Abstract
We derive a posteriori estimates for single-step methods, including Runge–Kutta and Galerkin methods for the time discretization of linear parabolic equations. We focus on the estimation of the error at the nodes and derive a posteriori estimates that show the full classical order (superconvergence order) in the interior of the spatial domain without any compatibility assumptions.
Similar content being viewed by others
Explore related subjects
Discover the latest articles and news from researchers in related subjects, suggested using machine learning.Avoid common mistakes on your manuscript.
1 Introduction
This paper studies a posteriori error estimation of the of one-step time discretization methods for linear parabolic differential equations. The objective is to give rigorous error bounds in terms of quantities derived from the computed solution, which should show the correct asymptotic order of accuracy as known from an a priori error analysis in terms of regularity bounds of the exact solution or the data. The expected order of accuracy at the time nodes is higher than the order expected in other points. In the case of Runge–Kutta methods, the maximal order at the nodal points is the classical order of the method. Such a higher order at nodal points is indeed attained provided compatibility conditions are satisfied, which are, however, unrealistic with all kinds of (Dirichlet or Neumann or Robin) boundary conditions except periodic boundary conditions. If the required compatibility conditions are not fully satisfied, an order reduction with respect to the classical order is observed [9, 13]. Remarkably, the order reduction remains localized near the boundary and the full classical order is attained in the interior of the domain [10].
In the present paper our goal is twofold: we give a new proof of a posteriori error bounds at the time nodes and we show that the order reduction does not occur in the a posteriori control of the error in the interior of the spatial domain.
We use the unified treatment of essentially all single-step time-stepping schemes of [3] and of the corresponding reconstructions. A key novel feature of our analysis is an error representation formula based on Duhamel’s principle. Through this expression a direct superconvergence analysis for Runge–Kutta and Galerkin time discretization schemes is possible. Our interior results are a posteriori analogs of the a priori estimates of [10].
For previous a posteriori results using various one step time discretization methods we refer, e.g., to [1–3, 5–8, 11, 14]. A posteriori time-superconvergence results for fully discrete schemes based on dG piecewise linear time discretization methods were derived in [5].
We consider linear parabolic equations in a Hilbert space setting: Let \(H\) be a Hilbert space with inner product \(\langle \cdot ,\cdot \rangle \) and corresponding norm \(|\cdot |\). The problem is to find \(u:[0,T] \rightarrow D(A)\) satisfying
with \(A\) a positive definite, self-adjoint, linear operator on \(H\) with domain \(D(A)\) dense in \(H\), and a given forcing term \(f : [0,T]\rightarrow H.\)
We remark that the results of Sect. 2 hold equally in a Banach space setting, whereas those of Sect. 3 require the above Hilbert space setting.
1.1 Discretization methods
We will use the notation and formalism of [3] to describe the numerical methods considered. We consider piecewise polynomial functions in arbitrary partitions \(0=t_0<t_1<\cdots <t_N=T\) of \([0,T],\) and let \(J_n:=(t_{n-1},t_n]\) and \(k_n:=t_n-t_{n-1}\). We denote by \(\mathcal{V }_q^\mathrm{d}, q\in \mathbb{N }_0,\) the space of functions that are piecewise polynomials of degree at most \(q\) in time in each subinterval \(J_n\) with coefficients in \(V=D(A^{1/2})\), without continuity requirements at the nodes \(t_n\). The elements of \(\mathcal{V }_q^\mathrm{d}\) are taken continuous to the left at the nodes \(t_n;\mathcal{V }_q(J_n)\) consist of the restrictions to \(J_n\) of the elements of \(\mathcal{V }_q^\mathrm{d}\). The spaces \({\mathcal{H }}_q^\mathrm{d}\) and \({\mathcal{H }}_q(J_n)\) are defined similarly by requiring that the coefficients are in \(H\). Let \(\mathcal{V }_q^\mathrm{c}\) and \(\mathcal{H }_q^\mathrm{c}\) be the spaces of the continuous elements of \(\mathcal{V }_q^\mathrm{d}\) and \(\mathcal{H }_q^\mathrm{d}\). For \(v\in \mathcal{V }_q^\mathrm{d}\) we let \(v^n:=v(t_n),\;v^{n+}:= \lim _{t\downarrow t_n} v(t)\).
To define the time stepping methods we introduce the operator \({\varPi }_{q-1}\) to be a projection operator to piecewise polynomials of degree \(q-1,\varPi _{q-1}: C([0,T];H) \rightarrow \oplus _{n=1}^N \mathcal{H }_{q-1}(J_n).\) Also, \(\widetilde{\varPi }_{q}:\mathcal{H }_q(J_n) \rightarrow \mathcal{H }_\ell (J_n)\) is an operator mapping polynomials of degree \(q\) to polynomials of degree \(\ell ,\) with \(\ell =q\) or \(\ell =q-1;\varPi _{q-1}\) and \(\widetilde{\varPi }_{q}\) are defined in a reference time interval and then transformed into \(J_{n}.\)
The time discrete approximation \(U\) to the solution \(u\) of (1.1) is defined as follows: We seek \(U\in \mathcal{V }_q^\mathrm{c}\) satisfying the initial condition \(U(0)=u^0\) as well as
where \(F(t,v)= Av - f(t).\) An equivalent Galerkin formulation is
for \(n=1,\ldots , N,\) see [3]. The above formalism covers a large class of one-step time discretization schemes. In particular, the continuous Galerkin (cG) method is
with \(\varPi _{q-1}:=P_{q-1}, \) with \(P_\ell \) denoting the \(L^2\) orthogonal projection operator onto \(\mathcal{H }_\ell (J_n).\) Furthermore, in [3] it was shown that (1.3) describes other important implicit single-step time stepping methods: the RK collocation methods (RK-C) with \(\varPi _{q-1}:=I_{q-1},\) the interpolation operator at the collocation points, and \(\widetilde{\varPi }_{q}\) the identity operator; all other interpolatory RK methods with \(\varPi _{q-1}:=I_{q-1},\) and appropriate \( \widetilde{\varPi }_{q}\) (with \(\ell =q\)); the discontinuous Galerkin (dG) method with \(\varPi _{q-1}:=P_{q-1}\) and \( \widetilde{\varPi }_{q}= I_{q-1},\) where \(I_{q-1}\) is the interpolation operator at the Radau points \(0<c_1<\cdots <c_q=1\) (so \(\ell =q-1\)).
1.2 Superconvergence—classical order
A key assumption for the time-discretization methods related to the accuracy at the time nodes is: We assume that the method (1.2) is associated to \(q\) pairwise distinct points \(c_1,\ldots , c_q\in [0, 1]\) with the property
This condition induces orthogonality conditions at each interval \(J_{n}\) with \(t_{n, i}:=t_{n-1}+c_i k_n, \) \( i=1,\ldots , q.\) These points will be associated to projections (or interpolants) used to define the method (1.2); see [3] for details. The superconvergence order or classical order \(p\) of the method at the nodes is denoted
which is equal to the order of the interpolatory quadrature with nodes \(c_i\).
2 Nodal error analysis in \(H\)
2.1 Reconstruction and main error equation
As in [3] we compare the solution \(u\) with the reconstruction \(\hat{U}\) of \(U\) defined through
where the projection operators \(\hat{\varPi }_q\) onto \(\mathcal{H }_q(J_n), n=1,\ldots , N,\) are chosen to that \(\hat{\varPi }_q w\) agree with \(\varPi _{q-1}\) at \(t_{n,i}\):
In view of (1.5) for \(v(\tau )=1\) and (2.2), we obtain \(\hat{U}(t_n)=U(t_n)\) and conclude that \(\hat{U}\) is continuous. Furthermore, \(\hat{U}\) satisfies
which has a similar structure to (1.2). The motivation for introducing \(\hat{U} \) goes back to [1, 2] and details for its various properties are discussed in [3]. The operator \(\hat{\varPi }_q\) is chosen as follows: for the cG and dG methods, \(\hat{\varPi }_q=P_q\); for the RK collocation methods, \(\hat{\varPi }_q\) is an interpolation operator at the \(q\) collocation points of \(J_n\) plus another point.
We mention two key properties of \(\hat{U}\). The first one is the orthogonality property which follows by (2.2):
for \(n=1,\ldots , N.\) The second one is a further assumption on \(\varPi _{q-1},\) namely for all \(V\in \mathcal{H }_q(J_n)\),
which, in view of (2.4), yields
Condition (2.5) is verified by both cG and dG methods, for which \(\varPi _{q-1}=P_{q-1}\), as well as by RK methods, for which \(\varPi _{q-1}=I_{q-1}.\) 1.1.
We state now the main error equation, which is the starting point of our analysis. Let \(\hat{R}\) be the residual of \(\hat{U},\)
Subtracting (2.7) from the differential equation in (1.1), we obtain the equation
for the error \(\hat{e}:=u-\hat{U},\) which we rewrite in the form
with
and
Notice that \(R_{\hat{\varPi }_q}\) vanishes when \(\hat{\varPi }_q\) is a projector over \(\mathcal{H }_q(J_n)\) whereas \(R_{\widetilde{\varPi }_{q}}\) vanishes when \( \widetilde{\varPi }_{q} =I.\)
We will estimate the error \(e:=u-U\) at the time grid points, where \(e(t_n)=\widehat{e}(t_n),\) which because of the error equation (2.8) can be bounded in terms of the residuals \(\widehat{R}(t)\) for \(0\le t \le t_n.\)
Remark 2.1
(Order of the residual) Theorem 2.2 in [2] states that
where \(\varphi _{q+1}\) is the polynomial of degree \(q+1\) with leading coefficient 1 with \(\varphi _{q+1}(0)=0\) and vanishing derivatives at \(c_1,\ldots ,c_q\). In view of using this formula and standard interpolation estimates in (2.9)–(2.11), we anticipate that in the case of a smooth solution the residual \(\hat{R}\) is of order \(O(k_n^{q+1})\) for the continuous Galerkin method cG(\(q\)), for the discontinuous Galerkin method dG(\(q\)), as well as for RK collocation methods with \(q\) arbitrary collocation points.
Some care must however be taken in which norm \(\widehat{R}\) is estimated. In Sect. 6 of [3] it is shown, for \(\rho \ge 1\), that \(U,\widehat{U}\in D(A^\rho )\) and \(\hat{R}\in D(A^{\rho -1})\) when \(U^0\in D(A^\rho )\) and \(f\in D(A^\rho )\), for the cases of the continuous and discontinuous Galerkin methods and for the RK collocation methods at, e.g., Radau points. However, if one does not assume \(f\in D(A)\) (which for the Laplacian with Dirichlet boundary conditions would require the compatibility condition that \(f\) vanishes at the boundary), then only \(A^{-1}\widehat{R}\) will be in \(H\), but not \(\widehat{R}\) itself. In the case of a spatially and temporally smooth solution we can then expect
which holds if \(|\widehat{U}^{(q+1)}| =O(1)\). This bound can indeed be shown rigorously if the error satisfies
which is obtained from standard a priori error bounds if the step sizes \(k_n\) are all of equal magnitude. For a finite element space discretization, the compatibility problem still shows up in that the \(H\)-norm of the spatially discrete residual is not \(O(k_n^{q+1})\) uniformly in the spatial grid size when \(f\notin D(A)\).
The difficulty of \(\hat{R}(t)\notin H\) does not occur for the heat equation on a cube with periodic boundary conditions or for the heat equation on the whole \(R^d\), where in the case of smooth data one obtains, for arbitrary \(\ell \ge 0\),
2.2 Error representation via Duhamel’s principle.
We now apply Duhamel’s principle to (2.8):
where \(E_A (t) = e^ {-At}\) is the solution operator of the homogeneous equation
i.e., \(v(t) = E_A (t) w.\) The family of operators \(E_A (t)\) is the semigroup of contractions on \(H\) with generator \(-A.\) The following properties are well known, cf., e.g., Crouzeix [4], Thomée [14],
and
Since \(A\) and \( E_A\) commute, (2.17) implies
when \(|E_A w| \le C_A t^{-m} |A^{-m} w|\).
Starting from (2.14) we derive now a different error representation formula involving time derivatives of \(E_A.\) In the interval \(t_{n-1}\le s \le t_n\) we define the scaled \(j\)th antiderivative of \(\hat{R} \) as
Then, one has,
Using (2.16), (2.20) and integrating by parts in (2.14) we obtain,
Further,
Thus, for any \(\rho , \)
Notice that, still for \(t \ge t_{n-1},\) and for \(s\in J_m\), \(E_A(t-s)=E_A(t-t_m)E_A(t_m-s),\) thus
Treating the last integral as (2.23) we have proved the following proposition.
Proposition 2.1
Let \(t\in J_n, \) then with \(\hat{R}^{[j]} _n\) defined by (2.19), the following error representation formula holds:
The error representation formula (2.25) will be the starting point of our analysis. We mainly consider \(t=t_n\), which leads to a posteriori error control at the time nodes. We will treat Galerkin schemes and Runge-Kutta methods separately. We use (2.18) in the above error representation formula to obtain
The terms in the above relation are treated differently. The expression (2.20) yields
and
Let
then we have
where \(\Vert A^{\rho -1} \hat{R}\Vert _{L^\infty (J_m,H)} = \sup _{s \in J_{m}} \big | A ^{\rho -1} \hat{R}( s ) \big | \). We note that at the expense of the logarithmic factor \(L_n\) we here reduced the regularity requirement of \(\hat{R}\) from \(D(A^\rho )\) to \(D(A^{\rho -1})\) as compared with a more straightforward estimation in (2.25). We are ready now to derive the main estimates of this section.
2.3 Nodal estimates for Galerkin schemes
In the case of Galerkin schemes (continuous or discontinuous) the error estimates are direct consequences of (2.27). The main point here is that the terms involving \( \hat{R}^{[j]}_m(t_m )\) all vanish due to the orthogonality. In this case we have the following result.
Theorem 2.1
Let \(q\ge 2\) and \(\hat{R} \in D(A^{\rho -1})\) hold for some \(1 \le \rho \le q-1.\) Then, the error of the continuous Galerkin method of degree \(q\) and of the discontinuous Galerkin method dG\((q-1)\) satisfies
where \(C_A\) is the stability constant in (2.17) and \(L_n\) is the logarithmic factor given in (2.26).
Proof
Recall that both methods are written in the form
where \({\widetilde{\varPi }_{q}}= I\) for the cG method, and for the dG method \( \widetilde{\varPi }_{q}= I_{q-1},\) where \(I_{q-1}\) is the interpolation operator at the Radau points. Notice that in both cases \(\varPi _{q-1}=P_{q-1}.\) In view of (2.3),
with \(\hat{\varPi }_q = P_q.\) Therefore, in the case of the cG method \(\hat{R}(t) = R_{\hat{U}}(t)+R_{\widetilde{\varPi }_{q}}(t) + R_{\hat{\varPi }_q}(t) + R_f(t) \) where
Hence by (2.4) we have
In view of the definition of \(\hat{R}^{[j]}_m\) of (2.20), we obtain
so that the terms involving \(\hat{R}^{[j]}_m(t_m)\) in (2.27) vanish and (2.35) follows.
In the case of the dG method, the properties of Gauss–Radau quadrature imply
Thus, given that in the expression for \(\hat{R}\) the difference to the cG case is that \(R_{\widetilde{\varPi }_{q}} + R_{\hat{\varPi }_q} = A(I_{q-1} U - U)\), (2.32) still holds in this case as well. The proof is thus complete. \(\square \)
Remark 2.2
(Order of convergence) One notices that the highest possible order of the residual \(\hat{R}\) for dG is \(q\) in (2.35), whereas it is \(q+1\) for cG. Hence the highest order in (2.35) is \(2q\) in for cG and \(2q-1\) for dG, as expected. The difference is due to the fact that in the dG case the residual \(\hat{R}\) contains an additional term of the form \(R_{\widetilde{\varPi }_{q}}= A(U-I_{q-1}U)\).
Note, however, that the full order is attained only if \(\hat{R} \in D(A^{q-2})\). This is usually not satisfied, since it requires unnatural compatibility conditions at the boundary when \(A\) is an elliptic operator with Dirichlet or Neumann boundary conditions and \(H=L_2(\Omega )\). In particular, in the case of Dirichlet conditions this would require that \(\widehat{U}\) vanishes on the boundary, which is not true in general when \(f\) does not vanish on the boundary. In this case one only has \(A^{-1}\widehat{R}\in H=L_2(\Omega )\).
We now turn to an a posteriori estimate (of order \(q+1\)) in terms of \(|A^{-1}\hat{R}(t)|\), which formally corresponds to the case \(\rho =0\) in the previous theorem.
Theorem 2.3
Let \(A^{-1}\hat{R} \in H\). Then, the error of the continuous Galerkin method of degree \(q\) and of the discontinuous Galerkin method dG\((q-1)\) satisfies
where \(C_A\) is the stability constant in (2.17) and \(L_n\) is the logarithmic factor given in (2.26).
Proof
At \(t=t_n\) we split the integral in the error formula (2.14) into the integrals from \(0\) to \(t_{n-1}\) and from \(t_{n-1}\) to \(t_n\) and use partial integration in the latter integral:
The result then follows on using (2.17).
2.4 Nodal estimates for collocation methods
In this section we establish a posteriori estimates for the nodal error for RK collocation methods. We recall that the classical order \(p\) of the RK-C method satisfies \(q+1\le p \le 2q, \) i.e., \(1\le \rho \le r = p-q-1.\) The main difference to the case of Galerkin schemes is that the terms involving \(\hat{R}^{[j]}_m (t_{m} ) \) give rise to non-zero expressions involving the inhomogeneity \(f.\) In this case we choose \(\hat{\varPi } _q = \widehat{I}_q ,\) [2, 3]. \(\widehat{I}_q \) is an extended interpolation operator defined on continuous functions \(v\) with the following two key properties
\(\widehat{I}_q\) interpolates \(v\) at one more point, either inside \(J_n\) either outside given that \(v\) is defined at an extended interval. This issue was discussed in detail in [3].
Now, as before we start from \(\hat{R}(t) = R_{\hat{U}}(t)+R_{\widetilde{\varPi }_{q}}(t) + R_{\hat{\varPi }_q}(t) + R_f(t) \) where
Therefore by the assumptions on \(\widehat{I}_q\) we have
Concerning the remaining term \(R_f\), we introduce the notation
This is just the quadrature error of the function \((t_n-\tau )^{j-1}/(j-1)! \cdot f(\tau )\) over the interval \(J_n\),
which is of optimal order \(O(k_n^{p+1})\) if \(f\) is \(p\)-times continuously differentiable. Then, in view of the definition of \(\hat{R}^{[j]}_n\) of (2.20) and due to (2.38), we have
With (2.27) we therefore obtain the following result. A similar result holds for perturbed collocation methods, [12], compare to [3].
Theorem 2.3
Let the classical order \(p\) of a \(q-\)stage Runge-Kutta collocation method satisfy \(p\ge q+2\) and let \(\hat{R},f\in D(A^{\rho -1})\) for \(1\le \rho \le r=p-q-1.\) Then the following a posteriori error estimate is valid at the nodes \(t_n\):
The full classical order \(p\) is attained when \(\hat{R},f\in D(A^{r-1})\), which for \(r>1\) again imposes unnatural compatibility conditions. Theorem 2.2 holds also for the collocation method, with the same proof.
Remark 2.3
The estimate in [3] for RK collocation methods is similar. The first term on the right hand side is the same but the term involving the quadrature errors \(E^{[j]}_{f,m}\) differs to the one of [3] which is
The auxiliary interpolator operators \(\widehat{I}_\ell \) are defined as follows: Let \(\hat{t}_{m,j} \in J_m\), with \(j=1,\ldots ,\rho \), be pairwise distinct and different from \(t_{m,i}\), with \(i=0,\ldots ,q\). The operator \(\widehat{I}_\ell \) is an interpolation operator of order \(\ell \) with \(\ell =q+1,\ldots ,q+\rho ,\) defined on continuous functions \(v\) on \([0,T]\) and values on \(\mathcal{H }_\ell (J_m)\):
Here, in contrast to [3] we have chosen not to include the non-homogeneous term in the argument involving the strong stability of \( E_A.\) For that reason our bound has one higher power of \(A.\) In both cases the required regularity of \(\hat{R}\) remains the same. Nevertheless, the second term in Theorem 2.3 can be controlled by the terms appearing in (2.40). To see why, notice that our assumptions imply
Then, with \(\widehat{I}_\ell \) as above we have,
The last integral is zero and therefore,
3 Interior a posteriori error bounds
We prove the following main result, which yields full-order a posteriori error bounds in the interior of the domain without requiring any compatibility conditions on the boundary. By \(H^{m}(S)\) we denote the standard Sobolev space of order \(m\) defined on a domain \(S.\)
Theorem 3.1
Let \(A\) be the negative Laplacian on a bounded Lipschitz domain \(\Omega \subset \mathbb{R }^d\), equipped with Dirichlet boundary conditions. Let \(\omega \subset \widehat{\omega }\subset \Omega \) be Lipschitz subdomains such that the boundaries of the three domains have pairwise distances of at least \(\delta >0\).
Let \(q\ge 2\) and \(\hat{R}|_{\widehat{\omega }} \in H^{2\rho }(\widehat{\omega })\) for some \(1 \le \rho \le r=p-q-1\), where \(p\) is the classical order of the method. Then, the error of the continuous Galerkin method of degree \(q\) and of the discontinuous Galerkin method dG\((q-1)\) satisfies
where \(C_1\) depends only on \(\Omega \) and \(\delta \).
Theorem 3.2
In the situation of Theorem 3.1, the error of a \(q\)-stage Runge–Kutta collocation method satisfies
where \(E_{f,m}^{[j]}\) is the quadrature error defined in (2.39) and \(C_1,C_2\) depend only on \(\Omega \) and \(\delta \).
The interior nodal error bounds are of optimal order \(p\) when \(\widehat{R}\) is sufficiently regular in a neighbourhood of the subdomain \(\widehat{\omega }\), which is the case if the solution is sufficiently regular on \(\widehat{\omega }.\) In this case we expect that a higher regularity version of (2.12) yields
It is important to note that the regularity away from \(\widehat{\omega }\) and the boundary behaviour on \(\partial \Omega \) play no role.
As the proof below shows, the dependence of \(C_1,C_2\) on the domain \(\Omega \) is only through the constants in Poincaré–Friedrichs inequalities. The dependence on \(\delta \) is obtained such that \(C_1,C_2=O(\delta ^{-4(\rho +2)})\).
The result could straightforwardly be generalized to any second-order elliptic differential operator with smooth coefficients and appropriate essential boundary conditions.
For the proof we consider a finite chain of domains
where \(\ell = 2\rho +4\) and the distance from \(\omega _j\) to the boundary of \(\omega _{j+1}\) is for all \(j\) bounded from below by a constant times \(\delta \). To these regions we associate smooth cutting functions \(\chi _j\) on \(\Omega \) such that
for \(j=0,1,\ldots ,\ell -1\), and \(\chi _{\ell }\equiv 1\) on \(\Omega \). Viewed as multiplication operators, these functions have the following property with respect to the norm \(|\cdot |\) of \(H=L_2(\Omega )\):
For \(A=-\Delta \), this bound is a consequence of the fact that the commutator \(A\chi _j - \chi _j A\) is a first-order differential operator: \( -\Delta (\chi _j\varphi ) + \chi _j\Delta \varphi = -(\Delta \chi _j)\varphi -2 \nabla \chi _j\cdot \nabla \varphi , \) and from this we also note that \(\beta \) is proportional to \(\delta ^{-2}\).
Lemma 3.1
If operators \(\chi _0,\ldots ,\chi _\ell \) satisfy (3.3) and \(\chi _\ell =\mathrm{id}\), then
Proof
We denote \(w(t)=E_A(t)v\) and \(B_j=\chi _j A - A \chi _j\). Since \(w(t)\) satisfies \(w^{\prime }+Aw=0\), \(w(0)=v\), we have
The standard parabolic energy estimate yields
and hence, by (3.3),
Since \(\chi _1w(t)\) solves \(\chi _1 w^{\prime } + A \chi _1 w = B_1w,\chi _1w(0)=\chi _1v,\) we obtain by the same argument
Continuing in this way, we have for \(j=1,\ldots , \ell -1\)
Since \(\chi _\ell =\mathrm{id}\), for \(j=\ell -1\) the last integral term equals
Concatenating the above estimates completes the proof. \(\square \)
Proof
(of Theorem 3.1) We work in the Hilbert space \(H=L_2(\Omega )\) with the norm \(|\cdot |=\Vert \cdot \Vert _{L_2(\Omega )}\). We begin by noting
and \(e(t_n)=\widehat{e}(t_n)\). For Galerkin methods we obtain from (2.25) and the Galerkin orthogonality (2.33) that
By Lemma 3.1 with \(\ell =2\rho +2\) we have, for \(w=\hat{R}_m^{[\rho ]}(s)\),
We now show that we can estimate
For this we use a duality argument:
Since the norm \(| A^{j/2} \cdot |\) is equivalent to the \(H^{j}(\Omega )\) Sobolev norm on \(C^\infty _0(\Omega )\), we have
Hence,
which is the desired estimate. Combining the above estimates, we obtain
which implies the error bound of Theorem 3.1 for the Galerkin methods. For the Runge–Kutta methods, there appear in addition the quadrature errors \(E_{f,m}^{[j]}\) of (2.39), which are treated in the same way.
Remark 3.1
While the results of Sect. 2 hold equally for an elliptic operator as well as for its finite element discretization (with \(A\) denoting the continuous and discrete operator, respectively), Theorem 3.1 is stated only for the spatially continuous, temporally discretized case. An analogous result would also hold for the full—space and time—discretization if the discrete operator \(A_h\) satisfied an estimate (3.3) with the operators given as \(\chi _{j,h}v_h= \Pi _h (\chi _jv_h)\) for \(v_h\in V_h\), where \(\Pi _h\) is a suitable projection to the finite element space \(V_h\). Though it may be expected that such an estimate would hold for time-independent quasi-uniform meshes with a constant \(\beta \) independent of the mesh size \(h\), a detailed analysis of this question is outside the scope of this paper, which is concerned only with the time discretization aspects.
References
Akrivis, G., Makridakis, Ch., Nochetto, R.H.: A posteriori error estimates for the Crank-Nicolson method for parabolic equations. Math. Comp. 75, 511–531 (2006)
Akrivis, G., Makridakis, Ch., Nochetto, R.H.: Optimal order a posteriori error estimates for a class of Runge-Kutta and Galerkin methods. Numer. Math. 114, 133–160 (2009)
Akrivis, G., Makridakis, Ch., Nochetto, R.H.: Galerkin and RungeKutta methods: unified formulation, a posteriori error estimates and nodal superconvergence. Numer. Math. 118, 429–456 (2011)
Crouzeix, M.: Sur l’approximation des équations différentielles opérationelles linéaires par des méthodes de Runge-Kutta. Université de Paris VI, Thèse (1975)
Eriksson, K., Johnson, C.: Adaptive finite element methods for parabolic problems. I. A linear model problem. SIAM J. Numer. Anal. 28, 43–77 (1991)
Eriksson, K., Johnson, C., Larsson, S.: Adaptive finite element methods for parabolic problems VI. Analytic semigroups. SIAM J. Numer. Anal. 35, 1315–1325 (1998)
Estep, D., French, D.: Global error control for the Continuous Galerkin finite element method for ordinary differential equations. RAIRO Modél. Math. Anal. Numér. 28, 815–852 (1994)
Lozinski, A., Picasso, M., Prachittham, V.: An anisotropic error estimator for the Crank-Nicolson method: application to a parabolic problem. SIAM J. Sci. Comp. 31, 2757–2783 (2009)
Lubich, Ch., Ostermann, A.: Runge–Kutta methods for parabolic equations and convolution quadrature. Math. Comp. 60, 105–131 (1993)
Lubich, Ch., Ostermann, A.: Interior estimates for time discretizations of parabolic equations. Appl. Numer. Math. 18, 241–251 (1995)
Makridakis, Ch., Nochetto, R.H.: A posteriori error analysis for higher order dissipative methods for evolution problems. Numer. Math. 104, 489–514 (2006)
Nørsett, S.P., Wanner, G.: Perturbed collocation and Runge–Kutta methods. Numer. Math. 38, 193–208 (1981)
Sanz-Serna, J.M., Verwer, J.G., Hundsdorfer, W.H.: Convergence and order reduction of Runge–Kutta schemes applied to evolutionary problems in partial differential equations. Numer. Math. 50, 405–418 (1987)
Thomée, V.: Galerkin Finite Element Methods for Parabolic Problems, 2nd edn., Springer, Berlin, 2006
Author information
Authors and Affiliations
Corresponding author
Additional information
Ch. Makridakis was partially supported by the FP7-REGPOT project “ACMAC- Archimedes Center for Modeling, Analysis and Computation” of the University of Crete.
Rights and permissions
About this article
Cite this article
Lubich, C., Makridakis, C. Interior a posteriori error estimates for time discrete approximations of parabolic problems. Numer. Math. 124, 541–557 (2013). http://doi.org/10.1007/s00211-013-0520-1
Received:
Revised:
Published:
Issue Date:
DOI: http://doi.org/10.1007/s00211-013-0520-1