ArticlePDF Available

Phase diagram of one-dimensional driven lattice gases with open boundaries

Authors:

Abstract and Figures

We consider the asymmetric simple exclusion process (ASEP) with open boundaries and other driven stochastic lattice gases of particles entering, hopping and leaving a one-dimensional lattice. The long-term system dynamics, stationary states, and the nature of phase transitions between steady states can be understood in terms of the interplay of two characteristic velocities, the collective velocity and the shock (domain wall) velocity. This interplay results in two distinct types of domain walls whose dynamics is computed. We conclude that the phase diagram of the ASEP is generic for one-component driven lattice gases with a single maximum in the current-density relation.
Content may be subject to copyright.
J. Phys. A: Math. Gen. 31 (1998) 6911–6919. Printed in the UK PII: S0305-4470(98)92448-9
Phase diagram of one-dimensional driven lattice gases with
open boundaries
Anatoly B Kolomeisky, Gunter M Sch¨
utz, Eugene B Kolomeisky§and
Joseph P Straleyk
Department of Chemistry, Baker Laboratory, Cornell University, Ithaca, NY 14853-1301, USA
IFF, Forschungszentrum J¨
ulich, 52425 J¨
ulich, Germany
§Department of Physics, University of Virginia, Charlottesville, VA 22901, USA
kDepartment of Physics and Astronomy, University of Kentucky, Lexington, KY 40506-0055,
USA
Received 13 March 1998
Abstract. We consider the asymmetric simple exclusion process (ASEP) with open boundaries
and other driven stochastic lattice gases of particles entering, hopping and leaving a one-
dimensional lattice. The long-term system dynamics, stationary states, and the nature of phase
transitions between steady states can be understood in terms of the interplay of two characteristic
velocities, the collective velocity and the shock (domain wall) velocity. This interplay results in
two distinct types of domain walls whose dynamics is computed. We conclude that the phase
diagram of the ASEP is generic for one-component driven lattice gases with a single maximum
in the current-density relation.
1. Introduction
Driven lattice gases with open boundaries are stochastic lattice models of particles which
hop preferentially in one direction and which are connected at their boundaries to particle
reservoirs of constant densities. As a result of this coupling, a stationary particle current
is maintained and the system never reaches an equilibrium state. A well known example,
which has become one of the standard models of nonequilibrium statistical mechanics, is
the asymmetric simple exclusion process (ASEP) where particles interact through hard-core
exclusion. This model is relevant to a variety of phenomena in physics and beyond [1]. The
one-dimensional version of this model was first introduced in 1968 to provide a qualitative
understanding of the kinetics of protein synthesis on RNA templates in terms of a ‘traffic
jam’ of ribosomes along the RNA [2, 3]. Other open systems in which transport of hard-
core particles through narrow (i.e. essentially one-dimensional) channels has been observed
in more recent experiments include certain types of zeolites [4]. Also recently, generalized
ASEPs have been succesfully used in traffic flow modelling [5]. On the mathematical side
[6, 7], the system is of interest in the theory of interacting particle systems because of its
tractability in conjunction with its rich and nontrivial behaviour.
The one-dimensional ASEP with open boundaries shows a complex behaviour exhibiting
nonequilibrium phase transitions that have no analogues in thermal equilibrium [1, 8] (see
section 2). There are intuitive arguments and evidence from the study of related problems
[7, 9, 10] that domain walls (shocks) and the nature of the boundary conditions imposed on
the system play a central role in understanding the physics of the system. The quantitative
0305-4470/98/336911+09$19.50 c
1998 IOP Publishing Ltd 6911
6912 A B Kolomeisky et al
clarification and development of the domain wall picture for the system with open boundaries
[11] constitute the primary goal of this paper. The totally asymmetric exclusion process
(TASEP) is the simplest case and we shall use this model to develop a theory of boundary-
induced phase transitions (section 3). The central ingredient of this theory is an interplay
of two characteristic shock velocities which can be derived from the (bulk) current density
relation j(ρ). We are lead to the observation of two different types of domain walls
whose structure and motion can be characterized to a large extent. The system dynamics,
stationary states, the nature and the location of the phase transitions can then be understood
quantitatively in terms of the domain wall picture. The basic mechanisms which generate
the phase transitions in this model will be shown to be universal and will thus eventually
lead us to an understanding of more generic cases of driven diffusive systems which have
a single maximum in the current density relation j(ρ). Moreover, we shall obtain—with
somewhat stronger assumptions on the coarse-grained domain wall dynamics—quantitative
results on the stationary density profile close to the phase transition lines (section 4). Some
further conclusions are drawn in section 5 where we also summarize our main results.
2. The TASEP with open boundaries
In the TASEP particles enter an N-site lattice from the left with rate αwhenever the first
site is empty. In the bulk particles can jump to the neighbouring right-hand site with
constant rate 1 provided that this site is empty (hard-core exclusion). On site Nparticles
exit the system with rate β. This choice of rates corresponds to a coupling of the system to
reservoirs of constant densities αand 1 β, respectively. The steady states in this system
(figure 1) have been determined exactly [11, 12]. For self-containedness we summarize the
main results.
The hard-core exclusion implies that the steady state particle current jis related to the
bulk density ρby j=ρ(1ρ) [13]. When particles are supplied at the left end with the
rate α>β, and removed at the right end not too fast, β<1
2
, there results a high-density
Figure 1. Phase diagram of the ASEP with stationary values of the particle density ρ, current
j, and allowed types of domain walls: low-density/high-density (0|1), maximal-current/high-
density (m|1), low-density/maximal-current (0|m).
Phase diagram of one-dimensional driven lattice gases 6913
phase (bulk density ρhigh =1β>1
2
) for which particle extraction is the limiting process.
When the particles are supplied not too fast, α<1
2and removed faster than supplied,
β>α, there results a low-density phase limited by particle supply; the bulk density is
ρlow =α. There is a discontinuous phase transition along the line α=β<1
2
. Here the
average particle density changes linearly across the system from αto 1α. When particles
are supplied and removed sufficiently rapidly, α>1
2
,β>1
2
, there results a continuous
phase transition into a maximal-current phase for which transport is bulk dominated; the
bulk density is ρm=1
2, and the current takes on its maximal value of 1
4.
Boundary density profiles in the high- and low-density phases decay exponentially with
finite localization length ξ; the localization length diverges as the maximal current phase is
entered. Within the maximal-current phase the decay is algebraic (equivalent to an infinite
correlation length). Remarkably, ξis composed of two individual localization lengths [11]
ξ1=|ξ
1
βξ
1
α|(1)
with ξ1
σ=−ln 4σ(1σ) for σ<1
2and ξ1
σ=0 for σ>1
2. A curious implication
of this equation is that the boundary density profile changes form within the high- and
low-density phases (for example, at α=1
2with β<1
2
), yet this is not accompanied by any
nonanalyticity in the current or bulk density. Along the first order transition line α=β< 1
2
the correlation length ξis infinite, but the individual localization lengths ξα,β are finite. The
mean-field solution [2, 14] for the density on this line has the form of a shock front or
domain wall, and the linear stationary state profile that is observed in the exact solution
can be interpreted as a superposition of domain walls present at any position [11, 15, 16].
Excepting the special case of this coexistence line the nature and motion of the domain
wall and the consequences for the phase diagram have not been elucidated generally and
explicitly for the problem with open boundaries.
3. Domain wall dynamics
From the exact solution we realize that understanding the origin and the physical meaning
of the localization lengths ξα,β holds the key to understanding the phase diagram of the
system. We start from two examples that provide insight into the physics of the TASEP.
Since the model has a particle–hole symmetry—it can be reformulated in terms of holes
entering the system from the right end—we only need to study the region α>β.
(1) Let us assume [16] that αand βare very small (αN 1,βN 1)and that the
initial distribution of particles is far from the true stationary state. The particles will travel
to the right end where they get stuck. At late times there will be a low-density region
at the left and a high-density region at the right (which we can present schematically as
00001111), with a domain wall between the low (0)- and high (1)-density segments. The
subsequent late-stage evolution of the system can be interpreted in terms of the motion of
this domain wall.
When a particle exits the system, the remaining particles rearrange themselves so that
the whole filled region shrinks by one lattice unit, and the domain wall moves one step to
the right. The conditions αN 1, βN 1 guarantee that while this rearrangement takes
place, no extra particle enters or leaves the system. Similarly a particle entering the system
from the left causes the domain wall to move one unit to the left. As in the Zel’dovich
theory of kinetics of first-order transitions [17], the domain wall motion can be understood
as diffusion of the ‘size of the high-density segment’. The ‘elementary processes’ that
change the length of the filled region consist of motion of the domain wall to the left at rate
6914 A B Kolomeisky et al
αor to the right at rate β, so that the domain wall does a biased random walk with drift
velocity V=βαand diffusion coefficient D= +β)/2. As a result, three physically
different situations are possible.
If α<β, then the domain wall is drifting to the right and will eventually reach the
end of the system; thereafter the system is in the low-density stationary state. If α>β,
the domain wall is travelling to the left, leading to the high-density stationary state. When
α=β, the domain wall position fluctuates with no net drift, and its rms displacement
increases with time as (Dt )1/2. Hence, at large times it can be anywhere in the system,
resulting in a linearly increasing stationary density profile as suggested earlier for finite rates
α, β on an intuitive basis. We note that following the dynamics of the wall also explains
why the phase transition between the cases α>βand α<βis discontinuous.
(2) Let us now assume that α>1
2and that only βis very small. We start from an
empty lattice. After a while, but before the true high-density stationary state is reached, the
system consists of two visually different segments which can be presented schematically as
mmmm1111. Near the left end of the system the high entering rate causes the formation
of a region closely resembling the maximal-current phase (m), while on the right there
is a high-density region (1) dominated by the small exit rate β. The expansion of the
high-density segment is again a biased random walk with some drift velocity and diffusion
coefficient, determined below.
In equilibrium phenomena a domain wall is a localized region where the order parameter
interpolates between degenerate ground states. In this nonequilibrium system a domain wall
is an object connecting two possible stationary states of the system. The domain wall picture
exhibited in the two examples is not specific to the case that one of the boundary rates is
small; we argue that everywhere in the low-/high-density phases there are two kinds of
domain walls: the (0|1)wall (for α< 1
2) connecting the stable high-density stationary state
to the metastable low-density state (as in example (1)), and the (m|1)wall [for α>1/2] that
connects the stable high-density stationary state to the maximal-current phase (example (2)).
This second, distinct type of domain wall is a new notion that we need to introduce here
for a full understanding of the system. The bulk densities far to the left and right low and
ρhigh, respectively) of the (0|1)domain wall are reached exponentially fast with length scale
ξ. As we increase the entering rate α(holding the exit rate β<1
2fixed), the localization
length ξαcharacterizing the low-density behaviour of the domain wall increases, going to
infinity [11, 12] at α=1
2: The (0|1)wall undergoes a continuous phase transition into the
maximal-current/high-density domain wall (m|1)described in our second example. Because
the maximal current phase is algebraic, the stationary density profile for α>1
2approaches
its bulk value not purely exponentially, but with an algebraic correction. Thus the domain
wall transition explains the non-analytical change of the stationary density profile within
the high-density phase at α=1
2for any value of β<1
2
.
We turn now to a quantitative analysis of the physical origin of these observations in
terms of the drift velocity of the domain wall and of the collective velocity of the lattice
gas, defined below. To this end we will now show how the late-stage dynamics of the
system and the approach to the true high-density stationary state is governed by the motion
of the two types of domain walls.
In the introductory examples the domain wall was easy to visualize because the wall
is sharp when the entering and exit rates are small. To compute the drift velocity in the
general case we first note that in any lattice gas (with conserved total particle number) the
local particle density ρ(x,t) =hn
x
(t) isatisfies a lattice continuity equation
d
dtρ(x,t) =jx1(t) jx(t ) (2)
Phase diagram of one-dimensional driven lattice gases 6915
with the local particle current jx(t). In the continuum limit this turns into the usual continuity
equation ∂ρ/∂ t +∂j /∂x =0. With a travelling wave solution of the form ρ(x Vt), and
integrating between minus and plus infinity, we find the domain wall velocity
V=j+j
ρ+ρ
.(3)
In a finite macroscopic system the assumption ρ(x,t) =ρ(x Vt) breaks down near
the boundaries, and therefore the parameters j+,and ρ+,(given in figure 1) should be
understood as bulk stationary state values of the current and density in the far left ()and
far right (+) parts of the domain wall.
For the low-density/high-density domain wall (0|1)one has j+=β(1β),ρ+=1β,
and j=α(1α),ρ=α(figure 1). Substituting these in (3) we find the drift velocity
Vfor the TASEP
V=βα(4)
which we expect to be valid for all α, β < 1
2and which is a well-established result for the
infinite system without boundaries [7, 13]. One realizes that the domain velocity changes
sign at α=β<1
2
, leading to the same scenario as in example 1 for small α, β.
For the maximal-current/high-density domain wall (m|1)we have j+=β(1β),
ρ+=1β, and j=1
4,ρ=1
2. Hence for an initially empty lattice
V=β1
2.(5)
The expression for Vchanges its functional form at α=1
2because the wall has a different
form beyond the phase transition (0|1)(m|1).
To understand why the transition takes place at α=1
2we consider the collective velocity
Vcoll d
dt
Pxxhnx(t)(n0(0)ρ)i
Pxhnx(n0(0)ρ)i(6)
of a lattice gas, averaged over a translationally invariant grand-canonical stationary
distribution of density ρ. The collective velocity measures the drift of the centre of mass of
a momentary local perturbation of the stationary distribution. For any lattice gas satisfying
a continuity equation (2) we obtain by taking the thermodynamic limit of a finite system
the exact nonequilibrium fluctuation-dissipation theorem
Vcoll =
∂ρ j(ρ). (7)
For the TASEP with j=ρ(1ρ) one finds Vcoll =12ρwhich changes sign at ρ=1
2
where the current takes its maximal value j=1
4. As in traffic flow a small perturbation (e.g.
an additional car which has just entered the road) will move with positive velocity (in the
direction of the flow) if the overall density is sufficienty low. However, in a high-density
regime, such a perturbation causes incoming particles to pile up behind the perturbation
(traffic jam) and thus leads to a negative collective velocity of the centre of mass of the
perturbation.
To appreciate the significance of the collective velocity for the phase diagram of the
TASEP consider first the low-density phase along a line with fixed β>1
2
. For a left
boundary density α<1
2a small perturbation of the stationary state (corresponding to a
fluctuation in the injection of particles) travels with positive speed into the bulk where
it will eventually dissipate. However, if the perturbation is maintained, i.e. the constant
left boundary density is increased by a small amount, the perturbation will continuously
penetrate into the bulk and lead to an increase of the bulk density. This happens until
6916 A B Kolomeisky et al
Figure 2. An ASEP with a domain wall present. The diagonal lines represent the instantaneous
position x(measured in lattice spacings) of every 10th particle at various times t(measured in
Monte Carlo steps per site). (a)α=β=0.2; (b)α=0.3, β=0.2; (c)α=0.7, β=0.2.
α=1
2. Further increase of the left boundary density results in a negative collective velocity
and the perturbation does not spread into the bulk. The system has entered the maximal-
current phase where it remains even if the left boundary density is further increased [8].
This phenomenon explains what may be intuitively described as the onset of an overfeeding
with particles [11].
For β<1
2the system does not enter the maximal-current phase (because of the negative
shock velocity), but the overfeeding still occurs for α>1
2and leads to the domain wall
transition (0|1)(m|1). The overfeeding implies that further increase of the left boundary
density beyond 1
2does not result in any change of the characteristic length scales in the
high-density phase. This is seen in the behaviour of the domain wall velocity V(4), (5)
and also in the divergence of the localization length ξα. The overfeeding originates in the
change of sign in the collective velocity. Particle–hole symmetry can be used to extend our
results to the low-density phase. Thus we can explain both the location of the second-order
phase transition lines in the TASEP and the nonanalytic changes in the density profile within
the low- and high-density phases.
We have performed numerical simulations of the ASEP, shown in figure 2, that give
support to our ideas. We chose N=1000 and on each Monte Carlo step attempted to
advance the particle on a randomly chosen site (if there was one present). The initial
configuration had particles placed independently and randomly, with a density step at the
middle or near one end of the line. Time is measured in units of Monte Carlo steps per
site. We represent the subsequent behaviour of the system by tracing the path of every
10th particle. The figure shows three cases: (a)α=β=0.2, describing a (0|1)domain
wall between phases that are particle–hole equivalents; (b)α=0.3, β=0.2, with a (0|1)
domain wall between inequivalent phases; (c)α=0.7, β=0.2, with a (m|1)domain wall
(choosing any α>0.5 would give exactly the same diagram). We verified that for each of
these, equations (4) or (5) accurately predicted the velocity Vof displacement of the wall.
Phase diagram of one-dimensional driven lattice gases 6917
4. Density profiles
We may go further and check this picture by considering the consequences of the fluctuations
in the domain wall position. The domain wall is a compromise between the particle injection
and extraction processes that attempt to enforce their distinct own stationary states. From
our random walk discussion of the domain wall dynamics we expect that a superposition of
the domain wall localized at the left boundary and uniform bulk density ρhigh =1βcapture
the physics of the high-density stationary state, in agreement with intuitive arguments [11].
For detailed analysis we adopt a coarse-grained point of view from which the motion of the
domain wall essentially becomes a Poisson process. For the domain wall the combinations
j+,/(ρ+ρ)which determine domain wall velocity (3) can then be interpreted as being
the effective jump rates DR,Lto the right (left). Thus the diffusion constant is given by
D=1
2
j++j
ρ+ρ
(8)
=1
2
β(1β) +α(1α)
1βαfor α, β < 1
2(9)
=4β(1β) +1
4(12β) for α>1
2<1
2
.(10)
We note that equation (8) correctly reproduces the diffusion coefficient of the TASEP in an
infinite system [7, 18] and that on the coexistence line α=β<1
2
, the expression reduces
to D=α(1α)/(12α) proposed earlier on the basis of current-fluctuation arguments
[16]. On approaching the maximal current phase (β1
2) the drift velocity Vvanishes but
the diffusion coefficient Ddiverges, signalling the failure of the domain wall picture for
β>1
2.
In the TASEP the position of a domain wall is sharp [19]. Thus it is tempting to derive
the localization length which determines the decay of the density profile directly from the
stationary distribution of a biased random walker in a large, but finite system. For a Poisson
process with right and left hopping rates DR,Lthis yields the exact equality ξ=ln (DR/DL).
For the domain wall motion this gives
ξ=|ln (j +/j )| .(11)
Somewhat surprisingly this simple-minded ansatz is in agreement with the exact expression
(1) in all phases and explains the origin of the two independent length scales ξα=ln j
and ξβ=ln j+in terms of the domain wall diffusion.
For a generic lattice gas we take again a coarse-grained approach and introduce a
localization time τthat characterizes the length of time the wall spends away from the
boundary. We argue that the drift distance |V|τand the diffusional wandering (Dτ )1/2each
are of the same order of magnitude as the localization length ξitself; then τ
=D/V 2and
ξ
=D/|V|.(12)
Thus we obtain not only a foundation in terms of the domain wall motion to the
phenomenological derivation of this result by Krug [8] for the transition from the low-
density phase to the maximal-current phase, but also an extension of the validity of (12)
to the other phase transition lines. This domain wall approach is legitimate whenever
the localization length ξis much bigger than the (unit) lattice spacing and than other,
internal bulk correlation lengths which may result from particle interactions in the lattice
gas. Substituting in (12) the expressions (4), (5) and (9), (10) respectively for Vand Dwe
get expressions ξ1,2which are much larger than unity (and thus trustworthy) in the vicinity
6918 A B Kolomeisky et al
of the coexistence line α=β<1
2(case 1) and close to the phase boundary with the
maximal-current phase β=1
2respectively (case 2). In these limits ξcoincides with the
exact expression (11).
Finally we note that by using scaling arguments the power-law decay of the density
profile in the maximal-current phase can be inferred from the superdiffusive spreading [20]
of the fluctuations in the centre-of-mass of a local perturbation [8]. The same result can
also be obtained from a renormalization group analysis of the TASEP with open boundaries
[21].
5. Conclusions
Our domain wall approach to the determination of the phase diagram makes little reference
to the microscopic details of the dynamics. The crucial ingredients, the domain wall velocity
(3) and the collective velocity (7) are generally valid. We conclude that the phase diagram of
the ASEP is universal in the sense that systems with a single maximum jmax in the current-
density relation j(ρ) have a similar phase diagram. Coupling to boundary reservoirs of left
density ρ=αand right density ρ+gives a low- and a high-density phase separated by a
coexistence line which is determined by j(ρ
)=j(ρ
+)<j
max. In the low-density phase
(j(ρ
)<j(ρ
+
)) the bulk density equals the left boundary density ρ, in the high-density
phase (j(ρ
)>j(ρ
+
)) the bulk density equals the right boundary density ρ+. Within these
phases there are domain-wall transitions at ρ±=ρwhich is the density that maximizes
the current and at which the collective velocity changes sign. For ρ+
the system
is in the maximal current phase where the bulk density takes the value ρ. The domain
wall velocity Vvanishes at all phase boundaries. Hence, given j(ρ), one finds the domain
wall velocity (3), the collective velocity (7) and thus the location of the phase transition
lines. We argue that also equation (12) is generally valid close to the phase transition lines.
Provided that the motion of the domain wall on a coarse-grained scale is a Poisson process
we can then predict the shape of the density profile from the fluctuations of the domain wall
motion. The diffusion coefficient Dis singular along second-order lines.
This picture is well supported not only by the exact solution of the TASEP, but also by
exact results on other exclusion processes [15, 22–24] for which the current-density relation
and location of phase transition lines is known. Moreover, from the specific form of the
current-density relations for these models we obtain from (3) and (7) new results for the
domain wall velocity and collective velocity of these models. The first-order transition is
consistent with the experimental data obtained for protein synthesis [2]. We note that data
obtained from automobile traffic flow [25] suggest a single maximum in the current-density
relation. Thus our approach may be used to predict the phase diagram of road segments
between junctions where cars can enter and leave the road.
Acknowledgments
ABK has completed this work in the research group of B Widom, and was supported by
the National Science Foundation and the Cornell University Materials Science Center. EBK
was supported by NSF grants DMR-9412561 and DMR-9531430. We thank C L Henley
and B Widom for useful discussions and M H Ernst for sending [24] prior to publication.
Phase diagram of one-dimensional driven lattice gases 6919
References
[1] Schmittmann B and ZiaRKP1995 Phase Transitions and Critical Phenomena vol 17, ed C Domb and
J Lebowitz (London: Academic)
[2] MacDonald J T, Gibbs J H and Pipkin A C 1968 Biopolymers 61
MacDonald J T and Gibbs J H 1969 Biopolymers 7707
[3] von Heijne G, Blomberg C and Liljenstr¨
om H 1987 J. Theor. Biol. 125 1
Sch¨
utz G M 1997 Int. J. Mod. Phys. B11 197
[4] Kukla V, Kornatowski J, Demuth D, Girnus I, Pfeifer H, Rees L, Schunk S, Unger K and K¨
arger J 1996
Science 272 702
[5] Nagel K and Schreckenberg M 1992 J. Physique I22221
Schreckenberg M, Schadschneider A, Nagel K and Ito N 1995 Phys. Rev. E51 2339
Nagatani T 1995 J. Phys. A: Math. Gen. 28 7079
[6] Ligget T 1985 Interacting Particle Systems (Berlin: Springer)
[7] Spohn H 1991 Large Scale Dynamics of Interacting Particles (New York: Springer)
[8] Krug J 1991 Phys. Rev. Lett. 67 1882
[9] Andjel E D, Bramson M and Ligget T M 1988 Prob. Theor. Relat. Fields 78 231
[10] Janowsky S A and Lebowitz J L 1992 Phys. Rev. A45 618
Sch¨
utz G 1993 J. Stat. Phys. 71 471
[11] Sch¨
utz G and Domany E 1993 J. Stat. Phys. 72 277
[12] Derrida B, Evans M R, Hakim V and Pasquier V 1993 J. Phys. A: Math. Gen. 26 1493
[13] Lebowitz J L, Presutti E and Spohn H 1988 J. Stat. Phys. 51 841 and references therein
[14] Derrida B, Domany E and Mukamel D 1992 J. Stat. Phys. 69 667
[15] Sch¨
utz G 1993 Phys. Rev. E47 4265
[16] Derrida B, Evans M R and Mallick K 1995 J. Stat. Phys. 79 833. For α=βour first example was mentioned
in this reference, however, we go farther in our conclusions.
[17] Landau L D and Lifshitz E M 1981 Physical Kinetics (Course of Theoretical Physics 10) ed E M Lifshitz
and L P Pitaevskii (New York: Pergamon)
[18] Ferrari P A and FontesLRG1994 Prob. Theor. Relat. Fields 99 305
[19] De Masi A, Kipnis C, Presutti E and Saada E 1989 Stoch. Stoch. Rep. 27 151
Ferrari P, Kipnis C and Saada E 1991 Ann. Prob. 19 226
[20] van Beijeren H, Spohn H and Kutner R 1985 Phys. Rev. Lett. 54 2026
van Beijeren H 1991 J. Stat. Phys. 63 47
[21] Janssen H K and Oerding K 1996 Phys. Rev. E53 4544
[22] Sandow S 1994 Phys. Rev. E50 2660
[23] Rajewsky N, Santen L, Schadschneider A and Schreckenberg M 1997 Preprint cond-mat 9710316
[24] Tilstra L G and Ernst M H 1998 J. Phys. A: Math. Gen. 31 5033. The authors have recently obtained the
localization length for the ASEP with parallel update and p=1. For small α, β this result is in agreement
with our predictions and helped us clarify the validity of our expression for the diffusion coefficient.
[25] Hall F L, Brian L A and Gunter M A 1986 Trans. Res. A20 197
... Lattice models of driven diffusive systems have been studied extensively in the last years, see e.g. [7,8]. The simplest model is the asymmetric simple exclusion process (ASEP) in one dimension, where particles hop on a one-dimensional lattice with a strong bias towards one direction (in the simplest case there are no backward steps at all) and the only interaction of the particles is hard core exclusion, i.e., steps to occupied lattice sites are forbidden. ...
... When coupled to open boundaries, this simple model already exhibits a complex phase diagram, see e.g. [8], which we will review below in some detail. The first model for the 1-dimensional ASEP was introduced more than 30 years ago by MacDonald et al. [9,10] in the context of protein synthesis by ribosomes on messenger RNA (mRNA). ...
... The formation of three different phases can be understood in terms of the underlying dynamics of domain walls and density fluctuations [8]. In the low density and high density phases, the selection of the stationary state is governed by domain wall motion. ...
Preprint
The traffic of molecular motors through open tube-like compartments is studied using lattice models. These models exhibit boundary-induced phase transitions related to those of the asymmetric simple exclusion process (ASEP) in one dimension. The location of the transition lines depends on the boundary conditions at the two ends of the tubes. Three types of boundary conditions are studied: (A) Periodic boundary conditions which correspond to a closed torus-like tube. (B) Fixed motor densities at the two tube ends where radial equilibrium holds locally; and (C) Diffusive motor injection at one end and diffusive motor extraction at the other end. In addition to the phase diagrams, we also determine the profiles for the bound and unbound motor densities using mean field approximations and Monte Carlo simulations. Our theoretical predictions are accessible to experiments.
... The hydrodynamic limit of the latter contributes to the theory of the differential equations of conservation laws [4,5]. If a driven system consists of particles of only one type (one species case), its dynamics can be well understood in terms of elementary excitations [6]. Pursuing further the approach of [6], one can explain and subsequently predict the stationary phase diagram for systems with arbitrary currentdensity relation [7]. ...
... If a driven system consists of particles of only one type (one species case), its dynamics can be well understood in terms of elementary excitations [6]. Pursuing further the approach of [6], one can explain and subsequently predict the stationary phase diagram for systems with arbitrary currentdensity relation [7]. However, multi-species models (i.e. ...
... where j Z is given by Eqs (5)(6)(7)(8). Here and below in this section we shall use ρ Z (x, t) for a continuously changing variable, not to be confused with constant ρ A , ρ B from section III. ...
Preprint
We consider classical hard-core particles hopping stochastically on two parallel chains in the same or opposite directions with an inter- and intra-chain interaction. We discuss general questions concerning elementary excitations in these systems, shocks and rarefaction waves. From microscopical considerations we derive the collective velocities and shock stability conditions. The findings are confirmed by comparison to Monte Carlo data of a multi-parameter class of simple two-lane driven diffusion models, which have the stationary state of a product form on a ring. Going to the hydrodynamic limit, we point out the analogy of our results to the ones known in the theory of differential equations of conservation laws. We discuss the singularity problem and find a dissipative term that selects the physical solution.
... Three phases can occur, the low-density, high-density, and maximum current states. Kolomeisky et al. [10] analyzed the formation of the steady-state phases in the mean-field equation by analyzing the dynamics of domain walls that can appear when two phases coexist in the same lane. This work also found that the boundary conditions are not always satisfied and there is no steady localized domain wall in the pure TASEP [11]. ...
... In the steady-state TASEP, the boundary conditions are not always satisfied at the boundary sites (or continuously approaching the boundary). The stability of the boundary density values was determined for the single-lane TASEP by Kolomeisky et al. [10], who worked out the speed at which a domain wall moves. When α and β < 1 2 , the domain wall velocity is V = jr−j l ρr−ρ l , where j r,l denotes the current at the right and left boundaries. ...
... A similar relation can be determined for matching a high-density region to a maximum-current region with a domain wall [10]. Suppose α > 1 2 and β < 1 2 , but a maximum-current phase appears on the left so that j l becomes 1 4 . ...
Preprint
Motor protein motion on biopolymers can be described by models related to the totally asymmetric simple exclusion process (TASEP). Inspired by experiments on the motion of kinesin-4 motors on antiparallel microtubule overlaps, we analyze a model incorporating the TASEP on two antiparallel lanes with binding kinetics and lane switching. We determine the steady-state motor density profiles using phase plane analysis of the steady-state mean field equations and kinetic Monte Carlo simulations. We focus on the the density-density phase plane, where we find an analytic solution to the mean-field model. By studying the phase space flows, we determine the model's fixed points and their changes with parameters. Phases previously identified for the single-lane model occur for low switching rate between lanes. We predict a new multiple coexistence phase due to additional fixed points that appear as the switching rate increases: switching moves motors from the higher-density to the lower-density lane, causing local jamming and creating multiple domain walls. We determine the phase diagram of the model for both symmetric and general boundary conditions.
... Let us point out that our system is markedly different from the open system proposed by Kolomeisky et. al [40] where all phases are triggered by boundaries rather than by a bottleneck and which essentially contains only free traffic (FT) and the HCT state. Specifically, one can associate the 'high-density' state of [40] with HCT, while the 'maximal-current' state corresponds to the outflow region from HCT and finally the 'low-density' state corresponds to FT. ...
... al [40] where all phases are triggered by boundaries rather than by a bottleneck and which essentially contains only free traffic (FT) and the HCT state. Specifically, one can associate the 'high-density' state of [40] with HCT, while the 'maximal-current' state corresponds to the outflow region from HCT and finally the 'low-density' state corresponds to FT. ...
Preprint
We generalize a wide class of time-continuous microscopic traffic models to include essential aspects of driver behaviour not captured by these models. Specifically, we consider (i) finite reaction times, (ii) estimation errors, (iii) looking several vehicles ahead (spatial anticipation), and (iv) temporal anticipation. The estimation errors are modelled as stochastic Wiener processes and lead to time-correlated fluctuations of the acceleration. We show that the destabilizing effects of reaction times and estimation errors can essentially be compensated for by spatial and temporal anticipation, that is, the combination of stabilizing and destabilizing effects results in the same qualitative macroscopic dynamics as that of the respectively underlying simple car-following model. In many cases, this justifies the use of simplified, physics-oriented models with a few parameters only. Although the qualitative dynamics is unchanged, multi-anticipation increase both spatial and temporal scales of stop-and-go waves and other complex patterns of congested traffic in agreement with real traffic data. Remarkably, the anticipation allows accident-free smooth driving in complex traffic situations even if reaction times exceed typical time headways.
... A basic insight that has emerged in the study of the effect of open boundaries already in the 1990s that was mentioned already above, viz., the appearance of boundary-induced phase transitions [73], result from interplay of shocks and rarefaction waves when they reach the boundary [91]. They result in two distinct types of non-equilibrium phase transitions: (i) First order, driven by domain-wall dynamics [72] where the fluctuations of the current become singular [38] and (ii) second-order to extremal-current phases which exhibit a universal density profile [73,88,61] and algebraically decaying correlations [37]. Given the stationary current for the infinite system, the phase diagram can be obtained through an extremal current principle [91]. ...
Preprint
We first survey some open questions concerning stochastic interacting particle systems with open boundaries. Then an asymmetric exclusion process with open boundaries that generalizes the lattice gas model of Katz, Lebowitz and Spohn is introduced and invariance of the one-dimensional Ising measure is proved. The stationary current is computed in explicit form and is shown to exhibit current reversal at some density. Based on the extremal-current principle for one-dimensional driven diffusive systems with one conservation law, the phase diagram for boundary-induced phase transitions is conjectured for this case. There are two extremal-current phases, unlike in the open ASEP (one extremal-current phase) or in the conventional KLS model (one or three extremal-current phases).
... However, when both entry and exit rates are above p/2, the current is maximal, i.e., J = p/4, which is achieved at ρ = 1/2. Note, that in the thermodynamic limit of L → ∞, such a phase diagram [ Fig. 11(b)] is exact [18][19][20]. ...
Article
Full-text available
We consider a family of totally asymmetric simple exclusion processes (TASEPs), consisting of particles on a lattice that require binding by a “token” in various physical configurations to advance over the lattice. Using a combination of theory and simulations, we address the following questions: (i) How does token binding kinetics affect the current-density relation on the lattice? (ii) How does this current-density relation depend on the scarcity of tokens? (iii) How do tokens propagate the effects of the locally imposed disorder (such as a slow site) over the entire lattice? (iv) How does a shared pool of tokens couple concurrent TASEPs running on multiple lattices? and (v) How do our results translate to TASEPs with open boundaries that exchange particles with the reservoir? Since real particle motion (including in biological systems that inspired the standard TASEP model, e.g., protein synthesis or movement of molecular motors) is often catalyzed, regulated, actuated, or otherwise mediated, the token-driven TASEP dynamics analyzed in this paper should allow for a better understanding of real systems and enable a closer match between TASEP theory and experimental observations. Published by the American Physical Society 2025
... Generically, for domain walls separating coexisting phases, the width of wall position fluctuations The Journal of Chemical Physics ARTICLE pubs.aip.org/aip/jcp grows diffusively as t 1/2 as long as t 1/2 /L ≪ 1 in a system of length L. [93][94][95][96] This scaling behavior corresponds to an effective random walk dynamics of the domain wall position. Indeed, this random walk property, which implies a Gaussian shape of the distribution of the domain wall position, has recently been proved rigorously 97 for some parameter manifolds in the ASEP. ...
Article
Full-text available
We analyze correlations between density fluctuations and between current fluctuations in a one-dimensional driven lattice gas with repulsive nearest-neighbor interaction and in single-file Brownian motion of hard spheres dragged across a cosine potential with constant force. By extensive kinetic Monte Carlo and Brownian dynamics simulations, we show that density and current correlation functions in nonequilibrium steady states follow the scaling behavior of the Kardar–Parisi–Zhang (KPZ) universality class. In a coordinate frame comoving with the collective particle velocity, the current correlation function decays as ∼−t−4/3 with time t. Density fluctuations spread superdiffusively as ∼t2/3 at long times, and their spatio-temporal behavior is well described by the KPZ scaling function. In the absence of the cosine potential, the correlation functions in the system of dragged hard spheres show scaling behavior according to the Edwards–Wilkinson universality class. In the coordinate frame comoving with the mean particle velocity, they behave as in equilibrium, with current correlations decaying as ∼−t−3/2 and density fluctuations spreading diffusively as ∼t1/2.
... The collective velocity v(ρ) = dJ(ρ)/dρ is zero in both coexisting phases in that case because the densities of these phases are extrema of the bulk current-density relationJ(ρ). Generically, for domain walls separating coexisting phases the width of wall position fluctuations grows diffusively as t 1/2 as long as t 1/2 /L ≪ 1 in a system of length L [89][90][91][92]. This scaling behavior corresponds to an effective random walk dynamics of the domain wall position. ...
Preprint
We analyze correlations between density fluctuations and between current fluctuations in a one-dimensional driven lattice gas with repulsive nearest-neighbor interaction and in single-file Brownian motion of hard spheres dragged across a cosine potential with constant force. By extensive kinetic Monte Carlo and Brownian dynamics simulations we show that density and current correlation functions in nonequilibrium steady states follow the scaling behavior of the Kardar-Parisi-Zhang (KPZ) universality class. In a coordinate frame comoving with the collective particle velocity, the current correlation function decays as t4/3\sim -t^{-4/3} with time t. Density fluctuations spread superdiffusively as t2/3\sim t^{2/3} at long times and their spatio-temporal behavior is well described by the KPZ scaling function. In the absence of the cosine potential, the correlation functions in the system of dragged hard spheres show scaling behavior according to the Edwards-Wilkinson universality class. In the coordinate frame comoving with the mean particle velocity, they behave as in equilibrium, with current correlations decaying as t3/2\sim -t^{-3/2} and density fluctuations spreading diffusively as t1/2\sim t^{1/2}.
... For q = p and different values of the boundary rates, the model exhibits a nontrivial phase diagram, with different density profiles and mean currents. 70 A representative trajectory in the maximum current phase 70 is shown in Fig. 5a. As a relatively low dimensional many-body system, SEP models represent an ideal testing ground for studying statistical errors in the computation of large deviation functions. ...
Preprint
Large deviation functions contain information on the stability and response of systems driven into nonequilibrium steady states, and in such a way are similar to free energies for systems at equilibrium. As with equilibrium free energies, evaluating large deviation functions numerically for all but the simplest systems is difficult, because by construction they depend on exponentially rare events. In this first paper of a series, we evaluate different trajectory-based sampling methods capable of computing large deviation functions of time integrated observables within nonequilibrium steady states. We illustrate some convergence criteria and best practices using a number of different models, including a biased Brownian walker, a driven lattice gas, and a model of self-assembly. We show how two popular methods for sampling trajectory ensembles, transition path sampling and diffusion Monte Carlo, suffer from exponentially diverging correlations in trajectory space as a function of the bias parameter when estimating large deviation functions. Improving the efficiencies of these algorithms requires introducing guiding functions for the trajectories.
Preprint
Full-text available
The availability of protein is an important factor for the determination of the size of the mitotic spindle. Involved in spindle-size regulation is kinesin-8, a molecular motor and microtubule (MT) depolymerase, which is known to tightly control MT length. Here, we propose and analyze a theoretical model in which kinesin-induced MT depolymerization competes with spontaneous polymerization while supplies of both tubulin and kinesin are limited. In contrast to previous studies where resources were unconstrained, we find that, for a wide range of concentrations, MT length regulation is bistable. We test our predictions by conducting in vitro experiments, and find that the bistable behavior manifests in a bimodal MT length distribution.
Article
Full-text available
We introduce a stochastic discrete automaton model to freeway traffic. Monte-Carlo simulations of the model show a transition from laminar traffic flow to start-stop-waves with increasing vehicle density, as is observed in real freeway traffic. For special cases analytical results can be obtained.
Article
Full-text available
Several recent works have shown that the one-dimensional fully asymmetric exclusion model, which describes a system of particles hopping in a preferred direction with hard core interactions, can be solved exactly in the case of open boundaries. Here the authors present a new approach based on representing the weights of each configuration in the steady state as a product of noncommuting matrices. With this approach the whole solution of the problem is reduced to finding two matrices and two vectors which satisfy very simple algebraic rules. They obtain several explicit forms for these non-commuting matrices which are, in the general case, infinite-dimensional. Their approach allows exact expressions to be derived for the current and density profiles. Finally they discuss briefly two possible generalizations of their results: the problem of partially asymmetric exclusion and the case of a mixture of two kinds of particles.
Article
Full-text available
The stochastic spreading of mass fluctuations in systems described by a fluctuating Burgers equation increases ast 2/3 with time. As a consequence the stochastic motion of a mass front, a point through which no excess mass current is flowing, is shown to increase ast 1/3. The same is true for the stochastic displacement of mass points and shock fronts with respect to their average drift, provided the initial configuration is fixed. An additional average over the stationary distribution of the initial configuration yields stochastic displacements, increasing with time ast 1/2.
Article
The integrable Heisenberg quantum chain with certain non-diagonal boundary fields is the generator of a Markov process known as asymmetric exclusion process with open boundary conditions. This is a driven lattice gas where particles hop randomly along a one-dimensional chain and are injected and absorbed at the boundaries. This model has been suggested in 1968 by MacDonald, Gibbs and Pipkin as a model for the kinetics of protein synthesis on nucleic acid templates. The exact solution of the steady state of the system (corresponding to the exact ground state of the Heisenberg chain) which was obtained recently is shown to be in qualitative agreement with experimental data. The exact solution supports some of the original conclusions drawn from a mean field treatment by MacDonald et al. but gives deeper insight into one important aspect.
Article
We study for a semi-infinite one dimensional initial distribution the asymptotic behaviour in the hydrodynamical limit at the shock. In this case the location of the shock is naturally identified by the position of the leftmost particle of the system for which we prove a central limit theorem. From this we deduce that at the shock local equilibrium does not hold
Article
The creation and annihilation of traffic jams are studied by a computer simulation. The one-dimensional (1D) fully-asymmetric exclusion model with open boundaries for parallel update is extended to take into account stochastic transition of particles (cars) where a particle moves ahead with transition probability pt if the forward nearest neighbour is not occupied. Near pt=1, the system is derived asymptotically into a steady state exhibiting a self-organized criticality. In the self-organized critical state, a traffic jam (start-stop wave) and an empty wave are created at the same time when a car stops temporarily. The traffic jam disappears by colliding with the empty wave. The coalescence process between traffic jams and empty waves is described by the ballistic annihilation process with pair creation. The resulting problem near pt=1 is consistent with the ballistic process in the context of 1D crystal growth studied by Krug and Spohn (1988). The typical lifetime of start-stop waves scales as approximately= Delta pt-0.54+or-0.04 where Delta pt=1-pt. It is shown that the cumulative distribution Nm( Delta pt) of lifetimes satisfies the scaling form Nm( Delta pt) approximately= Delta pt1.1f(m Delta pt0.54). Also, the typical interval between consecutive traffic jams scales as approximately= Delta pt-0.5+or-0.04. The cumulative interval distribution Ns( Delta pt) of traffic jams satisfies the scaling form Ns( Delta pt) approximately= Delta pt0.50g(s Delta pt0.50). For pt<1, no scaling holds.
Article
For a fully synchronous asymmetric exclusion process with open or closed boundaries only partial analytic results are known owing to the appearance of strong-short range correlations, which invalidate simple mean-field approximations. Here we present a new method for calculating basic properties of nonequilibrium steady states, and calculate densities, fluxes, travel times, spatial and temporal correlation functions, phase diagrams, profiles and widths of boundary layers and interfaces between phases in coexistence, as well as their microstructures. This paper is based on two new elements: (i) a microscopic characterization of order parameters and local configurations in the relevant phases, based on the microdynamics of the model, and (ii) an improved mean-field approximation, which neglects certain four-point - and higher-order correlation functions. It is conjectured that the density profiles, obtained here, are exact up to terms that are exponentially small in the system size.
Article
The kinetics of biosynthesis of polypeptides on polyribosomes is analyzed in accordance with a simple mathematical model. Each ribosome is assumed to block L adjacent. (m‐RNA) template sites but to move a distance of one, and only one, template site (nucleotide triplet) upon the addition of each monomer unit to the growing polypeptide chain bound to it. Solutions are sought, for the probability, n j (t), that a template possesses, at time t, a polypeptide chain that has reached degree of polymerization j. Several classes of steady‐state solutions are obtained via machine computation. These correspond to various choices of relative rates of initiation, polymerization along templates, and termination (release of completed chains from templates). Experimental data available from radioactive pulse labeling experiments are discussed. The data obtained by Dintzis, and by Winslow and Ingram, in studies of the synthesis of the α chain of rabbit hemoglobin and human hemoglobin, respectively, are consistent with steady‐state solutions obtained from the current theoretical calculations when the rates of initiation and termination are of comparable magnitude and rate‐determining. In this case, a (relatively) small number of chains are polymerized at a (relatively) fast rate each near the beginning of the template, and there exists a transition to a situation near the end of the template in which a (relatively) large number of chains are polymerized at a (relatively) slow rate each. For this solution the situation near the end of the template is entirely analogous to a traffic jam in automobile traffic.
Article
We review recent developments in the rigorous derivation of hydrodynamic-type macroscopic equations from simple microscopic models: continuous time stochastic cellular automata. The deterministic evolution of hydrodynamic variables emerges as the law of large numbers, which holds with probability one in the limit in which the ratio of the microscopic to the macroscopic spatial and temporal scales go to zero. We also study fluctuations in the microscopic system about the solution of the macroscopic equations. These can lead, in cases where the latter exhibit instabilities, to complete divergence in behavior between the two at long macroscopic times. Examples include Burgers' equation with shocks and diffusion-reaction equations with traveling fronts.