<<

in Heat Bath: Exact simulation of time-lapse-recorded data, exact analytical benchmark

Simon F. Nørrelykke Department of Molecular Biology, Princeton University, Princeton, New Jersey, USA. and Henrik Flyvbjerg Department of Micro- and Nanotechnology, Technical University of Denmark, Kongens Lyngby, Denmark. (Dated: May 29, 2018) The stochastic dynamics of the damped harmonic oscillator in a heat bath is simulated with an algorithm that is exact for time steps of arbitrary size. Exact analytical results are given for correlation functions and power spectra in the form they acquire when computed from experimental time-lapse recordings. Three applications are discussed: (i) Effects of finite sampling-rate and -time, described exactly here, are similar for other stochastic dynamical systems—e.g. motile micro-organisms and their time-lapse recorded trajectories. (ii) The same statistics is satisfied by any experimental system to the extent it is interpreted as a damped harmonic oscillator at finite temperature—such as an AFM cantilever. (iii) Three other models of fundamental interest are limiting cases of the damped harmonic oscillator at finite temperature; it consequently bridges their differences and describes effects of finite sampling rate and sampling time for these models as well. Finally, we give a brief discussion of nondimensionalization.

Keywords: Ornstein-Uhlenbeck process, power-spectral analysis, mean-squared displacement, discrete sam- pling, discretization effects, Langevin, Brownian motion, optical trap, leakage, non-dimensionalization, cor- relation function

I. INTRODUCTION dom motion, e.g. cell migration [8–10, 12–14, 25, 26] and commodity pricing [24]. Consequently, the treatment The damped harmonic oscillator in a heat bath is given here of its experimental statistics for time-lapse the archetypical bounded Brownian dynamical system recorded data should be useful in several ways: with inertia and the simplest possible of this kind. An- (i) All experimental statistics contain effects of finite alytically solvable, it offers insights that are valid also sampling rate, finite sampling time, and finite statis- for more complex systems. Its experimental correlation tics. They do so to various degrees, but the effects are functions and power spectra are given analytically here inherent in . We give exact analytical ex- in the form they take when computed from time-lapse pressions for these effects, for the model treated here. recorded trajectories. Such trajectories are generated Statistics for more complex systems contain the same and analyzed for illustration in an exact Monte Carlo effects, qualitatively, and quantitatively as well, by de- simulation. The results differ significantly from those grees we estimate. derived from the standard continuous-sampling formu- (ii) The statistics we describe below must be found lation. for any experimental system to the extent that system is interpreted as a damped harmonic oscillator at finite In mathematical terms, it is the Ornstein-Uhlenbeck temperature—e.g. an AFM cantilever to be calibrated (OU) process in a Hookean force field we treat, and the by interpretation of its thermal power spectrum. results discussed here extend the ones given in [19, 23] as (iii) Three other models are limiting cases of the well as the free case treated by Gillespie [11]. Three clas- model treated here: sical papers on the OU-process are [7, 29, 30]. Histori- cally, the model was prompted by a question by Smolu- 1. At vanishing mass, Einstein’s theory for Brownian chowski regarding how inertia might modify Einstein’s motion in a harmonic trap, which, e.g., is a min- theory of Brownian motion [20]. H. A. Lorentz soon af- imal model for the Brownian dynamics of a mi- arXiv:1102.0524v1 [physics.data-an] 2 Feb 2011 ter pointed out that Ornstein’s answer to that question crosphere held in an optical trap, with magnetic was insufficient for the case of classical Brownian mo- tweezers [28], or surface-tethered by DNA [1]. tion, i.e., in a fluid of density similar to the Brownian particle [16]: Hydrodynamical effects, entrainment and 2. At vanishing external force, the Ornstein- back-flow, were more important with the time resolu- Uhlenbeck model of free Brownian motion with tion that was available in the first century of Brownian inertia, which, e.g., is a minimal model for the motion. Inertia in classical Brownian motion became persistent random motion seen in trajectories of experimentally relevant only with precision calibration motile cells. of optical tweezers [2, 3] and was directly observed only 3. At vanishing mass and external force, Einstein’s in 2005 [17, 27]. original theory for Brownian motion in a fluid at The OU process remained and remains, however, an rest. important model for Brownian motion dominated by in- ertia, such as massive particles [5, 6, 15] or AFM can- The results for the harmonic oscillator, given below, tilevers [22] in air, and for other kinds of persistent ran- carry over to these three models. The limits are not 2 all obvious, but always enlightening, hence described which, for arbitrary positive ∆t, and with tj = j∆t, below. xj = x(tj), and vj = v(tj), gives us the recursive relation (iv) As the model treated here bridges the differences between the three limiting cases, the material pre-  x   x   ∆x  j+1 = e−M∆t j + j , (7) sented here is well suited for a pedagogical, hands-on vj+1 vj ∆vj computer-based introduction to the four dynamic sys- tems covered here: free/bound diffusion, with/without where inertia, their equilibrium behavior, correlations, and √ power spectra, and their transient behavior to equilib-   tj +∆t   2D Z 0 ∆xj 0 −M(tj +∆t−t ) 0 rium. = dt e 0 ∆vj η(t ) τ tj (8) and the time-independent matrix exponential can be II. EXACT DISCRETIZED written EINSTEIN-ORNSTEIN-UHLENBECK THEORY −M∆t − ∆t OF BROWNIAN MOTION IN HOOKEAN e = e 2τ [cos(ω∆t)I + sin(ω∆t)J] (9) FORCE FIELD with The Einstein-Ornstein-Uhlenbeck theory for the   1 1 ! Brownian motion of a damped harmonic oscillator in one 1 0 2ωτ ω I ≡ and J ≡ −ω2 . (10) dimension is simply Newton’s Second Law for the oscil- 0 1 0 −1 lator with a thermal driving force, a.k.a. the Langevin ω 2ωτ equation for this system, That the matrix exponential can be written this way,

mx¨(t) + γx˙(t) + κx(t) = Ftherm(t) (1) can be proven in several ways. We used the algebra of Pauli matrices. Alternatively, one may observe that F (t) = (2k T γ)1/2η(t) . (2) therm B the cosine and sine terms on the RHS are the even and Here x(t) is the coordinate of the oscillator as function odd parts of the LHS. The latter are straight-forwardly, of time t, m its inertial mass, γ its friction coefficient, κ if tediously, computed from the Taylor series for the exponential. Inserting Eq. (9) in Eq. (8) we see that is Hooke’s constant, and Ftherm is the thermal force on the oscillator. Equation (2) gives the amplitude of this √ Z tj+1 t −t thermal noise explicitly in terms of γ, the Boltzmann 2D − j+1 ∆x = dt e 2τ sin(ω(t − t)) η(t) (11) energy k T , and η(t), which is a normalized white-noise j j+1 B ωτ tj process, i.e., the time derivative of a Wiener process, η = dW/dt, hence √ tj+1 2D Z tj+1−t 0 0 − 2τ hη(t)i = 0 ; hη(t)η(t)i = δ(t − t ) for all t, t . (3) ∆vj = − 2 dt e sin(ω(tj+1 − t)) η(t) 2ωτ tj Equation (1) can be rewritten as two coupled first-order (12) √ differential equations, Z tj+1 t −t 2D − j+1 + dt e 2τ cos(ω(t − t)) η(t) √ τ j+1 d  x(t)   x(t)  2D  0  tj = −M + , (4) dt v(t) v(t) τ η(t) are two correlated random numbers from zero-mean where we have introduced Einstein’s relation D = Gaussian distributions. They can be written as a lin- k T/γ and the 2 × 2 matrix ear combination of two independent Gaussian variables: B The four parameters that determine this linear combi-  0 −1   0 −1  nation can be chosen at will, as long as the combination M = = . (5) κ γ ω2 1 has the same variance-covariance as the two original cor- m m 0 τ related variables. This is a direct consequence of the Gaussian distribution being completely determined by Here ω = pκ/m is the cyclic of the un- 0 its mean and variance-covariance. Thus, we can write damped oscillator, and τ = m/γ is the characteris- tic time of the exponential decrease with time that the momentum of the particle undergoes in the ab- ∆xj = σxx ξj (13) 2 p 2 4 2 sence of all but friction forces. Below, we shall also ∆vj = σxv/σxx ξj + σvv − σxv/σxx ζj , (14) need the cyclic frequency of the damped oscillator, p 2 2 p 2 2 ω = κ/m − γ /(4m ) = ω0 − 1/(4τ ), which is real where the σs are elements of the variance-covariance ma- for less than critical , γ2 < 4mκ. trix (see below), and ξ and ζ are independent random Equation (4) is solved by numbers with Gaussian distribution, unit variance, and √ zero mean. This particular choice of linear combina-   Z t   x(t) 2D 0 −M(t−t0) 0 tion mirrors the structure of Eqs. (11) and (12). Using = dt e 0 , (6) v(t) τ −∞ η(t ) Eqs. (3), (11), and (12) we calculate that the elements 3

of the variance-covariance matrix are, for ω 6= 0 2

1 2 2 D 2 2 σxx ≡ h(∆xj) i = 2 2 3 4ω τ + (15) 0 4ω ω0τ −1 − ∆t 2 2 

τ Positon (nm) e [cos(2ω∆t) − 2ωτ sin(2ω∆t) − 4ω0τ ] 6 −2 15 16 17 18 19 20 Time (ms) D σ2 ≡ h(∆v )2i = 4ω2τ 2+ (16) vv j 4ω2τ 3 4 ∆t  − τ 2 2 e [cos(2ω∆t) + 2ωτ sin(2ω∆t) − 4ω0τ ] Positon (nm) 2 D − ∆t 2 2 σ ≡ h∆x ∆v i = e τ sin (ω∆t) . (17) xv j j ω2τ 2

At critical damping (ω = 0) this variance-covariance 0 simplifies to

2 2 σxx ≡ h(∆xj) i (18)   −2 − ∆t 1 2 0 20 40 60 80 0 500 1000 = 4Dτ 1 − e τ [1 + ∆t/τ + (∆t/τ) ] 2 Time (ms) Count

2 2 FIG. 1. Positions of harmonically trapped massive parti- σvv ≡ h(∆vj) i (19)   cles in a thermal bath. Three different regimes are studied − ∆t 1 2 by varying the drag coefficient above and below the crit- = D/τ 1 − e τ [1 − ∆t/τ + (∆t/τ) ] 2 ical value, where ω = 0, by an order of magnitude: Un- derdamped (red, γ = γcrit/10), critically damped (green, ∆t crit 2 − τ 2 γ = γ , offset by 2 nm), and overdamped (blue, γ = σxv ≡ h∆xj∆vji = De (∆t/τ) . (20) 10 γcrit, offset by 4 nm). Simulation parameters in Eq. (7): 2 Figure 1 shows the simulated positions in the un- m = 1 ng, κ = 225 mg/s , T = 275 K, fsample = 65, 536 Hz, derdamped, critically damped, and overdamped regime. (x(0), v(0)) = (0, 0), and the same (ξj , ηj ) are used for the Note that all three regimes have the same mean (zero) three simulations. The values for the underdamped case are alike to those for an AFM cantilever in water, whereas the and variance (kBT/κ). Only at short times, as shown in the insert, does the difference between the three regi- critical and overdamped case corresponds to the same can- tilever in a more viscous environment or with a smaller spring mens reveal itself: The position coordinate of the under- constant. The insert shows a magnified portion of the trace, damped system oscillates, while the position coordinate revealing the oscillating, critical, and random nature of the of the overdamped system does not show discernible per- motion, respectively. On the right, histograms show the dis- sistence of motion. The position coordinate of the criti- tribution of the position data with kBT/κ-variance Gaus- cally damped system looks at times like it oscillates, but sians overlaid. actually only displays positively correlated random mo- tion, as revealed by the velocity auto-correlation func- tion given below and shown in Fig. 3B. 2 where hx i = kBT/κ, i.e., the MSD is a constant minus twice the auto-covariance of the position process (see III. MEAN-SQUARED-DISPLACEMENT Sec. IV). This is illustrated in Fig. 2 where we show the MSD for the three time series plotted in Fig. 1. When a time-series of positions is examined, the first measure applied is frequently the mean-squared- displacent (MSD). This has to do with its ease of com- putation, the existence of exact analytical results for the expectation value, and comparative robustness of IV. COVARIANCE the measure to experimental errors. Also, there are no discretization effects to worry about—the MSD for finite sampling frequency is simply the MSD for The stationary-state solution given in Eq. (6) oscil- continuous recording, evaluated at discrete time points. lates in resonating response to the thermal noise that More importantly, unlike the covariance and the power- drives it, which is seen by its oscillating auto-covariance spectra discussed below, the MSD is well-defined even functions, but its oscillations are randomly phase shifted if the process is not bounded. In the present example, by the same noise that drives oscillations, which causes both the position and velocity processes are bounded, the in auto-covariance functions, see and we therefore have a simple relation between the below. The result is that equal-time ensemble averages MSD and the auto-covariance: like hx(t)2i are constant in time. This is easily proven for quadratic expressions, either directly from the solu- MSD(t) ≡ h(x(t) − x(0))2i = 2hx2i − 2hx(t)x(0)i (21) tion given in Eq. (6), or by differentiation w.r.t. time 4

) 100 For ω real (underdamped system), these covariances os- 2 cillate, we see, with amplitudes that decrease exponen- tially in time with characteristic time 2τ. For ω imagi- nary (overdamped system), we rewrite Eq. (24) as

 hx(t)x(0)i hx(t)v(0)i  (25) 10 hv(t)x(0)i hv(t)v(0)i ( 1 ! −1 ! ) D ω2τ 1 −t/τ ω2τ −1 −t/τ = 0 + e − + 0 − e + . 2|ω|τ −1 −1 1 1 Underdamped, simulation τ− τ+ Critically damped, simulation Overdamped, simulation from which we see that the system decreases as a double- 1 Underdamped, theory Critically damped, theory exponential with characteristic times τ± = 2τ/(1 ± Overdamped, theory 2τ|ω|). At critical damping (ω = 0), the expressions Contiuous recording, theory for the covariances simplify to Mean − squared displacement (nm 0.1 1 10 Time−lag (ms)  hx(t)x(0)i hx(t)v(0)i  (26) hv(t)x(0)i hv(t)v(0)i FIG. 2. Mean-squared-displacements for harmonically 1+t/(2τ) ! D − t 2 t trapped massive particles. Same simulation conditions as = e 2τ ω0 . in Fig. 1, except tmsr = 8 s. The result of the stochastic τ −t 1 − t/(2τ) simulations are shown in red, green, and blue. Comple- mentary colors (cyan, magenta, and yellow) show the MSD, Figure 3 illustrates these dynamics for the normalized Eq. (21), evaluated at discrete time-points with the same pa- covariances, aka the correlation functions. rameter values as in the simulation. Thin black lines show As is the case for the MSD, there are no discretiza- the MSD Eq. (21) for continuous time. Red and blue lines tion effects to worry about—the covariance for finite show slopes of two and one, respectively. Notice how the sampling frequency is simply the covariance for con- MSDs all plateau at the same level for time-lags larger than tinuous recording, evaluated at discrete time points. 3 s. Only for shorter time-lags do the three regimes differ. However, the given velocity correlations, and thus also the position-velocity correlations, are only correct for the actual, instantaneous velocities: If the instanta- using It¯o’slemma: neous velocity is not directly measured, but instead es-  2  timated from measured positions as a “secant-velocity” d hx i  hxvi  = (22) (see Eq. (60) and Sec. VII A 1), then the correspond- dt hv2i ing secant-velocity correlation function should be cal-    2    culated from the position covariance function given in 0 2 0 hx i 0 Eq. (24). Figure 4 shows how such secant-velocity cor- 2 −1 0  −ω0 τ 1   hxvi  +   . relations deviate from the instantaneous-velocity corre- 2 −2 hv2i 2D 0 −2ω0 τ τ 2 lations, mainly for short time-lags. As is seen by inspection, this equation has the time- independent solution V. POWER SPECTRA

 2   kBT  hx i κ  hxvi  =  0  , (23) The power spectral density (PSD) of a variable x(t) hv2i kBT or v(t) is defined as the expectation value of the squared m modulus of its , with a normalization in accord with the equipartition theorem. This solution that differs between authors. We choose one in which is the unique attractor for the system’s dynamics: Left the power spectral density remains constant for increas- to itself, any discrepancy from this time-independent so- ing measurement time, tmsr lution will decrease to zero exponentially fast in time— 2 possibly while oscillating harmonically with cyclic fre- (x) 2 D/(2π ) P (fk) ≡ h|x˜k| i/tmsr = 2 2 2 2 2 (27) quency 2ω—as is seen from the fact that the 3×3-matrix (2πτ) (fk − f0 ) + fk in Eq. (23) has eigenvalues −1/τ and −1/τ ± i2ω, and 2 determinant −4ω0/τ. and Similar reasoning (or simply multiplying Eq. (7) with (v) 2 2 (x) P (fk) ≡ h|v˜k| i/tmsr = (2πfk) P (fk) (xj, vj), then taking the expectation value, and applying the equipartition theorem) gives the covariances 2D = 2 2 2 , (28) (2πτ) (fk − f /f) + 1  hx(t)x(0)i hx(t)v(0)i  0 (24) hv(t)x(0)i hv(t)v(0)i where 2πf0 = ω0 is the frequency of the sin ωt system. Here, we usedv ˜ = i2πx˜ , which is an ap- cos ωt+ sin ωt ! k k D t 2ωτ − 2 proximation that ignores contributions from the finite = e 2τ ω0 ω . τ − sin ωt sin ωt ω cos ωt − 2ωτ measurement-time. 5

1 1 Underdamped, secant theory Critically damped, secant theory A Overdamped, secant theory 0.5 0.5

0 0

1

−0.5 −0.5 0.5 Position autocorrelation Velocity autocorrelation

0

exp[−t/(2 )] 0 0.05 0.1 ! Time−lag (ms) −1 −1 0.1 1 10 0.1 1 10 Time−lag (ms) Time−lag (ms) 1 FIG. 4. Secant-velocity autocorrelation functions for massive B particles in a harmonic potential, driven by thermal noise. Thin black lines, red dashed lines, and simulation settings 0.5 are the same as in Fig. 3B. Additional symbols: Cyan-red, magenta-green, and yellow-blue squares show the autocorre- lations for wj = (xj − xj−1)/∆t calculated using hx(t)x(0)i 0 from Eq. (24). Insert shows the first 0.1 ms on a linear time- scale, revealing the very fast decay of the overdamped os- cillator and how the secant-velocity autocorrelations over- estimate the instantaneous-velocity autocorrelation decay- −0.5 times. As expected, the secant-velocity does a better job

Velocity autocorrelation estimating the instantaneous velocity for the less damped system because of its smoother particle-trajectory. exp[−t/(2!)] −1 0.1 1 10 Time−lag (ms) The finite-time Fourier transform (FTFT) used above is defined as 1 Z tmsr i2πfkt x˜k = dt e x(t), fk ≡ k/tmsr, k integer . C 0 (29) 0.5 In a similar manner we can calculate the finite-sampling- frequency PSDs by discrete Fourier transformation (DFT) of Eq. (7) 0  xˆ   xˆ   ∆ˆx  e−iπfk/fNyq k = e−M∆t k + k , (30) vˆk vˆk ∆ˆvk velocity crosscorrelation − −0.5 where

N 2 N X i2πjk/N X iπjfk/fNyq Position exp[−t/(2!)] zˆk = ∆t e zj = e zj , (31) −1 fNyq 0.1 1 10 j=1 j=1 Time−lag (ms) for z = x, v, ∆x, ∆v; xN+1 = x1 and vN+1 = v1;

FIG. 3. Correlation functions, Eq. (24), for massive par- fk = k∆f = k/tmsr (32) ticles in a harmonic potential, driven by thermal noise. k = 0, 1, 2,...,N − 1 (33) Color-coding and simulation settings are the same as in Fig. 2: Red, green, and blue show the result of a stochas- tmsr = N ∆t = N/fsample (34) tic simulation; complementary colors show the covariance N fsample f = ∆f = , (35) (not a fit), Eq. (24), normalized and evaluated at dis- Nyq 2 2 crete time-points; thin black lines show the covariance for continuous time (not a fit); and dashed red lines show and, the enveloping exponential exp(−t/(2τ))). Panel A: Po- ∗ 2 ∗ 2 h∆ˆx ∆ˆx 0 i = σ hξˆ ξˆ 0 i = σ ∆t t δ 0 (36) sition auto-correlations, hx(t)x(0)i/hx2i. Panel B: Veloc- k k xx k k xx msr k,k 2 ∗ 2 ity auto-correlations, hv(t)v(0)i/hv i. Panel C: Position- h∆ˆvk∆ˆvk0 i = σvv ∆t tmsr δk,k0 (37) p velocity cross-correlations, hx(t)v(0)i/ hx2ihv2i. The ∗ 2 h∆ˆxk∆ˆvk0 i = σxv ∆t tmsr δk,k0 . (38) velocity-position cross-correlation, hv(t)x(0)i/phx2ihv2i, (not shown) has the opposite sign but is otherwise identi- cal. 6

After isolating (ˆxk, vˆk) in Eq. (30), we find

 2 ∗   2 2  −12 A h|xˆk| i hxˆkvˆki −1 σx σxv † −1 10 ∗ 2 = Ak 2 2 (Ak) tmsr∆t hxˆkvˆki h|vˆk| i σxv σv −13 (39) 10

where the 2 × 2 matrix and its inverse are /kg) 2 −14 s 10 2 −1 αkI − βJ Ak = αkI + βJ ; A = (40) k 2 2 (m −15 αk + β ! 10 Underdamped, simulation with −16 Critically damped, simulation PSD/ 10 ∆t Overdamped, simulation −iπfk/fNyq − αk = e − e 2τ cos(ω∆t) (41) Underdamped, theory −17 − ∆t 10 Critically damped, theory β = −e 2τ sin(ω∆t) , (42) Overdamped, theory

−18 Continuous recording, theory 10 complex scalars (β/ω is real), Hermitian conjugation is 1 2 3 4 10 10 10 10 denoted by †, and complex conjugation by ∗. Frequency (Hz) For the discrete positional power spectrum we thus −3 10 get (see also [19, Appendix A]) Underdamped, simulation Critically damped, simulation (x) 2 Overdamped, simulation B Pk ≡ h|xˆk| i/tmsr (43) −4 10 Underdamped, theory n o 2 β 2 2 β β 2 β 2 Critically damped, theory |αk − 2ωτ | σxx − 2 ω Re αk − 2ωτ σxv + ω2 σvv = Overdamped, theory 2 2 2 −5 Continuous recording, theory |αk + β | fsample /kg) 10 2 Aliased theory

whereas the discrete velocity power spectrum is (m ! −6 (v) 2 10 Pk ≡ h|vˆk| i/tmsr (44) PSD/ β 2 2 2 β n β o 2 4 β2 2 |αk + | σvv + 2ω0 Re αk + σxv + ω0 2 σxx 2ωτ ω 2ωτ ω −7 10 = 2 2 2 . |αk + β | fsample

In case of critical damping (ω = 0) we simply in- −8 ∆t 10 − 1 2 3 4 sert β/ω = −e 2τ ∆t in Eqs. (43) and (44), and use 10 10 10 10 Eqs. (18,19,20) for the variance-covariances. Frequency (Hz) Figure 5 shows the power spectra that result from nu- merical simulations of AFM cantilever positions and ve- FIG. 5. Power spectra for positions and velocities of harmon- locities using Eq. (7) for three different drag coefficients, ically trapped massive particles in thermal bath, normalized as well as the corresponding analytical expressions as by γ. The color-scheme is the same as in Figs. 2 and 3: given in Eqs. (43,44). Synthetic data is shown in red, green, and blue; the aliased An alternative route to these discrete PSDs is to take theory, Eqs. (43) and (44), is shown in complementary col- the discrete Fourier transform of the covariance function ors; and the non-aliased theory, Eqs. (27) and (28), as thin from the previous section: Both the position and the black lines. Simulation parameters are as in Fig. 1 (giving velocity processes are stationary (the joint probability ω0 = 15 kHz) with nwin = 32 Hann windows applied to the density function for each is independent of time), so the tmsr = 8 s long time-series. Panel A: Power spectra for po- sitions, f −2 and f −4 behaviors are illustrated by the thin Wiener-Khinchin theorem applies and the Fourier trans- blue and red lines, respectively. Notice the disagreement form of the correlation function is equal to the PSD. For between the non-aliased (continuous recording) theory and vanishing spring constant, κ, the position process is un- data at high . Panel B: Power spectra for veloc- bounded (not stationary) and the covariance, as well ities. Dashed black lines show aliased versions of Eq. (28), as the PSD, are consequently ill-defined. This limit is (v,aliased) P∞ (v) i.e., P (f) = n=−∞ P (f + nfsample), [2], here treated in Sec. VII A. truncated to 201 terms. The non-aliased theory severely Note, the PSD for discretely measured (instanta- underestimates the power at high and low frequencies, and neous) velocities, Eq. (44), should not be confused with completely misses the data in the overdamped case. the discrete PSD for secant-velocities, Eq. (61). The latter is the correct expression to use in experiments 2 2 where the velocities are estimated from the measured 2D∆t/τ , and σ√xv = 0. That is, ∆xj = 0 and positions. ∆vj = σvvζj = 2D∆t/τ ζj. In other words, the ve- locity process is seen to be driving the position process. In this limit the positional PSD takes on the famil- VI. VANISHING ∆t: THE LIMIT OF iar form given in Eq. (27), and the velocity PSD is CONTINUOUS RECORDING given in Eq. (28). Comparing the continuous record- ing PSDs Eqs. (27,28) with their discrete sampling ana- As ∆t → 0 we find to first order in ∆t that the co- logues Eqs. (43,44) we see in Fig. 5A that they differ 2 2 variances in Eqs. (15–20) reduce to σxx = 0, σvv = substantially at high frequencies, and in Fig. 5B that 7 they differ everywhere but near f0 in the underdamped is not the friction between the cell and substrate but de- case. Thus, one should not attempt to fit a theoretical scribes the rate of memory-loss for the velocity process. PSD derived for “continuously recorded” trajectories to The structure of Eqs. (7,9,10,13,14) remain the same, an experimental PSD, which necessarily is obtained with the only change is ω0 = 0 in Eq. (10), hence ω = i/(2τ), time-lapse recording. More specifically, one should only and the covariances, Eqs. (15,16,17), consequently re- attempt to fit it to the low-frequency part of the experi- duce to mental spectrum for positions, or account for “aliasing” 2 2  before fitting, as described e.g. in [2, Appendix H], which σxx = Dτ 2∆t/τ − 4(1 − a) + (1 − a ) (48) 2 2 turns Eqs. (27) and (28) into Eqs. (43), respectively (44). σvv = D/τ 1 − a (49) The position-covariance and the MSD do not suffer σ2 = D (1 − a)2 , (50) from this dichotomy. None of them change their shape xv or form when switching between discrete and continuous where time: The discrete versions are equal to the continuous versions, evaluated at discrete times: a = e−∆t/τ , τ = m/γ . (51)

hxjxj+`i = hx(tj)x(tj+`)i = hx(0)x(t`)i (45) For the velocity and position processes we thus find v = av + ∆v (52) and j+1 j j xj+1 = xj + τ(1 − a)vj + ∆xj . (53) h(x −x )2i = h(x(t )−x(t ))2i = h(x(0)−x(t ))2i , j j+` j j+` ` That is, the velocity process has reduced to a stable (46) autoregressive model of order one, AR(1). Its time in- where, in the last step, we used that the process is sta- tegral, the position process, is unbounded—the particle tionary and the measures therefore invariant to time- is diffusing freely and without limits. Exact numeri- translations. That is, these measures are unaffected by cal update formulas for v and x have previously been the unavoidable discreteness of real-world data. j j given in [11], here we re-derived them for consistency of The velocity’s covariance is also unaffected by time- notation. lapse recording, if we can measure the instantaneous The discrete-time positional PSD, P (x), is obtain- velocity, i.e. the velocity vector that is tangential to k able from Eq. (43) with ω = 0 and the σs given in the trajectory of the position. If we cannot and time- 0 Eqs. (48,49,50); the expression does not simplify sig- lapse record only positions, we have a different situation, nificantly compared to Eq. (43). The discrete-time which is treated below. (v) velocity PSD, Pk , is much simpler and can be de- rived directly from the discrete-time velocity process, VII. VARIOUS PHYSICAL LIMITS: A Eq. (52); or by taking the κ = 0 limit of Eq. (44). From REFERENCE-SET OF FORMULAS these discrete-time PSDs the continuous-recording ex- pressions, P (x)(f) and P (v)(f), can be obtained by ex- panding to leading order in ∆t/τ and k/N = f /f ; Three physical limiting cases are of particular inter- k sample or they can be derived directly from the continuous-time est: (i) Vanishing spring constant, κ = 0, which is the equation of motion Eq. (47) by Fourier transformation Ornstein-Uhlenbeck theory for Brownian motion for a and, for the positional PSD, remembering thatx ˙ = v. free particle with inertia; (ii) vanishing mass, m = 0, The continuous-recording positional PSD and the two which is Einstein’s theory for Brownian motion of a par- velocity PSDs then read ticle trapped by a Hookean force, and a popular mini- malist model for the Brownian motion of a microsphere D/(2π2) P (x)(f ) = (54) in an optical trap; and (iii) vanishing mass and spring k (2πτ)2f 4 + f 2 constant, which is Einstein’s theory for Brownian mo- k k σ2 ∆t tion of a free particle. P (v) = vv (55) k 1 + a2 − 2a cos(2πk/N)

(v) 2D P (fk) = 2 . (56) A. Vanishing spring constant: 1 + (2πfkτ) The Ornstein-Uhlenbeck process Note, that the positional PSD diverges in fk = 0 due to the process being unbounded. This implies that we When there is no Hookean restoring force κ = 0, cannot use the Wiener-Khinchin theorem to obtain the Eq. (1) describes the free diffusion of a massive particle covariance by Fourier transforming the PSDs, and vice according to Ornstein and Uhlenbeck [29] versa. The velocity process is bounded, however, and mv˙(t) + γv(t) = F (t) . (47) the velocity correlation functions are straightforward to therm calculate, so we simply list the results for the discrete This velocity-process is known as the Ornstein- and the continuous case Uhlenbeck (OU) process [29]. When modeling other D hv v i = e−|i−j|∆t/τ (57) dynamical systems, such as migrating cells, m does not i j τ refer to the physical mass of the cell but rather its inertia D 0 hv(t)v(t0)i = e−|t−t |/τ . (58) to velocity-changes, or persistence of motion; likewise γ τ 8

As already mentioned, the time integral of the OU- −12 process is one of the instances where the mean-squared- 10 displacement provides a useful measure for the position process, while the positional covariance function is ill- defined. The mean-squared-displacement of the time integral of the OU-process is /s) 2 −13   10 h(x(t) − x(0))2i = 2Dτ t/τ + e−t/τ − 1 . (59)

2

For t  τ this MSD increases as t (ballistically) and for Power (m t  τ as t (diffusively), with an exponential cross-over OU velocity, simulation −14 between the two regimes with characteristic time τ. 10 OU secant velocity, simulation OU velocity, theory OU secant velocity, theory OU secant velocity, Eq. (68) theory

1. Secant-velocities 1 2 3 4 10 10 10 10 Frequency (Hz) In experiments, the velocity is typically not measured directly, but approximated from the measured positions FIG. 6. Power spectra of velocities and secant-velocities for as a “secant-velocity”: the OU process. Blue: Simulated velocities, Eq. (52). Red: Secant-velocities, Eq. (60), calculated from simulated posi- wj = (xj − xj−1)/∆t (60) tions, Eq. (53). Yellow: Aliased velocity theory, Eq. (55). with discrete power spectral density Cyan: Aliased secant-velocity theory, Eq. (61). Black line: Approximated aliased secant-velocity theory, Eq. (68). Sim- (w) 2 ulation settings are given the caption of Fig. 7. Notice the P ≡ h|wˆk| i/tmsr k discrepancy between the true velocity and the estimated (se- 2 (1 − cos(πf /f )) = k Nyq P (x) . (61) cant) velocity at high frequencies. (∆t)2 k This is a general result that follows from the definition of wj and is independent of the dynamic model. We now apply the Wiener-Khinchin theorem to get the For the OU-process, we can relate this PSD to the PSD as the Fourier transform of this approximated auto- PSD for the continuous velocity that it approximates, if correlation function and find we rewrite it as 2 P (w) = P (v) (τ/∆t)2(1 − a)2 + σ2 /∆t 1 ∆t k k xx P (w) ≈ P (v) − D. (68) ∗ k k 3 τ + τ(1 − a)(hvˆk ∆ˆxki + c.c.) . (62) Here the first term is proportional to P (v), the second term is a constant, and the last term is frequency depen- This approximation to the secant-velocity PSD, as well dent with c.c. the complex conjugate of the other term as the exact expression Eq. (61), are both shown in in the bracket. However, it is not easy to see from this Fig. 6, together with numerical simulation results and expression how much P (w) deviates from P (v). We can the PSD of the continuous velocity Eq. (55). Notice, get a good idea about the shape of the PSD from the that even with ∆t/τ = 0.27 < 1 (∆t = 1/fsample and (v) (w) auto-correlation function τ = m/γ), the relative difference between Pk and Pk  τ 2 is substantial for higher frequencies. Thus, proper care hw2i = 2hv2i (∆t/τ − 1 + a) (63) should be taken if empirical testing of a model includes j j ∆t the fitting of a theoretical velocity PSD to an experimen-  1  2 tal secant-velocity PSD, even when aliasing is properly ≈ hvj i 1 − ∆t/τ , (64) 3 accounted for. But, we now also know how to do this and for ` > 0 correctly: Either add 1/3(∆t/τ)2D to the experimental (v)  τ 2 secant-velocity PSD before fitting with Pk , or fit di- hwjwj+`i = 2hvjvj+`i (cosh(∆t/τ) − 1) (65) (w) ∆t rectly using the full theoretical expression for Pk given  1  in Eq. (61). In cases where this corrective constant is ≈ hv v i 1 + (∆t/τ)2 , (66) j j+` 12 much smaller than the velocity PSD for all frequencies fitted, it can obviously be ignored. where we used Eqs. (13,14,50,53,57) to derive We observe that the secant velocity is the time- Eqs. (63,65)—the latter two expressions are propor- average of the real velocity in the time interval spanned tional to the velocity autocorrelations they approximate, by the secant. Time-averaging is low-pass filtering, as is they only differ by multiplicative factors that are inde- well known [21, Chapter 13] and demonstrated in Fig. 6. pendent of the time-lag for ` > 0. To first order in ∆t/τ we thus have, for ` ≥ 0,  1 ∆t  hw w i = hv v i 1 − δ . (67) j j+` j j+` 3 τ 0,` B. Vanishing mass: 9

The optical trap limit hence

The position process in the limit of vanishing mass hξji = 0, hξiξji = δi,j , (81) is mathematically identical to the OU velocity process in the limit of vanishing trapping force treated above. and the velocity and position PSDs for discrete and con- Equation (1) reduces to tinuous sampling take on the simple forms: 2 (x) D/fsample γx˙(t) + κx(t) = Ftherm(t) (69) Pk = 2 (82) 2 sin (πfk/fsample) with solution (x) D P (fk) = 2 (83) t 2π2f 1 Z 0 k 0 −2πfc(t−t ) 0 x(t) = dt e Ftherm(t ) , (70) (v) (v) γ −∞ Pk = P (fk) = 2D. (84) where the corner frequency fc ≡ κ/(2πγ) is the fre- As was the case for the time integral of the OU process, quency where P (x)(f) = P (x)(0)/2, see below. Recy- the position PSDs are singular in fk = 0 because the cling the results from Section VII A we can directly write process is unbounded. That is, the meaningful measure down the autocovariance to study is not the covariance, but rather the mean- squared-displacement 0 hx(t)x(t0)i = hx2i e−2πfc|t−t | (71) h(x(t) − x(0))2i = 2Dt , (85) 2 with hx i = kBT/κ and the discrete update rules which is one of the few well-defined statistics for Ein- xj+1 = c xj + ∆xj (72) stein’s theory for Brownian motion: Because the trajec- c = exp(−κ∆t/γ) (73) tory of positions is a fractal, attempt at estimating the r average speed of Brownian motion from the displace- (1 − c2)Dγ ∆xj = ξj , (74) ment of position occurring in a given time-interval√ will κ depend on the duration, ∆t, of this interval as 1/ ∆t, where ξ are uncorrelated random numbers of unit vari- hence diverge when accuracy is sought improved by re- ance, zero mean, and Guassian distribution. Likewise, ducing ∆t. This was not appreciated before Einstein’s the position PSDs for discrete sampling, respectively, 1905 paper on the subject. continuous recording are found to be Figures 7 and 8 show the power spectra, respectively mean-squared-displacements obtained in numerical sim- 2 (x) D (1 − c ) ∆tγ/κ ulations of free diffusion and trapped diffusion, as well P = (75) k 1 + c2 − 2c cos(2πk/N) as the graphs of the corresponding analytical expres- D/(2π2) sions. At short time-scales, i.e., at high frequencies P (x)(f ) = . (76) k 2 2 in Fig. 7 and for small time-lags in Fig. 8, the ther- fc + fk mal forces dominate and the Hookean force has not had Finally, the mean-squared-displacement is time to influence the motion through its constant, but weak, confining effect. That is why the Brownian mo-   tion (green) and optical trap (red) data collapse in this h(x(t) − x(0))2i = 2Dγ/κ 1 − e−tγ/κ , (77) regime, whereas the Ornstein-Uhlenbeck process (blue) differ from the two due to inertial effects. Conversely, which shows an exponential cross-over from linear de- at long time scales, low frequencies in Fig. 7 and large pendence on t to the constant value 2Dγ/κ as t  κ/γ, time-lags in Fig. 8, inertia plays no role, so the Ornstein- see Fig. 2. Uhlenbeck process and Einstein’s theory for Brownian motion are indistinguishable, whereas the Hookean force has had time to exert its confining effect on the optical C. Vanishing mass and spring constant: trap data. Einstein’s Brownian motion

When both m = 0 and κ = 0, Eq. (1) reduces to VIII. NONDIMENSIONALIZATION

γx˙(t) = Ftherm(t) (78) Above, we worked with dimension-full equations be- so that cause they are physically intuitive and allow for dimen- √ sional checks of the calculations along the way. Extra xj+1 = xj + 2D ∆t ξj , (79) insight can be gained however, by nondimensionalizing the equations. where As an example, in the dynamic equation for the OT case 1 Z tj+1 ξj ≡ √ dt η(t) , (80) p ∆t tj γx˙ + κx = Ftherm(t) = 2kBT γ η(t) , (86) 10

−15 10 there are three dimension-full parameters and two dimension-full variables. Dividing through with one of the parameters there are only two left. From these we can form a characteristic time and a characteristic length. By expressing times and lengths in terms of

s) these, we have removed all parameters from the equa- 2 tion. In other words, there is only a single unique equa-

−20 tion. 10 What is a natural choice for the characteristic time EB, simulation

Power (m and length? By considering the physics, we see that OT, simulation OU, simulation switching off the driving-force will result in a solution to EB, theory the dynamics that is an exponentially decaying distance OT, theory from x(t) to x = 0. The characteristic time, γ/κ, for this OU, theory Continuous recording, theory decay is the characteristic time for the dynamics and it

1 2 3 4 makes sense to measure time in units of this time. With 10 10 10 10 Frequency (Hz) the driving-force on, a statistical equilibrium establishes itself. In this equilibrium the distribution of positions has a certain width that is temperature-dependent be- FIG. 7. Power spectra of positions. Green points: Freely dif- cause the driving-force is thermal. This width phx2i, is fusing massless particle (Einstein’s Brownian motion); Red a natural unit in which to measure distances and gives points: Trapped massless particle (OT limit, or OU veloc- ity process); and Blue points: Freely diffusing massive par- us a position variable that is dimensionless. ticles (time integral of OU process). Complementary col- The basic idea then, is to express the independent ors show the finite sampling frequency (aliased) theories, variable t, and dependent variable x in terms of di- whereas the continuous recording (non-aliased) theories are mensionless ones and then use the freedom we have shown as thin black lines. Green and blue lines indicate to choose parameters to make as many pre-factors as f −2 and f −4 behavior, respectively. Simulation parameters: possible equal to unity [4]. Above we used a physi- 2 D = 0.46 µ m /s, T = 275 K, m = 1 ng, and fc = 500 Hz, cal approach. In a more mathematical approach the f = 32, 768 Hz, N = 131, 072, n = 32 Hann win- sample win recipe is to write x = xcχ and t = tcθ where xc and tc dows. The simulation parameters for the OT case are those are dimension-full “characteristic” quantities, whereas of a 1 µm diameter polystyrene sphere held in an optical trap χ and θ are the new dimensionless dependent and in- in water at room temperature. The parameters for Einstein’s Brownian motion are the same, except κ = 0. For the OU dependent variables respectively. Dividing by γ we first process we increased the density of the sphere roughly 2,000 get times, which is not a physically realistic scenario but allows xc dχ κ 1 us to plot all power spectra with the same axes. + xcχ = ftherm(θ) (87) tc dθ γ γ where p ftherm(θ) = Ftherm(tcθ) = 2kBT γ/tc η˘(θ) (88)

) 1000 2 with √ η˘(θ) = tcη(tcθ) (89) and

hη˘(θ)˘η(θ0)i = δ(θ − θ0); hη˘i = 0 . (90) 100 Note, thatη ˘ is a dimensionless generalized function of EB, simulation the dimensionless parameter θ, whereas η is a general- OT, simulation √ OU, simulation ized function of t and has dimension 1/ t. Next, by EB, theory choosing OT, theory OU, theory p tc = γ/κ and xc = 2kBT/κ (91) Continuous recording, theory Mean − squared displacement (nm 1 0.1 1 10 we get Time−lag (ms) dχ + χ =η ˘(θ) , (92) FIG. 8. Mean-squared-displacement for the Ornstein- dθ Uhlenbeck, optical-trap, and Brownian-motion limits de- scribed in Sec. VII A, VII B, and VII C. Simulation settings as our dimensionless dynamic equation. Without hav- and legends are the same as in Fig. 7, except green/blue lines ing actually solved the original equation we have gained show slopes of one/two. three insights: (i) The system has a characteristic time- scale of γ/κ; (ii) the system has a characteristic length- p scale of 2kBT/κ (compare to the MSD); and (iii) there 11 is only one OT-equation, not a whole family for different IX. SUMMARY AND CONCLUSIONS values of γ and κ. Since the OU-process for the velocity We examined the dampened harmonic oscillator and three of its physical limits: The mass-less case (opti- cal trap), the free case (the Einstein-Ornstein-Uhlenbeck mv˙ + γv = F (t); v =x ˙ (93) therm theory for Brownian motion), and the mass-less free case (Einstein’s Brownian motion). By solving the system’s is mathematically identical to the OT process for the dynamical equations for an arbitrary time-lapse ∆t, ex- positions we directly have the dimensionless pendant act analytical expressions were derived for the changes in position and velocity during such a time-lapse. With dν these expressions, exact simulations of the dynamics are + ν =η ˘(θ) , (94) then possible—with an accuracy that is independent of dθ the duration of the time-lapse. In contrast, a numerical simulation, using Euler-integration or similar schemes, by writing v = vcν, t = tcθ, and choosing is exact only to first or second order in ∆t [18]. We gave exact analytical expressions for power- p tc = m/γ and vc = 2kBT/m . (95) spectral forms, mean-squared-displacements, and correlation-functions that can be fitted (see [19] before undertaking a least-squares-fit) to data obtained from For completeness, we also give the dimensionless ver- time-lapse recording of a system with dynamics similar sion of the original dynamics. The dimension-full equa- to the dampened harmonic oscillator or one of its three tion physical limits described here. The effect of finite sampling rates (aliasing) were also discussed. mx¨ + γx˙ + κx = F (t) (96) The effect on the power spectrum of velocity estima- therm tion from position-data (secant-velocity) was treated for the case of free diffusion of a massive particle. Approxi- can be rewritten as the universal oscillator equaition mate as well as exact corrective factors and expressions were given. Finally, we discussed some of the advan- d2χ dχ tages of and of nondimensionaliza- + 2λ + χ =η ˘(θ) (97) tion of the dynamic equations. Throughout, we pointed dθ2 dθ out when power spectral analysis makes sense (bounded process) or not, which of the statistical measures that √ 2 where λ√ = γ/ mκ is the damping ratio, xc = depend on the sampling-frequency, and which may be 3 2 2kBT γ/ mκ , and tc = m/κ. That is, there are described by the simpler continuous-time theory. three dimension-full parameters and a family of solu- tions parametrized by λ. To arrive at dimensionless versions of the discrete X. ACKNOWLEDGMENTS equations we can proceed to solve the above dimen- sionless equations as before. Alternatively, we can sim- S.F.N. gratefully acknowledges financial support from ply take the already known dimension-full results and the Carlsberg Foundation and the Lundbeck Foundati- nondimensionalize them. ion.

[1] J. F. Beausang, C. Zurla, L. Finzi, L. Sullivan, and P. C. [6] D. R. Burnham, P. J. Reece, and D. McGloin. Pa- Nelson. Elementary simulation of tethered Brownian rameter exploration of optically trapped liquid aerosols. motion. American Journal of Physics, 75(6):520–523, Phys. Rev. E, 82(5):051123, Nov 2010. 2007. [7] S. Chandrasekhar. Stochastic problems in physics and [2] K. Berg-Sørensen and H. Flyvbjerg. Power spectrum astronomy. Rev Mod Phys, 15:0001–0089, Dec 1943. analysis for optical tweezers. Rev Sci Instrum, 75:594– [8] E. A. Codling, M. J. Plank, and S. Benhamou. Random 612, Jan 2004. walk models in biology. Journal of The Royal Society [3] K. Berg-Sørensen, E. J. G. Peterman, T. Weber, C. F. Interface, 5(25):813–834, Apr 2008. Schmidt, and H. Flyvbjerg. Power spectrum analysis [9] U. Euteneuer and M. Schliwa. Persistent, directional for optical tweezers. II: Laser wavelength dependence of motility of cells and cytoplasmic fragments in the ab- parasitic filtering, and how to achieve high bandwidth. sence of microtubules. Nature, 310(5972):58–61, Jan Rev Sci Instrum, 77(6):063106, Jan 2006. 1984. [4] E. Buckingham. On Physically Similar Systems; Illus- [10] M. H. Gail and C. W. Boone. Locomotion of mouse trations of the Use of Dimensional Equations. Phys. fibroblasts in tissue culture. Biophysical Journal, Rev., 4(4):345–376, Oct 1914. 10(10):980–993, Jan 1970. [5] D. R. Burnham and D. McGloin. Radius measurements [11] D. T. Gillespie. Exact numerical simulation of the of optically trapped aerosols through Brownian motion. Ornstein-Uhlenbeck process and its integral. Phys. Rev. New J. Phys., 11(6):063022, Jun 2009. E, 54(2):2084–2091, 1996. 12

[12] R. L. Hall. Amoeboid movement as a correlated walk. [22] J. Sader. Frequency response of cantilever beams im- J Math Biol, 4(4):327–35, Oct 1977. mersed in viscous fluids with applications to the atomic [13] L. Li, E. C. Cox, and H. Flyvbjerg. “Dicty Dynam- force microscope. Journal of applied physics, 84:64, ics”: Dictyostelium motility as persistent random mo- 1998. tion. Submitted (2010). [23] L. Schimansky-Geier and C. Z¨ulicke. Harmonic noise: [14] L. Li, S. F. Nørrelykke, and E. C. Cox. Persistent Cell Effect on bistable systems. Zeitschrift f¨ur Physik B Con- Motion in the Absence of External Signals: A Search densed Matter, Jan 1990. Strategy for Eukaryotic Cells. PLoS ONE, 3(5):e2093, [24] E. Schwartz and J. Smith. Short-Term Variations and May 2008. Long-Term Dynamics in Commodity Prices. Manage- [15] T. Li, S. Kheifets, D. Medellin, and M. G. Raizen. Mea- ment Science, 46(7):893–911, Jul 2000. surement of the Instantaneous Velocity of a Brownian [25] D. Selmeczi, L. Li, L. I. I. Pedersen, S. F. Nørrelykke, Particle. Science, 328(5986):1673–1675, Jun 2010. P. H. Hagedorn, S. Mosler, N. B. Larsen, E. C. Cox, [16] H. A. Lorentz. Lessen over Theoretishe Natuurkunde, and H. Flyvbjerg. Cell motility as random motion: A volume V. E. J. Brill, Leiden, 1921. review. Eur Phys J-Spec Top, 157:1–15, Jan 2008. [17] B. Luki´c,S. Jeney, C. Tischer, A. J. Kulik, L. Forr´o,and [26] D. Selmeczi, S. Mosler, P. H. Hagedorn, N. B. Larsen, E.-L. Florin. Direct Observation of Nondiffusive Motion and H. Flyvbjerg. Cell motility as persistent random of a Brownian Particle. Phys. Rev. Lett., 95(16):160601, motion: Theories from experiments. Biophysical Jour- Oct 2005. nal, 89(2):912–31, Aug 2005. [18] R. Mannella. A gentle introduction to the integration of [27] D. Selmeczi, S. F. Toli´c-Nørrelykke, E. Schaffer, P. H. stochastic differential equations. Stochastic Processes in Hagedorn, S. Mosler, K. Berg-Sorensen, N. B. Larsen, Physics, Chemistry, and Biology, pages 353–364, 2000. and H. Flyvbjerg. Brownian motion after Einstein and [19] S. F. Nørrelykke and H. Flyvbjerg. Power spectrum Smoluchowski: Some new applications and new experi- analysis with least-squares fitting: Amplitude bias and ments. Acta Phys Pol B, 38(8):2407–2431, Jan 2007. its elimination, with application to optical tweezers and [28] A. J. W. te Velthuis, J. W. J. Kerssemakers, J. Lipfert, atomic force microscope cantilevers. Review of Scientific and N. H. Dekker. Quantitative Guidelines for Force Instruments, 81(7):075103, Jul 2010. Calibration through Spectral Analysis of Magnetic [20] L. S. Ornstein. On the Brownian Motion. Proc. Royal Tweezers Data. Biophys J, 99(4):1292–1302, Sep 2010. Acad. Amst., 21:96–108, 1919. [29] G. Uhlenbeck and L. Ornstein. On the theory of [21] W. H. Press, S. A. Teukolsky, W. T. Vetterling, and the Brownian motion. Physical Review, 36(5):823–841, B. P. Flannery. Numerical Recipes in FORTRAN: The 1930. Art of Scientific Computing. Cambridge University [30] M. C. Wang and G. E. Uhlenbeck. On the Theory of the Press, New York, NY, USA, second edition, 1992. Brownian Motion II. Rev Mod Phys, 17(2-3):323–342, Jan 1945.