Home Nonequilibrium Transport and Phase Transitions in Driven Diffusion of Interacting Particles
Article Publicly Available

Nonequilibrium Transport and Phase Transitions in Driven Diffusion of Interacting Particles

  • Dominik Lips ORCID logo , Artem Ryabov ORCID logo and Philipp Maass ORCID logo EMAIL logo
Published/Copyright: April 8, 2020

Abstract

Driven diffusive systems constitute paradigmatic models of nonequilibrium physics. Among them, a driven lattice gas known as the asymmetric simple exclusion process (ASEP) is the most prominent example for which many intriguing exact results have been obtained. After summarising key findings, including the mapping of the ASEP to quantum spin chains, we discuss the recently introduced Brownian ASEP (BASEP) as a related class of driven diffusive system with continuous space dynamics. In the BASEP, driven Brownian motion of hardcore-interacting particles through one-dimensional periodic potentials is considered. We study whether current–density relations of the BASEP can be considered as generic for arbitrary periodic potentials and whether repulsive particle interactions other than hardcore lead to similar results. Our findings suggest that shapes of current–density relations are generic for single-well periodic potentials and can always be attributed to the interplay of a barrier reduction, blocking, and exchange symmetry effect. This implies that in general up to five different phases of nonequilibrium steady states are possible for such potentials. The phases can occur in systems coupled to particle reservoirs, where the bulk density is the order parameter. For multiple-well periodic potentials, more complex current–density relations are possible, and more phases can appear. Taking a repulsive Yukawa potential as an example, we show that the effects of barrier reduction and blocking on the current are also present. The exchange symmetry effect requires hardcore interactions, and we demonstrate that it can still be identified when hardcore interactions are combined with weak Yukawa interactions. The robustness of the collective dynamics in the BASEP with respect to variations of model details can be a key feature for a successful observation of the predicted current–density relations in actual physical systems.

1 Introduction

Driven diffusive systems of interacting particles constitute an important class of systems to study fundamental aspects of nonequilibrium physics. This holds in particular for one-dimensional models, where exact analytical derivations are possible or reliable approximations are known, for example, when information about exact equilibrium properties can be utilised for the treatment of nonequilibrium states.

A prominent model in the field of driven diffusive systems is the asymmetric simple exclusion process (ASEP), where particles hop between nearest-neighbour sites of a lattice with a bias in one direction and where the sole interaction between particles is a mutual site exclusion, implying that a lattice site cannot be occupied by more than one particle [1], [2]. In the ASEP on a one-dimensional lattice with L sites and periodic boundary conditions, i.e. a ring of L sites, particles jump to vacant nearest-neighbour sites with rates Γ+ and Γ in clockwise and counterclockwise direction, respectively, where Γ+>Γ for a bias in clockwise direction. In a corresponding open system with L sites, where the leftmost and rightmost lattice site can exchange particles with reservoirs L and R, respectively, additional rates ΓinL, ΓinR and ΓoutL, ΓoutR specify the corresponding rates for particle injection and ejection. Many properties of the ASEP can be inferred from the even simpler totally asymmetric simple exclusion process (TASEP) with unidirectional transport (Γ=0).

Stochastic processes in driven lattice gases are described by a master equation for the probabilities of particle configurations, which can be viewed also as the occupation number representation of a Schrödinger equation in imaginary time [3], [4]. This leads to some interesting connections to quantum systems with in general non-Hermitian Hamilton operator H [2], [3], [5], [6]. As an example, we recapitulate in the Appendix the connection of the ASEP with periodic boundary conditions to the XXZ quantum spin chain with non-Hermitian boundary conditions [6]. Spin chains are often used to study fundamental aspects of nonequilibrium quantum physics. Several examples related to current problems, in particular to questions of equilibration in nonintegrable spin chain models, can be found in this special issue on the physics of non-equilibrium systems.

The ASEP has been intensively studied in the past. Let us summarise here some of the most important findings for the ASEP and variants of it:

  1. Using the Bethe ansatz for corresponding quantum spin chain models, or a construction in terms of matrix product states, exact results for microstate distributions in nonequilibrium steady states (NESSs) could be derived [1], [2], [7]. Matrix product states in principle exist for driven lattice gases with arbitrary nearest-neighbour interactions [8], although their explicit construction may be difficult.

  2. Based on the exact approaches for deriving distribution of microstates in NESS, large deviation functions for fluctuations of time-averaged densities and currents were derived [9], [10], [11]. They have been computed also for coarse-grained descriptions by the macroscopic fluctuation theory [12]. Large deviation functions are argued to play a similar role for time-averaged quantities in NESS as the free energy in equilibrium systems [13]. They can exhibit singularities [14], [15], [16], [17], sometimes referred to as “dynamical phase transitions,” which for certain classes of systems are caused by a violation of an “additivity principle” [18].

  3. The Bethe ansatz turned out to be a valuable tool also for deriving microstate distributions of nonsteady states [19], [20]. The propagator for the microstate time evolution in the ASEP was related to integrated Fredholm determinants [21] and led to the derivation of the Tracy–Widom distribution of random matrix theory for the asymptotic behaviour in case of a step initial condition [22]. This result generalised an earlier one derived for the TASEP [23] and proved that the propagation of density fluctuations in the ASEP belongs to the Kardar–Parisi–Zhang (KPZ) universality class [24].

  4. In open systems coupled to particle reservoirs, phase transitions between NESS occur [25], [26], [27], [28], where, upon change of control parameters characterising the system-reservoir couplings, the bulk density ρb changes discontinuously, or its derivative with respect to the control parameters. Knowing the density dependence of the steady-state bulk current jss(ρ), e.g. from results for a system with periodic boundary conditions, all possible NESS phases with bulk density ρb are predicted by the extremal current principles [25], [27], [29]:

    (1)ρb={argminρρρ+{jss(ρ)},ρρ+,argmaxρ+ρρ{jss(ρ)},ρ+ρ.

    Here ρ and ρ+ can be any densities bounding a monotonically varying region encompassing the plateau part with bulk density ρb (which may strictly exist only in the thermodynamic limit of infinite system size). Which of the phases predicted by (1) really occurs for a given control scheme of system-reservoir couplings is given by the dependence of ρ and ρ+ on respective control parameters.

    The extremal current principles can be reasoned based on the consideration of shock front motions [27], [29], [30] or by resorting to a decomposition of the steady-state current into its drift and diffusive part inside the region of monotonically varying density profile [25]. Because these reasonings do not require specific properties of the ASEP, they are quite generally valid for driven diffusive systems coupled to particle reservoirs. This includes driven lattice gases with interactions other than site exclusion [30], [31], [32], [33], systems with continuous space dynamics, and systems with periodic space structure and/or time-periodic driving, when considering period-averaged densities [34]. For specific system-reservoir couplings termed “bulk-adapted,” it is possible to parameterise the exchange of particles by reservoir densities such that all possible NESS phases must appear. The bulk-adapted couplings can be determined by a general method for driven lattice gases with short-range interactions [33], [34].

  5. For random and non-Poissonian hopping rates, Bose–Einstein–type condensations of vacancies can occur in front of the slowest particle with the smallest jump rate [35], [36].

  6. Coarse-grained continuum descriptions of the ASEP and of multilane variants [37] give rise to an infinite discrete family of nonequilibrium universality classes in nonlinear hydrodynamics, where density fluctuations spread in time by power laws with exponents given by the Kepler ratios of consecutive Fibonacci numbers [38]. This includes the KPZ class, for which an exact expression for the scaling function was derived [39]. Predictions of the theory of fluctuating nonlinear hydrodynamics were recently confirmed in an exact treatment by considering a two-species exclusion process [40].

As for applications, the ASEP appears as a basic building block in manifold descriptions of biological traffic [41], [42]. In fact, the ASEP was introduced first to describe protein synthesis by ribosomes [43], and it is frequently used in connection with the motion of motor proteins along microtubules or actin tracks [44], [45]. An in vitro study with fluorescently labelled single-headed kinesin motors moving along a microtubule provided experimental evidence for a state of coexisting phases with different motor densities [46]. Other applications concern vehicular traffic [47], [48], diffusion of ions through cell membranes [49] and of molecules through nanopores [50], [51], and electron transport along molecular wires in the incoherent classical limit [52], [53]. However, a direct experimental realisation of the ASEP is difficult, because of its discrete nature. Hence, it is important to see whether the nonequilibrium physics in the ASEP is reflected in models with continuous space dynamics.

For a single particle, it is well known that effective hopping transport emerges from an overdamped Brownian motion in a periodic potential with amplitude much larger than the thermal energy. The particle can be viewed to jump between neighbouring wells on a coarse-grained time scale with a rate determined by the inverse Kramers time [54]. One is thus led to ask whether the driven diffusion of many hardcore-interacting particles in a periodic energy landscape can reflect the driven lattice gas dynamics in the ASEP. To answer this question, we recently introduced a corresponding class of nonequilibrium processes termed Brownian ASEP (BASEP) [55], [56], where hard spheres with diameter σ are driven through a periodic potential with wavelength λ by a constant drag force f. For a sinusoidal external potential, we found that the current–density relation of the ASEP is indeed recaptured in the BASEP, but only for a limited range of particle diameters σ. For other σ, quite different behaviours are obtained.

The nonequilibrium physics of the BASEP should be explorable directly by experiment, for example, in setups utilising advanced techniques of microfluidics and optical and/or magnetic micromanipulation [57], [58], [59], [60], [61]. This includes arrangements where the particles are driven by travelling-wave (TW) potentials [62]. Many of the new collective transport properties seen in the BASEP can be even identified by studying local dynamics of individual transitions between potential wells [63].

In this work, we address the question how the current–density relations found for the BASEP in a sinusoidal external potential are affected when considering different external potentials and short-range interactions other than hardcore exclusions. Our investigation for the different external potentials is carried out based on the small-driving approximation introduced in [55] and [56]. With respect to short-range interactions other than hardcore exclusions, we focus on a Yukawa pair potential. It is shown that the current–density relation for single-well periodic potentials and for the Yukawa interaction has similar features as that of the BASEP. This suggests that the BASEP can serve as a reference model for a wide class of external periodic potentials and pair interactions.

In addition, we extend a former analysis to prove that current reversals cannot occur in systems driven by a constant drag and by travelling waves. These proofs are based on an exact calculation of the total entropy production in corresponding NESS for particles with arbitrary pair interactions. Current reversals refer to steady states, where particle flow is opposite to the external bias. They were reported for lattice models [34], [64], [65], [66], [67] and were recently found experimentally in a rocking Brownian motor [59]. Their absence in TW driven systems was conjectured based on simulation results and a perturbative expansion of the single-particle density in the NESS around its period-averaged value [68].

The article is organised as follows. In Section 2, Current–Density Relations: Analytical Results, we present an analytical treatment of densities and currents for the overdamped one-dimensional Brownian motion of particles with arbitrary pair interactions. This section partly summarises results presented earlier [55], [56] and introduces the small-driving approximation used subsequently for our investigation of hardcore-interacting particles. It also contains our proofs on the absence of current reversals for general pair interactions. In Section 3, Hardcore Interacting Particles in Harmonic Potential, we outline our findings for the BASEP with sinusoidal external potential, and in Section 4, Impact of External Periodic Potential, we contrast them with results for a Kronig–Penney and triple-well periodic potential. In Section 5, Impact of Interactions Other Than Hardcore Exclusions, we discuss our results for the Yukawa interaction. Section 6, Summary and Conclusions, concludes the article with a summary and outlook.

2 Current–Density Relations: Analytical Results

The overdamped single-file Brownian motion of N particles in a periodic potential U(x)=U(x+λ) with pair interaction under a constant drag force f is described by the Langevin equations:

(2)dxidt=μ(f+fiintU(xi)xi)+2Dηi(t),

where μ and D=μkBT are the bare mobility and diffusion coefficient, kBT is the thermal energy, and fiint is the interaction force on the ith particle. The ηi(t) are independent and δ-correlated Gaussian white noise processes with zero mean and unit variance, ηi=0 and ηi(t)ηj(t)=δijδ(tt). Unless noted otherwise, we consider closed systems with periodic boundary conditions, which means that the particles are dragged along a ring.

Hardcore interactions imply the boundary conditions |xixj|σ, i.e. overlaps between neighbouring particles are forbidden. For the BASEP with only hardcore interactions, these boundary conditions must be taken into account, while the interaction force fiint can be set to zero in (2). We define the density as a (dimensionless) filling factor of the potential wells, i.e. by ρ=N/M, where M denotes the total number of periods of U(x). The system length is L=Mλ, and the number density is ρ/λ. For hardcore-interacting particles of size σ, the filling factor ρ has the upper bound λ/σ.

The joint probability function (PDF) of the particle centre coordinates x=(x1,,xN) evolves in time according to the N-particle Smoluchowski equation[1],

(3)pN(x,t)t=J(x,t),

where the divergence operator acts on the probability current vector J(x,t) with the ith component given by

(4)Ji(x,t)=μ[fU(xi)xi+fiint(x)]pN(x,t)DpN(x,t)xi.

The first term describes the drift probability current caused by all forces acting on the ith particle, and the second term gives the diffusive current. The interaction force is assumed to be conservative and due to pair interactions uint(xi,xj), i.e. fiint(x)=Uint(x)/xi with

(5)Uint(x)=12ijNuint(xi,xj).

Additional hardcore interactions are not included in the potential (5) but are incorporated into the dynamics by requiring no-flux (reflecting) boundary conditions

(6)[Ji(x,t)Ji+1(x,t)]|xi+1=xi+σ=0,

if neighbouring particles hit each other. These boundary conditions ensure conservation of an initial ordering x1<x2<<xN of the particle positions for all times.

2.1 Exact Current–Density Relation

The local density is

(7)ϱ(x,t)=i=1Nδ[xxi(t)],

where the average is taken with respect to the solution of the Smoluchowski equation (3) subject to some initial condition. It satisfies the continuity equation

(8)ϱ(x,t)t=j(x,t)x,

with the particle current density given by [55]

(9)j(x,t)=μ[fext(x)+fint(x,t)]ϱ(x,t)Dϱ(x,t)x.

Here we introduced the total external force

(10)fext(x)=fU(x)x.

The local interaction force fint(x,t) in (9) is given by

(11)fint(x,t)=1ρ(x,t)0Ldyf2(x,y)ρ2(x,y,t),

where

(12)ϱ2(x,y,t)=ijNδ[xxi(t)]δ[yxj(t)]

is the two-point local density, and f2(x,y) is the interaction force of a particle at position y on a particle at position x. It can by expressed as a sum of two distinct contributions:

(13)f2(x,y)=kBT[δ(yx+σ)δ(xyσ)]uint(x,y)x.

The first term is due to a positive and a negative force, if a particle is in contact with other particles at positions xσ and x+σ, respectively. We note that the δ functions should not be interpreted as a derivative of a rectangular potential barrier of height kBT. Instead, they are a consequence of the noncrossing boundary conditions (6) [55]. The amplitude in front of the δ functions must be an energy on dimensional reasons, for which kBT is the only relevant scale. It corresponds to the typical collision energy due to the thermal noise. The second term in (13) is the force due to the interaction potential (5).

In the steady state of a closed system with periodic boundary conditions, the density profile is time-independent and periodic, ϱss(x+λ)=ϱss(x), and the current constant everywhere in the system. It follows directly from (9) [55]

(14)jss(ρ,σ)=μ[f+1λ0λdxfssint(x)]1λ0λdxϱss1(x).

Up to this point, no approximation has been made. The exact value of the steady-state current (14) depends on both ϱss(x) and the steady-state limit of the two-point density (12). However, a derivation of the two densities in NESS represents a challenging problem, which can be solved in a few special cases only. Therefore, to proceed further, we need to develop an appropriate approximate theory.

2.2 Small-Driving Approximation

For hardcore interactions, the small-driving approximation (SDA) turned out to be particularly successful in capturing qualitative behaviours of jss(ρ,σ) [56]. The approximation is carried out in two steps. First, we linearise the current (14) with respect to f,

(15)jss(ρ,σ)(1+χ)μf1λ0λdxϱeq1(x),f0,

where the response coefficient reads

(16)χ=f[1λ0λdxfssint(x)]|f=0,

and, second, we approximate the linear-response expression (15) by setting χ = 0.

The ad hoc χ = 0 approximation works well in an extended region of particle sizes except for a narrow range σλ/2 [56]. The equilibrium density profile is obtained by minimising the exact density functional for hard rods [69],

(17)Ω[ϱ(x)]=0λdxϱ(x){U(x)μchkBT[1ln(ϱ(x)1η(x))]},

where μch is the chemical potential, and

(18)η(x)=xσxdyϱ(y).

The minimisation yields the structure equation

(19)0=δΩ[ϱ]δϱ|ϱ=ϱeq=kBTln[ϱeq(x)1ηeq(x)]+kBTxx+σdyϱeq(y)1ηeq(y)+[U(x)μch],

which we discretised and solved numerically under periodic boundary conditions (ϱeq(x)=ϱeq(x+λ)). The chemical potential μch was adjusted to give the desired global density (filling factor) ρ=0λdxϱeq(x).

2.3 Entropy Production and Absence of Current Reversals

Theory for hardcore-interacting particles with Uint(x)=0 was the subject of our previous works on the BASEP [55], [56], [63]. Here we extend the analysis to nonzero Uint(x). We start with considerations related to the total entropy production:

(20)S˙tot(t)=S˙sys(t)+S˙med(t),

where S˙sys(t) and S˙med(t) are the entropy production in the system and surrounding medium.

For calculating the time derivative of Ssys(t)=kBdNxpN(x,t)lnpN(x,t), we can replace the time derivative of the PDF by the divergence of the current according to (3). After integrating by parts each individual term of the divergence, we get

(21)S˙sys(t)kB=ΩdNxJ(x,t)lnpN(x,t),

where Ω is the space of all system microstates consistent with the hardcore constraints. As the next step, we replace lnpN/xi via (4), which gives us two terms

(22)S˙sys(t)kB=ΩdNx|J(x,t)|2DpN(x,t)ΩdNxJ(x,t)F(x)kBT.

Here we have introduced the total force F(x) with components Fi(x)=fext(xi)+fiint(x). The first term is always positive and equal to the total entropy production [70]. The second term, proportional to the mean dissipated power, is the entropy production in the medium.

In the steady state, the system entropy is constant, S˙sys(t)=0, and the total entropy production equal to the entropy produced in the surrounding medium:

(23)0S˙tot=1TΩdNxJ(x)F(x)=1Ti=1NΩdNxJi(x)[fU(xi)xiUint(x)xi].

Here, each single-particle term simplifies after introducing the current density ji(x)=ΩdNxJi(x)δ(xix) of the ith particle, and by using that ji(x)=jss/N in the steady state, we obtain

(24)ΩdNxJi(x)[fU(xi)xi]=jssNLf.

The sum over all interaction forces in (23) yields, after integration by parts,

(25)i=1NΩdNxJi(x)Uint(x)xi=ΩdNxUint(x)J(x).

Because the divergence of the current is zero in the steady state, this term vanishes. From (23) and (24), we thus obtain the total entropy production in the Onsager form (current times thermodynamic force)

(26)S˙tot=jssfTL0.

It is extensive in the system size, and the numerator equals the mean heat dissipated at any point of the system in the steady state. As a consequence of the inequality in (26), the steady-state current must have the same sign as the drag force f.

2.4 Entropy Production in Travelling-Wave–Driven Systems and Current Bounds

A feasible way to verify BASEP current–density relations in a laboratory is to consider an equivalent ring system with the TW external periodic potential U(xvwt) and f = 0 [62], [71], [72], [73], [74]. In such a TW system, the ith component of the probability current vector is

(27)JiTW(x,t)=μ[U(xivwt)xi+fiint(x)]pNTW(x,t)DpNTW(x,t)xi.

Under a Galilean transformation

(28)xi(t)=xiTW(t)vwt.

the TW system maps to the corresponding BASEP with potential U(x) and constant drag force

(29)f=vwμ,

provided the pair interaction potential uint(x,y) is a function of the particle distance (xy) only. Local densities and currents of the two corresponding systems are related by [56]

(30)ϱTW(x,t)=ϱ(xvwt,t),
(31)jTW(x,t)=vwϱ(xvwt,t)+j(xvwt,t).

A remarkable aspect of this mapping that has not been discussed in our previous work relates to the fundamental difference of dissipations (their physical origins and their values) in the two pictures. In fact, hopping events that contribute positively to the dissipation (total entropy production) in one picture cause a decrease of the dissipation in the other.

In the BASEP, the dissipation equals the average work done by a constant nonconservative force on all the particles; see (26). In the TW system, there is no nonconservative force. Instead, each particle is acted upon by the time-dependent force [U(xvwt)], and the sum of these actions over all particles gives the total power input into the system. In the steady state, this power is dissipated into the ambient heat bath via friction. Therefore, in the TW model, the total entropy production averaged over one period τ=vw/λ reads (the bar denoting period-averaging in time)

(32)S˙totTW¯=1Ti=1N1τ0τdtU(xivwt)t.

After some algebra similar to that in Section 2.3, one obtains [56]

(33)S˙totTW¯=vwjssTW¯μTL0,

which means that the period-averaged stationary current jstTW¯ must have the same sign as vw. Hence, there are no current reversals in a TW system.

Furthermore, we can relate the TW current in (33) to the corresponding BASEP by taking the period-averaged form of (31) in the steady state:

(34)S˙totTW¯=vw2ρ+vwjssμTL0.

Here, the two terms in the numerator have clear physical meanings. The first contains the expression vw2/μ equal to the dissipated power by a particle moving at constant velocity vw in the fluid characterised by the friction coefficient 1/μ. This term gives the maximal possible dissipation in the TW system corresponding to the case with no jumps over potential barriers where the motion of each particle is exactly phase-locked with the TW potential U(xvwt). The second term contains the current in the corresponding BASEP and is negative because vw and f have opposite signs. Recalling (29), we see that the second term equals exactly the dissipation in the BASEP (26) up to the minus sign. It tells us that the total TW dissipation is diminished by the difference of the average number of jumps over potential barriers in and against bias direction.

Overall, the inequality in (34) implies

(35)0jss(ρ,σ)μfρ,

for f > 0; that is, we obtain the upper bound μfρ for the current, whereas the lower bound follows from (26) as already discussed.

3 Hardcore Interacting Particles in Harmonic Potential

The paradigmatic variant of the BASEP with hardcore-interacting particles diffusing in the external harmonic potential

(36)U(x)=U02cos(2πxλ)

has been studied thoroughly in our previous works [55], [56], [63]. Here, we review its basic properties that shall serve as a “reference case” for the following analysis.

In all illustrations, we fix units setting λ = 1 (defines units of length), λ2/D=1 (time), kBT=1 (energy); this implies that μ=D/(kBT)=1 also. We assume U01, which leads to a hopping-like motion between potential wells that resembles the dynamics on a lattice.

Four representative shapes of current–density relations are shown in Figure 1. In the low-density limit, all curves collapse to the linear behaviour j0(ρ)=v0ρ with the slope given by the velocity v0 of a single (noninteracting) particle. This is given by [75]

(37)v0=Dλ(1eβfλ)0λdxxx+λdyexp[β(U(y)fyU(x)+fx)],

where β=1/(kBT). Beyond the small-ρ region, the shapes change strongly with the particle size σ. This complex behaviour is caused by three competing collective effects:

Figure 1: Simulated steady-state current jss in the BASEP with cosine potential (36) as a function of the density ρ for different particle sizes σ. The solid black line marks the current of noninteracting particles j0(ρ)=v0ρ${j_{0}}(\rho)={v_{0}}\rho$; and the dashed line, the current–density relation jASEP(ρ)=v0ρ(1−ρ)${j_{{\text{ASEP}}}}(\rho)={v_{0}}\rho(1-\rho)$ of a corresponding ASEP [v0 = 0.043 from (37)].
Figure 1:

Simulated steady-state current jss in the BASEP with cosine potential (36) as a function of the density ρ for different particle sizes σ. The solid black line marks the current of noninteracting particles j0(ρ)=v0ρ; and the dashed line, the current–density relation jASEP(ρ)=v0ρ(1ρ) of a corresponding ASEP [v0 = 0.043 from (37)].

Figure 2: The different periodic externals potentials investigated for comparing current–density relations: (a) cosine (36), (b) Kronig–Penney (39), (c) piecewise linear (42), and (d) triple-well (43).
Figure 2:

The different periodic externals potentials investigated for comparing current–density relations: (a) cosine (36), (b) Kronig–Penney (39), (c) piecewise linear (42), and (d) triple-well (43).

  1. The barrier reduction effect leads to a current increase with ρ. It appears in multioccupied wells, where particles are pushing each other to regions of higher potential energy and thus decrease an effective barrier for a transition to neighbouring wells. The effect is best visible for small σ causing currents to be larger than j0(ρ)=v0ρ (solid black line in Fig. 1). Likewise, for small and moderate σ, the strong current increase at larger ρ is due to the occurrence of double-occupied wells.

  2. The blocking effect suppresses the current by reducing the number of transitions between neighbouring wells. It occurs for larger particle sizes: an extended particle is more easily blocked by another one occupying the neighbouring well (compared to smaller σ). To contrast with the most extreme case of blocking, the parabolic current–density relations jASEP(ρ)=v0ρ(1ρ) of a corresponding ASEP is shown as the dashed line in Figure 1.

  3. The exchange symmetry effect causes a deformation of the current–density relation towards the linear behaviour j0(ρ)=v0ρ if the particle size is close to σ=m, m=0, 1, 2,, that is, a multiple integer of λ. In the commensurate case σ=m, the current of interacting particles becomes equal to that of noninteracting ones. This effect is a consequence of the general relation

    (38)jss(ρ,σ)=(1mρ)jss(ρ1mρ,σmλ)

    that maps the stationary current in a system with particles of diameter σ and density ρ to that with particles of diameter σmλ and density ρ/(1mρ), where m=int(σ/λ) is the integer part of σ/λ.

4 Impact of External Periodic Potential

As discussed in the Introduction, we consider further external potentials, namely, the Kronig–Penney, a piecewise linear, and a triple-well potential. These potentials are plotted in Figure 2b–d together with the cosine potential of our reference system in Figure 2a.

4.1 Kronig–Penney Potential

The Kronig–Penney potential has the form

(39)U(x)={0,0x<λw,U0,λwx<λ,

where λw is width of the rectangular well, and λb=λλw is the width of the rectangular barrier. We are interested in the current–density relation for different λw in the limit of large U01. Specifically, we take the same value U0 = 6 as for the reference BASEP with cosine potential discussed in Section 3. In particular, we aim to clarify whether a current enhancement over that of noninteracting particles still occurs. As all particles dragged from one well to a neighbouring one have to surmount the same barrier height U0 now, it is not clear whether multiple occupations of wells lead to an effective barrier reduction. The blocking and exchange symmetry effect are expected to influence the current in an analogous manner.

Inserting the Kronig–Penney potential in (37) yields

(40a)v0(λw)=AB
(40b)A=Dλ(βfλ)2(eβfλ1)
(40c)B=2[eβfλw+eβf(λλw)eβfλ1]×[1cosh(βU0)]+βfλ(eβfλ1)
Figure 3: Drift velocity of a single particle in dependence of λw for the Kronig–Penney potential in (39); the velocity is normalised to f, i.e. its value in a flat (vanishing) external potential. The barrier height and drag force are U0 = 6 and f = 0.2.
Figure 3:

Drift velocity of a single particle in dependence of λw for the Kronig–Penney potential in (39); the velocity is normalised to f, i.e. its value in a flat (vanishing) external potential. The barrier height and drag force are U0 = 6 and f = 0.2.

Figure 4: Current jss, normalised with respect to v0(λw)${v_{0}}({\lambda_{\text{w}}})$ (Fig. 3), for the Kronig–Penney potential in the small-driving approximation as a function of ρ for different σ and (a) λw=0.1${\lambda_{\text{w}}}=0.1$, (b) 0.3, (c) 0.5, (d) 0.7, and (e) 0.9. The barrier height is U0 = 6.
Figure 4:

Current jss, normalised with respect to v0(λw) (Fig. 3), for the Kronig–Penney potential in the small-driving approximation as a function of ρ for different σ and (a) λw=0.1, (b) 0.3, (c) 0.5, (d) 0.7, and (e) 0.9. The barrier height is U0 = 6.

for the single-particle velocity. This result is plotted in Figure 3. As expected, v0(λw) approaches the mean drift velocity μf of a single particle in a flat potential in the limits λw0 (zero well width) and λw1 (zero barrier width). With increasing width of the wells (or of the barriers), v0(λw) rapidly decreases. Interestingly, (40c) implies the symmetry v0(λw)=v0(1λw)=v0(λb), which means the single-particle velocity remains unaltered if the barriers and wells are interchanged.

Current–density relations for hardcore-interacting particles calculated from the SDA (cf. Section 2.2) are shown in Figure 4 for five different values of λw. For each λw, we plotted jss/v0 vs. ρ for eight rod lengths σ analogous to our representation of current–density curves in Figure 1. As can be seen from the graphs, the shapes of the current–density relation are qualitatively comparable to that in Figure 1 for all λw, as well as their overall change with the diameter σ. This means that the interplay of the barrier reduction, blocking, and exchange symmetry is still present. As expected, the overall strengths of the effects in modifying the current of noninteracting particles become weaker with increasing λw; for λw1, the current jss(ρ,λw) indeed approaches j0(ρ,λw1)=ρv0(λw1)=μfρ.

The barrier reduction, however, can no longer be associated with a decrease of an effective barrier height, when two or more particles occupy a potential well. For the Kronig–Penney potential in (39), all particles in a multiple-occupied well have zero energy and need to overcome U0. Nevertheless, we can attribute the enhancement of the current compared to that of noninteracting particles with a barrier reduction. To see this, we analyse the potential of mean force

(41)Umf(x)=kBTlnϱeq(x)+C=Umf(x+λ)

for both the cosine and the Kronig–Penney potential, where the constant C is chosen to give a potential minimum equal to zero. If considering driven Brownian motion of noninteracting particles in the potential Umf(x), the current in the linear response limit would be equal to the many-particle current in the SDA. We therefore can interpret Umf as an effective barrier in the many-particle system. In Figure 5, we show Umf/U0 for both the cosine and the Kronig–Penney potential. At the maximum of the cosine potential at x = 0, we find Umf/U0<1, which means the barrier height is reduced. In contrast, the barrier at the step of the Kronig–Penney potential equals U0 (Umf(x=λw)/U0=1 in Figure 5). However, a barrier reduction is now clearly seen in the plateau part of the barrier in the range λw<x<1. We thus can distinguish between two types of barrier reduction, namely, the first type associated with a reduction of the barrier height and the second type associated with a lowering of the barrier plateau.

Figure 5: Potential of mean force Umf for the cosine potential (blue dashed line) and the Kronig–Penney potential with λw=0.2${\lambda_{\text{w}}}=0.2$ (solid red line). Parameters are σ = 0.2, ρ = 0.5, and U0 = 6 for both potentials.
Figure 5:

Potential of mean force Umf for the cosine potential (blue dashed line) and the Kronig–Penney potential with λw=0.2 (solid red line). Parameters are σ = 0.2, ρ = 0.5, and U0 = 6 for both potentials.

Generally, a single-well periodic potential can be characterised roughly by the widths of a valley and barrier part and the flanks in between these parts. A simple representation is given by the piecewise linear potential

(42)U(x)={0,0|x|λw2,U0λf(|x|λw2),λw2|x|λw2+λf,U0,λw2+λf|x|λ,

shown in Figure 2c, where λw, λf, and λb=λλw2λf are specifying the widths of the valley, barrier, and flanks. We performed additional calculations in the SDA for this potential. For various fixed λw and λf, we always found current–density relations with a behaviour similar to that found in Figure 1 for the BASEP with cosine potential. This model may thus be viewed as representative for Brownian single-file transport through single-well periodic potentials with barriers much larger than the thermal energy.

Current–density relations with different characteristics can, however, be obtained for multiple-well periodic potentials, as we discuss next.

4.2 Triple-Well Potential

The triple-well potential shown in Figure 2d is

(43)U(x)=U0j=13(j+1)cos(2πjxλ),

where we choose U0 = 1 here. For this potential, our calculations of jss(ρ) based on the SDA show a very sensitive dependence on σ. We concentrate here on one particle diameter σ = 0.323, where several local extrema occur; see Figure 6. In this figure, currents are shown up to a density (filling factor) N/M=ρ=2.98, corresponding to a coverage ρσ96% of the system by the hard rods. If ρ approaches its maximal value 1/σ3.10 corresponding to a complete coverage, the numerical calculation of the equilibrium density profile from (19) becomes increasingly difficult. Hence, we refrained to show current data for ρ > 2.98 due to a lack of sufficient numerical accuracy when calculating ϱeq(x). It is important to state in this context that the current for ρ1/σ is expected to approach that of noninteracting particles in a flat potential [56]; that is, it should hold jss(ρ,σ)μfρ for ρ1/σ (except for the singular point ρ=1/σ, where jss(1/σ,σ)=0). This means that the current in Figure 6 must steeply rise for ρ1/σ; that is, there must appear a further local minimum for ρ > 2.98. These arguments apply also to the currents shown in Figures 1 and 4.

Figure 6: Current jss, normalised with respect to f, for the triple-well potential in the small-driving approximation as a function of ρ for σ = 0.323. Densities at the local maxima are ρmax, 1=1.14${\rho_{{\text{max}},\;1}}=1.14$ and ρmax, 2=2.53${\rho_{{\text{max}},\;2}}=2.53$, and ρmin=1.98${\rho_{{\text{min}}}}=1.98$ at the local minimum.
Figure 6:

Current jss, normalised with respect to f, for the triple-well potential in the small-driving approximation as a function of ρ for σ = 0.323. Densities at the local maxima are ρmax, 1=1.14 and ρmax, 2=2.53, and ρmin=1.98 at the local minimum.

To explain the occurrence of the local extrema in Figure 6, we resort to (15) with χ = 0, that is, the SDA. This equation can be interpreted by considering μϱeq(x)dx/λ to be the “local conductivity” of a line segment dx. A serial connection of these segments implies that the “total conductivity” is given by the inverse of the sum of the inverse local conductivities, corresponding to a summation of the respective “local resistivities.” A stronger localisation of ρeq(x) around the minima of the potential leads to a smaller conductivity and hence a smaller current jss(ρ), whereas less localised density profiles lead to larger jss(ρ).

Using this picture, the occurrence of the first maximum in jss(ρ) can be traced back to an increasing occurrence of double-occupied wells for ρ ≳ 1. In double-occupied wells, particle motion is more restricted, leading to a stronger particle localisation at the two deeper minima at about x ≃ 0.2 and x ≃ 0.8; see Figure 2d. Accordingly, jss(ρ) starts to decrease with ρ for ρ ≳ 1 (ρmax, 11.14 in Fig. 6). The decrease of jss(ρ) continues up to a filling factor of about two (ρmin1.98 in Fig. 6), above which more than two particles occupy a well on average. With a significant appearance of triple-occupied wells goes along first a stronger spreading of the density, as the minimum of the potential at x=1/2 becomes occupied in wells containing three particles. The spreading of the density causes jss(ρ) to increase for ρ ≳ 2. A counteracting effect, however, is a strong particle localisation at all potential minima in neighbouring triple-occupied wells, where the hardcore constraints force the particles to become strongly localised around the potential minima. For ρ ≳ 2.5 (ρmax, 22.53 in Fig. 6), every second well is occupied by three particles on average, which lets jss(ρ) decrease again with further increasing ρ.

The more complex current–density relation in Figure 6 leads to a richer variety of NESS phases in open systems compared to the reference BASEP, which can exhibit up to five different phases [55], [56]. To identify all possible NESS phases, we consider the particle exchange with two reservoirs L and R at the left and right ends of an open system to be controlled by two parameters ρL and ρR. As discussed in connection with (1) in the Introduction, these control parameters can be considered as effective densities, or they can be associated with true reservoir densities for specific bulk-adapted couplings of the system to the reservoirs [33], [34].

Applying (1) with ρ=ρL and ρ+=ρR to the current–density relation in Figure 6 results in the diagram with seven different NESS phases I–VII shown in Figure 7. The colour coding shows the value of the bulk density ρb, that is, the order parameter of the phase transitions. Solid lines mark first-order, and dashed lines mark second-order phase transitions, which is reflected in the smooth (continuous) or sudden (jump-like) changes of the colour. The seven phases can be classified in two categories: boundary-matching phases, where ρb is equal to either ρL or ρR, and extremal current phases, where ρb is equal to one of the densities at which jss(ρ) has a local extremum in Figure 6. Specifically, phases I and V are left-boundary matching phases with ρb=ρL, phases III and VII are right-boundary matching phases with ρb=ρR, phase II is a maximal current phase with ρb=ρmax, 1, phase VI is a maximal current phase with ρb=ρmax, 2, and phase IV is a minimal current phase with ρb=ρmin. If one takes into account the existence of the further minimum for ρ > 2.98 in the current–density relation (see discussion above), then even more phases are possible. These additional phases, however, must appear in the two stripes 2.98<ρL,R1/σ3.1, that is, in a very narrow range of one of the two control parameters (marked in black in Fig. 7).

Figure 7: Phase diagram of NESS for the triple-well potential obtained by applying the extremal current principles to the current–density relation in Figure 6. The colour bar encodes the values of the bulk density ρb in an open system coupled to particle reservoirs. Phases I and V are left-boundary matching phases with ρb=ρL${\rho_{\text{b}}}={\rho_{\text{L}}}$, phases III and VII are right-boundary matching phases with ρb=ρR${\rho_{\text{b}}}={\rho_{\text{R}}}$, phases II and VI are maximal current phases with ρb=ρmax, 1${\rho_{\text{b}}}={\rho_{{\text{max}},\;1}}$ and ρb=ρmax, 2${\rho_{\text{b}}}={\rho_{{\text{max}},\;2}}$, respectively, and phase IV is a minimal current phase with ρb=ρmin${\rho_{\text{b}}}={\rho_{{\text{min}}}}$. Solid (dashed) lines mark first-(second-)order phase transitions. The dark black bars at the boundaries mark the two stripes 2.98<ρL,R≤1/σ≅3.1$2.98 < {\rho_{{\text{L}},{\text{R}}}}\leq 1/\sigma\cong 3.1$, where additional phases appear (see the discussion in Section 4.2, Triple-Well Potential).
Figure 7:

Phase diagram of NESS for the triple-well potential obtained by applying the extremal current principles to the current–density relation in Figure 6. The colour bar encodes the values of the bulk density ρb in an open system coupled to particle reservoirs. Phases I and V are left-boundary matching phases with ρb=ρL, phases III and VII are right-boundary matching phases with ρb=ρR, phases II and VI are maximal current phases with ρb=ρmax, 1 and ρb=ρmax, 2, respectively, and phase IV is a minimal current phase with ρb=ρmin. Solid (dashed) lines mark first-(second-)order phase transitions. The dark black bars at the boundaries mark the two stripes 2.98<ρL,R1/σ3.1, where additional phases appear (see the discussion in Section 4.2, Triple-Well Potential).

5 Impact of Interactions Other Than Hardcore Exclusions

In this section, we investigate the impact of other particle interactions beyond hardcore exclusion for the cosine external potential in (36). This is done in two different settings. First, we investigate the repulsive Yukawa potential

(44)uYint(r)=AYer/ξr/ξ,

between particles at distance r for a fixed small amplitude AY=1 and different decay lengths ξ. Second, we combine this Yukawa interaction with hardcore interactions. For obtaining current–density relations, we here employ Brownian dynamics simulations. This is because an exact density functional is not available for the Yukawa interaction, and the SDA with a precise determination of ϱeq(x) cannot be applied. For performing the Brownian dynamics simulations, we used a standard Euler integration scheme of the Langevin equations (2) with a time step Δt=104. To deal with the hardcore interactions, the algorithm developed in [76] was applied.

Figure 8: Simulated current–density relations for the external cosine potential (36) and particle interactions given by the Yukawa potential in (44). The amplitude of the Yukawa potential is AY=1${A_{Y}}=1$, and ξ values specify different decay lengths. The other simulation parameters are U0 = 6 and f = 2. The solid line marks the current j0(ρ)=v0ρ${j_{0}}(\rho)={v_{0}}\rho$ for noninteracting particles, and the dashed line indicates the current–density relation jASEP(ρ)=v0ρ(1−ρ)${j_{{\text{ASEP}}}}(\rho)={v_{0}}\rho(1-\rho)$ of a corresponding ASEP (same curves as in Fig. 1).
Figure 8:

Simulated current–density relations for the external cosine potential (36) and particle interactions given by the Yukawa potential in (44). The amplitude of the Yukawa potential is AY=1, and ξ values specify different decay lengths. The other simulation parameters are U0 = 6 and f = 2. The solid line marks the current j0(ρ)=v0ρ for noninteracting particles, and the dashed line indicates the current–density relation jASEP(ρ)=v0ρ(1ρ) of a corresponding ASEP (same curves as in Fig. 1).

Current–density relations for the Yukawa potential without additional hardcore interactions are shown in Figure 8 for six different values of ξ. These may be viewed to resemble effective particle diameters σ of hardcore-interacting systems. For small ξ, that is, ξ = 0.166 and 0.333 in Figure 8, jss shows an enhancement over that of noninteracting particles (solid black line) due to a prevailing barrier reduction effect similar to that in the reference BASEP for small σ. When enlarging ξ, the current is reduced for small ρ compared to that of noninteracting particles, while it rises strongly for large ρ. Again, this behaviour is analogous to that in the reference BASEP for increasing σ. Because the (effective) blocking effect is not so strong for AY=1, the current–density curves do not approach the limiting jASEP(ρ) as closely as for hardcore interactions. Nevertheless, one can say that the change of the current–density relation with varying ξ is reflecting the interplay of a barrier reduction and blocking effect as in the reference BASEP.

Figure 9: Simulated current–density relations for the cosine external potential and Yukawa interactions as in Figure 8, and additional hardcore interactions for commensurate particle size σ = 1. Parameters for the drag force, as well as for the amplitudes of the cosine and Yukawa potential, are chosen as in Figure 8. The solid line marks the current j0(ρ)=v0ρ${j_{0}}(\rho)={v_{0}}\rho$ for noninteracting particles (same line as in Figs. 1 and 8).
Figure 9:

Simulated current–density relations for the cosine external potential and Yukawa interactions as in Figure 8, and additional hardcore interactions for commensurate particle size σ = 1. Parameters for the drag force, as well as for the amplitudes of the cosine and Yukawa potential, are chosen as in Figure 8. The solid line marks the current j0(ρ)=v0ρ for noninteracting particles (same line as in Figs. 1 and 8).

However, one cannot find certain ξ values, where the current–density relation equals that of noninteracting particles for all ρ. The peculiar exchange symmetry effect in the BASEP for commensurate σ=m, m=0, 1, 2, is caused by the invariance of the stochastic particle dynamics against a specific coordinate transformation containing σ [55], [56]. Such coordinate transformation does not exist for the Yukawa potential.

For the Yukawa potential with additional hardcore interactions, we found changes of current–density curves caused by the barrier reduction and blocking effect as discussed above. But it is interesting to analyse now whether the relation jss(ρ)=j0(ρ)=v0ρ for commensurate σ=m and ξ = 0 is approximately reflected in current–density relations for ξ > 0, where the exchange symmetry effect is no longer strictly valid. One may expect that the hardcore-interacting system should be only weakly perturbed by the Yukawa potential if AY is of the order of the thermal energy and ξ not too large compared to σ. This is indeed confirmed by simulation results for σ = 1 shown in Figure 9. The data points for ξ = 0.166 and ξ = 0.333 lie almost directly on the curve j0(ρ) up to the highest simulated density ρ = 0.9. With increasing ξ, deviations from the linear behaviour are seen, which become the more pronounced the larger ρ. But even for ξ = 2, jss(ρ) follows j0(ρ) closely up to ρ = 0.4. We thus conclude that slight deviations from a perfect hardcore interaction, as they are always present in experiments, still allow an identification of the exchange symmetry effect.

6 Summary and Conclusions

To analyse how generic our previous findings are for the nonequilibrium physics of the BASEP in a sinusoidal potential, we have studied the driven Brownian motion of hardcore-interacting particles for other external periodic potentials. Our calculations were carried out based on a small-driving approximation, which refers to the linear response under neglect of a period-averaged mean interaction force. If the external periodic potential exhibits a single-well structure between barriers, that is, if there is only one local minimum per period, our results provide evidence that the various characteristic shapes of bulk current–density relations jss(ρ) for different particle sizes σ are always occurring. There are differences in the exact functional form and at which σ the type of shape is changing. For all single-well periodic potentials, it is the interplay of a barrier reduction, blocking, and exchange symmetry effect that causes a particular shape to appear. Even for a Kronig–Penney potential with alternating rectangular well and barrier parts, where the barrier reduction effect is not so obvious, we showed that an enhancement of the current over that of noninteracting particles occurs. For that potential, this enhancement can be attributed to an effective reduction of the barrier plateau parts. The generic behaviour of the bulk current–density relations implies that for single-well periodic potentials up to five different NESS phases appear in open BASEP systems coupled to particle reservoirs. This can be concluded by applying the extremal current principles [27], [29].

More complex shapes of jss(ρ) can occur in multiple-well periodic potentials. This was demonstrated for a particular triple-well potential, where our calculations yielded a current–density relation with two local maxima for a certain particle size. In that case, the extremal current principles predict more than five different NESS phases in an open system. When neglecting a narrow range of very high effective reservoir densities, which would be very difficult to realise by specific system-reservoir couplings in simulations or experiments, up to seven different NESS phases are possible. We point out that these results were obtained here for demonstration purposes. Systematic investigations of multiple-well external potentials should be performed in the future with a goal to reach a general classification similar as for the BASEP for single-well periodic potentials.

Current–density relations with several local maxima are particularly interesting in the case of “degenerate maxima,” that is, when the current at the maxima has the same value. In such situations, coexisting NESS phases of maximal current can occur in a whole connected region of the space spanned by the parameters controlling the coupling to the environment [77]. Such states of coexisting extremal current phases have not yet been studied in detail in the literature. Preliminary results for driven lattice gases indicate that fluctuations of interfaces separating extremal current phases exhibit an anomalous scaling with time and system length.[2] This is in contrast to the already well-studied interface fluctuations between the low- and high-density phases in the ASEP, which at long times show a simple random-walk behaviour.

We furthermore performed Brownian dynamics simulations of driven single-file diffusion through a cosine potential for a repulsive particle interaction other than hardcore exclusion. Specifically, we chose a Yukawa interaction with a small interaction amplitude equal to the thermal energy and studied the behaviour for different decay lengths ξ. Current–density relations for this system showed similar shapes as for the BASEP except for the effects implied by the exchange symmetry effect, which is absent for other interactions than hardcore. The change of shapes is solely determined by the interplay of a barrier reduction and effective blocking effect. If the hardcore interaction and the weak Yukawa interaction are combined, the consequences of the exchange symmetry effect can still be seen for particle sizes commensurate with the wavelength of the cosine potential. The current jss(ρ) follows closely that of noninteracting particles up to high densities even for large ξ. This means that deviations from a perfect hardcore interaction in experiments should still allow one to verify the exchange symmetry.

Acknowledgement

Financial support by the Czech Science Foundation (Funder Id: http://dx.doi.org/10.13039/501100001824, project no. 20-24748J) and the Deutsche Forschungsgemeinschaft (Funder Id: http://dx.doi.org/10.13039/501100001659, project no. 397157593) is gratefully acknowledged. We sincerely thank the members of the DFG Research Unit FOR 2692 for fruitful discussions.

Appendix: Example for Connection of ASEP to Quantum Spin Chain

Let p(n,t) denote the probability of configurations n={ni,i=1,,L} of occupation numbers ni{0, 1} in a single-species fermionic lattice gas at time t. Its time evolution is described by the master equation

(A.1)dp(n,t)dt=n[wnnp(n,t)wnnp(n,t)]=nHnnp(n,t)

where wnn is the transition rate from configuration n′ to n (for nn; otherwise, wnn=0), and Hnn=wnnδnnnwnn. This master equation corresponds to the occupation number representation of a Schrödinger equation

(A.2)d|pdt=H|p

in imaginary time [3], [4].

For the ASEP with periodic boundary conditions (nL+1=n1, n0=nL), the transitions rates wnn can be written as

(A.3)wnn=j=1Lδnn(j)[Γnj+1(1nj)+Γ+nj(1nj+1)]=j=1Lδnn(j)[Γ(1nj+1)nj+Γ+(1nj)nj+1],

where n(j) denotes the configuration n with the occupation numbers at sites j and (j+1) interchanged, i.e. nk(j)=nk for kj,(j+1), nj(j)=nj+1, and nj+1(j)=nj. Accordingly, the matrix elements Hnn=n|H|n=wnnδnnnwnn are

(A.4)Hnn=j=1L{δnn(j)[Γnj(1nj+1)+Γ+(1nj)nj+1]δnn[Γ(1nj)nj+1+Γ+nj(1nj+1)]}.

Because n|cjcj+1|n=δnn(j)nj(1nj+1) for creation and annihilation operators cj and cj of a particle at site j, the matrix elements in (A.4) are equal to those of the Hamiltonian

(A.5)H=j=1L{Γ[cjcj+1nj+1(1nj)]+Γ+[cj+1cjnj(1nj+1)]}

of spinless fermions, which for Γ+Γ is non-Hermitian. In a representation by Pauli matrices, one can write cj=σj+/2=(σjx+iσjy)/2, cj=σj/2=(σjxiσjy)/2, and nj=(1+σjz)/2, giving

(A.6)H=14j=1L[Γσj+σj+1+Γ+σj+1+σj+(Γ+Γ+)(σjzσj+1z1)].

The periodic boundary conditions imply σL+1±=σ1± and σL+1z=σ1z.

A transformed H=VHV1 with (nonsingular) operator V has the same spectrum as H, where eigenstates |φ and |φ of H and H′ to the same eigenvalue are related by |φ=V|φ. Such transformation can be used to symmetrise the non-Hermitian part (Γσj+σj+1+Γ+σj+1+σj) in (A.6) by choosing V=exp(αj=1Ljσjz) with some constant α, because Vσj±V1=e±2αjσj± [6]. With Vσj+σj+1V1=Vσj+VV1σj+1V1=e2ασj+σj+1 and Vσj+1+σjV1=e2ασj+1+σj, the symmetrisation is achieved by requiring Γe2α=Γ+e2α, that is, by setting eα=(Γ/Γ+)1/4. The transformed Hamiltonian is that of a quantum XXZ chain,

(A.7)H=ΓΓ+2j=1L(σjxσj+1x+σjyσj+1y+Δσjzσj+1zΔ)

with Δ=(Γ+Γ+)/(2ΓΓ+), but now non-Hermitian boundary conditions VσL+1±V1=Vσ1±V1, that is, σL+1±=e2αLσ1±=(Γ/Γ+)L/2σ1± and σL+1z=σ1z.

References

[1] B. Derrida, Phys. Rep. 301, 65 (1998).10.1016/S0370-1573(98)00006-4Search in Google Scholar

[2] G. M. Schütz, in: Phase Transitions and Critical Phenomena, Vol. 19 (Eds. C. Domb, J. Lebowitz), Academic Press, London 2001, p. 1.10.1016/S1062-7901(01)80015-XSearch in Google Scholar

[3] L.-H. Gwa and H. Spohn, Phys. Rev. A 46, 844 (1992).10.1103/PhysRevA.46.844Search in Google Scholar

[4] S. Sandow and S. Trimper, Europhys. Lett. (EPL) 21, 799 (1993).10.1209/0295-5075/21/8/001Search in Google Scholar

[5] S. Sandow, Phys. Rev. E 50, 2660 (1994).10.1103/PhysRevE.50.2660Search in Google Scholar

[6] M. Henkel and G. Schütz, Phys. A 206, 187 (1994).10.1016/0378-4371(94)90124-4Search in Google Scholar

[7] R. A. Blythe and M. R. Evans, J. Phys. A Math. Theor. 40, R333 (2007).10.1088/1751-8113/40/46/R01Search in Google Scholar

[8] K. Krebs and S. Sandow, J. Phys. A Math. Gen. 30, 3165 (1997).10.1088/0305-4470/30/9/024Search in Google Scholar

[9] B. Derrida and J. L. Lebowitz, Phys. Rev. Lett. 80, 209 (1998).10.1103/PhysRevLett.80.209Search in Google Scholar

[10] B. Derrida, J. L. Lebowitz, and E. R. Speer, Phys. Rev. Lett. 89, 030601 (2002).10.1103/PhysRevLett.89.030601Search in Google Scholar PubMed

[11] P. L. Krapivsky, K. Mallick, and T. Sadhu, Phys. Rev. Lett. 113, 078101 (2014).10.1103/PhysRevLett.113.078101Search in Google Scholar PubMed

[12] L. Bertini, A. De Sole, D. Gabrielli, G. Jona-Lasinio, and C. Landim, Rev. Mod. Phys. 87, 593 (2015).10.1103/RevModPhys.87.593Search in Google Scholar

[13] H. Touchette, Phys. Rep. 478, 1 (2009).10.1016/j.physrep.2009.05.002Search in Google Scholar

[14] L. Bertini, A. De Sole, D. Gabrielli, G. Jona-Lasinio, and C. Landim, Phys. Rev. Lett. 94, 030601 (2005).10.1103/PhysRevLett.94.030601Search in Google Scholar PubMed

[15] T. Bodineau and B. Derrida, Phys. Rev. E 72, 066110 (2005).10.1103/PhysRevE.72.066110Search in Google Scholar PubMed

[16] C. Appert-Rolland, B. Derrida, V. Lecomte, and F. van Wijland, Phys. Rev. E 78, 021122 (2008).10.1103/PhysRevE.78.021122Search in Google Scholar PubMed

[17] Y. Baek, Y. Kafri, and V. Lecomte, Phys. Rev. Lett. 118, 030604 (2017).10.1103/PhysRevLett.118.030604Search in Google Scholar PubMed

[18] T. Bodineau and B. Derrida, Phys. Rev. Lett. 92, 180601 (2004).10.1103/PhysRevLett.92.180601Search in Google Scholar PubMed

[19] G. M. Schütz, J. Stat. Phys. 88, 427 (1997).10.1007/BF02508478Search in Google Scholar

[20] C. A. Tracy and H. Widom, Commun. Math. Phys. 279, 815 (2008).10.1007/s00220-008-0443-3Search in Google Scholar

[21] C. A. Tracy and H. Widom, J. Stat. Phys. 132, 291 (2008).10.1007/s10955-008-9562-7Search in Google Scholar

[22] C. A. Tracy and H. Widom, Commun. Math. Phys. 290, 129 (2009).10.1007/s00220-009-0761-0Search in Google Scholar

[23] K. Johansson, Commun. Math. Phys. 209, 437 (2000).10.1007/s002200050027Search in Google Scholar

[24] M. Kardar, G. Parisi, and Y.-C. Zhang, Phys. Rev. Lett. 56, 889 (1986).10.1103/PhysRevLett.56.889Search in Google Scholar PubMed

[25] J. Krug, Phys. Rev. Lett. 67, 1882 (1991).10.1103/PhysRevLett.67.1882Search in Google Scholar PubMed

[26] G. Schütz and E. Domany, J. Stat. Phys. 72, 277 (1993).10.1007/BF01048050Search in Google Scholar

[27] A. B. Kolomeisky, G. M. Schütz, E. B. Kolomeisky, and J. P. Straley, J. Phys. A Math. Gen. 31, 6911 (1998).10.1088/0305-4470/31/33/003Search in Google Scholar

[28] A. Brzank and G. M. Schütz, J. Stat. Mech. Theor. Exp. 2007, P08028 (2007).10.1088/1742-5468/2007/08/P08028Search in Google Scholar

[29] V. Popkov and G. M. Schütz, Europhys. Lett. 48, 257 (1999).10.1209/epl/i1999-00474-0Search in Google Scholar

[30] J. S. Hager, J. Krug, V. Popkov, and G. M. Schütz, Phys. Rev. E 63, 056110 (2001).10.1103/PhysRevE.63.056110Search in Google Scholar PubMed

[31] T. Antal and G. M. Schütz, Phys. Rev. E 62, 83 (2000).10.1103/PhysRevE.62.83Search in Google Scholar

[32] M. Dierl, P. Maass, and M. Einax, Phys. Rev. Lett. 108, 060603 (2012).10.1103/PhysRevLett.108.060603Search in Google Scholar PubMed

[33] M. Dierl, M. Einax, and P. Maass, Phys. Rev. E 87, 062126 (2013).10.1103/PhysRevE.87.062126Search in Google Scholar PubMed

[34] M. Dierl, W. Dieterich, M. Einax, and P. Maass, Phys. Rev. Lett. 112, 150601 (2014).10.1103/PhysRevLett.112.150601Search in Google Scholar PubMed

[35] M. R. Evans, EPL 36, 13 (1996).10.1209/epl/i1996-00180-ySearch in Google Scholar

[36] R. J. Concannon and R. A. Blythe, Phys. Rev. Lett. 112, 050603 (2014).10.1103/PhysRevLett.112.050603Search in Google Scholar PubMed

[37] V. Popkov and M. Salerno, Phys. Rev. E 69, 046103 (2004).10.1103/PhysRevE.69.046103Search in Google Scholar PubMed

[38] V. Popkov, A. Schadschneider, J. Schmidt, and G. M. Schütz, PNAS 112, 12645 (2015).10.1073/pnas.1512261112Search in Google Scholar PubMed PubMed Central

[39] M. Prähofer and H. Spohn, J. Stat. Phys. 115, 255 (2004).10.1023/B:JOSS.0000019810.21828.fcSearch in Google Scholar

[40] Z. Chen, J. de Gier, I. Hiki, and T. Sasamoto, Phys. Rev. Lett. 120, 240601 (2018).10.1103/PhysRevLett.120.240601Search in Google Scholar PubMed

[41] A. Schadschneider, D. Chowdhury, and K. Nishinari, Stochastic Transport in Complex Systems: From Molecules to Vehicles, 3rd ed., Elsevier Science, Amsterdam, 2010.10.1016/B978-0-444-52853-7.00001-4Search in Google Scholar

[42] T. Chou, K. Mallick, and R. K. P. Zia, Rep. Prog. Phys. 74, 116601 (2011).10.1088/0034-4885/74/11/116601Search in Google Scholar

[43] C. T. MacDonald, J. H. Gibbs, and A. C. Pipkin, Biopolymers 6, 1 (1968).10.1002/bip.1968.360060102Search in Google Scholar PubMed

[44] A. B. Kolomeisky, J. Phys. Condens. Matter 25, 463101 (2013).10.1088/0953-8984/25/46/463101Search in Google Scholar PubMed PubMed Central

[45] C. Appert-Rolland, M. Ebbinghaus, and L. Santen, Phys. Rep. 593, 1 (2015).10.1016/j.physrep.2015.07.001Search in Google Scholar

[46] K. Nishinari, Y. Okada, A. Schadschneider, and D. Chowdhury, Phys. Rev. Lett. 95, 118101 (2005).10.1103/PhysRevLett.95.118101Search in Google Scholar PubMed

[47] C. Appert-Rolland, J. Cividini, and H. J. Hilhorst, J. Stat. Mech. Theor. Exp. 2011, P10014 (2011).10.1088/1742-5468/2011/10/P10014Search in Google Scholar

[48] M. E. Foulaadvand and P. Maass, Phys. Rev. E 94, 012304 (2016).10.1103/PhysRevE.94.012304Search in Google Scholar PubMed

[49] B. Hille, Ionic Channels of Excitable Membranes, 3rd ed., Sinauer Associates, Sunderland, MA 2001.Search in Google Scholar

[50] C.-Y. Cheng and C. R. Bowers, ChemPhysChem 8, 2077 (2007).10.1002/cphc.200700336Search in Google Scholar PubMed

[51] M. Dvoyashkin, H. Bhase, N. Mirnazari, S. Vasenkov, and C. R. Bowers, Anal. Chem. 86, 2200 (2014).10.1021/ac403868tSearch in Google Scholar PubMed

[52] A. Nitzan, Annu. Rev. Phys. Chem. 52, 681 (2001).10.1146/annurev.physchem.52.1.681Search in Google Scholar PubMed

[53] Y. A. Berlin, A. L. Burin, and M. A. Ratner, J. Am. Chem. Soc. 123, 260 (2001).10.1021/ja001496nSearch in Google Scholar PubMed

[54] H. Risken, The Fokker-Planck Equation: Methods of Solution and Applications, Springer-Verlag, Berlin 1985.10.1007/978-3-642-96807-5Search in Google Scholar

[55] D. Lips, A. Ryabov, and P. Maass, Phys. Rev. Lett. 121, 160601 (2018).10.1103/PhysRevLett.121.160601Search in Google Scholar PubMed

[56] D. Lips, A. Ryabov, and P. Maass, Phys. Rev. E 100, 052121 (2019).10.1103/PhysRevE.100.052121Search in Google Scholar PubMed

[57] A. V. Arzola, M. Villasante-Barahona, K. Volke-Sepúlveda, P. Jákl, and P. Zemánek, Phys. Rev. Lett. 118, 138002 (2017).10.1103/PhysRevLett.118.138002Search in Google Scholar PubMed

[58] M. J. Skaug, C. Schwemmer, S. Fringes, C. D. Rawlings, and A. W. Knoll, Science 359, 1505 (2018).10.1126/science.aal3271Search in Google Scholar PubMed

[59] C. Schwemmer, S. Fringes, U. Duerig, Y. K. Ryu, and A. W. Knoll, Phys. Rev. Lett. 121, 104102 (2018).10.1103/PhysRevLett.121.104102Search in Google Scholar PubMed

[60] R. Stoop, A. Straube, and P. Tierno, Nano Lett. 19, 433 (2019).10.1021/acs.nanolett.8b04248Search in Google Scholar PubMed

[61] K. Misiunas and U. F. Keyser, Phys. Rev. Lett. 122, 214501 (2019).10.1103/PhysRevLett.122.214501Search in Google Scholar PubMed

[62] A. V. Straube and P. Tierno, EPL 103, 28001 (2013).10.1209/0295-5075/103/28001Search in Google Scholar

[63] A. Ryabov, D. Lips, and P. Maass, J. Phys. Chem. C 123, 5714 (2019).10.1021/acs.jpcc.8b12081Search in Google Scholar

[64] K. Jain, R. Marathe, A. Chaudhuri, and A. Dhar, Phys. Rev. Lett. 99, 190601 (2007).10.1103/PhysRevLett.99.190601Search in Google Scholar PubMed

[65] F. Slanina, Phys. Rev. E 80, 061135 (2009).10.1103/PhysRevE.80.061135Search in Google Scholar PubMed

[66] F. Slanina, J. Stat. Phys. 135, 935 (2009).10.1007/s10955-009-9743-zSearch in Google Scholar

[67] D. Chaudhuri and A. Dhar, EPL (Europhys. Lett.) 94, 30006 (2011).10.1209/0295-5075/94/30006Search in Google Scholar

[68] D. Chaudhuri, A. Raju, and A. Dhar, Phys. Rev. E 91, 050103 (2015).10.1103/PhysRevE.91.050103Search in Google Scholar PubMed

[69] J. K. Percus, J. Stat. Phys. 15, 505 (1976).10.1007/BF01020803Search in Google Scholar

[70] U. Seifert, Rep. Prog. Phys. 75, 126001 (2012).10.1088/0034-4885/75/12/126001Search in Google Scholar PubMed

[71] M. Šiler, T. Čižmár, A. Jonáš, and P. Zemánek, New J. Phys. 10, 113010 (2008).10.1088/1367-2630/10/11/113010Search in Google Scholar

[72] R. Di Leonardo, S. Keen, J. Leach, C. D. Saunter, G. D. Love, et al., Phys. Rev. E 76, 061402 (2007).10.1103/PhysRevE.76.061402Search in Google Scholar PubMed

[73] A. Curran, A. Yao, G. Gibson, R. Bowman, J. Cooper, et al., J. Biophotonics 3, 244 (2010).10.1002/jbio.201000003Search in Google Scholar PubMed

[74] M. Šiler, T. Čižmár, and P. Zemánek, Appl. Phys. Lett. 100, 051103 (2012).10.1063/1.3680234Search in Google Scholar

[75] V. Ambegaokar and B. I. Halperin, Phys. Rev. Lett. 22, 1364 (1969).10.1103/PhysRevLett.22.1364Search in Google Scholar

[76] A. Scala, Phys. Rev. E 86, 026709 (2012).10.1103/PhysRevE.86.026709Search in Google Scholar PubMed

[77] P. Maass, M. Dierl, and M. Wolff, in: Diffusive Spreading in Nature, Technology and Society (Eds. A. Bunde, J. Caro, J. Kärger, and G. Vogl) Springer International Publishing, Cham 2018, Chap. 9, p. 147.10.1007/978-3-319-67798-9_9Search in Google Scholar

Received: 2020-01-24
Accepted: 2020-02-14
Published Online: 2020-04-08
Published in Print: 2020-05-26

© 2020 Walter de Gruyter GmbH, Berlin/Boston

Downloaded on 7.9.2025 from https://www.degruyterbrill.com/document/doi/10.1515/zna-2020-0028/html?lang=en
Scroll to top button