
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
[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
∂tϕi+∂x(ϕivi(ϕ))=0,i=1,…,N. | (1.1) |
Here
vi(ϕ)=vmaxiV(ϕ),i=1,…,N, | (1.2) |
where
V(0)=1,V′(ϕ)<0,V(ϕmax)=0. |
The simplest function having all these properties is the linear interpolation
V(ϕ)={Vf(ϕ)for 0⩽ϕ⩽ϕ∗,Vc(ϕ)for ϕ∗<ϕ⩽ϕmax,Vf∈C1[0,ϕ∗],Vc∈C1[ϕ∗,ϕmax],Vf(0)=1,V′f(ϕ)⩽0on [0,ϕ∗],V′c(ϕ)⩽0on [ϕ∗,ϕmax],Vf(ϕmax)=0,αV:=Vf(ϕ∗)−Vc(ϕ∗)>0. | (1.3) |
We consider (1.1) on the domain
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
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
∂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
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)T∈RN:ϕ1⩾0,…,ϕN⩾0,ϕ=ϕ1+⋯+ϕN⩽ϕmax} |
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.
Bulí
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
Towers [29] presented a finite difference scheme that implements a splitting consistent with the decomposition the flux
{{Un+1/2j=˜G−1(Unj−λgn+1/2j+1),j=M,M−1,…,1,gn+1/2j=(Un+1/2j−Unj+λgn+1/2j+1)/λ,j=M,M−1,…,1,Un+1j=Un+1/2j−λΔ−˜p(Un+1/2j+1−Un+1/2j),j=1,…,M, | (1.7) |
which can be written in conservation form as follows:
Un+1/2j=Unj−λ(gn+1/2j+1−gn+1/2j),Un+1=Un+1/2j−λΔ−˜p(Un+1/2j+1,Un+1/2j). |
The first part of the scheme is implicit and consistent with
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 (
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(ϕ∗−ϕ),pV(ϕ):=V(ϕ)−gV(ϕ), | (2.1) |
where
G(t)∈˜gV(s(t)), | (2.2) |
where we recall that
ϕ0(x)∈[0,ϕmax]for x∈(−L,L),ϕ0∈BV([−L,L]),gV(ϕ0)∈BV([−L,L]). |
The boundary functions
r(t),s(t)∈[0,ϕmax]for t∈[0,T],r,s∈BV([0,T]). |
We also assume that
Definition 2.1 (Weak solution [16]) A function
∫T0∫L−L(ϕψt+qψx)dxdt+∫L−Lϕ0(x)ψ(x,0)dx=0. |
The domain
ϕ0j=1Δx∫Ijϕ0(x)dx,j=1,…,M, |
and the boundary conditions with
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
Based on the form of the flux function of equations (2.4) and (2.5) and the properties of the functions
ϕn+1/2j=ϕnj−λ(ϕnjgn+1/2V,j+1−ϕnj−1gn+1/2V,j),ϕn+1j=ϕn+1/2j−λ(ϕn+1/2jpV(ϕn+1/2j+1)−ϕn+1/2j−1pV(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
ϕn+1/2j=ϕnj−λ(ϕnjgV(ϕn+1/2j+1)−ϕnj−1gV(ϕn+1/2j)) | (2.7) |
along with a known value
ϕn+1/2j−λϕnj−1gV(ϕn+1/2j)=ϕnj−λϕnjgn+1/2V,j+1. | (2.8) |
Let us now define the function
GV(z;ϕ):=z−λϕgV(z),z,ϕ∈[0,ϕmax] |
along with its multivalued version (with respect to
˜GV(z;ϕ):={z−λαVϕfor z∈[0,ϕ∗),[ϕ∗−λαVϕ,ϕ∗]for ϕ=ϕ∗,zfor z∈[ϕ∗,ϕmax], | (2.9a) |
˜G−1V(z;ϕ):={z+λαVϕfor z∈[−λαVϕ,ϕ∗−λαVϕ),ϕ∗for z∈[ϕ∗−λαVϕ,ϕ∗],zfor z∈[ϕ∗,ϕmax]. | (2.9b) |
Consequently, we may express (2.8) as
˜GV(ϕn+1/2j;ϕnj−1)=ϕnj−λϕnjgn+1/2V,j+1, |
which allows us to obtain
ϕn+1/2j=˜G−1V(ϕnj−λϕnjgn+1/2V,j+1;ϕnj−1). | (2.10) |
Now that
ϕn+1/2j=ϕnj−λ(ϕnjgn+1/2V,j+1−ϕnj−1gn+1/2V,j), | (2.11) |
provided that
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
do
if
else
endif
enddo
do
enddo
Output: approximate solution vector
Next, we demonstrate that the numerical scheme (2.11) is consistent with (2.4).
Lemma 2.1. Assume that
Proof. Let us first assume that
ϕn+1/2j−λϕnj−1gn+1/2V,j∈˜GV(ϕn+1/2j;ϕnj−1). |
Therefore, by a straightforward case-by-case study (of the cases arising in (2.9)) we conclude that
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+1/2j−Dn+1/2g,j−1/2Δ−ϕnj, | (2.12a) |
ϕn+1j=ϕn+1/2j+Cn+1/2p,j+1/2Δ+ϕn+1/2j−Dn+1/2p,j−1/2Δ−ϕn+1/2j | (2.12b) |
with the spatial difference operators
Cn+1/2g,j+1/2:={λϕnjgV(ϕn+1/2j)−gV(ϕn+1/2j+1)ϕn+1/2j+1−ϕn+1/2jif ϕn+1/2j+1≠ϕn+1/2j,0otherwise,Dn+1/2g,j−1/2:=λgV(ϕn+1/2j),Cn+1/2p,j+1/2:={λϕn+1/2jpV(ϕn+1/2j)−pV(ϕn+1/2j+1)ϕn+1/2j+1−ϕn+1/2jif ϕn+1/2j+1≠ϕn+1/2j,0otherwise,Dn+1/2p,j−1/2:=λpV(ϕn+1/2j). |
To have an
0⩽Dn+1/2p,j−1/2,Cn+1/2p,j+1/2⩽12,Cn+1/2g,j+1/2⩾0,0⩽Dn+1/2g,j−1/2⩽1for all j. |
First, we observe the following fact about
gV,1∈˜gV(z1),gV,2∈˜gV(z2)⟹gV,2−gV,1z2−z1⩽0. | (2.13) |
This property and Lemma 2.1 imply that
Dn+1/2g,j−1/2,Cn+1/2g,j+1/2⩾0for all j. |
Next, the properties of the function
Cn+1/2p,j+1/2,Dn+1/2p,j−1/2⩾0for all j. |
Finally, to enforce the inequalities
Dn+1/2p,j−1/2,Cn+1/2p,j+1/2⩽12andDn+1/2g,j−1/2⩽1for all j, |
we impose the CFL conditions
λ(ϕmaxmax1⩽j⩽M|p′V(ϕj)|+max1⩽j⩽MpV(ϕj))⩽12,λαV⩽1. | (2.14) |
The goal is to prove convergence of approximate solution to a weak solution of (1.6). The discrete solutions
ϕΔ(x,t)=N∑n=0M∑j=1χj(x)χn(t)ϕn+1/2j | (2.15) |
where
We start by proving an
Lemma 2.2. If
ϕnj,ϕn+1/2j∈[0,ϕmax]forallj=1,…,Mandn=1,…,N. | (2.16) |
Proof. Taking
ϕ1/2M=˜G−1V(ϕ0M−λϕ0Mg1/2V,M+1;ϕ0M−1). | (2.17) |
The boundary condition
−λαVϕ0M⩽ϕ0M−λϕ0Mg1/2V,M+1⩽ϕmax. |
Since
Lemma 2.3. The discrete approximate solutions generated by the scheme (2.12) satisfy the following spatial variation bounds:
M∑j=0|ϕnj+1−ϕnj|⩽TV(ϕ0)+TV(r)+TV(s),M∑j=0|ϕn+1/2j+1−ϕn+1/2j|⩽TV(ϕ0)+TV(r)+TV(s). | (2.18) |
Proof. Applying the operator
(1+Cn+1/2g,j+1/2)Δ+ϕn+1/2j=(1−Dn+1/2g,j+1/2)Δ+ϕnj+Cn+1/2g,j+3/2Δ+ϕn+1/2j+1+Dn+1/2g,j−1/2Δ+ϕnj−1. |
Taking absolute values, summing over
M−1∑j=1(1+Cn+1/2g,j+1/2)|Δ+ϕn+1/2j|⩽M−1∑j=1(1−Dn+1/2g,j+1/2)|Δ+ϕnj|+M−1∑j=1Cn+1/2g,j+3/2|Δ+ϕn+1/2j+1|+M−1∑j=1Dn+1/2g,j−1/2|Δ+ϕnj−1|. |
Cancelling telescoping terms we obtain
M−1∑j=1|Δ+ϕn+1/2j|+Cn+1/2g,3/2|Δ+ϕn+1/21|⩽M−1∑j=1|Δ+ϕnj|−Dn+1/2g,M−1/2|Δ+ϕnM−1|+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=(1−Dn+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,M−1/2Δ+ϕn+1/21. |
After taking absolute values in the two previous equations, we get
|Δ+ϕn+1/20|⩽(1−Dn+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,M−1/2|Δ+ϕn+1/2M−1|. | (2.20) |
From (2.19) and (2.20)
M∑j=0|Δ+ϕn+1/2j|⩽M∑j=0|Δ+ϕnj|. | (2.21) |
Reasoning in the same way as the proof of Lemma 5.2 in [29] we find that
M∑j=0|Δ+ϕn+1j|⩽M∑j=0|Δ+ϕnj|+|rn+1−rn|+|sn+1−sn|. |
It follows by induction that
M∑j=0|ϕnj+1−ϕnj|⩽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
M+1∑j=0|ϕn+1j−ϕn+1/2j|⩽Ω1,Ω1:=TV(ϕ0)+TV(r)+TV(s)+2ϕmax. |
Lemma 2.5. The following estimate holds:
M∑j=1|ϕ1/2j−ϕ0j|⩽Ω2,Ω2:=M∑j=1|gV(ϕ0j+1)−gV(ϕ0j)|+TV(ϕ0)+ϕmax. | (2.23) |
Proof. We define
ϕ1/2j−ϕ0j=λϕ0j−1(g1/2V,j−g0V,j)−λϕ0j(g1/2V,j+1−g0V,j+1)−λϕ0j(Δ+g0V,j)−λg0V,j(Δ+ϕ0j−1). |
Thus
ϕ1/2j−ϕ0j−λϕ0j−1(g1/2V,j−g0V,j)=λϕ0j(g1/2V,j+1−g0V,j+1)−λ(ϕ0jΔ+g0V,j+g0V,jΔ+ϕ0j−1). | (2.24) |
Taking absolute values in (2.24) and using (2.13) we find that
|ϕ1/2j−ϕ0j|+λϕ0j−1|g1/2V,j−g0V,j|⩽λϕ0j|g1/2V,j+1−g0V,j+1|+λϕ0j|Δ+g0V,j|+λg0V,j|Δ+ϕ0j−1|. |
Summing over
M∑j=1|ϕ1/2j−ϕ0j|+λϕ00|g1/2V,1−g0V,1|⩽λϕ0M|g1/2V,M+1−g0V,M+1|+λM∑j=1ϕ0j|Δ+g0V,j|+λM∑j=1g0V,j|Δ+ϕ0j−1|. | (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
M+1∑j=0|ϕn+1/2j−ϕn−1/2j|⩽Ω3forn⩾1. | (2.26) |
Proof. For
ϕn+1/2j−ϕn−1/2j−λϕn−1j−1(gn+1/2V,j−gn−1/2V,j)=(1−λgn+1/2V,j+1)(ϕnj−ϕn−1j)+λgn+1/2V,j(ϕnj−1−ϕn−1j−1)−λϕn−1j(gn+1/2V,j+1−gn−1/2V,j+1). |
Taking absolute values and applying the CFL condition (2.14) yields
|ϕn+1/2j−ϕn−1/2j−λϕn−1j−1(gn+1/2V,j−gn−1/2V,j)|⩽(1−λgn+1/2V,j+1)|ϕnj−ϕn−1j|+λgn+1/2V,j|ϕnj−1−ϕn−1j−1|+|λϕn−1j(gn+1/2V,j+1−gn−1/2V,j+1)|. |
From (2.13) we get
|ϕn+1/2j−ϕn−1/2j|+|λϕn−1j−1(gn+1/2V,j−gn−1/2V,j)|⩽(1−λgn+1/2V,j+1)|ϕnj−ϕn−1j|+λgn+1/2V,j|ϕnj−1−ϕn−1j−1|+|λϕn−1j(gn+1/2V,j+1−gn−1/2V,j+1)|. | (2.27) |
Summing over
M∑j=1|ϕn+1/2j−ϕn−1/2j|⩽M∑j=1|ϕnj−ϕn−1j|+λgn+1/2V,1|ϕn0−ϕn−10|+λϕn−1M|gn+1/2V,M+1−gn−1/2V,M+1|−λgn+1/2V,M+1|ϕnM−ϕn−1M|. |
The last inequality implies
M∑j=1|ϕn+1/2j−ϕn−1/2j|⩽M∑j=1|ϕnj−ϕn−1j|+|rn−rn−1|+|G(tn)−G(tn−1)|. |
We observe that
ϕnj−ϕn−1j=(1−Bn−1j+1/2−An−1j−1/2)(ϕn−1/2j−ϕn−3/2j)+An−1j+1/2(ϕn−1/2j+1−ϕn−3/2j+1)+Bn−1j−1/2(ϕn−1/2j−1−ϕn−3/2j−1), | (2.28) |
where
An−1j+1/2=−λ∫10∂1φ(θϕn−1/2j+1+(1−θ)θϕn−3/2j+1,θϕn−1/2j+1+(1−θ)θϕn−3/2j+1)dθ,Bn−1j+1/2=λ∫10∂2φ(θϕn−1/2j+1+(1−θ)θϕn−3/2j+1,θϕn−1/2j+1+(1−θ)θϕn−3/2j+1)dθ. |
Herein
0⩽An−1j+1/2,Bn−1j+1/2⩽12. | (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
Lemma 2.7. The functions
Proof. The proof is a standard argument using the
In order to show that the limit function
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
hΔ(x,t):=N∑n=0M∑j=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
M∑j=1|ϕn+1j−ϕnj|⩽Ω4,Ω4:=Ω2+TV(s)+TV(r), | (2.30) |
M∑j=1|ϕn+1/2j−ϕnj|⩽Ω1+Ω4. | (2.31) |
The following lemma is needed to establish a spatial variation bound on the approximations
Lemma 2.9. There exists a constant
M∑j=1ϕnj|Δ+gn+1/2V,j|⩽Ω5. |
Proof. From the first half-step of the scheme we get
ϕn+1/2j−ϕnj=−λ(ϕnjΔ+gn+1/2V,j+gn+1/2V,jΔ+ϕnj−1), |
which can be rerranged as
λϕnjΔ+gn+1/2V,j=−(ϕn+1/2j−ϕnj)−λgn+1/2V,jΔ+ϕnj−1. |
Taking absolute values and summing over
λM∑j=1ϕnj|Δ+gn+1/2V,j|⩽M∑j=1|ϕn+1/2j−ϕnj|+λM∑j=1gn+1/2V,j|Δ+ϕnj−1|. |
From Lemma 2.1 and the CFL condition (2.14) we have
M∑j=1ϕnj|Δ+gn+1/2V,j|⩽1λM∑j=1|ϕn+1/2j−ϕnj|+M∑j=1|Δ+ϕnj−1|. |
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
Lemma 2.10. There exists a constant
M∑j=1|hn+1/2j+1−hn+1/2j|⩽Ω6. | (2.32) |
Proof. The first part of scheme (2.6) can be written as
ϕn+1/2j=ϕnj−λ(ϕnjgn+1/2V,j+1+gn+1/2V,j(ϕn+1/2j−ϕnj−1))+λϕn+1/2jgn+1/2V,j. |
Applying the spatial difference operator to the above equation we get
Δ+ϕn+1/2j=(1−λgn+1/2V,j+1)Δ+ϕnj+λΔ+hn+1/2j−λϕnj+1Δ+gn+1/2V,j+1−λΔ+gn+1/2V,j(ϕn+1/2j−ϕnj)+λgn+1/2V,jΔ+ϕnj−1−λgn+1/2V,j+1Δ+ϕn+1/2j. |
Thus
λΔ+hn+1/2j=Δ+ϕn+1/2j−(1−λgn+1/2V,j+1)Δ+ϕnj+λϕnj+1Δ+gn+1/2V,j+1+λΔ+gn+1/2V,j(ϕn+1/2j−ϕnj)+λgn+1/2V,jΔ+ϕnj−1+λgn+1/2V,j+1Δ+ϕn+1/2j. |
After taking absolute values and using
λ|Δ+hn+1/2j|⩽2|Δ+ϕn+1/2j|+|Δ+ϕnj|+λϕnj+1|Δ+gn+1/2V,j+1|+|ϕn+1/2j−ϕnj|+|Δ+ϕnj−1|. |
Summing over
M∑j=1|Δ+hn+1/2j|⩽2λM∑j=1|Δ+ϕn+1/2j|+2λM∑j=1|Δ+ϕnj|+1λM∑j=1|ϕn+1/2j−ϕnj|+M∑j=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
Lemma 2.11. There exists a constant
ΔxM∑j=1N∑n=1ϕn−1j|gn+1/2V,j+1−gn−1/2V,j+1|⩽Ω7. | (2.33) |
Proof. From (2.27) we get
λϕn−1j|gn+1/2V,j+1−gn−1/2V,j+1|⩽(1−λgn+1/2V,j+2)|ϕnj+1−ϕn−1j+1|+λgn+1/2V,j+1|ϕnj−ϕn−1j|−|ϕn+1/2j+1−ϕn−1/2j+1|+λϕn−1j+1|gn+1/2V,j+2−gn−1/2V,j+2|. |
By induction we obtain
λϕn−1j|gn+1/2V,j+1−gn−1/2V,j+1|⩽M∑k=j+1|ϕnk−ϕn−1k|+|gn+1/2V,M+1−gn−1/2V,M+1|−M∑k=j+1|ϕn+1/2k−ϕn−1/2k|+|ϕnj−ϕn−1j|. | (2.34) |
Recalling (2.28) we have
M∑k=j+1|ϕnk−ϕn−1k|⩽M∑k=j+1(1−Bn−1k+1/2−An−1k−1/2)|ϕn−1/2k−ϕn−3/2k|+M∑k=j+1An−1k+1/2|ϕn−1/2k+1−ϕn−3/2k+1|+M∑k=j+1Bn−1k−1/2|ϕn−1/2k−1−ϕn−3/2k−1|. |
Cancelling telescoping terms and applying (2.29) yields
M∑k=j+1|ϕnk−ϕn−1k|⩽M∑k=j+1|ϕn−1/2k−ϕn−3/2k|+12|ϕn−1/2M+1−ϕn−3/2M+1|+12|ϕn−1/2j−ϕn−3/2j|. |
Then (2.34) becomes
λϕn−1j|gn+1/2V,j+1−gn−1/2V,j+1|⩽M∑k=j+1|ϕn−1/2k−ϕn−3/2k|−M∑k=j+1|ϕn+1/2k−ϕn−1/2k|+12|ϕn−1/2M+1−ϕn−3/2M+1|+12|ϕn−1/2j−ϕn−3/2j|+|ϕnj−ϕn−1j|+|gn+1/2V,M+1−gn−1/2V,M+1|. |
Summing over
ΔxM∑j=1N∑n=2ϕn−1j|gn+1/2V,j+1−gn−1/2V,j+1|⩽S1+⋯+S5, |
where we define
S1:=2LλM∑j=1|ϕ3/2j−ϕ1/2j|,S2:=LλN∑n=2|sn−1/2−sn−3/2|,S3:=ΔxλM∑j=1N∑n=2|ϕnj−ϕn−1j|,S4:=2LλN∑n=2|gn+1/2V,M+1−gn−1/2V,M+1|,S5:=Δx2λM∑j=1N∑n=2|ϕn−1/2j−ϕn−3/2j|. |
In view of the bounds established so far, there holds
S1⩽2LλΩ3,S2⩽LλTV(s),S3⩽Ω4T,S4⩽2LλTV(G),S5⩽Ω32T. |
These bounds in conjunction with
Lemma 2.12. There exists a constant
ΔtN∑n=0M∑j=1|hn+1/2j+1−hn+1/2j|+ΔxN∑n=1M∑j=1|hn+1/2j−hn−1/2j|⩽Ω8. | (2.35) |
Proof. In light of the spatial variation bound (2.32) we find that
ΔtN∑n=0M∑j=1|hn+1/2j+1−hn+1/2j|⩽Ω6T. |
The first part of (2.6) implies
ϕn+1/2j−ϕn−1/2j=(1−λgn+1/2V,j+1)(ϕnj−ϕn−1j)+λ(hn+1/2j−hn−1/2j)−λgn+1/2V,j(ϕn+1/2j−ϕnj−1)+λgn−1/2V,j(ϕn−1/2j−ϕn−1j−1)−λϕn−1j(gn+1/2V,j+1−gn−1/2V,j+1)=(1−λgn+1/2V,j+1)(ϕnj−ϕn−1j)+λ(hn+1/2j−hn−1/2j)−λgn+1/2V,j(ϕn+1/2j−ϕnj)−λgn+1/2V,jΔ+ϕnj−1+λgn−1/2V,jΔ+ϕn−1j−1+λgn−1/2V,j(ϕn−1/2j−ϕn−1j)−λϕn−1j(gn+1/2V,j+1−gn−1/2V,j+1). |
Consequently,
λ(hn+1/2j−hn−1/2j)=(ϕn+1/2j−ϕn−1/2j)−(1−λgn+1/2V,j+1)(ϕnj−ϕn−1j)+λgn+1/2V,j(ϕn+1/2j−ϕnj)+λgn+1/2V,jΔ+ϕnj−1−λgn−1/2V,jΔ+ϕn−1j−1−λgn−1/2V,j(ϕn−1/2j−ϕn−1j)+λϕn−1j(gn+1/2V,j+1−gn−1/2V,j+1). |
Taking absolute values and using the CFL condition (2.14) we get
λ|hn+1/2j−hn−1/2j|⩽|ϕn+1/2j−ϕn−1/2j|+|ϕnj−ϕn−1j|+|Δ+ϕnj−1|+|Δ+ϕn−1j−1|+|ϕn+1/2j−ϕnj|+|ϕn−1/2j−ϕn−1j|+λϕn−1j|gn+1/2V,j+1−gn−1/2V,j+1|. |
Multiplying this inequality by
ΔxN∑n=1M∑j=1|hn+1/2j−hn−1/2j|⩽U1+⋯+U5, |
where we define
U1:=ΔxλN∑n=1M∑j=1|ϕn+1/2j−ϕn−1/2j|,U2:=ΔxλN∑n=1M∑j=1|ϕnj−ϕn−1j|,U3:=2ΔxλN∑n=1M∑j=1|Δ+ϕnj−1|,U4:=2ΔxλN∑n=1M∑j=1|ϕn+1/2j−ϕnj|,U5:=ΔxN∑n=1M∑j=1ϕn−1j|gn+1/2V,j+1−gn−1/2V,j+1|. |
From (2.18), (2.26), (2.30), (2.31) and (2.33) we have
U1⩽Ω3T,U2⩽Ω4T,U3⩽2(TV(ϕ0)+TV(s)+TV(r))T,U4⩽2(Ω1+Ω4)T,U5⩽Ω7. |
Combining these bounds we see that there exists a constant
In the following lemma we state and prove convergence of the functions
Lemma 2.13. The functions
Proof. We observe that
w(x,t)∈[0,αVϕ∗]=˜Q(ϕ∗). |
In case
hΔ(x,t)=N∑n=0M∑j=1χj(x)χn(t)hn+1/2j=αVN∑n=0M∑j=1χj(x)χn(t)ϕn+1/2j=αVϕΔ(x,t). |
Thus
In the case
gΔV(x,t)=N∑n=0M∑j=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
Theorem 2.14 (Main result). The functions
Proof. The convergence is ensured by Lemma 2.7. It remains to prove that the limit
q(x,t)=ϕ∗pV(ϕ∗)+w(x,t)∈[ϕ∗pV(ϕ∗),ϕ∗pV(ϕ∗)+αVϕ∗]=˜f(ϕ∗). |
In either case
We note that the two steps of (2.6) imply
ϕn+1j−ϕnj+λ(ϕnjgn+1/2V,j+1−ϕnj−1gn+1/2V,j+ϕn+1/2jpn+1/2V,j+1−ϕn+1/2j−1pn+1/2V,j)=0. | (2.36) |
We now choose a test function
ΔxΔtN∑n=0M∑j=1ϕn+1j−ϕnjΔtψnj+ΔxΔtN∑n=0M∑j=1ϕnjgn+1/2V,j+1−ϕnj−1gn+1/2V,jΔxψnj+ΔxΔtN∑n=0M∑j=1ϕn+1/2jpn+1/2V,j+1−ϕn+1/2j−1pn+1/2V,jΔxψnj=0. |
A summation by parts yields
ΔxΔtN∑n=0M∑j=1ϕn+1jψn+1j−ψnjΔt+ΔxM∑j=1ϕ0jψ0j+ΔxΔtN∑n=0M∑j=1ψnj+1−ψnjΔxϕnjgn+1/2V,j+1+ΔxΔtN∑n=0M∑j=1ψnj+1−ψnjΔxϕn+1/2jpn+1/2V,j+1=0. | (2.37) |
An application of (2.12) yields, as
ΔxΔtN∑n=0M∑j=1ϕn+1jψn+1j−ψnjΔt=ΔxΔtN∑n=0M∑j=1ϕn+1/2jψn+1j−ψnjΔt+O(Δx). |
This equation and Lemma 2.3 imply that the two first sums in (2.37) converge to
∫T0∫L−Lϕψtdxdt+∫L−Lϕ0(x)ψ(x,0)dxdt. |
Concerning the last term in (2.37), we get
ΔxΔtN∑n=0M∑j=1ψnj+1−ψnjΔxϕn+1/2jpn+1/2V,j+1=ΔxΔtN∑n=0M∑j=1ψnj+1−ψnjΔxϕn+1/2jpn+1/2V,j+ΔxΔtN∑n=0M∑j=1ψnj+1−ψnjΔx(ϕn+1/2jpn+1/2V,j+1−ϕn+1/2jpn+1/2V,j). |
By properties of the function
ϕn+1/2jpn+1/2V,j+1−ϕn+1/2jpn+1/2V,j=ϕn+1/2j(pn+1/2V,j+1−pn+1/2V,j)⩽ϕmax‖p′V‖∞Δx. |
Thus
|ΔxΔtN∑n=0M∑j=1(ψnj+1−ψnj)Δx(ϕn+1/2jpn+1/2V,j+1−ϕn+1/2jpn+1/2V,j)|⩽2MTϕmaxΔx‖ψt‖∞‖p′V‖∞, |
which tends to zero as
∫T0∫L−LϕpV(ϕ)ψxdxdt |
as
ΔxΔtN∑n=0M∑j=1ψnj+1−ψnjΔxϕnjgn+1/2V,j+1=ΔxΔtN∑n=0M∑j=1ψnj+1−ψnjΔxϕn+1/2jgn+1/2V,j+ΔxΔtN∑n=0M∑j=1ψnj+1−ψnjΔxϕnjΔ+gn+1/2V,j+ΔxΔtN∑n=0M∑j=1ψnj+1−ψnjΔxgn+1/2V,j(ϕnj−ϕn+1/2j). |
Using Lemmas 2.9, 2.1, and 2.8 we get
|ΔxΔtN∑n=0M∑j=1ψnj+1−ψnjΔxϕnjΔ+gn+1/2V,j|⩽Δx‖ψx‖∞Ω5T,|ΔxΔtN∑n=0M∑j=1ψnj+1−ψnjΔxgn+1/2V,j(ϕnj−ϕn+1/2j)|⩽αVΔx‖ψx‖∞(Ω1+Ω4)T. |
Consequently, as
ΔxΔtN∑n=0M∑j=1ψnj+1−ψnjΔxϕnjΔ+gn+1/2V,j→0,ΔxΔtN∑n=0M∑j=1ψnj+1−ψnjΔxgn+1/2V,j(ϕnj−ϕn+1/2j)→0. |
Then substituting
∫T0∫L−Lwψxdxdt. |
Collecting the previous results we get
∫T0∫L−L(ϕψt+qψx)dxdt+∫L−Lϕ0(x)ψ(x,0)dx=0, |
so
Algorithm 2.1 cannot be applied directly in a component-wise manner for each class
ϕn+1/2i,j=ϕni,j−λvmaxi(ϕni,jgV(ϕn+1/2j+1)−ϕni,j−1gV(ϕn+1/2j)),ϕn+1i,j=ϕn+1/2i,j−λvmaxi(ϕn+1/2i,jpV(ϕn+1/2j+1)−ϕn+1/2i,j−1pV(ϕn+1/2j)),i=1,…,N,j=1,…,M, | (3.1) |
where the following quantity is an approximate value of the total density
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
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
ϕn+1/2j=ϕnj−λ(vmax)T(gn+1/2V,j+1Φnj−gV(ϕn+1/2j)Φnj−1). | (3.3) |
This can be rearranged as
ϕn+1/2j−λ(vmax)TΦnj−1gV(ϕn+1/2j)=ϕnj−λ(vmax)TΦnjgn+1/2V,j+1. | (3.4) |
Let us now define the function
˜GV(ϕn+1/2j;Φnj−1)=ϕnj−λ(vmax)TΦnjgn+1/2V,j+1 | (3.5) |
which allows us to obtain
ϕn+1/2j=˜G−1V(ϕnj−λ(vmax)TΦnjgn+1/2V,j+1;Φnj−1). |
Now that
ϕn+1/2j=ϕnj−λ(vmax)T(gn+1/2V,j+1Φnj−gn+1/2V,jΦnj−1) |
(cf. (3.3)). This yields
gn+1/2V,j=ϕn+1/2j−ϕnj+λgn+1/2V,j+1(vmax)TΦnjλ(vmax)TΦnj−1, |
provided that
Algorithm 3.1 (BCOV scheme, multiclass case).
Input: approximate solution vector
do
if
else
endif
enddo
do
do
enddo
enddo
do
do
enddo
enddo
Output: approximate solution vectors
Remark 3.1 We recall that
The problem of interest to us is to show that
ϕn+1j=ϕn+1/2j−λ(vmax)T(pV(ϕn+1/2j+1)n+1/2Φn+1/2j−pV(ϕn+1/2j)Φnj−1). |
The above equation can be written in incremental form as
ϕn+1j=ϕn+1/2j+Cn+1/2j+1/2Δ+ϕn+1/2j−Dn+1/2j−1/2Δ−ϕn+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+1−ϕn+1/2jif ϕn+1/2j+1≠ϕn+1/2j,0if ϕn+1/2j+1=ϕn+1/2j,Dn+1/2j−1/2:={λpV(ϕn+1/2j)(vmax)T(Φn+1/2j−Φn+1/2j−1)ϕn+1/2j−ϕn+1/2j−1if ϕn+1/2j≠ϕn+1/2j−1,0if ϕn+1/2j=ϕn+1/2j−1. |
Since
λϕmaxmax1⩽j⩽M|p′V(ϕnj)|⋅max1⩽i⩽Nvmaxi⩽12,λmax1⩽j⩽Mp(ϕnj)⋅max1⩽i⩽Nvmaxi⩽12. | (3.7) |
Lemma 3.1. Assume that
Proof. We claim that
Φn+1/2j∈Dfor all j=1,…,M⇒gn+1/2V,j∈[0,αV]for all j=1,…,M. | (3.8) |
In the case
ϕn+1/2j=˜G−1V(ϕnj−λ(vmax)TΦnjgn+1/2V,j+1;Φnj−1). |
Using (3.5) and (3.3) we find that
ϕn+1/2j−λ(vmax)TΦnj−1gn+1/2V,j∈˜GV(ϕn+1/2j;Φnj−1). |
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 (
V(ϕ)={1−ϕ/ϕmaxfor 0⩽ϕ⩽ϕ∗,−wf(1−ϕmax/ϕ)for ϕ∗<ϕ⩽ϕmax, |
where
In all numerical experiments computations are performed on a finite interval
We consider the Riemann problem for the scalar equation
ϕ0(x)={ϕLfor x<0.2,ϕRfor x⩾0.2 | (4.1) |
(no boundary conditions are involved). For
For
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
Towers | BCOV | Towers | BCOV | |
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 |
This example comes from [29,Example 6.2] and is designed to illustrate that when
To illustrate the invariant region property of the proposed scheme (Lemma 3.1), we consider the case
ϕ0(x)={(0.1,0.1,0.1)Tfor x<0.5,(0.4,0.5,0.1)Tfor x⩾0.5, |
with velocities
It is the purpose of this example to illustrate the boundary condition
gn+1/2V,M+1=G(ϕn+1/2M+1), |
where
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 x⩾0. |
Observe that
As in Example 3 we show that the solution depends on the boundary condition
In this example we consider
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
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 |
In this example we consider
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(x−2)2),ψ2(x):=exp(−50(x−1)2/4) |
for
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 |
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
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. ![]() |
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 |
Towers | BCOV | Towers | BCOV | |
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 |
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 |
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 |
Towers | BCOV | Towers | BCOV | |
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 |
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 |
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 |