
Citation: Sandrine Chifflet, Marc Tedetti, Hana Zouch, Rania Fourati, Hatem Zaghden, Boubaker Elleuch, Marianne Quéméneur, Fatma Karray, Sami Sayadi. Dynamics of trace metals in a shallow coastal ecosystem: insights from the Gulf of Gabès (southern Mediterranean Sea)[J]. AIMS Environmental Science, 2019, 6(4): 277-297. doi: 10.3934/environsci.2019.4.277
[1] | Paola Goatin, Matthias Mimault . A mixed system modeling two-directional pedestrian flows. Mathematical Biosciences and Engineering, 2015, 12(2): 375-392. doi: 10.3934/mbe.2015.12.375 |
[2] | Ziqiang Cheng, Jin Wang . Modeling epidemic flow with fluid dynamics. Mathematical Biosciences and Engineering, 2022, 19(8): 8334-8360. doi: 10.3934/mbe.2022388 |
[3] | Sebastien Motsch, Mehdi Moussaïd, Elsa G. Guillot, Mathieu Moreau, Julien Pettré, Guy Theraulaz, Cécile Appert-Rolland, Pierre Degond . Modeling crowd dynamics through coarse-grained data analysis. Mathematical Biosciences and Engineering, 2018, 15(6): 1271-1290. doi: 10.3934/mbe.2018059 |
[4] | Radu C. Cascaval, Ciro D'Apice, Maria Pia D'Arienzo, Rosanna Manzo . Flow optimization in vascular networks. Mathematical Biosciences and Engineering, 2017, 14(3): 607-624. doi: 10.3934/mbe.2017035 |
[5] | Jimin Yu, Jiajun Yin, Shangbo Zhou, Saiao Huang, Xianzhong Xie . An image super-resolution reconstruction model based on fractional-order anisotropic diffusion equation. Mathematical Biosciences and Engineering, 2021, 18(5): 6581-6607. doi: 10.3934/mbe.2021326 |
[6] | Azmy S. Ackleh, Rainey Lyons, Nicolas Saintier . Finite difference schemes for a structured population model in the space of measures. Mathematical Biosciences and Engineering, 2020, 17(1): 747-775. doi: 10.3934/mbe.2020039 |
[7] | Abdulhamed Alsisi, Raluca Eftimie, Dumitru Trucu . Non-local multiscale approach for the impact of go or grow hypothesis on tumour-viruses interactions. Mathematical Biosciences and Engineering, 2021, 18(5): 5252-5284. doi: 10.3934/mbe.2021267 |
[8] | Zhangli Peng, Andrew Resnick, Y.-N. Young . Primary cilium: a paradigm for integrating mathematical modeling with experiments and numerical simulations in mechanobiology. Mathematical Biosciences and Engineering, 2021, 18(2): 1215-1237. doi: 10.3934/mbe.2021066 |
[9] | O. E. Adebayo, S. Urcun, G. Rolin, S. P. A. Bordas, D. Trucu, R. Eftimie . Mathematical investigation of normal and abnormal wound healing dynamics: local and non-local models. Mathematical Biosciences and Engineering, 2023, 20(9): 17446-17498. doi: 10.3934/mbe.2023776 |
[10] | Boris Andreianov, Carlotta Donadello, Ulrich Razafison, Massimiliano D. Rosini . Riemann problems with non--local point constraints and capacity drop. Mathematical Biosciences and Engineering, 2015, 12(2): 259-278. doi: 10.3934/mbe.2015.12.259 |
This paper contributes to the macroscopic modelling of crowd movements. We consider the following initial-boundary value problem for a non-local scalar conservation law that describes the evolution of the local density ρ of pedestrians as a function of time t and position x=(x1,x2) on a crowd evolution domain Ω⊂R2:
{∂tρ+div (ρV(ρ)ν(x,I[ρ(t)](x)))=0,x∈Ω,t≥0,ρ(0,x)=ρ0(x),x∈Ω,ρ(t,x)=0,x∈∂Ω. | (1.1) |
Here ν=(ν1,ν2) is a vector field that (with slight abuse of notation) is defined as
ν(t,x):=ν(x,I[ρ(t)](x))=μ(x)+I[ρ(t)](x), | (1.2) |
where μ is the (normalized) fixed smooth vector field of preferred directions (e.g., given by the regularized solution of an eikonal equation), and I[ρ(t)] is a non-local correction term that depends on the current density distribution. This notation indicates a functional dependence, i.e., I depends on the function ρ(t):=ρ(t,⋅) as a whole. The function V:R+→R represents the average speed of pedestrians as a function of their density.
We assume that pedestrians move in a space surrounded by walls, and that the vector field ν points inward along the boundary ∂Ω of Ω, that is
ν(t,x)⋅n(x)≤0for all x∈∂Ω, t≥0, | (1.3) |
where n is the outward normal to Ω. (Of course, we may assume μ(x)⋅n(x)≤0 for all x∈∂Ω, then it is enough to ensure that also I[ρ(t)](x)⋅n(x)≤0 for all x∈∂Ω.) In this case, the condition ρ(t,x)=0 on ∂Ω corresponds to a zero-flux condition. If, for simulation reasons, we need to consider smaller domains and to add adsorbing conditions on the part of the boundary not corresponding to walls (and where the vector field points outwards), suitable modifications of the model are needed.
If supp(ρ0)⊂Ω and ν(t,x)⋅n(x)≤0 for all x∈∂Ω and t>0, then problem (1.1) is equivalent to the Cauchy problem
{∂tρ+div (ρV(ρ)ν(x,I[ρ(t)](x)))=0,x∈R2,t≥0,ρ(0,x)=ρ0(x),x∈R2. | (1.4) |
Following Colombo et al. [1], we consider a non-local interaction term of the form
I[ρ(t)](x)=−ε∇(η∗wρ(t))(x)√1+‖∇(η∗wρ(t))(x)‖2,0<ε<1, | (1.5) |
where η is a smooth non-negative kernel with compact support such that ∬ and \varepsilon is a model parameter modulating the magnitude of the interaction term. The term (1.5) models how pedestrians account for other pedestrian distribution close to them to correct their path. To better account for the reaction of pedestrians to densities ahead of them, one may consider anisotropic kernels \eta , see e.g., [1] and [2,Appendix D]. To account for the presence of boundaries, and walls (or obstacles) into boundaries, unlike [3,4], we modify the usual convolution product as follows:
\begin{equation} \bigl(\eta \ast_{w} \rho(t) \bigr) ( \mathbf{x}) = \iint_{ \mathbb R^2} \rho_w(t, \mathbf{y}) \eta( \mathbf{x} - \mathbf{y}) \, \mathrm{d} \mathbf{y}, \end{equation} | (1.6) |
where \rho_w: \mathbb R^2\to \mathbb R^+ is defined as
\begin{equation} \rho_w(t, \mathbf{x}): = \begin{cases} \rho(t, \mathbf{x}) & \text{if $ \mathbf{x}\in\Omega$, } \\ R_w &\text{if $ \mathbf{x}\in B(\Omega, d( \operatorname*{supp} \eta))\setminus\Omega$}, \\ 0&\text{elsewhere}, \end{cases} \end{equation} | (1.7) |
with R_{w} \geq R big enough so that (1.3) holds. Here we denote by
d(A) = \sup\bigl\{{\left|{ \mathbf{x}- \mathbf{y}}\right|}\colon \mathbf{x}, \mathbf{y}\in A \bigr\} |
the diameter of a set A\subset \mathbb R^2 and by
B(\Omega, \ell) = \left\{ \mathbf{x}\in \mathbb R^2\colon \inf\limits_{ \mathbf{y}\in\Omega}{\left|{ \mathbf{x}- \mathbf{y}}\right|}\leq \ell \right\} |
the "ball" of radius \ell around \Omega . Furthermore, we define
\begin{equation} M: = \iint_{ \mathbb R^2} \chi_{B(\Omega, d( \operatorname*{supp} \eta ))} ( \mathbf{y}) \, \mathrm{d} \mathbf{y}, \end{equation} | (1.8) |
which is finite if \Omega is bounded ( \chi denotes the characteristic function). See Figure 1.
The presence of high density values at the wall and obstacle locations included in (1.6) and (1.7) is intended to mimic their effect on the pedestrian dynamics. Indeed, in this way the non-local correction term (1.5) "sees" the presence of the wall and deviates the movement from the desired trajectory, thus acting as a discomfort term expressing the tendency of pedestrians to stay away from obstacles.
Remark 1. An explicit condition ensuring \nabla (\eta\ast_w\rho(t))(\mathbf{x})\cdot \mathbf{n}(\mathbf{x})\geq 0 for all \mathbf{x}\in \partial\Omega (and thus \boldsymbol\nu(t, \mathbf{x})\cdot \mathbf{n} (\mathbf{x})\leq 0 ) is the following:
R_w \iint_{\Omega^c} \nabla\eta( \mathbf{x} - \mathbf{y})\cdot \mathbf{n}( \mathbf{x}) \, \mathrm{d} \mathbf{y} \geq R \iint_{\Omega} \bigl(\nabla\eta( \mathbf{x} - \mathbf{y})\cdot \mathbf{n}( \mathbf{x})\bigr)^- \, \mathrm{d} \mathbf{y} \qquad \text{for all} \ \mathbf{x}\in \partial\Omega, |
where f(\mathbf{x})^-: = \max\{-f(\mathbf{x}), 0\} = -\min\{f(\mathbf{x}), 0\} denotes the negative part of a function f and \Omega^c: = \mathbb R^2\setminus \Omega is the complement of \Omega . Indeed we have
\begin{align*} \nabla \bigl(\eta\ast_w\rho(t) \bigr)( \mathbf{x})\cdot \mathbf{n}( \mathbf{x}) = \; & \mathbf{n}( \mathbf{x}) \cdot \iint_{ \mathbb R^2} \rho(t, \mathbf{y}) \nabla\eta( \mathbf{x}- \mathbf{y}) \, \mathrm{d} \mathbf{y} \\ = \; & \iint_{ \mathbb R^2} \rho(t, \mathbf{y}) \nabla\eta( \mathbf{x}- \mathbf{y}) \cdot \mathbf{n}( \mathbf{x}) \, \mathrm{d} \mathbf{y} \\ = \; & \iint_{\Omega^c} R_w \nabla\eta( \mathbf{x}- \mathbf{y}) \cdot \mathbf{n}( \mathbf{x}) \, \mathrm{d} \mathbf{y} + \iint_{\Omega} \rho(t, \mathbf{y}) \nabla\eta( \mathbf{x}- \mathbf{y}) \cdot \mathbf{n}( \mathbf{x}) \, \mathrm{d} \mathbf{y} \\ \geq\; & R_w \iint_{\Omega^c} \nabla\eta( \mathbf{x}- \mathbf{y}) \cdot \mathbf{n}( \mathbf{x}) \, \mathrm{d} \mathbf{y} - R \iint_{\Omega} \bigl(\nabla\eta( \mathbf{x}- \mathbf{y}) \cdot \mathbf{n}( \mathbf{x}) \bigr)^- \, \mathrm{d} \mathbf{y} . \end{align*} |
Macroscopic models of (vehicular and pedestrian) traffic flow are based on balance laws describing the spatio-temporal evolution of averaged quantities such as density and mean velocity. In analogy to vehicular traffic models, macroscopic crowd motion models were introduced starting from the beginning of this century based on scalar conservation laws [5,6], gas dynamics equations [7,8], gradient flow methods [9,10] and time evolving measures [11].
More recently, models consisting in non-local conservation laws in two-space dimension were proposed by different authors [1,3,4,12,13,14,15,16]. Conservation laws with non-local flux functions arise in a large variety of applications, such as traffic flow [17,18,19], sedimentation [20,21], and material flows on conveyor belts [22]. In particular, anisotropic kernels have been introduced in traffic models to account for downstream interactions, see e.g., [17,19]. Physically based derivations of vision-based interaction pedestrian dynamics can be found for example in [11,23,24,25]. We emphasize that none of the previously cited models (with the obvious exception of [1]) fits into the framework of the present approach, which is focused on the appropriate incorporation of boundary conditions and obstacles. That said, we are well aware of a vast body of literature on empirical and numerical studies of pedestrian flows in confined areas that call for the explicit incorporation of such boundaries, see for instance recent proceedings volumes of conferences on Pedestrian and Evacuation Dynamics [26,27,28]. (We refer in Section 4 to some specific contributions within these collections.)
The computation of numerical solutions for non-local models is challenging due to the high non-linearity of the system and the dependence of the flux function on the convolution product. In order to overcome the computational bottleneck, high order schemes have been developed for scalar equation [29] and systems [30,31] in one space dimension. For crowd dynamics models, first order finite volume approximations based on the Lax-Friedrichs scheme have been used in [1,3,4], aiming at demonstrating convergence properties for existence results. Here, we will consider the finite difference weighted essentially non-oscillatory (WENO) schemes developed in [32] to achieve high-resolution spatial accuracy. WENO schemes [33,34,35] are widely employed for the simulation of complex flow fields, their main advantage being their capability to achieve arbitrarily high order formal accuracy in smooth regions while maintaining stable, non-oscillatory and sharp discontinuity transitions.
The remainder of this paper is organized as follows. In Section 2, we deal with the well-posedness of problem (1.4), (1.5)–(1.7) under the hypothesis (1.3), so that (1.4), (1.5)–(1.7) is equivalent to (1.1), (1.5)–(1.7). Section 3 describes the WENO scheme used to compute approximate solutions and Section 4 collects three different numerical tests investigating the model characteristics. Conclusions and perspectives are elaborated in Section 5.
We summarize the assumptions concerning the domain \Omega and the functions V , \boldsymbol{\nu} , \boldsymbol{\mu} and \eta :
\mathbf{(I1)} The domain \Omega\subseteq \mathbb R^2 is a non-empty bounded open set with smooth boundary \partial\Omega , so that the outward normal \mathbf{n}(\mathbf{x}) is uniquely defined for all \mathbf{x}\in \partial\Omega .
\mathbf{(I2)} The speed function V\in\mathbf{C^{2}}(\mathbb R; \mathbb R) is non-increasing, V(0) = V_{\max} and V(R) = 0 for some constants V_{\max}, R > 0 .
\mathbf{(I3)} The vector field of preferred directions \boldsymbol\mu\in (\mathbf{C^{2}}\cap\mathbf{W^{{1, \infty}}})(\mathbb R^2; \mathbb R^2) is defined such that \mathinner{ \mathop{ {{\rm{div}}} } } \boldsymbol\mu \in (\mathbf{W^{{1, 1}}}\cap\mathbf{W^{{1, \infty}}})(\mathbb R^2; \mathbb R) .
\mathbf{(I4)} The vector field \boldsymbol\nu points inward along the boundary \partial\Omega of \Omega , i.e., \boldsymbol\nu(t, \mathbf{x})\cdot \mathbf{n} (\mathbf{x})\leq 0 for all \mathbf{x}\in \partial\Omega , t\geq 0 .
\mathbf{(I5)} The kernel function \eta\in\mathbf{C_c^{3}}(\mathbb R^2; \mathbb R^+) satisfies \iint_{ \mathbb R^2} \eta(\mathbf{x})\; \mathrm{d} \mathbf{x} = 1 .
Solutions of problem (1.4), (1.5)–(1.7), equivalently (1.1), (1.5)–(1.7), are intended in the following sense.
Definition 1. [1,Def. 2.1] For any T > 0 and \rho_0\in \mathbf{L^1}(\mathbb R^2, [0, R]) such that \operatorname*{supp} \rho_0 \subset\Omega , a function \rho\in\mathbf{C^{0}}([0, T], \mathbf{L^1}(\mathbb R^2;[0, R]) is said to be a weak entropy solution to (1.4), (1.5)–(1.7), if it is a Kružkov entropy solution to the Cauchy problem
\begin{equation} \begin{cases} \partial_t \rho + \text{div} \ \left( \rho V(\rho) \boldsymbol\nu(t, \mathbf{x})\right) = 0 & \text{in} \ \mathbb R^2, \\ \rho(0, \mathbf{x}) = \rho_0( \mathbf{x})& \text{in }\ \mathbb R^2, \end{cases} \end{equation} | (2.1) |
i.e., for all \kappa\in \mathbb R and all test functions \phi\in {\mathbf{C}}_{\mathbf{c}}^\infty (]-\infty, T[\, \times \mathbb R^2; \mathbb R^+) there holds
\begin{align} \begin{split} & \int_0^T \int_{ \mathbb R^2} \Bigl\{ {\left|{\rho-\kappa}\right|} \partial_t\phi + \text{sgn}(\rho-\kappa)\bigl(\rho V(\rho) -\kappa V(\kappa)\bigr) \boldsymbol\nu(t, \mathbf{x})\cdot\nabla\phi \Bigr\} \, \mathrm{d} \mathbf{x}\, \mathrm{d}t \\ & - \int_0^T\int_{ \mathbb R^2} \kappa V(\kappa) \text{div} \ \boldsymbol\nu (t, \mathbf{x}) \text{sgn}(\rho-\kappa) \phi \; \mathrm{d} \mathbf{x}\, \mathrm{d} t +\int_{ \mathbb R^2} {\left|{\rho_0( \mathbf{x})-\kappa}\right|}\phi(0, \mathbf{x}) \, \mathrm{d} \mathbf{x} \geq 0. \end{split} \end{align} | (2.2) |
Indeed, by mass conservation, we have that \operatorname*{supp} \rho(t, \cdot) \subset\Omega for all t > 0 , and therefore \rho(t, \mathbf{x}) = 0 for a.e. \mathbf{x}\in\Omega^c , see [1,Proposition 3.1]. Therefore, by abuse of notation, \rho\in{{\mathbf{L}}^\infty } ([0, T]\times\Omega; \mathbb R) can be seen as a Kružkov semi-entropy solution in the sense of [37,Def. 1]: namely, from (2.2) we have
\begin{align*} 0\leq & \int_0^T \int_{\Omega} \Bigl\{ (\rho-\kappa)^\pm \partial_t\phi + \text{sgn}^\pm(\rho-\kappa)\bigl(\rho V(\rho) -\kappa V(\kappa)\bigr) \boldsymbol\nu(t, \mathbf{x})\cdot\nabla\phi \Bigr\} \, \mathrm{d} \mathbf{x}\, \mathrm{d} t \\ & - \int_0^T\int_{\Omega} \kappa V(\kappa) \text{div} \ \boldsymbol\nu (t, \mathbf{x}) \text{sgn}^\pm(\rho-\kappa) \phi\; \, \mathrm{d} \mathbf{x}\, \mathrm{d} t +\int_{\Omega} \bigl(\rho_0( \mathbf{x})-\kappa \bigr)^\pm\phi(0, \mathbf{x}) \, \mathrm{d} \mathbf{x} \\ \leq & \int_0^T \int_{\Omega} \Bigl\{ (\rho-\kappa)^\pm \partial_t\phi + \text{sgn}^\pm(\rho-\kappa)\bigl(\rho V(\rho) -\kappa V(\kappa)\bigr) \boldsymbol\nu(t, \mathbf{x})\cdot\nabla\phi \Bigr\} \, \mathrm{d} \mathbf{x}\, \mathrm{d}t \\ & - \int_0^T\int_{\Omega} \kappa V(\kappa) \text{div} \ \boldsymbol\nu (t, \mathbf{x}) \text{sgn}^\pm(\rho-\kappa)\phi \, \mathrm{d} \mathbf{x}\, \mathrm{d} t +\int_{\Omega} \bigl(\rho_0( \mathbf{x})-\kappa\bigr)^\pm\phi(0, \mathbf{x}) \, \mathrm{d} \mathbf{x} \\ & + {\left\|{f'(\rho) \boldsymbol\nu}\right\|}_{{{\mathbf{L}}^\infty } ([0, T]\times\Omega\times[0, R])} \int_0^T\int_{ \partial\Omega} (-\kappa)^\pm \phi(t, \mathbf{x}) \, \mathrm{d} \mathbf{x}\, \mathrm{d}t, \end{align*} |
where f(\rho): = \rho V(\rho) , s^+ = \max\{s, 0\} , s^- = \max\{-s, 0\} , \text{sgn}^+(s) = \text{sgn}(s^+) and \text{sgn}^-(s) = -\text{sgn}(s^-) .
If \rho\in({{\mathbf{L}}^\infty } \cap \rm{BV})([0, T]\times\Omega; \mathbb R) , then the classical definition introduced by Bardos, Le Roux and Nédélec [38] holds:
\begin{align*} & \int_0^T \int_{\Omega} \Bigl\{ {\left|{\rho-\kappa}\right|} \partial_t\phi + \text{sgn}(\rho-\kappa)\bigl(\rho V(\rho) -\kappa V(\kappa)\bigr) \boldsymbol\nu(t, \mathbf{x})\cdot\nabla\phi \Bigr\} \, \mathrm{d} \mathbf{x}\, \mathrm{d} t \\ & - \int_0^T\int_{\Omega} \kappa V(\kappa) \text{div} \ \boldsymbol\nu (t, \mathbf{x}) \text{sgn}(\rho-\kappa) \phi \, \mathrm{d} \mathbf{x}\, \mathrm{d} t +\int_{\Omega} {\left|{\rho_0( \mathbf{x})-\kappa}\right|}\phi(0, \mathbf{x}) \, \mathrm{d} \mathbf{x} \\ & + \int_0^T\int_{ \partial\Omega} \text{sgn}(\kappa) \Bigl( \text{tr} \rho (t, \mathbf{x}) V \bigl(\text{tr} \rho(t, \mathbf{x}) \bigr) -\kappa V(\kappa) \Bigr) \boldsymbol\nu(t, \mathbf{x})\cdot \mathbf{n}( \mathbf{x}) \phi(t, \mathbf{x}) \, \mathrm{d} \mathbf{x}\, \mathrm{d} t \geq 0, \end{align*} |
where \text{tr}\rho denotes the trace of \rho at the boundary \partial\Omega .
We refer the reader to [36] for a discussion on the different notions of admissible solutions for scalar multi-dimensional initial-boundary value problems and their equivalence.
Under hypotheses (I1)–(I3), the non-local term {\mathcal I} defined by (1.5)-(1.6)-(1.7) satisfies the following estimates for \varepsilon as in (1.5), R_w as in (1.7) and M as in (1.8):
\begin{align} {\left\|{ {\mathcal I}[\rho]}\right\|}_{{{\mathbf{L}}^\infty } } &\leq \varepsilon R_w{\left\|{\nabla\eta}\right\|}_{\mathbf{L^1}}, \end{align} | (2.3) |
\begin{align} {\left\|{ {\mathcal I}[\rho]}\right\|}_{\mathbf{L^1}} &\leq \varepsilon R_w M{\left\|{\nabla\eta}\right\|}_{\mathbf{L^1}}, \end{align} | (2.4) |
\begin{align} {\left\|{\text{div} \ {\mathcal I}[\rho]}\right\|}_{{{\mathbf{L}}^\infty } } & \leq \varepsilon R_w{\left\|{\nabla\eta}\right\|}_{\mathbf{W^{{1, 1}}}}\left(1+R_w^2{\left\|{\nabla\eta}\right\|}_{\mathbf{L^1}}^2\right), \end{align} | (2.5) |
\begin{align} {\left\|{\text{div} \ {\mathcal I}[\rho]}\right\|}_{\mathbf{L^1}} & \leq \varepsilon R_w M {\left\|{\nabla\eta}\right\|}_{\mathbf{W^{{1, 1}}}}\left(1+R_w^2{\left\|{\nabla\eta}\right\|}_{\mathbf{L^1}}^2\right), \end{align} | (2.6) |
\begin{align} {\left\|{\nabla\text{div} \ {\mathcal I}[\rho]}\right\|}_{\mathbf{L^1}} &\leq \varepsilon R_w M{\left\|{\nabla\eta}\right\|}_{\mathbf{W^{{2, 1}}}}\left(1+8 R_w^2 {\left\|{\nabla\eta}\right\|}_{\mathbf{L^1}} {\left\|{\nabla\eta}\right\|}_{\mathbf{W^{{1, 1}}}}\right). \end{align} | (2.7) |
Moreover, for any r_1, r_2\in\mathbf{L^1}(\Omega; [0;R]) there hold
\begin{align} {\left\|{ {\mathcal I}[r_1]- {\mathcal I}[r_2]}\right\|}_{{{\mathbf{L}}^\infty } } &\leq \varepsilon \left(1+R_w^2{\left\|{\nabla\eta}\right\|}_{\mathbf{L^1}}^2\right) {\left\|{\nabla\eta}\right\|}_{{{\mathbf{L}}^\infty } }{\left\|{r_1-r_2}\right\|}_{\mathbf{L^1}}, \end{align} | (2.8) |
\begin{align} {\left\|{ {\mathcal I}[r_1]- {\mathcal I}[r_2]}\right\|}_{\mathbf{L^1}} &\leq \varepsilon \left(1+R_w^2{\left\|{\nabla\eta}\right\|}_{\mathbf{L^1}}^2\right) {\left\|{\nabla\eta}\right\|}_{\mathbf{L^1}}{\left\|{r_1-r_2}\right\|}_{\mathbf{L^1}}, \end{align} | (2.9) |
\begin{align} {\left\|{\text{div}( {\mathcal I}[r_1]- {\mathcal I}[r_2])}\right\|}_{\mathbf{L^1}} &\leq \varepsilon{\left\|{r_1-r_2}\right\|}_{\mathbf{L^1}}{\left\|{\nabla\eta}\right\|}_{\mathbf{W^{{1, 1}}}} \left(1+8R_w^2{\left\|{\nabla\eta}\right\|}_{\mathbf{W^{{1, 1}}}}^2+3R_w^4{\left\|{\nabla\eta}\right\|}_{\mathbf{W^{{1, 1}}}}^4\right), \end{align} | (2.10) |
see [1,Proof of Lemma 3.1].
Recalling that f(\rho) = \rho V(\rho) and following [1,Theorem 2.1], we have the following well-posedness result.
Theorem 1. Let (I1)–(I4) hold and \rho_0 \in (\mathbf{L^1}\cap \rm{BV})(\mathbb R^2;[0, R]) with \operatorname*{supp} \rho_0 \subset\Omega . Then there exists a unique weak entropy solution \rho\in\mathbf{C^{0}}(\mathbb R^+; \mathbf{L^1}(\mathbb R^2;[0, R])) to (1.4), (1.5)–(1.7) with \operatorname*{supp} \rho(t, \cdot) \subset\Omega for t > 0 . Moreover, \rho satisfies the following estimates
\begin{align} &{\left\|{\rho(t, \cdot)}\right\|}_{\mathbf{L^1}} = {\left\|{\rho_0}\right\|}_{\mathbf{L^1}} \quad \text{ for a.e.}\ t \gt 0, \\ & TV \bigl(\rho(t, \cdot)\bigr) \leq TV (\rho_0) \mathrm{e}^{ \mathcal{K} t} + {\pi} t \mathrm{e}^{ \mathcal{K} t} {\left\|{f}\right\|}_{{{\mathbf{L}}^\infty } ([0, R])} \bigl({\left\|{\nabla \mathinner{ \mathop{ {{\rm{div}}} } } \boldsymbol\mu}\right\|}_{\mathbf{L^1}}+C_M(R_w)\bigr), \end{align} | (2.11) |
where we define
\begin{align*} \mathcal{K}&: = 5{\left\|{f'}\right\|}_{{{\mathbf{L}}^\infty } ([0, R])}\left( {\left\|{\text{div} \ \boldsymbol\mu}\right\|}_{{{\mathbf{L}}^\infty } } + \varepsilon R_w{\left\|{\nabla\eta}\right\|}_{\mathbf{W^{{1, 1}}}}\left(1+R_w^2{\left\|{\nabla\eta}\right\|}_{\mathbf{L^1}}^2\right)\right), \\ C_M(R_w)&: = \varepsilon R_w M{\left\|{\nabla\eta}\right\|}_{\mathbf{W^{{2, 1}}}}\left(1+8 R_w^2 {\left\|{\nabla\eta}\right\|}_{\mathbf{L^1}} {\left\|{\nabla\eta}\right\|}_{\mathbf{W^{{1, 1}}}}\right). \end{align*} |
Stability with respect to \rho_0 , v and \boldsymbol\mu also holds from [1,Theorem 2.2].
PROOF OF THEOREM 1. Following [1], we have to check that we fit the required hypotheses. First of all, given any r\in\mathbf{C^{0}}([0, T]; \mathbf{L^1}(\Omega; [0, R])) , we verify that the scalar conservation law
\partial_t \rho + \text{div} \ \bigl( \rho V(\rho) \mathbf{w}(t, \mathbf{x})\bigr) = 0\qquad \text{in }\ \mathbb R^2, |
with \mathbf{w}(t, \mathbf{x}) = \boldsymbol\mu(\mathbf{x})+ {\mathcal I}[r(t)](\mathbf{x}) , satisfies the assumptions of [39,Theorem 5 and Sec. 5.4], and therefore admits a weak entropy solution \rho\in{{\mathbf{L}}^\infty } (\mathbb R^+; \mathbf{L^1}(\Omega; [0, R])) (see [1,Lemma 2.1]). Setting \varphi(t, \mathbf{x}, \rho) = f(\rho)\mathbf{w}(t, \mathbf{x}) , it is easy to check that \varphi, \partial_\rho\varphi, \partial^2_{x_i, \rho}\varphi, \partial^2_{x_i, x_j}\varphi\in\mathbf{C^{0}}([0, T]\times \mathbb R^2\times[0, R]) , \partial_\rho\varphi\in{{\mathbf{L}}^\infty } ([0, T]\times \mathbb R^2\times[0, R]) and \text{div} \ \varphi\in{{\mathbf{L}}^\infty } ([0, T]\times \mathbb R^2\times[0, R]) , thanks to (I2), (I3) and (I4). Moreover, by (2.6), we have that
{\left\|{\text{div} \ \mathbf{w}(t, \cdot)}\right\|}_{\mathbf{L^1}}\leq {\left\|{\text{div} \boldsymbol\mu}\right\|}_{\mathbf{L^1}} + \varepsilon R_w M {\left\|{\nabla\eta}\right\|}_{\mathbf{W^{{1, 1}}}}\left(1+R_w^2{\left\|{\nabla\eta}\right\|}_{\mathbf{L^1}}^2\right) \lt +\infty, |
which guarantees that \rho\in\mathbf{C^{0}}(\mathbb R^+; \mathbf{L^1}(\Omega; [0, R])) . Moreover, by (2.5),
{\left\|{\text{div} \ \partial_\rho\varphi}\right\|}_{{{\mathbf{L}}^\infty } } \leq {\left\|{f'}\right\|}_{{{\mathbf{L}}^\infty } ([0, R])} \left( {\left\|{\text{div} \ \boldsymbol\mu}\right\|}_{{{\mathbf{L}}^\infty } } + \varepsilon R_w{\left\|{\nabla\eta}\right\|}_{\mathbf{W^{{1, 1}}}}\left(1+R_w^2{\left\|{\nabla\eta}\right\|}_{\mathbf{L^1}}^2\right) \right), |
and by (2.7)
{\left\|{\nabla\text{div}( \boldsymbol\mu+ {\mathcal I}[r(t)])}\right\|}_{\mathbf{L^1}} \leq {\left\|{\nabla\text{div} \ \boldsymbol\mu}\right\|}_{\mathbf{L^1}} + C_M(R_w). |
Therefore, by [40,Theorem 2.2], we get (2.11).
Estimates (2.8)–(2.10) ensure that the same fixed point argument used in [1,Proof of Theorem 2.1] can be applied to prove the existence and uniqueness of solutions to (1.4).
We take \Omega = [x_1^{\min}, x_1^{\max}]\times[x_2^{\min}, x_2^{\max}] and denote by \rho:[0, \infty[\, \times\Omega\to[0, R] and
\begin{align*} \mathbf{f} = \mathbf{f} \bigl(t, \mathbf{x}, {\rho}, (\eta\ast_{w}{\rho}) \bigr): = \rho V(\rho) \boldsymbol\nu \bigl( \mathbf{x}, {\mathcal I}[\rho(t)]( \mathbf{x})\bigr) \end{align*} |
the solution and the flux function of problem (1.1)–(1.4). We use a uniform Cartesian grid with nodes (x_1^i, x_2^j) , i = 1, \dots, M_1 and j = 1, \dots, M_2 such that x_1^i = (i-1/2)h , x_2^j = (j-1/2)h , h = (x_1^{\max}-x_1^{\min})/M_1 = (x_2^{\max}-x_2^{\min})/M_2 . This corresponds to M_1\times M_2 grid points \mathbf{x}_{\mathbf{i}}: = (x_1^i, x_2^j) , where \mathbf{i} = (i, j)\in\mathcal{M}: = \{1, \dots, M_1\}\times\{1, \dots, M_2\}\subset {\mathbb N}^2 , and as in [9] we utilize two-dimensional unit vectors \mathbf{e}_1: = (1, 0) and \mathbf{e}_2 : = (0, 1) to address neighbouring grid points \mathbf{x}_{\mathbf{i}+\mathbf{e}_1} = (x_1^{i+1}, x_2^j) and \mathbf{x}_{\mathbf{i}+ \mathbf{e}_2} = (x_1^{i}, x_2^{j+1}) .
We define \mathbf{u}:[0, \infty)\to \mathbb R^{M_1\times M_2} as a solution computed at an instant t in the grid points where
\begin{align*} u_{\mathbf{i}}(t) = \rho(t, \mathbf{x}_{\mathbf{i}}), \quad \mathbf{f}_{\mathbf{i}} = \mathbf{f}\Bigl(t, \mathbf{x}_{\mathbf{i}}, {\rho}(t, \mathbf{x}_{\mathbf{i}}), \bigl(\eta\ast_{w}{\rho(t)}\bigr)(\mathbf{x}_{\mathbf{i}})\Bigr) \quad \text{for $\mathbf{i}\in\mathcal{M}$.} \end{align*} |
In Section 3.2 we will discuss the discretization of the convolution product. Using this notation, we may approximate the solution of (1.1)–(1.4) in semi-discrete form (that is, discrete in space but continuous in time) by a system of ODEs
\begin{equation} \frac{\mathrm{d} \mathbf{u} }{\mathrm{d} t} = \mathcal{C}(\mathbf{u}) \end{equation} | (3.1) |
where \mathcal{C}(\mathbf{u}) represents the spatial discretization of the convective term with entries given by \mathcal{C}(\mathbf{u}) = (\mathcal{C}(\mathbf{u})_{\mathbf{i}})_{\mathbf{i}\in\mathcal{M}} with
\begin{align} \mathcal{C}(\mathbf{u})_{{i}} = -\sum\limits_{l = 1}^2 \frac{1}{h}\Bigl(\hat{f}_{\mathbf{i}+\frac12\mathbf{e}_l}-\hat{f}_{\mathbf{i}-\frac12\mathbf{e}_l} \Bigr), \end{align} | (3.2) |
where \smash{\hat{f}_{\mathbf{i}+\frac12\mathbf{e}_1}} and \smash{\hat{f}_{\mathbf{i}+\frac12\mathbf{e}_2}} are the numerical fluxes, which in this paper will be a fifth-order version. To this end, we require the summands for l = 1 and l = 2 in (3.2) to be of the same order of approximation to \partial f /\partial{x_1} and \partial f /\partial {x_2} , respectively, at \mathbf{x} = \mathbf{x}_{\mathbf{i}} . For upwinding and stability, a flux function f(\rho) is split as follows:
\begin{align} f(\rho) = f^{+}(\rho)+f^{-}(\rho), \quad \text{with $ \frac{\mathrm{d}f^{+} (\rho)}{\mathrm{d}\rho} \geq0$ and $ \frac{\mathrm{d}f^{-} (\rho)}{\mathrm{d}\rho} \leq0$}. \end{align} | (3.3) |
Then each component is approximated separately using its own "wind direction" with respect to \boldsymbol{e}_l . The simple Lax-Friedrichs flux splitting
\begin{align*} f^{\pm}(\rho) = \frac12 \bigl(f(\rho)\pm\alpha \rho\bigr) \end{align*} |
with a suitable viscosity coefficient \alpha > 0 is used in this paper. We herein use
\begin{align*} \alpha_k = \max\limits_{\rho}{\left|{\partial_{\rho} \bigl(\rho V(\rho) \bigr)}\right|} \sup\limits_{ \mathbf{x}\in\Omega} {\left|{ \boldsymbol\nu( \mathbf{x})\cdot \boldsymbol{e}_k}\right|}, \quad k = 1, 2, \end{align*} |
in direction \boldsymbol{e}_k , k = 1, 2 . The numerical fluxes are split accordingly, i.e.,
\begin{equation} \hat{f}_{\boldsymbol{i}+\frac12\mathbf{e}_k} = \mathcal{R}^{+}\bigl( {f}^{+}_{\mathbf{i}+ (-r:r)\mathbf{e}_k}\bigr)+\mathcal{R}^{-}\bigl( {f}^{-}_{\mathbf{i}+ (-r+1:r+1)\mathbf{e}_k}\bigr) , \quad k = 1, 2, \end{equation} | (3.4) |
where \mathcal{R}^{\pm}\bigl({f}_{\mathbf{i}+ (-r:r)\mathbf{e}_k}\bigr) = \mathcal{R}^{\pm}\bigl({f}_{\mathbf{i}-r\mathbf{e}_k}, \dots, {f}_{\mathbf{i}+r\mathbf{e}_k} \bigr) denotes (2r-1) th-order WENO upwind-biased reconstructions for r = 2, 3, 4 , see [33,34,35].
Ghost cells are needed to compute numerical fluxes near the boundary. To handle these cases we use the boundary condition in (1.1) and set u_{\mathbf{i}} = 0 if \mathbf{i} \notin\mathcal{M} and as the vector field \boldsymbol\nu points inward along \partial\Omega , i.e., \boldsymbol\nu\cdot \mathbf{n} (\mathbf{x})\leq 0 for \mathbf{x}\in \partial\Omega , we set
\begin{equation} \hat{f}_{\mathbf{i}+\frac12\mathbf{e}_l} = \begin{cases} \mathcal{R}^{+}\bigl( {f}_{\mathbf{i}+ (-r:r)\mathbf{e}_l}\bigr) & \text{if $\mathbf{x}_{\mathbf{i}} \in \{x_1^{\min}\}\times [x_2^{\min}, x_2^{max}]\cup [x_1^{\min}, x_2^{max}] \times\{x_2^{\min}\}$}, \\ \mathcal{R}^{-}\bigl( {f}_{\mathbf{i}+ (-r:r)\mathbf{e}_l}\bigr) & \text{if $\mathbf{x}_{\mathbf{i}} \in \{x_1^{\max}\}\times [x_2^{\max}, x_2^{max}]\cup [x_1^{\min}, x_1^{max}] \times\{x_2^{\max}\}$}. \end{cases} \end{equation} | (3.5) |
For evacuation problems, to avoid dealing with extended domains, we have to handle a vector field \boldsymbol\nu which points outward at the exit door D\subset \partial\Omega , i.e., \boldsymbol\nu\cdot \mathbf{n} (\mathbf{x}) > 0 for \mathbf{x}\in D . In this case, we set
\begin{align*} \hat{f}_{\mathbf{i}+\frac12\mathbf{e}_l} = \mathcal{R}^{+}\bigl( {f}_{\boldsymbol{i}+ (-r:r)\mathbf{e}_l}\bigr) \quad \text{for $\mathbf{x}_{\mathbf{i}}\in D$}. \end{align*} |
For more details about the implementation of high-order finite difference WENO schemes for crowd dynamics see [9,32].
In order to evaluate the non-local term in (1.5), we take into account that \nabla(\eta\ast_w \rho) = \nabla\eta \ast_w \rho , where the convolution term \ast_w is defined by (1.6). The corresponding convolutions (\partial\eta/\partial x_1)\ast_w \rho and (\partial\eta/\partial x_2)\ast_w \rho are calculated approximately on the discrete grid via a quadrature formula, in our cases a composite Simpson rule. Since \mbox{supp}(\eta)\subset[-n_0h, n_0h]\times[-n_0h, n_0h] for n_0\in {\mathbb N} large enough, any convolution product is given by
\begin{equation*} \bigl(\eta\ast \rho(t) \bigr)(\mathbf{x}_{\mathbf{i}})\approx \sum\limits_{p = -{n_0}}^{n_0}\sum\limits_{q = -{n_0}}^{n_0} h^2c_p c_q \rho(t, \mathbf{x}_{\mathbf{i}-\mathbf{p}})\eta(\mathbf{x}_{\mathbf{p}}), \end{equation*} |
where c_p and c_q are the coefficients in the quadrature rule and \mathbf{p} = (p, q) . This formula for \mathbf{u} = (u_{\mathbf{i}})\in \mathbb R^{M_1\times M_2} and for the convolution product (1.6) can be written as
\begin{equation} (\eta\ast_w u)(\mathbf{x}_{\mathbf{i}}) = \sum\limits_{p = -{n_0}}^{n_0}\sum\limits_{q = -{n_0}}^{n_0} h^2c_pc_q u_{w, \mathbf{i}-\mathbf{p}} \eta(\mathbf{x}_{\mathbf{p}}), \end{equation} | (3.6) |
where u_{w, \mathbf{i}} is a discrete version of the function (1.7) defined by
\begin{equation} u_{w, \mathbf{i}} = \begin{cases} u_{\mathbf{i}}& \text{if $\mathbf{i} \in\mathcal{M}$, } \\ R_w &\text{if $ \mathbf{x}_{\mathbf{i}}\in B(\Omega, d( \operatorname*{supp} \eta))\setminus\Omega$, } \\ 0&\text{elsewhere}. \end{cases} \end{equation} | (3.7) |
Clearly, the discrete convolution (3.6) causes a computational bottleneck. This is a classical problem in scientific computing that is effectively handled by fast convolution algorithms, mainly based on Fast Fourier Transforms [41] (see also [13,14]).
Finally, the semi-discrete scheme (3.1) is discretized by a third-order TVD Runge-Kutta time discretization method [34] that can be specified as follows. Assume that \mathbf{u}^n is the vector of approximate solutions at t = t_n . Then the approximate values \mathbf{u}^{n+1} associated with t_{n+1} = t_n + \Delta t are calculated by
\begin{align} \begin{split} \mathbf{u}^{(1)} & = \mathbf{u}^n+\Delta t \, \mathcal{C}(\mathbf{u}^n), \\ \mathbf{u}^{(2)}& = \frac34 \mathbf{u}^n+\frac14 \mathbf{u}^{(1)} + \frac14 \Delta t\, \mathcal{C}(\mathbf{u}^{(1)}), \\ \mathbf{u}^{n+1}& = \frac13 \mathbf{u}^n+\frac23 \mathbf{u}^{(2)} + \frac23 \Delta t\, \mathcal{C}(\mathbf{u}^{(2)}). \end{split} \end{align} | (3.8) |
The combined space and time discretizations define a fully discrete scheme. Observe that the time evolution is only of third order of accuracy, while the spatial discretization is fifth-order accurate. To have fifth-order accuracy also in time, a two-step SSP Runge-Kutta method should be used.
We aim at investigating the effects of the non-local operator (1.5)–(1.7) from the crowd dynamics modelling point of view. To this end, in the following numerical examples, we solve numerically (1.4)–(1.7) for t\in[0, T] and \mathbf{x}\in\Omega by using the high-resolution numerical scheme described in Section 3. In particular, we consider here a FD-WENO5 scheme with fifth-order of accuracy in space, see [32] for a discussion of high-order finite difference WENO schemes with different orders of accuracy for crowd dynamics. For each iteration, the time step \Delta t in (3.8) is determined by the formula
\frac{\Delta t}{h}\max\{\alpha_1, \alpha_2\} = \frac12 C_{\mathrm{cfl}}. |
We choose C_{\mathrm{cfl}} as the largest multiple of 0.05 that yields oscillation-free numerical solutions. In all numerical tests we have used C_{\mathrm{cfl}} = 0.2 .
In contrast to (1.6)–(1.7), the model proposed in [3,4] uses the following definition of the convolution product in the non-local term (1.5)
\begin{equation} (\eta \ast_{\Omega} \rho(t)) ( \mathbf{x}) = \frac{1}{z( \mathbf{x})}\iint_{\Omega} \rho(t, \mathbf{y}) \eta( \mathbf{x} - \mathbf{y}) \, \mathrm{d} \mathbf{y}, \qquad\text{where}\quad z( \mathbf{x}): = \iint_{\Omega} \eta( \mathbf{x} - \mathbf{y}) \, \mathrm{d} \mathbf{y}. \end{equation} | (4.1) |
We remark that the model in [3,4] also displays non-locality in the speed, but this prevents a global maximum principle from holding. Therefore, here we keep the speed dependence on the density local.
We consider the evacuation problem proposed in [4,Section 2.1], where \Omega = R\setminus (C_1\cup C2) is composed by a room R = [0, 8]\times[-2, 2] with two symmetric columns C_1 = \, ]4.5, 7[\, \times\, ]0.8, 1.5[ and C_2 = \, ]4.5, 7[\, \times\, ]-1.5, -0.8[ that guide the evacuation through the exit door set at D = \{8\}\times\, ]-0.8, 0.8[. The functions and parameters are chosen as
\begin{equation} \begin{array}{ll} V(\rho) = 2\min\{1, \max\{0, (1-\rho)\}\}, & \rho_0( \mathbf{x}) = 0.9\chi_{[0.5, 3]\times[-1.8, 1.8]}( \mathbf{x}), \\ &\\ \varepsilon = 0.6 , \qquad R_w = 1.5 & \eta( \mathbf{x}) = \frac{315}{128\pi l^{18}}(l^4-\| \mathbf{x}\|^4)^4\chi_{[0, l]}( \mathbf{x}), \end{array} \end{equation} | (4.2) |
where \chi_{[0, l]}(\mathbf{x}) denotes the characteristic function. The fixed vector field \boldsymbol\mu(\mathbf{x}) is given by the unit vector tangent to the geodesic from \mathbf{x} to the exit door, see Figure 2a. Since the non-local term defined by (4.1) does not guarantee that the resulting direction of motion points inside the domain ( \boldsymbol\nu(t, \mathbf{x})\cdot \mathbf{n} (\mathbf{x})\leq 0 for \mathbf{x}\in \partial\Omega , t\geq 0 ), we need to add to \boldsymbol\mu a (fixed) discomfort vector field with maximal intensity along the walls as in [3,4], resulting in the vector field showed in Figure 2b.
We display numerical approximations computed with FD-WENO5 scheme with h = 1/80 at times T = 1, 3 and 6 for two kernel supports ( l = 0.45 in Figure 3 and l = 0.9 in Figure 4). We observe that the non-local correction term (1.5) allows pedestrians to "see" the presence of the wall and obstacles, and to deviate the movement from the desired trajectory. For l = 0.45 , we can observe that pedestrians can pass between the obstacles and the wall, as in the Colombo-Rossi model, however this effect is less remarkable for larger kernel supports like l = 0.9 . Indeed, comparing Figures 3 and 4, we can see that a larger kernel support corresponds to a wider discomfort effect, impacting the velocity vector field on larger portions of the walkable domain. More generally, qualitative differences between the two models depend on the parameter choices of R_w and the discomfort vector field. Nevertheless, we remark that our definition of the convolution (1.6)–(1.7) qualitatively captures the discomfort due to the presence of walls and obstacles delimiting the walking domain.
Figure 5 illustrates the effect of the parameter value R_w . Specifically, the scenario of the left column of Figure 5 is the same as in the left column of Figure 3, with l = 0.45 and R_w = 1.5 replaced by R_w = 2 . On the right column of Figure 5, we change also l to l = 0.9 and compare it to Figure 4, left. Larger values of R_w increase the "discomfort" effect exerted by walls and obstacles, as it becomes apparent if one compares, for instance, the widths of the streams of pedestrians to either side and between the obstacles: they are thinner for R_w = 2 than for R_w = 1.5 .
Finally, Figure 6 shows the impact of the parameters R_w and l on the evacuation time for both models. We can observe that the model is quite sensitive to changes in the convolution support since, as previously mentioned, this affects heavily the resulting velocity field near obstacles and walls. Furthermore, the evacuation process for a given value of l is slower for R_w = 2 than for R_w = 1.5 . We mention that the parameters and obstacles have been chosen to illustrate the model and the difference to [3,4], but do not correspond to a specific real-world scenario. There are however, abundant experimental studies providing data on pedestrian flows with obstacles that could form the basis for future comparison with simulations, see e.g., [42,43,44,45].
In this section, we consider that pedestrians have a limited vision field oriented in a given direction. We study a simple example of evacuation of a rectangular room \Omega = [0, 8]\times[-3, 3] , where the vector field \boldsymbol\mu(\mathbf{x}) = (1, 0) is fixed constantly oriented towards the right of the domain. We investigate the influence of a conic convolution kernel on the evacuation dynamics and the pattern formation. We take the kernel function \eta(\mathbf{x}) given in (4.2) and cut a conic section \eta(\mathbf{x})\chi_{\mathcal{S}(\mathbf{x}, l, \alpha, \boldsymbol{\gamma})}(\mathbf{x}) of angle 2\alpha oriented to direction \boldsymbol{\gamma}(\mathbf{x}) which is described by the region
\mathcal{S}( \mathbf{x}, l, \alpha, \boldsymbol{\gamma}) = \left\{ \mathbf{y}\in \mathbb R^2:{\left\|{ \mathbf{y}- \mathbf{x}}\right\|}\leq l, \frac{( \mathbf{y}- \mathbf{x})\cdot\boldsymbol{\gamma}( \mathbf{x})}{{\left\|{ \mathbf{y}- \mathbf{x}}\right\|}{\left\|{\boldsymbol{\gamma}( \mathbf{x})}\right\|}}\geq \cos \alpha\right\} |
(see Figure 7). The section \eta \chi_{\mathcal{S}(\mathbf{x}, l, \alpha, \boldsymbol{\gamma})} is smoothed by convolution with a Gaussian kernel g(\mathbf{x}) = \exp(-(\| \mathbf{x}\|^2/2\sigma)) with \sigma = 5\times 10^{-4} , then normalized and finally shifted by the distance 0.04 to ensure that the maximum of the normalized smoothed kernel is centered in (0, 0) . Other parameters are taken as
V(\rho) = 6\min\{1, \max\{0, (1-\rho)\}\}, \qquad\varepsilon = 0.6, \qquad l = 0.9, \qquad \rho_0( \mathbf{x}) = 0.9\chi_{[0.5, 4]\times[-1, 1]}( \mathbf{x}). |
We compare the dynamics given by different kernel orientations and different angle amplitudes. Besides the circular symmetric kernel \eta (i.e., \alpha = \pi ), we consider
● \boldsymbol{\gamma}(\mathbf{x}) = (-1, 0) , corresponding to forward interaction (by central symmetry of the convolution product), and \alpha = \pi/2, \; \pi/4 ;
● \boldsymbol{\gamma}(\mathbf{x}) = (1, 0) , corresponding to backward interaction, and \alpha = \pi/4 ;
● \boldsymbol{\gamma}(\mathbf{x}) = (0, -1) , modeling pedestrian reacting to the presence of other individuals (or obstacles) on their left, and \alpha = \pi/4 ;
● \boldsymbol{\gamma}(\mathbf{x}) = (-1, -1) , for forward-left interactions, and \alpha = \pi/4 .
The corresponding convolution kernels are depicted in Figure 7. In Figures 8–10, we display numerical approximations and the corresponding vector fields of preferred directions computed with FD-WENO5 scheme with h = 1/80 at times T = 0.1 , T = 0.2 and T = 0.4 . We observe that symmetric and half-circular downstream interaction kernels lead to similar patterns consisting of horizontal lanes (see Figure 8), while narrower angles \alpha lead to vertical patterns, both for forward and backward interactions (displayed in Figure 9). Finally, lateral interactions mostly lead to diagonal waves (Figure 10). Again, the scenario is hypothetical, but could be compared with experimental results from the literature [26,27,28,46].
In this section, we consider the problem of evacuating people in a room through a door. In particular, we are interested in studying the impact of the presence of obstacles in front of the door on the evacuation time. This problem has already been discussed by several authors, see for example [4,47,48] and references therein. In these works, the authors infer that, in some cases, the location and size of the obstacles may speed up the population to the exit. The interest in this phenomenon, sometimes called "faster-is-slower effect" [49,50], goes back to the well-known paper by Helbing et al. [51]. Delays during evacuation due to other effects have been investigated, for instance, in [52,53]. For this example, we consider the walking domain available to pedestrians is \Omega = R\setminus C_{i} , i = 1, 2, 3 , where the room R = [0, 8]\times[-3, 3] contains obstacles C_{i} . The door D , the functions v and \eta , and the parameter \varepsilon and R_w are the same as in (4.2). The initial condition is a linear combination of characteristic functions with values 0.9 in [0.5, 2]\times[-2.2, 0] , 0.6 in [0.5, 2.2]\times[0, 2.2] , 0.5 in [2,4]\times[-2.2, 0] and 0.8 in [2.2, 4]\times[0, 2.2] (see Figure 11). The obstacles and parameter l for the three different scenarios considered are
\begin{align} \begin{split} C_{1} & = \varnothing, \\ C_{2} & = \left] \, 7, 7.8 \, \right[ \times \left( \, \left] \, -1.8 , 1.3 \, \right[ \cup \left] \, 1.3 , 1.8 \, \right[ \, \right), \\ C_{3} & = \left] \, 5, 6 \, \right[ \times \left] \, -0.25 , 0.25\, \right[, \\ l& = 1. \end{split} \end{align} | (4.3) |
The numerical solutions for the three scenarios at three different times is displayed in Figure 12, where we have used the FD-WENO5 scheme to computed the approximate solutions.
Figure 13 shows the time evolution of the mass inside the room. We observe that in case C_1 the evacuation time is higher in comparison to cases C_2 and C_3 , where the evacuation time is lower, especially for case C_3 . This is confirmed by Figure 12, which shows the evacuation of the populations at times T = 6, 20 and 32 , for the cases C_1 , C_2 and C_3 respectively.
In this work, we have proposed and studied a non-local macroscopic pedestrian flow model accounting for the presence of walls and obstacles limiting the walking domain. In particular, the proposed model is able to capture pedestrians' discomfort near obstacles and walls. Under suitable regularity assumptions, the model turns out to be well-posed. Moreover, we analyzed the impact of different anisotropic kernels on the formation of patterns in the solutions. High-resolution numerical schemes of WENO type allow to perform accurate simulations, bypassing the computational bottleneck given by the dependence of the flux function on integral terms. Future research should focus on multi-population models accounting for groups with different characteristics and/or destinations, and on the theoretical analysis of the observed pattern formation.
This research was supported by the INRIA Associated Team "Efficient numerical schemes for non-local transport phenomena" (NOLOCO; 2018–2020). RB is supported by Fondecyt project 1170473 and CRHIAM, project ANID/FONDAP/15130015. LMV is supported by Fondecyt project 1181511. RB, DI and LMV are also supported by CONICYT/PIA/Concurso Apoyo a Centros Científicos y Tecnológicos de Excelencia con Financiamiento Basal AFB170001.
The authors declare there is no conflicts of interest in this paper.
[1] |
Cobelo-Garcia A, Prego R, Labandeira A (2004) Land inputs of trace metals, major elements, particulate organic carbon and suspended solids to an industrial coastal bay of the NE Atlantic. Water Res 38: 1753–1764. doi: 10.1016/j.watres.2003.12.038
![]() |
[2] | Duarte B, Gilda S, Costa JL, et al. (2014) Heavy metal distribution and partitioning in the vicinity of discharge areas of Lisbon drainage basin (Tagus Estuary, Portugal) J Sea Res 93: 101–111. |
[3] |
Oursel B, Garnier C, Pairaud I, et al. (2014) Behaviour and fate of urban particles in coastal waters: settling rate, size distribution and metals contamination characterization. Estuar Coast Shelf S 138: 14–26. doi: 10.1016/j.ecss.2013.12.002
![]() |
[4] |
Eggleton J, Thomas KV (2004) A review of factors affecting the release and bioavailability of contaminants during sediment disturbance events. Environ Int 30: 973–980. doi: 10.1016/j.envint.2004.03.001
![]() |
[5] | Morillo J, Usero J, Gracia I (2007) Potential mobility of metals in polluted coastal sediments in two bays of Southern Spain. J Coastal Res 23: 352–61. |
[6] | Pérez-López R, Álvarez-Valero AM, Nieto JM, et al. (2008) Use of sequential extraction procedure for assessing the environmental impact at regional scale of the São Domingos Mine (Iberian Pyrite Belt) Appl Geochem 23: 3452–63. |
[7] |
Ayata SD, Irisson JO. Aubert A, et al. (2018) Regionalisation of the Mediterranean basin, a MERMEX synthesis. Pr Oceanogr 163: 7–20. doi: 10.1016/j.pocean.2017.09.016
![]() |
[8] |
Bel Hassen M, Drira Z, Hamza A, et al. (2009) Phytoplankton dynamics related to water mass properties in the Gulf of Gabes: ecological implications. J Marine Sys 75: 216–226. doi: 10.1016/j.jmarsys.2008.09.004
![]() |
[9] | Meddeb S (2014) GEF: Governance and knowledge Generation Socio-economic Evaluation of Maritime Activities. Plan Bleu Project ID :P118145, Borrower/Bid N°:FC006, 103p. |
[10] |
Caçador I, Costa JL, Duarte B, et al. (2012) Macroinvertebrates and fishes as biomonitors of heavy metal concentration in the Seixal Bay (Tagus estuary): which species perform better? Ecol Indic 19: 184–190. doi: 10.1016/j.ecolind.2011.09.007
![]() |
[11] | Gargouri D, Azri C, Serbaji MM, et al. (2011) Heavy metal concentrations in the surface marine sediments of Sfax Coast, Tunisia. Environ Monit Assess 175: 514–530. |
[12] |
Ghannem N, Azri C, Serbaji MM, et al. (2011) Spatial distribution of heavy metals in the coastal zone of "Sfax-Kerkennah" plateau, Tunisia. Environ Progress Sustain Energy 30: 221–233. doi: 10.1002/ep.10462
![]() |
[13] |
Ghannem N, Gargouri D, Serbaji MM, et al. (2014) Metal contamination of surface sediments of the Sfax–Chebba coastal line, Tunisia. Environ Earth S 72: 3419–3427. doi: 10.1007/s12665-014-3248-z
![]() |
[14] | Ben Salem Z, Habib A (2016) Assessment of heavy metal contamination levels and toxicity in sediments and fishes from Mediterranean Sea (southern coast of Sfax, Tunisia) Environ Sci Pollut Res 23: 13954–13963. |
[15] |
Sammari C, Koutitonsky VG, Moussa M (2006) Sea level variability and tidal resonance in the Gulf of Gabès, Tunisia. Cont Shelf Res 26: 338–350. doi: 10.1016/j.csr.2005.11.006
![]() |
[16] | Hattour MJ, Sammari C, Ben Nassrallah S (2010) Hydrodynamics of the Gulf of Gabès deducted from the observations of currents and rivers levels. Revue Paralia 3: 13–24. |
[17] | Ben Mustapha KB, Komatsu T, Sammari Ch, et al. (2002) Tunisian megabenthos from infra (posidonia meadows) and circalittoral (coralligenous) sites. Bulletin de l'Institut National des Sciences et Technologies de la Mer de Salammbô 29: 24–36. |
[18] | Boudouresque CF, Bernard G, Pergent G, et al. (2009) Regression of Mediterranean seagrasses caused by natural processes and anthropogenic disturbances and stress: a critical review. Bot Mar 52: 395–418. |
[19] | D'Ortenzio F, Ribera d'Alcalà M (2009) On the trophic regimes of the Mediterranean Sea: a satellite analysis. Biogeosciences 6: 1–10. |
[20] |
Feki-Sahnoun W, Hamza A, Mahfoudi M, et al. (2014) Long-term microphytoplankton variability patterns using multivariate analyses: ecological and management implications. Environ Sci Pollut Res 21: 1–19. doi: 10.1007/s11356-013-1996-z
![]() |
[21] | Fourati R, Tedetti M, Guigue C, et al. (2018) Sources and spatial distribution of dissolved aliphatic and polycyclic aromatic hydrocarbons in surface coastal waters from the Gulf of Gabès (Tunisia, southern Mediterranean Sea) Prog Oceanogr 163: 232–247. |
[22] |
Chamtouri I, Abida H, Khanfir H, et al. (2008) Impact of at-site wastewater disposal systems on the groundwater aquifer in arid regions: case of Sfax city, Southern Tunisia. Environ Geol 55: 1123–1133. doi: 10.1007/s00254-007-1060-8
![]() |
[23] | Dahri N, Atoui A, Abida H, 2014. Environmental impact assessment of a flood control channel in Sfax city, Tunisia. Int J Sci Engin 7: 23–29. |
[24] | Callaert B, Bogaert JVD. (2010) The Taparura project: sustainable coastal remediation and development at Sfax, Tunisia. Terra et Aqua 118: 1–7. |
[25] | Houda B, Dorra G, Chafai A, et al. (2011) Impact of a mixed "industrial and domestic" wastewater effluent on the southern coastal sediments of Sfax (Tunisia) in the Mediterranean Sea. Int J Env Res 5: 691–704. |
[26] |
Naifar I, Pereira F, Zmemla R, et al. (2018) Spatial distribution and contamination assessment of heavy metals in marine sediments of the southern coast of Sfax, Gabes Gulf, Tunisia. Mar Pollut Bull 131: 53–62. doi: 10.1016/j.marpolbul.2018.03.048
![]() |
[27] | Louati A, Elleuch B, Kallel A, et al. (2001) Hydrocarbon contamination of coastal sediments from the Sfax area (Tunisia), Mediterranean Sea. Mar Pollut Bull 42: 445–452. |
[28] |
Quevauviller Ph (1998) Operationally defined extraction procedures for soil and sediment analysis: standardization. TrAC Trend Anal Chem 17: 289–298. doi: 10.1016/S0165-9936(97)00119-2
![]() |
[29] |
Rauret G, Lopez-Sanchez JF, Sahuquillo A, et al. (1999) Improvement of the BCR three step sequential extraction procedure prior to the certification of new sediment and soil reference materials. J Environ Monitor 1: 57–61. doi: 10.1039/a807854h
![]() |
[30] |
Field MP, Cullen JT, Sherrell RM. (1999) Direct determination of 10 trace metals in 50 µL samples of coastal seawater using desolvating micronebulization sector field ICP-MS. J Anal Atom Spectrom 14: 1425–1431. doi: 10.1039/A901693G
![]() |
[31] | Chester R, Stoner JH. (1973) Pb in particulates from the lower atmosphere of the Eastern Atlantic. Nature 245: 27–28. |
[32] |
Duce RA, Hoffman GL, ZoUer WH (1975) Atmospheric trace metals at remote Northern and Southern hemisphere sites: pollution or natural. Science 187: 59–61. doi: 10.1126/science.187.4171.59
![]() |
[33] | Alves-Martins MV, Laut L, Dumeba W, et al. (2017) Sediment quality and possible uses of dredged materials: the ria de Aveiro lagoon mouth area (Portugal) J Sediment Environ 2: 149–166. |
[34] | Müller G. (1969) Index of geoaccumulation in sediments of the Rhine river. Geol J 2: 108–118. |
[35] | Tomlinson D, Wilson J, Harris C, et al. (1980) Problem in Heavy Metals in Estuaries and the Formation of Pollution Index. Helgoländer Meeresunltersuchung 33: 566–575. |
[36] | Reimann C, de Caritat P (1998) Chemical elements in environment. Springer-Verlag, Berlin, 398p. |
[37] |
Reimann C, de Caritat P (2000) Intrinsic flaws of element enrichment factors (EFs) in environmental geochemistry. Envir Sci Tech 34: 5084–5091. doi: 10.1021/es001339o
![]() |
[38] |
Ergin M, Saydam C, Bastürk Ö, et al. (1991) Heavy metal concentrations in surface sediments from the two coastal inlets (Golden Horn Estuary and Izmit Bay) of the northeastern sea of Marmara. Chem Geol 91: 269–285. doi: 10.1016/0009-2541(91)90004-B
![]() |
[39] |
Schiff KC, Weisberg SB (1999) Iron as a reference element for determining trace metal enrichment in Southern California coast shelf sediments. Mar Environ Res 48: 161–176. doi: 10.1016/S0141-1136(99)00033-1
![]() |
[40] |
Karageorgis AP, Katsanevakis S, Kaberi H (2009) Use of enrichment factor for the assessment of Koumoundourou lake, Greece. Water Air Soil Pollut 204: 243–258. doi: 10.1007/s11270-009-0041-9
![]() |
[41] |
Sekabira K, Oryem Origa H, Basamba T, et al. (2010) Assessment of heavy metal pollution in urban stream sediments and its tributaries. Inter J Environ Sci Tech 7: 435–446. doi: 10.1007/BF03326153
![]() |
[42] | Shaw PJA (2003) Multivariate statistic for environmental sciences. Arnold, London, 233. |
[43] | Schaule BK, Patterson CC (1983) Perturbation of the natural lead depth profile in the Sargasso Sea by industrial lead. In: Wong C.S, Boyle E, Bruland K.W. (Eds.), Trace Metals in Sea Water Plenum Press, New York, 497–504. |
[44] |
Duce RA, Liss PS, Merill JT, et al. (1991) The atmospheric input of trace species to the world ocean. Global Biogeochem Cy 5: 193–259. doi: 10.1029/91GB01778
![]() |
[45] |
D'Amore JJ, Al-Abed RS, Scheckel KG, et al. (2005) Methods for speciation of metals in soils: a review. J Environ Qual 34: 1707–1745. doi: 10.2134/jeq2004.0014
![]() |
[46] |
Bacon JR, Davidson CM (2008) Is there a future for sequential chemical extraction? The analyst 133: 25–46. doi: 10.1039/B711896A
![]() |
[47] |
Tessier A, Campbell PGC, Bisson M (1979) Sequential extraction procedures for the speciation of particulate trace metals. Anall Chem 51: 844–851. doi: 10.1021/ac50043a017
![]() |
[48] |
Bryan GW, Langstone WJ (1992) Bioavailability, accumulation and effects of heavy metals in sediments with special reference to United Kingdom estuaries: a review. Environ Pollut 76: 89–131. doi: 10.1016/0269-7491(92)90099-V
![]() |
[49] |
Forstner U, Ahlf W, Calmano W (1989) Studies on the transfer of heavy metals between sedimentary phases with a multi-chamber device: combined effects of salinity and redox potential. Mar Chem 28:145–58 doi: 10.1016/0304-4203(89)90192-8
![]() |
[50] | Calmano W, Hong J, Forstner U (1993) Binding and mobilisation of heavy metals in contaminated sediments affected by pH and redox potential. Water Sci Techno 28: 8–9, 223–235 |
[51] |
Zhuang Y, Allen HE, Fu G (1994) Effect of aeration of sediment on cadmium binding. Environ Toxicol Chem 13: 717–724. doi: 10.1002/etc.5620130505
![]() |
[52] | Ramos L, Hernandez M, Gonzalez MJ (1984) Sequential fractionation of copper, lead, cadmium and zinc in soils from or near Doñana National Park. J Environ Qual 23: 50–57. |
[53] | Lopez-Sanchez JF, Rubio R, Samitier C, et al. (1996) Trace metal partitioning in marine sediments and sludges deposited off the coast of Barcelona (Spain) Water Res 30: 153–159. |
[54] |
Delgado J, Barba-Brioso C, Nieto JM, et al. (2011) Speciation and ecological risk of toxic elements in estuarine sediments affected by multiple anthropogenic contributions (Guadiana saltmarshes, SW Iberian Peninsula): I. Surficial sediments. Sci Total Environ 409: 3666–3679. doi: 10.1016/j.scitotenv.2011.06.013
![]() |
[55] | Fernandez-Leborans G, Novillo A (1994) Experimental approach to cadmium effects on a marine protozoan community. Clean Soil Air Water 22: 19–27. |
[56] |
Fernandez-Leborans G, Olalla, Y. (1999) Toxicity and bioaccumulation of cadmium in marine protozoa communities. Ecotox Environ Safe 43: 292–300. doi: 10.1006/eesa.1999.1793
![]() |
[57] |
Hamzeh M, Ouddane B, Daye M, et al. (2014) Trace metal mobilization from surficial sediments of the Seine river estuary. Water air soil pollut 225: 1878–1893. doi: 10.1007/s11270-014-1878-0
![]() |
[58] | Massolo S, Bignasca A, Sarkar SK, et al. (2012) Geochemical fractionation of trace elements in sediments of Hugli River (Ganges) and Sundarban wetland (West Bengal, India) Environ Monitor Assess 1984: 7561–7577. |
[59] | Bruemer GW, Gerth J, Tiller KG (1988) Reaction kinetics of the adsorption and desorption of nickel, zinc and cadmium by goethite. I. Adsorption and diffusion of metals. J Soil Sci 39: 37–52. |
[60] |
Gerringa L (1990) Aerobic degradation of organic matter and the mobility of Cu, Cd, Ni, Pb, Zn, Fe and Mn in marine sediment slurries. Marine Chem 29: 355–374. doi: 10.1016/0304-4203(90)90023-6
![]() |
[61] | Waeles M, Tanguy V, Lespes G, et al. (2008) Behaviour of colloidal trace metals (Cu, Pb and Cd) in estuarine waters: an approach using frontal ultrafiltrtaion (UF) and stripping chronopotentiometric methods (SCP) Estuar Coast Shelf Sci 80: 538–544. |
[62] | Grousset FE, Quetel CR, Thomas B, et al. (1995) Anthropogenic vs. lithogenic origins of trace elements (As, Cd, Pb, Rb, Sb, Sc, Sn, Zn) in water column particles: northwestern Mediterranean Sea. Mar Chem 48: 291–310. |
[63] |
Dong D, Nelson YM, Lion LW, et al. (2000) Adsorption of Pb and Cd onto metal oxides and organic material in natural surface coatings as determined by selective extractions: new evidence for the importance of Mn and Fe oxides. Water Res 34: 427–36. doi: 10.1016/S0043-1354(99)00185-2
![]() |
[64] | Shul'kin VM, Bogdanova NN (1998) Mobilization of zinc, copper, cadmium and lead in aerated seawater from a suspension of bottom sediments. Mar Chem 38: 620–627. |
[65] |
Banerjee ADK (2003) Heavy metal and solid phase speciation in street dusts of Delhi, India. Environ Pollut 123: 95–105. doi: 10.1016/S0269-7491(02)00337-8
![]() |
[66] |
Crane M, Kwok KWH, Wells C, et al. (2007) Use of field data to support European Water Framework Directive Quality Standards for dissolved metals. Environ S Technol 41: 5014–5021. doi: 10.1021/es0629460
![]() |
[67] | EC (2011) Technical report on common implementation strategy for the Water Framework Directive (2000/60/EC) Technical guidance for deriving Environmental Quality Standards, Guidance document n°27, European Community, 204. |
[68] | OSPAR (2005) Analysis of synergies in assessment and monitoring of hazardous substances, eutrophication, radioactive substances and offshore industry in the North-East Atlantic. Ass Monitor Series, vol. I, n°2005/230, 67. |
[69] |
Papanicolaou F, Antaniou S, Pashalidis I (2009) Experimental and theoretical studies on physic-chemical parameters affecting the solubility of phosphogypsum. J Environ Radioactiv 100: 854–857. doi: 10.1016/j.jenvrad.2009.06.012
![]() |
[70] |
Kuryatnyk T, Angulski da Luz C, Pera JA (2008) Valorization of phosphogypsum as hydraulic binder. J Hazard Mater 160: 681–687. doi: 10.1016/j.jhazmat.2008.03.014
![]() |
[71] |
Tayibi H, Choura M, López FA, et al. (2009) Environmental impact and management of phosphogypsum. J Environ Manage 90: 2377–2386. doi: 10.1016/j.jenvman.2009.03.007
![]() |
[72] |
Hentati O, Abrantes N, Caetano AL, et al. (2015) Phosphogypsum as a soil fertilizer: ecotoxicity of amended soil and elutriates to bacteria, invertebrates, algae and plants. J Hazard Mater 294: 80–89. doi: 10.1016/j.jhazmat.2015.03.034
![]() |
[73] |
Masson M, Blanc G, Schafer J (2006) Geochemical signals and source contributions to heavy metal (Cd, Zn, Pb, Cu) fluxes into the Gironde Estuary via its major tributaries. Sci Total Environ 370: 133–146. doi: 10.1016/j.scitotenv.2006.06.011
![]() |