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

Splitting scheme for a macroscopic crowd motion model with congestion for a two-typed population

  • Received: 01 January 2022 Revised: 01 May 2022 Published: 17 June 2022
  • 35Q92, 49Q22, 92C17

  • We study the extension of the macroscopic crowd motion model with congestion to a population divided into two types. As the set of pairs of density whose sum is bounded is not geodesically convex in the product of Wasserstein spaces, the generic splitting scheme may be ill-posed. We thus analyze precisely the projection operator on the set of admissible densities, and link it to the projection on the set of measures of bounded density in the mono-type case. We then derive a numerical scheme to adapt the one-typed population splitting scheme.

    Citation: Félicien BOURDIN. Splitting scheme for a macroscopic crowd motion model with congestion for a two-typed population[J]. Networks and Heterogeneous Media, 2022, 17(5): 783-801. doi: 10.3934/nhm.2022026

    Related Papers:

    [1] Bertrand Maury, Aude Roudneff-Chupin, Filippo Santambrogio, Juliette Venel . Handling congestion in crowd motion modeling. Networks and Heterogeneous Media, 2011, 6(3): 485-519. doi: 10.3934/nhm.2011.6.485
    [2] Filippo Santambrogio . A modest proposal for MFG with density constraints. Networks and Heterogeneous Media, 2012, 7(2): 337-347. doi: 10.3934/nhm.2012.7.337
    [3] Matthias Erbar, Dominik Forkert, Jan Maas, Delio Mugnolo . Gradient flow formulation of diffusion equations in the Wasserstein space over a Metric graph. Networks and Heterogeneous Media, 2022, 17(5): 687-717. doi: 10.3934/nhm.2022023
    [4] Michael Fischer, Gaspard Jankowiak, Marie-Therese Wolfram . Micro- and macroscopic modeling of crowding and pushing in corridors. Networks and Heterogeneous Media, 2020, 15(3): 405-426. doi: 10.3934/nhm.2020025
    [5] Raimund Bürger, Christophe Chalons, Rafael Ordoñez, Luis Miguel Villada . A multiclass Lighthill-Whitham-Richards traffic model with a discontinuous velocity function. Networks and Heterogeneous Media, 2021, 16(2): 187-219. doi: 10.3934/nhm.2021004
    [6] Nastassia Pouradier Duteil . Mean-field limit of collective dynamics with time-varying weights. Networks and Heterogeneous Media, 2022, 17(2): 129-161. doi: 10.3934/nhm.2022001
    [7] Daniel Roggen, Martin Wirz, Gerhard Tröster, Dirk Helbing . Recognition of crowd behavior from mobile sensors with pattern analysis and graph clustering methods. Networks and Heterogeneous Media, 2011, 6(3): 521-544. doi: 10.3934/nhm.2011.6.521
    [8] Caterina Balzotti, Simone Göttlich . A two-dimensional multi-class traffic flow model. Networks and Heterogeneous Media, 2021, 16(1): 69-90. doi: 10.3934/nhm.2020034
    [9] Jan Friedrich, Simone Göttlich, Annika Uphoff . Conservation laws with discontinuous flux function on networks: a splitting algorithm. Networks and Heterogeneous Media, 2023, 18(1): 1-28. doi: 10.3934/nhm.2023001
    [10] Felisia Angela Chiarello, Paola Goatin . Non-local multi-class traffic flow models. Networks and Heterogeneous Media, 2019, 14(2): 371-387. doi: 10.3934/nhm.2019015
  • We study the extension of the macroscopic crowd motion model with congestion to a population divided into two types. As the set of pairs of density whose sum is bounded is not geodesically convex in the product of Wasserstein spaces, the generic splitting scheme may be ill-posed. We thus analyze precisely the projection operator on the set of admissible densities, and link it to the projection on the set of measures of bounded density in the mono-type case. We then derive a numerical scheme to adapt the one-typed population splitting scheme.



    The modeling of the collective motion of entities - as a human crowd, a population of cells in biology or self-propelled particles in physics - involves a large range of approaches based on the level of description of the particle and its interactions. On the one hand, microscopic models rely on the description of each particle's motion. The population is thus described as a collection of n points moving and interacting in the space. Formally, the system evolves according to a system of coupled ordinary differential equations

    dXidt=vi(t,X1,,Xn)i=1,,n. (1)

    It is said as "Lagrangian", since we follow each trajectory in time. It is tractable for a reasonable number of particles, but it becomes unusable when the size of the system increases since it involves the computation of many interactions (roughly increasing as the square of the total population).

    On the other hand, in macroscopic models the population is described by its density, that is a number of particles by surface unit. The system is thus a partial differential equation on this very density ρ, of the type

    tρ+(ρu)=0 (2)

    where u is the "Eulerian" velocity of the crowd in the space. It allows the description of a system of arbitrarily large number of particles, but does not provide the information of every trajectory.

    Congestion. When modeling the collecting motion of particles, one shall often introduce a congestion constraint in the description of the motion. This congestion encodes the impenetrability of the entities at a microscopic scale. Generically, one distinguishes between a desired velocity (or self-propulsion in physics) and an effective velocity, which is the resulting velocity due to the interactions ruled by congestion. There are two main approaches to introduce congestion on a crowd motion model. On the one hand, soft models rely on a relaxation of the constraint that affects the behaviour of the system as some configurations get close of the saturation. In microscopic space-continuous models, a highly repulsive potential of interaction at short-range is usually introduced in the model to prevent any overlapping between particles for instance in the context of pedestrians [11], or collision of solids, see [2], [23]. For discrete-space microscopic models of motion of pedestrians, cellular automaton based models place people on a given grid [3], [5]. In macroscopic models, the porous media equation [8], [13], [19] can be seen as a gradient flow for the functional

    ρρm(x)dx, (3)

    hence penalizing the high values of the density. A kernel can also be considered [17], by introducing a term of the form

    (Gρ), (4)

    where Gρ is the convolution between the density ρ and a repulsive kernel G(x,y) generically of the form V(xy), for instance V(r)=er22σ2 (see [16] for various choices of V).

    In this article, we will be focused on hard models of congestion, that are models that do not allow any overlapping or violation of the constraint. A microscopic hard model of crowd motion is introduced in [15], in which the desired velocities are projected at every time on the set of admissible velocities, that are the velocities that lead to non-overlapping configurations. This complex model (the particles are driven by a velocity that is the projection on a set that evolves in time) is shown to fit into the framework of the sweeping process, introduced by Moreau [18]. The well-posedness of the model is based on a catching-up, or splitting algorithm which has generically two steps:

    ● Step 1 (prediction): apply the unconstrained velocities during a fixed time step,

    ● Step 2 (correction): project the obtained positions on the set of admissible configurations.

    The macroscopic model of crowd motion introduced in [14] is built on the same approach of projection of the velocities. Given a density on a domain ΩR2 under maximal congestion ρ1, the set of admissible velocities writes

    Cρ={vL2(Ω),v0 where ρ=1}. (5)

    The models then reads

    {tρ+(ρu)=0uargminvCρvU22 (6)

    U being the desired velocity. The well-posedness of the model is established in [14], and it is shown in [15] that the Eq. (6) fits in a sweeping framework.

    Macroscopic multi-type models. As the microscopic models allow natively to consider any heterogeneity in the behaviours of the population, there are few extensions of macroscopic congestion models for a population composed by different types. Multiphase flows have been widely studied in physics, see [4] and [9], even under a total density constraint [6], but may not be applied to crowd (or cell) motion since they involve a phase separation and not a total mixture of the types. A separation of phase has been also considered in order to study the motion of cars structured in a multilane traffic [24]. Conversely, in the model we shall consider we assume a mixture between two densities ρ1,ρ2 with a constraint on the total density ρ1+ρ21, as considered in [13] and [15]. It can be seen as the formal limit as m of a cross-diffusion model with a common pressure term

    p=mm1(ρ1+ρ2)m1, (7)

    that reduces to the classical porous media equation

    tρΔρm=0 (8)

    in the absence of one of the two types.

    In what follows, we study the two-species adaptation of the crowd motion model (6), that writes

    {tρ1+(ρ1u1)=0tρ2+(ρ2u2)=0(u1,u2)=PCρ1,ρ2(U1,U2) (9)

    where U1,U2 are the spontaneous velocities, Cρ1,ρ2 is the set of admissible velocities that transport the densities such that ρ1+ρ2 remains smaller than 1 almost everywhere and PCρ1,ρ2 the orthogonal projection operator on Cρ1,ρ2.

    Optimal transport, Wasserstein spaces and admissible set of densities. In order to define a splitting scheme that shall be the macroscopic counterpart of the microscopic splitting scheme defined above, we need to introduce the formalism of Wasserstein spaces. Being given a convex compact domain Ω of Rd, and two probability measures μ,ν on Ω, one denotes Π(μ,ν) the set of transport plans from μ to ν. The 2-Wasserstein distance between μ and ν is then defined as

    W22(μ,ν)=infγΠ(μ,ν)Ωxy2dγ(x,y). (10)

    It defines a distance on P(Ω) that metrizes the weak convergence of probability measures, at makes P(Ω) a geodesic space, denoted by W2(Ω). We refer to [22] for a general introduction to Optimal Transport and Wasserstein spaces. Under this setting, the splitting scheme introduced in [15] for (6) writes, given a time step τ>0 and ρt a density at time t:

    ● Step 1 (prediction): Transport the density by the desired velocity during τ, with

    μt=(Id+τU)#ρt, (11)

    where # is the pushforward operator.

    ● Step 2 (correction): Project the intermediate density on the set of admissible densities

    K1={ρW2(Ω),ρ1}. (12)

    Numerical schemes for hard congestion macroscopic models. We shall end this introduction by saying a few words on schemes developed to approximate the macroscopic crowd motion model (6). In [14], existence of solutions to (6) is proven by considering the limit of a JKO-scheme. This scheme originally introduced in [12] approximates the model by the iterations

    ρk+1=argminρW2(Ω)ΩDdρ+1K1(ρ)+12τW22(ρ,ρk), (13)

    when the desired velocity is of the form D. As highlighted in [21], there are two main issues when using this algorithm as a numerical scheme. The first one is that it involves the computation of a Wasserstein distance, which is a linear problem under constraint and thus can be long. The use of regularized optimal transport and Sinkhorn algorithm can overcome this difficulty [10], [20]. The second problem is a freezing phenomenon: when implementing the algorithm on a given mesh, for small timesteps the Wasserstein part of (13) becomes large and forbids any mass displacement. One should then work under the inverse of a CFL condition of the type dx<Cτ. This condition forces to use small meshes, resulting on a high increase of the time of computation. The last constraint is that the use of a JKO scheme is possible only if the desired velocity has a specific form that defines a model that has a gradient flow structure, see [22].

    We shall consider in this paper an implementation of the splitting scheme (11) (12) which extends the implementation presented in [15] for a mono-type population. The prediction step (11) is computed by transportation of the cells of the mesh by the desired velocity, in a Lagrangian approach. The correction step (12) is computed by a stochastic algorithm: from each oversaturated cell a random walk starts and dispatches the excess of mass when it reaches a free cell where the density is lower than 1. The analysis of the convergence of this scheme is far beyond the scope of this paper, but it is easy to compute and allows to choose large time steps, providing a fast computation of the scheme for large total time.

    The rest of the paper is organized as follows. In section 2, the model is introduced and a splitting algorithm is derived. In section 3, we analyze the projection operator on the set of admissible pairs of densities and state a well-posedness theorem on the scheme. In section 4, we present an implementation of the numerical algorithm and show some simulations.

    Let Ω be a convex bounded domain of R2. In what follows, we identify a density measure to a function. We want to model the evolution of ρ1,ρ2W2(Ω) by integrating a total density constraint ρ1+ρ21 as it is done for the mono-type model in [14]. We denote by UjL2(Ω,R2) the desired velocity of the type j. Being given two admissible densities, the cone of the admissible velocities writes

    Cρ1,ρ2={u1,u2L2(Ω,R2),(ρ1u1+ρ2u2)0 where ρ1+ρ2=1} (14)

    where "ϕ0 where ρ1+ρ2=1" for ϕ(H1(Ω)) has a weak formulation: for every qH1ρ1+ρ2(Ω),

    ϕ,q0, (15)

    where

    H1ρ1+ρ2(Ω)={qH1(Ω),q0 a.e.,q(1ρ1ρ2)=0}. (16)

    We shall thus study the following model:

    Problem 1. For ρ01,ρ02W2(Ω) such that ρ01+ρ021 and for U1,U2L2(Ω,R2), we look for ρ1,ρ2L2([0,T],L2(Ω)) that satisfy (in a weak sense)

    {tρ1+(ρ1u1)=0tρ2+(ρ2u2)=0(u1,u2)=PCρ1,ρ2(U1,U2)ρ1(0,),ρ2(0,)=ρ01,ρ02, (17)

    where PCρ1,ρ2 is the projection for the L2 norm on Cρ1,ρ2.

    Let us first recall how the existence is proven in [14] in the case of a single type when the desired velocity is of the form U=D. A JKO scheme is introduced, that writes

    ρk+1=argminρW2(Ω)ΩDdρ+1K1(ρ)+12τW22(ρ,ρk), (18)

    with

    K1={ρW2(Ω),ρ1 a.e.} (19)

    the set of admissible densities. The existence of a unique minimizer is obtained by the convexity of K1 for generalized geodesics of base ρk joining two minimizers μ0 and μ1, of the form μt=(tr0+(1t)r1)#ρk (where ri is a transport plan from ρk to μi) and from the strict convexity of the functional defined by (13) along these generalized geodesics. In the 2-dimensional case, the set of admissible densities

    K2={ρ1,ρ2W2(Ω),ρ1+ρ21} (20)

    is no more convex along a generalized geodesic.

    Example 1. Consider for instance ρ1 and ρ2 being the indicator function of the balls B(0,ϵ), B(1,ϵ) for a small ϵ>0 (we can normalize in order to have total masses equal to 1). Let μ1=ν2=ρ1 and μ2=ν1=ρ2. The transports plans from (ρ1,ρ2) to (μ1,μ2) are

    r1(x,y)=r2(x,y)=(x,y) (21)

    and the transport plans from (ρ1,ρ2) to (ν1,ν2) are

    s1(x,y)=(x+1,y)s2(x,y)=(x1,y). (22)

    The generalized geodesic between (μ1,μ2) and (ν1,ν2) with respect to (ρ1,ρ2) is thus

    (αt1,αt2)=(1B(t,ϵ),1B(1t,ϵ)), (23)

    which is not admissible for t=1/2, as one can see on figure 1.

    Figure 1. 

    The interpolation between two opposite configurations of spheres along generalized geodesics. In particular, for t=0.45 the pair (αt1,αt2) is not in K2

    .

    Remark 1. As a matter of fact, K2 is not even geodesically convex in the product space W2(Ω)×W2(Ω) endowed with

    d((μ1,μ2),(ν1,ν2))2=d(μ1,ν1)2+d(μ2,ν2)2. (24)

    Indeed, let (μ1,μ2) and (ν1,ν2)K2, and consider two geodesics of constant speed xt:[0,1]W2(Ω), yt:[0,1]W2(Ω) respectively from μ1 to ν1 and from μ2 to ν2. Then zt=(xt,yt) is a constant speed geodesic from (μ1,μ2) to (ν1,ν2) in W2(Ω)×W2(Ω). In the previous example, the generalized geodesics are thus actual geodesics and do not remain into K2.

    Even in the absence of an existence theorem for problem 1, we can write a splitting scheme for Eq. (17).

    Definition 2.1. Given ρ01,ρ02W2(Ω) such that ρ01+ρ021, U1,U2L2(Ω,R2) and a time step τ>0, we define the splitting scheme as

    μk+1j=(Id+τUj)#ρkj (prediction) (ρk+11,ρk+12)argmin(ρ1,ρ2)K2W22(ρ1,μk+11)+W22(ρ2,μk+12) (correction)  (25)

    From its pushforward structure, the prediction step uniquely defines a pair of intermediate probability measures. It is less straightforward that the projection on K2 is unique. Indeed, even if the argmin of the correction step of (25) is not empty from the minimization of a continuous functional on a compact set in W2(Ω)×W2(Ω), it is unclear that it is a singleton. The next section is dedicated to a special analysis of this correction step in order to get well-posedness of the splitting scheme (25).

    The goal of this section is to show that under tractable assumptions, the splitting scheme (25) is uniquely defined.

    Let us start with a lemma that links the projection on K2 with the projection of the total density on K1.

    Lemma 3.1. Let μ1,μ2W2(Ω). We denote ρ the projection of μ1+μ2 on K1. Let

    (ρ1,ρ2)argmin(ν1,ν2)K2W22(ν1,μ1)+W22(ν2,μ2). (26)

    Then ρ1+ρ2=ρ and

    W22(ρ1+ρ2,μ1+μ2)=W22(ρ1,μ1)+W22(ρ2,μ2). (27)

    Proof. From the definition of ρ, one has

    W22(ρ,μ1+μ2)W22(ρ1+ρ2,μ1+μ2). (28)

    Let T the optimal transport map from ρ1+ρ2 to μ1+μ2, T1 the optimal transport map from ρ1 to μ1 and T2 from ρ2 to μ2. Let γ=(Id,T1)#ρ1+(Id,T2)#ρ2. The marginals of γ are respectively ρ1+ρ2 and μ1+μ2. The quadratic cost associated to γ is W22(ρ1,μ1)+W22(ρ2,μ2). One thus has

    W22(ρ1+ρ2,μ1+μ2)W22(ρ1,μ1)+W22(ρ2,μ2). (29)

    Let ζ be an optimal transport plan (in the sense of Kantorovich) from μ1+μ2 to ρ. One has μ1μ1+μ2, μ2μ1+μ2: denote f and g the corresponding Radon-Nikodym derivatives, and

    dζ1(x,y)=f(x)dζ(x,y)dζ2(x,y)=g(x)dζ(x,y). (30)

    One has ζ1+ζ2=ζ, πx#ζ1=μ1 and πx#ζ2=μ2 (πx and πy are the projection on the coordinates). By denoting νj=πy#ζj for j=1,2, we get

    W22(ρ,μ1+μ2)=Ω2|xy|2dζ=Ω2|xy|2dζ1+Ω2|xy|2dζ2W22(ν1,μ1)+W22(ν2,μ2). (31)

    By optimality of the pair (ρ1,ρ2), one has

    W22(ρ1,μ1)+W22(ρ2,μ2)W22(ρ,μ1+μ2). (32)

    By combining (28), (29), (32), we get W22(ρ,μ1+μ2)=W22(ρ1+ρ2,μ1+μ2). As the projection on K1 is uniquely defined (proposition 2 in [15]), we get ρ=ρ1+ρ2.

    Remark 2. From the previous lemma, we get that the sum of two densities that realize the distance to any pair of densities (μ1,μ2) can be computed by projecting μ1+μ2 on K1. However, there can be many pairs that realize this distance. Consider for instance in dimension 1 the case μ1=μ2=δ0. One has PK1(μ1+μ2)=1[1,1]. Let then be f,gL([1,1]) two nonnegative functions such that f+g=1 almost everywhere. Setting

    ρ1=fdλρ2=gdλ (33)

    where λ is Lebesgue measure on [1,1], (ρ1,ρ2) realizes the distance to (μ1,μ2) in the sense of (26).

    Nevertheless, when every density is absolutely continuous with respect to Lebesgue measure we have uniqueness of the projection.

    Proposition 1. Let μ1,μ2W2(Ω) two absolutely continuous measures with respect to Lebesgue measure. There exists a unique couple (ρ1,ρ2) in K2 that minimizes W22(ρ1,μ1)+W22(ρ2,μ2).

    Proof. As every measure is absolutely continuous, all the transport plans are transport maps (in the sense of Monge). Let T be the unique transport map from μ1+μ2 to ρ:=PK1(μ1+μ2) and ρ1,ρ2 that minimize the distance to K2. Let T1 and T2 the optimal transport maps from μ1 to ρ1 and from μ2 to ρ2. The Tj are uniquely defined on the supports of the μj.

    Let us show that T1=T2 almost everywhere. Out of supp(μ1)supp(μ2), one can modify T1 or T2. Let γ=(Id,T1)#μ1+(Id,T2)#μ2 : the marginals of γ are μ1+μ2 and ρ1+ρ2=ρ from lemma 3.1. The cost of γ is

    Ω|xy|2dγ=Ω|xT1(x)|2dμ1+Ω|xT2(x)|2dμ2=W22(ρ1,μ1)+W22(ρ2,μ2)=W22(ρ,μ1+μ2). (34)

    γ is thus an optimal transport plan between its marginals. By uniqueness one has γ=(Id,T)#(ρ1+ρ2). Let zsupp(ρ1)supp(ρ2). We have

    (z,T1(z))supp(γ)

    (z,T2(z))supp(γ)

    so T1=T2=T a.e., and ρj=T#μj, for j=1,2. The ρj are thus uniquely determined.

    From the previous proof, we get the following corollary that gives the form of the K2 projection from the K1 projection of the sum of the densities:

    Corollary 1. Let μ1,μ2W2(Ω) two absolutely continuous measures with respect to Lebesgue measure. Let T be an optimal transport map from μ1+μ2 to its projection on K1. Then the projection of (μ1,μ2) on K2 is given by (T#μ1,T#μ2).

    Provided that the predicted densities in (25) are absolutely continuous, the correction step is thus uniquely defined. The remaining question is the following: what condition should be assumed to have a pair of predicted densities that are absolutely continuous with respect to Lebesgue measure?

    Let us start with a convenient definition.

    Definition 3.2. We define the partial order on the set of positive measures on Ω by

    μ1μ2μ1μ2 and dμ1dμ21μ2a.e.  (35)

    In other words, μ1μ2 if μ2μ1 defines a positive measure. The following lemma controls the image of Lebesgue measure by the prediction step.

    Lemma 3.3. Let Ω be a (large enough) open set of R2, U a C1 velocity field on Ω and Ω0 a compact set included in Ω. Assume that the partial derivatives of U are bounded by a constant c>0. Then for every τ<(5c)1, denoting λ the Lebesgue measure on R2,

    (Id+τU)#λ|Ω0(1+5cτ)λ|Ω0τUL(Ω0). (36)

    where Ω0τUL(Ω0) is the set of points at distance lower than τUL(Ω0) from Ω0.

    Proof. Let τ>0 and

    f:{Ω0f(Ω0)xx+τU(x). (37)

    f is C1 and onto for τ<c1: let x,yΩ such that f(x)=f(y). One has

    xy=τ10JacU((1t)x+ty)(yx)dt (38)

    so xycτxy.

    Let A a borelian set of Ω.

    (Id+τU)#λ|Ω0(A)=Ω01A((Id+τU)(x))dλ=Ω01A(f(x))|det(Jacf(x))||det(Jacf(x))|dλ (39)

    One has |det(Jacf)|12cτ2c2τ2, and for τ<212c, one gets {112cτ2c2τ21+4cτ+4c2τ2}. If τ(4c)1, one has 1+4cτ+4c2τ21+5cτ and

    (Id+τU)#λ|Ω0(A)(1+5cτ)Ω01A(f(x))|det(Jacf(x))|dλ=(1+5cτ)f(Ω0)1A(x)dλ(1+5cτ)λ(AΩτUL(Ω0)0). (40)

    We are now able to state a well-posedness result for the splitting scheme at two types (25).

    Theorem 3.4. Let U1,U2C1(R2,R2) two velocity fields. Assume that they are bounded by c1>1, and their partial derivatives are bounded by a constant c2. Let ρ01,ρ02 two absolutely continuous measures whose support is included in a ball B(x,r0), τ>0 a time step smaller than (5c2)1, and a total time T. Then, setting Ω=B(x,R) for R large enough, the splitting scheme (25) is uniquely defined for n:=Tτ iterations.

    Proof. From lemma 3.3, the first predicted pair of densities is supported in B(x,r0+τc1) and its sum has a density lower than 1+5c2τ. From proposition 1, the first correction step is well-defined. We shall use the following lemma, whose proof is in appendix.

    Lemma 3.4. Let Ω be a compact set, and μ1,μ2 two absolutely continuous measures, such that μ1μ2. Then PK1(μ1)PK1(μ2).

    From this lemma and corollary 1, one gets that the corrected velocities are supported on the ball B(x,r1), with r1=1+5c2τ(r0+c1τ). One proceed by induction to show that the densities are well-defined at each step and supported at step k by a ball of radius

    rk=(1+5c2τ)k2(r0+c1τ1+5c2τ1)c1τ1+c2τ1. (41)

    Setting R=rn concludes the proof.

    Let us end this theoretical section with a continuity proposition for the projection on K2.

    Proposition 2. Assume that Ω is compact. Then the restriction of the projection on K2 to the set of absolutely continuous measures with respect to Lebesgue measure is continuous.

    Proof. Let (μn1,μn2)n that converges towards (μ1,μ2). Denote

    (ρn1,ρn2)=PK2(μn1,μn2)(ρ1,ρ2)=PK2(μ1,μ2). (42)

    Let Tn an optimal transport map from μn1+μn2 to ρn1+ρn2 and T from μ1+μ2 to ρ1+ρ2. As PK1 is continuous (proposition 2 in [15]), (ρn1+ρn2)n converges towards ρ1+ρ2. Theorem 1.50 in [22] ensures (as Ω is compact) that (Id,Tn)#(μn1+μn2) converges towards (Id,T)#(μ1+μ2). Let ν1 be an accumulation point of (Tn#μn1)n; we still denote (Tn#μn1)n the subsequence converging to ν1. One has

    πy#((Id,Tn)#(μn1+μn2))=Tn#μn1+Tn#μn2. (43)

    On the one hand, πy#((Id,Tn)#(μn1+μn2)) converges to ρ1+ρ2. On the other hand, (Tn#μn1)n converges towards ν1. Therefore,

    lim (44)

    One thus has, by using twice lemma 3.1:

    (45)

    Since is admissible, by uniqueness of the projection on , . Thus has one unique accumulation point in which is compact: it converges towards (and similarly converges towards ).

    In this section we detail the numerical scheme used to approximate the splitting scheme (25). It is largely inspired of the methods developed for the mono-type model in [15], [21].

    We present here two main methods to implement the prediction step. We consider only one density since the predictions are made independently for the two types.

    Finite volume step. In [21], a conservative finite volume scheme in introduced, with an upwind operator to discretize the flux. Given a mesh of size , we start by defining the velocities at the edges of the mesh, as on figure 2. We then introduce the upwind operator that computes the flux flowing through a given edge:

    (46)
    Figure 2. 

    The cell of the mesh in position . The densities are defined inside the cell, whereas the velocities are defined at the edges

    .

    The prediction step reads

    (47)

    This scheme is conservative, stable under the CFL condition . However, it forces choosing small time steps as the mesh shrinks, and it is diffusive as we shall see on the following example in dimension 1.

    Example 2. Consider for a single density subject to a velocity (so and ). One can pick a maximal time step to preserve positivity. At step , one can see that the cells at the right of have some mass. The front of the density is thus traveling at speed . The scheme thus presents a quicker diffusion than the model, that prescribes a front speed equal to .

    Lagrangian transport. In order to have a scheme that can use large time step, we present a Lagrangian version of the prediction step, introduced in [15]. The key idea is to transport the whole cell by computing the image of its center by . One then dispatches the transported cell

    (48)

    on the cells of the mesh, as on figure 3.

    Figure 3. 

    The distribution of the image of the cell by T. We first compute the image of the center, then draw a box of size . The lower left part is lifted on the cell , the upper left part on , the upper right part on and the lower right part on

    .

    Remark 3. This scheme is conservative, stable and preserves the positivity for any time step. The front speed of the density studied in example 2 is now for this scheme , so it can be reduced by considering a large time step or a small mesh size.

    Let us first recall how the projection on is obtained in the mono-type case in [15]. We shall then use corollary 1 to derive an algorithm to compute the projection on .

    Starting from a density discretized on a given squared mesh, the projection on is approximated by the following stochastic algorithm.

    ● Start from a random oversaturated cell and transport the excess of mass with a random walk,

    ● When an undersaturated cell is reached, deliver as much mass as possible,

    ● When the total mass has been distributed, restart the process starting from another oversaturated cell.

    We then average this random projection on a great number of realizations in order to limit the error introduced in the computation. We refer to [15] or the heuristics used to derive this scheme. Let us illustrate on a toy example that in practical situations, despite the randomness of this projection step, the obtained density does not fluctuate much as we compute it several times with independent replicates.

    Example 3. On a 2D squared grid of subdivisions along each direction, we computed several times the projection of an oversatured disc carrying a density on the area

    (49)

    For each computation, we run parallel computations and take the average. We estimate so different projections and then compute the 2-Wasserstein distances between the pairs of projected densities. The histogram of the computed distances is represented on figure 3. One sees that the distances lie in the range , whereas the total mass is and the total moved mass is . The error is thus small regarding the amount of mass moved.

    Whereas the convergence of the previous scheme is far beyond the scope of this paper and could be in itself an interesting research focus, let us recall a partial result of convergence in dimension 1 [21].

    Theorem 4.1 (Theorem 5.2.1 in [21]). Let be a positive density on , and . Denote the mean of the random projection of as the mesh size tends to (we assume it converges). Then

    (50)

    From this very partial convergence result and the stability of the projection illustrated in example 3, we expect this random method to prescribe a projection close to the actual projection and that does not depend much on the random realization of the process.

    The correction step is thus computed with the following scheme :

    Remark 4. Let us compute (51) when is derived from a transport map and . One can compute that in this case

    (52)

    so the exact projection is indeed given by the previous algorithm when one has access to the actual transport map. In the case of regularized transport computed with Sinkhorn's algorithm, the plan is generically not a transport map. However, it is known [7] that the regularized transport plan converges towards the exact transport map as the regularization parameter is small. One thus has that converges towards and therefore that the approximated projection does so.

    The splitting scheme allows the velocities to be more general than two time-constant velocity fields that are generic when using a JKO scheme. In the implementation we shall consider, we allow a mixture of different terms:

    ● a constant velocity field ,

    ● attraction towards an external potential: ,

    ● attraction towards a chemoattractant emitted by the particles: where the concentration satisfies a Segel equation with instantaneous diffusion, i.e.

    (53)

    ● attraction/repulsion between the particles: where is a potential of interaction.

    A full implementation of the previous scheme is available1 with a ready-to-use notebook demo that reproduces every simulation of this section. One can in particular find every parameter and initial condition used to generate these simulations.

    1https://gitlab.math.u-psud.fr/bourdin/macroscopic-cell-motion

    On figure 5 is depicted the case of two saturated discs of radius subject to constant fields that make the densities cross. One sees that the crowd finds a compromise to avoid oversaturation by spreading on a wider area while crossing. The desired velocities chosen are here constants

    (54)
    Figure 4. 

    Distribution of the 2-Wasserstein distances between pairs of estimated projections of the density in example 3

    .
    Figure 5. 

    The motion of two crossing discs. The first column represents the sum of the two densities, and the other two the separated densities. The total time of the simulation is for a timestep and a mesh size . The random projection step is averaged on experiments

    .

    On figure 6, chemoattraction is introduced. After spreading in order to cross, the two densities aggregate back into saturated areas. The desired velocities are taken here of the form

    (55)
    Figure 6. 

    The motion of two crossing discs in the presence of chemoattraction. We chose the same parameters that in the previous simulation, with

    .

    A third example of more complex behavior is illustrated on figure 7. In the absence of external velocities field or potential, the motion is only driven here by chemoattractraction and short-range interactions with the same type. The short range interaction potential has been chosen of the form

    (56)
    Figure 7. 

    Aggregation of a composite crowd driven by chemoattraction and short-range interactions. For , we see small numerical diffusion due to the stochastic projection on . The parameters here are , , , , . The random projection step is averaged on runs and the mesh size is . The time parameters are still and

    .

    We added a repulsion for the chemoattractant of the other type. One sees that the crowd reaches a compromise to separate into distinct phases.

    The desired velocities write here

    (57)

    Proof of lemma 3.4. The proof of the following lemma was suggested by Filippo Santambrogio.

    Lemma 3.4. Let be a compact set, and two absolutely continuous measures, such that . Then .

    Proof. Let two absolutely continuous measures. Let a sequence of strictly convex functions, such that , and . We choose such that converges towards . It is possible to choose , with , as represented on figure 8.

    Figure 8. 

    In solid line, the function . In dashed, its approximation by a strictly convex smooth function . In dotted line is displayed the common limit to and

    .

    We consider the following optimization problems

    (58)

    with the convention if has no density. The optimality conditions write

    (59)

    where is a Kantorovich potential from to and is a constant. We aim at showing . As is strictly increasing, it is sufficient to show that a.e. Denote

    (60)

    One can write

    (61)

    is realized on some such that:

    (62)

    Moreover, for one has

    (63)

    for the transport map from to is of the form . Assume . Then so . Thus,

    (64)

    On the one hand, one has

    (65)

    On the other hand, denote , , , . and are Hessian matrix of convex functions so belong to . Let

    (66)

    The derivative of can be expressed with the comatrix of : . It is nonnegative from (62). We get which is absurd with (64) and (65). One thus has a.e.

    Let us then show the convergence of the . By optimality of , one has

    (67)

    where .

    By comparing and (increasing towards ), we get:

    (68)

    Let an accumulation point of . satisfies:

    (69)

    Let us show that . One will then have by optimality and thus

    (70)

    Let and . Assume . One has

    (71)

    One can assume (by extracting)

    (72)

    Let , one has

    (73)

    For any , , so

    (74)

    We get

    (75)

    Let a probability measure whose density is lower than almost everywhere. By optimality, one has

    (76)

    which is absurd when . Thus and is admissible.



    [1]

    L. Ambrosio, N. Gigli and G. Savaré, Gradient Flows in Metricspaces and in the Space of Proba-bility Measures, Lectures in Mathematics ETH Zürich. Birkhäuser Verlag, Basel, 2005.

    [2]

    T. M. Blackwell and P. Bentley, Don't push me! Collision-avoiding swarms, Proceedings of the 2002 Congress on Evolutionary Computation. CEC'02 (Cat. No. 02TH8600), IEEE, 2 (2002).

    [3] Cellular automata microsimulation for modeling bi-directional pedestrian walkways. Transportation Research Part B: Methodological (2001) 35: 293-312.
    [4]

    C. E. Brennen, Fundamentals of Multiphase flow, 2005.

    [5]

    C. Burstedde, et al., Simulation of pedestrian dynamics using a two-dimensional cellular automaton, Physica A: Statistical Mechanics and its Applications, 295 (2001), 507–525.

    [6] Incompressible immiscible multiphase flows in porous media: A variational approach. Anal. PDE (2017) 10: 1845-1876.
    [7] Convergence of entropic schemes for optimal transport and gradient flows. SIAM J. Math. Anal. (2017) 49: 1385-1418.
    [8] Finite speed of propagation in porous media by mass transportation methods. C. R. Math. (2004) 338: 815-818.
    [9] (2005) Multiphase Flow Handbook.CRC press.
    [10] Sinkhorn distances: Lightspeed computation of optimal transport. Advances in Neural Information Processing Systems (2013) 26: 2292-2300.
    [11] Social force model for pedestrian dynamics. Physical review E (1995) 51: 4282.
    [12] The variational formulation of the Fokker–Planck equation. SIAM J. Math. Anal. (1998) 29: 1-17.
    [13] On nonlinear cross-diffusion systems: An optimal transport approach. Calc. Var. Partial Differential Equations (2018) 57: 1-40.
    [14] A macroscopic crowd motion model of gradient flow type. Math. Models Methods Appl. Sci. (2010) 20: 1787-1821.
    [15] Handling congestion in crowd motion modeling. Netw. Heterog. Media (2011) 6: 485-519.
    [16] A non-local model for a swarm. J. Math. Biol. (1999) 38: 534-570.
    [17] An interacting particle system modelling aggregation behavior: From individuals to populations. J. Math. Biol. (2005) 50: 49-66.
    [18] Evolution problem associated with a moving convex set in a Hilbert space. J. Differential Equations (1977) 26: 347-374.
    [19] The geometry of dissipative evolution equations: The porous medium equation. Comm. Partial Differential Equations (2001) 26: 101-174.
    [20] Entropic approximation of Wasserstein gradient flows. SIAM J. Imaging Sci. (2015) 8: 2323-2351.
    [21]

    A. Roudneff-Chupin, Modélisation macroscopique de mouvements de foule, Phdthesis, PhD Thesis, Université Paris-Sud XI, 2011.

    [22]

    F. Santambrogio, Optimal Transport for Applied Mathematicians, Birkhäuser/Springer, Cham, 2015.

    [23] Focusing in collision problems in solids. J. Appl. Physics (1957) 28: 1246-1250.
    [24] Macroscopic dynamics of multilane traffic. Physical Review E (1999) 29: 6328.
  • Reader Comments
  • © 2022 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(902) PDF downloads(165) Cited by(0)

Figures and Tables

Figures(8)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog