bioRxiv preprint doi: https://doi.org/10.1101/359240; this version posted June 29, 2018. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license. Biophysical Journal Volume: 00 Month Year 1–12 1

Anomalous in inverted variable-lengthscale fluorescence correlation spectroscopy

M. Stolle and C. Fradin

Abstract Using fluorescence correlation spectroscopy (FCS) to distinguish between different types of diffusion processes is often a per- ilous undertaking, as the analysis of the resulting autocorrelation data is model-dependant. Two recently introduced strategies, however, can help move towards a model-independent interpretation of FCS experiments: 1) the obtention of correlation data at different length-scales and 2) its inversion to retrieve the mean-squared displacement associated with the process under study. We use computer simulations to examine the signature of several biologically relevant diffusion processes (simple diffusion, continuous-time , caged diffusion, obstructed diffusion, two-state diffusion and diffusing diffusivity) in variable-lengthscale FCS. We show that, when used in concert, lengthscale variation and data inversion permit to identify non-Gaussian processes and, regardless of Gaussianity, to retrieve their mean-squared displacement over several orders of magnitude in time. This makes unbiased discrimination between different classes of diffusion models possible. Received for publication xxx and in final form xxx.

1 Introduction characteristic decay time of the FCS autocorrelation func- tion (ACF), τ (5). The diffusion law provides a proxy Quantifying the motions of macromolecules in cells, while 1/2 for the particles MSD. Initially, VLS-FCS was achieved by an important task, is complicated. The cellular environment enlarging the usually diffraction-limited point-spread func- is crowded and heterogenous, and many biomolecules tran- tion (PSF) that sets the value of w (5–7). Since then, several siently interact with others. Because of this, macromolecular methods have been proposed to extend the range of avail- diffusion in cells can take many forms and is seldom sim- able w below the diffraction limit (8–10). Another interest- ply Brownian. It may exhibit a mean-squared displacement ing development has been the generation of VLS-FCS data (MSD) that is not linear in time, a distribution of displace- from imaging modalities (total internal reflection, single- ments that is not Gaussian, or both (1). Langowski’s study of plane illumination), in which case the observation volume the diffusion of the green fluorescent in cells, using size can be varied by binning pixels (11, 12). fluorescence correlation spectroscopy (FCS), was one of the The ultimate limit on FCS spatial resolution, however, first to put this anomalous behaviour in the spotlight (2). is not set by the size of the observation volume, but by the FCS, a technique based on the statistical analysis of the temporal resolution of the signal from which the ACF is fluorescence signal recorded from a well-defined observa- computed (13–15). The displacements of fluorophores over tion volume (radius w) is perfectly suited to the range of distance smaller than w still result in changes in fluorescence concentrations (1 to 100 nM) and diffusion coefficients (1 intensity - with low contrast -, and they are therefore cap- to 100 µm2/s) typically encountered for soluble in tured in the short-time regime of the ACF. Thus even with cells (3, 4). Further, the systematic variation of w has made a fixed-size observation volume, FCS can resolve motions this technique especially useful for the study of anomalous over a range of lengthscales on either sides of the diffrac- diffusion processes, which are often lengthscale-dependent. tion limit. This capacity was demonstrated in a cluster of This scheme, known as spot variation FCS or more generally studies exploring DNA segments dynamics, where the ACF variable-lengthscale FCS (VLS-FCS), allows the construc- was inverted to directly recover the MSD of the segments tion of the so-called ”diffusion law”, that is the relationship (16–18). The same ACF inversion procedure was later used between projected area of the observation volume, w2, and to study the effect of crowding in membranes and polymer

© 2013 The Authors 0006-3495/08/09/2624/12 $2.00 doi: 10.1529/biophysj.106.090944 bioRxiv preprint doi: https://doi.org/10.1101/359240; this version posted June 29, 2018. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license. 2 2 THEORY solutions (19, 20). A similar idea was later implemented for 2.2 ACF for an isotropic Gaussian diffusive process image correlation spectroscopy, using a different mathemat- If the propagator is both Gaussian and isotropic, it can be ical scheme to extract the MSD from spatiotemporal ACFs, expressed as a function of the MSD, hr2i(τ) : and used to characterize the anomalous behaviour of GFP  −3/2 2 diffusion in cells (15, 21). Of note, both these MSD recov- 2π − 3r p(r, τ) = hr2i(τ) e 2hr2i(τ) , (2) ery schemes work under the assumption of a process with a 3 Gaussian propagator. Following analytical work by Hofling¨ & Franosch (1, which leads to: 22), we recently showed that variation of the FCS obser- −1 − 1 G(τ, w)  hr2i(τ)   hr2i(τ)  2 vation volume size can be combined with ACF inversion = 1 + 2 1 + 2 . to obtain the MSD of tracer particles for over 5 orders of G(0, w) 3w /2 3(Sw) /2 (3) magnitude in time (20). The two model systems we studied The above equation is valid for any lag time τ at which (crowded dextran solutions and agarose gels) behaved very the propagator is Gaussian. It is the basis for the inver- differently in regard to the superimposition of the apparent sion procedure used in this work, and illustrated in Fig. 1, MSDs extracted from ACFs obtained at different length- ˜2 scales (20). This dissimilarity reflects a qualitative differ- where hr i(τ) is obtained from G(τ, w)/G(0, w) by invert- ence in the nature of the propagator (a.k.a. distribution of ing Eq. 3 (the tilde is used as a reminder that the apparent displacements) of the underlying process. Since only pro- MSD extracted from the ACF might differ from the actual cesses with Gaussian propagators are expected to lead to a MSD when the propagator is not Gaussian). perfect superimposition of the apparent MSDs, the combi- nation of VLS-FCS with ACF inversion provides a test of 2.3 Relationship between ACF and MSD at short lag times the Gaussianity of the diffusive process (20, 23). Here we study the signature of different biologically rel- We show in this section that for short τ there is a simple lin- evant diffusive processes (continuous time random walk, ear relationship between G(τ, w)/G(0, w) − 1 and hr2i(τ), two-component diffusion, diffusing diffusivity, obstructed whether the propagator is Gaussian or not. The only neces- diffusion, caged diffusion). Since analytical solutions are sary assumption is that the propagator is isotropic, in which available only in a small number of limiting cases, we per- case it can be written p(r, τ). At lag times much smaller formed 3D single agent simulations for all these processes, than the ACF characteristic decay time (τ  τ1/2), particles and obtained apparent MSDs from the inversion of ACFs have not yet diffused over distances comparable to the obser- simulated for a range of w. By highlighting the different vation volume radius, in other words hr2i(τ)  w2. This signatures of each diffusion process, our work provides a means that p(r, τ) ' 0 for r > w, in which case O(~r + ~ρ) benchmark for model-independent interpretation of inverted can be replaced in Eq. 1 by the first two even terms of its VLS-FCS experiments. It also refines our understanding of Taylor series expansion in ρ/w (the odd terms are left out the relationship between apparent MSD and real MSD. We as they disappear when integrating over ~ρ). Performing the show in particular that the inversion procedure yields the cor- integration over ~r in Eq. 1 then yields: rect MSD at short lag times regardless of the nature of the G(τ, w) ZZZ ρ2 ρ2 ρ2 propagator, suggesting possible refinements for VLS-FCS x y z ' d~ρp(ρ, τ)(1 − 2 )(1 − 2 )(1 − 2 2 ). experiments. G(0, w) w w S w (4) Again ignoring higher order terms in (ρ/w)2, and using the 2 2 2 1 2 2 Theory equality hrxi = hryi = hrz i = 3 hr i valid for an isotropic 2 2 propagator, we get (for τ  τ1/2, or hr i  w ): 2.1 General form of the ACF G(τ, w) (2 + 1/S2) hr2i(τ) The fluorescence signal I(t) obtained in an FCS experiment ' 1 − 2 . (5) is correlated in time to give the ACF, G(τ) = hδI(t)δI(t + G(0, w) 3 w τ)i/hIi2, where δI(t) = I(t) − hIi. For a transport pro- For a Gaussian process, this linear relationship between ACF cess with propagator p(~r, τ), and a 3D Gaussian observation and MSD can be recovered directly from Eq. 3, by per- volume (radius w, aspect ratio S) with normalized profile forming a first-order Taylor expansion in hr2i/w2. A more 2 2 2 2 2 2 O(~r) = e−2x /w e−2y /w e−2z /(Sw) , the ACF becomes: general form of this equation has been derived in Ref. (24).

3 2 2 Z 2x2 2y 2z 2 − − − 2 G(τ, w) = d~re w2 e w2 e (Sw) hciπ3S2w6 (1) 2.4 Normalization of the ACF 2 2 2 Z 2(x+ρx) 2(y+ρy ) 2(z+ρz ) − − − 2 d~ρp(~ρ,τ)e w2 e w2 e (Sw) , When inverting an experimentally obtained ACF to obtain the apparent MSD, the first step is to normalize its ampli- where hci is the average fluorophore concentration. tude to obtain G(τ, w)/G(0, w). When the actual value of

Biophysical Journal 00(00) 1–12 bioRxiv preprint doi: https://doi.org/10.1101/359240; this version posted June 29, 2018. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license. 2.5 ACF for a truncated detection volume 3

2 2 to be large enough for hr i(τ)  hr i(τmin). This is illus- trated in Fig. 1 for a simple diffusive process: hr˜2i deviates 2 from the actual hr i by less than 5% as long as τ > 20τmin.

2.5 ACF for a truncated detection volume In the simulations presented below, the observation volume profile was truncated in all three space directions to reduce computational times. Truncated observation volumes have been considered before, for either reflective or absorbing boundary conditions (25, 26). Here we consider the case that corresponds to our simulations, where particles become invisible when leaving the rectangular volume with dimen- sions 2bw ×2bw ×2bSw centred on the observation volume, but allowed to diffuse in and out. For a Gaussian propagator and Gaussian detection profile, the ACF is then given by:

GT (τ,w,b) = gT (τ,w,b)gT (τ,w,b)gT (τ,Sw,Sb), (7) GT (0,w,b) x y z where:

−1 Z +bw 2 T 1 a (τ, w) − 2u (1+a−2(τ,w)) g (τ,w,b) = du e w2 u 1/2 √ 2 π w erf  2b −bw      bwa2(τ, w) − u bwa2(τ, w) + u × erf  q  + erf  q  2 2 2 2 a(τ, w) 3 hr i(τ) a(τ, w) 3 hr i(τ) (8)

and:  4hr2i(τ)1/2 a(τ, w) = 1 + . (9) 3w2 Figure 1: (a) Simple diffusion ACFs (analytical form), normal- ized by amplitude at shortest lag time (black line: τmin = 0, orange Although there is no simple analytical expression for T symbols: τmin = τD/500, purple symbols: τmin = τD/50, green G (τ,w,b), Eq. 8 can be integrated numerically. The trun- symbols: τmin = τD/5). (b) Apparent MSDs obtained by inversion cated ACF (Eq. 7) becomes indistinguishable from the ACF of the ACFs shown in (a), showing a departure from the real MSD obtained in the absence of truncation (Eq. 1) for b ≥ 1 at lag times close to τmin. Schematic representations of the detection (Fig. 2). Since in our simulations b = 7.5, we consider in volume illustrate particle displacements shorter and larger than the the following that GT (τ,w,b) ' G(τ, w). diffraction limit, for lag times below and above τD, respectively. (c) Apparent and actual MSDs after normalization by 6Dτ. 2.6 ACF for a regular array of detection volumes To make our VLS-FCS simulations more efficient, we calcu- G(0, w) is unknown, the most straightforward solution is to lated the signal from an array of observation volumes spaced perform the inversion on G(τ, w)/G(τmin, w), where τmin is by 2bw in the focal plane and 2bSw along the optical axis the shortest lag time at which a reliable value of the ACF (Fig. 2). In this case the ACF takes the form: can be obtained (Fig. 1). Obviously, for the inversion to work A properly, one needs τmin  τ1/2. If this condition is fulfilled, G (τ,w,b) = then at short τ (when Eq. 5 is valid) the apparent MSD is: R R X X d~r d~ρ Oijk(~r)p(~ρ,τ) Olmn(~r + ~ρ) 1 i,j,k l,m,n 2  hr i(τ ) 2 . hr˜2i(τ) = hr2i(τ) × 1 − min . (6) hci R d~rO(~r) hr2i(τ) (10)

This highlights an additional necessary condition for the where Oijk(~r) is the intensity profile of the detection volume inversion procedure to work properly, which is that τ needs centred at {2bwi, 2bwj, 2bSwk}. Eq. 10 can be simplified by

Biophysical Journal 00(00) 1–12 bioRxiv preprint doi: https://doi.org/10.1101/359240; this version posted June 29, 2018. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license. 4 3 METHODS

3 Methods 3.1 Generation of particle trajectories We used a generic 3D single-agent continuum simulation procedure, written in C++, to generate particle trajectories and simulate VLS-FCS experiments with observation vol- ume varying in size between w = 0.3 and 30 µm. Particles (usually 32) were placed at random positions within a 3D rectangular simulation box with periodic boundary condi- tions. The box dimensions, a × a × 5a with a = 450 µm, were chosen to be 15 times larger than those of the largest observation volume considered. A time step ∆t = 100 ns was selected to ensure particle trajectories with sufficient resolution, even in the smallest observation volume consid- ered. Simulations were typically run for 1011 time steps. For simple diffusion, displacements at each step and in each spa- tial direction√ were drawn from a normal distribution with variance 2D∆t (diffusion coefficient D = 500 µm2/s).

Figure 2: (a) Respective sizes of the 7 different 3D-Gaussian ellip- soidal observation volumes used in our simulations. (b) Array of observation volumes. (c) ACFs expected for simple diffusion in a 3.2 Anomalous and obstructed trajectories single observation volume (black line, Eq. 3) or for a regular array The algorithm used to simulate simple Brownian diffusion of detection volumes (green symbols: b = 2.5, purple symbols: was modified to simulate 5 different diffusive processes, b = 5, orange symbols: b = 7.5, Eq. 11). The lower panel shows illustrated in Fig. 3. i) Simulations of continuous random the corresponding MSD (hr2i/(6Dτ) as obtained by inversion of walks (CTRW) were carried out following Refs. (28–30). the ACFs shown in the upper panel). Step lengths were drawn from the same Gaussian distribu- tion as for simple diffusion, but after each step a wait time τw was added, with duration drawn from a Pareto distribu- α 1+α −9 recognizing that only adjacent detection volumes are likely tion, p(τw) = ατmin/τw (τmin = 10 s, α = 0.8). To to record correlated events. Thus for a n × n × n array: allow for a range of wait times, the time step was changed for this diffusion process to 10−9s and simulations run for 1013 steps. ii) Two-component diffusion was simulated by A 3 G (τ,w,b) ' n × (G000(τ,w,b) allowing tracer particles to switch between a fast (D = 2 0 2 + 4G100(τ,w,b) + 4G110(τ,w,b) + 2G001(τ,w,b) (11) 500µm /s) and a slow diffusive state (D = 50µm /s). Transitions between states were assumed to be Poisson pro- + 8G101(τ,w,b) + 8G111(τ,w,b)), cesses with constant rates, kon = koff = 0 or kon = koff = 500 s−1. iii) Diffusing diffusivity was simulated following where Gopq denotes the correlation between two detection Chubynsky & Slater (31). The diffusion coefficient of each volumes spaced by ∆R~ = {2bwo, 2bwp, 2bSwq}: particle was allowed to undergo a 1D random walk between 2 2 Dmin = 0µm /s and Dmax = 500µm /s (with a random initial value within that range), with a ”diffusion coefficient” 2 2 2 − 4[(ob) +(pb) ] − 4(qSb) −17 4 3 a2(τ,w) a2(τ,Sw) (12) d = 5.5 × 10 m /s , chosen such that the full range Gopq(τ,w,b) = G(τ, w)e e , of possible diffusion coefficients was explored by a given particle over a time comparable to its characteristic diffusion where a2(τ, w) is given by Eq. 9. This expression can be time through one of the smaller observation volumes. iv) The derived in the same way as the ACF for two-focus FCS obstructed diffusion of particles hindered by the presence experiments (27). For an elongated observation volume (S = of fixed obstacles was simulated as previously done in 2D 5), the first three terms in Eq. 11 dominate, where the max- by Saxton (28). Fixed reflective cubic obstacles with dimen- ima for G100(τ,w,b) and G110(τ,w,b) are much higher and sions 0.75 µm were placed randomly on a cubic lattice, at a occur at shorter lag times than for the other terms. As can be volume fraction φ = 0.310 just below the limit expected, GA(τ,w,b) ' G(τ,w,b) for τ < (bw)2/(4D) (φ∗ = 0.3116). v) Caged diffusion in cubic 1µm corrals sep- (Fig. 2c). Thus in practice simulations done for an array arated by semi-permeable barriers was simulated as done in of detection volumes can be used for ACF inversion up to 2D in Refs. (5, 32). Particles could cross barriers only with 2 τ = b τ1/2 (in our case, ' 50τ1/2). probability p = 0.005 and were reflected otherwise. For the

Biophysical Journal 00(00) 1–12 bioRxiv preprint doi: https://doi.org/10.1101/359240; this version posted June 29, 2018. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license. 3.3 MSD and non-Gaussian parameter 5

Figure 3: Illustration of the different models considered, showing short two-dimensional trajectories. Successive positions are marked by black dots, while pink dots signal a wait time before taking the next step. The halos around each position represent the Gaussian distribution from which the next step is drawn. The colour of the halos represent either the value of the diffusion coefficient (a-e) or in which cage the particle is at that time (f). obstructed and caged diffusion models, the value of b was particle was calculated using Eq. 14, and individual particle changed very slightly for the smallest observation volumes, intensities were summed to give the total intensity, Fi(s∆τ). to avoid always placing their center in the same position Up to 107 steps (1 s), the simulated intensity was saved at within the cubic cells used to generate the obstacles. each step, after which only binned data (1 ms bins) was saved. At the end of the simulation, both the original and binned data were correlated using a discrete fast Fourier 3.3 MSD and non-Gaussian parameter transform, to obtain the ACF between 10−7 and 1 s, and 3 Both the second moment (hr2i, mean-squared displacement) 1 ms and 10 s, respectively. A symmetric normalization and fourth moment (hr4i) of the particle displacement were procedure was used to correct for the discrete nature of the calculated from particle trajectories using two separate sam- correlation, imposing that the numerator and denominator of pling times. For the first 107 steps, particle positions were the ACF are calculated from the same number of points (33). saved at each step (every 100 ns), and from these hr2i(τ) and hr4i(τ) calculated up to τ = 1s. After 107 steps, 3.5 ACF inversion positions were saved only every 1 ms, and hr2i(τ) and hr4i(τ) were calculated from this parsed data for τ > An apparent MSD, hr˜2i(τ), was calculated from each ACF 1 ms. The moments calculated with different sampling times using an inversion procedure justified in the case of Gaussian were stitched together to give data sets spanning the full diffusion (16, 19). The amplitude of the ACFs was first nor- explored time range, τ = 100 ns to 104 s. The non-Gaussian malized to 1 using G(∆τ,w), after which Eq. 3 was solved parameter, β, was then calculated according to its definition: for hr2i at each lag time, τ, to obtain hr˜2i(τ). d hr4i(τ) ACFs were also fitted to the general expression derived β(τ) = − 1, (13) for a Gaussian anomalous process with hr2i ∝ τ α (1, 34): d + 2 [hr2i(τ)]2 where d = 3 is the number of spatial dimensions (1). 1   τ α−1  1  τ α−1/2 G(τ) = 1 + 1 + 2 . N τD S τD (15) 3.4 Computation of the ACFs A set of 7 ACFs was calculated from the same particle tra- jectories, for 3D Gaussian detection volumes with 1/e2 radii, 4 Results wi (i = 1 to 7), ranging from 300 nm to 30 µm and equally We performed 3D single-agent simulations for different dif- spaced on a log scale. The aspect ratio was kept the same for fusive processes and characterized the obtained particle tra- all detection volumes, S = 5. In order to increase the number jectories in two different ways. First, we calculated the of particle trajectories passing through the smaller detection mean-squared displacement of the particles, hr2i(τ), the volumes, the simulation box was split into smaller boxes of non-Gaussian parameter, β(τ), and the distribution of dis- size 2bwi × 2bwi × 2bSwi, with b = 7.5, and a detection placements, P (x, τ), directly from the trajectories (repre- volume placed at the centre of each of these smaller boxes sented by black lines or symbols in Figs. 4-12). Second, (Fig. 2b). For each wi, the fluorescence intensity collected we simulated the result of VLS-FCS experiments, gener- for a particle at (x,y,z) was thus calculated as: ating ACFs for observation volumes ranging in size from

2 2 2 −2Mod(x,bwi) −2Mod(y,bwi) −2Mod(z,bSwi) w = 0.3 to 30 µm (ACFs and derived parameters are rep- w2 w2 S2w2 fi(x, y, z) = e i e i e i , resented by colored symbols in Figs. 4-12). Each ACF was (14) inverted according to the procedure suggested by Shuster- ˜2 where Mod(x, bwi) is the remainder of the division of x man et al. (16, 17), to yield an apparent MSD, hr i(τ), to be by bwi. At each simulation step, s, and for each detection compared to the actual MSD. ACFs were also fitted to the volume size, wi, the fluorescence intensity emitted by each general expression often used to assess anomalous diffusion

Biophysical Journal 00(00) 1–12 bioRxiv preprint doi: https://doi.org/10.1101/359240; this version posted June 29, 2018. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license. 6 4 RESULTS

of detection volumes spaced by 15w, a small positive corre- lation (∼ 0.1% of the ACF amplitude), resulting in a dip ˜2 2 in hr i(τ)/6τ, is observed around (15w) /4D = 225τD (see section 2.6 and Fig. 2). iii) The apparent MSD hr˜2i differs from hr2i at short lag times (τ < 50∆τ) because of the imperfect normalization of the ACFs, which were just divided by their value at ∆τ = 10−7 s (see section 2.4 and Fig. 1). Despite these limitations, considering only the inverted ACFs between 50∆τ and 10τD for each ACF, results in a apparent MSD which is equal to the actual MSD for 5 decades in time, from below 10−5 s to beyond 10−1 s.

4.2 Continuous time random walk We next simulated an anomalous diffusion model, a CTRW, where a wait time τw drawn from a power law distribu- 1+α Figure 4: Simple diffusion (D = 500 µm/s2). (a) ACFs obtained tion ∝ 1/τw (α = 0.8) was introduced between each for different detection volume sizes (coloured symbols: simula- step. CTRW processes are non-ergodic (29, 35). As a con- tions; black lines: fit of Eq. 15 to the simulated data). The same sequence, the MSD obtained by performing both a time and ACF, with lag time normalized by τ1/2, are shown in the right ensemble average is linear in time (Fig. 5). However, when panel. (b) Apparent diffusion coefficient, hr˜2i/(6τ), calculated performing only an ensemble average, the MSD is anoma- from the inversion of the ACF (solid coloured symbols) and com- lous, hr2i ∝ τ α (Fig. 6). Our simulation shows that the 2 pared with the actual diffusion coefficient, hr i/(6τ), obtained non-Gaussian parameter also depends on how averages are directly from the particle trajectories (black line). Empty symbols 2 performed: it is high (because of a large number of immobi- show the diffusion law (i.e. the value of w /(4τ1/2) as a function lized particles) and decays as a power law when performing of τ for each ACF). One of the inverted curves is also shown 1/2 a time average (Fig. 5d), but has a low and almost constant on its own and shifted upwards. (c) Apparent anomalous exponent obtained from the fit of the ACFs in (a). The dashed line is a guide value when performing only an ensemble average (Fig. 6d). for the eye showing the expected α = 1 value for simple diffu- To check whether ergodicity can also be detected from FCS sion. (d) Non-gaussian parameter calculated from the trajectories experiments, we calculated ACFs in two ways: first in the (black line) and expected β = 0 value for simple diffusion (dashed usual way (performing a time average on the signal, which line). (e) Distributions of displacements for τ = 10−6s (left) and already represents an ensemble average over the different τ = 10−2s (right). Lines are fit to a Gaussian distribution. particles present in the observation volume, Fig. 5), and sec- ond by averaging over many different repeats of the experi- ment instead of over time (ensemble average, Fig. 6, as also done in Ref. (35)). Whereas ACFs calculated in these two processes, Eq. 15, in order to construct the diffusion law and different ways are similar for ergodic processes, we see a recover an apparent anomalous exponent, α. clear difference for the CTRW process. The time-averaged ACFs never stabilize (their shape depends on the length of 4.1 Simple Brownian diffusion the measurement), a reflection of the ageing of the sample (36). Their shape also depends on the size of the observation We first considered simple Brownian diffusion, to check volume (with α increasing from 0.5 to 1 as w increases). The how faithfully our simulations could reproduce analytical apparent MSD extracted by inversion of these ACFs differs results (Fig. 4). As calculated from particle trajectories, the from the actual linear MSD and from one another (consis- 2 expected hr i = 6Dτ and β = 0 (black lines in Fig. 4b tent with the fact that the process is strongly non-Gaussian), and d) are obtained with good precision up to τ = 10 s. with hr˜2i ∝ t0.4 over a large range of lag times. Of note, The ACFs obtained from the simulated VLS-FCS data have however, apparent and actual MSD coincide when τ  τ1/2 the expected self-similar form, with a normal diffusion law (a regime that can be observed only for the larger detection 2 (w ∝ τ1/2, Fig. 4b) and α ' 1 (Fig. 4c). Performing volumes), as expected from the calculations presented in sec- the ACF inversion in this simple Gaussian case, where we tion 2.3. In contrast, the ensemble-average ACFs (Fig. 6) are ˜2 2 should have hr i = hr i at all τ, shows that three separate well-behaved and self-similar, with an apparent anomalous factors affect the quality of hr˜2i. i) Due to the finite simula- exponent α ' 0.7 close to the actual α = 0.8. Accordingly, tion time, for τ > 1 s the ACFs (and therefore hr˜2i) become the inversion of the ensemble-averaged ACFs gives apparent noisy. ii) Because simulations were done for regular arrays MSD that largely capture the power-law dependence of the

Biophysical Journal 00(00) 1–12 bioRxiv preprint doi: https://doi.org/10.1101/359240; this version posted June 29, 2018. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license. 4.3 Two-component diffusion 7

Figure 5: CTRW, where the ACFs (a), MSD (b), non-Gaussian Figure 6: CTRW, where the ACFs (a), MSD (b), non-Gaussian parameter (d) and distribution of displacements (e) have been cal- parameter (d) and distribution of displacements (e) have been cal- culated in the same way as for all the other models, that is taking culated using only an ensemble average over particles. Panels are both a time and an ensemble average over time and particles. Panels as in Fig. 4. are as in Fig. 4.

with the process is non-Gaussian (it is the sum of two Gaus- actual MSD (except for those curves with very short τ1/2 for sians). The non-Gaussian parameter has a constant value, 0 2 2 which the normalization does not work well). β = f(1 − f)(D − D ) /Davg. The ACFs (a weighted sum Our simulations confirm the result from previous 2D of the ACFs that would be obtained from either population simulations, which had shown that for a CTRW, the anoma- of tracers) are self-similar, with an apparent α ' 0.8 in the lous exponent recovered from FCS experiments using Eq. 15 conditions of our simulation. Because of the non-Gaussian can be significantly different than the real anomalous expo- nature of the process, hr˜2i(τ) coincides with hr2i(τ) only ˜2 nent (29, 37). In addition, our simulations emphasize that the for τ  τ1/2. For τ  τ1/2, all the hr i(τ) instead apparent α depends strongly on the size of the observation approach a simple diffusion MSD with diffusion coefficient 0 3/2 03/2 2/3 volume (Fig. 5c), a phenomenon directly linked to the scale- D∞ = DD /((1 − f)D + fD ) . The diffusion dependence of the non-Gaussian parameter (Fig. 5d). Calcu- law is linear in time, with an apparent diffusion coefficient lating the ACF as an ensemble-average only (as first done in comprised between Davg and D∞. Ref. (35)), although not necessarily easy to achieve exper- We then considered the general case, where particles can imentally, leads to much better behaved results, with sta- switch back and forth between two simple diffusion modes −1 ble self-similar ACFs, and an apparent anomalous exponent with Poisson statistics (kon = koff = 500 s ). In this approaching the actual one. scenario (one of the few considered before in the context of VLS-FCS (38, 39)), we expect a cross-over around the relaxation time, τc = 1/(kon + koff). In the ”fast diffusion” 4.3 Two-component diffusion regime (τ  τc) tracers remains in the same state while We next examined three cases in which tracer particles could crossing the observation volume and behave as two sepa- rate populations. In the ”fast reaction” regime (τ  τ ) switch (with constant rates kon and koff) between two modes c with diffusion coefficients D and D0. tracers switch state many times while crossing the observa- The simplest case is that of two separate populations of tion volume, and appear to be undergoing simple diffusion tracers (fractions f and f 0 = 1 − f) that are not allowed to with diffusion coefficient Davg. Meanwhile, β passes from 0 2 2 interchange (k = k = 0). Although a simple analytical f(1−f)(D−D ) /Davg to 0 around τc, and α becomes sub- on off p solution exists in this case, we still performed a simulation, diffusive below wc = 4Davgτc (Fig. 8). As in the previous whose results are shown in Fig. 7. For such a process the case, the actual MSD is linear in time at all lag times. How- 2 MSD is linear in time, hr i = 4Davgτ at all lag times with ever, this time the propagator is Gaussian for τ > τc, thus 0 ˜2 2 Davg = fD + (1 − f)D . Yet the propagator associated hr i coincides with hr i both below τ1/2 and above τc, and

Biophysical Journal 00(00) 1–12 bioRxiv preprint doi: https://doi.org/10.1101/359240; this version posted June 29, 2018. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license. 8 4 RESULTS

Figure 7: Two-component diffusion for separate populations of Figure 9: Two-component diffusion with D = 500 µm2/s and 2 0 2 0 2 tracers (kon = koff = 0, D = 500 µm /s, D = 50 µm /s, D = 0 µm /s (stick-and-diffuse). The transition rates were kon = 2 −1 2 f = 0.5). In this case, one expects Davg = 275 µm /s, D∞ = koff = 500 s , thus f = 0.5. Then Davg = 250 µm /s, 2 2 78 µm /s and β = 0.67. D∞ = 0 µm /s, and β = 1 at small τ. The characteristic time is τc = 1 ms, corresponding to wc ' 1 µm.

immobilization times is exponential. Its VLS-FCS signature is no different from that of the general case, with a crossover from anomalous and non-Gaussian to Gaussian around the crossover time τc = 1 ms (Fig. 9), but with higher non- Gaussianity (β = (1 − f)/f at small τ), lower Davg = fD and D∞ = 0.

4.4 Diffusing diffusivity As another example of a diffusive process with linear MSD but non-Gaussian propagator, we considered the diffusive diffusivity model (31). Particles were given a diffusion coef- ficient which varied in time according to a 1D random walk (with diffusion coefficient d = 5.5 × 107 µm4/s3) 2 2 between Dmax = 500 µm /s and Dmin = 0 µm /s. We then Figure 8: Two-component diffusion with D = 500µm2/s, expect - and observe - a crossover around τc = (Dmax − D0 = 50µm2/s k = k = 500 s−1 , and on off . In these condi- D )2/(2d) = 2 ms. The MSD is linear at all lag time, with tions, a crossover between non-Gaussian and Gaussian diffusion min apparent diffusion coefficient Dapp = (Dmax − Dmin)/2, and is expected around τc = 1 ms and wc ' 1 µm. a switch from non-Gaussian (β ' 0.35) to Gaussian (β = 0) behaviour around τc (Fig. 10). Below τc, the distribution of displacements approaches an exponential distribution, as tends towards D∞ in between. Notably, for large w > wc previously noted by Chubynsky & Slater (31). Accordingly, ˜2 2 (i.e. τ1/2 > τc), hr i = hr i at all lag times. the ACFs display an anomalous shape for w < wc ' 1 µm, Finally, we considered the limiting case in which tracers with α approaching 0.9 at the smallest detection volumes. transiently experience immobilization (D0 = 0). The signa- As for the two-component models, hr˜2i coincide with hr2i ture of this ”stick-and-diffuse” model has been considered both for τ  τ1/2 and τ  τc. Overall, the signature of dif- in the case of single-scale FCS experiments (40). It dif- fusing diffusivity resembles that of a two-component model, fers significantly from a CTRW because the distribution of but with less marked anomalous features.

Biophysical Journal 00(00) 1–12 bioRxiv preprint doi: https://doi.org/10.1101/359240; this version posted June 29, 2018. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license. 4.5 Obstructed diffusion 9

Figure 11: Obstructed diffusion, with D = 500 µm2/s, and cubic L = 0.75 µ φ = 0.31 Figure 10: Diffusing diffusivity, with a diffusion coefficient dif- obstacles of size m at a volume fraction . 2 2 fusing between Dmin = 0 µm /s and Dmax = 500 µm /s with a diffusion coefficient d = 5.5 × 107 µm4/s3. 1 µm, and a probability p = 0.005 to cross the barri- ers between cages when encountering them (Fig. 12). We observe the expected crossover between short-term and long- 4.5 Obstructed diffusion term diffusion coefficient, with a MSD that more or less Another model often invoked to account for anomalous dif- follows the approximate 2D analytical expression derived by fusion in complex media is obstructed diffusion, where the Destainville et al. (i.e. a switch from microscopic to effec- motion of the tracers is restricted by the presence of immo- tive diffusion coefficient around the relaxation time of the bile obstacles. If the obstacle concentration is below the particles in the cages) (32, 44, 46). Interestingly, the non- percolation limit, anomalous diffusion occurs as a transient Gaussian parameter displays two small amplitude peaks. The 2 regime between unhindered short-scale diffusion and large- first, found below τc = (L) /6D = 0.3 ms, corresponds scale effective diffusion. We simulated obstructed diffusion to the particles equilibrating in a cage. The second corre- using randomly placed cubic obstacles of size L at a volume sponds to the particles leaving the cage. Yet as β remains fraction φ slightly below the percolation limit (φ∗ = 0.3116 small, hr˜2i ' hr2i at all lag times (the strongest deviation is in this geometry). As seen before in 2D simulations (22, 41), observed just above τc). ACFs with w ' L reflect the imper- the apparent diffusion coefficient of the tracers switches fect confinement of the particles in the cage by displaying from their actual diffusion coefficient (D = 500 µm2/s) two separate timescales, and therefore α < 1. 2 to an effective value (D∞ = 360 µm /s in our condi- An interesting particular case is that of impermeable 2 2/3 tions) around τc = L /(6Dφ ) ' 0.4 ms (Fig. 11). This cages (p = 0). In practice, FCS experiments would be dif- transient anomalous regime is visible in the ACFs around ficult to perform for such a system, as confined particles 1/3 wc = L/φ ' 1 µm, where α has a minimum. Just above photobleach rapidly, yet it is interesting to think about the τc, a small peak is observed for the non-Gaussianity fac- signature of such processes. In this case, the apparent diffu- tor (a similarly weak non-Gaussianity was shown for 2D sion coefficient falls all the way to 0 above τc, and instead of simulation of obstructed diffusion at large lag time (42)). a second peak, β goes to negative values (data not shown). Accordingly, hr˜2i deviates very slightly from hr2i in that The distribution of displacements has a hard limit at x = L, region, for curves for which τ1/2 is around or below τc. eventually assuming a triangular shape at large τ. As w increases, the ACFs rapidly assume a shape reflecting con- finement, with α > 1 and with a characteristic decay time 4.6 Caged diffusion τ1/2 that no longer depends on w. The last considered model was caged diffusion, where tracer particles diffuse through an array of semi-permeable cages. 5 Discussion As for obstructed diffusion, a transitory anomalous regime visible in the ACFs is expected (43–45). We simulated caged Each of the models considered in this work is representa- diffusion for regularly arranged cubic cages of size L = tive of a class of diffusion processes relevant to the cellular

Biophysical Journal 00(00) 1–12 bioRxiv preprint doi: https://doi.org/10.1101/359240; this version posted June 29, 2018. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license. 10 5 DISCUSSION

introduced by Krichevsky et al. to obtain the MSD from the ACF (see section 3.5) (16, 17), and its combination with lengthscale variation as first considered by Hofling¨ and Fra- nosch (22), and later experimentally demonstrated by us (20), is so powerful. It uses the full range of information contained in the ACFs to allow a precise characterization of the diffusion process, over many decade in times, in a model-independent manner. By using the same simulation framework to study differ- ent classes of diffusion models and by simulating VLS-FCS experiments, we provide here a library of inverted ACFs, and thus an unbiased way to interpret the results of such experiments. Although comparison of different anomalous diffusion models using simulations are available (notably a comparison of fractional and CTRW (29)), none has ever been examined from the point of view of inverted VLS-FCS. All the considered models (except Figure 12: Caged diffusion (small escape probability), with D = for simple diffusion) resulted in ACFs with ”anomalous” 2 500 µm /s, L = 1 µm and p = 0.005. behaviour, i.e. with α < 1 for at least some of the observation volume sizes considered. However, they var- ied greatly with respect to several essential features that environment. CTRW and two-component diffusion are mod- can be accessed through a VLS-FCS experiment, namely: els of diffusing molecules interacting with slow or immobile 1) self-similarity vs. presence of a characteristic timescale binding partners, while diffusing diffusivity, obstructed dif- (visible as a crossover in the inverted ACFs), 2) linearity fusion and caged diffusion reflect different crowding scenar- of the MSD, 3) Gaussianity of the distribution of displace- ios. In cells, crowding and molecular interactions both play ment. All these models are non-Gaussian, an important char- a part in protein mobility. Our goal was thus to determine acteristic observed in many biological systems (51), e.g. how well the influence of these different scenarios could be RNA-protein particle diffusion in cells (52). Of particular distinguished in VLS-FCS experiments. interest, the models involving interchange between differ- Until recently, the information contained in VLS-FCS ent diffusion modes with Poisson statistics (two-component experiments has been exploited via the dependence of τ1/2 models) or via a continuous diffusive process (diffusing dif- on w2 (diffusion law), initially introduced to distinguish fusivity) exhibit an ”anomalous, yet Brownian” behaviour 2 simple diffusive processes (τ1/2 ∝ w ) from photophys- (non-Gaussian propagators associated with a ”normal” lin- ical processes (τ1/2 independent of w) (47). This model- ear MSD), a feature observed in many biologically relevant independent approach has also proved useful to help distin- contexts, for example diffusion in crowded polymer solu- guishing between different types of diffusion (5, 7, 13, 48). tions (20) or actin networks (53). This linear behaviour of For example, in 2D, caged diffusion and dynamic partition- the MSD can be observed in inverted VLS-FCS, whereas the ing into domains result in negative and positive τ = 0 inter- diffusion law has an apparent power-law dependence on time cepts of the diffusion law, respectively (13, 43). However, (Figs. 8-10). our 3D VLS-FCS simulations illustrate an important limita- It has been pointed out that a limitation of the inversion tion of the diffusion law, which is that it coincides with the procedure (the fact that it faithfully returns the MSD only MSD only for Gaussian or near-Gaussian processes. In more for Gaussian processes), could be turned to an advantage in complex cases (CTRW, two-component diffusion), the diffu- the case of VLS-FCS, as it can be used as a test for Gaus- sion law relates to the MSD in a non-trivial and ill-defined sianity (20, 23). Indeed, for all the models considered here, way, thus one must make assumptions about the underlying at those lag times for which the process is not Gaussian (i.e. diffusion process to extract information from it (a problem where β(τ) 6= 0), we observe a spread in the value of the already pointed out in the context of imaging FCS (49)). inverted ACFs obtained for different w (i.e. hr˜2i depends on Ultimately, the issue with the diffusion law is that it col- w). Moreover, the further β(τ) is from 0, the further the fam- lapses the rich information contained in the shape of the ACF ily of inverted ACFs deviate from one another. More than a into a single value, τ1/2. In contrast, analysis of the detailed simple test for Gaussianity, an inverted VLS-FCS can thus shape of the ACFs obtained at different w can give a lot inform on the range of lag times over which the process is of information about the underlying process (as shown for non-Gaussian, and on how far from Gaussian the process is. obstructed diffusion (22), two-component diffusion (38, 39) Maybe the most interesting result from our study is that, or diffusion in phase separated membranes (50)). However, regardless of Gaussianity, the inversion procedure based on it is model-dependent. This is why the inversion procedure

Biophysical Journal 00(00) 1–12 bioRxiv preprint doi: https://doi.org/10.1101/359240; this version posted June 29, 2018. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license. 11

Eq. 3 returns the actual MSD if τ  τ1/2. This can be proven 2002. Biological and chemical applications of fluorescence by performing a Taylor expansion of the ACF in hr2i/w2 (as correlation spectroscopy: a review. Biochemistry 41:697–705. shown in section 2.3), and can be seen in for all the models 4. Rigler, R., and E. S. Elson, 2012. Fluorescence correlation simulated here. The linear relationship existing between the spectroscopy: theory and applications, volume 65. Springer ACF and the MSD at short lag time has been noted before Science & Business Media. (24). However, we show here that this linear regime can be 5. Wawrezinieck, L., H. Rigneault, D. Marguet, and P.-F. Lenne, 2005. Fluorescence correlation spectroscopy diffusion laws to greatly extended by increasing w. Performing a single-point probe the submicron organization. Biophys. J. FCS experiment for a large detection volume, will allow 89:4029–4042. recovering the actual MSD from ≈ 20∆τ (where ∆τ is the 6. Masuda, A., K. Ushida, and T. Okamoto, 2005. New flu- experiment’s time resolution) up to ' τ1/2/10, whether the orescence correlation spectroscopy enabling direct observa- underlying process is Gaussian or not. Of course, this will tion of spatiotemporal dependence of diffusion constants as hold true only if good statistics can be achieved (i.e. the stud- an evidence of anomalous transport in extracellular matrices. ied system is stable enough to allow long measurements) and Biophys. J. 88:3584–3591. if the photophysics of the dye used to tag the tracer is not 7. Humpol´ıckovˇ a,´ J., E. Gielen, A. Benda, V. Fagulova, J. Ver- an issue at short τ. The MSD can then be retrieved for all cammen, M. VandeVen, M. Hof, M. Ameloot, and Y. Engel- kinds of diffusive processes, over several orders of magni- borghs, 2006. Probing diffusion laws within cellular tude in time (something which we have previously observed membranes by Z-scan fluorescence correlation spectroscopy. Biophys. J. 91:L23–L25. for fluorescent beads diffusing in gels (20)). Remarkably, 8. Kastrup, L., H. Blom, C. Eggeling, and S. W. Hell, 2005. the MSD can then extend below the diffraction limit (a point Fluorescence fluctuation spectroscopy in subdiffraction focal which has been made before eloquently (14, 54)). Thus, in volumes. Phys. Rev. Lett. 94:178104. FCS experiments, the best strategy to retrieve information 9. Samiee, K., J. Moran-Mirabal, Y. Cheung, and H. Craig- on processes at short time scale may not be to physically head, 2006. Zero mode waveguides for single-molecule achieve subdiffraction-limited detection volume, but instead spectroscopy on lipid membranes. Biophys. J. 90:3288–3299. to work on achieving high quality PSF for larger detection 10. Vobornik, D., D. S. Banks, Z. Lu, C. Fradin, R. Taylor, and volumes with a shape as close as possible to a 3D Gaussian L. J. Johnston, 2008. Fluorescence correlation spectroscopy profile (ensuring the inversion procedure faithfully returns with sub-diffraction-limited resolution using near-field optical the MSD at short lag time). probes. Appl. Phys. Lett. 93:163904. 11. Bag, N., D. H. X. Yap, and T. Wohland, 2014. Tem- perature dependence of diffusion in model and live cell 6 Author Contributions membranes characterized by imaging fluorescence correlation spectroscopy. BBA-Biomembranes 1838:802–813. MS designed and performed simulations, helped write the 12. Krieger, J. W., A. P. Singh, N. Bag, C. S. Garbe, T. E. Saunders, manuscript and prepared figures. CF helped design simula- J. Langowski, and T. Wohland, 2015. Imaging fluorescence tions and wrote the manuscript. (cross-) correlation spectroscopy in live cells and organisms. Nat. Protoc. 10:1948–1974. 13. Ruprecht, V., S. Wieser, D. Marguet, and G. J. Schutz,¨ 7 Acknowledgments 2011. Spot variation fluorescence correlation spectroscopy allows for superresolution chronoscopy of confinement times This article is dedicated to Jorg¨ Langowski, whose encour- in membranes. Biophys. J. 100:2839–2845. agements have been an inspiration for our work, and who 14. Krichevsky, O., 2013. Comment on ”Polymer dynamics, flu- was very passionate about the capacity of FCS to resolve orescence correlation spectroscopy, and the limits of optical subdiffraction motions. We also thank Felix Hofling¨ for resolution”. Phys. Rev. Lett. 110:159801. many helpful discussions. This work was funded by the Nat- 15. Di Rienzo, C., E. Gratton, F. Beltram, and F. Cardarelli, 2013. ural Sciences and Engineering Research Council of Canada Fast spatiotemporal correlation spectroscopy to determine pro- (NSERC), and enabled by the use of computing resources tein lateral diffusion laws in live cell membranes. Proc. Nat. provided by SHARCNET and Compute Canada. Acad. Sci. USA 110:12307–12312. 16. Shusterman, R., S. Alon, T. Gavrinyov, and O. Krichevsky, 2004. Monomer dynamics in double-and single-stranded DNA References polymers. Phys. Rev. Lett. 92:048303. 17. Shusterman, R., T. Gavrinyov, and O. Krichevsky, 2008. 1. Hofling,¨ F., and T. Franosch, 2013. Anomalous transport Internal Dynamics of Superhelical DNA. Phys. Rev. Lett. in the crowded world of biological cells. Rep. Prog. Phys. 100:098102. 76:046602. 18. Wocjan, T., J. Krieger, O. Krichevsky, and J. Langowski, 2009. 2. Wachsmuth, M., W. Waldeck, and J. Langowski, 2000. Dynamics of a fluorophore attached to superhelical DNA: FCS Anomalous diffusion of fluorescent probes inside living cell experiments simulated by Brownian dynamics. Phys. Chem. nuclei investigated by spatially-resolved fluorescence correla- Chem. Phys. 11:10671–10681. tion spectroscopy. J. Mol. Biol. 298:677–689. 19. Horton, M. R., F. Hofling,¨ J. O. Radler,¨ and T. Franosch, 3. Hess, S. T., S. Huang, A. A. Heikal, and W. W. Webb,

Biophysical Journal 00(00) 1–12 bioRxiv preprint doi: https://doi.org/10.1101/359240; this version posted June 29, 2018. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license. 12 7 ACKNOWLEDGMENTS

2010. Development of anomalous diffusion among crowding Anomalous subdiffusion is a measure for cytoplasmic crowd- proteins. Soft Matter 6:2648–2656. ing in living cells. Biophys. J. 87:3518–3524. 20. Banks, D. S., C. Tressler, R. D. Peters, F. Hofling,¨ and 38. Vagias, A., R. Raccis, K. Koynov, U. Jonas, H.-J. Butt, C. Fradin, 2016. Characterizing anomalous diffusion in G. Fytas, P. Kosovan,ˇ O. Lenz, and C. Holm, 2013. Complex crowded polymer solutions and gels over five decades in tracer diffusion dynamics in polymer solutions. Phys. Rev. time with variable-lengthscale fluorescence correlation spec- Lett. 111:088301. troscopy. Soft matter 12:4190–4203. 39. Ipina,˜ E. P., and S. P. Dawson, 2013. From free to effective 21. Di Rienzo, C., V. Piazza, E. Gratton, F. Beltram, and F. Car- diffusion coefficients in fluorescence correlation spectroscopy darelli, 2014. Probing short-range protein Brownian motion in experiments. Phys. Rev. E 87:022706. the of living cells. Nat. Commun. 111:679–685. 40. Yeung, C., M. Shtrahman, and X.-l. Wu, 2007. Stick-and- 22. Hofling,¨ F., K.-U. Bamberg, and T. Franosch, 2011. Anoma- diffuse and caged diffusion: a comparison of two models of lous transport resolved in space and time by fluorescence synaptic vesicle dynamics. Biophys. J. 92:2271–2280. correlation spectroscopy. Soft Matter 7:1358–1363. 41. Saxton, M. J., 1994. Anomalous diffusion due to obstacles: a 23. Kubecka,ˇ J., F. Uhl´ık, and P. Kosovan,ˇ 2016. Mean squared Monte Carlo study. Biophys. J. 66:394–401. displacement from fluorescence correlation spectroscopy. Soft 42. Ernst, D., J. Kohler,¨ and M. Weiss, 2014. Probing the type Matter 12:3760–3769. of anomalous diffusion with single-particle tracking. Phys. 24. Khadem, S., C. Hille, H.-G. Lohmannsr¨ oben,¨ and I. Sokolov, Chem. Chem. Phys. 16:7686–7691. 2016. What information is contained in the fluorescence 43. Lenne, P.-F., L. Wawrezinieck, F. Conchonaud, O. Wurtz, correlation spectroscopy curves, and where. Phys. Rev. E A. Boned, X.-J. Guo, H. Rigneault, H.-T. He, and D. Mar- 94:022407. guet, 2006. Dynamic molecular confinement in the plasma 25. Gennerich, A., and D. Schild, 2000. Fluorescence correlation membrane by microdomains and the cytoskeleton meshwork. spectroscopy in small cytosolic compartments depends criti- EMBO J. 25:3245–3256. cally on the diffusion model used. Biophys. J. 79:3294–3306. 44. Destainville, N., 2008. Theory of fluorescence correlation 26. Fradin, C., A. Abu-Arish, R. Granek, and M. Elbaum, 2003. spectroscopy at variable observation area for two-dimensional Fluorescence correlation spectroscopy close to a fluctuating diffusion on a meshgrid. Soft Matter 4:1288–1301. membrane. Biophysical journal 84:2005–2020. 45. Enderlein, J., 2012. Polymer dynamics, fluorescence correla- 27. Dertinger, T., V. Pacheco, I. von der Hocht, R. Hartmann, tion spectroscopy, and the limits of optical resolution. Phys. I. Gregor, and J. Enderlein, 2007. Two-focus fluorescence cor- Rev. Lett. 108:108101. relation spectroscopy: A new tool for accurate and absolute 46. Destainville, N., A. Sauliere,` and L. Salome,´ 2008. Comment diffusion measurements. ChemPhysChem 8:433–443. to the article by Michael J. Saxton: A biological interpreta- 28. Saxton, M. J., 1996. Anomalous diffusion due to binding: a tion of transient anomalous subdiffusion. i. qualitative model. Monte Carlo study. Biophys. J. 70:1250–1262. Biophys. J. 95:3117. 29. Szymanski, J., and M. Weiss, 2009. Elucidating the origin 47. Haupts, U., S. Maiti, P. Schwille, and W. W. Webb, 1998. of anomalous diffusion in crowded fluids. Phys. Rev. Lett. Dynamics of fluorescence fluctuations in green fluorescent 103:038102. protein observed by fluorescence correlation spectroscopy. 30. Klafter, J., and I. M. Sokolov, 2011. First steps in random Proc. Nat. Acad. Sci. USA 95:13573–13578. walks: from tools to applications. Oxford University Press. 48. Eggeling, C., C. Ringemann, R. Medda, G. Schwarzmann, 31. Chubynsky, M. V., and G. W. Slater, 2014. Diffusing Diffusiv- K. Sandhoff, S. Polyakova, V. N. Belov, B. Hein, C. von Mid- ity: A Model for Anomalous, yet Brownian, Diffusion. Phys. dendorff, A. Schonle,¨ et al., 2008. Direct observation of the Rev. Lett. 113:098302. nanoscale dynamics of membrane lipids in a living cell. Nature 32. Meilhac, N., L. Le Guyader, L. Salome, and N. Destainville, 457:1159–1162. 2006. Detection of confinement and jumps in single-molecule 49. Veerapathiran, S., and T. Wohland, 2017. The imaging FCS membrane trajectories. Phys. Rev. E 73:011915. diffusion law in the presence of multiple diffusive modes. 33. Shi, X., and T. Wohland, 2010. Fluorescence correlation spec- Methods 140-141:140–150. troscopy. In A. Diaspro, editor, Nanoscopy and multidimen- 50. Burns, M. C., M. Nouri, and S. L. Veatch, 2016. Spot size sional optical fluorescence microscopy, CRC Press. variation FCS in simulations of the 2D Ising model. J. Phys. 34. Schwille, P., J. Korlach, and W. W. Webb, 1999. Fluorescence D Appl. Phys. 49:214001. correlation spectroscopy with single-molecule sensitivity on 51. Metzler, R., 2017. Gaussianity fair: The riddle of anomalous cell and model membranes. Cytometry 36:176–182. yet non-Gaussian diffusion. Biophys. J 112:413. 35. Lubelski, A., and J. Klafter, 2009. Fluorescence correla- 52. Lampo, T. J., S. Stylianidou, M. P. Backlund, P. A. Wiggins, tion spectroscopy: the case of subdiffusion. Biophys. J. and A. J. Spakowitz, 2017. Cytoplasmic RNA-protein parti- 96:2055–2063. cles exhibit non-Gaussian subdiffusive behavior. Biophys. J. 36. Metzler, R., J.-H. Jeon, A. G. Cherstvy, and E. Barkai, 112:532–542. 2014. Anomalous diffusion models and their properties: non- 53. Wang, B., S. M. Anthony, S. C. Bae, and S. Granick, 2009. stationarity, non-ergodicity, and ageing at the centenary of Anomalous yet Brownian. Proc. Nat. Acad. Sci. USA single particle tracking. Phys. Chem. Chem. Phys. 16:24128– 106:15160–15164. 24164. 54. Di Rienzo, C., E. Gratton, F. Beltram, and F. Cardarelli, 2016. 37. Weiss, M., M. Elsner, F. Kartberg, and T. Nilsson, 2004. Spatiotemporal fluctuation analysis: A powerful tool for the future nanoscopy of molecular processes. Biophys. J. 5:1–8.

Biophysical Journal 00(00) 1–12