Loading [MathJax]/jax/element/mml/optable/GeneralPunctuation.js

A multiclass Lighthill-Whitham-Richards traffic model with a discontinuous velocity function

  • Received: 01 October 2020 Revised: 01 December 2020 Published: 18 January 2021
  • Primary: 65M06; Secondary: 35L45, 35L65, 76A99

  • The well-known Lighthill-Whitham-Richards (LWR) kinematic model of traffic flow models the evolution of the local density of cars by a nonlinear scalar conservation law. The transition between free and congested flow regimes can be described by a flux or velocity function that has a discontinuity at a determined density. A numerical scheme to handle the resulting LWR model with discontinuous velocity was proposed in [J.D. Towers, A splitting algorithm for LWR traffic models with flux discontinuities in the unknown, J. Comput. Phys., 421 (2020), article 109722]. A similar scheme is constructed by decomposing the discontinuous velocity function into a Lipschitz continuous function plus a Heaviside function and designing a corresponding splitting scheme. The part of the scheme related to the discontinuous flux is handled by a semi-implicit step that does, however, not involve the solution of systems of linear or nonlinear equations. It is proved that the whole scheme converges to a weak solution in the scalar case. The scheme can in a straightforward manner be extended to the multiclass LWR (MCLWR) model, which is defined by a hyperbolic system of N conservation laws for N driver classes that are distinguished by their preferential velocities. It is shown that the multiclass scheme satisfies an invariant region principle, that is, all densities are nonnegative and their sum does not exceed a maximum value. In the scalar and multiclass cases no flux regularization or Riemann solver is involved, and the CFL condition is not more restrictive than for an explicit scheme for the continuous part of the flux. Numerical tests for the scalar and multiclass cases are presented.

    Citation: Raimund Bürger, Christophe Chalons, Rafael Ordoñez, Luis Miguel Villada. A multiclass Lighthill-Whitham-Richards traffic model with a discontinuous velocity function[J]. Networks and Heterogeneous Media, 2021, 16(2): 187-219. doi: 10.3934/nhm.2021004

    Related Papers:

    [1] Raimund Bürger, Christophe Chalons, Rafael Ordoñez, Luis Miguel Villada . A multiclass Lighthill-Whitham-Richards traffic model with a discontinuous velocity function. Networks and Heterogeneous Media, 2021, 16(2): 187-219. doi: 10.3934/nhm.2021004
    [2] Helge Holden, Nils Henrik Risebro . Follow-the-Leader models can be viewed as a numerical approximation to the Lighthill-Whitham-Richards model for traffic flow. Networks and Heterogeneous Media, 2018, 13(3): 409-421. doi: 10.3934/nhm.2018018
    [3] Mauro Garavello . A review of conservation laws on networks. Networks and Heterogeneous Media, 2010, 5(3): 565-581. doi: 10.3934/nhm.2010.5.565
    [4] Paola Goatin, Sheila Scialanga . Well-posedness and finite volume approximations of the LWR traffic flow model with non-local velocity. Networks and Heterogeneous Media, 2016, 11(1): 107-121. doi: 10.3934/nhm.2016.11.107
    [5] Michael Herty, Lorenzo Pareschi, Mohammed Seaïd . Enskog-like discrete velocity models for vehicular traffic flow. Networks and Heterogeneous Media, 2007, 2(3): 481-496. doi: 10.3934/nhm.2007.2.481
    [6] Edward S. Canepa, Alexandre M. Bayen, Christian G. Claudel . Spoofing cyber attack detection in probe-based traffic monitoring systems using mixed integer linear programming. Networks and Heterogeneous Media, 2013, 8(3): 783-802. doi: 10.3934/nhm.2013.8.783
    [7] Simone Göttlich, Ute Ziegler, Michael Herty . Numerical discretization of Hamilton--Jacobi equations on networks. Networks and Heterogeneous Media, 2013, 8(3): 685-705. doi: 10.3934/nhm.2013.8.685
    [8] Ye Sun, Daniel B. Work . Error bounds for Kalman filters on traffic networks. Networks and Heterogeneous Media, 2018, 13(2): 261-295. doi: 10.3934/nhm.2018012
    [9] Michael Herty, J.-P. Lebacque, S. Moutari . A novel model for intersections of vehicular traffic flow. Networks and Heterogeneous Media, 2009, 4(4): 813-826. doi: 10.3934/nhm.2009.4.813
    [10] 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
  • The well-known Lighthill-Whitham-Richards (LWR) kinematic model of traffic flow models the evolution of the local density of cars by a nonlinear scalar conservation law. The transition between free and congested flow regimes can be described by a flux or velocity function that has a discontinuity at a determined density. A numerical scheme to handle the resulting LWR model with discontinuous velocity was proposed in [J.D. Towers, A splitting algorithm for LWR traffic models with flux discontinuities in the unknown, J. Comput. Phys., 421 (2020), article 109722]. A similar scheme is constructed by decomposing the discontinuous velocity function into a Lipschitz continuous function plus a Heaviside function and designing a corresponding splitting scheme. The part of the scheme related to the discontinuous flux is handled by a semi-implicit step that does, however, not involve the solution of systems of linear or nonlinear equations. It is proved that the whole scheme converges to a weak solution in the scalar case. The scheme can in a straightforward manner be extended to the multiclass LWR (MCLWR) model, which is defined by a hyperbolic system of N conservation laws for N driver classes that are distinguished by their preferential velocities. It is shown that the multiclass scheme satisfies an invariant region principle, that is, all densities are nonnegative and their sum does not exceed a maximum value. In the scalar and multiclass cases no flux regularization or Riemann solver is involved, and the CFL condition is not more restrictive than for an explicit scheme for the continuous part of the flux. Numerical tests for the scalar and multiclass cases are presented.



    The multiclass Lighthill-Whitham-Richards (MCLWR) model is a generalization of the well-known Lighthill-Whitham-Richards (LWR) model [25,28] to multiple classes of drivers and was formulated independently by Wong and Wong [31] and Benzoni-Gavage and Colombo [1]. The model is given by the following system of conservation laws in one space dimension, where the sought unknowns are the densities ϕi=ϕi(x,t) of vehicles of class i, i=1,,N, as a function of distance x and time t [1,31]:

    tϕi+x(ϕivi(ϕ))=0,i=1,,N. (1.1)

    Here ϕ=ϕ1++ϕN denotes the total density of vehicles. The velocity function vi is assumed to depend on ϕ, where we assume that

    vi(ϕ)=vmaxiV(ϕ),i=1,,N, (1.2)

    where vmax1<vmax2<<vmaxN are the maximum velocities of the N classes of vehicles and V is a hindrance function that models the drivers' attitude to reduce speed in the presence of other cars. This function is usually assumed to be continuous and piecewise smooth on an interval [0,ϕmax], where ϕmax>0 denotes a maximum vehicle density, with

    V(0)=1,V(ϕ)<0,V(ϕmax)=0.

    The simplest function having all these properties is the linear interpolation V(ϕ)=1ϕ/ϕmax. However, equation (1.1) is studied herein under the assumption that V is piecewise continuous with one decreasing jump at a density value ϕ(0,ϕmax), that is

    V(ϕ)={Vf(ϕ)for 0ϕϕ,Vc(ϕ)for ϕ<ϕϕmax,VfC1[0,ϕ],VcC1[ϕ,ϕmax],Vf(0)=1,Vf(ϕ)0on [0,ϕ],Vc(ϕ)0on [ϕ,ϕmax],Vf(ϕmax)=0,αV:=Vf(ϕ)Vc(ϕ)>0. (1.3)

    We consider (1.1) on the domain ΠT:=(L,L)×(0,T), where L>0 and T>0, along with the initial and boundary conditions

    ϕi(x,0)=ϕi,0(x)[0,ϕmax],x(L,L),ϕi(L,t)=ri(t)[0,ϕmax],t(0,T),ϕi(L,t)=si(t)[0,ϕmax],t(0,T),i=1,,N; (1.4a)
    F(t)(vmax)Ts(t)˜V(s(t)),t(0,T);vmax:=(vmax1,,vmaxN)T. (1.4b)

    Here and throughout the paper, we denote by a tilde the multivalued version of a given discontinuous function. The non-standard boundary condition (1.4b) on the total density is required in case that s(t)=ϕ, where s(t):=(s1(t),,sN(t))T and s(t)=s1(t)++sN(t). This implies that we assign values to F(t) according to

    F(t)={(vmax)Ts(t)V(ϕ)if the traffic ahead of x=L is free-flowing,(vmax)Ts(t)V(ϕ+)if the traffic ahead of x=L is congested. (1.5)

    This assumption is motivated in a wider sense by models of phase transitions between free and congested traffic flow regimes [13,14], and more specifically by treatments of the single-class scalar version of (1.1)–(1.4). In the scalar case the model can be formulated as the following initial-boundary value problem for a scalar conservation law defined on ΠT:

    tϕ+xf(ϕ)=0,(x,t)ΠTf(ϕ)=vmaxϕV(ϕ),ϕ(x,0)=ϕ0(x)[0,ϕmax],x(L,L),ϕ(L,t)=r(t)[0,ϕmax],t(0,T),ϕ(L,t)=s(t)[0,ϕmax],F(t)˜f(s(t))t(0,T), (1.6)

    with a jump in V or equivalently, in the flux f, see [29,30], where F(t)˜f(s(t)) represents the non-standard boundary condition of the flux discontinuity, see [29].

    It is the purpose of the present contribution to introduce a numerical scheme for the MCLWR model with discontinuous flux (1.1)–(1.3) that is based on the available treatment [29] of the scalar model (1.6). The scalar version of the scheme slightly differs from that of [29] but we prove that it produces approximations that also converge to a weak solution. Numerical experiments provide evidence that it approximates the same solutions as the scheme of [29]. In the multiclass case we prove satisfaction of an invariant region principle, that is, numerical solutions assume values in

    D:={(ϕ1,,ϕN)TRN:ϕ10,,ϕN0,ϕ=ϕ1++ϕNϕmax}

    under corresponding assumptions on the initial and boundary data.

    The MCLWR model (1.1) has been studied intensively in recent years. The system (1.1), (1.2) has some interesting properties and in particular admits a separable entropy function for an arbitrary number of driver classes. We refer to [1,2,5,7,9,10,11,19,20,31,32,33,34,35,36,37] for numerical and analytical treatments and emphasize that to our knowledge, a velocity function discontinuous in the unknowns has not been considered so far for the MCLWR model.

    Conservation laws with discontinuous flux function arise in many physical applications including flow in porous media [22], sedimentation [8,18], and the LWR traffic model [26,30]. Here we limit the discussion to analyses where the flux is a discontinuous function of the unknown (as opposed to the more widely studied discontinuous dependence on spatial position). This property implies that standard numerical methods cannot be applied directly due to the presence of waves that travel at infinite speed, namely so-called zero waves. These waves carry information about the flux but this value is transported instantaneously, which excludes applying explicit schemes due to the lack of regularity of the flux. Gimse [21] was the first to present a solution to this problem. He studied a conservation law where the flux function has a single jump. He discussed the existence of the zero wave, generalized the method of convex hull construction, and solved the Riemann problem using a front tracking algorithm.

    Carrillo [12] studied conservation laws with a discontinuous flux function where the flux is allowed to have discontinuities on a finite subset of real numbers. The proof of existence of solutions is based on the comparison principle and an entropy inequality involving a version of semi-Kružkov entropies. Dias and Figueira [15] studied a related problem by using a mollification technique to smooth out discontinuities. They showed that solutions to a suitably regularized problem converge to solutions of the original problem in the limit. They also defined the notions of weak solution and weak entropy solution. The mollification technique was implemented in [16,17]. Moreover, Dias and Figueira [16] proposed a numerical scheme for Riemann problem. Specifically, they introduced a procedure to obtain a new Lipschitz continuous flux function with the same lower convex envelope of the original flux, and then a standard Lax-Friedrichs method is employed.

    Martin and Vovelle [27] considered the problem of numerical approximation in the Cauchy-Dirichlet problem for a scalar conservation law with a flux function having finitely many discontinuities. The well-posedness of this problem had been proved by Carrillo. An implicit finite volume scheme is constructed in [27] and Newton's method is employed to solve the resulting system of nonlinear equations. Furthermore, convergence to the unique entropy solution is shown.

    Lu et al. [26] explicitly constructed the entropy solutions for the LWR traffic flow model with a piecewise quadratic flow-density relationship. Their approach is based on constructions of entropy solutions to a sequence of approximate problems in which the flow-density relationship is continuous but tends to the discontinuous flux when a small parameter in this sequence tends to zero.

    Bulíˇcek et al. [3] introduced new concepts of entropy weak and measure-valued solutions that are consistent with the standard ones if the flux is continuous. They identified a given discontinuous flux function with a continuous curve that consists of the graph of this flux and abscissae that fill the jumps. Consequently, instead of a discontinuous flux function of the unknown, they deal with an implicit relation that represents a curve. One has one degree of freedom to set up the "optimal" unknown (independent variable). These ideas are combined in [4], where the authors treat the case of a flux function discontinuous in spatial position and the unknown. Through appropriate estimates for entropy measure-valued solutions well-posedness is shown.

    Wiens et al. [30] applied Dias and Figueira's mollification approach to solving a conservation law with a piecewise linear flux function in which there is a single discontinuity at a critical point. They introduced a mollified function and then the analytical solution to the corresponding Riemann problem is derived in the limit. Furthermore they constructed a Riemann solver that forms the basis for a high-resolution finite volume scheme of Godunov type and used an alternate approach that eliminates the severe CFL constraint by incorporating the effect of zero waves directly into the local Riemann solver.

    Towers [29] presented a finite difference scheme that implements a splitting consistent with the decomposition the flux f(u)=p(u)+g(u), where p is a Lipschitz continuous function and g is a function of Heaviside type that includes the jumps of f. The scheme has the form (see [29,Eq. (3.11)])

    {{Un+1/2j=˜G1(Unjλgn+1/2j+1),j=M,M1,,1,gn+1/2j=(Un+1/2jUnj+λgn+1/2j+1)/λ,j=M,M1,,1,Un+1j=Un+1/2jλΔ˜p(Un+1/2j+1Un+1/2j),j=1,,M, (1.7)

    which can be written in conservation form as follows:

    Un+1/2j=Unjλ(gn+1/2j+1gn+1/2j),Un+1=Un+1/2jλΔ˜p(Un+1/2j+1,Un+1/2j).

    The first part of the scheme is implicit and consistent with ut+g(u)x=0, but the resulting equations can be solved by evaluation of a piecewise linear function. Hence, an iterative solver like Newton's method is not required. The second part of the scheme is consistent with ut+p(u)x=0 and is explicit, and can be solved by any scheme suitable for a scalar conservation law with Lipschitz continuous flux. Towers [29] focused on the Godunov flux for the explicit part but also presented a simple flux-limited Lax-Wendroff-type modification to the Godunov scheme.

    The remainder of the paper is organized as follows. In Section 2 we present a numerical scheme for the LWR traffic flow model. We first introduce some assumptions and the notion of weak solution in Section 2.1. Next, Section 2.2 is devoted to the presentation of our scheme for the scalar case (N=1) and we imposed the appropriate CFL condition. Then, in Section 2.3, we prove that under the CFL condition it satisfies uniform L and TVD properties. Moreover we prove some kind time continuity estimates and to the end this section we prove convergence of our numerical solutions to a weak solution. In Section 3 we extend the algorithm to the multiclass case (N>1) and prove that the scheme preserves the invariant region D. In Section 4 we present several numerical examples to confirm all the results mentioned before. Section 5 collects some conclusions.

    Before describing the numerical scheme we introduce some assumptions and the definition of weak solutions proposed in [16], which is employed herein.

    To outline the basic idea, and to make the comparison with [29] transparent, we define the functions

    gV(ϕ):=αVH(ϕϕ),pV(ϕ):=V(ϕ)gV(ϕ), (2.1)

    where pV is a Lipschitz continuous, piecewise smooth and decreasing function, while gV is a non-negative and decreasing function, see Figure 1. Furthermore, as in [29], we can equivalently specify

    Figure 1.  (a) Piecewise continuous velocity function V(ϕ) with discontinuity at ϕ=ϕ, (b) continuous and discontinuous portions pV(ϕ) (solid line) and gV(ϕ) (dashed line).
    G(t)˜gV(s(t)), (2.2)

    where we recall that ˜gV denotes the multivalued version of gV. Moreover, we assume that the initial density function ϕ0 satisfies

    ϕ0(x)[0,ϕmax]for x(L,L),ϕ0BV([L,L]),gV(ϕ0)BV([L,L]).

    The boundary functions r and s are assumed to satisfy

    r(t),s(t)[0,ϕmax]for t[0,T],r,sBV([0,T]).

    We also assume that G(t)[0,αV] for all t[0,T], and GBV([0,T]).

    Definition 2.1 (Weak solution [16]) A function ϕL(ΠT) is said to be a weak solution to the initial-boundary value problem (1.6) if there exists a function qL(ΠT) satisfying q(x,t)˜f(ϕ(x,t)) a.e. such that for all test functions ψC10([L,L]×[0,T)),

    T0LL(ϕψt+qψx)dxdt+LLϕ0(x)ψ(x,0)dx=0.

    The domain ΠT is discretized as follows. We choose a partition {Ij}Mj=1 of [L,L] composed of uniform cells Ij=[xj1/2,xj+1/2), where xj+1/2=xj+Δx/2, that are centered in xj and have length |Ij|=Δx=2L/M. Then, for Δt>0, we let tn=nΔt for n=0,,N, where N is an integer such that T[tN,tN+Δt). The unknowns ϕnj approximate the cell average of the exact solution ϕ(,tn) in the cell Ij. The initial condition is discretized by

    ϕ0j=1ΔxIjϕ0(x)dx,j=1,,M,

    and the boundary conditions with F(t)˜f(s) are discretized as follows:

    ϕn+1/20=ϕn0=r(tn)=rn,ϕn+1/2M+1=ϕnM+1=s(tn)=sn,rn[0,ϕmax],sn[0,ϕmax],gn+1/2M+1[0,αV],gn+1/2M+1=gnM+1=G(sn)={αVif sn<ϕ,αVif sn=ϕ and traffic ahead of x=L is free-flowing,0if sn=ϕ and traffic ahead of x=L is congested,0if sn>ϕ. (2.3)

    Before proposing our scheme we recall that the basic idea of a splitting scheme consists in solving within each time step, first the PDE

    tϕ+x(vmaxϕgV(ϕ))=0, (2.4)

    followed by the solution of the conservation law with continuous flux

    tϕ+x(vmaxϕpV(ϕ))=0. (2.5)

    Note that in the scalar case the constant vmax is immaterial. For the remainder of the analysis of the scalar case we assume that t or x are rescaled so that vmax=1.

    Based on the form of the flux function of equations (2.4) and (2.5) and the properties of the functions gV and pV, we may write a numerical scheme for (1.6) that is motivated by Scheme 4 of [6] in the following form:

    ϕn+1/2j=ϕnjλ(ϕnjgn+1/2V,j+1ϕnj1gn+1/2V,j),ϕn+1j=ϕn+1/2jλ(ϕn+1/2jpV(ϕn+1/2j+1)ϕn+1/2j1pV(ϕn+1/2j)),j=1,,M. (2.6)

    The first half-step in (2.6) is semi-implicit and is consistent with (2.4) whereas the second half-step is explicit and consistent with (2.5). Scheme 4 of [6] exploits the density times velocity structure of the flux by calculating the numerical flux by evaluating density on the left cell and velocity (if non-negative) on the right cell adjacent to a cell interface. This idea goes back to a discrete traffic model proposed by Hilliges and Weidlich [23].

    In order to evaluate the first line in (2.6), we start by computing the values gn+1/2V,j from j=M+1 to j=1 (in decreasing order). This is motivated by the following argument, where we start from the semi-implicit equation

    ϕn+1/2j=ϕnjλ(ϕnjgV(ϕn+1/2j+1)ϕnj1gV(ϕn+1/2j)) (2.7)

    along with a known value G(ϕn+1/2M+1) arising from the boundary condition. Next, we write gV(ϕn+1/2j+1) as gn+1/2V,j+1 and then rearrange (2.7) as

    ϕn+1/2jλϕnj1gV(ϕn+1/2j)=ϕnjλϕnjgn+1/2V,j+1. (2.8)

    Let us now define the function

    GV(z;ϕ):=zλϕgV(z),z,ϕ[0,ϕmax]

    along with its multivalued version (with respect to z) ˜GV(;ϕ). Then ˜GV is strictly increasing and has a unique inverse z˜G1V(z;ϕ), see Figure 2. Explicitly, we get

    Figure 2.  (a) function z˜GV(z;ϕ) given by (2.9a) with λvmax=1/2, αV=0.3, and ϕ=0.8, (b) its inverse z˜G1V(z;ϕ) given by (2.9b).
    ˜GV(z;ϕ):={zλαVϕfor z[0,ϕ),[ϕλαVϕ,ϕ]for ϕ=ϕ,zfor z[ϕ,ϕmax], (2.9a)
    ˜G1V(z;ϕ):={z+λαVϕfor z[λαVϕ,ϕλαVϕ),ϕfor z[ϕλαVϕ,ϕ],zfor z[ϕ,ϕmax]. (2.9b)

    Consequently, we may express (2.8) as

    ˜GV(ϕn+1/2j;ϕnj1)=ϕnjλϕnjgn+1/2V,j+1,

    which allows us to obtain ϕn+1/2j by applying ˜G1V(z;ϕ) to both sides, that is

    ϕn+1/2j=˜G1V(ϕnjλϕnjgn+1/2V,j+1;ϕnj1). (2.10)

    Now that ϕn+1/2j is available, we solve for gn+1/2V,j the equation

    ϕn+1/2j=ϕnjλ(ϕnjgn+1/2V,j+1ϕnj1gn+1/2V,j), (2.11)

    provided that ϕnj1>0. If ϕnj1=0, we define directly

    gn+1/2V,j=gV(ϕn+1/2j).

    The numerical scheme can be summarized in Algorithm 2.1:

    Algorithm 2.1 (BCOV scheme, scalar case).

    Input: approximate solution vector {ϕnj}Mj=1 for t=tn

    gn+1/2V,M+1G(ϕn+1/2M+1) (using (2.3))

    do j=M,M1,,1

          ϕn+1/2j˜G1V(ϕnjλϕnjgn+1/2V,j+1;ϕnj1)

          if ϕnj10 then

              gn+1/2V,jϕn+1/2jϕnj+λgn+1/2V,j+1ϕnjλϕnj1

          else

              gn+1/2V,jgV(ϕn+1/2j)

          endif

    enddo

    do j=1,2,,M

          ϕn+1jϕn+1/2jλ(ϕn+1/2jpV(ϕn+1/2j+1)ϕn+1/2j1pV(ϕn+1/2j))

    enddo

    Output: approximate solution vector {ϕn+1j}Mj=1 for t=tn+1=tn+Δt

    Next, we demonstrate that the numerical scheme (2.11) is consistent with (2.4).

    Lemma 2.1. Assume that ϕn+1/2j[0,ϕmax] for all j. Then gn+1/2V,j˜gV(ϕn+1/2j) for all j. In particular gn+1/2V,j[0,αV] for all j.

    Proof. Let us first assume that ϕj1=0. Then the result follows from the definition of the function gV and the corresponding assignment to gn+1/2V,j in Algorithm 2.1. If ϕj10, then (2.10) and (2.9) imply that

    ϕn+1/2jλϕnj1gn+1/2V,j˜GV(ϕn+1/2j;ϕnj1).

    Therefore, by a straightforward case-by-case study (of the cases arising in (2.9)) we conclude that gn+1/2V,j˜gV(ϕn+1/2j).

    Now, to derive CFL conditions, we write the scheme (2.6) in incremental form

    ϕn+1/2j=ϕnj+Cn+1/2g,j+1/2Δ+ϕn+1/2jDn+1/2g,j1/2Δϕnj, (2.12a)
    ϕn+1j=ϕn+1/2j+Cn+1/2p,j+1/2Δ+ϕn+1/2jDn+1/2p,j1/2Δϕn+1/2j (2.12b)

    with the spatial difference operators Δ+Vj:=Vj+1Vj and ΔVj:=VjVj1 and the incremental coefficients

    Cn+1/2g,j+1/2:={λϕnjgV(ϕn+1/2j)gV(ϕn+1/2j+1)ϕn+1/2j+1ϕn+1/2jif ϕn+1/2j+1ϕn+1/2j,0otherwise,Dn+1/2g,j1/2:=λgV(ϕn+1/2j),Cn+1/2p,j+1/2:={λϕn+1/2jpV(ϕn+1/2j)pV(ϕn+1/2j+1)ϕn+1/2j+1ϕn+1/2jif ϕn+1/2j+1ϕn+1/2j,0otherwise,Dn+1/2p,j1/2:=λpV(ϕn+1/2j).

    To have an L estimate (Lemma 2.2 below) and the Total Variation Diminishing (TVD) property (Lemma 2.3 below) sufficient conditions are

    0Dn+1/2p,j1/2,Cn+1/2p,j+1/212,Cn+1/2g,j+1/20,0Dn+1/2g,j1/21for all j.

    First, we observe the following fact about ˜gV. If z1,z2[0,ϕmax] and z1z2, then

    gV,1˜gV(z1),gV,2˜gV(z2)gV,2gV,1z2z10. (2.13)

    This property and Lemma 2.1 imply that

    Dn+1/2g,j1/2,Cn+1/2g,j+1/20for all j.

    Next, the properties of the function pV ensure that

    Cn+1/2p,j+1/2,Dn+1/2p,j1/20for all j.

    Finally, to enforce the inequalities

    Dn+1/2p,j1/2,Cn+1/2p,j+1/212andDn+1/2g,j1/21for all j,

    we impose the CFL conditions

    λ(ϕmaxmax1jM|pV(ϕj)|+max1jMpV(ϕj))12,λαV1. (2.14)

    The goal is to prove convergence of approximate solution to a weak solution of (1.6). The discrete solutions {ϕn+1/2j} constructed via the scheme (2.6) are extended to the whole domain ΠT by defining the piecewise constant function

    ϕΔ(x,t)=Nn=0Mj=1χj(x)χn(t)ϕn+1/2j (2.15)

    where Δ=(Δx,Δt), and χj(x) and χn(t) are the characteristic functions of cell Ij and the time interval [tn,tn+Δt), respectively. The ratio λ=Δt/Δx is always kept constant, so the limits Δt0, Δx0, and Δ0 are equivalent.

    We start by proving an L estimate on ϕΔ. In the remainder of this section it is always assumed that the CFL condition (2.14) is in effect.

    Lemma 2.2. If ϕ0j[0,ϕmax] for j=1,,M, then

    ϕnj,ϕn+1/2j[0,ϕmax]forallj=1,,Mandn=1,,N. (2.16)

    Proof. Taking n=0 and j=M in (2.10) yields

    ϕ1/2M=˜G1V(ϕ0Mλϕ0Mg1/2V,M+1;ϕ0M1). (2.17)

    The boundary condition g1/2V,M+1=G(t0)[0,αV] together with the assumption implies that

    λαVϕ0Mϕ0Mλϕ0Mg1/2V,M+1ϕmax.

    Since ˜G1V(;ϕ) is a nondecreasing function and maps [λαVϕ,ϕmax] onto [0,ϕmax], (2.17) implies that ϕ1/2M[0,ϕmax]. It follows from (2.1) that g1/2V,M[0,αV]. Reasoning in this way for j=M1,M2,,1 yields ϕ1/2j[0,ϕmax] for j=1,,M. Since ϕ1/20,ϕ1/2M+1[0,ϕmax] by (2.3), and taking into account (2.12), we find that ϕ1j is a convex combination of ϕ1/2j1, ϕ1/2j and ϕ1/2j+1. Thus, ϕ1j[0,ϕmax] for j=1,,M. Repeating this argument inductively for n=1,,N we obtain (2.16).

    Lemma 2.3. The discrete approximate solutions generated by the scheme (2.12) satisfy the following spatial variation bounds:

    Mj=0|ϕnj+1ϕnj|TV(ϕ0)+TV(r)+TV(s),Mj=0|ϕn+1/2j+1ϕn+1/2j|TV(ϕ0)+TV(r)+TV(s). (2.18)

    Proof. Applying the operator Δ+ to (2.12a) and rearranging yields

    (1+Cn+1/2g,j+1/2)Δ+ϕn+1/2j=(1Dn+1/2g,j+1/2)Δ+ϕnj+Cn+1/2g,j+3/2Δ+ϕn+1/2j+1+Dn+1/2g,j1/2Δ+ϕnj1.

    Taking absolute values, summing over j=1,,M1 and using (2.14) we get

    M1j=1(1+Cn+1/2g,j+1/2)|Δ+ϕn+1/2j|M1j=1(1Dn+1/2g,j+1/2)|Δ+ϕnj|+M1j=1Cn+1/2g,j+3/2|Δ+ϕn+1/2j+1|+M1j=1Dn+1/2g,j1/2|Δ+ϕnj1|.

    Cancelling telescoping terms we obtain

    M1j=1|Δ+ϕn+1/2j|+Cn+1/2g,3/2|Δ+ϕn+1/21|M1j=1|Δ+ϕnj|Dn+1/2g,M1/2|Δ+ϕnM1|+Cn+1/2g,M+1/2|Δ+ϕn+1/2M|+Dn+1/2g,1/2|Δ+ϕn+1/20|. (2.19)

    The boundary condition implies

    Δ+ϕn+1/20=(1Dn+1/2g,1/2)Δ+ϕn0+Cn+1/2g,3/2Δ+ϕn+1/21,(1+Cn+1/2g,M+1/2)Δ+ϕn+1/2M=Δ+ϕnM+Dn+1/2g,M1/2Δ+ϕn+1/21.

    After taking absolute values in the two previous equations, we get

    |Δ+ϕn+1/20|(1Dn+1/2g,1/2)|Δ+ϕn0|+Cn+1/2g,3/2|Δ+ϕn+1/21|,(1+Cn+1/2g,M+1/2)|Δ+ϕn+1/2M||Δ+ϕnM|+Dn+1/2g,M1/2|Δ+ϕn+1/2M1|. (2.20)

    From (2.19) and (2.20)

    Mj=0|Δ+ϕn+1/2j|Mj=0|Δ+ϕnj|. (2.21)

    Reasoning in the same way as the proof of Lemma 5.2 in [29] we find that

    Mj=0|Δ+ϕn+1j|Mj=0|Δ+ϕnj|+|rn+1rn|+|sn+1sn|.

    It follows by induction that

    Mj=0|ϕnj+1ϕnj|TV(ϕ0)+TV(r)+TV(s). (2.22)

    From (2.21) and (2.22) we get (2.18).

    Now, we prove some time continuity estimates. The proof of the first of them is very similar to that of [29,Lemma 5.5], so we omit the details.

    Lemma 2.4. The following discrete L1 time continuity estimate holds for n0:

    M+1j=0|ϕn+1jϕn+1/2j|Ω1,Ω1:=TV(ϕ0)+TV(r)+TV(s)+2ϕmax.

    Lemma 2.5. The following estimate holds:

    Mj=1|ϕ1/2jϕ0j|Ω2,Ω2:=Mj=1|gV(ϕ0j+1)gV(ϕ0j)|+TV(ϕ0)+ϕmax. (2.23)

    Proof. We define g0V,j=gV(ϕ0j). The first equation in (2.6) with n=0 implies

    ϕ1/2jϕ0j=λϕ0j1(g1/2V,jg0V,j)λϕ0j(g1/2V,j+1g0V,j+1)λϕ0j(Δ+g0V,j)λg0V,j(Δ+ϕ0j1).

    Thus

    ϕ1/2jϕ0jλϕ0j1(g1/2V,jg0V,j)=λϕ0j(g1/2V,j+1g0V,j+1)λ(ϕ0jΔ+g0V,j+g0V,jΔ+ϕ0j1). (2.24)

    Taking absolute values in (2.24) and using (2.13) we find that

    |ϕ1/2jϕ0j|+λϕ0j1|g1/2V,jg0V,j|λϕ0j|g1/2V,j+1g0V,j+1|+λϕ0j|Δ+g0V,j|+λg0V,j|Δ+ϕ0j1|.

    Summing over j=1,,M and cancelling telescoping terms yields

    Mj=1|ϕ1/2jϕ0j|+λϕ00|g1/2V,1g0V,1|λϕ0M|g1/2V,M+1g0V,M+1|+λMj=1ϕ0j|Δ+g0V,j|+λMj=1g0V,j|Δ+ϕ0j1|. (2.25)

    Applying the boundary condition in (2.25), Lemmas 2.1, 2.2, and the CFL condition (2.14) we get (2.23).

    Lemma 2.6. There exists a constant Ω3 that is independent of Δ such that the following time continuity estimate holds:

    M+1j=0|ϕn+1/2jϕn1/2j|Ω3forn1. (2.26)

    Proof. For n2 and subtracting from the first half-step of (2.6) the corresponding formula for ϕn1/2j and rearranging terms we get

    ϕn+1/2jϕn1/2jλϕn1j1(gn+1/2V,jgn1/2V,j)=(1λgn+1/2V,j+1)(ϕnjϕn1j)+λgn+1/2V,j(ϕnj1ϕn1j1)λϕn1j(gn+1/2V,j+1gn1/2V,j+1).

    Taking absolute values and applying the CFL condition (2.14) yields

    |ϕn+1/2jϕn1/2jλϕn1j1(gn+1/2V,jgn1/2V,j)|(1λgn+1/2V,j+1)|ϕnjϕn1j|+λgn+1/2V,j|ϕnj1ϕn1j1|+|λϕn1j(gn+1/2V,j+1gn1/2V,j+1)|.

    From (2.13) we get

    |ϕn+1/2jϕn1/2j|+|λϕn1j1(gn+1/2V,jgn1/2V,j)|(1λgn+1/2V,j+1)|ϕnjϕn1j|+λgn+1/2V,j|ϕnj1ϕn1j1|+|λϕn1j(gn+1/2V,j+1gn1/2V,j+1)|. (2.27)

    Summing over j and cancelling telescoping terms we obtain

    Mj=1|ϕn+1/2jϕn1/2j|Mj=1|ϕnjϕn1j|+λgn+1/2V,1|ϕn0ϕn10|+λϕn1M|gn+1/2V,M+1gn1/2V,M+1|λgn+1/2V,M+1|ϕnMϕn1M|.

    The last inequality implies

    Mj=1|ϕn+1/2jϕn1/2j|Mj=1|ϕnjϕn1j|+|rnrn1|+|G(tn)G(tn1)|.

    We observe that

    ϕnjϕn1j=(1Bn1j+1/2An1j1/2)(ϕn1/2jϕn3/2j)+An1j+1/2(ϕn1/2j+1ϕn3/2j+1)+Bn1j1/2(ϕn1/2j1ϕn3/2j1), (2.28)

    where

    An1j+1/2=λ101φ(θϕn1/2j+1+(1θ)θϕn3/2j+1,θϕn1/2j+1+(1θ)θϕn3/2j+1)dθ,Bn1j+1/2=λ102φ(θϕn1/2j+1+(1θ)θϕn3/2j+1,θϕn1/2j+1+(1θ)θϕn3/2j+1)dθ.

    Herein φ(ϕj+1,ϕj)=ϕjpV(ϕj+1) and iφ denotes the partial derivative of φ with respect to the i-th argument (i=1,2). Since ϕ,pV(ϕ)0 and pV(ϕ)0, the function φ(ϕj+1,ϕj) is nonincreasing with respect to ϕj+1 and nondecreasing with respect to ϕj. This implies (together with the CFL condition)

    0An1j+1/2,Bn1j+1/212. (2.29)

    The remainder of the proof is similar to the proof of Lemma 5.6 in [29]. Details are omitted.

    Now, we are ready to prove the convergence of ϕΔ as Δ0.

    Lemma 2.7. The functions ϕΔ defined by (2.15) converge in L1(ΠT) and boundedly a.e. a along subsequence to a limit function ϕC([0,T],L1(L,L))L(ΠT)).

    Proof. The proof is a standard argument using the L estimate (Lemma 2.2), the uniform spatial variation bound (Lemma 2.3), and the L1 Lipschitz continuity in time estimate (Lemma 2.6).

    In order to show that the limit function ϕ identified in Lemma 2.7 is a weak solution in the sense of Definition 2.1, we must also prove the convergence of the flux approximations. Instead of showing that the approximations {gn+1/2V,j} converge we show that the approximations {hn+1/2j} converge, where we define

    hn+1/2j:=ϕn+1/2jgn+1/2V,jfor all j=1,,M and n=0,,N,

    and extend these quantities to functions defined on ΠT by

    hΔ(x,t):=Nn=0Mj=1χj(x)χn(t)hn+1/2j.

    Now, we require additional time continuity estimates, which is the contents of the following lemma. Its proof is very similar to that of Lemmas 5.8 and 5.9 in [29], and is therefore omitted.

    Lemma 2.8. The following uniform estimates hold for n1, where the constant Ω4 is independent of Δ:

    Mj=1|ϕn+1jϕnj|Ω4,Ω4:=Ω2+TV(s)+TV(r), (2.30)
    Mj=1|ϕn+1/2jϕnj|Ω1+Ω4. (2.31)

    The following lemma is needed to establish a spatial variation bound on the approximations hn+1/2j.

    Lemma 2.9. There exists a constant Ω5 that is independent of Δ such that

    Mj=1ϕnj|Δ+gn+1/2V,j|Ω5.

    Proof. From the first half-step of the scheme we get

    ϕn+1/2jϕnj=λ(ϕnjΔ+gn+1/2V,j+gn+1/2V,jΔ+ϕnj1),

    which can be rerranged as

    λϕnjΔ+gn+1/2V,j=(ϕn+1/2jϕnj)λgn+1/2V,jΔ+ϕnj1.

    Taking absolute values and summing over j=1,,M we get

    λMj=1ϕnj|Δ+gn+1/2V,j|Mj=1|ϕn+1/2jϕnj|+λMj=1gn+1/2V,j|Δ+ϕnj1|.

    From Lemma 2.1 and the CFL condition (2.14) we have

    Mj=1ϕnj|Δ+gn+1/2V,j|1λMj=1|ϕn+1/2jϕnj|+Mj=1|Δ+ϕnj1|.

    The result is obtained from (2.31) in Lemma 2.8 and Lemma 2.3.

    We are now ready to bound the spatial variation of the approximations hn+1/2j.

    Lemma 2.10. There exists a constant Ω6 that is independent of Δ such that

    Mj=1|hn+1/2j+1hn+1/2j|Ω6. (2.32)

    Proof. The first part of scheme (2.6) can be written as

    ϕn+1/2j=ϕnjλ(ϕnjgn+1/2V,j+1+gn+1/2V,j(ϕn+1/2jϕnj1))+λϕn+1/2jgn+1/2V,j.

    Applying the spatial difference operator to the above equation we get

    Δ+ϕn+1/2j=(1λgn+1/2V,j+1)Δ+ϕnj+λΔ+hn+1/2jλϕnj+1Δ+gn+1/2V,j+1λΔ+gn+1/2V,j(ϕn+1/2jϕnj)+λgn+1/2V,jΔ+ϕnj1λgn+1/2V,j+1Δ+ϕn+1/2j.

    Thus

    λΔ+hn+1/2j=Δ+ϕn+1/2j(1λgn+1/2V,j+1)Δ+ϕnj+λϕnj+1Δ+gn+1/2V,j+1+λΔ+gn+1/2V,j(ϕn+1/2jϕnj)+λgn+1/2V,jΔ+ϕnj1+λgn+1/2V,j+1Δ+ϕn+1/2j.

    After taking absolute values and using |Δ+gn+1/2V,j|αV, Lemma 2.1 and the CFL condition (2.14) we get

    λ|Δ+hn+1/2j|2|Δ+ϕn+1/2j|+|Δ+ϕnj|+λϕnj+1|Δ+gn+1/2V,j+1|+|ϕn+1/2jϕnj|+|Δ+ϕnj1|.

    Summing over j=1,,M we get

    Mj=1|Δ+hn+1/2j|2λMj=1|Δ+ϕn+1/2j|+2λMj=1|Δ+ϕnj|+1λMj=1|ϕn+1/2jϕnj|+Mj=1ϕnj+1|Δ+gn+1/2V,j+1|.

    Finally, the result follows from Lemma 2.3, (2.31) in Lemma 2.8, and Lemma 2.9.

    The following lemma is required to prove the L1 Lipschitz continuity in time and spatial variation bounds on {hn+1/2j}.

    Lemma 2.11. There exists a constant Ω7 that is independent of Δ such that

    ΔxMj=1Nn=1ϕn1j|gn+1/2V,j+1gn1/2V,j+1|Ω7. (2.33)

    Proof. From (2.27) we get

    λϕn1j|gn+1/2V,j+1gn1/2V,j+1|(1λgn+1/2V,j+2)|ϕnj+1ϕn1j+1|+λgn+1/2V,j+1|ϕnjϕn1j||ϕn+1/2j+1ϕn1/2j+1|+λϕn1j+1|gn+1/2V,j+2gn1/2V,j+2|.

    By induction we obtain

    λϕn1j|gn+1/2V,j+1gn1/2V,j+1|Mk=j+1|ϕnkϕn1k|+|gn+1/2V,M+1gn1/2V,M+1|Mk=j+1|ϕn+1/2kϕn1/2k|+|ϕnjϕn1j|. (2.34)

    Recalling (2.28) we have

    Mk=j+1|ϕnkϕn1k|Mk=j+1(1Bn1k+1/2An1k1/2)|ϕn1/2kϕn3/2k|+Mk=j+1An1k+1/2|ϕn1/2k+1ϕn3/2k+1|+Mk=j+1Bn1k1/2|ϕn1/2k1ϕn3/2k1|.

    Cancelling telescoping terms and applying (2.29) yields

    Mk=j+1|ϕnkϕn1k|Mk=j+1|ϕn1/2kϕn3/2k|+12|ϕn1/2M+1ϕn3/2M+1|+12|ϕn1/2jϕn3/2j|.

    Then (2.34) becomes

    λϕn1j|gn+1/2V,j+1gn1/2V,j+1|Mk=j+1|ϕn1/2kϕn3/2k|Mk=j+1|ϕn+1/2kϕn1/2k|+12|ϕn1/2M+1ϕn3/2M+1|+12|ϕn1/2jϕn3/2j|+|ϕnjϕn1j|+|gn+1/2V,M+1gn1/2V,M+1|.

    Summing over n2 and j=1,,M, cancelling telescoping terms and multiplying the result by Δx we get

    ΔxMj=1Nn=2ϕn1j|gn+1/2V,j+1gn1/2V,j+1|S1++S5,

    where we define

    S1:=2LλMj=1|ϕ3/2jϕ1/2j|,S2:=LλNn=2|sn1/2sn3/2|,S3:=ΔxλMj=1Nn=2|ϕnjϕn1j|,S4:=2LλNn=2|gn+1/2V,M+1gn1/2V,M+1|,S5:=Δx2λMj=1Nn=2|ϕn1/2jϕn3/2j|.

    In view of the bounds established so far, there holds

    S12LλΩ3,S2LλTV(s),S3Ω4T,S42LλTV(G),S5Ω32T.

    These bounds in conjunction with |g3/2V,jg1/2V,j|αV imply that there exists a constant Ω7 such that (2.33) is valid.

    Lemma 2.12. There exists a constant Ω8 that is independent of Δ such that

    ΔtNn=0Mj=1|hn+1/2j+1hn+1/2j|+ΔxNn=1Mj=1|hn+1/2jhn1/2j|Ω8. (2.35)

    Proof. In light of the spatial variation bound (2.32) we find that

    ΔtNn=0Mj=1|hn+1/2j+1hn+1/2j|Ω6T.

    The first part of (2.6) implies

    ϕn+1/2jϕn1/2j=(1λgn+1/2V,j+1)(ϕnjϕn1j)+λ(hn+1/2jhn1/2j)λgn+1/2V,j(ϕn+1/2jϕnj1)+λgn1/2V,j(ϕn1/2jϕn1j1)λϕn1j(gn+1/2V,j+1gn1/2V,j+1)=(1λgn+1/2V,j+1)(ϕnjϕn1j)+λ(hn+1/2jhn1/2j)λgn+1/2V,j(ϕn+1/2jϕnj)λgn+1/2V,jΔ+ϕnj1+λgn1/2V,jΔ+ϕn1j1+λgn1/2V,j(ϕn1/2jϕn1j)λϕn1j(gn+1/2V,j+1gn1/2V,j+1).

    Consequently,

    λ(hn+1/2jhn1/2j)=(ϕn+1/2jϕn1/2j)(1λgn+1/2V,j+1)(ϕnjϕn1j)+λgn+1/2V,j(ϕn+1/2jϕnj)+λgn+1/2V,jΔ+ϕnj1λgn1/2V,jΔ+ϕn1j1λgn1/2V,j(ϕn1/2jϕn1j)+λϕn1j(gn+1/2V,j+1gn1/2V,j+1).

    Taking absolute values and using the CFL condition (2.14) we get

    λ|hn+1/2jhn1/2j||ϕn+1/2jϕn1/2j|+|ϕnjϕn1j|+|Δ+ϕnj1|+|Δ+ϕn1j1|+|ϕn+1/2jϕnj|+|ϕn1/2jϕn1j|+λϕn1j|gn+1/2V,j+1gn1/2V,j+1|.

    Multiplying this inequality by Δx and summing over j and n we get

    ΔxNn=1Mj=1|hn+1/2jhn1/2j|U1++U5,

    where we define

    U1:=ΔxλNn=1Mj=1|ϕn+1/2jϕn1/2j|,U2:=ΔxλNn=1Mj=1|ϕnjϕn1j|,U3:=2ΔxλNn=1Mj=1|Δ+ϕnj1|,U4:=2ΔxλNn=1Mj=1|ϕn+1/2jϕnj|,U5:=ΔxNn=1Mj=1ϕn1j|gn+1/2V,j+1gn1/2V,j+1|.

    From (2.18), (2.26), (2.30), (2.31) and (2.33) we have

    U1Ω3T,U2Ω4T,U32(TV(ϕ0)+TV(s)+TV(r))T,U42(Ω1+Ω4)T,U5Ω7.

    Combining these bounds we see that there exists a constant Ω8 that is independent of Δ such that (2.35) is valid.

    In the following lemma we state and prove convergence of the functions hΔ as Δ0. To this end, we define Q(ϕ):=ϕgV(ϕ) and denote by ˜Q the multivalued version of Q.

    Lemma 2.13. The functions hΔ converge in L1(ΠT) and boundedly a.e. along subsequence to some limit function wL1(ΠT)L(ΠT). Moreover, by a suitable choice of a subsequence, we have w(x,t)˜Q(ϕ(x,t)) a.e. in ΠT, where ϕ(x,t) is the limit of Lemma 2.7.

    Proof. We observe that |hn+1/2j|ϕmaxαV. Then by Helly's theorem [24] there exists a function wL1(ΠT) such that hΔw along a subsequence in L1(ΠT) and boundedly a.e. in ΠT. To prove the second assertion, we assume (by extracting further subsequences if necessary) that ϕΔϕ,hΔw in L1(ΠT) and fix a point (x,t)ΠT where ϕΔ(x,t)ϕ(x,t) and hΔ(x,t)w(x,t) as Δ0. First, we consider the case ϕ(x,t)=ϕ. Lemma 2.1 implies that 0hΔ(x,t)αVϕΔ(x,t). Then passing to the limit in the above inequality we obtain

    w(x,t)[0,αVϕ]=˜Q(ϕ).

    In case ϕ(x,t)ϕ first we consider ϕ(x,t)<ϕ, then ˜Q(ϕ(x,t))=αVϕ(x,t). For sufficiently small Δ the inequality ϕΔ(x,t)<ϕ implies that ϕn+1/2j<ϕ and ˜gV(ϕn+1/2j)={αV}. Then, by Lemma 2.1 we get

    hΔ(x,t)=Nn=0Mj=1χj(x)χn(t)hn+1/2j=αVNn=0Mj=1χj(x)χn(t)ϕn+1/2j=αVϕΔ(x,t).

    Thus w(x,t)=limhΔ(x,t)=αVlimϕΔ(x,t)=αVϕ(x,t)=˜Q(ϕ(x,t)).

    In the case ϕ(x,t)>ϕ there holds ˜Q(ϕ(x,t))=0. In this case it is necessary extend {gn+1/2V,j} to functions defined on ΠT by

    gΔV(x,t)=Nn=0Mj=1χj(x)χn(t)gn+1/2V,j,

    and we need to utilize the following consequence of Lemma 2.1:

    gΔV(x,t)˜gV(ϕΔ(x,t)),(x,t)ΠT.

    For sufficiently small Δ, ϕΔ(x,t)>ϕ implies that ˜gV(ϕΔ(x,t))={0}, hence gΔV(x,t)=0. Finally observe that 0hΔ(x,t)ϕΔ(x,t)gΔV(x,t)=0 for sufficiently small Δ. Hence w(x,t)=˜Q(ϕ(x,t))=0.

    Theorem 2.14 (Main result). The functions ϕΔ converge in L1(ΠT) and boundedly a.e. along subsequence to some ϕC([0,T],L1(L,L))L(ΠT). The limit function ϕ(x,t) is a weak solution in sense of Definition 2.1.

    Proof. The convergence is ensured by Lemma 2.7. It remains to prove that the limit ϕ is a weak solution. Let us fix a point (x,t)ΠT, then Lemma 2.13 implies that w(x,t)˜Q(ϕ(x,t)) a.e. in ΠT. If ϕ(x,t)ϕ, then ˜Q(ϕ(x,t))=Q(ϕ(x,t)). Thus w(x,t)=Q(ϕ(x,t)), then we define q(x,t)=ϕpV(ϕ)+Q(ϕ(x,t))=f(ϕ(x,t)). In the case where ϕ(x,t)=ϕ we take w(x,t)[0,αVϕ] and define

    q(x,t)=ϕpV(ϕ)+w(x,t)[ϕpV(ϕ),ϕpV(ϕ)+αVϕ]=˜f(ϕ).

    In either case q(x,t)˜f(ϕ(x,t)).

    We note that the two steps of (2.6) imply

    ϕn+1jϕnj+λ(ϕnjgn+1/2V,j+1ϕnj1gn+1/2V,j+ϕn+1/2jpn+1/2V,j+1ϕn+1/2j1pn+1/2V,j)=0. (2.36)

    We now choose a test function ψC10((L,L)×[0,T)) and define ψnj:=ψ(xj,tn). Multiplying (2.36) by Δxψnj and summing the result over j and n yields

    ΔxΔtNn=0Mj=1ϕn+1jϕnjΔtψnj+ΔxΔtNn=0Mj=1ϕnjgn+1/2V,j+1ϕnj1gn+1/2V,jΔxψnj+ΔxΔtNn=0Mj=1ϕn+1/2jpn+1/2V,j+1ϕn+1/2j1pn+1/2V,jΔxψnj=0.

    A summation by parts yields

    ΔxΔtNn=0Mj=1ϕn+1jψn+1jψnjΔt+ΔxMj=1ϕ0jψ0j+ΔxΔtNn=0Mj=1ψnj+1ψnjΔxϕnjgn+1/2V,j+1+ΔxΔtNn=0Mj=1ψnj+1ψnjΔxϕn+1/2jpn+1/2V,j+1=0. (2.37)

    An application of (2.12) yields, as Δx,Δt0,

    ΔxΔtNn=0Mj=1ϕn+1jψn+1jψnjΔt=ΔxΔtNn=0Mj=1ϕn+1/2jψn+1jψnjΔt+O(Δx).

    This equation and Lemma 2.3 imply that the two first sums in (2.37) converge to

    T0LLϕψtdxdt+LLϕ0(x)ψ(x,0)dxdt.

    Concerning the last term in (2.37), we get

    ΔxΔtNn=0Mj=1ψnj+1ψnjΔxϕn+1/2jpn+1/2V,j+1=ΔxΔtNn=0Mj=1ψnj+1ψnjΔxϕn+1/2jpn+1/2V,j+ΔxΔtNn=0Mj=1ψnj+1ψnjΔx(ϕn+1/2jpn+1/2V,j+1ϕn+1/2jpn+1/2V,j).

    By properties of the function pV we get the estimate

    ϕn+1/2jpn+1/2V,j+1ϕn+1/2jpn+1/2V,j=ϕn+1/2j(pn+1/2V,j+1pn+1/2V,j)ϕmaxpVΔx.

    Thus

    |ΔxΔtNn=0Mj=1(ψnj+1ψnj)Δx(ϕn+1/2jpn+1/2V,j+1ϕn+1/2jpn+1/2V,j)|2MTϕmaxΔxψtpV,

    which tends to zero as Δx0. Therefore the last term in (2.37) converges to

    T0LLϕpV(ϕ)ψxdxdt

    as Δx0. The second term in (2.37) can be written as

    ΔxΔtNn=0Mj=1ψnj+1ψnjΔxϕnjgn+1/2V,j+1=ΔxΔtNn=0Mj=1ψnj+1ψnjΔxϕn+1/2jgn+1/2V,j+ΔxΔtNn=0Mj=1ψnj+1ψnjΔxϕnjΔ+gn+1/2V,j+ΔxΔtNn=0Mj=1ψnj+1ψnjΔxgn+1/2V,j(ϕnjϕn+1/2j).

    Using Lemmas 2.9, 2.1, and 2.8 we get

    |ΔxΔtNn=0Mj=1ψnj+1ψnjΔxϕnjΔ+gn+1/2V,j|ΔxψxΩ5T,|ΔxΔtNn=0Mj=1ψnj+1ψnjΔxgn+1/2V,j(ϕnjϕn+1/2j)|αVΔxψx(Ω1+Ω4)T.

    Consequently, as Δx0,

    ΔxΔtNn=0Mj=1ψnj+1ψnjΔxϕnjΔ+gn+1/2V,j0,ΔxΔtNn=0Mj=1ψnj+1ψnjΔxgn+1/2V,j(ϕnjϕn+1/2j)0.

    Then substituting gΔV(xj,tn)=gn+1/2V,j and applying the dominated convergence theorem we obtain that the second term in (2.37) converges to

    T0LLwψxdxdt.

    Collecting the previous results we get

    T0LL(ϕψt+qψx)dxdt+LLϕ0(x)ψ(x,0)dx=0,

    so ϕ is a weak solution in sense of Definition 2.1.

    Algorithm 2.1 cannot be applied directly in a component-wise manner for each class i in the multiclass case (1.1)–(1.4), but we can first solve for the total density ϕ and then update the densities ϕ1,,ϕN for each class. The multiclass version of the scalar scheme (1.6) can be written as

    ϕn+1/2i,j=ϕni,jλvmaxi(ϕni,jgV(ϕn+1/2j+1)ϕni,j1gV(ϕn+1/2j)),ϕn+1i,j=ϕn+1/2i,jλvmaxi(ϕn+1/2i,jpV(ϕn+1/2j+1)ϕn+1/2i,j1pV(ϕn+1/2j)),i=1,,N,j=1,,M, (3.1)

    where the following quantity is an approximate value of the total density ϕ:

    ϕn+1/2j:=ϕn+1/21,j++ϕn+1/2N,j.

    In order to solve (3.1), we need to impose the non-standard boundary condition (1.4b). Recalling that V(ϕ)=gV(ϕ)+pV(ϕ) we can equivalently specify for the multiclass case the condition (2.2). The correspondence when s(t)=ϕ is

    F(t)=(vmax)Ts(t)V(ϕ)G(t)=αV,F(t)=(vmax)Ts(t)V(ϕ+)G(t)=0. (3.2)

    Coming back to (3.1), we define Φnj:=(ϕn1,j,,ϕnN,j)T. Summing over i, assuming that gV is evaluated at the new time step, and replacing gV(ϕn+1/2j+1) by gn+1/2V,j+1 yields

    ϕn+1/2j=ϕnjλ(vmax)T(gn+1/2V,j+1ΦnjgV(ϕn+1/2j)Φnj1). (3.3)

    This can be rearranged as

    ϕn+1/2jλ(vmax)TΦnj1gV(ϕn+1/2j)=ϕnjλ(vmax)TΦnjgn+1/2V,j+1. (3.4)

    Let us now define the function GV(z;Φ):=zλ(vmax)TΦgV(z) and denote by ˜GV(;Φ) its multivalued version (with respect to z). Then ˜G is strictly increasing and has a unique inverse z˜G1V(z;Φ). Expressing (3.4) as

    ˜GV(ϕn+1/2j;Φnj1)=ϕnjλ(vmax)TΦnjgn+1/2V,j+1 (3.5)

    which allows us to obtain ϕn+1/2j by applying ˜G1V(;Φnj1) to both sides, that is

    ϕn+1/2j=˜G1V(ϕnjλ(vmax)TΦnjgn+1/2V,j+1;Φnj1).

    Now that ϕn+1/2j is available, we solve for gn+1/2V,j the equation

    ϕn+1/2j=ϕnjλ(vmax)T(gn+1/2V,j+1Φnjgn+1/2V,jΦnj1)

    (cf. (3.3)). This yields

    gn+1/2V,j=ϕn+1/2jϕnj+λgn+1/2V,j+1(vmax)TΦnjλ(vmax)TΦnj1,

    provided that Φnj10. If Φnj1=0 then we set gn+1/2V,j=gV(ϕn+1/2j). The numerical scheme for the multiclass model can be summarized in the following algorithm.

    Algorithm 3.1 (BCOV scheme, multiclass case).

    Input: approximate solution vector {ϕni,j}Mj=1, i=1,,N for t=tn

    gn+1/2V,M+1G(ϕn+1/2M+1) (using (2.3) and (3.2))

    do j=M,M1,,1

          ϕn+1/2j˜G1V(ϕnjλgn+1/2V,j+1(vmax)TΦnj;Φnj1)

          if Φnj10 then

              gn+1/2V,jϕn+1/2jϕnj+λgn+1/2V,j+1(vmax)TΦnjλ(vmax)TΦnj1

          else

              gn+1/2V,jgV(ϕn+1/2j)

          endif

    enddo

    do j=1,,M

          do i=1,,N

              ϕn+1/2i,jϕni,jλvmaxi(ϕni,jgn+1/2V,j+1ϕni,j1gn+1/2V,j)

          enddo

    enddo

    do j=1,,M

          do i=1,,N

    ϕn+1i,jϕn+1/2i,jλvmaxi(ϕn+1/2i,jpV(ϕn+1/2j+1)ϕn+1/2i,j1pV(ϕn+1/2j))

          enddo

    enddo

    Output: approximate solution vectors {ϕn+1i,j}Mj=1, i=1,,N for t=tn+1=tn+Δt

    Remark 3.1 We recall that gn+1/2V,M+1=G(ϕn+1/2M+1), the boundary condition that appears in Algorithm 3.1, is defined using (2.3) for the total density ϕn+1/2M+1. We illustrate this boundary condition in Section 4.5.

    The problem of interest to us is to show that D is an invariant region of the scheme. To this end we first consider the evolution of the total density ϕ. Summing over i=1,,N the second equation in (3.1) yields

    ϕn+1j=ϕn+1/2jλ(vmax)T(pV(ϕn+1/2j+1)n+1/2Φn+1/2jpV(ϕn+1/2j)Φnj1).

    The above equation can be written in incremental form as

    ϕn+1j=ϕn+1/2j+Cn+1/2j+1/2Δ+ϕn+1/2jDn+1/2j1/2Δϕn+1/2j, (3.6)

    where we define

    Cn+1/2j+1/2:={λ(vmax)TΦn+1/2jpV(ϕn+1/2j)pV(ϕn+1/2j+1)ϕn+1/2j+1ϕn+1/2jif ϕn+1/2j+1ϕn+1/2j,0if ϕn+1/2j+1=ϕn+1/2j,Dn+1/2j1/2:={λpV(ϕn+1/2j)(vmax)T(Φn+1/2jΦn+1/2j1)ϕn+1/2jϕn+1/2j1if ϕn+1/2jϕn+1/2j1,0if ϕn+1/2j=ϕn+1/2j1.

    Since pV(ϕ) is a non-increasing positive function we have Cn+1/2j+1/2,Dn+1/2j1/20. To ensure that |Cn+1/2j+1/2|1/2 and |Dn+1/2j1/2|1/2 we impose the CFL condition

    λϕmaxmax1jM|pV(ϕnj)|max1iNvmaxi12,λmax1jMp(ϕnj)max1iNvmaxi12. (3.7)

    Lemma 3.1. Assume that Φ0jD for j=1,,M. Then Φnj,Φn+1/2jD for j=1,,M.

    Proof. We claim that

    Φn+1/2jDfor all j=1,,Mgn+1/2V,j[0,αV]for all j=1,,M. (3.8)

    In the case Φnj1=0 the result follows from the definition of the function gV and (2.1). Suppose that Φnj10, summing over i=1,,N the first equation in (3.1) yields

    ϕn+1/2j=˜G1V(ϕnjλ(vmax)TΦnjgn+1/2V,j+1;Φnj1).

    Using (3.5) and (3.3) we find that

    ϕn+1/2jλ(vmax)TΦnj1gn+1/2V,j˜GV(ϕn+1/2j;Φnj1).

    Thus, a straightforward case-by-case study and (3.5) prove that (3.8) is valid. The remainder of the proof is similar to the proof of Lemma 2.2.

    We now present some numerical simulations to illustrate the behaviour of solutions to system (1.1) by using Algorithms 2.1 and 3.1 for the scalar and multiclass case, respectively. In the scalar case, we compare numerical approximations with those generated by the scheme (1.7) proposed by Towers in [29]. In all numerical examples for both the scalar (N=1) and system (N2) cases we use the discontinuous velocity function

    V(ϕ)={1ϕ/ϕmaxfor 0ϕϕ,wf(1ϕmax/ϕ)for ϕ<ϕϕmax,

    where ϕ=0.5, wf=0.2, ϕmax=1, and αV=0.3.

    In all numerical experiments computations are performed on a finite interval [1,1] that is subdivided into M subintervals of length Δx=2/M, and the time step is computed by Δt=Δx/2 in the scalar case (N=1) and Δt=Δx/(2max{vmax1,,vmaxN}) in the multiclass case (N2). These choices ensure that the respective CFL conditions (2.14) and (3.7) are satisfied.

    We consider the Riemann problem for the scalar equation tϕ+x(ϕV(ϕ))=0 with initial data

    ϕ0(x)={ϕLfor x<0.2,ϕRfor x0.2  (4.1)

    (no boundary conditions are involved). For ϕL=0.3 and ϕR=0.9, the solution consists of two shock waves with negative velocities of propagation, namely a shock wave connecting ϕL with ϕ that travels at velocity σ1=0.55 and another shock wave connecting ϕ with ϕR with velocity σ2=0.2. Figure 3(a) shows the numerical approximations to the solution of this problem computed with M=800 for both schemes at simulated time T=1.8.

    Figure 3.  Example 1: numerical solution with M=800 and comparison with the exact solution of the Riemann problem (a) with ϕL=0.3 and ϕR=0.9 at simulated time T=1.8, (b) with ϕL=0.9 and ϕR=0.3 at simulated time T=1.5. Here and in Figures 4 and 5 we label with 'Towers scheme' the scheme (1.7) proposed in [29] and by 'BCOV scheme' the scheme of Algorithm 2.1 advanced in the present work.

    For ϕL=0.9 and ϕR=0.3, the solution consists of a shock wave connecting ϕL with ϕ that travels at velocity σ1=0.575 and a rarefaction wave connecting ϕ with ϕR. In Figure 3(b) we display the numerical solutions computed with M=800 for both schemes at simulated time T=1.5. In both scenarios, all waves are approximated correctly by both schemes.

    In this example we compare numerical approximations for equation (1.1) obtained by both schemes (Towers scheme (1.7) and Algorithm 2.1), starting from the initial function ϕ0(x)=exp((x+0.2)2/(0.04)) for x[1,1]. Numerical approximations are computed at simulated times T=0.1 and T=0.3 with discretizations M=100×2l, l=0,1,,4. Table 1 displays the corresponding approximate L1 errors obtained by utilizing a reference solution computed by the Towers scheme with Mref=12800. We observe that the approximate L1 errors decrease as the grid is refined. In Figure 4 we display the numerical approximations for M=100 and compare them with the reference solution.

    Table 1.  Example 2: approximate L1 errors eM(u) with Δx=2/M.
    T=0.1 T=0.3
    Towers BCOV Towers BCOV
    M eM(ϕΔ) eM(ϕΔ) eM(ϕΔ) eM(ϕΔ)
    100 1.32e-2 1.76e-2 1.63e-2 2.39e-2
    200 6.55e-3 9.22e-3 8.59e-3 1.31e-2
    400 3.29e-3 4.46e-3 4.25e-3 6.46e-3
    800 1.72e-3 2.403-3 2.12e-3 3.31e-3
    1600 8.00e-4 1.18e-3 9.29e-4 1.563-3

     | Show Table
    DownLoad: CSV
    Figure 4.  Example 2: numerical solutions for M=100 at simulated times (a) T=0.1, (b) T=0.3.

    This example comes from [29,Example 6.2] and is designed to illustrate that when s(tn)=ϕ, the solutions depend on the boundary condition F(t)˜f(ϕ). For this example we consider the Riemann problem with initial data (4.1) with ϕL=1/4 and ϕR=1/2. We compute the solution twice, once using G(t)=αV (equivalently, F(t)=1/2), and the second time using G(t)=0 (equivalently, F(t)=1/4). As shown in Figure 5, in the first case the solution corresponds to a shock wave connecting ϕL with ϕR with speed of propagation σ=1, and in the second case the solution corresponds to a stationary shock (σ=0) connecting ϕL with ϕR.

    Figure 5.  Example 3: numerical solutions depending on the boundary conditions F(t)˜f(ϕ) with M=1600 at simulated time T=0.5, with (a) F(t)˜f(ϕ) (free flow), (b) F(t)˜f(ϕ+) (congested flow).

    To illustrate the invariant region property of the proposed scheme (Lemma 3.1), we consider the case N=3 and the Riemann initial data

    ϕ0(x)={(0.1,0.1,0.1)Tfor x<0.5,(0.4,0.5,0.1)Tfor x0.5, 

    with velocities vmax=(1,3,10)T. The solution consists of a stationary shock plus two shock waves that travel with negative velocities. The numerical simulation at three simulated times is displayed in Figure 6. The profile for each class and the total density are displayed in this figure. Furthermore we can see that the profile of the total density in Figure 6 looks like the profile of Figure 3(a).

    Figure 6.  Example 4: density profiles simulated with M=1600 at (a) T=0.2, (b) T=0.4, (c) T=0.6.

    It is the purpose of this example to illustrate the boundary condition

    gn+1/2V,M+1=G(ϕn+1/2M+1),

    where G() is specified in (2.3), that appears within Algorithm 3.1. To this end consider N=3 and the velocities and Riemann initial data

    vmax=(1,3,6)T,Φ(x,0)=Φ0(x)={ΦL=(0.05,0.08,0.12)Tfor x<0,ΦR=(0.14,0.16,0.2)Tfor x0. 

    Observe that ϕR=ϕ=s(t), where ϕR is the total density of the right state ΦR.

    As in Example 3 we show that the solution depends on the boundary condition F(t)(vmax)Ts(t)˜V(s(t)). We start with the initial condition shown in Figure 7(a) and compute the solution twice, once using G(t)=αV, and the second time using G(t)=0. In Figures 7(b) and (c) we display the profile for each class and total density for the first case G(t)=αV at two different simulated times. We can see that in this case a free-flow regime is produced, which is verified in Figure 8(a). In Figure 9 we display the profiles for each class and total density for the second case G(t)=0 at two different simulated times. In contrast to the previous cases, a congested flow regime is produced, as is illustrated in Figure 8(b).

    Figure 7.  Example 5: numerical solution for a free-flow regime (G(t)=αV): (a) initial condition, (b, c) density profiles with M=1600 at simulated times (b) T=0.1, (c) T=0.2.
    Figure 8.  Example 5: simulated total density computed with BCOV scheme with N=3 and M=1600: (a) free flow (G(t)=αV), (b) congested flow (G(t)=0).
    Figure 9.  Example 5: numerical solution for a congested flow regime (G(t)=0): density profiles with M=1600 at simulated times (a) T=0.1, (b) T=0.2. The initial condition is the same as in Figure 7(a).

    In this example we consider N=5, the velocities vmax=(1,2,3,4,5)T, and the initial condition

    Φ(x,0)=Φ0(x)=(0.15,0.2,0.3,0.2,0.15)Tψ(x),ψ(x)=exp(50(x+2)2/3).

    We display in Figure 10 numerical approximation computed with M=1600 at simulation times T=0.02 and T=0.12. We observe the dynamics of each individual densities ϕi and the total density ϕ, which exhibits a shock wave due to the discontinuity in the flux. This behaviour is similar to that presented in Figure 4. In Figure 11 we display the evolution of ϕΔ(,t) for t[0,0.12], and we compare the solution with the approximation of the continuous problem (where αV=0). For the discontinuous case the shock is more clearly observed than in the continuous case. In Figures 12 and 13 we compare the numerical approximation computed with M=100, with a reference solution at simulated times T=0.02 and T=0.12. In Table 2 we compute the approximate L1 error based on a reference solution obtained by the BCOV scheme with Mref=12800. We observe that the approximate L1 errors decrease as the grid is refined.

    Figure 10.  Example 6: numerical solutions obtained with BCOV scheme with N=5 and M=1600 at simulated times (a) T=0.02, (b) T=0.12.
    Figure 11.  Example 6: simulated total density obtained with BCOV scheme with N=5 and M=1600: (a) discontinuous problem, (b) continuous problem.
    Figure 12.  Example 6: comparison of reference solution (Mref=12800) with approximate solutions computed by BCOV scheme with M=100 at simulated time T=0.02.
    Figure 13.  Example 6: comparison of reference solution (Mref=12800) with approximate solutions computed by BCOV scheme with M=100 at simulated time T=0.02.
    Table 2.  Example 6: approximate L1 errors eM(u) with Δx=2/M.
    T=0.02 T=0.12
    M eM(ϕΔ) eM(ϕΔ)
    100 1.39e-2 3.87e-2
    200 7.90e-3 2.47e-2
    400 4.20e-3 1.55e-2
    800 2.00e-3 9.20e-3
    1600 1.00e-3 5.10e-3

     | Show Table
    DownLoad: CSV

    In this example we consider N=5, the velocity vector vmax=(1,1.5,2,6,7)T, and the initial condition

    Φ(x,0)=Φ0(x)=(0.17,0.17,0.16,0,0)Tψ1(x)+(0,0,0,0.245,0.245)Tψ2(x),

    where we define

    ψ1(x):=exp(10(x2)2),ψ2(x):=exp(50(x1)2/4)

    for x[0,5]. We compute numerical approximation at simulated times T=0.1, T=0.2 and T=0.3 with different discretizations by using M=100×2l and l=0,1,,4. In Table 3 we compute the L1 error comparing with respect to a reference solution computed by the BCOV scheme with Mref=12800. We observe that the approximate L1 errors decrease as the grid is refined. Figure 14 shows results for M=Mref=12800. The numerical results of Figure 14 indicate that jumps in the total density ϕ only occur from smaller to higher values in an increasing x-direction. This phenomenon occurs because the speeds of the last two classes are greater than the first three. Furthermore, in Figure 15 we show the simulated total density computed by the BCOV scheme with N=5 and M=1600.

    Table 3.  Example 7: Approximate L1 errors eM(u) with Δx=5/M.
    T=0.1 T=0.2 T=0.3
    M eM(ϕΔ) eM(ϕΔ) eM(ϕΔ)
    100 7.42e-2 9.50e-2 1.06e-1
    200 4.12e-2 5.50e-2 6.49e-2
    400 2.27e-2 3.34e-2 3.88e-2
    800 1.24e-2 1.97-2 2.35e-2
    1600 6.50e-3 1.10e-2 1.35e-2

     | Show Table
    DownLoad: CSV
    Figure 14.  Example 7: numerical solution computed with BCOV scheme with N=5 and M=12800 at simulated times (a) T=0.1, (b) T=0.2 and (c) T=0.3.
    Figure 15.  Example 7: simulated total density computed with BCOV scheme with N=5 and M=1600.

    We have proposed a numerical scheme for an MCLWR model with a velocity function that is discontinuous in the solution variable. The treatment is motivated and in part based on the numerical scheme proposed by Towers [29]. However, in contrast to that approach we assume that the discontinuity is present in the velocity function (not in the flux); this observation makes it possible to construct an alternative scheme based on Scheme 4 of [6]. Furthermore, we have seen that our scheme can easily be extended to the multiclass case. We have proved for the scalar case that the numerical approximations convergence to a weak solution and for the multiclass case that the scheme preserves an invariant region. Examples 1 to 3 indicate that the scheme converges to the same weak solution as that of [29], and all numerical examples indicate that our scheme converges in both the scalar and multiclass cases.

    The present analysis and numerical method can be extended in several directions. Concerning the model itself, at the moment a certain shortcoming is the limitation to the initial-boundary value problem on a fixed road segment. This is due to the particular boundary condition (1.5). It seems desirable to obtain a formulation for a closed road with periodic boundary conditions (a configuration that is commonly studied in traffic modeling to analyze, say, the formation of stop-and-go waves; cf., e.g., [5,11]). However, it is not obvious whether the way the boundary condition is posed allows "gluing together" the ends of the computational domain to create a "seamless" closed circuit. Open issues also include the incorporation of discontinuities in spatial position (akin to the treatment in [9]), and the discussion of the notion of entropicity. In fact, the issue of convergence to an entropy solution is an open problem even in the scalar case for both the scheme advanced in [29] as well as the present approach. Likewise, we recall that for general N the MCLWR model with a Lipschitz continuous function V admits a separable entropy function (see [1,2]) that can be utilized, for instance, to construct entropy stable schemes [10]. It remains to be explored whether these concepts are meaningful for the MCLWR model with a discontinuous velocity function V. Finally, it is clear that the numerical method is formally first-order accurate and can possibly improved by known techniques (e.g., weighted essentially non-oscillatory (WENO) reconstructions in combination with higher-order time integrators).

    The authors are supported by the INRIA Associated Team "Efficient numerical schemes for non-local transport phenomena" (NOLOCO; 2018–2020). RB, RO and LMV also acknowledge support by project CMM, ANID-/PIA/AFB170001. In addition, LMV is supported by Fondecyt project 1181511 and RB by projects Fondecyt 1170473 and CRHIAM, ANID/FONDAP/15130015.



    [1] An n-populations model for traffic flow. Eur. J. Appl. Math. (2003) 14: 587-612.
    [2] Measure valued solutions to conservation laws motivated by traffic modelling. Proc. Royal Soc. A (2006) 462: 1791-1803.
    [3] On scalar hyperbolic conservation laws with a discontinuous flux. Math. Models Methods Appl. Sci. (2011) 21: 89-113.
    [4] Multi-dimensional scalar conservation laws with fluxes discontinuous in the unknown and the spatial variable. Math. Models Methods Appl. Sci. (2013) 23: 407-439.
    [5] R. Bürger, C. Chalons and L. M. Villada, Anti-diffusive and random-sampling Lagrangian-remap schemes for the multi-class Lighthill-Whitham-Richards traffic model, SIAM J. Sci. Comput., 35 (2013), B1341–B1368. doi: 10.1137/130923877
    [6] A family of numerical schemes for kinematic flows with discontinuous flux. J. Eng. Math. (2008) 60: 387-425.
    [7] Difference schemes, entropy solutions, and speedup impulse for an inhomogeneous kinematic traffic flow model. Netw. Heterog. Media (2008) 3: 1-41.
    [8] Second-order schemes for conservation laws with discontinuous flux modelling clarifier-thickener units. Numer. Math. (2010) 116: 579-617.
    [9] On some difference schemes and entropy conditions for a class of multi-species kinematic flow models with discontinuous flux. Netw. Heterog. Media (2010) 5: 461-485.
    [10] An entropy stable scheme for the multiclass Lighthill-Whitham-Richards traffic model. Adv. Appl. Math. Mech. (2019) 11: 1022-1047.
    [11] A diffusively corrected multiclass Lighthill-Whitham-Richards traffic model with anticipation lengths and reaction times. Adv. Appl. Math. Mech. (2013) 5: 728-758.
    [12] Conservation law with discontinuous flux function and boundary condition. J. Evol. Equ. (2003) 3: 283-301.
    [13] Godunov scheme and sampling technique for computing phase transitions in traffic flow modeling. Interf. Free Bound. (2008) 10: 197-221.
    [14] Hyperbolic phase transitions in traffic flow. SIAM J. Appl. Math. (2002) 63: 708-721.
    [15] On the Riemann problem for some discontinuous systems of conservation laws describing phase transitions. Commun. Pure Appl. Anal. (2004) 3: 53-58.
    [16] On the approximation of the solutions of the Riemann problem for a discontinuous conservation law. Bull. Braz. Math. Soc. New Ser. (2005) 36: 115-125.
    [17] Solutions to a scalar discontinuous conservation law in a limit case of phase transitions. J. Math. Fluid Mech. (2005) 7: 153-163.
    [18] A conservation law with point source and discontinuous flux function. SIAM J. Math. Anal. (1996) 56: 388-419.
    [19] Characteristic-based schemes for multi-class Lighthill-Whitham-Richards traffic models. J. Sci. Comput. (2008) 37: 233-250.
    [20] A secular equation for the Jacobian matrix of certain multi-species kinematic flow models. Numer. Methods Partial Differential Equations (2010) 26: 159-175.
    [21] Conservation laws with discontinuous flux functions. SIAM J. Numer. Anal. (1993) 24: 279-289.
    [22] Solution to the Cauchy problem for a conservation law with a discontinuous flux function. SIAM J. Math. Anal. (1992) 23: 635-648.
    [23] A phenomenological model for dynamic traffic flow in networks. Transp. Res. B (1995) 29: 407-431.
    [24] H. Holden and N. H. Risebro, Front Tracking for Hyperbolic Conservation Laws, 2nd edition, Springer-Verlag, Berlin, 2015. doi: 10.1007/978-3-662-47507-2
    [25] On kinematic waves: II. A theory of traffic flow on long crowded roads. Proc. Royal Soc. A (1955) 229: 317-345.
    [26] The entropy solutions for the Lighthill-Whitham-Richards traffic flow model with a discontinuous flow-density relationship. Transp. Sci. (2009) 43: 511-530.
    [27] Convergence of the finite volume method for scalar conservation law with discontinuous flux function. ESAIM Math. Model. Numer. Anal. (2008) 42: 699-727.
    [28] Shock waves on the highway. Oper. Res. (1956) 4: 42-51.
    [29] J. D. Towers, A splitting algorithm for LWR traffic models with flux discontinuities in the unknown, J. Comput. Phys., 421 (2020), 109722, 30 pp. doi: 10.1016/j.jcp.2020.109722
    [30] Riemann solver for a kinematic wave traffic model with discontinuous flux. J. Comput. Phys. (2013) 242: 1-23.
    [31] A multi-class traffic flow model–-an extension of LWR model with heterogeneous drivers. Transp. Res. A (2002) 36: 827-841.
    [32] Hyperbolicity and kinematic waves of a class of multi-population partial differential equations. Eur. J. Appl. Math. (2006) 17: 171-200.
    [33] A weighted essentially non-oscillatory numerical scheme for a multi-class Lighthill-Whitham-Richards traffic flow model. J. Comput. Phys. (2003) 191: 639-659.
    [34] A weighted essentially non-oscillatory numerical scheme for a multi-class traffic flow model on an inhomogeneous highway. J. Comput. Phys. (2006) 212: 739-756.
    [35] Hyperbolicity and kinematic waves of a class of multi-population partial differential equations. Eur. J. Appl. Math. (2006) 17: 171-200.
    [36] A note on the weighted essentially non-oscillatory numerical scheme for a multi-class Lighthill-Whitham-Richards traffic flow model. Commun. Numer. Meth. Eng. (2009) 25: 1120-1126.
    [37] A hybrid scheme for solving a multi-class traffic flow model with complex wave breaking. Comput. Methods Appl. Mech. Engrg. (2008) 197: 3816-3827.
  • This article has been cited by:

    1. Jan Friedrich, Simone Göttlich, Annika Uphoff, Conservation laws with discontinuous flux function on networks: a splitting algorithm, 2022, 18, 1556-1801, 1, 10.3934/nhm.2023001
    2. Ved Prakash Dubey, Devendra Kumar, Hashim M. Alshehri, Sarvesh Dubey, Jagdev Singh, Computational Analysis of Local Fractional LWR Model Occurring in a Fractal Vehicular Traffic Flow, 2022, 6, 2504-3110, 426, 10.3390/fractalfract6080426
    3. Bhawna Pokhriyal, Pranay Goswami, A generalized local fractional LWR model of vehicular traffic flow and its solution, 2023, 46, 0170-4214, 18899, 10.1002/mma.9598
  • 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(2443) PDF downloads(315) Cited by(3)

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog