Loading [MathJax]/jax/output/SVG/jax.js
Research article Special Issues

Numerical schemes for a class of nonlocal conservation laws: a general approach

  • In this work we present a rather general approach to approximate the solutions of nonlocal conservation laws. In a first step, we approximate the nonlocal term with an appropriate quadrature rule applied to the spatial discretization. Then, we apply a numerical flux function on the reduced problem. We present explicit conditions which such a numerical flux function needs to fulfill. These conditions guarantee the convergence to the weak entropy solution of the considered model class. Numerical examples validate our theoretical results and demonstrate that the approach can be applied to other nonlocal problems.

    Citation: Jan Friedrich, Sanjibanee Sudha, Samala Rathan. Numerical schemes for a class of nonlocal conservation laws: a general approach[J]. Networks and Heterogeneous Media, 2023, 18(3): 1335-1354. doi: 10.3934/nhm.2023058

    Related Papers:

    [1] Raimund Bürger, Harold Deivi Contreras, Luis Miguel Villada . A Hilliges-Weidlich-type scheme for a one-dimensional scalar conservation law with nonlocal flux. Networks and Heterogeneous Media, 2023, 18(2): 664-693. doi: 10.3934/nhm.2023029
    [2] Abraham Sylla . Influence of a slow moving vehicle on traffic: Well-posedness and approximation for a mildly nonlocal model. Networks and Heterogeneous Media, 2021, 16(2): 221-256. doi: 10.3934/nhm.2021005
    [3] 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
    [4] Jan Friedrich, Oliver Kolb, Simone Göttlich . A Godunov type scheme for a class of LWR traffic flow models with non-local flux. Networks and Heterogeneous Media, 2018, 13(4): 531-547. doi: 10.3934/nhm.2018024
    [5] Dong Li, Tong Li . Shock formation in a traffic flow model with Arrhenius look-ahead dynamics. Networks and Heterogeneous Media, 2011, 6(4): 681-694. doi: 10.3934/nhm.2011.6.681
    [6] Paola Goatin, Chiara Daini, Maria Laura Delle Monache, Antonella Ferrara . Interacting moving bottlenecks in traffic flow. Networks and Heterogeneous Media, 2023, 18(2): 930-945. doi: 10.3934/nhm.2023040
    [7] Raimund Bürger, Stefan Diehl, M. Carmen Martí, Yolanda Vásquez . A difference scheme for a triangular system of conservation laws with discontinuous flux modeling three-phase flows. Networks and Heterogeneous Media, 2023, 18(1): 140-190. doi: 10.3934/nhm.2023006
    [8] Michael Herty, Niklas Kolbe, Siegfried Müller . Central schemes for networked scalar conservation laws. Networks and Heterogeneous Media, 2023, 18(1): 310-340. doi: 10.3934/nhm.2023012
    [9] Maya Briani, Emiliano Cristiani . An easy-to-use algorithm for simulating traffic flow on networks: Theoretical study. Networks and Heterogeneous Media, 2014, 9(3): 519-552. doi: 10.3934/nhm.2014.9.519
    [10] Alexander Kurganov, Anthony Polizzi . Non-oscillatory central schemes for traffic flow models with Arrhenius look-ahead dynamics. Networks and Heterogeneous Media, 2009, 4(3): 431-451. doi: 10.3934/nhm.2009.4.431
  • In this work we present a rather general approach to approximate the solutions of nonlocal conservation laws. In a first step, we approximate the nonlocal term with an appropriate quadrature rule applied to the spatial discretization. Then, we apply a numerical flux function on the reduced problem. We present explicit conditions which such a numerical flux function needs to fulfill. These conditions guarantee the convergence to the weak entropy solution of the considered model class. Numerical examples validate our theoretical results and demonstrate that the approach can be applied to other nonlocal problems.



    Scalar conservation laws with a nonlocal transport term appear in a variety of applications and can be formally summarized by

    tρ+xF(t,x,ρ,R)=0(t,x)R+×R, (1.1)

    where ρ is the state variable, t the time, x the space variable, F a flux function and R an integral evaluation over the space. Typically, R is a convolution involving the state variable ρ over a possible compact space. Such nonlocal terms can be found in various physical applications such as supply chains [15], sedimentation [4,6], conveyor belts [25,36], crowd motion [13], and traffic flow [5,8,9,10,19,21,22,31]. Thereby, these equations are studied theoretically for existence and uniqueness [2,3,28,30]. The challenge here is that classical approaches for local conservation laws cannot be applied, e.g., due to the nonlocal term, no Riemann-solvers are currently available. But one can use numerical schemes [2,3] and Kruzkov's doubling of variables technique [32] or fixed point approaches based on the method of characteristics [28,30,31] to prove uniqueness and existence of solutions. Solutions must be understood in the weak sense by removing the strong regularity conditions on the solution as discontinuities may be present. For specific flux functions, i.e., of the form F(t,x,ρ,R)=ρv(R) with a suitable nonlocal term R and velocity function v, weak solutions are already unique [28], but for more general flux functions entropy conditions are used to single out the physically correct solutions [2,3]. Nevertheless, in the aforementioned works and in all the aforementioned applications, mostly specific forms of the involved flux functions and nonlocal terms are assumed. In its whole generality the well-or ill-posedness (1.1) is not proved yet.

    From a numerical point of view the situation is very similar. In most of the applications tailored Lax-Friedrichs-type numerical schemes have been used as a tool to prove the existence of solutions. However, it is well known that Lax-Friedrichs type numerical schemes produce diffusive behavior. To avoid such a problem, a Godunov-type scheme for a specific traffic flow model has been introduced in [21]. Furthermore, higher-order schemes have been considered [7,20], but they rely on already derived numerical flux functions. Numerical flux functions can also be derived by using Lagrangian-remap schemes as in [1,11]. So far, these approaches are tailored to specific flux functions* and nonlocal terms, too. An alternative is to use numerical schemes without dissipation which are based on the method of characteristics [29, Section 5], although the flux function must allow for a unique weak solution, i.e., be of the form F(t,x,ρ,R)=ρv(R).

    *We note that [1] deals with a similar flux function as we do in the following. However, for the rigorous proof of convergence they consider a simpler version of the flux.

    Let us compare this situation with a scalar one-dimensional (local) hyperbolic conservation law of the form

    ρt+f(ρ)x=0,(t,x)R+×R,ρ(0,x)=ρ0(x),xR. (1.2)

    Here, not only the existence and uniqueness theory is well understood, but also the construction of numerical solutions for local conservation laws. For more details, we refer to [17,23,26,34,35,37]. In particular, to solve the equation (1.2) numerically, one can use the most familiar three-point schemes in conservation form, i.e.,

    ρn+1j=ρnjΔtΔx(G(ρnj,ρnj+1)G(ρnj1,ρnj)), (1.3)

    where ρnjρ(tn,xj), numerical flux G is Lipschitz continuous function and is consistent with the physical flux f=f(u), i.e., G(u,u)=f(u). If the numerical flux G is monotone, then the scheme (1.3) is total variation diminishing (TVD) and satisfies the maximum principle [14,17,35]. By choosing a suitable monotone numerical flux function G, the numerical solution converges to a weak entropy solution of the conservation law (1.2). In particular, there are several choices of monotone numerical flux functions, so one can choose the flux functions that are the best suited for each individual modeling problem.

    As aforementioned for nonlocal conservation laws, such a general class of numerical schemes does not currently exist. One difficulty is that in general the schemes are no longer monotone due to the nonlocal term, see e.g., [5,27]. Recently, a first attempt to consider a more general class of numerical schemes is considered in [27], however they consider very specific assumptions on the flux function, see Remark 3.5 for a more detailed discussion. In this work we present a different approach to construct a general class of numerical schemes than in [27]. In particular, we consider a rather general nonlocal flux function to further fill the gap in the current literature. The derived schemes will thereby share some similarities with the class of monotone schemes for local conservation laws.

    We organize this paper as follows: In Section 2, we recall the main results of [9] on the existence and uniqueness of entropy solutions for a class of nonlocal conservation laws. Then in Section 3 we present an approach to construct numerical schemes for the considered model class. This includes necessary assumptions on the numerical flux function. Furthermore, we present several numerical schemes belonging to this class, e.g., a newly derived Engquist-Osher type scheme. Section 4 presents our main result on the convergence of the calls of numerical schemes. The proof is given in the remaining section and we discuss some important properties of the proposed schemes, such as a maximum principle, bounded variation (BV) and time-continuity estimates. In Section 5, we present numerical examples and compare the L1-accuracy of some exemplary schemes belonging to the proposed class. Furthermore, we calculate the numerical convergence rates and demonstrate how the approach can be adapted to other classes of nonlocal conservation laws. Finally, concluding remarks are given in Section 6.

    The goal of this work is to present a general approach to construct numerical schemes for nonlocal conservation laws. Therefore, we rely on an already well-posed class of nonlocal conservation laws. Here, we briefly recall the model class and the main result of ref. [9]. In [9], the following nonlocal conservation law is studied:

    tρ+x(g(ρ)V(t,x))=0, (2.1)
    V(t,x):=v(x+ηxωη(yx)ρ(t,y)dy). (2.2)

    This model considers a mean downstream density and is commonly used in various traffic flow models. For example, it can be used to describe a nonlocal LWR model or Arrhenius-type look-ahead dynamics. To prove well-posedness, the authors of [9] need the following hypotheses on the functions involved:

    Assumption 2.1. We assume the following hypotheses:

    1. ρ(0,x)=ρ0(x)BV(R,I),I=[ρmin,ρmax]R+,

    2. gC1(I;R+),

    3. vC2(I;R+),v0,

    4. ωηC1([0,η]),R+) with ω0,η0ωη(x)dx=1η>0,limηωη(0)=0,

    where ρmin:=minR(ρ0) and ρmax:=maxR(ρ0).

    Note that, unlike [9], we consider normalized kernel functions ωη, which means that the L1-norm is one and the interval for the density is [0,1]. Due to the presence of the function g, which can be nonlinear, the results of e.g., [28] cannot be applied. These results hold only for the special case where g is the identity. Hence, instead of weak solutions the authors of [9] consider, as usual for local conservation laws, weak entropy solutions of the problem (2.1). These are intended in the following sense, see [9, Definition 1.1]:

    Definition 2.2. A function ρ(L1LBV)(R+×R;R) is a weak entropy solution of Eq (2.1) with initial condition ρ(0,x)=ρ0(x), if

    +0+(|ρk|ϕt+|ρk|Vϕxsgn(ρk)kVxϕ)(t,x)dxdt++|ρ0(x)k|ϕ(0,x)dx0 (2.3)

    for all ϕC1c(R2;R)andkR.

    Now, we can state the main result of [9]:

    Theorem 2.3 (compare to Theorem 1.2 of [9]). Let the Assumptions 2.1 hold, then the Cauchy problem

    {tρ(t,x)+x(g(ρ(t,x))V(t,x))=0,xR,t>0,ρ(0,x)=ρ0(x),xR

    with V(t,x) defined as in Eq (2.2) admits a unique weak entropy solution in the sense of Definition 2.2, such that

    ρmin=minR(ρ0)ρ(t,x)maxR(ρ0)=ρmax,for a.e.xR,t>0.

    In [9] a Lax-Friedrichs-type numerical scheme is used as a tool to prove the existence of solutions. Our main goal now is to derive a complete class of numerical schemes, which provides the flexibility to choose or develop numerical schemes that are the best suited for each application. In particular, the numerical schemes may be less diffusive than a Lax-Friedrichs-type numerical scheme.

    In the following we will introduce a more general class of numerical schemes for the problem (2.1). From now on, unless otherwise stated, defines for simplicity the L norm over the underlying space, e.g., v:=vL([ρmin, ρmax]).

    We discretize space and time by an equidistant grid, where Δx is the step size in space and Δt is the step size in time. Hence, tn=nΔt with nN describes the time grid and xj=jΔx, jZ the cell centers of the space grid with the cell interfaces xj12 and xj+12. To construct a finite volume approximation ρΔx such that ρΔx(t,x)=ρnj for (t,x)[tn,tn+1)×[xj12,xj+12), we approximate the initial data by

    ρ0j=1Δxxj+12xj12ρ0(x)dx,jZ.

    The scheme is then given by

    ρn+1j=ρnjλ(Fnj+12(ρnj,ρnj+1)Fnj12(ρnj1,ρnj))withλ:=ΔtΔx, (3.1)

    with a suitable numerical flux Fnj+12(a,b) which needs to be determined.

    Therefore, we want to consider a rather general approach to derive the numerical flux function that approximates the flux passing through the cell interface xj+12. Our main idea is as follows: We consider the piecewise constant reconstruction of the density ρΔx. Then, the nonlocal term (2.2) evaluated at the cell boundary xj+12 and time tn is a fixed value which we call Vnj, see Figure 1. Hence, we can apply an appropriate numerical flux function similar to those known for local conservation laws to approximate the reduced problem with a fixed nonlocal term, i.e., the flux reduces to

    F(tn,xj+12,ρ)g(ρ)Vnj. (3.2)
    Figure 1.  Space discretization and approximation of the nonlocal term for η=3Δx with the densities used to calculated Vnj highlighted in gray.

    Now let us look more closely at the approximation of the nonlocal term and the requirements for a suitable numerical flux function.

    Definition 3.1 (Approximation of the nonlocal term). We approximate the nonlocal term (2.2) by

    V(tn,xj+12)v(xj+12+ηxj+12ωη(yx)ρΔx(tn,y)dy)=v(Nη1k=0γkρnj+k+1)=:Vnjwithγk=(k+1)ΔxkΔxωη(x)dxandNη:=η/Δx. (3.3)

    The weights γk must be computed exactly, e.g., by an appropriate quadrature rule.

    An illustration of the nonlocal term for η=3Δx can be seen in Figure 1. To ensure the positivity of the nonlocal term it is important that the weights are computed without any numerical error. For the reduced flux (3.2) Vnj needs to be considered as a constant and its dependence on ρnj+1,,ρnj+Nη will be neglected in the following. Then a suitable numerical flux function Fnj+12(a,b) must satisfy the following definition:

    Definition 3.2 (Numerical flux function). The numerical flux function Fnj+12 satisfies the following conditions:

    (i) The flux function can be written as

    Fnj+12(a,b)=G(a,b)Vnj

    with G only depending on g, ρmin, ρmax.

    (ii) Consistency:

    Fnj+12(ρ,ρ)=g(ρ)Vnj,or equivalently ,G(ρ,ρ)=g(ρ)ρ[ρmin,ρmax].

    (iii) The map (a,b)Fnj+12(a,b)=G(a,b)Vnj from [ρmin,ρmax]2 to R is nondecreasing with respect to a and nonincreasing with respect to b.

    (iv) Fnj+12(a,b) satisfies the following sort of weak Lipschitz continuity:

    |Fnj+12(a,b)Fnj+12(b,b)|L1|ab|and|Fnj+12(a,b)Fnj+12(a,a)|L2|ab| (3.4)

    for (a,b)[ρmin,ρmax]2 and every Vnj[v,v]. We will call L1>0 and L2>0 the Lipschitz constants.

    Remark 3.3. Note that the assumptions imply that Fnj+12 and G are bounded and that G satisfies a statement similar to Eq (3.4). The latter follows directly from Fnj+12 satisfying (3.4) for every Vnj[v,v] and the bound can be obtained by

    Fnj+12(a,b)=Fnj+12(a,b)Fnj+12(b,b)+Fnj+12(b,b)g(b)Vnj+L1|ab|gv+L1(ρmaxρmin).

    The assumptions on the flux function after approximating the nonlocal term are very similar to those for a monotone flux scheme of local conservation laws, cf. [17, Definition 5.6]. The only additional assumption is that we assume the numerical flux to be decomposable as in (i). Since the flux function of Eq (2.1) is decomposable into a nonlinear term and nonlocal term, i.e., F(t,x,ρ)=g(ρ)V(t,x), this is a rather natural assumption. The other assumptions (ii)–(iv) are also quite common. Assumption (ii) is necessary for the consistency with the flux of the modeling equation, while (iii) gives our scheme a monotone like behavior. Although the complete scheme is not monotone in every argument, as here the dependence of the nonlocal term on ρj+1,,ρj+Nη is not considered. Finally, (iv) is a slightly weaker form of the usual assumed local Lipschitz continuity. One could also assume that the flux function is locally Lipschitz continuous from R2 to R as typical for monotone flux schemes, cf. [17, Definition 5.6], but during the convergence analysis in Section 4 it will turn out that the weak Lipschitz continuity (3.4) is sufficient to guarantee convergence.

    Remark 3.4 (Generalization of the flux). The above strategy can be applied to other classes of nonlocal conservation laws that have a similar structure. For simplicity, we will perform the proofs only for the equation (2.1). We will discuss this further in the numerical examples in Section 5.2. Note that the Definition 3.2 also covers the case of Vnj<0, although this does not occur in the model (2.1).

    Let us present some examples for the equation (2.1) which are inspired by classical monotone schemes. In Section 4, we will see that Vnj0 holds due to Definition 3.1 and the Assumptions 2.1. Hence, we will already use Vnj0 to simplify the schemes in the following. Whenever we use this property, we denote it with ().

    (local) Lax-Friedrichs-type scheme: For a diffusion parameter α>0 and inspired by the classical Lax-Friedrichs scheme [33], we have

    Fnj+12(ρnj,ρnj+1)=Vnj2(g(ρnj)+g(ρnj+1)+α(ρnjρnj+1)). (3.5)

    To satisfy the Definition 3.2 αg needs to hold.

    Note that this scheme differs to the usual choice of the nonlocal Lax-Friedrichs-type scheme as used e.g. in [9]. Here, we consider g(ρnj+1)Vnj instead of g(ρnj+1)Vnj+1.

    Godunov-type scheme: Inspired by the Godunov scheme introduced in [24] we deduce the following flux function

    Fnj+12(ρnj,ρnj+1)={min[ρnj,ρnj+1]g(ρ)Vj,ρnjρnj+1,max[ρnj+1,ρnj]g(ρ)Vj,ρnj>ρnj+1,()=Vnj{min[ρnj,ρnj+1]g(ρ),ρnjρnj+1,max[ρnj+1,ρnj]g(ρ),ρnj>ρnj+1. (3.6)

    Engquist-Osher-type scheme: Further, we consider an Engquist-Osher-type scheme following the ideas of [16]:

    Fnj+12(ρnj,ρnj+1)=12(g(ρnj)Vnj+g(ρnj+1)Vnjρnj+1ρnj|g(ρ)Vnj|dρ)()=Vnj2(g(ρnj)+g(ρnj+1)ρnj+1ρnj|g(ρ)|dρ). (3.7)

    Note that in case of a strictly concave function g, i.e., g(ρ)>0 and g(c)=0 for c[ρmin,ρmax], the flux simplifies to

    Fnj+12(ρnj,ρnj+1)=Vnj(g(min{ρnj,c})+g(max{ρnj+1,c})g(c)),

    compare [23, p. 229].

    All these schemes are inspired by already well studied schemes for local conservation laws, in particular in the local case they all belong to the class of monotone schemes, see [17]. Hence, they also satisfy Definition 3.2. Note that the Godunov-type scheme has already been considered for similar problems in [10,12,21]. In contrast, the Lax-Friedrichs-type scheme is slightly different from the usual ones used in the literature, e.g., [5,9,27]. To the best of our knowledge this is the first time that an Engquist-Osher-type scheme is proposed for nonlocal conservation laws. Furthermore, an Upwind-type scheme does not in general satisfy Definition 3.2 for a nonlinear g, since it is not monotone. As in the local case it may converge to the wrong solution. Nevertheless, if g is linear, e.g., g(ρ)=ρ, also an Upwind-type scheme would converge to the correct solution. In particular, in this case the scheme coincides with the Godunov- and Engquist-Osher-type schemes.

    Remark 3.5 (Comparison to other general approaches in the literature). As aforementioned typically schemes tailored to each application indiviually are studied. In a recent work [27], the authors consider a more general approach for a class of numerical schemes. Nevertheless, they restrict themselves to a specific subcase of Eq (2.1), i.e., g(ρ)=ρ and v(ρ)=1ρ. The goal in [27] is to derive asymptotically compatible numerical schemes and hence they assume further restrictions on the initial data (strictly positive without negative jumps) and an upper bound on the nonlocal range η. Besides considering a more general flux function, our approach to derive a numerical scheme also differs since we first approximate the nonlocal term, leaving us with only one nonlocal term in the flux, while in [27] there may be an additional nonlocal term in the numerical flux function. Hence, our scheme might be a special case of their scheme. Nevertheless, some conditions for convergence are different, e.g., we assume Lipschitz type conditions on the numerical flux function in contrast to conditions on the partial derivatives. Some conditions are also similar, such as the consistency or the first order derivatives with respect to the local variables. For more details on the numerical scheme we refer to [27, Assumption 4].

    Remark 3.6 (Higher order extensions). The scheme (3.1) together with a numerical flux function satisfying Definition 3.2 is at most a first order approximation of the modeling equation (2.1), see Section 5 for numerical convergence rates. Nevertheless, using a semi-discrete approach and a more accurate approximation of the space, e.g., by a CWENO reconstruction procedure as in [20], together with an appropriate time stepping scheme and an accurate approximation of the nonlocal term, allows a higher order approximation of the solution. One can simply follow the same approach as in [20] with a numerical flux function that satisfies Definition 3.2.

    In this section we prove that the assumptions on the numerical fluxes in Definition 3.2 together with the appropriate approximation of the nonlocal term in Definition 3.1 and an appropriate Courant-Friedrichs-Levy (CFL) condition are sufficient to guarantee the convergence of the numerical scheme to the physically correct entropy solution. Thereby, we consider the following CFL condition:

    λ:=ΔtΔx1Gvγ0+L1+L2. (4.1)

    Now we are ready to present our main result:

    Theorem 4.1. Let the Assumptions 2.1 hold. Then a numerical scheme (3.1) with a numerical flux function fulfilling Definition 3.2 and the nonlocal term being approximated as in Definition 3.1 converges under the CFL condition (4.1) for (Δt,Δx)0 to the unique weak entropy solution of (2.1) in the sense of Definition 2.2.

    Moreover the sequence of approximate solutions fulfills the maximum principle for a given initial datum ρ0j, jZ

    infjZρ0jρnjsupjZρ0j, jZ, nN.

    The proof consists of proving several properties fulfilled by all schemes (3.1) satisfying the Definitions 3.1 and 3.2.

    We start by proving a discrete maximum principle. Due to the involved nonlocal term, the numerical fluxes of Definition 3.2 are not monotone in every argument, in particular for ρnj+k, k=2,,Nη1, see also [5,27]. Hence, we will have to rely on a different approach to show the maximum principle, similar to [8,18,21].

    Proposition 4.2. Under the Assumptions 2.1 and the CFL condition (4.1), for a given initial datum ρ0j, jZ with ρm=infjZρ0j and ρM=supjZρ0j, the scheme (3.1) satisfying the Definitions 3.1 and 3.2 fulfills

    ρmρnjρM, jZ, nN.

    Proof. We prove the claim by induction. For n=0 the claim is obvious, so we suppose

    ρmρnjρM, jZ

    holds for a fixed nN.

    Next, we consider

    Vnj1Vnj=v(Nη1k=0γkρnj+k)v(Nη1k=0γkρnj+k+1)=v(ξj)(Nη1k=1(γkγk1)ρnj+kγNη1ρnj+Nη+γ0ρnj),

    where ξj comes from the mean value theorem. Since γkγk1 holds, we get

    Vnj1Vnjv(ξj)(Nη1k=1(γkγk1)ρMγNη1ρM+γ0ρnj)vγ0(ρMρnj). (4.2)

    Then, by the assumptions (iii) and (iv) on the numerical flux in Definition 3.2 and Eq (4.2) we have

    Fnj12Fnj+12G(ρM,ρnj)Vnj1G(ρnj,ρM)Vnj=G(ρM,ρnj)(Vnj1Vnj)(G(ρnj,ρM)G(ρM,ρnj))VnjGvγ0(ρMρnj)Fnj+12(ρnj,ρM)±Fnj+12(ρM,ρM)+Fnj+12(ρM,ρnj)(Gvγ0+L1+L2)(ρMρnj).

    Hence, under the CFL condition (4.1) we obtain

    ρn+1jρM.

    Analogously to Eq (4.2), we obtain

    Vnj1Vnjvγ0(ρmρnj)

    and

    Fnj12Fnj+12(Gvγ0+L1+L2)(ρmρnj)

    to obtain under the CFL condition (4.1)

    ρn+1jρm.

    An immediate consequence of the maximum principle and in particular the positivity of the approximate solutions is that the discrete L1-norm is conserved.

    Corollary 4.3. Under the Assumptions 2.1 and the CFL condition (4.1) the approximate solutions computed by the numerical scheme (3.1) satisfying the Definitions 3.1 and 3.2 preserve the L1 norm, i.e.

    ΔxjZ|ρnj|=ρ0L1(R).

    Remark 4.4. Proposition 4.2 together with an approximation of the nonlocal term as in Definition 3.1 guarantees the following bounds on the nonlocal term:

    0v(ρM)Vnjv(ρm)v(0).

    Next, we want to derive bounded variation (BV) estimates in space and time on the approximate solutions. Therefore, we recall the definition of the piecewise constant function

    ρΔx(t,x):=nNjZρnjχ[tn,tn+1)×[xj12,xj+12)(t,x). (4.3)

    As for the analytical solution obtained in [9] we only recover estimates which increase exponentially in time.

    Proposition 4.5 (BV estimate in space). Let the Assumptions 2.1 and the CFL condition (4.1) hold. Let ρΔx be defined as in (4.3) with a scheme (3.1) satisfying the Definitions 3.1 and 3.2, then we obtain for any finite time tn>0 the following bound on the total variation in space

    TV(ρΔ(tn,))=jZ|ρnj+1ρnj|exp(tnωη(0)(v(2g+gρ)+2vgρ))TV(ρ0).

    Proof. We can rewrite the scheme as

    ρn+1j=ρnj+λ(G(ρnj1,ρnj)Vnj1G(ρnj,ρnj+1)Vnj)=ρnj+anj12(ρnj1ρnj)+bnj+12(ρnj+1ρnj)+λg(ρnj)(Vnj1Vnj)

    with

    anj12={λVnj1G(ρnj1,ρnj)G(ρnj,ρnj)ρnj1ρnj,if ρnj1ρnj,0,if ρnj1=ρnj, (4.4)
    bnj+12={λVnjG(ρnj,ρnj+1)G(ρnj,ρnj)ρnjρnj+1,if ρnj+1ρnj,0,if ρnj+1=ρnj, (4.5)

    with (due to the montonicity and weak Lipschitz continuity) 0anj12L1λ and 0anj12L2λ. Next we define

    Δnj=ρnj+1ρnj.

    Then, we obtain

    Δn+1j=(1anj+12bnj+12)Δnj+bnj+32Δnj+1+anj12Δnj1+λ(g(ρnj+1)(VnjVnj+1)g(ρnj)(Vnj1Vnj)).

    Adding a zero the last term can be written as

    g(ρnj+1)(VnjVnj+1)g(ρnj)(Vnj1Vnj)=g(ζj)Δnj(VnjVnj1)+g(ρnj)(Vnj+1+2VnjVnj1),

    with ζj coming form the mean value theorem such that g(ζj)=g(ρnj+1)g(ρnj). Now, we consider the difference in the velocities (with ξj also coming from the mean value theorem) and obtain

    VnjVnj1=v(ξj)Nη1k=0γkΔnj+kandVnj+1Vnj=v(ξj+1)Nη1k=0γkΔnj+k+1.

    This leads us to

    Vnj+1+2VnjVnj1=γ0v(ξj)ΔnjγNη1v(ξj+1)Δnj+Nη+Nη1k=1(γkv(ξj)γk1v(ξj+1))Δnj+k=γ0v(ξj)ΔnjγNη1v(ξj+1)Δnj+Nη+Nη1k=1(γkγk1)v(ξj)Δnj+k+Nη1k=1γk1(v(ξj)v(ξj+1))Δnj+k.

    As several terms are positive due to the CFL condition (4.1), we can put everything together by using also Eq (4.2) to obtain

    |Δn+1j|(1anj+12bnj+12+λγ0v(g+gρ))|Δnj|+bnj+32|Δnj+1|+anj12|Δnj1|+γNη1vgλ|Δnj+Nη|+Nη1k=0(γk1γk)vgλ|Δnj+k|+gλNη1k=1γk1|v(ξj)v(ξj+1)||Δnj+k|.

    Summing over jZ, rearranging the indices and using λγ0Δtωη(0) leaves us with

    jZ|Δn+1j|jZ(1+Δtωη(0)v(2g+gρ)+λvgNη1k=1γk1|ξjkξj+1k|)|Δnj|.

    We need to estimate |ξjξj+1|. As we used the mean value theorem we know that ξj is between

    Rnj:=Nη1k=0γkρnj+kandRnj+1.

    Hence, we obtain

    |ξjξj+1||ξjRnj+1|+|Rnj+1ξj+1||Rnj+1Rnj|+|Rnj+2Rnj+1|2γ0ρ. (4.6)

    This follows by similar calculations as done to obtain (4.2).

    Finally, this leaves us with

    jZ|Δn+1j|(1+Δtωη(0)(v(2g+gρ)+2vgρ))jZ|Δnj|.

    Therefore, we can conclude

    TV(ρΔx(tn+1,))=jZ|Δn+1j|(1+Δtωη(0)(v(2g+gρ)+2vgρ))tn+1/ΔtjZ|Δ0j|exp(tn+1ωη(0)(v(2ggρ)+2vgρ))jZ|Δ0j|exp(tn+1ωη(0)(v(2g+gρ)+2vgρ))TV(ρ0),

    where we used [17, Remark 5.4] for the last inequality.

    Remark 4.6. Due to the presence of g(ρ), we cannot rewrite the scheme directly into the form proposed in [1, Theorem A.1]. Nevertheless, our proof is similar to the one there and also to the one proposed in [17] for the local case.

    Remark 4.7. Note that the obtained estimate on the total variation is very similar to the one in [9, Proposition 3.2]. In fact, the main difference comes from the estimate (4.6), which gives the factor 2vgρ. In contrast, the factor 7/2vg is obtained in [9]. Hence, for ρ=ρ07/4, which is the case for most applications since ρ01 normally holds, our estimate provides a slightly smaller upper bound on the analytical solution.

    Now, we are able to provide an estimate on the continuity in time.

    Proposition 4.8 (Time continuity estimates). Let the Assumptions 2.1 and the CFL condition (4.1) hold. Let ρΔx be defined as in (4.3) with a scheme (3.1) satisfying the Definitions 3.1 and 3.2, then we obtain the following estimate

    ΔxjZ|ρn+1jρnj|Δt(Gv+L1+L2)jZ|ρnj+1ρnj|

    Proof. We observe

    ρn+1jρnj=λ(G(ρnj1,ρnj)Vnj1G(ρnj,ρnj+1)Vnj)=λ(G(ρnj1,ρnj)(Vnj1Vnj)+(G(ρnj1,ρnj)±G(ρnj,ρnj)G(ρnj,ρnj+1))Vnj).

    Hence, we obtain

    |ρn+1jρnj|λ(GvNη1k=0γk|ρnj+k+1ρnj+k|+L2|ρnj+1ρnj|+L1|ρnjρnj1|).

    Summing over jZ and rearranging the indices yields then the claim.

    From the two latter propositions we can obtain an estimate on the bounded variation in space and time:

    Corollary 4.9. Let the Assumptions 2.1 and the CFL condition (4.1) hold. Let ρΔx be defined as in (4.3) with a scheme (3.1) satisfying the Definitions 3.1 and 3.2, then we obtain for any nN

    n1m=0jZΔt|ρnj+1ρnj|+Δx|ρn+1jρnj|nΔt(1+Gv+L1+L2)exp(tnωη(0)(v(2g+gρ)+2vgρ))TV(ρ0).

    To prove the convergence against the entropy solution we need an equivalent discrete formulation. For a constant kR we denote

    Fkj+1/2(u,w)=Fnj+12(uk,wk)Fnj+12(uk,wk)

    with ab=max(a,b) and ab=min(a,b).

    Proposition 4.10. A numerical scheme (3.1) satisfying the Definitions 3.1 and 3.2, fulfills

    |ρn+1jk||ρnjk|+λ(Fkj+1/2(ρnj,ρnj+1)Fkj1/2(ρnj1,ρnj))+λ2sgn(ρn+1jk)g(k)(VnjVnj1)0.

    Proof. We define

    H(u,w,z)(u,w,z)=wλ(Fnj+12(w,z)Fnj12(u,w)),

    which is monotone in each variable due to our assumptions on the numerical flux in Definition 3.2. Now, we can follow [5,9,21] to obtain the desired claim. We omit the details.

    Finally, we are ready to state the proof of convergence for the considered class of numerical schemes.

    Proof of Theorem 4.1. Due to Proposition 4.2 and Corollary 4.9, we can apply Helly's theorem as presented in [17, Lemma 5.6] and conclude the existence of a subsequence of approximate solutions that converge to some ρ(LBV)(R+×R;R). In addition, thanks to Corollary 4.3, we have ρL1(R+×R;R). Using the discrete entropy inequality from Proposition 4.10 and Lax-Wendroff type arguments similar to [5] shows that the limit function ρ is a weak entropy solution of Eq (2.1) in the sense of Definition 2.2. The discrete maximum principle is given directly in Proposition 4.2

    In this section we complement the theoretical results with numerical examples and consider the numerical convergence rates of several numerical schemes satisfying Definition 3.2. In Section 5.1 we focus on a traffic flow model with Arrhenius look-ahead dynamics belonging to Eq (2.1). Section 5.2 considers our approach for a nonlocal sedimentation model which has a different convolution than Eq (2.2).

    We first consider a numerical example to compare the schemes derived above, namely the Lax-Friedrichs-type scheme (3.5), the Godunov-type scheme (3.6), and the Engquist-Osher-type scheme (3.7). We also compare the numerical schemes with the Lax-Friedrichs-type scheme of [9], which is slightly different from our numerical scheme (3.5). We choose the diffusion parameter α of [9] for a fair numerical comparison between the two Lax-Friedrichs-type schemes. Furthermore, the CFL condition is given as in [9], since this is the most restrictive one. We will use the following example:

    g(ρ)=ρ(1ρ),v(ρ)=exp(ρ),ωη(x)=2(ηx)/η2,η=0.1.

    Hence, we consider the Arrhenius-type look-ahead model with a linear decreasing function. In particular, in contrast to the case of the nonlocal LWR model, i.e., g(ρ)=ρ, all the schemes (3.5)–(3.7) differ here. We consider the discontinuous initial condition:

    ρ0(x)={0.8x[3/4,5/4],0else

    at the final time T=0.5. In Figure 2 we see the approximate solutions on the left and a zoom into the spatial domain on the right, together with a reference solution computed by the Godnuov-type scheme and Δx=0.0126. It can be seen that the Engquist-Osher and Godunov-type schemes are already closer to the reference solution, while the Lax-Friedrichs-type schemes have a higher numerical diffusion. However, the newly derived Lax-Friedrichs-scheme (3.5) is closer to the reference solution than the Lax-Friedrichs scheme of [9]. In addition, the zoom on the right side shows that the solutions of the two Lax-Friedrichs-type schemes, but also of the Engquist-Osher and Godunov-type schemes, differ.

    Figure 2.  Approximate solutions obtained by the different schemes from [9] and Eqs (3.5)–(3.7) with Δx=0.01 and a reference solution (ref. sol.), left, and a zoom into the spatial domain, right.

    Furthermore, we test the numerical convergence rate in the L1-norm in comparison to the reference solution, still computed by the Godnuov type scheme and Δx=0.0126. We choose Δx=0.012n for n=0,,5. Theoretically, we expect a convergence rate of one and Table 1 shows the corresponding errors. The numerical convergence rates suggest the expected order of one. We note that the errors of the Godunov-type scheme (3.6) are the smalles.

    Table 1.  L1-errors of the different schemes in comparison to a reference solution and the numerical convergence rates for Δx=0.012n.
    L1-errors convergence rate
    n LxF [9] LxF (3.5) Godunov (3.6) EO (3.7) LxF [9] LxF (3.5) Godunov (3.6) EO (3.7)
    0 0.0343 0.0248 0.0085 0.0085
    1 0.0178 0.0127 0.0026 0.0033 0.9482 0.9739 1.6770 1.3738
    2 0.0091 0.0064 0.0013 0.0016 0.9644 0.9792 1.0711 1.0578
    3 0.0049 0.0035 6.6881e-04 8.2489e-04 0.8882 0.8619 0.9133 0.9280
    4 0.0026 0.0019 3.4622e-04 4.2017e-04 0.9196 0.9004 0.9499 0.9730
    5 0.0014 0.0010 1.8495e-04 2.1174e-04 0.9203 0.8974 0.9045 0.9887

     | Show Table
    DownLoad: CSV

    In this part we want to show that the proposed strategy can be applied to other nonlocal problems and that it works well. However, we note that we will not provide a rigorous convergence proof as before. We will focus on a sedimentation model.

    But first, let us mention that the strategy was already successfully used to derive a Godunov-type scheme in [21] for another traffic flow model including a rigorous convergence proof. In addition, in [12] a Godunov scheme for a nonlocal Burgers equation is derived and numerically evaluated in a similar manner. In both these models the resulting Engquist-Osher scheme is the same as the Godunov-type scheme. Furthermore, in [6] a Hilliges-Weidlich-type scheme is derived for a specific nonlocal problem appearing for example in sedimentation. Due to the modeling equations, the nonlocal term is different from (2.2), but after approximating the nonlocal term, the numerical flux function of [6] satisfies Definition 3.2.

    Now, we want to apply the approach to the nonlocal sedimentation model presented in [4]. In this case we have

    g(ρ)=ρ(1ρ),v(ρ)=(1ρ)4

    but the nonlocal term is

    (ρωη)(x)=x+2ηx2ηρ(t,y)ωη(yx)dx,

    with a symmetric kernel. Hence, it does not belong to the class of Eq (2.1). Nevertheless, the well-posedness can be proved. For further details we refer to [4]. To construct numerical flux functions we apply our approach and approximate the nonlocal term by

    Vnj=v(Nη1k=Nηγkρnj+k+1)

    with Nη=2η/Δx and the weights γk are defined as before. As proved in [4], the solution inherits a maximum principle between [0,1]. Hence, as long as it is kept by the numerical schemes, the velocity of the nonlocal term remains positive. So we can apply the numerical flux functions (3.5)–(3.7). We repeat the numerical test of [4, Example 2] for the initial data:

    ρ0(x)={0.01x>0.2,0elseandρ0(x)={0.6x>0.2,0else.

    The kernel is set to

    ωη(x)=K(x/η)ηK(x)=38(1x24)for |x|<2 and K(x)=0 else.

    We compare our numerical schemes with the Lax-Friedrichs scheme used in [4]. For the latter one the convergence is proven. As in [4] we choose λ=0.2. Table 2 shows the error terms at T=3 for the first initial condition. The reference solution is computed using the Lax-Friedrichs scheme of [4] with Δx=0.229. The Engquist-Osher and Godunov type schemes obtain their expected order of convergence. We note that they coincide in this example. The numerical convergence rates for the Lax-Friedrichs type schemes are lower. Nevertheless, the error terms of all schemes suggest the convergence to the correct entropy solution. The solution with Δx=0.225 can be seen in Figure 3, left. The difference between the two Lax-Friedrichs type schemes is not visible.

    Table 2.  L1-errors of the different schemes in comparison to a reference solution and the numerical convergence rates for Δx=0.22n at T=3 for the first initial condition.
    L1-errors convergence rate
    n LxF [4] LxF (3.5) Godunov (3.6) EO (3.7) LxF [4] LxF (3.5) Godunov (3.6) EO (3.7)
    0 0.0134 0.0133 0.0055 0.0055
    1 0.0090 0.0089 0.0033 0.0033 0.5634 0.5715 0.7239 0.7239
    2 0.0063 0.0062 0.0022 0.0022 0.5248 0.5210 0.5588 0.5588
    3 0.0043 0.0043 0.0015 0.0015 0.5455 0.5469 0.5702 0.5702
    4 0.0029 0.0029 9.3067e-04 9.3067e-04 0.5725 0.5740 0.7010 0.7010
    5 0.0019 0.0019 4.5755e-04 4.5755e-04 0.5885 0.5864 1.0243 1.0243

     | Show Table
    DownLoad: CSV
    Figure 3.  Approximate solutions obtained by the different schemes from [4] and Eqs (3.5)–(3.7) with Δx=0.225 and a reference solution (ref. sol.) at T=3 for the first initial condition, left and at T=1 for the second initial condition, right.

    Table 3 presents the error terms for the second initial condition at T=1. The reference solution is computed using the Lax-Friedrichs scheme of [4] with Δx=0.2210. Again, the error terms indicate the convergence to the entropy solution. In this example, we see that the newly derived Lax-Friedrichs scheme (3.5) is closer to the reference solution than the scheme from [4]. Furthermore, the Engquist-Osher and Godunov type schemes differ. The numerical convergence rate of the newly derived Lax-Friedrichs scheme (3.5) seems to converge to one. The convergence rates of the Engquist-Osher and Godunov type schemes drop below one for Δx=0.225. This behavior occurs because these schemes have less numerical viscosity than the Lax-Friedrichs scheme of [4]. Hence, they appear to provide a more accurate solution than the reference solution, even for a much larger step size. This can be seen in Figure 3, right. It can also be seen that the newly derived Lax-Friedrichs scheme (3.5) outperforms the Lax-Friedrichs scheme of [4] in terms of accuracy.

    Table 3.  L1-errors of the different schemes in comparison to a reference solution and the numerical convergence rates for Δx=0.22n at T=1 for the second initial condition.
    L1-errors convergence rate
    n LxF [4] LxF (3.5) Godunov (3.6) EO (3.7) LxF [4] LxF (3.5) Godunov (3.6) EO (3.7)
    0 0.4479 0.2005 0.1164 0.1163
    1 0.3182 0.1429 0.0407 0.0423 0.4933 0.4891 1.5145 1.4597
    2 0.2118 0.0905 0.0333 0.0333 0.5876 0.6580 0.2915 0.3454
    3 0.1463 0.0583 0.0140 0.0140 0.5337 0.6344 1.2450 1.2453
    4 0.0997 0.0335 0.0060 0.0060 0.5534 0.8021 1.2192 1.2190
    5 0.0661 0.0168 0.0040 0.0036 0.5917 0.9973 0.5890 0.7582

     | Show Table
    DownLoad: CSV

    In this work we have presented a general approach to treat a specific class of nonlocal conservation laws. The nonlocal term is approximated by an appropriate quadrature rule, and then explicit conditions for a numerical flux function on the reduced problem are presented that guarantee convergence. These conditions share some similarities with the class of monotone schemes for local conservation laws. Numerical examples show that the approach can be applied to other classes of nonlocal conservation laws. Furthermore, other schemes from the literature use the same approach for different nonlocal conservation laws.

    Future work could include generalizing the concept to other nonlocal terms as well as nonlocal multidimensional balance laws. Another interesting topic would be the behavior of the schemes for η0. Depending on the modeling equation considered, a convergence to a local conservation law can be obtained on the analytic level. Similar to the work done in [27] it would be interesting to see, if the derived class of numerical schemes is asymptotically compatible, i.e. converges to the correct local solution for η0.

    J. F. is supported by the German Research Foundation (DFG) under grant HE 5386/18-1, 19-2, 22-1, 23-1. S. R. is supported by IIPE, Visakhapatnam, India, under the IRG grant number IIPE/DORD/IRG/001 and NBHM, DAE, India (Ref. No. 02011/46/2021 NBHM(R.P.)/R & D II/14874).

    The authors declare there is no conflict of interest.



    [1] E. Abreu, R. De la cruz, J. Juajibioy, W. Lambert, Lagrangian-eulerian approach for nonlocal conservation laws, J Dyn Differ Equ, (2022), 1–47.
    [2] A. Aggarwal, R. M. Colombo, P. Goatin, Nonlocal systems of conservation laws in several space dimensions, SIAM J. Numer. Anal., 53 (2015), 963–983. https://doi.org/10.1137/140975255 doi: 10.1137/140975255
    [3] P. Amorim, R. M. Colombo, A. Teixeira, On the numerical integration of scalar nonlocal conservation laws, ESAIM Math. Model. Numer. Anal., 49 (2015), 19–37. https://doi.org/10.1051/m2an/2014023 doi: 10.1051/m2an/2014023
    [4] F. Betancourt, R. Bürger, K. H. Karlsen, E. M. Tory, On nonlocal conservation laws modelling sedimentation, Nonlinearity, 24 (2011), 855–885. https://doi.org/10.1088/0951-7715/24/3/008 doi: 10.1088/0951-7715/24/3/008
    [5] S. Blandin, P. Goatin, Well-posedness of a conservation law with non-local flux arising in traffic flow modeling, Numer. Math., 132 (2016), 217–241. https://doi.org/10.1007/s00211-015-0717-6 doi: 10.1007/s00211-015-0717-6
    [6] R. Bürger, H. Contreras, L. Villada, A hilliges-weidlich-type scheme for a one-dimensional scalar conservation law with nonlocal flux, Netw. Heterog. Media., 18 (2023), 664–693. https://doi.org/10.3934/nhm.2023029 doi: 10.3934/nhm.2023029
    [7] C. Chalons, P. Goatin, L. M. Villada, High-order numerical schemes for one-dimensional nonlocal conservation laws, SIAM J. Sci. Comput., 40 (2018), A288–A305. https://doi.org/10.1137/16M110825X doi: 10.1137/16M110825X
    [8] F. A. Chiarello, J. Friedrich, P. Goatin, S. Göttlich, O. Kolb, A non-local traffic flow model for 1-to-1 junctions, Eur J. Appl. Math., 31 (2020), 1029–1049. https://doi.org/10.1017/S095679251900038X doi: 10.1017/S095679251900038X
    [9] F. A. Chiarello, P. Goatin, Global entropy weak solutions for general non-local traffic flow models with anisotropic kernel, ESAIM Math. Model. Numer. Anal., 52 (2018), 163–180. https://doi.org/10.1051/m2an/2017066 doi: 10.1051/m2an/2017066
    [10] F. A. Chiarello, P. Goatin, Non-local multi-class traffic flow models, Netw. Heterog. Media., 14 (2019), 371–387. https://doi.org/10.3934/nhm.2019015 doi: 10.3934/nhm.2019015
    [11] F. A. Chiarello, P. Goatin, L. M. Villada, Lagrangian-antidiffusive remap schemes for non-local multi-class traffic flow models, Comput. Appl. Math., 39 (2020), 1–22. https://doi.org/10.1007/s40314-019-0964-8 doi: 10.1007/s40314-019-0964-8
    [12] M. Colombo, G. Crippa, M. Graff, L. V. Spinolo, On the role of numerical viscosity in the study of the local limit of nonlocal conservation laws, Esaim Math Model Numer Anal, 55 (2021), 2705–2723. https://doi.org/10.1051/m2an/2021073 doi: 10.1051/m2an/2021073
    [13] R. M. Colombo, M. Garavello, M. Lécureux-Mercier, A class of nonlocal models for pedestrian traffic, Math. Models Methods Appl. Sci., 22 (2012), 1150023. https://doi.org/10.1142/S0218202511500230 doi: 10.1142/S0218202511500230
    [14] M. G. Crandall, A. Majda, Monotone difference approximations for scalar conservation laws, Math. Comp., 34 (1980), 1–21. https://doi.org/10.1090/S0025-5718-1980-0551288-3 doi: 10.1090/S0025-5718-1980-0551288-3
    [15] C. D'Apice, S. Göttlich, M. Herty, B. Piccoli, Modeling, simulation, and optimization of supply chains, Philadelphia: Society for Industrial and Applied Mathematics (SIAM), 2010.
    [16] B. Engquist, S. Osher, One-sided difference approximations for nonlinear conservation laws, Math Comput, 36 (1981), 321–351. https://doi.org/10.1090/S0025-5718-1981-0606500-X doi: 10.1090/S0025-5718-1981-0606500-X
    [17] R. Eymard, T. Gallouët, R. Herbin, Finite volume methods, Handb. Numer. Anal., 7 (2000), 713–1020. https://doi.org/10.1016/S1570-8659(00)07005-8
    [18] J. Friedrich, S. Göttlich, M. Osztfalk, Network models for nonlocal traffic flow, Esaim Math Model Numer Anal, 56 (2022), 213–235. https://doi.org/10.1051/m2an/2022002 doi: 10.1051/m2an/2022002
    [19] J. Friedrich, S. Göttlich, E. Rossi, Nonlocal approaches for multilane traffic models, Commun. Math. Sci., 19 (2021), 2291–2317. https://doi.org/10.4310/CMS.2021.v19.n8.a10 doi: 10.4310/CMS.2021.v19.n8.a10
    [20] J. Friedrich, O. Kolb, Maximum principle satisfying CWENO schemes for nonlocal conservation laws, SIAM J. Sci. Comput., 41 (2019), A973–A988. https://doi.org/10.1137/18M1175586 doi: 10.1137/18M1175586
    [21] J. Friedrich, O. Kolb, S. Göttlich, A Godunov type scheme for a class of LWR traffic flow models with non-local flux, Netw. Heterog. Media., 13 (2018), 531–547. https://doi.org/10.3934/nhm.2018024 doi: 10.3934/nhm.2018024
    [22] P. Goatin, S. Scialanga, Well-posedness and finite volume approximations of the lwr traffic flow model with non-local velocity, Netw. Heterog. Media., 11 (2016), 107–121. https://doi.org/10.3934/nhm.2016.11.107 doi: 10.3934/nhm.2016.11.107
    [23] E. Godlewski, P. A. Raviart, Numerical approximation of hyperbolic systems of conservation laws, New York: Springer-Verlag, 1996.
    [24] S. K. Godunov, A difference method for numerical calculation of discontinuous solutions of the equations of hydrodynamics, Mat. Sb. (N.S.), 47 (1959), 271–306.
    [25] S. Göttlich, S. Hoher, P. Schindler, V. Schleper, A. Verl, Modeling, simulation and validation of material flow on conveyor belts, Appl. Math. Model., 38 (2014), 3295–3313. https://doi.org/10.1016/j.apm.2013.11.039 doi: 10.1016/j.apm.2013.11.039
    [26] H. Holden, N. H. Risebro, Front Tracking for Hyperbolic Conservation Laws, Berlin: Springer, 2015.
    [27] K. Huang, Q. Du, Asymptotically compatibility of a class of numerical schemes for a nonlocal traffic flow model, arXiv: 2301.00803, [Preprint], (2023) [cited 2023 May 15 ]. Available from: https://doi.org/10.48550/arXiv.2301.00803
    [28] A. Keimer, L. Pflug, Existence, uniqueness and regularity results on nonlocal balance laws, J. Differ. Equ., 263 (2017), 4023–4069. https://doi.org/10.1016/j.jde.2017.05.015 doi: 10.1016/j.jde.2017.05.015
    [29] A. Keimer, L. Pflug, Nonlocal balance laws–an overview over recent results, Handbook of Numerical Analysis, Amsterdam: Elsevier, 2023.
    [30] A. Keimer, L. Pflug, M. Spinola, Existence, uniqueness and regularity of multi-dimensional nonlocal balance laws with damping, J. Math. Anal. Appl., 466 (2018), 18–55. https://doi.org/10.1016/j.jmaa.2018.05.013 doi: 10.1016/j.jmaa.2018.05.013
    [31] A. Keimer, L. Pflug, M. Spinola, Nonlocal scalar conservation laws on bounded domains and applications in traffic flow, SIAM J. Math. Anal., 50 (2018), 6271–6306. https://doi.org/10.1137/18M119817X doi: 10.1137/18M119817X
    [32] S. N. Kružkov, First order quasilinear equations with several independent variables, Mat. Sb. (N.S.), 81 (1970), 228–255. https://doi.org/10.1130/0016-7606(1970)81[255:DOMSTI]2.0.CO;2 doi: 10.1130/0016-7606(1970)81[255:DOMSTI]2.0.CO;2
    [33] P. D. Lax, Weak solutions of nonlinear hyperbolic equations and their numerical computation, Comm. Pure Appl. Math., 7 (1954), 159–193. https://doi.org/10.1002/cpa.3160070112 doi: 10.1002/cpa.3160070112
    [34] R. J. LeVeque, Numerical methods for conservation laws, Berlin: Springer, 1992.
    [35] R. J. LeVeque, Finite volume methods for hyperbolic problems, Cambridge: Cambridge university press, 2002.
    [36] E. Rossi, J. Weißen, P. Goatin, S. Göttlich, Well-posedness of a non-local model for material flow on conveyor belts, Esaim Math Model Numer Anal, 54 (2020), 679–704. https://doi.org/10.1051/m2an/2019062 doi: 10.1051/m2an/2019062
    [37] J. W. Thomas, Numerical partial differential equations: finite difference methods, Berlin: Springer Science & Business Media, 2013.
  • This article has been cited by:

    1. Jan Friedrich, Simone Göttlich, Alexander Keimer, Lukas Pflug, 2024, Chapter 30, 978-3-031-55263-2, 347, 10.1007/978-3-031-55264-9_30
    2. F. A. Chiarello, J. Friedrich, S. Göttlich, A non-local traffic flow model for 1-to-1 junctions with buffer, 2024, 19, 1556-1801, 405, 10.3934/nhm.2024018
    3. Kuang Huang, Qiang Du, Asymptotic Compatibility of a Class of Numerical Schemes for a Nonlocal Traffic Flow Model, 2024, 62, 0036-1429, 1119, 10.1137/23M154488X
    4. Veerappa Gowda G. D., Sudarshan Kumar Kenettinkara, Nikhil Manoj, Convergence of a second-order scheme for non-local conservation laws, 2023, 57, 2822-7840, 3439, 10.1051/m2an/2023080
    5. Eduardo Abreu, Richard De la cruz, Juan Juajibioy, Wanderson Lambert, Weak asymptotic analysis approach for first order scalar conservation laws with nonlocal flux, 2025, 85, 14681218, 104378, 10.1016/j.nonrwa.2025.104378
  • Reader Comments
  • © 2023 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(2149) PDF downloads(121) Cited by(5)

Figures and Tables

Figures(3)  /  Tables(3)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog