1.
Introduction
Performing accurate simulations of complex fluid flows that match real-world observations or experiments typically requires highly precise knowledge of the initial data. However, such data is often known in very sparsely-distributed locations, which is the case in, e.g., weather observation, ocean monitoring, etc. Thus, accurate, deterministic simulations based on initial data are often impractical. Data assimilation is a collection of methods that works around this difficulty by incorporating incoming data into the simulation to increase accuracy, hence data assimilation techniques are highly desirable to incorporate into simulations. However, the underlying physical equations often suffer from stability issues which can reduce the accuracy gained by using data assimilation. While there are many ways to stabilize numerical simulations, it is far from obvious how to adapt data assimilation techniques to combine them with cutting-edge stabilization methods. Therefore it becomes worthwhile to seek new ways to incorporate data assimilation into stabilized schemes. In this article, we propose and analyze a new approach to this problem which combines continuous data assimilation with velocity-vorticity stabilization.
Since Kalman's seminal paper [29] in 1960, a wide variety of data assimilation algorithms have arisen (see, e.g., [14,32]). In [5], Azouani, Olson, and Titi proposed a new algorithm known as continuous data assimilation (CDA), also referred to as the AOT algorithm. Their approach revived the so-called "nudging" methods of the 1970's (see, e.g., [4,26]), but with the addition of a spatial interpolation operator. This seemingly minor change had profound impacts, and the authors of [5] were able to prove that using only sparse observations, the CDA algorithm applied to the 2D Navier-Stokes equations converges to the correct solution exponentially fast in time, independent of the choice initial data. This stimulated a large amount of recent research on the CDA algorithm; see, e.g., [3,6,7,10,11,13,17,18,19,20,21,22,27,31,30,35,40,41] and the references therein. The recent paper [15] showed that CDA can be effectively used for weather prediction, showing that it can indeed be a powerful tool on practical large scale problems. Convergence of discretizations of CDA models was studied in [30,41,27,21], and found results similar to those at the continuous level. Our interest in the CDA algorithm arises from its adaptability to a wide range of nonlinear problems, as well as its small computational cost and straight-forward implementation. These qualities make it an ideal candidate for combining data assimilation with stabilization techniques; in particular, with the recently developed velocity-vorticity stabilization, described below.
Flows of incompressible, viscous Newtonian fluids are modeled by the Navier-Stokes equations (NSE), which take the form
together with suitable boundary and initial conditions. Here, u denotes a velocity vector field, p is pressure, f is external (given) force, and ν>0 represents the kinematic viscosity which is inversely proportional to the Reynolds number. Solving the NSE is important in many applications, however it is well known that doing so can be quite difficult, especially for small ν. Many different tools have been used for more accurate numerical simulations of the NSE, for example using NSE formulations tailored to particular application problems [23,12,37,33] or discretization and stabilization methods [39,28], and more recently using observed data to improve simulation [5,30,43,44,8].
We consider in this paper discretizations of a continuous data assimilation (CDA) enhancement applied to the following velocity-vorticity (VV) formulation of the 2D NSE:
Here, ω represents the (scalar) vorticity, P:=p+12|u|2 is the Bernoulli pressure, and rot is the 2D curl operation: rot(f1f2):=∂f2∂x−∂f1∂y. In the NSE, the velocity and vorticity are coupled via the relationship ω=rotu (or equivalently, the Biot-Savart Law). However, the VV formulation typically does not enforce this relationship, so u and ω are only coupled via the evolution equations in (2), and the relationship ω=rotu is recovered a posteriori, so that at the continuous level, (2) is formally equivalent to (1). However, in practice, discretizations of VV can behave quite differently from typical discretizations of NSE, providing better stability as well as accuracy (especially for vorticity) for vortex dominated or strongly rotating flows, see [36,37,34,2] and references therein. A very interesting property of (2) was recently shown in [25], where it was proven that the system (2) when discretized with standard finite elements and a decoupling backward Euler or BDF2 temporal discretization was unconditionally long-time stable in both L2 and H1 norms for both velocity and vorticity; no such analogous result is known for velocity-pressure discretizations/schemes. Hence the scheme itself is stabilizing, even though it is still formally consistent with the NSE. The recent work in [2] showed that these unconditionally long-time stable schemes also provide optimal vorticity accuracy, yielding a vorticity solution that is one full order of spatial accuracy better than for an analogous velocity-pressure scheme.
We consider herein CDA applied to (2), which yields a model of the form
where IH is an appropriate interpolation operator, IH(u) and IH(ω) are assumed known from measurement data (i.e. u and ω are known at some points in space), and μ1,μ2≥0 are nudging parameters. If μ2=0, then vorticity is not nudged and IH(ω) need not be assumed known. Due to the success of (2) in recent papers [25,2,36] and that of CDA in the works mentioned above, combining these ideas and studying (2) is a natural next step to see whether CDA will provide optimal long-time accuracy for the VV schemes already known to be unconditionally long-time stable. Herein, we do find that CDA provides convergence of (3), with any initial condition, to the true NSE solution (up to optimal discretization error) and moreover that CDA preserves the long-time stability.
This paper is organized as follows. In Section 2, we introduce the necessary notation and preliminaries needed in the analysis. In Section 3, we propose and analyze a fully discrete scheme for (3), and show that for nudging velocity and vorticity together and nudging just velocity, algorithms are long-time stable in L2 and H1 norms and long-time optimally accurate in L2 velocity and vorticity (under the usual CDA assumptions on the coarse mesh and nudging parameter). In Section 4, we illustrate the theory with numerical tests, and finally draw conclusions in section 5.
2.
Notation and preliminaries
We now provide notation and mathematical preliminaries to allow for a smooth analysis to follow. We consider the domain Ω⊂R2 to be the 2π-periodic box, with the L2(Ω) norm and inner product denoted by ‖⋅‖ and (⋅,⋅) respectively, while all other norms will be appropriately labeled.
For simplicity, we use herein periodic boundary conditions for velocity and vorticity. Extension to full nonhomogeneous Dirichlet conditions can be performed by following analysis in [34], although for no-slip velocity together with the more physically consistent natural vorticity boundary condition studied in [36,38] more work would be needed to handle the boundary integrals. We denote the natural corresponding function spaces for velocity, pressure, and vorticity by
In X (and W), we have the Poincaré inequality: there exists a constant CP depending only on Ω such that for any ϕ∈X (or W),
We define the skew-symmetric trilinear operator b∗:X×W×W→R to use for the nonlinear term in the vorticity equation, by
The following lemma is proven in [30], and is useful in our analysis.
Lemma 2.1. Suppose constants r and B satisfy r>1, B≥0. Then if the sequence of real numbers {an} satisfies
we have that
2.1. Discretization preliminaries
Denote by τh a regular, conforming triangulation of the domain Ω, and let Xh⊂X, Qh⊂Q be velocity-pressure spaces that satisfy the inf-sup condition. We will assume the use of Xh=X∩Pk(τh) and Qh=Q∩Pk−1(τh) Taylor-Hood or Scott-Vogelius elements (on appropriate meshes and/or polynomial degrees, see [24] and references therein). The discrete vorticity space is defined as Wh:=W∩Pk(τh). Define the discretely divergence free subspace by
We will assume the mesh is sufficiently regular so that the inverse inequality holds in Xh: There exists a constant C such that
The discrete Laplacian operator is defined as: For ϕ∈H1(Ω)2, Δhϕ∈Xh satisfies
The definition for Δh is written the same way when applied in Wh, since this is simply the above definition restricted to a single component.
The discrete Stokes operator Ah is defined as: For ϕ∈H1(Ω)2, find Ahϕ∈Vh such that for all vh∈Vh,
By the definition of discrete Laplace and Stokes operators, we have the Poincaré inequalities
We recall the following discrete Agmon inequalities and discrete Lp bounds [25]:
We note that all bounds above for Xh trivially hold in Wh, since Wh functions can be considered as components of functions in Xh.
A function space for measurement data interpolation is also needed. Hence we require another regular conforming mesh τH, and define XH=Pr(τH)2 and WH=Pr(τH) for some polynomial degree r. We require that the coarse mesh interpolation operator IH used for data assimilation satisfies the following bounds: for any w∈H1(Ω)d,
These are key properties for the interpolation operator that allow for both mathematical theory as well as providing guidance on how small H should be (i.e. how many measurement points are needed). We note the same IH operator is used for vector functions and scalar functions, with it being applied component-wise for vector functions.
3.
Analysis of a CDA-VV scheme
We consider now a discretization of (3) that uses a finite element spatial discretization and backward Euler temporal discretization. The backward Euler discretization is chosen only for simplicity of analysis; all results extend to the analogous BDF2 scheme following analysis in [2,25]. One difference of our scheme below compared to other discretizations of CDA is that IH is also applied to the test functions in the nudging terms. This was first proposed by the authors in [41], and allows for a simpler stability analysis as well as to the use of special types of efficient interpolation operators.
Algorithm 3.1. Given v0h∈Vh and w0h∈Wh, find (vn+1h,wn+1h,Pn+1h) ∈ (Xh,Wh,Qh) for n=0,1,2,..., satisfying
for all (χh,ψh,rh)∈(Xh,Wh,Qh), where IH(un+1), IH(rotun+1) are assumed known for all n≥0.
We begin our analysis with long-time stability estimates, followed by long-time accuracy.
3.1. Stability analysis of Algorithm 3.1
In this subsection, we prove that Algorithm 3.1 is unconditionally long-time L2 and H1 stable for both velocity and vorticity. This property was proven for the scheme without nudging in [25], and so these results show that CDA preserves this important property that is (seemingly) unique to VV schemes of this form.
Lemma 3.2 (L2 stability of velocity and vorticity). Let f∈L∞(0,∞;L2) and u∈L∞(0,∞;H1). Then, for any Δt>0, any integer n>0, and nudging parameters μ1,μ2≥0, velocity and vorticity solutions to Algorithm 3.1 satisfy
where α=1+νC−2PΔt.
Proof. Begin by choosing χh=2Δtvn+1h in (13), which vanishes the nonlinear and pressure terms, and leaves
The first right hand side term is bounded using the dual norm and Young's inequality via
and for the interpolation term, we use Cauchy-Schwarz, the interpolation property (12) and Young's inequality to get
Combining the above estimates and dropping ‖vn+1h−vnh‖2 and ‖IH(vn+1h)‖2 from the left hand side produces the bound
and thanks to the Poincaré inequality, we obtain
Defining α=1+C−2PΔtν>1, and then applying Lemma 2.1 reveals the L2 stability bound (16) for the velocity solution of Algorithm 3.1.
Applying similar analysis to the above will produce the stated L2 vorticity bound.
Lemma 3.3 (H1 stability of velocity and vorticity). Let f∈L∞(0,∞;H1) and u∈L∞(0,∞;H1). Then, for any Δt>0, any integer n>0, and nudging parameters μ1,μ2≥0, velocity and vorticity solutions to Algorithm 3.1 satisfy
where α=1+νC−2PΔt.
Proof. After testing the velocity equation (13) with χh=2ΔtAhvn+1h, we obtain
We now bound the right hand side terms. First, the forcing term is bounded by Cauchy-Schwarz and Young's inequalities via
Then, for the nonlinear terms, we again apply Hölder, discrete Agmon (9) and generalized Young inequalities, and the result of Lemma 3.2 to get
Lastly, the interpolation term is bounded using Cauchy-Schwarz and interpolation property 12, followed by Young's inequality and the result of Lemma 3.2 to obtain
Combining all these bounds for right hand side terms and dropping nonnegative term ‖∇(vn+1h−vnh)‖2 on left hand side give us that
By the Poincaré inequality (7), we now get
where α=1+νC−2PΔt. Finally, we apply Lemma 2.1 and reveal (18).
For the vorticity estimate, choose ψh=2ΔtΔhwn+1h in (15) to get
From here, the proof follows the same strategy as the H1 velocity proof above, except the nonlinear term is handled slightly differently. We use the discrete Agmon inequality (9), the discrete Sobolev inequality (10), the result of Lemma 3.1, the H1 stability bound for vorticity (18) proven above, and the generalized Young's inequality, as follows.
Now proceeding as in the velocity H1 bound will produce the H1 vorticity stability bound (19).
3.2. Long-time accuracy of Algorithm 3.1
We now consider the difference between the solutions of (13) - (15) to the NSE solution. We will show that the algorithm solution converges to the true solution, up to an optimal O(Δt+hk+1) discretization error, independent of the initial condition, provided a restriction on the coarse mesh width and nudging parameters. We will give two results, the first for μ2>0 and the second for μ2=0; while they both provide optimal long-time accuracy, when μ2>0 the convergence to the true solution occurs more rapidly in time.
In our theory below for long-time accuracy of Algorithm 3.1, we assume the use of Scott-Vogelius elements. This is done for simplicity, as for non-divergence-free elements like Taylor-Hood elements, similar optimal results can be obtained (although with some additional terms and different constants) but require more technical details; see, e.g., [21]. We define the following projection operator, PV, which will be used in the following accuracy analysis: Given ϕ∈H1(Ω) with ∇⋅ϕ=0, PV(ϕ)∈Vh satisfies (∇(ϕ−PV(ϕ)),∇vh)=0 for all vh∈Vh. We also define PW:H1(Ω)→Wh in the same way, with Vh replaced by Wh. The operators PV and PW are known to have optimal approximation properties on the divergence free subspace of X and on W, respectively, in both L2 and H1 norms, provided some commonly assumed properties of the domain [9,42].
Theorem 3.4 (Long-time L2 accuracy of Algorithm 3.1 with μ1>0 and μ2>0). Let true solutions u∈L∞(0,∞;Hk+2(Ω)), p∈L∞(0,∞;Hk(Ω)) where k≥1 and ut,utt∈L∞(0,∞;H1). Then, assume that time step Δt is sufficiently small, and that μ1 and μ2 satisfy
where H is chosen so that this inequality holds. Then, for any time tn, n=0,1,2,..., we have for solutions of Algorithm 3.1 using Scott-Vogelius elements,
where
and λ=min{μ14+νC−2P4,μ24+νC−2P4} with C independent of Δt, h and H.
Proof. The true NSE solution satisfies the VV system
where un is the velocity at time tn, Pn the Bernoulli pressure, ωn:=rotun, and t∗,t∗∗∈[tn,tn+1]. Note that by Taylor expansion, we can write ωn−ωn+1=−Δtωt(s∗) where s∗∈[tn,tn+1].
The difference equations are obtained by subtracting the solutions to Algorithm 3.1 from (22)-(24) after testing them with test functions from χh∈Xh, rh∈Qh and ψh∈Wh, respectively. We define the differences between velocity and vorticity as env:=vnh−un and enw:=wnh−ωn, respectively. Next, we will decompose the error into a term that lies in the discrete space Vh and one outside. To do so, add and subtract the discrete Stokes projection of un, denoted PVun, to env and let ηnv:=PVun−un, ϕnh,v:=vnh−PVun. Then env=ϕnh,v+ηnv and ϕnh,v∈Vh. In a similar manner, by taking the H10 projection of rotun into Wh denoted by PWωn, we obtain enw=ϕnh,w+ηnw with ϕnh,w∈Vh.
For velocity, since (∇ηn+1v,∇ϕn+1h,v)=0, the difference equation becomes
and similarly for vorticity, we have
where in (25) we have added and subtracted ϕn+1h,v to write it in the form found above using
and similarly for (26).
Next, we bound the terms on right hand side of difference equations, starting with the velocity difference equation (25). The first three right hand side terms are bounded using Cauchy-Schwarz and Young's inequalities, via
where s∗∈[tn,tn+1].
For nonlinear terms in (25), first we add and subtract en+1w in the first component, and un+1 in second component to get
The all resulting terms are bounded by Hölder's and Young's inequalities to obtain
Then, for the last nonlinear term, we apply Hölder's, Poincaré's and Young's inequalities and get
Next, the first interpolation term on the right hand side of (25) will be bounded with Cauchy-Schwarz inequality and (11) to obtain
For the second interpolation term, we apply inequality (12), which yields
Finally, the last interpolation term will be bounded using Cauchy-Schwarz, (12), and Young's inequality to get the bound
We now move on to the vorticity difference equation, (26). All the linear terms are majorized in a similar manner as in the velocity case, and so we show below the bounds only for the nonlinear terms. Due to the use of Scott-Vogelius elements, the skew-symmetric form reduces to the usual convective form, so b∗(u,v,w)=(u⋅∇v,w), with ‖∇⋅u‖=0. To bound the first nonlinear term on the right hand side of (26), we begin by breaking up the velocity error term, then apply Hölder's and Young's inequalities, yielding
For the second nonlinear term, we use Hölder's, Póincare's and Young's inequalities, which gives
For the last nonlinear term, we begin by breaking up the velocity error term, then apply Hölder's and Young's inequalities to get
Replacing the right hand sides of (25) and (26) with the computed bounds and dropping nonnegative terms with ‖ϕn+1h,v−ϕnh,v‖2 yields the bound
Using the assumptions on H and the nudging parameters, the time step restriction, and smoothness of the true solution, this reduces to
Now define
Using this in the inequality after applying Poincare's inequality and multiplying each side by 2Δt, we get
Then, with R:=(μ−11Δt2+ν−1Δt2+ν−1h2k+2+μ1h2k+2+μ2h2k+2) and λ:=min{λ1,λ2}, we obtain the bound
By Lemma 2.1, this implies
Lastly, applying triangle inequality completes the proof.
Theorem 3.5 (Long-time L2 accuracy of Algorithm 3.1 with μ1>0,μ2=0). Let true solution u∈L∞(0,∞;Hk+2(Ω)), p∈L∞(0,∞;Hk(Ω)) where k≥1 and ut,utt,∈L∞(0,∞;H1). Then, assume that time step Δt is sufficiently small, μ2=0, and that μ1 satisfies
where H is chosen so that this inequality holds. Then for any time tn, n=0,1,2,..., solutions of of Algorithm 3.1 using Scott-Vogelius elements satisfy
where
and λ=min{μ14+C−2Pν4,C−2Pν4} with C independent of Δt, h and H.
Remark 1. Algorithm 3.1 converges to the true solutions up to optimal discretization error in both cases μ1,μ2>0 and μ1>0,μ2=0. The key difference between two cases is that when μ2=0, the convergence in time to reach optimal accuracy is much slower since λ does not scale with the nudging parameters. This phenomena is illustrated in our numerical tests.
Proof. We follow the same steps with the proof of Theorem 3.4. The difference equation for velocity is already the same with (25), and just two nonlinear terms in the velocity difference equation are bounded with differently in this case. By Hölder, Poincaré and Young's inequalities, we get the bounds
All terms on the right hand side of vorticity difference equation for Theorem 3.4 are bounded identically. Proceeding as in the previous proof, we arrive at
Provided Δt is sufficiently small and the restriction
holds, then applying Poincaré inequality to the terms on left hand side and using
and assumptions on the true solution, we obtain
From here, the proof is finished in the same way as the previous theorem.
3.3. Second order temporal discretization
We now present results for a second order analogue of the first order algorithm studied above.
Algorithm 3.6. Find (vn+1h,wn+1h,qn+1h)∈(Xh,Wh,Qh) for n=0,1,2,..., satisfying
for all (χh,ψh,rh)∈(Xh,Wh,Qh), with v0∈X and IH(un+1), IH(rotun+1) given.
Stability and convergence results follow in the same manner as the first order scheme results above, using G-stability theory as in [1,2,30].
Theorem 3.7 (Long-time stability and accuracy of Algorithm 3.6 with μ1>0 and μ2>0). For any time step Δt>0, and any time tn, n=0,1,2,..., we have that solutions of Algorithm 3.6 satisfy
with C independent of n, Δt, h, H.
Furthermore, if we suppose the true solution u∈L∞(0,∞;Hk+2(Ω)), p∈L∞(0,∞;Hk(Ω)) where k≥1 and ut,uttt,∈L∞(0,∞;H1), that time step Δt is sufficiently small, Scott-Vogelius elements are used, and that μ1 and μ2 satisfy C(u)≤μ1,μ2≤CνH2, we have the bound
where
and λ=min{μ14+νC−2P4,μ24+νC−2P4} with C independent of Δt, h and H.
4.
Numerical experiments
In this section, we illustrate the above theory with two numerical tests, both using Algorithm 3.6. Our first test is for convergence rates on a problem with analytical solution, and the second test is for flow past a flat plate. For both tests, we report results only for (P2,P1) Taylor-Hood elements for velocity and pressure, and P2 for vorticity; however we also tried Scott-Vogelius elements on barycenter refined meshes that produced similar numbers of degrees of freedom, and results were very similar to those of Taylor-Hood. The coarse velocity and vorticity spaces XH and WH are defined to be piecewise constants on the same mesh used for the computations. The interpolation operator IH was taken to be the L2 projection operator onto XH (or WH), which is known to satisfy (11)-(12) [16].
4.1. Experiment 1: Convergence rate test
For our first test, we investigate the theory above for Algorithm 3.6. Here we use the analytic solution
on the unit square domain Ω=(0,1)2 with kinematic viscosity ν=1.0, and use the NSE to determine f and boundary conditions. We take the final time T=1, and choose initials conditions for Algorithm 3.6's velocity and vorticity to be 0. For the discretization, (P2,P1) Taylor-Hood elements are used for velocity and pressure, P2 for vorticity, and a time step size of Δt=0.001. From Section 3, we expect third order spatial convergence rate in the L2 norm for large enough times. Results are presented below for two cases, μ2>0 and μ2=0.
4.1.1. Results for μ1>0 and μ2>0
To test this case, we first calculated spatial convergence rates at the final time T=1 with the L2 error, using successively refined uniform meshes and μ1=μ2=100. Errors and rates are shown in table 1, and show clear third order spatial convergence of both velocity and vorticity. Deterioration of the rates for the smallest h is expected since the time step Δt is fixed while the spatial mesh width decreases.
We next consider convergence to the true solution exponentially in time (up to discretization error). Here we take h=1/32, and compute solutions using μ1=μ2=μ, with μ=1, 10, 100, 1000. Results are shown in figure 1, as L2 error versus time for velocity and vorticity. We observe exponential convergence in time of both velocity and vorticity, up to about 10−5, which is consistent with the choices of h and Δt and the accuracy of the method. We note that as μ is increased, convergence is faster in time, which is consistent with our theory for the case of μ1>0 and μ2>0.
4.1.2. Results for μ1>0 and μ2=0
We now consider the same tests as above, but with μ2=0. This is an important case, since it is not always practical to obtain accurate vorticity measurement data. Just as in the first case, we first calculated spatial convergence rates at the final time T=1 for the L2 error, on the same successively refined uniform meshes, but now with μ1=100 and μ2=0. Errors and rates are shown in table 2, and show clear third order spatial convergence of both velocity and vorticity. Deterioration of the rates for the smallest h is expected since the time step Δt was fixed at 0.001, although the vorticity errors are slightly worse than for the case of μ2=100 shown in table 1, and the deterioration of the rates occurs a bit earlier. Hence we observe essentially the same velocity errors and rates compared to the case of μ2=100, and slightly worse vorticity error but still with optimal L2 accuracy.
To test exponential convergence in time for the case of μ2=0, we again take h=1/32, and compute solutions using μ1=1, 10, 100, 1000. Results are shown in figure 2, as L2 error versus time for velocity and vorticity. Although we do observe exponential convergence in time of both velocity and vorticity, up to about 10−5 which is the same accuracy reached when μ2=100 above. An important difference here compared to when μ2=100 is that the convergence of vorticity to the true solution is independent of μ1, and the convergence of velocity is slower for larger choices of μ1. This reduced dependence of the convergence on the nudging parameters when μ2=0 is consistent with our theory. Hence without vorticity nudging, long-time optimal accuracy is still achieved, but it takes longer in time to get there.
4.2. Experiment 2: Flow past a normal flat plate
To test Algorithm 3.6 on a more practical problem, we consider flow past a flat plate with Re=50. The domain of this problem is [−7,20]×[−10,10] rectangular channel with a 0.125×1 plate fixed ten units into the channel from the left, vertically centered. The inflow velocity is uin=⟨0,1⟩, no-slip velocity and the corresponding natural vorticity boundary condition from [36] are used on the walls and plate, and homogeneous Neumann conditions are enforced weakly at the outflow. A setup for the domain is shown in figure 3. There is no external forcing applied, f=0. The viscosity is taken to be ν=1/50 which is inversely proportional to Re, based on the height of the plate. The end time for the test is T=80. A DNS was run until for 160 time units (from t = -80 to t = 80), and for t>0 measurement data for the VV-DA simulation was sampled from the DNS.
We computed solutions using a Delaunay generated triangular meshes that provided 27,373 total degree of freedom with (P2,P1,P2) velocity-pressure-vorticity elements, and time step Δt=0.02. We first compared convergence in time to the DNS solution, for two cases: μ1=μ2>0 and μ1>0, μ2=0. Plots of L2 velocity and vorticity error for both of these cases are shown in figure 4, with varying nudging parameters. There is a clear advantage seen in the plots for the simulations with μ2>0: when vorticity is nudged in addition to velocity, convergence to the true solution is much faster in time. The convergence when μ2=0 appears to still be occurring, but is much slower and even by t=80 the L2 vorticity error is barely smaller than O(1). We note that just like in the analytical test problem, when μ2=0 the vorticity convergence in time is independent of μ1.
To further illustrate the convergence of the DNS, we show contour plots of the DNS solution, the VVDA solution, and their difference, in figures 5-8. For these simulations, we used μ1=μ2=10 in figures 5-6, but used μ2=0 for figures 7-8. As expected due to the plots in figure 4, when μ1=μ2=10 we observe rapid convergence in the plots for VV-DA velocity and vorticity to the DNS velocity and vorticity, and by t=1 the contour plots are visually indistinguishable. This is not the case, however, when μ2=0. In this case, while the velocity plots do agree with DNS velocities by t=1 (in the eyeball norm), the vorticity error remains observable at t=10 and even at t=20 there are some small difference. The contour plots of the errors at early times for vorticity show the largest errors occur near vortex centers, indicating that the VV-DA method is not accurately predicting the strength of the vortices.
4.2.1. Re=100
We also tested Re=100 for flow past a flat plate, using the same discretization parameters as above for the Re=50 case, and overall see similar results as for the Re=50 case. When both velocity and vorticity are nudged, convergence to the true solution is exponential for both velocity and vorticity in the L2(Ω) norm, and we observe that at early times the larger μ convergence curves are steeper, but at later times μ=1 converges more rapidly. When only velocity is nudged, larger μ provides faster convergence of the velocity, but the vorticity converges nearly independent of μ.
5.
Conclusions
We have analyzed a VV scheme for NSE enhanced with CDA, using linearized backward Euler or BDF2 in time and finite elements in space. We proved that applying CDA preserves the unconditional stability properties of the scheme, and also yields optimal long-time accuracy if both velocity and vorticity are nudged, or velocity-only. If only velocity is nudged, then the convergence in time to the true solution is slower, but still exponentially fast in time. Numerical tests illustrate the theory, including the difference between nudging only velocity or also nudging vorticity.
For future directions, since nudging vorticity is difficult in practice due to accurate measurement data not typically being available, one may try to obtain better results for the velocity-only-nudging by penalizing the difference between wh and rotvh in the vorticity equation. That is, by setting μ2=0 and adding the term γ(w−rotu) to the vorticity equation (3), it may be possible to analytically prove a convergence result resembling Theorem 3.4. Determining whether this is possible, and if so then for what values of γ, and whether it works in practice (i.e. how large are associated constants), would need a detailed further study which the authors plan to undertake.