This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

An efficient Chorin-Temam projection proper orthogonal decomposition based reduced-order model for nonstationary Stokes equations

Xi Li111School of Mathematics, Sichuan University, Chengdu, Sichuan 610064, China ([email protected]). The work of this author was supported by the National Natural Science Foundation of China(Grant No. 11971337).,  Yan Luo222School of Mathematical Sciences, University of Electronic Science and Technology of China, Chengdu, Sichuan 611731, China ([email protected]). The work of this author was supported by the Young Scientists Fund of the National Natural Science Foundation of China(Grant No. 11901078) and by the Fundamental Research Funds for the Central Universities(Grant No. ZYGX2020J021).  and Minfu Feng333Corresponding author. School of Mathematics, Sichuan University, Chengdu, Sichuan 610064, China ([email protected]). The work of this author was supported by the National Natural Science Foundation of China(Grant No. 11971337).
Abstract

In this paper, we propose an efficient proper orthogonal decomposition based reduced-order model(POD-ROM) for nonstationary Stokes equations, which combines the classical projection method with POD technique. This new scheme mainly owns two advantages: the first one is low computational costs since the classical projection method decouples the reduced-order velocity variable and reduced-order pressure variable, and POD technique further improves the computational efficiency; the second advantage consists of circumventing the verification of classical LBB/inf-sup condition for mixed POD spaces with the help of pressure stabilized Petrov-Galerkin(PSPG)-type projection method, where the pressure stabilization term is inherent which allows the use of non inf-sup stable elements without adding extra stabilization terms. We first obtain the convergence of PSPG-type finite element projection scheme, and then analyze the proposed projection POD-ROM’s stability and convergence. Numerical experiments validate out theoretical results.
Keywords: Projection method; PSPG stabilization; Proper orthogonal decomposition.

1 Introduction

Reduced Order Models(ROMs) have become widespread in the scientific community and owns constant attention in academic community because of its strength and efficiency in alleviating the huge computational costs needed in many modern engineering applications and academic purposes. As one of the most popular ROM approaches, the proper orthogonal decomposition(POD) strategy aims to use singular value decomposition(SVD) to extract the dominant modes(in the sense of energy) from a high-resolution numerical scheme(commonly referred to full order models(FOMs)) to form POD bases/modes. Onto these reduced-order bases, a Galerkin projection can be used to project into the space spanned by these bases/modes to obtain the reduced-order numerical solution, this pioneering work is done by Kunisch and Volkwein in [19, 20] which has laid the foundation of numerical analysis for Galerkin-POD. Other references about Galerkin-POD can be found in [30, 22, 21, 32] and so on. Applications of Galerkin-POD on incompressible flows, especially with the spatial variables discretized by finite element method(FE-POD), has a long history, from Luo et al.[22] obtaining FE-POD-ROM for nonstationary Navier-Stokes equations based on finite difference method and mixed finite element method to discretize temporal and spatial variables, Ravindran[26] analyzing FE-POD-ROM for nonstationary Boussinesq equations, to other finite element methods, like discontinuous Galerkin and hybrid discontinuous Galerkin, for spatial discrete are applied to get FE-POD-ROM[12].
Traditional mixed finite element POD-ROM had encountered a problem, that is, when considering the classical mixed finite element scheme as the snapshot source of the POD, the pressure variable is eliminated because of the classical LBB/inf-sup condition, so only the velocity finite element solution is used as the snapshot of the POD technique to construct the POD bases/modes, which makes the final obtained FE-POD-ROM does not contain the pressure variable. This makes it improper to use POD technique directly to reduce the model order in some engineering and academic models that need the pressure variables. Among that problem, the pressure field must be reconstructed a posterior using some pressure recovery techniques, like pressure Poisson equation[23], Supremizer stabilized technique[2] and so on. However, just as stated in [11] and [5], there are some deficiencies in these a posterior pressure recovery techniques, such as the usage of those techniques still need the fulfillment of LBB/inf-sup condition which restricts some flexible pairs of mixed finite element spaces, and the un-physical and unclear boundary condition of pressure Poisson equation poses a challenge for the utilization of this technique. With those in mind, recently, a new research direction related to FE-POD has aroused some researchers’ interest, we refer as stabilized FE-POD, which draws inspiration from the successful applications of stabilized techniques in finite element method. Caiazzo et al.[5] firstly proposed to combine stabilization technique, like residual-based streamline-upwind Petrov-Galerkin(SUPG) stabilization, with FE-POD to preserve the finite element pressure solution, so that the POD reduced-order pressure can be obtained from the snapshot of finite element pressure solution. From then on, other stabilized techniques were introduced into FE-POD, like local projection stabilization(LPS) FE-POD[27, 24], artificial compressible(AC) FE-POD[11], streamline diffusion(SD) FE-POD[1] and so on.
In this paper, our aim is to combine the classical projection method and POD technique to propose an efficient projection POD-ROM. As one of the fast decoupled numerical algorithms for solving incompressible fluid, projection method has become a classical one after it was proposed by Chorin and Temam [8, 31] and developed by Shen and Guermond [29, 15, 28, 14] etc. Apart from that, as mentioned by Rannacher in [25], another remarkable feature of this method is that the original Chorin-Temam projection method owns an inherent mechanism of pressure stability. This mechanism was proposed by Rannacher [25] and was realized by Frutos et al. [10]. These two advantages of the classical projection method are merged into the POD technique to get our projection POD-ROM. The main difference in our approach with respect to those stabilized FE-POD-ROM is that the pressure stabilized Petrov-Galerkin(PSPG)-like FE-POD-ROM has the inherent pressure stabilization, which means no additional pressure stabilization term is needed to circumvent the LBB/inf-sup condition of mixed POD spaces which is still an open problem to verify.
The main contribution of this paper is that a more efficient ROM numerical scheme for solving incompressible fluid is obtained by combining the advantageous computational efficiency of both projection method and POD technique. This POD-ROM scheme stems from replacing the finite element functions in the fully discrete scheme of the projection method with the POD reduced-order functions. To this end, we analyze the fully discrete scheme of classical projection method in the form of with inherent pressure stabilized term, and numerically verify some theoretical results concerning with this finite element scheme. Furthermore, we take the finite element solution as snapshots in order to use method of snapshots to construct POD modes and POD spaces, and obtain our projection POD scheme. The stability and convergence of this scheme are analyzed, and which are also validated by some numerical experiments.
The outline of this paper is as follows: In section 2, we introduce some essential notations. In section 3, the classical projection method will be presented, followed by its stability and convergence analysis. In section 4, we propose the projection POD numerical scheme and prove its stability and convergence. In section 5, some numerical experiments will be investigated to confirm the theoritical results analyzed before. Finally, in section 6, we draw conclusions to complete this paper.

2 Preliminaries and notations

we consider the nonstationary Stokes equations

{t𝒖νΔ𝒖+p=𝒇,inΩ,𝒖=0,inΩ,𝒖=𝟎,onΩ,𝒖(0,x)=𝒖0(x),inΩ,\left\{\begin{aligned} \partial_{t}\bm{u}-\nu\Delta\bm{u}+\nabla p&=\bm{f},\quad\text{in}\;\Omega,\\ \nabla\cdot\bm{u}&=0,\quad\text{in}\;\Omega,\\ \bm{u}&=\bm{0},\quad\text{on}\;\partial\Omega,\\ \bm{u}(0,x)&=\bm{u}_{0}(x),\quad\text{in}\;\partial\Omega,\end{aligned}\right. (2.1)

where Ω\Omega is an open bounded domain in 2\mathbb{R}^{2} with a sufficiently smooth boundary Ω\partial\Omega. The unknowns are the vector function 𝒖\bm{u}(velocity) and the scalar function pp(pressure).
In order to state the above problem properly, we need some settings. Let Lp(Ω), 1pL^{p}(\Omega),\;1\leq p\leq\infty denote the space of standard p-th power absolutely integrable functions with respect to the Lebesgue measure. In particular, L2(Ω)L^{2}(\Omega) is a Hilbert space endowed with the scalar product (,)(\cdot,\cdot) and its induced norm 0\|\cdot\|_{0}. Moreover, we use 2(L2)2:=Δtn=1N02\|\cdot\|^{2}_{\ell^{2}(L^{2})}:=\Delta t\sum^{N}_{n=1}\|\cdot\|^{2}_{0} to denote the discrete integral in time with respect to 02\|\cdot\|^{2}_{0}. The Sobolev space Wm,p(Ω)W^{m,p}(\Omega) and its norm m,p\|\cdot\|_{m,p} are also standard in the sense of [9]. Furthermore, we use the abbreviation Hm(Ω):=Wm,2(Ω)H^{m}(\Omega):=W^{m,2}(\Omega) and m:=m,2\|\cdot\|_{m}:=\|\cdot\|_{m,2}, it’s a Hilbert space with a scalar product. Moreover, we denote by H01(Ω)H^{1}_{0}(\Omega) the space of functions of H1(Ω)H^{1}(\Omega) with vanishing trace on Ω\partial\Omega and by H1(Ω)H^{-1}(\Omega) its dual space. The symbol ,\langle\cdot,\cdot\rangle in general denotes the duality pairing between the space and it’s dual one. Vector analogues of the Sobolev spaces along with vector valued functions are denoted by boldface letters, for instance 𝑯m(Ω):=(Hm(Ω))2\bm{H}^{m}(\Omega):=\left(H^{m}(\Omega)\right)^{2}. We all know that by Poincaré inequality[3, p.135], the seminorm |u|1=u0|u|_{1}=\|\nabla u\|_{0} in H1(Ω)H^{1}(\Omega) is a norm in H01(Ω)H^{1}_{0}(\Omega).
Throughout this paper, we use CC to denote a positive constant independent of Δt\Delta t, hh, not necessarily the same at each occurrence.
Let us introduce some convenient spaces. The first one is

𝑯(div;Ω):={𝒖𝑳2(Ω)|𝒖𝑳2(Ω)},\bm{H}({\rm div};\Omega):=\{\bm{u}\in\bm{L}^{2}(\Omega)|\;\nabla\cdot\bm{u}\in\bm{L}^{2}({\Omega})\},

which is a Hilbert space with norm 𝒖div:=𝒖0+𝒖0\|\bm{u}\|_{{\rm div}}:=\|\bm{u}\|_{0}+\|\nabla\cdot\bm{u}\|_{0}, and

𝑯0(div;Ω):={𝒖𝑯(div;Ω)|𝒖𝒏|Ω=0}.\bm{H}_{0}({\rm div};\Omega):=\{\bm{u}\in\bm{H}({\rm div};\Omega)|\;\bm{u}\cdot\bm{n}|_{\partial\Omega}=0\}.

For future use, we also need the following spaces, which play a key role in Helmholtz decomposition,

𝑱0:={𝒗𝑯01(Ω):𝒗=0},\bm{J}_{0}:=\left\{\bm{v}\in\bm{H}^{1}_{0}(\Omega):\;\nabla\cdot\bm{v}=0\right\},

and

𝑱1:={𝒗𝑳2(Ω):𝒗=0,and𝒗𝒏|Ω=0}.\bm{J}_{1}:=\left\{\bm{v}\in\bm{L}^{2}(\Omega):\;\nabla\cdot\bm{v}=0,\;\text{and}\;\bm{v}\cdot\bm{n}|_{\partial\Omega}=0\right\}.

By means of spaces above, we give the classical Helmholtz decomposition[13, p.29] as

𝑳2(Ω)=𝑱1(Ω)[𝑱1(Ω)]=𝑱1(Ω){q:qH1(Ω)}.\bm{L}^{2}(\Omega)=\bm{J}_{1}(\Omega)\oplus[\bm{J}_{1}(\Omega)]^{\perp}=\bm{J}_{1}(\Omega)\oplus\{\nabla q:q\in H^{1}(\Omega)\}. (2.2)

In order to give a variational formulation of problem (2.1), we consider the velocity space 𝑽:=𝑯01\bm{V}:=\bm{H}^{1}_{0} and the pressure space Q:=L02={qL2(Ω)|Ωqdx=0}Q:=L^{2}_{0}=\{q\in L^{2}(\Omega)|\int_{\Omega}q\,{\rm d}x=0\}. For a given time constant TT, a weak formulation of the nonstationary Stokes equations are expressed as follows: for almost t(0,T]\forall t\in(0,T], find (𝒖,p):(0,T]𝑽×Q(\bm{u},p):(0,T]\rightarrow\bm{V}\times Q, such that

{(t𝒖,𝒗)+ν(𝒖,𝒗)+(p,𝒗)=𝒇,𝒗,𝒗𝑽,(𝒖,q)=0,qQ.𝒖(x,0)=𝒖0(x).\left\{\begin{aligned} (\partial_{t}\bm{u},\bm{v})+\nu(\nabla\bm{u},\nabla\bm{v})+(\nabla p,\bm{v})&=\langle\bm{f},\bm{v}\rangle,\quad\forall\bm{v}\in\bm{V},\\ (\nabla\cdot\bm{u},q)&=0,\quad\forall q\in Q.\\ \bm{u}(x,0)&=\bm{u}^{0}(x).\end{aligned}\right. (2.3)

Let {𝒯h}\{\mathcal{T}_{h}\} be a uniformly regular family of triangulation of Ω¯\overline{\Omega}(see [9, p.111]) and h:=maxK𝒯h{hK|hK:=diam(K)}h:=\max_{K\in\mathcal{T}_{h}}\{h_{K}|\;h_{K}\\ :={\rm diam}(K)\}. Since we want to construct the POD space where snapshots are deriving from both the finite element approximated velocity and also pressure, contrary to the general choice which is to choose the inf-sup stable mixed finite element spaces, we choose equal-order mixed finite element spaces and then it’s obviously not stable in the sense of the classical inf-sup condition, as

Yhl:={vhC0(Ω¯)|vh|Kl(K),K𝒯h},l1,Y^{l}_{h}:=\{v_{h}\in C^{0}(\overline{\Omega})|\;v_{h}|_{K}\in\mathbb{P}_{l}(K),\;\forall K\in\mathcal{T}_{h}\},\;l\geq 1,

where l(K)\mathbb{P}_{l}(K) is the space of polynomials up to degree ll on KK. Since there are two types of velocities in the temporal semi-discrete projection scheme, i.e., intermediate velocity 𝒖~n+1\widetilde{\bm{u}}^{n+1} and end-of-step velocity 𝒖n+1\bm{u}^{n+1}, we need the three finite element spaces to locate three full discrete variables of three unknowns. To this end, we define the equal-order finite element spaces for end-of-step velocity, intermediate velocity and pressure

𝑽h:=𝒀hl𝑽,𝑾h:=𝒀hl𝑱1,Qh:=YhlQ.\bm{V}_{h}:=\bm{Y}^{l}_{h}\cap\bm{V},\quad\bm{W}_{h}:=\bm{Y}^{l}_{h}\cap\bm{J}_{1},\quad Q_{h}:=Y^{l}_{h}\cap Q.

Since the triangulations are assumed to be shape regular, the following inverse inequality holds for each 𝒗h𝑽h\bm{v}_{h}\in\bm{V}_{h} on each mesh cell K{𝒯h}K\in\{\mathcal{T}_{h}\}(see [3, Theorem 4.5.11]),

𝒗hm,p,KCinvhKlm+2(1p1q)𝒗hl,q,K,\|\bm{v}_{h}\|_{m,p,K}\leq C_{{\rm inv}}h^{l-m+2(\frac{1}{p}-\frac{1}{q})}_{K}\|\bm{v}_{h}\|_{l,q,K}, (2.4)

where 0lm10\leq l\leq m\leq 1, 0qp0\leq q\leq p\leq\infty and hkh_{k} is the size of the mesh cell K{𝒯h}K\in\{\mathcal{T}_{h}\}. Then, IhI_{h}(resp. JhJ_{h}) denotes a bounded linear interpolation operator Ih:𝑾t,q(K)𝑽hI_{h}:\;\bm{W}^{t,q}(K)\rightarrow\bm{V}_{h}(resp. Jh:Wt,q(K)QhJ_{h}:\;W^{t,q}(K)\rightarrow Q_{h}), where 1q1\leq q\leq\infty, 0stl+10\leq s\leq t\leq l+1, that satisfies the following optimal bounds(see [3, Theorem 4.4.4])

𝒗Ih𝒗s,q,K\displaystyle\|\bm{v}-I_{h}\bm{v}\|_{s,q,K} ChKts𝒗t,q,K,\displaystyle\leq Ch_{K}^{t-s}\|\bm{v}\|_{t,q,K}, (2.5)
qJhqs,q,K\displaystyle\|q-J_{h}q\|_{s,q,K} ChKtsqt,q,K,\displaystyle\leq Ch_{K}^{t-s}\|q\|_{t,q,K},

and also interpolation operator JhJ_{h} has following stability

Jhq0Cq0.\|\nabla J_{h}q\|_{0}\leq C\|\nabla q\|_{0}. (2.6)

The Sobolev inclusion L2(Ω)H1(Ω)L^{2}(\Omega)\hookrightarrow H^{-1}(\Omega) relation implies the following inequality holds:

v1Cv0,vL2(Ω).\|v\|_{-1}\leq C\|v\|_{0},\quad\forall v\in L^{2}(\Omega). (2.7)

3 Classical projection method

3.1 classical projection scheme

The original projection method consists of firstly finding an intermediate velocity 𝒖~n+1\widetilde{\bm{u}}^{n+1} from the momentum equation without the pressure term, and then utilizing the classical Helmholtz decomposition (2.2) to get an end-of-step velocity 𝒖n+1\bm{u}^{n+1} which is solenoidal and its normal component 𝒖n+1𝒏\bm{u}^{n+1}\cdot\bm{n} vanishes. Meanwhile, as a “by-product” of the above decomposition, a discrete pressure pn+1p^{n+1} can be also get for which some investigations have been made to confirm that this “by-product” pressure is indeed the proper approximation of the exact pressure p(tn+1)p(t_{n+1}), see [25].
For a given timestep Δt\Delta t, we consider a uniform discretization of the time interval [0,T][0,T] as 0=t0<t1<<tN0=t_{0}<t_{1}<\cdots<t_{N} where N=T/ΔtN=\lfloor T/\Delta t\rfloor is the rounding T/ΔtT/\Delta t down. In mathematics, the above procedure can be interpreted as the following classical Chorin-Teman projection method[7, 31]:
(Scheme 1 - Semi-discrete of classical projection method) For n=0,1,,N1n=0,1,\ldots,N-1,
Step 1- Given 𝒖n𝑱0\bm{u}^{n}\in\bm{J}_{0}, find 𝒖~n+1𝑽\widetilde{\bm{u}}^{n+1}\in\bm{V} that satisfies

𝒖~n+1𝒖nΔtνΔ𝒖~n+1=𝒇n+1,inΩ.\frac{\widetilde{\bm{u}}^{n+1}-\bm{u}^{n}}{\Delta t}-\nu\Delta\widetilde{\bm{u}}^{n+1}=\bm{f}^{n+1},\quad\text{in}\;\Omega.

Step 2- Given 𝒖~n+1𝑽\widetilde{\bm{u}}^{n+1}\in\bm{V}, compute (𝒖n+1,pn+1)𝑯0(div,Ω)×[QH1(Ω)](\bm{u}^{n+1},p^{n+1})\in\bm{H}_{0}({\rm div},\Omega)\times\left[Q\cap H^{1}(\Omega)\right] from

{𝒖n+1+Δtpn+1=𝒖~n+1,inΩ,𝒖n+1=0,inΩ,𝒖n+1𝒏=0,onΩ.\left\{\begin{aligned} \bm{u}^{n+1}+\Delta t\nabla p^{n+1}&=\widetilde{\bm{u}}^{n+1},\quad\text{in}\;\Omega,\\ \nabla\cdot\bm{u}^{n+1}&=0,\quad\text{in}\;\Omega,\\ \bm{u}^{n+1}\cdot\bm{n}&=0,\quad\text{on}\;\partial\Omega.\end{aligned}\right.

The usage of classical Helmholtz decomposition is reflected in decomposing the intermediate velocity 𝒖~n+1\widetilde{\bm{u}}^{n+1} from step 1 to get a solenoidal velocity 𝒖n+1\bm{u}^{n+1} and a gradient of a H1H^{1}-function, which can be stated as:
𝐒𝐭𝐞𝐩 2\mathbf{Step\ 2^{\prime}}- Projecting 𝒖~n+1\widetilde{\bm{u}}^{n+1} onto the space 𝑱0\bm{J}_{0}

𝒖n+1=P𝑱0(𝒖~n+1).\bm{u}^{n+1}=P_{\bm{J}_{0}}(\widetilde{\bm{u}}^{n+1}).
Remark 3.1.

Let us explain the relation between 𝐒𝐭𝐞𝐩𝟐\bm{Step2} and 𝐒𝐭𝐞𝐩𝟐\bm{Step2^{\prime}}. Followed by (2.2), decomposing 𝐮~n+1\widetilde{\bm{u}}^{n+1} would directly gets:

𝒖~n+1=𝒖n+1+ϕn+1,\widetilde{\bm{u}}^{n+1}=\bm{u}^{n+1}+\nabla\phi^{n+1}, (3.8)

where 𝐮n+1𝐉1\bm{u}^{n+1}\in\bm{J}_{1} implies 𝐮n+1=0\nabla\cdot\bm{u}^{n+1}=0 and 𝐮n+1𝐧|Ω=0\bm{u}^{n+1}\cdot\bm{n}|_{\partial\Omega}=0, which is exactly the second and third equations in 𝐒𝐭𝐞𝐩 2\bm{Step\ 2}. Multiplying the previous decomposition by 1/Δt1/\Delta t and adding it to the first equation in 𝐒𝐭𝐞𝐩 1\bm{Step\ 1}, we get (𝐮n+1𝐮n)/ΔtνΔ𝐮~n+1+(ϕn+1/Δt)=𝐟n+1(\bm{u}^{n+1}-\bm{u}^{n})/\Delta t-\nu\Delta\widetilde{\bm{u}}^{n+1}+\nabla(\phi^{n+1}/\Delta t)=\bm{f}^{n+1}, which means ϕn+1/Δt\phi^{n+1}/\Delta t can be regarded as a proper approximation of p(tn+1)p(t_{n+1}), i.e., we can define pn+1:=ϕn+1/Δtp^{n+1}:=\phi^{n+1}/\Delta t. Thus, on the one hand, we can compute ϕn+1\phi^{n+1} in the former decomposition (3.8) and using pn+1:=ϕn+1/Δtp^{n+1}:=\phi^{n+1}/\Delta t to get pn+1p^{n+1}; on the other hand, we can also substitute ϕn+1\phi^{n+1} with Δtpn+1\Delta tp^{n+1} in (3.8) to directly compute pn+1p^{n+1}, which is exactly what the first equation in step 2 does.

Or, Step 2 is equivalent to the following pressure Poisson equation using 𝒖n+1=0\nabla\cdot\bm{u}^{n+1}=0:
𝐒𝐭𝐞𝐩 2′′\mathbf{Step\ 2^{\prime\prime}}- Given 𝒖~n+1𝑽\widetilde{\bm{u}}^{n+1}\in\bm{V}, compute pn+1QH2(Ω)p^{n+1}\in Q\cap H^{2}(\Omega) from

{Δpn+1=1Δt𝒖~n+1,inΩ,pn+1𝒏=0,onΩ.\left\{\begin{aligned} \Delta p^{n+1}&=\frac{1}{\Delta t}\nabla\cdot\widetilde{\bm{u}}^{n+1},\quad\text{in}\;\Omega,\\ \frac{\partial p^{n+1}}{\partial\bm{n}}&=0,\quad\text{on}\;\partial\Omega.\end{aligned}\right.

Based on the previous finite element spaces, we derive the fully discrete scheme of projection method.
(Scheme 2 - Fully-discrete of classical projection scheme) Let 𝒖h0\bm{u}^{0}_{h} be the Lagrangian interpolation or Ritz projection onto 𝑽h\bm{V}_{h} of 𝒖0\bm{u}_{0}, then for n=0,1,,N1n=0,1,\ldots,N-1, we can compute (𝒖~hn+1,𝒖hn+1,phn+1)(\widetilde{\bm{u}}^{n+1}_{h},\bm{u}^{n+1}_{h},p^{n+1}_{h})\in 𝑽h×𝑾h×Qh\bm{V}_{h}\times\bm{W}_{h}\times Q_{h} iteratively by

{(𝒖~hn+1𝒖hnΔt,𝒗h)+ν(𝒖~hn+1,𝒗h)=𝒇n+1,𝒗h,𝒗h𝑽h,(𝒖~hn+1,qh)+Δt(phn+1,qh)=0,qhQh,𝒖hn+1=𝒖~hn+1Δtphn+1.\left\{\begin{aligned} \left(\frac{\widetilde{\bm{u}}^{n+1}_{h}-\bm{u}^{n}_{h}}{\Delta t},\bm{v}_{h}\right)+\nu(\nabla\widetilde{\bm{u}}^{n+1}_{h},\nabla\bm{v}_{h})&=\langle\bm{f}^{n+1},\bm{v}_{h}\rangle,\quad\forall\bm{v}_{h}\in\bm{V}_{h},\\ (\nabla\cdot\widetilde{\bm{u}}^{n+1}_{h},q_{h})+\Delta t(\nabla p^{n+1}_{h},\nabla q_{h})&=0,\quad\forall q_{h}\in Q_{h},\\ \bm{u}^{n+1}_{h}&=\widetilde{\bm{u}}^{n+1}_{h}-\Delta t\cdot\nabla p^{n+1}_{h}.\end{aligned}\right. (3.9)

As pointed in [25], it is not necessary to compute the projected velocities {𝒖hn}n=0N\{\bm{u}^{n}_{h}\}^{N}_{n=0} since these quantities can be eliminated. To see this, replacing 𝒖hn\bm{u}^{n}_{h} in the first equation with the third equation at tnt_{n} in Scheme 3, we reach the following Proj-PSPG-FE scheme.
(Scheme 3 - Proj-PSPG-FE) Let 𝒖h0\bm{u}^{0}_{h} be the Lagrangian interpolation or Ritz projection onto 𝑽h\bm{V}_{h} of 𝒖0\bm{u}_{0} and ph0p^{0}_{h} denote some proper approximation of initial pressure p(t0)p(t_{0}), we set 𝒖~h0=𝒖h0+Δtph0\widetilde{\bm{u}}^{0}_{h}=\bm{u}^{0}_{h}+\Delta t\nabla p^{0}_{h} and replace 𝒖hn\bm{u}^{n}_{h} with the third equality in Scheme 2, we obtain the following equivalent scheme: for n=0,1,,N1n=0,1,\ldots,N-1, find (𝒖~hn+1,phn+1)𝑽h×Qh(\widetilde{\bm{u}}^{n+1}_{h},p^{n+1}_{h})\in\bm{V}_{h}\times Q_{h}

{(𝒖~hn+1𝒖~hnΔt,𝒗h)+ν(𝒖~hn+1,𝒗h)+(phn,𝒗h)=𝒇n+1,𝒗h,𝒗h𝑽h,(𝒖~hn+1,qh)+Δt(phn+1,qh)=0,qhQh.\left\{\begin{aligned} \left(\frac{\widetilde{\bm{u}}^{n+1}_{h}-\widetilde{\bm{u}}^{n}_{h}}{\Delta t},\bm{v}_{h}\right)+\nu(\nabla\widetilde{\bm{u}}^{n+1}_{h},\nabla\bm{v}_{h})+(\nabla p^{n}_{h},\bm{v}_{h})&=\langle\bm{f}^{n+1},\bm{v}_{h}\rangle,\quad\forall\bm{v}_{h}\in\bm{V}_{h},\\ (\nabla\cdot\widetilde{\bm{u}}^{n+1}_{h},q_{h})+\Delta t(\nabla p^{n+1}_{h},\nabla q_{h})&=0,\quad\forall q_{h}\in Q_{h}.\end{aligned}\right. (3.10)
Remark 3.2.

For the initial discrete pressure ph0p^{0}_{h}, we simply take ph0=0p^{0}_{h}=0 in later numerical experiments, the reasons behind this are twofold: firstly, it can simplify the solving of initial discrete pressure ph0=0p^{0}_{h}=0, compared to other techniques, like pressure Poisson equation used in [16, 17]; secondly, it can also directly get the initial intermediate discrete velocity 𝐮~h0\widetilde{\bm{u}}^{0}_{h} since at this point 𝐮~h0=𝐮h0\widetilde{\bm{u}}^{0}_{h}=\bm{u}^{0}_{h}. We may also observe that this rude treatment would make the first few time steps discrete pressure appear oscillation. Indeed, numerical experiments have showed that the first few time steps discrete pressure values have to be abandoned because of big discrete errors; nevertheless, we will see in later numerical experiments that, only after a few time steps(for example, from n=6n=6 on), the discrete pressure values can be adopted as the appropriate discrete approximation of the exact pressure.

Remark 3.3.

We note that although the obtained finite element velocity solution in scheme 3 is 𝐮~hn+1\widetilde{\bm{u}}^{n+1}_{h}, rather than 𝐮hn+1\bm{u}^{n+1}_{h}, we will see in latter numerical experiment that when choosing 𝐕h×Qh=P1\bm{V}_{h}\times Q_{h}=P^{1}-P1P^{1}, the solved 𝐮~hn+1\widetilde{\bm{u}}^{n+1}_{h} can be used as a proper finite element approximation for the exact velocity 𝐮n+1\bm{u}^{n+1}. If one wants to obtain 𝐮hn+1\bm{u}^{n+1}_{h}, we can multiply by the last equation in scheme 2 some test function 𝐰h𝐖h=𝐘h1𝐉1\bm{w}_{h}\in\bm{W}_{h}=\bm{Y}_{h}^{1}\cap\bm{J}_{1} to get

(𝒖hn+1,𝒘h)=(𝒖~hn+1,𝒘h)Δt(phn+1,𝒘h).(\bm{u}^{n+1}_{h},\bm{w}_{h})=(\widetilde{\bm{u}}^{n+1}_{h},\bm{w}_{h})-\Delta t(\nabla p^{n+1}_{h},\bm{w}_{h}). (3.11)

then, solving above equation can finally obtain 𝐮hn+1𝐖h=𝐘h1𝐉1\bm{u}^{n+1}_{h}\in\bm{W}_{h}=\bm{Y}_{h}^{1}\cap\bm{J}_{1}.

We observe that, in the classical projection scheme (3.9) or (3.10), if we set Δt=𝒪(h2)\Delta t=\mathcal{O}(h^{2}), the existence of PSPG-stabilized term Δt(phn+1,qh)\Delta t(\nabla p^{n+1}_{h},\nabla q_{h}) makes the spatial convergence order of the error can not achieve more than second order in L2L^{2} norm of the velocity and first order of the error in the L2L^{2} norm of the pressure, so P1P^{1}-P1P^{1} for finite element spaces and (𝒖,p)𝑯2×H1(\bm{u},p)\in\bm{H}^{2}\times H^{1} seem to be the best choice concerning about the efficiency of computation and the regularity for the exact solutions. Thus, We set l=1l=1 in the definition of YhlY^{l}_{h}, thus the finite element spaces for velocity and pressure would be 𝑽h=Yh1𝑽\bm{V}_{h}=Y^{1}_{h}\cap\bm{V} and Qh=Yh1QQ_{h}=Y^{1}_{h}\cap Q, respectively.
In the sequel we assume Δt=𝒪(h2)\Delta t=\mathcal{O}(h^{2}). Specifically, we assume there exist two positive constants C1,C2C_{1},C_{2}, such that

C1h2ΔtC2h2.C_{1}h^{2}\leq\Delta t\leq C_{2}h^{2}. (3.12)

The following modified inf-sup condition relaxes the classical inf-sup condition and makes many mixed finite element pairs “stable” in the sense of this modified one, whose detailed proof can be found in [4, Lemma 3] or [17, Lemma 2.1].

Lemma 3.1.

(modified inf-sup condition) Assuming (3.12) holds, then for n=0,1,,N1n=0,1,\ldots,N-1, we have the following pressure stability

βqh0Chqh0+sup𝒗h𝑽h(qh,𝒗h)𝒗h1,qhQh.\beta\|q_{h}\|_{0}\leq Ch\|\nabla q_{h}\|_{0}+\sup_{\bm{v}_{h}\in\bm{V}_{h}}\frac{(q_{h},\nabla\cdot\bm{v}_{h})}{\|\bm{v}_{h}\|_{1}},\quad\forall q_{h}\in Q_{h}. (3.13)

3.2 convergence of classical projection scheme

In this subsection, we will cite some error estimation results of the finite element solution in the classical projection scheme and omit those proofs, since those estimates have been analyzed in [10] in detail. The necessity of citing is based on the fact that the following POD snapshots are the finite element solutions, which means that when we analyze the discretization error of the POD-based reduced-order solutions, apart from the ROM truncation error, the discretization error also involve spatial and temporal discretization errors.

Lemma 3.2.

(error estimates for Proj-PSPG-FE) For n=1,,Nn=1,\cdots,N, Let (𝐮n,pn)𝐗×Q(\bm{u}^{n},p^{n})\in\bm{X}\times Q be the solution of continuous variational form (2.3) at discrete time t=tnt=t_{n}, (𝐮~hn,phn)𝐗h×Qh(\widetilde{\bm{u}}^{n}_{h},p^{n}_{h})\in\bm{X}_{h}\times Q_{h} is the solution obtained with the PSPG-like projection scheme (3.10), and assuming (3.12) holds, i.e. Δt=𝒪(h2)\Delta t=\mathcal{O}(h^{2}), then we have

𝒖n𝒖~hn0+hpnphn0+hΔt(pnphn)0\displaystyle\|\bm{u}^{n}-\widetilde{\bm{u}}^{n}_{h}\|_{0}+h\|p^{n}-p^{n}_{h}\|_{0}+h\sqrt{\Delta t}\|\nabla(p^{n}-p^{n}_{h})\|_{0} C(h2+Δt).\displaystyle\leq C(h^{2}+\Delta t). (3.14)
(𝒖𝒖~h)l2(L2)+pphl2(L2)\displaystyle\|\nabla(\bm{u}-\bm{\widetilde{u}}_{h})\|_{l^{2}(L^{2})}+\|p-p_{h}\|_{l^{2}(L^{2})} C(h+Δt).\displaystyle\leq C(h+\sqrt{\Delta t}).

In addition to the error estimates about the intermediate velocity 𝒖~hn\widetilde{\bm{u}}^{n}_{h} analyzed in [10], we can also prove the L2L^{2} error estimates about the end-of-step velocity 𝒖hn\bm{u}^{n}_{h} which is not made in [10]. In other words, we can obtain

Lemma 3.3.

For n=1,,Nn=1,\cdots,N, 𝐮hn\bm{u}^{n}_{h} is the velocity solution obtained in Scheme 2, and 𝐮n\bm{u}^{n} is the velocity solution of continuous variational form (2.3) at discrete time t=tnt=t_{n}, then

𝒖n𝒖hn0C(h2+Δt).\|\bm{u}^{n}-\bm{u}^{n}_{h}\|_{0}\leq C(h^{2}+\Delta t). (3.15)
Proof.

According to the convergent results given by Lemma 3.2, we can prove the stability of phn0\|\nabla p^{n}_{h}\|_{0}. That is, the first inequality in (3.14) implies

(pnphn)0C(h/Δt+Δt/h)C,\|\nabla(p^{n}-p^{n}_{h})\|_{0}\leq C(h/\sqrt{\Delta t}+\sqrt{\Delta t}/h)\leq C,

where we have used (3.12). Then

phn0(pnphn)0+pn0C,\|\nabla p^{n}_{h}\|_{0}\leq\|\nabla(p^{n}-p^{n}_{h})\|_{0}+\|\nabla p^{n}\|_{0}\leq C,

where the constant CC can be chose to stay nearly fixed with the decrease of hh and Δt\Delta t. So, based on the result, testing 𝒘h=𝒖~hn+1𝒖hn+1\bm{w}_{h}=\widetilde{\bm{u}}^{n+1}_{h}-\bm{u}^{n+1}_{h} in (3.11) and rearranging to get

𝒖~hn+1𝒖hn+10Δtphn+10CΔt.\|\widetilde{\bm{u}}^{n+1}_{h}-\bm{u}^{n+1}_{h}\|_{0}\leq\Delta t\|\nabla p^{n+1}_{h}\|_{0}\leq C\Delta t.

Finally, utilizing triangle inequality and Lemma 3.2,

𝒖n+1𝒖hn+10𝒖n+1𝒖~hn+10+𝒖~hn+1𝒖hn+10C(h2+Δt).\|\bm{u}^{n+1}-\bm{u}^{n+1}_{h}\|_{0}\leq\|\bm{u}^{n+1}-\widetilde{\bm{u}}^{n+1}_{h}\|_{0}+\|\widetilde{\bm{u}}^{n+1}_{h}-\bm{u}^{n+1}_{h}\|_{0}\leq C(h^{2}+\Delta t).

4 Projection POD

4.1 POD

In this subsection, we will briefly give some essential ingredients for constructing POD method, more details for constructing POD space and POD-based reduced-order models can be found in [21, 32]. As we remarked in Remark 3.2, the numerical pressure oscillation in the first few time steps caused by setting ph0=0p^{0}_{h}=0 makes us have to discard the finite element solutions in the previous time steps, so when using finite element solution obtained by solving Scheme 4 to construct POD spaces, we also need to consider this situations. In view of the fact that the different choice of initial pressure values will affect the discrete error of finite element pressure in the previous time steps, we uniformly denote n=n0,n01n=n_{0},n_{0}\geq 1 as the starting time steps of taking snapshots in Scheme 4; therefore, for a given positive integer MM, we choose the finite element solution (𝒖~hi,phi),n0in0+M1(\widetilde{\bm{u}}^{i}_{h},p^{i}_{h}),\;n_{0}\leq i\leq n_{0}+M-1 in Scheme 4, and its difference quotients(DQs) (𝒖~hi,phi),n0+1in0+M1(\partial\widetilde{\bm{u}}^{i}_{h},\partial p^{i}_{h}),\;n_{0}+1\leq i\leq n_{0}+M-1, where the DQs fi\partial f^{i} are defined by fi:=(fifi1)/Δt\partial f^{i}:=(f^{i}-f^{i-1})/\Delta t for some function ff at discrete time t=tit=t_{i}, to formulate the snapshot spaces

𝒰~\displaystyle\widetilde{\mathcal{U}} :=𝒖~hn0,𝒖~hn0+1,,𝒖~hM+n01,𝒖~hn0+1,𝒖~hn0+2,,𝒖~hM+n01,\displaystyle:=\langle\widetilde{\bm{u}}^{n_{0}}_{h},\widetilde{\bm{u}}^{n_{0}+1}_{h},\cdots,\widetilde{\bm{u}}^{M+n_{0}-1}_{h},\partial\widetilde{\bm{u}}^{n_{0}+1}_{h},\partial\widetilde{\bm{u}}^{n_{0}+2}_{h},\cdots,\partial\widetilde{\bm{u}}^{M+n_{0}-1}_{h}\rangle,
𝒫\displaystyle\mathcal{P} :=phn0,phn0+1,,phM+n01,phn0+1,phn0+2,,phM+n0+1.\displaystyle:=\langle p^{n_{0}}_{h},p^{n_{0}+1}_{h},\cdots,p^{M+n_{0}-1}_{h},\partial p^{n_{0}+1}_{h},\partial p^{n_{0}+2}_{h},\cdots,\partial p^{M+n_{0}+1}_{h}\rangle.

where S\langle S\rangle denotes the space spanned by the set SS and we denote by Ns=2M1N_{s}=2M-1 the number of snapshots. Let du~,dpd_{\widetilde{u}},d_{p} be the dimensions of the spaces 𝒰~\widetilde{\mathcal{U}} and 𝒫\mathcal{P}, respectively. The symbols Ku~,KpK_{\widetilde{u}},K_{p} are correlation matrices corresponding to the snapshots Ku~=((Ki,ju~))Ns×NsK_{\widetilde{u}}=\left((K^{\widetilde{u}}_{i,j})\right)\in\mathbb{R}^{N_{s}\times N_{s}} and Kp=((Ki,jp))Ns×NsK_{p}=\left((K^{p}_{i,j})\right)\in\mathbb{R}^{N_{s}\times N_{s}} where

Ki,ju~:=(𝒖~hi,𝒖~hj),Ki,jp:=(phi,phj).K^{\widetilde{u}}_{i,j}:=\left(\widetilde{\bm{u}}^{i}_{h},\widetilde{\bm{u}}^{j}_{h}\right),\quad K^{p}_{i,j}:=\left(p^{i}_{h},p^{j}_{h}\right).

and (,)(\cdot,\cdot) is the L2L^{2} inner product. Just as in [19], a singular value decomposition(SVD) is carried out and the leading generalized eigenfunctions are chosen as bases, referred to as the POD bases. We denote by λ1λ2λdu~>0\lambda_{1}\geq\lambda_{2}\geq\cdots\lambda_{d_{\widetilde{u}}}>0 the positive eigenvalues of Ku~K_{\widetilde{u}} and by 𝒙1,𝒙2,,𝒙du~Ns\bm{x}_{1},\bm{x}_{2},\ldots,\bm{x}_{d_{\widetilde{u}}}\in\mathbb{R}^{N_{s}} the associated eigenvectors. Analogously, {γi,𝒚i}i=1dp\{\gamma_{i},\bm{y}_{i}\}^{d_{p}}_{i=1} are the eigen-pairs of KpK_{p}. Then, the two POD bases can be written explicitly as

𝝋i\displaystyle\bm{\varphi}_{i} =1λi(j=n0n0+M1xij𝒖~hj+j=n0+1n0+M1xij+M1𝒖~hj),\displaystyle=\frac{1}{\sqrt{\lambda_{i}}}\left(\sum^{n_{0}+M-1}_{j=n_{0}}x^{j}_{i}\widetilde{\bm{u}}^{j}_{h}+\sum^{n_{0}+M-1}_{j=n_{0}+1}x^{j+M-1}_{i}\partial\widetilde{\bm{u}}^{j}_{h}\right),\; fori=1,2,,du~,\displaystyle\text{for}\;i=1,2,\cdots,d_{\widetilde{u}}, (4.16)
ψi\displaystyle\psi_{i} =1γi(j=n0n0+M1yijphj+j=n0n0+M1yij+M1phj),\displaystyle=\frac{1}{\sqrt{\gamma_{i}}}\left(\sum^{n_{0}+M-1}_{j=n_{0}}y^{j}_{i}p^{j}_{h}+\sum^{n_{0}+M-1}_{j=n_{0}}y^{j+M-1}_{i}\partial p^{j}_{h}\right),\; fori=1,2,,dp.\displaystyle\text{for}\;i=1,2,\cdots,d_{p}.

In what follows we will denote by

𝑽r=𝝋1,𝝋2,,𝝋r,Qr=ψ1,ψ2,,ψr.\displaystyle\bm{V}_{r}=\langle\bm{\varphi}_{1},\bm{\varphi}_{2},\cdots,\bm{\varphi}_{r}\rangle,\quad Q_{r}=\langle\psi_{1},\psi_{2},\cdots,\psi_{r}\rangle.

i.e., we choose the first rr POD bases to span the POD spaces, and define the traditional L2L^{2} projection into the POD velocity space 𝑽r\bm{V}_{r} and POD pressure space QrQ_{r} as

Definition 4.1.

Let Πrv:L2(Ω)𝐕r\Pi^{v}_{r}:L^{2}(\Omega)\rightarrow\bm{V}_{r} and Πrq:L2(Ω)Qr\Pi^{q}_{r}:L^{2}(\Omega)\rightarrow Q_{r} such that

(𝒖Πrv𝒖,𝒗r)\displaystyle(\bm{u}-\Pi^{v}_{r}\bm{u},\bm{v}_{r}) =0𝒗r𝑽rand\displaystyle=0\quad\forall\bm{v}_{r}\in\bm{V}_{r}\quad and
(pΠrqp,qr)\displaystyle(p-\Pi^{q}_{r}p,q_{r}) =0qrQr.\displaystyle=0\quad\forall q_{r}\in Q_{r}.

Then, we can get the following Proj-POD-ROM scheme:
(Scheme 4 - Proj-POD-ROM) For some n01n_{0}\geq 1, let (𝒖~rn0,prn0)=(Πrv𝒖~hn0,Πrqphn0)=(i=1r(𝒖~hn0,𝝋i)𝝋i(\widetilde{\bm{u}}^{n_{0}}_{r},p^{n_{0}}_{r})=(\Pi^{v}_{r}\widetilde{\bm{u}}^{n_{0}}_{h},\Pi^{q}_{r}p^{n_{0}}_{h})=(\sum^{r}_{i=1}(\widetilde{\bm{u}}^{n_{0}}_{h},\bm{\varphi}_{i})\bm{\varphi}_{i}, i=1r(phn0,ψi)ψi)\sum^{r}_{i=1}(p^{n_{0}}_{h},\psi_{i})\psi_{i}) be the Galerkin projection of (𝒖~hn0,phn0)(\widetilde{\bm{u}}^{n_{0}}_{h},p^{n_{0}}_{h}) onto POD spaces 𝑽r×Qr\bm{V}_{r}\times Q_{r}, then for n=n0,n0+1,n0+2,,N1n=n_{0},n_{0}+1,n_{0}+2,\cdots,N-1, we can compute (𝒖~rn+1,prn+1)𝑽r×Qr(\widetilde{\bm{u}}^{n+1}_{r},p^{n+1}_{r})\in\bm{V}_{r}\times Q_{r} from

{(𝒖~rn+1𝒖~rnΔt,𝒗r)+ν(𝒖~rn+1,𝒗r)+(prn,𝒗r)=𝒇n+1,𝒗r,𝒗r𝑽r,(𝒖~rn+1,qr)+Δt(prn+1,qr)=0,qrQr.\left\{\begin{aligned} \left(\frac{\widetilde{\bm{u}}^{n+1}_{r}-\widetilde{\bm{u}}^{n}_{r}}{\Delta t},\bm{v}_{r}\right)+\nu(\nabla\widetilde{\bm{u}}^{n+1}_{r},\nabla\bm{v}_{r})+(\nabla p^{n}_{r},\bm{v}_{r})&=\langle\bm{f}^{n+1},\bm{v}_{r}\rangle,\quad\forall\bm{v}_{r}\in\bm{V}_{r},\\ (\nabla\cdot\widetilde{\bm{u}}^{n+1}_{r},q_{r})+\Delta t(\nabla p^{n+1}_{r},\nabla q_{r})&=0,\quad\forall q_{r}\in Q_{r}.\end{aligned}\right. (4.17)

4.2 stability and convergence analysis of projection-POD scheme

We will carry on some numerical analysis about the newly proposed Proj-POD-ROM. To this end, we first recall the classical conclusions concerning about the projection error between the snapshots solution and POD-based reduced-order solution(see [19])

Lemma 4.1.

(time-averaged projection error estimates) We have the following time-averaged projection error estimates.

1Nsj=n0n0+M1𝒖~hjk=1r(𝒖~hj,𝝋k)𝝋k02+1Nsj=n0+1n0+M1𝒖~hjk=1r(𝒖~hj,𝝋k)𝝋k02\displaystyle\frac{1}{{N_{s}}}\sum^{n_{0}+M-1}_{j=n_{0}}\left\|\widetilde{\bm{u}}^{j}_{h}-\sum^{r}_{k=1}\left(\widetilde{\bm{u}}^{j}_{h},\bm{\varphi}_{k}\right)\bm{\varphi}_{k}\right\|^{2}_{0}+\frac{1}{{N_{s}}}\sum^{n_{0}+M-1}_{j=n_{0}+1}\left\|\partial\widetilde{\bm{u}}^{j}_{h}-\sum^{r}_{k=1}\left(\partial\widetilde{\bm{u}}^{j}_{h},\bm{\varphi}_{k}\right)\bm{\varphi}_{k}\right\|^{2}_{0} =k=r+1du~λk,\displaystyle=\sum^{d_{\widetilde{u}}}_{k=r+1}\lambda_{k}, (4.18)
1Nsj=n0n0+M1phjk=1r(phj,ψk)ϕk02+1Nsj=n0+1n0+M1phjk=1r(phj,ψk)ϕk02\displaystyle\frac{1}{{N_{s}}}\sum^{n_{0}+M-1}_{j=n_{0}}\left\|p^{j}_{h}-\sum^{r}_{k=1}\left(p^{j}_{h},\psi_{k}\right)\phi_{k}\right\|^{2}_{0}+\frac{1}{{N_{s}}}\sum^{n_{0}+M-1}_{j=n_{0}+1}\left\|\partial p^{j}_{h}-\sum^{r}_{k=1}\left(\partial p^{j}_{h},\psi_{k}\right)\phi_{k}\right\|^{2}_{0} =k=r+1dpγk.\displaystyle=\sum^{d_{p}}_{k=r+1}\gamma_{k}.
Remark 4.1.

We remark that choosing L2L^{2} inner product to construct POD bases and by virtue of Parseval’s identity makes the above three identities actually represent L2L^{2} POD projection errors.

Apart from the above time-averaged projection error estimates, we also need the following up-to-data optimal in time error estimates(see, [18]), which is presented in the form of hypothesis in the previous papers and is proved to be valid in the presence of DQs in [18].

Lemma 4.2.

(POD Optimal pointwise in time error estimate) We have the following optimal in time projection errors

max0kN𝒖~hkΠrv𝒖~hk02\displaystyle\max_{0\leq k\leq N}\|\bm{\widetilde{u}}^{k}_{h}-\Pi^{v}_{r}\bm{\widetilde{u}}^{k}_{h}\|^{2}_{0} Ci=r+1du~λi,\displaystyle\leq C\sum^{d_{\widetilde{u}}}_{i=r+1}\lambda_{i},
max0kNphkΠrqphk02\displaystyle\max_{0\leq k\leq N}\|p^{k}_{h}-\Pi^{q}_{r}p^{k}_{h}\|^{2}_{0} Ci=r+1dpγi,\displaystyle\leq C\sum^{d_{p}}_{i=r+1}\gamma_{i},

where C=6max{1,T2}C=6\max\{1,T^{2}\}, and

max0kN(𝒖~hkΠrv𝒖~hk)02\displaystyle\max_{0\leq k\leq N}\|\nabla(\bm{\widetilde{u}}^{k}_{h}-\Pi^{v}_{r}\bm{\widetilde{u}}^{k}_{h})\|^{2}_{0} Ci=r+1du~λi𝝋i02,\displaystyle\leq C\sum^{d_{\widetilde{u}}}_{i=r+1}\lambda_{i}\|\nabla\bm{\varphi}_{i}\|^{2}_{0},
max0kN(phkΠrqphk)02\displaystyle\max_{0\leq k\leq N}\|\nabla(p^{k}_{h}-\Pi^{q}_{r}p^{k}_{h})\|^{2}_{0} Ci=r+1dpγiψi02,\displaystyle\leq C\sum^{d_{p}}_{i=r+1}\gamma_{i}\|\nabla\psi_{i}\|^{2}_{0},

where C=6max{1,T2}C=6\max\{1,T^{2}\}.

We also need in the later analysis the time-averaged optimal projection error estimate about DQs, which is derived in [19].

Lemma 4.3.

(POD Optimal time-averaged error estimate about DQs) For the difference quotients we have the estimate:

1Mk=n0n0+M1𝒖~hkΠrv𝒖~hk02Ci=r+1du~λi.\frac{1}{M}\sum^{n_{0}+M-1}_{k=n_{0}}\|\partial\bm{\widetilde{u}}^{k}_{h}-\Pi^{v}_{r}\partial\bm{\widetilde{u}}^{k}_{h}\|^{2}_{0}\leq C\sum^{d_{\widetilde{u}}}_{i=r+1}\lambda_{i}.

The following analogues of finite element inverse inequalities will be used in later analysis(see, [19])

Lemma 4.4.

(POD inverse estimates) Denote by Su~=((si,ju~))Ns×NsS_{\widetilde{u}}=\left((s^{\widetilde{u}}_{i,j})\right)\in\mathbb{R}^{{N_{s}}\times{N_{s}}} and Sp=((si,jp))Ns×NsS_{p}=\left((s^{p}_{i,j})\right)\in\mathbb{R}^{{N_{s}}\times{N_{s}}} are stiffness matrices for POD spaces 𝐕r,Qr\bm{V}_{r},Q_{r} respectively, with entries si,ju~=(𝛗i,𝛗j),si,jp=(ψi,ψj)s^{\widetilde{u}}_{i,j}=(\nabla\bm{\varphi}_{i},\nabla\bm{\varphi}_{j}),s^{p}_{i,j}=(\nabla\psi_{i},\nabla\psi_{j}), and by ||||||2{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|\cdot\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{2} the spectral norm of the matrix, then

𝒗~0\displaystyle\|\nabla\widetilde{\bm{v}}\|_{0} |𝑺u~|2𝒗~0,for𝒗~𝑽r,\displaystyle\leq\sqrt{{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|\bm{S}_{\widetilde{u}}\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{2}}\|\widetilde{\bm{v}}\|_{0},\ \text{for}\ \forall\widetilde{\bm{v}}\in\bm{V}_{r}, (4.19)
q0\displaystyle\|\nabla q\|_{0} |Sp|2q0,forqQr.\displaystyle\leq\sqrt{{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|S_{p}\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{2}}\|q\|_{0},\ \text{for}\;\forall q\in Q_{r}.
Remark 4.2.

Mass matrix MM, for both (𝛗i,𝛗j)(\bm{\varphi}_{i},\bm{\varphi}_{j}) and (ψi,ψj)(\psi_{i},\psi_{j}), equals identity II since the standard orthogonality of POD base in the sense of L2L^{2} inner product.

We can prove the unconditional stability of the Proj-POD-ROM scheme.

Theorem 4.1.

(Unconditional stability of Proj-POD-ROM) We have the following unconditional stability

𝒖~rN02+n=n0N1𝒖~rn+1𝒖~rn02\displaystyle\left\|\widetilde{\bm{u}}_{r}^{N}\right\|_{0}^{2}+\sum_{n=n_{0}}^{N-1}\left\|\widetilde{\bm{u}}_{r}^{n+1}-\widetilde{\bm{u}}_{r}^{n}\right\|_{0}^{2} +Δtn=n0N1(ν𝒖~rn+102+Δtprn+102,)\displaystyle+\Delta t\sum_{n=n_{0}}^{N-1}\left(\nu\left\|\nabla\widetilde{\bm{u}}_{r}^{n+1}\right\|_{0}^{2}+\Delta t\left\|\nabla p_{r}^{n+1}\right\|_{0}^{2},\right) (4.20)
\displaystyle\leq C(𝒖~r002+Δtνn=n0N1𝒇n+11).\displaystyle C\left(\left\|\widetilde{\bm{u}}_{r}^{0}\right\|_{0}^{2}+\frac{\Delta t}{\nu}\sum_{n=n_{0}}^{N-1}\|\bm{f}^{n+1}\|_{-1}\right).
Proof.

Taking (𝒗r,qr)=(𝒖~rn+1,prn)(\bm{v}_{r},q_{r})=(\widetilde{\bm{u}}^{n+1}_{r},p^{n}_{r}) in (4.17) and multiply by 2Δt2\Delta t, adding both equations and integrating by parts, adding and subtracting 2Δt2prn+1022\Delta t^{2}\|\nabla p_{r}^{n+1}\|^{2}_{0}, we get

𝒖~rn+102𝒖~rn02\displaystyle\left\|\widetilde{\bm{u}}_{r}^{n+1}\right\|_{0}^{2}-\left\|\widetilde{\bm{u}}_{r}^{n}\right\|_{0}^{2} +𝒖~rn+1𝒖~rn02+2νΔt𝒖~rn+102+2Δt2prn+102,\displaystyle+\left\|\widetilde{\bm{u}}_{r}^{n+1}-\widetilde{\bm{u}}_{r}^{n}\right\|_{0}^{2}+2\nu\Delta t\left\|\nabla\widetilde{\bm{u}}_{r}^{n+1}\right\|_{0}^{2}+2\Delta t^{2}\|\nabla p_{r}^{n+1}\|^{2}_{0}, (4.21)
CΔtν𝒇n+11+νΔt𝒖~rn+102+2Δt2(prn+1,(prn+1prn)).\displaystyle\leq C\frac{\Delta t}{\nu}\|\bm{f}^{n+1}\|_{-1}+\nu\Delta t\left\|\nabla\widetilde{\bm{u}}_{r}^{n+1}\right\|_{0}^{2}+2\Delta t^{2}\left(\nabla p_{r}^{n+1},\nabla(p_{r}^{n+1}-p_{r}^{n})\right).

For the last term above, we utilize the second equation in (4.17) twice to obtain

Δt(prn+1,(prn+1prn))=((𝒖~rn+1𝒖~rn),prn+1)=(𝒖~rn+1𝒖~rn,prn+1).\Delta t\left(\nabla p_{r}^{n+1},\nabla(p_{r}^{n+1}-p_{r}^{n})\right)=-(\nabla\cdot(\widetilde{\bm{u}}^{n+1}_{r}-\widetilde{\bm{u}}^{n}_{r}),p_{r}^{n+1})=(\widetilde{\bm{u}}^{n+1}_{r}-\widetilde{\bm{u}}^{n}_{r},\nabla p_{r}^{n+1}).

This equality implies

2Δt2(prn+1,(prn+1prn))23𝒖~rn+1𝒖~rn02+32Δt2prn+102.2\Delta t^{2}\left(\nabla p_{r}^{n+1},\nabla(p_{r}^{n+1}-p_{r}^{n})\right)\leq\frac{2}{3}\|\widetilde{\bm{u}}^{n+1}_{r}-\widetilde{\bm{u}}^{n}_{r}\|^{2}_{0}+\frac{3}{2}\Delta t^{2}\|\nabla p_{r}^{n+1}\|^{2}_{0}.

Inserting the above equality into (4.21) and adding nn from 0 to N1N-1 we finally obtain (4.20). ∎

We have the following truncation error estimate between the POD reduced-order solution and the snapshot data.

Lemma 4.5.

(POD truncation error) For n=1,2,,Nn=1,2,\cdots,N, (𝐮~hn,phn)(\widetilde{\bm{u}}^{n}_{h},p^{n}_{h}) is the solution of finite element projection scheme (3.10) and (𝐮~rn,prn)(\widetilde{\bm{u}}^{n}_{r},p^{n}_{r}) is the solution of POD projection scheme (4.17), then we have the following error estimate:

𝒖~hN𝒖~rN02\displaystyle\|\widetilde{\bm{u}}^{N}_{h}-\widetilde{\bm{u}}^{N}_{r}\|^{2}_{0} +νΔtn=n0N1(𝒖~hn+1𝒖~rn+1)02+Δt2n=n0N1(phn+1prn+1)02,\displaystyle+\nu\Delta t\sum^{N-1}_{n=n_{0}}\|\nabla(\widetilde{\bm{u}}^{n+1}_{h}-\widetilde{\bm{u}}^{n+1}_{r})\|^{2}_{0}+\Delta t^{2}\sum^{N-1}_{n=n_{0}}\|\nabla(p^{n+1}_{h}-p^{n+1}_{r})\|^{2}_{0}, (4.22)
C(i=r+1du~γi(1+(ν+1)φi02)+(Δt+1)i=r+1dpϵiψi02+Δt2).\displaystyle\leq C\left(\sum^{d_{\widetilde{u}}}_{i=r+1}\gamma_{i}\Big{(}1+(\nu+1)\|\nabla\varphi_{i}\|^{2}_{0}\Big{)}+(\Delta t+1)\sum^{d_{p}}_{i=r+1}\epsilon_{i}\|\nabla\psi_{i}\|^{2}_{0}+\Delta t^{2}\right).
Proof.

Finite element projection scheme (3.10) subtract POD projection scheme (4.17) to get:

{(𝒖~hn+1𝒖~hnΔt𝒖~rn+1𝒖~rnΔt,𝒗r)+ν((𝒖~hn+1𝒖~rn+1),𝒗r)+((phnprn),𝒗r)=0,((𝒖~hn+1𝒖~rn+1),qr)+Δt((phn+1prn+1),qr)=0.\left\{\begin{aligned} \left(\frac{\widetilde{\bm{u}}^{n+1}_{h}-\widetilde{\bm{u}}^{n}_{h}}{\Delta t}-\frac{\widetilde{\bm{u}}^{n+1}_{r}-\widetilde{\bm{u}}^{n}_{r}}{\Delta t},\bm{v}_{r}\right)+\nu(\nabla(\widetilde{\bm{u}}^{n+1}_{h}-\widetilde{\bm{u}}^{n+1}_{r}),\nabla\bm{v}_{r})+(\nabla(p^{n}_{h}-p^{n}_{r}),\bm{v}_{r})&=0,\\ (\nabla\cdot(\widetilde{\bm{u}}^{n+1}_{h}-\widetilde{\bm{u}}^{n+1}_{r}),q_{r})+\Delta t(\nabla(p^{n+1}_{h}-p^{n+1}_{r}),\nabla q_{r})&=0.\end{aligned}\right. (4.23)

Denoting

𝒖~hn+1𝒖~rn+1=𝒖~hn+1Πrv𝒖~hn+1+Πrv𝒖~hn+1𝒖~rn+1\displaystyle\widetilde{\bm{u}}^{n+1}_{h}-\widetilde{\bm{u}}^{n+1}_{r}=\widetilde{\bm{u}}^{n+1}_{h}-\Pi^{v}_{r}\widetilde{\bm{u}}^{n+1}_{h}+\Pi^{v}_{r}\widetilde{\bm{u}}^{n+1}_{h}-\widetilde{\bm{u}}^{n+1}_{r} :=𝜼~u,rn+1+𝒘~u,rn+1,\displaystyle:=\widetilde{\bm{\eta}}^{n+1}_{u,r}+\widetilde{\bm{w}}^{n+1}_{u,r},
phn+1prn+1=phn+1Πrqphn+1+Πrqphn+1prn+1\displaystyle p^{n+1}_{h}-p^{n+1}_{r}=p^{n+1}_{h}-\Pi^{q}_{r}p^{n+1}_{h}+\Pi^{q}_{r}p^{n+1}_{h}-p^{n+1}_{r} :=ηp,rn+1+wp,rn+1.\displaystyle:=\eta^{n+1}_{p,r}+w^{n+1}_{p,r}.

Testing (𝒗r,qr)=(𝒘~u,rn+1,wp,rn+1)(\bm{v}_{r},q_{r})=(\widetilde{\bm{w}}^{n+1}_{u,r},w^{n+1}_{p,r}), we obtain

{(𝒘~u,rn+1𝒘~u,rnΔt,𝒘~u,rn+1)+ν𝒘~u,rn+102+(wp,rn+1,𝒘~u,rn+1)=(𝜼~u,rn+1𝜼~u,rnΔt,𝒘~u,rn+1)+ν(𝜼~u,rn+1,𝒘~u,rn+1),+((phn+1phn),𝒘~u,rn+1)+((prnprn+1),𝒘~u,rn+1)+(ηp,rn+1,𝒘~u,rn+1),(𝒘~u,rn+1,wp,rn+1)+Δtwp,rn+102=(𝜼~u,rn+1,𝒘~u,rn+1)+Δt(ηp,rn+1,wp,rn+1).\left\{\begin{aligned} \left(\frac{\widetilde{\bm{w}}^{n+1}_{u,r}-\widetilde{\bm{w}}^{n}_{u,r}}{\Delta t},\widetilde{\bm{w}}^{n+1}_{u,r}\right)&+\nu\|\nabla\widetilde{\bm{w}}^{n+1}_{u,r}\|^{2}_{0}+(\nabla w^{n+1}_{p,r},\widetilde{\bm{w}}^{n+1}_{u,r})=\left(\frac{\widetilde{\bm{\eta}}^{n+1}_{u,r}-\widetilde{\bm{\eta}}^{n}_{u,r}}{\Delta t},\widetilde{\bm{w}}^{n+1}_{u,r}\right)+\nu(\nabla\widetilde{\bm{\eta}}^{n+1}_{u,r},\nabla\widetilde{\bm{w}}^{n+1}_{u,r}),\\ &\quad+(\nabla(p^{n+1}_{h}-p^{n}_{h}),\widetilde{\bm{w}}^{n+1}_{u,r})+(\nabla(p^{n}_{r}-p^{n+1}_{r}),\widetilde{\bm{w}}^{n+1}_{u,r})+(\nabla\eta^{n+1}_{p,r},\widetilde{\bm{w}}^{n+1}_{u,r}),\\ (\nabla\cdot\widetilde{\bm{w}}^{n+1}_{u,r},w^{n+1}_{p,r})&+\Delta t\|\nabla w^{n+1}_{p,r}\|^{2}_{0}=(\nabla\cdot\widetilde{\bm{\eta}}^{n+1}_{u,r},\widetilde{\bm{w}}^{n+1}_{u,r})+\Delta t(\nabla\eta^{n+1}_{p,r},\nabla w^{n+1}_{p,r}).\end{aligned}\right.

Adding both equations we have the following error equation

(𝒘~u,rn+1𝒘~u,rnΔt,𝒘~u,rn+1)\displaystyle\left(\frac{\widetilde{\bm{w}}^{n+1}_{u,r}-\widetilde{\bm{w}}^{n}_{u,r}}{\Delta t},\widetilde{\bm{w}}^{n+1}_{u,r}\right) +ν𝒘~u,rn+102+Δtwp,rn+102=(𝜼~u,rn+1𝜼~u,rnΔt,𝒘~u,rn+1),\displaystyle+\nu\|\nabla\widetilde{\bm{w}}^{n+1}_{u,r}\|^{2}_{0}+\Delta t\|\nabla w^{n+1}_{p,r}\|^{2}_{0}=\left(\frac{\widetilde{\bm{\eta}}^{n+1}_{u,r}-\widetilde{\bm{\eta}}^{n}_{u,r}}{\Delta t},\widetilde{\bm{w}}^{n+1}_{u,r}\right), (4.24)
+ν(𝜼~u,rn+1,𝒘~u,rn+1)+((phn+1phn),𝒘~u,rn+1)+((prnprn+1),𝒘~u,rn+1),\displaystyle+\nu(\nabla\widetilde{\bm{\eta}}^{n+1}_{u,r},\nabla\widetilde{\bm{w}}^{n+1}_{u,r})+(\nabla(p^{n+1}_{h}-p^{n}_{h}),\widetilde{\bm{w}}^{n+1}_{u,r})+(\nabla(p^{n}_{r}-p^{n+1}_{r}),\widetilde{\bm{w}}^{n+1}_{u,r}),
+(ηp,rn+1,𝒘~u,rn+1)+(𝜼~u,rn+1,𝒘~u,rn+1)+Δt(ηp,rn+1,wp,rn+1),\displaystyle+(\nabla\eta^{n+1}_{p,r},\widetilde{\bm{w}}^{n+1}_{u,r})+(\nabla\cdot\widetilde{\bm{\eta}}^{n+1}_{u,r},\widetilde{\bm{w}}^{n+1}_{u,r})+\Delta t(\nabla\eta^{n+1}_{p,r},\nabla w^{n+1}_{p,r}),
:=A1+A2+A3+A4+A5+A6+A7.\displaystyle:=A_{1}+A_{2}+A_{3}+A_{4}+A_{5}+A_{6}+A_{7}.

We will bound the seven terms above separately.

|A1|\displaystyle|A_{1}| =|(𝜼~u,rn+1𝜼~u,rnΔt,𝒘~u,rn+1)|C𝜼~u,rn+1𝜼~u,rnΔt02+110𝒘~u,rn+102.\displaystyle=\left|\left(\frac{\widetilde{\bm{\eta}}^{n+1}_{u,r}-\widetilde{\bm{\eta}}^{n}_{u,r}}{\Delta t},\widetilde{\bm{w}}^{n+1}_{u,r}\right)\right|\leq C\left\|\frac{\widetilde{\bm{\eta}}^{n+1}_{u,r}-\widetilde{\bm{\eta}}^{n}_{u,r}}{\Delta t}\right\|^{2}_{0}+\frac{1}{10}\|\widetilde{\bm{w}}^{n+1}_{u,r}\|^{2}_{0}.

For the first term above, we utilize the expression of 𝒖~hn\widetilde{\bm{u}}^{n}_{h} and Πrv𝒖~hn\Pi^{v}_{r}\widetilde{\bm{u}}^{n}_{h} in the form of POD bases 𝝋i\bm{\varphi}_{i} to get

𝜼~u,rn+1=𝒖~hn+1Πrv𝒖~hn+1=k=1du~(𝒖~hn+1,𝝋k)𝝋kk=1r(𝒖~hn+1,𝝋k)𝝋k=k=r+1du~(𝒖~hn+1,𝝋k)𝝋k.\displaystyle\widetilde{\bm{\eta}}^{n+1}_{u,r}=\widetilde{\bm{u}}^{n+1}_{h}-\Pi^{v}_{r}\widetilde{\bm{u}}^{n+1}_{h}=\sum^{d_{\widetilde{u}}}_{k=1}\left(\widetilde{\bm{u}}^{n+1}_{h},\bm{\varphi}_{k}\right)\bm{\varphi}_{k}-\sum^{r}_{k=1}\left(\widetilde{\bm{u}}^{n+1}_{h},\bm{\varphi}_{k}\right)\bm{\varphi}_{k}=\sum^{d_{\widetilde{u}}}_{k=r+1}\left(\widetilde{\bm{u}}^{n+1}_{h},\bm{\varphi}_{k}\right)\bm{\varphi}_{k}.

Then

𝜼~u,rn+1𝜼~u,rnΔt02\displaystyle\left\|\frac{\widetilde{\bm{\eta}}^{n+1}_{u,r}-\widetilde{\bm{\eta}}^{n}_{u,r}}{\Delta t}\right\|^{2}_{0} =1Δt(k=r+1du~(𝒖~hn+1,𝝋k)𝝋kk=r+1du~(𝒖~hn,𝝋k)𝝋k)02,\displaystyle=\left\|\frac{1}{\Delta t}\left(\sum^{d_{\widetilde{u}}}_{k=r+1}\left(\widetilde{\bm{u}}^{n+1}_{h},\bm{\varphi}_{k}\right)\bm{\varphi}_{k}-\sum^{d_{\widetilde{u}}}_{k=r+1}\left(\widetilde{\bm{u}}^{n}_{h},\bm{\varphi}_{k}\right)\bm{\varphi}_{k}\right)\right\|^{2}_{0},
=k=r+1du~(𝒖~hn𝒖~hnΔt,𝝋k)𝝋k02,\displaystyle=\left\|\sum^{d_{\widetilde{u}}}_{k=r+1}\left(\frac{\widetilde{\bm{u}}^{n}_{h}-\widetilde{\bm{u}}^{n}_{h}}{\Delta t},\bm{\varphi}_{k}\right)\bm{\varphi}_{k}\right\|^{2}_{0},
=k=r+1du~(𝒖~hn+1,𝝋k)𝝋k02,\displaystyle=\left\|\sum^{d_{\widetilde{u}}}_{k=r+1}\left(\partial\widetilde{\bm{u}}^{n+1}_{h},\bm{\varphi}_{k}\right)\bm{\varphi}_{k}\right\|^{2}_{0},
=𝒖~hn+1Πrv𝒖~hn+102.\displaystyle=\left\|\partial\widetilde{\bm{u}}^{n+1}_{h}-\Pi^{v}_{r}\partial\widetilde{\bm{u}}^{n+1}_{h}\right\|^{2}_{0}.

For other terms on the right-side hand of (4.24),

|A2|\displaystyle|A_{2}| =|ν(𝜼~u,rn+1,𝒘~u,rn+1)|Cν𝜼~u,rn+102+12ν𝒘~u,rn+102Cνi=r+1du~γiφi02+12ν𝒘~u,rn+102,\displaystyle=\left|\nu(\nabla\widetilde{\bm{\eta}}^{n+1}_{u,r},\nabla\widetilde{\bm{w}}^{n+1}_{u,r})\right|\leq C\nu\|\nabla\widetilde{\bm{\eta}}^{n+1}_{u,r}\|^{2}_{0}+\frac{1}{2}\nu\|\nabla\widetilde{\bm{w}}^{n+1}_{u,r}\|^{2}_{0}\leq C\nu\sum^{d_{\widetilde{u}}}_{i=r+1}\gamma_{i}\|\nabla\varphi_{i}\|^{2}_{0}+\frac{1}{2}\nu\|\nabla\widetilde{\bm{w}}^{n+1}_{u,r}\|^{2}_{0},
|A3|\displaystyle|A_{3}| =|((phn+1phn),𝒘~u,rn+1)|C(phn+1phn)02+110𝒘~u,rn+102CΔttntn+1(tph)02dt+110𝒘~u,rn+102,\displaystyle=\left|(\nabla(p^{n+1}_{h}-p^{n}_{h}),\widetilde{\bm{w}}^{n+1}_{u,r})\right|\leq C\|\nabla(p^{n+1}_{h}-p^{n}_{h})\|^{2}_{0}+\frac{1}{10}\|\widetilde{\bm{w}}^{n+1}_{u,r}\|^{2}_{0}\leq C\Delta t\int^{t_{n+1}}_{t_{n}}\|\nabla(\partial_{t}p_{h})\|^{2}_{0}\,{\rm d}t+\frac{1}{10}\|\widetilde{\bm{w}}^{n+1}_{u,r}\|^{2}_{0},
|A4|\displaystyle|A_{4}| =|((prnprn+1),𝒘~u,rn+1)|C(prn+1prn)02+110𝒘~u,rn+102CΔttntn+1(tpr)02dt+110𝒘~u,rn+102,\displaystyle=\left|(\nabla(p^{n}_{r}-p^{n+1}_{r}),\widetilde{\bm{w}}^{n+1}_{u,r})\right|\leq C\|\nabla(p^{n+1}_{r}-p^{n}_{r})\|^{2}_{0}+\frac{1}{10}\|\widetilde{\bm{w}}^{n+1}_{u,r}\|^{2}_{0}\leq C\Delta t\int^{t_{n+1}}_{t_{n}}\|\nabla(\partial_{t}p_{r})\|^{2}_{0}\,{\rm d}t+\frac{1}{10}\|\widetilde{\bm{w}}^{n+1}_{u,r}\|^{2}_{0},
|A5|\displaystyle|A_{5}| =|(ηp,rn+1,𝒘~u,rn+1)|Cηp,rn+102+110𝒘~u,rn+102Ci=r+1dpϵiψi02+110𝒘~u,rn+102,\displaystyle=\left|(\nabla\eta^{n+1}_{p,r},\widetilde{\bm{w}}^{n+1}_{u,r})\right|\leq C\|\nabla\eta^{n+1}_{p,r}\|^{2}_{0}+\frac{1}{10}\|\widetilde{\bm{w}}^{n+1}_{u,r}\|^{2}_{0}\leq C\sum^{d_{p}}_{i=r+1}\epsilon_{i}\|\nabla\psi_{i}\|^{2}_{0}+\frac{1}{10}\|\widetilde{\bm{w}}^{n+1}_{u,r}\|^{2}_{0},
|A6|\displaystyle|A_{6}| =|(𝜼~u,rn+1,𝒘~u,rn+1)|C|(𝜼~u,rn+1,𝒘~u,rn+1)|C𝜼~u,rn+102+110𝒘~u,rn+102Ci=r+1du~γiφ02+110𝒘~u,rn+102,\displaystyle=\left|(\nabla\cdot\widetilde{\bm{\eta}}^{n+1}_{u,r},\widetilde{\bm{w}}^{n+1}_{u,r})\right|\leq C|(\nabla\widetilde{\bm{\eta}}^{n+1}_{u,r},\widetilde{\bm{w}}^{n+1}_{u,r})|\leq C\|\nabla\widetilde{\bm{\eta}}^{n+1}_{u,r}\|^{2}_{0}+\frac{1}{10}\|\widetilde{\bm{w}}^{n+1}_{u,r}\|^{2}_{0}\leq C\sum^{d_{\widetilde{u}}}_{i=r+1}\gamma_{i}\|\nabla\varphi\|^{2}_{0}+\frac{1}{10}\|\widetilde{\bm{w}}^{n+1}_{u,r}\|^{2}_{0},
|A7|\displaystyle|A_{7}| =|Δt(ηp,rn+1,wp,rn+1)|12Δtηp,rn+102+12Δtwp,rn+102CΔti=r+1dpϵiψi02+12Δtwp,rn+102.\displaystyle=\left|\Delta t(\nabla\eta^{n+1}_{p,r},\nabla w^{n+1}_{p,r})\right|\leq\frac{1}{2}\Delta t\|\nabla\eta^{n+1}_{p,r}\|^{2}_{0}+\frac{1}{2}\Delta t\|\nabla w^{n+1}_{p,r}\|^{2}_{0}\leq C\Delta t\sum^{d_{p}}_{i=r+1}\epsilon_{i}\|\nabla\psi_{i}\|^{2}_{0}+\frac{1}{2}\Delta t\|\nabla w^{n+1}_{p,r}\|^{2}_{0}.

Combining with all the seven terms’ results, we have

(𝒘~u,rn+1𝒘~u,rnΔt,𝒘~u,rn+1)\displaystyle\left(\frac{\widetilde{\bm{w}}^{n+1}_{u,r}-\widetilde{\bm{w}}^{n}_{u,r}}{\Delta t},\widetilde{\bm{w}}^{n+1}_{u,r}\right) +12ν𝒘~u,rn+102+12Δtwp,rn+102𝒖~hn+1Πrv𝒖~hn+102,\displaystyle+\frac{1}{2}\nu\|\nabla\widetilde{\bm{w}}^{n+1}_{u,r}\|^{2}_{0}+\frac{1}{2}\Delta t\|\nabla w^{n+1}_{p,r}\|^{2}_{0}\leq\left\|\partial\widetilde{\bm{u}}^{n+1}_{h}-\Pi^{v}_{r}\partial\widetilde{\bm{u}}^{n+1}_{h}\right\|^{2}_{0},
+CΔttntn+1(tph02+tpr02)dt+C(ν+1)i=r+1du~γiφi02,\displaystyle\quad+C\Delta t\int^{t_{n+1}}_{t_{n}}\Big{(}\|\partial_{t}p_{h}\|^{2}_{0}+\|\partial_{t}p_{r}\|^{2}_{0}\Big{)}\,{\rm d}t+C(\nu+1)\sum^{d_{\widetilde{u}}}_{i=r+1}\gamma_{i}\|\nabla\varphi_{i}\|^{2}_{0},
+C(Δt+1)i=r+1dpϵiψi02+12𝒘~u,rn+102.\displaystyle\quad+C(\Delta t+1)\sum^{d_{p}}_{i=r+1}\epsilon_{i}\|\nabla\psi_{i}\|^{2}_{0}+\frac{1}{2}\|\widetilde{\bm{w}}^{n+1}_{u,r}\|^{2}_{0}.

Multiplying by Δt\Delta t, adding with nn from n0n_{0} to N1N-1 and rearranging, we get

𝒘~u,rN02\displaystyle\|\widetilde{\bm{w}}^{N}_{u,r}\|^{2}_{0} +Δtn=n0N1𝒘~u,rn+1𝒘~u,rn02+νΔtn=n0N1𝒘~u,rn+102+Δt2n=n0N1wp,rn+102,\displaystyle+\Delta t\sum^{N-1}_{n=n_{0}}\|\widetilde{\bm{w}}^{n+1}_{u,r}-\widetilde{\bm{w}}^{n}_{u,r}\|^{2}_{0}+\nu\Delta t\sum^{N-1}_{n=n_{0}}\|\nabla\widetilde{\bm{w}}^{n+1}_{u,r}\|^{2}_{0}+\Delta t^{2}\sum^{N-1}_{n=n_{0}}\|\nabla w^{n+1}_{p,r}\|^{2}_{0},
Δtn=n0N1𝒖~hn+1Πrv𝒖~hn+102+C(ν+1)i=r+1du~γiφi02+12Δtn=n0N1𝒘~u,rn+102,\displaystyle\quad\leq\Delta t\sum^{N-1}_{n=n_{0}}\left\|\partial\widetilde{\bm{u}}^{n+1}_{h}-\Pi^{v}_{r}\partial\widetilde{\bm{u}}^{n+1}_{h}\right\|^{2}_{0}+C(\nu+1)\sum^{d_{\widetilde{u}}}_{i=r+1}\gamma_{i}\|\nabla\varphi_{i}\|^{2}_{0}+\frac{1}{2}\Delta t\sum^{N-1}_{n=n_{0}}\|\widetilde{\bm{w}}^{n+1}_{u,r}\|^{2}_{0},
+C(Δt+1)i=r+1dpϵiψi02+CΔt2tn0T(tph02+tpr02)dt.\displaystyle\quad+C(\Delta t+1)\sum^{d_{p}}_{i=r+1}\epsilon_{i}\|\nabla\psi_{i}\|^{2}_{0}+C\Delta t^{2}\int^{T}_{t_{n_{0}}}\Big{(}\|\partial_{t}p_{h}\|^{2}_{0}+\|\partial_{t}p_{r}\|^{2}_{0}\Big{)}\,{\rm d}t.

For the first term on the right-hand side above, we use Lemma 4.3 to have

Δtn=n0N1𝒖~hn+1Πrv𝒖~hn+102C1Mk=1M𝒖~hkΠrv𝒖~hk02Ci=r+1du~γi.\displaystyle\Delta t\sum^{N-1}_{n=n_{0}}\left\|\partial\widetilde{\bm{u}}^{n+1}_{h}-\Pi^{v}_{r}\partial\widetilde{\bm{u}}^{n+1}_{h}\right\|^{2}_{0}\leq C\frac{1}{M}\sum^{M}_{k=1}\|\partial\widetilde{\bm{u}}^{k}_{h}-\Pi^{v}_{r}\partial\widetilde{\bm{u}}^{k}_{h}\|^{2}_{0}\leq C\sum^{d_{\widetilde{u}}}_{i=r+1}\gamma_{i}.

By discrete Gronwall inequality and using stability result to bound the last term above, we get

𝒘~u,rN02\displaystyle\|\widetilde{\bm{w}}^{N}_{u,r}\|^{2}_{0} +Δtn=n0N1𝒘~u,rn+1𝒘~u,rn02+νΔtn=n0N1𝒘~u,rn+102+Δt2n=n0N1wp,rn+102,\displaystyle+\Delta t\sum^{N-1}_{n=n_{0}}\|\widetilde{\bm{w}}^{n+1}_{u,r}-\widetilde{\bm{w}}^{n}_{u,r}\|^{2}_{0}+\nu\Delta t\sum^{N-1}_{n=n_{0}}\|\nabla\widetilde{\bm{w}}^{n+1}_{u,r}\|^{2}_{0}+\Delta t^{2}\sum^{N-1}_{n=n_{0}}\|\nabla w^{n+1}_{p,r}\|^{2}_{0}, (4.25)
C(i=r+1du~γi(1+(ν+1)φi02)+(Δt+1)i=r+1dpϵiψi02+Δt2).\displaystyle\quad\leq C\left(\sum^{d_{\widetilde{u}}}_{i=r+1}\gamma_{i}\Big{(}1+(\nu+1)\|\nabla\varphi_{i}\|^{2}_{0}\Big{)}+(\Delta t+1)\sum^{d_{p}}_{i=r+1}\epsilon_{i}\|\nabla\psi_{i}\|^{2}_{0}+\Delta t^{2}\right).

Finally, by triangle inequality, Lemma 3.2 and Lemma 4.5, we obtain the following theorem which states the convergence between continuous variational form and POD projection scheme.

Theorem 4.2.

(error estimate for Proj-POD-ROM) For n=n0,n0+1,,Nn=n_{0},n_{0}+1,\cdots,N, let (𝐮n,pn)(\bm{u}^{n},p^{n}) is the solution of (2.3) at t=tnt=t_{n}, (𝐮~rn,prn)(\widetilde{\bm{u}}^{n}_{r},p^{n}_{r}) denotes the POD-based reduced-order solutions obtained in (4.17), then we have the following convergence estimate:

𝒖N𝒖~rN02\displaystyle\left\|\bm{u}^{N}-\widetilde{\bm{u}}^{N}_{r}\right\|^{2}_{0} +νΔtn=n0N1(𝒖~n𝒖~rn)L22+Δt2n=n0N1(ppr)L22,\displaystyle+\nu\Delta t\sum^{N-1}_{n=n_{0}}\left\|\nabla(\widetilde{\bm{u}}^{n}-\widetilde{\bm{u}}^{n}_{r})\right\|^{2}_{L^{2}}+\Delta t^{2}\sum^{N-1}_{n=n_{0}}\left\|\nabla(p-p_{r})\right\|^{2}_{L^{2}}, (4.26)
C(i=r+1du~γi(1+(ν+1)φi02)+(Δt+1)i=r+1dpϵiψi02+Δt2+h2).\displaystyle\leq C\left(\sum^{d_{\widetilde{u}}}_{i=r+1}\gamma_{i}\Big{(}1+(\nu+1)\left\|\nabla\varphi_{i}\right\|^{2}_{0}\Big{)}+(\Delta t+1)\sum^{d_{p}}_{i=r+1}\epsilon_{i}\left\|\nabla\psi_{i}\right\|^{2}_{0}+\Delta t^{2}+h^{2}\right).

5 Numerical Tests

In this section, we present some numerical experiments to confirm the aprioria\,priori error estimates derived in Theorem 4.2 for the POD projection scheme 4 - Proj-POD-ROM. To this end, we first determine some necessary parameters used before and verify the convergence in Lemma 3.2 and Lemma 3.3 of the finite element projection scheme, since with which we formed our POD snapshot spaces, and also the convergence result in Theorem 4.2 is closely related to the one of the finite element scheme. The open-source finite element package iFEM [6] has been used to run the numerical experiments.

5.1 Problem setting

We take Ω=[0,1]×[0,1]2\Omega=[0,1]\times[0,1]\subset\mathbb{R}^{2} and the time interval (0,1](0,1] with the viscosity coefficient ν=1\nu=1 and the prescribed solution

𝒖\displaystyle\bm{u} =cos(t)(πsin(πx)2sin(2πy)πsin(2πx)sin(πy)2),\displaystyle=\cos(t)\cdot\left(\begin{array}[]{c}\pi\sin(\pi x)^{2}\sin(2\pi y)\\ -\pi\sin(2\pi x)\sin(\pi y)^{2}\\ \end{array}\right),
p\displaystyle p =cos(t)10cos(πx)cos(πy).\displaystyle=\cos(t)\cdot 10\cos(\pi x)\cos(\pi y).

The right-hand side, the Dirichlet boundary condition and the initial condition are chosen in accordance to the above prescribed solution.
We set Δt=0.1h2\Delta t=0.1h^{2} and all grids are regular N×NN\times N triangular grids with SWNE diagonals for different NN(i.e., diagonals coming from connecting the southwest and the northeast vertexes on all rectangles), and we take N=4,8,16,32,64N=4,8,16,32,64 sequentially. For simplicity, we use P1P^{1}-P1P^{1} element pair for spatial discretization, and we take the snapshots on the finest computational mesh, i.e., h=1/64, and show this mesh in Figure 1.
When constructing the snapshots spaces, we take into account the fact that the finite element projection scheme involves the initial pressure ph0p^{0}_{h} which is not the part of the definition of the problem, so we take ph0=0p^{0}_{h}=0 and it might be good to think of the numerical discrete errors of pressure pnphnL2\|p^{n}-p^{n}_{h}\|_{L^{2}} in the previous steps nn are so large that it is not suitable to take finite element solution on those steps into snapshot spaces, thus we choose the finite element solution (𝒖hn,phn)(\bm{u}^{n}_{h},p^{n}_{h}) and its difference quotients (𝒖hn,phn)(\partial\bm{u}^{n}_{h},\partial p^{n}_{h}) from n=n0=6n=n_{0}=6 to n=25n=25 to formulate snapshots spaces, which means we take M=20M=20 and thus the number of snapshots is Ns=2M1=39N_{s}=2M-1=39.
The finite element solution (𝒖hn,𝒖~jn,phn)(\bm{u}^{n}_{h},\widetilde{\bm{u}}^{n}_{j},p^{n}_{h}) and exact solution (𝒖n,pn)(\bm{u}^{n},p^{n}) at discrete termination time t=ΔtNt=\Delta t\cdot N on finest mesh are shown in Figure 2, and following POD bases are formed from snapshots via L2L^{2} inner product.

Refer to caption
Figure 1: Computational mesh with h=1/64.
Refer to caption
Figure 2: The exact solution and finite element solution at n=Nn=N on mesh of h=1/64h=1/64. On upper row, from left to right: (u1n,u~1,hn,u1,hn,pnu^{n}_{1},\widetilde{u}^{n}_{1,h},u^{n}_{1,h},p^{n}); on lower row, from left to right: (u2n,u~2,hn,u2,hn,phnu^{n}_{2},\widetilde{u}^{n}_{2,h},u^{n}_{2,h},p^{n}_{h}).

5.2 Convergence of the finite element projection scheme

In this subsection, numerical tests will be used to numerically verify two theoretical results, that is Lemma 3.2 and Lemma 3.3, with which we utilize to get convergence analysis Theorem 4.2 of POD scheme 4. Although Lemma 3.2 has been analyzed in [10], it lacks the corresponding numerical verifications, so we supplement the numerical tests about Lemma 3.2 and also provide that of the Lemma 3.3, which is newly analyzed in this paper. Since Δt=𝒪(h2)\Delta t=\mathcal{O}(h^{2}), then from the Table 1 and Table 2 we can see

(𝒖𝒖~h)l2(L2)+pphl2(L2)\displaystyle\|\nabla(\bm{u}-\bm{\widetilde{u}}_{h})\|_{l^{2}(L^{2})}+\|p-p_{h}\|_{l^{2}(L^{2})} =𝒪(h),\displaystyle=\mathcal{O}(h),
𝒖n𝒖~hn0+hpnphn0+hΔt(pnphn)0\displaystyle\|\bm{u}^{n}-\widetilde{\bm{u}}^{n}_{h}\|_{0}+h\|p^{n}-p^{n}_{h}\|_{0}+h\sqrt{\Delta t}\|\nabla(p^{n}-p^{n}_{h})\|_{0} =𝒪(h2).\displaystyle=\mathcal{O}(h^{2}).

which are what the theoretical results in Lemma 3.2 shows. Especially, we can see from Table 1, 𝒖n𝒖hnL2=𝒪(h2)\left\|\bm{u}^{n}-\bm{u}^{n}_{h}\right\|_{L^{2}}=\mathcal{O}(h^{2}), which is in accordance of Lemma 3.3.

Table 1: Errors and convergence orders of finite element solution 𝒖~hn,𝒖hn\widetilde{\bm{u}}^{n}_{h},\bm{u}^{n}_{h} with P1P^{1}-P1P^{1} pair.
1/h maxn𝒖n𝒖~hnL2\max\limits_{n}\left\|\bm{u}^{n}-\widetilde{\bm{u}}^{n}_{h}\right\|_{L^{2}} maxn𝒖n𝒖hnL2\max\limits_{n}\left\|\bm{u}^{n}-\bm{u}^{n}_{h}\right\|_{L^{2}} (𝒖𝒖~h)2(L2)\left\|\nabla(\bm{u}-\widetilde{\bm{u}}_{h})\right\|_{\ell^{2}(L^{2})}
error rate error rate error rate
4  5.3013e-01  -  4.6509e-01  -  4.8187e++00  -
8  1.6490e-01 1.7078 1.4931e-01 1.6392 2.6626e++00 0.85582
16 4.3368e-02 1.9259 4.0108e-02 1.8963 1.3785e++00 0.94976
32 1.0969e-02 1.9828 1.0527e-02 1.9298 7.1098e-01 0.95516
64 2.7499e-03 1.9960 2.8273e-03 1.8966 3.7409e-01 0.92642
Table 2: Errors and convergence orders of finite element pressure phnp^{n}_{h} with P1P^{1}-P1P^{1} pair.
1/h maxnpnphnL2\max\limits_{n}\|p^{n}-p^{n}_{h}\|_{L^{2}} pph2(L2)\left\|p-p_{h}\right\|_{\ell^{2}(L^{2})} Δt(pph)2(L2)\sqrt{\Delta t}\left\|\nabla(p-p_{h})\right\|_{\ell^{2}(L^{2})}
error rate error rate error rate
4  2.7636e++00  -  2.2987e++00  -  1.4286e++00  -
8 1.1144e++00 1.3103 8.8892e-01 1.3707 4.6814e-01 1.6096
16 3.6664e-01 1.6038 2.7275e-01 1.7045 1.3827e-01 1.7595
32 1.2463e-01 1.5567 8.1260e-02 1.7470 4.5158e-01 1.6144
64 4.6335e-02 1.4275 2.5152e-02 1.6919 1.5553e-01 1.5378

5.3 Convergence of projection POD

After having the snapshots spaces, we determine du~=rank(𝒰~)=20d_{\widetilde{u}}=\text{rank}(\widetilde{\mathcal{U}})=20(for which γi<1015,i>du~\gamma_{i}<10^{-15},i>d_{\widetilde{u}}), dp=rank(𝒫)=20d_{p}=\text{rank}(\mathcal{P})=20(for which ϵi<1013,i>dp\epsilon_{i}<10^{-13},i>d_{p}). Figure 3 shows the decay of POD eigenvalues(left) of velocity γi,i=1,,du~\gamma_{i},i=1,\cdots,d_{\widetilde{u}} and pressure ϵi,i=1,,dp\epsilon_{i},i=1,\cdots,d_{p}, together with the corresponding captured system’s energy(right) in the form of 100i=1rγi/i=1du~γi100\sum^{r}_{i=1}\gamma_{i}/\sum^{d_{\widetilde{u}}}_{i=1}\gamma_{i} for velocity and 100i=1rϵi/i=1dpϵi100\sum^{r}_{i=1}\epsilon_{i}/\sum^{d_{p}}_{i=1}\epsilon_{i} for pressure. We note that the first r=4r=4 POD modes already capture more than 99.99%99.99\% of the system’s velocity-pressure energy.
In order to confirm the computational efficiency advantage of the POD-ROM scheme over the finite element FOM scheme, we take the number of POD modes is r=4r=4, whereas the number of degree of freedoms(DOFs) for velocity and pressure in finite element scheme are 8450 and 4225. We show in Table 3 the comparison of cumulative time spent by the two schemes to run to some certain time in the process of time-stepping iterative, and the L2L^{2} numerical spatial discrete error of velocity and pressure in two schemes at that time. We can see from Table 3 that, compared with the finite element FOM scheme, POD-ROM scheme can effectively improve computational efficiency in the context of only consuming approximately 1/6 times to get the numerical solution with even higher accuracy.
Figure 4 plots the temporal evolution of the discrete L2L^{2} relative error(in semilogarithmic scale) of the reduced-order velocity and pressure with respect to the full order ones: 𝒖~hn𝒖~rnL2/𝒖~hnL2\|\widetilde{\bm{u}}^{n}_{h}-\widetilde{\bm{u}}^{n}_{r}\|_{L^{2}}/\|\widetilde{\bm{u}}^{n}_{h}\|_{L^{2}} and phnprnL2/phnL2\|p^{n}_{h}–p^{n}_{r}\|_{L^{2}}/\|p^{n}_{h}\|_{L^{2}} for different number of POD modes. As expected in Theorem 4.2, the errors would decrease as the number of POD modes rr increasing. For velocity, the left figure in Figure 4 numerically demonstrate this result, and for pressure, the error would increase slightly when rr from 66 to 88, but when we continue to increase rr, the error will decrease, which is consistent with the Theorem 4.2 as a whole.

Refer to caption
Figure 3: POD velocity-pressure eigenvalues(left) and captured system’s velocity-pressure energy(right).
Table 3: Comparison of errors and time consumed by time-stepping iterative between finite element scheme with P1P^{1}-P1P^{1} pair and POD scheme with r=4r=4 on the mesh of h=1/64h=1/64 and Δt=0.1h2\Delta t=0.1h^{2}.
n finite element scheme POD scheme
𝒖n𝒖~hnL2\|\bm{u}^{n}\!-\!\widetilde{\bm{u}}^{n}_{h}\|_{L^{2}} pnphnL2\|p^{n}\!-\!p^{n}_{h}\|_{L^{2}} CPU run time(s) 𝒖n𝒖~rnL2\|\bm{u}^{n}\!-\!\widetilde{\bm{u}}^{n}_{r}\|_{L^{2}} pnprnL2\|p^{n}\!-\!p^{n}_{r}\|_{L^{2}} CPU run time(s)
2500 2.3789e-03 2.9458e-02 689 2.2860e-03 2.7823e-02 113
5000 2.3929e-03 2.9253e-02 1338 2.3007e-03 2.7642e-02 227
7500 2.3740e-03 2.8975e-02 2005 2.2826e-03 2.7379e-02 339
10000 2.3452e-03 2.8591e-02 2668 2.2549e-03 2.7015e-02 452
20000 2.1443e-03 2.6010e-02 5339 2.0618e-03 2.4573e-02 914
30000 1.8163e-03 2.1886e-02 8060 1.7464e-03 2.0673e-02 1399
40000 1.3805e-03 1.6464e-02 10746 1.3274e-03 1.5547e-02 1900
Refer to caption
Figure 4: Temporal evolution of L2L^{2} relative error of velocity 𝒖rn\bm{u}^{n}_{r} with respect to 𝒖hn\bm{u}^{n}_{h}(left) and velocity prnp^{n}_{r} with respect to phnp^{n}_{h}(right) .

6 Conclusions

In this paper, we proposed an efficient projection POD-ROM, which combined the advantages of classical projection method and POD technique.
The main contribution of the present paper consisted of two aspects: the first one was high computational efficiency. Through auxiliary intermediate velocity variable, the classical projection method decoupled the velocity variable and pressure variable, meanwhile decoupled the saddle-point system arose from Stokes equations, so one strength of projection method lied in its high computational efficiency, or low computational costs; Furthermore, POD technique was utilized to get the ROM, which have made the newly proposed projection POD-ROM had high computational efficiency. The second contribution was based on the fact that, in the fully discrete scheme of the classical projection method, the original scheme could be rewritten into a PSPG-type pressure stabilization scheme by eliminating the end-of-step velocity, where the pressure stabilized term Δt(phn+1,qhn+1)\Delta t(\nabla p^{n+1}_{h},\nabla q^{n+1}_{h}), Δt=𝒪(h2)\Delta t=\mathcal{O}(h^{2}), was inherent, so that some flexible mixed finite element spaces pairs(for example, P1P^{1}-P1P^{1} pair) could be used without considering the classical LBB/inf-sup condition for mixed POD spaces, which was different from other stabilized FE-POD-ROM to overcome LBB/inf-sup condition by adding extra stabilization terms.
Numerical experiments have been conducted to confirm the convergence for both projection finite element scheme and projection POD scheme. We first numerically confirmed the PSPG-type classical projection owned the desired convergence orders consistent with theoretical analysis, and then, finite element solutions were taken as the snapshots to formulate POD bases/modes which are used to get reduced-order velocity and pressure variables, so in theory, the discrete error between exact solutions and reduced-order solutions should converge to zero as increasing the number of POD modes, we numerically verify this result. Apart from the projection POD-ROM convergence, we also conduct the experiment to compare the error and computational time between projection finite element FOM and projection POD-ROM, and the results revealed the fact that projection POD-ROM not only had less discretization error, but also less computational costs, compared with projection finite element FOM.
One future research direction will be the applied of projection POD on nonlinear nonstationary Navier-Stokes equations. Another investigation is validation of fulfillment of LBB/inf-sup condition for mixed POD spaces.

Acknowledgment

The authors thank the anonymous referees for their constructive comments and suggestions which improved the manuscript.

References

  • [1] Mejdi Azaïez, Tomás Chacón Rebollo, and Samuele Rubino. A cure for instabilities due to advection-dominance in POD solution to advection-diffusion-reaction equations. J. Comput. Phys., 425:Paper No. 109916, 27, 2021.
  • [2] Francesco Ballarin, Andrea Manzoni, Alfio Quarteroni, and Gianluigi Rozza. Supremizer stabilization of POD-Galerkin approximation of parametrized steady incompressible Navier-Stokes equations. Internat. J. Numer. Methods Engrg., 102(5):1136–1161, 2015.
  • [3] Susanne C. Brenner and L. Ridgway Scott. The mathematical theory of finite element methods, volume 15 of Texts in Applied Mathematics. Springer, New York, third edition, 2008.
  • [4] Erik Burman and Miguel A. Fernández. Analysis of the PSPG method for the transient Stokes’ problem. Comput. Methods Appl. Mech. Engrg., 200(41-44):2882–2890, 2011.
  • [5] Alfonso Caiazzo, Traian Iliescu, Volker John, and Swetlana Schyschlowa. A numerical investigation of velocity-pressure reduced order models for incompressible flows. J. Comput. Phys., 259:598–616, 2014.
  • [6] Long Chen. ifem: an innovative finite element methods package in matlab. Preprint, University of Maryland, 2008.
  • [7] Alexandre Joel Chorin. Numerical solution of the Navier-Stokes equations. Math. Comp., 22:745–762, 1968.
  • [8] Alexandre Joel Chorin. On the convergence of discrete approximations to the Navier-Stokes equations. Math. Comp., 23:341–353, 1969.
  • [9] Philippe G. Ciarlet. The finite element method for elliptic problems, volume 40 of Classics in Applied Mathematics. Society for Industrial and Applied Mathematics (SIAM), Philadelphia, PA, 2002.
  • [10] Javier de Frutos, Bosco García-Archilla, and Julia Novo. Error analysis of projection methods for non inf-sup stable mixed finite elements. The transient Stokes problem. Appl. Math. Comput., 322:154–173, 2018.
  • [11] Victor DeCaria, Traian Iliescu, William Layton, Michael McLaughlin, and Michael Schneier. An artificial compression reduced order model. SIAM J. Numer. Anal., 58(1):565–589, 2020.
  • [12] Guosheng Fu and Zhu Wang. POD-(H)DG method for incompressible flow simulations. J. Sci. Comput., 85(2):Paper No. 24, 20, 2020.
  • [13] Vivette Girault and Pierre-Arnaud Raviart. Finite element methods for Navier-Stokes equations - Theory and algorithms, volume 5 of Springer Series in Computational Mathematics. Springer-Verlag, Berlin, 1986.
  • [14] J. L. Guermond, P. Minev, and Jie Shen. An overview of projection methods for incompressible flows. Comput. Methods Appl. Mech. Engrg., 195(44-47):6011–6045, 2006.
  • [15] J.-L. Guermond and L. Quartapelle. On the approximation of the unsteady Navier-Stokes equations by finite element projection methods. Numer. Math., 80(2):207–238, 1998.
  • [16] John G. Heywood and Rolf Rannacher. Finite element approximation of the nonstationary Navier-Stokes problem. I. Regularity of solutions and second-order error estimates for spatial discretization. SIAM J. Numer. Anal., 19(2):275–311, 1982.
  • [17] Volker John and Julia Novo. Analysis of the pressure stabilized Petrov-Galerkin method for the evolutionary Stokes equations avoiding time step restrictions. SIAM J. Numer. Anal., 53(2):1005–1031, 2015.
  • [18] Birgul Koc, Samuele Rubino, Michael Schneier, John Singler, and Traian Iliescu. On optimal pointwise in time error bounds and difference quotients for the proper orthogonal decomposition. SIAM J. Numer. Anal., 59(4):2163–2196, 2021.
  • [19] K. Kunisch and S. Volkwein. Galerkin proper orthogonal decomposition methods for parabolic problems. Numer. Math., 90(1):117–148, 2001.
  • [20] K. Kunisch and S. Volkwein. Galerkin proper orthogonal decomposition methods for a general equation in fluid dynamics. SIAM J. Numer. Anal., 40(2):492–515, 2002.
  • [21] Zhendong Luo and Goong Chen. Proper orthogonal decomposition methods for partial differential equations. Mathematics in Science and Engineering. Elsevier/Academic Press, London, 2019.
  • [22] Zhendong Luo, Jing Chen, I. M. Navon, and Xiaozhong Yang. Mixed finite element formulation and error estimates based on proper orthogonal decomposition for the nonstationary Navier-Stokes equations. SIAM J. Numer. Anal., 47(1):1–19, 2008/09.
  • [23] Bernd R. Noack, Paul Papas, and Peter A. Monkewitz. The need for a pressure-term representation in empirical galerkin models of incompressible shear flows. J. Fluid Mech., 523:339–365, 2005.
  • [24] Julia Novo and Samuele Rubino. Error analysis of proper orthogonal decomposition stabilized methods for incompressible flows. SIAM J. Numer. Anal., 59(1):334–369, 2021.
  • [25] Rolf Rannacher. On Chorin’s projection method for the incompressible Navier-Stokes equations. In The Navier-Stokes equations II—theory and numerical methods (Oberwolfach, 1991), volume 1530 of Lecture Notes in Math., pages 167–183. Springer, Berlin, 1992.
  • [26] S. S. Ravindran. Error analysis for Galerkin POD approximation of the nonstationary Boussinesq equations. Numer. Methods Partial Differential Equations, 27(6):1639–1665, 2011.
  • [27] Samuele Rubino. Numerical analysis of a projection-based stabilized POD-ROM for incompressible flows. SIAM J. Numer. Anal., 58(4):2019–2058, 2020.
  • [28] Jie Shen. On error estimates of projection methods for Navier-Stokes equations: first-order schemes. SIAM J. Numer. Anal., 29(1):57–77, 1992.
  • [29] Jie Shen. On error estimates of some higher order projection and penalty-projection methods for Navier-Stokes equations. Numer. Math., 62(1):49–73, 1992.
  • [30] John R. Singler. New POD error expressions, error bounds, and asymptotic results for reduced order models of parabolic PDEs. SIAM J. Numer. Anal., 52(2):852–876, 2014.
  • [31] Roger Temam. Une méthode d’approximation de la solution des équations de Navier-Stokes. Bull. Soc. Math. France, 96:115–152, 1968.
  • [32] S. Volkwein. Model reduction using proper orthogonal decomposition. Lecture Notes, Faculty of Mathematics and Statistics, University of Konstanz, 2011.