Publisher’s Note: Inequivalence of time and ensemble averages in ergodic systems: Exponential...

8
arXiv:1202.6502v1 [cond-mat.stat-mech] 29 Feb 2012 Inequivalence of time and ensemble averages in ergodic systems: exponential versus power-law relaxation in confinement Jae-Hyung Jeon 1, and Ralf Metzler 2,1, 1 Department of Physics, Tampere University of Technology, FI-33101 Tampere, Finland 2 Institute for Physics & Astronomy, University of Potsdam, D-14476 Potsdam-Golm, Germany Single particle tracking has become a standard tool to investigate diffusive properties, especially in small systems such as biological cells. Usually the resulting time series are analyzed in terms of time averages over individual trajectories. Here we study confined normal as well as anomalous diffusion modeled by fractional Brownian motion and the fractional Langevin equation, and show that even for such ergodic systems time-averaged quantities behave differently from their ensemble averaged counterparts, irrespective of how long the measurement time becomes. Knowledge of the exact behavior of time averages is therefore fundamental for the proper physical interpretation of measured time series, in particular, for extraction of the relaxation time scale from data. PACS numbers: 05.40.-a,02.50.-r,05.70.Ln,87.15.Vv I. INTRODUCTION Due to recent advances in single particle tracking tech- niques, analyses based on single trajectory averages have been widely employed to study diffusion in complex sys- tems, e.g., of large biomolecules and tracers in living cells [1]. Examples include the motion in the cellular cytoplasm of messenger RNA molecules [2], chromoso- mal loci [3], lipid granules [4], , and viruses [5], telomeres in cell nuclei [6], or of protein channels in the cell mem- brane [7]. Under the assumption of ergodicity, i.e., the equivalence of (long) time averages (TA) with ensemble averages (EA), the physical interpretation is often based on the time series analysis of single trajectories. For in- stance, particle-to-particle diffusion properties are typi- cally studied via TA mean squared displacements (MSD) of individual time series x(t), δ 2 (Δ) = 1 T Δ T Δ 0 x(t + Δ) x(t) 2 dt, (1) where Δ is the lag time and T the length of the time se- ries. Invoking ergodicity arguments, it is tacitly assumed that δ 2 (Δ) corresponds to the EA MSD x 2 (t)with the identification t Δ, in the limit of long measurement times (i.e., T →∞). For free normal diffusion one can indeed show analytically that x 2 (t)= δ 2 (t)=2K 1 t as T →∞ [11, 12]. At finite T , the result for δ 2 (Δ) will gen- erally show trajectory-to-trajectory variations. However, a similar equivalence still holds when δ 2 is averaged over many individual trajectories: x 2 (t)= δ 2 (t)[11]. In what follows we use the symbol δ 2 when T →∞, and δ 2 for finite T , unless specified otherwise. For anomalous diffusion of the form x 2 (t)=2K α t α with anomalous diffusion constant K α of physical dimension cm 2 /sec α * Electronic address: [email protected] Electronic address: [email protected] and anomalous diffusion exponent α (0 <α< 2) [13], the same conclusion holds if the process is described by fractional Brownian motion (FBM) or the fractional Langevin equation (FLE) [14–16]. In contrast, disagreements between TA and EA are not surprising for non-ergodic processes. A prominent example is anomalous diffusion described by continuous time random walks (CTRW) with diverging characteris- tic waiting times [17–19]: while the EA MSD scales as x 2 (t)〉≃ t κ with 0 <κ< 1, the TA MSD grows linearly with the lag time, δ 2 (Δ) Δ for free motion [11, 12]. Under confinement one observes δ 2 (Δ) Δ 1κ instead of the saturation plateau of the EA [15, 20, 21]. Recently it was found that the TA MSD of tracers in living cells indeed exhibit such CTRW behavior [4, 7]. Here, we show that even for ergodic processes the TA may differ from the EA. This a priori unexpected discrep- ancy arises from the fact that generally dynamic variables are not well-defined in the TA sense, and therefore care is necessary when interpreting TA based on knowledge about the corresponding EA. We explicitly study this ef- fect for stochastic processes of regular Brownian, FBM, and FLE types, confined in an harmonic potential. Pro- cesses of the FBM and FLE kind are closely associated with the motion of tracer molecules in viscous environ- ments, single file diffusion, monomer motion in polymers, or the relative motion of aminoacids in proteins [8]. They have also been identified as stochastic mechanisms for the tracer motion in living cells and reconstituted crowding systems [3, 4, 9, 10]. Consider first an overdamped Brownian particle in the harmonic potential U (x)= kx 2 /2 of stiffness k. With initial position x(0) = 0 the EA MSD is x 2 (t)= (1 e 2kt/γ )/[βk], (2) while the EA taken over the TA MSD (1) becomes δ 2 ,T ) = 2 βk 1 e kΔ + γ 2k(T Δ) e kΔ1 2 e 2kT/γ e 2kΔ . (3)

Transcript of Publisher’s Note: Inequivalence of time and ensemble averages in ergodic systems: Exponential...

arX

iv:1

202.

6502

v1 [

cond

-mat

.sta

t-m

ech]

29

Feb

2012

Inequivalence of time and ensemble averages in ergodic systems: exponential versus

power-law relaxation in confinement

Jae-Hyung Jeon1, ∗ and Ralf Metzler2, 1, †

1Department of Physics, Tampere University of Technology, FI-33101 Tampere, Finland2Institute for Physics & Astronomy, University of Potsdam, D-14476 Potsdam-Golm, Germany

Single particle tracking has become a standard tool to investigate diffusive properties, especiallyin small systems such as biological cells. Usually the resulting time series are analyzed in termsof time averages over individual trajectories. Here we study confined normal as well as anomalousdiffusion modeled by fractional Brownian motion and the fractional Langevin equation, and showthat even for such ergodic systems time-averaged quantities behave differently from their ensembleaveraged counterparts, irrespective of how long the measurement time becomes. Knowledge of theexact behavior of time averages is therefore fundamental for the proper physical interpretation ofmeasured time series, in particular, for extraction of the relaxation time scale from data.

PACS numbers: 05.40.-a,02.50.-r,05.70.Ln,87.15.Vv

I. INTRODUCTION

Due to recent advances in single particle tracking tech-niques, analyses based on single trajectory averages havebeen widely employed to study diffusion in complex sys-tems, e.g., of large biomolecules and tracers in livingcells [1]. Examples include the motion in the cellularcytoplasm of messenger RNA molecules [2], chromoso-mal loci [3], lipid granules [4], , and viruses [5], telomeresin cell nuclei [6], or of protein channels in the cell mem-brane [7]. Under the assumption of ergodicity, i.e., theequivalence of (long) time averages (TA) with ensembleaverages (EA), the physical interpretation is often basedon the time series analysis of single trajectories. For in-stance, particle-to-particle diffusion properties are typi-cally studied via TA mean squared displacements (MSD)of individual time series x(t),

δ2(∆) =1

T −∆

∫ T−∆

0

[

x(t+∆)− x(t)]2

dt, (1)

where ∆ is the lag time and T the length of the time se-ries. Invoking ergodicity arguments, it is tacitly assumedthat δ2(∆) corresponds to the EA MSD 〈x2(t)〉 with theidentification t ↔ ∆, in the limit of long measurementtimes (i.e., T → ∞). For free normal diffusion one can

indeed show analytically that 〈x2(t)〉 = δ2(t) = 2K1t as

T → ∞ [11, 12]. At finite T , the result for δ2(∆) will gen-erally show trajectory-to-trajectory variations. However,a similar equivalence still holds when δ2 is averaged overmany individual trajectories: 〈x2(t)〉 = 〈δ2(t)〉 [11]. In

what follows we use the symbol δ2 when T → ∞, and 〈δ2〉for finite T , unless specified otherwise. For anomalousdiffusion of the form 〈x2(t)〉 = 2Kαt

α with anomalousdiffusion constant Kα of physical dimension cm2/secα

∗Electronic address: [email protected]†Electronic address: [email protected]

and anomalous diffusion exponent α (0 < α < 2) [13],the same conclusion holds if the process is describedby fractional Brownian motion (FBM) or the fractionalLangevin equation (FLE) [14–16].In contrast, disagreements between TA and EA are

not surprising for non-ergodic processes. A prominentexample is anomalous diffusion described by continuoustime random walks (CTRW) with diverging characteris-tic waiting times [17–19]: while the EA MSD scales as〈x2(t)〉 ≃ tκ with 0 < κ < 1, the TA MSD grows linearly

with the lag time, δ2(∆) ≃ ∆ for free motion [11, 12].

Under confinement one observes δ2(∆) ≃ ∆1−κ insteadof the saturation plateau of the EA [15, 20, 21]. Recentlyit was found that the TA MSD of tracers in living cellsindeed exhibit such CTRW behavior [4, 7].Here, we show that even for ergodic processes the TA

may differ from the EA. This a priori unexpected discrep-ancy arises from the fact that generally dynamic variablesare not well-defined in the TA sense, and therefore careis necessary when interpreting TA based on knowledgeabout the corresponding EA. We explicitly study this ef-fect for stochastic processes of regular Brownian, FBM,and FLE types, confined in an harmonic potential. Pro-cesses of the FBM and FLE kind are closely associatedwith the motion of tracer molecules in viscous environ-ments, single file diffusion, monomer motion in polymers,or the relative motion of aminoacids in proteins [8]. Theyhave also been identified as stochastic mechanisms for thetracer motion in living cells and reconstituted crowdingsystems [3, 4, 9, 10].Consider first an overdamped Brownian particle in the

harmonic potential U(x) = kx2/2 of stiffness k. Withinitial position x(0) = 0 the EA MSD is

〈x2(t)〉 = (1 − e−2kt/γ)/[βk], (2)

while the EA taken over the TA MSD (1) becomes⟨

δ2(∆, T )⟩

=2

βk

(

1− e−k∆/γ)

2k(T −∆)

(

ek∆/γ − 1)2 (

e−2kT/γ − e−2k∆/γ)

.(3)

2

100 101 102 103 104 105

0.01

0.1

1

<x2(t)> with <x0>th

<x2(t)> with x0=0

< 2( )> with x0=0 theory<x

2 (t)>/

(k)

-1 o

r <2 (

)>/(

k)-1

t or (timestep)

2

FIG. 1: EA and TA MSD for a Brownian particle in an har-monic potential. Solid lines: Eqs. (2), (3). Symbols: simu-lations with γ = 1, β = 1, k = 1, time step δt = 0.001, andmeasurement time T = 105. The flat curve corresponds tothermal initial conditions.

Here γ is the friction coefficient and β the Boltzmannfactor. Both quantities are identical initially, before con-finement effects come into play: 〈x2(t)〉 ∼ 2K1t ∼ 〈δ2(t)〉with K1 = 1/[βγ]. However, for T,∆ → ∞ (with

T −∆ → ∞) the TA MSD converges to δ2(∆) → 2/[βk],twice the thermal value 〈x2〉th = 1/[βk]. The differencebetween TA and EA is shown in Fig. 1 for both the ana-lytical results and simulations, with excellent agreement.Note the sudden dip of the TA MSD at ∆ ≈ T at thefinite measurement time T , at which the limiting behav-ior 〈δ2(∆ → T )〉 = 〈x2(T )〉 is observed. These featuresare generic for the definition of the TA MSD (1) underconfinement, compare Ref. [15, 20].What results will be obtained for more complicated,

non-Brownian motion? We analyze the case of anoma-lous diffusion governed by FBM and FLE and show thatthe entire relaxation dynamics is significantly differentfor the TA, despite the ergodic nature of these processes.Knowledge of the exact behavior of TA quantities is im-perative for the correct physical interpretation of timeseries, in particular, to extract the relaxation time.

II. FRACTIONAL BROWNIAN MOTION

FBM xα(t) in an external potential U follows theLangevin equation

dxα(t)

dt= −kxα(t) + ξα(t), (4)

driven by fractional Gaussian noise ξα(t) of zero mean〈ξα(t)〉 = 0 and slowly decaying, power-law autocorrela-tion (t 6= t′) [23, 24]

〈ξα(t)ξα(t′)〉 ≃ αKα(α− 1)|t− t′|α−2. (5)

In free space, 〈x2α(t)〉 = 2Kαt

α [14]. Note the change ofsign in Eq. (5) between antipersistent subdiffusion 0 <α < 1 and persistent superdiffusion 1 < α < 2. Differentto subdiffusive CTRW processes with diverging waitingtime scale, FBM does not exhibit ageing effects. In fact,the free space propagator is the Gaussian [25]

P (x, t) =

1

4πKαtαexp

(

−x2

4Kαtα

)

, (6)

whose smooth shape contrasts the pronounced cuspsat the initial position in subdiffusive CTRW processes[13, 18]. Moreover, the propagator (6) obeys a time-localdiffusion equation with time-dependent diffusivity [25].The formal solution of the FBM-Langevin equation

(4),

xα(t) =

∫ t

0

e−k(t−t′)ξα(t′)dt′, (7)

and Eq. (5) yield the position autocorrelation function

〈xα(t1)xα(t2)〉 = Kα

{

e−kt1tα2 + e−kt2tα1 − |t2 − t1|α}

+Kα

2kα

{

e−k|t2−t1|γ(α+ 1, kt1)

+ek|t2−t1|γ(α+ 1, kt2)

−ek|t2−t1|γ(α+ 1, k|t2 − t1|)}

+kKα

2(α+ 1)|t2 − t1|

α+1e−k|t2−t1|

×M(α+ 1;α+ 2; k|t2 − t1|)

−kKα

2(α+ 1)tα+11 e−k(t1+t2)M(α+ 1;α+ 2; kt1)

−kKα

2(α+ 1)tα+12 e−k(t1+t2)M(α+ 1;α+ 2; kt2). (8)

For the EA MSD we then find

〈x2α(t)〉 =

kαγ(α+ 1, kt) + 2Kαt

αe−kt

−kKα

α+ 1tα+1e−2ktM(α+ 1;α+ 2; kt), (9)

where γ(z, x) =∫ x

0 dte−ttz−1 is the incomplete γ functionand

M(a; b; z) =Γ(b)

Γ(b − a)Γ(a)

∫ 1

0

eztta−1(1− t)b−a−1dt

(10)is the Kummer function [22]. Fig. 2 shows simulationsof FBM in an harmonic potential for various α val-ues, demonstrating excellent agreement with result (9).Asymptotic expansion of Eq. (9) at short times t ≪ k−1

yields free anomalous diffusion 〈x2α(t)〉 ∼ 2Kαt

α. Closeto stationarity, we find

〈x2α(t)〉 ∼ 〈x2

α〉th −2

k2α(α− 1)Kαt

α−2e−kt, (11)

3

100 101 102 103

100

101

102

103

104 =0.1 =0.2 =0.5 =1.0 =1.4 =1.6 =1.8 Eq. (4)

<x2 (t)>

/K

t (timestep)

FIG. 2: EA MSD 〈x2

α(t)〉 for FBM in an harmonic potential.

Solid lines: Eq. (9). Symbols: simulations with parametersk = 0.01, T = 2048, and x0 = 0.

exponentially approaching the stationary value

〈x2α〉th =

kαΓ(α+ 1) (12)

with the single characteristic time scale k−1 in the expo-nential function: as observed in Fig. 2, beyond t > k−1

the stationary state is attained independent of α. Thisproperty enables one to study the confinement effect byanalyzing the relaxation of 〈x2

α(t)〉. We also note an in-teresting feature of the relaxation dynamics in the in-termediate timescale: somewhat counterintuitively thesubdiffusive particle overshoots 〈x2

α〉th before a depres-sion back to this value, while for superdiffusion we ob-serve a monotonic increase (the sign of the second termin Eq. (11) depends on α). Note that the α-dependenceof the plateau value (12), a reminder of the fact thatFBM is driven by an external noise and thus not sub-ject to the fluctuation dissipation theorem, in contrast toFLE motion discussed below. Phenomenologically, bothprocesses are very similar.For the TA MSD in the limit T → ∞ we obtain the

expression

δ2(∆) = 2KαΓ(α+ 1)/kα + 2Kα∆α

−Kα

kα{

ek∆Γ(α+ 1, k∆) + e−k∆Γ(α+ 1)}

−kKα

α+ 1∆α+1e−k∆M(α+ 1;α+ 2; k∆), (13)

where Γ(z, x) =∫∞

x dte−ttz−1 is the complementary in-complete γ function. Comparison with the EA MSD,Eq. (9), demonstrates a completely different functionalbehavior over all time scales, except in the short-timelimit for which confinement is negligible. In particular,at ∆ → ∞, we find δ2(∆) = 2〈x2

α〉th for all α.The fundamental difference of the relaxation dynamics

of δ2(∆) and 〈x2α(t)〉 is evidenced in Fig. 3, in excellent

100 101 102 103100

101

102

103

EA MSD (sim.) TA MSD (sim.) theory, Eq. (4) theory, Eq. (7)

<x(t)

>/K

or

2 ()/

K

t or (timestep)

FIG. 3: EA and TA MSD, 〈x2

α(t)〉/Kα and δ2(∆)/Kα for

FBM in an harmonic potential (α = 0.40, 1.0, 1.4, 1.60, bot-tom to top). Solid and dashed lines: analytical results (9)and (13). Symbols: simulations. Parameters as in Fig. 2.

agreement with Eqs. (9) and (13). In contrast to theexponential relaxation of Eq. (9), the TA MSD shows apower-law approach to the limiting value 2〈x2

α〉th, exceptfor the Brownian limit α = 1. This is manifested in theasymptotic form of δ2(∆) at ∆ → ∞,

δ2(∆) ∼ 2〈x2α〉th −

KαΓ(α+ 1)

k2e−k∆

−2α(α− 1)Kα

k2∆2−α. (14)

The transient second term becomes the leading order atα = 1. Surprisingly, in Eq. (14) the relaxation dynamicsis determined by the power exponent α − 2. Moreover,no characteristic time scale exists beyond which the MSDcould be regarded saturated. For subdiffusion, as the al-gebraic decay is relatively fast (∼ ∆−κ with 1 < κ < 2),the MSD appears saturated at sufficiently long measure-ment time. However, the superdiffusive MSD relaxes veryslowly as α is closer to 2 (∼ ∆−κ with 0 < κ < 1). Dueto this the corresponding MSD does not show satura-tion even at long measurement time T . Only in the limit∆ → T the TA dips back to the plateau value of the EA.In typical experiments, however, this feature is obscuredby poor statistics, and thus the relaxation time wouldlikely be overestimated from the TA MSD.

III. FRACTIONAL LANGEVIN EQUATION

.

The FLE describes ergodic anomalous diffusion andfulfills the fluctuation-dissipation theorem [14]. In thepotential U , the FLE motion yα(t) follows the dynamic

4

equation [14, 26, 28]

md2yα(t)

dt2= −γ

∫ t

0

dt′|t− t′|α−2 dyαdt′

− kyα(t)

+√

γ/[α(α− 1)βKα]ξα(t), (15)

where ξα(t) represents fractional Gaussian noise, m is theparticle mass and γ the generalized friction coefficient. Inthe FLE, the dynamic exponent of the noise is restrictedto 1 < α < 2. This persistent noise results in subdiffusivemotion of the FLE in the overdamped limit. For unbiasedmotion (k = 0), 〈y2α(t)〉 = δ2(t) at T → ∞ [14], and

δ2(∆) =2∆2

βmEα,3

[

−Γ(α− 1)γ

m∆α

]

, (16)

Eα,3(z) being a generalized Mittag-Leffler function. Thelatter is defined via its Laplace image

∫ ∞

0

e−utEρ,δ (−η∗tα) =1

uδ + η∗u1−α. (17)

In terms of a series expansion around z = 0 and z → ∞this function reads [27]

Eρ,δ(z) =∞∑

n=0

zn

Γ(δ + ρn)= −

∞∑

n=1

z−n

Γ(δ − ρn). (18)

The MSD (16) accordingly turns from ballistic motion∼ ∆2 at short ∆ to subdiffusion ∼ ∆2−α at long ∆[26, 28]. In the presence of the potential, the FLE (15)can be solved analytically in the overdamped limit, thestationary state yielding [28]

〈yα(t1)yα(t2)〉th =1

βkE2−α

[

−k

γΓ(α− 1)|t2 − t1|

2−α

]

,

(19)with E2−α(z) = E2−α,1. Thus, 〈y2α(t)〉 has the station-ary value 〈y2α〉th = 1/(βk) for any α, contrasting theα-dependent result (12) for FBM. Moreover for T → ∞we obtain the TA MSD

δ2(∆) = 2〈y2α〉th

(

1− E2−α

[

−k

γΓ(α− 1)∆2−α

])

(20)

for ∆ >∼ τc, where the momentum relaxation time is [26]

τc =

(

mΓ(α+ 3)

2Γ(α− 1)(2α+1 − 1)γ

)1/α

. (21)

The TA MSD (20) behaves distinctly different from itsEA counterpart as well as the TA MSD (13) for FBM:the TA MSD (20) grows like ∼ ∆2−α at intermediatelag time, and eventually converges to 2〈y2α〉th for all αas ∆ → ∞. Similar to our above observations, the long-time behavior of Eq. (20) exhibits a power-law relaxation,namely,

δ2(∆) ≈ 2〈y2α〉th

(

1−γ

k∆2−α

)

. (22)

100 101 102 103 104 105

0.01

0.1

12

<y2(t)>th

=1.2 =1.4 =1.6 =1.8 theory, Eq. (12) theory, Eq. (11)

2 (

) /(

k)-1

t or (timestep)

FIG. 4: TA MSD for FLE motion in an harmonic potential,for α = 1.2, 1.4, 1.6, and 1.8 (bottom to top), and T = 217.A representative EA MSD is included for α = 1.2. Symbols:Simulations with time step δt = 0.001, stiffness k = 100, massm = 1, friction coefficient γ = 100, and β = 1, with equilib-rium distribution of initial position yα(0) and velocity yα(0).Full lines: theoretical result (20). Dotted lines: unbiased mo-tion, Eq. (16), with momentum relaxation at α = 1.4 and1.8, illustrating ballistic scaling ∼ ∆2 at ∆ < τc.

As for FBM, the dynamic exponent of the TA MSD is in-dependent of the confinement (k). Intriguingly the speedof convergence is slower as the driving noise ξα is morepersistent (i.e., when α → 2). Therefore, opposite to thetendency shown in Fig. 3 for FBM, in the FLE case theslow particle appears not to approach 2〈y2α〉th.In Fig. 4, we further analyze FLE motion in an har-

monic potential in terms of the TA MSD for various α.For times ∆ >

∼ τc and in the overdamped limit, the ana-lytical form (20) agrees well with the simulations resultsfor all cases. While the TA MSD approaches the thermalvalue 2〈y2α〉th algebraically, in our simulation the slow-est subdiffusive case (corresponding to α = 1.8) does notshow saturation even for long measurement times. Attimes less than τc the TA MSD shows quadratic scalingδ2 ≃ ∆2. The dynamics within this time range is ex-plained well by the full solution of the free space motion(16), shown for α = 1.2 and 1.8. An important fea-ture resulting from this inertia effect are the oscillationsof δ2 that are particularly pronounced as α approaches2. These oscillations are intrinsic in the sense that theyoccur regardless of the confinement, due to the strongpersistence in ξα and inertia effects [29].

IV. DISTRIBUTION OF TIME AVERAGED

MEAN SQUARED DISPLACEMENTS

At finite sampling time T the TA MSD δ2(∆, T ) isa random variable, even for ergodic processes such asBrownian motion, FBM, and FLE motion. In practice,this means that δ2(∆, T ) shows pronounced trajectory-

5

to-trajectory variations. This stochasticity of δ2 is mea-sured by the scatter probability density φ(ξ), in whichthe dimensionless variable ξ is defined through [11]

ξ = δ2/⟨

δ2⟩

. (23)

Such scatter distributions are of Levy stable type for sub-diffusive CTRW processes with diverging characteristictime scale [11, 15, 20, 30, 31]. In Figs. 5-7 we show φfor free and confined FBM and FLE motion, with fixedbin size 0.1. Each graph shows the distributions at threedifferent lag times ∆ and α for given measurement timeT . In each graph two sets of curves were shifted upwardsfor comparison, the shift value is indicated in the graphs.

A. FBM in an harmonic potential

On the Left of Fig. 5 we present the distributions of δ2

for the data shown in Fig. 3. As expected from our pre-vious study [30], the distributions are centered aroundthe ergodic value ξ = 1 and become wider as the lagtime ∆ increases. The wider distribution at longer lagtime means that the single TA MSD trajectories tend tobe more erratic as ∆ approaches T . A new finding isthat at a fixed lag time ∆ > 1 the scatter distributionbecomes broader as the motion is faster (i.e., growingα). This behavior is mainly due to the inherent prop-erty of FBM itself, as the same tendency is also foundwithout potential (see left panel in Fig. 7). This depen-dence on α is attributed to the fact that FBM is a Gaus-sian stationary process in which the spatial displacementx for time difference ∆ is governed by the distribution∼ exp(−x2/[4KH∆α]). On the Right we show the cor-responding distributions when the measurement time isincreased to T = 217. The distributions now appear in-sensitive to α and ∆. The fact that they are less sharpthan the analytically predicted Gaussian is due to thefinite size effect of the binning, see below.

B. FLE motion in an harmonic potential

The left and right panels of Fig. 6 depict the distribu-tions corresponding to the parameters used in Fig. 4, withT = 211 and 217, respectively. Note that the variation ofα was restricted to the range [1.0, 2.0], as FLE motion isonly well defined for subdiffusion. The anomalous diffu-sion exponent in this case is given by κ = 2−α in terms ofthe scaling exponent α of the fractional Gaussian noise.Generally the distributions are bell-shaped. Notably, atfixed lag time the distribution of FLE motion tends tobe wider as the overdamped motion becomes slower (i.e.,for increasing α), as opposed to the case of FBM. Wealso observe that the distributions of FLE motion appeargenerally wider than those of FBM (e.g., Fig. 5 Left), forthe fact that the initial values of position and velocity for

FLE were chosen as the corresponding equilibrium distri-butions. For the case of long measurement time (Right),the distributions again appear insensitive to α and ∆ forthe given bin size.

C. FBM & FLE motion in free space

To appreciate the effect of confinement on the distri-bution, we simulated free FBM and FLE motion for thesame parameters as in Figs. 3 and 4. As shown in Fig. 7in both cases, the scatter distributions manifest featuresconsistent with the confined cases. Only the width ofthe distribution becomes narrower by the presence of theconfining potential, in particular, at longer lag times.

D. Influence of bin size

Typically experimental probing windows are limited,and meaningful quantitative analysis requires more orless coarse binning. In the context of the scatter plotsshown here this practically means that for a bin size of0.1 the peak cannot exceed the value 10, for reasons ofnormalization,

∫∞

0 φ(ξ)dξ = 1. For simulations data wecan arbitrarily increase the accuracy and thus reduce binsizes while still maintaining good statistics. Such a resultis shown in Fig. 8 for bin size 0.01. For this resolutionwe may compare the shape of the distribution φ(ξ) de-termined from simulations with the theoretical approxi-mation valid for short lag times ∆,

φ(ξ) ≈

T −∆

4πτ∗exp

(

−(ξ − 1)2(T −∆)

4τ∗

)

, (24)

as derived in Ref. [30]. Here, the scale τ∗ is only in-troduced to account for correct dimensions and can betaken to one [time unit], compare Ref. [30]. The graphshows nice agreement with the measured data from thesimulations. Interestingly the agreement is somewhatbetter in the confined case. At larger ∆ the distribu-tion becomes wider than predicted by Eq. (24), due tothe strong correlation in successive square displacements,[x(t+∆)− x(t)]2 (see Ref. [30]), and the curves split upfor the different α.

V. DISCUSSION

Studying the representative example of ergodic FBMand FLE motion in an external harmonic potential wedemonstrated that the TA MSD behaves significantly dif-ferent from the EA MSD. Thus, naive interpretation ofsingle trajectory time averages based on the knowledgeof the ensemble behavior may lead to false conclusions onthe physics underlying the observed motion. This so faroverlooked discrepancy is particularly relevant for the re-laxation behavior: while for the EA MSD the relaxation

6

0

2

4

6

8

10

0 0.5 1 1.5 2

φ(ξ)

ξ

∆=500

∆=10 (+3.0)

∆=1 (+6.0)

α=0.2α=1.0α=1.6

0

2

4

6

8

10

0 0.5 1 1.5 2

φ(ξ)

ξ

∆=1000

∆=20 (+3.0)

∆=1 (+6.0)

α=0.2α=1.0α=1.6

FIG. 5: Scatter distributions for fractional Brownian motion in an harmonic potential. Left: T = 211 = 2, 048. Right:T = 217 = 13, 1072. In each graph we compare the results for three different lag times. The two upper sets of curves wereshifted by the indicated amount, for clarity.

0

1

2

3

4

5

0 0.5 1 1.5 2

φ(ξ)

ξ

∆=500

∆=10 (+1.2)

∆=1 (+2.8)

α=1.4α=1.6α=1.8

0

2

4

6

8

10

0 0.5 1 1.5 2

φ(ξ)

ξ

∆=1000

∆=20 (+2.5)

∆=1 (+5.0)

α=1.4α=1.6α=1.8

FIG. 6: Scatter distributions for fractional Langevin equation motion in an harmonic potential. Left: T = 211. Right: T = 217.In each graph we compare the results for three different lag times. The two upper sets of curves were shifted by the indicatedamount, for clarity.

0

2

4

6

8

10

0 0.5 1 1.5 2

φ(ξ)

ξ

∆=500

∆=10 (+3.0)

∆=1 (+6.0)

α=0.2α=1.0α=1.6

0

2

4

6

8

10

0 0.5 1 1.5 2

φ(ξ)

ξ

∆=1000

∆=20 (+2.5)

∆=1 (+5.0)

α=1.4α=1.6α=1.8

FIG. 7: Scatter distributions for free fractional Brownian and Langevin equation motion. Left: T = 211. Right: T = 217. Ineach graph we compare the results for three different lag times. The two upper sets of curves were shifted by the indicatedamount, for clarity.

7

0

2

4

6

8

10

12

0 0.5 1 1.5 2

φ(ξ)

ξ

φ(ξ)α=0.2α=1.0α=1.6

0

2

4

6

8

10

12

0 0.5 1 1.5 2

φ(ξ)

ξ

φ(ξ)α=0.2α=1.0α=1.6

FIG. 8: Scatter distribution for confined (Left) and free (Right) FBM with T = 211, and ∆ = 1. Data are the same as shownin Figs. 5 and 7 to the Left, but analyzed with bin size 0.01. As shown by the grey line the measured scatter nicely agrees withthe predicted Gaussian distribution (24).

time can be read off directly, the corresponding TA MSDappears to suggest a scale-free behavior. Hence it is im-perative to compare to analytical or simulations resultsfor the TA of the system. We note that while here wefocused on an harmonic external potential, the findingsreported here also pertain to other forms of confinement.What is the reason for this disagreement between EA

and TA? We find that for stochastic processes convergingto a stationary state, as 〈[x(t+∆)−x(t)]2〉th only dependson ∆, the definition of the TA MSD (1) in the limit oflong-time measurement leads to the general relation

limT→∞

δ2(∆, T ) = 2〈x2〉th[1− Cx(∆)], (25)

which is independent of diffusion models and details ofconfinement. Here

Cx(∆) =〈x(t)x(t +∆)〉th

〈x2〉th(26)

is the normalized position autocorrelation function.Therefore, the time-averaged variable δ2(∆) in fact is anindicator of the correlation of the spatial displacement,not the TA MSD of a trajectory. For ergodic systemssatisfying the Khinchin theorem (Cx(∆ → ∞) = 0), the

TA MSD always saturates to δ2 → 2〈x2〉th, where therelaxation dynamics reflecting the spatial correlation canbe very slow although the system is already fully relaxed,as shown in this study. Accordingly, in performing singletrajectory analysis, one should be aware of the poten-tial pitfalls in using δ2(∆). In contrast to non-ergodicsystems [11, 15, 20], the anomalous diffusion exponentα and the anomalous diffusion constant Kα can be esti-mated from the log-log plot of the TA MSD at short lagtimes. Meanwhile, physical quantities associated withconfinement such as the effective confinement size andthe relaxation time could be incorrectly deduced fromthe long-∆ behavior of δ2(∆).What alternative definitions of the TA MSD could be

used to mend this problem? Instead of Eq. (1) it would

be a straightforward idea to consider

x2(t) =1

t

∫ t

0

x2(t′)dt′ (27)

For ergodic processes with a stationary state, x2(t) att → ∞ equals

〈x2〉th =

x2e−βU(x)dx/

e−βU(x)dx. (28)

However, the time dependence of x2(t) is different fromthat of 〈x2(t)〉. Even for free diffusion exhibiting

〈x2(t)〉 = 2Kαtα, the ensemble mean of x2(t) is 2Kα

α+1 tα,

and thus this definition does not even work for the Brow-nian case. If a dynamic variable like the MSD as functionof time is concerned, it appears that no systematic wayexists for defining a TA expectation compatible with theanalogous EA.

For finite measurement time T the TA MSD δ2 showstrajectory-to-trajectory variations, even for Brownianmotion. Consistent with previous findings [30] for er-godic processes, the distributions are centered aroundthe ergodic value ξ = 1. Importantly, in all cases the dis-tribution is almost independent of confinement, exceptfor some narrowing at long lag times. For both FBMand FLE at long T the distributions converge, while forshorter T a dependence on α prevails.

Concluding, the study of single trajectory averages is anon-trivial extension of the theory of stochastic processesknowledge of which is necessary to establish quantitativemodels for diffusion-limited processes in small complexsystems. The current work contributes to the develop-ment of such a theory, and to a toolbox of diagnosismethods of the exact stochastic mechanism underlyingexperimental single particle trajectories [4, 7, 10, 32–34].

8

Acknowledgments

We thank E. Barkai and O. Pulkkinen for stimulat-ing discussions. Financial support from the Academy of

Finland (FiDiPro scheme) is gratefully acknowledged.

[1] C. Brauchle, D. C. Lamb, and J. Michaelis, Single parti-

cle tracking and single molecule energy transfer, (Wiley-VCH, Weinheim, 2010); X. S. Xie et al., Annu. Rev. Bio-phys. 37, 417 (2008).

[2] I. Golding and E. C. Cox, Phys. Rev. Lett. 96 098102(2006).

[3] S. C. Weber, A. J. Spakowitz, and J. A. Theriot, Phys.Rev. Lett. 104, 238102 (2010).

[4] J.-H. Jeon, V. Tejedor, S. Burov, E. Barkai, C. Selhuber-Unkel, K. Berg-Sørensen, L. Oddershede, and R. Metzler,Phys. Rev. Lett. 106, 048103 (2011).

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

[6] I. Bronstein, Y. Israel, E. Kepten, Y. Shav-Tal, E. Barkai,and Y. Garini, Phys. Rev. Lett. 103, 018102 (2009).

[7] A. V. Weigel, B. Simon, M. M. Tamkun, and D. Krapf,Proc. Nat. Acad. Sci. USA 108, 6438 (2011).

[8] S. C. Kou and X. S. Xie, Phys. Rev. Lett. 93, 180603(2004); A. Taloni, A. V. Chechkin, and J. Klafter, Phys.Rev. Lett. 104, 160602 (2010) and Refs. therein; L.Lizana, T. Ambjornsson, A. Taloni, E. Barkai, and M.A. Lomholt, Phys. Rev. Lett. 81, 051118 (2010).

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

[10] M. Magdziarz, A. Weron, K. Burnecki, Phys. Rev. Lett.103, 180602 (2009).

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

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

[13] R. Metzler and J. Klafter, Phys. Rep. 339, 1 (2000); J.Phys. A 37, R161 (2004).

[14] W. Deng and E. Barkai, Phys. Rev. E. 79, 011112 (2009).[15] S. Burov, J.-H. Jeon, R. Metzler, and E. Barkai, Phys.

Chem. Chem. Phys. 13, 1800 (2011).[16] I. Goychuk, Phys. Rev. E 80, 046125 (2009).[17] E. W. Montroll and G. H. Weiss, J. Math. Phys. 6, 167

(1965).[18] H. Scher and E. W. Montroll, Phys. Rev. B 12, 2455

(1975).[19] The related concept of weak ergodicity breaking was de-

veloped by Bouchaud and subsequently by Barkai and

coworkers: J.-P. Bouchaud, J. Phys. I 2, 1705 (1992); G.Bel and E. Barkai, Phys. Rev. Lett. 94, 240602 (2005); A.Rebenshtok and E. Barkai, ibid. 99, 210601 (2007). Theconsequences for the mixing behavior between bound-aries and bulk diffusion were explored by M. A. Lomholt,I. M. Zaid, and R. Metzler, Phys. Rev. Lett. 98, 200603(2007).

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

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

[22] M. Abramowitz and I. A. Stegun, Handbook of Mathe-

matical Functions (Dover, New York, NY, 1972).[23] A. N. Kolmogorov, Dokl. Acad. Sci. USSR 26, 115

(1940); B. B. Mandelbrot and J. W. van Ness, SIAMRev. 1, 422 (1968).

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

[25] E. Lutz, Phys. Rev. E 64, 051106 (2001).[26] J.-H. Jeon and R. Metzler, Phys. Rev. E. 81, 021103

(2010).[27] A. Erdelyi, Editor, Higher Transcendental Functions,

Bateman Manuscript Project Vol. (McGraw Hill, NewYork, NY, 1954).

[28] S. C. Kou, Ann. Appl. Stat. 2, 501 (2008).[29] A variety of oscillatory behaviors occur in 〈yα(0)yα(t)〉 by

combined effects of the memory of fractional noise andconfinement [S. Burov and E. Barkai, Phys. Rev. Lett.100, 070601 (2008); Phys. Rev. E. 78, 031112 (2008)].

[30] J.-H. Jeon and R. Metzler, J. Phys. A 43, 252001 (2010).[31] I. M. Sokolov, E. Heinsalu, P. Hanggi, and I. Goychuk,

Europhys. Lett. 86, 30009 (2009).[32] V. Tejedor, O. Benichou, R. Voituriez, R. Jungmann,

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

[33] T. Akimoto, E. Yamamoto, K. Yasuoka, Y. Hirano, andM. Yasui, Phys. Rev. Lett. 107, 178103 (2011).

[34] S. Condamin, V. Tejedor, R. Voituriez, O. Benichou, andJ. Klafter, Proc. Natl. Acad. Sci. USA 105, 5675 (2008).