Loading [MathJax]/jax/element/mml/optable/GeneralPunctuation.js
Review Topical Sections

Overview of wind power development over the two past decades (2000-2019) and its role in the Taiwan's energy transition and sustainable development goals

  • Received: 11 December 2020 Accepted: 22 February 2021 Published: 03 March 2021
  • In response to the international agreement on the climate change and the energy supply diversification, the Taiwan government actively promoted the use of wind power since 2000. This paper highlighted the Taiwan's wind resources, and interactively analyzed the regulatory strategies for promoting wind power and trend analysis of wind power over the two past decades (2000–2019). In the first development stage, the electricity generation by onshore wind power rapidly grew from 1.38 gigawatt-hours (GW-h) in 2000 to 1,492.7 GW-h in 2011 because of the implementations of promotional measures and official subsidies. However, its growth rate became slow due to the restrictions on available land for the onshore wind power farms. With the promulgation of the Renewable Energy Development Act in 2009, the government not only declared preferable feed-in-tariff (FIT) rates, but also announced some incentive and promotional plans, focusing on offshore wind power farms from the early 2010s. Therefore, the electricity generation by offshore wind power significantly increased from 26.7 GW-h in 2018 to 175.4 GW-h in 2019. Using the data on the electricity generation by wind power in 2019 (i.e., 1,871.3 × 106 kW-h) and the average default value (i.e., 0.539 kg CO2/kW-h), the equivalent mitigation of CO2 emission from the wind power systems in Taiwan was thus estimated to be about one million metric tons. Based on the progressive development of wind power and other renewable resources over the past two decades, one of the sustainable development goals (SDGs) for nuclear-free homeland and 20% renewable electricity generation may be prospective in the energy transition by 2025. However, this goal has faced serious challenges because the corona virus disease 2019 (COVID-19) has caused the delays in the Taiwanese wind power industry.

    Citation: Wen-Tien Tsai. Overview of wind power development over the two past decades (2000-2019) and its role in the Taiwan's energy transition and sustainable development goals[J]. AIMS Energy, 2021, 9(2): 342-354. doi: 10.3934/energy.2021018

    Related Papers:

    [1] Abraham Sylla . Influence of a slow moving vehicle on traffic: Well-posedness and approximation for a mildly nonlocal model. Networks and Heterogeneous Media, 2021, 16(2): 221-256. doi: 10.3934/nhm.2021005
    [2] Maria Laura Delle Monache, Paola Goatin . Stability estimates for scalar conservation laws with moving flux constraints. Networks and Heterogeneous Media, 2017, 12(2): 245-258. doi: 10.3934/nhm.2017010
    [3] Martin Gugat, Alexander Keimer, Günter Leugering, Zhiqiang Wang . Analysis of a system of nonlocal conservation laws for multi-commodity flow on networks. Networks and Heterogeneous Media, 2015, 10(4): 749-785. doi: 10.3934/nhm.2015.10.749
    [4] Christophe Chalons, Paola Goatin, Nicolas Seguin . General constrained conservation laws. Application to pedestrian flow modeling. Networks and Heterogeneous Media, 2013, 8(2): 433-463. doi: 10.3934/nhm.2013.8.433
    [5] Frederike Kissling, Christian Rohde . The computation of nonclassical shock waves with a heterogeneous multiscale method. Networks and Heterogeneous Media, 2010, 5(3): 661-674. doi: 10.3934/nhm.2010.5.661
    [6] Alexandre M. Bayen, Alexander Keimer, Nils Müller . A proof of Kirchhoff's first law for hyperbolic conservation laws on networks. Networks and Heterogeneous Media, 2023, 18(4): 1799-1819. doi: 10.3934/nhm.2023078
    [7] 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
    [8] Felisia Angela Chiarello, Giuseppe Maria Coclite . Nonlocal scalar conservation laws with discontinuous flux. Networks and Heterogeneous Media, 2023, 18(1): 380-398. doi: 10.3934/nhm.2023015
    [9] Michael Herty, Niklas Kolbe, Siegfried Müller . Central schemes for networked scalar conservation laws. Networks and Heterogeneous Media, 2023, 18(1): 310-340. doi: 10.3934/nhm.2023012
    [10] Gabriella Bretti, Roberto Natalini, Benedetto Piccoli . Numerical approximations of a traffic flow model on networks. Networks and Heterogeneous Media, 2006, 1(1): 57-84. doi: 10.3934/nhm.2006.1.57
  • In response to the international agreement on the climate change and the energy supply diversification, the Taiwan government actively promoted the use of wind power since 2000. This paper highlighted the Taiwan's wind resources, and interactively analyzed the regulatory strategies for promoting wind power and trend analysis of wind power over the two past decades (2000–2019). In the first development stage, the electricity generation by onshore wind power rapidly grew from 1.38 gigawatt-hours (GW-h) in 2000 to 1,492.7 GW-h in 2011 because of the implementations of promotional measures and official subsidies. However, its growth rate became slow due to the restrictions on available land for the onshore wind power farms. With the promulgation of the Renewable Energy Development Act in 2009, the government not only declared preferable feed-in-tariff (FIT) rates, but also announced some incentive and promotional plans, focusing on offshore wind power farms from the early 2010s. Therefore, the electricity generation by offshore wind power significantly increased from 26.7 GW-h in 2018 to 175.4 GW-h in 2019. Using the data on the electricity generation by wind power in 2019 (i.e., 1,871.3 × 106 kW-h) and the average default value (i.e., 0.539 kg CO2/kW-h), the equivalent mitigation of CO2 emission from the wind power systems in Taiwan was thus estimated to be about one million metric tons. Based on the progressive development of wind power and other renewable resources over the past two decades, one of the sustainable development goals (SDGs) for nuclear-free homeland and 20% renewable electricity generation may be prospective in the energy transition by 2025. However, this goal has faced serious challenges because the corona virus disease 2019 (COVID-19) has caused the delays in the Taiwanese wind power industry.



    Partial differential equations (PDEs) on networks have a large number of applications, including fluid flow in pipelines, traffic flow on a network of roads, blood flow in vessels, data networks, irrigation channels and supply chains. A treatment of this wide range of applications can be found in the review articles [6,14] and the references therein. In this paper we will focus on scalar, one-dimensional conservation laws

    ut+f(u)x=0 (1.1)

    on a network. Here, u=u(x,t) is the unknown conserved variable and f is a scalar flux function defined either on R or some subinterval. We aim to make sense of the conservation law on a directed graph and obtain existence, uniqueness, stability and approximability results.

    Consider a network represented by a connected and directed graph. We tag the edges of this graph with an index k and impose on each edge a scalar conservation law

    ukt+fk(uk)x=0,xDk, t>0uk(x,0)=ˉuk(x),xDk (1.2)

    for some spatial domain DkR. (Here and in the remainder, a superscript k will refer to an edge or a vertex.) We may think of edges as pipes or roads and the vertices as intersections, with the convention that the direction of travel is in the positive x-direction, as shown in Figure 1.

    Figure 1. 

    A star shaped network with two ingoing and three outgoing edges

    .

    In this paper we will be interested in uniqueness and stability for nonlinear scalar conservation laws on a network, as well as in constructing a numerical approximation and proving convergence of the numerical scheme. As opposed to many existing results, where the flux function on each edge is the same [10,18], we want to allow for a different flux function fk on each edge Dk of the network. Assuming that each flux fk is continuous and independent of the space variable, the problem can be seen as a PDE with a discontinuous flux, with the points of discontinuity sitting on the vertices of the graph. In fact, if our network would be the trivial network with only one ingoing and one outgoing edge then this would be exactly the case of a conservation law on the real line with a flux function with one discontinuity located at the vertex. Because of the connection to the theory for conservation laws with discontinuous fluxes (see e.g. [3]), we will borrow several ideas from this theory. It is well-established that nonlinear hyperbolic conservation laws develop shocks in finite time. Therefore, solutions are always understood in the weak sense. Unfortunately, weak solutions to nonlinear hyperbolic conservation laws turn out to be non-unique, and additional conditions, usually referred to as entropy conditions, are imposed to select a unique solution. If the flux function is continuous then the theory of entropy solutions is covered by Kruzkhov's theory [21]. For conservation laws with discontinuous fluxes the choice of entropy conditions is not obvious, and different physical models might lead to different entropy conditions. Although suitable entropy conditions can yield uniqueness, different entropy conditions are known to yield different solutions; see [30,19,7,3] and references therein. In [30] the author shows convergence of a finite difference scheme scheme under the assumption of a strictly concave flux function with a single maximum. In a later paper [19] a uniqueness result was shown for degenerate parabolic convection-diffusion equations, of which hyperbolic conservation laws are a subcase. In this work, the flux function had to satisfy a so-called "crossing condition". Another convergence result for an Engquist–Osher type scheme was given in [7]. The flux functions f,g were assumed to have a single maximum and to satisfy f(u)=g(u)=0 for u=0,1. The study of different entropy conditions for conservation laws with discontinuous fluxes culminated in the paper by Andreianov, Karlsen and Risebro [3]. The authors relate the question of admissibility of a solution to the properties of a set of constant solutions, a so-called germ. Inspired by the entropy theory of Andreianov, Karlsen and Risebro, we investigate so-called stationary and discrete stationary solutions for our graph problem and thus derive an entropy theory for conservation laws on networks. Although our theory is influenced by the theory in [3], we have strived to make this paper as self-contained as possible.

    In [18] the authors show uniqueness and existence to the Riemann problem as well as existence of a weak solution of the Cauchy problem on a network of roads in the case that the flux functions on each edge are identical. In [9,8,1,10] the authors show well-posedness results for vanishing viscosity solutions. In [15] the authors investigate two entropy type conditions. However, in none of the existing literature can one find a general entropy condition which leads to uniqueness and stability of solutions. In the present work we aim to address this deficiency in the existing theory of conservation laws on networks.

    The second important question to address is existence of a solution. Our approach will be to construct an approximation of the exact entropy solution by constructing a finite volume scheme. We will prove convergence to an entropy solution, thereby also proving existence of a solution. Convergence to the unique entropy solution of numerical schemes has been shown for conservation laws with strictly concave flux functions. This was done for schemes which are implicit on the nodes in [1,Section 3.2] and [2]. Convergence of a fully explicit scheme for the strictly concave case was shown in [29]. For a general overview over numerical methods for conservation laws on graphs see [6,Section 6].

    In this article we focus on monotone fluxes – that is, each flux fk is either increasing or decreasing. For non-monotone fluxes only some of the techniques in the present paper are applicable; a traffic flow model, in which fluxes are assumed to be strictly concave, will be treated in a forthcoming work [13].

    This article is structured as follows: In Section 2 we define our mathematical framework. We show uniqueness of entropy solutions to our problem in Section 3. In Section 4 we define a finite difference scheme appropriate for our problem, and in Section 5 we prove that our numerical scheme converges towards the unique entropy solution. In Section 6 we show that a class of monotone flux functions fits in our general scheme. Numerical experiments for the monotone case are presented in Section 7.

    While the theory outlined in Sections 2 through 4 holds for conservation laws with general flux functions, the convergence theory in Sections 5 and 7 focuses on monotone flux functions and upwind numerical fluxes.

    Consider a network (or directed graph) of vertices and edges; for simplicity we will assume that the network contains a single vertex, along with NinN edges entering and NoutN edges exiting the vertex (see Figure 1). (The generalization to general networks will follow analogously, due to the finite speed of propagation of the equations considered.) We think of the Nin edges as being to the left of the vertex and the Nout edges to its right. The ingoing edges will be labelled kIin:={Nin,,1} and the outgoing edges kIout:={1,,Nout}. We denote N:=Nin+Nout and we let I:=IinIout denote the set of all edge indices. Placing the vertex at the origin x=0, the incoming edges have coordinates xR=(,0), while the outgoing edges have coordinates xR+=(0,); we will denote the k-th edge by

    Dk={Rfor kIin,R+for kIout.

    On each edge Dk we now impose the scalar conservation law (1.1), resulting in the N distinct equations

    ukt+fk(uk)x=0for xDk, kI. (2.1)

    The collection of functions u=(uk)kI can be thought of as a function u:ΩR, where

    Ω:=kIDk×{k}.

    On the Borel σ-algebra B(Ω){kIAk×{k} : AkB(Dk)} on Ω we define the measure λ=L×#, where L is the one-dimensional Lebesgue measure and # is the counting measure; thus, the integral of u=(uk)kI is

    Ωudλ=kIDkuk(x)dx. (2.2)

    The set of L-bounded, real-valued functions on Ω will be denoted by L(Ω;λ). We define the total variation of a function uL(Ω;λ) as the sum of the variations of its components:

    TV(u):=Ω|dudx|dλ=kIDk|dukdx(x)|dx. (2.3)

    Definition 2.1 (Weak Solution). We say that a function uL(R+;L(Ω;λ)) is a weak solution of (2.1) with initial data ˉuL(Ω;λ) if

    kI0Dkukφkt+fk(uk)φkxdxdt+kIDkˉuk(x)φk(x,0)dx=0 (2.4)

    for all φkCc(¯Dk×[0,)) satisfying φk(0,t)φ1(0,t) for all kI.

    Weak solutions automatically satisfy a Rankine–Hugoniot condition at the intersection:

    Proposition 2.2 (Rankine–Hugoniot condition). Let (uk)kI be a weak solution of (2.1) such that fkuk(,t) has a strong trace at x=0 for every kI and a.e. t>0. Then

    kIinfk(uk)(0,t)=kIoutfk(uk)(0,t)for a.e.t>0. (2.5)

    Proof. Define

    θε(x)={1ε(ε+x)if x[ε,0]1ε(εx)if x[0,ε]0if |x|>ε. (2.6)

    We define Φ(x,t):=θε(x)ψ(t) where ψCc([0,)). The partial derivatives of Φ are

    Φx(x,t)={1εψ(t)if x[ε,0]1εψ(t)if x[0,ε]0if |x|>εandΦt(x,t)=θε(x)ψ(t).

    By a density argument, Φ qualifies as an admissible test function. Thus, we can insert Φ into the weak formulation (2.4) to get

    0=kI0DkukΦkt+fk(uk)Φkxdxdt+kIDkˉuk(x)Φk(x,0)dx=kI0Dkukθε(x)ψ(t)dtdx+1εkI0Dk(ε,ε)sgn(k)fk(uk)ψ(t)dxdt1εkIDk(ε,ε)(εx)ˉuk(x,0)ψ(0)dxkI0sgn(k)fk(uk)ψ(t)dt

    as ε0, which is equivalent to (2.5).

    Definition 2.3 (Stationary Solution). A stationary solution of (2.1) is a weak solution of (2.1) which is constant in time and is a strong solution on each edge Dk. We see from (2.4) and (2.5) that the stationary solutions are precisely those satisfying uk(x,t)ckR for xDk, t0 and kI, and where ck satisfy the Rankine–Hugoniot condition

    kIinfk(ck)=kIoutfk(ck). (2.7)

    Thus, we can identify each stationary solution with a vector c=(ck)kIRN.

    Remark 2.4. Note that if we only required stationary solutions to be weak solutions on each edge Dk then they could exhibit arbitrarily many jump discontinuities. More precisely, if f is not injective then a "stationary weak solution" could jump arbitrarily often between values uk=ck,±, where f(ck,)=f(ck,+).

    Next, we formulate conditions that will single out a unique weak solution.

    Definition 2.5 (Kruzkov entropy pairs). The Kruzkov entropy pairs are the pairs of functions ηc(u)=|uc|, qkc(u)=sgn(uc)(fk(u)fk(c)) for cR.

    The Kruzkov entropy pairs lead to a consistency condition on sets of stationary solutions:

    Definition 2.6. A subset GRN consisting of stationary solutions of (2.1) is mutually consistent if

    kIinqkck(˜ck)kIoutqkck(˜ck) (2.8)

    for every pair c,˜cG, where qkc are the Kruzkov entropy flux functions. The set G is maximal if for every cRN, the condition that (2.8) holds for every ˜cG implies that also cG.

    The set of stationary solutions G will determine what class of initial data we can consider:

    Definition 2.7. Let GRN. We let Loco(G) be the set of L-bounded data of G,

    Loco(G)={uL(Ω;λ) :  c,dG s.t. ckuk(x)dk  (x,k)Ω.} (2.9)

    Example 2.8. If Nin=Nout=1 and fk(u)=f(u)=u2 then the stationary solutions are all cR2 of the form c=(c,c) or c=(c,c) for cR. Both G1={(c,c):cR} and G2={(c,c):c0} (as well as any subset of these) are mutually consistent, as is G1G2. Note that no point of the form (c,c) for c>0 can be added to any of these sets and remain mutually consistent. Similarly, no set containing both (c,c) and (d,d) for distinct c,d>0 can be mutually consistent. The set G1 stands out as the smallest closed set which is such that both components span all of R, i.e. the projection onto either component equals R. It is readily checked that Loco(G1)=L(Ω;λ), and that any strict subset of G1 will yield a strictly smaller set of initial data. It is similarly straightforward to check that G2 generates a very restrictive set of initial data:

    Loco(G2)={uL(Ω;λ) : u1c, u1c for some c0}.

    Thus, G1 is the smallest mutually consistent set of stationary solutions that allows for initial data in all of L(Ω;λ).

    Definition 2.9 (Entropy Solution). Let GRN be a mutually consistent set of stationary solutions of (2.1) and let ˉuL(Ω;λ). We say that a function uL(R+,L(Ω;λ)) is an entropy solution of (2.1) with respect to G with initial data ˉu if each uk is a Kruzkhov entropy solution on Dk for all kI (in the usual sense), and if

    kI0Dkηck(uk)φkt+qkck(uk)φkxdxdt+kI0ηck(ˉuk(x))φk(x,0)dx0 (2.10)

    for every cG and every 0φCc(Ω×[0,)) satisfying φk(0,t)φ1(0,t) for all kI.

    Audusse and Perthame [4] considered an entropy condition similar to (2.10), but in the context of spatially dependent, discontinuous flux functions.

    We show first that entropy solutions are invariant in the set Loco from Definition 2.7.

    Lemma 2.10. Let GRN be a mutually consistent set of stationary solutions of (2.1) and let uL(R+,L(Ω;λ)) be an entropy solution w.r.t. G with initial data ˉuLoco(G). Then u(t)Loco(G) for a.e. t>0.

    Proof. Select c,dG such that cˉud (cf. (2.9)). Add inequality (2.10) and equation (2.4) for both ck and uk to get

    kI0Dk(ckuk)+φkt+H(ckuk)(fk(ck)fk(uk))φkxdxdt+kI0(ckˉuk(x))+=0φk(x,0)dx0

    (where +=max(,0) and H=sgn+ is the Heaviside function). Replacing φ by a sequence of approximations of the identity function on the set Ω×[0,T] yields

    kIDk(ckuk(x,T))+dx0

    for a.e. T>0, whence u(T)c. It follows similarly that u(T)d, and hence, u(T)Loco(G).

    The above lemma enables us to show that entropy solutions have strong traces.

    Lemma 2.11. Let GRN be a mutually consistent set of stationary solutions of (2.1) and let uL(R+,L(Ω;λ)) be an entropy solution w.r.t. G with initial data ˉuLoco(G). Then the functions qkuk and fkuk admit strong traces on {x=0}, for any kI.

    Proof. It follows from Lemma 2.10 that uk is a Kruzkhov entropy solution on Dk, for any kI. We can therefore apply [25,Theorem 1.4] to obtain the desired conclusion.

    Proposition 2.12. Let GRN be a set of stationary solutions of (2.1). Let u be an entropy solution of (2.1) w.r.t. G such that qkckuk(,t) has a strong trace at x=0 for every kI and a.e. t>0. Then

    kIinqkck(uk)(0,t)kIoutqkck(uk)(0,t)for a.e.t>0 (2.11)

    for every cG.

    Proof. We take a positive test function 0ψCc((0,)). As in the proof of Proposition 2.2 we define Φ(x,t):=θε(x,t)ψ(t) where θε is given by (2.6). Now we insert Φ as test function into the entropy inequality (2.10) to get

    0kI0Dkηck(uk)θε(x)ψ(t)dxdt1εkI(0Dk(ε,ε)sgn(k)qkck(uk)ψ(t)dxdt+Dkηck(ˉuk(x))θε(x)ψ(t)dx)kI0sgn(k)qkck(uk(0,t))ψ(t)dt

    as ε0, which shows the desired inequality.

    Corollary 2.13. If G is maximal (cf. Definition 2.6), then the trace of any entropy solution lies in G.

    Theorem 3.1 (Entropy Solutions are L1 stable). Let GRN be a mutually consistent, maximal set of stationary solutions. Let u,v, be entropy solutions of (2.1) w.r.t. G with initial data ˉu,ˉvLoco(G)L1(Ω;λ). Let fk be Lipschitz continuous for all kI. Then

    kI

    for every t>0 . In particular, there exists at most one entropy solution for given initial data.

    Proof. From Lemma 2.10 it follows that {\mathbf{u}}(t), {\mathbf{v}}(t)\in L^\infty_{\mathrm{oco}}( \mathscr{G}) for a.e. t>0 . Let k\in {{\mathscr{I}}_{\mathrm{in}}} ; the case k\in {{\mathscr{I}}_{\mathrm{out}}} will follow analogously. The first step is a standard doubling of variables argument on each edge k\in {{\mathscr{I}}_{\mathrm{in}}} by selecting \varphi^k\in C_c^\infty(D^k\times[0, \infty)) and \varphi_l\equiv0 for l\neq k . Note that this is the point in the argument where we need the fluxes f^k, \ k\in {\mathscr{I}} to be Lipschitz continuous. The doubling of variables technique on a single edge gives:

    \begin{equation} \begin{split} \int_{D^k} \int_0^{\infty} \big|u^k(x, t) - v^k(x, t)\big| \varphi_t + q_v^k(u) \varphi_x \, dt \, dx& \\ + \int_{D^k} \big|\bar{u}^k(x) - \bar{v}^k(x)\big| \varphi(x, 0) \, dx& \geqslant 0. \end{split} \end{equation} (3.1)

    Next, for general \varphi^k\in C_c^\infty\big(\overline{D^k}\times[0, \infty)\big) , we cut off the functions near x = 0 and couple the terms (3.1) on each edge together by utilizing (2.11). For h>0 we define

    \mu_h (x) : = \begin{cases} 0 & x \in (-\infty, -2h) \\ \frac{1}{h}(x+2h) & x \in [-2h, -h) \\ 1 & x \in [-h, 0] \end{cases}

    and

    \Psi_h (x) : = 1 - \mu_h(x).

    The derivative of \Psi_h reads

    \Psi_h'(x) = \begin{cases} 0 & x \in (-\infty, -2h) \\ -\frac{1}{h} & x \in [-2h, -h) \\ 0 & x \in [-h, 0]. \end{cases}

    Define \varphi^k(x, t) : = \xi^k (x, t) \Psi_h (x) for a function \xi^k \in C_c^\infty \big(\overline{D^k} \times [0, \infty) \big) . We insert \varphi into equation (3.1) to get

    \begin{align*} \int_{D^k} \int_0^\infty \big|u^k(x, t) - v^k(x, t)\big| \xi_t^k \Psi_h + q^k_{v^k}(u^k) \xi_x^k \Psi_h \, dt \, dx& \\ + \int_{D^k} \int_0^\infty q^k_{v^k}(u^k) \xi^k \Psi_h' \, dt\, dx + \int_{D^k} \big|\bar{u}^k(x)-\bar{v}^k(x)\big| \xi^k \Psi_h \, dx& \geqslant 0. \end{align*}

    Sending h \downarrow 0 we get

    \begin{align*} \int_{D^k} \int_0^\infty \big|u^k(x, t) - v^k(x, t)\big| \xi_t^k + q^k_{v^k}(u^k) \xi_x^k \, dt\, dx + \int_{D^k} \big|\bar{u}^k(x) - \bar{v}^k(x)\big| \xi^k \, dx & \\ + \lim\limits_{h \downarrow 0} \int_{-2h}^{-h} \int_0^\infty q^k_{v^k}(u^k) \xi^k \Psi_h' \, dt\, dx & \geqslant 0. \end{align*}

    Since the traces of q^k (u^k) and q^k (v^k) exist, we get

    \begin{align*} -\lim\limits_{h \downarrow 0} \frac{1}{h} \int_0^T \int_{-2h}^{-h} q^k_{v^k}(u^k) \xi^k \, dx \, dt = -\int_0^T q_{v^{k-}}^k (u^k) \xi^k (0, t) \, dt. \end{align*}

    We therefore obtain

    \begin{equation} \begin{split} \int_{D^k} \int_0^\infty \big|u^k(x, t) - v^k(x, t)\big| \xi_t + q^k_{v^k}(u^k) \xi_x \, dt\, dx& \\ + \int_{D^k} \big|\bar{u}^k(x) - \bar{v}^k(x)\big| \, dx -\int_0^T q_{v^k}^k (u^k) \xi(0, t) \, dt & \geqslant 0. \end{split} \end{equation} (3.2)

    By an analogous argument we get

    \begin{align*} \int_{D^k} \int_0^\infty \big|u^k(x, t) - v^k(x, t)\big| \xi_t + q^k_{v^k}(u^k) \xi_x \, dt\, dx& \\ + \int_{D^k} \big|\bar{u}^k(x) - \bar{v}^k(x)\big| \, dx + \int_0^T q_{v^k}^k (u^k) \xi(0, t) \, dt& \geqslant 0 \end{align*}

    for k\in {{\mathscr{I}}_{\mathrm{out}}}. Fix s>0 , let r, \kappa>0 , and let \alpha_r: \mathbb{R}_-\to \mathbb{R} and \beta_\kappa\colon \mathbb{R}_+\to \mathbb{R} be given by

    \begin{align*} \alpha_r (x) & = \begin{cases} 0 & x\in(-\infty, -r-1] \\ x+r+1 & x\in(-r-1, -r) \\ 1 & x \in[-r, 0) \end{cases} \\ \beta_\kappa(t) & = \begin{cases} 1 & t \in [0, s] \\ \frac{1}{\kappa} (\kappa + s - t) & t \in ( s, s+\kappa) \\ 0 & t \in [s+\kappa, \infty). \end{cases} \end{align*}

    Via a standard regularization argument one can check that \varphi (x, t) = \alpha_r (x) \beta_\kappa (t) is an admissible test function. We compute the partial derivatives of \varphi :

    \varphi_t(x, t) = \begin{cases} 0 & t \in [ 0, s] \\ -\frac{1}{\kappa}\alpha_r (x) & t \in (s, s+\kappa] \\ 0 & t \in (s + \kappa, \infty) \end{cases}

    and

    \varphi_x(x, t) = \begin{cases} 0 & x\in(-\infty, -r-1) \\ \beta_\kappa (t) & x\in(-r-1, -r) \\ 0 & x\in(-r, 0). \end{cases}

    We insert this into (3.2) to get

    \begin{align*} -\frac{1}{\kappa} \int_s^{s+\kappa} \int_{-r-1}^0 \big|u^k(x, t) - v^k(x, t)\big| \alpha_r(x) \, dx \, dt& \\ + \int_0^{s+\kappa} \int_{-r-1}^{-r} q_{v^k}^k (u^k) \beta_\kappa(t) \, dx\, dt + \int_{-r-1}^0 \big|\bar{u}^k(x) - \bar{v}^k(x)\big| \alpha_r (x) \, dx& \\ - \int_0^{s+\kappa} q_{v^k}^k ( u^k (0, t)) \beta_\kappa (t) \, dt & \geqslant 0. \end{align*}

    Letting \kappa \to 0 and r\to\infty , we get

    \begin{equation*} \big\| u^k(x, t) - v^k(x, t) \big\|_{L^1 ( D^k)} \leqslant \big\| \bar{u}^k(x) - \bar{v}^k(x)\big\|_{L^1 ( D^k)} - \int_0^s q_{v^k}^k ( u^k(0, t)) \, dt. \end{equation*}

    An analogous inequality holds for k\in {{\mathscr{I}}_{\mathrm{out}}} . We sum over all edges to get

    \begin{align*} &\sum\limits_{k\in {\mathscr{I}}} \big\|u^k(x, t) - v^k(x, t)\big\|_{L^1(D^k)} \\ & \leqslant \sum\limits_{k\in {\mathscr{I}}} \int_{D^k} \big|\bar{u}^k(x) - \bar{v}^k(x)\big| \, dx +\underbrace{\int_0^s \sum\limits_{k\in {\mathscr{I}}} {\rm sgn}(k) q^k_{v^k}(u^k)}_{\text{$ \leqslant 0$ by (2.11) and Corollary 2.13}} \\ & \leqslant \sum\limits_{k\in {\mathscr{I}}} \big\|\bar{u}^k(x) - \bar{v}^k(x)\big\|_{L^1(D^k)}. \end{align*}

    In this section we construct a finite volume numerical approximation for (2.1) and prove stability and convergence properties of the method. The numerical method is rather standard for hyperbolic conservation laws, but an important feature of the method is that the vertex is discretized as a separate control volume. Although this control volume vanishes as the mesh parameter {\Delta x} is passed to zero, its presence will ensure that entropy is correctly dissipated at the vertex, even in the limit {\Delta x}\to0 . As opposed to the method presented in [1], where the problem is an implicit one on the vertex, our method is completely explicit.

    Let {\Delta t}, {\Delta x}>0 be given discretization parameters. We define the index sets

    \begin{align*} D_ \mathrm{disc}^+ : = \mathbb{N}, \qquad D_ \mathrm{disc}^- : = - \mathbb{N}, \qquad D_ \mathrm{disc}^k : = D_ \mathrm{disc}^{{\rm sgn}(k)}, \qquad D_ \mathrm{disc}^0 : = \{0\}. \end{align*}

    For n\in \mathbb{N}_0 we discretize1 t^n = n {\Delta t} , and for k\in {\mathscr{I}} and i\in \mathbb{Z} we let x_ {i+{1/2}} = ( {i+{1/2}}) {\Delta x} , and partition the physical domain into cells

    1In numerical experiments, the timestep {\Delta t} is chosen dynamically at each step n in order to comply with the CFL condition derived in Section 4. We use a uniform timestep for the sake of simplicity only.

    {\mathscr{C}}_i^k = D^k \cap \big(x_ {i-{1/2}}, \, x_ {i+{1/2}}\big).

    We define the mesh size at the vertex by {\Delta x}_0: = \sum_{k\in {\mathscr{I}}}| {\mathscr{C}}_0^k| = N {\Delta x}/2 , where |A| denotes the Lebesgue measure of A\subset \mathbb{R} . We make the finite volume approximation

    \begin{align*} u_i^{k, n} &\approx \frac{1}{ {\Delta x}}\int_{ {\mathscr{C}}_i^k}u^k(x, t^n)\, dx \qquad \text{for } i\in D^k_ \mathrm{disc}, \\ u_0^n &\approx \frac{1}{ {\Delta x}_0}\sum\limits_{k\in {\mathscr{I}}}\int_{ {\mathscr{C}}_0^k} u^k(x, t^n)\, dx. \end{align*}

    Fix some i \in D_ \mathrm{disc}^k , let \varphi^k(x, t) = \frac{1}{ {\Delta t} {\Delta x}} \mathbb{1}_{ {\mathscr{C}}_i^k}(x) \mathbb{1}_{[t^n, \, t^{n+1})}(t) and \varphi^l\equiv0 for l\neq k , and (after an approximation procedure) insert these into (2.4). We then obtain the numerical method

    \begin{equation} \frac{u_i^{k, n+1}-u_i^{k, n}}{ {\Delta t}} + \frac{F_ {i+{1/2}}^{k, n}-F_ {i-{1/2}}^{k, n}}{ {\Delta x}} = 0 \end{equation} (4.1a)

    where F_ {i+{1/2}}^{k, n} = F^k\big(u^{k, n}_i, u^{k, n}_{i+1}\big) is an approximation of the mean flux through x_ {i+{1/2}} over the time interval [t^n, t^{n+1}) ,

    F_ {i+{1/2}}^{k, n}\approx\frac{1}{ {\Delta t}}\int_{t^n}^{t^{n+1}}f^k\big(u^k(x_ {i+{1/2}}, t)\big)\, dt.

    For the special cell i = 0 we let \varphi^k(x, t) = \frac{1}{ {\Delta t} {\Delta x}_0} \mathbb{1}_{ {\mathscr{C}}_0^k}(x) \mathbb{1}_{[t^n, \, t^{n+1})}(t) for k\in {\mathscr{I}} to obtain

    \begin{equation} \frac{u_0^{n+1}-u_0^n}{ {\Delta t}} + \frac{1}{ {\Delta x}_0}\biggl(\sum\limits_{k\in {{\mathscr{I}}_{\mathrm{out}}}} F_ {1/2}^{k, n} - \sum\limits_{k\in {{\mathscr{I}}_{\mathrm{in}}}} F_{- {1/2}}^{k, n} \biggr) = 0. \end{equation} (4.1b)

    (This is opposed to the explicit method of Towers [29] where the vertex is modelled as having zero width for any {\Delta x}>0 .) We will use the notational convention that u_0^{k, n} \equiv u_0^n for all k\in {\mathscr{I}}. (We postpone the definition of the initial data u_i^{k, 0} until Section 4.3.)

    Given a numerically computed solution ( {\mathbf{u}}_i^n)_{i, n} , we define the piecewise constant function

    \begin{equation} {\mathbf{u}}_ {\Delta t}(x, k, t) = u_i^{k, n} \qquad \text{for } x\in {\mathscr{C}}_i^k, \ t\in[t^n, t^{n+1}). \end{equation} (4.2)

    We remark that the integral of {\mathbf{u}}_ {\Delta t} w.r.t. the measure \lambda (cf. (2.2)) can be written

    \begin{equation} \int_\Omega {\mathbf{u}}_ {\Delta t}(\cdot, t)\, d\lambda = \sum\limits_{k\in {\mathscr{I}}}\sum\limits_{i\in D_ \mathrm{disc}^k} u_i^{k, n}\, {\Delta x} + u_0^n\, {\Delta x}_0 \end{equation} (4.3)

    for any t\in[t^n, t^{n+1}) , and the total variation of {\mathbf{u}}_ {\Delta t} (cf. (2.3)) can be written

    \begin{equation} \begin{split} {\rm TV}( {\mathbf{u}}_ {\Delta t}(\cdot, t)) = & \sum\limits_{k\in {{\mathscr{I}}_{\mathrm{in}}}}\sum\limits_{i \in D_ \mathrm{disc}^k}\big|u_{i+1}^{k, n}-u_i^{k, n}\big| + \sum\limits_{k\in {{\mathscr{I}}_{\mathrm{out}}}}\sum\limits_{i \in D_ \mathrm{disc}^k}\big|u_{i}^{k, n}-u_{i-1}^{k, n}\big| \\ = & \sum\limits_{k\in {{\mathscr{I}}_{\mathrm{in}}}}\sum\limits_{i \in D_ \mathrm{disc}^k}\big|u_{i}^{k, n}-u_{i-1}^{k, n}\big| + \sum\limits_{k\in {{\mathscr{I}}_{\mathrm{out}}}}\sum\limits_{i \in D_ \mathrm{disc}^k}\big|u_{i+1}^{k, n}-u_{i}^{k, n}\big| \\ &+ \sum\limits_{k\in {{\mathscr{I}}_{\mathrm{in}}}}\big|u_0^{n}-u_{-1}^{k, n}\big| + \sum\limits_{k\in {{\mathscr{I}}_{\mathrm{out}}}}\big|u_0^{n}-u_{1}^{k, n}\big|. \end{split} \end{equation} (4.4)

    Note also that a numerical method of the form (4.1) is conservative in the sense that the total mass \sum_{k\in {\mathscr{I}}} \int_\Omega {\mathbf{u}}_ {\Delta t} \, d\lambda is independent of n :

    \begin{align*} \int_\Omega {\mathbf{u}}_ {\Delta t}(\cdot, t^{n+1}) \, d\lambda = {}& \sum\limits_{k\in {\mathscr{I}}} \sum\limits_{i \in D_ \mathrm{disc}^k} u_i^{k, n+1} {\Delta x} + u_0^{n+1} {\Delta x}_0 \\ = {}&\sum\limits_{k\in {\mathscr{I}}} \sum\limits_{i \in D_ \mathrm{disc}^k} u_i^{k, n} {\Delta x} - {\Delta t} \big(F_ {i+{1/2}}^{k, n} - F_ {i-{1/2}}^{k, n}\big) + u_0^n {\Delta x}_0 \\ &{{}- {\Delta t}} \biggl( \sum\limits_{k\in {{\mathscr{I}}_{\mathrm{out}}}} F_{ {1/2}}^k - \sum\limits_{k\in {{\mathscr{I}}_{\mathrm{in}}}} F_{- {1/2}}^k \biggr) \\ = {}& \sum\limits_{k\in {\mathscr{I}}} \sum\limits_{i\in D_ \mathrm{disc}^k} u_i^{k, n} {\Delta x} + u_0^n {\Delta x}_0 \\ = {}& \int_\Omega {\mathbf{u}}_ {\Delta t}(\cdot, t^{n}) \, d\lambda. \end{align*}

    As a shorthand for the scheme (4.1) we define the functions

    \begin{equation} G^k \big( u_{i-1}^k, u_i^k, u_{i+1}^k \bigr) : = u_i^{k} - \frac{ {\Delta t}}{ {\Delta x}}\big(F^k\bigl(u_i^k, u_{i+1}^k\bigr) - F^k \bigl( u_{i-1}^k, u_i^k\bigr)\big) \end{equation} (4.5a)

    for k\in {\mathscr{I}} and

    \begin{equation} \begin{split} &G^0\bigl( u_{-1}^{- {N_{\mathrm{in}}}}, \dots, u_{-1}^{-1}, u_0, u_1^1, \dots, u_1^{ {N_{\mathrm{out}}}}\bigr) \\ &\quad: = u_0 - \frac{ {\Delta t}}{ {\Delta x}_0} \biggl(\sum\limits_{k\in {{\mathscr{I}}_{\mathrm{out}}}} F^k\bigl( u_0, u_1^k\bigr) - \sum\limits_{k\in {{\mathscr{I}}_{\mathrm{in}}}} F^k\bigl(u_{-1}^k, u_0\bigr)\biggr), \end{split} \end{equation} (4.5b)

    enabling us to write (4.1) in the update form

    \begin{equation} \begin{split} u_i^{k, n+1} & = G^k \big( u_{i-1}^{k, n}, u_i^{k, n}, u_{i+1}^{k, n} \bigr) \qquad\qquad \text{for } i\in D_ \mathrm{disc}^k, \ k\in {\mathscr{I}} \\ u_0^{n+1} & = G^0\bigl( u_{-1}^{- {N_{\mathrm{in}}}, n}, \dots, u_{-1}^{-1, n}, u_0^n, u_1^{1, n}, \dots, u_1^{ {N_{\mathrm{out}}}, n}\bigr). \end{split} \end{equation} (4.6)

    As a shorthand for (4.6), we will sometimes use the notation

    \begin{equation} u_i^{k, n+1} = G^k\big( {\mathbf{u}}_{i-1}^n, {\mathbf{u}}_i^n, {\mathbf{u}}_{i+1}^n\big) \qquad \text{for }i\in D_ \mathrm{disc}^k, \ k\in {\mathscr{I}}_0, \end{equation} (4.6')

    where {\mathbf{u}}_i^n is the vector containing all numerical values at index i at time n .

    Definition 4.1 (Monotone scheme). The difference scheme (4.6') is monotone if

    {\mathbf{u}}^n \leqslant {\mathbf{v}}^n \quad \Rightarrow \quad {\mathbf{u}}^{n+1} \leqslant {\mathbf{v}}^{n+1},

    where {\mathbf{u}}^n \leqslant {\mathbf{v}}^n means that every component u_i^{k, n} of {\mathbf{u}}^n is no greater than the corresponding component of {\mathbf{v}}^n .

    We state a straightforward CFL-type condition which ensures monotonicity of the numerical scheme.

    Proposition 4.2. Consider a consistent finite volume method (4.1), where F^k is nondecreasing in the first variable and nonincreasing in the second. Then the method is monotone under the CFL condition

    \begin{equation} {\Delta t}\max\limits_{k, u, v}\Bigl|\frac{\partial F^k}{\partial u}(u, v)\Bigr| \leqslant {\Delta x}/2, \qquad {\Delta t}\max\limits_{k, u, v}\Bigl|\frac{\partial F^k}{\partial v}(u, v)\Bigr| \leqslant {\Delta x}/2. \end{equation} (4.7)

    Proof. We can calculate the derivatives to the update functions to get

    \begin{align*} \frac{\partial G^k}{\partial u_{i-1}^{k}} & = \frac{ {\Delta t}}{ {\Delta x}} \frac{\partial F_ {i-{1/2}}^k}{\partial u_{i-1}^{k}}, \qquad \frac{\partial G^k}{\partial u_{i+1}^{k}} = -\frac{ {\Delta t}}{ {\Delta x}} \frac{\partial F_ {i+{1/2}}^k}{\partial u_{i+1}^{k}} \\ \frac{\partial G^k}{\partial u_{i}^{k}} & = 1 - \frac{ {\Delta t}}{ {\Delta x}} \biggl( \frac{\partial F_ {i+{1/2}}^k}{\partial u_{i}^{k}} - \frac{\partial F_ {i-{1/2}}^k}{\partial u_{i}^{k}} \biggr), \end{align*}

    for each k\in {\mathscr{I}} , and

    \begin{align*} \frac{\partial G^0}{\partial u_{-1}^{k}} & = \frac{ {\Delta t}}{ {\Delta x}_0} \frac{\partial F_{- {1/2}}^k}{\partial u_{-1}^{k}} && \text{for } k\in {{\mathscr{I}}_{\mathrm{in}}}, \\ \frac{\partial G^0}{\partial u_{1}^{k}} & = -\frac{ {\Delta t}}{ {\Delta x}_0} \frac{\partial F_{ {1/2}}^k}{\partial u_{1}^{k}} && \text{for } k\in {{\mathscr{I}}_{\mathrm{out}}}, \\ \frac{\partial G^0}{\partial u_0} & = 1 - \frac{ {\Delta t}}{ {\Delta x}_0} \biggl(\sum\limits_{k\in {{\mathscr{I}}_{\mathrm{out}}}}\frac{\partial F_{ {1/2}}^k}{\partial u_0^n} - \sum\limits_{k\in {{\mathscr{I}}_{\mathrm{in}}}}\frac{\partial F_{- {1/2}}^k}{\partial u_0^n} \biggr) \end{align*}

    on the vertex. We would like these derivatives to be non-negative. The monotonicity of F^k guarantees that the first, second, fourth and fifth expressions are non-negative. Applying monotonicity of F^k to the third and sixth terms, we get

    \frac{\partial G^k}{\partial u_{i}^{k}} = \frac{ {\Delta t}}{ {\Delta x}} \biggl(\frac{ {\Delta x}}{ {\Delta t}} - \Bigl|\frac{\partial F_ {i+{1/2}}^k}{\partial u_{i}^{k}}\Bigr| - \Bigl|\frac{\partial F_ {i-{1/2}}^k}{\partial u_{i}^{k}} \Bigr|\biggr) \geqslant 0

    (by (4.7)) and

    \begin{align*} \frac{\partial G^0}{\partial u_0} & = \frac{ {\Delta t}}{ {\Delta x}_0}\biggl(\frac{ {\Delta x}_0}{ {\Delta t}} - \sum\limits_{k\in {{\mathscr{I}}_{\mathrm{out}}}}\Bigl|\frac{\partial F_{ {1/2}}^k}{\partial u_0^n}\Bigr| - \sum\limits_{k\in {{\mathscr{I}}_{\mathrm{in}}}}\Bigl|\frac{\partial F_{- {1/2}}^k}{\partial u_0^n}\Bigr|\biggr) \end{align*}

    (using {\Delta x}_0 = N {\Delta x}/2)

    \begin{align*} & \geqslant \frac{ {\Delta t}}{ {\Delta x}_0}\biggl(N\frac{ {\Delta x}}{2 {\Delta t}} - {N_{\mathrm{out}}}\max\limits_{k, u, v}\Bigl|\frac{\partial F^k}{\partial u}(u, v)\Bigr| - {N_{\mathrm{in}}}\max\limits_{k, u, v}\Bigl|\frac{\partial F^k}{\partial v}(u, v)\Bigr|\biggr)\\ & \geqslant 0 \end{align*}

    by (4.7).

    Remark 4.3. As opposed to the explicit method that Towers proposes in [29], where the CFL condition gets more restrictive as the number of roads grows, we don't face any issues with the time step with the allowable time step with a high number of roads.

    In the same way that stationary solutions are essential for the well-posedness of entropy solutions (cf. Section 3), they are essential to the stability and convergence of numerical methods on networks. Asserting that a numerical solution is both constant in time and on each edge yields the following definition.

    Definition 4.4 (Discrete Stationary Solution). Consider a consistent, conservative numerical method (4.1). A discrete stationary solution for (4.1) is a vector

    {\mathbf{c}}^ \mathrm{disc} : = ( c^{- {N_{\mathrm{in}}}}, \dots, c^{ {N_{\mathrm{out}}}}) \in \mathbb{R}^{N+1}

    satisfying the Rankine–Hugoniot condition

    \begin{equation} \sum\limits_{k\in {{\mathscr{I}}_{\mathrm{in}}}} f^k ( c^k ) = \sum\limits_{k\in {{\mathscr{I}}_{\mathrm{out}}}} f^k ( c^k ) \end{equation} (4.8)

    as well as the conditions

    \begin{align} F^k ( c^k, c^0 ) & = f^k ( c^k ) \qquad \text{for } k\in {{\mathscr{I}}_{\mathrm{in}}}, \end{align} (4.9a)
    \begin{align} F^k ( c^0, c^k ) & = f^k ( c^k ) \qquad \text{for }k\in {{\mathscr{I}}_{\mathrm{out}}}. \end{align} (4.9b)

    In the remainder, sets of discrete stationary solutions will be denoted with a superscript, \mathscr{G}^0 , to signal that they also include a value at the vertex i = 0 .

    Remark 4.5. Note that our definition of a discrete stationary solution is analogous to [1,Definition 2.1]. As opposed to our definition, the authors of [1] only include values on the edges. The value c_0 , which is called p in [1], is excluded from the vectors of stationary solutions there.

    Notation 4.6. We will sometimes index a discrete stationary solution as

    \begin{equation} {\mathbf{c}}_i = \begin{cases} \big(c^{- {N_{\mathrm{in}}}}, \dots, c^{-1}\big) & i < 0 \\ c^0 & i = 0 \\ \big(c^1, \dots, c^ {N_{\mathrm{out}}}\big) & i > 0 \end{cases} \end{equation} (4.10a)

    for i\in \mathbb{Z} and, by extension,

    \begin{equation} c_i^k = \begin{cases} c^k & i\neq 0 \\ c^0 & i = 0. \end{cases} \end{equation} (4.10b)

    Using the notation (4.6'), it is readily checked that discrete stationary solutions are precisely those that are constant on each edge and satisfy

    {\mathbf{c}}_i = G^k( {\mathbf{c}}_{i-1}, {\mathbf{c}}_i, {\mathbf{c}}_{i+1}) \qquad \forall\ i\in D_ \mathrm{disc}^k, \ k\in {\mathscr{I}}_0.

    Remark 4.7. The conditions (4.9) say that the numerical fluxes at the vertex reduce to the upwind flux on the in edges and the downwind flux on the out edges. This can be interpreted as information only flowing into the vertex, not out of it. This is consistent with the interpretation of the vertex as a stationary shock.

    Remark 4.8. Discrete stationary solutions {\mathbf{c}} = \big(c^{- {N_{\mathrm{in}}}}, \dots, c^{ {N_{\mathrm{out}}}}\big) fulfil a discrete version of the Rankine–Hugoniot type condition (2.7),

    \sum\limits_{k\in {{\mathscr{I}}_{\mathrm{in}}}} F^k\big( c^k, c^0\big) = \sum\limits_{k\in {{\mathscr{I}}_{\mathrm{out}}}} F^k\big( c^0, c^k\big).

    Lemma 4.9. Consider a consistent, conservative numerical scheme (4.1). Let {\mathbf{c}} = (c^k)_{k\in {\mathscr{I}}} be a stationary solution for (1.1) and let c^0\in \mathbb{R} . Then the vector {\mathbf{c}}_ \mathrm{disc} = (c^{- {N_{\mathrm{in}}}}, \dots, c^{-1}, c^0, c^1, \dots, c^{ {N_{\mathrm{out}}}})\in \mathbb{R}^{N+1} is a discrete stationary solution if and only if

    c^0 \in \bigcap\limits_{k\in {{\mathscr{I}}_{\mathrm{in}}}} ( H^k )^{-1} (\{f^k(c^k)\}) \mathrel{\bigcap}\bigcap\limits_{k\in {{\mathscr{I}}_{\mathrm{out}}}} ( J^k )^{-1} (\{f^k(c^k)\})

    where

    \begin{align*} H^k(c) &: = F^k(c^k, c) \quad \mathit{\text{for}}\; k\in {{\mathscr{I}}_{\mathrm{in}}}, \qquad J^k(c) : = F^k(c, c^k) \quad \mathit{\text{for}}\;k\in {{\mathscr{I}}_{\mathrm{out}}}. \end{align*}

    Proof. We can rewrite conditions (4.9a) and (4.9b) as

    \begin{align*} &(4.9{\rm{b}}) \quad \Leftrightarrow \quad H^k(c^0) = f^k(c^k) \quad \Leftrightarrow \quad c^0 \in ( H^k )^{-1} (\{f^k (c^k)\}) \end{align*}

    for k\in {{\mathscr{I}}_{\mathrm{in}}}, and

    \begin{align*} &(4.9{\rm{a}}) \quad \Leftrightarrow \quad J^k(c^0) = f^k(c^k) \quad \Leftrightarrow \quad c^0 \in ( J^k )^{-1} (\{f^k (c^k)\}) \end{align*}

    for k\in {{\mathscr{I}}_{\mathrm{out}}} . Hence, if (4.9a), (4.9b) are satisfied then c^0 must lie in all of the sets on the right hand side, and hence in their intersection. Conversely, if c^0 lies in the intersection, then (4.9a), (4.9b) are satisfied.

    We set out to prove an L^\infty bound, L^1 contractiveness and Lipschitz continuity in time for solutions computed with a general consistent, conservative, monotone finite volume method on a network. Our starting point will be a class of discrete stationary solutions \mathscr{G}_ \mathrm{disc}^0\subset \mathbb{R}^{N+1} for a conservative finite volume method (4.1). We take initial data \bar u\in L^\infty_{\mathrm{oco}}( \mathscr{G}_ \mathrm{disc}^0) (cf. (2.9)), we let {\mathbf{c}}\in \mathscr{G}_ \mathrm{disc}^0 be as specified in (2.9), and consider the finite volume method (4.1) initialized by

    \begin{equation} u_i^{k, 0} = \frac{1}{ {\Delta x}}\int_{ {\mathscr{C}}_i^k}\bar u^k(x)\, dx, \qquad u_0^0 = c^0. \end{equation} (4.11)

    (The value c^0 is chosen for convenience, and any value in [c^0, d^0] will have the desired effect.)

    Lemma 4.10. Consider monotone numerical flux functions F^k ( k\in {\mathscr{I}} ). Let {\mathbf{c}}, {\mathbf{d}} be discrete stationary solutions satisfying c^k \leqslant d^l for all k, l\in {\mathscr{I}} (cf. Definition 2.7). Then c^0, d^0 can be modified such that {\mathbf{c}}, {\mathbf{d}} remain discrete stationary solutions and such that c^0 \leqslant d^0 .

    Proof. Define

    I^k(c^k) : = \begin{cases} \big(F^k(c^k, \cdot)\big)^{-1} \big(\{ f^k(c^k) \}\big) & \text{for } k \in {{\mathscr{I}}_{\mathrm{in}}} \\ \big(F^k(\cdot, c^k)\big)^{-1}\big(\{ f^k (c^k) \}\big) & \text{for } k \in {{\mathscr{I}}_{\mathrm{out}}}. \end{cases}

    Since all F^k are monotone, each I^k(c^k) is a connected interval which contains c^k , and moreover, Lemma 4.9 says that c^0 \in \bigcap_{k\in {\mathscr{I}}} I^k(c^k) . This implies that [\![{c^0}, {c^k}]\!] \subset I^k(c^k) , where [\![{a}, {b}]\!] = [\min(a, b), \, \max(a, b)] . Since \bigcap_{k\in {\mathscr{I}}}[\![{c^0}, {c^k}]\!] is nonempty, the number

    \tilde{c}^0 : = \min \biggl(\bigcap\limits_{k\in {\mathscr{I}}}[\![{c^0}, {c^k}]\!]\biggr)

    exists and satisfies \tilde{c}^0 \leqslant \max_{k\in {\mathscr{I}}} c^k . Appealing again to Lemma 4.9, {\mathbf{c}} remains a discrete stationary solution if c^0 is replaced by \tilde{c}^0 . In a similar way we replace d^0 by

    \tilde{d}^0 : = \max \biggl(\bigcap\limits_{k\in {\mathscr{I}}}[\![{d^0}, {d^k}]\!]\biggr),

    which satisfies \tilde{d}^0 \geqslant \min_{k\in {\mathscr{I}}} d^k . By our hypothesis, it follows that \tilde c^0 \leqslant \tilde{d}^0 .

    Proposition 4.11. Consider a consistent, conservative, monotone finite volume method (4.1), (4.11) with a set of discrete stationary states \mathscr{G}_ \mathrm{disc}^0 . For any initial data \bar {\mathbf{u}} \in L^\infty_{\mathrm{oco}}( \mathscr{G}_ \mathrm{disc}^0) , the numerical solution is uniformly L^\infty bounded.

    Proof. Pick discrete stationary states {\mathbf{c}}, {\mathbf{d}}\in \mathscr{G}_ \mathrm{disc}^0 as in (2.9). It is clear that the initial data defined in (4.11) satisfy c^k \leqslant u_i^{k, 0} \leqslant d^k for all i\in D_ \mathrm{disc}^k and k\in {\mathscr{I}}_0 . If the same holds at some time step n\in \mathbb{N}_0 then (using the notation (4.6'), (4.10))

    u_i^{k, n+1} = G^k \bigl( {\mathbf{u}}_{i-1}^{n}, {\mathbf{u}}_i^{n}, {\mathbf{u}}_{i+1}^{n} \bigr) \geqslant G^k\bigl( {\mathbf{c}}_{i-1}, {\mathbf{c}}_i, {\mathbf{c}}_{i+1}\bigr) = c_i^k

    for all i\in D_ \mathrm{disc}^k and k\in {\mathscr{I}}_0 , and similarly, u_i^{k, n+1} \leqslant d_i^k .

    Definition 4.12 (L^1 contractive method). A numerical method (4.6') is L^1 contractive if

    \begin{align*} \bigl\| {\mathbf{u}}_ {\Delta t} ( \cdot, t ) - {\mathbf{v}}_ {\Delta t} ( \cdot, t ) \bigr\|_{L^1(\Omega;\lambda)} \leqslant \|\bar {\mathbf{u}} - \bar {\mathbf{v}}\|_{L^1(\Omega;\lambda)} \end{align*}

    for all t \geqslant 0 , where {\mathbf{u}}_ {\Delta t} and {\mathbf{v}}_ {\Delta t} are the projection of the numerical solution (cf. (4.2)) computed with initial data \bar {\mathbf{u}}, \bar {\mathbf{v}}\in L^\infty_{\mathrm{oco}}( \mathscr{G}_ \mathrm{disc}^0)\cap L^1(\Omega;\lambda) , respectively. (See (4.3) for the integral of {\mathbf{u}}_ {\Delta t} , {\mathbf{v}}_ {\Delta t} w.r.t. \lambda .)

    We state the well known Crandall–Tartar lemma which we will use in the following proof. Here and below, we use the notation a\vee b = \max(a, b) .

    Theorem 4.13. (Crandall–Tartar: [11,Proposition 1]). Let (\Omega, \lambda) be a measure space. Let C \subset L^1 (\Omega; \lambda) have the property that f, g \in C implies f \vee g \in C . Let V\colon C \to L^1(\Omega; \lambda) satisfy \int_\Omega V (f)\, d\lambda = \int_\Omega f\, d\lambda for f \in C . Then the following three properties of V are equivalent:

    (a) f, g \in C and f \leqslant g a.e. implies V(f) \leqslant V(g) a.e.,

    (b) \int_\Omega (V(f) - V(g))_+ \leqslant \int_\Omega (f - g)_+ for f, g \in C ,

    (c) \int_\Omega \big|V(f) - V(g)\big| \leqslant \int_\Omega |f - g| for f, g \in C .

    We can now prove L^1 -contractivity of our method.

    Theorem 4.14. Every conservative, consistent monotone method (4.1), (4.11) is L^1-contractive.

    Proof. Let C = {C_ {\Delta x}} be the set of piecewise constant functions,

    {C_ {\Delta x}} = \Bigl\{ {\mathbf{u}}\in L^1\cap L^\infty(\Omega;\lambda)\ :\ {\mathbf{u}}(x) = \sum\limits_{k\in {\mathscr{I}}_0}\sum\limits_{i\in D_ \mathrm{disc}^k}u_i^k \mathbb{1}_{ {\mathscr{C}}_i^k} \text{ for } u_i^k\in \mathbb{R} \Bigr\}.

    We define the operator V: {C_ {\Delta x}}\to {C_ {\Delta x}} mapping a numerical solution to the next time step,

    \begin{align*} V ( {\mathbf{u}} ) : = {}& \sum\limits_{k\in {\mathscr{I}}} \sum\limits_{i\in D_ \mathrm{disc}^k} \mathbb{1}_{ {\mathscr{C}}_i^k} \Bigl( u_i^{k} - \frac{ {\Delta t}}{ {\Delta x}} \big(F^k\big(u_i^k, u_{i+1}^k\big) - F^k\big(u_{i-1}^k, u_i^k\big)\big) \Bigr) \\ &+ \sum\limits_{k\in {\mathscr{I}}} \mathbb{1}_{ {\mathscr{C}}_0^k} \biggl(u_0^{0} - \frac{ {\Delta t}}{ {\Delta x}_0} \biggl( \sum\limits_{k\in {{\mathscr{I}}_{\mathrm{out}}}} F^k(u_0, u_1^k) - \sum\limits_{k\in {{\mathscr{I}}_{\mathrm{in}}}} F^k(u_{-1}^k, u_0) \biggr) \biggr). \end{align*}

    By the definition (2.2) of the measure \lambda (cf. also (4.3)), we have \int_{\Omega} V ({\mathbf{u}}) \, d\lambda = \int_{\Omega} {\mathbf{u}} \, d\lambda for all {\mathbf{u}}\in {C_ {\Delta x}} . We apply the Crandall–Tartar lemma to conclude L^1 -contractivity of the numerical solution operator V .

    From L^1 -contractivity we get continuity in time as a corollary:

    Corollary 4.15. Consider a consistent, conservative and monotone method (4.1). Let {\mathbf{u}}_ {\Delta t} be an approximate solution computed with this method and assume that all numerical fluxes F^k are Lipschitz continuous in both arguments. Then computed solutions are uniformly L^1 Lipschitz continuous in time, i.e.,

    \begin{align*} \big\| {\mathbf{u}}_ {\Delta t}(t^{n+1}) - {\mathbf{u}}_ {\Delta t}(t^n) \big\|_{L^1(\Omega;\lambda)} & \leqslant \big\| {\mathbf{u}}_ {\Delta t}(t^1) - {\mathbf{u}}_ {\Delta t}(t^0) \big\|_{L^1(\Omega;\lambda)} \\ & \leqslant {\Delta t} \big(C{\rm TV}( {\mathbf{u}}^0) + \bar M\big), \end{align*}

    where the constants C and \bar{M} depend on the flux functions and on the initial data.

    Proof. We compute

    \begin{align*} \bigl\|& {\mathbf{u}}_ {\Delta t}(t^{n+1}) - {\mathbf{u}}_ {\Delta t}(t^n) \bigr\|_{L^1(\Omega;\lambda)} \\ & = \big\|V( {\mathbf{u}}_ {\Delta t}(t^{n})) - V( {\mathbf{u}}_ {\Delta t}(t^{n-1})) \big\|_{L^1(\Omega;\lambda)} \end{align*}

    (using Theorem 4.13(c))

    \begin{align*} & \leqslant \big\| {\mathbf{u}}_ {\Delta t}(t^{n}) - {\mathbf{u}}_ {\Delta t}(t^{n-1}) \big\|_{L^1(\Omega;\lambda)} \leqslant \dots \leqslant \big\| {\mathbf{u}}_ {\Delta t}(t^1) - {\mathbf{u}}_ {\Delta t}(t^0) \big\|_{L^1(\Omega;\lambda)} \\ & = {\Delta x} \sum\limits_{k\in {\mathscr{I}}} \sum\limits_{i \in D_ \mathrm{disc}^k} | u_i^{k, 1} - u_i^0| + {\Delta x}_0|u_0^{k, 1} - u_0^0| \\ & = {\Delta t} \sum\limits_{k\in {\mathscr{I}}} \sum\limits_{i \in D_ \mathrm{disc}^k} \big|F_ {i+{1/2}}^{k, 0} - F_ {i-{1/2}}^{k, 0}\big| + {\Delta t} \biggl|\sum\limits_{k\in {{\mathscr{I}}_{\mathrm{out}}}} F_ {1/2}^{k, 0} - \sum\limits_{k\in {{\mathscr{I}}_{\mathrm{in}}}} F_{- {1/2}}^{k, 0}\biggr| \\ & = {\Delta x} \lambda \sum\limits_{k\in {\mathscr{I}}} \sum\limits_{i\in D_ \mathrm{disc}^k}\big|u_i^{k, 0} - u_{i-1}^{k, 0}\big| \\ &\quad+ {\Delta t} \biggl|\!\begin{aligned}[t]&\sum\limits_{k\in {{\mathscr{I}}_{\mathrm{out}}}} F_ {1/2}^{k, 0} - F^{k, 0} \big(u_0^0, u_0^0\big) - \sum\limits_{k\in {{\mathscr{I}}_{\mathrm{in}}}} F_{- {1/2}}^{k, 0} - F^{k, 0} (u_0^0, u_0^0) \\ &+\overbrace{\sum\limits_{k\in {{\mathscr{I}}_{\mathrm{out}}}} f^k \big(u_0^0\big)}^{ = : f_ {\mathrm{out}}(u_0^0)} - \overbrace{\sum\limits_{k\in {{\mathscr{I}}_{\mathrm{in}}}} f^k\big(u_0^0\big)}^{ = : f_ {\mathrm{in}}(u_0^0)}\biggr|\end{aligned} \end{align*}
    \begin{align*} & \leqslant {\Delta t} \sum\limits_{k\in {\mathscr{I}}} \sum\limits_{i\in D_ \mathrm{disc}^k} L^k \bigl(\big|u_i^{k, 0} - u_{i-1}^{k, 0}\big| + \big|u_{i+1}^{k, 0} - u_i^{k, 0}\big| \bigr) \\ &\quad + {\Delta t} \underbrace{\Biggl(\sum\limits_{k\in {{\mathscr{I}}_{\mathrm{out}}}} L^k \big|u_1^{k, 0} - u_0^0\big| + \sum\limits_{k\in {{\mathscr{I}}_{\mathrm{in}}}} L^k \big|u_0^0 - u_{-1}^{k, 0}\big| + \overbrace{\big|f_ {\mathrm{out}}\big(u_0^0\big) - f_ {\mathrm{in}}\big(u_0^0\big)\big|}^{ \leqslant M}\Biggr)}_{ = :\bar M} \\ & \leqslant {\Delta t} (C {\rm TV}( {\mathbf{u}}^0) + \bar M ), \end{align*}

    where we collect all constants into the global constant C . We can bound \big|f_ {\mathrm{out}} \big(u_0^0\big) - f_ {\mathrm{in}} \big(u_0^0\big)\big| \leqslant \bar M with a constant \bar M \in \mathbb{R} since f_ {\mathrm{in}}, f_ {\mathrm{out}} are continuous and u_0^0 \in L^\infty .

    We are now in place to prove convergence in the case where the flux functions f^k are strictly monotone. We do this by using the upwind method where the numerical flux functions are defined by

    F^k(u, v) = \begin{cases} f^k(u) & \text{if $f^k$ is increasing, } \\ f^k(v) & \text{if $f^k$ is decreasing.} \end{cases}

    We shall show that the set of discrete approximations is compact in L^\infty\big([0, \infty);L^1(\Omega; \lambda)\big) , and that any limit is an entropy solution. In particular, this convergence result establishes existence of an entropy solution. We show convergence to a weak solution by proving a Lax–Wendroff type theorem:

    Theorem 5.1 (Lax–Wendroff theorem). Fix T > 0 . Assume that each flux function f^k is locally Lipschitz continuous and strictly monotone. Let \mathscr{G}_ \mathrm{disc}^0 be a mutually consistent class of discrete stationary solutions for the upwind method and let {\mathbf{u}}_ {\Delta t} be computed from the upwind method with initial data \bar {\mathbf{u}} \in L^\infty_{\mathrm{oco}}( \mathscr{G}_ \mathrm{disc}^0) \cap L^1(\Omega;\lambda) . Consider a subsequence \big({\mathbf{u}}_{ {\Delta t}_{\ell}}\big)_{\ell\in \mathbb{N}} such that {\Delta t}_{\ell} \to 0 and {\mathbf{u}}_{ {\Delta t}_{\ell}}\to {\mathbf{u}} in L^\infty([0, T]; L^1(\Omega; \lambda)) as k\to\infty . Then the limit {\mathbf{u}} is the unique entropy solution to (2.1), that is, {\mathbf{u}} satisfies (2.10).

    Remark 5.2. The existence of a non-trivial mutually consistent germ \mathscr{G}_ \mathrm{disc}^0 for monotone flux functions will be shown in 6.

    Proof. We write {\Delta x} and {\Delta t} rather than {\Delta x}_{\ell} and {\Delta t}_{\ell} , and we shall show that u satisfies the entropy condition (2.10) for every {\mathbf{c}}\in \mathscr{G}_ \mathrm{disc}^0 . Choosing stationary solutions {\mathbf{c}}, {\mathbf{d}}\in \mathscr{G}_ \mathrm{disc}^0 such that {\mathbf{c}} \leqslant {\mathbf{u}}_ {\Delta t} \leqslant {\mathbf{d}} (cf. Proposition 4.11) in particular shows that u is a weak solution.

    Let {\mathbf{c}}\in \mathscr{G}_ \mathrm{disc}^0 and consider the Crandall–Majda numerical entropy fluxes

    \begin{align*} Q_ {i+{1/2}}^{k, n} = F^k \bigl( u_i^{k, n} \vee c^k_i, u_{i+1}^{k, n} \vee c^k_i \bigr) - F^k \bigl( u_i^{k, n} \wedge c^k, u_{i+1}^{k, n} \wedge c^k \bigr) \end{align*}

    for i = 0, 1, \dots when k\in {{\mathscr{I}}_{\mathrm{out}}} , and for i = \dots, -2, -1 when k\in {{\mathscr{I}}_{\mathrm{in}}} , and

    \begin{align*} Q_{- {1/2}}^n = \sum\limits_{k\in {{\mathscr{I}}_{\mathrm{in}}}} Q_{- {1/2}}^{k, n}, \qquad Q_{ {1/2}}^n = \sum\limits_{k\in {{\mathscr{I}}_{\mathrm{out}}}} Q_{ {1/2}}^{k, n} \end{align*}

    (cf. Notation 4.6 for the definition of c_i^k ). Recalling the definition (4.5) of the update functions G^k , we see that

    \begin{align*} & G^k \bigl( u_{i-1}^{k, n} \vee c^k, u_i^{k, n} \vee c^k , u_{i + 1}^{k, n} \vee c^k \bigr) - G^k \bigl( u_{i-1}^{k, n} \wedge c^k, u_i^{k, n} \wedge c^k , u_{i + 1}^{k, n} \wedge c^k \bigr)\\ & \qquad = \bigl|u_i^{k, n} - c^k\bigr| - \frac{ {\Delta t}}{ {\Delta x}} \bigl( Q_ {i+{1/2}}^{k, n} - Q_ {i-{1/2}}^{k, n} \bigr), \end{align*}

    for k\in {\mathscr{I}} and i \in D_ \mathrm{disc}^k . Hence,

    \begin{equation} \begin{split} \big|u_i^{k, n+1}-c^k\big| & = u_i^{k, n+1}\vee c^k - u_i^{k, n+1}\wedge c^k \\ & = G^k \bigl(u_{i-1}^{k, n}, u_i^{k, n}, u_{i + 1}^{k, n}\bigr)\vee c^k - G^k \bigl(u_{i-1}^{k, n}, u_i^{k, n}, u_{i + 1}^{k, n}\bigr)\wedge c^k \\ & \leqslant G^k \bigl( u_{i-1}^{k, n} \vee c^k, u_i^{k, n} \vee c^k , u_{i + 1}^{k, n} \vee c^k \bigr) \\ & \quad- G^k \bigl( u_{i-1}^{k, n} \wedge c^k, u_i^{k, n} \wedge c^k , u_{i + 1}^{k, n} \wedge c^k \bigr) \\ & = \bigl|u_i^{k, n} - c^k\bigr| - \frac{ {\Delta t}}{ {\Delta x}} \bigl( Q_ {i+{1/2}}^{k, n} - Q_ {i-{1/2}}^{k, n} \bigr). \end{split} \end{equation} (5.1)

    Similarly, we find that

    \begin{equation} \bigl|u_0^{n+1} - c^0\bigr| - \bigl|u_0^n - c^0\bigr| + \frac{ {\Delta t}}{ {\Delta x}_0} \bigl( Q_ {1/2}^n - Q_{- {1/2}}^n \bigr) \leqslant 0 \end{equation} (5.2)

    We choose T = M {\Delta t} for a natural number M , multiply the above N+1 inequalities with a test function \varphi and sum up to get

    \begin{align*} 0& \geqslant\sum\limits_{n = 0}^{M}\sum\limits_{k\in {\mathscr{I}}} \sum\limits_{i \in D_ \mathrm{disc}^k} \varphi_i^{k, n}\biggl( \bigl( \big|u_i^{k, n+1} - c^k\big| - \big|u_i^{k, n} - c^k\big| \bigr) + \frac{ {\Delta t}}{ {\Delta x}} \bigl( Q_ {i+{1/2}}^{k, n} - Q_ {i-{1/2}}^{k, n} \bigr)\biggr) \\ &\quad +\sum\limits_{n = 0}^{M} \varphi_0^n\biggl( \frac{N}{2} \bigl( \big|u_0^{n+1} - c^0\big| - \big|u_0^n - c^0\big| \bigr) + \frac{ {\Delta t}}{ {\Delta x}} \biggl( \sum\limits_{k\in {{\mathscr{I}}_{\mathrm{out}}}} Q_{ {1/2}}^{k, n} - \sum\limits_{k\in {{\mathscr{I}}_{\mathrm{in}}}} Q_{- {1/2}}^{k, n} \biggr) \biggr), \end{align*}

    where \varphi_i^{k, n} = \varphi^k(x_i, t^n) . After summation by parts we get

    \begin{align*} 0& \geqslant-\sum\limits_{n = 1}^{M} \sum\limits_{k\in {\mathscr{I}}} \sum\limits_{i \in D_ \mathrm{disc}^k} \big|u_i^{k, n} - c^k\big| \bigl( \varphi_i^{k, n} - \varphi_i^{k, n-1} \bigr) - \sum\limits_{k\in {\mathscr{I}}} \sum\limits_{i \in D_ \mathrm{disc}^k} \big|u_i^{k, 0} - c^k\big| \varphi_i^{k, 0} \\ &\quad -\frac{ {\Delta t}}{ {\Delta x}} \sum\limits_{n = 0}^{M} \biggl(\begin{aligned}[t]& \sum\limits_{k\in {{\mathscr{I}}_{\mathrm{in}}}} \sum\limits_{i \in D_ \mathrm{disc}^k} Q_ {i-{1/2}}^{k, n} \bigl( \varphi_i^{k, n} - \varphi_{i-1}^{k, n} \bigr) \\ &+ \sum\limits_{k\in {{\mathscr{I}}_{\mathrm{out}}}} \sum\limits_{i \in D_ \mathrm{disc}^k} Q_ {i+{1/2}}^{k, n} \bigl( \varphi_{i+1}^{k, n} - \varphi_i^{k, n} \bigr) \\ &+ \sum\limits_{k\in {{\mathscr{I}}_{\mathrm{in}}}} \varphi_{-1}^{k, n} Q_{- {1/2}}^{k, n} - \sum\limits_{k\in {{\mathscr{I}}_{\mathrm{out}}}} \varphi_1^{k, n} Q_{ {1/2}}^{k, n} \biggr)\end{aligned} \\ &\quad-\frac{N}{2} \big|u_0^n - c^0\big| \varphi_0^0 - \frac{N}{2} \sum\limits_{n = 1}^{M} \big|u_0^n - c^0\big| \bigl( \varphi_0^n - \varphi_0^{n-1} \bigr) \\ &\quad+ \frac{ {\Delta t}}{ {\Delta x}} \sum\limits_{n = 0}^{M} \biggl(\sum\limits_{k\in {{\mathscr{I}}_{\mathrm{out}}}} Q_{ {1/2}}^{k, n} \varphi_0^n - \sum\limits_{k\in {{\mathscr{I}}_{\mathrm{in}}}} Q_{- {1/2}}^{k, n} \varphi_0^n \biggr). \end{align*}

    After shifting the i index on the second line we get

    \begin{align*} 0& \geqslant-\underbrace{ {\Delta t} {\Delta x} \sum\limits_{n = 1}^{M} \sum\limits_{k\in {\mathscr{I}}} \sum\limits_{i \in D_ \mathrm{disc}^k} \big|u_i^{k, n} - c^k\big| \biggl( \frac{ \varphi_i^{k, n} - \varphi_i^{k, n-1}}{ {\Delta t}} \biggr)}_{ = \, A_1} \end{align*}
    \begin{align*} &\quad- \underbrace{ {\Delta x} \sum\limits_{k\in {\mathscr{I}}} \sum\limits_{i \in D_ \mathrm{disc}^k} \big|u_i^{k, 0} - c^k\big| \varphi_i^{k, 0}}_{ = \, A_2} \\ &\quad- \underbrace{ {\Delta t} {\Delta x} \sum\limits_{n = 0}^{M} \biggl(\begin{aligned}[t]&\sum\limits_{k\in {{\mathscr{I}}_{\mathrm{in}}}} \sum\limits_{i \in D_ \mathrm{disc}^k} Q_ {i+{1/2}}^{k, n} \biggl( \frac{ \varphi_{i+1}^{k, n} - \varphi_{i}^{k, n}}{ {\Delta x}} \biggr) \\ &+ \sum\limits_{k\in {{\mathscr{I}}_{\mathrm{out}}}} \sum\limits_{i \in D_ \mathrm{disc}^k} Q_ {i-{1/2}}^{k, n} \biggl( \frac{ \varphi_{i}^{k, n} - \varphi_{i-1}^{k, n}}{ {\Delta x}} \biggr)\biggr)\end{aligned}}_{ = \, A_3} \\ &\quad- \underbrace{ {\Delta x} \frac{N}{2} \big|u_0^{0} - c^0\big| \varphi_0^0 - {\Delta t} {\Delta x} \frac{N}{2} \sum\limits_{n = 1}^{M} \big|u_0^n - c^0\big| \biggl( \frac{ \varphi_0^n - \varphi_0^{n-1}}{ {\Delta t}} \biggr)}_{ = \, A_4}. \end{align*}

    Taking limits we get for {\Delta t}, {\Delta x} \to 0

    \begin{align*} A_1 \to \sum\limits_{k\in {\mathscr{I}}} \int_0^{\infty} \int_{D^k} \big|u^k - c^k\big| \varphi_t^k \, dx \, dt, \end{align*}

    and for {\Delta x} \to 0

    \begin{align*} A_2 \to \sum\limits_{k\in {\mathscr{I}}} \big|u_0^k - c^k\big|(x) \varphi^k(x, 0) \, dx, \qquad A_4 \to 0. \end{align*}

    Thus, we are left with A_3 . Since the scheme is the upwind method, we can write

    \begin{align*} A_3 & = {\Delta t} {\Delta x} \sum\limits_{n = 0}^{M} \biggl(\begin{aligned}[t]&\sum\limits_{k\in {{\mathscr{I}}_{\mathrm{in}}}} \sum\limits_{i \in D_ \mathrm{disc}^k} q_{c^k}^k \bigl( u_i^{k, n} \bigr) \biggl( \frac{ \varphi_{i+1}^{k, n} - \varphi_{i}^{k, n}}{ {\Delta x}} \biggr) \\ &+ \sum\limits_{k\in {{\mathscr{I}}_{\mathrm{out}}}} \sum\limits_{i \in D_ \mathrm{disc}^k} q_{c^k}^k \bigl( u_{i}^{k, n} \bigr) \biggl( \frac{ \varphi_{i}^{k, n} - \varphi_{i-1}^{k, n}}{ {\Delta x}} \biggr)\biggr) \end{aligned}\\ &\to \sum\limits_{k\in {\mathscr{I}}} \int_0^T \int_{D^k} q_{c^k}^k \bigl( u^k(x, t) \bigr) \varphi_x (x, t) \, dx \, dt \end{align*}

    as {\Delta t}, {\Delta x} \to 0 , due to the a.e. pointwise convergence of {\mathbf{u}}_{ {\Delta t}} to {\mathbf{u}} .

    Now we have everything in place to proof a compactness theorem.

    Theorem 5.3 (Compactness and Convergence to a Weak Solution). Fix T > 0 . Assume that each flux function f^k is locally Lipschitz continuous and strictly monotone. Let \mathscr{G}_ \mathrm{disc}^0 be a set of discrete stationary states for the upwind method. Let {\mathbf{u}}_ {\Delta t} be computed from the upwind method with initial data \bar {\mathbf{u}} \in L^\infty_{\mathrm{oco}}( \mathscr{G}_ \mathrm{disc}^0) \cap L^1(\Omega;\lambda) , and assume that {\rm TV}(\bar {\mathbf{u}})<\infty . Then the numerical solution \{ {\mathbf{u}}_ {\Delta t}\}_{ {\Delta t}>0} converges in C([0, T], L_ {\mathrm{loc}}^1(\Omega;\lambda)) to a weak solution {\mathbf{u}} .

    Proof. We first show convergence of the sequence of functions {\mathbf{g}}_ {\Delta t}\colon\Omega\times[0, T]\to \mathbb{R} ,

    {\mathbf{g}}_ {\Delta t}(x, k, t): = f^k(u^k_ {\Delta t}(x, t)).

    The sequence {\mathbf{g}}_ {\Delta t} is uniformly L^\infty bounded, by Proposition 4.11, and it is Lipschitz continuous in time:

    \begin{align*} \int_{\Omega} \big| {\mathbf{g}}_ {\Delta t}(t^{n+1})- {\mathbf{g}}_ {\Delta t}(t^n)\big|\, d\lambda & \leqslant C_f\int_{\Omega}\big| {\mathbf{u}}_ {\Delta t}(t^{n+1})- {\mathbf{u}}_ {\Delta t}(t^n)\big|\, d\lambda \\ & \leqslant C_f(C{\rm TV}(\bar {\mathbf{u}})+\bar M) {\Delta t}, \end{align*}

    by Corollary 4.15. We can bound the total variation of {\mathbf{g}}_ {\Delta t} by

    \begin{align*} {\rm TV}( {\mathbf{g}}_ {\Delta t}(\cdot, t)) & = \sum\limits_{k\in {\mathscr{I}}}\sum\limits_{i \in D_ \mathrm{disc}^k} \big|f^k( u_i^{k, n}) - f^k( u_{i-1}^{k, n})\big| \\ &\quad + \sum\limits_{k\in {{\mathscr{I}}_{\mathrm{in}}}}\big|f^k(u_0^n)-f^k(u_{-1}^{k, n})\big| \\ & \leqslant \sum\limits_{k\in {\mathscr{I}}}\sum\limits_{i \in D_ \mathrm{disc}^k} \big|f^k( u_i^{k, n}) - f^k( u_{i-1}^{k, n})\big| + N\| {\mathbf{g}}_ {\Delta t}\|_{\infty} \\ & = \sum\limits_{k\in {\mathscr{I}}} \sum\limits_{i \in D_ \mathrm{disc}^k} \big|F_ {i+{1/2}}^{k, n} - F_ {i-{1/2}}^{k, n}\big| + N\| {\mathbf{g}}_ {\Delta t}\|_{\infty} \\ & = \frac{ {\Delta x}}{ {\Delta t}}\sum\limits_{k\in {\mathscr{I}}} \sum\limits_{i \in D_ \mathrm{disc}^k}\big|u_i^{k, n+1}-u_i^{k, n}\big| + N\| {\mathbf{g}}_ {\Delta t}\|_{\infty} \\ & \leqslant {\Delta x}(C{\rm TV}(\bar {\mathbf{u}}) + \bar M) + N\| {\mathbf{g}}_ {\Delta t}\|_{\infty}. \end{align*}

    Applying Ascoli's compactness theorem together with Helly's theorem, we get the existence of a subsequence {\Delta t}_\ell\to 0 such that {\mathbf{g}}_{ {\Delta t}_\ell} \to {\mathbf{g}} in C([0, T], L^1_ {\mathrm{loc}}(\Omega;\lambda)) for some function {\mathbf{g}} . The strict monotonicity of f^k implies that

    {\mathbf{u}}_ {\Delta t}(x, k, t) = ( f^k )^{-1}\big( {\mathbf{g}}_ {\Delta t}(x, k, t)\big),

    and hence, also {\mathbf{u}}_{ {\Delta t}_\ell} converges in C([0, T], L^1_ {\mathrm{loc}}(\Omega;\lambda)) to some function {\mathbf{u}} . Theorem 5.1 implies that {\mathbf{u}} is the entropy solution; since this solution is unique (Theorem 3.1), the entire sequence \{ {\mathbf{u}}_ {\Delta t}\}_{ {\Delta t}>0} must converge to {\mathbf{u}} .

    So far we have shown that if a sufficiently large class of stationary and discrete stationary solutions exists, then our equations on the network are well posed and the finite volume numerical approximations converge to the entropy solution. In this section we show that such classes exist in the case where either all fluxes f^k are strictly increasing or all are strictly decreasing. We also remark on the more general case.

    We henceforth assume that all fluxes are increasing; one can attain analogous results for decreasing fluxes following the same arguments. In the following we want to investigate the sets of discrete stationary solutions implied by the upwind method.

    We define

    \begin{equation*} f_ {\mathrm{in}}(u) : = \sum\limits_{k\in {{\mathscr{I}}_{\mathrm{in}}}} f^k(u), \qquad f_ {\mathrm{out}}(u) : = \sum\limits_{k\in {{\mathscr{I}}_{\mathrm{out}}}} f^k(u) \qquad \text{for } u\in \mathbb{R}. \end{equation*}

    It is clear that f_ {\mathrm{in}}, f_ {\mathrm{out}} are monotone by the monotonicity of their summand components. In particular, the two functions are invertible.

    For the upwind method the conditions (4.9a) and (4.9b) become

    \begin{align} f^k ( c^k ) & = f^k ( c^k ) \qquad \text{for } k\in {{\mathscr{I}}_{\mathrm{in}}}, \end{align} (6.1a)
    \begin{align} f^k ( c^0 ) & = f^k ( c^k ) \qquad \text{for }k\in {{\mathscr{I}}_{\mathrm{out}}}. \end{align} (6.1b)

    This is equivalent to

    \begin{align} c^k & = c^k \qquad \text{for } k\in {{\mathscr{I}}_{\mathrm{in}}}, \end{align} (6.2a)
    \begin{align} c^0 & = c^k \qquad \text{for }k\in {{\mathscr{I}}_{\mathrm{out}}}, \end{align} (6.2b)

    due to the invertibility of the flux functions f^k . It is obvious as well, that for two different discrete stationary solutions {\mathbf{c}}, {\mathbf{d}} satisfying c^k \leqslant d^k for k \in {\mathscr{I}} , we also have c^0 \leqslant d^0 . Henceforth, we denote

    \mathscr{G}_ \mathrm{disc}^0 : = \bigl\{\text{all discrete stationary states for the upwind method}\bigr\}

    and we let

    \mathscr{G}_ \mathrm{disc} : = \bigl\{\big(c^{-N_ {\mathrm{in}}}, \dots, c^{-1}, c^1, \dots, c^{N_ {\mathrm{out}}}\big) \mid {\mathbf{c}}\in \mathscr{G}_ \mathrm{disc}^0)\bigr\}.

    Although it might be too difficult to find a full characterization of the set L^\infty_{\mathrm{oco}}( \mathscr{G}) of admissible initial data, we will be able to characterize large subsets of L^\infty_{\mathrm{oco}}( \mathscr{G}) . Let

    I_ {\mathrm{in}} : = f_ {\mathrm{in}}^{-1} (R_ {\mathrm{in}}\cap R_ {\mathrm{out}}), \qquad I_ {\mathrm{out}} : = f_ {\mathrm{out}}^{-1} (R_ {\mathrm{in}}\cap R_ {\mathrm{out}})

    where

    R_ {\mathrm{in}} : = f_ {\mathrm{in}}( \mathbb{R}), \qquad R_ {\mathrm{out}} : = f_ {\mathrm{out}}( \mathbb{R}).

    By the continuity of f_ {\mathrm{in}}, f_ {\mathrm{out}} , the sets I_ {\mathrm{in}}, I_ {\mathrm{out}} are closed intervals.

    Theorem 6.1. We have \mathscr{L}\subset L^\infty_{\mathrm{oco}}( \mathscr{G}_ \mathrm{disc}) , where

    \mathscr{L}: = \Bigl\{ {\mathbf{u}}\in L^\infty(\Omega;\lambda) \mid u^k(x) \in I_ {\mathrm{in}}\ \forall\ k\in {{\mathscr{I}}_{\mathrm{in}}}, \ u^k(x)\in I_ {\mathrm{out}}\ \forall\ k\in {{\mathscr{I}}_{\mathrm{out}}} \Bigr\}.

    In particular, if f_ {\mathrm{in}}, f_ {\mathrm{out}} have the same range R_ {\mathrm{in}}, R_ {\mathrm{out}} , then L^\infty_{\mathrm{oco}}( \mathscr{G}_ \mathrm{disc}) = L^\infty(\Omega;\lambda) .

    Proof. Let {\mathbf{u}}\in\mathscr{L} . Since {\mathbf{u}}\in L^\infty(\Omega;\lambda) , and I_ {\mathrm{in}}, I_ {\mathrm{out}} are closed, we also have

    \underline{c}_ {\mathrm{in}}: = \inf\limits_{\substack{x\in D^k\\ k\in {{\mathscr{I}}_{\mathrm{in}}}}}u^k(x) \in I_ {\mathrm{in}}, \qquad \overline{c}_ {\mathrm{in}}: = \sup\limits_{\substack{x\in D^k\\ k\in {{\mathscr{I}}_{\mathrm{in}}}}}u^k(x) \in I_ {\mathrm{in}}

    and likewise for \underline{c}_ {\mathrm{out}}, \overline{c}_ {\mathrm{out}} . By continuity of f_ {\mathrm{in}}, f_ {\mathrm{out}} , there are \underline{d}_ {\mathrm{in}}\in I_ {\mathrm{in}} and \underline{d}_ {\mathrm{out}}\in I_ {\mathrm{out}} satisfying \underline{d}_ {\mathrm{in}} \leqslant\underline{c}_ {\mathrm{in}} and \underline{d}_ {\mathrm{out}} \leqslant \underline{c}_ {\mathrm{out}} so that f_ {\mathrm{in}}(\underline{d}_ {\mathrm{in}}) = f_ {\mathrm{out}}(\underline{d}_ {\mathrm{out}}) , that is, the vector \underline{ {\mathbf{d}}}: = \bigl(\underline{d}_ {\mathrm{in}}, \dots, \underline{d}_ {\mathrm{in}}, \underline{d}_ {\mathrm{out}}, \dots, \underline{d}_ {\mathrm{out}}\bigr) is a stationary solution. This stationary solution clearly satisfies (6.2), whence {\mathbf{d}}\in \mathscr{G}_ \mathrm{disc}^0 .

    In a similar way one finds a stationary solution \overline{ {\mathbf{d}}}: = \bigl(\overline{d}_ {\mathrm{in}}, \dots, \overline{d}_ {\mathrm{in}}, \overline{d}_ {\mathrm{out}}, \dots, \overline{d}_ {\mathrm{out}}\bigr) \in \mathscr{G}_ \mathrm{disc}^0 which bounds {\mathbf{u}} from above. Since now

    \underline{d}_ {\mathrm{in}} \leqslant u^k(x) \leqslant\overline{d}_ {\mathrm{in}} \;\forall\ k\in {{\mathscr{I}}_{\mathrm{in}}} \qquad \text{and} \qquad \underline{d}_ {\mathrm{out}} \leqslant u^k(x) \leqslant\overline{d}_ {\mathrm{out}} \;\forall\ k\in {{\mathscr{I}}_{\mathrm{out}}}

    we conclude that {\mathbf{u}}\in L^\infty_{\mathrm{oco}}( \mathscr{G}_ \mathrm{disc}) .

    Proposition 6.2. Consider a conservation law on a network with strictly increasing fluxes f^k . Let \mathscr{G}_ \mathrm{disc}^0 denote the set of all discrete stationary solutions for the upwind method. Then the set

    \mathscr{G}_ \mathrm{disc} : = \Big\lbrace \big(c^{- {N_{\mathrm{in}}}}, \dots, c^{-1}, c^1, \dots, c^{ {N_{\mathrm{out}}}}\big) \mid {\mathbf{c}} \in \mathscr{G}^0_ \mathrm{disc} \Big\rbrace

    is a mutually consistent and maximal set of stationary solutions.

    Proof. Every {\mathbf{c}}\in \mathscr{G}_ \mathrm{disc} is a stationary solution due to (4.8).

    To prove mutual consistency of \mathscr{G}_ \mathrm{disc} we plug a discrete stationary solution {\mathbf{d}}\in \mathscr{G}_ \mathrm{disc}^0 into (5.1) to get for n \in \mathbb{N} ,

    Q_{- {1/2}}^{k, n} \geqslant Q_{- {3/2}}^{k, n} \;\text{for }k \in {\mathscr{I}}_ {\mathrm{in}} \quad\text{and}\quad Q_{ {3/2}}^{k, n} \geqslant Q_ {1/2}^{k, n} \;\text{for } k \in {\mathscr{I}}_ {\mathrm{out}}.

    Since we are using the upwind scheme, this reduces to

    q_{c^k}^k (d^k) \geqslant q_{c^0}^k (d^0) \text{ for } k \in {\mathscr{I}}_ {\mathrm{in}} \qquad\text{and}\qquad q_{c^0}^k (d^0) \geqslant q^k_{c^k} (d^k) \text{ for } k \in {\mathscr{I}}_ {\mathrm{out}}.

    In the same manner, plugging d^0 into (5.2) gives us

    \sum\limits_{k\in {{\mathscr{I}}_{\mathrm{in}}}} Q_{- {1/2}}^{k, n} \geqslant \sum\limits_{k\in {{\mathscr{I}}_{\mathrm{out}}}} Q_ {1/2}^{k, n}.

    Combining these two observations, we get

    \sum\limits_{k\in {{\mathscr{I}}_{\mathrm{in}}}} q_{c^k}^k (d^k) \geqslant \sum\limits_{k\in {{\mathscr{I}}_{\mathrm{in}}}} q_{c^0}^k (d^0) \geqslant \sum\limits_{k\in {{\mathscr{I}}_{\mathrm{out}}}} q_{c^0}^k (d^0) \geqslant \sum\limits_{k\in {{\mathscr{I}}_{\mathrm{out}}}} q_{c^k}^k (d^k).

    As {\mathbf{c}}, {\mathbf{d}} were arbitrary, it follows that \mathscr{G}_ \mathrm{disc} is mutually consistent.

    If for some vector {\mathbf{d}} , the set \mathscr{G}_ \mathrm{disc} \cup \lbrace {\mathbf{d}} \rbrace is mutually consistent, then

    \sum\limits_{k\in {\mathscr{I}}_ {\mathrm{in}}} q^k_{c^k}\big(d^k\big) \geqslant \sum\limits_{k\in {\mathscr{I}}_ {\mathrm{out}}} q^k_{c^k}\big(d^k\big).

    We choose c^k = d^k\ \forall\ k \in {\mathscr{I}}_ {\mathrm{in}} and c^0 = (f_ {\mathrm{out}})^{-1}(\sum_{k\in {\mathscr{I}}_ {\mathrm{in}}} f^k(c^k)) . Since all f^k are monotonically increasing, the entropy flux reduces to q^k_{c^k}(d^k) = \big|f^k(c^k) - f^k(d^k)\big| and thus,

    0 \geqslant \sum\limits_{k\in {\mathscr{I}}_ {\mathrm{out}}} \big|f^k(c^0) - f^k(d^k)\big|,

    which implies d^k = c^0 for k \in {\mathscr{I}}_ {\mathrm{out}} , and thus, {\mathbf{d}}\in \mathscr{G}_ \mathrm{disc} . In other words, \mathscr{G}_ \mathrm{disc} is maximal.

    Although the framework presented in this manuscript is only applied to monotone flux functions, we remark here on the generalization of our results to more general choices of f^k . The two main ingredients are

    ● compactness of the sequence of approximations (here achieved via a TV bound on the (upwind) numerical flux);

    ● the existence of a maximal set of stationary states, and the consistency of the approximations with respect to that set.

    A TV bound on the numerical fluxes can be achieved in a more general setting, but that does not easily translate to compactness of the approximation itself. This can be achieved by a detour via the Temple functional [28]. The derivation of a maximal set of stationary states requires a careful design of the numerical method. We address both of these issues in the upcoming paper [13], where we prove convergence of an Engquist–Osher-type finite volume method for more general flux functions.

    We show numerical experiments for some example cases including results for linear and nonlinear as well as convex and concave fluxes. In all experiments we use a CFL number of {1/2} – that is, {\Delta t} is chosen so that there is equality in (4.7). In all experiments we compute the experimental order of convergence (EOC) as p \approx \dfrac{\log(e^{j+1}/e^{j})}{\log({\Delta x}_{j+1}/ {\Delta x}_j)} on a series of successive grids with 2^{j} cells, where e^j denotes the L^1 error on grid level j . The error is computed as the L^1 difference to a high-resolution reference solution. All errors and EOC are displayed in Table 1.

    Table 1. 

    L^1 errors and estimated orders of convergence (EOC) for a selection of examples

    .
    Example 7.3 Example 7.5 Example 7.4 Example 7.1 Example 7.2
    Grid level L^1 error EOC L^1 error EOC L^1 error EOC L^1 error EOC L^1 error EOC
    3 0.10877 - 0.11630 - 0.14459 - 0.07087 - 0.09904 -
    4 0.05496 0.98 0.07136 0.70 0.08016 0.85 0.0546 0.38 0.04913 1.01
    5 0.03649 0.59 0.04372 0.71 0.04651 0.79 0.03117 0.81 0.02844 0.79
    6 0.02629 0.47 0.02255 0.96 0.02711 0.78 0.01903 0.71 0.01627 0.81
    7 0.01830 0.52 0.01360 0.73 0.01495 0.86 0.01115 0.77 0.00919 0.82
    8 0.01255 0.54 0.00653 1.06 0.00925 0.69 0.00644 0.79 0.00527 0.80
    9 0.00883 0.51 0.00325 1.01 0.00480 0.95 0.00330 0.96 0.00268 0.98
    10 0.00625 0.50 0.00160 1.02 0.00295 0.70 0.00173 0.93 0.00150 0.84
    11 0.00442 0.50 0.00086 0.90 0.00152 0.96 0.00085 1.03 0.00084 0.84
    12 0.00312 0.50 0.00040 1.10 0.00081 0.91 0.00042 1.02 0.00047 0.84

     | Show Table
    DownLoad: CSV

    Example 7.1 (Burgers' equation with roundabout). In this example we include a roundabout – an edge whose endpoints meet at the same vertex, as shown in Figure 2. This case was not included in the theory but is interesting because it is analogous to a periodic boundary condition. We also include an ingoing edge and two outgoing edges, amounting to a total of two ingoing and three outgoing edges. As initial data we choose constants on the roundabout and the outgoing edges and two different constants on the independent ingoing edge. After a while the shock in the initial data on the independent ingoing edge will reach the edge and create new Riemann problems. We choose the initial data

    \begin{align*} \bar u^{-1}(x) = \begin{cases} 2 & \text{if } 0 \leqslant x < 0.5, \\ \sqrt{2} & \text{if } 0.5 \leqslant x, \end{cases}\qquad \bar u^{-2} = \bar u^1 = \bar u^2 = \bar u^3 \equiv 1. \end{align*}
    Figure 2. 

    A network with a periodic edge

    .

    We take all edges to have length 1 and choose zero Neumann boundary data on the outer boundaries. On the vertex we set u_0^0 = 1 . On the ingoing edge with index -1 we have a travelling shock wave

    \begin{equation*} u^{-1}(x, t) = \begin{cases} 2 & \text{if } 0 \leqslant x < \frac{1}{2-\sqrt{2}} t \\ \sqrt{2} & \text{if } \frac{1}{2-\sqrt{2}} t \leqslant x \end{cases} \end{equation*}

    which will hit the vertex at t^* = 1-\frac{1}{\sqrt{2}} . To compute the solution after t^* we compute the new vertex value c^0 = \sqrt{5/3} and therefore get the Riemann problem

    \begin{align*} u^k(x, t^*) = \begin{cases} \sqrt{5/3} & \text{if } x = 0, \\ 1 & \text{if } 0 < x \end{cases} \end{align*}

    for k = 1, 2, 3 , which results in a travelling shock wave with speed s = \frac{1}{\sqrt{5/3} - 1} . At time t^{**} : = \sqrt{5/3} - \frac{1}{\sqrt{2}} the travelling shock wave which originated on the roundabout edge hits the vertex once again, resulting in a new set of Riemann problems on the outgoing edge. This process will continue in a periodic fashion.

    We compute up to time T = 0.5 . A plot of the exact and approximate solution to this example at two different times is shown in Figure 3. The accuracy and order of convergence of the numerical approximation are shown in Table 1.

    Figure 3. 

    Initial state and state at t = 0.7 of a {Burgers-type equation} with travelling shock wave which hits the vertex at time t = 1-\frac{1}{\sqrt{2}} . Here, the graph includes a periodic edge

    .

    Example 7.2. We construct an example where we take the flux function from the traffic flow example in [18], f(u) = 4u(1-u) , but allow for different fluxes on different edges, f^k(u) = \alpha^k f \big(\frac{u}{\alpha^k} \big) for \alpha^k>0 , and compute on a star shaped graph with two ingoing edges and three outgoing edges like in Figure 1. The initial data is chosen so that all fluxes are strictly increasing over the range of \bar u ; thus, the fluxes f^k are in effect monotonously increasing functions. We choose constant solutions on the two ingoing roads and constant initial data on the outgoing roads which are chosen such that on one road a shock will develop, on one road the solution will stay constant over time and on one road a rarefaction wave will develop.

    Solving the conditions (4.8), (4.9) for c^0 yields

    c^0 = \frac{3 \pm \sqrt{9 - 4 \bigl(\frac{1}{\alpha_1} + \frac{1}{\alpha_2} + \frac{1}{\alpha_3}\bigr) \big( u^{-1} - \frac{1}{\alpha_{-1}} ( u^{-1} )^2 + u^{-2} - \frac{1}{\alpha_{-2}} ( u^{-2} )^2 \big)}}{2\big(\frac{1}{\alpha_1} + \frac{1}{\alpha_2} + \frac{1}{\alpha_3}\big)}.

    For the incoming edges to have a monotonically increasing flux we impose u_{-i} \leqslant \frac{1}{2} \alpha_{-i}, for i = 1, 2 and for outgoing edges c^0 \leqslant \frac{1}{2} \min \lbrace \alpha_1, \alpha_2, \alpha_3 \rbrace . We choose \alpha_{-1} = \alpha_{-2} = 1 , \alpha_1 = \alpha_2 = 4 and \alpha_3 = 2 with initial data

    \begin{align*} \bar u^{-1} = \bar u^{-2} \equiv 0.5, \qquad \bar u^1 \equiv 0, \qquad \bar u^2 \equiv \frac{1}{2} \big(3-\sqrt{7}\big), \qquad \bar u^3 \equiv 1. \end{align*}

    This gives us c^0 = \frac{1}{2}(3 - \sqrt{7}) . On the outer boundary we choose zero Neumann boundary conditions. For u^3 we will get a shock

    u^3(x, t) = \begin{cases} \frac{1}{2}(3 - \sqrt{7}) & \text{if } x < st, \\ 0 & \text{if } x \geqslant st, \end{cases}

    with speed s = \frac{(3-\sqrt{7})(2-\frac{3-\sqrt{7}}{4})-3}{\frac{3-\sqrt{7}}{2}-1}, and a rarefaction wave for u^1 of the form

    u^1(x, t) = \begin{cases} \frac{3-\sqrt{7}}{2} & \text{if }x < 2 (\sqrt{7}-1)t, \\ 1-\frac{x}{4t} & \text{if } 2(\sqrt{7}-1)t \leqslant x < 4t, \\ 0 & \text{if } 4t \leqslant x. \end{cases}

    On edge 2 we get the constant solution u^2 \equiv \frac{1}{2}(3 - \sqrt{7}) .

    We compute up to time T = 0.2 . A plot of the exact and approximate solution at two different timepoints is shown in Figure 4. Accuracy and order of convergence of the numerical approximation are shown in Table 1.

    Figure 4. 

    Initial state at t = 0 and state at t = 0.2 of a traffic flow problem with an initial shock at the vertex developing a different elementary wave on each outgoing edge

    .

    In addition to the examples described above we show errors and experimental order of convergence (EOC) for several additional examples in Table 1.

    Example 7.3 (EOC: Linear advection). We consider a linear advection equation with two ingoing edges and three outgoing edges as in Figure 1 with initial data

    \begin{align*} \bar u^{-1}(x, t) = \begin{cases} 2 & 0 \leqslant x < 0.8, \\ 1 & 0.8 \leqslant x, \end{cases}\qquad \bar u^{-2} \equiv 1, \qquad \bar u^1 = \bar u^2 = \bar u^3 \equiv \frac{2}{3}, \end{align*}

    and Dirichlet boundary conditions adapted to the edge values. We initialize the vertex node by u_0^0 = \frac{2}{3} . We compute up to time T = 0.5 .

    Example 7.4 (EOC: Burgers' equation with elementary waves). We choose \bar u^{-1} = \bar u^{-2} \equiv 1 as initial data on the ingoing roads and \bar u^1 \equiv 0 , \bar u^2 \equiv\sqrt{2/3} and \bar u^3 \equiv 2 on the outgoing edges of a star shaped graph as in Figure 1. The conditions on the numerical flux imply then 3(c^0)^2 = 2 \Leftrightarrow c^0 = \sqrt{3/2} . Thus, we get the following Riemann problems on the outgoing roads:

    \begin{align*} \bar u^1(x) = \begin{cases} \sqrt{2/3} & x = 0, \\ 0 & x > 0, \end{cases}\qquad \bar u^2 = \begin{cases} \sqrt{2/3} & x = 0, \\ \sqrt{2/3} & x > 0, \end{cases}\qquad \bar u^3 = \begin{cases} \sqrt{2/3} & x = 0, \\ 2 & x > 0, \end{cases} \end{align*}

    with zero Neumann boundary conditions at the outer edges. The solution to these problems are a shock, a constant solution and a rarefaction wave, respectively. We compute up to time T = 0.3 .

    Example 7.5 (EOC: Burgers' equation with travelling shock). We consider a Burgers-type equation with two ingoing edges and three outgoing edges as in Figure 1 with initial data

    \begin{align*} \bar u^{-1}(x) = \begin{cases} 2 & 0 \leqslant x < 0.8, \\ 1 & 0.8 \leqslant x, \end{cases}\qquad \bar u^{-2} \equiv 1, \qquad \bar u^1 = \bar u^2 = \bar u^3 \equiv \sqrt{\frac{2}{3}}, \end{align*}

    with Dirichlet boundary conditions of the same value as the associated edge. On the vertex node the initial condition is chosen as u_0^0 = \sqrt{\frac{2}{3}} . We compute up to T = 0.5 .

    Convergence order estimates for finite volume methods for nonlinear scalar conservation laws are due to Kuznetsov [22] for the continuous flux case and due to Badwaik, Ruf [5] for the case of monotone fluxes with points of discontinuity. In both of those cases the analytically proven convergence rate is at least \sqrt{ {\Delta x}} . Our numerical experiments indicate the same lower bound on the convergence rate for our numerical methods on graphs. Considering the fact that f_ {\mathrm{in}} and f_ {\mathrm{out}} from Section 6 are monotone it might be possible to generalize the result of Badwaik and Ruf to networks.

    In conclusion we have defined a framework for the analysis and numerical approximation of conservation laws on networks. We extended the concepts well known from the conventional case such as weak solution, entropy solution and monotone methods to make sense on a directed graph. We defined a reasonable entropy condition under which we have shown stability and uniqueness of an analytic solution. Existence is shown by convergence of a conservative, consistent, monotone difference scheme. In an upcoming work [13] we want to address convergence of a numerical method where the fluxes f^k are not monotone but concave, as is usually found in traffic flow models. This includes deriving a sufficiently large set of stationary and discrete stationary solutions for this case. Further, we want to extend our model to include boundary conditions and derive a convergence order estimate for numerical approximations. As for future work, a generalization to systems of conservation laws would be highly desirable. One could also try to construct numerical schemes for equations incorporating diffusive fluxes like it was done in [20] on the line. Generalized models would span more complex scenarios such as blood circulation [6] in a network of veins or a river delta by the means of Euler equations and shallow water equations, respectively.

    We would like to thank the referee for the valuable comments, helping to improve the quality of this work.



    [1] World Economic Outlook Databases (2019). Available from: https://www.imf.org/external/pubs/ft/weo/2019/02/weodata/index.aspx.
    [2] Ministry of Economic Affairs (2020). Energy statistics handbook-2019. MOEA: Taipei (Taiwan).
    [3] Wu JH, Huang YH (2006) Renewable energy perspectives and support mechanisms in Taiwan. Renew Energy 31: 1718-1732. doi: 10.1016/j.renene.2005.09.007
    [4] Chen FL, Lu SM, Chang YL (2007) Renewable energy in Taiwan: Its developing status and strategy. Energy 32: 1634-1646. doi: 10.1016/j.energy.2006.12.007
    [5] Hwang JJ (2010) Promotional policy for renewable energy development in Taiwan. Renew Sustain Energy Rev 14: 1079-1087. doi: 10.1016/j.rser.2009.10.029
    [6] Liou HM (2010) Policies and legislation driving Taiwan's development of renewable energy. Renew Sustain Energy Rev 14: 1763-1781. doi: 10.1016/j.rser.2010.02.013
    [7] Chen HH, Lee AHI (2014) Comprehensive overview of renewable energy development in Taiwan. Renew Sustain Energy Rev 37: 215-228. doi: 10.1016/j.rser.2014.04.055
    [8] Balat M (2010) Sustainable developments of wind power in Europe. Part 1: utilization and future prospects. Energy Sources B 5: 409-423. doi: 10.1080/15567240802532528
    [9] Pablo-Romero MP, Pozo-Barajas R (2017) Global changes in total and wind electricity (1990-2014). AIMS Energy 5: 290-312. doi: 10.3934/energy.2017.2.290
    [10] Soares-Ramos EPP, de Oliveira-Assis L, Sarrias-Mena R, et al. (2020) Current status and future trends of offshore wind power in Europe. Energy 202: 117787. doi: 10.1016/j.energy.2020.117787
    [11] Dawn S, Tiwari PK, Goswami AK, et al. (2019) Wind power: Existing status, achievements and government's initiative towards renewable power dominating India. Energy Strategy Rev 23: 178-199. doi: 10.1016/j.esr.2019.01.002
    [12] Zhang S, Wei J, Chen X, et al. (2020) China in global wind power development: Role, status and impact. Renew Sustain Energy Rev 127: 109881. doi: 10.1016/j.rser.2020.109881
    [13] Energy Statistics Information Query System (2020). Available from: https://www.moeaboe.gov.tw/wesnq/.
    [14] Guidelines on Energy Development (2017). Available from: https://adapt.epa.gov.tw/TCCIP-1-D/TCCIP-1-D-6.
    [15] Laws and Regulation Retrieving System (2021). Available from: https://law.moj.gov.tw/Eng/index.aspx.
    [16] Shieh CL, Wang CM, Chen YS, et al. (2010) An overview of disasters resulted from Typhoon Morakot in Taiwan. J Disaster Res 5: 236-244. doi: 10.20965/jdr.2010.p0236
    [17] Global Offshore Wind Speeds Rankings (2020). Available from: https://www.4coffshore.com/windfarms/windspeeds.aspx.
    [18] Fang HF (2014) Wind energy potential assessment for the offshore areas of Taiwan west coast and Penghu Archipelago. Renew Energy 67: 237-241. doi: 10.1016/j.renene.2013.11.047
    [19] Chang PC, Yang RY, Lai CM (2015) Potential of offshore wind energy and extreme wind speed forecasting on the west coast of Taiwan. Energies 8: 1685-1700. doi: 10.3390/en8031685
    [20] Valentin SV (2010) A STEP toward understanding wind power development policy barriers in advanced economies. Renew Sustain Energy Rev 14: 2796-2807. doi: 10.1016/j.rser.2010.07.043
    [21] Lu SM (2017) Renewable Energies in Taiwan: Reserve, Status, Policy; LAP LAMBERT Academic Publishing: Saarbrucken, Germany, 2017.
    [22] Lin CJ, Yu OS, Chang CL, et al. (2007) Challenges of wind farms connection to future power systems in Taiwan. Renew Energy 34: 1926-1930. doi: 10.1016/j.renene.2008.12.004
    [23] Lee YM, Tzeng YE (2008) Development and life-cycle inventory analysis of wind energy in Taiwan. J Energy Eng 134: 53-57. doi: 10.1061/(ASCE)0733-9402(2008)134:2(53)
    [24] Yue CD, Yang MH (2009) Exploring the potential of wind energy for a coastal state. Energy Policy 37: 3925-3940. doi: 10.1016/j.enpol.2009.04.055
    [25] Chen FL, Lu SM, Tseng KT, et al. (2010) Assessment of renewable energy reserves in Taiwan. Renew Sustain Energy Rev 14: 2511-2528. doi: 10.1016/j.rser.2010.06.021
    [26] Lee TL (2010) Assessment of the potential of offshore wind energy in Taiwan using fuzzy analytic hierarchy process. Open Civil Eng J 4: 96-104. doi: 10.2174/1874149501004010096
    [27] Chou CJ, Wu YK, Han GY, et al. (2012) Comparative evaluation of the HVDC and HVAC links integrated in a large offshore wind farm-an actual case study in Taiwan. IEEE Trans Ind Appl 48: 1639-1648. doi: 10.1109/TIA.2012.2209622
    [28] Kao SM, Pearre NS (2017) Administrative arrangement for offshore wind power developments in Taiwan: Challenges and prospects. Energy Policy 109: 463-472. doi: 10.1016/j.enpol.2017.07.027
    [29] Zhang Y, Zhang C, Chang YC, et al. (2017) Offshore wind farm in marine spatial planning and the stakeholders engagement: Opportunities and challenges for Taiwan. Ocean Coast Manag 149: 69-80. doi: 10.1016/j.ocecoaman.2017.09.014
    [30] Chien KH (2020) Pacing for renewable energy development: The development state in Taiwan's offshore wind power. Ann Am Assoc Geogr 110: 793-807.
    [31] Valentin SV (2010) Disputed wind directions: Reinvigorating wind power development in Taiwan. Energy Sustain Dev 14: 22-34. doi: 10.1016/j.esd.2009.12.005
    [32] Huang YH, Wu JH (2011) Assessment of the feed-in tariff mechanism for renewable energies in Taiwan. Energy Policy 39: 8106-8115. doi: 10.1016/j.enpol.2011.10.005
    [33] Tsai WT (2013) Prospects of renewable energy development in Taiwan. Energy Sources B 8: 263-269. doi: 10.1080/15567240903394281
    [34] Tsai WT (2014) Feed-in tariff promotion and innovative measures for renewable electricity: Taiwan case analysis. Renew Sustain Energy Rev 40: 1126-1132. doi: 10.1016/j.rser.2014.08.019
    [35] Chang CT, Lee HC (2016) Taiwan's renewable energy strategy and energy-intensive industrial policy. Renew Sustain Energy Rev 64: 456-465. doi: 10.1016/j.rser.2016.06.052
    [36] Kung CC, Zhang L, Chang MS (2017) Promotion policies for renewable energy and their effects in Taiwan. J Clean Prod 142: 965-975. doi: 10.1016/j.jclepro.2016.05.034
    [37] Kung CC, McCarl BA (2020) The potential role of renewable electricity generation in Taiwan. Energy Policy 138: 111227. doi: 10.1016/j.enpol.2019.111227
    [38] Electricity carbon emission factor (2020). Available from: https://www.moeaboe.gov.tw/ecw/populace/content/ContentDesc.aspx?menu_id=6989.
    [39] World Health Organization (2020). WHO coronavirus disease (COVID-19) dashboard. Available from: https://covid19.who.int/.
    [40] Cheval S, Adamescu CM, Georgiadis T, et al. (2020) Observed and potential impacts of the COVID-19 pandemic on the environment. Int J Environ Res Public Health 17: 4140. doi: 10.3390/ijerph17114140
    [41] International Energy Agency (2020) Covid-19 Impact on Renewable Energy Growth, Available from: https://www.iea.org/reports/renewable-energy-market-update/covid-19-impact-on-renewable-energy-growth.
    [42] Eroglu H (2020) Effects of Covid-19 outbreak on environment and renewable energy sector. Environ Dev Sustain (in press). Available from: https://doi.org/10.1007/s10668-020-00837-4.
    [43] Gillingham KT, Knittel CR, Li J, et al. (2020) The Short-run and Long-run Effects of Covid-19 on Energy and the Environment. Joule 4: 1337-1341. doi: 10.1016/j.joule.2020.06.010
  • This article has been cited by:

    1. Nicola De Nitti, Enrique Zuazua, On the Controllability of Entropy Solutions of Scalar Conservation Laws at a Junction via Lyapunov Methods, 2023, 51, 2305-221X, 71, 10.1007/s10013-022-00598-9
    2. Ulrik S. Fjordholm, Markus Musch, Nils H. Risebro, Well-Posedness and Convergence of a Finite Volume Method for Conservation Laws on Networks, 2022, 60, 0036-1429, 606, 10.1137/21M145001X
    3. P. Cardaliaguet, N. Forcadel, Microscopic Derivation of a Traffic Flow Model with a Bifurcation, 2024, 248, 0003-9527, 10.1007/s00205-023-01948-8
    4. Pierre Cardaliaguet, Nicolas Forcadel, Régis Monneau, A class of germs arising from homogenization in traffic flow on junctions, 2024, 21, 0219-8916, 189, 10.1142/S0219891624500073
    5. Alexandre M. Bayen, Alexander Keimer, Nils Müller, A proof of Kirchhoff's first law for hyperbolic conservation laws on networks, 2023, 18, 1556-1801, 1799, 10.3934/nhm.2023078
    6. Dilip Sarkar, Shridhar Kumar, Pratibhamoy Das, Higinio Ramos, Higher-order convergence analysis for interior and boundary layers in a semi-linear reaction-diffusion system networked by a k -star graph with non-smooth source terms, 2024, 19, 1556-1801, 1085, 10.3934/nhm.2024048
    7. Taras Mel’nyk, Christian Rohde, Asymptotic expansion for convection-dominated transport in a thin graph-like junction, 2024, 22, 0219-5305, 833, 10.1142/S0219530524500040
    8. Jon Asier Bárcena-Petisco, Márcio Cavalcante, Giuseppe Maria Coclite, Nicola De Nitti, Enrique Zuazua, Control of hyperbolic and parabolic equations on networks and singular limits, 2024, 0, 2156-8472, 0, 10.3934/mcrf.2024015
  • Reader Comments
  • © 2021 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

Metrics

Article views(4308) PDF downloads(194) Cited by(7)

Figures and Tables

Figures(2)  /  Tables(5)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog