1 2physics department, technical university of munich, d ... · arxiv:1009.4846v2...

15
arXiv:1009.4846v2 [cond-mat.stat-mech] 13 Oct 2010 22nd October 2018 Single particle tracking in systems showing anomalous diffusion: the role of weak ergodicity breaking Stas Burov, 1, Jae-Hyung Jeon, 2 Ralf Metzler, 2, and Eli Barkai 1, 1 Department of Physics, Bar Ilan University, Ramat-Gan 52900, Israel 2 Physics Department, Technical University of Munich, D-85747 Garching, Germany Anomalous diffusion has been widely observed by single particle tracking microscopy in complex systems such as biological cells. The resulting time series are usually evaluated in terms of time av- erages. Often anomalous diffusion is connected with non-ergodic behaviour. In such cases the time averages remain random variables and hence irreproducible. Here we present a detailed analysis of the time averaged mean squared displacement for systems governed by anomalous diffusion, con- sidering both unconfined and restricted (corralled) motion. We discuss the behaviour of the time averaged mean squared displacement for two prominent stochastic processes, namely, continuous time random walks and fractional Brownian motion. We also study the distribution of the time averaged mean squared displacement around its ensemble mean, and show that this distribution preserves typical process characteristic even for short time series. Recently, velocity correlation functions were suggested to distinguish between these processes. We here present analytucal ex- pressions for the velocity correlation functions. Knowledge of the results presented here are expected to be relevant for the correct interpretation of single particle trajectory data in complex systems. PACS numbers: 02.50.-r, 05.40.Fb, 05.10.Gg I. INTRODUCTION Single particle tracking microscopy provides the pos- ition time series r(t) of individual particle trajectories in a medium [1–4]. The information garnered by single particle tracking yields insights into the mechanisms and forces, that drive or constrain the motion of the particle. An early example of systematic single particle tracking is given by the work of Jean Perrin on diffusive motion [5]. Due to the relatively short individual trajectories, Perrin used an ensemble average over many trajectories to ob- tain meaningful statistics. A few years later, Nordlund conceived a method to record much longer time series [6], allowing him to evaluate individual trajectories in terms of the time average and thus to avoid averages over not perfectly identical particles. Today, single particle track- ing has become a standard tool to characterise the micro- scopic rheological properties of a medium [7], or to probe the active motion of biomolecular motors [8]. Particu- larly in biological cells and complex fluids single particle trajectory methods have become instrumental in uncov- ering deviations from normal Brownian motion of pass- ively moving particles [9–22]. Classical diffusion patterns are sketched in the left panel of Fig. 1. Accordingly, one may observe free diffu- sion, leading to a linear growth with time of the second moment [Line 1 in Fig. 1]. Brownian motion may also be restricted (corralled, confined). Confinement in a cell, for instance, could be due to the cell walls. In that case * Submitted to the special issue: Single Molecule Optical studies of Soft and Complex Matter Electronic address: [email protected] Electronic address: [email protected] 0.1 1 10 100 0.1 1 10 100 δ 2 () [a.u.] (1) (2) (3) 0.1 1 10 100 1000 [a.u.] (1’’) (2’) (1’) (3’) (2’’) Figure 1: Diffusion modes of the time averaged mean squared displacement (2) as function of the lag time ∆. Left: Nor- mal diffusion growing like δ 2 ∆ (1), restricted (confined) diffusion with a turnover from ∆ to 0 (2), drift dif- fusion 2 (3). Right: Ergodic subdiffusion α (1’), restricted ergodic subdiffusion turning over from α to 0 (2’), non-ergodic subdiffusion ∆ (1”), restricted non- ergodic subdiffusion turning over from ∆ to 1-α (2”), superdiffusion ∆ 1+α (3’). Here, 0 <α< 1. Note the double- logarithmic scale. the second moment initially grows linearly with time and eventually saturates to its thermal value equalling the second moment of the corresponding Boltzmann distri- bution [Line 2]. In the presence of a drift the second mo- ment grows with the square of time [Line 3]. Such results are typical for simple fluids. In more complex environ- ments, different patterns may be observed, as displayed on the right of Fig. 1. Here, subdiffusion may occur,

Upload: others

Post on 19-Nov-2020

2 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: 1 2Physics Department, Technical University of Munich, D ... · arXiv:1009.4846v2 [cond-mat.stat-mech] 13 Oct 2010 22nd October 2018 Single particle tracking in systems showing anomalous

arX

iv:1

009.

4846

v2 [

cond

-mat

.sta

t-m

ech]

13

Oct

201

022nd October 2018

Single particle tracking in systems showing anomalous diffusion: the role of weak

ergodicity breaking

Stas Burov,1, ∗ Jae-Hyung Jeon,2 Ralf Metzler,2, † and Eli Barkai1, ‡

1Department of Physics, Bar Ilan University, Ramat-Gan 52900, Israel2Physics Department, Technical University of Munich, D-85747 Garching, Germany

Anomalous diffusion has been widely observed by single particle tracking microscopy in complexsystems such as biological cells. The resulting time series are usually evaluated in terms of time av-erages. Often anomalous diffusion is connected with non-ergodic behaviour. In such cases the timeaverages remain random variables and hence irreproducible. Here we present a detailed analysis ofthe time averaged mean squared displacement for systems governed by anomalous diffusion, con-sidering both unconfined and restricted (corralled) motion. We discuss the behaviour of the timeaveraged mean squared displacement for two prominent stochastic processes, namely, continuoustime random walks and fractional Brownian motion. We also study the distribution of the timeaveraged mean squared displacement around its ensemble mean, and show that this distributionpreserves typical process characteristic even for short time series. Recently, velocity correlationfunctions were suggested to distinguish between these processes. We here present analytucal ex-pressions for the velocity correlation functions. Knowledge of the results presented here are expectedto be relevant for the correct interpretation of single particle trajectory data in complex systems.

PACS numbers: 02.50.-r, 05.40.Fb, 05.10.Gg

I. INTRODUCTION

Single particle tracking microscopy provides the pos-ition time series r(t) of individual particle trajectoriesin a medium [1–4]. The information garnered by singleparticle tracking yields insights into the mechanisms andforces, that drive or constrain the motion of the particle.An early example of systematic single particle tracking isgiven by the work of Jean Perrin on diffusive motion [5].Due to the relatively short individual trajectories, Perrinused an ensemble average over many trajectories to ob-tain meaningful statistics. A few years later, Nordlundconceived a method to record much longer time series [6],allowing him to evaluate individual trajectories in termsof the time average and thus to avoid averages over notperfectly identical particles. Today, single particle track-ing has become a standard tool to characterise the micro-scopic rheological properties of a medium [7], or to probethe active motion of biomolecular motors [8]. Particu-larly in biological cells and complex fluids single particletrajectory methods have become instrumental in uncov-ering deviations from normal Brownian motion of pass-ively moving particles [9–22].

Classical diffusion patterns are sketched in the leftpanel of Fig. 1. Accordingly, one may observe free diffu-sion, leading to a linear growth with time of the secondmoment [Line 1 in Fig. 1]. Brownian motion may alsobe restricted (corralled, confined). Confinement in a cell,for instance, could be due to the cell walls. In that case

∗Submitted to the special issue: Single Molecule Optical studies of

Soft and Complex Matter†Electronic address: [email protected]‡Electronic address: [email protected]

0.1

1

10

100

0.1 1 10 100

δ2 (∆)

∆ [a.u.]

(1)

(2)

(3)

0.1 1 10 100 1000∆ [a.u.]

(1’’)

(2’)

(1’)

(3’)

(2’’)

Figure 1: Diffusion modes of the time averaged mean squareddisplacement (2) as function of the lag time ∆. Left: Nor-

mal diffusion growing like δ2 ≃ ∆ (1), restricted (confined)diffusion with a turnover from ≃ ∆ to ≃ ∆0 (2), drift dif-fusion ≃ ∆2 (3). Right: Ergodic subdiffusion ≃ ∆α (1’),restricted ergodic subdiffusion turning over from ≃ ∆α to≃ ∆0 (2’), non-ergodic subdiffusion ≃ ∆ (1”), restricted non-ergodic subdiffusion turning over from ≃ ∆ to ≃ ∆1−α (2”),superdiffusion ∆1+α (3’). Here, 0 < α < 1. Note the double-logarithmic scale.

the second moment initially grows linearly with time andeventually saturates to its thermal value equalling thesecond moment of the corresponding Boltzmann distri-bution [Line 2]. In the presence of a drift the second mo-ment grows with the square of time [Line 3]. Such resultsare typical for simple fluids. In more complex environ-ments, different patterns may be observed, as displayedon the right of Fig. 1. Here, subdiffusion may occur,

Page 2: 1 2Physics Department, Technical University of Munich, D ... · arXiv:1009.4846v2 [cond-mat.stat-mech] 13 Oct 2010 22nd October 2018 Single particle tracking in systems showing anomalous

2

for which the second moment grows slower than linearlywith time [Line 1’]. Restricted subdiffusion would departfrom this behaviour to reach a plateau [Line 2’]. Drivenmotion may lead to a superdiffusive power-law form ofthe second moment with an exponent between 1 and 2[Line 3’]. However, as we demonstrate below, subdiffu-sion may also be non-ergodic, and the associated timeaveraged second moment may grow linearly with time[Line 1”]. Similarly strange behaviour may be observedfor restricted non-ergodic subdiffusion, which exhibits apower-law growth, not a saturation to a plateau [Line2”]. Non-ergodic processes come along with a significantscatter between individual trajectories. This is an effectof the ageing nature of the process that persists for longmeasurement times. In the following we discuss in detailthe behaviour of passive subdiffusive motion in terms oftime and ensemble averages and address the peculiarities,that may arise for non-ergodic systems.Non-ergodic behaviour of the above sense is indeed ob-

served experimentally. Fig. 2 shows the time averagedmean squared displacement for lipid granules in a liv-ing fission yeast cell. The motion is recorded by indir-ect tracking in an optical tweezers setup. Initially thegranule is located in the bottom of the laser trap poten-tial such that the granule moves freely. Eventually thegranule ventures away from the centre of the trap andexperiences the Hookean trap force. As demonstratedin a detailed analysis the granule motion indeed exhibitsweak ergodicity breaking, giving rise to the characteristicturnover from an initially linear scaling δ2 ≃ ∆ with thelag time ∆, to the power-law regime δ2 ≃ ∆1−α [23].Moreover a pronounced trajectory-to-trajectory scatteris observed, again typical for systems with weak ergodi-city breaking.Free diffusion is typically quantified in terms of the

second moment. The mean squared displacement

〈r2(t)〉 ≡

r2P (r, t)d3r (1)

is obtained as the spatial average over the probabilitydensity function P (r, t) to find the particle at positionr at time t. The quantity (1) therefore corresponds tothe ensemble averaged second moment of the particleposition, denoted by angular brackets, 〈·〉. In particu-lar, the time t enters into Eq. (1) only as a parameter.Conversely, single particle trajectories r(t) are usuallyevaluated in terms of the time averaged mean squareddisplacement defined as

δ2(∆, T ) ≡1

T −∆

∫ T−∆

0

(

r(t+∆)− r(t))2

dt, (2)

where we use an overline · to symbolise the time aver-age. Here ∆ is the so-called lag time constituting a timewindow swept along the time series, and T is the over-all measurement time. The time averaged mean squareddisplacement thus compares the particle positions alongthe trajectory as separated by the time difference ∆.

10-1 100 101 102 103

10-3

10-2

10-1

<>

()

(ms)

Figure 2: Time averaged mean squared displacement of lipidgranule motion in fission yeast cell S.pombe, measured by op-tical tweezers. As function of the lag time ∆ the initiallylinear scaling δ2 ≃ ∆ turns over to the power-law regimeδ2 ≃ ∆1−α, induced by the restoring force exerted on thegranules by the laser trap [23]. Note also the characteristicscatter between individual trajectories. Inset: Average beha-viour of the shown trajectories. In both graphs the unit ofδ2(∆, T ) is Volt2, i.e., the direct output voltage of the quad-rant photodiode. The voltage is directly proportional to thedistance from the centre of the laser trap.

In an ergodic system the time average of a certainquantity obtained from sufficiently long time series isequal to the corresponding ensemble mean [24, 25]. Forinstance, for the mean squared displacement ergodicitywould imply

limT→∞

δ2(∆ = t, T ) = 〈r2(t)〉. (3)

Brownian motion is ergodic, as well as certain station-ary processes leading to anomalous diffusion, such asfractional Brownian motion considered below. There ex-ist, however, non-ergodic processes, which are intimatelyconnected to ageing properties. In what follows we dis-cuss two prominent models for anomalous diffusion, thestationary fractional Brownian motion and the ageingcontinuous time random walk, and analyse in detail thefeatures of the associated time averaged mean squareddisplacement.Generally, anomalous diffusion denotes deviations from

the classical linear dependence of the mean squared dis-placement, 〈r2(t)〉 ≃ t. Such anomalies include ultraslow

diffusion of the form 〈r2(t)〉 ≃ logβ t [26]. In contrast,anomalous diffusion processes may become faster thanballistic, for instance for systems with correlated jumplengths or in systems governed by generalised Langevinequations [27, 28]. Here we are interested in anomalousdiffusion with power-law dependence on time [29],

〈r2(t)〉 ∼ 2dKα

Γ(1 + α)tα, (4)

Page 3: 1 2Physics Department, Technical University of Munich, D ... · arXiv:1009.4846v2 [cond-mat.stat-mech] 13 Oct 2010 22nd October 2018 Single particle tracking in systems showing anomalous

3

for which the anomalous diffusion exponent belongs tothe subdiffusive range 0 < α < 1, such that the limit α =1 corresponds to Brownian motion. The proportionalityfactor Kα in Eq. (4) is the anomalous diffusion coefficientof physical dimension cm2/secα. The embedding spatialdimension is d, and Eq. (4) includes the complete Gammafunction Γ(z).

Subdiffusion of the form (4) with 0 < α < 1 occurs inthe following biologically relevant systems. Fluorescentlylabelled mRNA in E.coli bacteria cells was observed tofollow δ2(∆, T ) ≃ ∆α with α ≈ 0.7 [9]. This result isconsistent with more recent findings according to whichfree RNA tracers in living cells exhibit α ≈ 0.8, whileDNA loci show α = 0.4 [10]. Telomeres in the nucleus ofmammalian cells were reported to follow anomalous dif-fusion with α ≈ 0.3 at shorter times and value α ≈ 0.5 atintermediate times, before a turnover to normal diffusionoccurs [11]. Also larger tracer particles show anomal-ous diffusion, such as adeno-associated viruses of radius≈ 15 nm in a cell with α = 0.5 . . .0.9 [12] and endo-genous lipid granules, of typical size of few hundred nmwith α ≈ 0.75 . . .0.85 [13–15]. It should be noted thatthis subdiffusion observed by single particle tracking mi-croscopy is consistent with results from other techniques,such as fluorescence correlation spectroscopy [17–20] ordynamic light scattering [21]. Subdiffusion of biopoly-mers larger than some 10 kD in living cells is due tomolecular crowding, the excluded volume effect in thesuperdense cellular environment [30–32]. Larger tracerparticles also experience subdiffusion due to interactionwith the semiflexible cytoskeleton [22].

Knowledge of the time or ensemble averaged meansquared displacement of an anomalous diffusion process isinsufficient to fully characterise the underlying stochasticmechanism, as the associated probability density P (r, t)is no longer necessarily Gaussian, and therefore no longerspecified by the first and second moments, only [29]. Thisproperty is in contrast to the universal Gaussian natureof Brownian motion which is effected by the central limittheorem. At the same time the very nature of the anom-alous diffusion process may result in decisively differentbehaviours for diffusional mixing and diffusion-limited re-actions [33]. In biological cells this would imply signific-ant differences for signalling and regulatory processes.For a better understanding of the dynamics in biologicalcells and other complex fluids knowledge of the underly-ing stochastic mechanism is therefore imperative.

Here we discuss the properties of continuous timerandom walk (CTRW) processes with diverging char-acteristic waiting time with respect to their time av-eraged behaviour, expanding on our earlier work [34–36]. We show that for free motion the lag time depend-ence of the time averaged mean squared displacement,δ2(∆, T ) ≃ ∆, is insensitive to the anomalous diffusionexponent α for CTRW processes. In contrast, for con-fined CTRW subdiffusion a universal scaling behaviouremerges, δ2(∆, T ) ≃ ∆1−α, with dynamic exponent 1−α[curves (1”) and (2”) in Fig. 1]. Subdiffusion governed

by fractional Brownian motion (FBM) leads to the scal-

ing δ2(∆, T ) ≃ ∆α of the time averaged mean squareddisplacement, turning over to a saturation plateau un-der confinement, δ2(∆, T ) ≃ ∆0 [curves (1’) and (2’) inFig. 1].

We particularly emphasise the irreproducible nature ofthe time averaged quantities and their associated scat-ter around the ensemble mean for CTRW subdiffusionprocesses. This randomness of the time averages is cap-tured by the distribution function of the amplitude ofthe time average. We show that even for relatively shorttrajectories this distribution is a good characteristic forthe underlying process. In contrast for FBM processesthe scatter typical of many single particle experiments isnot found in the long time limit.

In the remainder of this work, for simplicity we restrictthe discussion to the one-dimensional case (d = 1). Gen-eralisation to higher dimensions is straightforward. Thearticle is structured as follows. We start with a brief in-troduction to CTRW and FBM. We then consider thecases of unbounded motion and confined anomalous dif-fusion in the subsequent two Sections. The distributionof the time averages will be presented thereafter. Finally,we discuss the velocity autocorrelation functions for sub-diffusive CTRW and FBM processes, before presenting aconcluding discussion.

II. ANOMALOUS DIFFUSION PROCESSES

Although both CTRW and FBM give rise to an en-semble averaged mean squared displacement of the form(1), they are fundamentally different processes, as out-lined here. We note in passing that also in the randommotion on a fractal support subdiffusion arises [37, 38].We will not pursue this type of anomalous diffusion inthe following.

A. Continuous time random walk

CTRW theory dates back to the work of Montroll andWeiss [39], and was championed in the analysis of chargecarrier motion in amorphous semiconductors by Scherand Montroll [40]. CTRW has become a standard stat-istical tool to describe processes ranging from particlemotion in actin networks [22] to the tracer motion ingroundwater [41].

Each jump of a CTRW process is characterised by arandom jump length and a random waiting time elapsingbefore the subsequent jump. At each jump the jumplength and waiting time are chosen independently. Fora subdiffusive process we assume that the variance ofthe jump lengths is given by a finite value 〈δx2〉, andwe consider the unbiased case 〈δx〉 = 0. On a lattice ofspacing a, we would have 〈δx2〉 = a2. In contrast, the

Page 4: 1 2Physics Department, Technical University of Munich, D ... · arXiv:1009.4846v2 [cond-mat.stat-mech] 13 Oct 2010 22nd October 2018 Single particle tracking in systems showing anomalous

4

-1

0

1 x

(t)

-1

0

1

FBM

BM

x(t)

CTRW

0 10000 20000 30000 40000 50000 60000

-1

0

1

x(t)

t (a.u.)

Figure 3: Time series x(t) for Brownian motion (top), CTRW(middle), and FBM (bottom). The anomalous diffusion expo-nent for FBM and CTRW is α = 0.5. For CTRW the stallingevents are outstanding, while in the case of FBM strong anti-persistence occurs. Note that the stalling events and the anti-persistence are less pronounced for larger values of α. Typicalexperimental data include additional noise, such that the ap-pearance of measured data will not display the ideal behaviourshown here.

waiting times τ are drawn from the probability density

ψ(τ) ≃τα

|Γ(−α)|τ−1−α (5)

for large τ , with 0 < α < 1. This form of ψ is scale-free,that is, the average waiting time 〈τ〉 diverges, causingeffects such as ageing [42, 43] and weak ergodicity break-ing [44]. In Eq. (5), the quantity τ is a scaling factor.The anomalous diffusion constant in this case becomes[29, 45]

Kα =〈δx2〉

2τα. (6)

The scale-freeness of ψ(τ) allows individual waitingtimes τ to become quite large. No matter how long agiven time series is chosen, single τ values may becomeof the order of the length of the entire time span coveredin the trajectory. An example is shown in Fig. 3: thestalling events with large waiting times τ are quite dis-tinct. For values of the anomalous diffusion exponentα that are closer to 1 the stalling is less pronounced.Physically, the power-law form of the waiting time dis-tribution ψ(τ) may be related to comb models [37] orrandom energy landscapes [42, 46]; for more details, com-pare Refs. [47–50].

B. Fractional Brownian motion

FBM is a Gaussian process with stationary increments.Its position is defined in terms of the Langevin equation

dx(t)

dt= ξ(t), (7)

or, alternatively,

x(t) =

∫ t

0

ξ(t′)dt′. (8)

The motion is driven by stationary, fractional Gaussiannoise ξ(t) with zero mean 〈ξ(t)〉 and long-ranged noisecorrelation [51]

〈ξ(t1)ξ(t2)〉 = αK∗α(α− 1)|t1 − t2|

α−2

+2αK∗α|t1 − t2|

α−1δ(t1 − t2), (9)

contrasting the uncorrelated noise for normal diffusionα = 1: 〈ξ(t1)ξ(t2)〉 = 2K1δ(t1 − t2). In Eq. (9) theanomalous diffusion exponent is connected to the tra-ditionally used Hurst exponent by H = α/2, and weintroduce the abbreviation K∗

α = Kα/Γ(1 + α) for con-sistency with standard FBM notation. For subdiffusionthe fractional Gaussian noise is anticorrelated, decayinglike 〈ξ(t1)ξ(t2)〉 ∼ −K∗

αα|α− 1||t1− t2|α−2. This implies

that a given step is likely to go into the direction oppos-ite to the previous step. The corresponding oscillatingbehaviour is seen in Fig. 3. The position autocorrelationfunction of FBM becomes

〈x(t1)x(t2)〉 = K∗α

(

tα1 + tα2 − |t1 − t2|α)

, (10)

so that at equal times t1 = t2 we recover the meansquared displacement (4).If the fractional Gaussian noise is not considered ex-

ternal, but the validity of the fluctuation-dissipation the-orem is imposed, one obtains the generalised Langevinequation (GLE) [52],

md2x(t)

dt2= −γ

∫ t

0

(t− t′)β−2 dx

dt′dt′ + ηξ(t), (11)

where we write β instead of the exponent α in Eq. (9).In the GLE (11), we defined the coupling coefficient

η =√

kBT γ/[βK∗β(β − 1)] according to the fluctuation-

dissipation theorem, where kB is the Boltzmann constantand T the absolute temperature. In what follows we onlyconsider the overdamped limit, in which the inertia termmd2x(t)/dt2 can be neglected. The GLE then gives riseto the form

〈x2(t)〉 ≃ t2−β (12)

of the mean squared displacement. In contrast to FBM,that is, the GLE leads to subdiffusion for persistent noisewith 1 < β < 2, while 0 < β < 1 yields superdiffusion.

Page 5: 1 2Physics Department, Technical University of Munich, D ... · arXiv:1009.4846v2 [cond-mat.stat-mech] 13 Oct 2010 22nd October 2018 Single particle tracking in systems showing anomalous

5

FBM and the related GLE are used to describe pro-cesses such as long term storage capacity of water reser-voirs [53], climate fluctuations [54], economical marketdynamics [55], single file diffusion [56], and elastic mod-els [57]. Motion of this type has also been associatedwith the relative motion of aminoacids in proteins [58],and the free diffusion of biopolymers under molecularcrowding conditions [10, 19, 59].

III. FREE ANOMALOUS DIFFUSION

Let us begin with considering anomalous diffusion onan infinite domain and without drift. To find an ana-lytical expression for the time averaged mean squareddisplacement (2) we note that even a Brownian processrecorded over a finite time span T will show fluctuationsin the number of jumps performed during T . To averageout these trajectory-to-trajectory fluctuations we intro-duce the ensemble mean,

δ2(∆, T )⟩

=1

T −∆

∫ T−∆

0

(

x(t +∆)− x(t))2⟩

dt.

(13)We can then express the integrand in terms of the vari-ance of the jump lengths, 〈δx2〉, and the average numberof jumps n(t, t + ∆) in the time interval (t, t + ∆), asfollows:

(

x(t+∆)− x(t))2⟩

= 〈δx2〉n(t, t+∆). (14)

For a regular random walk on average every jump occursafter the waiting time 〈τ〉. Thus n(t, t + ∆) = ∆/〈τ〉,and

δ2(∆, T )⟩

= 2K1∆, (15)

where we defined the diffusion constant K1 =〈δx2〉/[2〈τ〉]. In the Brownian limit the time averagedmean squared displacement (15) in terms of the lag time∆ takes on exactly the same form as the ensemble aver-aged mean squared displacement (1) as function of time t.This is not surprising, as Brownian motion is an ergodicprocess. For sufficient duration T of the time records anytime average converges to the corresponding ensemble av-erage, and thus the ensemble average in expression (13)is no longer necessary.

A. Continuous time random walk

The number of jumps of a CTRW process with waitingtime distribution of the form (5) on average grows sublin-early with time, n(0, t) ∼ tα/[ταΓ(1+α)] [47]. This timeevolution translates into the mean squared displacement(1) with the anomalous diffusion coefficient (6). Combin-ing the time dependence of n(0, t) for CTRW subdiffusion

100 101 102 10310-5

10-4

10-3

10-2

10-1

100

101

()

Figure 4: Time averaged mean squared displacement for un-confined CTRW motion with α = 0.5, shown for 20 individualtrajectories. The overall measurement time is T = 105 (a.u.).Note the local changes of the slope in the trajectories as wellas the complete stalling in the lowest curve, all bearing wit-ness to the scale-free nature of the underlying waiting timedistribution ψ(τ ) ≃ τα/τ 1+α.

with the definition (13) we obtain the following result forthe time averaged mean squared displacement,

δ2(∆, T )⟩

∼ 2Kα∆

T 1−α(16)

in the limit ∆ ≪ T [34, 60]. This follows from expansionof the relation n(t, t + ∆) = n(0, t + ∆) − n(0, t). Thenoteworthy feature of this result is that the linear lagtime dependence of normal diffusion remains completelyunaffected by the anomalous nature of the stochastic pro-cess. Only the dependence on the overall measurementtime T witnesses the underlying subdiffusion. Eq. (16)can be understood if we notice that for normal diffusionwe have δ2(∆, T ) ≃ ∆/〈τ〉 [since according to EinsteinK1 ∝ 1/〈τ〉] where 〈τ〉 is the mean time between jumpevents. Now, for the subdiffusive case 〈τ〉 diverges and

must be replaced by∫ T

0ψ(τ)τdτ ∝ T 1−α, which explains

the term δ2 ∝ ∆/T 1−α. Fig. 4 shows 20 trajectories ofa CTRW process with α = 0.5, displaying the generaltrend δ2(∆, T ) ≃ ∆.Rewriting the result (16) in the form

δ2(∆, T )⟩

∼ 2K(T )∆ (17)

we see that the effective diffusion coefficient K(T ) de-cays as function of the measurement time T . The longerthe system evolves after its initial preparation, the lessmobile it appears, consistent with the ageing property ofCTRW subdiffusion. For instance, in the picture of thetrapping events, in the course of time deeper and deepertraps may be encountered by the diffusing particle, such

Page 6: 1 2Physics Department, Technical University of Munich, D ... · arXiv:1009.4846v2 [cond-mat.stat-mech] 13 Oct 2010 22nd October 2018 Single particle tracking in systems showing anomalous

6

that it gets more and more stuck. This increasing immob-ility is mirrored in the form K(T ). Note that Eqs. (16)and (17) suggest that from measuring the ∆ dependenceof an anomalous diffusion process one might draw the er-roneous conclusion that normal diffusion were observed.

The disparity between the time averaged mean squareddisplacement (16) and its ensemble averaged counterpart(4) demonstrates the weak ergodicity breaking charac-teristic of a process with diverging time scale [42, 44].Single or few, long waiting times are also responsiblefor the pronounced deviations between individual realisa-tions shown in Fig. 4. This apparent irreproducibility ofthe time averaged mean squared displacement is again in-timately coupled to the weak ergodicity breaking natureof the CTRW subdiffusion process. We will discuss thisfeature more quantitatively in terms of the distribution

φα(ξ) as function of the relative deviation ξ = δ2/⟨

δ2⟩

of the time averaged mean squared displacement aroundits ensemble mean in Section V.

B. Fractional Brownian motion

In contrast to the above behaviour of CTRW subdif-fusion processes, FBM is ergodic. Indeed, by help ofthe position autocorrelation (10) the time averaged meansquared displacement (13) becomes [61]

δ2(∆, T )⟩

= 2K∗α∆

α, (18)

exactly matching the ensemble averaged mean squareddisplacement, 〈x2(t)〉 = 2K∗

αtα. However, ergodicity is

reached algebraically slowly [61], see also below. In Fig. 5the comparatively minute scatter between individual tra-jectories supports the ergodic behaviour of FBM pro-cesses. The somewhat larger deviations at longer lagtimes are due to worsening statistics when ∆ → T .

IV. CONFINED ANOMALOUS DIFFUSION

Since the motion of tracer particles is typically con-fined, for example, by the cell walls or internal mem-branes, we now consider the important case of anomalousdiffusion in a bounded domain.

In a finite interval [−L,L] the mean squared displace-ment of a Brownian particle initially released well awayfrom the boundaries will grow linearly in time and even-tually turn over to a stationary plateau of magnitude〈x2〉st = L2/3. Similarly, if the particle evolves in theconfinement of an external harmonic potential of the formV (x) = 1

2mω2x2, the thermal value 〈x2〉th = kBT /[mω

2]will eventually be attained.

100 101 102 10310-2

10-1

100

() ~

Figure 5: Time averaged mean squared displacement for un-confined FBM with α = 0.5. The scatter between individualtrajectories is minute, mirroring the ergodicity of this process.

A. Continuous time random walk

How is this behaviour modified for a CTRW subdif-fusion process? Under the influence of an arbitrary ex-ternal potential V (x) = −

∫ xF (x′)dx′ defining the force

F (x), CTRW subdiffusion can be described by the frac-tional Fokker-Planck equation [29, 45], or, equivalently,in terms of the following coupled Langevin equations [62]:

dx(s)

ds=

K

kBTF (x) + η(s) (19a)

dt(s)

ds= ω(s). (19b)

Here the position x is expressed in terms of the parameters (the internal time), and driven by the white Gaussiannoise η(s). Thus, Eq. (19a) defines standard Brownianmotion x(s), where K is the diffusivity for the normaldiffusion process in internal time s. Laboratory time tis introduced by the so-called subordination through theprocess ω(s), given by the probability density function[29]

pt(s) =1

α

(

K

)1/αt

s1+1/αlα

(

[

K

]1/αt

s1+1/α

)

,

(20)where lα(z) is a one-sided Levy stable probability dens-ity with Laplace transform

∫∞

0lα(z) exp(−uz)dz =

exp(−uα). Thus, Eq. (19b) transforms the Brownianprocess x(s) with diffusivity K into the subdiffusive mo-tion with generalised diffusivity Kα. On the basis of thisscheme, the ensemble averaged position-position correla-tion in an arbitrary confining potential V (x) becomes

〈x(t1)x(t2)〉 =(

〈x2〉B − 〈x〉2B

)B(t1/t2, α, 1− α)

Γ(α)Γ(1 − α)+ 〈x〉2B ,

(21)

Page 7: 1 2Physics Department, Technical University of Munich, D ... · arXiv:1009.4846v2 [cond-mat.stat-mech] 13 Oct 2010 22nd October 2018 Single particle tracking in systems showing anomalous

7

valid in the limit t2 ≥ t1 ≫ (1/[Kαλ1])1/α, where λ1

is the smallest non-zero eigenvalue of the correspondingFokker-Planck operator [35]. Eq. (21) demonstrates that,despite the confinement, the process is non-stationary. Inresult (21) we used the incomplete Beta function

B(t1/t2, α, 1− α) ≡

∫ t1/t2

0

zα−1(1 − z)−αdz (22)

and defined the first and second Boltzmann moments,whose general definition is

〈xn〉B =1

Z

∫ ∞

−∞

xn exp

(

−V (x)

kBT

)

dx. (23)

The partition function reads

Z =

∫ ∞

−∞

exp

(

−V (x)

kBT

)

dx. (24)

At large time separation, t2 ≫ t1, the position autocor-relation decays algebraically,

〈x(t1)x(t2)〉 ∼(

〈x2〉B − 〈x〉2B

) (t1/t2)α

αΓ(1 + α)Γ(1 − α)+ 〈x〉2B , (25)

towards the value 〈x〉2B . The corresponding limiting be-haviour of the incomplete Beta function reads

B(t/(t+∆), α, 1 − α)

Γ(α)Γ(1 − α)∼ 1−

sin(πα)

(1 − α)π

(

t

)1−α

. (26)

Inserting expression (21) into the definition of the timeaveraged mean squared displacement (13) we obtain [35]

δ2(∆, T )⟩

=1

T −∆

∫ T−∆

0

[

〈x(t+∆)x(t +∆)〉

+〈x(t)x(t)〉 − 2〈x(t+∆)x(t)〉]

dt. (27)

Then, with relation (26) we arrive at the scaling beha-viour

δ2(∆, T )⟩

∼(

〈x2〉B − 〈x〉2B

) 2 sin(πα)

(1− α)πα

(

T

)1−α

,

(28)valid in the limits ∆/T ≪ 1 and ∆ ≫ (1/[Kαλ1])

1/α.Result (28) is quite remarkable: instead of the naively as-sumed saturation toward the stationary plateau a power-

law growth⟨

δ2(∆, T )⟩

∼ (∆/T )1−α is observed. Only

when the lag time ∆ approaches the overall measurementtime T the singularity in expression (13) causes a dip to-ward the plateau of the ensemble averaged mean squareddisplacement. Additionally Eqs. (21) and (28) are uni-versal in the sense that the exact form of the confiningpotential solely enters into the prefactor through the firsttwo moments of the Boltzmann distribution correspond-ing to the confining potential V (x). We note that the

10−4

10−2

100

102

104

106

10−8

10−6

10−4

10−2

100

〈δ2(∆

)〉/

〈x2〉 B

Figure 6: Simulated behaviour of⟨

δ2(∆)⟩

for an harmonic

binding potential V (x) = 2x2 (�) and a particle in a boxof length 2 (©), with α = 1/2, measurement time T = 107

(a.u.), kBT = 0.1, and K0.5 = 0.0892. Without fitting, thelines show the analytic results for the transition from ≃ ∆1

for short lag times according to Eq. (16) (−− and − · −)to ≃ ∆1−α for long lag times, Eq. (28) (—). For long ∆,⟨

δ2(∆)⟩

/〈x2〉B exhibits universal behaviour, in the sense

that the curve does not depend on the external field.

asymptotic scaling ≃ (∆/T )1−α is consistent with thenumerical analysis presented in Ref. [63].

Fig. 6 depicts the ensemble mean of the time averagedmean squared displacement (13) for two types of confine-ment, an harmonic potential and a box potential. Theparticle is initially placed at the bottom of the poten-tial well and in the middle of the box potential, respect-ively. This is why at short lag times the particle exhibits

the linear scaling⟨

δ2(∆, T )⟩

≃ ∆ typical for unconfined

motion, before turning over to the confinement-inducedscaling ≃ ∆1−α. Note that this plot is fit-free, i.e., thetheoretically calculated asymptotic behaviours nicely fallon top of the simulations results.

As shown in Fig. 7 individual trajectories still ex-hibit the pronounced scatter typical for CTRW subdiffu-sion. Visually the scatter does not change between theunbiased initial motion and the confinement-dominatedpart of the process. This is due to the fact that the scat-ter is caused by the scale-freeness of the waiting times.In our process waiting times and jump lengths are de-coupled, corresponding to the subordination property ofCTRW subdiffusion [29].

Page 8: 1 2Physics Department, Technical University of Munich, D ... · arXiv:1009.4846v2 [cond-mat.stat-mech] 13 Oct 2010 22nd October 2018 Single particle tracking in systems showing anomalous

8

10−4

10−2

100

102

104

106

10−8

10−6

10−4

10−2

100

δ2(∆

)/

〈x2〉 B

Figure 7: Scatter between individual trajectories in theharmonic potential V (x) = 2x2. The extent of the de-viations between the trajectories does not change qualit-atively between the initial unbiased motion and the laterconfinement-dominated regime. The squares (�) repres-

ent simulations results for the ensemble average⟨

δ2(∆, T )⟩

.

Same parameters as in Fig. 6.

B. Continuous time random walk with waiting

time cutoff

What happens if we introduce a cutoff in the power-lawof the waiting time distribution of the form

ψ(τ) =d

[

1−τα

(τ + τ)αexp

(

− τ/τ∗)

]

, (29)

in which a characteristic time scale τ∗ is introduced, ter-minating the power-law scaling? For the waiting timedensity (29) the characteristic waiting time

∫∞

0τψ(τ)dτ

becomes finite. At sufficiently short times one would ex-pect this process to still exhibit the features of CTRWsubdiffusion, while at times τ ≫ τ∗ the process shouldconverge to regular Brownian motion with a Gaussianpropagator. In Fig. 8 we demonstrate that for a suitablechoice of the cutoff time τ∗ the behaviour of the time av-eraged mean squared displacement δ2(∆, T ) under con-finement preserves the characteristic non-ergodic featuresof CTRW subdiffusion, i.e., the turnover from the initialscaling ≃ ∆ to the confinement-dominated scaling ∆1−α.Below we will show that at the same time the distribu-tion of the time average around its ensemble mean issignificantly altered.

C. Fractional Brownian motion

Ergodicity remains unaffected for FBM, that is con-fined to an interval of size [−L,L]. Namely, conver-

gence of δ2(∆, T ) is observed toward the stationary value

100 101 102 10310-1

100

101

102

2 ()

Figure 8: CTRW subdiffusion for a waiting time with ex-ponential cutoff, Eq. (29), in the box [−3, 3] with reflectingboundary conditions. The generic behaviour of initial andfinal scaling, δ2(∆, T ) ≃ ∆ and ≃ ∆1−α remains unaltered.We chose α = 0.8, τ = 1, τ∗ = 20 and overall measurementtime T = 1000 (a.u.).

〈x2〉st = L2/3 [64]. For FBM under the influence ofan harmonic potential V (x) = 1

2mω2x2 the position-

position correlation can be calculated exactly [65]. In-serting into the ensemble mean (13) of the time averagedmean squared displacement one can show that the ini-tial behaviour 〈x2(t)〉 ≃ tα turns over to the stationaryvalue 〈x2〉st = Γ(α + 1)kBT /[mω

2]. In general, the timeaveraged mean squared displacement for confined FBMconverges to a constant:

δ2(∆, T )⟩

∼ const. (30)

V. FLUCTUATIONS OF THE TIME AVERAGE

AND ERGODICITY BREAKING PARAMETER

The deviations of the time averaged mean squared dis-placement δ2(∆, T ) between individual trajectories canbe quantified in terms of the probability density functionφα(ξ) of the dimensionless ratio

ξ ≡δ2(∆, T )⟨

δ2(∆, T )⟩ (31)

of the time averaged mean squared displacement over itsensemble mean. For an ergodic process this distributionis necessarily sharp,

φerg(ξ) = δ(ξ − 1), (32)

for long measurement time T . Deviations from this formare expected for non-ergodic processes such as CTRWsubdiffusion, but also for relatively short trajectories. We

Page 9: 1 2Physics Department, Technical University of Munich, D ... · arXiv:1009.4846v2 [cond-mat.stat-mech] 13 Oct 2010 22nd October 2018 Single particle tracking in systems showing anomalous

9

here address both effects and demonstrate that the distri-bution φα(ξ) is a quite reliable means to distinguish dif-ferent stochastic processes even when the recorded timeseries are fairly short.

A. Continuous time random walk

For CTRW subdiffusion with power-law waiting timedistribution (5) the distribution assumes the form [34, 66]

φα(ξ) =Γ(1 + α)1/α

αξ1+1/αlα

(

Γ(1 + α)1/α

ξ1/α

)

(33)

for T → ∞, where lα(z) is again a one-sided Levy stabledistribution, whose Laplace transform is L {lα(z)} =exp (−uα). Special cases include the Brownian limit (32)for α = 1 and the Gaussian shape

φ1/2(ξ) =2

πexp

(

−ξ2

π

)

(34)

for α = 1/2. In fact, we observe an exponentially fastdecay of φα(ξ) with large ξ for all 0 < α < 1, compareAppendix A.For short trajectories the basic shape of the distribu-

tion φα(ξ) is surprisingly well preserved [67]. This isdemonstrated in Fig. 9, in which the characteristic asym-metric shape with respect to the ergodic value ξ = 1 forthe case α = 0.5 is reproduced for a process with T = 128,even for a lag time as long as ∆ = 100. No significant de-pendence on the confinement is observed. Also for largervalues of α the finite value for small ξ is similarly re-produced for short trajectories [67], pointing at the quiteremarkably reliability on the shape of the scatter distri-bution φα(ξ) for CTRW subdiffusion. As we show below,the distribution for FBM processes with its zero value atξ = 0 can be clearly distinguished from the CTRW form.A useful parameter to quantify the violation of ergodi-

city is the ergodicity breaking parameter [34]

EB = limT→∞

(

δ2)2⟩

−⟨

δ2⟩2

δ2⟩2 =

2Γ(1 + α)2

Γ(1 + 2α)− 1 (35)

EB varies from EB = 1 for α → 0 monotonically toEB = 0 in the Brownian limit α = 1. For the specialcase α = 1/2 one finds EB = π/2 − 1 ≈ 0.57, while forα = 0.75, EB ≈ 0.27.

B. Continuous time random walk with waiting

time cutoff

The scatter distribution is sensitive to few extremeevents. When these are lacking, the distribution shouldbe significantly different from the form discussed forCTRW subdiffusion with diverging characteristic time

0 1 2 3 4 5-0.5

0.0

0.5

1.0

1.5

2.0

0 1 2 3 4 5-0.5

0.0

0.5

1.0

1.5

2.0 =100 =100

()

=1 =10 =1 =10 theory

()

Figure 9: Distribution φα(ξ) of the time averaged mean

squared displacement, with ξ = δ2/⟨

δ2⟩

for a CTRW pro-

cess with α = 0.5, τ = 1, and T = 128 (a.u.). The filled andopen symbols, respectively, represent the unconfined case andconfined motion on the interval [−2, 2]. We see quite goodagreement with the expected Gaussian limiting distributionof the scatter, Eq. (34), centred around ξ = 0. In the insetwe show the case for the largest measured lag time, ∆ = 100,also in good agreement with the predicted shape of the dis-tribution.

scale. Indeed, as shown in Fig. 10 for the cutoff waitingtime distribution defined through Eq. (29), φα(ξ) dropsdown to zero around ξ = 0 and assumes an almost Gaus-sian shape around the ergodic value ξ = 1. Note that thesimulations were performed with the same parameters asfor Fig. 8, in which the breaking of ergodicity still per-sists, despite the cutoff. This demonstrates that differentquantities have different sensitivity to the cutoff.

C. Fractional Brownian motion

FBM is based on long-ranged correlations. However,we can obtain an approximate expression for the scatterdistribution [67]

φ(ξ) ≈

T −∆

4πτ†exp

(

−(ξ − 1)2(T −∆)

4τ†

)

. (36)

where τ† is an intrinsic time scale. Expression (36) isvalid for sufficiently small lag times ∆, for which thecorrelations are neglected [67]. Note that this distribu-tion is independent of α and centred around the ergodicvalue ξ = 1. In particular, in the long measurement timelimit T → ∞, the distribution converges to the sharpform φ(ξ) = δ(ξ − 1). Also this behaviour is surprisinglywell preserved for short trajectories, as demonstrated inFig. 11.

Page 10: 1 2Physics Department, Technical University of Munich, D ... · arXiv:1009.4846v2 [cond-mat.stat-mech] 13 Oct 2010 22nd October 2018 Single particle tracking in systems showing anomalous

10

0.0 0.5 1.0 1.5 2.0

0.0

0.5

1.0

1.5

2.0

2.5

3.0(

)

Figure 10: Scatter distribution φα(ξ) for CTRW subdiffusionwith cutoff, as defined in Eq. (29). The observed form isclearly different from the much more asymmetric shape inFig. 9. The value of φα(ξ) reaches zero for small values of ξ.An almost Gaussian shape centred around the ergodic valueξ = 1 is observed. Same parameters as in Fig. 8.

0.0 0.5 1.0 1.5 2.0

0

1

2

3

4

5

0.0 0.5 1.0 1.5 2.0012345

theory

Figure 11: Scatter distribution φ(ξ) of the time averaged

mean squared displacement, with ξ = δ2/⟨

δ2⟩

for FBM sub-

diffusion with α = 0.5 and T = 128 (a.u.). The filled and opensymbols, respectively, represent the cases for unconfined andconfined motion. For small values of the lag time ∆ we seequite good agreement with the expected Gaussian limitingdistribution of the scatter, Eq. (36), that is centred aroundξ = 1. For larger values of ∆ deviations are observed, how-ever, even for ∆ = 100 the value around ξ = 0 is consistentlyzero. A similar behaviour is found for larger values of α [67].Note that when T → ∞ the process is ergodic and φ(ξ) ap-proaches a δ function.

For FBM one can define a quantity similar to the er-godicity breaking parameter (35). Namely, without tak-ing the long measurement time limit, we obtain the nor-malised variance of the time averaged mean squared dis-placement

V =

(

δ2(∆, T ))2⟩

−⟨

δ2(∆, T )⟩2

δ2(∆, T )⟩2 . (37)

It turns out that the convergence to ergodicity is algeb-raically slow, and for subdiffusion 0 < α < 1 one obtainsa decay, which is inversely proportional to T [61]:

V ∼ k(α)∆

T. (38)

The coefficient k(α) here is defined as [61]

k(α) =

∫ ∞

0

(

(t+ 1)α + |t− 1|α − 2tα)

dt. (39)

It increases continuously from zero at α = 0 to k(1) =4/3.The quantity V is connected to the probability density

(36) by V = 2τ†/(T − ∆) ∼ 2τ†/T , obtained from ap-proximating that the values of ξ may range in the interval(−∞,∞), which appears reasonable given the sharp de-cay of φ(ξ) to 0 at ξ = 0 for FBM. Apart from the factthat V explicitly depends on α while the approximationleading to Eq. (36) loses the α dependence, both quant-ities decay ∼ 1/T , and the internal time scale τ† takeson a role similar to the lag time ∆.

VI. VELOCITY AUTOCORRELATION

FUNCTIONS

A typical quantity accessible from experimental datais the velocity autocorrelation function, which is definedthrough

C(ǫ)v (τ) =

1

ǫ2

(x(τ + ǫ)− x(τ)) (x(ǫ)− x(0))⟩

. (40)

Here the velocity is defined as the difference quotientv(τ) = ǫ−1[x(τ + ǫ)−x(τ)]. The velocity autocorrelation(40) was suggested as a tool to distinguish between dif-ferent subdiffusion models [10]. We show here that forconfined processes the shape of the velocity autocorrela-tion function does not allow for a significant distinctionbetween subdiffusive CTRW and FBM. Note that the cal-culation of C

(ǫ)v (τ) amounts to determine four two-point

correlation functions of the type 〈x(t1)x(t2)〉.

A. Continuous time random walk

For unbounded CTRW process starting with initialcondition x(0) = 0, the position correlation function is

Page 11: 1 2Physics Department, Technical University of Munich, D ... · arXiv:1009.4846v2 [cond-mat.stat-mech] 13 Oct 2010 22nd October 2018 Single particle tracking in systems showing anomalous

11

0 20 40 60 80 100

τ-0.2

0

0.2

0.4

0.6

0.8

1

Cv

(ε) (τ

)/Cv

(ε) (0

)Free CTRW SimulationFree CTRW Theory

Figure 12: Normalised velocity autocorrelation function

C(ǫ)v (τ ) for an unconfined CTRW process with α = 0.5. The

simulations (©) were performed for time 10000 (a.u) withǫ = 10 (a.u). The theoretical behaviour (—) is given byEq. (42).

given by the following expression [68],

〈x(t1)x(t2)〉 =2Kα

Γ(1 + α)

[

min{t1, t2}]α

. (41)

This result for free CTRW is due to the fact that thejump lengths in the interval (t1, t2) for t2 > t1 have zeromean. With Eq. (41) we find

C(ǫ)v (τ)

C(ǫ)v (0)

=

ǫα − τα

ǫατ ≤ ǫ

0 τ ≥ ǫ

(42)

The velocity autocorrelation function for an unboundedCTRW process does not yield negative values due to theabsence of correlations for different jumps. In this casethe velocity autocorrelations can be easily distinguishedfrom the behaviour of FBM processes, see below. Note

that C(ǫ)v (τ) in Eq. (42) is non-analytic at τ = ǫ, an ob-

servation that is confirmed in simulations. The behaviourof the velocity autocorrelations for CTRW subdiffusionare displayed in Fig. 12.For a confined CTRW process the situation is quite dif-

ferent. To explore the behaviour of the corresponding ve-locity autocorrelation function we use the general resultfor the correlation function of confined CTRW, Eq. (21).For simplicity we assume a symmetric potential such that〈x〉B = 0. With the initial condition x(0) = 0 we obtain

C(ǫ)v (τ)

C(ǫ)v (0)

=1

Γ(α)Γ(1 − α)

×

B(

ǫǫ+τ , α, 1− α

)

−B(

τǫ , α, 1− α

)

ǫ ≥ τ

B(

ǫǫ+τ , α, 1− α

)

−B(

ǫτ , α, 1− α

)

ǫ ≤ τ. (43)

Figs. 13 and 14 (for the absolute value) display excel-lent agreement of Eq. (43) with simulations of CTRW

0 20 40 60 80 100

τ

-0.4

-0.2

0

0.2

0.4

0.6

0.8

1

Cv

(ε) (τ

)/Cv

(ε) (0

)

Confined CTRW Theory

Confined CTRW SimulationFree FBM Theory

Figure 13: Normalised velocity autocorrelation function

C(ǫ)v (τ ) for a free FBM process and a confined CTRW process

with α = 0.5. The CTRW simulations (©) were performedover the time range 10000 (a.u), and the system correspondsto a particle on a lattice with ten lattice points and reflectingboundaries. We chose ǫ = 10 (a.u). The theoretical predic-tion for the CTRW (—) is given by Eq. (43), and for FBM (--) by Eq. (45).

subdiffusion for a lattice of size 10 with reflecting bound-ary conditions, and α = 1/2. We observe that for con-fined motion the CTRW velocity autocorrelation func-tion indeed attains negative values and has a minimumon τ = ǫ, as the confinement effectively induces correl-ations. For long τ the velocity autocorrelation functiondecays to zero (from the negative side) as the power-law

C(ǫ)v (τ)

C(ǫ)v (0)

∼ −1

Γ(α)Γ(1− α)

( ǫ

τ

)1+α

, (44)

valid for τ ≫ ǫ.

B. Fractional Brownian Motion

For free FBM the position correlation function (10)together with the definition (40) produce the result

C(ǫ)v (τ)

C(ǫ)v (0)

=(τ + ǫ)α − 2τα + |τ − ǫ|α

2ǫα. (45)

This function yields negative values for sufficiently longτ , and its minimum value (2α−1−1) is assumed at τ = ǫ.For long τ it decays toward zero from the negative sidein the power-law form

C(ǫ)v (τ)

C(ǫ)v (0)

∼ −α− α2

2

( ǫ

τ

)2−α

(46)

valid for τ ≫ ǫ.We see that both the velocity autocorrelation function

of the confined CTRW and the one for free FBM acquire

Page 12: 1 2Physics Department, Technical University of Munich, D ... · arXiv:1009.4846v2 [cond-mat.stat-mech] 13 Oct 2010 22nd October 2018 Single particle tracking in systems showing anomalous

12

0.01 0.1 1 10 100 1000

τ

0.001

0.01

0.1

1|C

v

(ε) (τ

)/Cv

(ε) (0

)|

Confined CTRW SimulationConfined CTRW Theory

Figure 14: Absolute value of the normalised velocity auto-

correlation function C(ǫ)v (τ ) for confined CTRW process with

α = 0.5. The CTRW simulations (©) are based on the para-meters from Fig. 13. The theory line (—) corresponds toEq. (43).

0 20 40 60 80 100

-0.50

-0.25

0.00

0.25

0.50

0.75

1.00 Confined FBM Free FBM Theory

Cv()/C

v(0)

Figure 15: Normalised velocity autocorrelation function

C(ǫ)v (τ ) for confined FBM with α = 0.5. The simulations

were performed on the interval [−2, 2], and ǫ = 10 (a.u.).For comparison, the velocity autocorrelation for free FBM isdrawn based on Eq. (45).

negative values and decay as power-laws for large τ , seebelow. Further both obtain a sharp minimum for τ =ǫ. Confined FBM behaves similarly to free FBM, seeFig. 15. From our analysis it becomes clear that theshape of the velocity autocorrelation function may not bea good diagnosis tool to distinguish between subdiffusiveCTRW and FBM processes. We note that for CTRWsubdiffusion the time averaged correlation functions willbe random variables, exactly like the time averaged meansquared displacement.

VII. DISCUSSION

The physical mechanisms leading to subdiffusion inbiological cells or other complex systems are varied.The anomalous diffusion of inert biopolymers larger thansome 10kD in living cells is due to molecular crowding,these are excluded volume effects in the viscoelastic, su-perdense cellular environment [30–32]. Potentially sim-ilar effects occur in single file diffusion: A tracer particlediffusing in a one dimensional channel and interactingwith other Brownian particles will slow down its ran-dom motion, and subdiffusion with α = 1/2 is observed[56]. What is the resulting motion like? Consider aparticle, that gets repeatedly trapped during its randomwalk. Such trapping may give rise to a broad distribu-tion of waiting times, as embodied in the CTRW model.Such broad distribution of trapping times could be dueto chemical binding of the tracer particle to its environ-ment on varied time scales, or due to active gelation/de-gelation of the crowding particles. Long-tailed trappingtime distributions were observed for µm-sized particlesdue to interaction with a semiflexible actin mesh [22].There, the anomalous diffusion exponent α depends onthe size of the tracer particles versus the typical actinmesh size. Alternatively, motion patterns of FBM typemay be due to coupling to the viscoelastic crowding envir-onment [19]. Yet other approaches are based on polymerdynamics. De Gennes’ reptation model of a polymer dif-fusing in a tube formed by foreign chains in a melt yieldssubdiffusion with α = 1/4, and was used as a possible ex-planation of the short time dynamics of telomere motion[11].

One may speculate whether the widely observed sub-diffusion of biomolecules and passive tracers is a merecoincidence of nature that should be attributed to thedense environment in the cell and/or to the wide distri-bution of obstacles and reactions in the cell. Or, maybe,subdiffusion is by itself a goal which was obtained viaevolution? There exist claims that sub-diffusion is use-ful in certain cellular search strategies [9, 69]. One mayalso view subdiffusion to be a sufficiently slow process tohelp maintaining the organisation of the genome in thenucleus of the cell without the need for physical com-partments [11]. Since reactions in many cases are con-trolled by diffusion, the emergence of slower-than-normaldiffusion has far-reaching implications to signalling andregulatory processes in the cell. The standard theor-ies of diffusion-controlled reactions under such circum-stances must be replaced by subdiffusion-controlled re-action models. While these issues are clearly important,we here focused more on the characterisation of traject-ories of single particles in the cell, in particular, as aquantitative diagnosis method to probe the nature of thestochastic motion in cells and other complex media.

Single particle tracking microscopy is a widely usedtechnique. It allows one to locally probe complex systemsin the liquid phase in situ. In particular it has becomeone of the standard tools in biophysical, colloidal, poly-

Page 13: 1 2Physics Department, Technical University of Munich, D ... · arXiv:1009.4846v2 [cond-mat.stat-mech] 13 Oct 2010 22nd October 2018 Single particle tracking in systems showing anomalous

13

meric, and gel-like environments. The motion of tracersin these systems is often anomalous. Typically the exper-imentally recorded time series are evaluated in terms ofthe time averaged mean squared displacement δ2(∆, T ).

Here we collected the behaviour of δ2(∆, T ) for the twomost prominent anomalous stochastic processes, the con-tinuous time random walk and fractional Brownian mo-tion.For CTRW subdiffusion, connected with ageing and

weak ergodicity breaking, in an unconfined system thetime averaged mean squared displacement scales linearlyin time, δ2(∆, T ) [Eq. (16)], renouncing the anomalousnature of the process. Only the dependence on the over-all measurement time T pays tribute to the underlyingsubdiffusion. This behaviour contrasts the anomalousscaling 〈x2(t)〉 = 2Kαt

α of the ensemble averaged meansquared displacement. Under confinement no plateauvalue is observed as in the ensemble average, instead,a power-law of the form δ2(∆, T ) ≃ ∆1−α, Eq. (28), isfound. Interestingly, this characteristic behaviour is ap-proximately preserved when an appropriate cutoff in thewaiting time is introduced. Conversely the distributionof the time averaged mean squared displacement aroundits ensemble mean becomes almost Gaussian in presenceof the cutoff, while it has an exponential decay and a fi-nite value at ξ = 0 when the system is non-ergodic andexhibits ageing.FBM, in contrast, is ergodic, although ergodicity is

reached algebraically slowly. Free FBM subdiffusionshows δ2(∆, T ) ≃ ∆α, Eq. (18), while under confinementthe plateau value of the ensemble average is reached,Eq. (30). For finite trajectories the distribution of thetime averaged mean squared displacement is approxim-ately Gaussian for short lag times.Our analysis demonstrates how different the two

stochastic processes are, despite sharing the same formof the ensemble averaged mean squared displacement.At the same time the velocity autocorrelation of con-fined CTRW subdiffusion is hardly distinguishable fromthat of FBM subdiffusion. Given a recorded time series

of anomalous diffusion from experiment it is importantto know more precisely which stochastic process is re-sponsible for the observed behaviour, in particular, withrespect to diffusion-limited reactions, general transportbehaviour, and related processes such as gene regulation.The analysis presented here, along with complementarytools discussed in Refs. [23, 59, 70], will be instrumentalin the classification of anomalous diffusion behaviour.

Acknowledgments

We thank C. Selhuber-Unkel and L. Oddershede forproviding the data used in Fig. 2, and for many usefuldiscussions. Partial financial support from the DeutscheForschungsgemeinschaft and the Israel Science Founda-tion is gratefully acknowledged.

Appendix A: Scatter distribution

Using the theory of Fox H-functions we obtain the ex-act form for the distribution φα(ξ) of the dimensionlessvariable ξ [71]:

φα(ξ) =1

α2ξH1,0

1,1

[

ξ1/α

Γ(1 + α)1/α

(0, 1)(0, 1/α)

]

. (A1)

This function has the series expansion

φα(ξ) =1

αξ

∞∑

n=0

(−1)n[Γ(1 + α)/ξ]n

n!Γ(−αn), (A2)

and the asymptotic form

φα(ξ) ≃(

ξ1/α)(1−α)/(2α)−α

× exp

(

−(1− α)

[

ααξ

Γ(1 + α)

]1/(1−α))

(A3)

valid at ξ ≫ Γ(1 + α).

[1] C. Brauchle, D. C. Lamb, and J. Michaelis, Editors,Single Particle Tracking and Single Molecule EnergyTransfer (Wiley-VCH, Weinheim, 2010).

[2] M. J. Saxton and K. Jacobson, Ann. Rev. Biophys. Bio-mol. Struct. 26, 373 (1997).

[3] M. J. Saxton, Biophys. J. 72, 1744 (1997).[4] H. Qian, M. P. Sheetz, and E. L. Elson, Biophys. J. 60,

910 (1991).[5] J. Perrin, Comptes Rendus (Paris) 146, 967 (1908); Ann.

Chim. Phys. 18, 5 (1909).[6] I. Nordlund, Z. Physik. Chemie 87, 40 (1914).[7] T. G. Mason and D. A. Weitz, Phys. Rev. Lett. 74, 1250

(1995).[8] W. J. Greenleaf, M. T. Woodside, and S. M. Block, Ann.

Rev. Biophys. Biomol. Struct. 36, 171 (2007).[9] I. Golding and E. C. Cox, Phys. Rev. Lett. 99, 098102

(2006).[10] S. C. Weber, A. J. Spakowitz, and J. A. Theriot, Phys.

Rev. Lett. 104, 238102 (2010).[11] I. Bronstein, Y. Israel, E. Kepten, S. Mai, Y. Shav-Tal,

E. Barkai, and Y. Garini, Phys. Rev. Lett. 103,018102(2009).

[12] G. Seisenberger, M. U. Ried, T. Endreß, H. Buning, M.Hallek, C. and Brauchle, Science 294, 1929 (2001).

[13] A. Caspi, R. Granek, and M. Elbaum, Phys. Rev. Lett.85, 5655 (2000).

[14] I. M. Tolic-Nørrelykke, E. L. Munteanu, G. Thon, L.Oddershede, and K. Berg-Sørensen, Phys. Rev. Lett. 93,078102 (2004).

[15] C. Selhuber-Unkel, P. Yde, K. Berg-Sørensen, and L. B.Oddershede, Phys. Biol. 6, 025015 (2009).

[16] N. Gal and D. Weihs, Phys. Rev. E 81, 020903(R) (2010).

Page 14: 1 2Physics Department, Technical University of Munich, D ... · arXiv:1009.4846v2 [cond-mat.stat-mech] 13 Oct 2010 22nd October 2018 Single particle tracking in systems showing anomalous

14

[17] D. Banks and C. Fradin, Biophys. J. 89, 2960 (2005).[18] M. Weiss, M. Elsner, F. Kartberg, and T. Nilsson, Bio-

phys. J. 87, 3518 (2004).[19] J. Szymanski and M. Weiss, Phys. Rev. Lett. 103, 038102

(2009).[20] J. Vercammen, G. Martens, and Y. Engelborghs,

Springer Ser. Fluoeresc. 4, 323 (2007).[21] W. Pan, L. Filobelo, N. D. Q. Pham, O. Galkin, V.

V. Uzunova, and P. G. Vekilov, Phys. Rev. Lett. 102,058101 (2009).

[22] I. Y. Wong, M. L. Gardel, D. R. Reichman, E. R. Weeks,M. T. Valentine, A. R. Bausch, and D. A. Weitz, Phys.Rev. Lett. 92, 178101 (2004).

[23] J.-H. Jeon, V. Tejedor, S. Burov, E. Barkai, C. Selhuber,K. Berg-Sørensen, L. Oddershede, and R. Metzler (un-published).

[24] N. G. van Kampen, Stochastic processes in physics andchemistry (Elsevier, Amsterdam, 2007).

[25] A. Y. Khinchin, Mathematical foundations of statist-ical mechanics (Dover Publications Inc., New York, NY,2003).

[26] Y. Sinai, Theor. Prob. Appl. 27, 256 (1982); J. Dragerand J. Klafter, Phys. Rev. Lett. 84, 5998 (2000).

[27] V. Tejedor and R. Metzler, J. Phys. A 43, 082002 (2010).[28] P. Siegle, I. Goychuk, and P. Hanggi, Phys. Rev. Lett.

105, 100602 (2010).[29] R. Metzler and J. Klafter, Phys. Rep. 339, 1 (2000); J.

Phys. A 37, R161 (2004); E. Barkai, Phys. Rev. E 63,046118 (2001).

[30] S. B. Zimmerman and S. O. Trach, J. Mol. Biol. 222,599 (1991).

[31] R. J. Ellis and A. P. Minton, Nature 425, 27 (2003); A.P. Minton, J. Cell Science 199, 2863 (2006).

[32] S. R. McGuffee and A. H. Elcock, PLoS Comput. Biol.6, e1000694 (2010).

[33] S. B. Yuste, G. Oshanin, K. Lindenberg, O. Benichou,and J. Klafter, Phys. Rev. E 78, 021105 (2008); E. Abad.S. B. Yuste, and K. Lindenberg, Phys. Rev. E 81, 031115(2010); I. M. Sokolov, S. B. Yuste, J. J. Ruiz-Lorenzo,and K. Lindenberg, Phys. Rev. E 79, 051113 (2009);D. Froemberg and I. M. Sokolov, Phys. Rev. Lett. 100,108304 (2008).

[34] Y. He, S. Burov, R. Metzler, and E. Barkai, Phys. Rev.Lett. 101, 058101 (2008).

[35] S. Burov, R. Metzler, and E. Barkai, Proc. Natl. Acad.Sci. USA 107, 13228 (2010).

[36] R. Metzler, V. Tejedor, J.-H. Jeon, Y. He, W. Deng, S.Burov, and E. Barkai, Acta Phys. Polonica B 40, 1315(2009).

[37] S. Havlin and D. ben-Avraham, Adv. Phys. 36, 695(1987).

[38] A. Klemm, R. Metzler, and R. Kimmich, Phys. Rev. E65, 021112 (2002); see also References therein.

[39] E. W. Montroll and G. H. Weiss, J. Math. Phys. 10, 753(1969).

[40] H. Scher and E. W. Montroll, Phys. Rev. B 12, 2455(1975).

[41] H. Scher, G. Margolin, R. Metzler, J. Klafter, and B.Berkowitz, Geophys. Res. Lett. 29, 1061 (2002); B.Berkowitz, A. Cortis, M. Dentz and H. Scher, Reviewsof Geophysics, 44, RG2003 (2006).

[42] C. Monthus and J.-P. Bouchaud, J. Phys. A 29, 3847(1996); G. Ben Arous, A. Bovier, and V. Gayrard Phys.Rev. Lett. 88, 087201 (2002)

[43] E. Barkai and Y. C. Cheng, J. Chem. Phys. 118, 6167(2003).

[44] J.-P. Bouchaud, J. Phys. (Paris) I, 2, 1705 (1992); G.Bel and E. Barkai, Phys. Rev. Lett. 94, 240602 (2005);A. Rebenshtok and E. Barkai, ibid. 99, 210601 (2007); M.A. Lomholt, I. M. Zaid, and R. Metzler, ibid., 98, 200603(2007); I. M. Zaid, M. A. Lomholt, and R. Metzler, Bio-phys. J. 97, 710 (2009); F. D. Stefani, J. P. Hoogenboom,and E. Barkai, Physics Today 62(2), 34 (2009); G. Mar-golin and E. Barkai, Phys. Rev. Lett. 94, 080601 (2005);G. Aquino, P. Grigolini, and B. J. West, EPL 80, 10002(2007).

[45] R. Metzler, E. Barkai, and J. Klafter, Phys. Rev. Lett.82, 3563 (1999); Europhys. Lett. 46, 431 (1999). E.Barkai, R. Metzler, and J. Klafter, Phys. Rev. E 61, 132(2000); R. Metzler, J. Klafter, and I. M. Sokolov, ibid.58, 1621 (1998).

[46] S. Burov and E. Barkai, Phys. Rev. Lett. 98, 250601(2007).

[47] B. D. Hughes Random Walks and Random Environ-

ments, Volume 1: Random Walks (Oxford UniversityPress, Oxford, 1995).

[48] J.-P. Bouchaud and A. Georges, Phys. Rep. 195, 127(1990).

[49] H. Scher, M. F. Shlesinger, and J. T. Bendler, Phys.Today 44, No. 1, 26 (1991).

[50] Anomalous transport: foundations and applications, ed-ited by R. Klages, G. Radons, and I. M. Sokolov (Wiley-VCH, Weinheim, 2007).

[51] A. N. Kolmogorov, Dokl. Acad. Sci. USSR 26, 115(1940); B. B. Mandelbrot and J. W. van Ness, SIAMRev. 1, 422 (1968); H. Qian, Fractional Brownian Mo-tion and Fractional Gaussian Noise. In G. Rangarajanand M.Z. Ding (eds), Processes with Long-Range Cor-

relations (Springer, Lecture Notes in Physics, Vol.621),pp.22-33.

[52] R. Kubo, M. Toda, and N. Hashitsume, Statistical Phys-ics II: Nonequilibrium Statistical Mechanics (Springer-Verlag, Heidelberg, 1995).

[53] H. E. Hurst, Trans. Am. Soc. Civ. Eng. 116, 400 (1951).[54] T. N. Palmer, G. J. Shutts, R. Hagedorn, F. J. Doblas-

Reyes, T. Jung, and M. Leutbecher, Annu. Rev. EarthPlanet Sci. 33, 163 (2005).

[55] I. Simonsen, Physica A 322, 597 (2003); N. E. Frangos,S. D. Vrontos, and A. N. Yannacopoulos, Appl. Stoch.,Models Bus. Ind. 23, 403 (2007).

[56] T. E. Harris, J. Appl. Probab. 2, 323 (1965); L. Liz-ana and T. Ambjornsson, Phys. Rev. Lett. 100, 200601(2008); L. Lizana, T. Ambjornsson, A. Taloni, E. Barkai,and M. A. Lomholt, Phys. Rev. E 81, 051118 (2010);E. Barkai and R. Silbey, Phys. Rev. Lett. 102, 050602(2009).

[57] A. Taloni, A. V. Chechkin, and J. Klafter, Phys. Rev.Lett. 104, 160602 (2010).

[58] S.C. Kou and X. S. Xie, Phys. Rev. Lett. 93, 180603(2004); W. Min, G. Luo, B. J. Cherayil1, S. C. Kou, andX. S. Xie, Phys. Rev. Lett. 94, 198302 (2005).

[59] M. Magdziarz, A. Weron, K. Burnecki, and J. Klafter,Phys. Rev. Lett. 103, 180602 (2009). Compare also M.Magdziarz and J. Klafter, Phys. Rev. E, at press; K.Burnecki and A. Weron, Phys. Rev. E 82, 021130 (2010).

[60] A. Lubelski, I. M. Sokolov, and J. Klafter, Phys. Rev.Lett. 100, 250602 (2008).

[61] W. H. Deng and E. Barkai, Phys. Rev. E 79, 011112

Page 15: 1 2Physics Department, Technical University of Munich, D ... · arXiv:1009.4846v2 [cond-mat.stat-mech] 13 Oct 2010 22nd October 2018 Single particle tracking in systems showing anomalous

15

(2009).[62] H. C. Fogedby, Phys. Rev. E 50, 1657 (1994); A. Baule

and R. Friedrich, Phys. Rev. E 71, 026101 (2005); M.Magdziarz, A. Weron, and J. Klafter, Phys. Rev. Lett.101, 210601 (2008); M. Magdziarz, A. Weron, and K.Weron, Phys. Rev. E 75, 016708 (2007); E. Heinsalu, M.Patriarca, I. Goychuk, G. Schmid, and P. Hanggi, Phys.Rev. E 73, 046133 (2006).

[63] T. Neusius, I. M. Sokolov, and J. C. Smith, Phys. Rev.E 80, 011109 (2009).

[64] J.-H. Jeon and R. Metzler, Phys. Rev. E 81, 021103(2009).

[65] O. Y. Sliusarenko, V. Y. Gonchar, A. V. Chechkin, I.M. Sokolov, and R. Metzler, Phys. Rev. E 81, 041119(2010).

[66] I. M. Sokolov, E. Heinsalu, P. Hanggi, and I. Goychuk,EPL 86, 30009 (2009).

[67] J.-H. Jeon and R. Metzler, J. Phys. A 43, 252001 (2010).[68] A. Baule and R. Friedrich Eur. Phys. Lett. 77, 10002

(2007).[69] G. Guigas and M. Weiss, Biophys. J. 94, 90 (2008).[70] V. Tejedor, O. Benichou, R. Voituriez, R. Jungmann, F.

Simmel, C. Selhuber, L. Oddershede, and R. Metzler,Biophys. J. 98, 1364 (2010).

[71] A. M. Mathai and R. K. Saxena, The H-function withApplications in Statistics and Other Disciplines (WileyEastern Ltd., New Delhi, 1978); A. M. Mathai, R. K.Saxena, and H. J. Haubold, The H-function, Theory andApplications (Springer, New York, 2010).