
This paper deals with a two-step explicit predictor-corrector approach so-called the two-step MacCormack formulation, for solving the one-dimensional nonlinear shallow water equations with source terms. The proposed two-step numerical scheme uses the fractional steps procedure to treat the friction slope and to upwind the convection term in order to control the numerical oscillations and stability. The developed scheme uses both forward and backward difference formulations in the predictor and corrector steps, respectively. The linear stability of the constructed technique is deeply analyzed using the Von Neumann stability approach whereas the convergence rate of the proposed method is numerically obtained in the L2-norm. A wide set of numerical examples confirm the theoretical results.
Citation: Rubayyi T. Alqahtani, Jean C. Ntonga, Eric Ngondiep. Stability analysis and convergence rate of a two-step predictor-corrector approach for shallow water equations with source terms[J]. AIMS Mathematics, 2023, 8(4): 9265-9289. doi: 10.3934/math.2023465
[1] | Kamel Mohamed, H. S. Alayachi, Mahmoud A. E. Abdelrahman . The mR scheme to the shallow water equation with horizontal density gradients in one and two dimensions. AIMS Mathematics, 2023, 8(11): 25754-25771. doi: 10.3934/math.20231314 |
[2] | Dalal Khalid Almutairi, Ioannis K. Argyros, Krzysztof Gdawiec, Sania Qureshi, Amanullah Soomro, Khalid H. Jamali, Marwan Alquran, Asifa Tassaddiq . Algorithms of predictor-corrector type with convergence and stability analysis for solving nonlinear systems. AIMS Mathematics, 2024, 9(11): 32014-32044. doi: 10.3934/math.20241538 |
[3] | Hasim Khan, Mohammad Tamsir, Manoj Singh, Ahmed Hussein Msmali, Mutum Zico Meetei . Numerical approximation of the time-fractional regularized long-wave equation emerging in ion acoustic waves in plasma. AIMS Mathematics, 2025, 10(3): 5651-5670. doi: 10.3934/math.2025261 |
[4] | Hassan Ranjbar, Leila Torkzadeh, Dumitru Baleanu, Kazem Nouri . Simulating systems of Itô SDEs with split-step (α,β)-Milstein scheme. AIMS Mathematics, 2023, 8(2): 2576-2590. doi: 10.3934/math.2023133 |
[5] | Douglas R. Anderson, Masakazu Onitsuka . A discrete logistic model with conditional Hyers–Ulam stability. AIMS Mathematics, 2025, 10(3): 6512-6545. doi: 10.3934/math.2025298 |
[6] | Mohammad Sajid, Biplab Dhar, Ahmed S. Almohaimeed . Differential order analysis and sensitivity analysis of a CoVID-19 infection system with memory effect. AIMS Mathematics, 2022, 7(12): 20594-20614. doi: 10.3934/math.20221129 |
[7] | Zhifang Li, Huihong Zhao, Hailong Meng, Yong Chen . Variable step size predictor design for a class of linear discrete-time censored system. AIMS Mathematics, 2021, 6(10): 10581-10595. doi: 10.3934/math.2021614 |
[8] | Danuruj Songsanga, Parinya Sa Ngiamsunthorn . Single-step and multi-step methods for Caputo fractional-order differential equations with arbitrary kernels. AIMS Mathematics, 2022, 7(8): 15002-15028. doi: 10.3934/math.2022822 |
[9] | Eric Ngondiep . An efficient two-level factored method for advection-dispersion problem with spatio-temporal coefficients and source terms. AIMS Mathematics, 2023, 8(5): 11498-11520. doi: 10.3934/math.2023582 |
[10] | Zongcheng Li, Jin Li . Linear barycentric rational collocation method for solving a class of generalized Boussinesq equations. AIMS Mathematics, 2023, 8(8): 18141-18162. doi: 10.3934/math.2023921 |
This paper deals with a two-step explicit predictor-corrector approach so-called the two-step MacCormack formulation, for solving the one-dimensional nonlinear shallow water equations with source terms. The proposed two-step numerical scheme uses the fractional steps procedure to treat the friction slope and to upwind the convection term in order to control the numerical oscillations and stability. The developed scheme uses both forward and backward difference formulations in the predictor and corrector steps, respectively. The linear stability of the constructed technique is deeply analyzed using the Von Neumann stability approach whereas the convergence rate of the proposed method is numerically obtained in the L2-norm. A wide set of numerical examples confirm the theoretical results.
Most open-channel flows of interest in the physical, hydrological, biological, engineering and social sciences are unsteady and can be considered to be one-dimensional (1D). In this paper, we are interested in the numerical solutions of one-dimensional shallow water equations with source terms introduced in [1,2] and still widely used in modeling flows in rivers, lakes and coastal areas as well as atmospheric and oceanic flows in some regimes. In the case of a prismatic channel, the shallow water equations with source terms read as follows
{∂A∂t+∂Q∂x=v,fort∈(0,T1) and x∈Ω=(0,L), ∂Q∂t+gAT∂A∂x+∂∂x(Q2A)=gA(S0−Sf),fort∈(0,T1) and x∈Ω=(0,L), | (1.1) |
where the bottom (or bed) slope (S0) and the friction slope (Sf) (see [1]) are defined as
S0=¯τPρgA and Sf=Q|Q|K2, | (1.2) |
where v=v(t,x) is the lateral inflow per unit length along the channel, T1 represents the time interval length, L denotes the rod interval length, A=A(t,x) and Q=Q(t,x) are the cross-section and the discharge, respectively, g denotes the acceleration of gravity. T=T(t,x) represents the top width assumed to be constant, ¯τ designates the average shear stress on the water from the channel boundary, ρ is the fluid density and P=P(x,y(t,x)) denotes the wetted perimeter (i.e., the length of the boundary of the cross-section that is underwater for a given height of water (y)). As in [1], the conveyance for a compact channel is defined as
K:=K(x,y)=1.49n1A(x,y)R(x,y)2/3, | (1.3) |
where R=A/P (see for example, [1,3]) denotes the hydraulic radius and n1 represents the manning's roughness coefficient.
Since the parameters g and T are positive constants, it is easy to see that gAT∂A∂x=∂∂x(gA22T). Using this equation together with equality ∂∂x(gA22T)+∂∂x(Q2A)=∂∂x(gA22T+Q2A), the second equation of the system (1.1) is equivalent to
∂Q∂t+∂∂x(gA22T+Q2A)=gA(S0−Sf), for t∈(0,T1) and x∈Ω=(0,L). |
This equation combined with the first equation in (1.1) results in
∂A∂t+∂Q∂x=v,fort∈(0,T1) and x∈Ω=(0,L), ∂Q∂t+∂∂x(gA22T+Q2A)=gA(S0−Sf),fort∈(0,T1) and x∈Ω=(0,L), |
which can be rewritten as
∂W∂t+∂F∂x=S, | (1.4) |
where W=(A,Q)T,F=(Q,gA22T+Q2A)T, and S=(v,gA(S0−Sf))T. Equation (1.4) emphasizes the conservative character of system (1.1).
The one-dimensional shallow water equations with source terms (1.4) are highly nonlinear and therefore do not have global analytical solutions [1]. When solving the system of balance laws (1.4) numerically, one typically faces several difficulties. One difficulty stems from the fact that many physically relevant solutions of (1.4) are small perturbations of steady-state solutions. So, using a wrong balance between the flux and geometric source term in Eq (1.4), the solution may develop spurious waves of a magnitude that can become larger than the exact solution. Another drawback occurs when the cross-section is very small. In that case, even small numerical oscillations in the computed solution can result in a very large discharge, which is not only physically irrelevant, but cause the numerical scheme to break down. To overcome these numerical challenges, one needs to use a numerical scheme that is both well-balanced and positivity-preserving.
A number of well-balanced and positivity-preserving numerical methods frequently used in the models based on the shallow water equations have been proposed in the literature [4,5], or on the boussinesq equations, which are reduced to shallow water equations, in order to simulate breaking waves [6,7,8,9]. Although the MacCormack scheme is less accurate than the more recent methods, it is commonly used for engineering problems due to its greater simplicity. So, we have to approximate the exact solution of problem (1.4) by a numerical method based on a two-level predictor-corrector scheme. This algorithm lies in the class of higher order finite difference methods (temporal second-order convergent and fourth-order accurate in space) which provide an effective way of joining the spectral method for accuracy and robust characteristics of finite difference schemes. For example, to compute unsteady flow specifically in the presence of discontinuity, inherent dissipation and stability, one such widely used method is the MacCormack method [10]. This technique has been used successfully to provide a time-accurate solution for fluid flow and aeroacoustics problems. The applications of this technique to 1D shock tubes and 2D acoustic scattering problems provide good results when compared with the analytical solution. The original MacCormack introduces a simpler variation of the Lax-Wendroff scheme which is basically a two-step scheme with second order Taylor series expansion in time and fourth order in spatial accuracy [10,11]. This algorithm is computationally efficient and easy to implement which can be appropriate to obtain reliable results. By using this scheme with two nodes, the flow field can be simulated for unsteady flows especially for shallow water problems in the presence of discontinuity and strict gradient conditions. Furthermore, to capture the fluid flow in transition over long periods of time and distance, numerical spatial derivative are required to be determined in a few grid points while error-controlled can be accurately computed. The authors [12,13,14,15,16,17] extended the MacCormack scheme [10] to an implicit-explicit scheme (by coupling the original MacCormack approach with Crank-Nicolson method), an implicit compact differencing scheme and a three-level time-split MacCormack procedure (by splitting the derivative operator of the scheme into one-sided forward and backward difference formulations). The one-sided nature of the time-split MacCormack approach is an essential advantage especially when severe gradients are present.
In [18,19,20] the authors compared the Lax-Wendroff scheme to many numerical methods of high order accuracy, such as, the linear Central Weighed Essential Non-Oscillatory (CWENO) scheme which is superior to full nonlinear CWENO method, to high-resolution TVD conservative procedures along with high order Central Schemes for hyperbolic systems of conservative Laws [3,21] and to Central-upwind schemes for the shallow water system [22]. In a search for stable and more accurate shock capturing numerical approach, they observed that the Lax-Wendroff technique is one of the most frequently encountered in the literature related to classical shock-capturing schemes. However, difficulties have been reported when trying to include source terms in the discretization and to keep the second order accuracy at the same time [23]. The proposed two-level solver which can be considered a predictor-corrector version of the Lax-Wendroff algorithm provides a reasonably good result at discontinuities. The developed method is more easier to apply than the Lax-Wendroff method because the Jacobian does not appear. For linear equations, both algorithms provide similar amplification factors and stability constraints (for instance, see [24], P. 202–206). It is worth noticing to mention that the solutions obtained for a given problem at the same courant number are different from those obtained using the Lax-Wendroff scheme. This situation is due to the switched differencing in the predictor, corrector and the nonlinear nature of the governing equations. It should be noted that reversing the differencing in the predictor and corrector steps leads to quite different results. Furthermore, the explicit MacCormack time discretization for nonlinear Burgers equations (which can serve as model equations for a wide set of nonlinear PDEs: Navier-Stokes equations, Stokes-Darcy equations, Parabolized Navier-Stokes equations, ...) give a suitable stability restriction which should be used with an appropriate safety factor (see [24], P. 227–228). As regards the proposed predictor-corrector formulation, like other explicit approaches, it requires a time-step limitation. In general, the maximum time-step allowable in the natural MacCormack scheme applied to linear hyperbolic equations is limited by the CFL condition, as for all explicit finite difference methods. However, the considered overland flow equations are nonlinear and a rigorous stability analysis of numerical techniques is exceedingly difficult. The source terms place additional and problem-dependent restrictions on the maximum admissible time-step for stability. Therefore, the CFL condition can only be considered as a general guideline here, and the maximum allowable time-step for any particular problem will be less than predicted by the CFL condition and determined by numerical experimentation (see [25], page 223).
Most recently, several researchers have deeply analyzed implicit/explicit (for example: Crank-Nicolson/MacCormack) approach, compact finite difference methods, two-level factored Crank-Nicolson scheme, three-level time-split MacCormack algorithms, central upwing schemes and Lax-Wendroff technique in the approximate solutions of mixed Stokes-Darcy's model, evolutionary reaction-diffusion equation, nonstationary convection-diffusion equation, time-dependent convection-diffusion-reaction problems, unsteady coupled Burgers'equations, Navier-Stokes equations, time fractional equations and shallow water problems. For more details, we refer the readers to [5,9,12,13,17,22,26,27,28,29,30,31,32,33,34,35,36,37,38,39,40,41]. In this paper, we are interested in a computed solution of a nonlinear system (1.1), but in the sense of a linear stability condition and convergence rate of the numerical scheme. In particular, we consider the case where the channel is prismatic and the interesting result is that the algorithm is second-order accurate in time and spatial fourth-order convergent, whereas the stability limitation is not similar to the CFL condition widely studied in the literature for hyperbolic PDEs. However, while the stability requirement is highly unusual, the result has a potential positive implication since the stability constraint controls the CFL condition. Indeed the nice feature is that, as required in a stability context, we normally find a linear stability requirement which can be considered as a necessary condition. In addition, it follows from this analysis that instability occurs when Δt is greater than Δtmax=(Δt)CFL. More specifically, the attention is focused in the following three items:
(i1) Full description of a two-step predictor-corrector method for solving one-dimensional shallow water equations with source terms.
(i2) Stability analysis of the proposed algorithm: this item together with item (i1) represent our original contributions and they improve some works studied in the literature [3,12,42,43].
(i3) Numerical experiments which considers the convergence rate of the method, the simulation of the numerical solution together with the analytical ones, and regarding the effectiveness of the proposed algorithm according to the theoretical analysis given in the first two items.
The paper is organized as follows. Section 2 deals with a full description of a two-step predictor-corrector formulation for one-dimensional shallow water equations with source terms. A necessary condition of stability of the proposed numerical scheme is deeply analyzed in Section 3. In Section 4, some numerical experiments which consider the convergence rate of the developed approach and some simulations are presented and discussed. We draw the general conclusion and present the future works in Section 5.
In this section, we give a detailed description of a two-step explicit predictor-corrector algorithm for the system of nonlinear Eq (1.4). The proposed scheme uses the forward difference in the predictor step while the corrector step considers the backward difference. Since the aim of this section is to analyze the linear stability of the method, without loss of generality we should use a constant time step Δt and mesh size Δx. However, this assumption does not compromise the result on the stability. Let K1 and M1 be two positive integers. Set xr=rΔx,ti=iΔt be the discrete points and let the superscript and subscript be the time level and space level, respectively, of the approximation. We denote Wir=(Air,Qir)T be the approximate solution of equations (1.4), provided by the constructed two-step MacCormack technique and W(ti,xr)=(A(ti,xr),Q(ti,xr))T be the analytical ones. Furthermore, the domain Ω=(0,L) is partitioned into M1+1 grid points {xr: r=0,1,...,M1}, whereas the time interval (0,T) is subdivided into K1+1 discrete points {ti: i=0,1,...,K1}. Using this, a full description of the proposed method for solving the system (1.4) reads: Given Wir, find an approximate solution wi+1r, for 0≤i≤K1−1 and 0≤r≤M1, satisfying
Predictor step: solve Eq (2.1) for predicted value
W¯i+1r=Wir−ΔtΔx(Fir+1−Fir)+ΔtSir. | (2.1) |
Corrector step: use the predicted value obtained in Eq (2.1) to compute the corrected one
Wi+1r=12[Wir+W¯i+1r−ΔtΔx(F¯i+1r−F¯i+1r−1)+ΔtS¯i+1r]. | (2.2) |
Lemma 2.1. Let K1 and M1 be two positive integers. Setting ti=iΔt, for i=0,1,...,K1, and xr=rΔx, for r=0,1,...,M1, where Δt and Δx are time step and mesh size, respectively. The proposed numerical scheme for solving the system of nonlinear Eq (1.4) is given by: for i=0,1,...,K1−1, and r=0,1,...,M1−1,
Predictor step: solve Eqs (2.3) and (2.4) for predicted values A¯i+1r and Q¯i+1r
A¯i+1r=Air−ΔtΔx(Qir+1−Qir)+Δtvir, | (2.3) |
Q¯i+1r=Qir−ΔtΔx{g2T((Air+1)2−(Air)2)+(Qir+1)2Air+1−(Qir)2Air}+gPΔt(¯τρg−n211.492P13Qir|Qir|(Air)73). | (2.4) |
Corrector step: use A¯i+1r and Q¯i+1r obtained in Eqs (2.3) and (2.4) to compute Ai+1r and Qi+1r
Ai+1r=12{Air+A¯i+1r−ΔtΔx(Q¯i+1r−Q¯i+1r−1)+Δtv¯i+1r}, | (2.5) |
Qi+1r=12{Qir+Q¯i+1r−ΔtΔx{g2T[(A¯i+1r)2−(A¯i+1r−1)2]+(Q¯i+1r)2A¯i+1r−(Q¯i+1r−1)2A¯i+1r−1}+ |
gPΔt(¯τρg−n211.492P13Q¯i+1r|Q¯i+1r|(A¯i+1r)73)}. | (2.6) |
Approximations (2.3)–(2.6) are subjects to appropriate initial and boundary conditions.
Proof. Since W=(A,Q)T, combining Eqs (2.1) and (2.2) to get
A¯i+1r=Air−ΔtΔx(Qir+1−Qir)+Δtvir, | (2.7) |
Q¯i+1r=Qir−ΔtΔx{g2T((Air+1)2−(Air)2)+(Qir+1)2Air+1−(Qir)2Air}+gΔtAir((S0)ir−(Sf)ir), | (2.8) |
Ai+1r=12{Air+A¯i+1r−ΔtΔx(Q¯i+1r−Q¯i+1r−1)+Δtv¯i+1r}, | (2.9) |
Qi+1r=12{Qir+Q¯i+1r−ΔtΔx{g2T[(A¯i+1r)2−(A¯i+1r−1)2]+(Q¯i+1r)2A¯i+1r−(Q¯i+1r−1)2A¯i+1r−1} |
+gΔtA¯i+1r((S0)¯i+1r−(Sf)¯i+1r)}. | (2.10) |
Substituting Eq (1.3) into relation (1.2) yields
Sf=n211.492Q|Q|A2R4/3=n211.492Q|Q|A10/3P4/3. | (2.11) |
In a similar way, substituting (1.2) and (2.11) into Eqs (2.8) and (2.10), respectively, results in
Q¯i+1r=Qir−ΔtΔx{g2T((Air+1)2−(Air)2)+(Qir+1)2Air+1−(Qir)2Air}+gPΔt(¯τρg−n211.492P13Qir|Qir|(Air)73), | (2.12) |
and
Qi+1r=12{Qir+Q¯i+1r−ΔtΔx{g2T[(A¯i+1r)2−(A¯i+1r−1)2]+(Q¯i+1r)2A¯i+1r−(Q¯i+1r−1)2A¯i+1r−1} |
+gPΔt(¯τρg−n211.492P13Q¯i+1r|Q¯i+1r|(A¯i+1r)73)}. | (2.13) |
A combination of Eqs (2.7),(2.9),(2.12) and (2.13) provides a full description of the developed two-step predictor-corrector scheme, that is,
A¯i+1r=Air−ΔtΔx(Qir+1−Qir)+Δtvir, |
Q¯i+1r=Qir−ΔtΔx{g2T((Air+1)2−(Air)2)+(Qir+1)2Air+1−(Qir)2Air}+gPΔt(¯τρg−n211.492P13Qir|Qir|(Air)73). |
Ai+1r=12{Air+A¯i+1r−ΔtΔx(Q¯i+1r−Q¯i+1r−1)+Δtv¯i+1r}, |
Qi+1r=12{Qir+Q¯i+1r−ΔtΔx{g2T[(A¯i+1r)2−(A¯i+1r−1)2]+(Q¯i+1r)2A¯i+1r−(Q¯i+1r−1)2A¯i+1r−1} |
+gPΔt(¯τρg−n211.492P13Q¯i+1r|Q¯i+1r|(A¯i+1r)73)}. |
Here, the terms A¯i+1 and Q¯i+1 are "predicted" values of A and Q, respectively, at the time level i+1. Assuming that the superscript ¯i+1 is a time-level, it is easy to see that the proposed MacCormack algorithm is a three-level method, so the initial conditions A0 and Q0 are needed to begin the algorithm. However, appropriate initial and boundary conditions must be specified. Further, the presence a cross-section in the denominator of several terms disallows zero cross sections. Therefore, a minimum cross section should be assigned to each node that is ponded. It is primarily the term A10/3 in the denominator of the friction slope term given by relation (2.11) that limits the magnitude of the minimum cross section and discharge. When the cross sections are very small, the friction slope is very large compared with the other terms in the second equation of system (1.1). As cross sections increase rapidly during the early stages of flow development, the friction slope term magnitude changes much faster than the other terms. This phenomenon renders the second equation in system (1.1) stiff and severely limits the maximum admissible time step for stability. Indeed, this phenomenon likely forced many researchers to consider the smallest time-steps relative to their mesh size (courant number ≪1) and keep lateral inflows and initial cross sections large [25].
This section deals with a linear stability analysis of a three-level MacCormack procedure in a numerical solution of the system of nonlinear PDEs (1.1). In the literature, a wide set of multi-level schemes for solving shallow water equations and unsteady transport problems have been used to advance the solution in time. The most popular of these techniques is the original explicit MacCormack approach, which was widely applied to obtain a time-accurate solution for fluid flow and aeroacoustic problems. Compared to a broad range of explicit methods such as, Lax-Wendroff formulation, the natural MacCormack scheme is computationally efficient, easy to implement and it provides less computing time. In fact, this algorithm lies in a class of higher order finite difference methods for solving unsteady flow specifically in the presence of discontinuity, inherent dissipative, dispersion and stability [10], whereas the Lax-Wendroff method indicates some difficulties to preserve the stability and the temporal second-order accuracy at the same time [23]. However, for linear problems both numerical schemes have provided almost the same amplification factor and stability restriction (for instance, see [24], P: 227–228). To analyze the stability of the proposed approach, we apply the Fourier stability method to the difference Eqs (2.3)–(2.6), by computing the amplification factor to obtain an algebraic criterion. Following the Von Neumann stability analysis for necessary condition of stability, we assume that the approximate solution Wir=(Air,Qir) can be expressed in the form of Fourier series as
Air=eatieˆirϕ and Qir=ebtieˆirϕ, | (3.1) |
where b=b1+ˆib2, a=a1+ˆia2, with aj,bj∈R, 1≤j≤2,ϕ=kΔx is the phase angle, k denotes the wave number, Δx represents the mesh grid and ˆi indicates the imaginary unit. For the sake of readability, we assume in the following that a2=b2 and the ratios |v||A|−1 and |v||Q|−1 are very small.
The analysis of the stability of the proposed method requires the following remark.
Remark 3.1. Since the considered problem is nonlinear, the solution may contain discontinuity even if the initial conditions are smooth enough. To overcome this numerical challenge, we should assume that the phase angle ϕ=kΔx satisfies |ϕ|<<π. In fact, the method could be generally stabilized by adding additional dissipation to the scheme without affecting the order of accuracy.
The following Lemma gives the "temporary" linear stability constraint of the two-step predictor-corrector algorithm described in section 2.
Lemma 3.1. A necessary condition of stability for the numerical scheme (2.5) is given by
Δt3Δx2(1+2Δt3Γ0μˆA−43)Γ0μ3ˆA−43|ϕ|2≤32(1−ϵ2), | (3.2) |
where 0≤ϵ<1, μ=max0≤i≤K1μi, with μi=|Qi||Ai|, ˆA=min0≤i≤K1|Ai|=min0≤i≤K1|ea1ti|≠0 (according to relation (3.1)), for any a1∈R, Γ0=gn211.492P43 and ϕ=kΔx, where k≠0 is the wave number.
Proof. Firstly, we should provide a simple expression of Ai+1. Combining relations (2.3) and (2.4), simple computations give
Q¯i+1r−Q¯i+1r−1=Qir−Qir−1−ΔtΔx{g2T[(Air+1)2−2(Air)2+(Air−1)2]+(Qir+1)2Air+1−2(Qir)2Air+(Qir−1)2Air−1} |
−Δtgn211.492P43(Qir|Qir|(Air)73−Qir−1|Qir−1|(Air−1)73), | (3.3) |
and
Air+A¯i+1r=2Air−ΔtΔx(Qir+1−Qir)+Δtvir. | (3.4) |
Substituting Eqs (3.3) and (3.4) into (2.5), this results in
Ai+1r=Air−Δt2Δx(Qir+1−Qir−1)+12(ΔtΔx)2{g2T[(Air+1)2−2(Air)2+(Air−1)2]+(Qir+1)2Air+1 |
−2(Qir)2Air+(Qir−1)2Air−1}+Δt22Δxgn211.492P43(Qir|Qir|(Air)73−Qir−1|Qir−1|(Air−1)73)+12Δt(vir+v¯i+1r). | (3.5) |
Neglecting the last term in (3.5), we obtain
Ai+1r=Air−Δt2Δx(Qir+1−Qir−1)+12(ΔtΔx)2{g2T[(Air+1)2−2(Air)2+(Air−1)2]+(Qir+1)2Air+1 |
−2(Qir)2Air+(Qir−1)2Air−1}+Δt22ΔxΓ0(Qir|Qir|(Air)73−Qir−1|Qir−1|(Air−1)73), | (3.6) |
where
Γ0=gn211.492P43. | (3.7) |
Indeed, since the terms |v(x,t)||A(x,t)|−1 and |v(x,t)||Q(x,t)|−1,∀(x,t)∈[0,L]×[0,T1], are too small, the tracking of the last term in (3.5) does not compromise the result.
Substituting Eq (3.1) into relation (3.6) to get
ea(ti+Δt)eˆikxr=eatieˆikxr−Δt2Δx(ebtieˆik(xr+Δx)−ebtieˆik(xr−Δx))+ |
12(ΔtΔx)2{g2T[e2atie2ˆik(xr+Δx)−2e2atie2ˆikxr+e2atie2ˆik(xr−Δx)]+e2(b−a)ti−2e2(b−a)ti+e2(b−a)ti} |
+Δt22ΔxΓ0(ebtieˆikxr|ebti|e73atie73ˆikxr−ebtieˆik(xr−Δx)|ebti|e73atie73ˆik(xr−Δx)). | (3.8) |
Multiplying both sides of relation (3.8) by e−atie−ˆikxr to obtain
eaΔt=1−Δt2Δx(eˆiϕ−e−ˆiϕ)e(b−a)ti+12(ΔtΔx)2g2T[eatieˆikxre2ˆiϕ−2eatieˆikxr+eatieˆikxre−2ˆiϕ] |
+Δt22ΔxΓ0|ebti|e(b−103a)tie−73ˆikxr(1−e43ˆiϕ). | (3.9) |
Using the identities: eˆiϕ−e−ˆiϕ=2ˆisin(ϕ),e2ˆiϕ−2+e−2ˆiϕ=−4sin2(ϕ) and 1−e43ˆiϕ=2sin2(23ϕ)+ˆisin(43ϕ), Eq (3.9) becomes
eaΔt=1−ˆiΔtΔxsin(ϕ)e(b1−a1)ti−(ΔtΔx)2gTsin2(ϕ)ea1ti[cos(a2ti+kxr)+ˆisin(a2ti+kxr)]−ˆiΔt2ΔxΓ0sin(23ϕ)cos(23ϕ)e(2b1−103a1)ti(cos(73[a2ti+kxr])−ˆisin(73[a2ti+kxr]))−2(ΔtΔx)2sin2(12ϕ)e2(b1−a1)ti=1−ˆiΔtΔxμisin(ϕ)−(ΔtΔx)2gTsin2(ϕ)ea1ti[cos(a2ti+kxr)+ˆisin(a2ti+kxr)]−ˆiΔt2ΔxΓ0sin(23ϕ)cos(23ϕ)e(2b1−103a1)ti(cos(73[a2ti+kxr])−ˆisin(73[a2ti+kxr]))−2(ΔtΔx)2(μi)2sin2(12ϕ), | (3.10) |
where
μi=|Qi||Ai|=eb1tiea1ti=e(b1−a1)ti. | (3.11) |
Setting
α2=73(a2ti+kxr); γ2=Γ0(μi)2e−43a1tie−ˆiα2; α3=a2ti+kxr γ1=μi; γ4=γ21; γ3=gTea1tieˆiα3. | (3.12) |
A combination of Eqs (3.12) and (3.10) yields
eaΔt=1−ˆiΔtΔxγ1sin(ϕ)−(ΔtΔx)2γ3sin2(ϕ)−ˆiΔt2Δxγ2sin(23ϕ)cos(23ϕ)−2(ΔtΔx)2γ4sin2(12ϕ)=1−Δt2Δx|γ2|sinα2sin(23ϕ)cos(23ϕ)−(ΔtΔx)2[|γ3|cosα3sin2(ϕ)+2|γ4|sin2(12ϕ)]−ˆi{ΔtΔx|γ1|sin(ϕ)+Δt2Δx|γ2|cosα2sin(23ϕ)cos(23ϕ)+(ΔtΔx)2|γ3|sinα3sin2(ϕ)}. | (3.13) |
Of course the aim of this report is to give a general picture of the necessary condition of stability. Since the formulae can become quite heavy, for the convenient of writing, the proof of our results will consider Remark 3.1. However, this leads to a linear stability condition which can be observed as a necessary condition of stability.
We start the analysis with some extreme cases: |ϕ|=π and ϕ=0.
∙ Case 1. For |ϕ|=π, it comes from Eq (3.13) that the amplification factor becomes
eaΔt=1+√3Δt24Δx|γ2|sinα2−2(ΔtΔx)2|γ4|−ˆi√3Δt24Δx|γ2|cosα2. |
The squared modulus of the amplification factor gives
|eaΔt|2=1+3Δt416Δx2|γ2|2+4(ΔtΔx)4|γ4|2+2(√3(Δt)24Δx|γ2|(1+Δt2|γ4|)sinα2−2(ΔtΔx)2|γ4|). |
But there exist values of α2 for which sinα2=1. So |eaΔt|2>1. Thus, the scheme is unconditionally unstable.
∙ Case ϕ=0. In that case, the amplification factor given by Eq (3.13) provides eaΔt=1 which implies |eaΔt|=1. Thus, the numerical scheme is stable. This suggests that the proposed technique is not dissipative in the sense of Kreiss [44] and when applied to the system of nonlinear Eq (1.1). That is, the computations should become unstable in certain circumstances. This instability is entirely due to the non-linearity of the equations, since the same scheme applied to linear shallow water equations without source terms does not diverge, although strong oscillations are generated (for example, see [10,17,45]).
∙ Case where 0<|ϕ|<<π. Using the Taylor expansion around ϕ=0, and neglecting high-order terms, the squared modulus of the amplification factor given by (3.13) is approximated as
|eaΔt|2=(1−2Δt23Δx|γ2|ϕsinα2)2+(ΔtΔx|γ1|+2Δt23Δx|γ2|cosα2)2ϕ2. | (3.14) |
For |eaΔt|2 to be less than one, the quantity (1−2Δt23Δx|γ2|ϕsinα2)2+(ΔtΔx|γ1|+2Δt23Δx|γ2|cosα2)2ϕ2 must be less than one. This implies
|1−2Δt23Δx|γ2|ϕsinα2|≤ϵ and ΔtΔx|γ1||1+2Δt3|γ2γ1|cosα2||ϕ|≤1−ϵ, | (3.15) |
for any value of α2, where 0≤ϵ<1. Using simple calculations, it is not hard to see that estimates given by (3.15) are equivalent to
32(1−ϵ)≤Δt2Δx|γ2|ϕsinα2≤32(1+ϵ) and 32(ϵ−1)≤ΔtΔx(32|γ1|+Δt|γ2|cosα2)|ϕ|≤32(1−ϵ), | (3.16) |
for every α2∈R. But, the second inequality in (3.16) implies 32(ϵ−1)≤ΔtΔx(32|γ1|+Δt|γ2|)|ϕ|≤32(1−ϵ). Since Δt2Δx|γ2|ϕsinα2≤Δt2Δx|γ2||ϕ|, these facts combined with the first estimate in (3.16) show that the numerical scheme (2.5) is stable if
32(1−ϵ)≤Δt2Δx|γ2||ϕ|≤32(1+ϵ) and 0≤ΔtΔx(32|γ1|+Δt|γ2|)|ϕ|≤32(1−ϵ). |
Multiplying these inequalities side by side and rearranging terms to get
(23|γ1|+4Δt9|γ2|)|γ2||ϕ|2Δt3Δx2≤1−ϵ2. | (3.17) |
Since |μi|=|QijAij|=e(b1−a1)ti, it comes from relation (3.12) that
|γ2|=Γ0(μi)2e−43a1ti and |γ1|=μi. | (3.18) |
But max0≤i≤K1|Ai|−1=[min0≤i≤K1|Ai|]−1, taking the maximum in both sides of estimates (3.17), for i=0,1,...,K1, the proof of Lemma 2 is completed thank to relations (3.18) and equality |Ai|=ea1ti.
Lemma 3.2. The numerical scheme (2.6) is linearly stable if the following estimate holds:
Δt(3W1(Δt,Δx)+1ΔxW2(Δt,Δx))≤1+√1−β∗, | (3.19) |
for i=0,1,...,K1, where β∗∈(0,1) and
W1(Δt,Δx)=max0≤i≤K1{32Pi+4(Δt+Δt2+ΔtΔx+Δt2Δx+Δt2Δx+(ΔtΔx)2)P2i}, | (3.20) |
W2(Δt,Δx)=max0≤i≤K1{|ϕ|(Pi+12Ri)+32(1+2[Δt+Δt2+ΔtΔx+Δt2Δx+Δt3Δx+(ΔtΔx)2])μiPi}, | (3.21) |
with
μi=|Qir||Air|=e(b1−a1)ti;Ri=gT|Air||μi|−1+μi;Pi=P¯τρ|Qir|−1+Γ0μi|Air|−43. | (3.22) |
Estimate (3.19) represents a necessary condition of stability for the numerical scheme (2.6).
Proof. Using the Taylor series expansion, the terms: 12(Qir+Q¯i+1r), g4T[(A¯i+1r)2−(A¯i+1r−1)2], (Q¯i+1r)2A¯i+1r−(Q¯i+1r−1)2A¯i+1r−1 and Q¯i+1r|Q¯i+1r|(A¯i+1r)73 can be approximated as:
12(Qij+Q¯i+1r)=Qir{1+Δt2ΔxC11ϕ+12ΔtC12+ˆi(Δt2Δx¯C11ϕ+12Δt¯C12)}+O(ϕ2), | (3.23) |
and
g4T[(A¯i+1r)2−(A¯i+1r−1)2]=Qir(C21+ˆi¯C21)ϕ+O(ϕ2), | (3.24) |
where
C11=gT|Ai||μi|−1sinα3; ¯C11=−gT|Ai||μi|−1cosα3−μi; C12=Pτρ|Qi|−1cosα3−Γ0μi|Ai|−43cosα2; |
¯C12=Pτρ|Qi|−1sinα3−Γ0μi|Ai|−43sinα2; C21=−g2T|Qi||μi|−1sinα3; ¯C21=g2T|Qi||μi|−1cosα3; | (3.25) |
where
α3=a2ti+kxr, and α2=73α3. | (3.26) |
(Q¯i+1r)2A¯i+1r−(Q¯i+1r−1)2A¯i+1r−1=Qirμi{H1−K21+K22+2K1K2(1−ΔtΔxμi)ϕ+ˆi(H2−2K1K2+ |
(K22−K21)(1−ΔtΔxμi)ϕ)}+O(ϕ2), | (3.27) |
where the functions H1, H2, K21, K22 and K1K2 are given by
H1=1+2ΔtC12+Δt2(C212−¯C212)+2Δt2Δx[C11C12−¯C11¯C12−μi(¯C12+ΔtC12¯C12)]ϕ, | (3.28) |
H2=Δt¯C12+Δt2¯C12C12+ΔtΔx[¯C11+μi+Δt(C11¯C12+¯C11C12+μi(2C12+Δt(C212−¯C212)))]ϕ, | (3.29) |
K21=1+(Δt2Δx)2[¯C221+4C21¯C21ϕ]+Δt2[¯C212−2¯C12¯C22ϕ]−ΔtΔx(¯C21+2C21ϕ)+ |
Δt(C12−2C22ϕ)−Δt2Δx[C12¯C21−(¯C21C22−2C12C21)ϕ], | (3.30) |
K22=(Δt2Δx)2(μi)2[C211+2C11(1−¯C11−μi)ϕ]+Δt2[C212−2C12¯C22ϕ]+ΔtΔxμiC11ϕ+ |
2Δt¯C12ϕ+Δt2Δxμi[C11¯C12−C11¯C22ϕ+¯C12(1−¯C11−μi)ϕ], | (3.31) |
and
K1K2=−{ϕ−(Δt2Δx)2μi[C11¯C21+[¯C21(1−¯C11−μi)+2C11C21]ϕ]+Δt[C12−(C12¯C22−C12)ϕ] |
+Δtμi2Δx[C11+(1−¯C11−μi−¯C21(μi)−1)ϕ]+Δt2μi2Δx[C11C12−¯C12¯C21(μi)−1+[C12(1−¯C11−μi) |
−C11C22+¯C21C22(μi)−1−2C21¯C12(μi)−1]ϕ]+Δt2[C12¯C12−(C12¯C22+¯C12C22)ϕ]}, | (3.32) |
Cls,¯Cls,l,s=1,2, are given by (3.25) and α2 and α3 follow from relation (3.26). Furthermore
C22=43Γ0μi|Air|−43sinα2 and ¯C22=43Γ0μi|Air|−43cosα2. | (3.33) |
Q¯i+1r|Q¯i+1r|(A¯i+1r)73=Qi|Qi|−43(μi)73{1+(ΔtΔx)2(μi)2ϕ2}−73{(1+ΔtΔxC11ϕ+2ΔtC12)2+ |
(ΔtΔx¯C11ϕ+2Δt¯C12)2}12{1+ΔtΔxC11ϕ+2ΔtC12+ˆi(ΔtΔx¯C11ϕ+2Δt¯C12)} |
{C31+ΔtΔx¯C32ϕ+ˆi[−¯C31+ΔtΔxC32ϕ]}73+O(ϕ2). | (3.34) |
Where the functions C1l and ˆC1l,l=1,2, are defined by relations (3.25) and (3.26),C3l and ¯C3l, l=1,2, are given by
C31=cosα2; ¯C31=−sinα2; C32=μicosα2; ¯C32=μisinα2. | (3.35) |
Combining Eqs (3.23), (3.24), (3.27) and (3.34), the amplification factor of the numerical scheme (2.6) is approximated as
ebΔt=1+Δt{12C12+C33−Γ0|Qir|−43(μi)73[1+(ΔtΔx)2(μi)2ϕ2]−73[(1+2ΔtC12+ΔtΔxC11ϕ)2 |
+(2Δt¯C12+ΔtΔx¯C11ϕ)2][(C31+ΔtΔx¯C32ϕ)2+(¯C31−ΔtΔxC32ϕ)2]76cos(θ1+73θ2)} |
−ΔtΔx{(C21−12C11)ϕ+12μi[H1−K21+K22+2K1K2(1−ΔtΔxμn)ϕ]} |
+ˆi{Δt{12¯C12−¯C33−Γ0|Qir|−43(μi)73[1+(ΔtΔx)2(μi)2ϕ2]−73[(1+2ΔtC12+ΔtΔxC11ϕ)2 |
+(2Δt¯C12+ΔtΔx¯C11ϕ)2][(C31+ΔtΔx¯C32ϕ)2+(¯C31−ΔtΔxC32ϕ)2]76sin(θ1+73θ2)} |
−ΔtΔx{(¯C21−12¯C11)ϕ+12μi[H2−2K1K2+(K22−K21)(1−ΔtΔxμi)ϕ]}}+O(ϕ2). | (3.36) |
Where ˆi is the unit complex number,
C33=P¯τρ|Qir|−1cosα3, ¯C33=P¯τρ|Qir|−1sinα3, | (3.37) |
H1, H2, K21, K22 and K1K2 are given by Eqs (3.28)–(3.32), respectively; Cls,¯Cls,s,l=1,2; come from (3.25); α2 and α3 follow from Eq (3.26) and C3l,¯C3l,l=1,2, are defined by relation (3.35). The functions θ1 and θ2 are given implicitly by relations
eˆiθ1=1+2ΔtC12+ΔtΔxC11ϕ+ˆi(2Δt¯C12+ΔtΔx¯C11ϕ)√(1+2ΔtC12+ΔtΔxC11ϕ)2+(2Δt¯C12+ΔtΔx¯C11ϕ)2 |
and
eˆiθ2=C31+ΔtΔx¯C32ϕ+ˆi(−¯C31+ΔtΔxC32ϕ)√(C31+ΔtΔx¯C32ϕ)2+(¯C31−ΔtΔxC32ϕ)2. |
Taking the squared modulus on both sides of approximation (3.36) and performing straightforward computations to get estimates (3.19)–(3.22) given in Lemma 3.2.
Using the above results (namely, Lemmas 3.1 and 3.2) we are ready to give a necessary stability restriction of the two-step predictor-corrector method (2.3)–(2.6) and to compare it with what is available in the literature (for example, Courant-Friedrich-Lewy condition for linear hyperbolic PDEs).
Theorem 3.1. The two-step predictor-corrector scheme (2.3)–(2.6) applied to nonlinear shallow water equations with source terms (1.4) is linearly stable if
Δt4Δx2(W1(Δt,Δx)+13ΔxW2(Δt,Δx))(1+2Δt3Γ0μˆA−43)Γ0μ3ˆA−43|ϕ|2≤12(1−ϵ2)(1+√1−β∗), | (3.38) |
where |ϕ|=|kΔx|<<π, μ=max0≤i≤K1μi, with μi=|Qi||Ai|, ˆA=min0≤i≤K1|Ai|=min0≤i≤K1|ea1ti|≠0, for any a1∈R, Γ0=gn211.492P43, 0<ϵ,β∗<1,W1(Δt,Δx) and W2(Δt,Δx) are given by relations (3.20) and (3.21), respectively.
Proof. The proof of Theorem 3.1 is obvious according to Lemma 3.1 and 3.2.
The Von Neumann stability approach, based on a Fourier analysis in the space domain has been developed for nonlinear one-dimensional shallow water equations with source terms. Although the stability condition has not to be derived analytically, we have analyzed the properties of the amplification factor numerically (by use of Taylor series expansion), which contain information on the dispersion and diffusion errors of the considered numerical scheme. It is worth noticing that we used a local, linearized stability analysis to obtain an estimate (3.38), which must be considered as a necessary condition of stability for the numerical scheme (2.3)–(2.6).
Some important remarks on stability analysis
This section considers some useful remarks based on the stability restrictions and compares them with what is known in the literature: CFL condition.
(i1) The linear stability condition (3.38) suggests that a small space step Δx cannot force the time step Δt to be more potentially small (indeed: |ϕ|2=k2Δx2). This improves the convergence speed of the proposed numerical scheme. Moreover, because consistency requires that ΔtiΔx(i≥1) approaches zero as Δt and Δx tend zero, an acceptable time step (not too small) should be applied to guarantee the stability condition (3.38). For this reason, the developed three-level MacCormack method is suitable for the calculation of steady solutions (where time accuracy is unimportant) and the unsteady ones.
(i2) The developed approach (2.3)–(2.6) for one-dimensional surface water equations has a linear stability limitation (3.38) that limits the maximum time step. This stability requirement does not coincide with the CFL condition obtained for linear hyperbolic PDEs (for example: linear advection equation, wave equation, linearized burgers equations, etc...) because the considered algorithm is applied to complex time-dependent PDEs. As a discussion on the stability restrictions one can refer to the stability analysis of the two-step Lax-Wendroff method and the MacCormack scheme applied to complete burgers equations (for example, see [24], P. 245–247). The linear stability condition (3.38) is highly unusual. Since we normally find this condition from a Fourier stability analysis, it follows from inequality (3.38) that instability occurs when |Δt| is greater than |Δt|max=(Δt)CFL. However, it comes from the linear stability restriction (3.38) that the empirical formula
Δt4(1+2Δt3Γ0μˆA−43)gT−1Γ0(μ)2ˆA−13k3≤3(1−ϵ2)(1+√1−β∗), |
can be used with an appropriate safety factor. It should be remembered that the "heuristic" stability analysis, i.e., estimates (3.38) can only provide a necessary condition for stability. Thus, for some finite difference algorithms, only partial information about the complete stability bound is obtained and for others (such as algorithms for the heat equation, wave equation and linearized Burgers equations) a more complete theory must be employed.
This section simulates the two-step explicit predictor-corrector scheme (2.3)–(2.6) for solving the one-dimensional shallow water Eq (1.4). We consider two examples described in [46] to demonstrate the efficiency and robustness of the proposed technique. A practical application of a shallow water flow deals with the Benoué river. The river is located in Cameroon with a 7000m long reach of the upstream part (altitude = 174.22 m). Furthermore, the characteristics of the flow consider a rather steep part in the first kilometers together with strong irregularities in the cross section and a low base discharge (708m3/s), altogether, produce a high velocity basic flow, transcritical in some parts. Specifically, the floods problem observed in this river in 2012 is discussed since it is a classical example of unsteady nonlinear flow with shocks to expect floods and to test conservation in numerical schemes. In addition, the considered problem is assumed to be generated by a one-dimensional shallow water equation for the ideal case of a flat and frictionless channel with a prismatic cross-section wetted perimeter (P=366,4 m) and top width (T=348 m). The initial data are provided by Eq (4.1).
Let w∈L2(0,T1;L2(0,L)), k=Δt and τ=Δx, we introduce the following discrete norm
‖|w|‖L2(0,T1;L2)=[kτK1∑i=0M1∑r=0|w(ti,xr)|2]1/2. |
Denote Wir=(Air,Qir), be the computed solution provided by the two-step approach (2.3)–(2.6) and w(ti,xr)=(A(ti,xr),Q(ti,xr)), be the exact one at the discrete point (ti,xr), the error at time ti and position xr is given by e(ti,xr)=w(ti,xr)−Wir=(A(ti,xr)−Air,Q(ti,xr)−Qir). Thus,
‖|eA|‖L2(0,T1;L2)=[kτK1∑i=0M1∑r=0|A(ti,xr)−Air|2]1/2, ‖|eQ|‖L2(0,T1;L2)=[kτK1∑i=0M1∑r=0|Q(ti,xr)−Qir|2]1/2. |
∙ Problem 1 (Dam break on a dry domain with friction). The analytical solution is computed by Dressler's dam break with friction [46]. In the literature different approaches are deeply analyzed. Dressler's analyzed Chézy friction law using a perturbation method in the Ritter's scheme, i.e., both velocity: u(m/s) and height: h(m), of the water are expanded as power series in the friction coefficient Cf=1/C2. The initial conditions are defined as
h(0,x)=h0(x)={hl>0,for 0≤x≤x0; 0,for x0<x≤L, u(0,x)=u0(x)={10−1,for 0≤x≤x0; 0,for x0<x≤L. | (4.1) |
We assume that C=40m1/2/s (Chézy coefficient), hl=5×10−3m, x0=L/2,T1=1s, and L=1m. Dressler's first order developments for the flow resistance give the following corrected height and velocity
{hc(t,x)=1g(23√ghl−x−x03t+g2C2α1t)2, uc(t,x)=23√ghl+2(x−x0)3t+g2C2α2t, | (4.2) |
where
α1=65(2−x−x0t√ghl)−23+4√3135(2−x−x0t√ghl)3/2, | (4.3) |
and
α2=122−x−x0t√ghl−83+8√3189(2−x−x0t√ghl)3/2−1087(2−x−x0t√ghl)2. | (4.4) |
Following Dressler's technique, four regions are considered: from upstream to downstream (a steady state region (hl,10−1) for x≤x1(t)); a corrected region ((hc,uc) for x1(t)≤x≤x2(t)); the tip region (for x2(t)≤x≤x3(t)) and the dry region ((0,0) for x3(t)≤x≤L). In the tip region, the friction term is preponderant thus (4.2) is no more valid. The velocity increases in the corrected region with x, thus Dressler assumes that the velocity reaches the maximum of uc at x2(t) and it is constant in space in the tip region.
utip(t)=maxx∈[x2(t),x3(t)]uc(t,x). | (4.5) |
Armed with these assumptions together with Eqs (4.2)–(4.5), the exact solution of the considered shallow water problem is defined as
h(t,x)={hl,for 0≤x≤x1(t) and t∈(0,T1], 1g(23√ghl−x−x03t+g2C2α1t)2,for x1(t)≤x≤x3(t) and t∈(0,T1], 0,for x3(t)≤x≤L and t∈(0,T1], | (4.6) |
and
u(t,x)={0,for 0≤x≤x1(t) and t∈(0,T1], 23√ghl+2(x−x0)3t+g2C2α2t,for x1(t)≤x≤x2(t) and t∈(0,T1], maxx∈[x2(t),x3(t)]uc(t,x),for x2(t)≤x≤x3(t) and t∈(0,T1], 0,for x3(t)≤x≤L and t∈(0,T1], | (4.7) |
where α1 and α2 are given by Eqs (4.3) and (4.4), respectively, x1(t)=x0−t√ghl,x3(t)=x0+2t√ghl and x2(t)∈[x1(t),x3(t)] is the point where the velocity uc(t,x) attains its maximum.
∙ Problem 2 (Dam break on a dry domain without friction). We consider the exact solution introduced by Dressler's dam break with friction [46]. We also consider Ritter's solution corresponding to an ideal dam break (case of a reservoir with a constant height hl) on a dry region. Similar to the Stoker's solution, the dam break is instantaneous, the bottom is flat and there is no friction. Using this fact, the initial condition (Riemann problem) is given by equation (4.1). In addition, we assume that C=40m1/2/s (Chézy coefficient), hl=5×10−3m, x0=L/2,T1=4s, and L=8m. At time t>0, the free surface is the constant water height (hl) at rest connected to a dry zone (hr) by a parabola. This parabola is limited upstream (resp. downstream) by the abscissa xA(t) (resp. xB(t)). The analytical solution is defined as
h(t,x)={hl,for 0≤x≤xA(t) and t∈(0,T1], 49g(√9ghl−x−x02t)2,if x1(t)≤x≤xB(t) and t∈(0,T1], 0,when xB(t)≤x≤L and t∈(0,T1], | (4.8) |
and
u(t,x)={0,if 0≤x≤xA(t) and t∈(0,T1], 23(x−x0t+√ghl),for xA(t)≤x≤xB(t) and t∈(0,T1], 0,when xB(t)≤x≤L and t∈(0,T1], | (4.9) |
where xA(t)=x0−t√ghl and xB(t)=x0+2t√ghl.
The analysis considers the case where the channel is prismatic with a constant top width (T) and average velocity (u) defined as: u(t,x)=Q(t,x)/A(t,x). Thus, the following formulas are satisfied
A(t,x)=Th(t,x) and Q(t,x)=Th(t,x)u(t,x). | (4.10) |
Since the water height is constant in the tip region, it follows from (4.10) that the cross section (A) and discharge (Q) are not modified in that region. A combination of Eqs (4.6)–(4.10) results in
∙ Problem 1 (Dam break on a dry domain with friction).
A(t,x)=Th(t,x)={Thl,for 0≤x≤x1(t) and t∈(0,T1], Tg(23√ghl−x−x03t+g2C2α1t)2,when x1(t)≤x≤x3(t) and t∈(0,T1], 0,if x3(t)≤x≤L and t∈(0,T1], |
Q(t,x)=Th(t,x)u(t,x)={0,for 0≤x≤x1(t) and t∈(0,T1], Thc(t,x)uc(t,x),when x1(t)≤x≤x2(t) and t∈(0,T1], Thc(t,x)utip(t,x),for x2(t)≤x≤x3(t) and t∈(0,T1], 0,if x3(t)≤x≤L and t∈(0,T1]. |
The table suggests that the proposed approach is second-order accurate in time and fourth-order convergent in space.
∙ Problem 2 (Dam break on a dry domain without friction).
A(t,x)=Th(t,x)={Thl,for 0≤x≤xA(t) and t∈(0,T1], 4T9g(√9ghl−x−x02t)2,if x1(t)≤x≤xB(t) and t∈(0,T1], 0,when xB(t)≤x≤L and t∈(0,T1], |
and
Q(t,x)=Th(t,x)u(t,x)={0,if 0≤x≤xA(t) and t∈(0,T1], 8T27g(x−x0t+√ghl)(√9ghl−x−x02t)2,for xA(t)≤x≤xB(t) and t∈(0,T1], 0,when xB(t)≤x≤L and t∈(0,T1]. |
The following values are used in the simulations: shear stress ¯τ=1.329N/m2; Top width T=348m, wetter perimeter P=366,4m; wavelength Kλ=2π≃6.28m, manning's number n1=0.025s/m1/3, the acceleration of gravity g=10m/s2, the rainfall intensity is described as
I(t,x)={1.18×10−5m/s,if (t,x)∈[0,T1]×[0,L],0,otherwise. | (4.11) |
We observe from this table that the proposed method is temporal second-order convergent and spatial fourth-order accurate.
The mathematical model for this ideal overland flow is the following: we consider a uniform plane catchment whose overall length in the direction of flow is L(m)∈{1,8}. The surface roughness and shear stress are assumed invariant in space and time. It comes from Eq (4.11) that the constant rainfall excess is defined as
v(t,x)={I(t,x),for (t,x)∈[t0,T1]×[0,L];0,otherwise. | (4.12) |
The space step Δx∈{2−l,l=2,⋯,7}, while the time step Δt varies in the range: 2−l,l=5,⋯,12.I is the rainfall intensity defined by relation (4.11),t0=0s and T1(s)∈{1,4}, are initial and final times, respectively, and L is the rod interval length. The numerical solutions given by Eqs (2.3) and (2.6) are displayed in Figures 1 and 2. Before 3 iterations are encountered, the discharge wave propagates with almost a perfectly constant value at different positions (Figures 1 and 2). Furthermore, after these iterations, the discharge wave approaches zero at different times (Figures 1 and 2). So, the graphs suggest that the computed solution cannot grow with time and should satisfy the necessary condition (3.38). Similar remarks are observed for the cross section. In addition, setting k=τ2:=h21, Tables 1 and 2 indicate that the developed numerical technique is second-order accurate in time and spatial fourth-order. Furthermore, the graphs show that the numerical solutions start to destroy after a fixed time. Thus, physical insight must be used when the stability restriction (3.38) of the constructed two-step predictor-corrector method (2.3)–(2.6) is investigated. Finally, both Tables 1 and 2 and Figures 1 and 2 indicate that the approximate solutions do not increase with time and converge to the analytical one. Specifically, they show that stability for the proposed numerical scheme is subtle. It is not unconditionally unstable, but stability depends on the parameters Δx and Δt.
k | ‖|A−A1|‖L2 | ‖|Q−Q1|‖L2 | r(A) | r(Q) |
2−6 | 4.5762×10−1 | 7.4876×10−5 | — | — |
2−8 | 1.1196×10−1 | 2.0111×10−5 | 1.9356 | 1.8965 |
2−10 | 2.8694×10−2 | 5.6473×10−6 | 2.0645 | 1.9867 |
2−12 | 6.7835×10−3 | 1.3628×10−6 | 2.1028 | 2.0684 |
τ | ‖|A−A1|‖L2 | ‖|Q−Q1|‖L2 | r(A) | r(Q) |
2−3 | 2.3578×10−2 | 1.8869×10−6 | — | — |
2−4 | 1.5372×10−3 | 1.4237×10−7 | 4.0000 | 4.1028 |
2−5 | 1.8358×10−4 | 8.0649×10−9 | 3.9873 | 4.1573 |
2−6 | 1.0623×10−5 | 4.3871×10−10 | 4.1187 | 4.2016 |
In this paper, we have developed a two-step explicit predictor-corrector approach (2.3)–(2.6) for solving the evolutionary nonlinear problem (1.4). A necessary condition of stability of the proposed numerical scheme has been deeply analyzed using the Von Neumann stability approach whereas the convergence rate of the algorithm is numerically computed using the L2-norm. The graphs (Figures 1 and 2) show that the considered method is both stable and convergent while Tables 1 and 2 suggest that the algorithm is temporal second-order convergent and fourth-order accurate in space. After a few number of iterations, the figures indicate that the numerical solutions strongly converge with the analytical ones. One should observe from Figures 1 and 2 that the only case where the exact solutions do not tend to zero corresponds to the initial condition. This follows from assumptions made by Dressler when constructing the analytical solutions. Our future investigations will develop a two-step explicit predictor-corrector scheme for the two-dimensional shallow water model.
The authors appreciate the valuable comments and remarks of anonymous referees which helped to greatly improve the quality of the paper.
The authors declare that they have no conflicts of interest.
[1] | D. D. Franz, C. S. Melching, Full equations (FEQ) model for the solution of the full, dynamic equations of motion for one-dimensional unsteady flow in open channels and through control structures, Michigan: U.S. Department of the Interior, U.S. Geological Survey, 1997. |
[2] | D. R. Basco, Computation of rapidly varied unsteady free surface flow, Michigan: U.S. Department of the Interior, U.S. Geological Survey, 1987. |
[3] | P. Brufau, J. Burguete, P. García-Navarro, J. Murillo, The shallow water equations: An example of hyperbolic system, Monografías de la Real Academia de Ciencias de Zaragoza, 31 (2008), 89–119. |
[4] | G. Cannata, L. Lasaponara, F. Gallerano, Nonlinear shallow water equations numerical integration on curvilinear boundary-conforning grids, WSEAS Trans. Fluids Mech., 10 (2015), 13–25. |
[5] |
G. Li, V. Caleffi, J. Gao, High-order well-balanced central WENO scheme for pre-balanced shallow water equations, Comput. Fluids, 99 (2014), 182–189. https://doi.org/10.1016/j.compfluid.2014.04.022 doi: 10.1016/j.compfluid.2014.04.022
![]() |
[6] |
F. Gallerano, G. Cannata, L. Lasaponara, A new numerical model for simulations of wave transformation, breaking and long-shore, currents in complex coastal regions, Int. J. Numer. Meth. Fluids, 80 (2016), 571–613. https://doi.org/10.1002/fld.4164 doi: 10.1002/fld.4164
![]() |
[7] |
Q. Zhou, J. Zhan, Y. Li, High-order finite-volume WENO schemes for Boussinesq modelling of nearshore wave processes, J. Hydraul. Res., 54 (2016), 646–662. https://doi.org/10.1080/00221686.2016.1175520 doi: 10.1080/00221686.2016.1175520
![]() |
[8] | A. J. C. Barré de Saint Venant, Théorie du mouvement non-permanent des eaux, avec application aux crues des rivières et à l'introduction des marées dans leur lit, Compte. Rendu de l'Académie des Sciences, 73 (1871), 147–154. |
[9] | A. Kurganov, G. Petrova, A second-order well-balanced positivity preserving central-upwind schemes for the Saint-Venant system, Commun. Math. Sci., 5 (2007), 133–160. |
[10] | R. W. Maccormack, The effect of viscosity in hypervelocity impact cratering, In: Frontiers of computational fluid dynamics, 2002, 27–43. |
[11] | P. D. Lax, B. Wendrof, Systems of conservation laws, J. Commun. Pure. Appl. Math., 13 (1959), 217–237. |
[12] |
F. T. Namio, E. Ngondiep, R. Ntchantcho, J. C. Ntonga, Mathematical models of complete shallow water equations with source terms, stability analysis of Lax-Wendroff scheme, J. Theor. Comput. Sci., 2 (2015), 1000132. 10.4172/2376-130X.1000132 doi: 10.4172/2376-130X.1000132
![]() |
[13] |
E. Ngondiep, Stability analysis of MacCormack rapid solver method for evolutionary Stokes-Darcy problem, J. Comput. Appl. Math., 345 (2019), 269–285. https://doi.org/10.1016/j.cam.2018.06.034 doi: 10.1016/j.cam.2018.06.034
![]() |
[14] | E. Ngondiep, A robust three-level time-split MacCormack scheme for solving two-dimensional unsteady convection-diffusion equation, J. Appl. Comput. Mech., 7 (2021), 559–577. |
[15] |
E. Ngondiep, An efficient three-level explicit time-split scheme for solving two-dimensional unsteady nonlinear coupled Burgers equations, Int. J. Numer. Methods Fluids, 92 (2020), 266–284. https://doi.org/10.1002/fld.4783 doi: 10.1002/fld.4783
![]() |
[16] |
E. Ngondiep, An efficient three-level explicit time-split approach for solving 2D heat conduction equations, Appl. Math. Inf. Sci., 14 (2020), 1075–1092. http://dx.doi.org/10.18576/amis/140615 doi: 10.18576/amis/140615
![]() |
[17] |
R. Hixon, E. Turkel, Compact implicit MacCormack type schemes with high accuracy, J. Comput. Phys., 158 (2000), 51–70. https://doi.org/10.1006/jcph.1999.6406 doi: 10.1006/jcph.1999.6406
![]() |
[18] |
G. S. Jiang, D. Levy, C. T. Lin, S. Osher, E. Tadmor, High-resolution non-oscillatory central schemes with Non-staggered grids for hyperbolic conservation laws, SIAM J. Numer. Anal., 35 (1998), 2147–2168. https://doi.org/10.1137/S0036142997317560 doi: 10.1137/S0036142997317560
![]() |
[19] |
D. Levy, G. Puppo, G. Russo, Central WENO schemes for hyperbolic systems of conservation laws, ESAIM Math. Model. Numer. Anal., 33 (1999), 547–571. https://doi.org/10.1051/m2an:1999152 doi: 10.1051/m2an:1999152
![]() |
[20] |
G. S. Jiang, C. W. Shu, Efficient implementation of weighted ENO schemes, J. Comput. Phys., 126 (1996), 202–228. https://doi.org/10.1006/jcph.1996.0130 doi: 10.1006/jcph.1996.0130
![]() |
[21] |
F. Bianco, G. Puppo, G. Russo, High order central schemes for hyperbolic systems of conservation laws, SIAM J. Sci. Comput., 21 (1999), 294–322. https://doi.org/10.1137/S10648275973249 doi: 10.1137/S10648275973249
![]() |
[22] |
A. Kurganov, D. Levy, Central-upwind schemes for the Saint-Venant system, ESAIM Math. Model. Numer. Anal., 36 (2002), 397–429. https://doi.org/10.1051/m2an:2002019 doi: 10.1051/m2an:2002019
![]() |
[23] |
R. Sanders, A. Weiser, A high resolution staggered mesh approach for nonlinear Hyperbolic systems of conservation laws, J. Comput. Phys., 101 (1992), 314–329. https://doi.org/10.1016/0021-9991(92)90009-N doi: 10.1016/0021-9991(92)90009-N
![]() |
[24] | F. A. Anderson, R. H. Pletcher, J. C. Tannehill, Computational fluid mechanics and Heat Transfer, New York: Taylor and Francis, 1997. |
[25] |
F. R. Fiedler, J. A. Ramirez, A numerical method for simulating discontinuous shallow flow over an infiltrating surface, Int. J. Numer. Meth. Fluids, 32 (2000), 219–240. https://doi.org/10.1002/(SICI)1097-0363(20000130)32:2<219::AID-FLD936>3.0.CO;2-J doi: 10.1002/(SICI)1097-0363(20000130)32:2<219::AID-FLD936>3.0.CO;2-J
![]() |
[26] | E. Ngondiep, A novel three-level time-split MacCormack method for solving two-dimensional viscous coupled Burgers equations, preprint paper, 2019. https://doi.org/10.48550/arXiv.1906.01544 |
[27] | E. Ngondiep, A novel three-level time-split approach for solving two-dimensional nonlinear unsteady convection-diffusion-reaction equation, J. Math. Comput. Sci., 26 (2022), 222–248. |
[28] |
E. Ngondiep, A fourth-order two-level factored implicit scheme for solving two-dimensional unsteady transport equation with time-dependent dispersion coefficients, Int. J. Comput. Method. Eng. Sci. Mech., 22 (2021), 253–264. https://doi.org/10.1080/15502287.2020.1856972 doi: 10.1080/15502287.2020.1856972
![]() |
[29] |
E. Ngondiep, Long time stability and convergence rate of MacCormack rapid solver method for nonstationary Stokes-Darcy problem, Comput. Math. Appl., 75 (2018), 3663–3684. https://doi.org/10.1016/j.camwa.2018.02.024 doi: 10.1016/j.camwa.2018.02.024
![]() |
[30] |
E. Ngondiep, A novel three-level time-split MacCormack scheme for two-dimensional evolutionary linear convection-diffusion-reaction equation with source term, Int. J. Comput. Math., 98 (2021), 47–74. https://doi.org/10.1080/00207160.2020.1726896 doi: 10.1080/00207160.2020.1726896
![]() |
[31] |
E. Ngondiep, A robust numerical two-level second-order explicit approach to predict the spread of COVID-2019 pandemic with undetected infectious cases, J. Comput. Appl. Math., 403 (2022), 113852. https://doi.org/10.1016/j.cam.2021.113852 doi: 10.1016/j.cam.2021.113852
![]() |
[32] |
E. Ngondiep, N. Kerdid, M. A. M. Abaoud, I. A. I. Aldayel, A three-level time-split MacCormack method for two-dimensional nonlinear reaction-diffusion equations, Int. J. Numer. Meth. Fluids, 92 (2020), 1681–1706. https://doi.org/10.1002/fld.4844 doi: 10.1002/fld.4844
![]() |
[33] | E. Ngondiep, A six-level time-split Leap-Frog/Crank-Nicolson approach for two-dimensional nonlinear time-dependent convection diffusion reaction equation, Int. J. Comput. Meth., 2023. https://doi.org/10.1142/S0219876222500645 |
[34] |
E. Ngondiep, Long time unconditional stability of a two-level hybrid method for nonstationary incompressible Navier-Stokes equations, J. Comput. Appl. Math., 345 (2019), 501–514. https://doi.org/10.1016/j.cam.2018.05.023 doi: 10.1016/j.cam.2018.05.023
![]() |
[35] | E. Ngondiep, Error estimate of MacCormack rapid solver method for 2D incompressible Navier-Stokes problems, preprint paper, 2019. https://doi.org/10.48550/arXiv.1903.10857 |
[36] | E. Ngondiep, Asymptotic growth of the spectral radii of collocation matrices approximating elliptic boundary problems, Int. J. Appl. Math. Comput., 4 (2012), 199–219. |
[37] |
E. Ngondiep, A two-level factored Crank-Nicolson method for two-dimensional nonstationary advection-diffusion equation with time dependent dispersion coefficients and source sink/term, Adv. Appl. Math. Mech., 13 (2021), 1005–1026. https://doi.org/10.4208/aamm.OA-2020-0206 doi: 10.4208/aamm.OA-2020-0206
![]() |
[38] |
E. Ngondiep, Unconditional stability over long time intervals of a two-level coupled MacCormack/Crank-Nicolson method for evolutionary mixed Stokes-Darcy model, J. Comput. Appl. Math., 409 (2022), 114148. https://doi.org/10.1016/j.cam.2022.114148 doi: 10.1016/j.cam.2022.114148
![]() |
[39] |
E. Ngondiep, A two-level fourth-order approach for time-fractional convection-diffusion-reaction equation with variable coefficients, Commun. Nonlinear Sci. Numer. Simul., 111 (2022), 106444. https://doi.org/10.1016/j.cnsns.2022.106444 doi: 10.1016/j.cnsns.2022.106444
![]() |
[40] | E. Ngondiep, Unconditional stability of a two-step fourth-order modified explicit Euler/Crank-Nicolson approach for folving time-variable fractional mobile-immobile advection-dispersion equation, preprint paper, 2022. https://doi.org/10.48550/arXiv.2205.05077 |
[41] |
K. Ye, Y. Zhao, F. Wu, W. Zhong, An adaptive artificial viscosity for the displacement shallow water wave equation, Appl. Math. Mech., 43 (2022), 247–262. https://doi.org/10.1007/s10483-022-2815-7 doi: 10.1007/s10483-022-2815-7
![]() |
[42] | M. D. Saiduzzaman, S. K. Ray, Comparison of numerical schemes for shallow water equation, Glob. J. Sci. Front. Res., 13 (2013), 28–46. |
[43] |
F. Wu, W. Zhong, On displacement shallow water wave equation and symplectic solution, Comput. Method. Appl. Mech. Eng., 318 (2017), 431–455. https://doi.org/10.1016/j.cma.2017.01.040 doi: 10.1016/j.cma.2017.01.040
![]() |
[44] | H. O. Kreiss, On difference approximations of the dissipative type for hyperbolic differential equations, Comm. Pure Appl. Math., 17 (1964), 335–353. |
[45] | R. Garcia, R. A. Kahawaita, Numerical solution of the Saint-Venant equations with MacCormack finite-difference scheme, Int. J. Numer. Meth. Fluids, 6 (1986), 259–274. |
[46] |
O. Delestre, C. Lucas, P. A. Ksinant, F. Darboux, C. Laguerre, T. N. Tuoi Vo, et al., SWASHES: a compilation of shallow water analytic solutions for hydraulic and environmental studies, Int. J. Numer. Meth. Fluids, 72 (2013), 269–300. https://doi.org/10.1002/fld.3741 doi: 10.1002/fld.3741
![]() |
1. | Eric Ngondiep, An efficient two-level factored method for advection-dispersion problem with spatio-temporal coefficients and source terms, 2023, 8, 2473-6988, 11498, 10.3934/math.2023582 | |
2. | Kaihong Zhao, Solvability, Approximation and Stability of Periodic Boundary Value Problem for a Nonlinear Hadamard Fractional Differential Equation with p-Laplacian, 2023, 12, 2075-1680, 733, 10.3390/axioms12080733 | |
3. | Weiping Wei, Youlin Shang, Hongwei Jiao, Pujun Jia, Shock stability of a novel flux splitting scheme, 2024, 9, 2473-6988, 7511, 10.3934/math.2024364 | |
4. | Eric Ngondiep, A fast three-step second-order explicit numerical approach to investigating and forecasting the dynamic of corruption and poverty in Cameroon, 2024, 10, 24058440, e38236, 10.1016/j.heliyon.2024.e38236 | |
5. | Eric Ngondiep, A high-order combined finite element/interpolation approach for multidimensional nonlinear generalized Benjamin–Bona–Mahony–Burgers equation, 2024, 215, 03784754, 560, 10.1016/j.matcom.2023.08.041 | |
6. | Eric Ngondiep, A posteriori error estimate of MacCormack rapid solver method for two-dimensional incompressible Navier–Stokes problems, 2024, 438, 03770427, 115569, 10.1016/j.cam.2023.115569 | |
7. | Eric Ngondiep, An efficient numerical approach for solving three‐dimensional Black‐Scholes equation with stochastic volatility, 2024, 0170-4214, 10.1002/mma.10576 | |
8. | Eric Ngondiep, An efficient high‐order two‐level explicit/implicit numerical scheme for two‐dimensional time fractional mobile/immobile advection‐dispersion model, 2024, 96, 0271-2091, 1305, 10.1002/fld.5296 | |
9. | Eric Ngondiep, A High‐Order Finite Element Method for Solving Two‐Dimensional Fractional Rayleigh–Stokes Problem for a Heated Generalized Second Grade Fluid, 2024, 0271-2091, 10.1002/fld.5361 | |
10. | Eric Ngondiep, A robust time-split linearized explicit/implicit technique for solving a two-dimensional hydrodynamic model: Case of floods in the far north region of Cameroon, 2025, 37, 1070-6631, 10.1063/5.0254027 |
k | ‖|A−A1|‖L2 | ‖|Q−Q1|‖L2 | r(A) | r(Q) |
2−6 | 4.5762×10−1 | 7.4876×10−5 | — | — |
2−8 | 1.1196×10−1 | 2.0111×10−5 | 1.9356 | 1.8965 |
2−10 | 2.8694×10−2 | 5.6473×10−6 | 2.0645 | 1.9867 |
2−12 | 6.7835×10−3 | 1.3628×10−6 | 2.1028 | 2.0684 |
τ | ‖|A−A1|‖L2 | ‖|Q−Q1|‖L2 | r(A) | r(Q) |
2−3 | 2.3578×10−2 | 1.8869×10−6 | — | — |
2−4 | 1.5372×10−3 | 1.4237×10−7 | 4.0000 | 4.1028 |
2−5 | 1.8358×10−4 | 8.0649×10−9 | 3.9873 | 4.1573 |
2−6 | 1.0623×10−5 | 4.3871×10−10 | 4.1187 | 4.2016 |
k | ‖|A−A1|‖L2 | ‖|Q−Q1|‖L2 | r(A) | r(Q) |
2−6 | 4.5762×10−1 | 7.4876×10−5 | — | — |
2−8 | 1.1196×10−1 | 2.0111×10−5 | 1.9356 | 1.8965 |
2−10 | 2.8694×10−2 | 5.6473×10−6 | 2.0645 | 1.9867 |
2−12 | 6.7835×10−3 | 1.3628×10−6 | 2.1028 | 2.0684 |
τ | ‖|A−A1|‖L2 | ‖|Q−Q1|‖L2 | r(A) | r(Q) |
2−3 | 2.3578×10−2 | 1.8869×10−6 | — | — |
2−4 | 1.5372×10−3 | 1.4237×10−7 | 4.0000 | 4.1028 |
2−5 | 1.8358×10−4 | 8.0649×10−9 | 3.9873 | 4.1573 |
2−6 | 1.0623×10−5 | 4.3871×10−10 | 4.1187 | 4.2016 |