A stabilized finite element method for inverse problems subject to the convection-diffusion equation. II: convection-dominated regime

Abstract

We consider the numerical approximation of the ill-posed data assimilation problem for stationary convection–diffusion equations and extend our previous analysis in [Numer. Math. 144, 451–477, 2020] to the convection-dominated regime. Slightly adjusting the stabilized finite element method proposed for dominant diffusion, we draw upon a local error analysis to obtain quasi-optimal convergence along the characteristics of the convective field through the data set. The weight function multiplying the discrete solution is taken to be Lipschitz and a corresponding super approximation result (discrete commutator property) is proven. The effect of data perturbations is included in the analysis and we conclude the paper with some numerical experiments.

Authors

Erik Burman
Department of Mathematics, University College London, Gower Street, London UK, WC1E 6BT

Mihai Nechita
Department of Mathematics, University College London, Gower Street, London UK, WC1E 6BT

Lauri Oksanen
Department of Mathematics, University College London, Gower Street, London UK, WC1E 6BT

Keywords

convection-diffusion equation, convection-dominated, unique continuation, ill-posed problem

Paper coordinates

E. Burman, M. Nechita, L. Oksanen, A stabilized finite element method for inverse problems subject to the convection-diffusion equation. II: convection-dominated regime, Numer. Math., 150:769-801, 2022, DOI: https://doi.org/10.1007/s00211-022-01268-1

PDF

About this paper

Journal

Numerische Mathematik

Publisher Name

Springer

Print ISSN

0029-599X

Online ISSN

0945-3245

[1] M. S. Alnæs, J. Blechta, J. Hake, A. Johansson, B. Kehlet, A. Logg, C. Richardson, J. Ring, M. E. Rognes, and G. N. Wells. The FEniCS Project Version 1.5. Archive of Numerical Software, 3(100), 2015.
[2] S. Bertoluzza. The discrete commutator property of approximation spaces. C. R. Acad. Sci.Paris Ser. I Math., 329(12):1097–1102, 1999.
[3]  E. Burman, J. Guzman, and D. Leykekhman. Weighted error estimates of the continuous interior penalty method for singularly perturbed problems. IMA J. Numer. Anal., 29(2):284– 314, 2009.
[4] E. Burman, P. Hansbo, and M. G. Larson. Solving ill-posed control problems by stabilized finite element methods: an alternative to Tikhonov regularization. Inverse Problems, 34:035004, 2018.
[5]  E. Burman, M. Nechita, and L. Oksanen. A stabilized finite element method for inverse problems subject to the convection–diffusion equation. I: diffusion-dominated regime. Numer. Math., 144(451–477), 2020.
[6] E. Burman. A unified analysis for conforming and nonconforming stabilized finite element methods using interior penalty. SIAM J. Numer. Anal., 43(5):2012–2033, 2005.
[7]  E. Burman. Stabilized finite element methods for nonsymmetric, noncoercive, and ill-posed problems. Part I: Elliptic equations. SIAM J. Sci. Comput., 35(6):A2752–A2780, 2013.
[8]  E. Burman. Stabilized finite element methods for nonsymmetric, noncoercive, and ill-posed problems. Part II: Hyperbolic equations. SIAM J. Sci. Comput., 36(4):A1911–A1936, 2014.
[9]  A. Ern and J.-L. Guermond. Theory and practice of finite elements, volume 159 of Applied Mathematical Sciences. Springer-Verlag, New York, 2004.
[10]  L. C. Evans. Partial differential equations, volume 19 of Graduate Studies in Mathematics. American Mathematical Society, 2010.
[11]  V. John, P. Knobloch, and J. Novo. Finite elements for scalar convection-dominated equations and incompressible flow problems: a never ending story? Comput. Vis. Sci., 19(5):47– 63, Dec 2018.
[12]  C. Johnson, U. Navert, and J. Pitkaranta. Finite element methods for linear hyperbolic problems. Comput. Methods Appl. Mech. Engrg., 45(1-3):285–312, 1984.
[13]  P. Monk and E. Suli. The adaptive computation of far-field patterns by a posteriori error estimation of linear functionals. SIAM J. Numer. Anal., 36(1):251–274, 1999.
[14]  H.-G. Roos and M. Stynes. Some open questions in the numerical analysis of singularly perturbed differential equations. Comput. Methods Appl. Math., 15(4):531–550, 2015.
[15]  E. M Stein. Singular integrals and differentiability properties of functions, volume 30 of Princeton Mathematical Series. Princeton University Press, 1970.

A STABILIZED FINITE ELEMENT METHOD FOR INVERSE PROBLEMS SUBJECT TO THE CONVECTION-DIFFUSION EQUATION. II: CONVECTION-DOMINATED REGIME

ERIK BURMAN, MIHAI NECHITA, AND LAURI OKSANEN
Abstract

We consider the numerical approximation of the ill-posed data assimilation problem for stationary convection-diffusion equations and extend our previous analysis in [Numer. Math. 144, 451-477, 2020] to the convection-dominated regime. Slightly adjusting the stabilized finite element method proposed for dominant diffusion, we draw upon a local error analysis to obtain quasi-optimal convergence along the characteristics of the convective field through the data set. The weight function multiplying the discrete solution is taken to be Lipschitz continuous and a corresponding super approximation result (discrete commutator property) is proven. The effect of data perturbations is included in the analysis and we conclude the paper with some numerical experiments.

1. Introduction

In this work, we consider a data assimilation problem for a stationary convection diffusion equation

u:=μΔu+βu=f in Ωn,\mathcal{L}u:=-\mu\Delta u+\beta\cdot\nabla u=f\quad\text{ in }\Omega\subset\mathbb{R}^{n}, (1)

when convection dominates, that is 0<μ|β|0<\mu\ll|\beta|, and complement the diffusion dominated case discussed in the first part [BNO20]. We assume that Ωn\Omega\subset\mathbb{R}^{n} is open, bounded and connected, and there exists a solution uH2(Ω)u\in H^{2}(\Omega) to (1). The problem under study is to approximate the solution uu given the source ff in Ω\Omega and the perturbed restriction U~ω=u|ω+δ\tilde{U}_{\omega}=\left.u\right|_{\omega}+\delta of the solution to an open subset ωΩ\omega\subset\Omega. The perturbation δ\delta is taken in L2(ω)L^{2}(\omega). Notice that we consider no boundary conditions on Ω\partial\Omega. This is a linear ill-posed problem also known as unique continuation.

To start with, let us briefly recall the main results obtained in the first part. Consider an open bounded set BΩB\subset\Omega that contains the data region ω\omega such that B\ωB\backslash\omega does not touch the boundary of Ω\Omega. For uH1(Ω)u\in H^{1}(\Omega), the following conditional stability estimate was proven for μ>0\mu>0 and βL(Ω)n\beta\in L^{\infty}(\Omega)^{n},

uL2(B)Cst(uL2(ω)+1μuH1(Ω))κuL2(Ω)1κ\|u\|_{L^{2}(B)}\leq C_{st}\left(\|u\|_{L^{2}(\omega)}+\frac{1}{\mu}\|\mathcal{L}u\|_{H^{-1}(\Omega)}\right)^{\kappa}\|u\|_{L^{2}(\Omega)}^{1-\kappa} (2)

where the Hölder exponent κ(0,1)\kappa\in(0,1) depends only on the geometric setting. In the case of simple geometric configurations, e.g. when ω,B,Ω\omega,B,\Omega are three concentric balls, the exponent κ(0,1)\kappa\in(0,1) can be given explicitly, see [BNO20, Remark 1]. The stability constant CstC_{st} is given explicitly in terms of the physical parameters

Cst=C1exp(C2(1+|β|μ)2),|β|:=βL(Ω)nC_{st}=C_{1}\exp\left(C_{2}\left(1+\frac{|\beta|}{\mu}\right)^{2}\right),\quad|\beta|:=\|\beta\|_{L^{\infty}(\Omega)^{n}} (3)
00footnotetext: Department of Mathematics, University College London, Gower Street, London UK, WC1E 6BT. E-mail addresses: {e.burman, mihai.nechita.16, l.oksanen }@ucl.ac.uk.
Date: February 22, 2022.

with constants C1,C2>0C_{1},C_{2}>0 depending only on the geometry. Note that the continuum estimate (2) is valid in both the diffusion-dominated and convection-dominated regimes, and that the stability constant CstC_{st} is uniformly bounded when diffusion dominates. However, when convection dominates CstC_{st} grows exponentially, rendering the stability estimate ineffective in practice. We also recall that for global unique continuation from ω\omega to the entire Ω\Omega the stability is no longer Hölder, but logarithmic, that is the modulus of continuity for the given data is not ||κ|\cdot|^{\kappa} any more, but |log()|κ|\log(\cdot)|^{-\kappa}.

On the discrete level, the continuum estimate (2) was combined with a stabilized linear finite element method to obtain convergence orders for the approximate solution. More precisely, for a mesh size hh, and defining the Péclet number

Pe(h):=|β|hμ,Pe(h):=\frac{|\beta|h}{\mu},

the following error bound [BNO20, Theorem 1] was proven for the approximation uhu_{h} in the diffusive regime Pe(h)<1Pe(h)<1,

uuhL2(B)Csthκ(uH2(Ω)+h1δL2(ω)),\left\|u-u_{h}\right\|_{L^{2}(B)}\leq C_{st}h^{\kappa}\left(\|u\|_{H^{2}(\Omega)}+h^{-1}\|\delta\|_{L^{2}(\omega)}\right), (4)

where the convergence order κ(0,1)\kappa\in(0,1) is the same as the Hölder exponent in (2) and the stability constant CstC_{st} is proportional to the one in (3). Under an additional assumption on the divergence of the convective field β\beta, similar error bounds were also proven in the H1H^{1}-norm, see [BNO20, Theorem 2].

The prototypical effect of dominating diffusion is shown in Figure 1, where the problem is set in the unit square and contour error plots are shown for data assimilation from a centered disk of radius 0.1 . One can notice oscillating errors that grow in size away from the data region towards the boundary. The exact solution in this example is u=2sin(5πx)sin(5πy)u=2\sin(5\pi x)\sin(5\pi y) where the factor 2 is taken to make its L2L^{2}-norm unitary. For the computation we used an unstructured mesh with 512 elements on a side and mesh size h0.0025h\approx 0.0025.
1.1. Objective and main results. We consider a stabilized finite element method for data assimilation subject to the convection-diffusion equation in the convection dominated regime. Since the behaviour of the physical system changes fundamentally when convection dominates and

Pe(h)1,Pe(h)\gg 1,

the goal of this second part paper is to reconsider the numerical method proposed in the first part [BNO20] and develop an error analysis that captures and exploits the governing transport phenomenon. This is illustrated in Figure 2 where the transition to the convection-dominated regime through an intermediate regime is made by decreasing the diffusion coefficient μ\mu. We aim to obtain sharper local error estimates along the characteristics of the convective field through the data region. Even though the error analysis that we perform herein is different in nature to the one in the first part, the numerical method itself is only slightly changed (see Remark 1 below). For the error localization technique we draw on ideas used for the streamline diffusion method in [JNP84], continuous interior penalty in [BGL09], and non-coercive hyperbolic problems in [Bur14].

From the definition of the Péclet number we see that the regime will also depend on the resolution of the computation besides the physical parameters. We can therefore expect the method to change behaviour as the resolution increases and Pe(h)Pe(h) decreases. This phenomenon was already observed computationally in [Bur13] and can now be explained theoretically.

Refer to caption
Figure 1: Figure 1. Absolute error contour plot in the diffusion-dominated case, μ=1,β=(1,0)\mu=1,\beta=(1,0). The domain is the unit square, data given in a centered disk of radius 0.1 for the exact solution u=2sin(5πx)sin(5πy)u=2\sin(5\pi x)\sin(5\pi y).
Refer to caption
Figure 2: Figure 2. Absolute error contour plot when convection becomes dominant, β=(1,0)\beta=(1,0). The domain is the unit square, data given in a centered disk of radius 0.1 for the exact solution u=2sin(5πx)sin(5πy)u=2\sin(5\pi x)\sin(5\pi y).

To make the presentation as simple as possible we consider a model case in the unit square Ω\Omega with constant convection

β:=(β1,0),β1,\beta:=\left(\beta_{1},0\right),\quad\beta_{1}\in\mathbb{R},

and the solution given in the subset

ω:=(0,x)×(y,y+)with x>h and y+y>h.\omega:=(0,x)\times\left(y^{-},y^{+}\right)\text{with }x>h\text{ and }y^{+}-y^{-}>h.

For the subset ωβΩ\omega_{\beta}\subset\Omega covered by the characteristics of β\beta that go through ω\omega, we introduce the stability region ωβωβ\stackrel_{\beta}\subset\omega_{\beta} by cutting off a crosswind layer of width 𝒪(h12|ln(h)|)\mathcal{O}\left(h^{\frac{1}{2}}|\ln(h)|\right) (see Section 2.1 for more details). We separate the convection-dominated and diffusion dominated regimes by introducing a constant Pelim >1Pe_{\text{lim }}>1 such that

Pe(h)>Pelim>1Pe(h)>Pe_{\lim}>1

To reduce the number of constants appearing in the analysis, we will write this as Pe(h)1Pe(h)\gtrsim 1. As suggested by Figure 2, we expect different results for data assimilation downstream vs upstream in an intermediate regime. We prove in Theorem 1 weighted error estimates that for β1>0\beta_{1}>0 essentially take the following form

uuhL2(ω^β)C(|β|12h32|u|H2(Ω)+|β|12h12δL2(ω)), when Pe(h)1.\left\|u-u_{h}\right\|_{L^{2}\left(\hat{\omega}_{\beta}\right)}\leq C\left(|\beta|^{\frac{1}{2}}h^{\frac{3}{2}}|u|_{H^{2}(\Omega)}+|\beta|^{\frac{1}{2}}h^{-\frac{1}{2}}\|\delta\|_{L^{2}(\omega)}\right),\text{ when }Pe(h)\gtrsim 1.

This is similar to the typical error estimates for piecewise linear stabilized FEMs for convection-dominated well-posed problems, such as local projection stabilization, dG methods, continuous interior penalty or Galerkin least squares. On general shape-regular meshes these methods have an 𝒪(h12)\mathcal{O}\left(h^{\frac{1}{2}}\right) gap to the best approximation convergence order. Taking this into account, our result is thus quasi-optimal. For a recent overview of challenges and open problems in the well-posed case, see e.g. [JKN18] and [RS15].

When going against the characteristics, i.e. β1<0\beta_{1}<0, we prove in Theorem 2 first the pre-asymptotic bound

uuhL2(ω˙β)C(|β|12h|u|H2(Ω)+h1δL2(ω)), when 1Pe(h)<h1,\left\|u-u_{h}\right\|_{L^{2}\left(\dot{\omega}_{\beta}\right)}\leq C\left(|\beta|^{\frac{1}{2}}h|u|_{H^{2}(\Omega)}+h^{-1}\|\delta\|_{L^{2}(\omega)}\right),\text{ when }1\lesssim Pe(h)<h^{-1},

followed by

uuhL2(ω˙β)C(|β|12h32|u|H2(Ω)+h12δL2(ω)), when Pe(h)>h1.\left\|u-u_{h}\right\|_{L^{2}\left(\dot{\omega}_{\beta}\right)}\leq C\left(|\beta|^{\frac{1}{2}}h^{\frac{3}{2}}|u|_{H^{2}(\Omega)}+h^{-\frac{1}{2}}\|\delta\|_{L^{2}(\omega)}\right),\text{ when }Pe(h)>h^{-1}.

It follows that when solving the data assimilation problem against the flow, the diffusivity reduces the convergence order in an intermediate regime. Only for very small diffusion coefficients μ<|β|h2\mu<|\beta|h^{2} do we get quasi-optimal bounds. This asymmetry of the error distribution for moderate Péclet numbers is clearly visible in the left plot of Figure 2.

Previous results on optimal control for stabilized convection-diffusion equations include [BV07, DQ05, HYZ09, YZ09]. We refer to the first part [BNO20] for a more detailed discussion.

In terms of notation, above and throughout the paper CC denotes a generic positive constant, not necessarily the same at each occurrence, that is independent of the coefficients μ,β\mu,\beta and the mesh size hh.

2. Discrete setting

Let VhH1(Ω)V_{h}\subset H^{1}(\Omega) be the conforming finite element space of piecewise affine 1\mathbb{P}_{1} functions defined on a computational mesh 𝒯h\mathcal{T}_{h} that consists of shape-regular triangular elements KK with diameter hKh_{K}. The mesh size hh is the maximum over hKh_{K} and we will assume that h<1h<1. The interior faces of all the elements are collected in the set i\mathcal{F}_{i} and the jump of a quantity across such a face FF is denoted by F\llbracket\cdot\rrbracket_{F}, omitting the subscript whenever the context is clear. We denote by nn the unit normal.

First we introduce the standard inner products with the induced norms

(vh,wh)Ξ:=Ξvhwhdx,vh,whΞ:=Ξvhwhds\left(v_{h},w_{h}\right)_{\Xi}:=\int_{\Xi}v_{h}w_{h}\mathrm{~d}x,\quad\left\langle v_{h},w_{h}\right\rangle_{\partial\Xi}:=\int_{\partial\Xi}v_{h}w_{h}\mathrm{~d}s

and the bilinear form in the weak formulation of (1)

ah(vh,wh):=(βvh,wh)Ω+(μvh,wh)Ωμvhn,whΩ.a_{h}\left(v_{h},w_{h}\right):=\left(\beta\cdot\nabla v_{h},w_{h}\right)_{\Omega}+\left(\mu\nabla v_{h},\nabla w_{h}\right)_{\Omega}-\left\langle\mu\nabla v_{h}\cdot n,w_{h}\right\rangle_{\partial\Omega}.

We will make use of the stabilizing bilinear forms

sΩ(vh,wh):=γFiFh(μ+|β|h)vhnwhndss_{\Omega}\left(v_{h},w_{h}\right):=\gamma\sum_{F\in\mathcal{F}_{i}}\int_{F}h(\mu+|\beta|h)\llbracket\nabla v_{h}\cdot n\rrbracket\cdot\llbracket\nabla w_{h}\cdot n\rrbracket\mathrm{~d}s

which will act on the discrete solution penalizing the jumps of its gradient across interior faces, and

s(vh,wh):=γ((|β|+μh1)vh,whΩ+(μvh,wh)Ω+sΩ(vh,wh)),s_{*}\left(v_{h},w_{h}\right):=\gamma_{*}\left(\left\langle\left(|\beta|+\mu h^{-1}\right)v_{h},w_{h}\right\rangle_{\partial\Omega}+\left(\mu\nabla v_{h},\nabla w_{h}\right)_{\Omega}+s_{\Omega}\left(v_{h},w_{h}\right)\right),

where γ\gamma and γ\gamma_{*} are positive constants that can be heuristically chosen at implementation. They do not play a role in the convergence of the method and most of the time we will include them in the generic constant CC. For the data assimilation term we consider the scaled inner product in the data set ω\omega given by

sω(vh,wh):=((|β|h1+μhζ)vh,wh)ω,ζ[0,2].s_{\omega}\left(v_{h},w_{h}\right):=\left(\left(|\beta|h^{-1}+\mu h^{-\zeta}\right)v_{h},w_{h}\right)_{\omega},\quad\zeta\in[0,2].

To this we add the stabilizing interior penalty term sΩs_{\Omega} to define for conciseness

s(vh,wh):=sΩ(vh,wh)+sω(vh,wh),s\left(v_{h},w_{h}\right):=s_{\Omega}\left(v_{h},w_{h}\right)+s_{\omega}\left(v_{h},w_{h}\right),

The idea behind the computational method follows the discretize-then-optimize approach: we first discretize and then formulate the data assimilation problem as a PDEconstrained optimization problem with additional stabilizing terms. Apart from their stabilizing intake, these terms are also chosen such that they vanish at optimal rates. For an overview on this approach to ill-posed problems and more details on the desired properties of discrete stabilizers, we refer the reader to [Bur13]. To be more precise, for an approximation uhVhu_{h}\in V_{h} and a discrete Lagrange multiplier zhVhz_{h}\in V_{h}, we consider the functional

Lh(uh,zh):=\displaystyle L_{h}\left(u_{h},z_{h}\right)= 12sω(uhU~ω,uhU~ω)+ah(uh,zh)(f,zh)Ω\displaystyle\frac{1}{2}s_{\omega}\left(u_{h}-\tilde{U}_{\omega},u_{h}-\tilde{U}_{\omega}\right)+a_{h}\left(u_{h},z_{h}\right)-\left(f,z_{h}\right)_{\Omega}
+12sΩ(uh,uh)12s(zh,zh)\displaystyle+\frac{1}{2}s_{\Omega}\left(u_{h},u_{h}\right)-\frac{1}{2}s_{*}\left(z_{h},z_{h}\right)

where the first term is measuring the misfit between uhu_{h} and the known perturbed restriction U~ω=u|ω+δ\tilde{U}_{\omega}=\left.u\right|_{\omega}+\delta, the next two terms are imposing the weak form of the PDE (1) as a constraint, and the last two terms have stabilizing role and act only on the discrete level.

We look for the saddle points of the Lagrangian LhL_{h} and analyse their convergence to the exact solution. Using the optimality conditions we obtain the finite element method for data assimilation subject to (1), which reads as follows: find (uh,zh)[Vh]2\left(u_{h},z_{h}\right)\in\left[V_{h}\right]^{2} such that

{ah(uh,wh)s(zh,wh)=(f,wh)Ωah(vh,zh)+s(uh,vh)=sω(U~ω,vh),(vh,wh)[Vh]2.\left\{\begin{array}[]{rl}a_{h}\left(u_{h},w_{h}\right)-s_{*}\left(z_{h},w_{h}\right)&=\left(f,w_{h}\right)_{\Omega}\\ a_{h}\left(v_{h},z_{h}\right)+s\left(u_{h},v_{h}\right)&=s_{\omega}\left(\tilde{U}_{\omega},v_{h}\right)\end{array},\quad\forall\left(v_{h},w_{h}\right)\in\left[V_{h}\right]^{2}.\right.

Notice that the exact solution uH2(Ω)u\in H^{2}(\Omega) (with noise δ0\delta\equiv 0 ) and the dual variable z0z\equiv 0 satisfy (5) since the gradient of the exact solution has no jumps across interior faces. Hence the Lagrange multiplier zhz_{h} should converge to zero.

Remark 1. The same finite element method (5) has been proposed in the first part [BNO20] for the diffusion-dominated case; sΩs_{\Omega} and ss_{*} are exactly the stabilizing terms introduced there. However, herein we have increased the penalty coefficient in the data term sωs_{\omega} from |β|h+μ|\beta|h+\mu to |β|h1+μhζ|\beta|h^{-1}+\mu h^{-\zeta}. We note that the bounds in [BNO20] hold also for this stronger penalty term, but the sensitivity to perturbations in data increases by a factor of h1h^{-1}.

Refer to caption
Figure 3: Figure 3. Data set ω\omega (gray) and the stability region $\dot{\omega

_{\beta}$ (hatched).}

Proposition 1. The finite element method (5) has a unique solution (uh,zh)[Vh]2\left(u_{h},z_{h}\right)\in\left[V_{h}\right]^{2} and the Euclidean condition number 𝒦2\mathcal{K}_{2} of the system matrix satisfies

𝒦2Ch4.\mathcal{K}_{2}\leq Ch^{-4}.

Proof. The proof given in [BNO20, Proposition 2] holds verbatim. \square
2.1. Stability region and weight functions. We will exploit the convective term of the PDE to obtain stability in the zone that connects through characteristics to the data region ω\omega. The objective is to obtain weighted L2L^{2}-estimates in this region that are independent of μ\mu (but not of the regularity of the exact solution) with the underlying assumption that μ|β|\mu\ll|\beta|. To this end we first define the subdomain where we can obtain stability (see Figure 3 for a sketch) and some weight functions that will be used to define weighted norms. These can be given in explicit form in the simple framework where β=(β1,0)\beta=\left(\beta_{1},0\right) and

ω:=(0,x)×(y,y+)with x>h and y+y>h.\omega:=(0,x)\times\left(y^{-},y^{+}\right)\text{with }x>h\text{ and }y^{+}-y^{-}>h.

Let ωβ\omega_{\beta} denote the closed set of all the points pΩ¯p\in\bar{\Omega} that can be reached through characteristics from ω\omega, i.e. for which there exists ss\in\mathbb{R} such that p+sβωp+s\beta\in\partial\omega. Similarly to the classical work [JNP84], we define the stability region ω˙β\dot{\omega}_{\beta} by cutting off a crosswind layer from ωβ\omega_{\beta}, namely

ωβ:={pωβ:dist(p,Ω\ωβ)cλh12ln(1/h)},\stackrel_{\beta}:=\left\{p\in\omega_{\beta}:\operatorname{dist}\left(p,\Omega\backslash\omega_{\beta}\right)\geq c_{\lambda}h^{\frac{1}{2}}\ln(1/h)\right\}, (6)

with the constant cλc_{\lambda} to be made precise in the following. In our setting, we simply have that ωβ=[0,1]×[y,y+]\stackrel_{\beta}=[0,1]\times\left[\stackrel^{-},\stackrel^{+}\right]for some y+>y>0\stackrel^{+}>\stackrel^{-}>0. The crosswind layer and its width are motivated by the subsequent construction of weight functions with a specific decay outside ωβ\omega_{\beta}.

We will consider different weight functions depending on the direction of the convection field. In the downstream case we let

ψ1(x,y):=ex, when β1>0,\psi_{1}(x,y):=e^{-x},\text{ when }\beta_{1}>0,

and in the upstream case

ψ1(x,y):=ex, when β1<0.\psi_{1}(x,y):=-e^{-x},\text{ when }\beta_{1}<0.

In both cases we have that ψ1=(ψ1,0)\nabla\psi_{1}=\left(-\psi_{1},0\right). Let then ψ2W1,(Ω)\psi_{2}\in W^{1,\infty}(\Omega) be a function satisfying

ψ2={1, in ωβ𝒪(h3), in Ω\ωβ,βψ2=0,|ψ2|Ch12.\psi_{2}=\left\{\begin{array}[]{ll}1,&\text{ in }\stackrel_{\beta}\\ \mathcal{O}\left(h^{3}\right),&\text{ in }\Omega\backslash\omega_{\beta}\end{array},\quad\beta\cdot\nabla\psi_{2}=0,\quad\left|\nabla\psi_{2}\right|\leq Ch^{-\frac{1}{2}}.\right.

Such a function can be obtained by taking a positive constant λ\lambda that will be specified later and letting

ψ2(x,y):={exp((y`+y)/(λh12)),y>y`+1,(x,y)ω`βexp((yy`)/(λh12)),y<y`.\psi_{2}(x,y):=\begin{cases}\exp\left(\left(\grave{y}^{+}-y\right)/\left(\lambda h^{\frac{1}{2}}\right)\right),&y>\grave{y}^{+}\\ 1,&(x,y)\in\grave{\omega}_{\beta}\\ \exp\left(\left(y-\grave{y}^{-}\right)/\left(\lambda h^{\frac{1}{2}}\right)\right),&y<\grave{y}^{-}.\end{cases}

Note that ψ2\psi_{2} is only piecewise continuously differentiable. For ψ2\psi_{2} to decrease sufficiently rapidly to 𝒪(h3)\mathcal{O}\left(h^{3}\right) outside of ωβ\omega_{\beta}, we can take

dist(ωβ,Ω\ωβ)=min(y+y+,yy)3λh12ln(1/h),\operatorname{dist}\left(\stackrel_{\beta},\Omega\backslash\omega_{\beta}\right)=\min\left(y^{+}-\stackrel^{+},\stackrel^{-}-y^{-}\right)\geq 3\lambda h^{\frac{1}{2}}\ln(1/h),

which corresponds to cλ=3λc_{\lambda}=3\lambda in the definition of ω˙β\dot{\omega}_{\beta} given in (6). We thus have that

|ψ2|λ1h12,\left|\nabla\psi_{2}\right|\leq\lambda^{-1}h^{-\frac{1}{2}},

and in the subsequent proofs the constant λ\lambda will be taken large enough.
We now define the weight function φW1,(Ω)\varphi\in W^{1,\infty}(\Omega) that will be used in the weighted norms. For the downstream case we take in Section 4.3

φ:=ψ1ψ2(0,1), when β1>0,\varphi:=\psi_{1}\psi_{2}\in(0,1),\text{ when }\beta_{1}>0, (8)

and for the upstream case in Section 4.4,

φ:=ψ1ψ2(1,0), when β1<0.\varphi:=\psi_{1}\psi_{2}\in(-1,0),\text{ when }\beta_{1}<0. (9)

Using the product rule and the fact that βψ2=0\beta\cdot\nabla\psi_{2}=0, it follows that in both cases we have

βφ=|β||φ|,\beta\cdot\nabla\varphi=-|\beta||\varphi|, (10)

and

|φ|(1+λ1h12)|φ|.|\nabla\varphi|\leq\left(1+\lambda^{-1}h^{-\frac{1}{2}}\right)|\varphi|. (11)

We will denote the inflow and outflow boundaries by Ω\partial\Omega^{-}and Ω+\partial\Omega^{+}, i.e. βn<0\beta\cdot n<0 on Ω\partial\Omega^{-}and βn>0\beta\cdot n>0 on Ω+\partial\Omega^{+}. We will also assume that βn=0\beta\cdot n=0 can only hold on the boundary of Ω\ωβ\Omega\backslash\omega_{\beta}, and that μPelim 1|βn|h\mu\leq\mathrm{Pe}_{\text{lim }}^{-1}|\beta\cdot n|h when βn0\beta\cdot n\neq 0. This is straightforward to verify in the model case of the unit square that we are considering.

3. Preliminaries and the discrete commutator property

We first collect several inequalities that will be used repeatedly. We recall the standard discrete inverse inequality

vhL2(K)Ch1vhL2(K),vh1(K),\left\|\nabla v_{h}\right\|_{L^{2}(K)}\leq Ch^{-1}\left\|v_{h}\right\|_{L^{2}(K)},\quad\forall v_{h}\in\mathbb{P}_{1}(K), (12)

see e.g. [EG04, Lemma 1.138], the continuous trace inequality

vL2(K)C(h12vL2(K)+h12vL2(K)),vH1(K),\|v\|_{L^{2}(\partial K)}\leq C\left(h^{-\frac{1}{2}}\|v\|_{L^{2}(K)}+h^{\frac{1}{2}}\|\nabla v\|_{L^{2}(K)}\right),\quad\forall v\in H^{1}(K), (13)

see e.g. [MS99], and the discrete trace inequality

vhnL2(K)Ch12vhL2(K),vh1(K).\left\|\nabla v_{h}\cdot n\right\|_{L^{2}(\partial K)}\leq Ch^{-\frac{1}{2}}\left\|\nabla v_{h}\right\|_{L^{2}(K)},\quad\forall v_{h}\in\mathbb{P}_{1}(K). (14)

We will use standard estimates for the L2L^{2}-projection πh:L2(Ω)Vh\pi_{h}:L^{2}(\Omega)\mapsto V_{h}, namely

πhuHm(Ω)\displaystyle\left\|\pi_{h}u\right\|_{H^{m}(\Omega)} CuHm(Ω),uHm(Ω),m=0,1,\displaystyle\leq C\|u\|_{H^{m}(\Omega)},\quad u\in H^{m}(\Omega),m=0,1,
uπhuHm(Ω)\displaystyle\left\|u-\pi_{h}u\right\|_{H^{m}(\Omega)} ChkmuHk(Ω),uHk(Ω),k=1,2.\displaystyle\leq Ch^{k-m}\|u\|_{H^{k}(\Omega)},\quad u\in H^{k}(\Omega),k=1,2.

Scaling the result in [BHL18, Lemma 2] we recall the Poincaré-type inequality

(μ12h+|β|12h32)vhH1(Ω)Cγ12s(vh,vh)12,vhVh.\left\|\left(\mu^{\frac{1}{2}}h+|\beta|^{\frac{1}{2}}h^{\frac{3}{2}}\right)v_{h}\right\|_{H^{1}(\Omega)}\leq C\gamma^{-\frac{1}{2}}s\left(v_{h},v_{h}\right)^{\frac{1}{2}},\quad\forall v_{h}\in V_{h}. (15)

Using (13) and approximation estimates we also have the jump inequality

sΩ(πhu,πhu)12Cγ12(μ12h+|β|12h32)|u|H2(Ω),uH2(Ω).s_{\Omega}\left(\pi_{h}u,\pi_{h}u\right)^{\frac{1}{2}}\leq C\gamma^{\frac{1}{2}}\left(\mu^{\frac{1}{2}}h+|\beta|^{\frac{1}{2}}h^{\frac{3}{2}}\right)|u|_{H^{2}(\Omega)},\quad\forall u\in H^{2}(\Omega). (16)

We also recall that for a Lipschitz domain KK - and hence for any element K𝒯hK\in\mathcal{T}_{h} - a function φ\varphi is Lipschitz continuous if and only if φW1,(K)\varphi\in W^{1,\infty}(K). This follows from the proof in [Eva10, Theorem 4, p. 294] where the extension operator in the third step of the proof is replaced by the extension operator in [Ste70, Theorem 5, p. 181]. This equivalence holds for more general domains satisfying the minimal smoothness property in [Ste70, p. 189]. The proof in [Eva10, Theorem 4, p. 294] also shows that the mean value theorem holds and for any x,yKx,y\in K,

|φ(x)φ(y)|CexhK|φ|W1,(K)|\varphi(x)-\varphi(y)|\leq C_{ex}h_{K}|\varphi|_{W^{1,\infty}(K)} (17)

where |φ|W1,(K):=φ,K|\varphi|_{W^{1,\infty}(K)}:=\|\nabla\varphi\|_{\infty,K} and the constant Cex>0C_{ex}>0 is the norm of the extension operator used.

Lemma 1. For all vhVhv_{h}\in V_{h} and K𝒯hK\in\mathcal{T}_{h}, the following inequalities hold

φ,KvhKCvhφK,\displaystyle\|\varphi\|_{\infty,K}\left\|v_{h}\right\|_{K}\leq C\left\|v_{h}\varphi\right\|_{K}, (18)
vhφKCh12vhφK,\displaystyle\left\|v_{h}\varphi\right\|_{\partial K}\leq Ch^{-\frac{1}{2}}\left\|v_{h}\varphi\right\|_{K}, (19)

assuming that ( h+λ1h12h+\lambda^{-1}h^{\frac{1}{2}} ) is small enough.
Proof. Let xKx^{*}\in K be such that |φ(x)|=φ,K\left|\varphi\left(x^{*}\right)\right|=\|\varphi\|_{\infty,K}. Using the triangle inequality we may write

φ,KvhKφvhK+(φ(x)φ)vhK.\|\varphi\|_{\infty,K}\left\|v_{h}\right\|_{K}\leq\left\|\varphi v_{h}\right\|_{K}+\left\|\left(\varphi\left(x^{*}\right)-\varphi\right)v_{h}\right\|_{K}.

By the mean value theorem (17) we have that

|φ(x)φ|Cexh|φ|W1,(K)\left|\varphi\left(x^{*}\right)-\varphi\right|\leq C_{ex}h|\varphi|_{W^{1,\infty}(K)}

and by (11) together with the assumption that Cex(h+λ1h12)<12C_{ex}\left(h+\lambda^{-1}h^{\frac{1}{2}}\right)<\frac{1}{2} we get

Cexh|φ|W1,(K)Cex(h+λ1h12)φ,K12φ,KC_{ex}h|\varphi|_{W^{1,\infty}(K)}\leq C_{ex}\left(h+\lambda^{-1}h^{\frac{1}{2}}\right)\|\varphi\|_{\infty,K}\leq\frac{1}{2}\|\varphi\|_{\infty,K}

It follows that

φ,KvhKφvhK+12φ,KvhK,\|\varphi\|_{\infty,K}\left\|v_{h}\right\|_{K}\leq\left\|\varphi v_{h}\right\|_{K}+\frac{1}{2}\|\varphi\|_{\infty,K}\left\|v_{h}\right\|_{K},

from which the claim (18) is immediate. Considering now (19), using the standard element-wise trace inequality (13) we have

h12vhφKC(vhφK+h(vhφ)K).\left\|h^{\frac{1}{2}}v_{h}\varphi\right\|_{\partial K}\leq C\left(\left\|v_{h}\varphi\right\|_{K}+h\left\|\nabla\left(v_{h}\varphi\right)\right\|_{K}\right).

We bound the gradient term using (11) and the inverse inequality (12),

h(vhφ)K\displaystyle h\left\|\nabla\left(v_{h}\varphi\right)\right\|_{K} hvhφK+hφvhK\displaystyle\leq h\left\|v_{h}\nabla\varphi\right\|_{K}+h\left\|\varphi\nabla v_{h}\right\|_{K}
(h+λ1h12)φ,KvhK+Cφ,KvhK.\displaystyle\leq\left(h+\lambda^{-1}h^{\frac{1}{2}}\right)\|\varphi\|_{\infty,K}\left\|v_{h}\right\|_{K}+C\|\varphi\|_{\infty,K}\left\|v_{h}\right\|_{K}.

We conclude by collecting the terms and using (18).
3.1. Discrete commutator property. We denote by ihi_{h} the Lagrange nodal interpolant. We herein consider a Lipschitz weight function and prove the following super approximation result, also known as the discrete commutator property. This result will be essential to derive local weighted estimates and is similar to the one proven in [Ber99] for smooth compactly supported weight functions. For an introduction to interior estimates we refer the reader to [NS74].
Lemma 2. Let vhVhv_{h}\in V_{h} and K𝒯hK\in\mathcal{T}_{h}. Then for any weight function φW1,(K)\varphi\in W^{1,\infty}(K)

vhφih(vhφ)K+h(vhφih(vhφ))KCh|φ|W1,(K)vhK\left\|v_{h}\varphi-i_{h}\left(v_{h}\varphi\right)\right\|_{K}+h\left\|\nabla\left(v_{h}\varphi-i_{h}\left(v_{h}\varphi\right)\right)\right\|_{K}\leq Ch|\varphi|_{W^{1,\infty}(K)}\left\|v_{h}\right\|_{K}

Proof. We will first show the L2L^{2}-norm estimate

vhφih(vhφ)KCh|φ|W1,(K)vhK\left\|v_{h}\varphi-i_{h}\left(v_{h}\varphi\right)\right\|_{K}\leq Ch|\varphi|_{W^{1,\infty}(K)}\left\|v_{h}\right\|_{K}

Let xKx^{*}\in K be such that |φ(x)|=φ,K\left|\varphi\left(x^{*}\right)\right|=\|\varphi\|_{\infty,K} and let Rφ=φφ(x)R_{\varphi}=\varphi-\varphi\left(x^{*}\right). Note that

(1ih)(vhφ)K=(1ih)(vhRφ)K\left\|\left(1-i_{h}\right)\left(v_{h}\varphi\right)\right\|_{K}=\left\|\left(1-i_{h}\right)\left(v_{h}R_{\varphi}\right)\right\|_{K}

Observe that ih(vhφ)=ih(vhihφ)i_{h}\left(v_{h}\varphi\right)=i_{h}\left(v_{h}i_{h}\varphi\right) and therefore

(1ih)(vhRφ)K=vhRφih(vhihRφ)K\left\|\left(1-i_{h}\right)\left(v_{h}R_{\varphi}\right)\right\|_{K}=\left\|v_{h}R_{\varphi}-i_{h}\left(v_{h}i_{h}R_{\varphi}\right)\right\|_{K}

By the triangle inequality

ih(vhihRφ)vhRφKih(vhihRφ)vhihRφK+vh(ihRφRφ)K\left\|i_{h}\left(v_{h}i_{h}R_{\varphi}\right)-v_{h}R_{\varphi}\right\|_{K}\leq\left\|i_{h}\left(v_{h}i_{h}R_{\varphi}\right)-v_{h}i_{h}R_{\varphi}\right\|_{K}+\left\|v_{h}\left(i_{h}R_{\varphi}-R_{\varphi}\right)\right\|_{K}

For the first term, since vhihRφH1(K)v_{h}i_{h}R_{\varphi}\in H^{1}(K) we have by standard interpolation that

ih(vhihRφ)vhihRφKCh(vhihRφ)K\left\|i_{h}\left(v_{h}i_{h}R_{\varphi}\right)-v_{h}i_{h}R_{\varphi}\right\|_{K}\leq Ch\left\|\nabla\left(v_{h}i_{h}R_{\varphi}\right)\right\|_{K}

and then

(vhihRφ)K|ihRφ|W1,(K)vhK+ihRφ,KvhK\left\|\nabla\left(v_{h}i_{h}R_{\varphi}\right)\right\|_{K}\leq\left|i_{h}R_{\varphi}\right|_{W^{1,\infty}(K)}\left\|v_{h}\right\|_{K}+\left\|i_{h}R_{\varphi}\right\|_{\infty,K}\left\|\nabla v_{h}\right\|_{K}

By inserting φ\nabla\varphi and φ\varphi, respectively, and using interpolation estimates in W1,(K)W^{1,\infty}(K) [EG04, Theorem 1.103] and the mean value theorem (17) we have the following approximation

h|ihRφ|W1,(K)+ihRφ,KCh|φ|W1,(K)h\left|i_{h}R_{\varphi}\right|_{W^{1,\infty}(K)}+\left\|i_{h}R_{\varphi}\right\|_{\infty,K}\leq Ch|\varphi|_{W^{1,\infty}(K)}

Combined with the previous estimate and the inverse inequality (12) this gives that

(vhihRφ)KC|φ|W1,(K)vhK\left\|\nabla\left(v_{h}i_{h}R_{\varphi}\right)\right\|_{K}\leq C|\varphi|_{W^{1,\infty}(K)}\left\|v_{h}\right\|_{K} (20)

For the second term, using again interpolation [EG04, Theorem 1.103] we have

vh(ihRφRφ)KihRφRφ,KvhKCh|φ|W1,(K)vhK,\left\|v_{h}\left(i_{h}R_{\varphi}-R_{\varphi}\right)\right\|_{K}\leq\left\|i_{h}R_{\varphi}-R_{\varphi}\right\|_{\infty,K}\left\|v_{h}\right\|_{K}\leq Ch|\varphi|_{W^{1,\infty}(K)}\left\|v_{h}\right\|_{K},

and thus we have shown that

vhφih(vhφ)KCh|φ|W1,(K)vhK\left\|v_{h}\varphi-i_{h}\left(v_{h}\varphi\right)\right\|_{K}\leq Ch|\varphi|_{W^{1,\infty}(K)}\left\|v_{h}\right\|_{K}

The approximation estimate for the gradient follows by the same arguments. Indeed,

(1ih)(vhφ)K\displaystyle\left\|\nabla\left(1-i_{h}\right)\left(v_{h}\varphi\right)\right\|_{K} =(1ih)(vhRφ)K=(vhRφ)ih(vhihRφ)K\displaystyle=\left\|\nabla\left(1-i_{h}\right)\left(v_{h}R_{\varphi}\right)\right\|_{K}=\left\|\nabla\left(v_{h}R_{\varphi}\right)-\nabla i_{h}\left(v_{h}i_{h}R_{\varphi}\right)\right\|_{K}
(vhRφ)(vhihRφ)K+(vhihRφ)ih(vhihRφ)K.\displaystyle\leq\left\|\nabla\left(v_{h}R_{\varphi}\right)-\nabla\left(v_{h}i_{h}R_{\varphi}\right)\right\|_{K}+\left\|\nabla\left(v_{h}i_{h}R_{\varphi}\right)-\nabla i_{h}\left(v_{h}i_{h}R_{\varphi}\right)\right\|_{K}.

We first use interpolation and the inverse inequality (12) to get

(vh(RφihRφ))K\displaystyle\left\|\nabla\left(v_{h}\left(R_{\varphi}-i_{h}R_{\varphi}\right)\right)\right\|_{K} vh(RφihRφ)K+(RφihRφ)vhK\displaystyle\leq\left\|v_{h}\nabla\left(R_{\varphi}-i_{h}R_{\varphi}\right)\right\|_{K}+\left\|\left(R_{\varphi}-i_{h}R_{\varphi}\right)\nabla v_{h}\right\|_{K}
C|φ|W1,(K)vhK.\displaystyle\leq C|\varphi|_{W^{1,\infty}(K)}\left\|v_{h}\right\|_{K}.

Then we use an inverse inequality followed by interpolation and (20) to obtain

(vhihRφ)ih(vhihRφ)KC(vhihRφ)KC|φ|W1,(K)vhK\left\|\nabla\left(v_{h}i_{h}R_{\varphi}\right)-\nabla i_{h}\left(v_{h}i_{h}R_{\varphi}\right)\right\|_{K}\leq C\left\|\nabla\left(v_{h}i_{h}R_{\varphi}\right)\right\|_{K}\leq C|\varphi|_{W^{1,\infty}(K)}\left\|v_{h}\right\|_{K}

4. A priori local error estimates

4.1. Consistency and continuity. The following consistency result holds exactly as in the diffusion-dominated case, see [BNO20, Lemma 4]. We give the proof for the sake of completeness.
Lemma 3 (Consistency). Let uH2(Ω)u\in H^{2}(\Omega) be the solution to (1) and (uh,zh)[Vh]2\left(u_{h},z_{h}\right)\in\left[V_{h}\right]^{2} the solution to (5), then

ah(πhuuh,wh)+s(zh,wh)=ah(πhuu,wh)a_{h}\left(\pi_{h}u-u_{h},w_{h}\right)+s_{*}\left(z_{h},w_{h}\right)=a_{h}\left(\pi_{h}u-u,w_{h}\right)

and

ah(vh,zh)+s(πhuuh,vh)=sΩ(πhuu,vh)+sω(πhuU~ω,vh)-a_{h}\left(v_{h},z_{h}\right)+s\left(\pi_{h}u-u_{h},v_{h}\right)=s_{\Omega}\left(\pi_{h}u-u,v_{h}\right)+s_{\omega}\left(\pi_{h}u-\tilde{U}_{\omega},v_{h}\right)

for all (vh,wh)[Vh]2\left(v_{h},w_{h}\right)\in\left[V_{h}\right]^{2}.
Proof. The first claim follows from the definition of aha_{h}, since

ah(uh,wh)s(zh,wh)=(f,wh)Ω=(μΔu+βu,wh)Ω=ah(u,wh)a_{h}\left(u_{h},w_{h}\right)-s_{*}\left(z_{h},w_{h}\right)=\left(f,w_{h}\right)_{\Omega}=\left(-\mu\Delta u+\beta\cdot\nabla u,w_{h}\right)_{\Omega}=a_{h}\left(u,w_{h}\right)

where in the last equality we integrated by parts. The second claim follows similarly from

ah(vh,zh)+s(uh,vh)=sω(U~ω,vh)a_{h}\left(v_{h},z_{h}\right)+s\left(u_{h},v_{h}\right)=s_{\omega}\left(\tilde{U}_{\omega},v_{h}\right)

which combined with the fact that sΩ(u,vh)=0s_{\Omega}\left(u,v_{h}\right)=0 leads to

ah(vh,zh)+s(πhuuh,vh)\displaystyle-a_{h}\left(v_{h},z_{h}\right)+s\left(\pi_{h}u-u_{h},v_{h}\right) =s(πhu,vh)sω(U~ω,vh)\displaystyle=s\left(\pi_{h}u,v_{h}\right)-s_{\omega}\left(\tilde{U}_{\omega},v_{h}\right)
=sΩ(πhuu,vh)+sω(πhuU~ω,vh)\displaystyle=s_{\Omega}\left(\pi_{h}u-u,v_{h}\right)+s_{\omega}\left(\pi_{h}u-\tilde{U}_{\omega},v_{h}\right)

We now introduce the stabilization norm on [Vh]2\left[V_{h}\right]^{2} by combining the primal and dual stabilizers

(vh,wh)s2:=s(vh,vh)+s(wh,wh)\left\|\left(v_{h},w_{h}\right)\right\|_{s}^{2}:=s\left(v_{h},v_{h}\right)+s_{*}\left(w_{h},w_{h}\right)

and the "continuity norm" defined on H32+ε(Ω)H^{\frac{3}{2}+\varepsilon}(\Omega), for any ε>0\varepsilon>0,

v:=|β|12h12vΩ+|β|12h12vΩ+h12μ12vnΩ.\|v\|_{\sharp}:=\left\||\beta|^{\frac{1}{2}}h^{-\frac{1}{2}}v\right\|_{\Omega}+\left\||\beta|^{\frac{1}{2}}h^{\frac{1}{2}}\nabla v\right\|_{\Omega}+\left\|h^{\frac{1}{2}}\mu^{\frac{1}{2}}\nabla v\cdot n\right\|_{\partial\Omega}.

From the jump inequality (16), standard approximation bounds for πh\pi_{h} and the trace inequality (13), it follows that for uH2(Ω)u\in H^{2}(\Omega)

(uπhu,0)s+uπhuC(μ12h+|β|12h32)|u|H2(Ω).\left\|\left(u-\pi_{h}u,0\right)\right\|_{s}+\left\|u-\pi_{h}u\right\|_{\sharp}\leq C\left(\mu^{\frac{1}{2}}h+|\beta|^{\frac{1}{2}}h^{\frac{3}{2}}\right)|u|_{H^{2}(\Omega)}. (21)

We also define the orthogonal space

Vh:={vH2(Ω):(v,wh)Ω=0,whVh}V_{h}^{\perp}:=\left\{v\in H^{2}(\Omega):\left(v,w_{h}\right)_{\Omega}=0,\quad\forall w_{h}\in V_{h}\right\}

Lemma 4. (Continuity) Let vVhv\in V_{h}^{\perp} and whVhw_{h}\in V_{h}, then

ah(v,wh)Cv(0,wh)s.a_{h}\left(v,w_{h}\right)\leq C\|v\|_{\sharp}\left\|\left(0,w_{h}\right)\right\|_{s}.

Proof. Integrating by parts in the convective term of aha_{h} and using β=0\nabla\cdot\beta=0 leads to

ah(v,wh)=(v,βwh)Ω+vβn,whΩ+(μv,wh)Ωμvn,whΩa_{h}\left(v,w_{h}\right)=-\left(v,\beta\cdot\nabla w_{h}\right)_{\Omega}+\left\langle v\beta\cdot n,w_{h}\right\rangle_{\partial\Omega}+\left(\mu\nabla v,\nabla w_{h}\right)_{\Omega}-\left\langle\mu\nabla v\cdot n,w_{h}\right\rangle_{\partial\Omega}

For the first term we recall the discrete approximation estimate that holds for any piecewise linear β\beta, see e.g. [Bur05, Theorem 2.2],

infxhVhh12(βwhxh)Ω\displaystyle\inf_{x_{h}\in V_{h}}\left\|h^{\frac{1}{2}}\left(\beta\cdot\nabla w_{h}-x_{h}\right)\right\|_{\Omega} C(FihβwhF2)12\displaystyle\leq C\left(\sum_{F\in\mathcal{F}_{i}}\left\|h\llbracket\beta\cdot\nabla w_{h}\rrbracket\right\|_{F}^{2}\right)^{\frac{1}{2}} (22)
C|β|12γ12sΩ(wh,wh)12\displaystyle\leq C|\beta|^{\frac{1}{2}}\gamma^{-\frac{1}{2}}s_{\Omega}\left(w_{h},w_{h}\right)^{\frac{1}{2}}

and use orthogonality to obtain

(v,βwh)Ωh12vΩinfxhVhh12(βwhxh)ΩCv(0,wh)s.-\left(v,\beta\cdot\nabla w_{h}\right)_{\Omega}\leq\left\|h^{-\frac{1}{2}}v\right\|_{\Omega}\inf_{x_{h}\in V_{h}}\left\|h^{\frac{1}{2}}\left(\beta\cdot\nabla w_{h}-x_{h}\right)\right\|_{\Omega}\leq C\|v\|_{\sharp}\left\|\left(0,w_{h}\right)\right\|_{s}.

For the remaining terms, applying the Cauchy-Schwarz inequality we see that

vβn,whΩ+(μv,wh)Ωμvn,whΩCv(0,wh)s.\left\langle v\beta\cdot n,w_{h}\right\rangle_{\partial\Omega}+\left(\mu\nabla v,\nabla w_{h}\right)_{\Omega}-\left\langle\mu\nabla v\cdot n,w_{h}\right\rangle_{\partial\Omega}\leq C\|v\|_{\sharp}\left\|\left(0,w_{h}\right)\right\|_{s}.

Note that the proof of the above continuity estimate holds for any divergence-free piecewise linear velocity field β\beta. To address the case of a general velocity field βW1,(Ω)\beta\in W^{1,\infty}(\Omega) one can use a similar argument by considering its piecewise linear approximation as in [BNO20, Lemma 5]. Assuming that β\beta is divergence-free, the constant would be proportional to hPe(h)12|β|1,/βhPe(h)^{\frac{1}{2}}|\beta|_{1,\infty}/\|\beta\|_{\infty}, otherwise it would be proportional to Pe(h)12|β|1,/β\operatorname{Pe}(h)^{\frac{1}{2}}|\beta|_{1,\infty}/\|\beta\|_{\infty}.
4.2. Convergence of regularization. We now prove optimal convergence for the stabilizing and data assimilation terms.
Proposition 2. (Convergence of regularization). Let uH2(Ω)u\in H^{2}(\Omega) be the solution of (1) and (uh,zh)[Vh]2\left(u_{h},z_{h}\right)\in\left[V_{h}\right]^{2} the solution to (5), then there holds

(πhuuh,zh)sC(μ12h+|β|12h32)(|u|H2(Ω)+h2δω).\left\|\left(\pi_{h}u-u_{h},z_{h}\right)\right\|_{s}\leq C\left(\mu^{\frac{1}{2}}h+|\beta|^{\frac{1}{2}}h^{\frac{3}{2}}\right)\left(|u|_{H^{2}(\Omega)}+h^{-2}\|\delta\|_{\omega}\right).

Proof. Denoting eh=πhuuhe_{h}=\pi_{h}u-u_{h} we have that

(eh,zh)s2=ah(eh,zh)+s(zh,zh)ah(eh,zh)+s(eh,eh).\left\|\left(e_{h},z_{h}\right)\right\|_{s}^{2}=a_{h}\left(e_{h},z_{h}\right)+s_{*}\left(z_{h},z_{h}\right)-a_{h}\left(e_{h},z_{h}\right)+s\left(e_{h},e_{h}\right).

Using both claims in Lemma 3 we may write

(eh,zh)s2=ah(πhuu,zh)+sΩ(πhuu,eh)+sω(πhuU~ω,eh).\left\|\left(e_{h},z_{h}\right)\right\|_{s}^{2}=a_{h}\left(\pi_{h}u-u,z_{h}\right)+s_{\Omega}\left(\pi_{h}u-u,e_{h}\right)+s_{\omega}\left(\pi_{h}u-\tilde{U}_{\omega},e_{h}\right).

Since πhuuVh\pi_{h}u-u\in V_{h}^{\perp} we have by Lemma 4 that

ah(πhuu,zh)Cπhuu(0,zh)s.a_{h}\left(\pi_{h}u-u,z_{h}\right)\leq C\left\|\pi_{h}u-u\right\|_{\sharp}\left\|\left(0,z_{h}\right)\right\|_{s}.

We bound the other terms using the Cauchy-Schwarz inequality
sΩ(πhuu,eh)+sω(πhuU~ω,eh)((πhuu,0)s+(|β|h1+μhζ)12δω)(eh,0)ss_{\Omega}\left(\pi_{h}u-u,e_{h}\right)+s_{\omega}\left(\pi_{h}u-\tilde{U}_{\omega},e_{h}\right)\leq\left(\left\|\left(\pi_{h}u-u,0\right)\right\|_{s}+\left(|\beta|h^{-1}+\mu h^{-\zeta}\right)^{\frac{1}{2}}\|\delta\|_{\omega}\right)\left\|\left(e_{h},0\right)\right\|_{s}.
Collecting the above bounds we have

(eh,zh)s2C(πhuu+(πhuu,0)s+(|β|h1+μhζ)12δω)(eh,zh)s\left\|\left(e_{h},z_{h}\right)\right\|_{s}^{2}\leq C\left(\left\|\pi_{h}u-u\right\|_{\sharp}+\left\|\left(\pi_{h}u-u,0\right)\right\|_{s}+\left(|\beta|h^{-1}+\mu h^{-\zeta}\right)^{\frac{1}{2}}\|\delta\|_{\omega}\right)\left\|\left(e_{h},z_{h}\right)\right\|_{s}

and the claim follows by applying the approximation inequality (21).
Remark 2. Compared to the result in the diffusion-dominated case [BNO20, Proposition 3], the sensitivity to data perturbations has increased by a factor of h1h^{-1}. This is due to the stronger penalty in the data term sωs_{\omega} (c.f. Remark 1).
4.3. Downstream estimates. In this case we consider β=(β1,0)\beta=\left(\beta_{1},0\right) with β1>0\beta_{1}>0 and the data set

ω=(0,x)×(y,y+)\omega=(0,x)\times\left(y^{-},y^{+}\right)

touching part of the inflow boundary Ω\partial\Omega^{-}. We aim to obtain control of the following weighted triple norm defined on VhV_{h},

vhφ2:=|β|12vhφ12Ω2+μ12vhφ12Ω2+|βn|12vhφ12Ω+2\left\|v_{h}\right\|_{\varphi}^{2}:=\left\||\beta|^{\frac{1}{2}}v_{h}\varphi^{\frac{1}{2}}\right\|_{\Omega}^{2}+\left\|\mu^{\frac{1}{2}}\nabla v_{h}\varphi^{\frac{1}{2}}\right\|_{\Omega}^{2}+\left\||\beta\cdot n|^{\frac{1}{2}}v_{h}\varphi^{\frac{1}{2}}\right\|_{\partial\Omega^{+}}^{2} (23)

where φ\varphi is given by (8). Since φ(0,1)\varphi\in(0,1), we will often use that φΩφ12Ω\|\cdot\varphi\|_{\Omega}\leq\left\|\cdot\varphi^{\frac{1}{2}}\right\|_{\Omega}. We first consider vhφv_{h}\varphi as a test function in the weak bilinear form aha_{h} and obtain the following bound.

Lemma 5. There exist α>0\alpha>0 and h0>0h_{0}>0 such that for all h<h0h<h_{0} and vhVhv_{h}\in V_{h} we have

αvhφ2ah(vh,vhφ)+C(vh,0)s2.\alpha\left\|v_{h}\right\|_{\varphi}^{2}\leq a_{h}\left(v_{h},v_{h}\varphi\right)+C\left\|\left(v_{h},0\right)\right\|_{s}^{2}.

Proof. We start with the convective term. Since β=0\nabla\cdot\beta=0, the divergence theorem leads to

2(βvh,vhφ)Ω=vhβn,vhφΩ(vh,vhβφ)Ω2\left(\beta\cdot\nabla v_{h},v_{h}\varphi\right)_{\Omega}=\left\langle v_{h}\beta\cdot n,v_{h}\varphi\right\rangle_{\partial\Omega}-\left(v_{h},v_{h}\beta\cdot\nabla\varphi\right)_{\Omega}

Then combining with (10) we have

(βvh,vhφ)Ω=12(vhβn,vhφΩ+|β|vhφ12Ω2).\left(\beta\cdot\nabla v_{h},v_{h}\varphi\right)_{\Omega}=\frac{1}{2}\left(\left\langle v_{h}\beta\cdot n,v_{h}\varphi\right\rangle_{\partial\Omega}+|\beta|\left\|v_{h}\varphi^{\frac{1}{2}}\right\|_{\Omega}^{2}\right).

We split the boundary term into inflow and outflow

vhβn,vhφΩ=|βn|12vhφ12Ω2+|βn|12vhφ12Ω+2,\left\langle v_{h}\beta\cdot n,v_{h}\varphi\right\rangle_{\partial\Omega}=-\left\||\beta\cdot n|^{\frac{1}{2}}v_{h}\varphi^{\frac{1}{2}}\right\|_{\partial\Omega^{-}}^{2}+\left\||\beta\cdot n|^{\frac{1}{2}}v_{h}\varphi^{\frac{1}{2}}\right\|_{\partial\Omega^{+}}^{2},

and write

12(|βn|12vhφ12Ω+2+|β|vhφ12Ω2)=(βvh,vhφ)Ω+12|βn|12vhφ12Ω2\frac{1}{2}\left(\left\||\beta\cdot n|^{\frac{1}{2}}v_{h}\varphi^{\frac{1}{2}}\right\|_{\partial\Omega^{+}}^{2}+|\beta|\left\|v_{h}\varphi^{\frac{1}{2}}\right\|_{\Omega}^{2}\right)=\left(\beta\cdot\nabla v_{h},v_{h}\varphi\right)_{\Omega}+\frac{1}{2}\left\||\beta\cdot n|^{\frac{1}{2}}v_{h}\varphi^{\frac{1}{2}}\right\|_{\partial\Omega^{-}}^{2}

Splitting now the inflow boundary with respect to the closed set ωβ\omega_{\beta} and using the discrete trace inequality (13) in ω\omega, and the weight decay (7) together with a standard global trace inequality for H1H^{1} functions outside, we have that

|βn|12vhφ12Ω\displaystyle\left\||\beta\cdot n|^{\frac{1}{2}}v_{h}\varphi^{\frac{1}{2}}\right\|_{\partial\Omega^{-}} C|β|12(vhφ12Ωωβ+vhφ12Ω\ωβ)\displaystyle\leq C|\beta|^{\frac{1}{2}}\left(\left\|v_{h}\varphi^{\frac{1}{2}}\right\|_{\partial\Omega^{-}\cap\omega_{\beta}}+\left\|v_{h}\varphi^{\frac{1}{2}}\right\|_{\partial\Omega^{-}\backslash\omega_{\beta}}\right)
C|β|12h12vhω+C|β|12h32vhH1(Ω)\displaystyle\leq C|\beta|^{\frac{1}{2}}h^{-\frac{1}{2}}\left\|v_{h}\right\|_{\omega}+C|\beta|^{\frac{1}{2}}h^{\frac{3}{2}}\left\|v_{h}\right\|_{H^{1}(\Omega)} (24)
Cγ12(vh,0)s\displaystyle\leq C\gamma^{-\frac{1}{2}}\left\|\left(v_{h},0\right)\right\|_{s}

where in the last step we used the Poincaré-type inequality (15). Hence we obtain control over the convective terms in the triple weighted norm

12(|βn|12vhφ12Ω+2+|β|vhφ12Ω2)(βvh,vhφ)Ω+Cγ1(vh,0)s2\frac{1}{2}\left(\left\||\beta\cdot n|^{\frac{1}{2}}v_{h}\varphi^{\frac{1}{2}}\right\|_{\partial\Omega^{+}}^{2}+|\beta|\left\|v_{h}\varphi^{\frac{1}{2}}\right\|_{\Omega}^{2}\right)\leq\left(\beta\cdot\nabla v_{h},v_{h}\varphi\right)_{\Omega}+C\gamma^{-1}\left\|\left(v_{h},0\right)\right\|_{s}^{2} (25)

Let us consider the terms in ah(vh,vhφ)a_{h}\left(v_{h},v_{h}\varphi\right) corresponding to the diffusion operator, starting with

(μvh,(vhφ))Ω=μ12vhφ12Ω2+(μvh,vhφ)Ω\left(\mu\nabla v_{h},\nabla\left(v_{h}\varphi\right)\right)_{\Omega}=\left\|\mu^{\frac{1}{2}}\nabla v_{h}\varphi^{\frac{1}{2}}\right\|_{\Omega}^{2}+\left(\mu\nabla v_{h},v_{h}\nabla\varphi\right)_{\Omega}

which we rearrange as

μ12vhφ12Ω2=(μvh,(vhφ))Ω(μvh,vhφ)Ω\left\|\mu^{\frac{1}{2}}\nabla v_{h}\varphi^{\frac{1}{2}}\right\|_{\Omega}^{2}=\left(\mu\nabla v_{h},\nabla\left(v_{h}\varphi\right)\right)_{\Omega}-\left(\mu\nabla v_{h},v_{h}\nabla\varphi\right)_{\Omega}

Bounding φ\nabla\varphi by (11) and using Cauchy-Schwarz together with μ|β|h\mu\leq|\beta|h we have that

|(μvh,vhφ)Ω|\displaystyle\left|\left(\mu\nabla v_{h},v_{h}\nabla\varphi\right)_{\Omega}\right| μ(|vhφ|,vh)Ω\displaystyle\leq\mu\left(\left|\nabla v_{h}\cdot\nabla\varphi\right|,v_{h}\right)_{\Omega}
μ(1+λ1h12)(|vh|φ12,vhφ12)Ω\displaystyle\leq\mu\left(1+\lambda^{-1}h^{-\frac{1}{2}}\right)\left(\left|\nabla v_{h}\right|\varphi^{\frac{1}{2}},v_{h}\varphi^{\frac{1}{2}}\right)_{\Omega}
13μ12vhφ12Ω2+C(h+λ2)|β|vhφ12Ω2\displaystyle\leq\frac{1}{3}\left\|\mu^{\frac{1}{2}}\nabla v_{h}\varphi^{\frac{1}{2}}\right\|_{\Omega}^{2}+C\left(h+\lambda^{-2}\right)|\beta|\left\|v_{h}\varphi^{\frac{1}{2}}\right\|_{\Omega}^{2}

We split the boundary term μvhn,vhφΩ\left\langle\mu\nabla v_{h}\cdot n,v_{h}\varphi\right\rangle_{\partial\Omega} into inflow and outflow again. Similarly to (24) we have that

μvhn,vhφΩChγ1(vh,0)s2\left\langle\mu\nabla v_{h}\cdot n,v_{h}\varphi\right\rangle_{\partial\Omega^{-}}\leq Ch\gamma^{-1}\left\|\left(v_{h},0\right)\right\|_{s}^{2}

For the outflow boundary term we use Cauchy-Schwarz and a trace inequality to obtain

μvhn,vhφΩ+\displaystyle\left\langle\mu\nabla v_{h}\cdot n,v_{h}\varphi\right\rangle_{\partial\Omega^{+}} μ12vhnφ12Ω+μ12vhφ12Ω+\displaystyle\leq\left\|\mu^{\frac{1}{2}}\nabla v_{h}\cdot n\varphi^{\frac{1}{2}}\right\|_{\partial\Omega^{+}}\left\|\mu^{\frac{1}{2}}v_{h}\varphi^{\frac{1}{2}}\right\|_{\partial\Omega^{+}}
Ch12μ12vhφ12ΩPelim12h12|βn|12vhφ12Ω+\displaystyle\leq Ch^{-\frac{1}{2}}\left\|\mu^{\frac{1}{2}}\nabla v_{h}\varphi^{\frac{1}{2}}\right\|_{\Omega}Pe_{\lim}^{-\frac{1}{2}}h^{\frac{1}{2}}\left\||\beta\cdot n|^{\frac{1}{2}}v_{h}\varphi^{\frac{1}{2}}\right\|_{\partial\Omega^{+}}
13μ12vhφ12Ω+2+Pelim1|βn|12vhφ12Ω+2\displaystyle\leq\frac{1}{3}\left\|\mu^{\frac{1}{2}}\nabla v_{h}\varphi^{\frac{1}{2}}\right\|_{\Omega^{+}}^{2}+Pe_{\lim}^{-1}\left\||\beta\cdot n|^{\frac{1}{2}}v_{h}\varphi^{\frac{1}{2}}\right\|_{\partial\Omega^{+}}^{2}

We denote the part of the boundary where βn=0\beta\cdot n=0 by Ω0\partial\Omega^{0} and use the assumption that Ω0\partial\Omega^{0} is away from ωβ\omega_{\beta}, meaning that the weight function φ\varphi is 𝒪(h3)\mathcal{O}\left(h^{3}\right) there. We use trace inequalities and (15) to bound

μvhn,vhφΩ0Cγ1(vh,0)s2.\left\langle\mu\nabla v_{h}\cdot n,v_{h}\varphi\right\rangle_{\partial\Omega^{0}}\leq C\gamma^{-1}\left\|\left(v_{h},0\right)\right\|_{s}^{2}.

Collecting the above bounds we obtain that

13μ12vhφ12Ω2\displaystyle\frac{1}{3}\left\|\mu^{\frac{1}{2}}\nabla v_{h}\varphi^{\frac{1}{2}}\right\|_{\Omega}^{2}\leq (μvh,(vhφ))Ωμvhn,vhφΩ\displaystyle\left(\mu\nabla v_{h},\nabla\left(v_{h}\varphi\right)\right)_{\Omega}-\left\langle\mu\nabla v_{h}\cdot n,v_{h}\varphi\right\rangle_{\partial\Omega}
+C(h+λ2+Pelim1)(|β|12vhφ12Ω2+|βn|12vhφ12Ω+2)\displaystyle+C\left(h+\lambda^{-2}+Pe_{\lim}^{-1}\right)\left(\left\||\beta|^{\frac{1}{2}}v_{h}\varphi^{\frac{1}{2}}\right\|_{\Omega}^{2}+\left\||\beta\cdot n|^{\frac{1}{2}}v_{h}\varphi^{\frac{1}{2}}\right\|_{\partial\Omega^{+}}^{2}\right)
+Cγ1(vh,0)s2.\displaystyle+C\gamma^{-1}\left\|\left(v_{h},0\right)\right\|_{s}^{2}.

We conclude by combining this with (25) and assuming that hh is small enough and Pelim Pe_{\text{lim }} are large enough (thus absorbing the convective terms from the rhs into the lhs). \square

Now we refine the control over the triple norm vhφ\left\|v_{h}\right\|_{\varphi} by taking the projection πh(vhφ)Vh\pi_{h}\left(v_{h}\varphi\right)\in V_{h} as a test function.
Corollary 1. There exist α>0\alpha>0 and h0>0h_{0}>0 such that for all h<h0h<h_{0} and vhVhv_{h}\in V_{h} we have

αvhφ2ah(vh,πh(vhφ))+C(vh,0)s2.\alpha\left\|v_{h}\right\|_{\varphi}^{2}\leq a_{h}\left(v_{h},\pi_{h}\left(v_{h}\varphi\right)\right)+C\left\|\left(v_{h},0\right)\right\|_{s}^{2}.

Proof. Since

ah(vh,πh(vhφ))=ah(vh,(πh1)(vhφ))+ah(vh,vhφ),a_{h}\left(v_{h},\pi_{h}\left(v_{h}\varphi\right)\right)=a_{h}\left(v_{h},\left(\pi_{h}-1\right)\left(v_{h}\varphi\right)\right)+a_{h}\left(v_{h},v_{h}\varphi\right),

we must bound ah(vh,(πh1)(vhφ))a_{h}\left(v_{h},\left(\pi_{h}-1\right)\left(v_{h}\varphi\right)\right) in a suitable way. Writing out the terms we have

ah(vh,(πh1)(vhφ))=\displaystyle a_{h}\left(v_{h},\left(\pi_{h}-1\right)\left(v_{h}\varphi\right)\right)= (βvh,(πh1)(vhφ))Ω+(μvh,(πh1)(vhφ))Ω\displaystyle\left(\beta\cdot\nabla v_{h},\left(\pi_{h}-1\right)\left(v_{h}\varphi\right)\right)_{\Omega}+\left(\mu\nabla v_{h},\nabla\left(\pi_{h}-1\right)\left(v_{h}\varphi\right)\right)_{\Omega}
μvhn,(πh1)(vhφ)Ω=I+II+III.\displaystyle-\left\langle\mu\nabla v_{h}\cdot n,\left(\pi_{h}-1\right)\left(v_{h}\varphi\right)\right\rangle_{\partial\Omega}=I+II+III.

Let us consider the convection term first, and use orthogonality combined with (22)

I=(βvh,(πh1)(vhφ))Ω\displaystyle I=\left(\beta\cdot\nabla v_{h},\left(\pi_{h}-1\right)\left(v_{h}\varphi\right)\right)_{\Omega} C|β|12γ12(vh,0)sh12(πh1)(vhφ)Ω\displaystyle\leq C|\beta|^{\frac{1}{2}}\gamma^{-\frac{1}{2}}\left\|\left(v_{h},0\right)\right\|_{s}h^{-\frac{1}{2}}\left\|\left(\pi_{h}-1\right)\left(v_{h}\varphi\right)\right\|_{\Omega}
C|β|12γ12(vh,0)sh12(ih1)(vhφ)Ω\displaystyle\leq C|\beta|^{\frac{1}{2}}\gamma^{-\frac{1}{2}}\left\|\left(v_{h},0\right)\right\|_{s}h^{-\frac{1}{2}}\left\|\left(i_{h}-1\right)\left(v_{h}\varphi\right)\right\|_{\Omega}

Integrating by parts and using that Δvh=0\Delta v_{h}=0 on every element KK we obtain by the trace inequality (13) and the assumption Pe(h)>1Pe(h)>1 that

II+III\displaystyle II+III =FiFμvhn(πh1)(vhφ)ds\displaystyle=\sum_{F\in\mathcal{F}_{i}}\int_{F}\mu\llbracket\nabla v_{h}\cdot n\rrbracket\left(\pi_{h}-1\right)\left(v_{h}\varphi\right)\mathrm{d}s
C|β|12γ12sΩ(vh,vh)12(h12(πh1)(vhφ)Ω+h12(πh1)(vhφ)Ω).\displaystyle\leq C|\beta|^{\frac{1}{2}}\gamma^{-\frac{1}{2}}s_{\Omega}\left(v_{h},v_{h}\right)^{\frac{1}{2}}\left(h^{-\frac{1}{2}}\left\|\left(\pi_{h}-1\right)\left(v_{h}\varphi\right)\right\|_{\Omega}+h^{\frac{1}{2}}\left\|\nabla\left(\pi_{h}-1\right)\left(v_{h}\varphi\right)\right\|_{\Omega}\right).

Notice that ih(vhφ)=πh(ih(vhφ))i_{h}\left(v_{h}\varphi\right)=\pi_{h}\left(i_{h}\left(v_{h}\varphi\right)\right) and the stability of the projection gives

(πhih)(vhφ)Ω=πh(1ih)(vhφ)ΩC(1ih)(vhφ)Ω,\left\|\nabla\left(\pi_{h}-i_{h}\right)\left(v_{h}\varphi\right)\right\|_{\Omega}=\left\|\nabla\pi_{h}\left(1-i_{h}\right)\left(v_{h}\varphi\right)\right\|_{\Omega}\leq C\left\|\nabla\left(1-i_{h}\right)\left(v_{h}\varphi\right)\right\|_{\Omega}, (26)

and hence

h12(πh1)(vhφ)Ω\displaystyle h^{\frac{1}{2}}\left\|\nabla\left(\pi_{h}-1\right)\left(v_{h}\varphi\right)\right\|_{\Omega} h12((πhih)(vhφ)Ω+(ih1)(vhφ)Ω)\displaystyle\leq h^{\frac{1}{2}}\left(\left\|\nabla\left(\pi_{h}-i_{h}\right)\left(v_{h}\varphi\right)\right\|_{\Omega}+\left\|\nabla\left(i_{h}-1\right)\left(v_{h}\varphi\right)\right\|_{\Omega}\right)
Ch12(ih1)(vhφ)Ω.\displaystyle\leq Ch^{\frac{1}{2}}\left\|\nabla\left(i_{h}-1\right)\left(v_{h}\varphi\right)\right\|_{\Omega}. (27)

Since

h12(πh1)(vhφ)ΩCh12(ih1)(vhφ)Ω,h^{-\frac{1}{2}}\left\|\left(\pi_{h}-1\right)\left(v_{h}\varphi\right)\right\|_{\Omega}\leq Ch^{-\frac{1}{2}}\left\|\left(i_{h}-1\right)\left(v_{h}\varphi\right)\right\|_{\Omega},

collecting the contributions above we see that

I+II+IIIC|β|12γ12(vh,0)s(h12(ih1)(vhφ)Ω+h12(ih1)(vhφ)ΩIV),I+II+III\leq C|\beta|^{\frac{1}{2}}\gamma^{-\frac{1}{2}}\left\|\left(v_{h},0\right)\right\|_{s}\underbrace{\left(h^{-\frac{1}{2}}\left\|\left(i_{h}-1\right)\left(v_{h}\varphi\right)\right\|_{\Omega}+h^{\frac{1}{2}}\left\|\nabla\left(i_{h}-1\right)\left(v_{h}\varphi\right)\right\|_{\Omega}\right.}_{IV}),

and hence

ah(vh,(πh1)(vhφ))=I+II+IIICγ1(vh,0)s2+|β|IV2.a_{h}\left(v_{h},\left(\pi_{h}-1\right)\left(v_{h}\varphi\right)\right)=I+II+III\leq C\gamma^{-1}\left\|\left(v_{h},0\right)\right\|_{s}^{2}+|\beta|IV^{2}.

The discrete commutator property Lemma 2 together with the φ\varphi-bounds (11) and (18) give that

IVCh12φ,ΩvhΩC(h12+λ1)vhφΩ.IV\leq Ch^{\frac{1}{2}}\|\nabla\varphi\|_{\infty,\Omega}\left\|v_{h}\right\|_{\Omega}\leq C\left(h^{\frac{1}{2}}+\lambda^{-1}\right)\left\|v_{h}\varphi\right\|_{\Omega}. (28)

Since φ(0,1)\varphi\in(0,1) and φ<φ12\varphi<\varphi^{\frac{1}{2}}, it follows that for hh small enough and λ\lambda large enough, given some α>0\alpha>0 we have

|β|IV2α2vhφ2|\beta|IV^{2}\leq\frac{\alpha}{2}\left\|v_{h}\right\|_{\varphi}^{2} (29)

Collecting the estimates for ah(vh,(πh1)(vhφ))a_{h}\left(v_{h},\left(\pi_{h}-1\right)\left(v_{h}\varphi\right)\right) and using Lemma 5 we see that

ah(vh,πh(vhφ))=ah(vh,(πh1)(vhφ))+ah(vh,vhφ)α2vhφ2Cγ1(vh,0)s2a_{h}\left(v_{h},\pi_{h}\left(v_{h}\varphi\right)\right)=a_{h}\left(v_{h},\left(\pi_{h}-1\right)\left(v_{h}\varphi\right)\right)+a_{h}\left(v_{h},v_{h}\varphi\right)\geq\frac{\alpha}{2}\left\|v_{h}\right\|_{\varphi}^{2}-C\gamma^{-1}\left\|\left(v_{h},0\right)\right\|_{s}^{2}

from which we conclude by renaming α/2\alpha/2 as α\alpha.
Lemma 6. For all vhVhv_{h}\in V_{h} there holds

(0,πh(vhφ))s2C(vhφ2+(vh,0)s2)\left\|\left(0,\pi_{h}\left(v_{h}\varphi\right)\right)\right\|_{s}^{2}\leq C\left(\left\|v_{h}\right\|_{\varphi}^{2}+\left\|\left(v_{h},0\right)\right\|_{s}^{2}\right)

Proof. First note that by triangle inequalities we have that up to a constant

(0,πh(vhφ))s\displaystyle\left\|\left(0,\pi_{h}\left(v_{h}\varphi\right)\right)\right\|_{s}\leq μ12(πh1)(vhφ))+Ωμ12(vhφ))Ω\displaystyle\left.\left.\|\mu^{\frac{1}{2}}\nabla\left(\pi_{h}-1\right)\left(v_{h}\varphi\right)\right)\left\|{}_{\Omega}+\right\|\mu^{\frac{1}{2}}\nabla\left(v_{h}\varphi\right)\right)\|_{\Omega}
+(|β|+μh1)12((πh1)(vhφ)Ω+vhφΩ)\displaystyle+\left(|\beta|+\mu h^{-1}\right)^{\frac{1}{2}}\left(\left\|\left(\pi_{h}-1\right)\left(v_{h}\varphi\right)\right\|_{\partial\Omega}+\left\|v_{h}\varphi\right\|_{\partial\Omega}\right)
+sΩ(πh(vhφ),πh(vhφ))12\displaystyle+s_{\Omega}\left(\pi_{h}\left(v_{h}\varphi\right),\pi_{h}\left(v_{h}\varphi\right)\right)^{\frac{1}{2}}

We bound these terms line by line. Using (27), (28), (11) and (18) we bound the first line by

|β|12h12(ih1)(vhφ)Ω+μ12vhφΩ+2|β|12(h12+λ1)vhφΩCvhφ.|\beta|^{\frac{1}{2}}h^{\frac{1}{2}}\left\|\nabla\left(i_{h}-1\right)\left(v_{h}\varphi\right)\right\|_{\Omega}+\left\|\mu^{\frac{1}{2}}\nabla v_{h}\varphi\right\|_{\Omega}+2|\beta|^{\frac{1}{2}}\left(h^{\frac{1}{2}}+\lambda^{-1}\right)\left\|v_{h}\varphi\right\|_{\Omega}\leq C\left\|v_{h}\right\|_{\varphi}.

For the second line, using a global trace inequality and the stability of the projection we have

(|β|+μh1)12(πh1)(vhφ)ΩC(|β|+μh1)12vhφΩC|β|12vhφ12Ω\left(|\beta|+\mu h^{-1}\right)^{\frac{1}{2}}\left\|\left(\pi_{h}-1\right)\left(v_{h}\varphi\right)\right\|_{\partial\Omega}\leq C\left(|\beta|+\mu h^{-1}\right)^{\frac{1}{2}}\left\|v_{h}\varphi\right\|_{\Omega}\leq C|\beta|^{\frac{1}{2}}\left\|v_{h}\varphi^{\frac{1}{2}}\right\|_{\Omega}

Splitting the boundary into inflow and outflow and using (24), we have

(|β|+μh1)12vhφΩC|β|12vhφΩC(vh,0)s+Cvhφ.\left(|\beta|+\mu h^{-1}\right)^{\frac{1}{2}}\left\|v_{h}\varphi\right\|_{\partial\Omega}\leq C|\beta|^{\frac{1}{2}}\left\|v_{h}\varphi\right\|_{\partial\Omega}\leq C\left\|\left(v_{h},0\right)\right\|_{s}+C\left\|v_{h}\right\|_{\varphi}.

For the contribution of the jump term, we insert ihi_{h} and bound

sΩ(πh(vhφ),πh(vhφ))12\displaystyle s_{\Omega}\left(\pi_{h}\left(v_{h}\varphi\right),\pi_{h}\left(v_{h}\varphi\right)\right)^{\frac{1}{2}}\leq sΩ((πhih)(vhφ),(πhih)(vhφ))12\displaystyle s_{\Omega}\left(\left(\pi_{h}-i_{h}\right)\left(v_{h}\varphi\right),\left(\pi_{h}-i_{h}\right)\left(v_{h}\varphi\right)\right)^{\frac{1}{2}}
+sΩ((ih1)(vhφ),(ih1)(vhφ))12\displaystyle+s_{\Omega}\left(\left(i_{h}-1\right)\left(v_{h}\varphi\right),\left(i_{h}-1\right)\left(v_{h}\varphi\right)\right)^{\frac{1}{2}} (30)
+sΩ(vhφ,vhφ)12\displaystyle+s_{\Omega}\left(v_{h}\varphi,v_{h}\varphi\right)^{\frac{1}{2}}

We first observe that using (14) and (26), we can bound the first term by

sΩ((πhih)(vhφ),(πhih)(vhφ))12\displaystyle s_{\Omega}\left(\left(\pi_{h}-i_{h}\right)\left(v_{h}\varphi\right),\left(\pi_{h}-i_{h}\right)\left(v_{h}\varphi\right)\right)^{\frac{1}{2}} |β|12h12(πhih)(vhφ)Ω\displaystyle\leq|\beta|^{\frac{1}{2}}h^{\frac{1}{2}}\left\|\nabla\left(\pi_{h}-i_{h}\right)\left(v_{h}\varphi\right)\right\|_{\Omega}
|β|12h12(ih1)(vhφ)Ω\displaystyle\leq|\beta|^{\frac{1}{2}}h^{\frac{1}{2}}\left\|\nabla\left(i_{h}-1\right)\left(v_{h}\varphi\right)\right\|_{\Omega}
C|β|12h12φ,ΩvhΩ\displaystyle\leq C|\beta|^{\frac{1}{2}}h^{\frac{1}{2}}\|\nabla\varphi\|_{\infty,\Omega}\left\|v_{h}\right\|_{\Omega}
C|β|12(h12+λ1)vhφΩ,\displaystyle\leq C|\beta|^{\frac{1}{2}}\left(h^{\frac{1}{2}}+\lambda^{-1}\right)\left\|v_{h}\varphi\right\|_{\Omega},

where for the last two inequalities we used the discrete commutator property Lemma 2 together with the φ\varphi-bounds (11) and (18). Since φ\varphi is Lipschitz continuous on K,φ|FK,\left.\varphi\right|_{F}
is also Lipschitz continuous, and so φ|FW1,(F)\left.\varphi\right|_{F}\in W^{1,\infty}(F). The restriction of the nodal interpolant on KK onto FF gives the nodal interpolant on FF, hence applying Lemma 2 to FF instead of KK we have the discrete commutator estimate

hn(ih1)(vhφ)FCh|φ|W1,(K)vhFC(h12+λ1)vhφK,h\left\|n\cdot\nabla\left(i_{h}-1\right)\left(v_{h}\varphi\right)\right\|_{F}\leq Ch|\varphi|_{W^{1,\infty}(K)}\left\|v_{h}\right\|_{F}\leq C\left(h^{\frac{1}{2}}+\lambda^{-1}\right)\left\|v_{h}\varphi\right\|_{K},

where in the last step we used (11) and (18) together with a discrete trace inequality. After summation we have that

sΩ((ih1)(vhφ),(ih1)(vhφ))12Cvhφ.s_{\Omega}\left(\left(i_{h}-1\right)\left(v_{h}\varphi\right),\left(i_{h}-1\right)\left(v_{h}\varphi\right)\right)^{\frac{1}{2}}\leq C\left\|v_{h}\right\|_{\varphi}.

Finally we use the trivial bound (since |φ|<1|\varphi|<1 )

sΩ(vhφ,vhφ)12sΩ(vh,vh)12.s_{\Omega}\left(v_{h}\varphi,v_{h}\varphi\right)^{\frac{1}{2}}\leq s_{\Omega}\left(v_{h},v_{h}\right)^{\frac{1}{2}}.

We conclude the proof by summing up the above contributions. \square

We can now prove the following error estimate showing that, in the zone ω˙β\dot{\omega}_{\beta} where we have stability, the convergence in the L2L^{2}-norm is of order 𝒪(h32)\mathcal{O}\left(h^{\frac{3}{2}}\right) on unstructured meshes, which is known to be optimal.

Theorem 1. Let uH2(Ω)u\in H^{2}(\Omega) be the solution of (1) and (uh,zh)[Vh]2\left(u_{h},z_{h}\right)\in\left[V_{h}\right]^{2} the solution to (5). Then there exists h0>0h_{0}>0 such that for all h<h0h<h_{0} with Pe(h)1\operatorname{Pe}(h)\gtrsim 1 there holds

uuhφC(|β|12h32|u|H2(Ω)+|β|12h12δω).\left\|u-u_{h}\right\|_{\varphi}\leq C\left(|\beta|^{\frac{1}{2}}h^{\frac{3}{2}}|u|_{H^{2}(\Omega)}+|\beta|^{\frac{1}{2}}h^{-\frac{1}{2}}\|\delta\|_{\omega}\right).

Proof. Let eh=πhuuhVhe_{h}=\pi_{h}u-u_{h}\in V_{h}, then uuh=uπhu+ehu-u_{h}=u-\pi_{h}u+e_{h}. By Corollary 1 there exists α>0\alpha>0 such that

αehφ2ah(eh,πh(ehφ))+Cγ1(eh,0)s2.\alpha\left\|e_{h}\right\|_{\varphi}^{2}\leq a_{h}\left(e_{h},\pi_{h}\left(e_{h}\varphi\right)\right)+C\gamma^{-1}\left\|\left(e_{h},0\right)\right\|_{s}^{2}.

By Cauchy-Schwarz combined with Lemma 6 and Young’s inequality

s(zh,πh(ehφ))Cε11(0,zh)s2+ε1(ehφ2+(eh,0)s2),-s_{*}\left(z_{h},\pi_{h}\left(e_{h}\varphi\right)\right)\leq C\varepsilon_{1}^{-1}\left\|\left(0,z_{h}\right)\right\|_{s}^{2}+\varepsilon_{1}\left(\left\|e_{h}\right\|_{\varphi}^{2}+\left\|\left(e_{h},0\right)\right\|_{s}^{2}\right),

for some 0<ε1<α/20<\varepsilon_{1}<\alpha/2, hence

α2ehφ2ah(eh,πh(ehφ))+s(zh,πh(ehφ))+Cε11(0,zh)s2+C(eh,0)s2.\frac{\alpha}{2}\left\|e_{h}\right\|_{\varphi}^{2}\leq a_{h}\left(e_{h},\pi_{h}\left(e_{h}\varphi\right)\right)+s_{*}\left(z_{h},\pi_{h}\left(e_{h}\varphi\right)\right)+C\varepsilon_{1}^{-1}\left\|\left(0,z_{h}\right)\right\|_{s}^{2}+C\left\|\left(e_{h},0\right)\right\|_{s}^{2}.

Applying the first equality of the consistency Lemma 3 we obtain

α2ehφ2ah(πhuu,πh(ehφ))+Cε11(0,zh)s2+C(eh,0)s2.\frac{\alpha}{2}\left\|e_{h}\right\|_{\varphi}^{2}\leq a_{h}\left(\pi_{h}u-u,\pi_{h}\left(e_{h}\varphi\right)\right)+C\varepsilon_{1}^{-1}\left\|\left(0,z_{h}\right)\right\|_{s}^{2}+C\left\|\left(e_{h},0\right)\right\|_{s}^{2}. (31)

Since πhuuVh\pi_{h}u-u\in V_{h}^{\perp} we may apply Lemma 4 to bound

ah(πhuu,πh(ehφ))Cπhuu(0,πh(ehφ))s.a_{h}\left(\pi_{h}u-u,\pi_{h}\left(e_{h}\varphi\right)\right)\leq C\left\|\pi_{h}u-u\right\|_{\sharp}\left\|\left(0,\pi_{h}\left(e_{h}\varphi\right)\right)\right\|_{s}.

From Lemma 6 and Young’s inequality we thus have that for some ε2>0\varepsilon_{2}>0,

ah(πhuu,πh(ehφ))C((1+ε21)πhuu2+(eh,0)s2+ε2ehφ2).a_{h}\left(\pi_{h}u-u,\pi_{h}\left(e_{h}\varphi\right)\right)\leq C\left(\left(1+\varepsilon_{2}^{-1}\right)\left\|\pi_{h}u-u\right\|_{\sharp}^{2}+\left\|\left(e_{h},0\right)\right\|_{s}^{2}+\varepsilon_{2}\left\|e_{h}\right\|_{\varphi}^{2}\right).

Taking ε2<α/4\varepsilon_{2}<\alpha/4 and combining the above bound with (31) we see that

α4ehφ2C((1+ε21)πhuu2+(1+ε11)(eh,zh)s2).\frac{\alpha}{4}\left\|e_{h}\right\|_{\varphi}^{2}\leq C\left(\left(1+\varepsilon_{2}^{-1}\right)\left\|\pi_{h}u-u\right\|_{\sharp}^{2}+\left(1+\varepsilon_{1}^{-1}\right)\left\|\left(e_{h},z_{h}\right)\right\|_{s}^{2}\right).

Since ε1,2\varepsilon_{1,2} are independent of hh we can absorb them in the generic constant CC and using the approximation inequality (21) together with Proposition 2, we conclude that

ehφ\displaystyle\left\|e_{h}\right\|_{\varphi} C(μ12h+|β|12h32)(|u|H2(Ω)+h2δω)\displaystyle\leq C\left(\mu^{\frac{1}{2}}h+|\beta|^{\frac{1}{2}}h^{\frac{3}{2}}\right)\left(|u|_{H^{2}(\Omega)}+h^{-2}\|\delta\|_{\omega}\right)
C(|β|12h32|u|H2(Ω)+|β|12h12δω),\displaystyle\leq C\left(|\beta|^{\frac{1}{2}}h^{\frac{3}{2}}|u|_{H^{2}(\Omega)}+|\beta|^{\frac{1}{2}}h^{-\frac{1}{2}}\|\delta\|_{\omega}\right),

where we used that Pe(h)>1Pe(h)>1. \square
4.4. Upstream estimates. In this case we consider β=(β1,0)\beta=\left(\beta_{1},0\right) with β1<0\beta_{1}<0 and the data set

ω=(0,x)×(y,y+)\omega=(0,x)\times\left(y^{-},y^{+}\right)

touching part of the outflow boundary Ω+\partial\Omega^{+}. We must choose the weight function differently and this time we take a negative φ\varphi given by (9)

φ:=ψ1ψ2(1,0)\varphi:=\psi_{1}\psi_{2}\in(-1,0)

It seems that in this case we can not simultaneously get control of the L2L^{2}-norm and the weighted H1H^{1}-norm and we have to sacrifice the latter since it is not uniform in μ\mu. We now take the weighted triple norm to be

vhφ2:=|β|12vh|φ|12Ω2+|βn|12vh|φ|12Ω2,\left\|v_{h}\right\|_{\varphi}^{2}:=\left\||\beta|^{\frac{1}{2}}v_{h}|\varphi|^{\frac{1}{2}}\right\|_{\Omega}^{2}+\left\||\beta\cdot n|^{\frac{1}{2}}v_{h}|\varphi|^{\frac{1}{2}}\right\|_{\partial\Omega^{-}}^{2}, (32)

and rederive the results obtained in Section 4.3, aiming for a local error estimate. Since φ(1,0)\varphi\in(-1,0), we will use that φΩ|φ|12Ω\|\cdot\varphi\|_{\Omega}\leq\left\|\cdot|\varphi|^{\frac{1}{2}}\right\|_{\Omega}.

We start with an analogue of Lemma 5 by taking vhφv_{h}\varphi as a test function in the weak bilinear form aha_{h} and notice that since φ<0\varphi<0 we now have that

12(|βn|12vh|φ|12Ω2+|β|vh|φ|12Ω2)=(βvh,vhφ)Ω+12|βn|12vh|φ|12Ω+2\frac{1}{2}\left(\left\||\beta\cdot n|^{\frac{1}{2}}v_{h}|\varphi|^{\frac{1}{2}}\right\|_{\partial\Omega^{-}}^{2}+|\beta|\left\|v_{h}|\varphi|^{\frac{1}{2}}\right\|_{\Omega}^{2}\right)=\left(\beta\cdot\nabla v_{h},v_{h}\varphi\right)_{\Omega}+\frac{1}{2}\left\||\beta\cdot n|^{\frac{1}{2}}v_{h}|\varphi|^{\frac{1}{2}}\right\|_{\partial\Omega^{+}}^{2}

Arguing as previously in (24) but now for the outflow boundary, we obtain the bound

|βn|12vh|φ|12Ω+\displaystyle\left\||\beta\cdot n|^{\frac{1}{2}}v_{h}|\varphi|^{\frac{1}{2}}\right\|_{\partial\Omega^{+}} C|β|12(vh|φ|12Ω+ωβ+vh|φ|12Ω+\ωβ)\displaystyle\leq C|\beta|^{\frac{1}{2}}\left(\left\|v_{h}|\varphi|^{\frac{1}{2}}\right\|_{\partial\Omega^{+}\cap\omega_{\beta}}+\left\|v_{h}|\varphi|^{\frac{1}{2}}\right\|_{\partial\Omega^{+}\backslash\omega_{\beta}}\right)
C|β|12h12vhω+C|β|12h32vhH1(Ω)\displaystyle\leq C|\beta|^{\frac{1}{2}}h^{-\frac{1}{2}}\left\|v_{h}\right\|_{\omega}+C|\beta|^{\frac{1}{2}}h^{\frac{3}{2}}\left\|v_{h}\right\|_{H^{1}(\Omega)} (33)
Cγ12(vh,0)s\displaystyle\leq C\gamma^{-\frac{1}{2}}\left\|\left(v_{h},0\right)\right\|_{s}

and thus

12(|β|vh|φ|12Ω2+|βn|12vh|φ|12Ω2)(βvh,vhφ)Ω+Cγ1(vh,0)s2.\frac{1}{2}\left(|\beta|\left\|v_{h}|\varphi|^{\frac{1}{2}}\right\|_{\Omega}^{2}+\left\||\beta\cdot n|^{\frac{1}{2}}v_{h}|\varphi|^{\frac{1}{2}}\right\|_{\partial\Omega^{-}}^{2}\right)\leq\left(\beta\cdot\nabla v_{h},v_{h}\varphi\right)_{\Omega}+C\gamma^{-1}\left\|\left(v_{h},0\right)\right\|_{s}^{2}. (34)

For the diffusion term we no longer have any positive contribution due to the change in sign of the weight function φ\varphi, since now

(μvh,(vhφ))Ω=μ12vh|φ|12Ω2+(μvh,vhφ)Ω\left(\mu\nabla v_{h},\nabla\left(v_{h}\varphi\right)\right)_{\Omega}=-\left\|\mu^{\frac{1}{2}}\nabla v_{h}|\varphi|^{\frac{1}{2}}\right\|_{\Omega}^{2}+\left(\mu\nabla v_{h},v_{h}\nabla\varphi\right)_{\Omega}

We must therefore control this entirely using the stabilization. Integrating by parts and using the weighted trace inequality (19)

(μvh,(vhφ))Ωμvhn,vhφΩ\displaystyle\left(\mu\nabla v_{h},\nabla\left(v_{h}\varphi\right)\right)_{\Omega}-\left\langle\mu\nabla v_{h}\cdot n,v_{h}\varphi\right\rangle_{\partial\Omega} =FiFμvhnvhφds\displaystyle=\sum_{F\in\mathcal{F}_{i}}\int_{F}\mu\llbracket\nabla v_{h}\cdot n\rrbracket v_{h}\varphi\mathrm{~d}s
Cγ12sΩ(vh,vh)12μ12h1vhφΩ\displaystyle\leq C\gamma^{-\frac{1}{2}}s_{\Omega}\left(v_{h},v_{h}\right)^{\frac{1}{2}}\mu^{\frac{1}{2}}h^{-1}\left\|v_{h}\varphi\right\|_{\Omega}
Cγ12sΩ(vh,vh)12μ12h1vh|φ|12Ω\displaystyle\leq C\gamma^{-\frac{1}{2}}s_{\Omega}\left(v_{h},v_{h}\right)^{\frac{1}{2}}\mu^{\frac{1}{2}}h^{-1}\left\|v_{h}|\varphi|^{\frac{1}{2}}\right\|_{\Omega}

To bound this by the triple norm we can simply use that |φ|<1|\varphi|<1 and μ|β|h\mu\leq|\beta|h, giving that μ12h1|φ|12|β|12h12\mu^{\frac{1}{2}}h^{-1}|\varphi|^{\frac{1}{2}}\leq|\beta|^{\frac{1}{2}}h^{-\frac{1}{2}}. Hence we have that for some ε>0\varepsilon>0,

|(μvh,(vhφ))Ωμvhn,vhφΩ|Cε1γ1h1sΩ(vh,vh)+Cεvhφ2.\left|\left(\mu\nabla v_{h},\nabla\left(v_{h}\varphi\right)\right)_{\Omega}-\left\langle\mu\nabla v_{h}\cdot n,v_{h}\varphi\right\rangle_{\partial\Omega}\right|\leq C\varepsilon^{-1}\gamma^{-1}h^{-1}s_{\Omega}\left(v_{h},v_{h}\right)+C\varepsilon\left\|v_{h}\right\|_{\varphi}^{2}.

However, when Pe(h)h>1Pe(h)h>1 one can obtain a better estimate due to μ12h1|φ|12|β|12\mu^{\frac{1}{2}}h^{-1}|\varphi|^{\frac{1}{2}}\leq|\beta|^{\frac{1}{2}}, which gives that

|(μvh,(vhφ))Ωμvhn,vhφΩ|Cε1γ1sΩ(vh,vh)+Cεvhφ2\left|\left(\mu\nabla v_{h},\nabla\left(v_{h}\varphi\right)\right)_{\Omega}-\left\langle\mu\nabla v_{h}\cdot n,v_{h}\varphi\right\rangle_{\partial\Omega}\right|\leq C\varepsilon^{-1}\gamma^{-1}s_{\Omega}\left(v_{h},v_{h}\right)+C\varepsilon\left\|v_{h}\right\|_{\varphi}^{2}

Summing these contributions we obtain the following result corresponding to Lemma 5.

Lemma 7. There exists α>0\alpha>0 such that for all vhVhv_{h}\in V_{h} we have

αvhφ2ah(vh,vhφ)+Ch1(vh,0)s2, when 1Pe(h)<h1,\alpha\left\|v_{h}\right\|_{\varphi}^{2}\leq a_{h}\left(v_{h},v_{h}\varphi\right)+Ch^{-1}\left\|\left(v_{h},0\right)\right\|_{s}^{2},\text{ when }1\lesssim Pe(h)<h^{-1},

and

αvhφ2ah(vh,vhφ)+C(vh,0)s2, when Pe(h)>h1.\alpha\left\|v_{h}\right\|_{\varphi}^{2}\leq a_{h}\left(v_{h},v_{h}\varphi\right)+C\left\|\left(v_{h},0\right)\right\|_{s}^{2},\text{ when }Pe(h)>h^{-1}.

Again, we can refine the control over the triple norm vhφ\left\|v_{h}\right\|_{\varphi} by taking the projection πh(vhφ)Vh\pi_{h}\left(v_{h}\varphi\right)\in V_{h} as a test function and we obtain corresponding results.
Corollary 2. There exists α>0\alpha>0 such that for all vhVhv_{h}\in V_{h} we have

αvhφ2ah(vh,πh(vhφ))+Ch1(vh,0)s2, when 1Pe(h)<h1,\alpha\left\|v_{h}\right\|_{\varphi}^{2}\leq a_{h}\left(v_{h},\pi_{h}\left(v_{h}\varphi\right)\right)+Ch^{-1}\left\|\left(v_{h},0\right)\right\|_{s}^{2},\text{ when }1\lesssim Pe(h)<h^{-1},

and

αvhφ2ah(vh,πh(vhφ))+C(vh,0)s2, when Pe(h)>h1.\alpha\left\|v_{h}\right\|_{\varphi}^{2}\leq a_{h}\left(v_{h},\pi_{h}\left(v_{h}\varphi\right)\right)+C\left\|\left(v_{h},0\right)\right\|_{s}^{2},\text{ when }Pe(h)>h^{-1}.

Proof. The argument in the proof of Corollary 1 remains valid with the remark that we now use the inequality |φ|<|φ|12|\varphi|<|\varphi|^{\frac{1}{2}}. \square

Lemma 8. For all vhVhv_{h}\in V_{h} there holds

(0,πh(vhφ))s2C(h1vhφ2+(vh,0)s2), when 1Pe(h)<h1,\left\|\left(0,\pi_{h}\left(v_{h}\varphi\right)\right)\right\|_{s}^{2}\leq C\left(h^{-1}\left\|v_{h}\right\|_{\varphi}^{2}+\left\|\left(v_{h},0\right)\right\|_{s}^{2}\right),\text{ when }1\lesssim Pe(h)<h^{-1},

and

(0,πh(vhφ))s2C(vhφ2+(vh,0)s2), when Pe(h)>h1.\left\|\left(0,\pi_{h}\left(v_{h}\varphi\right)\right)\right\|_{s}^{2}\leq C\left(\left\|v_{h}\right\|_{\varphi}^{2}+\left\|\left(v_{h},0\right)\right\|_{s}^{2}\right),\text{ when }Pe(h)>h^{-1}.

Proof. We follow the proof of Lemma 6 and we focus on the bounds that are now different. As before, by the triangle inequality we have that up to a constant

(0,πh(vhφ))s\displaystyle\left\|\left(0,\pi_{h}\left(v_{h}\varphi\right)\right)\right\|_{s}\leq μ12(πh1)(vhφ))Ω+μ12vhφΩ+μ12vhφΩ\displaystyle\left.\|\mu^{\frac{1}{2}}\nabla\left(\pi_{h}-1\right)\left(v_{h}\varphi\right)\right)\left\|{}_{\Omega}+\right\|\mu^{\frac{1}{2}}v_{h}\nabla\varphi\left\|{}_{\Omega}+\right\|\mu^{\frac{1}{2}}\nabla v_{h}\varphi\|_{\Omega}
+(|β|+μh1)12((πh1)(vhφ)Ω+vhφΩ)\displaystyle+\left(|\beta|+\mu h^{-1}\right)^{\frac{1}{2}}\left(\left\|\left(\pi_{h}-1\right)\left(v_{h}\varphi\right)\right\|_{\partial\Omega}+\left\|v_{h}\varphi\right\|_{\partial\Omega}\right)
+sΩ(πh(vhφ),πh(vhφ))12\displaystyle+s_{\Omega}\left(\pi_{h}\left(v_{h}\varphi\right),\pi_{h}\left(v_{h}\varphi\right)\right)^{\frac{1}{2}}

The first two terms can be bounded by CvhφC\left\|v_{h}\right\|_{\varphi} as previously. For the third one, we can use the inverse inequality (12) and (18) to obtain

μ12vhφΩCμ12h1φ,ΩvhΩCμ12h1vhφΩCμ12h1vh|φ|12Ω.\left\|\mu^{\frac{1}{2}}\nabla v_{h}\varphi\right\|_{\Omega}\leq C\mu^{\frac{1}{2}}h^{-1}\|\varphi\|_{\infty,\Omega}\left\|v_{h}\right\|_{\Omega}\leq C\mu^{\frac{1}{2}}h^{-1}\left\|v_{h}\varphi\right\|_{\Omega}\leq C\mu^{\frac{1}{2}}h^{-1}\left\|v_{h}|\varphi|^{\frac{1}{2}}\right\|_{\Omega}.

Hence we have that

μ12vhφΩCh12vhφ, when 1Pe(h)<h1,\left\|\mu^{\frac{1}{2}}\nabla v_{h}\varphi\right\|_{\Omega}\leq Ch^{-\frac{1}{2}}\left\|v_{h}\right\|_{\varphi},\text{ when }1\lesssim Pe(h)<h^{-1},

and

μ12vhφΩCvhφ, when Pe(h)>h1.\left\|\mu^{\frac{1}{2}}\nabla v_{h}\varphi\right\|_{\Omega}\leq C\left\|v_{h}\right\|_{\varphi},\text{ when }\operatorname{Pe}(h)>h^{-1}.

Arguing as previously, we can bound the second line by CvhφC\left\|v_{h}\right\|_{\varphi} using (33) instead of (24). We conclude the proof by recalling the estimate (30) for the jump term and the subsequent bounds. \square

We now prove the weighted error estimate in the upstream case showing that in the stability region ωβ\stackrel_{\beta} we have quasi-optimal convergence for high Péclet numbers and a reduction of the convergence order by 𝒪(h12)\mathcal{O}\left(h^{\frac{1}{2}}\right) in an intermediate regime.
Theorem 2. Let uH2(Ω)u\in H^{2}(\Omega) be the solution of (1) and (uh,zh)[Vh]2\left(u_{h},z_{h}\right)\in\left[V_{h}\right]^{2} the solution to (5), then there holds

uuhφC(|β|12h|u|H2(Ω)+|β|12h1δω), when 1Pe(h)<h1,\left\|u-u_{h}\right\|_{\varphi}\leq C\left(|\beta|^{\frac{1}{2}}h|u|_{H^{2}(\Omega)}+|\beta|^{\frac{1}{2}}h^{-1}\|\delta\|_{\omega}\right),\text{ when }1\lesssim Pe(h)<h^{-1},

and

uuhφC(|β|12h32|u|H2(Ω)+|β|12h12δω), when Pe(h)>h1.\left\|u-u_{h}\right\|_{\varphi}\leq C\left(|\beta|^{\frac{1}{2}}h^{\frac{3}{2}}|u|_{H^{2}(\Omega)}+|\beta|^{\frac{1}{2}}h^{-\frac{1}{2}}\|\delta\|_{\omega}\right),\text{ when }Pe(h)>h^{-1}.

Proof. We combine Lemma 7, Corollary 2 and Lemma 8 as in the proof of Theorem 1 and note that the argument holds verbatim when Pe(h)>h1Pe(h)>h^{-1}. Observe that when 1Pe(h)<h11\lesssim Pe(h)<h^{-1} we similarly obtain for some α>0\alpha>0 and 0<ε1<α/20<\varepsilon_{1}<\alpha/2,

α2ehφ2ah(πhuu,πh(ehφ))+Cε11(0,zh)s2+Ch1(eh,0)s2.\frac{\alpha}{2}\left\|e_{h}\right\|_{\varphi}^{2}\leq a_{h}\left(\pi_{h}u-u,\pi_{h}\left(e_{h}\varphi\right)\right)+C\varepsilon_{1}^{-1}\left\|\left(0,z_{h}\right)\right\|_{s}^{2}+Ch^{-1}\left\|\left(e_{h},0\right)\right\|_{s}^{2}. (35)

Since πhuuVh\pi_{h}u-u\in V_{h}^{\perp} we may apply Lemma 4 to bound

ah(πhuu,πh(ehφ))Cπhuu(0,πh(ehφ))s.a_{h}\left(\pi_{h}u-u,\pi_{h}\left(e_{h}\varphi\right)\right)\leq C\left\|\pi_{h}u-u\right\|_{\sharp}\left\|\left(0,\pi_{h}\left(e_{h}\varphi\right)\right)\right\|_{s}.

From Lemma 8 and Young’s inequality we thus have that for some ε2>0\varepsilon_{2}>0,

ah(πhuu,πh(ehφ))C((1+ε21h1)πhuu2+(eh,0)s2+ε2ehφ2).a_{h}\left(\pi_{h}u-u,\pi_{h}\left(e_{h}\varphi\right)\right)\leq C\left(\left(1+\varepsilon_{2}^{-1}h^{-1}\right)\left\|\pi_{h}u-u\right\|_{\sharp}^{2}+\left\|\left(e_{h},0\right)\right\|_{s}^{2}+\varepsilon_{2}\left\|e_{h}\right\|_{\varphi}^{2}\right).

Taking ε2<α/4\varepsilon_{2}<\alpha/4 and combining the above bound with (35) we see that

α4ehφ2C((1+ε21h1)πhuu2+ε11h1(eh,zh)s2).\frac{\alpha}{4}\left\|e_{h}\right\|_{\varphi}^{2}\leq C\left(\left(1+\varepsilon_{2}^{-1}h^{-1}\right)\left\|\pi_{h}u-u\right\|_{\sharp}^{2}+\varepsilon_{1}^{-1}h^{-1}\left\|\left(e_{h},z_{h}\right)\right\|_{s}^{2}\right).

Since ε1,2\varepsilon_{1,2} are independent of hh we can absorb them in the constant CC and conclude the proof by using the approximation inequality (21) and Proposition 2 to obtain that

ehφ\displaystyle\left\|e_{h}\right\|_{\varphi} C(μ12h12+|β|12h)(|u|H2(Ω)+h2δω)\displaystyle\leq C\left(\mu^{\frac{1}{2}}h^{\frac{1}{2}}+|\beta|^{\frac{1}{2}}h\right)\left(|u|_{H^{2}(\Omega)}+h^{-2}\|\delta\|_{\omega}\right)
C(|β|12h|u|H2(Ω)+|β|12h1δω)\displaystyle\leq C\left(|\beta|^{\frac{1}{2}}h|u|_{H^{2}(\Omega)}+|\beta|^{\frac{1}{2}}h^{-1}\|\delta\|_{\omega}\right)

The error bounds in this section contain a global Sobolev norm. This may be large in the presence of layers and it would be optimal to replace it with a local regularity measure. However, it is not clear how to reconcile this goal with the ill-posed character of the problem. Inserting everywhere the weight function as in the well-posed case [BGL09], would perturb the stability of the optimality system and, due to the lack of physical coercivity, the residual terms would then be hard to control. One could also consider a reduced transport equation (with zero diffusivity) and define the far field, where the solution is unknown, by a smooth extension, but is not obvious how this could be constructed in our context, and without requiring regularity for the right-hand side. Nonetheless, in numerical experiments we observe a good regularity behaviour, i.e. layers do not pollute the solution in the stability zone, as shown in Figures 15 and 16.

5. Numerical experiments

Refer to caption
Figure 4: Figure 4. Data set ω\omega and error measurement regions (hatched).

We let Ω\Omega be the unit square and illustrate the performance of the numerical method (5) for different locations of the data domain ω\omega and different regions of interest where we measure the approximation error. The computational domains are given in Figure 4 and the implementation is done using FEniCS [ABH+15]\left[\mathrm{ABH}^{+}15\right]. In all the examples below we have used uniform triangulations with alternating left/right diagonals. In the definition of sΩs_{\Omega} and ss_{*} we have taken the stabilization parameters γ=105\gamma=10^{-5} and γ=1\gamma_{*}=1, and ζ=2\zeta=2 for sωs_{\omega}. The effect of different combinations of γ\gamma and γ\gamma_{*} on the L2L^{2}-errors is shown in Figure 5 and Figure 6 when data is given in a centered disk. Similar results are obtained when the data set is near the inflow/outflow boundary. Notice that our choice is empirically close to being optimal both locally and globally.

Refer to caption
Figure 5: Figure 5. Varying the stabilization parameters γ\gamma and $\gamma_{*

$. Absolute L2L^{2}-errors downstream, computational domains in Figure 4a. β=(1,0)\beta=(1,0), exact solution u=2sin(5πx)sin(5πy)u=2\sin(5\pi x)\sin(5\pi y). Similar results in the upstream case.}

Refer to caption
Figure 6: Figure 6. Varying the stabilization parameters γ\gamma and $\gamma_{*

$. Absolute L2L^{2}-errors globally, computational domains in Figure 4a. β=(1,0)\beta=(1,0), exact solution u=2sin(5πx)sin(5πy)u=2\sin(5\pi x)\sin(5\pi y).}

We first show convergence plots both downstream and upstream from the data set when varying the diffusion coefficient μ\mu and keeping the convection field β\beta fixed. As
in the case of well-posed convection-dominated problems, the observed L2L^{2}-convergence order is typically 𝒪(h2)\mathcal{O}\left(h^{2}\right), surpassing by 𝒪(h12)\mathcal{O}\left(h^{\frac{1}{2}}\right) the weighted error estimates proven for general meshes (see Figure 7, for example).

Refer to caption
Figure 7: Figure 7. Absolute $Lˆ{2

$-errors against mesh size hh when varying the diffusion coefficient μ\mu for fixed β=(1,0)\beta=(1,0), exact solution u=2sin(5πx)sin(5πy)u=2\sin(5\pi x)\sin(5\pi y).}

Refer to caption
Figure 8: Figure 8. Absolute $Hˆ{1

$-errors against mesh size hh when varying the diffusion coefficient μ\mu for fixed β=(1,0)\beta=(1,0), exact solution u=2sin(5πx)sin(5πy)u=2\sin(5\pi x)\sin(5\pi y).}

5.1. Data set near the inflow/outflow boundary. We consider the data set ω\omega near the inflow and outflow boundaries of Ω\Omega, as assumed in Section 2.1. We observe in Figure 7 that as diffusion is reduced the convergence order for the L2L^{2}-errors increases, culminating with quadratic convergence when convection dominates. Confirming the theoretical analysis in Section 4.4, we note the presence of an intermediate regime for Péclet numbers in which the upstream convergence orders are reduced and the upstream errors are typically larger. This can also be seen in Figure 9 where we consider the diffusion coefficient μ=102\mu=10^{-2} and an interior data set. The errors in the H1H^{1}-seminorm are given in Figure 8 which shows almost linear convergence. This is probably due to the small contribution of the gradient term in the triple norm (23).

Refer to caption
Figure 9: Figure 9. Absolute $Lˆ{2

$-errors against mesh size hh, downstream vs upstream for μ=102,β=(1,0)\mu=10^{-2},\beta=(1,0), exact solution u=2sin(5πx)sin(5πy)u=2\sin(5\pi x)\sin(5\pi y).}

Refer to caption
Figure 10: Figure 10. Absolute $Lˆ{2

$-errors against mesh size hh, computational domains in Figure 4a. Varying the diffusion coefficient μ\mu for fixed β=(1,0)\beta=(1,0), exact solution u=2sin(5πx)sin(5πy)u=2\sin(5\pi x)\sin(5\pi y).}

5.2. Interior data set. Next we consider the setting of the example discussed in the Introduction (Figure 2), where data is given in the centre of the domain. We give the convergence of the L2L^{2}-errors in Figure 10 with the caveat that this location of the data set ω\omega is not rigorously covered by the theoretical analysis of the previous sections. Nonetheless, the experiments are in agreement with the proven results. Notice that the L2L^{2}-convergence is faster as μ\mu decreases and for high Péclet numbers (above 10) one has optimal quadratic convergence both downstream and upstream, with the distinction that in the upstream case the convergence order is reduced in an intermediate regime, in agreement with the theoretical results in Section 4. Also, as expected from the error estimates proven in the first part [BNO20], when diffusion is moderately small one can see the transition towards the diffusion-dominated regime as the mesh gets refined - the convergence changes from almost quadratic to sublinear as the Péclet number decreases below 1. Figure 11 shows almost linear convergence in the H1H^{1}-seminorm. We think this is observed due to the small contribution of the gradient term in the triple norm (23). We also remark almost no distinction between upstream and downstream for this
example, probably because the gradient term is controlled by the L2L^{2}-norm for small enough μ\mu.

Refer to caption
Figure 11: Figure 11. $Hˆ{1

$-errors against mesh size hh, computational domains in Figure 4a. Varying the diffusion coefficient μ\mu for fixed β=(1,0)\beta=(1,0), exact solution u=2sin(5πx)sin(5πy)u=2\sin(5\pi x)\sin(5\pi y).}

Refer to caption
Figure 12: Figure 12. $Lˆ{2

$-errors against mesh size hh for perturbations in data, computational domains in Figure 4a. Varying the diffusion coefficient μ\mu for fixed β=(1,0)\beta=(1,0), exact solution u=2sin(5πx)sin(5πy)u=2\sin(5\pi x)\sin(5\pi y).}

5.3. Data perturbations. We demonstrate the effect of data perturbations U~ω=u|ω+δ\tilde{U}_{\omega}=\left.u\right|_{\omega}+\delta in a downstream vs upstream setting by polluting the restriction of uu to each node of the mesh in ω\omega with uniformly distributed values in [h2,h2],[h,h]\left[-h^{2},h^{2}\right],[-h,h] and [h12,h12]\left[-h^{\frac{1}{2}},h^{\frac{1}{2}}\right], respectively. Comparing first Figure 12 to Figure 10 we see that perturbations of amplitude 𝒪(h2)\mathcal{O}\left(h^{2}\right) have no effect on the L2L^{2}-errors, as expected.

An 𝒪(h)\mathcal{O}(h) noise amplitude exhibits in Figure 13 the difference - proven in Theorem 1 and Theorem 2 - between the downstream and upstream scenarios. In the upstream case the noise has a strong effect for moderate Péclet numbers and the errors stagnate. Only for high Péclet numbers one has convergence of order 𝒪(h12)\mathcal{O}\left(h^{\frac{1}{2}}\right). In the downstream case one observes lower errors, faster convergence and almost no noise effect for high

Péclet numbers. The difference is also very clear for perturbations of amplitude 𝒪(h12)\mathcal{O}\left(h^{\frac{1}{2}}\right) shown in Figure 14. In the upstream case the errors stagnate and there seems to be no convergence, while in the downstream case the errors still convergence for high Péclet numbers.

Refer to caption
Figure 13: Figure 13. $Lˆ{2

$-errors against mesh size hh for perturbations in data, computational domains in Figure 4a. Varying the diffusion coefficient μ\mu for fixed β=(1,0)\beta=(1,0), exact solution u=2sin(5πx)sin(5πy)u=2\sin(5\pi x)\sin(5\pi y).}

Refer to caption
Figure 14: Figure 14. $Lˆ{2

$-errors against mesh size hh for perturbations in data, computational domains in Figure 4a. Varying the diffusion coefficient μ\mu for fixed β=(1,0)\beta=(1,0), exact solution u=2sin(5πx)sin(5πy)u=2\sin(5\pi x)\sin(5\pi y).}

5.4. Internal layer example. We now consider an exact solution u=sin(3πx)+tanh(100(y1/2))u=\sin(3\pi x)+\tanh(100(y-1/2)) having an internal layer at y=1/2y=1/2 and study qualitatively the transition from dominant diffusion to dominant convection. Data is given on both sides of the layer. The distribution of the absolute error is presented in Figure 15 for the diffusion-dominated regime and in Figure 16 for the intermediate and convection dominated regimes. Note that the width of the internal layer does not depend on the physical parameters. Initially, the errors oscillate away from the data sets and concentrate around the boundary of the domain. When convection dominates, the
approximation around the layer strongly deteriorates due to the crosswind position relative to the data sets. In this example the mesh is unstructured with 512 elements on a side and h0.0025h\approx 0.0025.

Refer to caption
Figure 15: Figure 15. Absolute error in the diffusion-dominated regime, μ=1,β=(1,0)\mu=1,\beta=(1,0). Data given in the four outlined boxes for the solution u=sin(3πx)+tanh(100(y1/2))u=\sin(3\pi x)+\tanh(100(y-1/2)).
Refer to caption
Figure 16: Figure 16. Absolute error in transition to the convection-dominated regime, β=(1,0)\beta=(1,0). Data given in the four outlined boxes for the solution u=sin(3πx)+tanh(100(y1/2))u=\sin(3\pi x)+\tanh(100(y-1/2)).

References

[ABH+15]\left[\mathrm{ABH}^{+}15\right] M. S. Alnæs, J. Blechta, J. Hake, A. Johansson, B. Kehlet, A. Logg, C. Richardson, J. Ring, M. E. Rognes, and G. N. Wells. The FEniCS Project Version 1.5. Archive of Numerical Software, 3(100):9-23, 2015.
[Ber99] S. Bertoluzza. The discrete commutator property of approximation spaces. C. R. Acad. Sci. Paris Sér. I Math., 329(12):1097-1102, 1999.
[BGL09] E. Burman, J. Guzmán, and D. Leykekhman. Weighted error estimates of the continuous interior penalty method for singularly perturbed problems. IMA J. Numer. Anal., 29(2):284314, 2009.
[BHL18] E. Burman, P. Hansbo, and M. G. Larson. Solving ill-posed control problems by stabilized finite element methods: an alternative to Tikhonov regularization. Inverse Problems, 34:035004, 2018.
[BNO20] E. Burman, M. Nechita, and L. Oksanen. A stabilized finite element method for inverse problems subject to the convection-diffusion equation. I: diffusion-dominated regime. NuNu- mer. Math., 144(451-477), 2020.
[Bur05] E. Burman. A unified analysis for conforming and nonconforming stabilized finite element methods using interior penalty. SIAM J. Numer. Anal., 43(5):2012-2033, 2005.
[Bur13] E. Burman. Stabilized finite element methods for nonsymmetric, noncoercive, and ill-posed problems. Part I: Elliptic equations. SIAM J. Sci. Comput., 35(6):A2752-A2780, 2013.
[Bur14] E. Burman. Stabilized finite element methods for nonsymmetric, noncoercive, and ill-posed problems. Part II: Hyperbolic equations. SIAM J. Sci. Comput., 36(4):A1911-A1936, 2014.
[BV07] R. Becker and B. Vexler. Optimal control of the convection-diffusion equation using stabilized finite element methods. Numer. Math., 106(3):349-367, 2007.
[DQ05] L. Dede’ and A. Quarteroni. Optimal control and numerical adaptivity for advectiondiffusion equations. M2AN Math. Model. Numer. Anal., 39(5):1019-1040, 2005.
[EG04] A. Ern and J.-L. Guermond. Theory and practice of finite elements, volume 159 of Applied Mathematical Sciences. Springer-Verlag, New York, 2004.
[Eva10] L. C. Evans. Partial differential equations, volume 19 of Graduate Studies in Mathematics. American Mathematical Society, 2010.
[HYZ09] M. Hinze, N. Yan, and Z. Zhou. Variational discretization for optimal control governed by convection dominated diffusion equations. J. Comput. Math., 27(2-3):237-253, 2009.
[JKN18] V. John, P. Knobloch, and J. Novo. Finite elements for scalar convection-dominated equations and incompressible flow problems: a never ending story? Comput. Vis. Sci., 19(5):4763, Dec 2018.
[JNP84] C. Johnson, U. Nävert, and J. Pitkäranta. Finite element methods for linear hyperbolic problems. Comput. Methods Appl. Mech. Engrg., 45(1-3):285-312, 1984.
[MS99] P. Monk and E. Süli. The adaptive computation of far-field patterns by a posteriori error estimation of linear functionals. SIAM J. Numer. Anal., 36(1):251-274, 1999.
[NS74] J. A. Nitsche and A. H. Schatz. Interior estimates for Ritz-Galerkin methods. Math. Comp., 28(128):937-958, 1974.
[RS15] H.-G. Roos and M. Stynes. Some open questions in the numerical analysis of singularly perturbed differential equations. Comput. Methods Appl. Math., 15(4):531-550, 2015.
[Ste70] E. M. Stein. Singular integrals and differentiability properties of functions, volume 30 of Princeton Mathematical Series. Princeton University Press, 1970.
[YZ09] N. Yan and Z. Zhou. A priori and a posteriori error analysis of edge stabilization Galerkin method for the optimal control problem governed by convection-dominated diffusion equation. J. Comput. Appl. Math., 223(1):198-217, 2009.

2022

Related Posts