Loading [MathJax]/jax/element/mml/optable/BasicLatin.js
Research article

An enhanced whale optimization algorithm for DNA storage encoding


  • Metaheuristic algorithms have the drawback that local optimal solutions are prone to precocious convergence. In order to overcome the disadvantages of the whale optimization algorithm, we propose an improved selective opposition whale optimization algorithm (ISOWOA) in this paper. Firstly, the enhanced quasi-opposition learning (EQOBL) is applied to selectively update the position of the predator, calculate the fitness of the population before and after, and retain optimal individuals as the food source position; Secondly, an improved time-varying update strategy for inertia weight predator position is proposed, and the position update of the food source is completed by this strategy. The performance of the algorithm is analyzed by 23 benchmark functions of CEC 2005 and 15 benchmark functions of CEC 2015 in various dimensions. The superior results are further shown by Wilcoxon's rank sum test and Friedman's nonparametric rank test. Finally, its applicability is demonstrated through applications to the field of biological computing. In this paper, our aim is to achieve access to DNA files and designs high-quantity DNA code sets by ISOWOA. The experimental results show that the lower bounds of the multi-constraint storage coding sets implemented in this paper equals or surpasses that of previous optimal constructions. The data show that the amount of the DNA storage cods filtered by ISOWOA increased 2–18%, which demonstrates the algorithm's reliability in practical optimization tasks.

    Citation: Sijie Wang, Shihua Zhou, Weiqi Yan. An enhanced whale optimization algorithm for DNA storage encoding[J]. Mathematical Biosciences and Engineering, 2022, 19(12): 14142-14172. doi: 10.3934/mbe.2022659

    Related Papers:

    [1] Matthew Gardner, Adam Larios, Leo G. Rebholz, Duygu Vargun, Camille Zerfas . Continuous data assimilation applied to a velocity-vorticity formulation of the 2D Navier-Stokes equations. Electronic Research Archive, 2021, 29(3): 2223-2247. doi: 10.3934/era.2020113
    [2] Caifeng Liu, Pan Liu . On Liouville-type theorem for the stationary compressible Navier–Stokes equations in R3. Electronic Research Archive, 2024, 32(1): 386-404. doi: 10.3934/era.2024019
    [3] Guoliang Ju, Can Chen, Rongliang Chen, Jingzhi Li, Kaitai Li, Shaohui Zhang . Numerical simulation for 3D flow in flow channel of aeroengine turbine fan based on dimension splitting method. Electronic Research Archive, 2020, 28(2): 837-851. doi: 10.3934/era.2020043
    [4] Jie Zhang, Gaoli Huang, Fan Wu . Energy equality in the isentropic compressible Navier-Stokes-Maxwell equations. Electronic Research Archive, 2023, 31(10): 6412-6424. doi: 10.3934/era.2023324
    [5] Jie Qi, Weike Wang . Global solutions to the Cauchy problem of BNSP equations in some classes of large data. Electronic Research Archive, 2024, 32(9): 5496-5541. doi: 10.3934/era.2024255
    [6] Xiaoxia Wang, Jinping Jiang . The uniform asymptotic behavior of solutions for 2D g-Navier-Stokes equations with nonlinear dampness and its dimensions. Electronic Research Archive, 2023, 31(7): 3963-3979. doi: 10.3934/era.2023201
    [7] Jay Chu, Xiaozhe Hu, Lin Mu . A pressure-robust divergence free finite element basis for the Stokes equations. Electronic Research Archive, 2024, 32(10): 5633-5648. doi: 10.3934/era.2024261
    [8] Wei Shi, Xinguang Yang, Xingjie Yan . Determination of the 3D Navier-Stokes equations with damping. Electronic Research Archive, 2022, 30(10): 3872-3886. doi: 10.3934/era.2022197
    [9] J. Bravo-Olivares, E. Fernández-Cara, E. Notte-Cuello, M.A. Rojas-Medar . Regularity criteria for 3D MHD flows in terms of spectral components. Electronic Research Archive, 2022, 30(9): 3238-3248. doi: 10.3934/era.2022164
    [10] Guochun Wu, Han Wang, Yinghui Zhang . Optimal time-decay rates of the compressible Navier–Stokes–Poisson system in R3. Electronic Research Archive, 2021, 29(6): 3889-3908. doi: 10.3934/era.2021067
  • Metaheuristic algorithms have the drawback that local optimal solutions are prone to precocious convergence. In order to overcome the disadvantages of the whale optimization algorithm, we propose an improved selective opposition whale optimization algorithm (ISOWOA) in this paper. Firstly, the enhanced quasi-opposition learning (EQOBL) is applied to selectively update the position of the predator, calculate the fitness of the population before and after, and retain optimal individuals as the food source position; Secondly, an improved time-varying update strategy for inertia weight predator position is proposed, and the position update of the food source is completed by this strategy. The performance of the algorithm is analyzed by 23 benchmark functions of CEC 2005 and 15 benchmark functions of CEC 2015 in various dimensions. The superior results are further shown by Wilcoxon's rank sum test and Friedman's nonparametric rank test. Finally, its applicability is demonstrated through applications to the field of biological computing. In this paper, our aim is to achieve access to DNA files and designs high-quantity DNA code sets by ISOWOA. The experimental results show that the lower bounds of the multi-constraint storage coding sets implemented in this paper equals or surpasses that of previous optimal constructions. The data show that the amount of the DNA storage cods filtered by ISOWOA increased 2–18%, which demonstrates the algorithm's reliability in practical optimization tasks.



    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

    utνΔu+(u)u+p=f,u=0, (1)

    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:

    utνΔu+ω×u+P=f,u=0,ωtνΔω+(u)ω=rotf. (2)

    Here, ω represents the (scalar) vorticity, P:=p+12|u|2 is the Bernoulli pressure, and rot is the 2D curl operation: rot(f1f2):=f2xf1y. 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

    vtνΔv+w×v+q+μ1IH(vu)=f,v=0,wtνΔw+(v)w+μ2IH(wω)=rotf, (3)

    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,μ20 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.

    We now provide notation and mathematical preliminaries to allow for a smooth analysis to follow. We consider the domain ΩR2 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

    X:=H1#(Ω)2={vH1loc(R)2, v is 2π-periodic in each direction, Ωv dx=0},Q:=L2#(Ω)={qL2loc(R), q is 2π-periodic in each direction, Ωq dx=0},W:=H1#(Ω)={vH1loc(R), v is 2π-periodic in each direction, Ωv dx=0}.

    In X (and W), we have the Poincaré inequality: there exists a constant CP depending only on Ω such that for any ϕX (or W),

    ϕCPϕ.

    We define the skew-symmetric trilinear operator b:X×W×WR to use for the nonlinear term in the vorticity equation, by

    b(u,ω,χ):=12((uω,χ)(uχ,ω)).

    The following lemma is proven in [30], and is useful in our analysis.

    Lemma 2.1. Suppose constants r and B satisfy r>1, B0. Then if the sequence of real numbers {an} satisfies

    ran+1an+B,

    we have that

    an+1a0(1r)n+1+Br1.

    Denote by τh a regular, conforming triangulation of the domain Ω, and let XhX, QhQ be velocity-pressure spaces that satisfy the inf-sup condition. We will assume the use of Xh=XPk(τh) and Qh=QPk1(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:=WPk(τh). Define the discretely divergence free subspace by

    Vh:={vhXh|(vh,qh)=0qhQh}.

    We will assume the mesh is sufficiently regular so that the inverse inequality holds in Xh: There exists a constant C such that

    χhCh1χhχhXh.

    The discrete Laplacian operator is defined as: For ϕH1(Ω)2, ΔhϕXh satisfies

    (Δhϕ,vh)=(ϕ,vh) vhXh. (4)

    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 ϕH1(Ω)2, find AhϕVh such that for all vhVh,

    (Ahϕ,vh)=(ϕ,vh). (5)

    By the definition of discrete Laplace and Stokes operators, we have the Poincaré inequalities

    χhCPΔhχh χhXh, (6)
    ϕhCPAhϕh ϕhVh. (7)

    We recall the following discrete Agmon inequalities and discrete Lp bounds [25]:

    vhLCvh1/2Ahvh1/2 vhVh, (8)
    vhLCvh1/2Δhvh1/2 vhXh, (9)
    vhL3Cvh1/3Δhvh2/3 vhXh. (10)

    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 wH1(Ω)d,

    IH(w)wCHw, (11)
    IH(w)Cw. (12)

    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.

    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 v0hVh and w0hWh, find (vn+1h,wn+1h,Pn+1h) (Xh,Wh,Qh) for n=0,1,2,..., satisfying

    1Δt(vn+1hvnh,χh)+(wnh×vn+1h,χh)(Pn+1h,χh)+ν(vn+1h,χh)+μ1(IH(vn+1hun+1),IH(χh))=(fn+1,χh), (13)
    (vn+1h,rh)=0, (14)
    1Δt(wn+1hwnh,ψh)+b(vn+1h,wn+1h,ψh)+ν(wn+1h,ψh)+μ2(IH(wn+1hrotun+1),IH(ψh))=(rotfn+1,ψh), (15)

    for all (χh,ψh,rh)(Xh,Wh,Qh), where IH(un+1), IH(rotun+1) are assumed known for all n0.

    We begin our analysis with long-time stability estimates, followed by long-time accuracy.

    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 fL(0,;L2) and uL(0,;H1). Then, for any Δt>0, any integer n>0, and nudging parameters μ1,μ20, velocity and vorticity solutions to Algorithm 3.1 satisfy

    vnh2αnv0h2+CC2Pν(ν1f2L(0,;H1)+μ1u2L(0,;L2))=:C1, (16)
    ωnh2αnω0h2+CC2Pν(ν1f2L(0,;L2)+μ2rotu2L(0,;L2))=:C2, (17)

    where α=1+νC2PΔt.

    Proof. Begin by choosing χh=2Δtvn+1h in (13), which vanishes the nonlinear and pressure terms, and leaves

    vn+1h2vnh2+vn+1hvnh2+2Δtνvn+1h2+2Δtμ1IH(vn+1h)2=2Δt(fn+1,vn+1h)+2Δtμ1(IH(un+1),IH(vn+1h)).

    The first right hand side term is bounded using the dual norm and Young's inequality via

    2Δt(fn+1,vn+1h)Δtν1fn+121+Δtνvn+1h2,

    and for the interpolation term, we use Cauchy-Schwarz, the interpolation property (12) and Young's inequality to get

    2Δtμ1(IH(un+1),IH(vn+1h))2Δtμ1IH(un+1)IH(vn+1h)CΔtμ1un+1IH(vn+1h)CΔtμ1un+12+Δtμ1IH(vn+1h)2.

    Combining the above estimates and dropping vn+1hvnh2 and IH(vn+1h)2 from the left hand side produces the bound

    vn+1h2+Δtνvn+1h2vnh2+Δtν1fn+121+CΔtμ1 un+12,

    and thanks to the Poincaré inequality, we obtain

    (1+C2PΔtν)vn+1h2vnh2+CΔt(ν1fn+121+μ1un+12).

    Defining α=1+C2PΔtν>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 fL(0,;H1) and uL(0,;H1). Then, for any Δt>0, any integer n>0, and nudging parameters μ1,μ20, velocity and vorticity solutions to Algorithm 3.1 satisfy

    vnh2αnv0h2+CC2Pν2(f2L(0,;L2)+C42C21ν2+μ21u2L(0,;L2)+μ21C21)=:˜C1, (18)
    wnh2αnw0h2+C2PCν2(rotfn+12+ν4˜C61C22+ν2˜C41C22+μ22rotun+12+μ22C22), (19)

    where α=1+νC2PΔt.

    Proof. After testing the velocity equation (13) with χh=2ΔtAhvn+1h, we obtain

    vn+1h2vnh2+(vn+1hvnh)2+2ΔtνAhvn+1h22Δt(fn+1,Ahvn+1h)+2Δt|(wnh×vn+1h,Ahvn+1h)|+2Δtμ1(IH(un+1vn+1h),IH(Ahvn+1h)).

    We now bound the right hand side terms. First, the forcing term is bounded by Cauchy-Schwarz and Young's inequalities via

    2Δt(fn+1,Ahvn+1h)CΔtν1fn+12+ν4ΔtAhvn+12. (20)

    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

    2Δt|(wnh×vn+1h,Ahvn+1h)|2Δtwnhvn+1hLAhvn+1hCΔtwnhvn+1h1/2Ahvn+1h3/2CΔtν3wnh4vn+1h2+ν8ΔtAhvn+1h2CC22C1Δtν3+ν8ΔtAhvn+1h2.

    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

    2Δtμ1(IH(un+1vn+1h),IH(Ahvn+1h))2Δtμ1(|(IHun+1,IHAhvn+1h)|+|(IHvn+1h,IHAhvn+1h)|)CΔtμ1IHun+1IHAhvn+1h+CΔtμ1IHvn+1hIHAhvn+1hCΔtμ21ν1un+12+CΔtμ21ν1C1+ν8ΔtAhvn+1h2. (21)

    Combining all these bounds for right hand side terms and dropping nonnegative term (vn+1hvnh)2 on left hand side give us that

    vn+1h2+ΔtνAhvn+1h2vnh2+CΔtν1fn+12+ν3ΔtCC22C1+Δtμ21ν1Cun+12+CΔtμ21ν1C1.

    By the Poincaré inequality (7), we now get

    αvn+1h2vnh2+CΔt(ν1fn+12+ν3C22C1+μ21ν1un+12+μ21ν1C1),

    where α=1+νC2PΔt. Finally, we apply Lemma 2.1 and reveal (18).

    For the vorticity estimate, choose ψh=2ΔtΔhwn+1h in (15) to get

    wn+1h2wnh2+(wn+1hwnh)2+2ΔtνΔhwn+1h22Δt|(rotfn+1,Δhwn+1h)|+2Δt|b(vn+1h,wn+1h,Δhwn+1h)|+2Δtμ2(IH(rotun+1wn+1h),IH(Δhwn+1h)).

    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.

    2Δt|b(vn+1h,wn+1h,Δhwn+1h)|2Δt(|(vn+1hwn+1h,Δhwn+1h)|+12|((vn+1h)wn+1h,Δhwn+1h)|)2Δtvn+1hL6wn+1hL3Δhwn+1h+Δtvn+1hwn+1hLΔhwn+1hCΔt˜C1wn+1h1/3Δhwn+1h5/3+CΔt˜C1wn+1h1/2Δhwn+1h3/2CΔt˜C1C1/32Δhwn+1h5/3+CΔt˜C1C1/22Δhwn+1h3/2CΔtν5˜C61C22+CΔtν3˜C41C22+ν3ΔtΔhwn+1h2.

    Now proceeding as in the velocity H1 bound will produce the H1 vorticity stability bound (19).

    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 ϕH1(Ω) with ϕ=0, PV(ϕ)Vh satisfies ((ϕPV(ϕ)),vh)=0 for all vhVh. 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 uL(0,;Hk+2(Ω)), pL(0,;Hk(Ω)) where k1 and ut,uttL(0,;H1). Then, assume that time step Δt is sufficiently small, and that μ1 and μ2 satisfy

    Cν1(ωn+12L+ωn+1PWωn+12L)μ1CνH2,
    Cν1(un+12L3+un+1PVun+12L3)μ2CνH2,

    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,

    vnhun2+wnhrotun2(1+λΔt)n(v0hu02+w0hrotu02)+Cλ1R,

    where

    R:=(μ11Δt2+μ12Δt2+ν1h2k+2+μ1h2k+2+μ2h2k+2),

    and λ=min with C independent of {\Delta t} , h and H .

    Proof. The true NSE solution satisfies the VV system

    \begin{align} \frac{1}{ {\Delta t}} (u^{n+1} - u^n) + \omega^{n} \times u^{n+1} + \nabla P^{n+1} - \nu \Delta u^{n+1} & = f^{n+1} + {\Delta t} u_{tt}(t^*) \\ & \ + (\omega^{n} - \omega^{n+1} )\times u^{n+1} , \end{align} (22)
    \begin{align} \nabla \cdot u^{n+1} & = 0, \end{align} (23)
    \begin{align} \frac{1}{ {\Delta t}}(\omega^{n+1} - \omega^n) + u^{n+1}\cdot \nabla \omega^{n+1} - \nu \Delta \omega^{n+1} & = {\rm{rot}} f^{n+1} + {\Delta t}\omega_{tt}(t^{**}), \end{align} (24)

    where u^n is the velocity at time t^n , P^n the Bernoulli pressure, \omega^n : = {\rm{rot}} u^n , and t^*, t^{**} \in [t^n, t^{n+1}] . Note that by Taylor expansion, we can write \omega^{n}-\omega^{n+1} = - {\Delta t} \omega_t (s^*) where s^* \in [t^n, t^{n+1}] .

    The difference equations are obtained by subtracting the solutions to Algorithm 3.1 from (22)-(24) after testing them with test functions from \chi_h\in X_h , r_h\in Q_h and \psi_h\in W_h , respectively. We define the differences between velocity and vorticity as e_v^{n} : = v_h^n - u^n and e_w^n : = w_h^n - \omega^n , respectively. Next, we will decompose the error into a term that lies in the discrete space V_h and one outside. To do so, add and subtract the discrete Stokes projection of u^n , denoted P_V u^{n} , to e_v^n and let \eta_v^n : = P_V u^{n} - u^n , \phi_{h,v}^n : = v_h^n - P_V u^{n} . Then e_v^n = \phi_{h,v}^n + \eta_v^n and \phi_{h,v}^{n} \in V_h . In a similar manner, by taking the H^1_0 projection of {\rm{rot}} u^n into W_h denoted by P_W \omega^{n} , we obtain e_w^n = \phi_{h,w}^n + \eta_w^n with \phi_{h,w}^n \in V_h .

    For velocity, since (\nabla \eta_v^{n+1},\nabla \phi_{h,v}^{n+1}) = 0 , the difference equation becomes

    \begin{align} & \frac{1}{2 {\Delta t}} [ \| \phi_{h,v}^{n+1}\|^2 - \| \phi_{h,v}^{n}\|^2 + \| \phi_{h,v}^{n+1} - \phi_{h,v}^n\|^2 ] + \nu \|\nabla \phi_{h,v}^{n+1}\|^2 + \mu_1 \| \phi_{h,v}^{n+1}\|^2 \\ & = {\Delta t}(u_{tt}(t^*), \phi_{h,v}^{n+1}) - \frac{1}{ {\Delta t}} ( \eta_{v}^{n+1}- \eta_{v}^n, \phi_{h,v}^{n+1}) - {\Delta t} (\omega_{t}(s^*), \phi_{h,v}^{n+1}) -(e_{\omega}^n \times v^{n+1}_h, \phi_{h,v}^{n+1}) \\ & \,\,\,\,\,\,\, -(\omega^{ n}\times \eta_{v}^{n+1}, \phi_{h,v}^{n+1}) - 2 \mu_1(I_H( \phi_{h,v}^{n+1}) - \phi_{h,v}^{n+1}, \phi_{h,v}^{n+1}) - \mu_1 \|I_H \phi_{h,v}^{n+1} - \phi_{h,v}^{n+1}\|^2 \\ & \,\,\,\,\,\,\, - \mu_1 (I_H \eta_{v}^{n+1}, I_H \phi_{h,v}^{n+1}) , \end{align} (25)

    and similarly for vorticity, we have

    \begin{align} &\frac{1}{2 {\Delta t}} [\| \phi_{h,w}^{n+1}\|^2 - \| \phi_{h,w}^{n}\|^2 + \| \phi_{h,w}^{n+1} - \phi_{h,w}^n\|^2] + \nu \|\nabla \phi_{h,w}^{n+1}\|^2 + \mu_2\| \phi_{h,w}^{n+1} \|^2 \\ & \,\,\, = {\Delta t}(\omega_{tt}(t^{**}), \phi_{h,w}^{n+1}) - \frac{1}{ {\Delta t}}( \eta_{w}^{n+1} - \eta_{w}^n, \phi_{h,w}^{n+1}) + b^*( e_v^{n+1}, \eta_{w}^{n+1}, \phi_{h,w}^{n+1}) \\ & \,\,\,\, + b^*(u^{n+1}, \eta_{w}^{n+1}, \phi_{h,w}^{n+1}) + b^*( e_v^{n+1}, \omega^{n+1}, \phi_{h,w}^{n+1}) -2 \mu_2(I_H( \phi_{h,w}^{n+1})- \phi_{h,w}^{n+1}, \phi_{h,w}^{n+1}) \\ & \,\,\,\, - \mu_2\|I_H \phi_{h,w}^{n+1}- \phi_{h,w}^{n+1}\|^2 - \mu_2(I_H \eta_{w}^{n+1}, I_H \phi_{h,w}^{n+1}), \end{align} (26)

    where in (25) we have added and subtracted \phi_{h,v}^{n+1} to write it in the form found above using

    \begin{align*} \mu_1& (I_H e_v^{n+1}, I_H \chi_h) \\ & = \mu_1(I_H \phi_{h,v}^{n+1}, I_H \chi_h) + \mu_1(I_H \eta_{v}^{n+1}, I_H \chi_h) \\ & = \mu_1(I_H \phi_{h,v}^{n+1} - \phi_{h,v}^{n+1} + \phi_{h,v}^{n+1}, I_H \chi_h - \phi_{h,v}^{n+1} + \phi_{h,v}^{n+1}) + \mu_1(I_H \eta_{v}^{n+1}, I_H \chi_h) \\ & = \mu_1 \| \phi_{h,v}^{n+1}\|^2 + \mu_1(I_H \phi_{h,v}^{n+1} - \phi_{h,v}^{n+1}, \phi_{h,v}^{n+1}) + \mu_1( \phi_{h,v}^{n+1}, I_H \chi_h - \phi_{h,v}^{n+1}) \\ & \,\,\,\, + \mu_1 (I_H \phi_{h,v}^{n+1} - \phi_{h,v}^{n+1}, I_H\chi_h - \phi_{h,v}^{n+1}) + \mu_1(I_H \eta_{v}^{n+1}, \chi_h), \end{align*}

    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

    \begin{align*} {\Delta t}(u_{tt}(t^*), \phi_{h,v}^{n+1}) &\leq {\Delta t} \|u_{tt}\|_{L^\infty(0,\infty;L^2(\Omega))} \| \phi_{h,v}^{n+1}\| \\ &\leq C {\Delta t}^2 \mu_1^{-1}\|u_{tt}\|^2_{L^\infty(0,\infty;L^2(\Omega))} + \frac{ \mu_1}{20}\| \phi_{h,v}^{n+1}\|^2, \end{align*}
    \begin{align*} | {\Delta t} (\omega_{tt}(s^*), \phi_{h,v}^{n+1}) | &\leq C {\Delta t} \|\omega_{tt}\|_{L^{\infty}(0,\infty;L^{2}(\Omega))} \|u^{n+1}\|_{L^{\infty}} \| \phi_{h,v}^{n+1}\| \\ &\leq C {\Delta t}^2 \mu_1^{-1} \|\omega_{tt}\|^2_{L^{\infty}(0,\infty;L^{2}(\Omega))} \|u^{n+1}\|^2_{L^{\infty}} + \frac{\mu_1}{20}\| \phi_{h,v}^{n+1}\|^2, \end{align*}
    \begin{align*} \frac{1}{ {\Delta t}} ( \eta_{v}^{n+1} - \eta_{v}^n, \phi_{h,v}^{n+1}) & = (\eta_{v,t}(s^*), \phi_{h,v}^{n+1}) \\ & \leq \|\eta_{v,t}(s^*)\| \| \phi_{h,v}^{n+1}\| \\ & \leq C \mu_1^{-1}\|\eta_{v,t}(s^*)\|^2 + \frac{\mu_1}{20}\| \phi_{h,v}^{n+1}\|^2, \end{align*}

    where s^* \in [t^n, t^{n+1}] .

    For nonlinear terms in (25), first we add and subtract e_w^{n+1} in the first component, and u^{n+1} in second component to get

    \begin{align*} ( e_w^n \times v^{n+1}_h, \phi_{h,v}^{n+1}) & = (( e_w^n- e_w^{n+1}) \times e_v^{n+1}, \phi_{h,v}^{n+1}) + ( e_w^{n+1} \times e_v^{n+1}, \phi_{h,v}^{n+1}) \\&\,\,\ + (( e_w^n- e_w^{n+1}) \times u^{n+1}, \phi_{h,v}^{n+1}) + ( e_w^{n+1} \times u^{n+1}, \phi_{h,v}^{n+1}) \\& = (( e_w^n- e_w^{n+1}) \times \eta_{v}^{n+1}, \phi_{h,v}^{n+1}) + ( e_w^{n+1} \times \eta_{v}^{n+1}, \phi_{h,v}^{n+1}) \\&\,\,\ + (( e_w^n- e_w^{n+1}) \times u^{n+1}, \phi_{h,v}^{n+1}) + ( e_w^{n+1} \times u^{n+1}, \phi_{h,v}^{n+1}). \end{align*}

    The all resulting terms are bounded by Hölder's and Young's inequalities to obtain

    \begin{align*} &(( e_w^n- e_w^{n+1}) \times \eta_{v}^{n+1}, \phi_{h,v}^{n+1}) \\ & \leq C\| \phi_{h,w}^n- \phi_{h,w}^{n+1}\| \| \eta_{v}^{n+1}\|_{L^{3}} \| \phi_{h,v}^{n+1}\|_{L^{6}}+ C\| \eta_{w}^n- \eta_{w}^{n+1}\|_{L^{\infty}} \| \eta_{v}^{n+1}\| \| \phi_{h,v}^{n+1}\| \\ &\leq C\nu^{-1}\| \phi_{h,w}^n- \phi_{h,w}^{n+1}\|^2 \| \eta_{v}^{n+1}\|^2_{L^{3}} + \frac{\nu}{8}\|\nabla \phi_{h,v}^{n+1}\|^2 \\ &\,\,\,\,\, + C\mu_1^{-1}\| \eta_{w}^n- \eta_{w}^{n+1}\|^2_{L^{\infty}} \| \eta_{v}^{n+1}\|^2+\frac{\mu_1}{20} \| \phi_{h,v}^{n+1}\|^2, \end{align*}
    \begin{align*} ( e_w^{n+1} \times \eta_{v}^{n+1}, \phi_{h,v}^{n+1})&\leq C \| \phi_{h,w}^{n+1}\| \| \eta_{v}^{n+1}\|_{L^{3}} \| \phi_{h,v}^{n+1}\|_{L^{6}} +C \| \eta_{w}^{n+1}\| \| \eta_{v}^{n+1}\|_{L^{\infty}} \| \phi_{h,v}^{n+1}\| \\ &\leq C\nu^{-1} \| \phi_{h,w}^{n+1}\|^2 \| \eta_{v}^{n+1}\|^2_{L^{3}} + \frac{\nu}{8}\|\nabla \phi_{h,v}^{n+1}\|^2 \\ &\,\,\,\,\, + C\mu_1^{-1}\| \eta_{w}^{n+1}\|^2 \| \eta_{v}^{n+1}\|^2_{L^{\infty}} + \frac{\mu_1}{20}\| \phi_{h,v}^{n+1}\|^2, \end{align*}
    \begin{align*} &(( e_w^n- e_w^{n+1}) \times u^{n+1}, \phi_{h,v}^{n+1}) \\ &\leq C\| \phi_{h,w}^n- \phi_{h,w}^{n+1}\| \|u^{n+1}\|_{L^{3}} \| \phi_{h,v}^{n+1}\|_{L^{6}} + C\| \eta_{w}^n- \eta_{w}^{n+1}\| \|u^{n+1}\|_{L^{\infty}} \| \phi_{h,v}^{n+1}\| \\ &\leq C\nu^{-1}\| \phi_{h,w}^n- \phi_{h,w}^{n+1}\|^2 \|u^{n+1}\|^2_{L^{3}} + \frac{\nu}{8} \|\nabla \phi_{h,v}^{n+1}\|^2 \\ &\,\,\,\,\, + C\mu_1^{-1}\| \eta_{w}^n- \eta_{w}^{n+1}\|^2 \|u^{n+1}\|^2_{L^{\infty}} + \frac{\mu_1}{20}\| \phi_{h,v}^{n+1}\|^2, \end{align*}
    \begin{align*} ( e_w^{n+1} \times u^{n+1}, \phi_{h,v}^{n+1})& \leq C \| \phi_{h,w}^{n+1}\| \|u^{n+1}\|_{L^{3}} \| \phi_{h,v}^{n+1}\|_{L^{6}} + C\| \eta_{w}^{n+1}\| \|u^{n+1}\|_{L^{\infty}} \| \phi_{h,v}^{n+1}\| \\ & \leq C\nu^{-1} \| \phi_{h,w}^{n+1}\|^2 \|u^{n+1}\|^2_{L^{3}} + \frac{\nu}{8} \|\nabla \phi_{h,v}^{n+1}\|^2 \\ &\,\,\,\,\, + C\mu_1^{-1}\| \eta_{w}^{n+1}\|^2 \|u^{n+1}\|_{L^{\infty}}^2 + \frac{\mu_1}{20}\| \phi_{h,v}^{n+1}\|^2. \end{align*}

    Then, for the last nonlinear term, we apply Hölder's, Poincaré's and Young's inequalities and get

    \begin{align*} ( \omega^{n} \times \eta_{v}^{n+1}, \phi_{h,v}^{n+1})&\leq \| \omega^{n} \|_{L^{\infty}} \| \eta_{v}^{n+1}\| \| \phi_{h,v}^{n+1}\| \\ & \leq C\nu^{-1}\| \omega^{n} \|^2_{L^{\infty}} \| \eta_{v}^{n+1}\|^2 + \frac{\nu}{8} \|\nabla \phi_{h,v}^{n+1}\|^2. \end{align*}

    Next, the first interpolation term on the right hand side of (25) will be bounded with Cauchy-Schwarz inequality and (11) to obtain

    \begin{align*} \mu_1(I_H( \phi_{h,v}^{n+1}) - \phi_{h,v}^{n+1}, \phi_{h,v}^{n+1}) &\leq \mu_1 \|I_H( \phi_{h,v}^{n+1}) - \phi_{h,v}^{n+1}\|\| \phi_{h,v}^{n+1}\| \\ & \leq \mu_1 C H \|\nabla \phi_{h,v}^{n+1}\|\| \phi_{h,v}^{n+1}\| \\ & \leq C \mu_1 H^2 \|\nabla \phi_{h,v}^{n+1}\|^2 + \frac{ \mu_1}{20}\| \phi_{h,v}^{n+1}\|^2. \end{align*}

    For the second interpolation term, we apply inequality (12), which yields

    \begin{align*} \mu_1\|I_H \phi_{h,v}^{n+1} - \phi_{h,v}^{n+1}\|^2 \leq C \mu_1 H^2 \|\nabla \phi_{h,v}^{n+1}\|^2 . \end{align*}

    Finally, the last interpolation term will be bounded using Cauchy-Schwarz, (12), and Young's inequality to get the bound

    \begin{align*} \mu_1(I_H \eta_{v}^{n+1},I_H \phi_{h,v}^{n+1}) & \leq C \mu_1\| \eta_{v}^{n+1}\|^2 + \frac{ \mu_1}{20}\| \phi_{h,v}^{n+1}\|^2 . \end{align*}

    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\cdot\nabla v,w) , with \| \nabla \cdot 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

    \begin{align*} b^*& ( e_v^{n+1}, \eta_{w}^{n+1}, \phi_{h,w}^{n+1}) = ( e_v^{n+1}\cdot \nabla \eta_{w}^{n+1}, \phi_{h,w}^{n+1}) \\ & = ( \phi_{h,v}^{n+1}\cdot \nabla \eta_{w}^{n+1}, \phi_{h,w}^{n+1}) + ( \eta_{v}^{n+1}\cdot \nabla \eta_{w}^{n+1}, \phi_{h,w}^{n+1}) \\ & = ( \phi_{h,v}^{n+1}\cdot \nabla \phi_{h,w}^{n+1}, \eta_{w}^{n+1}) + ( \eta_{v}^{n+1}\cdot \nabla \phi_{h,w}^{n+1}, \eta_{w}^{n+1}) \\ & \leq C \| \phi_{h,v}^{n+1}\| \|\nabla \phi_{h,w}^{n+1}\| \| \eta_{w}^{n+1}\|_{L^\infty} + C\| \eta_{v}^{n+1}\| \|\nabla \phi_{h,w}^{n+1}\| \| \eta_{w}^{n+1}\|_{L^\infty} \\ & \leq C\nu^{-1} \| \phi_{h,v}^{n+1}\|^2 \| \eta_{w}^{n+1}\|_{L^\infty}^2 + \frac{\nu}{10}\|\nabla \phi_{h,w}^{n+1}\|^2 + C\nu^{-1}\| \eta_{v}^{n+1}\|^2 \| \eta_{w}^{n+1}\|_{L^\infty}^2 \\ & \,\,\,\, + \frac{\nu}{10} \|\nabla \phi_{h,w}^{n+1}\|^2. \end{align*}

    For the second nonlinear term, we use Hölder's, Póincare's and Young's inequalities, which gives

    \begin{align*} b^*(u^{n+1}, \eta_{w}^{n+1}, \phi_{h,w}^{n+1}) & = (u^{n+1}\cdot \nabla \eta_{w}^{n+1}, \phi_{h,w}^{n+1}) \\ & = - (u^{n+1}\cdot \nabla \phi_{h,w}^{n+1}, \eta_{w}^{n+1}) \\ & \leq C \|u^{n+1}\|_{L^\infty } \| \nabla \phi_{h,w}^{n+1}\| \| \eta_{w}^{n+1} \| \\ & \leq C \nu^{-1} \|u^{n+1}\|_{L^\infty }^2 \| \eta_{w}^{n+1}\|^2 + \frac{\nu}{10} \| \nabla \phi_{h,w}^{n+1}\|^2. \end{align*}

    For the last nonlinear term, we begin by breaking up the velocity error term, then apply Hölder's and Young's inequalities to get

    \begin{align*} b^*& ( e_v^{n+1}, \omega^{n+1}, \phi_{h,w}^{n+1}) = ( e_v^{n+1}\cdot \nabla \omega^{n+1}, \phi_{h,w}^{n+1}) \\ & = ( \phi_{h,v}^{n+1}\cdot \nabla \omega^{n+1}, \phi_{h,w}^{n+1}) + ( \eta_{v}^{n+1}\cdot \nabla \omega^{n+1}, \phi_{h,w}^{n+1})\\ & = ( \phi_{h,v}^{n+1}\cdot \nabla \phi_{h,w}^{n+1},\omega^{n+1}) + ( \eta_{v}^{n+1}\cdot \nabla \phi_{h,w}^{n+1},\omega^{n+1}) \\ & \leq C\| \phi_{h,v}^{n+1}\| \|\nabla \phi_{h,w}^{n+1}\| \|\omega^{n+1}\|_{L^\infty} + C \| \eta_{v}^{n+1}\| \|\nabla \phi_{h,w}^{n+1}\| \|\omega^{n+1}\|_{L^\infty} \\ & \leq C \nu^{-1}\| \phi_{h,v}^{n+1}\|^2 \| \omega^{n+1}\|_{L^\infty}^2 + \frac{\nu}{10} \|\nabla \phi_{h,w}^{n+1}\|^2 +C\nu^{-1} \| \eta_{v}^{n+1}\|^2 \| \omega^{n+1}\|_{L^\infty}^2 \\ & \,\,\,\, + \frac{\nu}{10}\|\nabla \phi_{h,w}^{n+1}\|^2. \end{align*}

    Replacing the right hand sides of (25) and (26) with the computed bounds and dropping nonnegative terms with \| \phi_{h,v}^{n+1}- \phi_{h,v}^{n}\|^2 yields the bound

    \begin{align*} &\frac{1}{2 {\Delta t}} \left( \| \phi_{h,v}^{n+1}\|^2 + \| \phi_{h,w}^{n+1}\|^2 - \| \phi_{h,v}^n\|^2 - \| \phi_{h,w}^n\|^2 \right) \\ & + \left(\frac{1}{2 {\Delta t}}-C\nu^{-1} (\| \eta_{v}^{n+1}\|^2_{L^3} +\|u^{n+1}\|^2_{L^3} )\right) \| \phi_{h,w}^{n+1}- \phi_{h,w}^{n}\|^2 \\ &\ \ \ + \frac{\nu}{4} \|\nabla \phi_{h,v}^{n+1}\|^2 + \left( \frac{\nu}{4} - C \mu_1 H^2 \right)\|\nabla \phi_{h,v}^{n+1}\|^2 \\ &+ \frac{\nu}{4} \|\nabla \phi_{h,w}^{n+1}\|^2 + \left( \frac{\nu}{4} - C \mu_2 H^2 \right)\|\nabla \phi_{h,w}^{n+1}\|^2 \\ &\ \ \ + \frac{ \mu_1}{4} \| \phi_{h,v}^{n+1}\|^2 + \left( \frac{ \mu_1}{4}-C\nu^{-1}(\|\omega^{n+1}\|^2_{L^{\infty}}+\| \eta_{w}^{n+1}\|^2_{L^{\infty}}) \right) \| \phi_{h,v}^{n+1}\|^2 \\ &\ \ \ + \frac{ \mu_2}{4} \| \phi_{h,w}^{n+1}\|^2 + \left( \frac{ \mu_2}{4}-C\nu^{-1}(\|u^{n+1}\|^2_{L^{3}}+\| \eta_{v}^{n+1}\|^2_{L^{3}}) \right) \| \phi_{h,w}^{n+1}\|^2 \end{align*}
    \begin{align*} & \leq C {\Delta t}^2 \left( \mu_1^{-1}\|u_{tt}\|^2_{L^\infty(0,\infty ; L^2)} + \mu_1^{-1}\|\omega_{tt}\|^2_{L^\infty(0,\infty ; L^2)} \|u^{n+1}\|^2_{L^{\infty}} + \mu_2^{-1}\|\omega_{tt}\|^2_{L^\infty(0,\infty; L^2)} \right) \\ & \,\,\,\, + C \mu_1^{-1} \big( \|\eta_{v,t}\|_{L^\infty(0,\infty;L^2)}^2 + \| \eta_{w}^{n+1}- \eta_{w}^{n}\|^2_{L^{\infty}} \| \eta_{v}^{n+1}\|^2 \\ & + \| \eta_{w}^{n+1}\|^2 \| \eta_{v}^{n+1}\|^2_{L^{\infty}} + \| \eta_{w}^{n+1}- \eta_{w}^{n}\|^2 \|u^{n+1}\|^2_{L^{\infty}} \\ & \,\,\,\, + \| \eta_{w}^{n+1}\|^2 \|u^{n+1}\|^2_{L^{\infty}} \big) + C\mu_2^{-1} \|\eta_{w,t}\|_{L^\infty(0,\infty;L^2)}^2 \\ &+ C \nu^{-1} \big( \|\omega^{n}\|^2_{L^{\infty}} \| \eta_{v} ^{n+1}\|^2 + \| \eta_{v}^{n+1}\|^2\| \eta_{w}^{n+1}\|_{L^\infty}^2 \\ & \,\,\,\, + \| \eta_{v}^{n+1}\|^2 \|\omega^{n+1}\|_{L^\infty}^2 + \| \eta_{w}^{n+1}\|^2\|u^{n+1}\|_{L^\infty}^2 \big) + C \mu_1 \| \eta_{v}^{n+1}\|^2 + C \mu_2 \| \eta_{w}^{n+1}\|^2. \end{align*}

    Using the assumptions on H and the nudging parameters, the time step restriction, and smoothness of the true solution, this reduces to

    \begin{align*} \frac{1}{2 {\Delta t}}& \left( \| \phi_{h,v}^{n+1}\|^2 + \| \phi_{h,w}^{n+1}\|^2 - \| \phi_{h,v}^n\|^2 - \| \phi_{h,w}^n\|^2 \right) + \frac{\nu}{4} \|\nabla \phi_{h,v}^{n+1}\|^2 + \frac{\nu}{4} \|\nabla \phi_{h,w}^{n+1}\|^2 \\ & + \frac{ \mu_1}{4} \| \phi_{h,v}^{n+1}\|^2 + \frac{ \mu_2}{4} \| \phi_{h,w}^{n+1}\|^2 \\ & \leq C {\Delta t}^2 \left( \mu_1^{-1} + \mu_1^{-1} + \mu_2^{-1} \right) + C\mu_2^{-1} \|\eta_{w,t}\|_{L^\infty(0,\infty;L^2)}^2 \\ & \ + C \mu_1^{-1} \big( \|\eta_{v,t}\|_{L^\infty(0,\infty;L^2)}^2 + \| \eta_{w}^{n+1}- \eta_{w}^{n}\|^2_{L^{\infty}} \| \eta_{v}^{n+1}\|^2 \\ &+ \| \eta_{w}^{n+1}\|^2 \| \eta_{v}^{n+1}\|^2_{L^{\infty}} + \| \eta_{w}^{n+1}- \eta_{w}^{n}\|^2 + \| \eta_{w}^{n+1}\|^2 \big) \\ & \ +C \nu^{-1} \big( \| \eta_{v} ^{n+1}\|^2 + \| \eta_{v}^{n+1}\|^2\| \eta_{w}^{n+1}\|_{L^\infty}^2 + \| \eta_{v}^{n+1}\|^2 + \| \eta_{w}^{n+1}\|^2 \big) \\ & + C \mu_1 \| \eta_{v}^{n+1}\|^2 + C \mu_2 \| \eta_{w}^{n+1}\|^2. \end{align*}

    Now define

    \begin{align*} \lambda_1 : = & \frac{\mu_1}{4} + \frac{\nu C_P^{-2}}{4} ,\\ \lambda_2 : = & \frac{\mu_2}{4} + \frac{\nu C_P^{-2}}{4} . \end{align*}

    Using this in the inequality after applying Poincare's inequality and multiplying each side by 2 {\Delta t} , we get

    \begin{align*} &( 1 + {\Delta t}\lambda_1 )\| \phi_{h,v}^{n+1}\|^2 +(1 + {\Delta t}\lambda_2 ) \| \phi_{h,w}^{n+1}\|^2 \\ & \leq C {\Delta t} \left(\mu_1^{-1} {\Delta t}^2 + \mu_2^{-1} {\Delta t}^2 + \nu^{-1} h^{2k+2} + \mu_1 h^{2k+2} + \mu_2 h^{2k+2}\right) + \| \phi_{h,v}^{n}\|^2 + \| \phi_{h,w}^{n}\|^2. \end{align*}

    Then, with R: = \left(\mu_1^{-1} {\Delta t}^2 + \nu^{-1} {\Delta t}^2 + \nu^{-1} h^{2k+2} + \mu_1 h^{2k+2} + \mu_2 h^{2k+2}\right) and \lambda : = \min \left\{ \lambda_1, \lambda_2 \right\} , we obtain the bound

    \begin{align*} (1 +\lambda {\Delta t} )\left(\| \phi_{h,v}^{n+1}\|^2 +\| \phi_{h,w}^{n+1}\|^2\right) \leq C {\Delta t} R + \| \phi_{h,v}^{n}\|^2 + \| \phi_{h,w}^{n}\|^2. \end{align*}

    By Lemma 2.1, this implies

    \begin{align*} \| \phi_{h,v}^{n+1}\|^2 + \| \phi_{h,w}^{n+1}\|^2 & \leq C \lambda^{-1} R + (1 + \lambda {\Delta t} )^{-(n+1)}(\| \phi_{h,v}^{0}\|^2 + \| \phi_{h,w}^{0}\|^2). \end{align*}

    Lastly, applying triangle inequality completes the proof.

    Theorem 3.5 (Long-time L^2 accuracy of Algorithm 3.1 with \mu_1>0, \mu_2 = 0 ). Let true solution u \in L^\infty(0, \infty ; H^{k+2}(\Omega)) , p \in L^\infty(0, \infty ; H^{k}(\Omega)) where k\geq 1 and u_{t}, u_{tt}, \in L^\infty (0, \infty ; H^1) . Then, assume that time step {\Delta t} is sufficiently small, \mu_2 = 0 , and that \mu_1 satisfies

    \begin{align*} C& \nu^{-1} \max \{ (\|u^{n+1}\|^2_{L^{\infty}}+\| u^{n+1} - P_V u^{n+1}\|^2_{L^{\infty}}) , (\|\omega^{n+1}\|^2_{L^{\infty}}+\| \omega^{n+1} - P_W \omega^{n+1}\|^2_{L^{\infty}} ) \} \\ & \leq \mu_1 \leq \frac{C\nu}{H^2}, \end{align*}

    where H is chosen so that this inequality holds. Then for any time t^n , n = 0, 1, 2,... , solutions of of Algorithm 3.1 using Scott-Vogelius elements satisfy

    \begin{align} \|v_h^n-u^n\|^2 + \| w_h^n - {\rm{rot}} u^n\|^2 & \leq (1 + \lambda {\Delta t} )^{-n}(\|v_h^0-u^0\|^2 + \| w_h^0 - {\rm{rot}} u^0\|^2) + C\lambda^{-1} R, \end{align} (27)

    where

    R: = \left(\mu_1^{-1} {\Delta t}^2 + \nu^{-1} {\Delta t}^2 + \nu^{-1} h^{2k+2} + \mu_1 h^{2k+2} \right),

    and \lambda = \min \left\{ \frac{\mu_1}{4}+\frac{C_P^{-2} \nu}{4}, \frac{C_P^{-2}\nu}{4}\right\} with C independent of {\Delta t} , h and H .

    Remark 1. Algorithm 3.1 converges to the true solutions up to optimal discretization error in both cases \mu_1, \mu_2 >0 and \mu_1>0, \mu_2 = 0 . The key difference between two cases is that when \mu_2 = 0 , the convergence in time to reach optimal accuracy is much slower since \lambda 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

    \begin{align*} ( e_w^{n+1} \times \eta_{v}^{n+1}, \phi_{h,v}^{n+1})&\leq C \| \phi_{h,w}^{n+1}\| \| \eta_{v}^{n+1}\|_{L^{3}} \| \phi_{h,v}^{n+1}\|_{L^{6}} +C \| \eta_{w}^{n+1}\| \| \eta_{v}^{n+1}\|_{L^{\infty}} \| \phi_{h,v}^{n+1}\| \\ &\leq C \mu_1^{-1}\|\nabla \phi_{h,w}^{n+1}\|^2\| \eta_{v}^{n+1}\|^2_{L^\infty} + \frac{ \mu_1}{16}\| \phi_{h,v}^{n+1}\|^2 \\ &\,\,\,\,\, + C\mu_1^{-1}\| \eta_{w}^{n+1}\|^2 \| \eta_{v}^{n+1}\|^2_{L^{\infty}} + \frac{\mu_1}{20}\| \phi_{h,v}^{n+1}\|^2, \end{align*}
    \begin{align*} ( e_w^{n+1} \times u^{n+1}, \phi_{h,v}^{n+1})& \leq C \| \phi_{h,w}^{n+1}\| \|u^{n+1}\|_{L^{3}} \| \phi_{h,v}^{n+1}\|_{L^{6}} + C\| \eta_{w}^{n+1}\| \|u^{n+1}\|_{L^{\infty}} \| \phi_{h,v}^{n+1}\| \\ & \leq C \mu_1^{-1}\|\nabla \phi_{h,w}^{n+1}\|^2\|u^{n+1}\|_{L^\infty}^2 + \frac{ \mu_1}{16}\| \phi_{h,v}^{n+1}\|^2 \\ &\,\,\,\,\, + C\mu_1^{-1}\| \eta_{w}^{n+1}\|^2 \|u^{n+1}\|_{L^{\infty}}^2 + \frac{\mu_1}{20}\| \phi_{h,v}^{n+1}\|^2. \end{align*}

    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

    \begin{align*} & \frac{1}{2 {\Delta t}} \left( \| \phi_{h,v}^{n+1}\|^2 + \| \phi_{h,w}^{n+1}\|^2 - \| \phi_{h,v}^n\|^2 - \| \phi_{h,w}^n\|^2 \right) \nonumber \\ & + \left(\frac{1}{2 {\Delta t}}-C\nu^{-1} (\| \eta_{v}^{n+1}\|^2_{L^3} -\|u^{n+1}\|^2_{L^3} )\right) \| \phi_{h,w}^{n+1}- \phi_{h,w}^{n}\|^2 \\ & + \frac{ \mu_1}{4} \| \phi_{h,v}^{n+1}\|^2 + \left( \frac{ \mu_1}{4}-C\nu^{-1}(\|\omega^{n+1}\|^2_{L^{\infty}}+\| \eta_{w}^{n+1}\|^2_{L^{\infty}}) \right) \| \phi_{h,v}^{n+1}\|^2 + \frac{\nu}{4} \|\nabla \phi_{h,v}^{n+1}\|^2 \\ & + \left( \frac{\nu}{4} - C \mu_1 H^2 \right)\|\nabla \phi_{h,v}^{n+1}\|^2 + \frac{\nu}{4} \|\nabla \phi_{h,w}^{n+1}\|^2 \nonumber \\ & + \left( \frac{\nu}{4} -C\mu_1^{-1}(\|u^{n+1}\|_{L^{\infty}}+\| \eta_{v}^{n+1}\|_{L^{\infty}}) \right)\|\nabla \phi_{h,w}^{n+1}\|^2 \\ & \leq C {\Delta t}^2 \left( \mu_1^{-1}\|u_{tt}\|^2_{L^\infty(0,\infty ; L^2)} + \mu_1^{-1}\|\omega_{tt}\|^2_{L^\infty(0,\infty ; L^2)} \|u^{n+1}\|^2_{L^{\infty}} + \mu_2^{-1}\|\omega_{tt}\|^2_{L^\infty(0,\infty; L^2)} \right) \\ & \,\,\,\, + C \mu_1^{-1} \big( \|\eta_{v,t}\|_{L^\infty(0,\infty;L^2)}^2 + \| \eta_{w}^{n+1}- \eta_{w}^{n}\|^2_{L^{\infty}} \| \eta_{v}^{n+1}\|^2 + \| \eta_{w}^{n+1}\|^2 \| \eta_{v}^{n+1}\|^2_{L^{\infty}} \\& + \| \eta_{w}^{n+1}- \eta_{w}^{n}\|^2 \|u^{n+1}\|^2_{L^{\infty}} + \| \eta_{w}^{n+1}\|^2 \|u^{n+1}\|^2_{L^{\infty}} \big) + C \nu^{-1} \big( \|\omega^{n+1}\|^2_{L^{\infty}} \| \eta_{v}\|^2 \\ &+ \| \eta_{v}^{n+1}\|^2\| \eta_{w}^{n+1}\|_{L^\infty}^2 + \| \eta_{v}^{n+1}\|^2 \|\omega^{n+1}\|_{L^\infty}^2 + \| \eta_{w}^{n+1}\|^2\|u^{n+1}\|_{L^\infty}^2 \big) + C \mu_1 \| \eta_{v}^{n+1}\|^2. \end{align*}

    Provided {\Delta t} is sufficiently small and the restriction

    \max \left\{ C\nu^{-1}(\|u^{n+1}\|^2_{L^{\infty}}+\| \eta_{v}^{n+1}\|^2_{L^{\infty}}) , C\nu^{-1}(\|\omega^{n+1}\|^2_{L^{\infty}}+\| \eta_{w}^{n+1}\|^2_{L^{\infty}} ) \right\} \leq \mu_1 \leq \frac{C\nu}{H^2},

    holds, then applying Poincaré inequality to the terms on left hand side and using

    \begin{align*} \lambda_1 : = & \frac{\mu_1}{4}+\frac{C_P^{-2} \nu}{4},\\ \lambda_2 : = & \frac{C_P^{-2}\nu}{4}, \end{align*}

    and assumptions on the true solution, we obtain

    \begin{align*} (& 1 + {\Delta t}\lambda_1 )\| \phi_{h,v}^{n+1}\|^2 +(1 + {\Delta t}\lambda_2 ) \| \phi_{h,w}^{n+1}\|^2 \\ & \leq C {\Delta t} \left(\mu_1^{-1} {\Delta t}^2 + \nu^{-1} {\Delta t}^2 + \nu^{-1} h^{2k+2} + \mu_1 h^{2k+2} \right) + \| \phi_{h,v}^{n}\|^2 + \| \phi_{h,w}^{n}\|^2. \end{align*}

    From here, the proof is finished in the same way as the previous theorem.

    We now present results for a second order analogue of the first order algorithm studied above.

    Algorithm 3.6. Find (v_h^{n+1}, w_h^{n+1}, q_h^{n+1}) \in (X_h, W_h, Q_h) for n = 0,1,2,... , satisfying

    \begin{align} \frac{1}{2 \Delta t} \left( 3v_h^{n+1} - 4v_h^n + v_h^{n-1},\chi_h \right) + ( (2w_h^{n} - w_h^{n-1}) \times v_h^{n+1}, \chi_h) - (P_h^{n+1},\nabla \cdot \chi_h) & \\+ \nu (\nabla v_h^{n+1},\nabla \chi_h) + \mu_1 (I_H(v_h^{n+1} - u^{n+1}),I_H \chi_h) = (f^{n+1},\chi_h),& \end{align} (28)
    \begin{align} (\nabla \cdot v_h^{n+1},r_h) = 0,& \end{align} (29)
    \begin{align} \frac{1}{2\Delta t} \left( 3w_h^{n+1} - 4w_h^n - v_h^{n-1},\psi_h \right) + (v_h^{n+1}\cdot \nabla w_h^{n+1},\psi_h) + \nu (\nabla w_h^{n+1},\nabla \psi_h) & \\ + \mu_2 (I_H(w_h^{n+1} - {\rm{rot}} u^{n+1}),I_H(\psi_h)) = ({\rm{rot}} f^{n+1},\psi_h),& \end{align} (30)

    for all (\chi_h, \psi_h, r_h) \in (X_h,W_h,Q_h) , with v^0\in X and I_H(u^{n+1}) , I_H({\rm{rot}} u^{n+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 \mu_1>0 and \mu_2>0 ). For any time step {\Delta t}>0 , and any time t^n , n = 0, 1, 2,... , we have that solutions of Algorithm 3.6 satisfy

    \| v_h^n \| + \| w_h^n \| + \| \nabla v_h^n \| + \| \nabla w_h^n \| \le C,

    with C independent of n , {\Delta t},\ h,\ H .

    Furthermore, if we suppose the true solution u \in L^\infty(0, \infty ; H^{k+2}(\Omega)) , p \in L^\infty(0, \infty ; H^{k}(\Omega)) where k\geq 1 and u_{t}, u_{ttt}, \in L^\infty (0, \infty ; H^1) , that time step {\Delta t} is sufficiently small, Scott-Vogelius elements are used, and that \mu_1 and \mu_2 satisfy C(u) \leq \mu_1,\mu_2 \leq \frac{C\nu}{H^2} , we have the bound

    \begin{align*} & \|v_h^n-u^n\|^2 + \|\omega_h^n - {\rm{rot}} u^n\|^2 \leq \\ & (1 + \lambda {\Delta t} )^{-n}(\|v_h^0-u^0\|^2 + \|\omega_h^0 - {\rm{rot}} u^0\|^2 + \|v_h^1-u^1\|^2 + \|\omega_h^1 - {\rm{rot}} u^1\|^2) + C \lambda^{-1} R, \end{align*}

    where

    R: = \left(\mu_1^{-1} {\Delta t}^4 + \mu_2^{-1} {\Delta t}^4 + \nu^{-1} h^{2k+2} + \mu_1 h^{2k+2} + \mu_2 h^{2k+2}\right),

    and \lambda = \min \left\{ \frac{\mu_1}{4} + \frac{\nu C_P^{-2}}{4} , \frac{\mu_2}{4} + \frac{\nu C_P^{-2}}{4} \right\} with C independent of {\Delta t} , h and H .

    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 (P_2,P_1) Taylor-Hood elements for velocity and pressure, and P_2 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 X_H and W_H are defined to be piecewise constants on the same mesh used for the computations. The interpolation operator I_H was taken to be the L^2 projection operator onto X_H (or W_H ), which is known to satisfy (11)-(12) [16].

    For our first test, we investigate the theory above for Algorithm 3.6. Here we use the analytic solution

    u = \begin{bmatrix} \cos(\pi(y-t))\\\sin(\pi(x+t)) \end{bmatrix}, \quad \quad p = (1+t^2)\sin(x+y),

    on the unit square domain \Omega = (0,1)^2 with kinematic viscosity \nu = 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, (P_2,P_1) Taylor-Hood elements are used for velocity and pressure, P_2 for vorticity, and a time step size of \Delta t = 0.001 . From Section 3, we expect third order spatial convergence rate in the L^2 norm for large enough times. Results are presented below for two cases, \mu_2>0 and \mu_2 = 0 .

    To test this case, we first calculated spatial convergence rates at the final time T = 1 with the L^2 error, using successively refined uniform meshes and \mu_1 = \mu_2 = 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 \Delta t is fixed while the spatial mesh width decreases.

    Table 1.  Shown above are L^2 velocity and vorticity errors and convergence rates on varying mesh widths, at the final time T = 1 , using Algorithm 3.6 with \mu_1 = \mu_2 = 100 .
    h \|e_v(T) \| rate \|e_w(T)\| rate
    1/4 2.62008e-03 - 7.70647e-03 -
    1/8 3.20467e-04 3.0314 9.68456e-04 2.9923
    1/16 3.97307e-05 3.0146 1.20888e-04 3.0041
    1/32 4.94529e-06 3.0061 1.50809e-05 3.0029
    1/64 6.19332e-07 2.9973 1.99325e-06 2.9195
    1/128 8.13141e-08 2.9247 3.15236e-07 2.5855

     | Show Table
    DownLoad: CSV

    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 \mu_1 = \mu_2 = \mu , with \mu = 1,\ 10,\ 100,\ 1000 . Results are shown in figure 1, as L^2 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 \Delta t and the accuracy of the method. We note that as \mu is increased, convergence is faster in time, which is consistent with our theory for the case of \mu_1>0 and \mu_2>0 .

    Figure 1.  Shown above are L^2 velocity and vorticity errors for Algorithm 3.6 with \mu_1 = \mu_2 = \mu , with varying \mu>0 .

    We now consider the same tests as above, but with \mu_2 = 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 L^2 error, on the same successively refined uniform meshes, but now with \mu_1 = 100 and \mu_2 = 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 \Delta t was fixed at 0.001, although the vorticity errors are slightly worse than for the case of \mu_2 = 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 \mu_2 = 100 , and slightly worse vorticity error but still with optimal L^2 accuracy.

    Table 2.  Shown above are L^2 velocity and vorticity errors and convergence rates on varying mesh widths, at the final time T = 1 , using Algorithm 3.6 with \mu_1 = 100 and \mu_2 = 0 .
    h \|e_v(T)\| rate \|e_w(T)\| rate
    1/4 2.62003e-03 - 7.79431e-03 -
    1/8 3.20466e-04 3.0313 9.70492e-04 3.0056
    1/16 3.97175e-05 3.0123 1.20897e-04 3.0049
    1/32 4.94501e-06 3.0057 1.50883e-05 3.0023
    1/64 6.17406e-07 3.0017 2.08215e-06 2.8573
    1/128 8.11244e-08 2.9280 9.37122e-07 1.1518

     | Show Table
    DownLoad: CSV

    To test exponential convergence in time for the case of \mu_2 = 0 , we again take h = 1/32 , and compute solutions using \mu_1 = 1,\ 10,\ 100,\ 1000 . Results are shown in figure 2, as L^2 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 \mu_2 = 100 above. An important difference here compared to when \mu_2 = 100 is that the convergence of vorticity to the true solution is independent of \mu_1 , and the convergence of velocity is slower for larger choices of \mu_1 . This reduced dependence of the convergence on the nudging parameters when \mu_2 = 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.

    Figure 2.  Shown above are L^2 velocity and vorticity errors (from left to right) for Algorithm 3.6 with varying \mu_1 and \mu_2 = 0 .

    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]\times[-10,10] rectangular channel with a 0.125\times1 plate fixed ten units into the channel from the left, vertically centered. The inflow velocity is u_{in} = \langle0,1\rangle , 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 \nu = 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.

    Figure 3.  Setup for the flow past a normal flat plate.

    We computed solutions using a Delaunay generated triangular meshes that provided 27,373 total degree of freedom with (P_2, P_1, P_2) velocity-pressure-vorticity elements, and time step {\Delta t} = 0.02 . We first compared convergence in time to the DNS solution, for two cases: \mu_1 = \mu_2>0 and \mu_1>0,\ \mu_2 = 0 . Plots of L^2 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 \mu_2>0 : when vorticity is nudged in addition to velocity, convergence to the true solution is much faster in time. The convergence when \mu_2 = 0 appears to still be occurring, but is much slower and even by t = 80 the L^2 vorticity error is barely smaller than O(1) . We note that just like in the analytical test problem, when \mu_2 = 0 the vorticity convergence in time is independent of \mu_1 .

    Figure 4.  L^2 velocity and vorticity errors (from left to right) for Algorithm 3.6 with \mu_1 = \mu_2 = \mu >0 (top) and \mu_1 = \mu>0, \mu_2 = 0 (bottom).

    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 \mu_1 = \mu_2 = 10 in figures 5-6, but used \mu_2 = 0 for figures 7-8. As expected due to the plots in figure 4, when \mu_1 = \mu_2 = 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 \mu_2 = 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.

    Figure 5.  Contour plots of velocity for DNS (left), VV-DA with \mu_1 = \mu_2 = 10 (center), and their difference (right), for times t = 0,\ 0.1,\ 1,\ 10,\ 20,\ 80 (top to bottom).
    Figure 6.  Contour plots of vorticity for DNS (left), VV-DA with \mu_1 = \mu_2 = 10 (center), and their difference (right), for times t = 0,\ 0.1,\ 1,\ 10,\ 20,\ 80 (top to bottom).
    Figure 7.  Contour plots of velocity for DNS (left), VV-DA with \mu_1 = 10,\ \mu_2 = 0 (center), and their difference (right), for times t = 0,\ 0.1,\ 1,\ 10,\ 20,\ 80 (top to bottom).
    Figure 8.  Contour plots of vorticity for DNS (left), VV-DA with \mu_1 = 10,\ \mu_2 = 0 (center), and their difference (right), for times t = 0,\ 0.1,\ 1,\ 10,\ 20,\ 80 (top to bottom).
    Figure 9.  L^2 velocity and vorticity errors (from left to right) for Algorithm 3.6 with \mu_1 = \mu_2 = \mu >0 (top) and \mu_1 = \mu>0, \mu_2 = 0 (bottom), with 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 L^2(\Omega) norm, and we observe that at early times the larger \mu convergence curves are steeper, but at later times \mu = 1 converges more rapidly. When only velocity is nudged, larger \mu provides faster convergence of the velocity, but the vorticity converges nearly independent of \mu .

    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 w_h and {\rm{rot}} v_h in the vorticity equation. That is, by setting \mu_2 = 0 and adding the term \gamma (w - {\rm{rot}} u) 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 \gamma , 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.



    [1] S. Zhou, L. Xing, X. Zheng, N. Du, L. Wang, Q. Zhang, A self-adaptive differential evolution algorithm for scheduling a single batch-processing machine with arbitrary job sizes and release times, IEEE Trans. Cybern., 51 (2021), 1430–1442. https://doi.org/10.1109/TCYB.2019.2939219 doi: 10.1109/TCYB.2019.2939219
    [2] W. Deng, J. Xu, Y. Song, H. Zhao, Differential evolution algorithm with wavelet basis function and optimal mutation strategy for complex optimization problem, Appl. Soft Comput., 100 (2021), 1568–4946. https://doi.org/10.1016/j.asoc.2020.106724 doi: 10.1016/j.asoc.2020.106724
    [3] F. Zhao, X. He, L. Wang, A two-stage cooperative evolutionary algorithm with problem-specific knowledge for energy-efficient scheduling of no-wait flow-shop problem, IEEE Trans. Cybern., 51 (2020), 5291–5303. https://doi.org/10.1109/TCYB.2020.3025662 doi: 10.1109/TCYB.2020.3025662
    [4] A. Kaveh, A. D. Eslamlou, Water strider algorithm: A new metaheuristic and applications, Structures, 25 (2020), 520–541. https://doi.org/10.1016/j.istruc.2020.03.033 doi: 10.1016/j.istruc.2020.03.033
    [5] F. Zhao, L. Zhang, J. Cao, A cooperative water wave optimization algorithm with reinforcement learning for the distributed assembly no-idle flowshop scheduling problem, Comput. Ind. Eng., 153 (2021), 107082. https://doi.org/10.1016/j.cie.2020.107082 doi: 10.1016/j.cie.2020.107082
    [6] H. Yapici, N. Cetinkaya, A new meta-heuristic optimizer: Pathfinder algorithm, Appl. Soft Comput., 78 (2019), 545–568. https://doi.org/10.1016/j.asoc.2019.03.012 doi: 10.1016/j.asoc.2019.03.012
    [7] S. H. S. Moosavi, V. K. Bardsiri, Poor and rich optimization algorithm: A new human-based and multi populations algorithm, Eng. Appl. Artif. Intell., 86 (2019), 165–181. https://doi.org/10.1016/j.engappai.2019.08.025 doi: 10.1016/j.engappai.2019.08.025
    [8] A.W. Mohamed, A. A. Hadi, A. K. Mohamed, Gaining-sharing knowledge based algorithm for solving optimization problems: A novel nature-inspired algorithm, Int. J. Mach. Learn. Cybern., 11 (2020), 1501–1529. https://doi.org/10.1007/s13042-019-01053-x doi: 10.1007/s13042-019-01053-x
    [9] Y. Liu, R. Li, PSA: A photon search algorithm. J. Inf. Process. Syst., 16 (2020), 478–493. https://doi.org/10.3745/JIPS.04.0168 doi: 10.3745/JIPS.04.0168
    [10] L. Pflug, N. Bernhardt, M. Grieshammer, CSG: A new stochastic gradient method for the efficient solution of structural optimization problems with infinitely many states, Struct. Multidisc. Optim., 61 (2020), 2595–2611. https://doi.org/10.1007/s00158-020-02571-x doi: 10.1007/s00158-020-02571-x
    [11] Y. Yang, Y. Gao, S. Tan, An opposition learning and spiral modelling based arithmetic optimization algorithm for global continuous optimization problems, Eng. Appl. Artif. Intell., 113 (2022), 104981. https://doi.org/10.1016/j.engappai.2022.104981 doi: 10.1016/j.engappai.2022.104981
    [12] A. T. Khan, X. Cao, S. Li, Dual beetle antennae search system for optimal planning and robust control of 5-link biped robots, J. Comput. Sci., 60 (2022), 1877–7503. https://doi.org/10.1016/j.jocs.2022.101556 doi: 10.1016/j.jocs.2022.101556
    [13] W. L. Liu, Y. J. Gong, W. N. Chen, Z. Liu, H. Wang, J. Zhang, Coordinated charging scheduling of electric vehicles: a mixed-variable differential evolution approach, IEEE Trans. Intell. Transp., 21 (2020), 5094–5109. https://doi.org/10.1109/TITS.2019.2948596 doi: 10.1109/TITS.2019.2948596
    [14] F. Zhao, R. Ma, L. Wang, A self-learning discrete jaya algorithm for multi-objective energy-efficient distributed no-idle flow-shop scheduling problem in heterogeneous factory system, IEEE Trans. Cybern., 247 (2021), 1–12. https://doi.org/10.1109/TCYB.2021.3086181 doi: 10.1109/TCYB.2021.3086181
    [15] L. Pflug, N. Bernhardt, M. Grieshammer, CSG: A new stochastic gradient method for the efficient solution of structural optimization problems with infinitely many states, Struct. Multidisc. Optim., 61 (2020), 2595–2611. https://doi.org/10.1007/s00158-020-02571-x doi: 10.1007/s00158-020-02571-x
    [16] G. Dhiman, M. Garg, MoSSE: A novel hybrid multi-objective meta-heuristic algorithm for engineering design problems, Soft Comput., 24 (2020), 18379–18398. https://doi.org/10.1007/s00500-020-05046-9 doi: 10.1007/s00500-020-05046-9
    [17] F. Zhao, S. Di, J. Cao, J. Tang, Jonrinaldi, A Novel Cooperative Multi-Stage Hyper-Heuristic for Combination Optimization Problems, Complex Syst. Model. Simul., 1 (2021), 91-108. https://doi.org/10.23919/CSMS.2021.0010 doi: 10.23919/CSMS.2021.0010
    [18] Y. Yang, C. Qian, H. Li, An efficient DBSCAN optimized by arithmetic optimization algorithm with opposition-based learning, J. Supercomput., 440 (2022). https://doi.org/10.1007/s11227-022-04634-w doi: 10.1007/s11227-022-04634-w
    [19] S. Mirjalili, A. Lewis, The whale optimization algorithm, Adv. Eng. Software, 95 (2016), 51–67. https://doi.org/10.1016/j.advengsoft.2016.01.008 doi: 10.1016/j.advengsoft.2016.01.008
    [20] S. Chakraborty, A. K. Saha, S. Sharma, S. Mirjalili, R. Chakraborty, A novel enhanced whale optimization algorithm for global optimization, Comput. Ind. Eng., 153 (2021), 107086. https://doi.org/10.1016/j.cie.2020.107086 doi: 10.1016/j.cie.2020.107086
    [21] Q. Jin, Z. Xu, W. Cai, An improved whale optimization algorithm with random evolution and special reinforcement dual-operation strategy collaboration, Symmetry, 13 (2021), 238–256. https://doi.org/10.3390/sym13020238 doi: 10.3390/sym13020238
    [22] J. Luo, H. Chen, A. A. Heidari, Y. Xu, Q. Zhang, C. Li, Multi-strategy boosted mutative whale-inspired optimization approaches, Appl. Math. Modell., 73 (2019), 109–123. https://doi.org/10.1016/j.apm.2019.03.046 doi: 10.1016/j.apm.2019.03.046
    [23] Y. Yang, Z. Tao, C. Qian, A hybrid robust system considering outliers for electric load series forecasting, Appl. Intell., 52 (2022), 1630–1652. https://doi.org/10.1007/s10489-021-02473-5 doi: 10.1007/s10489-021-02473-5
    [24] Q. Fan, Z. Chen, W. Zhang, ESSAWOA: Enhanced whale optimization algorithm integrated with salp swarm algorithm for global optimization, Eng. Comput., 20 (2020). https://doi.org/10.1007/s00366-020-01189-3 doi: 10.1007/s00366-020-01189-3
    [25] M. Nguyen, T. Tran, T. Nguyen, G. Nguyen, Hybridization of galactic swarm and evolution whale optimization for global search problem, IEEE Access, 8 (2020), 74991–75010. https://doi.org/10.1109/ACCESS.2020.2988717 doi: 10.1109/ACCESS.2020.2988717
    [26] M. Abdel-Basset, V. Chang, R. Mohamed, HSMA_WOA: A hybrid novel Slime mould algorithm with whale optimization algorithm for tackling the image segmentation problem of chest X-ray images, Appl. Soft Comput., 95 (2020), 1568–4946. https://doi.org/10.1016/j.asoc.2020.106642 doi: 10.1016/j.asoc.2020.106642
    [27] J. Zhang, L. Hong, Q. Liu, An improved whale optimization algorithm for the traveling salesman problem, Symmetry, 13 (2020), 48–56. https://doi.org/10.3390/sym13010048 doi: 10.3390/sym13010048
    [28] M. Wang, H. Chen, Chaotic multi-swarm whale optimizer boosted support vector machine for medical diagnosis, Appl. Soft Comput., 88 (2020), 105946. https://doi.org/10.1016/j.asoc.2019.105946 doi: 10.1016/j.asoc.2019.105946
    [29] V. Hemmelmayr, K. F. Doerner, R. F. Hartl, S. Rath, A heuristic solution method for node routing based solid waste collection problems, J. Heuristics, 19 (2013), 129–156. https://doi.org/10.1177/0734242X18801186 doi: 10.1177/0734242X18801186
    [30] F. Kılıç, Y. Kaya, S. Yildirim, A novel multi population based particle swarm optimization for feature selection, Knowl.-Based. Syst., 219 (2021), 126–140. https://doi.org/10.1016/j.knosys.2021.106894 doi: 10.1016/j.knosys.2021.106894
    [31] H. R. Tizhoosh, Opposition-based learning: A new scheme for machine intelligence, in International conference on computational intelligence for modelling, control and automation and international conference on intelligent agents, web technologies and internet commerce (CIMCA-IAWTIC'06), 1 (2005), 695–701. https://doi.org/10.1109/CIMCA.2005.1631345
    [32] S. Dhargupta, M. Ghosh, S. Mirjalili, R. Sarkar, Selective opposition based grey wolf optimization, Expert. Syst. Appl., 151 (2020), 113389. https://doi.org/10.1016/j.eswa.2020.113389 doi: 10.1016/j.eswa.2020.113389
    [33] M. H. Qais, H. M. Hasanien, S. Alghuwainem, Enhanced whale optimization algorithm for maximum power point tracking of variable-speed wind generators, Appl. Soft Comput., 86 (2020), 105937. https://doi.org/10.1016/j.asoc.2019.105937 doi: 10.1016/j.asoc.2019.105937
    [34] S. P. Adam, S. N. Alexandropoulos, P. M. Pardalos, M. N. Vrahatis, No free lunch theorem: A review, Optim. Appl., 145 (2019), 57–82. https://doi.org/10.1007/978-3-030-12767-1_5 doi: 10.1007/978-3-030-12767-1_5
    [35] Q. Fan, Z. Chen, Z. Li, A new improved whale optimization algorithm with joint search mechanisms for high-dimensional global optimization problems, Eng. Comput., 34 (2020), 1851–1878. https://doi.org/10.1007/s00366-019-00917-8 doi: 10.1007/s00366-019-00917-8
    [36] S. Mirjalili, SCA: A sine cosine algorithm for solving optimization problems, Knowl.-Based. Syst., 96 (2016), 120–133. https://doi.org/10.1016/j.knosys.2015.12.022 doi: 10.1016/j.knosys.2015.12.022
    [37] S. Mirjalili, S. M. Mirjalili, A. Lewis, Grey wolf optimizer, Adv. Eng. Software, 69 (2014), 46–61. https://doi.org/10.1016/j.advengsoft.2013.12.007 doi: 10.1016/j.advengsoft.2013.12.007
    [38] S. Mirjalili, Moth-flame optimization algorithm: A novel nature-inspired heuristic paradigm. Knowl.-Based. Syst., 89 (2015), 228–249. https://doi.org/10.1016/j.knosys.2015.07.006 doi: 10.1016/j.knosys.2015.07.006
    [39] H. Chen, W. Li, X. Yang, A whale optimization algorithm with chaos mechanism based on quasi-opposition for global optimization problems, Expert. Syst. Appl., 158 (2020), 23–46. https://doi.org/10.1016/j.eswa.2020.113612 doi: 10.1016/j.eswa.2020.113612
    [40] W. Qiao, Z. Yang, Z. Kang, Short-term natural gas consumption prediction based on Volterra adaptive filter and improved whale optimization algorithm, Eng. Appl. Artif. Intell., 87 (2020), 103323. https://doi.org/10.1016/j.engappai.2019.103323 doi: 10.1016/j.engappai.2019.103323
    [41] T. Tiwari, M. Saraswat, A new firefly algorithm-based superpixel clustering method for vehicle segmentation, Appl. Soft Comput., 278 (2022), 1–14. https://doi.org/10.1007/s00500-022-07206-5 doi: 10.1007/s00500-022-07206-5
    [42] A. Darwish, Bio-inspired computing: Algorithms review, deep analysis, and the scope of applications, Future Comput. Inf. J., 3 (2018), 231–246, https://doi.org/10.1016/j.fcij.2018.06.001 doi: 10.1016/j.fcij.2018.06.001
    [43] S. Zhou, A quantum image encryption method based on DNACNot, IEEE Access, 8 (2020), 178336–178344. https://doi.org/10.1109/ACCESS.2020.3027964 doi: 10.1109/ACCESS.2020.3027964
    [44] H. Hong, L. Wang, H. Ahmad, J. Li, Y. Yang, C. Wu, Construction of DNA codes by using algebraic number theory, Finite Fields Appl., 37 (2016), 328–343. https://doi.org/10.1016/j.ffa.2015.10.008 doi: 10.1016/j.ffa.2015.10.008
    [45] X. Wang, W. Xue, J. An, Image encryption algorithm based on LDCML and DNA coding sequence, Multimedia Tools Appl., 80 (2021), 591–614. https://doi.org/10.1007/s11042-020-09688-7 doi: 10.1007/s11042-020-09688-7
    [46] W. Song, K. Cai, M. Zhang, C. Yuen, Codes with run-length and GC-content constraints for DNA-based data storage, IEEE Commun. Lett., 22 (2018), 2004–2007. https://doi.org/10.1109/LCOMM.2018.2866566 doi: 10.1109/LCOMM.2018.2866566
    [47] T. T. Nguyen, K. Cai, K. A. S. Immink, Constrained coding with error control for DNA-based data storage, in 2020 IEEE International Symposium on Information Theory (ISIT), (2020), 694–699. https://doi.org/10.1109/ISIT44484.2020.9174438
    [48] Y. Wang, M. Noor-A-Rahim, E. Gunawan, Y. L. Guan, C. L. Poh, Thermodynamically stable DNA code design using a similarity significance model, in 2020 IEEE International Symposium on Information Theory (ISIT), (2020), 21–26. https://doi.org/10.1109/ISIT44484.2020.9174468
    [49] B. Cao, X. Li, X. Zhang, B. Wang, X. Wei, Designing uncorrelated address constrain for DNA storage by DMVO algorithm, IEEE/ACM Trans. Comput. Biol. Bioinf., 11 (2020), 866–877. https://doi.org/10.1109/TCBB.2020.3011582 doi: 10.1109/TCBB.2020.3011582
    [50] D. Limbachiya, M. K. Gupta, V. Aggarwal, Family of constrained codes for archival DNA data storage, IEEE Commun. Lett., 22 (2018), 1–10. https://doi.org/10.1109/LCOMM.2018.2861867 doi: 10.1109/LCOMM.2018.2861867
    [51] B. Cao, Q. Zhang, B. Wu, Q. Wang, X. Zhang, X. Wei, Minimum free energy coding for DNA storage, IEEE Trans. Nanobiosci., 20 (2021), 212–222. https://doi.org/10.1109/TNB.2021.3056351 doi: 10.1109/TNB.2021.3056351
    [52] M. Levy, E. Yaakobi, Mutually uncorrelated codes for DNA storage, IEEE Trans. Inf. Theory, 65 (2018), 3671–3691. https://doi.org/10.1109/TIT.2018.2873138 doi: 10.1109/TIT.2018.2873138
  • This article has been cited by:

    1. Bo You, Qing Xia, Continuous Data Assimilation Algorithm for the Two Dimensional Cahn–Hilliard–Navier–Stokes System, 2022, 85, 0095-4616, 10.1007/s00245-022-09863-2
    2. Yuan Pei, Regularity and Convergence Results of the Velocity-Vorticity-Voigt Model of the 3D Boussinesq Equations, 2021, 176, 0167-8019, 10.1007/s10440-021-00453-y
    3. Gülnur Haçat, Mine Akbas, Aytekin Çıbık, Analysis of continuous data assimilation scheme for the Navier–Stokes equations using variational multiscale method, 2023, 66, 18777503, 101914, 10.1016/j.jocs.2022.101914
    4. Trenton Franz, Adam Larios, Collin Victor, The bleeps, the sweeps, and the creeps: Convergence rates for dynamic observer patterns via data assimilation for the 2D Navier–Stokes equations, 2022, 392, 00457825, 114673, 10.1016/j.cma.2022.114673
    5. Özge KAZAR, Meryem KAYA, Global Existence and Uniqueness of The Inviscid Velocity-Vorticity Model of The g-Navier-Stokes Equations, 2022, 1301-4048, 10.16984/saufenbilder.1097179
    6. Elizabeth Carlson, Joshua Hudson, Adam Larios, Vincent R. Martinez, Eunice Ng, Jared P. Whitehead, Dynamically learning the parameters of a chaotic system using partial observations, 2022, 42, 1078-0947, 3809, 10.3934/dcds.2022033
    7. Elizabeth Carlson, Adam Larios, Sensitivity Analysis for the 2D Navier–Stokes Equations with Applications to Continuous Data Assimilation, 2021, 31, 0938-8974, 10.1007/s00332-021-09739-9
    8. Yu Chen Peng, Liang Chun Wu, Ming-Cheng Shiue, Finite time synchronization of the continuous/discrete data assimilation algorithms for Lorenz 63 system based on the back and forth nudging techniques, 2023, 20, 25900374, 100407, 10.1016/j.rinam.2023.100407
    9. Amanda E. Diegel, Xuejian Li, Leo G. Rebholz, Analysis of continuous data assimilation with large (or even infinite) nudging parameters, 2025, 456, 03770427, 116221, 10.1016/j.cam.2024.116221
    10. Xuejian Li, Wei Gong, Xiaoming He, Tao Lin, Variational Data Assimilation and its Decoupled Iterative Numerical Algorithms for Stokes–Darcy Model, 2024, 46, 1064-8275, S142, 10.1137/22M1492994
    11. Adam Larios, Collin Victor, Continuous data assimilation for the 3D and higher-dimensional Navier–Stokes equations with higher-order fractional diffusion, 2024, 540, 0022247X, 128644, 10.1016/j.jmaa.2024.128644
    12. Elizabeth Carlson, Adam Larios, Edriss S. Titi, Super-Exponential Convergence Rate of a Nonlinear Continuous Data Assimilation Algorithm: The 2D Navier–Stokes Equation Paradigm, 2024, 34, 0938-8974, 10.1007/s00332-024-10014-w
    13. Yongqing Zhao, Wenjun Liu, Guangying Lv, Yuepeng Wang, Continuous data assimilation for the three dimensional primitive equations with magnetic field, 2024, 140, 18758576, 77, 10.3233/ASY-241912
    14. Animikh Biswas, Joshua Hudson, Determining the viscosity of the Navier–Stokes equations from observations of finitely many modes, 2023, 39, 0266-5611, 125012, 10.1088/1361-6420/ad065f
    15. Mustafa Aggul, Aytekin Çıbık, Fatma G. Eroglu, Songül Kaya, Deferred correction method for the continuous data assimilation model, 2023, 415, 00457825, 116259, 10.1016/j.cma.2023.116259
    16. Aseel Farhat, Adam Larios, Vincent R. Martinez, Jared P. Whitehead, Identifying the body force from partial observations of a two-dimensional incompressible velocity field, 2024, 9, 2469-990X, 10.1103/PhysRevFluids.9.054602
    17. Aytekin Çıbık, Rui Fang, William Layton, Farjana Siddiqua, Adaptive parameter selection in nudging based data assimilation, 2025, 433, 00457825, 117526, 10.1016/j.cma.2024.117526
    18. Xi Li, Youcai Xu, Minfu Feng, A Pressure-Stabilized Continuous Data Assimilation Reduced Order Model for Incompressible Navier–Stokes Equations, 2025, 103, 0885-7474, 10.1007/s10915-025-02828-x
    19. Xin Li, Tian You, Tiaoxia Wei, Yang Liu, Continuous data assimilation for the 3D Nernst-Planck-Stokes system, 2025, 0, 1531-3492, 0, 10.3934/dcdsb.2025042
  • Reader Comments
  • © 2022 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

Metrics

Article views(2535) PDF downloads(125) Cited by(2)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog