Keywords

These keywords were added by machine and not by the authors. This process is experimental and the keywords may be updated as the learning algorithm improves.

1 Continuous Problem

\(\varepsilon \) - spi1

Let \(\Omega \subset {\mathbb{R}}^{d},d = 1,2,3\) be a bounded open (polyhedral) domain. We treat the following nonlinear convection-diffusion problem: find \(u : \Omega \times (0,T) \rightarrow \mathbb{R}\) such that

$$\begin{array}{rcl} \mbox{ (a)}\quad \frac{\partial u} {\partial t} +\mathrm{ div}\,\mathbf{f}(u) = \varepsilon \varDelta u + g\quad \mbox{ in}\ \Omega \times (0,T),& &\end{array}$$
(1)
$$\begin{array}{rcl} \mbox{ (b)}\quad u\big|_{\Gamma _{D}\times (0,T)} = u_{D},\quad \varepsilon \,\frac{\partial u} {\partial n}\big|_{\Gamma _{N}\times (0,T)} = g_{N},& &\end{array}$$
(2)

along with the initial condition \(u(x,0) = {u}^{0}(x)\) in Ω. The diffusion coefficient \(\varepsilon \geq 0\) is a given constant, \(g,u_{D},g_{N}\), and u 0 are given functions.

We assume that the convective fluxes \(\mathbf{f} = (f_{1},\cdots \,,f_{d}) \in {(C_{b}^{2}(\mathbb{R}))}^{d} = {({C}^{2}(\mathbb{R}) \cap {W}^{2,\infty }(\mathbb{R}))}^{d}\), hence f and \({\mathbf{f}}^{{\prime}} = (f_{1}^{{\prime}},\cdots \,,f_{d}^{{\prime}})\) are globally Lipschitz continuous. For improved estimates via Remark 1, we shall assume \(\mathbf{f} \in {(C_{b}^{3}(\mathbb{R}))}^{d}\). In [4], the error analysis is extended, assuming only local properties, i.e. \(\mathbf{f} \in {({C}^{2}(\mathbb{R}))}^{d}\) and \(\mathbf{f} \in {({C}^{3}(\mathbb{R}))}^{d}\).

In our analysis, we need to assume that Γ N is an outflow boundary for either u or u h , i.e. e.g. for u, we assume \(\Gamma _{N}^{(t)} \subseteq \{ x \in \partial \Omega ;{\mathbf{f}}^{{\prime}}(u(x,t)).\mathbf{n} \geq 0\}\) and \(\Gamma _{D}^{(t)} := \partial \Omega \setminus \Gamma _{N}\).

2 Discretization

Let \(\mathcal{T}_{h}\) be (generally nonconforming) triangulation of \(\overline{\Omega }\). For \(K \in \mathcal{T}_{h}\) we set \(h = \mbox{ max}_{K\in \mathcal{T}_{h}}\mbox{ diam}(K)\). By \(\mathcal{F}_{h}\) we denote the system of all faces of all elements \(K \in \mathcal{T}_{h}\). By \(\mathcal{F}_{h}^{I},\mathcal{F}_{h}^{D},\mathcal{F}_{h}^{N},\mathcal{F}_{h}^{B}\) we denote the sets on interior, Dirichlet, Neumann and boundary edges, respectively. For each \(\Gamma \in \mathcal{F}_{h}\) we define a fixed unit normal n Γ , which has the same orientation as the outer normal to \(\partial \Omega \) if \(\Gamma \in \mathcal{F}_{h}^{B}\).

Over a triangulation \(\mathcal{T}_{h}\) we define the broken Sobolev spaces \({H}^{k}(\Omega ,\mathcal{T}_{h}) =\{ v;\,v\vert _{K} \in {H}^{k}(K),\,\forall K \in \mathcal{T}_{h}\}\). For \(\Gamma \in \mathcal{F}_{h}^{I}\) we have two neighbours \(K_{\Gamma }^{(L)},\,K_{\Gamma }^{(R)} \in \mathcal{T}_{h}\), where n Γ is the outer normal to \(K_{\Gamma }^{(L)}\). For \(v \in {H}^{1}(\Omega ,\mathcal{T}_{h})\) we define on \(\Gamma \in \mathcal{F}_{h}^{I}\): \(v\vert _{\Gamma }^{(L)} = \text{ the trace of }v\vert _{K_{\Gamma }^{(L)}}\text{ on }\Gamma ,\;v\vert _{\Gamma }^{(R)} = \text{ the trace of }v\vert _{K_{\Gamma }^{(R)}}\text{ on }\Gamma ,\;\langle v\rangle _{\Gamma } = \frac{1} {2}\big{(}v\vert _{\Gamma }^{(L)}+v\vert _{ \Gamma }^{(R)}\big{)}\)and \([v]_{\Gamma } = v\vert _{\Gamma }^{(L)} - v\vert _{\Gamma }^{(R)}.\) On \(\Gamma \in \mathcal{F}_{h}^{B}\) we set \(v_{\Gamma } = v\vert _{\Gamma }^{(L)} = \text{ the trace of }v\vert _{K_{\Gamma }^{(L)}}\text{ on }\Gamma \), while \(v\vert _{\Gamma }^{(R)} = u_{D}\) on Γ D , \(v\vert _{\Gamma }^{(R)} = v\vert _{\Gamma }^{(L)}\) on Γ N .

Let p ≥ 1 be an integer. The approximate solution will be sought in the space of discontinuous piecewise polynomial functions \(S_{h} =\{ v;\,v\vert _{K} \in {P}^{p}(K),\forall K \in \mathcal{T}_{h}\},\) where P p(K) are polynomials on K of degree ≤ p. By \((\cdot ,\cdot )\) we denote the L 2(Ω)-scalar product and by \(\|\cdot \|\) the L 2(Ω)-norm. By \(\|\cdot \|_{\infty }\), we denote the \({L}^{\infty }(\Omega )\)-norm.

We introduce the following forms defined for \(v,\varphi \in {H}^{2}(\Omega ,\mathcal{T}_{h})\). Diffusion form:

$$\begin{array}{ll} a_{h}(v,\varphi ) = \sum\limits_{K\in \mathcal{T}_{h}} \int \nolimits _{K}\nabla v\cdot \nabla \varphi \,\mathrm{d}x& -\int \nolimits _{\mathcal{F}_{h}^{I}}\langle \nabla v\rangle \cdot \mathbf{n}[\varphi ]\,\mathrm{d}S - \Theta \int \nolimits _{\mathcal{F}_{h}^{I}}\langle \nabla \varphi \rangle \cdot \mathbf{n}[v]\,\mathrm{d}S \\ & -\int \nolimits _{\mathcal{F}_{h}^{D}}\nabla v\cdot \mathbf{n}\varphi \,\mathrm{d}S - \Theta \int \nolimits _{\mathcal{F}_{h}^{D}}\nabla \varphi \cdot \mathbf{n}v\,\mathrm{d}S.\end{array}$$

Interior and boundary penalty jump terms:

$$J_{h}(v,\varphi ) = \int \nolimits _{\mathcal{F}_{h}^{I}}\sigma [v][\varphi ]\,\mathrm{d}S + \int \nolimits _{\mathcal{F}_{h}^{D}}\sigma v\varphi \,\mathrm{d}S.$$

Right-hand side form:

$$l_{h}(\varphi )(t)\,=\,\int \nolimits _{\Omega }g(t)\varphi \,\mathrm{d}x+\int \nolimits _{\mathcal{F}_{h}^{N}}g_{N}(t)\varphi \,\mathrm{d}S-\varepsilon \Theta \int \nolimits _{\mathcal{F}_{h}^{D}}\nabla \varphi \cdot \mathbf{n}u_{D}(t)\,\mathrm{d}S+\varepsilon \int \nolimits _{\mathcal{F}_{h}^{D}}\sigma u_{D}(t)\varphi \,\mathrm{d}S.$$

The parameter σ in the diffusion and right-hand side forms is defined by \(\sigma \vert _{\Gamma } = C_{W}\vert \Gamma {\vert }^{-1}\), where C W  > 0 is a constant, which is chosen large enough to ensure coercivity of the diffusion form – cf. Lemma 2. Depending on the value of Θ in the diffusion form, we get the symmetric (Θ = 1), incomplete (Θ = 0) and nonsymmetric interior penalty \((\Theta = -1)\) variants of the diffusion a right-hand side forms.

Finally we define the convective form

$$b_{h}(v,\varphi ) = -\sum\limits_{K\in \mathcal{T}_{h}} \int \nolimits _{K}\mathbf{f}(v)\cdot \nabla v\,\mathrm{d}x+\int \nolimits _{\mathcal{F}_{h}^{I}}H({v}^{(L)},{v}^{(R)},\mathbf{n})[\varphi ]\,\mathrm{d}S+\int \nolimits _{\mathcal{F}_{h}^{B}}H({v}^{(L)},{v}^{(R)},\mathbf{n}){\varphi }^{(L)}\,\mathrm{d}S.$$

The form b h approximates convective terms with the aid of a numerical flux H(v, w, n). We assume that H has the following standard properties: H is Lipschitz-continuous, consistent, conservative and H is an E-flux, i.e.

$$\begin{array}{rcl} \big{(}H(v,w,\mathbf{n}) -\mathbf{f}(q)\cdot \mathbf{n}\big{)}(v - w) \geq 0,\quad \forall v,\,w \in \mathbb{R},\ \mathbf{n} \in B_{1}\text{ and all }q\text{ between }v,w.& & \\ \end{array}$$

The E-flux condition was introduced as a generalization of monotone fluxes by Osher in [5]. Many numerical fluxes used in practice are E-fluxes, e.g. Lax-Friedrichs, Godunov, Engquist-Osher and the Roe flux with entropy fix, cf. [5].

Definition 1.

We say that \(u_{h} \in {C}^{1}([0,T];S_{h})\) is a DG solution of (1) and (2), if \(u_{h}(0) = u_{h}^{0} \approx {u}^{0}\) and for all \(\varphi _{h} \in S_{h},\) and t ∈ (0, T)

$$\frac{d} {dt}\big{(}u_{h}(t),\varphi _{h}\big{)} + b_{h}\big{(}u_{h}(t),\varphi _{h}\big{)} + \varepsilon J_{h}\big{(}u_{h}(t),\varphi _{h}\big{)} + \varepsilon a_{h}\big{(}u_{h}(t),\varphi _{h}\big{)} = l_{h}\big{(}\varphi _{h}\big{)}(t).$$
(7)

3 Some Necessary Results

We assume that the weak solution u is sufficiently regular, namely \(u_{t} := \tfrac{\partial u} {\partial t} \in {L}^{2}\big{(}0,T;{H}^{p+1}(\Omega )\big{)}\), \(u \in {L}^{\infty }(0,T;{W}^{1,\infty }(\Omega )),\) where p ≥ 1 is the degree of approximation. These conditions imply \(u \in C\big{(}[0,T];{H}^{p+1}(\Omega )\big{)}\).

As for the mesh assumptions, we consider a system \(\{\mathcal{T}_{h}\}_{h\in (0,h_{0})}\), h 0 > 0, of triangulations, which are shape regular and satisfy the inverse assumption, cf. [2].

Now, for \(v \in {L}^{2}(\Omega )\) we denote by Π h v the \({L}^{2}(\Omega )\)-projection of v on S h :

$$\Pi _{h}v \in S_{h},\quad \left (\Pi _{h}v - v,\,\varphi _{h}\right ) = 0,\qquad \forall \,\varphi _{h} \in S_{h}.$$

Let \(\eta _{h}(t) = u(t) - \Pi _{h}u(t) \in {H}^{p+1}(\Omega ,\mathcal{T}_{h})\) and \(\xi _{h}(t) = \Pi _{h}u(t) - u_{h}(t) \in S_{h}\) for t ∈ (0, T). Then we can write the error e h as \(e_{h}(t) := u(t) - u_{h}(t) = \eta _{h}(t) + \xi _{h}(t)\). Standard approximation results give us estimates for \(\eta _{h}(t)\) in terms of power of h, e.g. \(\vert \vert \eta \vert \vert _{{L}^{2}(\Omega )} \leq C{h}^{p+1}\vert u\vert _{{H}^{p+1}}\), cf. [2].

Lemma 1.

There exists a constant C ≥ 0 independent of h,t, such that

$$b_{h}\big{(}u_{h}(t),\xi _{h}(t)\big{)}-b_{h}\big{(}u(t),\xi _{h}(t)\big{)} \leq C\Big{(}1+ \frac{\|e_{h}(t)\|_{\infty }^{2}} {{h}^{2}} \Big{)}\big{(}{h}^{2p+1}\vert u(t)\vert _{{ H}^{p+1}}^{2}+\|\xi _{ h}{(t)\|}^{2}\big{)}.$$

Proof.

The proof follows the arguments of [7], where similar estimates are derived for periodic boundary conditions or compactly supported solutions. The proof for mixed Dirichlet-Neumann boundary conditions is contained in [4]. Here, we only note that the estimate is based on performing second order Taylor expansions of and using the E-flux properties for H. □ 

Remark 1.

We can improve Lemma (1), if we suppose \(\mathbf{f} \in {(C_{b}^{3}(\mathbb{R}))}^{d}\) and \(\Gamma _{N} = \varnothing \). Then we obtain a factor of \({h}^{-1}\|e_{h}\|_{\infty }^{2}\) instead of \({h}^{-2}\|e_{h}\|_{\infty }^{2}\) in the estimate of Lemma (1). This improved estimate will be useful in proving the resulting estimates for lower order polynomials and with a less restrictive CFL condition, cf. Remark 3.

Lemma 2 (Ellipticity and boundedness of A h , cf. [3]). 

Let the constant C W be large enough. Then the form A h is elliptic and bounded , i.e.

$$\begin{array}{ll} \|v\|_{DG}^{2} & \leq A_{h}(v,v),\quad \forall v \in {H}^{2}(\Omega ,\mathcal{T}_{h}), \\ A_{h}(v,w)& \leq \| v\|_{DG}\|w\|_{DG},\quad \forall v,w \in {H}^{2}(\Omega ,\mathcal{T}_{h}), \end{array}$$

where \(\Vert w\Vert _{DG}^{2} = \tfrac{1} {2}\big{(}\sum\limits_{K\in \mathcal{T}_{h}}\vert w\vert _{{H}^{k}(K)}^{2} + J_{h}(w,w)\big{)}\) and \(A_{h}(\cdot ,\cdot ) = a_{h}(\cdot ,\cdot ) + J_{h}(\cdot ,\cdot )\) .

4 Error Analysis for the Method of Lines

We proceed in a standard way. Due to Galerkin orthogonality, we subtract the equations for u and u h and set \(\varphi _{h} := \xi _{h}(t) \in S_{h}\). Since \(\big{(}\tfrac{\partial \xi _{h}} {\partial t} ,\,\xi _{h}\big{)} = \tfrac{1} {2}\, \tfrac{\mathrm{d}} {\mathrm{d}t}\,\Vert \xi {_{h}\|}^{2},\) we get

$$\begin{array}{ll} &\frac{1} {2}\, \frac{\mathrm{d}} {\mathrm{d}t}\,\|\xi _{h}{(t)\|}^{2} + \varepsilon A_{ h}\big{(}\xi _{h}(t),\xi _{h}(t)\big{)} \\ & = -\varepsilon A_{h}\big{(}\eta _{h}(t),\xi _{h}(t)\big{)} + b_{h}\big{(}u_{h}(t),\xi _{h}(t)\big{)} - b_{h}\big{(}u(t),\xi _{h}(t)\big{)} -\Big{(}\frac{\partial \eta _{h}(t)} {\partial t} ,\,\xi _{h}(t)\Big{)}.\end{array}$$

For the last right-hand side term, we use the Cauchy and Young’s inequalities and standard estimates for η. For the convective and diffusion terms we use Lemmas 1 and 2. Integration from 0 to t ∈ [0, T] yields

$$\begin{array}{rcl} & & \|\xi _{h}{(t)\|}^{2} + \int \nolimits _{0}^{t}\varepsilon \|\xi _{ h}({\vartheta})\|_{DG}^{2}\,\mathrm{d}{\vartheta} \\ & & \leq C\int \nolimits _{0}^{t}\Big{(}1 + \frac{\|e_{h}({\vartheta})\|_{\infty }^{2}} {{h}^{2}} \Big{)}\Big{(}\big{(}{h}^{2p+1} + \varepsilon {h}^{2p}\big{)}\vert u({\vartheta})\vert _{{ H}^{p+1}}^{2} + {h}^{2p+2}\vert u_{ t}({\vartheta})\vert _{{H}^{p+1}}^{2} +\| \xi _{ h}{({\vartheta})\|}^{2}\Big{)}\,\mathrm{d}{\vartheta}.\end{array}$$
(12)

For simplicity we have assumed ξ h (0) = 0, i.e. \(u_{h}^{0} = \Pi _{h}{u}^{0}\). Otherwise, we must assume e.g. \(\|\xi _{h}(0)\| = O({h}^{p+1/2})\) and include this term in (12). We notice that if we knew apriori that \(\|e_{h}\|_{\infty } = O(h)\) then the unpleasant term \({h}^{-2}\|e_{h}\|_{\infty }^{2}\) in (12) would be O(1). Thus we could simply apply the standard Gronwall inequality to obtain the desired error estimates.

Lemma 3.

Let t ∈ [0,T] and p ≥ d∕2. If \(\|e_{h}({\vartheta})\| \leq {h}^{1+d/2}\) for all \({\vartheta} \in [0,t]\) , then there exists a constant C T independent of h,t and \(\varepsilon \) such that

$$\max _{{\vartheta}\in [0,t]}\|e_{h}{({\vartheta})\|}^{2} + \int \nolimits _{0}^{t}\varepsilon \|e_{ h}({\vartheta})\|_{DG}^{2}\,\mathrm{d}{\vartheta} \leq C_{ T}^{2}\big{(}{h}^{2p+1} + \varepsilon {h}^{2p}\big{)}.$$
(13)

Proof.

The assumptions imply, using the inverse inequality and estimates of η, that

$$\begin{array}{ll} &\|e_{h}({\vartheta})\|_{\infty }\leq \| \eta _{h}({\vartheta})\|_{\infty } +\| \xi _{h}({\vartheta})\|_{\infty }\leq Ch\vert u(t)\vert _{{W}^{1,\infty }(\Omega )} + C_{I}{h}^{-d/2}\|\xi _{h}({\vartheta})\| \\ & \leq Ch + C_{I}{h}^{-d/2}\|e_{h}({\vartheta})\| + C_{I}{h}^{-d/2}\|\eta _{h}({\vartheta})\| \leq Ch + C{h}^{p+1-d/2}\vert u({\vartheta})\vert _{{H}^{p+1}} \leq Ch, \end{array}$$

where the constant C is independent of \(h,{\vartheta},t\). Using this estimate in (12) gives us

$$\begin{array}{ll} \|&\xi _{h}{(t)\|}^{2} + \int \nolimits _{0}^{t}\varepsilon \|\xi _{h}({\vartheta})\|_{DG}^{2}\,\mathrm{d}{\vartheta} \leq \tilde{ C}\big{(}{h}^{2p+1} + \varepsilon {h}^{2p}\big{)} + C\int \nolimits _{0}^{t}\|\xi _{h}{({\vartheta})\|}^{2}\,\mathrm{d}{\vartheta}, \end{array}$$

Applying Gronwall’s inequality gives us the desired estimate for ξ h , which along with similar estimates for η gives us (13). □ 

Now it remains to get rid of the apriori assumption \(\|e_{h}\|_{\infty } = O(h)\). In [7] this is done for an explicit scheme using mathematical induction. Starting from \(\|e_{h}^{0}\| = O({h}^{p+1/2})\), the following induction step is proved:

$$\|e_{h}^{n}\| = O({h}^{p+1/2})\quad \Longrightarrow\quad \|e_{ h}^{n+1}\|_{ \infty } = O(h)\quad \Longrightarrow\quad \|e_{h}^{n+1}\| = O({h}^{p+1/2}).$$
(16)

For the method of lines we have no discrete structure with respect to time and hence cannot use mathematical induction straightforwardly. However, we can divide [0, T] into a finite number of sufficiently small intervals \([t_{n},t_{n+1}]\) on which “e h does not change too much” and use induction with respect to n. This is essentially a continuous mathematical induction argument, a concept introduced in [1].

Remark 2.

Due to the regularity assumptions, \(u,u_{h} \in C([0,T];{L}^{2}(\Omega ))\). Since [0, T] is a compact set, \(e_{h}(\cdot )\) is a uniformly continuous function from [0, T] to L 2(Ω), i.e.

$$\forall \bar{\varepsilon } > 0\;\exists \varDelta > 0 :\; s,\bar{s} \in [0,T],\vert s -\bar{ s}\vert \leq \varDelta \Rightarrow \| e_{h}(s) - e_{h}(\bar{s})\| \leq \bar{ \varepsilon }.$$

Theorem 1 (Main theorem). 

Let \(p > 1 + d/2\) . Then there exists h 1 > 0 such that for all h ∈ (0,h 1 ] we have the estimate

$$\max\limits _{{\vartheta}\in [0,T]}\|e_{h}{({\vartheta})\|}^{2} + \int \nolimits _{0}^{T}\varepsilon \|e_{ h}({\vartheta})\|_{DG}^{2}\,\mathrm{d}{\vartheta} \leq C_{ T}^{2}\big{(}{h}^{2p+1} + \varepsilon {h}^{2p}\big{)}.$$

Proof.

We have \(p > 1 + d/2\), thus for all sufficiently small h, we have \(C_{T}({h}^{p+1/2} + \sqrt{\varepsilon }{h}^{p}) \leq \tfrac{1} {2}{h}^{1+d/2}\). We fix an arbitrary h. By Remark 2, there exists δ > 0, such that if \(s,\bar{s} \in [0,T],\vert s -\bar{ s}\vert \leq \varDelta \), then \(\|e_{h}(s) - e_{h}(\bar{s})\| \leq \tfrac{1} {2}{h}^{1+d/2}\). We define \(t_{i} = i\varDelta ,\,i = 0,1,\ldots \) and set \(N :=\max \{ i = 0,1,\ldots ;t_{i} < T\}\), \(t_{N+1} := T\). This defines a partition \(0 = t_{0} < t_{1} < \cdots < t_{N+1} = T\) of [0, T] into N + 1 intervals of length (at most) δ.

We shall now prove by induction that for all \(n = 1,\ldots ,N + 1\)

$$\max \limits_{{\vartheta}\in [0,t_{n}]}\|e_{h}{({\vartheta})\|}^{2} + \int \nolimits _{0}^{t_{n} }\varepsilon \|e_{h}({\vartheta})\|_{DG}^{2}\,\mathrm{d}{\vartheta} \leq C_{ T}^{2}\big{(}{h}^{2p+1} + \varepsilon {h}^{2p}\big{)}.$$
(19)

The desired error estimate is thus obtained by taking \(n := N + 1\) in (19).

  1. (i)

    n = 1: Since \(\|e_{h}(0)\| \leq \tfrac{1} {2}{h}^{1+d/2}\). By uniform continuity, we have for all s ∈ [0, t 1]

    $$\|e_{h}(s)\| \leq \| e_{h}(0)\| +\| e_{h}(s) - e_{h}(0)\| \leq \tfrac{1} {2}{h}^{1+d/2} + \tfrac{1} {2}{h}^{1+d/2} = {h}^{1+d/2}.$$

    Therefore, by Lemma 3 we obtain estimate (19) on [0, t 1], i.e. for n = 1.

  2. (ii)

    Induction step: We assume that (19) holds for general n < N + 1. Therefore \(\|e_{h}(t_{n})\| \leq C_{T}({h}^{p+1/2} + \sqrt{\varepsilon }{h}^{p}) \leq \tfrac{1} {2}{h}^{1+d/2}\). By uniform continuity, for all \(s \in [t_{n},t_{n+1}]\)

$$\|e_{h}(s)\| \leq \| e_{h}(t_{n})\| +\| e_{h}(s) - e_{h}(t_{n})\| \leq \tfrac{1} {2}{h}^{1+d/2} + \tfrac{1} {2}{h}^{1+d/2} = {h}^{1+d/2}.$$

This and the induction assumption imply that \(\|e_{h}(s)\| \leq {h}^{1+d/2}\) for \(s \in [0,t_{n}] \cup [t_{n},t_{n+1}] = [0,t_{n+1}]\). By Lemma 3, we obtain estimate (19) on [0, t n + 1]. □ 

Remark 3.

If we assume \(\mathbf{f} \in {(C_{b}^{3}(\mathbb{R}))}^{d}\) then by Remark 1 we get the improved assumption \(p > (1 + d)/2\) in Theorem 1. If \(\varepsilon = 0\) we need to assume only p > d ∕ 2.

Remark 4.

For the method of lines we can use a nonlinear Gronwall-type lemma to prove Theorem 1 directly, cf. [4]. As stated in Remark 6, this is not possible for an implicit scheme, since an analogous discrete Gronwall lemma cannot exist.

5 Error Estimates for a Fully Implicit Scheme

In this section, we shall introduce and analyze the DG scheme with a standard implicit Euler time discretization. Here we cannot use the approach of [7] for the explicit scheme, since we were unable to prove the first implication in the induction step (16). On the other hand, in Lemma 6 we prove that for the implicit Euler scheme we cannot use a discrete Gronwall-type lemma as mentioned in Remark 4.

We consider a partition \(0 = t_{0} < t_{1} < \cdots < t_{N+1} = T\) of [0, T] and set \(\tau _{n} = t_{n+1} - t_{n}\) for \(n = 0,\cdots \,,N\). The exact solution u(t n ) will be approximated by \(u_{h}^{n} \in S_{h}\).

Definition 2.

We say that \(\{u_{h}^{n}\}_{n=0}^{N} \subset S_{h}\) is an implicit Euler DGFE solution of the convection-diffusion problem (1) and (2), if \(u_{h}^{0} = \Pi _{h}{u}^{0}\) and for all \(\varphi _{h} \in S_{h},n = 0,\cdots \,,N\)

$$\Big{(}\frac{u_{h}^{n+1} - u_{h}^{n}} {\tau _{n}} ,\varphi _{h}\Big{)} + b_{h}\big{(}u_{h}^{n+1},\varphi _{ h}\big{)} + \varepsilon A_{h}\big{(}u_{h}^{n+1},\varphi _{ h}\big{)} = l_{h}\big{(}\varphi _{h}\big{)}(t_{n+1}).$$
(22)

Similarly as in Sect. 3, we define \(\eta _{h}^{n} = u(t_{n}) - \Pi _{h}u(t_{n}) \in {H}^{p+1}(\Omega ,\mathcal{T}_{h})\) and \(\xi _{h}^{n} = \Pi _{h}u(t_{n}) - u_{h}^{n} \in S_{h}\). Then we can write the error \(e_{h}^{n}\) as \(e_{h}^{n} := u(t_{n}) - u_{h}^{n} = \eta _{h}^{n} + \xi _{h}^{n}\).

First, we analyze problem (22), proving that \(u_{h}^{n+1}\) exists uniquely and depends continuously on τ n . To this end we define an abstract formulation of problem (22):

Definition 3.

(Auxiliary problem) Let t ∈ [0, T], τ ∈ [0, T] and \(U_{h} \in S_{h}\). We seek \(u_{\tau } \in S_{h}\) such that

$$\big{(}u_{\tau } - U_{h},\varphi _{h}\big{)} + \tau b_{h}\big{(}u_{\tau },\varphi _{h}\big{)} + \tau \varepsilon A_{h}\big{(}u_{\tau },\varphi _{h}\big{)} = \tau l_{h}\big{(}\varphi _{h}\big{)}(t),\quad \forall \varphi _{h} \in S_{h}.$$
(23)

Remark 5.

If we take \(\tau := \tau _{n},U_{h} := u_{h}^{n},t := t_{n+1}\) and define \(u_{h}^{n+1} := u_{\tau }\), the auxiliary problem (23) reduces to equation (22), which defines \(u_{h}^{n+1}\). If we take τ : = 0 the solution of (23) is \(u_{\tau } = u_{h}^{n}\). Between these two cases \(u_{\tau }\) depends continuously on τ:

Lemma 4.

There exist constants \(C_{1},C_{2} > 0\) independent of \(h,\tau ,t,\varepsilon \) , such that the following holds. Let \(t \in [0,T],h \in (0,h_{0}),U_{h} \in S_{h}\) and τ ∈ [0,τ 0 ), where \(\tau _{0} =\max \{ C_{1}\varepsilon ,C_{2}h\}\) . Then the solution u τ of (23) exists, is uniquely determined and \(\|u_{\tau }\|\) depends continuously on \(\tau \in [0,\tau _{0})\) .

Proof.

Problem (23) is a nonlinear equation for u τ on the finite-dimensional space S h . The statements follow from the nonlinear Lax-Milgram theorem, cf. [6]. For details of the proof, see [4]. □ 

Definition 4 (Continuated discrete solution). 

Let \(\tilde{u}_{h} : [0,T] \rightarrow S_{h}\) such that for \(s \in [t_{n},t_{n+1}]\) we set \(\tilde{u}_{h}(s) := u_{\tau }\), the solution of the auxiliary problem (23) with \(\tau := s - t_{n}\), t : = t n + 1 and \(U_{h} := u_{h}^{n}\). Furthermore, we define \(\tilde{e}_{h} := u -\tilde{ u}_{h}\) and \(\tilde{\xi }_{h} := \Pi _{h}u -\tilde{ u}_{h}\).

Under the assumptions of Lemma 4, \(\tilde{u}_{h},\tilde{e}_{h} \in C([0,T];{L}^{2}(\Omega ))\) and \(\tilde{u}_{h}\) is uniquely determined. Also, \(\tilde{u}_{h}(t_{n}) = u_{h}^{n}\) and \(\tilde{e}_{h}(t_{n}) = e_{h}^{n}\) for \(n = 0,\cdots \,,N\). Therefore, estimates of \(\tilde{e}_{h}(\cdot )\) imply estimates of \(e_{h}^{n}\). Since \(\tilde{u}_{h}\) is constructed using problem (23), which is essentially the implicit scheme (22) with special data, we can derive error estimates for \(\tilde{u}_{h}\) in a standard manner. For simplicity we assume a uniform partition of [0, T].

Lemma 5.

Let p > d∕2 and \(s \in (t_{n},t_{n+1}]\) for some \(n \in \{ 0,\cdots \,,N - 1\}\) . If \(\|\tilde{e}_{h}(s)\| \leq {h}^{1+d/2}\) and \(\|\tilde{e}_{h}(t_{k})\| \leq {h}^{1+d/2}\) for all \(k = 0,\cdots \,,n,\) then there exists C T > 0 independent of s,n,h,τ such that

$$\max \limits_{t\in \{t_{0},\cdots \,,t_{n},s\}}\|\tilde{e}_{h}{(t)\|}^{2}+\sum\limits_{k=1}^{n}\tau \varepsilon \|\tilde{e}_{ h}(t_{k})\|_{DG}^{2}+(s-t_{ n})\varepsilon \|\tilde{e}_{h}(s)\|_{DG}^{2} \leq C_{ T}^{2}\big{(}{h}^{2p+1}+\varepsilon {h}^{2p}+{\tau }^{2}\big{)}.$$

Proof.

We subtract (23) from the equation for the exact solution. Thus \(\tilde{e}_{h}(s)\) satisfies

$$\begin{array}{rcl} \big{(}\tilde{e}_{h}(s)-& \tilde{e}_{h}(t_{n}),\varphi _{h}\big{)} + (s - t_{n})\big{(}b_{h}(u(s),\varphi _{h}) - b_{h}(\tilde{u}_{h}(s),\varphi _{h})\big{)} + (s - t_{n})\varepsilon A_{h}(\tilde{e}_{h}(s),\varphi _{h})& \\ & = \big{(}u(s) - u(t_{n}) - (s - t_{n})u_{t}(s),\varphi _{h}\big{)}. &\end{array}$$
(25)

We set \(\varphi _{h} :=\tilde{ \xi }_{h}(s)\) and use the fact that \(2(a - b,a) =\| {a\|}^{2} -\| {b\|}^{2} +\| a - {b\|}^{2}\). We estimate the convective terms using Lemma 1 and the diffusion terms using Lemma 2. The right-hand side represents the temporal error and is estimated as usual. Thus

$$\begin{array}{rcl} & & \|\tilde{\xi }_{h}{(s)\|}^{2} -\|\tilde{ \xi }_{ h}{(t_{n})\|}^{2} +\|\tilde{ \xi }_{ h}(s) -\tilde{ \xi }_{h}{(t_{n})\|}^{2} + (s - t_{ n})\varepsilon \|\tilde{\xi }_{h}(s)\|_{DG}^{2} \\ & & \leq C\tau \Big{(}1 + \frac{\|\tilde{e}_{h}(s)\|_{\infty }^{2}} {{h}^{2}} \Big{)}\Big{(}\big{(}{h}^{2p+1} + \varepsilon {h}^{2p})\big{\vert }u\vert _{{ L}^{\infty }({H}^{p+1})}^{2} + {\tau }^{2}\|u_{ tt}\|_{{L}^{\infty }({L}^{2}(\Omega )))}^{2} +\|\tilde{ \xi }_{ h}{(s)\|}^{2}\Big{)}.\end{array}$$

The assumptions imply \(\|\tilde{e}_{h}(s)\|_{\infty }\leq Ch\), eliminating the factor h  − 2. Thus

$$\|\tilde{\xi }_{h}{(s)\|}^{2} + (s - t_{ n})\varepsilon \|\tilde{\xi }_{h}(s)\|_{DG}^{2} \leq \|\tilde{ \xi }_{ h}{(t_{n})\|}^{2} + C\tau \big{(}{h}^{2p+1} + \varepsilon {h}^{2p} + {\tau }^{2} +\|\tilde{ \xi }_{ h}{(s)\|}^{2}\big{)}.$$

Similarly, we may derive estimates at t k + 1:

$$\|\tilde{\xi }_{h}{(t_{k+1})\|}^{2} + \tau \varepsilon \|\tilde{\xi }_{ h}(t_{k+1})\|_{DG}^{2} \leq \|\tilde{ \xi }_{ h}{(t_{k})\|}^{2} + C\tau \big{(}{h}^{2p+1} + \varepsilon {h}^{2p} + {\tau }^{2} +\|\tilde{ \xi }_{ h}{(t_{k+1})\|}^{2}\big{)}.$$

Combining these estimates and using the discrete Gronwall lemma gives us the desired estimate for \(\tilde{\xi }_{h}\). Standard estimates for η give us the estimate for \(\tilde{e}_{h}\). □ 

Theorem 2 (Main theorem – implicit version). 

Let \(p > 1 + d/2\) . Let \(h_{1},\tau _{1} > 0\) be such that \(C_{T}(h_{1}^{p+1/2} + \sqrt{\varepsilon }h_{1}^{p} + \tau _{1}) = \tfrac{1} {2}h_{1}^{1+d/2}\) and \(\tau _{1} < \tau _{0}\) , where τ 0 is defined in Lemma  4. Then for all \(h \in (0,h_{1}),\tau \in (0,\tau _{1})\) we have the estimate

$$\max \limits_{n\in \{0,\cdots \,,N\}}\|e_{h}^{{n}}\|^{2}+\sum\limits_{n=1}^{N}\tau \Big{(}\varepsilon \|e_{ h}^{n}\|_{ DG}^{2}+\widetilde{J}_{ h}\big{(}e_{h}^{n},e_{ h}^{n}\big{)}\Big{)} \leq C_{ T}^{2}\big{(}{h}^{2p+1}+\varepsilon {h}^{2p}+{\tau }^{2}\big{)}.$$
(28)

Proof.

Again, \(\tilde{e}_{h}(\cdot )\) is a uniformly continuous function from [0, T] to \({L}^{2}(\Omega )\). This allows to use continuous mathematical induction to eliminate the apriori assumption \(\|\tilde{e}_{h}(t)\| = O({h}^{1+d/2})\) from Lemma 5. The proof thus follows that of Theorem 1. □ 

Remark 6.

The reason we introduced the continuation of \(u_{h}^{n}\) is that a more standard, straightforward approach is insufficient. Specifically, we prove in [4] that there does not exist a Gronwall-type lemma which could prove the desired error estimate (28) only from the error equation of the implicit scheme tested by \(\xi _{h}^{n+1}\) and the derived estimates of individual terms contained therein.

6 Conclusion

We have presented an analysis of the DG method for a nonlinear convection-diffusion problem. Building on results from [7], which dealt with an explicit time discretization, we proved apriori \({L}^{\infty }({L}^{2})\) error estimates independent of the diffusion coefficient for the method of lines and a fully implicit scheme. We have derived the key estimates for the case of mixed Dirichlet-Neumann boundary conditions, improving the results of [7]. For the method of lines, the error estimates are derived using a continuous mathematical induction argument or a nonlinear Gronwall lemma. For the implicit time discretization, we show that a similar discrete Gronwall lemma does not exist and prove the error estimates using continuous mathematical induction applied to a suitable continuation of the discrete solution. However, using this technique, we obtain an unnatural CFL-like condition for the implicit scheme. In [4], the presented results are extended to of a locally Lipschitz continuous f.