Location via proxy:   [ UP ]  
[Report a bug]   [Manage cookies]                
Skip to content
Publicly Available Published by De Gruyter December 5, 2017

A Second-Order Time-Stepping Scheme for Simulating Ensembles of Parameterized Flow Problems

  • Max Gunzburger , Nan Jiang and Zhu Wang EMAIL logo

Abstract

We consider settings for which one needs to perform multiple flow simulations based on the Navier–Stokes equations, each having different initial condition data, boundary condition data, forcing functions, and/or coefficients such as the viscosity. For such settings, we propose a second-order time accurate ensemble-based method that to simulate the whole set of solutions, requires, at each time step, the solution of only a single linear system with multiple right-hand-side vectors. Rigorous analyses are given proving the conditional stability and establishing error estimates for the proposed algorithm. Numerical experiments are provided that illustrate the analyses.

MSC 2010: 65M60; 76D05

1 Introduction

Many computational fluid dynamics applications require multiple simulations of a flow under different input conditions. For example, the ensemble Kalman filter approach used in data assimilation first simulates a forward model a large number of times by perturbing either the initial condition data, boundary condition data, or uncertain parameters, then corrects the model based on the model forecasts and observational data. A second example is the construction of low-dimensional surrogates for partial differential equation (PDE) solutions such as sparse-grid interpolants or proper orthogonal decomposition approximations, for which one has to first obtain expensive approximations of solutions corresponding to several parameter samples. Another example is sensitivity analyses of solutions for which one often has to determine approximate solutions for a number of perturbed inputs such as the values of certain physical parameters. In this paper, we consider such applications and develop a second-order time-stepping scheme for efficiently simulating an ensemble of flows. In particular, we consider the setting in which one wishes to determine the PDE solutions for several different choices of initial condition and boundary condition data, forcing functions, and physical parameters appearing in the PDE model.

The ensemble algorithm was first developed in [18] to find a set of J solutions of the Navier–Stokes equations (NSE) subject to different initial condition and forcing functions. The main idea is that, based on the introduction of an ensemble average and a special semi-implicit time discretization, the discrete systems for the multiple flow simulations share a common coefficient matrix. Thus, instead of solving J linear systems with J different matrices and right-hand sides, one only need solve a single linear system with J right-hand side vectors. This leads to very significant computational savings in linear solver costs when LU factorization (for small-scale systems) or a block iterative algorithm (for large-scale systems) are used. High-order ensemble algorithms were designed in [15, 16]. For high Reynolds number flows, ensemble regularization methods and a turbulence model based on ensemble averaging have been developed in [15, 19, 17, 24]. The method has also been extended to simulate MHD flows in [23] and to develop ensemble-based reduced-order modeling techniques in [9, 8]. In [10], the authors proposed a first-order ensemble algorithm that deals with a number of flow simulations subject to not only different initial condition, boundary conditions, and/or body force data, but also distinct viscosity coefficients appearing in the NSE model. In this paper, we follow the same direction and develop an ensemble scheme having higher accuracy.

To begin, consider an ensemble of incompressible flow simulations on a bounded domain subject to Dirichlet boundary conditions. The j-th member of the ensemble is a simulation associated with the positive viscosity coefficient νj, initial condition data uj0, boundary condition data gj, and body force fj. All of these data may vary from one simulation to another. Then, for j=1,,J, we need to solve

(1.1)

uj,t+ujuj-νjuj+pj=fj(x,t)in Ω×[0,),
uj=0in Ω×[0,),
uj=gj(x,t)on Ω,
uj(x,0)=uj0(x)in Ω.

There is a long list of work in developing time discretization methods for the NSE including explicit, implicit, and semi-implicit schemes, for example, [11, 13, 12, 14, 20, 22, 25]. In general, explicit schemes are easier to implement, but they suffer the severely restricted time step size from stability requirement. The fully implicit and semi-implicit schemes have better stability conditions, but the discretization would lead to a varying coefficient matrix of the system. As a result, a different linear system has to be solved for each member at every time step, thus totally J linear systems need to be solved per time step. To overcome this issue, we propose a new, second-order accurate in time, ensemble-based scheme that improves the computational efficiency. The scheme is semi-implicit that permits the use of a known quantity (the ensemble mean defined below), which is independent of the ensemble index j, in the advection term and, therefore, leads to a single coefficient matrix for all the ensemble members.

For keeping the exposition simple, we consider a uniform time step Δt and let tn=nΔt for n=0,1,. We then consider the ensemble of semi-discrete in time systems

(1.2)3ujn+1-4ujn+ujn-12Δt+u¯nujn+1+ujn(2ujn-ujn-1)+pjn+1-ν¯Δujn+1-(νj-ν¯)Δ(2ujn-ujn-1)=fjn+1,ujn+1=0,

where ujn, pjn and fjn denote approximations of uj(,tn), pj(,tn) and fj(,tn) of (1.1), respectively. In (1.2), u¯n and ν¯ denote the ensemble means of the velocity field and viscosity coefficient, respectively, defined by

u¯n:=1Jj=1J(2ujn-ujn-1)andν¯:=1Jj=1Jνj

and ujn represents the fluctuation defined by

ujn=2ujn-ujn-1-u¯n.

It is easy to see that the coefficient matrix in the spatial discretization of (1.2) does not depend on j. Thus, all the members in the ensemble do share a common coefficient matrix. To advance one time step, one only need solve a single linear system with J right-hand side vectors, which is more efficient than solving J individual simulations.

We assume homogeneous flow boundary conditions (gj=0) in the following derivation and analysis of the proposed ensemble algorithm. But flows with inhomogeneous essential boundary conditions are considered in our first numerical experiment presented in Section 5, where, in the implementation, the data gj at each time step is first replaced by its interpolant on the Lagrangian finite element space and then is enforced on the boundary nodes. The extension of our analysis to the inhomogeneous cases will follow the idea presented in [3], which will be discussed elsewhere.

In what follows, we present a rigorous theoretical analysis of the proposed scheme. In Section 2, we provide some notations and preliminaries; in Section 3, the stability conditions of the scheme are obtained; and in Section 4, an error estimate is derived. Then, several numerical experiments are presented in Section 5.

2 Notation and Preliminaries

Let Ω be an open, regular domain in d(d=2 or 3). The space L2(Ω) is equipped with the norm and inner product (,). Denote by Lp and Wpk, respectively, the norms for Lp(Ω) and the Sobolev space Wpk(Ω). Let Hk(Ω) be the Sobolev space W2k(Ω) equipped with the norm k. For functions v(x,t) defined on (0,T), we define (1m<)

v,k:=EssSup[0,T]v(t,)kandvm,k:=(0Tv(t,)km𝑑t)1m.

Given a time step Δt, let vn=v(tn) and define the discrete norms

|||v|||,k=max0nNvnkand|||v|||m,k:=(n=0NvnkmΔt)1m.

Denote by H-1(Ω) the dual space of bounded linear functionals defined on

H01(Ω)={vH1(Ω):v=0 on Ω}.

A norm for H-1(Ω) is given by

f-1=sup0vH01(Ω)(f,v)v.

We choose the velocity space X and pressure space Q to be

X:=(H01(Ω))dandQ:=L02(Ω)={qL2(Ω):Ωq𝑑x=0}.

The space of weakly divergence free functions is then

V:={vX:(v,q)=0 for all qQ}.

A weak formulation of (1.1) reads: find uj:[0,T]X and pj:[0,T]Q for a.e. t(0,T] satisfying, for j=1,,J,

(uj,t,v)+(ujuj,v)+νj(uj,v)-(pj,v)=(fj,v)for all vX,
(uj,q)=0for all qQ,

with uj(x,0)=uj0(x).

For the spatial discretization, we use a finite element (FE) method. However, the results can be extended to many other variational methods without much difficulty. Denote by XhX and QhQ the conforming velocity and pressure FE spaces on an edge to edge triangulation of Ω with h denoting the maximum diameter of the triangles. Assume that the pair of spaces (Xh,Qh) satisfy the discrete inf-sup (or LBBh) condition, that is required to guarantee the stability of FE approximations. We also assume that the FE spaces satisfy the following approximation properties [21]:

(2.1)infvhXhv-vhChk+1vk+1for all v[Hk+1(Ω)]d,
(2.2)infvhXh(v-vh)Chkvk+1for all v[Hk+1(Ω)]d,
(2.3)infqhQhq-qhChs+1qs+1for all qHs+1(Ω),

where the generic constant C>0 is independent of the mesh size h. One example for which the LBBh stability condition is satisfied is the family of Taylor–Hood Ps+1-Ps element pairs (i.e., k=s+1 in the definition of Xh), for s1 [7]. The discrete divergence free subspace of Xh is

Vh:={vhXh:(vh,qh)=0 for all qhQh}.

We assume the mesh and FE spaces satisfy the following standard inverse inequality (typical for locally quasi-uniform meshes and standard FE spaces, see, e.g., [2]): for all vhXh,

(2.4)hvhC(inv)vh.

Define the explicitly skew-symmetric trilinear form

b(u,v,w):=12(uv,w)-12(uw,v)

that satisfies the bounds [21]

(2.5)b(u,v,w)C(Ω)(uu)12vwfor all u,v,wX,
(2.6)b(u,v,w)C(Ω)uv(ww)12for all u,v,wX,

where C(Ω) is a constant depending on the domain. Denote the exact solution to (1.1) and the FE approximate solution to (2.7) at t=tn by ujn and uj,hn, respectively.

The fully discrete finite element discretization of (1.2) at tn+1 is as follows: given uj,hn, find uj,hn+1Xh and pj,hn+1Qh satisfying

(2.7){(3uj,hn+1-4uj,hn+uj,hn-12Δt,vh)+b(u¯hn,uj,hn+1,vh)+b(2uj,hn-uj,hn-1-u¯hn,2uj,hn-uj,hn-1,vh)-(pj,hn+1,vh)+ν¯(uj,hn+1,vh)+(νj-ν¯)((2uj,hn-uj,hn-1),vh)=(fjn+1,vh),vhXh,(uj,hn+1,qh)=0,qhQh.

This is a two-step method that requires uj,h0 and uj,h1 to start the time stepping; uj,h0 is determined by the initial condition and uj,h1 can be computed by the first-order ensemble algorithm developed by the authors in [10] (which is locally second-order accurate) or by using a standard, non-ensemble time-stepping method to compute each individual simulation at the very first time step. Compared to the second-order ensemble scheme developed in [15] for the NSEs without variations in the viscosity coefficient, the scheme (2.7) for parametrized flows introduces an additional average of the viscosity coefficients for the parameterized flow ensemble. As is shown in the next section, the deviation of the flow viscosity coefficients from the ensemble average will play an important role in the stability analysis of the scheme.

3 Stability Analysis

We begin by proving the conditional, nonlinear, long time stability of (2.7) under conditions on the time step and viscosity coefficient deviation: for any j=1,,J, there exist 0μ<1 and 0<ϵ2-2μ such that

(3.1)CΔtν¯huj,hn2(2-2μ-ϵ)μ2(μ+ϵ),
(3.2)|νj-ν¯|ν¯μ3,

where C denotes a generic constant depending on the domain and the minimum angle of the mesh.

Theorem 3.1 (Stability).

The ensemble scheme (2.7) is stable provided conditions (3.1)–(3.2) hold. In particular, for j=1,,J and for any N2, we have

14(uj,hN2+2uj,hN-uj,hN-12)+18n=1N-1uj,hn+1-2uj,hn+uj,hn-12+ν¯Δtμ+ϵ2-μ(μ22+ϵμ+ϵ-3|νj-ν¯|2ν¯)uj,hN2
+ν¯Δt3μ+ϵ2-μ(μ22+ϵμ+ϵ-3|νj-ν¯|2ν¯)uj,hN-12
n=1N-1μ+ϵ2ϵ(2-μ)Δtν¯fjn+1-12+14(uj,h12+2uj,h1-uj,h02)+ν¯Δtμ+ϵ2-μ(μ22+ϵμ+ϵ-3|νj-ν¯|2ν¯)uj,h12
(3.3)+ν¯Δt3μ+ϵ2-μ(μ22+ϵμ+ϵ-3|νj-ν¯|2ν¯)uj,h02.

Proof.

See Appendix A. ∎

Remark 3.2.

Observe that the stability conditions (3.1) and (3.2) are oppositional to each other. The upper bound for the relative deviation of the viscosity coefficient given in (3.2) must be less than μ3 whereas the upper bound in the time-step condition (3.1) must be less than 1-μ because this bound is increasing when ϵ is decreasing, and it approaches 1-μ as ϵ0. In practice, condition (3.2) is easy to check. If it does not hold, one could split the ensemble into smaller groups so that this condition holds for each group. Condition (3.1) can be satisfied by adjusting the time-step size.

4 Error Analysis

In this section we derive the numerical error estimate of the proposed ensemble scheme (2.7). We first give a lemma on the estimate of the consistency error of the backward differentiation formula that will be used in the error analysis for the fully discrete ensemble scheme.

Lemma 4.1.

For any uH3(0,T;L2(Ω)), we have that

(4.1)3un+1-4un+un-12Δt-utn+1274Δt3tn-1tn+1uttt2𝑑t.

Proof.

The proof is given in Appendix B. ∎

By assuming that Xh and Qh satisfy the LBBh condition, the ensemble scheme (2.7) is equivalent to: forn=1,,N-1, finduj,hn+1Vhsuch that

(4.2)

(3uj,hn+1-4uj,hn+uj,hn-12Δt,vh)+b(u¯hn,uj,hn+1,vh)+ν¯(uj,hn+1,vh)+(νj-ν¯)((2uj,hn-uj,hn-1),vh)
+b(2uj,hn-uj,hn-1-u¯hn,2uj,hn-uj,hn-1,vh)=(fjn+1,vh),vhVh.

To analyze the rate of convergence of the approximation, we assume the regularity assumptions on the NSE given by

ujH2(0,T;Hk+1(Ω))H3(0,T;H1(Ω)),
pjL2(0,T;Hs+1(Ω)),
fjL2(0,T;L2(Ω)).

Let ejn=ujn-uj,hn be the error between the true solution of (1.1) and the approximate solution determined from (4.2). We then have the following error estimates.

Theorem 4.2 (Error Estimate).

For any j=1,,J, under the stability conditions of (3.1)–(3.2) for some μ and ϵ satisfying 0μ<1 and 0<ϵ2-2μ, there exists a positive constant C independent of the time step Δt and mesh size h such that

14ejN2+C1ν¯ΔtejN2eCTν¯3{14(ej12+2ej1-ej02)+C1ν¯Δtej12+C2ν¯Δtej02+Cν¯-1h2k|||uj|||4,k+14
+Cν¯-1h2k+CΔt4|νj-ν¯|2ν¯uj,tt2,02+Cν¯-1Δt4uj,tt2,02+Cν¯-1h2k|||uj|||2,k+12
+ChΔt3uj,tt2,02+Ch2k+1Δt3uj,tt2,k+12+Cν¯-1h2s+2|||pj|||2,s+12
+Cν¯-1h2k+2uj,t2,k+12+Cν¯h2k|||uj|||2,k+12+C|νj-ν¯|2ν¯h2k|||uj|||2,k+12
+Cν¯-1Δt4uj,ttt2,02}+Ch2k+2|||uj|||2,k+1+Cν¯h2kΔt|||uj|||2,k+1,

with positive constants

C1=2C0+μ+ϵ2-μ(1-17C0-3|νj-ν¯|2ν¯),C2=C0+μ+ϵ3(2-μ)(1-17C0-3|νj-ν¯|2ν¯),C0=117ϵμ+ϵ(1-μ2).

Proof.

See Appendix C. ∎

It is well known that the Taylor–Hood Ps+1-Ps element pairs for which the LBBh stability condition and the approximation properties (2.1)–(2.3) are all satisfied [2, 7, 21]. In particular, when the popular P2-P1 Taylor–Hood FE is used (i.e., k=2 and s=1 in the definitions of Xh and Qh, respectively), we have the following optimal convergence results.

Corollary 4.2.1.

Suppose the P2-P1 Taylor–Hood FE pair is used for the spatial discretization and assume that the initial errors uj0-uj,h0, (uj0-uj,h0), uj1-uj,h1 and (uj1-uj,h1) are all at least O(h2) accurate. Then the approximation error of the ensemble scheme (2.7) at time tN satisfies

14ujN-uj,hN2+2C0ν¯Δt(ujN-uj,hN)2𝒪(h4+Δt4+hΔt3).

5 Numerical Experiments

The goal of the numerical experiments is two-fold:

  1. to numerically illustrate the convergence rate of the ensemble algorithm (2.7), that is, illustrate the second-order accuracy in time.

  2. to explore the stability of the algorithm.

In particular, the numerical results strongly indicate that the stability condition (3.2) is sharp.

5.1 Convergence Test

We illustrate the convergence rate of (2.7) by considering a test problem for the NSE from [6] that has an analytical solution. This solution preserves the spatial patterns of the Green–Taylor solution [1, 5] but the vortices do not decay as t. On the unit square Ω=[0,1]2, we define

uref=[-s(t)cosxsiny,s(t)sinxcosy]andpref=-14[cos(2x)+cos(2y)]s2(t)

with s(t)=sin(2t). We then have the corresponding source term

fref(x,y,t)=(s(t)+2νs(t))[-cosxsiny,sinxcosy]

and an inhomogeneous Dirichlet boundary condition with data gref0(x,y,t)=uref(x,y,t) for (x,y)Ω and zero initial condition data uref0(x,y)=uref(x,y,0)=[0,0].

To illustrate the convergence behavior, we consider an ensemble of two members with different viscosity coefficients and perturbed initial conditions. For the first member, we choose the viscosity coefficient ν1=0.2 and the exact solution is chosen as u1=(1+ϵ)uref whereas for the second member, we choose ν2=0.3 and u2=(1-ϵ)uref, where ϵ=10-3. The initial condition data, boundary condition data, and source terms are adjusted accordingly.

For this choice of parameters, we have |νj-ν¯|ν¯=15 for both j=1 and j=2; hence the stability condition (3.2) is satisfied. We first apply the ensemble algorithm (2.7) using the P2-P1 Taylor–Hood FE and evaluate the rates of convergence. The initial mesh size and time step size are chosen to be h=0.1 and Δt=0.05; both the spatial and temporal grids are uniformly refined. Numerical results are listed in Table 1 for which

jE,0=max0nNujn-uj,hnandjE2,0=Δtn=0N(ujn-uj,hn)2.

It is seen that the convergence rates for both u1 and u2 are second order, which matches our theoretical analysis.

Furthermore, we implement the two individual simulations separately and denote the corresponding numerical errors by jS,0 and jS2,0. Comparing the ensemble simulation solutions in Table 1 with the independent simulation results listed in Table 2, we observe that the former achieves the same order of accuracy as the latter.

Table 1

Approximation errors for ensemble simulations of two members with inputs ν1=0.2, u1=(1+10-3)urefand ν2=0.3, u2=(1-10-3)uref.

1h1E,0rate1E2,0rate2E,0rate2E2,0rate
101.02e-048.51e-048.02e-057.99e-04
202.60e-051.982.12e-042.002.03e-051.981.99e-042.00
406.54e-061.995.31e-052.005.12e-061.994.99e-052.00
801.64e-061.991.33e-052.001.28e-062.001.25e-052.00
Table 2

Approximation errors for two individual simulations: ν1=0.2, u1=(1+10-3)uref and ν2=0.3, u2=(1-10-3)uref.

1h1S,0rate1S2,0rate2S,0rate2S2,0rate
101.08e-048.79e-047.64e-057.79e-04
202.74e-051.982.20e-042.001.94e-051.981.94e-042.00
406.92e-061.995.50e-052.004.87e-061.994.85e-052.00
801.74e-061.991.38e-051.991.22e-062.001.21e-052.00
Table 3

Pressure approximation errors for the ensemble and individual simulations.

1h𝒫1E,0rate𝒫2E,0rate𝒫1S,0rate𝒫2S,0rate
102.09e-032.08e-032.08e-032.08e-03
205.27e-041.995.22e-041.995.27e-041.995.22e-041.99
401.32e-042.001.31e-041.991.32e-042.001.31e-041.99
803.30e-052.003.27e-052.003.30e-052.003.27e-052.00

Although the pressure error is not discussed in this paper, we determine the pressure approximation accuracy of the ensemble simulation using the same uniform mesh refinement strategy and then, in Table 3, provide results for 𝒫jE,0, the maximum values over all the time levels of the pressure errors in the L2 norm. Results for approximate solutions obtained by the ensemble method as well as through separate computations are given. It is observed that the ensemble-based scheme achieves second-order convergence in the pressure approximation and the associated numerical errors are nearly identical to those obtained from individual simulations, 𝒫jS,0.

5.2 Stability Tests

Two conditions, (3.1) and (3.2), guarantee the stability of the proposed scheme. Condition (3.2), in many applications, relates to the probability distribution of the uncertain physical parameters. This requirement on the parameter deviation ratio can be easily checked. If it is not fulfilled, one could divide the parameter sample set into smaller subsets so that it holds on each subset. Condition (3.1) depends on the nature of nonlinear problems. Its severity depends on the governing equations, domain, model parameters, initial/boundary conditions, forcing terms, etc. In practice, once condition (3.2) holds, condition (3.1) can be satisfied by making Δt sufficiently small and/or by dividing the ensemble into smaller ensembles. Of course, when the ensemble consists of high Reynolds number flows, this condition could easily fail due to the requirement of having an extremely small time-step size leading to a prohibitive computational cost. Condition (3.1) has been investigated extensively in [18, 19, 15]. Hence, in the following, we are mostly interested on the optimality of condition (3.2). However, we do consider the conditional stability due to (3.1) because we want to determine values of the parameter for which that condition is satisfied; this is not directly computable from (3.1) because of the generic constant appearing in that condition. Note that (3.2) contains no such constant so that we can directly study the sharpness of the condition.

We check the stability of our algorithm by using the problem of a flow between two offset circles (see, e.g., [18, 15, 19, 17]). The domain is a disk with a smaller off-center obstacle inside. By letting r1=1, r2=0.1, and c=(c1,c2)=(12,0), the domain is given by

Ω={(x,y):x2+y2r12 and (x-c1)2+(y-c2)2r22}.

The flow is driven by a counterclockwise rotational body force

(5.1)f(x,y,t)=[-6y(1-x2-y2),6x(1-x2-y2)]

with no-slip boundary conditions imposed on both circles. A von Kármán vortex street forms behind the inner circle and then re-interacts with that circle and with itself, generating complex flow patterns. We consider multiple numerical simulations of the flow with different viscosity coefficients using the ensemble-based algorithm (2.7). For spatial discretization, we apply the P2-P1 Taylor–Hood element pair on a triangular mesh that is generated by a Delaunay triangulation with 80 mesh points on the outer circle and 60 mesh points on the inner circle and with refinement near the inner circle, resulting in 18,638 degrees of freedom; see Figure 1.

Figure 1 Mesh for the flow between two offset cylinders.
Figure 1

Mesh for the flow between two offset cylinders.

In order to illustrate the stability analysis based on conditions (3.1) and (3.2), we design two numerical tests involving two different sets of viscosity coefficients within an ensemble of three members, keeping the rest of computational setting, including the initial and boundary conditions and body force, the same for all the members. In particular, the initial condition is generated by solving the steady Stokes problem with viscosity ν=0.03 and the same body force f(x,y,t) given by (5.1). We have two test cases:

  1. Case 1: ν1=0.021, ν2=0.030, ν3=0.039.

  2. Case 2: ν1=0.019, ν2=0.030, ν3=0.041.

Note that the viscosity coefficients ν1 and ν3 for Case 2 are obtained by making small perturbations from those for Case 1 with the average of the viscosity coefficients ν¯=0.03 being the same for in both cases. However, the stability condition (3.2) holds in the first case but breaks down in the second case. In fact, the parameter deviation ratios are given by

  1. Case 1: |ν1-ν¯|ν¯=310<13, |ν2-ν¯|ν¯=0<13, |ν3-ν¯|ν¯=310<13.

  2. Case 2: |ν1-ν¯|ν¯=1130>13, |ν2-ν¯|ν¯=0<13, |ν3-ν¯|ν¯=1130>13.

For the stability test, we use the kinetic energy as a criterion and compare the ensemble simulation results to the independent simulations using the same mesh and time-step size.

Case 1.

Condition (3.2) is satisfied so that this case illustrates the conditional stability due to (3.1). As mentioned above, we also use this test to determine a value for Δt for which (3.1) is satisfied so that, in Case 2, we can study the sharpness of condition (3.2). We first test the ensemble-based algorithm at a large time step size Δt=0.5. The corresponding evolutions of the energy of all the three members are plotted in Figure 2. It is seen that for Δt=0.5, the algorithm is unstable because the energy of the third member increases dramatically after t=4 and that of the first member after t=4.5. Although not shown in this figure, the energy of the second member also blows up but not until after t=20. This implies that the stability condition (3.1) does not hold. Therefore, we next decrease the time step size to Δt=0.05 and re-run the ensemble simulations. The associated evolutions of the energies are shown in Figure 2, indicating that the algorithm is now stable over the same time interval. Indeed, additional numerical experiments show that, for any time step size smaller than 0.05, the algorithm is always stable in Case 1 over a much longer time interval, for instance, [0,50]. When an even smaller time step size Δt=0.01 is selected, the comparison of the energy evolutions of ensemble-based simulations with the corresponding independent simulations over the time interval [0,5] is given in Figures 3. The ensemble simulation is obviously stable and the output energy approximations are very close to those of the independent simulations.

Figure 2 For the flow between two offset cylinders, Case 1, the energy evolution of the ensemble simulations for Δ⁢t=0.5{\Delta t=0.5}and Δ⁢t=0.05{\Delta t=0.05}.
Figure 2

For the flow between two offset cylinders, Case 1, the energy evolution of the ensemble simulations for Δt=0.5and Δt=0.05.

Figure 3 For the flow between two offset cylinders, Case 1, the energy evolution of the ensemble (Ens.) and independentsimulations (Ind.) for Δ⁢t=0.01{\Delta t=0.01}.
Figure 3

For the flow between two offset cylinders, Case 1, the energy evolution of the ensemble (Ens.) and independentsimulations (Ind.) for Δt=0.01.

Case 2.

We run ensemble simulations using the small time step size Δt=0.01 over the same time interval as that for Case 1. As we mentioned above, the viscosity coefficients in Case 2 are obtained by slightly perturbing those in Case 1; this is the only difference between the two computational settings. Since Δt is chosen small, we believe condition (3.1) still holds for Case 2. But condition (3.2) no longer holds. Therefore, we expect the ensemble simulation to be unstable even when using the small time step size Δt=0.01. The plots of energy evolutions in Figure 4 matches our expectation as they clearly indicate that the ensemble simulation is unstable in this case. In fact, the energy of the third member blows up after t=1.95 and then affects the other two members and results in their energy dramatically increasing after t=2.45.

Figure 4 For the flow between two offset cylinders, Case 2, the energy evolution of the ensemble (Ens.) and independent simulations (Ind.) for Δ⁢t=0.01{\Delta t=0.01}.
Figure 4

For the flow between two offset cylinders, Case 2, the energy evolution of the ensemble (Ens.) and independent simulations (Ind.) for Δt=0.01.

6 Conclusions

In this paper, we develop a second-order time-stepping ensemble scheme to compute a set of Navier–Stokes equations in which every member is subject to an independent computational setting including a distinct viscosity coefficient, initial condition data, boundary condition data, and/or body force. By using the ensemble algorithm, all ensemble members share a common coefficient matrix after discretization, although with different right-hand side vectors. Therefore, many efficient block iterative solvers such as the block CG and block GMRES can be applied to solve such a single linear system with multiple right-hand side vectors, leading to great savings in both storage and simulation time. A rigorous analysis shows the proposed algorithm is conditionally, nonlinearly and long-time stable provided two explicit conditions hold and is second-order accurate in time. Two numerical experiments are presented that illustrate our theoretical analysis. In particular, the first is a test problem having an analytic solution that serves to illustrate that the rate of convergence with respect to the time-step size is indeed second order whereas the second example is for a flow between two offset cylinders and shows that the stability condition is sharp. For future work, we plan to investigate the performance of the ensemble algorithm in data assimilation applications.

Funding statement: This research was partially supported by the U.S. Department of Energy under grants DE-SC0009324 and DE-SC0016540, the U.S. Air Force Office of Scientific Research grant FA9550-15-1-0001, a Defense Advanced Projects Agency contract administered under the Oak Ridge National Laboratory subcontract 4000145366, the U.S. National Science Foundation grants DMS-1522672 and DMS-1720001, and a University of Missouri Research Board grant.

A Proof of Theorem 3.1

Proof.

Setting vh=uj,hn+1 and qh=pj,hn+1 in (2.7) and multiplying the result by Δt gives

14(uj,hn+12+2uj,hn+1-uj,hn2)-14(uj,hn2+2uj,hn-uj,hn-12)+14uj,hn+1-2uj,hn+uj,hn-12
+ν¯Δtuj,hn+12+Δtb*(2uj,hn-uj,hn-1-u¯hn,2uj,hn-uj,hn-1,uj,hn+1)
=Δt(fjn+1,uj,hn+1)-(νj-ν¯)Δt((2uj,hn-uj,hn-1),uj,hn+1).

Applying Young’s inequality to the terms on the right-hand side yields, for any α,β1,β2>0,

(A.1)14(uj,hn+12+2uj,hn+1-uj,hn2)-14(uj,hn2+2uj,hn-uj,hn-12)+14uj,hn+1-2uj,hn+uj,hn-12+ν¯Δtuj,hn+12+Δtb*(2uj,hn-uj,hn-1-u¯hn,2uj,hn-uj,hn-1,uj,hn+1)αν¯Δt4uj,hn+12+Δtαν¯fjn+1-12+β1ν¯Δtuj,hn+12+(νj-ν¯)2Δtβ1ν¯uj,hn2+β2ν¯Δt4uj,hn+12+(νj-ν¯)2Δtβ2ν¯uj,hn-12.

Because the last four terms on the right-hand side of (A.1) need to be absorbed into ν¯Δtuj,hn+12 on the left-hand side, we minimize

β1ν¯Δtuj,hn+12+(νj-ν¯)2Δtβ1ν¯uj,hn2

by taking β1=|νj-ν¯|ν¯ and

β2ν¯Δt4uj,hn+12+(νj-ν¯)2Δtβ2ν¯uj,hn-12

by taking β2=2|νj-ν¯|ν¯. Then (A.1) becomes

(A.2)14(uj,hn+12+2uj,hn+1-uj,hn2)-14(uj,hn2+2uj,hn-uj,hn-12)+14uj,hn+1-2uj,hn+uj,hn-12+ν¯Δtuj,hn+12+Δtb*(2uj,hn-uj,hn-1-u¯hn,2uj,hn-uj,hn-1,uj,hn+1)αν¯Δt4uj,hn+12+Δtαν¯fjn+1-12+3|νj-ν¯|Δt2uj,hn+12+|νj-ν¯|Δtuj,hn2+|νj-ν¯|Δt2uj,hn-12.

Next, we bound the trilinear term using inequality (2.6) and the inverse inequality (2.4):

b*(2uj,hn-uj,hn-1-u¯hn,2uj,hn-uj,hn-1,uj,hn+1)=b*(2uj,hn-uj,hn-1-u¯hn,-uj,hn+1+2uj,hn-uj,hn-1,uj,hn+1)C(2uj,hn-uj,hn-1-u¯hn)uj,hn+1(uj,hn+1-2uj,hn+uj,hn-1)12uj,hn+1-2uj,hn+uj,hn-112Ch-12(2uj,hn-uj,hn-1-u¯hn)uj,hn+1uj,hn+1-2uj,hn+uj,hn-1.

Using Young’s inequality again gives

(A.3)Δt|b*(2uj,hn-uj,hn-1-u¯hn,2uj,hn-uj,hn-1,uj,hn+1)|CΔt2h(2uj,hn-uj,hn-1-u¯hn)2uj,hn+12+18uj,hn+1-2uj,hn+uj,hn-12.

Substituting (A.3) into (A.2) and combining like terms, we have

(A.4)14(uj,hn+12+2uj,hn+1-uj,hn2)-14(uj,hn2+2uj,hn-uj,hn-12)+18uj,hn+1-2uj,hn+uj,hn-12+ν¯Δt(1-α4-3|νj-ν¯|2ν¯)uj,hn+12Δtαν¯fjn+1-12+CΔt2h(2uj,hn-uj,hn-1-u¯hn)2uj,hn+12+|νj-ν¯|Δtuj,hn2+|νj-ν¯|Δt2uj,hn-12.

The second term on the right-hand side of (A.4), as well as the last two terms, need to be absorbed into the viscous term on the left-hand side. Thus we select an arbitrary number σ(0,1), decompose the positive viscous term into four parts, and move all the terms that need to be bounded on right-hand side to the left-hand side of the inequality, which gives

(A.5)14(uj,hn+12+2uj,hn+1-uj,hn2)-14(uj,hn2+2uj,hn-uj,hn-12)+18uj,hn+1-2uj,hn+uj,hn-12+ν¯Δtσ(1-α4-3|νj-ν¯|2ν¯)(uj,hn+12-uj,hn2)+ν¯Δt((1-σ)(1-α4-3|νj-ν¯|2ν¯)-CΔtν¯h(2uj,hn-uj,hn-1-u¯hn)2)uj,hn+12+ν¯Δt(23σ(1-α4-3|νj-ν¯|2ν¯)-|νj-ν¯|ν¯)uj,hn2+ν¯Δtσ3(1-α4-3|νj-ν¯|2ν¯)(uj,hn2-uj,hn-12)+ν¯Δt(σ3(1-α4-3|νj-ν¯|2ν¯)-|νj-ν¯|2ν¯)uj,hn-12Δtαν¯fjn+1-12.

Because α>0 is arbitrary, we take α=4-2(σ+1)σμ. To make sure that α is greater than 0, we need

σ>μ2-μ,where μ2-μ(0,1).

Now taking

σ=μ+ϵ2-μ,where ϵ(0,2-2μ),

inequality (A.5) becomes

(A.6)14(uj,hn+12+2uj,hn+1-uj,hn2)-14(uj,hn2+2uj,hn-uj,hn-12)+18uj,hn+1-2uj,hn+uj,hn-12+ν¯Δtσ(σ+12σμ-3|νj-ν¯|2ν¯)(uj,hn+12-uj,hn2)+ν¯Δt((1-σ)(σ+12σμ-3|νj-ν¯|2ν¯)-CΔtν¯h(2uj,hn-uj,hn-1-u¯hn)2)uj,hn+12+ν¯Δt((σ+1)(μ3-|νj-ν¯|ν¯))uj,hn2+ν¯Δtσ3(σ+12σμ-3|νj-ν¯|2ν¯)(uj,hn2-uj,hn-12)+ν¯Δtσ+12(μ3-|νj-ν¯|ν¯)uj,hn-12Δtαν¯fjn+1-12.

Stability follows if the following conditions hold:

σ+12σμ-3|νj-ν¯|2ν¯0,
(1-σ)(σ+12σμ-3|νj-ν¯|2ν¯)-CΔtν¯h(2uj,hn-uj,hn-1-u¯hn)20,
μ3-|νj-ν¯|ν¯0.

Under the assumption of (3.2), we have

μ3-|νj-ν¯|ν¯0andσ+12σμ-3|νj-ν¯|2ν¯μ(2-μ)2(μ+ϵ)0.

Together with the first assumption in (3.1), we have

(1-σ)(σ+12σμ-3|νj-ν¯|2ν¯)-CΔtν¯h(2uj,hn-uj,hn-1-u¯hn)2
(2-2μ-ϵ)μ2(μ+ϵ)-CΔtν¯h(2uj,hn-uj,hn-1-u¯hn)2
(2-2μ-ϵ)μ2(μ+ϵ)-(2-2μ-ϵ)μ2(μ+ϵ)=0.

Hence, we can draw the conclusion that the ensemble algorithm (2.7) is stable under conditions (3.1)–(3.2). Indeed, assuming both conditions (3.1)–(3.2) hold, (A.6) reduces to

(A.7)14(uj,hn+12+2uj,hn+1-uj,hn2)-14(uj,hn2+2uj,hn-uj,hn-12)+18uj,hn+1-2uj,hn+uj,hn-12+ν¯Δtμ+ϵ2-μ(μ22+ϵμ+ϵ-3|νj-ν¯|2ν¯)(uj,hn+12-uj,hn2)+ν¯Δt3μ+ϵ2-μ(μ22+ϵμ+ϵ-3|νj-ν¯|2ν¯)(uj,hn2-uj,hn-12)μ+ϵ2ϵ(2-μ)Δtν¯fjn+1-12.

Summing up (A.7) from n=1 to N-1 results in

14(uj,hN2+2uj,hN-uj,hN-12)+18n=1N-1uj,hn+1-2uj,hn+uj,hn-12+ν¯Δtμ+ϵ2-μ(μ22+ϵμ+ϵ-3|νj-ν¯|2ν¯)uj,hN2
+ν¯Δt3μ+ϵ2-μ(μ22+ϵμ+ϵ-3|νj-ν¯|2ν¯)uj,hN-12
n=1N-1μ+ϵ2ϵ(2-μ)Δtν¯fjn+1-12+14(uj,h12+2uj,h1-uj,h02)+ν¯Δtμ+ϵ2-μ(μ22+ϵμ+ϵ-3|νj-ν¯|2ν¯)uj,h12
(A.8)+ν¯Δt3μ+ϵ2-μ(μ22+ϵμ+ϵ-3|νj-ν¯|2ν¯)uj,h02.

This completes the proof. ∎

B Proof of Lemma 4.1

Proof.

To prove (4.1), we first rewrite

3(un+1-un)-(un-un-1)-2Δtutn+1
=3tntn+1ut𝑑t-tn-1tnut𝑑t-2Δtutn+1
=3([(t-tn)ut]tntn+1-tntn+1(t-tn)utt𝑑t)-([(t-tn-1)ut]tn-1tn-tn-1tn(t-tn-1)utt𝑑t)-2Δtutn+1
=3Δtutn+1-Δtutn-2Δtutn+1-3tntn+1ddt(12(t-tn)2)utt𝑑t+tn-1tnddt(12(t-tn-1)2)utt𝑑t
=Δtutn+1-Δtutn-3tntn+1ddt(12(t-tn)2)utt𝑑t+tn-1tnddt(12(t-tn-1)2)utt𝑑t
=Δttntn+1utt𝑑t-3([12(t-tn)2utt]tntn+1-tntn+112(t-tn)2uttt𝑑t)
+([12(t-tn-1)2utt]tn-1tn-tn-1tn12(t-tn-1)2uttt𝑑t)
=Δt([(t-tn)utt]tntn+1-tntn+1(t-tn)uttt𝑑t)-3(12Δt2uttn+1-tntn+112(t-tn)2uttt𝑑t)
+(12Δt2uttn-tn-1tn12(t-tn-1)2uttt𝑑t)
=(Δt2uttn+1-Δttntn+1(t-tn)uttt𝑑t)-3(12Δt2uttn+1-tntn+112(t-tn)2uttt𝑑t)
+(12Δt2uttn-tn-1tn12(t-tn-1)2uttt𝑑t)
=-12Δt2(uttn+1-uttn)-Δttntn+1(t-tn)uttt𝑑t+3tntn+112(t-tn)2uttt𝑑t-tn-1tn12(t-tn-1)2uttt𝑑t
=-12Δt2tntn+1uttt𝑑t-Δttntn+1(t-tn)uttt𝑑t+3tntn+112(t-tn)2uttt𝑑t-tn-1tn12(t-tn-1)2uttt𝑑t.

Then the L2 norm of the term of interest can be estimated as follows:

3un+1-4un+un-12Δt-utn+12
=14Δt2Ω|-12Δt2tntn+1uttt𝑑t-Δttntn+1(t-tn)uttt𝑑t+3tntn+112(t-tn)2uttt𝑑t-tn-1tn12(t-tn-1)2uttt𝑑t|2𝑑x
12Δt2Ω(14Δt4|tntn+1utttdt|2+Δt2|tntn+1(t-tn)utttdt|2+94|tntn+1(t-tn)2utttdt|2
+14|tn-1tn(t-tn-1)2utttdt|2)dx
12Δt2Ω(14Δt4|tntn+1uttt𝑑t|2+Δt4[tntn+1|uttt|𝑑t]2+94Δt4[tntn+1|uttt|𝑑t]2+14Δt4[tn-1tn|uttt|𝑑t]2)𝑑x
12Δt2Ω(14Δt5tntn+1|uttt|2𝑑t+Δt5tntn+1|uttt|2𝑑t+94Δt5tntn+1|uttt|2𝑑t+14Δt5tn-1tn|uttt|2𝑑t)𝑑x
74Δt3Ωtn-1tn+1|uttt|2𝑑t𝑑x
74Δt3tn-1tn+1uttt2𝑑t.

This completes the proof. ∎

C Proof of Theorem 4.2

Proof.

The true solution (uj,pj) of the NSE satisfies

(3ujn+1-4ujn+ujn-12Δt,vh)+b*(ujn+1,ujn+1,vh)+νj(ujn+1,vh)-(pjn+1,vh)
(C.1)=(fjn+1,vh)+Intp(ujn+1;vh)for all vhVh,

where Intp(ujn+1;vh) is defined as

Intp(ujn+1;vh)=(3ujn+1-4ujn+ujn-12Δt-uj,t(tn+1),vh).

Let

ejn=ujn-uj,hn=(ujn-Ihujn)+(Ihujn-uj,hn)=ηjn+ξj,hn,

where IhujnVh is the FE interpolant of ujn in Vh. Subtracting (4.2) from (C.1) gives

(3ξj,hn+1-4ξj,hn+ξj,hn-12Δt,vh)+b*(ujn+1,ujn+1,vh)+ν¯(ξj,hn+1,vh)+(νj-ν¯)((2ξj,hn-ξj,hn-1),vh)-b*(2uj,hn-uj,hn-1-uj,hn,uj,hn+1,vh)-b*(uj,hn,2uj,hn-uj,hn-1,vh)-(pjn+1,vh)=-(3ηjn+1-4ηjn+ηjn-12Δt,vh)-ν¯(ηjn+1,vh)+Intp(ujn+1;vh)+(ν¯-νj)((2ηjn-ηjn-1),vh)+(ν¯-νj)((ujn+1-2ujn+ujn-1),vh).

Setting vh=ξj,hn+1Vh and rearranging the nonlinear terms leads to

14Δt(ξj,hn+12+2ξ-j,hn+1ξj,hn2)-14Δt(ξj,hn2+2ξj,hn-ξj,hn-12)+14Δtξj,hn+1-2ξj,hn+ξj,hn-12+ν¯ξj,hn+12
=-b*(ujn+1,ujn+1,ξj,hn+1)+b*(2uj,hn-uj,hn-1,uj,hn+1,ξj,hn+1)+b*(uj,hn,2uj,hn-uj,hn-1-uj,hn+1,ξj,hn+1)
+(pjn+1,ξj,hn+1)-(3ηjn+1-4ηjn+ηjn-12Δt,ξj,hn+1)-ν¯(ηjn+1,ξj,hn+1)+Intp(ujn+1;ξj,hn+1)
+(ν¯-νj)((2ξj,hn-ξj,hn-1),ξj,hn+1)+(ν¯-νj)((2ηjn-ηjn-1),ξj,hn+1)
(C.2)+(ν¯-νj)((ujn+1-2ujn+ujn-1),ξj,hn+1).

We first bound the viscous terms on the right-hand side of (C):

(C.3)-(νj-ν¯)((ujn+1-2ujn+ujn-1),ξj,hn+1)14C0|νj-ν¯|2ν¯(ujn+1-2ujn+ujn-1)2+C0ν¯ξj,hn+12Δt34C0|νj-ν¯|2ν¯tn-1tn+1uj,tt2𝑑t+C0ν¯ξj,hn+12,

and

(C.4)-ν¯(ηjn+1,ξj,hn+1)ν¯4C0ηjn+12+C0ν¯ξj,hn+12,
(C.5)-2(νj-ν¯)(ηjn,ξj,hn+1)1C0|νj-ν¯|2ν¯ηjn2+C0ν¯ξj,hn+12,
(C.6)(νj-ν¯)(ηjn-1,ξj,hn+1)14C0|νj-ν¯|2ν¯ηjn-12+C0ν¯ξj,hn+12,
(C.7)-2(νj-ν¯)(ξj,hn,ξj,hn+1)1C1|νj-ν¯|2ν¯ξj,hn2+C1ν¯ξj,hn+12|νj-ν¯|ξj,hn2+|νj-ν¯|ξj,hn+12,
(C.8)(νj-ν¯)(ξj,hn-1,ξj,hn+1)14C2|νj-ν¯|2ν¯ξj,hn-12+C2ν¯ξj,hn+12|νj-ν¯|2ξj,hn-12+|νj-ν¯|2ξj,hn+12,

where, because the terms on the right-hand side of (C.7) and (C.8) need to be hidden on the left-hand side of the error equation, we took C1=|νj-ν¯|ν and C2=|νj-ν¯|2ν in order to minimize their summations.

Next, we analyze the nonlinear terms on the right-hand side of (C) one by one. The first two nonlinear terms can be rewritten as

-b*(ujn+1,ujn+1,ξj,hn+1)+b*(2uj,hn-uj,hn-1,uj,hn+1,ξj,hn+1)
=-b*(2ejn-ejn-1,ujn+1,ξj,hn+1)-b*(2uj,hn-uj,hn-1,en+1,ξj,hn+1)-b*(ujn+1-2ujn+ujn-1,ujn+1,ξj,hn+1)
=-b*(2ηjn-ηjn-1,ujn+1,ξj,hn+1)-b*(2ξj,hn-ξj,hn-1,ujn+1,ξj,hn+1)-b*(2uj,hn-uj,hn-1,ηjn+1,ξj,hn+1)
(C.9)-b*(ujn+1-(2ujn-ujn-1),ujn+1,ξj,hn+1).

and

-b*(2ηjn-ηjn-1,ujn+1,ξj,hn+1)C(2ηjn-ηjn-1)ujn+1ξj,hn+1
C0ν¯ξj,hn+12+C24C0ν¯(ηjn2+ηjn-12)ujn+12.

Since ujL(0,T;H1(Ω)), we have the estimates

(C.10)-2b*(ξj,hn,ujn+1,ξj,hn+1)Cξj,hn12ξj,hn12ujn+1ξj,hn+1Cξj,hn12ξj,hn12ξj,hn+1C(ϵξj,hn+12+1ϵξj,hnξj,hn)C(ϵξj,hn+12+1ϵ(δξj,hn2+1δξj,hn2))C0ν¯ξj,hn+12+C0ν¯ξj,hn2+CC0-3ν¯-3ξj,hn2.

Similarly,

(C.11)b*(ξj,hn-1,ujn+1,ξj,hn+1)Cξj,hn-112ξj,hn-112ujn+1ξj,hn+1Cξj,hn-112ξj,hn-112ξj,hn+1C(ϵξj,hn+12+1ϵξj,hn-1ξj,hn-1)C(ϵξj,hn+12+1ϵ(δξj,hn-12+1δξj,hn-12))C0ν¯ξj,hn+12+C0ν¯ξj,hn-12+CC0-3ν¯-3ξj,hn-12.

Also by inequality (2.5) and the stability result (3.3), i.e. uj,hn2C, we have

(C.12)-2b*(uj,hn,ηjn+1,ξj,hn+1)Cuj,hn12uj,hn12ηjn+1ξj,hn+1C0ν¯ξj,hn+12+C24C0ν¯uj,hnηjn+12,
(C.13)b*(uj,hn-1,ηjn+1,ξj,hn+1)Cuj,hn-112uj,hn-112ηjn+1ξj,hn+1C0ν¯ξj,hn+12+C24C0ν¯uj,hn-1ηjn+12,
(C.14)-b*(ujn+1-(2ujn-ujn-1),ujn+1,ξj,hn+1)C(ujn+1-2ujn+ujn-1)ujn+1ξj,hn+1C0ν¯ξj,hn+12+C24C0ν¯(ujn+1-2ujn+ujn-1)2uj,hn+12C0ν¯ξj,hn+12+C24C0ν¯Δt3ujn+12tn-1tn+1uj,tt2𝑑t.

Now we bound the third nonlinear term in (C):

-b*(uj,hn,uj,hn+1-2uj,hn+uj,hn-1,ξj,hn+1)=b*(uj,hn,ejn+1-2ejn+ejn-1,ξj,hn+1)-b*(uj,hn,ujn+1-2ujn+ujn-1,ξj,hn+1)
=b*(uj,hn,ξj,hn+1-2ξj,hn+ξj,hn-1,ξj,hn+1)+b*(uj,hn,ηjn+1-2ηjn+ηjn-1,ξj,hn+1)
(C.15)-b*(uj,hn,ujn+1-2ujn+ujn-1,ξj,hn+1).

By skew symmetry,

-b*(uj,hn,ξj,hn+1-2ξj,hn+ξj,hn-1,ξj,hn+1)=b*(uj,hn,ξj,hn+1,ξj,hn+1-2ξj,hn+ξj,hn-1).

Using (2.6) and inverse inequality (2.4) gives

(C.16)b*(uj,hn,2ξj,hn-ξj,hn-1-ξj,hn+1,ξj,hn+1)Cuj,hnξj,hn+1(ξj,hn+1-2ξj,hn+ξj,hn-1)12ξj,hn+1-2ξj,hn+ξj,hn-112Cuj,hnξj,hn+1(h-1/2)ξj,hn+1-2ξj,hn+ξj,hn-118Δtξj,hn+1-2ξj,hn+ξj,hn-12+CΔthuj,hn2ξj,hn+12,
(C.17)b*(uj,hn,ηjn+1-2ηjn+ηjn-1,ξj,hn+1)Cuj,hn(ηjn+1-2ηjn+ηjn-1)ξj,hn+1C0ν¯ξj,hn+12+CC0-1ν¯-1uj,hn2(ηjn+1-2ηjn+ηjn-1)2C0ν¯ξj,hn+12+CΔt3C0ν¯uj,hn2tn-1tn+1ηj,tt2𝑑t,
(C.18)b*(uj,hn,ujn+1-2ujn+ujn-1,ξj,hn+1)Cuj,hn(ujn+1-2ujn+ujn-1)ξj,hn+1C0ν¯ξj,hn+12+CC0-1ν¯-1uj,hn2(ujn+1-2ujn+ujn-1)2C0ν¯ξj,hn+12+CC0-1ν¯-1Δt3uj,hn2tn-1tn+1uj,tt2𝑑t.

For the pressure term in (C), since ξj,hn+1Vh, we have

(pjn+1,ξj,hn+1)=(pjn+1-qj,hn+1,ξj,hn+1)
dpjn+1-qj,hn+1ξj,hn+1
(C.19)d4C0ν¯-1pjn+1-qj,hn+12+C0ν¯ξj,hn+12.

The other terms are bounded as

(3ηjn+1-4ηjn+ηjn-12Δt,ξj,hn+1)C4C0ν¯-13ηjn+1-4ηjn+ηjn-12Δt2+C0ν¯ξj,hn+12
C4C0ν¯-11Δttn-1tn+1ηj,t𝑑t2+C0ν¯ξj,hn+12
C4C0ν¯Δttn-1tn+1ηj,t2𝑑t+C0ν¯ξj,hn+12

and

(C.20)Intp(ujn+1;ξj,hn+1)=(3ujn+1-4ujn+ujn-12Δt-uj,t(tn+1),ξj,hn+1)C3ujn+1-4ujn+ujn-12Δt-uj,t(tn+1)ξj,hn+1C0ν¯ξj,hn+12+C24C0ν¯3ujn+1-4ujn+ujn-12Δt-uj,t(tn+1)2C0ν¯ξj,hn+12+5C2Δt38C0ν¯tn-1tn+1uj,ttt2𝑑t.

Combining (C.3)–(C.20) and taking C0=117ϵμ+ϵ(1-μ2) with ϵ(0,2-2μ), we have for all σ such that 0<σ<1,

14Δt(ξj,hn+12+2ξ-j,hn+1ξj,hn2)-14Δt(ξj,hn2+2ξj,hn-ξj,hn-12)+18Δtξj,hn+1-2ξj,hn+ξj,hn-12
+2C0ν¯(ξj,hn+12-ξj,hn2)+C0ν¯(ξj,hn2-ξj,hn-12)
+ν¯σ(1-17C0-3|νj-ν¯|2ν¯)(ξj,hn+12-ξj,hn2)
+ν¯((1-σ)(1-17C0-3|νj-ν¯|2ν¯)-CΔtν¯huj,hn2)ξj,hn+12
+ν¯(23σ(1-17C0-3|νj-ν¯|2ν¯)-|νj-ν¯|ν¯)ξj,hn2
+ν¯σ3(1-17C0-3|νj-ν¯|2ν¯)(ξj,hn2-ξj,hn-12)
+ν¯(σ3(1-17C0-3|νj-ν¯|2ν¯)-|νj-ν¯|2ν¯)ξj,hn-12
CC0-3ν-3(ξj,hn2+ξj,hn-12)+C4C0ν¯uj,hnηjn+12+C4C0ν¯uj,hn-1ηjn+12
+Δt34C0|νj-ν¯|2ν¯tn-1tn+1uj,tt2𝑑t+CΔt34C0ν¯ujn+12tn-1tn+1uj,tt2𝑑t
+C4C0ν¯(ηjn2+ηjn-12)ujn+12+CΔt3C0ν¯uj,hn2tn-1tn+1uj,tt2𝑑t
+CΔt3C0ν¯uj,hn2tn-1tn+1ηj,tt2𝑑t+d4C0ν¯pjn+1-qj,hn+12
+C4C0ν¯Δttn-1tn+1ηj,t2𝑑t+ν¯4C0ηjn+12+1C0|νj-ν¯|2ν¯ηjn2
(C.21)+14C0|νj-ν¯|2ν¯ηjn-12+CΔt34C0ν¯tn-1tn+1uj,ttt2𝑑t,

where C on the right-hand side is a generic constant independent of Δt and h. Similar to the discussion in the stability proof, we take

σ=μ+ϵ2-μ.

By the viscosity deviation condition (3.2), we have

1-17C0-3|νj-ν¯|2ν¯=(2+ϵ)μ2(μ+ϵ)-3|νj-ν¯|2ν¯>(2+ϵ)μ2(μ+ϵ)-μ2=μ(2-μ)2(μ+ϵ)>0,
23σ(1-17C0-3|νj-ν¯|2ν¯)-|νj-ν¯|ν¯>23μ+ϵ2-μμ(2-μ)2(μ+ϵ)-μ3=0,
13σ(1-17C0-3|νj-ν¯|2ν¯)-|νj-ν¯|2ν¯>0.

Also, by the stability condition (3.2), we have

(1-σ)(1-17C0-3|νj-ν¯|2ν¯)-CΔtν¯huj,hn2=2-2μ-ϵ2-μμ(2-μ)2(μ+ϵ)-Ctν¯huj,hn2
>(2-2μ-ϵ)μ2(μ+ϵ)-(2-2μ-ϵ)μ2(μ+ϵ)=0.

Then (C.21) reduces to

14Δt(ξj,hn+12+2ξ-j,hn+1ξj,hn2)-14Δt(ξj,hn2+2ξj,hn-ξj,hn-12)+18Δtξj,hn+1-2ξj,hn+ξj,hn-12
+C1ν¯(ξj,hn+12-ξj,hn2)+C2ν¯(ξj,hn2-ξj,hn-12)
CC0-3ν¯-3(ξj,hn2+ξj,hn-12)+C4C0ν¯uj,hnηjn+12+C4C0ν¯uj,hn-1ηjn+12
+Δt34C0|νj-ν¯|2ν¯tn-1tn+1uj,tt2𝑑t+CΔt34C0ν¯ujn+12tn-1tn+1uj,tt2𝑑t
+C4C0ν¯(ηjn2+ηjn-12)ujn+12+CΔt3C0ν¯uj,hn2tn-1tn+1uj,tt2𝑑t
+CΔt3C0ν¯uj,hn2tn-1tn+1ηj,tt2𝑑t+d4C0ν¯pjn+1-qj,hn+12
+C4C0ν¯Δttn-1tn+1ηj,t2𝑑t+ν¯4C0ηjn+12+1C0|νj-ν¯|2ν¯ηjn2
(C.22)+14C0|νj-ν¯|2ν¯ηjn-12+CΔt34C0ν¯tn-1tn+1uj,ttt2𝑑t,

where

C1=2C0+σ(1-17C0-3|νj-ν¯|2ν¯)andC2=C0+σ3(1-17C0-3|νj-ν¯|2ν¯).

Summing (C.22) from n=1 to N-1, multiplying both sides by Δt and absorbing constants gives

14(ξj,hN2+2ξj,hN-ξj,hN-12)+n=1N-118ξj,hn+1-2ξj,hn+ξj,hn-12+C1ν¯Δtξj,hN2+C2ν¯Δtξj,hN-12
14(ξj,h12+2ξj,h1-ξj,h02)+C1ν¯Δtξj,h12+C2ν¯Δtξj,h02+CΔtν¯3n=0N-1ξj,hn2
+CΔtn=1N-1{ν¯-1uj,hnηjn+12+ν¯-1uj,hn-1ηjn+12+Δt3|νj-ν¯|2ν¯tn-1tn+1uj,tt2dt
+Δt3ν¯-1ujn+12tn-1tn+1uj,tt2𝑑t+ν¯-1(ηjn2+ηjn-12)ujn+12
+Δt3ν¯-1uj,hn2tn-1tn+1uj,tt2𝑑t+Δt3ν¯-1uj,hn2tn-1tn+1ηj,tt2𝑑t
+ν¯-1pjn+1-qj,hn+12+ν¯-1Δt-1tn-1tn+1ηj,t2𝑑t+ν¯ηjn+12
+|νj-ν¯|2ν¯ηjn2+|νj-ν¯|2ν¯ηjn-12+Δt3ν¯-1tn-1tn+1uj,ttt2dt}.

Using the interpolation inequality (2.2) and the bound on the time average norm of uj,hn+1 obtained from a slightly changed stability analysis by, in (A.1), splitting out 18αν¯Δtuj,hn+12 from the viscosity term on the left-hand side and modifying the first two terms on the right-hand side to be 18αν¯Δtuj,hn+12 and 2Δtαν¯fjn+1-12, i.e., Δtn=1N-1uj,hn+12C, we have

ν¯-1Δtn=1N-1uj,hnηjn+12ν¯-1h2kΔtn=1N-1uj,hnujn+1k+12ν-1h2k(Δtn=1N-1ujn+1k+14+Δtn=1N-1uj,hn2)ν-1h2k|||uj|||4,k+14+Cν-1h2k,
ν¯-1Δtn=1N-1uj,hn-1ηjn+12ν¯-1h2kΔtn=1N-1uj,hn-1ujn+1k+12ν-1h2k(Δtn=1N-1ujn+1k+14+Δtn=1N-1uj,hn-12)ν-1h2k|||uj|||4,k+14+Cν-1h2k.

Because ujL(0,T;H1(Ω)), we have ujn+12C. Using convergence condition (3.1) and applying interpolation inequalities (2.1), (2.2) and (2.3) gives

14(ξj,hN2+2ξj,hN-ξj,hN-12)+n=1N-118ξj,hn+1-2ξj,hn+ξj,hn-12+C1ν¯Δtξj,hN2+C2ν¯Δtξj,hN-12
14(ξj,h12+2ξj,h1-ξj,h02)+C1ν¯Δtξj,h12+C2ν¯Δtξj,h02
+C[Δtν¯3n=0N-1ξj,hn2+ν¯-1h2k|||uj|||4,k+14+ν¯-1h2k+Δt4|νj-ν¯|2ν¯uj,tt2,02
+ν¯-1Δt4uj,tt2,02+ν¯-1h2k|||uj|||2,k+12+hΔt3uj,tt2,02
+h2k+1Δt3uj,tt2,k+12+ν¯-1h2s+2|||pj|||2,s+12+ν¯-1h2k+2uj,t2,k+12
+ν¯h2k|||uj|||2,k+12+|νj-ν¯|2ν¯h2k|||uj|||2,k+12+ν¯-1Δt4uj,ttt2,02].

The next step uses an application of the discrete Gronwall inequality (Girault and Raviart [4, p. 176]):

14(ξj,hN2+2ξj,hN-ξj,hN-12)+n=1N-118ξj,hn+1-2ξj,hn+ξj,hn-12+C1ν¯Δtξj,hN2+C2ν¯Δtξj,hN-12
eCTν¯3{14(ξj,h12+2ξj,h1-ξj,h02)+C1ν¯Δtξj,h12+C2ν¯Δtξj,h02
+C[ν¯-1h2k|||uj|||4,k+14+ν¯-1h2k+Δt4|νj-ν¯|2ν¯uj,tt2,02+ν¯-1Δt4uj,tt2,02+ν¯-1h2k|||uj|||2,k+12
+hΔt3uj,tt2,02+h2k+1Δt3uj,tt2,k+12+ν¯-1h2s+2|||pj|||2,s+12+ν¯-1h2k+2uj,t2,k+12
(C.23)+ν¯h2k|||uj|||2,k+12+|νj-ν¯|2ν¯h2k|||uj|||2,k+12+ν¯-1Δt4uj,ttt2,02]}.

Recall that ejn=ηjn+ξj,hn. Using the triangle inequality on the error equation to split the error terms into the terms of ηjn and ξj,hn gives

14ejN2+C1ν¯ΔtejN214ξj,hN2+C1ν¯Δtξj,hN2+14ηjN2+C1ν¯ΔtηjN2,

and

14(ξj,h12+2ξj,h1-ξj,h02)+C1ν¯Δtξj,h12+C2ν¯Δtξj,h02
14(ej12+2ej1-ej02)+C1ν¯Δtej12+C2ν¯Δtej02
+14(ηj12+2ηj1-ηj02)+C1ν¯Δtηj12+C2ν¯Δtηj02.

Applying inequality (C.23), using the previous bounds for the ηjn terms, and absorbing constants into a new constant C, we have Theorem 4.2. ∎

References

[1] L. C. Berselli, On the large eddy simulation of the Taylor–Green vortex, J. Math. Fluid Mech. 7 (2005), no. 2, S164–S191. 10.1007/s00021-005-0152-zSearch in Google Scholar

[2] S. C. Brenner and L. R. Scott, The Mathematical Theory of Finite Element Methods, 3rd ed., Texts Appl. Math. 15, Springer, New York, 2008. 10.1007/978-0-387-75934-0Search in Google Scholar

[3] G. J. Fix, M. D. Gunzburger and J. S. Peterson, On finite element approximations of problems having inhomogeneous essential boundary conditions, Comput. Math. Appl. 9 (1983), 687–700. 10.1016/0898-1221(83)90126-8Search in Google Scholar

[4] V. Girault and P.-A. Raviart, Finite Element Approximation of the Navier–Stokes Equations, Lecture Notes in Math. 749, Springer, Berlin, 1979. 10.1007/BFb0063447Search in Google Scholar

[5] A. Green and G. Taylor, Mechanism of the production of small eddies from larger ones, Proc. Royal Soc. Lond. Ser. A Math. Phys. Sci. 158 (1937), 499–521. 10.1098/rspa.1937.0036Search in Google Scholar

[6] J.-L. Guermond and L. Quartapelle, On stability and convergence of projection methods based on pressure Poisson equation, Internat. J. Numer. Methods Fluids 26 (1998), no. 9, 1039–1053. 10.1002/(SICI)1097-0363(19980515)26:9<1039::AID-FLD675>3.0.CO;2-USearch in Google Scholar

[7] M. D. Gunzburger, Finite Element Methods for Viscous Incompressible Flows, Comput. Sci. Sci. Computing, Academic Press, Boston, 1989. 10.1016/B978-0-12-307350-1.50009-0Search in Google Scholar

[8] M. Gunzburger, N. Jiang and M. Schneier, A higher-order ensemble/proper orthogonal decomposition method for the nonstationary Navier–Stokes equations, Int. J. Numer. Anal. Model., to appear; https://arxiv.org/abs/1709.06422. Search in Google Scholar

[9] M. Gunzburger, N. Jiang and M. Schneier, An ensemble-proper orthogonal decomposition method for the nonstationary Navier–Stokes equations, SIAM J. Numer. Anal. 55 (2017), no. 1, 286–304. 10.1137/16M1056444Search in Google Scholar

[10] M. Gunzburger, N. Jiang and Z. Wang, An efficient algorithm for simulating ensembles of parameterized flow problems, preprint (2017), https://arxiv.org/abs/1705.09350. 10.1093/imanum/dry029Search in Google Scholar

[11] Y. He, Two-level method based on finite element and Crank–Nicolson extrapolation for the time-dependent Navier–Stokes equations, SIAM J. Numer. Anal. 41 (2003), no. 4, 1263–1285. 10.1137/S0036142901385659Search in Google Scholar

[12] Y. He, The Euler implicit/explicit scheme for the 2D time-dependent Navier–Stokes equations with smooth or non-smooth initial data, Math. Comp. 77 (2008), no. 264, 2097–2124. 10.1090/S0025-5718-08-02127-3Search in Google Scholar

[13] Y. He and W. Sun, Stability and convergence of the Crank–Nicolson/Adams–Bashforth scheme for the time-dependent Navier–Stokes equations, SIAM J. Numer. Anal. 45 (2007), no. 2, 837–869. 10.1137/050639910Search in Google Scholar

[14] J. G. Heywood and R. Rannacher, Finite-element approximation of the nonstationary Navier–Stokes problem. IV. Error analysis for second-order time discretization, SIAM J. Numer. Anal. 27 (1990), no. 2, 353–384. 10.1137/0727022Search in Google Scholar

[15] N. Jiang, A higher order ensemble simulation algorithm for fluid flows, J. Sci. Comput. 64 (2015), no. 1, 264–288. 10.1007/s10915-014-9932-zSearch in Google Scholar

[16] N. Jiang, A second-order ensemble method based on a blended backward differentiation formula timestepping scheme for time-dependent Navier–Stokes equations, Numer. Methods Partial Differential Equations 33 (2017), no. 1, 34–61. 10.1002/num.22070Search in Google Scholar

[17] N. Jiang, S. Kaya and W. Layton, Analysis of model variance for ensemble based turbulence modeling, Comput. Methods Appl. Math. 15 (2015), no. 2, 173–188. 10.1515/cmam-2014-0029Search in Google Scholar

[18] N. Jiang and W. Layton, An algorithm for fast calculation of flow ensembles, Int. J. Uncertain. Quantif. 4 (2014), no. 4, 273–301. 10.1615/Int.J.UncertaintyQuantification.2014007691Search in Google Scholar

[19] N. Jiang and W. Layton, Numerical analysis of two ensemble eddy viscosity numerical regularizations of fluid motion, Numer. Methods Partial Differential Equations 31 (2015), no. 3, 630–651. 10.1002/num.21908Search in Google Scholar

[20] H. Johnston and J.-G. Liu, Accurate, stable and efficient Navier–Stokes solvers based on explicit treatment of the pressure term, J. Comput. Phys. 199 (2004), no. 1, 221–259. 10.1016/j.jcp.2004.02.009Search in Google Scholar

[21] W. Layton, Introduction to the Numerical Analysis of Incompressible Viscous Flows, Comput. Sci. Eng. 6, Society for Industrial and Applied Mathematics, Philadelphia, 2008. 10.1137/1.9780898718904Search in Google Scholar

[22] M. Marion and R. Temam, Navier–Stokes equations: Theory and approximation, Handbook of Numerical Analysis. VI, North-Holland, Amsterdam (1998), 503–688. 10.1016/S1570-8659(98)80010-0Search in Google Scholar

[23] M. Mohebujjaman and L. G. Rebholz, An efficient algorithm for computation of MHD flow ensembles, Comput. Methods Appl. Math. 17 (2017), no. 1, 121–137. 10.1515/cmam-2016-0033Search in Google Scholar

[24] A. Takhirov, M. Neda and J. Waters, Time relaxation algorithm for flow ensembles, Numer. Methods Partial Differential Equations 32 (2016), no. 3, 757–777. 10.1002/num.22024Search in Google Scholar

[25] F. Tone, Error analysis for a second order scheme for the Navier–Stokes equations, Appl. Numer. Math. 50 (2004), no. 1, 93–119. 10.1016/j.apnum.2003.12.003Search in Google Scholar

Received: 2017-04-09
Revised: 2017-11-07
Accepted: 2017-11-09
Published Online: 2017-12-05
Published in Print: 2019-07-01

© 2019 Walter de Gruyter GmbH, Berlin/Boston

Downloaded on 10.10.2024 from https://www.degruyter.com/document/doi/10.1515/cmam-2017-0051/html
Scroll to top button