Loading [MathJax]/jax/output/SVG/jax.js
Special Issues

Efficient computations for linear feedback control problems for target velocity matching of Navier-Stokes flows via POD and LSTM-ROM

  • An efficient computing method for a target velocity tracking problem of fluid flows is considered. We first adopts the Lagrange multipliers method to obtain the optimality system, and then designs a simple and effective feedback control law based on the relationship between the control f and the adjoint variable w in the optimality system. We consider a reduced order modeling (ROM) of this problem for real-time computing. In order to improve the existing ROM method, the deep learning technique, which is currently being actively researched, is applied. We review previous research results and some computational results are presented.

    Citation: Hyung-Chun Lee. Efficient computations for linear feedback control problems for target velocity matching of Navier-Stokes flows via POD and LSTM-ROM[J]. Electronic Research Archive, 2021, 29(3): 2533-2552. doi: 10.3934/era.2020128

    Related Papers:

    [1] Hyung-Chun Lee . Efficient computations for linear feedback control problems for target velocity matching of Navier-Stokes flows via POD and LSTM-ROM. Electronic Research Archive, 2021, 29(3): 2533-2552. doi: 10.3934/era.2020128
    [2] Wei Shi, Xinguang Yang, Xingjie Yan . Determination of the 3D Navier-Stokes equations with damping. Electronic Research Archive, 2022, 30(10): 3872-3886. doi: 10.3934/era.2022197
    [3] Guoliang Ju, Can Chen, Rongliang Chen, Jingzhi Li, Kaitai Li, Shaohui Zhang . Numerical simulation for 3D flow in flow channel of aeroengine turbine fan based on dimension splitting method. Electronic Research Archive, 2020, 28(2): 837-851. doi: 10.3934/era.2020043
    [4] Matthew Gardner, Adam Larios, Leo G. Rebholz, Duygu Vargun, Camille Zerfas . Continuous data assimilation applied to a velocity-vorticity formulation of the 2D Navier-Stokes equations. Electronic Research Archive, 2021, 29(3): 2223-2247. doi: 10.3934/era.2020113
    [5] Jie Zhang, Gaoli Huang, Fan Wu . Energy equality in the isentropic compressible Navier-Stokes-Maxwell equations. Electronic Research Archive, 2023, 31(10): 6412-6424. doi: 10.3934/era.2023324
    [6] Derrick Jones, Xu Zhang . A conforming-nonconforming mixed immersed finite element method for unsteady Stokes equations with moving interfaces. Electronic Research Archive, 2021, 29(5): 3171-3191. doi: 10.3934/era.2021032
    [7] Jiwei Jia, Young-Ju Lee, Yue Feng, Zichan Wang, Zhongshu Zhao . Hybridized weak Galerkin finite element methods for Brinkman equations. Electronic Research Archive, 2021, 29(3): 2489-2516. doi: 10.3934/era.2020126
    [8] Caifeng Liu, Pan Liu . On Liouville-type theorem for the stationary compressible Navier–Stokes equations in $ \mathbb{R}^{3} $. Electronic Research Archive, 2024, 32(1): 386-404. doi: 10.3934/era.2024019
    [9] Guochun Wu, Han Wang, Yinghui Zhang . Optimal time-decay rates of the compressible Navier–Stokes–Poisson system in $ \mathbb R^3 $. Electronic Research Archive, 2021, 29(6): 3889-3908. doi: 10.3934/era.2021067
    [10] J. Bravo-Olivares, E. Fernández-Cara, E. Notte-Cuello, M.A. Rojas-Medar . Regularity criteria for 3D MHD flows in terms of spectral components. Electronic Research Archive, 2022, 30(9): 3238-3248. doi: 10.3934/era.2022164
  • An efficient computing method for a target velocity tracking problem of fluid flows is considered. We first adopts the Lagrange multipliers method to obtain the optimality system, and then designs a simple and effective feedback control law based on the relationship between the control f and the adjoint variable w in the optimality system. We consider a reduced order modeling (ROM) of this problem for real-time computing. In order to improve the existing ROM method, the deep learning technique, which is currently being actively researched, is applied. We review previous research results and some computational results are presented.



    The problem of fluid flow control has long been the focus of fluid mechanics research and is still an important research topic for many researchers. One of the most important fluid control problems is target velocity matching or tracking issues for fluid flow. The issue of target velocity matching or tracking problems for fluid flows has been a very active area of research e.g. [1,15,17,18,19,21,22,29,31] as well as others. The computation of optimally controlled flows based on solving an optimality system is expensive in terms of CPU time and memory space. This is due to the fact that such an approach involves a coupled system of state and adjoint equations with initial and final conditions. The system has to be solved on the entire space-time domain and cannot be solved by marching in time. Therefore, a feedback control design and/or reduced order modeling (ROM) is required for efficient and real time computing.

    Feedback control is a less expensive solution technique that gives qualitatively similar results to optimal control. Feedback control can effect good velocity tracking and at the same time the solution can be obtained step-by-step in time at the cost of a single flow solve. Mathematical theories and approximation techniques for optimal control problems for the Navier-Stokes equations have been developed in various areas of fluid dynamics; see, e.g., [1,17,21,22]. Some numerical methods for solving control problems for unsteady flows have been proposed and tested; see, e.g., [17,18,19,21,22] and references therein. Many of these deal with distributed controls and take different approaches to limiting the size of the control. Feedback controls for the Navier-Stokes equations have also been the subject of previous, mostly computational, studies; see, e.g., [11,16] and references therein. In this article, we consider desired states U in L2 which are not necessarily divergence-free, which is a feature not seen in previous work.

    Proper orthogonal decomposition (POD) is a common technique for extracting the dominant mode that contributes the most to the energy of the entire system [6]. POD combined with Galerkin projection (GP) have been used for many years to formulate ROMs for dynamic systems [5,8,9,32,24]. In such ROMs, the full-order set of equations is projected onto a reduced space, resulting in a dynamic system (modal coefficients) of much lower order than the full order model (FOM). However, in many situations there is a discrepancy between the governing equation and the observed system. This can be due to an approximation of the underlying phenomenon, incorrect parameterization, or insufficient information about the source terms contained in the system. This is especially evident in fluid flow systems where many complex phenomena and source terms interact in different ways. A common reduced order modeling (ROM) development procedure may be described by the following tasks:

    1. Reduced space and basis identification.

    2. Nonlinear dynamical system evolution in the reduced space.

    3. Reconstruction in full-order space for assessments.

    Machine learning (ML) tools have had considerable success in the fluid mechanics community, identifying basic structures and mimicking dynamics [3,7,25,26]. However, modeling with ML, especially deep learning, has faced strict opposition from both academia and industry alike because it can produce non-physical results due to its black box nature and its lack of interpretability and generalization [10,23]. In the course of conducting this study, we also experienced these points seriously. Even if the input and output data used to train the ML algorithm is physically accurate, the quantity interpolated with the ML approach can deviate significantly from a physically accurate solution. However, the fluid mechanics problem solving using deep learning is considered a promising research in the future, and we think it should be continued. A perspective on machine learning for advancing fluid dynamics can be found in a recent review articles [7] and references therein.

    The plan of the rest of the paper is as follows. In Section 2, we define and discuss the linear feedback control. We derive a time-space discretized version of the feedback control of Navier-Stokes equations. In Section 3, we explain the POD reduced basis and Galerkin Projection ROM. In Section 4, we introduce a closure model using deep neural networks. Finally, in section 5 some numerical results will be given.

    Let Ω be a bounded open set. We shall use the standard function spaces and their norms; for details see [2,13]. For any nonnegative integer m, we define the Sobolev spaces Hm(Ω) by

    Hm(Ω)={uL2(Ω):DαuL2(Ω)for0|α|m},

    where Dαu denotes the weak (or distributional) partial derivative and α is a multi-index, |α|=iαi. Note that H0(Ω)=L2(Ω). We equip Hm(Ω) with the norm

    ||u||2m=|α|m||Dαu||20

    The usual inner product associated with Hm(Ω) will be denoted by (,)m. Let Hm0(Ω) denote the closure of C0(Ω) under the norm ||||m and Hm0(Ω) be the dual space of Hm0(Ω).

    For vector-valued functions, we define the Sobolev space Hm(Ω) (in all cases, boldface indicates vector-valued) by

    Hm(Ω)={u|uiHm(Ω),i=1,2},

    where u=(u1,u2) and its associated norm by

    ||u||m=(2i=1||ui||2m)1/2.

    We also define particular subspace

    L20(Ω)={pL2(Ω):ΩpdΩ=0}.

    We introduce the solenoidal spaces

    V={uC0(Ω):u=0},V={uH10(Ω):u=0},W={uL2(Ω):u=0}.

    The spaces V and W are closure of V in H10(Ω) and L2(Ω), respectively. All subspaces are equipped with the norms inherited from the corresponding underlying spaces.

    In order to define a weak form of the Navier-Stokes equations, we introduce the continuous bilinear form

    a(u,v)=2ni,j=1ΩDij(u)Dij(v)dΩu,vH1(Ω), (1)
    b(v,q)=ΩqvdΩqL20,vH1(Ω), (2)

    where D(u)=(1/2)(u+uT) and the trilinear form

    c(w;u,v)=ni,j=1Ωwj(uixj)vidΩw,u,vH1(Ω). (3)

    A weak formulation of the Navier-Stokes equations is given by given gH1(Ω), find (u,p)H10(Ω)×L20(Ω) satisfying

    {ut,v+νa(u,v)+c(u;u,v)+b(v,p)=g,v,b(u,q)=0, (4)

    with initial velocity u0, where , denotes the duality paring between H10(Ω) and H10(Ω).

    We consider the target velocity fields UUad, where Uad is defined by

    {U=U(t,x)C([0,T];H2(Ω)H10(Ω)),U(t,x)=0xΩ. (5)

    We will consider the target velocity matching problem in which we want to minimize

    T0Ω|uU|2dΩdt.

    To do this, we define the following control and cost functional. Let fL2((0,T);L2(Ω)) denote the distributed control. Given T, we define the functional

    J(u,f)=T0Ω(α2|uU|2+β2|f|2)dΩdt+δ2Ω|u(T)U(T)|2dΩ. (6)

    The minimization of the T0Ω|uU|2dΩdt is the real goal of velocity matching problem, which is to keep the solution u to U over (0,T)×Ω. The T0Ωβ2|f|2dΩdt term is introduced in order to bound the control function and to prove the existence of the optimal control (regularization term). The Ω|u(T)U(T)|2dΩ term is necessary in order to keep the solution u to U near the time T. An optimal control problem is defined by

    seek (u,f)L2((0,T);H10(Ω))×L2((0,T);L2(Ω)) such that the cost functional (6) is minimized subject to constraints

    {ut,v+νa(u,v)+c(u;u,v)+b(v,p)=f,vvH10(Ω),b(u,q)=0L20(Ω), (7)

    with initial velocity u0H10(Ω).

    Using the Lagrange multipliers method, one can obtain the optimality system: seek uL2((0,T);H10(Ω)),pL2((0,T);L20(Ω)),wL2((0,T);H10(Ω)) and rL2((0,T);L20(Ω)) such that

    {ut,v+νa(u,v)+c(u;u,v)+b(v,p)=f,vvH10(Ω),b(u,q)=0L20(Ω), (8)

    with initial velocity u0H10(Ω) and the adjoint equations

    {wt,v+νa(w,v)+c(w;u,v)+c(u;w,v)+b(v,r)=α(uU),vvH10(Ω),b(w,q)=0L20(Ω), (9)

    with final velocity w(T,x)=δ(u(T)U(T)) and the optimality condition

    w=βf. (10)

    Several treatments of similar optimal control problems can be found in the literature, most notably in [1,17]. The numerical treatment of the velocity tracking problem is also an outstanding problem and other algorithms have been proposed. For example, a quasi-optimal control has been studied in [21,22].

    Remark 1. Thus the optimality system is a system of the coupled nonlinear partial differential equations (8)-(10). We already mentioned that solving an optimality system is very expensive in terms of CPU time and memory space. This is due to the fact that such an approach involves a coupled system of state and adjoint equations with initial and final conditions. The system has to be solved on the entire space-time domain and cannot be solved by marching in time. A gradient algorithm is needed and so the final algorithm is complex, involving multiple flow solutions, and the convergence may be slow.

    From (10), we see that a control f is related linearly with the adjoint variable w which is a solution of the adjoint equations (9) with f=(1/β)w. Also, we see that w is linearly dependent only on the source term of the adjoint equations (9) α(uU) since the adjoint equations are linear equations. So, we see that

    fαβ(uU). (11)

    Usually, β is very small real number to get a small value of the first term T0Ω|uU|2dΩdt in (6).

    f=γ(uU). (12)

    with very large value of γ. Remember that γ is closely related to β. The smaller β, the greater control f. Increasing control is not a problem in numerical simulation, but it is expensive in actual control problems. Therefore, γ can be set according to the actual problem.

    Now, we consider the feedback control problem

    {ut+(u)uν2u+p=γ(uU)in (0,T)×Ω,u=0in (0,T)×Ω,u(t,x)=0in (0,T)×Ω,u(0,x)=u0in Ω. (13)

    Remark 2. In [19], the control is achieved by means of a linear feedback law relating the body force to the velocity field, i.e., f=Fγ(uU) where U is in the set of the admissible target velocity Uad if U is a divergence free vector field in the set {c:cC((0,T);H2(Ω)H10(Ω)):tcC((0,T);H1(Ω))}. The corresponding body force generated by U is defined as

    F(t,u)=Ut(t,x)ν2U(t,x)+(U(t,x))U(t,x).

    In this article, UL((0,T);L2(Ω)) so F can not be defined and we do not include F in the control. We will compare our results with those in [19]. Our computational result is almost the same as that of [19] depending on the value of γ (see Figure 3).

    Figure 3.  ||u(t)U(t)||2, Compare between Figure 1 in [19] (left) and our results (right) where Re=1.

    Then, a linear feedback control problem of Navier-Stokes equations can be formulated by

    {(ut,v)+νa(u,v)+c(u;u,v)+b(v,p)γ(u,v)=γ(U,v)vH10(Ω),b(u,q)=0L20(Ω), (14)

    with initial velocity u0H10(Ω)

    A typical finite element approximation of (14) is defined as follows: we first choose conforming finite element subspaces VhH1(Ω) and ShL2(Ω) and then define Vh,0=VhH10(Ω), and Sh,0=ShL20(Ω). One then seeks uh(t,)Vh,0 and phSh,0 such that

    {(uht,vh)+νa(uh,vh)+c(uh;uh,vh)+b(vh,ph)γ(uh,vh)=γ(U,vh)for allvhVh,0,b(uh,qh)=0for allqhSh,0,uh(0,x)=u0,h(x)inΩ, (15)

    where u0,h(x)Vh,0 is an approximation, e.g., a projection, of u0(x). Discretization is completed by choosing a time-marching method such as the backward Euler scheme. See, e.g., [13] for details about finite element discretizations of the Navier-Stokes system.

    Let σN={tn}Nn=0 be a partition of [0,T] into equal intervals Δt=T/N with t0=0 and tN=T. Then a time discretization of (15) is given by

    {(u(n)hu(n1)hΔt,vh)+νa(u(n)h,vh)+c(u(n)h;u(n)h,vh)+b(vh,p(n)h)γ(u(n)h,vh)=γ(U,vh)for allvhVh,0,b(u(n)h,qh)=0for allqhSh,0,u(0)h(x)=u0,h(x)inΩ. (16)

    Since the trilinear term is nonlinear, a linearization of the discrete equations (16) must be introduced in order to solve it by a iteration method. We solve the following linearized equations:

    {1Δt(u(n)h(k),vh)+νa(u(n)h(k),vh)+σ˜c(u(n)h(k);u(n)h(k1),vh)+˜c(u(n)h(k1);u(n)h(k),vh)+b(vh,p(n)h(k))γ(u(n)h(k),vh)=1Δt(u(n1)h(k),vh)+σ˜c(u(n)h(k1);u(n)h(k1),vh)γ(U,vh)for allvhVh,0,b(u(n)h(k),qh)=0for allqhSh,0,u(0)h(x)=u0,h(x)inΩ, (17)

    where ˜c(u;v,w) is the modified trilinear form (see [12]) to preserve the antisymmetry of the trilinear form c(u;v,w) on the finite element spaces

    ˜c(u;v,w)=12{c(u;v,w)c(u;w,v)}u,v,wH10(Ω).

    Given the velocity u(n1)h and an integer m, one can generate the sequence {u(n)h(k), p(n)h} for k=1,2, by solving the linear problem with σ=0 for km and σ=1 for k>m. The algorithm terminates when the maximum value of

    |u(n)h(k)u(n1)h(k)||u(n)h(k)|

    is less or equal to τ on Ω. The global convergence of the simple iteration method (γ=0), and the rapid convergence of Newton's method (γ=1) are combined in this numerical algorithm. The value of m is set to be two or three, depending on the time step Δt.

    Let us briefly introduce the POD method following the formulation of Sirovich [30], Rathinam [28] and Burkardt et. al. [8,9]. POD provides a method for finding the best approximating subspace to a given set of data. Originally POD was used as data representation technique. Proper orthogonal decomposition (POD), also known as Karhunen–Lóeve decomposition or principal component analysis, provides a technique for analyzing multidimensional data. This method essentially provides an orthonormal basis for representing the given data in a certain least squares optimal sense.

    Given a discrete set of snapshot vectors W={wn}Nn=1 belonging to RJ, where N<J, we form the J×N snapshot matrix A whose columns are the snapshot vectors wn:

    A=(w1w2wN).

    Let

    UTAV=(Σ000),

    where U and V are J×J and N×N orthogonal matrices, respectively, and Σ=diag(σ1,,σ˜N) with σ1σ2σ˜N be the singular value decomposition of A. Here, ˜N is the rank of A, i.e., the dimension of the snapshot set W, which would be less than N whenever the snapshot set is linearly dependent. It is well known [14] that if

    U=(ϕ1ϕ2ϕJ)andV=(ψ1ψ2ψN),

    then

    Aψi=σiϕiandATϕi=σiψifor i=1,,˜N

    so that also

    ATAψi=σ2iψiandAATϕi=σ2iϕifor i=1,,˜N

    so that σ2i, i=1,,˜N, are the nonzero eigenvalues of ATA (and also of AAT) arranged in nondecreasing order. Note that the matrix C=ATA is simply the correlation matrix for the set of snapshot vectors W={wn}Nn=1, i.e., we have that Cmn=wTmwn.

    In the reduced-order modeling context, given a set of snapshots W={wn}Nn=1 belonging to RJ, the POD reduced-basis of dimension KN<J is the set {ϕk}Kk=1 of vectors also belonging to RJ consisting of the first K left singular vectors of the snapshot matrix A. Thus, one can determine the POD basis by computing the (partial) singular value decomposition of the J×N matrix A. Alternately, one can compute the (partial) eigensystem {σ2k,ψi}Ki=1 of the N×N correlation matrix C=ATA and then set ϕk=1σkAψk, k=1,,K.

    The K-dimensional POD basis has the obvious property of orthonormality. It also has several other important properties which we now mention. Let {sk}Kk=1 be an arbitrary set of K orthonormal vectors in RJ and let Πw denote the projection of a vector wRJ onto the subspace spanned by that set. Further, let

    E(s1,,sK)=Nn=1|wnΠwn|2,

    i.e., E is the sum of the squares of the error between each snapshot vector wn and its projection Πwn onto the span of {sk}Kk=1. Then, it can be shown that

    {the POD basis {ϕk}Kk=1 minimizes E over all possibleKdimensional orthonormal sets in RJ . (18)

    In fact, often the POD basis corresponding to a set of snapshots W={wn}Nn=1 and then its relation to the singular value decomposition of the matrix A or to the eigenvalue decomposition of ATA are derived properties. We note that E(ϕ1,,ϕK) is referred to as the "POD energy" or "error in the POD basis". Also, it can be shown that

    E(ϕ1,,ϕK)=˜Nk=K+1σ2k, (19)

    i.e., the error in the POD basis is simply the sum of the squares of the singular values corresponding to the neglected POD modes.

    We now show how a POD basis is used to define a reduced-order model for the Navier-Stokes system. For the sake of brevity, we only discuss the case for which the snapshot set is viewed as a set of finite element coefficient vectors; the case for which the snapshot set is a set of finite element functions proceeds in an almost identical manner.

    Let {ϕ}Kk=1 be a K-dimensional POD basis corresponding to the snapshot set {wn}Nn=1 and let

    UK=span{ϕk}Ki=1Vh.

    We then determine uKh(t,)UK from the discrete problem

    {(uKht,v)+νa(uKh,v)+c(uKh;uKh,v)γ(uKh,v)=γ(U,v)vUK(uKh(0,x),v)=(u0,h(x),v)vUK. (20)

    For a given reduced space UK, GP-ROM finds the approximation of the velocity field spanned by the low-dimensional space,

    uuKh(t,x)Kk=1ak(t)ϕk(x) (21)

    where {ak(t)}Kk=1 are the sought time-varying coefficients. The GP-ROM can be obtained by projecting the FOM (Full order finite element model) onto the POD space:

    {Kk=1ddtak(t)(ϕk,ϕ)+2νKk=1ak(t)(D(ϕk),D(ϕ))+(Km=1am(t)ϕmKk=1ak(t)ϕk,ϕ)γKk=1ak(t)(ϕk,ϕ)=γKk=1bk(ϕk,ϕ),Kk=1ak(0)(ϕk,ϕ)=(u0,ϕ),Kk=1bk(ϕk,ϕl)=(U,ϕ), (22)

    for =1,,K. Equivalently, we have the system of nonlinear ordinary differential equations that determine the coefficient functions {ak(t)}Kk=1:

    {Gddta(t)+Ka(t)+(a(t))TNa(t)γGa(t)=γGbGa(0)=a0, (23)

    or

    {ddta(t)+(G1KγI)a(t)+G1a(t)TNa(t)=γba(0)=G1a0, (24)

    where the Gram matrix G, stiffness matrix K, convection tensor N, and solution vector a(t) are respectively given by

    Gk=ΩϕkϕdΩIK,Kk=2νΩD(ϕk):D(ϕ)dΩ,Nmk=Ω(ϕm)ϕkϕdΩ,and(a)k=ak(t) (25)

    for k,,m=1,,K. Note that all of these matrices and tensors are full; however, since K will be chosen small, this does not cause any computational inefficiencies.

    It is easy to see that the computational cost of GP-ROM (24) is O(K3), which limits the number of modes to be used in the ROM. Several efforts have been devoted to introduce stabilization and closure techniques to account for the effects of truncated modes on ROM's dynamics. In this study, we utilize a LSTM closure model to improve the accuracy and efficiency for our computations.

    The model reduction error E ROM(t,x) is equal to

    E ROM(t,x)=u FOM(t,x)u ROM(t,x)=u FOM(t,x)u Proj(t,x)+u Proj(t,x)u ROM(t,x)=EUK(t,x)+EUK(t,x). (26)

    The first term EUK is the projection error that is a result of neglecting the components of u FOM(t,x) that lie in the space orthogonal to UK. This error is represented as EUK in Fig. 1. The second term EUK is the reduced order modeling error that results from solving a different dynamical system than the original. This error vector lies in the subspace UK spanned by the reduced order basis {ϕk}Ki=1. This error represents EUK in Fig. 1. For more details, one can refer the articles [3,4].

    Figure 1.  Discrepancy of the trajectories between the FOM, ROM and least squares projected solutions after Galerkin projection.

    The GP-ROM equations (24) with linear and nonlinear operators can be written as:

    ˙a=(LγI)aaTNaγb, (27)

    where L, I and N are the linear, identity and nonlinear operators in (22), respectively. The above equation (27) can be written more explicitly

    dak(t)dt=Ki=1(Likγ)ai(t)Ki=1Kj=1Nijkai(t)aj(t)γbk(t),k=1,2,,K, (28)

    which can be numerically solved using time-stepping integration,

    ak(tn+1)=ak(tn)Δtsq=0βqG(ak(tnq))βqγbk(t)n=0,,J1, (29)

    where s and βq depend on the numerical scheme. Since this problem is a non-stiff problem, we usually use the fourth-order Adams-Bashforth method, e.g., ode113 in Matlab. From (29), we see that

    G(ak(tn))=Ki=1(Likγ)ai(tn)Ki=1Kj=1Nijkai(tn)aj(tn). (30)

    To reduce the error E ROM, we use a closure model which can be written as:

    dak(t)dt=Ki=1(Likγ)ai(t)Ki=1Kj=1Nijkai(t)aj(t)γbk(t)+Ck(a1,,aK,α1,,αK), (31)

    where αk(t) is the Least-Squares (LS) projection modal coefficients which can be obtained by:

    αk(tn)=u(tn,x),ϕk,n=1,,J, (32)

    where the angle parentheses refer to the Euclidean inner product in RK. The LS projection modal coefficients include the hidden physics and its interaction with the dynamical core of the system. For simplicity, we assume that Ck depend only on ak and αk for each k. The correction function Ck can be defined in various way, especially and naturally,

    Ck(tn+1)=αk(tn)ak(tn) (33)

    and

    Ck(tn+1)=αk(tn+1)[αk(tn)+Δtsq=0βqG(αk(tnq))], (34)

    which are used in [25] and [26], respectively. In theory, they are different but we can not find the differences in computational performance. So, we will study more on these problem in the later articles. In this article, we adopt the equation (33) to the correction step in each time step tn. One can use any of the suitable machine learning algorithm to learn this correction term Ck such as ResNET and LSTM network, so on. In this article, we employ LSTM neural network algorithm to learn the mapping from LS projection modal coefficients to the correction term

    {α1,,αK}RK{C1,,CK}RK.

    Let σN={tn}Nn=0 be a partition of [0,T] into equal intervals Δt=T/N with t0=0 and tN=T. The finite element spaces are chosen to be piecewise quadratic for the velocity and linear for the pressure, i.e., the Taylor-Hood finite element pair, based on a rectangular mesh. We use FreeFEM++ [20].

    In this test we are interested in the convergence history for the parameters involved and so a simple stationary target velocity U=(U,V) is chosen, where

    U(x,y)=10dϕdy(0.4,x,y),V(x,y)=10dϕdx(0.4,x,y),ϕ(t,x,y)=(1cos(2πtx))(1x)2(1cos(2πty))(1y)2. (35)

    For the comparison purpose with the results in [19], we set up the initial velocity as follows:

    {u0(x,y)=10U(x,y),v0(x,y)=10V(x,y). (36)

    This initial velocity rotates in the opposite direction to the target velocity U. For this calculation, we use Δt=0.0001, h=1/64 and T=1.0. The number of unknowns (degree of freedom) is 49923.

    The flow evolution is given in Fig. 2. The controlled fluid is on the left and the desired flow is on the right. All the figures are normalized. At the beginning, we have a reduction in the magnitude, then we have a change in shape, and finally a change in magnitude again. This evolution is typical of linear feedback control. As shown in Fig. 2, the change in shape is so quick that it is difficult to see the full evolution without using very small time steps. The error ||u(t)U(t)|| rapidly goes to zero and a good match is reached at t=0.5. To compare computational results for our feedback law with those in [19], we compute ||u(t)U(t)||L2(Ω) for γ=10,30. Fig. 3 shows the L2 norm ||u(t)U(t)||L2(Ω) between the calculated discretized controlled flow (here denoted by u instead of uh) and the exact target flow U. We see that our feedback law is almost same effect to that of [19].

    Figure 2.  Controlled (left) and target (right) flows at t=0.0,0.050,0.057,0.059 (left pair of columns) and t=0.060,0.061, 0.064,0.5 (right pair of columns) for γ=10.

    Now, we set up the Reynolds number as Re=10 and the initial velocity for our computational experiments as follows:

    {u0(x,y)=U(x,y),v0(x,y)=V(x,y). (37)

    We generate the data snapshots for γ=[10,20,30,40] with a time-step of 0.001 from time t=0 to T=0.25. The ||u(t)U(t)||L2(Ω) for each cases of γ is shown in Fig. 4.

    Figure 4.  ||uU||L2(Ω) for different control values γ.

    The POD reduced basis is determined from the corresponding snapshot set as described in Section 3. Note that each basis function satisfies the discretized continuity equation, i.e., it is discretely solenoidal. The eight-dimensional POD basis functions are displayed in Fig. 5 (Upper two rows). We see that the difference of basis between those of γ=10 and γ=40.

    Figure 5.  The POD reduced basis of cardinality 8 (γ=10 Upper two rows and γ=40 Down two rows). Basis 1, 2, 3, 4 from the left to the right (Top) and Basis 5, 6, 7, 8 from the left to the right (Bottom) in each two rows.

    We can compute the energy retained by POD basis functions using a relative information content (RIC) formula as given below:

    RIC(K)=Ki=1σ2j˜Ni=1σ2j, (38)

    where σ's are singular values and ˜N is the total number of singular values. RIC represents the fraction of information of the the total data that can be recovered using K basis functions.

    Fig. 6 displays the convergence of relative information content with respect to the number of POD basis functions used to represent the reduced order system for different control number γ. For the snapshot set determined as described in Section 2.2 with γ=[10,20,30,40], the first 8 singular values of the corresponding snapshot matrix are listed in Table 1.

    Figure 6.  The singular values of the snapshot data matrix (right) for different control values γ.
    Table 1.  The first 8 singular values of the snapshot matrix.
    singular vales RIC(γ=40) singular vales RIC(γ=40)
    1 4.83907e+02 99.2476% 2 2.41493e+00 99.8174%
    3 3.98155e-01 99.9207% 4 2.02669e-01 99.9731%
    5 6.01101e-02 99.9909% 6 1.65390e-02 99.9964%
    7 6.20453e-03 99.9985% 8 2.51213e-03 99.9994%

     | Show Table
    DownLoad: CSV

    We retained eight basis functions (i.e., K=8) as they captured more than 99.997% of the energy for all control numbers γ=[10,20,30,40] (see Fig. 6 and Table 1 for γ=40). It can be seen that the singular values for the feedback control problem decreases rapidly. We consider here, a "small-dimensional" POD basis can capture most of the information contained in the snapshot set.

    Now, given the desired state U, we use K-dimensional system of of nonlinear ordinary differential equations (24) to determine reduced-order solutions of the Navier-Stokes system. Approximations of solutions a(n)k of the system of ordinary differential equation (27) are determined using a fourth order Adams-Moulton method. To quantify the performance of different frameworks, we defined the root mean squared error (RMSE) between the FOM velocity field and the velocity field computed with different ROM frameworks. The RMSE is defined as:

    RMSE(tn)=1nxnynxi=1nyj=1(u FOM(xi,yj,tn)u ROM(xi,yj,tn))2 (39)

    and

    RMSET=1ntntn=1RMSE(tn). (40)

    Fig. 7 shows the RMSE(t) over time for four cases γ=[10,20,30,40]. The LS Projection error is EUK(t,x)=u FOM(t,x)u Proj(t,x) and GP-ROM error is E ROM(t,x)=u FOM(t,x)u ROM(t,x). We calculate the LS projection modal coefficients by orthogonally projecting the state variables onto the reduced spaces. We will use the differences between RSME of LS projection and that of GP-ROM in the LSTM-ROM study, which will be studied in the next subsection.

    Figure 7.  RMSE of Galerkin ROM (Upper 4 lines) and LS Projection (Lower 4 lines).

    In this subsection, we test GP-LSTM closure model for the out-of-sample condition γ=50. For testing our GP-LSTM closure model, we have generated four cases in different values of parameters such that γ=[10,20,30,40].

    The LS projection modal coefficients include the hidden physics and its interaction with the dynamical core of the system. We can then define the correction term as:

    Correction:=C(n)k=α(n)ka(n+1)k (41)

    A supervised learning framework is applied to model the correction term C with information obtained from FOM and projection data. We train our LSTM neural network to learn the mapping from GP modal coefficients to the correction term. Since GP modal coefficients is used as input features to the LSTM network, the parameter γ governing the system's behavior is taken implicitly into account. We train the LSTM network with two hidden layers and 120 cells.

    Fig. 8 shows a sketch of Long Short-Term Memory unit and an architecture of LSTM network training. For more details for LSTM architecture, one can refer to the articles [27] and references therein. Programs obtained from GitHub, namely ETC-ROM Master, Hybrid-Modeling Master, UROM Master and mnni-rom Master, are adopted, modified and used in the learning process.

    Figure 8.  LSTM Cell (left) and LSTM Training where r=5.

    Once the model is trained, we could correct the GP modal coefficients with LSTM-based correction to approximate true projection modal coefficients.

    a(n+1)k=a(n+1)kΔt(Ki=1(Likγ)a(n)i+Ki=1Kj=1Nijka(n)ia(n)j+γb(n)k+C(n+1)k)

    We give the numerical procedure as follows:

    1. Data Generation: For γ=10,20,30,40 to use in learning algorithm and γ=50 to use a test.

    2. Basis Construction: For γ=10,20,30,40.

    3. The velocity modal coefficients a(n)k of Galerkin projection reduced order modeling calculation for γ=10,20,30,40 using each corresponding basis.

    4. LSTM Training using C(n+1)k and a(n+1)k. A summary of the adopted hyper-parameters is presented in Table 2. The training and validation loss during the learning procedure are in Fig. 9.

    Table 2.  A list of hyperparameters utilized to train the LSTM network for numerical experiments.
    Variables Hyperparameters
    Number of hidden layers 2
    Number of neurons in each hidden layer 120
    Number of lookbacks 5
    Batch size 32
    Epochs 1000
    Activation functions in the LSTM layers tanh
    Validation data set 20%
    Loss function MSE
    Optimizer ADAM

     | Show Table
    DownLoad: CSV
    Figure 9.  Training and validation loss.

    5. Prediction for γ=50 using the basis which is obtained from the snapshot sets with γ=40 (see Fig. 10).

    Figure 10.  Temporal evolution of the 8 POD modal coefficients predicted by LS Projection, GP-ROM and GP-LSTM for γ=50 using the basis from γ=40.

    6. Reconstruction and compare with GP-ROM and LS Projection (see Fig. 11).

    Figure 11.  RMSE of ROMs predicted by LS Projection, GP-ROM and GP-LSTM for γ=50 using the basis of γ=40.

    In step 5, interpolation techniques such as Grassman manifold interpolation [3,4,25], or the discrete empirical interpolation method, (DEIM) are typically applied to postprocess the results. For this work, however, such interpolations were not applied, so as to focus on the effects obtained from the Deep Learning technique. Thus, we use the same basis which is obtained from the snapshot sets with γ=40. The POD modal coefficients predicted by these approaches are in Figure 10. As shown in the figure 10, it can be seen that the error of GP-LSTM is much smaller than that of GP-ROM.

    Finally, we report the "online" computing time for our numerical experiments. We show the computational time as well as the RMSE of reconstructed fields at the final time t=0.25 for the test cases at γ=50 in Table 3. All calculations were performed on an iMac 3.6Ghz 8 Core. We observe that computing time of GP-LSTM is about twice that of GP-ROM(8) and the error of GP-LSTM(8) is smaller (about 1/10) than that of GP-ROM(8). If we use as much data as and use interpolations like in other papers, for example ([3,25,26]), one can reduce the error even more. However, in this papers, LSTM is applied in the simplest way, to see the possibility of applying the deep learning technique to the optimal control problem or UQ problem of fluid flows.

    Table 3.  CPU time (in second) comparison for the different ROM frameworks investigated in this study and RMSE at t=0.25.
    Framework Times RMSE(0.25)
    FOM 1695.7 0
    LS-Proj 1.34053e-05
    GP-ROM(8) 0.3268 3.77711e-03
    GP-ROM (16) 1.7012 1.28362e-03
    GP-LSTM(8) 0.6372 3.20970e-04

     | Show Table
    DownLoad: CSV

    In this study, a simple feedback rule is developed in order to reduce the amount of computation. It has been shown that our feedback law works very well. It is believed that mathematical proof will be possible without much difficulty. For real-time computation, a numerical experiment was performed by adopting a GP-ROM. In order to increase the GP-ROM's accuracy, the deep learning method, especially the LSTM method, which has been actively developed recently, was studied and applied. The ROM using deep learning such as LSTM performed in this study is considered to be worth continuing research. However, it is difficult to study systematically because the mathematical theory is not supported. We intend to apply GP-LSTM and GP-ResNet methods to the next studies, the optimal control problems and the uncertainty quantification problems of fluid flows.

    We thank the reviewer for the comments that helped improve the paper.



    [1] On some control problems in fluid mechanics. Theoret. Comput. Fluid Dynamics (1990) 1: 303-325.
    [2] (1975) Sobolev Spaces. Academic Press.
    [3] S. E. Ahmed, O. San, A. Rasheed and T. Iliescu, A long short-term memory embedding for hybrid uplifted reduced order models, Phys. D, 409 (2020), 132471, 16 pp. doi: 10.1016/j.physd.2020.132471
    [4] D. Amsallem, Interpolation on Manifolds of CFD-Based Fluid and Finite Element-Based Structural Reduced-Order Models for On-Line Aeroelastic Predictions, Ph. D. Thesis, Stanford University, 2010.
    [5] A survey of projection-based model reduction methods for parametric dynamical systems. SIAM Rev. (2015) 57: 483-531.
    [6] The proper orthogonal decomposition in the analysis of turbulent flows. Annual review of fluid mechanics (1993) 25: 539-575.
    [7] Machine Learning for Fluid Mechanics. Annu. Rev. Fluid Mech. (2020) 52: 477-508.
    [8] Centroidal voronoi tessellation-based reduced-order modeling of complex systems. SIAM J. Sci. Comput. (2006) 28: 459-484.
    [9] POD and CVT-based reduced-order modeling of Navier-Stokes flows. Comput. Methods Appl. Mech. Engrg. (2006) 196: 337-355.
    [10] Theory-guided data science for climate change. Computer (2014) 47: 74-78.
    [11] Active control of vortex shedding. J. Fluids Struct. (1989) S3: 115-122.
    [12] V. Girault and P. Raviart, Navier-Stokes Equations, North-Hollan, Amsterdam, 1979.
    [13] V. Girault and P.-A. Raviart, Finite Element Methods for Navier-Stokes Equations, Springer-Verlag, Berlin, 1986. doi: 10.1007/978-3-642-61623-5
    [14] G. H. Golub and C. F. Van Loan, Matrix Computations, Johns Hopkins University, Baltimore, 1996.
    [15] Finite-dimensional approximation of a class of constrained nonlinear optimal control problems. SIAM J. Control Optim. (1996) 34: 1001-1043.
    [16] Active control of vortex shedding. J. Appl. Mech. (1996) 63: 828-835.
    [17] Analysis and approximation of the velocity tracking problem for Navier-Stokes flows with distributed control. SIAM J. Numer. Anal. (2000) 37: 1481-1512.
    [18] The velocity tracking problem for Navier-Stokes flows with boundary control. SIAM J. Control Optim. (2000) 39: 594-634.
    [19] Analysis and approximation for linear feedback control for tracking the velocity in Navier-Stokes flows. Comput. Methods Appl. Mech. Engrg. (2000) 189: 803-823.
    [20] New development in FreeFem++. J. Numer. Math. (2012) 20: 251-265.
    [21] Dynamics for controlled Navier-Stokes systems with distributed controls. SIAM J. Control Optim. (1997) 35: 654-677.
    [22] Dynamics and approximations of a velocity tracking problem for the Navier-Stokes flows with piecewise distributed controls. SIAM J. Control Optim. (1997) 35: 1847-1885.
    [23] Theory-guided data science: A new paradigm for scientific discovery from data. IEEE Trans. Knowl. Data Eng. (2017) 29: 2318-2331.
    [24] Strategies for reduced-order models for predicting the statistical responses and uncertainty quantification in complex turbulent dynamical systems. SIAM Rev. (2018) 60: 491-549.
    [25] S. Pawar, S. Ahmed, O. San and A. Rasheed, An evolve-then-correct reduced order model for hidden fluid dynamics. Mathematics, Mathematics, 8 (2020), 570.
    [26] S. Pawar, S. E. Ahmed, O. San and A. Rasheed, Data-driven recovery of hidden physics in reduced order modeling of fluid flows, preprint, arXiv: 1910.13909 doi: 10.1063/5.0002051
    [27] M. Rahman, S. Pawar, O. San, A. Rasheed and T. Iliescu, A non-intrusive reduced order modeling framework for quasi-geostrophic turbulence, preprint, arXiv: 1906.11617
    [28] A new look at proper orthogonal decomposition. SIAM J. Numer. Anal. (2003) 41: 1893-1925.
    [29] L. Scarpa, Analysis and optimal velocity control of a stochastic convective Cahn-Hilliard equation, preprint, arXiv: 2007.14735
    [30] Turbulence and the dynamics of coherent structures, part ⅰ: Coherent structures; part ⅱ: symmetries and transformations; part ⅲ: Dynamics and scaling. Quart. Appl. Math. (1987) 45: 561-590.
    [31] M. Strazzullo, Z. Zainib, F. Ballarin and G. Rozza, Reduced order methods for parametrized non-linear and time dependent optimal flow control problems, towards applications in biomedical and environmental sciences, preprint, arXiv: 1912.07886
    [32] Kailai Xu, Bella Shi and Shuyi Yin, Deep Learning for Partial Differential Equations, Stanford University, 2018.
  • Reader Comments
  • © 2021 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

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

Metrics

Article views(2375) PDF downloads(310) Cited by(0)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog