Apparent anomalous diffusion and non-Gaussian distributions in a simple mobile–immobile transport model with Poissonian switching
Abstract
We analyse mobile–immobile transport of particles that switch between the mobile and immobile phases with finite rates. Despite this seemingly simple assumption of Poissonian switching, we unveil a rich transport dynamics including significant transient anomalous diffusion and non-Gaussian displacement distributions. Our discussion is based on experimental parameters for tau proteins in neuronal cells, but the results obtained here are expected to be of relevance for a broad class of processes in complex systems. Specifically, we obtain that, when the mean binding time is significantly longer than the mean mobile time, transient anomalous diffusion is observed at short and intermediate time scales, with a strong dependence on the fraction of initially mobile and immobile particles. We unveil a Laplace distribution of particle displacements at relevant intermediate time scales. For any initial fraction of mobile particles, the respective mean squared displacement (MSD) displays a plateau. Moreover, we demonstrate a short-time cubic time dependence of the MSD for immobile tracers when initially all particles are immobile.
1. Introduction
Already in the 1960s, there was considerable interest in the transport of chemical tracers, especially pesticides, nitrates and heavy metals through water-carrying layers of soil [1]. A typical description for such contaminant transport was the diffusion–advection equation (sometimes called the convective–dispersive equation) [2]
Motivated by concrete biological examples, we here study a seemingly simple version of the mobile–immobile transport model, in which particles switch between a freely diffusive phase and an immobile, stagnant phase. Even for the Poissonian switching dynamics considered here between the mobile and immobile phases and for biologically relevant parameters, we demonstrate the existence of a significant, transient anomalous–diffusive regime with a distinct non-Gaussian displacement distribution.
In fact, various components of biological cells, including tau proteins, synaptic vesicles in hippocampal neurons, glucocorticord receptors, calcium-sensing proteins and transcription factors at the junction of the endoplasmic reticulum and the plasma membrane, undergo diffusion with transient immobilization [17–24]. Another example is given by the diffusion and target search of DNA-binding proteins along DNA molecules. For instance, the core domain of the tumour suppressor p53, the damage detection complex Rad4-Rad23 and the architectural DNA-binding protein Fis repeatedly attach to and detach from the DNA during the target search [25–29]. We here focus on tau proteins, which transiently bind to microtubules in axons of neuronal cells and are immobilized in the bound state, as schematically depicted in figure 1. Tau proteins stabilize microtubules that give structure to cells [30]. Alzheimer’s disease is associated with tau proteins losing the ability to bind to microtubules [30,31]. This effectively destabilizes the microtubules and leads to neurodegeneration [30,31]. Owing to the extremely elongated shape of the axon, the motion of tau proteins can be effectively described in one dimension [17]. If the immobilization time follows an exponential distribution with mean τim and tracers immobilize with rate , i.e. a Poissonian dynamics, as assumed in [17], the motion can be described by the mobile–immobile model

Figure 1. Schematic of tau protein dynamics in axons of neuronal cells. Diffusing tau proteins bind to longitudinally aligned microtubules inside the axon with the rate . Upon binding, they remain immobile for the average duration τim and unbind with the rate . The green markers represent fluorescent proteins attached to the tau proteins. Because of the elongated shape of the axons, the tau protein dynamic can effectively be described in one dimension. In our model, we assume a homogeneous binding site density.
Such Fickian yet non-Gaussian diffusion has been shown to occur for the motion of colloidal beads on phospholipid bilayer tubes, molecules at surfaces and colloids in a dense matrix of micropillars, where the colloids can get trapped in pockets [33–35]. Fickian yet non-Gaussian diffusion with a finite correlation time beyond which the displacement probability density function (PDF) crosses over to a Gaussian with an effective diffusivity arises in diffusing–diffusivity models, in which the diffusivity of individual tracers varies stochastically over time [36–41]. Direct examples for such randomly evolving diffusion coefficients (mobilities) are indeed known from lipids in protein-crowded bilayer membranes [42], shape-shifting protein molecules [43] or (de)polymerizing oligomer chains [44,45]. In other systems, an intermittent plateau emerges in the MSD; for instance, for two-dimensional fluids confined in a random matrix of obstacles or a porous cavity, in which trapping in finite pockets plays a key role [46–48]. We also mention plateaus in the MSD of both two- and three-dimensional isotropic Lennard-Jones binary liquids [49]. In most of the systems mentioned here, the PDF crosses over from an exponential (Laplace) PDF to a Gaussian. In the following, we explicitly show how a Laplace distribution with fixed scale parameters arises at intermediate time scales in our mobile–immobile model, paired with transient anomalous diffusion.
In what follows, we consider three initial conditions: an equilibrium fraction of mobile tracers and a scenario in which initially all tracers are mobile or immobile. These experimentally feasible situations significantly change the diffusion at short and intermediate time scales, at which apparent anomalous diffusion arises with slow-down and plateau-like behaviour, or ballistic diffusion, respectively. Together with the transient non-Gaussian displacement PDF, this behaviour is remarkably rich, given the simplicity of the governing equation (1.2). We individually analyse the motion of the mobile and immobile population of tracers, made possible by the formulation of separate densities for mobile and immobile particles in this modelling approach. One physical incentive to do so is that the function of the tau proteins depends on their binding state [30]. Only bound tau proteins stabilize microtubules, or transcription factors modulate gene expression when bound to the DNA [21,30]. In some situations, only the mobile or immobile tracers can be measured. An example is given by combining total internal reflection fluorescence microscopy with fluorescently labelled single-stranded DNA, which binds to the microscope coverslip [50].
We present general results for the mobile and immobile concentrations and the MSD for arbitrary fractions of initially mobile tracers in §2. Sections 3–5 present concrete results and detailed discussions for different fractions of initial mobile particle concentrations; respectively, we start with the cases when all tracers are initially mobile and immobile and commence with an equilibrium fraction of mobile tracers. We conclude in §6.
2. Model and general solutions
We consider the mobile–immobile model equations (1.2) for the initial conditions nm(x, 0) = fmδ(x) and nim(x, 0) = fimδ(x), where fm and fim denote the fractions of initially mobile and immobile tracers, respectively, with the normalization fm + fim = 1. This formulation is suitable for typical single-particle tracking experiments used in biological and soft matter systems. They are also relevant for geophysical experiments, in which point-like injection of tracers is used. In this section, we keep the fractions fm and fim arbitrary and choose specific values in the following three sections.
In what follows, we use the concrete parameters D = 13.9 (μm)2 s−1, and from [17] in all figures and neglect the vanishingly small advection velocity v = 0.002 μm s−1.1 The values were obtained from experiments using the fluorescent decay after the photoactivation technique [17]. Let us briefly address the experimental origin of the time-scale separation between τm and τim. From single-particle tracking experiments of single-stranded DNA or tau proteins, immobilization times during the particle motion can be extracted [18,50]. The experiments for the tau proteins in [18] provided two-dimensional information and revealed relatively short residence times of the tau proteins on the microtubules, when compared with mobile times [18]. By contrast, the fluorescence decay after photoactivation (FDAP) experiment in one dimension along the axon direction, here denoted as the x variable, revealed long residence times and short mobile periods: τim ≈ 48τm [17]. This seeming contradiction can be resolved when examining more closely the two-dimensional trajectories in the electronic supplementary material of [18]. Namely, the microtubules inside the axon are aligned in parallel with the axon axis, as also shown in figure 1. While a single binding event is short, an unbound particle quickly rebinds to a parallel, nearby microtubule after a short distance covered by diffusion perpendicular to the axon axis. This perpendicular motion does not contribute to the one-dimensional motion in the x-direction and thus, while individual binding times are relatively short, effective binding times appear much longer in the projection to one dimension. Since we are only interested in the one-dimensional motion, we use the parameters of [17] and hence long immobilization times.
2.1. Mobile and immobile concentration profiles
We consider the Fourier–Laplace transform of the concentrations and solve for nm(k, s), nim(k, s) and ntot(k, s) in expressions (A 1) and (A 2), in which the Fourier wavenumber k corresponds to the distance x in real space and the Laplace variable s is conjugated to time t; see appendix A for details. We denote functions in Fourier or Laplace space solely by replacing the explicit dependencies on the respective arguments. The relations in the Fourier–Laplace domain can be Fourier-inverted, and we obtain the expressions in the Laplace domain,
2.2. Moments
In general, the fractions and of mobile and immobile tracers, initially fixed as fm and fim, change over time. To obtain the respective numbers, we integrate the tracer densities over space. This corresponds to setting k = 0 in the Fourier–Laplace transforms nm(k, s) and nim(k, s) of the densities. After Laplace inversion, we find
3. All tracers initially mobile
We now consider the initial condition when all tracers are mobile, i.e. nm(x, 0) = δ(x) and nim(x, 0) = 0. This initial condition does not correspond to the experiment carried out by Igaev et al. [17]. However, this situation could be realized experimentally, e.g. by using the method of injection of fluorescently labelled tau proteins [53]. In what follows, we repeatedly use the time-scale separation τm ≪ τim observed for tau proteins and also relevant to other systems.
3.1. Concentration
We calculate the densities at short, intermediate and long times. In B.1, we obtain the Gaussian

Figure 2. Concentration profiles for mobile initial conditions. The solid black line shows ntot(x, t) and the grey striped area nm(x, t), obtained via Laplace inversion of relations (2.1) and (2.3). Colours indicate the number of immobilization events of particles from a Brownian dynamics simulation with 5 × 106 trajectories in a stacked histogram. The striped area denotes mobile particles and the white dotted line denotes initially mobile tracers that have not yet been immobilized up to the indicated time t (3.2); this result almost coincides with the full concentration in the top left panel. For t = 0.5 s to 2 s, the white dashed line shows the Laplacian (3.5); for t = 50 s and 200 s, it shows the long-time Gaussian (2.4).
The concentration of freely diffusing particles that have not immobilized yet, i.e. have zero immobilization events Nim = 0, is given by the PDF of free Brownian motion multiplied by the probability of not having immobilized, i.e.
3.2. Mean squared displacement
From the general expression for the MSD (2.9) for immobile initial conditions, we obtain the expression

Figure 3. Second moments for different initial conditions on a log–log scale. In (a), all tracers are initially mobile, as in §3. After a linear growth, the second moment 〈x2(t)〉 of all tracers (equation (3.6)) shows a plateau for τm ≪ t ≪ τim. The second moment of the mobile particles (equation (3.10)) in (a) has a peak immediately before the total particle moment and the mobile particle moment reach a plateau value. Immobile tracers spread ∼Dt at short times, and the second moment (3.11) has a plateau at intermediate times. In (b), all tracers are initially immobile, as in §4. The second moment of all tracers (equation (4.3)) grows ∼Dt2/τim at short times, owing to the decaying number of particles located at x = 0. The immobile tracers spread ∼Dt3/(3τmτim) at short times, while the full expression is given in equation (4.3). The mobile tracers in (b) spread exactly like the immobile tracers in (a), where all tracers are initially mobile. (c) The equilibrium case, §5, in which the second moment grows like 2Dt/(1 + τim/τm) (equation (5.3)) for all times. The mobile and immobile moments exactly match the moments of the total distribution with mobile and immobile initial conditions, respectively.
When calculating the moments of the mobile and immobile tracers (2.7), the time-dependent normalizations of the tracer densities (2.6),
The immobile MSD (3.11) has the short-time behaviour 〈x2(t)〉im ∼ Dt for t ≪ τm, τim. The factor when compared with the mobile tracers arises because immobile tracers effectively average over the history of the mobile tracers. Namely, for t′ ≪ τm, τim, mobile particles immobilize with the constant rate p(t′) = 1/τm. A particle that immobilized at time t′ before moved for the duration t′ and thus contributes 2Dt′ to the second moment for t > t′; see figure 4a for a schematic drawing. When averaging over different mobile periods t′ and normalizing with the fraction of immobile tracers , we obtain

Figure 4. Schematic showing the short-time behaviour of tracers for mobile (a) and immobile initial conditions (b) at three snapshots of time. In both panels, the tracers change the mobilization state at times t′1 and t′2, respectively. For mobile initial conditions in (a), the number of immobile tracers grows ∼t/τm at short times. Namely, the later a tracer immobilizes, the longer it was previously mobile. In (b), the number of mobile tracers grows ∼t/τim. Namely, the earlier a tracer mobilizes in (b), the longer it is mobile.
4. All tracers initially immobile
We now discuss the case when all tracers are immobile at t = 0, nim(x, 0) = δ(x) and nm(x, 0) = 0.
4.1. Concentration
In B.1, we obtain the short-time behaviour

Figure 5. Concentration profiles for immobile initial conditions; for a description of the legend, see figure 2. The main difference from the case of mobile initial conditions poses the peak of immobile tracers at x = 0 that have not moved up to time t, as shown by the circle. In addition, we here find a pronounced relative increase of mobile particle numbers and the very slow spread of immobile tracers at short times. The short-time approximation (4.1) is shown as the black dashed line in the top left panel. For t = 0.5–10 s, the white dashed line shows the Laplacian (4.2) with growing weight; for t = 50 s and 200 s, it shows the long-time Gaussian (2.4).
4.2. Mean squared displacement
From the general expression for the MSD (2.9), we obtain the expression
5. Equilibrium initial fractions of initial mobile tracers
In this section, we use the equilibrium values and as initial conditions.
5.1. Concentration profiles
From the general expressions (2.1) and (2.3) for the densities nm(x, s) and ntot(x, s), we find that the mobile concentration of the equilibrium case discussed here is proportional to the total concentration for the mobile initial condition in §3 at all times. To understand why this is true, we note that both concentrations at all times contain mobile tracers that were initially mobile. Moreover, from equations (2.1) and (2.3), we see that the mobile concentration of the equilibrium case contains initially immobile tracers, while the total concentration contains immobile tracers that were initially mobile. In equations (2.1) and (2.3), the respective terms that appear in addition to the initially mobile fractions that are still mobile are proportional to each other at all times, as described in §2.1. An analogous relation holds between the immobile concentration with equilibrium initial conditions and the total concentration with immobile initial conditions, as can be seen in equations (2.1) and (2.3).
We consider the short-time approximation t ≪ τm, τim for which initially immobile tracers have not yet mobilized and initially mobile tracers have not yet been trapped. Therefore, we can neglect the terms with the rates and in (1.2) and solve nm(x, t) and nim(x, t) separately, yielding

Figure 6. Concentration profiles for equilibrium initial conditions. At t = 0, all tracers are at x = 0 and the equilibrium fraction τim/(τm + τim) is immobile. For a description of the legend, see figure 2. The top left panel shows the short-time behaviour consisting of a Gaussian and a δ-distribution (equation (5.1)). At t = 1, almost all initially mobile tracers immobilized at least once and the total concentration follows the Laplace distribution (5.2), as shown by the black–white striped line for t = 0.5–2 s. At longer times, after several immobilizations the concentration profiles cross over to a Gaussian, as witnessed by equation (2.4), shown as a black–white striped line for t = 50 s and t = 200 s.
5.2. Mean squared displacement
The number of mobile and immobile tracers remains constant for equilibrium initial conditions. At all times, the second moment of all tracers (2.9) thus simplifies to
In the long-time limit, all mobile and immobile second moments grow like the moments of the total concentration, i.e. 〈x2(t)〉m ∼ 〈x2(t)〉im ∼ 2Defft for t ≫ τim, τm.
6. Discussion and conclusion
We considered a quite simple mobile–immobile model according to which tracer particles switch between a mobile diffusing state and an immobilized state. On average, the tracers remain mobile for the duration τm and immobile for τim. We considered the particular case, motivated by experiments on tau proteins binding to and unbinding from microtubules in axons of dendritic cells [17], when the two time scales are separated, τm ≪ τim. We analysed three different initial conditions with varying fractions of mobile to immobile tracers at the origin, which can, in principle, all be realized in experiments. The initial condition of mobile tracers can be realized by injecting fluorescently labelled proteins [53]. Initially, immobile tracers could in principle be obtained in single-particle tracking experiments, by focusing on the tracks of immobile tracers. Equilibrium fractions of mobile tracers naturally occur when the tau proteins were in proximity to the microtubules for t ≫ τm, τim before the start of the data acquisition.
First, we studied the case when all tracers are initially mobile, as described in the experiment in [53]. Second, we assumed all tracers to be initially immobile. Third, we considered an equilibrium fraction, corresponding to the experiment in [17]. For non-equilibrium fractions of initially mobile tracers, we find anomalous diffusion at short and intermediate time scales, at which initially mobile tracers display a plateau in the MSD at intermediate times and initially immobile tracers spread ballistically at short times. At t ≪ τm and an initial equilibrium fraction, the tracer density consists of a Gaussian and a delta peak. Initially, mobile tracers follow a Gaussian distribution at short times. When all tracers are initially immobile, the short-time distribution consists of a delta peak and a non-Gaussian distribution. At intermediate times τm ≪ t ≪ τim, the distribution is made up of a Laplace distribution and a delta distribution of initially immobile tracers that have not moved yet. The coefficients of the two distributions depend on the specific initial conditions. We additionally obtain expressions for the densities that are valid for the whole range t ≪ τim. We stress that the distribution is non-Gaussian at intermediate times, regardless of the initial conditions. By contrast, the distribution asymptotically at long times matches a Gaussian for all initial conditions. The densities of mobile and immobile tracers with equilibrium initial conditions match the total tracer densities of mobile and immobile initial conditions, respectively, at all times. Moreover, the immobile tracer density from mobile initial conditions is proportional to the mobile tracer density from immobile initial conditions at all times. As a special case for equilibrium initial conditions, our model corresponds to the one-dimensional version of the model used in [32] to describe Fickian yet non-Gaussian diffusion. We find the same linear MSD for all times and obtain a closed expression for the Laplace distribution at intermediate time scales.
The model developed here is, of course, much more general. We provided the framework for any ratio of the characteristic time scales τm and τim, such that the model will be useful for scenarios ranging from geophysical experiments with Poissonian (im)mobilization statistics to molecular systems such as protein (un)binding to DNA in nanochannel set-ups. To the best of our knowledge, the transient Laplace distribution of tau proteins has not been observed yet. We now discuss possible experiments that could reveal the anomalous diffusion regime and the Laplace distribution, which depend on the time scales τm and τim. For the present analysis, we used the parameters and for tau proteins, which were obtained from an FDAP experiment [17]. FDAP experiments do not directly allow the observation of single-particle displacement densities and the moments thereof. However, a single-molecule tracking (SMT) study of tau proteins [18] with two-dimensional trajectories of length was conducted, where we expect the transient Laplace distribution to be visible in the marginal distribution, given that the sample size is large enough. From SMT experiments, the moments can be obtained, although, in [18], the moments of the distribution were not evaluated. Another example for a system with τim > τm is given by synaptic vesicles [19]. In [19], fluorescence correlation spectroscopy reveals and . In addition, glucocorticoid receptors show long immobilization events with compared with , as revealed by fluorescent recovery after a photobleaching experiment [20]. The Laplace distribution cannot be observed in this experiment, owing to the missing information on single tracers. SMT experiments of the transcription factor p53 [22] show a switched separation of time scales with and . Here, SMT allows us to measure the exponential binding time distribution corresponding to a single binding rate, as in our model. The second moment is measured for up to 0.6 s, where longer trajectories would allow for a comprehensive comparison with the moments calculated in this work. For τm > τim, the Laplace distribution does not arise. We now look at another SMT experiment in more detail. In [29], the architectural DNA-binding protein Fis was observed to have a linear MSD and a non-Gaussian displacement distribution, as depicted in figure 7. The authors of [29] fitted two Gaussians to the distribution and deduced the presence of two sliding modes of Fis on the DNA. Since the motion during the slow sliding mode is within experimental accuracy, it is plausable to assume that the non-Gaussian distribution emerges as a result of immobilization. In figure 7, we show fits with a Laplace distribution and a Gaussian distribution in a logarithmic presentation. The Laplace distribution matches the general shape with few exceptions around , while the Gaussian distribution does not capture the peak in the centre. We note that the Laplace distribution requires a single fitting parameter, compared with the two Gaussians with advection used in [29] requiring five parameters. The apparent Laplace distribution and linear MSD translate to equilibrium initial conditions in our model.
Figure 7. Displacement distribution of the Fis DNA-binding protein. The histogram depicts the data measured by Kamagata et al. [29]. The solid and dashed lines depict fits with a Laplace and a Gaussian distribution, respectively. The Laplace fit exp( − |x|/a)/2a yields . The Gaussian fit with yields , which is within the margin of error of the value of the fast diffusion coefficient reported in [29]. The data were extracted from the PDF file of [29]. Figure 8. Comparison of the Laplace inversion of nim(x, s) (expression (2.3)) and the analytic expression for nim(x, t) (equation (3.3)) that holds for t ≪ τim. Both overlap almost perfectly for t < τim = 7.7 s.
We note that (non-)exponential (im)mobilization has been studied using a Langevin equation with switching diffusivities [56,57] and the continuous time random walk framework, where the waiting time probability distribution function consists of a combination of two exponentials with different time scales [58]. It will be a topic of future research to study the effect of a drift velocity on the pre-asymptotic behaviour for different initial conditions, as well as what happens when non-exponential (im)mobilization is considered in a mobile–immobile model in connection with chemical reactions.
Data accessibility
This article has no addition data.
Authors' contributions
T.J.D.: formal analysis, investigation, methodology, validation, visualization, writing—original draft, writing—review and editing; A.V.C.: conceptualization, formal analysis, investigation, supervision, writing—review and editing; R.M.: conceptualization, funding acquisition, project administration, supervision, validation, writing—original draft, writing—review and editing. All authors gave final approval for publication and agreed to be held accountable for the work performed herein.
Conflict of interest declaration
We declare we have no competing interests.
Funding
We acknowledge funding from the German Science Foundation (DFG, grant no. ME 1535/12-1). A.V.C. acknowledges the support of the Polish National Agency for Academic Exchange (NAWA).
Footnotes
Endnotes
Appendix A. General equations
Starting with equation (1.2), we apply the Fourier–Laplace transform to the rate to obtain
Appendix B. Asymptotics calculated in Laplace space
We go from a short-time limit to a long-time limit.
B.1. Short-time limit
For t ≪ τm, τim, we obtain sτim ≫ 1 and sτim ≫ 1. This yields ϕ(s) ∼ s in this limit. With (A 1) for fm = 1 and fim = 0, we obtain the expression
B.2. Density at intermediate time scales
We now investigate the intermediate time τm ≪ t ≪ τim, corresponding to sτm ≪ 1 and sτim ≫ 1. In this regime, we have , yielding the expression
B.3. Density in the long-time limit
We obtain the long-time limit t ≫ τm, τim from ntot(k, s) (A 2) using s ≪ 1/τim, 1/τm and ϕ(s) ∼ s(1 + τim/τim). This yields the expression
B.4. Density at short to intermediate time scales
Here, we analyse the regime t ≪ τim. The case fm = 1 and fim = 0 is considered in §3. We consider the case fim = 1 and fm = 0 here. From n(x, s) (2.3), we obtain, with sτim ≫ 1 and ϕ(s) ∼ s + 1/τm,

Figure 9. All tracers initially immobile. Comparison of the exact Laplace inversion of (2.3), the short-time approximation (B 4), the intermediate-time approximation (B 7) and the short- to intermediate-time approximation (B 11).

Figure 10. Total concentration ntot(x, t) for fim = 3/10 and fm = 7/10. Expression (B 12) is shown as the blue line and the Laplace inversion of ntot(x, s) (2.3) is shown as the black line with markers. Both overlap over five decades in amplitude, for all times shown. The red marker with the grey edge at x = 0 denotes the initially immobile tracers that have not yet moved. At short times, the distribution consists of the particles at x = 0 and a Gaussian. At t = 1 s, the distribution follows a Laplace distribution (linear tails in the log-linear plot), on top of the particles at x = 0.