We consider the numerical approximation of the linear ill-posed problem of unique continuation for the Helmholtz equation. We first review the conditional stability of this problem and then discuss high-order conforming finite element methods, with regularization added on the discrete level using gradient jump penalty and Galerkin least squares. The method is shown to converge in terms of the stability of the problem, quantified by the Hölder exponent, and the polynomial degree of approximation. The analysis also takes into account possibly noisy data. Numerical examples illustrating the theory are presented for a Helmholtz version of the classical Hadamard example for two geometric configurations with different stability properties for the continuation problem, including data perturbations of different amplitudes.
Authors
Mihai Nechita
“Tiberiu Popoviciu” Institute of Numerical Analysis, Romanian Academy
[1] E. Burman, G. Delay, and A. Ern., A hybridized high-order method for unique continuation subject to the helmholtz equation, SIAM J. Numer. Anal., 59(5):2368–2392,2021.[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.
[2] S. M. Berge and E. Malinnikova, On the three ball theorem for solutions of the Helmholtz equation, Complex Anal. Synerg., 7(2):1–10, 2021.
[3] E. Burman, M. Nechita, and L. Oksanen, Unique continuation for the Helmholtz equation using stabilized finite element methods, J. Math. Pures Appl., 129:1–22, 2019.
[4] I. M. Babuska and S. A. Sauter, Is the pollution effect of the FEM avoidable for the Helmholtz equation considering high wave numbers? SIAM J. Numer. Anal., 34(6):2392–2423, 1997.
[5] 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.
[6] E. Burman, Stabilised finite element methods for ill-posed problems with conditional stability. In Building Bridges: Connections and Challenges in Modern Approaches to Numerical Partial Differential Equations, volume 114 of Lect. Notes Comput. Sci. Eng., pages 93–127. Springer, 2016.
[7] A. Ern and J.-L. Guermond, Finite elements I: Approximation and interpolation, volume 72 of Texts in Applied Mathematics, Springer Nature, 2021.
[8] T. Hrycak and V. Isakov, Increased stability in the continuation of solutions to the Helmholtz equation, Inverse Problems, 20(3):697–712, 2004.
[9] V. Isakov, Inverse problems for partial differential equations, volume 127 of Applied Mathematical Sciences, Springer, 3rd edition, 2017.
[10] J. M. Melenk and S. A. Sauter, Wavenumber explicit convergence analysis for Galerkin discretizations of the Helmholtz equation, SIAM J. Numer. Anal., 49(3):1210–1243, 2011.
[11] M. Nechita, Unique continuation problems and stabilised finite element methods, PhD thesis, University College London, 2020.
Paper (preprint) in HTML form
helmholtz_ho
On high-order conforming finite element methods for ill-posed Helmholtz problems
Mihai Nechita ^(1,2){ }^{1,2}^(1){ }^{1} Commedia, Inria Paris and Laboratoire Jacques-Louis Lions, Sorbonne Université, France^(2){ }^{2} Tiberiu Popoviciu Institute of Numerical Analysis, Romanian Academy, Cluj-Napoca, Romania
We consider the numerical approximation of the linear ill-posed problem of unique continuation for the Helmholtz equation. We first review the conditional stability of this problem and then discuss high-order conforming finite element methods, with regularization added on the discrete level using gradient jump penalty and Galerkin least squares. The method is shown to converge in terms of the stability of the problem, quantified by the Hölder exponent, and the polynomial degree of approximation. The analysis also takes into account possibly noisy data. Numerical examples illustrating the theory are presented for a Helmholtz version of the classical Hadamard example for two geometric configurations with different stability properties for the continuation problem, including data perturbations of different amplitudes.
1 Introduction
Acoustic waves are typically modeled by the Helmholtz equation when their propagation is considered in the frequency domain. For such partial differential equations (PDEs), we are interested in the numerical approximation of linear ill-posed problems appearing in the context of data assimilation, control theory and inverse problems. More precisely, we consider the unique continuation problem, in which no boundary conditions are known, but measurements of the solution are available in a subset of the domain. For an elliptic operator such data can be extended uniquely due to the unique continuation principle: our goal is to numerically approximate the solution in a target domain.
This ill-posed problem has been recently considered in [BNO19], with theoretical results addressing quantitative unique continuation (in norms suitable for numerical analysis) with explicit dependence on the frequency. The stability of the problem is conditional-one must a priori assume a certain bound on the solution-and has Hölder continuity, quantifying the degree of ill-posedness. The sensitivity of the stability estimates with respect to the frequency depends on the geometric configuration and is crucially determined by a convexity condition on the geometry of the target domain relative to the data domain. Roughly speaking, when the target domain is inside the convex hull of the data domain, the stability constant is robust in the frequency, or grows at most linearly, depending on the norm used ( L^(2)L^{2} or H^(1)H^{1} ). On the other hand, when the solution is estimated outside of the convex hull of the data domain, the stability is very sensitive to the frequency as the constants can grow superpolynomially.
From the computational point of view, [BNO19] introduced and analyzed a low-order finite element method (FEM). In brief, the Helmholtz unique continuation problem-formulated as a PDE-constrained optimization problem-is first discretized using standard piecewise affine elements and then regularization is added using established techniques from the literature on stabilized FEM. Convergence is proven using the conditional stability estimates from the continuous level. This discretize-then-regularize framework represents an alternative to the more standard Tikhonov regularization (see [BHL18] for a comparison) and, in particular, it provides a way of unifying the analysis of the discretization error and the data perturbation error. In this paper we adopt the same framework but consider high-order conforming approximations.
In general, (well-posed) Helmholtz problems are challenging to solve numerically: first, due to the oscillatory nature of the solutions, and secondly, due to the so-called pollution effect [BS97]: the constant bounding the error by the best-approximation error increases with the frequency, preventing the error estimate from being quasi-optimal. High-order methods are typically preferred for such problems since, besides the usual improvement in accuracy, the pollution effect can be reduced or, under some conditions, even eliminated when increasing the polynomial degree of the approximation [MS11]. When it comes to ill-posed problems for elliptic equations, high-order FEM were introduced in [Bur13] and subsequently analyzed in [Bur16] for conditionally stable problems (with a focus on the elliptic Cauchy problem), where it was illustrated numerically that, when dealing with such problems, faster convergence comes with a price: a higher sensitivity to possible perturbations in data. For the Helmholtz equation, a high-order hybridized discontinuous Galerkin method for the unique continuation problem was recently proposed in [BDE21]. Error estimates were proven in terms of the degree of illposedness, the polynomial order and perturbations in data (which get amplified as the order increases). However, a study of standard high-order conforming methods for such ill-posed Helmholtz problems is currently missing from the literature and the aim of this paper is to fill this gap.
2 Ill-posed problem
Figure 1: Domains for the unique continuation problem (1).
Let Omega subR^(d),d in{2,3}\Omega \subset \mathbb{R}^{d}, d \in\{2,3\}, be an open, bounded and connected set, and let omega sub Omega\omega \subset \Omega be a subset (also open, bounded and connected), as in Figure 1. For the Helmholtz equation with wave number k > 0k>0 (from now on called frequency, for convenience), we consider the unique continuation problem of finding u inH^(1)(Omega)u \in H^{1}(\Omega) such that
{[(1)-Delta u-k^(2)u=f" in "Omega","],[u=g" in "omega","]:}\left\{\begin{align*}
-\Delta u-k^{2} u=f & \text { in } \Omega, \tag{1}\\
u=g & \text { in } \omega,
\end{align*}\right.
with source term f inL^(2)(Omega)f \in L^{2}(\Omega) and interior datum g inH^(1)(omega)g \in H^{1}(\omega). Note that no boundary conditions are prescribed, and partial measurements are given in a subset of the domain. If problem (1) has a solution, then its uniqueness is guaranteed by the unique continuation principle for elliptic operators, see for example [Isa17]. We will assume that the data ff and gg are compatible, such that the function gg is the restriction on omega\omega of a solution to the Helmholtz equation with source term ff.
It is well-known, see for example [Isa17], that problem (1) is ill-posed in the sense of Hadamard: there is no uniform stability with respect to the data ff and gg. The reason for this is the fact that the operator mapping a function uu to the corresponding pair ( f,gf, g ) does not have a closed range, and its generalized inverse is, thus, discontinuous. This means that one can find functions in the kernel of the Helmholtz operator that are small in the data set omega\omega, but grow exponentially away from it, as in the example given in Section 4 concerning the Hadamard-type solutions (20). Nonetheless, assuming an additional a priori bound, the solution can be bounded by the data-this is known as conditional stability and in what follows we will present a local version of it.
2.1 Conditional stability and frequency dependence
We denote by B sub OmegaB \subset \Omega the target set containing omega\omega such that B\\ bar(omega)sub OmegaB \backslash \bar{\omega} \subset \Omega, that is B\\omegaB \backslash \omega does not touch the boundary of Omega\Omega. It is well-known, see for example [Isa17], that there exist constants C_(st)(k) > 0C_{s t}(k)>0 and alpha in(0,1]\alpha \in(0,1] such that the following Hölder stability estimate holds
for any u inH^(1)(Omega)u \in H^{1}(\Omega) satisfying (1), where the stability constant C_(st)(k)C_{s t}(k) depends on the frequency kk, typically in an implicit way. The exponent alpha in(0,1]\alpha \in(0,1] encodes the degree of ill-posedness for the continuation problem: note that alpha=1\alpha=1 would correspond to a well-posed problem with Lipschitz stability, and as alpha < 1\alpha<1 gets smaller the Hölder stability deteriorates. Both C_(st)C_{s t} and alpha\alpha depend on the geometric configuration in a nontrivial way.
Figure 2: Different frequency dependence for the stability constant. Data set omega\omega (dark grey) and target set BB (light grey). Omega\Omega is the large box.
How the stability constant C_(st)(k)C_{s t}(k) depends on the frequency kk is an important aspect for the stability of this ill-posed Helmholtz problem. For example, when there is a straight line that intersects BB but not bar(omega)\bar{\omega}, as in Figure 2a, it was proven in [BNO19, Example 4] that for any N inN,C_(st)(k) <= k^(N)N \in \mathbb{N}, C_{s t}(k) \leq k^{N} cannot hold uniformly in kk-in other words, the stability constant in (2)
grows superpolynomially in the frequency. Estimates with such a fast growth might be relevant in practice only for low frequencies, and it is thus of interest to consider obtaining estimates with a more favorable frequency dependence. Bounds that have a good behavior with respect to the frequency can be obtained under a convexity condition of the target domain BB relative to the data set omega\omega-essentially that BB is included in the convex hull of omega\omega, as for example in Figure 2b. Such a condition was first considered in [HI04], where is was shown that the stability of the solution in the L^(2)L^{2}-norm can actually improve in a certain sense as the frequency increases. In this vein, similar results that make use of a convexity condition were proven in [BNO19] for a particular geometric setting prototypical for continuation inside the convex hull of omega\omega. It was shown in [BNO19, Corollary 2] that the bound (2) is robust with respect to the frequency, i.e., it holds with the stability constant C_(st)C_{s t} independent of kk. This robust bound is also valid for other geometric configurations, as for example the one in Figure 2b.
In order to use such conditional stability estimates for the numerical analysis of a finite element method, the norms for measuring the data must be weaken: from L^(2)(Omega)L^{2}(\Omega) and H^(1)(omega)H^{1}(\omega) to H^(-1)(Omega)H^{-1}(\Omega) and L^(2)(omega)L^{2}(\omega), respectively. A result of this kind was proven in [BNO19, Corollary 3 and Lemma 2] under a geometric convexity condition. More precisely, there exist constants C > 0C>0 and alpha in(0,1]\alpha \in(0,1] such that
for any u inH^(1)(Omega)u \in H^{1}(\Omega) satisfying (1). Note that the bound is robust in the L^(2)L^{2}-norm, while the frequency dependence is linear for the H^(1)H^{1}-seminorm. Apart from the good dependence on the frequency, this kind of estimate is particularly suitable for numerical analysis since it can be directly applied to the finite element error equation.
As discussed above, when the domains do not satisfy the convexity condition, the stability constant can be sensitive to the frequency. To mark the difference, in general we write the stability estimate as follows: there exist constants C_(st)(k) > 0C_{s t}(k)>0 and alpha in(0,1]\alpha \in(0,1] such that
Let us also remark that for three-ball inequalities (where omega,B,Omega\omega, B, \Omega are concentric balls) it was recently shown in [BM21] that, when using the maximum norm, C_(st)(k)C_{s t}(k) grows exponentially in kk and this dependence is optimal.
3 Discretization, regularization and error analysis
3.1 Finite element preliminaries
From now on we will consider for simplicity that Omega subR^(d),d in{2,3}\Omega \subset \mathbb{R}^{d}, d \in\{2,3\}, is a convex polygonal/polyhedral domain. Let {T_(h)}_(h > 0)\left\{\mathcal{T}_{h}\right\}_{h>0} be a quasi-uniform family of shape-regular meshes covering Omega\Omega, each triangulation T_(h)\mathcal{T}_{h} containing elements KK with maximal diameter hh. Let pp denote the polynomial degree of approximation and let P_(p)(K)\mathbb{P}_{p}(K) be the set of polynomials of degree at most pp defined on an element KK. We consider the conforming finite element space
V_(h)^(p):={v_(h)inH^(1)(Omega):v_(h)|_(K)inP_(p)(K),AA K inT_(h)},V_{h}^{p}:=\left\{v_{h} \in H^{1}(\Omega):\left.v_{h}\right|_{K} \in \mathbb{P}_{p}(K), \forall K \in \mathcal{T}_{h}\right\},
with continuous, piecewise polynomial functions defined on the mesh T_(h)\mathcal{T}_{h}, and its subspace with homogeneous boundary conditions
For a given mesh, we collect all the interior edges/faces of the elements in the set F_(i)\mathcal{F}_{i} and denote the jump of a quantity across such an edge/face FF by [[*]]_(F)\llbracket \cdot \rrbracket_{F}, omitting the subscript whenever there is no confusion. For a discrete function v_(h)inV_(h)^(p)v_{h} \in V_{h}^{p}, we denote the jump of the normal gradient across F inF_(i)F \in \mathcal{F}_{i} by
with K_(1),K_(2)inT_(h)K_{1}, K_{2} \in \mathcal{T}_{h} being two elements such that K_(1)nnK_(2)=FK_{1} \cap K_{2}=F, and n_(1),n_(2)n_{1}, n_{2} the outward pointing normals of K_(1),K_(2)K_{1}, K_{2}, respectively.
We now recall some classical finite element results, see for example [EG21]. We will use the inverse inequality
{:(6)||v||_(L^(2)(del K)) <= C(h^(-(1)/(2))||v||_(L^(2)(K))+h^((1)/(2))||grad v||_(L^(2)(K)))","quad AA v inH^(1)(K).:}\begin{equation*}
\|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) . \tag{6}
\end{equation*}
We denote by pi_(h):L^(2)(Omega)rarrV_(h)^(p)\pi_{h}: L^{2}(\Omega) \rightarrow V_{h}^{p} the L^(2)L^{2}-projection, that satisfies the orthogonality
(u-pi_(h)u,v_(h))_(L^(2)(Omega))=0,quad AA u inL^(2)(Omega),AAv_(h)inV_(h)^(p).\left(u-\pi_{h} u, v_{h}\right)_{L^{2}(\Omega)}=0, \quad \forall u \in L^{2}(\Omega), \forall v_{h} \in V_{h}^{p} .
The L^(2)L^{2}-projection is stable in the L^(2)L^{2}-norm
||pi_(h)u||_(L^(2)(Omega)) <= ||u||_(L^(2)(Omega)),quad AA u inL^(2)(Omega)\left\|\pi_{h} u\right\|_{L^{2}(\Omega)} \leq\|u\|_{L^{2}(\Omega)}, \quad \forall u \in L^{2}(\Omega)
and, since the family {T_(h)}\left\{\mathcal{T}_{h}\right\} is quasi-uniform, it is also stable in the H^(1)H^{1}-norm
||pi_(h)u||_(H^(1)(Omega)) <= C||u||_(H^(1)(Omega)),quad AA u inH^(1)(Omega)\left\|\pi_{h} u\right\|_{H^{1}(\Omega)} \leq C\|u\|_{H^{1}(\Omega)}, \quad \forall u \in H^{1}(\Omega)
We also use the Scott-Zhang interpolant pi_(sz):H^(1)(Omega)rarrV_(h)^(p)\pi_{s z}: H^{1}(\Omega) \rightarrow V_{h}^{p} which preserves vanishing Dirichlet boundary conditions. This interpolant is stable in both the L^(2)L^{2} - and the H^(1)H^{1}-norm and enjoys the same approximation error estimate as the L^(2)L^{2}-projection
{:(7)||u-i_(h)u||_(L^(2)(Omega))+h||grad(u-i_(h)u)||_(L^(2)(Omega)) <= Ch^(m)|u|_(H^(m)(Omega))","quad AA u inH^(m)(Omega)","i_(h)=pi_(h)","pi_(sz).:}\begin{equation*}
\left\|u-i_{h} u\right\|_{L^{2}(\Omega)}+h\left\|\nabla\left(u-i_{h} u\right)\right\|_{L^{2}(\Omega)} \leq C h^{m}|u|_{H^{m}(\Omega)}, \quad \forall u \in H^{m}(\Omega), i_{h}=\pi_{h}, \pi_{s z} . \tag{7}
\end{equation*}
3.2 Discretization
We denote the standard inner product (with its induced norm) on a set SS by
(v_(h),w_(h))_(S):=int_(S)v_(h)w_(h)dx\left(v_{h}, w_{h}\right)_{S}:=\int_{S} v_{h} w_{h} \mathrm{~d} x
and define on V_(h)^(p)xxW_(h)^(p)V_{h}^{p} \times W_{h}^{p} the standard bilinear form corresponding to the Helmholtz equation
where the L^(2)L^{2}-distance to the interior datum is minimized under the constraint of the weak form of the differential equation. Introducing a test function as a Lagrange multiplier, one can consider the discrete Lagrangian functional
The solution to the above optimization problem is given by the saddle-points of L_(h)^(0)L_{h}^{0}, satisfying the Euler-Lagrange equations. However, the ill-posedness of problem (8) has not been alleviated in any way by this reformulation, and the optimality conditions lead to a linear system that might be singular.
3.3 Regularization
To mitigate the ill-posedness, one must regularize the problem by adding some penalty terms that provide well-posedness. The approach that we take herein has been introduced in [Bur13] and is a discretize-then-regularize one. We add the regularization terms on the discrete level, using techniques from stabilized FEM for numerically unstable well-posed problems: gradient jump penalty and Galerkin least squares. Apart from providing the needed stability, the regularizing terms should ideally perturb the problem in a weakly consistent way by converging to zero at optimal rates for smooth enough solutions.
For an approximation u_(h)inV_(h)^(p)u_{h} \in V_{h}^{p} and a Lagrange multiplier z_(h)inW_(h)^(p)z_{h} \in W_{h}^{p}, we consider the regularized Lagrangian
with ss and s^(**)s^{*} being the stabilization/regularization terms, to be specified later. The saddlepoints (u_(h),z_(h))inV_(h)^(p)xxW_(h)^(p)\left(u_{h}, z_{h}\right) \in V_{h}^{p} \times W_{h}^{p} of L_(h)L_{h} satisfy the optimality conditions: for any (v_(h),w_(h))inV_(h)^(p)xxW_(h)^(p)\left(v_{h}, w_{h}\right) \in V_{h}^{p} \times W_{h}^{p},
The stabilized finite element method is thus looking for (u_(h),z_(h))inV_(h)^(p)xxW_(h)^(p)\left(u_{h}, z_{h}\right) \in V_{h}^{p} \times W_{h}^{p} such that
The linear system associated to this problem can be represented in block matrix form as
[[M_(omega)+S,A^(T)],[A,-S^(**)]][(u_(h))/(z_(h))]=[[ tilde(g)+s_(u)],[f]].\left[\begin{array}{c|c}
M_{\omega}+S & A^{T} \\
\hline A & -S^{*}
\end{array}\right]\left[\frac{u_{h}}{z_{h}}\right]=\left[\begin{array}{c}
\tilde{g}+s_{u} \\
\hline f
\end{array}\right] .
To address the well-posedness of the discrete problem we must prove that it satisfies an inf-sup condition. For this, let us introduce the bilinear form
which is equivalent to the well-posedness of the discrete problem (10).
We now present some stabilization operators that satisfy the criteria discussed above. First, we define the gradient jump penalty, also known as continuous interior penalty (CIP),
J_(h)(u_(h),v_(h)):=sum_(F inF_(i))int_(F)h[[gradu_(h)*n]]_(F)[[gradv_(h)*n]]_(F)ds,\mathcal{J}_{h}\left(u_{h}, v_{h}\right):=\sum_{F \in \mathcal{F}_{i}} \int_{F} h \llbracket \nabla u_{h} \cdot n \rrbracket_{F} \llbracket \nabla v_{h} \cdot n \rrbracket_{F} \mathrm{~d} s,
to which we add the Galerkin least squares and define the bilinear form s:V_(h)^(p)xxV_(h)^(p)s: V_{h}^{p} \times V_{h}^{p} by
This represents the primal stabilization that will act on the approximation u_(h)inV_(h)^(p)u_{h} \in V_{h}^{p}. The constant parameter gamma > 0\gamma>0 is meant to tune the stabilization empirically and will be omitted from the analysis. The presence of the gradient term in ss is motivated by the a priori H^(1)H^{1} bound that the discrete solution must satisfy (uniformly in hh ) in order to be able to use the conditional stability from the continuous level, see also [Bur16]. This gradient term is not needed for piecewise affine approximations. The right-hand side term in (10) for the true solution uu is given by
With this choice, since the finite element space contains piecewise polynomials we have that ss induces a norm on V_(h)^(p)V_{h}^{p}, and also that s^(**)s^{*} induces a norm on W_(h)^(p)W_{h}^{p}. We will denote these norms by
Note that (11) is indeed a norm on V_(h)^(p)xxW_(h)^(p)V_{h}^{p} \times W_{h}^{p}, thus guaranteeing the inf-sup condition (12) is satisfied. This means that the discrete problem is well-posed.
From the standard approximation (7) and the trace inequality (6), we have the approximation bound in the stabilization norm
{:(16)||u-pi_(h)u||_(V_(h)^(p)) <= Ch^(p)||u||_(H^(p+1)(Omega)):}\begin{equation*}
\left\|u-\pi_{h} u\right\|_{V_{h}^{p}} \leq C h^{p}\|u\|_{H^{p+1}(\Omega)} \tag{16}
\end{equation*}
Let us now motivate the choice for the gradient jump penalty J_(h)\mathcal{J}_{h} and the Galerkin least squares stabilization in the primal stabilizer (13) by the following continuity result.
Lemma 1. Let u inH^(1)(Omega),u_(h)inV_(h)^(p),w inH_(0)^(1)(Omega)u \in H^{1}(\Omega), u_{h} \in V_{h}^{p}, w \in H_{0}^{1}(\Omega) and the norm ||w||_(#):=h^(-1)||w||_(Omega)+||grad w||_(Omega)\|w\|_{\#}:=h^{-1}\|w\|_{\Omega}+\|\nabla w\|_{\Omega}. Then
We start by proving that the stabilization/regularization converges optimally for smooth solutions in the norm ||(u_(h)-pi_(h)u,z_(h))||_(s)\left\|\left(u_{h}-\pi_{h} u, z_{h}\right)\right\|_{s}. Note that this is independent of the stability of the continuous problem.
Lemma 2. Let u inH^(p+1)(Omega)u \in H^{p+1}(\Omega) be the solution to (1) and let (u_(h),z_(h))inV_(h)^(p)xxW_(h)^(p)\left(u_{h}, z_{h}\right) \in V_{h}^{p} \times W_{h}^{p} be the finite element solution to (10). Then
||(u-u_(h),z_(h))||_(s) <= ||(u-pi_(h)u,0)||_(s)+||(u_(h)-pi_(h)u,z_(h))||_(s)\left\|\left(u-u_{h}, z_{h}\right)\right\|_{s} \leq\left\|\left(u-\pi_{h} u, 0\right)\right\|_{s}+\left\|\left(u_{h}-\pi_{h} u, z_{h}\right)\right\|_{s}
The first term is bounded from the approximation estimate (16) and we only need to focus on the second term. Let us start by expanding
{:[(17)||(u_(h)-pi_(h)u,z_(h))||_(s)^(2)=(u_(h),u_(h)-pi_(h)u)_(omega)+s(u_(h),u_(h)-pi_(h)u)+s^(**)(z_(h),z_(h))],[-(pi_(h)u,u_(h)-pi_(h)u)_(omega)-s(pi_(h)u,u_(h)-pi_(h)u)]:}\begin{align*}
\left\|\left(u_{h}-\pi_{h} u, z_{h}\right)\right\|_{s}^{2}= & \left(u_{h}, u_{h}-\pi_{h} u\right)_{\omega}+s\left(u_{h}, u_{h}-\pi_{h} u\right)+s^{*}\left(z_{h}, z_{h}\right) \tag{17}\\
& -\left(\pi_{h} u, u_{h}-\pi_{h} u\right)_{\omega}-s\left(\pi_{h} u, u_{h}-\pi_{h} u\right)
\end{align*}
For the first row of this expansion we use (10)_(1)(10)_{1} and we have that
Returning to the stabilization norm (17), we obtain that
{:(18)||(u_(h)-pi_(h)u,z_(h))||_(s)^(2)=(( tilde(g))-pi_(h)u,u_(h)-pi_(h)u)_(omega)+a(pi_(h)u-u,z_(h))+s(u-pi_(h)u,u_(h)-pi_(h)u):}\begin{equation*}
\left\|\left(u_{h}-\pi_{h} u, z_{h}\right)\right\|_{s}^{2}=\left(\tilde{g}-\pi_{h} u, u_{h}-\pi_{h} u\right)_{\omega}+a\left(\pi_{h} u-u, z_{h}\right)+s\left(u-\pi_{h} u, u_{h}-\pi_{h} u\right) \tag{18}
\end{equation*}
We now discuss the convergence of these three terms. First, by Cauchy-Schwarz and standard approximation (7) we bound
s(u-pi_(h)u,u_(h)-pi_(h)u) <= ||u-pi_(h)u||_(V_(h)^(p))||u_(h)-pi_(h)u||_(V_(h)^(p)) <= Ch^(p)||u||_(H^(p+1)(Omega))||u_(h)-pi_(h)u||_(V_(h)^(p)),s\left(u-\pi_{h} u, u_{h}-\pi_{h} u\right) \leq\left\|u-\pi_{h} u\right\|_{V_{h}^{p}}\left\|u_{h}-\pi_{h} u\right\|_{V_{h}^{p}} \leq C h^{p}\|u\|_{H^{p+1}(\Omega)}\left\|u_{h}-\pi_{h} u\right\|_{V_{h}^{p}},
where we used the approximation bound (16). Collecting the bounds for the terms in (18), we conclude that
||(u_(h)-pi_(h)u,z_(h))||_(s)^(2) <= C(h^(p)||u||_(H^(p+1)(Omega))+||delta g||_(omega))||(u_(h)-pi_(h)u,z_(h))||_(s)\left\|\left(u_{h}-\pi_{h} u, z_{h}\right)\right\|_{s}^{2} \leq C\left(h^{p}\|u\|_{H^{p+1}(\Omega)}+\|\delta g\|_{\omega}\right)\left\|\left(u_{h}-\pi_{h} u, z_{h}\right)\right\|_{s}
and hence
||(u_(h)-pi_(h)u,z_(h))||_(s) <= C(h^(p)||u||_(H^(p+1)(Omega))+||delta g||_(omega)).\left\|\left(u_{h}-\pi_{h} u, z_{h}\right)\right\|_{s} \leq C\left(h^{p}\|u\|_{H^{p+1}(\Omega)}+\|\delta g\|_{\omega}\right) .◻\square
We continue by proving that the dual norm of the residual can be controlled by the stabilization.
Lemma 3. Let u inH^(p+1)(Omega)u \in H^{p+1}(\Omega) be the solution to (1) and let (u_(h),z_(h))inV_(h)^(p)xxW_(h)^(p)\left(u_{h}, z_{h}\right) \in V_{h}^{p} \times W_{h}^{p} be the finite element solution to (10). Let r inH^(-1)(Omega)r \in H^{-1}(\Omega) be the residual (:r,w:)=a(u-u_(h),w),w inH_(0)^(1)(Omega)\langle r, w\rangle=a\left(u-u_{h}, w\right), w \in H_{0}^{1}(\Omega). Then
We will now prove an error estimate on the solution in a target domain BB, using the conditional stability from Section 2. We will write a general version, with implicit dependence on the frequency-this can be made explicit for particular geometric settings satisfying a convexity condition.
Theorem 1. Let u inH^(p+1)(Omega)u \in H^{p+1}(\Omega) be the solution to (1) and let (u_(h),z_(h))inV_(h)^(p)xxW_(h)^(p)\left(u_{h}, z_{h}\right) \in V_{h}^{p} \times W_{h}^{p} be the finite element solution to (10). Then
Proof. We consider the error u-u_(h)u-u_{h} together with the residual (:r,w:)=a(u-u_(h),w)\langle r, w\rangle=a\left(u-u_{h}, w\right), for w inH_(0)^(1)(Omega)w \in H_{0}^{1}(\Omega), and apply the conditional stability estimate (4)
Note that ||u-u_(h)||_(L^(2)(omega))\left\|u-u_{h}\right\|_{L^{2}(\omega)} is bounded from Lemma 2 and the residual ||r||_(H^(-1)(Omega))\|r\|_{H^{-1}(\Omega)} is bounded from Lemma 3. Bounding
Note that in Theorem 1 the convergence rate increases linearly with the polynomial degree pp of the approximation and in the case of well-posed problems (with alpha=1\alpha=1 ) we recover the optimal rate pp. In the presence of noise, the error bound can degenerate and perturbations in data get amplified as pp increases, meaning that in practical settings the advantages of high-order methods might be reduced. Also, one should consider a certain threshold after which refining the mesh becomes detrimental.
4 Numerical experiments
We present numerical experiments for the Helmholtz unique continuation problem (1) solved with the high-order conforming method (10). The test case we will focus on has been considered in [Nec20] and [BDE21], and represents the Helmholtz version of the classical Hadamard example for ill-posed elliptic equations. Let n inNn \in \mathbb{N} and consider the Cauchy problem
{[(19)Delta u+k^(2)u=0" in "Omega:=(0","pi)xx(0","1)","],[u(x","0)=0" for "x in[0","pi]","],[u_(y)(x","0)=sin(nx)" for "x in[0","pi]","]:}\left\{\begin{align*}
\Delta u+k^{2} u & =0 & & \text { in } \Omega:=(0, \pi) \times(0,1), \tag{19}\\
u(x, 0) & =0 & & \text { for } x \in[0, \pi], \\
u_{y}(x, 0) & =\sin (n x) & & \text { for } x \in[0, \pi],
\end{align*}\right.
For such Hadamard-type solutions, we will consider the interior datum g=u|_(omega)g=\left.u\right|_{\omega} and study the convergence in the target set BB for two geometric configurations of omega\omega and BB, as discussed in Section 2, namely
sketched in Figure 2a. To assess the effect of increasing the frequency, we will take n=5,k=1n=5, k=1 and n=11,k=10n=11, k=10, having similar values for sqrt(k^(2)-n^(2))\sqrt{k^{2}-n^{2}} in (20). Noisy data is considered by polluting the nodal values of the given restriction gg with uniformly distributed values in [-h,h][-h, h] or [ -h^(2),h^(2)-h^{2}, h^{2} ], for a noise amplitude O(h)\mathcal{O}(h) or O(h^(2))\mathcal{O}\left(h^{2}\right), respectively. The results shown below are obtained with the stabilization parameter gamma=10^(-3)\gamma=10^{-3} in the primal stabilizer (13).
For the domains in (21), data is given around three sides of the domain Omega\Omega and the stability of the continuation problem is expected to be close to Lipschitz. Indeed, for approximations of order p in{1,2,3}p \in\{1,2,3\} we observe in Figure 3 clear convergence with rate pp, indicating that the exponent alpha~~1\alpha \approx 1 in the conditional stability estimate (3). The convergence rates are in agreement with the theoretical error estimates. Also, in this case the target set BB is inside the convex hull of the data set omega\omega, and we notice that increasing the frequency has only a small effect on the size of the errors, showing the robustness of the stability constants.
Figure 3: Convergence rates for domains in (21), good stability.
Figure 4: Noisy data, convergence rates for domains in (21), n=5,k=1n=5, k=1.
In Figure 4 we consider noisy data for the domains in (21) with stability exponent alpha~~1\alpha \approx 1. We observe that for a noise amplitude of O(h^(2))\mathcal{O}\left(h^{2}\right) the perturbations have no impact for p in{1,2}p \in\{1,2\}, while for p=3p=3 one order of convergence is lost, as predicted by the theory. Increasing the data perturbations to O(h)\mathcal{O}(h) makes no difference for p=1p=1, while the convergence decreases by a factor of hh for p=2p=2, and by a factor of h^(2)h^{2} for p=3p=3. This confirms the theoretical results and shows that high-order methods are more sensitive to the size of the data perturbations.
Figure 5: Convergence rates for domains in (22), poor stability.
Figure 5 shows the numerical results for the domains (22), for which the stability of the continuation problem-quantified by the exponent alpha≪1-\alpha \ll 1- is expected to be poor. For k=1k=1, the convergence rate for p=1p=1 is close to 0.25 and increasing the frequency to k=10k=10 has a strong effect on the numerics: the convergence rate for p=1p=1 is now close to 0.1 and the size of the errors increase. This is due to the fact that continuation is done outside of the convex hull of the data region, and the stability constants increase very fast with the frequency. Nevertheless, in both plots we observe that the rates for quadratic and cubic approximations increase linearly with the polynomial order pp, as expected from Theorem 1.
Figure 6: Noisy data, convergence rates for domains in (22), convexity condition not satisfied.
Perturbations in data for the domains (22) are considered in Figure 6. We see that for a noise amplitude of O(h^(2))\mathcal{O}\left(h^{2}\right) the perturbations have no impact, while for an amplitude of O(h)\mathcal{O}(h) the method does not seem to converge for p=3p=3. The noise effect for p=2p=2 does not seem apparent for such coarse meshes, probably due to the large size of the stability constants. When using high-order methods for problems with poor stability and large perturbations in data, it is clear that one should consider stopping the mesh refinement before the accuracy of the approximation starts to deteriorate.
References
[BDE21] E. Burman, G. Delay, and A. Ern. A hybridized high-order method for unique continuation subject to the helmholtz equation. SIAM J. Numer. Anal., 59(5):2368-2392, 2021.
[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.
[BM21] S. M. Berge and E. Malinnikova. On the three ball theorem for solutions of the Helmholtz equation. Complex Anal. Synerg., 7(2):1-10, 2021.
[BNO19] E. Burman, M. Nechita, and L. Oksanen. Unique continuation for the Helmholtz equation using stabilized finite element methods. J. Math. Pures Appl., 129:1-22, 2019.
[BS97] I. M. Babuška and S. A. Sauter. Is the pollution effect of the FEM avoidable for the Helmholtz equation considering high wave numbers? SIAM J. Numer. Anal., 34(6):2392-2423, 1997.
[Bur13] E. Burman. Stabilized finite element methods for nonsymmetric, noncoercive, and ill-posed problems. Part I: Elliptic equations. SIAM J. Sci. Comput., 35(6):A2752A2780, 2013.
[Bur16] E. Burman. Stabilised finite element methods for ill-posed problems with conditional stability. In Building Bridges: Connections and Challenges in Modern Approaches
to Numerical Partial Differential Equations, volume 114 of Lect. Notes Comput. Sci. Eng., pages 93-127. Springer, 2016.
[EG21] A. Ern and J.-L. Guermond. Finite elements I: Approximation and interpolation, volume 72 of Texts in Applied Mathematics. Springer Nature, 2021.
[HI04] T. Hrycak and V. Isakov. Increased stability in the continuation of solutions to the Helmholtz equation. Inverse Problems, 20(3):697-712, 2004.
[Isa17] V. Isakov. Inverse problems for partial differential equations, volume 127 of Applied Mathematical Sciences. Springer, 3rd edition, 2017.
[MS11] J. M. Melenk and S. A. Sauter. Wavenumber explicit convergence analysis for Galerkin discretizations of the Helmholtz equation. SIAM J. Numer. Anal., 49(3):1210-1243, 2011.
[Nec20] M. Nechita. Unique continuation problems and stabilised finite element methods. PhD thesis, University College London, 2020.