Home Tensor-Product Space-Time Goal-Oriented Error Control and Adaptivity With Partition-of-Unity Dual-Weighted Residuals for Nonstationary Flow Problems
Article Publicly Available

Tensor-Product Space-Time Goal-Oriented Error Control and Adaptivity With Partition-of-Unity Dual-Weighted Residuals for Nonstationary Flow Problems

  • Julian Roth ORCID logo EMAIL logo , Jan Philipp Thiele ORCID logo , Uwe Köcher ORCID logo and Thomas Wick ORCID logo
Published/Copyright: May 31, 2023

Abstract

In this work, the dual-weighted residual method is applied to a space-time formulation of nonstationary Stokes and Navier–Stokes flow. Tensor-product space-time finite elements are being used to discretize the variational formulation with discontinuous Galerkin finite elements in time and inf-sup stable Taylor–Hood finite element pairs in space. To estimate the error in a quantity of interest and drive adaptive refinement in time and space, we demonstrate how the dual-weighted residual method for incompressible flow can be extended to a partition-of-unity based error localization. We substantiate our methodology on 2D benchmark problems from computational fluid mechanics.

MSC 2010: 65N30; 65N50; 76D05

1 Introduction

This work is devoted to the efficient numerical solution of the incompressible Navier–Stokes equations employing adaptive Galerkin finite elements in space and time. Adaptivity is based on a posteriori goal-oriented error control using a partition-of-unity dual-weighted residual (PU-DWR) approach. The closest studies to ours are on the one hand [15] as well as the PhD thesis [54] in which similar concepts for full space-time adaptivity for the Navier–Stokes equations were developed. The main differences are two-fold in that we work with tensor-product space-time finite elements and that we utilize a partition-of-unity for localization (originally proposed for stationary problems in [51]) rather than the filtering approach [18]. On the other hand, another closely related study is [58] in which the current PU-DWR method was developed and analyzed for parabolic problems. Using discontinuous Galerkin (dG) test functions in time decouples the formulation and yields a natural localization for which a PU in time is not mandatory. However, we view the PU as a systematic extension to full space-time settings that holds for discontinuous and continuous Galerkin time discretizations. For instance, full continuous Galerkin space-time settings in 4D as in the work of Schafelner [52], error control and adaptivity are naturally achieved with a space-time PU localization. Furthermore, higher-order time derivatives and integration by parts in time in the localization process can be included (as this was one of the main motivations of Richter–Wick [51]).

Employing (full) space-time adaptivity using the DWR method dates back to [55], which developed and tested the method for parabolic problems. Shortly afterward, the space-time DWR method was applied to the incompressible Navier–Stokes equations in [54, 15] and for second-order hyperbolic problems (i.e. the elastic wave equation) in [48, 7]. In this regard, the first space-time finite element formulation (without adaptivity) for incompressible flow was proposed in [45]. Employing space-time (therein called G 2 ) for turbulent flow and goal-oriented adaptivity for the efficient computation of the mean drag was done in [31].

Transport problems with coupled flow with an emphasis on adaptivity of the transport part was investigated in [12], and the extension to a multirate framework was suggested in [21]. Goal-oriented error control with space-time formulations, but only applied to temporal error control for parabolic problems, the Navier–Stokes equations and fluid-structure interaction was undertaken in [42, 43, 26], respectively. Space-time methods with residual-based error control and adaptivity can be found for instance in [38, 52]. Finally, [44] proposed goal-oriented space-time adaptivity with optimal control and [36, 37] developed residual-based adaptivity in optimal control.

The main objectives of this work are to develop space-time PU-DWR for the incompressible Navier–Stokes equations. Therein, we emphasize the following novelties in comparison to existing work. Our approach allows for natural higher-order time discretizations by setting the polynomial degrees of the trial and test basis functions since a general variational discontinuous Galerkin (dG) formulation in time is used and there is no need for the explicit derivation of time-stepping schemes (as for instance done in [55, 15]). Therefore, our framework is more generally extendable to other PDEs since time stepping formulations, e.g. a piecewise linear and discontinuous in time dG ( 1 ) , neither have to be explicitly derived nor implemented for coupled multiphysics problems. To this end, we consider piecewise quadratic time discretizations ( dG ( 2 ) ) for the dual problems without major methodological efforts. Precisely, the defining nodes for the piecewise polynomial basis functions, which are called support points, can be flexibly, arbitrarily and independently chosen within the temporal elements. Such support points might be the quadrature points of the used quadrature for the temporal integration or some other distribution in the temporal element such that the temporal polynomial degree holds. For instance, in this work, we apply (and compare) schemes generated by the Gauss–Lobatto and the Gauss–Legendre formulae. Finally, our implementation is done in the modern finite element library deal.II [6, 5]. These developments are accompanied with a detailed computational analysis in form of error reductions and studies of effectivity indices. To better understand the behavior, the (linear) Stokes equations are considered as well. Therein, Galerkin discretizations using tensor products in time and space are employed. In time, discontinuous Galerkin dG ( r ) elements with polynomial degree r 0 are used for the primal forward problem. The adjoint problem is approximated with globally higher-order finite elements. In space, the classical Taylor–Hood element (i.e. continuous Galerkin cG ( s ) , with s 1 with s = 2 for the velocities and s = 1 for the pressure) is employed for the primal problem and a higher-order approximation for the adjoint. A delicate issue are pressure-robust discretizations on dynamically changing adaptive meshes for which we employ a divergence-free projection from [16], but we make several remarks to more modern methods. These developments yield expressions for error indicators in time and space, which are then utilized to formulate a final adaptive algorithm.

The outline of this paper is as follows. In Section 2, the space-time formulation and discretization are introduced. Next, in Section 3, the space-time PU-DWR is developed and error estimators and error indicators are derived. Afterward, in Section 4, several benchmark tests are adopted to substantiate our algorithmic developments. Our work is summarized in Section 5.

2 Space-Time Formulation and Discretization

We model viscous fluid flow with constant density and temperature by the Navier–Stokes equations [27, 57, 49, 28, 59, 29].

Formulation 1

Formulation 1 (Incompressible Isothermal Navier–Stokes Equations)

Find the vector-valued velocity

v : ( 0 , T ) × Ω R d with d { 2 , 3 }

and the scalar-valued pressure p : ( 0 , T ) × Ω R such that

t v x σ + ( v x ) v = 0 in ( 0 , T ) × Ω , x v = 0 in ( 0 , T ) × Ω , v = v D on ( 0 , T ) × Ω , v ( 0 ) = v 0 in Ω ,

where we use the non-symmetric Cauchy stress tensor for the stress 𝝈 defined as

σ := σ ( v p ) = p I + ν x v .

Here, 𝜈 is the kinematic viscosity, v 0 is the initial velocity and v D is a possibly time-dependent Dirichlet boundary value.

Remark 1

By omitting the nonlinear convection term ( v x ) v , we obtain the Stokes problem.

We employ the following notation for the function spaces. Spatial function spaces are denoted by V := V 1 × V 2 , where V 1 is a vector-valued function space and V 2 is a scalar-valued function space. Without indices, the function space of the continuous problem statement is V ( Ω ) = H 0 1 ( Ω ) d × L 2 ( Ω ) . The space of lowest-order Taylor–Hood elements (quadratic FE for velocity, linear FE for pressure) is V h 2 / 1 ( T h ) , and the space of fourth-order FE for velocity and second-order FE for pressure on a given spatial mesh T h is V h 4 / 2 ( T h ) . Temporal function spaces are denoted by 𝑋. Without indices, 𝑋 and X ~ are the continuous level function spaces without and with discontinuities respectively. The space of discontinuous finite elements in time of 𝑟-th order is X k dG ( r ) , and the space of discontinuous finite elements in time of ( r + 1 ) -th order is X k dG ( r + 1 ) . We denote spatio-temporal function spaces by combining a spatial and a temporal function space,

X ( I , V ) := { ( v p ) | v L 2 ( I , V 1 ) , t v L 2 ( I , V 1 ) , p L 2 ( I , V 2 ) , Ω p ( t ) d x = 0 for all t I }

that will be the space-time Hilbert space on which we define the continuous weak formulation. Here, the dual space of V 1 is denoted by V 1 * = L ( V 1 , R ) . This is a good choice for the continuous level function space since we have a continuous embedding into a function space where the fluid velocity is continuous on I ¯ (see [22]). This ensures that the initial condition is well-defined for the velocity.

Remark 2

When we have homogeneous Neumann boundary conditions σ n = 0 on Γ out Ω , the so-called do-nothing (outflow) condition [30], the pressure does not require normalization for its uniqueness. As the Dirichlet constraints are only set on part of the boundary, i.e. Γ D = Ω Γ out , the continuous spatial function space is V = H 0 , Γ D 1 ( Ω ) d × L 2 ( Ω ) . Then we have the spatio-temporal function space

X ( I , V ) := { ( v p ) | v L 2 ( I , V 1 ) , t v L 2 ( I , V 1 ) , p L 2 ( I , V 2 ) } .

Finally, to simplify notation, we introduce

( f , g ) := ( f , g ) L 2 ( Ω ) := Ω f g d x , ( ( f , g ) ) := ( f , g ) L 2 ( I , L 2 ( Ω ) ) := I ( f , g ) d t .

In this notation, f g represents the Euclidean inner product if 𝑓 and 𝑔 are scalar- or vector-valued and the Frobenius inner product if 𝑓 and 𝑔 are matrices.

For a given initial value v 0 L 2 ( Ω ) d , the weak formulation reads as follows: find

U := ( v p ) ( v D 0 ) + X ( I , V ( Ω ) )

such that

A ( U ) ( Φ ) := ( ( t v , Φ v ) ) + a ( U ) ( Φ ) + ( v ( 0 ) v 0 , Φ v ( 0 ) ) = 0 for all Φ := ( Φ v Φ p ) X ( I , V ( Ω ) )

where, for the Navier–Stokes equations, we have

a ( U ) ( Φ ) := ( ( p , x Φ v ) ) + ν ( ( x v , x Φ v ) ) + ( ( ( v x ) v , Φ v ) ) + ( ( x v , Φ p ) ) ,

and for the Stokes equations, we have

a ( U ) ( Φ ) := ( ( p , x Φ v ) ) + ν ( ( x v , x Φ v ) ) + ( ( x v , Φ p ) ) .

2.1 Discretization in Time

Let T k := { I m , m = 1 , , M } , with I m := ( t m 1 , t m ) , be a partitioning of time, i.e. I ¯ = [ 0 , T ] = m = 1 M I ¯ m . Then, in an intermediate step similar to spatial dG methods, we can define a broken function space that allows for discontinuities at the temporal grid points,

(2.1) X ~ ( T k , V ) := { ( v p ) L 2 ( I , L 2 ( Ω ) d + 1 ) , ( v p ) | I m X ( I m , V ) for all I m T k } .

As local continuous embeddings into C ( I ¯ m , V 1 ) exist for 𝒗, we can define the limits of 𝒇 at time t m from above and from below for a function f X ~ ( T k , V ) as

f m ± := lim ϵ 0 f ( t m ± ϵ ) ,

and the jump of the function value of 𝒇 at time t m as [ f ] m := f m + f m . With this, we can introduce a weak formulation with discontinuities on the continuous level by interval-wise evaluation of the temporal integrals and inserting jumps between the temporal elements.

Find

U := ( v p ) ( v D 0 ) + X ~ ( T k , V ( Ω ) )

such that

(2.2) A ~ ( U ) ( Φ k ) := m = 1 M I m ( t v , Φ v ) ( p , x Φ v ) + ν ( x v , x Φ v ) d t + m = 1 M I m ( ( v x ) v , Φ v ) + ( x v , Φ p ) d t + m = 1 M 1 ( [ v ] m , Φ m v , + ) + ( v , 0 + ( 0 ) v 0 , Φ 0 v , + ( 0 ) ) = 0 for all Φ := ( Φ v Φ k p ) X ~ ( T k , V ( Ω ) ) .

As functions f X ( I , V ( Ω ) ) are continuous on 𝐼, [ f ] m = 0 holds for all m = 1 , , M . Therefore, it follows that X ( I , V ( Ω ) ) X ~ ( T k , V ( Ω ) ) holds as well and the above weak formulation is consistent. By discretizing the space-time function space (2.1) in time with the discontinuous Galerkin method of order r N 0 ( dG ( r ) ), we obtain the semi-discrete space as

X k dG ( r ) ( T k , V ) := { ( v k p k ) | v k | I m P r ( I m , V 1 ) , p k | I m P r ( I m , V 2 ) for all 1 m M , Ω p k ( t ) d x = 0 for all t I } X ~ ( T k , V ) .

Here, P r ( I m , Y ) is the space of polynomials of order 𝑟, which map from the time interval I m into the function space 𝑌. The dG ( r ) time discretization for the cases r = 0 and r = 1 is illustrated in Figure 1.

Figure 1

dG ( r ) time discretization.

(a) 
                     
                        
                           
                              
                                 
                                    dG
                                    
                                       (
                                       0
                                       )
                                    
                                 
                              
                              
                              \operatorname*{dG}(0)
(a)

dG ( 0 )

(b) 
                     
                        
                           
                              
                                 
                                    dG
                                    
                                       (
                                       1
                                       )
                                    
                                 
                              
                              
                              \operatorname*{dG}(1)
(b)

dG ( 1 )

It has been derived in [54] for the Navier–Stokes equations that we obtain the backward Euler scheme by using suitable numerical quadrature for the temporal integrals of the dG ( 0 ) discretization. Therefore, it can be seen as a variant of that scheme. Additionally, the corresponding time-stepping scheme of dG ( 1 ) has been formulated therein. The dG ( r ) time discretization of the Navier–Stokes equations reads as follows: find

U k := ( v k p k ) ( v D 0 ) + X k dG ( r ) ( T k , V ( Ω ) )

such that

A k ( U k ) ( Φ k ) := m = 1 M I m ( t v k , Φ k v ) ( p k , x Φ k v ) + ν ( x v k , x Φ k v ) d t + m = 1 M I m ( ( v k x ) v k , Φ k v ) + ( x v k , Φ k p ) d t + m = 1 M 1 ( [ v k ] m , Φ k , m v , + ) + ( v k , 0 + ( 0 ) v 0 , Φ k , 0 v , + ( 0 ) ) = 0 for all Φ k := ( Φ k v Φ k p ) X k dG ( r ) ( T k , V ( Ω ) ) .

2.2 Discretization in Space

Next, we describe the spatial discretization of the dG ( r ) formulation for the (Navier–)Stokes equations with continuous finite elements. Let T h be a shape-regular mesh of the domain Ω R d . The elements K T h are a non-overlapping covering of Ω ¯ , where the elements 𝐾 are quadrilaterals for d = 2 and hexahedrals for d = 3 . The discretization parameter ℎ is defined element-wise as h | K = h K = diam ( K ) . After adaptive refinement in space, h K can vary across elements and we need to account for hanging nodes in the assembly of the matrices and vectors. These are degrees of freedom that are located in the middle of the neighboring elements’ edge ( d = 2 ) or face ( d = 3 ). Furthermore, on different time intervals, we allow for changing spatial meshes, i.e. possibly T h ( I m ) T h ( I n ) , but for simplicity, we drop the dependence on the time interval and only mention it when it is not obvious from the context.

During the spatial discretization of the mixed system of the (Navier–)Stokes equations, we need to account for the inf-sup or Ladyzhenskaya–Babuska–Brezzi (LBB) condition

inf q L 2 ( Ω ) sup φ H 0 1 ( Ω ) d ( q , φ ) q L 2 ( Ω ) φ H 1 ( Ω ) d γ > 0 ,

which guarantees well-posedness (existence, uniqueness and stability) of the pressure solution under suitable boundary conditions and assumptions on the domain regularity; see e.g. [28, 50]. In the spatial discretization, we now replace L 2 ( Ω ) and H 0 1 ( Ω ) d with conforming finite element spaces that still fulfill a discrete version of this stability estimate. For this, we define the mixed finite element space

V h s v / s p ( T h ) := { ( v p ) C ( Ω ¯ ) d + 1 | v | K Q s v ( K ) d , p | K Q s p ( K ) for all K T h } ,

where Q s ( K ) is being constructed by mapping tensor-product polynomials of degree 𝑠 from the master element ( 0 , 1 ) d to the element 𝐾. In our computations, we use the function space V h 2 / 1 , the so-called Taylor–Hood elements [33] with quadratic finite elements for the velocity and linear finite elements for the pressure. Additionally, we use as V h 4 / 2 an enriched space with fourth-order finite elements for the velocity and second-order finite elements for the pressure. In R 2 , we could have also used V h 3 / 2 as an enriched space since Q 3 / Q 2 was shown to be an inf-sup stable Taylor–Hood pair [56], but this is not necessarily the case in 3D anymore. Hence, we approximate the enriched velocity by Q 4 instead of Q 3 finite elements in space. The fully discretized problem then reads as follows: find

U k h := ( v k h p k h ) ( v D 0 ) + X k dG ( r ) ( T k , V h s v / s p ( T h ) )

such that

(2.3) A k h ( U k h ) ( Φ k h ) := m = 1 M I m ( t v k h , Φ k h v ) ( p k h , x Φ k h v ) + ν ( x v k h , x Φ k h v ) d t + m = 1 M I m ( ( v k h x ) v k h , Φ k h v ) + ( x v k h , Φ k h p ) d t + m = 1 M 1 ( [ v k h ] m , Φ k h , m v , + ) + ( v k h , 0 + ( 0 ) v 0 , Φ k h , 0 v , + ( 0 ) ) = 0 for all Φ k h := ( Φ k h v Φ k h p ) X k dG ( r ) ( T k , V h s v / s p ( T h ) ) .

2.3 Tensor-Product Space-Time Discretization

The above space-time analysis is required for the theory of error estimation for the time-dependent partial differential equations. However, this formulation is usually transformed into a time-stepping scheme in the literature by applying a suitable numerical quadrature to the temporal integrals (see also Section 2.1). Instead, we will follow the approach of Bruchhäuser, Köcher and Bause [21, Section 4], which uses tensor-product spaces to directly work with the fully discrete weak formulation (2.3). In this approach, the space-time cylinder ( 0 , T ) × Ω is divided and discretized into space-time slabs

S l n := ( m = l n I m × T h m )

on which the weak formulation is solved sequentially. The spatial mesh is slabwise constant, i.e. T h l = = T h n .

For Hilbert function spaces H 1 ( I ) and H 2 ( Ω ) , the closure of the Hilbert space tensor product H 1 ( I ) ^ H 2 ( Ω ) is isometric to H 1 ( I , H 2 ( Ω ) ) , and the isometry also holds for tensor products of finite-dimensional subspaces [47, Chapter 1.2]. From this, it follows that f ( t ) g ( x ) with f H 1 ( I ) , g H 2 ( Ω ) can be identified with a function h ( t , x ) H 1 ( I , H 2 ( Ω ) ) . Hence, we can define the space-time tensor-product finite element basis on S l n by taking the tensor product of the temporal discontinuous Galerkin basis functions and the spatial continuous Galerkin basis functions. This allows for the extension of a simple finite element code for a stationary Stokes problem by creating a temporal mesh and adding temporal loops over elements, quadrature points and degrees of freedom in the assembly of the FEM matrices. Furthermore, unlike the time-stepping formulations of (2.3) where each polynomial degree in time needs to be derived and implemented separately, changing the temporal degree of the space-time discretization can be performed simply by changing the polynomial degree of the temporal finite elements and our code theoretically allows for hp-adaptivity in time since the temporal degree could be chosen on each slab individually. Another benefit of using space-time slabs for adaptive refinement is that the spatial mesh is fixed for the length of the slab which is sometimes desirable for numerical stability. On the other hand, using slabs with multiple time elements comes at the cost of solving larger linear equation systems.

Remark 3

Remark 3 (Support Points for the Temporal Elements)

As we use discontinuous elements in time, we are not limited to placing the outermost temporal degrees of freedom on the element edges. Instead, we can use arbitrary points without the need for recalculation of a time-stepping scheme. For spatial dG methods, Gauss–Legendre support points have been shown to provide a better discretization [23].

2.4 Linearization of the Semilinear Form

For the Navier–Stokes equations, we need to include the nonlinear convection term ( ( ( v x ) v , Φ v ) ) in our variational formulation A k h ( U k h ) ( Φ k h ) = 0 . Therefore, we need to apply linearization to obtain a solvable linear equation system. In this paper, we decided to solve this variational root finding problem by using the residual-based Newton method. The resulting nonlinear iteration step j + 1 N reads as follows: find

δ U k h := ( δ v k h δ p k h ) X k dG ( r ) ( T k , V h s v / s p ( T h ) )

such that

A k h , U ( U k h j ) ( δ U k h , Φ k h ) = A k h ( U k h j ) ( Φ k h ) for all Φ k h X k dG ( r ) ( T k , V h s v / s p ( T h ) ) .

Obtain the new iterate by

U k h j + 1 = U k h j + δ U k h .

Remark 4

Note that we will also need to assemble the directional derivative A k h , U for the adjoint problem (3.2) with the minor difference that there the trial and test functions are interchanged. This corresponds to transposing the system matrix from a Newton step.

The derivative of the complete semilinear form is given by

A k h , U ( U k h ) ( δ U k h , Φ k h ) = m = 1 M I m ( t δ v k h , Φ k h v ) ( δ p k h , x Φ k h v ) + ν ( x δ v k h , x Φ k h v ) d t + m = 1 M I m ( ( δ v k h x ) v k h + ( v k h x ) δ v k h , Φ k h v ) + ( x δ v k h , Φ k h p ) d t + m = 1 M 1 ( [ δ v k h ] m , Φ k h , m v , + ) + ( δ v k h , 0 + ( 0 ) , Φ k h , 0 v , + ( 0 ) ) .

The Newton algorithm was taken from [60] and the corresponding GitHub code[1], which could naturally be reused in our space-time setting. For the numerical tests, we used max line search = 10 as the maximum number of line search steps and λ = 0.6 as the line search damping parameter. We set the maximum number of Newton steps to max Newton = 10 and the tolerance TOL to 10 10 .

Since we are dealing with a time-dependent problem, the initial guess can be chosen as the solution of the previous slab evaluated at its end time, i.e. for the slabs S m 1 m 1 := ( t m 2 , t m 1 ) × T h and S m m := ( t m 1 , t m ) × T h , we define

U k h m , 0 U k h m 1 ( t m 1 ) .

Remark 5

Alternatively one could create the initial guess U k h m , 0 as the temporal extrapolant of the entire solution from the previous slab, e.g.

U k h m , 0 ( t ) := U k h m 1 ( t m 2 ) + ( t t m 2 ) U k h m 1 ( t m 1 ) U k h m 1 ( t m 2 ) t m 1 t m 2

for linear extrapolation in time.

Additionally, we need to prescribe the non-homogeneous Dirichlet conditions on the initial guess. However, the boundary conditions for the update δ U k h m become homogeneous, i.e. we enforce δ U k h m = 0 on the Dirichlet boundary.

In order to save some computational cost, to compute the directional derivative A k h , U ( U k h m , j ) ( δ U k h m , Φ k h ) at each step of the Newton method, in our program, we use the approximation

A k h , U ( U k h m , j + 1 ) ( , Φ k h ) A k h , U ( U k h m , j ) ( , Φ k h )

when the reduction rate

θ j := R ( U k h m , j + 1 ) R ( U k h m , j )

is sufficiently small. We use these so-called simplified-Newton steps when θ j θ max := 0.1 .

2.5 Divergence-Free Projection

In fluid dynamics, working with dynamically changing meshes leads to non-physical oscillations in the values of goal functionals, which has been proven in [16] for a Stokes model problem. Further analysis of the numerical artifacts caused by the violation of the inf-sup condition, due to the usage of time-varying spatial meshes, can be found in [19, 9, 2, 39]. To avoid the resulting numerical artifacts, Besier and Wollner [16] proposed to either repeat the last time step on the finite element space of the new slab or to use a divergence-free projection for the initial condition when switching to a different spatial mesh. Here, we decided to focus on the latter and computationally analyze the impact of a divergence-free L 2 and a divergence-free H 0 1 projection [16].

Remark 6

We notice that this approach uses a global correction, which is expensive and requires the solution of an additional problem. Using the Helmholtz decomposition, Linke [40] showed that the continuous velocity solution is invariant to irrotational forcings. However, conforming mixed finite elements lose these properties on the discrete level leading to non-physical solutions. Various fixes were proposed and studied using local projections to H ( div ) -conforming, i.e. discretely divergence-free, finite elements inside the weak formulation of the (Navier–)Stokes equations [40, 41, 35, 39]. However, these studies for fluid flow have all been done on triangular elements, and applying them with corresponding quadrilateral elements did not yield the desired results for our fluid flow formulations. In incompressible solid mechanics, quadrilateral elements with such pressure-robust schemes seem to work though [10, 11].

This problem can be formalized as follows. Let

S m m := ( t m 1 , t m ) × T h m and S m + 1 m + 1 := ( t m , t m + 1 ) × T h m + 1

be two space-time slabs with different spatial triangulations T h m T h m + 1 . Then we need to find a projection Π of the initial condition

U k h ( t m ) := ( v k h ( t m ) p k h ( t m ) ) ( v D ( t m ) 0 ) + V h s v / s p ( T h m )

into the function space of the next slab

( v D ( t m ) 0 ) + V h s v / s p ( T h m + 1 )

such that the projected initial condition is divergence free, i.e.

( x ( Π v k h ( t m ) ) , Φ h p ) = 0 for all Φ h p V h s p ( T h m + 1 ) .

To achieve this goal, Besier and Wollner [16] presented and analyzed a divergence-free L 2 projection and a divergence-free H 0 1 projection.

  • Divergence-free L 2 projection: find

    Π U k h ( t m ) = ( Π v k h ( t m ) Π p k h ( t m ) ) ( v D ( t m ) 0 ) + V h s v / s p ( T h m + 1 )

    such that

    ( Π v k h ( t m ) , Φ h v ) ( Π p k h ( t m ) , x Φ h v ) + ( x ( Π v k h ( t m ) ) , Φ h p ) = ( v k h ( t m ) , Φ h v ) for all Φ h := ( Φ h v Φ h p ) V h s v / s p ( T h m + 1 ) .

  • Divergence-free H 0 1 projection: find

    Π U k h ( t m ) = ( Π v k h ( t m ) Π p k h ( t m ) ) ( v D ( t m ) 0 ) + V h s v / s p ( T h m + 1 )

    such that

    ( x ( Π v k h ( t m ) ) , x Φ h v ) ( Π p k h ( t m ) , x Φ h v ) + ( x ( Π v k h ( t m ) ) , Φ h p ) = ( x v k h ( t m ) , x Φ h v ) for all Φ h := ( Φ h v Φ h p ) V h s v / s p ( T h m + 1 ) .

3 Space-Time Dual-Weighted Residual Error Estimation

In many practical applications, we are in general not interested in the entire solution of the Navier–Stokes equations but only in some quantity of interest. In this chapter, we will describe how we can use the Dual-Weighted Residual (DWR) method [14, 13] (see also other references [8, 55, 15, 1, 46]) to estimate the overall error of our finite element solution in this quantity of interest and how we can localize the error with a space-time partition of unity (PU) [51, 58] to refine the spatial and temporal meshes.

3.1 DWR for Parabolic PDEs

In the following analysis, we will work with homogeneous boundary conditions to simplify the presentation. Let us consider a goal functional J : X ~ ( I , V ( Ω ) ) R of the form

J ( U ) = 0 T J 1 ( U ( t ) ) d t + J 2 ( U ( T ) ) ,

which represents our physical quantity of interest. We now want to minimize the difference between the quantity of interest of the continuous solution 𝑼 and its FEM approximation U k h , i.e.

J ( U ) J ( U k h )

subject to the constraint that the variational formulation is being satisfied. As in constrained optimization, we define the Lagrange functionals

L : X ~ ( I , V ( Ω ) ) × X ~ ( I , V ( Ω ) ) R , ( U , Z ) J ( U ) A ( U ) ( Z ) , L k : X k dG ( r ) ( T k , V ( Ω ) ) × X k dG ( r ) ( T k , V ( Ω ) ) R , ( U k , Z k ) J ( U k ) A k ( U k ) ( Z k ) , L k h : X k dG ( r ) ( T k , V h s v / s p ( T h ) ) × X k dG ( r ) ( T k , V h s v / s p ( T h ) ) R , ( U k h , Z k h ) J ( U k h ) A k h ( U k h ) ( Z k h ) .

The stationary points ( U , Z ) , ( U k , Z k ) and ( U k h , Z k h ) of the Lagrange functionals ℒ, L k and L k h need to satisfy the Karush–Kuhn–Tucker first-order optimality conditions. Firstly, the stationary points are solutions to the equations

L Z ( U , Z ) ( δ Z ) = 0 for all δ Z X ~ ( I , V ( Ω ) ) , L k , Z ( U k , Z k ) ( δ Z k ) = 0 for all δ Z k X k dG ( r ) ( T k , V ( Ω ) ) , L k h , Z ( U k h , Z k h ) ( δ Z k h ) = 0 for all δ Z k h X k dG ( r ) ( T k , V h s v / s p ( T h ) ) .

We call these equations the primal problems and their solutions 𝑼, U k and U k h the primal solutions. Secondly, the stationary points must also satisfy the equations

L U ( U , Z ) ( δ U ) = 0 for all δ U X ~ ( I , V ( Ω ) ) , L k , U ( U k , Z k ) ( δ U k ) = 0 for all δ U k X k dG ( r ) ( T k , V ( Ω ) ) , L k h , U ( U k h , Z k h ) ( δ U k h ) = 0 for all δ U k h X k dG ( r ) ( T k , V h s v / s p ( T h ) ) .

These equations are called the adjoint or dual problems and their solutions 𝒁, Z k and Z k h are the adjoint solutions. To be able to decide whether the spatial or temporal mesh should be refined, we split the total discretization error into

J ( U ) J ( U k h ) = ( J ( U ) J ( U k ) ) + ( J ( U k ) J ( U k h ) ) η k + η h ,

where η k stands for the temporal error estimate and η h stands for the spatial error estimate. Following [15, Theorem 5.2] and neglecting the remainder terms, which are of higher order, we get the error estimates

(3.1) J ( U ) J ( U k ) 1 2 ( ρ ( U k ) ( Z Z k ) + ρ * ( U k , Z k ) ( U U k ) ) = : η k , J ( U k ) J ( U k h ) 1 2 ( ρ ( U k h ) ( Z k Z k h ) + ρ * ( U k h , Z k h ) ( U k U k h ) ) = : η h ,

where we denote the primal and adjoint residuals as

ρ ( U ) ( δ Z ) := L k h , Z ( U , Z ) ( δ Z ) , ρ * ( U , Z ) ( δ U ) := L k h , U ( U , Z ) ( δ U ) .

We now have two problems to solve: the primal and the adjoint problem. All the remaining variables in the error estimator can be computed via higher-order interpolation or via interpolation to a lower-order space.

3.2 Primal Problem

To derive the primal problem, we need to compute the Gâteaux derivatives of the Lagrange functionals ℒ, L k and L k h with respect to the adjoint solution 𝒁. Using the definition of the continuous (𝐴), the time-discrete ( A k ) and the fully discrete ( A k h ) variational formulations, we observe that they are linear in the test functions, and hence we have

L Z ( U , Z ) ( δ Z ) = A ( U ) ( δ Z ) = 0 for all δ Z X ~ ( I , V ( Ω ) ) , L k , Z ( U k , Z k ) ( δ Z k ) = A k ( U k ) ( δ Z k ) = 0 for all δ Z k X k dG ( r ) ( T k , V ( Ω ) ) , L k h , Z ( U k h , Z k h ) ( δ Z k h ) = A k h ( U k h ) ( δ Z k h ) = 0 for all δ Z k h X k dG ( r ) ( T k , V h s v / s p ( T h ) ) .

To get the primal solution, we simply need to solve the weak formulation of the Navier–Stokes problem.

3.3 Adjoint Problem

To derive the adjoint problem, we need to compute the Gâteaux derivatives of the Lagrange functionals ℒ, L k and L k h with respect to the primal solution 𝑼. We then get

L U ( U , Z ) ( δ U ) = J U ( U ) ( δ U ) A U ( U ) ( δ U , Z ) = 0 for all δ U X ~ ( I , V ( Ω ) ) , L k , U ( U k , Z k ) ( δ U k ) = J U ( U k ) ( δ U k ) A k , U ( U k ) ( δ U k , Z k ) = 0 for all δ U k X k dG ( r ) ( T k , V ( Ω ) ) , L k h , U ( U k h , Z k h ) ( δ U k h ) = J U ( U k h ) ( δ U k h ) A k h , U ( U k h ) ( δ U k h , Z k h ) = 0 for all δ U k h X k dG ( r ) ( T k , V h s v / s p ( T h ) ) .

Therefore, to obtain the adjoint solution, we need to solve an additional equation, the adjoint problem

(3.2) A k h , U ( U k h ) ( δ U k h , Z k h ) = J U ( U k h ) ( δ U k h ) .

Remark 7

For linear PDEs, like the Stokes equations, and linear goal functionals, like the mean drag functional, the adjoint problems simplify to

L U ( U , Z ) ( δ U ) = J ( δ U ) A ( δ U ) ( Z ) = 0 for all δ U X ~ ( I , V ( Ω ) ) , L k , U ( U k , Z k ) ( δ U k ) = J ( δ U k ) A k ( δ U k ) ( Z k ) = 0 for all δ U k X k dG ( r ) ( T k , V ( Ω ) ) , L k h , U ( U k h , Z k h ) ( δ U k h ) = J ( δ U k h ) A k h ( δ U k h ) ( Z k h ) = 0 for all δ U k h X k dG ( r ) ( T k , V h s v / s p ( T h ) ) .

In this scenario, we have the error identities

(3.3) J ( U ) J ( U k ) = A k ( U k ) ( Z Z k ) , J ( U k ) J ( U k h ) = A k h ( U k h ) ( Z k Z k h ) .

We show this error identity for the spatial error contribution. Using the linearity of the goal functional and the definition of the adjoint problem, we have

J ( U k ) J ( U k h ) = J ( U k U k h ) = A k ( U k U k h ) ( Z k ) = A k h ( U k U k h ) ( Z k ) = A k h ( U k h ) ( Z k ) = A k h ( U k h ) ( Z k Z k h ) .

For the temporal error identity, we can follow the same steps with the continuous level weak formulation including discontinuities A ~ (2.2).

By Remark 7, we can thus write the adjoint problem (3.2) for the Stokes problem as

A k h ( δ U k h ) ( Z k h ) = J ( δ U k h ) m = 1 M I m ( t δ v k h , Z k h v ) ( δ p k h , x Z k h v ) + ν ( x δ v k h , x Z k h v ) + ( x δ v k h , Z k h p ) d t + m = 1 M 1 ( [ δ v k h ] m , Z k h , m v , + ) + ( δ v k h , 0 + ( 0 ) , Z k h , 0 v , + ( 0 ) ) = J ( δ U k h ) .

To move the time derivative from the test function to the adjoint solution, we apply integration by parts in time and get

m = 1 M I m ( δ v k h , t Z k h v ) ( δ p k h , x Z k h v ) + ν ( x δ v k h , x Z k h v ) + ( x δ v k h , Z k h p ) d t m = 1 M ( δ v k h , m , [ Z k h v ] m ) + ( δ v k h , M ( T ) , Z k h , M v , ( T ) ) = J ( δ U k h ) .

Similarly, to solve the adjoint problem of the Navier–Stokes equations, we now need to solve the general adjoint problem (3.2), which due to the linearity of our goal functional can be simplified to

A k h , U ( U k h ) ( δ U k h , Z k h ) = J ( δ U k h ) .

Note that we still need to access the primal solution U k h to be able to solve the adjoint equation. The adjoint problem for the Navier–Stokes equations reads

m = 1 M I m ( δ v k h , t Z k h v ) ( δ p k h , x Z k h v ) + ν ( x δ v k h , x Z k h v ) d t + m = 1 M I m ( ( δ v k h x ) v k h + ( v k h x ) δ v k h , Z k h v ) + ( x δ v k h , Z k h p ) d t m = 1 M ( δ v k h , m , [ Z k h v ] m ) + ( δ v k h , M ( T ) , Z k h , M v , ( T ) ) = J ( δ U k h ) .

3.4 PU-DWR Navier–Stokes Error Estimator

For linear PDEs, like the Stokes problem, and linear goal functionals, we have the error identity

J ( U ) J ( U k h ) = [ J ( U ) J ( U k ) ] + [ J ( U k ) J ( U k h ) ] = A k ( U k ) ( Z Z k ) = : η k A k h ( U k h ) ( Z k Z k h ) = : η h = : η ,

which was derived in (3.3). For the nonlinear Navier–Stokes equations, we do not have the error identity (3.3), but need to use approximation (3.1) which contains the primal residual 𝜌 and the adjoint residual ρ . Including the adjoint residual yields an estimator which is accurate up to a remainder term of third order, compared to a purely primal residual based error estimator (see [14, Section 2] or [8, Chapter 6]) with a second-order remainder term only. For mildly nonlinear problems, such as incompressible Navier–Stokes (being semilinear in a laminar flow regime, i.e. nonlinearity in a low-order term) and linear goal functionals, the estimator asymptotically yields the same primal and adjoint residuals. As verification in the stationary setting, we adopted [25, Section 7.2] and evaluated 𝜌 and ρ separately, getting I eff primal I eff adjoint 1 . Clearly, for highly nonlinear problems (e.g. quasi-linear or fully nonlinear) and nonlinear goal functionals, both estimator parts are necessary as demonstrated in [25, Figure 4] and [24, Section 6.5]. Since being sufficiently accurate for such settings and being cheaper, the primal estimator only has been used in fluid applications; see e.g. [14, Section 8], [20] or [32]. Hence, for the estimation of the temporal and spatial error contributions, we employ

J ( U ) J ( U k ) A k ( U k ) ( Z Z k ) , J ( U k ) J ( U k h ) A k h ( U k h ) ( Z k Z k h ) .

To localize the spatial and temporal error, we now include the partition of unity { χ i m } i , m with m = 1 M i T h m χ i m 1 in the adjoint weights to localize the error. For the partition of unity, we choose linear finite elements in space ( cG ( 1 ) ) and piecewise discontinuous finite elements in time ( dG ( 0 ) ). This allows for the localization of the error to the individual temporal elements and the vertices of the spatial elements. We have

(3.4) η k = m = 1 M i T h m A k ( U k ) ( ( Z Z k ) χ i m ) = : η k i , m , η h = m = 1 M i T h m A k h ( U k h ) ( ( Z k Z k h ) χ i m ) = : η h i , m .

Exemplarily, we will show the formula for the computation of the temporal error indicator η k i , m by applying the product rule

η k i , m = A k ( U k ) ( ( Z Z k ) χ i m ) = I m ( t v k , ( Z v Z k v ) χ i m ) ( p k , x { ( Z v Z k v ) χ i m } ) + ν ( x v k , x { ( Z v Z k v ) χ i m } ) d t I m ( ( v k x ) v k , ( Z v Z k v ) χ i m ) + ( x v k , ( Z p Z k p ) χ i m ) d t ( [ v k ] m 1 , ( Z m 1 v , + Z k , m 1 v , + ) χ i m ) = I m ( t v k , ( Z v Z k v ) χ i m ) ( p k , { x ( Z v Z k v ) } χ i m + ( Z v Z k v ) ( x χ i m ) ) d t I m ν ( x v k , { x ( Z v Z k v ) } χ i m + ( Z v Z k v ) ( x χ i m ) ) d t I m ( ( v k x ) v k , ( Z v Z k v ) χ i m ) + ( x v k , ( Z p Z k p ) χ i m ) d t ( [ v k ] m 1 , ( Z m 1 v , + Z k , m 1 v , + ) χ i m ) .

Here, v w stands for the tensor product of two vectors 𝒗 and 𝒘, i.e. ( v w ) i , j := v i w j and the initial condition is included as [ v k ] 0 := v + ( 0 ) v 0 . With this partition of unity approach, the spatial and temporal error can now be localized to each space-time degree of freedom of the cG ( 1 ) dG ( 0 ) space-time discretization and will later indicate which spatial or temporal elements have the largest error and should be refined. However, to test the quality of our error estimator, we will initially work with globally refined space-time meshes and monitor how well the estimated error coincides with the true error by computing the effectivity index

I eff := | η J ( U ) J ( U k h ) | = | η k + η h J ( U ) J ( U k h ) | .

We asymptotically expect I eff k , h 0 1 for problems that satisfy a saturation assumption with a constant in the upper bound that converges to 0 for k , h 0 (see [58]) (based on the ideas from the spatial case [25]).

3.5 Adaptive Algorithm

We will use the adaptive algorithm proposed in [58], which is a time-slabbing adaptation of the algorithm proposed in [55] to balance the temporal and the spatial discretization error.

Algorithm 1

Algorithm 1 (MARK and REFINE for the time-slabbing approach)

In our first numerical experiments, we chose a sufficiently big equilibration factor 𝑐 ( c = 10 7 ) to always refine both in space and time. In the future, we want to use a smaller equilibration constant 𝑐 to only refine either the temporal or the spatial discretization which dominates the total error. A computational analysis of the choice of the equilibration factor 𝑐 can be found in Appendix A.

After having computed the temporal error indicators for each temporal element and the spatial error indicators for each spatial element for each slab, we mark spatial and temporal elements for refinement according to Algorithm 1.

3.6 Evaluation of the Error Indicators

Until now, we have not discussed the practical realization of the PU-DWR error estimator (3.4). To make the error indicators computable, we need to make some approximations and get the unknown solutions by post-processing of the primal and adjoint solutions. In the works of Besier (and co-workers) [55, 54, 15], the linearization point in the temporal error has been replaced by the primal solution, i.e.

η k = A k ( U k ) ( Z Z k ) A k ( U k h ) ( Z Z k ) ,

and he has discussed that this additional approximation error is small in comparison to the total discretization error [55, Remark 3.2]. Therefore, we will use this simplification in the same way. We still need to think about the approximation of the weights Z Z k and Z k Z k h . This approximation depends on the choice of our finite element space for the adjoint problem, and in this work, we consider only a mixed-order error estimator.

For mixed order, the adjoint solution is of higher order in space and time than the primal solution. More concretely, we choose the primal solution to be Q 2 / Q 1 in space and dG ( 1 ) in time, whereas the adjoint solution is Q 4 / Q 2 in space and dG ( 2 ) in time. This setting has already been discussed in [12, Section 4], and in this work, we approximate the adjoint weights Z Z k and Z k Z k h using the adjoint solution Z k h ( 2 ) X k dG ( 2 ) ( T k , V h 4 / 2 ( T h ) ) by

Z Z k ( id I k dG ( 1 ) ) Z k h ( 2 ) , Z k Z k h ( id I h 2 / 1 ) I k dG ( 1 ) Z k h ( 2 ) .

Here, I h 2 / 1 : X ( I , V h 4 / 2 ( T h ) ) X ( I , V h 2 / 1 ( T h ) ) denotes the interpolation in space from the enriched spatial function space to the primal spatial function space and I k dG ( 1 ) : X k dG ( 2 ) ( T k , V ( Ω ) ) X k dG ( 1 ) ( T k , V ( Ω ) ) denotes the interpolation in time from the enriched temporal function space to the primal temporal function space.

From an implementational point of view, this approach is very convenient since it only requires interpolation of the adjoint solution to a lower-order function in space. This can be done by evaluating the adjoint solution at the space-time degrees of freedom of the lower-order space.

4 Numerical Tests

In this section, we perform some numerical experiments to verify the validity of our space-time error estimation framework. In the first two experiments, we consider two-dimensional time-dependent laminar flow around a cylinder, which is being modeled by the Stokes or Navier–Stokes equations. In a further experiment, we model the two-dimensional time-dependent laminar backward-facing step problem and do adaptive refinement using a nonlinear functional. The code is written in the C++ based FEM library deal.II [6, 5].

4.1 Stokes Flow around a Cylinder

In the first numerical experiment, we consider the test case 2D-3[2] from the Schäfer–Turek Navier–Stokes benchmarks [53]. The 2D-3 benchmark problem describes the two-dimensional laminar flow around a circular cylinder with a time-dependent inflow condition. In Figure 2, the domain Ω and the different boundary components Γ in , Γ out , Γ wall and Γ circle are depicted.

Figure 2 
                  Domain of the 2D Navier–Stokes benchmark problem.
Figure 2

Domain of the 2D Navier–Stokes benchmark problem.

Here, the kinematic viscosity is ν = 10 3 . The domain is defined as Ω := ( 0 , 2.2 ) × ( 0 , 0.41 ) B r ( 0.2 , 0.2 ) with r = 0.05 , where B r ( x , y ) denotes the ball around ( x , y ) with radius 𝑟 and we denote the time interval by I := ( 0 , 8 ) . On the boundary Ω , we prescribe Dirichlet and Neumann boundary conditions. We apply the no-slip condition, i.e. homogeneous Dirichlet boundary conditions v = 0 on Γ wall Γ circle , where

Γ wall = ( 0 , 2.2 ) × { 0 } ( 0 , 2.2 ) × { 0.41 } and Γ circle = B r ( 0.2 , 0.2 ) .

On the boundary Γ in = { 0 } × ( 0 , 0.41 ) , we enforce a time-dependent parabolic inflow profile, which is a type of inhomogeneous Dirichlet boundary condition, i.e. v = v D on Γ in . For the 2D-3 benchmark, the inflow parabola is given by

v D ( t , 0 , y ) = ( sin ( π t 8 ) 6 y ( 0.41 y ) 0.41 2 0 ) .

For brevity, we denote the Dirichlet boundary as

Γ D := Γ in Γ wall Γ circle .

The outflow boundary Γ out = { 2.2 } × ( 0 , 0.41 ) has Neumann boundary conditions. To get a correct outflow profile of the pressure, it has been discussed in [30] that the (modified) do-nothing condition ν ( x v ) n p n = 0 needs to be enforced on Γ out . For the non-symmetric stress tensor, this corresponds to homogeneous Neumann boundary conditions, i.e. σ n = 0 on Γ out . It can be shown that, for this problem, the Reynolds number satisfies 0 Re ( t ) 100 .

At first, we consider a simplification of this benchmark problem, where the nonlinearity is omitted such that we solve the Stokes equations. As mentioned in Remark 7, this results in an error identity for our primal error estimator and we expect I eff 1 under uniform refinement of the space-time meshes. For our calculations, we use the mean drag goal functional

J drag ( U ) := 1 8 0 8 c D ( t ) d t = 20 8 0 8 Γ circle σ ( U ) n e 1 d s d t 0.40284197 ,

where e 1 = ( 1 , 0 ) T , c D denotes the drag coefficient and the reference value has been calculated on a fine spatio-temporal grid with 10,240 temporal and 1,480,000 spatial degrees of freedom.

4.1.1 Uniform Refinement

Using our mixed-order error estimator, we perform a grid search over the number of primal temporal degrees of freedom N k { 20 , 40 , 80 , 160 , 320 , 640 } and over the number of primal spatial degrees of freedom

N h { 445 , 1610 , 6100 , 23720 , 93520 } .

Furthermore, we compare Gauss–Legendre and Gauss–Lobatto quadrature for the temporal support points of the dG ( r ) method.

Gauss–Legendre Support Points in Time

We observe in Table 1 that the effectivity index for the Stokes test case is almost constant under temporal refinement. This can be explained by Table 2 and Table 3, where we find that the discretization error consists mainly of the spatial discretization error, while the temporal discretization error is a few orders of magnitude smaller for this test case.

Looking at the columns of Table 2, i.e. keeping the number of temporal degrees of freedom fixed and only refining uniformly in space, we see that the temporal error estimator remains almost constant. This verifies that our temporal error estimator does not depend on spatial refinement.

Looking at the rows of Table 3, i.e. keeping the number of spatial degrees of freedom fixed and only refining uniformly in time, we see that the spatial error remains almost constant. This verifies that our spatial error estimator does not depend on temporal refinement. Additionally, we observe quadratic convergence of the spatial error with respect to the spatial mesh size ℎ.

Table 1

Effectivity indices of mixed order for Stokes 2D-3 with Gauss–Legendre support points in time.

N h \ N k 20 40 80 160 320
445 0.61 0.61 0.61 0.61 0.61
1,610 0.68 0.68 0.68 0.68 0.68
6,100 0.73 0.72 0.72 0.72 0.72
23,720 0.77 0.75 0.75 0.75 0.75
93,520 0.90 0.81 0.80 0.80 0.80
Table 2

Temporal error estimator of mixed order for Stokes 2D-3 with Gauss–Legendre support points in time.

N h \ N k 20 40 80 160 320
445 2.7901 10 6 5.2875 10 7 1.0091 10 7 1.7218 10 8 2.5813 10 9
1,610 1.0153 10 5 1.5887 10 6 2.2938 10 7 3.1912 10 8 4.3840 10 9
6,100 1.1674 10 5 1.9094 10 6 2.8545 10 7 4.0021 10 8 5.3819 10 9
23,720 1.1797 10 5 1.9447 10 6 2.9447 10 7 4.1974 10 8 5.7309 10 9
93,520 1.1806 10 5 1.9471 10 6 2.9515 10 7 4.2178 10 8 5.7855 10 9
Table 3

Spatial error estimator of mixed order for Stokes 2D-3 with Gauss–Legendre support points in time.

N h \ N k 20 40 80 160 320
445 3.7394 10 2 3.7397 10 2 3.7398 10 2 3.7398 10 2 3.7398 10 2
1,610 1.3111 10 2 1.3110 10 2 1.3110 10 2 1.3110 10 2 1.3110 10 2
6,100 4.0683 10 3 4.0674 10 3 4.0673 10 3 4.0673 10 3 4.0673 10 3
23,720 1.1460 10 3 1.1457 10 3 1.1457 10 3 1.1457 10 3 1.1457 10 3
93,520 3.0495 10 4 3.0486 10 4 3.0485 10 4 3.0485 10 4 3.0484 10 4
Table 4

Effectivity indices of mixed order for Stokes 2D-3 with Gauss–Lobatto support points in time.

N h \ N k 20 40 80 160 320
445 0.58 0.61 0.61 0.61 0.61
1610 0.58 0.65 0.67 0.68 0.68
6100 0.45 0.63 0.70 0.72 0.72
23720 0.23 0.49 0.66 0.73 0.75
93520 0.08 0.25 0.52 0.70 0.77
Table 5

Temporal error estimator of mixed order for Stokes 2D-3 with Gauss–Lobatto support points in time.

N h \ N k 20 40 80 160 320
445 1.0450 10 3 3.3808 10 4 8.4236 10 5 2.6087 10 4 3.2106 10 4
1,610 2.4703 10 4 1.6773 10 4 8.1390 10 5 1.2743 10 5 2.5248 10 5
6,100 4.2596 10 5 2.9024 10 5 2.2523 10 5 1.5217 10 5 7.6783 10 6
23,720 1.7540 10 5 4.5460 10 6 2.5516 10 6 2.0765 10 6 1.6699 10 6
93,520 1.5636 10 5 2.4541 10 6 4.8653 10 7 2.0288 10 7 1.5701 10 7
Table 6

Spatial error estimator of mixed order for Stokes 2D-3 with Gauss–Lobatto support points in time.

N h \ N k 20 40 80 160 320
445 3.8122 10 2 3.7657 10 2 3.7294 10 2 3.7132 10 2 3.7076 10 2
1,610 1.3234 10 2 1.3249 10 2 1.3185 10 2 1.3121 10 2 1.3085 10 2
6,100 4.0609 10 3 4.0857 10 3 4.0874 10 3 4.0820 10 3 4.0749 10 3
23,720 1.1383 10 3 1.1456 10 3 1.1473 10 3 1.1476 10 3 1.1473 10 3
93,520 3.0247 10 4 3.0437 10 4 3.0485 10 4 3.0496 10 4 3.0499 10 4
Gauss–Lobatto Support Points in Time

In contrast to Table 1, we observe in Table 4 that the effectivity indices with Gauss–Lobatto support points in time are smaller for coarse temporal meshes. However, for fine temporal scales, the effectivity indices from Table 4 converge to the effectivity indices of Table 1.

In Table 5, we observe that the temporal error estimator with Gauss–Lobatto support points is larger than in Table 2 for Gauss–Legendre. Therefore, from now on, we only use Gauss–Legendre support points in time.

Finally, in Table 6, we get similar spatial error estimates as in Table 3. The only difference is that the spatial error estimator changes slightly less under temporal refinement for Gauss–Legendre support points in time, which further motivates our preferred usage of this quadrature formula.

4.1.2 Adaptive Refinement

We now apply the space-time adaptivity, Algorithm 1, starting from the coarsest spatial mesh with N h = 445 primal DoFs and M = 20 slabs with one temporal element each, i.e. we have N k = 40 initial temporal degrees of freedom due to the dG ( 1 ) discretization. This results in 17,800 space-time degrees of freedom.

As a marking strategy in space, we use the averaging strategy, which marks all elements K T h such that

η h | K > α | T h | K T h η h | K .

In our computations, we used α = 1.1 . As a marking strategy in time, we use a fixed rate strategy (fixed number), where the ε % of temporal cells with the largest error are being refined. In our computations, we used ε = 75 .

Table 7

Adaptive refinement of mixed order on dynamic meshes for Stokes 2D-3.

#DoF(primal) #DoF(adjoint) 𝑀 η k η h 𝜂 J ( U ) J ( U k h ) I eff
17,800 96,600 20 5.2875 10 7 3.7397 10 2 3.7397 10 2 6.0872 10 2 0.61
55,328 303,288 36 5.6502 10 8 1.5619 10 2 1.5619 10 2 2.1453 10 2 0.73
259,712 1,468,032 64 3.8869 10 8 4.1840 10 3 4.1840 10 3 5.7545 10 3 0.73
902,838 5,137,110 113 8.8754 10 9 1.1668 10 3 1.1668 10 3 1.5598 10 3 0.75
3,438,496 19,668,018 199 7.5194 10 9 3.1632 10 4 3.1632 10 4 4.2411 10 4 0.75
Table 8

Adaptive refinement of mixed order on dynamic meshes for Stokes 2D-3 with divergence-free L 2 projection.

#DoF(primal) #DoF(adjoint) 𝑀 η k η h 𝜂 J ( U ) J ( U k h ) I eff
17,800 96,600 20 1.7268 10 6 3.7635 10 2 3.7634 10 2 5.9702 10 2 0.63
55,092 301,932 36 4.7644 10 8 1.5736 10 2 1.5736 10 2 2.1001 10 2 0.75
259,712 1,468,032 64 4.2556 10 8 4.2011 10 3 4.2011 10 3 5.3248 10 3 0.79
902,838 5,137,110 113 1.0787 10 8 1.1719 10 3 1.1718 10 3 1.3084 10 3 0.90
3,437,792 19,663,926 199 7.5608 10 9 3.1791 10 4 3.1790 10 4 2.7098 10 4 1.17
Table 9

Adaptive refinement of mixed order on dynamic meshes for Stokes 2D-3 with divergence-free H 0 1 projection.

#DoF(primal) #DoF(adjoint) 𝑀 η k η h 𝜂 J ( U ) J ( U k h ) I eff
17,800 96,600 20 1.9364 10 6 3.7704 10 2 3.7702 10 2 5.9309 10 2 0.64
55,092 301,932 36 9.5524 10 8 1.5745 10 2 1.5745 10 2 2.0791 10 2 0.76
259,712 1,468,032 64 4.7248 10 8 4.1966 10 3 4.1966 10 3 5.0362 10 3 0.83
902,838 5,137,110 113 1.1303 10 8 1.1696 10 3 1.1696 10 3 1.1248 10 3 1.04
3,439,256 19,672,314 199 7.3748 10 9 3.1672 10 4 3.1672 10 4 1.5370 10 4 2.06

In Table 7, we show the results from five adaptive refinement loops with a mixed-order error estimator for the Stokes 2D-3 problem. We do not use a divergence-free projection here, which leads to oscillations in the goal functionals (cf. Figure 4). However, the error estimates for the Stokes problem are still acceptable and the effectivity index does not change much.

In Table 8, we have the same setup as in Table 7 and additionally employ a divergence-free L 2 projection from Section 2.5. We observe that ensuring the incompressibility of the snapshot on the new mesh yields a smaller true error in the goal functional between the reference solution and the finite element solution than in Table 7. In contrast to Table 8, we use a divergence-free H 0 1 projection in Table 9, but we make similar observations when it comes to the true error.

Figure 3 
                     Comparison of the error in the mean drag for the Stokes problem for uniform and adaptive spatio-temporal refinement.
Figure 3

Comparison of the error in the mean drag for the Stokes problem for uniform and adaptive spatio-temporal refinement.

In Figure 3, we start with a common coarse spatio-temporal mesh and iteratively refine uniformly in space and time or adaptively refine in space and time based on Algorithm 1. As expected the dual-weighted residual method driven mesh refinement leads to faster convergence in the error in the mean drag functional. In particular, for adaptive refinement, a lot fewer degrees of freedom are needed to reach a prescribed accuracy in the goal functional, e.g. to achieve a tolerance of 2 10 3 , using uniform refinement requires 7,590,400 space-time degrees of freedom, whereas adaptive refinement only needs 902,838 space-time degrees of freedom for the primal problem.

Figure 4

Trajectory of the drag/lift coefficient for the Stokes problem for adaptive refinement.

(a) 
                        Drag coefficient
(a)

Drag coefficient

(b) 
                        Lift coefficient
(b)

Lift coefficient

In Figure 4, we compare the trajectory for the drag and lift coefficients over time for the Stokes 2D-3 problem and the different adaptive refinement methods. The drag trajectory is mostly smooth with only minor sawtooth-like oscillations near the maximal drag. However, the lift coefficient displays large spikes when using adaptive refinement without ensuring the incompressibility of the solution on the next spatial mesh. This underlines the necessity of methods like the divergence-free L 2 and H 0 1 projections from Besier and Wollner [16], which lead to smoother lift coefficient trajectories that more closely match the uniformly refined solution.

4.2 Navier–Stokes Flow around a Cylinder

We now include the nonlinear term in the variational formulation and solve the Navier–Stokes equations. In contrast to the Stokes problem from Section 4.1, there are plenty of reference results for the Navier–Stokes 2D-3 benchmark problem in the literature [53, 34], and we decided to use the publicly available results of the FeatFlow software[3] of the group at the TU Dortmund around Stefan Turek. The FeatFlow authors discretize in space with Q 2 / P 1 disc elements, and in time, they discretize with the Crank–Nicolson scheme. For their finest calculations, they use a fixed time step size of k = 0.000625 and 667,264 degrees of freedom in space. We use their mean drag

J drag ( U ) 1.6031368

as our reference solution. As marking strategies, we again used the averaging strategy in space and the fixer rate strategy in time with the same parameters as for the linear case ( α = 1.1 and ε = 75 ).

Table 10

Uniform refinement of mixed order on dynamic meshes for Navier–Stokes 2D-3.

#DoF(primal) #DoF(adjoint) 𝑀 η k η h 𝜂 J ( U ) J ( U k h ) I eff
17,800 96,600 20 8.4213 10 6 3.8557 10 1 3.8557 10 1 5.5869 10 1 0.69
128,800 732,000 40 1.3424 10 3 2.8297 10 1 2.8431 10 1 2.9690 10 1 0.96
976,000 5,692,800 80 6.8970 10 2 1.3391 10 1 2.0288 10 1 1.3978 10 1 1.45
7,590,400 44,889,600 160 5.5834 10 3 2.4517 10 2 3.0100 10 2 2.5428 10 2 1.18
Table 11

Adaptive refinement of mixed order on dynamic meshes for Navier–Stokes 2D-3.

#DoF(primal) #DoF(adjoint) 𝑀 η k η h 𝜂 J ( U ) J ( U k h ) I eff
17,800 96,600 20 8.4213 10 6 3.8557 10 1 3.8557 10 1 5.5869 10 1 0.69
63,336 350,244 36 2.3497 10 7 2.6516 10 1 2.6516 10 1 2.6719 10 1 0.99
228,568 1,285,878 64 1.7174 10 3 1.0803 10 2 9.0853 10 3 1.2438 10 1 0.07
835,942 4,749,933 113 1.3694 10 1 1.7656 10 1 3.9619 10 2 2.7110 10 2 1.46
3,008,930 17,274,966 199 3.9451 10 3 1.1779 10 2 1.5724 10 2 1.1074 10 2 1.42
Table 12

Adaptive refinement of mixed order on dynamic meshes for Navier–Stokes 2D-3 with divergence-free L 2 projection.

#DoF(primal) #DoF(adjoint) 𝑀 η k η h 𝜂 J ( U ) J ( U k h ) I eff
17,800 96,600 20 9.3298 10 6 3.8542 10 1 3.8541 10 1 5.5752 10 1 0.69
63,454 350,922 36 1.4015 10 7 2.6505 10 1 2.6505 10 1 2.6642 10 1 0.99
230,032 1,294,482 64 8.9182 10 4 1.2571 10 2 1.1679 10 2 1.2586 10 1 0.09
828,744 4,706,883 113 1.1615 10 1 7.6888 10 2 3.9265 10 2 2.5449 10 2 1.54
3,004,686 17,251,722 199 4.3194 10 3 1.9094 10 2 2.3414 10 2 1.9674 10 2 1.19
Table 13

Adaptive refinement of mixed order on dynamic meshes for Navier–Stokes 2D-3 with divergence-free H 0 1 projection.

#DoF(primal) #DoF(adjoint) 𝑀 η k η h 𝜂 J ( U ) J ( U k h ) I eff
17,800 96,600 20 9.1676 10 6 3.8548 10 1 3.8547 10 1 5.5768 10 1 0.69
63,690 352,278 36 1.4710 10 6 2.6493 10 1 2.6493 10 1 2.6667 10 1 0.99
234,878 1,322,502 64 7.5795 10 4 4.4232 10 3 3.6653 10 3 1.2633 10 1 0.03
834,710 4,741,881 113 2.3546 10 3 7.6972 10 2 7.9327 10 2 1.9651 10 2 4.04
3,044,708 17,485,449 199 3.0977 10 3 9.0900 10 3 1.2188 10 2 6.5227 10 3 1.87

In Table 10, we observe that the effectivity indices are close to 1 under uniform spatio-temporal refinement for the Navier–Stokes 2D-3 problem. We want to remark again that, for this nonlinear problem, our primal error estimator no longer satisfies the error identity, but we can computationally verify its accuracy.

In Table 11, we display the results from five DWR refinement loops with a mixed-order error estimator for the Navier–Stokes 2D-3 problem. Although we do not employ a divergence-free projection, the effectivity indices hint at a good match between the true error and the error estimator. However, in the third loop, a significant underestimation of the error in the mean drag can be observed which nevertheless does not persist after further refinement.

In Table 12, we observe that the accuracy of the error estimator is improved compared to Table 11 by employing a divergence-free L 2 projection. Additionally, this reduces the true error and thus yields a faster convergence toward the true solution.

In Table 13, we have the same setup as in Table 12 but now use a divergence-free H 0 1 projection. In comparison to Table 12, the effectivity indices are now slightly worse, but the true error is lower.

Figure 5 
                  Comparison of the error in the mean drag for the Navier–Stokes problem for uniform and adaptive spatio-temporal refinement.
Figure 5

Comparison of the error in the mean drag for the Navier–Stokes problem for uniform and adaptive spatio-temporal refinement.

In Figure 5, we visualize the true error from Tables 10, 11, 12 and 13. Analogous to Figure 3, we detect that the error of our adaptively refined solution converges faster than a uniformly refined solution.

Figure 6

Trajectory of the drag/lift coefficient for the Navier–Stokes problem for adaptive refinement.

(a) 
                     Drag coefficient
(a)

Drag coefficient

(b) 
                     Lift coefficient
(b)

Lift coefficient

In Figure 6, we compare the trajectory of the drag and lift coefficient for the Navier–Stokes 2D-3 benchmark problem under different kinds of adaptive refinement. Like we observed in Figure 4, the divergence-free projections create less oscillatory drag and lift trajectories. This can be seen especially for the lift coefficient, where the projections help to remove the large spikes caused by using dynamic meshes. At the same time, this causes a slight damping between t = 4 s and t = 6 s .

Figure 7 
                  Snapshots of the velocity magnitude of the primal Navier–Stokes problem after four adaptive refinements with divergence-free 
                        
                           
                              
                                 L
                                 2
                              
                           
                           
                           L^{2}
                        
                      projection.
Figure 7

Snapshots of the velocity magnitude of the primal Navier–Stokes problem after four adaptive refinements with divergence-free L 2 projection.

In Figure 7, we show some dynamic adaptive spatial meshes together with the snapshots of the velocity magnitude of the primal Navier–Stokes problem after four adaptive refinements with divergence-free L 2 projection. We observe that mainly the region around the cylinder is being refined, which is to be expected given that the drag coefficient is a boundary integral. Further away from the cylinder, e.g. close to the outflow boundary, the solution does not need to be resolved with the same accuracy. Moreover, since we are using dynamic meshes, the spatial mesh refinement can change over time as shown in Figure 7.

Figure 8 
                  Temporal mesh in the final DWR loop for Navier–Stokes 2D-3 with divergence-free 
                        
                           
                              
                                 L
                                 2
                              
                           
                           
                           L^{2}
                        
                      projection.
Figure 8

Temporal mesh in the final DWR loop for Navier–Stokes 2D-3 with divergence-free L 2 projection.

In Figure 8, we show the temporal mesh in the fifth DWR loop for the Navier–Stokes 2D-3 benchmark when using a divergence-free L 2 projection. The temporal mesh is also being refined adaptively, and we can see that, in some regions, the diameter of the temporal elements is smaller, especially in the second half of the temporal domain.

4.3 Navier–Stokes Flow over Backward Facing Step

In the third numerical experiment, we consider a version of the two-dimensional laminar backward-facing step benchmark [4, 17] with a time-dependent inflow condition. The domain is shown in Figure 9.

Figure 9 
                  Domain of the 2D backward-facing step benchmark.
Figure 9

Domain of the 2D backward-facing step benchmark.

The domain is defined as Ω := ( 1 2 , 0 ) × ( 1 2 , 1 ) ( 0 , 2 ) × ( 0 , 1 ) . The kinematic viscosity is ν = 1 , and we consider the time interval I = ( 0 , 8 ) . We prescribe homogeneous Neumann boundary conditions on the outflow boundary Γ out := { 2 } × ( 0 , 1 ) . On the inflow boundary Γ in := { 1 2 } × ( 1 2 , 1 ) , we enforce the time-dependent parabolic inflow profile

v D ( t , 1 2 , y ) = ( sin ( π t 8 ) 80 ( 1 + ( 3 2 y ) y ) 0 ) .

On all remaining boundaries, homogeneous Dirichlet boundary conditions are being applied. In contrast to the previous test cases, we now consider the mean squared vorticity

J vorticity ( U ) = 1 8 0 8 Ω ( × v ) 2 d x d t = 1 8 0 8 Ω ( 1 v 2 2 v 1 ) 2 d x d t

as our goal functional. Note that this is a nonlinear goal functional and the right-hand side of the dual problem is given by

J vorticity , U ( U ) ( δ U ) = 2 8 0 8 Ω ( × v ) ( × δ v ) d x d t .

Using 80 initial temporal degrees of freedom and a spatial grid with 1,439 primal degrees of freedom, we obtained the series of functional values from Table 14. We can see that this series converges from below and that the accuracy is increasing with each refinement step.

Table 14

Mean vorticity value under uniform spatio-temporal refinement for the backward-facing step test case.

#temporal DoF(primal) #spatial DoF(primal) J ( U k h )
80 1,439 506.20900857749609
160 5,467 510.27887058685025
320 21,299 512.24259375586087
640 84,067 513.18333123881553
1,280 334,019 513.63122326525161
2,560 1,331,587 513.84343972465513

For the sake of brevity, we will only discuss adaptive results in detail. For this test case, we used the fixed rate marking strategy both in space and time with ε = 30 .

Table 15

Signed effectivity indices for the different projections with a once coarsened spatio-temporal mesh (coarse) compared to the finest reference mesh (fine). Positive I eff means the computed functional value is below the reference value.

DWR loop Coarse I eff Fine I eff Coarse I eff L 2 Fine I eff L 2 Coarse I eff H 0 1 Fine I eff H 0 1
1 1.06 1.03 1.05 1.03 1.06 1.03
2 1.15 1.08 1.15 1.08 1.10 1.04
3 1.34 1.17 1.34 1.17 1.83 1.53
4 2.02 1.37 2.21 1.47 −2.51 −3.97
5 −1.11 2.06 −1.53 2.47 −6.46 −0.70

Table 15 shows the resulting effectivity indices for the different projection strategies (no projection, L 2 projection, H 0 1 projection). To investigate the effect of the reference value on the approximated real error, we compare the results with the last (fine) and second-to-last (coarse) mean vorticity as reference values J vorticity ( U ) . As long as the vorticity on the adaptive mesh is below the reference value, we can see that we get a better effectivity with the finer reference value, suggesting that our estimator is a good predictor of the true error. Additionally, we see a bad effectivity index for the loop in which we have the same h min as in the reference grid. However, this is not surprising as it is generally not advisable to analyze adaptive results on the reference mesh level.

To overcome the issue of the unreliable reference solution and since the functional values in Table 14 converge nicely, we extrapolated a new value from the two finest values using the formula in [3, Section 17.7.5] with h 2 = h 1 2 , obtaining

J extr. := ( J 2 1 4 J 1 ) 4 3 = 513.9141785444563 .

Additionally, we analyzed the differences between consecutive functional values, i.e. d J n := J n J n 1 , where J n is the functional value at grid level 𝑛 starting with J 0 = 506.209 . As these differences are roughly halved in each uniform refinement step, we applied linear regression to the five log ( J n ) values and obtained

log ( d J ( n ) ) 2.148469873234862 0.738556050505728 n .

With this, we calculated new functional values by

J reg. ( n ) = J 0 + k = 1 n exp ( log ( d J ( k ) ) ) .

Using this formula, we obtained two new functional values

J reg. ( 6 ) = 513.9587146059994 and J reg. ( 25 ) = 514.0520354056016 .

Table 16

Signed effectivity indices for the different projections with an extrapolated reference value (extr.) vs. reference values J reg. ( 6 ) and J reg. ( 25 ) obtained by regression of the log-differenced function values. Positive I eff means the computed functional value is below the reference value.

Loop Extr. I eff Reg. 6 I eff Reg. 25 I eff Extr. I eff L 2 Reg. 6 I eff L 2 Reg. 25 I eff L 2 Extr. I eff H 0 1 Reg. 6 I eff H 0 1 Reg. 25 I eff H 0 1
1 1.02 1.02 1.00 1.01 1.01 1.00 1.02 1.02 1.00
2 1.06 1.05 1.02 1.06 1.04 1.02 1.02 1.01 0.98
3 1.12 1.09 1.03 1.12 1.09 1.03 1.45 1.40 1.31
4 1.23 1.16 1.04 1.32 1.24 1.11 −4.94 −5.81 −9.30
5 1.54 1.33 1.03 1.78 1.51 1.15 −0.72 −0.73 −0.76

Table 16 shows that extrapolation of the reference value greatly improves the effectivity indices. In particular, recursively applying the extrapolation formula a few times (cf. reg. 25) yields effectivity indices much closer to 1 when using no projection or an L 2 projection. However, the H 0 1 projection results still do not look good on the finer meshes.

Figure 10 
                  Comparison of the true error in the mean squared vorticity for the backward-facing step for uniform and adaptive spatio-temporal refinement using 
                        
                           
                              
                                 
                                    J
                                    reg.
                                 
                                 ⁢
                                 
                                    (
                                    25
                                    )
                                 
                              
                           
                           
                           J_{\textup{reg.}}(25)
                        
                      as the reference value.
Figure 10

Comparison of the true error in the mean squared vorticity for the backward-facing step for uniform and adaptive spatio-temporal refinement using J reg. ( 25 ) as the reference value.

In Figure 10, we visualize the true error in the mean squared vorticity for the backward-facing step for uniform and adaptive spatio-temporal refinement. Analogous to Figure 3 and Figure 5 for the (Navier–)Stokes 2D-3 benchmark, we detect that the error of our adaptively refined solution converges with a higher rate than a uniformly refined solution. In particular, when comparing the fourth refinement cycle, we observe that adaptive refinement uses only 1 6 of the number of space-time degrees of freedom as uniform refinement to reach a tolerance of 0.7 in the goal functional.

Figure 11 
                  Snapshot of the velocity magnitude and streamlines of the primal backward-facing step problem at 
                        
                           
                              
                                 t
                                 ≈
                                 4
                              
                           
                           
                           t\approx 4
                        
                     .
Figure 11

Snapshot of the velocity magnitude and streamlines of the primal backward-facing step problem at t 4 .

Figure 12 
                  Snapshot of the vorticity magnitude of the primal backward-facing step problem at 
                        
                           
                              
                                 t
                                 ≈
                                 4
                              
                           
                           
                           t\approx 4
                        
                     .
Figure 12

Snapshot of the vorticity magnitude of the primal backward-facing step problem at t 4 .

Figure 13 
                  Spatial grid at 
                        
                           
                              
                                 t
                                 ≈
                                 4
                              
                           
                           
                           t\approx 4
                        
                      for the backward-facing step problem after 4 adaptive refinements with divergence-free 
                        
                           
                              
                                 L
                                 2
                              
                           
                           
                           L^{2}
                        
                      projection.
Figure 13

Spatial grid at t 4 for the backward-facing step problem after 4 adaptive refinements with divergence-free L 2 projection.

In Figure 11, we show the snapshot of the velocity magnitude of the primal backward-facing step problem at time t 4 and its streamlines. Looking at the streamlines, we observe a corner vortex in the lower left corner after the step. In Figure 12, we plot the snapshot of the vorticity magnitude of the primal backward-facing step problem at time t 4 . It can be seen that the vorticity magnitude is largest at the reentrant corner x = ( 0 , 1 2 ) .

In Figure 13, we display a four times adaptively refined grid with divergence-free L 2 projection. On the one hand, we observe that the grid is mostly being refined close to the reentrant corner where the vorticity magnitude is at its maximum. On the other hand, closer to the inflow or outflow boundaries, the mesh is much coarser.

5 Conclusions and Outlook

In this work, we formulated a space-time PU-DWR approach for goal-oriented error control and local mesh adaptivity for the incompressible Navier–Stokes equations. Therein, our focus was temporal and spatial error control, which has been computationally analyzed for different benchmark problems such as the Schäfer–Turek 1996 benchmark. In terms of convergence as well as evaluation of effectivity indices, we observed excellent performances for different goal functionals. In ongoing work, an interesting aspect would be to implement Linke’s pressure-robust discretizations rather than the global corrections proposed by Besier and Wollner.

Award Identifier / Grant number: 433082294

Funding statement: The first and fourth authors acknowledge the funding of the German Research Foundation (DFG; http://dx.doi.org/10.13039/501100001659) within the framework of the International Research Training Group on Computational Mechanics Techniques in High Dimensions GRK 2657 under Grant Number 433082294.

A Computational Tests of the Equilibration Factor in Algorithm 1

A.1 Changing Equilibration Factor with a Fixed Number of Initial Temporal Elements

To test the impact of the equilibration constant 𝑐 on the adaptive results, we repeated the simulation from Table 11 – the adaptive refinement of Navier–Stokes 2D-3 without divergence-free projection – with

c { 1 , 5 , 10 3 , 10 5 , 10 7 } .

As shown in these tables, we need to increase the equilibration constant 𝑐 to refine both the spatial and temporal meshes. In particular, for c = 1 (cf. Table 17) and c = 5 (cf. Table 18), the temporal mesh is not being refined at all since the spatial error dominates the temporal error. If we then increase the equilibration constant to c = 10 3 (cf. Table 19) or c = 10 5 (cf. Table 20), then we get some temporal refinement, but not in every refinement cycle. This motivates larger values for the equilibration constant, e.g. c = 10 7 (cf. Table 21 and Table 11), to refine both the spatial and temporal meshes.

Table 17

Adaptive refinement of Navier–Stokes 2D-3 with c = 1 .

#DoF(primal) #DoF(dual) 𝑀 η k η h 𝜂 J ( U ) J ( U k h ) I eff
17,800 96,600 20 8.4213 10 6 3.8557 10 1 3.8557 10 1 5.5869 10 1 0.69
35,428 195,978 20 3.7918 10 5 2.6520 10 1 2.6516 10 1 2.6716 10 1 0.99
72,440 407,643 20 3.9093 10 4 1.4886 10 2 1.4495 10 2 1.2598 10 1 0.12
143,692 815,421 20 1.5155 10 4 2.2525 10 2 2.2374 10 2 5.1573 10 2 0.43
285,260 1,633,776 20 2.1708 10 5 3.1083 10 3 3.0866 10 3 3.1490 10 2 0.10
Table 18

Adaptive refinement of Navier–Stokes 2D-3 with c = 5 .

#DoF(primal) #DoF(dual) 𝑀 η k η h 𝜂 J ( U ) J ( U k h ) I eff
17,800 96,600 20 8.4213 10 6 3.8557 10 1 3.8557 10 1 5.5869 10 1 0.69
35,428 195,978 20 3.7918 10 5 2.6520 10 1 2.6516 10 1 2.6716 10 1 0.99
72,440 407,643 20 3.9093 10 4 1.4886 10 2 1.4495 10 2 1.2598 10 1 0.12
143,692 815,421 20 1.5155 10 4 2.2525 10 2 2.2374 10 2 5.1573 10 2 0.43
285,260 1,633,776 20 2.1708 10 5 3.1083 10 3 3.0866 10 3 3.1490 10 2 0.10
Table 19

Adaptive refinement of Navier–Stokes 2D-3 with c = 10 3 .

#DoF(primal) #DoF(dual) 𝑀 η k η h 𝜂 J ( U ) J ( U k h ) I eff
17,800 96,600 20 8.4213 10 6 3.8557 10 1 3.8557 10 1 5.5869 10 1 0.69
35,428 195,978 20 3.7918 10 5 2.6520 10 1 2.6516 10 1 2.6716 10 1 0.99
72,440 407,643 20 3.9093 10 4 1.4886 10 2 1.4495 10 2 1.2598 10 1 0.12
263,200 1,494,210 36 2.7724 10 1 2.4822 10 1 2.9025 10 2 5.1403 10 2 0.56
995,472 5,718,699 64 1.4090 10 1 2.5893 10 1 1.1804 10 1 3.8993 10 2 3.03
Table 20

Adaptive refinement of Navier–Stokes 2D-3 with c = 10 5 .

#DoF(primal) #DoF(dual) 𝑀 η k η h 𝜂 J ( U ) J ( U k h ) I eff
17,800 96,600 20 8.4213 10 6 3.8557 10 1 3.8557 10 1 5.5869 10 1 0.69
63,336 350,244 36 2.3497 10 7 2.6516 10 1 2.6516 10 1 2.6719 10 1 0.99
129,164 726,756 36 4.0473 10 4 4.1378 10 2 4.0973 10 2 1.2605 10 1 0.33
468,778 2662221 64 1.7039 10 + 0 2.8312 10 + 0 1.1274 10 + 0 5.2671 10 2 21.4
1,719,320 9,873,924 113 2.4233 10 2 7.5433 10 4 2.3479 10 2 2.2151 10 2 1.06
Table 21

Adaptive refinement of Navier–Stokes 2D-3 with c = 10 7 .

#DoF(primal) #DoF(dual) 𝑀 η k η h 𝜂 J ( U ) J ( U k h ) I eff
17,800 96,600 20 8.4213 10 6 3.8557 10 1 3.8557 10 1 5.5869 10 1 0.69
63,336 350,244 36 2.3497 10 7 2.6516 10 1 2.6516 10 1 2.6719 10 1 0.99
228,568 1,285,878 64 1.7174 10 3 1.0803 10 2 9.0853 10 3 1.2438 10 1 0.07
835,942 4,749,933 113 1.3694 10 1 1.7656 10 1 3.9619 10 2 2.7110 10 2 1.46
3,008,930 17,274,966 199 3.9451 10 3 1.1779 10 2 1.5724 10 2 1.1074 10 2 1.42

A.2 Changing Number of Initial Temporal Elements with Fixed Equilibration Factor

We deliberately started with a very coarse temporal mesh, which might lead to an underestimation of the temporal error. Therefore, we performed additional tests with finer temporal meshes and c = 1 as well as c = 5 to see whether the equilibration plays a more important role when the error is better resolved in time. Overall, the behavior is similar to M = 20 (cf. Table 17 and Table 18) with most steps only refining the spatial grids. For M = 40 (cf. Table 22 and Table 23) and M = 80 (cf. Table 24 and Table 25), the last step is different with only temporal refinement for c = 1 and temporal as well as spatial refinement for c = 5 . Comparing actual errors as well as the number of DoFs, we can see that sufficient temporal refinement (roughly more than 120 intervals; see Tables 24, 25, 26, 27) is needed to get an error of around 0.01 in the fourth refinement step. So both enforced global refinement as well as equilibration with a finer initial grid led to good error reduction. However, while the number of degrees of freedom in the final step is smaller with equilibration, the total cost is higher in comparison since solving and estimating have to be performed for more time steps. Additionally, the effectivity indices for enforced global refinement seem to stabilize in the last two steps.

Table 22

Adaptive refinement of Navier–Stokes 2D-3 with c = 1 and M = 40 .

#DoF(primal) #DoF(dual) 𝑀 η k η h 𝜂 J ( U ) J ( U k h ) I eff
35,600 193,200 40 3.3813 10 8 3.8553 10 1 3.8553 10 1 5.5866 10 1 0.69
70,880 392,145 40 2.5936 10 6 2.6511 10 1 2.6511 10 1 2.6762 10 1 0.99
147,986 833,289 40 2.3636 10 4 3.8774 10 2 3.8538 10 2 1.2625 10 1 0.31
292,392 1661,022 40 3.1185 10 1 1.0598 10 1 2.0587 10 1 5.5182 10 2 3.73
513,362 2915,817 71 2.8892 10 1 2.3645 10 1 5.2464 10 2 3.4048 10 2 1.54
Table 23

Adaptive refinement of Navier–Stokes 2D-3 with c = 5 and M = 40 .

#DoF(primal) #DoF(dual) 𝑀 η k η h 𝜂 J ( U ) J ( U k h ) I eff
35,600 193,200 40 3.3813 10 8 3.8553 10 1 3.8553 10 1 5.5866 10 1 0.69
70,880 392,145 40 2.5936 10 6 2.6511 10 1 2.6511 10 1 2.6762 10 1 0.99
147,986 833,289 40 2.3636 10 4 3.8774 10 2 3.8538 10 2 1.2625 10 1 0.31
292,392 1,661,022 40 3.1185 10 1 1.0598 10 1 2.0587 10 1 5.5182 10 2 3.73
1,092,116 6,271,314 71 8.2739 10 2 6.5925 10 2 1.6814 10 2 3.4252 10 2 0.49
Table 24

Adaptive refinement of Navier–Stokes 2D-3 with c = 1 and M = 80 .

#DoF(primal) #DoF(dual) 𝑀 η k η h 𝜂 J ( U ) J ( U k h ) I eff
71,200 386,400 80 4.9319 10 7 3.8556 10 1 3.8556 10 1 5.5865 10 1 0.69
141,360 782,082 80 2.4853 10 6 2.6658 10 1 2.6658 10 1 2.6925 10 1 0.99
295,970 1,666,374 80 6.3941 10 3 8.3012 10 2 8.9406 10 2 1.2361 10 1 0.72
602,996 3,427,926 80 2.4988 10 1 1.3495 10 1 1.1493 10 1 3.1269 10 2 3.68
1,053,362 5,987,769 141 2.9715 10 3 2.0812 10 2 1.7841 10 2 1.2436 10 2 1.43
Table 25

Adaptive refinement of Navier–Stokes 2D-3 with c = 5 and M = 80 .

#DoF(primal) #DoF(dual) 𝑀 η k η h 𝜂 J ( U ) J ( U k h ) I eff
71,200 386,400 80 4.9319 10 7 3.8556 10 1 3.8556 10 1 5.5865 10 1 0.69
141,360 782,082 80 2.4853 10 6 2.6658 10 1 2.6658 10 1 2.6925 10 1 0.99
295,970 1,666,374 80 6.3941 10 3 8.3012 10 2 8.9406 10 2 1.2361 10 1 0.72
602,996 3,427,926 80 2.4988 10 1 1.3495 10 1 1.1493 10 1 3.1269 10 2 3.68
2,160,878 12,413,085 141 6.3265 10 3 6.4954 10 3 1.2822 10 2 9.8795 10 3 1.30
Table 26

Adaptive refinement of Navier–Stokes 2D-3 with c = 1 and M = 160 .

#DoF(primal) #DoF(dual) 𝑀 η k η h 𝜂 J ( U ) J ( U k h ) I eff
142,400 772,800 160 1.2458 10 7 3.8558 10 1 3.8558 10 1 5.5865 10 1 0.69
282,110 1,560,498 160 1.2875 10 6 2.6660 10 1 2.6660 10 1 2.6935 10 1 0.99
589,760 3,320,313 160 8.6094 10 5 1.4437 10 1 1.4428 10 1 1.1088 10 1 1.30
1,221,266 6,944,181 160 3.6799 10 3 2.8981 10 2 2.5301 10 2 4.8499 10 3 5.22
2,439,656 14,009,478 160 4.1653 10 3 9.1947 10 3 5.0294 10 3 1.2499 10 2 0.40
Table 27

Adaptive refinement of Navier–Stokes 2D-3 with c = 5 and M = 160 .

#DoF(primal) #DoF(dual) 𝑀 η k η h 𝜂 J ( U ) J ( U k h ) I eff
142,400 772,800 160 1.2458 10 7 3.8558 10 1 3.8558 10 1 5.5865 10 1 0.69
282,110 1,560,498 160 1.2875 10 6 2.6660 10 1 2.6660 10 1 2.6935 10 1 0.99
589,760 3,320,313 160 8.6094 10 5 1.4437 10 1 1.4428 10 1 1.1088 10 1 1.30
1,221,266 6,944,181 160 3.6799 10 3 2.8981 10 2 2.5301 10 2 4.8499 10 3 5.22
2,439,656 14,009,478 160 4.1653 10 3 9.1947 10 3 5.0294 10 3 1.2499 10 2 0.40

Acknowledgements

We thank Hendrik Fischer (Hannover/Paris), Marius Paul Bruchhäuser (Hamburg), Bernhard Endtmayer (Hannover), Johannes Lankeit (Hannover) and Ludovic Chamoin (Paris) for fruitful discussions and comments.

References

[1] M. Ainsworth and J. T. Oden, A Posteriori Error Estimation in Finite Element Analysis, Pure Appl. Math. (New York), Wiley-Interscience, New York, 2000. 10.1002/9781118032824Search in Google Scholar

[2] F. Alauzet and M. Mehrenberger, P 1 -conservative solution interpolation on unstructured triangular meshes, Internat. J. Numer. Methods Engrg. 84 (2010), no. 13, 1552–1588. 10.1002/nme.2951Search in Google Scholar

[3] S. Amstutz and T. Wick, Refresher course in maths and a project on numerical modeling done in twos, Institutionelles Repositorium der Leibniz Universität Hannover, Universität Hannover, Hannover (2022), https://doi.org/10.15488/11629. Search in Google Scholar

[4] B. Armaly, F. Durst, Jose Pereira and B. Schönung, Experimental and theoretical investigation of backward-facing step flow, J. Fluid Mech. 127 (1983), 473–496. 10.1017/S0022112083002839Search in Google Scholar

[5] D. Arndt, W. Bangerth, D. Davydov, T. Heister, L. Heltai, M. Kronbichler, M. Maier, J.-P. Pelteret, B. Turcksin and D. Wells, The deal.II finite element library: Design, features, and insights, Comput. Math. Appl. 81 (2021), 407–422. 10.1016/j.camwa.2020.02.022Search in Google Scholar

[6] D. Arndt, W. Bangerth, M. Feder, M. Fehling, R. Gassmöller, T. Heister, L. Heltai, M. Kronbichler, M. Maier, P. Munch, J.-P. Pelteret, S. Sticko, B. Turcksin and D. Wells, The deal.II library, version 9.4., J. Numer. Math. 30 (2022), no. 3, 231–246. 10.1515/jnma-2022-0054Search in Google Scholar

[7] W. Bangerth, M. Geiger and R. Rannacher, Adaptive Galerkin finite element methods for the wave equation, Comput. Methods Appl. Math. 10 (2010), no. 1, 3–48. 10.2478/cmam-2010-0001Search in Google Scholar

[8] W. Bangerth and R. Rannacher, Adaptive Finite Element Methods for Differential Equations, Lectures in Math. ETH Zürich, Birkhäuser, Basel, 2003. 10.1007/978-3-0348-7605-6Search in Google Scholar

[9] E. Bänsch, F. Karakatsani and C. G. Makridakis, A posteriori error estimates for fully discrete schemes for the time dependent Stokes problem, Calcolo 55 (2018), no. 2, Paper No. 19. 10.1007/s10092-018-0259-2Search in Google Scholar

[10] S. Basava, K. Mang, M. Walloth, T. Wick and W. Wollner, Adaptive and pressure-robust discretization of incompressible pressure-driven phase-field fracture, Non-Standard Discretisation Methods in Solid Mechanics, Lect. Notes Appl. Comput. Mech. 98, Springer, Cham, (2022), 191–215. 10.1007/978-3-030-92672-4_8Search in Google Scholar

[11] S. R. Basava and W. Wollner, Gradient robust mixed methods for nearly incompressible elasticity, J. Sci. Comput. 95 (2023), Article No. 93. 10.1007/s10915-023-02227-0Search in Google Scholar

[12] M. Bause, M. P. Bruchhäuser and U. Köcher, Flexible goal-oriented adaptivity for higher-order space-time discretizations of transport problems with coupled flow, Comput. Math. Appl. 91 (2021), 17–35. 10.1016/j.camwa.2020.08.028Search in Google Scholar

[13] R. Becker and R. Rannacher, A feed-back approach to error control in finite element methods: Basic analysis and examples, East-West J. Numer. Math. 4 (1996), no. 4, 237–264. Search in Google Scholar

[14] R. Becker and R. Rannacher, An optimal control approach to a posteriori error estimation in finite element methods, Acta Numer. 10 (2001), 1–102. 10.1017/S0962492901000010Search in Google Scholar

[15] M. Besier and R. Rannacher, Goal-oriented space-time adaptivity in the finite element Galerkin method for the computation of nonstationary incompressible flow, Internat. J. Numer. Methods Fluids 70 (2012), no. 9, 1139–1166. 10.1002/fld.2735Search in Google Scholar

[16] M. Besier and W. Wollner, On the pressure approximation in nonstationary incompressible flow simulations on dynamically varying spatial meshes, Internat. J. Numer. Methods Fluids 69 (2012), no. 6, 1045–1064. 10.1002/fld.2625Search in Google Scholar

[17] G. Biswas, M. Breuer and F. Durst, Backward-facing step flows for various expansion ratios at low and moderate Reynolds numbers, J. Fluids Eng.-Trans. ASME 126 (2004), 362–374. 10.1115/1.1760532Search in Google Scholar

[18] M. Braack and A. Ern, A posteriori control of modeling errors and discretization errors, Multiscale Model. Simul. 1 (2003), no. 2, 221–238. 10.1137/S1540345902410482Search in Google Scholar

[19] M. Braack, J. Lang and N. Taschenberger, Stabilized finite elements for transient flow problems on varying spatial meshes, Comput. Methods Appl. Mech. Engrg. 253 (2013), 106–116. 10.1016/j.cma.2012.08.006Search in Google Scholar

[20] M. Braack and T. Richter, Solutions of 3D Navier–Stokes benchmark problems with adaptive finite elements, Comput. Fluids 35 (2006), no. 4, 372–392. 10.1016/j.compfluid.2005.02.001Search in Google Scholar

[21] M. P. Bruchhäuser, U. Köcher and M. Bause, On the implementation of an adaptive multirate framework for coupled transport and flow, J. Sci. Comput. 93 (2022), no. 3, Paper No. 59. 10.1007/s10915-022-02026-zSearch in Google Scholar

[22] R. Dautray and J.-L. Lions, Mathematical Analysis and Numerical Methods for Science and Technology. Vol. 5, Springer, Berlin, 2000. Search in Google Scholar

[23] J. D. De Basabe, M. K. Sen and M. F. Wheeler, The interior penalty discontinuous Galerkin method for elastic wave propagation: Grid dispersion, Geophys. J. Internat. 175 (2008), no. 1, 83–93. 10.1111/j.1365-246X.2008.03915.xSearch in Google Scholar

[24] B. Endtmayer, U. Langer and T. Wick, Multigoal-oriented error estimates for non-linear problems, J. Numer. Math. 27 (2019), no. 4, 215–236. 10.1515/jnma-2018-0038Search in Google Scholar

[25] B. Endtmayer, U. Langer and T. Wick, Two-side a posteriori error estimates for the dual-weighted residual method, SIAM J. Sci. Comput. 42 (2020), no. 1, A371–A394. 10.1137/18M1227275Search in Google Scholar

[26] L. Failer and T. Wick, Adaptive time-step control for nonlinear fluid-structure interaction, J. Comput. Phys. 366 (2018), 448–477. 10.1016/j.jcp.2018.04.021Search in Google Scholar

[27] G. P. Galdi, An Introduction to the Mathematical Theory of the Navier–Stokes Equations: Steady-State Problems, Springer Monogr. Math., Springer, New York, 2011. 10.1007/978-0-387-09620-9Search in Google Scholar

[28] V. Girault and P.-A. Raviart, Finite Element Methods for Navier–Stokes Equations, Springer Ser. Comput. Math. 5, Springer, Berlin, 1986. 10.1007/978-3-642-61623-5Search in Google Scholar

[29] R. Glowinski, Finite element methods for incompressible viscous flow, Numerical Methods for Fluids (Part 3), Handb. Numer. Anal. 9, North-Holland, Amsterdam (2003), 3–1176. 10.1016/S1570-8659(03)09003-3Search in Google Scholar

[30] J. G. Heywood, R. Rannacher and S. Turek, Artificial boundaries and flux and pressure conditions for the incompressible Navier–Stokes equations, Internat. J. Numer. Methods Fluids 22 (1996), no. 5, 325–352. 10.1002/(SICI)1097-0363(19960315)22:5<325::AID-FLD307>3.0.CO;2-YSearch in Google Scholar

[31] J. Hoffman, Efficient computation of mean drag for the subcritical flow past a circular cylinder using general Galerkin G2, Internat. J. Numer. Methods Fluids 59 (2009), no. 11, 1241–1258. 10.1002/fld.1865Search in Google Scholar

[32] J. Hoffman and C. Johnson, Adaptive finite element methods for incompressible fluid flow, Error Estimation and Adaptive Discretization Methods in Computational Fluid Dynamics, Lect. Notes Comput. Sci. Eng. 25, Springer, Berlin (2003), 97–157. 10.1007/978-3-662-05189-4_3Search in Google Scholar

[33] P. Hood and C. Taylor, Navier–Stokes equations using mixed interpolation, Finite Element Methods in Flow Problems, John Wiley, New York (1974), 121–132. Search in Google Scholar

[34] V. John, Reference values for drag and lift of a two-dimensional time-dependent flow around a cylinder, Internat. J. Numer. Methods Fluids 44 (2004), no. 7, 777–788. 10.1002/fld.679Search in Google Scholar

[35] V. John, A. Linke, C. Merdon, M. Neilan and L. G. Rebholz, On the divergence constraint in mixed finite element methods for incompressible flows, SIAM Rev. 59 (2017), no. 3, 492–544. 10.1137/15M1047696Search in Google Scholar

[36] U. Langer and A. Schafelner, Adaptive space-time finite element methods for non-autonomous parabolic problems with distributional sources, Comput. Methods Appl. Math. 20 (2020), no. 4, 677–693. 10.1515/cmam-2020-0042Search in Google Scholar

[37] U. Langer and A. Schafelner, Adaptive space-time finite element methods for parabolic optimal control problems, J. Numer. Math. 30 (2022), no. 4, 247–266. 10.1515/jnma-2021-0059Search in Google Scholar

[38] U. Langer and O. Steinbach, Space-Time Methods—Applications to Partial Differential Equations, Radon Ser. Comput. Appl. Math. 25, De Gruyter, Berlin, 2019. 10.1515/9783110548488Search in Google Scholar

[39] P. L. Lederer, A. Linke, C. Merdon and J. Schöberl, Divergence-free reconstruction operators for pressure-robust Stokes discretizations with continuous pressure finite elements, SIAM J. Numer. Anal. 55 (2017), no. 3, 1291–1314. 10.1137/16M1089964Search in Google Scholar

[40] A. Linke, On the role of the Helmholtz decomposition in mixed methods for incompressible flows and a new variational crime, Comput. Methods Appl. Mech. Engrg. 268 (2014), 782–800. 10.1016/j.cma.2013.10.011Search in Google Scholar

[41] A. Linke, G. Matthies and L. Tobiska, Robust arbitrary order mixed finite element methods for the incompressible Stokes equations with pressure independent velocity errors, ESAIM Math. Model. Numer. Anal. 50 (2016), no. 1, 289–309. 10.1051/m2an/2015044Search in Google Scholar

[42] D. Meidner and T. Richter, Goal-oriented error estimation for the fractional step theta scheme, Comput. Methods Appl. Math. 14 (2014), no. 2, 203–230. 10.1515/cmam-2014-0002Search in Google Scholar

[43] D. Meidner and T. Richter, A posteriori error estimation for the fractional step theta discretization of the incompressible Navier–Stokes equations, Comput. Methods Appl. Mech. Engrg. 288 (2015), 45–59. 10.1016/j.cma.2014.11.031Search in Google Scholar

[44] D. Meidner and B. Vexler, Adaptive space-time finite element methods for parabolic optimization problems, SIAM J. Control Optim. 46 (2007), no. 1, 116–142. 10.1137/060648994Search in Google Scholar

[45] S. Mittal, A. Ratner, D. Hastreiter and T. E. Tezduyar, Space-time finite element computation of incompressible flows with emphasis on flow involving oscillating cylinders, Internat. Video J. Engrg. Res. 1 (1991), 83–86. Search in Google Scholar

[46] J. T. Oden, Adaptive multiscale predictive modelling, Acta Numer. 27 (2018), 353–450. 10.1017/S096249291800003XSearch in Google Scholar

[47] R. Picard and D. McGhee, Partial Differential Equations, De Gruyter Exp. Math. 55, Walter de Gruyter, Berlin, 2011. 10.1515/9783110250275Search in Google Scholar

[48] A. Rademacher, Adaptive finite element methods for nonlinear hyperbolic problems of second order, PhD thesis, Technische Universität Dortmund, 2009. Search in Google Scholar

[49] R. Rannacher, Finite element methods for the incompressible Navier–Stokes equations, Fundamental Directions in Mathematical Fluid Mechanics, Adv. Math. Fluid Mech., Birkhäuser, Basel (2000), 191–293. 10.1007/978-3-0348-8424-2_6Search in Google Scholar

[50] R. Rannacher, Probleme der Kontinuumsmechanik und ihre numerische Behandlung, Heidelberg University, Heidelberg, 2017. Search in Google Scholar

[51] T. Richter and T. Wick, Variational localizations of the dual weighted residual estimator, J. Comput. Appl. Math. 279 (2015), 192–208. 10.1016/j.cam.2014.11.008Search in Google Scholar

[52] A. Schafelner, Space-time finite element methods, PhD thesis, Johannes Kepler University, Linz, 2021. Search in Google Scholar

[53] M. Schäfer, S. Turek, F. Durst, E. Krause and R. Rannacher, Benchmark computations of laminar flow around a cylinder, Flow Simulation with High-Performance Computers II, Vieweg & Teubner, Wiesbaden (1996), 547–566. 10.1007/978-3-322-89849-4_39Search in Google Scholar

[54] M. Schmich, Adaptive finite element methods for computing nonstationary incompressible flows, Doctoral Thesis, Heidelberg University, 2009. Search in Google Scholar

[55] M. Schmich and B. Vexler, Adaptivity with dynamic meshes for space-time finite element discretizations of parabolic equations, SIAM J. Sci. Comput. 30 (2007/08), no. 1, 369–393. 10.1137/060670468Search in Google Scholar

[56] R. Stenberg, Error analysis of some finite element methods for the Stokes problem, Math. Comp. 54 (1990), no. 190, 495–508. 10.1090/S0025-5718-1990-1010601-XSearch in Google Scholar

[57] R. Temam, Navier–Stokes Equations: Theory and Numerical Analysis, AMS Chelsea, Providence, 2001. 10.1090/chel/343Search in Google Scholar

[58] J. P. Thiele and T. Wick, Variational partition-of-unity localizations of space-time dual weighted residual estimators for parabolic problems, preprint (2022), https://arxiv.org/abs/2207.04764. Search in Google Scholar

[59] S. Turek, Efficient Solvers for Incompressible Flow Problems: An Algorithmic and Computational Approach, Lect. Notes Comput. Sci. Eng. 6, Springer, Berlin, 1999. 10.1007/978-3-642-58393-3Search in Google Scholar

[60] T. Wick, Solving monolithic fluid-structure interaction problems in arbitrary Lagrangian Eulerian coordinates with the deal.II library, Arch. Numer. Softw. 1 (2013), 10.11588/ans.2013.1.10305. 10.11588/ans.2013.1.10305Search in Google Scholar

Received: 2022-10-06
Revised: 2023-02-20
Accepted: 2023-04-26
Published Online: 2023-05-31
Published in Print: 2024-01-01

© 2023 Walter de Gruyter GmbH, Berlin/Boston

Downloaded on 23.9.2025 from https://www.degruyterbrill.com/document/doi/10.1515/cmam-2022-0200/html
Scroll to top button