Research article

Finite-in-time flocking of the thermodynamic Cucker–Smale model

  • Received: 27 March 2024 Revised: 26 April 2024 Accepted: 07 May 2024 Published: 16 May 2024
  • We illustrate finite-in-time flocking in the thermodynamic Cucker–Smale (TCS) model. First, we extend the original TCS model to allow for a continuous vector field with a locally Lipschitz continuity. Then, within this system, we derive appropriate dissipative inequalities concerning the position-velocity-temperature using several preparatory estimates. Subsequently, based on initial data and system parameters, we formulate sufficient conditions to guarantee the desired finite-time flocking in each case where the communication weight conditions are divided into two scenarios: one with a positive lower bound and another with nonnegativity and monotonicity. Finally, we provide several numerical simulations and compare them with the analytical results.

    Citation: Hyunjin Ahn, Se Eun Noh. Finite-in-time flocking of the thermodynamic Cucker–Smale model[J]. Networks and Heterogeneous Media, 2024, 19(2): 526-546. doi: 10.3934/nhm.2024023

    Related Papers:

    [1] 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
    [2] 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
    [3] Maria Laura Delle Monache, Paola Goatin . Stability estimates for scalar conservation laws with moving flux constraints. Networks and Heterogeneous Media, 2017, 12(2): 245-258. doi: 10.3934/nhm.2017010
    [4] Felisia Angela Chiarello, Giuseppe Maria Coclite . Nonlocal scalar conservation laws with discontinuous flux. Networks and Heterogeneous Media, 2023, 18(1): 380-398. doi: 10.3934/nhm.2023015
    [5] 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
    [6] Markus Musch, Ulrik Skre Fjordholm, Nils Henrik Risebro . Well-posedness theory for nonlinear scalar conservation laws on networks. Networks and Heterogeneous Media, 2022, 17(1): 101-128. doi: 10.3934/nhm.2021025
    [7] Gabriella Bretti, Roberto Natalini, Benedetto Piccoli . Numerical approximations of a traffic flow model on networks. Networks and Heterogeneous Media, 2006, 1(1): 57-84. doi: 10.3934/nhm.2006.1.57
    [8] Christophe Chalons, Paola Goatin, Nicolas Seguin . General constrained conservation laws. Application to pedestrian flow modeling. Networks and Heterogeneous Media, 2013, 8(2): 433-463. doi: 10.3934/nhm.2013.8.433
    [9] 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
    [10] Martin Gugat, Alexander Keimer, Günter Leugering, Zhiqiang Wang . Analysis of a system of nonlocal conservation laws for multi-commodity flow on networks. Networks and Heterogeneous Media, 2015, 10(4): 749-785. doi: 10.3934/nhm.2015.10.749
  • We illustrate finite-in-time flocking in the thermodynamic Cucker–Smale (TCS) model. First, we extend the original TCS model to allow for a continuous vector field with a locally Lipschitz continuity. Then, within this system, we derive appropriate dissipative inequalities concerning the position-velocity-temperature using several preparatory estimates. Subsequently, based on initial data and system parameters, we formulate sufficient conditions to guarantee the desired finite-time flocking in each case where the communication weight conditions are divided into two scenarios: one with a positive lower bound and another with nonnegativity and monotonicity. Finally, we provide several numerical simulations and compare them with the analytical results.



    Delle Monache and Goatin developed in [20] a macroscopic model aiming at describing the situation in which a slow moving large vehicle – a bus for instance – reduces the road capacity and thus generates a moving bottleneck for the surrounding traffic flow. Their model is given by a Cauchy problem for Lightwill-Whitham-Richards scalar conservation law in one space dimension with local point constraint. The constraint is prescribed along the slow vehicle trajectory (y(t),t), the unknown y being coupled to the unknown ρ of the constrained LWR equation. Point constraints were introduced in [19,17] to account for localized in space phenomena that may occur at exits and which act as obstacles. The constraint in the model of [20] depends upon the slow vehicle speed ˙y, where its position y verifies the following ODE

    ˙y(t)=ω(ρ(y(t)+,t)). (A)

    Above, ρ=ρ(x,t)[0,R] is the traffic density and ω:[0,R]R+ is a nonincreasing Lipschitz continuous function which links the traffic density to the slow vehicle velocity. Delle Monache and Goatin proved an existence result for their model in [20] with a wave-front tracking approach in the BV framework. Adjustments to the result were recently brought by Liard and Piccoli in [28]. Despite the step forward made in [21], the uniqueness issue remained open for a time. Indeed, the appearance of the trace ρ(y(t)+,t) makes it fairly difficult to get a Lipschitz continuous dependency of the trajectory y=y(t) from the solution ρ=ρ(x,t). Nonetheless, a highly nontrivial uniqueness result was achieved by Liard and Piccoli in [27]. To describe the influence of a single vehicle on the traffic flow, the authors of [26] proposed a PDE-ODE coupled model without constraint on the flux for which they proposed in [9] two convergent schemes. In the present paper, we consider a modified model where the point constraint becomes nonlocal, making the velocity of the slow vehicle depend on the mean density evaluated in a small vicinity ahead the driver. More precisely, instead of A, we consider the relation

    ˙y(t)=ω(Rρ(x+y(t),t)μ(x)dx), (B)

    where μBV(R;R+) is a weight function used to average the density. From the mathematical point of view, this choice makes the study of the new model easier. Indeed, the authors of [3,4,5] put forward techniques for full well-posedness analysis of similar models with nonlocal point constraints. From the modeling point of view, considering B makes sense for several reasons outlined in Section 3.5.

    The paper is organized as follows. Sections 2 and 3 are devoted to the proof of the well-posedness of the model. In Section 4 we introduce the numerical finite volume scheme and prove its convergence. An important step of the reasoning is to prove a BV regularity for the approximate solutions. It serves both in the existence proof and it is central in the uniqueness argument. In that optic, the appendix is essential. Indeed, it is devoted to the proof of a BV regularity for entropy solutions to a large class of limited flux models. Let us stress that we highlight the interest of the BVloc discrete compactness technique of Towers [33] in the context of general discontinuous-flux problems. In the numerical section 5, first we perform numerical simulations to validate our model. Then we investigate both qualitatively and quantitatively the proximity between our model – in which we considered B – as δμ0+ and the model of [20] in which the authors considered A.

    Note that we find it convenient to study the problem in the bus frame, which means setting X=xy(t) in the model of Delle Monache and Goatin in [20]. Keeping in mind what we said above about the nonlocal constraint, the problem we consider takes the following form:

    {tρ+x(F(˙y(t),ρ))=0R×(0,T)ρ(x,0)=ρ0(x+y0)xRF(˙y(t),ρ)|x=0Q(˙y(t))t(0,T)˙y(t)=ω(Rρ(x,t)μ(x)dx)t(0,T)y(0)=y0. (1)

    Above, ρ=ρ(x,t) denotes the traffic density, of which maximum attainable value is R>0, and

    F(˙y(t),ρ)=f(ρ)˙y(t)ρ

    denotes the normal flux through the curve x=y(t). We assume that the flux function f:[0,R]R is Lipschitz continuous and bell-shaped, which are commonly used assumptions in traffic dynamics:

    f(ρ)0, f(0)=f(R)=0, ! ¯ρ(0,R), f(ρ)(¯ρρ)>0 for a.e. ρ(0,R). (2)

    In [20], the authors chose the function Q(s)=α×(1s2)2 to prescribe the maximal flow allowed through a bottleneck located at x=0. The parameter α(0,1) was giving the reduction rate of the road capacity due to the presence of the slow vehicle. We use the s variable to stress that the value of the constraint is a function of the speed of the slow vehicle. In the sequel the s variable will refer to quantities related to the slow vehicle velocity. Regarding the function Q, we can allow for more general choices. Specifically,

    Q:[0,ωL]R+

    can be any Lipschitz continuous function. It is a well known fact that in general, the total variation of an entropy solution to a constraint Cauchy problem may increase (see [17,Section 2] for an example). However, this increase can be controlled if the constraint level does not reach the maximum level. A mild assumption on Q – see Assumption (17) below – will guarantee availability of BV bounds, provided we suppose that

    ρ0L1(R;[0,R])BV(R).

    Throughout the paper, we denote by

    Φ(a,b)=sign(ab)(f(a)f(b))andΦ˙y(t)(a,b)=Φ(a,b)˙y(t)|ab|

    the entropy fluxes associated with the Kružkov entropy ρ|ρκ|, for all κ[0,R], see [25]. Following [20,17,6,15], we give the following definition of solution for Problem (1).

    Definition 2.1. A couple (ρ,y) with ρL(R×(0,T)) and yW1,((0,T)) is an admissible weak solution to (1) if

    (ⅰ) the following regularity is fulfilled:

    ρC([0,T];L1loc(R)); (3)

    (ⅱ) for all test functions φCc(R×R+),φ0 and κ[0,R], the following entropy inequalities are verified for all 0τ<τT:

    ττR|ρκ|tφ+Φ˙y(t)(ρ,κ)xφdxdt+R|ρ(x,τ)κ|φ(x,τ)dxR|ρ(x,τ)κ|φ(x,τ)dx+2ττR˙y(t)(κ,q(t))φ(0,t)dt0, (4)

    where

    R˙y(t)(κ,q(t))=F(˙y(t),κ)min{F(˙y(t),κ),q(t)}andq(t)=Q(˙y(t));

    (ⅲ) for all test functions ψC([0,T]),ψ0 and some given φCc(R) which verifies φ(0)=1, the following weak constraint inequalities are verified for all 0τ<τT:

    ττR+ρt(φψ)+F(˙y(t),ρ)x(φψ)dxdtR+ρ(x,τ)φ(x)ψ(τ)dx+R+ρ(x,τ)φ(x)ψ(τ)dxττq(t)ψ(t)dt; (5)

    (ⅳ) the following weak ODE formulation is verified for all t[0,T]:

    y(t)=y0+t0ω(Rρ(x,s)μ(x)dx)ds. (6)

    Definition 2.2. We will call BV-regular solution any admissible weak solution (ρ,y) to the Problem (1) which also verifies

    ρL((0,T);BV(R)).

    Remark 1. It is more usual to formulate (4) with φCc(R×[0,T)), τ=0 and τ=T. The equivalence between the two formulations is due to the regularity (3).

    Remark 2. As it happens, the time-continuity regularity (3) is actually a consequence of inequalities (4). Indeed, we will use the result [12,Theorem 1.2] which states that if Ω is an open subset of R and if for all test functions φCc(Ω×[0,T)),φ0 and κ[0,R], ρ satisfies the following entropy inequalities:

    T0Ω|ρκ|tφ+Φ˙y(t)(ρ,κ)xφdxdt+Ω|ρ0(x)κ|φ(x,0)dx0,

    then ρC([0,T];L1loc(Ω)). Moreover, since ρ is bounded and ¯ΩΩ has a Lebesgue measure 0, ρC([0,T];L1loc(¯Ω)). We will use this remark several times in the sequel of the paper, with Ω=R.

    Remark 3. Any admissible weak solution (ρ,y) to Problem (1) is also a distributional solution to the conservation law in (1). Therefore, inequalities (5) imply the following ones for all 0τ<τT:

    ττRρt(φψ)+F(˙y(t),ρ)x(φψ)dxdt+Rρ(x,τ)φ(x)ψ(τ)dxRρ(x,τ)φ(x)ψ(τ)dxττq(t)ψ(t)dt,

    where φ and ψ are such as described in Definition 2.1 (ⅲ).

    The interest of weak formulations (5)-(6) for the flux constraint and for the ODE governing the slow vehicle lies in their stability with respect to ρ. Formulation (4) – (6) is well suited for passage to the limit of a.e. convergent sequences of exact or approximate solutions.

    In this section, we prove stability with respect to the initial data and uniqueness for BV-regular solutions to Problem (1). We start with the

    Lemma 2.3. If (ρ,y) is an admissible weak solution to Problem (1), then ˙yW1,((0,T)). In particular, ˙yBV([0,T]).

    Proof. Denote for all t[0,T],

    s(t)=ω(Rρ(x,t)μ(x)dx).

    Since μL1(R)L(R) and ρC([0,T];L1loc(R)), s is continuous on [0,T]. By definition, y satisfies the weak ODE formulation (6). Consequently, for a.e. t(0,T), ˙y(t)=s(t). We are going to prove that s is Lipschitz continuous on [0,T], which will ensure that ˙yW1,((0,T)). Since μBV(R), there exists a sequence (μn)nNBV(R)Cc(R) such that:

    μnμL1 n+ 0andTV(μn) n+ TV(μ).

    Introduce for all nN and t[0,T], the function

    ξn(t)=Rρ(x,t)μn(x)dx.

    Fix ψCc((0,T)). Since ρ is a distributional solution to the conservation law in (1), we have for all nN,

    T0ξn(t)˙ψ(t)dt=T0Rρt(ψμn)dxdt=T0RF(˙y(t),ρ)x(ψμn)dxdt=T0(RF(˙y(t),ρ)μn(x)dx)ψ(t)dt,

    which means that for all nN, ξn is differentiable in the weak sense, and that for a.e. t(0,T),

    ˙ξn(t)=RF(˙y(t),ρ)μn(x)dx.

    In particular, since both the sequences (μnL1)n and (TV(μn))n are bounded – say by C>0 – we also have for all nN,

    ξnLRCand˙ξnLC(fL+ωLR).

    Therefore, the sequence (ξn)n is bounded in W1,((0,T)). Now, for all t,τ[0,T] and nN, triangle inequality yields:

    |s(t)s(τ)|2ωLRμnμL1+ωL|R(ρ(x,t)ρ(x,τ))μn(x)dx|=2ωLRμnμL1+ωL|ξn(t)ξn(τ)|2ωLRμnμL1+CωL(fL+ωLR)K|tτ|.

    Letting n+, we get that for all t,τ[0,T], |s(t)s(τ)|K|tτ|, which proves that s is Lipschitz continuous on [0,T]. The proof of the statement is completed.

    Before stating the uniqueness result, we make the following additional assumption:

    s[0,ωL], argmaxρ[0,R] F(s,ρ)>0. (7)

    This ensures that for all s[0,ωL], the function F(s,) verifies the bell-shaped assumptions (53). For example, when considering the flux f(ρ)=ρ(Rρ), (7) reduces to ωL<R, which only means that the maximum velocity of the slow vehicle is lesser than the maximum velocity of the cars.

    Theorem 2.4. Suppose that f satisfies (2) and (7). Fix ρ10,ρ20L1(R;[0,R])BV(R) and y10,y20R. We denote by (ρ1,y1) a BV-regular solution to Problem (1) corresponding to initial data (ρ10,y10), and by (ρ2,y2) an admissible weak solution with initial data (ρ20,y20). Then there exist constants α,β,γ>0 such that

    for a.e. t(0,T), ρ1(t)ρ2(t)L1(|y10y20|TV(ρ10)+ρ10ρ20L1)exp(αt) (8)

    and

    t[0,T], |y1(t)y2(t)||y10y20|+(β|y10y20|+γρ10ρ20L1)(exp(αt)1). (9)

    In particular, Problem (1) admits at most one BV-regular solution.

    Proof. Since (ρ1,y1) is a BV-regular solution to Problem (1), there exists C0 such that

    t[0,T], TV(ρ1(t))C.

    Lemma 2.3 ensures that ˙y1,˙y2BV([0,T];R+). We can use result (54) to obtain that for a.e. t(0,T),

    ρ1(t)ρ2(t)L1|y10y20|TV(ρ10)+ρ10ρ20L1+(2QL+2R+C)t0|˙y1(s)˙y2(s)|ds. (10)

    Moreover, since for a.e. t(0,T),

    |˙y1(t)˙y2(t)|ωLμLρ1(t)ρ2(t)L1,

    Gronwall's lemma yields (8) with α=(2QL+2R+C)ωLμL. Then for all t[0,T],

    |y1(t)y2(t)||y10y20|+t0|˙y1(s)˙y2(s)|ds|y10y20|+ωLμLt0ρ1(s)ρ2(s)L1ds|y10y20|+(β|y10y20|+γρ10ρ20L1)(exp(αt)1),

    where

    β=TV(ρ10)2QL+2R+Candγ=12QL+2R+C.

    The uniqueness of a BV-regular solution is then clear.

    Remark 4. Up to inequality (10), our proof was very much following the one of [21,Theorem 2.1]. However, the authors of [21] faced an issue to derive a Lipschitz stability estimate between the car densities and the slow vehicle velocities starting from

    |ω(ρ1(0+,t))ω(ρ2(0+,t))|.

    For us, due to the nonlocality of our problem, it was straightforward to obtain the bound

    |ω(Rρ1(x,t)μ(x)dx)ω(Rρ2(x,t)μ(x)dx)|ωLμLρ1(t)ρ2(t)L1.

    Remark 5. A noteworthy consequence of Theorem 2.4 is that existence of a BV-regular solution will ensure uniqueness of an admissible weak one.

    In [20], to prove existence for their problem, the authors took a wave-front tracking approach. We choose here to use a time-splitting technique. The main advantage of this technique is that it relies on a ready-to-use theory. More precisely, at each time step, we will deal with exact solutions to a conservation law with a flux constraint, which have now become standard, see [17,6,15].

    Fix ρ0L1(R;[0,R]) and y0R. Let δ>0 be a time step, NN such that T[Nδ,(N+1)δ) and denote for all n{0,,N+1}, tn=nδ. We initialize with

    tR, ρ0(t)=ρ0(+y0)andt[0,T], y0(t)=y0.

    Fix n{1,,N+1}. First, we define for all t(tn1,tn],

    σn(t)=ω(Rρn1(x,tδ)μ(x)dx), sn=σn(tn) and qn=Q(sn).

    Since both qn and ρn1(,tn1) are bounded, [6,Theorem 2.11] ensures the existence and uniqueness of a solution ρnL(R×[tn1,tn]) to

    {tρ+x(F(sn,ρ))=0R×(tn1,tn)ρ(x,tn1)=ρn1(x,tn1)xRF(sn,ρ)|x=0qnt(tn1,tn),

    in the sense that ρn satisfies entropy/constraint inequalities analogous to (4)-(5) with suitable flux/constraint function and initial data, see Definition A.1. Taking also into account Remark 2, ρnC([tn1,tn];L1loc(R)). We then define the following functions:

     ρδ(t)=ρ01R(t)+N+1n=1ρn(t)1(tn1,tn](t) σδ(t),qδ(t),sδ(t)=σn(t),qn,snif t(tn1,tn] yδ(t)=y0+t0σδ(u)du.

    First, let us prove that (ρδ,yδ) solves an approximate version of Problem (1).

    Proposition 1. The couple (ρδ,yδ) is an admissible weak solution to

    {tρδ+x(F(sδ(t),ρδ))=0R×(0,T)ρδ(x,0)=ρ0(x+y0)xRF(sδ(t),ρδ)|x=0qδ(t)t(0,T)˙yδ(t)=ω(Rρδ(x,tδ)μ(x)dx)t(0,T)yδ(0)=y0, (11)

    in the sense that ρδC([0,T];L1loc(R)) and satisfies entropy/constraint inequalities analogous to (4)-(5) with flux F(sδ(),), constraint qδ, and initial data ρ0(+y0); and yδ satisfies, instead of (6), the following weak ODE formulation:

    t[0,T], yδ(t)=y0+t0ω(Rρδ(x,sδ)μ(x)dx)ds.

    Proof. By construction, for all n{1,,N+1}, ρnC([tn1,tn];L1loc(R)). Combining this with the "stop-and-restart" conditions ρn(,tn1)=ρn1(,tn1), one ensures that ρδC([0,T];L1loc(R)). Let t[0,T] and n{1,,N+1} such that t[tn1,tn). Then,

    yδ(t)y0=n1k=1tktk1σk(s)ds+ttn1σn(s)ds=n1k=1tktk1ω(Rρk1(x,sδ)ρδ(x,sδ)μ(x)dx)ds+ttn1ω(Rρn1(x,sδ)ρδ(x,sδ)μ(x)dx)ds=t0ω(Rρδ(x,sδ)μ(x)dx)ds, (12)

    which proves that ˙yδ solves the ODE in (11) in the weak sense. Fix now φCc(R×R+),φ0 and κ[0,R]. By construction of the sequence ((ρk,yk))k, we have for all n,m{0,,N+1},

    tmtnR|ρδκ|tφ+Φsδ(t)(ρδ,κ)xφdxdt=mk=n+1tktk1R|ρkκ|tφ+Φsk(ρk,κ)xφdxdtmk=n+1{R|ρk(x,tk)κ|φ(x,tk)dxR|ρk(x,tk1)ρk1(x,tk1)κ|φ(x,tk1)dx2tktk1Rsk(κ,qk)φ(0,t)dt}=R|ρδ(x,tm)κ|φ(x,tm)dxR|ρδ(x,tn)κ|φ(x,tn)dx2tmtnRsδ(t)(κ,qδ(t))φ(0,t)dt.

    It is then straightforward to prove that for all 0τ<τT,

    ττR|ρδκ|tφ+Φsδ(t)(ρδ,κ)xφ dxdt+R|ρδ(x,τ)κ|φ(x,τ)dxR|ρδ(x,τ)κ|φ(x,τ)dx+2ττRsδ(t)(κ,qδ(t))φ(0,t)dt0. (13)

    Proving that ρδ satisfies constraint inequalities is very similar so we omit the details. One has to start from

    ττR+ρδt(φψ)+F(sδ(t),ρδ)x(φψ) dxdt

    and make use once again of the construction of the sequence ((ρk,yk))k to obtain

    ττR+ρδt(φψ)+F(sδ(t),ρδ)x(φψ) dxdtR+ρδ(x,τ)φ(x)ψ(τ)dx+R+ρδ(x,τ)φ(x)ψ(τ)dxττqδ(t)ψ(t)dt. (14)

    This concludes the proof.

    Remark 6. Remark that we have for all δ>0,

    σδLωLandyδL|y0|+TωL.

    This means that the sequence (yδ)δ is bounded in W1,((0,T)). Then the compact embedding of W1,((0,T)) in C([0,T]) yields a subsequence of (yδ)δ, which we do not relabel, which converges uniformly on [0,T] to some yC([0,T]).

    At this point, we propose two ways to obtain compactness for the sequence (ρδ)δ, which will lead to two existence results.

    Theorem 3.1. Fix ρ0L1(R;[0;R]) and y0R. Suppose that f is Lipschitz continuous, satisfies (2)-(7) and the following nondegeneracy assumption

    for a.e. s(0,ωL), mes{ρ[0,R] | f(ρ)s=0}=0. (15)

    Then Problem (1) admits at least one admissible weak solution.

    Proof. Condition (15) combined with the obvious uniform L bound

    δ>0, (x,t)R×[0,T], ρδ(x,t)[0,R],

    and the results proved by Panov in [30,31] ensure the existence of a subsequence – which we do not relabel – that converges in L1loc(R×(0,T)) to some ρL1loc(R×(0,T)); and a further extraction yields the almost everywhere convergence on R×(0,T) and also the fact that ρL(R×(0,T);[0,R]). We now show that the couple (ρ,y) constructed above is an admissible weak solution to (1) in the sense of Definition 2.1.

    For all δ>0 and t[0,T],

    yδ(t)y0=t0ω(Rρδ(x,sδ)μ(x)dx)ds=tδδω(Rρδ(x,s)μ(x)dx)ds=t0ω(Rρδ(x,s)μ(x)dx)ds+(0δttδ)ω(Rρδ(x,s)μ(x)dx)ds.

    The last term vanishes as δ0 since ω is bounded. Then, Lebesgue theorem combined with the continuity of ω gives, for all t[0,T],

    yδ(t) δ0 y0+t0ω(Rρ(x,s)μ(x)dx)ds.

    This last quantity is also equal to y(t) due to the uniform convergence of (yδ)δ to y. This proves that y verifies (6). Now, we aim at passing to the limit in (13) and (14). With this in mind, we prove the a.e. convergence of the sequence (σδ)δ towards ˙y. Since μBV(R), there exists a sequence of smooth functions (μn)nNBV(R)Cc(R) such that:

    μnμL1 n+ 0andTV(μn) n+ TV(μ).

    Introduce for every δ>0 and nN, the function

    ξnδ(t)=Rρδ(x,t)μn(x)dx.

    Since for all δ>0, ρδ is a distributional solution to the conservation law in (11), one can show – following the proof of Lemma 2.3 for instance – that for every nN, ξnδW1,((0,T)), and that for a.e. t(0,T),

    ˙ξnδ(t)=RF(sδ(t),ρδ)μn(x)dx.

    Moreover, since both the sequences (μnL1)n and (TV(μn))n are bounded, it is clear that (ξnδ)δ,n is uniformly bounded in W1,((0,T)), therefore so is (ω(ξnδ))δ,n. Consequently, for all nN, δ>0 and almost every t(0,T), triangle inequality yields:

    |σδ(t)ω(Rρ(x,t)μ(x)dx)|2ωLRμnμL1+δ supnN ω(ξnδ)W1,+ωL|R(ρδ(x,t)ρ(x,t))μ(x)dx|δ0n+0,

    which proves that (σδ)δ converges a.e. on (0,T) to ˙y. To prove the time-continuity regularity, we first apply inequality (13) with τ=0, τ=T (which is licit since ρδ is continuous in time), φCc(R×[0,T)),φ0 and κ[0,R]:

    T0R|ρδκ|tφ+Φσδ(t)(ρδ,κ)xφ dxdt+R|ρ0(x+y0)κ|φ(x,0)dx0.

    Then, we let δ0 to get

    T0R|ρκ|tφ+Φ˙y(t)(ρ,κ)xφdxdt+R|ρ0(x+y0)κ|φ(x,0)dx0.

    Consequently, ρC([0,T];L1loc(R)), see Remark 2. Finally, the a.e. convergences of (σδ)δ and (ρδ)δ to ˙y and ρ, respectively, are enough to pass to the limit in (13). This ensures that for all test functions φCc(R×R+),φ0 and κ[0,R], the following inequalities hold for a.e. 0τ<τT:

    ττR|ρκ|tφ+Φ˙y(t)(ρ,κ)xφ dxdt+R|ρ(x,τ)κ|φ(x,τ)dxR|ρ(x,τ)κ|φ(x,τ)dx+2ττR˙y(t)(κ,q(t))φ(0,t)dt0.

    Observe that the expression in the left-hand side of the previous inequality is a continuous function of (τ,τ) which is almost everywhere greater than the continuous function 0. By continuity, this expression is everywhere greater than 0, which proves that ρ satisfies the entropy inequalities (4). Using similar arguments, we show that ρ satisfies the constraint inequalities (5). This proves the couple (ρ,y) is an admissible weak solution to Problem (1), and this concludes the proof.

    In this section, we proved an existence result for L initial data, but we have no guarantee of uniqueness since a priori we have no information regarding the BV regularity of such solutions.

    Assumption (15) ensures the compactness for sequences of entropy solutions to conservation laws with flux function F. However, it prevents us from using flux functions with linear parts – which corresponds to constant traffic velocity for small densities – whereas such fundamental diagrams are often used in traffic modeling. The results of the next section will extend to this interesting case, under the extra BV assumption on the data.

    To obtain compactness for (ρδ)δ, an alternative to the setting of Section 3.2 is to derive uniform BV bounds.

    Theorem 3.2. Fix ρ0L1(R;[0,R])BV(R) and y0R. Suppose that f satisfies (2) and (7). Suppose also that

    s[0,ωL], F(s,)C1([0,R]{¯ρs}), (16)

    where ¯ρs=argmaxρ[0,R] F(s,ρ). Finally assume that Q satisfies the condition

    ε>0, s[0,ωL], Q(s)maxρ[0,R] F(s,ρ)ε. (17)

    Then Problem (1) admits a unique admissible weak solution, which is also BV-regular.

    Proof. Fix δ>0. Recall that (ρδ,yδ) is an admissible weak solution to (11). In particular, ρδ is an admissible weak solution to the constrained conservation law in (11), in the sense of Definition A.1. It is clear from the splitting construction that for a.e. t(0,T),

    σδ(t)=ω(Rρδ(x,tδ)μ(x)dx).

    Following the steps of the proof of Lemma 2.3, we can show that for all δ>0, σδBV([0,T];R+). Even more than that, by doing so we show that the sequence (TV(σδ))δ is bounded. Therefore, the sequence (TV(sδ))δ is bounded as well. Moreover, since Q verifies (17), all the hypotheses of Corollary 2 are fulfilled. Combining this with Remark 15, we get the existence of a constant Cε=Cε(sFL) such that for all t[0,T],

    TV(ρδ(t))TV(ρ0)+4R+Cε(TV(qδ)+TV(sδ))TV(ρ0)+4R+Cε(1+QL)TV(sδ). (18)

    Consequently for all t[0,T], the sequence (ρδ(t))δ is bounded in BV(R). A classical analysis argument – see [24,Theorem A.8] – ensures the existence of ρC([0,T];L1loc(R)) such that

    t[0,T], ρδ(t) δ0 ρ(t) in L1loc(R).

    With this convergence, we can follow the proof of Theorem 3.1 to show that (ρ,y) is an admissible weak solution to (1). Then, when passing to the limit in (18), the lower semi-continuity of the BV semi-norm ensures that (ρ,y) is also BV-regular. By Remark 5, it ensures uniqueness and concludes the proof.

    To end this section, we now study the stability of Problem (1) with respect to the weight function μ. More precisely, let (μ)BV(R;R+) be a sequence of weight functions that converges to μ in the weak L1 sense:

    gL(R), Rg(x)μ(x)dx + Rg(x)μ(x)dx. (19)

    Let (y0)R be a sequence of real numbers that converges to some y0 and let (ρ0)L1(R;[0,R]) be a sequence of initial data that converges to ρ0 in the strong L1 sense. We suppose that the flux function f satisfies Assumptions (2)-(7)-(15). Theorem 3.1 allows us to define or all N, the couple (ρ,y) as an admissible weak solution to the problem

    {tρ+x(F(˙y(t),ρ))=0R×(0,T)ρ(x,0)=ρ0(x+y0)xRF(˙y(t),ρ)|x=0Q(˙y(t))t(0,T)˙y(t)=ω(Rρ(x,t)μ(x)dx)t(0,T)y(0)=y0.

    Remark 7. Using the same arguments as in Remark 6 and as in the proof of Theorem 3.1, we get that up to the extraction of a subsequence, (y) converges uniformly on [0,T] to some yC([0,T]) and (ρ) converges a.e. on R×(0,T) to some ρL(R×(0,T)).

    Theorem 3.3. The couple (ρ,y) constructed above is an admissible weak solution to Problem (1).

    Proof. The sequence (μ) converges in the weak L1 sense and is bounded in L1(R); by the Dunford-Pettis theorem, this sequence is equi-integrable:

    ε>0, α>0, AB(R), mes(A)<α N, Aμ(x)dxε (20)

    and

    ε>0, X>0, N, |x|Xμ(x)dxε. (21)

    Fix t(0,T) and ε>0. Fix α,X>0 given by (20) and (21). Egoroff theorem yields the existence of a measurable subset Et[X,X] such that

    mes([X,X]Et)<α and ρ(,t)ρ(,t) uniformly on Et.

    For a sufficiently large N,

    |Rρ(x,t)μ(x)dxRρ(x,t)μ(x)dx||x|X|ρρ|μdx+|Et(ρρ)μdx|+|[X,X]Et(ρρ)μdx|+|RρμdxRρμdx|Rε+ρρL(Et)Etμ(x)dx+R[X,X]Etμ(x)dx+ε2(R+1)ε,

    which proves that for a.e. t(0,T),

    Rρ(x,t)μ(x)dx + Rρ(x,t)μ(x)dx. (22)

    We get that y verifies the weak ODE formulation (6) by passing to the limit in

    y(t)=y0+t0ω(Rρ(x,s)μ(x)dx)ds.

    By definition, for all N, the couple (ρ,y) satisfies the analogue of entropy/constraint inequalities (4)-(5) with suitable flux/constraint functions. Applying these inequalities with τ=0, τ=T, φCc(R×[0,T)),φ0 and κ[0,R], we get

    T0R|ρκ|tφ+Φ˙y(t)(ρ,κ)xφdxdt+R|ρ0(x+y0)κ|φ(x,0)dx0.

    The continuity of ω and the convergence (22) ensure that (˙y) converges a.e. to ˙y. This combined with the a.e. convergence of (ρ) to ρ and Riesz-Frechet-Kolmogorov theorem – (ρ0) being strongly compact in L1(R) – is enough to show that when letting + in the inequality above, we get, up to the extraction of a subsequence, that

    T0R|ρκ|tφ+Φ˙y(t)(ρ,κ)xφdxdt+R|ρ0(x+y0)κ|φ(x,0)dx0.

    Consequently ρC([0,T];L1loc(R)), see Remark 2. Finally, the combined a.e. convergences of (˙y) and (ρ) to ˙y and ρ, respectively, guarantee that (ρ,y) verifies inequalities (4)-(5) for almost every 0τ<τT. The same continuity argument we used in the proof Theorem 3.1 holds here to ensure that (ρ,y) actually satisfies the inequalities for all 0τ<τT. This concludes the proof of our stability claim.

    The last section concludes the theoretical analysis of Problem (1). The nonlocality in space of the constraint delivers an easy proof of stability with respect to the initial data in the BV framework. Although a proof of existence using a fixed point theorem was possible (cf. [5]), we chose to propose a proof based on a time-splitting technique. The stability with respect to μ is a noteworthy feature, which shows a certain sturdiness of the model. However, the case we had in mind – namely μδ0+ – is not reachable with the assumptions we used to prove the stability, especially (19). We will explore this singular limit numerically, after having built a robust convergent numerical scheme for Problem (1). Let us also underline that unlike in [27,28] where the authors required a particular form for the function ω to prove well-posedness for their model, our result holds as long as ω is Lipschitz continuous.

    As evoked earlier, the nonlocality in space of the constraint makes the mathematical study of the model easier. But in the modeling point of view, this choice also makes sense for several reasons. First of all, one can think that the velocity ˙y of the slow moving vehicle – unlike its acceleration – is a rather continuous value. Even if the driver of the slow vehicle suddenly applies the brakes, the vehicle will not decelerate instantaneously. Note that the LWR model allows for discontinuous averaged velocity of the agents, however while modeling the slow vehicle we are concerned with an individual agent and can model its behavior more precisely. Moreover, considering the mean value of the traffic density in a vicinity ahead of the driver could be seen at taking into account both the driver anticipation and a psychological effect. For example, if the driver sees – several dozens of meters ahead of him/her – a speed reduction on traffic, he/she will start to slow down. This observation can be related to the fact that, compared to the fluid mechanics models where the typical number of agents is governed by the Avogadro constant, in traffic models the number of agents is at least 1020 times less. Therefore, a mild nonlocality (evaluation of the downstream traffic flow via averaging over a handful of preceding cars) is a reasonable assumption in the macroscopic traffic models inspired by fluid mechanics. This point of view is exploited in the model of [16]. Note that it is feasible to substitute the basic LWR equation on ρ by the nonlocal LWR introduced in [16] in our nonlocal model for the slow vehicle. Such mildly nonlocal model remains close to the basic local model of [20]. It can be studied combining the techniques of [16] and the ones we developed in this section.

    In this section, we aim at constructing a finite volume scheme and at proving its convergence toward the BV-regular solution to (1). We will use the notations:

    ab=max{a,b}andab=min{a,b}.

    Fix ρ0L1(R;[0,R]) and y0R.

    For a fixed spatial mesh size Δx and time mesh size Δt, let xj=jΔx, tn=nΔt. We define the grid cells Kj+1/2=(xj,xj+1). Let NN such that T[tN,tN+1). We write

    R×[0,T]Nn=0jZPnj+1/2,Pnj+1/2=Kj+1/2×[tn,tn+1).

    We choose to discretize the initial data ρ0(+y0) and the weight function μ with (ρ0j+1/2)jZ and (μj+1/2)jZ where for all jZ, ρ0j+1/2 and μj+1/2 are their mean values on the cell Kj+1/2.

    Remark 8. Others choice could be made, for instance in the case ρ0C(R) such that lim|x|+ρ0(x) exists (in which case, the limit is zero due to the integrability assumption), the values ρ0j+1/2=ρ0(xj+1/2+y0) can be used. The only requirements are

    jZ, ρ0j+1/2[0,R]andρ0Δ=jZρ0j+1/21Kj+1/2 Δx0 ρ0(+y0) in L1loc(R).

    Fix n{0,,N1}. At each time step we first define an approximate velocity of the slow vehicle sn+1 and a constraint level qn+1:

    sn+1=ω(jZρnj+1/2μj+1/2Δx),qn+1=Q(sn+1). (23)

    With these values, we update the approximate traffic density with the marching formula for all jZ:

    ρn+1j+1/2=ρnj+1/2ΔtΔx(Fn+1j+1(ρnj+1/2,ρnj+3/2)Fn+1j(ρnj1/2,ρnj+1/2)), (24)

    where, following the recipe of [6,15],

    Fn+1j(a,b)={Fn+1(a,b)if j0min{Fn+1(a,b),qn+1}if j=0, (25)

    Fn+1 being a monotone consistent and Lipschitz numerical flux associated to F(sn+1,). We will also use the notation

    ρn+1j+1/2=Hn+1j(ρnj1/2,ρnj+1/2,ρnj+3/2), (26)

    where Hn+1j is given by the expression in the right-hand side of (24). We then define the functions

     ρΔ(x,t)=Nn=0jZρnj+1/21Pnj+1/2(x,t) sΔ(t),qΔ(t)=sn+1,qn+1if t[tn,tn+1) yΔ(t)=y0+t0sΔ(u)du.

    Let Δ=(Δx,Δt). For our convergence analysis, we will assume that Δ0, with λ=Δt/Δx verifying the CFL condition

    λsups[0,ωL] (FsaL+FsbL)L1, (27)

    where Fs=Fs(a,b) is the numerical flux, associated to F(s,), we use in (24).

    Remark 9. When considering the Rusanov flux or the Godunov one, (27) is guaranteed when

    2λ(fL+ωL)1.

    Proposition 2 (L stability). The scheme (26) is

    (i) monotone: for all n{0,,N1} and jZ, Hn+1j is nondecreasing with respect to its three arguments;

    (ii) stable:

    n{0,,N}, jZ,ρnj+1/2[0,R]. (28)

    Proof. (i) In the classical case – j{1,0} – we simply differentiate the Lipschitz function Hn+1j and make use of both the CFL condition (27) and the monotonicity of Fn+1. For j{1,0}, note that the authors of [6] pointed out (in Proposition 4.2) that the modification done in the numerical flux (25) does not change the monotonicity of the scheme.

    (ii) The L stability is a consequence of the monotonicity and also of the fact that 0 and R are stationary solutions of the scheme. Indeed, as in [6,Proposition 4.2] for all n{0,,N} and jZ,

    Hn+1j(0,0,0)=0,Hn+1j(R,R,R)=R.

    In order to show that the limit of (ρΔ)Δ – under the a.e. convergence up to a subsequence – is a solution of the conservation law in (1), we derive discrete entropy inequalities. These inequalities also contain terms that will help to pass to the limit in the constrained formulation of the conservation law, as soon as the sequence (qΔ)Δ of constraints is proved convergent as well.

    Proposition 3 (Discrete entropy inequalities). The numerical scheme (26) fulfills the following inequalities for all n{0,,N1}, jZ and κ[0,R]:

    (|ρn+1j+1/2κ||ρnj+1/2κ|)Δx+(Φnj+1Φnj)ΔtRsn+1(κ,qn+1)Δt δj{1,0}+(Φn0Φnint)Δt(δj=1δj=0), (29)

    where Φnj and Φnint denote the numerical entropy fluxes:

    Φnj=Fn+1(ρnj1/2κ,ρnj+1/2κ)Fn+1(ρnj1/2κ,ρnj+1/2κ),Φnint=min{Fn+1(ρn1/2κ,ρn1/2κ),qn+1}min{Fn+1(ρn1/2κ,ρn1/2κ),qn+1}

    and

    Rsn+1(κ,qn+1)=F(sn+1,κ)min{F(sn+1,κ),qn+1}.

    Proof. This result is a direct consequence of the scheme monotonicity. When the constraint does not enter the calculations i.e. j{1,0}, the proof follows [23,Lemma 5.4]. The key point is not only the monotonicity, but also the fact that in the classical case, all the constants κ[0,R] are stationary solutions of the scheme. This observation does not hold when the constraint enters the calculations. For example if j=1,

    Hn+11(κ,κ,κ)=κ+λRsn+1(κ,qn+1).

    Consequently, we have both

    ρn+11/2κHn+11(ρn3/2κ,ρn1/2κ,ρn1/2κ)

    and

    ρn+11/2κHn+11(ρn3/2κ,ρn1/2κ,ρn1/2κ)λRsn+1(κ,qn+1).

    By substracting these last two inequalities, we get

    |ρn+11/2κ|=ρn+11/2κρn+11/2κHn+11(ρn3/2κ,ρn1/2κ,ρn1/2κ)Hn+11(ρn3/2κ,ρn1/2κ,ρn1/2κ)+λRsn+1(κ,qn+1)=|ρn1/2κ|λ(min{Fn+1(ρn1/2κ,ρn1/2κ),qn+1}Fn+1(ρn1/2κ,ρn1/2κ))+λ(min{Fn+1(ρn1/2κ,ρn1/2κ),qn+1}Fn+1(ρn1/2κ,ρn1/2κ))+λRsn+1(κ,qn+1)=|ρn1/2κ|λ(Φn0Φn1)+λ(Φn0Φnint)+λRsn+1(κ,qn+1),

    which is exactly (29) in the case j=1. The case j=0 is similar so we omit the details of the proof for this case.

    Starting from (24) and (29), we can obtain approximate versions of (4) and (5). Let us introduce the functions:

    ΦΔ(ρΔ,κ)=Nn=0jZΦnj1Pnj+1/2;FΔ(sΔ,ρΔ)=Nn=0jZFn+1(ρnj1/2,ρnj+1/2)1Pnj+1/2.

    Proposition 4 (Approximate entropy/constraint inequalities). (i) Fix φCc(R×R+),φ0 and κ[0,R]. Then there exists a constant Cφ1=Cφ1(R,T,L), nondecreasing with respect to its arguments, such that the following inequalities hold for all 0τ<τT:

    ττR|ρΔκ|tφ+ΦΔ(ρΔ,κ)xφdxdt+R|ρΔ(x,τ)κ|φ(x,τ)dxR|ρΔ(x,τ)κ|φ(x,τ)dx+2ττRsΔ(t)(κ,qΔ(t))φ(0,t)dtCφ1(Δt+Δx). (30)

    (ii) Fix ψC([0,T]),ψ0 and φCc(R) such that φ(0)=1. Then there exists a constant Cφ,ψ2=Cφ,ψ2(R,T,L,QL), nondecreasing with respect to its arguments, such that for all 0τ<τT:

    ττR+ρΔt(φψ)+FΔ(sΔ,ρΔ)x(φψ)dxdtR+ρΔ(x,τ)φ(x)ψ(τ)dx+R+ρΔ(x,τ)φ(x)ψ(τ)dxττqΔ(t)ψ(t)dt+Cφ,ψ2(Δx+Δt). (31)

    Proof. Fix k,mN such that τ[tk,tk+1) and τ[tm,tm+1).

    (i) Define for all nN and jZ, φnj+1/2=1ΔxΔtPnj+1/2φ(x,t)dxdt. Multiplying the discrete entropy inequalities (29) by φnj+1/2, then summing over n{k,,m1} and jZ, one obtains after reorganization of the sums (using in particular the Abel/"summation-by-parts" procedure)

    A+B+C+D+E0, (32)

    with

    A=m1n=k+1jZ|ρnj+1/2κ|(φnj+1/2φn1j+1/2)ΔxB=m1n=kjZΦnj(φnj+1/2φnj1/2)ΔtC=jZ|ρkj+1/2κ|φkj+1/2ΔxjZ|ρmj+1/2κ|φm1j+1/2ΔxD=m1n=kRsn+1(κ,qn+1)(φn1/2+φn1/2)ΔtE=m1n=k(Φn0Φnint)(φn1/2φn1/2)Δt.

    Inequality (30) follows from (32) with

    Cφ1=R(T maxt[0,T]2ttφ(,t)L1+4 maxt[0,T] tφ(,t)L1)+RL(T maxt[0,T] 2xxφ(,t)L1+2 maxt[0,T] xφ(,t)L1+4φL+2TxφL),

    making use of the bounds:

    |AττR|ρΔκ|tφdxdt|R(T maxt[0,T] 2ttφ(,t)L1+2 maxt[0,T] tφ(,t)L1)Δt,|BττRΦΔ(ρΔ,κ)xφdxdt|RL(T maxt[0,T] 2xxφ(,t)L1Δx,+2 maxt[0,T] xφ(,t)L1Δt)|CR|ρΔ(x,τ)κ|φ(x,τ)dx+R|ρΔ(x,τ)κ|φ(x,τ)dx|2R maxt[0,T] tφ(,t)L1Δt,|D2ττRsΔ(t)(κ,qΔ(t))φ(0,t)dt|RL(4φLΔt+TxφLΔx)and|E|2RTLxφLΔx.

    (ii) In this case, the constant Cφ,ψ2 reads

    Cφ,ψ2=RφL1(TψL+4ψL)+QLψL(2+TφL)+RLψL(2φL1+TφL1+TφL1).

    Following the proof of (30), define for all nN and jZ,

    ψn=1Δttn+1tnψ(t)dt,andφj+1/2=1Δxxj+1xjφ(x)dx,

    multiply the scheme (24) by φj+1/2ψn, then take the sum over n{k,,m1} and j0. Since the proof is very similar to the one of (i), we omit the details.

    The final step is to obtain compactness for the sequences (ρΔ)Δ and (yΔ)Δ in order to pass to the limit in (30)-(31). We start with (yΔ)Δ.

    Proposition 5. For all t[0,T],

    yΔ(t)=y0+t0ω(RρΔ(x,u)μ(x)dx)du. (33)

    Consequently, there exists yC([0,T]) such that up to an extraction, (yΔ)Δ converges uniformly to y on [0,T].

    Proof. For all t[0,T], if t[tn,tn+1) for some n{0,,N}, then we can write

    yΔ(t)y0=n1k=0tk+1tksk+1du+ttnsn+1du=n1k=0tk+1tkω(jZRρkj+1/2μj+1/2Δx)du+ttnω(jZRρnj+1/2μj+1/2Δx)du=t0ω(RρΔ(x,u)μ(x)dx)du.

    Let us also point out that from (23), we get that for all Δ and almost every t(0,T),

    sΔ(t)=ω(RρΔ(x,t)μ(x)dx). (34)

    Combining (33) and (34), we obtain that for all Δ,

    ˙yΔL=sΔLωLandyΔL|y0|+TωL.

    The sequence (yΔ)Δ is therefore bounded in W1,((0,T)). Making use of the compact embedding of W1,((0,T)) in C([0,T]), we get the existence of yC([0,T]) such that up to the extraction of subsequence, (yΔ)Δ converges uniformly to y on [0,T].

    The presence of a time dependent flux in the conservation law of (1) complicates the obtaining of compactness for (ρΔ)Δ. In particular, the techniques used in [10,11] to derive localized BV estimates don't apply here since our problem lacks time translation invariance. In the present situation, it would be possible to derive weak BV estimates ([6,23]). We choose different options. Similarly to what we did in Section 3, we propose two ways to obtain compactness, which will lead to two convergence results.

    First, we choose to adapt techniques and results put forward by Towers in [33]. With this in mind, we suppose in this section that fC2([0,R]) and strictly concave. Therefore,

    α>0, ρ[0,R],f(ρ)α. (35)

    Though this assumption is stronger than the nondegeneracy one (15), since f is bell-shaped, these two assumptions are similar in their spirit. We will also assume, following [33], that

    the numerical flux chosen in (24) is eitherthe Engquist-Osher one or the Godunov one. (36)

    To be precise, the choice made for the numerical flux at the interface – i.e. when j=0 in (25) – does not play any role. What is important is that away from the interface, one chooses either the Engquist-Osher flux or the Godunov one. We denote for all n{0,,N+1} and jZ,

    Dnj=max{ρnj1/2ρnj+1/2,0}.

    We will also use the notation

    ˆZ=Z{1,0,1}.

    In [33], the author dealt with a discontinuous in both time and space flux and the specific "vanishing viscosity" coupling at the interface. The discontinuity in space was localized along the curve {x=0}. Here, we deal with only a discontinuous in time flux, but we also have a flux constraint along the curve {x=0} since we work in the bus frame. The applicability of the technique of [33] for our case with moving interface and flux-constrained interface coupling relies on the fact that one can derive a bound on Dnj as long as the "interface" does not enter the calculations for Dnj i.e. jˆZ. This is what the following lemma points out under Assumptions (35)-(36). For readers' convenience and in order to highlight the generality of the technique of Towers [33], let us provide the key elements of the argumentation leading to compactness.

    Lemma 4.1. Let n{0,,N1} and jˆZ. Then if a=λα4, we have

    Dn+1jmax{Dnj1,Dnj,Dnj+1}a(max{Dnj1,Dnj,Dnj+1})2 (37)

    and

    Dn+1j1min{|j|1,n+1}a. (38)

    Proof. (Sketched) Inequality (38) is an immediate consequence of inequality (37), see [33,Lemma 4.3]. Obtaining inequality (37) however, is less immediate. Let us give some details of the proof.

    First, note that by introducing the function ψ:zzaz2, inequality (37) can be stated as:

    Dn+1jψ(max{Dnj1,Dnj,Dnj+1}). (39)

    Then, one can show – only using the monotonicity of both the scheme and the function ψ – that under the assumption inequality (39) holds when

     (ρnj+3/2ρnj+1/2), (ρnj1/2ρnj3/2)0, (40)

    it follows that inequality (39) holds for all cases. And finally in [33,Page 23], the author proves that if the flux considered is either the Engquist-Osher flux or the Godunov flux, then (40) holds.

    The following lemma is an immediate consequence of inequality (38).

    Lemma 4.2. Fix 0<ε<X. Let i,JN such that εKi+1/2 and XKJ1/2. Then if Δx/ε is sufficiently small, there exists a constant B=B(R,X,1a,1ε), nondecreasing with respect to its arguments, such that for all ni1,

    J1j=i|ρnj+1/2ρnj1/2|,i1j=J+1|ρnj+1/2ρnj1/2|B (41)

    and

    J2j=i|ρn+1j+1/2ρnj+1/2|,i2j=J+1|ρn+1j+1/2ρnj+1/2|2λLB. (42)

    Proposition 6. There exists ρL(R×(0,T)) such that up to the extraction of a subsequence, (ρΔ)Δ converges almost everywhere to ρ in R×(0,T).

    Proof. Fix 0<ε<X and t>λε. Denote by Ω(X,ε)=(X,ε)(ε,X). Introduce i,J,nN such that εKi+1/2, XKJ1/2 and t[tn,tn+1). Remark that

    (n+1)Δt>t>λελ(i Δx)=iΔt,

    i.e. ni1. Then if we suppose that Δx/ε is sufficiently small, we can use Lemma 4.2. From (41), we get

    TV(ρΔ(t)|Ω(X,ε))2B (43)

    and from (42), we deduce

    Ω(X,ε)|ρΔ(x,t+Δt)ρΔ(x,t)|dx4LBΔt. (44)

    Combining (43)-(44) and the L bound (28), a functional analysis result ([24,Theorem A.8]) ensures the existence of a subsequence which converges almost everywhere to some ρ on Ω(X,ε)×(λε,T). By a standard diagonal process we can extract a further subsequence (which we do not relabel) such that (ρΔ)Δ converges almost everywhere to ρ on R×(0,T).

    Theorem 4.3. Fix ρ0L1(R;[0,R]) and y0R. Suppose that fC2 satisfies Assumptions (2)-(7)-(35). Suppose also that in (25), we use the Engquist-Osher flux or the Godunov one when j0 and any other monotone consistent and Lipschitz numerical flux when j=0. Then under the CFL condition (27), the scheme (23) – (25) converges to an admissible weak solution to Problem (1).

    Proof. We have shown that – up to the extraction of a subsequence – yΔ converges uniformly on [0,T] to some yC([0,T]) and that ρΔ converges a.e. on R×(0,T) to some ρL(R×(0,T)). We now prove that this couple (ρ,y) is an admissible weak solution to Problem (1) in the sense of Definition 2.1.

    Recall that for all Δ and t[0,T],

    yΔ(t)=y0+t0ω(RρΔ(x,u)μ(x)dx)du.

    When letting Δ0, the dominated convergence theorem ensures that y satisfies (6). Apply inequality (30) with τ=0, τ=T, φCc(R×[0,T)),φ0 and κ[0,R] to obtain

    T0R|ρΔκ|tφ+ΦΔ(ρΔ,κ)xφdxdt+R|ρ0Δκ|φ(x,0)dxCφ1(Δx+Δt).

    Then the a.e. convergence of (sΔ)Δ to ˙y – coming from (34) – and the a.e. convergence of (ρΔ)Δ to ρ ensure that when letting Δ0, we get

    T0R|ρκ|tφ+Φ˙y(t)(ρ,κ)xφdxdt+R|ρ0(x+y0)κ|φ(x,0)dx0,

    and consequently ρC([0,T];L1loc(R)), see Remark 2. Now, we pass to the limit in (30) and (31) using the a.e. convergence of (sΔ)Δ to ˙y and of (ρΔ)Δ to ρ as well as the continuity of Q and ω. Consequently, for all test functions φCc(R×R+),φ0 and κ[0,R], the following inequalities hold for almost every 0τ<τT:

    ττR|ρκ|tφ+Φ˙y(t)(ρ,κ)xφdxdt+R|ρ(x,τ)κ|φ(x,τ)dxR|ρ(x,τ)κ|φ(x,τ)dx+2ττR˙y(t)(κ,q(t))φ(0,t)dt0.

    To conclude, note that the expression in the left-hand side of the previous inequality is a continuous function of (τ,τ) which is almost everywhere greater than the continuous function 0. By continuity, this expression is everywhere greater than 0, which proves that ρ satisfies the entropy inequalities (4). Using similar arguments, one shows that ρ also satisfies the constraint inequalities (5). This shows that the couple (ρ,y) is an admissible weak solution to (1), and that concludes the proof of convergence.

    We proved than in the L framework, the scheme converges to an admissible weak solution, but note that there is no guarantee of uniqueness in this construction. Also stress that we cannot extend this result to general consistent monotone numerical fluxes beyond hypothesis (36).

    The following result is the discrete version of Lemma 2.3 so it is consistent that the proof uses the discrete analogous arguments of the ones we used in the proof of Lemma 2.3.

    Lemma 4.4. Introduce for all Δ>0 the function ξΔ defined for all t[0,T] by

    ξΔ(t)=RρΔ(x,t)μ(x)dx.

    Then ξΔ has bounded variation and consequently, so does sΔ.

    Proof. Since μBV(R), there exists a sequence of smooth functions (μ)NBV(R)Cc(R) such that

    μμL1 + 0andTV(μ) + TV(μ).

    Introduce for all N and t[0,T], the function ξΔ,(t)=RρΔ(x,t)μ(x)dx and let K>0 such that

    N,μL1,TV(μ)K.

    For all N and t,s[0,T], if t[tk,tk+1) and s[tm,tm+1), we have

    |ξΔ,(t)ξΔ,(s)|=|ξΔ,(tk)ξΔ,(tm)|=|RρΔ(x,tk)μ(x)dxRρΔ(x,tm)μ(x)dx|=|jZ(ρkj+1/2ρmj+1/2)μj+1/2Δx|,μj+1/2=1Δxxj+1xjμ(x)dx=|jZk1τ=m(ρτ+1j+1/2ρτj+1/2)μj+1/2Δx|=|k1τ=mjZ(Fτ+1j(ρτj1/2,ρτj+1/2)Fτ+1j+1(ρτj+1/2,ρτj+3/2))μj+1/2Δt|=|k1τ=mjZFτ+1j+1(ρτj+1/2,ρτj+3/2)(μj+3/2μj+1/2)Δt|RLk1τ=mTV(μ)ΔtRLK(|ts|+2Δt).

    Consequently, for all N, Δ>0 and t,τ[0,T], the triangle inequality yields:

    |ξΔ(t)ξΔ(τ)|2RμμL1+RLK(|tτ|+2Δt).

    Letting +, we get that for all Δ>0 and t,τ[0,T],

    |ξΔ(t)ξΔ(τ)|RLK(|tτ|+2Δt),

    which leads to

    TV(ξΔ)=Nk=0|ξΔ(tk+1)ξΔ(tk)|3RLK(T+Δt).

    This proves that ξΔBV([0,T]). Since ω is Lipschitz continuous, sΔ also has bounded variation.

    Theorem 4.5. Fix ρ0L1(R;[0,R])BV(R) and y0R. Suppose that f satisfies (2)-(7)-(16) and that Q satisfies (17). Suppose also that in (25), we use the Godunov flux when j=0 and any other monotone consistent and Lipschitz numerical flux when j0. Then under the CFL condition (27), the scheme (23) – (25) converges to a BV-regular solution to Problem (1).

    Proof. All the hypotheses of Lemma A.4 are fulfilled. Consequently, there exists a constant Cε>0 such that for all n{0,,N1},

    TV(ρΔ(tn+1))TV(ρ0)+4R+Cε(nk=0|qk+1qk|+nk=0|sk+1sk|)TV(ρ0)+4R+Cε(1+QL)nk=0|sk+1sk|. (45)

    Making use of Lemma 4.4, we obtain that for all n{0,,N},

    nk=0|sk+1sk|=nk=0|sΔ(tk+1)sΔ(tk)|ωLnk=0|ξΔ(tk+1)ξΔ(tk)|3RLKωL(T+Δt).

    where the constant K was introduced in the proof of Lemma 4.4. The two last inequalities imply that for all t[0,T], we have

    TV(ρΔ(t))TV(ρ0)+4R+3Cε(1+QL)ωLRLK(T+Δt). (46)

    Therefore, the sequence (ρΔ)Δ is uniformly in time bounded in BV(R). Using [22,Appendix], we get the existence of ρC([0,T];L1loc(R)) such that

    t[0,T], ρΔ(t) Δ0 ρ(t) in L1loc(R).

    Following the proof of Theorem 4.3, we show that (ρ,y) is an admissible weak solution. Then passing to the limit in (46), the lower semi-continuity of the BV semi-norm ensures that (ρ,y) is also BV-regular.

    Remark 10. Note the complementarity of the hypotheses made in the above theorem with the ones of Theorem 4.3. Recall that in Theorem 4.3, we needed the Godunov flux only away from the interface.

    In this section we present some numerical tests performed with the scheme analyzed in Section 4. In all the simulations we take the uniformly concave flux f(ρ)=ρ(1ρ) (the maximal car velocity and the maximal density are assumed to be equal to one). Following the hypotheses of Theorem 4.5, we choose the Godunov flux at the interface, and the Rusanov one away from the interface. We will use weight functions of the kind

    μk(x)=2k1[0;12k](x),

    for one (in Section 5.1) or several (in Section 5.3) values of kN.

    In this section, consider a two-lane road on which a bus travels with a speed given by the function

    ω(ρ)={α(β+ρ)2if0ρ0.61ρif0.6ρ1,

    where α and β are chosen so that ω(0)=0.7 and ω(0.6)=0.4, as illustrated in Figure 1 (left). The set-up of the experiment is the following. Consider a domain of computation [0,11], the weight function μ4 and the following data:

    ρ0(x)=0.51[0.5;1](x),y0=1.5,Q(s)=0.75×(1s2)2.
    Figure 1.  Evolution in time of the subjective density ξ and the bus velocity ˙y (1280 cells).

    The idea behind the choice of Q is that in average (between the two lanes), the presence of the slow vehicle reduces by 25% the maximum traffic flow. As we can see in Figure 1 (right), the slow vehicle nearly always travels at maximum velocity. It makes sense because even though we can see that cars are overtaking it (Figure 1, right and Figure 2), the density ξ ahead of it is never sufficiently important to make it go slower.

    Figure 2.  The numerical solution at different fixed times.

    Remark 11. The function ω we chose above is not of the form as required in [27,28]. Once again, let us stress that the particular form ω(ρ)=min{Vbus;1ρ}, where Vbus is the maximum bus velocity, is crucial for the well-posedness result of [27,28] to hold. Indeed, it is essential in the analysis of [27,28] that the velocity of the bus be constant (equal to Vbus) across the nonclassical shocks. Our nonlocal model is not bound to this restriction.

    We also perform a convergence analysis for this test. In Table 1, we computed the relative errors

    EΔρ=ρΔρΔ/2L1((0,T);L1(R))andEΔy=yΔyΔ/2L,
    Table 1.  Measured errors at time T=13.
    Number of cells EΔρ (×102) EΔy (×103)
    160 24.053 48.0643
    320 15.731 15.939
    640 9.647 7.698
    1280 6.197 3.715
    2560 3.226 1.777
    5120 1.936 0.889
    10240 1.055 0.443

     | Show Table
    DownLoad: CSV
    Figure 3.  Rates of convergence at time T=13.

    for different number of space cells at the final time T=13. We see that those ratio converge with convergence orders approximately equal to 0.76 for the car density and approximately equal to 1.1 for the slow moving vehicle position.

    Now we confront the numerical tests performed with our model with the tests done by the authors in [14] approximating the original problem of [20]. We deal with a road of length 1 parametrized by the interval [0,1] and choose the weight function μ3. Moreover,

    ω(ρ)=min{0.3; 1ρ}andQ(s)=0.6×(1s2)2.

    First, consider the initial datum

    ρ0(x)={0.4ifx<0.50.5ifx>0.5 y0=0.5. (47)

    The numerical solution is composed of two classical shocks separated by a nonclassical discontinuity, as illustrated in Figure 4 (left).

    Figure 4.  Evolution in time of the numerical density corresponding to initial data (47) (left) and (48) (right), with 1280 cells.

    Next, we choose

    ρ0(x)={0.8ifx<0.50.5ifx>0.5 y0=0.5. (48)

    The values of the initial condition create a rarefaction wave followed by a nonclassical and classical shocks, as illustrated in Figure 4 (right).

    Finally, still following [14], we consider

    ρ0(x)={0.8ifx<0.50.4ifx>0.5 y0=0.4. (49)

    Here the solution is composed of a rarefaction wave followed by nonclassical and classical shocks on the density that are created when the slow vehicle approaches the rarefaction and initiates a moving bottleneck, as illustrated in Figure 5.

    Figure 5.  Evolution in time of the numerical density corresponding to initial data (49) (1280 cells).

    With these three tests, we can already see – in a qualitative way – the resemblance between the numerical approximations to the solutions to our model and the numerical approximations of [14]. One way to quantify their proximity is for example to evaluate the L1 error between the car densities and the L error between the bus positions. More precisely, denote by (ρΔ,yΔ) the approximation of the BV-regular solution to (1) obtained with the scheme (23) – (25), and denote by (¯ρΔ,¯yΔ) the couple obtained with this same scheme but

    replacingsn+1=ω(jZρnj+1/2μj+1/2Δx)bysn+1=ω(ρn1/2).

    Let us precise that this is not the scheme the authors of [14] proposed. However, this scheme is consistent with the problem

    {tρ+x(F(˙y(t),ρ))=0R×(0,T)ρ(x,0)=ρ0(x+y0)xRF(˙y(t),ρ)|x=0Q(˙y(t))t(0,T)˙y(t)=ω(ρ(0+,t))t(0,T)y(0)=y0 (50)

    and behaves in a stable way in the calculations we performed. Therefore, the couple (¯ρΔ,¯yΔ) is expected to give a reasonable approximation of the solution to (50). With this in mind, for the case (49) and still with the weight function μ3, we computed in Table 2 the measured errors

    EΔL1=ρΔ¯ρΔL1((0,T);L1(R))andEΔL=yΔ¯yΔL.
    Table 2.  Measured errors at time T=0.7245.
    Number of cells EΔL1 (×104) EΔL (×103)
    160 32.672 18.519
    320 14.236 7.341
    640 5.837 3.701
    1280 3.833 4.879
    2560 3.207 6.405
    5120 2.922 7.144
    10240 2.776 7.501
    20480 2.698 7.674
    40960 2.658 7.759

     | Show Table
    DownLoad: CSV

    These calculations indicate that for a sufficiently large number of cells J40960,

    EΔL12.7×104andEΔL7.6×103.

    This indicates the discrepancy between our nonlocal and the local model (50) of [20]. The idea is now to fix the number of cells J=40960 and to make the length of the weight function support go to zero. In Table 3, we have computed, for different weight functions, the error between the approximations of the two models. This error corresponds, as in the above calculation, to the residual error observed starting from a sufficiently small Δx.

    Table 3.  Measured errors at time T=0.7245.
    weight function EΔL1 EΔL
    μ1 6.810×103 5.489×102
    μ2 1.105×103 1.972×102
    μ3 2.658×104 7.759×103
    μ4 9.232×105 2.913×103
    μ5 6.190×105 9.110×104

     | Show Table
    DownLoad: CSV

    Remark 12. The previous simulations show a closeness between our model as μδ0+ and (50). Let us however point that the nonlocality in space for the slow vehicle introduces an undesirable artefact into the model. In the rarefaction regime one may observe that the large vehicle may move a bit faster that the surrounding flow. The situation where this effect becomes truly perceptible is when considering initial data of the type

    ρ0(x)={1ifx<xb0ifx>xb. (51)

    Indeed, for such data, there exists a small time interval [0,δ] in which ˙y(t)>v(ρ(y(t)+))=0, which would suggest that the slow vehicle moves forward while the cars in front of it do not. This time interval is in fact quite small due to the narrowness of the support of the weight function. The local model does not develop such phenomena. This qualitative artefact precludes us from giving a microscopic interpretation to the model, which main output is the global influence of the slow vehicle on the flow; however, let us stress that the phenomenon becomes quantitatively negligible for larger times. Indeed, Ole\u{ı}nik estimate on decay of positive waves ensures that data of the type (51) evolve into rarefaction waves and do not appear while driving: the classical LWR model precludes the formation of rarefaction waves focused at positive time. The modification of the classical LWR brought by the constraint may produce nonclassical waves at positive times; while these waves are downward jumps in density like in (51), they are situated precisely at the location of the constraint and not slightly behind it, like in (51).

    Even if we are unable, at this time, to rigorously link our problem (1) with μδ0+ and the original problem (50) of the authors in [20], this last experiment corroborates the conjecture that the local model (50) is the singular limit of our model in the case ω is of the form ω(ρ)=min{Vbus;1ρ}. The other interesting question is whether the local model is well posed beyond this particular choice of ω.

    The author is most grateful to Boris Andreianov for his constant support and many enlightening discussions.

    We focus on the study of the following class of models:

    {tρ+x(F(s(t),ρ))=0R×(0,T)ρ(x,0)=ρ0(x)xRF(s(t),ρ)|x=0q(t)t(0,T), (52)

    where sBV([0,T];[0,Σ]) for some Σ>0 and qBV([0,T];R+). We suppose that FC1([0,Σ]×[0,R]) and that for all s[0,Σ], F(s,) is bell-shaped i.e.

    s[0,Σ], F(s,0)=0, F(s,R)0and! ¯ρs(0,R), ρF(s,ρ)(¯ρsρ)>0 for a.e. ρ(0,R). (53)

    This framework covers the particular case when F takes the form:

    F(s(t),ρ)=f(ρ)s(t)ρ,

    with bell-shaped f:[0,R]R+, which our model (1) is based on. This class of models is well known, especially when the flux function is not time dependent, cf. [17,6]. In this appendix, we establish in passing the well-posedness of Problem (52), but our main interest lies in the BV in space regularity of the solutions. More precisely, we aim at obtaining a bound on the total variation of the solutions to (52), using a finite volume approximation which allows for sharp control of the variation at the constraint. Note that the alternative offered by wave-front tracking would be cumbersome because of the explicit time-dependency in (52). In the general case, entropy solutions to limited flux problems like (52) do not belong to L((0,T);BV(R)), see [1]. We will show that it is the case under a mild assumption on the constraint function q – see Assumption (59) below – and provided that

    ρ0L1(R;[0,R])BV(R).

    Throughout the appendix, for all s[0,Σ] and a,b[0,R], we denote by

    Φs(a,b)=sign(ab)(F(s,a)F(s,b))

    the classical Kružkov entropy flux associated with the Kružkov entropy ρ|ρκ|, for all κ[0,R], see [25].

    Let us first recall the following definition.

    Definition A.1. A bounded function ρL(R×(0,T)) is an admissible weak solution to (52) if

    (ⅰ) the following regularity is fulfilled: ρC([0,T];L1loc(R));

    (ⅱ) for all test functions φCc(R×R+),φ0 and κ[0,R], the following entropy inequalities are verified for all 0τ<τT:

    ττR|ρκ|tφ+Φs(t)(ρ,κ)xφdxdt+R|ρ(x,τ)κ|φ(x,τ)dxR|ρ(x,τ)κ|φ(x,τ)dx+2ττRs(t)(κ,q(t))φ(0,t)dt0,

    where

    Rs(t)(κ,q(t))=F(s(t),κ)min{F(s(t),κ),q(t)};

    (ⅲ) for all test functions ψC([0,T]),ψ0 and some given φCc(R) which verifies φ(0)=1, the following weak constraint inequalities are verified for all 0τ<τT:

    ττR+ρt(φψ)+F(s(t),ρ)x(φψ) dxdtR+ρ(x,τ)φ(x)ψ(τ)dx+R+ρ(x,τ)φ(x)ψ(τ)dxττq(t)ψ(t)dt.

    Definition A.2. An admissible weak solution ρ will be called BV-regular if it verifies ρL((0,T);BV(R)).

    As we pointed out before, this notion of solution is well suited for passage to the limit of a.e. convergent sequences of exact or approximate solutions. However, it is not so well-adapted to prove uniqueness. An equivalent notion of solution, based on explicit treatment of traces of ρ at the constraint, was introduced by the authors of [7]. This notion of solution leads to the following stability estimate.

    Theorem A.3. Fix s1,s2BV([0,T];[0,Σ]), ρ10,ρ20L1(R;[0,R])BV(R) and q1,q2BV([0,T];R+). Denote by ρ1 a BV-regular solution to (52) with data ρ10,q1,s1 and ρ2 an admissible weak solution to (52) with data ρ20,q2,s2. Suppose that the flux functions (t,ρ)F(s1(t),ρ),F(s2(t),ρ) satisfy (53). Then for a.e. t(0,T), we have:

    ρ1(t)ρ2(t)L1ρ10ρ20L1+2t0|q1(τ)q2(τ)|dτ+2t0F(s1(τ),)F(s2(τ),)Ldτ+t0||ρF(s1(τ),)ρF(s2(τ),)||LTV(ρ1(τ))dτ. (54)

    In particular, Problem (52) admits at most one BV-regular solution.

    Proof. Since our interest to details lies rather on the numerical approximation point of view, we do not fully prove this statement but we give the essential steps leading to this stability result.

    Definition of solution. First, the authors of [7] introduce a subset of R2 called germ, which can be seen as the set of all the possible traces of a solution to (52). Then, they say that ρ is a solution to (52) if it satisfies entropy inequalities away from the interface – i.e. with φCc(R×R+) in the entropy inequalities – and if the couple constituted of left-side and the right-side traces of ρ belongs to this so-called germ.

    Equivalence of the two definitions. The next step is to prove that this latter definition of solution is equivalent to Definition A.1. This part is done using good choices of test functions, see [7,Theorem 3.18] or [6,Proposition 2.5,Theorem 2.9].

    First stability estimate. One first shows that if s1=s2, then for a.e. t(0,T), one has

    ρ1(t)ρ2(t)L1ρ10ρ20L1+2t0|q1(τ)q2(τ)|dτ. (55)

    The proof starts with the classical doubling of variables method of Kružkov [25,Theorem 1] and then uses the germ structure, what the authors of [7] called L1-dissipativity, see [7,Definition 3.1] and [6,Lemma 2.7].

    Proof of estimate (54). The proof is based upon estimate (55) and elements borrowed from [8,18]. Most details can be found in the proof of [21,Theorem 2.1].

    Remark 13. Though the definition of solutions with the germ explicitly involves the traces of ρ, we did not discuss the existence of such traces. A first way to ensure such existence is to deal with BV-regular solutions. That way, traces do exist and are to be understood in the sense of BV functions. Outside the BV framework, existence of strong traces for solutions to (52) is ensured provided an assumption on the fundamental diagram like (15), see [2,29]. Finally, if one does not want to impose such a condition on the flux, (which is our case in this appendix), one can follow what the authors of [7] proposed (in Section 2) and consider the "singular mapping traces."

    We now turn to the proof of the existence of BV-regular solutions by the means of a finite volume scheme.

    Fix ρ0L1(R;[0,R]). For a fixed spatial mesh size Δx and time mesh size Δt, let xj=jΔx, tn=nΔt. We define the grid cells Kj+1/2=(xj,xj+1) and NN such that T[tN,tN+1). We write

    R×[0,T]Nn=0jZPnj+1/2,Pnj+1/2=Kj+1/2×[tn,tn+1).

    We choose to discretize the initial data ρ0 and the functions s,q with (ρ0j+1/2)j, (sn)n and (qn)n where for all jZ and n{0,,N}), ρ0j+1/2, sn and qn are their mean values on each cell Kj+1/2 and [tn,tn+1). Following [6], the marching formula of the scheme is the following: for all n{0,,N1} and jZ:

    ρn+1j+1/2=ρnj+1/2ΔtΔx(Fnj+1(ρnj+1/2,ρnj+3/2)Fnj(ρnj1/2,ρnj+1/2)), (56)

    where

    Fnj(a,b)={Fn(a,b)if j0min{Fn(a,b),qn)}if j=0, (57)

    Fn being a monotone consistent and Lipschitz numerical flux associated to F(sn,). We then define

    ρΔ(x,t)=ρnj+1/2 if (x,t)Pnj+1/2andsΔ(t),qΔ(t)=sn,qn if t[tn,tn+1).

    Let Δ=(Δx,Δt). For the convergence analysis, we will assume that Δ0, with λ=Δt/Δx, verifying the CFL condition

    λsups[0,Σ](FsaL+FsbL)L1, (58)

    where Fs=Fs(a,b) is the numerical flux – associated to F(s,) – we use in the scheme (56). From now, the analysis of the scheme follows the same path as in Section 4. In that order, we prove that the scheme (56)-(57) is L stable, satisfies discrete entropy inequalities similar to (29) and approximate entropy/constraint inequalities similar to (30)-(31). Only the compactness for (ρΔ)Δ is left to obtain since the L1loc compactness for the sequences (sΔ)Δ and (qΔ)Δ is clear. One way to do so is to derive uniform BV bounds.

    Lemma A.4. We suppose that ρ0L1(R;[0,R])BV(R) and that q verifies the assumption

    ε>0, t[0,T], s[0,Σ], q(t)maxρ[0,R]F(s,ρ)ε:=qε(s). (59)

    Then there exists a constant Cε=Cε(sFL) nondecreasing with respect to its argument such that for all n{0,,N1},

    TV(ρΔ(tn+1))TV(ρ0)+4R+Cε(nk=0|qk+1qk|+nk=0|sk+1sk|), (60)

    where ρΔ=(ρnj+1/2)n,j is the finite volume approximation constructed with the scheme (56)-(57), using the Godunov numerical flux when j=0 in (57).

    Proof. Fix n{0,N1}. With this set up we can follow the proofs of [13,Section 2] to obtain the following estimate:

    jZ|ρn+1j+1/2ρn+1j1/2|TV(ρ0)+4R+2nk=0|(ˆρsk+1(qk+1)ˆρsk(qk))(˘ρsk+1(qk+1)˘ρsk(qk))|,

    where for all k{0,,n}, the couple (ˆρsk(qk),˘ρsk(qk))[0,R]2 is uniquely defined by the conditions

    F(sk,ˆρsk(qk))=F(sk,˘ρsk(qk))=qkandˆρsk(qk)>˘ρsk(qk).

    Denote by Ω(ε) the open subset

    Ω(ε)=s[0,Σ]Ωs(ε)

    where for all s[0,Σ], Ωs(ε)=(˘ρs(qε(s)),ˆρs(qε(s))). By Assumption (59), the continuous function (s,ρ)|ρF(s,ρ)| is positive on the compact subset [0,Σ]×[0,R]Ω(ε). Hence, it attains its minimal value C0>0. Consequently, for all s[0,Σ], if one denotes by Is:[0,˘ρs(qε(s))][0,qε(s)] the increasing part of F(s,), this function carries out a C1-diffeomorphism. Moreover,

    q[0,qε(s)], |(I1s)(q)|1C0.

    Then, for all k{0,,n},

    |˘ρsk+1(qk+1)˘ρsk(qk)|=|(I1sk+1)(qk+1)˘ρsk(qk)|1C0|qk+1qk|+|(I1sk+1)(qk)˘ρsk(qk)|=1C0|qk+1qk|+|(I1sk+1)(qk)(I1sk+1)  Isk+1(˘ρsk(qk))|1C0(|qk+1qk|+|qkIsk+1(˘ρsk(qk))|)=1C0(|qk+1qk|+|F(sk,˘ρsk(qk))F(sk+1,˘ρsk(qk))|)1C0(|qk+1qk|+sFL|sk+1sk|)1+sFLC0(|qk+1qk|+|sk+1sk|).

    Using the same techniques, one can show that the same inequality holds when considering |ˆρsk+1(qk+1)ˆρsk(qk)|. Therefore, inequality (60) follows with

    Cε=4×(1+sFLC0).

    Remark 14. Recall we suppose that F:[0,Σ]×[0,R] is continuously differentiable, but if we look in the details of the proof above, we actually need F=F(s,ρ) to be continuously differentiable with respect to s and

    s[0,Σ], F(s,)C1([0,R]{¯ρs}),¯ρs=argmaxρ[0,R] F(s,ρ).

    Corollary 1. Fix ρ0L1(R;[0,R])BV(R), sBV([0,T],[0,Σ]) and qBV([0,T],R+). Suppose that q verifies Assumption (59). Let ρΔ=(ρnj+1/2)n,j be the finite volume approximate solution constructed with the scheme (56)-(57), using the Godunov numerical flux when j=0 in (57), and any other monotone consistent and Lipschitz numerical flux when j0. Then there exists ρC([0,T];L1loc(R)) such that

    t[0,T], ρΔ(t) Δ0 ρ(t) in L1loc(R).

    Proof. Since s and q have bounded variation, inequality (60) leads to an uniform in time BV bound for the sequence (ρΔ)Δ. Then the result from [22,Appendix] establish the compactness statement.

    Theorem A.5. Fix ρ0L1(R;[0,R])BV(R), sBV([0,T];[0,Σ]), FC1([0,Σ]×[0,R]) verifying (53) and qBV([0,T];R+). Suppose that in (57), we use the Godunov flux when j=0 and any other monotone consistent and Lipschitz numerical flux when j0. Finally, suppose that q satisfies (59). Then under the CFL condition (58), the scheme (56)-(57) converges to an admissible weak solution ρ, to (52), which is also BV-regular. More precisely, there exists a constant Cε=Cε(sFL) nondecreasing with respect to its argument such that

    t[0,T], TV(ρ(t))TV(ρ0)+4R+Cε(TV(q)+TV(s)). (61)

    Proof. From the scheme (56), one can derive approximate entropy/constraint inequalities analogous to (30)-(31) of Section 4. Let ρ be the limit to the finite volume scheme, the compactness of (ρΔ)Δ coming from the last corollary. We already know that ρC([0,T];L1loc(R)). By passing to the limit in the approximate entropy/constraint inequalities verified by (ρΔ)Δ we get that ρ satisfies the entropy/constraint inequalities of Definition A.1. This shows that ρ is an admissible weak solution to Problem (52). Finally, from (60), the lower semi-continuity of the BV semi-norm ensures that ρL([0,T];BV(R)) and verifies (61). This concludes the proof.

    Corollary 2. Fix ρ0L1(R;[0,R])BV(R), sBV([0,T];[0,Σ]), FC1([0,Σ]×[0,R]) verifying (53) and qBV([0,T];R+). Suppose that q satisfies Assumption (59). Then Problem (52) admits a unique BV-regular solution ρ. Moreover, ρ satisfies the bound (61).

    Proof. Uniqueness comes from Theorem 54, the existence and the BV bound comes from Theorem A.5.

    Remark 15. Under the hypotheses of Corollary 2, if we prove the existence of an other admissible weak solution ¯ρ to (11) (by another method, splitting for instance), then Theorem A.3 ensures that ¯ρ=ρ.



    [1] H. Ahn, Emergent behaviors of thermodynamic Cucker–Smale ensemble with a unit-speed constraint, Discrete Continuous Dyn Syst Ser B, 28 (2023), 4800–4825. https://doi.org/10.3934/dcdsb.2023042 doi: 10.3934/dcdsb.2023042
    [2] H. Ahn, Uniform stability of Cucker–Smale and thermodynamic Cucker–Smale ensembles with singular kernels, Netw. Heterog. Media, 17 (2022), 753–782. https://doi.org/10.3934/nhm.2022025 doi: 10.3934/nhm.2022025
    [3] H. Ahn, S. Y. Ha, W. Shim, Emergent behaviors of the discrete thermodynamic Cucker–Smale model on Riemannian manifolds, J. Math. Phys., 62 (2021), 122701. https://doi.org/10.1063/5.0058616 doi: 10.1063/5.0058616
    [4] H. Ahn, S. Y. Ha, W. Shim, Emergent dynamics of a thermodynamic Cucker–Smale ensemble on complete Riemannian manifolds, Kinet. Relat. Models., 14 (2021), 323–351. https://doi.org/10.3934/krm.2021007 doi: 10.3934/krm.2021007
    [5] J. A. Carrillo, Y. P. Choi, P. B. Muncha, J. Peszek, Sharp conditions to avoid collisions in singular Cucker–Smale interactions, Nonlinear Anal Real World Appl, 37 (2017), 317–328. https://doi.org/10.1016/j.nonrwa.2017.02.017 doi: 10.1016/j.nonrwa.2017.02.017
    [6] J. A. Carrillo, M. Fornasier, J. Rosado, G. Toscani, Asymptotic flocking dynamics for the kinetic Cucker–Smale model, SIAM. J. Math. Anal., 42 (2010), 218–236. https://doi.org/https://doi.org/10.1137/090757290 doi: 10.1137/090757290
    [7] P. Cattiaux, F. Delebecque, L. Pedeches, Stochastic Cucker–Smale models: old and new, Ann. Appl. Probab., 28 (2018), 3239–3286. https://doi.org/10.1214/18-AAP1400 doi: 10.1214/18-AAP1400
    [8] H. Cho, J. G. Dong, S. Y. Ha, Emergent behaviors of a thermodynamic Cucker–Smale flock with a time-delay on a general digraph, Math. Methods Appl. Sci., 45 (2021), 164–196. https://doi.org/https://doi.org/10.1002/mma.7771 doi: 10.1002/mma.7771
    [9] S. H. Choi, S. Y. Ha, Emergence of flocking for a multi-agent system moving with constant speed, Commun. Math. Sci., 14 (2016), 953–972. https://doi.org/10.4310/CMS.2016.v14.n4.a4 doi: 10.4310/CMS.2016.v14.n4.a4
    [10] Y. P. Choi, S. Y. Ha, J. Kim, Propagation of regularity and finite-time collisions for the thermomechanical Cucker–Smale model with a singular communication, Netw. Heterog. Media, 13 (2018), 379–407. https://doi.org/10.3934/nhm.2018017 doi: 10.3934/nhm.2018017
    [11] Y. P. Choi, D. Kalise, J. Peszek, A. A Peters, A collisionless singular Cucker–Smale model with decentralized formation control, SIAM J. Appl. Dyn. Syst., 18 (2019), 1954–1981. https://doi.org/10.1137/19M1241799 doi: 10.1137/19M1241799
    [12] J. Cho, S. Y. Ha, F. Huang, C. Jin, D. Ko, Emergence of bi-cluster flocking for the Cucker–Smale model, Math. Models Methods Appl. Sci., 26 (2016), 1191–1218. https://doi.org/10.1142/S0218202516500287 doi: 10.1142/S0218202516500287
    [13] F. Cucker, J. G. Dong, A conditional, collision-avoiding, model for swarming, Discrete Contin. Dyn. Syst., 34 (2014), 1009–1020. https://doi.org/10.3934/dcds.2014.34.1009
    [14] F. Cucker, S. Smale, Emergent behavior in flocks, IEEE Trans. Automat. Control, 52 (2007), 852–862. https://doi.org/10.1109/TAC.2007.895842
    [15] J. G. Dong, S. Y. Ha, D. Kim, From discrete Cucker–Smale model to continuous Cucker–Smale model in a temperature field, J. Math. Phys., 60 (2019), 072705. https://doi.org/10.1063/1.5084770 doi: 10.1063/1.5084770
    [16] J. G. Dong, S. Y. Ha, D. Kim, Emergent behaviors of continuous and discrete thermomechanical Cucker–Smale models on general digraphs, Math. Models Methods Appl. Sci., 29 (2019), 589–632. https://doi.org/10.1142/S0218202519400013 doi: 10.1142/S0218202519400013
    [17] A. Figalli, M. J. Kang, A rigorous derivation from the kinetic Cucker–Smale model to the pressureless Euler system with nonlocal alignment, Anal. PDE., 12 (2019), 843–866. https://doi.org/10.2140/apde.2019.12.843 doi: 10.2140/apde.2019.12.843
    [18] S. Y. Ha, J. Jeong, S. E. Noh, Q. Xiao, X. Zhang, Emergent dynamics of Cucker–Smale flocking particles in a random environment, J. Differ. Equ., 262 (2017), 2554–2591. https://doi.org/10.1016/j.jde.2016.11.017 doi: 10.1016/j.jde.2016.11.017
    [19] S. Y. Ha, M. J. Kang, J. Kim, W. Shim, Hydrodynamic limit of the kinetic thermomechanical Cucker–Smale model in a strong local alignment regime, Commun. Pure Appl. Anal., 19 (2020), 1233–1256. https://doi.org/10.3934/cpaa.2020057 doi: 10.3934/cpaa.2020057
    [20] S. Y. Ha, D. Kim, J. Lee, S. E. Noh, Emergent dynamics of an orientation flocking model for multi-agent system, Discrete Contin. Dyn. Syst., 40 (2020), 2037–2060. https://doi.org/10.3934/dcds.2020105 doi: 10.3934/dcds.2020105
    [21] S. Y. Ha, J. Kim, C. Min, T. Ruggeri, X. Zhang, Uniform stability and mean-field limit of a thermodynamic Cucker–Smale model, Quart. Appl. Math., 77 (2019), 131–176. https://doi.org/10.1090/qam/1517 doi: 10.1090/qam/1517
    [22] S. Y. Ha, J. Kim, T. Ruggeri, Emergent behaviors of thermodynamic Cucker–Smale particles, SIAM J. Math. Anal., 50 (2019), 3092–3121. https://doi.org/10.1137/17M111064 doi: 10.1137/17M111064
    [23] S. Y. Ha, J. G. Liu, A simple proof of Cucker–Smale flocking dynamics and mean-field limit, Commun. Math. Sci., 7 (2009), 297–325. https://doi.org/10.4310/CMS.2009.v7.n2.a2 doi: 10.4310/CMS.2009.v7.n2.a2
    [24] S. Y. Ha, T. Ruggeri, Emergent dynamics of a thermodynamically consistent particle model, Arch. Ration. Mech. Anal., 223 (2017), 1397–1425. https://doi.org/10.1007/s00205-016-1062-3 doi: 10.1007/s00205-016-1062-3
    [25] S. Y. Ha, E. Tadmor, From particle to kinetic and hydrodynamic description of flocking, Kinet. Relat. Models., 1 (2008), 415–435. https://doi.org/10.3934/krm.2008.1.415 doi: 10.3934/krm.2008.1.415
    [26] H. Gayathri, P. M. Aparna, A. Verma, A review of studies on understanding crowd dynamics in the context of crowd safety in mass religious gatherings, Int. J. Disaster Risk Sci., 25 (2017), 82–91. https://doi.org/10.1016/j.ijdrr.2017.07.017 doi: 10.1016/j.ijdrr.2017.07.017
    [27] T. K. Karper, A. Mellet, K. Trivisa, Hydrodynamic limit of the kinetic Cucker–Smale flocking model, Math. Models Methods Appl. Sci., 25 (2015), 131–163. https://doi.org/10.1142/S0218202515500050 doi: 10.1142/S0218202515500050
    [28] T. K. Karper, A. Mellet, K. Trivisa, Existence of weak solutions to kinetic flocking models, SIAM. J. Math. Anal., 45 (2013), 215–243. https://doi.org/10.1137/12086682 doi: 10.1137/12086682
    [29] Z. Li, X. Xue, Cucker–Smale flocking under rooted leadership with fixed and switching topologies, SIAM J. Appl. Math., 70 (2010), 3156–3174. https://doi.org/10.1137/100791774 doi: 10.1137/100791774
    [30] P. B. Mucha, J. Peszek, The Cucker–Smale equation: singular communication weight, measure-valued solutions and weak-atomic uniqueness, Arch. Rational Mech. Anal., 227 (2018), 273–308. https://doi.org/10.1007/s00205-017-1160-x doi: 10.1007/s00205-017-1160-x
    [31] J. Park, H. J. Kim, S. Y. Ha, Cucker–Smale flocking with inter-particle bonding forces, IEEE Trans. Automat. Control, 55 (2010), 2617–2623. https://doi.org/10.1109/TAC.2010.2061070 doi: 10.1109/TAC.2010.2061070
    [32] L. Perea, G. Gomez, P. Elosegui, Extension of the Cucker–Smale control law to space flight formations, J. Guid. Control, 32 (2009), 527–537. https://doi.org/10.2514/1.36269 doi: 10.2514/1.36269
    [33] J. Peszek, Discrete Cucker–Smale flocking model with a weakly singular kernel, SIAM J. Math. Anal., 47 (2015), 3671–3686. https://doi.org/10.1137/15M1009299 doi: 10.1137/15M1009299
    [34] B. Piccoli, F. Rossi, E. Trélat, Control to flocking of the kinetic Cucker–Smale model, SIAM J. Math. Anal., 47 (2014), 4685–4719. https://doi.org/10.1137/140996501 doi: 10.1137/140996501
    [35] L. Ru, X. Li, Y. Liu, X. Wang, Finite-time flocking of Cucker–Smale model with unknown intrinsic dynamics, Discrete Continuous Dyn Syst Ser B, 28 (2023), 3680–3696. https://doi.org/10.3934/dcdsb.2022237 doi: 10.3934/dcdsb.2022237
    [36] C. M. Topaz, A. L. Bertozzi, Swarming patterns in a two-dimensional kinematic model for biological groups, SIAM J. Appl. Math., 65 (2004), 152–174. https://doi.org/10.1137/S0036139903437424 doi: 10.1137/S0036139903437424
    [37] Y. Z. Sun, F. Liu, W. Li, H. J. Shi, Finite-time flocking of Cucker–Smale systems, 34th Chinese Control Conference (CCC), (2015), 7016–7020. https://doi.org/10.1109/ChiCC.2015.7260749
  • This article has been cited by:

    1. Paola Goatin, Chiara Daini, Maria Laura Delle Monache, Antonella Ferrara, Interacting moving bottlenecks in traffic flow, 2023, 18, 1556-1801, 930, 10.3934/nhm.2023040
    2. Abraham Sylla, A LWR model with constraints at moving interfaces, 2022, 56, 2822-7840, 1081, 10.1051/m2an/2022030
    3. Boris Andreianov, Abraham Sylla, Existence analysis and numerical approximation for a second-order model of traffic with orderliness marker, 2022, 32, 0218-2025, 1295, 10.1142/S0218202522500294
    4. Boris Andreianov, Carlotta Donadello, Massimiliano D. Rosini, Entropy solutions for a two-phase transition model for vehicular traffic with metastable phase and time depending point constraint on the density flow, 2021, 28, 1021-9722, 10.1007/s00030-021-00689-5
    5. Boris Andreianov, Abraham Sylla, Finite volume approximation and well-posedness of conservation laws with moving interfaces under abstract coupling conditions, 2023, 30, 1021-9722, 10.1007/s00030-023-00857-9
    6. Boris Andreianov, Massimiliano D. Rosini, Graziano Stivaletta, On existence, stability and many-particle approximation of solutions of 1D Hughes' model with linear costs, 2023, 369, 00220396, 253, 10.1016/j.jde.2023.06.004
    7. Boris Andreianov, Theo Girard, Existence of Solutions for a Class of One-Dimensional Models of Pedestrian Evacuations, 2024, 56, 0036-1410, 3386, 10.1137/23M1550256
    8. Malaya Mohanty, Madhu Lisha Pattanaik, Satya Ranjan Samal, Shivam Shivam, Assessing the Impact of Three-Wheelers on Traffic Flow in India: A Case Study Using Ann, 2024, 13354205, 10.26552/com.C.2024.033
  • Reader Comments
  • © 2024 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(1297) PDF downloads(129) Cited by(0)

Figures and Tables

Figures(2)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog