Skip to main content

REVIEW article

Front. Phys., 03 August 2016
Sec. Biophysics
Volume 4 - 2016 | https://doi.org/10.3389/fphy.2016.00033

Macroscopic Transport Equations in Many-Body Systems from Microscopic Exclusion Processes in Disordered Media: A Review

Marta Galanti1,2,3 Duccio Fanelli1,3* Francesco Piazza2
  • 1Dipartimento di Fisica e Astronomia and CSDC, Università degli Studi di Firenze, Firenze, Italia
  • 2Centre de Biophysique Moléculaire, Centre National de la Recherche Scientifique-UPR4301, Université d'Orléans, Orléans, France
  • 3Istituto Nazionale di Fisica Nucleare (INFN), Sezione di Firenze, Florence, Italy

Describing particle transport at the macroscopic or mesoscopic level in non-ideal environments poses fundamental theoretical challenges in domains ranging from inter and intra-cellular transport in biology to diffusion in porous media. Yet, often the nature of the constraints coming from many-body interactions or reflecting a complex and confining environment are better understood and modeled at the microscopic level. In this paper we review the subtle link between microscopic exclusion processes and the mean-field equations that ensue from them in the continuum limit. We show that in an inhomogeneous medium, i.e., when jumps are controlled by site-dependent hopping rates, one can obtain three different nonlinear advection-diffusion equations in the continuum limit, suitable for describing transport in the presence of quenched disorder and external fields, depending on the particular rule embodying site inequivalence at the microscopic level. In a situation that might be termed point-like scenario, when particles are treated as point-like objects, the effect of crowding as imposed at the microscopic level manifests in the mean-field equations only if some degree of inhomogeneity is enforced into the model. Conversely, when interacting agents are assigned a finite size, under the more realistic extended crowding framework, exclusion constraints persist in the unbiased macroscopic representation.

1. Introduction

Diffusive transport is central in many areas of physics, chemistry, biology, and soft matter [14]. However, while the mathematics of diffusive processes in dilute and simple media is fairly well developed and understood [1], many interesting and relevant diffusive processes take place in strongly non-ideal conditions. These include a wealth of different highly dense media, from non-ideal plasmas [5] to biological membranes [6], media with complex topological structures, including porous media [79] and living cells [10, 11] and strongly confining environments [3, 1216].

Crowding and confinement effects on diffusion-influenced phenomena still pose fundamental yet unanswered questions. Several computational [17] and experimental [18, 19] evidences exist of anomalous diffusion under dense crowded conditions [20]. Anomalous diffusion refers to phenomena that lead to a nonlinear growth of particles' mean square displacement. Deviation from the standard linear behavior arise, for example, when obstacles impede the motion of particles [17, 21, 22] or when distinguishable species compete for the available spatial resources [10, 2327]. These conditions are certainly met when studying the molecular mobility inside a cell [28, 29]. Cells are occupied for over 30% of their volume by membrane delimited organelles and different sorts of cytoskeletal structures. In this respect, living cells behave much like fractal or otherwise disordered systems [30, 31]. However, strong evidences also exist in favor of normal (Brownian) diffusion, crowding, and confinement resulting in this scenario in (often nontrivial) modifications of the diffusion coefficient [10, 27, 32, 33]. Moreover, as shown in Galanti et al. [34], the effect of crowding can result in crossovers between normal and anomalous diffusion, leading to different descriptive scenarios which appear to depend on the selected initial conditions and on the specific time scale of observation.

Another related issue is that of diffusion-limited reactions [35], which are ubiquitous in many domains in biology and chemistry, touching upon problems such as association, folding and stability of proteins [13, 36] and bimolecular reactions in solution [3741], including enzyme kinetics [42], but also the dynamics of active agents [43, 44]. Many theoretical studies have tackled these and related problems under different angles [13, 39, 42, 4447]. Nevertheless, a full theoretical comprehension of transport in non-ideal media remains an elusive task, Fick's law itself and the very notion of effective diffusion coefficient being questionable in a disordered medium [31].

In this paper we review the subtle link between macroscopic transport equations, such as the diffusion equation, and microscopic processes, modeling the stochastic dynamics of some agents. For a classical and comprehensive account on diffusion in disordered media the reader can refer to Haus and Kehr [48] and references therein [4956].

The purpose of our study is to contribute to the debate with a two-fold approach. One the one hand, we wish to understand in greater depth the delicate procedure of obtaining mean-field transport equations from microscopic, agent-based stochastic models. The idea is that sometimes it may prove simpler or more effective to describe a complex transport process (or a simple one occurring in a complex medium) at the microscopic level. On the contrary, it is sometimes better to deal with macroscopic equations. It is thus important to investigate how the two levels of description interface with each other. Furthermore, we will elaborate on the reasons why considerable information can be eventually lost when passing from the microscopic stochastic process to the macroscopic mean-field description, and draw attention on the distinct notions of point-like and extended crowding, this latter bearing potentially interesting applications, only partially explored in the relevant literature.

The paper is organized as follows. In Section 2 we discuss the general framework of simple exclusion processes (SEPs), which constitute the basic tool of the microscopic description, as well as the process of obtaining mean-field equations from SEPs in the continuum limit. We show that SEPs in inhomogeneous media can be constructed in apparently equivalent manners as for the site-dependence of the hopping rates. However, different advection-diffusion equations can ensue in the continuum limit, rendering this operation a rather subtle one in disordered systems [57]. Moreover, we show that the effect of crowding, enforced in the microscopic description through an excluded volume effect, is recognizable in the continuum limit only if some degree of inhomogeneity is introduced. In Section 3, we move a step forward and consider microscopic exclusion processes involving agents characterized by a finite size, as opposed to standard SEPs. In the last section we draw the conclusions and we summarize the different extent to which the crowding fine-tunes deviation from the classical picture.

2. From Microscopic Processes to Macroscopic Equations

SEPs are space-discrete, agent-based stochastic processes modeling some kind of transport according to specific rules and bound to the constraint that no two agents can ever occupy the same site. SEPs occupy a central role in non-equilibrium statistical physics [58, 59]. While the first theoretical ideas underlying such processes can be traced back to Boltzmann's works [60], SEPs were introduced and widely studied in the 70s as simplified models of one-dimensional transport for phenomena like hopping conductivity [61] and kinetics of biopolymerization [5]. Along the same lines, the asymmetric exclusion process (ASEP), originally introduced by Spitzer [62], has become a paradigm in non-equilibrium statistical physics [6366] and has now found many applications, such as the study of molecular motors [67], transport through nano-channels [68] and depolymerization of microtubules [69].

The most general SEP in one dimension is described by a stochastic jump process on a 1D lattice with inequivalent sites in the presence of a field

ni(k+1)-ni(k)=zi-1+ni-1(k)[1-ni(k)]                                + zi+1-ni+1(k)[1-ni(k)]                                - zi+ni(k)[1-ni+1(k)]                                - zi-ni(k)[1-ni-1(k)]    (1)

Equation (1) is to be regarded as the update rule for a Monte Carlo process, where ni(k) is the occupancy of site i at time t = kΔt, which can be either zero or one. The quantities zi± are variables which have the value 0 or 1 according to a random number ξi which has a uniform distribution between 0 and 1. By defining the jump probabilities qj± (j = i, i ± 1) one can formally write:

zi-1+=θ(ξi)-θ(ξi-qi-1+)zi+1-=θ(ξi-qi-1+)-θ(ξi-qi-1+-qi+1-)  zi+=θ(ξi-qi-1+-qi+1-)-θ(ξi-qi-1+-qi+1--qi+)  zi-=θ(ξi-qi-1+-qi+1--qi+)-θ(ξi-1)    (2)

where θ(·) stands for the Heaviside step function and where we are assuming that qi-1++qi+1-+qi++qi-=1. Note that the ordering of appearance of the qj± in the above expressions is arbitrary. Equations (2) entail that zj±=qj±, where 〈·〉 indicates an average over many values of ξi, for a given configuration {ni}. The above process is fully determined by the fields qi±, specifying the probability of jumping from site i to site i+1 (qi+) or to site i − 1 (qi-) in a time interval Δt.

A (discrete-time) master equation for the above SEP can be obtained by averaging over many Monte Carlo cycles performed according to rule (1)

Pi(k+1)-Pi(k)=qi-1+[Pi-1(k)-Pi,i-1(k)]                                + qi+1-[Pi+1(k)-Pi,i+1(k)]                                - qi+[Pi(k)-Pi,i+1(k)]                                - qi-[Pi(k)-Pi,i-1(k)]    (3)

where we have defined the one-body and two-body site occupancy probabilities

Pi(k)=ni(k)    (4)
Pi,i±1(k)=ni(k)ni±1(k)    (5)

Here 〈〈·〉〉 denotes averages over many independent Monte Carlo cycles performed until time kΔt, starting from the same initial condition. We emphasize that the same equation has been derived through a slightly different procedure by Richards in 1977 [61].

2.1. Mean-Field Equations

With the aim of deriving macroscopic transport equations from the microscopic stochastic process described by Equation (1), it is customary to assume a mean-field (MF) factorization,

Pi,i±1(k)ni(k)ni±1(k)=ni(k)ni±1(k)                     =Pi(k)Pi±1(k)    (6)

With the help of Equation(6), the master Equation (3) becomes

Pi(k+1)-Pi(k)=qi-1+Pi-1(k)[1-Pi(k)]                                + qi+1-Pi+1(k)[1-Pi(k)]                                - qi+Pi(k)[1-Pi+1(k)]                                - qi-Pi(k)[1-Pi-1(k)]    (7)

Nonlinear mean-field equations for exclusion process of this type have been used since the 70s to investigate one-dimensional transport in solids [70]. In fact, despite mean-field descriptions for the inhomogeneous ASEP are known to provide imperfect descriptions of certain non-equilibrium observables in one dimension, e.g., the current-density relation and critical exponents [71], continuum descriptions can be employed reliably to track the time-evolution of large-wavelength density fluctuations [34, 7278]

2.2. Site-Dependent Rates: Three Jump Processes Yield Three Transport Equations

Let a be the lattice spacing and let us define a reversal probability ϵi, such that

qi+=Qi  qi-=Qi-ϵi    (8)

The condition (Equation 8) (with ϵi > 0) amounts to considering a field introducing a bias in the positive x direction. In order to take the continuum limit lima,Δt0Pi(k)=P(x,t), we must require

lima,Δt0Qia2Δt=D(x)    (9)
lima,Δt0ϵiaΔt=v(x)    (10)

Equation (9) defines the position-dependent diffusion coefficient, while Equation (10) defines the field-induced drift velocity. Note that we are assuming that the reversal probability vanishes linearly with a.

A pondering pause is required at this point before carrying out the continuum limit. In fact, a moment's thought is enough to realize that there are different ways one can enforce quenched disorder, corresponding to spatially varying hopping rates, in a jump process with the aim of modeling propagation in a inhomogenous medium. The master Equation (7) reflects only one of the possible choices. This observation has been made and thoroughly discussed in Painter and Sherratt [57]. In this paper the authors investigate the movements of cells capable of sensing strategies which depend on environmental factors.

To illustrate this interesting mathematical property, let us consider the jump from site i to site i + 1. The probability of an agent taking an ii + 1 leap can be equally well taken as (1) proportional to qi+, (2) proportional to qi+1+ or (3) proportional to (qi++qi+1+)2. All three cases correspond to the same space-dependent function D(x) in the continuum limit, as prescribed by Equation (9). However, as we shall see in the following, depending on whether rule (1), (2), or (3) is chosen, one is led to totally different advection-diffusion equations in the continuum limit.

2.2.1. No-Exclusion Processes

In order to illustrate this subtle point, let us start with jump processes in the presence of quenched disorder but with no exclusion constraints on the allowed moves. Following the same reasoning that led us to Equation (7), it is not difficult to realize that the three possible choices (1), (2), and (3) referred above lead to the following master equations

Pi(k+1)-Pi(k)=qi+1-Pi+1(k)+qi-1+Pi-1(k)                                - (qi++qi-)Pi(k)    (11)
Pi(k+1)-Pi(k)=qi-Pi+1(k)+qi+Pi-1(k)                                -(qi+1++qi-1-)Pi(k)    (12)
Pi(k+1)-Pi(k)=(qi+1-+qi-2)Pi+1(k)                                +(qi-1++qi+2)Pi-1(k)                                - Pi(k)[(qi+1++qi+2)+(qi-1-+qi-2)](13)

In the first case, Equation (11) (already obtained above), the rate for a given jump depends on the starting site, while the second, Equation (12), is the opposite, the rate depends on the index of the target site. The third case, Equation (13) is an intermediate, symmetric situation where jump rates are associated with links rather than with nodes. In the continuum limit a, Δt → 0 one has

Qi±1Q(x)±Qxa+122Qx2a2+    (14)
ϵi±1ϵ(x)±ϵxa+    (15)
Pi±1(k)P(x,t)±Pxa+122Px2a2+    (16)

Therefore, recalling Equations (9) and (10), we have from Equations (11), (12) and (13), respectively,

P(x,t)t=-J1(x,t)x    (17)
P(x,t)t=-J2(x,t)x    (18)
P(x,t)t=-J(x,t)x    (19)

where

J1(x,t)=-D(x)P(x,t)x+[v(x)-D(x)x]P(x,t)    (20)
J2(x,t)=-D(x)P(x,t)x+[v(x)+D(x)x]P(x,t)    (21)
J(x,t)=-D(x)P(x,t)x+v(x)P(x,t)    (22)

We see that the stochastic processes (11) and (12) correspond to diffusion with drift (or, equivalently, in an external potential). The drift velocity comprises two contributions: the difference between the right-bound and left-bound jump rate fields and a contribution arising from the spatial variation of the diffusion coefficient. Interestingly, the latter term has the opposite sign depending on whether jumps at the microscopic level are controlled by the rates evaluated at the start or target sites. This means that in these cases, even a symmetric jump process results in diffusion with drift. Conversely, considering symmetrized jump rates does not result in the appearance of such additional term in the drift velocity. Working in this setting one hence recover the standard diffusion equation.

2.2.2. Enforcing the Excluded-Volume Constraint

We can now come back to our original aim, i.e., taking the continuum limit of the master Equation (1). It is now clear that, if we want to consider an inhomogeneous medium, we must not restrict to the prescription leading to Equation (1), but we must also consider the other two kinds of processes described above in the absence of exclusion. Extending the reasoning that led us to Equation (1) and recalling Equations (12) and (13), the three master equations with excluded-volume constraints read

Pi(k+1)Pi(k)=[qi+1Pi+1(k)+qi1+Pi1(k)][1Pi(k)]                                Pi(k){qi+[1Pi+1(k)]                                +qi[1Pi1(k)]}    (23)
Pi(k+1)Pi(k)=[qiPi+1(k)+qi+Pi1(k)][1Pi(k)]                                Pi(k){qi+1+[1Pi+1(k)]                                +qi1[1Pi1(k)]}    (24)
Pi(k+1)-Pi(k)=(qi+1-+qi-2)Pi+1(k)[1-Pi(k)]                                +(qi-1++qi+2)Pi-1(k)[1-Pi(k)]                                -(qi+1++qi+2)Pi(k)[1-Pi+1(k)]                                -(qi-1-+qi-2)Pi(k)[1-Pi-1(k)]    (25)

The mean field limit a, Δt → 0 of the above master equations is readily obtained by introducing as above the Taylor expansions of q(x) and P(x, t). By doing this, we find

P(x,t)t=-J1(x,t)x    (26)
P(x,t)t=-J2(x,t)x    (27)
P(x,t)t=-J3(x,t)x    (28)

where

J1(x,t)=-[1-P(x,t)]x[D(x)P(x,t)]- D(x)P(x,t)P(x,t)x                  + v(x)P(x,t)[1-P(x,t)]    (29)
J2(x,t)=P(x,t)x{D(x)[1-P(x,t)]}                  - D(x)[1-P(x,t)]P(x,t)x                  + v(x)P(x,t)[1-P(x,t)]    (30)
J3(x,t)=- D(x)P(x,t)x+ v(x)P(x,t)[1-P(x,t)]    (31)

We see that even in the case of excluded-volume interactions, the mean-field equations can be cast in the form of continuity equations with suitably defined currents given by Equations (29), (30) and (31). We note that, as it happens with the symmetric exclusion process without quenched disorder, the microscopic exclusion constraint disappears in taking the continuum limit of the master Equation (25) in the absence of external fields, which yields a transport equation, Equation (28), identical to its counterpart with no exclusion, Equation (19) in the case v(x) ≡ 0.

Equations (26), (27) and (28) are nonlinear advection-diffusion equation, appropriate for describing the continuum limit of a microscopic exclusion process occurring on a lattice of inequivalent sites in the presence of a field. It is interesting to note that in the case of equivalent sites, which translates to a constant diffusion coefficient, the diffusive parts become linear, i.e., the microscopic exclusion rule is lost in the diffusive part. In the case of zero field, one then simply recovers the ordinary diffusion equation for the three jump processes, which, as it is widely known, can be derived from a microscopic jump process with no exclusion rules. This curious observation has been first reported by Huber [70]. If both the diffusion coefficient and the drift velocity are constant, Equations (26), (27) and (28) all reduce to

Pt=D2P-vx[P(1-P)]    (32)

an equation already obtained recently in Simpson et al. [78].

Equations (26), (27) and (28) contain the single-particle probability field P(x, t), which is a number between zero and one. The value P = 1 should correspond to the maximum density ρM allowed in the system. Thus, more physical equations can be obtained by introducing the agent density

ρ(x,t)ρMP(x,t)=ϕMv1(σ2)P(x,t)    (33)

where

v1(r)=(π12r)dΓ(1+d2)    (34)

is the volume of a d-dimensional sphere1 of radius r and ϕM is the maximum packing fraction for systems of d-dimensional hard spheres, ϕM = 1 (d = 1), ϕM=π120.907 (d = 2) and ϕM=π180.740 (d = 3) [79]. With these definitions, and using a more general vector notation, Equations (26), (27) and (28) become

ρ(x,t)t=-·J1(x,t)    (35)
ρ(x,t)t=-·J2(x,t)    (36)
ρ(x,t)t=-·J3(x,t)    (37)

with

J1(x,t)=-(1-ρρM)[D(x)ρ(x,t)]- D(x)(ρρM)ρ(x,t)                  + v(x)ρ(x,t)(1-ρρM)    (38)
J2(x,t)=ρ(x,t)[D(x)(1-ρρM)]                  - D(x)(1-ρρM)ρ(x,t)                  + v(x)ρ(x,t)(1-ρρM)    (39)
J3(x,t)=- D(x)ρ(x,t)+ v(x)ρ(x,t)(1-ρρM).    (40)

3. Extended Crowding

In the previous section we described how to derive macroscopic mean-field equations starting from microscopic master equations which account for exclusion effects. In our description we did not take into account size and shape of the agents by prescribing that each molecule occupies a single lattice site and can not move if the target site is already occupied by another particle. To recover a macroscopic continuous description, we performed the limit for vanishing lattice spacing. This strategy amounts to considering agents of vanishing size in the continuum limit. We term this peculiar situation in the macroscopic world point-like crowding. As we have shown in the previous section, considerable microscopic information is lost in the continuum limit with point-like agents. The point-like characterization has to be the reason why the mean-field approximation loses the memory of the microscopic exclusion constraint and the diffusion equation is recovered for equivalent sites in the absence of a field. Starting from these premises, we set to work in the extended crowding framework, where the finite size of the particles is explicitly accounted for. Operating within this scenario, the exclusion constraint shall be detectable in the mean-field limit even for unbiased motion and homogeneous domain. A reason for employing the extended crowding philosophy is that in many biological contexts one has to model the interaction between agents displaying complicated shapes, which are not well represented by spherical particle occupying a single lattice site. Biological macromolecules diffusing in the cytoplasm, or even proliferating cells, for example, are often elongated and rod-shaped, a property that inspired the derivation of microscopic models using hard rods as the individual units [80, 81]. This is for example the case of the human peritoneal mesothelial cells modeled in Simpson et al. [80] as hard rods of aspect ratio equal to four. Moreover, when modeling the interaction between multiple species diffusing in the same environment, the differences in shape and dimension of the agents do have an influence on the extent of the collective motion.

The first model for the diffusion of extended particles with exclusion interactions on a one-dimensional lattice was described in Schönherr and Schütz [82] for a general process involving symmetric, as well as asymmetric, hopping dynamics of the rods (the theory is named L-ASEP). Referring to rods of length L, where L has to be interpreted as the aspect ratio between the dimensions of the elongated agent, the authors derive a mean-field equation for the one-dimensional exclusion process. In the absence of a field and for equivalent sites the equation for the density of particles reads

ρt=D02x2[ρ1-(L-1)ρ].    (41)

The previous equation shows that, at least in one dimension, the extended-crowding procedure yields a modified diffusion term even in the absence of external fields or spatial inhomogeneities. By defining a density-dependent diffusion coefficient, which has to be regarded as a collective diffusivity D(ρ)=D0/[(1(L1)ρ]2, eq. (41) can be reformulated as a nonlinear transport equation:

ρt=x[D(ρ)ρx].    (42)

The nonlinear diffusion Equation (41) has been derived through an ingenious but complicate change of variables based on a quantitative mapping between the L-ASEP and the zero-range process [82]. However, it is interesting to note that it can be regarded as the local-density approximation (LDA) of a simple general property of one-dimensional exclusion processes. As pointed out in 1967 by Lebowitz and Percus [83] concerning bulk properties:

For many purposes, however, adding a finite diameter does not introduce any new complications; it merely requires the replacement in certain expressions of the actual volume per particle ρ−1 by the reduced volume ρ−1 − σ, i.e., ρ → ρ∕(1 − σρ).”

In that sense, the quantity ρ∕[1 − (L − 1)ρ] is recognized as an effective density in Ferreira and Alcaraz [84] within the analysis of the velocity of finite sized particles which occupy L units of lattice spacing in one dimension. By performing the substitution

ρ(x,t)ρ(x,t)[1-(L-1)ρ(x,t)]

in Fick's law, one recovers (41). Point-like crowding in the mean field approximation corresponds to systems of fully penetrable spheres, while extended-size crowding yields a transport equation suitable for systems of totally impenetrable (hard) objects. We stress that the case of point-like crowding is recovered for agents of aspect ratio L = 1. The discussion above which leads to Equation (41) applies to one dimensional systems. Starting from this setting, one can raise the question whether similar arguments might be employed to obtain a modified nonlinear equation accounting for excluded volume effects in the diffusion of hard spheres in two and three dimensions. Unfortunately, the strategy used to recover Equation (41) in Schönherr and Schütz [82] can not be employed to provide a description of the extended crowding in higher dimensions. Several other models have been proposed for two and three dimensional domains [76, 80, 85], starting from stochastic processes enforced with different microscopic rules. Depending on the shape of the agents (hard rods or hard spheres), on the prescribed hopping rules, on the allowed mechanisms for changing the orientation of the agents (rotation or reptation), and on the mean field assumptions made to recover the macroscopic picture, different equations are derived for the density of extended particles in higher dimensions. All models yield however the same qualitative behavior at low densities, suggesting that the diffusion coefficient should increase linearly with the concentration amount.

4. Summary and Discussion

Crowding and confinement certainly affect diffusion-driven phenomena, and potentially impact a large plethora of distinct applications. Particularly relevant is the study of mobility inside the cell: the complex cellular environment is in fact populated by obstacles and compartments, which impede the particles flowing. Moreover, the cytoplasm is quite densely packed, different molecular species competing for the available spatial resources. Given these premises, it is important to elaborate on plausible theoretical pictures that enable one to adequately capture the constraints imposed by many-body interactions or reflecting the specificity of the embedding, confining environment.

In this paper, we have reviewed a general framework allowing to obtain macroscopic transport equation accounting for excluded volume effects starting from microscopic stochastic exclusion processes. The aim of this procedure is to derive mean-field equations suitable for describing transport processes in many-body systems in highly non-ideal conditions. Two strategies can be identified to carry out this task. The first route, termed point-like crowding, leads from standard SEPs to the mean-field Equations (35), (36) and (37). For a homogenous medium in the absence of a field these reduce to a simple diffusion equation, which is why this scenario can be dubbed point-like crowding. Only for inequivalent sites and/or in the presence of a field the microscopic exclusion constraint does survive in the mean-field limit. Different is the case of multiple species [86] or recognizable agents [34], where cross-terms appear in the mean field limit due to the fact that particles are made distinguishable. In general, we can state that crowding reflects in the macroscopic limit of the point-like description only if some degree of inhomogeneity is enforced at the microscopic level. Moreover, we have highlighted that in the presence of quenched disorder, i.e., inequivalent sites, there are (at least) three different ways to assign the site dependence of the hopping rates. These correspond to different microscopic stochastic processes and lead to as many different nonlinear advection-diffusion equations in the continuum limit.

The second strategy, named extended-crowding, takes inspiration from a modified microscopic exclusion process in one dimension involving extended agents, the so-called L-ASEP [82]. Qualitatively similar equations for the continuous density are found in higher dimensions at moderate concentrations for different shapes of the individual units. Remarkably, for extended objects characterized by a finite size, the blueprint of the crowding persists at the macroscopic level, even for systems defined on homogeneous domains and without external potentials. This observation marks a clear distinction between the two approaches to which we alluded above and points to the potential interest of the extended crowding setting, which, to the best of our knowledge, remains to be largely explored.

Author Contributions

All authors listed, have made substantial, direct and intellectual contribution to the work, and approved it for publication.

Conflict of Interest Statement

The authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

Footnotes

1. ^We emphasize that we use the general terminology of d-dimensional hard spheres. Obviously, these are hard rods in one dimension and hard disks in two.

References

1. Crank J. The Mathematics of Diffusion. New York, NY: Oxford University Press (1975).

Google Scholar

2. Dhont J. An Introduction to Dynamics of Colloids. Amsterdam: Elsevier (1996).

Google Scholar

3. McGuffee SR, Elcock AH. Diffusion, crowding & protein stability in a dynamic molecular model of the bacterial cytoplasm. PLoS Comput Biol. (2010) 6:e1000694. doi: 10.1371/journal.pcbi.1000694

PubMed Abstract | CrossRef Full Text | Google Scholar

4. Caspi A, Granek R, Elbaum M. Diffusion and directed motion in cellular transport. Phys Rev E (2002) 66:011916. doi: 10.1103/PhysRevE.66.011916

PubMed Abstract | CrossRef Full Text | Google Scholar

5. Kremp D, Schlanges M, Bonitz M, Bornath T. Reaction and diffusion in dense nonideal plasmas. Phys Fluids B Plasma Phys. (1993) 5:216–29.

Google Scholar

6. Saffman PG, Delbrück M. Brownian motion in biological membranes. Proc Natl Acad Sci USA. (1975) 72:3111–3.

PubMed Abstract | Google Scholar

7. Talmon Y, Shtirberg L, Harneit W, Rogozhnikova OY, Tormyshev V, Blank A. Molecular diffusion in porous media by PGSE ESR. Phys Chem Chem Phys. (2010) 12:5998–6007. doi: 10.1039/b922060g

PubMed Abstract | CrossRef Full Text | Google Scholar

8. Mitra PP, Sen PN, Schwartz LM, Le Doussal P. Diffusion propagator as a probe of the structure of porous media. Phys Rev Lett. (1992) 68:3555–8.

PubMed Abstract | Google Scholar

9. Whitaker S. Diffusion and dispersion in porous media. AIChE J. (1967) 13:420–7.

Google Scholar

10. Dix JA, Verkman AS. Crowding effects on diffusion in solutions and cells. Annu Rev Biophys. (2008) 37:247–63. doi: 10.1146/annurev.biophys.37.032807.125824

PubMed Abstract | CrossRef Full Text | Google Scholar

11. Phair RD, Misteli T. High mobility of proteins in the mammalian cell nucleus. Nature (2000) 404:604–9. doi: 10.1038/35007077

PubMed Abstract | CrossRef Full Text | Google Scholar

12. Condamin S, Benichou O, Tejedor V, Voituriez R, Klafter J. First-passage times in complex scale-invariant media. Nature (2007) 450:77–80. doi: 10.1038/nature06201

PubMed Abstract | CrossRef Full Text | Google Scholar

13. Zhou H-X. Protein folding and binding in confined spaces and in crowded solutions. J Mol Recognit. (2004) 17:368–75. doi: 10.1002/jmr.711

PubMed Abstract | CrossRef Full Text | Google Scholar

14. Minton AP. Implications of macromolecular crowding for protein assembly. Curr Opin Struct Biol. (2000) 10:34–9. doi: 10.1016/S0959-440X(99)00045-7

PubMed Abstract | CrossRef Full Text | Google Scholar

15. Zimmerman SB, Minton AP. Macromolecular crowding: biochemical, biophysical, and physiological consequences. Annu Rev Biophys Biomol Struct. (1993) 22:27–65.

PubMed Abstract | Google Scholar

16. Ando T, Skolnick J. Crowding and hydrodynamic interactions likely dominate in vivo macromolecular motion. Proc Natl Acad Sci USA. (2010) 107:18457–62. doi: 10.1073/pnas.1011354107

PubMed Abstract | CrossRef Full Text | Google Scholar

17. Ghosh SK, Cherstvy AG, Metzler R. Non-universal tracer diffusion in crowded media of non-inert obstacles. Phys Chem Chem Phys. (2015) 17.3:1847–58. doi: 10.1039/c4cp03599b

PubMed Abstract | CrossRef Full Text | Google Scholar

18. Pastor I, Vilaseca E, Madurga S, Garcés JL, Cascante M, Mas F. Diffusion of α-chymotrypsin in solution-crowded media. A fluorescence recovery after photobleaching study J Phys Chem B (2010) 114.11:4028–34. doi: 10.1021/jp910811j

PubMed Abstract | CrossRef Full Text | Google Scholar

19. Axelrod D, Koppel DE, Schlessinger J, Elson E, Webb WW. Mobility measurement by analysis of fluorescence photobleaching recovery kinetics. Biophys J. (1976) 16.9:1055.

PubMed Abstract | Google Scholar

20. Metzler R, Jeon JH, Cherstvy A, Barkai E. Anomalous diffusion models and their properties: non-stationarity, non-ergodicity, and ageing at the centenary of single particle tracking newblock. Phys Chem Chem Phys. (2014) 16:24128–64. doi: 10.1039/C4CP03465A

PubMed Abstract | CrossRef Full Text | Google Scholar

21. Sekhar Burada P, Hänggi P, Marchesoni F, Schmid G, Talkner P. Diffusion in confined geometries. ChemPhysChem. (2009) 10.1:45–54. doi: 10.1002/cphc.200800526

PubMed Abstract | CrossRef Full Text | Google Scholar

22. Stauffer D, Schulze C, Heermann DW. Superdiffusion in a model for diffusion in a molecularly crowded environment. J Biol Phys. (2007) 33.4:305–312. doi: 10.1007/s10867-008-9075-2

PubMed Abstract | CrossRef Full Text | Google Scholar

23. Foffi G, Pastore A, Piazza F, Temussi PA. Macromolecular crowding: chemistry and physics meet biology. Phys Biol. (2013) 10.4:040301. doi: 10.1088/1478-3975/10/4/040301

PubMed Abstract | CrossRef Full Text | Google Scholar

24. Rivas G, Ferrone F, Herzfeld J. Life in a crowded world. Phys Biol. (2004) 5.1:23–7. doi: 10.1038/sj.embor.7400056

CrossRef Full Text | Google Scholar

25. Ellis RJ. Macromolecular crowding: obvious but underappreciated. Trends Biochem Sci. (2001) 26.10:597–604. doi: 10.1016/S0968-0004(01)01938-7

PubMed Abstract | CrossRef Full Text | Google Scholar

26. Zhou HX, Rivas G, Minton AP. Macromolecular crowding and confinement: biochemical, biophysical, and potential physiological consequences. Annu Rev Biophys. (2008) 37:375. doi: 10.1146/annurev.biophys.37.032807.125817

PubMed Abstract | CrossRef Full Text | Google Scholar

27. Konopka MC, Shkel IA, Cayley S, Record MT, Weisshaar JC. Crowding and confinement effects on protein diffusion in vivo. J Bacteriol. (2006) 188:6115–23. doi: 10.1128/JB.01982-05

PubMed Abstract | CrossRef Full Text | Google Scholar

28. Balbo J, Mereghetti P, Herten DP, Wade RC. The shape of protein crowders is a major determinant of protein diffusion. Biophys J. (2013) 104:1576–84. doi: 10.1016/j.bpj.2013.02.041

PubMed Abstract | CrossRef Full Text | Google Scholar

29. Weiss M, Elsner M, Kartberg F, Nilsson T. Anomalous subdiffusion is a measure for cytoplasmic crowding in living cells. Biophys J. (2004) 87:3518–24. doi: 10.1529/biophysj.104.044263

PubMed Abstract | CrossRef Full Text | Google Scholar

30. Ben-Avraham D, Havlin S. Diffusion and Reactions in Fractals and Disordered Systems. Cambridge, UK: Cambridge University Press (2000).

Google Scholar

31. Bouchaud J-P, Georges A. Anomalous diffusion in disordered media: statistical mechanisms, models and physical applications. Phys Rep. (1990) 195:127–293.

Google Scholar

32. Novak IL, Kraikivski P, Slepchenko BM. Diffusion in cytoplasm: effects of excluded volume due to internal membranes and cytoskeletal structures. Biophys J. (2009) 97:758–67. doi: 10.1016/j.bpj.2009.05.036

PubMed Abstract | CrossRef Full Text | Google Scholar

33. Nicholson C, Phillips JM. Ion diffusion modified by tortuosity and volume fraction in the extracellular microenvironment of the rat cerebellum. J Physiol. (1981) 321:225–57.

PubMed Abstract | Google Scholar

34. Galanti M, Fanelli D, Maritan A, Piazza F. Diffusion of tagged particles in a crowded medium. Europhys Lett. (2014) 107:20006. doi: 10.1209/0295-5075/107/20006

CrossRef Full Text | Google Scholar

35. von Smoluchowski M. Drei vortrage ubër diffusion brownsche molekular bewegung und koagulation von kolloidteichen. Phys Z (1916) 17:557–71.

Google Scholar

36. Cheung MS, Klimov D, Thirumalai D. Molecular crowding enhances native state stability and refolding rates of globular proteins. Proc Natl Acad Sci USA. (2005) 102:4753–8. doi: 10.1073/pnas.0409630102

PubMed Abstract | CrossRef Full Text | Google Scholar

37. Dzubiella J, McCammon JA. Substrate concentration dependence of the diffusion-controlled steady-state rate constant. J Chem Phys. (2005) 122:184902. doi: 10.1063/1.1887165

PubMed Abstract | CrossRef Full Text | Google Scholar

38. Piazza F, Rios PdeL, Fanelli D, Bongini L, Skoglund U. Anticooperativity in diffusion-controlled reactions with pairs of anisotropic domains: a model for the antigen-antibody encounter. Eur Biophys J. (2005) 34:899–911. doi: 10.1007/s00249-005-0460-5

PubMed Abstract | CrossRef Full Text

39. Fanelli D, McKane AJ. Diffusion in a crowded environment. Phys Rev E (2010) 82:021113. doi: 10.1103/PhysRevE.82.021113

PubMed Abstract | CrossRef Full Text | Google Scholar

40. Schmit JD, Kamber E, Kondev J. Lattice model of diffusion-limited bimolecular chemical reactions in confined environments. Phys Rev Lett. (2009) 102:218302. doi: 10.1103/PhysRevLett.102.218302

PubMed Abstract | CrossRef Full Text | Google Scholar

41. Tachiya M, Traytak SD. Concentration dependence of fluorescence quenching by ionic reactants. J Phys Condens Matter (2007) 19:065111. doi: 10.1088/0953-8984/19/6/065111

CrossRef Full Text | Google Scholar

42. Agrawal M, Santra SB, Anand R, Swaminathan R. Effect of macromolecular crowding on the rate of diffusion-limited enzymatic reaction. PRAMANA-J Phys. (2008) 71:359–68. doi: 10.1007/s12043-008-0169-y

CrossRef Full Text | Google Scholar

43. Ricci N, Barbanera F, Erra F. A quantitative approach to movement, displacement, and mobility of protozoa. J Eukaryot Microbiol. (1998) 45:606–11.

Google Scholar

44. Galanti M, Fanelli D, Piazza F. Conformation-controlled binding kinetics of antibodies. Sci Rep. (2016) 6:18976. doi: 10.1038/srep18976

PubMed Abstract | CrossRef Full Text | Google Scholar

45. Cecconi F, De Los Rios P, Piazza F. Diffusion-limited unbinding of small peptides from PDZ domains. J Phys Chem B (2007) 111:11057–063. doi: 10.1021/jp0730390

PubMed Abstract | CrossRef Full Text | Google Scholar

46. Dorsaz N, De Michele C, Piazza F, Foffi G. Inertial effects in diffusion-limited reactions. J Phys (2010) 22:104116. doi: 10.1088/0953-8984/22/10/104116

PubMed Abstract | CrossRef Full Text | Google Scholar

47. Felderhof BU, Deutch JM. Concentration dependence of the rate of diffusion-controlled reactions. J Chem Phys (1976) 64:4551. doi: 10.1063/1.432087

CrossRef Full Text | Google Scholar

48. Haus JW, Kehr KW. Diffusion in regular and disordered lattices. Phys Rep. (1987) 150:263–406.

Google Scholar

49. Shlesinger MF, West BJ. Random Walks and Their Applications in the Physical and Biological Sciences. New York, NY: American Institute of Physics (1984).

Google Scholar

50. Spitzer F. Principles of Random Walk. The University Series in Higher Mathematics. Princeton, NJ; Toronto, ON; London: D. Van Nostrand Co., Inc. (1984).

Google Scholar

51. Goel NS, Richter-Dyn N. Stochastic Models in Biology. Elsevier (2016).

Google Scholar

52. Montroll EW, Weiss GH. Random walks on lattices. J Math Phys (1965) 6.2:167–81. doi: 10.1063/1.1704269. Available online at: http://scitation.aip.org/content/aip/journal/jmp/6/2/10.1063/1.1704269

CrossRef Full Text

53. Lennard-Jones JE. Processes of adsorption and diffusion on solid surfaces. Trans Faraday Soc (1932) 28:333–59.

Google Scholar

54. Bernasconi J, Alexander S, Orbach R. Classical diffusion in one-dimensional disordered lattice. Phys Rev Lett. (1978) 41.3:185.

Google Scholar

55. Van Beijeren H. Transport properties of stochastic Lorentz models. Rev Mod Phys. (1982) 54.1:195.

Google Scholar

56. Termonia Y. Random walk on a strange attractor. Phys Rev Lett. (1984) 53.14:1356.

Google Scholar

57. Painter KJ, Sherratt JA. Modelling the movement of interacting cell populations. J Theor Biol. (2003) 225:327–39. doi: 10.1016/S0022-5193(03)00258-3

PubMed Abstract | CrossRef Full Text | Google Scholar

58. Privman V, editor. Nonequilibrium Statistical Mechanics in One Dimension. Cambridge University Press (1997).

Google Scholar

59. Liggett TM. Stochastic Interacting Systems: Contact, Voter and Exclusion Processes. Berlin; Heidelberg: Springer-Verlag (1999).

Google Scholar

60. Boltzmann L. Vorlesungen über Gastheorie (Lectures on Gas Theory). Vol. I and II. Transl. by S. G. Brush in 1966. Berkeley: University of California Press (1896).

61. Richards PM. Theory of one-dimensional hopping conductivity and diffusion. Phys Rev B (1977) 16:1393–409.

Google Scholar

62. Spitzer F. Interaction of Markov processes. Adv Math. (1970) 5:246–90.

63. Derrida B, Evans MR, Hakim V, Pasquier V. Exact solution of a 1D asymmetric exclusion model using a matrix formulation. J Phys A Math Gen. (1993) 26:1493.

Google Scholar

64. Schütz G, Domany E. Phase transitions in an exactly soluble one-dimensional exclusion process. J Stat Phys. (1993) 72:277–96.

Google Scholar

65. Derrida B. An exactly soluble non-equilibrium system: the asymmetric simple exclusion process. Phys Rep. (1998) 301:65–83.

Google Scholar

66. Harris RJ, Schütz GM. Fluctuation theorems for stochastic dynamics. J Stat Mech Theory Exp. (2007) 2007.07:P07020. doi: 10.1088/1742-5468/2007/07/P07020

CrossRef Full Text | Google Scholar

67. Golubeva N, Imparato A. Efficiency at maximum power of interacting molecular machines. Phys Rev Lett. (2012) 109:190602. doi: 10.1103/PhysRevLett.109.190602

PubMed Abstract | CrossRef Full Text | Google Scholar

68. Zilman A, Bel G. Crowding effects in non-equilibrium transport through nano-channels. J Phys Condens Matter (2010) 22:454130. doi: 10.1088/0953-8984/22/45/454130

PubMed Abstract | CrossRef Full Text | Google Scholar

69. Reese L, Melbinger A, Frey E. Crowding of molecular motors determines microtubule depolymerization. Biophys J. (2011) 101:2190–200. doi: 10.1016/j.bpj.2011.09.009

PubMed Abstract | CrossRef Full Text | Google Scholar

70. Huber DL. Particle kinetics on one-dimensional lattices with inequivalent sites. Phys Rev B (1977) 15:533–8.

Google Scholar

71. Nossan JS. Disordered exclusion process revisited: some exact results in the low-current regime. J Phys A Math Theor. (2013) 46:315001. doi: 10.1088/1751-8113/46/31/315001

CrossRef Full Text | Google Scholar

72. Lakatos G, O'Brien J, Chou T. Hydrodynamic mean-field solutions of 1D exclusion processes with spatially varying hopping rates. J Phys A Math Gen. (2006) 39:2253–64. doi: 10.1088/0305-4470/39/10/002

CrossRef Full Text | Google Scholar

73. Fernando AE, Landman KA, Simpson MJ. Nonlinear diffusion and exclusion processes with contact interactions. (2010) Phys Rev E 81:011903. doi: 10.1103/PhysRevE.81.011903

PubMed Abstract | CrossRef Full Text | Google Scholar

74. Landman KA, Fernando AE. Myopic random walkers and exclusion processes: single and multispecies. Phys A Stat Mech Appl. (2011) 390:3742–53. doi: 10.1016/j.physa.2011.06.034

CrossRef Full Text | Google Scholar

75. Penington CJ, Hughes BD, Landman KA. Building macroscale models from microscale probabilistic models: a general probabilistic approach for nonlinear diffusion and multispecies phenomena. Phys Rev E (2011) 84:041120. doi: 10.1103/PhysRevE.84.041120

PubMed Abstract | CrossRef Full Text | Google Scholar

76. Penington CJ, Korvasová K, Hughes BD, Landman KA. Collective motion of dimers. Phys Rev E (2012) 86:051909. doi: 10.1103/PhysRevE.86.051909

PubMed Abstract | CrossRef Full Text | Google Scholar

77. Penington CJ, Hughes BD, Landman KA. Interacting motile agents: taking a mean-field approach beyond monomers and nearest-neighbor steps. Phys Rev E (2014) 89:032714. doi: 10.1103/PhysRevE.89.032714

PubMed Abstract | CrossRef Full Text | Google Scholar

78. Simpson MJ, Landman KA, Hughes BD. Pathlines in exclusion processes. Phys Rev E (2009) 79:031920. doi: 10.1103/PhysRevE.79.031920

PubMed Abstract | CrossRef Full Text | Google Scholar

79. Torquato S. Random Heterogenous Materials. Interdisciplinary Applied Mathematics. New York, NY: Springer-Verlag (2002).

80. Simpson MJ, Baker RE, McCue SW. Models of collective cell spreading with variable cell aspect ratio: a motivation for degnerate diffusion models. Phys Rev E (2011) 83:021901. doi: 10.1103/PhysRevE.83.021901

PubMed Abstract | CrossRef Full Text | Google Scholar

81. Baker RE, Simpson MJ. Models of collective cell motion for cell populations with different aspect ratio: diffusion, proliferation and travelling waves. Phys A Stat Mech Appl. (2012) 391:3729–50. doi: 10.1016/j.physa.2012.01.009

CrossRef Full Text | Google Scholar

82. Schönherr G, Schütz GM. Exclusion process for particles of arbitrary extension: hydrodynamic limit and algebraic properties. J Phys A Math Gen. (2004) 37:8215. doi: 10.1088/0305-4470/37/34/002

CrossRef Full Text | Google Scholar

83. Lebowitz JL, Percus JK. Kinetic equations and density expansions: exactly solvable one-dimensional system. Phys Rev. (1967) 155:122–38.

Google Scholar

84. Ferreira AA, Alcaraz FC. Anomalous tag diffusion in the asymmetric exclusion model with particles of arbitrary sizes. Phys Rev E (2002) 65:052102. doi: 10.1103/PhysRevE.65.052102

PubMed Abstract | CrossRef Full Text | Google Scholar

85. Bruna M, Chapman SJ. Excluded-volume effects in the diffusion of hard spheres. Phys Rev E (2012) 85:011103. doi: 10.1103/PhysRevE.85.011103

PubMed Abstract | CrossRef Full Text | Google Scholar

86. Fanelli D, McKane AJ, Pompili G, Tiribilli B, Vassalli M, Biancalani T. Diffusion of two molecular species in a crowded environment: theory and experiments. Phys Biol. (2013) 10:045008. doi: 10.1088/1478-3975/10/4/045008

PubMed Abstract | CrossRef Full Text | Google Scholar

Keywords: transport equations, nonlinear diffusion, macromolecular crowding, non-ideal fluids, stochastic processes

PACS numbers: 02.50.Ey, 05.60.Cd, 87.10.Ed, 66.10.cg, 61.20.-p

Citation: Galanti M, Fanelli D and Piazza F (2016) Macroscopic Transport Equations in Many-Body Systems from Microscopic Exclusion Processes in Disordered Media: A Review. Front. Phys. 4:33. doi: 10.3389/fphy.2016.00033

Received: 24 March 2016; Accepted: 12 July 2016;
Published: 03 August 2016.

Edited by:

Francisco Monroy, Complutense University of Madrid, Spain

Reviewed by:

Andrey Cherstvy, University of Potsdam, Germany
Luis Diambra, National University of La Plata, Argentina

Copyright © 2016 Galanti, Fanelli and Piazza. This is an open-access article distributed under the terms of the Creative Commons Attribution License (CC BY). The use, distribution or reproduction in other forums is permitted, provided the original author(s) or licensor are credited and that the original publication in this journal is cited, in accordance with accepted academic practice. No use, distribution or reproduction is permitted which does not comply with these terms.

*Correspondence: Duccio Fanelli, duccio.fanelli@gmail.com

Download