Processing math: 100%

From discrete to continuous Wardrop equilibria

  • Received: 01 November 2011 Revised: 01 March 2012
  • Primary: 90C46, 90C35; Secondary: 49N15.

  • The notion of Wardrop equilibrium in congested networks has been very popular in congested traffic modelling since its introduction in the early 50's, it is also well-known that Wardrop equilibria may be obtained by some convex minimization problem. In this paper, in the framework of $\Gamma$-convergence theory, we analyze what happens when a cartesian network becomes very dense. The continuous model we obtain this way is very similar to the continuous model of optimal transport with congestion of Carlier, Jimenez and Santambrogio [6] except that it keeps track of the anisotropy of the network.

    Citation: Jean-Bernard Baillon, Guillaume Carlier. From discrete to continuous Wardrop equilibria[J]. Networks and Heterogeneous Media, 2012, 7(2): 219-241. doi: 10.3934/nhm.2012.7.219

    Related Papers:

    [1] Boyu Wang . A splitting lattice Boltzmann scheme for (2+1)-dimensional soliton solutions of the Kadomtsev-Petviashvili equation. AIMS Mathematics, 2023, 8(11): 28071-28089. doi: 10.3934/math.20231436
    [2] Lingxiao Li, Jinliang Zhang, Mingliang Wang . The travelling wave solutions of nonlinear evolution equations with both a dissipative term and a positive integer power term. AIMS Mathematics, 2022, 7(8): 15029-15040. doi: 10.3934/math.2022823
    [3] Pengshe Zheng, Jing Luo, Shunchu Li, Xiaoxu Dong . Elastic transformation method for solving ordinary differential equations with variable coefficients. AIMS Mathematics, 2022, 7(1): 1307-1320. doi: 10.3934/math.2022077
    [4] A. S. Mohamed . Fibonacci collocation pseudo-spectral method of variable-order space-fractional diffusion equations with error analysis. AIMS Mathematics, 2022, 7(8): 14323-14337. doi: 10.3934/math.2022789
    [5] Sheza. M. El-Deeb, Gangadharan Murugusundaramoorthy, Kaliyappan Vijaya, Alhanouf Alburaikan . Certain class of bi-univalent functions defined by quantum calculus operator associated with Faber polynomial. AIMS Mathematics, 2022, 7(2): 2989-3005. doi: 10.3934/math.2022165
    [6] Ali Khalouta, Abdelouahab Kadem . A new computational for approximate analytical solutions of nonlinear time-fractional wave-like equations with variable coefficients. AIMS Mathematics, 2020, 5(1): 1-14. doi: 10.3934/math.2020001
    [7] Muhammad Imran Liaqat, Sina Etemad, Shahram Rezapour, Choonkil Park . A novel analytical Aboodh residual power series method for solving linear and nonlinear time-fractional partial differential equations with variable coefficients. AIMS Mathematics, 2022, 7(9): 16917-16948. doi: 10.3934/math.2022929
    [8] Cheng Chen . Hyperbolic function solutions of time-fractional Kadomtsev-Petviashvili equation with variable-coefficients. AIMS Mathematics, 2022, 7(6): 10378-10386. doi: 10.3934/math.2022578
    [9] Abdulah A. Alghamdi . Analytical discovery of dark soliton lattices in (2+1)-dimensional generalized fractional Kundu-Mukherjee-Naskar equation. AIMS Mathematics, 2024, 9(8): 23100-23127. doi: 10.3934/math.20241123
    [10] Lin Fan, Shunchu Li, Dongfeng Shao, Xueqian Fu, Pan Liu, Qinmin Gui . Elastic transformation method for solving the initial value problem of variable coefficient nonlinear ordinary differential equations. AIMS Mathematics, 2022, 7(7): 11972-11991. doi: 10.3934/math.2022667
  • The notion of Wardrop equilibrium in congested networks has been very popular in congested traffic modelling since its introduction in the early 50's, it is also well-known that Wardrop equilibria may be obtained by some convex minimization problem. In this paper, in the framework of $\Gamma$-convergence theory, we analyze what happens when a cartesian network becomes very dense. The continuous model we obtain this way is very similar to the continuous model of optimal transport with congestion of Carlier, Jimenez and Santambrogio [6] except that it keeps track of the anisotropy of the network.


    In this paper, Backward Difference Lattice Boltzmann (BD-LB) method is used to simulate the Burgers equation with variable coefficient in space and time. It is very difficult to construct its analytical solution directly for the partial differential equation (PDE) with variable coefficient in space and time, whereas finding its numerical solution is an effective way. The traditional macroscopic numerical simulation methods include finite difference (FD) method, finite volume (FV) method, and finite element (FE) method. Compared with these methods, lattice Boltzmann method (LBM) is a mesoscopic numerical approach with many advantages, such as clear physical background, simple algorithm, easy parallel computing, and easy to implement program and handle complex boundary conditions [1,2,3].

    The Boltzmann equation [4], a complex microintegral equation, is the fundamental equation of the gas kinetics theory. The right-end term is called the collision term, which is referred to as $ \Omega \left (f \right) $. The existence of $ \Omega \left (f \right) $ brings great difficulties to solve Eq (1.1).

    $ ft+ξfr+afξ=(ff1ff1)d2D|g|cosθdΩdξ1.
    $
    (1.1)

    However, the collision operators used in the LBM are generally based on the much simpler Bhatnagar-Gross-Krook (BGK) collision operator. Chen and Qian et al. [5,6] put forward to single-relaxation-time lattice Boltzmann (SRT-LB) model. The model controls the speed of different particles near the equilibrium state by using the same time relaxation coefficient.

    $ fj(x+cjδt,t+δt)fj(x,t)=1τ[fj(x,t)feqj(x,t)],
    $
    (1.2)

    where $ c_{j} $ is the discrete lattice velocity, $ \tau $ stands for the relaxation time, and $ f_{j}^ {eq} $ represents the distribution function of local equilibrium state.

    Qin et al. [7] adds a body force term in Eq (1.2) to simulate the incompressible Navier-Stokes flow and then investigate aqueous humor dynamics in human eye. Hu and Lan et al. [8,9] add the modified function $ h_{i} $ to the Eq (1.2) to simulate the Gardner equation and KdV-Burgers equation with time-dependent variable coefficient respectively. Chai et al. [10,11] add the auxiliary distribution function $ G_{j}\left (\bf{x}, t \right) $ and a source term $ F_{j} \left (\bf{x}, t \right) $, proposing a multi-relaxed time lattice Boltzmann (MRT-LB) model to solve the Navier-Stokes equation and the convective diffusion equations. Gerasim V. Krivovichev [12] uses the Eq (1.2) formula to analyze the parameterized higher-order finite difference schemes of the linear advection equations. These schemes are based on a linear combinations of the spatial approximations of the convective term at the characteristic directions.

    We use a FD method to solve the PDE problems in computational mathematics, but the forward difference format is often conditionally stable and the backward difference format is unconditionally stable. In this paper, we do not add correction function and auxiliary function, but we adopt the backward difference format and then rewrite the Eq (1.2) as follow:

    $ fj(x,t)fj(xcjδt,tδt)=1τ[fj(xciδt,tδt)feq j(xcjδt,tδt)].
    $
    (1.3)

    The Burgers equation with variable coefficient is generally used to simulate the formation and decay of non-plane shock waves. The Burgers equation with the nonlinear term $ a(t) $ and the dispersion term $ b(t) $ can simulate the propagation of long shock waves in two shallow liquids [13].

    $ ut+a(t)uux+b(t)2ux2=0.
    $
    (1.4)

    It is shown in the literature that when $ b(t) $ is equal to 1, the shock wave solution reverses its velocity and collapses after $ a(t) $ changes the critical point of its symbol. In literature, the soliton-type solutions are constructed by using $ B\ddot{a}cklund $ transformation for the given form of $ a(t) $ and $ b(t) $. But these methods are very difficult for solving the variable coefficient partial differential equations, and we study the following Burgers equation with variable coefficient by using BD-LB method [14].

    $ ut+ux[a(x,t)u]+2x2[b(x,t)u]=0.
    $
    (1.5)

    Equation (1.5) can represent a variety of physical models widely used in the fields of solid state materials, plasmas and fluid. $ u $ is amplitude function about time $ t $ and space $ x $. $ a(x, t) $ and $ b(x, t) $ are both analytical functions about $ x $ and $ t $. The subscripts represent partial derivatives.

    This paper is organized as follows: In Section 2, the LB method for the Burgers equation of variable coefficient is recovered by using both the Chapman-Enskog (CE) analysis and the direct Taylor expansion (DTE) method. In Section 3, we first test Example 3.1 and obtain some reasonable lattice parameters. Then we apply these parameters to other variable coefficient examples and compare the numerical result to the analytical solution. In Section 4, we discuss the obtained results and draw a conclusion.

    The Eq (1.3) is rewrote as follow:

    $ fj=feqj,n1+fneqj,n1.
    $
    (2.1)

    We will use the following notations in this paper:

    ● $ f_{j}: = f_{j}\left(\boldsymbol{x}, t\right) $;

    ● $ f_{j, n-1}: = f_{j}\left(\boldsymbol{x}-c_{j} \delta t, t-\delta t\right) $;

    ● $ f_{j, n-1}^{e q}: = f_{j}^{e q}\left(\boldsymbol{x}-c_{j} \delta t, t-\delta t\right) $;

    ● $ f_{j, n-1}^{neq}: = \left(1-\frac{1}{\tau}\right)(f_{j, n-1}-f_{j, n-1}^{e q}) $;

    ● $ \delta $ is the Kronecker delta;

    ● $ \nabla $ is first-order spatial partial derivative.

    In Eq (2.1), $ f_{j, n-1}^{e q} $ is the equilibrium distribution function, $ f_{j, n-1}^{neq} $ is the non-equilibrium distribution function. We define the macroscopic variable $ u $ as the sum of $ \sum_{i} f_{i} $.

    Judging from Eq (2.1), we can get the following equation:

    $ fjfeqj,n1=fj,n1feqj,n11τ(fj,n1feqj,n1).
    $
    (2.2)

    According to the equation proposed by Eq (1.2),

    $ fjfj,n1=1τ(fjfeqj,n1),
    $
    (2.3)

    and then

    $ feqj,n1=fj,n1+τ(fjfj,n1)=fj,n1[1+τfj,nfj,n1fj,n1]=fj,n1{1+τ[fj,nfj,n11]}.
    $
    (2.4)

    Equation (2.4) can be written as follow:

    $ feqj=fj{1+τ[fj,n+1fj1]},
    $
    (2.5)

    we delimit $ f_{j, n+1} = e^{D_{j}}f_{j} $, where $ D_{j} $ is the difference operator $ D_{j} = \partial_{t}+c_{j} \nabla $.

    The Eq (2.1) can be rewritten as follow:

    $ fj(x,t)=[1+τ(eDj1)]1feqj(x,t).
    $
    (2.6)

    In the $ CE $ expansion, the LB equation is expanded by a dimensionless parameter $ \epsilon $, which is proportional the Knudsen number ($ Kn = \lambda /L $), $ \lambda $ is the mean free path and $ L $ is the feature length. We choose the local equilibrium distribution function in the following form [2,10,15,16,17,18,20]:

    $ feqj(u)=wj,0r(u)+wj,1cjs(u)+wj,2(c2jc2S,2)t(u),
    $
    (2.7)

    we adopt the weight family $ w_{j, a} $ selection satisfy

    $ jwj,a=1,jwj,acj=0,jwj,ac2j=c2S,a,
    $
    (2.8)

    here, $ c_{S, a} $ is the lattice sound speed and the expansion of the function, $ r, s, t $ are expanded by $ CE $ as follows:

    $ r=n=0ϵnrn(u),s=n=0ϵnsn(u),t=n=0ϵntn(u).
    $
    (2.9)

    Correspondingly, the expansion of the equilibrium state distribution function is as follow:

    $ feqj(u)=n=0ϵnf(eq,l)j(u),
    $
    (2.10)

    that is to say, the equilibrium state distribution function varies when it is close to the scaling limit.

    We define the second moment of $ u $ in Eq (2.7) as follows:

    $ Rn(u)=jf(eq,n)j,Sn(u)=jf(eq,n)jcj,Tn(u)=jf(eq,n)jc2j,
    $
    (2.11)

    here, $ j > 0 $ breaks the conservation of $ u $ through the collision process because $ R_{n} \left (u \right) = r_{n} \left (u \right) $ and the $ r_{0} \left (u\right) = u $. In this paper, we only consider Eq (1.5) with a conservation $ u $, so we adopt $ r_{j} = 0 $ when $ j > 0 $. Furthermore, we hypothesize that $ u $ is differentiable in the analysis.

    For $ \epsilon $, the temporal and spatial scale expansion as $ \partial_{t} = \sum_{k = 1}^{\infty} \epsilon^{k} \partial_{t_{k}} $, $ \nabla = \epsilon \nabla_{1} $, $ t_{k} $ is expressed as $ k $ time scales, $ \nabla_{k} $ is expressed as $ k $ space scales, respectively.

    $ Dj=k=1ϵkDj,k,
    $
    (2.12)

    in which $ D_{j, k}: = \partial_{t_{k}}+\delta_{k, 1} c_{j} \nabla_{1} $. The solution of the Eq (2.6) can be written in the following form:

    $ fj(x,t)=k=0ϵkf(k)j(x,t),
    $
    (2.13)

    where

    $ f(0)j(x,t)=f(eq,0)j,f(1)j(x,t)=τDj,1f(eq,0)j+f(eq,1)j,f(2)j(x,t)=τ[Dj,2(τ12)D2j,1]f(eq,0)jτDj,1f(eq,1)j+f(eq,2)j.
    $
    (2.14)

    The Burgers equation has a second-order spatial derivative. It needs to sort the $ \epsilon^{2} $ by using the formalized $ u $, the results are summarized in Table 1 and we choose the following equation in order to make the results equivalent to the Burgers equation.

    $ J0=0,J1=a(x,t)u2/2,K0=b(x,t)u/(τ1/2).
    $
    (2.15)
    Table 1.  Equations of all order with regarding to the parameter $ \epsilon $.
    $ \epsilon $ order Equations of motion
    1 $ \partial_{t_{1}} u+ \partial_{x_{1}} J_{0}=0, $
    2 $ \partial_{t_{2}}u=\partial_{x_{1}}\left\{\left(\tau-\frac{1}{2}\right)\partial_{x_{1}} K_{0} \right\}- \partial_{x_{1}}J_{1}. $

     | Show Table
    DownLoad: CSV

    We rewrite $ f_{j, n-1}^{neq} $ in Eq (2.1) into Eq (2.16) [19]:

    $ fneqj,n1=k=1(11τ)k{feqj(x(k+1)cjδt,t(k+1)δt)feqj(xkcjδt,tkδt)}.
    $
    (2.16)

    Here, we assume $ f_{j}^{e q} $ as follow, and the equilibrium state distribution function $ f_{j}^{eq} $ is an analytic function.

    $ feqj=u[w(0)j+Kw(2)j]+u2Jw(1)j,
    $
    (2.17)

    the moments of $ w_{j} $ are shown in Eq (2.17) as in Table 2. Explicit forms for these weights are presented in Eqs (2.28) and (2.30).

    Table 2.  The moments of $ w_{j} $.
    Order $ w_{j}^{(0)} $ $ w_{j}^{(1)} $ $ w_{j}^{(2)} $
    0 1 0 0
    1 0 1 0
    2 0 0 1

     | Show Table
    DownLoad: CSV

    In Eq (2.1), for $ j $ sum easy to calculate, on the left of the equation is $ u\left(x, t\right) $, and the equilibrium part to the right of Eq (2.1) changes to

    $ jfeqj,n1=jm=0(δt)mm!(t+(cjx))mfeqj(x,t)=uδt(ut+u[Ju]x)+12(δt)22[Ku]x2+O(3ux3,2u2xt,2ut2).
    $
    (2.18)

    At the same time, for $ f_{j, n-1}^{neq} $ in Eq (2.16), one obtain

    $ jk=1(11τ)k{feqj(x(k+1)cjδt,t(k+1)δt)feqj(xkcjδt,tkδt)}=δtT1(ut+u[Ju]x)+12(δt)22[Ku]x2T2+O(3ux3,2u2xt,2ut2),
    $
    (2.19)

    where $ \mathcal{T}_{1} = \tau -1, \; \mathcal{T}_{2} = 2\tau^{2}-\tau-1, \; \mathcal{T}_{3} = 6\tau^{3}-6\tau^{2}+\tau-1 $ and $ \tau > 1/2 $ [20].

    Putting Eqs (2.18) and (2.19) into Eq (2.1), we can get

    $ ut=u[Ju]x+δt2!2[Ku]x2T2+1T1+1+O(3ux3,2u2xt,2ut2).
    $
    (2.20)

    In order to recover the macroscopic Burgers equation, the parameters are defined as

    $ J=a(x,t)/2,K=b(x,t)/(τ1/2).
    $
    (2.21)

    Compared with Eqs (2.15) and (2.21), the two analytical equations produce the same results for non-linear equations. These two methods are very different, but they recover the macroscopic Burgers equation.

    In numerical simulation, we assume the physical space $ X = \delta x\cdot x $ and $ T = \delta t\cdot t $, and then put the $ X, \; T $ into Eq (2.20). Therefore, Eq (2.21) is as follows:

    $ J=a(x,t)δt/(2δx),K=b(x,t)δt/[(δx)2(τ1/2)].
    $
    (2.22)

    At the same time, the leading truncation error term at $ (\delta t)^{0} $ of Eq (2.20) is the fourth spatial derivative term whose coefficients involve $ \mathcal{K} $. This error term is

    $ (δx)2(T4+1)2(T2+1)4uX4.
    $
    (2.23)

    For the D1Q5 lattice, the format of $ f_{j}^{e q} $ is Eq (2.17). In order to remove the truncation error, the following $ \delta f_{j}^{eq} $ is added to $ f_{j}^{eq} $ in the D1Q7 lattice,

    $ δfeqj=2(T1+1)δt(T2+1)(δx)2w(4)j.
    $
    (2.24)

    For the D1Q7 lattice, the following $ f_{j}^{e q} $ is employed:

    $ feqj=u[w(0)j+Kw(2)j2(T1+1)δt(T2+1)(δx)2w(4)j]+u2Jw(1)j.
    $
    (2.25)

    We adopt the notation standardized in the LBM literature, where DdQq [6] refers to the d spatial dimensional model with q kinetic velocity. For the D1Q5 and D1Q7 models, a set of discrete weights having only the the unit $ n $ moment, is provided. These set of weights have the following properties:

    $ icpiw(n)i=δp,n.
    $
    (2.26)

    In the case of D1Q5, when $ c_{i} = \left \{0, \pm 1, \pm 2 \right \} $, $ w_{i}^{(n)} $ can be obtained by inverting the matrix:

    $ (111110112201144011880111616).
    $
    (2.27)

    The $ w_{i}^{(n)} $ is

    $ (w(0)iw(1)iw(2)iw(3)iw(4)i)=({1,0,0}{0,±23,112}{54,23,124}{0,16,±112}{14,16,124}).
    $
    (2.28)

    In the case of D1Q7, when $ c_{i} = \left \{0, \pm 1, \pm 2, \pm 3 \right \} $, $ w_{i}^{(n)} $ can be obtained by inverting the matrix:

    $ (11111110112233011449901188272701116168181011882432430111616729729).
    $
    (2.29)

    The $ w_{i}^{(n)} $ is

    $ (w(0)iw(1)iw(2)iw(3)iw(4)iw(5)iw(6)i)=({1,0,0,0,0}{0,±34,320,±160}{4936,34,340,1180}{0,1348,±16,148}{718,1348,112,1144}{0,±148,160,±1240}{136,148,1120,1720}).
    $
    (2.30)

    In this paper, the global relative error (GRE) is used to verify the effectiveness of the lattice Boltzmann model and the least-squares fitting is used to calculate the accuracy of the model.

    $ GRE = \frac{ { \sum_{k = 1}}\left | u\left ( x_{k} , t \right )-u^{*} \left ( x_{k}, t \right ) \right | }{ { \sum_{k = 1} \left | u^{*} \left ( x_{k} , t \right ) \right | } }, $

    among these, $ u\left (x, t \right) $ and $ u^ {\ast} \left (x, t \right) $ respectively represent the numerical solution and analytical solution.

    Example 3.1. For $ a\left (x, t \right) = C_{1}, b\left (x, t \right) = C_{2} $, Eq (1.5) and the initial boundary conditions are taken as follows:

    $ ut+C1uux=C22ux2,0x1,0tT;u(x,0)=2C2πsin(πx)C3+cos(πx),0x1;u(0,t)=u(1,t)=0,0tT,
    $

    analytical solution [21]

    $ u(x, t) = \frac{2 C_{2} \pi e^{-\pi^{2} C_{2} t} sin (\pi x)}{C_{3}+e^{-\pi^{2} C_{2} t} cos (\pi x)}. $

    We assume $ C_{1} = 1, \; C_{2} = 0.01, \; C_{3} = 2, \; T = 1 $. Figures 13 depict the images when $ \delta x = 0.05 $, $ \delta x = 0.1 $, $ \delta x = 0.2 $, and the left-hand plots in each figure show the minimum stability values of $ \tau $ and $ \delta t $ about the D1Q5 and D1Q7 lattices. We compare them with the reference [19], and the fitting about the images is very good. The value of $ \delta t $ is shown in Table 3, $ \delta t $ displays the maximum time increments to maintain stability with $ \tau = 1 $.

    Figure 1.  the relaxation time $ \tau $ (a) and the GRE (b) are presented for $ \delta x = 0.05 $.
    Figure 2.  the relaxation time $ \tau $ (a) and the GRE (b) are presented for $ \delta x = 0.1 $.
    Figure 3.  the relaxation time $ \tau $ (a) and the GRE (b) are presented for $ \delta x = 0.2 $.
    Table 3.  The value of $ \delta t $ corresponding to $ \delta x $.
    $ \delta x $ 0.05 0.10 0.20
    $ \delta t $(D1Q5) $ 1.26 \times 10^{-6} $ $ 5.75 \times 10^{-6} $ $ 1.20 \times 10^{-4} $
    $ \delta t $(D1Q7) $ 9.66 \times 10^{-7} $ $ 4.07 \times 10^{-6} $ $ 7.24 \times 10^{-5} $

     | Show Table
    DownLoad: CSV

    The plots on the right-hand in Figures 13 show the relationship of GRE and $ \delta t $ about D1Q5 and D1Q7 lattices. The crossed points show corresponding results between the D1Q5 and D1Q7 models. In Figure 4, the numerical results agree very well with the analytical solution for any t.

    Figure 4.  the D1Q5 scheme (a) and the D1Q7 scheme (b) are used to compare with the analytical solution for $ \delta x = 0.2 $.

    Example 3.2. For $ a\left (x, t \right) = sech^{2} \left (t \right), \; b\left (x, t \right) = C_{1} sech^{2} \left (t \right) $, Eq (1.5) becomes

    $ \frac{\partial u}{\partial t} +u\frac{\partial }{\partial x} \left [ sech^{2}\left ( t \right )u \right ]+\frac{\partial ^{2} }{\partial x^{2} }\left [ C_{1} sech^{2}\left ( t \right )u \right ] = 0, $

    analytical solution [22]

    $ u(x, t) = \frac{C_{3}}{C_{2}} \pm \frac{C_{1} C_{2} \sqrt{r^{2}-1}}{r+cosh (k(x)+c(t)+l)}+\frac{C_{1} C_{2} sinh (k(x)+c(t)+l)}{r+cosh (k(x)+c(t)+l)}, $

    where $ k(x) = C_{2} x, \; c(t) = -C_{3} tanh\left (t \right) $, $ C_{2} $ and $ C_{3} $ are any constant, $ r^{2} \ge 1 $.

    Here, macro parameters are assumed to be $ r = 2, \; l = 0, \; C_{1} = 2, \; C_{2} = C_{3} = 1 $ and lattice parameters are assumed $ \delta x = 0.2, \; \tau = 1 $, and the value of $ \delta t $ is shown in Table 3. The calculation area is fixed at $ \left [-10, 10 \right] $. Figure 5 shows the comparison plot of numerical results and analytical solution at different moments. The space-time plots of the numerical result from $ t = 0 $ to $ t = 10 $ are shown in Figure 6.

    Figure 5.  Comparison with the analytical solution for each t using the basic D1Q5 scheme (a) and the proposed D1Q7 scheme (b) for $ \delta x = 0.2 $.
    Figure 6.  The numerical results for D1Q5 (a) and D1Q7 (b) for propagation of the soliton from $ t = 0 $ to $ t = 10 $.

    Remark 3.1. The colormap default in the MATLAB color box represents the solution of the D1Q5 simulation and the colormap jet stands for the solution of the D1Q7 simulation.

    Then we do some numerical accuracy experiments. Several simulations are performed at different lattice resolutions $ \delta x = \left \{ 0.05, 0.1, 0.15, 0.2 \right \} $, and the value of $ \delta t $ is $ 1.0 \times 10^{-5} $. Based on the GRE at $ t = 1 $ and $ t = 2 $, the slopes of the fitting lines are very close to 2 in Figure 7, which indicates all of three models have a second-order accuracy in space. When $ \delta t = \left \{ 1.0 \times 10^{-5}, 5.0 \times 10^{-5}, 1.0 \times 10^{-4} \right \} $, $ \delta x = 0.2 $. Based on the GRE at $ x = 1 $ and $ x = 2 $, the slopes of the fitting lines are very close to 1 in Figure 8, which indicates all of three models have a first-order accuracy in time. The results are the same as Eq (2.20).

    Figure 7.  The numerical spatial accuracy diagram of D1Q5 (a) and D1Q7 (b) in Example 3.2.
    Figure 8.  The numerical time accuracy chart of D1Q5 (a) and D1Q7 (b) in Example 3.2.

    Example 3.3. For $ a\left (x, t \right) = 2sech^{2} \left (t \right) ln(cosh(\frac{x}{2})), \; b\left (x, t \right) = C_{1} sech^{2} \left (t \right) $. Equation (1.5) becomes

    $ \frac{\partial u}{\partial t} +u\frac{\partial }{\partial x} \left [ 2sech^{2} \left ( t \right ) ln(cosh(\frac{x}{2} ))u \right ] +\frac{\partial ^{2} }{\partial x^{2} } \left [ C_{1}sech^{2} \left ( t \right )u \right ] = 0, $

    analytical solution [22]

    $ u(x,t)=C3C2±C1C2r21r+cosh(k(x)+c(t)+l)+C1C2sinh(k(x)+c(t)+l)r+cosh(k(x)+c(t)+l),
    $

    where

    $ k(x)=C1C2a0[ln(cosh((C1C4a0x)C3C1a0C2))ln(cosh(C4C3a0C2))],
    $
    $ \quad c(t) = -C_{3} \int_{0}^{t} \alpha(\tau) \mathrm{d} \tau, \alpha \left ( t\right ) = sech ^{2}\left ( t\right ). $

    When the analytical solution is positive, and the parameters are $ a_{0} = 1, \; r = 2 $, $ C_{1} = 2, \; C_{2} = C_{3} = 1 $, $ C_{4} = l = 0 $, lattice parameters are assumed to be $ \delta x = 0.2, \; \tau = 1 $, and the calculation area is fixed at $ \left[-10, 10 \right] $.

    We present the comparison between detailed numerical results and analytical solution. Figure 9 shows the two-dimensional visual comparisons at some different times. The space-time evolution graph of the numerical results is shown in Figure 10. The numerical results show that the scheme has good long-time numeric simulation for the Burgers equation with variable coefficient in space and time. All of them clearly show that the numerical results agree with the analytical solutions well.

    Figure 9.  Comparison with analytical solution for each t using the D1Q5 (a) and D1Q7 (b) when $ \delta x = 0.2 $.
    Figure 10.  The numerical results for D1Q5 (a) and D1Q7 (b) for propagation of the soliton from $ t = 0 $ to $ t = 10 $.

    The calculation cost is given in Table 4 so as to compare the calculation time of the proposed scheme on D1Q7 and the basic scheme on D1Q5. We find that D1Q7 costs less than D1Q5. This improvement is even more significant as spatial step $ \delta x $ is increased.

    Table 4.  The calculation cost is compared with D1Q5 and D1Q7 in various $ \delta x $.
    $ \delta_x $ D1Q5 D1Q7
    0.05 $ 4534.35\pm 5.38 $ $ 675.75\pm2.15 $
    0.1 $ 180.45\pm 2.43 $ $ 56.00\pm1.60 $
    0.2 $ 19.30\pm1.05 $ $ 8.00\pm0.00 $

     | Show Table
    DownLoad: CSV

    We derive the nonlinear Burgers equation with variable coefficient from the LB equation by using the CE analysis and the DTE methods. Based on comparative observations in Section 2, we suggest that when we derive LB models for macro equation, it is best to start with using the CE analysis of general equilibrium states. After obtaining some results on the equilibrium distribution function, we can apply DTE method to conduct error analysis to improve the stability and accuracy of the model.

    In this study, we derive the LB model of Burgers by using a SRT-LB model and compare the LB solutions of the model with the corresponding analysis, which verifies the accuracy of the model. What's more, the improvement of accuracy by increasing lattice speeds can be regarded as a compensation for the deteriorated precision due to increased $ \tau $. Increasing the spatial step size $ \delta x $ can reduce the computational cost. In the future, we prepare to use LB model to simulate more non-linear PDEs with variable coefficient.

    This study is supported by National Natural Science Foundation of China (Grant Nos. 11761005, 11861003), the Natural Science Foundation of Ningxia (2021AAC03206), Postgraduate Innovation Project of North Minzu University (YCX21156) and the First-Class Disciplines Foundation of Ningxia (Grant No. NXYLXK2017B09).

    All authors declare no conflicts of interest in this paper.

    [1] J.-B. Baillon and R. Cominetti, Markovian traffic equilibrium, Math. Prog., 111 (2008), 33-56. doi: 10.1007/s10107-006-0076-2
    [2] M. Beckmann, C. McGuire and C. Winsten, "Studies in Economics of Transportation," Yale University Press, New Haven, 1956.
    [3] F. Benmansour, G. Carlier, G. Peyré and F. Santambrogio, Numerical approximation of continuous traffic congestion equilibria, Netw. Heterog. Media, 4 (2009), 605-623.
    [4] A. Braides, "$\Gamma$-Convergence for Beginners," Oxford Lecture Series in Mathematics and its Applications, 22, Oxford University Press, Oxford, 2002.
    [5] L. Brasco, G. Carlier and F. Santambrogio, Congested traffic dynamics, weak flows and very degenerate elliptic equations, Journal de Mathématiques Pures et Appliquées (9), 93 (2010), 652-671.
    [6] G. Carlier, C. Jimenez and F. Santambrogio, Optimal transportation with traffic congestion and Wardrop equilibria, SIAM J. Control Optim., 47 (2008), 1330-1350. doi: 10.1137/060672832
    [7] G. Dal Maso, "An Introduction to $\Gamma-$Convergence," Progress in Nonlinear Differential Equations and their Applications, 8, Birkhäuser Boston, Inc., Boston, MA, 1993.
    [8] C. Villani, "Topics in Optimal Transportation," Graduate Studies in Mathematics, 58, American Mathematical Society, Providence, RI, 2003.
    [9] J. G. Wardrop, Some theoretical aspects of road traffic research, Proc. Inst. Civ. Eng., 2 (1952), 325-378.
  • This article has been cited by:

    1. Yingxin Feng, Houping Dai, Xuedan Wei, Numerical solutions to the Sharma–Tasso–Olver equation using Lattice Boltzmann method, 2023, 95, 0271-2091, 1546, 10.1002/fld.5219
  • Reader Comments
  • © 2012 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(4168) PDF downloads(100) Cited by(7)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog