (2008).[11] R. Metzler et al., Acta Phys. Polon. B 40, 1315 (2009).[12] Y. He, S. Burov, R. Metzler, and E. Barkai, Phys. Rev.

Lett. 101, 058101 (2008).[13] T. Neusius, I.M. Sokolov, and J. C. Smith, Phys. Rev. E

80, 011109 (2009).[14] J. Szymanski and M. Weiss, Phys. Rev. Lett. 103, 038102

(2009).[15] M.M. Meerschaert, D. A. Benson, H. P. Scheffler, and B.

Baeumer, Phys. Rev. E 65, 041103 (2002).[16] M. Magdziarz, A. Weron, and K. Weron, Phys. Rev. E 75,

016708 (2007).[17] E. Gudowska-Nowak, B. Dybiec, P. F. Gora, and R.

Zygadlo, Acta Phys. Polon. B 40, 1263 (2009).[18] A. Janicki and A. Weron, Simulation and Chaotic

Behaviour of �-Stable Stochastic Processes (MarcelDekker, New York, 1994).

[19] T. F. Chan and J. Shen, Image Processing. Variational,PDE, Wavelet and Stochastic Methods (SIAM,Philadelphia, 2005).

[20] L. C. G. Rogers, Math. Finance 7, 95 (1997).[21] M. Magdziarz, Path Properties of Subdiffusion—A

Martingale Approach, report, 2009.[22] A. Weron and M. Magdziarz, Europhys. Lett. 86, 60010

(2009).[23] J. Beran, Statistics for Long-Memory Processes (Chapman

& Hall, New York, 1994).[24] Sz. Mercik, K. Weron, K. Burnecki, and A. Weron, Acta

Phys. Polon. B 34, 3773 (2003).[25] See EPAPS Document No. E-PRLTAO-103-028946 for

the p variation test of the experimental data. For moreinformation on EPAPS, see http://www.aip.org/pubservs/epaps.html.

FIG. 2 (color online). Panel (a) shows the 1=H variation

Vð1=HÞn ðtÞ of one sample trajectory taken from Golding-Cox

empirical data (with H ¼ 0:35 as in [3]). Parameters: n ¼ 10(blue line); n ¼ 9 (red line); n ¼ 8 (green line); n ¼ 7 (blackline). We observe that the 1=H variation does not exhibit any

trend, meaning that Vð1=HÞn ðtÞ neither increases nor decreases with

increasing n. Similar behavior is observed for simulated trajec-tories of the FBM with the same number of points. In panel (b)

the 2 variation Vð2Þn ðtÞ of the analyzed trajectory is presented.

Parameters as in panel (a). The 2 variation increases withincreasing n, which confirms that the 2 variation is not finite.Thus, the data does not follow CTRW model.

PRL 103, 180602 (2009) P HY S I CA L R EV I EW LE T T E R Sweek ending

30 OCTOBER 2009

180602-4