Loading [MathJax]/jax/output/SVG/jax.js

Properties of the LWR model with time delay

  • Received: 01 March 2020 Revised: 01 October 2020 Published: 08 December 2020
  • 35L65, 90B20, 65M06

  • In this article, we investigate theoretical and numerical properties of the first-order Lighthill-Whitham-Richards (LWR) traffic flow model with time delay. Since standard results from the literature are not directly applicable to the delayed model, we mainly focus on the numerical analysis of the proposed finite difference discretization. The simulation results also show that the delay model is able to capture Stop & Go waves.

    Citation: Simone Göttlich, Elisa Iacomini, Thomas Jung. Properties of the LWR model with time delay[J]. Networks and Heterogeneous Media, 2021, 16(1): 31-47. doi: 10.3934/nhm.2020032

    Related Papers:

    [1] Simone Göttlich, Elisa Iacomini, Thomas Jung . Properties of the LWR model with time delay. Networks and Heterogeneous Media, 2021, 16(1): 31-47. doi: 10.3934/nhm.2020032
    [2] Xiaoqian Gong, Alexander Keimer . On the well-posedness of the "Bando-follow the leader" car following model and a time-delayed version. Networks and Heterogeneous Media, 2023, 18(2): 775-798. doi: 10.3934/nhm.2023033
    [3] Jan Friedrich, Oliver Kolb, Simone Göttlich . A Godunov type scheme for a class of LWR traffic flow models with non-local flux. Networks and Heterogeneous Media, 2018, 13(4): 531-547. doi: 10.3934/nhm.2018024
    [4] Michael Burger, Simone Göttlich, Thomas Jung . Derivation of second order traffic flow models with time delays. Networks and Heterogeneous Media, 2019, 14(2): 265-288. doi: 10.3934/nhm.2019011
    [5] Caterina Balzotti, Simone Göttlich . A two-dimensional multi-class traffic flow model. Networks and Heterogeneous Media, 2021, 16(1): 69-90. doi: 10.3934/nhm.2020034
    [6] Tong Li . Qualitative analysis of some PDE models of traffic flow. Networks and Heterogeneous Media, 2013, 8(3): 773-781. doi: 10.3934/nhm.2013.8.773
    [7] Maya Briani, Rosanna Manzo, Benedetto Piccoli, Luigi Rarità . Estimation of NO$ _{x} $ and O$ _{3} $ reduction by dissipating traffic waves. Networks and Heterogeneous Media, 2024, 19(2): 822-841. doi: 10.3934/nhm.2024037
    [8] Matteo Piu, Gabriella Puppo . Stability analysis of microscopic models for traffic flow with lane changing. Networks and Heterogeneous Media, 2022, 17(4): 495-518. doi: 10.3934/nhm.2022006
    [9] Didier Georges . Infinite-dimensional nonlinear predictive control design for open-channel hydraulic systems. Networks and Heterogeneous Media, 2009, 4(2): 267-285. doi: 10.3934/nhm.2009.4.267
    [10] John D. Towers . The Lax-Friedrichs scheme for interaction between the inviscid Burgers equation and multiple particles. Networks and Heterogeneous Media, 2020, 15(1): 143-169. doi: 10.3934/nhm.2020007
  • In this article, we investigate theoretical and numerical properties of the first-order Lighthill-Whitham-Richards (LWR) traffic flow model with time delay. Since standard results from the literature are not directly applicable to the delayed model, we mainly focus on the numerical analysis of the proposed finite difference discretization. The simulation results also show that the delay model is able to capture Stop & Go waves.



    Nowadays traffic models have become an indispensable tool in the urban and extraurban management of vehicular traffic. Understanding and developing an optimal transport network, with efficient movement of traffic and minimal traffic congestions, will have a great socio- economical impact on the society. This is why in the last decades an intensive research activity in the field of traffic flow modelling flourished.

    Literature about traffic flow is quite large and many methods have been developed resorting to different approaches. Starting from the natural idea of tracking every single vehicle, several microscopic models, based on the idea of Follow-the-Leader, grew-up for computing positions, velocities and accelerations of each car by means of systems of ordinary differential equations (ODEs) [1,8,19,21,44]. Other ways go from kinetic [28,34,45] to macroscopic fluid-dynamic and measures approaches [2,11,12,24,33,38], focusing on averaged quantities, such as the traffic density and the speed of the traffic flow, by means of systems of hyperbolic partial differential equations (PDEs), in particular conservation laws. In this way we loose the detailed level of vehicles' description, indeed they become indistinguishable from each other. The choice of the scale of observation mainly depends on the number of the involved vehicles, the size of the network and so on.

    In this paper we deal with the macroscopic scale, in particular we will focus on first order macroscopic models. The most relevant model in this framework is the LWR model, introduced by Lighthill, Whitham [31] and Richards [36] in the '50. The main idea underlying this approach is that the total mass has to be preserved, since cars can not disappear. Moreover, in this model the mean velocity is supposed to be dependent on the density, thus is closing the equation. On the other hand the lacks of the LWR model are well-known. For example, it fails to generate capacity drop, hysteresis, relaxation, platoon diffusion, or spontaneous congestions like Stop & Go waves, that are typical features of traffic dynamics. These drawbacks are due to the fact that the LWR model represents a simplification of the reality, assuming that accelerations are instantaneous and traffic is described only at the equilibrium.

    In order to overcome these issues, second order models have been proposed, see [1,2,46]. They take into account the non-equilibria states, assuming that accelerations are not instantaneous. To do this, the equation that describes the variation of the velocity in time has to be added to the system, replacing the typical given law of the fist order models. Other ways are also possible to improve first order models, just considering phase transition models [6,14], non-local traffic models [5,15,23,25,29,43] or multi-scale approaches [16,18]. Instead of switching to second order models, we propose a first order macroscopic model with a time delay term in the flux function, for taking into account that the velocity can not change instantaneously. In this framework the delay represents the reaction time of both drivers and vehicles.

    At a microscopic level, a model with time delay appears for the first time in the work done by Newell [32], then similar models are presented in [3,13]. The mathematical tools needed in this framework are not systems of ODEs anymore, but systems of delay differential equations (DDEs), particular differential equations in which the derivative of the unknown function at a certain time is given in terms of the values of the function at previous times. Macroscopic models can be derived from microscopic description following a well-known procedure described in [1,17,21]. Depending on how to treat the delay term, one can recover different macroscopic models, as in [10] or [42], in which a Taylor's approximation is applied to the delay term and the obtained model is a diffusive LWR type model. On the other hand, we want to keep the delay in the explicit form, and therefore avoid the diffusion approximation. The model derived in [9] will be studied in details in the following, investigating carefully its theoretical and numerical features.

    Several delayed-systems are presented in literature, since many phenomena need some transient to become visible or effective: the study of the evolution of the HIV in medicine [20,39], cell population dynamics in biology [26,35], the feedback control loops in control engineering [30], and many applications in mechanics and economics [4], but to the authors best knowledge, they are closer to delayed parabolic partial differential equations or to delayed ordinary differential equations, i.e. they are studied only at a microscopic level.

    In this work instead we deal with a delayed hyperbolic partial differential equation. We will point out similarities and differences with the undelayed model in order to catch the effect of the delay on traffic dynamics, both from theoretical and numerical points of view. Moreover, since we are interested in reproducing real traffic phenomena, the numerical tests are mainly focused on traffic instabilities. In particular we investigate the phenomenon of Stop & Go (S & G) waves, which are a typical feature of congested traffic and represent a real danger for drivers. They lead not only to safety hazard, but they also have a negative impact on fuel consumption and pollution. Indeed a S & G wave is detected when vehicles stop and restart without any apparent reason, generating a wave that travels backward with respect to the cars' trajectories. Since modeling properly this phenomenon is crucial for developing techniques aimed at reducing it, a considerable literature is growing up on this topic. This means that a lot of models have been developed in the last years, i.e. [7,22,28,37,42], and also several real experiments took place, just see [41,47].

    In this framework, our aim is to investigate if our delayed model is able to capture the S & G phenomena and, therefore, to present an easy to use algorithm able to reproduce S & G waves at a macroscopic level. Indeed from the numerical point of view, just an altered Lax Friedrichs numerical scheme will be employed to compute the evolution of the density. In order to validate our model, several numerical tests will be provided for comparing our delayed model with the existing ones.

    Paper organization. In Section 2, we introduce the delayed model and investigate its theoretical properties, as the conservation of mass, the positivity and the boundedness of the solution. After that, we focus on the numerical aspects, presented in Section 3, proposing a suitable numerical scheme and checking the theoretical features still hold. Section 4 is completely devoted to the numerical tests.

    In macroscopic models [24], traffic is described in terms of macroscopic variables such as density ρ=ρ(x,t), that is the number of vehicles per unit length, and the mean velocity V=V(x,t) at the point xR at time t>0.

    The LWR model, introduced by Lighthill, Whitham [31] and Richards [36], is one of the oldest and still most relevant first order macroscopic models for traffic flow. The natural assumption that the total mass is conserved along the road is closed by the assumption that the velocity V=V(ρ) is given as function of the density ρ:

    {tρ(x,t)+x(ρ(x,t) V(ρ(x,t)))=0ρ(x,0)=ρ0(x). (1)

    A lot of possible choices for the function V(ρ) are available in the literature, i.e. the Greenshields function [27] which proposes a linear velocity function:

    V(ρ)=Vmax(1ρρmax). (2)

    In order to simplify the notation, we will consider the normalized quantities ρmax=Vmax=1. Aiming to overcome the drawbacks of LWR model 1 presented in the introduction, in 3 we propose a first order macroscopic model with a time delay term in the flux function, for taking into account that the velocity can not change instantaneously.

    In this framework the delay represents the reaction time of both drivers and vehicles. Such a model has been recovered from a delayed microscopic model, as shown in [9] keeping the delay in the explicit form. Assuming T0 as the time delay, we consider:

    tρ(x,t)+x(ρ(x,t) V(ρ(x,tT)))=0. (3)

    We will call this model delayed LWR model. Note that in the limit case of T=0 the classical LWR model is recovered and therefore, it can be interpreted as a generalization of the LWR model. On the other hand, if the delay is too large and there are suitable initial conditions, cars can overtake or crash each other, for example when a vehicle suddenly brakes and the following car is not reacting in time to slow down.

    Note that in order to guarantee the well-posedness of the problem, we have to provide an initial history function as initial data defined on [T,0], thus we need ρ0(x,t) defined on t[T,0] when starting at t=0.

    As far as it concerns the existence and the uniqueness of the solution for time delayed model, it might be possible to apply the results presented in [29] for non-local conservation laws with time delay. Indeed, as the authors in [29] said, the existence of solutions as well as uniqueness can only be obtained for smooth initial datum and only on a significantly small time horizon. However, the main difference with the model presented here is that we consider a delayed model of local type and convergence results for non-local to local traffic flow models are still missing.

    After introducing the delayed model, we want to investigate its properties. Since this model can be seen as a generalization of the classical LWR model, i.e. when T=0, it is natural to investigate how its properties differ from the undelayed model.

    In the framework of conservation laws and traffic flow models the conservation of the total mass is a crucial property which has to be guaranteed. For the LWR model 1, we have one equation and one conserved quantity, i.e. ρ. Indeed cars do not appear or disappear, they can only enter and leave at the boundaries. We note that the structure of the equation in the delayed and undelayed model stays the same and that we have a flux function that is now dependent on two variables.

    Lemma 2.1. The delayed LWR model 3 conserves the quantity ρ(x,t).

    Proof. We integrate the equation 3 over an arbitrary space interval [a,b] and get

    ddtbaρ(x,t)dx=bax(ρ(x,t)V(ρ(x,tT)))dx=ρ(a,t)V(ρ(a,tT))ρ(b,t)V(ρ(b,tT)).

    Since baρ(x,t)dx is the amount of density in the interval [a,b], ddtbaρ(x,t)dx denotes the change over time for the density. Therefore, the density only changes due to the flux at the boundaries a and b for every space interval.

    We see that the density is still conserved in the delayed model, which is very important for its reliability. The introduction of an explicit time delay therefore does not destroy this property.

    Another property one would ensure is the positivity of the solution. Indeed we want the density to stay positive, as negative densities have no physical meaning.

    Lemma 2.2. Assume we have initial data with non-negative density ρ. For the delayed LWR model, then the density stays non-negative.

    Proof. We rewrite 3 as

    tρ(x,t)=(ρ(x,t)xV(ρ(x,tT))+V(ρ(x,tT))xρ(x,t)). (4)

    For the density to become negative, we need to have ρ=0 and tρ<0.

    We show now that this is not possible. Let us fix a time t such that the density is positive for t<t. We can assume that due to the hypothesis the initial density is positive. Then, at time t, there exists a point x in which ρ(x,t)=0. Since ρ(x,t)>0 for t<t, the density has a minimum in (x,t), therefore xρ(x,t)=0.

    Plugging ρ(x,t)=0 and xρ(x,t)=0 into 4, we obtain:

    tρ(x,t)=(0 xV(ρ(x,tT))+V(ρ(x,tT)) 0)=0.

    We have therefore shown that ρ can not become negative.

    Remark 1. The velocity V in this model is a function of ρ and can be chosen and altered depending on the needs. The properties regarding the velocity in the first order model can therefore be acquired by choosing a suitable function V. For example, we can have lower and upper bounds for the velocity by defining V to be cut at the bounds. This means that infinite (as in the Greenberg model) or negative velocities should be truncated.

    The last property we want to investigate is the boundedness of the solution. In particular, we want to know if there is a maximal density. For the undelayed model, this is guaranteed. For the delayed model, we need to check if this still true.

    Lemma 2.3. Assume V is monotone decreasing and V(ρmax)=0 for ρmax, the maximal density in the classical LWR model. The delayed first order model 3 has no maximal density ρmax.

    Proof. Assume we have a maximal density ρmax. The velocity function V is chosen in such a way that V(ρmax)=0 and monotone decreasing. Then, for an arbitrary point (x,t) where ρ(x,t)=ρmax we have

    tρ(x,t)=ρ(x,t)xV(ρ(x,tT))V(ρ(x,tT))xρ(x,t).

    Since ρ(x,t) is the maximal density, xρ(x,t)=0 if the derivative exists and we have

    tρ(x,t)=ρ(x,t)xV(ρ(x,tT))

    left. We know ρ(x,t)0 and this means that the sign of tρ(x,t) is only dependent on xV(ρ(x,tT)).

    In the undelayed case, we know that xV(ρ(x,t))>0, since V is monotone decreasing and ρ(x,t) is the maximal ρ.

    In the delayed case, we do not have knowledge if ρ(x,tT) is maximal, so we can in general say nothing about xV(ρ(x,tT)). This means, in general, ρ>ρmax is possible.

    Remark 2. Regarding the positivity, we claimed that the choice of V in the first order model is a key to guarantee a positive velocity. We here see, due to the fact that ρ overshoots any ρmax, that the classical choices for V must be altered to avoid negative velocities, i.e. we need to cut the function, for example assuming that V(ρ)=0 when ρ>ρmax.

    Remark 3. If the density ρ>ρmax, the model is not reliable any more. On the other hand this situation could not be avoided since rear-end collisions are actually possible in real situations.

    After the investigations on the analytical properties of the delayed LWR model, let us focus on its numerical counterpart.

    Since 3 is a hyperbolic partial differential equation, we can employ the Lax-Friedrichs method for the numerical approximation. To do that, we first introduce space and time steps Δx, Δt>0 and a grid in space {xj=jΔx,jZ} and time {tn=nΔt,nN}. Since we are in the framework of finite difference approximation, discretized variables are expressed by ρnj=ρ(xj,tn), where i is the space and n the time index. Also, we have ΔtT to be able to treat the delay.

    The Lax-Friedrichs numerical scheme for 1 is stated by:

    ρn+1j=12(ρnj+1+ρnj1)Δt2Δx(f(ρnj+1)f(ρnj1)).

    Now using the structure of 3, we can identify a flux function f(ρ(x,tT),ρ(x,t))=V(ρ(x,tT))ρ(x,t). Plugging this into the Lax-Friedrichs numerical scheme, we end up with an altered Lax-Friedrichs numerical scheme

    ρn+1j=12(ρnj+1+ρnj1)Δt2Δx(f(ρnTΔj+1,ρnj+1)f(ρnTΔj1,ρnj1)), (5)

    where jZ,nN and TΔ is the number of steps that make up the time delay T. In order to guarantee the well-posedness of the discrete problem, we have to provide an initial history function as initial data defined on [T,0], as we said above for the continuous problem. The simplest choice one can do is to consider ρ0(x,t) as a constant function on t[T,0] when starting at t=0. In the following we will assume that ρ0(x,0) is constant in t for t[T,0]. The Lax-Friedrichs numerical scheme has a CFL condition in the classical case, which is given as ΔtΔxmaxk(λk), where λk are the eigenvalues of the jacobian matrix of f. We also expect to find a CFL condition in the delayed case, but a priori it is not clear how this condition may look like. In the following, we want to investigate some properties of this method, and in this process we will find an appropriate CFL condition. For the sake of the calculations, we assume the velocity function V to be the Greenshields function, or a cut variation of it, where we have |V(ρ)|ρmax with Vmax=1.

    First, we check if the conservation property is preserved from the numerical scheme. Here, we assume the density to be on a compact support, so we do not have infinite density initially. We get

    Δxjρn+1j=Δxj12(ρnj+1+ρnj1)Δt2Δx(V(ρnTΔj+1)ρnj+1V(ρnTΔj1)ρnj1), (6)

    where the part

    j12(ρnj+1+ρnj1)=jρnj

    and the part

    jΔt2Δx(V(ρnTΔj+1)ρnj+1V(ρnTΔj1)ρnj1)

    is a telescope sum and equals zero due to the compact support. This gives us

    Δxjρn+1j=Δxjρnj

    and therefore conservation.

    We show that Lemma 2.2 holds also at the discrete level under a certain CFL condition. Starting with 5, we see that 12(ρnj+1+ρnj1) is always positive, since we assume ρn to be positive. Indeed, if ρn=0, ρn+1j will also be zero. So to guarantee positivity, we need to guarantee

    Δt2Δx(V(ρnTΔj+1)ρnj+1V(ρnTΔj1)ρnj1)12(ρnj+1+ρnj1). (7)

    We compute a CFL-condition, namely ΔtΔxmax{|ρn|,|ρnTΔ|}. Therefore, we get for the left-hand-side of 7

    12max{|ρn|,|ρnTΔ|}(V(ρnTΔj+1)ρnj+1V(ρnTΔj1)ρnj1)=V(ρnTΔj+1)2max{|ρn|,|ρnTΔ|}ρnj+1V(ρnTΔj1)2max{|ρn|,|ρnTΔ|}ρnj1max{|ρn|,|ρnTΔ|}2max{|ρn|,|ρnTΔ|}ρnj+1+max{|ρn|,|ρnTΔ|}2max{|ρn|,|ρnTΔ|}ρnj1=12(ρnj+1+ρnj1), (8)

    which shows the positivity for this CFL-condition. Here we use that |V(ρ)|max{|ρn|,|ρnTΔ|} if ρ is positive.

    The crucial role played by the new CFL condition is explained at the beginning of the Section 4.

    Focusing on the boundedness of the discrete solution, we look for an estimate in the norm ||L. Since we have positivity, only an upper bound for ρ in 5 is required. We assume that the data at time tn has an upper bound which we denote with ρnmax. We further denote V(ρnTΔj)ρnj=f(ρnTΔj,ρnj) and get for a special ξ=(ξ1,ξ2) using the mean value theorem:

    |ρn+1j|=|12(ρnj+1+ρnj1)Δt2Δx(f(ρnTΔj+1,ρnj+1)f(ρnTΔj1,ρnj1))|=|12(ρnj+1+ρnj1)+Δt2Δx(f(ξ1,ξ2)(ρnj1ρnj+1ρnTΔj1ρnTΔj+1))||12(ρnj+1+ρnj1)|+Δt2Δx|(V(ξ2)(ρnj1ρnj+1)+ξ1(ρnTΔj+1ρnTΔj1))|CFL12|ρnj+1+ρnj1|+12|ρnj1ρnTΔj1+ρnTΔj+1ρnj+1|12|2ρnj1ρnTΔj1+ρnTΔj+1|ρnmax+12|ρnTΔj+1ρnTΔj1|2max{|ρn|,|ρnTΔ|}. (9)

    We use again the positivity and the CFL-condition, estimating the velocity by the maximal density value and V(ξ1)1.

    Remark 4. We can find a different estimate for ρn+1j, depending on T. Note that the estimate in 9 is more accurate. We start in the same way as above, and have

    |ρn+1j|12|2ρnj1ρnTΔj1+ρnTΔj+1|.

    Rearranging gives us

    12(ρnj1+ρnj1ρnTΔj1+ρnTΔj+1)=12(ρnj1ρnTΔj1+ρnTΔj+1+ρnj1),

    and by again using the mean value theorem for ρj1(t) we get

    ρn+1j12(Ttρj1+ρnj1+ρnTΔj+1)max{|ρn|,|ρnTΔ|}+12T||tρ||L. (10)

    Let us look for an estimate on the Total Variation for the method 5. The first thing we check is the difference between the velocity function in two cells.

    V(ρnTΔj+1)V(ρnTΔj)=(1ρnTΔj+1)(1ρnTΔj)=ρnTΔjρnTΔj+1=ΔnTΔj+12. (11)

    Then, we look at the difference between two neighboring cells, where we use 5, 11 and the notation Δnj+12=ρnj+1ρnj:

    Δn+1j+12=ρn+1j+1ρn+1j=12(Δnj+32+Δnj12)Δt2Δx(V(ρnTΔj+2)Δnj+32V(ρnTΔj)Δnj12ρnj+1ΔnTΔj+32+ρnj1ΔnTΔj12). (12)

    The Total Variation at time tn+1 (denoted by TV(ρn+1Δ)) is given as j|Δn+1j+12|. For the next step and with 12 as well as shifting the indices, we get

    j|Δn+1j+12|=j|ρn+1j+1ρn+1j|=j(1+2|Δt2ΔxρnTΔj+1|+2|Δt2Δxρnj+1|)max{|Δnj+12|,|ΔnTΔj+12|}. (13)

    If we introduce a CFL-condition of the type Δt2Δxmax{|ρnTΔj|,|ρnj|} (i.e. not as strong as above), the dependency on j disappears and we get

    (5+1max{|ρnTΔj|,|ρnj|})jmax{|Δnj+12|,|ΔnTΔj+12|}.

    By assumption, TV(ρnΔ) and TV(ρnTΔΔ) are finite, so we can estimate further

    TV(ρn+1Δ)2(5+1max{|ρnTΔj|,|ρnj|})max{TV(ρnΔ),TV(ρnTΔΔ)}. (14)

    For the estimate in time, we look at

    j|ρn+1jρnj|.

    By simply plugging in 5 and using the same CFL-condition, we can write

    j|ρn+1jρnj|j12(|ρnj+1|+|ρnj1|)+Δt2Δx(|V(ρnTΔj+1)||ρnj+1|+|V(ρnTΔj1)||ρnj1|)+|ρnj|j2max{|ρnTΔj|,|ρnj|}+2+2max{|ρnTΔj|,|ρnj|}=j4max{|ρnTΔj|,|ρnj|}+2.

    We now have a BV-Bound in space as well as in time, which gives us all the desired BV estimates.

    With the CFL-condition we introduced, the time step can become smaller every step, since we have no maximum principle. Here, we want to see if we can actually reach every time horizon. Therefore, we plug our L-bound into the CFL and get

    ΔtΔxmax{|ρn|,|ρnTΔ|}Δx(32)n||ρ0||L.

    Now, the time horizon we reach with n time steps is given by

    tn=ni=1Δt=Δx||ρ0||Lni=1(23)i.

    So for infinite time steps n, we end up with a geometric series which converges to 3Δx||ρ0||L. That is the time horizon we can guarantee with this estimate.

    With the alternative estimate that depends on T, the time horizon we can reach is dependent also on T. With n time steps, we can now reach the time horizon

    tn=Δxni=11ρ0max+i12||tρ||T.

    This is basically a harmonic series shifted and with a factor, but it is divergent to if the factor is not zero. This also means, that for small T, the steps can be larger. Furthermore, with this second estimate, we can guarantee to reach every time horizon.

    This section is devoted to the numerical simulation results for the model presented above, focusing in particular on the S & G waves phenomenon, a typical feature of congested traffic, detected when vehicles stop and restart without any apparent reason, generating a wave that travels backward with respect to the cars' trajectories.

    Starting from empirical observations and the work done in [18,47], let us assume the velocity function as follow:

    V(ρ)={Vmaxρρfα (1ρ1ρc)ρf<ρ<ρc0ρρc (15)

    where α>0 is a parameter, ρc(0,ρmax] and ρf[0,ρmax) are two density thresholds. In particular ρc represents the so-called safe distance at the macroscopic level: if the density is higher than ρc, vehicles do not respect the safe distance so the desired velocity has to be 0, indeed they should stop.

    On the other hand, if the density is very low, which means that vehicles are far enough from each others, the desired velocity is the maximum one.

    Note that 15 respects the hypothesis |V(ρ)|ρmax choosing Vmax=1. Moreover, depending on the choice made for α, the velocity function can be discontinuous.

    For the discretization, let us assume the space interval [a,b]=[0,1], Δx=0.02 and periodic boundary conditions. Moreover, the time step Δt is chosen in such a way the CFL condition is satisfied. We denote with Tf the final time of the simulation. The density thresholds are ρc=0.75 and ρf=0.2 as real data suggests. The delay term depends on the CFL condition and the initial data, for this reason each numerical test has its time delay interval which ensures the reliability of the model, i.e. ρρmax. However, in general, one has to assume TΔ one order of magnitude greater than Δt to see the effect, i.e. TΔ(10Δt,20Δt).

    Remark 5. Let us point out the relevance of the new CFL condition. If we compute the solution of 5 with the classical CFL condition, we will end up with spurious oscillations as in Fig. 1(left)-2(left). Applying the computed CFL condition, these oscillations disappear, Fig. 1(right)-2(right). The parameters are Δx=0.1, Tf=10 TΔ=1 and the time step is chosen Δt=0.5Δx2max{|ρn|,|ρnTΔ|} in the latter and Δt=1.5Δx2ρmax for the former case. In this way we can focus only on the CFL condition, without any influence from the delay term, since the value of the delay is fixed and not depending on the time step.

    Figure 1.  Comparison between density profiles computed with (right) and without(left) CFL condition in case of a rarefaction wave.
    Figure 2.  Comparison between density profiles computed with (right) and without (left) CFL condition in a shock framework.

    In order to be more comprehensive as possible in reproducing S & G waves, let us describe first the backward propagation of the perturbation. After that we will focus also on the triggering of this phenomenon.

    Test 0. In order to point out the crucial role played by the delay in this framework, let us compare the evolution of the density obtained with the delayed model and the classical LWR model, or, in other words, when TΔ=0.

    Assume as initial data ρ0(x)=58+18sin(2πx). Moreover the time step is Δt=0.01, the delay TΔ=15Δt in the delayed case and the final time of the simulation is Tf=10.

    It is evident how the LWR model smears out the perturbations in the initial data and after a certain time the density becomes constant on the whole road, see Fig. 3(right). On the other hand, the delayed model preserves the perturbations and also makes them increase as usually happens in traffic evolution, Fig. 3(left).

    Figure 3.  Test 0: Comparing the density evolution computed by the delayed model (left) and the LWR model (right).

    Remark 6. Unfortunately, we have no definition of weak solutions for the continuous delayed problem. Therefore, we can not show the convergence of the numerical scheme 5 to any weak solution of 3. On the other hand, we provide a comparison with the reference solution to show the numerical convergence. For reference solution we mean the solution computed with the same numerical scheme but with a high resolution of the mesh, i.e. with Δx and Δt very small. In Fig. 4, we show the comparison between three density profiles corresponding to different grid steps size, i.e. Δx=102,103,104 and Δt=0.5102,103,104 respectively, at the same time, with TΔ=5Δt. We can note that the difference between the density profiles is decreasing as the grid is more and more refined.

    Figure 4.  Test 0: Comparison between density profiles corresponding to different grid steps size.

    Remark 7. We have to be very careful in choosing the delay term. Indeed if the delay is too small, we recover a situation very similar to the LWR model, but, on the other hand, if the delay is too high, i.e. TΔ=18Δt, the hypothesis on the model are no longer satisfied and the density grows more than 1, so the model has no sense anymore, see Fig. 5.

    Figure 5.  Test 0: Density evolution and profile, at time T=13Tf, in case of a too high delay.

    In the following, we will consider the delay as the maximum allowed by the model feasibility.

    Test 1. In this numerical simulation we want to reproduce with our model the tests presented in [7]. Starting from the same initial data our aim is to recover a similar behaviour for the density. In [7], a nonlinear 2-equations discrete velocity model is implemented to compute the density evolution:

    tρ+xq=0tq+Hq1ρxρ+(1Hq1ρxq)=1ε(qF(ρ)),

    where q is the flux, F is the fundamental diagram and H is a measure for the look ahead and the nonlocality of the equations. Such a model converges to the LWR type equations in the relaxation limit but shows also similarities to the Aw-Rascle model. The main difference to the delayed model is that at the macroscopic level they are composed of a system of two PDEs while the delayed model is described only by a conservation law with a time delay in the velocity term.

    Let us assume ρ0(x,k)=58+18sin(2kπx) for k=1,2, Tf=10 and Δt=0.01. We consider TΔ=16Δt and TΔ=22Δt, respectively if k=1,2.

    In Fig. 6, the density values on the (x,t)-plane are represented by the colors as shown in the colorbar. We obtain persistent waves where the number of waves is directly related to the perturbations in the initial condition, as in [7]. Numerically, we recover this result if TΔ(11Δt,16Δt) for k=1 and TΔ(18Δt,22Δt) for k=2.

    Figure 6.  Test 1: Reproducing the simulation presented in [7], with ρ0(x,1) on the left and ρ0(x,2) on the right.

    Test 2. We consider the discrete delayed model 5 with the velocity function 15 and initial data:

    ρ0(x)={0.6x<0.50.1x0.5. (16)

    Let us assume the time delay as TΔ=10Δt, Tf=3.5 and the time step Δt=0.01 such that the CFL condition is satisfied. Moreover, the coefficient α is chosen in such a way as to make the velocity function continuous. Looking at Fig. 7(left), one can immediately note that the initial slowdown in the first half of the road is increasing in time until vehicles completely stop since ρρc, as the colorbar suggests. This is a typical S & G wave behaviour, as we can see also in [42], where they compute the density evolution considering another extension of the LWR model. Indeed they assume that the velocity term depends on ρ, the reaction time τ and the derivative in space of the velocity itself xV(ρ):

    tρ+x(ρV(ρ1τxV(ρ)))=0. (17)
    Figure 7.  Test 2: Density values in the (x,t)-plane (left) and density profile at time T=Tf (right).

    Investigating the density profile at the end of the simulation, we are able to recognize a well-defined S & G wave profile, see Fig. 7(right), as described in [22].

    Moreover, let us note that if the delay is smaller, i.e. TΔ=4Δt, the model is no longer able to preserve the perturbation and the density profile becomes smoother, see Fig. 8.

    Figure 8.  Test 2: Density values in the (x,t)-plane with low delay term, TΔ=4Δt, (left) and density profile at T=Tf (right).
    Figure 9.  Triggering of Stop & Go waves: Density values in the (x,t)-plane.

    Choosing the delay term 7Δt<TΔ<11Δt, we are able to recover the S & G behaviour.

    Let us now focus on the triggering of S & G waves. Starting from a small perturbation, i.e. a slowdown, in the initial data, our aim is to find out if it is possible to recover a S & G wave. In this direction, let us consider the example presented in [18], in which the initial data is given by:

    ρ0(x)={0.351.34x1.3420.2elsewhere, (18)

    and Dirichlet boundary conditions. The initial data is modelling a small region, a cell, where vehicles are moving slower than elsewhere and therefore the density is higher in that cell. This slowdown can be caused by the presence of sags, road sections in which gradient changes significantly from downwards to upwards [37], or the presence of a school zone in which the velocity has to be reduced, [40].

    In [18], the density evolution is computed by coupling the LWR model with a second order microscopic model, specifically conceived to reproduce S & G waves. Instead of switching to multiscale models in which we have to manage the microscopic model too, let us see if we can recover the density evolution with the delayed model.

    Assuming Δt=0.009, Tf=5 and TΔ=21Δt, we note that the initial perturbation increases and moves backward as the time increases, as it happens in the multiscale case.

    Therefore, the delayed model is able to reproduce the triggering of S & G waves too, not only their backward propagation.

    In this paper, we have introduced theoretical and numerical properties of the delayed LWR traffic model. While the derivation of the model has been done in [9], the numerical behavior of the delayed model has not been studied intensively before.

    Starting from the undelayed scenario, we investigated the theoretical features of the delayed model to point out the numerical properties of the scheme and we proposed an altered Lax-Friedrichs numerical scheme to compute the evolution of the density. The key observation therefore is that the delayed model is really able to reproduce Stop & Go waves for the right choice of parameters. Comparisons to already existing results from the literature also underline this characteristic.

    Future works will include the extension to networks as well as parameter estimation techniques to determine the time delay.



    [1] Derivation of continuum flow traffic models from microscopic follow-the-leader models. SIAM J. Appl. Math. (2002) 63: 259-278.
    [2] Resurrection of "second order" models of traffic flow?. SIAM J. Appl. Math. (2000) 60: 916-938.
    [3] Dynamical model of traffic congestion and numerical simulation. Physical review E (1995) 51: 10-35.
    [4] C. Bianca, M. Ferrara and L. Guerrini, The time delays' effects on the qualitative behavior of an economic growth model, Abstract and Applied Analysis, 2013 (2013). doi: 10.1155/2013/901014
    [5] Well-posedness of a conservation law with non-local flux arising in traffic flow modeling. Numer. Math. (2016) 132: 217-241.
    [6] A general phase transition model for vehicular traffic. SIAM Journal on Applied Mathematics (2011) 71: 107-127.
    [7] A nonlinear discrete velocity relaxation model for traffic flow. SIAM Journal on Applied Mathematics (2018) 78: 2891-2917.
    [8] M. Braskstone and M. McDonald, Car following: A historical review, transportation research part f. 2., Pergamon, 2000.
    [9] Derivation of a first order traffic flow model of Lighthill-Whitham-Richards type. IFAC-PapersOnLine (2018) 51: 49-54.
    [10] Derivation of second order traffic flow models with time delays. Netw. Heterog. Media (2019) 14: 265-288.
    [11] S. Cacace, F. Camilli, R. De Maio and A. Tosin, A measure theoretic approach to traffic flow optimisation on networks, European Journal of Applied Mathematics, (2018), 1-23. doi: 10.1017/S0956792518000621
    [12] Measure-valued solutions to nonlocal transport equations on networks. Journal of Differential Equations (2018) 264: 7213-7241.
    [13] Traffic dynamics: Studies in car following. Operations Research (1958) 6: 165-184.
    [14] Hyperbolic phase transitions in traffic flow. SIAM Journal on Applied Mathematics (2003) 63: 708-721.
    [15] Non local balance laws in traffic models and crystal growth. ZAMM Z. Angew. Math. Mech. (2007) 87: 449-461.
    [16] A mixed ode-pde model for vehicular traffic. Mathematical Methods in the Applied Sciences (2015) 38: 1292-1302.
    [17] R. M. Colombo and E. Rossi, On the micro-macro limit in traffic flow, Rendiconti del Seminario Matematico della Università di Padova, 131 (2014), 217-236. doi: 10.4171/RSMUP/131-13
    [18] E. Cristiani and E. Iacomini, An interface-free multi-scale multi-order model for traffic flow, Discrete & Continuous Dynamical Systems-Series B, 25 (2019). doi: 10.3934/dcdsb.2019135
    [19] On the micro-to-macro limit for first-order traffic flow models on networks. Netw. Heterog. Media (2016) 11: 395-413.
    [20] A delay-differential equation model of hiv infection of cd4+ t-cells. Mathematical Biosciences (2000) 165: 27-39.
    [21] Many particle approximation of the Aw-Rascle-Zhang second order model for vehicular traffic. Math. Biosci. Eng. (2017) 14: 127-141.
    [22] Self-sustained nonlinear waves in traffic flow. Phys. Rev. E (2009) 79: 56-113.
    [23] A Godunov type scheme for a class of LWR traffic flow models with non-local flux. Netw. Heterog. Media (2018) 13: 531-547.
    [24] M. Garavello and B. Piccoli, Traffic Flow on Networks, AIMS Series on Applied Mathematics, 2006.
    [25] Well-posedness and finite volume approximations of the LWR traffic flow model with non-local velocity. Netw. Heterog. Media (2016) 11: 107-121.
    [26] D. Green Jr. and H. W. Stech, Diffusion and hereditary effects in a class of population models, in Differential Equations and Applications in Ecology, Epidemics, and Population Problems, Elsevier, 1981, 19-28.
    [27] B. D. Greenshields, A study in highway capacity, Highway Research Board Proc., 1935, (1935), 448-477.
    [28] The bgk approximation of kinetic models for traffic. Kinetic & Related Models (2020) 13: 279-307.
    [29] A. Keimer and L. Pflug, Nonlocal conservation laws with time delay, Nonlinear Differential Equations and Applications NoDEA, 26 (2019), 54. doi: 10.1007/s00030-019-0597-z
    [30] Feedback stabilization of linear systems with delayed control. IEEE Transactions on Automatic control (1980) 25: 266-269.
    [31] On kinematic waves Ⅱ. A theory of traffic flow on long crowded roads. Proceedings of the Royal Society of London. Series A. Mathematical and Physical Sciences (1955) 229: 317-345.
    [32] Nonlinear effects in the dynamics of car following. Operations Research (1961) 9: 209-229.
    [33] B. Piccoli and A. Tosin, Vehicular traffic: A review of continuum mathematical models, in Encyclopedia of Complexity and Systems Science, Springer, New York, NY, 2009, 9727-9749. doi: 10.1007/978-1-4614-1806-1_112
    [34] G. Puppo, M. Semplice, A. Tosin and G. Visconti, Kinetic models for traffic flow resulting in a reduced space of microscopic velocities, Kinetic & Related Models, 10 (2016), 823. doi: 10.3934/krm.2017033
    [35] A. D. Rey and C. Mackey, Multistability and boundary layer development in a transport equation, Canadian Applied Mathematics Quarterly, (1993), 61-81.
    [36] Shock waves on the highway. Operations Research (1956) 4: 42-51.
    [37] Empirical analysis of the causes of stop-and-go waves at sags. IET Intell. Transp. Syst. (2014) 8: 499-506.
    [38] M. D. Rosini, Macroscopic Models for Vehicular Flows and Crowd Dynamics: Theory and Applications, Understanding Complex Systems, Springer International Publishing, 2013. doi: 10.1007/978-3-319-00155-5
    [39] H. L. Smith, An Introduction to Delay Differential Equations with Applications to the Life Sciences, 57, Springer, New York, 2011. doi: 10.1007/978-1-4419-7646-8
    [40] A second order traffic flow model with lane changing. Journal of Scientific Computing (2019) 81: 1429-1445.
    [41] R. E. Stern, et al., Dissipation of stop-and-go waves via control of autonomous vehicles: Field experiments, Transportation Res. Part C, 89 (2018), 205-221.
    [42] From traffic and pedestrian follow-the-leader models with reaction time to first order convection-diffusion flow models. SIAM Journal on Applied Mathematics (2018) 78: 63-79.
    [43] M. Treiber, A. Hennecke and D. Helbing, Derivation, properties, and simulation of a gas-kinetic-based, nonlocal traffic model, Physical Review E, 59 (1999), 239.
    [44] M. Treiber and A. Kesting, Traffic flow dynamics, in Traffic Flow Dynamics: Data, Models and Simulation, Springer-Verlag Berlin Heidelberg, 2013. doi: 10.1007/978-3-642-32460-4
    [45] Multivalued fundamental diagrams of traffic flow in the kinetic Fokker-Planck limit. Multiscale Model. Simul. (2017) 15: 1267-1293.
    [46] A non-equilibrium traffic model devoid of gas-like behavior. Transportation Research Part B: Methodological (2002) 36: 275-290.
    [47] A unified follow-the-leader model for vehicle, bicycle and pedestrian traffic. Transportation Res. Part B (2017) 105: 315-327.
  • This article has been cited by:

    1. NGUYEN Dinh Dung, LAX - FRIEDRICHS REGULARIZATION DIFFERENCE ALGORITHM FOR TRAFFIC DENSITY FORECASTING PROBLEM WITH DELAY, 2023, 53, 30304563, 54, 10.56824/vujs.2023a144
    2. Zhaobin Mo, Xu Chen, Xuan Di, Elisa Iacomini, Chiara Segala, Michael Herty, Mathieu Lauriere, A Game-Theoretic Framework for Generic Second-Order Traffic Flow Models Using Mean Field Games and Adversarial Inverse Reinforcement Learning, 2024, 0041-1655, 10.1287/trsc.2024.0532
    3. Ilaria Ciaramaglia, Paola Goatin, Gabriella Puppo, Non-local traffic flow models with time delay: Well-posedness and numerical approximation, 2024, 0, 1531-3492, 0, 10.3934/dcdsb.2024113
  • Reader Comments
  • © 2021 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

Metrics

Article views(2405) PDF downloads(297) Cited by(2)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog