• Aucun résultat trouvé

Microscopic and macroscopic models for vehicular and pedestrian flows

N/A
N/A
Protected

Academic year: 2021

Partager "Microscopic and macroscopic models for vehicular and pedestrian flows"

Copied!
29
0
0

Texte intégral

(1)

HAL Id: hal-02491071

https://hal.archives-ouvertes.fr/hal-02491071

Submitted on 25 Feb 2020

HAL is a multi-disciplinary open access

archive for the deposit and dissemination of

sci-entific research documents, whether they are

pub-lished or not. The documents may come from

teaching and research institutions in France or

abroad, or from public or private research centers.

L’archive ouverte pluridisciplinaire HAL, est

destinée au dépôt et à la diffusion de documents

scientifiques de niveau recherche, publiés ou non,

émanant des établissements d’enseignement et de

recherche français ou étrangers, des laboratoires

publics ou privés.

To cite this version:

(2)

Chapter 1

Microscopic and macroscopic models

for vehicular and pedestrian flows

Massimiliano D. Rosini

Dipartimento di Matematica e Informatica, Universit`a degli Studi di Ferrara, Via Machiavelli 35, 44121 Ferrara, Italy

and

Uniwersytet Marii Curie - Sk lodowskiej,

Plac Marii Curie - Sk lodowskiej 1, 20 - 031 Lublin, Poland, massimilianodaniele.rosini@unife.it

In this chapter we develop and study microscopic and macroscopic models for car and pedestrian flows. After a general introduction on the basic features of the flows under consideration, we develop our models. We first introduce at the microscopic level few basic and well established interaction rules. We then encode them in analytical models and obtain at the microscopic level a first order follow-the-leader model and at the macroscopic level a scalar conservation law (an hyperbolic partial differential equation). These models are then generalized to the cases of heterogeneous flows and flows across point-wise inhomogeneities characterized by reduced capacity, such as doors or toll gates. At last we also consider a crowd in panic situations. Some examples are proposed and serve as benchmarks to point out the ability of the models to reproduce realistic scenarios. The solutions to these examples are constructed by applying a wave-front tracking algorithm based on Riemann solvers. We also show how macroscopic models can be obtained from a microscopic model as many-particle limit.

Contents

1. Introduction . . . 2

2. A first order microscopic model . . . 4

3. A first order macroscopic model . . . 6

3.1. Riemann problem . . . 7

3.2. Wave-front tracking . . . 10

4. Links between microscopic and macroscopic models . . . 13

5. Heterogeneous traffic . . . 15 5.1. Microscopic model . . . 15 5.2. Macroscopic model . . . 16 6. Bottlenecks . . . 19 6.1. Microscopic model . . . 19 6.2. Macroscopic model . . . 20

7. A macroscopic model for pedestrian flows with panic . . . 24

References . . . 25

(3)

1. Introduction

In this chapter we consider both car and pedestrian flows. To some extent, they are well known from every day life experiences. For this reason, there is no need to give a technical description, as some general insights are already well-known to the reader. Nevertheless, there are some counter intuitive features that appear as paradoxes and need some comments.

This chapter aims at indicating research perspectives on car and pedestrian flows motivated by insights into their complexity features. In fact drivers and crowds have to be looked at as living complex systems, namely systems of agents who are able to choose actively their individual dynamics. As a consequence, the emergent collective be-havior and self-organization are not generated by a pas-sive reaction of the whole system to external stimuli, but they stem from one-to-one, or at most one-to-few, inter-actions among the agents.

Consider the evacuation of a room through an exit door as in Fig. 1. The presence of an exit door has a

Fig. 1.: Capacity drop of an exit door. The dots represent a crowd evacuating a 2D room with a sin-gle exit door. The dots marked in gray are clogging the exit door.

big impact on the evacuation time. An important prop-erty of the door is its capacity Q, which gives the maxi-mum number of pedestrians that can flow through it in a given time interval. Despite the common belief, the capacity of a door is not constant. In fact high densi-ties upstream of an exit door may determine its capacity drop. This is clear by observing that two or more pedes-trians may block each other and not move further. In real situations these kind of clogging are avoided by plac-ing a further door or column in the upstream vicinity of the door. Indeed, the presence of such obstacles may pre-vent the crowd density from reaching high values and may

help to minimize the evacuation time, since in a moder-ate density regime the full capacity of the exit door can be exploited. The decrease of the evacuation time may seem unexpected, as some of the pedestrians are forced to choose a longer path to reach the exit. This is usually referred to as Braess’ paradox for pedestrians.1

A further phenomenon linked to the capacity drop of the exit doors is the so-called faster-is-slower effect.2 Faster-is-slower effect refers to the jamming and clogging at the doors, that result in an increase of the evacua-tion time when the velocity of a crowd is high.3,4 This partly counter-intuitive phenomenon is explained by the fact that if a huge number of pedestrians reach the exit door faster, then the capacity of the exit door falls earlier. In fact, the importance of these paradoxes stems from the fact that they serve as benchmarks to point out the ability of the models to reproduce realistic scenarios. We aim to develop microscopic and macroscopic models able to reproduce both the Braess’ paradox and the faster-is-slower effect. In the literature, there are two main ap-proaches to do so. The first just couples the equations for the evolution of the crowd, with boundary conditions cor-responding to obstacles, such as walls and doors, see for instance Ref. [5–7]. The second approach adds a further condition which accounts for the capacity of the bottle-necks, as in the present chapter and in Ref. [8–11].

In Section 6 we reproduce the effects of an exit doors by applying the theory of point constraint on the flow first introduced in Ref. [8]. More precisely, in Examples 22 and 23 we reproduce the Braess’ paradox and the faster-is-slower effect.

Since traffic dynamics are essentially microscopic, it is much easier to physically motivate the microscopic rather than some macroscopic assumptions. Thus, we first encode the interacting rules in a first order micro-scopic follow-the-leader (FTL) model. We then construct the first order macroscopic Lighthill-Whitham-Richards (LWR) model12,13taking advantage of analogies with the first order FTL model. We also recall that in Ref. [14] it is rigorously proved that LWR model can be obtained as many-particle limit of FTL model approximations, sim-ilar to taking the hydrodynamical limit of Boltzmann equations

(4)

charac-terized in terms of length and maximal speed, we show how to deduce at the microscopic level a FTL model and at the macroscopic level the Aw-Rascle-Zhang (ARZ) model.15,16 We point out the analogies between the two models and recall that in Ref. [17] it is rigorously proved that ARZ model can be obtained as many-particle limit of FTL model approximations. We then consider the presence of pointwise inhomogenities. From the model-ing point of view, pointwise inhomogenities correspond, for instance, to a narrow exit in crowd modeling, or to a toll gate in vehicular traffic. This leads us to consider point constraints on the flow. At the macroscopic level this is encoded by simply adapting the velocity function. At the macroscopic level, we rely on the paradigm of pas-sage to the limit from microscopic to macroscopic level proposed in Ref. [18]. In particular we give two exam-ples that show the ability of the model to reproduce two paradoxes linked to the presence of bottlenecks, namely Braess’ paradox and faster-is-slower effect. At last we also develop a macroscopic model for crowd dynamics in case of panic,8 which is able to reproduce the arise of panic.

Vehicular and pedestrian flows share most of their main features. The main difference is the panic behavior that characterizes only crowd dynamics. Panic is one of the main cause of crowd accidents. We make clear some basic features of such events. First, during a crowd acci-dent compressive forces occur from both vertical stacking and horizontal pushing, moreover the domino effect of people leaning against each other has to be taken into account. Evidence of bent steel railings after fatal crowd accidents show that forces of more than 4, 500N occurred. For instance, at Ibrox Park soccer stadium in UK, after the crowd accident occurred in 1971, the pile of bodies was 3m high, showing that people were moving over each other. Second, deaths from human stampedes occur pri-marily from compressive asphyxiation and not trampling. For instance after the crowd accident occurred at River-front Coliseum in Cincinnati, USA, on 1979, a line of bodies was found 9m from a wall. This indicates that crowd pressures came from both directions as rear ranks pressed forward and front ranks pushed off the wall.

Crowd accidents are caused by an irrational behavior of pedestrians due to panic. According to sociology scien-tists, panic is a maladaptive reaction of flight stemming from intense fear. Moreover, panic is infectious and easily

spreads to other people nearby. Such definition cannot be easily exploited in an analytical model. For this reason in this chapter panic simply corresponds to densities higher than (7 people)/m2. This choice stems from the fact that

we want to avoid deaths that result from panic, namely from high densities.

Crowd behaviour is not always the primary cause of accidents. Architects and city planners try to provide a safe environment for places of public assembly by using analytical models to determine and prevent congestion. A good example is the Jamarat Bridge, Saudi Arabia. Following the 2006 accident with 363 deaths, see also Ta-ble 1, the old bridge was demolished and construction be-gan on a new 4-level bridge. The new bridge is designed to accommodate 5 million pilgrims. It has 10 entrances and 12 exits to allow the flow of 300, 000 pilgrims per hour. It is also equipped with surveillance technology to help authorities intervene in case of any stampedes. As a result, so far no crowd accidents occur on the Jamarat Bridge since 2006.a

Table 1.: Some notable crowd accidents occurred at the Jamarat Bridge, Saudi Arabia.

Year 2006 2004 2001 1998 1997 1994 1990

Dead 363 249 35 118 22 266 1,426

It is therefore possible to avoid or to mitigate crowd accidents. Nowadays, all public entertainment venues are equipped with doors that open outwards using crash bar latches that open when pushed. Moreover, it is well known since ancient times, that a tall column placed in the upstream of the door exit may reduce the inter-pedestrian pressure, decreasing the magnitude of clogging and making the overall outflow higher and more regular, this is the so called Braess’ paradox that we have already mentioned.

Vehicular and pedestrian flows can be described at three levels: microscopic, mesoscopic and macroscopic. At the microscopic level we work in the framework of Newtonian mechanics and the dynamics of each

pedes-aIn 2015 around 1, 470 deaths occur in the camps close to the

(5)

trian moving under the action of the surrounding pedes-trians and of the environment is represented by an ordi-nary differential equation. Examples of microscopic mod-els are follow-the-leader modmod-els,19AI-based models,20 so-cial force model,21cellular automata models.22,23 At the mesoscopic level we work in the framework of kinetic the-ory of gases and introduce a Boltzmann-type evolution equation for the statistical distribution function of the position and velocity of the pedestrians, see for example Ref. [24–26] and references therein. At the macroscopic level we work in the framework of the hydrodynamics and the flow is described in terms of averaged quantities, which are assumed to be continuous and evolve according to partial differential equations. The first macroscopic modeling attempt is due to Hughes.27 Current macro-scopic models use non linear conservation laws with non classical shocks,8gas dynamics equations,28,29time evolv-ing measures30 and gradient flow methods.31

We concentrate our attention on microscopic and macroscopic models. On one hand there are a posteri-ori some advantages in using macroscopic models. For instance, it is very powerful in the description of queue tails as they correspond to the position of shock waves. Furthermore it is suitable with very large number of cars or pedestrians as it is less computationally intensive than microscopic models and can allow fast simulations also for large-scale traffic networks. Moreover it is easy to validate and implement due to the low number of param-eters. At last is it suitable to real time prediction, estima-tion, optimization and management. On the other hand, the number of cars or pedestrians is far lower than that of molecules, for instance, in gas dynamics. Therefore the continuum assumption is not justified and the macro-scopic formulation is not a priori justified. In this respect, a validation of the macroscopic approach is provided by the micro-macro limit rigorously proved in Ref. [14].

At the macroscopic level the flow of cars or pedestrians is seen as a unique stream, in analogy with the flow of fluids or gases, and its dynamics are described by means of aggregate variables, such as density, mean speed and flow. Hence macroscopic models describe the traffic as a fluid of “thinking” particles. However we have to keep in mind the following differences between the flow of cars or pedestrians and the flow of fluids:

(1) As we already noticed, the continuum assumption is

not justified in the framework of car or pedestrian flows, because the number of vehicles or pedestrians is far lower than that of molecules, hence vehicles or pedestrians do not form a continuum.

(2) A fluid particle is isotropic (i.e., responds to stim-uli from the surrounding particles), whereas a vehicle or a pedestrian is anisotrpic (i.e., responds only to frontal stimuli).

(3) Vehicles and pedestrians don’t have negative speed. (4) In vehicular and pedestrian flows there is no

conser-vation of momentum.

(5) A driver or a pedestrian is a living system since he/she has a personality (e.g., aggressive or timid) and this remains unchanged by motion (e.g., a slow driver is unaffected by its interaction with faster drivers passing it, or queueing behind it).

This chapter is structured as follows. In the next sec-tion we construct a first order microscopic model start-ing from basic considerations. In Section 3 we construct a first order macroscopic model, that is the LWR model, and show how the wave-front tracking algorithm can con-struct a weak solution on the basis of a Riemann solver. We give two examples and show that the obtained weak solutions are realistic. We conclude the section by point-ing out the links between the two first order models. In Section 5 we generalize the previous two models to the case of an heterogeneous traffic and obtain a second or-der FTL model and the ARZ model. As in the previous section, we give an example and show that the obtained weak solution is realistic. In Section 6 we take into con-sideration the effects of pointwise bottlenecks, both at microscopic and macroscopic levels. We reproduce the Braess’ paradox and the faster-is-slower effect with two examples. We also point out further possible generaliza-tions. At last in Section 7 we propose a way to construct a macroscopic model for the evacuation of a narrow cor-ridor in the case of panic behavior.

2. A first order microscopic model

(6)

Consider the simplest case of a straight one-lane ho-mogeneous road, along which overtaking is not allowed. The road can be parametrized by x ∈ R. If we have n cars, then traffic is described by the position of each car at each time, namely by

x1(t) , x2(t) , . . . , xi(t) , . . . , xn(t) ∈ R.

More precisely xi+1(t) < xi(t) and xi(t) corresponds to

the rear of the car. It is not restrictive to assume that cars move with positive speed, see Fig. 2.

xn(t) xi(t) x2(t) x1(t)

x . . . .

Fig. 2.: A car traffic.

The simplest case is when the traffic is given by just one car, i.e. n = 1. In this case such car has free road, it moves at the maximum speed vmax> 0 and its equation

of motion is

˙

x1(t) = vmax, (1)

where ˙x1 is the time derivative of position x1.

Equa-tion (1) is a first order linear ordinary differential equa-tion (ODE). In order to know the (exact) posiequa-tion of the car at each time, it is necessary to know its position at an arbitrarily fixed time. If for instance we know that

x1(t0) = ¯x1, (2)

where ¯x1 ∈ R is the position of the car at time t0 ∈ R,

then its position at time t ∈ R is given by

x1(t) = ¯x1+ (t − t0) vmax. (3)

We just solved the Cauchy problem (1), (2), indeed its (unique) solution is given by (3).

We consider now a second car which at time t0 is at

¯

x2∈ R, namely

x2(t0) = ¯x2. (4)

It is not restrictive to assume that ¯x2< ¯x1(otherwise we

just change the labels of the cars). More precisely, if both cars have length `, then we assume that

¯

x1− ¯x2> `.

In principle we could think that also the second car moves at the maximum speed vmax. This would not lead

to any contradiction as cars will not overtake each other.

What’s more, this would definitely optimize the traffic flow! However this is in contrast with reality, as (non-autonomous) cars move taking into account their distance from the car ahead, which is called in jargon its leader. For this reason it is more reasonable to consider an equa-tion of moequa-tion of the form

˙ x2(t) = v  1 x1(t) − x2(t)  , (5)

for a “well-chosen” function v. Equation (5) is a first order non-linear ODE.

The explicit expression for v can be obtained from experimental data. Bruce D. Greenshields was the first to

v

ρ vmax

ρmax

Fig. 3.: A linear velocity function.

carry out tests to measure traffic flow, traffic density and speed using photographic measurement methods in the 1930s. He postulated a linear relationship between speed and traffic density as in Fig. 3. Nowadays technology allows to have more advanced and detailed experimental data. As a result it is nowadays clear that the speed-density diagram is in general given by a cloud of points, it depends on the lane of the road and also on the time interval of the day under consideration.

Since we aim to develop a reasonably general model that can be applied to different cases, we rather have to understand which conditions must be imposed on v in order to obtain a realistic model. Let’s start from the following basic considerations:

(1) overtaking is not allowed and the minimal distance between two cars is `, hence the argument of v in (5) ranges in [0, ρmax], where ρmax= 1/`;

(2) speeds cannot be negative and are bounded from above by vmax, hence v takes values in [0, vmax];

(3) higher velocities correspond to higher distances from the car ahead, hence v is decreasing;

(4) a car is at rest if and only if it is “bumper-to-bumper” with the car ahead, see Fig. 4, namely x1(t) − x2(t) =

(7)

x

Fig. 4.: A “bumper-to-bumper” traffic condition. As a consequence, we are led to assume the following hypothesis

v : [0, ρmax] → [0, vmax] is a

de-creasing function such that v(0) = vmax and v(ρmax) = 0.

(V)

Clearly, a linearly decreasing velocity as in Fig. 3 satisfies (V).

In summary, the model for a traffic given by two cars is described by Equations (1), (2), (4) and (5), namely by the Cauchy problem

           ˙ x1(t) = vmax, ˙ x1(t) = v  1 x1(t) − x2(t)  , xi(0) = ¯xi, i ∈ {1, 2},

with v satisfying (V). By iterating (4) and (5) it is now clear that the traffic given by n > 2 cars is described by the Cauchy problem for a first order system of non-linear ODEs            ˙ x1(t) = vmax, ˙ xi(t) = v  1 xi−1(t) − xi(t)  , i ∈ {2, . . . , n}, xi(0) = ¯xi, i ∈ {1, . . . , n}. (6)

Model (6) is called first order follow-the-leader (FTL) model. Its name recalls that the speed of each car de-pends only on its distance from the car ahead (its leader) and because it is a system of first order ODEs.

3. A first order macroscopic model

The macroscopic variables are the density ρ, that is the number of vehicles per unit length, the velocity v, that is the space covered per unit time, and the flow f , that is the number of vehicles per unit time. We show now how to formally derive the relations between ρ, v and f . Example 1. Consider cars with the same length ` and moving equally spaced with the same velocity v, see Fig. 5. Both the distance between vehicles and the den-sity ρ do not change in time. Thus the number of vehicles passing an observer at x0 ∈ R in the time interval [0, τ ]

x x0

x0− τ v

Fig. 5.: The case considered in Example 1. is the number of vehicles initially in the space interval [x0− τ v, x0] and therefore

f = ρ x0− (x0− τ v) 

τ =⇒ f = ρ v. (7)

Example 2. Consider in the time interval [t1, t2] a

sec-tion of the road [a, b] with neither entries nor exits. Clearly, the number of cars in [a, b] at time t = t2 is

equal to the number of cars in [a, b] at time t = t1, plus

the number of cars that are entering in the time inter-val [t1, t2] through x = a, minus the number of cars that

are exiting in the time interval [t1, t2] through x = b, see

Fig. 6. In formulas this means that Z b a ρ(t2, y) dy = Z b a ρ(t1, y) dy + Z t2 t1 f (t, a) dt − Z t2 t1 f (t, b) dt. x t t2 t1 a b

Fig. 6.: The case considered in Example 2. Then, at least formally, we derive that

Z t2 t1 Z b a  ∂tρ + ∂xf (ρ)  dx dt = 0.

Since a, b, t1and t2are arbitrary, the integrand must be

null, namely

∂tρ + ∂xf = 0. (8)

We just obtained a scalar conservation law expressing the conservation of the number of cars. It is a partial differ-ential equation (PDE).

(8)

The simplest way to close the system is by assuming that the velocity of each vehicle depends on the density alone

v = v(ρ).

The considerations done in the previous section lead us to assume that the velocity function v : [0, ρmax] → [0, vmax]

satisfies (V). It is standard by now to quantify macro-scopic properties of traffic flow in terms of the fundamen-tal diagram, which relates the traffic flow measured in cars per unite time and the traffic density measured in cars per unit length. The reason is that a fundamental diagram shows beside the flow f (ρ), also the velocity as v(ρ) is by definition the slope of the line joining the ori-gin and the point ρ, f (ρ) on the fundamental diagram, see Fig. 7. Below we display such diagrams for different cases.

As a result, we obtain the scalar conservation law

∂tρ + ∂x ρ v(ρ) = 0. (9)

We just deduced the Lighthill-Whitham-Richards (LWR) model.12,13 v ρ vmax ρmax f ρ ρmax fmax ρm vmax v( ρ) ρ Fig. 7.: On the left a linear velocity function. On the right the corresponding strictly concave funda-mental diagram. Notice that the slope of the tan-gent to the fundamental diagram is vmax. Moreover

v(ρ) is the slope of the line crossing the fundamental diagram at (0, 0) and ρ, f (ρ).

In the remaining of this section we show how to ap-proximate solutions to Cauchy problems for LWR model

(9) by considering simple Riemann problems. This

method is called wave-front tracking method.32–34 We recall that a Riemann problem is a Cauchy problem with piecewise constant initial datum with (at most) only one jump.

For simplicity we assume below that v : [0, ρmax] →

[0, vmax] is C222 and such that f (ρ) = ρ v(ρ) is strictly

concave, namely

f00(ρ) = 2v0(ρ) + ρ v00(ρ) < 0 ∀ρ ∈ [0, ρmax].

3.1. Riemann problem

Consider the Riemann problem for LWR model        ∂tρ + ∂xf (ρ) = 0, ρ(0, x) = ( ρL if x < 0, ρR if x > 0, (10)

where f (ρ) = ρ v(ρ) and ρL, ρR∈ [0, ρmax] are constants.

The invariance of both PDE and initial datum under the change of coordinates

(t, x) 7→ (α t, α x), α > 0,

suggests to look for solutions invariant under such change of coordinates, namely that are constant along the rays x = m t, m ∈ R. Hence we restrict ourself to self-similar solutions, namely to solutions of the form

ρ(t, x) = φ(x/t).

This change of variables formally reduces the PDE in (10) to the following ODE

0 = −x t2φ +˙ 1 t f 0(φ) ˙φ =1 t  f0(φ) −x t ˙φ. By defying ξ = x/t, Riemann problem (10) becomes then the limiting problem

       f0(φ) − ξ˙ φ = 0, φ(−∞) = ρL, φ(+∞) = ρR. (11)

We need to distinguish the following cases to construct a solution to (11):

• If ρL = ρR, then the constant function φ ≡ ρL,R is

the solution.

• If ρL 6= ρR, then ˙φ 6≡ 0 and f0(φ(ξ)) = ξ. By this

last condition and the strict concavity of f we have f0φ f0(ρ)

(9)

con-cavity of f we have then

ρL= φ(−∞) = (f0)−1(−∞)>

> (f0)−1(+∞) = φ(+∞)= ρR.

In conclusion we get the continuous and piecwise C111

-solution φ(ξ) =        ρL if ξ < f0(ρL), (f0)−1(ξ) if f0 L) 6 ξ < f0(ρR), ρR if ξ > f0(ρR),

at least in the case ρL> ρR.

• If ρL < ρR, then we have to look for

discontinu-ous solutions and consider the weak formulation of f0(φ) − ξ˙ φ = 0, that is Z R  f (φ)ϕ0−φ d dξ ξ ϕ  dξ = 0 ∀ϕ ∈ C∞∞∞c (R2 ; R). (12) The expression of the initial datum suggests to look for a weak solution with a single jump. By introduc-ing φ(ξ) = ( ρL if ξ < σ ρR if ξ > σ (13) in (12) we obtain the constant σ ∈ R as follows:

+ Z σ −∞  f (ρL)ϕ0− ρL d dξ ξ ϕ  dξ + Z ∞ σ  f (ρR)ϕ0− ρR d dξ ξ ϕ  dξ = 0 ⇔ f (ρL) − f (ρR) − ρL− ρR σ  ϕ(σ) = 0 ⇔ σ =f (ρL) − f (ρR) ρL− ρR .

Therefore (13) is a piecewise constant weak solution with σ given above in the case with ρL < ρR.

Remark 3. The above computations for the speed of propagation of a discontinuity hold also in the case ρL>

ρR. For this reason, any discontinuity (ρL, ρR) with ρL6=

ρRhave speed of propagation

σ(ρL, ρR) =

f (ρL) − f (ρR)

ρL− ρR

. (14)

Such condition is the so called Rankine-Hugoniot condi-tion.35

We just obtained the Riemann solver RLWR, which is

an operator that associates to any ρL, ρR∈ [0, ρmax] the

weak solution φ(ξ) = RLWR[ρL, ρR](ξ) of the

correspond-ing Riemann problem (11) constructed above; we define it more rigorously in the following definition. Below we denote by BV the space of functions with bounded vari-ation. We can assume that any function in BV is right continuous by possibly changing its values at countably many points.

Definition 4. The Riemann solver RLWR: [0, ρmax]2 →

BV(R; [0, ρmax]) associated to LWR model (9) is defined

as follows:

(1) If ρL= ρR then RLWR[ρL, ρR] ≡ ρL,R.

(2) If ρL> ρRthen RLWR[ρL, ρR] is the rarefaction wave

RLWR[ρL, ρR](ξ) =        ρL if ξ < f0(ρL) t, (f0)−1(ξ) if f0(ρL) 6 ξ < f0(ρR), ρR if ξ > f0(ρR).

(3) If ρL< ρR then RLWR[ρL, ρR] is the shock wave

RLWR[ρL, ρR](ξ) =

(

ρL if ξ < σ(ρL, ρR),

ρR if ξ > σ(ρL, ρR),

where σ(ρL, ρR) is the speed of propagation of the

discontinuity and is given by (14).

f ρ ρL ρR x x t f0(ρL) ¯t ¯ t f0(ρR) ¯t ˙x= f0 (ρ L) ˙x= f 0(ρ R) x ρ(¯t, ·) ρL ρR

(10)

f ρ σ ρL ρR x x t σ ¯t ¯ t ˙x = σ x ρ(¯t, ·) ρL ρR

Fig. 9.: A shock wave. On the left we represent the fundamental diagram. On the right we give the solution in the (x, t)-plane, above, and its profile x 7→ ρ(¯t, x) computed at a time ¯t > 0, below.

Summing up, for any ρL, ρR∈ [0, ρmax] Riemann problem

(10) admits ρ(t, x) = RLWR[ρL, ρR](x/t) as self-similar

weak solution, see Fig. 8 and Fig. 9.

The mathematical construction given above for weak self-similar solutions to (10) may seem quite artificial. The following two examples show that such solutions have a clear physical counterpart: an acceleration corresponds to a rarefaction wave and braking produces a shock wave. Example 5 (Rarefaction wave). Consider cars that are bumper-to-bumper in the upstream of a traffic light placed at x = 0, which turns from red to green at time t = 0. The corresponding model is given by Riemann problem (10) with ρL = ρmax and ρR = 0. The solution

is then the rarefaction

ρ(t, x) =        ρmax if x < f0(ρmax) t, (f0)−1(x/t) if f0(ρmax) t 6 x 6 f0(0) t, 0 if x > f0(0) t.

According to this solution, see Fig. 10, the car initially at x0< 0 is at rest until time t1= x0/f0(ρmax) > 0,

x(t) = x0, t ∈ [0, t1],

and then starts to move according to the law ˙ x(t) = v ρ(t, x) = v (f0)−1 x(t)/t, t > t 1. f ρ ρL= ρmax ρR= 0 x t ρmax 0 x0 t1

Fig. 10.: The rarefaction considered in Example 5. The dashed curve corresponds to the trajectory of a car.

Notice that for any t > t1 its trajectory belongs to the

cone

f0(ρmax) t 6 x 6 f0(0) t

simply because f0(0) = vmax.

This corresponds to what we actually observe in real life: the cars closer to the traffic light start to move earlier and then accelerate.

Example 6 (Shock wave). Consider cars that are

equally spaced and moving with the same velocity in the upstream of a traffic light placed at x = 0, which is red and stays red. This is equivalent to place cars bumper-to-bumper in the downstream of the traffic light. The corresponding model is given by Riemann problem (10) with ρL∈ (0, ρmax) and ρR= ρmax. The solution is then

the shock wave ρ(t, x) = ( ρL if x < σ t ρmax if x > σ t with σ =f (ρL) − f (ρmax) ρL− ρmax . According to this solution, see Fig. 11, the car initially

f ρ ρR= ρmax ρL σ x t ρL ρmax x0 t1

Fig. 11.: The shock considered in Example 6. The dashed curve corresponds to the trajectory of a car. at x0 < 0 moves with velocity v = v(ρL) > 0 until time

t1= −x0/ v(ρL) − σ > 0,

(11)

and then stops to move ˙

x(t) = 0, t > t1.

This corresponds to what we actually observe in real life: once a car reaches the tail of a queue it suddenly stops.

3.2. Wave-front tracking

A solution of the Cauchy problem for LWR model (9) (

∂tρ + ∂xf (ρ) = 0

ρ(0, x) = ¯ρ(x) (15)

can be obtained by applying the wave-front tracking al-gorithm. Roughly speaking, approximate solutions to (15) are obtained by gluing together approximate solu-tions to Riemann problems corresponding to the jumps of an approximate initial datum and to the interac-tions between waves. We give below more details about this construction. For simplicity we assume that ¯ρ ∈ BV ∩ L1

(R; [0, ρmax]).

As a first step we construct an approximate initial datum ¯ρn as follows:

• Fix n ∈ N sufficiently large and define in [0, ρmax] the

mesh

Mn=ρ0, ρ1, . . . , ρn

where ρi= εni with εn = ρmax/n. Notice that ρ0= 0

and ρn = ρmax.

• Approximate the initial datum ¯ρ by a piecewise con-stant function ¯ρn: R → [0, ρmax], see Fig. 12, with a

finite number of jumps and such that ¯

ρn(R) ⊆ Mn, TV( ¯ρn) 6 TV(¯ρ),

lim

n→+∞k¯ρn− ¯ρkL1(R)= 0.

(16)

The second step is to construct a piecewise constant approximate solution ρn: R → Mn to the approximate

Cauchy problem corresponding to the approximate initial datum ¯ρn: R → Mn

(

∂tρ + ∂xf (ρ) = 0,

ρ(0, x) = ¯ρn(x).

(17)

Remark 7. For t > 0 sufficiently small, the exact so-lution ρex to (17) can be obtained by gluing the exact

solutions of the Riemann problems corresponding to the

x ρ ¯ ρn ¯ ρ

Fig. 12.: An approximate initial datum ¯ρn of the

initial datum ¯ρ.

discontinuity points of ¯ρn. This means that if ¯ρn has

a discontinuity at x0 ∈ R and lim x→x±0 ¯ ρn = ρ±0 ∈ Mn, then ρex(t, x) = RLWR[ρ−0, ρ + 0] (x − x0)/t  at least for t > 0 and |x − x0| sufficiently small. Such

construc-tion for ρexholds until an interaction occurs between two

waves started from two adjacent discontinuity points of ¯

ρn. The exact solution ρex can be then prolonged by

applying again the Riemann solver RLWR at the

inter-action if and only if ρex is piecewise constant. However

this is not in general the case. Indeed RLWR[ρL, ρR] is

piecewise constant if and only if ρL6 ρR. On the other

hand, if ρL> ρR then RLWR[ρL, ρR] is a rarefaction and

RLWR[ρL, ρR](R) = [ρR, ρL].

The considerations in Remark 7 motivate the dis-cretization of the rarefactions and the introduction of the Riemann solver RnLWR: M2n → BV(R; Mn) defined as

follows:

(1) If ρL, ρR∈ Mnare such that ρL6 ρR, then we define

Rn

LWR[ρL, ρR] ≡ RLWR[ρL, ρR].

(2) If ρL, ρR∈ Mnare such that ρL> ρR, then we define

Rn LWR[ρL, ρR](ξ) =        ρL if ξ 6 σ(ρL, ρh+1), ρj if σ(ρj−1, ρj) < ξ 6 σ(ρj, ρj+1), h < j < k, ρR if ξ > σ(ρk−1, ρR),

where {ρh, . . . , ρk} = [ρR, ρL] ∩ Mn are such that

ρj > ρj+1.

Notice that the discontinuities of the discretized rar-efaction satisfy the Rankine-Hugoniot condition (14) and therefore (t, x) 7→ Rn

LWR[ρL, ρR](x/t) is a weak solution

(12)

A piecewise constant approximate solution ρn: R →

Mnto (17) is then obtained by juxtaposing the solutions

of the Riemann problems corresponding to the disconti-nuities of ¯ρn and to the interactions between waves, and

solved by applying Rn LWR, see Fig. 13. f ρ ρ0 ρ1 ρ2 ρ3 ρ4 ρ5 σ1 σ2 σ3 σ 4 σ 5 x t ˙x= σ1 ˙x= σ 5 ˙ x = 0 ˙x = σ 5 ˙x = σ 4 ˙ x = σ3 ˙x= σ2 ˙x= σ1

Fig. 13.: Partial construction of a piecewise con-stant approximate solution ρn to (17) with initial

datum ¯ρn represented in Fig. 12.

Such procedure can be iterated as long as both the number of waves and the number of interactions are bounded on any bounded time interval; see Fig. 14 for an hypothetical example for which this is not the case.

x t

T

Fig. 14.: An example of infinitely many interac-tions occurring in a bonded time interval [0, T ], see Ref. [36] for its construction.

The global existence of the approximate solution ρn

is ensured by the following proposition.

Proposition 8. After any interaction the number of waves decreases by at least one and the total variation does not increase.

Proof. Consider an interaction between m > 2 waves. Let ρi and ρi+1 be the left and right states of the i-th

wave, i ∈ {1, . . . , m}. We distinguish the following cases: f ρ σ1 σ2 σ 3 σ4 ρ1 ρ2 ρ3 ρ4 t x σ1 σ2 σ 3 σ 4 ρ

Fig. 15.: An interaction involving three shock waves separating ρ1and ρ2, ρ2and ρ3, and ρ3and ρ4. On

the left we represent in the fundamental diagram both the states, ρi, and the speeds of propagation,

σi. Indeed, for instance, σ1= σ(ρ1, ρ2) is the slope

of the segment joining (ρ1, f (ρ1)) to (ρ2, f (ρ2)), as

well as the speed of propagation of the shock wave separating the states ρ1 and ρ2. On the right we

represent the shock waves in the (x, t)-plane.

f ρ σ1 σ2 σ3 σ 4 σ5 ρ1 ρ3ρ2 ρ5ρ4 t x σ1 σ2 σ 3 σ 4 σ5 ρ

Fig. 16.: An interaction involving four waves sepa-rating ρ1and ρ2, ρ2and ρ3, ρ3 and ρ4, and ρ4 and

ρ5. On the left we represent in the fundamental

di-agram both the states, ρi, and the speeds of

prop-agation, σi. Indeed, for instance, σ1 = σ(ρ1, ρ2)

is the slope of the segment joining (ρ1, f (ρ1)) to

(ρ2, f (ρ2)), as well as the speed of propagation of

the shock wave separating the states ρ1and ρ2. On

the right we represent the shock waves in the (x, t)-plane.

(1) Assume that all the involved waves are shock waves, see Fig. 15. In this case ρi < ρi+1 and the result of

the interaction is a single shock wave between ρ1and

ρm+1. Therefore the number of waves decreases by

(13)

the interaction, see Fig. 16. If the i-th wave is a rarefaction wave, then ρi− ρi+1 = εn; moreover if

i > 1 then the (i − 1)-th wave is a shock wave with ρi − ρi−1 > 2εn and if i < m then the (i + 1)-th

wave is a shock wave with ρi+2 − ρi+1 > 2εn. As

in the previous case the result of the interaction is a single shock wave between ρ1and ρm+1, the number

of waves decreases by at least one and, differently from the previous case, the total variation strictly decreases.

Since the above cases encompasses all the possible inter-actions, the proof is complete.

By Proposition 8 the approximate solution ρncan be

con-structed globally in time. Moreover if th< th+1 are two

consecutive interaction times, then for any t ∈ (th, th+1)

the approximate solution ρn(t, ·) takes the form

ρn(t, x) =+ρh1 2 ·1(−∞,x h 1(t))(t) + Nh−1 X i=1 ρhi+1 2 ·1[x h i(t),xhi+1(t))(t) +ρhNh+1 2 ·1[xh N(t),+∞)(t), (18)

where1I is the indicator function of the interval I ⊂ R,

ρh 1 2 , . . . , ρh Nh+1 2

∈ Mn are constant with ρhi−1 2 6= ρh i+1 2 , xh

i(t) ∈ R is a discontinuity point of ρn(t, ·) with ˙xhi(t) =

σ ρh i−1 2 , ρh i+1 2 , and Nh = Nh (t) ∈ N is the number of discontinuity points of ρn(t, ·). We stress that t 7→ Nh(t)

is piecewise constant and non-increasing by Proposition 8.

The next step is to prove that (up to a subsequence) {ρn}n converges in L1loc([0, +∞) × R). To do so we need

the estimates provided by the following proposition. Proposition 9. There exists a constant C > 0 such that for any t, s > 0 and n ∈ N we have

TV ρn(t, ·) 6 TV(¯ρ), kρn(t, ·)kL∞(R)6 ρmax,

kρn(t, ·) − ρn(s, ·)kL1(R)6 C |t − s|.

(19) Proof. The first estimate follows from Proposition 8 and (16)2. By construction ρntakes values in Mn⊂ [0, ρmax],

hence the second estimate holds true. The last es-timate follows from the fact that f is Lipschitz con-tinuous and its Lipschitz constant L = sup[0,ρmax]|f0|

bounds the speed of propagation of any wave. Let’s

s t x1(s) σ1 x1(t) x2(s) σ2 x2(t) x3(s) σ3 x3(t) ρi−3 2 ρi− 1 2 ρi+ 1 2 ρi+ 3 2 ρi−3 2 ρi−1 2 ρi+1 2 ρi+3 2

Fig. 17.: The profiles of ρn(t, ·) and ρn(s, ·) in the

case no interaction occurs in the time interval [s, t]. prove it. Let th < th+1 be two consecutive

interac-tion times. Then ρn takes the form (18) in the time

interval (th, th+1). Fix s, t ∈ (th, th+1). Notice that

xh i(t) = xhi(s) + σ ρhi−1 2 (t), ρh i+1 2 (t) · (t − s) for any i ∈ [1, Nh

] ∩ N. Hence, see Fig. 17, we have

kρn(t, ·) − ρn(s, ·)kL1(R)= Nh X i=1 (t − s) · σ ρhi−1 2 − ρh i+1 2 · ρ h i−1 2 − ρh i+1 2  6 L · |t − s| · Nh X i=1 ρhi−1 2 − ρh i+1 2 = L · |t − s| · TV ρn(t, ·) 6 L · |t − s| · TV(¯ρ).

The case when one or more interactions take place for times between t and s is similar, because by the finite speed of propagation of the waves the map t 7→ ρn(t, ·)

is L1-continuous across interaction times. Therefore it is

sufficient to take C = TV( ¯ρ) · L.

By (19) we can apply Helly’s theorem in the form [33, Theorem 2.4], and obtain that (up to a subsequence) {ρn}nconverges in L1loc([0, +∞) × R) to a function ρ and

that

TV ρ(t, ·) 6 TV(¯ρ), kρ(t, ·)kL∞(R)6 ρmax,

kρ(t, ·) − ρ(s, ·)kL1(R)6 C |t − s|.

At last we have to prove that ρ is actually a weak solution of the original Cauchy problem (15).

Proposition 10. ρ is a weak solution of Cauchy problem (15), namely ρ(0+, x) = ¯

ρ(x) for almost every x ∈ R and

(14)

Proof. The initial condition is satisfied almost every-where because, by (16)3, (19)3and the L1

loc-convergence

of ρn to ρ, for any a < b we have that

Z b a ρ(t, x) − ¯ρ(x) dx 6 Z b a ρ(t, x) − ρn(t, x) dx + Z b a ρn(t, x) − ¯ρn(x) dx + Z b a ¯ρn(x) − ¯ρ(x) dx t→0+ −−−−−→ n→+∞ 0.

Take ϕ in C∞∞∞c (0, +∞) × R; R. Choose T > 0 such that ϕ(t, ·) ≡ 0 for any t > T . Since ρn is uniformly

bounded and f is uniformly continuous on bounded sets, it is sufficient to prove that

Z T 0 Z R  ρn∂tϕ + f (ρn) ∂xϕ  dx dt −−−−−→ n→+∞ 0. (20)

By (18) the approximate solution ρn(t, ·) has a jump at

x1(t), . . . , xN (t)(t) with xi(t) < xi+1(t). The polygonal

lines x = xi(t) subdivide the strip [0, T ] × R into finitely

many regions Γi where ρn is constant. Introducing the

vector

Φ ≡ ϕ ρn, ϕ f (ρn),

by the Gauss-Green formula the double integral in (20) can be written as X i Z Z Γi div Φ(t, x) dx dt =X i Z ∂Γi Φ · nnn ds. (21) Above ∂Γi is the oriented boundary of Γi, while nnn is

the outer unit normal vector. Observe that nnn ds = ±( ˙xi, −1) dt along each polygonal line x = xi(t), while

ϕ(t, x) = 0 along the lines t = 0 and T = 0. By (21) the double integral in (20) is equal to

Z T 0 X i  ˙ xi  ρi+1 2 − ρi− 1 2  −f ρi+1 2 − f ρi− 1 2   ϕ t, xi(t) dt. By construction ˙xi= σ ρi+1 2, ρi+ 1 2, hence ˙ xi  ρi+1 2 − ρi−12  −f ρi+1 2 − f ρi−12  = 0 and therefore (20) is trivial.

It can be proved that ρ is not the only weak solution of Cauchy problem (15), but it is the unique weak solution satisfying the Kruzhkov37,38 condition

Z R Z +∞ 0  |ρ(t, x) − k| ∂tϕ(t, x) + F ρ(t, x), k∂xϕ(t, x)  dt dx > 0 ∀ϕ ∈ C∞∞∞c (0, +∞) × R; [0, +∞), k ∈ [0, ρmax], (22) where F (ρ, k) = sign ρ − kf (ρ) − f (k). (23) In poor words, Kruzhkov condition (22) selects the only weak solution with only physically reasonable shocks, that are the discontinuities across which the density in-creases. Indeed only braking produces a shock wave and only an increase of the density produces braking, there-fore the density has to increase across a shock wave. We recall that Kruzhkov condition (22) can be deduced by applying the vanishing viscosity method, namely by con-sidering the solutions ρεof the viscous approximations

∂tρ + ∂xf (ρ) = ε∂x2ρ, ε > 0,

which converge to ρ as ε goes to zero. Furthermore in Ref. [37] the uniqueness is proved by applying the dou-bling of variables method.

4. Links between microscopic and macro-scopic models

In this section we give some basic ideas on how it is pos-sible to derive the first order microscopic FTL model (6) from the scalar macroscopic LWR model (9) and vicev-ersa.

We start by showing how it is possible to heuristically derive a first order FTL model from the LWR model:

• Let ρ be the unique weak solution of LWR model (9) satisfying Kruzhkov condition (22).

• Let F be the cumulative distribution of ρ F (t, x) =

Z x

−∞

ρ(t, y) dy. • Let X be the pseudo inverse of F

X(t, z) = inf{x ∈ R : F (x) > z}, z ∈ [0, L]. • Formally, X(t, z) satisfies F (t, X(t, z)) = z and

there-fore ∂xF = ρ ∂tF = −ρ v(ρ) ⇒ 1 = ∂zF t, X(t, z) = (∂xF )(∂zX) = ρ (∂zX) 0 = ∂tF t, X(t, z) = ∂tF + (∂xF )(∂tX) = ρ ∂tX − v(ρ)  ⇒ Xt= v(1/Xz).

• Approximating the obtained equation via a forward z-finite difference with step ` gives

(15)

The obtained system corresponds to the first order FTL model (6), which is therefore a discrete Lagrangian ver-sion of (9).

Conversely, below we show the main steps to approxi-mate the unique weak solution of the Cauchy problem for LWR model (15) satisfying Kruzhkov condition (22) via a particle system governed by the first order FTL model (6):

(1) Fix an initial datum ρ0 and let L = kρ0kL1(R)be the

total length of the traffic.

(2) Split ¯ρ in n ∈ N platoons of (possible fractional) cars with length ` = L/n and with (ordered) end points

¯

xn < ¯xn−1< . . . < ¯x1< ¯x0.

(3) “Particles” with mass ` are initially placed in ¯

x0, . . . , ¯xn and evolve according to the first order

mi-croscopic FTL model            ˙ x0(t) = vmax, ˙ xi(t) = v  ` xi−1(t) − xi(t)  , i ∈ {1, . . . , n}, xi(0) = ¯xi, i ∈ {0, . . . , n}.

(4) Introduce the discrete density ρn(t, x) = n X i=1 ` xi(t) − xi−1(t) 1[xi−1(t),xi(t))(x). (24) We have then the following result.

Theorem 11. Let ρ be the unique weak solution of the Cauchy problem for LWR model (15)

(

∂tρ + ∂xf (ρ) = 0,

ρ(0, x) = ¯ρ(x),

satisfying Kruzhkov condition (22). Assume that ¯ρ is in L1∩ L∞

(R), and v is in C1([0, +∞); R) and is strictly decreasing. If one of the following conditions is satisfied

¯

ρ ∈ BV(R),

ρ 7→ ρ v0(ρ) is non-increasing,

then (up to a subsequence) {ρn}n given by (24) converges

to ρ almost everywhere and in L1loc on R+× R.

Proof. Below we give only the main ideas of the proof. The rigorous proof is quite technical and can be found in Ref. [14,39,40].

(1) Prove by contradiction a maximum principle which ensures that the particles preserve their order and the global existence for the solution of FTL model (6).

(2) Prove by induction a uniform BV estimate for ρn(t, ·).

(a) If ¯ρ is in BV, then prove a uniform BV contrac-tion property for ρn(t, ·).

(b) If ρ 7→ ρ v0(ρ) is non-increasing, then prove a uni-form discrete Oleinik condition for ρn(t, ·).

(3) Prove that {ρn}n converges to a measure ρ in

L1

loc( ¯R+; dL,1), where dL,1 is the rescaled

Wasser-stein distance defined by

dL,1(ρ1, ρ2) = kFρ1− Fρ2kL1(R;R)

for any non-negative Radon measures ρ1, ρ2in R with

compact support and such that ρ1(R) = L = ρ2(R),

where Fρi(x) = ρi (−∞, x], i ∈ {1, 2}.

(a) Prove a continuity

esti-mate for ρn in L1loc(R+; dL,1) uniform in t and

n, namely for Xn in L1loc(R+× [0, L]).

(b) Apply a generalised Aubin-Lions lemma. (4) Prove that ρn is a weak solution and then use the

strong L1-compactness to pass to the limit and get

that ρ is a weak solution.

(5) Prove that ρ satisfy Kruzhkov condition (22) by care-fully estimating several terms appearing after intro-ducing ρn in (22).

For completeness we recall the generalised Aubin-Lions lemma.

Theorem 12 (Generalized Aubin-Lions lemma). Let T, L > 0 be fixed constants and I ⊂ R be a bounded open interval. Assume w : R → R is a Lipschitz contin-uous and strictly monotone function. Let {ρn}n∈N be a

sequence in L∞ (0, T ) × R ∩ ML such that:

(1) ρn: (0, T ) → L1(R) is measurable for all n ∈ N;

(2) spt ρn(t) ⊆ I for all t ∈ (0, T ) and n ∈ N;

(3) sup n∈N Z T 0  w ρn(t)  L1(I)+ TV  w ρn(t)  dt < +∞; (4) there exists a constant C depending only on T such

that

(16)

Then, {ρn}n∈N is strongly relatively compact in

L1 (0, T ) × R.

The importance of the micro-macro limit described above is twofold: first, it justifies at the microscopic level the use of macroscopic models; second, it allows to de-velop macroscopic models on the basis of microscopic in-teraction rules. In particular, this procedure leads to se-lect a unique Riemann solver for a macroscopic model, which is not necessarily the classical Kruzhkov Riemann solver; see for instance Ref. [14,17,18,39–59] for macro-scopic models justified or developed on the basis of this procedure.

5. Heterogeneous traffic

We motivate the need to introduce new models by listing the main drawbacks of both FTL model (6) and LWR model (9):

• Traffic is typically heterogeneous: assuming that all cars have the same length and maximal velocity is sometimes a severe restriction.

• Unrealistic behavior of the drivers, that adjust in-stantaneously their velocities according to the densi-ties they are experiencing.

• Experimental data show that the fundamental dia-gram (ρ, ρ v) is given by a cloud of points rather than the support of a map ρ 7→ ρ v(ρ).

• They do not reproduce the amplification of small perturbations (phantom traffic jams), due to the so called “over reaction”.

• They do not reproduce hysteresis processes (acceler-ation and deceler(acceler-ation follow two different branches in the density-flow diagram).

In this section we generalize both models (6) and (9) to the case of an heterogeneous traffic, mainly given by trucks and cars, see Fig. 18. We characterize each

ve-xn(t) xi(t) x2(t) x1(t) x . . . .

Fig. 18.: An heterogeneous vehicle traffic. hicle by its length and maximal speed. Since trucks are

longer and (typically) slower than cars, we can assume that longer vehicles are slower. Moreover, it is reasonable to assume that length and maximal speed of a vehicle do not change in time.

We encode this information (maximal speed and a length) by expressing the velocity v in terms of La-grangian marker w and density ρ as follows

v(ρ, w) = w − p(ρ), (25)

with p(ρ) = ργ, where γ > 0 is a parameter of the model.

Here p plays the role of an anticipation factor, taking into account divers’ reactions to the state of traffic in front of them.

Remark 13. In general it is sufficient to assume for p : [0, ρmax] → R that

• p(0) = 0,

• p0(ρ) > 0 and 2p0(ρ) + p00(ρ)ρ > 0 for every ρ > 0.

The choice p(ρ) = ργ, γ > 0, is made here only to avoid

technicalities linked to working in a general framework. At the same time, we avoid to use the explicit expression of p whenever it is not necessary.

Equation (25) means that a vehicle characterized by La-grangian marker w moves with velocity v = w − p(ρ) if it experiences the density ρ. As a consequence, a vehicle with Lagrangian marker w has maximal speed v(0, w) = w. Moreover, since a vehicle is at rest if and only if it is “bumper-to-bumper”, the length ` of a vehi-cle with Lagrangian marker w satisfies v(1/`, w) = 0 and therefore

` = 1/p−1(w).

Notice that the length ` is lower for higher values of the maximal speed w.

For simplicity in the exposition we consider separately the microscopic and the macroscopic models.

5.1. Microscopic model

(17)

where wiis the maximal speed of the i-th vehicle. System

(26) can be written as the Cauchy problem for a system of second order non-linear ODEs

                                   ˙ xi(t) = vi(t), i ∈ {1, . . . , n}, ˙vi(t) = p0  1 xi−1(t) − xi(t)  v i−1(t) − vi(t) xi−1(t) − xi(t) 2, i ∈ {2, . . . , n}, xi(0) = ¯xi, i ∈ {1, . . . , n}, v1(0) = w1, vi(0) = v  1 ¯ xi−1− ¯xi , wi  , i ∈ {2, . . . , n}.

The above model is called second order FTL model. 5.2. Macroscopic model

By introducing the velocity function (25) in the conser-vation law (8) we obtain

∂tρ + ∂x ρ v(ρ, w) = 0. (27)

Since we have two unknown variables, ρ and w, we need a second PDE. By assumption the maximal speed and length of each vehicle do not change in time. This implies that w is constant along any trajectory t 7→ x(t) and therefore

d

dtw t, x(t) = ∂tw + ˙x ∂xw = ∂tw + v(ρ, w) ∂xw = 0. We just obtained the transport equation

∂tw + v(ρ, w) ∂xw = 0. (28)

By coupling the conservation law (27) together with the transport equation (28) we obtain the 2 × 2 system of PDEs

(

∂tρ + ∂x(ρ v) = 0,

∂tw + v ∂xw = 0.

(29) This is a second order macroscopic model and is the Aw-Rascle-Zhang (ARZ) model.15,16

ARZ model can be interpreted as a generalization of the LWR model to a heterogeneous traffic. In poor words, according to LWR model (ρ, f ) belongs to a unique funda-mental diagram ρ 7→ f (ρ) = ρ v(ρ), whereas ARZ model allows to consider a one parameter family of fundamen-tal diagrams ρ 7→ fw(ρ) = fw(ρ) = ρ v(ρ, w), where

v(ρ, w) = w − p(ρ). Notice that if we consider only one

type of vechicle, then w is constant and ARZ model (29) reduces to LWR model (9) with velocity function given by v(ρ) = w − p(ρ), which satisfies (V).

As for LWR model, we construct a Riemann solver for ARZ model. Consider the Riemann problem for ARZ model (29)                      ∂tρ + ∂x ρ v(ρ, w) = 0, ∂tw + v(ρ, w) ∂xw = 0, ρ(0, x) = ( ρL if x < 0, ρR if x > 0, w(0, x) = ( wL if x < 0, wR if x > 0, (30)

where WL= (ρL, wL), WR= (ρR, wR) ∈ Ω are constant,

and Ω is the physical domain defined by

Ω = {(ρ, w) ∈ [0, +∞) × (0, +∞) : w > p(ρ)} . Notice that v(W ) > 0 for any W = (ρ, w) belonging to Ω.

We need to distinguish the following cases to construct a solution to (30).

• If WL = WR, then the constant function (ρ, w) ≡

WL,Ris the solution.

• If wL = wR, then all the vehicles have the same

length and maximal speed, hence ARZ model re-duces to LWR model. Indeed, introducing (ρ, wL,R)

in (30) we obtain (10) with f (ρ) = ρ wL,R −

p(ρ). For what we have already seen in Section 3.1, the solution to (30) is then given by W (t, x) = RLWR[ρL, ρR](x/t), wL,R and it can be either a

rar-efaction (if ρL> ρR) or a shock (if ρL< ρR).

• If wL 6= wR, then we have two families of vehicles,

one in (−∞, 0) and one in [0, +∞). By assumption overtaking is not allowed, hence the two families do not mix each other. The vehicles are anisotropic (respond only to frontal stimuli), hence the vehicles with Lagrangian marker wRare not influenced by the

other vehicles; for this reason they move with speed vR = v(WR) = wR− p(ρR). As a consequence the

w-component of the solution to (30) is w(t, x) =

(

wL if x < vRt,

wR if x > vRt.

In x < vRt we have only vehicles with Lagrangian

marker wL and therefore their dynamics can be

(18)

them moves with speed vL= wL− p(ρL) may lead to

a contradiction because their velocity cannot exceed vR (overtaking is not allowed). Also assuming that

each of them moves with speed vRmay lead to a

con-tradiction because their velocity cannot exceed their maximal speed wL. For this reason we introduce the

intermediate density ρ ∈ [0, pe −1(wL)] implicitly

de-fined by v(ρ, we L) = min{wL, vR} and take ρ(t, x) = ( RLWR[ρL,ρ](x/t)e if x < vRt ρR if x > vRt

as ρ-component of the solution to (30).

The above construction leads to the following defini-tion.

Definition 14. The Riemann solver RARZ: Ω2 →

BV(R; Ω) associated to ARZ model (29) is defined by RARZ[WL, WR] ≡ (φ1, φ2), where φ1(ξ) = ( RLWR[ρL,ρ](ξ)e if ξ < vR, ρR if ξ > vR, φ2(ξ) = ( wL if ξ < vR, wR if ξ > vR, e

ρ = p−1(wL− min{wL, vR}) and RLWR is the Riemann

solver associated to LWR model (9) with v(ρ) = wL−

p(ρ), see Definition 4.

Proposition 15. For any WL = (ρL, wL), WR =

(ρR, wR) ∈ Ω, we have that

W (t, x) = RARZ[WL, WR](x/t)

is a self-similar weak solution of Riemann problem (30). Proof. The initial condition in (30) is clearly satisfied. Hence we are left to show that W satisfies ARZ model (29), at least in a weak sense.

The case wL = wR is straightforward because then

the w-component of W is constant, w ≡ wL,R, and W

satisfies (29) because its ρ-component satisfies (9) with v(ρ) = wL,R− p(ρ).

Assume wL 6= wR. The simplest way to obtain a

weak formulation of (29) is to rewrite it as a 2 × 2 system

of conservation laws. Let y = ρ w be the generalized momentum of the vehicles. By introducing in ARZ model (29) the conservative variables U = (ρ, y), we obtain the 2 × 2 system of conservation laws

∂tU + ∂xF (U ) =

0 0 

, (31)

where, with a slight abuse of notations, F (U ) = U v(U ) with

v(U ) = y

ρ− p(ρ).

We stress that the above expression of the velocity in terms of the conservative variables U is not well defined at the vacuum ρ = 0; for this reason (29) and (31) are equivalent only away from the vacuum ρ = 0. The weak formulation of (31) is        Z R Z +∞ 0 (ρ ∂tϕ + v ρ ∂xϕ) dt dx = 0 Z R Z +∞ 0 w (ρ ∂tϕ + v ρ ∂xϕ) dt dx = 0 ∀ϕ ∈ C∞∞∞c (0, +∞) × R; R. (32)

Consider a test function ϕ and let D ⊂ R2be its compact

support. We distinguish the following cases: • If D ⊂ {(t, x) ∈ (0, +∞) × R : x > vRt}, then

W ≡ ρR wR



in D.

Hence W is constant in D and therefore satisfies (32). • If D ⊂ {(t, x) ∈ (0, +∞) × R : x < vRt}, then

W ≡ ρ

wL



in D,

where ρ(t, x) = RLWR[ρL, ρR](x/t) is a weak solution

of the Riemann problem for LWR model (10) with f (ρ) = ρ wL− p(ρ). Therefore also in this case (32)

easily follows. • Assume that

D−= D ∩ {(x, t) ∈ R × (0, +∞) : x < vRt} 6= ∅,

D+= D ∩ {(x, t) ∈ R × (0, +∞) : x > vRt} 6= ∅.

By construction W is piecewise C111 and, beside x =

vRt, there exist at most other two straight lines

across which W is not C111. For this reason it is not

(19)

x t x = v Rt D− D+ t2 t1 x1 x2

Fig. 19.: Notations used in the proof of Proposi-tion 15.

that W is C111 both in D

− and D+. Fix x1, x2 ∈ R

and t1, t2> 0, see Fig. 19, such that

D ⊆ [x1, x2] × [t1, t2]. We have then Z Z D− ρ ∂tϕ + v(ρ, w) ρ ∂xϕ dx dt = Zt2 t1 ZvRt x1 ρ ∂tϕ + v(ρ, wL) ρ ∂xϕ dx dt = Zt2 t1  d dt ZvRt x1 ρ ϕ dx  − vRρ ϕx=vRt− Z vRt x1 ∂tρ ϕ dx  dt + Zt2 t1 ZvRt x1 ∂x v(ρ, wL) ρ ϕ − ∂x v(ρ, wL) ρϕ dx dt = − Zt2 t1 vReρ ϕ(t, vRt) dt − Zt2 t1 ZvRt x1 ∂tρ ϕ dx dt + Zt2 t1 v(ρ, we L)eρ ϕ(t, vRt) dt − Zt2 t1 ZvRt x1 ∂x v(ρ, wL) ρϕ dx dt = Zt2 t1 v(ρ, we L) − vR  e ρ ϕ(t, vRt) dt − Zt2 t1 ZvRt x1  ∂tρ + ∂x v(ρ, wL) ρ  ϕ dx dt = 0

because ρ is a classical C111-solution of (9) with v(ρ) =

v(ρ, wL) in D−, and because either v(ρ, we L) = vR, or v(ρ, we L) = wL andρ = 0. Furthermore we have thate

Z Z D+ (ρ ∂tϕ + v(ρ, w) ρ ∂xϕ) dx dt = Zt2 t1 Zx2 vRt ρR∂tϕ + v(WR) ρR∂xϕ dx dt = Zt2 t1  d dt Z x2 vRt ρRϕ dx  + vRρRϕ(t, vRt) − v(WR) ρRϕ(t, vRt)  dt = vR− v(WR)ρR Z t2 t1 ϕ(t, vRt) dt = 0

because by definition v(WR) = vR. At last, from the

above computations easily follow that Z Z D− w(ρ ∂tϕ + v(ρ, w) ρ ∂xϕ) dx dt = wL Z Z D− (ρ ∂tϕ + v(ρ, wL) ρ ∂xϕ) dx dt = 0, Z Z D+ w(ρ ∂tϕ + v(ρ, w) ρ ∂xϕ) dx dt = wR Z Z D+ (ρ ∂tϕ + v(ρ, wR) ρ ∂xϕ) dx dt = 0.

This concludes the proof.

Notice that if wL = wR then ρ = ρe R and the weak solution W has a rarefaction if ρL > ρR, or a shock if

ρL< ρR. On the other hand, if wL6= wR, then the weak

solution W has a discontinuity across x = vRt, which is

called contact discontinuity. We recall that an accelera-tion corresponds to a rarefacaccelera-tion wave and braking pro-duces a shock wave. About the contact discontinuities, they simply separate vehicles with different Lagrangian markers. We clarify this point with the following exam-ple. f ρ σ ρL e ρ ρR x ˙x= vR x t fwL0 (ρL) ¯t ¯ t fwL0 (0) ¯t vR¯t ˙x = f0 w L L ) ˙x= f 0 wL (0) x ρ(¯t, ·) ρL e ρ ρR x w(¯t, ·) wL wR

Fig. 20.: A rarefaction wave followed by a contact discontinuity. On the left we represent the funda-mental diagram. On the right we give the solution in the (x, t)-plane, above, and the two components of its profiles x 7→ ρ(¯t, x) and x 7→ w(¯t, x) computed at a time ¯t > 0, below.

(20)

time t = 0 are respectively in (−∞, 0) and [0, +∞). More precisely, assume that the vehicles in x < 0 have Lagrangian marker wL and are “bumper-to-bumper”,

whereas those in x > 0 have Lagrangian marker wR and

velocity vR, with 0 < wL< vR< wR. The corresponding

model is given by Riemann problem (30) with ρL= p−1(wL), ρR= p−1(wR− vR).

In this case ρ = pe −1(w

L− min{wL, vR}) = p−1(0) = 0.

The solution, see Fig. 20, has a rarefaction followed by contact discontinuity, has ρ-component

ρ(t, x) =            ρL if x < a(ρL) t, (fw0 L) −1(x/t) if f0 wL(ρL)t 6 x < f 0 wL(0) t, 0 if fw0L(0) t 6 x < vRt, ρR if x > vRt, where fwL(ρ) = ρ v(ρ, wL) = ρ wL − p(ρ), and w-component w(t, x) = ( wL if x < vRt, wR if x > vRt.

Remark 17. Some comments on the above example are in order. Even if the vacuum is not initially present, it arises immediately after t = 0. This has a physical mo-tivation: the slow vehicle cannot reach the speed vR of

the fast vehicle because wL < vR, for this reason a gap

is created between the two families of vehicles. As a con-sequence, the total variation of the ρ-component of the solution increases instantaneously at time t = 0+. On the

contrary, it is easy to see that the total variation of the solution does not increase in the Riemann invariant coor-dinates (v, w). For this reason existence results for ARZ model are based on estimates obtained in the Riemann invariant coordinates. At last, it is worth to mention that, at least away from the vacuum, 2 × 2 system of conser-vation laws (31) is a Temple60system: the Lax curves in the Riemann invariant coordinates are straight lines. Remark 18. In Ref. [61] it is proved that no Kruzhkov type condition can select a unique solution for ARZ model. Roughly speaking, the Kruzhkov type condi-tions can prohibit “wrong shocks” but it cannot pro-hibit “wrong contact discontinuities” adjacent to vacuum states. As a consequence, if for instance WL, WR ∈ Ω

have wL = wR, ρL 6= 0 and ρR = 0, then both the weak

solutions W1(t, x) = (RLWR[ρL, ρR](x/t), wL,R), W2(t, x) = ( WL if x < vLt, WR if x > vLt,

satisfy any Kruzhkov type condition with the equality, even if only W1is physically reasonable.

A possible way to uniquely select a solution is by re-quiring that it can be obtained as a many-particle limit analogous to that described in Section 4 for LWR model, see Ref. [17].

6. Bottlenecks

In this section we generalize both models (6) and (9) to reproduce the effects of an exit door placed at x = 0 on the evacuation of a narrow corridor, see Fig. 21.

x 0

Fig. 21.: Evacuation of a corridor through an exit door.

The main feature of a door is given by its capacity Q, which is the maximum number of pedestrians that can flow through it in a given time interval. For this reason at the door position we impose the constraint on the flow

f (ρ) 6 Q. (33)

It is reasonable to assume that Q ∈ (0, fmax),

where fmax= max{f (ρ) : ρ ∈ [0, ρmax]}.

6.1. Microscopic model

At the microscopic level we have to rewrite condition (33) on the flow in terms of velocity function. This can be achieved by considering velocity function

vε(x, ρ) =

( Q

fmaxv(ρ) if |x| 6 ε,

v(ρ) otherwise,

where 2ε > 0 is the size of the door. Indeed (7) leads then to the flux function

(21)

which attains values in [0, Q] for any x ∈ [−ε, ε]. By introducing such velocity function in (6) we obtain the FTL model            ˙ x1(t) = vmax, ˙ xi(t) = vε  xi(t), 1 xi−1(t) − xi(t)  , i ∈ {2, . . . , n}, xi(0) = ¯xi, i ∈ {1, . . . , n}. (34) 6.2. Macroscopic model

Differently from the microscopic level, at the macroscopic level it is not reasonable to consider the size of the door. For this reason condition (33) has to be imposed at x = 0. As a result, the corresponding macroscopic model is given by LWR model (9) coupled with the point constraint on the flow

f (t, 0) 6 Q. (35)

Notice that the classical solution selected by Kruzhkov condition (22) does not in general satisfy (35), as the following example shows.

Example 19. Fix ρ0 ∈ (0, ρmax) and consider the

Cauchy problem for LWR model (9) with constant ini-tial condition

ρ(0, x) = ρ0. (36)

Its unique weak solution satisfying Kruzhkov condition (22) is the constant function ρ ≡ ρ0. However, if f (ρ0) >

Q then such solution does not satisfy (35).

For this reason we introduce a new selection criterion of weak solution, and a new Riemann solver RQ, which we

call constrained Riemann solver. To do so we can adapt the limiting process described in Section 4 as follows. On the basis of microscopic model (34), we consider

           ˙ x0(t) = vmax, ˙ xi(t) = vε  xi(t), ` xi−1(t) − xi(t)  , i ∈ {1, . . . , n}, xi(0) = ¯xi, i ∈ {0, . . . , n},

where x0, . . . , xn ∈ R represent the positions of n + 1

pedestrians, ¯xn < . . . < ¯x0 are their initial positions,

and 2ε > 0 is the size of the door. Introduce then the corresponding discrete density

ρnε(t, x) = n X i=1 ` xi(t) − xi−1(t) 1 [xi−1(t),xi(t))(x).

By letting n go to infinity and ε go to zero, we obtain a a weak solution satisfying both (9) and (35), and we call it constrained solution.

The corresponding constrained Riemann solver RQ

is described below. Let ˇρQ, ˆρQ ∈ [0, ρmax] be implicitly

defined by

f ( ˆρQ) = Q = f ( ˇρQ), 0 6 ˇρQ 6 ρm6 ˆρQ,

where ρm ∈ (0, ρmax) is the maximum of f , namely

f (ρm) = fmax. Then RQ: [0, ρmax]2 → BV(R; [0, ρmax])

is given by RQ[ρL, ρR](ξ) . = ( RLWR[ρL, ˆρQ](ξ) if ξ < 0, RLWR[ ˇρQ, ρR](ξ) if ξ > 0.

A case by case study shows that RQ[ρL, ρR] does

not coincide with RLWR[ρL, ρR] if and only if

f (RLWR[ρL, ρR])(0) > Q.

The constrained solution can also be obtained by ap-plying a WFT algorithm analogous to that described in Section 3.2, but with RQ replacing RLWR any time an

interaction occurs at the constraint position x = 0, see Ref. [62] for the proof.

It is possible to show that the constrained solution is actually the unique weak solution of (9) satisfying the pointwise constraint condition (35) for a.e. t > 0, together with the Kruzhkov type condition

Z R Z+∞ 0 h |ρ(t, x) − k| ∂tϕ(t, x) + F ρ(t, x), k∂xϕ(t, x) i dt dx +2 Z R+ h f (k) − min {f (k), Q} i f (k) ϕ(t, 0) dt > 0 ∀ϕ ∈ C∞∞∞ c (0, +∞) × R; [0, +∞), k ∈ [0, ρmax], where F is given by (23).

In the next example we apply RQ to construct the

constrained solution of the problem considered in Exam-ple 19.

Example 20. Fix ρ0 ∈ (0, ρmax) such that f (ρ0) > Q.

Consider again the Cauchy problem for LWR model (9) with initial condition (36). By applying the constrained Riemann solver RQ we obtain the constrained solution

ρ(t, x) = RQ[ρ0, ρ0](x/t),

(22)

f ρ ˆ ρ ˇ ρ ρ0 Q x x t ˆ σ ¯t ¯ t ˇ σ ¯t x ρ(¯t, ·) ρ0 ˆ ρ ˇ ρ ρ0

Fig. 22.: The constrained solution of the LWR model with point constraint (9), (35). Above ˆ

ρ = ˆρQ(ρ0, ρ0), ˇρ = ˇρQ(ρ0, ρ0), ˆσ = σ(ρ0, ˆρQ) and

ˇ

σ = σ( ˇρQ, ρ0).

maximum principle. Moreover, its total variation instan-taneously increases at time t = 0+. We point out that

the authors of Ref. [62] proved that the total variation of a constrained solution does not increase in the coordinate ψ(ρ) =Rρ

0 |f

0(r)| dr.

In the next two examples we show how Braess’ para-dox and faster-is-slower effects can be reproduced by the constrained LWR model (9), (35) in the framework of pedestrian flows.

Example 22. Fix a < b < 0 and 0 < q < Q < fmax.

Consider a narrow corridor with an exit door placed at x = 0. Assume that the capacity of the door is q ∈ (0, fmax). Consider a crowd to be evacuated, which at

time t = 0 is uniformly distributed in [a, b] with density ρmax. The resulting model is the Cauchy problem for

LWR model (9) with point constraint on the flow (35) and initial condition

ρ(0, x) = ρmax1[a,b](x),

where1[a,b]is the indicator function of the interval [a, b] ⊂

R. The approximate solution can be constructed by ap-plying a WFT analogous to that described in Section 3.2. Below we show some details of this construction for the most representative case, see Fig. 23.

• Fix n ∈ N sufficiently big. Introduce the mesh

Mn = n j ρˇq n : j ∈ [0, n] ∩ Z o a b ˙x = σ(0 ,ˆρ q ) T1 x t 0 0 0 0 ˆ ρq ˇ ρq t1 t2 a b ˙ x= σ(0 ,ρ Qˆ ) ˙x= σ(0 ,ˆρ q) T2 x 0 t c ˆ ρQ ˇ ρQ 0 0 0 0 ˆ ρq ˇ ρq t1 t2 t3

Fig. 23.: Left: Evacuation of a narrow corridor (−∞, 0) with an exit door placed at x = 0 and a crowd initially uniformly distributed in [a, b] with density ρ = ρmax. At time t = t1 the first

pedes-trian crosses the door. At time t = t2 the capacity

q of the exit door is reached and a queue appears in the upstream of the exit door. At time t = T1

the corridor is evacuated. Right: Evacuation of a narrow corridor (−∞, 0) with two doors placed at x = 0 and x = c, and a crowd initially uniformly distributed in [a, b] with density ρ = ρmax. At time

t = t1the first pedestrian crosses the door. At time

t = t2the capacity q of the exit door is reached and

a queue appears in the upstream of the exit door. At time t = t3the capacity Q of the door is reached

and a queue appears in the upstream of the door at x = c. At time t = T2 the corridor is evacuated.

Notice that the evacuation time is higher in the case considered on the left, T1> T2.

∪nρˇq+ j ˆ ρq− ˇρq n : j ∈ (0, n) ∩ N o ∪nρˆq+ j ρmax− ˆρq n : j ∈ [0, n] ∩ Z o . Let ρj ∈ [0, ρmax] be such that Mn = {ρj : j ∈

[0, 3n] ∩ Z} and ρj < ρj+1. Observe that

ρ0= 0, ρn = ˇρq, ρ2n= ˆρq, ρ3n= ρmax.

• Let Rn

LWR and R

n

q be the approximate Riemann

solvers obtained from RLWR and Rq by discretizing

Références

Documents relatifs

The proposed work is based on two methodological ingredients to overcome the difficulties described above: the Sensitivity Equation Method (SEM) for unsteady flows on one side,

Central in the anal- ysis is the question of which mechanism ultimately controls the AMOC response: changes in net evaporation over the At- lantic basin, local processes within

L’archive ouverte pluridisciplinaire HAL, est destinée au dépôt et à la diffusion de documents scientifiques de niveau recherche, publiés ou non, émanant des

Our data on soap-film flows, as well as the available data on pipe flows, are, however, consistent with the predictions of a recently proposed theory of the frictional drag 14–16..

Nous allons montrer que toutes les conditions d’acceptation que nous avons définies sont MSO-définissables et, par des techniques classiques, nous montrerons que tout langage

Enfin, on a réalisé une simulation du temps de chute d’une grappe de commande dans un assemblage combustible en configuration déformée, et montré une meilleure

Secondly, we consider bivariate samples generated from six unimodal distributions that strongly differ in terms of skewness and tails heaviness, and we study the quality of