You are on page 1of 9

Available online at www.sciencedirect.

com

ScienceDirect
Transportation Research Procedia 2 (2014) 87 – 95

The
The Conference
Conference on Pedestrian and
in Pedestrian and Evacuation
Evacuation Dynamics
Dynamics 2014
2014 (PED2014)
(PED2014)

The dynamics of waiting: the exclusive queueing process


Chikashi Arita a , Andreas Schadschneider b,∗
a Theoretische Physik, Universität des Saarlandes, 66041 Saarbrücken, Germany
b Institut für Theoretische Physik, Universität zu Köln, 50937 Köln, Germany

Abstract
The dynamics of pedestrian crowds has been studied intensively in recent years, both theoretically and empirically. However, in
many situations pedestrian movement is rather static, e.g. due to jamming near bottlenecks or queueing at ticket counters or the su-
permarket checkout. Classically such queues are described by a M/M/1 queue which neglects the internal structure (density profile)
of the queue by focussing on the queue length as the only dynamical variable. This is different in the exclusive queueing process
(EQP) which considers the queue on a microscopic level. It is equivalent to a totally asymmetric exclusion process (TASEP) of
varying length for which many exact results can be obtained. The EQP has a surprisingly rich phase diagram with respect to the
arrival probability α and the service probability β. The behaviour at the phase transition line is much more complex than for the
TASEP. It is nonuniversal and depends strongly on the update procedure used.
© 2014 The Authors. Published by Elsevier B.V. This is an open access article under the CC BY-NC-ND license

c 2014 The Authors. Published by Elsevier B.V.
(http://creativecommons.org/licenses/by-nc-nd/3.0/).
Peer-review under
Peer-review underresponsibility
responsibilityof of
Department
PED2014.of Transport & Planning Faculty of Civil Engineering and Geosciences
Delft University of Technology
Keywords: queueing theory; exclusion process; phase diagram; density profile; phase transition
PACS: 45.70.Vn, 02.50.Ey, 05.70.Fh, 89.75.-k
2000 MSC: 60K25, 90B20, 82C22

1. Introduction

Pedestrian dynamics usually considers moving pedestrians. However, from a practical point of view, we often have
to deal with situations where motion is almost impossible. In these situations quite generally queues are formed which
can have different structures. One of the simplest cases are queues at a supermarket checkout or ticket counter. Also
the behavior of crowds at bottlenecks might well be described by queueing dynamcis. Therefore a better understanding
of the dynamics of queues is desirable. It might help to optimize queueing processes, e.g. by minimizing the waiting
time. Queueing theory has originally been developed to describe problems of telecommunication (Erlang (1909)), but
has been applied in many other fields, such as traffic flow, biology and supply chains, see e.g. Medhi (2003), Saaty
(1961), Hopp and Spearman (2008) and Arita and Schadschneider (2014).
The M/M/1 queueing process is the simplest case in the classical queueing theory, which describes the dynamics
of a single queue. Arrival and service are modeled by Poisson processes (Medhi (2003); Saaty (1961)). Customers

∗ Corresponding author.
E-mail address: as@thp.uni-koeln.de

2352-1465 © 2014 The Authors. Published by Elsevier B.V. This is an open access article under the CC BY-NC-ND license
(http://creativecommons.org/licenses/by-nc-nd/3.0/).
Peer-review under responsibility of Department of Transport & Planning Faculty of Civil Engineering and Geosciences Delft University of Technology
doi:10.1016/j.trpro.2014.09.012
88 Chikashi Arita and Andreas Schadschneider / Transportation Research Procedia 2 (2014) 87 – 95

1
® ¯ C
¯
D
L 0 ® 1

Fig. 1. Definition of M/M/1 queue (left) and its phase diagram (right). When the arrival probability α is larger than service probability β, the queue
diverges (D). It converges (C) when α < β.

(= particles) arrive with probability α at the end of the queue and are serviced (= removed) with probability β at the
front of the queue (“server”) as illustrated in Fig. 1. The queue length Lt is the only variable, which we define as the
number of customers at time t including one at the server. The behavior of the length in the limit t → ∞ depends on
the relation between α and β:



⎨∞ for α ≥ β,
limt→∞ Lt  = ⎪
⎪ (1)
⎩ α(1−β) for α < β.
β−α

We call here the line separating the two phases (as shown in Fig. 1) “critical line”.
The M/M/1 queue neglects the spatial structure of queues and the particles in the queues do not interact with
each other. Thus the density along the queue is always constant. Although this works well in many cases, it seems
not suitable for many pedestrian queues. Therefore an extension of the classical M/M/1 queueing process has been
introduced recently in Arita (2009) and Yanagisawa et al. (2010), which is called the Exclusive Queueing Process
(EQP). It takes into account particle interactions through the excluded-volume effect and leads to nontrivial density
profiles of the queue. Since the excluded-volume effect was introduced by using the Totally Asymmetric Simple
Exclusion Process (TASEP), we shall first review the essential aspects of the TASEP in the next section. Then we
shall define and show results of the EQP which can be interpreted as a driven-diffusion model with varying system
length.

2. Totally asymmetric simple exclusion process

The Totally Asymmetric Exclusion Process (TASEP) is one of the paradigmatic models of nonequilibrium physics,
which is one of lattice gas models, see e.g. Schütz (2001); Blythe and Evans (2007); Schadschneider et al. (2010) for
reviews. The rule for the motion of particles is very simple: a particle at site j moves (rightward) to site j + 1 with
probability p if site j + 1 is not occupied by another particle. Let us consider hydrodynamic behavior of the density ρ
of the TASEP on Z Krapivsky et al. (2010). The following two types of waves are “time-invariant”, by rescaling the
position j as j/t.



⎨ρleft ( j < v s t),
Shock wave (ρleft < ρright ) ρ( j/t) = ⎪
⎪ (2)
⎩ρright ( j > v s t),



⎪ ρright ( j < t f (ρright )),


⎨ −1
Rarefaction wave (ρleft > ρright ) ρ( j/t) = ⎪
⎪ f ( j/t) (t f (ρleft ) < j < t f (ρright )), (3)



⎩ρleft ( j < t f (ρleft )).

Here the shock velocity v s is determined by

J(ρleft ) − J(ρright )
vs = , (4)
ρleft − ρright

and the function f is defined as f (ρ) = dJ


dρ , where the fundamental diagram J(ρ) (the relation between the current J
and the density ρ) can be derived by analysis e.g. on periodic boundary condition. The fundamental diagram depends
Chikashi Arita and Andreas Schadschneider / Transportation Research Procedia 2 (2014) 87 – 95 89

1
p p
MC
® ¯ ¯ LD
¯c
HD
1 2 3 ... L 0 ®c 1
®


Fig. 2. Definition of the TASEP (left) and its exact phase diagram for the parallel update (right). The critical points are αc = βc = 1 − 1 − p.

on the update rules (which we shall explain later), for example


⎧ √


⎪ 1− 1−4pρ(1−ρ)
⎨ (parallel update),
J(ρ) = ⎪

2 (5)

⎩ pρ(1−ρ) (backward-sequential update).
1−pρ

The “open boundary condition” on the L-site chain is also a well-studied case (Fig. 2). At the boundaries j = 1
and j = L the TASEP is coupled to particle reservoirs. If site 1 is empty, a particle is inserted with probability α. A
particle on site L is removed from the system with probability β.
As a function of the boundary parameters α and β three phases can be distinguished (Fig. 2). In the Low-Density
(LD) phase the current J = n j (1 − n j+1 ) through the system depends only on the input rate α. The input is less
efficient than the transport in the bulk of the system or the output and therefore dominates the behavior of the whole
system. In the High-Density (HD) phase the output is the least efficient part of the system and the current depends
only on β. In the Maximal-Current (MC) phase, input and output are more efficient than the transport in the bulk of
the system. Here the current has reached the largest possible value corresponding to the maximum of the fundamental
diagram (5). The phase diagram was first derived rigorously in the works by Schütz and Domany (1993) and Derrida
et al. (1993) for the continuous-time case. In particular, Derrida et al. (1993) introduced matrices to construct the
exact stationary solution. The basic idea is to make a matrix product with replacing occupied and unoccupied sites by
matrices. This matrix product Ansatz has been widely applied to many one-dimensional interacting particle systems,
see the review by Blythe and Evans (2007). The matrix product representation for the TASEP with parallel update was
by Evans et al. (1999), see also de Gier and Nienhuis (1999) for a slightly different approach. The phase transitions can
be also understood by physical arguments using the domain wall theory for driven systems developed by Kolomeisky
et al. (1998). In this technique, where the phases are distinguished based on the sign of the shock velocity (4). The
collective velocity given by the function f is relevant for the stability of the phases.

3. Exclusive queueing process

The Exclusive Queueing Process (EQP) is defined on a semi-infinite chain where sites are labeled by natural
numbers from right to left (Fig. 3). The dynamics of the model is defined as follows:

(i) input: A new particle is inserted with probability α on the site just behind the last particle in the queue. If there
is no particle in the system, a new particle is inserted directly to the site 1 with the same probability.
(ii) hopping: Particles behind an empty site move forward with probability p
(iii) output: A particle at site 1 is serviced (i.e. removed) with probability β.

For the parallel update these rules are applied simultaneously to all sites. In the case of backward-sequential update,
first (i) and (iii) are carried out. Then (ii) is applied sequentially to all sites starting at site j = 1. The dynamics of the
particle hopping is the same as in the TASEP introduced in Sec. 2.
We define the length L of the queue as the position of the last particle, and thus a new particle is inserted at site
L + 1. Note that this boundary condition for the left end (i) is different from the TASEP case, whereas (iii) is the same
as the TASEP. Therefore the EQP can be interpreted as a TASEP of variable length.
By changing the input and output probabilities α and β the EQP shows boundary-induced phase transitions, which
are classified according to different criteria:
90 Chikashi Arita and Andreas Schadschneider / Transportation Research Procedia 2 (2014) 87 – 95

® p p
¯

L ... 3 2 1

Fig. 3. Definition of the Exclusive Queueing Process (EQP). The length L is defined by the position of the last (leftmost) particle.

• Queueing classification – convergent or divergent queue length (see Sec. 4.1).


In the convergent phase (C) the average queue length converges to a finite value for long times. In the divergent
phase (D), the average length grows infinitely, being proportional to time t. The phase diagram of the M/M/1
case (Fig. 1) is changed due to the excluded-volume effect (Arita and Schadschneider (2011b)).
• TASEP classification – form of the outflow (see Sec. 4.2).
By definition, the inflow of particles is given by α. However, the outflow is not identical to β because the last
site (server) can be empty. The form of the non-stationary flow Jout is identical to the form for the MC or HD
phases of the open TASEP (see Arita and Schadschneider (2011a)) which is natural since the same boundary
condition for the right end is imposed for both the EQP and the open TASEP. In the maximal current phase, the
current Jout of particles going through the right end is independent of both α and β. In the high-density phase
the current depends only on β, but is independent of α.
• Classification according to density profile (see Sec. 4.3).
The divergent phase can be divided into subphases according to the number of plateaus in the density profile
(Arita and Schadschneider (2012)). The rescaled profile has the form of a rarefaction wave (Krapivsky et al.
(2010)).

4. Phase diagram of the EQP

4.1. Convergent and divergent phases

For the parallel EQP the stationary state can be obtained exactly using the matrix product Ansatz (Arita and
Yanagisawa (2010)). Due to the close relation with the parallel TASEP some exact results for the latter can be used
for the parallel EQP as well. However, a similar reasoning does not apply to the EQP with backward update although
the open TASEP with backward update has a matrix product stationary state (Evans (1997), Rajewsky et al. (1996),
Rajewsky et al. (1998)).
From the exact solution of the parallel EQP it can be shown that for


⎪ β(p−β)
⎨α < αc = p−β

√2
(β ≤ βc ),


⎪  (6)
⎩α ≤ αc = 1− 1−p
(β > βc = 1 − 1 − p)
2

a proper stationary state exists (Arita and Yanagisawa (2010)). The parameter region α < αc is called “convergent
phase” since in this region the queue length and the number of particles converge to finite values
αp(R − p + 2(1 − α)) α(1 − α)(p − 2αp + R)
L∞  = , N∞  = (7)
R(R − p + 2(1 − α)β) R(R − p + 2(1 − α)β)

with R = p(p − 4α(1 − α)).
 diverge. The critical line is given by αc (β). It
When α ≥ αc , the average queue length and the number of particles
consists of two parts, a straight-line part α = αc and β > βc = 1 − 1 − p, and a curved part αc (β) for β ≤ βc .
For the backward case the exact stationary state is not known. The phase diagram can be determined by Monte
Carlo simulations which lead to the conjecture


⎪ β(p−β)
⎨α < αc = p(1−β)
⎪ √
(0 < β ≤ βc )

⎪ , (8)

⎩α < αc = (1− 1−p)2
(βc < β < 1)
p
Chikashi Arita and Andreas Schadschneider / Transportation Research Procedia 2 (2014) 87 – 95 91

600 1200

500
MC-C 1000 MC-D
®=0.2 ®=0.75
400 ¯=0.8 800 ¯=0.8
hL t i hL t i
p=0.84 p=0.84
300 600
hNt i hNt i
200 400

100 200

0 1000 2000 3000 4000 0 200 400 600 800 1000


t t
400 1000
HD-C HD-D
®=0.2 800 ®=0.75
300
¯=0.4 hL t i ¯=0.4
hL t i p=0.84 600
p=0.84
200
hNt i hNt i
400

100
200

0 1000 2000 3000 4000 5000 6000 0 200 400 600 800 1000
t t

Fig. 4. The length and the number of particles vs. time for the parallel EQP. The markers × and  present data for Nt  and Lt  obtained from
Monte Carlo simulations, where 1000 samples are averaged. The solid lines correspond to Eqs. (9), (11) with (10), (13).

for the convergent phase. Explicit forms of the average values like (7) are not known up to now.

4.2. The outflow

The non-stationary properties of the EQP were first analyzed in Arita and Schadschneider (2011a) by a similar
argument to the domain wall theory Kolomeisky et al. (1998) for the TASEP.
The inflow Jtin of the EQP is always α due to the fact that the site where particles enter is by definition never
blocked. Assuming that the outflow Jout = βτ1  is constant, particle conservation implies
Nt  = (Jin − Jout )t + N0 . (9)
Under this assumption, the average queue length diverges (converges) linearly in the divergent (convergent) phases
(Fig. 4). From Monte Carlo simulations, we found that the outflow Jout is given as
⎧ β(p−β) ⎧ β(p−β)


⎪ (β ≤ βc ), ⎪

⎪ (β ≤ βc ),
⎨ (p−β√2 ) ⎨ p(1−β)

parallel : Jout = ⎪
⎪ , backward : J = ⎪
⎪ (10)

⎩ (1− 1−p)
(β > βc ),
out

⎩ (1− 1−p)2
(βc < β).
2 p

In analogy to the TASEP (Sec. 2) these phases are called High-Density (HD) phase for β ≤ βc and Maximal-Current
(MC) phase for βc < β. Note that the form of the outflow is identical to the critical value αc defined in Eqs. (6), (8).
The phase diagram of the EQP can now be understood as follows: when Jin < Jout , the number of particles increases
(resp. decreases for Jin > Jout ). Assuming that the global density Nt /Lt  = ρ is constant, the queue length shows an
analogous behavior (see (9)),
Lt  = Vt + L0 , V = (Jin − Jout )/ρ. (11)
In Fig. 4, the solid lines for Lt  are drawn under the assumption that ρ is the solution to J(ρ) = Jout . This assumption
is not always true in the divergent phase, which can be observed in Fig. 4. This fact implies that the global density
profile is not always constant in the divergent phase, which we shall see in the next subsection. On the other hand, in
the convergent phase, Eqns. (9), (11) become no longer true and the outflow becomes α after a “shock” (i.e. the left
end of the density profile) reaches the server.
Based on the queueing and TASEP classifications, the parameter space can be divided into 4 regions, the MC-C,
MC-D, HD-C and HD-D phases. In the divergent phases subphases can be distinguished based on the form of the
density profile.
92 Chikashi Arita and Andreas Schadschneider / Transportation Research Procedia 2 (2014) 87 – 95

MC-C MC-D-II MC-D-III

¯
1 { 1{ p
1 HD-D-II
HD-C
HD-D-III
HD-D-I

x 0 1 { 1{ p p 1
V 0 2 ®
1 1

MC-C MC-D-II
MC
-C MC-D-II MC-D-III
¯ ¯
1 { 1{ p
HD-D-II
HD-C
p { p(1 { p) 1 { 1{ p
HD-D-I HD HD-D-II
-C HD-D-III
HD-D-I
2 2 p
0 (1 { 1 { p ) 1 0 (1 { 1 { p ) 1
p p 1 {p
® ®

Fig. 5. A schematic picture for the density profile in the divergent phase (top-left), where x is the rescaled position j/t, and the phase diagrams
of the EQP with parallel (top-right) and backward-sequential dynamics for p < 1/2 (bottom-left) and p > 1/2 (bottom-right). According to the
injection probability (rate) α, the rarefaction wave is “cut” by the leftmost particle (x = V) and the server (x = 0). The end of the queue can be
in three different regimes, (plateau at density ρ < 1, regime of increasing density or plateau at density ρ = 1), which defines the regions I, II, III.,
respectively. The density profile shown here belongs to II.

4.3. Density profiles and subphases of the divergent phase

From simulation results, the density profile in the divergent phase is well described by cutting the rarefaction wave,
see Fig. 5. More precisely, the density of site j (Vt > j > 0) at time t is written as ρ(− j/t) as in (3) where the negative
sign is due to the fact that the sites are labeled from right to left for the EQP. V is the velocity of the left end, i.e.
Lt  Vt. ρleft is given by

ρleft = 1 (12)

which can be understood by considering the extreme case where α ≈ 1, β ≈ 0 and p ≈ 0. In this case, new particles
are always injected to the system, and newly injected particles cannot move immediately. The density ρright at the
service end is given by
⎧ p−β ⎧ p−β

⎪ ⎪

⎪ (β ≤ βc )
⎨ p−β2 (β ≤ βc ) ⎨ p(1−β)

parallel : ρright =⎪
⎪ , backward : ρright =⎪
⎪ (13)
⎩1 (β > βc ) ⎪
⎩ 1− p1−p (β > βc )
2

which is compatible with (10) via the fundamental diagram (5). According to the values of parameters α, β, p, the
position at which the profile is cut off changes. Depending on whether the cut-off is in the right plateau ρright , the
non-constant region desribed by f −1 or the left plateau ρleft , three different regimes can be distinguished. Thus the
velocity V has three different analytic forms, see Arita and Schadschneider (2012, 2014). In the phase diagrams in
Fig. 5, they are distinguished by roman numbers I, II and III.
Chikashi Arita and Andreas Schadschneider / Transportation Research Procedia 2 (2014) 87 – 95 93

500 500
(p,¯)
(0.84,0.3)
(0.64,0.2) 100 »t 1/2
100 (0.36,0.1) 1/2
50 »t (p,¯) (0.84,0.8)
10 (0.84,0.3) (0.64,0.7)
hLt i hLt i (0.64,0.2) (0.36,0.6)
10 1/4 (0.36,0.1) (0.19,0.55)
»t
5 1
(0.64,0.7)
(0.84,0.8) (0.36,0.6)
1 3 0.1 3
10 10 4
105 106 10 104 105 106
t t
Fig. 6. Time-dependence of average system length Lt  on the critical line for parallel dynamics (left) and backward dynamics (right).

To summarize, we have found up to 5 subphases in the divergent case, according to the classification based on the
forms of the outflow and the velocity, as shown in Fig. 5. The shape of the global density profile changes depending
on the parameters.

4.4. Critical line: Non-universal behavior

In the divergent phase the average length Lt  and the average number of particles Nt  diverges linearly in time.
On the phase transition line separating the convergent and divergent phases the growth is slower than linear, i.e.
Lt  = O(tγ ), Nt  = O(tγ ) with an critical exponent γ smaller than 1.
Fig. 6 shows the time-dependence of the average system length obtained by Monte Carlo simulations. As one can
observe in these log-log plots, the slopes depend both on the update type and the location on the critical line (curved
part β < βc or straight part β > βc ). Depending on the update rule, the exponents have different values:

⎧ ⎪

⎪1/2 (for β < βc ),


⎨1/2 (for β < βc ) ⎪


parallel : γ=⎪ ⎪ , backward : γ=⎪⎪g(p) (for β > βc , p < pc ) (14)
⎩1/4 (for β > βc ) ⎪


⎩0 (for β > βc , p > pc )
with some function g(p) ∈ (0, 1/4), whose explicit form is not known. The nonuniversal behavior of the backward
case and the existence of the critical point pc for the backward case has been tested by simulations (t  106 , averaged
over up to 106 samples, see Arita and Schadschneider (2013) for details). We think that this is the most reasonable
interpretation, although one could consider other scenarios on the straight part of the critical line α = αc , β > βc for
the backward case, e.g. a length that always converges, but with extremely slow relaxation to the stationary length for
small p.

5. Interacting queues: EQP with Langmuir kinetics

The most natural application of the EQP is queueing of pedestrians, e.g. at a supermarket checkout. This can
be generalized in a straightforward way to multiple queues where customers might jump from one queue to another.
Focussing on the dynamics of one of the interacting queues, customers changing from or to other queues can be
interpreted as particles that are created or annihilated in the queue. In physics such dynamics is usually called Lang-
muir kinetics. Its effects have previously been studied for the TASEP leading to the TASEP-LK which is relevant for
several biological processes like intracellular transport (see e.g. Parmeggiani et al. (2003); Evans et al. (2003)). The
TASEP-LK has a rich phase diagram. Therefore we have extended the EQP with parallel update in a similar way
(EQP-LK). The EQP-LK can be interpreted as an effective model for interacting queues where the other queues are
considered to act as reservoir for the EQP-LK.
In the presence of Langmuir kinetics, particles are detached with probability ωD , and for each empty site j(≤ L)
a particle is attached with probability ωA (see Fig. 7). As in the TASEP-LK (Parmeggiani et al. (2003) and Evans
et al. (2003)), the attachment and detachment probabilities are scaled as ωA = ΩA /L and ωD = ΩD /L which leads to
94 Chikashi Arita and Andreas Schadschneider / Transportation Research Procedia 2 (2014) 87 – 95

00 5000
1 00
1000
L0
1000

500 500
50
500

500
1000
!D !A !D !D !A !D 350
100
® p p

hLt i
hLt i
0

hhLt i
5
¯

10

250
0.05 0.5
0 t 6¢104

L 3 2 1

1
0 104 2¢104 0 104 2¢104
t t

Fig. 7. Exclusive queueing process with Langmuir kinetics (left), and behaviors of Lt  starting from various initial lengths, for (p, α, β, ΩA , ΩD ) =
(0.8, 0.2, 0.2, 0.2, 0.35) (middle) and (p, α, β, ΩA , ΩD ) = (0.8, 0.3, 0.2, 0.1, 0.9) (right). We have set the initial density as ΩA /(ΩD + ΩA ), and
averaged over 103 samples. In the middle figure, we observe all the average lengths with initial lengths L0 = 0, 100, 350, 500, 1000 converge to a
stationary value. In the right figure, we find that the behavior of the average length depends on the initial length. Furthermore in the inset of (b),
the length exhibits non-monotonic behavior.

a competition between bulk and boundary dynamics. We do not impose attachment and detachment when the system
length is 0. Note that, in contrast to the TASEP-LK, the system length Lt of the EQP-LK varies, and thus probabilities
ωA , ωD depend on the current state. In each time step, first the configuration is updated according to the rule of the
EQP with parallel update. Then the Langmuir kinetics is applied.
Preliminary studies have found that the EQP-LK has surprising properties (Schultens et al.). It shows a strong
dependence of the behavior on the initial condition, see Fig. 7. In fact, in a certain parameter region, some samples
appear to converge to a finite length whereas other samples appear to diverge within the simulation time (Schultens
et al.; Arita and Schadschneider (2014)). The percentage of apparently divergent samples depends strongly on the
initial length L0 = Lt=0 of the queue. It is very small for small L0 and becomes large for large L0 .
This surprising behavior is related to the length dependence of the attachment and detachment probabilities. Once
the queue has become short it is difficult to escape from Lt = 0 after reaching Lt = 0 since the detachment rate is
large. Thus, starting from a short queue L0 < L∗ , the length tends to remain finite. On the other hand, starting from
long queues L0 > L∗ the chain tends to grow further because the detachment rate is small. These two observations
contradict the general theory of Markov processes. However, we can interpret the convergence from short queues as
“quasi stationary,” and a very long time is required to reach e.g. L = 500. In this sense, the ergodicity of the EQP-LK
is effectively broken.

6. Discussion

The Exclusive Queueing Process (EQP) can be considered as a minimal model of pedestrian queues which takes
into account the internal dynamics of the queue. We have found that the EQP has a rich phase diagram. Surprisingly,
it shows a strong dependence of its phase diagram and critical properties on the update scheme. This is rather different
from the TASEP with a fixed system length. The order of the phase transition between the diverging and converging
phases can also be different. The strong dependence of the phase diagram on the update type indicates that the
queueing dynamics depends sensitively on the details of the update. This indicates that appropriate control measures
can be used for optimization, e.g. a reduction of the waiting time.
A generalization where the probability of moving depends on the gap to the next customer in front was studied in
Behlau (2013). This might be realistic e.g. for queues at an airport check-in where the passengers have to pick up their
luggage when moving forward. Since this is uncomfortable they typically wait until a critical gap to the preceding
passenger has opened. Recently, de Gier and Finn (2014) have introduced a queueing process with two types of
customers. In this Priority Exclusion Process (PEP) high priority customers can overtake low priority customers.
The EQP-LK which is an EQP with additional Langmuir kinetics is a single chain queueing process that can be
interpreted as an effective model for a multi-chain process. The interaction between the chains through exchange
of particles is modeled by Langmuir kinetics which allows the change of the particle number within the bulk of
Chikashi Arita and Andreas Schadschneider / Transportation Research Procedia 2 (2014) 87 – 95 95

the queue. We have found that this kind of interaction between queues might lead to “unpredictable” results with
effectively broken ergodicity.
Of course, the EQP can also be extended to a genuine multi-chain model. Here different rules for the changing of
pedestrians from one queue to another can be implemented, depending on the application. Of interest are also merging
queues which e.g. allows to compare the effectivity of different methods of queue control.
One advantage of the EQP and its relatives for applications is that it is an intrinsically microscopic model where
the different “units” can be distinguished. Therefore they can have different properties (e.g. average velocities) in a
natural way. We are currently working on such extension. Results will be reported elsewhere.

Acknowledgements

AS is partially supported by Deutsche Forschungsgemeinschaft (DFG) under grant “Scha 636/8-1”. The authors
are grateful to Christoph Behlau, Christian Borghardt, Martin R. Evans, Alexander Lück, Kirone Mallick, Ludger
Santen, Gunter Schütz, Christoph Schultens and Daichi Yanagisawa for useful discussions.

References

Arita, C., 2009. Queueing process with excluded-volume effect. Phys. Rev. E 80, 051119.
Arita, C., Schadschneider, A., 2011a. Dynamical analysis of the exclusive queueing process. Phys. Rev. E 83, 051128.
Arita, C., Schadschneider, A., 2011b. Exact dynamical state of the exclusive queueing process with deterministic hopping. Phys. Rev. E 84,
051127.
Arita, C., Schadschneider, A., 2012. Density profiles of the exclusive queueing process. J. Stat. Mech. , P12004.
Arita, C., Schadschneider, A., 2013. Critical behavior of the exclusive queueing process. EPL 104, 30004.
Arita, C., Schadschneider, A., 2014. Exclusive queueing processes and their application to traffic systems. Math. Model Meth. Appl. Sci. Submitted.
Arita, C., Yanagisawa, D., 2010. Exclusive queueing process with discrete time. J. Stat. Phys. 141, 829.
Behlau, C., 2013. Bachelor thesis. Cologne University.
Blythe, R., Evans, M., 2007. Nonequilibrium steady states of matrix product form: a solver’s guide. J. Phys. A: Math. Gen. 40, R333.
Derrida, B., Evans, M.R., Hakim, V., Pasquier, V., 1993. Exact solution of a 1d asymmetric exclusion model using a matrix formulation. J. Phys.
A 26, 1493.
Erlang, A., 1909. The theory of probabilities and telephone conversations. Nyt Tidsskrift for Matematik B 20, 33.
Evans, M., 1997. Exact steady states of disordered hopping particle models with parallel and ordered sequential dynamics. J. Phys. A 30, 5669.
Evans, M., Rajewsky, N., Speer, E., 1999. Exact solution of a cellular automaton for traffic. J. Stat. Phys. 95, 45.
Evans, M.R., Juhász, R., Santen, L., 2003. Shock formation in an exclusion process with creation and annihilation. Phys. Rev. E 68, 026117.
de Gier, J., Finn, C., 2014. Exclusion in a priority queue. J. Stat. Mech. , P07014.
de Gier, J., Nienhuis, B., 1999. Exact stationary state for an ASEP with fully parallel dynamics. Phys. Rev. E 59, 4899.
Hopp, W., Spearman, M., 2008. Factory Physics. McGraw-Hill, Boston.
Kolomeisky, A., Schütz, G., Kolomeisky, E., Straley, J., 1998. Phase diagram of one-dimensional driven lattice gases with open boundaries. J.
Phys. A: Math. Gen. 31, 6911.
Krapivsky, P., Redner, S., Ben-Naim, E., 2010. A Kinetic View of Statistical Physics. Cambridge University Press, Cambridge.
Medhi, J., 2003. Stochastic Models in Queueing Theory. Academic Press, San Diego.
Parmeggiani, A., Franosch, T., Frey, E., 2003. Phase coexistence in driven one-dimensional transport. Phys. Rev. Lett. 90, 086601.
Rajewsky, N., Santen, L., Schadschneider, A., Schreckenberg, M., 1998. The asymmetric exclusion process: Comparison of update procedures. J.
Stat. Phys. 92, 151.
Rajewsky, N., Schadschneider, A., Schreckenberg, M., 1996. The asymmetric exclusion model with sequential update. J. Phys. A 29, L305.
Saaty, T., 1961. Elements of Queueing Theory With Applications. Dover Publ.
Schadschneider, A., Chowdhury, D., Nishinari, K., 2010. Stochastic Transport in Complex Systems: From Molecules to Vehicles. Elsevier Science,
Amsterdam.
Schultens, C., Borghardt, C., Schadschneider, A., Arita, C., . In preparation.
Schütz, G., 2001. Exactly solvable models for many-body systems far from equilibrium, in: Domb, C., Lebowitz, J. (Eds.), Phase Transitions and
Critical Phenomena, Vol 19.. Academic Press, San Diego.
Schütz, G., Domany, E., 1993. Phase transitions in an exactly solvable one-dimensional exclusion process. J. Stat. Phys. 72, 277.
Yanagisawa, D., Tomoeda, A., Jiang, R., Nishinari, K., 2010. Excluded volume effect in queueing theory. JSIAM Lett. 2, 61.

You might also like