Skip to main content
Log in

Reduced-order control using low-rank dynamic mode decomposition

  • Original Article
  • Published:
Theoretical and Computational Fluid Dynamics Aims and scope Submit manuscript

Abstract

In this work, we perform full-state LQR feedback control of fluid flows using non-intrusive data-driven reduced-order models. We propose a model reduction method called low-rank dynamic mode decomposition (lrDMD) that solves for a rank-constrained linear representation of the dynamical system. lrDMD is shown to have lower data reconstruction error compared to standard optimal mode decomposition (OMD) and dynamic mode decomposition (DMD), but with an increased computational cost arising from solving a non-convex matrix optimization problem. We demonstrate model order reduction in the complex linearized Ginzburg–Landau equation in the globally unstable regime and on the unsteady flow over a flat plate at a high angle of attack. In both cases, low-dimensional full-state feedback controller is constructed using reduced-order models constructed using DMD, OMD and lrDMD. It is shown that lrDMD stabilizes the Ginzburg–Landau system with a lower- order controller and is able to suppress vortex shedding from an inclined flat plate at a cost lower than either DMD or OMD. It is further shown that lrDMD yields an improved estimate of the adjoint system, for a given rank, relative to DMD and OMD.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8

Similar content being viewed by others

References

  1. Kim, J., Bodony, D.J., Freund, J.B.: Adjoint-based control of loud events in a turbulent jet. J. Fluid Mech. 741, 28–59 (2014)

    Article  MathSciNet  Google Scholar 

  2. Rowley, C.W., Dawson, S.T.M.: Model reduction for flow analysis and control. Annu. Rev. Fluid Mech. 49, 387–417 (2017)

    Article  MathSciNet  MATH  Google Scholar 

  3. Carlberg, K., Bou-Mosleh, C., Farhat, C.: Efficient non-linear model reduction via a least-squares petrov-Galerkin projection and compressive tensor approximations. Int. J. Numer. Methods Eng. 86(2), 155–181 (2011)

    Article  MathSciNet  MATH  Google Scholar 

  4. Willcox, K., Peraire, J.: Balanced model reduction via the proper orthogonal decomposition. AIAA J. 40(11), 2323–2330 (2002)

    Article  Google Scholar 

  5. Rowley, C.W.: Model reduction for fluids, using balanced proper orthogonal decomposition. Int. J. Bifurc. Chaos 15(3), 997–1013 (2005)

    Article  MathSciNet  MATH  Google Scholar 

  6. Bagheri, S., Henningson, D.S., Hoepffner, J., Schmid, P.J.: Input-output analysis and control design applied to a linear model of spatially developing flows. Appl. Mech. Rev. 62(2), 020803 (2009)

    Article  Google Scholar 

  7. Semeraro, O., Bagheri, S., Brandt, L., Henningson, D.S.: Feedback control of three-dimensional optimal disturbances using reduced-order models. J. Fluid Mech. 677, 63–102 (2011)

    Article  MathSciNet  MATH  Google Scholar 

  8. Semeraro, O., Bagheri, S., Brandt, L., Henningson, D.S.: Transition delay in a boundary layer flow using active control. J. Fluid Mech. 731, 288–311 (2013)

    Article  MATH  Google Scholar 

  9. Illingworth, S.J.: Model-based control of vortex shedding at low reynolds numbers. Theor. Comput. Fluid Dyn. 30(5), 429–448 (2016)

    Article  Google Scholar 

  10. Kim, J., Bewley, T.R.: A linear systems approach to flow control. Annu. Rev. Fluid Mech. 39, 383–417 (2007)

    Article  MathSciNet  MATH  Google Scholar 

  11. Huang, S.-C., Kim, J.: Control and system identification of a separated flow. Phys. Fluids 20(10), 101509 (2008)

    Article  MATH  Google Scholar 

  12. Hervé, A., Sipp, D., Schmid, P.J., Samuelides, M.: A physics-based approach to flow control using system identification. J. Fluid Mech. 702, 26–58 (2012)

    Article  MATH  Google Scholar 

  13. Akaike, H.: Fitting autoregressive models for prediction. Ann. Inst. Stat. Math. 21(1), 243–247 (1969)

    Article  MathSciNet  MATH  Google Scholar 

  14. Van Overschee, P., De Moor, B.: N4sid: subspace algorithms for the identification of combined deterministic-stochastic systems. Automatica 30(1), 75–93 (1994)

    Article  MathSciNet  MATH  Google Scholar 

  15. Van Overschee, P., De Moor, B.L.: Subspace Identification for Linear Systems: Theory Implementation Applications. Springer, Berlin (2012)

    MATH  Google Scholar 

  16. Iñigo, J.G., Sipp, D., Schmid, P.J.: A dynamic observer to capture and control perturbation energy in noise amplifiers. J. Fluid Mech. 758, 728–753 (2014)

    Article  MathSciNet  Google Scholar 

  17. Qin, S.J.: An overview of subspace identification. Comput. Chem. Eng. 30(10–12), 1502–1513 (2006)

    Article  Google Scholar 

  18. Juang, J.-N., Pappa, R.S.: An eigensystem realization algorithm for modal parameter identification and model reduction. J. Guid. Control Dyn. 8(5), 620–627 (1985)

    Article  MATH  Google Scholar 

  19. Brunton, S.L., Dawson, S.T.M., Rowley, C.W.: State-space model identification and feedback control of unsteady aerodynamic forces. J. Fluids Struct. 50, 253–270 (2014)

    Article  Google Scholar 

  20. Flinois, T.L.B., Morgans, A.S.: Feedback control of unstable flows: a direct modelling approach using the eigensystem realisation algorithm. J. Fluid Mech. 793, 41–78 (2016)

    Article  MathSciNet  MATH  Google Scholar 

  21. Ahuja, S., Rowley, C.W.: Feedback control of unstable steady states of flow past a flat plate using reduced-order estimators. J. Fluid Mech. 645, 447–478 (2010)

    Article  MathSciNet  MATH  Google Scholar 

  22. Illingworth, S.J., Morgans, A.S., Rowley, C.W.: Feedback control of cavity flow oscillations using simple linear models. J. Fluid Mech. 709, 223–248 (2012)

    Article  MathSciNet  MATH  Google Scholar 

  23. Belson, B.A., Semeraro, O., Rowley, C.W., Henningson, D.S.: Feedback control of instabilities in the two-dimensional blasius boundary layer: the role of sensors and actuators. Phys. Fluids 25(5), 054106 (2013)

    Article  Google Scholar 

  24. Illingworth, S.J., Naito, H., Fukagata, K.: Active control of vortex shedding: an explanation of the gain window. Phys. Rev. E 90(4), 043014 (2014)

    Article  Google Scholar 

  25. Schmid, P.J.: Dynamic mode decomposition of numerical and experimental data. J. Fluid Mech. 656, 5–28 (2010)

    Article  MathSciNet  MATH  Google Scholar 

  26. Tu, J.H., Rowley, C.W., Luchtenburg, D.M., Brunton, S.L., Kutz, J.N.: On dynamic mode decomposition: theory and applications. arXiv preprint arXiv:1312.0041 (2013)

  27. Goulart, P.J., Wynn, A., Pearson, D.: Optimal mode decomposition for high dimensional systems. In: 2012 IEEE 51st Annual Conference on Decision and Control (CDC), pp. 4965–4970. IEEE (2012)

  28. Chen, K.K., Rowley, C.W.: \(\text{ H }_2\) optimal actuator and sensor placement in the linearised complex Ginzburg–Landau system. J. Fluid Mech. 681, 241–260 (2011)

    Article  MathSciNet  MATH  Google Scholar 

  29. Hemati, M.S., Williams, M.O., Rowley, C.W.: Dynamic mode decomposition for large and streaming datasets. Phys. Fluids 26(11), 111701 (2014)

    Article  Google Scholar 

  30. Hemati, M.S., Rowley, C.W., Deem, E.A., Cattafesta, L.N.: De-biasing the dynamic mode decomposition for applied Koopman spectral analysis of noisy datasets. Theor. Comput. Fluid Dyn. 31(4), 349–368 (2017)

    Article  Google Scholar 

  31. Dawson, S.T.M., Hemati, M.S., Williams, M.O., Rowley, C.W.: Characterizing and correcting for the effect of sensor noise in the dynamic mode decomposition. Exp. Fluids 57(3), 42 (2016)

    Article  Google Scholar 

  32. Kramer, B., Peherstorfer, B., Willcox, K.: Feedback control for systems with uncertain parameters using online-adaptive reduced models. SIAM J. Appl. Dyn. Syst. 16(3), 1563–1586 (2017)

    Article  MathSciNet  MATH  Google Scholar 

  33. Deem, E.A., Cattafesta, L.N., Yao, H., Hemati, M., Zhang, H., Rowley, C.W.: Experimental implementation of modal approaches for autonomous reattachment of separated flows. In: 2018 AIAA Aerospace Sciences Meeting, p. 1052 (2018)

  34. Hemati, M., Deem, E., Williams, M., Rowley, C.W., Cattafesta, L.N.: Improving separation control with noise-robust variants of dynamic mode decomposition. In: 54th AIAA Aerospace Sciences Meeting, p. 1103 (2016)

  35. Bhattacharjee, D., Hemati, M., Klose, B., Jacobs, G.: Optimal actuator selection for airfoil separation control. In: 2018 Flow Control Conference, p. 3692 (2018)

  36. Sashittal, P., Bodony, D.: Low-rank dynamic mode decomposition using Riemannian manifold optimization. In: 2018 IEEE Conference on Decision and Control (CDC), pp. 2265–2270 (2018)

  37. Héas, P., Herzet, C.: Low-rank approximation and dynamic mode decomposition. arXiv preprint arXiv:1610.02962 (2016)

  38. Wynn, A., Pearson, D.S., Ganapathisubramani, B., Goulart, P.J.: Optimal mode decomposition for unsteady flows. J. Fluid Mech. 733, 473–503 (2013)

    Article  MathSciNet  MATH  Google Scholar 

  39. Absil, P.-A., Mahony, R., Sepulchre, R.: Optimization Algorithms on Matrix Manifolds. Princeton University Press, Princeton (2009)

    MATH  Google Scholar 

  40. Edelman, A., Arias, T.A., Smith, S.T.: The geometry of algorithms with orthogonality constraints. SIAM J. Matrix Anal. Appl. 20(2), 303–353 (1998)

    Article  MathSciNet  MATH  Google Scholar 

  41. Boumal, N., Mishra, B., Absil, P.-A., Sepulchre, R., et al.: Manopt, a matlab toolbox for optimization on manifolds. J. Mach. Learn. Res. 15(1), 1455–1459 (2014)

    MATH  Google Scholar 

  42. Absil, P.-A., Baker, C.G., Gallivan, K.A.: Trust-region methods on Riemannian manifolds. Found. Comput. Math. 7(3), 303–330 (2007)

    Article  MathSciNet  MATH  Google Scholar 

  43. Kwakernaak, H., Sivan, R.: Linear Optimal Control Systems, vol. 1. Wiley, New York (1972)

    MATH  Google Scholar 

  44. Simoncini, V.: Analysis of the rational Krylov subspace projection method for large-scale algebraic Riccati equations. SIAM J. Matrix Anal. Appl. 37(4), 1655–1674 (2016)

    Article  MathSciNet  MATH  Google Scholar 

  45. Alla, A., Simoncini, V.: Order reduction approaches for the algebraic Riccati equation and the LQR problem. arXiv preprint arXiv:1711.01077 (2017)

  46. Schmid, P.J., Henningson, D.S.: Stability and Transition in Shear Flows, vol. 142. Springer, Berlin (2012)

    MATH  Google Scholar 

  47. Schmid, P.J., Brandt, L.: Analysis of fluid systems: stability, receptivity, sensitivitylecture notes from the flow-nordita summer school on advanced instability methods for complex flows, Stockholm, Sweden, 2013. Appl. Mech. Rev. 66(2), 024803 (2014)

    Article  Google Scholar 

  48. Luchini, P., Bottaro, A.: Adjoint equations in stability analysis. Annu. Rev. Fluid Mech. 46, 493–517 (2014)

    Article  MathSciNet  MATH  Google Scholar 

  49. Natarajan, M., Freund, J.B., Bodony, D.J.: Actuator selection and placement for localized feedback flow control. J. Fluid Mech. 809, 775–792 (2016)

    Article  MathSciNet  MATH  Google Scholar 

  50. Colonius, T., Taira, K.: A fast immersed boundary method using a nullspace approach and multi-domain far-field boundary conditions. Comput. Methods Appl. Mech. Eng. 197(25–28), 2131–2146 (2008)

    Article  MathSciNet  MATH  Google Scholar 

  51. Ahuja, S.: Reduction Methods for Feedback Stabilization of Fluid Flows. PhD thesis, PhD thesis, Princeton University, NJ (2009)

  52. Proctor, J.L., Brunton, S.L., Kutz, J.N.: Dynamic mode decomposition with control. SIAM J. Appl. Dyn. Syst. 15(1), 142–161 (2016)

    Article  MathSciNet  MATH  Google Scholar 

  53. Gugercin, S., Antoulas, A.C.: A survey of model reduction by balanced truncation and some new results. Int. J. Control 77(8), 748–766 (2004)

    Article  MathSciNet  MATH  Google Scholar 

  54. Schönemann, P.H.: A generalized solution of the orthogonal procrustes problem. Psychometrika 31(1), 1–10 (1966)

    Article  MathSciNet  MATH  Google Scholar 

Download references

Acknowledgements

The authors would like to thank Mr. Daniel Floryan and Dr. Clarence Rowley for their help with the code used to simulate flow past an inclined flat plate. This work was sponsored, in part, by the Office of Naval Research (ONR) as part of the Multidisciplinary University Research Initiatives (MURI) Program, under Grant Number N00014-16-1-2617.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Palash Sashittal.

Additional information

Communicated by Jeff D. Eldredge.

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Appendices

Appendix A: Numerical methods

In this section, we present the two methods to solve the lrDMD optimization problem.

A.1 Subspace projection method

In this method, we use iterative subspace projection to find a good approximation of the optimal solution. We first make the observation that \(Q_R\) is an orthogonal projection matrix in the column space of \(X^TR\). This means that there exists an orthogonal matrix \(C_R\) such that \(C_RC_R^T = Q_R\) and \(\text {Im}(C_R) = \text {Im}(Q_R)\). Substituting \(Q_R\) with \(C_RC_R^T\), we get the following cost function,

$$\begin{aligned} G(L,R) = -\left||L^TYC_R\right||_F^2. \end{aligned}$$

with the constraint that \(\text {Im}(C_R) = \text {Im}(Q_R)\). If \(C_R\) is fixed, the optimal solution for L is given by the left singular vectors of \(YC_R\). Finding \(C_R\) for a fixed L under the given constraint is not as trivial. If the left singular vectors of \(X^TR\) span the same space as the left singular vectors of \(Y^TL\), then optimal \(C_R\) under the constraint will just be the left singular vectors of \(X^TR\). As an approximation, we try to find R that minimizes the distance between \(X^TR\) and \(Y^TL\) by solving the following ‘Orthogonal Procrustes Problem’ [54]

$$\begin{aligned} \min _R \left||Y^TL - X^TR\right||_F^2. \end{aligned}$$

The closed form optimal solution to this problem is \(R = UV^T\) where U and V are the left and the right singular vector matrices, respectively, from the singular vector decomposition of \(XY^TL\). \(C_R\) is given by the orthogonal basis of the column space of \(X^TR\) denoted by \(\Pi (X^TR)\). Algorithm 1 describes all the steps for this method.

The solution provided by this algorithm relies heavily on the initial guess. We only need an initial guess for R or, effectively, \(C_R\). In our study, we choose the r leading left singular vectors of X as the initial guess for R. Note that this initial guess is the same as the projection subspace used in DMD. Due to the optimal choice of L for a fixed \(C_R\), the chosen initial guess ensures that the algorithm provides a solution with a reconstruction error at most as high as the error in DMD reconstruction of the same rank.

figure a

A.2 Gradient descent method

In this section, we describe the Riemannian gradient descent method employed to solve the optimization problem. For a thorough review on Riemannian optimization on the Grassmanian manifold, the reader is referred to [40]. For a function F(L) defined on the Grassmanian manifold where \(L\in \mathcal {G}_{r,m}\), the gradient \(\text {grad}\, F(L) \in \mathcal {T}_{L}\mathcal {G}_{r,m}\) (tangent space to the manifold at the point L) is defined as

$$\begin{aligned} \text {grad}\,F(L)&= \nabla F(L) - LL^T\nabla F(L) \end{aligned}$$
(17)
$$\begin{aligned} \text {where}\quad (\nabla F(L))_{ij}&= \frac{d F}{dL_{ij}} \end{aligned}$$
(18)

Similarly, the action of the Hessian on any tangent vector \(dL \in \mathcal {T}_L\mathcal {G}_{r,m}\) is defined as

$$\begin{aligned} \text {Hess}\,F(L)[dL]&= (I_m - LL^T)\nabla ^2F(L)[dL] - dL(L^T\nabla F(L)) \end{aligned}$$
(19)
$$\begin{aligned} \text {where}\quad (\nabla ^2 F(L))_{ij,kl}&= \frac{d^2F}{dL_{ij}dL_{kl}}. \end{aligned}$$
(20)

Further details of the method and its implementation can be seen in [36]. We employ the trust-region algorithm [42] to solve the optimization problem using the MATLAB package ManOpt [41].

Appendix B: lrDMD with control

The lrDMD framework can easily be extended to incorporate the effects of control. Consider a dynamical system with state vector \(x\in \mathbb {R}^m\) such that

$$\begin{aligned} x_{k+1} = f(x_k,u_k), \end{aligned}$$

where \(u \in \mathbb {R}^p\) is the control input and the subscripts denote the time iteration. We have access to a sequence of time snapshots of the state vector which we represent as a data matrices \(X, Y \in \mathbb {R}^{m\times n}\) formed by n pairs of data snapshots as follows,

$$\begin{aligned} X:=(x_1|\cdots | x_n), \quad Y:=(x_2|\cdots |x_{n+1}). \end{aligned}$$

We additionally have access to a sequence of control inputs, which we arrange in the following matrix

$$\begin{aligned} \Omega :=(u_1|\cdots | u_n). \end{aligned}$$

Our goal is to use matrices XY and \(\Omega \) to obtain a low-order approximation of the function \(f(\cdot )\). To this end, we construct a linear approximation of f in both the state x and u such that

$$\begin{aligned} f(x,u) = \hat{A}x + \hat{B}u. \end{aligned}$$

Along with a rank-constraint on the system dynamical matrix, we get the following optimization problem

$$\begin{aligned} \min _{L,D,R,B} \left||Y - LDR^TX - LB\Omega \right||_F^2, \end{aligned}$$
(21)

where \(D\in \mathbb {R}^{r\times r}\), \(B\in \mathbb {R}^{r\times p}\), \(L,R\in \mathbb {R}^{m\times r}\) and \(L^TL=R^TR=I_r\) (\(r\times r\) identity matrix) such that \(\hat{A} = LDR^T\) is the r-ranked matrix approximating the state dynamics and \(\hat{B} = LB\) accounts for the effect of control input on the state evolution.

For a fixed L and R, we observe that the objective function is convex for both D and B. The optimality conditions for the solution of D and B for fixed L and R are

$$\begin{aligned} D(R^TXX^TR) + B(\Omega X^TR)&= L^TYX^TR, \\ B(\Omega \Omega ^T) + D(R^TX\Omega ^T)&= L^TY\Omega ^T, \end{aligned}$$

which can be rewritten as

$$\begin{aligned} \left[ D \;\;\;\; B \right] \left[ \begin{array}{cc} R^TXX^TR \;\;\;\;&{} \Omega X^TR \\ R^TX\Omega ^T \;\;\;\;&{} \Omega \Omega ^T \end{array}\right] = \left[ L^TYX^TR \;\;\;\; L^TY\Omega ^T \right] \end{aligned}$$
(22)

We propose two ways to solve this problem. Both methods work with an initial guess for the optimal LR which can be obtained from leading singular vectors of the data matrices as shown in the main text. The first method is alternative minimization by solving for optimal (DB) for fixed (LR) by solving Eq. () followed by gradient-based minimization of the objective function () for fixed (DB). The second method is to use Schur complements to get closed form expressions for the optimal \((D^*(L,R),B^*(L,R))\), substitute that in the objective function () and use gradient-based methods to find optimal (LR) solution. We will explore these avenues in future studies.

Appendix C: Projection error comparison with ERA

ERA [18] is a system identification method proposed for linear systems. Consider a linear system with state variable \(x \in \mathbb {R}^{m}\) and control input \(u\in \mathbb {R}^{p}\) governed by the equation

$$\begin{aligned} x_{k+1} = Ax_{k} + Bu_{k}, \end{aligned}$$

where subscripts denote the time iteration, \(A\in \mathbb {R}^{m \times m}\) is the state-transition matrix and \(B\in \mathbb {R}^{m\times p}\) captures the effect of control on the state variable. The data matrix \(X\in \mathbb {R}^{m\times p(m_c+1)}\) using impulse response of this system will be

$$\begin{aligned} X = \left[ B\,\,AB\,\,\cdots \,\,A^{m_c}B\right] \end{aligned}$$

where \(m_c + 1\) is the number of snapshots.

The first step of the ERA method is to form Hankel matrices from the impulse response data of the system. We construct a generalized Hankel matrix \(H\in \mathbb {R}^{m(m_o+1)\times p(m_c+1)}\)

$$\begin{aligned} H = \left[ \begin{array}{cccc} B &{} AB &{} \cdots &{} A^{m_c}B \\ AB &{} A^2B &{} \cdots &{} A^{m_c + 1}B \\ \vdots &{} \vdots &{} \ddots &{} \vdots \\ A^{m_o}B &{} A^{m_o+1}B &{} \cdots &{} A^{m_o + m_c}B \end{array}\right] \end{aligned}$$

where \(m_c\) and \(m_o\) are chosen such that \(m_c + m_o \le m\). We then compute the singular value decomposition of H to obtain left and right singular vectors \(U\in \mathbb {R}^{m(m_o+1)\times p(m_c+1)}\) and \(V\in \mathbb {R}^{p(m_c+1)\times p(m_c+1)}\), and the diagonal matrix with decreasing singular values \(\Sigma \in \mathbb {R}^{p(m_c+1)\times p(m_c+1)}\), such that

$$\begin{aligned} H = U\Sigma V^T. \end{aligned}$$

The primal modes \(\phi \in \mathbb {R}^{m\times p(m_c+1)}\) and adjoint modes \(\psi \in \mathbb {R}^{p(m_c+1)}\) are given by

$$\begin{aligned} \phi&= XV\Sigma ^{-1/2} \\ \psi&= \phi (\phi ^T\phi )^{-1}. \end{aligned}$$

To compare the projection error for r-ranked reduced-order model, we consider the r leading columns \(\phi \) and \(\psi \) to get \(\phi _r\) and \(\psi _r\), respectively. The projection errors for eigenmode v and adjoint mode w are given by \(\epsilon (v,\phi _r)\) and \(\epsilon (w,\psi _r)\) where \(\epsilon (\cdot )\) is defined in Equation (14). For consistency with Sect. 5.1.1 we use \(m_c = 14\), so that we use the same number of snapshots in the data matrix X. \(m_o\) is chosen to be 4. The comparison of projection error of lrDMD, OMD and DMD with ERA is shown in Fig. 9. Even though ERA has the same projection error for the unstable eigenmode compared to DMD, it shows significant improvement in the unstable adjoint mode projection error. As shown in Fig. 9b, lrDMD outperforms ERA only at ranks higher than 5. However, Fig. 10 shows that increasing \(m_o\) improves ERA performance in adjoint projection error and even outperforms lrDMD although having the same higher error in eigenmode projection. This shows that adding delay coordinates by increasing \(m_o\) can significantly decrease the adjoint projection error but does not affect the eigenmode projection.

Fig. 9
figure 9

a Projection error for the unstable eigenmode and b projection error for the unstable adjoint mode for DMD (red, circle), OMD (blue, cross), lrDMD (green, square) and ERA (black, diamond) for different rank approximations (colour figure online)

Fig. 10
figure 10

a Projection error for the unstable eigenmode and b projection error for the unstable adjoint mode for lrDMD (green, square), ERA with \(m_o =4 \) (black, diamond) and ERA with \(m_o = 15\) for different rank approximations (colour figure online)

Fig. 11
figure 11

\(C_D\) versus time for \(35^{\circ }\) inclined flat plate with the actuator activated at \(t_0 = 210\) and control applied every 20 timesteps using DMD-, OMD- and lrDMD-based reduced-order controllers learned using snapshots 10 timesteps apart

Appendix D: Effect of different learning and control timesteps

In this section, examine what happens when the learning and control timesteps differ for DMD, OMD and lrDMD. Let the resulting reduced- order model using any one of the three methods be \(\widehat{A}\). We generate controllers that would be applied to the flow every 20 timepsteps, and therefore, we build reduced-order controllers using the model \(\widehat{A}^2\). We do this for DMD, OMD and lrDMD while keeping all the parameters the same as in Sect. 5.2.3. Figure 11 shows the controller performance when the controllers are switched on at \(t_0 = 210\). This time, DMD and OMD outperform lrDMD in stabilizing the system and driving it to the steady state of the flow. This is because, while DMD and OMD keep dynamics confined to one low-dim subspace, lrDMD has different input and output subspaces that are optimal for dynamics discretized for \(T = 10\Delta t\). Applying the reduced-order model for 20 timestep requires projection from subspace L to R since

$$\begin{aligned} (LDR^T)^2 = L(DR^TLD)R^T. \end{aligned}$$

We think this projection leads to a deterioration of the controller performance. In practice, the input and output subspace learned from snapshots that are 10 timesteps apart should be used as an initial condition to get optimal subspaces for dynamics discretized for \(T = 20\Delta t\).

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Sashittal, P., Bodony, D.J. Reduced-order control using low-rank dynamic mode decomposition. Theor. Comput. Fluid Dyn. 33, 603–623 (2019). https://doi.org/10.1007/s00162-019-00508-9

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00162-019-00508-9

Keywords

Navigation