
We propose a finite difference method based on the Lax-Friedrichs scheme for a model of interaction between multiple solid particles and an inviscid fluid. The single-particle version has been studied extensively during the past decade. The model studied here consists of the inviscid Burgers equation with multiple nonconservative moving source terms that are singular and account for drag force interaction between the fluid and the particles. Each particle trajectory satisfies a differential equation that ensures conservation of momentum of the entire system. To deal with the singular source terms we discretize a model that associates with each particle an advection PDE whose solution is a shifted Heaviside function. This alternative model is well known but has not previously been used in numerical methods. We propose a definition of entropy solution which directly generalizes the previously defined single-particle notion of entropy solution. We prove convergence (along a subsequence) of the Lax-Friedrichs approximations, and also prove that if the set of times where the particle paths intersect has Lebesgue measure zero, then the limit is an entropy solution. We also propose a higher resolution version of the scheme, based on MUSCL processing, and present the results of numerical experiments.
Citation: John D. Towers. The Lax-Friedrichs scheme for interaction between the inviscid Burgers equation and multiple particles[J]. Networks and Heterogeneous Media, 2020, 15(1): 143-169. doi: 10.3934/nhm.2020007
[1] | John D. Towers . The Lax-Friedrichs scheme for interaction between the inviscid Burgers equation and multiple particles. Networks and Heterogeneous Media, 2020, 15(1): 143-169. doi: 10.3934/nhm.2020007 |
[2] | Boris Andreianov, Frédéric Lagoutière, Nicolas Seguin, Takéo Takahashi . Small solids in an inviscid fluid. Networks and Heterogeneous Media, 2010, 5(3): 385-404. doi: 10.3934/nhm.2010.5.385 |
[3] | Jan Friedrich, Oliver Kolb, Simone Göttlich . A Godunov type scheme for a class of LWR traffic flow models with non-local flux. Networks and Heterogeneous Media, 2018, 13(4): 531-547. doi: 10.3934/nhm.2018024 |
[4] | Caterina Balzotti, Simone Göttlich . A two-dimensional multi-class traffic flow model. Networks and Heterogeneous Media, 2021, 16(1): 69-90. doi: 10.3934/nhm.2020034 |
[5] | Raimund Bürger, Harold Deivi Contreras, Luis Miguel Villada . A Hilliges-Weidlich-type scheme for a one-dimensional scalar conservation law with nonlocal flux. Networks and Heterogeneous Media, 2023, 18(2): 664-693. doi: 10.3934/nhm.2023029 |
[6] | Tong Yan . The numerical solutions for the nonhomogeneous Burgers' equation with the generalized Hopf-Cole transformation. Networks and Heterogeneous Media, 2023, 18(1): 359-379. doi: 10.3934/nhm.2023014 |
[7] | Verónica Anaya, Mostafa Bendahmane, David Mora, Ricardo Ruiz Baier . On a vorticity-based formulation for reaction-diffusion-Brinkman systems. Networks and Heterogeneous Media, 2018, 13(1): 69-94. doi: 10.3934/nhm.2018004 |
[8] | Li-Bin Liu, Limin Ye, Xiaobing Bao, Yong Zhang . A second order numerical method for a Volterra integro-differential equation with a weakly singular kernel. Networks and Heterogeneous Media, 2024, 19(2): 740-752. doi: 10.3934/nhm.2024033 |
[9] | Anya Désilles, Hélène Frankowska . Explicit construction of solutions to the Burgers equation with discontinuous initial-boundary conditions. Networks and Heterogeneous Media, 2013, 8(3): 727-744. doi: 10.3934/nhm.2013.8.727 |
[10] | Jie Gu, Lijuan Nong, Qian Yi, An Chen . Two high-order compact difference schemes with temporal graded meshes for time-fractional Black-Scholes equation. Networks and Heterogeneous Media, 2023, 18(4): 1692-1712. doi: 10.3934/nhm.2023074 |
We propose a finite difference method based on the Lax-Friedrichs scheme for a model of interaction between multiple solid particles and an inviscid fluid. The single-particle version has been studied extensively during the past decade. The model studied here consists of the inviscid Burgers equation with multiple nonconservative moving source terms that are singular and account for drag force interaction between the fluid and the particles. Each particle trajectory satisfies a differential equation that ensures conservation of momentum of the entire system. To deal with the singular source terms we discretize a model that associates with each particle an advection PDE whose solution is a shifted Heaviside function. This alternative model is well known but has not previously been used in numerical methods. We propose a definition of entropy solution which directly generalizes the previously defined single-particle notion of entropy solution. We prove convergence (along a subsequence) of the Lax-Friedrichs approximations, and also prove that if the set of times where the particle paths intersect has Lebesgue measure zero, then the limit is an entropy solution. We also propose a higher resolution version of the scheme, based on MUSCL processing, and present the results of numerical experiments.
This paper concerns a one-dimensional model of solid-fluid interaction:
{∂tu+∂xf(u)=K∑k=1λk(h′k(t)−u)δ(x−hk(t)),(x,t)∈R×(0,T):=ΠT,mkhk″(t)=λk(u(hk(t),t)−h′k(t)),t∈(0,T),k=1,…,K,u(x,0)=u0(x),(hk(0),h′k(0))=(hk,0,vk,0),k=1,…,K. | (1.1) |
Here
The fluid velocity is governed by the inviscid Burgers equation
There are some difficulties associated with (1.1), in addition to the well-known ones associated with a nonlinear conservation law. The source terms on the right side of the first equation are nonconservative products of distributions; their meaning is not immediately clear. The differential equations appearing in the second line are coupled to the conservation law. Due to discontinuities in
Notwithstanding these difficulties there has been much progress on the single-particle version of (1.1). A notion of solution has been developed, well-posedness has been proven, and numerical algorithms have been designed whose approximations are known to converge to the unique solution. In this paper we focus on the multiple-particle problem, which has not been studied as thoroughly. We propose a notion of entropy solution suitable for multiple particles, present a Lax-Friedrichs difference scheme for the multiple-particle problem, and prove that the resulting approximations converge to an entropy solution. This is accomplished under the assumption that the particle paths do not intersect except possibly at a set of times whose Lebesgue measure is zero.
Reference [4] developed a unifying framework for the jump conditions that hold across a spatial flux discontinuity for a conservation law with discontinuous flux, using the theory of
Definition 1.1 (the germ
G(λ,c)=(c,c)+{(a,b)∈R2|b=a−λ}∪{(a,b)∈R2|a≥0,b≤0,−λ≤a+b≤λ}. | (1.2) |
Reference [6] gives a definition of entropy solution for the single-particle version of (1.1). The following is a direct generalization of that definition to the multiple-particle problem.
Definition 1.2 (entropy solution).
(ⅰ) Given
(u(hk(t)−,t),u(hk(t)+,t))∈G(λk(t),h′k(t)),k=1,…,K. | (1.3) |
(ⅱ) A function
mkh″k(t)=(12u(hk(t)−,t)2−h′k(t)u(hk(t)−,t))−(12u(hk(t)+,t)2−h′k(t)u(hk(t)+,t)). | (1.4) |
(ⅲ) With the notation
Remark 1. Definition 1.2 requires strong one-sided traces
Assumption 1.1. The initial data satisfies
Above we have used the notation
TV(ρ):=sup{M∑i=1|ρ(ξi)−ρ(ξi−1)|}<∞, |
where the
Theorem 1.3 (Main theorem). The Lax-Friedrichs scheme described in Section 2 produces approximations that converge as the mesh size approaches zero, along a subsequence, to a pair
As mentioned above, there has been significant progress on the single-particle version of (1.1) [1,5,6,7,11]. The study of (1.1) started with reference [11]. Among other things the authors completely solved the Riemann problem for
In reference [5], the authors introduce two finite volume methods for computing approximate solutions. One is a Glimm-like scheme, and the other is a well-balanced scheme that uses nonrectangular space-time cells near the interface. These methods employ random sampling for placing the particle at a mesh interface at each time step. The nonconservative source term is handled by using a certain well-balanced scheme that was analyzed in [7]. They avoid the use of a moving mesh, and also avoid the use of a Riemann solver for the full model. The case of multiple particles is addressed, and is handled via a splitting method.
Reference [14] presents a finite volume scheme that is based on the well-balanced scheme of [5,7], but uses an adaptive stencil as an alternative to using a moving grid. The multiple-particle case is handled by splitting.
Reference [7] proves well-posedness for the problem
ut+(u2/2)x=−λuδ(x),u(x,0)=u0(x). | (1.5) |
This is a simplification of (1.1), but its analysis provides an important step in analyzing the full problem. As mentioned above the germ
Reference [6] proves well-posedness of the model (1.1) for
Reference [1] presents a class of finite volume schemes for (1.1) when
References [2] and [3] concern a generalized version of (1.1) (again, for
Reference [10] specifically deals with a multiple-particle problem. The authors prove well-posedness for a version of (1.1) where the particle paths
Let
{∂tu+∂x(u2/2)=K∑k=1λk(h′k(t)−u)∂xwk,(x,t)∈ΠT,∂twk+h′k(t)∂xwk=0,(x,t)∈ΠT,k=1,…,K,mkhk″(t)=λk(u(hk(t),t)−h′k(t)),t∈(0,T),k=1,…,K,u(x,0)=u0(x),(hk(0),h′k(0))=(hk,0,vk,0),k=1,…,K,wk(x,0)=H(x−hk,0),k=1,…,K. | (1.6) |
Although the splitting approach for multiple particles used in [5] and [14] gives good numerical results, extending the convergence analysis from the single-particle to the multiple-particle problem seems difficult. Various bounds required for convergence are not preserved by the splitting steps. The numerical schemes in those papers are based on the model (1.1). In this paper we instead discretize (1.6), using Lax-Friedrichs differencing for each of the PDEs. The advantage of this approach is that the case of multiple particles is accommodated without splitting. This makes it possible to obtain a number of estimates which taken together give a convergence proof for the multiple-particle model. On the other hand, while the schemes of [1], [5], and [14] give very sharply resolved shocks at the particle locations, our Lax-Friedrichs method results in a substantial amount of smearing. With this in mind, we additionally propose a higher resolution version of the scheme, based on MUSCL processing.
The rest of the paper is organized as follows. In Section 2 we describe the Lax-Friedrichs scheme mentioned above. In Section 3 we prove convergence, modulo a subsequence, of the approximations for
We use a uniform spatial mesh size
xj=jΔx,j∈Z,tn=nΔt,0≤n≤N, | (2.1) |
where the integer
Δ+Qnj=Qnj+1−Qnj,Δ−Qnj=Qnj−Qnj−1,ˆQnj=12(Qnj−1+Qnj+1),Qnmin=infj∈ZQnj,Qnmax=supj∈ZQnj,‖Qn‖∞=supj∈Z|Qnj|. | (2.2) |
Let
infx∈Ijv0(x)≤V0j≤supx∈Ijv0(x),and∑j∈Zχj(x)V0j→v0(x)inL1loc(R)asΔx→0. | (2.3) |
With the notation
We extend
uΔ(x,t)=N∑n=0∑j∈Zχj(x)χn(t)Unj,wΔk(x,t)=N∑n=0∑j∈Zχj(x)χn(t)Wnk,j. | (2.4) |
Similarly,
cΔk(t)=N∑n=0χn(t)cnk,hΔk(t)=N∑n=0χn(t)(hnk+(t−tn)cnk), | (2.5) |
where
Let
{Un+1j=Unj−μΔ−ˉfnj+1/2+K∑k=1λkμ2(cnk−ˆUnj)(Wnk,j+1−Wnk,j−1),Wn+1k,j=Wnk,j−μΔ−ˉgnk,j+1/2,cn+1k=cnk−1mk∑j∈ZΔtλk2(cnk−ˆUnj)(Wnk,j+1−Wnk,j−1),hn+1k=hnk+cnkΔt. | (2.6) |
Here
ˉfnj+1/2=ˉf(Unj+1,Unj)=12((Unj+1)2/2+(Unj)2/2)−q2μ(Unj+1−Unj),ˉgnk,j+1/2=12(cnkWnk,j+1+cnkWnk,j)−q2μ(Wnk,j+1−Wnk,j), | (2.7) |
where
Remark 2. The scheme (2.6) preserves solutions where the fluid velocity and particle velocities are equal to the same constant:
Remark 3. Some explanation of the third equation of (2.6) is in order. Based on the third equation of (1.6), the third equation of (2.6) should be (approximately) equivalent to
cn+1k=cnk−1mkΔtλkcnk+1mkΔtλk˜u(hk(tn),tn), |
where
cn+1k=cnk−1mkΔtλkcnk+1mkΔtλk(1/2)∑j∈ZˆUnj(Wnk,j+1−Wnk,j−1). |
Thus, by defining
˜u(hk(tn),tn):=(1/2)∑j∈ZˆUnj(Wnk,j+1−Wnk,j−1)≈∫Ru(x,tn)δ(x−hk(tn))dx, |
we have the desired approximation
From the first two equations of (1.1) it follows that, at least formally, the total momentum of the system is conserved:
ddt(∫Ru(x,t)dx+K∑k=1mkh′k(t))=0. | (2.8) |
The scheme (2.6) enforces a discrete version of (2.8).
Proposition 1. Assume that there is a
Mn=Δx∑j∈ZUnj+K∑k=1mkcnk. | (2.9) |
The discrete momentum is conserved:
Proof. Multiplying by
Δx∑j∈ZUn+1j=Δx∑j∈ZUnj+∑j∈ZK∑k=1λkΔt2(cnk−ˆUnj)(Wnk,j+1−Wnk,j−1). | (2.10) |
Multiplying the third equation of (2.6) by
K∑k=1mkcn+1k=K∑k=1mkcnk−K∑k=1∑j∈ZΔtλk2(cnk−ˆUnj)(Wnk,j+1−Wnk,j−1). | (2.11) |
The proof is completed by adding (2.10) and (2.11).
Define
Znj=Unj+K∑k=1λkWnk,j,zΔ(x,t)=N∑n=0∑j∈Zχj(x)χn(t)Znj. | (2.12) |
Lemma 2.1.
Zn+1j=Znj−μΔ−ˉf(Znj+1,Znj)+μ2K∑k=1λkˆWnk,j(Znj+1−Znj−1), | (2.13) |
Zn+1j=Znj+12(q−μˆUnj)Δ+Znj−12(q+μˆUnj)Δ−Znj. | (2.14) |
Remark 4. From (1.6) and the definition
∂tz+∂xf(z)=K∑k=1λkwk∂xz. | (2.15) |
Evidently (2.13) is a discretization of (2.15).
Remark 5. It is clear by inspection of either (2.13) or (2.14) that the scheme (2.6) preserves solutions of the form
Proof. Using (2.12) and (2.6) we find that
Zn+1j=Unj−μΔ−ˉfnj+1/2+K∑k=1λkμ2(cnk−ˆUnj)(Wnk,j+1−Wnk,j−1)+K∑k=1λk(Wnk,j−μΔ−ˉgnk,j+1/2)=Znj−μΔ−ˉfnj+1/2+K∑k=1λkμ2(cnk−ˆUnj)(Wnk,j+1−Wnk,j−1)−μK∑k=1λkΔ−ˉgnk,j+1/2. | (2.16) |
Next we use
Δ−ˉfnj+1/2=12ˆUnj(Unj+1−Unj−1)−q2μΔ+Δ−Unj,Δ−ˉgnk,j+1/2=12cnk(Wnk,j+1−Wnk,j−1)−q2μΔ+Δ−Wnk,j. | (2.17) |
Substituting (2.17) into (2.16) and canceling
Zn+1j=Znj−μ2ˆUnj(Unj+1−Unj−1)+q2Δ+Δ−Unj−μ2ˆUnjK∑k=1λk(Wnk,j+1−Wnk,j−1)+q2K∑k=1λkΔ+Δ−Wnk,j=Znj−μ2ˆUnj(Znj+1−Znj−1)+q2Δ+Δ−Znj=Znj−μ2ˆUnj(Δ+Znj+Δ−Znj)+q2(Δ+Znj−Δ−Znj). | (2.18) |
The identity (2.14) is immediate from (2.18).
For the proof of (2.13), we start from the second equality of (2.18) and substitute
Zn+1j=Znj−μ2(ˆZnj−K∑k=1λkˆWnk,j)(Znj+1−Znj−1)+q2Δ+Δ−Znj=Znj−μ2ˆZnj(Znj+1−Znj−1)+μ2(K∑k=1λkˆWnk,j)(Znj+1−Znj−1)+q2Δ+Δ−Znj=Znj−μ2(f(Znj+1)−f(Znj−1))+q2Δ+Δ−Znj+μ2(K∑k=1λkˆWnk,j)(Znj+1−Znj−1). | (2.19) |
The identity (2.13) now follows directly from (2.19).
Let
μmax(max1≤k≤K|c0k|,‖z0‖∞+K∑k=1λk,‖u0‖∞+K∑k=1λk)≤q≤1/2. | (3.1) |
Additionally we assume that
Δt≤mk/λk,k=1,…,K, | (3.2) |
which will be satisfied automatically for
Define
Lemma 3.1. The following properties hold:
z0min≤Znj≤z0max,‖Zn‖∞≤‖z0‖∞, | (3.3) |
u0min−K∑k=1λk≤Unj≤u0max+K∑k=1λk,‖Un‖∞≤‖u0‖∞+K∑k=1λk, | (3.4) |
Wnk,j∈[0,1],Δ+Wnk,j≥0,∑j∈ZΔ+Wnk,j=1, | (3.5) |
|cnk|≤max(|c0k|,‖u0‖∞+K∑k=1λk). | (3.6) |
Proof. The proof is by induction on
μ(‖Zn‖∞+K∑k=1λk)≤q,μ|cnk|≤q,k=1,…,K. | (3.7) |
To prove that (3.3) holds at time step
Zn+1j=Znj+Cnj+1/2Δ+Znj−Dnj−1/2Δ−Znj, | (3.8) |
where
Cnj+1/2=12(q−μˆUnj),Dnj−1/2=12(q+μˆUnj). | (3.9) |
Using
Zn+1j=(1−Cnj+1/2−Dnj−1/2)Znj+Cnj+1/2Znj+1+Dnj−1/2Znj−1. | (3.10) |
From (3.10) it is clear that
Next we prove that (3.5) holds for
Wn+1k,j=(1−αnk−βnk)Wnk,j+αnkWnk,j+1+βnkWnk,j−1, | (3.11) |
where
αnk=12(q−μcnk),βnk=12(q+μcnk). | (3.12) |
By (3.7) we have
Δ+Wn+1k,j=(1−αnk−βnk)Δ+Wnk,j+αnkΔ+Wnk,j+1+βnkΔ+Wnk,j−1. | (3.13) |
Invoking the induction hypothesis again yields
To prove (3.4) holds at
z0min−K∑k=1λkWn+1k,j≤Un+1j≤z0max−K∑k=1λkWn+1k,j. | (3.14) |
It is readily verified that
u0min−K∑k=1λkWn+1k,j≤Un+1j≤u0max+K∑k=1λk−K∑k=1λkWn+1k,j. | (3.15) |
Recalling that
To verify that (3.6) holds for
cn+1k=(1−Δtλk2mk∑j∈Z(Wnk,j+1−Wnk,j−1))cnk+Δtλk2mk∑j∈Z(Wnk,j+1−Wnk,j−1)ˆUnj. | (3.16) |
The induction hypothesis yields
|cn+1k|≤(1−Δtλkmk)|cnk|+Δtλk2mk∑j∈Z(Wnk,j+1−Wnk,j−1)|ˆUnj|≤(1−Δtλkmk)|cnk|+Δtλk2mk∑j∈Z(Wnk,j+1−Wnk,j−1)(‖u0‖∞+K∑k=1λk)=(1−Δtλkmk)|cnk|+Δtλkmk(‖u0‖∞+K∑k=1λk)≤(1−Δtλkmk)max(|c0k|,‖u0‖∞+K∑k=1λk)+Δtλkmk(‖u0‖∞+K∑k=1λk), | (3.17) |
from which the desired inequality follows readily.
Lemma 3.2.
∑j∈Z|Δ+Znj|≤TV(u0)+K∑k=1λk, | (3.18) |
and
∑j∈Z|Δ+Unj|≤TV(u0)+2K∑k=1λk. | (3.19) |
Proof. We claim that the scheme is a so-called Total Variation Decreasing (TVD) scheme with respect to the variable
∑j∈Z|Δ+Zn+1j|≤∑j∈Z|Δ+Znj|. | (3.20) |
To prove the claim we use (3.8). We have shown that
Cnj+1/2+Dnj+1/2=q−μ4(Unj+1+Unj−1)+μ4(Unj+2+Unj)≤q+μ‖Un‖∞≤q+μ(‖u0‖∞+K∑k=1λk)≤2q. | (3.21) |
Here we have used (3.1) to get the last inequality. The desired bound then results by recalling that
∑j∈Z|Δ+Znj|≤∑j∈Z|Δ+Z0j|≤TV(z0). | (3.22) |
It is readily verified using (2.12) that
∑j∈Z|Unj+1−Unj|−K∑k=1λk≤∑j∈Z|Znj+1−Znj|≤∑j∈Z|Unj+1−Unj|+K∑k=1λk. | (3.23) |
Then (3.18) follows from (3.22) and the
Lemma 3.3. The following time continuity estimate holds:
∑j∈Z|Un+1j−Unj|≤B, | (3.24) |
where the constant
Proof. Rearranging the first equation of (2.6), and using (2.17) to rewrite
Un+1j−Unj=12(q−μˆUnj)Δ+Unj−12(q+μˆUnj)Δ−Unj+μ2K∑k=1λk(cnk−ˆUnj)(Wnk,j+1−Wnk,j−1). | (3.25) |
After taking absolute values, applying the triangle inequality, then using the bounds on
∑j∈Z|Un+1j−Unj|≤B1∑j∈Z|Δ+Unj|+B2K∑k=1∑j∈Z|Wnk,j+1−Wnk,j−1|, | (3.26) |
where
Lemma 3.4. The particle velocity approximations satisfy the following bound:
|cn+1k−cnk|≤λkΔtmk(max(|c0k|,‖u0‖∞+K∑k=1λk)+‖u0‖∞+K∑k=1λk). | (3.27) |
Proof.
We start with the third formula of (2.6). Subtracting
|cn+1k−cnk|≤1mk∑j∈ZΔtλk2|cnk−ˆUnj|(Wnk,j+1−Wnk,j−1)≤1mk∑j∈ZΔtλk2(|cnk|+‖Un‖∞)(Wnk,j+1−Wnk,j−1)=Δtλkmk(|cnk|+‖Un‖∞). | (3.28) |
The proof of (3.27) is completed using (3.4) and (3.6).
Lemma 3.5. The approximations
Proof. The proof is a standard argument (e.g., the proof of Proposition 2.4 of [1]) using Lemmas 3.1, 3.2, and 3.3 for the
Remark 6. In Sections 5 and 6 we will assume that the particle trajectories do not intersect except possibly on a subset of
In what follows
Lemma 4.1.
∑j∈Z|Δ+Wnk,j|=1,∑j∈Z|Wn+1k,j−Wnk,j|≤1/2. | (4.1) |
Proof. The first part of (4.1) is evident from (3.5). For the second part of (4.1), we write (3.11) in the form
Wn+1k,j−Wnk,j=αnkΔ+Wnk,j−βnkΔ−Wnk,j. | (4.2) |
Taking absolute values, and recalling from the proof of Lemma 3.1 that
|Wn+1k,j−Wnj|≤αnk|Δ+Wnk,j|+βnk|Δ−Wnk,j|. | (4.3) |
Then summing over
Lemma 4.2. As
Proof. Lemma 4.1 along with
A standard Lax-Wendroff calculation [9] proves that
∂twk+h′k(t)∂xwk=0,wk(x,0)=H(x−hk(0)). | (4.4) |
One such weak solution is
∫T0∫R(˜wk−wk){ϕt+h′k(t)ϕx}dxdt=∫T0(˜wk−wk)ϕ(x,T)dt. | (4.5) |
Fix
ϕ(x,t)=∫tTψ(x−hk(t)+hk(σ),σ)dσ. | (4.6) |
It is readily verified that
∫T0∫R(˜wk−wk)ψ(x,t)dxdt=0. | (4.7) |
Since (4.7) holds for any
The following lemma is a direct consequence of (2.12), Lemma 3.5, and Lemma 4.2.
Lemma 4.3. Define
In this section we verify that the subsequential limit
Here and in Section 6 we will employ the test function
ψ′δ(x)={ηδ(x+δ/2),x≤0,−ηδ(x−δ/2),x≥0, | (5.1) |
where
supp(ηδ)=[−δ/2,δ/2],ηδ(x)≥0∀x∈R,∫Rηδ(x)dx=1. | (5.2) |
Assumption 5.1. Assume that the particle trajectories do not intersect except possibly on a subset
Remark 7. The set
Fi,j:={t∈(0,T)|hi(t)=hj(t)}. |
Since each of the particle paths
Lemma 5.1. Define
Proof. The partial derivatives of
∂G∂Unj=1−q,∂G∂Unj+1=−μ2Unj+1+q2,∂G∂Unj−1=μ2Unj−1+q2. | (5.3) |
Clearly
∂G∂Unj±1≥12(q−μ‖Un‖∞)≥12(q−μ(‖u0‖∞+K∑k=1λk)). | (5.4) |
In view of (5.4) and (3.1) it is clear that
For
∂Zn+1j∂Znj=1−q,∂Zn+1j∂Znj+1=q2−μ2Znj+1+μ2K∑k=1λkˆWnk,j,∂Zn+1j∂Znj−1=q2+μ2Znj−1−μ2K∑k=1λkˆWnk,j. | (5.5) |
It is readily verified that each of these partial derivatives is nonnegative using (3.1) and the fact that
The following lemma is a straightforward consequence of (3.5) and Lemma 4.2.
Lemma 5.2. Define
Snj=K∑k=1λkˆWnk,j,SΔ(x,t)=N∑n=0∑j∈Zχj(x)χn(t)Snj. | (5.6) |
0≤Snj≤K∑k=1λk,Δ+Snj≥0,∑j∈ZΔ+Snj=K∑k=1λk, | (5.7) |
and as
Lemma 5.3. The following discrete entropy inequalities hold for all
Zn+1j∨κ≤Znj∨κ−μΔ−ˉf(Znj+1∨κ,Znj∨κ)+μ2Snj(Znj+1∨κ−Znj−1∨κ),Zn+1j∧κ≥Znj∧κ−μΔ−ˉf(Znj+1∧κ,Znj∧κ)+μ2Snj(Znj+1∧κ−Znj−1∧κ). | (5.8) |
Proof. Writing (2.13) in the form
Lemma 5.4. The limit solution
Proof. We start with the first inequality in (5.8), and use the identity
Aj(Bj+1−Bj−1)=Δ+(AjBj)−Bj+1Δ+Aj+Δ−(AjBj)−Bj−1Δ−Aj. | (5.9) |
This results in
Zn+1j∨κ≤Znj∨κ−μΔ−(ˉf(Znj+1∨κ,Znj∨κ)−12Snj+1(Znj+1∨κ)−12Snj(Znj∨κ))−μ2((Znj+1∨κ)Δ+Snj+(Znj−1∨κ)Δ−Snj). | (5.10) |
Since
(Znj+1∨κ)Δ+Snj≥κΔ+Snj,(Znj−1∨κ)Δ−Snj≥κΔ−Snj, | (5.11) |
and so we can replace (5.10) by
Zn+1j∨κ≤Znj∨κ−μΔ−(ˉf(Znj+1∨κ,Znj∨κ)−12Snj+1(Znj+1∨κ)−12Snj(Znj∨κ))−μκ2(Snj+1−Snj−1). | (5.12) |
Following the proof of the Lax-Wendroff theorem [9], let
ΔxΔt∑j∈Z∑n≥0(Zn+1j∨κ)ϕn+1j−ϕnjΔt+ΔxΔt∑j∈Z∑n≥0(ˉf(Znj+1∨κ,Znj∨κ)−12Snj(Znj∨κ)−12Snj+1(Znj+1∨κ))Δ+ϕnjΔx+ΔxΔtκ∑j∈Z∑n≥0SnjΔ+ϕnjΔx≥0. | (5.13) |
Letting
∫T0∫R(z∨κ)ϕtdxdt+∫T0∫R(f(z∨κ)−K∑l=1λlH(x−hl(t))(z∨κ))ϕxdxdt+κ∫T0∫RK∑l=1λlH(x−hl(t))ϕxdxdt≥0. | (5.14) |
After simplifying the last integral the result is
∫T0∫R(z∨κ)ϕtdxdt+∫T0∫R(f(z∨κ)−K∑l=1λlH(x−hl(t))(z∨κ))ϕxdxdt−κK∑l=1λl∫T0ϕ(hl(t),t)dt≥0. | (5.15) |
A similar calculation starting from the second inequality of (5.8) yields
∫T0∫R(z∧κ)ϕtdxdt+∫T0∫R(f(z∧κ)−K∑l=1λlH(x−hl(t))(z∧κ))ϕxdxdt−κK∑l=1λl∫T0ϕ(hl(t),t)dt≤0. | (5.16) |
Recalling Assumption 5.1 and Remark 7, fix an interval
h1(t)<h2(t)<⋯<hk(t)<⋯<hK(t),t∈Im. | (5.17) |
Let
∫Im{f(z−∨κ)−ck(z−∨κ)−(f(z+∨κ)−ck(z+∨κ)−λk(z+∨κ))−γk(z−∨κ−z+∨κ)−λkκ}ρ(t)dt≥0, | (5.18) |
where
∫Im{f(z−∧κ)−ck(z−∧κ)−(f(z+∧κ)−ck(z+∧κ)−λk(z+∧κ))−γk(z−∧κ−z+∧κ)−λkκ}ρ(t)dt≤0. | (5.19) |
Continuing with the abbreviation
f(z−∨κ)−ck(z−∨κ)−(f(z+∨κ)−ck(z+∨κ)−λk(z+∨κ))−γk(z−∨κ−z+∨κ)−λkκ≥0, | (5.20) |
f(z−∧κ)−ck(z−∧κ)−(f(z+∧κ)−ck(z+∧κ)−λk(z+∧κ))−γk(z−∧κ−z+∧κ)−λkκ≤0. | (5.21) |
Fix a time
z−≤κ≤z+⟹{f(z+)−f(κ)≤(λk+˜ck)(z+−κ),f(z−)−f(κ)≤˜ck(z−−κ). | (5.22) |
where
z+≤κ≤z−⟹{f(z+)−f(κ)≥(λk+˜ck)(z+−κ),f(z−)−f(κ)≥˜ck(z−−κ). | (5.23) |
Plugging
z++z−≤2(λk+˜ck). | (5.24) |
The second inequality of (5.22) (for
z−≥˜ck. | (5.25) |
Substituting
z++z−≥2˜ck. | (5.26) |
Finally, with
z+≤λk+˜ck. | (5.27) |
Thus either
u+−ck=u−−ck−λk, | (5.28) |
or
u−−ck≥0,u+−ck≤0,−λk≤(u−−ck)+(u+−ck)≤λk. | (5.29) |
Recalling Definition 1.1, and that
(u−,u+)∈G(λk,ck)=G(λk,h′k(t)), | (5.30) |
and this holds for a.e.
Lemma 5.5. The following discrete entropy inequality holds for each
|Un+1j−κ|≤|Unj−κ|−μΔ−ˉF(Unj+1,Unj)+μ2K∑k=1λk|cnk−ˆUnj|(Wnk,j+1−Wnk,j−1), | (5.31) |
where
Proof. First assume that
Un+1j=G(Unj+1,Unj,Unj−1)+Qnj, | (5.32) |
where
Vn+1j:=G(Unj+1,Unj,Unj−1)=Unj−μΔ−ˉfnj+1/2,Qnj=μ2K∑k=1λk(cnk−ˆUnj)(Wnk,j+1−Wnk,j−1). | (5.33) |
Invoking the monotonicity of
|Vn+1j−κ|≤|Unj−κ|−μΔ−ˉF(Unj+1,Unj), | (5.34) |
for
Now take the case where
Un+1j≤Unj−μΔ−ˉfnj+1/2+|Qnj|. | (5.35) |
which, recalling the first equation of (2.6), is clearly satisfied. The case where
Lemma 5.6. The limit
Proof. Define
∫T0∫R(|u−κ|ϕt+F(u,κ)ϕx)dxdt+∫R|u0−κ|ϕ(x,0)dx≥0 | (5.36) |
for every
The proof is based on the discrete entropy inequality (5.31). Due to the bounds on
μ2K∑k=1λk|cnk−ˆUnj|(Wnk,j+1−Wnk,j−1)≤μ2BK∑k=1λk(Wnk,j+1−Wnk,j−1). | (5.37) |
Substituting into (5.31) the result is
|Un+1j−κ|≤|Unj−κ|−μΔ−ˉF(Unj+1,Unj)+μ2BK∑k=1λk(Wnk,j+1−Wnk,j−1). | (5.38) |
Multiplying by
ΔxΔtN∑n=0∑j∈Z{|Un+1j−κ|(ϕn+1j−ϕnj)/Δt+ˉF(Unj+1,Unj)(ϕnj+1−ϕnj)/Δx}−BK∑k=1λkΔxΔtN∑n=0∑j∈ZWnk,j12(ϕnj+1−ϕnj−1)/Δx+Δx∑j∈Z|U0j−κ|ϕ0jdx≥0. | (5.39) |
Letting
∫T0∫R(|u−κ|ϕt+F(u,κ)ϕx)dxdt−BK∑k=1λk∫T0∫RH(x−hk(t))ϕxdxdt+∫R|u0(x)−κ|dx≥0. | (5.40) |
The proof is finished by observing that
In this section we prove that the limit
Lemma 6.1. The limit
Proof. Fix a particle with index
mkank=−∑j∈Zλk2(cnk−ˆUnj)(Wnk,j+1−Wnk,j−1). | (6.1) |
Define
mkank=−λk2∑j∈Z(cnk−ˆUnj)(Wnk,j+1−Wnk,j−1)ψnj−λk2∑j∈Z(cnk−ˆUnj)(Wnk,j+1−Wnk,j−1)(1−ψnj). | (6.2) |
Next we multiply by
mkΔt∑n≥0ankξn=−λk2Δt∑n≥0∑j∈Z(cnk−ˆUnj)(Wnk,j+1−Wnk,j−1)ψnjξn−λk2Δt∑n≥0∑j∈Z(cnk−ˆUnj)(Wnk,j+1−Wnk,j−1)(1−ψnj)ξn. | (6.3) |
We solve for
(cnk−ˆUnj)(Wnk,j+1−Wnk,j−1)=2λkμ(Un+1j−Unj+μΔ−ˉfnj+1/2)−1λk∑l≠kλl(cnl−ˆUnj)(Wnl,j+1−Wnl,j−1), | (6.4) |
and substitute into the first sum on the right side of (6.3). The result is
mkΔt∑n≥0ankξn=−Δx∑n≥0∑j∈Z(Un+1j−Unj+μΔ−ˉfnj+1/2)ψnjξn⏟S1+12Δt∑n≥0∑j∈Z∑l≠kλl(cnl−ˆUnj)(Wnl,j+1−Wnl,j−1)ψnjξn⏟S2−λk2Δt∑n≥0∑j∈Z(cnk−ˆUnj)(Wnk,j+1−Wnk,j−1)(1−ψnj)ξn⏟S3. | (6.5) |
Summing the left side of (6.5) by parts, we find that
mkΔt∑n≥0ankξn=−mkΔt∑n≥0cn+1kξn+1−ξnΔt. | (6.6) |
Letting
mkΔt∑n≥0ankξn→−mk∫T0h′k(t)ξ′(t)dt, | (6.7) |
and for
S1→∫T0∫R{u∂t(ψδ(x−hk(t))ξ(t))+f(u)∂x(ψδ(x−hk(t))ξ(t))}dxdt. | (6.8) |
We next estimate
S2,l=12Δt∑n≥0∑j∈Zλl(cnl−ˆUnj)(Wnl,j+1−Wnl,j−1)ψnjξn. | (6.9) |
Since
|S2,l|≤BΔt∑n≥0|ξn|∑j∈Z(Wnl,j+1−Wnl,j−1)ψnj | (6.10) |
where
∑j∈Z(Wnl,j+1−Wnl,j−1)ψnj=∑j∈Z(Wnl,j+1ψnj+1−Wnl,j−1ψnj−1)−∑j∈Z(Wnl,j+1+Wnl,j)Δ+ψnj. | (6.11) |
The first sum on the right is telescoping and is equal to zero. Thus, referring back to (6.10) we have
|S2,l|≤−BΔt∑n≥0|ξn|∑j∈Z(Wl,j+1+Wl,j)Δ+ψnj=−2BΔxΔt∑n≥0|ξn|∑j∈Z12(Wl,j+1+Wl,j)Δ+ψnj/Δx. | (6.12) |
Letting
lim supΔ→0|S2,l|≤−2B∫T0|ξ(t)|∫Rwl(x,t)∂xψδ(x−hk(t))dxdt. | (6.13) |
Recalling that
∫Rwl(x,t)∂xψδ(x−hk(t))dx=∫∞x=hl(t)∂xψδ(x−hk(t))dx=−ψδ(hl(t)−hk(t)). | (6.14) |
Substituting into (6.13) yields the desired estimate of
lim supΔ→0|S2,l|≤2B∫T0|ξ(t)|ψδ(hl(t)−hk(t))dt. | (6.15) |
We claim that
|S3|≤BΔt∑n≥0|ξn|∑j∈Z(Wnk,j+1−Wnk,j−1)(1−ψnj). | (6.16) |
where
∑j∈Z(Wnk,j+1−Wnk,j−1)(1−ψnj)=∑j∈Z(Wnk,j+1(1−ψnj+1)−Wnk,j−1(1−ψnj−1))+∑j∈Z(Wnk,j+1+Wnk,j)Δ+ψnj. | (6.17) |
In the second term on the right side we have used
|S3|≤2BΔt∑n≥0|ξn|+BΔt∑n≥0|ξn|∑j∈Z(Wk,j+1+Wk,j)Δ+ψnj=2BΔt∑n≥0|ξn|+2BΔtΔx∑n≥0|ξn|∑j∈Z12(Wk,j+1+Wk,j)Δ+ψnj/Δx. | (6.18) |
Letting
lim supΔ→0|S3|≤2B∫T0|ξ(t)|dt+2B∫T0|ξ(t)|∫Rwk(x,t)∂xψδ(x−hk(t)dxdt. | (6.19) |
Substituting
∫Rwk(x,t)∂xψδ(x−hk(t))dx=∫∞x=hk(t)∂xψδ(x−hk(t))dx=−1. | (6.20) |
Plugging (6.20) into (6.19) completes the proof of the claim.
Combining
−mk∫T0h′k(t)ξ′(t)dt=∫T0∫R{u(ψδ(x−hk(t))ξ(t))t+f(u)(ψδ(x−hk(t))ξ(t))x}dxdt+∫Ru0(x)ψδ(x−hk(0))ξ(0)dx+Rk, | (6.21) |
where
|Rk|≤2B∑l≠k∫T0|ξ(t)|ψδ(hl(t)−hk(t))dt. | (6.22) |
Next we consider the limit when
ψδ(hl(t)−hk(t))→0fora.e.t∈(0,T), | (6.23) |
with the result that
[u(hk(t),t)]=u(hk(t)+,t)−u(hk(t)−,t),[f(u(hk(t),t))]=f(u(hk(t)+,t))−f(u(hk(t)−,t)). | (6.24) |
A straightforward calculation using (5.1), (5.2) gives
∫T0∫R{u(ψδ(x−hk(t))ξ(t))t+f(u)(ψδ(x−hk(t))ξ(t))x}dxdt→∫T0{h′k(t)[u(hk(t),t)]−[f(u(hk(t),t))]}ξ(t)dt, | (6.25) |
and
∫Ru0(x)ψδ(x−hk(0))ξ(0)dx→0. | (6.26) |
The result is that (6.21) becomes
−mk∫T0h′k(t)ξ′(t)dt=∫T0{h′k(t)[u(hk(t),t)]−[f(u(hk(t),t))]}ξ(t)dt. | (6.27) |
After integrating the left side by parts the result is
∫T0{mkh″k(t)−[u(hk(t),t)]h′k(t)−[f(u(hk(t),t))]}ξ(t)dt=0, | (6.28) |
implying that (1.4) holds for a.e.
The observation that for all
Proof of the main theorem.
Proof. Lemma 3.5 provides the convergence portion of Theorem 1.3. That the limit
Remark 8. For the single-particle case, Theorem 8 of [6] states that Definition 1.2 is sufficient for uniqueness. Thus if
It is possible to somewhat reduce the excessively diffusive nature of Lax-Friedrichs differencing without adding too much complexity by using the MUSCL approach. Our incorporation of MUSCL processing is standard [12]. Let
M(a,b)=12(sgn(a)+sgn(b))min(|a|,|b|). | (7.1) |
We replace the numerical fluxes
˜fnj+1/2=12((Un,−j+1)2/2+(Un,+j)2/2)−q2μ(Un,−j+1−Un,+j),˜gnk,j+1/2=12(cnkWn,−k,j+1+cnkWn,+k,j)−q2μ(Wn,−k,j+1−Wn,+k,j), | (7.2) |
where
Un,±j=Unj±12M(Δ+Unj,Δ−Unj),Wn,±k,j=Wnk,j±12M(Δ+Wnk,j,Δ−Wnk,j). | (7.3) |
We do not presently have any convergence results or even stability estimates for the resulting scheme with MUSCL processing incorporated. A moderate amount of numerical experience indicates that the algorithm produces approximations that converge to the same solution as the basic algorithm of Section 2.
Following are a few numerical examples. We refer to the scheme of Section 2 as the basic scheme, and the modified scheme of Section 7 as the MUSCL scheme. We used
Example 8.1. This is a single-particle Riemann problem, with
(uL,uR)=(.15,−.15),(h(0),h′(0))=(0,.65),λ=.5,m=2. | (8.1) |
The exact solution is available for comparison, using the results of [11]. See Figure 1. The approximations appear to improve when the mesh size is halved, as expected. It is also apparent that the MUSCL scheme is more accurate than the basic one.
The sharp transition at
Example 8.2. This is another single-particle Riemann problem with
(uL,uR)=(.25,.75),(h(0),h′(0))=(0,.65),λ=.5,m=1. | (8.2) |
As in the previous example the exact solution is available via [11]. This example displays a spurious kink, see Figure 2, that appears in some cases where a particle's velocity
Example 8.3. This is a two-particle example with
h″k+λkmkh′k=σk(t),k=1,2. | (8.3) |
Here
σk(t)=λkˆz−λ2k/2mk+pk(t), | (8.4) |
where
p1(t)={0,h1(t)<h2(t),−λ1λ2m1,h1(t)>h2(t),p2(t)={−λ1λ2m2,h1(t)<h2(t),0,h1(t)>h2(t). | (8.5) |
Assume that the particle trajectories do not intersect except for a finite set of times
hk(t)=hk(τν)+h′k(τν)rk(1−exp(−rk(t−τν)))−σkr2k(1−exp(−rk(t−τν)))+σkrk(t−τν). | (8.6) |
The parameters used in this example are
m1=.025,m2=.02,(h1(0),h′1(0))=(.2,1.2),(h2(0),h′2(0))=(.3,0.9),λ1=.75,λ2=.5,ˆz=.5. | (8.7) |
See Figures 3 and 4. From Figure 3 it appears that the MUSCL scheme is more accurate than the basic scheme, as expected. We also see that the discrete
Example 8.4. This is another two-particle example. This time the particles are initially heading toward each other, and the fluid is initially at rest. Unlike the previous example the true solution is not known. In Figure 5 we show the particle trajectories at three levels of grid refinement. It appears that the particle trajectories are converging as the mesh size is refined. The MUSCL scheme is better able to resolve the fine details of the trajectory, especially after the first crossing of trajectories.
The initial fluid velocity is zero,
m1=.04,m2=.02,(h1(0),h′1(0))=(.1,−2),(h2(0),h′2(0))=(−.1,4),λ1=λ2=1. | (8.8) |
The author thanks an anonymous referee for providing the now improved version of Assumption 5.1, and sharing ideas about how to weaken Assumption 5.1 for future efforts to address much more general particle interaction scenarios.
[1] |
Convergence of finite volume schemes for the coupling between the inviscid Burgers equation and a particle. Math. Comp. (2017) 86: 157-196. ![]() |
[2] |
Riemann problem for a particle-fluid coupling. Math. Models Methods Appl. Sci. (2015) 25: 39-78. ![]() |
[3] |
N. Aguillon, Numerical simulations of a fluid-particle coupling, in Finite Volumes for Complex Applications Ⅶ-Elliptic, Parabolic and Hyperbolic Problems. Springer Proceedings in Mathematics & Statistics (eds. J. Fuhrmann, M. Ohlberger M. and C. Rohde), Springer, 78 (2014), 759–767. doi: 10.1007/978-3-319-05591-6_76
![]() |
[4] |
A theory of L1-dissipative solvers for scalar conservation laws with discontinuous flux. Arch. Ration. Mech. Anal. (2011) 201: 27-86. ![]() |
[5] |
Small solids in an inviscid fluid. Netw. Heterog. Media (2010) 5: 385-404. ![]() |
[6] |
Well-posedness for a one-dimensional fluid-particle interaction model. SIAM J. Math. Anal. (2014) 46: 1030-1052. ![]() |
[7] |
Analysis of a Burgers equation with singular resonant source term and convergence of well-balanced schemes. Discrete Contin. Dyn. Syst. (2012) 32: 1939-1964. ![]() |
[8] | Monotone difference approximations for scalar conservation laws. Math. Comp. (1980) 34: 1-21. |
[9] |
H. Holden and N. H. Risebro, Front Tracking for Hyperbolic Conservation Laws, Springer-Verlag, New York, 2002. doi: 10.1007/978-3-642-56139-9
![]() |
[10] | C. Klingenberg, J. Klotzky and N. Seguin, On well-posedness for a multi-particle fluid model, in Theory, Numerics and Applications of Hyperbolic Problems Ⅱ. HYP 2016. Springer Proceedings in Mathematics & Statistics, (eds. C. Klingenberg and M. Westdickenberg), Springer, 237 (2018), 167–177. |
[11] |
A simple 1D model of inviscid fluid-solid interaction. J. Differ. Equ. (2008) 245: 3503-3544. ![]() |
[12] |
(2002) Finite Volume Methods for Hyperbolic Problems. Cambridge, UK: Cambridge University Press. ![]() |
[13] |
Existence of strong traces for quasi-solutions of multidimensional conservation laws. J. Hyperbolic Differ. Equ. (2007) 4: 729-770. ![]() |
[14] |
A fixed grid, shifted stencil scheme for inviscid fluid-particle interaction. Appl. Numer. Math. (2016) 110: 26-40. ![]() |
1. | Boris Andreianov, Abraham Sylla, Finite volume approximation and well-posedness of conservation laws with moving interfaces under abstract coupling conditions, 2023, 30, 1021-9722, 10.1007/s00030-023-00857-9 |