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

High dimensional Riesz space distributed-order advection-dispersion equations with ADI scheme in compression format


  • Received: 30 December 2021 Revised: 15 February 2022 Accepted: 13 March 2022 Published: 23 March 2022
  • A second order alternating direction implicit scheme for time-dependent Riesz space distributed-order advection-dispersion equations is applied to higher dimensions with the Tensor-Train decomposition technique. The solutions are solved in compressed format, the Tensor-Train format, and the errors accumulated due to compressions are analyzed to ensure convergence. Problems with low-rank data are tested, the results illustrated a steeper growth in the ranks of the numerical solutions than that in related works.

    Citation: Lot-Kei Chou, Siu-Long Lei. High dimensional Riesz space distributed-order advection-dispersion equations with ADI scheme in compression format[J]. Electronic Research Archive, 2022, 30(4): 1463-1476. doi: 10.3934/era.2022077

    Related Papers:

    [1] Robert Reynolds, Allan Stauffer . Extended Prudnikov sum. AIMS Mathematics, 2022, 7(10): 18576-18586. doi: 10.3934/math.20221021
    [2] Mohamed Akel, Muajebah Hidan, Salah Boulaaras, Mohamed Abdalla . On the solutions of certain fractional kinetic matrix equations involving Hadamard fractional integrals. AIMS Mathematics, 2022, 7(8): 15520-15531. doi: 10.3934/math.2022850
    [3] Robert Reynolds, Allan Stauffer . Note on an integral by Fritz Oberhettinger. AIMS Mathematics, 2021, 6(1): 564-568. doi: 10.3934/math.2021034
    [4] Robert Reynolds, Allan Stauffer . Derivation of some integrals in Gradshteyn and Ryzhik. AIMS Mathematics, 2021, 6(2): 1816-1821. doi: 10.3934/math.2021109
    [5] Maliha Rashid, Amna Kalsoom, Maria Sager, Mustafa Inc, Dumitru Baleanu, Ali S. Alshomrani . Mellin transform for fractional integrals with general analytic kernel. AIMS Mathematics, 2022, 7(5): 9443-9462. doi: 10.3934/math.2022524
    [6] Robert Reynolds, Allan Stauffer . A quintuple integral involving the product of Hermite polynomial $ H_{n}(\beta x) $ and parabolic cylinder function $ D_{v}(\alpha t) $: derivation and evaluation. AIMS Mathematics, 2022, 7(5): 7464-7470. doi: 10.3934/math.2022418
    [7] Shahid Mubeen, Rana Safdar Ali, Iqra Nayab, Gauhar Rahman, Thabet Abdeljawad, Kottakkaran Sooppy Nisar . Integral transforms of an extended generalized multi-index Bessel function. AIMS Mathematics, 2020, 5(6): 7531-7547. doi: 10.3934/math.2020482
    [8] D. L. Suthar, A. M. Khan, A. Alaria, S. D. Purohit, J. Singh . Extended Bessel-Maitland function and its properties pertaining to integral transforms and fractional calculus. AIMS Mathematics, 2020, 5(2): 1400-1410. doi: 10.3934/math.2020096
    [9] Robert Reynolds . A short note on a extended finite secant series. AIMS Mathematics, 2023, 8(11): 26882-26895. doi: 10.3934/math.20231376
    [10] Waleed Mohamed Abd-Elhameed, Ahad M. Al-Sady, Omar Mazen Alqubori, Ahmed Gamal Atta . Numerical treatment of the fractional Rayleigh-Stokes problem using some orthogonal combinations of Chebyshev polynomials. AIMS Mathematics, 2024, 9(9): 25457-25481. doi: 10.3934/math.20241243
  • A second order alternating direction implicit scheme for time-dependent Riesz space distributed-order advection-dispersion equations is applied to higher dimensions with the Tensor-Train decomposition technique. The solutions are solved in compressed format, the Tensor-Train format, and the errors accumulated due to compressions are analyzed to ensure convergence. Problems with low-rank data are tested, the results illustrated a steeper growth in the ranks of the numerical solutions than that in related works.



    In this paper we deal with microscopic modeling of traffic flow, focusing on lane changing dynamics. In particular we study a second order model for one lane that combines two different interaction terms and we describe the extension to the multi-lane case giving particular attention at the two-lane case.

    The interest in the dynamics of traffic flow dates back to the first half of the twentieth century and the related mathematical literature is quite large. An overall view can be found, for instance, in the book by Haberman [10] and in the survey paper by Helbing [11].

    There are various points of view for modeling traffic flow. In this paper we concentrate on the microscopic approach that is based on the dynamics of individual vehicles considering the individual behaviour of each driver. A typical microscopic model is the Car Following model or Follow the Leader model (FtL) based on the idea that the dynamics of each vehicle (follower) depends on the vehicle in front (leader) and therefore the other vehicles do not affect it. These models are normally for single-lane roads [4,6,14]. A typical Follow the Leader model can be described as follows. In a single-lane with N vehicles where overtaking is not allowed, we are interested in study the position xn(t) and the velocity vn(t) of each vehicle n=1,,N at different times t. This dynamics can be described by a system of ordinary differential equations:

    {˙xn(t)=vn(t)n=1,,N˙vn(t)=a(xn(t),xn+1(t),vn(t),vn+1(t))n=1,,N1˙vN=w(t) (1)

    where a() is a given acceleration function and w() is the dynamics of the leader vehicle, independent from the other vehicles (followers).

    Many single-lane car following models have been developed and applied to study traffic dynamics. Here we recall some models that will be useful in the following.

    The Follow the Leader model, introduced in [26,27], assumes that each vehicle modifies its velocity based on the distance (headway) xn+1xn to the vehicle ahead, the n+1-th, and to the difference in velocities between its own velocity vn and the velocity of the vehicle ahead vn+1, multiplied by appropriate coefficients βn. This model can be described by the following system

    {˙xn(t)=vn˙vn(t)=βnvn+1vn(xn+1xn)2. (2)

    The optimal velocity model (OVM) of Bando et al. [3,2] in which a driver aims to a desired velocity function V that depends on the headway with the vehicle ahead. The equation of this model is given by

    {˙xn(t)=vn˙vn(t)=αn(V(xn+1xn)vn) (3)

    with appropriate coefficients αn.

    We mention also some interesting works. Pipes proposed [25] a traffic model in which each vehicle maintains a certain prescribed "following distance" from the preceding vehicle; the generalized force model (GFM) by Helbing and Tilch [13] in which the optimal velocity function is obtained calibrating the parameters with the observed data; the full velocity difference model (FVDM) by Jiang et al. [17] that predicts delay time of car motion and kinematic wave speed at jam density; the optimal velocity difference model (OVDM) by Peng et al. [24] where a new term is introduced involving the optimal velocity functions and the vehicles n,n+1,n+2. Aw et al. [1] studied the derivation of a continuum model starting from the FtL model. We mention an analytical study for the OVM with a stepwise specification of the optimal velocity function and a simple kind of perturbation in [12].

    Another type of microscopic model is given by lane changing models which provide for the possibility of changing lanes according to the analysis of some factors that intervene in the decision process, for example the need, opportunity and safety of a lane change [7,29]. The interest in modeling vehicle lane changing is due to the effects that it induces in traffic flow, for instance in bottleneck discharge rate and in the stop & go oscillations. Here we recall some works. Cassidy and Rudjanakanoknad [5] showed that when traffic density upstream of a busy merge increases beyond a critical value, vehicles manoeuvre toward faster lanes causing traffic breakdown and "capacity drop" of the road; Zheng et al. [30] showed that lane changing are responsible for transforming subtle localized oscillations to substantial disturbances; Klar and Wegener [21,20] developed a model based on reaction thresholds from which they derived a kinetic model; Song and Karni [28] proposed a macroscopic model in which the acceleration terms take lead from microscopic car-following models, and yield a non-linear hyperbolic system with viscous and relaxation terms; Herty et al. [15] proposed a macroscopic model, which accounts for lane-changing on motorway, based on a two-dimensional extension of the Aw and Rascle and Zhang macroscopic model for traffic flow; Gong et al. [9] presented a finite dimensional hybrid system based on the continuous Bando-Follow-the-Leader dynamics coupled with discrete events due to lane changing; Goatin and Rossi [8] developed a macroscopic model for multi-lane road networks with discontinuities both in the speed law and in the number of lanes; Hodas and Jagota presented in [16] a microscopic model for multi-lane dynamics where each car experiences a force resulting from a combination of the desire of the driver to attain a certain velocity and change of the force due to cars interactions; Kesting et al. [19] proposed a general model to derive lane changing rules for discretionary and mandatory lane changes for a wide class of car following models; Lv et al. in [22] extended the continuous single-lane models to simulate the lane changing behaviour on an urban roadway with three lanes and in [23] proposed a model where lane changing is not instantaneous but is a continuing process which can affect the following cars; Zheng et al. in [31] analysed the effects of lane changing in the driver behaviour.

    This paper proposes the study of a second order microscopic model combining models 2 and 3 for reproducing traffic flow and its extension to the multi-lane case with simple lane changing conditions in order to study its stability under perturbations. In Section 2 we introduce the model for a single-lane and we study its stability in the linearized case, then we show numerical tests making comparisons with model 3. In section 3 we describe the extension of the model to the two-lane case studying its stability around the equilibrium when a lane is perturbed. We present some numerical tests that confirm the predictions of the linear stability analysis. Finally, in section 4, we illustrate the generalization of the model to the generic multi-lane case.

    In this section we describe the main mathematical model we use in this paper. Consider a homogeneous population of NN vehicles, and denote by xn=xn(t) and vn=vn(t) the position and the velocity of the n-th vehicle at time tR+. We want to describe the traffic flow in a road with a single-lane where overtaking is not allowed.

    The dynamical equations of the system are obtained combining two interaction terms. The first one is the interaction term related to the model 3 [2,3]. It is a relaxation term towards a desired velocity function V() that depends only on the headway Δxn=xn+1xn>0 between the vehicle n and the vehicle ahead with index n+1, as shown in Fig. 1. The acceleration of each vehicle is regulated by the difference between its velocity and the optimal velocity. The optimal velocity function is typically a monotonically increasing function of the headway and it is bounded. It tends to zero for small headways and to a maximum value Vmax for large headways. Furthermore we assume that V is non-negative. This term is multiplied by a parameter αn denoting the speed of response of each driver, with dimensions one over time. The second term is the classical Follow-the-Leader interaction term [26,27] from model 2, multiplied by a parameter βn with dimensions length square over time. In this term the acceleration of a vehicle is directly proportional to the difference between the velocity of the vehicle in front and its own and is inversely proportional to their mutual distance.

    Figure 1. 

    Vehicles in single-lane road

    .

    Since we are considering identical vehicles we assume αn=α and βn=β for all n=1,,N.

    The model is given by

    {˙xn=vn˙vn=α(V(Δxn)vn)+βΔvn(Δxn)2 (4)

    with Δxn=xn+1xn and Δvn=vn+1vn.

    In our study we usually refer to a circular road which means to solve 4 with periodic boundary conditions, in this way the vehicle with index n=N+1 coincides with vehicle with index n=1. If we deal with a straight road we simply add an equation describing the dynamics of the leader vehicle, which must be known.

    Let us characterize the equilibrium for the single-lane model.

    Proposition 1. The equilibrium of the system 4 is given if all vehicles are equally spaced and move with the same constant velocity.

    In fact, let us indicate with h=LN the constant spacing of two successive vehicles, where L>0 is the length of the road. Then solving 4 with initial conditions

    {xn+1(0)xn(0)=hvn(0)=V(h)for   n=1,,N (5)

    with xN+1()=x1() by boundary conditions, we easily obtain the solution of the system that represents the steady state described above:

    ˉxn(t)=hn+V(h)t. (6)

    Note that the equation depends parametrically by the given number N of vehicles which is constant due to the periodic boundary conditions.

    Now we study the stability of model 4 around the equilibrium 6 by linearizing the original system. Let yn be a small perturbation from the steady state 6 and consider

    xn=ˉxn+yn. (7)

    Disregarding terms higher than O(y2n) we obtain the linearized equation of 4

    ¨yn=α(V(h)Δyn˙yn)+βΔ˙ynh2 (8)

    where Δyn=yn+1yn and Δ˙yn=˙yn+1˙yn, again vehicle with index n=N+1 coincides with the vehicle with index n=1.

    We solve 8 looking for solutions

    yk(n,t)=exp{iakn+zt} (9)

    where eiakn is the Fourier coefficient with ak=2πNk, k=0,,N1 and zC. Substituting in 8 we obtain an equation for z=u+iv

    z2+z(αβh2(eiak1))αV(h)(eiak1)=0. (10)

    If the amplitude of yk(n,t) blows up in time then the solution is unstable, so in order to find stable solutions we require that (z)=u<0.

    Let us write the two solutions of 10 as zj=uj+ivj for j=1,2, then the following relations holds:

    (z1+z2)=u1+u2=α+βh2(cos(ak)1)(z1+z2)=v1+v2=βh2sin(ak)(z1z2)=u1u2v1v2=αV(h)(cos(ak)1)(z1z2)=u1v2v1u2=αV(h)sin(ak).

    The boundary of the stability region is obtained when u1=0 then

    v1=αV(h)sin(ak)α+βh2(cos(ak)1).

    After some algebraic manipulations we get

    V(h)=α2cos2(ak2)+βh2+2tan2(ak2)βh2(βαh2+1). (11)

    We can study this problem with polar coordinates in the (αk,V(h)) plane as shown in Fig. 2. The plane (V(h),ak) can be divided into two regions: a stable region (u<0) and an unstable one (u>0) by the critical curve u(ak,V(h))=0 express by 11. We observe that equation 11 coincides with the curve found in [3] if β=0. The curve 11 is represented by the red line while the black curve is the critical curve of model 3.

    Figure 2. 

    Red: curve 11 in the (αk,V(h)) polar coordinate plane. Black: critical curve of model 3

    .

    Thus we have proved the following result.

    Proposition 2. If

    V(h)<α2+βh2 (12)

    the steady state 6 of model 4 is stable, because for all k we have u<0; if V(h)=α2+βh2 we have a marginal state; while for V(h)>α2+βh2 the model is unstable, because there exists at least one index k such that u>0.

    For β=0 the condition 12 is consistent with the stability condition derived in [3]. Remembering that h=LN the previous condition expresses that we gain more stability with a large number of vehicles.

    Now we present some numerical tests of model 4 using the Runge Kutta 5 method, with time step Δt=0.1 s.

    Let us fix α=1 s1, β=100 m2/s, L=1500 m, and consider the desired velocity function expressed by

    V(Δx)=max{0,VHT(Δx)} (13)

    see Fig 3, where

    VHT(Δx)=V1+V2tanh(C1(Δxlc)C2) (14)
    Figure 3. 

    V() function

    .

    is the function given by Helbing and Tilch in [13] where they carried out a calibration of model 3 respect to the empirical data, obtaining the optimal parameter values V1=6.75 m/s, V2=7.91 m/s, C1=0.13 m1, C2=1.57 and lc=5 m is the length of the vehicles. Velocity parameters V1,V2 determine the minimum expected speed V1V2 and the maximal expected speed V1+V2, while C1,C2 are calibration parameters. Thus Vmax=14.66 m/s.

    From condition 12 we obtain that the model 4 with velocity 13 is stable if h<10.14 m and h>24 m as shown in Fig. 4. In terms of number of vehicles along the circular road we have stability for N<68 and N>100. Note that, with the same parameters, the model 3 is stable for N<62 and N>147.

    Figure 4. 

    V() function in blue, model 4 stability condition in red, model 3 stability condition in black

    .

    In the next two simulations we show a comparison between model 4 and model 3, perturbing the system adding or removing a vehicle. The initial number of vehicles is chosen in such a way that the model 4 is stable while the model 3 is unstable according their stability condition.

    In this simulation we consider N=120 vehicles at the equilibrium 6, equispaced with distance LN=12.5 m and with velocities equal to V(LN). At time t=0 s we perturb the system adding a one new vehicle inserting it in the position 12(xN+L) with initial velocity equal to V(LN). The final time is T=1000 s.

    Model 4:

    Figure 5. 

    On the left: all vehicles trajectories, on the right: velocity of vehicle 1

    .

    Model 3:

    Figure 6. 

    On the left: all vehicles trajectories, on the right: velocity of vehicle 1

    .

    We can see how the perturbation is absorbed in the in first model while it causes a creation of stop & go waves in the second model.

    In this simulation we consider again N=120 vehicles at the equilibrium 6, equispaced with distance LN and with velocities equal to V(LN). At time t=0 s we perturb the system removing one vehicle choosing the one with index N. We set the final time T=1000 s.

    Model 4:

    Figure 7. 

    On the left: all vehicles trajectories, on the right: velocity of vehicle 1

    .

    Model 3:

    Figure 8. 

    On the left: all vehicles trajectories, on the right: velocity of vehicle 1

    .

    Also in this test we can observe the differences when a perturbation occurs in the two models.

    In this simulation we start with N=90 vehicles at the equilibrium 6, equispaced with distance LN16.66 m and with velocities equal to V(LN). At time t=0 s we perturb the system adding a new vehicle as in the previous simulations. We set T=1000 s.

    Model 4:

    Figure 9. 

    On the left: all vehicles trajectories, on the right: velocity of vehicle 1

    .

    Model 3:

    Figure 10. 

    On the left: all vehicles trajectories, on the right: velocity of vehicle 1

    .

    An example of instability for both models is reported. Although stop & go waves occur we can appreciate the differences in the oscillations of the velocity in the two models and the lack of region with zero speed in model 4.

    Here we study the extension of model 4 to a road with two lanes, where lane changing is allowed: lane 1 is the driving lane, while lane 2 is the fast lane. We consider a single population of homogeneous vehicles and we assume that the coefficients α,β are the same for both lanes and for all vehicles.

    Let N be the total number of vehicles in the road and Nj=Nj(t) the number of vehicles in lane j=1,2 at time t; we have for all t,N1(t)+N2(t)=N; we recall we are assuming periodic boundary conditions. Each vehicle is identified by an index n{1,,N}, and it is associated with a vector Nn=(j,p1n,p2n,s1n,s2n) whose components are: the current index lane j{1,2}, and the indices sjn of the vehicle in front of vehicle n in the lane j (successive vehicle) and pjn of the vehicle behind vehicle n in the lane j (previous vehicle) as shown in Fig. 11. If the n-th vehicle does not have a successive or a previous vehicle in lane j we set sjn=1 or pjn=1 respectively. In other words, the index 1 signifies that there is no such vehicle; for instance s1n=1 means that the vehicle n has no vehicle in front in lane 1. Whenever a lane change occurs, e.g. if the n-th vehicle changes lane, the vectors Nk for k{n,s1n,s2n,p1n,p2n} affected by the change are updated with the new indices.

    Figure 11. 

    Components of the vector Nn, containing information on cell neighbours of the n-th vehicle

    .

    Assuming that vehicle n is currently in lane j then Δxjn=xsjnxn and Δvjn=vsjnvn denote the difference of positions and the difference of the velocities between vehicle n and its successive in the same lane. Moreover we denote with Ij(t)=Ij the set of indices of vehicles ordered by their position in lane j at time t. Note that it is sufficient to update this set only after each lane changing.

    The model can be written for j=1,2 as

    {˙xn=vn˙vn=α(Vj(Δxjn)vn)+βΔvjn(Δxjn)2nIj + lane changing conditions (15)

    where Vj() is the desired velocity function for lane j=1,2 with Vmax2Vmax1. In particular we assume that the velocity functions are equal to zero up to a security distance, then they monotonically increase up to their maximum value:

    V1(Δx)=V2(Δx)=0Δxds (security distance)V1(Δx)V2(Δx)otherwise. (16)

    The parameter ds is a fixed security distance that must be held by the vehicles in order to avoid collisions.

    The lane changing rules are based according essentially on two criteria: a vehicle may change lane if it would travel at a faster speed in the new lane, which means that is would have a higher acceleration (incentive criterion); and the changing action must be safe in order to avoid collisions with the vehicles in the adjacent lane, which means to held the security distance in every movement (security criterion).

    For simplicity we introduce the compact notations:

    d(n,m)=xmxn,aj(n,m)=α(Vj(d(n,m))vn)+βvmvn(d(n,m))2 (17)

    to denote the difference of positions between vehicles with indices n and m, and the acceleration of vehicle n where vehicle m is its successive vehicle in lane j.

    Thus the lane changing rules from lane j to lane j can be expressed as

    aj(n,sjn)>aj(n,sjn)(incentive criterion)d(n,sjn)>dsandd(pjn,n)>ds(security criterion) (18)

    In particular cases we have:

    ● if sjn=1 we consider only the security criterion;

    ● if pjn=1 we consider only the incentive criterion;

    ● if sjn=1 and pjn=1 we decide to change lane;

    ● if sjn=1 we decide to do not change lane.

    Note that in this model lane changes are instantaneous and the velocity of the vehicle remains the same after the changing action. The vehicles following in the new lane adjust their velocities according to the distance from the new vehicle.

    In order to reproduce a realistic description of traffic flow, we introduce a physical timer for lane changing because, as reported by experimental studies [18], lane changing is not frequent. In other words, although a vehicle might have the opportunity and the advantage in changing lane, most often drivers prefer not to change lane. Therefore we set an expected number of lane changes per second Nc and we pick randomly Nc vehicles per second uniformly distributed on the set of vehicles.

    In the following we will use to this characterization of a steady state of model 15.

    Proposition 3. A steady state of model 15 is obtained when both lanes are in equilibrium and there are no lane changing. The equilibrium velocity is given by the optimal velocity functions.

    It is easy to show that such steady state for the two-lanes model 15 is given when the vehicles moves with the same uniform headways hj=LNj, for lane j=1,2 respectively, and with the optimal velocities Vj(hj). We also need to link the velocities for preserve lane changes; the condition is satisfied provided

    V1(h1)=V2(h2). (19)

    Recalling that N=N1+N2, where N is constant, we can write h2 in terms of h1 as

    h2=Lh1Nh1L (20)

    and if the equilibrium velocity is less than Vmax1 we can find a unique value for h1 from equation 19 that we denote by ˉh1. Let ¯N1 be the number of vehicles in lane 1 with headways ˉh1 and in the same way we define ˉh2 and ˉN2. Thus

    Veq:=V1(ˉh1)=V2(ˉh2). (21)

    Now we prove that if 21 holds we have no lane changes and both lanes remain at equilibrium. Consider model 15 with ¯N1 vehicles in lane 1 and with ˉN2 vehicles in lane 2, with initial conditions

    nI1{xn(0)equally spaced with distanceˉh1vn(0)=VeqnI2{xn(0)equally spaced with distanceˉh2vn(0)=Veq. (22)

    For the lane change from lane 1 to lane 2 we can show that the condition

    a2(n,s2n)>a1(n,s1n) (23)

    is never verified because

    a2(n,s2n)a1(n,s1n)==α(V2(xs2nxn)vn)+βvs2nvn(xs2nxn)2+Xα(V1(xs1nxn)vn)βvs1nvn(xs1nxn)2=V2(xs2nxn)V1(xs1nxn). (24)

    Moreover ˉh2<ˉh1 so the distance xs2nxn(ds,ˉh2ds), but from the monotonicity of the function we obtain that V2(h)<V1(ˉh1)h(ds,ˉh2ds). In conclusion 24 is always negative. Similarly we can prove that there are not lane changes from lane 2 to lane 1.

    We have proved the following result.

    Proposition 4. Consider the system 15 with initial conditions 21-22, then no lane changing occurs.

    In the following we study the stability of this equilibrium solution perturbing the initial headways in a lane and analysing the possibility of lane changing in both lanes. We start perturbing the slow lane (lane 1) and then the fast lane (lane 2). Thus we start from an initial condition in which lane 1 is in a local equilibrium but does not satisfy the global equilibrium we described above. This means that we consider a uniform perturbation ε in the headways in lane 1 where we fix an initial constant headway equal to ˉh1+ε and initial velocities equal to V1(ˉh1+ε). In lane 2 we consider initial headways ˉh2 and initial velocities V2(h2). We would like to study how this perturbation influences the equilibrium 22.

    We study the possibility of lane changes from lane 1 to lane 2. Let us consider a vehicle with index n in lane 1, we wonder if the acceleration in lane 2 could be greater than the acceleration in lane 1

    a2(n,s2n)?>a1(n,s1n)V2(d2)V1(ˉh1+ε)+γd22(V2(ˉh2)V1(¯h1+ε))?>0 (25)

    where d2=d(n,s2n) and γ=βα. If ε>0 we do not have lane changes because the previous inequality is always false, in fact it means that in lane 1 there is now a smaller number of vehicles.

    Consider now the case ε<0, assuming V1 is an invertible function, and denoted with V11 its inverse, we can write

    ε<V11(V2(d2)+γd22V2(ˉh2)1+γd22)ˉh1. (26)

    Recalling the security criterion we have that an admissible distance d2 must satisfy d2(ds,ˉh2ds) and therefore the maximum of 26 is reached when d2 tends to ˉh2ds. We get so this threshold for ε:

    ε<V11(V2(ˉh2ds)+γ(ˉh2ds)2V2(ˉh2)1+γ(ˉh2ds)2)ˉh1<0. (27)

    Using a Taylor expansion for V1 and disregarding terms of order O(ε2) we can also obtain an approximation at the first order of the threshold 27. In fact the relation

    V2(d2)V2(ˉh2)ε(1+γd22)V1(ˉh1)?>0 (28)

    is satisfied provided

    ε<V2(d2)V2(ˉh2)(1+γd22)V1(ˉh1). (29)

    Then using the monotonicity of the velocity function we get this a priori bound, approximated at the first order respect to ε

    ε<V2(ˉh2ds)V2(ˉh2)(1+γ(ˉh2ds)2)V1(ˉh1)<0. (30)

    So if ε is smaller than this value we have lane changes from lane 1 to lane 2.

    Consider a vehicle with index n in lane 2 as in Fig. 13. This vehicle will change to lane 1 if the following condition is satisfied

    a1(n,s1n)?>a2(n,s2n). (31)
    Figure 12. 

    Lane change from 1 to 2

    .
    Figure 13. 

    Lane change from 2 to 1

    .

    In this case clearly we will not have lane changes if ε<0. Thus we consider only the case ε>0 and we obtain

    V1(d1)V1(ˉh1)+γd21V1(ˉh1)ε?>0 (32)

    where d1=d(n,s1n) with admissible distance d1(ds,ˉh1+εds). If d1>ˉh1 the previous relation is always verified, while if d1ˉh1 considering the security criterion we can conclude that the perturbation must be greater than the safety distance in order to activate lane changes:

    ε>ds. (33)

    In fact the arrival of a vehicle from lane 2 modifies the initial perturbation ε, decreasing the headway in lane 1, and we go back to the case 1.

    Now we repeat the same analysis adding a perturbation ε in the initial headways in lane 2 starting from the equilibrium 22. Thus we consider an initial condition where vehicles in lane 1 have initial headways ˉh1 and initial velocities V1(ˉh1), and vehicles in lane 2 have initial headways ˉh2+ε and initial velocities V2(ˉh2+ε).

    In this case a vehicle in lane 1 could clearly have a greater acceleration from lane 1 to the lane perturbed if ε>0, but from the security criterion the perturbation must be satisfy the condition

    ε>ds (34)

    as seen in case 2.

    If the perturbation ε is positive we expect no lane changes of this type. Therefore let us consider the case ε<0. Let n be the index of a vehicle in lane 1 we wonder if

    a1(n,s1n)?>a1(n,s2n)V1(d1)V2(ˉh2+ε)+γd21(V1(ˉh1)V2(ˉh2)+ε))?>0 (35)

    with admissible distance d1(ds,ˉh1ds). Consider the maximum distance d1=ˉh1ds, the previous inequality is satisfy if

    ε<V12(V1(d1)+γ(d1)2V1(ˉh1)1+γ(d1)2)ˉh2<0 (36)

    which can be linear approximated by

    ε<V1(d1)V1(ˉh1)(1+γd21)V2(ˉh2). (37)

    Then using the monotonicity of the velocity function we get this a priori bound, approximated at the first order respect to ε

    ε<V1(ˉh1ds)V1(ˉh1)(1+γ(ˉh1ds)2)V2(ˉh2)<0. (38)

    We can summarize the results in the following proposition.

    Proposition 5. Starting from the equilibrium, lane changing for system 15 are activated if a perturbation ε in the headways satisfies the thresholds in Tab. 1. Therefore there are perturbations that do not affect the equilibrium of the system.

    Table 1. 

    Thresholds and perturbations

    .
    from lane 1 to lane 2 from lane 2 to lane 1
    perturbation ε in lane 1 (slow lane) ε<V2(ˉh2ds)V2(ˉh2)(1+γ(ˉh2ds)2)V1(ˉh1)<0 ε>ds>0
    perturbation ε in lane 2 (fast lane) ε>ds>0 ε<V1(ˉh1ds)V1(ˉh1)(1+γ(ˉh1ds)2)V2(ˉh2)<0

     | Show Table
    DownLoad: CSV

    Here we present some numerical tests for the two-lane model 15, using the Runge Kutta 5 method. In the following simulations we set a maximum number of lane changes per second equal to Nc=1 and we fix Δt=0.1 s.

    Let us set L=1500 m, α=5 s1, β=100 m2/s. We use the two optimal velocity functions defined in 13 with parameters V1=0, V2=5, C1=0.02 m1, C2=0, lc=5 m, thus

    V1(h)={5tanh(0.02(h5))if h>ds0otherwiseV2(h)=2V1(h). (39)

    with ds=5 m. We make this choice in order to verify the stability condition 12 in both single lanes for every value of N. We are interesting to study the stability of the model due to the lane changes.

    In this simulation we want to study the perturbation of the lane one from the equilibrium state. Let us fix N=100. Solving equation 19 we get the values ˉh1=45.4 m and ˉh2=22.4 m for which the system remains at the equilibrium if we start from the corresponding steady state. In this case ˉN1=33 m and ˉN2=67 m.

    Now we want to perturb the lane 1 adding new vehicles. From bound 30 we obtain that the perturbation ε in the headways of lane 1 that enables lane changing from lane 1 to lane 2 must satisfy ε<16.5 m, which means that lane changes occur only if N1>51.7.

    Thus fix ˜ε=16.59 m in order to have N1(0)=52 and set N2(0)=ˉN2. We consider the following initial data

    nI1{xn(0)equally spaced with distanceˉh1+˜εvn(0)=V1(ˉh1+˜ε)nI2{xn(0)equally spaced with distanceˉh2vn(0)=V1(ˉh2) (40)

    Fig. 15 shows the simulation for T=500 s. We can see how the perturbation in lane 1 causes lane changes to lane 2 as expected, until the number of vehicles in lane 1 is such that the headways become smaller than the value ˉh1+ε for which we cannot have any more lane changes. In this particular case a new equilibrium is reached with N1(T)=48 and the corresponding headways in lane 1 are equal to LN1(T)=31.25=ˉh113.48 m. This corresponds to a perturbation with ε=14.15 m which is greater than the threshold above. Thus no more lane changes are expected and the system has acquired a new equilibrium with h1=31.25 m h2=21.13 m and V1(h1)=2.41 m/s, V2(h2)=3.12 m/s.

    Figure 14. 

    Optimal velocity functions

    .
    Figure 15. 

    Top: vehicle trajectories in the two lanes. Bottom: number of vehicles versus time

    .

    Whit this simulation we want to study the possibility of lane changes from lane 2 to lane 1. We consider again the equilibrium found in Test 1, and we focus attention to perturb the headways in lane 1 with a positive value of ε, which means to remove some vehicles from the initial value ˉN1.

    From 33 we know that a perturbation that activates lane changes from lane 2 to lane 1 must be greater that the security distance. In our case this is verify if we consider N1<29.73 vehicles at initial time. Therefore we fix ˜ε=6.27 m in order to have N1(0)=29 and set N2(0)=ˉN2. Thus the initial conditions are given by

    nI1{xn(0)equally spaced with distanceˉh1+˜εvn(0)=V1(ˉh1+˜ε)nI2{xn(0)equally spaced with distanceˉh2vn(0)=V1(ˉh2) (41)

    Fig. 16 shows the simulation for T=500 s. We can see how the perturbation in lane 1 causes lane changes from lane 2 to lane 1 as predicted. A new equilibrium is reached with N1(T)=31 and the corresponding headways in lane 1 are equal to LN1(T)=48.39=ˉh1+2.99 m. This corresponds to a perturbation with ε=2.99 m which is smaller than the threshold above. Thus no more lane changes are expected and the system has acquired a new equilibrium with h1=48.38 m h2=23.07 m and V1(h1)=3.50 m/s, V2(h2)=3.46 m/s.

    Figure 16. 

    Top: vehicle trajectories in the two lanes. Bottom: number of vehicles versus time

    .

    In this simulation we study the evolution towards equilibrium. We start with the same number of vehicles in both lanes N1(0)=N2(0)=50. At the initial time all vehicles are equally spaced with zero velocity.

    Fig. 17 shows the simulation for T=1000 s. We can see the presence of an initial phase where vehicles change lane more frequently until arriving in a phase with few lane changes that let the traffic more regular. Initially all vehicles accelerate and lane 1 is partially defected by lane changes towards lane 2 until N1(T)=38,N2(T)=62. In this simulation lane changes from lane 1 to lane 2 are the 92.8% of the total lanes changes.

    Figure 17. 

    Top: vehicle trajectories in the two lanes. Bottom: number of vehicles versus time

    .

    In this simulation we use the two velocity functions as in 13 in order to consider also the instability due to the number of vehicles as seen in the single-lane case. We fix α=1,β=100 and

    V1(Δx)={6.75+7.91tanh(0.13(Δx5)1.57)Δx>50otherwiseV2(Δx)=2V1(Δx).

    The stability conditions for the single-lane 12 are in this case: for lane 1 stability for N<68 and N>100, while for lane 2 we have stability for N<57 and N>130.

    We start with the same number of vehicles in both lanes N1(0)=N2(0)=90; lane 2 is at the equilibrium while in lane 1 we add random perturbations rn in the initial positions of the vehicles. Thus we have

    nI1{xn(0)xn1(0)=LN1(0)+rnvn(0)=V1(LN1(0))nI2{xn(0)xn1(0)=LN2(0)vn(0)=V2(LN2(0)) (42)

    Fig. 18 shows the simulation for T=500 s. We can see the creation of stop & go waves in both lanes due to the frequently lane changes and to the instability of the model.

    Figure 18. 

    Top: vehicle trajectories in the two lanes. Bottom: number of vehicles versus time

    .

    The model 15 can be easily generalized to the multi-lane case with a generic number of lanes. We can differentiate the lanes by attributing different profiles of desired velocity, therefore let J be the number of lanes, we consider the velocities functions V1(),,VJ() with the property Vi()Vj() for i<j.

    The model can be written as

    {˙xn=vn˙vn=α(Vj(Δxjn)vn)+βΔvjn(Δxjn)2nIj + lane changing conditionsfor j=1,,J. (43)

    We adopt the lane changing conditions as in 18. Note that, except for the cases j=1 or j=J, if j>2 a vehicle might have the possibility to changes from lane j to lane j1 or from lane j to lane j+1. Consequently if both changes are possible we choose the most advantageous one in terms of acceleration.

    As we done for the two-lane model we can define the steady state of model 43 in which all lane are at the equilibrium and lane changes do not occur. This is provided for the values of the headways

    ˉh1,,ˉhJ (44)

    that verify the condition

    V1(ˉh1)==VJ(ˉhJ). (45)

    In order to find this equilibrium we require also that the equilibrium velocity defined in 45 must be smaller than the value Vmax1, that is the maximum velocity value allowed in the slower lane (j=1).

    Let us consider a three-lane road (J=3). The steady state is given by the three values of the headways ˉh1,ˉh2,ˉh3 such that Veq:=V1(ˉh1)=V2(ˉh2)=V3(ˉh3). Using the same previous techniques can be show that with these conditions no lane changes occur and the system remains at the equilibrium.

    We are now interested to add a perturbation in the middle lane and to study the possibility of lane changing. More specifically let us consider the initial conditions

    nI1{xn(0)equally spaced with distanceˉh1vn(0)=VeqnI2{xn(0)equally spaced with distanceˉh2+εvn(0)=V2(ˉh2+ε)nI3{xn(0)equally spaced with distanceˉh3vn(0)=Veq. (46)

    We can observe that the system is comparable to two subsystems: lane 1 - lane 2 and lane 2 - lane 3 where the lane changes are regulated by the thresholds in Table 1. More specifically for the subsystem lane 2 - lane 3 we consider the case of a perturbation in the slow lane (first row of the table) while for the subsystem lane 1 - lane 2 we refer to the case of a perturbation in the fast lane (second row of the table). We add to this framework the possibility of choosing the best advantageous change for a vehicle in the middle lane that might have two possibilities for change lane. The thresholds that enable lane changes can be obtained from Table 1 with the appropriate modifications. We have

    Here we propose a numerical example with a three-lane road, using the Runge Kutta 5 method. Consider the velocity function V1(h) as in 39 and define V2(h)=32V1(h) and V3(h)=2V1(h). From the value ˉh1=50 m we obtain that ˉh2=31 m, ˉh3=23.7 m and Veq=3.58 m/s as shown in Fig. 19. The corresponding number of vehicles are: ˉN1=30, ˉN2=48, ˉN3=63. In order to add a perturbation in lane 2 we find the values of the perturbation that allow lane changing. From Table 2 we obtain: ε>5 m for lane changes from lanes 1 and 3 to lane 2, ε<2.25 m for lane changes from lane 2 to lane 1 and ε<7.74 m for lane changes from lane 2 to lane 3. In the following numerical tests we use the initial conditions 46 with ε=2.68 m in the test (a) and with ε=7.91 m in the test (b). We can observe that in the test (a) the perturbation has produced lane changes from lane 2 to lane 1 while in the test (b) lane changes from lane 2 to lane 3 occurred.

    Figure 19. 

    Desired velocity functions

    .
    Table 2. 

    Thresholds and perturbations

    .
    12 & 32 21 23
    pert. ε in lane 2 ε>ds>0 ε<V1(ˉh1ds)V1(ˉh1)(1+γ(ˉh1ds)2)V2(ˉh2)<0 ε<V3(ˉh3ds)V3(ˉh3)(1+γ(ˉh3ds)2)V2(ˉh2)<0

     | Show Table
    DownLoad: CSV
    Figure 20. 

    Test (a) - Top: vehicle trajectories in the three lanes. Bottom: number of vehicles versus time

    .
    Figure 21. 

    Test (b) - Top: vehicle trajectories in the three lanes. Bottom: number of vehicles versus time

    .

    In the following test we show an example of instability, comparing the results with the test in section 3.3.4. Let us consider the function V1(h) as in the aforementioned test, and define V2(h)=32V1(h) and V3(h)=2V1(h). We consider N1(0)=N3(0)=90 and N2(0)=0 with initial conditions with random perturbations rn.

    nI1{xn(0)xn1(0)=LN1(0)+rnvn(0)=V1(LN1(0))nI3{xn(0)xn1(0)=LN3(0)vn(0)=V2(LN3(0)) (47)

    From Fig. 22 we can see that lane 1 gradually empties into lane 2. Due to frequent lane changes, more pronounced stop & go waves occur in fast lanes, while slow lane tends to stabilize. In test 3.3.4 we recall that the instabilities were evident in both lanes.

    Figure 22. 

    Top: vehicle trajectories in the three lanes. Bottom: number of vehicles versus time

    .

    In this paper we have studied a microscopic model 4 for lane changing proposing simple lane changing rules. We have computed global steady states and we have investigated the linear stability of such solutions. The global steady state of the multi-lane model is parametrized by the total number N of vehicles in the road. All lanes are coupled by the lane changing conditions, and the equilibrium is reached only when the crowding of each single lane is such that no lane changing is convenient anymore. At that point the system can reach the equilibrium lane by lane. We have proved that the model for the single-lane case has a larger stability region than the model 3. In the multi-lane case we have proved that is possible to determine conditions on perturbations in which the equilibrium of the steady state is preserved and lane changing does not occur. We plan to derive a macroscopic version of this model where each lane would be described by its own equation and the lane changes would appear as source terms for the macroscopic equations. This study can be useful in applications for instance in the design of velocity profiles to minimize lane changes in order to avoid jams and car accidents.

    Both authors are members of the INdAM Research group GNCS. This work was supported in part by Progetto di Ateneo 2019, n. 1622397 and 2020 n. 2023082 (Sapienza - Università di Roma), and PRIN 2017KKJP4X.



    [1] A. V. Chechkin, J. Klafter, I. M. Sokolov, Fractional Fokker-Planck equation for ultraslow kinetics, Europhys. Lett., 63 (2003), 326–332. https://doi.org/10.1209/epl/i2003-00539-0 doi: 10.1209/epl/i2003-00539-0
    [2] Y. G. Sinai, The limiting behavior of a one-dimensional random walk in a random medium, Theory Probab. Appl., 27 (1983), 256–268. https://doi.org/10.1137/1127028 doi: 10.1137/1127028
    [3] M. M. Meerschaert, E. Nane, P. Vellaisamy, Distributed-order fractional diffusions on bounded domains, J. Math. Anal. Appl., 379 (2011), 216–228. https://doi.org/10.1016/j.jmaa.2010.12.056 doi: 10.1016/j.jmaa.2010.12.056
    [4] M. Caputo, Distributed order differential equations modelling dielectric induction and diffusion, Fract. Calc. Appl. Anal., 4 (2001), 421–442.
    [5] X. Hu, F. Liu, V. Anh, I. Turner, A numerical investigation of the time distributed-order diffusion model, Anziam J., 5 (2014), C464–C478. https://doi.org/10.21914/anziamj.v55i0.7888 doi: 10.21914/anziamj.v55i0.7888
    [6] J. Jia, H. Wang, A fast finite difference method for distributed-order space-fractional partial differential equations on convex domains, Comput. Math. Appl., 75 (2018), 2031–2043. https://doi.org/10.1016/j.camwa.2017.09.003 doi: 10.1016/j.camwa.2017.09.003
    [7] W. Fan, F. Liu, A numerical method for solving the two-dimensional distributed order space-fractional diffusion equation on an irregular convex domain, Appl. Math. Lett., 77 (2018), 114–121. https://doi.org/10.1016/j.aml.2017.10.005 doi: 10.1016/j.aml.2017.10.005
    [8] X. Wang, F. Liu, X. Chen, Novel second-order accurate implicit numerical methods for the Riesz space distributed-order advection-dispersion equations, Adv. Math. Phys., 2015 (2015), 1–14. https://doi.org/10.1155/2015/590435 doi: 10.1155/2015/590435
    [9] I. Oseledets, Tensor-train decomposition, SIAM J. Sci. Comput., 33 (2011), 2295–2317. https://doi.org/10.1137/090752286 doi: 10.1137/090752286
    [10] D. Bertaccini, F. Durastante, Block structured preconditioners in tensor form for the all-at-once solution of a finite volume fractional diffusion equation, Appl. Math. Lett., 95 (2019), 92–97. https://doi.org/10.1016/j.aml.2019.03.028 doi: 10.1016/j.aml.2019.03.028
    [11] T. Breiten, V. Simoncini, M. Stoll, Low-rank solvers for fractional differential equations, Electron. Trans. Numer. Anal., 45 (2016), 107–132. https://doi.org/10.17617/2.2270973 doi: 10.17617/2.2270973
    [12] S. Dolgov, J. Pearson, D. Savostyanov, M. Stoll, Fast tensor product solvers for optimization problems with fractional differential equations as constraints, Appl. Math. Comput., 273 (2016), 604–623. https://doi.org/10.1016/j.amc.2015.09.042 doi: 10.1016/j.amc.2015.09.042
    [13] I. Oseledets, E. Tyrtyshnikov, N. Zamarashkin, Tensor-train ranks for matrices and their inverses, Comput. Methods Appl. Math., 11 (2011), 394–403. https://doi.org/10.2478/cmam-2011-0022 doi: 10.2478/cmam-2011-0022
    [14] V. Kazeev, B. Khoromskij, E. Tyrtyshnikov, Multilevel Toeplitz matrices generated by tensor-structured vectors and convolution with logarithmic complexity, SIAM J. Sci. Comput., 35 (2013), A1511–A1536. https://doi.org/10.1137/110844830 doi: 10.1137/110844830
    [15] L. Chou, S. Lei, Tensor-train format solution with preconditioned iterative method for high dimensional time-dependent space-fractional diffusion equations with error analysis, J. Sci. Comput., 80 (2019), 1731–1763. https://doi.org/10.1007/s10915-019-00994-3 doi: 10.1007/s10915-019-00994-3
    [16] L. Chou, S. Lei, Finite volume approximation with ADI scheme and low-rank solver for high dimensional spatial distributed-order fractional diffusion equations, Comput. Math. Appl., 89 (2021), 116–126. https://doi.org/10.1016/j.camwa.2021.02.014 doi: 10.1016/j.camwa.2021.02.014
    [17] W. Deng, B. Li, W. Tian, P. Zhang, Boundary problems for the fractional and tempered fractional operators, Multiscale Model. Simul., 16 (2018), 125–149. https://doi.org/10.1137/17M1116222 doi: 10.1137/17M1116222
    [18] C. Çelik, M. Duman, Crank-Nicolson method for the fractional diffusion equation with the Riesz fractional derivative, J. Comput. Phys., 231 (2012), 1743–1750. https://doi.org/10.1016/j.jcp.2011.11.008 doi: 10.1016/j.jcp.2011.11.008
    [19] I. Gohberg, V. Olshevsky, Circulants, displacements and decompositions of matrices, Integr. Equat. Oper. Theory, 15 (1992), 730–743. https://doi.org/10.1007/bf01200697 doi: 10.1007/bf01200697
    [20] R. Chan, M. Ng, Conjugate gradient methods for Toeplitz systems, SIAM Rev., 38 (1996), 427–482. https://doi.org/10.1137/S0036144594276474 doi: 10.1137/S0036144594276474
  • This article has been cited by:

    1. Mengyuan Jin, Qun Li, Jifen Liu, Jing Sun, Fang Hu, 2024, Multimodal Feature Fusion-Based Node Representation Learning for Drug-Drug Interaction Prediction, 979-8-3503-7854-2, 507, 10.1109/CEI63587.2024.10871727
  • 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(2127) PDF downloads(69) Cited by(0)

Figures and Tables

Tables(3)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog