Skip to main content

Advertisement

Log in

Feedback control of transitional shear flows: sensor selection for performance recovery

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

Abstract

The choice and placement of sensors and actuators is an essential factor determining the performance that can be realized using feedback control. This determination is especially important, but difficult, in the context of controlling transitional flows. The highly non-normal nature of the linearized Navier–Stokes equations makes the flow sensitive to small perturbations, with potentially drastic performance consequences on closed-loop flow control. Full-information controllers, such as the linear quadratic regulator (LQR), have demonstrated some success in reducing transient energy growth and suppressing transition; however, sensor-based output feedback controllers with comparable performance have been difficult to realize. In this study, we propose two methods for sensor selection that enable sensor-based output feedback controllers to recover full-information control performance: one based on a sparse controller synthesis approach, and one based on a balanced truncation procedure for model reduction. Both approaches are investigated within linear and nonlinear simulations of a sub-critical channel flow with blowing and suction actuation at the walls. We find that sensor configurations identified by both approaches allow sensor-based static output feedback LQR controllers to recover full-information LQR control performance, both in reducing transient energy growth and suppressing transition. Further, our results indicate that both the sensor selection methods and the resulting controllers exhibit robustness to Reynolds number variations.

Graphic abstract

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
Fig. 9
Fig. 10
Fig. 11
Fig. 12
Fig. 13
Fig. 14
Fig. 15
Fig. 16
Fig. 17

Similar content being viewed by others

Notes

  1. Note that what we refer to as “streamwise disturbances” in this study are spanwise constant in nature and exhibit waviness in the streamwise direction. Likewise, what we refer to as “spanwise disturbances” are streamwise constant and exhibit waviness in the spanwise direction.

References

  1. Schmid, P.J., Henningson, D.S.: Stability and Transition in Shear Flows. Springer, New York (2001)

    Book  Google Scholar 

  2. Schmid, P.J.: Nonmodal stability theory. Annu. Rev. Fluid Mech. 39(1), 129–162 (2007)

    Article  MathSciNet  Google Scholar 

  3. Landahl, M.T.: A note on an algebraic instability of inviscid parallel shear flows. J. Fluid Mech. 98(2), 243–251 (1980). https://doi.org/10.1017/S0022112080000122

    Article  MathSciNet  MATH  Google Scholar 

  4. Farrell, B.F., Ioannou, P.J.: Optimal excitation of three-dimensional perturbations in viscous constant shear flow. Phys. Fluids A 5(6), 1390–1400 (1993). https://doi.org/10.1063/1.858574

    Article  MATH  Google Scholar 

  5. Reddy, S.C., Henningson, D.S.: Energy growth in viscous channel flows. J. Fluid Mech. 252, 209–238 (1993)

    Article  MathSciNet  Google Scholar 

  6. Trefethen, L.N., Trefethen, A.E., Reddy, S.C., Driscoll, T.A.: Hydrodynamic stability without eigenvalues. Science 261(5121), 578–584 (1993)

    Article  MathSciNet  Google Scholar 

  7. Jovanović, M.R., Bamieh, B.: Componentwise energy amplification in channel flows. J. Fluid Mech. 534, 145–183 (2005)

    Article  MathSciNet  Google Scholar 

  8. Bamieh, B., Dahleh, M.: Energy amplification in channel flows with stochastic excitation. Phys. Fluids 13(11), 3258–3269 (2001). https://doi.org/10.1063/1.1398044

    Article  MathSciNet  MATH  Google Scholar 

  9. Butler, K.M., Farrell, B.F.: Three-dimensional optimal perturbations in viscous shear flow. Phys. Fluids A 4, 1637–1650 (1992)

    Article  Google Scholar 

  10. Hemati, M.S., Yao, H.: Performance limitations of observer-based feedback for transient energy growth suppression. AIAA J. 56(6), 2119–2123 (2018). https://doi.org/10.2514/1.J056877

    Article  Google Scholar 

  11. Högberg, M., Bewley, T.R., Henningson, D.S.: Linear feedback control and estimation of transition in plane channel flow. J. Fluid Mech. 481, 149–175 (2003)

    Article  MathSciNet  Google Scholar 

  12. Ilak, M., Rowley, C.W.: Feedback control of transitional channel flow using balanced proper orthogonal decomposition. In: 5th AIAA Theoretical Fluid Mechanics Conference (2008)

  13. Martinelli, F., Quadrio, M., McKernan, J., Whidborne, J.F.: Linear feedback control of transient energy growth and control performance limitations in subcritical plane Poiseuille flow. Phys. Fluids 23(1), 014103 (2011)

    Article  Google Scholar 

  14. Sun, Y., Hemati, M.S.: Feedback control for transition suppression in direct numerical simulations of channel flow. Energies 12(21), 4127 (2019)

    Article  Google Scholar 

  15. Kalur, A., Hemati, M.S.: Control-oriented model reduction for minimizing transient energy growth in shear flows. AIAA J. 58, 1034–1045 (2019)

    Article  Google Scholar 

  16. Yao, H., Hemati, M.S.: Revisiting the separation principle for improved transition control. In: 2018 Flow Control Conference. AIAA paper (2018)

  17. Bewley, T.R., Liu, S.: Optimal and robust control and estimation of linear paths to transition. J. Fluid Mech. 365, 305–349 (1998)

    Article  MathSciNet  Google Scholar 

  18. Hœpffner, J., Chevalier, M., Bewley, T.R., Henningson, D.S.: State estimation in wall-bounded flow systems. part 1 perturbed laminar flows. J. Fluid Mech. 534, 263–294 (2005). https://doi.org/10.1017/S0022112005004210

    Article  MathSciNet  MATH  Google Scholar 

  19. Toivonen, H.T., Mäkilä, P.M.: A descent Anderson-Moore algorithm for optimal decentralized control. Automatica 21(6), 743–744 (1985). https://doi.org/10.1016/0005-1098(85)90048-2

    Article  MATH  Google Scholar 

  20. Syrmos, V.L., Abdallah, C.T., Dorato, P., Grigoriadis, K.: Static output feedback: a survey. Automatica 33(2), 125–137 (1997). https://doi.org/10.1016/S0005-1098(96)00141-0

    Article  MathSciNet  MATH  Google Scholar 

  21. Cao, Y.-Y., Lam, J., Sun, Y.-X.: Static output feedback stabilization: an ILMI approach. Automatica 34(12), 1641–1645 (1998). https://doi.org/10.1016/S0005-1098(98)80021-6

    Article  MATH  Google Scholar 

  22. Yao, H., Hemati, M.S.: Advances in output feedback control of transient energy growth in a linearized channel flow. In: AIAA Scitech 2019 Forum. AIAA paper (2019)

  23. Whidborne, J.F., McKernan, J.: On the minimization of maximum transient energy growth. IEEE Trans. Autom. Control 52(9), 1762–1767 (2007). https://doi.org/10.1109/TAC.2007.900854

    Article  MathSciNet  MATH  Google Scholar 

  24. Yao, H., Sun, Y., Mushtaq, T., Hemati, M.S.: Reducing transient energy growth in a channel flow using static output feedback control. AIAA J. (2022). https://doi.org/10.2514/1.J061345

    Article  Google Scholar 

  25. Willcox, K.: Unsteady flow sensing and estimation via the gappy proper orthogonal decomposition. Comput. Fluids 35(2), 208–226 (2006). https://doi.org/10.1016/j.compfluid.2004.11.006

    Article  MATH  Google Scholar 

  26. Manohar, K., Brunton, B.W., Kutz, J.N., Brunton, S.L.: Data-driven sparse sensor placement for reconstruction: demonstrating the benefits of exploiting known patterns. IEEE Control Syst. Mag. 38(3), 63–86 (2018). https://doi.org/10.1109/MCS.2018.2810460

    Article  MathSciNet  MATH  Google Scholar 

  27. Manohar, K., Kutz, J.N., Brunton, S.L.: Optimal sensor and actuator placement using balanced model reduction. arXiv pre-print arXiv:1812.01574 (2018)

  28. Clark, E., Askham, T., Brunton, S.L., Kutz, J.N.: Greedy sensor placement with cost constraints. IEEE Sens. J. 19(7), 2642–2656 (2019)

    Article  Google Scholar 

  29. Saito, Y., Nonomura, T., Yamada, K., Asai, K., Sasaki, Y., Tsubakino, D.: Determinant-based fast greedy sensor selection algorithm. arXiv pre-print arXiv:1911.08757 (2020)

  30. Yamada, K., Saito, Y., Nankai, K., Nonomura, T., Asai, K., Tsubakino, D.: Fast greedy optimization of sensor selection in measurement with correlated noise. arXiv pre-print arXiv:1912.01776 (2020)

  31. Natarajan, M., Freund, J.B., Bodony, D.J.: Actuator selection and placement for localized feedback flow control. J. Fluid Mech. 809, 775–792 (2016). https://doi.org/10.1017/jfm.2016.700

    Article  MathSciNet  MATH  Google Scholar 

  32. Chen, K.K., Rowley, C.W.: H2 optimal actuator and sensor placement in the linearised complex Ginzburg–Landau system. J. Fluid Mech. 681, 241–260 (2011). https://doi.org/10.1017/jfm.2011.195

    Article  MathSciNet  MATH  Google Scholar 

  33. Oehler, S.F., Illingworth, S.J.: Sensor and actuator placement trade-offs for a linear model of spatially developing flows. J. Fluid Mech. 854, 34–55 (2018). https://doi.org/10.1017/jfm.2018.590

    Article  MATH  Google Scholar 

  34. Bhattacharjee, D., Klose, B., Jacobs, G.B., Hemati, M.S.: Data-driven selection of actuators for optimal control of airfoil separation. Theor. Comput. Fluid Dyn. 34, 557–575 (2020). https://doi.org/10.1007/s00162-020-00526-y

    Article  MathSciNet  Google Scholar 

  35. McKernan, J., Papadakis, G., Whidborne, J.F.: A linear state-space representation of plane Poiseuille flow for control design: a tutorial. Int. J. Model. Ident. Control 1(4), 272–280 (2006)

    Article  Google Scholar 

  36. Butler, K.M., Farrell, B.F.: Three-dimensional optimal perturbations in viscous shear flow. Phys. Fluids A 4(8), 1637–1650 (1992). https://doi.org/10.1063/1.858386

    Article  Google Scholar 

  37. Whidborne, J.F., Amar, N.: Computing the maximum transient energy growth. BIT Numer. Math. 51(2), 447–457 (2011)

    Article  MathSciNet  Google Scholar 

  38. Brogan, W.L.: Modern Control Theory. Prentice Hall, Upper Saddle River (1991)

    MATH  Google Scholar 

  39. Rautert, T., Sachs, E.W.: Computational design of optimal output feedback controllers. SIAM J. Optim. 7(3), 837–852 (1997)

    Article  MathSciNet  Google Scholar 

  40. Lin, F., Fardad, M., Jovanović, M.R.: Design of optimal sparse feedback gains via the alternating direction method of multipliers. IEEE Trans. Autom. Control 58(9), 2426–2431 (2013). https://doi.org/10.1109/TAC.2013.2257618

    Article  MathSciNet  MATH  Google Scholar 

  41. Polyak, B.T., Khlebnikov, M.V., Shcherbakov, P.S.: Sparse feedback in linear control systems. Autom. Remote. Control 75(12), 2099–2111 (2014). https://doi.org/10.1134/S0005117914120029

    Article  MathSciNet  MATH  Google Scholar 

  42. Moore, B.: Principal component analysis in linear systems: controllability, observability, and model reduction. IEEE Trans. Autom. Control 26(1), 17–32 (1981). https://doi.org/10.1109/TAC.1981.1102568

    Article  MathSciNet  MATH  Google Scholar 

  43. Antoulas, A.C.: Approximation of Large-Scale Dynamical Systems. Society for Industrial and Applied Mathematics, Philadelphia (2005). https://doi.org/10.1137/1.9780898718713

    Book  MATH  Google Scholar 

  44. Laub, A., Heath, M., Paige, C., Ward, R.: Computation of system balancing transformations and other applications of simultaneous diagonalization algorithms. IEEE Trans. Autom. Control 32(2), 115–122 (1987). https://doi.org/10.1109/TAC.1987.1104549

    Article  MATH  Google Scholar 

  45. Gibson, J.F., Halcrow, J., Cvitanović, P.: Visualizing the geometry of state space in plane Couette flow. J. Fluid Mech. 611, 107–130 (2008)

    Article  MathSciNet  Google Scholar 

  46. Gibson, J.F.: Channelflow: A spectral Navier–Stokes simulator in C++. Technical report, U. New Hampshire (2014). Channelflow.org

  47. Reddy, S.C., Schmid, P.J., Baggett, J.S., Henningson, D.S.: On stability of streamwise streaks and transition thresholds in plane channel flows. J. Fluid Mech. 365, 269–303 (1998)

    Article  MathSciNet  Google Scholar 

  48. Hunt, J.C.R., Wray, A.A., Moin, P.: Eddies, streams, and convergence zones in turbulent flows. In: Proceedings of the Summer Program, pp. 193–208. Center of Turbulence Research (1988)

  49. Leibfritz, F., Lipinski, W.: Description of the benchmark examples in compleib 1.0, technical report. Univ. of Trier (2003). www.complib.de

Download references

Acknowledgements

This material is based upon work supported by the Air Force Office of Scientific Research under award number FA9550-19-1-0034, monitored by Dr. Gregg Abate, and the National Science Foundation under award number CBET-1943988, monitored by Dr. Ronald D. Joslin.

Author information

Authors and Affiliations

Authors

Corresponding authors

Correspondence to Huaijin Yao or Maziar S. Hemati.

Ethics declarations

Conflict of interest

The authors report no conflict of interest.

Additional information

Publisher's Note

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

Appendices

Appendix A Anderson-Moore algorithm with Armijo-type adaptation

Algorithm 3 presents the Anderson-Moore algorithm with Armijo-type adaptation that can be used to solve an SOF-LQR controller [16, 24]. The SOF-LQR problem consists of a coupled set of nonlinear matrix equations whose solution may not yield a global optimal solution. The Anderson-Moore algorithm is an iterative method for computing a locally optimal solution. The algorithm requires an initial stabilizing SOF control, which can influence the resulting SOF-LQR controller. In this work, Algorithm 3 is initialized using a SOF control gain determined using an iterative linear matrix inequality approach [21, 24]. Subsequently, the complexity of a single iteration of the Anderson-Moore algorithm with Armijo type adaptations is of \({\mathcal {O}}(n^3)\). Additional details regarding the SOF-LQR synthesis methods used here can be found in [24].

figure d

Appendix B A comparison of column-norm evaluation with convex-optimization-based sparse control synthesis

The column-norm evaluation (CE) method proposed in Sect. 3.1 is a simple heuristic that facilitates sensor selection for large-scale problems. The CE heuristic was inspired by the convex-optimization-based methods proposed in [41]. Here we will provide justification for the CE method by comparing with a convex-optimization-based design on a system for which convex synthesis is computationally tractable. Although the CE method is still heuristic in nature, this example will establish that it is at least a reasonable heuristic to consider when computationally intensive convex-optimization-based methods are not tractable.

Consider the linear state-space system,

$$\begin{aligned} {\dot{x}}&= Ax + Bu \end{aligned}$$
(B1)
$$\begin{aligned} y&=Cx, \end{aligned}$$
(B2)

where

$$\begin{aligned} \begin{aligned} A&= \left[ \begin{array}{c@{\quad }c@{\quad }c@{\quad }c@{\quad }c@{\quad }c@{\quad }c@{\quad }c} -0.0046 &{} -0.1978 &{} 0.0039 &{} 0.0133 &{} 0.0127 &{} -0.0285 &{} 0 &{} 0 \\ 0.0380 &{} -0.5667 &{} -0.0029 &{} -0.0014 &{} -0.0100 &{} -0.0232 &{} 0 &{} 0 \\ 0.3259 &{} 0.3570 &{} -0.2947 &{} -0.4076 &{} -0.8152 &{} 0.1064 &{} 1.0000 &{} 0 \\ -0.0045 &{} -0.0378 &{} 0.0070 &{} -0.0654 &{} -0.0397 &{} 0.0709 &{} 0 &{} 0 \\ -0.4020 &{} -0.2149 &{} 0.2266 &{} -0.4093 &{} -0.8210 &{} -0.2786 &{} 0 &{} 1.0000 \\ -0.0730 &{} 0.5683 &{} 0.0148 &{} 0.2674 &{} 0.1442 &{} -0.7396 &{} 0 &{} 0 \\ -9.8100 &{} 0 &{} 0 &{} 0 &{} 0 &{} 0 &{} 0 &{} 0 \\ 0 &{} 0 &{} 0 &{} 9.8100 &{} 0 &{} 0 &{} 0 &{} 0 \end{array} \right] , \\ B&= I_{8\times 8},\\ C&= \left[ \begin{array}{c@{\quad }c@{\quad }c@{\quad }c@{\quad }c@{\quad }c@{\quad }c@{\quad }c} 0.0676 &{} -1.1151 &{} 0.0062 &{} -0.0170 &{} -0.0129 &{} 0.1390 &{} 0 &{} 0 \\ 0.1221 &{} 0.1055 &{} -0.0682 &{} 0.0049 &{} 0.0106 &{} 0.0059 &{} 0 &{} 0 \\ -0.0001 &{} 0.0039 &{} 0.0010 &{} 0.1067 &{} 0.2227 &{} 0.0326 &{} 0 &{} 0 \\ -0.0016 &{} 0.0035 &{} -0.0035 &{} 0.1692 &{} 0.1430 &{} -0.4070 &{} 0 &{} 0 \end{array} \right] . \end{aligned} \end{aligned}$$

The dual of this system was used to demonstrate row-sparse controller synthesis in [41] and was reported as benchmark problem HE3 [49] corresponding to a linearized 8-state model of the Bell 201A-1 Helicopter. We work with this modified version of the problem here so as to study column-sparse controller synthesis, corresponding to sensor selection and the CE method.

Our first step will be to compute an optimal controller according to [41]. Since \(B=I\), this turns out to be a convex problem that can be solved via (B3) in the sparse LQR step of Algorithm B, to be described momentarily. The initial condition is set to \(x_0 = [1,1,...,1]^T\), following the choice in [41]. Solving for the exact optimal control gain \(F^{exact}\), we obtain

$$\begin{aligned} F^\text {exact} = \left[ \begin{array}{c@{\quad }c@{\quad }c@{\quad }c} -0.0181 &{} -0.9803 &{} 0.1325 &{} 0.0028 \\ 0.7672 &{} -0.1281 &{} -0.0674 &{} 0.0205 \\ -1.8853 &{} 19.6594 &{} -0.6169 &{} 1.5326 \\ 0.0896 &{} -0.1614 &{} -0.7120 &{} -0.4899 \\ 0.2274 &{} -0.1401 &{} -6.6221 &{} -3.0483 \\ 0.0154 &{} 0.0517 &{} -0.8353 &{} 0.1169 \\ -1.4943 &{} 21.1077 &{} -4.2401 &{} 0.0429 \\ 0.9987 &{} -4.3681 &{} -11.9121 &{} -6.6448 \end{array} \right] . \end{aligned}$$

The optimal cost associated with this solution is \(J^\text {exact}=1.32 \times 10^3\).

Table 3 Column norm values of the SOF gain matrix F

In order to apply the CE heuristic, we evaluate the column norms of the gain matrix \(F^\text {exact}\) (see Table 3). According to the CE heuristic, the first column of \(F^\text {exact}\) (i.e., the sensor associated with the first row of C) is least important to the closed-loop performance, and so can be removed. By the CE approach, the next candidate for removal would be the second column of \(F^\text {exact}\), or the second row of C. Finally, the last candidate for removal would be column 3 of \(F^\text {exact}\), or the sensor associated with the third row of C. In our example, we opt to remove two sensors. Following the CE sequence identified above, we have

$$\begin{aligned} C^\text {CE} = \left[ \begin{array}{c@{\quad }c@{\quad }c@{\quad }c@{\quad }c@{\quad }c@{\quad }c@{\quad }c} 0.1221 &{} 0.1055 &{} -0.0682 &{} 0.0049 &{} 0.0106 &{} 0.0059 &{} 0 &{} 0 \\ -0.0001 &{} 0.0039 &{} 0.0010 &{} 0.1067 &{} 0.2227 &{} 0.0326 &{} 0 &{} 0 \end{array} \right] . \end{aligned}$$

We now re-solve for the optimal control via (B3), which yields

$$\begin{aligned} F^\text {CE} = \left[ \begin{array}{c@{\quad }c} -0.9823 &{} 0.1772 \\ -0.0393 &{} 0.0313 \\ 19.8284 &{} -0.3111 \\ -0.1582 &{} -0.7959 \\ -0.2287 &{} -7.7153 \\ 0.1156 &{} -0.8048 \\ 21.3852 &{} -4.6557 \\ -4.5979 &{} -13.9553 \end{array} \right] . \end{aligned}$$

The cost associated with this solution will be \(J^\text {CE} = 1.39 \times 10^3 \approx 1.05J^\text {exact}\). Note that to achieve a sparse sensor solution, a small (5%) price was paid on the performance.

Now to compare with a more rigorous (i.e., non-heuristic) solution approach to establish some justification for the CE heuristic. A column-sparse controller can be obtained by working with Algorithm 4, listed at the end of this appendix. This is a convex-optimization-based controller synthesis that seeks a column-sparse design. The row-sparse control synthesis version of this algorithm was proposed in [41]. The basic idea is to relax the original objective function J to allow for some sparsity in the resulting control design. This is reflected in the condition (B4), where \(\xi \) is a user-selected design parameter. A larger \(\xi \) will allow for a sparser controller, while \(\xi \rightarrow 1\) will tend toward the original optimal control solution. For additional details on the derivation of this algorithm, we refer the reader to [41].

In applying Algorithm 4, we set \(\xi = 1.25\) in (B4). This was the same value used in the example reported in [41]. In step 3, we use \(10^{-8}\) for the zeroing threshold. The resulting controller gain is

$$\begin{aligned} F^\text {convex} = \left[ \begin{array}{c@{\quad }c@{\quad }c@{\quad }c} 0 &{} -0.9662 &{} 0.1954 &{} 0 \\ 0 &{} -0.0314 &{} 0.0404 &{} 0 \\ 0 &{} 19.5898 &{} -0.5827 &{} 0 \\ 0 &{} -0.1424 &{} -0.7770 &{} 0 \\ 0 &{} -0.1352 &{} -7.6064 &{} 0 \\ 0 &{} 0.1383 &{} -0.7835 &{} 0 \\ 0 &{} 21.2576 &{} -4.8017 &{} 0 \\ 0 &{} -4.3871 &{} -13.7127 &{} 0 \end{array} \right] . \end{aligned}$$

The associated cost of this column sparse controller is \(J^\text {convex} = 1.39 \times 10^3 \approx 1.05J^\text {exact}\).

Two observations regarding these solutions provide justification for the CE method: (1) the same set of sensors is removed/retained between the CE method and the convex-optimization-based synthesis, and (2) the associated cost function for both the CE and convex-optimization-based methods are comparable (\(1.05J^\text {exact}\)). We note that the specific gain values are different between the two methods. This is an artifact of the manner in which the gains were computed. In the convex-optimization-based control synthesis, the zero rows in the C matrix were not removed, whereas in the CE method they were handling the C matrix in the same manner.

Since the same sparse sensor solution is obtained, a re-design of \(F^\text {convex}\) with those zero rows removed does yield an identical control solution \(F^\text {CE}\) as obtained by the CE method. The converse is also true. We note that the correspondence between the CE method and the convex-optimization-based method is dependent on the user-specified design variable \(\xi \). As such, this correspondence between solutions is not guaranteed. Nonetheless, as we saw in this example, it is possible for the CE method to achieve a comparable solution as the convex-optimization-based approach for particular values of the design parameter \(\xi \).

While we have established some justification for the CE method, this justification is by no means universal. The example problem selected here for demonstration was chosen to be convex. This was done to facilitate comparisons with a unique solution—removing the complications arising from iterative methods and multiple local minima in the context of non-convex problems. Even in this simple context of convex problems, the CE method can have tremendous computational advantages over the convex-optimization-based for larger-scale problems such as fluid flows. The semi-definite programming problem underlying Algorithm 4 will scale with \({\mathcal {O}}(n^6)\) in general. Without a computationally tractable alternative design algorithm, one must begin to consider model reduction (see e.g., [15]). When the problem is non-convex—which is the more realistic case in flow control because the actuation will also be sparse—these computational challenges compound. As such, an efficient heuristic method—such as the CE approach—is most welcome.

figure e
Fig. 18
figure 18

Grid resolution study of transient kinetic energy density E for a streamwise-wave and b spanwise-wave disturbance cases

Appendix C Grid resolution study

Grid resolution studies have been performed for streamwise- and spanwise-wave disturbances, as shown in Fig. 18. The coarse and refined meshes contain \(64\times 101 \times 64\) and \(128\times 101 \times 64\) grid points in the x-, y- and z-direction, respectively. The grids have been tested using different optimal disturbance amplitudes. At the transient energy growth and laminar-to-turbulent transition stages, the results from the coarse and refined meshes overlap before the flows have completely become turbulent. The results in Fig. 18 indicate that the grid resolution is sufficient for the flow condition considered in the present work, which does not require fully resolving the turbulent flow. Because we further examine the transition mechanism in the channel flow, the refined mesh is adopted in the direct numerical simulation to better capture relatively small-scale structures present during the transition process. Moreover, because the flow with oblique-wave disturbance has similar characteristics to the one with streamwise-wave disturbance, the grid selected for streamwise-wave disturbance is sufficient to resolve the flow response to the oblique-wave disturbance.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Yao, H., Sun, Y. & Hemati, M.S. Feedback control of transitional shear flows: sensor selection for performance recovery. Theor. Comput. Fluid Dyn. 36, 597–626 (2022). https://doi.org/10.1007/s00162-022-00616-z

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00162-022-00616-z

Keywords

Navigation